Comprehensive Physiology Wiley Online Library

Energetics of Contraction

Full Article on Wiley Online Library



ABSTRACT

Muscles convert energy from ATP into useful work, which can be used to move limbs and to transport ions across membranes. The energy not converted into work appears as heat. At the start of contraction heat is also produced when Ca2+ binds to troponin‐C and to parvalbumin. Muscles use ATP throughout an isometric contraction at a rate that depends on duration of stimulation, muscle type, temperature and muscle length. Between 30% and 40% of the ATP used during isometric contraction fuels the pumping Ca2+ and Na+ out of the myoplasm. When shortening, muscles produce less force than in an isometric contraction but use ATP at a higher rate and when lengthening force output is higher than the isometric force but rate of ATP splitting is lower. Efficiency quantifies the fraction of the energy provided by ATP that is converted into external work. Each ATP molecule provides 100 zJ of energy that can potentially be converted into work. The mechanics of the myosin cross‐bridge are such that at most 50 zJ of work can be done in one ATP consuming cycle; that is, the maximum efficiency of a cross‐bridge is ∼50%. Cross‐bridges in tortoise muscle approach this limit, producing over 90% of the possible work per cycle. Other muscles are less efficient but contract more rapidly and produce more power. © 2015 American Physiological Society. Compr Physiol 5:961‐995, 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. The dependence of molar enthalpy change for PCr hydrolysis (ΔHPCr) on pH and temperature when [Mg2+] is 1 mmol L−1. Intracellular pH in resting muscle is typically between 7 and 7.2 which correspond to ΔHPCr values of 35 and 37 kJ mol−1, respectively. In that pH range temperature has little effect on ΔHPCr. Reprinted from Ref. (258) with permission from Elsevier.
Figure 2. Figure 2. Comparison of time courses of enthalpy output (•, heat + work) and the enthalpy output expected from the extent of PCr breakdown (O). The vertical separation of the lines and the shaded area represent the magnitude of unexplained enthalpy. Enthalpy production is normalized by muscle dry weight; in the study from which these records were taken (64), the ratio of wet weight to dry weight was 6.3, therefore 500 mJ (g dry)−1 ≈ 80 mJ (g wet)−1. Reprinted from Ref. (64) with permission from John Wiley Inc., ©1979 The Physiological Society.
Figure 3. Figure 3. (A) Time course of heat production from a rat soleus muscle during and after an isometric tetanus. The main figure shows the complete time course of heat production associated with a 6 s tetanus and includes the initial heat output (qi), produced during the contraction, and recovery heat output (qR), produced mainly during the 4 min following the contraction. The inset shows the time course of the initial heat production, the rate of which is constant (∼15 mW g−1), and the force output during stimulation. The rate of recovery heat production reached a maximum ∼30 s after the start of stimulation. (B) Magnitude of PCr breakdown, determined using NMR, during and after 9 s of isometric contraction. For the first 80 s after the end of stimulation the change in [PCr] follows an exponential time course as PCr is regenerated (inset to Fig. 3B). At ∼100 s after the end of stimulation further PCr splitting occurred. Note that PCr concentration was normalized by muscle wet weight and heat output was normalized by dry weight; the ratio of wet weight to dry wet was taken to be 5 in that study. Experiment details: temperature, 20°C; stimulus frequency/duration, 35 Hz/ 6 s or 9 s. Both figures from Ref. (200) with permission from John Wiley Inc., ©1993 The Physiological Society.
Figure 4. Figure 4. Comparison of the measured and predicted time courses of heat production from a rat soleus muscle during and after an isometric tetanus. The upper line is the measured heat output and the lower line the heat output predicted from the extent of PCr breakdown, measured using NMR, and assuming ΔHPCr = 36 kJ mol−1. Throughout the recording the measured heat is ∼40% greater than that expected on the basis of the measured biochemical change. This indicates that an unknown process is contributing to the initial heat output. Adapted from Ref. (200) with permission from John Wiley Inc., ©1993 The Physiological Society.
Figure 5. Figure 5. Isometric ATP turnover of mammalian muscle and dependencies on temperature and fiber type. (A) ATP turnover in intact preparations of rat and mouse muscle with ATP turnover expressed relative to whole muscle mass; rates include ATP use by cross‐bridge cycling and ion pumps. Open symbols, slow‐twitch muscles; closed symbols, fast‐twitch muscles. Symbols labeled “R” are from rat muscle; other data points for mouse muscle. Points connected by lines were from the same study of temperature dependence of isometric energy turnover (24). Numeric values are given in Table 4. (B) ATP turnover from skinned fibers from rat, rabbit, and human muscle. Labels adjacent to symbols indicate fiber type. Note the difference in the y‐scales in A and B. The inset shows data from rabbit fibers with a fluorescent Pi indicator to determine ATP turnover (107). The rates of ATP breakdown measured in that experiment were an order of magnitude greater than the data from the other skinned fiber experiments shown in the figure but are similar to those for intact fibers. Numeric values given in Table 7.
Figure 6. Figure 6. Records of the time courses of force production (top) and heat output (bottom) of fast‐twitch EDL (black lines) and slow‐twitch soleus (gray lines) muscles from the mouse during isometric contraction. The peak force outputs from the two muscles were similar but the rate of heat output from the fast‐twitch muscle (average, 170 ± 10.5 mW g−1) was five times that from the slow‐twitch muscle (31.8 ± 2.6 mW g−1). These rates correspond to ATP turnovers of 4.7 and 0.88 μmol g−1 s−1 for EDL and soleus, respectively, assuming ΔHPCr was 36 mJ μmol−1 (258). Experiment details: temperature, 20°C; stimulus frequency/duration, EDL, 70 Hz/1 s; Soleus, 40 Hz/1.5 s. Adapted from Ref. (168) with permission from the American Physiological Society.
Figure 7. Figure 7. Measurement of activation heat in frog semitendinosus muscle at 0°C. The upper panel shows records of twitch force and heat production at 4 different muscle lengths from l0 (a) and in progressive steps to 1.5 × l0 in d. The graph shows heat output plotted as a function of twitch force for multiple measurements at different lengths >l0 for one muscle. Note that force decreased as muscle length was increased. The intercept on the y‐axis is the activation heat. Adapted from Ref. (129) with permission from John Wiley Inc., ©1972 The Physiological Society.
Figure 8. Figure 8. The dependence of force output and rate of ATP hydrolysis on sarcomere length. Data from skinned fibers from rabbit psoas muscle at 10°C. The only ATP‐consuming process in the fibers was cross‐bridge cycling. Therefore, when filament overlap was reduced by increasing sarcomere length above 2.5 μm force output and rate of ATP splitting were reduced to the same extent and in proportion to the degree of reduction in overlap. At short sarcomere lengths (i.e., <2.5 μm), force output declined markedly but the decline in rate of ATP utilization was much smaller. Adapted from Ref. (111) with permission from John Wiley Inc., ©2001 The Physiological Society.
Figure 9. Figure 9. (A) Records of force (top), change in sarcomere length (middle), and heat output (bottom) from a frog (R. pipiens) sartorius muscle at 0°C during a contraction containing a period of shortening. Isovelocity shortening at 0.8 l0 s−1 started after 0.75 s isometric contraction (indicated by dashed line). The amplitude of shortening was ∼15% l0. During shortening, force output is lower than the isometric value and rate of heat output is higher (compare record labeled e (contraction with shortening) to record f (from an isometric contraction). The spikes on the heat records are artifacts arising from the stimulus pulses. Muscle length 11.9 mm, mass 77.8 mg. Reprinted from Ref. (123) with permission from John Wiley Inc., ©1983 The Physiological Society. (B) Shortening velocity dependence of power and enthalpy outputs. Power output (solid symbols) and rate of enthalpy output (open symbols) from frog muscle at 0°C. Figure drawn using data taken from Ref. (114) with permission from the Royal Society.
Figure 10. Figure 10. The shortening velocity dependence of enthalpy efficiency of frog muscle at 0°C. Enthalpy efficiency is ratio of work output to enthalpy output. In the example shown maximum efficiency was 0.43 and occurred at a shortening velocity of ∼20% vmax. The inset shows the same efficiency data plotted as a function of afterload. Figure adapted from Fig. 1 in Ref. (115) with permission of the Royal Society.
Figure 11. Figure 11. The shortening velocity dependencies of power output, enthalpy output, and efficiency during steady shortening of fast‐ and slow‐twitch mouse muscles. Data are shown for mouse EDL (•) and soleus (O) muscles at 30°C. (A) Peak power output of soleus was lower and occurred at a lower absolute shortening velocity than for EDL. Mean vmax: soleus, 5.2 l0 s−1; EDL, 8.9 l0 s−1. (B) Rate of enthalpy output was higher at all shortening velocities in EDL than soleus. At velocities >50% vmax, rate of enthalpy output declined with increasing velocity. (C) Enthalpy efficiency was calculated as w·/(w·+h·). Maximum efficiency of soleus (mean ± S.E, 0.46 ± 0.01, n = 5) was higher than that of EDL (0.30 ± 0.01, n = 5) but efficiency of EDL was close to its maximum value across a broad velocity range. Adapted from Ref. (24) with permission from John Wiley Inc., ©2010 The Physiological Society.
Figure 12. Figure 12. Records of force output and heat output from a frog single muscle fiber at 1°C during a contraction containing a period of constant velocity lengthening. Lengthening (indicated by upper trace) started after 0.5 s isometric contraction. During lengthening, the force output was ∼50% greater than the isometric force and the rate of heat output was greater than that during the isometric phase. While lengthening, the rate of enthalpy change (heat + work) was negative; that is, energy was being absorbed so not all of the work being done on the muscle (lower record) appeared as heat. Adapted from Ref. (175) with permission from John Wiley & Sons Inc., Copyright ©2014 The Physiological Society.
Figure 13. Figure 13. Variations in force, fraction of cross‐bridges attached and cross‐bridge strain with velocity for frog muscle at 0°C. Fraction of cross‐bridges attached and cross‐bridge strain calculated from fiber stiffness measurements corrected for filament compliance. For each variable, values are expressed as a percentage of the value during isometric contraction. For shortening velocities between 0 and ∼50% vmax and for most lengthening velocities, force output (•) reflects variation in only the number of attached cross‐bridges (□); that is, variation in the number of attached cross‐bridges determines muscle force output across most of the force‐velocity relationship (202). Only at high shortening velocities was there a decrease in average cross‐bridge strain (▵). The increase in force during lengthening corresponds to a proportional increase in the number of attached cross‐bridges, consistent with the idea that the second cross‐bridge of each pair attaches during lengthening (35). Reprinted from Ref. (23) with permission from Elsevier.
Figure 14. Figure 14. (A) Records of muscle length, force output, and energy output during a cyclic contraction protocol. Records from a fiber bundle from mouse soleus muscle at 31°C. The period of stimulation in each cycle is indicated by the vertical dotted lines. Note that more than half the work performed in each cycle (middle record, lower panel) is performed after the end of stimulation. It is proposed that performance of work during relaxation increases efficiency. (B) Variations in power output, heat output, and rate of enthalpy output (labeled “Total”) with cycle frequency. Rate of heat output did not vary with contraction frequency. Maximum enthalpy efficiency (power/rate of enthalpy output) was 0.52 at a frequency of 3 Hz. Reprinted from Ref. (10) with permission from the Company of Biologists.
Figure 15. Figure 15. Temperature dependence of the apparent equilibrium constant for creatine kinase (K'). There is a linear dependence of log10(K') on 1/temperature. Temperatures in °C are shown alongside each point and absolute values of K' are shown adjacent to the vertical axis. Adapted, with permission, from Ref. (233) with permission from The American Society for Biochemistry and Molecular Biology.
Figure 16. Figure 16. T2 curves can be used to estimate the maximum work a cross‐bridge can perform in one attachment cycle. (A) T2 force is the force reached after the rapid force recovery following application of a fast, small change in fiber length to a contracting muscle. The upper record shows the change in fiber length and the lower record the force response. The quick recovery of force to the T2 level is thought to be due to redevelopment of force by cross‐bridges attached to the thin filament when the step was applied. Adapted from Fig. 6B of Ref. (81) with permission from John Wiley Inc., ©1977 The Physiological Society. (B) T2 forces for muscle fibers from frog, dogfish and rat. T2 forces are expressed relative to maximum isometric force (P0). Forces are plotted against the amplitude of the fiber length change, after correction for filament compliance; that is, the x‐axis indicates the step amplitude transmitted to the attached cross‐bridges (ΔL). To calculate ΔL, filament compliance was taken to be 0.012 μm kPa−1 h−1 for frog fibers (23), 0.017 μm kPa−1 h−1 for dogfish fibers (197) and 0.015 μm kPa−1 h−1 for rat fibers (172). The solid curve is a third‐order polynomial fitted through all the data. The area under that curve between ΔL of +2.75 nm h−1 and −11 nm h−1 was taken to be the maximum work that a cross‐bridge could perform in one attachment cycle; that area is 9.95 T0 nm. Adapted, with permission, from Ref. (24) with permission from John Wiley Inc., ©2010 The Physiological Society.


Figure 1. The dependence of molar enthalpy change for PCr hydrolysis (ΔHPCr) on pH and temperature when [Mg2+] is 1 mmol L−1. Intracellular pH in resting muscle is typically between 7 and 7.2 which correspond to ΔHPCr values of 35 and 37 kJ mol−1, respectively. In that pH range temperature has little effect on ΔHPCr. Reprinted from Ref. (258) with permission from Elsevier.


Figure 2. Comparison of time courses of enthalpy output (•, heat + work) and the enthalpy output expected from the extent of PCr breakdown (O). The vertical separation of the lines and the shaded area represent the magnitude of unexplained enthalpy. Enthalpy production is normalized by muscle dry weight; in the study from which these records were taken (64), the ratio of wet weight to dry weight was 6.3, therefore 500 mJ (g dry)−1 ≈ 80 mJ (g wet)−1. Reprinted from Ref. (64) with permission from John Wiley Inc., ©1979 The Physiological Society.


Figure 3. (A) Time course of heat production from a rat soleus muscle during and after an isometric tetanus. The main figure shows the complete time course of heat production associated with a 6 s tetanus and includes the initial heat output (qi), produced during the contraction, and recovery heat output (qR), produced mainly during the 4 min following the contraction. The inset shows the time course of the initial heat production, the rate of which is constant (∼15 mW g−1), and the force output during stimulation. The rate of recovery heat production reached a maximum ∼30 s after the start of stimulation. (B) Magnitude of PCr breakdown, determined using NMR, during and after 9 s of isometric contraction. For the first 80 s after the end of stimulation the change in [PCr] follows an exponential time course as PCr is regenerated (inset to Fig. 3B). At ∼100 s after the end of stimulation further PCr splitting occurred. Note that PCr concentration was normalized by muscle wet weight and heat output was normalized by dry weight; the ratio of wet weight to dry wet was taken to be 5 in that study. Experiment details: temperature, 20°C; stimulus frequency/duration, 35 Hz/ 6 s or 9 s. Both figures from Ref. (200) with permission from John Wiley Inc., ©1993 The Physiological Society.


Figure 4. Comparison of the measured and predicted time courses of heat production from a rat soleus muscle during and after an isometric tetanus. The upper line is the measured heat output and the lower line the heat output predicted from the extent of PCr breakdown, measured using NMR, and assuming ΔHPCr = 36 kJ mol−1. Throughout the recording the measured heat is ∼40% greater than that expected on the basis of the measured biochemical change. This indicates that an unknown process is contributing to the initial heat output. Adapted from Ref. (200) with permission from John Wiley Inc., ©1993 The Physiological Society.


Figure 5. Isometric ATP turnover of mammalian muscle and dependencies on temperature and fiber type. (A) ATP turnover in intact preparations of rat and mouse muscle with ATP turnover expressed relative to whole muscle mass; rates include ATP use by cross‐bridge cycling and ion pumps. Open symbols, slow‐twitch muscles; closed symbols, fast‐twitch muscles. Symbols labeled “R” are from rat muscle; other data points for mouse muscle. Points connected by lines were from the same study of temperature dependence of isometric energy turnover (24). Numeric values are given in Table 4. (B) ATP turnover from skinned fibers from rat, rabbit, and human muscle. Labels adjacent to symbols indicate fiber type. Note the difference in the y‐scales in A and B. The inset shows data from rabbit fibers with a fluorescent Pi indicator to determine ATP turnover (107). The rates of ATP breakdown measured in that experiment were an order of magnitude greater than the data from the other skinned fiber experiments shown in the figure but are similar to those for intact fibers. Numeric values given in Table 7.


Figure 6. Records of the time courses of force production (top) and heat output (bottom) of fast‐twitch EDL (black lines) and slow‐twitch soleus (gray lines) muscles from the mouse during isometric contraction. The peak force outputs from the two muscles were similar but the rate of heat output from the fast‐twitch muscle (average, 170 ± 10.5 mW g−1) was five times that from the slow‐twitch muscle (31.8 ± 2.6 mW g−1). These rates correspond to ATP turnovers of 4.7 and 0.88 μmol g−1 s−1 for EDL and soleus, respectively, assuming ΔHPCr was 36 mJ μmol−1 (258). Experiment details: temperature, 20°C; stimulus frequency/duration, EDL, 70 Hz/1 s; Soleus, 40 Hz/1.5 s. Adapted from Ref. (168) with permission from the American Physiological Society.


Figure 7. Measurement of activation heat in frog semitendinosus muscle at 0°C. The upper panel shows records of twitch force and heat production at 4 different muscle lengths from l0 (a) and in progressive steps to 1.5 × l0 in d. The graph shows heat output plotted as a function of twitch force for multiple measurements at different lengths >l0 for one muscle. Note that force decreased as muscle length was increased. The intercept on the y‐axis is the activation heat. Adapted from Ref. (129) with permission from John Wiley Inc., ©1972 The Physiological Society.


Figure 8. The dependence of force output and rate of ATP hydrolysis on sarcomere length. Data from skinned fibers from rabbit psoas muscle at 10°C. The only ATP‐consuming process in the fibers was cross‐bridge cycling. Therefore, when filament overlap was reduced by increasing sarcomere length above 2.5 μm force output and rate of ATP splitting were reduced to the same extent and in proportion to the degree of reduction in overlap. At short sarcomere lengths (i.e., <2.5 μm), force output declined markedly but the decline in rate of ATP utilization was much smaller. Adapted from Ref. (111) with permission from John Wiley Inc., ©2001 The Physiological Society.


Figure 9. (A) Records of force (top), change in sarcomere length (middle), and heat output (bottom) from a frog (R. pipiens) sartorius muscle at 0°C during a contraction containing a period of shortening. Isovelocity shortening at 0.8 l0 s−1 started after 0.75 s isometric contraction (indicated by dashed line). The amplitude of shortening was ∼15% l0. During shortening, force output is lower than the isometric value and rate of heat output is higher (compare record labeled e (contraction with shortening) to record f (from an isometric contraction). The spikes on the heat records are artifacts arising from the stimulus pulses. Muscle length 11.9 mm, mass 77.8 mg. Reprinted from Ref. (123) with permission from John Wiley Inc., ©1983 The Physiological Society. (B) Shortening velocity dependence of power and enthalpy outputs. Power output (solid symbols) and rate of enthalpy output (open symbols) from frog muscle at 0°C. Figure drawn using data taken from Ref. (114) with permission from the Royal Society.


Figure 10. The shortening velocity dependence of enthalpy efficiency of frog muscle at 0°C. Enthalpy efficiency is ratio of work output to enthalpy output. In the example shown maximum efficiency was 0.43 and occurred at a shortening velocity of ∼20% vmax. The inset shows the same efficiency data plotted as a function of afterload. Figure adapted from Fig. 1 in Ref. (115) with permission of the Royal Society.


Figure 11. The shortening velocity dependencies of power output, enthalpy output, and efficiency during steady shortening of fast‐ and slow‐twitch mouse muscles. Data are shown for mouse EDL (•) and soleus (O) muscles at 30°C. (A) Peak power output of soleus was lower and occurred at a lower absolute shortening velocity than for EDL. Mean vmax: soleus, 5.2 l0 s−1; EDL, 8.9 l0 s−1. (B) Rate of enthalpy output was higher at all shortening velocities in EDL than soleus. At velocities >50% vmax, rate of enthalpy output declined with increasing velocity. (C) Enthalpy efficiency was calculated as w·/(w·+h·). Maximum efficiency of soleus (mean ± S.E, 0.46 ± 0.01, n = 5) was higher than that of EDL (0.30 ± 0.01, n = 5) but efficiency of EDL was close to its maximum value across a broad velocity range. Adapted from Ref. (24) with permission from John Wiley Inc., ©2010 The Physiological Society.


Figure 12. Records of force output and heat output from a frog single muscle fiber at 1°C during a contraction containing a period of constant velocity lengthening. Lengthening (indicated by upper trace) started after 0.5 s isometric contraction. During lengthening, the force output was ∼50% greater than the isometric force and the rate of heat output was greater than that during the isometric phase. While lengthening, the rate of enthalpy change (heat + work) was negative; that is, energy was being absorbed so not all of the work being done on the muscle (lower record) appeared as heat. Adapted from Ref. (175) with permission from John Wiley & Sons Inc., Copyright ©2014 The Physiological Society.


Figure 13. Variations in force, fraction of cross‐bridges attached and cross‐bridge strain with velocity for frog muscle at 0°C. Fraction of cross‐bridges attached and cross‐bridge strain calculated from fiber stiffness measurements corrected for filament compliance. For each variable, values are expressed as a percentage of the value during isometric contraction. For shortening velocities between 0 and ∼50% vmax and for most lengthening velocities, force output (•) reflects variation in only the number of attached cross‐bridges (□); that is, variation in the number of attached cross‐bridges determines muscle force output across most of the force‐velocity relationship (202). Only at high shortening velocities was there a decrease in average cross‐bridge strain (▵). The increase in force during lengthening corresponds to a proportional increase in the number of attached cross‐bridges, consistent with the idea that the second cross‐bridge of each pair attaches during lengthening (35). Reprinted from Ref. (23) with permission from Elsevier.


Figure 14. (A) Records of muscle length, force output, and energy output during a cyclic contraction protocol. Records from a fiber bundle from mouse soleus muscle at 31°C. The period of stimulation in each cycle is indicated by the vertical dotted lines. Note that more than half the work performed in each cycle (middle record, lower panel) is performed after the end of stimulation. It is proposed that performance of work during relaxation increases efficiency. (B) Variations in power output, heat output, and rate of enthalpy output (labeled “Total”) with cycle frequency. Rate of heat output did not vary with contraction frequency. Maximum enthalpy efficiency (power/rate of enthalpy output) was 0.52 at a frequency of 3 Hz. Reprinted from Ref. (10) with permission from the Company of Biologists.


Figure 15. Temperature dependence of the apparent equilibrium constant for creatine kinase (K'). There is a linear dependence of log10(K') on 1/temperature. Temperatures in °C are shown alongside each point and absolute values of K' are shown adjacent to the vertical axis. Adapted, with permission, from Ref. (233) with permission from The American Society for Biochemistry and Molecular Biology.


Figure 16. T2 curves can be used to estimate the maximum work a cross‐bridge can perform in one attachment cycle. (A) T2 force is the force reached after the rapid force recovery following application of a fast, small change in fiber length to a contracting muscle. The upper record shows the change in fiber length and the lower record the force response. The quick recovery of force to the T2 level is thought to be due to redevelopment of force by cross‐bridges attached to the thin filament when the step was applied. Adapted from Fig. 6B of Ref. (81) with permission from John Wiley Inc., ©1977 The Physiological Society. (B) T2 forces for muscle fibers from frog, dogfish and rat. T2 forces are expressed relative to maximum isometric force (P0). Forces are plotted against the amplitude of the fiber length change, after correction for filament compliance; that is, the x‐axis indicates the step amplitude transmitted to the attached cross‐bridges (ΔL). To calculate ΔL, filament compliance was taken to be 0.012 μm kPa−1 h−1 for frog fibers (23), 0.017 μm kPa−1 h−1 for dogfish fibers (197) and 0.015 μm kPa−1 h−1 for rat fibers (172). The solid curve is a third‐order polynomial fitted through all the data. The area under that curve between ΔL of +2.75 nm h−1 and −11 nm h−1 was taken to be the maximum work that a cross‐bridge could perform in one attachment cycle; that area is 9.95 T0 nm. Adapted, with permission, from Ref. (24) with permission from John Wiley Inc., ©2010 The Physiological Society.
References
 1. Abbott BC , Howarth JV . Heat studies in excitable tissues. Physiol Rev 53: 120‐158, 1973.
 2. Alberty RA . Calculation of standard transformed formation properties of biochemical reactants and standard apparent reduction potentials of half reactions. Arch Biochem Biophys 358: 25‐39, 1998.
 3. Alberty RA . Thermodynamics of the hydrolysis of adenosine triphosphate as a function of temperature, pH, pMg, and ionic strength. J Phys Chem B 107: 12324‐12330, 2003.
 4. Amara CE , Shankland EG , Jubrias SA , Marcinek DJ , Kushmerick MJ , Conley KE . Mild mitochondrial uncoupling impacts cellular aging in human muscles in vivo. Proc Natl Acad Sci U S A 104: 1057‐1062, 2007.
 5. Ardevol A , Adan C , Remesar X , Fernandez‐Lopez JA , Alemany M . Hind leg heat balance in obese Zucker rats during exercise. Pflügers Arch 435: 454‐464, 1998.
 6. Ariano MA , Armstrong RB , Edgerton VR . Hindlimb muscle fiber populations of five mammals. J Histochem Cytochem 21: 51‐55, 1973.
 7. Ashley CC , Moisescu DG . Effect of changing the composition of the bathing solutions upon the isometric tension‐pCa relationship in bundles of crustacean myofibrils. J Physiol 270: 627‐652, 1977.
 8. Askew GN , Tregear RT , Ellington CP . The scaling of myofibrillar actomyosin ATPase activity in apid bee flight muscle in relation to hovering flight energetics. J Exp Biol 213: 1195‐1206, 2010.
 9. Aubert X , Gilbert SH . Variation in the isometric maintenance heat rate with muscle length near that of maximum tension in frog striated muscle. J Physiol 303: 1‐8, 1980.
 10. Barclay CJ . Efficiency of fast‐ and slow‐twitch muscles of the mouse performing cyclic contractions. J Exp Biol 193: 65‐78, 1994.
 11. Barclay CJ . Mechanical efficiency and fatigue of fast and slow muscles of the mouse. J Physiol 497: 781‐794, 1996.
 12. Barclay CJ . Initial mechanical efficiency in cyclic contractions of mouse skeletal muscle. J Appl Biomech 13: 418‐422, 1997.
 13. Barclay CJ . Estimation of cross‐bridge stiffness from maximum thermodynamic efficiency. J Muscle Res Cell Motil 19: 855‐864, 1998.
 14. Barclay CJ . A weakly coupled version of the Huxley crossbridge model can simulate energetics of amphibian and mammalian skeletal muscle. J Muscle Res Cell Motil 20: 163‐176, 1999.
 15. Barclay CJ . Modelling diffusive O2 supply to isolated preparations of mammalian skeletal and cardiac muscle. J Muscle Res Cell Motil 26: 225‐235, 2005.
 16. Barclay CJ . Quantifying Ca2+ release and inactivation of Ca2+ release in fast‐ and slow‐twitch muscles. J Physiol 590: 6199‐6212, 2012.
 17. Barclay CJ , Arnold PD , Gibbs CL . Fatigue and heat production in repeated contractions of mouse skeletal muscle. J Physiol 488: 741‐752, 1995.
 18. Barclay CJ , Constable JK , Gibbs CL . Energetics of fast‐ and slow‐twitch muscles of the mouse. J Physiol 472: 61‐80, 1993.
 19. Barclay CJ , Curtin NA , Woledge RC . Changes in crossbridge and non‐crossbridge energetics during moderate fatigue of frog muscle fibres. J Physiol 468: 543‐556, 1993.
 20. Barclay CJ , Lichtwark GA , Curtin NA . The energetic cost of activation in mouse fast‐twitch muscle is the same whether measured using reduced filament overlap or N‐benzyl‐p‐toluenesulphonamide. Acta Physiol 193: 381‐391, 2008.
 21. Barclay CJ , Weber CL . Slow skeletal muscles of the mouse have greater initial efficiency than fast muscles but the same net efficiency. J Physiol 559: 519‐533, 2004.
 22. Barclay CJ , Woledge RC , Curtin NA . Energy turnover for Ca2+ cycling in skeletal muscle. J Muscle Res Cell Motil 28: 259‐274, 2007.
 23. Barclay CJ , Woledge RC , Curtin NA . Inferring crossbridge properties from skeletal muscle energetics. Prog Biophys Mol Biol 102: 53‐71, 2010.
 24. Barclay CJ , Woledge RC , Curtin NA . Is the efficiency of mammalian (mouse) skeletal muscle temperature dependent? J Physiol 588: 3819‐3831, 2010.
 25. Barclay JK . Energetics of mammalian fast and slow twitch muscles. Fed Proc 41: 189‐191, 1982.
 26. Barsotti RJ , Butler TM . Chemical energy usage and myosin light chain phosphorylation in mammalian skeletal muscle. J Muscle Res Cell Motil 5: 45‐64, 1984.
 27. Baskin RJ . The variation of muscle oxygen consumption with velocity of shortening. J Gen Physiol 49: 9‐15, 1965.
 28. Baylor SM , Chandler WK , Marshall MW . Optical measurements of intracellular pH and magnesium in frog skeletal muscle fibres. J Physiol 331: 105‐137, 1982.
 29. Baylor SM , Chandler WK , Marshall MW . Sarcoplasmic reticulum calcium release in frog skeletal muscle fibres estimated from Arsenazo III calcium transients. J Physiol 344: 625‐666, 1983.
 30. Baylor SM , Hollingworth S . Sarcoplasmic reticulum calcium release compared in slow‐twitch and fast‐twitch fibres of mouse muscle. J Physiol 551: 125‐138, 2003.
 31. Benzinger T , Kitzinger C , Hems R , Burton K . Free‐energy changes of the glutaminase reaction and the hydrolysis of the terminal pyrophosphate bond of adenosine triphosphate. Biochem J 71: 400‐407, 1959.
 32. Block BA . Thermogenesis in muscle. Annu Rev Physiol 56: 535‐577, 1994.
 33. Bottinelli R , Canepari M , Reggiani C , Stienen G . Myofibrillar ATPase activity during isometric contraction and isomyosin composition in rat single skinned muscle fibres. J Physiol 481: 663‐675, 1994.
 34. Brooks SV , Faulkner JA , McCubbrey DA . Power outputs of slow and fast skeletal muscles of mice. J Appl Physiol 68: 1282‐1285, 1990.
 35. Brunello E , Reconditi M , Elangovan R , Linari M , Sun YB , Narayanan T , Panine P , Piazzesi G , Irving M , Lombardi V . Skeletal muscle resists stretch by rapid binding of the second motor domain of myosin to actin. Proc Natl Acad Sci U S A 104: 20114‐20119, 2007.
 36. Burchfield DM , Rall JA . Temperature dependence of the crossbridge cycle during unloaded shortening and maximum isometric tetanus in frog skeletal muscle. J Muscle Res Cell Motil 7: 320‐326, 1986.
 37. Buschman H , van der Laarse W , Stienen G , Elzinga G . Force‐dependent and force‐independent heat production in single slow‐ and fast‐twitch muscle fibres from Xenopus laevis . J Physiol 496: 503‐519, 1996.
 38. Buschman HPJ , Linari M , Elzinga G , Woledge RC . Mechanical and energy characteristics during shortening in isolated type‐1 muscle fibres from Xenopus laevis studied at maximal and submaximal activation. Pflügers Arch 435: 145‐150, 1997.
 39. Cain DF , Davies RE . Breakdown of adenosine triphosphate during a single contraction of working muscle. Biochem Biophys Res Commun 8: 361‐366, 1962.
 40. Cain DF , Infante AA , Davies RE . Chemistry of muscle contraction. Adenosine triphosphate and phosphorylcreatine as energy supplies for single contractions of working muscle. Nature 196: 214‐217, 1962.
 41. Cain SM , Davies RE . Rapid arrest of metabolism with melting freon. In: Chance B , Eisenhardt RH , Gibson QH , Lonberg‐Holm KK , editors. Rapid Mixing and Sampling Techniques in Biochemistry. New York: Academic Press, 1964, pp. 229‐237.
 42. Canfield P , Lebacq J , Marechal G . Energy balance in frog sartorius muscle during an isometric tetanus at 20°C. J Physiol 232: 467‐483, 1973.
 43. Carey FG . Fishes with warm bodies. Sci Am 228: 36‐44, 1973.
 44. Carlson FD , Hardy D , Wilkie DR . The relation between heat produced and phosphorylcreatine split during isometric contraction of frog's muscle. J Physiol 189: 209‐235, 1967.
 45. Chapman JB . Fluorometric studies of oxidative metabolism in isolated papillary muscle of the rabbit. J Gen Physiol 59: 135‐154, 1972.
 46. Chapman JB , Gibbs CL , Loiselle DS . Myothermic, polarographic, and fluorometric data from mammalian muscles: Correlations and an approach to a biochemical synthesis. Fed Proc 41: 176‐184, 1982.
 47. Cheung A , Dantzig JA , Hollingworth S , Baylor SM , Goldman YE , Mitchison TJ , Straight AF . A small‐molecule inhibitor of skeletal muscle myosin II. Nat Cell Biol 4: 83‐88, 2002.
 48. Cieslar J , Huang MT , Dobson GP . Tissue spaces in rat heart, liver, and skeletal muscle in vivo. Am J Physiol Regul Integr Comp Physiol 275: R1530‐1536, 1998.
 49. Clausen T , Overgaard K , Nielsen OB . Evidence that the Na+‐K+ leak/pump ratio contributes to the difference in endurance between fast‐ and slow‐twitch muscles. Acta Physiol Scand 180: 209‐216, 2004.
 50. Close R . Force: Velocity properties of mouse muscles. Nature 206: 718‐719, 1965.
 51. Close RI . Dynamic properties of mammalian skeletal muscles. Physiol Rev 52: 129‐197, 1972.
 52. Conley KE , Amara CE , Jubrias SA , Marcinek DJ . Mitochondrial function, fibre types and ageing: New insights from human muscle in vivo . Exp Physiol 92: 333‐339, 2007.
 53. Constable JK , Barclay CJ , Gibbs CL . Energetics of lengthening in mouse and toad skeletal muscles. J Physiol 505: 205‐215, 1997.
 54. Cooke R . Actomyosin interaction in striated muscle. Physiol Rev 77: 671‐697, 1997.
 55. Cooke R . The role of the myosin ATPase activity in adaptive thermogenesis by skeletal muscle. Biophys Rev 3: 33‐45, 2011.
 56. Crabtree B , Nicholson BA . Thermodynamics and metabolism. In: Jones MN , editor. Biochemical Thermodynamics. Amsterdam: Elsevier Scientific, 1988, pp. 347‐395.
 57. Crow MT , Kushmerick MJ . Chemical energetics of slow‐ and fast‐twitch muscles of the mouse. J Gen Physiol 79: 147‐166, 1982.
 58. Curtin NA , Davies RE . Chemical and mechanical changes during stretching of activated frog skeletal muscle. Cold Spring Harbor Symp Quant Biol 37: 619‐626, 1973.
 59. Curtin NA , Davies RE . Very high tension with very little ATP breakdown by active skeletal muscle. J Mechanochem Cell Motil 3: 147‐154, 1975.
 60. Curtin NA , Gilbert C , Kretzschmar KM , Wilkie DR . The effect of the performance of work on total energy output and metabolism during muscular contraction. J Physiol 238: 455‐472, 1974.
 61. Curtin NA , Howarth JV , Rall JA , Wilson MG , Woledge RC . Absolute values of myothermic measurements on single muscle fibres from frog. J Muscle Res Cell Motil 7: 327‐332, 1986.
 62. Curtin NA , Woledge RC . A comparison of the energy balance in two successive isometric tetani of frog muscle. J Physiol 270: 455‐471, 1977.
 63. Curtin NA , Woledge RC . Energy changes and muscular contraction. Physiol Rev 58: 690‐761, 1978.
 64. Curtin NA , Woledge RC . Chemical change and energy production during contraction of frog muscle: How are their time courses related? J Physiol 288: 353‐366, 1979.
 65. Curtin NA , Woledge RC . Effect of muscle length on energy balance in frog skeletal muscle. J Physiol 316: 453‐468, 1981.
 66. Curtin NA , Woledge RC . Efficiency of energy conversion during shortening of muscle fibres from the dogfish Scyliorhinus canicula . J Exp Biol 158: 343‐353, 1991.
 67. Curtin NA , Woledge RC . Efficiency of energy conversion during sinusoidal movement of red muscle fibres from the dogfish Scyliorhinus canicula . J Exp Biol 185: 195‐206, 1993.
 68. Curtin NA , Woledge RC . Efficiency of energy conversion during sinusoidal movement of white muscle fibers from the dogfish Scyliorhinus canicula . J Exp Biol 183: 137‐147, 1993.
 69. Curtin NA , Woledge RC . Power at the expense of efficiency in contraction of white muscle fibres from dogfish Scyliorhinus canicula. J Exp Biol 199: 593‐601, 1996.
 70. Davies RE , Cain D , Delluva AM . The energy supply for muscle contraction. Ann N Y Acad Sci 81: 468‐476, 1959.
 71. Dawson MJ , Gadian DG , Wilkie DR . Contraction and recovery of living muscles studies by 31P nuclear magnetic resonance. J Physiol 267: 703‐735, 1977.
 72. Dawson MJ , Gadian DG , Wilkie DR . Muscular fatigue investigated by phosphorus nuclear magnetic resonance. Nature 274: 861‐866, 1978.
 73. Dawson MJ , Gadian DG , Wilkie DR . Mechanical relaxation rate and metabolism studied in fatiguing muscle by phosphorus nuclear magnetic resonance. J Physiol 299: 465‐484, 1980.
 74. DeFuria RR , Kushmerick MJ . ATP utilization associated with recovery metabolism in anaerobic frog muscle. Am J Physiol Cell Physiol 232: C30‐36, 1977.
 75. Dzeja PP , Terzic A . Phosphotransfer networks and cellular energetics. J Exp Biol 206: 2039‐2047, 2003.
 76. Edwards RH , Hill DK , Jones DA . Heat production and chemical changes during isometric contractions of the human quadriceps muscle. J Physiol 251: 303‐315, 1975.
 77. Elzinga G , Peckham M , Woledge RC . The sarcomere length dependence of the rate of heat production during isometric tetanic contraction of frog muscles. J Physiol 357: 495‐504, 1984.
 78. Fenn WO . A quantitative comparison between the energy liberated and the work performed by isolated sartorius muscle of the frog. J Physiol 58: 175‐203, 1923.
 79. Ferenczi MA , He ZH , Chillingworth RK , Brune M , Corrie JE , Trentham DR , Webb MR . A new method for the time‐resolved measurement of phosphate release in permeabilized muscle fibers. Biophys J 68: 191S‐192S; discussion 192S‐193S, 1995.
 80. Ferenczi MA , Homsher E , Trentham DR . The kinetics of magnesium adenosine triphosphate cleavage in skinned muscle fibres of the rabbit. J Physiol 352: 575‐599, 1984.
 81. Ford LE , Huxley AF , Simmons RM . Tension responses to sudden length change in stimulated frog muscle fibres near slack length. J Physiol 269: 441‐515, 1977.
 82. Galler S , Hilber K . Tension/stiffness ratio of skinned rat skeletal muscle fibre types at various temperatures. Acta Physiol Scand 162: 119‐126, 1998.
 83. Galler S , Hilber K , Pette D . Force responses following stepwise length changes of rat skeletal muscle fibre types. J Physiol 493: 219‐227, 1996.
 84. Gibbs CL , Barclay CJ . Efficiency of skeletal and cardiac muscle. Adv Exp Med Biol 453: 527‐535, 1998.
 85. Gibbs CL , Gibson WR . Energy production of rat soleus muscle. Am J Physiol 223: 864‐871, 1972.
 86. Gilbert C , Kretzschmar KM , Wilkie DR , Woledge RC . Chemical change and energy output during muscular contraction. J Physiol 218: 163‐193, 1971.
 87. Gilbert C , Kretzschmar KM , Wilkie DR , Woledge RC . Chemical change and energy output during muscular contraction. J Physiol 218: 163‐193, 1971.
 88. Gilbert SH , Ford LE . Heat changes during transient tension responses to small releases in active frog muscle. Biophys J 54: 611‐617, 1988.
 89. Gilbert SH , Mathias RT . Analysis of diffusion delay in a layered medium. Application to heat measurements from muscle. Biophys J 54: 603‐610, 1988.
 90. Gillis GB , Biewener AA . Hindlimb muscle function in relation to speed and gait: In vivo patterns of strain and activation in a hip and knee extensor of the rat (Rattus norvegicus). J Exp Biol 204: 2717‐2731, 2001.
 91. Gillis JM . Relaxation of vertebrate skeletal muscle. A synthesis of the biochemical and physiological approaches. Biochim Biophys Acta 811: 97‐145, 1985.
 92. Gillis JM , Thomason D , Lefevre J , Kretsinger RH . Parvalbumins and muscle relaxation: A computer simulation study. J Muscle Res Cell Motil 3: 377‐398, 1982.
 93. Glaser BW , You G , Zhang M , Medler S . Relative proportions of hybrid fibres are unaffected by 6 weeks of running exercise in mouse skeletal muscles. Exp Physiol 95: 211‐221, 2010.
 94. Glyn H , Sleep J . Dependence of adenosine triphosphatase activity of rabbit psoas muscle fibres and myofibrils on substrate concentration. J Physiol 365: 259‐276, 1985.
 95. Golding EM , Teague WE , Dobson GP . Adjustment of K' to varying pH and pMg for the creatine kinase, adenylate kinase and ATP hydrolysis equilibria permitting quantitative bioenergetic assessment. J Exp Biol 198: 1775‐1782, 1995.
 96. Goldspink G . An attempt at estimating extrafiber fluid in small skeletal muscles by a simple physical method. Can J Physiol Pharmacol 44: 765‐775, 1966.
 97. Gonzalez‐Alonso J , Quistorff B , Krustrup P , Bangsbo J , Saltin B . Heat production in human skeletal muscle at the onset of intense dynamic exercise. J Physiol 524(Pt 2): 603‐615, 2000.
 98. Gordon AM , Huxley AF , Julian FJ . The variation in isometric tension with sarcomere length in vertebrate muscle fibres. J Physiol 184: 170‐192, 1966.
 99. Gower D , Kretzschmar KM . Heat production and chemical change during isometric contraction of rat soleus muscle. J Physiol 258: 659‐671, 1976.
 100. Guzun R , Kaambre T , Bagur R , Grichine A , Usson Y , Varikmaa M , Anmann T , Tepp K , Timohhina N , Shevchuk I , Chekulayev V , Boucher F , Dos Santos P , Schlattner U , Wallimann T , Kuznetsov AV , Dzeja P , Aliev M , Saks V . Modular organization of cardiac energy metabolism:energy conversion, transfer and feedback regulation. Acta Physiol (Oxf) 213: 84‐106, 2014.
 101. Haiech J , Derancourt J , Pechere JF , Demaille JG . Magnesium and calcium binding to parvalbumins: Evidence for differences between parvalbumins and an explanation of their relaxing function. Biochemistry 18: 2752‐2758, 1979.
 102. Harkins AB , Kurebayashi N , Baylor SM . Resting myoplasmic free calcium in frog skeletal muscle fibers estimated with fluo‐3. Biophys J 65: 865‐881, 1993.
 103. Harrison GJ , van Wijhe MH , de Groot B , Dijk FJ , van Beek JH . CK inhibition accelerates transcytosolic energy signaling during rapid workload steps in isolated rabbit hearts. Am J Physiol Heart Circ Physiol 276: H134‐140, 1999.
 104. Harwood CL , Young IS , Tikunov BA , Hollingworth S , Baylor SM , Rome LC . Paying the piper: The cost of Ca2+ pumping during the mating call of toadfish. J Physiol 589: 5467‐5484, 2011.
 105. Haseler LJ , Richardson RS , Videen JS , Hogan MC . Phosphocreatine hydrolysis during submaximal exercise: The effect of FIO2. J Appl Physiol 85: 1457‐1463, 1998.
 106. He ZH , Bottinelli R , Pellegrino MA , Ferenczi MA , Reggiani C . ATP consumption and efficiency of human single muscle fibers with different myosin isoform composition. Biophys J 79: 945‐961, 2000.
 107. He ZH , Chillingworth RK , Brune M , Corrie JE , Trentham DR , Webb MR , Ferenczi MA . ATPase kinetics on activation of rabbit and frog permeabilized isometric muscle fibres: A real time phosphate assay. J Physiol 501: 125‐148, 1997.
 108. He ZH , Chillingworth RK , Brune M , Corrie JE , Webb MR , Ferenczi MA . The efficiency of contraction in rabbit skeletal muscle fibres, determined from the rate of release of inorganic phosphate. J Physiol 517: 839‐854, 1999.
 109. Heglund NC , Cavagna GA . Mechanical work, oxygen consumption, and efficiency in isolated frog and rat muscle. Am J Physiol 223: C22‐C29, 1987.
 110. Heizmann CW , Berchtold MW , Rowlerson AM . Correlation of parvalbumin concentration with relaxation speed in mammalian muscles. Proc Natl Acad Sci U S A 79: 7243‐7247, 1982.
 111. Hilber K , Sun YB , Irving M . Effects of sarcomere length and temperature on the rate of ATP utilisation by rabbit psoas muscle fibres. J Physiol 531: 771‐780, 2001.
 112. Hill AV . Heat of shortening and the dynamic constants of muscle. Proc R Soc Lond B Biol Sci 126: 136‐195, 1938.
 113. Hill AV . The mechanical efficiency of frog muscle. Proc R Soc Lond B Biol Sci 127: 434‐451, 1939.
 114. Hill AV . The effect of load on the heat of shortening of muscle. Proc R Soc Lond B Biol Sci 159: 297‐318, 1964.
 115. Hill AV . The efficiency of mechanical power development during muscular shortening and its relation to load. Proc R Soc Lond B Biol Sci 159: 319‐324, 1964.
 116. Hill AV. Trails and Trials in Physiology. London: Arnold, 1965.
 117. Hill AV , Howarth JV . The reversal of chemical reactions in contracting muscle during an applied stretch. Proc R Soc Lond B Biol Sci 151: 169‐193, 1959.
 118. Hill AV , Woledge RC . An examination of absolute values in myothermic measurements. J Physiol 162: 311‐333, 1962.
 119. Hill DK . The time course of oxygen consumption of stimulated frog's muscle. J Physiol 98: 207‐227, 1940.
 120. Holroyd SM , Gibbs CL , Luff AR . Shortening heat in slow‐ and fast‐twitch muscles of the rat. Am J Physiol 270: C293‐297, 1996.
 121. Holt NC , Askew GN . The effects of asymmetric length trajectories on the initial mechanical efficiency of mouse soleus muscles. J Exp Biol 215: 324‐330, 2012.
 122. Homsher E . Muscle enthalpy production and its relationship to actomyosin ATPase. Annu Rev Physiol 49: 673‐690, 1987.
 123. Homsher E , Irving M , Lebacq J . The variation in shortening heat with sarcomere length in frog muscle. J Physiol 345: 107‐121, 1983.
 124. Homsher E , Irving M , Wallner A . High‐energy phosphate metabolism and energy liberation associated with rapid shortening in frog skeletal muscle. J Physiol 321: 423‐436, 1981.
 125. Homsher E , Kean CJ . Skeletal muscle energetics and metabolism. Annu Rev Physiol 40: 93‐131, 1978.
 126. Homsher E , Kean CJ . Unexplained enthalpy production in contracting skeletal muscles. Fed Proc 41: 149‐154, 1982.
 127. Homsher E , Kean CJ , Wallner A , Garibian‐Sarian V . The time‐course of energy balance in an isometric tetanus. J Gen Physiol 73: 553‐567, 1979.
 128. Homsher E , Lacktis J , Yamada T , Zohman G . Repriming and reversal of the isometric unexplained enthalpy in frog skeletal muscle. J Physiol 393: 157‐170, 1987.
 129. Homsher E , Mommaerts WF , Ricchiuti NV , Wallner A . Activation heat, activation metabolism and tension‐related heat in frog semitendinosus muscles. J Physiol 220: 601‐625, 1972.
 130. Homsher E , Mommaerts WFHM , Ricchiuti NV . Energetics of shortening muscles in twitches and tetanic contractions: II. Force‐determined shortening heat. J Gen Physiol 62: 677‐692, 1973.
 131. Homsher E , Rall JA , Wallner A , Ricchiuti NV . Energy liberation and chemical change in frog skeletal muscle during single isometric tetanic contractions. J Gen Physiol 65: 1‐21, 1975.
 132. Homsher E , Yamada T , Wallner A , Tsai J . Energy balance studies in frog skeletal muscles shortening at one‐half maximal velocity. J Gen Physiol 84: 347‐359, 1984.
 133. Hou TT , Johnson JD , Rall JA . Parvalbumin content and Ca2+ and Mg2+ dissociation rates correlated with changes in relaxation rate of frog muscle fibres. J Physiol 441: 285‐304, 1991.
 134. Hou TT , Johnson JD , Rall JA . Effect of temperature on relaxation rate and Ca2+, Mg2+ dissociation rates from parvalbumin of frog muscle fibres. J Physiol 449: 399‐410, 1992.
 135. Huxley AF . A note suggesting that the cross‐bridge attachment during muscle contraction may take place in two stages. Proc R Soc Lond B Biol Sci 183: 83‐86, 1973.
 136. Huxley AF . Muscular contraction. J Physiol 243: 1‐43, 1974.
 137. Huxley AF . Biological motors: Energy storage in myosin molecules. Curr Biol 8: R485‐488, 1998.
 138. Huxley AF , Simmons RM . Proposed mechanism of force generation in striated muscle. Nature 233: 533‐538, 1971.
 139. Huxley AF , Tideswell S . Filament compliance and tension transients in muscle. J Muscle Res Cell Motil 17: 507‐511, 1996.
 140. Huxley SA. Relections on Muscle. Liverpool: Liverpool University Press, 1980.
 141. Imaizumi M , Tanokura M , Yamada K . Calorimetric studies on calcium and magnesium binding by troponin C from bullfrog skeletal muscle. J Biochem, Tokyo 107: 127‐132, 1990.
 142. Infante AA , Klaupiks D , Davies RE . Adenosine triphosphate: Changes in muscles doing negative work. Science 144: 1577‐1578, 1964.
 143. Iotti S , Frassineti C , Sabatini A , Vacca A , Barbiroli B . Quantitative mathematical expressions for accurate in vivo assessment of cytosolic [ADP] and DeltaG of ATP hydrolysis in the human brain and skeletal muscle. Biochim Biophys Acta 1708: 164‐177, 2005.
 144. Irving M , Woledge RC . The energy liberation of frog skeletal muscle in tetanic contractions containing two periods of shortening. J Physiol 321: 401‐410, 1981.
 145. Jacobus WE . Theoretical support for the heart phosphocreatine energy transport shuttle based on the intracellular diffusion limited mobility of ADP. Biochem Biophys Res Commun 133: 1035‐1041, 1985.
 146. Josephson RK . Mechanical power output from striated muscle during cyclic contraction. J Exp Biol 114: 493‐512, 1985.
 147. Josephson RK , Malamud JG , Stokes DR . The efficiency of an asynchronous flight muscle from a beetle. J Exp Biol 204: 4125‐4139, 2001.
 148. Kean C , Homsher E , Sarian‐Garibian V . The energy balance of crossbridge cycling in frog skeletal muscle. Biophys J 17: 202a, 1977.
 149. Kemp GJ , Meyerspeer M , Moser E . Absolute quantification of phosphorus metabolite concentrations in human muscle in vivo by 31P MRS: A quantitative review. NMR Biomed 20: 555‐565, 2007.
 150. Kindig CA , Howlett RA , Stary CM , Walsh B , Hogan MC . Effects of acute creatine kinase inhibition on metabolism and tension development in isolated single myocytes. J Appl Physiol 98: 541‐549, 2005.
 151. Klotz IM . Energy Changes in Biochemical Reactions. New York: Academic Press, 1967.
 152. Kotsias BA , Venosa RA . Sodium influx during action potential in innervated and denervated rat skeletal muscles. Muscle Nerve 24: 1026‐1033, 2001.
 153. Kretzschmar KM , Wilkie DR . A new approach to freezing tissues rapidly. J Physiol 202: 66P‐67P, 1969.
 154. Kushmerick MJ . The chemical energetics of muscle contraction. II. The chemistry, efficiency and power of maximally working sartorius muscles: Appendix: Free energy and enthalpy of ATP hydrolysis in the sarcoplasm. Proc R Soc Lond B Biol Sci 174: 348‐353, 1969.
 155. Kushmerick MJ . Energy balance in muscle contraction: A biochemical approach. Curr Top Bioenerg 6: 1‐37, 1978.
 156. Kushmerick MJ . Energetics of muscle contraction. In: Peachey LE , editor. Handbook of Physiology: Skeletal Muscle. Bethesda, Maryland: American Physiological Society, 1983, pp. 198‐236.
 157. Kushmerick MJ , Krasner B . Force and ATPase rate in skinned skeletal muscle fibers. Fed Proc 41: 2232‐2237, 1982.
 158. Kushmerick MJ , Meyer RA . Chemical changes in rat leg muscle by phosphorus nuclear magnetic resonance. Am J Physiol 248: C542‐C549, 1985.
 159. Kushmerick MJ , Meyer RA , Brown TR . Phosphorus NMR spectroscopy of cat biceps and soleus muscles. Adv Exp Med Biol 159: 303‐325, 1983.
 160. Kushmerick MJ , Moerland TS , Wiseman RW . Mammalian skeletal muscle fibers distinguished by contents of phosphocreatine, ATP, and Pi. Proc Natl Acad Sci U S A 89: 7521‐7525, 1992.
 161. Kushmerick MJ , Moerland TS , Wiseman RW . Two classes of mammalian skeletal muscle fibres distinguished by metabolite content. Adv Exp Med Biol 332: 749‐760, 1993.
 162. Kushmerick MJ , Paul RJ . Aerobic recovery metabolism following a single isometric tetanus in frog sartorius muscle at 0 degrees C. J Physiol 254: 693‐709, 1976.
 163. Kushmerick MJ , Paul RJ . Relationship between initial chemical reactions and oxidative recovery metabolism for single isometric contractions of frog sartorius at 0°C. J Physiol 254: 711‐727, 1976.
 164. Kushmerick MJ , Paul RJ . Chemical energetics in repeated contractions of frog sartorius muscles at 0°C. J Physiol 267: 249‐260, 1977.
 165. Lawson JW , Veech RL . Effects of pH and free Mg2+ on the Keq of the creatine kinase reaction and other phosphate hydrolyses and phosphate transfer reactions. J Biol Chem 254: 6528‐6537, 1979.
 166. Lee RS , Tikunova SB , Kline KP , Zot HG , Hasbun JE , Minh NV , Swartz DR , Rall JA , Davis JP . Effect of Ca2+ binding properties of troponin C on rate of skeletal muscle force redevelopment. Am J Physiol Cell Physiol 299: C1091‐1099, 2010.
 167. Leijendekker WJ , Elzinga G . Metabolic recovery of mouse extensor digitorum longus and soleus muscle. Pflügers Arch 416: 22‐27, 1990.
 168. Leijendekker WJ , van Hardeveld C , Elzinga G . Heat production during contraction in skeletal muscle of hypothyroid mice. Am J Physiol Endocrinol Metabol 253: E214‐220, 1987.
 169. Lewis DB , Barclay CJ . Efficiency and cross‐bridge work output of skeletal muscle is decreased at low levels of activation. Pflügers Arch 466: 599‐609, 2014.
 170. Limouze J , Straight AF , Mitchison T , Sellers JR . Specificity of blebbistatin, an inhibitor of myosin II. J Muscle Res Cell Motil 25: 337‐341, 2004.
 171. Linari M , Caremani M , Lombardi V . A kinetic model that explains the effect of inorganic phosphate on the mechanics and energetics of isometric contraction of fast skeletal muscle. Proc R Soc Lond B Biol Sci 277: 19‐27, 2010.
 172. Linari M , Caremani M , Piperio C , Brandt P , Lombardi V . Stiffness and fraction of myosin motors responsible for active force in permeabilized muscle fibers from rabbit psoas. Biophys J 92: 2476‐2490, 2007.
 173. Linari M , Dobbie I , Reconditi M , Koubassova N , Irving M , Piazzesi G , Lombardi V . The stiffness of skeletal muscle in isometric contraction and rigor: The fraction of myosin heads bound to actin. Biophys J 74: 2459‐2473, 1998.
 174. Linari M , Woledge RC . Efficiency during ramp and staircase shortening in intact single fibers from frog muscle. Pflügers Arch 428: R5‐R5, 1994.
 175. Linari M , Woledge RC , Curtin NA . Energy storage during stretch of active single fibres from frog skeletal muscle. J Physiol 548: 461‐474, 2003.
 176. Little SC , Tikunova SB , Norman C , Swartz DR , Davis JP . Measurement of calcium dissociation rates from troponin C in rigor skeletal myofibrils. Front Physiol 2: 1‐12, 2011.
 177. Lombardi V , Piazzesi G . The contractile response during steady lengthening of stimulated frog muscle fibres. J Physiol 431: 141‐171, 1990.
 178. Lou F , Curtin NA , Woledge RC . Activation heat in white muscle fibres isolated from dogfish. J Physiol 494P: P131‐P131, 1996.
 179. Lou F , Curtin NA , Woledge RC . Contraction with shortening during stimulation or during relaxation: How do the energetic costs compare? J Muscle Res Cell Motil 19: 797‐802, 1998.
 180. Lou F , Van der Laarse WJ , Curtin NA , Woledge RC . Heat production and oxygen consumption during metabolic recovery of white muscle fibres from the dogfish Scyliorhinus canicula . J Exp Biol 203: 1201‐1210, 2000.
 181. Luff AR , Atwood HL . Changes in the sarcoplasmic reticulum and transverse tubular system of fast and slow skeletal muscles of the mouse during postnatal development. J Cell Biol 51: 369‐383, 1971.
 182. Mahler M . Diffusion and consumption of oxygen in the resting frog sartorius muscle. J Gen Physiol 71: 533‐557, 1978.
 183. Mahler M . Kinetics of oxygen consumption after a single isometric tetanus of frog sartorius muscle at 20 degrees C. J Gen Physiol 71: 559‐580, 1978.
 184. Mahler M . First‐order kinetics of muscle oxygen consumption, and an equivalent proportionality between QO2 and phosphorylcreatine level. Implications for the control of respiration. J Gen Physiol 86: 135‐165, 1985.
 185. Mahler M , Louy C , Homsher E , Peskoff A . Reappraisal of diffusion, solubility, and consumption of oxygen in frog skeletal muscle, with applications to muscle energy balance. J Gen Physiol 86: 105‐134, 1985.
 186. Makinose M , Hasselbach W . ATP synthesis by the reverse of the sarcoplasmic calcium pump. FEBS Lett 12: 271‐272, 1971.
 187. Meyer RA , Brown TR , Kushmerick MJ . Phosphorus nuclear magnetic resonance of fast‐ and slow‐twitch muscle. Am J Physiol 248: C279‐287, 1985.
 188. Meyer RA , Sweeney HL , Kushmerick MJ . A simple analysis of the “phosphocreatine shuttle”. Am J Physiol 246: C365‐C377, 1984.
 189. Milton RL , Behforouz MA . Na channel density in extrajunctional sarcolemma of fast and slow twitch mouse skeletal muscle fibres: Functional implications and plasticity after fast motoneuron transplantation on to a slow muscle. J Muscle Res Cell Motil 16: 430‐439, 1995.
 190. Mobley BA , Eisenberg BR . Sizes of components in frog skeletal muscle measured by methods of stereology. J Gen Physiol 66: 31‐45, 1975.
 191. Moisescu DG . Activation of isolated bundles of frog myofibrils in Ca‐buffered solutions. J Physiol 263: 161P‐162P, 1976.
 192. Mommaerts WFHM . Muscular Contraction. A Topic in Molecular Physiology. New York: Intersciences Publishers, Inc., 1950.
 193. Mommaerts WFHM . Energetics of muscular contraction. Physiol Rev 49: 427‐508, 1969.
 194. Nicholls DG , Ferguson SJ . Bioenergetics 4. London: Academic Press, 2013.
 195. Nielsen OB , Harrison AP . The regulation of the Na+,K+ pump in contracting skeletal muscle. Acta Physiol Scand 162: 191‐200, 1998.
 196. Ogawa Y . Calcium binding to troponin C and troponin: Effects of Mg2+, ionic strength and pH. J Biochem, Tokyo 97: 1011‐1023, 1985.
 197. Park‐Holohan S , Linari M , Reconditi M , Fusi L , Brunello E , Irving M , Dolfi M , Lombardi V , West TG , Curtin NA , Woledge RC , Piazzesi G . Mechanics of myosin function in white muscle fibres of the dogfish, Scyliorhinus canicula . J Physiol 590: 1973‐1988, 2012.
 198. Paul RJ . Physical and biochemical energy balance during an isometric tetanus and steady state recovery in frog sartorius at 0°C. J Gen Physiol 81: 337‐354, 1983.
 199. Phillips RC , George P , Rutman RJ . Thermodynamic data for the hydrolysis of adenosine triphosphate as a function of pH, Mg2+ ion concentration, and ionic strength. J Biol Chem 244: 3330‐3342, 1969.
 200. Phillips SK , Takei M , Yamada K . The time course of phosphate metabolites and intracellular pH using 31P NMR compared to recovery heat in rat soleus muscle. J Physiol 460: 693‐704, 1993.
 201. Piazzesi G , Francini F , Linari M , Lombardi V . Tension transients during steady lengthening of tetanized muscle fibres of the frog. J Physiol 445: 659‐711, 1992.
 202. Piazzesi G , Reconditi M , Linari M , Lucii L , Bianco P , Brunello E , Decostre V , Stewart A , Gore DB , Irving TC , Irving M , Lombardi V . Skeletal muscle performance determined by modulation of number of myosin motors rather than motor force or stroke size. Cell 131: 784‐795, 2007.
 203. Potma EJ , van Graas IA , Stienen GJ . Influence of inorganic phosphate and pH on ATP utilization in fast and slow skeletal muscle fibers. Biophys J 69: 2580‐2589, 1995.
 204. Potter JD , Gergely J . The calcium and magnesium binding sites on troponin and their role in the regulation of myofibrillar adenosine triphosphatase. J Biol Chem 250: 4628‐4633, 1975.
 205. Potter JD , Hsu FJ , Pownall HJ . Thermodynamics of Ca2+ binding to troponin‐C. J Biol Chem 252: 2452‐2454, 1977.
 206. Rall JA . Relationship of isometric unexplained energy production to parvalbumin content in frog skeletal muscle. Prog Clin Biol Res 315: 117‐126, 1989.
 207. Rall JA , Homsher E , Wallner A , Mommaerts WF . A temporal dissociation of energy liberation and high energy phosphate splitting during shortening in frog skeletal muscles. J Gen Physiol 68: 13‐27, 1976.
 208. Rall JA , Schottelius BA . Energetics of contraction in phasic and tonic skeletal muscles of the chicken. J Gen Physiol 62: 303‐323, 1973.
 209. Rall JA , Woledge RC . Influence of temperature on mechanics and energetics of muscle contraction. Am J Physiol 259: R197‐203, 1990.
 210. Reggiani C , Potma EJ , Bottinelli R , Canepari M , Pellegrino MA , Stienen GJM . Chemo‐mechanical energy transduction in relation to myosin isoform composition in skeletal muscle fibres of the rat. J Physiol 502: 449‐460, 1997.
 211. Rich PR . The molecular machinery of Keilin's respiratory chain. Biochem Soc Trans 31: 1095‐1105, 2003.
 212. Robbins EA , Boyer PD . Determination of the equilibrium of the hexokinase reaction and the free energy of hydrolysis of adenosine triphosphate. J Biol Chem 224: 121‐135, 1957.
 213. Robertson SP , Johnson JD , Potter JD . The time‐course of Ca2+ exchange with calmodulin, troponin, parvalbumin, and myosin in response to transient increases in Ca2+ . Biophys J 34: 559‐569, 1981.
 214. Rome LC , Klimov AA . Superfast contractions without superfast energetics: ATP usage by SR‐Ca2+ pumps and crossbridges in toadfish swimbladder muscle. J Physiol 526: 279‐286, 2000.
 215. Rome LC , Kushmerick MJ . Energetics of isometric contractions as a function of muscle temperature. Am J Physiol 244: C100‐109, 1983.
 216. Rosing J , Slater EC . The value of ΔGo for the hydrolysis of ATP. Biochim Biophys Acta 267: 275‐290, 1972.
 217. Sandberg JA , Carlson FD . The dependence of phosphorylcreatine hydrolysis during an isometric tetanus. Biochem Z 345: 212‐231, 1966.
 218. Segal SS , Faulkner JA . Temperature‐dependent physiological stability of rat skeletal muscle in vitro. Am J Physiol 248: C265‐C270, 1985.
 219. Segal SS , Faulkner JA , White TP . Skeletal muscle fatigue in vitro is temperature dependent. J Appl Physiol 61: 660‐665, 1986.
 220. Semafuko WE , Bowie WC . Papillary muscle dynamics: In situ function and responses of the papillary muscle. Am J Physiol 228: 1800‐1807, 1975.
 221. Sheff MF , Zacks SI . Interstitial space of mouse skeletal muscle. J Physiol 328: 507‐519, 1982.
 222. Smith IC , Bombardier E , Vigna C , Tupling AR . ATP consumption by sarcoplasmic reticulum Ca2+ pumps accounts for 40‐50% of resting metabolic rate in mouse fast and slow twitch skeletal muscle. PloS one 8: e68924, 2013.
 223. Smith ICH . Energetics of activation in frog and toad muscle. J Physiol 220: 583‐599, 1972.
 224. Smith NP , Barclay CJ , Loiselle DS . The efficiency of muscle contraction. Prog Biophys Mol Biol 88: 1‐58, 2005.
 225. Smith SJ , Woledge RC . Thermodynamic analysis of calcium binding to frog parvalbumin. J Muscle Res Cell Motil 6: 757‐768, 1985.
 226. Stainsby WN , Barclay JK . Exercise metabolism: O2 deficit, steady level O2 uptake and O2 uptake for recovery. Med Sci Sports 2: 177‐181, 1970.
 227. Stainsby WN , Barclay JK . Oxygen uptake for brief tetanic contractions of dog skeletal muscle in situ. Am J Physiol 223: 371‐375, 1972.
 228. Stephenson DG , Stewart AW , Wilson GJ . Dissociation of force from myofibrillar MgATPase and stiffness at short sarcomere lengths in rat and toad skeletal muscle. J Physiol 410: 351‐366, 1989.
 229. Stephenson DG , Williams DA . Calcium‐activated force responses in fast‐ and slow‐twitch skinned muscle fibres of the rat at different temperatures. J Physiol 317: 281‐302, 1981.
 230. Sun YB , Hilber K , Irving M . Effect of active shortening on the rate of ATP utilisation by rabbit psoas muscle fibres. J Physiol 531: 781‐791, 2001.
 231. Takashi R , Putnam S . A fluorimetric method for continuously assaying ATPase: Application to small specimens of glycerol‐extracted muscle fibers. Anal Biochem 92: 375‐382, 1979.
 232. Tanokura M , Imaizumi M , Yamada K . A calorimetric study of Ca2+ binding by the parvalbumin of the toad (Bufo): Distinguishable binding sites in the molecule. FEBS Lett 209: 77‐82, 1986.
 233. Teague WE , Dobson GP . Effect of temperature on the creatine kinase equilibrium. J Biol Chem 267: 14084‐14093, 1992.
 234. Teague WE , Golding EM , Dobson GP . Adjustment of the K' for the creatine kinase, adenylate kinase and ATP hydrolysis equilibria to varying temperature and ionic strength. J Exp Biol 199: 509‐512, 1996.
 235. Veech RL , Lawson JW , Cornell NW , Krebs HA . Cytosolic phosphorylation potential. J Biol Chem 254: 6538‐6547, 1979.
 236. Walsh B , Howlett RA , Stary CM , Kindig CA , Hogan MC . Measurement of activation energy and oxidative phosphorylation onset kinetics in isolated muscle fibers in the absence of cross‐bridge cycling. Am J Physiol 290: R1707‐1713, 2006.
 237. Walsh TH , Woledge RC . Heat production and chemical change in tortoise muscle. J Physiol 206: 457‐469, 1970.
 238. Wendt IR , Barclay JK . Effects of dantrolene on the energetics of fast‐ and slow‐ twitch muscles of the mouse. Am J Physiol Cell Physiol 238: C56‐C61, 1980.
 239. Wendt IR , Chapman JB . Fluorometric studies of recovery metabolism of rat fast‐ and slow‐twitch muscles. Am J Physiol 230: 1644‐1649, 1976.
 240. Wendt IR , Gibbs CL . Energy production of rat extensor digitorum longus muscle. Am J Physiol 224: 1081‐1086, 1973.
 241. Wendt IR , Gibbs CL . Energy production of mammalian fast‐ and slow‐twitch muscles during development. Am J Physiol 226: 642‐647, 1974.
 242. Wendt IR , Gibbs CL . Recovery heat production of mammalian fast‐ and slow‐twitch muscles. Am J Physiol 230: 637‐643, 1976.
 243. Wendt IR , Stephenson DG . Effects of caffeine on Ca‐activated force production in skinned cardiac and skeletal muscle fibres of the rat. Pflügers Arch 398: 210‐216, 1983.
 244. West TG , Curtin NA , Ferenczi MA , He ZH , Sun YB , Irving M , Woledge RC . Actomyosin energy turnover declines while force remains constant during isometric muscle contraction. J Physiol 555: 27‐43, 2004.
 245. West TG , Toepfer CN , Woledge RC , Curtin NA , Rowlerson A , Kalakoutis M , Hudson P , Wilson AM . Power output of skinned skeletal muscle fibres from the cheetah (Acinonyx jubatus). J Exp Biol 216: 2974‐2982, 2013.
 246. Westerblad H , Allen DG . Myoplasmic free Mg2+ concentration during repetitive stimulation of single fibres from mouse skeletal muscle. J Physiol 453: 413‐434, 1992.
 247. Westerblad H , Allen DG . Myoplasmic Mg2+ concentration in Xenopus muscle fibres at rest, during fatigue and during metabolic blockade. Exp Physiol 77: 733‐740, 1992.
 248. Wilkie DR . Thermodynamics and the interpretation of biological heat measurements. Prog Biophys Biophys Chem 10: 259‐298, 1960.
 249. Wilkie DR . Heat work and phosphorylcreatine break‐down in muscle. J Physiol 195: 157‐183, 1968.
 250. Wilkie DR . The efficiency of muscular contraction. J Mechanochem Cell Motil 2: 257‐267, 1974.
 251. Wilkie DR , Dawson MJ , Edwards RH , Gordon RE , Shaw D . 31P NMR studies of resting muscle in normal human subjects. Adv Exp Med Biol 170: 333‐347, 1984.
 252. Wilkie DR , Woledge RC . The application of irreversible thermodynamics to muscular contraction. Comments on a recent theory by S. R. Caplan. Proc R Soc Lond B Biol Sci 169: 17‐29, 1967.
 253. Woledge RC . The thermoelastic effect of change of tension in active muscle. J Physiol 155: 187‐208, 1961.
 254. Woledge RC . The energetics of tortoise muscle. J Physiol 197: 685‐707, 1968.
 255. Woledge RC . Heat production and chemical change in muscle. Prog Biophys Mol Biol 22: 39‐74, 1971.
 256. Woledge RC , Barclay CJ , Curtin NA . Temperature change as a probe of muscle crossbridge kinetics: A review and discussion. Proc R Soc Lond B Biol Sci 276: 2685‐2695, 2009.
 257. Woledge RC , Curtin NA , Homsher E . Energetic Aspects of Muscle Contraction. London: Academic Press, 1985.
 258. Woledge RC , Reilly PJ . Molar enthalpy change for hydrolysis of phosphorylcreatine under conditions in muscle cells. Biophys J 54: 97‐104, 1988.
 259. Yamada T , Homsher E . The dependence on the distance of shortening of the energy output from frog skeletal muscle shortening at velocities of Vmax, 1/2Vmax and 1/4Vmax . Adv Exp Med Biol 170: 883‐885, 1984.
 260. Yates LD , Greaser ML . Troponin subunit stoichiometry and content in rabbit skeletal muscle and myofibrils. J Biol Chem 258: 5770‐5774, 1983.

Related Articles:

Chemical Mechanism of Myosin‐Catalyzed ATP Hydrolysis

Contact Editor

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

* Required Field

How to Cite

C. J. Barclay. Energetics of Contraction. Compr Physiol 2015, 5: 961-995. doi: 10.1002/cphy.c140038