A novel method reveals that solvent water favors polyproline II over β‐strand conformation in peptides and unfolded proteins: conditional hydrophobic accessible surface area (CHASA)

In aqueous solution, the ensemble of conformations sampled by peptides and unfolded proteins is largely determined by their interaction with water. It has been a long‐standing goal to capture these solute‐water energetics accurately and efficiently in calculations. Historically, accessible surface area (ASA) has been used to estimate these energies, but this method breaks down when applied to amphipathic peptides and proteins. Here we introduce a novel method in which hydrophobic ASA is determined after first positioning water oxygens in hydrogen‐bonded orientations proximate to all accessible peptide/protein backbone N and O atoms. This conditional hydrophobic accessible surface area is termed CHASA. The CHASA method was validated by predicting the polyproline‐II (PII) and β‐strand conformational preferences of non‐proline residues in the coil library (i.e., non‐α‐helix, non‐β‐strand, non‐β‐turn library derived from X‐ray elucidated structures). Further, the method successfully rationalizes the previously unexplained solvation energies in polyalanyl peptides and compares favorably with published experimentally determined PII residue propensities.

[1]  Amedeo Caflisch,et al.  Calculation of conformational transitions and barriers in solvated systems: Application to the alanine dipeptide in water , 1999 .

[2]  K A Dill,et al.  Solvation: how to obtain microscopic energies from partitioning and solvation experiments. , 1997, Annual review of biophysics and biomolecular structure.

[3]  P. Y. Chou,et al.  Prediction of the secondary structure of proteins from their amino acid sequence. , 2006 .

[4]  B. Lee,et al.  The interpretation of protein structures: estimation of static accessibility. , 1971, Journal of molecular biology.

[5]  H. Scheraga,et al.  Accessible surface areas as a measure of the thermodynamic parameters of hydration of peptides. , 1987, Proceedings of the National Academy of Sciences of the United States of America.

[6]  B. Montgomery Pettitt,et al.  Conformational free energy of hydration for the alanine dipeptide: Thermodynamic analysis , 1988 .

[7]  R. Srinivasan,et al.  A physical basis for protein secondary structure. , 1999, Proceedings of the National Academy of Sciences of the United States of America.

[8]  M. Swindells,et al.  Intrinsic phi, psi propensities of amino acids, derived from the coil regions of known structures. , 1995, Nature structural biology.

[9]  J. Richardson,et al.  The penultimate rotamer library , 2000, Proteins.

[10]  G. N. Ramachandran,et al.  Stereochemistry of polypeptide chain configurations. , 1963, Journal of molecular biology.

[11]  J. Apostolakis,et al.  Comparison of a GB Solvation Model with Explicit Solvent Simulations: Potentials of Mean Force and Conformational Preferences of Alanine Dipeptide and 1,2-Dichloroethane , 1998 .

[12]  T. N. Bhat,et al.  The Protein Data Bank , 2000, Nucleic Acids Res..

[13]  K. Griebenow,et al.  Preferred peptide backbone conformations in the unfolded state revealed by the structure analysis of alanine-based (AXA) tripeptides in aqueous solution. , 2004, Proceedings of the National Academy of Sciences of the United States of America.

[14]  Paul E. Smith,et al.  THE ALANINE DIPEPTIDE FREE ENERGY SURFACE IN SOLUTION , 1999 .

[15]  Charles L. Brooks,et al.  Simulations of peptide conformational dynamics and thermodynamics , 1993 .

[16]  G. Rose,et al.  Turns in peptides and proteins. , 1985, Advances in protein chemistry.

[17]  Olga Kennard,et al.  Hydrogen-bond geometry in organic crystals , 1984 .

[18]  T. Creamer,et al.  Solvation energies of amino acid side chains and backbone in a family of host-guest pentapeptides. , 1996, Biochemistry.

[19]  Angel E. Garcia,et al.  Characterization of non-alpha helical conformations in Ala peptides , 2004 .

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

[21]  R. S. Spolar,et al.  Use of liquid hydrocarbon and amide transfer data to estimate contributions to thermodynamic functions of protein folding from the removal of nonpolar and polar surface from water. , 1992, Biochemistry.

[22]  G. Rose,et al.  A simple model for polyproline II structure in unfolded states of alanine‐based peptides , 2002, Protein science : a publication of the Protein Society.

[23]  N. Metropolis,et al.  Equation of State Calculations by Fast Computing Machines , 1953, Resonance.

[24]  K. P. Murphy,et al.  Structural energetics of peptide recognition: Angiotensin II/antibody binding , 1993, Proteins.

[25]  Mihaly Mezei,et al.  Unfolded state of polyalanine is a segmented polyproline II helix , 2004, Proteins.

[26]  George I. Makhatadze,et al.  Hydration of the peptide backbone largely defines the thermodynamic propensity scale of residues at the C′ position of the C-capping box of α-helices , 2001, Proceedings of the National Academy of Sciences of the United States of America.

[27]  Patrick J Fleming,et al.  Ab Initio Protein Folding Using LINUS , 2004, Numerical Computer Methods, Part D.

[28]  Christian Bartels,et al.  Multidimensional adaptive umbrella sampling: Applications to main chain and side chain peptide conformations , 1997, J. Comput. Chem..

[29]  V. Hilser,et al.  The enthalpy change in protein folding and binding: Refinement of parameters for structure‐based calculations , 1996, Proteins.

[30]  R J Read,et al.  Crystallography & NMR system: A new software suite for macromolecular structure determination. , 1998, Acta crystallographica. Section D, Biological crystallography.

[31]  R. L. Baldwin,et al.  Energetics of the interaction between water and the helical peptide group and its role in determining helix propensities. , 2000, Proceedings of the National Academy of Sciences of the United States of America.

[32]  R. Srinivasan,et al.  LINUS: A hierarchic procedure to predict the fold of a protein , 1995, Proteins.

[33]  M. Lewis,et al.  Calculation of the free energy of association for protein complexes , 1992, Protein science : a publication of the Protein Society.

[34]  Lorna J. Smith,et al.  Toward a Description of the Conformations of Denatured States of Proteins. Comparison of a Random Coil Model with NMR Measurements , 1996 .

[35]  J M Thornton,et al.  Analysis of main chain torsion angles in proteins: prediction of NMR coupling constants for native and random coil conformations. , 1996, Journal of molecular biology.

[36]  P. Privalov,et al.  Contribution of hydration to protein folding thermodynamics. II. The entropy and Gibbs energy of hydration. , 1993, Journal of molecular biology.

[37]  J. Andrew McCammon,et al.  Conformational sampling with Poisson–Boltzmann forces and a stochastic dynamics/Monte Carlo method: Application to alanine dipeptide , 1997 .

[38]  J. Apostolakis,et al.  Evaluation of a fast implicit solvent model for molecular dynamics simulations , 2002, Proteins.

[39]  Robert B. Hermann,et al.  Theory of hydrophobic bonding. II. Correlation of hydrocarbon solubility in water with solvent cavity surface area , 1972 .

[40]  Guoli Wang,et al.  PISCES: a protein sequence culling server , 2003, Bioinform..

[41]  J. D. de Pablo,et al.  Configurational temperature density of states simulations of proteins. , 2003, Biophysical journal.

[42]  T. Creamer,et al.  Short sequences of non-proline residues can adopt the polyproline II helical conformation. , 2004, Biochemistry.

[43]  Alan Grossfield,et al.  Role of solvent in determining conformational preferences of alanine dipeptide in water. , 2004, Journal of the American Chemical Society.

[44]  R. Levy,et al.  Enthalpy−Entropy and Cavity Decomposition of Alkane Hydration Free Energies: Numerical Results and Implications for Theories of Hydrophobic Solvation , 2000 .

[45]  A. Doig,et al.  Rotamer strain energy in protein helices - quantification of a major force opposing protein folding. , 2001, Journal of molecular biology.

[46]  M. Rao,et al.  On the force bias Monte Carlo simulation of water: methodology, optimization and comparison with molecular dynamics , 1979 .

[47]  Charles L. Brooks,et al.  Conformational equilibrium in the alanine dipeptide in the gas phase and aqueous solution : a comparison of theoretical results , 1992 .

[48]  A. Shrake,et al.  Environment and exposure to solvent of protein atoms. Lysozyme and insulin. , 1973, Journal of molecular biology.

[49]  P. Privalov,et al.  Contribution of hydration to protein folding thermodynamics. I. The enthalpy of hydration. , 1993, Journal of molecular biology.

[50]  Mihaly Mezei,et al.  Polyproline II helix is the preferred conformation for unfolded polyalanine in water , 2004, Proteins.

[51]  Alexander D. MacKerell,et al.  All-atom empirical potential for molecular modeling and dynamics studies of proteins. , 1998, The journal of physical chemistry. B.

[52]  J Hermans,et al.  Microfolding: Conformational probability map for the alanine dipeptide in water from molecular dynamics simulations , 1988, Proteins.

[53]  Kentaro Shimizu,et al.  Molecular Dynamics Simulations in Aqueous Solution: Application to Free Energy Calculation of Oligopeptides , 1998 .

[54]  J. Hermans,et al.  Comparison of a QM/MM force field and molecular mechanics force fields in simulations of alanine and glycine “dipeptides” (Ace‐Ala‐Nme and Ace‐Gly‐Nme) in water in relation to the problem of modeling the unfolded peptide backbone in solution , 2003, Proteins.

[55]  Distance-scaled Force Biased Monte Carlo Simulation for Solutions containing a Strongly Interacting Solute , 1991 .

[56]  W. Delano The PyMOL Molecular Graphics System , 2002 .

[57]  L. Serrano,et al.  H‐bonding in protein hydration revisited , 2004, Protein science : a publication of the Protein Society.

[58]  K. P. Murphy,et al.  Prediction of binding energetics from structure using empirical parameterization. , 1998, Methods in enzymology.

[59]  G. Rose,et al.  Side-chain entropy opposes alpha-helix formation but rationalizes experimentally determined helix-forming propensities. , 1992, Proceedings of the National Academy of Sciences of the United States of America.

[60]  Interaction between water and polar groups of the helix backbone: an important determinant of helix propensities. , 1999, Proceedings of the National Academy of Sciences of the United States of America.

[61]  Vincent J Hilser,et al.  Thermodynamics of binding to SH3 domains: the energetic impact of polyproline II (PII) helix formation. , 2004, Biochemistry.

[62]  M Mezei,et al.  The sensitivity of conformational free energies of the alanine dipeptide to atomic site charges. , 1997, Biopolymers.

[63]  Thermodynamic mechanism and consequences of the polyproline II (PII) structural bias in the denatured states of proteins. , 2004, Biochemistry.

[64]  R. L. Baldwin,et al.  Role of backbone solvation and electrostatics in generating preferred peptide backbone conformations: Distributions of phi , 2003, Proceedings of the National Academy of Sciences of the United States of America.

[65]  Greg L. Hura,et al.  Hydration dynamics near a model protein surface. , 2004, Biophysical journal.

[66]  M. Swindells,et al.  Intrinsic φ,ψ propensities of amino acids, derived from the coil regions of known structures , 1995, Nature Structural Biology.

[67]  A. Miele,et al.  Free energy of burying hydrophobic residues in the interface between protein subunits. , 1998, Proceedings of the National Academy of Sciences of the United States of America.

[68]  Sándor Suhai,et al.  Theoretical Study of Aqueous N-Acetyl-l-alanine N‘-Methylamide: Structures and Raman, VCD, and ROA Spectra , 1998 .

[69]  Robert Bell Hermann,et al.  Theory of Hydrophobic Bonding , 1974 .