Hostname: page-component-8448b6f56d-sxzjt Total loading time: 0 Render date: 2024-04-18T12:19:28.774Z Has data issue: false hasContentIssue false

Ca2+ and activation mechanisms in skeletal muscle

Published online by Cambridge University Press:  17 March 2009

Christopher C. Ashley
Affiliation:
University Laboratory of Physiology, Parks Road, Oxford OXi 3PT, UK
Ian P. Mulligan
Affiliation:
University Laboratory of Physiology, Parks Road, Oxford OXi 3PT, UK
Trevor J. Lea
Affiliation:
University Laboratory of Physiology, Parks Road, Oxford OXi 3PT, UK

Extract

It has been known for a number of years that calcium ions play a crucial role in excitation-contraction (e-c) coupling (Sandow, 1952). The majority of the calcium required for this process is derived, at least in vertebrate striated muscle fibres, from discrete intracellular stores located at sites within the cell: the terminal cysternae (tc)/junctional SR of the sarcoplasmic reticulum (SR) (Fig. 1 a). These storage sites not only form a compartment that is distinct from the sarcoplasm of the fibre, but they are also closely associated with the contractile elements, the myofibrils. The SR release sites are activated following the spread of electrical activity (Huxley and Taylor, 1958) along the transverse (T) tubular system (Eisenberg and Gage, 1967; Adrian et al. 1969a, b; Peachey, 1973) from the surface membrane (Bm).

Type
Research Article
Copyright
Copyright © Cambridge University Press 1991

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Adams, S. R., Kao, J. P. Y. & Tsien, R. Y. (1986). Photolabile chelators that ‘cage’ calcium with improved speed of release and pre-photolysis affinity. J. gen. Physiol. 88, 9a10a.Google Scholar
Adams, S. R., Kao, J. P. Y. & Tsien, R. Y. (1989). Biological useful chelators that take up Ca2+ upon illumination. J. Am. Chem. Soc. 111, 79577968.CrossRefGoogle Scholar
Adams, S. R., Kao, J. P. Y., Grynkiewicz, G., Minta, A. & Tsien, R. Y. (1988). Biologically useful chelators that release Ca2+ upon illumination. J. Am. Chem. Soc. 110, 32123220.Google Scholar
Adams, B. A., Tanabe, T., Mikami, A., Numa, S. & Beam, K. G. (1990). Intramembrane charge movement restored in dysgenic skeletal muscle by injection of dihydro-pyridine receptor cDNAs. Nature 346, 569572.Google Scholar
Adrian, R. H. & Huang, C.-H. (1984). Experimental analysis of the relationship between charge movement components in skeletal muscle of Rana temporaria. J. Physiol. 353, 419434.Google Scholar
Adrian, R. H., Chandler, W. K. & Hodgkin, A. L. (1969a). The kinetics of mechanical activation in frog muscle. J. Physiol. 204, 207230.CrossRefGoogle ScholarPubMed
Adrian, R. H., Costantin, L. L. & Peachey, L. D. (1969b). Radial spread of contraction in frog muscle fibres. J. Physiol. 204, 231257.CrossRefGoogle ScholarPubMed
Allen, D. G. & Blinks, J. R. (1978). Calcium transients in aequorin-injected cardiac muscle. Nature 273, 509513.CrossRefGoogle ScholarPubMed
Allen, D. G., Blinks, J. R. & Pendergast, F. G. (1977). Aequorin luminescence: relation of light emission to calcium concentration– a calcium-independent component. Science 195, 996998.Google Scholar
Anderson, K., Grunwald, R., El-Hashem, A., Sealock, R. & Meissner, G. (1990). High affinity ryanodine and PN200–110 binding to rabbit skeletal muscle triads. Biophys. J. 57, 171a.Google Scholar
Armstrong, C. M., Bezanilla, F. & Horowicz, P. (1972). Twitches in the presence of EGTA. Biochim. biophys. Acta 267, 605608.Google Scholar
Ashley, C. C. (1970). An estimate of calcium concentration changes during the contraction of single muscle fibres. J. Physiol. 210, 133134P.Google Scholar
Ashley, C. C. (1978). Calcium ion regulation in barnacle muscle fibers and its relation to force development. Ann. N. Y. Acad. Sci. 307, 308329.Google Scholar
Ashley, C. C. (1983). Calcium in muscle. In Calcium in Biology, (ed. Spiro, T. G.), pp. 109173. New York: Wiley.Google Scholar
Ashley, C. C. & Caldwell, P.C. (1974). Calcium movements in relation to contraction. Biochem. Soc. Symp. 39, 2950.Google Scholar
Ashley, C. C. & Campbell, A. K. (1979). Detection and Measurement of free Ca2+ in Cells. Amsterdam: Elsevier North Holland Biomedical Press.Google Scholar
Ashley, C. C. & Moisescu, D. G. (1972a). Model for the action of calcium in muscle. Nature (New Biol.), 237, 208211.CrossRefGoogle ScholarPubMed
Ashley, C. C. & Moisescu, D. G. (1972b). Tension changes in isolated muscle fibres as predicted by the free calcium concentration. J. Physiol. 226, 8284P.Google ScholarPubMed
Ashley, C. C. & Moisescu, D. G. (1973 a). Tension changes in isolated bundles of frog and barnacle myofibrils in response to sudden changes in the external free calcium concentration. J. Physiol. 233, 89P.Google ScholarPubMed
Ashley, C. C. & Moisescu, D. G. (1973b). The mechanism of the free calcium change in single muscle fibres. J. Physiol. 231, 2325P.Google ScholarPubMed
Ashley, C. C. & Moisescu, D. G. (1974). The influence of Mg2+ concentration and of pH upon the relationship between steady-state isometric tension and Ca2+ concentration in isolated bundles of barnacle myofibrils. J. Physiol. 239, 112114P.Google Scholar
Ashley, C. C. & Moisescu, D. G. (1975). The part played by Ca2+ in the contraction of isolated bundles of myofibrils. In Calcium Transport in Contraction and Secretion (ed. Carafoli, E. et al. ), pp. 517525. Amsterdam: North-Holland.Google Scholar
Ashley, C. C. & Moisescu, D. G. (1977). Effects of changing the composition of the bathing solutions upon the isometric tension-pCa relationship in bundles of crustacean myofibrils. J. Physiol. 270, 627652.Google Scholar
Ashley, C.C. & Ridgway, E. B. (1970). On the relationship between membrane potential, calcium transient and tension in single barnacle muscle fibres. J. Physiol. 209, 105130.Google Scholar
Ashley, C. C., Moisescu, D. G. & Rose, R. M. (1974a). Aequorin-light and tension responses from bundles of myofibrils following a sudden change in free Ca2+. J. Physiol. 241, 104106P.Google Scholar
Ashley, C. C., Moisescu, D. G. & Rose, R. M. (1974b). Kinetics of calcium during contraction: myofibrillar and SR fluxes during a single response of a skeletal muscle fibre. In Calcium Binding Proteins (ed. Drabikowski, W. et al. ), pp. 609642. Amsterdam: Elsevier.Google Scholar
Ashley, C. C., Rink, T. J. & Tsien, R. Y. (1978). Changes in free calcium during muscle contraction, measured with an intracellular Ca2+-sensitive electrode. J. Physiol. 280, 27P.Google Scholar
Ashley, C. C., Barsotti, R. J., Ferenczi, M. A., Lea, T. J. & Mulligan, I. P. (1987a). Fast activation of skinned muscle fibres from the frog by photolysis of caged-calcium. J. Physiol. 394, 24P.Google Scholar
Ashley, C. C., Barsotti, R. J., Ferenczi, M. A., Lea, T. J., Mulligan, I. P. & Tsien, R. Y. (1987b). Caged-calcium photolysis activates demembranated muscle fibres from the rabbit. J. Physiol. 390, 144P.Google Scholar
Ashley, C. C., Barsotti, R. J., Ferenczi, M. A., Lea, T. J. & Mulligan, I. P. (1988a). Simultaneous photolysis of‘caged’-ATP and ‘caged’-calcium in permeabilised single muscle fibres from the frog. J. Physiol. 398, 71P.Google Scholar
Ashley, C. C., Griffiths, P. J. & Potter, J. D. (1988b). The mobility of TnCDANZ following injection into barnacle muscle fibres. J. Physiol. 399, 20P.Google Scholar
Ashley, C. C., Barsotti, R. J., Ferenczi, M. A., Lea, T. J. & Mulligan, I. P. (1989a). Thin filament activation by photolysis of caged calcium in skinned muscle fibres. In Biochemical Approaches to Cellular Calcium, vol. 19 (ed. Reid, E., Cook, G. M. W. and Luzio, J. P.), pp. 131132. London: Royal Society of Chemistry.Google Scholar
Ashley, C. C., Lea, T. J., Mulligan, I. P.Timmerman, M. P. (1989b). Calciuminduced calcium release from the sarcoplasmic reticulum of Balanus striated muscle using laser-induced photolysis of nitr-5. J. Physiol. 414, 50P.Google Scholar
Ashley, C. C., Mulligan, I. P. & Palmer, R. (1990). ADP slows the relaxation of single permeabilised muscle fibres from frog following flash photolysis of the caged Ca2+ chelator, diazo-2. J. Physiol. 426, 31P.Google Scholar
Bagni, M. A., Cecchi, G. & Schoenberg, M. (1988). A model for force production that explains the lag between cross-bridge attachment and force after electrical stimulation of striated muscle fibres. Biophys. J. 54, 11051114.Google Scholar
Bailey, K. (1948). Tropomyosin: a new asymmetric protein component of the muscle fibril. Biochem. J. 43, 271279.Google Scholar
Baker, P. F. (1972). Transport and metabolism of calcium ions in nerve. Prog. Biophys. molec. Biol. 24, 177223.CrossRefGoogle ScholarPubMed
Baltrop, J. A., Plant, P. J. & Schofield, P. (1966). Photosensitive protecting groups. Chem. Commun., 822823.Google Scholar
Barsotti, R. J., Kentish, J. C., Lea, T. J. & Mulligan, I. P. (1988). Laser-induced photolysis of nitr-5 triggers Ca2+ release from the sarcoplasmic reticulum of saponintreated muscles from guinea-pig ventricle. J. Physiol. 396, 80P.Google Scholar
Baylor, S. M. & Hollingworth, S. (1988). Fura-2 calcium transients in frog skeletal muscle fibers. J. Physiol. 403, 151192.CrossRefGoogle Scholar
Baylor, S. M., Chandler, W. K. & Marshall, M. W. (1982). Optical measurements of intracellular pH and Mg in frog skeletal muscle fibers. J. Physiol. 331, 105137.CrossRefGoogle Scholar
Baylor, S. M., Chandler, W. K. & Marshall, M. W. (1983). Sarcoplasmic reticulum Ca2+ release in frog skeletal muscle fibres estimated from ASIII calcium transients. J. Physiol. 344, 625666.CrossRefGoogle Scholar
Baylor, S. M., Hollingworth, S. & Marshall, M. W. (1988). Effect of intracellular ruthenium red on excitation-contraction coupling in skeletal muscle. Biophys. J. 53, 647a.Google Scholar
Beam, K. G., Knudson, C. M. & Powell, J. A. (1986). A lethal mutation in mice eliminates the slow calcium current in skeletal muscle cells. Nature 320, 168170.Google Scholar
Binder, K. (ed). (1983). Applications of the Monte Carlo Method in Statistical Physics. Topics in Current Physics, vol. 11. Berlin: Springer-Verlag.Google Scholar
Binder, K. (ed). (1986). Monte Carlo Method in Statistical Physics. Topics in Current Physics, vol. 7, 2nd edn. Berlin: Springer-Verlag.Google Scholar
Blinks, J. R., Rudel, R. & Taylor, S. R. (1978). Calcium transients in isolated muscle fibres: detection with aequorin. J. Physiol. 277, 291323.Google Scholar
Blinks, J. R., Wier, G., Hess, P. & Prendergast, F. G. (1982). Measurement of Ca2+ in living cells. Prog. Biophys. molec. Biol. 40, 1114.Google Scholar
Blinks, J. R., Cai, Y.-D. & Lee, N. K. M. (1987). Ins(1, 4, 5)P3 causes Ca2+ release in frog skeletal muscle only when transverse tubules have been interrupted. J. Physiol. 394, 23P.Google Scholar
Block, B. A., Imagawa, T., Campbell, K. P. & Franzini-Armstrong, C. (1988). Structural evidence for direct interaction between the molecular components of the transverse tubule/SR junction in skeletal muscle. J. Cell Biol. 107, 25872600.CrossRefGoogle Scholar
Brandt, P. W., Cox, R. N. & Kawai, M. (1980). Can the binding of Ca2+ to two regulatory sites on troponin C determine the steep pCa/tension relationship of skeletal muscle? Proc. natn. Acad. Sci. USA 77, 47174720.CrossRefGoogle ScholarPubMed
Brandt, P. W., Cox, R. N., Kawai, M. & Robinson, T. (1982). Effect of cross-bridge kinetics on apparent Ca2+ sensitivity. J. gen. Physiol. 79, 9971016.Google Scholar
Brandt, P. W., Diamond, M. S. & Rutchik, J. S. (1987). Co-operative interactions between troponin-tropomyosin units extend the length of the thin filament in skeletal muscle. J. molec. Biol. 195, 885896.Google Scholar
Brandt, N. R., Caswell, A. H., Wen, S.-R. & Talvenheimo, J. A. (1990). Molecular interactions of the junctional foot protein and dihydropyridine receptor in skeletal muscle tetrads. J. Membr. Biol. 113, 237252.Google Scholar
Bremel, R. D. & Weber, A. (1972). Cooperative behaviour within the functional unit of the actin filament in vertebrate skeletal muscle. Nature (New Biol.), 238, 97101.CrossRefGoogle Scholar
Brum, G., Stefani, E. & Rios, E. (1986). Simultaneous measurements of Ca2+ currents, and intracellular Ca2+ concentrations in single skeletal muscle fibers of the frog. Can. J. Physiol. Pharmacol. 65, 681685.CrossRefGoogle Scholar
Caldwell, P. C. (1958). Studies of the internal pH of large muscle and nerve fibres. J. Physiol. 142, 2262.Google Scholar
Cannell, M. B. (1986). Effect of tetanus duration on the free calcium during the relaxation of frog skeletal muscle fibres. J. Physiol. 376, 203218.CrossRefGoogle ScholarPubMed
Cannell, M. B. & Allen, D. G. (1984). Model of calcium movements during activation in the sarcomere of frog skeletal muscle. Biophys. J. 45, 913925.CrossRefGoogle ScholarPubMed
Castellani, L., Franzini-Armstrong, C. & Loesser, K. (1989). Shape, size and disposition of feet injunctions between transverse tubules and sarcoplasmic reticulum of Bivalvia, Insecta, Crustacea and Arachnida. J. Physiol. 418, 118P.Google Scholar
Caswell, A. H. & Brandt, N. R. (1989). Does muscle activation occur by direct mechanical coupling of transverse tubules to sarcoplasmic reticulum? TIBS 14, 161165.Google Scholar
Cecchi, G., Griffiths, P. J. & Taylor, S. R. (1982). Muscular contraction: kinetics of cross-bridge attachment studied by high frequency stiffness measurements. Science, 217, 7072.Google Scholar
Cecchi, G., Griffiths, P. J., Bagni, M. A., Ashley, C. C. & Maeda, Y. (1990). Radial component of cross-bridge force detected by lattice spacing changes in intact single fibres. Science, 250, 14091411.CrossRefGoogle Scholar
Chadwick, C. C. & Fleischer, S. (1990). Purification of the inositol 1.4.5-trisphosphate receptor (IP3REC) from smooth muscle microsomes. Biophys. J. 57, 285a.Google Scholar
Chalovich, J. M., Chock, P. B. & Eisenberg, E. (1981). Mechanism of action of troponin-tropomyosin. J. biol. Chem. 256, 575578.CrossRefGoogle ScholarPubMed
Chen, Y. D. & Hill, T. L. (1983). Use of Monte Carlo calculations in the study of microtubule subunit kinetics. Proc. natn. Acad. Sci. USA 80, 75207523.Google Scholar
Collins, J. H., Potter, J. D., Horn, M. J., Wiltshire, G. & Jackson, N. (1973). The amino acid sequence of rabbit skeletal muscle troponin C. FEBS Lett. 36, 268272.CrossRefGoogle ScholarPubMed
Collins, J. H., Greaser, M. L., Potter, J. D. & Horn, M. J. (1977). Determination of the amino acid sequence of troponin C from rabbit skeletal muscle. J. biol. Chem. 252, 63566362.CrossRefGoogle ScholarPubMed
Donaldson, S. K., Goldberg, N. D., Walseth, T. F. & Huetteman, D. A. (1987). Inositol trisphosphate stimulates Ca2+ release from peeled skeletal muscle fibers. Biochim. biophys. Acta 927, 9299.CrossRefGoogle ScholarPubMed
Donaldson, S. K., Goldberg, N. D., Walseth, T. F. & Huetteman, D. A. (1988). Voltage dependence of Ins(1, 4, 5)P3-induced Ca2+ release in peeled skeletal muscle fibers. Proc. natn. Acad. Sci. USA 85, 57495753.Google Scholar
Ebashi, S. (1963). Third component participating in the superprecipitation of natural ‘actomyosin’. Nature 22, 10101012.CrossRefGoogle Scholar
Ebashi, S. & Ebashi, F. (1964). A new protein component participating in the superprecipitation of myosin B. J. Biochem. (Tokyo), 55, 604613.CrossRefGoogle ScholarPubMed
Ebashi, S., Kodama, A. & Ebashi, F. (1968). Troponin. I: Preparation and physiological function. J. Biochem. (Tokyo), 64, 465477.Google Scholar
Ebashi, S., Endo, M. & Ohsuki, I. (1969). Control of muscle contraction. Q. Rev. Biophys. 2, 351384.Google Scholar
Edwards, C. & Lorkowic, H. (1967). The role of Ca2+ in excitation-contraction coupling in various muscles of the frog, mouse and barnacle. Am. Zool. 7, 615622.Google Scholar
Ecelman, E. H. (1985). The structure of F-actin. J. Muscle Res. Cell Motil. 6, 129151.Google Scholar
Ehrlich, B. E. & Watras, J. (1988). Inositol 1, 4, 5-trisphosphate activates a channel from smooth muscle SR. Nature 336, 583586.Google Scholar
Eigen, M. (1963). Fast elementary steps in chemical reaction mechanisms. Pure Appl. Chem. 6, 97115.CrossRefGoogle Scholar
Eisenberg, R. S. & Gage, P. W. (1967). Frog skeletal muscle fibres: changes in electrical properties after T system disruption. Science, 158, 17001701.Google Scholar
El-Saleh, S. C., Warber, K. D. & Potter, J. D. (1986). The role of tropomyosintroponin in the regulation of skeletal muscle contraction. J. Muscle Res. Cell. Motil. 7, 387404.Google Scholar
Endo, M. (1973). Length dependence of activation of skinned muscle fibres by calcium. Cold Spring Harbor Symp. Quant. Biol. 37, 505510.Google Scholar
Endo, M. (1977). Calcium release from the sarcoplasmic reticulum. Physiol. Rev., 57, 71108.Google Scholar
Endo, M. (1981). Mechanism of calcium-induced calcium release in the SR membrane. In The Mechanism of Gated Ca2+ Transport across Biological Membranes (ed. Ohnishi, S. T. and Endo, M.), pp. 257264. New York: Academic Press.CrossRefGoogle Scholar
Endo, M., Tanaka, M. & Ogawa, Y. (1970). Ca-induced Ca2+ release from the sarcoplasmic reticulum of skinned skeletal muscle fibers. Nature 228, 3436.Google Scholar
Engels, J. & Schlaeger, E. J. (1977). Synthesis, structure and reactivity of adenosine cyclic 3′, 5′ -phosphate benzyl triesters. J. Med. Chem. 20, 907911.Google Scholar
Fabiato, A. (1983). Calcium-induced Ca2+ release from the cardiac sarcoplasmic reticulum. Am. J. Physiol. 245, C114.Google Scholar
Fabiato, A. (1985). Calcium-induced release of calcium from the sarcoplasmic reticulum of skinned fibers from the frog semitendinosus. Biophys. J. 47, 195a.Google Scholar
Fabiato, A. & Fabiato, F. (1978). Effect of pH on the myofilaments and the SR of skinned cells from cardiac and skeletal muscles, J. Physiol. 276, 233255.CrossRefGoogle Scholar
Ferris, C. D., Huganir, R. L., Supattapone, S. & Snyder, S. H. (1989). Purified inositol 1, 4, 5-trisphosphate receptor mediates Ca2+ flux in reconstituted lipid vesicles. Nature 342, 8789.CrossRefGoogle Scholar
Fill, M. & Coronado, R. (1988). Ryanodine receptor channel of sarcoplasmic reticulum. TINS 11, 453457.Google Scholar
Fill, M., Coronado, R., Mickelson, J. R., Kilven, J., MA, J., Jacobsen, B. A. & Lonis, C. F. (1990). Abnormal ryanodine receptor channels in malignant hyperthermia. Biophys. J. 57, 471475.CrossRefGoogle ScholarPubMed
Fleischer, S. & Inui, M. (1989). Biochemistry and biophysics of excitation-contraction coupling. Ann. Rev. Biophys. biophys. Chem. 18, 333364.Google Scholar
Fleischer, S., Ogunbunni, E. M., Dixon, M. C. & Fleer, E. A. (1985). Localisation of Ca2+ release channels with ryanodine in junctional terminal cisternae of SR of fast skeletal muscle. Proc. natn. Acad. Sci. USA 82, 72567259.Google Scholar
Ford, L. E. & Podolsky, R. J. (1970). Regenerative calcium release within muscle cells. Science 167, 5859.Google ScholarPubMed
Ford, L. E., Huxley, A. F. & Simmons, R. M. (1977). Tension responses to sudden length changes in stimulated frog muscle fibres near slack length. J. Physiol. 269, 441515.Google Scholar
Fosset, M., Jaimovich, E., Delpont, E. & Lazdunski, M. (1983). [3H]nitrendipine receptors in skeletal muscle. J. biol. Chem. 258, 60866092.Google Scholar
Franzini-Armstrong, C. (1970). Studies of the triad. I: Structure of the junction in frog twitch fibres. J. Cell Biol. 47, 488499.Google Scholar
Franzini-Armstrong, C. (1975). Membrane particles and transmission at the triad. Fed. Proc. 34, 13821389.Google ScholarPubMed
Fryer, M. W., Lamb, G. D. & Neering, I. R. (1989). The action of ryanodine on rat fast and slow intact skeletal muscles. J. Physiol. 414, 399413.CrossRefGoogle ScholarPubMed
Fuchs, F. (1985). The binding of calcium to detergent-extracted rabbit psoas muscle fibres during relaxation and force generation. J. Muscle Res. Cell Motil. 6, 477486.CrossRefGoogle ScholarPubMed
Furuichi, Y., Yoshikawa, S., Mlyawaki, A., Wada, K., Maeda, N. & Mlkoshiba, K. (1989). Primary structure and functional expression of the inositol 1, 4, 5-trisphosphatebinding protein P400. Nature 342, 3238.CrossRefGoogle ScholarPubMed
Garcia, J. & Stefani, E. (1990). Calcium transients in rat skeletal muscle: evidence for a Ca2+-regulated Ca2+ release process. Biophys. J. 57, 400a.Google Scholar
Garcia, J., Pizzaro, G., Rios, E. & Stefani, E. (1990). Depletion of the SR reduces the delayed charge movement of frog skeletal muscle. Biophys. J. 57, 341a.Google Scholar
Gill, D. L. (1989). Receptor kinships revealed. Nature 342, 1618.Google ScholarPubMed
Gillis, J. M. (1985). Relaxation of vertebrate skeletal muscle. A synthesis of the biochemical and physiological approaches. Biochim. biophys. Acta 811, 97145.CrossRefGoogle ScholarPubMed
Gillis, J. M., Thomason, D., Lefevre, J. & Kretsinger, R. H. (1982). Parvalbumins and muscle relaxation: a computer simulation study. J. Muscle Res. Cell Motil. 3, 377398.CrossRefGoogle ScholarPubMed
Goblet, C. & Mounier, Y. (1986). Calcium-induced Ca2+ release mechanism from the SR in skinned crab muscle fibers. Cell Calcium 7, 6172.CrossRefGoogle Scholar
Goldman, Y. E., Hibberd, M. G., McCray, J. A. & Trentham, D. R. (1982). Relaxation of muscle fibres by photolysis of caged ATP. Nature 300, 701705.CrossRefGoogle ScholarPubMed
Gordon, A. M. & Ridgway, E. B. (1987). Extra Ca2+ on shortening in barnacle muscle. J. gen. Physiol. 90, 321340.Google Scholar
Grabarek, Z. & Gergely, J. (1983). On the applicability of Hill type analysis to fluorescence data. J. biol. Chem. 258, 1410314105.CrossRefGoogle ScholarPubMed
Grabarek, Z., Grabarek, J., Leavis, P. C. & Gergely, J. (1983). Cooperative binding to the Ca2+-specific sites of troponin C in regulated actin and actomyosin. J. biol. Chem. 258, 1409814102.Google Scholar
Grabarek, Z., Tan, R.-Y., Wang, J., Tao, T. & Gergely, J. (1990). Inhibition of mutant troponin C activity by intra-domain disulphide bond. Nature 345, 132135.Google ScholarPubMed
Greaser, M. L. & Gergely, J. (1971). Reconstitution of troponin activity from three protein components. J. biol. Chem. 246, 42264233.Google Scholar
Greaser, M. L. & Gergely, J. (1973). Purification and properties of the components from troponin. J. biol. Chem. 248, 21252133.Google Scholar
Greene, L. E. & Eisenberg, E. (1980). Cooperative binding of myosin subfragment to the actin-troponin-tropomyosin complex. Proc. natn. Acad. Sci. USA 77, 26162620.Google Scholar
Griffiths, P. J., Potter, J. D., Coles, B., Strang, P. & Ashley, C. C. (1984). Fluorescence changes from single striated muscle fibres injected with labelled troponin C (TnCDANZ). FEBS Lett. 176, 144150.Google Scholar
Griffiths, P. J., Potter, J. D., Maeda, Y. & Ashley, C. C. (1988). Transient kinetics and time resolved X-ray diffraction studies in isolated single muscle fibres. Adv. Exp. Med. Biol. 226, 113129.Google Scholar
Griffiths, P. J., Duchateau, J. J., Maeda, Y., Potter, J. D. & Ashley, C. C. (1990). Mechanical characteristics of skinned intact muscle fibres from the giant barnacle. B. nubilus. Pfl¨g. Archiv 415, 554565.Google Scholar
Grynkiewiez, G., Poenie, M. & Tsien, R. Y. (1985). A new generation of Ca2+ indicators with greatly improved fluorescence properties. J. biol. Chem. 260, 34403450.Google Scholar
Gurney, A. M., Tsien, R. Y. & Lester, H. A. (1987). Activation of a potassium current by rapid photochemically generated step increases of intracellular calcium in rat sympathetic neurones. Proc. natn. Acad. Sci. USA 84, 34963500.Google Scholar
G¨th, K. & Potter, J. D. (1987). Effect of rigor and cycling cross-bridges on the structure of troponin C and on the Ca2+ affinity of the Ca2+-specific regulatory sites in skinned rabbit psoas fibers. . biol. Chem. 262, 1362713635.Google Scholar
Harada, Y., Noguchi, A., Kishino, A. & Yanagida, T. (1987). Sliding movement of single actin filaments on one-headed myosin filaments. Nature 326, 805806.Google Scholar
Hartshorne, D. J., Theiner, M. & Mueller, H. (1968). Studies on troponin. Biochim. biophys. Acta 175, 320330.Google Scholar
Haselgrove, J. C. (1973). X-ray evidence for a conformation change in the actin containing filaments of vertebrate striated muscle. Cold Spring Harbor Symp. Quant. Biol. 37, 341352.CrossRefGoogle Scholar
Heilbrunn, L. V. & Wiercinski, F. J. (1947). The action of various cations on muscle protoplasm. J. Cell Comp. Physiol. 29, 1532.Google ScholarPubMed
Hellam, D. C. & Podolsky, R. J. (1969). Force measurements in skinned muscle fibres. J. Physiol. 200, 807819.CrossRefGoogle ScholarPubMed
Herzberg, O. & James, M. N. G. (1985). Structure of the calcium regulatory muscle protein Troponin C at 2.8 Å resolution. Nature, 313, 653659.Google Scholar
Hess, P., Metzger, P. & Weingart, R. (1982). Free magnesium in sheep, ferret and frog striated muscle at rest measured with ion-selective micro-electrodes. J. Physiol. 329, 173188.Google Scholar
Hess, P., Lansman, J. B. & Tsien, R. W. (1984). Different modes of Ca2+ channel gating behaviour followed by DHP Ca2+ agonists and antagonists. Nature 311, 538544.Google Scholar
Hidalgo, C. & Jaimovich, E. (1989). Inositol trisphosphate and E-C coupling in skeletal muscle. J. Bioenerg. Biomembr. 21, 267281.CrossRefGoogle Scholar
Hilkert, R. J., Zaidi, N. F., Lagenaur, C. F. & Salama, G. (1990). Immunoaffinity purified 106 kDa protein from sarcoplasmic reticulum (SR) is a Ca2+ release channel modulated by agents that alter Ca2+ release. Biophys. J. 57, 275a.Google Scholar
Hill, A. V. (1910). The possible effects of the aggregation of the molecules of haemoglobin on its dissociation curve. J. Physiol. 40, 47.Google Scholar
Hill, A. V. (1948). On the time required for diffusion and its relation to processes in muscle. Proc. Roy. Soc. B135, 446453.Google Scholar
Hill, T. L. (1983). Two elementary models for the regulation of skeletal muscle contraction by calcium. Biophys. J. 44, 383396.Google Scholar
Hill, T. L. (1985). Co-operativity Theory in Biochemistry: Steady State and Equilibrium Systems. Berlin and Heidelberg: Springer-Verlag.Google Scholar
Holmes, K. C., Popp, D., Gebhard, W. & Kabsch, W. (1990). Atomic model of the actin filament. Nature 347, 4449.CrossRefGoogle ScholarPubMed
Holroyde, M. J., Robertson, S. P., Johnson, J. D., Solaro, R. J. & Potter, J. D. (1980). The calcium and magnesium binding sites on cardiac troponin and their role in regulation of myofibrillar adenosine triphosphatase. J. biol. Chem. 255, 11681193.CrossRefGoogle ScholarPubMed
Housmans, P. R., Lee, N. K. M. & Blinks, J. R. (1983). Active shortening retards the decline of the intracellular calcium transient in mammalian heart muscle. Science 221, 159161.CrossRefGoogle ScholarPubMed
Huxley, A. F. (1957). Muscle structure and theories of contraction. Prog. Biophys. 7, 255318.Google ScholarPubMed
Huxley, A. F. & Simmons, R. M. (1971). Proposed mechanism for force generation in striated muscle. Nature 233, 533538.Google Scholar
Huxley, A. F. & Taylor, R. E. (1958). Local activation of striated muscle fibres. J. Physiol. 144, 426441.Google Scholar
Huxley, H. E. (1969). The mechanism of muscular contraction. Science, 164, 13561366.Google Scholar
Huxley, H. E. (1973). Structural changes in the actin and myosin containing filaments during contraction. Cold Spring Harbor Symp. Quant. Biol. 37, 361376.Google Scholar
Huxley, H. E. & Kress, M. (1985). Cross-bridge behaviour during muscle contraction. J. Muscle Res. Cell Motil. 6, 153161.Google Scholar
Hymel, L., Inui, M., Fleischer, S. & Schindler, H. (1988). Purified ryanodine receptor of skeletal muscle SR forms Ca2+-activated oligomeric Ca2+ channels in planar bilayers. Proc. natn. Acad. Sci. USA 85, 441445.Google Scholar
Ikemoto, N., Antoniu, B. & Kim, D. H. (1984). Rapid Ca2+ release from isolated sarcoplasmic reticulum is triggered via the attached transverse tubular system. J. biol. Chem. 259, 1315113158.Google Scholar
Ikemoto, N., Ronjat, M. & Meszaros, L. G. (1989). Kinetic analysis of excitationcontraction coupling. J. Bioenerg. Biomemb. 21, 247266.Google Scholar
Ildefonse, M., Jacquemond, V., Rougier, O., Renand, J. F., Fosset, M. & Lazdunski, M. (1985). Excitation-contraction coupling in skeletal muscle: evidence for a role of slow Ca2+ channels using Ca2+ channel activators and inhibitors in the dihydropyridine series. Biochem. biophys. Res. Comm. 129, 904909.Google Scholar
Irving, M. (1988). Molecular motor mechanics. Nature 334, 1112.CrossRefGoogle ScholarPubMed
Irving, M., Maylie, J., Sizto, N. L. & Chandler, W. K. (1989). Simultaneous monitoring of changes in Mg2+ and Ca2+ concentrations in frog cut twitch fibers containing antipyrylazo III. J gen. Physiol. 93, 585608.Google Scholar
Irving, M., Maylie, J., Sizto, N. L. & Chandler, W. K. (1990). Intracellular diffusion in the presence of mobile buffers. Biophys. J. 57, 717721.Google Scholar
Jackson, A. P., Timmerman, M. P., Bagshaw, C. R. & Ashley, C. C. (1987). The kinetics of calcium binding to fura-2 and indo-1. FEBS Lett. 216, 3539.Google Scholar
Jenden, D. J. & Fairhurst, A. S. (1969). The pharmacology of ryanodine. Pharm. Rev. 21, 125.Google Scholar
Johnson, J. D., Collins, J. H. & Potter, J. D. (1978). Dansylaziridine-labelled troponin C. A fluorescent probe of Ca2+-binding to the Ca2+-specific regulatory sites. J. biol. Chem. 253, 64516458.CrossRefGoogle Scholar
Johnson, J. D., Charlton, S. C. & Potter, J. D. (1979). A fluorescence stopped-flow analysis of Ca2+ exchange with troponin-C. J. biol. Chem. 254, 34973502.CrossRefGoogle ScholarPubMed
Johnson, J. D., Collins, J. H., Robertson, S. P. & Potter, J. D. (1980). A fluorescent probe study of Ca2+-exchange with the Ca2+-specific sites of cardiac troponin and troponin C. J. biol. Chem. 255, 96359640.CrossRefGoogle Scholar
Johnson, J. D., Robinson, D. E., Robertson, S. P., Schwartz, A. & Potter, J. D. (1981). Ca2+ exchange with troponin and the regulation of muscle contraction. In The Regulation of Muscle Contraction: Excitation-Contraction Coupling (ed. Grinnell, A. D.), pp. 241259. New York: Academic Press.Google Scholar
Julian, F. (1971). The effect of calcium on the force-velocity relation of briefly glycerinated frog muscle fibres, J. Physiol. 218, 117145.Google Scholar
Kalos, M. H. & Whitlock, P. A. (1986). Monte Carlo Methods. New York: Wiley.CrossRefGoogle Scholar
Kao, J. P. Y. & Tsien, R. Y. (1988). Ca2+ binding kinetics of fura-2 and azo-i from temperature jump relaxation measurements. Biophys. J. 53, 635639.CrossRefGoogle Scholar
Kaplan, J. & Ellis-Davies, G. (1988). Properties and applications of DM-nitrophen, a new caged-Ca2+. Biophys. J. 53, 36a.Google Scholar
Kaplan, J. H. & Somlyo, A. P. (1989). Flash photolysis of caged compounds: new tools for cellular physiology. TINS 12, 5459.Google ScholarPubMed
Kaplan, J. H., Forbush, B. & Hoffman, J. F. (1978). Rapid photolytic release of adenosine 5-triphosphate from a protected analogue: Utilization by the Na:K pump of human red blood cell ghosts. Biochemistry 17, 19291935.CrossRefGoogle ScholarPubMed
Kentish, J. C., Barsotti, R. J., Lea, T. J., Mulligan, I. P., Patel, J. R. & Ferenczi, M. A. (1990). Calcium release from cardiac sarcoplasmic reticulum induced by photorelease of calcium or Ins(1, 4, 5)P3. Am. J. Physiol. 258, H610–615.Google Scholar
Kishino, A. & Yanagida, T. (1988). Force measurements by micromanipulation of a single actin filament by glass needles. Nature 334, 7476.Google Scholar
Kobayashi, S., Somlyo, A. V. & Somlyo, A. P. (1988). Heparin inhibits the inositol 1, 4, 5-trisphosphate-dependent, but not the independent, calcium release induced by guanine nucleotide in vascular smooth muscle. Biochem. biophys. Res. Comm. 153, 625631.Google Scholar
Kress, M., Huxley, H. E., Faruqi, A. R. & Hendrix, J. (1986). Structural changes during activation of frog muscle studied by time-resolved X-ray diffraction. J. molec. Biol. 188, 325342.Google Scholar
Kretsinger, R. H. & Nockolds, C. E. (1973). Carp muscle calcium-binding. II. Structure determination and general description. J. biol. Chem. 248, 33133326.Google ScholarPubMed
Kwok, W. M. & Best, P. M. (1990). Ryanodine sensitivity and multiple conductance states of the Ca2+ release channel from native SR membrane. Biophys. J. 57, 168a.Google Scholar
Lai, F. A. & Meissner, G. (1989). The muscle ryanodine receptor and its intrinsic Ca2+ channel activity, J. Bioenerg. Biomemb. 21, 227246.Google Scholar
Lai, F. A., Erickson, H. P., Rousseau, E., Liu, Q. Y. & Meissner, G. (1988). Purification and reconstitution of the calcium release channel from skeletal muscle. Nature 331, 315319.Google Scholar
Lamb, G. D. (1986). Components of charge movement in rabbit skeletal muscle: the effect of tetracaine and nifedipine. J. Physiol. 376, 85100.Google Scholar
Lea, T. J. & Ashley, C. C. (1981). Carbon dioxide or bicarbonate ions release Ca2+ from internal stores in crustacean myofibrillar bundles. J. Membr. Biol. 61, 115125.Google Scholar
Lea, T. J. & Ashley, C. C. (1989). Ca2+-induced Ca2+ release from the sarcoplasmic reticulum of isolated myofibrillar bundles of barnacle muscle fibres. Pfl¨g. Archiv. 413 401406.Google Scholar
Lea, T. J. & Ashley, C. C. (1990). Ca2+ release from the sarcoplasmic reticulum of barnacle myofibrillar bundles initiated by photolysis of caged Ca2+. J, Physiol. 427, 435453.Google Scholar
Lea, T. J., Fenton, M. J., Potter, J. D. & Ashley, C. C. (1990). Rapid activation by photolysis of nitr-5 in skinned fibres of the striated adductor muscle from scallop. Biochim. biophys. Acta 1034, 186194.CrossRefGoogle ScholarPubMed
Lea, T. J., Griffiths, P. J., Tregear, R. T. & Ashley, C. C. (1986). An examination of the ability of InsP3 to induce Ca2+ release and tension development in skinned muscle fibres of frog and Crustacea. FEBS Lett. 207, 153161.Google Scholar
Leavis, P. C. & Kraft, E. L. (1978). Calcium binding to cardiac TnC. Archiv. Biochem. Biophys. 186, 411415.CrossRefGoogle Scholar
Lester, H. A. & Nerbonne, J. M. (1982). Physiological and pharmacological manipulations with light flashes. Ann. Rev. Biophys. Bioeng. 11, 151175.Google Scholar
Leung, A. R., Imagawa, T., Block, B., Franzini-Armstrong, C. & Campbell, K. P. (1988). Biochemical and ultrastructural characterisation of the dihydropyridine receptor from rabbit skeletal muscle. J. biol. Chem. 263, 9941001.Google Scholar
Liu, Q. Y., Lai, F. A., Xu, L. L., Jones, R. V., Ladine, J. K. & Meissner, G. (1989). Comparison of the mammalian and amphibian skeletal muscle ryanodine receptor- Ca2+ release channel complexes. Biophys. J. 55, 85a.Google Scholar
McCray, J. A. & Trentham, D. R. (1989). Properties and uses of photoreactive caged compounds. Ann. Rev. Biophys. biophys. Chem. 18, 239270.Google Scholar
Ma, J., Fill, M., Knudson, M. C., Campbell, K. P. & Coronado, R., (1988). Ryanodine receptor of skeletal muscle is a gap junction-type channel. Science N. Y. 242, 99102.CrossRefGoogle ScholarPubMed
Maclennan, D. H., Brandl, C. J., Koryack, B. & Green, N. M. (1985). Amino-acid sequence of a Ca2+ + Mg2+-dependent ATPase from rabbit muscle SR, deduced from its complementary DNA sequence. Nature 316, 696700.CrossRefGoogle Scholar
Maclennan, D. H. & Wong, P. T. S. (1971). Isolation of a calcium-sequestering protein from sarcoplasmic reticulum. Proc. natn. Acad. Sci. USA 68, 12311235.Google Scholar
Marban, E., Rink, T., Tsien, R. W. & Tsien, R. Y. (1980). Free calcium in heart muscle at rest and during contraction measured with Ca2+-sensitive microelectrodes. Nature 286, 845850.Google Scholar
Martonosi, A. N. (1984). Mechanisms of Ca2+ release from sarcoplasmic reticulum of skeletal muscle. Physiol. Rev. 64, 12401320.Google Scholar
Maruyama, K. & Maclennan, D. H. (1988). Mutation of aspartic acid-351, lysine-352 and lysine-515 alters the Ca2+-transport activity of the Ca2+-ATPase expressed in COS-i cells. Proc. natn. Acad. Sci. USA 85, 33143318.Google Scholar
Mathias, R. T., Levis, R. A. & Eisenberg, R. S. (1980). Electrical models of excitation contraction coupling and charge movement in skeletal muscle, J. gen. Physiol. 76, 131.Google Scholar
Melzer, W., Rios, E. & Schneider, M. F. (1986). The removal of myoplasmic free calcium following calcium release in frog skeletal muscle. J. Physiol. 372, 261292.Google Scholar
Metropolis, N. & Ulam, S. (1949). The Monte Carlo Method. J. Amer. Stat. Assoc. 44. 335–380.Google Scholar
Metropolis, N., Rosenbluth, A., Rosenbluth, M., Teller, A. & Teller, E. (1953). Equation of state calculations by fast computing machines, J. phys. Chem. 21, 10871092.Google Scholar
Mikami, A., Imoto, K., Tanabe, T., Niidome, T., Mori, Y., Takeshima, H., Narumiya, S. & Numa, S. (1989). Primary structure and functional expression of the cardiac dihydropyridine-sensitive calcium channel. Nature 340, 230233.Google Scholar
Miqnery, G. A., Sudhof, T. C., Takei, K. & De Camilli, P. (1989). Putative receptor for Ins(1, 4, 5) P3 similar to ryanodine receptor. Nature 342, 192195.Google Scholar
Moisescu, D. G. (1976). Kinetics of reaction in calcium-activated skinned muscle fibres. Nature 262, 610613.Google Scholar
Monod, J., Wyman, J. & Changeux, J.-P. (1965). On the nature of allosteric transitions: a plausible model. J. molec. Biol. 12, 88118.Google Scholar
Moore, P. B., Huxley, H. E. & Derosier, D. J. (1970). Three-dimensional reconstruction of F-Actin, thin filaments and decorated thin filaments. J. molec. Biol., 50, 279295.Google Scholar
Moss, R. L., Giulian, G. G. & Greaser, M. L. (1985). The effects of partial extraction of TnC upon the tension–pCa relationship in rabbit skinned skeletal muscle fibers. J. gen. Physiol. 86, 585600.CrossRefGoogle ScholarPubMed
Mounier, Y. & Goblet, C. (1987). Role of the different Ca2+ sources in the excitationcontraction coupling in crab muscle fibers. Can. J. Physiol. Pharmacol. 65, 667671.Google Scholar
Mulligan, I. P. (1989). Mechanical studies on skinned muscle fibres using caged ATP and caged calcium. Ph.D. Thesis, University of Oxford, UK.Google Scholar
Mulligan, I. P. & Ashley, C. C. (1989). Rapid relaxation of single frog skeletal muscle fibres following laser flash photolysis of the caged calcium chelator, diazo-2. FEBS Lett. 255, 196200.CrossRefGoogle ScholarPubMed
Mulligan, I. P., Adams, S. R., Tsien, R. Y., Potter, J. D. & Ashley, C. C. (1990). Flash photolysis of the caged calcium-chelator, diazo-2 produces rapid relaxation of single skeletal muscle fibres. Biophys. J. 57, 541a.Google Scholar
Nakamura, Y. & Schwartz, A. (1972). The influence of hydrogen ion concentration on calcium binding and release by skeletal muscle SR. J. gen. Physiol. 59, 2232.Google Scholar
Natori, R. (1954). The property and contraction process of isolated myofibrils. Jikeikai Med. J. 1, 119126.Google Scholar
Oetliker, H. (1982). An appraisal of the evidence for a sarcoplasmic reticulum membrane potential and its relation to calcium release in skeletal muscle. J. Muscle Res. Cell Motil. 3, 247272.Google Scholar
Ogawa, Y. (1985). Calcium binding to troponin C and troponin: effects of Mg2+, ionic strength and pH. J. Biochem. Tokyo 97, 10111023.Google Scholar
Ohkusa, T., Kang, J. J., Heemstra, V., & Ikemoto, N. (1990). Induction of Ca2+ release from sarcoplasmic reticulum (SR) is mediated by conformational changes of the foot protein (FP). Biophys. J. 57, 276a.Google Scholar
Palmer, R., Mulligan, I. P., Nunn, C. & Ashley, C. C. (1990). Striated scallop muscle relaxation: fast force transients produced by photolysis of diazo-2. Biochem. biophys. Res. Commun. 168, 295300.Google Scholar
Pan, B. S. & Solaro, R. J. (1987). Calcium-binding properties of troponin C in detergent-skinned heart muscle fibers. J. biol. Chem. 262, 7839–7849.Google Scholar
Pape, P. C., Konishi, M., Baylor, S. M. & Somlyo, A. P. (1988). Excitation-contraction coupling in skeletal muscle fibers injected with the InsP3 blocker, heparin. FEBS Lett. 235, 5762.Google Scholar
Pape, P. C., Konishi, M., Hollingworth, S. & Baylor, S. M. (1989). Myoplasmic pH and calcium transients from intact frog skeletal muscle fibers simultaneously injected with phenol red and fura-2. Biophys. J. 55, 412a.Google Scholar
Patchornik, A., Amit, B. & Woodward, R. B. (1970). Photosensitive protecting groups. J. Am. Chem. Soc. 92, 63336335.CrossRefGoogle Scholar
Peachey, L. D. (1965). The sarcoplasmic reticulum and transverse tubules of the frog's sartorius. J. Cell Biol. 25, 209231.Google Scholar
Peachey, L. D. (1973). Electrical events in the T-system of frog skeletal muscle. Cold Spring Harbor Symp. Quant. Biol. 37, 479487.CrossRefGoogle Scholar
Perez-Reyes, E., Kim, H. S., Lacerda, A. E., Horner, W., Wei, X., Rampe, D., Campbell, K. P., Brown, A. M. & Birnbaumer, L. (1989). Induction of calcium currents by the expression of the alpha1-subunit of the dihydropyridine receptor from skeletal muscle. Nature 340, 233236.Google Scholar
Perry, S. V. (1975). Troponin and activation in muscle. In Contraction and Relaxation in the Myocardium, (ed. Nayler, W.), pp. 2943. New York: Academic Press.Google Scholar
Pessah, I. N., Francini, A. O., Scales, D. J., Waterhouse, A. L. & Casida, J. E. (1986). Calcium ryanodine receptor complex. J. biol. Chem. 261, 86438648.CrossRefGoogle ScholarPubMed
Pillai, V. N. (1980). Photoremovable protecting groups in organic synthesis. Synthesis 126.Google Scholar
Potter, J. D. (1974). The content of troponin, tropomyosin, actin and myosin in rabbit skeletal muscle myofibrils. Archiv. Biochem. Biophys. 162, 436441.Google Scholar
Potter, J. D. & Gergely, J. (1975). The calcium and magnesium binding sites of troponin and their role in the regulation of myofibrillar adenosinetriphosphatase. J. biol. Chem. 250, 46284633.Google Scholar
Potter, J. D. & Johnson, J. D. (1982). Troponin. In Calcium and Cell Function, vol. 11 (ed. Cheung, W.), pp. 145173. New York: Academic Press.Google Scholar
Potter, J. D., Robertson, S. P. & Johnson, J. D. (1981). Magnesium and the regulation of muscle contraction. Fed. Proc. 40, 26532656.Google Scholar
Putkey, J. A., Carrol, S. L. & Means, A. R. (1987). The non-transcribed chicken calmodulin pseudogene cross-hybridises with mRNA from the slow muscle troponin C gene. Mol. Cell Biol. 7, 15491553.Google Scholar
Putkey, J. A., Sweeney, H. L. & Campbell, S. T. (1989). Site-directed mutation of the trigger calcium-binding sites in cardiac troponin C. J. biol. Chem. 264, 1237012378.Google ScholarPubMed
Ridgway, E. B. & Ashley, C. C. (1967). Calcium transients in single muscle fibers. Biochem. biophys. Res. Commun. 29, 229233.Google Scholar
Ridgway, E. B. & Gordon, A. (1984). Muscle Ca2+ transient: effect of poststimulus length changes in single fibres, J. gen. Physiol. 83, 75103.Google Scholar
Rios, E. & Brum, G. (1987). Involvement of dihydropyridine receptors in excitationcontraction coupling in skeletal muscle. Nature 325, 717720.Google Scholar
Robertson, S. P., Johnson, J. D. & Potter, J. D. (1981). The time course of Ca2+ exchange with calmodulin, troponin, parvalbumin and myosins to transient increases in Ca2+. Biophys. J. 34, 559569.Google Scholar
Rojas, C. & Hidalgo, C. (1990). Inositol trisphosphate binds to heavy sarcoplasmic reticulum membranes isolated from frog skeletal muscle. Biophys. J. 57, 342a.Google Scholar
Rojas, C. & Jaimovich, E. (1990). Calcium release modulated by inositol trisphosphate in ruptured fibres from frog skeletal muscle. Pflüg. Archiv 416, 110.Google Scholar
Rojas, E., Nasser-Gentina, V., Luxoro, M., Pollard, M. E. & Carrasco, M. A. (1987). InsP3-induced Ca2+ release from the SR and contraction in crustacean muscle. Can. J. Physiol. Pharmacol. 65, 672679.Google Scholar
Rosenfeld, S. S. & Taylor, E. W. (1985a). Kinetic studies of calcium and magnesium binding to troponin C. J. biol. Chem. 260, 242251.Google Scholar
Rosenfeld, S. S. & Taylor, E. W. (1985b). Kinetic studies of calcium binding to regulatory complexes from skeletal muscle, J. biol. Chem. 260, 252261.Google Scholar
Rüegg, J. C. (1986). Calcium in Muscle Activation. Berlin: Springer-Verlag.Google Scholar
Saito, A., Inui, M., Radermacher, M., Frank, J. & Fleischer, S. (1988). Ultrastructure of the Ca2+ release channel of the sarcoplasmic reticulum. J. Cell Biol. 107, 211219.CrossRefGoogle Scholar
Saito, A., Chadwick, C. C. & Fleischer, S. (1990). Ultrastructure of the inositol trisphosphate receptor (IP3REC) from smooth muscle. Biophys. J. 57, 285a.Google Scholar
Sandow, A. (1952). Excitation contraction coupling in muscular response. Yale J. Biol. Med. 25, 176201.Google ScholarPubMed
Satyshur, K. A., Rao, S. T., Pyzalska, D., Drendel, W., Greaser, M. & Sunaralingam, M. (1988). Refined structure of skeletal muscle troponin C in the 2 calcium state at 2 Å resolution. J. biol. Chem. 263, 16281647.Google Scholar
Schaub, M. C. & Perry, S. V. (1969). The relaxing protein system of striated muscle. Biochem. J. 115, 9931004.Google Scholar
Schneider, M. F. & Chandler, W. K. (1973). Voltage-dependent charge movement in skeletal muscle: a possible step in excitation-contraction coupling. Nature 242, 244246.Google Scholar
Schoenberg, M. (1980a). Geometrical factors influencing muscle force development I. The effect of filament spacing upon axial forces. Biophys. J. 30, 5167.Google Scholar
Schoenberg, M. (1980b). Geometrical factors influencing muscle force development II. Radial forces. Biophys. J. 30, 6977.Google Scholar
Schwartz, L. M., McCleskey, E. W. & Almers, W. (1985). Dihydropyridine receptors in muscle are voltage-dependent but most are not functional calcium channels. Nature 314, 747750.Google Scholar
Shimomura, O., Johnson, F. H. & Saigi, Y. (1962). Extraction, purification and properties of aequorin: a bioluminescent protein from the luminous hydromedusan, Aequorea. J. cell. Comp. Physiol. 59, 223239.Google Scholar
Shiner, J. S. & Solaro, R. J. (1982). Activation of thin-filament-regulated muscle by calcium ion: considerations based on nearest-neighbor lattice statistics. Proc. natn. Acad. Sci. USA 79, 46374641.Google Scholar
Shiner, J. S. & Solaro, R. J. (1984). The Hill coefficient for the Ca2+-activation of striated muscle contraction. Biophys. J. 46, 541543.Google Scholar
Shoshan, V., Maclennan, D. H. & Wood, D. S. (1981). A proton gradient controls a calcium-release channel in SR. Proc. natn. Acad. Sci. USA 78, 48284832.Google Scholar
Simon, B. J., Klein, M. G. & Schneider, M. F. (1989). Caffeine slows turn-off of calcium release in voltage clamped muscle fibers. Biophys. J. 55, 793797.Google Scholar
Smith, J. S., Coronado, R. & Meissner, G. (1985). SR contains adenine nucleotideactivated calcium channels. Nature 316, 446449.Google Scholar
Smith, J. S.Coronado, R. & Meissner, G. (1986). Single channel measurements of the Ca2+ release channel from skeletal muscle SR. J. gen. Physiol. 88, 573588.Google Scholar
Smith, J. S., Imagawa, T., Ma, J., Fill, M., Campbell, K. P. & Coronado, R. (1988). Purified ryanodine receptor from rabbit skeletal muscle is the calcium-release channel of sarcoplasmic reticulum. J. gen. Physiol. 92, 126.Google Scholar
Somlyo, A. P., Walker, J. W., Goldman, Y. E., Trentham, D. R., Kobayashi, S., Kitazawa, T. & Somlyo, A. V. (1988). Inositol trisphosphate, calcium and muscle contraction. Phil. Trans. R. Soc. Lond. B320, 399414.Google Scholar
Somlyo, A. V., Gonzalez-Serratos, H., Shuman, H., McClellan, G. & Somlyo, A. P. (1981). Calcium release and ionic changes in the sarcoplasmic reticulum of tetanized muscle: an electron-probe study. J. Cell Biol. 90, 577594.Google Scholar
Spiecker, W. &Luttgau, H. C. (1979). Extracellular calcium and excitation–contraction coupling. Nature 280, 158160.Google Scholar
Squire, J. (1981). Muscle regulation: a decade of the steric blocking model. Nature 291, 614615.Google Scholar
Stein, P. & Palade, P. (1988). Sarcoballs: direct access to sarcoplasmic reticulum Ca2+ channels in skinned frog muscle fibers. Biophys. J. 54, 357363.Google Scholar
Stein, R. B., Bobet, J., Otuztöreli, M. N. & Fryer, M. (1988). The kinetics relating calcium and force in skeletal muscle. Biophys. J. 54, 705717.Google Scholar
Suarez-Isla, B. A., Irribarra, V., Oberhauser, A., Larralde, L., Bull, R., Hidalgo, C. & Jaimovich, E. (1988). Inositol (1, 4, 5)-trisphosphate activates a calcium channel in isolated SR membranes. Biophys. J. 54, 737741.Google Scholar
Sundaralingam, M., Bergstrom, R., Strasburg, G., Rao, S. I., Roychowdhury, P., Greaser, M. L. & Wang, B. C. (1985). Molecular structure of troponin C from chicken skeletal muscle at 3 Å resolution. Science, N. Y. 227, 945948.Google Scholar
Szent-Györgyi, A. (1953). Chemical Physiology of Contraction in Body and Heart Muscle. New York: Academic Press.Google Scholar
Takeshima, H., Nishimura, S., Matsumoto, T., Ishida, H., Kangawa, K., Minamino, N., Matsuo, H., Ueda, M., Hanaoka, M., Hirose, T. & Numa, S. (1989). Primary structure and expression from complementary DNA of skeletal muscle ryanodine receptor. Nature 339, 439445.CrossRefGoogle ScholarPubMed
Tanabe, T., Beam, K. G., Adams, B. A., Niidome, T. & Numa, S. (1990b). Regions of the skeletal muscle dihydropyridine receptor critical for excitation–contraction coupling. Nature 346, 567569.Google Scholar
Tanabe, T., Takeshima, H., Mikami, A., Flockerzi, V., Takahashi, H., Kangawa, K., Kojima, M., Matsuo, H., Hirose, T. & Numa, S. (1987). Primary structure of the receptor for calcium channel blockers from skeletal muscle. Nature 328, 313318.Google Scholar
Tanabe, T., Mikami, A., Numa, S. & Beam, K. G. (1990a). Cardiac-type excitation–contraction in dysgenic skeletal muscle injected with cardiac dihydropyridine receptor DNA. Nature 344, 451453.Google Scholar
Taylor, K. A. & Amos, L. A. (1981). A new model for the geometry of the binding of myosin crossbridges to muscle thin filaments. J. molec. Biol. 147, 297324.Google Scholar
Timmerman, M. P. & Ashley, C. C. (1986). Fura-2 diffusion and its use as an indicator of transient free calcium changes in single striated muscle cells. FEBS Lett. 209, 18.Google Scholar
Timmerman, M. P., Godber, J. F., Walton, A. & Ashley, C. C. (1990). Imaging spatial distribution of release in single muscle fibres from B. nubilus using image intensification. Cell Calcium 11, 211220.Google Scholar
Toyoshima, Y. Y., Kron, S. J., McNally, E. M., Niebling, K. R., Toyoshima, C. & Spudich, J. A. (1987). Myosin subfragment-1 is sufficient to move actin filaments in vivo. Nature 328, 536539.Google Scholar
Tsien, R. Y. (1980). New calcium indicators and buffers with high selectivity against magnesium and protons: design synthesis and properties of prototype structures. Biochemistry 19, 23962404.Google Scholar
Tsien, R. Y. & Rink, T. J. (1980). Neutral carrier ion selective electrodes for measurement of intracellular free calcium. Biochem. biophys. Acta 599, 623638.Google Scholar
Tsien, R. Y. & Zucker, R. S. (1986). Control of cytoplasmic calcium with photolabile tetracarboxylate 2-nitrobenzhydrol chelators. Biophys. J. 50, 843853.Google Scholar
Tsien, R. Y., Rink, T. J. & Poenie, M. (1985). Measurement of cytosolic free Ca2+ in individual small cells using fluorescent microscopy with dual excitation wavelengths. Cell Calcium 6, 145157.Google Scholar
Valdeolmillos, M., O'Neill, S. C., Smith, G. L. & Eisner, D. A. (1989). Calciuminduced Ca2+ release activates contraction in intact cardiac cells. Pflüg. Archiv 413, 676678.Google Scholar
Van Eerd, J. P. & Takahashi, K. (1976). Determination of the complete amino acid sequence of bovine cardiac troponin C. Biochemistry 15, 11711180.Google Scholar
Vergara, J., Tsien, R. Y. & Delay, M. (1985). Inositol 1, 4, 5-trisphosphate: a possible chemical link in EC coupling in muscle. Proc. natn. Acad. Sci. USA 82, 63526356.Google Scholar
Vilven, J. & Coronado, R. (1988). Opening of dihydropyridine Ca2+ channels in skeletal muscle membranes by InsP3. Nature 336, 587589.Google Scholar
Volpe, P., Salviatti, G., Di Virgilio, F. & Pozzan, T. (1985). Inositol 1, 4, 5- trisphosphate induces Ca2+ release from SR of skeletal muscle. Nature, 316, 347349.Google Scholar
Wagenknecht, T., Grassucci, R., Frank, J., Saito, A., Inui, M. & Fleischer, S (1989). Three-dimensional architecture of the calcium channel/foot structure of the sarcoplasmic reticulum. Nature 338, 167170.Google Scholar
Wakabayashi, T., Huxley, H. E., Amos, L. A. & Klug, A. (1975). Three-dimensional image reconstruction of actin–;tropomyosin complex and actin–tropomyosin–troponin T–troponin I complex. J. molec. Biol. 93, 477497.Google Scholar
Walker, J. W., Somlyo, A. V., Goldman, Y. E., Somlyo, A. P. & Trentham, D. R. (1987). Kinetics of smooth and skeletal muscle activation by laser pulse photolysis of caged inositol 1, 4, 5-trisphosphate. Nature 327, 249252.Google Scholar
Weber, A. & Herz, R. (1963). The binding of calcium to actomyosin systems in relation to their biological activity. J. biol. Chem. 238, 599605.Google Scholar
Weber, A., Herz, R. & Reiss, I. (1964). The regulation of myofibrillar activity by calcium. Proc. Roy. Soc. B160, 489501.Google Scholar
Whitehead, E. A. (1979). The structure of steady state enzyme kinetic equations: a graph-theoretical algorithm for obtaining conditions for reduction in degree by common-factor cancellations. J. theor. Biol. 80, 355381.Google Scholar
Winegrad, S. (1965). The location of muscle calcium with respect to the myofibrils. J. gen. Physiol., 48, 9971002.Google Scholar
Wnuk, W., Schoechlin, M. & Stein, E. A. (1984). Regulation of actomyosin ATPase by a single calcium-binding site on troponin C from crayfish. J. biol. Chem. 259, 90179023.Google Scholar
Zorzato, F., Fujii, J., Otsu, K., Phillips, M., Green, N. M., Lai, F. A., Meissner, G., & Maclennan, D. H. (1990). Molecular cloning of cDNA encoding human and rabbit forms of the Ca2+ release channel (ryanodine receptor) of skeletal muscle sarcoplasmic reticulum. J. biol. Chem. 265, 22442256.Google Scholar
Zot, A. S. & Potter, J. D. (1987a). Structural aspects of troponin-tropomyosin regulation of skeletal muscle contraction. Ann. Rev. Biophys. biophys. Chem. 16, 535539.Google Scholar
Zot, A. S. & Potter, J. D. (1987b). The effect of [Mg2+’ on the Ca2+-dependence of ATPase and tension development of fast skeletal muscle: the role of the Ca2+ specific sites of troponin C. J. biol. Chem. 262, 19661969.Google Scholar
Zot, H. G. & Potter, J. D. (1987c). Calcium binding and fluorescence measurements of dansylaziridine-labelled troponin C in reconstituted thin filaments. J. Muscle Res. Cell Motil. 8, 428436.Google Scholar
Zot, H. G., Güth, K. & Potter, J. D. (1986). Fast skeletal muscle skinned fibers and myofibrils reconstituted with N-terminal fluorescent analogues of troponin C. J. biol. Chem. 261, 1588315890.Google Scholar
Zot, H. G., Iida, S. & Potter, J. D. (1983). Thin filament interactions and Ca2+ binding to Tn. Chem. Scr. 21, 133136.Google Scholar