International
Tables for
Crystallography
Volume F
Crystallography of biological macromolecues
Edited by M. G. Rossmann and E. Arnold

International Tables for Crystallography (2006). Vol. F. ch. 22.3, pp. 553-557   | 1 | 2 |
https://doi.org/10.1107/97809553602060000712

Chapter 22.3. Electrostatic interactions in proteins

K. A. Sharpa*

aE. R. Johnson Research Foundation, Department of Biochemistry and Biophysics, University of Pennsylvania, Philadelphia, PA 19104-6059, USA
Correspondence e-mail: sharp@crystal.med.upenn.edu

References

First citation Antosiewicz, J., McCammon, J. A. & Gilson, M. K. (1994). Prediction of pH-dependent properties of proteins. J. Mol. Biol. 238, 415–436.Google Scholar
First citation Åqvist, J., Luecke, H., Quiocho, F. A. & Warshel, A. (1991). Dipoles localized at helix termini of proteins stabilize charges. Proc. Natl Acad. Sci. USA, 88, 2026–2030.Google Scholar
First citation Bacquet, R. & Rossky, P. (1984). Ionic atmosphere of rodlike polyelectrolytes. A hypernetted chain study. J. Phys. Chem. 88, 2660.Google Scholar
First citation Bashford, D. & Karplus, M. (1990). pKa's of ionizable groups in proteins: atomic detail from a continuum electrostatic model. Biochemistry, 29, 10219–10225.Google Scholar
First citation Beroza, P., Fredkin, D., Okamura, M. & Feher, G. (1991). Protonation of interacting residues in a protein by a Monte-Carlo method. Proc. Natl Acad. Sci. USA, 88, 5804–5808.Google Scholar
First citation Bharadwaj, R., Windemuth, A., Sridharan, S., Honig, B. & Nicholls, A. (1994). The fast multipole boundary-element method for molecular electrostatics – an optimal approach for large systems. J. Comput. Chem. 16, 898–913.Google Scholar
First citation Bruccoleri, R. E., Novotny, J., Sharp, K. A. & Davis, M. E. (1996). Finite difference Poisson–Boltzmann electrostatic calculations: increased accuracy achieved by harmonic dielectric smoothing and charge anti-aliasing. J. Comput. Chem. 18, 268–276.Google Scholar
First citation Davis, M. E. (1994). The inducible multipole solvation model – a new model for solvation effects on solute electrostatics. J. Chem. Phys. 100, 5149–5159.Google Scholar
First citation Davis, M. E. & McCammon, J. A. (1989). Solving the finite difference linear Poisson Boltzmann equation: comparison of relaxational and conjugate gradient methods. J. Comput. Chem. 10, 386–395.Google Scholar
First citation Gilson, M. (1993). Multiple-site titration and molecular modeling: 2. Rapid methods for computing energies and forces for ionizable groups in proteins. Proteins Struct. Funct. Genet. 15, 266–282.Google Scholar
First citation Gilson, M., Davis, M., Luty, B. & McCammon, J. (1993). Computation of electrostatic forces on solvated molecules using the Poisson–Boltzmann equation. J. Phys. Chem. 97, 3591–3600.Google Scholar
First citation Gilson, M. & Honig, B. (1986). The dielectric constant of a folded protein. Biopolymers, 25, 2097–2119.Google Scholar
First citation Gilson, M., McCammon, J. & Madura, J. (1995). Molecular dynamics simulation with continuum solvent. J. Comput. Chem. 16, 1081–1095.Google Scholar
First citation Gilson, M., Sharp, K. A. & Honig, B. (1988). Calculating the electrostatic potential of molecules in solution: method and error assessment. J. Comput. Chem. 9, 327–335.Google Scholar
First citation Holst, M., Kozack, R., Saied, F. & Subramaniam, S. (1994). Protein electrostatics – rapid multigrid-based Newton algorithm for solution of the full nonlinear Poisson–Boltzmann equation. J. Biomol. Struct. Dyn. 11, 1437–1445.Google Scholar
First citation Holst, M. & Saied, F. (1993). Multigrid solution of the Poisson–Boltzmann equation. J. Comput. Chem. 14, 105–113.Google Scholar
First citation Jayaram, B., Fine, R., Sharp, K. A. & Honig, B. (1989). Free energy calculations of ion hydration: an analysis of the Born model in terms of microscopic simulations. J. Phys. Chem. 93, 4320–4327.Google Scholar
First citation Jayaram, B., Sharp, K. A. & Honig, B. (1989). The electrostatic potential of B-DNA. Biopolymers, 28, 975–993.Google Scholar
First citation Jean-Charles, A., Nicholls, A., Sharp, K., Honig, B., Tempczyk, A., Hendrickson, T. & Still, C. (1990). Electrostatic contributions to solvation energies: comparison of free energy perturbation and continuum calculations. J. Am. Chem. Soc. 113, 1454–1455.Google Scholar
First citation Langsetmo, K., Fuchs, J. A., Woodward, C. & Sharp, K. A. (1991). Linkage of thioredoxin stability to titration of ionizable groups with perturbed pKa. Biochemistry, 30, 7609–7614.Google Scholar
First citation Lee, F., Chu, Z. & Warshel, A. (1993). Microscopic and semimicroscopic calculations of electrostatic energies in proteins by the Polaris and Enzymix programs. J. Comput. Chem. 14, 161–185.Google Scholar
First citation Lee, F. S., Chu, Z. T., Bolger, M. B. & Warshel, A. (1992). Calculations of antibody–antigen interactions: microscopic and semi-microscopic evaluation of the free energies of binding of phosphorylcholine analogs to Mcpc603. Protein Eng. 5, 215–228.Google Scholar
First citation Misra, V., Hecht, J., Sharp, K., Friedman, R. & Honig, B. (1994). Salt effects on protein–DNA interactions: the lambda cI repressor and Eco R1 endonuclease. J. Mol. Biol. 238, 264–280.Google Scholar
First citation Misra, V. & Honig, B. (1995). On the magnitude of the electrostatic contribution to ligand–DNA interactions. Proc. Natl Acad. Sci. USA, 92, 4691–4695.Google Scholar
First citation Misra, V., Sharp, K., Friedman, R. & Honig, B. (1994). Salt effects on ligand–DNA binding: minor groove antibiotics. J. Mol. Biol. 238, 245–263.Google Scholar
First citation Mohan, V., Davis, M. E., McCammon, J. A. & Pettitt, B. M. (1992). Continuum model calculations of solvation free energies – accurate evaluation of electrostatic contributions. J. Phys. Chem. 96, 6428–6431.Google Scholar
First citation Murthy, C. S., Bacquet, R. J. & Rossky, P. J. (1985). Ionic distributions near polyelectrolytes. A comparison of theoretical approaches. J. Phys. Chem. 89, 701.Google Scholar
First citation Nakamura, H., Sakamoto, T. & Wada, A. (1988). A theoretical study of the dielectric constant of a protein. Protein Eng. 2, 177–183.Google Scholar
First citation Nicholls, A. & Honig, B. (1991). A rapid finite difference algorithm utilizing successive over-relaxation to solve the Poisson–Boltzmann equation. J. Comput. Chem. 12, 435–445.Google Scholar
First citation Oberoi, H. & Allewell, N. (1993). Multigrid solution of the nonlinear Poisson–Boltzmann equation and calculation of titration curves. Biophys. J. 65, 48–55.Google Scholar
First citation Olmsted, M. C., Anderson, C. F. & Record, M. T. (1989). Monte Carlo description of oligoelectrolyte properties of DNA oligomers. Proc. Natl Acad. Sci. USA, 86, 7766–7770.Google Scholar
First citation Olmsted, M. C., Anderson, C. F. & Record, M. T. (1991). Importance of oligoelectrolyte end effects for the thermodynamics of conformational transitions of nucleic acid oligomers. Biopolymers, 31, 1593–1604.Google Scholar
First citation Pack, G., Garrett, G., Wong, L. & Lamm, G. (1993). The effect of a variable dielectric coefficient and finite ion size on Poisson–Boltzmann calculations of DNA–electrolyte systems. Biophys. J. 65, 1363–1370.Google Scholar
First citation Pack, G. R. & Klein, B. J. (1984). The distribution of electrolyte ions around the B- and Z-conformers of DNA. Biopolymers, 23, 2801.Google Scholar
First citation Pack, G. R., Wong, L. & Prasad, C. V. (1986). Counterion distribution around DNA. Nucleic Acids Res. 14, 1479.Google Scholar
First citation Rashin, A. A. (1990). Hydration phenomena, classical electrostatics and the boundary element method. J. Phys. Chem. 94, 725–733.Google Scholar
First citation Record, T., Olmsted, M. & Anderson, C. (1990). Theoretical studies of the thermodynamics of ion interaction with DNA. In Theoretical biochemistry and molecular biophysics. New York: Adenine Press.Google Scholar
First citation Reiner, E. S. & Radke, C. J. (1990). Variational approach to the electrostatic free energy in charged colloidal suspensions. J. Chem. Soc. Faraday Trans. 86, 3901.Google Scholar
First citation Schaeffer, M. & Frommel, C. (1990). A precise analytical method for calculating the electrostatic energy of macromolecules in aqueous solution. J. Mol. Biol. 216, 1045–1066.Google Scholar
First citation Sharp, K. & Honig, B. (1990). Calculating total electrostatic energies with the non-linear Poisson–Boltzmann equation. J. Phys. Chem. 94, 7684–7692.Google Scholar
First citation Sharp, K. A., Friedman, R., Misra, V., Hecht, J. & Honig, B. (1995). Salt effects on polyelectrolyte–ligand binding: comparison of Poisson–Boltzmann and limiting law counterion binding models. Biopolymers, 36, 245–262.Google Scholar
First citation Simonson, T. & Brooks, C. L. (1996). Charge screening and the dielectric-constant of proteins: insights from molecular-dynamics. J. Am. Chem. Soc. 118, 8452–8458.Google Scholar
First citation Simonson, T. & Brünger, A. (1994). Solvation free energies estimated from a macroscopic continuum theory. J. Phys. Chem. 98, 4683–4694.Google Scholar
First citation Simonson, T. & Perahia, D. (1995). Internal and interfacial dielectric properties of cytochrome c from molecular dynamics in aqueous solution. Proc. Natl Acad. Sci. USA, 92, 1082–1086.Google Scholar
First citation Sitkoff, D., Sharp, K. & Honig, B. (1994). Accurate calculation of hydration free energies using macroscopic solvent models. J. Phys. Chem. 98, 1978–1988.Google Scholar
First citation Slagle, S., Kozack, R. E. & Subramaniam, S. (1994). Role of electrostatics in antibody–antigen association: anti-hen egg lysozyme/lysozyme complex (HyHEL–5/HEL). J. Biomol. Struct. Dyn. 12, 439–456.Google Scholar
First citation Smith, P., Brunne, R., Mark, A. & van Gunsteren, W. (1993). Dielectric properties of trypsin inhibitor and lysozyme calculated from molecular dynamics simulations. J. Phys. Chem. 97, 2009–2014.Google Scholar
First citation Still, C., Tempczyk, A., Hawley, R. & Hendrickson, T. (1990). Semianalytical treatment of solvation for molecular mechanics and dynamics. J. Am. Chem. Soc. 112, 6127–6129.Google Scholar
First citation Takashima, S. & Schwan, H. P. (1965). Dielectric constant measurements on dried proteins. J. Phys. Chem. 69, 4176.Google Scholar
First citation Warshel, A. & Åqvist, J. (1991). Electrostatic energy and macromolecular function. Annu. Rev. Biophys. Biophys. Chem. 20, 267–298.Google Scholar
First citation Warshel, A. & Russell, S. (1984). Calculations of electrostatic interactions in biological systems and in solutions. Q. Rev. Biophys. 17, 283.Google Scholar
First citation Warwicker, J. (1994). Improved continuum electrostatic modelling in proteins, with comparison to experiment. J. Mol. Biol. 236, 887–903.Google Scholar
First citation Warwicker, J. & Watson, H. C. (1982). Calculation of the electric potential in the active site cleft due to alpha-helix dipoles. J. Mol. Biol. 157, 671–679.Google Scholar
First citation Wendoloski, J. J. & Matthew, J. B. (1989). Molecular dynamics effects on protein electrostatics. Proteins, 5, 313.Google Scholar
First citation Yang, A., Gunner, M., Sampogna, R., Sharp, K. & Honig, B. (1993). On the calculation of pKa in proteins. Proteins Struct. Funct. Genet. 15, 252–265.Google Scholar
First citation Yoon, L. & Lenhoff, A. (1992). Computation of the electrostatic interaction energy between a protein and a charged suface. J. Phys. Chem. 96, 3130–3134.Google Scholar
First citation Zauhar, R. & Morgan, R. J. (1985). A new method for computing the macromolecular electric potential. J. Mol. Biol. 186, 815–820.Google Scholar
First citation Zhou, H. X. (1994). Macromolecular electrostatic energy within the nonlinear Poisson–Boltzmann equation. J. Phys. Chem. 100, 3152–3162.Google Scholar