Skip to main content
Protein Science : A Publication of the Protein Society logoLink to Protein Science : A Publication of the Protein Society
. 1997 Jun;6(6):1293–1301. doi: 10.1002/pro.5560060617

On the calculation of binding free energies using continuum methods: application to MHC class I protein-peptide interactions.

N Froloff 1, A Windemuth 1, B Honig 1
PMCID: PMC2143728  PMID: 9194189

Abstract

This paper describes a methodology to calculate the binding free energy (delta G) of a protein-ligand complex using a continuum model of the solvent. A formal thermodynamic cycle is used to decompose the binding free energy into electrostatic and non-electrostatic contributions. In this cycle, the reactants are discharged in water, associated as purely nonpolar entities, and the final complex is then recharged. The total electrostatic free energies of the protein, the ligand, and the complex in water are calculated with the finite difference Poisson-Boltzmann (FDPB) method. The nonpolar (hydrophobic) binding free energy is calculated using a free energy-surface area relationship, with a single alkane/water surface tension coefficient (gamma aw). The loss in backbone and side-chain configurational entropy upon binding is estimated and added to the electrostatic and the nonpolar components of delta G. The methodology is applied to the binding of the murine MHC class I protein H-2Kb with three distinct peptides, and to the human MHC class I protein HLA-A2 in complex with five different peptides. Despite significant differences in the amino acid sequences of the different peptides, the experimental binding free energy differences (delta delta Gexp) are quite small (< 0.3 and < 2.7 kcal/mol for the H-2Kb and HLA-A2 complexes, respectively). For each protein, the calculations are successful in reproducing a fairly small range of values for delta delta Gcalc (< 4.4 and < 5.2 kcal/mol, respectively) although the relative peptide binding affinities of H-2Kb and HLA-A2 are not reproduced. For all protein-peptide complexes that were treated, it was found that electrostatic interactions oppose binding whereas nonpolar interactions drive complex formation. The two types of interactions appear to be correlated in that larger nonpolar contributions to binding are generally opposed by increased electrostatic contributions favoring dissociation. The factors that drive the binding of peptides to MHC proteins are discussed in light of our results.

Full Text

The Full Text of this article is available as a PDF (3.5 MB).

Selected References

These references are in PubMed. This may not be the complete list of references from this article.

  1. Ajay, Murcko M. A. Computational methods to predict binding free energy in ligand-receptor complexes. J Med Chem. 1995 Dec 22;38(26):4953–4967. doi: 10.1021/jm00026a001. [DOI] [PubMed] [Google Scholar]
  2. Altuvia Y., Schueler O., Margalit H. Ranking potential binding peptides to MHC molecules by a computational threading approach. J Mol Biol. 1995 Jun 2;249(2):244–250. doi: 10.1006/jmbi.1995.0293. [DOI] [PubMed] [Google Scholar]
  3. Andrews P. R., Craik D. J., Martin J. L. Functional group contributions to drug-receptor interactions. J Med Chem. 1984 Dec;27(12):1648–1657. doi: 10.1021/jm00378a021. [DOI] [PubMed] [Google Scholar]
  4. Bash P. A., Singh U. C., Brown F. K., Langridge R., Kollman P. A. Calculation of the relative change in binding free energy of a protein-inhibitor complex. Science. 1987 Jan 30;235(4788):574–576. doi: 10.1126/science.3810157. [DOI] [PubMed] [Google Scholar]
  5. Bernstein F. C., Koetzle T. F., Williams G. J., Meyer E. F., Jr, Brice M. D., Rodgers J. R., Kennard O., Shimanouchi T., Tasumi M. The Protein Data Bank: a computer-based archival file for macromolecular structures. J Mol Biol. 1977 May 25;112(3):535–542. doi: 10.1016/s0022-2836(77)80200-3. [DOI] [PubMed] [Google Scholar]
  6. Beveridge D. L., DiCapua F. M. Free energy via molecular simulation: applications to chemical and biomolecular systems. Annu Rev Biophys Biophys Chem. 1989;18:431–492. doi: 10.1146/annurev.bb.18.060189.002243. [DOI] [PubMed] [Google Scholar]
  7. Bluestone J. A., Jameson S., Miller S., Dick R., 2nd Peptide-induced conformational changes in class I heavy chains alter major histocompatibility complex recognition. J Exp Med. 1992 Dec 1;176(6):1757–1761. doi: 10.1084/jem.176.6.1757. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Brünger A. T., Karplus M. Polar hydrogen positions in proteins: empirical energy placement and neutron diffraction comparison. Proteins. 1988;4(2):148–156. doi: 10.1002/prot.340040208. [DOI] [PubMed] [Google Scholar]
  9. Böhm H. J. The development of a simple empirical scoring function to estimate the binding constant for a protein-ligand complex of known three-dimensional structure. J Comput Aided Mol Des. 1994 Jun;8(3):243–256. doi: 10.1007/BF00126743. [DOI] [PubMed] [Google Scholar]
  10. Catipović B., Dal Porto J., Mage M., Johansen T. E., Schneck J. P. Major histocompatibility complex conformational epitopes are peptide specific. J Exp Med. 1992 Dec 1;176(6):1611–1618. doi: 10.1084/jem.176.6.1611. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Catipović B., Talluri G., Oh J., Wei T., Su X. M., Johansen T. E., Edidin M., Schneck J. P. Analysis of the structure of empty and peptide-loaded major histocompatibility complex molecules at the cell surface. J Exp Med. 1994 Nov 1;180(5):1753–1761. doi: 10.1084/jem.180.5.1753. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Doig A. J., Sternberg M. J. Side-chain conformational entropy in protein folding. Protein Sci. 1995 Nov;4(11):2247–2251. doi: 10.1002/pro.5560041101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Elliott T., Cerundolo V., Elvin J., Townsend A. Peptide-induced conformational change of the class I heavy chain. Nature. 1991 May 30;351(6325):402–406. doi: 10.1038/351402a0. [DOI] [PubMed] [Google Scholar]
  14. Erickson H. P. Co-operativity in protein-protein association. The structure and stability of the actin filament. J Mol Biol. 1989 Apr 5;206(3):465–474. doi: 10.1016/0022-2836(89)90494-4. [DOI] [PubMed] [Google Scholar]
  15. Fremont D. H., Stura E. A., Matsumura M., Peterson P. A., Wilson I. A. Crystal structure of an H-2Kb-ovalbumin peptide complex reveals the interplay of primary and secondary anchor positions in the major histocompatibility complex binding groove. Proc Natl Acad Sci U S A. 1995 Mar 28;92(7):2479–2483. doi: 10.1073/pnas.92.7.2479. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Friedman R. A., Honig B. A free energy analysis of nucleic acid base stacking in aqueous solution. Biophys J. 1995 Oct;69(4):1528–1535. doi: 10.1016/S0006-3495(95)80023-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Gilson M. K., Given J. A., Bush B. L., McCammon J. A. The statistical-thermodynamic basis for computation of binding affinities: a critical review. Biophys J. 1997 Mar;72(3):1047–1069. doi: 10.1016/S0006-3495(97)78756-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Gilson M. K., Honig B. Calculation of the total electrostatic energy of a macromolecular system: solvation energies, binding energies, and conformational analysis. Proteins. 1988;4(1):7–18. doi: 10.1002/prot.340040104. [DOI] [PubMed] [Google Scholar]
  19. Ippolito J. A., Alexander R. S., Christianson D. W. Hydrogen bond stereochemistry in protein structure and function. J Mol Biol. 1990 Oct 5;215(3):457–471. doi: 10.1016/s0022-2836(05)80364-x. [DOI] [PubMed] [Google Scholar]
  20. Jackson R. M., Sternberg M. J. A continuum model for protein-protein interactions: application to the docking problem. J Mol Biol. 1995 Jul 7;250(2):258–275. doi: 10.1006/jmbi.1995.0375. [DOI] [PubMed] [Google Scholar]
  21. Lee B., Richards F. M. The interpretation of protein structures: estimation of static accessibility. J Mol Biol. 1971 Feb 14;55(3):379–400. doi: 10.1016/0022-2836(71)90324-x. [DOI] [PubMed] [Google Scholar]
  22. Matsumura M., Saito Y., Jackson M. R., Song E. S., Peterson P. A. In vitro peptide binding to soluble empty class I major histocompatibility complex molecules isolated from transfected Drosophila melanogaster cells. J Biol Chem. 1992 Nov 25;267(33):23589–23595. [PubMed] [Google Scholar]
  23. McCammon J. A. Computer-aided molecular design. Science. 1987 Oct 23;238(4826):486–491. doi: 10.1126/science.3310236. [DOI] [PubMed] [Google Scholar]
  24. Miyamoto S., Kollman P. A. Absolute and relative binding free energy calculations of the interaction of biotin and its analogs with streptavidin using molecular dynamics/free energy perturbation approaches. Proteins. 1993 Jul;16(3):226–245. doi: 10.1002/prot.340160303. [DOI] [PubMed] [Google Scholar]
  25. Myers J. K., Pace C. N. Hydrogen bonding stabilizes globular proteins. Biophys J. 1996 Oct;71(4):2033–2039. doi: 10.1016/S0006-3495(96)79401-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Nicholls A., Sharp K. A., Honig B. Protein folding and association: insights from the interfacial and thermodynamic properties of hydrocarbons. Proteins. 1991;11(4):281–296. doi: 10.1002/prot.340110407. [DOI] [PubMed] [Google Scholar]
  27. Novotny J., Bruccoleri R. E., Saul F. A. On the attribution of binding energy in antigen-antibody complexes McPC 603, D1.3, and HyHEL-5. Biochemistry. 1989 May 30;28(11):4735–4749. doi: 10.1021/bi00437a034. [DOI] [PubMed] [Google Scholar]
  28. Page M. I., Jencks W. P. Entropic contributions to rate accelerations in enzymic and intramolecular reactions and the chelate effect. Proc Natl Acad Sci U S A. 1971 Aug;68(8):1678–1683. doi: 10.1073/pnas.68.8.1678. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Pickett S. D., Sternberg M. J. Empirical scale of side-chain conformational entropy in protein folding. J Mol Biol. 1993 Jun 5;231(3):825–839. doi: 10.1006/jmbi.1993.1329. [DOI] [PubMed] [Google Scholar]
  30. Richards F. M. Areas, volumes, packing and protein structure. Annu Rev Biophys Bioeng. 1977;6:151–176. doi: 10.1146/annurev.bb.06.060177.001055. [DOI] [PubMed] [Google Scholar]
  31. Rohren E. M., McCormick D. J., Pease L. R. Peptide-induced conformational changes in class I molecules. Direct detection by flow cytometry. J Immunol. 1994 Jun 1;152(11):5337–5343. [PubMed] [Google Scholar]
  32. Ruppert J., Sidney J., Celis E., Kubo R. T., Grey H. M., Sette A. Prominent role of secondary anchor residues in peptide binding to HLA-A2.1 molecules. Cell. 1993 Sep 10;74(5):929–937. doi: 10.1016/0092-8674(93)90472-3. [DOI] [PubMed] [Google Scholar]
  33. STEINBERG I. Z., SCHERAGA H. A. Entropy changes accompanying association reactions of proteins. J Biol Chem. 1963 Jan;238:172–181. [PubMed] [Google Scholar]
  34. Sharp K. A., Honig B. Electrostatic interactions in macromolecules: theory and applications. Annu Rev Biophys Biophys Chem. 1990;19:301–332. doi: 10.1146/annurev.bb.19.060190.001505. [DOI] [PubMed] [Google Scholar]
  35. Sharp K. A., Nicholls A., Fine R. F., Honig B. Reconciling the magnitude of the microscopic and macroscopic hydrophobic effects. Science. 1991 Apr 5;252(5002):106–109. doi: 10.1126/science.2011744. [DOI] [PubMed] [Google Scholar]
  36. Shen J., Wendoloski J. Binding of phosphorus-containing inhibitors to thermolysin studied by the Poisson-Boltzmann method. Protein Sci. 1995 Mar;4(3):373–381. doi: 10.1002/pro.5560040303. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Sherman L. A., Chattopadhyay S., Biggs J. A., Dick R. F., 2nd, Bluestone J. A. Alloantibodies can discriminate class I major histocompatibility complex molecules associated with various endogenous peptides. Proc Natl Acad Sci U S A. 1993 Aug 1;90(15):6949–6951. doi: 10.1073/pnas.90.15.6949. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Smith K. C., Honig B. Evaluation of the conformational free energies of loops in proteins. Proteins. 1994 Feb;18(2):119–132. doi: 10.1002/prot.340180205. [DOI] [PubMed] [Google Scholar]
  39. Stanfield R. L., Wilson I. A. Protein-peptide interactions. Curr Opin Struct Biol. 1995 Feb;5(1):103–113. doi: 10.1016/0959-440x(95)80015-s. [DOI] [PubMed] [Google Scholar]
  40. Straatsma T. P., McCammon J. A. Theoretical calculations of relative affinities of binding. Methods Enzymol. 1991;202:497–511. doi: 10.1016/0076-6879(91)02025-5. [DOI] [PubMed] [Google Scholar]
  41. Tidor B., Karplus M. The contribution of vibrational entropy to molecular association. The dimerization of insulin. J Mol Biol. 1994 May 6;238(3):405–414. doi: 10.1006/jmbi.1994.1300. [DOI] [PubMed] [Google Scholar]
  42. Vajda S., Weng Z., Rosenfeld R., DeLisi C. Effect of conformational flexibility and solvation on receptor-ligand binding free energies. Biochemistry. 1994 Nov 29;33(47):13977–13988. doi: 10.1021/bi00251a004. [DOI] [PubMed] [Google Scholar]
  43. Warwicker J., Watson H. C. Calculation of the electric potential in the active site cleft due to alpha-helix dipoles. J Mol Biol. 1982 Jun 5;157(4):671–679. doi: 10.1016/0022-2836(82)90505-8. [DOI] [PubMed] [Google Scholar]
  44. Wendoloski J. J., Shen J., Oliva M. T., Weber P. C. Biophysical tools for structure-based drug design. Pharmacol Ther. 1993 Nov;60(2):169–183. doi: 10.1016/0163-7258(93)90005-x. [DOI] [PubMed] [Google Scholar]
  45. Williams D. H., Searle M. S., Mackay J. P., Gerhard U., Maplestone R. A. Toward an estimation of binding constants in aqueous solution: studies of associations of vancomycin group antibiotics. Proc Natl Acad Sci U S A. 1993 Feb 15;90(4):1172–1178. doi: 10.1073/pnas.90.4.1172. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Yang A. S., Honig B. Free energy determinants of secondary structure formation: I. alpha-Helices. J Mol Biol. 1995 Sep 22;252(3):351–365. doi: 10.1006/jmbi.1995.0502. [DOI] [PubMed] [Google Scholar]
  47. Zhang T., Koshland D. E., Jr Computational method for relative binding energies of enzyme-substrate complexes. Protein Sci. 1996 Feb;5(2):348–356. doi: 10.1002/pro.5560050219. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Protein Science : A Publication of the Protein Society are provided here courtesy of The Protein Society

RESOURCES