Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2008 Jun 16.
Published in final edited form as: J Chem Phys. 2007 Mar 28;126(12):124114. doi: 10.1063/1.2714528

Polarizable Atomic Multipole Solutes in a Poisson-Boltzmann Continuum

Michael J Schnieders 1, Nathan A Baker 2, Pengyu Ren 3, Jay W Ponder 2,*
PMCID: PMC2430168  NIHMSID: NIHMS40640  PMID: 17411115

Abstract

Modeling the change in the electrostatics of organic molecules upon moving from vacuum into solvent, due to polarization, has long been an interesting problem. In vacuum, experimental values for the dipole moments and polarizabilities of small, rigid molecules are known to high accuracy; however, it has generally been difficult to determine these quantities for a polar molecule in water. A theoretical approach introduced by Onsager used vacuum properties of small molecules, including polarizability, dipole moment and size, to predict experimentally known permittivities of neat liquids via the Poisson equation. Since this important advance in understanding the condensed phase, a large number of computational methods have been developed to study solutes embedded in a continuum via numerical solutions to the Poisson-Boltzmann equation (PBE). Only recently have the classical force fields used for studying biomolecules begun to include explicit polarization in their functional forms. Here we describe the theory underlying a newly developed Polarizable Multipole Poisson-Boltzmann (PMPB) continuum electrostatics model, which builds on the Atomic Multipole Optimized Energetics for Biomolecular Applications (AMOEBA) force field. As an application of the PMPB methodology, results are presented for several small folded proteins studied by molecular dynamics in explicit water as well as embedded in the PMPB continuum. The dipole moment of each protein increased on average by a factor of 1.27 in explicit water and 1.26 in continuum solvent. The essentially identical electrostatic response in both models suggests that PMPB electrostatics offers an efficient alternative to sampling explicit solvent molecules for a variety of interesting applications, including binding energies, conformational analysis, and pKa prediction. Introduction of 150 mM salt lowered the electrostatic solvation energy between 2–13 kcal/mole, depending on the formal charge of the protein, but had only a small influence on dipole moments.

I. INTRODUCTION

Modeling the change in the electrostatic moments of organic molecules upon moving from vacuum to solvent has a long history, with an important initial contribution from Onsager, who in 1936 identified the difference between the cavity field and reaction field.1 The approach used was to treat the solvent as a high dielectric continuum surrounding a spherical, low dielectric solute with a dipole moment, which was considered to be a sum of permanent and induced contributions. Using the vacuum dipole moment, molecular polarizability, and an estimate of molecular size, a prediction of the experimentally observable liquid permittivity was achieved for a range of molecules. Through the use of computers, this approach has been extended in order to treat solutes with arbitrary geometry and charge distributions by numerically solving the Poisson-Boltzmann equation using finite-difference, finite element and boundary element methods.2 An advantage of using a continuum solvent over explicit representation of solvent molecules is alleviation of the need to sample over water degrees of freedom in order to determine the mean solvent response. Applications that have benefited from using continuum solvent approaches include predictions of pKas, redox potentials, binding energies, molecular design, and conformational preferences.

An accurate implicit solvation model does not necessitate any loss of solute thermodynamic information.3 For most implicit solvent models, the total solvation energy Δ Wsolvation as a function of solute conformation is separated into contributions due to nonpolar Δ Wnp and electrostatic Δ Welec effects

ΔWsolvation=ΔWnp+ΔWelec. (1)

Here we concentrate on the electrostatic contribution, motivated by recent work on improving the accuracy of force field electrostatic models through the incorporation of polarizable multipoles.47 This requires revisiting the theory underlying the electrostatic component of implicit solvent models, including those based on solving the linearized Poisson-Boltzmann equation (LPBE)

[ε(r)Φ(r)]κ¯2(r)Φ(r)=4πρ(r), (2)

where the coefficients are a function of position r, Φ(r) is the potential, ε(r) the permittivity, κ̄2(r) the modified Debye-Hückel screening factor and ρ(r) is the solute charge density. For an introduction to Poisson-Boltzmann based methodology and applications, see the review of Honig and Nichols8 or those of Baker.2,9

In spite of the fact that the dipole moment of a polar solute can increase by 30% or more during transfer from gas to aqueous phase, empirical potentials have typically neglected explicit treatment of polarization for reasons of computational efficiency. However, ab initio implicit solvent models, including the Polarizable Continuum Model (PCM) introduced in 1981 by Miertus, Scrocco and Tomasi10,11, the Conductor-Like Screening Model (COSMO) of Klamt12, the distributed multipole approach of Rinaldi et al.13 and the SMx series of models introduced in the early 90s by Cramer and Truhlar14 have long incorporated self-consistent reaction fields (SCRF).10,15 These models allow the solute, described using a range of ab initio or semi-empirical levels of theory, and continuum solvent to relax self-consistently based on their mutual interaction. The principle advantage of the present PMPB model over these existing formulations is computational savings resulting from a purely classical representation of the solute Hamiltonian, which facilitates the study of large systems. However, each of these models demonstrates that there continues to be broad interest in coupling highly accurate solute potentials with a continuum treatment of the environment.

For example, there is growing evidence that current goals of computational protein design, including incorporation of catalytic activity and protein-protein recognition, may require a more accurate description of electrostatics than has been achieved by fixed partial charge force fields used in conjunction with implicit solvents.16 For example, it has been shown that both Poisson-Boltzmann and generalized Born models used with the CHARMM2217 potential tend to favor burial of polar residues over non-polar ones.18 It is important to note that this behavior may not be directly due to the treatment of solvation electrostatics, but could result from inaccuracies in the underlying protein force field or the nonpolar component of the solvation model. However, the fixed charge nature of traditional protein potentials may also contribute to such discrepancies. Polar residues elicit a solvent field that increases their dipole moment via polarization relative to a nonpolar environment, which has a favorable energetic consequence that cannot be captured by a fixed charge force field even when using explicit water.

Along with the AMOEBA force field, other efforts toward developing polarizable force fields for biomolecular modeling are also under way, for example see the review of Ponder and Case.7 Here we comment more thoroughly on the Polarizable Force Field (PFF) of Maple and coworkers,19 since it has recently been incorporated into a continuum environment. Specifically, there are a number of salient differences between AMOEBA and PFF, including facets of the underlying polarization model and the use of permanent quadrupoles in AMOEBA. Significantly, AMOEBA allows mutual polarization between atoms with 1–2, 1–3 and 1–4 bonding arrangements, which is crucial for reproducing molecular polarizabilities.

The current work addresses a number of issues raised in the description of the PFF solvation model. First, discretization procedures previously reported for mapping partial charges onto a source grid are inadequate for higher order moments. Instead, we present a multipole discretization procedure using B-splines that leads to essentially exact energy gradients. Furthermore, we provide a rigorous demonstration of the numerical precision of our approach, similar in spirit to the work of Im et al. with respect to partial charge models.20 We also show that divergence of the polarization energy is not possible due to use in AMOEBA of Thole-style damping at short range.21

Formulation of consistent energies and gradients based on the LPBE has been reported previously for partial charge force fields.20,2224 Gilson et al. pointed out limitations in previous approaches using variational differentiation of an electrostatic free energy density functional.23 Later Im et al. showed that it is possible to begin the derivation based upon the underlying finite-difference calculation used to solve the LPBE.20 This approach leads to a formulation with optimal numerical consistency and will be adopted here.

II. BACKGROUND

In order to describe the total electrostatic energy of the PMPB model, specifically of an AMOEBA solute in a LPBE continuum solvent, it is first necessary to present the energy in vacuum using a convenient notation. We will also summarize the previously developed procedure for determining robust energies and gradients for fixed charge force fields in a LPBE continuum.20,23 Given this background, the theoretical infrastructure required to implement the PMPB model is motivated and presented in a self-contained fashion.

A. AMOEBA vacuum electrostatic energy

Following Ren and Ponder5, each permanent atomic multipole site can be considered as a vector of coefficients including charge, dipole and quadrupole components

Mi=[qi,di,x,di,y,di,z,Θi,xx,Θi,xy,Θi,xz,,Θi,zz]t, (3)

where the superscript t denotes the transpose. The interaction energy between two sites i and j separated by distance sij can then be represented in tensor notation as

U(sij)=MitTijMj=[qidi,xdi,ydi,zΘi,xx]t[1xjyjzjxi2xixj2xiyj2xizjyi2yixj2yiyj2yizjzi2zixj2ziyj2zizj]1sij[qjdj,xdj,ydj,zΘj,xx]. (4)

Each site may also be polarizable, such that an induced dipole μi proportional to the strength of the local field is present

μi=αiEi=αi(jiTd,ij(1)Mj+kiTik(1)μk). (5)

Here αi is an isotropic atomic polarizability and Ei is the total field, which can be decomposed into contributions from permanent multipole sites and induced dipoles, and the summations are over Ns multipole sites. Later, this expression will be modified to include the solvent reaction field. The interaction tensors Td,ij(1) and Tik(11) are, respectively,

Td,ij(1)=[xj2xixj2xiyj2xizjyi2yixj2yiyj2yizjzi2zixj2ziyj2zizj]1sij, (6)

and

Tik(11)=[2xixk2xiyk2xizk2yixk2yiyk2yizk2zixk2ziyk2zizk]1sik. (7)

where the subscript d in Td,ij(1) indicates that masking rules for the AMOEBA group-based polarization model are applied.47 This linear system of equations can be solved via a number of approaches, including direct matrix inversion or iterative schemes such as successive over-relaxation (SOR). Note at short range the field is damped via the Thole model, which is not included above for clarity and is discussed elsewhere.5 The total vacuum electrostatic energy Uelecv includes pairwise permanent multipole interactions and many-body polarization

Uelecv=12[MtT(μv)tTp(1)]M, (8)

where the factor of one-half avoids double-counting of permanent multipole interactions in the first term and accounts for the cost of polarizing the system in the second term. Furthermore, M is a column vector of 13Ns multipole components

M=[M1M2MNs], (9)

T is a Ns × Ns supermatrix with Tij as the off-diagonal elements

T=[0T12T13T210T23T31T320], (10)

μv is a 3Ns column vector of converged induced dipole components in vacuum

μv=[μ1,xμ1,yμ1,zμNs,z], (11)

and Tp(1) is a 3Ns × 13Ns supermatrix with Tp,ij(1) as off-diagonal elements

Tp(1)=[0Tp,12(1)Tp,13(1)Tp,21(1)0Tp,23(1)Tp,31(1)Tp,32(1)0]. (12)

The subscript p denotes a tensor matrix that operates on the permanent multipoles to produce the electric field in which the polarization energy is evaluated, while the subscript d was used above to specify an analogous tensor matrix that produces the field that induces dipoles. The differences between the two are masking rules that scale short-range through-bond interactions in the former case and use the AMOEBA group-based polarization scheme for the later.4

B. LPBE energies and gradients for fixed partial charge force fields

LPBE solvation energies and gradients have been determined previously by a number of groups for fixed partial charge force fields.22,23 We will briefly restate the results of Im et al. to introduce the approach extended here for use with the PMPB model.20 The solvation free energy Δ G of a permanent charge distribution is

ΔGelec=12(ΦstΦvt)q, (13)

where q is a column vector of fractional charges, Φst is the transpose of a column vector containing the electrostatic potential of the solvated system and Φvt is the corresponding vacuum potential. The number of components in each vector is equal to the number of grid points used to represent the system. In the present work the grid will always be cubic, and as discussed in the section on multipole discretization must use equal grid spacing in each dimension, although the number of grid points along each axis can vary. The potential can be determined numerically using a finite-difference representation of the LPBE20,25,26, which can be formally defined as a linear system of equations

AΦ=4πq, (14)

where A is a symmetric matrix that represents the linear operator (differential and linear term). An equivalent, but more cumbersome representation that makes clear the underlying finite-difference formalism is given in the Appendix. Solving Eq. (14) for the potential

Φ=4πA1q, (15)

highlights that A−1 is the Green’s function with dimensions Nr × Nr, where Nr is the number of grid points. By defining the Green’s function for the solvated As1 and Av1 homogeneous cases, the electrostatic hydration free energy is

ΔGelec=12(4πqt)(As1Av1)q (16)

The derivative with respect to movement of the γ coordinate of atom j is

ΔGelecsj,γ=2π[qtsj,γ(As1Av1)q+qt(As1Av1)sj,γq+qt(As1Av1)qsj,γ]. (17)

This expression can be simplified by noting that the derivative of the homogeneous Green’s function is zero everywhere

Av1sj,γ=0, (18)

and the derivative of the solvated Green’s function can be substituted for using a standard relationship of matrix algebra.

As1As=IAs1sj,γAs+As1Assj,γ=0As1sj,γ=As1Assj,γAs1 (19)

Finally, due to the symmetry of the Green’s function, the first and third terms of Eq. (17) are equivalent

qtsj,γ(As1Av1)q=qt(As1Av1)qsj,γ. (20)

Using the relationships in Eqs. (1820), Eq. (17) becomes

ΔGelecsj,γ=4πqt(As1Av1)qsj,γ+18π(4πqtAs1)A2sj,γ(4πAs1q). (21)

Finally, the two products of Green’s functions with source charges in the second term on the right-hand side can be replaced by the resulting potentials using Eq. (15) to give

ΔGelecsj,γ=(ΦsΦv)tqsj,γ+18πΦstAssj,γΦs. (22)

In the limit of infinitesimal grid spacing and infinite grid size, it was shown that this solution is equivalent to the forces derived by Gilson et al.20,23 To generalize this result for permanent atomic multipoles, the derivative ∂q/∂sj,γ remains to be defined and is discussed below. Additionally, all moments except the monopole are subject to torques, which are equivalent to forces on the local multipole frame defining sites.

III. METHODOLOGICAL DEVELOPMENTS

Based on the AMOEBA electrostatic energy in vacuum and previous work in obtaining the energy and gradients for a solute represented by a fixed partial charge force field described above20, we now derive the formulation needed to describe the electrostatic solvation energy within the PMPB model. First, we consider the steps necessary to express the LPBE on a grid, including discretization of the source multipoles or induced dipoles, assignment of the permittivity, assignment of the modified Debye-Hückel screening factor and estimation of the potential at the grid boundary. A variety of techniques are available to solve the algebraic system of equations that result, although this work uses an efficient multigrid approach implemented in PMG26 and used via the APBS software package.27 Second, given the electrostatic potential solution to the LPBE, we describe how to determine the electrostatic solvation energy and its gradient. In fact, at least four LPBE solutions are required to determine the PMPB electrostatic solvation energy, and at least six to determine energy gradients. For comparison, fixed charged models typically require at most two LPBE solutions, the vacuum and solvated states, as outlined in the previous background section, although formulations that eliminate the self-energy exist.28 The reasons and implications for the increased number of solutions of the LPBE required for the PMPB model will be discussed below.

A. Atomic multipoles as the source charge density

An important first step to expressing the LPBE in finite-difference form is discretization of an ideal point multipole onto the source charge grid. We begin by recalling that an ideal multipole arises from a Taylor series expansion of the potential in vacuum at a location R due to n charges near the origin, each with a magnitude and position denoted by ci and ri, respectively.29

V(R)=i=1nciRri (23)

In performing the expansion, the convention for repeated summation over subscripts is utilized and we truncate after second order

V(R)=i=1nci[1Rri,αα1R+12ri,αri,βαβ1R], (24)

where the α and β subscripts each denote an x-, y- or z-component of a position vector or differentiation with respect to that coordinate. Based on this expansion, the monopole q, dipole d, and traceless quadrupole Θ moments are defined as

q=i=1ncidα=i=1nri,αciΘαβ=i=1n32ri,αri,βci12ri2δαβ. (25)

There are various ways to define the quadrupole moment because only five quadrupole components are independent. This particular formulation ensures that the diagonal components are traceless, which simplifies many formulae because summations over the trace vanish.29 Substitution into the potential gives

V(R)=q(1R)dαα(1R)+13Θαβαβ(1R). (26)

The reverse operation, representation of an ideal multipole by partial charges at grid sites (or charge density over finite volumes), is degenerate. However, some necessary properties reduce the space of practical solutions. These include local support (region of non-zero values on the grid) and smooth derivatives for the change in charge magnitude due to movement of a multipole site with respect to the grid. For fixed partial charges a normalized cubic basis spline, or B-spline, has been used successfully for discretizing monopole charge distributions (delta functions) on finite difference grids. For quadrupole moments at least 4th order continuity is required such that a normalized 5th order B-spline N5 (x), which is a piecewise polynomial (Fig. 1), is appropriate.30

Figure 1.

Figure 1

Normalized 5th order B-spline on the interval [0, 5].

N5(x)={124x4,0x118+16x+14(x1)2+16(x1)316(x1)4,1x21324+12x14(x2)212(x2)314(x2)4,2x3472412x14(x3)212(x3)316(x3)43x4172416x14(x4)216(x4)3+124(x4)4,4x50,otherwise (27)

The sum of this function evaluated at any five evenly spaced points between 0 and 5 is unity. The second part of the Appendix gives a rigorous demonstration that B-splines satisfy the properties of the delta functional and therefore can be used to implement a gradient operator.

To illustrate this approach, the fraction of charge that a grid point with coordinates ri will receive from a charge site with coordinates sj will now be described. We use r to denote an Nr × 3 matrix containing all grid coordinates, while s is an Ns × 3 matrix containing the coordinates of multipole sites. Elements of both matrices will be specified using two subscripts, the first is an index and the second is a dimension. We first consider the x-dimension, which requires the relative distance of nearby y-z planes from the charge site in dimensionless grid units (ri,xsj,x)/h, where h is the grid spacing. The B-spline domain is centered over the charge site by shifting its domain from [0,5] to [−2.5,2.5] by adding 2.5 to its argument. Therefore, the weights of the 5 closest y-z planes to the charge site will be nonzero and sum to 1, where each weight is given by

W(ri,x,sj,x)=N5(ri,xsj,xh+2.5). (28)

If the charge site is located on a y-z grid plane, then the maximum of the B-spline will be assigned to that plane. Repeated partitioning in the y- and z-dimensions leads to a tensor product description of the charge density

B(ri,sj)=W(ri,x,sj,x)W(ri,y,sj,y)W(ri,z,sj,z). (29)

A further useful property of nth order B-splines is that their derivative can be formulated as a linear combination of n-1 order B-splines.30

Nn(x)x=Nn1(x)Nn1(x1) (30)

For example, the first derivative of the normalized 5th order B-spline can be constructed from two of 4th order, suggesting a dipole basis (or gradient stencil) for determining the electric field from the potential grid as shown in Fig. 2.

Figure 2.

Figure 2

The sum of two 4th order B-splines (dashed) equal the first derivative of a normalized 5th order B-spline (solid).

N5(x)x=N4(x)N4(x1) (31)

Similarly, the 2nd derivative can be constructed from a linear combination of 3rd order B-splines, suggesting an axial quadrupole basis as well as an axial 2nd potential gradient stencil as in Fig. 3.

Figure 3.

Figure 3

The sum of three 3rd order B-splines (dashed) equal the second derivative of a normalized 5th order B-spline (solid).

2N5(x)x2=N3(x)2N3(x1)+N3(x2) (32)

For notational convenience we define a matrix B with dimension Nr × Ns which is used to convert a collection of Ns permanent atomic multipole sites into grid charge density over the Nr grid points

B=[B(r1,s1)B(r1,sNs)B(rNr,s1)B(rNr,sNs)]. (33)

Only 125 entries per column will have non-zero coefficients, due to each multipole being partitioned locally among 53 grid points.

Given the matrix B, the charge density at all grid points due to the permanent multipoles of an AMOEBA solute is

qM=Bq1hαBdα+13h2αβBΘαβ, (34)

where q, dx, dy, dz, Θxx, Θxy, Θxz, …, Θzz are column vectors and h normalizes for grid size. This is, in effect, the inverse operation to the original Taylor expansion by which the multipole moments were defined from a collection of point charges. All atomic multipole moments are exactly conserved to numerical precision as long as equal grid spacing is used in each dimension. While most finite difference methods can be generalized to non-uniform Cartesian meshes26, the nature of traceless multipoles requires uniform Cartesian mesh discretizations. If unequal grid spacing is used, then the trace will be nonzero due to inconsistent coupling between the axial quadrupole components.

The gradient of the charge density at grid sites with respect to an atomic coordinate can be written

qMsj,γ=Bsj,γq1hαBsj,γdα+13h2αβBsj,γΘαβ. (35)

A more compact notation is needed for derivations presented below, and can be achieved by defining a matrix TB of size Nr × 13Ns

TB=[B(r1,s1)1hB(r1,s1)s1,x13h22B(r1,s1)s1,z213h22B(r1,sNs)sNs,z2B(rNr,s1)1hB(rNr,s1)s1,x13h22B(rNr,s1)s1,z213h22B(rNr,sNs)sNs,z2]. (36)

The matrix product Φt TB, where Θ is a column vector of length Nr containing the potential from a numerical solution to the LPBE, produces the same tensor components (i.e., the potential, field and field gradient) as Mt T in Eq. (8) for the AMOEBA vacuum electrostatic energy. Using this notation allows manipulation of reaction potentials and intramolecular potentials to be handled on equal footing. The same approach is appropriate for induced dipoles.

There is a trade-off between higher order B-splines and the goal of maintaining the smallest possible support for the multipoles. As the support grow s, the charge density is less representative of the ideal multipole limit. Additionally, placement of solute charge density outside the low-dielectric cavity should be avoided. This restriction of charge density to the solute interior places an upper bound on acceptable grid spacings for use with finite difference discretizations of higher-order B-splines. If, for example, the solute cavity for a hydrogen atom ends approximately 1.2 Å from its center, then the maximum recommended grid spacing when using 5th order B-splines is 0.48 Å (1.2/2.5), whereas for third order B-splines a value of 0.80 Å is reasonable (1.2/1.5). Therefore, the use of quintic B-splines requires a smaller upper bound on grid spacing than cubic B-splines.

B. Permittivity and modified Debye-Hückel screening factor

The permittivity ε(r) and modified Debye-Hückel screening factor κ̄2(r) functions are defined through a characteristic function H(ri,s), where ri represents the coordinates of a grid point and s the coordinates of all multipole sites. Inside the solute cavity the characteristic function is 0, while in the solvent it is 1. For the homogeneous calculation the permittivity is set to unity over all space, while for the solvated state it takes the value 1 inside the solute, εs in solvent, and intermediate values over a transition region

ε(ri)=1+(εs1)H(ri,s). (37)

The modified Debye-Hückel screening factor is zero everywhere for the vacuum calculation and for the solvated calculation is defined by

κ¯2(ri)=κ¯b2H(ri,s), (38)

where κ¯b2=εsκb2 is the modified bulk screening factor and is related to the ionic strength and I=12iqi2ci via κb2=8πI/εskBT. Here qi ci are the charge and number concentration of mobile ion species i, respectively, kB is the Boltzmann constant and T the absolute temperature. The characteristic function itself can be formulated as the product of a radially symmetric function applied to each solute atom

H(ri,s)=j=1,NsHj(risj) (39)

where Hj(r) must allow for a smooth transition across the solute-solvent boundary to achieve stable Cartesian energy gradients.20,31 This is a result of terms that depend on the gradient of the characteristic function with respect to an atomic displacement. A successful approach to defining a differentiable boundary is the use of a polynomial switch Sn(r) of order n, although other definitions have been suggested based on atom centered Gaussians.32 For any atom j, Hj(r) takes the form

Hj(r)={0,rbSn(r,b,e)b<r<e1,er, (40)

where b is the beginning of the transition between solute and solvent and e the end of the transition. For the permittivity, the switch begins and ends at

b=σjwe=σjw, (41)

where σj is the radius of atom j and w indicates how far the smoothing window extends radially inward and outward. For the modified Debye-Hückel screening factor the radius of the largest ionic species σion is also taken into account

b=σj+σionwe=σj+σionw. (42)

For fixed partial charge force fields, a cubic switch S3 has been used with success. However, a characteristic function with higher order continuity has been found to improve energy conservation at a given grid spacing. Table V reports representative examples of this effect. By using a 7th order polynomial switch S7,

TABLE V.

The norm of the gradient sum over all atoms (kcal/mole/Å) for three different solutes is shown for cubic, quintic and heptic characteristic functions at two grid spacings. A norm of zero, indicating perfect conservation of energy, is nearly achieved for acetamide and ethanol at 0.11 Å grid spacing using a heptic characteristic function. Conservation of energy is improved by reducing grid spacing and also by increasing the continuity of the solute-solvent boundary via the characteristic function.

Acetamide
Ethanol
CRN
Grid Spacing 0.21 0.11 0.22 0.11 0.32 0.18
Cubic 2.48 0.46 0.77 0.29 18.82 2.58
Quintic 1.93 0.23 0.32 0.15 9.38 2.42
Heptic 0.88 0.06 0.21 0.06 6.27 1.26
S7(r,b,e)=(c7r7+c6r6+c5r5+c4r4+c3r3+c2r2+c1r+c0)b7+7b6e21b5e2+35b4e335b3e4+21b2e57be6+e7,c0=b4(b3+7b2e21be2+35e3),c1=140b3e3,c2=210b2e2(b+e),c3=140be(b2+3be+e2),c4=35(b3+9b2+9be2+e3),c5=84(b2+3be+e2),c6=70(b+e),c7=20 (43)

the first three derivatives of the characteristic function can be constrained to zero at the beginning and end of the switching region. The cubic, quintic and heptic volume exclusions functions are shown in Fig. 4.

Figure 4.

Figure 4

Comparison of cubic, quintic and heptic characteristic functions for an atom with radius 3 Å using a total window width of 0.6 Å.

C. Boundary conditions

Single Debye-Hückel (SDH) and multiple Debye-Hückel (MDH) boundary conditions for a solute are two common approximations to the true potential used to specify Dirichlet boundary conditions for non-spherical solutes described by a collection of atomic multipoles.2 SDH assumes that all atomic multipole sites are collected into a single multipole at the center of the solute, which is approximated by a sphere. MDH assumes the superposition of the contribution of each atomic multipole considered in the absence of all other sites that displace solvent. Therefore, to construct the Dirichlet problem for a solute described by an arbitrary number of atomic multipole sites, the potential outside a solvated multipole located at the center of a sphere is required.

For solvent described by the LPBE

2Φ(r)=κb2Φ(r), (44)

a solution was first formulated by Kirkwood.33 This form of the LPBE is simplified relative to Eq. (2) since there is no fixed charge distribution and no spatial variation in either the permittivity or Debye-Hückel screening factor. Inside the cavity, the Poisson equation is obeyed

2Φ(r)=4πρ(r)ε, (45)

where the charge density ρ(r) may contain moments of arbitrary order. The boundary conditions are enforced by requiring the potential and dielectric displacement to be continuous across the interface between the solute and solvent via

Φ(r)in=Φ(r)out, and (46)
εΦ(r)inr=εoutΦ(r)outr, (47)

respectively. Additional requirements on the solution are that it be bounded at the origin and approach an arbitrary constant at infinity, usually chosen to be zero.

In presenting the solution, it is convenient for our purposes to continue using Cartesian multipoles, rather than switching to spherical harmonics, although there is a well-known equivalence between the two approaches.34 The potential at r due to a symmetric, traceless multipole in a homogeneous dielectric ε is

Φε(rij)=(Tε)tMj=([1xyz132xx](1εrij))t[qjμx,jμy,jμz,jΘxx,j], (48)

where rij = risj might be the difference between a grid location and a multipole site. The potential inside the spherical cavity is the superposition of the homogeneous potential and the reaction potential

Φin(rij)=[(I+Rin)Tε]tMj, (49)

where I is the identity matrix and Rin is a diagonal matrix with diagonal elements

[cin(0),cin(1),cin(1),cin(1),cin(2),] (50)

that are based on coefficients for multipoles of order n to be determined by the boundary conditions

cin(n)=βn(rija)2n+1. (51)

Similarly, the potential outside the cavity is

Φout(rij)=(RoutTε)tMj, (52)

where Rout is a diagonal matrix with diagonal elements

[cout(0),cout(1),cout(1),cout(1),cout(2),] (53)

based on a second set of coefficients for multipoles of order n also determined by the boundary conditions

cout(n)=εεoutκrijαnkn(κrij)(rija)n. (54)

where kn(x) is the modified spherical Bessel function of the third kind

kn(x)=πex2xi=0n(n+i)!i!(ni)!(2x)i. (55)

Kirkwood first solved Eqs. (44) and (45) subject to the boundary conditions in Eqs. (46) and (47) to determine αn and βn as

αn=(2n+1)/κankn(κa)/ε^κakn(κa) (56)

and

βn=(n+1)kn(κa)/ε^+κakn(κa)kn(κa)/ε^κakn(κa), (57)

where kn(x) is the derivative of kn(x) and ε̂ is the ratio of the permittivity in solvent to that inside the sphere εout/ε.33 We only require the potential outside the cavity to construct SDH and MDH boundary conditions and therefore we provide specific values of αn and kn through quadrupole order as shown in Table I. As the ionic strength goes to zero, the Laplace equation is obeyed in solvent. For multipoles through quadrupole order, the difference between the LPBE and Laplace potentials outside the cavity are summarized in Table II.

TABLE I.

Explicit values for the functions αn(x)and kn(x) up to quadrupole order.

n αn(x)/(2exp(x)π) kn(x)/(π2exp(x))
0 1(1+x) 1x
1 3ε^x1+x+ε^(2+2x+x2) 1+xx2
2 5ε^x22(3+3x+x2)+ε^(9+9x+4x2+x3) (3+3x+x2)x3

TABLE II.

Explicit values of the coefficients used to calculate the potential at the grid boundary of LPBE and PE calculations, respectively, under the SDH or MDH approximation. The LPBE coefficients reduce to the PE coefficients as salt concentration goes to zero.

n κrαn(κa)kn(κr)(ra)n limκ0[κrαn(κa)kn(κr)(ra)n]
0 exp(κ(ar))1+κa 1
1 3ε^exp(κ(ar))(1+κr)1+κa+ε^(2+2κα+(κa)2) 3ε^1+2ε^
2 5ε^exp(κ(ar))(3+3κr+(κr)2)2(3+3κa+(κa)2)+ε^(9+9κa+4(κa)2+(κa)3) 5ε^2+3ε^

D. Permanent atomic multipole solvation energy and gradient

The PMPB permanent atomic multipole (PAM) solvation energy and gradient are very similar to those for fixed partial charge force fields. Based on Eqs. (15) and (34) the PAM vacuum, solvated and reaction potentials are, respectively,

ΦvM=4πAv1qMΦsM=4πAs1qMΦM=ΦsMΦvM. (58)

The expression for the permanent electrostatic solvation energy is then identical to that for a fixed partial charge force field given in Eqs. (13) and (16), except the source charge density is based on PAM via qM

ΔGM=12(ΦM)tqM=12(4πqM)t(As1Av1)qM. (59)

Derivation of the energy gradient is identical to Eqs. (1722) and yields

ΔGMsj,γ=(ΦM)tqMsj,γ+18π(ΦsM)Assj,γΦsM. (60)

There are, however, some important differences between achieving smooth gradients for a fixed partial charge force field and one based on PAM. First, as discussed in the section on multipole discretization, quadrupoles require at least 5th order B-splines to guarantee continuous derivatives of the source charge density with respect to movement of a multipole site

qMsj,γ=TBsj,γM. (61)

Second, we have found that if a third order polynomial is used to define the transition between solute and solvent for purposes of assigning the permittivity and the modified Debye-Hückel screening factor, energy conservation is achieved only for very fine grid spacing. As discussed earlier, use of a 7th order polynomial improves energy conservation for coarser grids. Details of the numerical realization of Eq. (60), including torques, are presented in part 3 of the Appendix.

E. Self-consistent reaction field

An SCRF protocol is used to achieve numerical convergence of the coupling between a polarizable solute and continuum solvent. The starting point of the iterative convergence is the total “direct” field Ed at each polarizable site. This is defined by the sum of the PAM intramolecular field

Ed=Td(1)M, (62)

where Td(1) is analogous to the tensor matrix used in deriving the AMOEBA vacuum energy in Eq. (12), and the PAM reaction field

ERFM=DBtΦM, (63)

where DB is a matrix of B-spline derivatives of size Nr by 3Ns

DB=1h[B(r1,s1)s1,xB(r1,s1)s1,yB(r1,s1)s1,zB(r1,sNs)sNs,zB(rNr,s1)s1,xB(rNr,s1)s1,yB(rNr,s1)s1,zB(rNr,sNs)sNs,z], (64)

that produces the reaction field at induced dipoles sites given a potential grid. The induced dipoles are determined as the product of the direct field Ed with a vector of isotropic atomic polarizabilities α:

μd=αEd=α(Td(1)MDBtΦM). (65)

We define the direct model of polarization to consist of induced dipoles not acted upon by each other or their own reaction field. Although this is a nontrivial approximation, the direct model requires little more work to compute energies than a fixed partial charge force field since the limiting factor in both cases is two numerical LPBE solutions. Energy gradients under the direct model require three pairs of LPBE solutions and are therefore a factor of 3 more expensive than for a fixed charge solute. The direct model is expected to be quite useful for many applications. For example, a geometry optimization might utilize the direct polarization model initially, then switch to the more expensive mutual polarization model described below as the minimum is approached.

In contrast to the direct model, the total solvated field E has two additional contributions due to the induced dipoles and their reaction field,

E=Td(1)M+T(11)μDBt(ΦM+Φμ), (66)

for a sum of 4 contributions. The procedure for determining the vacuum, solvated and reaction potential, respectively, due to the induced dipoles is identical to that of the PAM

Φvμ=4πAv1qμΦsμ=4πAs1qμΦμ=ΦsμΦvμ, (67)

except the source charge density is

qμ=DBμ. (68)

The induced dipoles

μ=α[Td(1)M+T(11)μDBt(ΦM+Φμ)], (69)

can be determined in an iterative fashion using successive over-relaxation (SOR) to accelerate convergence.35 The SCRF is usually deemed to have converged when the change in the induced dipoles is less than 10−2 RMS debye between steps. This generally requires 4–5 cycles and therefore the mutual polarization model necessitates 8–10 additional numerical solutions of the LPBE to determine the PMPB solvation energy. Although calculation of the direct polarization energy is no more expensive than that for fixed multipoles, mutual polarization energies that depend on SCRF convergence are approximately a factor of 5 more costly.

F. PMPB electrostatic solvation energy

Having described the PAM solvation energy and gradients and our approach for determining the induced dipoles, it is now possible to discuss the total solvated electrostatic energy for the PMPB model,

Uelec=12[MtTμtTp(1)+ΦtTB]M, (70)

where Φ is the LPBE reaction potential for the converged solute charge distribution

Φ=ΦM+Φμ. (71)

The total electrostatic energy in solvent is similar to the vacuum electrostatic energy of Eq. (8), with an important difference. The vacuum induced dipoles μv change in the presence of a continuum solvent by an amount represented by μΔ, such that the SCRF induced moments μ can be decomposed into a sum

μ=μv+μΔ. (72)

The change in the potential, field, etc. within the solute is not only a result of the solvent response, but also due to changes in intramolecular polarization. By definition, the electrostatic solvation energy ΔGelec is the change in total electrostatic energy due to moving from vacuum to solvent

ΔGelec=UelecUelecv=12[(μΔ)tTp(1)+μtTB]M. (73)

In practice, it is convenient to compute the total solvated electrostatic energy Uelec and vacuum electrostatic energy Uelecv using the SCRF μ and vacuum μv induced dipole moments, respectively. The electrostatic solvation energy ΔGelec is then determined as the difference.

G. Polarization energy and gradient

As described previously, the induced dipoles are determined using an iterative SOR procedure until a predetermined convergence criterion is achieved. Since this is a linear system, it is possible to solve for the induced dipoles directly, which facilitates derivation of the polarization energy gradient with respect to atomic displacements. Substitution of the reaction potential due to the induced dipoles μν from Eq. (67) into the expression for the induced dipoles in Eq. (69) makes clear all dependencies on μ.

μ=α[Td(1)MDBtΦM+T(11)μ4πDBt(As1Av1)DB]. (74)

Collecting all terms containing the induced dipoles on the left hand side gives

[α1T(11)+4πDBt(As1Av1)DB]μ=Td(1)MDBtΦM. (75)

For convenience, a matrix C is defined as

C=[α1T(11)+4πDBt(As1Av1)DB], (76)

which is substituted into Eq. (75) to show the induced dipoles are a linear function of the PAM M

μ=C1(Td(1)MDBtΦM)=C1(Ed+ERFM). (77)

The first term results from the intramolecular interaction tensor Td(1) that implicitly contains the AMOEBA group based polarization scheme, and the second term is the permanent reaction field.

The polarization energy can now be described in terms of the permanent reaction field and permanent intramolecular solute field Ep

Uμ=12(Ep+ERFM)tμ. (78)

To find the polarization energy gradient, we wish to avoid terms that rely on the change in induced dipoles with respect to atomic displacement. Therefore, the induced dipoles in Eq. (78) are replaced using Eq. (77) to yield

Uμ=12(Ep+ERFM)tC1(Ed+ERFM). (79)

By the chain rule, the polarization energy gradient is

Uμsj,γ=12[(Epsj,γ+ERFMsj,γ)tC1(Ed+ERFM)+(Ep+ERFM)tC1sj,γ(Ed+ERFM)+(Ep+ERFM)tC1(Edsj,γ+ERFMsj,γ)] (80)

For convenience a mathematical quantity _ is defined as

ν=(Ep+ERFM)C1, (81)

which is similar to μ. We can now greatly simplify Eq. (80) using Eqs. (77) and (81) along with the identity C1sj,γ=C1C1sj,γC1 to give

Uμsj,γ=12[(Epsj,γ)tμ+νtEdsj,γ+(ERFMsj,γ)μ+νtERFMsj,γνtCsj,γμ]. (82)

G II. Direct polarization gradient

Under the direct polarization model, C is an identity matrix whose derivative is zero, and therefore Eq. (82) simplifies to

Uμdsj,γ=12[(Epsj,γ)tμ+νtEdsj,γ+(ERFMsj,γ)μ+νtERFMsj,γ]. (83)

The first two terms on the RHS appear in the polarization energy gradient even in the absence of a continuum reaction field and are described elsewhere.4 The third and fourth terms are specific to LPBE calculations and will now be discussed. The derivative of the LPBE reaction field due to permanent multipoles with respect to movement of any atom has a similar form to the analogous derivative of the potential. Substitution for the field using Eq. (63) into the third term of Eq. (83) gives

(ERFMsj,γ)tμ=DBtΦMsj,γμ. (84)

Substitution of the permanent multipole potential from Eq. (58) into Eq. (84) yields

(ERFMsj,γ)tμ=4πsj,γ[DBt(As1Av1)TBM]μ, (85)

which is differentiated by applying the chain rule

(ERFMsj,γ)tμ=4π[DBtsj,γ(As1Av1)TBM+DBtAs1sj,γTBM+DBt(As1Av1)TBsj,γM]μ. (86)

The same simplifications described in Eqs. (18-22) are applied to Eq. (86), except that in this case the first and third terms are not equivalent and cannot be combined

(ERFMsj,γ)tμ=(Φμ)tTBsj,γM14π(Φsμ)Assj,γΦsMμtDBtsj,γΦM. (87)

The fourth term on the RHS of Eq. (83) leads to a result analogous to Eq. (87) using similar arguments

νtERFMsj,γ=(Φν)tTBsj,γM14π(Φsμ)Assj,γΦsMνtDBtsj,γΦM. (88)

G III. Mutual polarization gradient

In addition to the implicit difference due to the induced dipoles being converged self-consistently, the full mutual polarization gradient includes an additional contribution beyond the direct polarization gradient. Specifically, the derivative of the matrix C leads to four terms

νtCsj,γμ=νt[T(11)sj,γ+4πDBtsj,γ(As1Av1)DB+4πDBtAs1sj,γDB+4πDBt(As1Av1)DBsj,γ]μ. (89)

The first term on the RHS occurs in vacuum and is described elsewhere4, while the final three terms are specific to LPBE calculations. Using the simplifications described in Eqs. (1822) results in

νtCsj,γμ=νtT(11)sj,γμνtDBtsj,γΦμ14π(Φsν)tAssj,γΦsμ(Φν)tDBsj,γμ. (90)

Substitution of Eqs. (87), (88) and (90) into Eq. (82) gives the total mutual polarization energy gradient for an AMOEBA solute interacting self-consistently with the PMPB continuum.

Uμsj,γ=12[(Epsj,γ)tμ+νtEdsj,γ]12νtT(11)sj,γμ+12[(Φμ)tTBsj,γM+μtDBtsj,γΦM]+18π(Φsμ)tAssj,γΦsM+12[(Φν)tTBsj,γM+νtDBtsj,γΦM]+18π(Φsν)tAssj,γΦsM+12[(Φν)tDBsj,γμ+νtDBtsj,γΦμ]+18π(Φsν)tAssj,γΦsμ (91)

The first two terms (where each set of square brackets will be considered a single term) are evaluated even in the absence of continuum solvent, although in this case μ and ν have been converged in a self-consistent field that includes continuum contributions. The remaining terms are analogous to those found in Poisson-Boltzmann calculations involving only permanent electrostatics. The number of LPBE calculations required for evaluation of the energy gradient includes two for the permanent multipoles and two each for μ and ν at each SOR convergence step. Further details on the numerical implementation of Eq. (91) can be found in the third section of the Appendix.

IV. VALIDATION AND APPLICATION

This section presents useful benchmarks for demonstrating the expected numerical precision of the present work. Our first goal is to compare against analytical results for a source charge distribution described by a single charge, dipole, polarizable dipole or quadrupole located at the center of a low dielectric sphere in high dielectric solvent. The transition between solute and solvent is initially specified using a step function, and then subsequently using a smooth transition described by a heptic polynomial. We then compare analytic gradients to those determined using finite-differences of the energy for a variety of two sphere systems to isolate the reaction field, dielectric boundary and ionic boundary gradients for the permanent multipole solvation energy, the direct polarization model and the mutual polarization model. Finally, the method is applied to a series of proteins and comparisons are made to corresponding simulations in explicit water.

A. Energy

The numerical accuracy of the multipole discretization procedure was studied by comparison to analytical solutions of the Poisson equation for a monopole, dipole, polarizable dipole and quadrupole located within a spherical cavity of radius 3.0 Å. The monopole case, or Born ion36, has a well-known analytical solution

Uq=12(1ε1)q2a (92)

where q is the charge magnitude, a the cavity radius and ε is the solvent dielectric. For a permanent dipole, the analogous solution was first determined by Onsager1,

Ud=12(2(ε1)1+2ε)dda3, (93)

where d is the dipole vector. For a polarizable dipole, the energy is the sum of two contributions, the cost of polarization and the energy of the total dipole in the total reaction field37

Uα,d=12fdd1fα, (94)

where α is the polarizability, d is the permanent dipole and f is the reaction field factor

f=1a3[2(ε1)1+2ε]. (95)

The analytic solution for the self-energy of a traceless Cartesian quadrupole can be derived beginning from the energy of a quadrupole in an electric field gradient

UΘγδΘγδ3δγΦ, (96)

where we are summing over the subscripts γ and δ, and the factor of 1/3 is due to use of traceless quadrupoles.29 To determine the needed reaction field gradient, which for the moment will be assumed to come from any quadrupole component and not necessarily be a self-interaction, we begin from the reaction potential inside the cavity

ΦΘαβ=Θαβ3[3(ε1)2+3ε]3αβa5 (97)

and take the first derivative

γΦΘαβ=Θαβ3[3(ε1)2+3ε]3(βδαγ+αδβγ)a5, (98)

followed by a second differentiation to achieve the reaction field gradient

δγΦΘαβ=Θαβ3[3(ε1)2+3ε]3(δαγδβδ+δβγδαδ)a5. (99)

Substituting Eq. (98) into Eq. (95) and taking into account that half the energy is lost due to polarizing the continuum gives the self-energy of a traceless quadrupole in its own reaction field gradient

UΘαβ,Θγδ=12[3(ε1)2+3ε]Θαβ(δαγδβδ+δβγδαδ)Θγδ3a5. (100)

From Eq. (100), it is seen that all non-self interactions, for example Θxx with Θyy, are zero, and the quadrupole self-energy is simply the sum of nine terms,

UΘαβ=12[3(ε1)2+3ε]2Θαβ23a5. (101)

The first series of numerical tests used a step function at the dielectric boundary rather than the smooth transition that is required for continuous energy gradients described previously. This simplification is necessary in order to compare the known analytic results directly with the numerical solver. In each case, the solution domain was a 10.0 Å cube with the low-dielectric sphere located at the center. In Table III it is shown that each test case converges toward the analytic result as grid spacing is decreased.

TABLE III.

As grid spacing decreases, the numerical solution to the PE approaches the analytic solution for four canonical test cases including a charge, dipole, polarizable dipole and quadrupole. Each test case involved a 3 Å sphere of dielectric 1 and solvent dielectric of 78.3 with a step-function transition between solute and solvent (kcal/mole).

Grid Points Grid Spacing Charge Dipole Polarizable Dipole Quadrupole
33 × 33 × 33 0.313 −55.6514 −5.3556 −5.8011 −1.8487
65 × 65 × 65 0.156 −54.9150 −5.1450 −5.5548 −1.7211
129 × 129 × 129 0.078 −54.8024 −5.1134 −5.5180 −1.7038
225 × 225 × 225 0.045 −54.7236 −5.0915 −5.4925 −1.6912
Analytic - −54.6355 −5.0675 −5.4645 −1.6783

Our next goal was to determine the energy change due to introduction of a smooth dielectric boundary with a window width of 0.6 Å. Using a grid spacing slightly less than 0.1 Å, it can be seen in Table IV that the smooth dielectric boundary increases the solvation energy over the analogous step function boundary. By increasing the radius of the low dielectric cavity by approximately 0.2 Å, the energy of the charge, dipole, polarizable dipole and quadrupole can be adjusted to simultaneously mimic the known analytic results.

TABLE IV.

The tests from Table III are repeated using 129 grid points (0.078 Å spacing), however, the transition between solute and solvent is defined by a 7th order polynomial, which acts over a total window width of 0.6 Å. Increasing the radius of the low dielectric sphere by approximately 0.2 Å raises the energies to mimic the step function transition results (kcal/mole).

Radius Increase Charge Dipole Polarizable Dipole Quadrupole
0.0 −58.4926 −6.2126 −6.8202 −2.3555
0.1 −56.4762 −5.5922 −6.0798 −1.9767
0.2 −54.5941 −5.0518 −5.4463 −1.6687
Step Function −54.8024 −5.1134 −5.5180 −1.7038

B. Energy gradient

Our first goal is to show that the energy gradient is continuous for higher order moments as a result of using 5th order B-splines. This is seen in Figs. (5) through (7) for a charge, dipole and quadrupole interacting with a neutral cavity, respectively. It is also clear that the sum of the forces between the neutral and charged site (and a third reference site that defines the local multipole frame in the cases of the dipole and quadrupole) is zero, indicating conservation of energy. Similarly, the reaction field and dielectric boundary gradients of the polarization energy for both the direct and mutual models are smooth and demonstrate conservation of energy, as shown in Figs. (8) and (9), respectively. Finally, it is clear that polarization catastrophes are avoided even when a charged site is moved toward superimposition with a polarizable site, due to use of a modified Thole model that damps mutual polarization at short range.21

C. The electrostatic response of solvated proteins

As described in the introduction, a motivation for the current work is study of polar macromolecules by an improved electrostatic model within an empirical molecular mechanics framework. From explicit water simulations it is possible to measure the total dipole moment of a solvated protein in a fixed folded conformation by sampling over the water degrees of freedom. The resulting ensemble average electrostatic response can then be directly compared to the PMPB model.

Simulations of five proteins taken from the Protein Databank38 (1CRN39, 1ENH40, 1FSV41, 1PGB42 and 1VII43) were equilibrated under NPT conditions (1 atm, 298 K) using a standard protocol. Formal charge and system size are given in Table VI. A single snapshot for each protein system was taken from equilibrated molecular dynamics simulations using the AMOEBA force field. The protein coordinates were frozen, and sampling of the solvent degrees of freedom continued for 150 ps under the same NPT conditions, with the first 50 ps discarded prior to analysis. For all simulations the Berendsen weak coupling thermostat and barostat were employed with time constants of 0.1 and 2.0 ps, respectively.44 Long range electrostatics were treated using particle mesh Ewald (PME) summation with a cutoff for real space interactions of 7.0 Å and an Ewald coefficient of 0.54 Å−1.45 The PME methodology used tinfoil boundary conditions, a 54 × 54 × 54 charge grid and 6th order B-spline interpolation. van der Waals interactions were smoothly truncated to zero at 12.0 Å using a switching window of width 1.2 Å. Simulations were run using TINKER version 4.2.46

TABLE VI.

Synopsis of the protein systems studied in explicit and continuum solvent.

Number of Atoms
Protein Formal Charge Protein Protein +Water
CRN 0 642 4980
ENH +7 947 5039
FSV +5 504 6435
PGB −4 855 6143
VII +2 596 4271

The same conformation of each protein studied in explicit water was examined using the LPBE methodology developed in this work at a range of grid spacings using the direct and mutual polarization models. In addition, 150 mM electrolyte was used in conjunction with the mutual polarization model to determine the relative effect of salt on the electrostatic response. The results are summarized in Table VII. Similar to the analytic test cases, as grid spacing is reduced the total electrostatic energy rises monotonically toward the converged solution. The total dipole moments are less sensitive to grid spacing than are the energies, with little change observed in moving from 0.3 Å to 0.2 Å. Adding 150 mM salt lowers the electrostatic energy by 1.7–12.5 kcal/mole at the smallest grid spacing studied, with CRN (neutral) and ENH (+7) showing the smallest and largest response, respectively. The magnitude of the energetic change indicates that salt concentration plays an important role in protein energetics, especially for highly charged species. For these calculations we have chosen an ionic radius of 2.0 Å, however smaller or larger values increase or decrease the energetic response, respectively.

TABLE VII.

The energy (kcal/mole) and dipole moment (debye) of each protein system was studied using a range of grid spacings under the direct polarization model, mutual polarization model, and mutual polarization model with 150 mM salt. The cavity was defined using AMOEBA Rmin values for each atom and smooth dielectric and ionic boundaries via a total window width of 0.6 Å.

Direct Polarization
Mutual Polarization
150 mM Salt
Protein Grid Spacing Energy μ Energy μ Energy μ
CRN 0.61 −597.4 83.9 −679.1 81.0 −680.9 81.6
0.31 −563.3 83.4 −641.3 80.6 −643.0 81.1
0.18 −554.7 83.4 −632.1 80.6 −633.8 81.1
ENH 0.63 −1892.6 265.2 −2055.1 265.1 −2067.5 266.8
0.32 −1851.1 265.8 −2008.8 265.8 −2021.2 267.4
0.18 −1834.9 265.7 −1991.1 265.7 −2003.6 267.3
FSV 0.66 −1207.0 208.1 −1293.8 215.7 −1301.0 216.3
0.33 −1184.3 208.4 −1269.3 216.0 −1276.5 216.6
0.19 −1173.1 208.4 −1257.1 215.9 −1264.3 216.5
PGB 0.71 −1327.7 128.4 −1453.5 132.7 −1458.9 133.4
0.36 −1275.7 127.8 −1400.5 132.0 −1405.9 132.6
0.20 −1259.3 127.7 −1380.3 131.9 −1385.7 132.5
VII 0.62 −902.4 194.2 −1009.8 197.1 −1014.3 198.1
0.31 −866.0 194.4 −970.6 197.3 −975.0 198.3
0.18 −858.3 194.3 −962.0 197.2 −966.5 198.2

Finally, we compare the increase in dipole moment between the explicit water simulations and the continuum LPBE environment for each protein. As shown in Table VIII, both the direct and mutual models lead to total moments that are in good agreement with those found by molecular dynamics sampling of explicit water degrees of freedom. On average, the dipole moment increased by a factor of 1.27 in explicit water and 1.26 using the mutual polarization model. This result, which was achieved without detailed parameterization of atomic radii (AMOEBA Buffered-14-7 Rmin values were used), indicates that at the length scale of whole proteins the continuum assumption is justified. Timings and memory requirements for the LPBE calculations as a function of grid size are shown in Table IX.

TABLE VIII.

The dipole moment (debye) of each protein in vacuum mv, under the direct and mutual polarization models interacting with a continuum of permittivity 78.3, and in explicit water. Ensemble averages were taken over 100 psec trajectories and each has a std. err. of less than ± 0.3. The ratio of the solvated to vacuum dipole moment is given in each case. The cavity was defined using AMOEBA Rmin values for each atom and smooth dielectric and ionic boundaries via a total window width of 0.6 Å.

Vacuum
Direct Polarization
Mutual Polarization
Explicit Water
Protein μv μ μ/μv μ μ/μv < μ > < μ>/μv
CRN 62.1 83.4 1.34 80.6 1.30 81.8 1.32
ENH 208.3 265.7 1.28 265.7 1.28 267.0 1.28
FSV 184.7 208.4 1.13 215.9 1.17 213.5 1.16
PGB 101.4 127.7 1.26 131.9 1.30 134.3 1.32
VII 158.3 194.4 1.23 197.3 1.25 197.7 1.25

Average 143.0 175.9 1.25 178.3 1.26 178.9 1.27

TABLE IX.

Memory requirements and wall clock timings for each protein system are shown. All calculations were run on a 2.4 Ghz Opteron.

Protein Cubic Box Size Grid Points Memory (MB) Direct (s) Mutual (s)
CRN 39.31 65 84 6.9 50.2
129 487 34.5 276.8
225 2027 189.3 1414.7
ENH 40.50 65 81 9.2 80.3
129 491 45.5 414.7
225 1983 253.0 2457.3
FSV 42.35 65 73 5.7 42.7
129 487 31.9 234.6
225 2040 188.1 1463.8
PGB 45.48 65 80 8.2 60.7
129 375 30.6 230.9
225 1880 156.9 1176.5
VII 39.47 65 73 6.4 65.2
129 487 37.2 360.6
225 2040 194.8 2062.8

V. CONCLUSION

We have presented methodology required to determine the energy and gradient for the AMOEBA force field in conjunction with numerical solutions to the LPBE, which captures the electrostatic response of solvent by treating it as a dielectric continuum. The PMPB model was then applied to a series of proteins that were also studied using explicit water simulations. The resulting increases in dipole moment found using each approach were in excellent agreement. This indicates that the continuum assumption is a reasonable approximation at the length scale of the systems studied here.

Future work will include detailed parameterization of the solute cavity to reproduce small molecule electrostatic solvation energies. All of the methodology described in this paper is implemented in the TINKER molecular modeling package. This will facilitate use of the PMPB model using a variety of algorithms including molecular dynamics, Monte Carlo and a range of optimization methods. Additionally, parallelization of the LPBE calculations using existing approaches in APBS, which have been applied to fixed partial charge models, would be an important improvement in terms of speed and increasing the size of systems that can be routinely studied.

The methodology presented here is also expected to be useful for the development of continuum electrostatics models for coarse grain potentials. For example, Golubkov and Ren have recently described a generalized coarse grain model based on point multipoles and Gay-Berne potentials, which saves several orders of magnitude over all atom models.47 We have also begun to use the PMPB model as a standard of accuracy in the testing of a polarizable analytic approximation similar in spirit to the generalized Born model, which was recently reviewed by Feig and Brooks.48

Figure 5.

Figure 5

Analytic and finite-difference gradients for a neutral cavity fixed at the origin and a sphere with unit positive charge vs. separation. Both spheres have a radius of 3.0 Å and the solvent dielectric is 78.3. The gradient of the neutral cavity is due entirely to the dielectric boundary force and cancels exactly the force on the charged sphere.

Figure 6.

Figure 6

Analytic and finite-difference gradients for a neutral cavity fixed at the origin and a sphere with dipole moment components of (2.54, 2.54, 2.54) debye vs. separation. Both spheres have a radius of 3.0 Å and movement of the dipole is along the x-axis. The gradient of the neutral cavity is due entirely to the dielectric boundary force and cancels exactly the sum of the forces on the dipole and a third site (that has no charge density or dielectric properties) that defines the local coordinate system of the dipole.

Figure 7.

Figure 7

Analytic and finite-difference gradients for a neutral cavity fixed at the origin and a sphere with quadrupole moment components of (5.38, 2.69, 2.69, 2.69, −2.69, 2.69, 2.69, 2.69, −2.69) Buckinghams vs. separation. Both spheres have a radius of 3.0 Å and movement of the quadrupole is along the x-axis. The gradient of the neutral cavity cancels exactly the sum of the forces on the quadrupole and a third site (that has no charge density or dielectric properties) that defines the local coordinate system of the quadrupole.

Figure 8.

Figure 8

Analytic and finite-difference gradients for a neutral, polarizable cavity fixed at the origin and a sphere with unit positive charge vs. separation using the direct polarization model. Both spheres have a radius of 3.0 Å. The gradient can be seen to approach zero at a number of points, notably when the spheres are separated by approximately 1.5 Å leading to a maximum in the reaction field produced by the charge at the polarizable site, and again when the spheres are superimposed and the reaction field is zero at the polarizable site.

Figure 9.

Figure 9

Analytic and finite-difference gradients for a neutral, polarizable cavity fixed at the origin and a polarizable sphere with unit positive charge vs. separation using the mutual polarization model. Both spheres have a radius of 3.0 Å and a polarizability of 1.0 Å −3. Note that the mutual polarization gradients are smaller than those of Figure 8 for the otherwise equivalent direct polarization model.

Figure 10.

Figure 10

The dielectric of the solvent and test spheres are both set to 1 in this case, while a salt concentration of 150 mM is used to isolate the ionic boundary gradients. Analytic and finite-difference gradients for a neutral, polarizable cavity fixed at the origin (3.0 Å radius) and a polarizable sphere with a unit positive charge (1.0 Å radius) vs. separation using the mutual polarization model. Both spheres have a polarizability of 1.0 Å −3, and the ionic radius is set to zero Å.

Acknowledgments

We thank Sergio Urahata for making available to us snapshots from explicit water simulations of the proteins systems examined here; and Todd Dolinsky and Dave Gohara for helping to port portions of the PMPB model into APBS. MJS was partially supported by a Computational Biology Training Grant from the NIH. NAB was supported by NIH grant GM069702. JWP was supported by NSF grant CHE-0535675 and NIH grant GM069553.

APPENDIX A: FINITE-DIFFERENCE REPRESENTATION OF THE LPBE

The finite-difference representation of the LPBE for uniform grid spacing is 20,25,26

εx(i,j,k)[Φ(i+1,j,k)Φ(i,j,k)]+εx(i1,j,k)[Φ(i1,j,k)Φ(i,j,k)]+εy(i,j,k)[Φ(i,j+1,k)Φ(i,j,k)]+εy(i,j1,k)[Φ(i,j1,k)Φ(i,j,k)]+εz(i,j,k)[Φ(i,j,k+1)Φ(i,j,k)]+εz(i,j,k1)[Φ(i,j,k1)Φ(i,j,k)],+κ¯2(i,j,k)Φ(i,j,k)h2=4πq(i,j,k)h (A1)

where h is the grid spacing, Φ (i, j, k) is the electrostatic potential, κ̄2 (i, j, k) is the modified Debye-Hückel screening factor and q (i, j, k) is the fractional charge. The permittivity is specified by three separate arrays, εx, εy and εz, where each is shifted along its respective grid branch such that εx(i, j, k) represents the location (xi + h/2, yi, zi) for the grid point ( xi, yi, zi). Eq. (A1) is the basis for formulating the LPBE as a linear system of equations, which are represented compactly by Eq. (14).

APPENDIX B: DEFINITION OF THE DELTA FUNCTIONAL USING B-SPLINES

The delta functional δ is defined by

δ(xa)dx=1 (A2)

and δ(xa) = 0 for xa. An approximate discrete 1-dimensional realization of this definition (approximate because the width is not infinitesimally small) is

i=15W(xi,a)=1, (A3)

where the function W has been defined in Eq. (28) via 5th order B-splines and {x1,…,x5} are the 5 closest grid points to a. In the limit of infinitesimal grid spacing, the properties of the Delta functional are met exactly by expressing Eq. (A3) above as a continuous integral

εεW(x,a)dx=1, (A4)

where ε > 0. The value at a of any function known to be defined over the grid can then be determined as

εεW(x,a)f(x)dx=f(a), (A5)

and the negative of its gradient as

εε(W(x,a))f(x)dx=W(x,a)f(x)εεεεW(x,a)f(x)dx=f(a). (A6)

Further differentiations can be found in an analogous fashion, limited only by the continuity of the B-spline.

APPENDIX C: IMPLEMENTATION OF THE PERMANENT, DIRECT POLARIZATION AND MUTUAL POLARIZATION FORCES

After solving the linear system, the permanent electrostatic solvation forces are determined via Eq. (60), which in the limit of infinitesimal grid spacing becomes 20

Fi,γ=ΔGMsi,γ=V[ΦMρMsI,γ+18πΦsM(εssi,γΦsM)18π(ΦsM)2κ¯2si,γ]d3r, (A7)

where γ represents differentiation with respect to either the x-, y- or z-coordinate of atom i. The three terms on the RHS of Eq. (A7) are usually referred to as the reaction field force, dielectric boundary force and ionic boundary force, respectively. We briefly review the implementation of these forces, in order to develop the foundation necessary to discuss additional details of realizing the polarization forces.

C I. Permanent reaction field force and torque

The γ -component of the “Permanent Reaction Field Force” FiPerm RF for atom i is

Fi,γPerm RF=si,γ[qiBidi,ααBi+Θi,αβ3αβBi]ΦM, (A8)

where Bi is a single column of the B-spline matrix in Eq. (33), di,α is the α -component of the permanent dipole, Φi,αβ is the αβ component of the quadrupole and the convention for summation over the α and β subscripts is implied. There is also an associated “Permanent Reaction Field Torque” τiPerm RF, whose x-component is

τi,xPerm q-Phi=di,yERF,i,zMdi,zERF,i,yM23[Θi,yααERF,i,zMΘi,zααERF,i,yM], (A9)

where ERF,i,αM is the α -component of the permanent multipole reaction field. The y- and z- components are analogous, and we note that all torques are equivalent to forces on neighboring atoms that define the local frame of the multipole.

C II. Direct polarization reaction field force and torque

Similarly, the third term of the polarization gradient given in Eq. (91) results in a “Direct Polarization Reaction Field Force”

Fi,γDirect RF=12si,γ[(μi,ααBi)ΦM+(qiBidi,ααBi+Θi,αβ3αβBi)Φμ], (A10)

while the fifth term we label the “Non-Local Direct Polarization Reaction Field Force”

Fi,γNLDirect RF=12si,γ[(νi,ααBi)ΦM+(qiBidi,ααBi+Θi,αβ3αβBi)Φν], (A11)

respectively. The label “non-local” is used to denote that the term σ results from omitting or scaling the contribution to the intramolecular field of permanent multipoles that are in a 1-5 connected or closer, as opposed to the induced dipoles μ that result from the AMOEBA group based polarization scheme. Additionally, the x-component of the torques, τiDirect RF and τiNL-Direct RF, on the permanent moments due to the continuum reaction field of μ and ν are, respectively,

τi,xDirect RF=12[di,yERF,i,zμdσ,zERF,i,yμ23(Θi,yααERF,i,zμΘi,zααERF,i,yμ)], (A12)

and

τi,xNLDirect RF=12[di,yERF,i,zνdσ,zERF,i,yν23(Θi,yααERF,i,zνΘi,zααERF,i,yν)]. (A13)

C III. Mutual polarization reaction field force

The last reaction field force results from the seventh term of Eq. (91) and is due to mutual polarization

Fi,γMutual RF=12si,γ(μi,ααBiΦν+νi,ααBiΦμ). (A14)

C IV. Permanent dielectric boundary force

The second term in Eq. (A7), the “Permanent Dielectric Boundary Force”, is determined from Eq. (A1) as

Fi,γPerm DB=h8i,j,kΦsM(i,j,k){εx(i,j,k)ri,γ[ΦsM(i+1,j,k)ΦsM(i,j,k)]+εx(i1,j,k)ri,γ[ΦsM(i1,j,k)ΦsM(i,j,k)]+εy(i,j,k)ri,γ[ΦsM(i,j+1,k)ΦsM(i,j,k)]+εy(i,j1,k)ri,γ[ΦsM(i,j1,k)ΦsM(i,j,k)]+εz(i,j,k)ri,γ[ΦsM(i,j,k+1)ΦsM(i,j,k)]+εz(i,j,k1)ri,γ[ΦsM(i,j,k1)ΦsM(i,j,k)]}, (A15)

where the partial derivatives of the permittivity depend on Eqs. (37), (39) and the heptic characteristic function presented in Eqs. (4043).

εx(i,j,k)ri,γ=(1εs)Hx(i,j,k)Hxi(i,j,k)Hx(i,j,k)ri,γ=[εx(i,j,k)1]Hxi(i,j,k)Hxi(i,j,k)ri,γri, (A16)

where Hx and Hxi are the characteristic function of the solute and atom i for the x-branch of the cubic grid at (xi + h/2, yj, zk), respectively, and the vector ri is the distance from the atomic center to the grid point.

C V. Direct and mutual polarization dielectric boundary forces

The fourth, sixth and eighth terms in Eq. (91) result in dielectric boundary force components. For example, the “Direct Polarization Dielectric Boundary Force” is

Fi,γDirect DB=h8i,j,kΦsμ(i,j,k){εx(i,j,k)ri,γ[ΦsM(i+1,j,k)ΦsM(i,j,k)]+εx(i1,j,k)ri,γ[ΦsM(i1,j,k)ΦsM(i,j,k)]+εy(i,j,k)ri,γ[ΦsM(i,j+1,k)ΦsM(i,j,k)]+εy(i,j1,k)ri,γ[ΦsM(i,j1,k)ΦsM(i,j,k)]+εz(i,j,k)ri,γ[ΦsM(i,j,k+1)ΦsM(i,j,k)]+εz(i,j,k1)ri,γ[ΦsM(i,j,k1)ΦsM(i,j,k)]} (A17)

or

Fi,γDirect DB=h8i,j,kΦsM(i,j,k){εx(i,j,k)ri,γ[Φsμ(i+1,j,k)Φsμ(i,j,k)]+εx(i1,j,k)ri,γ[Φsμ(i1,j,k)Φsμ(i,j,k)]+εy(i,j,k)ri,γ[Φsμ(i,j+1,k)Φsμ(i,j,k)]+εy(i,j1,k)ri,γ[Φsμ(i,j1,k)Φsμ(i,j,k)]+εz(i,j,k)ri,γ[Φsμ(i,j,k+1)Φsμ(i,j,k)]+εz(i,j,k1)ri,γ[Φsμ(i,j,k1)Φsμ(i,j,k)]} (A18)

where the superscript on the solvated potentials have been exchanged between Eq. (A17) and Eq. (A18). In other words, both Eqs. (A17) and (A18) are equivalent to numerical precision and either may be implemented. Analogous expressions for the sixth and eighth terms in Eq. (91) are referred to as the “Non-Local Direct Polarization Dielectric Boundary Force” and the “Mutual Polarization Dielectric Boundary Force”, respectively.

C VI. Permanent ionic boundary force

The last term in Eq. (A7), the “Permanent Ionic Boundary Force”, is determined from Eq. (A1) as

Fi,γPerm IB=h38πi,j,kΦsM(i,j,k)2κ¯2(i,j,k)ri,γ (A19)

using

κ¯2(i,j,k)ri,γ=κ¯b2H(i,j,k)Hi(i,j,k)Hi(i,j,k)ri,γ=κ¯2(i,j,k)Hi(i,j,k)Hi(i,j,k)ri,γri (A20)

where H and Hi are the characteristic function of the solute and atom i, respectively, and the vector ri is the distance from the atomic center to the grid point.

C VII. Direct and mutual polarization ionic boundary force

The fourth, sixth and eighth terms in Eq. (91) result in ionic boundary force components. For example, the “Direct Polarization Ionic Boundary Force” is

Fi,γDirect IB=h38πi,j,kΦsM(i,j,k)Φsμ(i,j,k)κ¯2(i,j,k)ri,γ (A21)

Analogous expressions for the sixth and eighth terms in Eq. (91) are termed the “Non-Local Direct Polarization Ionic Boundary Force” and the “Mutual Polarization Ionic Boundary Force”, respectively. Note the difference between Eqs. (A19) and (A21); specifically the potential is squared in Eq. (A19), but is asymmetric in Eq. (A21).

References

  • 1.Onsager L. J Am Chem Soc. 1936;58:1486. [Google Scholar]
  • 2.Baker NA. Method Enzymol. 2004;383:94. doi: 10.1016/S0076-6879(04)83005-2. [DOI] [PubMed] [Google Scholar]
  • 3.Roux B, Simonson T. Biophys Chem. 1999;78:1. doi: 10.1016/s0301-4622(98)00226-9. [DOI] [PubMed] [Google Scholar]
  • 4.Ren P, Ponder JW. J Comput Chem. 2002;23:1497. doi: 10.1002/jcc.10127. [DOI] [PubMed] [Google Scholar]
  • 5.Ren P, Ponder JW. J Phys Chem B. 2003;107:5933. [Google Scholar]
  • 6.Ren P, Ponder JW. J Phys Chem B. 2004;108:13427. [Google Scholar]
  • 7.Ponder JW, Case DA. Adv Prot Chem. 2003;66:27. doi: 10.1016/s0065-3233(03)66002-x. [DOI] [PubMed] [Google Scholar]
  • 8.Honig B, Nicholls A. Science. 1995;268:1144. doi: 10.1126/science.7761829. [DOI] [PubMed] [Google Scholar]
  • 9.Baker NA. Curr Opin Struc Biol. 2005;15:137. doi: 10.1016/j.sbi.2005.02.001. [DOI] [PubMed] [Google Scholar]
  • 10.Tomasi J. Theor Chem Acc. 2004;112:184. [Google Scholar]
  • 11.Cances E, Mennucci B, Tomasi J. J Chem Phys. 1997;107:3032. [Google Scholar]; Cances E, Mennucci B. J Math Chem. 1998;23:309. [Google Scholar]; Cammi R, Tomasi J. J Comput Chem. 1995;16:1449. [Google Scholar]; Mierts S, Scrocco E, Tomasi J. Chem Phys. 1981;55:117. [Google Scholar]
  • 12.Klamt A, Schuurmann G. J Chem Soc Perk T. 1993;2:799. [Google Scholar]; Dolney DM, Hawkins GD, Winget P, Liotard DA, Cramer CJ, Truhlar DG. J Comput Chem. 2000;21:340. [Google Scholar]; Klamt A. J Phys Chem. 1995;99:2224. [Google Scholar]; Baldridge K, Klamt A. J Chem Phys. 1997;106:6622. [Google Scholar]
  • 13.Rinaldi D, Bouchy A, Rivail JL, Dillet V. J Chem Phys. 2004;120:2343. doi: 10.1063/1.1635355. [DOI] [PubMed] [Google Scholar]; Rinaldi D, Bouchy A, Rivail JL. Theor Chem Acc. 2006;116:664. [Google Scholar]
  • 14.Hawkins GD, Cramer CJ, Truhlar DG. Chem Phys Lett. 1995;246:122. [Google Scholar]; Hawkins GD, Cramer CJ, Truhlar DG. J Phys Chem. 1996;100:19824. [Google Scholar]; Kelly CP, Cramer CJ, Truhlar DG. J Chem Theory Comput. 2005;1:1133. doi: 10.1021/ct050164b. [DOI] [PubMed] [Google Scholar]; Kelly CP, Cramer CJ, Truhlar DG. J Phys Chem B. 2006;110:16066. doi: 10.1021/jp063552y. [DOI] [PubMed] [Google Scholar]; Li JB, Zhu TH, Hawkins GD, Winget P, Liotard DA, Cramer CJ, Truhlar DG. Theor Chem Acc. 1999;103:9. [Google Scholar]; Li JB, Hawkins GD, Cramer CJ, Truhlar DG. Chem Phys Lett. 1998;288:293. [Google Scholar]; Thompson JD, Cramer CJ, Truhlar DG. J Phys Chem A. 2004;108:6532. [Google Scholar]; Thompson JD, Cramer CJ, Truhlar DG. Theor Chem Acc. 2005;113:107. [Google Scholar]; Zhu TH, Li JB, Hawkins GD, Cramer CJ, Truhlar DG. J Chem Phys. 1998;109:9117. [Google Scholar]; Hawkins GD, Cramer CJ, Truhlar DG. J Phys Chem B. 1998;102:3257. [Google Scholar]; Giesen DJ, Hawkins GD, Liotard DA, Cramer CJ, Truhlar DG. Theor Chem Acc. 1997;98:85. [Google Scholar]; Chambers CC, Hawkins GD, Cramer CJ, Truhlar DG. J Phys Chem. 1996;100:16385. [Google Scholar]; Giesen DJ, Cramer CJ, Truhlar DG. J Phys Chem. 1995;99:7137. [Google Scholar]; Cramer CJ, Truhlar DG. J Comput Aid Mol Des. 1992;6:629. doi: 10.1007/BF00126219. [DOI] [PubMed] [Google Scholar]; Cramer CJ, Truhlar DG. Science. 1992;256:213. doi: 10.1126/science.256.5054.213. [DOI] [PubMed] [Google Scholar]; Cramer CJ, Truhlar DG. J Am Chem Soc. 1991;113:8305. [Google Scholar]
  • 15.Cramer CJ, Truhlar DG. Chem Rev. 1999;99:2161. doi: 10.1021/cr960149m. [DOI] [PubMed] [Google Scholar]
  • 16.Vizcarra CL, Mayo SL. Curr Opin Chem Biol. 2005;9:622. doi: 10.1016/j.cbpa.2005.10.014. [DOI] [PubMed] [Google Scholar]
  • 17.MacKerell AD, Bashford D, Bellott M, Dunbrack RL, Evanseck JD, Field MJ, Fischer S, Gao J, Guo H, Ha S, Joseph-McCarthy D, Kuchnir L, Kuczera K, Lau FTK, Mattos C, Michnick S, Ngo T, Nguyen DT, Prodhom B, Reiher WE, Roux B, Schlenkrich M, Smith JC, Stote R, Straub J, Watanabe M, Wiorkiewicz-Kuczera J, Yin D, Karplus M. J Phys Chem B. 1998;102:3586. doi: 10.1021/jp973084f. [DOI] [PubMed] [Google Scholar]
  • 18.Jaramillo A, Wodak SJ. Biophys J. 2005;88:156. doi: 10.1529/biophysj.104.042044. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Maple JR, Cao YX, Damm WG, Halgren TA, Kaminski GA, Zhang LY, Friesner RA. J Chem Theory Comput. 2005;1:694. doi: 10.1021/ct049855i. [DOI] [PubMed] [Google Scholar]; Cortis CM, Langlois JM, Beachy MD, Friesner RA. J Chem Phys. 1996;105:5472. [Google Scholar]; Friedrichs M, Zhou RH, Edinger SR, Friesner RA. J Phys Chem B. 1999;103:3057. [Google Scholar]
  • 20.Im W, Beglov D, Roux B. Comput Phys Commun. 1998;111:59. [Google Scholar]
  • 21.Thole BT. Chem Phys. 1981;59:341. [Google Scholar]
  • 22.Davis ME, McCammon JA. J Comput Chem. 1990;11:401. [Google Scholar]; Sharp K. Biophys J. 1990;57:A10. [Google Scholar]; Niedermeier C, Schulten K. Mol Simulat. 1992;8:361. [Google Scholar]; Gilson MK. J Comput Chem. 1995;16:1081. [Google Scholar]
  • 23.Gilson MK, Davis ME, Luty BA, McCammon JA. J Phys Chem. 1993;97:3591. [Google Scholar]
  • 24.Micu AM, Bagheri B, Ilin AV, Scott LR, Pettitt BM. J Comput Phys. 1997;136:263. [Google Scholar]
  • 25.Klapper I, Hagstrom R, Fine R, Sharp K, Honig B. Proteins. 1986;1:47. doi: 10.1002/prot.340010109. [DOI] [PubMed] [Google Scholar]; Warwicker J, Watson HC. J Mol Biol. 1982;157:671. doi: 10.1016/0022-2836(82)90505-8. [DOI] [PubMed] [Google Scholar]
  • 26.Holst M, Saied F. J Comput Chem. 1993;14:105. [Google Scholar]
  • 27.Baker NA, Sept D, Joseph S, Holst MJ, McCammon JA. P Natl Acad Sci USA. 2001;98:10037. doi: 10.1073/pnas.181342398. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Rocchia W, Sridharan S, Nicholls A, Alexov E, Chiabrera A, Honig B. J Comput Chem. 2002;23:128. doi: 10.1002/jcc.1161. [DOI] [PubMed] [Google Scholar]; Zhou ZX, Payne P, Vasquez M, Kuhn N, Levitt M. J Comput Chem. 1996;17:1344. doi: 10.1002/(SICI)1096-987X(199608)17:11<1344::AID-JCC7>3.0.CO;2-M. [DOI] [PubMed] [Google Scholar]
  • 29.Stone AJ. The Theory of Intermolecular Forces. Clarendon Press; Oxford: 1996. [Google Scholar]
  • 30.de Boor C. A Practical Guide to Splines. Springer; New York: 2001. [Google Scholar]
  • 31.Gilson MK, Davis ME, Luty BA, Madura JD, McCammon JA. Biophys J. 1993;64:A354. doi: 10.1016/S0006-3495(93)81335-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Prabhu NV, Zhu PJ, Sharp KA. J Comput Chem. 2004;25:2049. doi: 10.1002/jcc.20138. [DOI] [PubMed] [Google Scholar]; Grant JA, Pickup BT, Nicholls A. J Comput Chem. 2001;22:608. [Google Scholar]
  • 33.Kirkwood JG. J Chem Phys. 1934;2:351. [Google Scholar]
  • 34.Applequist J. J Phys A-Math Gen. 1989;22:4303. [Google Scholar]; Applequist J. Theor Chem Acc. 2002;107:103. [Google Scholar]
  • 35.Young DM. Iterative Solutions of Large Linear Systems. Academic Press; New York, NY: 1971. [Google Scholar]
  • 36.Born M. Z Phys. 1920;1:45. [Google Scholar]
  • 37.Bottcher CJF. Dielectrics in Static Fields. 1993. [Google Scholar]
  • 38.Berman HM, Westbrook J, Feng Z, Gilliland G, Bhat TN, Weissig H, Shindyalov IN, Bourne PE. Nucleic Acids Res. 2000;28:235. doi: 10.1093/nar/28.1.235. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Teeter MM. P Natl Acad Sci USA. 1984;81:6014. doi: 10.1073/pnas.81.19.6014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Clarke ND, Kissinger CR, Desjarlais J, Gilliland GL, Pabo CO. Prot Sci. 1994;3:1779. doi: 10.1002/pro.5560031018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Dahiyat BI, Mayo SL. Science. 1997;278:82. doi: 10.1126/science.278.5335.82. [DOI] [PubMed] [Google Scholar]
  • 42.Gallagher T, Alexander P, Bryan P, Gilliland GL. Biochemistry. 1994;33:4721. [PubMed] [Google Scholar]
  • 43.McKnight CJ, Matsudaira PT, Kim PS. Nat Struct Biol. 1997;4:180. doi: 10.1038/nsb0397-180. [DOI] [PubMed] [Google Scholar]
  • 44.Berendsen HJC, Postma JPM, van Gunsteren WF, DiNola A, Haak JR. J Chem Phys. 1984;81:3684. [Google Scholar]
  • 45.Sagui C, Pedersen LG, Darden TA. J Chem Phys. 2004;120:73. doi: 10.1063/1.1630791. [DOI] [PubMed] [Google Scholar]
  • 46.Ponder JW. TINKER: Software Tools for Molecular Design, Version 4.2. Washington University; Saint Louis, MO: 2004. [Google Scholar]
  • 47.Golubkov PA, Ren P. J Chem Phys. 2006;125:64103. doi: 10.1063/1.2244553. [DOI] [PubMed] [Google Scholar]
  • 48.Feig M, Brooks CL., III Curr Opin Struc Biol. 2004;14:217. doi: 10.1016/j.sbi.2004.03.009. [DOI] [PubMed] [Google Scholar]

RESOURCES