Comprehensive Physiology Wiley Online Library

Force Generation and Shortening in Skeletal Muscle

Full Article on Wiley Online Library



Abstract

The sections in this article are:

1 Experimental Preparations
2 States of the Fiber
3 Properties of Activated Fibers
3.1 Force‐Velocity Relation
3.2 Viscoelastic Theory
3.3 Fenn Effect
3.4 Quick Releases
3.5 Discovery of Sliding Filaments and Cross Bridges
3.6 Force‐Length Relation
3.7 Kinetic Properties of Cross Bridges
3.8 Cross‐Bridge Model of A. F. Huxley
3.9 High‐Time‐Resolution Mechanical Measurements
3.10 Force Steps
3.11 Length Steps
4 Muscle Biochemistry
4.1 Other Experiments: Some Puzzling Results
4.2 Recent Biochemical Studies
4.3 Hill Formalism
4.4 Application of Hill Formalism
5 Current Frontiers
6 Conclusion
Figure 1. Figure 1.

Double array of myofilaments in a sarcomere. Array of thick filaments in center of sarcomere forms the A band. Thin filaments extend into the thick‐filament array from the dense Z line. The part of the thin‐filament array that does not interdigitate with thick filaments forms the I band. The part of the thick‐filament array that does not interdigitate with thin filaments forms the H zone. × 67,000. [From Huxley 36.]

Figure 2. Figure 2.

Length‐tension relation and filament dispositions at various sarcomere lengths. Upper diagram: length‐tension relation for frog muscle. Middle diagram: sarcomere—a, thick‐filament length; b, thinfilament length; c, length of central bare region of thick filament; Z, width of the Z band. Lower diagrams: filament dispositions at various striation spacings. Number at left of each diagram corresponds to numbered parts of length‐tension relation. Note that 1,2,3, and 5 are close to discontinuities in length‐tension relation. [From Gordon et al. 18.]

Figure 3. Figure 3.

Force‐pCa relation for skinned frog muscle fiber. [From Hellam and Podolsky 23.]

Figure 4. Figure 4.

Contraction kinetics of a skinned frog muscle fiber at different free‐calcium concentrations. Top trace, displacement; middle trace, force; bottom trace, zero force. Fiber is fully activated at pCa 5 and half‐activated at pCa 6.4. Dashed line on displacement trace is back extrapolation of steady motion. The 2 displacement traces, which are juxtaposed in inset, are almost the same. [From Gulati and Podolsky 20.]

Figure 5. Figure 5.

Rate functions for cross‐bridge interaction in the 1957 model of A. F. Huxley. Here f is the rate function for cross‐bridge formation, g is the rate function for cross‐bridge dissociation, x is the distance between the actual position of the actin site and the actin position at which the cross bridge exerts zero force, and h is the value of x at which f reaches a maximum value.

From Huxley 30, © 1957, with permission from Pergamon Press, Ltd
Figure 6. Figure 6.

Cross‐bridge distributions at different steady speeds. Note that area of distribution function (and therefore instantaneous number of cross bridges) decreases as contraction speed increases.

From Huxley 30, © 1957, with permission from Pergamon Press, Ltd
Figure 7. Figure 7.

Cross‐bridge mechanism in which force is generated by a very rapid configurational change. The S1 moiety of the myosin molecule is attached to the thick filament at M through a flexible region that can be extended to length l0 by thermal energy. It is assumed that S1 attaches to an actin site A in one configuration (solid line) and then very rapidly and irreversibly changes to another configuration (broken line) that can stretch the flexible region beyond l0 and produce force. A: change in state stretches the flexible region to length l and force is proportional to ll0. B: flexible region is not completely extended when S1 attached to A, and stretched length of flexible region is less than l. C: stretched length of flexible region is l0 and force is zero. D: change in configuration of S1 moiety does not stretch flexible region beyond l0 and force is zero. If S1 remains attached to A, flexible region can be stretched by motion of the thin filament; negative force is produced when x/h < −0.8. E: force function for this mechanism. Note that stiffness is not a measure of number of attached cross bridges in this mechanism, because cross bridges in flat region of force curve do not show stiffness.

Figure 8. Figure 8.

Response of frog muscle fibers to sudden changes in load at 3°C. Lower traces, force records; upper traces, displacement. a, Isometric tension; b, record after sudden change in load, slow time scale; c‐k, changes in load, rapid time scale. Force step as fraction of P0 is given along side force trace. Arrows mark points at which actual motion intersects back extrapolation of steady motion. Note that force is steady ∼2–6 ms after force step, but displacement transient lasts 10–40 ms. Displacement scale bar is 4 nm per half sarcomere in panels c through g and 8 nm per half sarcomere in the other panels. [From Civan and Podolsky 3.]

Figure 9. Figure 9.

Rate functions for cross‐bridge interaction in the model of Podolsky and Nolan. Upper panel: f is rate function for cross‐bridge formation; g is rate function for cross‐bridge dissociation. Lower panel: k is force function for cross bridge.

From Podolsky and Nolan 52, © 1971, reprinted by permission of Prentice‐Hall, Inc., Englewood Cliffs, NJ
Figure 10. Figure 10.

Cross‐bridge distributions in forcestep experiments. Top panel shows distribution during isometric contraction. When load is suddenly changed to P < P0, isometric distribution shifts to the left as shown in lower panels. Steady‐state distribution for each load is given by solid line. Note that area of steady‐state distribution function (and therefore instantaneous number of cross bridges) increases as load decreases and contraction speed increases.

From Podolsky and Nolan 52, © 1971, reprinted by permission of Prentice‐Hall, Inc., Englewood Cliffs, NJ
Figure 11. Figure 11.

Transient changes in tension exerted by a stimulated frog muscle fiber when suddenly stretched (top panel) or shortened (middle panels). Bottom panel shows typical release. Number next to each record shows size of corresponding length change per half sarcomere (in nm). [From Huxley 31.]

Figure 12. Figure 12.

Relation between velocity transient and tension transient. Phase 1 represents an instantaneous elasticity. Phase 2 is a rapid shortening in velocity transient or a rapid tension recovery in tension transient. Phase 3 is a marked reduction of either shortening speed or tension recovery. Phase 4 is steady shortening in the velocity transient or a very slow recovery of tension in the tension transient. Note that duration of phases is not the same in the two types of transient. Phases 1 and 2 of the tension transient are shown on a faster time scale in Figure 11. [From Huxley 31.]

Figure 13. Figure 13.

Behavior of the cross‐bridge head and compliance (spring) in the model of Huxley and Simmons during tension development and during an isometric transient. There are 3 attached states: state 1 (A), state 2(B and C), and state 3(D). Step length change of muscle occurs between B and C.

Adapted from Huxley and Simmons 35
Figure 14. Figure 14.

Effect of step amplitude size on T1 (extreme tension) and T2 (tension approached during rapid recovery phase), both as a fraction of To, which is isometric tension immediately before the step. [From Huxley 31.]

Figure 15. Figure 15.

One possible free‐energy diagram corresponding to the biochemical cycle AMD·Pi AM MT MD·Pi AMD·Pi shown in text. Free‐energy levels of unattached states are independent of x, a measure of strain in an attached cross bridge, more rigorously defined in Cross‐Bridge Model of A. F. Huxley, p. 177. Free‐energy curves of attached states are parabolas. Minimum free energy of AM state is to left of that for state AMD·Pi under the assumption that equilibrium configuration for AM state is more acutely angled than that for AMD·Pi (i.e., equilibrium configuration for AMD·Pi looks like Fig. 13B and AM more like Fig. 13D). Other states, such as M and AMT, are considered unimportant in this simple model.

Adapted from Eisenberg and Hill 10


Figure 1.

Double array of myofilaments in a sarcomere. Array of thick filaments in center of sarcomere forms the A band. Thin filaments extend into the thick‐filament array from the dense Z line. The part of the thin‐filament array that does not interdigitate with thick filaments forms the I band. The part of the thick‐filament array that does not interdigitate with thin filaments forms the H zone. × 67,000. [From Huxley 36.]



Figure 2.

Length‐tension relation and filament dispositions at various sarcomere lengths. Upper diagram: length‐tension relation for frog muscle. Middle diagram: sarcomere—a, thick‐filament length; b, thinfilament length; c, length of central bare region of thick filament; Z, width of the Z band. Lower diagrams: filament dispositions at various striation spacings. Number at left of each diagram corresponds to numbered parts of length‐tension relation. Note that 1,2,3, and 5 are close to discontinuities in length‐tension relation. [From Gordon et al. 18.]



Figure 3.

Force‐pCa relation for skinned frog muscle fiber. [From Hellam and Podolsky 23.]



Figure 4.

Contraction kinetics of a skinned frog muscle fiber at different free‐calcium concentrations. Top trace, displacement; middle trace, force; bottom trace, zero force. Fiber is fully activated at pCa 5 and half‐activated at pCa 6.4. Dashed line on displacement trace is back extrapolation of steady motion. The 2 displacement traces, which are juxtaposed in inset, are almost the same. [From Gulati and Podolsky 20.]



Figure 5.

Rate functions for cross‐bridge interaction in the 1957 model of A. F. Huxley. Here f is the rate function for cross‐bridge formation, g is the rate function for cross‐bridge dissociation, x is the distance between the actual position of the actin site and the actin position at which the cross bridge exerts zero force, and h is the value of x at which f reaches a maximum value.

From Huxley 30, © 1957, with permission from Pergamon Press, Ltd


Figure 6.

Cross‐bridge distributions at different steady speeds. Note that area of distribution function (and therefore instantaneous number of cross bridges) decreases as contraction speed increases.

From Huxley 30, © 1957, with permission from Pergamon Press, Ltd


Figure 7.

Cross‐bridge mechanism in which force is generated by a very rapid configurational change. The S1 moiety of the myosin molecule is attached to the thick filament at M through a flexible region that can be extended to length l0 by thermal energy. It is assumed that S1 attaches to an actin site A in one configuration (solid line) and then very rapidly and irreversibly changes to another configuration (broken line) that can stretch the flexible region beyond l0 and produce force. A: change in state stretches the flexible region to length l and force is proportional to ll0. B: flexible region is not completely extended when S1 attached to A, and stretched length of flexible region is less than l. C: stretched length of flexible region is l0 and force is zero. D: change in configuration of S1 moiety does not stretch flexible region beyond l0 and force is zero. If S1 remains attached to A, flexible region can be stretched by motion of the thin filament; negative force is produced when x/h < −0.8. E: force function for this mechanism. Note that stiffness is not a measure of number of attached cross bridges in this mechanism, because cross bridges in flat region of force curve do not show stiffness.



Figure 8.

Response of frog muscle fibers to sudden changes in load at 3°C. Lower traces, force records; upper traces, displacement. a, Isometric tension; b, record after sudden change in load, slow time scale; c‐k, changes in load, rapid time scale. Force step as fraction of P0 is given along side force trace. Arrows mark points at which actual motion intersects back extrapolation of steady motion. Note that force is steady ∼2–6 ms after force step, but displacement transient lasts 10–40 ms. Displacement scale bar is 4 nm per half sarcomere in panels c through g and 8 nm per half sarcomere in the other panels. [From Civan and Podolsky 3.]



Figure 9.

Rate functions for cross‐bridge interaction in the model of Podolsky and Nolan. Upper panel: f is rate function for cross‐bridge formation; g is rate function for cross‐bridge dissociation. Lower panel: k is force function for cross bridge.

From Podolsky and Nolan 52, © 1971, reprinted by permission of Prentice‐Hall, Inc., Englewood Cliffs, NJ


Figure 10.

Cross‐bridge distributions in forcestep experiments. Top panel shows distribution during isometric contraction. When load is suddenly changed to P < P0, isometric distribution shifts to the left as shown in lower panels. Steady‐state distribution for each load is given by solid line. Note that area of steady‐state distribution function (and therefore instantaneous number of cross bridges) increases as load decreases and contraction speed increases.

From Podolsky and Nolan 52, © 1971, reprinted by permission of Prentice‐Hall, Inc., Englewood Cliffs, NJ


Figure 11.

Transient changes in tension exerted by a stimulated frog muscle fiber when suddenly stretched (top panel) or shortened (middle panels). Bottom panel shows typical release. Number next to each record shows size of corresponding length change per half sarcomere (in nm). [From Huxley 31.]



Figure 12.

Relation between velocity transient and tension transient. Phase 1 represents an instantaneous elasticity. Phase 2 is a rapid shortening in velocity transient or a rapid tension recovery in tension transient. Phase 3 is a marked reduction of either shortening speed or tension recovery. Phase 4 is steady shortening in the velocity transient or a very slow recovery of tension in the tension transient. Note that duration of phases is not the same in the two types of transient. Phases 1 and 2 of the tension transient are shown on a faster time scale in Figure 11. [From Huxley 31.]



Figure 13.

Behavior of the cross‐bridge head and compliance (spring) in the model of Huxley and Simmons during tension development and during an isometric transient. There are 3 attached states: state 1 (A), state 2(B and C), and state 3(D). Step length change of muscle occurs between B and C.

Adapted from Huxley and Simmons 35


Figure 14.

Effect of step amplitude size on T1 (extreme tension) and T2 (tension approached during rapid recovery phase), both as a fraction of To, which is isometric tension immediately before the step. [From Huxley 31.]



Figure 15.

One possible free‐energy diagram corresponding to the biochemical cycle AMD·Pi AM MT MD·Pi AMD·Pi shown in text. Free‐energy levels of unattached states are independent of x, a measure of strain in an attached cross bridge, more rigorously defined in Cross‐Bridge Model of A. F. Huxley, p. 177. Free‐energy curves of attached states are parabolas. Minimum free energy of AM state is to left of that for state AMD·Pi under the assumption that equilibrium configuration for AM state is more acutely angled than that for AMD·Pi (i.e., equilibrium configuration for AMD·Pi looks like Fig. 13B and AM more like Fig. 13D). Other states, such as M and AMT, are considered unimportant in this simple model.

Adapted from Eisenberg and Hill 10
References
 1. Abbott, R. H., and G. J. Steiger. Temperature and amplitude dependence of tension transients in glycerinated skeletal and insect fibrillar muscle. J. Physiol. London 266: 13–42, 1977.
 2. Chock, S. P., P. B. Chock, and E. Eisenberg. Pre‐steady state kinetic evidence for a cyclic interaction of myosin subfragment‐1 with actin during the hydrolysis of ATP. Biochemistry 15: 3224–3253, 1976.
 3. Civan, M. M., and R. J. Podolsky. Contraction kinetics of striated muscle fibres following quick changes in load. J. Physiol. London 184: 511–534, 1966.
 4. Craig, R., A. G. Szent‐Gyorgyi, L. Beese, P. Flicker, P. Vibert, and C. Cohen. Electron microscopy of thin filaments decorated with a Ca2+‐regulated myosin. J. Mol. Biol. 140: 35–55, 1980.
 5. Curtin, N. A., and R. E. Davies. Chemical and mechanical changes during stretching of activated frog skeletal muscle. Cold Spring Harbor Symp. Quant. Biol. 37: 619–626, 1972.
 6. Eastwood, A. B., D. S. Wood, K. L. Bock, and M. M. Sorenson. Chemically skinned mammalian skeletal muscle. I. The structure of skinned rabbit psoas. Tissue Cell 11: 553–566, 1979.
 7. Ebashi, S., and M. Endo. Calcium ion and muscle contraction. Prog. Biophys. Mol. Biol. 18: 123–183, 1968.
 8. Edman, K. A. P. The velocity of unloaded shortening and its relation to sarcomere length and isometric force in vertebrate muscle fibres. J. Physiol. London 291: 143–159, 1979.
 9. Eisenberg, E., and L. E. Greene. The relation of muscle biochemistry to muscle physiology. Annu. Rev. Physiol. 42: 293–309, 1980.
 10. Eisenberg, E., and T. L. Hill. A cross‐bridge model of muscle contraction. Prog. Biophys. Mol. Biol. 33: 55–82, 1978.
 11. Eisenberg, E., T. L. Hill, and Y. Chen. Cross‐bridge model of muscle contraction. Quantitative analysis. Biophys. J. 29: 195–227, 1980.
 12. Eisenberg, E., and W. W. Kielley. Evidence for a refractory state of heavy meromyosin and subfragment‐1 unable to bind to actin in the presence of ATP. Cold Spring Harbor Symp. Quant. Biol. 37: 145–152, 1972.
 13. Fenn, W. O. A quantitative comparison between the energy liberated and the work performed by the isolated sartorius muscle of the frog. J. Physiol. London 58: 175–203, 1923.
 14. Fenn, W. O. The relation between the work performed and the energy liberated in muscular contraction. J. Physiol. London 58: 373–395, 1924.
 15. Ford, L. E., A. F. Huxley, and R. M. Simmons. Tension responses to sudden length change in stimulated frog muscle fibres near slack length. J. Physiol. London 269: 441–515, 1977.
 16. Ford, L. E., A. F. Huxley, and R. M. Simmons. The relation between stiffness and filament overlap in stimulated frog muscle fibres. J. Physiol. London 311: 219–249, 1981.
 17. Goldman, Y., and R. M. Simmons. Active and rigor muscle stiffness. J. Physiol. London 269: 55P–57P, 1977.
 18. Gordon, A. M., A. F. Huxley, and F. J. Julian. The variation in isometric tension with sarcomere length in vertebrate muscle fibres. J. Physiol. London 184: 170–192, 1966.
 19. Greene, L. E., and E. Eisenberg. Cooperative binding of myosin subfragment‐1 to the actin‐troponin‐tropomyosin complex. Proc. Natl. Acad. Sci. USA 77: 2616–2620, 1980.
 20. Gulati, J., and R. J. Podolsky. Contraction transients of skinned muscle fibers: effects of calcium and ionic strength. J. Gen. Physiol. 72: 701–716, 1978.
 21. Gulati, J., and R. J. Podolsky. Isotonic contraction of skinned muscle fibers on a slow time base. Effects of ionic strength and calcium. J. Gen. Physiol. 78: 233–257, 1981.
 22. Haselgrove, J. C., and H. E. Huxley. X‐ray evidence for radial cross‐bridge movement and for the sliding filament model in actively contracting muscle. J. Mol. Biol. 77: 549–568, 1973.
 23. Hellam, D. C., and R. J. Podolsky. Force measurements in skinned muscle fibres. J. Physiol. London 200: 807–819, 1969.
 24. Herzig, J. W., T. Yamamoto, and J. C. Ruegg. Dependence of force and immediate stiffness on sarcomere length and Ca2+ activation in frog muscle fibres. Pfluegers Arch. 389: 97–103, 1981.
 25. Hill, A. V. The maximum work and mechanical efficiency of human muscles and their most economical speed. J. Physiol. London 56: 19–41, 1922.
 26. Hill, A. V. The heat of shortening and the dynamic constants of muscle. Proc. R. Soc. London Ser. B 126: 136–195, 1938.
 27. Hill, D. K. Tension due to interaction between the sliding filaments in resting striated muscle: the effect of stimulation. J. Physiol. London 199: 637–684, 1968.
 28. Hill, T. L. Theoretical formalism for the sliding filament model of contraction of striated muscle. Part I. Prog. Biophys. Mol. Biol. 28: 267–340, 1974.
 29. Hill, T. L. Theoretical formalism for the sliding filament model of contraction of striated muscle. Part II. Prog. Biophys. Mol. Biol. 29: 105–159, 1974.
 30. Huxley, A. F. Muscle structure and theories of contraction. Prog. Biophys. Biophys. Chem. 7: 255–318, 1957.
 31. Huxley, A. F. Muscular contraction. J. Physiol. London 243: 1–43, 1974.
 32. Huxley, A. F. Reflections on Muscle. Princeton, NJ: Princeton Univ. Press, 1980.
 33. Huxley, A. F., and R. Niedergerke. Interference microscopy of living muscle fibres. Nature London 173: 971–973, 1954.
 34. Huxley, A. F., and L. D. Peachey. The maximum length for contraction in vertebrate striated muscle. J. Physiol. London 156: 150–165, 1961.
 35. Huxley, A. F., and R. M. Simmons. Proposed mechanism of force generation in striated muscle. Nature London 233: 533–538, 1971.
 36. Huxley, H. E. The double array of filaments in cross‐striated muscle. J. Biophys. Biochem. Cytol. 3: 631–648, 1957.
 37. Huxley, H. E. The mechanism of muscular contraction. Science 164: 1356–1366, 1969.
 38. Huxley, H. E., Time resolved X‐ray diffraction studies on muscle. In: Cross‐Bridge Mechanism in Muscle Contraction, edited by H. Sugi and G. H. Pollack. Tokyo: Univ. of Tokyo Press, 1979, p. 391–401.
 39. Huxley, H. E., and J. Hanson. Changes in the cross‐striations of muscle during contraction and stretch and their structural interpretation. Nature London 173: 973, 1954.
 40. Jewell, B. R., and D. R. Wilkie. An analysis of the mechanical components in frog's striated muscle. J. Physiol. London 143: 515–540, 1958.
 41. Julian, F. J. The effect of calcium on the force‐velocity relation of briefly glycerinated frog muscle fibres. J. Physiol. London 218: 117–145, 1971.
 42. Julian, F. J., and M. R. Sollins. Variation of muscle stiffness with force at increasing speeds of shortening. J. Gen. Physiol. 66: 287–302, 1975.
 43. Kawai, M. Head rotation or dissociation? A study of exponential rate processes in chemically skinned rabbit muscle fibers when MgATP concentration is changed. Biophys. J. 22: 97–103, 1978.
 44. Kawai, M., and P. W. Brandt. Sinusoidal analysis: a high resolution method for correlating biochemical reactions with physiological processes in activated skeletal muscles of rabbit, frog, and crayfish. J. Muscle Res. Cell Motil. 1: 279–303, 1980.
 45. Lowey, S., H. S. Slayter, A. G. Weeds, and H. Baker. Substructure of the myosin molecule. J. Mol. Biol. 43: 1–29, 1969.
 46. Lymn, R. W., and E. W. Taylor. Mechanism of adenosine triphosphate hydrolysis by actomyosin. Biochemistry 10: 4617–4624, 1971.
 47. Marston, S. B., C. D. Rodger, and R. T. Tregear. Changes in muscle cross bridges when β, γ‐imido‐ATP binds to myosin. J. Mol. Biol. 104: 263–276, 1976.
 48. Marston, S. B., R. T. Tregear, C. D. Rodger, and M. L. Clark. Coupling between the enzymatic site of myosin and the mechanical output of muscle. J. Mol. Biol. 128: 111–126, 1979.
 49. Natori, R. The property and contraction process of isolated myofibrils. Jikeikai Med. J. 1: 119–126, 1954.
 50. Podolsky, R. J. Kinetics of muscular contraction: the approach to the steady state. Nature London 188: 666–668, 1960.
 51. Podolsky, R. J. The maximum sarcomere length for contraction of isolated myofibrils. J. Physiol. London 170: 110–123, 1964.
 52. Podolsky, R. J., and A. C. Nolan. Cross‐bridge properties derived from physiological studies of frog muscle fibers. In: Contractility of Muscle Cells, edited by R. J. Podolsky. Engle‐wood Cliffs, NJ: Prentice‐Hall, 1971, p. 247–260.
 53. Podolsky, R. J., R. St. Onge, L. Yu, and R. W. Lymn. X‐ray diffraction of actively shortening muscle. Proc. Natl. Acad. Sci. USA 73: 813–817, 1976.
 54. Podolsky, R. J., and L. E. Teichholz. The relation between calcium and contraction in skinned muscle fibres. J. Physiol. London 211: 19–35, 1970.
 55. Pringle, J. W. S. The contractile mechanism of insect fibrillar muscle. Prog. Biophys. Biophys. Chem. 17: 1–60, 1967.
 56. Ramsey, R. W., and S. F. Street. The isometric length‐tension diagram of isolated skeletal muscle fibers of the frog. J. Cell. Comp. Physiol. 15: 11–34, 1940.
 57. Reedy, M. K., K. C. Holmes, and R. T. Tregear. Induced changes in orientation of the cross bridges of glycerinated insect flight muscle. Nature London 207: 1276–1280, 1965.
 58. Sleep, J. A., and R. L. Hutton. Actin mediated release of ATP from a myosin‐ATP complex. Biochemistry 17: 5423–5430, 1978.
 59. Stein, L. A., R. P. Schwarz, P. B. Chock, and E. Eisenberg. The mechanism of actomyosin ATPase: evidence that ATP hydrolysis can occur without dissociation of the actomyosin complex. Biochemistry 18: 3894–3909, 1979.
 60. Szent‐Gyorgyi, A. Free‐energy relations and contractions of actomyosin. Biol. Bull. Woods Hole Mass. 96: 140–161, 1949.
 61. Taylor, S. R., and R. Rudel. Striated muscle fibers: inactivation of contraction induced by shortening. Science 167: 882–884, 1970.
 62. Yu, L. C., J. E. Hartt, and R. J. Podolsky. Equatorial X‐ray intensities and isometric force levels in frog sartorius muscle. J. Mol. Biol. 132: 53–67, 1979.

Contact Editor

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

* Required Field

How to Cite

Richard J. Podolsky, Mark Schoenberg. Force Generation and Shortening in Skeletal Muscle. Compr Physiol 2011, Supplement 27: Handbook of Physiology, Skeletal Muscle: 173-187. First published in print 1983. doi: 10.1002/cphy.cp100106