ebook img

Predicting Hydration Gibbs Energies of Alkyl-aromatics Using Molecular Simulation PDF

54 Pages·2011·1.54 MB·English
by  
Save to my drive
Quick download
Download
Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.

Preview Predicting Hydration Gibbs Energies of Alkyl-aromatics Using Molecular Simulation

Page 1 of 54 Physical Chemistry Chemical Physics Predicting Hydration Gibbs Energies of Alkyl-aromatics Using Molecular Simulation: A Comparison of Current Force Fields and the Development of a New Parameter Set for Accurate Solvation Data Nuno M. Garrido1, Miguel Jorge1, António J. Queimada1, José R. B. Gomes2, Ioannis G. Economou3 and Eugénia A. Macedo1,* 1. LSRE Laboratory of Separation and Reaction Engineering, Departamento de Engenharia Química, Faculdade de Engenharia, Universidade do Porto, Rua do Dr. Roberto Frias, 4200 - 465 Porto, Portugal 2. CICECO, Department of Chemistry, University of Aveiro, Campus Universitário de Santiago, Aveiro 3810-193, Portugal 3. The Petroleum Institute, Department of Chemical Engineering, PO Box 2533, Abu Dhabi, United Arab Emirates *Author to whom all correspondence should be addressed at: [email protected] Manuscript submitted for publication in Physical Chemistry Chemical Physics April 2011 1 Physical Chemistry Chemical Physics Page 2 of 54 ABSTRACT The Gibbs energy of hydration is an important quantity to understand the molecular behavior in aqueous systems at constant temperature and pressure. In this work we review the performance of some popular force fields, namely TraPPE, OPLS-AA and Gromos, in reproducing the experimental Gibbs energies of hydration of several alkyl- aromatic compounds – benzene, mono-, di- and tri-substituted alkylbenzenes – using molecular simulation techniques. In the second part of the paper, we report a new model that is able to improve such hydration energy predictions, based on LJ parameters from the recent TraPPE-EH force field and atomic partial charges obtained from natural population analysis of density functional theory calculations. We apply a scaling factor determined by fitting the experimental hydration energy of only two solutes, and then present a simple rule to generate atomic partial charges for different substituted alkyl- aromatics. This rule has the added advantages of eliminating the unnecessary assumption of fixed charge on every substituted carbon atom and providing a simple guideline for extrapolating the charge assignment to any multi-substituted alkyl- aromatic molecule. The point charges derived here yield excellent predictions of experimental Gibbs energies of hydration, with an overall absolute average deviation of less than 0.6 kJ/mol. This new parameter set can also give good predictive performance for other thermodynamic properties and liquid structural information. 2 Page 3 of 54 Physical Chemistry Chemical Physics 1. Introduction Over the last decades, free energy calculations have continuously become more important in the chemical and biochemical fields, as noted in several literature reviews1-4. Free energy studies allow scientists and engineers to achieve microscopic- level insights into the behavior of complex systems that may be helpful in chemical process design, e.g. to estimate important properties such as solubility5,6 or partition coefficients7,8. Accurate predictions of the Gibbs energy of solvation, also called solvation free energy, have the capacity to revolutionize several fields, predominantly in the pharmaceutical industry and/or in biological chemistry9,10. In this context, molecular simulation has emerged as a highly promising method to estimate free energies of drug- like molecules10,11. The importance and at the same time the difficulty of connecting classical atomistic models to experimental results in order to predict the Gibbs energy of solvation is also illustrated in a recent blind challenge12. One of the major drawbacks of free energy predictions from molecular simulations is that they tend to be quite sensitive to the force field and in some cases to the simulation parameters (cutoff radii, treatment of long-range electrostatics, etc.).13 For this reason, several studies have focused on comparing existing force fields14-17 or on fine-tuning existing ones18-20 for solvation Gibbs energy predictions. Some of those studies have shown that the electrostatic contribution to the calculated Gibbs energy, commonly dependent on an adequate assignment of partial charges to each atom or group of atoms, is particularly prone to uncertainties, and sometimes hinders the transferability of the force field parameters21,22. In fact, an important shortcoming of fixed-charge force fields is that they do not normally take into account the different polarization environment when a molecule is transferred from the liquid phase to the gas phase (or vice-versa)23,24. Recently, our group has also tested the performance of three force-fields in predicting the Gibbs energy of hydration for the n-alkane series7 and some representative polar compounds25. One of the conclusions was that “generic” force fields were able to provide reasonable predictions of Gibbs energy of solvation of simple and monofunctional compounds, but they generally failed for more complex polyfunctional molecules, where the deviation from experimental data can be larger than 10 kJ/mol. This justifies the need for a more detailed study of the effects of force field parameters on the Gibbs energy of solvation, most notably of partial atomic charges, since the electrostatic contribution usually represents a very large percentage of 3 Physical Chemistry Chemical Physics Page 4 of 54 the total hydration energy for polar solutes21,22. Our strategy is to systematically study a homologous series of compounds, rather than to apply the usual approach of testing the force fields against a large database of solutes with many different functional moieties. In this way, we hope to identify more easily the reasons for the success or failure of certain parameter sets, and to establish a well-defined set of rules for force field development. In the present work, we focus on the aqueous behavior of benzene and several alkyl-substituted aromatics: mono-, di- and tri-substituted alkylbenzenes. Such aromatic fragments are widely present in chemical and pharmaceutical compounds, biological assemblies and petrochemical systems. Xylenes are a particular family of di-substituted alkylbenzene isomers that are used as large scale industrial solvents and intermediates for many derivatives, e.g, p-xylene is used in the polyester industry. Besides their importance as industrial chemicals, aromatic compounds were also found to play a key role in biochemistry26-28. This class of compounds is also particularly interesting from a theoretical point of view, as aromatic interactions are usually difficult to quantify experimentally29. To the best of our knowledge there is no published systematic study about the prediction of the Gibbs energy of hydration of alkyl-aromatic compounds from molecular simulation. Thus, in the first part of this work we predict the Gibbs energies of hydration of benzene and five mono-substituted alkyl-benzenes using five current force field parameter sets from the TraPPE, Gromos and OPLS families. When modeling polyatomic molecules, two approaches are generally followed to represent the dispersion and repulsive forces: i) all-atom (AA) models; ii) united-atom (UA) models. In AA models, a separate force center is assigned to each atom, located on its nucleus. In UA models specific groups of atoms, such as CH, CH2 or CH3, are treated as single force centers. The advantage of AA models is that they generally give a better account of molecular geometry and structure, as well as dynamic properties30,31. The counterpart is that they require a great deal of computer time because of the larger number of force centers, bending angles and torsion angles involved. UA methods generally neglect the hydrogen atoms connected to carbon atoms, and their influence is considered implicitly in the parameterisation of potential parameters for the pseudo- atoms. A “hybrid” approach is also possible, where aromatic hydrogens are treated explicitly, while aliphatic hydrogens are treated within a UA approximation. In this paper, we will call this an explicit-hydrogen (EH) model. 4 Page 5 of 54 Physical Chemistry Chemical Physics The TraPPE force field was originally developed under a UA philosophy, where both aromatic CH and aliphatic CH groups were treated as pseudo-atoms whose x molecular weights are equivalent to those of the whole fragment (henceforth denoted by TraPPE-UA31-34). More recently, an EH version of the TraPPE force-field was proposed for benzene35 and we have also decided to test this version of TraPPE (henceforth denoted by TraPPE-EH). However, this force field has not yet been extended to alkyl- aromatic compounds. Likewise, the Gromos biomolecular force field was originally developed as a united atom force field36 (Gromos-UA). However, recent developments have provided parameters for an EH description of the aromatic CH groups (Gromos- EH, version 53A537), which were parameterized to reproduce, among other properties, the Gibbs energy of hydration of amino acids containing aromatic groups37. Finally, although starting with a UA description, the Optimized Potential for Liquid Simulations (OPLS)30 force field quickly became popular with its AA description of the solute molecules. Moreover, for aromatic molecules the OPLS-UA model was not parameterized. One of the purposes of this paper is to understand how accurate the current parameter sets are in reproducing experimental hydration energies, and to evaluate the need for using optimized parameters to correctly achieve hydration data predictions. In the second part of this paper, we attempt to improve the prediction of hydration energies, by generating a new charge set for an EH model for alkylbenzenes, based on the recent TraPPE-EH force field35, using information from natural population analysis (NPA) of density functional theory (DFT) calculations. NPA is a method for deriving charges and orbital populations of atoms in a molecular system, based on a natural bond orbital analysis38. The new charge set is obtained by scaling the calculated gas-phase charges to fit the experimental Gibbs energy of solvation of benzene and of 1,3,5- trimethylbenzene. A new rule for assigning partial atomic charges for multisubstituted alkyl-aromatic compounds is also presented. The ability of this new charge set to predict solvation energies of several other alkyl-aromatic compounds is examined. The remainder of this paper is organized as follows: in Section 2 the methodology to calculate hydration energies and thermodynamic properties of the pure components is explained together with the main computational details. In Section 3.1 we test the ability of current force field parameter sets to predict the Gibbs energy of hydration. Section 5 Physical Chemistry Chemical Physics Page 6 of 54 3.2 is dedicated to improving force field parameters for better reproducing experimental hydration energies. Finally, our main conclusions are summarized in Section 4. 2. Theoretical and Computational Details 2.1. Calculation of the Gibbs energy of hydration The Gibbs energy of hydration can be seen as the total reversible work required to transfer a solute molecule from the ideal gas phase to water, at constant pressure and temperature, representing the infinite dilution chemical potential. Free energies can be calculated for an appropriate molecular model by carrying out molecular dynamics (MD) simulations and a thermodynamic integration (TI) procedure. Based on a thermodynamic cycle, the Gibbs energy of hydration, ∆ G, is estimated from hyd simulations in which solute-solvent interactions are progressively turned off. This has normally been accomplished by carrying out two separate simulations, one in solution (where both intra- and inter-molecular contributions are turned-off) and another one in vacuum (to subtract the intra-molecular contribution)39. Details of this methodology can be found in a previous publication13. However, a new feature in the MD GROMACS 4.0 suite40 allows for these intermolecular solute-solvent interactions to be quantified by solely running a liquid phase simulation, thus avoiding the need for a vacuum step. The latter approach was adopted here. The TI algorithm, free of hysteresis, has been described in detail in previous publications41-43. Briefly, if we consider any two generic well-defined states, the total Hamiltonian of the system can be made a function of a coupling parameter, λ, and used to describe the transition between those two states – in this case, an initial fully interacting system and a final “dummy” state where all solute-solvent interactions are turned off. Considering several discrete and independent λ values, equilibrium averages can be used to evaluate derivatives of the Gibbs energy with respect to λ. One then integrates these derivatives along a continuous path connecting the initial and final states to obtain the Gibbs energy between them. It is both physically insightful and methodologically convenient to separate the Gibbs energy into a sum of two separate contributions: a non-polar Lennard-Jones (LJ) contribution (that includes the cavity formation work and favorable solute-solvent dispersion interactions) and an electrostatic (Coulombic) term. In our protocol, in order 6 Page 7 of 54 Physical Chemistry Chemical Physics to avoid charge-fusion effects14,44 we have first turned off the electrostatic interactions and then the LJ interactions. The total Gibbs energy of hydration can then be estimated from: 1 ∂H (λ) Total 1 ∂H (λ) Elec 1 ∂H (λ) LJ  ∆ G =−∫ dλ=−∫ dλ+∫ dλ (1) hyd ∂λ  ∂λ ∂λ  0 λ 0 λ 0 λ  A MD simulation was performed for each discrete λ value, ranging from λ = 0 (fully interacting solute) to λ = 1 (non-interacting solute). We have used 5 λ values for the electrostatic decoupling and 15 λ points to evaluate the LJ contribution. As demonstrated in a previous work, this choice of intermediate states ensures sufficient accuracy in the free energy calculations45. The integration of the LJ Hamiltonian derivatives was carried out by fitting the data to a physically-based approximation to the cavity formation and dispersion interaction terms and then integrating the curve analytically45. The fitting function is: ∂H (λ) A A   = Aλ2 + Aλ− 2 + 2 (2) ∂λ 0 1 λ2 −Aλ+ A A   LJ 3 4 4 where A are fitting parameters45. In all cases we have employed non-linear weighted i least squares fittings of the simulation data. Correlation coefficients above 0.99 and a global RMS error of 0.3 kJ/mol were obtained. This integration procedure was shown to increase the precision of the calculated hydration energies while decreasing the number of necessary intermediate points45. Simpson’s rule was used for the integration of the electrostatic component. During the decoupling process, the electrostatic interactions were linearly interpolated between neighboring states while the LJ interactions were interpolated via soft-core interactions46, with parameters given in our previous publications13. This soft-core dependence eliminates singularities in the calculation as the LJ interactions are turned off 47. 2.2. Molecular Dynamics Simulations We have carried out MD simulations both for the calculation of the Gibbs energy of hydration, as described in the previous section, and for computing pure-fluid thermodynamic properties. All simulations were performed with version 4.0.7 of the 7 Physical Chemistry Chemical Physics Page 8 of 54 GROMACS software40 using the leap-frog Verlet integration algorithm48 with a time step of 2 fs to integrate Newton’s equations of motion. Simulations were performed using periodic boundary conditions in all directions. Covalent bonds involving hydrogen atoms were constrained using the LINCS algorithm49 while the water geometry was fixed with the SETTLE algorithm50. For efficiency reasons (see reference51 for details), the reaction-field method52, which approximates the medium beyond a cut-off distance of 1 nm by a dielectric continuum of uniform permittivity, was used to handle long-range electrostatics. The dielectric constant was adjusted to be the permittivity of each solvent or pure component. The Ewald summation method was applied in a few test cases and the resulting hydration energies were practically the same as with the reaction field method (difference of at most 0.2 kJ/mol). The remaining cut- off radii were 1 nm for the short-range neighbor list and a 0.8-0.9 nm switched cut-off for the LJ interactions. We have also applied long range dispersion corrections for energy and pressure14. In the case of free energy calculations, solvated systems consisted of one solute molecule and 500 water molecules at 298 K and 1 bar. As discussed in several previous studies, the choice of the water model may have implications in the final predicted Gibbs energy15,17,21,53. In the present work we have decided to use the Modified Extended Simple Point Charge (MSPC/E)54 model for the simulation of water. MSPC/E is an accurate force field for pure water and aqueous phase equilibria thermodynamic properties, and includes a polarization correction expected to improve hydration energy predictions. A comparison of the free energy results predicted with the MSPC/E and SPC/E55 water models was reported previously51. Nevertheless, we have also performed some tests with the TIP4P56 water model in this paper, as discussed in Section 3.2. Langevin stochastic dynamics57 was used to control the temperature, with a frictional constant of 1 ps-1 while for constant pressure runs the Berendsen barostat58, with a time constant of 0.5 ps and an isothermal compressibility of 4.5×10−5bar-1, was used to enforce pressure coupling. For each simulation, we first run an energy minimization (the Limited-memory Broyden-Fletcher-Goldfarb-Shanno algorithm59 during 5000 steps followed by a steepest descent minimization during 500 steps were used) and afterwards a constant volume equilibration (100 ps), a constant pressure equilibration (500 ps), enough to obtain complete equilibration of the box volume, and 8 Page 9 of 54 Physical Chemistry Chemical Physics finally a 5 ns NpT production stage. It is worth noticing that 5 ns is considered to be enough to observe several transitions between stable configurations, indicating sufficient sampling of torsional degrees of freedom, even for complex solutes.13 Nevertheless, this point was carefully confirmed in this work for the longer n- alkylbenzenes, where several transitions between different torsional configurations were verified. All simulations were performed in triplicate, or in the case of the larger molecules in quintuplicate, in order to obtain statistically meaningful results. For the estimation of the pure component properties we have employed the Nosé- Hoover thermostat60,61 for temperature coupling using a time constant of 1 ps, and the Parrinello-Rahman approach62 with a time constant of 2 ps to enforce pressure coupling. The reference pressure was set to 1 bar and the compressibility was set according to the fluid under evaluation (data from Cibulka and Takagi63). Each simulation box contained 512 molecules and had equilibrated box dimensions ranging from 4.19 to 5.14 nm. Liquid densities were directly obtained from the GROMACS suite using the g_energy64 tool, while heats of vaporization were estimated by taking the difference of enthalpy in the vapor and liquid phases: ∆ H = E −E +RT (3) vap g L where, E is the total energy in the gas phase and E is the total energy per mole in the g L liquid phase. Although other approaches for computing the heat of vaporization exist65, equation (3) is sufficiently accurate for our purposes. Vacuum (gas phase) simulations were conducted without cutoffs or periodic boundary conditions for 50 ns. 2.3. Density Functional Theory Calculations The gas-phase geometries of all benzene derivatives considered in this work were optimized, using very tight convergence criteria (opt=verytight and int=ultrafine keywords), with the B3LYP hybrid DFT approach66,67 as included in the Gaussian 03 package68. The 6-31+G(d,p) basis set was used to describe the electronic densities of all atoms69,70. The optimized structures were characterized as true minima on the potential energy surface by calculation of the vibrational frequencies at the same level of theory (zero imaginary frequencies). The computational approach adopted above yielded accurate structures and thermochemical data for several different compounds derived 9 Physical Chemistry Chemical Physics Page 10 of 54 from benzene71,72. The atomic charges were calculated i) from a natural population analysis73 (NPA) of the natural atomic orbitals (NAO) obtained by using the atomic blocks of the density matrix averaged over the spatial directions, e.g. p , p , p orbitals, x y z and ii) from the application of the CHelpG scheme74, based on the fitting of charges to the electrostatic potential calculated in a regularly spaced grid of points around the molecule under the constraint that the dipole moment is preserved. All charges thus calculated are provided in the Supporting Information (Tables S5 to S12). 3. Results and Discussion 3.1. Prediction of ∆∆∆∆ G using current force field parameter sets hyd The Gibbs energy of hydration of benzene (BZ) and several mono-substituted alkylbenzenes, namely methylbenzene (MB), ethylbenzene (EB), propylbenzene (PB), butylbenzene (BB) and pentylbenzene (PeB), were initially estimated using the five parameter sets already mentioned: TraPPE-UA/EH, Gromos-UA/EH and OPLS-AA. Gibbs energy of hydration results using these force fields are presented in Tables S1-S4 of the Supporting Information and illustrated in Figure 1, while absolute average deviations (AAD) are summarized in Table 1. From these results one may generally conclude that with the exception of Gromos-EH version 53A6, which was specifically reparameterized to fit hydration energies, none of the current force field parameter sets are able to accurately predict the Gibbs energy of hydration of our molecular test set. Moreover, it is clear that OPLS-AA, in spite of its increased number of degrees of freedom arising from the all-atom description, is not able to accurately describe the Gibbs energy of hydration. Although it provides a reasonable prediction for BZ, it significantly overestimates the increase in free energy with the degree of substitution. Similar results were obtained for the hydration of alkanes7 suggesting that the shortcoming may arise from a deficiency in the parameterization of the LJ component in this force field. Both of the UA force fields yield Gibbs energies of hydration that are too positive by a large amount (AAD of 5.9 and 9.4 kJ/mol for Gromos-UA and TraPPE-UA, respectively). More precisely, all predictions for the Gibbs energies of hydration with UA models are positive, while the experimental values are all negative. This 10

Description:
about the prediction of the Gibbs energy of hydration of alkyl-aromatic (16) Mobley, D. L.; Bayly, C. I.; Cooper, M. D.; Shirts, M. R.; Dill, K. A. J.
See more

The list of books you might like

Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.