Comprehensive Physiology Wiley Online Library

Structure of Vertebrate Striated Muscle as Determined by X‐ray‐Diffraction Studies

Full Article on Wiley Online Library



Abstract

The sections in this article are:

1 Muscle Structure: An Overview
1.1 Description of Low‐Angle Diffraction Patterns
2 Interpretation of X‐Ray Patterns
2.1 Theoretical Aspects
2.2 Experimental Approaches
3 Filament Lattice
3.1 Lattice Spacing
3.2 Filament Orientations
3.3 Radial Electron‐Density Distribution
4 Actin‐Containing Filaments
4.1 Helix Parameters
4.2 Thin‐Filament Regulation of Contraction
5 Myosin‐Filament Backbone
5.1 Axial Packing of Myosin
5.2 Helical Packing of Myosin
5.3 Axial Repeat of C Protein
6 Cross Bridges
6.1 Relaxed Muscle
6.2 Rigor Muscle
6.3 Contracting Muscle
7 Analogues of ATP
Figure 1. Figure 1.

Electron micrographs of the filament lattice in vertebrate striated muscle. A: longitudinal section parallel to the 1.1 plane (see also Fig. 5). Thick myosin‐containing filaments lie in longitudinal register and thinner actin‐containing filaments extend on either side of the Z line and interdigitate with myosin filaments. Cross bridges project from the thick filaments. Different zones are marked. B: transverse section through A bands of several myofibrils. Thick filaments are located at lattice points of a hexagonal lattice with actin filaments at trigonal positions, equidistant from 3 myosin filaments (see Fig. 5). [From H. E. Huxley, unpublished observations.]

Figure 2. Figure 2.

A 4‐step, 4‐state model of cross‐bridge action showing mechanical states with corresponding predominant biochemical species. Cross‐bridge orientations in states C and D are often depicted as 45° angles; both A and B are shown as 90° angles. Although states A and D can be distinguished biochemically, differences in physical orientation are still unclear. The hinge region between head and tail of the myosin molecule is thought to be very flexible. A: myosin cross bridge, with cleaved ATP‐hydrolysis products still bound to it, is not yet attached to actin. B: cross bridge attaches to actin monomer at approximate 90° angle. C: cross bridge‐actin angle changes to 45°, pulling the filaments past each other while the cleaved nucleotide products dissociate from myosin. Resultant state is the rigor cross‐link. D: ATP binds to rigor cross‐link causing myosin cross bridge to dissociate from actin filament. Subsequent hydrolysis of ATP leaves the cross bridge in original state (A), ready to bind to the next available actin monomer.

Figure 3. Figure 3.

Low‐angle X‐ray‐diffraction patterns of frog sartorius muscles at rest length, displayed as if taken with muscle axis vertical. The vertical axis of symmetry is called the meridian; the horizontal axis is the equator. Right (in each picture), axial position of myosin meridional reflections (14.3 nm); left, axial position of actin layer line (5.9 nm). A: relaxed muscle. Pattern is dominated by layer lines arising from helical arrangement of cross bridges around myosin filament. B: relaxed muscle, meridional region (very low angle pattern). T, doublet near 38.5 nm arising from troponin repeat; C, doublet near 43.0 nm arising from C protein. C: rigor muscle. Relaxed myosin layer lines have disappeared, replaced by a new series of layer lines arising from cross bridges attached to actin. D: contracting muscle. The layer lines arising from myosin cross bridges are extremely weak and the pattern is dominated by layer lines from actin filament. [With permission from Haselgrove 46. Copyright by Academic Press Inc. (London) Ltd.]

Figure 4. Figure 4.

Schematic representation of principal X‐ray reflections from rest‐length VSM. Right, spacings of myosin reflections; left, spacing of actin reflections. [From Haselgrove and Rodger 51.]

Figure 5. Figure 5.

Hexagonal filament lattice with myosin filaments at lattice positions, actin filaments at trigonal positions. (See also Fig. LA.) A and B: different sets of planes have different crystallographic notations and their spacing determines the spacing of corresponding equatorial reflections. When the sarcomere's electron‐density distribution is projected along the lattice planes, the resulting one‐dimensional distribution can be described in terms of sine waves. the amplitude (F) of the fundamental sine wave (with a wavelength equal to the separation of the planes) controls and is proportional to the equatorial reflection amplitude for those planes. Transferring cross‐bridge mass from the vicinity of thick filaments to that of thin filaments decreases intensity F2 of the 1.0 reflection and increases intensity of the 1.1 reflection. Relaxed muscle (A): myosin cross bridges (hatched area) lie close to myosin filaments. Rigor muscle (B): myosin cross bridges lie close to actin filaments. C, D, E: radial electron‐density maps for rest‐length muscle constructed by Fourier summation from the amplitudes of the 1.0 and 1.1 equatorial reflections with phases of 0° for both reflections. Amplitudes taken from patterns of relaxed (C), contracting (D), and rigor (E) muscle. F: radial electron‐density distribution of relaxed muscle calculated by Fourier summation; amplitudes calculated from intensities in Table 3. The 1.0, 1.1, and 3.0 reflections had phases of 0°; the 2.0 and 2.1 reflections had phases of 180°. Calculations assuming 6‐fold symmetry gave rise to 6 projections, but no fine details can be usefully interpreted in terms of cross‐bridge shape and postion. [C‐E with permission from Haselgrove and Huxley 49. Copyright by Academic Press Inc. (London) Ltd.; F from Haselgrove et al. 52.]

Figure 6. Figure 6.

A: equatorial diffraction pattern of relaxed frog sartorius muscle at rest length, showing the first 5 equatorial reflections arising from the filament lattice. From right to left from the origin, they are 1.0, 1.1, 2.0, 2.1, 3.0. A weak reflection lying between 1.0 and 1.1 is attributed to the Z line 183. Relative reflection intensities are shown in Table 3. Center, meridional reflection at 14.3 nm is indicated. To right of backstop, higher‐order reflections obscured by part of the camera. B‐G: densitometer tracings from equator of relaxed, contracting, and rigor muscles at different sarcomere lengths. Thick line, 1.0 reflection; thin line, 1.1 reflection. Tracings are not all to the same vertical scale.

A from Haselgrove et al. 48, © 1977 by The Institute of Physics. B‐G with permission from Haselgrove and Huxley 49. Copyright by Academic Press Inc. (London) Ltd
Figure 7. Figure 7.

A: relation between myosin‐filament spacing (d) and sarcomere length (s) in intact living muscle fiber. Data fit a line of the form d2 ∝ 1/s. Thus lattice volume is independent of sarcomere length. B: comparison of filament lattice behavior of intact and skinned fibers. The lattice is constrained by the membrane in intact living fiber. Filled circles, relative unit cell areas of living intact fibers plotted against the inverse of the osmolarity of the bathing medium; filled triangles, area of skinned fibers plotted against the inverse of the corresponding ionic strength. Open symbols, mean values for muscles in normal physiological conditions. [A from April 3; B from April 2.]

Figure 8. Figure 8.

A: X‐ray pattern showing myosin layer lines in relaxed muscle sampled by Bragg reflections, arrows, from the myosin‐filament lattice. Sampling on 1st and 2nd layer lines occurs at different radial positions from the equator and 3rd layer line, indicating that the lattice is not a simple one with all myosin filaments having the same orientation. B: superlattice proposed by H.E. Huxley and Brown to explain the X‐ray sampling. (This illustration is for a 2‐stranded filament, but the logic may be applied to filaments of any symmetry.) Numbers on the cross bridges represent the level (in steps of 14.3 nm) along the filament at which the cross bridges project in the direction shown. Nearest neighbors are rotated through 1/3 or 2/3 revolution with respect to each other. Superlattice, broken line, links adjacent filaments with the same orientation. In cross section, myosin filaments have triangular profiles and on such a superlattice all triangles would point in the same direction. C: Squire's superlattice in which filaments face in only 2 directions. [A from Haselgrove et al. 48; B with permission from Huxley and Brown 70. Copyright by Academic Press Inc. (London) Ltd.; C adapted from Luther and Squire 83.]

Figure 9. Figure 9.

Plot of equatorial intensity ratio I1.0/I1.1 for different muscle states. All data are for frog striated muscle except the open circles for live relaxed rabbit psoas muscle. [From Huxley 66.] A, relaxed living muscle; B and C, isometrically contracting muscle; D, rigor muscle. Vertical line, ratios depend on tension for muscles that, although contracting isometrically, generate only a fraction of the maximum isometric tension. Tensions of 0%, 25%, 50%, and 100% are shown 182. Open triangle, isometric contraction 105. Two reflections arise from lattice planes across the muscle (see Fig. 5). Ratio I1.0/I1.1 falls when electron density at actin position relative to myosin position rises. This occurs 1) if actin filaments are more ordered at short as opposed to long sarcomere lengths or 2) if cross bridges move from the myosin‐filament region to that of actin when the muscle contracts or passes into rigor. [Data for A, B, D from Haselgrove and Huxley 49; data for C from Podolsky et al. 129.]

Figure 10. Figure 10.

A: the actin helix may be considered either a two‐stranded structure with a half‐pitch of about 37.5 nm (top part of helix) or a single‐stranded genetic helix with a pitch of about 5.9 nm (lower part). Because the pitches of the actin and myosin helices are different in VSM, actin‐myosin interaction does not repeat exactly; note the difference in attachment geometry between the 1st and 4th cross bridges. Here only the helical structure of the actin filament is emphasized, although attachment positions of myosin cross bridges to actin are determined by the helical structures of both filaments. B: structure of the actin filament with tropomyosin and troponin. [A adapted from Haselgrove and Reedy 50.]

Figure 11. Figure 11.

A and B: diagrams of two actin filaments with different axial repeats. A repeats after 2 × 35.5 nm whereas B repeats after 2 × 40.9 nm. Superficially the filaments appear very similar, but differences are clear if one holds the page at eye level and sights from left to right. C and D: low‐angle diffraction patterns anticipated from the structures of A and B, respectively. By sighting across the page one can see that changes in the long helix pitch make only very slight changes in layer‐line positions. Since layer lines tend to be broad parallel to the meridian it is impossible to measure their axial spacing with any accuracy or, therefore, to obtain a reliable measure of the long helix pitch to an accuracy of better than ∼3% (2 × 1 nm in 2 × 38.5 nm).

Figure 12. Figure 12.

X‐ray patterns printed by enhancement technique [see 166] to show actin‐filament layer lines (cf. Fig. 11). Positions of first 3 layer lines, influenced by tropomyosin, are marked between the 2 patterns. A: relaxed muscle; 3rd layer line is stronger than 2nd because tropomyosin molecule is positioned near the edge of the actin groove (see also Fig. 13). B: rigor muscle; 2nd layer line is stronger than 3rd because tropomyosin has moved nearer to center of groove. [From Vibert et al. 166. Copyright by Academic Press Inc. (London) Ltd.]

Figure 13. Figure 13.

Schematic diagrams illustrating movement of tropomyosin when thin filaments are activated, and possible correlation to myosin S1‐ac‐tin binding. All diagrams drawn as if looking along the filament from M line to Z line. A and B: the 3‐D reconstruction data of Moore et al. indicate that myosin S1 binds on each filament strand approximately as shown. X‐ray data show that tropomyosin moves toward the center of the helix groove when the relaxed filament (solid circle position) is activated (broken circle position); X‐ray data, however, cannot be used to determine on which side of the filament tropomyosin sits. One is, therefore, unable to distinguish between 2 possible geometries, A and B. A, original steric‐blocking mechanism; B, tropomyosin position as deduced by Seymour and O'Brien. In the latter model, tropomyosin position cannot control S1. C: recent data indicating that S1 probably extends to a position where tropomyosin can physically prevent attachment. TMa, position of tropomyosin in activated muscle; TMb, position in relaxed muscle. [C with permission from Taylor and Amos 157. Copyright by Academic Press Inc. (London) Ltd.]

Figure 14. Figure 14.

A: schematic representation of layer lines from myosin cross‐bridge array. Layer lines are indexed on a helix with a subunit pitch of 14.3 nm and a helix pitch of N × 14.3 where N = rotational symmetry of the filament. B, C, D: nets of surface lattices that can give rise to layer lines. The N values of myosin VSM filament are thought to be 2, 3, or 4. Note that each lattice has an axial repeat subunit of 14.3 nm and repeats exactly after 42.9 nm, although the basic helical strand, thin line, has a pitch of N × 42.9 nm. Perfect lattices like those in B, C, and D would give meridional reflections at orders of 14.3 nm only; the presence of forbidden reflections (see Tables 1 and 2; Fig. 3) indicates that the lattice is not perfect. A possible perturbation might be for the axial subunit repeat to be nonuniform. E: a 2‐stranded filament with regular repeat of 14.3 nm between cross bridges. F: sequence of 16, 11, and 16 nm. Note that since the filament's rotational symmetry is not known, any suitable symmetry may be used for illustrating general structural features. [B‐D with permission from Luther and Squire 83. Copyright by Academic Press Inc. (London) Ltd.; E and F from Yagi et al. 180.]

Figure 15. Figure 15.

Possible cross‐bridge orientations on backbone of relaxed muscle. Cross‐bridge shape is described by Elliott and Offer 26. A: view perpendicular to filament axis. Two parts of the cross bridges, perhaps representing 2 S1 heads, are tilted slightly along the axis. B: view looking down on one crown of a 3‐stranded model. Heads lie close to filament surface, and one head of each molecule lies behind the other in this view.

Adapted from Haselgrove 47
Figure 16. Figure 16.

Electron micrographs of insect flight muscle, IFM, in rigor (cf. Fig. 1.). A: longitudinal section, 2 cross bridges attach to actin every 38.5 nm along all filaments throughout the fibril. B: insert, optical‐diffraction pattern of longitudinal section. C: transverse section, 4 cross bridges connect each myosin filament to 4 of the adjacent actin filaments. Because this section is cut at a slight angle, the axial level of sectioning changes across the picture: flared X made by the cross bridges rotates clockwise by about 60° from left to right. Note that IFM actin filaments are in different lattice positions than those in VSM. [From Haselgrove and Reedy 50.]

Figure 17. Figure 17.

X‐ray pattern of semitendinosus muscle in rigor at non‐overlap length (cf. Fig. 3C). Myosin layer lines have disappeared, but because the actin filaments have been withdrawn from the A band, cross bridges do not attach to actin and no rigor layer lines are seen. [With permission from Haselgrove 46. Copyright by Academic Press Inc. (London) Ltd.]

Figure 18. Figure 18.

Kinetic data following time course of changing X‐ray reflections during tension development and relaxation. Open squares, intensity; filled circles, tension. A and B: curves plotted as percentage of maximal change. A: 42.9‐nm layer line during a single twitch. Three hundred separate twitches were summed for a total integrated time of 3 s for each 10‐ms time interval. Note that intensity decreased by 90%. B: 14.3‐nm meridional reflection and tension during a twitch at 10°C. Total exposure is 1.5 s in each 10‐ms time slot. C: time course of change in intensity of the 14.3‐nm reflection during a quick release in a tetanus. Release was 1.3% of initial length. [A and B from Huxley et al. 72; C from Huxley et al. 75.]



Figure 1.

Electron micrographs of the filament lattice in vertebrate striated muscle. A: longitudinal section parallel to the 1.1 plane (see also Fig. 5). Thick myosin‐containing filaments lie in longitudinal register and thinner actin‐containing filaments extend on either side of the Z line and interdigitate with myosin filaments. Cross bridges project from the thick filaments. Different zones are marked. B: transverse section through A bands of several myofibrils. Thick filaments are located at lattice points of a hexagonal lattice with actin filaments at trigonal positions, equidistant from 3 myosin filaments (see Fig. 5). [From H. E. Huxley, unpublished observations.]



Figure 2.

A 4‐step, 4‐state model of cross‐bridge action showing mechanical states with corresponding predominant biochemical species. Cross‐bridge orientations in states C and D are often depicted as 45° angles; both A and B are shown as 90° angles. Although states A and D can be distinguished biochemically, differences in physical orientation are still unclear. The hinge region between head and tail of the myosin molecule is thought to be very flexible. A: myosin cross bridge, with cleaved ATP‐hydrolysis products still bound to it, is not yet attached to actin. B: cross bridge attaches to actin monomer at approximate 90° angle. C: cross bridge‐actin angle changes to 45°, pulling the filaments past each other while the cleaved nucleotide products dissociate from myosin. Resultant state is the rigor cross‐link. D: ATP binds to rigor cross‐link causing myosin cross bridge to dissociate from actin filament. Subsequent hydrolysis of ATP leaves the cross bridge in original state (A), ready to bind to the next available actin monomer.



Figure 3.

Low‐angle X‐ray‐diffraction patterns of frog sartorius muscles at rest length, displayed as if taken with muscle axis vertical. The vertical axis of symmetry is called the meridian; the horizontal axis is the equator. Right (in each picture), axial position of myosin meridional reflections (14.3 nm); left, axial position of actin layer line (5.9 nm). A: relaxed muscle. Pattern is dominated by layer lines arising from helical arrangement of cross bridges around myosin filament. B: relaxed muscle, meridional region (very low angle pattern). T, doublet near 38.5 nm arising from troponin repeat; C, doublet near 43.0 nm arising from C protein. C: rigor muscle. Relaxed myosin layer lines have disappeared, replaced by a new series of layer lines arising from cross bridges attached to actin. D: contracting muscle. The layer lines arising from myosin cross bridges are extremely weak and the pattern is dominated by layer lines from actin filament. [With permission from Haselgrove 46. Copyright by Academic Press Inc. (London) Ltd.]



Figure 4.

Schematic representation of principal X‐ray reflections from rest‐length VSM. Right, spacings of myosin reflections; left, spacing of actin reflections. [From Haselgrove and Rodger 51.]



Figure 5.

Hexagonal filament lattice with myosin filaments at lattice positions, actin filaments at trigonal positions. (See also Fig. LA.) A and B: different sets of planes have different crystallographic notations and their spacing determines the spacing of corresponding equatorial reflections. When the sarcomere's electron‐density distribution is projected along the lattice planes, the resulting one‐dimensional distribution can be described in terms of sine waves. the amplitude (F) of the fundamental sine wave (with a wavelength equal to the separation of the planes) controls and is proportional to the equatorial reflection amplitude for those planes. Transferring cross‐bridge mass from the vicinity of thick filaments to that of thin filaments decreases intensity F2 of the 1.0 reflection and increases intensity of the 1.1 reflection. Relaxed muscle (A): myosin cross bridges (hatched area) lie close to myosin filaments. Rigor muscle (B): myosin cross bridges lie close to actin filaments. C, D, E: radial electron‐density maps for rest‐length muscle constructed by Fourier summation from the amplitudes of the 1.0 and 1.1 equatorial reflections with phases of 0° for both reflections. Amplitudes taken from patterns of relaxed (C), contracting (D), and rigor (E) muscle. F: radial electron‐density distribution of relaxed muscle calculated by Fourier summation; amplitudes calculated from intensities in Table 3. The 1.0, 1.1, and 3.0 reflections had phases of 0°; the 2.0 and 2.1 reflections had phases of 180°. Calculations assuming 6‐fold symmetry gave rise to 6 projections, but no fine details can be usefully interpreted in terms of cross‐bridge shape and postion. [C‐E with permission from Haselgrove and Huxley 49. Copyright by Academic Press Inc. (London) Ltd.; F from Haselgrove et al. 52.]



Figure 6.

A: equatorial diffraction pattern of relaxed frog sartorius muscle at rest length, showing the first 5 equatorial reflections arising from the filament lattice. From right to left from the origin, they are 1.0, 1.1, 2.0, 2.1, 3.0. A weak reflection lying between 1.0 and 1.1 is attributed to the Z line 183. Relative reflection intensities are shown in Table 3. Center, meridional reflection at 14.3 nm is indicated. To right of backstop, higher‐order reflections obscured by part of the camera. B‐G: densitometer tracings from equator of relaxed, contracting, and rigor muscles at different sarcomere lengths. Thick line, 1.0 reflection; thin line, 1.1 reflection. Tracings are not all to the same vertical scale.

A from Haselgrove et al. 48, © 1977 by The Institute of Physics. B‐G with permission from Haselgrove and Huxley 49. Copyright by Academic Press Inc. (London) Ltd


Figure 7.

A: relation between myosin‐filament spacing (d) and sarcomere length (s) in intact living muscle fiber. Data fit a line of the form d2 ∝ 1/s. Thus lattice volume is independent of sarcomere length. B: comparison of filament lattice behavior of intact and skinned fibers. The lattice is constrained by the membrane in intact living fiber. Filled circles, relative unit cell areas of living intact fibers plotted against the inverse of the osmolarity of the bathing medium; filled triangles, area of skinned fibers plotted against the inverse of the corresponding ionic strength. Open symbols, mean values for muscles in normal physiological conditions. [A from April 3; B from April 2.]



Figure 8.

A: X‐ray pattern showing myosin layer lines in relaxed muscle sampled by Bragg reflections, arrows, from the myosin‐filament lattice. Sampling on 1st and 2nd layer lines occurs at different radial positions from the equator and 3rd layer line, indicating that the lattice is not a simple one with all myosin filaments having the same orientation. B: superlattice proposed by H.E. Huxley and Brown to explain the X‐ray sampling. (This illustration is for a 2‐stranded filament, but the logic may be applied to filaments of any symmetry.) Numbers on the cross bridges represent the level (in steps of 14.3 nm) along the filament at which the cross bridges project in the direction shown. Nearest neighbors are rotated through 1/3 or 2/3 revolution with respect to each other. Superlattice, broken line, links adjacent filaments with the same orientation. In cross section, myosin filaments have triangular profiles and on such a superlattice all triangles would point in the same direction. C: Squire's superlattice in which filaments face in only 2 directions. [A from Haselgrove et al. 48; B with permission from Huxley and Brown 70. Copyright by Academic Press Inc. (London) Ltd.; C adapted from Luther and Squire 83.]



Figure 9.

Plot of equatorial intensity ratio I1.0/I1.1 for different muscle states. All data are for frog striated muscle except the open circles for live relaxed rabbit psoas muscle. [From Huxley 66.] A, relaxed living muscle; B and C, isometrically contracting muscle; D, rigor muscle. Vertical line, ratios depend on tension for muscles that, although contracting isometrically, generate only a fraction of the maximum isometric tension. Tensions of 0%, 25%, 50%, and 100% are shown 182. Open triangle, isometric contraction 105. Two reflections arise from lattice planes across the muscle (see Fig. 5). Ratio I1.0/I1.1 falls when electron density at actin position relative to myosin position rises. This occurs 1) if actin filaments are more ordered at short as opposed to long sarcomere lengths or 2) if cross bridges move from the myosin‐filament region to that of actin when the muscle contracts or passes into rigor. [Data for A, B, D from Haselgrove and Huxley 49; data for C from Podolsky et al. 129.]



Figure 10.

A: the actin helix may be considered either a two‐stranded structure with a half‐pitch of about 37.5 nm (top part of helix) or a single‐stranded genetic helix with a pitch of about 5.9 nm (lower part). Because the pitches of the actin and myosin helices are different in VSM, actin‐myosin interaction does not repeat exactly; note the difference in attachment geometry between the 1st and 4th cross bridges. Here only the helical structure of the actin filament is emphasized, although attachment positions of myosin cross bridges to actin are determined by the helical structures of both filaments. B: structure of the actin filament with tropomyosin and troponin. [A adapted from Haselgrove and Reedy 50.]



Figure 11.

A and B: diagrams of two actin filaments with different axial repeats. A repeats after 2 × 35.5 nm whereas B repeats after 2 × 40.9 nm. Superficially the filaments appear very similar, but differences are clear if one holds the page at eye level and sights from left to right. C and D: low‐angle diffraction patterns anticipated from the structures of A and B, respectively. By sighting across the page one can see that changes in the long helix pitch make only very slight changes in layer‐line positions. Since layer lines tend to be broad parallel to the meridian it is impossible to measure their axial spacing with any accuracy or, therefore, to obtain a reliable measure of the long helix pitch to an accuracy of better than ∼3% (2 × 1 nm in 2 × 38.5 nm).



Figure 12.

X‐ray patterns printed by enhancement technique [see 166] to show actin‐filament layer lines (cf. Fig. 11). Positions of first 3 layer lines, influenced by tropomyosin, are marked between the 2 patterns. A: relaxed muscle; 3rd layer line is stronger than 2nd because tropomyosin molecule is positioned near the edge of the actin groove (see also Fig. 13). B: rigor muscle; 2nd layer line is stronger than 3rd because tropomyosin has moved nearer to center of groove. [From Vibert et al. 166. Copyright by Academic Press Inc. (London) Ltd.]



Figure 13.

Schematic diagrams illustrating movement of tropomyosin when thin filaments are activated, and possible correlation to myosin S1‐ac‐tin binding. All diagrams drawn as if looking along the filament from M line to Z line. A and B: the 3‐D reconstruction data of Moore et al. indicate that myosin S1 binds on each filament strand approximately as shown. X‐ray data show that tropomyosin moves toward the center of the helix groove when the relaxed filament (solid circle position) is activated (broken circle position); X‐ray data, however, cannot be used to determine on which side of the filament tropomyosin sits. One is, therefore, unable to distinguish between 2 possible geometries, A and B. A, original steric‐blocking mechanism; B, tropomyosin position as deduced by Seymour and O'Brien. In the latter model, tropomyosin position cannot control S1. C: recent data indicating that S1 probably extends to a position where tropomyosin can physically prevent attachment. TMa, position of tropomyosin in activated muscle; TMb, position in relaxed muscle. [C with permission from Taylor and Amos 157. Copyright by Academic Press Inc. (London) Ltd.]



Figure 14.

A: schematic representation of layer lines from myosin cross‐bridge array. Layer lines are indexed on a helix with a subunit pitch of 14.3 nm and a helix pitch of N × 14.3 where N = rotational symmetry of the filament. B, C, D: nets of surface lattices that can give rise to layer lines. The N values of myosin VSM filament are thought to be 2, 3, or 4. Note that each lattice has an axial repeat subunit of 14.3 nm and repeats exactly after 42.9 nm, although the basic helical strand, thin line, has a pitch of N × 42.9 nm. Perfect lattices like those in B, C, and D would give meridional reflections at orders of 14.3 nm only; the presence of forbidden reflections (see Tables 1 and 2; Fig. 3) indicates that the lattice is not perfect. A possible perturbation might be for the axial subunit repeat to be nonuniform. E: a 2‐stranded filament with regular repeat of 14.3 nm between cross bridges. F: sequence of 16, 11, and 16 nm. Note that since the filament's rotational symmetry is not known, any suitable symmetry may be used for illustrating general structural features. [B‐D with permission from Luther and Squire 83. Copyright by Academic Press Inc. (London) Ltd.; E and F from Yagi et al. 180.]



Figure 15.

Possible cross‐bridge orientations on backbone of relaxed muscle. Cross‐bridge shape is described by Elliott and Offer 26. A: view perpendicular to filament axis. Two parts of the cross bridges, perhaps representing 2 S1 heads, are tilted slightly along the axis. B: view looking down on one crown of a 3‐stranded model. Heads lie close to filament surface, and one head of each molecule lies behind the other in this view.

Adapted from Haselgrove 47


Figure 16.

Electron micrographs of insect flight muscle, IFM, in rigor (cf. Fig. 1.). A: longitudinal section, 2 cross bridges attach to actin every 38.5 nm along all filaments throughout the fibril. B: insert, optical‐diffraction pattern of longitudinal section. C: transverse section, 4 cross bridges connect each myosin filament to 4 of the adjacent actin filaments. Because this section is cut at a slight angle, the axial level of sectioning changes across the picture: flared X made by the cross bridges rotates clockwise by about 60° from left to right. Note that IFM actin filaments are in different lattice positions than those in VSM. [From Haselgrove and Reedy 50.]



Figure 17.

X‐ray pattern of semitendinosus muscle in rigor at non‐overlap length (cf. Fig. 3C). Myosin layer lines have disappeared, but because the actin filaments have been withdrawn from the A band, cross bridges do not attach to actin and no rigor layer lines are seen. [With permission from Haselgrove 46. Copyright by Academic Press Inc. (London) Ltd.]



Figure 18.

Kinetic data following time course of changing X‐ray reflections during tension development and relaxation. Open squares, intensity; filled circles, tension. A and B: curves plotted as percentage of maximal change. A: 42.9‐nm layer line during a single twitch. Three hundred separate twitches were summed for a total integrated time of 3 s for each 10‐ms time interval. Note that intensity decreased by 90%. B: 14.3‐nm meridional reflection and tension during a twitch at 10°C. Total exposure is 1.5 s in each 10‐ms time slot. C: time course of change in intensity of the 14.3‐nm reflection during a quick release in a tetanus. Release was 1.3% of initial length. [A and B from Huxley et al. 72; C from Huxley et al. 75.]

References
 1. Amemiya, Y., H. Sugi, and H. Hashizume. Effect of slow stretch on the cross‐bridges in active frog skeletal muscle. In: Proc. Int. Symp. Biophys., 4th, Ibaraki, 1978. Ibaraki, Japan: Taniguchi Found., 1978, p. 152–163.
 2. April, E. W. Liquid‐crystalline characteristics of the thick filament lattice of striated muscle. Nature London 257: 139–141, 1975.
 3. April, E. W. The myofilament lattice: studies on isolated fibers. IV. Lattice equilibria in striated muscle. J. Mechanochem. Cell Motil. 3: 111–121, 1975.
 4. April, E. W., and P. W. Brandt. The myofilament lattice: studies on isolated fibers. 3. The effect of myofilament spacing upon tension. J. Gen. Physiol. 61: 490–508. 1973.
 5. April, E. W., P. W. Brandt, and G. F. Elliott. The myofilament lattice: studies on isolated fibers. II. The effects of osmotic strength, ionic concentration, and pH upon the unit‐cell volume. J. Cell Biol. 53: 53–65, 1972.
 6. April, E. W., and D. Wong. Non‐isovolumetric behavior of the unit cell of skinned striated muscle fibers. J. Mol. Biol. 101: 107–114, 1976.
 7. Armitage, P., A. Miller, C. D. Rodger, and R. T. Tregear. The structure and function of insect muscle. Cold Spring Harbor Symp. Quant. Biol. 37: 379–387, 1972.
 8. Armitage, P. M., R. T. Tregear, and A. Miller. Effect of activation on the X‐ray diffraction pattern from insect flight muscle. J. Mol. Biol. 92: 39–53, 1975.
 9. Astbury, W. T. On the structure of biological fibres and the problem of muscle. Proc. R. Soc. London 134: 303–328, 1947.
 10. Barrington‐Leigh, J., R. S. Goody, W. Hofman, K. Holmes, H. G. Mannhertz, G. Rosenbaum, and R. T. Tregear. The interpretation of X‐ray diffraction from glycerinated flight muscle fibre bundles: new theoretical and experimental approaches. In: Insect Flight Muscle, edited by R. T. Tregear. Amsterdam: Elsevier/North‐Holland, 1977, p. 137–146.
 11. Barrington‐Leigh, J., K. C. Holmes, H. G. Mannhertz, F. Eckstein, and R. Goody. Effects of ATP analogues on the low‐angle X‐ray diffraction pattern of insect flight muscle. Cold Spring Harbor Symp. Quant. Biol. 37: 443–448, 1972.
 12. Bear, R. S. Small‐angle X‐ray diffraction studies on muscle. J. Am. Chem. Soc. 67: 1625, 1945.
 13. Bear, R. S., and C. C. Selby. The structure of paramyosin fibrils according to X‐ray diffraction. J. Biophys. Biochem. Cytol. 2: 55–69, 1956.
 14. Bernal, J. D., and I. Fankuchen. X‐ray and crystallographic studies of plant virus preparations. J. Gen. Physiol. 25: 111–146, 1941.
 15. Blinks, J. R. Influence of osmotic strength on cross section and volume of isolated single muscle fibres. J. Physiol. London 177: 42–47, 1965.
 16. Carlson, F. D., R. Bonner, and A. Fraser. Intensity fluctuation autocorrelation studies of contracting frog sartorius muscle. Cold Spring Harbor Symp. Quant. Biol. 37: 389–396, 1972.
 17. Cohen, C., and J. Hanson. An X‐ray diffraction study of F actin. Biochim. Biophys. Acta 21: 177–178, 1956.
 18. Craig, R. Structure of A‐segments from frog and rabbit skeletal muscle. J. Mol. Biol. 109: 69–81, 1977.
 19. Craig, R., and G. Offer. Axial arrangement of crossbridges in thick filaments of vertebrate skeletal muscle. J. Mol. Biol. 102: 325–332, 1976.
 20. Craig, R., and G. Offer. The location of C‐protein in rabbit skeletal muscle. Proc. R. Soc. London Ser. B 192: 451–461, 1976.
 21. 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.
 22. Depue, R. H., and R. Rice. F‐actin is a right‐handed helix. J. Mol. Biol. 12: 302–303, 1965.
 23. Eames, C. H., and A. J. Rowe. Frictional properties and molecular weight of native and synthetic myosin filaments from vertebrate skeletal muscle. Biochim. Biophys. Acta 537: 125–144, 1978.
 24. Ebashi, S., M. Endo, and I. Ohtsuki. Control of muscle contraction. Q. Rev. Biophys. 2: 351–384, 1969.
 25. Elliott, A., and J. Lowy. A model for the coarse structure of paramyosin filaments. J. Mol. Biol. 53: 181–203, 1970.
 26. Elliott, A., and G. Offer. The shape and flexibility of the myosin molecule. J. Mol. Biol. 123: 505–519, 1978.
 27. Elliott, G. F. X‐ray diffraction studies on striated and smooth muscles. Proc. R. Soc. London Ser. B 160: 467–472, 1964.
 28. Elliott, G. F. Force‐balances and stability in hexagonallypacked polyelectrolyte systems. J. Theor. Biol. 21: 71–87, 1968.
 29. Elliott, G. F., J. Lowy, and B. M. Millman. X‐ray diffraction from living striated muscle during contraction. Nature London 206: 1357–1358, 1965.
 30. Elliott, G. F., J. Lowy, and B. M. Millman. Low‐angle x‐ray diffraction studies of living striated muscle during contraction. J. Mol. Biol. 25: 31–45, 1967.
 31. Elliott, G. F., J. Lowy, and C. R. Worthington. An x‐ray and light diffraction study of the filament lattice of striated muscle in the living state and in rigor. J. Mol. Biol. 6: 295–305, 1963.
 32. Elliott, G. F., E. M. Rome, and M. Spencer. A type of contraction hypothesis applicable to all muscles. Nature London 226: 417–420, 1970.
 33. Faruqi, A. R., and H. E. Huxley. A review of techniques for time‐resolved X‐ray studies on muscle. In: Scattering Techniques Applied to Supramolecular and Nonequilibrium Systems, edited by S. H. Chem, B. Chu, and R. Nossall. New York: Plenum, 1981. (Nato Adv. Study Inst. Ser. B Phys.).
 34. Gayton, D. C., and G. F. Elliott. Structural and osmotic studies of single giant fibres of barnacle muscle. J. Muscle Res. Cell Motil. 1: 391–407, 1980.
 35. Gillis, J. M., and E. J. O'brien. The effect of calcium on the structure of reconstituted muscle thin filaments. J. Mol. Biol. 99: 445–459, 1975.
 36. Goody, R. S., J. Barrington‐Leigh, H. G. Mannhertz, R. T. Tregear, and G. Rosenbaum. X‐ray titration of binding of beta, gamma‐imido‐ATP to myosin in insect flight muscle. Nature London 262: 613–615, 1976.
 37. Goody, R. S., and F. Eckstein. Thiophosphate analogues of nucleoside di‐ and tri‐phosphates. J. Am. Chem. Soc. 93: 6252, 1971.
 38. Goody, R. S., K. C. Holmes, H. G. Mannhertz, J. Barrington‐Leigh, and G. Rosenbaum. Cross bridge conformation as revealed by X‐ray diffraction studies of insect flight muscles with ATP analogues. Biophys. J. 15: 687–705, 1975.
 39. Gordon, A. M., A. F. Huxley, and F. J. Julian. Tension development in highly stretched vertebrate muscle fibres. J. Physiol. London 184: 143–169, 1966.
 40. Hanson, J. Axial period of actin filaments. Nature London 213: 353–356, 1967.
 41. Hanson, J. Recent X‐ray diffraction studies of muscle. Q. Rev. Biophys. 1: 177–216, 1968.
 42. Hanson, J., V. Lednev, E. J. O'Brien, and P. M. Bennett. Structure of the actin‐containing filaments in vertebrate skeletal muscle. Cold Spring Harbor Symp. Quant. Biol. 37: 311–318, 1972.
 43. Hanson, J., and J. Lowy. The structure of F‐actin and of actin filaments isolated from muscle. J. Mol. Biol. 6: 46–60, 1963.
 44. Haselgrove, J. C. X‐ray diffraction studies on muscle. Cambridge, UK: Univ. of Cambridge, 1970. Dissertation.
 45. Haselgrove, J. C. X‐ray evidence for a conformational change in the actin‐containing filaments of vertebrate striated muscles. Cold Spring Harbor Symp. Quant. Biol. 37: 341–352, 1972.
 46. Haselgrove, J. C. X‐ray evidence for conformational changes in the myosin filaments of vertebrate striated muscle. J. Mol. Biol. 92: 113–143, 1975.
 47. Haselgrove, J. C. A model of myosin cross‐bridge structure consistent with the low‐angle X‐ray diffraction pattern from vertebrate muscle. J. Muscle Res. Cell Motil. 2: 177–191, 1980.
 48. Haselgrove, J. C., A. R. Faruqi, H. E. Huxley, and V. W. Arndt. The design and use of a camera for low‐angle x‐ray diffraction experiments with synchrotron radiation. J. Phys. E 10: 1035–1046, 1977.
 49. 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.
 50. Haselgrove, J. C., and M. K. Reedy. Modeling rigor cross‐bridge patterns in muscle: initial studies of the rigor lattice of insect flight muscle. Biophys. J. 24: 713–728, 1978.
 51. Haselgrove, J. C., and C. Rodger. The interpretation of X‐ray diffraction patterns from vertebrate striated muscle. J. Muscle Res. Cell Motil. 1: 371–390, 1980.
 52. Haselgrove, J. C., M. Stewart, and H. E. Huxley. Cross‐bridge movement during muscle contraction. Nature London 261: 606–608, 1976.
 53. Holmes, K. C., The myosin cross‐bridge as revealed by structure studies. In: Myocardial Failure, edited by G. Reiker, A. Weber, and J. Goodwin. New York: Springer‐Verlag, 1977, p. 16–27.
 54. Holmes, K. C., and D. M. Blow. The Use of X‐Ray Diffraction in the Study of Protein and Nucleic Acid Structure. New York: Wiley, 1966.
 55. Holmes, K. C., R. S. Goody, H. G. Mannhertz, J. Barrington‐Leigh, and G. Rosenbaum. An investigation of the cross‐bridge cycle using ATP analogues and low‐angle X‐ray diffraction from glycerinated fibres of insect flight muscle. In: Molecular Basis of Motility, edited by L. Heilmeyer, J. C. Ruegg, and T. H. Wieland. Heidelberg: Springer‐Verlag, 1976, p. 26–41. (Ges. Biol. Chem., 26th, Mosbach, Germany, April 10–12, 1975.).
 56. Holmes, K. C., R. T. Tregear, and J. Barrington‐Leigh. Interpretation of the low‐angle diffraction from insect flight muscle. Proc. R. Soc. London Ser. B 207: 13–33, 1980.
 57. Huxley, A. F. Muscle structure and theories of contraction. Prog. Biophys. Biophys. Chem. 7: 255–318, 1957.
 58. Huxley, A. F. Muscular contraction. J. Physiol. London 243: 1–43, 1974.
 59. Huxley, A. F., and R. Niedergeerke. Structural changes in muscle during contraction. Nature London 173: 171–173, 1954.
 60. Huxley, A. F., and R. M. Simmons. Proposed mechanism of force generation in striated muscle. Nature London 233: 533–538, 1971.
 61. Huxley, A. F., and R. M. Simmons. Mechanical transients and the origin of muscular force. Cold Spring Harbor Symp. Quant. Biol. 37: 669–680, 1972.
 62. Huxley, H. E. X‐Ray Diffraction Studies on Muscle. Cambridge, UK: Univ. of Cambridge, 1952. Dissertation.
 63. Huxley, H. E. X‐ray analysis and the problem of muscle. Proc. R. Soc. London Ser. B 141: 59–62, 1953.
 64. Huxley, H. E. The double array of filaments in cross‐striated muscle. J. Biophys. Biochem. Cytol. 3: 631–647, 1957.
 65. Huxley, H. E. Electron microscope studies of the structure of natural and synthetic protein filaments from striated muscle. J. Mol. Biol. 7: 281–308, 1963.
 66. Huxley, H. E. Structural difference between resting and rigor muscle: evidence from intensity changes in the low‐angle equatorial X‐ray diagram. J. Mol. Biol. 37: 507–520, 1968.
 67. Huxley, H. E. The mechanism of muscle contraction. Science 164: 1356–1366, 1969.
 68. Huxley, H. E. Structural changes in the actin‐ and myosin‐containing filaments during contraction. Cold Spring Harbor Symp. Quant. Biol. 37: 361–376, 1972.
 69. 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.
 70. Huxley, H. E., and W. Brown. The low‐angle x‐ray diagram of vertebrate striated muscle and its behaviour during contraction and rigor. J. Mol. Biol. 30: 383–434, 1967.
 71. Huxley, H. E., W. Brown, and K. C. Holmes. Constancy of axial spacings in frog sartorius muscle during contraction. Nature London 206: 1358, 1965.
 72. Huxley, H. E., A. R. Faruqi, J. Bordas, M. H. J. Koch, and J. R. Milch. The use of synchrotron radiation in time‐resolved X‐ray diffraction studies of myosin layer‐line reflections during muscle contraction. Nature London 284: 140–143, 1980.
 73. Huxley, H. E., and J. Hanson. Changes in the cross‐striations of muscle during contraction and stretch and their structural interpretation. Nature London 173: 973–974, 1954.
 74. Huxley, H. E., and J. C. Haselgrove. The structural basis of contraction in muscle and its study by rapid X‐ray diffraction methods. In: Myocardial Failure, edited by G. Reiker, A. Weber, and J. Goodwin. New York: Springer‐Verlag, 1977, p. 4–15.
 75. Huxley, H. E., R. M. Simmons, A. R. Faruqi, M. Kress, and J. Bordas. Millisecond time‐resolved changes in X‐ray reflections from contracting muscle during rapid mechanical transients, recorded using synchrotron radiation. Proc. Natl. Acad. Sci. USA 78: 2297–2301, 1981.
 76. Katsura, I., and H. Noda. Assembly of myosin molecules into the structure of thick filaments of muscle. Adv. Biophys. 5: 177–202, 1973.
 77. Koretz, J. F. Effects of C‐protein on synthetic myosin filament structure. Biophys. J. 27: 433–446, 1979.
 78. Lamvik, M. K. Muscle thick filament mass measured by electron scattering. J. Mol. Biol. 122: 55–68, 1978.
 79. Lowy, J., F. R. Poulsen, and P. J. Vibert. Myosin filaments in vertebrate smooth muscle. Nature London 225: 1053–1054, 1970.
 80. Lowy, J., and P. J. Vibert. Structure and organisation of actin in a molluscan smooth muscle. Nature London 215: 1254–1255, 1967.
 81. Lowy, J., and P. J. Vibert. Studies of the low‐angle X‐ray diffraction patterns of a molluscan smooth muscle during tonic contraction and rigor. Cold Spring Harbor Symp. Quant. Biol. 37: 353–359, 1972.
 82. Lowy, J., P. J. Vibert, J. C. Haselgrove, and F. R. Poulsen. The structure of the myosin elements in vertebrate smooth muscles. Philos. Trans. R. Soc. London Ser. B 265: 191–196, 1973.
 83. Luther, P. K., and J. Squire. Three‐dimensional structure of the vertebrate muscle A band. J. Mol. Biol. 141: 409–439, 1980.
 84. Lymn, R. W. Equatorial X‐ray reflections and cross arm movement in skeletal muscle. Nature London 258: 770–772, 1975.
 85. Lymn, R. W. Low‐angle x‐ray diagrams from skeletal muscle: the effect of AMP‐PNP, a non‐hydrolyzed analogue of ATP. J. Mol. Biol. 99: 567–582, 1975.
 86. Lymn, R. W. Myosin subfragment‐1 attachment to actin. Expected effect on equatorial reflections. Biophys. J. 21: 92–98, 1978.
 87. Lymn, R. W., and H. E. Huxley. X‐ray diagrams from skeletal muscle in the presence of ATP analogues. Cold Spring Harbor Symp. Quant. Biol. 37: 449–453, 1972.
 88. Lymn, R. W., and E. W. Taylor. Mechanism of adenosine triphosphate hydrolysis by actomyosin. Biochemistry 10: 4617–4624, 1971.
 89. Maeda, Y. X‐ray diffraction patterns from molecular arrangements with 38‐nm periodicities around muscle thin filaments. Nature London 277: 670–672, 1979.
 90. Maeda, Y., I. Matsubara, and N. Yagi. Structural changes in thin filaments of crab striated muscle. J. Mol. Biol. 127: 191–201, 1979.
 91. Magid, A., and M. K. Reedy. X‐ray diffraction observations of chemically skinned frog skeletal muscle processed by an improved method. Biophys. J. 30: 27–40, 1980.
 92. Mannhertz, H. G., J. Barrington‐Leigh, K. Holmes, and G. Rosenbaum. Identification of the transitory complex myosin‐ATP by the use of α, β‐methylene‐ATP. Nature London New Biol. 241: 226, 1973.
 93. Marston, S. B., C. D. Rodger, and R. T. Tregear. Changes in muscle crossbridges when beta, gamma‐imido‐ATP binds to myosin. J. Mol. Biol. 104: 263–276, 1976.
 94. Marston, S. B., and R. T. Tregear. Evidence for a complex between myosin and ADP in relaxed muscle fibres. Nature London New Biol. 235: 23–24, 1972.
 95. Marston, S. B., R. T. Tregear, C. D. Rodger, and M. L. Clarke. Coupling between the enzymatic site of myosin and the mechanical output of muscle. J. Mol. Biol. 28: 111–126, 1979.
 96. Maruyama, K., and A. Weber. Binding of ATP to myofibrils during contraction and relaxation. Biochemistry 11: 2990–2998, 1972.
 97. Matsubara, I. X‐ray diffraction studies of the heart. Annu. Rev. Biophys. Bioeng. 9: 81–105, 1980.
 98. Matsubara, I., and G. F. Elliott. X‐ray diffraction studies on skinned single fibres of frog skeletal muscle. J. Mol. Biol. 72: 657–669, 1972.
 99. Matsubara, I., A. Kamiyama, and H. Suga. X‐ray diffraction study of contracting heart muscle. J. Mol. Biol. 111: 121–128, 1977.
 100. Matsubara, I., and B. M. Millman. X‐ray diffraction patterns from mammalian heart muscle. J. Mol. Biol. 82: 527–536, 1974.
 101. Matsubara, I., H. Suga, and N. Yagi. An X‐ray diffraction study of the cross‐circulated canine heart. J. Physiol. London 270: 311–320, 1977.
 102. Matsubara, I., and N. Yagi. A time‐resolved X‐ray diffraction study of muscle during twitch. J. Physiol. London 278: 297–307, 1978.
 103. Matsubara, I., N. Yagi, and M. Endoh. Behaviour of myosin projections during the staircase phenomenon of heart muscle. Nature London 273: 67, 1978.
 104. Matsubara, I., N. Yagi, and M. Endoh. Movement of myosin heads during a heart beat. Nature London 278: 474–476, 1979.
 105. Matsubara, I., N. Yagi, and H. Hashizume. Use of an X‐ray television for diffraction of the frog striated muscle. Nature London 255: 728–729, 1975.
 106. Maw, M. C., and A. Rowe. Fraying of A‐filaments into three subfilaments. Nature London 286: 412–414, 1980.
 107. Miller, A., and R. T. Tregear. Evidence concerning cross‐bridge attachment during muscle contraction. Nature London 226: 1060–1061, 1970.
 108. Miller, A., and R. T. Tregear. X‐ray studies on the structure and function of invertebrate muscle. In: Contractility of Muscle Cells and Related Processes, edited by R. J. Podolsky. London: Prentice‐Hall, 1971.
 109. Miller, A., and R. T. Tregear. Structure of insect fibrillar flight muscle in the presence and absence of ATP. J. Mol. Biol. 70: 85–104, 1972.
 110. Miller, A., and J. Woodhead‐Galloway. Long range forces in muscle. Nature London 229: 470–473, 1971.
 111. Millman, B., and P. M. Bennett. Structure of the cross‐striated adducter muscle of the scallop. J. Mol. Biol. 103: 439–467, 1976.
 112. Millman, B. M., G. F. Elliott, and J. Lowy. Axial period of actin filaments: X‐ray diffraction studies. Nature London 213: 356–358, 1967.
 113. Millman, B. M., T. J. Racey, and I. Matsubara. Effects of hyperosmotic solutions on the filament lattice of intact frog skeletal muscle. Biophys. J. 33: 189–202, 1981.
 114. Moore, P. B., H. E. Huxley, and D. J. DeRosier. Three‐dimensional reconstruction of F‐actin, thin filaments and decorated thin filaments. J. Mol. Biol. 50: 279–295, 1970.
 115. Moos, C., G. Offer, R. Starr, and P. Bennett. Interaction of C‐protein with myosin, myosin rod and light meromyosin. J. Mol. Biol. 97: 1–9, 1975.
 116. Morimoto, K., and W. F. Harrington. Substructure of the thick filament of vertebrate striated muscle. J. Mol. Biol. 83: 83–97, 1974.
 117. Namba, K., K. Wakabayashi, and T. Mitsui. X‐ray structure analysis of the thin filament of crab striated muscle in the rigor state. J. Mol. Biol. 138: 1–26, 1980.
 118. O'Brien, E. J., P. M. Bennett, and J. Hanson. Optical diffraction studies of myofibrillar structure. J. Mol. Biol. 83: 83–97, 1971.
 119. Offer, G. C‐protein and periodicity in the thick filaments of vertebrate skeletal muscle. Cold Spring Harbor Symp. Quant. Biol. 37: 87–95, 1972.
 120. Offer, G., and A. Elliott. Can a myosin molecule bind to two actin filaments? Nature London 271: 325–329, 1978.
 121. Offer, G., C. Moss, and R. Starr. A new protein of the thick filaments of vertebrate skeletal myofibrils. J. Mol. Biol. 74: 653–676, 1973.
 122. Ohtsuki, I. T., Y. Masaki, Y. Nonomura, and S. Ebashi. Periodic distribution of troponin along the thin filament. J. Biochem. Tokyo 61: 817–819, 1967.
 123. Page, S., and H. E. Huxley. Filament lengths in striated muscle. J. Cell Biol. 19: 369–370, 1963.
 124. Parry, D. A. D., and J. M. Squire. Structural role of tropomyosin in muscle regulation: analysis of the X‐ray diffraction patterns from relaxed and contracting muscles. J. Mol. Biol. 75: 33–55, 1973.
 125. Pepe, F. A. The myosin filament. I. Structural organization from antibody staining observed in electron microscopy. J. Mol. Biol. 27: 203–225, 1967.
 126. Pepe, F. A., F. T. Ashton, P. Dowben, and M. Stewart. The myosin filament. VII. Changes in internal structure along the length of the filament. J. Mol. Biol. 145: 421–440, 1981.
 127. Pepe, F. A., and B. Druker. The myosin filament. J. Mol. Biol. 130: 379–393, 1979.
 128. Podolsky, R. J., and A. C. Nolan. Muscle contraction transients, cross‐bridge kinetics and the Fenn effect. Cold Spring Harbor Symp. Quant. Biol. 37: 661–668, 1972.
 129. Podolsky, R. J., H. St. Onge, L. Yu, and R. W. Lymn. X‐ray diffraction of actively shortening muscle. Proc. Natl. Acad. Sci. USA 73: 813–817, 1976.
 130. Potter, J. The content of troponin, tropomyosin, actin and myosin in rabbit skeletal muscle myofibrils. Arch. Biochem. Biophys. 162: 436–441, 1974.
 131. Pringle, J. W. S., The mechanical characteristics of insect fibrillar muscle. In: Insect Flight Muscle, edited by R. T. Tregear. Amsterdam: Elsevier/North‐Holland, 1977, p. 177–196.
 132. Reedy, M. K. Ultrastructure of insect flight muscle. J. Mol. Biol. 31: 155–176, 1968.
 133. Reedy, M. K., G. F. Bahr, and D. A. Fischman. How many myosins per cross bridge? I. Flight muscle myofibrils from the blowfly. Cold Spring Harbor Symp. Quant. Biol. 37: 397–422, 1972.
 134. 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.
 135. Rodger, C. D., and R. T. Tregear. Crossbridge angle when ADP is bound to myosin. J. Mol. Biol. 86: 495–497, 1974.
 136. Rome, E. Light and X‐ray diffraction studies of the filament lattice of glycerol‐extracted rabbit psoas muscle. J. Mol. Biol. 27: 591–602, 1967.
 137. Rome, E. X‐ray diffraction studies of the filament lattice of striated muscle in various bathing media. J. Mol. Biol. 37: 331–344, 1968.
 138. Rome, E. Relaxation of glycerinated muscles: low‐angle X‐ray diffraction studies. J. Mol. Biol. 65: 331–345, 1972.
 139. Rome, E. Structural studies by X‐ray diffraction of striated muscle permeated with certain ions and proteins. Cold Spring Harbor Symp. Quant. Biol. 37: 331–339, 1972.
 140. Rome, E. M., T. Hirabayashi, and S. V. Perry. X‐ray diffraction of muscle labelled with antibody to troponin‐C. Nature London New Biol. 244: 154–155, 1973.
 141. Rome, E. M., G. Offer, and F. A. Pepe. X‐ray diffraction of muscle labelled with antibody to C‐protein. Nature London New Biol. 244: 152–154, 1973.
 142. Rüdel, R., and F. Zite‐Ferenczy. Interpretation of light diffraction by cross‐striated muscle as Bragg reflection of light by the lattice of contractile proteins. J. Physiol. London 290: 317–330, 1979.
 143. Selby, C., and R. S. Bear. The structure of actin‐rich filaments of muscles according to X‐ray diffraction. J. Biophys. Biochem. Cytol. 2: 71–85, 1956.
 144. Seymour, J., and E. J. O'Brien. The position of tropomyosin in muscle thin filaments. Nature London 283: 680–681, 1980.
 145. Shear, D. B. Electrostatic forces in muscle contraction. J. Theor. Biol. 28: 531–546, 1970.
 146. Shoenberg, C. F., and J. C. Haselgrove. Filaments and ribbons in vertebrate smooth muscle. Nature London 249: 152–154, 1974.
 147. Sjöström, M., and J. M. Squire. Fine structure of the A‐band in cryo‐sections. The structure of the A‐band of human skeletal muscle fibres from ultra‐thin cryo‐sections negatively stained. J. Mol. Biol. 109: 49–68, 1977.
 148. Slayter, H. S., and S. Lowey. Substructure of the myosin molecule as visualized by electron microscopy. Proc. Natl. Acad. Sci. USA 58: 1611–1618, 1967.
 149. Spudich, J. A., H. E. Huxley, and J. T. Finch. Regulation of skeletal muscle contraction. II. Structural studies of the interaction of the tropomyosin‐troponin complex with actin. J. Mol. Biol. 72: 619–632, 1972.
 150. Spudich, J. A., and S. Watt. The regulation of rabbit skeletal muscle contraction. J. Biol. Chem. 246: 4866–4871, 1971.
 151. Squire, J. M. General model for the structure of all myosin‐containing filaments. Nature London 233: 457–462, 1971.
 152. Squire, J. M. General model of myosin filament structure. II. Myosin filaments and cross‐bridge interactions in vertebrate striated and insect flight muscles. J. Mol. Biol. 72: 125–138, 1972.
 153. Squire, J. M. General model of myosin filament structure. III. Molecular packing arrangements in myosin filaments. J. Mol. Biol. 77: 291–323, 1973.
 154. Squire, J. M. Muscle filament structure and muscle contraction. Annu. Rev. Biophys. Bioeng. 4: 137–163, 1975.
 155. Sugi, H., Y. Amemiya, and H. Hashizume. Time‐resolved X‐ray diffraction from skeletal muscle during the course of an after‐loaded isotonic twitch. In: Proc. Int. Symp. Biophys., 4th, Ibaraki, 1978. Ibaraki, Japan: Taniguchi Found., 1978, p. 164–176.
 156. Szent‐Gyorgyi, A. G., C. Cohen, and D. E. Philpott. Light meromyosin fraction I: a helical molecule from myosin. J. Mol. Biol. 2: 133–142, 1960.
 157. Taylor, K. A., and L. A. Amos. A new model for the geometry of the binding of the myosin crossbridges to muscle thin filaments. J. Mol. Biol. 147: 297–324, 1981.
 158. Thomas, D. D., and R. Cooke. The measurement of myosin head orientation in muscle fibers using nitroxide spin labels. Biophys. J. 25: 19a, 1979.
 159. Thomas, D. D., and R. Cooke. Orientation of spin labelled myosin heads in glycerinated muscle fibers. Biophys. J. 32: 891–906, 1980.
 160. Tregear, R. T. (editor). Insect Flight Muscle. Amsterdam: Elsevier/North‐Holland, 1977.
 161. Tregear, R. T., J. R. Milch, R. S. Goody, K. C. Holmes, and C. D. Rodger. The use of some novel X‐ray diffraction techniques to study the effect of nucleotides on cross‐bridges in insect flight 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.
 162. Tregear, R. T., and A. Miller. Evidence of crossbridge movement during contraction of insect flight muscle. Nature London 222: 1184–1185, 1969.
 163. Tregear, R. T., and J. M. Squire. Myosin content and filament structure in smooth and striated muscle. J. Mol. Biol. 77: 279–290, 1973.
 164. Ullrick, W. A theory of contraction for striated muscle. J. Theor. Biol. 15: 53–69, 1976.
 165. Vainshtein, B. K. Diffraction of X‐Rays by Chain Molecules. Amsterdam: Elsevier, 1966.
 166. Vibert, P. J., J. C. Haselgrove, J. Lowy, and F. R. Poulsen. Structural changes in actin‐containing filaments of muscle. J. Mol Biol. 71: 757–767, 1972.
 167. Vibert, P. J., J. Lowy, J. C. Haselgrove, and F. R. Poulsen. Structural changes in actin filaments of muscle. Nature London New Biol. 236: 182–183, 1972.
 168. Vibert, P., A. G. Szent‐Gyorgyi, R. Craig, J. Wray, and C. Cohen. Changes in crossbridge attachment in a myosin‐regulated muscle. Nature London 273: 64–66, 1978.
 169. Wakabayashi, T., H. E. Huxley, L. A. Amos, and A. Klug. Three‐dimensional image reconstruction of actin‐tropomyosin complex and actin‐tropomyosin‐troponin T‐troponin I complex. J. Mol. Biol. 93: 477–497, 1975.
 170. Weber, A., and R. D. Bremel. Regulation of contraction and relaxation in the myofibril. In: Contractility of Muscle Cells and Related Processes, edited by R.J. Podolsky. London: Prentice‐Hall, 1971, p. 37–53.
 171. Worthington, C. R. Large axial spacings in striated muscle. J. Mol. Biol. 1: 398–401, 1959.
 172. Wray, J. S. Structure of the backbone in myosin filaments of muscle. Nature London 277: 37–40, 1979.
 173. Wray, J. S. Filament geometry and the activation of flight muscles. Nature London 280: 325–326, 1980.
 174. Wray, J. S., P. J. Vibert, and C. Cohen. Cross‐bridge arrangements in Limulus muscle. J. Mol. Biol. 88: 343–348, 1974.
 175. Wray, J. S., P. J. Vibert, and C. Cohen. Diversity of cross‐bridge configurations in invertebrate muscles. Nature London 257: 561–564, 1975.
 176. Wray, J., P. J. Vibert, and C. Cohen. Actin filaments in muscle: pattern of myosin and tropomyosin/troponin attachments. J. Mol. Biol. 124: 501–521, 1978.
 177. Yagi, N., M. H. Ito, H. Nakajima, T. Izumi, and I. Matsubara. Return of myosin heads to thick filaments after muscle contraction. Science 197: 685–687, 1977.
 178. Yagi, N., and I. Matsubara. Recent progress in fast X‐ray diffraction of muscle. In: Proc. Int. Symp. Biophys., 4th, Ibaraki, 1978. Ibaraki, Japan: Taniguchi Found., 1978, p. 142–151.
 179. Yagi, N., and I. Matsubara. Myosin heads do not move on activation in highly stretched vertebrate striated muscle. Science 207: 307–308, 1980.
 180. Yagi, N., E. J. O'Brien, and I. Matsubara. Changes of thick filament structure during contraction of frog striated muscle. Biophys. J. 33: 121–138, 1981.
 181. Yount, R. G., D. Ojala, and D. Babcock. Interaction of P‐N‐P and P‐C‐P analogues of adenosine triphosphate with heavy meromyosin, myosin and actomyosin. Biochemistry 10: 2490–2496, 1971.
 182. 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.
 183. Yu, L. C., R. W. Lymn, and R. J. Podolsky. Characterization of a non‐indexable equatorial X‐ray reflection from frog sartorius muscle. J. Mol. Biol. 115: 455–464, 1977.

Contact Editor

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

* Required Field

How to Cite

John C. Haselgrove. Structure of Vertebrate Striated Muscle as Determined by X‐ray‐Diffraction Studies. Compr Physiol 2011, Supplement 27: Handbook of Physiology, Skeletal Muscle: 143-171. First published in print 1983. doi: 10.1002/cphy.cp100105