William W. Muir Gillespie College of Veterinary Medicine (LMU‐CVM) at Lincoln Memorial University, Harrogate, Tennessee, USA The cardiovascular system is a closed tubular arrangement that consists of the heart, blood vessels, and blood. The purpose of the cardiovascular system is to circulate blood and other essential materials, especially oxygen, remove waste products, and maintain homeostatic control of the body’s temperature, pH, and cellular environment. The purpose of the heart is to pump blood in quantities sufficient to meet the body’s oxygen demands. The purpose of the vasculature is to distribute blood to tissues, maintain adequate parenchymal perfusion, and facilitate the cellular exchange of nutrients for waste products. Blood transports oxygen, carbon dioxide, nutrients, waste, and hormones; regulates pH, temperature, and fluid balance; and protects the body from blood loss and foreign invaders via clotting mechanisms and the immune response, respectively. Schematically, the cardiovascular system of mammals is a circuit comprised of two circulations (i.e., systemic and pulmonary) in series (Fig. 33.1). Oxygen is essential for maintaining normal cellular metabolic function in all tissues. Relatively brief periods (2–3 min) of cerebral or cardiac anoxia, for example, can produce devastating consequences that threaten life even if blood flow is restored. Oxygenated blood returning from the lungs enters the left atrium and then the left ventricle and is pumped into the aorta and coronary circulation. The aorta, in conjunction with an elaborate array of large arteries, distributes blood throughout the systemic circulation to the tissue beds of each body organ (e.g., brain, heart, kidney, skeletal muscle, etc.). Large arteries transition to smaller arteries and eventually into arterioles (i.e., resistance vessels) that distribute and regulate blood flow and oxygen delivery to capillary beds (i.e., exchange vessels) based upon their oxygen demand. Capillaries provide a surface area that is over three orders of magnitude larger than the aorta and are the primary sites for tissue oxygen, carbon dioxide, and nutrient exchange. Capillary beds coalesce to form venules that return deoxygenated blood, carbon dioxide, and metabolic waste products to small and large veins that transition into the venae cavae. The venae cavae return deoxygenated blood to the right atrium and subsequently the right ventricle and pulmonary artery. Pulmonary capillaries exchange carbon dioxide for oxygen and return oxygenated blood to the left atrium thus completing the circuit. The body’s circulatory system also includes a network of lymph vessels, nodes, and organs (e.g., thymus and tonsils) in parallel with the vascular system that actively maintain interstitial fluid homeostasis by regulating filtered plasma (i.e., lymph) accumulation, preventing infection (i.e., immune response), and removing or destroying bacteria and toxins (Fig. 33.1). Figure 33.1 The mammalian circulatory systems consist of the heart, two circulations, pulmonary and systemic, and the lymphatic system. Blood returning to the right atrium (RA) from the systemic veins (70% blood volume) enters the right ventricle (RV) and is pumped into the pulmonary artery and oxygenated in the capillaries of the lung (15% blood volume). Oxygenated blood returns to the left atrium (LA) enters the left ventricle (LV), is pumped into the aorta and arterial conduits (10% blood volume), and distributed to capillaries (5% blood volume). Deoxygenated blood (blue); oxygenated blood (red); lymph (green). The lymphatic system regulates tissue fluid homeostasis by transporting interstitial fluid (i.e., lymph) to the venous system via the thoracic duct. The circulation of fluid (i.e., blood and lymph) throughout the body depends on a normally functioning heart, blood vessels, blood volume, and lymphatic system. Since blood flow is responsible for the uptake, delivery, and elimination of all drugs including anesthetic drugs, a functional appreciation of cardiovascular anatomy and circulatory dynamics are required for safe anesthetic practice. This chapter reviews the anatomy and physiology of the cardiovascular and lymphatic systems of mammals and summarizes the general effects of anesthesia and anesthetic drugs on their function. All vertebrate animals have a heart and circulatory systems (i.e., pulmonary, systemic, and lymphatic) although this can vary substantially among genera. The heart exhibits intrinsic electrical and mechanical properties that ensure the delivery of blood to closed (i.e., blood is contained within blood vessels) systemic and pulmonary circulations (Table 33.1). The lymphatic circulatory system is an open network of tissues, vessels, and organs that moves filtered interstitial fluid (i.e., lymph) back into the bloodstream. The American Veterinary Medical Association lists fish, ferrets, rabbits, hamsters, birds, gerbils, rodents, frogs, turtles, snakes, and lizards among others as exotic species. Several of these species have unique circulatory systems and not all have a four‐chambered heart (Fig. 33.2) [1,2]. Current opinions regarding circulatory system architectures, “open” or “closed” configurations, and different species’ unique circulatory characteristics are beyond the scope of this chapter; however, the opinion that the heart and circulation of amphibians and reptiles (i.e., three chambers with cardiac and vascular shunts) are functionally inefficient intermediate steps to the four‐chambered heart of birds and mammals is considered obsolete [1–3]. It is more likely that cardiac and vascular shunts represent an adaptive phenotypic trait providing various genera with the ability to regulate blood flow depending upon cellular respiration and environmental requirements. Table 33.1 Intrinsic properties of the heart. * tropy = change. Figure 33.2 The circulatory systems of vertebrates. Birds and mammals have a closed vascular system and complete ventricular septum. The hearts of turtles, lizards, and snakes consist of two atrial chambers and two ventricular chambers separated by an incomplete ventricular septum. The heart of crocodilians is completely divided into four chambers but retains a dual aortic arch system. The foramen of Panizza located near the RV and LV outflow tracts allows blood to bypass the lungs when submerged. Note that the ventricular septum is absent or incomplete in fish, amphibians, turtles, lizards, and snakes. Source: Adapted from Hicks [2]. The fish heart consists of four parts that includes two primary chambers, an entrance, and an exit, and its cardiovascular system is considered to consist of a single‐cycle (i.e., one “circulation”) closed‐loop circulatory system (Fig. 33.2). Amphibians have a three‐chambered heart (i.e., two atria and one ventricle) and are considered to have a “double” circulatory system (arterial and venous). The ventricle is incompletely separated by a partial septum into two pumps that results in less mixing of oxygenated and deoxygenated blood. Blood returning to the heart is pumped into two circulations: one carries blood to the lungs and skin for oxygenation, and the other delivers oxygenated blood to the rest of the body (Fig. 33.2). The systemic circulation of reptiles, like other vertebrates, consists of arterial, venous, and lymphatic vessels. The hearts of many reptiles including turtles, snakes, and lizards consist of two atria and a single ventricle that delivers blood to two circulatory routes; however, blood is only oxygenated in the lungs. The single ventricle has an incomplete muscular ridge that acts to divide the ventricle into two major chambers minimizing the mixing of oxygenated and deoxygenated blood (Fig. 33.2) [2,4]. The pulmonary artery and right and left aortic arches originate from specific anatomical sites (i.e., cavum pulmonale and cavum venosum) [5]. The sinus venosus is occasionally considered to be an additional chamber classifying the non‐crocodilian heart as an atypical “four‐chambered” organ [5]. The pulmonary artery is equipped with a muscular sphincter that when contracted diverts blood flow through the incomplete ventricular septum into the left ventricle and out the aorta producing a right‐to‐left (R‐L) shunt estimated to equal 60–70% of the venous return. Both pulmonary and systemic arterial vascular resistance and the degree of contraction of the muscular ridge control the magnitude and direction of cardiac shunting in reptiles [2]. For example, deoxygenated blood mixes with oxygenated blood during diastole but not during systole when the contraction process causes the ventricle to function as a dual chamber pressure pump. Factors that control the direction and magnitude of R‐L shunting during diving and breath‐holding include the size of the shunt orifice, the pressure gradient between the chambers or vessels that produce the shunt, and the “downstream” resistance to flow (Table 33.2) [2]. Both diving and breath‐holding also increase peripheral vascular resistance resulting in bradycardia, a normal response in reptiles, leading to the redirection of blood flow to the brain and the heart. Reptiles switch from aerobic to anaerobic glycolysis during extended breath‐holding resulting in restricted pulmonary blood flow and R‐L shunting of blood to ensure that blood flow continues to the systemic circulation [4]. The resumption of normal breathing reverses these events, decreasing pulmonary resistance, increasing heart rate, and reducing R‐L shunting. These physiologic changes can be pharmacologically induced by dissociative anesthetics (e.g., ketamine and tiletamine), α2‐adrenergic receptor agonists (e.g., medetomidine and dexmedetomidine), or propofol and should be anticipated [4]. Snakes possess a vertebral venous plexus comprised of a network of spinal veins located within and around the vertebral column. The plexus is supported by the surrounding bones providing a route for venous return and the maintenance of cerebral blood supply [5]. A renal portal system is found in birds, amphibians, reptiles, and fish, suggesting that drugs metabolized or excreted predominantly by the liver will undergo a greater hepatic first‐pass effect (e.g., opioids and most anesthetics) [6]. Venous blood returning from the tail and hindlimbs is filtered through the kidneys and through the liver. Valves located between the abdominal and femoral veins regulate blood flow through the kidneys especially during times of water conservation. Parenteral drug administration of potentially nephrotoxic drugs into the tail or caudal extremities is generally not recommended because of the potential for nephrotoxicity [7]. The administration of non‐nephrotoxic drugs into the caudal extremities does not pose a risk [6]. Reptiles have a lymphatic system but lack lymph nodes. The lymph vessels have muscular dilations referred to as “lymph hearts” that propel lymph into the venous system [5]. Table 33.2 Physiologic effects of cardiac shunting in mammals. R‐L, right‐to‐left shunt; L‐R, left‐to‐right shunt. The circulatory anatomy of crocodilians is unique among reptiles and more like birds and mammals (i.e., four separate heart chambers and a complete ventricular septum), thus preventing intracardiac R‐L shunting. However, crocodilians retain the dual aortic arch system found in many reptiles [2]. The left aortic arch originates from the right ventricle next to the pulmonary artery, and the right aortic arch originates from the left ventricle. Importantly, the left and right aortic arches of crocodilians communicate via the foramen of Panizza located close to the heart (Fig. 33.2). This anatomic arrangement permits systemic venous blood to bypass the pulmonary circulation. The consequences of R‐L and left‐to‐right (L‐R) shunts can have important consequences for anesthesia (Box 33.1). An R‐L shunt bypasses the lungs diverting deoxygenated blood back into the systemic circulation, whereas an L‐R shunt recirculates pulmonary venous (oxygenated) blood back into the pulmonary circulation. Physiologic L‐R shunting in reptiles is not thought to contribute to pulmonary circulation overload due to the ability of blood to be shunted either L‐R or R‐L, although it can contribute to pulmonary overcirculation in mammals. Autonomic tone, neurohumoral substances, and the activity of pulmonary stretch and chemoreceptors control pulmonary and systemic vascular resistances in reptiles and the direction and magnitude of cardiac shunting. Ventilatory status, particularly in diving animals, thermoregulation, feeding, and digestion may also influence cardiac shunting. For example, most reptiles breathe intermittently with long periods of apnea. Parasympathetic tone increases during apnea resulting in bradycardia and increased pulmonary vascular resistance promoting the development of R‐L shunting [2]. The heart of birds and mammals is comprised of four chambers: two thin‐walled atria separated by an interatrial septum, and two thick‐walled ventricles separated by a complete interventricular septum. The unique respiratory system of birds includes the presence of air sacs that facilitate gas exchange throughout the inspiratory and expiratory respiratory cycle. Intrapulmonary or intracardiac (congenital malformation) shunts can affect the onset and elimination of anesthetic drug effects in birds and mammals, particularly when inhalant anesthetics are administered (Box 33.1). The speed of inhalant anesthetic induction is determined by how fast the anesthetic partial pressure is reached in the brain, which is determined by the rate of anesthetic inflow into the lungs, the rate of anesthetic transfer to the blood, and the rate of anesthetic transfer from the arterial blood to the brain. Inhalant anesthetics with a low blood‐gas partition coefficient (low solubility) are generally expected to produce a more rapid induction and recovery in normal mammals. For example, anesthetic induction and washout in healthy pigs is more rapid with the less soluble desflurane than with sevoflurane, which is more rapid than with isoflurane [8]. An L‐R cardiac shunt has minimal or no effect on the onset or elimination of injectable or inhalant anesthetics in mammals (Box 33.1). The rate of inhalant anesthetic induction is unchanged even though recirculation of blood through the lungs promotes a more rapid rate of rise in alveolar partial pressure. The speed of induction increases following the administration of an intravenous anesthetic to animals that have an R‐L shunt due to its immediate transfer to the brain. An R‐L shunt slows the rate of inhalant anesthetic induction and elimination, and this effect is more pronounced with less soluble anesthetics (e.g., isoflurane, sevoflurane, and desflurane) compared to a more soluble anesthetic (e.g., halothane). This effect is dependent upon the size of the shunt and the dilutional effect of shunted blood, which contains no volatile anesthetic. One study demonstrated that the rate of rise of the end‐tidal concentration of the comparatively soluble anesthetic halothane was not significantly affected by an R‐L intracardiac shunt in children [9]. The boundaries of the cardiac chambers in mammals are easily defined by the great veins (i.e., cranial and caudal venae cavae) that return blood to the right atrium, the smaller pulmonary veins that return oxygenated blood from the lung to the left atrium, the coronary sulcus (i.e., groove) that demarcates the atria from the ventricles, and the anterior (i.e., ventral) and posterior (i.e., dorsal) interventricular sulci that contain branches of the left circumflex coronary artery that separate the right and left ventricles (Fig. 33.3). All four chambers of the heart are easily visualized (Fig. 33.3A). The ventral interventricular branch of the left main coronary artery, occasionally referred to as the “left anterior descending coronary artery” in humans, provides blood supply to the ventricular septum and left ventricular free wall. The dorsal interventricular artery, generally an extension of the left circumflex coronary artery (the other major branch of the left main coronary artery), provides blood supply to most of the left ventricle. The right coronary artery provides blood supply to the right ventricular free wall and occasionally a large portion of the left ventricle (Fig. 33.3B). Anatomical differences in the anatomy of the coronary circulation among species account for dissimilarities in the source and distribution of myocardial blood flow. The left interventricular (i.e., anterior descending) coronary artery and its diagonal branches supply a larger mass of myocardium than the circumflex in humans, swine, rats, and nonhuman primates. The circumflex artery supplies the largest mass of cardiac muscle in dogs, cats, horses, and cattle although there is considerable species variation. Dogs have an extensive collateral coronary circulation (i.e., a network of tiny blood vessels that are not open under normal conditions but open during reduced blood flow), while pigs, monkeys, and humans do not. Cardiac veins empty into the great cardiac vein, which empties into the right atrium via the coronary sinus and accounts for over 90% of the venous return from the heart. Small capillary‐like vessels (Thebesian veins) are present throughout the endocardium and drain directly into the cardiac chambers accounting for approximately 4% of the heart’s venous return. Blood returning from the systemic circulation (i.e., venous return) and pulmonary veins empties into the right and left atrium, respectively. The ventricles are the major pumping chambers of the heart and are separated from the atria by a fibrous skeleton of dense connective tissue (i.e., fibrous trigone) that anchors the atrioventricular (AV) (i.e., tricuspid and mitral) and semilunar (i.e., pulmonic and aortic) heart valves. The tricuspid valve is located between the right atrium and right ventricle and the mitral valve between the left atrium and left ventricle. The pulmonic valve is located between the right ventricle and pulmonary artery and the aortic valve between the left ventricle and aorta. Once the process of cardiac contraction is initiated, almost simultaneous contraction of the atria is followed by nearly synchronous and simultaneous contraction of the ventricles. Cardiac contraction produces pressure changes that are responsible for AV and semilunar valve opening and closing and the production of audible heart sounds. Fibrous bands (i.e., chordae tendineae) originating from papillary muscles located on the inner wall of the ventricular chambers are attached to the free edges of the AV valve leaflets and limit valve prolapse into the atria during ventricular contraction thereby preventing the regurgitation of blood into the atrium (Fig. 33.3A). Alterations in heart chamber geometry (e.g., stretch, dilation, or hypertrophy) produced by changes in blood volume, ventricular deformation (pericardial tamponade), incompetent heart valves, or disease (e.g., tumors and scarring) can have profound effects on ventricular pressures, blood flow, and myocardial function. Figure 33.3 A. The mammalian heart is comprised of two thin‐walled (i.e., right, and left atria) and two thick‐walled chambers (i.e., right and left ventricles) and contains a highly reactive and diffuse vascular network (i.e., coronary circulation). The thin‐walled right atrium (RA) and left atrium (LA) are separated from each other by a membranous septum. The interventricular muscular septum (IVS) lies between the right and left ventricles. The tricuspid valve (TV) and mitral valve (MV) lie between the RA and RV and the LA and LV, respectively. The pulmonary and aortic (AV) valves separate the RV and LV from the pulmonary artery (PA) and aorta (Ao), respectively. B. A fibrous trigone located at the center of the heart provides the scaffolding for the heart valves, atria, and ventricles. The right and left coronary arteries emerge from the base of the aorta and supply blood to all the chambers of the heart including the heart valves. The maintenance of normal cardiac electrical and mechanical activity is dependent upon the energy (i.e., adenosine triphosphate [ATP]) derived from multiple metabolic pathways. Mitochondria generate more than 95% of the ATP used by the heart, in addition to regulating intracellular calcium concentrations, cell signaling, and coordinated cell death. A detailed description of these pathways is beyond the scope of this chapter and is more appropriately described in reviews and texts specifically dedicated to this subject [10,11]. The heart produces ATP when mitochondrial oxidative phosphorylation is fueled by electrons generated by dehydrogenation of carbon fuels (i.e., fats and carbohydrates) and the oxidation (i.e., electron transfer) of nicotinamide adenine dinucleotide (NADH) and flavin adenine dinucleotide (FADH2). NADH and FADH2 are primarily produced by fatty acid oxidation, the citric acid cycle, and to a minor extent, by the pyruvate dehydrogenase reaction and glycolysis (Fig. 33.4). Oxidative phosphorylation is responsible for approximately 60–90% of the ATP formation in the heart during normal non‐ischemic conditions with the remainder derived from the citric acid cycle. The citric acid cycle is dependent upon acetyl‐CoA formed by the decarboxylation of pyruvate and the oxidation of fatty acids (Fig. 33.4) [10]. Approximately 60–70% of ATP hydrolysis is consumed by cardiac contraction, while the remaining 30–40% is used to maintain the ion pumps and sarcoplasmic reticulum Ca2+ ATPase. The rate of oxidative phosphorylation is directly linked to the rate of ATP hydrolysis so that ATP concentrations remain constant during intense exercise or changes in neurohumoral activity (e.g., variations in catecholamine concentrations) [11]. Figure 33.4 ATP is primarily produced (60–90%) by mitochondrial oxidative phosphorylation of free fatty acids but can also be generated by glycolysis (10–40%). Dehydrogenation reactions produced by the fatty acid β‐oxidation pathway, citric acid cycle, and glycolysis generate reducing equivalents (i.e., NADH) that deliver electrons to the electron transport chain resulting in ATP formation. Adenosine triphosphate (ATP) and its metabolite (adenosine diphosphate) provide energy for myocardial contraction, membrane pump activity, and ion homeostasis, NADH, reduced nicotinamide adenine dinucleotide. A highly integrated series of electrical, metabolic, and mechanical events are required to produce normal cardiac contractile function (i.e., excitation–contraction coupling). Notably, myocardial contraction is preceded by, and will not occur without, electrical activation although normal or near‐normal electrical activity is possible without myocardial contraction (e.g., electrical–mechanical uncoupling, electrical–mechanical dissociation, and pulseless electrical activity [PEA]). PEA is the presence of organized cardiac electrical activity, other than ventricular fibrillation or ventricular tachycardia, that should produce a pulse, but does not. The cardiac cell membrane (sarcolemma) is a highly specialized lipid bilayer that contains protein‐associated channels, pumps, enzymes, and exchangers in an architecturally organized, yet fluid (i.e., reorganizable and movable) medium [12]. The molecular composition and fluidity (i.e., freely movable membrane protein and lipid constituents) of cardiac membranes determine their ion transport/exchange and membrane‐associated electrical properties, as well as the ability to generate and propagate electrical impulses [12,13]. The unequal distribution (i.e., concentrations) of various ions (i.e., sodium, potassium, chloride, and calcium) across the sarcolemma and its permeability to these ions are responsible for the development of the resting membrane potential. Resting membrane potential can be estimated by the Nernst equation or more accurately determined by the Goldman–Hodgkin–Katz constant‐field equation [12] as shown below. Nernst equation for potassium ions: where E is electromotive force; 2.303 is the conversion of ln to log10; Ci and Co are ion concentrations inside (i) and outside (o) the cell membrane; R is the gas constant; T is absolute temperature in Kelvin; z is valence of the ion (in this case, 1); F is the Faraday constant; and mV is millivolts at 37 °C. Goldman–Hodgkin–Katz equation: where Em is the resting membrane potential and PNa (0.04), PK (1.0), and PCl (0.45) are the relative permeabilities of the membrane at rest, and other terms are defined above. Transmembrane changes in the electric potential (i.e., action potential) generated by cardiac cells are the result of transmembrane ion fluxes (active properties) that generate currents through voltage‐ and ligand‐gated membrane channels (Fig. 33.5; Table 33.3) [12,14,15]. The term “gated” refers to the opening (activation) or closing (deactivation or inactivation) of ion channels. Ion channels are characterized by their ionic selectivity, conductance, gating characteristics, and density. The channel gating mechanisms control ion passage and are composed of both activation and inactivation gates, which are voltage‐ and frequently time‐dependent (e.g., sodium channels). The functional configuration of the gates determines channel state: activated or open, inactivated or closed, and resting (i.e., capable of being activated) [13]. The directional movement (inward or outward) of the various ions ultimately depends on channel state and the electrochemical driving force (i.e., equilibrium potential minus membrane potential) for each ion. The electrochemical driving force, as illustrated by the Nernst equation, is composed of an electric force and the transmembrane ion concentration gradient. It should be noted that most local anesthetics (e.g., lidocaine, bupivacaine, etc.) demonstrate use‐ (i.e., frequency‐) dependent block [16,17]. Use‐dependent block is the phenomenon exhibited by cardiac cells when, in the presence of a drug, increases in stimulation rate (i.e., chronotropy) produce a more pronounced drug effect because more drug molecules are able to enter the channel and cause more inactivation than during slower rates of stimulation [16]. Excitability (i.e., bathmotropy), or the ability of the cardiac cell membrane to be activated, generate, and propagate an electric potential (action potential), is a fundamental intrinsic property of cardiac cells [12,13]. The cardiac action potential (AP) morphology and duration vary considerably from that of nerves and skeletal muscle (Fig. 33.6A) [15]. The cardiac muscle AP arises from a more negative membrane potential (–90 versus –70 mV), may be of similar magnitude, but is much longer in duration (150–300 versus 1 ms) than APs recorded from nerve or skeletal muscle (Fig. 33.6A). Cardiac AP magnitude and duration also vary depending on the type of cardiac cell that generates them (Fig. 33.6B). Five characteristic phases of the cardiac AP are discernible in most cardiac cells: phase 0, or the phase of rapid depolarization, is caused by the rapid and relatively large influx of sodium ions (i.e., fast inward current) into the cell; phase 1, the early repolarization phase, is caused by the transient outward movement of potassium ions; phase 2, the plateau phase, is attributed to the continued, but decreased, entry of sodium ions and a large, but slower, influx of calcium ions (i.e., slow inward current) into the cell; phase 3 is the phase of membrane repolarization during which the membrane potential returns to its resting value primarily due to potassium efflux (i.e., outward current) from the cell; and phase 4 is a resting phase in atrial and ventricular muscle cells during which membrane pumps (i.e., Na+–K+ ATPase, Ca2+ pumps), and exchange mechanisms maintain the negative membrane potential and keep intracellular ion concentrations more or less constant prior to initiation of the next AP. For example, the Na+–Ca2+ exchanger (NCX) or antiporter membrane protein removes calcium from cells helping to restore normal transmembrane electrolyte concentrations (Fig. 33.5; Table 33.3) but is reversible, moving Ca2+ out or into the cell depending on the prevailing electrochemical driving forces for Ca2+ and Na+. Notably, a great deal of interest has been focused upon repolarization currents (IK) due to their importance in determining membrane resting potential, AP duration, and their involvement in repolarization abnormalities (i.e., inhomogeneities or dispersion of the QT interval) including long QT syndrome [18,19]. Long QT syndrome is an uncommon hereditary or drug‐induced condition (e.g., select antiarrhythmics, antidepressants, antipsychotics, antimicrobials, and opioids), in which AP duration becomes longer and may become highly variable among ventricular myocytes leading to electrical instability and the risk of reentrant induced torsades de pointes and ventricular fibrillation [20–22]. Figure 33.5 The normal action potential has five phases: Resting (4), Upstroke or depolarization (0), Early repolarization (1), Plateau (2), and Late or final repolarization (3). Gradual depolarization occurs at the end of phase 3 during phase 4 (blue dotted line) in automatic tissues (e.g., sinus node; Purkinje fibers). The inward currents, INa, ICa, If, and the sodium‐calcium exchanger (NCX) are shown in bold black lined boxes. NCX is electrogenic and can generate an inward or outward current depending upon the membrane potential. Outward currents, IKACh, IK1, Ito, IKur, IKr, and IKs are shown in light gray lined boxes. The action potential duration (APD) in ventricular muscle cells is usually > 100 milliseconds. Source: Grant [15], Nature Publishing Group. Table 33.3 Cardiac action potential (AP) currents. Figure 33.6 A. Nerve cell action potentials (light blue) are much shorter in duration (e.g., sodium based 1–2 ms) than cardiac muscle cell action potentials (black). Automatic cells determine heart rate (e.g., atrial pacemaker: orange). B. Action potential magnitudes and durations from various cardiac tissues. Table 33.4 Intrinsic rate, action potential duration, and conduction velocity in cardiac tissues of larger mammalsa. APD, action potential duration. a APD decreases and conduction velocity increases as intrinsic rate increases. b Atrial and ventricular muscle is not normally automatic. The speed of conduction of the cardiac electrical impulse (i.e., dromotropy) is determined by the magnitude and rate of sodium influx into the cardiac muscle cells and is directly related to the rate of voltage change (dV/dt) during phase 0 (Fig. 33.5; Table 33.4) [23]. A large dV/dt indicates more rapid depolarization of the cell membrane and more rapid conduction of the cardiac impulse through cardiac tissue [12]. The change (d) in transmembrane potential (V) is determined by the transmembrane ionic current (Iion) and the membrane capacitance (Cm; 1 μF/cm2) provided by the charge separation across the lipid bilayer: Ion movement occurs due to the activity of voltage‐gated ion channels, pumps, and exchangers [13]. A negative Iion indicates the inward flow of positive ions into the cell and produces a positive dV/dt, which depolarizes (i.e., makes the inside of the cells more positive) the membrane potential. A positive Iion indicates an outward flow of positive ions, which repolarizes the membrane potential by generating a negative dV/dt. The generation of the cardiac AP results from the time‐, voltage‐, and concentration‐dependent evolution of Iion and represents the contribution of multiple ion‐selective mechanisms for ion movement across the cardiac cell membrane (Fig. 33.5) [23]. Current flows from a depolarized cell to its neighboring less depolarized cells via intercellular resistive pathways known as “gap junctions.” Gap junctions are hydrophilic intercellular protein channels (i.e., connexons) composed of six connexin proteins that allow ions and electrical impulses to pass between cells through regulated hydrophilic channels [23]. Voltage‐gated ionic channels and gap junctions are depressed (i.e., resistance is increased) by anesthetic drugs through direct interaction with their protein subunits, thereby interfering with cardiac excitability, transmission, and propagation of electrical impulses [24–26]. Calcium enters cardiac cells through L‐type calcium channels during phase 2 of the AP triggering intracellular calcium release from the sarcoplasmic reticulum, which is important for normal cellular contraction. Since calcium enters the cell more slowly than sodium (i.e., slow inward calcium current) and from a less negative membrane potential, cardiac cells with a reduced resting membrane potential (e.g., the sinoatrial [SA] and atrioventricular [AV] nodes) demonstrate a considerably smaller phase 0 dV/dt and slower conduction velocity compared to those with a more negative resting membrane potential (e.g., atrial and ventricular muscle and Purkinje cells) [11–13]. Potassium efflux from cardiac cells is controlled by a variety of mechanisms, including the concentration difference of potassium across the membrane and changes in membrane permeability to potassium as the membrane potential changes (Table 33.3). The channels responsible for repolarization (phase 3) are major determinants of cardiac AP duration and cardiac cell refractoriness (Fig. 33.7). The duration of the cardiac AP and refractoriness have important clinical implications relative to the amount of calcium that enters the cell during depolarization (i.e., longer AP duration permits more calcium entry) and the potential for arrhythmia development, respectively [12,27]. Arrhythmias develop if there are large differences (i.e., inhomogeneities) in myocyte AP duration (e.g., > 20–40 ms) among groups of adjacent cardiac cells because of the potential for electrical impulses to re‐enter cells that have rapidly repolarized and regained excitability and conduction capabilities (i.e., reentry). Cardiac cells with particularly long AP durations are thought to predispose to dispersion of refractoriness and afterdepolarizations (i.e., afterpotentials) that leads to electrical instability and cardiac arrhythmias [27–29]. Phase 4, diastolic depolarization, endows the heart with the unique property of automaticity (i.e., self‐excitation). SA automaticity is produced by depolarizing membrane currents and is influenced by local factors, including temperature, pH, and blood gases (PO2 and PCO2), extracellular potassium concentration, catecholamines, and various hormones (Fig. 33.8). Cells in the SA and AV nodes and specialized atrial and ventricular muscle cells (i.e., Purkinje) cells are automatic (Table 33.4) [30]. The resting membrane potential of these cells depolarizes toward a threshold potential, which when reached triggers the development of an AP. The ionic processes responsible for phase 4 diastolic depolarization vary among the various specialized tissues of the heart primarily because of differences in their resting membrane potential and cell type (e.g., SA node, AV node, and Purkinje cell), but all are dependent upon the funny current (If), a mixed sodium–potassium current that activates upon hyperpolarization [31]. The funny current, first described in SA node myocytes as an inward current activated during hyperpolarization, determines the end of an AP, the steepness of phase 4 depolarization, and the frequency of AP firing [31]. Cells in the SA and AV nodes have comparatively less negative maximum diastolic potentials (–65 mV) compared to cardiac muscle and Purkinje cells (≥ –90 mV) and depend upon If, the entry of calcium ions (i.e., inward calcium current), and a progressive decrease in membrane permeability to potassium efflux for their automaticity [12,13,31,32]. Automatic cells in atrial specialized pathways and the ventricular Purkinje network have a more negative maximum diastolic potential (≥ –90 mV) and depend more upon If [32]. The principal mechanisms responsible for altering automaticity are changes in the threshold potential, the rate of phase 4 depolarization, and the maximum diastolic potential following repolarization. The cardiac tissue with the most rapid rate of phase 4 depolarization (normally the SA node) is termed the “pacemaker” and determines the heart rate (HR). The cardiac pacemaker normally activates (i.e., overdrives) and, therefore, resets the automaticity of slower or subsidiary pacemakers, thereby preventing other potential pacemakers from controlling HR (i.e., overdrive suppression; Fig. 33.9A) [32]. Overdrive suppression of subsidiary pacemakers is caused by activation of the sodium–potassium (Na+–K+) pump, leading to membrane hyperpolarization and a longer time required to reach threshold (Fig. 33.9B) [13]. Subsidiary pacemakers are depressed by fast HRs because the Na+–K+ pump is more active at faster rates, resulting in a more negative maximum diastolic potential. Clinically, the administration of anticholinergic drugs (e.g., atropine and glycopyrrolate) can produce increases in sinus rate that eliminate (i.e., overdrive suppress) infrequent or slower automatic ventricular arrhythmias. Alternatively, ivabradine, a drug that selectively and specifically inhibits the cardiac pacemaker current If, slows HR [33]. Figure 33.7 Absolute (ARP), effective (ERP), and relative (RRP) refractory periods. During the ARP the cell is incapable of generating an electrical response (i.e., action potential). ERP: An electrical response can be generated during the ERP but is not conducted. RRP: A greater than normal stimulus can generate a conducted response. The functional refractory period (FRP; not shown) is the shortest interval between two consecutively conducted responses. Supranormal period: the period following the RRP during which a smaller than normal stimulus can generate a conducted action potential. Figure 33.8 The transmembrane potential of the sinoatrial node (i.e., normal pacemaker) is characterized by a maximum diastolic potential ≅ 60 mV that depolarizes toward threshold (i.e., phase 4 diastolic depolarization), a slow phase 0 caused by ICa, and a relatively rapid repolarization due to IK. The rate of phase 4 diastolic depolarization (i.e., automaticity) can be increased by increases in temperature, calcium, mild oxygen deficiency, catecholamines, decreases in extracellular potassium concentration, and anticholinergic drugs (e.g., atropine). Increases in extracellular potassium concentration and If inhibitors (e.g., ivabradine) decrease diastolic depolarization. The average of the voltage vectors produced by all the APs produced by activation of each cardiac cell is responsible for the body surface electrocardiogram (ECG; a graphical representation of the electrical depolarization and repolarization of the heart) (Fig. 33.10). Initiation of an electric impulse in the SA node is followed by rapid transmission of a wave of depolarization throughout the atria, giving rise to the P wave. Repolarization of the atria gives rise to the Ta wave, which is small and difficult to identify in smaller animals (e.g., rats, rabbits, dogs, cats, etc.) but becomes obvious in larger animals (e.g., horses and cattle) because the total atrial tissue mass is substantial enough to generate an electromotive force that can be detected electrocardiographically from the body surface. Similarly, depolarization of the SA and AV nodes does not normally generate a large enough electromotive force to be recorded at the body surface. Once the wave of depolarization reaches the AV node, conduction is slowed because of its less negative resting membrane potential (i.e., –40 to –60 mV) and reduced slope (dV/dt) of phase 0 (Fig. 33.6B). This is especially evident when the AV node is stimulated at higher rates due to a phenomenon called “decremental conduction,” which is characterized by poor propagation of the electrical impulse due to progressive decreases in membrane potential and AP dV/dt. Increased parasympathetic tone also produces marked slowing of AV nodal conduction, leading to first‐degree, second‐degree, and, occasionally, third‐degree AV block. Many drugs used for anesthesia, including opioids, α2‐adrenergic receptor agonists, and occasionally acepromazine, can increase parasympathetic tone, causing sinus slowing and bradyarrhythmias. The use of anticholinergic drugs (e.g., atropine and glycopyrrolate) is generally effective therapy in these situations but not when AV block is caused by inflammation or structural disease (e.g., fibrosis and calcification). Caution is indicated when atropine or glycopyrrolate are co‐administered with α2‐adrenergic receptor agonists due to their potential to produce or exacerbate ventricular arrhythmias. α2‐Adrenergic receptor agonist–anticholinergic drug combinations that produce cardiac arrhythmias are likely due to increases in the rate–pressure product and myocardial oxygen consumption, or myocardial ischemia in conjunction with functional impairment of autonomic tone (i.e., dysautonomia) [34,35]. Figure 33.9 A. Overdrive suppression occurs when a higher frequency of activation suppresses automaticity. Usually interpreted to mean that increase SA node firing suppresses other pacemaker sites. B. Overdrive suppression is mediated by enhanced Na+–K+ ATPase pump activity. Increased frequency of depolarizations increases intracellular Na+ stimulating the electrogenic (i.e., moves more Na+ outward than K+ inward) Na+–K+ ATPase pump thereby generating an outward (hyperpolarizing) current across the cell membrane and a more negative membrane potential, counteracting the depolarizing If current, thus slowing the rate of phase 4 diastolic depolarization and suppressing spontaneous impulse initiation. Figure 33.10 The electrocardiogram (ECG) is the average of the voltage vectors produced by the electrical activation (i.e., action potentials) of the heart during a cardiac cycle. SA, sinoatrial node; AV, atrioventricular node; brs, branches; int, interval; LA, left atrium; LV, left ventricle; m, muscle; RA, right atrium; RV, right ventricle; seg, segment. Figure 33.11 A. The specialized conducting pathways of the heart include the sinoatrial node, the atrial internodal pathways, the atrioventricular node and His bundle, and the right and left bundle branches. Bachmann’s bundle facilitates electrical signal transmission to the left atrium. The conduction velocity of the electrical signal varies in different cardiac tissues. B. The sequence of electrical conduction in the atrial and ventricles (1–4) from initiation (yellow) to termination (dark blue) in relation to the development of the ECG. The conduction of the electrical impulse through the AV node provides time for the atria to contract (i.e., atrial priming or kick) prior to activation and contraction of the ventricles. The delay in ventricular depolarization is responsible for the PR or PQ interval of the ECG (Fig. 33.10). This delay is functionally important, particularly at faster HRs, because it enables atrial contraction to contribute to ventricular filling. It is worth remembering that cells of the AV node depend upon both sodium and calcium ions for the generation and conduction of their AP. Thus, cells of the AV node are extremely sensitive to drugs that block transmembrane calcium flux, including excessive doses of anesthetic drugs and calcium channel antagonists (e.g., verapamil and diltiazem). Excessive doses of injectable or inhalant anesthetic drugs can produce post‐repolarization refractoriness, a phenomenon wherein cardiac cells remain refractory to electric activation after complete repolarization. This phenomenon is likely to produce AV block as HR increases. In addition, increases in parasympathetic tone in the presence of drugs or diseases (e.g., inflammation and ischemia) that interfere with conduction of the electric impulse through the AV node can lead to ECG sinus bradycardia and first‐degree (prolonged PR interval), second‐degree (blocked P wave), or third‐degree (dissociation of P from the QRS complex) AV block. Once the electrical impulse has traversed the AV node, it is rapidly transmitted via specialized pathways to ventricular muscle cells. Bundles of Purkinje cells (the right and left bundle branches) transmit the electrical impulse to the ventricular septum and the right and left ventricular wall papillary muscles, respectively, via the moderator band in the right ventricle and the left anterior and posterior divisions of the left bundle branch, in the left ventricle (Fig. 33.11A). Purkinje fibers conduct the electric impulse at relatively rapid speeds (3–4 m/s) throughout the ventricles. Purkinje cells and midmyocardial (M) cells located at the terminal ends of the Purkinje bundle branches and in the middle of the ventricular walls, respectively, have the longest AP durations and, therefore, serve as physiological “gates” preventing reentry or recycling of electric impulses into the ventricular myocardium [36,37]. The comparatively long duration of Purkinje and M cell APs is also considered to be one cause for the development of U waves in the ECG (Fig. 33.10). It is important to remember that the electrical transmission of the cardiac AP ultimately depends on spatial variation in myocardial cell refractoriness and the maintenance of uniform (isotropic) cell‐to‐cell resistive and capacitive (passive force) membrane properties that are largely determined by low‐resistance gap or nexus junctions between cells [38,39]. The interval beginning immediately after the terminal S wave of the QRS complex (J point) and the ventricular repolarization T wave is referred to as the “ST segment.” Elevation or depression of the ST segment (±0.2 mV or greater) from the isoelectric line is indicative of myocardial hypoxia or ischemia caused by low cardiac output (CO), hypotension, anemia, pericarditis, or cardiac contusion, and suggests the potential for arrhythmias to develop. The ST segment is followed by, and may slur into, the T wave [40]. The configuration and magnitude of the T wave vary considerably among species and are particularly influenced by changes in HR, blood temperature, and the extracellular potassium concentration. Hyperkalemia, for example, produces an increase in membrane conductance to potassium and shortens repolarization producing large magnitude, spiked, or pointed T waves and a shortened QT interval. U waves, when present, can be distinguished immediately following the T wave (Fig. 33.10) [36,41]. Like Ta waves, U waves are more frequently observed in larger species of animals (horses and cattle). As described previously, the ECG is produced by recording the electrically synchronized depolarization and repolarization of the atria and ventricles. The morphology of the various waveforms (i.e., P, QRS, and T waves) that characterize atrial and ventricular depolarization and repolarization is dependent upon the different patterns of depolarization, electrode placement, the size of the cardiac chambers (e.g., hypertrophy increases the electromotive force generated), the orthogonal direction of the dominant electromotive force, resistance to electrical signal transmission (i.e., fat, fluid, and air), and a host of electrode, equipment, and environmental factors (e.g., electrical noise). P wave morphology will vary depending on where within the crescent‐shaped SA node electrical activation originates and spreads into the right atrium. The electrical signal is then transmitted via intra‐atrial and interatrial specialized pathways that depolarize the right and left atrial (e.g., Bachmann’s bundle) muscle fibers, respectively (Fig. 33.11A). The right atrium is normally activated before the left atrium since the electrical signal normally originates in the SA node. This sequence of activation (i.e., right atrium then left atrium) may be hard to appreciate in smaller animals (e.g., rats, rabbits, cats, and small dogs) since their atrial muscle mass is relatively small and the electrical activation process is completed in a shorter time compared to larger animals (e.g., pigs, horses, goats, and cattle). The larger atrial muscle mass in large animals or in larger smaller species (e.g., dogs) with an enlarged left atrium can often be identified by a bifid P wave. Once the atrial electrical signal traverses the AV node, it activates the rapidly conducting Purkinje fiber network. Bundles of Purkinje fibers form the right and left bundle branches, which distribute the wave of depolarization to the respective papillary muscles and ventricular muscle cells of the right and left ventricles generating the QRS wave that represents ventricular depolarization (Figs. 33.10 and 33.11B). Purkinje fibers penetrate a short distance from the subendocardium in smaller animal species (Category I: rats, rabbits, cats, and dogs) while larger animal species have evolved a more diffuse Purkinje fiber distribution that extends close to the subepicardium (Category II: sheep, goats, cattle, pigs, and horses). This difference in Purkinje fiber anatomical distribution ensures the synchrony of ventricular activation and contraction. Fiber‐to‐fiber spread of the ventricular myocardial electrical signal occurs from the endpoints of the Purkinje fibers in the subendocardium of Category I (i.e., smaller) species (Fig. 33.12A) or wherever Purkinje fibers terminate in Category II (i.e., larger) species (Fig. 33.12B). Orthogonal (i.e., X, Y, Z; ECG leads: I, aVf, V10) ECG lead placement illustrates differences in the dominant ventricular electrical activation forces, spatial orientation, and pattern of ventricular depolarization generated by animals in Category I (e.g., cats, dogs, and humans) and Category II (Category IIA: sheep, goats, and cattle; Category IIB: pigs and horses) (Fig. 33.13). Several inhalant (e.g., halothane, isoflurane, and enflurane) and injectable (e.g., thiopental, propofol, and opioids) anesthetic drugs are known to sensitize the myocardium to catecholamines or facilitate torsadogenicity resulting in the development of cardiac arrhythmias [42–45]. Cardiac sensitization to anesthetic drugs has been linked to cardiac α1‐adrenerigc receptor stimulation and is likely produced by drug‐related alterations in intracellular calcium cycling, alterations in electrical impulse propagation, and muscle cell excitability [46,47]. Anesthetic drugs, particularly volatile anesthetics, are also known to interact with multiple cardiac ion channels. Halothane, isoflurane, and sevoflurane depress ion channels at different minimum alveolar anesthetic concentration (MAC) multiples: sodium (> 2 MAC), calcium (~30% at > 2 MAC), potassium (~20% at > 1 MAC), and chloride (> 2 MAC) [48]. Inhibition of these channels may alter AP shape (i.e., triangulation) and conduction velocity and these are the two most common proarrhythmic changes in the cardiac AP associated with volatile anesthetics [48–51]. Triangulation allows more time for Na+–Ca2+ exchange, reactivation of the sodium current, reduced synchronization of APs, and facilitation of re‐excitation predisposing to the development of early afterdepolarizations [50,51]. Slowed conduction and tissue refractoriness are key factors for determining whether reentrant arrhythmias occur. In addition to these mechanisms, autonomic nervous system imbalances lead to events that predispose to cardiac repolarization and conduction abnormalities, coronary vasospasm, and unifocal or multifocal ventricular tachycardias. Autonomic imbalance may be responsible for the cardiac arrhythmias attributed to the concurrent administration of anticholinergic (e.g., atropine) and α2‐adrenerigc receptor agonist (e.g., medetomidine and dexmedetomidine) drugs [34]. γ‐Aminobutyric acid type A (GABAA) anesthetics (i.e., propofol, etomidate, and alfaxalone) acting on various GABAA subunits (α, β, γ) may also inhibit postganglionic vagal input (i.e., reduce M2 receptor activity) to the heart provoking sinus tachycardia while several opioids, particularly methadone, are known to lengthen the QT interval predisposing to ventricular arrhythmias [52,53]. In summary, the active (i.e., ion movement) and passive (resistive and capacitive) properties of cardiac cell membranes determine the heart’s excitability (i.e., bathmotropy), automaticity (i.e., chronotropy), conduction (i.e., dromotropy), rhythmicity, and refractoriness. Injectable and inhalant anesthetic drugs produce alterations in cardiac ion channels and transmembrane ion flux that may impair mechanical contraction or induce cardiac arrhythmias [42,46,47,54,55]. Excitation–contraction coupling refers to the process whereby electrical activation of myocardial cells is transformed into muscle contraction [56–59]. The process begins with depolarization of the cardiac cell membrane and ends with sarcomere and subsequently muscle fiber shortening. The normal extracellular [Ca2+] is 10‐3 M compared with an intracellular [Ca2+] of 10‐7 M. The electrical activation of the sarcolemma and its extensions (i.e., transverse [T] tubules) into the cell interior initiates an influx of calcium ions through L‐type calcium channels that trigger the release (i.e., calcium‐induced calcium release [CICR]) of calcium from ryanodine‐sensitive Ca2+ release channels in the sarcoplasmic reticulum [59]. CICR raises the intracellular calcium concentration from 10‐7 to 10‐5 M resulting in myosin–actin cycling and sarcomere contraction (Fig. 33.14). While most calcium ions enter the cardiac cell through voltage‐dependent L‐type calcium channels (i.e., dihydropyridine channel or DHP channel), some calcium ions enter via the Na+–Ca2+ exchanger (NCX) mechanism [37]. Voltage‐dependent calcium channels are of two types (Table 33.3): a fast but transient (i.e., T‐type) channel that is activated earlier and at more negative potentials than L‐type channels and is believed to play a role in activation of pacemaker cells; and a slow, long‐lasting (i.e., L‐type) channel that is opened by cellular depolarization and is responsible for excitation–contraction coupling. T‐type channels are insensitive to sodium channel blockers (e.g., lidocaine or tetrodotoxin) and calcium antagonists [57]. L‐type channels are sensitive to specific types of calcium antagonists (e.g., verapamil, diltiazem, and nifedipine‐like compounds) [57,58]. The L‐type channels are more prevalent in atrial and ventricular muscle cells than are T‐type channels, open at less negative potentials, and account for the latter phases of calcium channel opening [56]. Both channels are physiologically linked via specialized bridging or spanning (i.e., “foot”) proteins that connect them to the ryanodine‐sensitive Ca2+ release channels in the sarcoplasmic reticulum. The foot structures are part of a high‐molecular‐weight protein complex termed the “ryanodine receptor” (RyR) because of its affinity for the insecticide ryanodine [60]. Multiple RyR isoforms exist and cardiac RyRs are of the RyR2 type [60]. Any drug that prolongs channel opening triggered by cytoplasmic Ca2+ will also promote RyR activation and increase cytoplasmic [Ca2+]. For example, mutations in the skeletal muscle isoform of the ryanodine receptor Ca2+‐release channel (RyR1) are responsible for susceptibility to malignant hyperthermia and can be triggered by inhalant anesthetics (e.g., halothane, isoflurane, sevoflurane, and succinylcholine) [61,62]. Membrane‐bound inositol 1,4,5‐trisphosphate receptors (IP3R) provide a second pathway for internal Ca2+ release [63]. Their subcellular localization in atrial, ventricular, and Purkinje fibers is about 50‐fold less than RyR2s in ventricular myocytes. They are believed to modulate transcription, amplify RyR2 Ca2+ signals, and provide independent cellular activation through diverse pathways that generate IP3. The activation of IP3Rs by select agonists (e.g., angiotensin II, endothelin, and norepinephrine) is believed to be important in development of cardiac arrhythmias, hypertrophy and heart failure [63]. Notably, inhalant (i.e., halothane, isoflurane, sevoflurane, and desflurane) and injectable (i.e., propofol, etomidate, and midazolam) anesthetics produce dose‐dependent inhibition of the cardiac L‐type calcium current by interacting with and inhibiting DHP and RyR calcium channels thereby reducing CICR and the force of cardiac contraction [64–67]. Figure 33.12 A. Depolarization reaches every point in the myocardium by traveling rapidly in specialized pathways and slowly from muscle fiber to muscle fiber. Hearts from larger animals (e.g., horses, cattle) or hearts from species with large ancestors (e.g., sheep and goats) require more complete penetration of Purkinje fibers to achieve adequate synchronization of contraction. B. Conduction velocity along His–Purkinje fibers (red arrows) is ~2–4 m/s for all animals. Conduction patterns within the ventricular myocardium from the endpoint of the Purkinje fiber network (red dot) are responsible for different patterns of ventricular activation and body surface ECG patterns: Category 1 (e.g., guinea pig, rabbits, dog, cat, and humans); Category IIA (e.g., goat and cow); Category IIB (e.g., horse, pig, dolphins). Source: Dr. Robert Hamlin, with permission. Figure 33.13 QRS patterns in orthogonal (i.e., X, Y. Z: I, aVf, V10) ECG leads in Category I (e.g., cat, dog, and human) and Category II (e.g., Category IIA: sheep, goat, and cattle; Category IIB: pigs and horses) animals. Source: Dr. Robert Hamlin, with permission. Figure 33.14 Cardiac muscle cell excitation–contraction coupling. Cardiac muscle cells (cardiomyocytes) are surrounded by a surface membrane with tubular invaginations (T‐tubules). The T‐tubular network assists action potential propagation into the cytoplasm of cardiac muscle cells. Calcium enters the cardiac cell during phase 2 of the action potential via L‐type calcium channels (i.e., dihydropyridine receptors). The junctin, triadin, calsequestrin, and ryanodine receptor form a quaternary complex (i.e., foot protein) that triggers normal intracellular Ca2+ release from the sarcoplasmic reticulum (SR) by a process termed “calcium‐induced calcium release” (CICR). Increases in intracellular calcium (10−7 to 10−5 M) activates the regulatory protein troponin (Tn) complex (i.e., Tn‐C on the thin (i.e., actin) filament allowing binding of thick filament myosin heads to attach to exposed binding sites on the thin filament (inset) resulting in filament sliding, sarcomere shortening and muscle contraction. Calcium is transported back into the sarcoplasmic reticulum (SR) by the SR‐Ca2+ ATPase pump (SERCA) and out of the cell by the sodium‐calcium exchanger (NCX), located in the surface membrane. Phospholamban modulates Ca2+ pump reuptake of calcium into the SR. Metabolism of adenosine triphosphate (ATP) and formation of adenosine diphosphate and inorganic phosphate, and reuptake of calcium by the SR cause sarcomere relaxation. Small quantities of cytosolic Ca2+ stimulate mitochondrion resynthesis of ATP. Cardiac muscle cells (myocytes) are composed of repeating core contractile units called “sarcomeres” (Fig. 33.15A). Sarcomeres contain the contractile proteins actin (i.e., thin) and myosin (i.e., thick) filaments, which are the smallest functional units of striated cardiac muscle cells. The thin actin filaments have plus and minus ends and are attached at their plus end to structural proteins (i.e., Z bands) that separate each sarcomere (Fig. 33.15B) [68]. Each actin filament contains two helical strands of actin intertwined with tropomyosin, which has periodic troponin complexes. Increases in intracellular calcium‐ion concentration initiated during phase 2 of the cardiac AP and amplified by CICR serve as the catalyst for actin–myosin interaction and sarcomere shortening. More specifically, calcium ions bind to the regulatory protein troponin C (Tn‐C; C for calcium) and remove the inhibitory function of troponin I (Tn‐I; I for inhibitor) on the chemical interaction between actin and myosin thick filaments, allowing actin–myosin crossbridges to form (Fig. 33.15C). Myosin heavy chain (MHC) is the actin‐activated molecular motor of muscle thick filaments that forms crossbridges with actin pulling the thin actin filament over the thicker myosin filament toward the center of the sarcomere, thereby shortening the sarcomere [68,69]. Two functionally distinct isoforms of myosin, alpha (α) and beta (β), are present in mammalian hearts both containing ATPase and actin‐binding domains. The isoforms are structurally similar and are key determinants of sarcomere force generation and shortening velocity [70]. The α isoform is more highly expressed in the atria and ventricles of smaller mammals (i.e., mice and rats), whereas the β isoform predominates in the ventricles of larger mammals (i.e., pig, dogs, and humans) [71]. β‐Myosin has lower ATPase activity and produces slower actin sliding velocities than α‐myosin but is able to generate greater force [72]. Cardiac MHCs are comprised of two domains: heavy meromyosin (HMM) and light meromyosin (LMM). HMM is subdivided into subfragments (S1 and S2) (Fig. 33.15D). The S1 subfragment is comprised of two actin‐activated ATPase globular heads (the motor domain) followed by a converter domain and a lever arm. The essential light chain (ELC) and regulatory light chain (RLC) comprise the lever arm. Both the ELC and RLC in conjunction with myosin‐binding protein C (MyBPC) modulate sarcomere activity, intermolecular interactions between the two myosin S1 heads, and the conversion of chemical energy (ATP) into heat (i.e., Fenn effect) and mechanical work (Fig. 33.15C) [73–75]. The force produced by myosin motors is fueled by ATP hydrolysis resulting in the sequential release of inorganic phosphate (Pi) and ADP, strong actin–myosin binding, and the swing of the lever arm. The swing of the lever arm pulls the actin (i.e., power stroke) causing the sarcomere to shorten. The lever arm merges with the coiled‐coil tail of HMM (i.e., S2 subfragment) and connects to LMM molecules that are packed with the tails of other LMM molecules forming the backbone of the thick filament (Fig. 33.15C) [68,76]. Pairs of myosin S1 heads encircle the thick filament in a helical or quasi‐helical fashion and exist in one of three transitional states: active, disordered relaxed (DRX; approximately 50–60%), and super‐relaxed (SRX; approximately 40–50%) [77,78]. The SRX state is “off” or “hibernating.” SRX myosin S1 heads are folded back (“parked” or “docked”) along the thick filament. Parked S1 heads have markedly reduced ATPase activity, cannot bind to actin, and provide a method for regulating cardiac contractility in response to changes in loading conditions (i.e., preload and afterload) [79,80]. Activation of cardiac muscle only involves free heads in the DRX state. Myosin S1 heads in the DRX state display a flexible or free (“swaying”) position (Fig. 33.15C). Their partnered blocked myosin S1 head is docked onto its S2 tail and cannot bind to actin. Intermolecular interactions between myosin S1 heads and their S2 tails, ELC, RLC, and MyBPC favor the DRX or SRX states and are collectively referred to as the “interacting‐heads motif” (IHM) [68]. The regulatory protein MyBPC modulates force generation by maintaining the SRX (“off”) state [79,81–83]. Phosphorylation of the RLC releases one (i.e., free head) or both of the two myosin S1 heads, and increases myosin S1 step size, the velocity of sarcomeric shortening, isometric force, and power generation [68]. These interactions modulate the kinetics of actin–myosin crossbridge formation, myosin ATPase activity, and calcium utilization thereby acting as molecular regulators of muscle force and velocity generation [84]. Notably, stress sensed by the myosin thick filament destabilizes the SRX state releasing myosin S1 heads from their docked state and transitioning them from an “off” to an active state (i.e., positive cooperativity) [68,78,85]. Cycling of the myosin S1 heads (i.e., velocity of lever‐arm swinging; number that synchronously swing and form strong attachments) produces the force, velocity, and power needed for contraction [86,87]. Important steps in the actin‐activated myosin chemomechanical ATPase cycle (i.e., “swinging lever‐arm” model) include (Fig. 33.16): (1) ATP hydrolysis to ADP and Pi and adoption of a bent or “cocked” (pre‐power stroke) position by the myosin S1 head; (2) weak binding of the myosin S1 head containing ADP and Pi to actin; (3) transition of the myosin S1 head from a weak to strong binding state upon Pi release and lever‐arm swinging pulling the actin filament toward the center of the sarcomere (“swinging lever‐arm” model); (4) adoption of a post‐power stroke position associated with ADP release; and (5) binding of ATP to the myosin S1 head weakening its interaction with actin. The rate of Pi release is the rate limiting step that determines the duration that myosin S1 heads are strongly bound to actin. The velocity of actin filament sliding until the myosin S1 heads detach is determined by ADP release (i.e., detachment rate [kdet]) [84,87,88]. Cyclic asynchronous ATP‐driven actin–myosin mechanosensing and biased Brownian motion driven by increased ionic strength are responsible for the lever‐arm swing, myosin displacement, sarcomere shortening, and force development [89,90]. Maximum velocity of contraction (Vmax) is derived from the unitary displacement divided by the myosin head detachment rate. Thus, the load‐dependent attachment rate, overall cycle rate, amount of time that myosin is attached to actin, and the total number of myosin S1 heads in the active state (i.e., duty ratio) determine the developed force (F), sliding velocity (Vmax), and amount of ATP consumed [90–93]. Increases in intracellular calcium facilitate increased myosin ATPase activity. Therefore, intracellular calcium concentration serves as the principal factor for determining the rate at which crossbridges attach and detach. The rate of crossbridge detachment and the number of interacting crossbridges (e.g., baseline: 20–30% of total crossbridges) is the basis for the force–velocity relationship that determines cardiac contractility [84,87,94,95]. The concept of optimal sarcomere length serves as the explanation for the Frank–Starling law of the heart and predicts an increase in contractile force when sarcomeres are stretched (e.g., increased ventricular volume) to their optimal length (i.e., approximately 2.2 μm) [94–96]. It is unlikely, however, that this explanation provides an adequate explanation of the Frank–Starling effect since sarcomeres rarely change length even during dilated forms of heart failure [37]. The more probable reason for Starling’s law of the heart is that sarcomere loading increases troponin C affinity for calcium and RLC phosphorylation leading to increased activation of the myofilament and sarcomere shortening without increases in sarcomere length or additional increases in intracellular calcium [96,97]. Figure 33.15 A. Cardiac muscle cells contain sarcomeres, mitochondria, and a sarcoplasmic reticulum that occupy ~60%, 35%, and 5% of the cell volume, respectively. Sarcomeres contain an overlapping array of thin (actin) and thick (myosin) protein filaments that produces a striated or banded (i.e., A, H, I) appearance. Sarcomeres are separated by Z‐discs (i.e., bands). B. The left half of a sarcomere illustrates the polarized (‐/+) thin actin filament connected to the Z‐disc and the myosin thick filament connected to the center of the sarcomere (M‐line). The mechanosensing protein titin regulates myosin length and connects myosin to the M‐line and Z‐Disc. Myosin‐binding protein C (MyBPC) modulates the force of contraction. C. Actin–myosin crossbridges are formed when myosin heads bind to myosin‐binding sites on the actin filament. Myosin access to actin‐binding sites is regulated by the troponin (Tn) complex comprised of three regulatory proteins: troponin I, troponin T, troponin C. An increase in intracellular calcium concentration increases calcium binding to Tn‐C repositioning tropomyosin (Tm) and exposing myosin‐binding sites on actin filaments to myosin heads. Myosin heads exist in free, disorganized (DRX), and folded‐back, super relaxed (SRX), or docked (i.e., parked) states, collectively referred to as the “interacting‐heads motif” (IHM). Myosin head binding is regulated by interactions among the MyBPC, the converter domain (dark dot), essential light chain (ELC) and the regulatory light chain (RLC). D. The myosin filament is composed of heavy meromyosin (HMM) that includes subfragment 1 (myosin S1 head [S1], essential light chain [ELC], regulatory light chain [RLC]) and subfragment 2 (S2) that transitions to a coiled‐coil tail and then into light meromyosin (LMM). S2 tails combine with other light meromyosin (LMM) tails to form the thick myosin filament. Figure 33.16 The actin‐activated myosin chemomechanical ATPase cycle (i.e., ATP binding, hydrolysis, Pi release, ADP release). A. ATP binds to the ATPase containing the myosin S1 globular head (“molecular motor”) weakening its interaction with actin thereby freeing the myosin S1 head so that the lever arm can cock into a pre‐power stroke DRX position. B. One myosin S1 head is in the folded back SRX (docked on the coiled‐coil S2 tail) state and the other is in the post‐stroke DRX state (1). ATP hydrolysis transitions the myosin S1 head from a weak to strong binding state with actin (2). C. Pi release is associated with a leftward shift (leftward arrow) of the lever arm and movement of the actin filament to the left (3). The swing of the lever arm is responsible for force production (power stroke). D. ADP dissociates, forming the rigor state (4). Binding of ATP to myosin S1 head dissociates it from actin forming the post‐rigor state (5). Decreased interaction between actin and myosin filaments is directly related to a decrease in [Ca2+]i and signals the beginning of the actin–myosin uncoupling process and myocardial relaxation [98]. Three principal mechanisms are responsible for reducing intracellular calcium‐ion concentration and the subsequent decrease in cardiac contractile force. Depolarization‐triggered increases in intracellular calcium increase the activity of the calcium regulatory protein calmodulin. Calmodulin serves as an intracellular calcium sensor and, when activated (i.e., calmodulin‐calcium complex), stimulates the active extrusion of calcium by pumps in the sarcolemma. Phosphorylation of the phospholamban‐modulated calcium pump increases calcium uptake by the sarcoplasmic reticulum and enhances the activity of the Na+–Ca2+ exchanger (NCX) (Fig. 33.14) [99]. Calcium transported by the sarcoendoplasmic reticulum calcium transport ATPase (SERCA) lowers the cytosolic calcium concentration and removes calcium from Tn‐C [100]. The reduced intracellular calcium induces a conformational change in the troponin complex allowing Tn‐I inhibition of the actin‐binding site. Upon completion of the chemomechanical cycle, a new ATP binds to the myosin S1 head displacing ADP and restoring the initial sarcomere length, and the entire cycle is ready to be repeated (Fig. 33.16). Most intravenous anesthetic drugs (e.g., barbiturates, ketamine, and propofol), and in particular the inhalation anesthetics, produce dose‐dependent decreases in cardiac contractility by decreasing calcium influx through L‐type channels, decreasing calcium release from the sarcoplasmic reticulum, and decreasing Tn‐C sensitivity to calcium [101–110]. Anesthetic drugs also interfere with the reuptake of calcium by the sarcoplasmic reticulum ultimately leading to depletion of intracellular calcium [98,106,107,111,112]. Notably, the effects of both injectable and inhalant anesthetics on the chemomechanical mechanisms responsible for myocardial contraction and relaxation require reexamination using updated contemporary experimental technologies. The function of the vascular system (arteries, capillaries, and veins) is to transport blood, nutrients, and oxygen to tissues, remove waste products (e.g., carbon dioxide), protect tissues, and stop bleeding. Blood vessels larger than 100–150 μm comprise the macrocirculation, while those that are smaller than 100 μm constitute the microcirculation [113]. The largest blood vessels are not embedded in organs and function as conduits for the delivery of blood to and from the heart, lungs, and peripheral tissues. Small vessels (i.e., < 100 μm) are embedded in organs and are actively involved in regulating blood flow or exchange processes. The pulmonary artery and aorta deliver blood to the pulmonary and systemic circulations, respectively. The pulmonary circulation receives its blood supply from the right ventricle via the pulmonary artery while the nutrient supply of the lung parenchyma is provided by the bronchial arteries. The systemic circulation receives its blood supply from the left ventricle via the aorta. The left ventricle, aorta, and other large arteries comprise the high‐pressure portion of the systemic circulation while the right ventricle, pulmonary artery, left atrium, and all the veins are considered low pressure (i.e., < 20–30 mmHg). Vessels of the systemic circulation undergo repeated division into smaller and smaller parallel vascular beds that terminate in arterioles (i.e., the smallest arteries at 10–30 μm) that subdivide and transition into capillaries (5–10 μm). Blood vessels can be functionally categorized as primarily elastic, compliant conduits (e.g., large arteries: aorta), muscular arteries (e.g., distributive arteries: femoral artery), sphincter or resistance vessels (e.g., arterioles), exchange vessels (e.g., capillaries), capacitance vessels (e.g., venules and veins), large conduit veins, and shunt vessels (e.g., arteriovenous anastomoses) [114]. All vessels have an endothelial surface and all but capillaries contain varying proportions of elastic fibers, smooth muscle, and fibrous tissue (Fig. 33.17). The tunica media is composed mostly of smooth muscle and elastic connective tissue, and the outer layer, the tunica externa (i.e., tunica adventitia), contains fibrous collagen fibers. The proportion of elastic connective tissue to smooth muscle determines the vessel’s principal function (i.e., conduit, resistive, or capacitive). Larger arteries possess a high proportion of elastic tissue in comparison to smooth muscle and fibrous tissues found in veins (Fig. 33.17). These structural differences allow the aorta to stretch during ventricular contraction and the ejection of blood. The potential (stored) energy imparted to the elastic fibers of the stretched proximal aorta is returned as kinetic (motion) energy assisting blood flow (i.e., the Windkessel effect) [115]. Thus, the highly elastic architecture of the aorta aids in converting pulsatile (non‐uniform) flow to continuous blood flow (Fig. 33.18). The Windkessel effect is believed to be responsible for up to 50% of peripheral blood flow in most animals during normal resting HRs. Figure 33.17 Blood vessel types, diameter, and wall thickness. Note the differences in vessel diameter, elastic tissue, and smooth muscle between arteries and veins. Figure 33.18 Aortic distensibility (i.e., compliance) and elasticity (i.e., ability to resume its normal shape after being stretched) are often conflated with stiffness. Aortic elasticity is responsible for the Windkessel effect wherein the aorta is distended during systole and recoils during diastole. Large distensible elastic arteries briefly store blood (i.e., hydraulic accumulator) during systole and distribute blood during diastole. Figure 33.19 A. The endothelial surface layer (ESL) is a multilayered structure that covers the endothelial surface of most blood vessels. It is composed of water, a dense inner layer of sulfated proteoglycans (e.g., syndecans, glypicans) and a less dense outer layer of covalently bound glycosaminoglycans (GAGs: heparan sulfate, chondroitin sulfate, and hyaluronan), glycoproteins, and plasma proteins. Transcytosis of albumin via caveolae is an important means for its transport across the endothelium. B. The electroviscous effects of the glycocalyx attenuate blood cell–vessel wall interactions and influences the permeability of larger negatively charged molecules (e.g., albumin). The luminal surface of all blood vessels is lined by endothelial cells (i.e., tunica intima; Fig. 33.17) that are covered with a multilayered endothelial surface layer (Fig. 33.19A). The glycocalyx is a thin layer of negatively charged (anionic) material composed of syndecans, glypicans, glycoproteins, and polysaccharides that extends from the surface of vascular endothelial cells and binds plasma proteins and soluble glycosaminoglycans (Fig. 33.19A) [116,117]. Endothelial cells are dynamic and have both metabolic and synthetic functions exerting autocrine, paracrine, and endocrine actions that modulate smooth muscle contraction and relaxation, platelet and leukocyte adherence, thrombosis, and thrombolysis. Endothelial cells generate and maintain the glycocalyx [117,118], a voluminous water‐phobic (negatively charged) intravascular compartment that plays an important role in maintaining vascular wall homeostasis, red and white blood cell movement, and vessel wall permeability in capillaries (Fig. 33.19B) [117–120]. Pathologic loss of the glycocalyx initiates breakdown of the vascular barrier and has been linked to ischemia, systemic inflammatory response, sepsis, and volume overload [117,121–123]. Large elastic arteries serve as conduits through which blood is transported to the periphery (Fig. 33.17). The elasticity of large arteries opposes the stretching effect that the blood pressure produces following ventricular contraction. For example, the initial stretching of the aorta produced by ventricular ejection is opposed by the elastic tissue in the vessel walls, which returns the aorta and large arteries to their original dimension once the pressure imposed by ventricular ejection subsides (i.e., the Windkessel effect) [115]. The degree to which the larger arteries can be stretched depends on the ratio of elastic to collagen fibers and fibrous tissue. Systemic veins are more than 30 times more distensible (i.e., compliant) than systemic arteries (Fig. 33.20A) and the lumped systemic circulation (i.e., arteries, capillaries, and veins) is more compliant than the lumped pulmonary circulation (Fig. 33.20B) [124]. Peripheral muscular arteries contain greater percentages of smooth muscle compared to elastic tissue, thereby providing greater control over vessel diameter, vascular resistance, and the regulation of blood flow (Fig. 33.17). The amount of smooth muscle determines the vessel’s resting tone (i.e., myogenic basal tone), the ability to respond to an increase in blood pressure, and the amount of stress relaxation (i.e., vasoconstriction followed by delayed relaxation; Bayliss effect), and reverse stress relaxation (i.e., vessel contraction due to fall in blood volume) [125,126]. Stress relaxation is characterized by a rapid initial increase in resting tone caused by an increase in pressure that declines gradually during the next several minutes. The blood pressure decreases because of smooth muscle myofilament rearrangement. Reverse stress relaxation is the reverse of this process [125]. The most distal branches of peripheral feed arteries terminate in “resistance vessels” that control the distribution of blood flow. These vessels contain a predominance of smooth muscle, are densely innervated by the sympathetic nervous system, and include arterioles, metarterioles, and arteriovenous anastomoses. Resistance vessels include the muscular arterioles, metarterioles (i.e., short vessels that link arterioles and capillaries), and arteriovenous anastomoses (Fig. 33.17). The small arteries and arterioles represent the primary vessels involved in the distribution of blood flow and regulation of arterial blood pressure. Notably, approximately 50–60% of the pressure drop between the aorta and capillaries occurs across terminal arterioles and metarterioles (Fig. 33.21). Functionally, precapillary resistance vessels (i.e., sphincter vessels) help to regulate the number of open capillaries and, therefore, the size of the capillary bed that is available for exchange processes. The relatively thick‐walled muscular arterioles and sphincter vessels are regulated by a variety of neural, humoral, and local metabolic factors and are the principal determinants of the total amount of blood flow distributed to all tissues of the body (Table 33.5). Arteriovenous anastomoses bypass the capillary by connecting arterioles to small veins. These vessels reduce or totally interrupt blood flow to capillaries and are densely populated by both α1‐ and α2‐adrenergic receptors [127]. Arteriovenous anastomoses possess smooth muscle cells throughout their entire length and are found in the greatest numbers in the skin and extremities (ears, feet, and hooves) of most species [127,128]. When open, they shunt arterial blood to venous plexuses in the limbs and play an important role in temperature regulation. Their identification in the intestine, kidney, liver, and skeletal muscles suggests their importance as a separate blood flow regulatory mechanism for controlling nutrient blood flow in these tissues. α2‐Adrenergic receptor agonists have been demonstrated to cause redistribution of CO away from less vital organs [129]. Figure 33.20 A. Lumped (i.e., arteries, capillaries, veins) pulmonary and systemic compliance (i.e., V/P). Lumped pulmonary compliance is less than lumped systemic compliance. A small increase in pulmonary blood volume causes a much larger increase in pulmonary pressure. B. Systemic arteries are much less compliant than systemic veins. Source: Adapted from Green [124]. Figure 33.21 Blood pressure (mmHg), flow velocity (cm/s), and vascular cross‐sectional area (cm2) in the systemic vasculature. As blood approaches the capillaries, blood pressure and blood flow velocity decrease and cross‐sectional area increase. Table 33.5 Factors producing vasodilation or vasoconstriction. The microcirculation includes the smallest blood vessels in the body and consists of the terminal arterioles, the capillary network (4–8 μm) and venules (≲ 10–100 μm) (Fig. 33.17). The microcirculation is embedded within organs and responsible for the exchange of oxygen (O2), carbon dioxide (CO2), nutrients, metabolites, and fluid between blood and tissue. Exchange occurs by three processes: passive diffusion, filtration, and pinocytosis (e.g., fluid endocytosis). A single layer of endothelial cells, normally covered by the glycocalyx, separates the intravascular volume from the interstitial fluid volume (Fig. 33.19A) [117]. The glycocalyx, part liquid and part solid, becomes denser closer to the endothelial cell surface and serves as the major barrier to transcapillary flow of macromolecules larger than 15 nm [130]. The dense inner region is composed of a network of membrane‐bound and membrane‐attached proteoglycans, glycosaminoglycans, glycoproteins, and adherent plasma proteins extending out from the endothelial surface (Fig. 33.19A) [115]. The capillary wall, therefore, is a three‐layered structure consisting of the glycocalyx on the luminal surface, the basement membrane on the abluminal surface, and endothelial cells in between. The intravascular volume consists of the glycocalyx, plasma volume, and red cell volume. The glycocalyx is normally the dynamic and active interface between plasma and the endothelial cell surface acting as a semipermeable molecular sieve with respect to anionic macromolecules such as albumin and other plasma proteins, whose size and structure determine their ability to penetrate the layer (Fig. 33.19B) [117–121]. The normal glycocalyx is impermeable to red blood cells (RBCs) and molecules larger than 70 kDa and semipermeable to albumin (molecular weight ~69 kDa; diameter ~7 nm; reflection coefficient 0.79–0.9) [129]. Increasing evidence, however, suggests that albumin transport from blood plasma to the interstitial fluid may occur via plasmalemmal vesicles (i.e., transcytosis) thereby serving as large pore (i.e., 20–70 nm) equivalents (Fig. 33.19A). For example, hydrostatic pressure‐dependent activation of albumin transcytosis in the lung microvasculature increases interstitial colloid osmotic pressure (COP) enhancing fluid flux across the endothelium, reducing intravascular pressure. Functionally, amphiphilic (i.e., having both hydrophilic and hydrophobic parts) molecules like albumin help to sustain and regulate the vascular and glycocalyx permeability [130,131]. There are three types of capillaries: continuous non‐fenestrated, fenestrated, and discontinuous or sinusoidal (Fig. 33.22A) [132]. Their distribution, porosity, and numbers vary in different tissue beds depending on tissue metabolism (O2 requirements) and the importance of fluid exchange. Continuous non‐fenestrated capillaries with tight junctions are present in all tissues of the body except epithelia and cartilage. They have a functional pore size of approximately 5 nm that permits the diffusion of water, small solutes, and lipid‐soluble materials into the interstitial fluid. The term “pore” is not a simple opening but refers to any path through (e.g., transcytosis) or between (e.g., intercellular cleft) endothelial cells that allows fluid to pass. Specialized continuous non‐fenestrated capillaries are found throughout most of the central nervous system (CNS), enteric nervous system, retina, and in the thymus (Fig. 33.22B). The endothelial cells are bound together by tight junctions with an effective pore size of < 1 nm [132]. Endothelial cells in the brain and spinal cord are tightly opposed by zonula occludens (i.e., tight junctions) with few interendothelial breaks and are responsible for the blood–brain barrier. The blood–brain barrier is only permeable to the smallest non‐lipid‐soluble molecules. Breaks within the interendothelial cell junctions produced by trauma or inflammation are the primary path for increased porosity and transvascular fluid filtration. Aquaporins (i.e., water channels) are present within continuous, but not fenestrated, vascular endothelial cells [133]. Aquaporins selectively facilitate the rapid transport of fluid across epithelial and endothelial cells but are also found in other tissues such as muscle and nerve cells. Aquaporins mediate osmotic water transport across plasma membranes and function as membrane channels for water alone or for water plus small molecules [132]. Anesthetic drugs (e.g., propofol, isoflurane, and sevoflurane) are known to influence aquaporin expression and water homeostasis, but additional research is required to determine the clinical relevance of these findings [134,135]. Fenestrated (i.e., windowed) capillaries with diaphragmed fenestrae and small gaps between adjacent endothelial cells are present in skin, connective tissue, kidney, intestinal mucosa, endocrine and exocrine glands, and the choroid plexus (Fig. 33.22B). The basement membrane of these capillaries is continuous. Their upper pore size is in the range of 6–12 nm [132]. Fenestrated capillaries with open fenestrae (i.e., “windows,”) span the endothelial lining (Fig. 33.22B). Windows permit the rapid exchange of water and solutes. Open fenestrated capillaries are present in the kidney cortex and medulla, the gastrointestinal mucosa, and the lymph nodes. Sinusoidal capillaries (sinusoids) resemble fenestrated capillaries but are discontinuous and are characterized by gaps between adjacent endothelial cells. The basement membrane and glycocalyx are absent, and interstitial fluid is part of the plasma volume in sinusoidal tissues (spleen, liver, bone marrow, and endocrine organs; Fig. 33.22B). Plasma proteins (e.g., albumin) secreted by liver cells pass easily through the sinusoids into the bloodstream through pores as large as 200–280 nm [132]. Phagocytic cells monitor the passing blood in sinusoidal tissues, engulfing damaged RBCs, pathogens, and cellular debris. Capillary fluid exchange is governed by two primary processes: diffusion and filtration. Diffusion is the movement of a specific molecule from an area of high concentration to a low concentration. Fick’s law of diffusion describes solute exchange (Js) as where D is the diffusion coefficient, A is the capillary surface area, MT is the membrane thickness, and CD is the transmembrane concentration gradient or difference. The diffusion coefficient is determined by the diffusion medium and the qualities characteristic of the diffusion particle such as the molecular weight, molecular shape, ionic charge, and lipid solubility. Filtration is the movement of fluid and molecules from an area of high pressure to one of low pressure. Fluid exchange (i.e., fluid flux [Jv]) by filtration is determined by four primary factors (Pc, Pi, πc, and πi) according to a dynamic equilibrium originally hypothesized by Starling but formulated by others (i.e., Starling principle of capillary fluid exchange), wherein where, in the classical Starling equation, Jv is fluid flux across the capillary (positive for filtration, and negative for reabsorption), Pc and Pi are capillary and interstitial hydrostatic pressures, πc and πi are the plasma and interstitial colloid osmotic pressures, Kf is the capillary filtration coefficient, and σ is the osmotic reflection coefficient for all plasma proteins [136–139]. The filtration coefficient (Kf) indicates the resistance of the capillary wall to fluid flow and is determined by the surface area, the number and radius of capillary pores, the capillary wall thickness, and the viscosity of the fluid being filtered. The osmotic reflection coefficient (σ) is an indicator of transvascular protein transport (i.e., a reflection coefficient of 0 implies that the substance is freely permeable and 1 implies that the substance does not pass through the membrane). The initial osmotic reflection coefficient for albumin and most colloid solutions (i.e., > 70 kDa) is close to 1 in normal animals since most capillary membranes are impermeable to large colloids. Providing that all factors can be accurately measured or approximated, net fluid flux across the capillary wall can be estimated by the Starling equation [139]. Historically, Starling’s principle of capillary fluid flux suggested fluid filtration at the arterial end of the capillary and fluid reabsorption at the venous end of the capillary. Lymph vessels were assumed to carry excess interstitial fluid back to the venous circulation. It is now understood, however, that non‐fenestrated capillaries normally filter fluid throughout their entire length, fluid flux to the interstitial space is under a dominant hydrostatic pressure gradient (capillary pressure Pc minus Pi), and that the effect of πc on transvascular fluid exchange is much less than predicted [119]. The low protein concentration within the subglycocalyx intercellular spaces accounts for low Jv and lymph flow in most tissues during normal physiologic conditions. Absorption through capillaries and venules does not occur except in abnormal physiologic conditions or disease states. During normal conditions, πc opposes but does not reverse filtration, and most of the filtered plasma returns to the circulation as lymph [120]. The glycocalyx covers the endothelial intercellular clefts in fenestrated capillaries, separating plasma from an almost protein‐free subglycocalyx space (Fig. 33.19). The subglycocalyx COP (πg) replaces πi as a determinant of transcapillary fluid flux (Jv). The fluid at the abluminal side of the glycocalyx is separated from the pericapillary interstitial fluid by the tortuous path through the intercellular clefts. Plasma proteins, including albumin, can escape to the interstitial space through the intercellular clefts or via transcytosis (i.e., large pores) and are responsible for the increased Jv observed during endothelial inflammatory states. Importantly, Jv can be modified by many factors including drugs and intravenous fluid replacement regimens that produce fluid overload. A more accurate depiction of transcapillary fluid flux is represented by the revised Starling equation (Fig. 33.23):
33
Cardiovascular Physiology
Introduction
Comparative cardiovascular morphology and shunting
Term
Description
Chronotropy*
Heart rate
Bathmotropy
Excitability
Dromotropy
Conduction
Inotropy
Force
Lusitropy
Relaxation
Clinotropy
Velocity
Physiologic effect
Direction of shunt
Decreased arterial oxygen content
R‐L
Decreased carbon dioxide removal
R‐L
Decreased lung plasma filtration
R‐L
Erythrocytosis
R‐L
Increased heart rate
R‐L
Increased arterial oxygen content
L‐R
Increased carbon dioxide removal
L‐R
Reduced ventilation–perfusion mismatch
L‐R
Pulmonary overperfusion
L‐R
Renin–angiotensin system activation
L‐R
The mammalian cardiovascular system
Anatomy
Metabolism
Electrophysiology and electrocardiogram
Current
Abbreviation
Effect
AP phase
Sodium
I Na
Rapid depolarization
Phase 0
Transient outward
I to
Transient repolarization
Phase 1
Delayed rectifier
I Kur
Rapid repolarization
Phase 1
L‐type calcium
I CaL
Plateau depolarizations
Phase 2
Delayed rectifier
I Ks
Slow repolarization
Phase 3
Delayed rectifier
I Kr
Rapid repolarization
Phase 3
Inward rectifier
I K1
Resting membrane potential
Phases 3, 4
Muscarinic potassium
I KAch
Inwardly rectifying SA node pacemaker
Phase 4
Pacemaker
I f
Hyperpolarization‐activated inward pacemaker current
Phase 4
Other Currents
T‐type calcium
I CaT
Transient depolarization
Phase 2
ADP activated potassium
I KATP
Inward rectification
Phases 1, 2
Na+–Ca2+ exchanger
NCX
Electrogenic exchange; reversible
All phases
Cardiac tissue
Intrinsic rate
APD (ms)
Conduction velocity (m/second)
Sinoatrial node
50–80
80–120
0.01–0.05
Atrial muscleb
–
120–200
1.0
Atrioventricular node
20–40
80–120
0.01–0.05
Bundle of His
20–40
120–200
1.5
Purkinje fibers
5–20
300–450
2–4
Ventricular muscleb
–
150–250
1.0
Excitation–contraction coupling
The actin‐activated myosin chemomechanical ATPase cycle
Myocardial contraction and relaxation
The vascular system
Large arteries and veins
Small arteries
Resistance vessels
Vasoconstriction
Vasodilation
Systemic
Increased noradrenergic discharge
Circulating catecholamines
Angiotensin II
Arginine vasopressin
Serotonin (5‐HTB receptors)
Endothelin A
Neuropeptide Y
Thromboxane A2
PgA2, PGF2α
Reactive oxygen species (superoxide, hydroxyl radical)
Activation of cholinergic dilators in skeletal muscle
Bradykinin
Histamine
Kallikrein
Endothelin B
Substance P (axon reflex)
PGI2, PGE2
Local
Decreased temperature
Increased temperature
Increased PO2
Decreased PO2
Increased H+, K+, lactic acid, H2S
Increased NO
Adenosine
Capillaries
Transcapillary fluid exchange

Stay updated, free articles. Join our Telegram channel

Full access? Get Clinical Tree

