Hostname: page-component-7c8c6479df-94d59 Total loading time: 0 Render date: 2024-03-29T04:38:13.661Z Has data issue: false hasContentIssue false

The Hofmeister effect and the behaviour of water at interfaces

Published online by Cambridge University Press:  17 March 2009

Kim D. Collins
Affiliation:
Department of Biochemistry, The Johns Hopkins University, School of Hygiene and Public Health, 615 North Wolfe Street, Baltimore, MD 21205, USA
Michael W. Washabaugh
Affiliation:
Department of Biochemistry, The Johns Hopkins University, School of Hygiene and Public Health, 615 North Wolfe Street, Baltimore, MD 21205, USA

Summary

Starting from known properties of non-specific salt effects on the surface tension at an air–water interface, we propose the first general, detailed qualitative molecular mechanism for the origins of ion-specific (Hofmeister) effects on the surface potential difference at an air–water interface; this mechanism suggests a simple model for the behaviour of water at all interfaces (including water–solute interfaces), regardless of whether the non-aqueous component is neutral or charged, polar or non-polar. Specifically, water near an isolated interface is conceptually divided into three layers, each layer being 1 water-molecule thick. We propose that the solute determines the behaviour of the adjacent first interfacial water layer (I1); that the bulk solution determines the behaviour of the third interfacial water layer (I3), and that both I1 and I3 compete for hydrogen-bonding interactions with the intervening water layer (I2), which can be thought of as a transition layer. The model requires that a polar kosmotrope (polar water-structure maker) interact with I1 more strongly than would bulk water in its place; that a chaotrope (water-structure breaker) interact with I1 somewhat less strongly than would bulk water in its place; and that a non-polar kosmotrope (non-polar water-structure maker) interact with I1 much less strongly than would bulk water in its place.

We introduce two simple new postulates to describe the behaviour of I1 water molecules in aqueous solution. The first, the ‘relative competition’ postulate, states that an I1 water molecule, in maximizing its free energy (—δG), will favour those of its highly directional polar (hydrogen-bonding) interactions with its immediate neighbours for which the maximum pairwise enthalpy of interaction (—δH) is greatest; that is, it will favour the strongest interactions. We describe such behaviour as ‘compliant’, since an I1 water molecule will continually adjust its position to maximize these strong interactions. Its behaviour towards its remaining immediate neighbours, with whom it interacts relatively weakly (but still favourably), we describe as ‘recalcitrant’, since it will be unable to adjust its position to maximize simultaneously these interactions. The second, the ‘charge transfer’ postulate, states that the strong polar kosmotrope–water interaction has at least a small amount of covalent character, resulting in significant transfer of charge from polar kosmotropes to water–especially of negative charge from Lewis bases (both neutral and anionic); and that the water-structuring effect of polar kosmotropes is caused not only by the tight binding (partial immobilization) of the immediately adjacent (I1) water molecules, but also by an attempt to distribute among several water molecules the charge transferred from the solute. When extensive, cumulative charge transfer to solvent occurs, as with macromolecular polyphosphates, the solvation layer (the layer of solvent whose behaviour is determined by the solute) can become up to 5- or 6-water-molecules thick.

We then use the ‘relative competition’ postulate, which lends itself to simple diagramming, in conjunction with the ‘charge transfer’ postulate to provide a new, startlingly simple and direct qualitative explanation for the heat of dilution of neutral polar solutes and the temperature dependence of relative viscosity of neutral polar solutes in aqueous solution. This explanation also requires the new and intriguing general conclusion that as the temperature of aqueous solutions is lowered towards o °C, solutes tend to acquire a non-uniform distribution in the solution, becoming increasingly likely to cluster 2 water molecules away from other solutes and surfaces (the driving force for this process being the conversion of transition layer water to bulk water). The implications of these conclusions for understanding the mechanism of action of general (gaseous) anaesthetics and other important interfacial phenomena are then addressed.

Type
Research Article
Copyright
Copyright © Cambridge University Press 1985

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

REFERENCES

1Acharya, A. S. & Sussman, L. G. (1984). The reversibility of the ketoamine linkages of aldoses with proteins. J. biol. Chem. 259, 43724378.CrossRefGoogle ScholarPubMed
2Adams, D. M., Blandamer, M. J., Symons, M. C. R. & Waddington, D. (1971). Solvation spectra. Part 39. Infra-red and Raman studies of aqueous and non-aqueous solutions containing per-chlorates. Trans. Faraday Soc. 67, 611617.CrossRefGoogle Scholar
3Adams, P. A. (1980). The interpretation of entropy/enthalpy compensation phenomena in the deacylation of acyl-α-chymotrypsins. Biochem. J. 191, 653655.CrossRefGoogle ScholarPubMed
4Adams, R. L. P., Burdon, R. H., Campbell, A. M. & Smellie, R. M. S. (1976). Davidson's The Biochemistry of the Nucleic Acids, 8th ed., p. 70. New York: Academic Press.Google Scholar
5Adamson, A. W. (1982). Physical Chemistry of Surfaces, 4th ed.New York: John Wiley.Google Scholar
6Agustsson, I. & Strom, A. R. (1981). Biosynthesis and turnover of trimethylamine oxide in the teleost cod, Gadus morhua. J. biol. Chem. 256, 80458049.CrossRefGoogle ScholarPubMed
7Ahmad, F. & McPhie, P. (1978). Thermodynamics of the denaturation of pepsinogen by urea. Biochemistry 17, 241246.CrossRefGoogle ScholarPubMed
8Ahmad, F. & McPhie, P. (1980). The intrinsic viscosity of glyco-proteins. Int. J. Biochem. 11, 9196.CrossRefGoogle Scholar
9Aird, R. B. (1983). The importance of seizure-inducing factors in the control of refractory forms of epilepsy. Epilepsia 24, 567583.CrossRefGoogle ScholarPubMed
10Alber, T., Hartman, F. C., Johnson, R. M., Petsko, G. A. & Tsernoglou, D. (1981). Crystallization of yeast triose phosphate isomerase from polyethylene glycol. Protein crystal formation following phase separation. J. biol. Chem. 256, 13561361.CrossRefGoogle ScholarPubMed
11Albert, A. & Serjeant, E. P. (1984). The Determination of Ionization Constants. A Laboratory Manual, 3rd ed., p. 163. New York: Chapman and Hall.CrossRefGoogle Scholar
12Alexander, R., Ko, E. C. F., Parker, A. J. & Broxton, T. J.(1968). Solvation of ions. XIV. Protic-dipolar aprotic solvent effects on rates of bimolecular reactions. Solvent activity coefficients of reactants and transition states at 25°. J. Am. chem. Soc. 90, 50495069.CrossRefGoogle Scholar
13Alkana, R. L., Finn, D. A., Galleisky, G. G., Syapin, P. J. & Malcolm, R. D. (1985). Ethanol withdrawal in mice precipitated and exacerbated by hyperbaric exposure. Science 229, 772774.CrossRefGoogle ScholarPubMed
14Alkana, R. L. & Malcolm, R. D. (1981). Low-level hyperbaric ethanol antagonism in mice. Dose and pressure response. Pharmacology 22, 199208.CrossRefGoogle ScholarPubMed
15Allen, A. (1983). Mucus – a protective secretion of complexity. Trends Biochem. Sci. 8, 169173.CrossRefGoogle Scholar
16Aloni, Y., Delmer, D. P. & Benziman, M. (1982). Achievement of high rates of in vitro synthesis of 1,4-β-D-glucan: activation by cooperative interaction of the Acetobacter xylinum enzyme system with GTP, polyethylene glycol, and a protein factor. Proc. natn. Acad. Sci. U.S.A. 79, 64486452.CrossRefGoogle Scholar
17Altekar, W. (1975). Fluorescence studies of the interactions of neutral salts with proteins. Indian J. biochem. Biophys. 12, 397399.Google Scholar
18Altekar, W. (1977 a). Fluorescence of proteins in aqueous neutral salt solutions. I. Influence of anions. Biopolymers 16, 341368.CrossRefGoogle ScholarPubMed
19Altekar, W. (1977 b). Fluorescence of proteins in aqueous neutral salt solutions. II. Influence of monovalent cation chlorides, particularly cesium chloride. Biopolymers 16, 369386.CrossRefGoogle ScholarPubMed
20Anderson, P. M. (1981). Purification and properties of the glutamine-and N–acetyl-L-glutamate-dependent carbamoyl phosphate synthetase from liver of Sgualus acanthias. J. biol. Chem. 256, 1222812235.CrossRefGoogle ScholarPubMed
21Anderson, R. G. & Symons, M. C. R. (1969). Solvation spectra. Part 28. N.m.r. studies of aqueous tertiary butyl alcohol in the presence and absence of various solutes. Trans. Faraday Soc. 65, 25502558.CrossRefGoogle Scholar
22Angell, C. A. (1983). Supercooled water. Ann. Rev. Phys. Chem. 34, 593630.CrossRefGoogle Scholar
23Aniya, Y. & Matsusaki, K. (1983). Effects of neutral salts on hepatic microsomal drug-metabolizing enzyme system in rats. Japan J. Pharmacol. 33, 647653.CrossRefGoogle ScholarPubMed
24Arakawa, T. & Timasheff, S. N. (1982 a). Stabilization of protein structure by sugars. Biochemistry 21, 65366544.CrossRefGoogle ScholarPubMed
25Arakawa, T. & Timasheff, S. N. (1982 b). Preferential interactions of proteins with salts in concentrated solutions. Biochemistry 21, 65456552.CrossRefGoogle ScholarPubMed
26Arakawa, T. & Timasheff, S. N. (1983). Preferential interactions of proteins with solvent components in aqueous amino acid solutions. Arch. biochem. Biophys. 224, 169177.CrossRefGoogle ScholarPubMed
27Arakawa, T. & Timasheff, S. N. (1984 a). The mechanism of action of Na glutamate, lysine HC1, and piperazine-N, N′ -bis(2-ethanesulphonic acid) in the stabilization of tubulin and microtubule formation. J. biol. Chem. 259, 49794986.CrossRefGoogle Scholar
28Arakawa, T. & Timasheff, S. N. (1984 b). Mechanism of protein salting in and salting out by divalent cation salts: balance between hydration and salt binding. Biochemistry 23, 59125923.CrossRefGoogle ScholarPubMed
29Arakawa, T. & Timasheff, S. N. (1984 c). Protein stabilization and destabilization by guanidinium salts. Biochemistry 23, 59245929.CrossRefGoogle ScholarPubMed
30Arakawa, T. & Timasheff, S. N. (1985 a). Mechanism of poly(ethylene glycol) interaction with proteins. Biochemistry 24, 67566762.CrossRefGoogle ScholarPubMed
31Arakawa, T. & Timasheff, S. N. (1985 b). Theory of protein solubility. Methods Enzymol. 114, 4977.CrossRefGoogle ScholarPubMed
32Argos, P., Rossmann, M. G., Grau, U. M., Zuber, H., Frank, G. & Tratschin, J. D. (1979). Thermal stability and protein structure. Biochemistry 18, 56985703.Google Scholar
33Arnett, E. M., Bentrude, W. G., Burke, J. J. & Duggleby, P. McC. (1965). Solvent effects in organic chemistry. V. Molecules, ions, and transition states in aqueous ethanol. J. Am. chem. Soc. 87, 15411553.CrossRefGoogle Scholar
34Arnett, E. McC., Duggleby, P. McC. & Burke, J. J. (1963). The importance of ground state solvation in the solvolysis of t–butyl chloride. J. Am. chem. Soc. 85, 13501352.CrossRefGoogle Scholar
35Arnett, E. M. & McKelvey, D. R. (1965). A large solvation enthalpy effect in highly aqueous t–butyl alcohol solutions. J. Am. chem. Soc. 87, 13931394.CrossRefGoogle Scholar
36Arnett, E. M. & McKelvey, D. R. (1966). A large enhancement of solvation enthalpy by the reduction of temperature. J. Am. chem. Soc. 88, 50315033.CrossRefGoogle Scholar
37Arshadi, M. & Kebarle, P. (1970). Hydration of OH and O2 in the gas phase. Comparative solvation of OH by water and the hydrogen halides. Effects of acidity. J. phys. Chem. 74, 14831485.CrossRefGoogle Scholar
38Arshadi, M., Yamdagni, R. & Kebarle, P. (1970). Hydration of the halide negative ions in the gas phase. II. Comparison of hydration energies for the alkali positive and halide negative ions. J. phys. Chem. 74, 14751482.CrossRefGoogle Scholar
39Asahina, E. (1965). Freezing process and injury in isolated animal cells. Fedn Proc. 24, S183–S187.Google ScholarPubMed
40Asakura, T., Adachi, K. & Schwartz, E. (1978). Stabilizing effect of various organic solvents on protein. J. biol. Chem. 253, 64236425.CrossRefGoogle ScholarPubMed
41Ashwood-Smith, M. J. (1967). Radioprotective and cryoprotective properties of dimethyl sulfoxide in cellular systems. Ann. N. Y. Acad. Sci. 141, 4562.CrossRefGoogle ScholarPubMed
42Atha, D. H. & Ingham, K. C. (1981). Mechanism of precipitation of proteins by polyethylene glycols. Analysis in terms of excluded volume. J. biol. Chem. 256, 1210812117.CrossRefGoogle ScholarPubMed
43Aveyard, R. & Lawrence, A. S. C. (1964). Calorimetrie studies on n–aliphatic alcohol + water and n–aliphatic alcohol + water detergent systems. Trans. Faraday Soc. 60, 22652278.CrossRefGoogle Scholar
44Aviram, I. (1973 a). The interaction of chaotropic anions with proteins at acid pH. Eur. J. Biochem. 40, 631636.CrossRefGoogle ScholarPubMed
45Aviram, I. (1973 b). The interaction of chaotropic anions with acid ferricytochrome c. J. biol. Chem. 248, 18941896.CrossRefGoogle ScholarPubMed
46Avron, M. (1986). The osmotic components of halotolerant algae. Trends Biochem. Sci. 11, 56.CrossRefGoogle Scholar
47Back, J. F., Oakenfull, D. & Smith, M. B. (1979). Increased thermal stability of proteins in the presence of sugars and polyols. Biochemistry 18, 51915196.CrossRefGoogle ScholarPubMed
48Bailey, F. E. Jr. & Callard, R. W. (1959). Some properties of poly(ethylene oxide) in aqueous solution. J. Applied Polymer Science 1, 5662.CrossRefGoogle Scholar
49Baird, J. K., Sandford, P. A. & Cottrell, I. W. (1983). Industrial applications of some new microbial polysaccharides. Biotechnology 1, 778783CrossRefGoogle Scholar
50Baker, E. N. & Dodson, E. J. (1980). Crystallographic refinement of the structure of actinidin at 1·7 Å resolution by fast Fourier least-squares methods. Acta Cryst. A36, 559572.CrossRefGoogle Scholar
51Baker, E. N. & Hubbard, R. E. (1984). Hydrogen bonding in globular proteins. Prog. biophys. molec. Biol. 44, 97179.CrossRefGoogle ScholarPubMed
52Balasubramanian, D. & Mitra, P. (1979). Critical solution temperatures of liquid mixtures and the hydrophobic effect. J. phys. Chem. 83, 27242727.CrossRefGoogle Scholar
53Ball, C. D., Hardt, C. R. & Duddles, W. J. (1943). The influence of sugars on the formation of sulfhydryl groups in heat denaturation and heat coagulation of egg albumin. J. biol. Chem. 151, 163169.CrossRefGoogle Scholar
54Bamberger, S., Seaman, G. V. F., Brown, J. A. & Brooks, D. E. (1984). The partition of sodium phosphate and sodium chloride in aqueous dextran poly(ethylene glycol) two-phase systems. J. Colloid and Interface Sci. 99, 187193.CrossRefGoogle Scholar
55Barclay, D. J. (1968). Chemical softness and specific adsorption at electrodes. J. electroanal. Chem. 19, 318321.CrossRefGoogle Scholar
56Barclay, I. M. & Butler, J. A. V. (1938). The entropy of solution. Trans. Faraday Soc. 34, 14451454.CrossRefGoogle Scholar
57Bark, L. S., Graham, R. J. T. & McCormick, D. (1966). Chromatography of halide ions on thin layers of cellulose. Anal. chim. Acta 35, 268270.CrossRefGoogle Scholar
58Barone, G., Rizzo, E. & Vitagliano, V. (1970). Opposite effect of urea and some of its derivatives on water structure. J. phys. Chem. 74, 22302232.CrossRefGoogle Scholar
59Barrett, E. L. & Kwan, H. S. (1985). Bacterial reduction of Tri-methylamine oxide. Ann. Rev. Microbiol. 39, 131149.CrossRefGoogle Scholar
60Barry, P. H. & Diamond, J. M. (1984). Effects of unstirred layers on membrane phenomena. Physiological Reviews 64, 763872.CrossRefGoogle ScholarPubMed
61Baux, G., Simonneau, M. & Tauc, L. (1979). Transmitter release: ruthenium red used to demonstrate a possible role of sialic acid-containing substrates. J. Physiol. 291, 161178.CrossRefGoogle ScholarPubMed
62Beaudette, N. V., Okabayashi, H. & Fasman, G. D. (1982). Conformational effects of organic solvents on histone complexes. Biochemistry 21, 17651772.CrossRefGoogle ScholarPubMed
63Beauregard, D. V. & Barrett, R. E. (1968). Ultrasonics and water structure in urea solutions. J. chem. Phys. 49, 52415244.CrossRefGoogle Scholar
64Beilinsson, A. (1929). Stabilization of protein solutions to heat with sucrose and glycerol. Biochem. Z. 213, 399405.Google Scholar
65Bell, G. M. & Rangecroft, P. D. (1971). Theory of surface tension for a 2–1 electrolyte solution. Trans. Faraday Soc. 67, 649659.Google Scholar
66Bell, R. P. (1937). Relations between the energy and entropy of solution and their significance. Trans. Faraday Soc. 33, 496501.CrossRefGoogle Scholar
67Bell, R. P. (1973). The Proton in Chemistry, 2nd. ed., p. 82. Ithaca: Cornell Univ. Press.CrossRefGoogle Scholar
68Bello, J. & Bello, H. R. (1962). Evidence from model peptides relating to the denaturation of proteins by lithium salts. Nature 194, 681682.CrossRefGoogle Scholar
69Bello, J., Haas, D. & Bello, H. R. (1966). Interactions of protein-denaturing salts with model amides. Biochemistry 5, 25392548.CrossRefGoogle ScholarPubMed
70Bello, J., Riese, H. C. A. & Vinograd, J. R. (1956). Mechanism of gelation of gelatin. Influence of certain electrolytes on the melting points of gels of gelatin and chemically modified gelatins. J. phys.Chem. 60, 12991306.Google Scholar
71Ben-Amotz, A. & Avron, M. (1983). Accumulation of metabolites by halotolerant algae and its industrial potential. Ann. Rev. Microbiol. 37, 95119.Google Scholar
72Ben-Naim, A. (1967). Solubility and thermodynamics of solution of argon in water-methanol system. J. phys. Chem. 71, 40024007.CrossRefGoogle Scholar
73Ben-Naim, A. (1968). Solubility and thermodynamics of solution of argon in the water–ethylene glycol system. J. phys. Chem. 72, 29983001.CrossRefGoogle Scholar
74Ben-Naim, A. & Baer, S. (1964). Solubility and thermodynamics of solution of argon in water + ethanol system. Trans. Faraday Soc. 60, 17361741.CrossRefGoogle Scholar
75Ben-Naim, A. & Moran, G. (1965). Solubility and thermodynamics of solution of argon in water + p–dioxane system. Trans. Faraday Soc. 61, 821825.Google Scholar
76Bergqvist, M. S. & Forslind, E. (1962). Proton magnetic resonance study of the water lattice distortions in aqueous alkali halide solutions. Acta chem. scand. 16, 20692086.CrossRefGoogle Scholar
77Bernal, J. D. & Fowler, R. H. (1933). A theory of water and ionic solution, with particular reference to hydrogen and hydroxyl ions. J. chem. Phys. 1, 515548.CrossRefGoogle Scholar
78Berry, R. S. & Reimann, C. W. (1963). Absorption spectrum of gaseous F and electron affinities of the halogen atoms. J. chem. Phys. 38, 15401543.CrossRefGoogle Scholar
79Bigelis, R. & Umbarger, H. E. (1975). Purification of yeast α-isopropylmalate isomerase. High ionic strength hydrophobicchromatography. J. biol. Chem. 250, 43154321.CrossRefGoogle Scholar
80Bigelis, R. & Umbarger, H. E. (1976). Yeast α-isopropylmalate isomerase. Factors affecting stability and enzyme activity. J. biol. Chem. 251, 35453552.CrossRefGoogle ScholarPubMed
81Bingham, E. C. (1941). Fluidity of electrolytes. J. phys. Chem. 45, 885903.CrossRefGoogle Scholar
82Bio-Rad, (1984). The Bio-Gel phenyl-5PW column: reversed phase without denaturation. Bulletin 1153, Bio-Rad, Richmond, California.Google Scholar
83Blake, C. C. F., Pulford, W. C. A. & Artymiuk, P. J. (1983). X-ray studies of water in crystals of lysozyme. J. molec. Biol. 167, 693723.CrossRefGoogle ScholarPubMed
84Blandamer, M. J. (1973). Acoustic properties. In Water. A Comprehensive Treatise, vol. 2: Water in Crystalline Hydrates. Aqueous Solutions of Simple Nonelectrolytes (ed. Franks, F.), pp. 495528. New York: Plenum Press.Google Scholar
85Blandamer, M. J., Clarke, D. E., Claxton, T. A., Fox, M. F., Hidden, N. J., Oakes, J., Symons, M. C. R., Verma, G. S. P. & Wootten, M. J. (1967 a). Application of spectroscopic and relaxation techniques to the study of t–butyl alcohol–water mixtures. Chem. Comm. 273274.Google Scholar
86Blandamer, M. J. & Fox, M. F. (1970). Theory and applications of charge-transfer-to-solvent spectra. Chem. Rev. 70, 5993.CrossRefGoogle Scholar
87Blandamer, M. J., Hidden, N. J., Symons, M. C. R. & Treloar, N. C. (1969). Ultrasonic absorption properties of solutions. Part 7. Mixtures of water + substituted alcohols. Trans. Faraday Soc. 65, 18051809.CrossRefGoogle Scholar
88Bockris, J. O'M., Devanathan, M. A. V. & Muller, K. (1963). On the structure of charged interfaces. Proc. R. Soc. Lond. A274, 5579.Google Scholar
89Bode, W. & Schwager, P. (1975). The refined crystal structure of bovine β-trypsin at 1·8 Å resolution. II. Crystallographic refinement, calcium binding site, benzamidine binding site and active site at pH 7·0. J. Molec. Biol. 98, 693717.CrossRefGoogle ScholarPubMed
90Boggs, J. M., Yoong, T. & Hsia, J. C. (1976). Site and mechanism of anesthetic action. I. Effect of anesthetics and pressure on fluidity of spin-labeled lipid vesicles. Mol. Pharmacol. 12, 127135.Google ScholarPubMed
91Bohon, R. L. & Claussen, W. F. (1951). The solubility of aromatic hydrocarbons in water. J. Am. chem. Soc. 73, 15711578.Google Scholar
92Bonner, O. D., Arisman, R. K. & Jumper, C. F. (1977). The effect of organic solutes on water structure. Z. phys. Chem. (Leipzig) 258, 4958.Google Scholar
93Bonner, O. D. & Woolsey, G. B. (1968). The effect of solutes and temperature on the structure of water. J phys. Chem. 72, 899905.CrossRefGoogle Scholar
94Borak, J. (1978). Chromatographic behaviour of inorganic salts on hydroxyethyl methacrylate gels. J. Chromat. 155, 6982.CrossRefGoogle Scholar
95Borejdo, J., Linder, S. & Werber, M. M. (1984). Hydrophobic interaction chromatography of myosin fragments: potential use in purification. Arch. biochem. Biophys. 231, 193201.CrossRefGoogle ScholarPubMed
96Bourgarit, J.-J. (1975). Random immunoglobulin. I. Non-specific adsorption of the fluorescent dye rhodamine B on human immunoglobulin is resolved from fluorescence quenching titration in chao-tropic media. Ann. Immunol. (Inst. Pasteur) 126C, 639652.Google Scholar
97Bower, V. E. & Robinson, R. A. (1963). The thermodynamics of the ternary system: urea–sodium chloride–water at 25°. J. phys. Chem. 67, 15241527.CrossRefGoogle Scholar
98Bradbury, S. L. & Jakoby, W. B. (1972). Glycerol as an enzyme-stabilizing agent: effects on aldehyde dehydrogenase. Proc. natn. Acad. Sci. U.S.A. 69, 23732376.Google Scholar
99Braibanti, A. (ed.) (1979). Bioenergetics and Thermodynamics: Model Systems. Boston: D. Reidel.Google Scholar
100Brandts, J. F. (1964). The thermodynamics of protein denaturation. I. The denaturation of chymotrypsinogen. J. Amer. Chem. Soc. 86, 42914301.CrossRefGoogle Scholar
101Brandts, J. F. (1967). Heat effects on proteins and enzmes. In Thermobiology (ed. Rose, A. H.), pp. 2572. New York: Academic Press.Google Scholar
102Brandts, J. F., Fu, J. & Nordin, J. H. (1970). The low temperature denaturation of chymotrypsinogen in aqueous solution and in frozen aqueous solution. In The Frozen Cell (ed. Wolstenholme, G. E. W. and O'Connor, M.), pp. 189212. Ciba Foundation Symposium, London: J. & A. Churchill.Google Scholar
103Brandts, J. F. & Hunt, L. (1967). The thermodynamics of protein denaturation. III. The denaturation of ribonuclease in water and in aqueous urea and aqueous ethanol mixtures. J. Am. chem. Soc. 89, 48264838.Google Scholar
104Britton, H. G., Rubio, V. & Grisolia, S. (1981). Synthesis of carbamoyl phosphate by carbamoyl phosphate synthetase I in the absence of acetylglutamate. Activation of the enzyme by cryoprotectants. Biochem. biophys. Res. Comm. 99, 11311137.CrossRefGoogle ScholarPubMed
105Brooks, D. E. (1973 a). The effect of neutral polymers on the electrokinetic potential of cells and other charged particles. II. A model for the effect of adsorbed polymer on the diffuse double layer. J. Colloid and Interface Sci. 43, 687699.Google Scholar
106Brooks, D. E. (1973 b). The effect of neutral polymers on the electrokinetic potential of cells and other charged particles. III. Experimental studies on the dextran/erythrocyte system. J. Colloid and Interface Sci. 43, 700713.Google Scholar
107Brooks, D. E., Greig, R. G. & Janzen, J. (1980). Mechanisms of erythrocyte aggregation. In Erythrocyte Mechanics and Blood Flow (ed. Cokelet, G. R., Meiselman, H. J. and Brooks, D. E.), pp. 119140. Kroc. Found. Ser. 13, New York: Alan R. Liss.Google Scholar
108Brooks, D. E. & Seaman, G. V. F. (1973). The effect of neutral polymers on the electrokinetic potential of cells and other charged particles. I. Models for the zeta potential increase. J. Colloid and Interface Sci. 43, 670686.CrossRefGoogle Scholar
109Brown, A. D. (1976). Microbial water stress. Bact. Rev. 40, 803846.CrossRefGoogle ScholarPubMed
110Brown, J. E. & Klee, W. A. (1971). Helix-coil transition of the isolated amino terminus of ribonuclease. Biochemistry 10, 470476.CrossRefGoogle ScholarPubMed
111Brown, R. F. (1962). The linear enthalpy–entropy effect. J. Org. Chem. 27, 30153026.CrossRefGoogle Scholar
112Bruins, E. M. (1932). Numerical definition of the lyotropic series. Proc. Acad. Sci. Amsterdam 35, 107115.Google Scholar
113Bruins, E. M. (1934). The lyotropic number and its elucidation. Ree. trav. chim. 53, 292307.CrossRefGoogle Scholar
114Buchner, E. H. (1934). A quantitative relationship in the lyotropic series. Rec. trav. chim. 53, 288291.Google Scholar
115Buchner, E. H. (1936). Quantitative relations in the lyotropic series. Kolloid-Zeitschrift 75, 19.Google Scholar
116Buchner, E. H. (1942). The lyotropic effect. Chem. Weekblad 39, 402404.Google Scholar
117Bull, H. B. (1981). Protein hydration: a sucrose probe. Arch. biochem. Biophys. 208, 229232.CrossRefGoogle ScholarPubMed
118Bull, H. B. & Breese, K. (1978). Interaction of alcohols with proteins. Biopolymers 17, 21212131.CrossRefGoogle Scholar
119Bush, C. A., Ralapati, S., Matson, G. M., Yamasaki, R. B., Osuga, D. T., Yeh, Y. & Feeney, R. E. (1984). Conformation of the antifreeze glycoprotein of polar fish. Arch. biochem. Biophys. 232, 624631.CrossRefGoogle ScholarPubMed
120Buswell, A. M., Gore, R. C. & Rodebush, W. H. (1939). Effect of ions of the lyotropic series on the infrared absorption spectrum of water. J. phys. Chem. 43, 11811184.CrossRefGoogle Scholar
121Butler, J. A. V. (1937). The energy and entropy of hydration of organic compounds. Trans. Faraday Soc. 33, 229238.CrossRefGoogle Scholar
122Bywater, R. P. & Marsden, N. V. B. (1983). Gel chromatography. In Journal of Chromatography Library 22A, Chromatography Fundamentals and Applications of Chromatographic and Electro-phoretic Methods, part A: Fundamentals and Techniques (ed. Heftmann, E.), pp. A257–A330. New York: Elsevier Scientific Publishing Co.CrossRefGoogle Scholar
123Calhoun, D. B. & Englander, S. W. (1985). Internal protein motions, concentrated glycerol, and hydrogen exchange studied in myoglobin. Biochemistry, 24, 20952100.CrossRefGoogle ScholarPubMed
124Cantwell, F. F. & Puon, S. (1979). Mechanism of Chromatographic retention of organic ions on a nonionic adsorbent. Analyt. Chem. 51, 623632.CrossRefGoogle Scholar
125Caplan, A. I. (1984). Cartilage. Scientific American 251, 8494.CrossRefGoogle ScholarPubMed
126Cargill, R. W. & Morrison, T. J. (1975). Solubility of argon in water + alcohol systems. J. Chem. Soc. Faraday Trans. 1 71, 618624.CrossRefGoogle Scholar
127Carlson, S. S. & Kelly, R. B. (1983). A highly antigenic proteoglycan-like component of cholinergic synaptic vesicles. J. biol. Chem. 258, 1108211091.CrossRefGoogle ScholarPubMed
128Carrelli, A. & Caggia, L. D. (1959). On the dielectric constant of some aqueous solutions. Il Nuovo Cimento 14, 161167.CrossRefGoogle Scholar
129Casassa, E. F. & Eisenberg, H. (1961). Partial specific volumes and refractive index increments in multicomponent systems. J. phys. Chem. 65, 427433.CrossRefGoogle Scholar
130Casassa, E. F. & Eisenberg, H. (1964). Thermodynamic analysis of multicomponent solutions. Adv. Protein Chem. 19, 287395.CrossRefGoogle ScholarPubMed
131Chadwell, H. M. & Asnes, B. (1930). The viscosities of several aqueous solutions of organic substances II. J Am. chem. Soc. 52, 35073518.CrossRefGoogle Scholar
132Chandra, B. R. S., Prakash, V. & Rao, M. S. N. (1986). Partial specific volume and interaction of glycinin with solvent components in urea and guanidine hydrochloride. Int. J. Peptide Protein Res. 27, 138144.CrossRefGoogle Scholar
133Chandrasekhar, J., Spellmeyer, D. C. & Jorgensen, W. L. (1984). Energy component analysis for dilute aqueous solutions of Li+, Na+, F, and Cl ions. J Am. chem. Soc. 106, 903910.CrossRefGoogle Scholar
134Charest, R. & Dunn, A. (1984). Chromatographic separation of choline, trimethylamine, trimethylamine oxide, and betaine from tissues of marine fish. Analyt. Biochem. 136, 421424.CrossRefGoogle ScholarPubMed
135Cheesman, D. F. & King, A. (1940). The electrical double layer in relation to the stabilisation of emulsions with electrolytes. Trans. Faraday Soc. 36, 241247.Google Scholar
136Chien, S. (1980). Aggregation of red blood cells: an electrochemical and colloid chemical problem. In Bioelectrochemistry: ions, surfaces, membranes (ed. Blank, M.), Adv. Chem. Ser. no. 188, pp. 338, Washington, D.C.: American Chemical Society.CrossRefGoogle Scholar
137Chilson, O. P., Costello, L. A. & Kaplan, N. O. (1965). Effects of freezing on enzymes. Fedn Proc. 24, S55–S65.Google ScholarPubMed
138Choppin, G. R. & Buijs, K. (1963). Near-infrared studies of the structure of water. II. Ionic solutions. J. chem. Phys. 39, 20422050.Google Scholar
139Ciferri, A. & Orofino, T. A. (1966). Phase separation of poly-L-proline in salt solutions. J. phys. Chem. 70, 32773285.Google Scholar
140Clark, M. E., Hinke, J. A. M. & Todd, M. E. (1981). Studies on water in barnacle muscle fibres. II. Role of ions and organic solutes in swelling of chemically-skinned fibres. J. exp. Biol. 90, 4363.CrossRefGoogle Scholar
141Clark, M. E. & Zounes, M. (1977). The effects of selected cell osmolytes on the activity of lactate dehydrogenase from the eury-haline polychaete, Nereis succinea. Biol. Bull. 153, 468484.CrossRefGoogle Scholar
142Clarke, G. A. & Taft, R. W. (1962). The effects of aqueous electrolytes on the activity coefficient of t–butyl chloride and of its solvolysis transition state. J. Am. chem. Soc. 84, 22952303.CrossRefGoogle Scholar
143Cohen, A. S., Shirahama, T., Sipe, J. D. & Skinner, M. (1984). Amyloid proteins, precursors, mediator, and enhancer. Adv. Biol. Dis. 1, 8386.Google Scholar
144Combes, D. & Monsan, P. (1984). Effect of polyhydric alcohols on invertase stabilization. Ann. N.Y. Acad. Sci. 434, 6163.Google Scholar
145Comper, W. D. & Laurent, T. C. (1978). Physiological function of connective tissue polysaccharides. Physiol. Reviews 58, 255315.Google Scholar
146Conde, O., Teixeira, J. & Papon, P. (1982). Analysis of sound velocity in supercooled H2O, D2O, and water–ethanol mixtures. J. chem. Phys. 76, 37473753.CrossRefGoogle Scholar
147Conway, B. E. (1966). Electrolyte solutions: solvation and structural aspects. Ann. Rev. phys. Chem. 17, 481528.CrossRefGoogle Scholar
148Conway, B. E. (1975). Ion hydration near air/water interfaces and the structure of liquid surfaces. J. Electroanal. Chem. 65, 491504.CrossRefGoogle Scholar
149Conway, B. E. (1977). The state of water and hydrated ions at interfaces. Adv. in Colloid and Interface Sci. 8, 91212.CrossRefGoogle Scholar
150Conway, B. E. (1978). The evaluation and use of properties of individual ions in solution. J. Sol. Chem. 7, 721770.CrossRefGoogle Scholar
151Conway, B. E. (1981). Ionic Hydration in Chemistry and Biophysics. Studies in Physical and Theoretical Chemistry, vol. 12. New York: Elsevier Scientific Publishing Co.Google Scholar
152Conway, B. E., Desnoyers, J. E. & Smith, A. C. (1964). On the hydration of simple ions and polyions. Phil. Trans. R. Soc. Lond. A 256, 389437.Google Scholar
153Conway, B. E. & Laliberte, L. H. (1968). H2O–D2O isotope effect in partial molal volumes of alkali metal and tetraalkylammonium salts. J. phys. Chem. 72, 43174319.CrossRefGoogle Scholar
154Cordeiro, R. F. & Savarese, T. M. (1984). Reversal by L-cysteine of the growth inhibitory and glutathione-depleting effects of N–Methylformamide and N, N -dimethylformamide. Biochem. biophys. Res. Comm. 122, 798803.Google Scholar
155Costerton, J. W., Irvin, R. T. & Cheng, K.-J. (1981). The bacterial glycocalyx in nature and disease. Ann. Rev. Microbiol. 35, 299324.CrossRefGoogle ScholarPubMed
156Cotton, F. A. & Wilkinson, G. (1980). Advanced Inorganic Chemistry. A Comprehensive Text, p. 543. New York: John Wiley.Google Scholar
157Couture, A. M. & Laidler, K. J. (1956). The partial molal volumes of ions in aqueous solution. I. Dependence on charge and radius. Can. J. Chem. 34, 12091216.Google Scholar
158Couture, A. M. & Laidler, K. J. (1957). The partial molal volumes of ions in aqueous solution. II. An empirical equation for oxy-anions. Can. J. Chem. 35, 207210.CrossRefGoogle Scholar
159Cowie, J. M. G. & Toporowski, P. M. (1961). Association in the binary liquid system dimethyl sulphoxide–water. Can. J. Chem. 39, 22402243.Google Scholar
160Cowley, A. C., Fuller, N. L., Rand, R. P. & Parsegian, V. A. (1978). Measurement of repulsive forces between charged phospholipid bilayers. Biochemistry 17, 31633168.CrossRefGoogle ScholarPubMed
161Cox, B. G. (1973). Free energies, enthalpies, and entropies of transfer of non-electrolytes from water to mixtures of water and dimethyl sulphoxide, water and acetonitrile, and water and dioxan. J. Chem. Soc. Perkin Trans. II, 607610.CrossRefGoogle Scholar
162Cox, W. M. & Wolfenden, J. H. (1934). The viscosity of strong electrolytes measured by a differential method. Proc. R. Soc. Lond. 145, 475488.Google Scholar
163Creighton, T. E. (1980). Role of the environment in the refolding of reduced pancreatic trypsin inhibitor. J. Molec. Biol. 144, 521550.CrossRefGoogle ScholarPubMed
164Cupane, A., Giacomazza, D. & Cordone, L. (1982). Kinetics of thermal denaturation of met-hemoglobin in perturbed solvent: relevance of bulk-electrostatic and hydrophobic interactions. Biopolymers 21, 10811092.CrossRefGoogle ScholarPubMed
165Cupo, P., El-Deiry, W., Whitney, P. L. & Awad, W. M. Jr., (1980). Stabilization of proteins by guanidination. J. bol. Chem. 255, 1082810833.CrossRefGoogle ScholarPubMed
166Damodaran, S. & Kinsella, J. E. (1980). Stabilization of proteins by solvents. Effect of pH and anions on the positive cooperativity of 2-nonanone binding to bovine serum albumin. J. biol. Chem. 255, 85038508.Google Scholar
167Damodaran, S. & Kinsella, J. E. (1981). The effects of neutral salts on the stability of macromolecules. A new approach using a protein-ligand binding system. J. biol. Chem. 256, 33943398.Google Scholar
168Damodaran, S. & Kinsella, J. E. (1983). Dissociation of nucleoprotein complexes by chaotropic salts. FEBS Lett. 158, 5357.CrossRefGoogle ScholarPubMed
169Da Silva, J. J. J. R. F. & Williams, R. J. P. (1976). The uptake of elements by biological systems. Structure and Bonding 29, 67121.CrossRefGoogle Scholar
170Davidson, D. W. (1973). Clathrate hydrates. In Water. A Comprehensive Treatise, vol. 2: Water in Crystalline Hydrates; Aqueous Solutions of Simple Nonelectrolytes (ed. Franks, F.), pp. 115234. New York: Plenum Press.Google Scholar
171Davidson, D. W., Garg, S. K., Gough, S. R., Handa, Y. P., Ratcliffe, C. L., Tse, J. S. & Ripmeester, J. A. (1984). Some structural and thermodynamic studies of clathrate hydrates. J. Inclusion Phenomena 2, 231238.CrossRefGoogle Scholar
172Davidson, S. J. & Jencks, W. P. (1969). The effect of concentrated salt solutions on a merocyanine dye, a Vinylogous amide. J. Am. chem. Soc. 91, 225234.CrossRefGoogle Scholar
173Davies, J., Ormondroyd, S. & Symons, M. C. R. (1971 a). Solvation spectra. Part 41 – absolute proton magnetic resonance shifts for water protons induced by cations and anions in aqueous solutions. Trans. Faraday Soc. 67, 34653473.Google Scholar
174Davies, J., Ormondroyd, S. & Symons, M. C. R. (1971 b). Proton nuclear magnetic resonance shifts for water containing alkylammonium ions. Chem. Comm. 12041205.CrossRefGoogle Scholar
175Davies, P. L., Hough, C., Scott, G. K., Ng, N., White, B. N. & Hew, C. L. (1984). Antifreeze protein genes of the winter flounder. J. biol. Chem. 259, 92419247.CrossRefGoogle ScholarPubMed
176Davson, H. (1940). The influence of the lyotropic series of anions on cation permeability. Biochem. J. 34, 917925.CrossRefGoogle ScholarPubMed
177De Visser, C., Perron, G. & Desnoyers, J. E. (1977). The heat capacities, volumes, and expansibilities of tert–butyl alcohol–water mixtures from 6 to 65 °C. Can. J. Chem. 55, 856862.CrossRefGoogle Scholar
178Debye, P. (1954). The electric field of ions and salting-out [Z. Physik. Chem. 130, 5664 (1927)]; The Collected Papers of Peter J. W. Debye, pp. 366373. New York: Interscience.Google Scholar
179Decastel, M., Bourrillon, R. & Frenoy, J.-P. (1981). Cryo-insolubility of peanut agglutinin. Effect of saccharides and neutral salts. J. biol. Chem. 256, 90039008.CrossRefGoogle ScholarPubMed
180Deguchi, T. (1975). Separation of halide anions by gel chromatography. J. Chromat. 108, 409414.CrossRefGoogle Scholar
181Deguchi, T., Hisanaga, A. & Nagai, H. (1977). Chromatographic behaviour of inorganic anions on a sephadex G-15 column. J. Chromatog. 133, 173179.Google Scholar
182DePuy, C. H., Della, E. W., Filley, J., Grabowski, J. J. & Bier-Baum, V. M. (1983). Absence of an α-effect in the gas-phase nucleophilic reactions of HOO. J. Am. chem. Soc. 105, 24812482.CrossRefGoogle Scholar
183Desnoyers, J. E., Arel, M., Perron, G. & Jolicoeur, C. (1969). Apparent molal volumes of alkali halides in water at 25°. Influence of structural hydration interactions on the concentration dependence. J. phys. Chem. 73, 33463351.Google Scholar
184Desnoyers, J. E. & Jolicoeur, C. (1969). Hydration effects and thermodyamic properties of ions. In Modern Aspects of Electrochemistry, no. 5 (ed. Bockris, J. O'M. and Conway, B. E.), pp. 189. New York: Plenum Press.Google Scholar
185Determann, H. & Walter, I. (1968). Source of aromatic affinity to ‘Sephadex’ dextran gels. Nature 219, 604605.Google Scholar
186DeTrobriand, A., Ceccaldi, M., Monique, H., Marciacq-Rousselot, M. M. & Lucas, M. (1972). Infrared spectroscopic study of the effect of addition of quaternary ammonium salts to water at different temperatures. C. R. Acad. Sci. Paris Sér. C 274, 919922.Google Scholar
187DeVries, A. L. (1983). Antifreeze peptides and glycopeptides in cold-water fishes. Ann. Rev. Physiol. 45, 245260.CrossRefGoogle ScholarPubMed
188Diamond, D. A., Berry, S. J., Jewett, H. J., Eggleston, J. C. & Coffey, D. S. (1982). A new method to assesses metastatic potential of human prostate cancer: relative nuclear roundness. J. Urology 128, 729734.Google Scholar
189DiGuiseppi, J. & Fridovich, I. (1982). Oxygen toxicity in Streptococcus sanguis. The relative importance of Superoxide and hydroxyl radicals. J. biol. Chem. 257, 40464051.CrossRefGoogle ScholarPubMed
190Dimmling, W. & Lange, E. (1951). Heats of dilution and solution of n–propyl alcohol and isopropyl alcohol in water at 25°. Z. Elektro-chem. 55, 322327.Google Scholar
191Di Paolo, T., Kier, L. B. & Hall, L. H. (1977). Molecular connectivity and structure–activity relationship of general anesthetics. Molec. Pharmacol. 13, 3137.Google ScholarPubMed
192Doan, P. E. & Drago, R. S. (1984). Requirements and interpretation of linear free energy relations. J. Am. chem. Soc. 106, 27722774.Google Scholar
193Doebbler, G. F. & Rinfret, A. P. (1962). The influence of protective compounds and cooling and warming conditions on hemolysis of erythrocytes by freezing and thawing. Biochim. biophys. Acta 58, 449458.Google Scholar
194Doebbler, G. F. & Rinfret, A. P. (1965). Rapid freezing of human blood. Physical and chemical considerations of injury and protection. Cryobiology 1, 205211.Google Scholar
195Donovan, J. W. (1977). A study of the baking process by differential scanning calorimetry. J. Sci. Fd. Agric. 28, 571578.CrossRefGoogle Scholar
196Duman, J. & Horwath, K. (1983). The role of hemolymph proteins in the cold tolerance of insects. Ann. Rev. Physiol. 45, 261270.Google Scholar
197Dundee, J. W. (1970). Intravenous ethanol anesthesia: a study of dosage and blood levels. Anesth. Analg. 49, 467475.CrossRefGoogle ScholarPubMed
198Duran, N., Baeza, J., Freer, J., Brunet, J. E., Gonzalez, G. A., Sotomayor, C. P. & Faljoni-Alario, A. (1981). Dimethyl sulfoxide as chemical and biological probe: conformational effect on peroxidase systems. Biochem. biophys. Res. Commun. 103, 131138.CrossRefGoogle ScholarPubMed
199Dwivedi, P. C. & Rao, C. N. R. (1970). Spectroscopic studies of the charge-transfer interactions of halide ions. Spectrochimica acta 26A, 15351543.CrossRefGoogle Scholar
200Merck, E., (1985). Fractogel® TSK butyl. In Fractogel® TSK Polymeric Media for Biochromatography, pp. 3135. Darmstadt: E. Merck.Google Scholar
201Eagland, D. (1975). Nucleic acids, peptides, and proteins. In Water. A Comprehensive Treatise, vol. 4: Aqueous Solutions of Amphiphiles and Macromolecules (ed. Franks, F.), pp. 305518. New York: Plenum Press.Google Scholar
202Eagland, D. & Allen, A. P. (1977). The influence of hydration upon the potential at the shear plane (zeta potential) of a hydrophobic surface in the presence of various electrolytes. J. Colloid and Interface Sci. 58, 230241.Google Scholar
203Earp, H. S., Rubin, R. A., Austin, K. S. & Dy, R. C. (1983). DMSO increases tyrosine residue phosphorylation in membranes from murine erythroleukemia cells. Biochem. biophys. Res. Comm. 112, 413418.CrossRefGoogle ScholarPubMed
204Ebert, G. & Kuroyanagi, Y. (1983). Conformational studies of alternating copoly(Leu-Lys) and copoly(Leu-Orn) in alcohol–water solvent mixtures. Int. J. Biol. Macromol. 5, 106110.CrossRefGoogle Scholar
205Edmond, E. & Ogston, A. G. (1968). An approach to the study of phase separation in ternary aqueous systems. Biochem. J. 109, 569576.Google Scholar
206Edsall, J. T. (1979). The development of the physical chemistry of proteins 1898–1940. Ann. N.Y. Acad. Sci. 325, 5374.CrossRefGoogle ScholarPubMed
207Edsall, J. T. (1981). Edwin J. Cohn and the physical chemistry of proteins. Trends Biochem. Sci. 6, 335337.CrossRefGoogle Scholar
208Edsall, J. T. & Gutfreund, H. (1983). Biothermodynamics. The Study of Biochemical Processes at Equilibrium. New York: John Wiley & Sons.Google Scholar
209Edsall, J. T. & McKenzie, H. A. (1978). Water and proteins. I. The significance and structure of water; its interaction with electrolytes and non-electrolytes. Adv. Biophys. 10, 137207.Google ScholarPubMed
210Edsall, J. T. & McKenzie, H. A. (1983). Water and proteins. II. The location and dynamics of water in protein systems and its relation to their stability and properties. Adv. Biophys. 16, 53183.Google Scholar
211Edsall, J. T. & Wyman, J. (1958). Biophysical Chemistry, vol. 1: Thermodynamics, Electrostatics, and the Biological Significance of the properties of Matter. New York: Academic Press.Google Scholar
212Edwards, J. O. & Pearson, R. G. (1962). The factors determining nucleophilic reactivities. J. Am. chem. Soc. 84, 1624.Google Scholar
213Eftink, M. R., Anusiem, A. C. & Biltonen, R. L. (1983). Enthalpy–Entropy compensation and heat capacity changes for protein–ligand interactions: general thermodynamic models and data for the binding of nucleotides to ribonuclease A. Biochemistry 22, 38843896.Google Scholar
214Egan, B. Z. (1968). Selectivity of polyacrylamide and dextran gels for simple cations and anions. J. Chromatog. 34, 382388.Google Scholar
215Egan, E. P. & Luff, B. B. (1966). Heat of solution, heat capacity, and density of aqueous urea solutions at 25 °C. J. Chem. Eng. Data 11, 192194.Google Scholar
216Eger, E. I. II, (1974). Anesthetic Uptake and Action. Baltimore: Williams & Wilkins Co.Google Scholar
217Eger, E. I. II, Lundgren, C., Miller, S. L. & Stevens, W. C. (1969). Anesthetic potencies of sulfur hexafluoride, carbon tetrafluoride, chloroform and ethrane in dogs: correlation with the hydrate and lipid theories of anesthetic action. Anesthesiology 30, 129135.CrossRefGoogle ScholarPubMed
218Eger, E. I., Saidman, L. J. & Brandstater, B. (1965). Temperature dependence of halothane and cyclopropane anesthesia in dogs: correlation with some theories of anesthetic action. Anesthesiology 26, 764770.CrossRefGoogle ScholarPubMed
219Ehrenfeld, G. M., Francis, T. A. & Hecht, S. M. (1983). Loss of positional specificity in the aminoacylation of Escherichia coli tRNAGLY*. J. biol. Chem. 258, 1174511750.CrossRefGoogle ScholarPubMed
220Eigen, M. (1952). Theory of heat conductivity of aqueous electrolyte solutions. Z. Elektrochem. 56, 836840.Google Scholar
221Eigen, M. & Wicke, E. (1951). Hydration of ions and specific heats of aqueous solutions of electrolytes. Z. Elektrochem. 55, 354363.Google Scholar
222Eisenberg, D. & Kauzmann, W. (1969). The Structure and Properties of Water, p. 218. New York: Oxford University Press.Google Scholar
223Eisenberg, H. (1976). Biological Macromolecules and Polyelectrolytes in Solution. Oxford: Clarendon.Google Scholar
224Eisenberg, H., Haik, Y., Ifft, J. B., Leicht, W., Mevarech, M. & Pundak, S. (1978). Interactions of proteins and nucleic acids with solutes in concentrated solutions of monovalent salts, relating to hydration, structural transitions and inactivation of halophilic malate and glutamate dehydrogenase. In Energetics and Structure of Halophilic Microorganisms (ed. Caplan, S. R. and Ginzburg, M.), pp. 1332. New York: Elsevier/North Holland Biomedical.Google Scholar
225Eisenman, G. (1969). Theory of membrane electrode potentials: an examination of the parameters determining the selectivity of solid and liquid ion exchangers and of neutral ion-sequestering molecules. In Ion-Selective Electrodes (ed. Durst, R. A.), pp. 156. National Bureau of Standards Special Publication 314. Washington, D.C.: U.S. Govt. Printing Off.Google Scholar
226Eley, D. D. & Evans, M. G. (1938). Heats and entropy changes accompanying the solution of ions in water. Trans. Faraday Soc. 34, 10931112.Google Scholar
227Ellerton, H. D. & Dunlop, P. J. (1966). Activity coefficients for the systems water–urea and water–urea–sucrose at 25 °C from isopiestic measurements. J. phys. Chem. 70, 18311837.Google Scholar
228Elworthy, P. H. (1961). The adsorption of water vapour by lecithin and lysolecithin, and the hydration of lysolecithin micelles. J. Chem. Soc. 53855389.CrossRefGoogle Scholar
229Enderby, J. E. & Neilson, G. W. (1980). Structural properties of ionic liquids. Adv. Phys. 29, 323365.Google Scholar
230Endom, L., Hertz, H. G., Thul, B. & Zeidler, M. D. (1967). A microdynamic model of electrolyte solutions as derived from nuclear relaxation and self-diffusion data. Deutsche Bunsenges. Phys. Chem. 71, 10081031.Google Scholar
231Engbersen, J. F. J. & Engberts, J. B. F. N. (1975). Water structure and its kinetic effects on the neutral hydrolysis of two acyl activated esters. J. Amer. Chem. Soc. 97, 15631568.CrossRefGoogle Scholar
232Engberts, J. B. F. N. (1979). Mixed aqueous solvent effects on kinetics and mechanisms of organic reactions. In Water. A Comprehensive Treatise, vol. 6: Recent Advances (ed. Franks, F.), pp. 139237. New York: Plenum Press.CrossRefGoogle Scholar
233Engel, G. & Hertz, H. G. (1968). On the negative hydration. A nuclear magnetic relaxation study. Ber. der Bunsengesellschaft 72, 808834.Google Scholar
234Epton, R., Holloway, C. & McLaren, J. V. (1976). Enzacryl® gel packings. Inorganic salt compatibility, pH stability and ion-exchange derivatives. J. Chromatog. 117, 245255.CrossRefGoogle Scholar
235Eucken, A. (1948). Ion hydrates in aqueous solution. Z. Elektrochem. 52, 624.Google Scholar
236Evans, M. G. & Polanyi, M. (1936). Further considerations on the thermodynamics of chemical equilibria and reaction rates. Trans. Faraday Soc. 32, 13331360.CrossRefGoogle Scholar
237Exner, O. (1964). On the enthalpy–entropy relationship. Collection of Czechoslovak Chem. Comm. 29, 10941113.Google Scholar
238Eyring, H. (1935). The activated complex in chemical reactions. J. Chem. Phys. 3, 107115.CrossRefGoogle Scholar
239Fabricand, B. P. & Goldberg, S. (1961). Proton resonance shifts in alkali halide solutions. J. chem. Phys. 34, 16241628.Google Scholar
240Farquhar, M. G., Courtoy, P. J., Lemkin, M. C. & Kanwar, Y. S. (1982). Current knowledge of the functional architecture of the glomerular basement membrane. In New Trends in Basement Membrane Research (ed. Kuehn, K., Schoene, H. and Timpl, R.), pp. 929. New York: Raven Press.Google Scholar
241Farrah, S. R., Shah, D. O. & Ingram, L. O. (1981). Effects of chaotropic and antichaotropic agents on elution of poliovirus adsorbed on membrane filters. Proc. natn. Acad. Sci. U.S.A. 78, 12291232.CrossRefGoogle ScholarPubMed
242Farrant, J. (1977). Water transport and cell survival in cryobio-logical procedures. Phil. Trans. R. Soc. Lond. B278, 191205.Google Scholar
243Fausnaugh, J. L., Pfannkoch, E., Gupta, S. & Regnier, F. E. (1984). High-performance hydrophobic interaction chromatography of proteins. Anal. Biochem. 137, 464472.Google Scholar
244Feakins, D. & Watson, P. (1963). Studies in ion solvation innon-aqueous solvents and their aqueous mixtures. Part II. Properties of ion constituents. J. Chem. Soc. 47344741.CrossRefGoogle Scholar
245Featherstone, R. M. & Muehlbaecher, C. A. (1963). The current role of inert gases in the search for anesthesia mechanisms. Pharmacological Reviews 15, 97121.Google Scholar
246Feeney, R. E., Osuga, D. T. & Yeh, Y. (1979). Anomalous depression of the freezing temperature by blood-serum proteins of fishes. In Proteins at Low Temperatures (ed. Fennema, O.), pp. 83107. Advances in Chemistry Series, no. 180. Washington, D.C.: American Chemical Society.CrossRefGoogle Scholar
247Feeney, R. E. & Yeh, Y. (1978). Antifreeze proteins from fish bloods. Adv. in Protein Chem. 32, 191282.Google Scholar
248Feierman, D. E. & Cederbaum, A. I. (1983). The effect of EDTA and iron on the oxidation of hydroxyl radical scavenging agents and ethanol by rat liver microsomes. Biochem. biophys. Res. Comm. 116, 765770.CrossRefGoogle ScholarPubMed
249Finer, E. G., Franks, F. & Tait, M. J. (1972). Nuclear magnetic resonance studies of aqueous urea solutions. J. Am. chem. Soc. 94, 44244429.Google Scholar
250Finney, J. L. (1977). The organization and function of water in protein crystals. Phil. Trans. R. Soc. Lond. B278, 332.Google ScholarPubMed
251Finney, J. L. (1979). The organization and function of water in protein crystals. In Water. A Comprehensive Treatise, vol. 6: Recent Advances (ed. Franks, F.), pp. 47122. New York: Plenum Press.Google Scholar
252Fischer, L. (1980). Gel Filtration Chromatography. Laboratory Techniques in Biochemistry and Molecular Biology (ed. Work, T. S. & Burdon, R. H.). New York: Elsevier/North Holland Biomedical Press.Google Scholar
253Fishbein, W. N. & Winkert, J. W. (1979). Parameters of freezing damage to enzymes. In Proteins at Low Temperatures (ed. Fennema, O.), pp. 5582, Advances in Chemistry Series, vol. 180. Washington, D.C.: American Chemical Society.Google Scholar
254Fister, F. & Hertz, H. G. (1967). O17-NMR study of aqueous electrolyte and non-electrolyte solutions. Deutsche Bunsenges. Phys. Chem. 71, 10321040.CrossRefGoogle Scholar
255Flory, P. J. (1949). The configuration of real polymer chains. J. chem. Phys. 17, 303310.Google Scholar
256Flory, P. J. (1953). Principles of Polymer Chemistry. Ithaca: Cornell.Google Scholar
257Flory, P. J. (1969). Statistical Mechanics of Chain Molecules. New York: John Wiley & Sons.Google Scholar
258Formisano, S., Johnson, M. L. & Edelhoch, H. (1978). Effects of Hofmeister salts on the self-association of Glucagon. Biochemistry 17, 14681473.Google Scholar
259Fox, T. G. Jr, & Flory, P. J. (1949). Intrinsic viscosity–molecular weight relationships for polyisobutylene. J. phys. and Colloid Chem. 53, 197212.Google Scholar
260Frank, H. S. & Evans, M. W. (1945). Free volume and entropy in condensed systems. III. Entropy in binary liquid mixtures: partial molal entropy in dilute solutions; structure and thermodynamics in aqueous electrolytes. J. chem. Phys. 13, 507532.Google Scholar
261Frank, H. S. & Robinson, A. L. (1940). The entropy of dilution of strong electrolytes in aqueous solutions. J chem. Phys. 8, 933938.CrossRefGoogle Scholar
262Frank, H. S. & Wen, W.-Y. (1957). III. Ion-solvent interaction. Structural aspects of ion-solvent interaction in aqueous solutions: a suggested picture of water structure. Discuss. of the Faraday Soc. 24, 133140.CrossRefGoogle Scholar
263Franks, F. (1975). The hydrophobic interaction. In Water. A Comprehensive Treatise, vol. 4: Aqueous Solutions of Amphiphiles and Macromolecules (ed. Franks, F.), pp. 194. New York: Plenum Press.CrossRefGoogle Scholar
264Franks, F. (1977). Solvation and conformational effects in aqueous solutions of biopolymer analogues. Phil. Trans. R. Soc. Lond. B 278, 3357.Google ScholarPubMed
265Franks, F. (ed.) (1982 a). Water. A Comprehensive Treatise, vol. 7: Water and Aqueous Solutions at Subzero Temperatures. New York: Plenum Press.Google Scholar
266Franks, F. (1982 b). The nucleation of ice in undercooled aqueous solutions. Cryo-Letters 2, 2731.Google Scholar
267Franks, F. (1982 c). Apparent osmotic activities of water soluble polymers used as cryoprotectants. Cryo-Letters 3, 115120.Google Scholar
268Franks, F. (1985). Biophysics and Biochemistry at Low Temperatures. New York: Cambridge University Press.Google Scholar
269Franks, F. & Eagland, D. (1975). The role of solvent interactions in protein conformation. CRC Critical Reviews in Biochemistry 3, 165219.Google Scholar
270Franks, F. & Ives, D. J. G. (1960). The adsorption of alcohols at hydrocarbon–water interfaces. J. Chem. Soc. 741754.Google Scholar
271Franks, F. & Ives, D. J. G. (1966). The structural properties of alcohol–water mixtures. Q. Rev. 20, 144.CrossRefGoogle Scholar
272Franks, F. & Johnson, H. H. (1962). Accurate evaluation of partial molar properties. Trans. Faraday Soc. 58, 656661.CrossRefGoogle Scholar
273Franks, F., Mathias, S. F., Parsonage, P. & Tang, T. B. (1983). Differential scanning calorimetric study on ice nucleation in water and in aqueous solutions of hydroxyethyl starch. Thermochimica acta 61, 195202.CrossRefGoogle Scholar
274Franks, F. & Reid, D. S. (1973). Thermodynamic properties. In Water. A Comprehensive Treatise, vol. 2: Water in Crystalline Hydrates. Aqueous Solutions of Simple Nonelectrolytes (ed. Franks, F.), pp. 323380. New York: Plenum Press.Google Scholar
275Franks, F. & Smith, H. T. (1968). Volumetric properties of alcohols in dilute aqueous solutions. Trans. Faraday Soc. 64, 29622972.CrossRefGoogle Scholar
276Freed, K. F. (1985). Polymer excluded volume and the renormalization group. Acc. Chem. Res. 18, 3845.CrossRefGoogle Scholar
277Frey, P. A. & Sammons, R. D. (1985). Bond order and charge localization in nucleoside phosphorothioates. Science 228, 541545.Google Scholar
278Fridovich, I. (1963). Inhibition of acetoacetic decarboxylase by anions. J. biol. Chem. 238, 592598.CrossRefGoogle ScholarPubMed
279Fried, M. & Chun, P. W. (1971). Water-soluble nonionic polymers in protein purification. Meth. Enzymol. 22, 238248.CrossRefGoogle Scholar
280Fried, M. G. & Bloomfield, V. A. (1984). DNA gelation in concentrated solutions. Biopolymers 23, 21412155.Google Scholar
281Friedman, H. L. & Krishnan, C. V. (1973). Thermodynamics of ion hydration. In Water. A Comprehensive Treatise, vol. 3: Aqueous Solutions of Simple Electrolytes (ed. Franks, F.), pp. 1118. New York: Plenum Press.Google Scholar
282Friedman, M. E. & Scheraga, H. A. (1965). Volume changes in hydrocarbon–water systems. Partial molal volumes of alcohol–water solutions. J. phys. Chem. 69, 37953800.CrossRefGoogle Scholar
283Frumkin, A. (1924). Phase-boundary forces and adsorption at the interface air: solutions of inorganic salts. Z. physik. Chem. 109, 3448.CrossRefGoogle Scholar
284Frumkin, A., Reichstein, S. & Kulvarskaja, R. (1926). Ion absorption in the water surface. Kolloid-Z, 40, 911.CrossRefGoogle Scholar
285Frumkin, A. N., Yofa, Z. A. & Gerovich, M. A. (1956). On the potential difference at the water–air interface. Zh. Fiz. Khim. 30, 14551468.Google Scholar
286Fujio, M., McLver, R. T. Jr, & Taft, R. W. (1981). Effects on the acidities of phenols from specific substituent–solvent interactions. Inherent substituent parameters from gas-phase acidities. J. Am. chem. Soc. 103, 40174029.CrossRefGoogle Scholar
287Fuller, R. S., Kaguni, J. M. & Kornberg, A. (1981). Enzymatic replication of the origin of the Escherichia coli chromosome. Proc. Natn. Acad. Sci. U.S.A. 78, 73707374.CrossRefGoogle ScholarPubMed
288Fulton, A.B. (1982). How crowded is the cytoplasm? Cell 30, 345347.CrossRefGoogle ScholarPubMed
289Geiger, A., Rahman, A. & Stillinger, F. H. (1979). Molecular dynamics study of the hydration of Lennard–Jones solutes. J. chem. Phys. 70, 263276.Google Scholar
290Geigert, J., DeWitt, S. K., Neidleman, S. L., Lee, G., Dalietos, D. J. & Moreland, M. (1983). DMSO is a substrate for chloro-peroxidase. Biochem. biophys. Res. Comm. 116, 8285.Google Scholar
291Gekko, K. (1982). Mechanism of protein stabilization by polyols:thermodynamics of transfer of amino acids and proteins from water to aqueous polyol solutions. In Ions and Molecules in Solution (ed. Tanaka, N., Ohtaki, H. and Tamamushi, R.), pp. 339358. New York: Elsevier Science Publishers.Google Scholar
292Gekko, K. & Morikawa, T. (1981 a). Preferential hydration of bovine serum albumin in polyhydric alcohol–water mixtures. J. Biochem. 90, 3950.CrossRefGoogle ScholarPubMed
293Gekko, K. & Morikawa, T. (1981 b). Thermodynamics of polyol-induced thermal stabilization of chymotrypsinogen. J. Biochem. 90, 5160.CrossRefGoogle ScholarPubMed
294Gekko, K. & Timasheff, S. N. (1981 a). Mechanism of protein stabilization by glycerol: preferential hydration in glycerol–water mixtures. Biochem. 20, 46674676.Google Scholar
295Gekko, K. & Timasheff, S. N. (1981 b). Thermodynamic and kinetic examination of protein stabilization by glycerol. Biochem. 20, 46774686.CrossRefGoogle ScholarPubMed
296George, H., McMahan, J., Bowler, K. & Elliott, M. (1969). Stabilization of lactate and malate dehydrogenase by organic solvents. Biochim. biophys. Acta 191, 466468.Google Scholar
297George, M. F. & Burke, M. J. (1984). Supercooling of tissue water to extreme low temperature in overwintering plants. Trends in Biochemical Sciences 7, 211214.CrossRefGoogle Scholar
298Gerlsma, S. Y. (1968). Reversible denaturation of ribonuclease in aqueous solutions as influenced by polyhydric alcohols and some other additives. J. biol. Chem. 243, 957961.CrossRefGoogle ScholarPubMed
299Gerlsma, S. Y. (1970). The effects of polyhydric and monohydric alcohols on the heat-induced reversible denaturation of chymotryp-sinogen A. Eur. J. Biochem. 14, 150153.CrossRefGoogle Scholar
300Gerlsma, S. Y. & Stuur, E. R. (1972). The effect of polyhydric and monohydric alcohols on the heat-induced reversible denaturation of lysozyme and ribonuclease. Int. J. Peptide Protein Res. 4, 377383.CrossRefGoogle ScholarPubMed
301Gerlsma, S. Y. & Stuur, E. R. (1974). The effects of combining two different alcohols on the heat-induced reversible denaturation of ribonuclease. Int. J. Peptide Protein Res. 6, 6574.CrossRefGoogle ScholarPubMed
302Gershon, N. D., Porter, K. R. & Trus, B. L. (1985). The cyto-plasmic matrix: its volume and surface area and the diffusion of molecules through it. Proc. Natn. Acad. Sci. U.S.A. 82, 50305034.Google Scholar
303Gibbs, J. W. (1928). The Collected Works of J. Willard Gibbs, Ph.D., LL.D., vol. 1: Thermodynamics. New York: Longmans, Green.Google Scholar
304Giese, K., Kaatze, U. & Pottel, R. (1970). Permittivity and dielectric and proton magnetic relaxation of aqueous solutions of the alkali halides. J. phys. Chem. 74, 37183725.Google Scholar
305Gill, S. J., Dec, S. F., Olofsson, G. & Wadso, I. (1985). Anomalous heat capacity of hydrophobic solvation. J. phys. Chem. 89, 37583761.Google Scholar
306Gill, S. J., Nichols, N. F. & Wadso, I. (1976). Calorimetric determination of enthalpies of solution of slightly soluble liquids. II. Enthalpy of solution of some hydrocarbons in water and their use in establishing the temperature dependence of their solubilities. J. Chem. Thermodynamics 8, 445452.CrossRefGoogle Scholar
307Gilliland, G. L. & Davies, D. R. (1984). Protein crystallization: the growth of large-scale single crystals. Meth. Enzymol. 104, 370381.CrossRefGoogle ScholarPubMed
308Gjerde, D. T. & Fritz, J. S. (1981). Sodium and potassium benzoate and benzoic acid as eluents for ion chromatography. Anal. Chem. 53, 23242327.CrossRefGoogle Scholar
309Gjerde, D. T., Fritz, J. S. & Schmuckler, G. (1979). Anion chromatography with low-conductivity eluents. J. Chromatog. 186, 509519.Google Scholar
310Gjerde, D. T., Schmuckler, G. & Fritz, J. S. (1980). Anion chromatography with low-conductivity eluents. II. J. Chromatog. 187, 3545.Google Scholar
311Glasel, J. A. (1972). Nuclear magnetic resonance studies on water and ice. In Water. A Comprehensive Treatise, vol. 1: The Physics and Physical Chemistry of Water (ed. Franks, F.), pp. 215254. New York: Plenum Press.Google Scholar
312Glew, D. N., Mak, H. D. & Rath, N. S. (1968). Aqueous non-electrolyte solutions. Part VII. Water shell stabilization by interstitial nonelectrolytes. In Hydrogen-Bonded Solvent Systems (ed. Covington, A. K. and Jones, P.), pp. 195210. London: Taylor & Francis.Google Scholar
313Click, R. E., Stewart, W. E. & Tewari, K. C. (1966). Proton nuclear magnetic resonance solvent shifts in aqueous solutions of alkali halides. J. chem. Phys. 45, 40494052.Google Scholar
314Goard, A. K. (1925). Negative adsorption. The surface tensions and activities of some aqueous salt solutions. J. Chem. Soc. 127, 24512458.CrossRefGoogle Scholar
315Goldammer, E. V. & Hertz, H. G. (1970). Molecular motion and structure of aqueous mixtures with nonelectrolytes as studied by nuclear magnetic relaxation methods. J. phys. Chem. 74, 37343755.Google Scholar
316Gomez-Puyou, M. T., De, Ayala, G., Darszon, A. & Gomez-Puyou, A. (1984). Oxidative phosphorylation and the Pi-ATP exchange reaction of submitochondrial particles under the influence of organic solvents. J. biol. Chem. 259, 94729478.CrossRefGoogle Scholar
317Gonzalez, R. G., Barnett, P., Aguayo, J., Cheng, H.-M. & Chylack, L. T. Jr., (1984). Direct measurement of polyol pathway activity in the ocular lens. Diabetes 33, 196199.Google Scholar
318Good, W. (1961 b). The haemolysis of human erythrocytes in relation to the lattice structure of water. III. The effect of non-electrolytes on malonamide-induced haemolysis. Biochim. biophys. Acta 50, 485493.Google Scholar
319Good, W. (1961 c). The haemolysis of human erythrocytes in relation to the lattice structure of water. IV. Rapid haemolysis in solutions of erythrocyte-permeable substances. Biochim. biophys. Acta 52, 545551.CrossRefGoogle Scholar
320Good, W. (1962). The haemolysis of human erythrocytes in relation to the lattice structure of water. VI. Osmotic haemolysis in solution of non-electrolytes. Biochim. biophys. Acta 57, 104110.Google Scholar
321Gooding, D., Schmuck, M. & Gooding, K. (1984). Analysis of proteins with new, mildly hydrophobic high-performance liquid chromatography packing materials. J. Chromatog. 296, 107114.Google Scholar
322Gorbunoff, M. J. (1984). The interaction of proteins with hydroxyapatite. 1. Role of protein charge and structure. Analyt. Biochem. 136, 425432.Google Scholar
323Gordon, J. E. (1975). The Organic Chemistry of Electrolyte Solutions. New York: Wiley-Interscience.Google Scholar
324Gordon, J. E. & Thorne, R. L. (1969). Proton nuclear magnetic resonance solvent shifts in aqeuous electrolyte solutions. II. Mixtures of two salts, additivity and nonlinearity of shifts. J. phys. Chem. 73, 36523660.CrossRefGoogle Scholar
325Gorevic, P. D. & Franklin, E. C. (1981). Amyloidosis. Ann. Rev. Med. 32, 261271.Google Scholar
326Gould, G. W. & Measures, J. C. (1977). Water relations in single cells. Phil. Trans. R. Soc. Lond. 6278, 151166.Google Scholar
327Grady, S. R., Wang, J. K. & Dekker, E. E. (1981). Steady-state kinetics and inhibition studies of the aldol condensation reaction catalyzed by bovine liver and Escherichia coli 2-keto-4-hydroxy-glutarate aldolase. Biochemistry 20, 24972502.CrossRefGoogle ScholarPubMed
328Gratzer, W. B. & Beaven, G. H. (1969). Effect of protein denaturation on micelle stability. J. phys. Chem. 73, 22702273.Google Scholar
329Graves, D. J., Sealock, R. W. & Wang, J. H. (1965). Cold inactivation of glycogen phosphorylase. Biochemistry 4, 290296.Google Scholar
330Greaney, G. S. & Somero, G. N. (1979). Effects of anions on the activation thermodynamics and fluorescence emission spectrum of alkaline phosphatase: evidence for enzyme hydration changes during catalysis. Biochemistry 18, 53225332.CrossRefGoogle ScholarPubMed
331Greaney, G. S. & Somero, G. N. (1980). Contributions of binding and catalytic rate constants to evolutionary modifications in Km of NADH for muscle-type (M4) lactate dehydrogenases. J. comp. Physiol. 137, 115121.Google Scholar
332Gregor, H. P., Belle, J. & Marcus, R. A. (1955). Studies on ion-exchange resins. XIII. Selectivity coefficients of quaternary base anion-exchange resins toward univalent anions. J. Am. chem. Soc. 77, 27132719.Google Scholar
333Greyson, J. (1967). The influence of the alkali halides on the structure of water. J. phys. Chem. 71, 22102213.Google Scholar
334Gross, P. & Schwarz, K. (1930). The salting-out action. Monatsh. 55, 287306.Google Scholar
335Gruen, D. W. R., Marcelja, S. & Parsegian, V. A. (1984). Water structure near the membrane surface. In Cell Surface Dynamics. Concepts and Models (ed. Perelson, A. S., DeLisi, C. and Wiegel, F. W.), pp. 5991. New York: Dekker.Google Scholar
336Grunwald, E., Baughman, G. & Kohnstam, G. (1960). The solvation of electrolytes in dioxane–water mixtures, as deduced from the effect of solvent change on the standard partial molar free energy. J. Am. chem. Soc. 82, 58015811.Google Scholar
337Gucker, F. T. Jr., Gage, F. W. & Moser, C. E. (1938). The densities of aqueous solutions of urea at 25 and 30° and the apparent molal volume of urea. J. Am. chem. Soc. 60, 25822588.Google Scholar
338Gucker, F. T. Jr., & Moser, C. E. (1939). The coefficient of expansibility of aqueous solutions of urea at 27·5° calculated from the densities at 25 and 30°. J. Am. chem. Soc. 61, 15581559.Google Scholar
339Gucker, F. T. Jr., & Pickard, H. B. (1940). The heats of dilution, heat capacities, and activities of urea in aqueous solutions from the freezing points to 40°. J. Am. chem. Soc. 62, 14641472.Google Scholar
340Gucker, F. T. Jr., Pickard, H. B. & Planck, R. W. (1939). A new micro-calorimeter: the heats of dilution of aqueous solutions of sucrose at 20 and 30° and their heat capacities at 25°. J. Am. chem. Soc. 61, 459470.Google Scholar
341Gurney, R. W. (1953). Ionic Processes in Solution. New York: McGraw-Hill.Google Scholar
342Gusarsky, E. & Treinin, A. (1965). The relation between electrochemical and spectroscopic properties of the halide and pseudohalide ions in solution. J. phys. Chem. 69, 31763177.Google Scholar
343Gutmann, V. (1978). The Donor-Acceptor Approach to Molecular Interactions. New York: Plenum Press.Google Scholar
344Haas, D. J. (1964). Interactions between lithium salts and a model peptide: crystal structure of lithium chloride-N-methylacetamide complex. Nature 201, 6465.CrossRefGoogle Scholar
345Haberland, H., Langosch, H., Schindler, H.-G. & Worsnop, D. R. (1984). Negatively charged water clusters: mass spectra of (H2O)n and (D2O)n. J. phys. Chem. 88, 39033904.Google Scholar
346Hade, E. P. K. (1972). The isopiestic method. Meth. Enzymol. 26, 177181.Google Scholar
347Hade, E. P. K. & Tanford, C. (1967). Isopiestic compositions as a measure of preferential interactions of macromolecules in two-component solvents. Application to proteins in concentrated aqueous cesium chloride and guanidine hydrochloride. J. Am. chem. Soc. 89, 50345040.Google Scholar
348Haggis, G. H., Hasted, J. B. & Buchanan, T. J. (1952). The dielectric properties of water in solutions. J. chem. Phys. 20, 14521465.Google Scholar
349Haglund, A. C. & Marsden, N. V. B. (1980). Hydrophobie and polar contributions to solute affinity for a highly crosslinked water-swollen (Sephadex) gel. J. Polymer Sci. (Polymer Lett. ed.) 18, 271279.Google Scholar
350Haglund, A. C. & Marsden, N. V. B. (1984 a). The partitioning of 1-alkanols and other compounds in Sephadex G-10 when eluted with formamide or aqueous solutions of urea, guanidinium salts and some simple electrolytes. J. Chromatog. 301, 365376.Google Scholar
351Haglund, A. C. & Marsden, N. V. B. (1984 b). Partitioning of aliphatic alcohols in Sephadex G-15 at 25°C with water as solvent. J. Chromatog. 301, 4755.CrossRefGoogle Scholar
352Hahne, G. & Hoffmann, F. (1984). Dimethyl sulfoxide can initiate cell divisions of arrested callus protoplasts by promoting cortical microtubule assembly. Proc. natn. Acad. Sci. U.S.A. 81, 54495453.Google Scholar
353Haire, R. N. & Hedlund, B. E. (1983). Hemoglobin function in the water–ethylene glycol cosolvent system: linkage between oxygen binding and hydration. Biochem. 22, 327334.Google Scholar
354Haldane, J. B. S. (1930). Enzymes, p. 182. New York: Longmans, Green.Google Scholar
355Halfpap, B. F. & Sorensen, C. M. (1982). The viscosity of supercooled aqueous solutions of ethanol and hydrazine. J. chem. Phys. 77, 466471.Google Scholar
356Hallenga, K., Grigera, J. R. & Berendsen, H. J. C. (1980). Influence of hydrophobic solutes on the dynamic behavior of water. J. phys. Chem. 84, 23812390.CrossRefGoogle Scholar
357Halliwell, H. F. & Nyburg, S. C. (1960). The reaction of the benzenediazonium ion with certain anions in aqueous acid solution. J. Chem. Soc. 46034608.CrossRefGoogle Scholar
358Halliwell, H. F. & Nyburg, S. C. (1963). Enthalpy of hydration of the proton. Trans. Faraday Soc. 59, 11261140.Google Scholar
359Halsey, M. J. (1974). Mechanisms of general anesthesia. In Anesthetic Uptake and Action (ed. Eger, E. I., II), pp. 4576, Baltimore: Williams & Wilkins.Google Scholar
360Halsey, M. J. & Smith, E. B. (1970). Effects of anesthetics on luminous bacteria. Nature 227, 13631365.Google Scholar
361Hamabata, A., Chang, S. & Von Hippel, P. H. (1973 a). Model studies on the effects of neutral salts on the conformational stability of biological macromolecules. III. Solubility of fatty acid amides in ionic solutions. Biochemistry 12, 12711278.CrossRefGoogle Scholar
362Hamabata, A., Chang, S. & Von Hippel, P. H. (1973 b). Model studies on the effects of neutral salts on the conformational stability of biological macromolecules. IV. Properties of fatty acid amide micelles. Biochemistry 12, 12781282.Google Scholar
363Hamabata, A. & Von Hippel, P. H. (1973). Model studies on the effects of neutral salts on the conformational stability of biological macromolecules. II. Effects of vicinal hydrophobic groups on the specificity of binding of ions to amide groups. Biochemistry 12, 12641271.Google Scholar
364Hamaguchi, K. & Geiduschek, E. P. (1962). The effect of electrolytes on the stability of the deoxyribonucleate helix. J. Am. chem. Soc. 84, 13291338.Google Scholar
365Hamer, W. J. (1959). The Structure of Electrolytic Solutions. New York: John Wiley & Sons Inc.; London: Chapman & Hall Limited.CrossRefGoogle Scholar
366Hammes, G. G. & Schimmel, P. R. (1967). An investigation of water–urea and water–urea–polyethylene glycol interactions. J. Am. chem. Soc. 89, 442446.Google Scholar
367Hammett, L. P. (1970). Physical Organic Chemistry. Reaction Rates, Equilibria, and Mechanisms, 2nd ed.New York: McGraw-Hill.Google Scholar
368Han, C.-C., Dodd, J. A. & Brauman, J. I. (1986). Structure and reactivity in ionic reactions. J. phys. Chem. 90, 471477.CrossRefGoogle Scholar
369Hand, S. C. & Somero, G. N. (1982). Urea and methylamine effects on rabbit muscle phosphofructokinase. Catalytic stability and aggregation state as a function of pH and temperature. J. biol. Chem. 257, 734741.CrossRefGoogle ScholarPubMed
370Hansch, C., Vittoria, A., Silipo, C. & Jow, P. Y. C. (1975). Partition coefficients and the structure-activity relationship of the anesthetic gases. J. med. Chem. 18, 546548.Google Scholar
371Hardingham, T. E. (1984). Structure and associations of proteoglycans in cartilage. In Molecular Biophysics of the Extracellular Matrix (ed. Arnott, S., Rees, D. A. & Morris, E. R.), pp. 119, Clifton, N.J.: Humana Press.Google Scholar
372Hardt, C. R., Huddleson, I. F. & Ball, C. D. (1946). An electro-phoretic analysis of changes produced in blood serum and plasma proteins by heat in the presence of sugars. J. biol. Chem. 163, 211220.Google Scholar
373Harkins, W. D. & Gilbert, E. C. (1926). The structure of films of water on salt solutions. II. The surface tension of calcium chloride solutions at 25°. J. Am. chem. Soc. 48, 604607.Google Scholar
374Harkins, W. D. & McLaughlin, H. M. (1925). The structure of films of water on salt solutions. I. Surface tension and adsorption for aqueous solutions of sodium chloride. J. Chem. Soc. 47, 20832089.Google Scholar
375Harned, H. S. & Owen, B. B. (1958). Physical Chemistry of Electrolyte Solutions, 3rd ed.New York: Reinhold Publishing Corp.Google Scholar
376Harris, F. E. & O'Konski, C. T. (1957). Dielectric properties of aqueous ionic: solutions at microwave frequencies. J. phys. Chem. 61, 310319.Google Scholar
377Harrison, P. R. (1976). Analysis of erythropoeisis at the molecular level. Nature 262, 353356.Google Scholar
378Hart, G. W. (1982 a). Corneal proteoglycans. In Cell Biology of the Eye (ed. McDevitt, D. S.), pp. 152. New York: Academic Press.Google Scholar
379Hartman, K. A. Jr., (1966). The structure of water and the stability of the secondary structure in biological molecules. An infrared and proton magnetic resonance study. J. phys. Chem. 70, 270276.Google Scholar
380Harvey, A. M., Johns, R. J., McKusick, V. A., Owens, A. H. & Ross, R. S. (1984). The Principles and Practice of Medicine, 21st ed.Norwalk: Appleton-Century-Crofts.Google Scholar
381Harvey, J. M., Jackson, S. E. & Symons, M. C. R. (1977). Interactions in water-alcohol mixtures studied by NMR and infrared spectroscopy. Chem. Phys. Lett. 47, 440441.Google Scholar
382Hascall, V. C. (1981). Proteoglycans: structure and function. In Biology of Carbohydrates, vol. 1 (ed. Ginsburg, V., and Robbins, P.), pp. 149. New York: Wiley-Interscience.Google Scholar
383Hasted, J. B. (1972). Liquid water: dielectric properties. In Water. A Comprehensive Treatise, vol. 1: The Physics and Physical Chemistry of Water (ed. Franks, F.), pp. 255332. New York: Plenum Press.Google Scholar
384Hatefi, Y. & Hanstein, W. G. (1969). Solubilization of particulate proteins and nonelectrolytes by chaotropic agents. Proc. Natn. Acad. Sci. U.S.A. 62, 11291136.Google Scholar
385Hauser, H. (1975). Lipids. In Water. A Comprehensive Treatise, vol. 4: Aqueous Solutions of Amphiphiles and Macromolecules (ed. Franks, F.), pp. 209303. New York: Plenum Press.Google Scholar
386Hawkes, R., Grutter, M. G. & Schellman, J. (184). Thermodynamic stability and point mutations of bacteriophage T4 lysozyme. J. molec. Biol. 175, 195212.Google Scholar
387Hawkes, S. & Wang, J. L. (eds) (1982). Extracellular Matrix. New York: Academic Press.Google Scholar
388Hay, E. D. (ed.) (1981). Cell Biology of Extracellular Matrix. New York: Plenum Press.Google Scholar
389Hendry, E. B. (1952). Delayed hemolysis of human erythrocytes in solutions of glucose. J. gen. Physiol. 35, 605616.Google Scholar
390Hepler, L. G. (1963). Effects of substituents on acidities of organic acids in water: thermodynamic theory of the Hammett equation. J. Am. chem. Soc. 85, 30893092.Google Scholar
391Hepler, L. G. & O'hara, W. F. (1961). Thermodynamic theory of acid dissociation of methyl substituted phenols in aqueous solution. J. phys. Chem. 65, 811814.Google Scholar
392Herskovits, T. T., Carberry, S. E. & San George, R. C. (1983). Subunit structure and dissociation of Homarus americanus hemocyanin. Effects of salts and ureas on the acetylated and unmodified hexamers. Biochemistry 22, 41074112.CrossRefGoogle Scholar
393Herskovits, T. T., Cavanagh, S. M. & San George, R. C. (1977). Light-scattering investigations of the subunit dissociation of human hemoglobin A. Effects of various salts. Biochemistry 16, 57955801.Google Scholar
394Herskovits, T. T., Gadegbeku, B. & Jaillet, H. (1970). On the structural stability and solvent denaturation of proteins. I. Denaturation by the alcohols and glycols. J. biol. Chem. 245, 25882598.CrossRefGoogle ScholarPubMed
395Herskovits, T. T. & Russell, M. W. (1984). Light-scattering investigation of the subunit structure and dissociation of Helix pomatia hemocyanin. Effects of salts and ureas. Biochemistry 23, 28122819.Google Scholar
396Herskovits, T. T., San George, R. C. & Erhunmwunsee, L. J. (1981). Light-scattering investigation of the subunit dissociation of Homarus americanus hemocyanin. Effects of salts and ureas. Biochemistry 20, 25802587.Google Scholar
397Herskovits, T. T. & Solli, N. J. (1975). Studies of the conformation of apomyoglobin in aqueous solutions and denaturing organic solvents. Biopolymers 14, 319334.CrossRefGoogle ScholarPubMed
398Hertz, H. G. (1963). Kernresonanzuntersuchungen and Elektrolytlosungen. Ber. Bunsenges. physik. Chem. 67, 311327.Google Scholar
399Hertz, H. G. (1967). Microdynamic behaviour of liquids as studied by NMR relaxation times. Prog. Nuclear Magn. Reson. Spectrosc. 3, 159230.Google Scholar
400Hertz, H. G. (1973 a). Nuclear magnetic relaxation spectroscopy. In Water. A Comprehensive Treatise, vol. 3: Aqueous Solutions of Simple Electrolytes (ed. Franks, F.), pp. 301399. New York: Plenum Press.Google Scholar
401Hertz, H. G. (1974). N.M.R. methods, aqueous electrolyte solutions. In Structure of Water and Aqueous Solutions (ed. Luck, W. A. P.), pp. 339460. New York: Verlag Chemie.Google Scholar
402Hertz, H. G. & Radle, C. (1973). The orientation of the water molecules in the hydration sphere of F and in the hydrophobic hydration sphere. Ber. Bunsenges. Phys. Chem. 77, 521531.Google Scholar
403Hertz, H. G. & Radle, C. (1974). Fluoride–fluoride association in aqueous solution from F19 nuclear magnetic relaxation data. Ber. Bunsenges. Phys. Chem. 78, 509514.Google Scholar
404Hertz, H. G. & Spalthoff, W. (1959). Chemical shifts of proton resonances of water in electrolyte solutions. Z. Elektrochem. 63, 10961110.Google Scholar
405Hertz, H. G., Tutsch, R. & Versmold, H. (1971). Molecular motion and structure around the hydrated ions Li+ and Al3+. Deutsche Bunsenges. Phys. Chem. 75, 11771191.Google Scholar
406Hertz, H. G., Versmold, H. & Yoon, C. (1983). The effect of added salts on the proton exchange rate of water as studied by 17O-NMR. Ber. Bunsenges. Phys. Chem. 87, 577582.Google Scholar
407Hertz, H. G. & Zeidler, M. D. (1963). Elementary processes in the hydrated shells of ions from proton-relaxation and deuteron-relaxation times. Ber. Bunsenges. Phys. Chem. 67, 774786.Google Scholar
408Hertz, H. G. & Zeidler, M. D. (1976). Nuclear magnetic relaxation in hydrogen bonded liquids. In Hydrogen Bond: Recent Developments in Theory and Experiments, vol. 3 (ed. Schuster, P., Zundel, G. and Sandorfy, C.), pp. 10271061. New York: North Holland.Google Scholar
409Hey, M. J., Clough, J. M. & Taylor, D. J. (1976). Ion effects on macromolecules in aqueous solution. Nature 262, 807809.Google Scholar
410Hibbard, L. S. & Tulinsky, A. (1978). Expression of functionality of α-chymotrypsin. Effects of guanidine hydrochloride and urea in the onset of denaturation. Biochemistry 17, 54605468.Google Scholar
411Hildebrand, J. H. (1979). Is there a ‘hydrophobic effect’? Proc. natn. Acad. Sci. U.S.A. 76, 194.Google Scholar
412Hindman, J. C. (1962). Nuclear magnetic resonance effects in aqueous solutions of 1–1 electrolytes. J. chem. Phys. 36, 10001015.Google Scholar
413Hirtz, D. G. & Nelson, K. B. (1983). The natural history of febrile seizures. Ann. Rev. Med. 34, 453471.Google Scholar
414Hjerten, S. (1981 a). Hydrophobic interaction chromatography of proteins, nucleic acids, viruses, and cells on noncharged amphiphilic gels. Meth. Biochem. Anal. 27, 89110.Google Scholar
415Hjerten, S. (1981 b). Hydrophobic interaction chromatography. Adv. in Chromatography 19, 111123.Google Scholar
416Hodgkinson, S. C. & Lowry, P. J. (1981). Hydrophobic-interaction chromatography and anion-exchange chromatography in the presence of acetonitrile. Biochem. J. 199, 619627.Google Scholar
417Hofmeister, F. (1888). On the understanding of the effect of salts. Second report. On regularities in the precipitating effect of salts and their relationship to their physiological behavior. Naunyn-Schmiedebergs Archiv fuer Experimentelle Pathologie und Pharmakologie (Leipzig) 24, 247260.Google Scholar
418Hollander, W. (1976). Unified concept on the role of acid muco-polysaccharides and connective tissue proteins in the accumulation of lipids, lipoproteins, and calcium in the atheroschlerotic plaque. Expl. Mol. Path. 25, 106120.Google Scholar
419Holmberg, L. (1983). Structural investigation of epichlorohydrin crosslinked polysaccharide gels. Ph.D. thesis, Swedish University of Agricultural Sciences, Uppsala.Google Scholar
420Holmes, W. M., Hurd, R. E., Reid, B. R., Rimerman, R. A. & Hatfield, G. W. (1975). Separation of transfer ribonucleic acid by Sepharose chromatography using reverse salt gradients. Proc. natn. Acad. Sci. U.S.A. 72, 10681071.Google Scholar
421Hook, M., Kjellen, L., Johansson, S. & Robinson, J. (1984). Cell-surface glycosaminoglycans. Ann. Rev. Biochem. 53, 847869.CrossRefGoogle ScholarPubMed
422Hopfinger, A. J. (1977). Intermodular Interactions and Biomolecular Organization, pp. 251277. New York: Wiley-Interscience.Google Scholar
423Horne, R. A. (1972). Water and Aqueous Solutions. Structure, Thermodynamics, and Transport Processes. New York: Wiley-Interscience.Google Scholar
424Hsu, D. K., Huang, Y. Y. & Kimura, T. (1984). Thermodynamic properties of the cholesterol transfer reaction from liposomes to cytochrome P450scc: an enthalpy–entropy compensation effect. Biochem. biophys. Res. Comm. 118, 877884.CrossRefGoogle ScholarPubMed
425Hupe, D. J. & Jencks, W. P. (1977). Nonlinear structure–reactivity correlations. Acyl transfer between sulfur and oxygen nucleophiles. J. Am. chem. Soc. 99, 451464.Google Scholar
426Hupe, D. J. & Pohl, E. R. (1984). On the magnitude of primary isotope effects for proton abstraction from carbon. J. Am. chem. Soc. 106, 56345640.Google Scholar
427Hupe, D. J. & Wu, D. (1977). The effect of solvation on Bronsted β values for proton transfer reactions. J. Am. chem. Soc. 99, 76537659.Google Scholar
428Hupe, D. J., Wu, D. & Shepperd, P. (1977). The effect of solvation on β values for nucleophilic reactions. J. Am. chem. Soc. 99, 76597662.Google Scholar
429Hvidt, A., Moss, R. & Nielsen, G. (1978). Volume properties of aqueous solutions of tert–butyl alcohol at temperatures between 5 and 25°C. Acta chem. scand. 832, 274280.Google Scholar
430Hyne, J. B. (1960). Specific solvation in binary solvent mixtures. Part I. Variations in activation energy of reactions in mixed solvents. J. Am. chem. Soc. 82, 51295135.Google Scholar
431Hyne, J. B., Wills, R. & Wonka, R. E. (1962). Specific solvation in binary solvent mixtures. Part II. The dependence of activation energy of solvolysis of benzyl chloride in ethanol–water mixtures on temperature and ring substitution. J. Am. chem. Soc. 84, 29142919.Google Scholar
432Imai, Y. & Sato, R. (1967). Conversion of P-450 to P-420 by neutral salts and some other reagents. Eur. J. Biochem. 1, 419426.Google Scholar
433Ingham, K. (1978). Precipitation of proteins with polyethylene glycol: characterization of albumin. Arch. Biochem. Biophys. 186, 106113.Google Scholar
434Ingham, K. C. (1984). Protein precipitation with polyethylene glycol. Meth. Enzymol. 104, 351356.Google Scholar
435Irias, J. J., Olmsted, M. R. & Utter, M. F. (1969). Pyruvate carboxylase. Reversible inactivation by cold. Biochemistry 8, 51365148.Google Scholar
436Irving, R. J., Nelander, L. & Wadso, I. (1964). Thermodynamics of the ionization of some thiols in aqueous solution. Acta chem. scand. 18, 769787.Google Scholar
437Iskandarani, Z. & Pietrzyk, D. J. (1982 a). Ion interaction chromatography of organic anions on a poly(styrene-divinylbenzene) adsorbent in the presence of tetraalkylammonium salts. Analyt. Chem. 54, 10651071.Google Scholar
438Iskandarani, Z. & Pietrzyk, J. (1982 b). Ion interaction chromatography of inorganic anions on a poly(styrene-divinylbenzene) adsorbent in the presence of tetraalkylammonium salts. Anal. Chem. 54, 2427–2431.Google Scholar
439Ives, D. J. G. & Marsden, P. D. (1965). The ionisation functions of diisopropylcyanoacetic acid in relation to hydration equilibria and the compensation law. J. Chem. Soc. 649676.Google Scholar
440Janoff, A. S. & Miller, K. W. (1982). A critical assessment of the lipid theories of general anaesthetic action. In Biological Membranes, vol. 4 (ed. Chapman, D.), pp. 417476. New York: Academic Press.Google Scholar
441Janson, J.-C. (1967). Adsorption phenomena on Sephadex®. J. Chromatogr. 28, 1220.Google Scholar
442Jarabak, J., Seeds, A. E. Jr., & Talalay, P. (1966). Reversible cold inactivation of a 17β-hydroxysteroid dehydrogenase of human placenta: protective effect of glycerol. Biochemistry 5, 12691279.Google Scholar
443Jarvis, N. L. & Scheiman, M. A. (1968). Surface potentials of aqueous electrolyte solutions. J. phys. Chem. 72, 7478.Google Scholar
444Jasra, R. V. & Ahluwalia, J. C. (1983). Enthalpies and heat capacities of transfer of some sugars from water to aqueous urea solutions. J. Chem. Soc. Faraday Trans, I 79, 13031309.CrossRefGoogle Scholar
445Jeffrey, G. A. (1969). Water structure in organic hydrates. Accts. Chem. Res. 2, 344352.Google Scholar
446Jellinek, H. H. G. (1972). The ice interface. In Water and Aqueous Solutions (ed. Horne, R. A.), pp. 65107. New York: Wiley-Interscience.Google Scholar
447Jencks, W. P., Brant, S. R., Gandler, J. R., Fendrich, G. & Nakamura, C. (1982). Nonlinear Bronsted correlations: the roles of resonance, solvation, and changing transition-state structure. J. Am. chem. Soc. 104, 70457051.Google Scholar
448Jencks, W. P. & Gilchrist, M. (1962). The nucleophilic reactivity of alcoholate anions toward p–nitrophenyl acetate. J. Am. chem. Soc. 84, 29102913.Google Scholar
449Jendrasiak, G. L. & Hasty, J. H. (1974). The hydration of phospholipids. Biochim. biophys. Acta 337, 7991.Google Scholar
450Jensen, W. B. (1980). The Lewis Acid-Base Concepts. An Overview. New York: John Wiley.Google Scholar
451Jezorek, J. R. & Mark, H. B. Jr., (1970). The effect of water as a proton donor on the decay of anthracene and naphthalene anion radicals in aqueous mixtures of acetonitrile, dimethylformamide, and dimethyl sulfoxide. J. phys. Chem. 74, 16271633.Google Scholar
452Johansson, K. & Eriksson, J. C. (1974). γ and dγ/dT measurements on aqueous solutions of 1,1-electrolytes. J. Colloid and Interface Sci. 49, 469480.Google Scholar
453Johnson, F. H. & Flagler, E. A. (1950). Hydrostatic pressure reversal of narcosis in tadpoles. Science 112, 9192.Google Scholar
454Johnson, K. J. & Ward, P. A. (1984). Newer concepts in the pathogenesis of immune complex-induced tissue injury. Adv. Biol. Dis. 1, 104112.Google Scholar
455Jones, D. & Symons, M. C. R. (1971). Solvation spectra. Part 38. – E.s.r. spectra of m–dinitrobenzene anions in water + t–butyl alcohol mixtures: asymmetric hydration. Trans. Faraday Soc. 67, 961965.Google Scholar
456Jones, G. & Dole, M. (1929). The viscosity of aqueous solutions of strong electrolytes with special reference to barium chloride. J. Am. chem. Soc. 51, 29502964.Google Scholar
457Jones, G. & Ray, W. A. (1941). The surface tension of solutions of electrolytes as a function of concentration II. J. Am. chem. Soc. 63, 288294.Google Scholar
458Jorgensen, W. L., Gao, J. & Ravimohan, C. (1985). Monte Carlo simulations of alkanes in water: hydration numbers and the hydrophobic effect. J. phys. Chem. 89, 34703473.Google Scholar
459Jorgensen, W. L. & Swenson, C. J. (1985 a). Optimized intermolecular potential functions for amides and peptides. Structure and properties of liquid amides. J. Am. chem. Soc. 107, 569578.Google Scholar
460Jorgensen, W. L. & Swenson, C. J. (1985 b). Optimized intermolecular potential functions for amides and peptides. Hydration of amides. J. Am. chem. Soc. 107, 14891496.Google Scholar
461Kador, P. F., Robison, W. G. Jr., & Kinoshita, J. H. (1985). The pharmacology of aldose reductase inhibitors. Ann. Rev. Pharmacol. Toxicol. 25, 691714.Google Scholar
462Kaminsky, M. (1957). Ion-solvent interaction and the viscosity of strong-electrolyte solutions. Discuss. of the Faraday Soc. 24, 171179.CrossRefGoogle Scholar
463Karplus, M. & Rossky, P. J. (1980). Solvation: a molecular dynamics study of a dipeptide in water. In Water in Polymers (ed. Rowland, S. P.), pp. 2342. ACS Symposium Series, vol. 127. Washington, D.C.: American Chemical Society.Google Scholar
464Karplus, M. & Rossky, P. J. (1981). Solvation of a dipeptide by water. Ann. N.Y. Acad. Sci. 367, 151161.Google Scholar
465Kasahara, M. & Penefsky, H. S. (1978). High affinity binding of monovalent P1 by beef heart mitochondrial adenosine triphosphatase. J. biol. Chem. 253, 41804187.Google Scholar
466Katagiri, M., Takemori, S., Nakazawa, K., Suzuki, H. & Akagi, K. (1967). Benzyl alcohol dehydrogenase, a new alcohol dehydrogenase from Pseudomonas sp. Biochim. biophys. Acta 139, 173176.Google Scholar
467Katz, J. R. & Muschter, F. J. F. Jr., (1933). The lyotropic series in swelling and its extension to organic, non-ionizable substances. I. Influence of inorganic salts with different anions on the swelling of potato starch. Biochem. Z. 257, 385396.Google Scholar
468Kaufman, R. D. (1977). Biophysical mechanisms of anesthetic action: historical perspective and review of current concepts. Anesthesiology 46, 4962.Google Scholar
469Kay, R. L. (1968). The effect of water structure on the transport properties of electrolytes. In Trace Inorganics in Water (ed. Baker, R. A.), Advances in Chemistry Series 73, pp. 117. Washington, D.C.: American Chemical Society.Google Scholar
470Kay, R. L. (1973). Ionic transport in water and mixed aqueous solvents. In Water. A Comprehensive Treatise, vol. 3: Aqueous Solutions of Simple Electrolytes (ed. Franks, F.), pp. 173209. New York: Plenum Press.Google Scholar
471Kay, R. L., Cunningham, G. P. & Evans, D. F. (1968). The effect of solvent structure on ionic mobilities in aqueous solvent mixtures. In Hydrogen-Bonded Solvent Systems (ed. Covington, A. K. and Jones, P.), p. 255. London: Taylor and Francis Ltd.Google Scholar
472Kay, R. L. & Evans, D. F. (1965). The conductance of the tetraalkylammonium halides in deuterium oxide solutions at 25°. J. phys. Chem. 69, 42164221.Google Scholar
473Kella, N. K. D. & Poola, I. (1985). Sugars decrease the thermal denaturation and aggregation of arachin. Int. J. Peptide Protein Res. 26, 390399.Google Scholar
474Kenttamaa, J. & Lindberg, J. J. (1960 a). Volumes and heats of mixing of dimethyl sulphoxide–water solutions. Suomen Kemistilehti B33, 3235.Google Scholar
475Kenttamaa, J. & Lindberg, J. J. (1960 b). Thermodynamic excess functions of the system dimethyl sulfoxide–water. Suomen Kemistilehti 833, 98100.Google Scholar
476Kenttamaa, J., Tommila, E. & Martti, M. (1959). Some thermodynamic properties of the system t–butanol + water. Suomal. Tiedeakat. Toim. Annales Academie Scientiarum Fennicae, Series AII. Chemica 93, 320.Google Scholar
477Kerekes, M. (1963). Protective action of certain sugars against heat denaturation of serum proteins. Acad. Rep. Populare Romine, Studii Cercetari Biochim. 6, 245–9; Chem. Abstr. 59, 142194 (1963).Google Scholar
478Kingston, B. & Symons, M. C. R. (1973). Solvation spectra. Part 44 – Nuclear magnetic resonance study of binary solvent mixtures: water structural effects. J. Chem. Soc. Faraday Trans. II, 69, 978992.Google Scholar
479Kinoshita, J. H. (1974). Mechanisms initiating cataract formation. Invest. Ophthalmol. 13, 713724.Google Scholar
480Kleeberg, H., Kocack, O. & Luck, W. A. P. (1982). Comparison of calorimetrie and IR-spectroscopic data for alcohols and alcoholic solutions. J. Solution Chem. 11, 611624.Google Scholar
481Kleeberg, H. & Luck, W. A. P. (1983). Infrared study of anionwater interactions in dichloromethane. J. Sol. Chem. 12, 369381.Google Scholar
482Klein, D. J. & Seitz, W. A. (1983). Self-similar self-avoiding structures: Models for polymers. Proc. natn. Acad. Sci. U.S.A. 80, 31253128.Google Scholar
483Klein, S. M., Cohen, G. & Cederbaum, A. I. (1981). Production of formaldehyde during metabolism of dimethyl sulfoxide by hydroxyl radical generating systems. Biochemistry 20, 60066012.Google Scholar
484Kleinman, H. K., McGarvey, M. L., Hassell, J. R., Star, V. L., Cannon, F. B., Laurie, G. W. & Martin, G. R. (1986). Basement membrane complexes with biological activity. Biochemistry 25, 312318.Google Scholar
485Klibanov, A. M. (1979). Enzyme stabilization by immobilization. Anal. Biochem. 93, 125.Google Scholar
486Klotz, I. M. (1965). Role of water structure in macromolecules. Fedn. Proc. 24, 824833.Google Scholar
487Knoll, D. & Hermans, J. (1983). Polymer-protein interactions. Comparison of experiment and excluded volume theory. J. biol. Chem. 258, 5710–5715.Google Scholar
488Ko, H. C., O'Hara, W. F., Hu, T. & Hepler, L. G. (1964). lonization of substituted phenols in aqueous solution. J. Am. chem. Soc. 86, 10031004.Google Scholar
489Kohno, T. & Roth, J. (1979). Electrolyte effects on the activity of mutant enzymes in vivo and in vitro. Biochem. 18, 13861392.Google Scholar
490Kono, N. & Uyeda, K. (1973). Chicken liver phosphofructokinase II. Cold inactivation. J. biol. Chem. 248, 86038609.Google Scholar
491Korber, C. & Scheiwe, M. W. (1980). The cryoprotective properties of hydroxyethyl starch investigated by means of differential thermal analysis. Cryobiology 17, 5465.Google Scholar
492Kozloff, L. M., Lute, M. & Westaway, D. (1984). Phosphatidyl-inositol as a component of the ice nucleating site of Pseudomonas syringae and Erwinia hersicola. Science 226, 845846.Google Scholar
493Kremmer, T. & Boross, L. (1979). Gel Chromatography. New York: John Wiley & Sons.Google Scholar
494Kresheck, G. C. (1969). An alternate interpretation of the mechanism of self-association of urea in dilute aqueous solution. J. phys. Chem. 73, 24412443.Google Scholar
495Kresheck, G. G. & Scheraga, H. A. (1965). The temperature dependence of the enthalpy of formation of the amide hydrogen bond; the urea model. J. phys. Chem. 69, 17041706.Google Scholar
496Krestov, G. A. (1962 a). The thermodynamic characteristics of structural changes in water accompanying the hydration of ions. J. struct. Chem. 3, 125130.Google Scholar
497Krestov, G. A. (1962 b). The thermodynamic characteristics of structural changes in water connected with the hydration of polyatomic and complex ions. J. Struct. Chem. 3, 391398.Google Scholar
498Krestov, G. A. (1965 a). Entropy changes in the hydration of monatomic ions. Theor. Expl. Chem. 1, 313318.Google Scholar
499Krestov, G. A. (1965 b). Entropy characteristics of small and large ionic hydration spheres. Bulletin of the institutions of higher education, chemistry, and chemical technology 8, 734740; Chem. Abstr. 64, 10478a (1966).Google Scholar
500Krishnan, C. V. & Friedman, H. L. (1970). Solvation enthalpies of various ions in water and heavy water. J. phys. Chem. 74, 23562362.Google Scholar
501Krishnan, C. V. & Friedman, H. L. (1971). Solvation enthalpies of electrolytes in methanol and dimethylformamide. J. phys. Chem. 75, 36063612.Google Scholar
502Kruyt, H. R. & Robinson, C. (1926). Lyotropy. Verslag Akad. Wetenschappen Amsterdam 35, 812818.Google Scholar
503Kuehn, K., Schoene, H. & Timpl, R. (1981). New Trends in Basement Membrane Research. New York: Raven Press.Google Scholar
504Kura, G., Koyama, A. & Tarutani, T. (1977). Chromatographic study of some inorganic ions on Sephadex gel in thiocyanate media. J. Chromatog. 144, 245252.Google Scholar
505Kurtz, J. & Harrington, W. F. (1966). Interaction of poly-L-proline with lithium bromide. J. molec. Biol. 17, 440455.Google Scholar
506Kurz, J. L. (1963). Transition state characterized for catalyzed reactions. J. Am. chem. Soc. 85, 987991.Google Scholar
507Laidler, K. J. (1959). Thermodynamics of ionization processes in aqueous solution. Part 1.– general theory of substituent effects. Trans. Faraday Soc. 55, 17251730.Google Scholar
508Lakshmi, T. S. & Nandi, P. K. (1976 a). Effects of sugar solutions on the activity coefficients of aromatic amino acids and their N–acetyl ethyl esters. J. phys. Chem. 80, 249252.Google Scholar
509Lakshmi, T. S. & Nandi, P. K. (1976 b). Hydrophobie interaction in sugar solutions. Results from gel interaction study. J. Chromat. 116, 177179.Google Scholar
510Landis, B. H., Koehler, K. A. & Fenton, J. W. II., (1981). Human thrombins. Group IA and IIA salt-dependent properties of α-thrombin. J. biol. Chem. 256, 46044610.Google Scholar
511Lange, E. (1959). In The Structure of Electrolytic Solutions (ed. Hamper, W. J.), p. 144. New York: John Wiley & Sons; London: Chapman & Hall.Google Scholar
512Lange, E. & Markgraf, H. G. (1950). Heats of dilution of non-electrolytes in aqueous solution up to great dilution. Z. Elektrochem. 54, 7376.Google Scholar
513Lange, E. & Mohring, K. (1953). The integral heat of dilution of some nonelectrolytes in water and octamethyltetrasiloxane at small concentration. Z. Elektrochem. 57, 660662.Google Scholar
514Langer, L. J. & Engel, L. L. (1958). Human placental estradiol 17β dehydrogenase. I. Concentration, characterization and assay. J. biol. Chem. 233, 583588.Google Scholar
515Langmuir, I. (1917). The constitution and fundamental properties of solids and liquids. II. Liquids. J. Am. chem. Soc. 39, 18481906.Google Scholar
516Lanyi, J. K. (1974). Salt-dependent properties of proteins from extremely halophilic bacteria. Bad. Rev. 38, 272290.Google Scholar
517Lanyi, J. K. & Stevenson, J. (1970). Studies of the electron transport chain of extremely halophilic bacteria. IV. Role of hydrophobic forces in the structure of menadione reductase. J. biol. Chem. 245, 40744080.Google Scholar
518Lapanje, S. (1978). Physicochemical Aspects of Protein Denaturation. New York: John Wiley.Google Scholar
519Lapanje, S., Lunder, M., Vlachy, V. & Skerjanc, J. (1977). Thermodynamics of the isothermal interaction of βlactoglobulin with guanidiniurn chloride and urea. Biochim. biophys. Acta 491, 482490.Google Scholar
520Larson, J. W. & Hepler, L. G. (1969). Heats and entropies of ionization. In Solute-Solvent Interactions (ed. Coetzee, J. F. and Ritchie, C. D.), pp. 144. New York: Dekker.Google Scholar
521Lash, J. W. & Vasan, N. S. (1983). Glycosaminoglycans of cartilage. In Cartilage, vol. 1: Structure, Function, and Biochemistry (ed. Hall, B. K.), pp. 215251. New York: Academic Press.Google Scholar
522Latimer, W. M., Pitzer, K. S. & Slansky, C. M. (1939). The free energy of hydration of gaseous ions, and the absolute potential of the normal calomel electrode. J. chem. Phys. 7, 108111.Google Scholar
523Lauffer, M. A. (1964). Polymerization-depolymerization of tobacco mosaic virus protein. II. Theory of protein hydration. Biochemistry 3, 731736.Google Scholar
524Lauffer, M. A. (1975). Entropy-Driven Processes in Biology. Polymerization of Tobacco Mosaic Virus Protein and Similar Reactions. Molecular Biology, Biochemistry and Biophysics, vol. 20. New York: Springer-Verlag.Google Scholar
525Laurent, T. C. (1963 a). The interaction between polysaccharides and other macromolecules VI. Further studies on the solubility of proteins in dextran solutions. Acta chem. scand. 17, 26642668.Google Scholar
526Laurent, T. C. (1963 b). The interaction between polysaccharides and other macromolecules 5. The solubility of proteins in the presence of dextran. Biochem. J. 89, 253257.Google Scholar
527Laurent, T. C. & Ogston, A. G. (1963). The interaction between polysaccharides and other macromolecules 4. The osmotic pressure of mixtures of serum albumin and hyaluronic acid. Biochem. J. 89, 249253.Google Scholar
528Lee, J. C., Frigon, R. P. & Timasheff, S. N. (1975). Structural stability of calf brain microtubule protein. Ann. N. Y. Acad. Sci. 253, 284291.Google Scholar
529Lee, J. C., Gekko, K. & Timasheff, S. N. (1979). Measurements of preferential solvent interactions by densimetric techniques. Meth. Enzymol. 61, 2649.Google Scholar
530Lee, J. C. & Lee, L. L. Y. (1981). Preferential solvent interactions between proteins and polyethylene glycols. J. biol. Chem. 256, 625631.Google Scholar
531Lee, J. C. & Timasheff, S. N. (1974). Partial specific volumes and interactions with solvent components of proteins in guanidine hydrochloride. Biochemistry 13, 257265.Google Scholar
532Lee, J. C. & Timasheff, S. N. (1979). The calculation of partial specific volumes of proteins in 6 M guanidine hydrochloride. Meth. Enzmol. 61, 4957.Google Scholar
533Lee, J. C. & Timasheff, S. N. (1981). The stabilization of proteins by sucrose. J. biol. Chem. 256, 71937201.Google Scholar
534Leffler, J. E. (1955). The enthalpy–entropy relationship and its implications for organic chemistry. J. Org. Chem. 20, 12021231.Google Scholar
535Lehmann, M. S., Mason, S. A. & McIntyre, G. J. (1985). Study of ethanol–lysozyme interactions using neutron diffraction. Biochemistry 24, 58625869.Google Scholar
536Lehmann, M. S. & Zaccai, G. (1984). Neutron small-angle scattaring studies of ribonuclease in mixed aqueous solutions and determination of the preferentially bound water. Biochemistry 23, 19391942.Google Scholar
537Lehrer, S. S. & Leavis, P. C. (1978). Solute quenching of protein fluorescence. Meth. Enzymol. 49, 222236.Google Scholar
538Leneveu, D. M., Parsegian, V. A. & Gingell, D. (1977). Measurement and modification of forces between lecithin bilayers. Biophs. J. 18, 209230.Google Scholar
539LeNeveu, D. M., Rand, R. P. & Parsegian, V. A. (1976). Measurement of forces between lecithin bilayers. Nature 259, 601603.Google Scholar
540Le Rudulier, D., Strom, A. R., Dandekar, A. M., Smith, L. T. & Valentine, R. C. (1984). Molecular biology of osmoregulation. Science 224, 10641068.Google Scholar
541Leung, P. S. & Safford, G. J. (1970). A neutron inelastic scattering investigation of the concentration and anion dependence of low frequency motions of H2O molecules in ionic solutions. J. phys. Chem. 74, 36963709.Google Scholar
542Lever, M. J., Miller, K. W., Paton, W. D. M. & Smith, E. B. (1971). Pressure reversal of anesthesia. Nature 231, 368371.CrossRefGoogle Scholar
543Levine, A. S., Hasegawa, F. & Murayama, M. (1975). The influence of solutes and solvent structure on gelation and aggregation of deoxy-sickle cell hemoglobin. J. Molecular Med. 1, 1926.Google Scholar
544Li, Z. Q., Giege, R., Jacrot, B., Oberthur, R., Thierry, J.-C. & Zaccai, G. (1983). Structure of phenylalanine-accepting transfer ribonucleic acid and of its environment in aqueous solvents with different salts. Biochemistry 22, 43804388.Google Scholar
545Liang, S.-M. & Liu, T.-Y. (1982). Studies on the limulus coagulation system: inhibition of activation of the proclotting enzyme by dimethyl sulfoxide. Biochem. biophys. Res. Comm. 105, 553559.Google Scholar
546Lienhard, G. E. (1973). Enzymatic catalysis and transition-state theory. Science 180, 149154.Google Scholar
547Lienhard, G. E., Secemski, I. I., Koehler, K. A. & Lindquist, R. N. (1972). Enzymatic catalysis and the transition state theory of reaction rates: transition state analogs. Cold Spring Harbor symp. quant. Biol. 36, 4561.Google Scholar
548Lilley, T. H. (1973). Raman spectroscopy of aqueous electrolyte solutions. In Water. A Comprehensive Treatise, vol. 3: Aqueous Solutions of Simple Electrolytes (ed. Franks, F.), pp. 265299. New York: Plenum Press.Google Scholar
549Lindqvist, B. (1962). The separation of buffer salts at gel filtration. Acta chem. scand. 16, 17941798.Google Scholar
550Lis, L. J., Parsegian, V. A. & Rand, R. P. (1981 a). Binding of divalent cations to dipalmitoylphosphatidylchloline bilayers and its effect on bilayer interaction. Biochemistry 20, 17611770.Google Scholar
551Lis, L. J., Lis, W. T., Parsegian, V. A. & Rand, R. P. (1981 b). Adsorption of divalent cations to a variety of phosphatidylcholine bilayers. Biochemistry 20, 17711777.Google Scholar
552Loewus, F. A. & Loewus, M. W. (1983). Myo-inositol: its biosynthesis and metabolism. Ann. Rev. Plant Physiol. 34, 137161.Google Scholar
553Loftfield, R. B., Eigner, E. A., Pastuszyn, A., Lovgren, T. N. E. & Jakubowski, H. (1980). Conformational changes during enzyme catalysis: role of water in the transition state. Proc. natn. Acad. Sci. U.S.A. 77, 33743378.Google Scholar
554Long, F. A. & McDevit, W. F. (1952). Activity coefficients of nonelectrolyte solutes in aqueous salt solutions. Chem. Rev. 51, 119169.Google Scholar
555Lonsdale, K. (1958). The structure of ice. Proc. R. Soc. Land. A.247, 424434.Google Scholar
556Loosley-Millman, M. E., Rand, R. P. & Parsegian, V. A. (1982). Effects of monovalent ion binding and screening on measured electrostatic forces between charged phospholipid bilayers. Biophys. J. 40, 221232.Google Scholar
557Lovelock, J. E. (1953 a). The mechanism of the protective action of glycerol against haemolysis by freezing and thawing. Biochim. biophys. Acta 11, 2836.Google Scholar
558Lovelock, J. E. (1953 b). The haemolysis of human red blood-cells by freezing and thawing. Biochim. biophys. Acta 10, 414426.Google Scholar
559Low, P. S. & Somero, G. N. (1975 a). Activation volumes in enzymatic catalysis: their sources and modification by low-molecular-weight solutes. Proc. natn. Acad. Sci. U.S.A. 72, 30143018.Google Scholar
560Low, P. S. & Somero, G. N. (1975 b). Protein hydration changes during catalysis: a new mechanism of enzymic rate-enhancement and ion activation/inhibition of catalysis. Proc. natn. Acad. Sci. 72, 33053309.Google Scholar
561Low, R. L., Kaguni, J. M. & Kornberg, A. (1984). Potent catenation of supercoiled and gapped DNA circles by topoisomerase I in the presence of a hydrophilic polymer. J. biol. Chem. 259, 45764581.Google Scholar
562Luck, W. A. P. (1974 a). Infrared fundamental region. In Structure of Water and Aqueous Solutions (ed. Luck, W. A. P.), pp. 221245. New York: Verlag Chemie.Google Scholar
563Luck, W. A. P. (1974 b). Infrared overtone region. In Structure of Water and Aqueous Solutions (ed. Luck, W. A. P.), pp. 247284. New York: Verlag Chemie.Google Scholar
564Luck, W. A. P. (1976 a). Water in biological systems. Topics in Current Chemistry 64, 115180.Google Scholar
565Luck, W. A. P. (1976 b). The angle dependence of hydrogen bond interactions. In The Hydrogen Bond. Recent Developments in Theory and Experiments, vol. 11: Structure and Spectroscopy (ed. Schuster, P., Zundel, G. and Sandorfy, C.), pp. 527562. New York: North-Holland.Google Scholar
566Luck, W. A. P. (1980 a). The structure of aqueous systems and the influence of electrolytes. In Water in Polymers (ed. Rowland, S. P.), pp. 4371. Washington, D.C.: American Chemical Society.Google Scholar
567Luck, W. A. P. (1980 b). A model of hydrogen-bonded liquids. Angew. Chem. Int. Ed. Engl. 19, 2841.Google Scholar
568Luck, W. A. P. (1985). The influence of ions on water structure and on aqueous systems. In Water and Ions in Biological Systems (ed. Pullman, A., Vasilescu, V. & Packer, L.), pp. 95126. New York: Plenum Press.Google Scholar
569Luck, W. A. P. & Ditter, W. (1970). Approximate methods for determining the structure of H2O and HOD using near-infrared spectroscopy. J. phys. Chem. 74, 36873695.Google Scholar
570Lumry, R. & Rajender, S. (1970). Enthalpy–entropy compensation phenomena in water solutions of proteins and small molecules: a ubiquitous property of water. Biopolymers 9, 11251227.Google Scholar
571Lusena, C. V. (1955). Ice propagation in systems of biological interest III. Effect of solutes on nucleation and growth of ice crystals. Arch. biochem. Biophys. 57, 277284.Google Scholar
572Lusena, C. V. & Rose, D. (1956). Effect of rate of ice-crystal on hemolysis of erythocytes. Arch. biochem. Biophys. 65, 534544.Google Scholar
573Luyet, B. & Rapatz, G. (1970). A review of basic researches on the cryopreservation of red blood cells. Cryobiology 6, 425482.Google Scholar
574Luzar, A., Svetina, S. & Zeks, B. (1983). The contribution of hydrogen bonds to the surface tension of water. Chem. phys. Lett. 96, 485490.Google Scholar
575Lyons, M. S. & Thomas, J. V. (1950). Diffusion studies on dilute aqueous glycine solutions at 1 and 25°with the Gouy interference method. J. Am. chem. Soc. 72, 45064511.Google Scholar
576Ma, R. J. & Wang, C. H. (1983). Studies of the protein–protein interaction of lysozyme in dimethyl sulfoxide-water solutions by quasielastic light scattering. J. phys. Chem. 87, 679682.Google Scholar
577MacDonald, J. C., Serphillips, J. & Guerrera, J. J. (1973). Effect of urea concentration upon the activation parameters for fluidity of water. J. phys. Chem. 77, 370372.Google Scholar
578MacKenzie, A. P. (1977). Non-equilibrium freezing behaviour of aqueous systems. Phil. Trans. R. Soc. Land. B 278, 167189.Google Scholar
579MacKenzie, A. P. (1981). Modelling the ultra-rapid freezing of cells and tissues. In Microprobe Analysis of Biological Systems (ed. Hutchinson, T. E. and Somlyo, A. P.), pp. 397421. New York: Academic Press.Google Scholar
580Magnera, T. F., Caldwell, G., Sunner, J., Ikuta, S. & Kebarle, P. (1984). Solvation of the halide anions in dimethyl sulfoxide. Factors involved in enhanced reactivity of negative ions in dipolar aprotic solvents. J. Amer. chem. Soc. 106, 61406146.Google Scholar
581Makinen, M. W. & Fink, A. L. (1977). Reactivity and Cryoenzymology of enzymes in the crystalline state. Ann. Rev. biophys. Bioeng. 6, 301343.Google Scholar
582Marcus, Y. (1977). Enthalpy of mixing of water-aprotic solvent mixtures at 298 K as a function of the composition. In Introduction to Liquid State Chemistry, p. 203. New York: John Wiley & Sons.Google Scholar
583Marra, J. & Israelachvili, J. (1985). Direct measurements of forces between phosphatidylcholine and phosphatidylethanolamine bilayers in aqueous electrolyte solutions. Biochemistry 24, 46084618.Google Scholar
584Marsden, N. V. B. (1973). Anionic dominance in ionic interactions with aqueous dextran gel systems. Naturwissenschaften 60, 257.Google Scholar
585Marsden, N. V. B. & Haglund, A. C. (1984). Inclusion phenemona in a system designed for exclusion – a cavitary problem and a model. J. Inclusion Phenom. 2, 2134.Google Scholar
586Marsh, K. N. & Richards, A. E. (1980). Excess volumes for ethanol + water mixtures at 10-K intervals from 278·15 to 338·15 K. Aust. J. Chem. 33, 21212132.Google Scholar
587Martin, D. & Hauthal, H. G. (1971). Dimethyl Sulphoxide. New York: John Wiley & Sons.Google Scholar
588Martin, G. J. & Martin, M. L. (1964). Nuclear magnetic resonance and infrared absorption spectra of pure and dissolved vinyl bromides. J. chim. Phys. 61, 12221230.Google Scholar
589Martin, G. R., Rohrbach, D. H., Terranova, V. P. & Liotta, L. A. (1983). Structure, function, and pathology of basement membranes. In Connective Tissue Diseases (ed. Wagner, B. M., Fleischmajer, R. and Kaufman, N.), pp. 1630. Baltimore: Williams & Wilkins.Google Scholar
590Martin, M. (1962). Study of complexes by nuclear magnetic resonance. Chemical shift and basicities. Ann. Phys. 7, 3555.Google Scholar
591Martinez-Harnandez, A. & Amenta, P. S. (1984). The basement membrane in pathology. Adv. Biol. Dis. 1, 5273.Google Scholar
592Mason, L. S., Offutt, W. F. & Robinson, A. L. (1949). The heats of dilution of aqueous solutions of four amino acids at 25°. J. Am. chem. Soc. 71, 14631468.Google Scholar
593Masterson, W. L. (1954). Partial molai volumes of hydrocarbons in water solution. J. chem. Phys. 22, 18301833.Google Scholar
594Mastroianni, M. J., Pikal, M. J. & Lindenbaum, S. (1972). Effect of dimethyl sulfoxide, urea, guanidine hydrochloride, and sodium chloride on hydrophobic interactions. Heats of dilution of tetrabutylammonium bromide and lithium bromide in mixed aqueous solvent systems. J. phys. Chem. 76, 30503057.Google Scholar
595Mathias, S. F., Franks, F. & Trafford, K. (1984). Nucleation and growth of ice in deeply undercooled erythrocytes. Cryobiology 21, 123132.Google Scholar
596Mathieson, J. G. & Conway, B. E. (1974). Partial molai compressibilities of salts in aqueous solution and assignment of ionic contributions. J. Sol. Chem. 3, 455477.Google Scholar
597Matsumoto, J. J. (1979). Denaturation of fish muscle proteins during frozen storage. In Proteins at Low Temperatures (ed. Fennema, O.), pp. 205224. Washington, D.C.: American Chemical Society.Google Scholar
598McBain, J. W. (1950). The lyotropic series of ions. In Colloid Science, pp. 131140. Boston: D. C. Heath and Company.Google Scholar
599McCabe, W. C. & Fisher, H. F. (1970). A near-infrared spectroscopic method for investigating the hydration of a solute in aqueous solution. J. phys. Chem. 74, 29902998.Google Scholar
600McCall, D. W. & Douglass, D. C. (1965). The effect of ions on the self-diffusion of water. I. Concentration dependence. J. phys. Chem. 69, 20012011.Google Scholar
601McCreery, M. J. & Hunt, W. A. (1978). Physico-chemical correlates of alcohol intoxication. Neuropharmacology 17, 451461.Google Scholar
602McDevit, W. F. & Long, F. A. (1952). The activity coefficient of benzene in aqueous salt solutions. J. Amer. chem. Soc. 74, 17731777.Google Scholar
603McGann, L. E. (1978). Differing actions of penetrating and nonpenetrating cryoprotective agents. Cryobiology 15, 382390.Google Scholar
604McPherson, A. Jr., (1976). Crystallization of proteins from polyethylene glycol. J. biol. Chem. 251, 63006303.Google Scholar
605McPherson, A. (1982). Preparation and Analysis of Protein Crystals. New York: John Wiley.Google Scholar
606McQuarrie, I. & Peeler, D. B. (1931). The effects of sustained pituitary antidiuresis and forced water drinking in epileptic children. A diagnostic and etiologic study. J. Clin. Invest, 10, 915940.Google Scholar
607Meis, L. De & Inesi, G. (1982). ATP synthesis by sarcoplasmic reticulum ATPase following Ca+2, pH, temperature, and water activity jumps. J. biol. Chem. 257, 12891294.Google Scholar
608Melander, W. & Horvath, C. (1977). Salt effects on hydrophobic interactions in precipitation and chromatography of proteins: an interpretation of the lyotropic series. Arch. biochem. Biophys. 183, 200215.Google Scholar
609Merkler, D. J., Farrington, G. K. & Wedler, F. C. (1981). Protein thermostability. Int. J. Peptide Protein Res. 18, 430442.Google Scholar
610Meryman, H. T. (1956). Mechanics of freezing in living cells and tissues. Science 124, 515521.Google Scholar
611Meryman, H. T. (1970). The exceeding of a minimum tolerable cell volume in hypertonic suspension as a cause of freezing injury. In The Frozen Cell (ed. Wolstenholme, G. E. and O'Connor, M.), pp. 5167. London: J. & A. Churchill.Google Scholar
612Meryman, H. T. (1977). The influence of the solute environment on membrane properties. In Mammalian Cell Membranes, vol. 5: Responses of Plasma Membrames (ed. Jamieson, G. A. and Robinson, D. M.), pp. 2946. London: Butterworths.Google Scholar
613Mevarech, M., Leicht, W. & Werber, M. M. (1976). Hydrophobic chromatography and fractionation of enzymes from extremely halophilic bacteria using decreasing concentration gradients of ammonium sulfate. Biochemistry 15, 23832387.Google Scholar
614Meyerstein, D. & Treinin, A. (1962). The relation between lyotropic and spectroscopic properties of anions in solution. J. phys. Chem. 66, 446450.Google Scholar
615Mhatre, S. S., Chetty, K. G. & Pradhan, D. S. (1983). Uncoupling of oxidative phosphorylation in rat liver mitochondria following the administration of dimethyl sulphoxide. Biochem. biophys. Res. Comm. 110, 325331.Google Scholar
616Michelmore, R. W. & Franks, F. (1982). Nucleation rates of ice in undercooled water and aqueous solutions of polyethylene glycol. Cryobiology 19, 163171.Google Scholar
617Middaugh, C. R., Tisel, W. A., Haire, R. N. & Rosenberg, A. (1979). Determination of the apparent thermodynamic activities of saturated protein solutions. J. biol. Chem. 254, 367370.Google Scholar
618Mikhailov, V. A. (1961). Changes of structure in aqueous solutions of nonelectrolytes. J. struct. Chemistry 2, 625628.Google Scholar
619Miller, J. & Parker, A. J. (1961). Dipolar aprotic solvents in bimolecular aromatic nucleophilic substitution reactions. J. Am. chem. Soc. 83, 117123.Google Scholar
620Miller, J. C. & Miller, K. W. (1975). Approaches to the mechanisms of action of general anaesthetics. In Physiological and Pharmacological Biochemistry (ed. Blaschko, H. K. F.), pp. 3376. MTP International Review of Science, Biochemistry series i, vol. 12. London: Butterworths and Baltimore: University Park Press.Google Scholar
621Miller, K. W. (1969). How do anesthetics work? Anesthesiology 30, 127128.Google Scholar
622Miller, K. W., Paton, W. D. M. & Smith, E. B. (1965). Site of action of general anaesthetics. Nature 206, 574577.Google Scholar
623Miller, K. W., Paton, W. D. M., Smith, E. B. & Smith, R. A. (1972). Physicochemical approaches to the mode of action of general anesthetics. Anesthesiology 36, 339351.Google Scholar
624Miller, K. W., Paton, W. D. M., Smith, R. A. & Smith, E. B. (1973). The pressure reversal of general anesthesia and the critical volume hypothesis. Molec. Pharmacol. 9, 131143.Google Scholar
625Miller, K. W. & Smith, E. B. (1973). Intermolecular forces and the pharmacology of simple molecules. In A Guide to Molecular Pharmacology-Toxicology, part 2 (ed. Featherstone, R. M.), pp. 427475. New York: Dekker.Google Scholar
626Miller, S. L. (1961). A theory of gaseous anesthetics. Proc. natn. Acad. Sci. U.S.A. 47, 15151524.Google Scholar
627Minotani, N., Sekiguchi, T., Bautista, J. G. & Nosoh, Y. (1979). Basis of thermostability in pig heart lactate dehydrogenase treated with O–methylisourea. Biochim. biophys. Acta 581, 334341.Google Scholar
628Minton, A. P. (1981). Excluded volume as a determinant of macromolecular structure and reactivity. Biopolymers 20, 20932120.Google Scholar
629Minton, A. P. (1983). The effect of volume occupancy upon the thermodynamic activity of proteins: some biochemical consequences. Molec. Cell. Biochem. 55, 119140.Google Scholar
630Minton, A. P. & Wilf, J. (1981). Effect of macromolecular crowding upon the structure and function of an enzyme: glyceralde-hyde-3-phosphate dehydrogenase. Biochemistry 20, 48214826.Google Scholar
631Mishra, A. K. & Ahluwalia, J. C. (1981). Enthalpies, heat ca pacities and apparent molai volumes of transfer of some amino acids from water to aqueous t–butanol. J. Chem. Soc. Faraday Trans., I 77, 14691483.Google Scholar
632Mishra, A. K. & Ahluwalia, J. C. (1983). Alcohol induced conformational transitions of proteins and polypeptides. Thermodynamic studies of some model compounds. Internat. J. Peptide Protein Res. 21, 322330.Google Scholar
633Mitchell, A. G. & Wynne-Jones, W. F. K. (1953). Thermodynamic and other properties of solutions involving hydrogen bonding. Discuss. of the Faraday Soc. 15, 161168.Google Scholar
634Mollenhauer, A., Schmitt, J. M., Coughlan, S. & Heber, U. (1983). Loss of membrane proteins from thylakoids during freezing. Biochimica et biophysica acta 728, 331338.Google Scholar
635Molyneux, P. (1983). Water-Soluble Synthetic Polymers: Properties and behavior, vol. 1. Boca Raton: CRC Press, Inc.Google Scholar
636Molyneux, P. (1984). Water-Soluble Synthetic Polymers: Properties and behavior, vol. 2. Boca Raton: CRC Press, Inc.Google Scholar
637Monsan, P. & Combes, D. (1984). Effect of water activity on enzyme action and stability. Ann. N. Y. Acad. Sci. 434, 4860.Google Scholar
638Morris, D. F. C. (1959). Lyotropic numbers of the formate and acetate ions and related thermodynamic properties. Recueil des Travaux Chimiques des Pays-Bas 78, 150160.Google Scholar
639Morris, D. F. C. (1968). Ionic radii and enthalpies of hydration of ions. Structure and Bonding 4, 6382.Google Scholar
640Moss, J., Stanley, S. J. & Osborne, J. C. (1981). Effect of Self-association on activity of an ADP-ribosyltransferase from turkey erythrocytes. Conversion of inactive oligomers to active protomers by chaotropic salts. J. biol. Chem. 256, 1145211456.Google Scholar
641Murthy, A. S. N. & Rao, C. N. R. (1968). Spectroscopic studies of the hydrogen bond. Applied Spectroscopy Reviews 2, 69191.Google Scholar
642Myers, E. R. & Mow, V. C. (1983). Biomechanics of cartilage and its response to biomechanical stimuli. In Cartilage, vol. 1: Structure, Function and Biochemistry (ed. Hall, B. K.), pp. 313341. New York: Academic Press.Google Scholar
643Na, G. C. & Timasheff, S. N. (1981). Interaction of calf brain tubulin with glycerol. J. molec. Biol. 151, 165178.Google Scholar
644Nagy, B. & Jencks, W. P. (1965). Depolymerization of F-actin by concentrated solution of salts and denaturing agents. J. Am. chem. Soc. 87, 24802488.Google Scholar
645Nakanishi, K. (1960). Partal molal volumes of butyl alcohols and of related compounds in aqueous solution. Bull. chem. Soc. Jap. 33, 793797.Google Scholar
646Nandi, P. K. & Edelhoch, H. (1984). The effects of lyotropic (Hofmeister) salts on the stability of clathrin coat structure in coated vesicles and baskets. J. biol. Chem. 259, 1129011296.Google Scholar
647Nandi, P. K. & Robinson, D. R. (1972 a). The effects of salts on the free energy of the peptide group. J. Am. chem. Soc. 94, 12991308.Google Scholar
648Nandi, P. K. & Robinson, D. R. (1972 b). The effects of salts on the free energies of nonpolar groups in model peptides. J. Am. chem. Soc. 94, 13081315.Google Scholar
649Narten, A. H. (1970). Diffraction pattern and structure of aqueous ammonium halide solutions. J. phys. Chem. 74, 765768.Google Scholar
650Narten, A. H. & Lindenbaum, S. (1969). Diffraction pattern and structure of the system tetra-n–butylammonium fluoride-water at 25 °C. J. chem. Phys. 51, 11081114.Google Scholar
651Neddermeyer, P. A. & Rogers, L. B. (1968). Gel filtration behavior of inorganic salts. Analyt. Chem., 40, 755762.Google Scholar
652Neddermeyer, P. A. & Rogers, L. B. (1969). Column efficiency and electrolyte effects of inorganic salts in aqueous gel chromatography. Analyt. Chem. 41, 94102.Google Scholar
653Neujahr, H. Y. (1983). Effect of anions, chaotropes, and phenol on the attachment of flavin adenine dinucleotide to phenol hydroxylase. Biochemistry 22, 580584.Google Scholar
654Nightingale, E. R. Jr., (1959). Phenomenological theory of ion solvation. Effective radii of hydrated ions. J. phys. Chem. 63, 13811387.Google Scholar
655Northrop, J. H. & Kunitz, M. (1926). The swelling and osmotic pressure of gelatin in salt solutions. J. gen. Physiology 8, 317337.Google Scholar
656Noyes, R. M. (1964). Assignment of individual ionic contributions to properties of aqueous ions. J. Amer. chem. Soc. 86, 971979.Google Scholar
657Nyns, E. J. & Wiaux, A. L. (1969). Stabilization with organic solvents of alcohol dehydrogenase from Candida lipolytica grown on n–hexadecane. Arch. Internat. physiol. Biochim. 77, 393395.Google Scholar
658Oguni, M. & Angell, C. A. (1983). Hydrophobic and hydrophilic solute effects on the homogeneous nucleation temperature of ice from aqueous solutions. J. phys. Chem. 87, 18481851.Google Scholar
659O'Hara, W. F. & Hepler, L. G. (1961). Thermodynamics of lonization of aqueous meta-chlorophenol. J. phys. Chem. 65, 21072108.Google Scholar
660Olmstead, W. N. & Brauman, J. I. (1977). Gas-phase nucleophilic displacement reactions. J. Am. chem. Soc. 99, 42194228.Google Scholar
661Onsager, L. & Samaras, N. N. T. (1934). The surface tension of Debye-Huckel electrolytes. J. chem. Phys. 2, 528536.Google Scholar
662Out, D. J. P. & Los, J. M. (1980). Viscosity of aqueous solutions of univalent electrolytes from 5 to 95 °C. J. Sol. Chem. 9, 1935.Google Scholar
663Pace, N. C. (1975). The stability of globular proteins. Crit. Rev. Biochem. 3, 143.Google Scholar
664Page, M. I. & Jencks, W. P. (1971). Entropic contributions to rate accelerations in enzymic and intramolecular reactions and the chelate effect. Proc. natn. Acad. Sci. U.S.A. 68, 16781683.Google Scholar
665Pahlman, S., Rosengren, J. & Hjerten, S. (1977). Hydrophobic interaction chromatography on uncharged sepharose derivatives. Effects of neutral salts on the adsorption of proteins. J. Chromat. 131, 99108.Google Scholar
666Paquette, J. & Jolicoeur, C. (1977). A near-infrared study of the hydration of various ions and nonelectrolytes. J. Sol. Chem. 6, 403428.Google Scholar
667Parker, A. J. (1962). The effects of solvation on the properties of anions in dipolar aprotic solvents. Quarterly Reviews 16, 163187.Google Scholar
668Parker, A. J. (1969). Protic-dipolar aprotic solvent effects on rates of bimolecular reactions. Chem. Rev. 69, 132.Google Scholar
669Parodi, R. M., Bianchi, E. & Ciferri, A. (1973). Thermodynamics of unfolding of lysozyme in aqueous alcohol solutions. J. biol. Chem. 248, 40474051.Google Scholar
670Parsegian, V. A., Fuller, N. & Rand, R. P. (1979). Measured work of deformation and repulsion of lecithin bilayers. Proc. natn. Acad. Sci. U.S.A. 76, 27502754.Google Scholar
671Parsegian, V. A. & Rand, R. P. (1983). Membrane interaction and deformation. Ann. N.Y. Acad. Sci. 416, 112.Google Scholar
672Pater, A. & Pater, M. M. (1977). Simultaneous separation of halo-philic proteins and nucleic acids after adsorption onto agarose gels. Can. J. Biochem. 55, 904907.Google Scholar
673Pauling, L. (1946). Molecular architecture and biological reactions. Chem. Eng. News 24, 13751377.Google Scholar
674Pauling, L. (1948). Nature of forces between large molecules of biological interest. Nature 161, 707709.Google Scholar
675Pauling, L. (1960). The Nature of the Chemical Bond, 3rd ed., p. 90. Ithaca, New York: Cornell University Press.Google Scholar
676Pauling, L. (1961). A molecular theory of general anesthesia. Science 134, 1521.Google Scholar
677Pauling, L. (1964). The hydrate microcrystal theory of general anesthesia. Anest. Analg. curr. Res. 43, 110.Google Scholar
678Payzant, J. D., Yamdagni, R. & Kebarle, P. (1971). Hydration of CN, NO2, NO3 and OH in the gas phase. Can. J. Chem. 49, 33083314.Google Scholar
679Peacock, C. K. & Nickless, G. (1969). The dissociation constants of some phosphorus (V) acids. Z. Naturfosch. Ser. A 24, 245.Google Scholar
680Pecsok, R. L. & Saunders, D. (1968). On the mechanism of gel chromatography of inorganic salts. Separation Science 3, 325355.Google Scholar
681Pedley, T. J. (1983). Calculation of unstirred layer thickness in membrane transport experiments: a survey. Q. Rev. Biophys 16, 115150.Google Scholar
682Penefsky, H. S. & Warner, R. C. (1965). Partial resolution of the enzymes catalyzing oxidative phosphorylation. VI. Studies on the mechanism of cold inactivation of mitochondrial adenosine tri-phosphatase. J. biol. Chem. 240, 46944702.Google Scholar
683Persidsky, M. & Richards, V. (1962). Mode of protection with polyvinylpyrrolidone in freezing of bone marrow. Nature 196, 585586.Google Scholar
684Person, W. B. (1962). Thermodynamic properties of donor-acceptor complexes. J. Am. chem. Soc. 84, 536540.Google Scholar
685Petersdorf, R. G., Adams, R. D., Braunwald, E., Isselbacher, K. J., Martin, J. B. & Wilson, J. D. (1983). Harrison's Principles of Internal Medicine, 10th ed.New York: McGraw-Hill.Google Scholar
686Petsko, G. A. (1975). Protein crystallography at sub-zero temperatures: cryoprotective mother liquors for protein crystals. J. molec. Biol. 96, 381392.Google Scholar
687Pharmacia, (1970). Sephadex, Gel Filtration in Theory and Practice. Uppsala: Pharmacia Fine Chemicals AB.Google Scholar
688Philip, P. R. & Jolicoeur, C. (1973). Near-infrared study of the state of water in aqueous solutions of tetraalkylammonium and -phosphonium bromides and alkali halides at 10, 25 and 40°. J. phys. Chem. 77, 30713077.Google Scholar
689Philip, P. R., Perron, G. & Desnoyers, J. E. (1974). Apparent molal volumes and heat capacities of urea and methyl-substituted ureas in H2O and D2O at 25 °C. Can. J. Chem. 52, 17091713.Google Scholar
690Piez, K. A. & Reddi, A. H. (eds) (1984). Extracellular Matrix Biochemistry. New York: Elsevier.Google Scholar
691Pittz, E. P. & Timasheff, S. N. (1978). Interaction of ribonuclease A with aqueous 2-methyl-2,4-pentanediol at pH 5·8. Biochemistry 17, 615623.Google Scholar
692Plyler, E. K. & Barr, E. S. (1938). The change in absorption of water at 4·7 μ due to solutions. J. chem. Phys. 6, 316318.Google Scholar
693Pohl, E. R., Wu, D. & Hupe, D. J. (1980). Effect of solvation on β values for formyl, acetyl, and pivaloyl transfer between sulfur and oxygen nucleophiles. J. Am. chem. Soc. 102, 27592768.Google Scholar
694Pohl, F. M. (1968). Kinetics of reversible denaturation of trypsin in water and water–ethanol mixtures. Eur. J. Biochem. 7, 146152.Google Scholar
695Poillon, W. N. & Bertles, J. F. (1979). Deoxygenated sickle hemoglobin. Effects of lyotropic salts on its solubility. J. Biol. Chem. 254, 34623467.Google Scholar
696Polge, C., Smith, A. U. & Parkes, A. S. (1949). Revival of spermatozoa after vitrification and dehydration at low temperatures. Nature 164, 666.Google Scholar
697Pollard, A. & Wyn Jones, R. G. (1979). Enzyme activities in concentrated solutions of glycinebetaine and other solutes. Planta 144, 291298.Google Scholar
698Porath, J., Sundberg, L., Fornstedt, N. & Olsson, I. (1973). Salting-out in amphiphilic gels as a new approach to hydrophobic adsorption. Nature 245, 465466.Google Scholar
699Pottel, R. (1973). Dielectric properties. In Water. A Comprehensive Treatise, vol. 3: Aqueous Solutions of Simple Electrolytes (ed. Franks, F.), pp. 401455. New York: Plenum Press.Google Scholar
700Pottel, R., Giese, K. & Kaatze, U. (1974). Dielectric relaxation of water in aqueous solutions. In Structure of Water and Aqueous Solutions (ed. Luck, W. A. P.), pp. 391407. New York: Verlag Chemie.Google Scholar
701Pozner, R. I., Shepard, M. L. & Cocks, F. H. (1977). The equilibrium and non-equilibrium thermal behaviour of aqueous ternary Solutions based on complex physiological support media, containing NaCl, and dimethyl sulphoxide or glycerol. J. materials Sci. 12, 299304.Google Scholar
702Prakash, V., Loucheux, C., Scheufele, S., Gorbunoff, M. J. & Timasheff, S. N. (1981). Interactions of proteins with solvent components in 8 M urea. Arch. biochem. Biophys. 210, 455464.Google Scholar
703Prakash, V. & Nandi, P. K. (1977). Association–dissociation behavior of sesame α-globulin in electrolyte solutions. J. biol. Chem. 252, 240243.Google Scholar
704Prakash, V. & Timasheff, S. N. (1981). The calculation of partial specific volumes of proteins in 8 M urea solution. Analyt. Biochem. 117, 330335.Google Scholar
705Prakash, V. & Timasheff, S. N. (1985). Calculation of partial specific volumes of proteins in 8 M urea solution. Meth. Enzymol. 117, 5360.Google Scholar
706Preisler, H. D. & Lyman, G. (1975). Differentiation of erythroleukemia cells in vitro: properties of chemical inducers. Cell Differentiation 4, 179185.Google Scholar
707Pulito, V. L., Miller, D. L., Sassa, S. & Yamane, T. (1983). DNA fragments in friend erythroleukemia cells induced by dimethylsulfoxide. Proc. natn. Acad. Sci. U.S.A. 80, 59125915.Google Scholar
708Pundak, S., Aloni, H. & Eisenberg, H. (1981). Structure and activity of malate dehydrogenase from the extreme halophilic bacteria of the Dead Sea. 2. Inactivation, dissociation and unfolding at NaCl concentrations below 2 M. Salt, salt concentration and temperature dependence of enzyme stability. Eur. J. Biochem. 118, 471477.Google Scholar
709Radnai, T., Palinkas, G., Szasz, Gy. I. & Heinzinger, K. (1981). The second hydration shell of Li+ in aqueous Lil from X-ray and MD studies. Z. Naturforsch. 36A, 10761081.Google Scholar
710Rahmann, H. (1978). Gangliosides and thermal adaption in vertebrates. Japan J. Exp. Med. 48, 8596.Google Scholar
711Rahmann, H. (1979). The possible functional role for gangliosides in synaptic transmission and memory formation. In Biological Aspects of Learning, Memory Formation and Ontogeny of the CNS (ed. Matthies, H., Krug, M. and Popov, N.), pp. 83110. Berlin: Akademie-Verlag.Google Scholar
712Rallo, F., Rodante, F. & Silvestroni, P. (1970). Calorimetric determination of partial molar enthalpies of solution of water and dimethylsulfoxide in their mixtures. Thermochimica acta 1, 311316.Google Scholar
713Ramshaw, J. A. M., Bateman, J. F. & Cole, W. G. (1984). Precipitation of collagens by polyethylene glycols. Analyt. Biochem. 141, 361365.Google Scholar
714Rand, R. P. (1981). Interacting phospholipid bilayers: measured forces and induced structural changes. Ann. Rev. biophys. Bioeng. 10, 277314.Google Scholar
715Rand, R. P., Parsegian, V. A., Henry, J. A. C., Lis, L. J. & McAlister, M. (1980). The effect of cholesterol on measured interaction and compressibility of dipalmitoylphosphatidylcholine bilayers. Can. J. Biochem. 58, 959968.Google Scholar
716Randles, J. E. B. (1957). Ionic hydration and the surface potential of aqueous electrolytes. Discuss. Faraday Soc. 24, 194199.Google Scholar
717Handles, J. E. B. (1963). The interface between aqueous electrolyte solutions and the gas phase. In Advances in Electrochemistry and Electrochemical Engineering, vol. 3: Electrochemistry (ed. Delahay, P.), pp. 130. New York: Wiley-Interscience.Google Scholar
718Randles, J. E. B. & Schiffrin, D. J. (1965). The temperature-dependence of the surface potential of aqueous electrolytes. J. Electroanal. Chem. 10, 480484.Google Scholar
719Rasmussen, D. H. & Mackenzie, A. P. (1972). Effect of solute on ice-solution interfacial free energy: calculation from measured homogeneous nucleation temperatures. In Water Structure at the Water–Polymer Interface (ed. Jellinek, H. H. G.), pp. 126145. New York: Plenum Press.Google Scholar
720Rau, D. C., Lee, B. & Parsegian, V. A. (1984). Measurement of the repulsive force between polyelectrolyte molecules in ionic solution: hydration forces between parallel DNA double helices. Proc. natn. Acad. Sci. U.S.A. 81, 26212625.Google Scholar
721Ray, A. & Nemethy, G. (1971). Effects of ionic protein denaturants on micelle formation by nonionic detergents. J. Amer. Chem. Soc. 93, 67876793.Google Scholar
722Rees, D. C., Lewis, M. & Lipscomb, W. N. (1983). Refined crystal structure of carboxypeptidase A at 1·54 Å resolution. J. molec. Biol. 168, 367387.Google Scholar
723Reiners, G., Lorenz, W. J. & Hertz, H. G. (1978). Tracer diffusion coefficients in aqueous electrolyte solutions of various structure forming and breaking ions. Ber. Bunsenges. Phys. Chem. 82, 738744.Google Scholar
724Reisler, E. & Eisenberg, H. (1969). Interpretation of Equilibrium sedimentation measurements of proteins in guanidine hydrochloride solutions. Partial volumes, density increments, and the molecular weight of the subunits of rabbit muscle aldolase. Biochemistry 8, 45724578.Google Scholar
725Richards, C. D. (1978). Anesthetics and membranes. In Biochemistry of Cell Walls and Membranes, vol. 11 (ed. Metcalfe, J. C.), pp. 157220. International Review of Biochemistry, vol. 19. Baltimore: University Park Press.Google Scholar
726Riebesehl, W., Tomlinson, E. & Grunbauer, H. J. M. (1984). Thermodynamics of solute transfer between alkanes and water. J. phys. Chem. 88, 47754779.Google Scholar
727Rill, R. L., Hilliard, P. R. Jr, & Levy, G. C. (1983). Spontaneous ordering of DNA. Effects of intermolecular interactions on DNA motional dynamics monitored by 13C and 31P nuclear magnetic resonance spectroscopy. J. biol. Chem. 258, 250256.Google Scholar
728Rimerman, R. A. & Hatfield, G. W. (1973). Phosphate-induced protein chromatography. Science 182, 12681270.Google Scholar
729Rinfret, A. P. (1963). Some aspects of preservation of blood by rapid freeze-thaw procedures. Fedn Proc. 22, 94101.Google Scholar
730Roberts, N. K. (1976). Proton diffusion and activity in the presence of electrolytes. J. phys. Chem. 80, 11171120.Google Scholar
731Roberts, N. K. & Northey, H. L. (1974). Proton and deuteron mobility in normal and heavy water solutions of electrolytes. J. Chem. Soc. Faraday Trans. 1, 253262.Google Scholar
732Robinson, D. R. (1972). The determination of activity coefficients from distribution measurements. Meth. Enzymol. 26, 365380.Google Scholar
733Robinson, D. R. & Grant, M. E. (1966). The effects of aqueous salt solutions on the activity coefficients of purine and pyrimidine bases and their relation to the denaturation of deoxyribonucleic acid by salts. J. biol. Chem. 241, 40304042.Google Scholar
734Robinson, D. R. & Jencks, W. P. (1965). The effect of compounds of the urea-guanidinium class on the activity coefficient of acetyltetraglycine ethyl ester and related compounds. J. Am. chem. Soc. 87, 24622470.Google Scholar
735Robinson, J. B., Strottmann, J. M. & Stellwagen, E. (1981). Prediction of neutral salt elution profiles for affinity chromatography. Proc. natn. Acad. Sci. U.S.A. 78, 22872291.Google Scholar
736Robinson, R. A. & Stokes, R. H. (1959). Electrolyte Solutions. The Measurement and Interpretation of Conductance, Chemical Potential and Diffusion in Solutions of Simple Electrolytes, 2nd ed.London: Butterworths Scientific Publications.Google Scholar
737Robinson, R. A. & Stokes, R. H. (1961). Activity coefficients in aqueous solutions of sucrose, mannitol and their mixtures at 25°. J. Phys. Chem. 65, 19541958.Google Scholar
738Roden, L. (1980). Structure and metabolism of connective tissue proteoglycans. In The Biochemistry of Glycoproteins andProteoglycans (ed. Lennarz, W. J.), pp. 267371. New York: Plenum Press.Google Scholar
739Rohdewald, P. & Moldner, M. (1973). Dielectric constants of amide-water systems. J. phys. Chem. 77, 373377.Google Scholar
740Romer, H. & Rahmann, H. (1979). Effects of exogenous neuraminidase on unit activity in frog spinal cord and fish optic tectum. Exp. Brain Res. 34, 4958.Google Scholar
741Roseman, M. & Jencks, W. P. (1975). Interactions of urea and other polar compounds in water. J. Am. chem. Soc. 97, 631640.Google Scholar
742Rosenberg, P. H. (1979). Effects of halothane, lidocaine and 5-hydroxytryptamine on fluidity of synaptic plasma membranes, myelin membranes and synaptic mitochondrial membranes. Naynyn-Schmiedeberg's Arch. Pharmacol. 307, 199206.Google Scholar
743Rosenberg, P. H. (1980). Synaptosomal studies of fluidity changes caused by anesthetics. In Molecular Mechanisms of Anesthesia (ed. Fink, B. R.), pp. 325334. (Progress in Anesthesiology, vol. 2.) New York: Raven Press.Google Scholar
744Rosenberg, P. H., Eibl, H. & Stier, A. (1975). Biphasic effects of halothane on phospholipid and synaptic plasma membranes: a spin label study. Molec. Pharmacol. 11, 879882.Google Scholar
745Rosengren, J., Pahlman, S., Glad, M. & Hjerten, S. (1975). Hydrophobic interaction chromatography on non-charged Sepharose® derivatives. Binding of a model protein, related to ionic strength, hydrophobicity of the substituent, and degree of substitution (determined by NMR). Biochim. biophys. Acta 412, 5161.Google Scholar
746Rossky, P. J. & Karplus, M. (1979). Solvation. A molecular dynamics study of a dipeptide in water. J. Am. chem. Soc. 101, 19131937.Google Scholar
747Rothstein, F., Rosenoer, V. M. & Hughes, W. L. (1977). Current concepts concerning albumin purification. In Albumin Structure, Function and Uses (ed. Rosenoer, V. M., Oratz, M., and Rothschild, M. A.), pp. 725. New York: Pergamon Press.Google Scholar
748Roux, G., Roberts, D., Perron, G. & Desnoyers, J. E. (1980). Microheterogeneity in aqueous-organic solutions: heat capacities, volumes and expansibilities of some alcohols, aminoalcohol and tertiary amines in water. J. Solution Chem. 9, 629647.Google Scholar
749Rubin, R. A. & Earp, H. S. (1983). Solubilization of EGF receptor with trition X-100 alters stimulation of tyrosine residue phos-phorylation by EGF and dimethyl sulfoxide. J. biol. Chem. 258, 51775182.Google Scholar
750Russegger, P., Lischka, H. & Schuster, P. (1972). Model calculations of the solvation of oneatomic ions (LCAO-MO-investigations of molecular structures VIII). Theoret. chim. acta (Beri.) 24, 191200.Google Scholar
751Ruthven, D. M. (1984). Principles of adsorption and adsorption processes. New York: Wiley-Interscience.Google Scholar
752Samoilov, O. Ya. (1957 a) (1965). Structure of Aqueous Electrolyte Solutions and the Hydration of Ions. New York: Consultants Bureau.Google Scholar
753Samoilov, O. Ya. (1957 b). A new approach to the study of hydration of ions in aqueous solutions. Discuss. Faraday Soc. 24, 141146.Google Scholar
754Samoilov, O. Ya. (1972). Residence times of ionic hydration. In Water and Aqueous Solutions. Structure, Thermodynamics, and Transport Processes (ed. Home, R. A.), pp. 597612. New York: Wiley-Interscience.Google Scholar
755Saunders, D. & Pecsok, R. L. (1968). Calculation of distribution coefficients in inorganic gel chromatography. Analyt. Chem. 40, 4448.Google Scholar
756Savage, J. J. & Wood, R. H. (1976). Enthalpy of dilution of aqueous mixtures of amides, sugars, urea, ethylene glycol, and pentaerythritol at 25 °C: enthalpy of interaction of the hydrocarbon, amide, and hydroxyl functional groups in dilute aqueous solutions. J. Sol. Chem. 5, 733750.Google Scholar
757Saylor, J. H., Whitten, A. I., Claiborne, I. & Gross, P. M. (1952). The solubilities of benzene, nitrobenzene and ethylene chloride in aqueous salt solutions. J. Am. chem. Soc. 74, 17781781.Google Scholar
758Scatchard, G., Hamer, W. J. & Wood, S. E. (1938). Isotonic solutions. I. The chemical potential of water in aqueous solutions of sodium chloride, potassium chloride, sulfuric acid, sucrose, urea and glycerol at 25°. J. Am. chem. Soc. 60, 30613070.Google Scholar
759Schaffer, S. W., Ahmed, A. K. & Wetlaufer, D. B. (1975). Salt effects in the glutathione-facilitated reactivation of reduced bovine pancreatic ribonuclease. J. biol. Chem. 250, 84838486.Google Scholar
760Schauer, R. (1982 a). Chemistry, metabolism, and biological functions of sialic acids. Adv. in Carbohydrate Chem. and Biochem. 40, 131234.Google Scholar
761Schauer, R. (ed.) (1982 b). Sialic Acids: Chemistry, Metabolism and function (Cell Biology Monographs, vol. 10). New York: Springer-Verlag.Google Scholar
762Schellman, J. A. (1953). The thermodynamics of urea solutions and the heat of formation of the peptide hydrogen bond. C.R. Trav. Lab. Carlsberg 29, 223229.Google Scholar
763Scher, W., Preisler, H. D. & Friend, C. (1973). Hemoglobin synthesis in murine virus-induced leukemic cells in vitro III. Effects of 5-bromo-2′-deoxyuridine, dimethylformamide and dimethylsul-foxide. J. cell. Physiol. 81, 6370.Google Scholar
764Schleich, T., Gentzler, R. & Von Hippel, P. H. (1968). Proton exchange of N–methylacetamide in concentrated aqueous electrolyte solutions I. Acid catalysis. J. Am. chem. Soc., 90, 59545960.Google Scholar
765Schleich, T., Rollefson, B. & Von Hippel, P. H. (1971). Proton exchange of N–methylacetamide in concentrated aqueous electrolyte solutions. II. Acid catalysis in water-dioxane mixtures and base catalysis. J. Am. chem. Soc. 93, 70707074.Google Scholar
766Schleich, T. & Von Hippel, P. H. (1969). Specific ion effects on the solution conformation of poly-L-proline. Biopolymers 7, 861877.Google Scholar
767Schobert, B. (1982). Evidence for a protein stabilizing mechanism in plant cells under water stress. In Biophysics of Water (ed. Franks, F. and Mathias, S. F.), pp. 309311. New York: Wiley-Interscience.Google Scholar
768Schoffeniels, E. (1976). Adaptions with respect to salinity. Biochem. Soc. Symp. 41, 179204.Google Scholar
769Schott, H., Royce, A. E. & Han, S. K. (1984). Effect of inorganic additives on solutions of nonionic surfactants VII. Cloud point shift values of individual ions. J. Colloid and Interface Sci. 98, 196201.Google Scholar
770Schrier, E. E., Ingwall, R. T. & Scheraga, H. A. (1965). The effect of aqueous alcohol solutions on the thermal transition of ribonuclease. J. phys. Chem. 69, 298303.Google Scholar
771Schrier, E. E. & Schrier, E. B. (1967). The salting-out behavior of amides and its relation to the denaturation of proteins by salts. J. phys. Chem. 71, 18511860.Google Scholar
772Schryvers, A. & Weiner, J. H. (1981). The anaerobic sn-glycerol-3-phosphate dehydrogenase of Escherichia coli. Purification and characterization. J. biol. Chem. 256, 99599965.Google Scholar
773Schultz, J. W. & Hornig, D. F. (1961). The effect of dissolved alkali halides on the raman spectrum of water. J. phys. Chem., 65, 21312138.Google Scholar
774Seeman, P. M. (1972). The membrane actions of anesthetics and tranquilizers. Pharmac. Rev. 24, 583655.Google Scholar
775Seeman, P. M. (1966). Membrane stabilization by drugs: tranquilizers, steroids, and anesthetics. Int. Rev. Neurobiol. 9, 145221.Google Scholar
776Sellin, S., Aronsson, A.-C. & Mannervik, B. (1980). A method based on the use of methanol as a stabilizing agent to prepare metal-free glyoxalase I and to reconstitute activity by addition of bivalent metal ions. Acta chem. scand. 634, 541543.Google Scholar
777Sergeeva, V. F. (1965). Salting-out and salting-in of non-electrolytes. Russ. chem. Revs 34, 309318.Google Scholar
778Setschenow, J. (1889). Über die Konstitution der Salzlosungen auf Grund ihres Verhaltens zu Kohlensaure. Z. phys. Chem. 4, 117125.Google Scholar
779Setschenow, J. (1891). Action de l'acide carbonique sur les solutions des sels a acides forts. Annales de Chimie et du Physique 25, 226270.Google Scholar
780Shames, S. L. & Byers, L. D. (1981 a). Acyl substituent effects on rates of acyl transfer to thiolate, hydroxide, and oxy dianions. J. Am. chem. Soc. 103, 61706177.Google Scholar
781Shaner, S. L., Piatt, D. M., Wensley, C. G., Yu, H., Burgess, R. R. & Record, M. T. Jr., (1982). Aggregation equilibria of Escherichia coli RNA polymerase: evidence for anion-linked conformational transitions in the protomers of core and holoenzyme. Biochemistry 21, 55395551.Google Scholar
782Shibata, A., Yamashita, S. & Yamashita, T. (1982). Specific anion effects on water structure at polypeptide monolayer-water interface. Bull. chem. Soc. Japan 55, 28142819.Google Scholar
783Shin, M., Sakihama, N., Oshino, R. & Sasaki, H. (1984). Butyltoyopearl 650 as a new hydrophobic adsorbent for water-soluble enzyme proteins. Analyt. Biochem. 138, 259261.Google Scholar
784Shoolery, J. N. & Alder, B. J. (1955). Nuclear magnetic resonance in concentrated aqueous electrolytes. J. chem. Phys. 23, 805811.Google Scholar
785Shukuya, R. & Schwert, G. W. (1960). Glutamic acid decarboxylase III. The inactivation of the enzyme at low temperatures. J. biol. Chem. 235, 16581661.Google Scholar
786Sigel, H. & Martin, R. B. (1982). Coordinating properties of the amide bond. Stability and structure of metal ion complexes of peptides and related ligands. Chem. Rev. 82, 385426.Google Scholar
787Sihag, R. K. & Deutscher, M. P. (1983). Perturbation of the aminoacyl-tRNA synthetase complex by salts and detergents. Importance of hydrophobic interactions and possible involvement of lipids. J. biol. Chem. 258, 1184611850.Google Scholar
788Silberberg, A. (1962 a). The adsorption of flexible macromolecules. Part II. The shape of the adsorbed molecule; the adsorption isotherm surface tension, and pressure. J. phys. Chem. 66, 18841907.Google Scholar
789Silberberg, A. (1962 b). The adsorption of flexible macromolecules Part I. The isolated macromolecule at a plane interface. J. phys. Chem. 66, 18721883.Google Scholar
790Simpson, R. B. & Kauzmann, W. (1953). The kinetics of protein denaturation I. The behavior of the optical rotation of ovalbumin in urea solutions. J. Am. chem. Soc. 75, 51395152.Google Scholar
791Sinibaldi, M. & Lederer, M. (1975). Adsorption of inorganic anions on Sephadex gels. J. Chromat. 107, 210212.Google Scholar
792Skinner, M. & Cohen, A. S. (1983). Amyloidosis: clinical, pathologic, and biochemical characteristics. In Connective Tissue Diseases (ed. Wagner, B. M., Fleischmajer, R. and Kaufman, N.), pp. 97119. Baltimore: Williams & Wilkins.Google Scholar
793Smith, M. & Symons, M. C. R. (1958). Solvation spectra. Part I – The effect of environmental changes upon the ultra-violet absorption of solvated iodide ions. Trans. Faraday Soc. 54, 338345.Google Scholar
794Smith, R. L., Macara, I. G., Levenson, R., Housman, D. & Cantley, L. (1982). Evidence that a Na+/Ca2+ antiport system regulates murine erythroleukemia cell differentiation. J. biol. Chem. 257, 773780.Google Scholar
795Sohnel, O. & Novotny, P. (1985). Densities of Aqueous Solutions of Inorganic Substances. New York: Elsevier.Google Scholar
796Somero, G. N., Neubauer, M. & Low, P. S. (1977). Neutral salt effects on the velocity and activation volume of the lactate dehydrogenase reaction: evidence for enzyme hydration changes during catalysis. Arch. Biochem. Biophys. 181, 438446.Google Scholar
797Soupart, P. & Clewe, T. H. (1965). Sperm penetration of rabbit zona pellucida inhibited by treatment of ova with neuraminidase. Pert. Steril. 16, 677689.Google Scholar
798Span, J. & Lapanje, S. (1973). Solvation of β-lactoglobulin and chymotrypsinogen A in aqueous urea solutions. Biochim. biophys. Acta 295, 371378.Google Scholar
799Span, J., Lenarcic, S. & Lapanje, S. (1974). Solvation of lysozyme and β-lactoglobulin in aqueous guanidine hydrochloride solutions. Biochim. biophys. Acta 359, 311319.Google Scholar
800Srere, P. A. (1980). The infrastructure of the mitochondrial matrix. Trends Biochem. Sci. 5, 120121.Google Scholar
801Srere, P. A. (1981). Protein crystals as a model for mitochondrial matrix proteins. Trends Biochem. Sci. 6, 47.Google Scholar
802Srere, P. A. (1984). Why are enzymes so big? Trends Biochem. Sci. 9, 387390.Google Scholar
803Srinivasan, R. & Ruckenstein, E. (1980). Role of physical forces in hydrophobic interaction chromatography. Separation and Purification Methods 9, 267370.Google Scholar
804Stein, J. H., Cline, M. J., Daly, W. J., Easton, J. D., Hutton, J. J., Kohler, P. O., O'Rourke, R. A., Sande, M. A., Trier, J. S. & Zvaifler, N. J. (eds.) (1983). Internal Medicine, 1st ed.Boston: Little Brown.Google Scholar
805Stern, J. H. & Hubler, P. M. (1984). Hydrogen bonding in aqueous solutions of D-ribose and 2-deoxy-D-ribose. J. phys. Chem. 88, 16801681.Google Scholar
806Stewart, R. (1985). The Proton: Applications to Organic Chemistry. New York: Academic Press.Google Scholar
807Stillinger, F. H. (1980). Water revisited. Science 209, 451456.Google Scholar
808Stokes, R. H. (1967). Thermodynamics of aqueous urea solutions. Aust. J. Chem. 20, 20872100.Google Scholar
809Stokes, R. H. & Mills, R. (1965). Viscosity of Electrolytes and Related Properties. New York: Pergamon Press.Google Scholar
810Storey, K. B. & Storey, J. M. (1983). Biochemistry of Freeze Tolerance in Terrestrial Insects. Trends Biochem. Sci. 8, 242245.Google Scholar
811St. Pierre, T. & Jencks, W. P. (1969). Interactions of salts and denaturing agents with a polyacrylamide gel. Arch. biochem. Biophys. 133, 99102.Google Scholar
812Streitwieser, A. Jr., (1985). Driving force and nucleophilicity in SN2 displacements. Proc. natn. Acad. Sci. U.S.A. 82, 82888290.Google Scholar
813Subramanian, S. & Fisher, H. F. (1972). Near-infrared spectral studies on the effects of perchlorate and tetrafluoroborate ions on water structure. J. phys. Chem. 76, 8489.Google Scholar
814Suda, S., Enomoto, S., Abe, E. & Suda, T. (1984). Inhibition by 1α,25-dihydroxyvitamin D3 of dimethyl sulfoxide-induced differentiation of friend erythroleukemia cells. Biochem. biophys. Res. Comm. 119, 807813.Google Scholar
815Sudholter, E. J. R. & Engberts, J. B. F. N. (1979). Salt effects on the critical micellar concentration, iodide counterion binding, and surface micropolarity of 1-methyl-4-dodecylpyridinium iodide micelles. J. Phys. Chem. 83, 18541859.Google Scholar
816Sunner, J. & Kebarle, P. (1984). Ion-solvent molecule interactions in the gas phase. The potassium ion and Me2SO, DMA, DMF and acetone. J. Am. chem. Soc. 106, 61356139.Google Scholar
817Swain, C. G. & Bader, R. F. W. (1960). The nature of the structure difference between light and heavy water and the origin of the solvent isotope effect-I. Tetrahedron 10, 182199.Google Scholar
818Swain, C. G., Swain, M. S., Powell, A. L. & Alunni, S. (1983). Solvent effects on chemical reactivity. Evaluation of anion and cation solvation components. J. Am. chem. Soc. 105, 502513.Google Scholar
819Swann, A. C. (1983). Brain (Na+, Ka+)-ATPase. Opposite effects of ethanol and dimethyl sulfoxide on temperature dependence of enzyme conformation and univalent cation binding. J. biol. Chem. 258, 1178011786.Google Scholar
820Swezey, R. R. & Somero, G. N. (1982). Polymerization thermodynamics and structure stabilities of skeletal muscle actins from vertebrates adapted to different temperatures and hydrostatic pressures. Biochemistry 21, 44964503.Google Scholar
821Symons, M. C. R. (1975). Water structure and hydration. Phil. Trans. R. Soc. Lond. B272 1328.Google Scholar
822Symons, M. C. R. (1981). Water structure and reactivity. Accts. Chem. Res. 14, 179187.Google Scholar
823Symons, M. C. R. (1983). Structure in solvents and solutions-NMR and vibrational spectroscopic studies. Chem. Soc. Reviews 12, 134.Google Scholar
824Symons, M. C. R. & Blandamer, M. J. (1968). Spectroscopic and ultrasonic relaxation studies of the water-t–butyl alcohol system. In Hydrogen-Bonded Solvent Systems (ed. Covington, A. K. and Jones, P.), pp. 211220. London: Taylor and Francis.Google Scholar
825Symons, M. C. R., Harvey, J. M. & Jackson, S. E. (1980 b). Spectroscopic studies of water-aprotic-solvent interactions in the waterrich region. J. Chem. Soc. Faraday Trans. I 76, 256265.Google Scholar
826Szasz, Gy. I. & Heinzinger, K. (1983). A molecular dynamics study of the translational and rotational motions in an aqueous Lil solution. J. chem. Phys. 79, 34673473.Google Scholar
827Szasz, Gy. I., Heinzinger, K. & Palinkas, G. (1981 a). The structure of the hydration shell of the lithium ion. Chem. phys. Lett. 78, 194196.Google Scholar
828Szasz, Gy. I., Heinzinger, K. & Riede, W. O. (1981 b). Structural properties of an aqueous Lil solution derived from a molecular dynamics simulation. Z. Naturforsch. 36A, 10671075.Google Scholar
829Szasz, Gy. I., Heinzinger, K. & Riede, W. O. (1981 c). Self-diffusion and reorientational motion in an aqueous Lil solution. A molecular dynamics study. Ber. Bunsenges. Phys. Chem. 85, 10561059.Google Scholar
830Taborsky, G. (1979). Protein alterations at low temperatures: an overview. In Proteins at Low Temperatures (ed. Fennema, O.), pp. 126. Advances in Chemistry Series, vol. 180. Washington, D.C.: American Chemical Society.Google Scholar
831Tabushi, I., Kobuke, Y. & Imuta, J. (1981). Lipophilic diammonium cation having a rigid structure complementary to pyrophosphate dianions of nucleotides. Selective extraction and transport of nucleotides. J. Am. chem. Soc. 103, 61526157.Google Scholar
832Takemori, S., Furuya, E., Suzuki, H. & Katagiri, M. (1967). Stabilization of enzyme activity by an organic solvent. Nature 215, 417419.Google Scholar
833Tamas, J. & Ujszaszy, K. (1966). The diffusion of H218O molecules in concentrated aqueous solutions of salts II. Acta Chim. Hung. 49, 377393.Google Scholar
834Tan, K. H. & Lovrien, R. (1972). Enzymology in aqueous-organic cosolvent binary mixtures. J. biol. Chem. 247, 32783285.Google Scholar
835Tanabe, K. (1984). Raman spectroscopic study of hydrophobic hydration of organic molecules in aqueous solution. J. Inclusion Phenomena 2, 267273.Google Scholar
836Tanaka, H., Nakanishi, K. & Nishikawa, K. (1984). Clathrate-like structure of water around some nonelectrolytes in dilute solution as revealed by computer simulation and X-ray diffraction studies. J. Inclusion Phenomena 2, 119126.Google Scholar
837Tanaka, M., Levy, J., Terada, M., Breslow, R., Rifkind, R. A. & Marks, P. A. (1975). Induction of erythroid differentiation in murine virus infected erythroleukemia cells by highly polar com pounds. Proc. natn. Acad. Sci. U.S.A. 72, 10031006.Google Scholar
838Tanford, C. (1968). Protein denaturation. Adv. Prot. Chem. 23, 121282.Google Scholar
839Tanford, C. (1979). Interfacial free energy and the hydrophobic effect. Proc. natn. Acad. Sci. U.S.A. 76, 41754176.Google Scholar
840Taylor, J. B. & Rowlinson, J. S. (1955). The thermodynamic properties of aqueous solutions of glucose. Trans. Faraday Soc. 51, 11831192.Google Scholar
841Taylor, R. P. & Kuntz, I. D. Jr., (1972). Proton acceptor abilities of anions and possible relevance to the Hofmeister series. J. Amer. Chem. Soc. 94, 79637965.Google Scholar
842Teeter, M. M. (1984). Water structure of a hydrophobic protein at atomic resolution: pentagon rings of water molecules in crystals of crambin. Proc. natn. Acad. Sci. U.S.A. 81, 60146018.Google Scholar
843Teissie, J., Prats, M., Soucaille, P. & Tocanne, J. F. (1985). Evidence for conduction of protons along the interface between water and a polar lipid monolayer. Proc. natn. Acad. Sci. U.S.A. 82, 32173221.Google Scholar
844Tellam, R. L., Sculley, M. J., Nichol, L. W. & Wills, P. R. (1983). The influence of poly(ethylene glycol) 6000 on the properties of skeletal-muscle actin. Biochem. J. 213, 651659.Google Scholar
845Tessman, J. R., Kahn, A. H. & Shockley, W. (1953). Electronic polarizabilities of ions in crystals. Phys. Rev. 92, 890895.Google Scholar
846Tettamanti, G., Preti, A., Cestaro, B., Masserini, M., Son-Nino, S. & Ghidoni, R. (1980). Gangliosides and associated enzymes at the nerve-ending membranes. In Cell Surface Glycolipids (ed. Sweeley, C. C.), pp. 321343. Washington, D.C.: American Chemical Society.Google Scholar
847Timasheff, S. N. (1970). Protein-solvent interactions and protein conformation. Accts. Chem. Res. 3, 6268.Google Scholar
848Timasheff, S. N. & Inoue, H. (1968). Preferential binding of solvent components to proteins in mixed water-organic solvent systems. Biochemistry 7, 25012513.Google Scholar
849Timasheff, S. N., Lee, J. C., Pittz, E. P. & Tweedy, N. (1976). The interaction of tubulin and other proteins with structure-stabilizing solvents. J. Colloid and Interface Sci. 55, 658663.Google Scholar
850Traube, J. (1910). The attraction pressure. J. Physical Chem. 14, 452470.Google Scholar
851Tsopanakis, A., Tsopanakis, C. & Shall, S. (1978). Useof subzero temperatures and aqueous/organic solvent systems to increase the stability of labile enzymes. Biochemical Soc. Trans. 6, 12821285.Google Scholar
852Ueberreiter, K. (1980 a). Change of water structure by polyalcohols. Density and viscosity measurements, 1. Water/aliphatic alcohol solutions. Makromol. Chem. Rapid. Commun., 1, 139142.Google Scholar
853Ueberreiter, K. (1980 b). Change of water structure by polyalcohols. Density and viscosity measurements, 2. Water/polyalcohol solutions. Makromol. Chem. Rapid. Commun., 1, 143147.Google Scholar
854Ueda, M., Katayama, A., Kuroki, N. & Urahata, T. (1976 a). Effect of urea and p–toluenesulfonic acid on the solubility of toluene in water-ethylene glycol mixture. Colloid & Polymer Sci. 254, 417420.Google Scholar
855Ueda, M., Katayama, A., Kuroki, N. & Urahata, T. (1976 b). Effect of glycerol on the solubilities of benzene and toluene in water. Colloid & Polymer Sci. 254, 532533.Google Scholar
856Ueda, M., Katayama, A., Urahata, T. & Kuroki, N. (1978 a). Effect of sorbitol and inositol on the solubility of toluene in water. Colloid & Polymer Sci. 256, 10321033.Google Scholar
857Ueda, M., Katayama, A., Kuroki, N. & Urahata, T. (1978 b). Effect of urea on the solubility of benzene and toluene in water. Progr. in Colloid & Polymer Sci. 63, 116119.Google Scholar
858Ueda, M., Lim, Y., Urahata, T. & Kuroki, N. (1980). Effects of ureas on the aqueous solubilities of aromatic hydrocarbons. Kagaku to Kogyo (Osaka) 54, 104108.Google Scholar
859Uedaira, H. (1980). Structure and function of water in biological systems. In Water and Metal Cations in Biological Systems (ed. Pullman, B. and Yagi, K.), pp. 4756. Tokyo: Japan Scientific Societies Press.Google Scholar
860Ueno, Y., Yoza, N. & Ohashi, S. (1970). Gel Chromatographic behavior of some metal ions. J. Chromat. 52, 321327.Google Scholar
861Ujimoto, K. & Kurihara, H. (1981). Mechanism of separation of aliphatic alcohols in aqueous dextran gel systems. J. Chromat. 208, 183200.Google Scholar
862Varma, R., Michos, G. A., Mesmer, R. E., Varma, R. S. & Shirey, R. E. (1983). Beta-glucuronidase in sera of patients with epileptic seizure activity, diabetes and some other disease states. Neuroscience Letters 39, 105111.Google Scholar
863Varma, R., Michos, G. A., Varma, R. S., Stolar, J., Mesmer, R. & Joy, C. R. (1982). Serum beta-glucuronidase in epilepsy. Res. Commun. PsychoL., Psychiat. & Behav. 7, 377380.Google Scholar
864Varma, S. D. & Kinoshita, J. H. (1974). Sorbitol pathway in diabetic and galactosemic rat lens. Biochim. biophys. Acta 338, 632640.Google Scholar
865Varma, R. & Varma, R. S. (1983). Mucopolysaccharides – Glyco-saminoglycans – of Body Fluids in Health and Disease. New York: Walter de Gruyter.Google Scholar
866Veis, A. & Nawrot, C. F. (1970). Basicity differences among peptide bonds. J. Am. chem. Soc. 92, 39103914.Google Scholar
867Verkman, A. S. & Dix, J. A. (1984). Effect of unstirred layers on binding and reaction kinetics at a membrane surface. Anal. Biochem. 142, 109116.Google Scholar
868Verrall, R. E. (1973). Infrared spectroscopy of aqueous electrolyte solutions. In Water. A Comprehensive Treatise, vol. 3: Aqueous Solutions of Simple Electrolytes (ed. Franks, F.), pp. 211264. New York: Plenum Press.Google Scholar
869Voelkel, J. (1981). Salt effect during the swelling and dissolution of poly(vinyl alcohol). Influence on the nature of ions. Polish J. Chem. 55, 445455.Google Scholar
870Voet, A. (1936). Ionic radii and heat of hydration. Trans. Faraday Soc. 32, 13011304.Google Scholar
871Voet, A. (1937). Quantitative lyotropy. Chem. Rev. 20, 169179.Google Scholar
872Von Der Haar, F. (1976). Purification of proteins by fractional interfacial salting out on unsubstituted agarose gels. Biochem. biophys. Res. Comm. 70, 10091013.Google Scholar
873Von Der Haar, F. (1978). Interfacial salting out and the ligand induced solubility shift: another affinity technique in purification of proteins. In Theory and Practice in Affinity Techniques (ed. Sundaram, P. V. and Eckstein, F.), pp. 113. New York: Academic Press.Google Scholar
874Von Hippel, P. H. (1975). Neutral salt effects on the conformational stability of biological macromolecules. In Protein–Ligand Interactions (ed. Sund, H. and Blauer, G.), pp. 452471. New York: Walter de Gruyter.Google Scholar
875Von Hippel, P. H. & Hamabata, A. (1973). Model studies on the effects of neutral salts on the conformational stability of biological macromolecules. J. Mechanochem. Cell Mobility 2, 127138.Google Scholar
876Von Hippel, P. H., Peticolas, V., Schack, L. & Karlson, L. (1973). Model studies on the effects of neutral salts on the conformational stability of biological macromolecules. I. Ion binding to polyacrylamide and polystyrene columns. Biochemistry 12, 12561264.Google Scholar
877Von Hippel, P. H. & Schleich, T. (1969 a). The effects of neutral salts on the structure and conformational stability of macromolecules in solution. In Structure and Stability of Biological Macromolecules (ed. Timasheff, S. N., and Pasman, G. D.), pp. 417574. New York: Marcel Dekker.Google Scholar
878Von Hippel, P. H. & Schleich, T. (1969 b). Ion effects on the solution structure of biological macromolecules. Accts. Chem. Res. 2, 257265.Google Scholar
879Von Hippel, P. H. & Wong, K.-Y. (1962). The effect of ions on the kinetics of formation and the stability of the collagen-fold. Biochemistry 1, 664674.Google Scholar
880Von Hippel, P. H. & Wong, K.-Y. (1965). On the conformational stability of globular proteins. The effects of various electrolytes and nonelectrolytes on the thermal ribonuclease transition. J. biol. Chem. 240, 39093923.Google Scholar
881Wagner, B. M., Fleischmajer, R. & Kaufman, N. (eds.) (1983). Connective Tissue Diseases. Baltimore: Williams & Wilkins.Google Scholar
882Wagner, C. (1924). The surface tension of dilute solutions of electrolytes. Physik. Z. 25, 474477.Google Scholar
883Waldron, R. D. (1957). Infrared spectra of HDO in water and ionic solutions. J. chem. phys. 26, 809814.Google Scholar
884Walrafen, G. E. (1962). Raman spectral studies of the effects of electrolytes on water. J. chem. Phys. 36, 10351042.Google Scholar
885Walrafen, G. E. (1966). Raman spectral studies of the effects of urea and sucrose on water structure. J. chem. Phys. 44, 37263727.Google Scholar
886Walrafen, G. E. (1970). Raman spectral studies of the effects of perchlorate ion on water structure. J. chem. Phys. 52, 41764198.Google Scholar
887Walrafen, G. E. (1971). Raman spectral studies of the effects of solutes and pressure on water structure. J. chem. Phys. 55, 768792.Google Scholar
888Warren, J. C. & Cheatum, S. G. (1966). Effect of neutral salts on enzyme activity and structure. Biochemistry 5, 17021706.Google Scholar
889Warren, J. C., Stowring, L. & Morales, M. F. (1966). The effect of structure-disrupting ions on the activity of myosin and other enzymes. J. biol. Chem. 241, 309316.Google Scholar
890Warshel, A. (1978 a). A microscopic model for calculations of chemical processes in aqueous solutions. Chem. Phys. Lett. 55, 454458.Google Scholar
891Warshel, A. (1978 b). Energetics of enzyme catalysis. Proc. natn. Acad. Sci. U.S.A. 75, 52505254.Google Scholar
892Warshel, A. (1979). Calculations of chemical processes in solutions. J. Phys. Chem. 83, 16401652.Google Scholar
893Warshel, A. (1980). An empirical valence bond approach for comparing reactions in solutions and in enzymes. J. Am. chem. Soc. 102, 62186226.Google Scholar
894Warshel, A. (1981). Electrostatic basis of structure-function correlation in proteins. Accts. Chem. Res. 14, 284290.Google Scholar
895Warshel, A. & Russell, R. T. (1984). Calculations of electrostatic interactions in biological systems and in solutions. Quarterly Rev. Biophys. 17, 283422.Google Scholar
896Washabaugh, M. W. & Collins, K. D. (1983). Purification of aqueous ethylene glycol. Analyt. Biochem. 134, 144152.Google Scholar
897Washabaugh, M. W. & Collins, K. D. (1986 a). The systematic characterization by aqueous column chromatography of solutes which affect protein stability. J. Biol. Chem. 261, in press.Google Scholar
898Washabaugh, M. W. & Collins, K. D. (1986 b). Dihydroorotase from Escherichia coli. Sulfhydryl group – metal ion interactions. J. biol. Chem. 261, 59205929.Google Scholar
899Watenpaugh, K. D., Margulis, T. N., Sieker, L. C. & Jensen, L. H. (1978). Water structure in a protein crystal: rubredoxin at 1·2 Å resolution. J. molec. Biol. 122, 175190.Google Scholar
900Weeks, J. L., Meaburn, G. M. A. C. & Gordon, S. (1963). Absorption coefficients of liquid water and aqueous solutions in the far ultraviolet. Radiation Research 19, 559567.Google Scholar
901Weiss, L. (1965). Studies on cell deformability. I. Effect of surface charge. J. Cell Biol. 26, 735739.Google Scholar
902Wen, W.-Y. (1982). Hydration of some solutes in aqueous solutions. In Ions and Molecules in Solution; Studies in Physical and Theoretical Chemistry, vol. 27 (ed. Tanaka, N., Ohtaki, H. and Tamamushi, R.), pp. 4559. Amsterdam: Elsevier Science Publishers.Google Scholar
903Weston, R. E. Jr. (1962). Raman spectra of electrolyte solutions in light and heavy water. Spectrochimica acta 18, 12571277.Google Scholar
904Wetlaufer, D. B., Malik, S. K., Stoller, L. & Coffin, R. L. (1964). Nonpolar group participation in the denaturation of proteins by urea and guanidinium salts. Model compound studies. J. Am. chem. Soc. 86, 508513.Google Scholar
905Wheaton, R. M. & Bauman, W. C. (1951). Properties of strongly basic anion exchange resins. Ind. Eng. Chem. 43, 10881093.Google Scholar
906White, D. C. & Dundas, C. R. (1970). Effect of anaesthetics on emission of light by luminous bacteria. Nature 226, 456458.Google Scholar
907White, J. C. (1985). The risks of excessive water drinking in epileptic patients. New Eng. J. Med. 312, 246247.Google Scholar
908White, S. H. & King, G. I. (1985). Molecular packing and area compressibility of lipid bilayers. Proc. natn. Acad. Sci. U.S.A. 82, 65326536.Google Scholar
909Wilf, J. & Minton, A. P. (1981). Evidence for protein self-association induced by excluded volume myoglobin in the presence of globular proteins. Biochim. biophys. Acta 670, 316322.Google Scholar
910Wilhelm, E., Battino, R. & Wilcock, R. J. (1977). Low-pressure solubility of gases in liquid water. Chem. Rev. 77, 219262.Google Scholar
911Williams, R. J. & Harris, D. (1977). The distribution of cryoprotective agents into lipid interfaces. Cryobiology 14, 670680.Google Scholar
912Williams, W. J. (1979). Handbook of Anion Determination. London: Butterworths.Google Scholar
913Wilson, N. & Greenhouse, V. Y. (1976). Chromatography of Na 125I on Sephadex gels. J. Chromat. 118, 7582.Google Scholar
914Winstein, S. & Fainberg, A. H. (1957). Correlation of solvolysis rates. IV. Solvent effects on enthalpy and entropy of activation for solvolysis of t–butyl chloride. J. Am. chem. Soc. 79, 59375950.Google Scholar
915Winstein, S., Savedoff, L. G., Smith, S., Stevens, I. D. R. & Gall, J. S. (1960). Ion pairs, nucleophilicity and salt effects in bimolecular nucleophilic substitution. Tetrahedron Lett. 2430.Google Scholar
916Wolfenden, R. (1969). Transition state analogues for enzyme catalysis. Nature 223, 704705.Google Scholar
917Wolfenden, R. (1976). Transition state analog inhibitors and enzyme catalysis. Ann. Rev. biophys. Bioeng. 5, 271306.Google Scholar
918Wolfenden, R. (1978). Interaction of the peptide bond with solvent water: a vapor phase analysis. Biochemistry 17, 201204.Google Scholar
919Wolfenden, R. (1983). Waterlogged molecules. Science 222, 10871093.Google Scholar
920Worley, J. D. & Klotz, I. M. (1966). Near-infrared spectra of H2O-D2O solutions. J. chem. Phys. 45, 28682871.Google Scholar
921Wyn Jones, R. G. & Pollard, A. (1982). Towards a physical chemical characterization of compatible solutes. In Biophysics of Water (ed. Franks, F. and Mathias, S. F.), pp. 335339. New York: Wiley-Interscience.Google Scholar
922Wynne-Jones, W. F. K. & Eyring, H. (1935). The absolute rate of reactions in condensed phases. J. chem. Phys. 3, 492502.Google Scholar
923Yamashita, T., Shibata, A. & Yamashita, S. (1978). Effect of water structure on poly-є-benzyloxycarbonyl-L-lysine monolayers. Chemistry Letters 1112.Google Scholar
924Yancey, P. H., Clark, M. E., Hand, S. C., Bowlus, R. D. & Somero, G. N. (1982). Living with water stress: evolution of os-molyte systems. Science 217, 12141222.Google Scholar
925Yang, C.-H., Brown, J. N. & Kopple, K. D. (1979). Peptide-water association in peptide crystals. Int. J. Peptide Protein Res. 14, 1220.Google Scholar
926Yano, Y. & Janado, M. (1980). Hydrophobic interaction chromatography of aliphatic alcohols on unsubstituted Sephadex gels with high dextran concentrations. J. Chromat. 200, 125136.Google Scholar
927Yeh, Y. & Feeney, R. E. (1978). Anomalous depression of the freezing temperatures in a biological system. Aces. Chem. Res. 11, 129135.Google Scholar
928Yon, R. J. & Simmonds, R. J. (1975). Protein chromatography on adsorbents with hydrophobic and ionic groups. Some properties of N–(3-carboxypropionyl)aminodecyl-sepharose and its interaction with wheat-germ aspartate transcarbamoylase. Biochem. J. 151, 281290.Google Scholar
929Yoshida, T., Mori, T. & Ueda, I. (1983). Giant planar lipid bilayer. I. Capacitance and its biphasic response to inhalation anesthetics. J. Colloid and Interface Sci. 96, 3947.Google Scholar
930Yoza, N., Ogata, T., Ueno, Y. & Ohashi, S. (1971). The application of a thermal detector to the gel chromatography of inorganic compounds. J. Chromat. 61, 295305.Google Scholar
931Yu, C. & Levy, G. C. (1983). Solvent and intramolecular proton dipolar relaxation of the three phosphates of ATP: a heteronuclear 2D NOE study. J. Am. chem. Soc. 105, 69946996.Google Scholar
932Zeitler, H.-J. & Stadler, E. (1972). Determination of parameters in dextran gel filtration. J. chromat. 74, 5971.Google Scholar
933Zillig, W., Stetter, K. O. & Tobien, M. (1978). DNA-dependent RNA polymerase from Halobacterium halobium. Eur. J. Biochem. 91, 193199.Google Scholar
934Zimmerman, S. B. & Pheiffer, B. H. (1983). Macromolecular crowding allows blunt-end ligation by DNA ligases from rat liver or Escherichia coli. Proc. natn. Acad. Sci. U.S.A. 80, 58525856.Google Scholar
935Zwingelstein, G., Tapiero, H., Portoukalian, J. & Fourcade, A. (1982). The effect of dimethylsulfoxide on the lipid composition of inducible and non inducible friend leukemia cells. Biochem. biophys. Res. Comm. 108, 437446.Google Scholar