k θ (θ θ 0 ) 2 angles r i j r i j

Similar documents
Lecture 11: Potential Energy Functions

Molecular mechanics. classical description of molecules. Marcus Elstner and Tomáš Kubař. April 29, 2016

Force fields, thermo- and barostats. Berk Hess

All-atom Molecular Mechanics. Trent E. Balius AMS 535 / CHE /27/2010

CE 530 Molecular Simulation

Molecular Mechanics. Yohann Moreau. November 26, 2015

Structural Bioinformatics (C3210) Molecular Mechanics

Chapter 11 Molecular Mechanics

Molecular Modelling. part of Bioinformatik von RNA- und Proteinstrukturen. Sonja Prohaska. Leipzig, SS Computational EvoDevo University Leipzig

Dihedral Angles. Homayoun Valafar. Department of Computer Science and Engineering, USC 02/03/10 CSCE 769

arxiv: v1 [cond-mat.stat-mech] 6 Jan 2014

Example questions for Molecular modelling (Level 4) Dr. Adrian Mulholland

Molecular Mechanics. I. Quantum mechanical treatment of molecular systems

Biomolecular modeling I

Biochemistry,530:,, Introduc5on,to,Structural,Biology, Autumn,Quarter,2015,

Molecular Dynamic Simulation Study of the Volume Transition of PNIPAAm Hydrogels

Peptide folding in non-aqueous environments investigated with molecular dynamics simulations Soto Becerra, Patricia

Molecular Mechanics / ReaxFF

Biomolecular modeling I

3rd Advanced in silico Drug Design KFC/ADD Molecular mechanics intro Karel Berka, Ph.D. Martin Lepšík, Ph.D. Pavel Polishchuk, Ph.D.

Computational Chemistry. An Introduction to Molecular Dynamic Simulations

Hyeyoung Shin a, Tod A. Pascal ab, William A. Goddard III abc*, and Hyungjun Kim a* Korea

T6.2 Molecular Mechanics

Force Fields for Classical Molecular Dynamics simulations of Biomolecules. Emad Tajkhorshid

This semester. Books

Supporting Information

Nanotube AFM Probe Resolution

ICCP Project 2 - Advanced Monte Carlo Methods Choose one of the three options below

POTENTIAL FUNCTIONS FOR HYDROGEN BONDS IN PROTEIN STRUCTURE PREDICTION AND DESIGN. By ALEXANDRE V. MOROZOV* AND TANJA KORTEMME {

Force fields in computer simulation of soft nanomaterials

Bioengineering 215. An Introduction to Molecular Dynamics for Biomolecules

A Molecular Dynamics Simulation of a Homogeneous Organic-Inorganic Hybrid Silica Membrane

The Dominant Interaction Between Peptide and Urea is Electrostatic in Nature: A Molecular Dynamics Simulation Study

Ion-Gated Gas Separation through Porous Graphene

Introduction to molecular dynamics

An introduction to Molecular Dynamics. EMBO, June 2016

Intermolecular Forces in Density Functional Theory

Session 1. Introduction to Computational Chemistry. Computational (chemistry education) and/or (Computational chemistry) education

Molecular Aggregation

74 these states cannot be reliably obtained from experiments. In addition, the barriers between the local minima can also not be obtained reliably fro

An Informal AMBER Small Molecule Force Field :

Molecular Mechanics. C. David Sherrill School of Chemistry and Biochemistry Georgia Institute of Technology. January 2001

Solutions to Assignment #4 Getting Started with HyperChem

Why study protein dynamics?

Reactive Empirical Force Fields

Force Fields in Molecular Mechanics

510 Subject Index. Hamiltonian 33, 86, 88, 89 Hamilton operator 34, 164, 166

Structural and dynamical properties of Polyethylenimine in explicit water at different protonation states: A Molecular Dynamics Study

Molecular Dynamics, Monte Carlo and Docking. Lecture 21. Introduction to Bioinformatics MNW2

Interatomic Potentials. The electronic-structure problem

Close agreement between the orientation dependence of hydrogen bonds observed in protein structures and quantum mechanical calculations

UB association bias algorithm applied to the simulation of hydrogen fluoride

3. Solutions W = N!/(N A!N B!) (3.1) Using Stirling s approximation ln(n!) = NlnN N: ΔS mix = k (N A lnn + N B lnn N A lnn A N B lnn B ) (3.

Supporting Information

Dynamics. capacities of. (A.I.1) (E b ), bond. E total stretching (A.I.2) (A.I.3) A.I.1. Ebond force (A.I.4) (A.I.5)

Molecular Dynamics Simulations. Dr. Noelia Faginas Lago Dipartimento di Chimica,Biologia e Biotecnologie Università di Perugia

The Effect of Model Internal Flexibility Upon NEMD Simulations of Viscosity

Introduction to Classical Molecular Dynamics. Giovanni Chillemi HPC department, CINECA

Why Is Molecular Interaction Important in Our Life

New Six-site Acetonitrile Model for Simulations of Liquid Acetonitrile and its Aqueous Mixtures

Electronic excitations in conjugated. Many-Body Green's Functions. Behnaz Bagheri Varnousfaderani. Behnaz Bagheri Varnousfaderani

Water models in classical simulations

Statistical Mechanics for Proteins

Equations of State. Equations of State (EoS)

Atoms & Their Interactions

Scuola di Chimica Computazionale

Advanced Quantum Chemistry III: Part 6

Universal Repulsive Contribution to the. Solvent-Induced Interaction Between Sizable, Curved Hydrophobes: Supporting Information

The Potential Energy Surface (PES) And the Basic Force Field Chem 4021/8021 Video II.iii

Polypeptide Folding Using Monte Carlo Sampling, Concerted Rotation, and Continuum Solvation

Density Functional Theory: from theory to Applications

The Molecular Dynamics Method

Chapter 2 Structure and Properties of Organic Molecules. Advanced Bonding: Review

Advanced Electronic Structure Theory Density functional theory. Dr Fred Manby

Non-covalent force fields computed ab initio

Proteins polymer molecules, folded in complex structures. Konstantin Popov Department of Biochemistry and Biophysics

Exercise 1: Structure and dipole moment of a small molecule

Supplemntary Infomation: The nanostructure of. a lithium glyme solvate ionic liquid at electrified. interfaces

Molecular Simulation II. Classical Mechanical Treatment

Computational Methods. Chem 561

Figure 1: Transition State, Saddle Point, Reaction Pathway

Chapter 6 Cyclic urea - a new central unit in bent-core compounds

Advanced Quantum Chemistry III: Part 6

4 th Advanced in silico Drug Design KFC/ADD Molecular Modelling Intro. Karel Berka, Ph.D.

DFT calculations of NMR indirect spin spin coupling constants

Potentials, periodicity

Coarse-Grained Models!

CHEMISTRY 4021/8021 MIDTERM EXAM 1 SPRING 2014

Potential Energy (hyper)surface

Aqueous solutions. Solubility of different compounds in water

Molecular Mechanics, Dynamics & Docking

Computer simulation methods (2) Dr. Vania Calandrini

Free Radical-Initiated Unfolding of Peptide Secondary Structure Elements

Force Fields for Classical Molecular Dynamics simulations of Biomolecules. Emad Tajkhorshid

Reactive potentials and applications

CHAPTER-IV. FT-IR and FT-Raman investigation on m-xylol using ab-initio HF and DFT calculations

arxiv: v1 [cond-mat.soft] 22 Oct 2007

Force Fields for MD simulations

Wave Properties of Electrons. Chapter 2 Structure and Properties of Organic Molecules. Wave Interactions. Sigma Bonding

Transition states and reaction paths

Transcription:

1 Force fields 1.1 Introduction The term force field is slightly misleading, since it refers to the parameters of the potential used to calculate the forces (via gradient) in molecular dynamics simulations. The underlying idea is to create a certain number of atom types upon which any bonds, angles, impropers, dihedrals and long-range interactions may be described. More atomtypes than elements are necessary, since the chemical surroundings greatly influence the parameters. The angle between carbons in alkyne side chains will obviously differ from that of carbons in a benzene ring, so two different atom types are used to describe them. Since all simultations rely greatly upon the correctness of these parameters, it is important to know, where these parameters come from and how they may be derived in case some are lacking. One of the most commonly used force fields is the OPLS (Optimized Potential for Liquid Simulations) force field, developed by William L. Jorgensen at Purdue University and later at Yale University. It shall be our main example here. 1.2 A word on AMBER The functional form of the OPLS force field is very similar to that of AMBER (Assisted Model Building and Energy Refinement), originally developed by Peter Kollman s group at the University of California San Francisco, which has the form [1, 2]: V ({ r N }) = bonds + torsions N 1 + i=1 j=i+1 1 2 k b(r r 0 ) 2 + 1.3 Functional form of OPLS angles k θ (θ θ 0 ) 2 1 2 V n[1 + cos(nω γ)] { [ N (σij ) 12 ( ) ] } 6 σij 4ǫ ij + + q iq j r i j r i j 4πǫ 0 r ij Based upon AMBER, Jorgensen et al. created the OPLS force field using the following functional form [3, 4, 5, 6, 7]: V ({ r N }) = 1 2 k b(r r 0 ) 2 + k θ (θ θ 0 ) 2 bonds + torsions N 1 + { V1 N i=1 j=i+1 angles 2 [1 + cos(φ)] + V 2 2 [1 cos(2φ)] + V 3 2 [1 + cos(3φ)] + V } 4 [1 cos(4φ)] 2 { [ (σij ) 12 ( ) ] } 6 σij 4ǫ ij + + q iq j e 2 f ij r i j r i j r ij

1.4 Derivation of the parameters Bonds and angles are described by harmonic potentials, since they are very strong and fluctuate only slightly around their equilibrium values at room temperature. In many simulations, bonds and even angles are constrained to fixed values to save computational effort. The dihedral potential is described by a cosine expansion and may take any value within 360 depending on the height of the barrier between the low energy conformations, which makes the precision of the dihedral potential barrier crucial for many polymer properties. Dihedral potentials always possess a symmetry around 180. The long range interactions are only counted for atoms three or more bonds apart. They are consist of Coulomb and Lennard Jones two-body interaction terms. The Lennard Jones potential is a combination of attractive van der Waals forces due to dipole-dipole interactions and empirical repulsive forces due to Pauli repulsion. The scaling factor f ij equals 0.5 for 1-4 interactions and one otherwise. 1.4 Derivation of the parameters Since most molecular dynamics publications simply cite the force field used or give a list of parameters without further explanation, this section shall shed some light on how these parameters are found [1]. It should be noted, that in general any force field derivation must start with very small molecules, such as CH 3, CH 4, or maybe benzene, to limit the number of variables. Larger molecules can only be considered once the force field parameters for most constituents are considered to be known. 1.4.1 Bonds and angles The equilibrium values for bonds as well as angles are usually simply taken from x-ray data. The values for the force constants are derived by fitting to experimental vibrational frequency data. While one should actually go through a more complex procedure to ensure that the geometries of simple molecules match experimental data as good as possible after energy minimiztion, it is believed that in most cases the difference is negligible. Fig. 1: Bond and angle between atoms i,j,k including corresponding force constants.

1.4 Derivation of the parameters 1.4.2 Dihedrals First off, there are two different approaches as to how to calculate dihedral potentials. One is to optimize the dihedral potential for the simplest possible molecule and then apply it to larger ones containing the same dihedral, the other is to optimize the dihedral parameters to best describe a large number of different molecules. Note that, while the latter method might sound more accurate at first, it also leads to a dependence on the set of chosen molecules. In both cases, the dihedral parameters are computed from ab initio methods as follows [8]: i. Ab initio calculations Scan dihedral (or improper) of interest Optimize geometry at each step Calculate change in potential energy For this purpose, either perturbation (MP2), restricted Hartree-Fock (RHF) or hybrid methods between Hartree-Fock and density functional theory (B3LYP) are used. The basis sets chosen for the geometry optimization are at least 6-31g and may go up to 6-311g**, depending on the size of the molecule in question. Fig. 2: Dihedral and improper between atoms i,j,k,l including corresponding force constants. ii. Potential energy according to MD Set dihedral parameters to zero Compute potential energy of each optimized configuration Note that this requires the knowledge of all other force field parameters in the molecule. iii. Fitting of the parameters Subtract MD from ab initio results

1.4 Derivation of the parameters Obtain parameters from fitting a dihedral function to the resulting curve The difference between ab initio and MD results corresponds exactly to the influence of the dihedral, since all other interactions are already included in the force field. The most common mathematical representations of dihedrals are: Proper dihedrals Ryckaert-Belleman dihedrals where φ = ψ 180. iv. Checking the results V ({r}) = C n (1 + cos(nφ)) V ({r}) = 5 C n cos n (ψ) n=0 Rerun MD simulations using new parameters Compare MD and ab initio results 1.4.3 Charges To obtain the partial charges of a molecule in question, whose values will no longer be altered during the MD runs (unless a polarizable force field is used), it is important to make sure the charge calculations are done in the equilibrium, conformation. This may be achieved by optimizing the geometry using any ab inito method with a basis set better than or equal to 6-31G. The charges may then be obtained by fitting to the electrostatic potential, i.e. adjusting the partial charges at the centers of the nuclei in such a fashion that the electrostatic potential given by the wave functions is best reproduced. One of the favored methods is CHELPG (CHarges from ELectrostatic Potentials using a Grid based method) [9]. Charge calculations are often done using higher-level methods and basis sets than the geometry optimization, e.g. B3LYP/cc-pVTZ or HF/6-31G**. There are two known problems with this straigh-forward approach however: Firstly, oftentimes considerable variation is seen when charges are computed for different conformations of a molecules, which is especially problematic for molecules with multiple low energy conformations such as propylamine [10]. Secondly, standard electrostatic potential (ESP) charges tend to miscalculate charges of buried atoms, since they are statistically underdetermined and often assume too large values for nonpolar atoms. For these reasons, the restricted ESP charge model was developed [11]. It consist of a least-squares fit of the charges to the electrostatic potential (as before), but with hyperbolic restraints on heavy atom charges. This is followed by a second fitting stage, needed to fit methyl groups which require equivalent charges on hydrogen atoms which are not equivalent by molecular symmetry.

1.5 Testing the force field 1.4.4 Van der Waals parameters This is the area where the difference between OPLS/AMBER and other force fields comes into play: Jorgensen et al. pioneered on creating force field parameters for organic molecules focusing on systems explicitely taking the solvent into account. As a matter of fact, the OPLS/AMBER force field for peptides and proteins takes most bond, angle and dihedral parameters from the force field developed by Weiner et al [12]. This is achieved by carrying out Monte Carlo (MC) simulations of organic liquids, e.g. CH 4 or C 2 H 6, and then empirically adjusting the Lennard Jones (σ and ǫ) parameters to match the experimental densities and enthalpies of vaporization. A difficult issue is the factor to scale down the Lennard Jones 1-4 interactions. It has to be done, since otherwise the r 12 term would lead to unphysically high repulsions. The value chosen is somewhat arbitrary and force field dependent, however. The same holds true for the combination rules for the Lennard Jones parameters, i.e. the choice of LJ parameters for the interaction between two different atom types. There are two main competing methods: Lorentz-Bertelot Geometrical average σ ij = 1 2 (σ ii + σ jj ) ǫ ij = (ǫ ii ǫ jj ) 1/2 σ ij = (σ ii σ jj ) 1/2 ǫ ij = (ǫ ii ǫ jj ) 1/2 The method should be chosen in accordance with the force field to be used. 1.5 Testing the force field The validity of the force field can be tested in multiple ways. For proteins, it is important that the Ramachandran plots for the dihedral angles correspond to those obtained experimentally to ensure sampling of the correct secondary structures in simulations. For liquids, the radial distribution function as well as the density should be correctly reproduced. The radius of gyration and the persistence length are checked for polymer chains. 1.6 Some details for use with Gromacs When one has chosen a force field to use with Gromacs it is important that some input parameters for the simulations are in accordance with this force field. All important settings may be found in the headers of the itp and rtp files. They are as follows:

1.6 Some details for use with Gromacs 1.6.1 itp-file nbfunc type of non-bonded function to be used 1: Lennard Jones (OPLS, AMBER) 2: Buckingham comb-rule combination rule to be used for non-bonded interactions 1: Given C 6 and C N Cij M = Ci M Cj M 2: σ ij = 1 2 (σ 1 + σ j ) and ǫ ij = ǫ i ǫ j (AMBER) 3: σ ij = σ 1 σ j and ǫ ij = ǫ i ǫ j (OPLS) genpairs generate LJ 1-4 pairs? yes (OPLS, AMBER) no Note: do not add [pairtypes] section, if genpairs = yes fudgelj scaling factor for LJ 1-4 interactions 0.5 (OPLS, AMBER) fudgeqq scaling factor for Coulomb 1-4 interactions 1.6.2 rtp-file 0.8333 (AMBER) 0.5 (OPLS-UA) 1/8 (OPLS-AA) bonds function used to describe bonded potential 1: bond (OPLS, AMBER) 2: G96 bond 3: morse 4: cubic

1.6 Some details for use with Gromacs 5: connection 6: harmonic potential 7: FENE angles function used to describe angular potential 1: angle (OPLS, AMBER) 2: G96 angle 3: quadratic angle dihedrals function used to describe dihedral potential 1: proper 2: improper 3: Ryckaert-Belleman (OPLS, AMBER) impropers function used to describe improper potential 1: proper (OPLS, AMBER) 2: improper 3: Ryckaert-Belleman all dihedrals generate all dihedrals? 0: only for heavy atoms 1: yes (OPLS, AMBER) nrexcl number of excluded neighbors for non-bonded interactions 3 (default, should not be changed) HH14 generate 1-4 interaction pairs between hydrogens? 0: no 1: yes (OPLS, AMBER) RemoveDih remove propers over the same bond as a improper?

REFERENCES 0: no (OPLS, AMBER) 1: yes All above informations are based on Gromacs version 3.3.1. References [1] W. Cornell, P. Cieplak, P.A. Kollmann et al., J. Am. Chem. Soc. 117, 5179-5197 (1995). [2] Y. Duan, C. Wu, P. Kollmann et al., J. Comp. Chem. 24, 1999-2012 (2003). [3] W.L. Jorgensen and J. Tirado-Rives, J. Am. Chem. Soc. 110, 1657-1666 (1988). [4] W.L. Jorgensen, D.S. Maxwell, and J. Tirado-Rives, J. Am. Chem. Soc. 118, 11225-11236 (1996). [5] W.L. Jorgensen, Encyclopedia of Comp. Chem., 1986-1988 (1998). [6] W.L. Jorgensen, Encyclopedia of Comp. Chem., 1754-1762 (1998). [7] W.L. Jorgensen and J. Tirado-Rives, PNAS 102, 6665-6670 (2005). [8] A.D. Mackerell, J. Comp. Chem. 25, 1584-1604 (2004). [9] C.M. Breneman,and K.B. Wiberg, J. Comp. Chem. 11, 361-373 (1989). [10] W. Cornell, P. Cieplak, C. Bayly, and P.A. Kollmann, J. Am. Chem. Soc. 115, 9620-9631 (1993). [11] C. Bayly, P. Cieplak, W. Cornell, and P.A. Kollmann, J. Phys. Chem. 97, 10269-10280 (1993). [12] S.J. Weiner, P.A. Kollmann, P.J. Weiner et al., J. Am. Chem. Soc. 106, 765-784 (1984).