Comprehensive Physiology Wiley Online Library

Assessment of Cardiac Function—Basic Principles and Approaches

Full Article on Wiley Online Library



ABSTRACT

Increased access and ability to visualize the heart has provided a means to measure a myriad of cardiovascular parameters in real or near real time. However, without fundamental knowledge regarding the basis for cardiac contraction and how to evaluate cardiac function in terms of loading conditions and inotropic state, appropriate interpretation of these cardiovascular parameters can be difficult and can lead to misleading conclusions regarding the functional state of the cardiac muscle. Thus, in this series of Comprehensive Physiology, the basic properties of cardiac muscle function, the cardiac cycle, and determinants of pump function will be reviewed. These basic concepts will then be integrated by presenting approaches in which the effects of preload, afterload, and myocardial contractility can be examined. Moreover, the utility of the pressure‐volume relation in terms of assessing both myocardial contractility as well as critical aspects of diastolic performance will be presented. Finally, a generalized approach for the assessment and interpretation of cardiac function within the intact cardiovascular system will be presented. © 2015 American Physiological Society. Compr Physiol 5:1911‐1946, 2015.

Comprehensive Physiology offers downloadable PowerPoint presentations of figures for non-profit, educational use, provided the content is not modified and full credit is given to the author and publication.

Download a PowerPoint presentation of all images


Figure 1. Figure 1.

(A) Electron micrograph of the myofibril arrangement in a cardiac myocyte where the fundamental contractile units, sarcomeres, are visible as the component contained within two Z‐bands. The letters on this figure refer to other histological banding patterns that are due to different composition and structural alignment of the contractile elements, the myofilaments. In this example, the myocardium was immersion fixed in situ at a LV filling pressure of 40 mmHg. This resulted in the myocardium being fixed under a filling pressure and demonstrated the sarcomere length to be approximately 2.2 to 2.4 μm. Reproduced, with permission, from (169).

(B) Sections of myocardium were perfusion fixed and then subjected to maceration digestion and scanning electron microscopy to remove cellular constituents and provide a greater relief of the fibrillar collagen matrix. The fibrillar collagen weave surrounding where individual myocyte profiles existed can be readily appreciated through this process. Moreover, the high degree of complexity of this three‐dimensional ECM network can be appreciated. Efficient transduction of sarcomere shortening into an overall muscle contraction, and ultimately LV ejection, requires significant intracellular organization of the myocyte with transmembrane receptors mediating attachment of the ECM. Reproduced, with permission, from Rossi M.A. Connective tissue skeleton in the normal left ventricle and in hypertensive left ventricular hypertrophy and chronic chagasic myocarditis. Med Sci Monit 7: 820‐832, 2001.

Figure 2. Figure 2. (A) Schematic of sarcomere ultrastructure with an emphasis on actin‐myosin interactions. The fundamental contractile unit of the cardiac myocyte is the sarcomere, and histological designations of different regions of the sarcomere were identified in Figure 1 and carried forward in this illustration. The interaction between actin and a myosin head is termed a cross‐bridge, and the shortening of the sarcomere is determined by the number of actin‐myosin cross‐bridges formed. The degree and extent of overlap between the actin filaments and the myosin heads are important determinants of cross‐bridge formation. That is, the resting sarcomere length as defined by the distance between Z‐lines is reflective of the degree of actin‐myosin overlap that exists at the onset of contraction. Thus, increasing the distance between Z‐lines, or increasing resting sarcomere length, will reduce the extent of this overlap and thus increase the number of cross‐bridges that can form with the onset of contraction. Reproduced, with permission, from Walker JW. Kinetics of the actin‐myosin interaction. Supplement 6: Handbook of Physiology, The Cardiovascular System, The Heart. 2002, p. 241. (B) The effects of cardiac muscle tension with different resting sarcomere lengths as determined by laser‐diffraction microscopy (open and closed circles). With an increase in initial sarcomere length, there is an increase in cardiac muscle tension development. This demonstrates the concept of preload at the level of the fundamental contractile unit, the sarcomere. The other important concept demonstrated in these measurements was the effects of increasing Ca2+ concentrations in these cardiac muscle preparations. Since these were “skinned” preparations, that is, the sarcolemma was removed, then this reflects absolute changes in Ca2+ concentrations at the level of the myofilaments. The increase in Ca2+ at the level of the myofilaments caused an increase in muscle tension development, which is the fundamental definition of increased contractility. As can be seen with the addition of preload, or an increase in sarcomere length, the entire Ca2+‐muscle tension curve shifts upward and to the left. This demonstrates the additive and independent effects of preload. Adapted, with permission, from (75).
Figure 3. Figure 3. (A) A schematic demonstrating the relationship between resting sarcomere length and tension development. As resting sarcomere length is increased, the resting tension on the cardiac muscle is increased (lower curve) and with subsequent contraction, the total tension or force, is increased (top curve). An example of this relationship is shown when moving from a resting sarcomere length of 1.9 to 2.2 μm (moving from point A to point B). Note that at longer sarcomere lengths, more than ∼2.3 μm, there is no net gain in muscle performance and at longer resting lengths can actually decline. (B) A schematic relationship of muscle force development as a function of resting muscle tension or force. A similar curvilinear relationship can be observed as that of resting sarcomere lengths. In the cardiac‐muscle function relationship, the resting length that achieves maximal force development, or at the point that this relationship hits a plateau (point B), is defined as maximal muscle length: L max. It is important to note that at muscle lengths beyond L max, force development does not increase, and with excessive resting tension, can actually decrease. These relationships emphasize at the level of the cardiac muscle the relationship of preload to force development. That is, over a very specific range of preloads, an almost linear relationship occurs with respect to force development, which at higher preloads plateaus. This curvilinear relationship will persist when examining the effects of preload on LV stroke volume in the intact cardiac preparation.
Figure 4. Figure 4. (A) Schematic of the effects of changes in resistive load (afterload) placed upon a contracting cardiac muscle preparation. With increased afterload, there is a fall in the velocity of muscle shortening which approximates a negative exponential relation. Extrapolation of this relationship to the y‐axis intercept yields a velocity of cardiac muscle shortening under a theoretical no afterload condition: V max. At the level of the cardiac muscle, V max is reflective of the intrinsic capacity of the muscle to contract, that is, contractility. (B) The effect of preloading a cardiac muscle preparation with changes in afterload reflects a shift in this relationship upward and to the right. Thus, at any given afterload, increased preload will yield a higher velocity of muscle shortening. This is because the resting sarcomere length has been increased and hence an improved mechanical advantage for cross‐bridge formation at the onset of contraction. The increase in preload, however, does not change the intrinsic contractile state of the cardiac muscle itself as evidenced by no change in V max. (C) The effect of increasing the inotropic state of the cardiac muscle by exposure to the beta‐receptor agonist norepinephrine. In this instance, the velocity of cardiac muscle shortening is increased at a given afterload, much like what was observed with increasing preload. However, there is a robust shift up and to the right of the exponential curve, resulting in an increased y‐intercept (V max). The increased V max is reflective that the intrinsic contractile state of the cardiac muscle has increased. (D) The comparative effects of both preload and inotropic state on cardiac muscle velocity of shortening, while both interventions will increase velocity of shortening at a given afterload, only an increase in inotropy will cause an increase in V max.
Figure 5. Figure 5. An idealized cardiac cycle that presents changes in aortic pressure, LV pressure and volumes, and atrial pressure with respect to the ECG. Moving from left to right, the P wave of the ECG reflects atrial depolarization, which is quickly followed by atrial contraction as reflected by the “a” wave on the atrial trace (a similar waveform can be detected within the venous system of large veins). This atrial contraction will result in final LV filling and thus maximal LV volumes: end‐diastolic volume. This will be followed by closure of the mitral valve (heart sound 1) and the R wave, indicative of LV depolarization. There is a temporal lag between the R wave and the beginning of LV pressure development as significant myocardial depolarization must occur and the process of excitation contraction coupling to ensue. Thus, it is common to utilize the R wave of the ECG not as an index of LV systole but actually as the indication of the end of LV diastole. Thus, the LV volume that coincides with the R wave is by convention considered the LV end‐diastolic volume. LV pressure development against the closed mitral and aortic valve is the isovolumetric phase of systole as there is no change in LV volume. The maximal rate of rise of LV pressure, termed peak +dP/dt, is determined during this phase of systole (see also Fig. 6). During LV isovolumetric contraction, LV pressure development will cause bulging of the mitral valve, which will be a reflected wave on the atrial trace, the “c” wave. When LV pressure exceeds that of the aortic diastolic pressure, the aortic valve opens and the ejection phase of systole starts. During this period, blood is continuously returning to the atrium with a closed mitral valve and will result in a progressive increase in atrial pressure. With the opening of the aortic valve, and under normal conditions, the LV and aortic pressure become superimposable. With the end of ejection and the beginning of LV pressure decline, the aortic valve closes and causes a reflective pressure wave, the dicrotic notch. This is associated with heart sound 2, and the dicrotic notch is a commonly used arterial waveform for identifying and synchronizing cardiac events. Specifically, the LV volume at the dicrotic notch is conventionally called the LV end‐systolic volume. The change in LV volume from end‐diastole to end‐systole reflects the amount of blood ejected from the LV, the stroke volume. It is also notable that it is during the LV ejection phase that myocardial repolarization occurs, designated by the T wave on the ECG, which further demonstrates the temporal differences in myocardial electrical events to mechanical events. With the closure of the aortic valve and the mitral valve remaining closed, the LV enters the first phase of diastole, the isovolumetric phase. This is considered the active relaxation phase of diastole as this is when Ca2+ is removed from the myofilaments and taken back up by the SR or moved across the sarcolemma—both energy‐dependent processes. Thus, defects in active relaxation will be reflected as a prolonged rate of LV pressure decline (see Fig. 6). As LV pressure declines, there has been a concomitant rise in left atrial pressure, creating an LV‐atrial pressure gradient and ultimately opening of the mitral valve. Just at the opening of the mitral valve will be the crescendo of the atrial pressure, which then rapidly falls with LV filling, the “v” wave. The initial opening of the mitral valve will result in significant and rapid LV filling due to the LV‐atrial pressure gradient and the continuous relaxation of the LV myocardium, which can create “suction.” As the LV and atrial pressures equilibrate, atrial contraction occurs and the cycle begins again. Adapted, with permission, from (93).
Figure 6. Figure 6. (Inset) A stylized tracing of Ca2+ transient and cardiac myocyte shortening under normal conditions and with abnormal conditions, such as LV failure due to defects in SR function (53,57). A prolongation of Ca2+ resequestration by the SR will result in a prolongation of the cardiac myocyte to return to resting length. This will be translated at the level of the LV as a prolongation of the active phase of relaxation, as depicted in the LV pressure curves shown. The time points by which peak +dP/dt (isovolumetric contraction) and peak –dP/dt (isovolumetric relaxation) is determined from mathematical differentiation of the LV pressure values. The relative rate of LV pressure decline, –dP/dt, is prolonged with abnormalities in active relaxation. The log transform of this –dP/dt signal results in a linear decay relationship, and the slope of this linear decay is termed the time constant of isovolumetric relaxation, tau. Defects in active relaxation therefore can be detected by changes in tau. For example, diminished rates of Ca2+ will result in a prolonged tau and can be used as in index for identifying potential defects in energy dependent processes involved in Ca2+ removal from the myofilament apparatus. Adapted, with permission, from (201).
Figure 7. Figure 7. (A) Scanning electron micrographs revealing the fibrillar collagen weave of the ECM in LV myocardial samples taken from normal rodents and following the development of pressure overload hypertrophy (POH). The fibrillar collagen weave of the ECM is increased with hypertrophy and has been termed myocardial “fibrosis.” This is a common structural change within the myocardium with a prolonged LV pressure overload due to either hypertension or aortic stenosis. Parallel cardiac muscle samples were treated with the serine protease, plasmin, which is a known activator of endogenous matrix metalloproteinases, a fundamental proteolytic pathway for collagen degradation (78). Plasmin treatment reduced fibrillar collagen surrounding individual myocytes in both normal and hypertrophy myocardium. (B) One of the common approaches to measure intrinsic myocardial stiffness is through developing a strain‐stress relationship. This is usually an exponential relationship, and coefficient of this relationship, beta is reflective of changes in myocardial stiffness properties and is often termed the myocardial stiffness constant. Representative strain‐stress relationships are shown for an untreated normal cardiac muscle and with plasmin treatment. A shift downward and to the right of the exponential curve, indicative of a reduced slope or beta, is reflective of reduced passive myocardial stiffness. (C) Summary results for the myocardial stiffness constant, beta, from normal and POH cardiac muscle samples with and without plasmin treatment. Activation of endogenous ECMC proteolytic enzymes, which reduced collagen content, caused a reduction in myocardial stiffness in both normal and POH cardiac muscle. These studies underscore the importance of ECM content and composition within the LV myocardium in terms of passive stiffness properties and hence passive LV filling during diastole. Adapted, with permission, from (174).
Figure 8. Figure 8. (Top) A schematic of left atrial and LV pressures that have been amplified to demonstrate the relation between atrial contraction and LV pressure at the end of diastole. Atrial contraction will propel blood across the mitral valve and cause an increase in LV end‐diastolic volume and pressure. This is characterized by an increase in the rate of LV diastolic pressure developed just before mitral valve closure and the onset of LV isovolumetric contraction. This is termed the “atrial kick” and under normal conditions and ambient heart rates, this may only contribute a minor proportion of the total LV end‐diastolic volume. However, as the duration of passive filling is decreased, such as with exercise or tachycardia, then the contribution of the atrial contraction becomes much more important. This is also true when there is an increase in passive myocardial stiffness properties (see Fig. 7), which will impair the rate and magnitude of passive LV filling in early diastole. Indeed, the contribution of the “atrial kick” becomes much more important in conditions such as with aging or hypertrophy where passive LV filling is impaired. (Bottom) Schematic illustrations of LV filling under normal conditions, with LV hypertrophy (LVH) and increased myocardial stiffness, and with LV systolic failure resulting in increased LV volumes, notably increased end‐diastolic volume. The arrows indicate the direction and magnitude of blood flow from the atrium to the LV during diastole. The bottom panels are representative color Doppler velocity profiles of blood flow through the mitral annulus, where the relative velocity of flow has been indicated by a white line. In normal conditions, rapid filling as shown by the steep slope of the velocity profile occurs as a function of active relaxation and normal myocardial stiffness. With LVH and increased myocardial stiffness, the velocity of flow during passive filling is slowed and can be seen by the fall in the slope of the flow velocity profile. Similarly in LV failure, notably systolic heart failure, there are high LV residual volumes at the end of ejection and thus higher intracavitary pressure. As a consequence, passive LV filling will be impaired as shown by a reduced slope of the velocity flow profile. While the causes for the impaired LV passive filling are distinctly different in these disease states, both result in a much greater reliance on atrial contraction and the “atrial kick.” Bottom panels are reproduced, with permission, from (106).
Figure 9. Figure 9. (A) The drawing from the work by Starling illustrating how LV preload was controlled by changing venous inflow volumes and then measuring the ejected LV stroke volume. The designated abbreviations have been defined as they were in the original report. This was landmark work as it began to define the basic relationship of increasing LV loading conditions prior to ejection‐preload, whereby LV afterload was held constant. (B) The stylized Frank‐Starling Law of the Heart, or the LV preload‐stroke work relation. In normal subjects, LV preload is maintained within the region that is almost linear. As such, small changes in LV preload (arrow) will result in significant increases in LV stroke volume. It should be noted that this relationship plateaus at higher LV preloads. Thus, excessive volume loading of the LV will not yield any further stroke volume and can be deleterious. Panel A reproduced, with permission, from (171).
Figure 10. Figure 10. Utilization and interpretation of the Frank‐Starling Law of the Heart. (A) The LV preload‐stroke volume relationship was determined in conscious dogs before (control) and after the development of pacing induced LV failure. Changes in LV preload, as measured by LV end‐diastolic volume, were achieved by acute volume loading. This allowed for plotting the LV preload‐stroke volume relation. The first important observation is that under these physiological loading conditions, this relation is linear. The second observation is that under ambient conditions prior to changes in LV load, there is a shift downward and to the right (arrow A). The third observation is that with an equivalent LV preload (equivalent LV end‐diastolic volume), LV stroke volume is greatly reduced with LV failure (arrow B). (B) Idealized LV preload‐stroke volume relationships have been plotted for the purposes of interpretation with key events. First, arrow A indicates a movement downward in this relation, whereby at an identical LV preload, LV stroke volume is reduced. This is commonly caused by a reduction in intrinsic myocardial contractility, which often accompanies chronic systolic heart failure. Second, arrow B indicates a shift upward in this relation, whereby at an equivalent preload, LV stroke volume is increased. This can be interpreted as an increase in myocardial contractility, most commonly achieved through increased inotropic state. Third, a shift downward and to the left on the LV preload‐stroke volume relationship implies that a reduction in LV preload has occurred, not a change in underlying myocardial contractility. This relation emphasizes the importance of considering the underlying ambient LV preload conditions when evaluating changes in LV stroke volume. Panel A reproduced, with permission, from (94).
Figure 11. Figure 11. Recordings obtained by Starling demonstrating in the isolated heart preparation that acute changes in aortic pressure (afterload) caused a direct change in LV stroke volume. In this example, beat by beat LV stroke volume was measured as the volume of blood ejected into a calibrated column, and cardiometer (“C”), aortic pressure (“BP”), and central venous pressure (“VP”) were measured by mercury columns. As can be seen, a step‐wise increase in aortic pressure caused a proportional decline in LV stroke volume, hence demonstrating the acute effects of increased LV afterload on LV ejection performance. Adapted, with permission, from (171).
Figure 12. Figure 12. There are a number of limitations of LV ejection fraction (EF) in terms of evaluating LV pump function. An example of the effects of LV ejection fraction in the context of incompetent valves, such as mitral regurgitation (MR), is exemplified here. Under normal conditions, the entire LV stroke volume is ejected through the aortic valve (green arrow) and hence there is no flow of ejected blood into the left atrium (LA) during contraction, and as such, there is no fraction of ejected blood being regurgitated into the LA (regurgitant fraction (RF) = 0%). In this example, 65 mL of blood is ejected and is the LV stroke volume, and with an LV end‐diastolic volume of 100 mL, the LV EF = 50%. However with MR, a proportion of blood is ejected into the LA, and in this example, the RF is 50%. With MR, LV end‐diastolic volume and total stroke volume commonly increase, and in this example, increased to 120 and 100 mL, respectively, with a resultant EF computation of 83%. Thus, a “supra‐normal” EF computation can occur with MR, which fails to reflect any underlying change in myocardial function/contractility. Figure adapted, with permission, from (59).
Figure 13. Figure 13. (A) Length‐tension curves as a function of stimulation frequency, whereby the rate of tension development (dT/dt) was measured at increasing muscle lengths and stimulation frequencies. Consistent with the concept of the muscle preload/force relationship (see Fig. 2 and 3), an increase in resting muscle length increased muscle tension development. With increased stimulation frequency, the cardiac muscle length‐tension curve shifted upward and to the left indicative of an increase in inotropy, an intrinsic change in Ca2+ dynamics. (B) The rate of LV pressure development (dP/dt) measured at an equivalent LV developed systolic pressure of 40 mmHg (DP 40) was used as an index of force development in the intact LV. This index of LV pressure development was plotted as a function of heart rate in normal pigs and in pigs following the development of pacing induced heart failure. Measurements were performed with infusion of the beta‐adrenergic agonist, dobutamine. In the normal condition, LV pressure development increased as a function of rate but was significantly blunted with the development of LV failure. Panel A adapted, with permission, from (100); Panel B adapted, with permission, from (46).
Figure 14. Figure 14. Through altering LV preload, in this case LV end‐diastolic volume by transient inferior cava occlusion, a series of LV end‐diastolic‐stroke work values can be obtained for a linear relation—the preload recruitable stroke work relation. In this example, the LV preload recruitable stroke work relation was computed under control, ambient conditions, and then determined again following infusion of calcium chloride. A shift upward and to the left of this linear relationship is reflective of a change in intrinsic myocardial contractility, consistent with the effects of increasing Ca2+ availability to the myofilaments with contraction. Adapted, with permission, from (62).
Figure 15. Figure 15. (A) The LV end‐systolic wall stress‐fractional shortening relationship was determined in normal control pigs by infusion of the alpha agonist phenylephrine. A negative linear relationship was observed between increased LV end‐systolic wall stress and fractional shortening, and the 95% confidence interval obtained from these referent normal preparations is shown. Since increased arterial pressure by phenylephrine infusion will cause a reduction in heart rate (via baroreceptor reflex), a constant heart rate was achieved by electrical pacing. With the development of LV failure secondary to chronic rapid pacing, the isochronal LV end‐systolic stress‐fractional shortening values were outside of the normal referent range and fell to the right, suggestive of impaired myocardial contractility. With recovery from this form of LV failure, LV end‐systolic stress‐fractional shortening values returned to within referent normal limits. (B) Schematic interpretation of the LV end‐systolic stress‐fractional shortening relation. As LV afterload is increased, as shown in this example as an increase in LV end‐systolic walls stress (moving from point A to point B), then a fall in LV fractional shortening will occur. This does not imply that a change in LV myocardial contractility has occurred but simply a change in LV afterload. However, if LV fractional shortening is reduced under an equivalent afterload, as shown from the movement of the LV end‐systolic wall stress‐shortening value from point A to point C, then this would be indicative of a fall in myocardial contractility. Conversely, a shift upward in this relation, as shown as movement from point A to point D would be suggestive of increased myocardial contractility. Panel A adapted, with permission, from (183).
Figure 16. Figure 16. A stylized LV pressure‐volume loop that identifies key points in this relation in terms of the cardiac cycle. Point A reflects the end of diastole, mitral valve closure, and the onset of isovolumetric contraction. Point B reflects the point whereby LV pressure development just exceeds that necessary for aortic valve opening and the onset of ejection. Point C reflects the end of systole and aortic valve closure. This point, the end‐systolic pressure‐volume point, is used to develop the end‐systolic pressure volume relation. Point C also reflects the onset of isovolumetric relaxation (active relaxation). Point D reflects the end of the isovolumetric relaxation phase, opening of the mitral valve, and the onset of the filling phase of diastole. This phase of diastole can be used to develop the diastolic pressure‐volume relation. Both the end‐systolic and end‐diastolic pressure‐volume relationships are determined by plotting values with acute changes in LV loading conditions to develop a family of LV pressure‐volume loops and are shown here for illustrative purposed only.
Figure 17. Figure 17. The effects of acute changes in LV preload or afterload on the pressure‐volume relation. Left: With an acute infusion of a volume of physiological solution, an increase in LV preload occurs as reflected as a shift to the right in LV end‐diastolic volume. The increased LV preload will result in a greater stroke volume, hence a larger area within the pressure‐volume loops. Right: An acute increase in LV afterload, which will result in increased LV pressure development, will also cause a shift to the right of the pressure‐volume relation. In this case, an initial increase in LV afterload will cause a reduction in LV stroke volume for that contraction, which will then be translated into a higher LV end‐diastolic volume at the subsequent diastolic phase. As a result, a higher LV preload will occur with the increase in LV afterload, which in turn will result in a normalization of LV stroke volume. These beat‐beat adjustments are an example of the interaction between LV afterload and preload to maintain stroke volume. However, the pressure‐volume area is increased significantly, which means that while stroke volume may be normalized, this is being achieved under much greater work load and with greater myocardial oxygen demand/energy requirements. Figure adapted, with permission, from (175).
Figure 18. Figure 18. Example of a shift in the end‐systolic pressure volume relation following stimulation of the beta adrenergic system by epinephrine in the beating heart preparation as described by Suga and Sagawa (177). The direction and key points of the LV pressure volume loop are shown by the arrows and letters, respectively. With infusion of the positive inotrope epinephrine, a shift to the left in the slope of this relation was demonstrated, indicative of increased myocardial contractility. Adapted, with permission, from (177).
Figure 19. Figure 19. Adult pigs were instrumented with high fidelity, MRI compatible LV microtransducers to measure LV pressure (panel A) and simultaneously LV volumes by MRI (panel B). A series of LV pressure‐volume loops were obtained by a transient reduction in LV preload (vena cava occlusion) under normal resting conditions (rest) and following a dobutamine infusion, thus causing increased beta adrenergic receptor stimulation (stress). Representative LV pressure volume loops are shown in panel C, whereby a shift in the end‐systolic pressure volume relation (ESPVR) was observed with dobutamine infusion, consistent with an increase in inotropic state and hence contractile function. The lower blue points indicate LV diastole and the end‐diastolic pressure volume relation (EDPVR). A cross‐section of the heart by MRI (white line) is shown in panel D, whereby the dimensions across this single plane (“m‐mode”) is shown in panel E, demonstrating the high fidelity of measuring changes in LV volume with transient occlusion of the vena cava. Reproduced, with permission, from (196).
Figure 20. Figure 20. (A) Schematic of the lower portion of the LV pressure‐volume loop whereby the passive filling phase of diastole has been amplified. A curve‐linear relation exists between LV diastolic pressure and volume, whereby a shift upward and to the right demonstrates that an increase in relative LV chamber stiffness occurred. In this example, arrows indicate that at a specific LV diastolic volume, a much higher diastolic pressure is obtained with increased LV chamber stiffness. However, it should be emphasized that using a single measurement of LV diastolic volume and pressure will not allow for interpretation of LV chamber stiffness but must be determined from a set of points. (B) Using a curve fitting algorithm, the relative slope of the change in LV diastolic pressure and volume (dP/dV) is computed and has been termed the LV chamber stiffness modulus, or Kc. However, the dP/dV relation can change as a function of operating volumes and pressures as shown moving from point A to point B. In this example, computing Kc under filling volumes and pressures will yield different values. Thus, when comparing Kc across subjects or across time, it is important to compute this index of LV chamber stiffness under equivalent diastolic volume and pressures as shown moving from point A to point C. In this example, an absolute increase in LV chamber stiffness has occurred. Figures adapted, with permission, from (201).
Figure 21. Figure 21. A number of factors can influence the LV diastolic pressure‐volume relation and thereby will influence the ability to execute the computations necessary for determining LV chamber stiffness. Abnormalities in active relaxation properties, which will therefore prolong the isovolumetric relaxation phase of diastole, can cause shifts in the LV diastolic pressure‐volume relation. This can be due to defects in Ca2+ resequestration or Ca2+ overload, diminished energy/metabolism pathways, alterations in protein structure or composition involved in active relaxation, or a combination of these factors. Another factor that can influence this relationship is an extrinsic one, which is pericardial restraint. This can be due to thickening and/or adhesions of the pericardium, which can occur with chronic inflammatory diseases for example. Another cause for a shift in this relation, and one that is taking on greater import with the high incidence of hypertensive heart disease (aka diastolic heart failure: heart failure with a preserved ejection fraction), is increased LV chamber stiffness (91,107,203,204). Finally, chronic changes in LV geometry, such as increased chamber volumes, will shift this relationship usually upward and to the right as shown. Adapted, with permission, from (29).
Figure 22. Figure 22. Three‐dimensional wire model reconstructed from bi‐lane ventriculography of the LV (red) and RV (blue). The significant differences in geometry of these ventricles can be appreciated, which also translates into differences in the mechanical form of contraction. The LV generates force by a compressive twisting motion whereas the RV is more of a bellows pump type of action. The RV operates under much lower afterload as the pulmonary circuit is a low resistance, high compliance circuit. Thus, the RV can be much more sensitive to increases in afterload, which will cause RV dilation as a compensatory mechanism. The RV dilation can affect LV filling through changes in interventricular septal motion as well as through the pericardium. Thus, LV‐RV interactions can significantly affect overall cardiac performance. Adapted, with permission, from (168).
Figure 23. Figure 23. A schematic demonstrating the relationship between an increase in different determinants of cardiac output to that of myocardial oxygen consumption. The changes have been placed in arbitrary units for the purposes of demonstrating the differences in myocardial oxygen consumption with changes in LV load, heart rate, and finally, contractility. While an equivalent cardiac output may be measured between subjects, the factors that contribute to this value may be distinctly different. Thus, myocardial efficiency, that is the energy cost for a given LV ejection or to maintain total flow (cardiac output), may be very different. An evaluation of cardiac function should not only consider LV ejection but the underlying determinants that govern ejection and flow, as this will identify potential differences in myocardial efficiency.
Figure 24. Figure 24. The assessment of cardiac function in terms of cardiac output can be considered a multivariable equation where LV load, contractility, and rate are independent variables. The measurement of these independent variables will provide for a more comprehensive view of overall cardiac performance and allow for assessment of changes in cardiac function across subjects, time, or treatment. Adapted, with permission, from (9).


Figure 1.

(A) Electron micrograph of the myofibril arrangement in a cardiac myocyte where the fundamental contractile units, sarcomeres, are visible as the component contained within two Z‐bands. The letters on this figure refer to other histological banding patterns that are due to different composition and structural alignment of the contractile elements, the myofilaments. In this example, the myocardium was immersion fixed in situ at a LV filling pressure of 40 mmHg. This resulted in the myocardium being fixed under a filling pressure and demonstrated the sarcomere length to be approximately 2.2 to 2.4 μm. Reproduced, with permission, from (169).

(B) Sections of myocardium were perfusion fixed and then subjected to maceration digestion and scanning electron microscopy to remove cellular constituents and provide a greater relief of the fibrillar collagen matrix. The fibrillar collagen weave surrounding where individual myocyte profiles existed can be readily appreciated through this process. Moreover, the high degree of complexity of this three‐dimensional ECM network can be appreciated. Efficient transduction of sarcomere shortening into an overall muscle contraction, and ultimately LV ejection, requires significant intracellular organization of the myocyte with transmembrane receptors mediating attachment of the ECM. Reproduced, with permission, from Rossi M.A. Connective tissue skeleton in the normal left ventricle and in hypertensive left ventricular hypertrophy and chronic chagasic myocarditis. Med Sci Monit 7: 820‐832, 2001.



Figure 2. (A) Schematic of sarcomere ultrastructure with an emphasis on actin‐myosin interactions. The fundamental contractile unit of the cardiac myocyte is the sarcomere, and histological designations of different regions of the sarcomere were identified in Figure 1 and carried forward in this illustration. The interaction between actin and a myosin head is termed a cross‐bridge, and the shortening of the sarcomere is determined by the number of actin‐myosin cross‐bridges formed. The degree and extent of overlap between the actin filaments and the myosin heads are important determinants of cross‐bridge formation. That is, the resting sarcomere length as defined by the distance between Z‐lines is reflective of the degree of actin‐myosin overlap that exists at the onset of contraction. Thus, increasing the distance between Z‐lines, or increasing resting sarcomere length, will reduce the extent of this overlap and thus increase the number of cross‐bridges that can form with the onset of contraction. Reproduced, with permission, from Walker JW. Kinetics of the actin‐myosin interaction. Supplement 6: Handbook of Physiology, The Cardiovascular System, The Heart. 2002, p. 241. (B) The effects of cardiac muscle tension with different resting sarcomere lengths as determined by laser‐diffraction microscopy (open and closed circles). With an increase in initial sarcomere length, there is an increase in cardiac muscle tension development. This demonstrates the concept of preload at the level of the fundamental contractile unit, the sarcomere. The other important concept demonstrated in these measurements was the effects of increasing Ca2+ concentrations in these cardiac muscle preparations. Since these were “skinned” preparations, that is, the sarcolemma was removed, then this reflects absolute changes in Ca2+ concentrations at the level of the myofilaments. The increase in Ca2+ at the level of the myofilaments caused an increase in muscle tension development, which is the fundamental definition of increased contractility. As can be seen with the addition of preload, or an increase in sarcomere length, the entire Ca2+‐muscle tension curve shifts upward and to the left. This demonstrates the additive and independent effects of preload. Adapted, with permission, from (75).


Figure 3. (A) A schematic demonstrating the relationship between resting sarcomere length and tension development. As resting sarcomere length is increased, the resting tension on the cardiac muscle is increased (lower curve) and with subsequent contraction, the total tension or force, is increased (top curve). An example of this relationship is shown when moving from a resting sarcomere length of 1.9 to 2.2 μm (moving from point A to point B). Note that at longer sarcomere lengths, more than ∼2.3 μm, there is no net gain in muscle performance and at longer resting lengths can actually decline. (B) A schematic relationship of muscle force development as a function of resting muscle tension or force. A similar curvilinear relationship can be observed as that of resting sarcomere lengths. In the cardiac‐muscle function relationship, the resting length that achieves maximal force development, or at the point that this relationship hits a plateau (point B), is defined as maximal muscle length: L max. It is important to note that at muscle lengths beyond L max, force development does not increase, and with excessive resting tension, can actually decrease. These relationships emphasize at the level of the cardiac muscle the relationship of preload to force development. That is, over a very specific range of preloads, an almost linear relationship occurs with respect to force development, which at higher preloads plateaus. This curvilinear relationship will persist when examining the effects of preload on LV stroke volume in the intact cardiac preparation.


Figure 4. (A) Schematic of the effects of changes in resistive load (afterload) placed upon a contracting cardiac muscle preparation. With increased afterload, there is a fall in the velocity of muscle shortening which approximates a negative exponential relation. Extrapolation of this relationship to the y‐axis intercept yields a velocity of cardiac muscle shortening under a theoretical no afterload condition: V max. At the level of the cardiac muscle, V max is reflective of the intrinsic capacity of the muscle to contract, that is, contractility. (B) The effect of preloading a cardiac muscle preparation with changes in afterload reflects a shift in this relationship upward and to the right. Thus, at any given afterload, increased preload will yield a higher velocity of muscle shortening. This is because the resting sarcomere length has been increased and hence an improved mechanical advantage for cross‐bridge formation at the onset of contraction. The increase in preload, however, does not change the intrinsic contractile state of the cardiac muscle itself as evidenced by no change in V max. (C) The effect of increasing the inotropic state of the cardiac muscle by exposure to the beta‐receptor agonist norepinephrine. In this instance, the velocity of cardiac muscle shortening is increased at a given afterload, much like what was observed with increasing preload. However, there is a robust shift up and to the right of the exponential curve, resulting in an increased y‐intercept (V max). The increased V max is reflective that the intrinsic contractile state of the cardiac muscle has increased. (D) The comparative effects of both preload and inotropic state on cardiac muscle velocity of shortening, while both interventions will increase velocity of shortening at a given afterload, only an increase in inotropy will cause an increase in V max.


Figure 5. An idealized cardiac cycle that presents changes in aortic pressure, LV pressure and volumes, and atrial pressure with respect to the ECG. Moving from left to right, the P wave of the ECG reflects atrial depolarization, which is quickly followed by atrial contraction as reflected by the “a” wave on the atrial trace (a similar waveform can be detected within the venous system of large veins). This atrial contraction will result in final LV filling and thus maximal LV volumes: end‐diastolic volume. This will be followed by closure of the mitral valve (heart sound 1) and the R wave, indicative of LV depolarization. There is a temporal lag between the R wave and the beginning of LV pressure development as significant myocardial depolarization must occur and the process of excitation contraction coupling to ensue. Thus, it is common to utilize the R wave of the ECG not as an index of LV systole but actually as the indication of the end of LV diastole. Thus, the LV volume that coincides with the R wave is by convention considered the LV end‐diastolic volume. LV pressure development against the closed mitral and aortic valve is the isovolumetric phase of systole as there is no change in LV volume. The maximal rate of rise of LV pressure, termed peak +dP/dt, is determined during this phase of systole (see also Fig. 6). During LV isovolumetric contraction, LV pressure development will cause bulging of the mitral valve, which will be a reflected wave on the atrial trace, the “c” wave. When LV pressure exceeds that of the aortic diastolic pressure, the aortic valve opens and the ejection phase of systole starts. During this period, blood is continuously returning to the atrium with a closed mitral valve and will result in a progressive increase in atrial pressure. With the opening of the aortic valve, and under normal conditions, the LV and aortic pressure become superimposable. With the end of ejection and the beginning of LV pressure decline, the aortic valve closes and causes a reflective pressure wave, the dicrotic notch. This is associated with heart sound 2, and the dicrotic notch is a commonly used arterial waveform for identifying and synchronizing cardiac events. Specifically, the LV volume at the dicrotic notch is conventionally called the LV end‐systolic volume. The change in LV volume from end‐diastole to end‐systole reflects the amount of blood ejected from the LV, the stroke volume. It is also notable that it is during the LV ejection phase that myocardial repolarization occurs, designated by the T wave on the ECG, which further demonstrates the temporal differences in myocardial electrical events to mechanical events. With the closure of the aortic valve and the mitral valve remaining closed, the LV enters the first phase of diastole, the isovolumetric phase. This is considered the active relaxation phase of diastole as this is when Ca2+ is removed from the myofilaments and taken back up by the SR or moved across the sarcolemma—both energy‐dependent processes. Thus, defects in active relaxation will be reflected as a prolonged rate of LV pressure decline (see Fig. 6). As LV pressure declines, there has been a concomitant rise in left atrial pressure, creating an LV‐atrial pressure gradient and ultimately opening of the mitral valve. Just at the opening of the mitral valve will be the crescendo of the atrial pressure, which then rapidly falls with LV filling, the “v” wave. The initial opening of the mitral valve will result in significant and rapid LV filling due to the LV‐atrial pressure gradient and the continuous relaxation of the LV myocardium, which can create “suction.” As the LV and atrial pressures equilibrate, atrial contraction occurs and the cycle begins again. Adapted, with permission, from (93).


Figure 6. (Inset) A stylized tracing of Ca2+ transient and cardiac myocyte shortening under normal conditions and with abnormal conditions, such as LV failure due to defects in SR function (53,57). A prolongation of Ca2+ resequestration by the SR will result in a prolongation of the cardiac myocyte to return to resting length. This will be translated at the level of the LV as a prolongation of the active phase of relaxation, as depicted in the LV pressure curves shown. The time points by which peak +dP/dt (isovolumetric contraction) and peak –dP/dt (isovolumetric relaxation) is determined from mathematical differentiation of the LV pressure values. The relative rate of LV pressure decline, –dP/dt, is prolonged with abnormalities in active relaxation. The log transform of this –dP/dt signal results in a linear decay relationship, and the slope of this linear decay is termed the time constant of isovolumetric relaxation, tau. Defects in active relaxation therefore can be detected by changes in tau. For example, diminished rates of Ca2+ will result in a prolonged tau and can be used as in index for identifying potential defects in energy dependent processes involved in Ca2+ removal from the myofilament apparatus. Adapted, with permission, from (201).


Figure 7. (A) Scanning electron micrographs revealing the fibrillar collagen weave of the ECM in LV myocardial samples taken from normal rodents and following the development of pressure overload hypertrophy (POH). The fibrillar collagen weave of the ECM is increased with hypertrophy and has been termed myocardial “fibrosis.” This is a common structural change within the myocardium with a prolonged LV pressure overload due to either hypertension or aortic stenosis. Parallel cardiac muscle samples were treated with the serine protease, plasmin, which is a known activator of endogenous matrix metalloproteinases, a fundamental proteolytic pathway for collagen degradation (78). Plasmin treatment reduced fibrillar collagen surrounding individual myocytes in both normal and hypertrophy myocardium. (B) One of the common approaches to measure intrinsic myocardial stiffness is through developing a strain‐stress relationship. This is usually an exponential relationship, and coefficient of this relationship, beta is reflective of changes in myocardial stiffness properties and is often termed the myocardial stiffness constant. Representative strain‐stress relationships are shown for an untreated normal cardiac muscle and with plasmin treatment. A shift downward and to the right of the exponential curve, indicative of a reduced slope or beta, is reflective of reduced passive myocardial stiffness. (C) Summary results for the myocardial stiffness constant, beta, from normal and POH cardiac muscle samples with and without plasmin treatment. Activation of endogenous ECMC proteolytic enzymes, which reduced collagen content, caused a reduction in myocardial stiffness in both normal and POH cardiac muscle. These studies underscore the importance of ECM content and composition within the LV myocardium in terms of passive stiffness properties and hence passive LV filling during diastole. Adapted, with permission, from (174).


Figure 8. (Top) A schematic of left atrial and LV pressures that have been amplified to demonstrate the relation between atrial contraction and LV pressure at the end of diastole. Atrial contraction will propel blood across the mitral valve and cause an increase in LV end‐diastolic volume and pressure. This is characterized by an increase in the rate of LV diastolic pressure developed just before mitral valve closure and the onset of LV isovolumetric contraction. This is termed the “atrial kick” and under normal conditions and ambient heart rates, this may only contribute a minor proportion of the total LV end‐diastolic volume. However, as the duration of passive filling is decreased, such as with exercise or tachycardia, then the contribution of the atrial contraction becomes much more important. This is also true when there is an increase in passive myocardial stiffness properties (see Fig. 7), which will impair the rate and magnitude of passive LV filling in early diastole. Indeed, the contribution of the “atrial kick” becomes much more important in conditions such as with aging or hypertrophy where passive LV filling is impaired. (Bottom) Schematic illustrations of LV filling under normal conditions, with LV hypertrophy (LVH) and increased myocardial stiffness, and with LV systolic failure resulting in increased LV volumes, notably increased end‐diastolic volume. The arrows indicate the direction and magnitude of blood flow from the atrium to the LV during diastole. The bottom panels are representative color Doppler velocity profiles of blood flow through the mitral annulus, where the relative velocity of flow has been indicated by a white line. In normal conditions, rapid filling as shown by the steep slope of the velocity profile occurs as a function of active relaxation and normal myocardial stiffness. With LVH and increased myocardial stiffness, the velocity of flow during passive filling is slowed and can be seen by the fall in the slope of the flow velocity profile. Similarly in LV failure, notably systolic heart failure, there are high LV residual volumes at the end of ejection and thus higher intracavitary pressure. As a consequence, passive LV filling will be impaired as shown by a reduced slope of the velocity flow profile. While the causes for the impaired LV passive filling are distinctly different in these disease states, both result in a much greater reliance on atrial contraction and the “atrial kick.” Bottom panels are reproduced, with permission, from (106).


Figure 9. (A) The drawing from the work by Starling illustrating how LV preload was controlled by changing venous inflow volumes and then measuring the ejected LV stroke volume. The designated abbreviations have been defined as they were in the original report. This was landmark work as it began to define the basic relationship of increasing LV loading conditions prior to ejection‐preload, whereby LV afterload was held constant. (B) The stylized Frank‐Starling Law of the Heart, or the LV preload‐stroke work relation. In normal subjects, LV preload is maintained within the region that is almost linear. As such, small changes in LV preload (arrow) will result in significant increases in LV stroke volume. It should be noted that this relationship plateaus at higher LV preloads. Thus, excessive volume loading of the LV will not yield any further stroke volume and can be deleterious. Panel A reproduced, with permission, from (171).


Figure 10. Utilization and interpretation of the Frank‐Starling Law of the Heart. (A) The LV preload‐stroke volume relationship was determined in conscious dogs before (control) and after the development of pacing induced LV failure. Changes in LV preload, as measured by LV end‐diastolic volume, were achieved by acute volume loading. This allowed for plotting the LV preload‐stroke volume relation. The first important observation is that under these physiological loading conditions, this relation is linear. The second observation is that under ambient conditions prior to changes in LV load, there is a shift downward and to the right (arrow A). The third observation is that with an equivalent LV preload (equivalent LV end‐diastolic volume), LV stroke volume is greatly reduced with LV failure (arrow B). (B) Idealized LV preload‐stroke volume relationships have been plotted for the purposes of interpretation with key events. First, arrow A indicates a movement downward in this relation, whereby at an identical LV preload, LV stroke volume is reduced. This is commonly caused by a reduction in intrinsic myocardial contractility, which often accompanies chronic systolic heart failure. Second, arrow B indicates a shift upward in this relation, whereby at an equivalent preload, LV stroke volume is increased. This can be interpreted as an increase in myocardial contractility, most commonly achieved through increased inotropic state. Third, a shift downward and to the left on the LV preload‐stroke volume relationship implies that a reduction in LV preload has occurred, not a change in underlying myocardial contractility. This relation emphasizes the importance of considering the underlying ambient LV preload conditions when evaluating changes in LV stroke volume. Panel A reproduced, with permission, from (94).


Figure 11. Recordings obtained by Starling demonstrating in the isolated heart preparation that acute changes in aortic pressure (afterload) caused a direct change in LV stroke volume. In this example, beat by beat LV stroke volume was measured as the volume of blood ejected into a calibrated column, and cardiometer (“C”), aortic pressure (“BP”), and central venous pressure (“VP”) were measured by mercury columns. As can be seen, a step‐wise increase in aortic pressure caused a proportional decline in LV stroke volume, hence demonstrating the acute effects of increased LV afterload on LV ejection performance. Adapted, with permission, from (171).


Figure 12. There are a number of limitations of LV ejection fraction (EF) in terms of evaluating LV pump function. An example of the effects of LV ejection fraction in the context of incompetent valves, such as mitral regurgitation (MR), is exemplified here. Under normal conditions, the entire LV stroke volume is ejected through the aortic valve (green arrow) and hence there is no flow of ejected blood into the left atrium (LA) during contraction, and as such, there is no fraction of ejected blood being regurgitated into the LA (regurgitant fraction (RF) = 0%). In this example, 65 mL of blood is ejected and is the LV stroke volume, and with an LV end‐diastolic volume of 100 mL, the LV EF = 50%. However with MR, a proportion of blood is ejected into the LA, and in this example, the RF is 50%. With MR, LV end‐diastolic volume and total stroke volume commonly increase, and in this example, increased to 120 and 100 mL, respectively, with a resultant EF computation of 83%. Thus, a “supra‐normal” EF computation can occur with MR, which fails to reflect any underlying change in myocardial function/contractility. Figure adapted, with permission, from (59).


Figure 13. (A) Length‐tension curves as a function of stimulation frequency, whereby the rate of tension development (dT/dt) was measured at increasing muscle lengths and stimulation frequencies. Consistent with the concept of the muscle preload/force relationship (see Fig. 2 and 3), an increase in resting muscle length increased muscle tension development. With increased stimulation frequency, the cardiac muscle length‐tension curve shifted upward and to the left indicative of an increase in inotropy, an intrinsic change in Ca2+ dynamics. (B) The rate of LV pressure development (dP/dt) measured at an equivalent LV developed systolic pressure of 40 mmHg (DP 40) was used as an index of force development in the intact LV. This index of LV pressure development was plotted as a function of heart rate in normal pigs and in pigs following the development of pacing induced heart failure. Measurements were performed with infusion of the beta‐adrenergic agonist, dobutamine. In the normal condition, LV pressure development increased as a function of rate but was significantly blunted with the development of LV failure. Panel A adapted, with permission, from (100); Panel B adapted, with permission, from (46).


Figure 14. Through altering LV preload, in this case LV end‐diastolic volume by transient inferior cava occlusion, a series of LV end‐diastolic‐stroke work values can be obtained for a linear relation—the preload recruitable stroke work relation. In this example, the LV preload recruitable stroke work relation was computed under control, ambient conditions, and then determined again following infusion of calcium chloride. A shift upward and to the left of this linear relationship is reflective of a change in intrinsic myocardial contractility, consistent with the effects of increasing Ca2+ availability to the myofilaments with contraction. Adapted, with permission, from (62).


Figure 15. (A) The LV end‐systolic wall stress‐fractional shortening relationship was determined in normal control pigs by infusion of the alpha agonist phenylephrine. A negative linear relationship was observed between increased LV end‐systolic wall stress and fractional shortening, and the 95% confidence interval obtained from these referent normal preparations is shown. Since increased arterial pressure by phenylephrine infusion will cause a reduction in heart rate (via baroreceptor reflex), a constant heart rate was achieved by electrical pacing. With the development of LV failure secondary to chronic rapid pacing, the isochronal LV end‐systolic stress‐fractional shortening values were outside of the normal referent range and fell to the right, suggestive of impaired myocardial contractility. With recovery from this form of LV failure, LV end‐systolic stress‐fractional shortening values returned to within referent normal limits. (B) Schematic interpretation of the LV end‐systolic stress‐fractional shortening relation. As LV afterload is increased, as shown in this example as an increase in LV end‐systolic walls stress (moving from point A to point B), then a fall in LV fractional shortening will occur. This does not imply that a change in LV myocardial contractility has occurred but simply a change in LV afterload. However, if LV fractional shortening is reduced under an equivalent afterload, as shown from the movement of the LV end‐systolic wall stress‐shortening value from point A to point C, then this would be indicative of a fall in myocardial contractility. Conversely, a shift upward in this relation, as shown as movement from point A to point D would be suggestive of increased myocardial contractility. Panel A adapted, with permission, from (183).


Figure 16. A stylized LV pressure‐volume loop that identifies key points in this relation in terms of the cardiac cycle. Point A reflects the end of diastole, mitral valve closure, and the onset of isovolumetric contraction. Point B reflects the point whereby LV pressure development just exceeds that necessary for aortic valve opening and the onset of ejection. Point C reflects the end of systole and aortic valve closure. This point, the end‐systolic pressure‐volume point, is used to develop the end‐systolic pressure volume relation. Point C also reflects the onset of isovolumetric relaxation (active relaxation). Point D reflects the end of the isovolumetric relaxation phase, opening of the mitral valve, and the onset of the filling phase of diastole. This phase of diastole can be used to develop the diastolic pressure‐volume relation. Both the end‐systolic and end‐diastolic pressure‐volume relationships are determined by plotting values with acute changes in LV loading conditions to develop a family of LV pressure‐volume loops and are shown here for illustrative purposed only.


Figure 17. The effects of acute changes in LV preload or afterload on the pressure‐volume relation. Left: With an acute infusion of a volume of physiological solution, an increase in LV preload occurs as reflected as a shift to the right in LV end‐diastolic volume. The increased LV preload will result in a greater stroke volume, hence a larger area within the pressure‐volume loops. Right: An acute increase in LV afterload, which will result in increased LV pressure development, will also cause a shift to the right of the pressure‐volume relation. In this case, an initial increase in LV afterload will cause a reduction in LV stroke volume for that contraction, which will then be translated into a higher LV end‐diastolic volume at the subsequent diastolic phase. As a result, a higher LV preload will occur with the increase in LV afterload, which in turn will result in a normalization of LV stroke volume. These beat‐beat adjustments are an example of the interaction between LV afterload and preload to maintain stroke volume. However, the pressure‐volume area is increased significantly, which means that while stroke volume may be normalized, this is being achieved under much greater work load and with greater myocardial oxygen demand/energy requirements. Figure adapted, with permission, from (175).


Figure 18. Example of a shift in the end‐systolic pressure volume relation following stimulation of the beta adrenergic system by epinephrine in the beating heart preparation as described by Suga and Sagawa (177). The direction and key points of the LV pressure volume loop are shown by the arrows and letters, respectively. With infusion of the positive inotrope epinephrine, a shift to the left in the slope of this relation was demonstrated, indicative of increased myocardial contractility. Adapted, with permission, from (177).


Figure 19. Adult pigs were instrumented with high fidelity, MRI compatible LV microtransducers to measure LV pressure (panel A) and simultaneously LV volumes by MRI (panel B). A series of LV pressure‐volume loops were obtained by a transient reduction in LV preload (vena cava occlusion) under normal resting conditions (rest) and following a dobutamine infusion, thus causing increased beta adrenergic receptor stimulation (stress). Representative LV pressure volume loops are shown in panel C, whereby a shift in the end‐systolic pressure volume relation (ESPVR) was observed with dobutamine infusion, consistent with an increase in inotropic state and hence contractile function. The lower blue points indicate LV diastole and the end‐diastolic pressure volume relation (EDPVR). A cross‐section of the heart by MRI (white line) is shown in panel D, whereby the dimensions across this single plane (“m‐mode”) is shown in panel E, demonstrating the high fidelity of measuring changes in LV volume with transient occlusion of the vena cava. Reproduced, with permission, from (196).


Figure 20. (A) Schematic of the lower portion of the LV pressure‐volume loop whereby the passive filling phase of diastole has been amplified. A curve‐linear relation exists between LV diastolic pressure and volume, whereby a shift upward and to the right demonstrates that an increase in relative LV chamber stiffness occurred. In this example, arrows indicate that at a specific LV diastolic volume, a much higher diastolic pressure is obtained with increased LV chamber stiffness. However, it should be emphasized that using a single measurement of LV diastolic volume and pressure will not allow for interpretation of LV chamber stiffness but must be determined from a set of points. (B) Using a curve fitting algorithm, the relative slope of the change in LV diastolic pressure and volume (dP/dV) is computed and has been termed the LV chamber stiffness modulus, or Kc. However, the dP/dV relation can change as a function of operating volumes and pressures as shown moving from point A to point B. In this example, computing Kc under filling volumes and pressures will yield different values. Thus, when comparing Kc across subjects or across time, it is important to compute this index of LV chamber stiffness under equivalent diastolic volume and pressures as shown moving from point A to point C. In this example, an absolute increase in LV chamber stiffness has occurred. Figures adapted, with permission, from (201).


Figure 21. A number of factors can influence the LV diastolic pressure‐volume relation and thereby will influence the ability to execute the computations necessary for determining LV chamber stiffness. Abnormalities in active relaxation properties, which will therefore prolong the isovolumetric relaxation phase of diastole, can cause shifts in the LV diastolic pressure‐volume relation. This can be due to defects in Ca2+ resequestration or Ca2+ overload, diminished energy/metabolism pathways, alterations in protein structure or composition involved in active relaxation, or a combination of these factors. Another factor that can influence this relationship is an extrinsic one, which is pericardial restraint. This can be due to thickening and/or adhesions of the pericardium, which can occur with chronic inflammatory diseases for example. Another cause for a shift in this relation, and one that is taking on greater import with the high incidence of hypertensive heart disease (aka diastolic heart failure: heart failure with a preserved ejection fraction), is increased LV chamber stiffness (91,107,203,204). Finally, chronic changes in LV geometry, such as increased chamber volumes, will shift this relationship usually upward and to the right as shown. Adapted, with permission, from (29).


Figure 22. Three‐dimensional wire model reconstructed from bi‐lane ventriculography of the LV (red) and RV (blue). The significant differences in geometry of these ventricles can be appreciated, which also translates into differences in the mechanical form of contraction. The LV generates force by a compressive twisting motion whereas the RV is more of a bellows pump type of action. The RV operates under much lower afterload as the pulmonary circuit is a low resistance, high compliance circuit. Thus, the RV can be much more sensitive to increases in afterload, which will cause RV dilation as a compensatory mechanism. The RV dilation can affect LV filling through changes in interventricular septal motion as well as through the pericardium. Thus, LV‐RV interactions can significantly affect overall cardiac performance. Adapted, with permission, from (168).


Figure 23. A schematic demonstrating the relationship between an increase in different determinants of cardiac output to that of myocardial oxygen consumption. The changes have been placed in arbitrary units for the purposes of demonstrating the differences in myocardial oxygen consumption with changes in LV load, heart rate, and finally, contractility. While an equivalent cardiac output may be measured between subjects, the factors that contribute to this value may be distinctly different. Thus, myocardial efficiency, that is the energy cost for a given LV ejection or to maintain total flow (cardiac output), may be very different. An evaluation of cardiac function should not only consider LV ejection but the underlying determinants that govern ejection and flow, as this will identify potential differences in myocardial efficiency.


Figure 24. The assessment of cardiac function in terms of cardiac output can be considered a multivariable equation where LV load, contractility, and rate are independent variables. The measurement of these independent variables will provide for a more comprehensive view of overall cardiac performance and allow for assessment of changes in cardiac function across subjects, time, or treatment. Adapted, with permission, from (9).
References
 1. Abbott BC , Lowy J . Stress relaxation in muscle. Proc R Soc Lond B Biol Sci 146(923): 281‐288, 1956.
 2. Abrahams C , Janicki JS , Weber KT . Myocardial hypertrophy in Macaca fascicularis. Structural remodeling of the collagen matrix. Lab Invest 56(6): 676‐683, 1987.
 3. Allessie MA , Rensma PL , Brugada J , Smeets JLRM , Penn OC , Kirchhof CJHS . Pathophysiology of atrial fibrillation. In: Cardiac Electrophysiology: From Cell to Bedside. Philadelphia, PA: WB Saunders Company, pp. 548‐559, 1990.
 4. Armour JA , Randall WC . Structural basis for cardiac function. Am J Physiol 218(6): 1517‐1523, 1970.
 5. Armstrong WE . Echocardiography. In: Bonow RO , Mann DL , Zipes DP , Libby P , editors. Braunwald's Heart Disease: A Textbook of Cardiovascular Medicine (7th ed). Philadelphia: Elsevier Saunders, pp. 187‐270, 2005.
 6. Baan J , Jong TT , Kerkhof PL , Moene RJ , van Dijk AD , van der Velde ET , Koops J . Continuous stroke volume and cardiac output from intra‐ventricular dimensions obtained with impedance catheter. Cardiovasc Res 15(6): 328‐334, 1981.
 7. Bakos AC . The question of the function of the right ventricular myocardium: An experimental study. Circulation 1: 724‐732, 1950.
 8. Barbee RW , Perry BD , Ré RN , Murgo JP . Microsphere and dilution techniques for the determination of blood flows and volumes in conscious mice. Am J Physiol 263(3 Pt 2): R728‐R733, 1992.
 9. Barrett KE , Barman SM , Boitano S , Brooks HL . The Heart as a Pump. In: Weitz M , Brown RY , editors. Ganong's Review of Medical Physiology (23rd ed). USA: The McGraw Hill Companies, Inc, Figure 31‐5, p. 515, 2010.
 10. Bashey RI , Martinez‐Hernandez A , Jimenez SA . Isolation, characterization, and localization of cardiac collagen type VI. Associations with other extracellular matrix components. Circ Res 70(5): 1006‐1017, 1992.
 11. Bellenger NG , Burgess MI , Ray SG , Lahiri A , Coats AJ , Cleland JG , Pennell DJ . Comparison of left ventricular ejection fraction and volumes in heart failure by echocardiography, radionuclide ventriculography and cardiovascular magnetic resonance; are they interchangeable? Eur Heart J 21(16): 1387‐1396, 2000.
 12. Bennett ED , Barclay SA , Davis AL , Mannering D , Mehta N . Ascending aortic blood velocity and acceleration using Doppler ultrasound in the assessment of left ventricular function. Cardiovasc Res 18(10): 632‐638, 1984.
 13. Bers DM . Excitation‐contraction coupling. In: Bers DM , editor. Excitation‐Contraction Coupling and Cardiac Contractile Force. Norwell MA: Kluwer Academic Publishers, pp. 155‐158, 1991.
 14. Bers DM . Regulation of cellular calcium in cardiac myocytes. In: Page E , Fozzard HA , Solaro RJ , editors. Handbook of Physiology: The Cardiovascular System, Volume 1: The Heart. Oxford: University Press, pp. 335‐387, 2002.
 15. Bhave NM , Lang RM . Evaluation of left ventricular structure and function by three‐dimensional echocardiography. Curr Opin Crit Care 19(5): 387‐396, 2013.
 16. Bishop VS , Horwitz LD , Stone HL , Stegall HF , Engelken EJ . Left ventricular internal diameter and cardiac function in conscious dogs. J Appl Physiol 27(5): 619‐623, 1969.
 17. Blinks JR , Endoh M . Modification of myofibrillar responsiveness to Ca++ as an inotropic mechanism. Circulation 73(3 Pt 2): III85‐III98, 1986.
 18. Boettcher DH , Vatner SF , Heyndrickx GR , Braunwald E . Extent of utilization of the Frank‐Starling mechanism in conscious dogs. Am J Physiol 234(4): H338‐H345, 1978.
 19.Bowditch, Henry P . About the peculiarities of excitability which the show muscle fibers of the heart. Arb Physiol Aust Leipzig 6: 139‐176, 1871.
 20. Brady AJ . Active state in cardiac muscle. Physiol Rev 48(3): 570‐600, 1968.
 21. Brady AJ . Mechanical properties of isolated cardiac myocytes. Physiol Rev 71(2): 413‐428, 1991.
 22. Britman NA , Levine HJ . Contractile element work: A major determinant of myocardial oxygen consumption. J Clin Invest 43: 1397‐1408, 1964.
 23. Buchalter MB1 , Weiss JL , Rogers WJ , Zerhouni EA , Weisfeldt ML , Beyar R , Shapiro EP . Noninvasive quantification of left ventricular rotational deformation in normal humans using magnetic resonance imaging myocardial tagging. Circulation 81(4): 1236‐1244, 1990.
 24. Buckberg GD , Hoffman JI , Coghlan HC , Nanda NC . Ventricular structure‐function relations in health and disease: Part I. The normal heart. Eur J Cardiothorac Surg pii: ezu278, 2014.
 25. Burgeson RE . New collagens, new concepts. Annu Rev Cell Biol 4: 551‐577, 1988.
 26. Burns JW , Covell JW , Ross J Jr . Mechanics of isotonic left ventricular contractions. Am J Physiol 224(4): 725‐732, 1973.
 27. Cannon RO 3rd , Butany JW , McManus BM , Speir E , Kravitz AB , Bolli R , Ferrans VJ . Early degradation of collagen after acute myocardial infarction in the rat. Am J Cardiol 52(3): 390‐395, 1983.
 28. Carabello BA . Progress in mitral and aortic regurgitation. Prog Cardiovasc Dis 43(6): 457‐475, 2001.
 29. Carroll JD , Lang RM , Neumann AL , Borow KM , Rajfer SI . The differential effects of positive inotropic and vasodilator therapy on diastolic properties in patients with congestive cardiomyopathy. Circulation 74(4): 815‐825, 1986.
 30. Caulfield JB , Wolkowicz P . Inducible collagenolytic activity in isolated perfused rat hearts. Am J Pathol 131(2): 199‐205, 1988.
 31. Chan W , Ellims AH , Duffy SJ , Kaye DM , Taylor AJ . Principles, current status and clinical implications of ischaemic heart disease assessment by cardiac magnetic resonance imaging. Intern Med J 42(1): 7‐17, 2012.
 32. Cheung YF . The role of 3D wall motion tracking in heart failure. Nat Rev Cardiol 9(11): 644‐657, 2012.
 33. Clark JE , Marber MS . Advancements in pressure‐volume catheter technology ‐ stress remodelling after infarction. Exp Physiol 98(3): 614‐621, 2013.
 34. Colan SD , Borow KM , Neumann A . Left ventricular end‐systolic wall stress‐velocity of fiber shortening relation: A load‐independent index of myocardial contractility. J Am Coll Cardiol 4(4): 715‐724, 1984.
 35. Covell JW , Ross J Jr . Systolic and diastolic function (mechanics) of the intact heart. In: Page E , Fozzard HA , Solaro RJ , editors. Handbook of Physiology, The Cardiovascular System, The Heart. Oxford: University Press, 2002. pp. 741‐785.
 36. Covell JW , Ross J Jr , Sonnenblick EH , Braunwald E . Comparison of the force‐velocity relation and the ventricular function curve as measures of the contractile state of the intact heart. Circ Res 19(2): 364‐372, 1966.
 37. Covell JW , Ross J Jr , Taylor R , Sonnenblick EH , Braunwald E . Effects of increasing frequency of contraction on the force velocity relation of left ventricle. Cardiovasc Res 1(1): 2‐8, 1967.
 38. Dalen H , Haugen BO , Graven T . Feasibility and clinical implementation of hand‐held echocardiography. Expert Rev Cardiovasc Ther 11(1): 49‐54, 2013.
 39. Davidson CJ , Bonow RO . Cardiac catheterization. In: Braunwald's Heart Disease: A Textbook of Cardiovascular Medicine (7th ed). Philadelphia: Elsevier Saunders, pp. 395‐422, 2005.
 40. Davidson DM , Covell JW , Malloch CI , Ross J Jr . Factors influencing indices of left ventricle contractility in the conscious dog. Cardiovasc Res 8(3): 299‐312, 1974.
 41. de Simone G , Wallerson DC , Volpe M , Devereux RB . Echocardiographic measurement of left ventricular mass and volume in normotensive and hypertensive rats. Necropsy validation. Am J Hypertens 3(9): 688‐696, 1990.
 42.del Monte F , Hajjar RJ , Harding SE . Overwhelming evidence of the beneficial effects of SERCA gene transfer in heart failure. Circ Res 88(11): E66‐E67, 2001.
 43. Dell'Italia LJ . Anatomy and physiology of the right ventricle. Cardiol Clin 30(2): 167‐187, 2012.
 44. Edman KAP , Johannsson M . The contractile state of rabbit papillary muscle in relation to stimulation frequency. J Physiol 254(3): 565‐581, 1976.
 45. Eising GP , Hammond HK , Helmer GA , Gilpin E , Ross J Jr . Force‐frequency relations during heart failure in pigs. Am J Physiol 267(6 Pt 2): H2516‐H2522, 1994.
 46. Fabiato A , Fabiato F . Myofilament‐generated tension oscillations during partial calcium activation and activation dependence of the sarcomere length‐tension relation of skinned cardiac cells. J Gen Physiol 72(5): 667‐699, 1978.
 47. Feldman DS , Carnes CA , Abraham WT , Bristow MR . Mechanisms of disease: Beta‐adrenergic receptors–alterations in signal transduction and pharmacogenomics in heart failure. Nat Clin Pract Cardiovasc Med 2(9): 475‐483, 2005.
 48. Fick A . Ueber die Messung des Blutquantums in den Herzventrikeln. Proceedings of the Physical Medical Society to Wurzburg. XVI, 1870.
 49. Fomovsky GM , Thomopoulos S , Holmes JW . Contribution of extracellular matrix to the mechanical properties of the heart. J Mol Cell Cardiol 48(3): 490‐496, 2010.
 50. Frank O . For the dynamics of the heart muscle. Z Biol. 32:370‐437, 1895.
 51. Frank O . Thermodynamics of the muscle. Reviews of physiology. 3: 348‐412, 1904.
 52. Freeman GL . Effects of increased afterload on left ventricular function in closed‐chest dogs. Am J Physiol 259(2 Pt 2): H619‐H625, 1990.
 53. Freeman GL , Colston JT . Evaluation of long‐term variance of left ventricular performance indexes in closed‐chest dogs. Am J Physiol 257(1 Pt 2): H70‐H78, 1989.
 54. Freeman GL , LeWinter MM . Pericardial adaptations during chronic cardiac dilation in dogs. Circ Res 54(3): 294‐300, 1984.
 55. Freeman GL , Little WC , O'Rourke RA . Influence of heart rate on left ventricular performance in conscious dogs. Circ Res 61(3): 455‐464, 1987.
 56. Fukuta H , Little WC . Contribution of systolic and diastolic abnormalities to heart failure with a normal and a reduced ejection fraction. Prog Cardiovasc Dis 49(4): 229‐240, 2007.
 57. Furnival CM , Linden RJ , Snow HM . Inotropic changes in the left ventricle: The effect of changes in heart rate, aortic pressure and end‐diastolic pressure. J Physiol 211(2): 359‐387, 1970.
 58. Gaasch WH , Delorey DE , St John Sutton MG , Zile MR . Patterns of structural and functional remodeling of the left ventricle in chronic heart failure. Am J Cardiol 102(4): 459‐462, 2008.
 59. Gaudio C , Tanzilli G , Mazzarotto P , Motolese M , Romeo F , Marino B , Reale A . Comparison of left ventricular ejection fraction by magnetic resonance imaging and radionuclide ventriculography in idiopathic dilated cardiomyopathy. Am J Cardiol 67(5): 411‐415, 1991.
 60. Gibbons WR , Zygmunt AC . Excitation‐contraction coupling in heart. In: Page E , Fozzard HA , Solaro RJ , editors. The Heart and Cardiovascular System, Vol 2. New York: Raven Press, pp. 1249‐1279, 1991.
 61. Glower DD , Spratt JA , Snow ND , Kabas JS , Davis JW , Olsen CO , Tyson GS , Sabiston DC Jr , Rankin JS . Linearity of the Frank‐Starling relationship in the intact heart: The concept of preload recruitable stroke work. Circulation 71(5): 994‐1009, 1985.
 62. Goldstein MA , Schroeter JP . Ultrastructure of the heart. In: Reich M , Lindsey M , editors. Handbook of Physiology: The Cardiovascular System, Volume 1: The Heart. Oxford: University Press, pp. 3‐74, 2002.
 63. Golestani R , Wu C , Tio RA , Zeebregts CJ , Petrov AD , Beekman FJ , Dierckx RA , Boersma HH , Slart RH . Small‐animal SPECT and SPECT/CT: Application in cardiovascular research. Eur J Nucl Med Mol Imaging 37(9): 1766‐1777, 2010.
 64. Gorman JH III , Gupta KB , Streicher JT , Gorman RC , Jackson BM , Ratcliffe MB , Bogen DK , Edmunds LH Jr . Dynamic three‐dimensional imaging of the mitral valve and left ventricle by rapid sonomicrometry array localization. J Thorac Cardiovasc Surg 112(3): 712‐726, 1996.
 65. Grimm AF , Whitehorn WV . Characteristics of resting tension of myocardium and localization of its elements. Am J Physiol 210(6): 1362‐1368, 1966.
 66. Grossman W , Jones D , McLaurin LP . Wall stress and patterns of hypertrophy in the human left ventricle. J Clin Invest 56(1): 56‐64, 1975.
 67. Grossman W . Blood flow measurement: The cardiac output. In: Goolsby J , editor. Cardiac Catheterization, Angiography, and Intervention. Philadelphia: Lippincott, Williams, & Wilkins, pp. 101‐117, 2000.
 68. Hains AD , Al‐Khawaja I , Hinge DA , Lahiri A , Raftery EB . Radionuclide left ventricular ejection fraction. Br Heart J 57(3): 242‐246, 1987.
 69. Hales S . Statical Essays Containing Haemastatics (2nd ed). London: W. Innys and R. Manby, 1733.
 70. Hamilton WF , Moore JW , Kinsman JM . Studies on the circulation. IV. Further analysis of the injection method and of changes in hemodynamics under physiological and pathological conditions. Am J Physiol 99: 534, 1932.
 71. Hammond HK1 , White FC , Bhargava V , Shabetai R . Heart size and maximal cardiac output are limited by the pericardium. Am J Physiol 263(6 Pt 2): H1675‐H1681, 1992.
 72. Hawthorne EW . Instantaneous dimensional changes of the left ventricle in dogs. Circ Res 9: 110‐119, 1961.
 73. Hein S , Arnon E , Kostin S , Schönburg M , Elsässer A , Polyakova V , Bauer EP , Klövekorn WP , Schaper J . Progression from compensated hypertrophy to failure in the pressure‐overloaded human heart: Structural deterioration and compensatory mechanisms. Circulation 107(7): 984‐991, 2003.
 74. Hibberd MG , Jewell BR . Calcium‐ and length‐ dependent force production in rat ventricular muscle. J Physiol. 1982; 329: 527‐540
 75. Hoffmann R , Barletta G , von Bardeleben S , Vanoverschelde JL , Kasprzak J , Greis C , Becher H . Analysis of left ventricular volumes and function: A multicenter comparison of cardiac magnetic resonance imaging, cine ventriculography, and unenhanced and contrast‐enhanced two‐dimensional and three‐dimensional echocardiography. J Am Soc Echocardiogr 27(3): 292‐301, 2014.
 76. Hoffman BF , Bassett AL , Bartelstone HJ . Some mechanical properties of isolated mammalian cardiac muscle. Circ Res 23(2): 291‐312, 1968.
 77. Hryshko LV . Cardiac Na+‐Ca2+exchanger. Handbook of Physiology, The Cardiovascular System, The Heart. Oxford: University Press, 2011.
 78. Hryshko LV . The cardiac Na+‐Ca2+ exchanger. In: Page E , Fozzard HA , Solaro RJ , editors. Handbook of Physiology: The Cardiovascular System, Volume 1: The Heart. Oxford: University Press, pp. 388‐419, 2002.
 79. Iwano H , Little WC . Heart failure: What does ejection fraction have to do with it? J Cardiol 62(1): 1‐3, 2013.
 80. Julian FJ , Sollins MR , Moss RL . Absence of a plateau in length‐tension relationship of rabbit papillary muscle when internal shortening is prevented. Nature 26: 340‐342. 1976.
 81. Kahn ML , Kavaler F , Fisher VJ . Frequency‐force relationships of mammalian ventricular muscle in vivo and in vitro. Am J Physiol 230(3): 631‐636, 1976.
 82. Kambayashi M , Miura T , Oh BH , Rockman HA , Murata K , Ross J Jr . Enhancement of the force‐frequency effect on myocardial contractility by adrenergic stimulation in conscious dogs. Circulation 86(2): 572‐580, 1992.
 83. Karliner JS , Gault JH , Eckberg D , Mullins CB , Ross J Jr . Mean velocity of fiber shortening. A simplified measure of left ventricular myocardial contractility. Circulation 44(3): 323‐333, 1971.
 84. Kass DA . Alterations in ventricular function: Systolic heart failure. In: Mann D , editor. Heart Failure: A Companion to Braunwald's Heart Disease. St. Louis: Saunders, Elsevier, p. 198‐212, 2011.
 85. Kass DA , Bronzwaer JG , Paulus WJ . What mechanisms underlie diastolic dysfunction in heart failure? Circ Res 94(12): 1533‐1542, 2004.
 86. Keller RS , Shai SY , Babbitt CJ , Pham CG , Solaro RJ , Valencik ML , Loftus JC , Ross RS . Disruption of integrin function in the murine myocardium leads to perinatal lethality, fibrosis, and abnormal cardiac performance. Am J Pathol 158(3): 1079‐1090, 2001.
 87. Khunti K . Systematic review of open access echocardiography for primary care. Eur J Heart Fail 6(1): 79‐83, 2004.
 88. Kimura BJ , DeMaria AN . Technology insight: Hand‐carried ultrasound cardiac assessment–evolution, not revolution. Nat Clin Pract Cardiovasc Med 2(4): 217‐223, 2005
 89. Kitzman DW , Little WC , Brubaker PH , Anderson RT , Hundley WG , Marburger CT , Brosnihan B , Morgan TM , Stewart KP . Pathophysiological characterization of isolated diastolic heart failure in comparison to systolic heart failure. JAMA 288(17): 2144‐2150, 2002.
 90. Kleber AG , Janse MJ , Fast VG . Normal and abnormal conduction in the heart. In: Page E , Fozzard HA , Solaro RJ , editors. Handbook of Physiology: The Cardiovascular System, Volume 1: The Heart. Oxford: University Press, pp. 455‐530, 2002.
 91. Koeppen BM , Stanton BA . Elements of cardiac function. In: Koeppen BM , Stanton BA , editors. Berne and Levy Physiology (6th ed). Philadelphia: Mosby Elsevier, p. 321, 2008.
 92. Komamura K , Shannon RP , Ihara T , Shen YT , Mirsky I , Bishop SP , Vatner SF . Exhaustion of Frank‐Starling mechanism in conscious dogs with heart failure. Am J Physiol 265(4 Pt 2): H1119‐H1131, 1993.
 93. Kranias EG , Hajjar RJ . Modulation of cardiac contractility by the phospholamban/SERCA2a regulatome. Circ Res 110(12): 1646‐1660, 2012.
 94. Krayenbuehl HP , Hess OM , Monrad ES , Schneider J , Mall G , Turina M . Left ventricular myocardial structure in aortic valve disease before, intermediate, and late after aortic valve replacement. Circulation 79(4): 744‐755, 1989.
 95. Kresh JY , Chopra A . Intercellular and extracellular mechanotransduction in cardiac myocytes. Pflugers Arch 462(1): 75‐87, 2011.
 96. Lakatta EG . Changes in cardiovascular function with aging. Eur Heart J 11(Suppl C): 22‐29, 1990.
 97. Lakatta EG . Length modulation of muscle performance. In: Page E , Fozzard HA , Solaro RJ , editors. The Heart and Cardiovascular System, Vol 2. New York: Raven Press, pp. 1325‐1351, 1991.
 98. Lakatta EG , Jewell BR . Length‐dependent activation: Its effect on the length‐tension relation in cat ventricular muscle. Circ Res 40(3): 251‐257, 1977.
 99. Lang RM , Bierig M , Devereux RB , Flachskampf FA , Foster E , Pellikka PA , Picard MH , Roman MJ , Seward J , Shanewise JS , Solomon SD , Spencer KT , Sutton MS , Stewart WJ ; Chamber Quantification Writing Group; American Society of Echocardiography's Guidelines and Standards Committee; European Association of Echocardiography. Recommendations for chamber quantification: A report from the American Society of Echocardiography's Guidelines and Standards Committee and the Chamber Quantification Writing Group, developed in conjunction with the European Association of Echocardiography, a branch of the European Society of Cardiology. J Am Soc Echocardiogr 18(12): 1440‐1463, 2005.
 100. Lew WY . Mechanisms of volume‐induced increase in left ventricular contractility. Am J Physiol 265(5 Pt 2): H1778‐H1786, 1993.
 101. LeWinter MM , Fabian J , Bell SP . Left ventricular restoring forces: Modulation by heart rate and contractility. Basic Res Cardiol 93(Suppl 1): 143‐147, 1998.
 102. LeWinter MM , Kabbani S . Pericardial diseases. In: Bonow RO , Mann DL , Zipes DP , Libby P , editors. Braunwald's Heart Disease: A Textbook of Cardiovascular Medicine (7th ed). Philadelphia: Elsevier Saunders, pp. 1757‐1780, 2005.
 103. Li AH , Liu PP , Villarreal FJ , Garcia RA . Dynamic changes in myocardial matrix and relevance to disease: Translational perspectives. Circ Res 114(5): 916‐927, 2014.
 104. Little WC . Diastolic dysfunction beyond distensibility: Adverse effects of ventricular dilatation. Circulation 112(19): 2888‐2890, 2005.
 105. Little WC , Kitzman DW , Cheng CP . Diastolic dysfunction as a cause of exercise intolerance. Heart Fail Rev 5(4): 301‐306, 2000.
 106. Loiselle DS , Crampin EJ , Niederer SA , Smith NP , Barclay CJ . Energetic consequences of mechanical loads. Prog Biophys Mol Biol 97(2‐3): 348‐366, 2008.
 107. Ludin G . Mechanical properties of cardiac muscle. Acta Physiol Scand 7: 1‐85, 1944.
 108. MacKenna DA , Omens JH , McCulloch AD , Covell JW . Contribution of collagen matrix to passive left ventricular mechanics in isolated rat hearts. Am J Physiol 266(3 Pt 2): H1007‐H1018, 1994.
 109. MacKenna DA , Vaplon SM , McCulloch AD . Microstructural model of perimysial collagen fibers for resting myocardial mechanics during ventricular filling. Am J Physiol 273(3 Pt 2): H1576‐H1586, 1997.
 110. MacLennan DH , Kranias EG . Phospholamban: A crucial regulator of cardiac contractility. Nat Rev Mol Cell Biol 4(7): 566‐577, 2003.
 111. Mall FP . On the muscular architecture of the ventricles of the human heart. Am J Anat 11: 211‐266, 1911.
 112. Markwalder J , Starling EH . A note on some factors which determine the blood‐flow through the coronary circulation. J Physiol 47(4‐5): 275‐285, 1913.
 113. Markwalder J , Starling EH . On the constancy of the systolic output under varying conditions. J Physiol 48(4): 348‐356, 1914.
 114. Mason DT . Usefulness and limitations of the rate of rise of intraventricular pressure (dp‐dt) in the evaluation of myocardial contractility in man. Am J Cardiol 23(4): 516‐527, 1969.
 115. Maughan WL , Sunagawa K , Burkhoff D , Sagawa K . Effect of arterial impedance changes on the end‐systolic pressure‐volume relation. Circ Res 54(5): 595‐602, 1984.
 116. Meier GD , Ziskin MC , Santamore WP , Bove AA . Kinematics of the beating heart. IEEE Trans Biomed Eng 27(6): 319‐329, 1980.
 117. Mirsky I . Left ventricular stresses in the intact human heart. Biophys J 9(2): 189‐208, 1969.
 118. Mirsky I , Parmley WW . Force‐velocity studies in isolated and intact heart muscle. In: Mirsky M , Ghista DH , Sandler H , editors. Cardiac Mechanics: Physiological, Clinical, and Mathematical Considerations. New York: Wiley, pp. 87‐112, 1974.
 119. Mirsky I , Tajimi T , Peterson KL . The development of the entire end‐systolic pressure‐volume and ejection fraction‐afterload relations: A new concept of systolic myocardial stiffness. Circulation 76(2): 343‐356, 1987.
 120. Knowlton AA and Liu TT . Mitochondrial dynamics in heart failure. Comp Physiol (in press).
 121. Møgelvang J , Stokholm KH , Saunamäki K , Reimer A , Stubgaard M , Thomsen C , Fritz‐Hansen P , Henriksen O . Assessment of left ventricular volumes by magnetic resonance in comparison with radionuclide angiography, contrast angiography and echocardiography. Eur Heart J 13(12): 1677‐1683, 1992.
 122. Moss RL , Buck SH . Regulation of cardiac contraction by calcium. In: Page E , Fozzard HA , Solaro RJ , editors. Handbook of Physiology: The Cardiovascular System, Volume 1: The Heart. Oxford: University Press, pp. 420‐454, 2002.
 123. Moss RL , Buck SH . The regulation of cardiac contraction. Handbook of Physiology. Oxford: University Press, 2011.
 124. Nakano K , Swindle MM , Spinale FG , Ishihara K , Kanazawa F , Smith A , Biederman RWW, Clamp L , Hamada Y , Zile MR , Carabello BA . Depressed contractile function due to canine mitral regurgitation improves following correction of the volume overload. J Clin Invest, 87: 2077‐2086, 1991
 125. Ono S , Bhargava V , Ono S , Mao L , Hagan G , Rockman HA , Ross J Jr . In vivo assessment of left ventricular remodelling after myocardial infarction by digital video contrast angiography in the rat. Cardiovasc Res 28(3): 349‐357, 1994.
 126. Owen AA , Spinale FG . Myocardial basis for heart failure: Role of the cardiac interstitium. In: Mann DL , editor. Heart Failure: A Companion to Braunwald's Heart Disease. St. Louis: Saunders, Elsevier, pp. 73‐84, 2004.
 127. Palakodeti V , Oh S , Oh BH , Mao L , Hongo M , Peterson KL , Ross J Jr . Force‐frequency effect is a powerful determinant of myocardial contractility in the mouse. Am J Physiol 273(3 Pt 2): H1283‐H1290, 1997.
 128. Parry DA . The molecular and fibrillar structure of collagen and its relationship to the mechanical properties of connective tissue. Biophys Chem 29(1‐2): 195‐209, 1988.
 129. Patel CN , Chowdhury FU , Scarsbrook AF . Hybrid SPECT/CT: The end of “unclear” medicine. Postgrad Med J 85(1009): 606‐613, 2009.
 130. Patterson SW , Piper H , Starling EH . The regulation of the heartbeat. J Physiol 48(6): 465‐513, 1914.
 131. Patterson WW , Starling EH . On the mechanical factors which determine the output of the ventricles. J Physiol 48(5): 357‐379, 1914.
 132. Pennell D . Cardiovascular magnetic resonance. In: Bonow RO , Mann DL , Zipes DP , Libby P , editors. Braunwald's Heart Disease: A Textbook of Cardiovascular Medicine (7th ed). Philadelphia: Elsevier Saunders, pp. 335‐353, 2005.
 133. Periasamy M , Bhupathy P , Babu GJ . Regulation of sarcoplasmic reticulum Ca2+ ATPase pump expression and its relevance to cardiac muscle physiology and pathology. Cardiovasc Res 77(2): 265‐273, 2008.
 134. Peterson KL , Skloven D , Ludbrook P , Uther JB , Ross J Jr . Comparison of isovolumic and ejection phase indices of myocardial performance in man. Circulation 49(6): 1088‐1101, 1974.
 135. Pouleur H , Lefèvre J , Van Mechelen H , Charlier AA . Free‐wall shortening and relaxation during ejection in the canine right ventricle. Am J Physiol 239(5): H601‐H613, 1980.
 136. Quinones MA , Gaasch WH , Alexander JK . Influence of acute changes in preload, afterload, contractile state and heart rate on ejection and isovolumic indices of myocardial contractility in man. Circulation 53(2): 293‐302, 1976.
 137. Quinones MA , Waggoner AD , Reduto LA , Nelson JG , Young JB , Winters WL Jr, Ribeiro LG , Miller RR . A new, simplified and accurate method for determining ejection fraction with two‐dimensional echocardiography. Circulation 64(4): 744‐753, 1981.
 138. Rausch MK , Famaey N , Shultz TO , Bothe W , Miller DC , Kuhl E . Mechanics of the mitral valve: A critical review, an in vivo parameter identification, and the effect of prestrain. Biomech Model Mechanobiol 12(5): 1053‐1071, 2013.
 139. Ritter M , Hess OM , Murakami T , Jenni R , Egloff L , Nonogi H , Schneider J , Krayenbuehl HP . Left ventricular systolic series elastic properties in aortic stenosis before and after valve replacement. Cardiovasc Res 22(11): 759‐767, 1988.
 140. Robb JS , Robb RC . The normal heart. Anatomy and physiology of the structural units. Am Heart J 23: 455‐467, 1942.
 141. Robinson TF , Geraci MA , Sonnenblick EH , Factor SM . Coiled perimysial fibers of papillary muscle in rat heart: Morphology, distribution, and changes in configuration. Circ Res 63(3): 577‐592, 1988.
 142. Ross J Jr, Braunwald E . Aortic stenosis. Circulation 38(1 Suppl): 61‐67, 1968.
 143. Ross J Jr, Covell JW , Sonnenblick EH , Braunwald E . Contractile state of the heart characterized by force‐velocity relations in variably afterloaded and isovolumic beats. Circ Res. 18: 149‐163, 1966.
 144. Rudy Y . Cardiac ventricular action potential. Handbook of Physiology, The Cardiovascular System, The Heart. Oxford: University Press, 2011.
 145. Sagawa K . The ventricular pressure‐volume diagram revisited. Circ Res 43(5): 677‐687, 1978.
 146. Sagawa K , Suga H , Shoukas AA , Bakalar KM . End‐systolic pressure/volume ratio: A new index of ventricular contractility. Am J Cardiol 40(5): 748‐753, 1977.
 147. Santamore WP , Lynch PR , Heckman JL , Bove AA , Meier GD . Left ventricular effects on right ventricular developed pressure. J Appl Physiol 41(6): 925‐930, 1976.
 148. Sarnoff SJ , Mitchell JH . The regulation of the performance of the heart. Am J Med 30: 747‐771, 1961.
 149. Sasaki T , Inui M , Kimura Y , Kuzuya T , Tada M . Molecular mechanism of regulation of Ca2+ pump ATPase by phospholamban in cardiac sarcoplasmic reticulum. Effects of synthetic phospholamban peptides on Ca2+ pump ATPase. J Biol Chem 267(3): 1674‐1679, 1992.
 150. Schiller NB , Acquatella H , Ports TA , Drew D , Goerke J , Ringertz H , Silverman NH , Brundage B , Botvinick EH , Boswell R , Carlsson E , Parmley WW . Left ventricular volume from paired biplane two‐dimensional echocardiography. Circulation 60(3): 547‐555, 1979.
 151. Schmidt U , Hajjar RJ , Helm PA , Kim CS , Doye AA , Gwathmey JK . Contribution of abnormal sarcoplasmic reticulum ATPase activity to systolic and diastolic dysfunction in human heart failure. J Mol Cell Cardiol 30(10): 1929‐1937, 1998.
 152. Schmidt U , Hajjar RJ , Kim CS , Lebeche D , Doye AA , Gwathmey JK . Human heart failure: cAMP stimulation of SR Ca(2+)‐ATPase activity and phosphorylation level of phospholamban. Am J Physiol 277(2 Pt 2): H474‐H480, 1999.
 153. Schwinn DA , Caron MG , Lefkowitz RJ . The beta‐adrenergic receptor as a model for molecular structure‐function relationships in G protein‐coupled receptors. In: Page E , Fozzard HA , Solaro RJ , editors. The Heart and Cardiovascular System, Vol 2. New York: Raven Press, pp. 1657‐1684, 1991.
 154. Sechtem U , Sommerhoff BA , Markiewicz W , White RD , Cheitlin MD , Higgins CB . Regional left ventricular wall thickening by magnetic resonance imaging: Evaluation in normal persons and patients with global and regional dysfunction. Am J Cardiol 59(1): 145‐151, 1987.
 155. Sengupta PP , Khandheria BK , Narula J . Twist and untwist mechanics of the left ventricle. Heart Fail Clin 4(3): 315‐324, 2008.
 156. Sengupta PP , Tajik AJ , Chandrasekaran K , Khandheria BK . Twist mechanics of the left ventricle: Principles and application. JACC Cardiovasc Imaging 1(3): 366‐376, 2008.
 157. Shabetai R . Pericardial and cardiac pressure. Circulation 77(1): 1‐5, 1988.
 158. Shai SY , Harpf AE , Babbitt CJ , Jordan MC , Fishbein MC , Chen J , Omura M , Leil TA , Becker KD , Jiang M , Smith DJ , Cherry SR , Loftus JC , Ross RS . Cardiac myocyte‐specific excision of the beta1 integrin gene results in myocardial fibrosis and cardiac failure. Circ Res 90(4): 458‐464, 2002.
 159. Sheehan FH , Feneley MP , DeBruijn NP , Rankin JS , Davis JW , Bolson EL , Glass PS , Clements FM . Quantitative analysis of regional wall thickening by transesophageal echocardiography. J Thorac Cardiovasc Surg 103(2): 347‐354, 1992.
 160. Shirato K , Shabetai R , Bhargava V , Franklin D , Ross J Jr . Alteration of the left ventricular diastolic pressure‐segment length relation produced by the pericardium. Effects of cardiac distension and afterload reduction in conscious dogs. Circulation 57(6): 1191‐1198, 1978.
 161. Smith VE , Zile MR . Relaxation and diastolic properties of the heart. In: Page E , Fozzard HA , Solaro RJ , editors. The Heart and Cardiovascular System, Vol 2. New York: Raven Press, pp. 1353‐1367, 1992.
 162. Solaro RJ . Modulation of cardiac myofilament activity by protein phosphorylation. In: Page E , Fozzard HA , Solaro RJ , editors. Handbook of Physiology: The Cardiovascular System, Volume 1: The Heart. Oxford: University Press, pp. 264‐300, 2002.
 163. Sonnenblick EH , Ross J Jr , Covell JW , Braunwald E . Alterations in resting length‐tension relations of cardiac muscle induced by changes in contractile force. Circ Res. 19: 980‐988, 1966.
 164. Spinale FG . Myocardial matrix remodeling and the matrix metalloproteinases: Influence on cardiac form and function. Physiol Rev 87(4): 1285‐1342, 2007.
 165. Spinale FG , Crawford FA , Carabello BA . Right ventricular function and modeling using computer aided design. J Appl Physiol 68: 1707‐1716, 1990.
 166. Spotnitz HM , Sonnenblick EH , Spiro D . Relation of ultrastructure to function in the intact heart: Sarcomere structure relative to pressure volume curves of intact left ventricles of dog and cat. Circ Res 18(1): 49‐66, 1966.
 167. Spray DC , Suadicani SO , Srinivas M , Gutstein DE , Fishman GI . Gap junctions in the cardiovascular system. In: Page E , Fozzard HA , Solaro RJ , editors. Handbook of Physiology: The Cardiovascular System, Volume 1: The Heart. Oxford: University Press, pp. 169‐212, 2002.
 168. Starling EH . The Linacre Lecture on the Law of the Heart, Given at Cambridge. London: Longmans, Green and Co, 1918.
 169. Starr I , Collins LH , Wood FC . Studies of the basal work and output of the heart in clinical conditions. J Clin Invest 12(1): 13‐43, 1933.
 170. Streeter DD Jr . Gross morphology and fiber geometry of the heart. In: Berne RM , Geiger SR , Sperelakis N , editors. Handbook of Physiology. Section 2: The Cardiovascular System. Baltimore: American Physiological Society, pp. 61‐112, 1979.
 171. Stroud JD , Baicu CF , Barnes MA , Spinale FG , Zile MR . Viscoelastic properties of pressure overload hypertrophied myocardium: Effect of serine protease treatment. Am J Physiol Heart Circ Physiol 282(6): H2324‐H2335, 2002.
 172. Suga H . Ventricular energetics. Physiol Rev 70(2): 247‐277, 1990.
 173. Suga H , Hisano R , Goto Y , Yamada O , Igarashi Y . Effect of positive inotropic agents on the relation between oxygen consumption and systolic pressure volume area in canine left ventricle. Circ Res 53(3): 306‐318, 1983.
 174. Suga H , Sagawa K . Instantaneous pressure‐volume relationships and their ratio in the excised, supported canine left ventricle. Circ Res 35(1): 117‐126, 1974.
 175. Susanni EE , Vatner DE , Homcy CJ . The beta‐adrenergic receptor/adenylyl cyclase system. In: Page E , Fozzard HA , Solaro RJ , editors. The Heart and Cardiovascular System, Vol 2. New York: Raven Press, pp. 1685‐1708, 1991.
 176. Tada M . Cardiac sarcoplasmic reticulum Ca2+‐ATPase. Handbook of Physiology, The Cardiovascular System, The Heart. Oxford: University Press, 2011.
 177. Tada M , Toyofuku T . Cardiac sarcoplasmic reticulum Ca2+‐ATPase. In: Page E , Fozzard HA , Solaro RJ , editors. Handbook of Physiology: The Cardiovascular System, Volume 1: The Heart. Oxford: University Press, pp. 308‐334, 2002.
 178. Taylor RR , Covell JW , Ross J Jr . Influence of the thyroid state on left ventricular tension‐velocity relations in the intact, sedated dog. J Clin Invest 48(4): 775‐784, 1969.
 179.ter Keurs HE , Rijnsburger WH , van Heuningen R , Nagelsmit MJ . Tension development and sarcomere length in rat cardiac trabeculae. Evidence of length‐dependent activation. Circ Res 46(5): 703‐714, 1980.
 180. Tomita M , Spinale FG , Crawford FA , Zile MR . Changes in left ventricular volume, mass, and function during the development and regression of supraventricular tachycardia‐induced cardiomyopathy. Disparity between recovery of systolic versus diastolic function. Circulation 83(2): 635‐644, 1991.
 181. Trautwein W , Hescheler J . Regulation of cardiac L‐type calcium current by phosphorylation and G proteins. Annu Rev Physiol 52: 257‐274, 1990.
 182. Tsutsui H , Spinale FG , Nagatsu M , Nagatsu M , Schmid PG , Ishihara K , DeFreyte G , Cooper G , Carabello B . The effects of chronic b‐adrenergic blockade on the left ventricular and cardiocyte abnormalities of chronic mitral regurgitation. J Clin Invest 93: 2639‐2648, 1994.
 183. Udelson JE , Dilsizian V , Bonow RO . Nuclear cardiology. Braunwald's Heart Disease: A Textbook of Cardiovascular Medicine (7th ed). Philadelphia: Elsevier Saunders, pp. 287‐333, 2005.
 184. Vandsburger MH , Epstein FH . Emerging MRI methods in translational cardiovascular research. J Cardiovasc Transl Res 4(4): 477‐492, 2011.
 185. van Heerebeek L , Franssen CP , Hamdani N , Verheugt FW , Somsen GA , Paulus WJ . Molecular and cellular basis for diastolic dysfunction. Curr Heart Fail Rep 9(4): 293‐302, 2012.
 186. Wackers FJ , Berger HJ , Johnstone DE , Goldman L , Reduto LA , Langou RA , Gottschalk A , Zaret BL . Multiple gated cardiac blood pool imaging for left ventricular ejection fraction: Validation of the technique and assessment of variability. Am J Cardiol 43(6): 1159‐1166, 1979.
 187. Waldman LK , Fung YC , Covell JW . Transmural myocardial deformation in the canine left ventricle. Normal in vivo three‐dimensional finite strains. Circ Res 57(1): 152‐163, 1985.
 188. Warbasse JR , Braunwald E , Aygen MM . Starling's law of the heart. VII. Ventricular function in closed‐chest unanesthetized dogs. Am J Physiol 204: 439‐445, 1963.
 189. Weber KT . Cardiac interstitium in health and disease: The fibrillar collagen network. J Am Coll Cardiol 13(7): 1637‐1652, 1989.
 190. Wei CL , Shih MH . Calibration capacity of the conductance‐to‐volume conversion equations for the mouse conductance catheter measurement system. IEEE Trans Biomed Eng 56(6): 1627‐1634, 2009.
 191. Wei CL , Valvano JW , Feldman MD , Nahrendorf M , Peshock R , Pearce JA . Volume catheter parallel conductance varies between end‐systole and end‐diastole. IEEE Trans Biomed Eng 54(8): 1480‐1489, 2007.
 192. Wiggers CJ , Katz LN . The contour of the ventricular volume curves under different conditions. Am J Physiol. 58: 439‐475, 1922.
 193. Witschey WR , Contijoch F , McGarvey JR , Ferrari VA , Hansen MS , Lee ME , Takebayashi S , Aoki C , Chirinos JA , Yushkevich PA , Gorman JH III , Pilla JJ , Gorman RC . Real‐time magnetic resonance imaging technique for determining left ventricle pressure‐volume loops. Ann Thorac Surg 97(5): 1597‐1603, 2014.
 194. Wood PW , Choy JB , Nanda NC , Becher H . Left ventricular ejection fraction and volumes: It depends on the imaging method. Echocardiography 31(1): 87‐100, 2014.
 195. Yarbrough WM , Mukherjee R , Ikonomidis JS , Zile MR , Spinale FG . Myocardial remodeling with aortic stenosis and after aortic valve replacement: Mechanisms and future prognostic implications. J Thorac Cardiovasc Surg 143(3): 656‐664, 2012.
 196. Yellin EL , Meisner JS . Physiology of diastolic function and transmitral pressure‐flow relations. Cardiol Clin 18(3): 411‐433, 2000.
 197. Young AA , Cowan BR . Evaluation of left ventricular torsion by cardiovascular magnetic resonance. J Cardiovasc Magn Reson 14: 49, 2012.
 198. Zile MR , Baicu CF . Alterations in ventricular function: Diastolic heart failure. Heart Failure: A Companion to Braunwald's Heart Disease. St. Louis: Saunders, Elsevier, pp. 213‐231, 2011.
 199. Zile MR , Baicu CF , Bonnema DD . Diastolic heart failure: Definitions and terminology. Prog Cardiovasc Dis 47(5): 307‐133, 2005.
 200. Zile MR , Brutsaert DL . New concepts in diastolic dysfunction and diastolic heart failure: Part I: Diagnosis, prognosis, and measurements of diastolic function. Circulation 105(11): 1387‐1393, 2002.
 201. Zile MR , Spinale FG . Integrating the myocardial matrix into heart failure recognition and management. Circ Res 113(6): 725‐738, 2013.
 202. Zile MR , Tanaka R , Lindroth JR , Spinale FG , Carabello BA , Mirsky I . Left ventricular volume determined echocardiographically by using a constant LV epicardial long axis to short axis dimension ratio throughout the cardiac cycle. J Am Coll Cardiol 20(4): 986‐983, 1992.

Contact Editor

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

* Required Field

How to Cite

Francis G. Spinale. Assessment of Cardiac Function—Basic Principles and Approaches. Compr Physiol 2015, 5: 1911-1946. doi: 10.1002/cphy.c140054