Coordination number of zinc ions in the phosphotriesterase active site by molecular dynamics and quantum mechanics

We have run several molecular dynamics (MD) simulations on zinc‐containing phosphotriesterase (PTE) with two bound substrates, sarin and paraoxon, and with the substrate analog diethyl 4‐methylbenzylphosphonate. A standard nonbonded model was employed to treat the zinc ions with the commonly used charge of +2. In all the trajectories, we observed a tightly bound water (TBW) molecule in the active site that was coordinated to the less buried zinc ion. The phosphoryl oxygen of the substrate/inhibitor was found to be coordinated to the same zinc ion so that, considering all ligands, the less buried zinc was hexa‐coordinated. The hexa‐coordination of this zinc ion was not seen in the deposited X‐ray pdb files for PTE. Several additional MD simulations were then performed using different charges (+1, +1.5) on the zinc ions, along with ab initio and density functional theory (DFT) calculations, to evaluate the following possibilities: the crystal diffraction data were not correctly interpreted; the hexa‐coordinated zinc ion in PTE is only present in solution and not in the crystal; and the hexa‐coordinated zinc ion in PTE is an artifact of the force field used. A charge of +1.5 leads to a coordination number (CN) of 5 on both zinc ions, which is consistent with the results from ab initio and DFT calculations and with the latest high resolution X‐ray crystal structure. The commonly used charge of +2 produces a CN of 6 on the less buried zinc. The CN on the more buried zinc ion is 5 when the substrate/inhibitor is present in the simulation, and increases to 6 when the substrate/inhibitor is removed prior to the simulation. The results of both of the MD and quantum mechanical calculations lead to the conclusion that the zinc ions in the PTE active site are both penta‐coordinated, and that the MD simulations performed with the charge of +2 overestimate the CN of the zinc ions in the PTE active site. The overall protein structures in the simulations remain unaffected by the change in zinc charge from +2 to +1.5. The results also suggest that the charge +1.5 is the most appropriate for the molecular dynamics simulations on zinc‐containing PTE when a nonbonded model is used and no global thermodynamic conclusion is sought. We also show that the standard nonbonded model is not able to properly treat the CN and energy at the same time. A preliminary, promising charge‐transfer model is discussed with the use of the zinc charge of +1.5. © 2003 Wiley Periodicals, Inc. J Comput Chem 24: 368–378, 2003

[1]  T. Darden,et al.  Particle mesh Ewald: An N⋅log(N) method for Ewald sums in large systems , 1993 .

[2]  Georg Johansson,et al.  Structures of Complexes in Solution Derived from X-Ray Diffraction Measurements , 1994 .

[3]  Kenneth M. Merz,et al.  Solvent Dynamics and Mechanism of Proton Transfer in Human Carbonic Anhydrase II , 1999 .

[4]  Toshio Yamaguchi,et al.  X-Ray Diffraction Studies of the Structures of Hydrated Divalent Transition-Metal Ions in Aqueous Solution , 1976 .

[5]  D. Christianson,et al.  Structural biology of zinc. , 1991, Advances in protein chemistry.

[6]  F. Raushel,et al.  Three-dimensional structure of the zinc-containing phosphotriesterase with the bound substrate analog diethyl 4-methylbenzylphosphonate. , 1996, Biochemistry.

[7]  R. Ornstein,et al.  Determination of Two Structural Forms of Catalytic Bridging Ligand in Zinc−Phosphotriesterase by Molecular Dynamics Simulation and Quantum Chemical Calculation , 1999 .

[8]  Charles W. Bock,et al.  Hydration of Zinc Ions: A Comparison with Magnesium and Beryllium Ions , 1995 .

[9]  A. Becke Density-functional thermochemistry. III. The role of exact exchange , 1993 .

[10]  R. Ornstein,et al.  Theoretical Determination of Two Structural Forms of the Active Site in Cadmium-Containing Phosphotriesterases , 2002 .

[11]  F. Raushel,et al.  Perturbations to the active site of phosphotriesterase. , 1997, Biochemistry.

[12]  R. Ornstein,et al.  Mobility of the active site bound paraoxon and sarin in zinc-phosphotriesterase by molecular dynamics simulation and quantum chemical calculation. , 2001, Journal of the American Chemical Society.

[13]  G. Corongiu,et al.  Monte Carlo simulations of water clusters around Zn++ and a linear Zn++⋅CO2 complex , 1980 .

[14]  Nack-Do Sung,et al.  STRUCTURE AND REACTIVITY OF DINUCLEAR COBALT(III) COMPLEXES WITH PEROXIDE AND PHOSPHATE DIESTER ANALOGUES BRIDGING THE METAL IONS , 1998 .

[15]  B. Vallee,et al.  Functional zinc-binding motifs in enzymes and DNA-binding proteins. , 1992, Faraday discussions.

[16]  P. Kollman,et al.  A Second Generation Force Field for the Simulation of Proteins, Nucleic Acids, and Organic Molecules , 1995 .

[17]  A. Warshel,et al.  Computer simulation of the initial proton transfer step in human carbonic anhydrase I. , 1992, Journal of molecular biology.

[18]  Parr,et al.  Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. , 1988, Physical review. B, Condensed matter.

[19]  M Bolognesi,et al.  Evolutionary constraints for dimer formation in prokaryotic Cu,Zn superoxide dismutase. , 1999, Journal of molecular biology.

[20]  F. Raushel,et al.  Mechanism-based inhibitors for the inactivation of the bacterial phosphotriesterase. , 1997, Biochemistry.

[21]  L. Nilsson,et al.  Structural and dynamic differences of the estrogen receptor DNA-binding domain, binding as a dimer and as a monomer to DNA: molecular dynamics simulation studies , 1999, European Biophysics Journal.

[22]  Orlando Tapia,et al.  MOLECULAR DYNAMICS SIMULATION OF A ZIF268-DNA COMPLEX IN WATER. SPATIAL PATTERNS AND FLUCTUATIONS SENSED FROM A NANOSECOND TRAJECTORY , 1998 .

[23]  G. Ciccotti,et al.  Numerical Integration of the Cartesian Equations of Motion of a System with Constraints: Molecular Dynamics of n-Alkanes , 1977 .

[24]  Johan Åqvist,et al.  Modelling of ion-ligand interactions in solutions and biomolecules , 1992 .

[25]  Thomas V. O'Halloran,et al.  Transition metals in control of gene expression. , 1993, Science.

[26]  F. Raushel,et al.  Metal-substrate interactions facilitate the catalytic activity of the bacterial phosphotriesterase. , 1996, Biochemistry.

[27]  F. Raushel,et al.  Stereochemical constraints on the substrate specificity of phosphotriesterase. , 1999, Biochemistry.

[28]  W. Lipscomb,et al.  Binding of substrate CO2 to the active site of human carbonic anhydrase II: a molecular dynamics study. , 1990, Proceedings of the National Academy of Sciences of the United States of America.

[29]  Kenneth M. Merz,et al.  Force Field Design for Metalloproteins , 1991 .

[30]  Y. Pang Successful molecular dynamics simulation of two zinc complexes bridged by a hydroxide in phosphotriesterase using the cationic dummy atom method , 2001, Proteins.

[31]  B Chance,et al.  X-ray absorption fine structure study of the active site of zinc and cobalt carboxypeptidase A in their solution and crystalline forms. , 1992, Biochemistry.

[32]  F. Raushel,et al.  High resolution X-ray structures of different metal-substituted forms of phosphotriesterase from Pseudomonas diminuta. , 2001, Biochemistry.

[33]  Brian W. Matthews,et al.  Structural basis of the action of thermolysin and related zinc peptidases , 1988 .

[34]  M Karplus,et al.  Zinc binding in proteins and solution: A simple but accurate nonbonded representation , 1995, Proteins.

[35]  F. Raushel,et al.  Characterization of the zinc binding site of bacterial phosphotriesterase. , 1992, The Journal of biological chemistry.

[36]  A. Guy Orpen,et al.  A new class of chiral bridged metallocene: Synthesis, structure, and olefin (co)polymerization behavior of rac- and meso-1,2-CH2CH2{4-(7-Me-indenyl)}(2)ZrCl2 , 1998 .

[37]  F. Raushel,et al.  The Binding of Substrate Analogs to Phosphotriesterase* , 2000, The Journal of Biological Chemistry.