The OPLS [optimized potentials for liquid simulations] potential functions for proteins, energy minimizations for crystals of cyclic peptides and crambin.

A complete set of intermolecular potential functions has been developed for use in computer simulations of proteins in their native environment. Parameters are reported for 25 peptide residues as well as the common neutral and charged terminal groups. The potential functions have the simple Coulomb plus Lennard-Jones form and are compatible with the widely used models for water, TIP4P, TIP3P, and SPC. The parameters were obtained and tested primarily in conjunction with Monte Carlo statistical mechanics simulations of 36 pure organic liquids and numerous aqueous solutions of organic ions representative of subunits in the side chains and backbones of proteins. Bond stretch, angle bend, and torsional terms have been adopted from the AMBER united-atom force field. As reported here, further testing has involved studies of conformational energy surfaces and optimizations of the crystal structures for four cyclic hexapeptides and a cyclic pentapeptide. The average root-mean-square deviation from the X-ray structures of the crystals is only 0.17 A for the atomic positions and 3% for the unit cell volumes. A more critical test was then provided by performing energy minimizations for the complete crystal of the protein crambin, including 182 water molecules that were initially placed via a Monte Carlo simulation. The resultant root-mean-square deviation for the non-hydrogen atoms is still ca. 0.2 A and the variation in the errors for charged, polar, and nonpolar residues is small. Improvement is apparent over the AMBER united-atom force field which has previously been demonstrated to be superior to many alternatives. Computer simulations are undoubtedly destined to became an increasingly important means for investigating the structures and dynamics of biomolecular systems.' At the heart of such theoretical calculations are the force fields that describe the interatomic interactions and the mechanics of deformations of the molecules.* There is also little doubt that there will be a continual evolution in force fields with added complexity and improved performance paralleling the availability of computer resources. Our own efforts in this area over the last few years have resulted in the OPLS potential functions for proteins whose development and performance are summarized here. These potential functions have a simple form and they have been parametrized directly to reproduce experimental thermodynamic and structural data on fluids. Consequently, they are computationally efficient and their description of proteins in solution or crystalline environments should be superior to many alterantives that have been developed with limited condensed-phase data. The latter point is pursued here primarily through calculations on the crystal structures for four cyclic hexapeptides, a cyclic pentapeptide, and the protein crambin. Improvements are apparent in comparison to the AMBER united-atom force field3 which has previously been shown to be superior to many alternative^.^ (1) Beveridge, D. L., Jorgensen, W. L., Eds. Ann. N.Y. Acad. Sci. 1986, 482. ( 2 ) For reviews, see: (a) Levitt, M. Annu. Reu. Biophys. Eioeng. 1982, 11, 251. (b) McCammon, J. A. Rep. Prog. Phys. 1984, 47, 1. (3) Weiner, S. J.; Kollman, P. A.; Case, D. A,; Singh, U. C.; Ghio, C.; Alagona, G.; Profeta, S.; Weiner, P. J. Am. Chem. SOC. 1984, 106, 765. Parametrization The peptide residues of proteins contain readily identifiable organic subunits such as amides, hydrocarbons, alcohols, thioethers, etc. In view of this and since data are available on the corresponding pure organic liquids, our approach to developing a force field for proteins was to build it up from parameters demonstrated to yield good descriptions of organic liquids. U1timately, the force field would need to treat both intramolecular terms for bond stretches, angle bends, and torsions, as well as the intermolecular and intramolecular nonbonded interactions. The latter are generally accepted to be the most difficult part of the problem and have been our focus.3 A simple, computationally efficient form was chosen to represent the nonbonded interactions through Coulomb and Lennard-Jones terms interacting between sites centered on nuclei (eq 1). Thus, the intermolecular inter-

[1]  K. Kuchitsu Electron Diffraction Investigation on the Molecular Structure of n-Butane , 1959 .

[2]  John A. Nelder,et al.  A Simplex Method for Function Minimization , 1965, Comput. J..

[3]  Formation and Thermal Decomposition of Bicyclo[1.1.0]butane-2-exo-d11 , 1966 .

[4]  E. P. Blanchard,et al.  Bicyclo[1.1.0]butane Chemistry. I. The Synthesis and Reactions of 3-Methylbicyclo[1.1.0]butanecarbonitriles , 1966 .

[5]  T. Kitagawa,et al.  Energy Difference between Rotational Isomers of Methyl Ethyl Ether , 1968 .

[6]  P. Pfeffer,et al.  The steric course of the thermal rearrangements of methylbicyclobutanes , 1968 .

[7]  M. Dreyfus,et al.  Molecular orbital calculations on the conformation of polypeptides and proteins. I. Preliminary investigations and simple dipeptides. , 1970, Journal of theoretical biology.

[8]  J. W. Gibson,et al.  The conformation and crystal structure of the cyclic polypeptide -gly-gly-D-ala-D-ala-gly-gly .3H2O. , 1970, Journal of the American Chemical Society.

[9]  Irwin D. Kuntz,et al.  Hydration of macromolecules. III. Hydration of polypeptides , 1971 .

[10]  William F. Murphy,et al.  Rotational isomerism. XI. Raman spectra of n-butane, 2-methylbutane, and 2, 3-dimethylbutane , 1974 .

[11]  Acetylglycine-N-methylamide , 1974 .

[12]  MINDO/3 study of thermolysis of bicyclobutane. Allowed and stereoselective reaction that is not concerted , 1975 .

[13]  Application of the principle of least motion to organic reactions. 4. More complex molecular rearrangements , 1976 .

[14]  J. Mulder Orbital symmetry and the benzene .dblarw. benzvalene interconversion , 1977 .

[15]  Norman L. Allinger,et al.  Conformational analysis. 130. MM2. A hydrocarbon force field utilizing V1 and V2 torsional terms , 1977 .

[16]  Harold A. Scheraga,et al.  Energy parameters in polypeptides. 8. Empirical potential energy algorithm for the conformational analysis of large molecules , 1978 .

[17]  I. Karle,et al.  Crystal structure and conformation of cyclo-(glycylprolylglycyl-D-alanylprolyl) containing 4 .fwdarw. 1 and 3 .fwdarw. 1 intramolecular hydrogen bonds , 1978 .

[18]  J. Aihara Aromaticity-Based Theory of Pericyclic Reactions , 1978 .

[19]  M. Hossain,et al.  Conformation and crystal structures of two cycloisomeric hexapeptides: cyclo-(L-alanyl-L-alanylglycylglycyl-L-alanylglycyl) monohydrate (I) and cyclo-(L-alanyl-L-alanylglycyl-L-alanylglycylglycyl) dihydrate (II) , 1978 .

[20]  Kazuko Oyanagi,et al.  Molecular structure and conformation of ethyl methyl sulfide as studied by gas electron diffraction. , 1978 .

[21]  W. E. Thiessen,et al.  Crystal structure and molecular conformation of the cyclic hexapeptide cyclo-(Gly-L-Pro-Gly)2 , 1979 .

[22]  U. Burkert Ab initio calculations of the rotational potential functions for propanol and ethyl methyl ether , 1980 .

[23]  William F. Murphy,et al.  Low-frequency Raman spectrum and asymmetric potential function for internal rotation of gaseous n-butane , 1980 .

[24]  Wayne A. Hendrickson,et al.  Structure of the hydrophobic protein crambin determined directly from the anomalous scattering of sulphur , 1981, Nature.

[25]  W. L. Jorgensen Quantum and statistical mechanical studies of liquids. 10. Transferable intermolecular potential functions for water, alcohols, and ethers. Application to liquid water , 2002 .

[26]  H. Berendsen,et al.  Interaction Models for Water in Relation to Protein Hydration , 1981 .

[27]  Structures and properties of organic liquids: n-alkyl ethers and their conformational equilibriums , 1981 .

[28]  William L. Jorgensen,et al.  Monte Carlo simulation of n‐butane in water. Conformational evidence for the hydrophobic effect , 1982 .

[29]  L. Schäfer,et al.  Ab initio studies of structural features not easily amenable to experiment. 23. Molecular structures and conformational analysis of the dipeptide N‐acetyl‐N′‐methyl glycyl amide and the significance of local geometries for peptide structures , 1982 .

[30]  H. Scheraga,et al.  Computed conformational states of the 20 naturally occurring amino acid residues and of the prototype residue α-aminobutyric acid , 1983 .

[31]  W. L. Jorgensen,et al.  Comparison of simple potential functions for simulating liquid water , 1983 .

[32]  F. Momany,et al.  Ab initio studies of molecular geometries. 27. Optimized molecular structures and conformational analysis of N.alpha.-acetyl-N-methylalaninamide and comparison with peptide crystal data and empirical calculations , 1983 .

[33]  U. Singh,et al.  A NEW FORCE FIELD FOR MOLECULAR MECHANICAL SIMULATION OF NUCLEIC ACIDS AND PROTEINS , 1984 .

[34]  William L. Jorgensen,et al.  Optimized intermolecular potential functions for liquid hydrocarbons , 1984 .

[35]  M. Teeter,et al.  Water structure of a hydrophobic protein at atomic resolution: Pentagon rings of water molecules in crystals of crambin. , 1981, Proceedings of the National Academy of Sciences of the United States of America.

[36]  David Hall,et al.  An appraisal of molecular force fields for the representation of polypeptides , 1984 .

[37]  F. Momany,et al.  Conformational transitions and geometry differences between low‐energy conformers of N‐acetyl‐N′‐methyl alanineamide: An ab initio study at the 4‐21G level with gradient relaxed geometries , 1984 .

[38]  M. Dewar Multibond reactions cannot normally be synchronous , 1984 .

[39]  Krishnan Raghavachari,et al.  Rotational potential surface for alkanes: Basis set and electron correlation effects on the conformations of n‐butane , 1984 .

[40]  William L. Jorgensen,et al.  Monte Carlo simulations of alkanes in water: hydration numbers and the hydrophobic effect , 1985 .

[41]  William L. Jorgensen,et al.  Temperature and size dependence for Monte Carlo simulations of TIP4P water , 1985 .

[42]  William L. Jorgensen,et al.  Additions and Corrections - Optimized Intermolecular Potential Functions for Amides and Peptides. Hydration of Amides. , 1985 .

[43]  William L. Jorgensen,et al.  Optimized intermolecular potential functions for amides and peptides. Structure and properties of liquid amides , 1985 .

[44]  F. Momany,et al.  Local geometry maps and conformational transitions between low-energy conformers of N-acetyl-N′-methyl glycine amide: An ab initio study at the 4–21g level with gradient relaxed geometries , 1985 .

[45]  W. L. Jorgensen,et al.  Monte Carlo simulation of differences in free energies of hydration , 1985 .

[46]  M. Whitlow,et al.  An empirical examination of potential energy minimization using the well-determined structure of the protein crambin , 1986 .

[47]  William L. Jorgensen,et al.  Intermolecular potential functions and Monte Carlo simulations for liquid sulfur compounds , 1986 .

[48]  K. B. Wiberg The Concept of Strain in Organic Chemistry , 1986 .

[49]  William L. Jorgensen,et al.  Optimized intermolecular potential functions for liquid alcohols , 1986 .

[50]  R. G. Snyder,et al.  Experimental determination of the trans–gauche energy difference of gaseous n‐pentane and diethyl ether , 1986 .

[51]  Ab initio study of structures and binding energies for anion-water complexes , 1986 .

[52]  William L. Jorgensen,et al.  Monte Carlo simulations of the hydration of ammonium and carboxylate ions , 1986 .

[53]  P. Kollman,et al.  An all atom force field for simulations of proteins and nucleic acids , 1986, Journal of computational chemistry.

[54]  L. Berliner,et al.  1H NMR characterization of two crambin species , 1987 .