Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-pftt2 Total loading time: 0 Render date: 2024-05-09T09:08:25.749Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  07 November 2020

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
Publisher: Cambridge University Press
Print publication year: 2020

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, BA and Beam, KG (1990). Muscular dysgenesis in mice: a model system for studying excitation-contraction coupling. FASEB J 4, 28092816.Google Scholar
Adams, DJ and Gage, PW (1979). Sodium and calcium gating currents in an Aplysia neurone. J Physiol 291, 467481.CrossRefGoogle Scholar
Adrian, RH (1956). The effect of internal and external potassium concentration on the membrane potential of frog muscle. J Physiol 133, 631658.Google Scholar
Adrian, RH (1964). The rubidium and potassium permeability of frog muscle membrane. J Physiol 175, 134159.CrossRefGoogle ScholarPubMed
Adrian, ED and Lucas, K (1912). On the summation of propagated disturbances in nerve and muscle. J Physiol 44, 68124.Google Scholar
Adrian, RH and Almers, W (1976a). Charge movement in the membrane of striated muscle. J Physiol 254, 339360.Google Scholar
Adrian, RH and Almers, W (1976b). The voltage dependence of membrane capacity. J Physiol 254, 317338.Google Scholar
Adrian, RH and Bryant, SH (1974). On the repetitive discharge in myotonic muscle fibres. J Physiol 240, 505515.Google Scholar
Adrian, RH and Peachey, LD (1973). Reconstruction of the action potential of frog sartorius muscle. J Physiol 235, 103131.Google Scholar
Adrian, RH, Chandler, WK and Hodgkin, AL (1970). Voltage clamp experiments in striated muscle fibres. J Physiol 208, 607644.Google Scholar
Adrian, RH, Costantin, LL and Peachey, LD (1969). Radial spread of contraction in frog muscle fibres. J Physiol 204, 231257.Google Scholar
Ahmad, S, Valli, H, Edling, CE, et al. (2017). Effects of ageing on pro-arrhythmic ventricular phenotypes in incrementally paced murine Pgc-1β–/– hearts. Pflügers Arch 469, 15791590.CrossRefGoogle ScholarPubMed
Ahmad, S, Valli, H, Smyth, R, et al. (2019). Reduced cardiomyocyte Na+ current in the age-dependent murine Pgc-1β–/– model of ventricular arrhythmia. J Cell Physiol 234, 39213932.Google Scholar
Aidley, D (1998). The Physiology of Excitable Cells 4/e. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Allen, DG, Lamb, GD and Westerblad, H (2008). Skeletal muscle fatigue: cellular mechanisms. Physiol Rev 88, 287332.CrossRefGoogle ScholarPubMed
Almers, W, Stanfield, PR and Stühmer, W (1983). Lateral distribution of sodium and potassium channels in frog skeletal muscle: measurements with a patch-clamp technique. J Physiol 336, 261284.CrossRefGoogle ScholarPubMed
Anderson, BR and Granzier, HL (2012). Titin-based tension in the cardiac sarcomere: molecular origin and physiological adaptations. Prog Biophys Mol Biol 110, 204217.Google Scholar
Armstrong, CM and Bezanilla, F (1973). Currents related to movement of the gating particles of the sodium channels. Nature 242, 459461.Google Scholar
Armstrong, CM and Bezanilla, F (1974). Charge movement associated with the opening and closing of the activation gates of the Na channels. J Gen Physiol 63, 533552.Google Scholar
Armstrong, CM and Bezanilla, F (1977). Inactivation of the sodium channel. II. Gating current experiments. J Gen Physiol 70, 567590.Google Scholar
Ashley, CC and Ridgway, EB (1968). Simultaneous recording of membrane potential, calcium transient and tension in single muscle fibres. Nature 219, 11681169.Google Scholar
Ashley, CC and Ridgway, EB (1970). On the relationships between membrane potential, calcium transient and tension in single barnacle muscle fibres. J Physiol 209, 105130.CrossRefGoogle ScholarPubMed
Bagshawe, C (1993). Muscle Contraction 2/e. London: Chapman & Hall.Google Scholar
Baker, PF, Hodgkin, AL and Shaw, TI (1962). The effects of changes in internal ionic concentrations on the electrical properties of perfused giant axons. J Physiol 164, 355374.Google Scholar
Balasubramaniam, R, Chawla, S, Grace, AA and Huang, CL-H (2005). Caffeine-induced arrhythmias in murine hearts parallel changes in cellular Ca2+ homeostasis. Am J Physiol Heart Circ Physiol 289, H1584H1593.Google Scholar
Balasubramaniam, R, Chawla, S, Mackenzie, L, et al. (2004). Nifedipine and diltiazem suppress ventricular arrhythmogenesis and calcium release in mouse hearts. Pflügers Arch 449, 150158.Google Scholar
Balasubramaniam, R, Grace, A, Saumarez, R, Vandenberg, J and Huang, CL-H (2003). Electrogram prolongation and nifedipine-suppressible ventricular arrhythmias in mice following targeted disruption of KCNE1. J Physiol 552, 535546.Google Scholar
Barnard, EA, 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 London - Biol Sci 215, 241246.Google Scholar
Baruscotti, M, Bottelli, G, Milanesi, R, DiFrancesco, JC and DiFrancesco, D (2010). HCN-related channelopathies. Pflügers Arch 460, 405415.CrossRefGoogle ScholarPubMed
Baylor, SM, Chandler, WK and Marshall, MW (1983). Sarcoplasmic reticulum calcium release in frog skeletal muscle fibres estimated from arsenazo III calcium transients. J Physiol 344, 625666.Google Scholar
Belardinelli, L, Giles, W, Rajamani, S, Karagueuzian, H and Shryock, J (2015). Cardiac late Na+ current: Proarrhythmic effects, roles in long QT syndromes, and pathological relationship to CaMKII and oxidative stress. Heart Rhythm 12, 440448.Google Scholar
Ben-Johny, M, Yang, PS, Niu, J, et al. (2014). Conservation of Ca2+/Calmodulin Regulation across Na and Ca2+ channels. Cell 157, 16571670.Google Scholar
Bers, DM (2001). Excitation-Contraction Coupling and Cardiac Contractile Force 2/e. Dordrecht, The Netherlands: Kluwer Academic Publishers.Google Scholar
Bezanilla, F (2018). Gating currents. J Gen Physiol 150, 911932.Google Scholar
Blaauw, B, Schiaffino, S and Reggiani, C (2013). Mechanisms modulating skeletal muscle phenotype. Compr Physiol 3, 16451687.Google Scholar
Bliss, TVP and Cooke, SF (2011). Long-term potentiation and long-term depression: a clinical perspective. Clinics 66, 317.Google Scholar
Bliss, TVP and Lomo, T (1973). Long-lasting potentiation of synpatic transmission in the dentate area of the anaesthetized rabbit following stimulation of the perforant path. J Physiol 232, 331356.Google Scholar
Block, BA, Imagawa, T, Campbell, KP 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, 25872600.Google Scholar
Boyle, PJ and Conway, EJ (1941). Potassium accumulation in muscle and associated changes. J Physiol 100, 163.Google Scholar
Brock, LG, Coombs, JS and Eccles, JC (1952). The recording of potentials from motoneurones with an intracellular electrode. J Physiol 117, 431460.Google Scholar
Brodal, P (2016). The Central Nervous System 5/e. New York: Oxford University Press.Google Scholar
Brugada, R (2000). Use of intravenous antiarrhythmics to identify concealed Brugada syndrome. Curr Control Trials Cardiovasc Med 1, 4547.CrossRefGoogle ScholarPubMed
Bruns, D and Jahn, R (1995). Real-time measurement of transmitter release from single synaptic vesicles. Nature 377, 6265.Google Scholar
Bülbring, E (1979). Post junctional adrenergic mechanisms. Brit Med Bull 35, 285294.Google Scholar
Buller, A (1975). The Contractile Behaviour of Mammalian Skeletal Muscle (Oxford Biology Reader No. 36). London: Oxford University Press.Google Scholar
Cain, DF, Infante, AA and Davies, RE (1962). Chemistry of muscle contraction: adenosine triphosphate and phosphorylcreatine as energy supplies for single contractions of working muscle. Nature 196, 214217.CrossRefGoogle ScholarPubMed
Caldwell, PC and Keynes, R (1957). The utilization of phosphate bond energy for sodium extrusion from giant axons. J Physiol 137, 12P13P.Google Scholar
Caldwell, PC, Hodgkin, AL, Keynes, RD and Shaw, TI (1960). The effects of injecting “energy-rich” phosphate compounds on the active transport of ions in the giant axons of Loligo. J Physiol 152, 561590.Google Scholar
Cannon, SC (2018). Sodium channelopathies of skeletal muscle In: Handbook of Experimental Pharmacology. Berlin: Springer, 309330.Google Scholar
Catterall, WA (1992). Cellular and molecular biology of voltage-gated sodium channels. Physiol Rev 72, S15S48.Google Scholar
Chadda, KR, Ahmad, S, Valli, H et al. (2017a). The effects of ageing and adrenergic challenge on electrocardiographic phenotypes in a murine model of long QT syndrome type 3. Sci Rep 7, 11070.Google Scholar
Chadda, KR, Jeevaratnam, K, Lei, M, Huang, CL-H (2017b). Sodium channel biophysics, late sodium current and genetic arrhythmic syndromes. Pflugers Arch 469, 629641.Google Scholar
Chandler, WK, Rakowski, RF and Schneider, MF (1976). A non‐linear voltage dependent charge movement in frog skeletal muscle. J Physiol 254, 245283.Google Scholar
Chawla, S, Skepper, JN, Hockaday, AR and Huang, CL-H (2001). Calcium waves induced by hypertonic solutions in intact frog skeletal muscle fibres. J Physiol 536, 351359.Google Scholar
Chawla, S, Skepper, JN and Huang, CL-H (2002). Differential effects of sarcoplasmic reticular Ca2+-ATPase inhibition on charge movements and calcium transients in intact amphibian skeletal muscle fibres. J Physiol 539, 869882.CrossRefGoogle ScholarPubMed
Chawla, S, Vanhoutte, P, Arnold, FJL, Huang, CL-H and Bading, H (2003). Neuronal activity-dependent nucleocytoplasmic shuttling of HDAC4 and HDAC5. J Neurochem 85, 151159.Google Scholar
Cheng, EP, Yuan, C, Navedo, MF, et al. (2011). Restoration of normal L-type Ca2+ channel function during Timothy syndrome by ablation of an anchoring protein. Circ Res 109, 255261.Google Scholar
Cheng, H, Lederer, WJ and Cannell, MB (1993). Calcium sparks: elementary events underlying excitation-contraction coupling in heart muscle. Science 262, 740744.Google Scholar
Chiamvimonvat, N, Kargacin, ME, Clark, RB and Duff, HJ (1995). Effects of intracellular calcium on sodium current density in cultured neonatal rat cardiac myocytes. J Physiol 483, 307318.Google Scholar
Clausen, T (2003). Na+-K+ pump regulation and skeletal muscle contractility. Physiol Rev 83, 12691324.Google Scholar
Cole, KS (1941). Rectification and inductance in the squid giant axon. J Gen Physiol 25, 2951.CrossRefGoogle ScholarPubMed
Cole, KS and Curtis, HJ (1939). Electric impedance of the squid giant axon during activity. J Gen Physiol 22, 649670.Google Scholar
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, 501557.Google Scholar
Conway, EJ (1957). Nature and significance of concentration relations of potassium and sodium ions in skeletal muscle. Physiol Rev 37, 84132.Google Scholar
Coombs, JS, Eccles, JC and Fatt, P (1955a). Excitatory synaptic action in motoneurones. J Physiol 130, 374395.Google Scholar
Coombs, JS, Eccles, JC and Fatt, P (1955b). The specific ionic conductances and the ionic movements across the motoneuronal membrane that produce the inhibitory post-synaptic potential. J Physiol 130, 326373.CrossRefGoogle ScholarPubMed
Corrias, A, Giles, W and Rodriguez, B (2011). Ionic mechanisms of electrophysiological properties and repolarization abnormalities in rabbit Purkinje fibers. Am J Physiol - Heart Circ Physiol 300, H1806H1813.Google Scholar
Costantin, LL (2011). Activation in striated muscle. Compr Physiol 215–259.Google Scholar
Cousins, HM, Edwards, FR, Hickey, H, Hill, CE and Hirst, GDS (2003). Electrical coupling between the myenteric interstitial cells of Cajal and adjacent muscle layers in the guinea-pig gastric antrum. J Physiol 550, 829844.Google Scholar
Dale, HH, Feldberg, W and Vogt, M (1936). Release of acetylcholine at voluntary motor nerve endings. J Physiol 86, 353380.Google Scholar
Davies, L, Jin, J, Shen, W, et al. (2014). Mkk4 is a negative regulator of the transforming growth factor beta 1 signaling associated with atrial remodeling and arrhythmogenesis with age. J Am Heart Assoc 3, 119.Google Scholar
Del Castillo, J and Katz, B (1954). Quantal components of the end‐plate potential. J Physiol 124, 560573.Google Scholar
Del Castillo, J and Katz, B (1955). On the localization of acetylcholine receptors. J Physiol 128, 157181.Google Scholar
Del Castillo, J and Moore, JW (1959). On increasing the velocity of a nerve impulse. J Physiol 148, 665670.Google Scholar
DiFrancesco, D (1993). Pacemaker mechanisms in cardiac tissue. Annu Rev Physiol 55, 455471.CrossRefGoogle ScholarPubMed
DiFrancesco, D and Noble, D (1985). A model of cardiac electrical activity incorporating ionic pumps and concentration changes. Philos Trans R Soc Lond B Biol Sci 307, 353398.Google Scholar
Doyle, DA, Cabral, JM, Pfuetzner, RA, et al. (1998). The structure of the potassium channel: Molecular basis of K+ conduction and selectivity. Science 280, 6977.Google Scholar
Dutka, TL and Lamb, GD (2000). Effect of lactate on depolarization-induced Ca2+ release in mechanically skinned skeletal muscle fibers. Am J Physiol Cell Physiol 278, C517C525.Google Scholar
Ebashi, S and Endo, M (2003). Calcium and muscle contraction. Prog Biophys Mol Biol 18, 123183.Google Scholar
Eccles, JC (1964). The Physiology of Synapses. Berlin: Springer-Verlag.Google Scholar
Edling, CE, Fazmin, IT, Saadeh, K, et al. (2019). Molecular basis of arrhythmic substrate in ageing murine peroxisome proliferator-activated receptor γ co-activator deficient hearts modelling mitochondrial dysfunction. Biosci Rep. 39, BSR20190403.Google Scholar
Eglen, RM, Reddy, H, Watson, N and Challiss, RAJ (1994). Muscarinic acetylcholine receptor subtypes in smooth muscle. Trends Pharmacol Sci 15, 114119.Google Scholar
Einthoven, W (1924). The string galvanometer and the measurement of the action currents of the heart, Nobel lecture 1925 In: Nobel Lectures, Physiology or Medicine 1921–41. Republished 1965. Amsterdam: Elsevier 287306.Google Scholar
Emslie-Smith, D, Paterson, C, Scratcherd, T and Read, NW (1988). Textbook of Physiology 11/e. Edinburgh: Churchill-Livingstone.Google Scholar
Erlanger, J and Gasser, HS (1937). Electrical Signs of Nervous Activity. Philadelphia: University of Pennsylvania Press.Google Scholar
Fabritz, L, Kirchhof, P, Franz, MR, et al. (2003). Prolonged action potential durations, increased dispersion of repolarization, and polymorphic ventricular tachycardia in a mouse model of proarrhythmia. Basic Res Cardiol 98, 2532.Google Scholar
Fatt, P and Katz, B (1951). An analysis of the end‐plate potential recorded with an intra‐cellular electrode. J Physiol 115, 320370.Google Scholar
Fatt, P and Katz, B (1952). Spontaneous subthreshold activity at motor nerve endings. J Physiol 117, 109128.Google Scholar
Fawcett, DW and McNutt, NS (1969). The ultrastructure of the cat myocardium. I. Ventricular papillary muscle. J Cell Biol 42, 145.Google Scholar
Ferenczi, EA, Fraser, JA, Chawla, S, et al. (2004). Membrane potential stabilization in amphibian skeletal muscle fibres in hypertonic solutions. J Physiol 555, 423438.Google Scholar
Ferenczi, EA, Tan, X and Huang, CL-H (2019). Principles of optogenetic methods for application to cardiac experimental systems. Front Physiol 10, 1096.Google Scholar
Fermini, B and Fossa, AA (2003). The impact of drug-induced QT interval prolongation on drug discovery and development. Nat Rev Drug Discov 2, 439447.Google Scholar
Field, AC, Hill, C and Lamb, GD (1988). Asymmetric charge movement and calcium currents in ventricular myocytes of neonatal rat. J Physiol 406, 277297.Google Scholar
Filatov, GN, Pinter, MJ and Rich, MM (2009). Role of Ca2+ in injury-induced changes in sodium current in rat skeletal muscle. Am J Physiol Cell Physiol 297, C352C359.Google Scholar
Finck, BN and Kelly, DP (2006). PGC-1 coactivators: inducible regulators of energy metabolism in health and disease. J Clin Invest 116, 615622.Google Scholar
Frankenhaeuser, B and Hodgkin, AL (1957). The action of calcium on the electrical properties of squid axons. J Physiol 137, 218244.Google Scholar
Fraser, JA and Huang, CL-H (2004). A quantitative analysis of cell volume and resting potential determination and regulation in excitable cells. J Physiol 559, 459478.Google Scholar
Fraser, JA and Huang, CL-H (2007). Quantitative techniques for steady-state calculation and dynamic integrated modelling of membrane potential and intracellular ion concentrations. Prog Biophys Mol Biol 94, 336372.Google Scholar
Fraser, JA, Huang, CL-H and Pedersen, TH (2011). Relationships between resting conductances, excitability, and t-system ionic homeostasis in skeletal muscle. J Gen Physiol 138, 95116.Google Scholar
Fraser, JA, Middlebrook, CE, Usher-Smith, JA, Schwiening, CJ and Huang, CL-H (2005). The effect of intracellular acidification on the relationship between cell volume and membrane potential in amphibian skeletal muscle. J Physiol 563, 745764.Google Scholar
Fraser, JA, Skepper, JN, Hockaday, AR and Huang, CL-H (1998). The tubular vacuolation process in amphibian skeletal muscle. J Muscle Res Cell Motil 19, 613629.Google Scholar
Frommeyer, G and Eckardt, L (2016). Drug-induced proarrhythmia: risk factors and electrophysiological mechanisms. Nat Rev Cardiol 13, 3647.Google Scholar
Gattuso, JM, Davies, AH, Glasby, MA, Gschmeissner, SE and Huang, CL-H (1988). Peripheral nerve repair using muscle autografts: recovery of transmission in primates. J Bone Joint Surg Br 70, 524529.Google Scholar
Glasby, MA, Gschmeissner, S, Hitchcock, R and Huang, CL-H (1986a). Regeneration of the sciatic nerve in rats: the effect of muscle basement membrane. J Bone Joint Surg Br 68, 829833.Google Scholar
Glasby, MA, Gschmeissner, SG, Hitchcock, RJI and Huang, CL-H (1986b). The dependence of nerve regeneration through muscle grafts in the rat on the availability and orientation of basement membrane. J Neurocytol 15, 479510.Google Scholar
Goddard, CA, Ghais, NS, Zhang, Y, et al. (2008). Physiological consequences of the P2328S mutation in the ryanodine receptor (RyR2) gene in genetically modified murine hearts. Acta Physiol 194, 123140.Google Scholar
Goldman, DE (1943). Potential, impedance and rectification in membranes. J Gen Physiol 27, 3760.Google Scholar
Gordon, AM, Huxley, AF and Julian, FJ (1966). The variation in isometric tension with sarcomere length in vertebrate muscle fibres. J Physiol 184, 170192.Google Scholar
Grabner, M, Dirksen, RT, Suda, N and Beam, KG (1999). The II-III loop of the skeletal muscle dihydropyridine receptor is responsible for the bidirectional coupling with the ryanodine receptor. J Biol Chem 274, 2191321919.Google Scholar
Granzier, H and Labeit, S (2002). Cardiac titin: an adjustable multi-functional spring. J Physiol 541, 335342.Google Scholar
Granzier, H and Labeit, S (2007). Structure-function relations of the giant elastic protein titin in striated and smooth muscle cells. Muscle and Nerve 36, 740755.Google Scholar
Greeff, NG, Keynes, RD and Van Helden, DF (1982). Fractionation of the asymmetry current in the squid giant axon into inactivating and non-inactivating components. Proc R Soc London - Biol Sci 215, 375389.Google Scholar
Guibert, C, Ducret, T and Savineau, JP (2008). Voltage-independent calcium influx in smooth muscle. Prog Biophys Mol Biol 98, 1023.Google Scholar
Gulbis, JM and Doyle, DA (2004). Potassium channel structures: do they conform? Curr Opin Struct Biol 14, 440446.Google Scholar
Gundersen, K (2011). Excitation-transcription coupling in skeletal muscle: the molecular pathways of exercise. Biol Rev 86, 564600.Google Scholar
Gurung, I, Medina-Gomez, G, Kis, A, et al. (2011). Deletion of the metabolic transcriptional coactivator PGC1-β induces cardiac arrhythmia. Cardiovasc Res 92, 2938.Google Scholar
Gussak, I and Antzelevich, C (2003). Cardiac Repolarization: Bridging Basic and Clinical Science. Totowa, NJ: Humana Press, Inc.CrossRefGoogle Scholar
Guzadhur, L, Jiang, W, Pearcey, SM, et al. (2012). The age-dependence of atrial arrhythmogenicity in Scn5a+/- murine hearts reflects alterations in action potential propagation and recovery. Clin Exp Pharmacol Physiol 39, 518527.Google Scholar
Hakim, P, Gurung, IS, Pedersen, TH, et al. (2008). Scn3b knockout mice exhibit abnormal ventricular electrophysiological properties. Prog Biophys Mol Biol 98, 251266.Google Scholar
Hamill, OP, Marty, A, Neher, E, Sakmann, B and Sigworth, FJ (1981). Improved patch-clamp techniques for high-resolution current recording from cells and cell-free membrane patches. Pflügers Arch 391, 85100.Google Scholar
Head, CE, Balasubramaniam, R, Thomas, G, et al. (2005). Paced electrogram fractionation analysis of arrhythmogenic tendency in DeltaKPQ Scn5a mice. J Cardiovasc Electrophysiol 16, 13291340.Google Scholar
Heatwole, CR and Moxley, RT 3rd (2007). The nondystrophic myotonias. Neurotherapeutics 4, 238251.Google Scholar
Hernandez-Ochoa, EO, Pratt, SJP, Lovering, RM and Schneider, MF (2016). Critical role of intracellular RyR1 calcium release channels in skeletal muscle function and disease. Front Physiol 6, 111.Google Scholar
Hibino, H, Inanobe, A, Furutani, K, et al. (2010). Inwardly rectifying potassium channels: their structure, function, and physiological roles. Physiol Rev 90, 291366.Google Scholar
Hill, AV (1938). The heat of shortening and the dynamic constants of muscle. Proc R Soc London - Biol Sci 126, 136195.Google Scholar
Hill, AV (1950a). The dimensions of animals and their muscular dynamics. Sci Prog Lond 38, 209230.Google Scholar
Hill, AV (1950b). A challenge to biochemists. Biochim Biophys Acta 4, 411.Google Scholar
Hill, AV and Hartree, W (1920). The four phases of heat production of muscle. J Physiol 54, 84128.Google Scholar
Hille, B (1971). The hydration of sodium ions crossing the nerve membrane. Proc Natl Acad Sci 68, 280282.Google Scholar
Hirst, GDS and Edwards, FR (2001). Generation of slow waves in the antral region of guinea-pig stomach - a stochastic process. J Physiol 535, 165180.Google Scholar
Hodgkin, AL (1939). The relation between conduction velocity and the electrical resistance outside a nerve fibre. J Physiol 94, 560570.Google Scholar
Hodgkin, AL (1951). The ionic basis of electrical activity in nerve and muscle. Biol Rev 26, 339409.CrossRefGoogle Scholar
Hodgkin, AL (1958). The Croonian Lecture: ionic movements and electrical activity in giant nerve fibres. Proc R Soc London - Biol Sci 148, 137.Google Scholar
Hodgkin, AL and Horowicz, P (1957). The differential action of hypertonic solutions on the twitch and action potential of a muscle fibre. J Physiol 137, 17P18P.Google Scholar
Hodgkin, AL and Horowicz, P (1959). The influence of potassium and chloride ions on the membrane potential of single muscle fibres. J Physiol 148, 127160.Google Scholar
Hodgkin, AL and Huxley, AF (1952). A quantitative description of membrane current and its application to conduction and excitation in nerves. J Physiol 117, 500544.Google Scholar
Hodgkin, AL and Katz, B (1949). The effect of sodium ions on the electrical activity of the giant axon of the squid. J Physiol 108, 3777.Google Scholar
Hodgkin, AL and Keynes, RD (1955a). The potassium permeability of a giant nerve fibre. J Physiol 128, 6188.Google Scholar
Hodgkin, AL and Keynes, RD (1955b). Active transport of cations in giant axons from Sepia and Loligo. J Physiol 128, 2860.Google Scholar
Hodgkin, AL, Huxley, AF and Katz, B (1949). The effect of sodium ions on the electrical activity of the giant axon of the squid. J Physiol 108, 3777.Google Scholar
Hodgkin, AL, Huxley, AF and Katz, B (1952). Measurement of current-voltage relations in the membrane of the giant axon of Loligo . J Physiol 116, 424448.Google Scholar
Hoffman, B and Cranefield, P (1960). Electrophysiology of the Heart. New York: McGrawHill.Google Scholar
Hollingworth, S and Marshall, MW (1981). A comparative study of charge movement in rat and frog skeletal muscle fibres. J Physiol 321, 583602.Google Scholar
Hollingworth, S, Peet, J, Chandler, WK and Baylor, SM (2002). Calcium sparks in intact skeletal muscle fibers of the frog. J Gen Physiol 118, 653678.Google Scholar
Homsher, E (1987). Muscle enthalpy production and its relationship to actomyosin ATPase. Annu Rev Physiol 49, 673690.CrossRefGoogle ScholarPubMed
Hothi, SS, Gurung, IS, Heathcote, JC, et al. (2008). Epac activation, altered calcium homeostasis and ventricular arrhythmogenesis in the murine heart. Pflügers Arch 457, 253270.Google Scholar
Hothi, SS, Thomas, G, Killeen, MJ, Grace, AA and Huang, CL-H (2009). Empirical correlation of triggered activity and spatial and temporal re-entrant substrates with arrhythmogenicity in a murine model for Jervell and Lange-Nielsen syndrome. Pflügers Arch 458, 819835.Google Scholar
Huang, CL-H (1982). Pharmacological separation of charge movement components in frog skeletal muscle. J Physiol 324, 375387.Google Scholar
Huang, CL-H (1984). Analysis of “off” tails of intramembrane charge movements in skeletal muscle of Rana temporaria. J Physiol 356, 375390.Google Scholar
Huang, CL-H (1988). Intramembrane charge movements in skeletal muscle. Physiol Rev 68, 11971247.Google Scholar
Huang, CL-H (1990). Voltage-dependent block of charge movement components by nifedipine in frog skeletal muscle. J Gen Physiol 96, 535557.Google Scholar
Huang, CL-H (1993). Intramembrane Charge Movements in Striated Muscle. Monographs of the Physiological Society, No. 44. Oxford: Clarendon Press.Google Scholar
Huang, CL-H (1994). Charge conservation in intact frog skeletal muscle fibres in gluconate-containing solutions. J Physiol 474, 161171.Google Scholar
Huang, CL-H (1996). Kinetic isoforms of intramembrane charge in intact amphibian striated muscle. J Gen Physiol 107, 515534.Google Scholar
Huang, CL-H (1998a). The influence of caffeine on intramembrane charge movements in intact frog striated muscle. J Physiol 512, 707721.Google Scholar
Huang, CL-H (1998b). The influence of perchlorate ions on complex charging transients in amphibian striated muscle. J Physiol 506, 699714.Google Scholar
Huang, CL-H (2017). Murine electrophysiological models of cardiac arrhythmogenesis. Physiol Rev 97, 283409.Google Scholar
Huang, CL-H (2018). Roger Yonchien Tsien. 1 February 1952 – 24 August 2016. Biogr Mem Fellows R Soc 65, 405428.Google Scholar
Huang, CL-H and Hockaday, AR (1988). Development of myotomal cells in Xenopus laevis larvae. J Anat 159, 129136.Google Scholar
Huang, CL-H and Peachey, LD (1989). Anatomical distribution of voltage-dependent membrane capacitance in frog skeletal muscle fibers. J Gen Physiol 93, 565584.Google Scholar
Huang, CL-H, Pedersen, TH and Fraser, JA (2011). Reciprocal dihydropyridine and ryanodine receptor interactions in skeletal muscle activation. J Muscle Res Cell Motil 32, 171202.Google Scholar
Huang, CL-H, Wu, L, Jeevaratnam, K and Lei, M (2020). Update on antiarrhythmic drug pharmacology. J Cardiovasc Electrophysiol 31, 579592.Google Scholar
Huxley, AF and Niedergerke, R (1954). Structural changes in muscle during contraction: interference microscopy of living muscle fibres. Nature 173, 971973.Google Scholar
Huxley, AF and Stämpfli, R (1949). Evidence for saltatory conduction in peripheral myelinated nerve fibres. J Physiol 108, 315339.Google Scholar
Huxley, AF and Taylor, RE (1958). Local activation of striated muscle fibres. J Physiol 144, 426441.Google Scholar
Huxley, HE (1963). Electron microscope studies on the structure of natural and synthetic protein filaments from striated muscle. J Mol Biol 7, 281308.Google Scholar
Huxley, HE (1976). The structural basis of contraction and regulation in skeletal muscle In: Molecular Basis of Motility. (Eds.). Heilmeyer, L Jr, Ruegg, J and Wieland, T, Berlin: Springer Verlag. 925.Google Scholar
Huxley, HE (1990). Sliding filaments and molecular motile systems. J Biol Chem 265, 83478350.Google Scholar
Huxley, HE and Hanson, J (1954). Changes in the cross-striations of muscle during contraction and stretch and their structural interpretation. Nature 173, 973976.Google Scholar
Jack, J, Noble, D and Tsien, R (1983). Electric Current Flow in Excitable Cells. Oxford: Oxford University Press.Google Scholar
James, MF, Smith, MI, Bockhorst, KHJ, et al. (1999). Cortical spreading depression in the gyrencephalic feline brain studied by magnetic resonance imaging. J Physiol 519, 415425.Google Scholar
January, CT and Riddle, JM (1989). Early afterdepolarizations: mechanism of induction and block. A role for L-type Ca2+ current. Circ Res 64, 977990.Google Scholar
Jeevaratnam, K, Guzadhur, L, Goh, YM, Grace, AA and Huang, CL-H (2016). Sodium channel haploinsufficiency and structural change in ventricular arrhythmogenesis. Acta Physiol 216, 186202.Google Scholar
Jeevaratnam, K, Rewbury, R, Zhang, Y, et al. (2012). Frequency distribution analysis of activation times and regional fibrosis in murine Scn5a+/- hearts: the effects of ageing and sex. Mech Ageing Dev 133, 591599.Google Scholar
Johnston, L, Sergeant, GP, Hollywood, MA, Thornbury, KD and McHale, NG (2005). Calcium oscillations in interstitial cells of the rabbit urethra. J Physiol 565, 449461.Google Scholar
Jurkat-Rott, K, Holzherr, B, Fauler, M and Lehmann-Horn, F (2010). Sodium channelopathies of skeletal muscle result from gain or loss of function. Pflügers Arch 460, 239248.Google Scholar
Karbat, I and Reuveny, E (2019). Voltage sensing comes to rest. Cell 178, 776778.Google Scholar
Kato, G (1936). On the excitation, conduction, and narcotization of single nerve fibers. Cold Spr Harb Symp Quant Biol 4, 202213.Google Scholar
Katz, B (1962). The Croonian Lecture - the transmission of impulses from nerve to muscle, and the subcellular unit of synaptic action. Proc R Soc London - Biol Sci 155, 455477.Google Scholar
Katz, B and Miledi, R (1969). Spontaneous and evoked activity of motor nerve endings in calcium Ringer. J Physiol 203, 689706.Google Scholar
Keating, MT and Sanguinetti, MC (2001). Molecular and cellular mechanisms of cardiac arrhythmias. Cell 104, 569580.Google Scholar
Keynes, RD (1951). The ionic movements during nervous activity. J Physiol 114, 119150.Google Scholar
Keynes, RD (1963). Chloride in the squid giant axon. J Physiol 169, 690705.Google Scholar
Keynes, RD and Elinder, F (1998). On the slowly rising phase of the sodium gating current in the squid giant axon. Proc R Soc London - Biol Sci 265, 255262.Google Scholar
Keynes, RD and Elinder, F (1999). The screw-helical voltage gating of ion channels. Proc R Soc London - Biol Sci 266, 843852.Google Scholar
Keynes, RD and Kimura, JE (1983). Kinetics of activation of the sodium conductance in the squid giant axon. J Physiol 336, 621634.Google Scholar
Keynes, RD and Lewis, PR (1951). The sodium and potassium content of cephalopod nerve fibers. J Physiol 114, 151182.Google Scholar
Keynes, RD and Martins-Ferreira, H (1953). Membrane potentials in the electroplates of the electric eel. J Physiol 119, 315351.Google Scholar
Keynes, RD and Ritchie, JM (1984). On the binding of labelled saxitoxin to the squid giant axon. Proc R Soc London - Biol Sci 222, 147153.Google Scholar
Keynes, RD and Rojas, E (1973). Characteristics of the sodium gating current in the squid giant axon. J Physiol 233, 28P30P.Google Scholar
Keynes, RD and Rojas, E (1974). Kinetics and steady‐state properties of the charged system controlling sodium conductance in the squid giant axon. J Physiol 239, 393434.Google Scholar
Keynes, RJ, Hopkins, WG and Huang, CL-H (1984). Regeneration of mouse peripheral nerves in degenerating skeletal muscle: guidance by residual muscle fibre basement membrane. Brain Res 295, 275282.Google Scholar
Killeen, MJ, Gurung, IS, Thomas, G, et al. (2007a). Separation of early afterdepolarizations from arrhythmogenic substrate in the isolated perfused hypokalaemic murine heart through modifiers of calcium homeostasis. Acta Physiol 191, 4358.Google Scholar
Killeen, MJ, Sabir, IN, Grace, AA and Huang, CL-H (2008a). Dispersions of repolarization and ventricular arrhythmogenesis: lessons from animal models. Prog Biophys Mol Biol 98, 219229.Google Scholar
Killeen, MJ, Thomas, G, Gurung, IS, et al. (2007b). Arrhythmogenic mechanisms in the isolated perfused hypokalaemic murine heart. Acta Physiol 189, 3346.Google Scholar
Killeen, MJ, Thomas, G, Sabir, IN, Grace, AA and Huang, CL-H (2008b). Mouse models of human arrhythmia syndromes. Acta Physiol 192, 455469.Google Scholar
Kim, YC, Koh, SD and Sanders, KM (2002). Voltage-dependent inward currents of interstitial cells of Cajal from murine colon and small intestine. J Physiol 541, 797810.Google Scholar
King, J, Huang, CL-H and Fraser, JA (2013a). Determinants of myocardial conduction velocity: implications for arrhythmogenesis. Front Physiol 4, 154.Google Scholar
King, J, Wickramarachchi, C, Kua, K, et al. (2013b). Loss of Nav1.5 expression and function in murine atria containing the RyR2-P2328S gain-of-function mutation. Cardiovasc Res 99, 751759.Google Scholar
King, J, Zhang, Y, Lei, M, et al. (2013c). Atrial arrhythmia, triggering events and conduction abnormalities in isolated murine RyR2-P2328S hearts. Acta Physiol 207, 308323.Google Scholar
Koeppen, B and Stanton, B (2009). Berne and Levy: Principles of Physiology 6/e. New York: Mosby.Google Scholar
Kovács, L, Ríos, E and Schneider, MF (1979). Calcium transients and intramembrane charge movement in skeletal muscle fibres. Nature 279, 391396.Google Scholar
Kovacs, L, Rios, E and Schneider, MF (1983). Measurement and modification of free calcium transients in frog skeletal muscle fibres by a metallochromic indicator dye. J Physiol 343, 161196.Google Scholar
Kuffler, S (1980). Slow synaptic responses in autonomic ganglia and the pursuit of a peptidergic transmitter. J Exp Biol 89, 257286.Google Scholar
Kyte, J and Doolittle, RF (1982). A simple method for displaying the hydropathic character of a protein. J Mol Biol 157, 105132.Google Scholar
Lammers, WJEP (2015). Normal and abnormal electrical propagation in the small intestine. Acta Physiol 213, 349359.Google Scholar
Lammers, WJEP, Al-Bloushi, HM, Al-Eisaei, SA, et al. (2011). Slow wave propagation and plasticity of interstitial cells of Cajal in the small intestine of diabetic rats. Exp Physiol 96, 10391048.Google Scholar
Laver, DR (2018). Regulation of the RyR channel gating by Ca2+ and Mg2+. Biophys Rev 10, 10871095.Google 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, 387400.Google Scholar
Lei, M, Wu, L, Terrar, DA and Huang, CL-H (2018). Modernized classification of cardiac antiarrhythmic drugs. Circulation 138, 18791896.Google Scholar
Lei, M, Zhang, H, Grace, AA and Huang, CL-H (2007). SCN5A and sinoatrial node pacemaker function. Cardiovasc Res 74, 356365.Google Scholar
Lichtman, JW, Magrassi, L and Purves, D (1987). Visualization of neuromuscular junctions over periods of several months in living mice. J Neurosci 7, 12151222.Google Scholar
Lillie, RS (1925). Factors affecting transmission and recovery in the passive iron nerve model. J Gen Physiol 7, 473507.Google Scholar
Lin, J, Handschin, C and Spiegelman, BM (2005). Metabolic control through the PGC-1 family of transcription coactivators. Cell Metab 1, 361370.Google Scholar
Liu, M, Sanyal, S, Gao, G, et al. (2009). Cardiac Na+ current regulation by pyridine nucleotides. Circ Res 105, 737745.Google Scholar
Loewi, O (1921). Über humorale Übertragbarkeit der Herzner-venwirkung. Pflügers Arch 189, 239242.Google Scholar
London, B (2001). Cardiac arrhythmias: from (transgenic) mice to men. J Cardiovasc Electrophysiol 12, 10891091.Google Scholar
London, B, Baker, LC, Petkova-Kirova, P, et al. (2007). Dispersion of repolarization and refractoriness are determinants of arrhythmia phenotype in transgenic mice with long QT. J Physiol 578, 115129.Google Scholar
Lu, Y, Mahaut-Smith, MP, Huang, CL-H and Vandenberg, JI (2003). Mutant MiRP1 subunits modulate HERG K+ channel gating: a mechanism for pro-arrhythmia in long QT syndrome type 6. J Physiol 551, 253262.Google Scholar
Lu, Y, Mahaut-Smith, MP, Varghese, A, et al. (2001). Effects of premature stimulation on HERG K+ channels. J Physiol 537, 843851.Google Scholar
Lukas, A and Antzelevitch, C (1996). Phase 2 reentry as a mechanism of initiation of circus movement reentry in canine epicardium exposed to simulated ischemia. Cardiovasc Res 32, 593603.Google Scholar
Lüttgau, HC and Spiecker, W (1979). The effects of calcium deprivation upon mechanical and electrophysiological parameters in skeletal muscle fibres of the frog. J Physiol 296, 411429.Google Scholar
MacLennan, DH (2000). Ca2+ signalling and muscle disease. Eur J Biochem 267, 52915297.Google Scholar
Magleby, KL and Stevens, CF (1972). A quantitative description of end-plate currents. J Physiol 223, 173197.Google Scholar
Makowski, L, Caspar, DL, Phillips, WC, Baker, TS and Goodenough, DA (1984). Gap junction structures. VI. Variation and conservation in connexon conformation and packing. Biophys J 45, 208218.Google Scholar
Mangoni, ME and Nargeot, J (2008). Genesis and regulation of the heart automaticity. Physiol Rev 88, 919982.Google Scholar
Marmont, G (1949). Studies on the axon membrane. I. A new method. J Cell Comp Physiol 34, 351382.Google Scholar
Martin, CA, Grace, AA and Huang, CL-H (2011a). Spatial and temporal heterogeneities are localized to the right ventricular outflow tract in a heterozygotic Scn5a mouse model. Am J Physiol Heart Circ Physiol 300, H605H616.Google Scholar
Martin, CA, Guzadhur, L, Grace, AA, Lei, M and Huang, CL-H (2011b). Mapping of reentrant spontaneous polymorphic ventricular tachycardia in a Scn5a+/- mouse model. Am J Physiol Heart Circ Physiol 300, H1853H1862.Google Scholar
Martin, CA, Huang, CL-H and Grace, AA (2010). Progressive conduction diseases. Card Electrophysiol Clin North Am 2, 509519.Google Scholar
Martin, CA, Siedlecka, U, Kemmerich, K, et al. (2012). Reduced Na+ and higher K+ channel expression and function contribute to right ventricular origin of arrhythmias in Scn5a+/- mice. Open Biol 2, 120072.Google Scholar
Martins-Ferreira, H, Nedergaard, M and Nicholson, C (2000). Perspectives on spreading depression. Brain Res Rev 32, 215234.Google Scholar
Matthews, GDK, Guzadhur, L, Grace, AA and Huang, CL-H (2012). Nonlinearity between action potential alternans and restitution, which both predict ventricular arrhythmic properties in Scn5a+/- and wild-type murine hearts. J Appl Physiol 112, 18471863.Google Scholar
Matthews, GDK, Guzadhur, L, Sabir, IN, Grace, AA and Huang, CL-H (2013). Action potential wavelength restitution predicts alternans and arrhythmia in murine Scn5a+/- hearts. J Physiol 591, 41674188.Google Scholar
Matthews, GDK, Huang, CL-H, Sun, L and Zaidi, M (2011). Translational musculoskeletal science: is sarcopenia the next clinical target after osteoporosis? Ann N Y Acad Sci 1237, 95105.Google Scholar
Matthews, GDK, Martin, CA, Grace, AA, Zhang, Y and Huang, CL-H (2010). Regional variations in action potential alternans in isolated murine Scn5a+/- hearts during dynamic pacing. Acta Physiol 200, 129146.Google Scholar
Matthews, HR, Tan, SRX, Shoesmith, JA, et al. (2019). Sodium current inhibition following stimulation of exchange protein directly activated by cyclic-3′,5′-adenosine monophosphate (Epac) in murine skeletal muscle. Sci Rep 9, 1927.Google Scholar
Maylie, J, Irving, M, Sizto, N and Chandler, WK (1987). Calcium signals recorded from cut frog twitch fibers containing antipyrylazo III. J Gen Physiol 89, 83143.Google Scholar
McHale, NG, Hollywood, M, Sergeant, G and Thornbury, K (2006). Origin of spontaneous rhythmicity in smooth muscle. J Physiol 570, 2328.Google Scholar
Mickelson, JR and Louis, CF (1996). Malignant hyperthermia: excitation-contraction coupling, Ca2+ release channel, and cell Ca2+ regulation defects. Physiol Rev 76, 537592.Google Scholar
Miller, DJ (2004). Sydney Ringer; physiological saline, calcium and the contraction of the heart. J Physiol 555, 585587.Google Scholar
Miyawaki, A, Llopis, J, Heim, R, et al. (1997). Fluorescent indicators for Ca2+ based on green fluorescent proteins and calmodulin. Nature 388, 882887.Google Scholar
Morad, M and Orkand, RK (1971). Excitation-concentration coupling in frog ventricle: evidence from voltage clamp studies. J Physiol 219, 167–89.Google Scholar
Mulroy, AD, Glasby, MA and Huang, CL-H (1990). Regeneration of unmyelinated nerve fibers through skeletal muscle autografts in rat sciatic nerve. Neuro-Orthopedics 10, 114.Google Scholar
Namadurai, S, Yereddi, NR, Cusdin, FS, et al.(2015). A new look at sodium channel β subunits. Open Biol 5, 140192.Google Scholar
Neher, E and Sakmann, B (1976). Single-channel currents recorded from membrane of denervated frog muscle fibres. Nature 260, 799802.Google Scholar
Nicholls, CG, Fuchs, PA, Martin, AR and Wallace, B (2001). From Neuron to Brain: A Cellular Approach to the Function of the Nervous System 3/e. Sunderland, USA: Sinauer Associates Inc.Google Scholar
Nielsen, OB, De Paoli, F and Overgaard, K (2001). Protective effects of lactic acid on force production in rat skeletal muscle. J Physiol 536, 161166.Google Scholar
Ning, F, Luo, L, Ahmad, S, et al. (2016). The RyR2-P2328S mutation downregulates Nav1.5 producing arrhythmic substrate in murine ventricles. Pflügers Arch 468, 655665.Google Scholar
Noble, D (1979). The Initiation of the Heartbeat 2/e. Oxford: Oxford University Press.Google Scholar
Noda, M, Shimizu, S, Tanabe, T, et al. (1984). Primary structure of electrophorus electricus sodium channel deduced from cDNA sequence. Nature 312, 121127.Google Scholar
Noda, M, Takahashi, H, Tanabe, T, et al. (1982). Primary structure of alpha-subunit precursor of Torpedo californica acetylcholine receptor deduced from cDNA sequence. Nature 299, 793797.Google Scholar
Nuyens, D, Stengl, M, Dugarmaa, S, et al. (2001). Abrupt rate accelerations or premature beats cause life-threatening arrhythmias in mice with long-QT3 syndrome. Nat Med 7, 10211027.Google Scholar
O’Malley, HA and Isom, LL (2015). Sodium channel β subunits: emerging targets in channelopathies. Annu Rev Physiol 77, 481504.Google Scholar
Offer, G (1974). The molecular basis of muscular contraction In: Companion to Biochemistry. (Eds.). Bull, A, Lagnado, J, Thomas, J and Tipton, K, London: Longman. 623671.Google Scholar
Ördög, T (2008). Interstitial cells of Cajal in diabetic gastroenteropathy. Neurogastroenterol Motil 20, 818.Google Scholar
Ostrowski, J, Kjelsberg, MA, Caron, MG and Lefkowitz, RJ (1992). Mutagenesis of the beta 2-adrenergic receptor: how structure elucidates function. Annu Rev Pharmacol Toxicol 32, 167183.Google Scholar
Ozaki, H, Gerthoffer, WT, Publicover, NG, Fusetani, N and Sanders, KM (1991a). Time-dependent changes in Ca2+ sensitivity during phasic contraction of canine antral smooth muscle. J Physiol 440, 207224.Google Scholar
Ozaki, H, Stevens, RJ, Blondfield, DP, Publicover, NG and Sanders, KM (1991b). Simultaneous measurement of membrane potential, cytosolic Ca2+, and tension in intact smooth muscles. Am J Physiol Physiol 260, C917C925.Google Scholar
Padmanabhan, N and Huang, CL-H (1990). Separation of tubular electrical activity in amphibian skeletal muscle through temperature change. Exp Physiol 75, 721724.Google Scholar
Pan, X, Li, Z, Zhou, Q, et al. (2018). Structure of the human voltage-gated sodium channel Nav1.4 in complex with β1. Science 362, eaau2486.Google Scholar
Park, KJ, Hennig, GW, Lee, H-T, et al. (2005). Spatial and temporal mapping of pacemaker activity in interstitial cells of Cajal in mouse ileum in situ. Am J Physiol Cell Physiol 290, C1411C1427.Google Scholar
Peachey, LD (1965). The sarcoplasmic reticulum and transverse tubules of the frog’s sartorius. J Cell Biol 25, 209231.Google Scholar
Peachey, LD and Eisenberg, BR (1978). Helicoids in the T system and striations of frog skeletal muscle fibers seen by high voltage electron microscopy. BiophysJ 22, 145154.Google Scholar
Pedersen, TH, Huang, CL-H and Fraser, JA (2011). An analysis of the relationships between subthreshold electrical properties and excitability in skeletal muscle. J Gen Physiol 138, 7393.CrossRefGoogle ScholarPubMed
Pedersen, TH, Macdonald, WA, de Paoli, FV, Gurung, IS and Nielsen, OB (2009). Comparison of regulated passive membrane conductance in action potential-firing fast- and slow-twitch muscle. J Gen Physiol 134, 525525.Google Scholar
Pedersen, TH, Nielsen, OB, Lamb, GD and Stephenson, DG (2004). Intracellular acidosis enhances the excitability of working muscle. Science 305, 11441147.Google Scholar
Porter, KR and Palade, GE (1957). Studies on the endoplasmic reticulum: III. Its form and distribution in striated muscle cells. J Cell Biol 3, 269300.Google Scholar
Powell, T, Terrar, DA and Twist, VW (1980). Electrical properties of individual cells isolated from adult rat ventricular myocardium. J Physiol 302, 131153.Google Scholar
Putney, JW, Broad, LM, Braun, FJ, Lievremont, JP and Bird, GS (2001). Mechanisms of capacitative calcium entry. J Cell Sci 114, 22232229.Google Scholar
Querol, L and Illa, I (2013). Myasthenia gravis and the neuromuscular junction. Curr Opin Neurol 26, 459465.Google Scholar
Rayment, I and Holden, HM (1994). The three-dimensional structure of a molecular motor. Trends Biochem Sci 19, 129134.Google Scholar
Rayment, I (1996). The active site of myosin. Annu Rev Physiol 58, 671702.Google Scholar
Rios, E and Brum, G (1987). Involvement of dihydropyridine receptors in excitation-contraction coupling in skeletal muscle. Nature 325, 717720.Google Scholar
Ritchie, JM and Rogart, RB (1977). The binding of saxitoxin and tetrodotoxin to excitable tissue. Rev Physiol Biochem Pharmacol 79, 150.Google Scholar
Robertson, D (1960). The molecular structure and contact relationships of cell membranes. Prog Biophys Mol Biol 10, 343418.Google Scholar
Roden, D (2004). Drug-induced prolongation of the QT interval. N Engl J Med 350, 10131022.Google Scholar
Rushton, WAH (1933). Lapicque’s theory of curarization. J Physiol 77, 337364.Google Scholar
Ryall, RW (1979). Mechanisms of Drug Action on the Nervous System. Cambridge: Cambridge University Press.Google Scholar
Saadeh, K, Achercouk, Z, Fazmin, IT, et al. (2020). Protein expression profiles in murine ventricles modeling catecholaminergic polymorphic ventricular tachycardia: effects of genotype and sex. Ann NY Acad Sci. DOI: 10.1111/nyas.14426.Google Scholar
Sabir, IN, Fraser, JA, Cass, TR, Grace, AA and Huang, CL-H (2007). A quantitative analysis of the effect of cycle length on arrhythmogenicity in hypokalaemic Langendorff-perfused murine hearts. Pflügers Arch 454, 925936.Google Scholar
Sabir, IN, Killeen, MJ, Grace, AA and Huang, CL-H (2008a). Ventricular arrhythmogenesis: insights from murine models. Prog Biophys Mol Biol 98, 208218.Google Scholar
Sabir, IN, Li, LM, Grace, AA and Huang, CL-H (2008b). Restitution analysis of alternans and its relationship to arrhythmogenicity in hypokalaemic Langendorff-perfused murine hearts. Pflügers Arch 455, 653666.Google Scholar
Sabir, IN, Ma, N, Jones, VJ, et al. (2010). Alternans in genetically modified Langendorff-perfused murine hearts modeling catecholaminergic polymorphic ventricular tachycardia. Front Physiol 1, 126.Google Scholar
Salvage, SC, Gallant, E, Beard, N, et al. (2019a). Ion channel gating in cardiac ryanodine receptors from the arrhythmic RyR2-P2328S mouse. J Cell Sci 132, jcs229039.Google Scholar
Salvage, SC, Huang, CL-H, and Jackson, AP (2020a). Cell-adhesion properties of auxiliary β-subunits in the regulation of cardiomyocyte sodium channels. Special Issue: Trafficking of Cardiac Ion Channels ? Mechanisms and Alterations Leading to Disease. Ed. M. Verges. Biomolecules 10, 989.Google Scholar
Salvage, SC, King, JH, Chandrasekharan, KH, et al. (2015). Flecainide exerts paradoxical effects on sodium currents and atrial arrhythmia in murine RyR2-P2328S hearts. Acta Physiol 214, 361375.Google Scholar
Salvage, SC, Rees, J, McStea, A, et al. (2020b). Supramolecular clustering of the cardiac sodium channel Nav1.5 in HEK293F cells, with and without the auxiliary β3-subunit. FASEB J 34, 35373553.Google Scholar
Salvage, SC, Zhu, W, Habib, Z, et al. (2019b). Gating control of the cardiac sodium channel Nav1.5 by its β3-subunit: distinct roles for a transmembrane glutamic acid and the extracellular domain. J Biol Chem 94, 1975219763Google Scholar
Samsó, M (2017). A guide to the 3D structure of the ryanodine receptor type 1 by cryoEM. Protein Sci 26, 5268.Google Scholar
Sanchez, JA and Stefani, E (1978). Inward calcium current in twitch muscle fibres of the frog. J Physiol 283, 197209.Google Scholar
Sanders, KM, Hwang, SJ and Ward, SM (2010). Neuroeffector apparatus in gastrointestinal smooth muscle organs. J Physiol 588, 46214639.Google Scholar
Sanders, KM, Koh, SD and Ward, SM (2006). Interstitial cells of Cajal as pacemakers in the gastrointestinal tract. Annu Rev Physiol 68, 307343.Google Scholar
Sanguinetti, MC, Jiang, C, Curran, ME and Keating, MT (1995). A mechanistic link between an inherited and an acquired cardiac arrhythmia: HERG encodes the IKr potassium channel. Cell 81, 299307.Google Scholar
Sarbjit-Singh, S, Matthews, HR and Huang, CL-H (2020). Ryanodine receptor modulation by caffeine challenge modifies Na+ current properties in intact murine skeletal muscle fibres. Sci Rep 10, 2199.Google Scholar
Sarquella-Brugada, G, Campuzano, O, Arbelo, E, Brugada, J and Brugada, R (2016). Brugada syndrome: clinical and genetic findings. Genet Med 18, 312.Google Scholar
Scher, A (1965). Mechanical events in the cardiac cycle In: Physiology and Biophysics. (Eds.). Ruch, T and Patton, H, Philadelphia, USA: Saunders.Google Scholar
Scheuer, T and Gilly, WF (1986). Charge movement and depolarization-contraction coupling in arthropod vs. vertebrate skeletal muscle. Proc Natl Acad Sci U S A 83, 87998803.Google Scholar
Schmidt-Nielsen, K (1990). Animal Physiology 4/e. Cambridge: Cambridge University Press.Google Scholar
Schneider, MF and Chandler, WK (1973). Voltage dependent charge movement in skeletal muscle: a possible step in excitation-contraction coupling. Nature 242, 244246.Google Scholar
Schwartz, LM, McCleskey, EW and Almers, W (1985). Dihydropyridine receptors in muscle are voltage-dependent but most are not functional calcium channels. Nature 314, 747751.Google Scholar
Sejersted, OM and Sjøgaard, G (2000). Dynamics and consequences of potassium shifts in skeletal muscle and heart during exercise. Physiol Rev 80, 14111481.Google Scholar
Sheikh, SM, Skepper, JN, Chawla, S, et al. (2001). Normal conduction of surface action potentials in detubulated amphibian skeletal muscle fibres. J Physiol 535, 579590.Google Scholar
Shotton, DM, Heuser, JE, Reese, BF and Reese, TS (1979). Postsynaptic membrane folds of the frog neuromuscular junction visualized by scanning electron microscopy. Neuroscience 4, 427435.Google Scholar
Singer, SJ and Nicolson, GL (1972). The fluid mosaic model of the structure of cell membranes. Science 175, 720731.Google Scholar
Skou, J (1998). The identification of the sodium pump. Biosci Rep 24, 436451.Google Scholar
Smith, JM, Bradley, DP, James, MF and Huang, CL-H (2006). Physiological studies of cortical spreading depression. Biol Rev Camb Philos Soc 81, 457481.Google Scholar
Smith, JM, James, MF, Bockhorst, KHJ, et al. (2001). Investigation of feline brain anatomy for the detection of cortical spreading depression with magnetic resonance imaging. J Anat 198, 537554.Google Scholar
Smith, PL, Baukrowitz, T and Yellen, G (1996). The inward rectification mechanism of the HERG cardiac potassium channel. Nature 379, 833836.Google Scholar
Spudich, JA (1994). How molecular motors work. Nature 372, 515518.Google Scholar
Spudich, JA, Finer, J, Simmons, B, et al. (1995). Myosin structure and function. Cold Spr Harb Symp Quant Biol 60, 783791.Google Scholar
Squire, J (1986). Muscle: Design, Diversity and Disease. California: Benjamin/Cummings, Menlo Park.Google Scholar
Takeshima, H, Nishimura, S, Matsumoto, T, et al. (1989). Primary structure and expression from complementary DNA of skeletal muscle ryanodine receptor. Nature 339, 439445.Google Scholar
Takeuchi, A and Takeuchi, N (1959). Active phase of frog’s end-plate potential. J Neurophysiol 22, 395411.Google Scholar
Takeuchi, A and Takeuchi, N (1960). On the permeability of end‐plate membrane during the action of transmitter. J Physiol 154, 5267.Google Scholar
Takla, M, Huang, CL-H, Jeevaratnam, K (2020). The cardiac CaMKII-Nav1.5 relationship: From physiology to pathology. J Mol Cell Cardiol 139, 190200.Google Scholar
Tanabe, T, Beam, KG, Adams, BA, Niidome, T and Numa, S (1990). Regions of the skeletal muscle dihydropyridine receptor critical for excitation-contraction coupling. Nature 346, 567569.Google Scholar
Tanaka, Y, Horinouchi, T and Koike, K (2005). New insights into β-adrenoceptors in smooth muscle: distribution of receptor subtypes and molecular mechanisms triggering muscle relaxation. Clin Exp Pharmacol Physiol 32, 503514.Google Scholar
Tasaki, I (1953). Nervous Transmission. Springfield, Illinois: Charles C. Thomas.Google Scholar
Task Force of the Working Group on Arrhythmias of the European Society of Cardiology (1991). The Sicilian gambit: a new approach to the classification of antiarrhythmic drugs based on their actions on arrhythmogenic mechanisms. Circulation 84, 18311851.Google Scholar
Thomas, G, Gurung, IS, Killeen, MJ, et al. (2007a). Effects of L-type Ca2+ channel antagonism on ventricular arrhythmogenesis in murine hearts containing a modification in the Scn5a gene modelling human long QT syndrome 3. J Physiol 578, 8597.Google Scholar
Thomas, G, Killeen, MJ, Grace, AA and Huang, CL-H (2008). Pharmacological separation of early afterdepolarizations from arrhythmogenic substrate in DeltaKPQ Scn5a murine hearts modelling human long QT 3 syndrome. Acta Physiol 192, 505517.Google Scholar
Thomas, G, Killeen, MJ, Gurung, IS, et al. (2007b). Mechanisms of ventricular arrhythmogenesis in mice following targeted disruption of KCNE1 modelling long QT syndrome 5. J Physiol 578, 99114.Google Scholar
Thornton, CA (2014). Myotonic dystrophy. Neurol Clin 32, 705719.Google Scholar
Tsien, RW and Malinow, R (1991). Changes in presynaptic function during long-term potentiation. Ann NY Acad Sci 635, 208220.Google Scholar
Turner, C and Hilton-Jones, D (2014). Myotonic dystrophy. Curr Opin Neurol 27, 599606.Google Scholar
Unwin, N (2003). Structure and action of the nicotinic acetylcholine receptor explored by electron microscopy. FEBS Lett 555, 9195.Google Scholar
Unwin, N (2014). Nicotinic acetylcholine receptor and the structural basis of neuromuscular transmission: insights from Torpedo postsynaptic membranes. Q Rev Biophys 46, 283322.Google Scholar
Usher-Smith, JA, Fraser, JA, Bailey, PSJ, Griffin, JL and Huang, CL-H (2006a). The influence of intracellular lactate and H+ on cell volume in amphibian skeletal muscle. J Physiol 573, 799818.Google Scholar
Usher-Smith, JA, Huang, CL-H and Fraser, JA (2009). Control of cell volume in skeletal muscle. Biol Rev 84, 143159.Google Scholar
Usher-Smith, JA, Skepper, JN, Fraser, JA and Huang, CL-H (2006b). Effect of repetitive stimulation on cell volume and its relationship to membrane potential in amphibian skeletal muscle. Pflügers Arch 452, 231239.Google Scholar
Valli, H, Ahmad, S, Chadda, K, et al. (2017). Age-dependent atrial arrhythmic phenotype secondary to mitochondrial dysfunction in Pgc-1β deficient murine hearts. Mech Ageing Dev 167, 3045.Google Scholar
Valli, H, Ahmad, S, Jiang, AY, et al. (2018a). Cardiomyocyte ionic currents in intact young and aged murine Pgc-1β–/–atrial preparations. Mech Ageing Dev 169, 19.Google Scholar
Valli, H, Ahmad, S, Sriharan, S, et al. (2018b). Epac-induced ryanodine receptor type 2 activation inhibits sodium currents in atrial and ventricular murine cardiomyocytes. Clin Exp Pharmacol Physiol 45, 278292.Google Scholar
Vandenberg, JI, Varghese, A, Lu, Y, et al. (2006). Temperature dependence of human ether-a-go-go-related gene K+ currents. Am J Physiol Cell Physiol 291, C165C175.Google Scholar
Vandenberg, JI, Walker, BD and Campbell, TJ (2001). HERG K+ channels: friend and foe. Trends Pharmacol Sci 22, 240246.Google Scholar
Vaughan Williams, E (1975). Classification of antidysrhythmic drugs. Pharmacol Ther B 1, 115138.Google Scholar
Veeraraghavan, R, Gourdie, RG, Poelzing, S (2014). Mechanisms of cardiac conduction: a history of revisions. Am J Physiol Heart Circ Physiol 306, H619–627.Google Scholar
Veeraraghavan, R, Larsen, AP, Torres, NS, Grunnet, M and Poelzing, S (2013). Potassium channel activators differentially modulate the effect of sodium channel blockade on cardiac conduction. Acta Physiol 207, 280289.Google Scholar
Verheule, S, Sat, T, Everett, T IV, et al. (2004). Increased vulnerability to atrial fibrillation in transgenic mice with selective atrial fibrosis caused by overexpression of TGF-β1. Circ Res 94, 14581465.Google Scholar
Verschuuren, J, Strijbos, E and Vincent, A (2016). Neuromuscular junction disorders In: Handbook of Clinical Neurology. Vol 133. (Eds.). Pittock, SJ and Vincent, A, New York: Elsevier. 447466.Google Scholar
Vinogradova, TM, Lyashkov, AE, Zhu, W, et al. (2006). High basal protein kinase A-dependent phosphorylation drives rhythmic internal Ca2+ store oscillations and spontaneous beating of cardiac pacemaker cells. Circ Res 98, 505514.Google Scholar
Wang, Y, Cheng, J, Joyner, RW, Wagner, MB and Hill, JA (2006). Remodeling of early-phase repolarization: a mechanism of abnormal impulse conduction in heart failure. Circulation 113, 18491856.Google Scholar
Wang, Y, Tsui, H, Ke, Y, et al. (2014). Pak1 is required to maintain ventricular Ca2+ homeostasis and electrophysiological stability through SERCA2a regulation in mice. Circ Arrhythm Electrophysiol 7, 938948.Google Scholar
Waxman, S (2017). Clinical Neuroanatomy 28/e. New York: McGrawHill.Google Scholar
Weidmann, S (1956). Elektrophysiologie der Herzmuskelfaser. Berne: Huber.Google Scholar
Weidmann, S (1952). The electrical constants of Purkinje fibres. J Physiol 118, 348360.Google Scholar
Wilkie, DR (1968). Heat work and phosphorylcreatine break‐down in muscle. J Physiol 195, 157183.Google Scholar
Wisedchaisri, G, Tonggu, L, McCord, E, et al. (2019). Resting-state structure and gating mechanism of a voltage-gated sodium channel. Cell 178, 9931003.Google Scholar
Yan, GX and Antzelevitch, C (1998). Cellular basis for the normal T wave and the electrocardiographic manifestations of the long-QT syndrome. Circulation 98, 19281936.Google Scholar
Yan, Z, Zhou, Q, Wang, L, et al. (2017). Structure of the Nav1.4-β1 complex from electric eel. Cell 170, 470482.Google Scholar
Yarov-Yarovoy, V, DeCaen, PG, Westenbroek, RE, et al. (2012). Structural basis for gating charge movement in the voltage sensor of a sodium channel. Proc Natl Acad Sci 109, E93E102.Google Scholar
Yoneda, S, Takano, H, Takaki, M and Suzuki, H (2002). Properties of spontaneously active cells distributed in the submucosal layer of mouse proximal colon. J Physiol 542, 887897.Google Scholar
Young, J (1949). Factors influencing the regeneration of nerves. Advan Surg 1, 165220.Google Scholar
Zaleska, M, Salvage, S, Thompson, A, et al. (2018). The voltage-dependent sodium channel family In: Oxford Handbook of Neuronal Ion Channels. (Ed.). Bhattacharjee, A, Oxford: Oxford University Press.Google Scholar
Zhang, MM, Wilson, MJ, Azam, L, et al. (2013a). Co-expression of NaVβ subunits alters the kinetics of inhibition of voltage-gated sodium channels by pore-blocking μ-conotoxins. Br J Pharmacol 168, 15971610.Google Scholar
Zhang, Y, Fraser, JA, Jeevaratnam, K, et al. (2011). Acute atrial arrhythmogenicity and altered Ca2+ homeostasis in murine RyR2-P2328S hearts. Cardiovasc Res 89, 794804.Google Scholar
Zhang, Y, Guzadhur, L, Jeevaratnam, K, et al. (2014). Arrhythmic substrate, slowed propagation and increased dispersion in conduction direction in the right ventricular outflow tract of murine Scn5a+/- hearts. Acta Physiol 211, 559573.Google Scholar
Zhang, Y, Schwiening, C, Killeen, MJ, et al. (2009). Pharmacological changes in cellular Ca2+ homeostasis parallel initiation of atrial arrhythmogenesis in murine Langendorff-perfused hearts. Clin Exp Pharmacol Physiol 36, 969980.Google Scholar
Zhang, Y, Wu, J, Jeevaratnam, K, et al. (2013b). Conduction slowing contributes to spontaneous ventricular arrhythmias in intrinsically active murine RyR2-P2328S hearts. J Cardiovasc Electrophysiol 24, 210218.Google Scholar
Zheng, H, Park, KS, Koh, SD and Sanders, KM (2014). Expression and function of a T-type Ca2+ conductance in interstitial cells of Cajal of the murine small intestine. Am J Physiol Physiol 306, C705C713.Google Scholar
Zhou, J, Brum, G, González, A, et al. (2003). Ca2+ sparks and embers of mammalian muscle. Properties of the sources. J Gen Physiol 122, 95114.Google Scholar
Zhu, MH, Kim, TW, Ro, S, et al. (2009). A Ca2+-activated Cl conductance in interstitial cells of Cajal linked to slow wave currents and pacemaker activity. J Physiol 587, 49054918.Google Scholar
Zhu, MH, Sung, TS, Kurahashi, M, et al. (2016). Na+-K+-Cl cotransporter (NKCC) maintains the chloride gradient to sustain pacemaker activity in interstitial cells of Cajal.. Am J Physiol Gastrointest Liver Physiol 311, G1037G1046.Google Scholar

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.

  • References
  • Christopher L.-H. Huang, University of Cambridge
  • Book: Keynes & Aidley's Nerve and Muscle
  • Online publication: 07 November 2020
  • Chapter DOI: https://doi.org/10.1017/9781108860789.018
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.

  • References
  • Christopher L.-H. Huang, University of Cambridge
  • Book: Keynes & Aidley's Nerve and Muscle
  • Online publication: 07 November 2020
  • Chapter DOI: https://doi.org/10.1017/9781108860789.018
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.

  • References
  • Christopher L.-H. Huang, University of Cambridge
  • Book: Keynes & Aidley's Nerve and Muscle
  • Online publication: 07 November 2020
  • Chapter DOI: https://doi.org/10.1017/9781108860789.018
Available formats
×