Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-9pm4c Total loading time: 0 Render date: 2024-04-25T09:03:16.621Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  05 June 2012

Richard D. Keynes
Affiliation:
University of Cambridge
David J. Aidley
Affiliation:
University of East Anglia
Christopher L.-H. Huang
Affiliation:
University of Cambridge
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Nerve and Muscle , pp. 170 - 177
Publisher: Cambridge University Press
Print publication year: 2011

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

Adrian, E. D. and Lucas, K. (1912). On the summation of propagated disturbances in nerve and muscle. J. Physiol., Lond. 44, 68–124.CrossRefGoogle ScholarPubMed
Adrian, R.H. and Almers, W. (1976a). The voltage dependence of membrane capacity. J. Physiol., Lond. 254, 317–338.CrossRefGoogle ScholarPubMed
Adrian, R.H. and Almers, W. (1976b). Charge movement in the membrane of striated muscle. J. Physiol., Lond. 254, 339–360.CrossRefGoogle ScholarPubMed
Adrian, R. H. and Bryant, S. H. (1974). On the repetitive discharge in myotonic muscle fibres. J. Physiol. 240, 505–515.CrossRefGoogle ScholarPubMed
Adrian, R. H. and Peachey, L.D. (1973). Reconstruction of the action potential of frog sartorius muscle. J. Physiol. 235, 103–131.CrossRefGoogle ScholarPubMed
Ahern, C. A. and Horn, R. (2005). Focused electric field across the voltage sensor of potassium channels. Neuron 48, 25–29.CrossRefGoogle ScholarPubMed
Aidley, D. J. (1998). The Physiology of Excitable Cells, 4th edn. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Allen, D. G., Lamb, G. D. and Westerblad, H. (2008). Skeletal muscle fatigue: cellular mechanisms. Physiol. Rev. 88, 287–332.CrossRefGoogle ScholarPubMed
Armstrong, C. M. and Bezanilla, F. M. (1973). Current related to the movement of the gating particle of the sodium channels. Nature 242, 459–461.CrossRefGoogle Scholar
Ashley, C. C. and Ridgeway, E. B. (1968). Simultaneous recording of membrane potential, calcium transient and tension in single muscle fibres. Nature 219, 1168–1169.CrossRefGoogle Scholar
Bagshaw, C. R. (1993). Muscle Contraction, 2nd edn. London: Chapman & Hall.CrossRefGoogle Scholar
Baker, P. F., Hodgkin, A. L. and Shaw, T. I. (1962). The effect of changes in internal ionic concentrations on the electrical properties of perfused giant axons. J. Physiol. 164, 355–374.CrossRefGoogle Scholar
Barnard, E. A., Miledi, R. and Sumikawa, K. (1982). Translation of exogenous messenger RNA coding for nicotinic acetylcholine receptors produces functional receptors in Xenopus oocytes. Proc. R. Soc. Lond. B215, 241–246.CrossRefGoogle Scholar
Block, B. A., Imagawa, T., Campbell, K. P. and Franzini-Armstrong, C. (1988). Structural evidence for direct interaction between the molecular components of the transverse tubule/sarcoplasmic reticulum junction in skeletal muscle. J. Cell Biol. 107, 2587–2600.CrossRefGoogle ScholarPubMed
Boyle, P. J. and Conway, E. J. (1941). Potassium accumulation in muscle and associated changes. J. Physiol. 100, 1–63.CrossRefGoogle ScholarPubMed
Brock, L. G., Coombs, J. S. and Eccles, J. C. (1952). The recording of potentials from motoneurones with an intracellular electrode. J. Physiol. 117, 431–460.CrossRefGoogle ScholarPubMed
Broomand, A. and Elinder, F. (2008). Large-scale movement within the voltage-sensor paddle of a potassium channel-support for a helical-screw motion. Neuron 59, 770–777.CrossRefGoogle ScholarPubMed
Bülbring, E. (1979). Post junctional adrenergic mechanisms. Brit. Med. Bull. 35, 285–294.CrossRefGoogle Scholar
Buller, A. J. (1975). The Contractile Behaviour of Mammalian Skeletal Muscle (Oxford Biology Reader No. 36). London: Oxford University Press.Google Scholar
Cain, D. F., Infante, A. A. and Davies, R. E. (1962). Chemistry of muscle contraction. Adenosine triphosphate and phosphoryl creatine as energy supplies for single contractions of working muscle. Nature 196, 214–217.CrossRefGoogle Scholar
Caldwell, P. C., Hodgkin, A. L., Keynes, R. D. and Shaw, T. I. (1960). The effects of injecting ‘energy-rich’ phosphate compounds on the active transport of ions in the giant axons ofLoligo. J. Physiol. 152, 561–590.CrossRefGoogle Scholar
Caldwell, P. C. and Keynes, R. D. (1957). The utilization of phosphate bond energy for sodium extrusion from giant axons. J. Physiol. 137, 12–13P.Google ScholarPubMed
Catterall, W. A. (1986). Molecular properties of voltage-sensitive sodium channels. Ann. Rev. Biochem. 55, 953–985.CrossRefGoogle ScholarPubMed
Catterall, W. A. (1992). Cellular and molecular biology of voltage-gated ion channels. Physiol. Rev. 72, S15–S48.CrossRefGoogle Scholar
Catterall, W. A. (2001). A one-domain voltage-gated sodium channel in bacteria. Science 294, 2306–2308.CrossRefGoogle ScholarPubMed
Chandler, W. K. and Meves, H. (1970). Evidence for two types of sodium conductance in axons perfused with sodium fluoride solution. J. Physiol. 211, 653–678.CrossRefGoogle ScholarPubMed
Clausen, T. (2003). Na+-K+ pump regulation and skeletal muscle contractility. Physiol. Rev. 83, 1269–1324.CrossRefGoogle ScholarPubMed
Cole, K. S. and Curtis, H. J. (1939). Electric impedance of the squid giant axon during activity. J. Gen. Physiol. 22, 649–670.CrossRefGoogle ScholarPubMed
Colquhoun, D. and Sakmann, B. (1985). Fast events in single-channel currents activated by acetylcholine and its analogues at the frog muscle end-plate. J. Physiol. 369, 501–557.CrossRefGoogle ScholarPubMed
Conway, E. J. (1957). Nature and significance of concentration relations of potassium and sodium ions in skeletal muscle. Physiol. Rev. 37, 84–132.CrossRefGoogle ScholarPubMed
Coombs, J. S., Eccles, J. C. and Fatt, P. (1955a). Excitatory synaptic action in motoneurones. J. Physiol. 130, 374–395.CrossRefGoogle ScholarPubMed
Coombs, J. S., Eccles, J. C. and Fatt, P. (1955b). The specific ionic conductances and the ionic movements across the motoneuronal membrane that produce the inhibitory postsynaptic potential. J. Physiol. 130, 326–373.CrossRefGoogle Scholar
Dale, H. H., Feldburg, W. and Vogt, M. (1936). Release of acetylcholine at voluntary motor nerve endings. J. Physiol. 86, 353–380.CrossRefGoogle ScholarPubMed
Davson, H. and Danielli, J. F. (1943). The Permeability of Natural Membranes. Cambridge: Cambridge University Press.Google Scholar
del Castillo, J. and Katz, B. (1954). Quantal components of the end-plate potential. J. Physiol. 124, 560–573.CrossRefGoogle ScholarPubMed
del Castillo, J. and Katz, B. (1955). On the localization of acetylcholine receptors. J. Physiol. 128, 157–181.CrossRefGoogle ScholarPubMed
del Castillo, J. and Moore, J. W. (1959). On increasing the velocity of a nerve impulse. J. Physiol. 148, 665–670.CrossRefGoogle ScholarPubMed
DiFrancesco, D. (1993). Pacemaker mechanisms in cardiac tissue. Ann. Rev. Physiol. 55, 455–472.CrossRefGoogle ScholarPubMed
DiFrancesco, D. and Noble, D. (1985). A model of cardiac electrical activity incorporating ionic pumps and concentration changes. Phil. Trans. R. Soc. Lond. B307, 353–398.CrossRefGoogle Scholar
Doyle, D. A., Cabral, J. M., Pfuetzen, R. A. et al. (1998). The structure of the potassium channel: molecular basis of K+ conduction and selectivity. Science 280, 69–77.CrossRefGoogle ScholarPubMed
Eccles, J. C. (1964). The Physiology of Synapses. Berlin: Springer Verlag.CrossRefGoogle Scholar
Einthoven, W. (1924). The string galvanometer and measurement of the action currents of the heart. Nobel Lecture. Republished in 1965 in Nobel Lectures, Physiology or Medicine 1921–41. Amsterdam: Elsevier.Google Scholar
Emslie-Smith, D., Paterson, C.R., Scratcherd, T. and Read, N.W. (1988). Textbook of Physiology, 11th edn. Edinburgh: Churchill-Livingstone.
Erlanger, J. and Gasser, H. S. (1937). Electrical Signs of Nervous Activity. Philadelphia: University of Pennsylvania Press.Google Scholar
Fatt, P. and Katz, B. (1951). An analysis of the end-plate potential recorded with an intracellular electrode. J. Physiol. 115, 320–369.CrossRefGoogle ScholarPubMed
Fatt, P. and Katz, B. (1952). Spontaneous subthreshold activity at motor nerve endings. J. Physiol. 117, 109–128.Google ScholarPubMed
Fawcett, D.W. and McNutt, N. S. (1969). The ultrastructure of cat myocardium. I. Ventricular papillary muscle. J. Cell Biol. 42, 1–45.CrossRefGoogle ScholarPubMed
Ferenczi, E. A., Fraser, J. A., Chawla, S.et al. (2004). Membrane potential stabilization in amphibian skeletal muscle fibres in hypertonic solutions. J. Physiol. 555, 423–438.CrossRefGoogle ScholarPubMed
Frankenhaeuser, B. and Hodgkin, A. L. (1957). The action of calcium on the electrical properties of squid axons. J. Physiol. 137, 218–244.CrossRefGoogle ScholarPubMed
Franzini-Armstrong, C. and Jorgensen, A. O. (1994). Structure and development of e–c coupling units in skeletal muscle. Ann. Rev. Physiol. 56, 509–534.CrossRefGoogle ScholarPubMed
Fraser, J. A. and Huang, C. L.-H. (2004). A quantitative analysis of cell volume and resting potential determination and regulation in excitable cells. J. Physiol. 559, 459–478.CrossRefGoogle ScholarPubMed
Fraser, J. A. and Huang, C. L.-H. (2007). Quantitative techniques for steady-state calculation and dynamic integrated modelling of membrane potential and intracellular ion concentrations. Prog. Biophys. Mol. Biol. 94, 336–372.CrossRefGoogle ScholarPubMed
Fraser, J. A., Middlebrook, C. E., Usher-Smith, J. A., Schweining, C. J. and Huang, C. L.-H. (2005). The effect of intracellular acidification on the relationship between cell volume and membrane potential in amphibian skeletal muscle. J. Physiol. 563, 745–764.CrossRefGoogle ScholarPubMed
Fraser, J. A., Skepper, J. N., Hockaday, A. R. and Huang, C. L.-H. (1998). The tubular vacuolation process in amphibian skeletal muscle. J. Musc. Res. Cell Motility 19, 613–629.CrossRefGoogle ScholarPubMed
Gordon, A. M., Huxley, A. F. and Julian, F. J. (1966). The variation in isometric tension with sarcomere length in vertebrate muscle fibres. J. Physiol. 184, 170–192.CrossRefGoogle ScholarPubMed
Gulbis, J. M. and Doyle, D. A. (2004). Potassium channel structures: do they conform? Curr. Opin. Struct. Biol. 14, 440–446.CrossRefGoogle ScholarPubMed
Gussak, I. and Antzelevich, C (2003). Cardiac Repolarization: Bridging Basic and Clinical Science. Totowa NJ: Humana Press Inc.CrossRef
Guy, H. R. (1988). A model relating the structure of the sodium channel to its functions. Curr. Topics Membr. Transp. 33, 289–308.CrossRefGoogle Scholar
Hamill, O. P., Marty, A., Neher, E., Sakmann, B. and Sigworth, F. J. (1981). Improved patch-clamp techniques for high-resolution current recording from cells and cell-free membrane patches. Pflügers Arch. 391, 85–100.CrossRefGoogle ScholarPubMed
Hill, A. V. (1938). The heat of shortening and the dynamic constants of muscle. Proc. R. Soc. Lond. B126, 136–195.CrossRefGoogle Scholar
Hill, A. V. (1950a). A challenge to biochemists. Biochim. Biophys. Acta 4, 4–11.CrossRefGoogle ScholarPubMed
Hill, A. V. (1950b). The dimensions of animals and their muscular dynamics. Sci. Prog. Lond. 38, 209–230.Google Scholar
Hill, A. V. and Hartree, W. (1920). The four phases of heat production of muscle. J. Physiol. 54, 84–128.CrossRefGoogle ScholarPubMed
Hille, B. (1971). The hydration of sodium ions crossing the nerve membrane. Proc. Nat. Acad. Sci. USA. 68, 280–282.CrossRefGoogle ScholarPubMed
Hodgkin, A. L. (1939). The relation between conduction velocity and the electrical resistance outside a nerve fibre. J. Physiol. 94, 560–70.CrossRefGoogle ScholarPubMed
Hodgkin, A. L. (1951). The ionic basis of electrical activity in nerve and muscle. Biol. Rev. 26, 339–409.CrossRefGoogle Scholar
Hodgkin, A. L. (1958). Ionic movements and electrical activity in giant nerve fibres. Proc. R. Soc. Lond. B148, 1–37.CrossRefGoogle Scholar
Hodgkin, A. L. (1975). The optimum density of sodium channels in an unmyelinated nerve. Phil. Trans. R. Soc. Lond. B270, 297–300.CrossRefGoogle Scholar
Hodgkin, A. L. and Horowicz, P. (1957). The differential action of hypertonic solutions on the twitch and action potential of a muscle fibre. J. Physiol. 136, 17–18P.Google Scholar
Hodgkin, A. L. and Horowicz, P. (1959). The influence of potassium and chloride ions on the membrane potential of single muscle fibres. J. Physiol. 148, 127–160.CrossRefGoogle ScholarPubMed
Hodgkin, A. L. and Horowicz, P. (1960). Potassium contractures in single muscle fibres. J. Physiol. 153, 386–403.CrossRefGoogle ScholarPubMed
Hodgkin, A. L. and Huxley, A. F. (1952). A quantitative description of membrane current and its application to conduction and excitation in nerve. J. Physiol. 117, 500–544.CrossRefGoogle ScholarPubMed
Hodgkin, A. L., Huxley, A. F. and Katz, B. (1952). Measurement of current voltage relations in the membrane of the giant axon ofLoligo. J. Physiol. 116, 424–448.CrossRefGoogle Scholar
Hodgkin, A. L. and Katz, B. (1949). The effects of sodium ions on the electrical activity of the giant axon of the squid. J. Physiol. 108, 37–77.CrossRefGoogle Scholar
Hodgkin, A. L. and Keynes, R. D. (1955a). Active transport of cations in giant axons from, Sepia and Loligo. J. Physiol. 128, 28–60.CrossRefGoogle ScholarPubMed
Hodgkin, A. L. and Keynes, R. D. (1955b). The potassium permeability of a giant nerve fibre. J. Physiol. 128, 61–88.CrossRefGoogle ScholarPubMed
Hoffman, B.F. and Cranefield, P.F. (1960). Electrophysiology of the Heart. New York: McGrawHill.Google Scholar
Homsher, E. (1987). Muscle enthalpy production and its relationship to actomyosin ATPase. Ann. Rev. Physiol. 49, 673–690.CrossRefGoogle ScholarPubMed
Huang, C.L.-H. (1982). Pharmacological separation of charge movement components in frog skeletal muscle. J. Physiol. 324, 375–387.CrossRefGoogle ScholarPubMed
Huang, C. L.-H. (1990). Voltage-dependent block of charge movement components by nifedipine in frog skeletal muscle. J. Gen. Physiol. 96, 535–557.CrossRefGoogle ScholarPubMed
Huang, C.L.-H. (1994) Charge conservation in intact frog skeletal muscle fibres in gluconate-containing solutions. J. Physiol. 474, 161–171.CrossRefGoogle ScholarPubMed
Huang, C. L.-H. (1996). Kinetic isoforms of intramembrane charge in intact amphibian striated muscle. J. Gen. Physiol. 107, 515–534.CrossRefGoogle ScholarPubMed
Huang, C. L.-H. (1998). The influence of caffeine on intramembrane charge movements in intact frog striated muscle. J. Physiol. 512, 707–721.CrossRefGoogle ScholarPubMed
Huang, C. L.-H. and Peachey, L. D. (1989). The anatomical distribution of voltage-dependent membrane capacitance in frog skeletal muscle fibres. J. Gen. Physiol. 93, 565–584.CrossRefGoogle Scholar
Huxley, A. F. and Niedergerke, R. (1954). Structural changes in muscle during contraction. Interference microscopy of living muscle fibres. Nature 173, 971–973.CrossRefGoogle ScholarPubMed
Huxley, A. F. and Stämpfli, R. (1949). Evidence for saltatory conduction in peripheral myelinated nerve fibres. J. Physiol. 108, 315–339.CrossRefGoogle ScholarPubMed
Huxley, A. F. and Taylor, R. E. (1958). Local activation of striated muscle fibres. J. Physiol. 144, 426–441.CrossRefGoogle ScholarPubMed
Huxley, H. E. (1963). Electron microscope studies on the structure of natural and synthetic protein filaments from striated muscle. J. Mol. Biol. 7, 281–308.CrossRefGoogle ScholarPubMed
Huxley, H. E. (1976). The structural basis of contraction and regulation in skeletal muscle. In Molecular Basis of Motility, ed. Heilmeyer, L. M. G. Jr, Ruegg, J. C. and Wieland, Th.. Berlin: Springer-Verlag.Google Scholar
Huxley, H. E. (1990). Sliding filaments and molecular motile systems. J. Biol. Chem. 265, 8347–8350.Google ScholarPubMed
Huxley, H. E. and Hanson, J. (1954). Change in the cross-striations of muscle during contraction and stretch and their structural interpretation. Nature 173, 973–976.CrossRefGoogle Scholar
Jiang, Y., Lee, A., Chen, J. et al. (2003). X-ray structure of a voltage-dependent K+ channel. Nature 423, 33–41.CrossRefGoogle ScholarPubMed
Kato, G. (1936). On the excitation, conduction, and narcotization of single nerve fibers. Cold Spr. Harb. Symp. Quant. Biol. 4, 202–213.CrossRefGoogle Scholar
Katz, B. and Miledi, R. (1965). The measurement of synaptic delay, and the time course of acetylcholine release at the neuromuscular junction. Proc. R. Soc. Lond. B 161, 483–495.CrossRefGoogle ScholarPubMed
Katz, B. and Miledi, R. (1967). The timing of calcium action during neuromuscular transmission. J. Physiol. 189, 535–544.CrossRefGoogle ScholarPubMed
Keynes, R. D. (1951). The ionic movements during nervous activity. J. Physiol. 114, 119–150.CrossRefGoogle ScholarPubMed
Keynes, R. D. (1963). Chloride in the squid giant axon. J. Physiol. 169, 690–705.CrossRefGoogle ScholarPubMed
Keynes, R. D. (1994). The kinetics of voltage-gated ion channels. Q. Rev. Biophys. 27, 339–434.CrossRefGoogle ScholarPubMed
Keynes, R. D. and Elinder, F. (1998a). On the slowly rising phase of the sodium gating current in the squid giant axon. Proc. R. Soc. Lond. B 265, 255–262.CrossRefGoogle ScholarPubMed
Keynes, R. D. and Elinder, F. (1998b). Modelling the activation, opening, inactivation and reopening of the voltage-gated sodium channel. Proc R. Soc. Lond. B: Biol Sci. 265, 263–270.CrossRefGoogle ScholarPubMed
Keynes, R. D. and Elinder, F. (1999). The screw-helical voltage gating of ion channels. Proc. R. Soc. Lond. B 266, 843–852.CrossRefGoogle ScholarPubMed
Keynes, R. D. and Lewis, P. R. (1951). The sodium and potassium content of cephalopod nerve fibres. J. Physiol. 114, 151–182.CrossRefGoogle Scholar
Keynes, R. D. and Martins-Ferreira, H. (1953). Membrane potentials in the electroplates of the electric eel. J. Physiol. 119, 315–351.CrossRefGoogle ScholarPubMed
Keynes, R. D. and Meves, H. (1993). Properties of the voltage sensor for the opening and closing of the sodium channels in the squid giant axon. R. Soc. Lond. B: Biol Sci. 253, 61–68.CrossRefGoogle ScholarPubMed
Keynes, R. D. and Ritchie, J. M. (1984). On the binding of labelled saxitoxin to the squid giant axon. Proc. R. Soc. Lond. B 222, 147–153.CrossRefGoogle ScholarPubMed
Keynes, R. D. and Rojas, E. (1973). Characteristics of the sodium gating current in the squid giant axon. J. Physiol. 233, 28–30PGoogle ScholarPubMed
Keynes, R. D. and Rojas, E. (1974). Kinetics and steady-state properties of the charged system controlling sodium conductance in the squid giant axon. J Physiol. 239, 393–434.CrossRefGoogle ScholarPubMed
Killeen, M. J., Sabir, I. N., Grace, A. A. and Huang, C. L.-H. (2008). Dispersions of repolarization and ventricular arrhythmogenesis: lessons from animal models. Prog. Biophys. Mol. Biol. 98, 219–229.CrossRefGoogle ScholarPubMed
Koeppen, B. M. and Stanton, B. A. (2009). Berne & Levy: Principles of Physiology, 6th edn. New York: Mosby, 848pp.Google Scholar
Kovacs, L., Rios, E. and Schneider, M. F. (1979). Calcium transients and intramembrane charge movement in skeletal muscle fibres. Nature 279, 391–396.CrossRefGoogle ScholarPubMed
Kuffler, S. W. (1980). Slow synaptic responses in the autonomic ganglia and the pursuit of a peptidergic transmitter. J. Exp. Biol. 89, 257–286.Google ScholarPubMed
Kuffler, S. W. and Yoshikami, D. (1975). The number of transmitter molecules in a quantum: an estimate from iontophoretic application of acetylcholine at the neuromuscular synapse. J. Physiol. 251, 465–482.CrossRefGoogle Scholar
Kyte, J. and Doolittle, R. F. (1982). A simple method for displaying the hydropathic character of a protein. J. Mol. Biol. 157, 105–132.CrossRefGoogle ScholarPubMed
Lamb, J. F., Ingram, C. G., Johnston, I. A. and Pitman, R. M. (1991). Essential Physiology. 3rd edn. Oxford: WileyBlackwell.Google Scholar
Larsson, H. P., Baker, O. S., Dhillon, D. S. and Isacoff, E. Y. (1996). Transmembrane movement of the Shaker K+ channel S4. Neuron 16, 387–397.CrossRefGoogle Scholar
Lei, M., Goddard, C., Liu, J.et al. (2005). Sinus node dysfunction following targeted disruption of the murine cardiac sodium channel gene, SCN5A. J. Physiol. 567, 387–400.CrossRefGoogle ScholarPubMed
Levy, M. N., Koeppen, B. M and Stanton, B. A. (2005). Berne & Levy: Principles of Physiology, 4th edn. New York: Mosby.Google Scholar
Loewi, O. (1921). Über humorale Übertragbarkeit der Herzner-venwirkung. Pflügers Arch. Ges. Physiol. 189, 239–242.CrossRefGoogle Scholar
Makowski, L., Caspar, D. L. D., Phillips, W. C., Baker, T. S. and Goodenough, D. A. (1984). Gap junction structures VI. Variation and conservation in connexon conformation and packing. Biophys. J. 45, 208–218.CrossRefGoogle ScholarPubMed
Maylie, J., Irving, M., Sizto, N. L. and Chandler, W. K. (1987). Calcium signals recorded from cut frog twitch fibers containing Antipyrylazo III. J. Gen. Physiol. 89, 83–143.CrossRefGoogle ScholarPubMed
McCleskey, E. W. (1999). Calcium channel permeation: a field in flux. J. Gen. Physiol. 113, 765–772.CrossRefGoogle ScholarPubMed
Merton, P. A. (1954). Voluntary strength and fatigue. J. Physiol. 128, 553–564.CrossRefGoogle Scholar
Merton, P. A., Hill, D. K. and Morton, H. B. (1981). Indirect and direct stimulation of fatigued human muscle. In Human Muscle Fatigue: Physiological Mechanisms, ed. Porter, R. and Whelan, J., pp. 120–126. London: Pitman Medical.Google Scholar
Neher, E. and Sakmann, B. (1976). Single-channel currents recorded from membrane of denervated frog muscle cells. Nature 260, 799–802.CrossRefGoogle Scholar
Nielsen, O. B., Paoli, F., and Overgaard, K. (2001). Protective effects of lactic acid on force production in rat skeletal muscle. J. Physiol. 536: 161–166.CrossRefGoogle ScholarPubMed
Noble, D. (1979). The Initiation of the Heartbeat, 2nd edn. Oxford: Oxford University Press.Google Scholar
Noda, M., Shimizu, S., Tanabe, T.et al. (1984). Primary structure ofElectrophorus electricus sodium channel deduced from cDNA sequence. Nature 312, 121–127.Google ScholarPubMed
Noda, M., Takahashi, H., Tanabe, T.et al. (1982). Primary structure of α-subunit precursor of Torpedo californica acetylcholine receptor deduced from cDNA sequence. Nature 279, 793–797.CrossRefGoogle Scholar
Offer, G. (1974). The molecular basis of muscular contraction. In Companion to Biochemistry, ed. Bull, A. T., Lagnado, J. R., Thomas, J. O. and Tipton, K. F., pp. 623–671. London: Longman.Google Scholar
Ostrowski, J., Kjelsberg, M. A., Caron, M. G. and Lefkowitz, R. J. (1992). Mutagenesis of the β2-adrenergic receptor: how structure elucidates function. Ann. Rev. Pharmacol. Toxicol. 32, 167–183.CrossRefGoogle Scholar
Padmanabhan, N. and Huang, C. L.-H. (1990). Separation of tubular electrical activity in amphibian skeletal muscle through temperature change. Exp. Physiol. 75, 721–724.CrossRefGoogle ScholarPubMed
Papadatos, G. A., Wallerstein, P. M. R., Head, C. E. G.et al. (2002). Slowed conduction and ventricular tachycardia after targeted disruption of the cardiac sodium channel Scn5a. Proc. Nat. Acad. Sci. 99, 6210–6215.CrossRefGoogle Scholar
Peachey, L. D. (1965). The sarcoplasmic reticulum and transverse tubules of the frog's sartorius. J. Cell Biol. 25, 209–232.CrossRefGoogle ScholarPubMed
Pedersen, T. H, Macdonald, W. A., dePaoli, F. V., Gurung, I. S. and Nielsen, O. B. (2009). Comparison of regulated passive membrane conductance in action potential firing fast and slow-twitch muscle. J. Gen. Physiol. 134, 323–337.CrossRefGoogle ScholarPubMed
Pedersen, T. H., Nielsen, O. B., Lamb, G. D., and Stephenson, D. G. (2004). Intracellular acidosis enhances the excitability of working muscle. Science 305, 1144–1147.CrossRefGoogle ScholarPubMed
Porter, K. R. and Palade, G. E. (1957). Studies on the endoplasmic reticulum. III. Its form and distribution in striated muscle cells. J. Biophys. Biochem. Cytol. 3, 269–300.CrossRefGoogle ScholarPubMed
Rayment, I. and Holden, H. M. (1994). The three-dimensional structure of a molecular motor. Trends Biochem. Sci. 19, 129–134.CrossRefGoogle ScholarPubMed
Rayment, I., Smith, C. and Yount, R. G. (1996). The active site of myosin. Ann. Rev. Physiol. 58, 671–702.CrossRefGoogle ScholarPubMed
Rios, E. and Brum, G. (1987). Involvement of dihydropyridine receptors in excitation-contraction coupling in skeletal muscle. Nature 325, 717–720.CrossRefGoogle ScholarPubMed
Ritchie, J. M. and Rogart, R. B. (1977). The binding of saxitoxin and tetrodotoxin to excitable tissue. Rev. Physiol. Biochem. Pharmacol. 79, 1–50.CrossRefGoogle ScholarPubMed
Robertson, J. D. (1960). The molecular structure and contact relationships of cell membranes. Prog. Biophys. 10, 343–418.Google ScholarPubMed
Rushton, W. A. H. (1933). Lapicque's theory of curarization. J. Physiol. 77, 337–364.CrossRefGoogle ScholarPubMed
Ryall, R. W. (1979). Mechanisms of Drug Action on the Nervous System. Cambridge: Cambridge University Press.Google Scholar
Scher, A. M. (1965). Electrical correlates of the cardiac cycle. In Physiology and Biophysics, ed. Ruch, T. C. and Patten, H. D.. Philadelphia: Saunders, pp. 565–599.Google Scholar
Schmidt-Nielsen, K. (1990). Animal Physiology, 4th edn. Cambridge: Cambridge University Press.Google Scholar
Schneider, M.F. and Chandler, W. K. (1973). Voltage-dependent charge in skeletal muscle: a possible step in excitation-contraction coupling. Nature 242, 244–246.CrossRefGoogle ScholarPubMed
Schwartz, L. M., McCleskey, E. W. and Almers, W. (1985). Dihydropyridine receptors in muscle are voltage-dependent but most are not functional calcium channels. Nature. 314, 747–751.CrossRefGoogle Scholar
Sejersted, O. M. and Sjøgård, G. (2000). Dynamics and consequences of potassium shifts in skeletal muscle and heart during exercise. Physiol. Rev. 80, 1411–1481.CrossRefGoogle ScholarPubMed
Sheikh, S. M., Skepper, J. N., Chawla, S.et al. (2001). Normal conduction of surface action potentials in detubulated amphibian skeletal muscle fibres. J. Physiol. 535, 579–590.CrossRefGoogle ScholarPubMed
Skou, J. C. (1957). The influence of some cations on an adenosine triphosphatase from peripheral nerves. Biochim. Biophys. Acta 23, 394–401.CrossRefGoogle ScholarPubMed
Skou, J. C. (1998). Nobel Lecture. The identification of the sodium pump. Biosci Rep. 18, 155–169.CrossRefGoogle ScholarPubMed
Spudich, J. A. (1994). How molecular motors work. Nature 372, 515–518.CrossRefGoogle ScholarPubMed
Spudich, J. A., Finer, J., Simmons, B.et al. (1995). Myosin structure and function. Cold Spr. Harb. Symp. Quant. Biol. 60, 783–71.CrossRefGoogle ScholarPubMed
Squire, J. M. (1986). Muscle: Design, Diversity and Disease. Menlo Park, California: Benjamin/Cummings.Google Scholar
Stokoe, K. S., Thomas, G., Goddard, C. A.et al. (2007). Effects of flecainide and quinidine on arrhythmogenic properties of Scn5a+/Δ murine hearts modelling long QT syndrome 3. J. Physiol. 578, 69–84.CrossRefGoogle ScholarPubMed
Takeshima, H., Nishimura, S., Matsumoto, T.et al. (1989). Primary structure and expression from complementary DNA of skeletal muscle ryanodine receptor. Nature 339, 439–445.CrossRefGoogle ScholarPubMed
Takeuchi, A. and Takeuchi, N. (1959). Active phase of frog's end-plate potential. J. Neurophysiol. 22, 395–411.CrossRefGoogle ScholarPubMed
Takeuchi, A. and Takeuchi, N. (1960). On the permeability of the end-plate membrane during the action of the transmitter. J. Physiol. 154, 52–67.CrossRefGoogle ScholarPubMed
Tasaki, I. (1953). Nervous Transmission. Springfield, Illinois: Charles C. Thomas.Google Scholar
Unwin, N. (1993). Nicotinic acetylcholine receptor at 9 Å resolution. J. Mol. Biol. 229, 1101–1124.CrossRefGoogle ScholarPubMed
Unwin, N. (1995). Acetylcholine receptor channel imaged in the open state. Nature 373, 37–43.CrossRefGoogle ScholarPubMed
Usher-Smith, J. A., Fraser, J. A., Bailey, P. S. J., Griffin, J. L., and Huang, C. L-H. (2006a). The influence of intracellular lactate and H+ on cell volume in amphibian skeletal muscle. J. Physiol. 573, 799–818.CrossRefGoogle ScholarPubMed
Usher-Smith, J. A., Huang, C. L.-H. and Fraser, J. A. (2009). Control of cell volume in skeletal muscle. Biol Rev Camb Philos Soc. 84, 143–159.CrossRefGoogle ScholarPubMed
Usher-Smith, J. A., Skepper, J. N., Fraser, J. A. and Huang, C. L.-H. (2006b). Effect of repetitive stimulation on cell volume and its relationship to membrane potential in amphibian skeletal muscle. Eur. J. Physiol. 452: 231–239.CrossRefGoogle ScholarPubMed
Weidmann, S. (1956). Elektrophysiologie der Herzmuskelfaser. Huber: Berne.Google Scholar
Whittaker, V. P. (1984). The structure and function of cholinergic synaptic vesicles. Biochem. Soc. Trans. 12, 561–576.CrossRefGoogle ScholarPubMed
Wilkie, D. R. (1968). Heat work and phosphorylcreatine breakdown in muscle. J. Physiol. 195, 157–183.CrossRefGoogle ScholarPubMed
Wilkie, D. R. (1976). Energy transformation in muscle. In Molecular Basis of Motility, ed. Heilmeyer, L. M. G. Jr., Ruegg, J. C. and Wieland, Th., pp. 69–80. Berlin: Springer-Verlag.Google Scholar
Yang, N., George, A. L. and Horn, R. (1996). Molecular basis of charge movement in voltage-gated sodium channels. Neuron 16, 113–122.CrossRefGoogle ScholarPubMed
Zhang, F., Wang, L. P., Boyden, E. S. and Deisseroth, K. (2006). Channelrhodopsin-2 and optical control of excitable cells. Nature Methods. 3, 785–792.CrossRefGoogle ScholarPubMed

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×