Skip to main content Accessibility help
×
Hostname: page-component-6d856f89d9-mhpxw Total loading time: 0 Render date: 2024-07-16T06:24:49.831Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  24 May 2010

Meyer B. Jackson
Affiliation:
University of Wisconsin, Madison
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: 2006

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

Abbott, A. J. and Nelsestuen, G. L. (1988). The collisional limit: an important consideration for membrane-associated enzymes and receptors. Faseb J., 2, 2858–2866.CrossRefGoogle ScholarPubMed
Accardi, A. and Miller, C. (2004). Secondary active transport mediated by a prokaryotic homologue of ClC Cl− channels. Nature, 427, 803–807.CrossRefGoogle ScholarPubMed
Adam, G. and Delbrück, M. (1968). Reduction of dimensionality in biological diffusion processes. In Structural Chemistry and Molecular Biology, ed. Rich, A. and Davidson, N.. San Francisco: Freeman, pp. 198–215.Google Scholar
Adams, D. J., Dwyer, T. M. and Hille, B. (1980). The permeability of endplate channels to monovalent and divalent metal cations. J. Gen. Physiol., 75, 493–510.CrossRefGoogle ScholarPubMed
Adams, S. A. and DeFelice, L. J. (2002). Flux coupling in the human serotonin transporter. Biophys. J., 83, 3268–3282.CrossRefGoogle ScholarPubMed
Ahern, C. A. and Horn, R. (2004). Stirring up controversy with a voltage sensor paddle. TINS, 27, 303–307.Google ScholarPubMed
Aicken, C. C. (1990). Chloride transport across the sarcolemma of vertebrate smooth and skeletal muscle. In Chloride Channels and Carriers in Nerve, Muscle, and Glial Cells, ed. Alvarez-Leefmans, F. J. and Russell, J. M.. New York: Plenum Press, pp. 209–249.CrossRefGoogle Scholar
Aidley, D. J. (1978). The Physiology of Excitable Cells. Cambridge: Cambridge University Press.Google Scholar
Alber, T., Dao-pin, S., Wilson, K., Wozniak, J. A., Cook, S. P. and Matthews, B. W. (1987). Contributions of hydrogen bonds of Thr 157 to the thermodynamic stability of phage T4 lysozyme. Nature, 330, 41–46.CrossRefGoogle ScholarPubMed
Albery, W. J. and Knowles, J. R. (1976). Evolution of enzyme function and the development of catalytic efficiency. Biochemistry, 15, 5631–5640.CrossRefGoogle ScholarPubMed
Aldrich, R. W., Getting, P. A. and Thompson, S. H. (1979). Mechanism of frequency-dependent broadening of molluscan neurone soma spikes. J. Physiol., 291, 531–544.CrossRefGoogle ScholarPubMed
Allen, T. W., Andersen, O. S. and Roux, B. (2004). Energetics of ion conduction through the gramicidin channel. Proc. Natl Acad. Sci., 101, 117–122.CrossRefGoogle ScholarPubMed
Almers, W. and McCleskey, E. W. (1984). The non-selective conductance in calcium channels of frog muscle: calcium selectivity in a single-file pore. J. Physiol., 353, 585–608.CrossRefGoogle Scholar
Andersen, O. S. (1983). Ion movement through gramicidin A channels. Single-channel measurements at very high potentials. Biophys. J., 41, 119–133.CrossRefGoogle ScholarPubMed
Anderson, C. R. and Stevens, C. F. (1973). Voltage-clamp analysis of acetylcholine produced end plate current fluctuations at frog neuromuscular junction. J. Physiol., 235, 655–691.CrossRefGoogle ScholarPubMed
Antzelevitch, C. (2001). Basic mechanisms of reentrant arrhythmias. Curr. Opin. Cardiol., 16, 1–7.CrossRefGoogle ScholarPubMed
Åqvist, J. and Luzhkov, V. (2000). Ion permeation mechanism of the potassium channel. Nature, 404, 881–884.CrossRefGoogle ScholarPubMed
Ashcroft, F. M. (2000). Ion Channels and Disease. San Diego: Academic Press.Google Scholar
Aurora, R., Creamer, T. P., Srinivasan, R. and Rose, G. D. (1997). Local interactions in protein folding: lessons from the α-helix. J. Biol. Chem., 272, 1413–1416.CrossRefGoogle ScholarPubMed
Aveyard, R. and Haydon, D. A. (1973). An Introduction to the Principles of Surface Chemistry. Cambridge: Cambridge University Press, p. 231.Google Scholar
Axe, D. D., Foster, N. W. and Fersht, A. R. (1996). Active barnase variants with completely random hydrophobic cores. Proc. Natl Acad. Sci., 93, 5590–5594.CrossRefGoogle ScholarPubMed
Baldwin, R. L. (1996). How Hofmeister ion interactions affect protein stability. Biophys. J., 71, 2056–2063.CrossRefGoogle ScholarPubMed
Barrett, J. N. and Crill, W. E. (1974). Specific membrane properties of cat motoneurons. J. Physiol., 239, 301–324.CrossRefGoogle Scholar
Bashford, C. L. and Pasternak, C. A. (1986). Plasma membrane potential of some animal cells is generated by ion pumping, not by gradients. Trends Biochem. Sci., 11, 113–116.CrossRefGoogle Scholar
Bass, R. B., Strop, P., Barclay, M. and Rees, D. C. (2002). Crystal structure of Escherichia coli MscS, a voltage-modulated and mechanosensitive channel. Science, 298, 1582–1587.CrossRefGoogle ScholarPubMed
Bedzek, M. J., Bommarito, G. M., Caffrey, M. and Penner, T. L. (1990). Diffuse-double layer at a membrane-aqueous interface measured with X-ray standing waves. Science, 248, 52–56.CrossRefGoogle Scholar
Beece, D., Eisenstein, L., Frauenfelder, H.et al. (1980). Solvent viscosity and protein dynamics. Biochemistry, 19, 5147–5157.CrossRefGoogle ScholarPubMed
Ben-Shaul, A., Ben-Tal, N. and Honig, B. (1996). Statistical thermodynamic analysis of peptide and protein insertion into lipid membranes. Biophys. J., 71, 130–137.CrossRefGoogle ScholarPubMed
Benedek, G. B. and Villars, F. M. H. (2000). Physics with Illustrative Examples from Medicine and Biology. New York: Springer-Verlag.Google Scholar
Benz, R. and Läuger, P. (1977). Transport kinetics of dipicrylamine through lipid bilayer membranes. Biochem. Biophys. Acta, 468, 245–258.CrossRefGoogle ScholarPubMed
Berezin, I. V., Kazanskaya, N. F. and Klyosov, A. A. (1971). Determination of the individual rate constants of α-chymotrypsin-catalyzed hydrolysis with the added nucleophilic agent 1,4-butanediol. FEBS Lett., 15, 121–124.CrossRefGoogle ScholarPubMed
Berg, H. C. (1983). Random Walks in Biology. Princeton: Princeton University Press. [Cited in Chapters 6 and 8. This is an accessible and clear introduction to Brownian motion and diffusion with good biological examples.]Google Scholar
Berg, H. C. and Purcell, E. M. (1977). Physics of chemoreception. Biophys. J., 20, 193–219. [Cited in Chapter 8. A seminal paper that introduces many important concepts in diffusion-limited encounters.]CrossRefGoogle ScholarPubMed
Berg, O. G. and Hippel, P. H. (1985). Diffusion-controlled macromolecular interactions. Ann. Rev. Biophys. Biophys. Chem., 14, 131–160.CrossRefGoogle ScholarPubMed
Berniche, S. and Roux, B. (2000). Molecular dynamics of the KcsA K+ channel in a bilayer membrane. Biophys. J., 78, 2900–2917.CrossRefGoogle Scholar
Betz, S. F., Bryson, J. W. and DeGrado, W. F. (1995). Native-like and structurally characterized designed α-helical bundles. Curr. Opin. Struct. Biol., 5, 457–463.CrossRefGoogle ScholarPubMed
Bezanilla, F. (2000). The voltage sensor in voltage-dependent ion channels. Physiol. Rev., 80, 555–592.CrossRefGoogle ScholarPubMed
Blacklow, S. C., Raines, R. T., Lim, W. A., Zamore, P. D. and Knowles, J. R. (1988). Triosephosphate isomerase catalysis is diffusion controlled. Biochemistry, 27, 1158–1167.CrossRefGoogle ScholarPubMed
Blangy, D., Buc, H. and Monod, J. (1968). Kinetics of the allosteric interactions of phosphofructokinase from Escherichia coli. J. Mol. Biol., 31, 13–35. Cited in Chapter 5. This provides an example of how the MWC theory can synthesize a broad range of observations.]CrossRefGoogle ScholarPubMed
Bloomfield, S. A., Hamos, J. E. and Sherman, S. M. (1987). Passive cable properties and morphological correlates of neurones in the lateral geniculate nucleus of the cat. J. Physiol., 383, 653–692.CrossRefGoogle ScholarPubMed
Bloomfield, V. A., Crothers, D. M. and Tinoco, I. (1974). The Physical Chemistry of Nucleic Acids. New York: Harper and Row, Publishers, Inc.Google Scholar
Boyle, P. J. and Conway, E. J. (1941). Potassium accumulation in muscle and associated changes. J. Physiol., 100, 1–63.CrossRefGoogle ScholarPubMed
Breslow, E. and Gurd, F. R. N. (1962). Reactivity of sperm whale metmyoglobin towards hydrogen ions and p-nitrophenyl acetate. J. Biol. Chem., 237, 371–381.Google ScholarPubMed
Brooks, B. R., Bruccoleri, R. E., Olafson, B. D., States, D. J., Swaminathan, S. and Karplus, M. (1983). CHARMM: A program for macromolecular energy, minimization, and dynamics calculations. J. Comput. Chem., 4, 187–217.CrossRefGoogle Scholar
Brouwer, A. C. and Kirsch, J. F. (1982). Investigation of diffusion-limited rates of chymotrypsin reactions by viscosity variation. Biochemistry, 21, 1302–1307.CrossRefGoogle ScholarPubMed
Bruice, T. C. (1970). Proximity effects and enzyme catalysis. In The Enzymes II, ed. Boyer, P. D.. New York: Academic Press, pp. 217–279.Google Scholar
Bryant, R. G. (1996). The dynamics of water–protein interactions. Ann. Rev. Biophys. Biomol. Struct., 25, 29–53.CrossRefGoogle ScholarPubMed
Cai, M. and Jordan, P. C. (1990). How does vestibule surface charge affect ion conduction and toxin binding in a sodium channel. Biophys. J., 57, 883–891.CrossRefGoogle Scholar
Camacho, C. J. and Thirumalai, D. (1993). Minimum energy compact structures of random sequences of heteropolymers. Phys. Rev. Lett., 71, 2505–2508.CrossRefGoogle ScholarPubMed
Cantor, P. R. and Schimmel, P. R. (1980). Biophysical Chemistry. San Francisco: W. H. Freeman and Co. [Cited in Chapter 3. This thorough three-volume set has excellent chapters on conformational statistics of polymers and the helix–coil transition.]Google Scholar
Cardinale, G. J. and Abeles, R. H. (1968). Purification and mechanism of action of proline racemase. Biochemistry, 7, 3970–3978.CrossRefGoogle ScholarPubMed
Carra, J. H., Murphy, E. C. and Privalov, P. L. (1996). Thermodynamic effects of mutations on the denaturation of T4 lysozyme. Biophys. J., 71, 1994–2001.CrossRefGoogle ScholarPubMed
Carslaw, H. S. and Jaeger, J. C. (1959). Conduction of Heat in Solids. Oxford: Oxford University Press.Google Scholar
Cassidy, C. S., Lin, J. and Frey, P. A. (1997). A new concept for the mechanism of action of chymotrypsin: the role of the low-barrier hydrogen bond. Biochemistry, 36, 4576–4584.CrossRefGoogle ScholarPubMed
Caterall, W. A., Chandy, G. K. and Gutman, G. A. (eds.) (2002). The IUPHAR Compendium of Voltage-Gated Ion Channels. Leeds: IUPHAR Media.Google Scholar
Chan, H. S. and Dill, K. A. (1991). Polymer principles in protein structure and stability. Ann. Rev. Biophys. Biophys. Chem., 20, 447–490. [Cited in Chapter 3. This article draws many interesting connections between basic polymer theory and protein structure.]CrossRefGoogle ScholarPubMed
Chandrasekhar, S. (1943). Stochastic problems in physics and astronomy. Rev. Mod. Phys., 15, 1–89. [Cited in Chapter 6. This is a clear account of Brownian motion, and is also included in an excellent selection of related papers, edited by N. Wax and published by Dover.]CrossRefGoogle Scholar
Changeux, J. P. (1984). Acetylcholine receptor: an allosteric protein. Science, 225, 1335–1345.CrossRefGoogle ScholarPubMed
Chapman, M. L., Donger, H. M. A. and Donger, A. M. J. (1997). Activation-dependent subconductance levels in the drk 1 K channel suggest a subunit basis for ion permeation and gating. Biophys. J., 72, 708–719.CrossRefGoogle Scholar
Checover, S., Nachliel, E., Dencher, N. A. and Gutman, M. (1997). Mechanism of proton entry into the cytoplasmic section of the proton-conducting channel of bacteriorhodopsin. Biochemistry, 36, 13 919–13 928.CrossRefGoogle ScholarPubMed
Chen, Y. -D. and Hill, T. L. (1973). Fluctuations and noise in kinetic systems: Applications to K+ channels in the squid axon. Biophys. J., 13, 1276–1295.CrossRefGoogle Scholar
Chung, S. -H., Allen, T. W., Hoyles, M. and Kuyucak, S. (1999). Permeation of ions across the potassium channel: Brownian dynamics studies. Biophys. J., 77, 2517–2533. [Cited in Chapter 14. A lucid treatment of K+ channel permeation with modern computational methods.]CrossRefGoogle ScholarPubMed
Chung, S. -H., Allen, T. W. and Kuyucak, S. (2002). Conducting-state properties of the KcsA potassium channel from molecular and Brownian dynamics simulations. Biophys. J., 82, 628–645.CrossRefGoogle ScholarPubMed
Cleland, W. W. (1970). Steady state kinetics. In The Enzymes, ed. Boyer, P. D.. New York: Academic Press, Vol. ii, pp. 1–65.Google Scholar
Cleland, W. W., Frey, P. A. and Gerlt, J. A. (1998). The low barrier hydrogen bond in enzymatic catalysis. J. Biol. Chem., 273, 25 529–25 532.CrossRefGoogle ScholarPubMed
Clements, J. D. and Redman, S. J. (1989). Cable properties of cat spinal motoneurons measured by combining voltage clamp, current clamp and intracellular staining. J. Physiol., 409, 63–87.CrossRefGoogle ScholarPubMed
Cohn, E. J. and Edsall, J. T. (1943). Proteins, Amino Acids and Peptides as Ions and Dipolar Ions. New York: Reinhold Publishing Corporation.Google Scholar
Colquhoun, D. and Hawkes, A. G. (1982). On the stochastic properties of bursts of single ion channel openings and of clusters of bursts. Phil. Trans. R. Soc. Lond., B300, 1–59.CrossRefGoogle ScholarPubMed
Colquhoun, D. and Hawkes, A. G.(1995). A Q-matrix cookbook. In Single-Channel Recording, ed. Sakmann, B. and Neher, E.. New York: Plenum, pp. 589–633.Google Scholar
Connelly, P., Ghosaini, L., Hu, C. -Q., Kitamura, S., Tanaka, A. and Sturtevant, J. M. (1991). A differential scanning calorimetric study of the thermal unfolding of seven mutant forms of phage T4 lysozyme. Biochem., 30, 1887–1891.CrossRefGoogle ScholarPubMed
Connor, J. A. and Stevens, C. F. (1971a). Inward and delayed outward membrane current in isolated neural somata under voltage clamp. J. Physiol., 213, 1–19.CrossRefGoogle Scholar
Connor, J. A. and Stevens, C. F.(1971b). Prediction of repetitive firing behavior from voltage clamp data on an isolated neurone soma. J. Physiol., 213, 31–53.CrossRefGoogle Scholar
Cooper, A. (1976). Thermodynamic fluctuations in protein molecules. Proc. Natl Acad. Sci., 73, 2740–2741.CrossRefGoogle ScholarPubMed
Coronado, R., Rosenberg, R. L. and Miller, C. (1980). Ionic selectivity, saturation, and block in a K+-selective channel from sarcoplasmic reticulum. J. Gen. Physiol., 76, 425–446.CrossRefGoogle Scholar
Cox, D. R. and Miller, H. D. (1965). The Theory of Stochastic Processes. New York: Chapman and Hall.Google Scholar
Crank, J. (1975). The Mathematics of Diffusion. Oxford: Oxford University Press.Google Scholar
Crothers, D. M., Drak, J., Kahn, J. D. and Levene, S. D. (1992). DNA bending, flexibility, and helical repeat by cyclization kinetics. Methods Enzymol., 212, 3–29.CrossRefGoogle ScholarPubMed
Cubero, E., Luque, F. J. and Orozco, M. (1998). Is polarization important in cation–π interactions?Proc. Natl Acad. Sci., 95, 5976–5980.CrossRefGoogle ScholarPubMed
Curtis, H. J. and Cole, K. S. (1942). Membrane resting and action potentials from the squid giant axon. J. Cell. Comp. Physiol., 19, 135–144.CrossRefGoogle Scholar
Daggett, V. and Fersht, A. R. (2003). Is there a unifying mechanism for protein folding? TIBS, 28, 18–25.Google Scholar
DeFelice, L. J. (1981). Introduction to Membrane Noise. New York: Plenum Press.CrossRefGoogle Scholar
DeFelice, L. J.(2004). Transporter structure and mechanism. TINS, 27, 352–359.Google ScholarPubMed
Gennes, P. G. (1972). Exponents for the excluded volume problem as derived by the Wilson method. Phys. Letts, 38A, 339–340.CrossRefGoogle Scholar
Lean, A., Stadel, J. M. and Lefkowitz, R. J. (1980). A ternary complex model explains the agonist-specific binding properties of the adenylate cyclase-coupled b-adrenergic receptor. J. Biol. Chem., 255, 7108–7117.Google Scholar
Schutter, E. (1986). Alternative equations for the molluscan ion currents described by Connor and Stevens. Brain Res., 382, 134–138.CrossRefGoogle ScholarPubMed
Dill, K. A. (1990). Dominant forces in protein folding. Biochemistry, 29, 7133–7155. [Cited in Chapter 2. This is an excellent review with an emphasis on the hydrophobic effect.]CrossRefGoogle ScholarPubMed
Dill, D. A., Bromberg, S., Yue, K.et al. (1995). Principles of protein folding – a perspective from simple exact models. Protein Sci., 4, 561–602.CrossRefGoogle ScholarPubMed
Dinner, A. R. and Karplus, M. (2001). Comment on the communication “The key to solving the protein-folding problem lies in an accurate description of the denatured state” by van Gunsteren et al. Angew. Chem. Int. Ed., 40, 4615–4616.3.0.CO;2-H>CrossRefGoogle Scholar
Doyle, D. A., Cabral, J. M., Pfuetzner, R. A.et al. (1998). The structure of the potassium channel: molecular basis of K+ conduction and selectivity. Science, 280, 69–77. [Cited in Chapter 14. The first crystal structure of a selective channel and a major advance.]CrossRefGoogle ScholarPubMed
Dunitz, J. D. (1994). The entropic cost of bound water in crystals and biomolecules. Science, 264, 670. [Cited in Chapter 4. This is a lucid account of the thermodynamics of water association with proteins.]CrossRefGoogle ScholarPubMed
Eaton, W. A., Henry, E. R. and Hofrichter, J. (1991). Application of linear free energy relations to protein conformational changes: the quaternary structural change of hemoglobin. Proc. Natl Acad. Sci., 88, 4472–4475.CrossRefGoogle ScholarPubMed
Edwards, S., Corry, B., Kuyucak, S. and Chung, S. -H. (2002). Continuum electrostatics fails to describe ion permeation in the gramicidin channel. Biophys. J., 83, 1348–1360.CrossRefGoogle ScholarPubMed
Ehrenstein, G., Blumenthal, R., Latorre, R. and Lecar, H. (1974). Kinetics of the opening and closing of individual excitability-inducing material channels in a lipid bilayer. J. Gen. Physiol., 63, 707–721.CrossRefGoogle Scholar
Eigen, M. and Hammes, G. G. (1963). Elementary steps in enzyme reactions. Adv. Enyzmol., 25, 1–38.Google ScholarPubMed
Einstein, A. (1956). Investigations on the Theory of Brownian Movement. New York: Dover. [Cited in Chapter 6. This small book contains five papers published by Einstein from 1905 to 1908. This work is timeless and still worthy of careful study by students of biophysics.]Google Scholar
Eisenman, G. and Horn, R. (1983). Ionic selectivity revisited: the role of kinetic and equilibrium processes in ion permeation through channels. J. Membrane Biol., 76, 197–225.CrossRefGoogle ScholarPubMed
Elson, E. L. and Magde, D. (1974). Fluorescence correlation spectroscopy: I. Conceptual basis. Biopolymers, 13, 1–27.CrossRefGoogle Scholar
Falke, J. J., Drake, S. K., Hazard, A. L. and Peersen, O. B. (1994). Molecular tuning of ion binding to calcium signaling proteins. Q. Rev. Biophys., 27, 219–290.CrossRefGoogle ScholarPubMed
Fersht, A. R. (1987). The hydrogen bond in molecular recognition. Trends Biochem. Sci., 12, 301–304.CrossRefGoogle Scholar
Fersht, A. R.(1998). Structure and Mechanism in Protein Science. New York: Freeman. [Cited in Chapters 7 and 10. This is an excellent resource for kinetics and enzyme mechanisms.]Google Scholar
Fersht, A. R., Shi, J. -P., Knill-Jones, J.et al. (1985). Hydrogen bonding and biological specificity analysed by protein engineering. Nature, 314, 235–238.CrossRefGoogle ScholarPubMed
Fersht, A. R., Itzhaki, L. S., El Masry, N. F., Matthews, J. M. and Otzen, D. E. (1994). Single versus parallel pathways of protein folding and fractional formation of structure in the transition state. Proc. Natl Acad. Sci., 91, 10 426–10 429. [Cited in Chapter 7. This provides a very nice example of how to use Φ plots to address the mechanism of protein unfolding.]CrossRefGoogle ScholarPubMed
Fielder, E. M., Roberts, P. B., Bray, R. C.et al. (1974). The mechanism of action of superoxide dismutase from pulse radiolysis and electron paramagnetic resonance. Biochem. J., 139, 49–60.CrossRefGoogle Scholar
Finkel, A. S. and Redman, S. J. (1983). The synaptic current evoked in cat spinal motoneurons by impulse in single group Ia axons. J. Physiol., 342, 615–632.CrossRefGoogle Scholar
Finkelstein, A. (1987). Water Movement Through Lipid Bilayers, Pores, and Plasma Membranes. New York: Wiley-Interscience.Google Scholar
Finkelstein, A. and Andersen, O. S. (1981). The gramicidin A channel: A review of its permeability characteristics with special reference to the single-file aspect of transport. J. Membrane Biol., 59, 155–171.CrossRefGoogle Scholar
Finkelstein, A. V. and Janin, J. (1989). The price of lost freedom: entropy of bimolecular complex formation. Protein Engineering, 3, 1–3.CrossRefGoogle Scholar
Flory, P. J. (1969). Statistical Mechanics of Chain Molecules. New York: Interscience Publishers.Google Scholar
Fredkin, D. R., Montal, M. and Rice, J. A. (1985). Identification of aggregated Markovian models: application to the nicotinic acetylcholine receptor. In Proceedings of the Berkeley Conference in Honor of Jerzy Neyman and Jack Kiefer, ed. Carn, L. M. and Olshen, R. A.. Wadsworth, Vol. 1, pp. 269–289.Google Scholar
Gallivan, J. P. and Dougherty, D. A. (1999). Cation–π interactions in structural biology. Proc. Natl Acad. Sci., 96, 9459–9464.CrossRefGoogle ScholarPubMed
Gallivan, J. P. and Dougherty, D. A.(2000). A computational study of cation-π interactions vs salt bridges in aqueous media: Implications for protein engineering. J. Amer. Chem. Soc., 122, 870–874.CrossRefGoogle Scholar
Garofoli, S. and Jordan, P. C. (2003). Modeling permeation energetics in the KcsA potassium channel. Biophys. J., 84, 2814–2830.CrossRefGoogle ScholarPubMed
Gavish, B. and Werber, M. M. (1979). Viscosity-dependent structural fluctuations in enzyme catalysis. Biochemistry, 18, 1269–1275.CrossRefGoogle ScholarPubMed
Gentet, L. J., Stuart, G. J. and Clements, J. D. (2000). Direct measurement of specific membrane capacitance in neurons. Biophys. J., 79, 314–320.CrossRefGoogle ScholarPubMed
Gilson, M. K., Given, J. A., Bush, B. L. and McCammon, J. A. (1997). The statistical-thermodynamic basis for computation of binding affinities: a critical review. Biophys. J., 72, 1047–1069. [Cited in Chapter 4. This is an excellent review of the statistical mechanics of molecular associations.]CrossRefGoogle ScholarPubMed
Goldman, L. and Albus, J. S. (1968). Computation of impulse conduction in myelinated fibers; theoretical basis of the velocity-diameter relation. Biophys. J., 8, 596–607.CrossRefGoogle ScholarPubMed
Goldstein, S. S. and Rall, W. (1974). Changes of action potential shape and velocity for changing core conductor geometry. Biophys. J., 14, 731–757.CrossRefGoogle ScholarPubMed
Goychuk, I. and Hänggi, P. (2002). Ion channel gating: a first-passage time analysis of the Kramers type. Proc. Natl Acad. Sci., 99, 3552–3556.CrossRefGoogle ScholarPubMed
Green, W. N., Weiss, L. B. and Andersen, O. S. (1987). Batrachotoxin-modified sodium channels in planar lipid bilayers. J. Gen. Physiol., 89, 841–872.CrossRefGoogle ScholarPubMed
Griko, Y. V. and Privalov, P. L. (1992). Calorimetric study of the heat and cold denaturation of beta-lactoglobulin. Biochemistry, 31, 8810–8815.CrossRefGoogle ScholarPubMed
Grosman, C. (2003). Free-energy landscapes of ion-channel gating are malleable: changes in the number of bound ligands are accompanied by changes in the location of the transition state of the acetylcholine-receptor channels. Biochemistry, 42 (50), 14 977–14 987.CrossRefGoogle ScholarPubMed
Grosman, C., Zhou, M. and Auerbach, A. (2000). Mapping the conformational wave of acetylcholine receptor channel gating. Nature, 403, 773–776. [Cited in Chapter 7. This is an important advance mapping the gating transition of an ion channel and an excellent illustration of the use of Φ plots in biophysics.]CrossRefGoogle ScholarPubMed
Gurney, R. W. (1953). Ionic Processes in Solution. New York: McGraw-Hill.Google Scholar
Gutman, M. and Nachliel, E. (1997). Time-resolved dynamics of proton transfer in proteinous systems. Ann. Rev. Phys. Chem., 48, 329–356.CrossRefGoogle ScholarPubMed
Hagen, S. J. and Eaton, W. A. (1996). Nonexponential structural relaxations in proteins. J. Chem. Phys., 104, 3395–3398.CrossRefGoogle Scholar
Hall, J. E., Mead, C. A. and Szabo, G. (1973). A barrier model for current flow in lipid bilayer membranes. J. Membr. Biol., 11, 75–97. [Cited in Chapter 14. This is an early example of how barriers shape the current–voltage behavior of a membrane.]CrossRefGoogle Scholar
Hammes, G. G. (1978). Principles of Chemical Kinetics. New York: Academic Press.Google Scholar
Han, W. -G., Jalkanen, K. J., Elstner, M. and Suhai, S. (1998). Theoretical study of aqueous N-acetyl-L-alanine N′-methylamine: structures and Raman, VCD, and ROA spectra. J. Phys. Chem. B, 102, 2487–2602.CrossRefGoogle Scholar
Hänggi, P., Talkner, P. and Borkovec, M. (1990). Reaction-rate theory: fifty years after Kramers. Rev. Modern Phys., 62, 251–341. [Cited in Chapter 7. This is an excellent review of modern approaches to rate theory.]CrossRefGoogle Scholar
Hardy, L. W. and Kirsch, J. F. (1984). Diffusion-limited components of reactions catalyzed by bacillus cereus b-lactanase I. Biochemistry, 23, 1275–1282.CrossRefGoogle Scholar
Hecht, S., Schlaer, S. and Pirenne, M. (1942). Energy, quanta, and vision. J. Gen. Physiol., 25, 819–840.CrossRefGoogle ScholarPubMed
Hedstrom, L., Perona, J. J. and Rutter, W. J. (1994). Converting trypsin to chymotrypsin: Residue 172 is a substrate specificity determinant. Biochemistry, 33, 8757–8763.CrossRefGoogle ScholarPubMed
Hess, P. and Tsien, R. W. (1984). Mechanism of ion permeation through calcium channels. Nature, 309, 453–456.CrossRefGoogle ScholarPubMed
Hestrin, S., Nicoll, R. A., Perkel, D. J. and Sah, P. (1990). Analysis of excitatory synaptic action in pyramidal cells using whole-cell recording from rat hippocampal slices. J. Physiol., 422, 203–225.CrossRefGoogle ScholarPubMed
Hill, T. L. (1960). An Introduction to Statistical Thermodynamics. Reading: Addison-Wesley. [This is highly recommended as a resource for statistical mechanics.]
Hille, B. (1977). Ionic basis of resting and action potentials. In Handbook of Physiology. The Nervous System. Cellular Biology of Neurons, ed. Brookhart, J. M. and Mountcastle, V. B.. Bethesda: American Physiological Society, pp. 99–136.Google Scholar
Hille, B.(1991). Ion Channels of Excitable Membranes. Sunderland: Sinauer Associates.Google Scholar
Hille, B. and Schwarz, W. (1978). Potassium channels as multi-ion single-file pores. J. Gen. Physiol., 72, 409–442. [Cited in Chapter 14. A full treatment of single-file models and an excellent summary of their basic properties.]CrossRefGoogle ScholarPubMed
Hines, M. L. and Carnevale, N. T. (1997). The NEURON simulation environment. Neural Computation, 9, 1179–1209.CrossRefGoogle ScholarPubMed
Hodgkin, A. L. (1964). The Conduction of Nervous Impulse. Springfield, IL: Charles Thomas.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 Huxley, A. F. (1952). A quantitative description of membrane current and its application to conduction and excitation in nerve. J. Physiol., 117, 500–544. [Cited in Chapter 16. This is the culminating paper in the seminal series on membrane excitability.]CrossRefGoogle ScholarPubMed
Hodgkin, A. L. and Keynes, R. D. (1955). The potassium permeability of a giant nerve fibre. J. Physiol., 128, 61–88. [Cited in Chapter 14. This is the classical development of the single-file theory of permeation.]CrossRefGoogle ScholarPubMed
Hodgkin, A. L. and Rushton, W. A. H. (1946). The electrical constants of a crustacean nerve fiber. Proc. R. Soc. Lond., B133, 444–479.CrossRefGoogle Scholar
Hodgkin, A. L., Huxley, A. F. and Katz, B. (1952). Measurement of current voltage relations in the membrane of the giant axon of Loligo. J. Physiol., 116, 424–448.CrossRefGoogle ScholarPubMed
Hol, W. G. J., Duijnen, P. T. and Berendsen, H. J. C. (1978). The α-helix dipole and the properties of proteins. Nature, 273, 443–446.CrossRefGoogle ScholarPubMed
Horn, R. and Lange, K. (1983). Estimating kinetic constants from single channel data. Biophys. J., 43, 207–223.CrossRefGoogle ScholarPubMed
Horn, R. and Vandenberg, C. A. (1984). Statistical properties of single sodium channels. J. Gen. Physiol., 84, 505–534.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
Imoto, K., Busch, C., Sakmann, B.et al. (1988). Rings of negatively charged amino acids determine the acetylcholine receptor channel conductance. Nature, 335, 645–648.CrossRefGoogle ScholarPubMed
Itzhaki, L. S., Otzen, D. E. and Fersht, A. R. (1995). The structure of the transition state for protein folding of chymotrypsin inhibitor 2 analysed by protein engineering methods: evidence for a nucleation condensation mechanism for protein folding. J. Mol. Biol., 254, 260–288.CrossRefGoogle ScholarPubMed
Jack, J. J. B. and Redman, S. J. (1971). An electrical description of the motoneuron and its application to the analysis of synaptic potentials. J. Physiol., 215, 321–352.CrossRefGoogle ScholarPubMed
Jack, J. J. B., Noble, D. and Tsien, R. W. (1983). Electric Current Flow in Excitable Cells. Oxford: Oxford University Press. [Cited in Chapters 15 and 16. This is a comprehensive textbook that covers cable theory and excitability very well.]Google Scholar
Jackson, J. D. (1975). Classical Electrodynamics. New York: John Wiley & Sons.Google Scholar
Jackson, M. B. (1985). The stochastic behavior of a many channel membrane system. Biophys. J., 47, 129–137.CrossRefGoogle ScholarPubMed
Jackson, M. B.(1992). Cable analysis with the whole-cell patch clamp: theory and experiment. Biophys. J., 61, 756–766.CrossRefGoogle ScholarPubMed
Jackson, M. B.(1993a). Passive current flow and morphology in the terminal arborizations of the posterior pituitary. J. Neurophysiol., 69, 692–702.CrossRefGoogle Scholar
Jackson, M. B.(1993b). Binding specificity of receptor chimeras revisited. Biophys. J., 63, 1443–1444.CrossRefGoogle Scholar
Jackson, M. B.(1993c). On the time scale and time course of protein conformational changes. J. Chem. Phys., 99, 7253–7259.CrossRefGoogle Scholar
Jackson, M. B.(1994). Single channel currents in the nicotinic receptor: A direct demonstration of allosteric transitions. TIBS, 19, 396–399.Google ScholarPubMed
Jackson, M. B.(1997a). Adding up the energies in the acetylcholine receptor channel: relevance to allosteric theory. In The Nicotinic Acetylcholine Receptor: Current Views and Future Trends, ed. Barrantes, F.. Austin: Landes Bioscience, pp. 61–84.Google Scholar
Jackson, M. B.(1997b). Inversion of Markov processes to determine rate constants from single channel data. Biophys. J., 73, 1382–1394.CrossRefGoogle Scholar
Jackson, M. B.(1998). Allosteric mechanisms in the activation of ligand-gated channels. Biophysics Textbook of the Biophysical Society. Rockville: Biophysical Society. [Cited in Chapter 5. A discussion of allosteric mechanisms using ligand-gated channels to illustrate important concepts.]Google Scholar
Jackson, M. B. and Zhang, S. J. (1995). Action potential propagation and propagation block by GABA in rat posterior pituitary nerve terminals. J. Physiol., 483(3), 597–611.CrossRefGoogle ScholarPubMed
Jackson, M. B., Wong, B. M., Morris, C. E., Lecar, H. and Christian, C. N. (1983). Successive openings of the same acetylcholine receptor channel are correlated in open time. Biophys. J., 42, 109–114.CrossRefGoogle ScholarPubMed
Jackson, M. B., Konnerth, A. and Augustine, G. J. (1991). Action potential broadening and frequency-dependent facilitation of calcium signals in pituitary nerve terminals. Proc. Natl Acad. Sci., 88, 380–384.CrossRefGoogle ScholarPubMed
Jacobson, K., Ishihara, A. and Inman, R. (1987). Lateral diffusion of proteins in membranes. Ann. Rev. Physiol., 49, 163–175.CrossRefGoogle ScholarPubMed
Jeffrey, G. A. and Saenger, W. (1991). Hydrogen Bonding in Biological Structures. Berlin: Springer–Verlag.CrossRefGoogle Scholar
Jencks, W. P. (1975). Binding energy, specificity, and enzyme catalysis: the Circe effect. Adv. Enzymol., 43, 219–410.Google ScholarPubMed
Jencks, W. P. and Carriuolo, J. (1961). General base catalysis of ester hydrolysis. Journal of the American Chemical Society, 83, 1743–1750.CrossRefGoogle Scholar
Jentsch, T. J., Stein, V., Weinreich, F. and Zdebik, A. A. (2002). Molecular structure and physiological function of chloride channels. Physiol. Rev., 82, 503–568.CrossRefGoogle ScholarPubMed
Jones, S. W. (1989). On the resting potential of isolated frog sympathetic neurons. Neuron, 3, 153–161.CrossRefGoogle ScholarPubMed
Jordan, P. C. (1990). Ion-water and ion-polypeptide interactions in a gramicidin-like channel. A molecular dynamics study. Biophys. J., 58, 1133–1156.CrossRefGoogle Scholar
Jordan, P. C., Bacquet, R. J., McCammon, A. J. and Tran, P. (1989). How electrolyte shielding influences the electrical potential in transmembrane ion channels. Biophys. J., 55, 1041–1052.CrossRefGoogle ScholarPubMed
Kallenbach, N. (2001). Breaking open a protein barrel. Proc. Natl Acad. Sci., 98, 2958–2960. [Cited in Chapter 2. This presents a clear presentation of the oil-droplet versus jigsaw puzzle pictures of a protein interior.]CrossRefGoogle Scholar
Kao, J. P. Y. and Tsien, R. Y. (1988). Ca2 + binding kinetics of fura-2 and azo-1 from temperature jump relaxation measurements. Biophys. J., 53, 635–639.CrossRefGoogle ScholarPubMed
Karplus, M. (2002). Molecular dynamics simulations of biomolecules. Acc. Chem. Res., 35, 321–323. [Cited in Chapter 2. This is the editorial for a special issue that covers a wide range of applications.]CrossRefGoogle ScholarPubMed
Katz, B. (1966). Nerve, Muscle, and Synapse. New York: McGraw-Hill.Google Scholar
Kaya, H. and Chan, H. S. (2000). Polymer principles of protein calorimetric two-state cooperativity. Proteins: Structure, Function, and Genetics, 40, 637–661.3.0.CO;2-4>CrossRefGoogle ScholarPubMed
Keizer, J. (1987). Diffusion effects on rapid bimolecular chemical reactions. Chem. Rev., 87, 167–180.CrossRefGoogle Scholar
Kell, M. J. and DeFelice, L. J. (1988). Surface charge near the cardiac inward-rectifier channel measured from single-channel conductance. J. Membrane Biol., 102, 1–10.CrossRefGoogle ScholarPubMed
Kellermayer, M. S. Z., Smith, S. B., Granzier, H. L. and Bustamante, C. (1997). Folding–unfolding transitions in single titin molecules characterized with laser tweezers. Science, 276, 1112–1116.CrossRefGoogle ScholarPubMed
Khanin, R., Parnas, H. and Segel, L. (1994). Diffusion cannot govern the discharge of neurotransmitter in fast synapses. Biophys. J., 67, 966–972.CrossRefGoogle ScholarPubMed
Kijima, S. and Kijima, H. (1987). Statistical analysis of channel current from a membrane patch I. Some stochastic properties of ion channels or molecular systems at equilibrium. J. Theor. Biol., 128, 423–434.CrossRefGoogle Scholar
Kittel, C. (1958). Elementary Statistical Physics. New York: John Wiley & Sons.Google Scholar
Kolinski, A., Godzik, A. and Skolnick, J. (1993). A general method for the prediction of the three dimensional structure and folding pathway of globular proteins: Application to designed helical proteins. J. Chem. Phys., 98, 7420–7433.CrossRefGoogle Scholar
Kolinski, A., Galazka, W. and Skolnick, J. (1996). On the origin of the co-operativity of protein folding: implications from model simulations. Proteins: Structure, Function, and Genetics, 26, 271–287.3.0.CO;2-H>CrossRefGoogle Scholar
Koshland, D. E., Nemethy, G. and Filmer, D. (1966). Comparison of experimental binding data and theoretical models in proteins containing subunits. Biochemistry, 5, 365–384.CrossRefGoogle ScholarPubMed
Kramers, H. A. (1940). Brownian motion in a field of force. Physica, 7, 284–304.CrossRefGoogle Scholar
Kuyucak, S., Andersen, O. S. and Chung, S. -H. (2001). Models of permeation in ion channels. Reports of Progress in Physics, 64, 1427–1472.CrossRefGoogle Scholar
Latorre, R., Labarca, P. and Naranjo, D. (1992). Surface charge effects on ion conduction in ion channels. In Ion Channels (Methods in Enzymology), ed. L. Iverson and B. Rudy, vol. 207, pp. 471–501. [Cited in Chapter 11. A very clear overview of surface charge effects in ion channels.]CrossRef
Läuger, P. (1973). Ion transport through pores: a rate-theory analysis. Biophysica and Biochimica Acta, 311, 423–441. [Cited in Chapter 14. A thorough development of barrier models for permeation.]CrossRefGoogle ScholarPubMed
Lecar, H. and Sachs, F. (1981). Membrane noise analysis. In Excitable Cells in Tissue Culture, ed. Nelson, P. G. and Lieberman, M.. New York: Plenum, pp. 137–172.CrossRefGoogle Scholar
Lee, A. W., Karplus, M., Poyart, C. and Bursaux, E. (1988). Analysis of proton release in oxygen binding by hemoglobin: implications for cooperative mechanism. Biochemistry, 27, 1285–1301.CrossRefGoogle ScholarPubMed
Lesk, A. M., Lo Conte, L. and Hubbard, T. J. P. (2001). Assessment of novel fold targets in CASP4: Predictions of three-dimensional structures, secondary structures, and interresidue contacts. Proteins: Structure, Function, and Genetics, 45(S5), 98–118.CrossRefGoogle Scholar
Levinthal, C. (1968). Are there pathways for protein folding. J. Chem. Phys., 65, 44–45.Google Scholar
Levitt, D. G (1978a). Electrostatic calculations for an ion channel. I. Energy and potential profiles and interactions between ions. Biophys. J., 22, 209–219.CrossRefGoogle Scholar
Levitt, D. G(1978b). Electrostatic calculations for an ion channel II. Kinetic behavior of the gramicidin A channel. Biophys. J., 22, 221–248.CrossRefGoogle Scholar
Levitt, M., Sander, C. and Stern, P. S. (1985). Protein normal-mode dynamics: trypsin inhibitor, crambin, ribonuclease and lysozyme. J. Mol. Biol., 181, 423–447.CrossRefGoogle ScholarPubMed
Levitt, M., Hirshberg, M., Sharon, R. and Daggett, V. (1995). Potential energy function and parameters for simulations of the molecular dynamics of proteins and nucleic acids in solution. Comput. Phys. Communs, 91, 215–231.CrossRefGoogle Scholar
Lewis, C. A. (1979). Ion-concentration dependence of the reversal potential and the single channel conductance of ion channels at the frog neuromuscular junction. J. Physiol., 286, 417–445.CrossRefGoogle ScholarPubMed
Lifson, S. and Roig, A. (1963). On the theory of helix–coil transition in polypeptides. J. Chem. Phys., 34, 1961–1974. [Cited in Chapter 3. This is an elegent and clear development of the mathematical theory of helix–coil transitions.]Google Scholar
Lifson, S. and Warshel, A. (1968). Consistent force field for calculations of conformations, vibrational spectra, and enthalpies of cycloalkane and n-alkane molecules. J. Chem. Phys., 49, 5119–5129.CrossRefGoogle Scholar
Lin, J., Cassidy, C. S. and Frey, P. A. (1998). Correlations of the basicity of His 57 with transition state analogue binding, substrate reactivity, and the strength of the low-barrier hydrogen bond in chymotrypsin. Biochemistry, 37, 11 940–11 948.CrossRefGoogle ScholarPubMed
Linderstrøm-Lang, K. (1924). On the ionization of proteins. Compt. rend. trav. Carlsberg, 17, 1–29. (See also Linderstrøm-Lang, K. (1962) Selected Papers. New York: Academic Press.)Google Scholar
Ma, J. C. and Dougherty, D. A. (1997). The cation–π interaction. Chem. Rev., 97, 1303–1324.CrossRefGoogle Scholar
MacInnes, D. A. (1961). The Principles of Electrochemistry. New York: Dover.Google Scholar
MacKinnon, R., Latorre, R. and Miller, C. (1989). Role of surface electrostatics in the operation of a high-conductance Ca2 +-activated K+ channel. Biochemistry, 28, 8092–8099.CrossRefGoogle ScholarPubMed
Magee, J. C. and Cook, E. P. (2000). Somatic EPSP amplitude is independent of synapse location in hippocampal pyramidal neurons. Nature Neurosci., 3, 895–903.CrossRefGoogle ScholarPubMed
Mainen, Z. F., Joerges, J., Huguenard, J. R. and Sejnowski, T. J. (1995). A model of spike initiation in neocortical pyramidal neurons. Neuron, 15, 1427–1439.CrossRefGoogle ScholarPubMed
Major, G., Evan, J. D. and Jack, J. B. (1993). Solutions for transients in arbitrarily branching cables: I. Voltage recording with a somatic shunt. Biophys. J., 65, 423–449. [Cited in Chapter 15. This paper explores the cable equation in complicated dendritic arbors.]CrossRefGoogle ScholarPubMed
Manning, G. S. (1969). Limiting laws and counterion condensation in polyelectrolyte solutions. I. Colligative properties. J. Chem. Phys., 51, 924–933.CrossRefGoogle Scholar
Manning, G. S.(1978). The molecular theory of polyelectrolyte solutions with applications to the electrostatic properties. Q. Rev. Biophys., 11, 179–246.CrossRefGoogle ScholarPubMed
Marcus, R. A. (1964). Chemical and electrochemical electron-transfer theory. Ann. Rev. Phys. Chem., 15, 155–196.CrossRefGoogle Scholar
Marcus, R. A.(1968). Theoretical relations among rate constants, barriers, and Brønsted slopes of chemical reactions. J. Phys. Chem., 72, 891–899.CrossRefGoogle Scholar
Martinez, M. B., Flickinger, M. C. and Nelsestuen, G. L. (1996). Accurate kinetic modeling of alkaline phosphatase in the Escherichia coli periplasm: implications for enzyme properties and substrate diffusion. Biochemistry, 35, 1179–1186.CrossRefGoogle ScholarPubMed
Marty, A, and Neher, E. (1995). Tight-seal whole-cell recording. In Single-Channel Recording, ed. Sakmann, B. and Neher, E.. New York: Plenum, pp. 31–51.Google Scholar
McCammon, A. J., Wolynes, P. G. and Karplus, M. (1979). Picosecond dynamics of tyrosine side chains in proteins. Biochemistry, 18, 927–942. [Cited in Chapter 10. This is a seminal paper on the internal dynamics of proteins.]CrossRefGoogle ScholarPubMed
McLaughlin, S. (1989). The electrostatic properties of membranes. Ann. Rev. Biophys. Biophys. Chem., 18, 113–136.CrossRefGoogle ScholarPubMed
McQuarrie, D. A., (1976). Statistical Mechanics. New York: Harper & Row. [This is highly recommended as a resource for statistical mechanics.]Google Scholar
Meyer, E., Müller, C. O. and Fromherz, P. (1997). Cable properties of dendrites in hippocampal neurons of the rat mapped by a voltage-sensitive dye. Eur. J. Neurosci., 9, 778–785.CrossRefGoogle ScholarPubMed
Miedema, H. (2002). Surface potentials and the calculated selectivity of ion channels. Biophys. J., 82, 156–159.CrossRefGoogle ScholarPubMed
Migliore, M., Hoffman, D. A., Magee, J. C. and Johnston, D. (1999). Role of an A-type K+ conductance in the back-propagation of action potentials in the dendrites of hippocampal pyramidal neurons. J. Comp. Neurosci., 7, 5–15.CrossRefGoogle ScholarPubMed
Millar, J. A., Barrett, L., Southan, A. P., Page, K. M., Fyffe, R. E. W. and Robertson, B. (2000). A functional role for the two-pore domain potassium channel TASK-1 in cerebellar granule neurons. Proc. Natl Acad. Sci., 97, 3614–3618.CrossRefGoogle ScholarPubMed
Miller, B. G. and Wolfenden, R. (2002). Catalytic proficiency: The unusual case of OMP decarboxylase. Ann. Rev. Biochem., 71, 847–885.CrossRefGoogle ScholarPubMed
Moczydlowski, E., Alvarez, O., Vergara, C. and Latorre, R. (1985). Effect of phospholipid surface charge on the conductance and gating of a Ca2 +-activated K+ channel in planar lipid bilayers. J. Membrane Biol., 83, 273–282.CrossRefGoogle ScholarPubMed
Monod, J., Wyman, J. and Changeux, J. -P. (1965). On the nature of allosteric transitions: A plausible model. J. Mol. Biol., 12, 88–118. [Cited in Chapter 5. This is the seminal paper on allosteric regulation of proteins. A remarkable conceptual advance and still well worth reading.]CrossRefGoogle ScholarPubMed
Moore, W. J. (1972). Physical Chemistry. Englewood Cliffs: Prentice-Hall. [This is highly recommended as a resource for statistical mechanics.]Google Scholar
Morais-Cabral, J. H., Zhou, Y. and MacKinnon, R. (2001). Energetic optimization of ion conduction rate by the K+ selectivity filter. Nature, 414, 37–42.CrossRefGoogle ScholarPubMed
Morris, C. E. and Lecar, H. (1981). Voltage oscillations in the barnacle giant muscle fiber. Biophys. J., 35, 193–213. [Cited in Chapter 16. A clear and illuminating account of membrane oscillations.]CrossRefGoogle ScholarPubMed
Moy, G., Corry, B., Kuyucak, S. and Chung, S. -H. (2000). Tests of continuum theories as models of ion channels. I. Poisson–Boltzmann theory versus Brownian dynamics. Biophys. J., 78, 2349–2363.CrossRefGoogle ScholarPubMed
Myers, J. K. and Pace, C. N. (1996). Hydrogen bonding stabilizes globular proteins. Biophys. J., 71, 2033–2039.CrossRefGoogle ScholarPubMed
Nakajima, Y., Nakajima, S. and Inoue, M. (1988). Pertussis toxin-insensitive G protein mediates substance P-induced inhibition of potassium channels in brain neurons. Proc. Natl Acad. Sci., 85, 3643–3647.CrossRefGoogle ScholarPubMed
Nakatani, H. and Dunford, H. B. (1979). Meaning of diffusion-controlled association rate constants in enzymology. J. Phys. Chem., 83, 2662–2665.CrossRefGoogle Scholar
Neher, E. and Steinbach, J. H. (1978). Local anesthetics transiently block currents through single acetylcholine-receptor channels. J. Physiol., 277, 153–176. [Cited in Chapter 9. This is a striking example of the power of single-channel kinetics in the analysis of mechanisms of drug action.]CrossRefGoogle ScholarPubMed
Nelsestuen, G. L. and Martinez, M. B. (1997). Steady state enzyme velocities that are independent of [enzyme]: an important behavior in many membrane and particle-bound states. Biochemistry, 36, 9081–9086.CrossRefGoogle Scholar
Nolte, H. -J., Rosenberry, T. L. and Neumann, E. (1980). Effective charge on acetylcholinesterase active sites determined from the ionic strength dependence of association rate constants with cationic ligands. Biochemistry, 19, 3705–3711.CrossRefGoogle ScholarPubMed
Noskov, S. Y., Berniche, S. and Roux, B. (2004). Control of ion selectivity in potassium channels by electrostatic and dynamic properties of carbonyl ligands. Nature, 431, 830–834.CrossRefGoogle ScholarPubMed
Obaid, A. L. and Salzberg, B. M. (1996). Micromolar 4-aminopyridine enhances invasion of a vertebrate neurosecretory terminal arborization. J. Gen. Physiol., 107, 353–368.CrossRefGoogle ScholarPubMed
Oberhauser, A. F. and Fernandez, J. M. (1995). Hydrophobic ions amplify the capacitance currents used to measure exocytotic fusion. Biophys. J., 69, 451–459.CrossRefGoogle ScholarPubMed
O'Mara, M., Barry, P. H. and Chung, S. -H. (2003). A model of the glycine receptor deduced from Brownian dynamics studies. Proc. Natl Acad. Sci., 100, 4310–4315.CrossRefGoogle ScholarPubMed
O'Neil, K. T. and DeGrado, W. F. (1990). A thermodynamic scale for the helix-forming tendencies of the commonly occurring amino acids. Science, 250, 646–651.CrossRefGoogle ScholarPubMed
Oosawa, F. (1971). Polyelectrolytes. New York: Marcel Dekker, Inc.Google Scholar
Overbeek, J. Th. G. (1952). The electrochemistry of the double layer. In Colloid Science, Vol. 1, ed. Kruyt, H. R.. Amsterdam: Elsevier Publishing Company, pp. 115–193.Google Scholar
Overbeek, J. T. G. and Wiersema, P. H. (1967). The interpretation of electrophoretic mobilities. In Electrophoresis, ed. Bier, M.. New York: Academic Press, pp. 1–52.Google Scholar
Pace, C. N. (1992). Contributions of the hydrophobic effect to globular protein stability. J. Mol. Biol., 226, 29–35.CrossRefGoogle Scholar
Papazian, D. M., Timpe, L. C., Jan, Y. N. and Jan, L. Y. (1991). Alteration of voltage-dependence of Shaker potassium channel by mutations in the S4 sequence. Nature, 349, 305–310.CrossRefGoogle ScholarPubMed
Parsegian, V. A. (1969). Energy of an ion crossing a low dielectric membrane: Solutions to four relevant electrostatics problems. Nature, 221, 844–846. [Cited in Chapters 2 and 14. This is an early study that defined the basic energetic parameters of permeation.]CrossRefGoogle Scholar
Parsegian, V. A.(1973). Long-range physical forces in the biological milieu. Ann. Rev. Biophys. Bioeng., 2, 221–255.CrossRefGoogle ScholarPubMed
Patlak, C. S. (1960). Derivation of an equation for the diffusion potential. Nature, 188, 944–945.CrossRefGoogle ScholarPubMed
Perkel, D. H., Mulloney, B. and Budelli, R. W. (1981). Quantitative methods for predicting neuronal behavior. Neuroscience, 5, 823–837.CrossRefGoogle Scholar
Perutz, M. F., Wilkenson, A. J., Paoli, M. and Dodson, D. D. (1998). The stereochemical mechanism of the cooperative effects in hemoglobin revisited. Ann. Rev. Biophys. Biomol. Structure, 27, 1–34. [Cited in Chapter 5. This contains a clear and thorough discussion of the evidence favoring a two-state description of hemoglobin.]CrossRefGoogle ScholarPubMed
Peters, R. and Cherry, R. J. (1982). Lateral and rotational diffusion of bacteriorhodopsin in lipid bilayers: experimental test of the Saffman–Delbrück equations. Proc. Natl Acad. Sci., 79, 4317–4321.CrossRefGoogle ScholarPubMed
Piek, T. (1975). Ionic and electrical properties. In Insect Muscle, ed. Usherwood, P. N. R.. New York: Academic Press, pp. 275–336.Google Scholar
Plowman, K. M. (1972). Enzyme Kinetics. New York: McGraw Hill.Google Scholar
Pokarowski, P., Kolinski, A. and Skolnick, J. (2003). A minimal physically realistic protein-like lattice model: designing an energy landscape that ensures all-or-none folding to a unique native state. Biophys. J., 84, 1518–1526.CrossRefGoogle ScholarPubMed
Poland, D. and Scheraga, H. A. (1970). Theory of Helix–Coil Transitions. New York: Academic Press.Google Scholar
Privalov, P. L. (1979). Stability of proteins. Adv. Protein Chem., 33, 167–241.CrossRefGoogle ScholarPubMed
Privalov, P. L.(1982). Stability of proteins: Proteins which do not present a single cooperative system. Adv. Protein Chem., 35, 1–104. [Cited in Chapters 2 and 3. This reference presents thorough reviews of the thermodynamics of thermal transitions in proteins.]CrossRefGoogle ScholarPubMed
Prod'hom, B., Peitrobon, D. and Hess, P. (1987). Direct measurement of proton transfer rates to a group controlling the dihydropyridine-sensitive Ca2 + channel. Nature, 329, 243.CrossRefGoogle ScholarPubMed
Pumphrey, R. J. and Young, J. Z. (1938). The rates of conduction of nerve fibres of various diameters in cephalopods. J. Exp. Biol., 14, 453–466.Google Scholar
Putnam, S. J., Coulson, A. F., Farley, I. R., Ridd Leston, B. and Knowles, J. R. (1972). Specificity and kinetics of triose phosphate isomerase from chicken muscle. Biochem. J., 129, 301–310.Google Scholar
Raleigh, D. P. and DeGrado, W. F. (1992). A de novo designed protein shows a thermally induced transition from a native to a molten globule-like state. J. Amer. Chem. Soc., 114, 10 079–10 081.CrossRefGoogle Scholar
Rall, W. (1959) Branching dendritic trees and motoneuron membrane resistivity. Exp. Neurol., 1, 491–527. [Cited in Chapter 15. This provides a remarkable insight into how to simplify the cable analysis of dendrites.]CrossRefGoogle ScholarPubMed
Rall, W.(1967). Distinguishing theoretical synaptic potentials computed for different soma-dendritic distributions of synaptic input. J. Neurophysiol., 30, 1138–1168.CrossRefGoogle ScholarPubMed
Rall, W.(1969). Time constants and electrotonic length of membrane cylinders and neurons. Biophys. J., 9, 1483–1508. [Cited in Chapter 15. A thorough and clear exposition of practical aspects of cable analysis.]CrossRefGoogle ScholarPubMed
Rall, W.(1977). Core conductor theory and cable properties of neurons. In Handbook of Physiology. The Nervous System. Cellular Biology of Neurons, ed. Brookhart, J. M. and Mountcastle, V. B.. Bethesda: American Physiological Society, pp. 39–97.Google Scholar
Rall, W., Burke, R. E., Smith, T. G., Nelson, P. G. and Frank, K. (1967). Dendritic location of synapses and possible mechanisms for the monosynaptic EPSPs in motoneurons. J. Neurophysiol., 30, 1169–1193.CrossRefGoogle Scholar
Ramachandran, G. N. and Sasisekharan, V. (1968). Conformation of polypeptides and proteins. Adv. Protein Chem., 23, 284–437.Google ScholarPubMed
Ramachandran, G. N., Venkatachalam, C. M. and Krimm, S. (1966). Stereochemical criteria for polypeptide and protein chain conformations. 3. Helical and hydrogen-bonded polypeptide chains. Biophys. J., 6(6), 849–872.CrossRefGoogle ScholarPubMed
Rand, R. P. (1981). Interacting phospholipid bilayers: measured forces and induced structural changes. Ann. Rev. Biophys. Bioengin., 10, 288–314.CrossRefGoogle ScholarPubMed
Rapp, M., Yarom, Y. and Segev, I. (1996). Modeling back propagating action potentials in weakly excitable dendrites of neocortical pyramidal cells. Proc. Natl Acad. Sci., 93, 11 985–11 990.CrossRefGoogle ScholarPubMed
Rashin, A. A. and Honig, B. (1985). Reevaluation of the Born model of ion hydration. J. Phys. Chem., 89, 5588–5593.CrossRefGoogle Scholar
Record, M. T. (1975). Effects of Na+ and Mg++ ions on the helix–coil transition of DNA. Biopolymers, 14, 2137–2158.CrossRefGoogle Scholar
Record, M. T., Mazur, S. J., Melancon, P., Roe, J. -H., Shaner, S. L. and Unger, L. (1981). Double helical DNA: conformations, physical properties, and interactions with ligands. Ann. Rev. Biochem., 50, 997–1024.CrossRefGoogle ScholarPubMed
Redman, S. and Walmsley, B. (1983). The time course of synaptic potentials evoked in cat spinal motoneurons at identified group 1a synapses. J. Physiol., 343, 117–133.CrossRefGoogle Scholar
Reed, A. E. and Weinhold, F. (1991). Natural bond orbital analysis of internal rotation barriers and related phenomena. Isr. J. Chem., 31, 277–285.CrossRefGoogle Scholar
Rees, D. C., DeAntonio, L. and Eisenberg, D. (1989). Hydrophobic organization of membrane proteins. Science, 245, 510–513.CrossRefGoogle ScholarPubMed
Richard, J. P. (1998). The enhancement of enzymatic rate accelerations by Brønsted acid–base catalysis. Biochemistry, 37, 4305–4309.CrossRefGoogle ScholarPubMed
Rigler, R. and Elson, E. L. (eds.) (2001). Fluorescence Correlaton Spectroscopy: Theory and Applications. Berlin: Springer.CrossRefGoogle Scholar
Rogawski, M. A. (1985). The A-current: how ubiquitous a feature of excitable cells is it? Trends Neurosci., 8, 214–219.CrossRefGoogle Scholar
Rose, G. D., Gesolowitz, A. R., Lesser, G. J., Lee, R. H. and Zehfus, M. H. (1985). Hydrophobicity of amino acid residues in globular proteins. Science, 229 (4716), 834–838.CrossRefGoogle ScholarPubMed
Roseman, M. A. (1988). Hydrophobicity of the peptide C = O … H–N hydrogen-bonded group. J. Mol. Biol., 201, 621–623.CrossRefGoogle Scholar
Rosenthal, L., Rabolt, J. F. and Hummel, J. (1982). An investigation of the conformational equilibrium of n-butane in a solvent using Raman spectroscopy. J. Chem. Phys., 76, 817–820.CrossRefGoogle Scholar
Rothberg, B. S. and Magleby, K. L. (2001). Testing for detailed balance (microscopic reversibility) in ion channel gating. Biophys. J., 80, 3025–3026.CrossRefGoogle Scholar
Roux, B. and MacKinnon, R. (1999). The cavity and pore helices in the KcsA K+ channel: electrostatic stabilization of monovalent cations. Science, 285, 100–102.CrossRefGoogle ScholarPubMed
Roux, B., Berniche, S. and Im, W. (2000). Ion channels, permeation, and electrostatics: insight into the function of KcsA. Biochemistry, 39, 13 295–13 306.CrossRefGoogle ScholarPubMed
Rushton, W. A. H. (1951). A theory of the effects of fibre size in medullated nerve. J. Physiol., 115, 101–122.CrossRefGoogle ScholarPubMed
Saffman, P. G. and Delbrück, M. (1975). Brownian motion in biological membranes. Proc. Natl Acad. Sci., 72, 3111–3113.CrossRefGoogle ScholarPubMed
Sanchez, I. C. (1979). Phase transition behavior of the isolated polymer chain. Macromolecules, 12, 980–988.CrossRefGoogle Scholar
Saxton, M. J. and Jacobson, K. (1997). Single-particle tracking: applications to membrane dynamics. Ann. Rev. Biophys. Biomol. Struct., 26, 373–399.CrossRefGoogle ScholarPubMed
Scatchard, G. (1949). The attractions of proteins for small molecules and ions. Ann. N. Y. Acad. Sci., 51, 660–671.CrossRefGoogle Scholar
Schafmeister, C. E., LaPorte, S. L., Miercke, L. J. W. and Stroud, R. M. (1997). A designed four helix bundle protein with native-like structure. Nature Struct. Biol., 4, 1039–1046.CrossRefGoogle ScholarPubMed
Scheer, A., Fanelli, F., Costa, T., Benedetti, P. G. and Cotecchia, S. (1997). The activation process of the α1B-adrenergic receptor: potential role of protonation and hydrophobicity of a highly conserved aspartate. Proc. Natl Acad. Sci., 94, 808–818.CrossRefGoogle ScholarPubMed
Schirmer, T. and Evans, P. R. (1990). Structural basis of the allosteric behavior of phosphofructokinase. Nature, 343, 140–145.CrossRefGoogle ScholarPubMed
Schoppa, N. E., McCormack, K., Tanouye, M. A. and Sigworth, F. J. (1992). The size of the gating charge in wild-type and mutant Shaker potassium channels. Science, 255, 1712–1715. [Cited in Chapter 1. This is an excellent synthesis of the charge and steepness in a voltage-induced protein transition.]CrossRefGoogle ScholarPubMed
Schultz, P. G. and Lerner, R. A. (1995). From molecular diversity to catalysis: Lessons from the immune system. Science, 269: 1835–1842.CrossRefGoogle ScholarPubMed
Schumaker, M. F. and MacKinnon, R. (1990). A simple model for multi-ion permeation: single vacancy conduction in a simple pore model. Biophys. J., 58, 975–984.CrossRefGoogle Scholar
Segev, I. (1990). Computer study of presynaptic inhibition controlling the spread of action potentials into nerve terminals. J. Neurophys., 63, 987–997.CrossRefGoogle Scholar
Segev, I., Fleshman, J. W. and Burke, R. E. (1989). Compartmental models of complex neurons. In Methods in Neural Modeling, ed. Koch, C. and Segev, I.. Cambridge: MIT Press, pp. 63–96.Google Scholar
Serrano, L., Matouschek, A. and Fersht, A. R. (1992). The folding of an enzyme III. Structure of the transition state of barnase analysed by a protein engineering procedure. J. Mol. Biol., 224, 805–818.CrossRefGoogle ScholarPubMed
Setlow, R. B., and Pollard, E. C. (1962). Molecular Biophysics, chapter 6. Palo Alto: Addison-Wesley Publishing Co. Inc. [Cited in Chapter 2. This forgotten text contains a lucid presentation of molecular forces.]Google Scholar
Shi, Z., Krantz, B. A., Kallenbach, N. and Sosnick, T. R. (2002a). Contribution of hydrogen bonding to protein stability estimated from isotope effects. Biochemistry, 41, 2120–2129.CrossRefGoogle Scholar
Shi, Z., Olson, C. A. and Kallenbach, N. R. (2002b). Cation–π interaction in model α-helical peptides. J. Amer. Chem. Soc., 124, 3284–3291.CrossRefGoogle Scholar
Shi, Z., Olson, C. A., Rose, G. D., Baldwin, R. L. and Kallenbach, N. R. (2002c). Polyproline II structure in a sequence of seven alanine residues. Proc. Natl Acad. Sci., 99, 9190–9195.CrossRefGoogle Scholar
Shoup, D., Lipari, G. and Szabo, A. (1981). Diffusion-controlled bimolecular reaction rates. Biophys. J., 36, 697–714.CrossRefGoogle ScholarPubMed
Sigworth, F. J. (1994). Voltage gating of ion channels. Q. Rev. Biophys., 27, 1–27.CrossRefGoogle ScholarPubMed
Silverman, D. N. (2000). Marcus rate theory applied to enzymatic proton transfer. Biochim. Biophys. Acta, 1458, 88–103.CrossRefGoogle ScholarPubMed
Silverman, D. N., Tu, C., Chen, X., Tanhauser, S. M., Kresge, A. J. and Laipis, P. J. (1993). Rate-equilibria relationships in intramolecular proton transfer in human carbonic anhydrase III. Biochemistry, 32, 10 757–10 761.CrossRefGoogle ScholarPubMed
Silverman, J. A., Balakrishnan, R. and Harbury, P. B. (2001). Reverse engineering the (β/α)8 barrel fold. Proc. Natl Acad Sci., 98, 3092–3097.CrossRefGoogle ScholarPubMed
Sine, S. M., Claudio, T. and Sigworth, F. (1990). Activation of Torpedo acetylcholine receptors expressed in mouse fibroblasts. J. Gen. Physiol., 96, 395–437.CrossRefGoogle ScholarPubMed
Spolar, R. S. and Record, M. T. (1994). Coupling of local folding to site-specific binding of proteins to DNA. Science, 263, 777–784. [Cited in Chapter 4. This is a study that considers many different contributions to the free energy of association.]CrossRefGoogle Scholar
Spolar, R. S., Livingstone, J. R. and Record, M. T. (1992). Use of liquid hydrocarbon and amide transfer data to estimate contributions to thermodynamic functions of protein folding from the removal of nonpolar and polar surfaces from water. Biochemistry, 31, 3947–3955.CrossRefGoogle Scholar
Steinberg, I. Z. (1987). Relationship between statistical properties of single ion channel recordings and thermodynamic state of the channels. J. Theor. Biol., 124, 71–87.CrossRefGoogle ScholarPubMed
Steinberg, I. Z. and Scheraga, H. A. (1963). Entropy changes accompanying association reactions of proteins. J. Biol. Chem., 238, 172–181.Google ScholarPubMed
Stigter, D. and Dill, K. A. (1990). Charge effects on folded and unfolded proteins. Biochemistry, 29, 1262–1271.CrossRefGoogle ScholarPubMed
Stillinger, F. H. (1980). Water revisited. Science, 209, 451–457.CrossRefGoogle ScholarPubMed
Stockbridge, N. (1988). Etiology of the supernormal period. Biophys. J., 54, 777–780.CrossRefGoogle ScholarPubMed
Stuart, G. J. and Sakmann, B. (1994). Active propagation of somatic action potentials into neocortical pyramidal cell dendrites. Nature, 367, 69–72.CrossRefGoogle ScholarPubMed
Stuart, G., Spruston, N. and Hausser, M. (2000). Dendrites. Oxford: Oxford University Press.Google Scholar
Stuehmer, W., Conti, F., Suzuki, H.et al. (1989). Structural parts involved in activation and inactivation of the sodium channel. Nature, 339, 597–603.CrossRefGoogle Scholar
Sukharev, S., Blount, P., Martinac, B. and Kung, C. (1997). Mechanosensitive channels of Escherichia coli: the MscL gene, protein, and activities. Ann. Rev. Physiol., 59, 633–657.CrossRefGoogle ScholarPubMed
Sukharev, S., Durell, S. R. and Guy, H. R. (2001). Structural models of the MscL gating mechanism. Biophys. J., 81 (2), 917–936.CrossRefGoogle ScholarPubMed
Sussman, J. L., Harel, M., Frolow, F.et al. (1991). Atomic structure of acetylcholinesterase from Torpedo californica: a prototypic acetylcholine-binding protein. Science, 253, 872–879.CrossRefGoogle ScholarPubMed
Swadlow, H. A., Kocsis, J. D. and Waxman, S. G. (1980). Modulation of impulse conduction along the axonal tree. Ann. Rev. Biophys. Bioeng., 9, 143–179.CrossRefGoogle ScholarPubMed
Szabo, A. and Karplus, M. (1972). A mathematical model for structure-function relations in hemoglobin. J. Mol. Biol., 72, 163–197. [Cited in Chapter 5. This is an important theoretical effort to relate ideas of allosteric regulation to a more detailed picture of protein structure.]CrossRefGoogle ScholarPubMed
Tainer, J. A., Getzoff, E. D., Richardson, J. S. and Richardson, D. C. (1983). Structure and mechanism of copper, zinc superoxide dismutase. Nature, 306, 284–287.CrossRefGoogle ScholarPubMed
Tanford, C. (1955). Hydrogen ion titration curves of proteins. In Electrochemistry in Biology and Medicine (ed. Shedlovsky, T.). New York: John Wiley & Sons, pp. 248–265.Google Scholar
Tanford, C.(1961). Physical Chemistry of Macromolecules. New York: John Wiley & Sons.Google Scholar
Tanford, C.(1968). Protein denaturation. Adv. Protein Chem., 23, 121–282.CrossRefGoogle ScholarPubMed
Tanford, C.(1970). Protein denaturation Part C. Theoretical models for the mechanisms of denaturation. Adv. Protein Chem., 24, 1–95.CrossRefGoogle Scholar
Tauc, L. (1962). Site of origin and propagation of spike in the giant neuron of Aplysia. J. Gen. Physiol., 45, 1077–1097.CrossRefGoogle ScholarPubMed
Terada, S., Kinjo, M. and Hirokawa, N. (2000). Oligomeric tubulin in large transporting complex is transported via kinesin in squid giant axons. Cell, 103, 141–155.CrossRefGoogle ScholarPubMed
Tian, F. and Cross, T. A. (1999). Cation transport: an example of structural based selectivity. J. Mol. Biol., 285, 1993–2003.CrossRefGoogle ScholarPubMed
Tidor, B. and Karplus, M. (1994). The contribution of vibrational entropy to molecular association: the dimerization of insulin. J. Mol. Biol., 238, 405–414.CrossRefGoogle ScholarPubMed
Tucek, S. (1997). Is the R and R∗ dichotomy real? TIPS, 18, 414–416.Google Scholar
Tytgat, J. and Hess, P. (1992). Evidence for cooperative interactions in potassium channel gating. Nature, 359, 420–423.CrossRefGoogle ScholarPubMed
Ussing, H. H. (1949). The distinction by means of tracers between active transport and diffusion. Acta Physiologica Scand., 19, 43–56.CrossRefGoogle Scholar
Kampen, N. (1981). Stochastic Processes in Physics and Chemistry. New York: North Holland.Google Scholar
Wallace, B. A. (1990). Gramicidin channels and pores. Ann. Rev. Biophys. Biophys. Chem., 19, 127–157.CrossRefGoogle ScholarPubMed
Wang, W., Donini, O., Reyes, C. M. and Kollman, P. A. (2001). Biomolecular simulations: Recent developments in force fields, simulations of enzyme catalysis, protein–ligand, protein–protein, and protein–nucleic acid noncovalent interactions. Ann. Rev. Biophys. Biomol. Struct., 30, 211–243.CrossRefGoogle ScholarPubMed
Waxman, S. G. and Swadlow, H. A. (1977). The conduction properties of axons in central white matter. Progress Neurobiol., 8, 297–324.CrossRefGoogle ScholarPubMed
Warshel, A. and Levitt, M. (1976). Theoretical studies of enzymatic reactions: dielectric, electrostatic and steric stabilization of the carbonium ion in the reaction of lysozyme. J. Mol. Biol., 103, 227–249. [Cited in Chapter 10. A seminal paper using computational methods to investigate the energetics of enzyme catalysis.]CrossRefGoogle Scholar
Weiner, M. C. and White, S. H. (1992). Structure of a fluid dioleoylphosphatidylcholine bilayer determined by joint refinement of x-ray and neutron diffraction data III. Complete structure. Biophys. J., 61, 434–447.CrossRefGoogle Scholar
Weiner, S. J., Kollman, P. A., Nguyen, D. T. and Case, D. A. (1986). An all atom force field for simulations of proteins and nucleic acids. J. Comput. Chem., 7, 230–252.CrossRefGoogle ScholarPubMed
Weinhold, F. (1997). Nature of H-bonding in clusters, liquids, and enzymes: an ab initio, natural bond perspective. J. Mol. Struct. (Theochem)., 398–399, 181–197.CrossRefGoogle Scholar
Wells, T. N. C. and Fersht, A. R. (1986). Use of binding energy in catalysis analyzed by mutagenesis of the tyrosyl-tRNA synthetase. Biochemistry, 25, 1881.CrossRefGoogle ScholarPubMed
Wess, J., Gdula, D. and Brann, M. R. (1990). Site-directed mutagenesis of the m3 muscarinic receptor: identification of a series of threonine and tyrosine residues involved in agonist but not antagonist binding. EMBO J., 10, 3729–3734.Google Scholar
Wilson, E. B., Decius, J. C. and Cross, P. C. (1955). Molecular Vibrations, chapter 8. New York: Dover Publications, Inc.Google Scholar
Wu, N., Mo, Y., Gao, J. and Pai, E. F. (2000). Electrostatic stress in catalysis: structure and mechanism of the enzyme orotidine monophosphate decarboxylase. Proc. Natl Acad. Sci., 97 (5), 2017–2022.CrossRefGoogle ScholarPubMed
Yue, K., Fiebig, K. M., Thomas, P. D., Chan, H. S., Shakhnovich, E. I. and Dill, K. A. (1995). A test of lattice protein folding algorithms. Proc. Natl Acad. Sci., 92, 325–329.CrossRefGoogle ScholarPubMed
Zeltwanger, S., Wang, F., Wang, G. -T., Gilles, K. D. and Hwang, T. -C. (1999). Gating of the cystic fibrosis transmembrane conductance regulator chloride channels by adenosine triphosphate hydrolysis: quantitative analysis of a cyclic gating scheme. J. Gen. Physiol., 113, 541–554.CrossRefGoogle ScholarPubMed
Zerangue, N. and Kavanaugh, M. P. (1996) Flux coupling in a neuronal glutamate transporter. Nature, 383, 634–637.CrossRefGoogle Scholar
Zhang, S. J. and Jackson, M. B. (1995). Properties of the GABAa receptor of rat posterior pituitary nerve terminals. J. Neurophysiol., 73, 1135–1144.CrossRefGoogle ScholarPubMed
Zhong, W., Gallivan, J. P., Zhang, Y., Li, L., Lester, H. A. and Dougherty, D. A. (1998). From ab initio quantum mechanics to molecular neurobiology: a cation–π binding site in the nicotinic receptor. Proc. Natl Acad. Sci., 95, 12 088–12 093.CrossRefGoogle ScholarPubMed
Zimm, B. H. and Bragg, J. K. (1959). Theory of the phase transition between helix and random coil in polypeptide chains. J. Chem. Phys., 31, 526–533.CrossRefGoogle Scholar
Abbott, A. J. and Nelsestuen, G. L. (1988). The collisional limit: an important consideration for membrane-associated enzymes and receptors. Faseb J., 2, 2858–2866.CrossRefGoogle ScholarPubMed
Accardi, A. and Miller, C. (2004). Secondary active transport mediated by a prokaryotic homologue of ClC Cl− channels. Nature, 427, 803–807.CrossRefGoogle ScholarPubMed
Adam, G. and Delbrück, M. (1968). Reduction of dimensionality in biological diffusion processes. In Structural Chemistry and Molecular Biology, ed. Rich, A. and Davidson, N.. San Francisco: Freeman, pp. 198–215.Google Scholar
Adams, D. J., Dwyer, T. M. and Hille, B. (1980). The permeability of endplate channels to monovalent and divalent metal cations. J. Gen. Physiol., 75, 493–510.CrossRefGoogle ScholarPubMed
Adams, S. A. and DeFelice, L. J. (2002). Flux coupling in the human serotonin transporter. Biophys. J., 83, 3268–3282.CrossRefGoogle ScholarPubMed
Ahern, C. A. and Horn, R. (2004). Stirring up controversy with a voltage sensor paddle. TINS, 27, 303–307.Google ScholarPubMed
Aicken, C. C. (1990). Chloride transport across the sarcolemma of vertebrate smooth and skeletal muscle. In Chloride Channels and Carriers in Nerve, Muscle, and Glial Cells, ed. Alvarez-Leefmans, F. J. and Russell, J. M.. New York: Plenum Press, pp. 209–249.CrossRefGoogle Scholar
Aidley, D. J. (1978). The Physiology of Excitable Cells. Cambridge: Cambridge University Press.Google Scholar
Alber, T., Dao-pin, S., Wilson, K., Wozniak, J. A., Cook, S. P. and Matthews, B. W. (1987). Contributions of hydrogen bonds of Thr 157 to the thermodynamic stability of phage T4 lysozyme. Nature, 330, 41–46.CrossRefGoogle ScholarPubMed
Albery, W. J. and Knowles, J. R. (1976). Evolution of enzyme function and the development of catalytic efficiency. Biochemistry, 15, 5631–5640.CrossRefGoogle ScholarPubMed
Aldrich, R. W., Getting, P. A. and Thompson, S. H. (1979). Mechanism of frequency-dependent broadening of molluscan neurone soma spikes. J. Physiol., 291, 531–544.CrossRefGoogle ScholarPubMed
Allen, T. W., Andersen, O. S. and Roux, B. (2004). Energetics of ion conduction through the gramicidin channel. Proc. Natl Acad. Sci., 101, 117–122.CrossRefGoogle ScholarPubMed
Almers, W. and McCleskey, E. W. (1984). The non-selective conductance in calcium channels of frog muscle: calcium selectivity in a single-file pore. J. Physiol., 353, 585–608.CrossRefGoogle Scholar
Andersen, O. S. (1983). Ion movement through gramicidin A channels. Single-channel measurements at very high potentials. Biophys. J., 41, 119–133.CrossRefGoogle ScholarPubMed
Anderson, C. R. and Stevens, C. F. (1973). Voltage-clamp analysis of acetylcholine produced end plate current fluctuations at frog neuromuscular junction. J. Physiol., 235, 655–691.CrossRefGoogle ScholarPubMed
Antzelevitch, C. (2001). Basic mechanisms of reentrant arrhythmias. Curr. Opin. Cardiol., 16, 1–7.CrossRefGoogle ScholarPubMed
Åqvist, J. and Luzhkov, V. (2000). Ion permeation mechanism of the potassium channel. Nature, 404, 881–884.CrossRefGoogle ScholarPubMed
Ashcroft, F. M. (2000). Ion Channels and Disease. San Diego: Academic Press.Google Scholar
Aurora, R., Creamer, T. P., Srinivasan, R. and Rose, G. D. (1997). Local interactions in protein folding: lessons from the α-helix. J. Biol. Chem., 272, 1413–1416.CrossRefGoogle ScholarPubMed
Aveyard, R. and Haydon, D. A. (1973). An Introduction to the Principles of Surface Chemistry. Cambridge: Cambridge University Press, p. 231.Google Scholar
Axe, D. D., Foster, N. W. and Fersht, A. R. (1996). Active barnase variants with completely random hydrophobic cores. Proc. Natl Acad. Sci., 93, 5590–5594.CrossRefGoogle ScholarPubMed
Baldwin, R. L. (1996). How Hofmeister ion interactions affect protein stability. Biophys. J., 71, 2056–2063.CrossRefGoogle ScholarPubMed
Barrett, J. N. and Crill, W. E. (1974). Specific membrane properties of cat motoneurons. J. Physiol., 239, 301–324.CrossRefGoogle Scholar
Bashford, C. L. and Pasternak, C. A. (1986). Plasma membrane potential of some animal cells is generated by ion pumping, not by gradients. Trends Biochem. Sci., 11, 113–116.CrossRefGoogle Scholar
Bass, R. B., Strop, P., Barclay, M. and Rees, D. C. (2002). Crystal structure of Escherichia coli MscS, a voltage-modulated and mechanosensitive channel. Science, 298, 1582–1587.CrossRefGoogle ScholarPubMed
Bedzek, M. J., Bommarito, G. M., Caffrey, M. and Penner, T. L. (1990). Diffuse-double layer at a membrane-aqueous interface measured with X-ray standing waves. Science, 248, 52–56.CrossRefGoogle Scholar
Beece, D., Eisenstein, L., Frauenfelder, H.et al. (1980). Solvent viscosity and protein dynamics. Biochemistry, 19, 5147–5157.CrossRefGoogle ScholarPubMed
Ben-Shaul, A., Ben-Tal, N. and Honig, B. (1996). Statistical thermodynamic analysis of peptide and protein insertion into lipid membranes. Biophys. J., 71, 130–137.CrossRefGoogle ScholarPubMed
Benedek, G. B. and Villars, F. M. H. (2000). Physics with Illustrative Examples from Medicine and Biology. New York: Springer-Verlag.Google Scholar
Benz, R. and Läuger, P. (1977). Transport kinetics of dipicrylamine through lipid bilayer membranes. Biochem. Biophys. Acta, 468, 245–258.CrossRefGoogle ScholarPubMed
Berezin, I. V., Kazanskaya, N. F. and Klyosov, A. A. (1971). Determination of the individual rate constants of α-chymotrypsin-catalyzed hydrolysis with the added nucleophilic agent 1,4-butanediol. FEBS Lett., 15, 121–124.CrossRefGoogle ScholarPubMed
Berg, H. C. (1983). Random Walks in Biology. Princeton: Princeton University Press. [Cited in Chapters 6 and 8. This is an accessible and clear introduction to Brownian motion and diffusion with good biological examples.]Google Scholar
Berg, H. C. and Purcell, E. M. (1977). Physics of chemoreception. Biophys. J., 20, 193–219. [Cited in Chapter 8. A seminal paper that introduces many important concepts in diffusion-limited encounters.]CrossRefGoogle ScholarPubMed
Berg, O. G. and Hippel, P. H. (1985). Diffusion-controlled macromolecular interactions. Ann. Rev. Biophys. Biophys. Chem., 14, 131–160.CrossRefGoogle ScholarPubMed
Berniche, S. and Roux, B. (2000). Molecular dynamics of the KcsA K+ channel in a bilayer membrane. Biophys. J., 78, 2900–2917.CrossRefGoogle Scholar
Betz, S. F., Bryson, J. W. and DeGrado, W. F. (1995). Native-like and structurally characterized designed α-helical bundles. Curr. Opin. Struct. Biol., 5, 457–463.CrossRefGoogle ScholarPubMed
Bezanilla, F. (2000). The voltage sensor in voltage-dependent ion channels. Physiol. Rev., 80, 555–592.CrossRefGoogle ScholarPubMed
Blacklow, S. C., Raines, R. T., Lim, W. A., Zamore, P. D. and Knowles, J. R. (1988). Triosephosphate isomerase catalysis is diffusion controlled. Biochemistry, 27, 1158–1167.CrossRefGoogle ScholarPubMed
Blangy, D., Buc, H. and Monod, J. (1968). Kinetics of the allosteric interactions of phosphofructokinase from Escherichia coli. J. Mol. Biol., 31, 13–35. Cited in Chapter 5. This provides an example of how the MWC theory can synthesize a broad range of observations.]CrossRefGoogle ScholarPubMed
Bloomfield, S. A., Hamos, J. E. and Sherman, S. M. (1987). Passive cable properties and morphological correlates of neurones in the lateral geniculate nucleus of the cat. J. Physiol., 383, 653–692.CrossRefGoogle ScholarPubMed
Bloomfield, V. A., Crothers, D. M. and Tinoco, I. (1974). The Physical Chemistry of Nucleic Acids. New York: Harper and Row, Publishers, Inc.Google Scholar
Boyle, P. J. and Conway, E. J. (1941). Potassium accumulation in muscle and associated changes. J. Physiol., 100, 1–63.CrossRefGoogle ScholarPubMed
Breslow, E. and Gurd, F. R. N. (1962). Reactivity of sperm whale metmyoglobin towards hydrogen ions and p-nitrophenyl acetate. J. Biol. Chem., 237, 371–381.Google ScholarPubMed
Brooks, B. R., Bruccoleri, R. E., Olafson, B. D., States, D. J., Swaminathan, S. and Karplus, M. (1983). CHARMM: A program for macromolecular energy, minimization, and dynamics calculations. J. Comput. Chem., 4, 187–217.CrossRefGoogle Scholar
Brouwer, A. C. and Kirsch, J. F. (1982). Investigation of diffusion-limited rates of chymotrypsin reactions by viscosity variation. Biochemistry, 21, 1302–1307.CrossRefGoogle ScholarPubMed
Bruice, T. C. (1970). Proximity effects and enzyme catalysis. In The Enzymes II, ed. Boyer, P. D.. New York: Academic Press, pp. 217–279.Google Scholar
Bryant, R. G. (1996). The dynamics of water–protein interactions. Ann. Rev. Biophys. Biomol. Struct., 25, 29–53.CrossRefGoogle ScholarPubMed
Cai, M. and Jordan, P. C. (1990). How does vestibule surface charge affect ion conduction and toxin binding in a sodium channel. Biophys. J., 57, 883–891.CrossRefGoogle Scholar
Camacho, C. J. and Thirumalai, D. (1993). Minimum energy compact structures of random sequences of heteropolymers. Phys. Rev. Lett., 71, 2505–2508.CrossRefGoogle ScholarPubMed
Cantor, P. R. and Schimmel, P. R. (1980). Biophysical Chemistry. San Francisco: W. H. Freeman and Co. [Cited in Chapter 3. This thorough three-volume set has excellent chapters on conformational statistics of polymers and the helix–coil transition.]Google Scholar
Cardinale, G. J. and Abeles, R. H. (1968). Purification and mechanism of action of proline racemase. Biochemistry, 7, 3970–3978.CrossRefGoogle ScholarPubMed
Carra, J. H., Murphy, E. C. and Privalov, P. L. (1996). Thermodynamic effects of mutations on the denaturation of T4 lysozyme. Biophys. J., 71, 1994–2001.CrossRefGoogle ScholarPubMed
Carslaw, H. S. and Jaeger, J. C. (1959). Conduction of Heat in Solids. Oxford: Oxford University Press.Google Scholar
Cassidy, C. S., Lin, J. and Frey, P. A. (1997). A new concept for the mechanism of action of chymotrypsin: the role of the low-barrier hydrogen bond. Biochemistry, 36, 4576–4584.CrossRefGoogle ScholarPubMed
Caterall, W. A., Chandy, G. K. and Gutman, G. A. (eds.) (2002). The IUPHAR Compendium of Voltage-Gated Ion Channels. Leeds: IUPHAR Media.Google Scholar
Chan, H. S. and Dill, K. A. (1991). Polymer principles in protein structure and stability. Ann. Rev. Biophys. Biophys. Chem., 20, 447–490. [Cited in Chapter 3. This article draws many interesting connections between basic polymer theory and protein structure.]CrossRefGoogle ScholarPubMed
Chandrasekhar, S. (1943). Stochastic problems in physics and astronomy. Rev. Mod. Phys., 15, 1–89. [Cited in Chapter 6. This is a clear account of Brownian motion, and is also included in an excellent selection of related papers, edited by N. Wax and published by Dover.]CrossRefGoogle Scholar
Changeux, J. P. (1984). Acetylcholine receptor: an allosteric protein. Science, 225, 1335–1345.CrossRefGoogle ScholarPubMed
Chapman, M. L., Donger, H. M. A. and Donger, A. M. J. (1997). Activation-dependent subconductance levels in the drk 1 K channel suggest a subunit basis for ion permeation and gating. Biophys. J., 72, 708–719.CrossRefGoogle Scholar
Checover, S., Nachliel, E., Dencher, N. A. and Gutman, M. (1997). Mechanism of proton entry into the cytoplasmic section of the proton-conducting channel of bacteriorhodopsin. Biochemistry, 36, 13 919–13 928.CrossRefGoogle ScholarPubMed
Chen, Y. -D. and Hill, T. L. (1973). Fluctuations and noise in kinetic systems: Applications to K+ channels in the squid axon. Biophys. J., 13, 1276–1295.CrossRefGoogle Scholar
Chung, S. -H., Allen, T. W., Hoyles, M. and Kuyucak, S. (1999). Permeation of ions across the potassium channel: Brownian dynamics studies. Biophys. J., 77, 2517–2533. [Cited in Chapter 14. A lucid treatment of K+ channel permeation with modern computational methods.]CrossRefGoogle ScholarPubMed
Chung, S. -H., Allen, T. W. and Kuyucak, S. (2002). Conducting-state properties of the KcsA potassium channel from molecular and Brownian dynamics simulations. Biophys. J., 82, 628–645.CrossRefGoogle ScholarPubMed
Cleland, W. W. (1970). Steady state kinetics. In The Enzymes, ed. Boyer, P. D.. New York: Academic Press, Vol. ii, pp. 1–65.Google Scholar
Cleland, W. W., Frey, P. A. and Gerlt, J. A. (1998). The low barrier hydrogen bond in enzymatic catalysis. J. Biol. Chem., 273, 25 529–25 532.CrossRefGoogle ScholarPubMed
Clements, J. D. and Redman, S. J. (1989). Cable properties of cat spinal motoneurons measured by combining voltage clamp, current clamp and intracellular staining. J. Physiol., 409, 63–87.CrossRefGoogle ScholarPubMed
Cohn, E. J. and Edsall, J. T. (1943). Proteins, Amino Acids and Peptides as Ions and Dipolar Ions. New York: Reinhold Publishing Corporation.Google Scholar
Colquhoun, D. and Hawkes, A. G. (1982). On the stochastic properties of bursts of single ion channel openings and of clusters of bursts. Phil. Trans. R. Soc. Lond., B300, 1–59.CrossRefGoogle ScholarPubMed
Colquhoun, D. and Hawkes, A. G.(1995). A Q-matrix cookbook. In Single-Channel Recording, ed. Sakmann, B. and Neher, E.. New York: Plenum, pp. 589–633.Google Scholar
Connelly, P., Ghosaini, L., Hu, C. -Q., Kitamura, S., Tanaka, A. and Sturtevant, J. M. (1991). A differential scanning calorimetric study of the thermal unfolding of seven mutant forms of phage T4 lysozyme. Biochem., 30, 1887–1891.CrossRefGoogle ScholarPubMed
Connor, J. A. and Stevens, C. F. (1971a). Inward and delayed outward membrane current in isolated neural somata under voltage clamp. J. Physiol., 213, 1–19.CrossRefGoogle Scholar
Connor, J. A. and Stevens, C. F.(1971b). Prediction of repetitive firing behavior from voltage clamp data on an isolated neurone soma. J. Physiol., 213, 31–53.CrossRefGoogle Scholar
Cooper, A. (1976). Thermodynamic fluctuations in protein molecules. Proc. Natl Acad. Sci., 73, 2740–2741.CrossRefGoogle ScholarPubMed
Coronado, R., Rosenberg, R. L. and Miller, C. (1980). Ionic selectivity, saturation, and block in a K+-selective channel from sarcoplasmic reticulum. J. Gen. Physiol., 76, 425–446.CrossRefGoogle Scholar
Cox, D. R. and Miller, H. D. (1965). The Theory of Stochastic Processes. New York: Chapman and Hall.Google Scholar
Crank, J. (1975). The Mathematics of Diffusion. Oxford: Oxford University Press.Google Scholar
Crothers, D. M., Drak, J., Kahn, J. D. and Levene, S. D. (1992). DNA bending, flexibility, and helical repeat by cyclization kinetics. Methods Enzymol., 212, 3–29.CrossRefGoogle ScholarPubMed
Cubero, E., Luque, F. J. and Orozco, M. (1998). Is polarization important in cation–π interactions?Proc. Natl Acad. Sci., 95, 5976–5980.CrossRefGoogle ScholarPubMed
Curtis, H. J. and Cole, K. S. (1942). Membrane resting and action potentials from the squid giant axon. J. Cell. Comp. Physiol., 19, 135–144.CrossRefGoogle Scholar
Daggett, V. and Fersht, A. R. (2003). Is there a unifying mechanism for protein folding? TIBS, 28, 18–25.Google Scholar
DeFelice, L. J. (1981). Introduction to Membrane Noise. New York: Plenum Press.CrossRefGoogle Scholar
DeFelice, L. J.(2004). Transporter structure and mechanism. TINS, 27, 352–359.Google ScholarPubMed
Gennes, P. G. (1972). Exponents for the excluded volume problem as derived by the Wilson method. Phys. Letts, 38A, 339–340.CrossRefGoogle Scholar
Lean, A., Stadel, J. M. and Lefkowitz, R. J. (1980). A ternary complex model explains the agonist-specific binding properties of the adenylate cyclase-coupled b-adrenergic receptor. J. Biol. Chem., 255, 7108–7117.Google Scholar
Schutter, E. (1986). Alternative equations for the molluscan ion currents described by Connor and Stevens. Brain Res., 382, 134–138.CrossRefGoogle ScholarPubMed
Dill, K. A. (1990). Dominant forces in protein folding. Biochemistry, 29, 7133–7155. [Cited in Chapter 2. This is an excellent review with an emphasis on the hydrophobic effect.]CrossRefGoogle ScholarPubMed
Dill, D. A., Bromberg, S., Yue, K.et al. (1995). Principles of protein folding – a perspective from simple exact models. Protein Sci., 4, 561–602.CrossRefGoogle ScholarPubMed
Dinner, A. R. and Karplus, M. (2001). Comment on the communication “The key to solving the protein-folding problem lies in an accurate description of the denatured state” by van Gunsteren et al. Angew. Chem. Int. Ed., 40, 4615–4616.3.0.CO;2-H>CrossRefGoogle Scholar
Doyle, D. A., Cabral, J. M., Pfuetzner, R. A.et al. (1998). The structure of the potassium channel: molecular basis of K+ conduction and selectivity. Science, 280, 69–77. [Cited in Chapter 14. The first crystal structure of a selective channel and a major advance.]CrossRefGoogle ScholarPubMed
Dunitz, J. D. (1994). The entropic cost of bound water in crystals and biomolecules. Science, 264, 670. [Cited in Chapter 4. This is a lucid account of the thermodynamics of water association with proteins.]CrossRefGoogle ScholarPubMed
Eaton, W. A., Henry, E. R. and Hofrichter, J. (1991). Application of linear free energy relations to protein conformational changes: the quaternary structural change of hemoglobin. Proc. Natl Acad. Sci., 88, 4472–4475.CrossRefGoogle ScholarPubMed
Edwards, S., Corry, B., Kuyucak, S. and Chung, S. -H. (2002). Continuum electrostatics fails to describe ion permeation in the gramicidin channel. Biophys. J., 83, 1348–1360.CrossRefGoogle ScholarPubMed
Ehrenstein, G., Blumenthal, R., Latorre, R. and Lecar, H. (1974). Kinetics of the opening and closing of individual excitability-inducing material channels in a lipid bilayer. J. Gen. Physiol., 63, 707–721.CrossRefGoogle Scholar
Eigen, M. and Hammes, G. G. (1963). Elementary steps in enzyme reactions. Adv. Enyzmol., 25, 1–38.Google ScholarPubMed
Einstein, A. (1956). Investigations on the Theory of Brownian Movement. New York: Dover. [Cited in Chapter 6. This small book contains five papers published by Einstein from 1905 to 1908. This work is timeless and still worthy of careful study by students of biophysics.]Google Scholar
Eisenman, G. and Horn, R. (1983). Ionic selectivity revisited: the role of kinetic and equilibrium processes in ion permeation through channels. J. Membrane Biol., 76, 197–225.CrossRefGoogle ScholarPubMed
Elson, E. L. and Magde, D. (1974). Fluorescence correlation spectroscopy: I. Conceptual basis. Biopolymers, 13, 1–27.CrossRefGoogle Scholar
Falke, J. J., Drake, S. K., Hazard, A. L. and Peersen, O. B. (1994). Molecular tuning of ion binding to calcium signaling proteins. Q. Rev. Biophys., 27, 219–290.CrossRefGoogle ScholarPubMed
Fersht, A. R. (1987). The hydrogen bond in molecular recognition. Trends Biochem. Sci., 12, 301–304.CrossRefGoogle Scholar
Fersht, A. R.(1998). Structure and Mechanism in Protein Science. New York: Freeman. [Cited in Chapters 7 and 10. This is an excellent resource for kinetics and enzyme mechanisms.]Google Scholar
Fersht, A. R., Shi, J. -P., Knill-Jones, J.et al. (1985). Hydrogen bonding and biological specificity analysed by protein engineering. Nature, 314, 235–238.CrossRefGoogle ScholarPubMed
Fersht, A. R., Itzhaki, L. S., El Masry, N. F., Matthews, J. M. and Otzen, D. E. (1994). Single versus parallel pathways of protein folding and fractional formation of structure in the transition state. Proc. Natl Acad. Sci., 91, 10 426–10 429. [Cited in Chapter 7. This provides a very nice example of how to use Φ plots to address the mechanism of protein unfolding.]CrossRefGoogle ScholarPubMed
Fielder, E. M., Roberts, P. B., Bray, R. C.et al. (1974). The mechanism of action of superoxide dismutase from pulse radiolysis and electron paramagnetic resonance. Biochem. J., 139, 49–60.CrossRefGoogle Scholar
Finkel, A. S. and Redman, S. J. (1983). The synaptic current evoked in cat spinal motoneurons by impulse in single group Ia axons. J. Physiol., 342, 615–632.CrossRefGoogle Scholar
Finkelstein, A. (1987). Water Movement Through Lipid Bilayers, Pores, and Plasma Membranes. New York: Wiley-Interscience.Google Scholar
Finkelstein, A. and Andersen, O. S. (1981). The gramicidin A channel: A review of its permeability characteristics with special reference to the single-file aspect of transport. J. Membrane Biol., 59, 155–171.CrossRefGoogle Scholar
Finkelstein, A. V. and Janin, J. (1989). The price of lost freedom: entropy of bimolecular complex formation. Protein Engineering, 3, 1–3.CrossRefGoogle Scholar
Flory, P. J. (1969). Statistical Mechanics of Chain Molecules. New York: Interscience Publishers.Google Scholar
Fredkin, D. R., Montal, M. and Rice, J. A. (1985). Identification of aggregated Markovian models: application to the nicotinic acetylcholine receptor. In Proceedings of the Berkeley Conference in Honor of Jerzy Neyman and Jack Kiefer, ed. Carn, L. M. and Olshen, R. A.. Wadsworth, Vol. 1, pp. 269–289.Google Scholar
Gallivan, J. P. and Dougherty, D. A. (1999). Cation–π interactions in structural biology. Proc. Natl Acad. Sci., 96, 9459–9464.CrossRefGoogle ScholarPubMed
Gallivan, J. P. and Dougherty, D. A.(2000). A computational study of cation-π interactions vs salt bridges in aqueous media: Implications for protein engineering. J. Amer. Chem. Soc., 122, 870–874.CrossRefGoogle Scholar
Garofoli, S. and Jordan, P. C. (2003). Modeling permeation energetics in the KcsA potassium channel. Biophys. J., 84, 2814–2830.CrossRefGoogle ScholarPubMed
Gavish, B. and Werber, M. M. (1979). Viscosity-dependent structural fluctuations in enzyme catalysis. Biochemistry, 18, 1269–1275.CrossRefGoogle ScholarPubMed
Gentet, L. J., Stuart, G. J. and Clements, J. D. (2000). Direct measurement of specific membrane capacitance in neurons. Biophys. J., 79, 314–320.CrossRefGoogle ScholarPubMed
Gilson, M. K., Given, J. A., Bush, B. L. and McCammon, J. A. (1997). The statistical-thermodynamic basis for computation of binding affinities: a critical review. Biophys. J., 72, 1047–1069. [Cited in Chapter 4. This is an excellent review of the statistical mechanics of molecular associations.]CrossRefGoogle ScholarPubMed
Goldman, L. and Albus, J. S. (1968). Computation of impulse conduction in myelinated fibers; theoretical basis of the velocity-diameter relation. Biophys. J., 8, 596–607.CrossRefGoogle ScholarPubMed
Goldstein, S. S. and Rall, W. (1974). Changes of action potential shape and velocity for changing core conductor geometry. Biophys. J., 14, 731–757.CrossRefGoogle ScholarPubMed
Goychuk, I. and Hänggi, P. (2002). Ion channel gating: a first-passage time analysis of the Kramers type. Proc. Natl Acad. Sci., 99, 3552–3556.CrossRefGoogle ScholarPubMed
Green, W. N., Weiss, L. B. and Andersen, O. S. (1987). Batrachotoxin-modified sodium channels in planar lipid bilayers. J. Gen. Physiol., 89, 841–872.CrossRefGoogle ScholarPubMed
Griko, Y. V. and Privalov, P. L. (1992). Calorimetric study of the heat and cold denaturation of beta-lactoglobulin. Biochemistry, 31, 8810–8815.CrossRefGoogle ScholarPubMed
Grosman, C. (2003). Free-energy landscapes of ion-channel gating are malleable: changes in the number of bound ligands are accompanied by changes in the location of the transition state of the acetylcholine-receptor channels. Biochemistry, 42 (50), 14 977–14 987.CrossRefGoogle ScholarPubMed
Grosman, C., Zhou, M. and Auerbach, A. (2000). Mapping the conformational wave of acetylcholine receptor channel gating. Nature, 403, 773–776. [Cited in Chapter 7. This is an important advance mapping the gating transition of an ion channel and an excellent illustration of the use of Φ plots in biophysics.]CrossRefGoogle ScholarPubMed
Gurney, R. W. (1953). Ionic Processes in Solution. New York: McGraw-Hill.Google Scholar
Gutman, M. and Nachliel, E. (1997). Time-resolved dynamics of proton transfer in proteinous systems. Ann. Rev. Phys. Chem., 48, 329–356.CrossRefGoogle ScholarPubMed
Hagen, S. J. and Eaton, W. A. (1996). Nonexponential structural relaxations in proteins. J. Chem. Phys., 104, 3395–3398.CrossRefGoogle Scholar
Hall, J. E., Mead, C. A. and Szabo, G. (1973). A barrier model for current flow in lipid bilayer membranes. J. Membr. Biol., 11, 75–97. [Cited in Chapter 14. This is an early example of how barriers shape the current–voltage behavior of a membrane.]CrossRefGoogle Scholar
Hammes, G. G. (1978). Principles of Chemical Kinetics. New York: Academic Press.Google Scholar
Han, W. -G., Jalkanen, K. J., Elstner, M. and Suhai, S. (1998). Theoretical study of aqueous N-acetyl-L-alanine N′-methylamine: structures and Raman, VCD, and ROA spectra. J. Phys. Chem. B, 102, 2487–2602.CrossRefGoogle Scholar
Hänggi, P., Talkner, P. and Borkovec, M. (1990). Reaction-rate theory: fifty years after Kramers. Rev. Modern Phys., 62, 251–341. [Cited in Chapter 7. This is an excellent review of modern approaches to rate theory.]CrossRefGoogle Scholar
Hardy, L. W. and Kirsch, J. F. (1984). Diffusion-limited components of reactions catalyzed by bacillus cereus b-lactanase I. Biochemistry, 23, 1275–1282.CrossRefGoogle Scholar
Hecht, S., Schlaer, S. and Pirenne, M. (1942). Energy, quanta, and vision. J. Gen. Physiol., 25, 819–840.CrossRefGoogle ScholarPubMed
Hedstrom, L., Perona, J. J. and Rutter, W. J. (1994). Converting trypsin to chymotrypsin: Residue 172 is a substrate specificity determinant. Biochemistry, 33, 8757–8763.CrossRefGoogle ScholarPubMed
Hess, P. and Tsien, R. W. (1984). Mechanism of ion permeation through calcium channels. Nature, 309, 453–456.CrossRefGoogle ScholarPubMed
Hestrin, S., Nicoll, R. A., Perkel, D. J. and Sah, P. (1990). Analysis of excitatory synaptic action in pyramidal cells using whole-cell recording from rat hippocampal slices. J. Physiol., 422, 203–225.CrossRefGoogle ScholarPubMed
Hill, T. L. (1960). An Introduction to Statistical Thermodynamics. Reading: Addison-Wesley. [This is highly recommended as a resource for statistical mechanics.]
Hille, B. (1977). Ionic basis of resting and action potentials. In Handbook of Physiology. The Nervous System. Cellular Biology of Neurons, ed. Brookhart, J. M. and Mountcastle, V. B.. Bethesda: American Physiological Society, pp. 99–136.Google Scholar
Hille, B.(1991). Ion Channels of Excitable Membranes. Sunderland: Sinauer Associates.Google Scholar
Hille, B. and Schwarz, W. (1978). Potassium channels as multi-ion single-file pores. J. Gen. Physiol., 72, 409–442. [Cited in Chapter 14. A full treatment of single-file models and an excellent summary of their basic properties.]CrossRefGoogle ScholarPubMed
Hines, M. L. and Carnevale, N. T. (1997). The NEURON simulation environment. Neural Computation, 9, 1179–1209.CrossRefGoogle ScholarPubMed
Hodgkin, A. L. (1964). The Conduction of Nervous Impulse. Springfield, IL: Charles Thomas.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 Huxley, A. F. (1952). A quantitative description of membrane current and its application to conduction and excitation in nerve. J. Physiol., 117, 500–544. [Cited in Chapter 16. This is the culminating paper in the seminal series on membrane excitability.]CrossRefGoogle ScholarPubMed
Hodgkin, A. L. and Keynes, R. D. (1955). The potassium permeability of a giant nerve fibre. J. Physiol., 128, 61–88. [Cited in Chapter 14. This is the classical development of the single-file theory of permeation.]CrossRefGoogle ScholarPubMed
Hodgkin, A. L. and Rushton, W. A. H. (1946). The electrical constants of a crustacean nerve fiber. Proc. R. Soc. Lond., B133, 444–479.CrossRefGoogle Scholar
Hodgkin, A. L., Huxley, A. F. and Katz, B. (1952). Measurement of current voltage relations in the membrane of the giant axon of Loligo. J. Physiol., 116, 424–448.CrossRefGoogle ScholarPubMed
Hol, W. G. J., Duijnen, P. T. and Berendsen, H. J. C. (1978). The α-helix dipole and the properties of proteins. Nature, 273, 443–446.CrossRefGoogle ScholarPubMed
Horn, R. and Lange, K. (1983). Estimating kinetic constants from single channel data. Biophys. J., 43, 207–223.CrossRefGoogle ScholarPubMed
Horn, R. and Vandenberg, C. A. (1984). Statistical properties of single sodium channels. J. Gen. Physiol., 84, 505–534.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
Imoto, K., Busch, C., Sakmann, B.et al. (1988). Rings of negatively charged amino acids determine the acetylcholine receptor channel conductance. Nature, 335, 645–648.CrossRefGoogle ScholarPubMed
Itzhaki, L. S., Otzen, D. E. and Fersht, A. R. (1995). The structure of the transition state for protein folding of chymotrypsin inhibitor 2 analysed by protein engineering methods: evidence for a nucleation condensation mechanism for protein folding. J. Mol. Biol., 254, 260–288.CrossRefGoogle ScholarPubMed
Jack, J. J. B. and Redman, S. J. (1971). An electrical description of the motoneuron and its application to the analysis of synaptic potentials. J. Physiol., 215, 321–352.CrossRefGoogle ScholarPubMed
Jack, J. J. B., Noble, D. and Tsien, R. W. (1983). Electric Current Flow in Excitable Cells. Oxford: Oxford University Press. [Cited in Chapters 15 and 16. This is a comprehensive textbook that covers cable theory and excitability very well.]Google Scholar
Jackson, J. D. (1975). Classical Electrodynamics. New York: John Wiley & Sons.Google Scholar
Jackson, M. B. (1985). The stochastic behavior of a many channel membrane system. Biophys. J., 47, 129–137.CrossRefGoogle ScholarPubMed
Jackson, M. B.(1992). Cable analysis with the whole-cell patch clamp: theory and experiment. Biophys. J., 61, 756–766.CrossRefGoogle ScholarPubMed
Jackson, M. B.(1993a). Passive current flow and morphology in the terminal arborizations of the posterior pituitary. J. Neurophysiol., 69, 692–702.CrossRefGoogle Scholar
Jackson, M. B.(1993b). Binding specificity of receptor chimeras revisited. Biophys. J., 63, 1443–1444.CrossRefGoogle Scholar
Jackson, M. B.(1993c). On the time scale and time course of protein conformational changes. J. Chem. Phys., 99, 7253–7259.CrossRefGoogle Scholar
Jackson, M. B.(1994). Single channel currents in the nicotinic receptor: A direct demonstration of allosteric transitions. TIBS, 19, 396–399.Google ScholarPubMed
Jackson, M. B.(1997a). Adding up the energies in the acetylcholine receptor channel: relevance to allosteric theory. In The Nicotinic Acetylcholine Receptor: Current Views and Future Trends, ed. Barrantes, F.. Austin: Landes Bioscience, pp. 61–84.Google Scholar
Jackson, M. B.(1997b). Inversion of Markov processes to determine rate constants from single channel data. Biophys. J., 73, 1382–1394.CrossRefGoogle Scholar
Jackson, M. B.(1998). Allosteric mechanisms in the activation of ligand-gated channels. Biophysics Textbook of the Biophysical Society. Rockville: Biophysical Society. [Cited in Chapter 5. A discussion of allosteric mechanisms using ligand-gated channels to illustrate important concepts.]Google Scholar
Jackson, M. B. and Zhang, S. J. (1995). Action potential propagation and propagation block by GABA in rat posterior pituitary nerve terminals. J. Physiol., 483(3), 597–611.CrossRefGoogle ScholarPubMed
Jackson, M. B., Wong, B. M., Morris, C. E., Lecar, H. and Christian, C. N. (1983). Successive openings of the same acetylcholine receptor channel are correlated in open time. Biophys. J., 42, 109–114.CrossRefGoogle ScholarPubMed
Jackson, M. B., Konnerth, A. and Augustine, G. J. (1991). Action potential broadening and frequency-dependent facilitation of calcium signals in pituitary nerve terminals. Proc. Natl Acad. Sci., 88, 380–384.CrossRefGoogle ScholarPubMed
Jacobson, K., Ishihara, A. and Inman, R. (1987). Lateral diffusion of proteins in membranes. Ann. Rev. Physiol., 49, 163–175.CrossRefGoogle ScholarPubMed
Jeffrey, G. A. and Saenger, W. (1991). Hydrogen Bonding in Biological Structures. Berlin: Springer–Verlag.CrossRefGoogle Scholar
Jencks, W. P. (1975). Binding energy, specificity, and enzyme catalysis: the Circe effect. Adv. Enzymol., 43, 219–410.Google ScholarPubMed
Jencks, W. P. and Carriuolo, J. (1961). General base catalysis of ester hydrolysis. Journal of the American Chemical Society, 83, 1743–1750.CrossRefGoogle Scholar
Jentsch, T. J., Stein, V., Weinreich, F. and Zdebik, A. A. (2002). Molecular structure and physiological function of chloride channels. Physiol. Rev., 82, 503–568.CrossRefGoogle ScholarPubMed
Jones, S. W. (1989). On the resting potential of isolated frog sympathetic neurons. Neuron, 3, 153–161.CrossRefGoogle ScholarPubMed
Jordan, P. C. (1990). Ion-water and ion-polypeptide interactions in a gramicidin-like channel. A molecular dynamics study. Biophys. J., 58, 1133–1156.CrossRefGoogle Scholar
Jordan, P. C., Bacquet, R. J., McCammon, A. J. and Tran, P. (1989). How electrolyte shielding influences the electrical potential in transmembrane ion channels. Biophys. J., 55, 1041–1052.CrossRefGoogle ScholarPubMed
Kallenbach, N. (2001). Breaking open a protein barrel. Proc. Natl Acad. Sci., 98, 2958–2960. [Cited in Chapter 2. This presents a clear presentation of the oil-droplet versus jigsaw puzzle pictures of a protein interior.]CrossRefGoogle Scholar
Kao, J. P. Y. and Tsien, R. Y. (1988). Ca2 + binding kinetics of fura-2 and azo-1 from temperature jump relaxation measurements. Biophys. J., 53, 635–639.CrossRefGoogle ScholarPubMed
Karplus, M. (2002). Molecular dynamics simulations of biomolecules. Acc. Chem. Res., 35, 321–323. [Cited in Chapter 2. This is the editorial for a special issue that covers a wide range of applications.]CrossRefGoogle ScholarPubMed
Katz, B. (1966). Nerve, Muscle, and Synapse. New York: McGraw-Hill.Google Scholar
Kaya, H. and Chan, H. S. (2000). Polymer principles of protein calorimetric two-state cooperativity. Proteins: Structure, Function, and Genetics, 40, 637–661.3.0.CO;2-4>CrossRefGoogle ScholarPubMed
Keizer, J. (1987). Diffusion effects on rapid bimolecular chemical reactions. Chem. Rev., 87, 167–180.CrossRefGoogle Scholar
Kell, M. J. and DeFelice, L. J. (1988). Surface charge near the cardiac inward-rectifier channel measured from single-channel conductance. J. Membrane Biol., 102, 1–10.CrossRefGoogle ScholarPubMed
Kellermayer, M. S. Z., Smith, S. B., Granzier, H. L. and Bustamante, C. (1997). Folding–unfolding transitions in single titin molecules characterized with laser tweezers. Science, 276, 1112–1116.CrossRefGoogle ScholarPubMed
Khanin, R., Parnas, H. and Segel, L. (1994). Diffusion cannot govern the discharge of neurotransmitter in fast synapses. Biophys. J., 67, 966–972.CrossRefGoogle ScholarPubMed
Kijima, S. and Kijima, H. (1987). Statistical analysis of channel current from a membrane patch I. Some stochastic properties of ion channels or molecular systems at equilibrium. J. Theor. Biol., 128, 423–434.CrossRefGoogle Scholar
Kittel, C. (1958). Elementary Statistical Physics. New York: John Wiley & Sons.Google Scholar
Kolinski, A., Godzik, A. and Skolnick, J. (1993). A general method for the prediction of the three dimensional structure and folding pathway of globular proteins: Application to designed helical proteins. J. Chem. Phys., 98, 7420–7433.CrossRefGoogle Scholar
Kolinski, A., Galazka, W. and Skolnick, J. (1996). On the origin of the co-operativity of protein folding: implications from model simulations. Proteins: Structure, Function, and Genetics, 26, 271–287.3.0.CO;2-H>CrossRefGoogle Scholar
Koshland, D. E., Nemethy, G. and Filmer, D. (1966). Comparison of experimental binding data and theoretical models in proteins containing subunits. Biochemistry, 5, 365–384.CrossRefGoogle ScholarPubMed
Kramers, H. A. (1940). Brownian motion in a field of force. Physica, 7, 284–304.CrossRefGoogle Scholar
Kuyucak, S., Andersen, O. S. and Chung, S. -H. (2001). Models of permeation in ion channels. Reports of Progress in Physics, 64, 1427–1472.CrossRefGoogle Scholar
Latorre, R., Labarca, P. and Naranjo, D. (1992). Surface charge effects on ion conduction in ion channels. In Ion Channels (Methods in Enzymology), ed. L. Iverson and B. Rudy, vol. 207, pp. 471–501. [Cited in Chapter 11. A very clear overview of surface charge effects in ion channels.]CrossRef
Läuger, P. (1973). Ion transport through pores: a rate-theory analysis. Biophysica and Biochimica Acta, 311, 423–441. [Cited in Chapter 14. A thorough development of barrier models for permeation.]CrossRefGoogle ScholarPubMed
Lecar, H. and Sachs, F. (1981). Membrane noise analysis. In Excitable Cells in Tissue Culture, ed. Nelson, P. G. and Lieberman, M.. New York: Plenum, pp. 137–172.CrossRefGoogle Scholar
Lee, A. W., Karplus, M., Poyart, C. and Bursaux, E. (1988). Analysis of proton release in oxygen binding by hemoglobin: implications for cooperative mechanism. Biochemistry, 27, 1285–1301.CrossRefGoogle ScholarPubMed
Lesk, A. M., Lo Conte, L. and Hubbard, T. J. P. (2001). Assessment of novel fold targets in CASP4: Predictions of three-dimensional structures, secondary structures, and interresidue contacts. Proteins: Structure, Function, and Genetics, 45(S5), 98–118.CrossRefGoogle Scholar
Levinthal, C. (1968). Are there pathways for protein folding. J. Chem. Phys., 65, 44–45.Google Scholar
Levitt, D. G (1978a). Electrostatic calculations for an ion channel. I. Energy and potential profiles and interactions between ions. Biophys. J., 22, 209–219.CrossRefGoogle Scholar
Levitt, D. G(1978b). Electrostatic calculations for an ion channel II. Kinetic behavior of the gramicidin A channel. Biophys. J., 22, 221–248.CrossRefGoogle Scholar
Levitt, M., Sander, C. and Stern, P. S. (1985). Protein normal-mode dynamics: trypsin inhibitor, crambin, ribonuclease and lysozyme. J. Mol. Biol., 181, 423–447.CrossRefGoogle ScholarPubMed
Levitt, M., Hirshberg, M., Sharon, R. and Daggett, V. (1995). Potential energy function and parameters for simulations of the molecular dynamics of proteins and nucleic acids in solution. Comput. Phys. Communs, 91, 215–231.CrossRefGoogle Scholar
Lewis, C. A. (1979). Ion-concentration dependence of the reversal potential and the single channel conductance of ion channels at the frog neuromuscular junction. J. Physiol., 286, 417–445.CrossRefGoogle ScholarPubMed
Lifson, S. and Roig, A. (1963). On the theory of helix–coil transition in polypeptides. J. Chem. Phys., 34, 1961–1974. [Cited in Chapter 3. This is an elegent and clear development of the mathematical theory of helix–coil transitions.]Google Scholar
Lifson, S. and Warshel, A. (1968). Consistent force field for calculations of conformations, vibrational spectra, and enthalpies of cycloalkane and n-alkane molecules. J. Chem. Phys., 49, 5119–5129.CrossRefGoogle Scholar
Lin, J., Cassidy, C. S. and Frey, P. A. (1998). Correlations of the basicity of His 57 with transition state analogue binding, substrate reactivity, and the strength of the low-barrier hydrogen bond in chymotrypsin. Biochemistry, 37, 11 940–11 948.CrossRefGoogle ScholarPubMed
Linderstrøm-Lang, K. (1924). On the ionization of proteins. Compt. rend. trav. Carlsberg, 17, 1–29. (See also Linderstrøm-Lang, K. (1962) Selected Papers. New York: Academic Press.)Google Scholar
Ma, J. C. and Dougherty, D. A. (1997). The cation–π interaction. Chem. Rev., 97, 1303–1324.CrossRefGoogle Scholar
MacInnes, D. A. (1961). The Principles of Electrochemistry. New York: Dover.Google Scholar
MacKinnon, R., Latorre, R. and Miller, C. (1989). Role of surface electrostatics in the operation of a high-conductance Ca2 +-activated K+ channel. Biochemistry, 28, 8092–8099.CrossRefGoogle ScholarPubMed
Magee, J. C. and Cook, E. P. (2000). Somatic EPSP amplitude is independent of synapse location in hippocampal pyramidal neurons. Nature Neurosci., 3, 895–903.CrossRefGoogle ScholarPubMed
Mainen, Z. F., Joerges, J., Huguenard, J. R. and Sejnowski, T. J. (1995). A model of spike initiation in neocortical pyramidal neurons. Neuron, 15, 1427–1439.CrossRefGoogle ScholarPubMed
Major, G., Evan, J. D. and Jack, J. B. (1993). Solutions for transients in arbitrarily branching cables: I. Voltage recording with a somatic shunt. Biophys. J., 65, 423–449. [Cited in Chapter 15. This paper explores the cable equation in complicated dendritic arbors.]CrossRefGoogle ScholarPubMed
Manning, G. S. (1969). Limiting laws and counterion condensation in polyelectrolyte solutions. I. Colligative properties. J. Chem. Phys., 51, 924–933.CrossRefGoogle Scholar
Manning, G. S.(1978). The molecular theory of polyelectrolyte solutions with applications to the electrostatic properties. Q. Rev. Biophys., 11, 179–246.CrossRefGoogle ScholarPubMed
Marcus, R. A. (1964). Chemical and electrochemical electron-transfer theory. Ann. Rev. Phys. Chem., 15, 155–196.CrossRefGoogle Scholar
Marcus, R. A.(1968). Theoretical relations among rate constants, barriers, and Brønsted slopes of chemical reactions. J. Phys. Chem., 72, 891–899.CrossRefGoogle Scholar
Martinez, M. B., Flickinger, M. C. and Nelsestuen, G. L. (1996). Accurate kinetic modeling of alkaline phosphatase in the Escherichia coli periplasm: implications for enzyme properties and substrate diffusion. Biochemistry, 35, 1179–1186.CrossRefGoogle ScholarPubMed
Marty, A, and Neher, E. (1995). Tight-seal whole-cell recording. In Single-Channel Recording, ed. Sakmann, B. and Neher, E.. New York: Plenum, pp. 31–51.Google Scholar
McCammon, A. J., Wolynes, P. G. and Karplus, M. (1979). Picosecond dynamics of tyrosine side chains in proteins. Biochemistry, 18, 927–942. [Cited in Chapter 10. This is a seminal paper on the internal dynamics of proteins.]CrossRefGoogle ScholarPubMed
McLaughlin, S. (1989). The electrostatic properties of membranes. Ann. Rev. Biophys. Biophys. Chem., 18, 113–136.CrossRefGoogle ScholarPubMed
McQuarrie, D. A., (1976). Statistical Mechanics. New York: Harper & Row. [This is highly recommended as a resource for statistical mechanics.]Google Scholar
Meyer, E., Müller, C. O. and Fromherz, P. (1997). Cable properties of dendrites in hippocampal neurons of the rat mapped by a voltage-sensitive dye. Eur. J. Neurosci., 9, 778–785.CrossRefGoogle ScholarPubMed
Miedema, H. (2002). Surface potentials and the calculated selectivity of ion channels. Biophys. J., 82, 156–159.CrossRefGoogle ScholarPubMed
Migliore, M., Hoffman, D. A., Magee, J. C. and Johnston, D. (1999). Role of an A-type K+ conductance in the back-propagation of action potentials in the dendrites of hippocampal pyramidal neurons. J. Comp. Neurosci., 7, 5–15.CrossRefGoogle ScholarPubMed
Millar, J. A., Barrett, L., Southan, A. P., Page, K. M., Fyffe, R. E. W. and Robertson, B. (2000). A functional role for the two-pore domain potassium channel TASK-1 in cerebellar granule neurons. Proc. Natl Acad. Sci., 97, 3614–3618.CrossRefGoogle ScholarPubMed
Miller, B. G. and Wolfenden, R. (2002). Catalytic proficiency: The unusual case of OMP decarboxylase. Ann. Rev. Biochem., 71, 847–885.CrossRefGoogle ScholarPubMed
Moczydlowski, E., Alvarez, O., Vergara, C. and Latorre, R. (1985). Effect of phospholipid surface charge on the conductance and gating of a Ca2 +-activated K+ channel in planar lipid bilayers. J. Membrane Biol., 83, 273–282.CrossRefGoogle ScholarPubMed
Monod, J., Wyman, J. and Changeux, J. -P. (1965). On the nature of allosteric transitions: A plausible model. J. Mol. Biol., 12, 88–118. [Cited in Chapter 5. This is the seminal paper on allosteric regulation of proteins. A remarkable conceptual advance and still well worth reading.]CrossRefGoogle ScholarPubMed
Moore, W. J. (1972). Physical Chemistry. Englewood Cliffs: Prentice-Hall. [This is highly recommended as a resource for statistical mechanics.]Google Scholar
Morais-Cabral, J. H., Zhou, Y. and MacKinnon, R. (2001). Energetic optimization of ion conduction rate by the K+ selectivity filter. Nature, 414, 37–42.CrossRefGoogle ScholarPubMed
Morris, C. E. and Lecar, H. (1981). Voltage oscillations in the barnacle giant muscle fiber. Biophys. J., 35, 193–213. [Cited in Chapter 16. A clear and illuminating account of membrane oscillations.]CrossRefGoogle ScholarPubMed
Moy, G., Corry, B., Kuyucak, S. and Chung, S. -H. (2000). Tests of continuum theories as models of ion channels. I. Poisson–Boltzmann theory versus Brownian dynamics. Biophys. J., 78, 2349–2363.CrossRefGoogle ScholarPubMed
Myers, J. K. and Pace, C. N. (1996). Hydrogen bonding stabilizes globular proteins. Biophys. J., 71, 2033–2039.CrossRefGoogle ScholarPubMed
Nakajima, Y., Nakajima, S. and Inoue, M. (1988). Pertussis toxin-insensitive G protein mediates substance P-induced inhibition of potassium channels in brain neurons. Proc. Natl Acad. Sci., 85, 3643–3647.CrossRefGoogle ScholarPubMed
Nakatani, H. and Dunford, H. B. (1979). Meaning of diffusion-controlled association rate constants in enzymology. J. Phys. Chem., 83, 2662–2665.CrossRefGoogle Scholar
Neher, E. and Steinbach, J. H. (1978). Local anesthetics transiently block currents through single acetylcholine-receptor channels. J. Physiol., 277, 153–176. [Cited in Chapter 9. This is a striking example of the power of single-channel kinetics in the analysis of mechanisms of drug action.]CrossRefGoogle ScholarPubMed
Nelsestuen, G. L. and Martinez, M. B. (1997). Steady state enzyme velocities that are independent of [enzyme]: an important behavior in many membrane and particle-bound states. Biochemistry, 36, 9081–9086.CrossRefGoogle Scholar
Nolte, H. -J., Rosenberry, T. L. and Neumann, E. (1980). Effective charge on acetylcholinesterase active sites determined from the ionic strength dependence of association rate constants with cationic ligands. Biochemistry, 19, 3705–3711.CrossRefGoogle ScholarPubMed
Noskov, S. Y., Berniche, S. and Roux, B. (2004). Control of ion selectivity in potassium channels by electrostatic and dynamic properties of carbonyl ligands. Nature, 431, 830–834.CrossRefGoogle ScholarPubMed
Obaid, A. L. and Salzberg, B. M. (1996). Micromolar 4-aminopyridine enhances invasion of a vertebrate neurosecretory terminal arborization. J. Gen. Physiol., 107, 353–368.CrossRefGoogle ScholarPubMed
Oberhauser, A. F. and Fernandez, J. M. (1995). Hydrophobic ions amplify the capacitance currents used to measure exocytotic fusion. Biophys. J., 69, 451–459.CrossRefGoogle ScholarPubMed
O'Mara, M., Barry, P. H. and Chung, S. -H. (2003). A model of the glycine receptor deduced from Brownian dynamics studies. Proc. Natl Acad. Sci., 100, 4310–4315.CrossRefGoogle ScholarPubMed
O'Neil, K. T. and DeGrado, W. F. (1990). A thermodynamic scale for the helix-forming tendencies of the commonly occurring amino acids. Science, 250, 646–651.CrossRefGoogle ScholarPubMed
Oosawa, F. (1971). Polyelectrolytes. New York: Marcel Dekker, Inc.Google Scholar
Overbeek, J. Th. G. (1952). The electrochemistry of the double layer. In Colloid Science, Vol. 1, ed. Kruyt, H. R.. Amsterdam: Elsevier Publishing Company, pp. 115–193.Google Scholar
Overbeek, J. T. G. and Wiersema, P. H. (1967). The interpretation of electrophoretic mobilities. In Electrophoresis, ed. Bier, M.. New York: Academic Press, pp. 1–52.Google Scholar
Pace, C. N. (1992). Contributions of the hydrophobic effect to globular protein stability. J. Mol. Biol., 226, 29–35.CrossRefGoogle Scholar
Papazian, D. M., Timpe, L. C., Jan, Y. N. and Jan, L. Y. (1991). Alteration of voltage-dependence of Shaker potassium channel by mutations in the S4 sequence. Nature, 349, 305–310.CrossRefGoogle ScholarPubMed
Parsegian, V. A. (1969). Energy of an ion crossing a low dielectric membrane: Solutions to four relevant electrostatics problems. Nature, 221, 844–846. [Cited in Chapters 2 and 14. This is an early study that defined the basic energetic parameters of permeation.]CrossRefGoogle Scholar
Parsegian, V. A.(1973). Long-range physical forces in the biological milieu. Ann. Rev. Biophys. Bioeng., 2, 221–255.CrossRefGoogle ScholarPubMed
Patlak, C. S. (1960). Derivation of an equation for the diffusion potential. Nature, 188, 944–945.CrossRefGoogle ScholarPubMed
Perkel, D. H., Mulloney, B. and Budelli, R. W. (1981). Quantitative methods for predicting neuronal behavior. Neuroscience, 5, 823–837.CrossRefGoogle Scholar
Perutz, M. F., Wilkenson, A. J., Paoli, M. and Dodson, D. D. (1998). The stereochemical mechanism of the cooperative effects in hemoglobin revisited. Ann. Rev. Biophys. Biomol. Structure, 27, 1–34. [Cited in Chapter 5. This contains a clear and thorough discussion of the evidence favoring a two-state description of hemoglobin.]CrossRefGoogle ScholarPubMed
Peters, R. and Cherry, R. J. (1982). Lateral and rotational diffusion of bacteriorhodopsin in lipid bilayers: experimental test of the Saffman–Delbrück equations. Proc. Natl Acad. Sci., 79, 4317–4321.CrossRefGoogle ScholarPubMed
Piek, T. (1975). Ionic and electrical properties. In Insect Muscle, ed. Usherwood, P. N. R.. New York: Academic Press, pp. 275–336.Google Scholar
Plowman, K. M. (1972). Enzyme Kinetics. New York: McGraw Hill.Google Scholar
Pokarowski, P., Kolinski, A. and Skolnick, J. (2003). A minimal physically realistic protein-like lattice model: designing an energy landscape that ensures all-or-none folding to a unique native state. Biophys. J., 84, 1518–1526.CrossRefGoogle ScholarPubMed
Poland, D. and Scheraga, H. A. (1970). Theory of Helix–Coil Transitions. New York: Academic Press.Google Scholar
Privalov, P. L. (1979). Stability of proteins. Adv. Protein Chem., 33, 167–241.CrossRefGoogle ScholarPubMed
Privalov, P. L.(1982). Stability of proteins: Proteins which do not present a single cooperative system. Adv. Protein Chem., 35, 1–104. [Cited in Chapters 2 and 3. This reference presents thorough reviews of the thermodynamics of thermal transitions in proteins.]CrossRefGoogle ScholarPubMed
Prod'hom, B., Peitrobon, D. and Hess, P. (1987). Direct measurement of proton transfer rates to a group controlling the dihydropyridine-sensitive Ca2 + channel. Nature, 329, 243.CrossRefGoogle ScholarPubMed
Pumphrey, R. J. and Young, J. Z. (1938). The rates of conduction of nerve fibres of various diameters in cephalopods. J. Exp. Biol., 14, 453–466.Google Scholar
Putnam, S. J., Coulson, A. F., Farley, I. R., Ridd Leston, B. and Knowles, J. R. (1972). Specificity and kinetics of triose phosphate isomerase from chicken muscle. Biochem. J., 129, 301–310.Google Scholar
Raleigh, D. P. and DeGrado, W. F. (1992). A de novo designed protein shows a thermally induced transition from a native to a molten globule-like state. J. Amer. Chem. Soc., 114, 10 079–10 081.CrossRefGoogle Scholar
Rall, W. (1959) Branching dendritic trees and motoneuron membrane resistivity. Exp. Neurol., 1, 491–527. [Cited in Chapter 15. This provides a remarkable insight into how to simplify the cable analysis of dendrites.]CrossRefGoogle ScholarPubMed
Rall, W.(1967). Distinguishing theoretical synaptic potentials computed for different soma-dendritic distributions of synaptic input. J. Neurophysiol., 30, 1138–1168.CrossRefGoogle ScholarPubMed
Rall, W.(1969). Time constants and electrotonic length of membrane cylinders and neurons. Biophys. J., 9, 1483–1508. [Cited in Chapter 15. A thorough and clear exposition of practical aspects of cable analysis.]CrossRefGoogle ScholarPubMed
Rall, W.(1977). Core conductor theory and cable properties of neurons. In Handbook of Physiology. The Nervous System. Cellular Biology of Neurons, ed. Brookhart, J. M. and Mountcastle, V. B.. Bethesda: American Physiological Society, pp. 39–97.Google Scholar
Rall, W., Burke, R. E., Smith, T. G., Nelson, P. G. and Frank, K. (1967). Dendritic location of synapses and possible mechanisms for the monosynaptic EPSPs in motoneurons. J. Neurophysiol., 30, 1169–1193.CrossRefGoogle Scholar
Ramachandran, G. N. and Sasisekharan, V. (1968). Conformation of polypeptides and proteins. Adv. Protein Chem., 23, 284–437.Google ScholarPubMed
Ramachandran, G. N., Venkatachalam, C. M. and Krimm, S. (1966). Stereochemical criteria for polypeptide and protein chain conformations. 3. Helical and hydrogen-bonded polypeptide chains. Biophys. J., 6(6), 849–872.CrossRefGoogle ScholarPubMed
Rand, R. P. (1981). Interacting phospholipid bilayers: measured forces and induced structural changes. Ann. Rev. Biophys. Bioengin., 10, 288–314.CrossRefGoogle ScholarPubMed
Rapp, M., Yarom, Y. and Segev, I. (1996). Modeling back propagating action potentials in weakly excitable dendrites of neocortical pyramidal cells. Proc. Natl Acad. Sci., 93, 11 985–11 990.CrossRefGoogle ScholarPubMed
Rashin, A. A. and Honig, B. (1985). Reevaluation of the Born model of ion hydration. J. Phys. Chem., 89, 5588–5593.CrossRefGoogle Scholar
Record, M. T. (1975). Effects of Na+ and Mg++ ions on the helix–coil transition of DNA. Biopolymers, 14, 2137–2158.CrossRefGoogle Scholar
Record, M. T., Mazur, S. J., Melancon, P., Roe, J. -H., Shaner, S. L. and Unger, L. (1981). Double helical DNA: conformations, physical properties, and interactions with ligands. Ann. Rev. Biochem., 50, 997–1024.CrossRefGoogle ScholarPubMed
Redman, S. and Walmsley, B. (1983). The time course of synaptic potentials evoked in cat spinal motoneurons at identified group 1a synapses. J. Physiol., 343, 117–133.CrossRefGoogle Scholar
Reed, A. E. and Weinhold, F. (1991). Natural bond orbital analysis of internal rotation barriers and related phenomena. Isr. J. Chem., 31, 277–285.CrossRefGoogle Scholar
Rees, D. C., DeAntonio, L. and Eisenberg, D. (1989). Hydrophobic organization of membrane proteins. Science, 245, 510–513.CrossRefGoogle ScholarPubMed
Richard, J. P. (1998). The enhancement of enzymatic rate accelerations by Brønsted acid–base catalysis. Biochemistry, 37, 4305–4309.CrossRefGoogle ScholarPubMed
Rigler, R. and Elson, E. L. (eds.) (2001). Fluorescence Correlaton Spectroscopy: Theory and Applications. Berlin: Springer.CrossRefGoogle Scholar
Rogawski, M. A. (1985). The A-current: how ubiquitous a feature of excitable cells is it? Trends Neurosci., 8, 214–219.CrossRefGoogle Scholar
Rose, G. D., Gesolowitz, A. R., Lesser, G. J., Lee, R. H. and Zehfus, M. H. (1985). Hydrophobicity of amino acid residues in globular proteins. Science, 229 (4716), 834–838.CrossRefGoogle ScholarPubMed
Roseman, M. A. (1988). Hydrophobicity of the peptide C = O … H–N hydrogen-bonded group. J. Mol. Biol., 201, 621–623.CrossRefGoogle Scholar
Rosenthal, L., Rabolt, J. F. and Hummel, J. (1982). An investigation of the conformational equilibrium of n-butane in a solvent using Raman spectroscopy. J. Chem. Phys., 76, 817–820.CrossRefGoogle Scholar
Rothberg, B. S. and Magleby, K. L. (2001). Testing for detailed balance (microscopic reversibility) in ion channel gating. Biophys. J., 80, 3025–3026.CrossRefGoogle Scholar
Roux, B. and MacKinnon, R. (1999). The cavity and pore helices in the KcsA K+ channel: electrostatic stabilization of monovalent cations. Science, 285, 100–102.CrossRefGoogle ScholarPubMed
Roux, B., Berniche, S. and Im, W. (2000). Ion channels, permeation, and electrostatics: insight into the function of KcsA. Biochemistry, 39, 13 295–13 306.CrossRefGoogle ScholarPubMed
Rushton, W. A. H. (1951). A theory of the effects of fibre size in medullated nerve. J. Physiol., 115, 101–122.CrossRefGoogle ScholarPubMed
Saffman, P. G. and Delbrück, M. (1975). Brownian motion in biological membranes. Proc. Natl Acad. Sci., 72, 3111–3113.CrossRefGoogle ScholarPubMed
Sanchez, I. C. (1979). Phase transition behavior of the isolated polymer chain. Macromolecules, 12, 980–988.CrossRefGoogle Scholar
Saxton, M. J. and Jacobson, K. (1997). Single-particle tracking: applications to membrane dynamics. Ann. Rev. Biophys. Biomol. Struct., 26, 373–399.CrossRefGoogle ScholarPubMed
Scatchard, G. (1949). The attractions of proteins for small molecules and ions. Ann. N. Y. Acad. Sci., 51, 660–671.CrossRefGoogle Scholar
Schafmeister, C. E., LaPorte, S. L., Miercke, L. J. W. and Stroud, R. M. (1997). A designed four helix bundle protein with native-like structure. Nature Struct. Biol., 4, 1039–1046.CrossRefGoogle ScholarPubMed
Scheer, A., Fanelli, F., Costa, T., Benedetti, P. G. and Cotecchia, S. (1997). The activation process of the α1B-adrenergic receptor: potential role of protonation and hydrophobicity of a highly conserved aspartate. Proc. Natl Acad. Sci., 94, 808–818.CrossRefGoogle ScholarPubMed
Schirmer, T. and Evans, P. R. (1990). Structural basis of the allosteric behavior of phosphofructokinase. Nature, 343, 140–145.CrossRefGoogle ScholarPubMed
Schoppa, N. E., McCormack, K., Tanouye, M. A. and Sigworth, F. J. (1992). The size of the gating charge in wild-type and mutant Shaker potassium channels. Science, 255, 1712–1715. [Cited in Chapter 1. This is an excellent synthesis of the charge and steepness in a voltage-induced protein transition.]CrossRefGoogle ScholarPubMed
Schultz, P. G. and Lerner, R. A. (1995). From molecular diversity to catalysis: Lessons from the immune system. Science, 269: 1835–1842.CrossRefGoogle ScholarPubMed
Schumaker, M. F. and MacKinnon, R. (1990). A simple model for multi-ion permeation: single vacancy conduction in a simple pore model. Biophys. J., 58, 975–984.CrossRefGoogle Scholar
Segev, I. (1990). Computer study of presynaptic inhibition controlling the spread of action potentials into nerve terminals. J. Neurophys., 63, 987–997.CrossRefGoogle Scholar
Segev, I., Fleshman, J. W. and Burke, R. E. (1989). Compartmental models of complex neurons. In Methods in Neural Modeling, ed. Koch, C. and Segev, I.. Cambridge: MIT Press, pp. 63–96.Google Scholar
Serrano, L., Matouschek, A. and Fersht, A. R. (1992). The folding of an enzyme III. Structure of the transition state of barnase analysed by a protein engineering procedure. J. Mol. Biol., 224, 805–818.CrossRefGoogle ScholarPubMed
Setlow, R. B., and Pollard, E. C. (1962). Molecular Biophysics, chapter 6. Palo Alto: Addison-Wesley Publishing Co. Inc. [Cited in Chapter 2. This forgotten text contains a lucid presentation of molecular forces.]Google Scholar
Shi, Z., Krantz, B. A., Kallenbach, N. and Sosnick, T. R. (2002a). Contribution of hydrogen bonding to protein stability estimated from isotope effects. Biochemistry, 41, 2120–2129.CrossRefGoogle Scholar
Shi, Z., Olson, C. A. and Kallenbach, N. R. (2002b). Cation–π interaction in model α-helical peptides. J. Amer. Chem. Soc., 124, 3284–3291.CrossRefGoogle Scholar
Shi, Z., Olson, C. A., Rose, G. D., Baldwin, R. L. and Kallenbach, N. R. (2002c). Polyproline II structure in a sequence of seven alanine residues. Proc. Natl Acad. Sci., 99, 9190–9195.CrossRefGoogle Scholar
Shoup, D., Lipari, G. and Szabo, A. (1981). Diffusion-controlled bimolecular reaction rates. Biophys. J., 36, 697–714.CrossRefGoogle ScholarPubMed
Sigworth, F. J. (1994). Voltage gating of ion channels. Q. Rev. Biophys., 27, 1–27.CrossRefGoogle ScholarPubMed
Silverman, D. N. (2000). Marcus rate theory applied to enzymatic proton transfer. Biochim. Biophys. Acta, 1458, 88–103.CrossRefGoogle ScholarPubMed
Silverman, D. N., Tu, C., Chen, X., Tanhauser, S. M., Kresge, A. J. and Laipis, P. J. (1993). Rate-equilibria relationships in intramolecular proton transfer in human carbonic anhydrase III. Biochemistry, 32, 10 757–10 761.CrossRefGoogle ScholarPubMed
Silverman, J. A., Balakrishnan, R. and Harbury, P. B. (2001). Reverse engineering the (β/α)8 barrel fold. Proc. Natl Acad Sci., 98, 3092–3097.CrossRefGoogle ScholarPubMed
Sine, S. M., Claudio, T. and Sigworth, F. (1990). Activation of Torpedo acetylcholine receptors expressed in mouse fibroblasts. J. Gen. Physiol., 96, 395–437.CrossRefGoogle ScholarPubMed
Spolar, R. S. and Record, M. T. (1994). Coupling of local folding to site-specific binding of proteins to DNA. Science, 263, 777–784. [Cited in Chapter 4. This is a study that considers many different contributions to the free energy of association.]CrossRefGoogle Scholar
Spolar, R. S., Livingstone, J. R. and Record, M. T. (1992). Use of liquid hydrocarbon and amide transfer data to estimate contributions to thermodynamic functions of protein folding from the removal of nonpolar and polar surfaces from water. Biochemistry, 31, 3947–3955.CrossRefGoogle Scholar
Steinberg, I. Z. (1987). Relationship between statistical properties of single ion channel recordings and thermodynamic state of the channels. J. Theor. Biol., 124, 71–87.CrossRefGoogle ScholarPubMed
Steinberg, I. Z. and Scheraga, H. A. (1963). Entropy changes accompanying association reactions of proteins. J. Biol. Chem., 238, 172–181.Google ScholarPubMed
Stigter, D. and Dill, K. A. (1990). Charge effects on folded and unfolded proteins. Biochemistry, 29, 1262–1271.CrossRefGoogle ScholarPubMed
Stillinger, F. H. (1980). Water revisited. Science, 209, 451–457.CrossRefGoogle ScholarPubMed
Stockbridge, N. (1988). Etiology of the supernormal period. Biophys. J., 54, 777–780.CrossRefGoogle ScholarPubMed
Stuart, G. J. and Sakmann, B. (1994). Active propagation of somatic action potentials into neocortical pyramidal cell dendrites. Nature, 367, 69–72.CrossRefGoogle ScholarPubMed
Stuart, G., Spruston, N. and Hausser, M. (2000). Dendrites. Oxford: Oxford University Press.Google Scholar
Stuehmer, W., Conti, F., Suzuki, H.et al. (1989). Structural parts involved in activation and inactivation of the sodium channel. Nature, 339, 597–603.CrossRefGoogle Scholar
Sukharev, S., Blount, P., Martinac, B. and Kung, C. (1997). Mechanosensitive channels of Escherichia coli: the MscL gene, protein, and activities. Ann. Rev. Physiol., 59, 633–657.CrossRefGoogle ScholarPubMed
Sukharev, S., Durell, S. R. and Guy, H. R. (2001). Structural models of the MscL gating mechanism. Biophys. J., 81 (2), 917–936.CrossRefGoogle ScholarPubMed
Sussman, J. L., Harel, M., Frolow, F.et al. (1991). Atomic structure of acetylcholinesterase from Torpedo californica: a prototypic acetylcholine-binding protein. Science, 253, 872–879.CrossRefGoogle ScholarPubMed
Swadlow, H. A., Kocsis, J. D. and Waxman, S. G. (1980). Modulation of impulse conduction along the axonal tree. Ann. Rev. Biophys. Bioeng., 9, 143–179.CrossRefGoogle ScholarPubMed
Szabo, A. and Karplus, M. (1972). A mathematical model for structure-function relations in hemoglobin. J. Mol. Biol., 72, 163–197. [Cited in Chapter 5. This is an important theoretical effort to relate ideas of allosteric regulation to a more detailed picture of protein structure.]CrossRefGoogle ScholarPubMed
Tainer, J. A., Getzoff, E. D., Richardson, J. S. and Richardson, D. C. (1983). Structure and mechanism of copper, zinc superoxide dismutase. Nature, 306, 284–287.CrossRefGoogle ScholarPubMed
Tanford, C. (1955). Hydrogen ion titration curves of proteins. In Electrochemistry in Biology and Medicine (ed. Shedlovsky, T.). New York: John Wiley & Sons, pp. 248–265.Google Scholar
Tanford, C.(1961). Physical Chemistry of Macromolecules. New York: John Wiley & Sons.Google Scholar
Tanford, C.(1968). Protein denaturation. Adv. Protein Chem., 23, 121–282.CrossRefGoogle ScholarPubMed
Tanford, C.(1970). Protein denaturation Part C. Theoretical models for the mechanisms of denaturation. Adv. Protein Chem., 24, 1–95.CrossRefGoogle Scholar
Tauc, L. (1962). Site of origin and propagation of spike in the giant neuron of Aplysia. J. Gen. Physiol., 45, 1077–1097.CrossRefGoogle ScholarPubMed
Terada, S., Kinjo, M. and Hirokawa, N. (2000). Oligomeric tubulin in large transporting complex is transported via kinesin in squid giant axons. Cell, 103, 141–155.CrossRefGoogle ScholarPubMed
Tian, F. and Cross, T. A. (1999). Cation transport: an example of structural based selectivity. J. Mol. Biol., 285, 1993–2003.CrossRefGoogle ScholarPubMed
Tidor, B. and Karplus, M. (1994). The contribution of vibrational entropy to molecular association: the dimerization of insulin. J. Mol. Biol., 238, 405–414.CrossRefGoogle ScholarPubMed
Tucek, S. (1997). Is the R and R∗ dichotomy real? TIPS, 18, 414–416.Google Scholar
Tytgat, J. and Hess, P. (1992). Evidence for cooperative interactions in potassium channel gating. Nature, 359, 420–423.CrossRefGoogle ScholarPubMed
Ussing, H. H. (1949). The distinction by means of tracers between active transport and diffusion. Acta Physiologica Scand., 19, 43–56.CrossRefGoogle Scholar
Kampen, N. (1981). Stochastic Processes in Physics and Chemistry. New York: North Holland.Google Scholar
Wallace, B. A. (1990). Gramicidin channels and pores. Ann. Rev. Biophys. Biophys. Chem., 19, 127–157.CrossRefGoogle ScholarPubMed
Wang, W., Donini, O., Reyes, C. M. and Kollman, P. A. (2001). Biomolecular simulations: Recent developments in force fields, simulations of enzyme catalysis, protein–ligand, protein–protein, and protein–nucleic acid noncovalent interactions. Ann. Rev. Biophys. Biomol. Struct., 30, 211–243.CrossRefGoogle ScholarPubMed
Waxman, S. G. and Swadlow, H. A. (1977). The conduction properties of axons in central white matter. Progress Neurobiol., 8, 297–324.CrossRefGoogle ScholarPubMed
Warshel, A. and Levitt, M. (1976). Theoretical studies of enzymatic reactions: dielectric, electrostatic and steric stabilization of the carbonium ion in the reaction of lysozyme. J. Mol. Biol., 103, 227–249. [Cited in Chapter 10. A seminal paper using computational methods to investigate the energetics of enzyme catalysis.]CrossRefGoogle Scholar
Weiner, M. C. and White, S. H. (1992). Structure of a fluid dioleoylphosphatidylcholine bilayer determined by joint refinement of x-ray and neutron diffraction data III. Complete structure. Biophys. J., 61, 434–447.CrossRefGoogle Scholar
Weiner, S. J., Kollman, P. A., Nguyen, D. T. and Case, D. A. (1986). An all atom force field for simulations of proteins and nucleic acids. J. Comput. Chem., 7, 230–252.CrossRefGoogle ScholarPubMed
Weinhold, F. (1997). Nature of H-bonding in clusters, liquids, and enzymes: an ab initio, natural bond perspective. J. Mol. Struct. (Theochem)., 398–399, 181–197.CrossRefGoogle Scholar
Wells, T. N. C. and Fersht, A. R. (1986). Use of binding energy in catalysis analyzed by mutagenesis of the tyrosyl-tRNA synthetase. Biochemistry, 25, 1881.CrossRefGoogle ScholarPubMed
Wess, J., Gdula, D. and Brann, M. R. (1990). Site-directed mutagenesis of the m3 muscarinic receptor: identification of a series of threonine and tyrosine residues involved in agonist but not antagonist binding. EMBO J., 10, 3729–3734.Google Scholar
Wilson, E. B., Decius, J. C. and Cross, P. C. (1955). Molecular Vibrations, chapter 8. New York: Dover Publications, Inc.Google Scholar
Wu, N., Mo, Y., Gao, J. and Pai, E. F. (2000). Electrostatic stress in catalysis: structure and mechanism of the enzyme orotidine monophosphate decarboxylase. Proc. Natl Acad. Sci., 97 (5), 2017–2022.CrossRefGoogle ScholarPubMed
Yue, K., Fiebig, K. M., Thomas, P. D., Chan, H. S., Shakhnovich, E. I. and Dill, K. A. (1995). A test of lattice protein folding algorithms. Proc. Natl Acad. Sci., 92, 325–329.CrossRefGoogle ScholarPubMed
Zeltwanger, S., Wang, F., Wang, G. -T., Gilles, K. D. and Hwang, T. -C. (1999). Gating of the cystic fibrosis transmembrane conductance regulator chloride channels by adenosine triphosphate hydrolysis: quantitative analysis of a cyclic gating scheme. J. Gen. Physiol., 113, 541–554.CrossRefGoogle ScholarPubMed
Zerangue, N. and Kavanaugh, M. P. (1996) Flux coupling in a neuronal glutamate transporter. Nature, 383, 634–637.CrossRefGoogle Scholar
Zhang, S. J. and Jackson, M. B. (1995). Properties of the GABAa receptor of rat posterior pituitary nerve terminals. J. Neurophysiol., 73, 1135–1144.CrossRefGoogle ScholarPubMed
Zhong, W., Gallivan, J. P., Zhang, Y., Li, L., Lester, H. A. and Dougherty, D. A. (1998). From ab initio quantum mechanics to molecular neurobiology: a cation–π binding site in the nicotinic receptor. Proc. Natl Acad. Sci., 95, 12 088–12 093.CrossRefGoogle ScholarPubMed
Zimm, B. H. and Bragg, J. K. (1959). Theory of the phase transition between helix and random coil in polypeptide chains. J. Chem. Phys., 31, 526–533.CrossRefGoogle 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
  • Meyer B. Jackson
  • Book: Molecular and Cellular Biophysics
  • Online publication: 24 May 2010
  • Chapter DOI: https://doi.org/10.1017/CBO9780511754869.024
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
  • Meyer B. Jackson
  • Book: Molecular and Cellular Biophysics
  • Online publication: 24 May 2010
  • Chapter DOI: https://doi.org/10.1017/CBO9780511754869.024
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
  • Meyer B. Jackson
  • Book: Molecular and Cellular Biophysics
  • Online publication: 24 May 2010
  • Chapter DOI: https://doi.org/10.1017/CBO9780511754869.024
Available formats
×