Cardiovascular Anatomy, Physiology, Pathophysiology, and Anesthesia Management

Chapter 23


Cardiovascular Anatomy, Physiology, Pathophysiology, and Anesthesia Management




Cardiovascular System


Knowledge of anatomy and physiology of the cardiovascular system is essential to anesthesia practice. Every anesthetic agent has either a direct or an indirect effect on the cardiovascular system. Therefore whether the clinical concern is about a sympathectomy that causes a decrease in blood pressure during neuraxial blockade or about myocardial depression during inhalation anesthesia, a thorough understanding of these effects and their implications with regard to human physiology is vital if excellent anesthesia care is to be provided.


The cardiovascular system is composed of the heart and the vasculature that carries blood to provide nutrients to all cells in the body. In addition, the cardiovascular system transports substances such as hormones and electrolytes from one part of the body to another.


At the center of this network is the heart. The heart pumps unoxygenated blood to the lungs and then supplies oxygenated blood to all parts of the body. This chapter describes the anatomic and physiologic characteristics, as well as pathophysiologic changes, associated with the cardiovascular system.



Heart





Pericardium


The heart is situated within the mediastinum and surrounded by a fibrous, double-walled sac called the pericardium,2 which envelops the heart and the roots of the great vessels. It consists of a visceral portion, which is in intimate contact with the outer surface of the heart (epicardium), and an outer parietal portion, and adheres to the fibrous pericardium (Figure 23-1).



image


FIGURE 23-1 The pericardium.


The fibrous pericardium is pierced superiorly by the aorta, the pulmonary trunk, and the superior vena cava. The base of the fibrous pericardium is fused with the central tendon of the diaphragm. The visceral pericardium and parietal pericardium are separated by a thin potential space known as the pericardial cavity. This space normally contains approximately 10 to 25 mL of serous fluid, which provides lubrication for the free movement of the heart within the mediastinum. In disease states, the pericardial space can fill with blood and or serosanguinous fluid, compress the heart, and decrease cardiac output (CO). In acute cardiac tamponade, the volume rapidly increases, producing myocardial dysfunction. In contrast, in chronic cardiac tamponade, the degree of pressure exerted on the heart increases slowly because the pericardial sac stretches over time to accommodate the blood that accumulates. However, the pressure may eventually increase as much as 10-fold before symptoms of cardiac tamponade occur.3


The pericardium receives its arterial blood supply from the branches of the internal thoracic arteries and through the bronchial, esophageal, and superior phrenic arteries. Venous drainage from the pericardium occurs through the azygos system and the pericardiophrenic veins, which anastomose with the internal thoracic veins. Nervous innervation to the pericardium is derived from the vagus nerve, the phrenic nerves, and the sympathetic trunks.



Surface Anatomy


The atria are separated from the ventricles by the coronary sulcus (AV sulcus), as seen in Figure 23-2, A. The right coronary artery travels within this sulcus. The circumflex artery arises from the left coronary artery and travels in the coronary sulcus until it branches posteriorly. The RV and LV are separated by the interventricular sulci, which descend from the coronary sulcus to the apex. The interventricular sulci are composed of an anterior interventricular sulcus and a posterior interventricular sulcus. The anterior interventricular sulcus contains the left anterior descending (LAD) artery, which courses over the interventricular septum and continues in the posterior interventricular sulcus.



The crux of the heart is the place at which the coronary and the posterior interventricular sulci meet. Internally, it is where the atrial and ventricular septa meet (Figure 23-2, B). This anatomic crux is important in determining coronary artery dominance.



Cardiac Skeleton


Essential to a discussion of the chambers of the heart is a description of the fibrous skeleton, the annulus fibrosus (Figure 23-3). Tough fibrous rings surround the AV valves and act as points of attachment for the valves. Two additional fibrous annuli develop in relation to the bases of the aorta and the pulmonary trunk. The aortic fibrous annulus is connected to the pulmonary annulus by a fibrous band called the tendon of the conus. The aortic annulus is connected to the AV annuli by the small left fibrous trigone and the larger right fibrous trigone, also called the central fibrous body. The four annuli and their interconnections constitute the fibrous cardiac skeleton.



The annulus fibrosus is the fixation point for the cardiac musculature and plays an important role in the structure, function, and efficiency of the heart. The annulus acts as an insulator to prevent aberrant electrical conduction from the atria to the ventricles so that AV conduction moves through one pathway only: the AV node to the bundle of His. This element increases the electromechanical efficiency of the heart and helps prevent dysrhythmias.



Chambers of the Heart



Right Atrium

The atria act as the priming chambers for the ventricles. As such, the RA acts as a reservoir for the RV and has unique anatomic characteristics. It has a muscle wall thickness of approximately 2 mm.


The RA receives blood from several sources: the superior vena cava, the inferior vena cava, and the coronary sinus (Figure 23-4). The RA consists of two parts: an anterior, thin-walled trabeculated portion and a posterior, smooth-walled portion called the sinus venarum. The sinus venarum receives blood from the vena cava and the coronary sinus. The auricle projects to the left from the root of the superior vena cava and overlaps the root of the ascending aorta.



The superior vena cava returns blood to the RA from the upper body. The inferior vena cava returns blood to the RA from the lower body. The entrance of the inferior vena cava into the RA is protected by a rudimentary valve called the eustachian valve.4


The entrance from the coronary sinus into the RA is located between the AV orifice and the valve of the inferior vena cava. This opening is protected in part by a rudimentary valve of the coronary sinus called the thebesian valve.5 Other distinguishing structures in the RA include the interatrial septum and the fossa ovalis cordis, which is the remnant of the fetal foramen ovale within the septum.



Right Ventricle

The RV ejects blood into the pulmonary arterial system for oxygenation and removal of carbon dioxide by the lungs. The RV communicates with the RA through the AV orifice, which is separated by the tricuspid valve. The RV also communicates with the pulmonary outflow tract through the pulmonary orifice, which is guarded by the pulmonic valve (see Figure 23-4).


The walls of the RV are much thicker (4 to 5 mm) than those of the RA because of the increased pressures required to generate forward blood flow into the pulmonary circulation. The superior portion of the RV as it approaches the pulmonary orifice has a conical appearance and is called the conus arteriosus or infundibulum.6


The inner wall of the conus is smooth, but the remainder of the right ventricular wall has a rough appearance because of the presence of several irregular muscular bundles called the papillary muscles and the trabeculae carneae. One of the trabeculae carneae (the moderator band) crosses the cavity of the ventricles and carries the right branch of the AV bundle. The papillary muscles have attachments to the ventricular walls and to the chordae tendineae. The chordae tendineae are attached to the cusps of the tricuspid valve; together with the papillary muscles, they help prevent eversion of the tricuspid valve into the RA during ventricular systole.7



Left Atrium

The left atrium (LA) acts as a reservoir for oxygenated blood as it receives blood from the four pulmonary veins and also serves as a pump during atrial systole. It provides a 20% to 30% increase in left ventricular end-diastolic volume (LVEDV), which is known as the atrial kick. A person who has normal myocardial performance does not rely on this increase in ventricular filling to achieve adequate CO. However, in certain cardiovascular or respiratory pathologic conditions, compromised patients do rely on this atrial kick to maintain an adequate CO. The LA is located superiorly and posteriorly to the other cardiac chambers. The walls of the LA are slightly thicker (3 µm) than those of the RA. The LA connects to the LV through the left AV orifice, which contains the mitral valve. The atrial septum is smooth but may contain a central depression that corresponds to the location of the fossa ovalis cordis.



Left Ventricle

The apex of the LV is positioned within the mediastinum in an anterior and inferior orientation. The LV receives blood from the LA and ejects it into the aorta. Left ventricular wall thickness is approximately 8 to 15 mm, or two to three times the thickness of the RV. This additional muscle mass is required to overcome the systemic vascular resistance (SVR), or afterload, to maintain CO.


The ventricular septum separates the right and left ventricular cavities.8 The upper third of the septum is smooth endocardium. The remaining two thirds of the septum and the rest of the ventricular wall are covered with trabeculae carneae.


Two large papillary muscles are present within the LV. The anterior papillary muscle attaches to the anterior part of the left ventricular wall, and the posterior papillary muscle arises from the posterior aspect of the inferior wall. The chordae tendineae of each muscle are attached to the cusps of the mitral valve and prevent eversion of the valve during ventricular systole.



Myocardium


The cardiac musculature is arranged in three distinct layers: an outer epicardium, a middle muscular myocardium, and an inner endocardium. The epicardium is composed of mesothelium, connective tissue, and fat. The middle muscular myocardium consists of two muscle layers—a superficial and a deep layer. These layers are arranged in a spiral fashion and appear on cross section to run at right angles to each other. It has been postulated that the superficial and deep layers of the myocardium are not two separate layers but one tortuous and continuous layer. The arrangement of the muscle layers provides strength during contraction of the myocardium and efficient propulsion of blood toward the semilunar valves. The endocardium consists of endothelium and a layer of connective tissue.



Valves


The cardiac valves increase the heart’s efficiency by ensuring a one-way flow of blood through the circuit. They open and close in response to pressure gradients that exist above or below the valves. These valves may be categorized as AV or semilunar in configuration.


One of the most accurate ways to determine the presence of valvular pathology is by calculating valve area. The standard method for determining valve area is by cardiac catheterization. A cardiologist is able to determine valve gradients using the Gorlin formula or its correction, which can provide information regarding the degree of pathology that exists.9


image


Where valve area is expressed in cm2, blood flow across the mitral valve is expressed in mL/s, K is a hydraulic pressure constant, and mean transvalvular gradient is expressed in mmHg.


Echocardiography is a noninvasive method of determining valve area and is used in the diagnosis of valvular heart disease. The use of echocardiography is discussed later in this chapter.



Atrioventricular Valves




Semilunar Valves

The configuration of the aortic and pulmonary valves is similar. The cusps of the aortic valve are slightly thicker because it is subjected to greater pressures, which are created by left ventricular ejection. The semilunar valves are situated within the outflow tracts of their corresponding ventricles. Each valve is composed of three cusps. Above the aortic valve is a dilation known as the sinus of Valsalva, which allows the valve to open efficiently without occluding the coronary ostia or openings that communicate with the coronary arteries. Eddy currents form behind the valve leaflets and prevent contact between the valve leaflets and the walls of the aorta. Normal aortic valve area is 2.5 to 3.5 cm2. Reduction of the valve area by one third to one half is associated with an increase in the symptoms caused by aortic stenosis.



Coronary Circulation


The heart is an aerobic organ that depends on a constant supply of oxygen to meet its high metabolic demand. It requires an elaborate arterial and venous network to ensure that myocytes are adequately supplied with oxygen. The arterial system consists of epicardial and subendocardial vessels. The epicardial vessels are located superficially and most commonly become obstructed at areas of bifurcation where the blood flow is turbulent rather than laminar. Significant obstruction (50% to 70% reduction in luminal diameter) can result in myocardial ischemia or infarction as a result of increased resistance to flow across the stenotic areas.



Coronary Arteries

The coronary ostia are the entrance points by which blood flows through the coronary circulation and they are located behind the aortic cusps near the superior part of the sinus of Valsalva. The ostium of the left coronary artery is superior and posterior to the right coronary ostium. The coronary arteries act as end arteries, and each supplies blood to its respective capillary bed12 (Figure 23-5).







Right Coronary Artery: The right coronary artery supplies blood to the SA and AV nodes, the RA and RV, the posterior third of the interventricular septum, the posterior fascicle of the left bundle branch, and the interatrial septum. In approximately 90% of the population, the right coronary artery leaves the right coronary sinus and descends in the right AV groove. At the crux, the right coronary artery courses inferiorly in the posterior AV groove and terminates as a left ventricular branch.


The branches of the right coronary artery include the conus artery, the sinus node artery (50% to 60% of the population), several anterior right ventricular branches, the right atrial branches, the acute marginal branch, the AV node artery (90% of the population), the proximal bundle branches, the posterior descending artery, and the terminal branches to the LA and LV.




Venous Drainage

An extensive venous system exists in the heart. The three major systems include the coronary sinus, the anterior cardiac veins, and the thebesian veins (Figure 23-6).



The coronary sinus is located in the posterior AV groove near the crux. It collects approximately 85% of the blood from the LV, and for this reason it is catheterized when metabolic studies of the LV are performed. It may also be cannulated during cardiopulmonary bypass to deliver cardioplegia. The coronary sinus receives blood from the great, middle, and small cardiac veins; the posterior left ventricular veins; and the left atrial vein of Marshall.


Two to four anterior cardiac veins drain the anterior right ventricular wall. These veins may enter the RA directly, or they may empty into the coronary sinus.


The thebesian veins traverse the myocardium and drain into the various cardiac chambers, especially the RA, the RV, and to a lesser extent the LV. The thebesian veins may carry up to 40% of the blood that is returned to the RA.



Cardiac Innervation


The autonomic nervous system is divided into the sympathetic and parasympathetic nervous systems. Efferent impulses are transmitted from the brainstem and hypothalamus to numerous body systems, including the heart. The neurologic innervation to the heart originates from the autonomic nervous system, as well as from sensory fibers. The myocardium also has a specialized conduction system that is discussed in this chapter.


Increased sympathetic nervous system tone increases heart rate (chronotropic), force of myocardial contraction (inotropic), and rate of sinus node discharge (dromotropic). Sympathetic nervous system activation results in the mobilization of myocardial fat-free acids and glycogen for energy use by the myocardial cells. The preganglionic sympathetic nervous system fibers originate from the cells in the intermediolateral columns of the higher thoracic segments of the spinal cord and synapse at the first through the fourth or fifth thoracic paravertebral ganglia. These spinal cord segments are known as the cardioaccelerator fibers. The postganglionic fibers then travel as the superior, middle, and inferior cardiac nerves and the thoracic visceral nerves. These fibers form an epicardial plexus and are distributed over the entire ventricular myocardium. There is greater distribution of sympathetic nerves that innervate the ventricles, resulting in increased ventricular contractility, as is shown in Figure 23-7. Catecholamines released during sympathetic stimulation bind with adrenergic receptors on the heart (primarily B1) and change the biochemical properties within the myocyte. This process is discussed later in this chapter.



Some of these postganglionic sympathetic fibers also join with the postganglionic parasympathetic fibers from the cardiac plexus and primarily innervate the SA and AV nodes and the atrial myocardium. Suppression or blockade of this thoracic portion of the spinal cord by regional anesthesia causes bradycardia and hypotension as a result of the blockade of these sympathetic ganglia parasympathetic nervous system predominance.


The preganglionic parasympathetic fibers originate in the dorsal motor nucleus of the medulla. Short postganglionic fibers primarily innervate the SA and AV nodes and the atrial muscle fibers (see Figure 23-7). For this reason, increased parasympathetic tone decreases heart rate (HR). The function of the parasympathetic nervous system is primarily to slow the HR and secondarily to decrease contractility. In fact, maximal vagal (parasympathetic) stimulation reduces contractility by only 30%, whereas maximal sympathetic stimulation increases contractility by 100%. Acetylcholine is the neurotransmitter of the parasympathetic nervous system. Acetylcholine binds to muscarinic receptors on the heart and decreases the rate of sinus node discharge and slows conduction velocity through the AV node. The physiologic effects of parasympathetic nervous system stimulation occur because of increased permeability of cardiac muscle cell membranes to potassium, resulting in hyperpolarization. As a result, SA and AV node cells are less excitable.


Sensory innervation to the heart originates in the nerve endings in the walls of the heart, the coronary artery adventitia, and the pericardium. These nerve endings synapse with ascending fibers in the posterior gray columns of the spinal cord, where the fibers synapse with second-order neurons. From these neurons, the fibers ascend in the ventral spinothalamic tract and terminate in the posteroventral nucleus of the thalamus.



Cardiac Conduction System


Within the myocardium lies the specialized conduction system whose purpose is to automatically initiate and coordinate the cardiac rhythm. The cells of this system differ from the other myocardial cells because they are more variable in shape, contain fewer myofibrils, and have a characteristic pale staining of the cytoplasm. The conductive system consists of the following components: the sinoatrial (SA) node, the internodal tracts, the AV node, the AV bundle, and the Purkinje system (Figure 23-8).





Internodal Tracts

The internodal tracts are located within the atria and are the preferential conduction pathways between the SA and the AV nodes. They are composed of a combination of closely packed parallel myocardial fibers and large pale-staining cells with a perinuclear clear zone. They have large nuclei and sparse myofibrils that resemble the Purkinje cells. Like the SA node, the internodal tracts contain P cells and transitional cells.


Three major internodal tracts exist: the anterior, middle, and posterior internodal tracts. The anterior internodal tract, or Bachmann bundle, sends fibers to the LA and then travels down through the atrial septum to the AV node. The middle internodal tract, or Wenckebach tract, curves behind the superior vena cava before descending to the AV node. Finally, the posterior internodal tract, or Thorel tract, continues along the terminal crest to enter the atrial septum and then passes to the AV node.






Structural and Regulatory Proteins

The myocardium has characteristics of both skeletal and smooth muscle. Like smooth muscle, cardiac muscle fibers are interconnected (syncytial), which allows for an action potential to rapidly spread to adjacent cells. Because of this characteristic of cardiac muscle fibers, action potential propagation and muscle contraction occur as an “all-or-none” response.


The myocardial cell is similar to skeletal muscle in that it is composed of sarcomeres (Figure 23-9). These sarcomeres contain all the microfilaments and structures that are consistent with the skeletal muscle sarcomere. The sarcomere stretches from Z line to Z line. The A bands consist of the actin filaments, which contain a bilayer filament of F-actin and tropomyosin. Along the actin filament, many active sites exist that can attach to the head of the myosin molecule. A troponin complex is necessary to inhibit actin and myosin from interacting and initiating muscle contraction. The other microfilament in the cardiac muscle is the myosin molecule. This molecule is made up of two major parts: a light meromyosin chain and a heavy meromyosin chain. The heavy meromyosin chain consists of two hinged ends and a head that plays a role in the “ratchet theory” of muscle contraction (Figure 23-10).1




During sympathetic nervous system stimulation, catecholamines (primarily epinephrine and norepinephrine) are released from the central nervous system and the adrenal medulla. Increased cardiac conduction velocity, increased force of contraction, and increased heart rate are primarily mediated by beta11)-adrenergic receptors. When catecholamine hormones interact with β1 receptors, they stimulate G protein activation. Adenyl cyclase activity increases and catalyzes the formation of cyclic adenosine monophosphate (cAMP). A specific protein kinase is formed, and phosphorylation occurs, increasing myocardial cell permeability to calcium and sodium. Threshold potential is reached, and depolarization occurs, which increases the concentration of calcium from the sarcoplasmic reticulum and the transverse tubular system. Calcium interacts with the troponin tropomyosin complex to initiate cardiac contraction. The force of myocardial contraction is dependent on the quantity of calcium present in the cardiac cell.


Evidence indicates that the troponin-tropomyosin complex inhibits the binding of the heads of the myosin filaments with the active sites on the actin molecule (Figure 23-11). During the initiation of contraction, calcium is released from the sarcoplasmic reticulum. Calcium binds to the troponin-tropomyosin complex and causes a conformational change so that the active binding sites on the actin filaments become exposed. The myosin cross bridges bind to the active filament and move along the actin filament by alternately attaching and detaching from the active sites, thereby causing shortening of the Z lines (Figure 23-12). This is known as the sliding filament theory. When the actin filaments and myosin cross bridges intermingle, muscle contraction occurs. Inhibition of calcium influx into the cardiac muscle cells is the proposed mechanism whereby the inhaled anesthetic agents cause depression of myocardial contractility. Cellular energy or adenosine triphosphate (ATP) is required for this process, known as excitation-contraction coupling, to occur, as shown (Figure 23-13). For muscle contraction to cease, calcium reuptake into the sarcoplasmic reticulum occurs as a result of active transport. The troponin-tropomyosin complex reinhibits the interaction between actin and myosin. That this process occurs constantly and within milliseconds is the primary reason for the heart’s high metabolic demands (see Figure 23-13). With a limited supply of oxygen and substrate (primarily fatty acids), such as with patients with coronary artery disease or an increased energy demand caused by tachycardia from sympathetic nervous system stimulation, myocardial dysfunction and infarction can occur.





Similar length-force relationships exist within the myocardium and in skeletal muscle. The resting sarcomere length at which the muscle cell is most efficient is 2.0 to 2.4 µm. At greater lengths, the interdigitation of the actin and myosin is compromised, and at shorter lengths the sarcomere is unable to generate an efficient contraction.


Clinically this concept is demonstrated by having the ideal filling pressure of the LV necessary to achieve adequate CO. Filling pressures are used to reflect the filling volumes of the ventricles and (indirectly) the amount of stretch on the ventricular muscle at rest. Filling pressures are measured by the use of the pulmonary capillary wedge pressure (PCWP) or the pulmonary artery diastolic pressure. It has been demonstrated that at excessively high filling pressures (as in congestive heart failure) and at excessively low filling pressures (as in hypovolemia), the CO can be compromised as a result of either excessive or inadequate stretch of the left ventricular myocardium. The greater the degree of stretch of myocardial muscle fibers, the greater the number of actin filaments and myosin cross bridges that are more completely approximated. This will result in an increase in the force of cardiac contraction, as shown in Figure 23-14. This concept is the basis for the Frank-Starling law of the heart, which is discussed later in this chapter.




Differences Between Skeletal and Cardiac Muscle Cells

Several differences exist between myocardial muscle cells and skeletal muscle cells (see Figure 23-9). At the junctions between the fibers in the myocardial muscle mass, many branching, interconnected fibers are intercalated disks and gap junctions, or nexi. Areas of low resistance facilitate the conduction of the action potential from one myocardial cell to another.


The myocardial sarcomeres also contain higher concentrations of mitochondria than do other types of muscle cells. The cardiac cells are aerobic and cannot tolerate oxygen deficiency. Skeletal muscles can function both aerobically and anaerobically.


The myocardial sarcomere system has a rich capillary blood supply (one capillary per fiber) that allows for efficient diffusion and perfusion. The T-tubular system and the sarcoplasmic reticulum are extensive within the cardiac sarcomere. This situation allows for the rapid release and reabsorption of calcium from the cells. It also serves to highlight the important role extracellular calcium plays in the contractile process of the myocardial cell.



Generation of Membrane Potentials



Resting Membrane Potentials

The myocardial sarcomere is not merely a contractile entity. It also possesses properties common to neural tissue, such as the generation of a resting membrane potential, the ability to generate an action potential, and the conduction of the action potential from one sarcomere to the next.


The resting cell membrane is relatively permeable to potassium and relatively impermeable to both sodium and calcium. The resting membrane potential is caused by a chemical force, an electrostatic counterforce, and the sodium-potassium active transport pump.


The chemical force relies on the potential difference in ion concentration between one side of the cell membrane and the other. The ions primarily responsible for this force are sodium, potassium, and calcium.13 The electrostatic counterforce results from the negative potential generated by the ion difference of the interior of the cell. This force can pull ions into the cell, especially potassium.


The sodium-potassium pump requires an energy source (active transport) and involves the magnesium-dependent enzyme adenosine triphosphatase located in the cell membrane. Three molecules of sodium are pumped out of the cell into the extracellular fluid for every two molecules of potassium pumped into the intracellular fluid.


Calculation of the equilibrium potential (Em, measured in millivolts) has been accomplished by examining the concentration of an ion inside the cell versus outside the cell (see the Nernst equation that follows). Table 23-1 lists the equilibrium potentials of the most physiologically important ions. The ion most responsible for the resting membrane potential is potassium.


image


where R is a gas constant, T is temperature in Kelvin, F is Faraday’s constant, [K]i is the intracellular concentration of potassium, and [K]o is the extracellular concentration of potassium.



If a temperature of 310° K is assumed for a living human, the Nernst equation reduces to the following in a human heart:


image


The Nernst potential is useful only in discussions of a single ion. The membrane potentials are generated because the cell membrane is permeable to several different ions. Three factors affect the calculation of the effect of these different ions on the resting membrane potential: the electric charge of each ion, the permeability of the membrane to each ion, and the concentration gradient across the membrane. The following equation, the Goldman-Hodgkin-Katz equation, is a modification of the Nernst equation that accounts for these factors.


image


where [K]i is the intracellular ion concentration of potassium, [K]o is the extracellular ion concentration of potassium, PNa is the membrane permeability of sodium, PCl is the membrane permeability of chlorine, and PK (when calculated) is the membrane permeability of potassium.



Ventricular Muscle Fiber Action Potential




Phases of the Action Potential

The action potentials of the various parts of the conduction system vary according to their locations and functions.14 The action potential of the ventricular muscle fiber is separated into five phases (Figure 23-15). Phase 0, or upstroke, is represented by depolarization and involves the fast sodium channels. The fast sodium channel activation gates (M gates) open between −70 and −65 mV (threshold potential). At 0 mV, both the activation and the inactivation gates (H gates) are open. The rapid upstroke velocity of phase 0 gives a relative indication of the conductivity of the myocardial cell. Local anesthetics such as lidocaine have an inhibitory effect on phase 0 by decreasing the influx of sodium. Table 23-2 describes the phases and events of a cardiac action potential.




At phase 1, or early rapid repolarization (+2 to +30 mV), the sodium gates close, the rapid influx of sodium stops, and the slower influx of calcium begins. Also, potassium gates open and potassium moves out of the interior of the cell into the extracellular fluid.


One of the characteristics of the action potential unique to ventricular muscle is phase 2, or the plateau phase. The plateau phase exists because the slow calcium channels open at −30 to −40 mV and allow an influx of calcium. This inward calcium flux delays repolarization and prolongs the absolute refractory period. Toward the end of phase 2, a decreased permeability to potassium occurs that accounts for a small outward leakage of potassium balanced by the calcium and sodium influx that maintains a membrane potential near 0 mV. Calcium channel blockers exert their pharmacologic effect during phase 2. The physiologic effects include decreased contractility, decreased heart rate, and decreased cardiac conduction velocity.15


The terminal repolarization phase, phase 3, is initiated as the slow calcium channels become inactivated, and this phenomenon is sustained by an accelerated potassium efflux. These events return the transmembrane potential to its resting membrane value.


The sodium-potassium pump, which is dependent of ATP, reestablishes the proper intracellular-to-extracellular ionic concentrations during phase 4 (diastolic repolarization phase). Phase 4 lasts from the completion of repolarization to the next action potential. Lidocaine lengthens the duration of phase 4 by decreasing the cardiac cell membrane’s permeability to potassium ion, thereby decreasing the efflux of potassium and delaying the onset of the resting membrane potential.15


The cardiac glycoside digoxin inhibits the sodium-potassium ATP-dependent pump, decreasing sodium efflux into the extracellular fluid. As intracellular sodium concentrations increase, the exchange between sodium and calcium is decreased. The result is a higher concentration of calcium remains within the cardiac cell, and this effect is believed to be responsible for the increased inotropic effect.15



Refractory Periods

The extended duration of the action potential of the myocardial cell protects it against premature excitation. This period of quiescence is known as the refractory period and can be divided into absolute and relative periods. The refractory periods are a result of the properties of the sodium channels during the action potential.


The term effective or absolute refractory period is used to describe the time during which a conducted action potential may not be evoked, even if an active response is elicited by a stimulus at the cellular level. This period lasts from phase 0 to the middle of phase 3, when the membrane potential drops below −60 mV. The relative refractory period is the time during the action potential when a second stimulus can result only in an action potential with decreased amplitude, upstroke velocity, and conduction velocity. The relative refractory period extends from this middle part of the phase 3 range to the beginning of phase 4, when the membrane potential ranges from −60 to −90 mV. This information can be clinically related to synchronized cardioversion. The shock will not be delivered during the T wave on the electrocardiogram, which represents ventricular repolarization. The relative refractory period occurs during the T wave, and electricity delivered to the chest during this time can cause electrical disorganization in cardiac cells, resulting in ventricular tachycardia or ventricular fibrillation.



Sinoatrial Node Action Potential

The myocardium has among its characteristics contractility, automaticity, and conductivity. Each of the various myocardial masses has its own intrinsic automaticity and rate of action-potential initiation. The SA node is the primary pacemaker of the heart and has several unique characteristics (Figure 23-16). As a result of its higher resting membrane potentials, the SA node membrane is more permeable to sodium than other atrial myocardial cells. This “leakiness” gradually raises the membrane potential closer to threshold potential (−55 to −60 mV), at which point an action potential may be initiated. Therefore the action potential originating within the SA node differs from the action potential generated within the ventricular muscle mass. For this reason, the SA node is the primary pacemaker of the heart.



The SA node and the other automatic cells exhibit only phase 4, phase 0, and phase 3 of the action potential. Because rapid depolarization does not occur, phase 1 or phase 2 (plateau phase) does not occur.


If the SA node fails, the area of the heart with the next highest intrinsic rate, the AV node, replaces the SA node as the pacemaker of the heart. The intrinsic firing rate of the AV node is 40 to 60 beats per minute. If both the SA and the AV nodes fail, the ventricular cells take over and become automatic, firing at a rate of 15 to 30 beats per minute.



Physiology of the Heart



Cardiac Cycle

To understand the cardiac cycle, one must have a firm understanding of the basics of the anatomy of the heart and the pressures and volumes generated within the various chambers during the cardiac cycle. An appreciation for the valves and their positions during the phases of the cycle is essential (Figure 23-17). Additionally, notice that the electrocardiogram impulse generation precedes the mechanical action of the heart. This delay between the electrical impulse and the mechanical event occurs because time is needed for the wave of depolarization to spread across the myocardium before contraction can begin. In relation to the electrocardiogram, the P wave represents atrial systole, the QRS complex signifies ventricular systole, and the T wave represents ventricular repolarization.



The cardiac cycle extends from one ventricular contraction to the next. It may be divided into two main phases—systole and diastole. Usually the cardiac cycle is described in relation to the left side of the heart. However, similar conclusions about the function of the right side of the heart may be drawn in the absence of cardiopulmonary pathology.



Diastole: During ventricular systole, the atria fill and blood returns from the venous system to the right side of the heart and from the pulmonary circulation to the left side of the heart. The first phase of diastole is the period of isovolumetric relaxation. The ventricular muscle mass relaxes, and the aortic and mitral valves are closed as long as the ventricular pressure remains higher than the atrial pressure. The true filling phase is divided into three periods: (1) rapid inflow, or diastasis, (2) reduced inflow, and (3) atrial systole. Once the ventricular pressure drops below atrial pressure, the mitral valve opens, and the period of rapid inflow to passively fill the ventricle begins. The second period of diastole is diastasis, in which minimal changes occur in volume and in pressure.


Atrial systole provides another period of rapid filling that is commonly referred to as the atrial kick. This phenomenon increases ventricular filling by 20% to 30%. In patients who have severe mitral stenosis, the atrial kick may be responsible for up to 40% of the ventricular filling. During periods of strenuous exercise or in patients with many pathologic conditions such as shock or congestive heart failure, the additional ventricular filling is critical to maintaining CO.



Systole: After atrial systole, the isovolumetric phase, or isovolumetric contraction, which is the phase at the beginning of ventricular contraction, occurs. The myocardial fibers shorten, and pressure is generated within the ventricle but only enough to close the mitral valve. Therefore during this period, an increase in left ventricular pressure occurs without a change in ventricular diastolic volume. Isovolumetric contraction begins with closure of the mitral valve and lasts until opening of the mitral valve.


Systolic ejection begins with the opening of the aortic valve and occurs when the ventricular pressure exceeds the aortic pressure. This phase of the cardiac cycle is divided into two periods, with the period of rapid ejection taking the first third of systole and the period of reduced ejection taking the last two thirds of systole. During rapid ejection, ventricular systolic pressures reach their maximum, and the largest amount of volume is ejected. Therefore systole is composed of isovolumetric contraction, rapid ejection, and reduced ejection. The dicrotic notch or incisura on the arterial pressure tracing occurs within the period of isovolumetric relaxation. This segment represents retrograde blood flow back into the LV before aortic valve closure. Three waveform segments are present on the left atrial pressure tracing: the a wave, c wave, and v wave. The specific waveforms correspond to their position within the cardiac cycle. The a wave represents atrial systole as it ends just before mitral valve closure. The c wave represents ventricular contraction and is produced by bulging of the mitral valve caused by increasing left ventricular pressure. The v wave represents increased pressure in the LA caused by blood return from the pulmonary artery before mitral valve opening.



Physiology of Coronary Circulation

The anatomy of the coronary circulation has already been discussed. A description of the physiologic determinants of coronary blood flow follows.



Coronary Blood Flow: The rate of blood flow is determined by a change in the pressure within the vessel divided by resistance of the system. Alterations of the radius of a vessel change the flow to the fourth power of the radius. This phenomenon is an extension of Poiseuille’s law, which determines the flow of a fluid through a tube.


At rest, approximately 4% to 5% of the CO, or 225 mL/min of blood, passes through the coronary vasculature. Phasic changes have been documented during coronary blood flow. A greater amount of coronary flow in the LV occurs during diastole. During systole, left coronary artery blood flow ceases to the subendocardium due to compression of the subendocardial vessels by the myocardium; flow through the epicardial vessels is not affected during systole to this extent. The flow to the left coronary artery is greatest during diastole as a result of the decreased resistance to flow from decreased myofibril tension that occurs as the intracavitary pressure decreases (Figure 23-18).




Control of Coronary Circulation and Oxygen Supply and Demand: Coronary blood flow is regulated by intrinsic and extrinsic factors that affect coronary artery tone. Intrinsic factors include the anatomic arrangement and perfusion pressure of the coronary vessels. Extrinsic factors include compressive factors within the myocardium, as well as metabolic, neural, and humoral factors. Blood flow through the coronary circulation is primarily controlled by the factors that determine oxygen demand and oxygen supply. Myocardial oxygen supply is determined by arterial blood content, diastolic blood pressure, diastolic time as determined by HR, oxygen extraction, and coronary blood flow. Myocardial oxygen demand is determined by preload, afterload, contractility, and HR (Figure 23-19).



The factors that increase myocardial oxygen consumption (mimageo2) are listed in Table 23-3. Notice in Figure 23-19 that heart rate appears on both the supply (diastolic time) side and demand side of the balance. Increasing heart rate not only increases demand but also decreases diastolic time, which is when 80% to 90% of coronary filling and myocardial perfusion occurs. Increased heart rate is the most important factor that negatively affects mimageo2. Doubling the heart rate doubles mimageo2.16 This phenomenon most dramatically affects patients with coronary artery disease, because the supply of blood is compromised and may not be able to meet the oxygen demands caused by tachycardia. As a result, myocardial dysfunction or infarction can occur. By slowing HR and decreasing contractility, β-blocking medications increase supply and decrease demand, protecting the heart from ischemia.17



Because at rest the myocardium extracts 65% to 70% of the available oxygen, the only way to increase oxygen delivery to the myocardium is by increasing blood flow. During periods of increased myocardial oxygen demands, flow through the coronary arteries can increase by three to four times. Several vasodilator substances have been identified and are released from the myocardium in response to decreased oxygen delivery or concentration. Among these substances are adenosine, adenosine phosphate compounds, potassium ions, hydrogen ions, carbon dioxide, bradykinin, and prostaglandin. Experts believe that adenosine is the primary substance responsible for coronary vasodilation.


The normal physiologic parameters of the heart are given in Table 23-4. The determinants of mimageo2 include myocardial contractility, myocardial wall tension (preload), HR, and mean arterial pressure (MAP; afterload). Oxygen extraction is determined by measurement of the difference between the oxygen tension in the pulmonary arterial blood and that in the coronary sinus.



Oxygen supply relies on the blood oxygen content (see following equation), which is affected by both the oxygen carried on the hemoglobin (Hgb) molecule (1.34 mL of oxygen per gram of Hgb) and to a lesser extent the oxygen dissolved in the plasma (0.003 mL of oxygen per milliliter of plasma).


image


Other factors that have an influence on coronary circulation include the direct and indirect effects of the sympathetic nervous system and the effect of certain substrates of cardiac metabolism.



Autoregulation: Under normal physiologic conditions, the coronary circulation, like other tissue beds in the body, exhibits autoregulation, which is the ability to maintain coronary blood flow through a range of MAPs by dilating or constricting. Coronary blood flow is maintained at a constant rate through a MAP range of 60 to 140 mmHg. When arterial blood pressure is less than or exceeds these pressure limits, coronary blood flow becomes pressure dependent. Therefore during hypotension, when the coronary arteries are maximally dilated, coronary blood flow is determined by the MAP minus the right atrial pressure.


A method for directly estimating coronary perfusion pressure (CPP) can be calculated by subtracting left ventricular end-diastolic pressure (LVEDP) from diastolic blood pressure (DBP)—that is, CPP = DBP − LVEDP. Because under normal conditions LVEDP (10 mmHg) is significantly less than DBP (80 mmHg), the major determinant of CPP is DBP.


Coronary vascular reserve is the difference between the maximal flow and the autoregulated flow. The closer these two values, the lower the coronary reserve of the patient. Factors that increase myocardial oxygen demand and limit supply decrease coronary reserve flow and can result in myocardial dysfunction.


The concept of “coronary steal” has emerged, especially in reference to the use of agents such as adenosine, nitroglycerin, and isoflurane. If vasodilator treatment is used in a patient who has both an ischemic area of the heart that is supplying a stenotic vessel with collateral flow and another area that has an intact autoregulated vessel, only the autoregulated vessel dilates further and has the ability to increase its flow. Constant maximal coronary artery dilation exists in an area where stenosis is present. Therefore only the areas of the heart with intact autoregulation respond to vasodilators and receive preferential flow over the stenotic area. The existence of this phenomenon is questionable. As long as adequate CPP is maintained, coronary steal and myocardial ischemia caused by isoflurane do not occur.18,19 A second factor that could result in this phenomenon is coronary steal–prone anatomy. This has been defined as complete occlusion of one coronary artery and at least 50% occlusion of a second coronary artery that supplies collateral blood flow to the area in which the complete occlusion exists.18 In addition, recent evidence suggests that the inhaled anesthetics isoflurane, desflurane, and sevoflurane produce myocardial protection during periods of ischemia in humans by decreasing the formation of free radicals, preserving myocardial ATP stores, and inhibiting increased intracellular calcium.19,20 This is referred to as anesthetic preconditioning.



Cardiac Output: Cardiac output is the amount of blood ejected from the LV during 1 minute. Comparing various CO values among several patients requires a method for calculating output in relation to the size of the patient. The CO is measured in liters per minute. Cardiac output is indexed, because a CO of 3.5 L/min may be adequate for a patient who is 5 feet tall and weighs 95 lb, but it is less than optimal for a patient who is 6 feet 7 inches tall and weighs 300 lb. The average CO is 5 L/min, and the average cardiac index (CI) is 2.5 L/min or more per square meter of body surface area (BSA). The formula for this relationship is CI = CO/BSA.


The primary determinants of CO are stroke volume (SV) and HR. CO is derived by using the equation CO = HR × SV. The SV is the amount of blood ejected from the LV with each beat. The average SV is approximately 70 mL. If the average HR is 70 to 80 beats per minute, a CO of 5 L/min results.


Several key factors affect SV, including preload, afterload, and myocardial contractility. Preload is the effective tension of the blood on the ventricle or the wall tension at the end of diastole. Preload can either be passive (the flow of blood from the atria to the ventricles during diastole) or active (the volume contributed by the atrial kick). With increased preload, there is an associated increase in contractility. This phenomenon is known as the Frank-Starling law of the heart, which states that the greater the wall tension (preload), the greater the compensatory increase in myocardial contractility. This mechanism allows the heart to immediately compensate for increased preload and avoid overdistention of the cardiac chambers by increasing SV, which facilitates chamber emptying. However, there is a point at which progressive increases in preload no longer increase contractility but can contribute to decreased myocardial performance. Increased preload increases myocardial oxygen demand. Clinically, preload can be estimated by using the PCWP and the pulmonary artery diastolic pressure. In patients with normal mitral valve and ventricular muscle function, either of these measures provides an estimate of the preload or LVEDP and volume.


Cardiac output can be determined indirectly by applying the Fick principle (see the following equation). Assuming normal respiratory function, CO is equal to the amount of oxygen absorbed by the lungs divided by the arteriovenous oxygen difference.


image


Right-sided heart pressures (central venous pressure) that are obtained clinically can be estimates of left ventricular volumes in patients with good left ventricular function.21


Afterload is the wall tension the myocardium needs to overcome to eject the SV. It is the pressure within the LV during peak systole. Factors affecting LV afterload include the state of the ventricular chamber and the compliance of the arterial vasculature. The shape, size, and wall thickness of the ventricle play an important role in afterload. The vascular component of the afterload includes SVR and MAP because these variables relate to the vascular compliance of the aorta. Thus increases in blood pressure increase afterload.


Afterload is most often estimated clinically by determining the SVR. The SVR may be calculated once the CO and the difference between the MAP and the central venous pressure (CVP) are known (see the following equation). The normal SVR is 800 to 1500 dyn.s/cm5


image


The problem with equating afterload with SVR is that ventricular wall tension, which is an integral part of the afterload, is not considered.


Contractility of the myocardium is the state of inotropy that is independent of either preload or afterload. It may be altered by many cardiovascular disease states. Factors such as rate of pressure changes over time (dP/dt [first derivative of pressure measured over time]), force-velocity or Starling ventricular function curves, pressure-volume loops, ejection fraction (EF), and velocity of circumferential fiber shortening have all been used to estimate contractility.


Left ventricular dP/dt measurements require a high-fidelity recording system, and for this reason these measures are not readily available in the clinical setting. A wide range of normal values exists (800 to 1700 mmHg/sec), making patient-to-patient comparisons difficult. Assessment of the acute change in contractility of a single patient over time is still the common method of using this measurement.


Ventricular function curves22 (Figure 23-20), define the relationship between the left ventricular filling pressure (left ventricular diastolic pressure, left atrial pressure, PCWP) and the left ventricular stroke work index (LVSWI), which is calculated by use of the following equation:



image


where SVI = CI/HR.


Each left ventricular function curve has a steep upstroke that has a plateau at higher filling pressures. To apply the Frank-Starling mechanism to Figure 23-20, notice in the “normal” and “hyperdynamic” curves that as pressure (horizontal axis) increases, so does LV output (vertical axis). However, at the top of these curves, there is a plateau where increasing the filling pressures no longer increases performance and can then decrease ventricular output. Symptoms may be elicited by either high or low filling pressures. On the “failure” curve, with compromised cardiac function, increases in filling pressures do not dramatically increase myocardial performance and can lead to cardiogenic shock. The clinical determination of LVSWI is worthwhile because it contains many of the factors that contribute to CO, and it gives measures of both systolic and diastolic performance.


Left ventricular pressure-volume loops have been mentioned before as conceptual models depicting phases of the cardiac cycle. They may also be used as tools to determine myocardial performance. Left ventricular pressure-volume loops simultaneously measure chamber pressures and the resultant volumes (Figure 23-21). Movement from left to right on the horizontal axis represents increased volume. Movement from right to left on the horizontal axis represents decreased volume. Movement up and down on the vertical axis represents increases and decreases in pressure, respectively. The distinct phases of the left ventricular pressure-volume loop are represented in Figure 23-22.




The interior of the curve (distance between the two vertical lines of the LV pressure-volume loop) is representative of SV. In this diagram, SV is calculated by subtracting end-systolic volume (ESV) from end-diastolic volume (EDV), or EDV (110 mL) − ESV (40 mL). Thus in this example, stroke volume is 70 mL. Ejection fraction (EF) can then be estimated using the following equations:


image


or


image


The EF is the percentage of the end-diastolic volume ejected during systole, as seen in Figure 23-23. The normal EF is 60% to 70%. An EF of less than 40% is associated with significant left ventricular impairment.



Deviations from the normal left ventricular pressure-volume loops occur as a result of many causes. Factors that alter the normal loop include increases and decreases in LV preload, LV afterload, and LV contractility. These factors can be acute and transient, such as during the administration of vasoactive medications, or chronic as a result of myocardial compensation caused by valvular heart disease. In Figure 23-24, these changes are consistent with valvular heart disease. Because many factors contribute to variations in pressure and volume from beat to beat within the LV, each pressure volume loop is distinct, represents one LV systolic ejection, and is different for each contraction. A discussion of these pathologic curves is presented later in this chapter.



A clinically useful tool is two-dimensional transesophageal echocardiography (TEE). When this technology is appropriately used, real-time movement of all four chambers of the heart, as well as that of the valves, may be visualized.


TEE can be used to detect valvular function and blood flows in both regurgitant and stenotic lesions. It is also useful in determining areas of hypokinesis, dyskinesis, or akinesis caused by myocardial infarction, ischemia, or injury. It can be useful in determining myocardial contractility, and it is a more direct measure of intraventricular volume status than pulmonary artery catheter pressure measures. TEE has also proven useful in directing fluid and pharmacologic therapy in patients who have undergone coronary bypass and other surgical procedures. Ventricular dysfunction or reperfusion injury, as well as the presence of intraventricular air, can be determined with TEE. This diagnostic tool is the gold standard for assessing intraoperative myocardial performance. The practical problems associated with the clinical use of TEE entail acquiring the skills necessary for accurate interpretation of the visual data, the possibility of esophageal rupture, and the cost associated with its use.



Cardiovascular Reflexes




Valsalva Maneuver

The Valsalva maneuver occurs as a result of forced expiration against a closed glottis. The reflex is mediated through the baroreceptors located near the bifurcation of the internal and external carotid arteries (carotid sinus) and the aortic arch. The afferent pathway is directed via Hering’s nerve and either the glossopharyngeal nerve (carotid sinus) or the vagus nerve (aortic arch), as shown in Figure 23-25. Stimulation of either of these areas inhibits the vasomotor center in the medulla. The response inhibits the sympathetic nervous system and stimulates the parasympathetic nervous system, producing a decrease in HR, a decrease in myocardial contractility, and vasodilation, resulting in a decrease in blood pressure. The Valsalva maneuver also increases intrathoracic pressure, which decreases venous return and thereby decreases CO.





Oculocardiac Reflex

Traction on the extraocular muscles (especially the medial rectus), conjunctiva, or orbital structures causes hypotension and a reflex slowing of the HR, as well as arrhythmias. The oculocardiac reflex may also be elicited during retrobulbar block, ocular trauma, or pressure on the tissue that remains after enucleation. The afferent path of the reflex is mediated by the long and short ciliary nerves to the ciliary ganglion of the oculomotor nerve and then the ophthalmic division of the trigeminal nerve (cranial nerve V) to the gasserian ganglion. The efferent branch of the reflex is mediated by the vagus nerve (cranial nerve X). This reflex may be blunted by the use of retrobulbar block or the release of the offending stimulus. The resulting vagal response to the heart can be inhibited by an anticholinergic agent (atropine or glycopyrrolate).



May 31, 2016 | Posted by in ANESTHESIA | Comments Off on Cardiovascular Anatomy, Physiology, Pathophysiology, and Anesthesia Management

Full access? Get Clinical Tree

Get Clinical Tree app for offline access