RM1: A reparameterization of AM1 for H, C, N, O, P, S, F, Cl, Br, and I

Twenty years ago, the landmark AM1 was introduced, and has since had an increasingly wide following among chemists due to its consistently good results and time‐tested reliability—being presently available in countless computational quantum chemistry programs. However, semiempirical molecular orbital models still are of limited accuracy and need to be improved if the full potential of new linear scaling techniques, such as MOZYME and LocalSCF, is to be realized. Accordingly, in this article we present RM1 (Recife Model 1): a reparameterization of AM1. As before, the properties used in the parameterization procedure were: heats of formation, dipole moments, ionization potentials and geometric variables (bond lengths and angles). Considering that the vast majority of molecules of importance to life can be assembled by using only six elements: C, H, N, O, P, and S, and that by adding the halogens we can now build most molecules of importance to pharmaceutical research, our training set consisted of 1736 molecules, representative of organic and biochemistry, containing C, H, N, O, P, S, F, Cl, Br, and I atoms. Unlike AM1, and similar to PM3, all RM1 parameters have been optimized. For enthalpies of formation, dipole moments, ionization potentials, and interatomic distances, the average errors in RM1, for the 1736 molecules, are less than those for AM1, PM3, and PM5. Indeed, the average errors in kcal · mol−1 of the enthalpies of formation for AM1, PM3, and PM5 are 11.15, 7.98, and 6.03, whereas for RM1 this value is 5.77. The errors, in Debye, of the dipole moments for AM1, PM3, PM5, and RM1 are, respectively, 0.37, 0.38, 0.50, and 0.34. Likewise, the respective errors for the ionization potentials, in eV, are 0.60, 0.55, 0.48, and 0.45, and the respective errors, in angstroms, for the interatomic distances are 0.036, 0.029, 0.037, and 0.027. The RM1 average error in bond angles of 6.82° is only slightly higher than the AM1 figure of 5.88°, and both are much smaller than the PM3 and PM5 figures of 6.98° and 9.83°, respectively. Moreover, a known error in PM3 nitrogen charges is corrected in RM1. Therefore, RM1 represents an improvement over AM1 and its similar successor PM3, and is probably very competitive with PM5, which is a somewhat different model, and not fully disclosed. RM1 possesses the same analytical construct and the same number of parameters for each atom as AM1, and, therefore, can be easily implemented in any software that already has AM1, not requiring any change in any line of code, with the sole exception of the values of the parameters themselves. © 2006 Wiley Periodicals, Inc. J Comput Chem 27: 1101–1111, 2006

[1]  Guojie Li,et al.  From the Editor-in-Chief , 1995, Journal of Computer Science and Technology.

[2]  F. Cichos,et al.  Coupled molecular dynamics/semiempirical simulation of organic solutes in polar liquids. I. Naphthalene in acetonitrile , 2001 .

[3]  P. Kollman,et al.  Encyclopedia of computational chemistry , 1998 .

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

[5]  Alistair P. Rendell,et al.  Comparison of linear‐scaling semiempirical methods and combined quantum mechanical/molecular mechanical methods for enzymic reactions. II. An energy decomposition analysis , 2002, J. Comput. Chem..

[6]  Y. Inoue,et al.  Effects of hydration on the electronic structure of an enzyme: implications for the catalytic function. , 2001, Journal of the American Chemical Society.

[7]  Walter Thiel,et al.  Orthogonalization corrections for semiempirical methods , 2000 .

[8]  Sándor Suhai,et al.  Self-consistent-charge density-functional tight-binding method for simulations of complex materials properties , 1998 .

[9]  William L. Jorgensen,et al.  PDDG/PM3 and PDDG/MNDO: Improved semiempirical methods , 2002, J. Comput. Chem..

[10]  Minoru Sakurai,et al.  Effect of Protein Environment on pKa Shifts in the Active Site of Photoactive Yellow Protein , 2003 .

[11]  Volkhard Helms,et al.  Mechanism of phosphoryl transfer in kinases investigated by semiempirical calculations , 2003 .

[12]  Eamonn F. Healy,et al.  Development and use of quantum mechanical molecular models. 76. AM1: a new general purpose quantum mechanical molecular model , 1985 .

[13]  Victor M. Anisimov,et al.  Accuracy assessment of semiempirical molecular electrostatic potential of proteins , 2003 .

[14]  Marc Albe,et al.  The dynamo library for molecular simulations using hybrid quantum mechanical and molecular mechanical potentials , 2000 .

[15]  Noriyuki Kurita,et al.  Electronic properties of DNA binding site in transcription OmpR family: semiempirical molecular orbital calculations , 2003 .

[16]  Minoru Sakurai,et al.  Quantum chemical study of the pKa control mechanism for the active center in bacteriorhodopsin and its M intermediate , 2003 .

[17]  Amedeo Caflisch,et al.  Computational combinatorial chemistry for de novo ligand design: Review and assessment , 1995 .

[18]  Fabio Pichierri,et al.  Computation of the permanent dipole moment of α-chymotrypsin from linear-scaling semiempirical quantum mechanical methods , 2003 .

[19]  Alistair P. Rendell,et al.  Comparison of semiempirical and ab initio QM decomposition analyses for the interaction energy between molecules , 2002 .

[20]  Yonezo Morino,et al.  Average structures of butadiene, acrolein, and glyoxal determined by gas electron diffraction and spectroscopy , 1968 .

[21]  Fabio Pichierri,et al.  A quantum mechanical study on phosphotyrosyl peptide binding to the SH2 domain of p56lck tyrosine kinase with insights into the biochemistry of intracellular signal transduction events. , 2004, Biophysical chemistry.

[22]  Vladimir B. Sulimov,et al.  Semiempirical calculations of binding enthalpy for protein-ligand complexes , 2004 .

[23]  Juha Linnanto,et al.  Semiempirical PM5 molecular orbital study on chlorophylls and bacteriochlorophylls: Comparison of semiempirical, ab initio, and density functional results , 2004, J. Comput. Chem..

[24]  H Luecke,et al.  Molecular mechanism of spectral tuning in sensory rhodopsin II. , 2001, Biochemistry.

[25]  Fabio Pichierri,et al.  Elastic properties of the poly-l-glycine α-helix from periodic SCF-LMO calculations , 2000 .

[26]  James J. P. Stewart,et al.  Application of localized molecular orbitals to the solution of semiempirical self‐consistent field equations , 1996 .

[27]  Alistair P. Rendell,et al.  The Potassium Ion Channel: Comparison of Linear Scaling Semiempirical and Molecular Mechanics Representations of the Electrostatic Potential , 2001 .

[28]  J. Stewart Optimization of parameters for semiempirical methods II. Applications , 1989 .

[29]  Dale G. Drueckhammer,et al.  Computational Study of the Citrate Synthase Catalyzed Deprotonation of Acetyl-Coenzyme A and Fluoroacetyl-Coenzyme A: Demonstration of a Layered Quantum Mechanical Approach , 2003 .

[30]  Alfredo M. Simas,et al.  MESOIONIC RINGS AS EFFICIENT ASYMMETRIC BRIDGES FOR THE DESIGN OF COMPOUNDS WITH LARGE OPTICAL NONLINEARITIES , 1996 .

[31]  S. Lias,et al.  Ionization Potential and Appearance Potential Measurements, 1971-1981, , 1982 .

[32]  J. Stewart Optimization of parameters for semiempirical methods I. Method , 1989 .

[33]  Alistair P. Rendell,et al.  Comparison of linear-scaling semiempirical methods and combined quantum mechanical/molecular mechanical methods applied to enzyme reactions , 2000 .

[34]  Minoru Sakurai,et al.  Application of an integrated MOZYME+DFT method to pKa calculations for proteins , 2001 .

[35]  A. L. McClellan,et al.  Tables of experimental dipole moments , 1963 .

[36]  M Elstner,et al.  Quantum mechanics simulation of protein dynamics on long timescale , 2001, Proteins.