Topical Review

Biomolecular electrostatics—I want your solvation (model)*

Published 30 November 2012 © 2012 IOP Publishing Ltd
, , Citation Jaydeep P Bardhan 2013 Comput. Sci. Discov. 5 013001 DOI 10.1088/1749-4699/5/1/013001

1749-4699/5/1/013001

Abstract

We review the mathematical and computational foundations for implicit-solvent models in theoretical chemistry and molecular biophysics. These models are valuable theoretical tools for studying the influence of a solvent, often water or an aqueous electrolyte, on a molecular solute such as a protein. Detailed chemical and physical aspects of implicit-solvent models have been addressed in numerous exhaustive reviews, as have numerical algorithms for simulating the most popular models. This work highlights several important conceptual developments, focusing on selected works that spotlight the need for research at the intersections between chemical, biological, mathematical, and computational physics. To introduce the field to computational scientists, we begin by describing the basic theoretical ideas of implicit-solvent models and numerical implementations. We then address practical and philosophical challenges in parameterization, and major advances that speed up calculations (covering continuum theories based on Poisson as well as faster approximate theories such as generalized Born). We briefly describe the main shortcomings of existing models, and survey promising developments that deliver improved realism in a computationally tractable way, i.e. without increasing simulation time significantly. The review concludes with a discussion of ongoing modeling challenges and relevant trends in high-performance computing and computational science.

Export citation and abstract BibTeX RIS

1. Introduction

Biology happens in solution—mostly water, with tightly regulated concentrations of ions such as sodium and chloride. These solvent molecules strongly influence the structure and function of molecules, affecting protein folding and stability, as well as the affinity and specificity with which biomolecules bind. Therefore, a mechanistic understanding of the physiological function and control of proteins, cell membranes and so forth requires a theoretical model of the effects of solvation. Molecular dynamics (MD) simulations use fully atomistic models for biomolecules surrounded by tens of thousands of solvent molecules, providing very detailed views of molecular solvation [2], but at significant computational cost associated with the solvent degrees of freedom, even though in most cases our primary interest is their collective influence (or other statistical properties). Faster, less detailed models eliminate these explicit solvent molecules and account for their effects implicitly, and are therefore called implicit-solvent models. These models can be orders of magnitude faster than their atomistic counterparts, and are widely used in a range of applications for which an approximate treatment of solvent is acceptable. However, the drastic simplification from explicit to implicit solvent means that not all implicit-solvent models are equally well suited to all applications, and research in this area continually strives to broaden the generality of these faster models.

In this spirit, the present review addresses the basic mathematical concepts and recent progress in model development and computation. The importance of implicit-solvent models, and the wide variety of models and applications, has led to the publication of a large number of excellent, encyclopedic reviews already, spanning chemistry, biology, physics, and mathematics [323]. The present review aims to avoid recapitulating these accounts of progress in the field; instead we focus on two other goals. The first is to illustrate the significant progress in more sophisticated modeling approaches; future improvements will require the adoption of correspondingly sophisticated mathematical and computational methods. Our second, significantly related purpose is to introduce applied mathematicians and computational scientists to the lively and fascinating challenges in developing implicit-solvent models and numerical algorithms for them. The citations are designed to support these purposes, and therefore our primary interest is in being reasonably comprehensive regarding important concepts and research directions; however, the bibliography is far from exhaustive and we apologize for unintentional omissions.

From the perspective of computational science, the following themes and trends seem particularly important.

  • 1.  
    The community has developed a diversity of implicit-solvent models that range tremendously in detail and in computational cost. These models are generally related via physical insight and scientific analysis that certain approximations hold reasonably well. In contrast, in other areas of computational physics such as computational fluid dynamics or electromagnetics, the domains of applicability of the underlying equations (the Navier–Stokes equation and the Maxwell equations, respectively) are known with significantly more rigor and detail. Therefore, in these disciplines, the trade-off between speed and accuracy exists on a spectrum within the same model, and those communities hold a standard of rigorous error analysis, bounding, convergence properties etc. This comparison to classical physics is intended as aspirational rather than critical; after all, the Navier–Stokes and Maxwell equations were developed many decades before our capacity to study molecular water in nearly as much detail. Although we do not yet have such a complete theory for implicit-solvent models, new approaches hold great promise to move us towards such a goal.
  • 2.  
    Computational cost is virtually always a severe limitation, regardless of the investigation. This constraint motivates an ongoing interplay and feedback loop between the underlying model and the numerical algorithms used to compute it. In other words, many times a 'new model' is presented and justified on the basis that numerical computation is faster than was possible with the original theory. It is important to be explicit whether:
    • (a)  
      the proposed approach is actually a numerical approximation of the original model, in which case it must converge to the exact answer given sufficient resolution/time;
    • (b)  
      it is an approximation that is exact only in limiting cases, for instance in the sphere;
    • (c)  
      the new model is physically distinct, that is, it satisfies neither of the preceding two conditions.
    The introduction of smoothly varying permittivities across the solute–solvent boundary represents a popular example. Originally motivated by computational considerations [24, 25], the new model becomes the traditional two-dielectric model only in the limit as the transition region approaches zero width. Clarity on these issues can be an important aid to cross-disciplinary communication and the development of powerful (and suitably general) high-performance software.
  • 3.  
    The comparative lack of mathematical analysis of formal relationships between models arises from justifiable skepticism that any one model gives a reasonable answer in all cases. For instance, Poisson-based models miss solvent structure, and generalized-Born (GB) models only approximate Poisson or Poisson–Boltzmann (PB). In this author's opinion, implicit-solvent models would develop more efficiently by outreach to researchers in applied and computational mathematics.
  • 4.  
    Numerous arguments demonstrate the power of developing mathematical frameworks where rigorous approximations enable a model to span the spectrum from fast analytical models (for dynamics) down to expensive and realistic calculations to test recently proposed physics; in engineering terms, mathematical frameworks are enabling technologies. For example, in Poisson-based theories, one may add nonlinear or nonlocal dielectric effects without sacrificing analytical or computational tractability. Secondly, the computational tools developed for one specific instance in a class can be brought to new, related models in that framework. The importance of this argument increases constantly, especially as high-performance computing platforms such as graphics processing units (GPUs) make scalable software more difficult to program by non-experts. Thirdly, advanced computational methods hold exciting potential to accelerate model development within each theoretical framework (for instance, models based on Poisson-type partial-differential equations (PDEs)).
  • 5.  
    Analytically solvable problems have been underutilized despite the substantial impact that these analyses usually have on the field's development. The most outstanding example is probably Kirkwood's 1934 paper on electrostatics in a sphere [26], which continues to be cited almost 80 years after its publication. Many significant improvements to GB models have come from analytical studies of the sphere problem, for example—but only after an extended period of largely empirical modifications that provided only modest improvements. A subtle related point is that exponential growth in computing capabilities allows us to consider some problems as 'exactly solvable' via large-scale simulation; here it is useful to consider the important paper by Onufriev et al [27] that used a large number of (slow) Poisson calculations to show that the main failings of GB models had been the calculation of effective Born radii.
  • 6.  
    The community's modeling challenges create important opportunities to drive the development of improved mathematical and computational methods. Success stories include motivating early fast multipole methods [28, 29], high-quality models for parallel computing (e.g. in the software NAMD [30]) and multigrid methods [3133]. These advances derive not from the underlying challenges themselves, but from a willingness to encourage researchers in other fields to learn about these challenges, and to contribute as they are able. As implicit-solvent models grow more advanced, the interdisciplinary challenges will become more pronounced and make this spirit of collaboration even more vital for efficient progress.

Implicit-solvent models fall basically into two categories. The first represents essentially empirical models: fitting parameters are designed to reproduce experimental data such as solvation free energies [34, 35]. Such models are extremely valuable in a range of applications for drug design and protein engineering. Models in the second category are designed to reproduce solvation interactions as calculated by more detailed models such as all-atom MD, using a potential-of-mean-force (PMF) approach [36]. Ultimately, of course, these two categories should converge to a single consistent model, but currently, the specific application of interest largely dictates which type of model should be selected. Although models from these two classes often use the same set of equations, e.g. the Poisson equation for electrostatics, it is important to be clear which approach has been adopted in any given work, because the choice of approach affects the interpretation and validation of one's results.

For several reasons, this review focuses on the PMF-based interpretation of implicit-solvent models. Firstly, the PMF concept offers immediate application in multiscale modeling such as MD-based estimates of binding free energies (e.g. MM/PBSA [3746]) or Monte Carlo simulations [4759]. Secondly, a primary goal of this review is to focus on the methodology of model improvement. This challenge is more readily understood in the PMF context, where computation's role as a surrogate for experiment enables model strengths and weaknesses to be probed in more breadth and depth. Thirdly, semi-empirical models pose important and serious complications, such as the controversial question of the meaning of the dielectric constant [60, 61]. Focusing on the PMF question allows us to set aside many of these complications temporarily—not in any way to minimize their importance, but instead to illustrate the numerous difficulties that remain even after their removal! Finally, the computational task of parameterizing implicit-solvent models to match finer-grained simulations is a substantial challenge, with very few standardized tools available, at least to the author's knowledge. Hopefully, this review will spark further developments for computational scientists interested in large-scale inverse problems for molecular biophysics. Also, this review emphasizes mainly implicit-solvent models for molecular mechanics (i.e. classical force fields) rather than those specialized for electronic structure, whose popularity has led to extensive reviews in that specific area, e.g. [7, 18]. Table 1 provides a complete list of the acronyms used throughout the paper, many of which may be new to the computational mathematics or biophysics communities.

Table 1. Acronyms used in the text and their definitions.

Acronym Definition
(L/NL) PB Linear or nonlinear Poisson–Boltzmann
PDE Partial-differential equation
GB Generalized Born
MD Molecular dynamics
PMF Potential of mean force
MM/PBSA Molecular-mechanics/Poisson–Boltzmann plus surface area
MC Monte Carlo
BIE Boundary-integral equation
FDM Finite-difference method
FEM Finite-element method
BEM Boundary-element method
MIB Matched interface and boundary
FFT Fast Fourier transform
SVD Singular value decomposition
SOR Successive over-relaxation
PCM Polarizable continuum model
FMM Fast multipole method
API Application programming interface
CFA Coulomb-field approximation
GBMV Generalized Born with molecular volume
ACE Analytical continuum electrostatics
BIBEE Boundary-integral-based electrostatics estimation
VCFA Variational Coulomb-field approximation
DiSCO Discrete surface charge optimization
GPU Graphics processing unit
FEP Free-energy perturbation
PDLD Protein-dipole–Langevin-dipole
LD Langevin–Debye
DFT Density-functional theory

The remainder of this review is organized as follows. The next section presents the basic mathematical structure of a broad class of implicit-solvent models, along with a brief historical perspective. We then present the main computational methods for these models in sections 3 and 4; sections 3 and 4 address the PDE-based models and several fast approximations to PDE models. The following two sections discuss challenges and weaknesses of current models: section 5 details parameterization and section 6 addresses the limitations of popular models. These weaknesses have spurred the design of advanced models for water (section 7) and also for the behavior of aqueous electrolytes (section 8). The next section discusses new developments and ongoing challenges (section 9) and section 10 concludes the review with a brief summary.

2. Implicit-solvent models

Implicit-solvent models can be viewed as simple approximations to the PMF exerted by actual solvent on the solute molecule(s) of interest [36, 6264]. Our introduction follows Roux and Simonson's excellent early review [36]. Readers unfamiliar with this formalism should consult Roux and Simonson's review directly, because it includes essential discussions about thermodynamic state dependence and time-dependence issues for dynamical simulations; we omit these here to focus on the implicit-solvent models themselves.

The statistical–mechanical argument for deriving the exact PMF begins with a probability density function for the state of a solute surrounded by a solvent,

Equation (1)

where β = (kBT)−1, the solute atom positions are denoted by $\mathbf {X}=\left \{\mathbf {x}_1,\mathbf {x}_2,\ldots \right \}$ , those of the solvent atoms by $\mathbf {Y}=\left \{\mathbf {y}_1,\mathbf {y}_2,\ldots \right \}$ , U(X,Y) is the total potential energy, and the integration is over all allowable atom positions. By 'integrating out' the solvent degrees of freedom, we obtain a marginal probability density function

Equation (2)

We then define a function W(X) such that

Equation (3)

and W(X) is called the PMF because its gradient (with respect to the solute atom positions) is identified with the average force

Equation (4)

and the subscript to the angle brackets denotes the ensemble average taken over all solvent atom positions Y for a fixed solute configuration X. Given W(X), the average value of any function that depends only on X is (by definition) expressed exactly as

Equation (5)

One of the most popular applications of the PMF formalism is in the evaluation of molecular binding free energies, e.g. drug–protein and protein–protein binding (more details of these issues are addressed elsewhere [65]).

Most potential functions in atomic simulations, for example in MD force fields, can be written as

Equation (6)

where Um(X) denotes the intramolecular potential (solute van der Waals interactions, bond stretch energies etc), Uss(Y) denotes the solvent–solvent interactions, and the two terms of most interest are the nonpolar interactions (e.g. van der Waals) between the solute and the solvent, U(np)ms(X,Y), and the solute–solvent electrostatic interactions Uesms(X,Y). Then the desired PMF can be written as

Equation (7)

but it is important to note that this decomposition into a sum of energetic terms is critically associated with the order of operations. Envision the process of 'growing' the solute into a solvent bath, i.e. by slowly changing the solute atoms' van der Waals terms and charges from zero (the initial state, solvent with no solute) to their actual value (the final state is the solute in water): the overall change in free energy cannot depend on whether one turns on the charges first and then the van der Waals interactions, or vice versa. In actual computations, turning on the charges first leads to difficulties associated with solvent molecules overlapping bare point charges; instead, one 'turns on' the nonpolar interactions before the electrostatic ones [36].

In this setting, the whole challenge of implicit-solvent modeling is to find rapidly computable but highly accurate approximations to ΔW(np)(X) and ΔW(es)(X).

2.1. Nonpolar component of implicit-solvent models

The nonpolar term is often modeled as proportional to the solute surface area, i.e. as

Equation (8)

where γ is the surface tension and A(X) is the surface area—a function that is heuristic, in the sense that it does not have an exactly defined form which can be approximated numerically to an arbitrary controlled precision. Roux and Simonson [36] present a simple, intuitive explanation of the widespread use of this functional form, and we will not address the nonpolar term further except to note that the subject remains an active research area [6672]. Wagoner and Baker [73, 74] used MD simulations to illustrate the challenges with existing models, and Isele-Holder et al [75] have discussed calculating surface areas from MD.

2.2. Electrostatic component of implicit-solvent models

Early studies of solvent influence on molecules demonstrated that macroscopic dielectric theory could be surprisingly useful in modeling the electrostatic component of solvation [26]. In this approach, the electrostatic potential φ(r) obeys a Poisson equation, which in its simplest form can be written as

Equation (9)

where r is the spatial position, ε(r) is a spatially varying dielectric constant, ε0 is the permittivity of free space and ρsolute(r) is the solute charge distribution (usually a set of discrete point charges situated at the atom positions). Equation (9) is usually presented as the fundamental continuum law, which is unfortunate, because dielectric theory offers substantially more sophistication.

In particular, for a pure-water solvent (i.e. with no mobile ions), equation (9) is just Gauss's law, which is a conservation law for the electric flux,

Equation (10)

where ρbound(r), the second term on the right-hand side, represents the density of induced charge that arises in a dielectric medium in response to an imposed electric field. It is defined by

Equation (11)

where the polarization density field P(r) is defined by the medium's constitutive relation. Equation (9) assumes that the constitutive relation is

Equation (12)

where the electric susceptibility χ(r) is related to the dielectric constant by ε(r) = 1 + χ(r). Often these expressions are simplified by the definition of the electric displacement field D and its relationship to the electric field E(r) = −∇φ(r),

Equation (13)

Of course, as with all other forms of differential equations, the mathematical problem is not well posed until the boundary conditions are specified.

Very commonly, a simple dielectric model is used in which space is separated into distinct solute and solvent regions, and then each is modeled as a uniform dielectric. In other words, the solvent is usually modeled with water's macroscopic dielectric constant of about 80, and the solute permittivity is assigned a value between 1 and 10; the definition of the solute dielectric constant remains controversial (e.g. [60]) and depends on the application. For that matter, in reality, the solvent dielectric constant (relative permittivity) also depends on the ion concentrations and temperature. In the PMF view, the appropriate dielectric constant is usually 1 (giving intramolecular interactions the same exact value they would have in the corresponding MD simulation) or 2. The dielectric constant ε(r) is discontinuous as r crosses the boundary Ω separating these two regions.

From the Maxwell equations, the potential is continuous across the boundary, as is the normal component of the displacement:

Equation (14)

Equation (15)

and defining

Equation (16)

we can derive a boundary-integral equation (BIE) that is exactly equivalent to the mixed-dielectric Poisson problem of equation (9):

Equation (17)

where σ(r) is the infinitesimally thin layer of charge that develops at a dielectric boundary in response to an applied field; consider that the polarization field P(r) changes discontinuously across the boundary. Equation (17) is well known in several areas of science and engineering [54, 7679]. We refer readers unfamiliar with BIEs to [80, 81].

The situation is orders of magnitude more complicated still if the solvent includes mobile ions—which is to say, every biologically relevant solvent. A simplistic approximation assumes that the ions' effect can be modeled using concentration fields that are proportional to the Boltzmann factor for an ion placed at the given point, i.e.

Equation (18)

where K denotes the number of ionic species, kB is the Boltzmann constant, T is the temperature and ci and qi represent the bulk concentration and charge of the ith ion species. Equation (18) is a particularly simple form of the famous PB equation. Explicit-atom approaches are particularly challenging for dilute salt solutions; in contrast, continuum models capture these effects naturally. Nevertheless, the PB equation has its own severe limitations, neglecting numerous chemically and biologically important phenomena (see section 6). Furthermore, because the PB model is still a nonlinear PDE, it is far from trivial to analyze, and its linearized form (the LPB equation) has become a popular alternative. Sharp and Honig [82] have addressed many subtle aspects, as do numerous reviews [14] . Many implicit-solvent models rely on a still more drastic simplification by linearizing the right-hand side of equation (18), which amounts to assuming that the electrostatic potential is always very small (so that qiφ(r)/kBT ≪ 1) or that the bulk ion concentrations are vanishingly small.

3. Numerical methods for the standard continuum Poisson model

The Poisson, PB and linearized Poisson–Boltzmann (LPB) problems belong to a class of mathematical models called elliptic PDEs, which mathematicians have studied for over 200 years due to their important applications in the study of celestial mechanics, fluid mechanics, elasticity, electromagnetic theory and quantum mechanics. Before the rise of high-speed digital computers, analytical solutions, e.g. in separable geometries such as spheres, represented the primary tool for theoretical investigation. Consider that even today, papers cite Kirkwood's 1934 paper [26] deriving the analytical solution for charges in a spherical model protein in a solvent obeying the LPB equation. More complicated geometries required arduous amounts of manual arithmetic, for instance via the Ritz (or Rayleigh–Ritz) method, which is essentially the finite-element method (FEM) we know today.

The wide range of applications of elliptic PDEs has driven the development of numerical algorithms for solving arbitrary problems, and in general, methods for solving Poisson equations may be classified as belonging to one of two classes. Volume-based methods solve the PDE directly to obtain the unknown (here, generally the electrostatic potential) throughout space. In contrast, surface-based methods can be used to reformulate the PDE as an equivalent BIE, or system of coupled BIEs, and find an unknown function (or set of unknown functions) on the boundaries between regions, e.g. the molecule and the solvent. Such techniques are less readily applied to nonlinear PDEs (e.g. the actual PB equation) or problems with smoothly varying dielectric constants. Note that BIEs pose a different set of challenges from volume integral equations, whether for the Poisson problem [83] or for statistical–mechanical theories of solvent structure [84].

Our discussion of numerical methods is relatively brief—for details, see the recent reviews of Lu et al [21] and Shi and Koehl [85], especially for methods of solving the Poisson and PB equation in its differential form. In addition, we assume somewhat greater familiarity with computational mathematics than in other sections of the review.

3.1. Volume-based methods

The first numerical calculation of the mixed-dielectric Poisson problem equation (9) for a protein was performed in 1982 [86] using a finite-difference method (FDM). Careful work by Honig and co-workers in their development of DelPhi software [8790] established many important details for numerical accuracy. Important advances addressed the need to approximate the boundary condition at infinity, ways to represent the complicated dielectric boundaries and methods to represent the protein charge distribution (a set of discrete point charges) using the grid points for the FDM [91, 92]. One particularly important advantage of volume-based methods is that they can be extended easily to solve the PB equation in either its linear or nonlinear forms, with the latter requiring more delicate analysis [82]. A second important capability is that the underlying dielectric model is easily changed from a piecewise-constant form (i.e. εprotein in the protein and εwater outside) to one with multiple regions or ion-exclusion regions around the molecule (Stern layer) [93, 94], or with a permittivity that varies smoothly from εprotein to εwater, e.g. [25].

The general aim of efforts to improve volume-based methods has been, of course, to allow fast and accurate calculations, which means high accuracy at low grid resolutions [91]. Here, we cover improvements that apply equally well to both static and dynamic calculations (Luo et al have studied extensively issues of dynamics [9598]). An important recent development was the introduction of a post-processing step to compute the induced surface charge σ(r) from the grid potentials [99]; it was found that this approach offers substantially more accurate reaction potentials at the charge locations, for a given grid resolution.

One of the most valuable theoretical developments of the last few years has been the decomposition of the electrostatic potential into separate Coulombic and reaction terms [33, 100102]:

Equation (19)

The first term contains all the singularities due to the use of point charges, and the function φreac is harmonic, i.e. it satisfies the Laplace equation. These properties make the reaction potential suitable for rigorous mathematical analysis [33], and calculations exhibit superior accuracy at lower grid resolutions because the (smooth) reaction potential is easier to approximate accurately than the total potential which has sharp peaks corresponding to the singularities [101, 102]. We note that this is the same reason that smoothed dielectric boundaries [25] allow reduced grid resolutions. Wei and co-workers have combined this decomposition with the matched interface and boundary method to develop a second-order accurate solver [100, 101, 103, 104], and the Luo group has characterized several ways to improve the convergence rate of the iterative method [98, 105], including using a smooth dielectric boundary [98]. A decomposition similar to equation (19), proposed by Chern et al [106] and developed thoroughly by Holst et al [107], splits the total potential into three terms rather than two. The extra term derives from further decomposing the reaction potential φreac in equation (19), in order to address an instability that arises for volumetric numerical methods. Holst and co-workers [108] have also applied this decomposition to the Poisson–Nernst–Planck model for diffusion–reaction problems.

Early work on FEM simulations of elliptic PDEs [109113] established the value of FEM for molecular electrostatics, and later work by Baker and Holst provided the mathematical and algorithmic foundations for the widely used APBS software [114, 115]. It is instructive to note the substantial advantages of the software's rigorous and general mathematical basis, which allow APBS to be a scalable and flexible framework for numerous continuum models, and Holst and co-workers [116] continue to advance finite-element modeling in a variety of ways, including meshing (as have several others recently, including in particular [85]). Very recently, adaptive grid FDMs [117] and novel finite-element approaches have also been introduced [118], and it will be interesting to see how their advantages can be leveraged among the many applications of implicit-solvent models.

3.2. Surface-based methods

The simplest example of a surface method is the BIE of equation (17) for the mixed-dielectric Poisson problem. BIEs have several advantages and disadvantages compared to PDE methods. An important advantage is that the boundary conditions at infinity are handled exactly, with all of the unknowns on the (finite) dielectric boundary itself, rather than the potential throughout all space and towards infinity. A closely related advantage, which we will not discuss further here, is that BIEs can analytically account for other semi-infinite domains, such as a planar membrane, using what are essentially image charges. A second related advantage is that the unknowns in BIE problems, being distributions on a surface, are only 'two-dimensional unknowns', rather than the three-dimensional unknown electrostatic potential. This means that, in principle, BIEs can be more efficient due to a smaller number of unknowns. Finally, the potential due to the protein charge distribution is captured exactly, because one computes its influence on the boundary (or boundaries) directly. This property is one reason why surface approaches have been comparatively popular in electronic structure methods, where the charge distribution (i.e. the electron density) is very complex compared to a set of point charges; furthermore, the smooth potential throughout the solute volume is the main subject of interest.

The main disadvantages of BIE are of three types. Firstly, implementing numerical calculations, using what are known as boundary-element methods (BEM), is more challenging. Secondly, the usual BIE approaches are limited to problems with piece-wise-constant dielectric constants, and also to linear problems. The latter issue is severely restrictive when exploring the space of possible theories, and especially given that standard BIE can be applied to linearized PB problems, but not the fully NLPB equation.

Expanding on the first class of challenges: numerical difficulties include primarily the fact that most approaches lead to a dense matrix equation. That is, from BEM one obtains a linear system of equations Ax = b where every entry of A is non-zero; thus, the memory required to store A grows quadratically with the number of unknowns n, and the time required to compute x using LU factorization increases cubically, i.e. as O(n3). In contrast, because finite-difference, finite-volume and finite-element methods approximate the differential form of the Poisson equation, they produce very sparse matrices.

The dense nature of A is easy to conceptualize if one imagines representing the induced boundary charge σ(r) as a set of discrete point charges on the boundary—every point charge produces an electric field everywhere else. Happily, this same conceptual picture leads to a much faster approach, a class of techniques known as fast BEM solvers, which allow these problems to be solved using time and memory that scale as O(n). First of all, the N2 interaction can be computed using fast N-body algorithms such as the fast-multipole method [29, 119121], tree-codes [122], multiple grid methods [123], P3M [124], precorrected-FFT [125], kernel-independent fast multipole method [126] or FFTSVD [127]. These algorithms allow one to perform accurate matrix–vector multiplication using time and memory that scale as O(n) or O(nlogn) rather than the normal O(n2) time and memory. The second component of the fast solver approach is to use the Krylov-subspace iterative methods [128, 129] such as GMRES [130] to compute an approximate solution $\hat {x}\approx A^{-1} b$ that lies in the Krylov subspace 〈b,Ab,A2b,...〉. It should be noted that iterative methods are widely used in volume-based methods as well, with successive over-relaxation among the most popular [90].

The second major challenge for BEM calculations is that the matrix entries are comparatively difficult to compute accurately, because they involve calculating integrals of singular or near singular functions [131]. Calculation accuracy can depend quite strongly on implementation details that might not seem important [132, 133]. The third disadvantage is that dynamics have been much more difficult to address, because most boundary-integral methods require some kind of mesh (see below), which must be regenerated along with the discretized matrix at each time step. This disadvantage is not fundamental, however: Totrov and Abagyan [134] demonstrated that BEM simulations could be fast enough for Monte Carlo simulations of protein folding, and Wang et al [135] introduced methods for stochastic dynamics.

3.2.1. Boundary-element methods.

One of the first applications of a surface approach was designed for electronic-structure calculations for molecules in solvent [78]. In the quantum-chemistry community, equation (17) is called the polarizable continuum model, and numerous advances enable treatment of LPB models [136, 137], anisotropic dielectrics [138140] and interfaces [141], among other extensions. For details, see the extensive reviews available on these methods [7, 9, 10, 18]. Equation (17) was also derived by Shaw [79] for molecular mechanics and point charge models, and later significantly extended by Zauhar and Morgan [142145]; as a side note, this BIE applies to many other problems in science and engineering [76, 77, 146]. Other BEM advances were made by Rashin and Namboodiri [147, 148], Pratt et al [149], Zhou [150], Purisima and Nilar [151, 152], Scheraga and co-workers [153], Vorobjev et al [154157] and Honig and co-workers [158].

A major challenge was including LPB models in the solvent, which is substantially easier for volume-based methods. The first reported boundary-integral only approach was Yoon and Lenhoff's formulation [159] of the LPBE problem. Previous methods had discussed hybrid boundary/volume methods for solving the NLPB. Juffer et al introduced another LPBE formulation in 1991 [160], and demonstrated a simplified method for the Poisson problem in the appropriate limit (i.e. as the salt concentration goes to zero). Other formulations have been presented [161, 162]. Not all formulations are independent of one another—the relationships between them, and several equivalences, have been discussed [131, 136, 137]. An interesting mathematical challenge arises for electronic structure problems in which the charge distribution is usually non-zero outside the dielectric boundary [163165].

A second major challenge that strongly limited surface methods' popularity during their early development was that LPB problems could not be solved efficiently. Early fast methods allowed fast calculation only for the mixed-dielectric Poisson problem [158, 166]. Fenley et al made an important (and, unfortunately, frequently overlooked) contribution in their development of fast multipole methods for LPB [167169]. Greengard et al [120, 170] developed a consistent and rigorous FMM for LPB, which combined the Laplace and Helmholtz FMM methods. The complications of deriving a different FMM for every PDE, or more precisely for the associated Green's function, motivated the development of several fast algorithms that are said to be Green's-function- or kernel-independent [125127]. With these techniques available, it became possible to compute LPB problems using BEM [169, 171174]. Recently, the boundary-integral community endeavored to find still faster simulation approaches that do not require Krylov iterative methods but instead directly apply an approximate inverse of the matrix. These approaches are known as fast direct methods [175177], and promise to accelerate certain calculations by orders of magnitude, including pKa calculations (see, e.g., the formalism in [178]) and electrostatic optimization in molecular design [179181].

As our mathematical models of solvation grow more sophisticated, the generality of volume-based approaches seems at first to offer a steadily increasing advantage over boundary-integral methods. Nevertheless, the system response is dictated by the shape of the molecular boundary (whatever it may be) and the associated boundary conditions. As a result, surface methods will still hold important possibilities for characterizing solvation free energies, even if the mathematical methods for obtaining them require substantial sophistication [182].

3.2.2. Generating a discretized boundary representation (meshing).

Boundary-integral methods require some method to describe the surface unknowns. One simple approach is to assume that these unknown functions can be represented using a set of discrete point charges at the dielectric interface (or ion-exclusion surface) [157, 183187]. Such a discretization is readily implemented, and numerous software packages exist to generate the needed set of surface points and associated weights, e.g. [157]. The price one pays for this simplicity, however, is that accurate calculations then demand a large number of unknowns; strictly speaking, the unknown is actually a function, or set of functions, on the surface, and thus not well approximated by a discrete number of sources. Most work therefore employs the next simplest method, which is to describe the boundary using a set of flat (planar) triangles or quadrilaterals, e.g. [159, 160]. Such triangularizations or meshes can be obtained using any of several software applications, e.g. [188, 189].

Mesh generation remains a challenging problem, especially for extremely large systems such as the ribosome or viral capsids [189], and accordingly a large number of meshing algorithms and methods have been presented [157, 160, 188208]. More than one algorithm includes slight modifications to the surface definition, for varying reasons (e.g. computing the mesh becomes easier or faster; sharp features on the surface are smoothed out) [145, 157, 209]. In addition, certain BEM formulations are sensitive to accuracy loss if the boundary elements take on pathological shapes—for example, infinitesimally thin triangles. In addition, the standard Connolly molecular surface often has singularities, or 'cusps' where the surface normal is discontinuous. These pathological cases not only complicate the theoretical analysis of the underlying BIEs, but also cause unpredictable failures in mesh generation and may degrade simulation accuracy. A particularly promising approach has been advanced by Edelsbrunner, Shi, Koehl and co-workers in their development of an extensive, rigorous understanding of the mathematics of molecular surfaces with provable smoothness properties [199, 210216].

3.2.3. Higher-order methods and other approaches.

As with any numerical simulation, boundary-integral methods for implicit-solvent models ask the user to choose a trade-off between calculation accuracy and resource use (computer time and memory). One way to improve accuracy involves using higher-order methods, for instance, allowing the charge distribution on each element to vary linearly or quadratically, instead of approximating it as constant [159, 172]; other methods have also been introduced [217]. However, the impact on accuracy of such extensions is limited as long as one continues to approximate curved molecular surfaces using flat boundary elements, because the error is then dominated by the incorrect representation of the dielectric boundary.

Early work on curved-element BEM set a high standard. Zauhar introduced an exact discretization method for the solvent-accessible surface (a union of spheres, in contrast to the Richards molecular surface which includes spherical patches as well as patches of tori [204, 218, 219]). Liang and Subramaniam employed Edelsbrunner's theory of alpha shapes to define and discretize the surface [210] and a rigorous numerical method for computing the needed singular and near-singular integrals [220]. Later work by Bordner and Huber [161] employed quadratically curved elements and linear basis functions that have been thoroughly studied in the BEM community [221]. These types of curved elements and others cannot represent the usual surface definitions exactly (van der Waals, solvent-accessible or molecular), which can complicate understanding whether the geometry error or basis-set error dominates in most protein calculations. To test the influence of geometry error, the author and collaborators developed a specialized fast BEM algorithm, new curved elements, and surface meshing algorithms so that the discretized surface was exact at every level of discretization [127, 174, 222]. The mesh generation process was computationally challenging, even though it was based on leading FEM/BEM mesh software [223]. These studies demonstrated that for the common accuracy requirements, curved element BEM required approximately ten times fewer basis functions to achieve the same accuracy, because the surface was better represented. The authors also showed that with curved-element BEM, it became possible to obtain much more converged solvation energies of proteins than had been demonstrated previously [174]. However, these simulations required expensive numerical integrations that make the approach suitable for a comparatively limited set of applications, for example, in molecular design [174].

In contrast, Bajaj et al [192, 196, 224] have developed curved-element methods for surface discretizations using B-splines. Because splines are widely used in graphics and computational science, and many algorithms are known for computations on splines, the minor inexactness in surface representation seems to be a small disadvantage in practice. Furthermore, the availability of high-quality meshing algorithms recommends their use for higher-order BEM simulations. To demonstrate this advantage and others, Bajaj and co-workers have adapted the kernel-independent fast multipole method [126] for BEM simulation of the LPB problem [225]. Other advanced boundary-integral approaches include spectral boundary-integral methods, which have recently been applied to proteins by Kuo et al  [226]. The advantage of spectral methods is that they require a very small number of unknowns to achieve high accuracy; Kuo et al demonstrated that for a small molecule, three digits of accuracy could be obtained with 20 times fewer unknowns than standard, triangular-element BEM. The price for this remarkable improvement is the need to find appropriate surface parameterizations, an area which mathematicians or computational mesh-generation specialists may be able to contribute substantially. Hybrid boundary/volume methods have been proposed for solving the full NLPB problem [227]; because the nonlinearities tend to be important only in a relatively thin region around the protein surface, coupled methods might be efficient. Simulations coupling BEM and FEM/FDM are more widely performed in other areas of computational physics than in implicit-solvent models [228], and could be a promising area for interdisciplinary collaboration.

Last but certainly not least, time-dependent solvation problems have been studied using BEM, first by Song et al [229, 230], and later by Hofinger and Simonson  [185]; the standard approach is to solve multiple problems in the frequency domain, using known models of water's dielectric response at different frequencies (i.e. one substitutes ε(r,ω) into equation (17)). Boundary-integral methods have an interesting advantage for such problems, in that the cost to simulate a large system using a fast BEM algorithm is dominated by the setup time for computing a sparsified representation of the dense BEM matrix. Because the same representation is used for each of the frequencies simulated, the setup cost should be largely amortized.

3.3. A perspective on software

Clearly, there is no shortage of numerical methods for solving the Poisson equation or its relatives. Popular high-quality software packages already offer high-speed calculations, thorough documentation and developer support. Some of these packages are freely available as open-source software (e.g. APBS [32, 114, 231]), whereas others are available to academic institutions for very modest fees. Almost always, such software is used in 'production environments'; in other words, the user has significant trust in the software and its developers, has extensively tested it in his or her own laboratory and finally a community of users (perhaps led by the developers themselves) has created software infrastructure to interface it with her other simulation code. For example, Baker et al have developed interfaces between their APBS application and CHARMM (Chemistry at Harvard Macromolecular Mechanics), as well as visual molecular dynamics (VMD). New algorithms and software will have minimal impact in the biophysical community unless the developers provide a variety of support mechanisms to make their new approach competitive as a production-ready code. We list a few below.

Firstly, new software must offer a reasonable interface to existing file formats in computational chemistry, e.g. files for the MD software CHARMM and AMBER (Assisted Model Building with Energy Refinement) or the PQR (Point, Charge(Q), Radius) format [232]. Secondly, the software source code must be readable and modifiable by other groups, or at least offer a high-quality application programming interface, because an interface that goes deeper than a 'black box' can lead to dramatic performance improvements. For instance, coupling fast BEM codes to electrostatic optimization theory can speed up calculations for drug analysis and design by several orders of magnitude [181, 233]. Thirdly, the numerical algorithm and software should be extensible to other models; that is, it should provide a means to solve other models. For instance, Orland and co-workers have implemented an efficient solver for their dipolar Poisson model [234237] using APBS. Fourthly, developers must be willing to offer significant user support. Implicit-solvent models are applied to a diverse and growing range of problems in molecular science and engineering, and no single piece of software can cover them all without some modification, or at the least a moderate amount of assistance.

4. Fast approximate methods for electrostatic free energies

Even the fastest available solvers for the Poisson and PB problems are too computationally expensive for some applications. Therefore, there has been substantial interest in finding much faster approximations to continuum models. Many such applications involve long dynamical simulations and therefore require that the energy be differentiable with respect to atomic positions for force calculations. Computational chemists and biophysicists have developed a variety of mathematical models that give electrostatic free energies in remarkable agreement with solutions of the Poisson or PB equations. However, looking ahead to the development of more advanced models, many of these approximations fail in their capacity for extension to better physics, unless one resorts to ad hoc modifications of existing approximate models.

Here, we review what the community has learned regarding such approximation schemes. The most popular Poisson-based models are undoubtedly the GB models, which we address in detail in section 4.3. Firstly, however, we discuss a key idea that has driven much of GB theory: the Coulomb-field approximation (CFA). We also comment on a recent mathematical formulation regarding an assumption that has been problematic and the subject of recent GB advances, namely the dependence of the solvation free energy on the dielectric constants. Then, after discussing GB, section 4.4 describes two relatively recent approaches based on fast models that employ systematic approximations of the Poisson problem. For a more detailed and mathematically oriented review of GB-type methods, see [23] and the other excellent surveys [22, 238, 239].

4.1. The Coulomb-field approximation

As mentioned in section 2.2, macroscopic dielectric theory assumes that a microscopic polarization charge density P(r) develops in response to an electric field E(r), with a purely local relationship determined by the susceptibility χ(r):

Equation (20)

The electric displacement field D(r) is then

Equation (21)

so that

Equation (22)

To motivate the CFA, consider the computational cost of the Poisson model as arising from the need to find a self-consistent set of fields P and E, because the polarization field induces an extra potential. The CFA makes an approximation to the self-consistency requirement by assuming that

Equation (23)

and thus the Coulomb field directly determines every other field; that is, the polarization charge density does not itself create any further polarization charge. This approximation is extensively discussed in an early paper by Kharkats et al  [240], who address its use in both local and nonlocal dielectric theories, and note its earlier use in modeling Marcus electron transfer. A later development of the theory by Schaefer and Froemmel emphasized its interpretation as a means of approximating the electrostatic energy density [241], and these authors introduced a quasi-image-charge approach for fast modeling. Like other works before and after [242], they decomposed the energy calculation into two summations, one over the atomic 'self energies' and the other over the pair-wise interactions. Luo et al [243] also developed a CFA-like approximation for Warshel's solvent model, which uses a lattice of Langevin dipoles rather than a macroscopic dielectric.

Ghosh et al [244] noted the possible advantages for surface integration over volume integration, and applied the divergence theorem to convert the CFA volume integral to a surface integral. They noted that the resulting surface integral can be viewed as an operator approximation to the BIE of equation (17), and commented that the CFA's exactness for a sphere with central charge can be attributed to symmetry. However, this explanation is not the entire story for the exactness, as discussed in section 4.4. The authors developed two sets of empirical correction terms that improved agreement with Poisson calculations, both short-range and long-range. An early and important review of Bashford and Case [22] employed a detailed analysis of the semi-infinite planar-boundary case to illustrate the CFA's weaknesses, and discuss the introduction by Srinivasan et al [245] of a GB-style approximation to include salt effects via LPB theory. Bashford and Case note, however, that 'there is no good reason to expect GB models to overcome ... intrinsic limitations of continuum models'. As implicit-solvent models add complexity to mitigate their inaccuracies, it will be imperative to heed to this consideration and aim at developing rigorous approaches to allow simple and easy transfer of new physics from advanced models to faster approximate ones.

4.2. The notion of separability between geometry and material properties

We recently observed [246] an interesting common feature of most fast Poisson approximations: the reaction field is assumed to take a separable form in which the dielectric constants and the geometry are decoupled, i.e. the reaction potential due to a charge at rj obeys the ansatz

Equation (24)

The classic GB model follows this approximation, as do some fast models based on approximating equation (17) [247]. Failures associated with this assumption have been noted repeatedly [248], but not explained from the separability perspective. For example, Lee et al [249] raised the possibility that parameters could be dependent on the dielectric constants, and the subject has been discussed in a recent review [239]. A more detailed analysis of this assumption may enable simple and physically rigorous improvements to GB models and others.

4.3. Generalized Born models

The main idea of the GB model dates back to more than five decades ago (its early history is characterized in a paper detailing the pair-wise-descreening approximation [250]). One separates the electrostatic solvation free energy of the quadratic expression

Equation (25)

where q is the vector of atomic charges and L is the reaction-potential operator, into contributions from the diagonal and off-diagonal components of L:

Equation (26)

where the first sum accounts for the individual charges' 'self-energies', and the double summation includes the screened interactions between different charges. This decomposition then readily leads to the identification of the first term in terms of effective Born radii, which is defined as the radius of a sphere which would have the same solvation free energy (Lii). Still et al [251] gave a simple empirical functional form to approximate Lij in terms of the distance rij and the effective radii Ri and Rj. Even though more than 20 years have passed since their paper [251], only a few attempts have been made to modify this form, and none have made enormous advances in accuracy. In fact, a insightful mathematical analysis by Onufriev et al [248] showed that for charges in a spherical molecule, a strikingly similar function to Still's may be obtained. This work provided a much stronger basis for the GB approach than had been known previously.

In 2002, two important publications significantly changed the direction of GB research [27, 249]. Firstly, Onufriev et al [27] published a paper of significant importance for both practical and computational reasons; they used extensive PB calculations to calculate the 'perfect Born radii' and assessed the accuracy of the GB model with these ideal parameters. This process, by removing 'effective Born radius error' due to CFA or other heuristics, focuses distinctly on one central question about GB: the adequacy of the underlying Still equation. As a practical matter, the study was important because it demonstrated that GB is extremely accurate for Poisson problems, giving confidence in the method's wide use. Furthermore, their results indicated that the emphasis in model development could focus on designing methods to more accurately estimate radii (not just solvation energies). From the philosophical point of view, their work provided a new understanding of the limitations of GB models, namely that the GB should not, except in the cases of overparameterization (the 'fitting an elephant' problem), perform better than the results indicated in that paper. This work shows how an accurate numerical computation can provide a substitute for 'exact results', and thus give much deeper insights for the general problem than are available for analytically solvable cases. We also note that the authors have commented that the system Green's function is the basic Coulomb contribution plus the reaction-field component; we will return to this point in greater detail in section 4.4.

In a second paper, Lee et al [249] reported a major breakthrough, apparently serendipitous, in finding systematic corrections to the CFA's weaknesses. They noted that the CFA involves the energy-density volume integral $\int 1/r^4\, \mathrm {d}V$ , where r is the distance from the charge location, so the electric field falls as 1/r2 and the energy density |E|2 falls as 1/r4. The 1/r4 integrand is of the same form as the integration between a monopole and an induced dipole. This observation, combined with the clear need for more accurate self-energies (beyond those computed via the CFA), suggested that higher-order interactions might significantly improve accuracy. They found that a 1/r5 volume integral (and the square root of the result) accomplished precisely this effect, although as the authors noted it is empirical, i.e. it lacks the physical interpretation of the CFA. Furthermore, this correction involves only a single parameter, which can be set so that the computed Born radii are exact for the sphere. The authors presented two versions of this much improved model, called GBMV. One was a slower version that employs a grid-based numerical integration (i.e. it actually computes the volume integral). The other was an approximate method based on analytically evaluated integrals. Although the former naturally proved to be more accurate than the latter, the latter's speed advantage and ready implementation in dynamics argued for its adoption in implicit-solvent MD over the grid-based method.

The work of Lee et al spurred several further analyses. The same group developed an improved version (called GBMV2) that used the fourth root of a 1/r7 volume integral and a modified version of the Still equation [252]. Also, Grycuk [253] and Wojciechowski and Lesyng [254] analyzed the sphere case extensively. Grycuk suggested a 1/r6 volume integral, and explicitly assumed that εprotein ≪ εwater (the work of the Onufriev group addressed the limitations of this common assumption [248]). Another development 'beyond the CFA' presented a generalization of the volume-integral methods [255], using what the authors termed an 'expansion of energy' approach (again, rather than potential). In particular, they suggest computing self-energies of several of these volume integrals,

Equation (27)

with the self-energy defined by

Equation (28)

and An are determined by parameter fitting, and obviously the sum in practice has a finite number of terms (in practice they use 4–7). Note that this energy is still separable in epsilons and geometry. However, the authors do discuss the non-separable aspects of dependence and point out that their approach improves accuracy over a range of dielectric constants.

Bajaj and Zhao recently published a sophisticated implementation of surface GB [224], which combines a number of advanced numerical algorithms. Motivated by the computational cost of approximating Born radii from perfect ones [27, 238] and the need to compute them accurately to have stable simulations, their software comprises four key steps. Firstly, they triangulate the surface (they use a Gaussian definition [256]); secondly, they improve the mesh quality; thirdly, they represent the mesh using a splice surface (i.e. approximate curvature with high accuracy); finally, they use a fast, non-uniform FFT to perform the needed surface integrations. Their detailed analysis of numerical error and convergence should be a model for other work in the area, and allows them to discuss with confidence the application of their methods to the recent GBr6NL model [257]. An important point not discussed in that work, however, is that empirical correction terms such as in the original surface GB [244] are not always suitable for fast, scalable computation.

4.4. Fast methods based on operator approximations

Whereas GB models have been developed to give at the end a simple, analytical functional form for the charge–charge interactions, relatively few publications have addressed the problem as one of mathematical approximation. The exception, in general, has been studies using spheres or other model geometries as limiting cases, and the fast approximations are not always exact even in simple geometries. This is unfortunate, because the formal problem statement is extremely simple, and as we shall see it provides a novel framework for developing fast continuum-model approximations: the challenge is to find a rapidly computed, and hopefully analytical approximation to the system Green's function, which furthermore should be differentiable with respect to the atom positions.

The readers are, of course, familiar with the notion that in a vacuum, a point charge at r0 gives rise to the Coulomb potential

Equation (29)

this Green's function is known as the free-space Green's function, i.e. φ(r) = GFS(r,r0) satisfies

Equation (30)

where δ(·) is the Dirac delta function, and the boundary condition is that the potential decays to zero at infinity. This is just a special case of the system Green's function Gsys(r,r0), which is the response at r to a point source at r0 in a specified system with boundary conditions, i.e.

Equation (31)

where ε(r) is just the relative permittivity throughout space. Crucially, the system Green's function is the sum of the direct Coulomb potential and the reaction potential—this is the same decomposition that has proven valuable in volume-based numerical methods (equation (19)). The function Gsys essentially inverts the PDE—including the dielectric constants, the shape of the boundary, the Debye screening parameter κ for LPB problems—so that given an arbitrary charge distribution ρ(r') in the molecule, one obtains the potential via the convolution

Equation (32)

To obtain the reaction free energy, we make use of the fact that Gsys = GFS + Greac to obtain the familiar relation

Equation (33)

which simplifies for point charges to

Equation (34)

Roux and Simonson made essentially this point in early reviews [36, 63], but apparently its import went unrecognized. A likely explanation for this missed opportunity is that the usual decomposition for the electrostatic solvation free energy is as a pair of discrete summations over the atom charges, i.e.

Equation (35)

and we use Roux and Simonson's notation F(ri,rj) to highlight that these two equivalent expressions, equations (34) and (35), do not suggest the same approaches to modeling and computation.

The first term is appealing because it connects with the Born expression for self-energy of an ion; in this author's view, the self-energy is just a red herring, because it obscures the need to approximate the reaction field potential as a function. It should be noted that much early work, particularly by Schaefer et al [241, 242], others (SEDO etc), addressed the problem of the total energy density. However, the singular nature of the energy density (due to the Coulomb term in the total potential) necessitated complicated re-arrangements, such as integrating over volumes outside the atomic one. Our intention in highlighting this observation is not to dispute that self-energy and charge–charge interaction terms can indeed be treated separately, given the appropriate care, as the ACE model [242] demonstrated and as Sigalov and Onufriev showed more recently [248]. However, recent work has demonstrated convincingly [247, 258] that separating the two is not necessary and, in fact, that treating them together in a field approach presents new opportunities for mathematical analysis.

Using the PDE form of the Poisson problem, Borgis et al [258] developed a functional approach to the electrostatic free energy. Their work begins by developing a local approximation of the polarization field; this approximation appears to be equivalent to the BIBEE/CFA method described below, although more analysis is needed. As noted in work on BIBEE [247], this approximation is generally less accurate for problems dominated by the dipolar component of the total charge distribution. Borgis et al developed an effective re-scaling for these contributions using mathematical analysis and found excellent agreement for many conformations of small molecules. The authors then introduce what they term the variational CFA (VCFA), which approximates the polarization field by a set of 'virtual charges' situated at the charge locations; the variational approach implies that they obtain the best approximation possible for the given basis set, and note that the virtual charge distribution could also include multipoles centered at the charge locations. They show, furthermore, that the VCFA is exact for a dipole inside a sphere, which suggests that this approximate model automatically goes beyond the separability approximation. Their framework offers significant modeling power, as they note that their framework can naturally allow dynamics via Car–Parrinello-type methods, that different definitions of the dielectric boundary (sharp and smoothed) pose no difficulty, that they can improve GB radii estimation and that the variational approximation is readily extended to more sophisticated solvation models. In light of these advantages, Borgis's formalism appears to have a lot of potential (pun intended).

The possible value of fictitious charges for reducing computational complexity is not new, of course, and here we detail a connection between Borgis's framework and the DiSCO multiscale model of Beard and Schlick [259]. The DiSCO model was developed to study extremely large macromolecular systems (chromatin folding processes) for which efficient computational methods are an absolute necessity. Citing earlier 'reduced charge' models, including Stigter's pre-HPC line-of-charge approach for long DNA [260] and the work of Gabdoulline and Wade [261], the authors proposed to assemble a set of discrete LPB 'point charges' situated on a fictitious surface surrounding each solute of interest. Charge values are found by numerically minimizing the difference between a computed PB field and the approximated field. The fictitious point charge approach recalls the early CFA work of Schaefer and Froemmel [241], as well as the basic ideas of potential theory and BIEs. However, Borgis's VCFA approach, or a derivative thereof, may provide a rigorous unifying methodology for applications that span in scale from small molecules (ions or drugs) to large macromolecules. Beard and Schlick performed careful studies to validate the accuracy of their model, and a solid theoretical foundation such as VCFA should allow the design of rigorous error estimators to allow seamless adaptivity with well-characterized error control.

Bardhan et al have developed a complementary operator approximation based on the BIE of equation (17) [246, 247, 262]. This approximation, called boundary-integral-based electrostatics estimation (BIBEE), relies on the spectral properties of the boundary-integral operators. The first BIBEE study showed that the CFA can be interpreted in terms of the extremal eigenvalue and eigenvector of the boundary-integral operator, which are associated with the overall monopole of the solute charge distribution [247]. The same boundary-integral analysis illustrated that the CFA is exact not only for a sphere with central charge, but also for any charge distribution that generates a uniform potential (see a related discussion in [240]). A different approximation is called BIBEE/P, because it relies on BEM preconditioning, which is a way to accelerate the convergence of the iterative method for solving a full BEM problem. The BIBEE/P variant employs the other end of the operator spectrum, and later work proved that the two variants provide upper and lower bounds on the electrostatic solvation free energy [262]. More precisely, BIBEE/CFA provides an upper bound for all shapes, and BIBEE/P is a provable bound only for spheres and prolate spheroids; however, tests on hundreds of proteins (the Feig et al test set [263]) did not provide an example of its failure as an effective bound.

One advantage of the BIBEE algorithm is that the main computational expense is in computing the solute charges' Coulomb potentials and electric fields at the dielectric boundaries. As a result, existing scalable algorithms (e.g. parallel fast multipole methods [264]) are directly applicable without modification [187]. By developing an approximate model that uses only 'algorithmic primitives', it becomes possible to use external libraries such that the same BIBEE code can run on a single or multiple CPUs, on one GPU or multiple GPUs on a CPU, or even on a large cluster containing hundreds of GPUs [187]. More recently, the method's accuracy was substantially improved by studying the operator eigenvalues more carefully, using the analytically solved case of a sphere [246]. The new one-parameter model (called BIBEE/I) provided 4% accuracy on the Feig test set, and was competitive with GBMV for an ensemble of peptide conformations. However, to date the BIBEE model has not yet been applied to LPB problems, and there exists no implementation that might be suitable for dynamics.

5. Parameterization

As mentioned in the introduction, implicit-solvent models are generally either designed semi-empirically, i.e. to match experimental measurements such as solvation free energies or pKa shifts, or, in accordance with the PMF interpretation, to approximate the solvent–solute PMF from MD or other data.

Some of the first empirical parameter sets focused naturally on the relatively simple problems, such as ion solvation thermodynamics, to allow important simplifications. For example, one could assume that the solute takes a simple spherical shape, that the molecular charge distribution is particularly simple (assumed to be a point charge at the sphere center, of value equal to the ion valence) and that the polar component of solvation was the dominant energy. Earlier studies had employed ionic radii, which led to significant errors because calculated solvation enthalpies were substantially larger than experimental data. Rashin and Honig's derivation of the Born expression indicated that the correct radius would be that of the cavity formed in the particular solvent, in this case water [34]. For cations, this suggested covalent radii, and for anions this suggested ionic radii; however, to minimize a systematic error, these radii had to be scaled up empirically by 7%. The final model predicted solvation enthalpies that were remarkably accurate—to within a few per cent for 31 ions, which ranged in charge from −1e to +4e where e is the electron charge. For more complex molecules, the Honig group introduced in 1994 the widely used PARSE (parameters for solvation energy) parameters, atomic radii and charges for many chemical groups relevant to biological simulations [35]. These parameters were defined so that calculated solvation free energies of small molecules matched experimental measurements. The dielectric constants for the molecular interior were set to 1, although Honig et al discussed the use of ε = 2 for continuum calculations.

Somewhat later, as MD-based free-energy methods [265] and binding calculations [65] became more popular, it became imperative to develop implicit-solvent models that were properly balanced (i.e. consistent) with MD force fields. One shortcoming of empirical implicit-solvent models is that the solvation free energies are not necessarily suitable for comparison to intramolecular energies such as bond stretching and van der Waals interactions. The continual increase in computer processing power has made explicit-solvent MD free-energy perturbation (FEP) methods a popular approach to calculating PMFs, which are then used as reference data to fit the parameters for implicit-solvent models. The first FEP-derived parameter set, by Nina et al [266, 267], addressed proteins in the CHARMM force field [2], but similar strategies have since been employed for other classes of solutes and force fields [268271]. Chen et al [271] devoted extensive attention to the question of 'balancing' in their optimization of a GB model, following the standard PMF approach [36, 62, 272].

5.1. Challenges

The most elementary challenge in parameterization is, of course, to avoid overfitting. This problem should rarely, if ever, arise in implicit-solvent model parameterizations based on other computational results, because one can always generate more reference data. The more difficult task is evaluating what data the model must be able to predict and what it should be able to predict and what it should not be able to predict (that is, that success would have to be fortuitous). Because parameter fitting is inherently an ill-posed inverse problem, parameters must be fit to the most discriminating test cases that are available. For example, when gradients of potentials are of interest (forces [273, 274] or electric fields [259]), it is better (more stable) to parameterize against the gradients rather than the potentials themselves, because differentiation amplifies numerical error. Furthermore, semi-phenomenological approaches make transferability a serious problem; loosely speaking, a transferable parameter is one that can be fit for one application and still give accurate answers for others.

Although solvation free energies have been the most popular source of reference data, recent evidence suggests that more challenging data are needed if models are to reliably predict binding free energies. A study on estimating the electrostatic contributions to binding free energies illustrated that parameter variations can have substantial impact on the change in the electrostatic solvation free energy, also known as the 'desolvation penalty' [275]. Scarsi and Caflisch [276] recommended that parameterization should be conducted using binding free energies directly. Their work is especially noteworthy for its use of the Poisson model as the reference for the parameterization of faster approximate models.

Rankin et al [277] demonstrated the severity of the problem by conducting an exhaustive search in a reduced parameter space with three variables: a single scale factor for atomic radii (using AMBER as the reference radii), the solute dielectric constant, and, for the nonpolar solvation component, one coefficient of proportionality for the surface-area term. Using the experimental solvation free energies of over 200 small molecules and their BEM software for the Poisson model [152], they found that a surprising range of parameter values fit the data equally well; in fact, the nonpolar term could have been dropped entirely without substantially altering their results. A clear example of problems with transferability could be found in their search space, which included the AMBER charges with PARSE radii; predictions were substantially poorer than with other models.

More generally, for a fixed value of the nonpolar coefficient, there exists a compensating effect between reduced radii and increased dielectric constant that leads to a meaningful 'valley' in the parameter space, i.e. instead of a deep local minimum, there is a region of parameter space in which every point gives essentially the same agreement with experimental solvation free energies. However, these points give very different predictions for binding free energies, spanning almost 25 kcal mol−1 in predicted affinity. These empirical results showed, following earlier work, that the electrostatic contribution to binding depends much more strongly on the dielectric constant than does the solvation free energy. The authors offered an incisive demonstration of this property using the analytically solvable system of two spherical molecules binding, so the smaller is completely buried in the bound state. Their results led them to state, 'Parameter sets that yield equivalent results for transfer free energies yield highly disparate results for binding free energies. Hence, parameterization of a solvation model against hydration data, although a necessary condition, is not sufficient to validate its use for calculating binding free energies in solution' (p 959 of [277]).

Similar concerns have been raised for GB models [278]. Michel et al used genetic algorithms to search a high-dimensional parameter space and generate multiple candidates with equivalently good fits. These calculations led to parameter sets with noticeably different parameters, but almost equally good fits for solvation free energies, finding that some 'better' models actually achieved their accuracy at the expense of predicting nonphysical positive electrostatic solvation free energies. As the authors note that GB models are often used in molecular binding applications, they also tested how well their optimized parameters reproduced PMFs of the approach between small molecules; again, equally good parameterizations for solvation free energies were insufficient for accuracy in the binding-type application.

Finally, MD force fields continue to evolve—for instance, adding atomic polarizability [279282]—so it is important that the community develop standardized, automated methods and software for parameterizing the large number of implicit-solvent models. A still more complex challenge, with important applications in drug design and analysis, is the problem of developing general parameter sets suitable for drug-like compounds, which are many times more diverse than the most commonly simulated biological molecules (proteins, nucleic acids, lipids and carbohydrates) [283]. MD force fields for this enormous space of molecules have emerged relatively recently, including OPLS-AA, AMBER and CHARMM force fields [284287]. Therefore, large-scale optimization methods such as implicit-solvent parameterization ought to be an active area, particularly in testing new and improved solvation models.

6. Room for improvement: model limitations

The simplistic form of most implicit-solvent models makes it impossible, of course, to capture all types of solute–solvent interactions exactly. Although many models have been validated to semi-quantitative agreement against a range of data [35, 288290] (see also the review [291]), there remain several controversies regarding whether implicit-solvent models can capture certain phenomena even qualitatively correctly (e.g. [292] and references in [16]). Here, we address these briefly, mostly to provide some background for non-experts in introducing advanced solvent models in the following sections.

As noted in the introductory sections, this review has focused on implicit-solvent models as PMF-based approximations of explicit solvent models. In this setting, the solute dielectric constant should be either 1 or 2, depending on the treatment of polarization [35], because conformational relaxation is to be modeled explicitly. Models with more empirical character must address a much more complex set of questions about what phenomena the protein dielectric constant is supposed to address. Schutz and Warshel have reviewed these issues in detail [60], as have other reviews, e.g. [293], to establish that the choice of dielectric constant is problem-dependent. Numerous groups have argued that the notion of a uniform, isotropic dielectric is overly simplistic [294, 295]. A prominent example of this challenge may be found in the calculation of protein pKa shifts [61, 178, 296], where numerous groups have found that large dielectric constants (εp > 8) are needed to improve accuracy compared to experiments. This phenomenon is sometimes attributed to buried water [61] or to conformational flexibility, i.e. the empirical view of the model (see the discussion in [296]). Inclusion of explicit flexibility has improved predictions in many cases [297299], but has not yet resolved the problems entirely [300]. Lund et al [301] have presented an interesting analysis of the fundamental issues using spherical model geometries. Similar to the pKa problem, high dielectric constants and non-uniform dielectric constants have been noted as important for accurate modeling of electron transfer [185].

The definition of the dielectric boundary is also reasonably arbitrary. Of course, it is possible to fit atomic radii to reproduce electrostatic solvation free energies from experiments [35] or from MD [266]. However, these energies can be highly sensitive to the choice of the definition [16, 266]. The model protein–protein system of barnase and barstar provides a specific example of the type of controversies that can emerge from different approaches to the boundary definition [302305]. In a much more general study, Papazyan and Warshel [306, 307] used their dipole–lattice model to illustrate several problematic aspects of using cavity definitions, also relating their dipole–lattice model to Poisson-based approaches [308]. Much more recently, a standard Poisson study by Tjong and Zhou showed that the differences in solvation free energies computed using vdW and molecular surfaces are size dependent [309]. These issues have motivated the development of several energy-functional implicit-solvent models in which the 'boundary' is not defined explicitly at all, but is instead an output of the model (e.g. [72, 310, 311]).

Many model comparisons have been made to validate new models or establish relative strengths and weaknesses of popular ones [73, 263, 312325]. We omit cataloging details of these extensive comparisons because several reviews have addressed their findings previously [11, 12, 15, 16, 21].

6.1. Linear-response theory

As noted in equation (9), the standard Poisson theory treats water as a linear dielectric. Numerous results from both experiment and MD suggest that this assumption is adequate in many cases, with numerous failure modes of varying subtlety and means for remedy. In particular, for monovalent ions, linear response appears to hold quite well [34, 288, 289, 326]; Rick and Berne [288] also comment on the problems with arbitrary cavity definitions, with findings similar to that discussed by Papazyan and Warshel [306].

Aqvist and Hansson approached their analysis of linear response in implicit-solvent models from the PMF perspective [327]. They follow Zwanzig's approach to obtain the free-energy difference between the uncharged and charged states of the solute. They make an important observation that the assumption of linear response implies that the electrostatic potential in the uncharged solute is zero (or negligible), and go beyond this point to derive thermodynamic invariants for linear-response theory and for the case when the potential in the uncharged solute is not zero [327]. To illustrate their theoretical analysis they study electrostatic solvation free energies of numerous solutes (both polar and ionic) in several solvents. Also, in discussing the implications of the linear-response assumption, they note that hydration-layer waters do exhibit net orientation [288, 326], and the arbitrariness of the cavity definition. They also make the important comment, 'Even if violations of the linear response assumption are found, the deviations might turn out to behave in a systematic way which could still be utilized for simplified free-energy calculations' (p 9512). Challenging this assumption, Cerutti et al [328] found via extensive MD simulations that the net electrostatic potential in an uncharged protein is significant, rather than close to zero; their results seem to hold several important implications for implicit-solvent models, and merit much more study.

Lin and Pettitt [329] have also addressed the assumption of zero potential in the uncharged state. Their work provided a complementary source of simulation data, as they studied amino-acid side-chain analogs using highly accurate continuum-model calculations, replica-exchange methods and an explicit-solvent linear-response approximation. They found noticeable deviations for highly polar molecules (analogs for asparagine and histidine), which occurred near the uncharged states. However, the total energies estimated by linear-response methods were not grossly in error, and furthermore, the relative energies were quite good; as the authors noted, these relative energies are more often the relevant quantities for applications involving protein mutations or molecular design. Matyushov et al [330] have also shown that the potential inside an uncharged cavity is not zero, using atomistic simulations and a model spherical solute called a 'Rossky cavity'. The authors suggested that the standard Maxwell dielectric boundary conditions are in error because the solvent molecules are free to rotate (for instance, in water, to improve hydrogen bonding). A small applied field—in their case, applied externally—does not necessarily re-orient the waters as macroscopic dielectric theory holds them. Their MD simulations indicated deviations from the induced surface charge predicted by standard Poisson theory, and the authors indicate that their results better match a 'Lorentz virtual cavity' model. It is unclear whether one might find modified Poisson boundary conditions for capturing these effects. Kimura et al [331] have used a novel discretecharge approximation to address related shortcomings, but more work is needed. One of the notable arguments raised by Matyushov and co-workers [330] is the case for a wider range of macroscopic observables in the validation of models: they show that the standard Poisson model and their Lorentz-cavity approach predict quite different results for dielectric constants of mixtures.

6.2. Lack of molecular detail

The standard Poisson model is obviously incapable of predicting atomic-level interactions between solvent and solute. For example, it cannot address how water molecules can 'bridge' two protein side chains by hydrogen bonding to both, thereby stabilizing them at a distance beyond direct contact but before continuum-level theory would be appropriate. Still more complex failures can occur in confined spaces such as enzyme active sites; even at nearly planar interfaces such as lipid bilayers, water structure leads to failures of linear response and local dielectric response [332]. Many results indicate that only a few layers of surrounding solvent are needed to substantially improve model realism [332], which has motivated the development of several semi-implicit solvation models to include these few layers. Gouda et al [42] demonstrated the impact of first-solvation-shell errors on binding free-energy estimation. Reduced screening due to solvent structure has motivated the development of nonlocal dielectric theory (see the following section). However, even these theories assume zero potential in the charged state, and as the work of Cerutti et al [328] demonstrates, solvent structure can indeed create a sizable electrostatic potential inside an uncharged protein. We note that solvent structure is captured in a natural way by more complex statistical–mechanical integral-equation theories [84, 333335], even though such models are outside the scope of this review.

6.3. Ionic solutions

The PB equation's key limitations are well known, e.g. [336, 337], and easily stated. Firstly, the model assumes that the ions are point charges, which they obviously are not; secondly, that the ions are in equilibrium (thus the Boltzmann weighting for their concentrations) with all correlations between them neglected. An important consequence of assuming ions to be point charges is that predicted ion concentrations can be unphysically large [338]. A classical empirical correction, due to Stern, is to enforce an ion-exclusion layer of the ion radius near charged surfaces [339]. Lastly, the linearized form of the PB equation is only a justifiable approximation for very small electrostatic potentials, a condition which is violated near highly charged species such as DNA or protein active sites.

7. Advanced models for water behavior

Clearly, real water response to a molecular charge distribution includes nonlinear effects as well as nonlocal ones. The familiar Poisson framework is linear and local, so the theoretical challenge is to identify simple reduced models that are (a) accurate enough for detailed biological science, e.g. understanding protein function or how mutations lead to pathogenesis; (b) fast enough to be used in dynamics or Monte Carlo simulations; and (c) based on a mathematical framework suitable for multi-scale modeling, parallel computing or molecular design.

Warshel and Levitt's protein-dipole–Langevin-dipole (PDLD) model represents an early advanced implicit-solvent theory [340] that includes both nonlinear and nonlocal effects. In later work, Warshel and colleagues have made a number of important advances in solvation modeling [340345], demonstrated the implications of solvation for molecular function and engineering [346348] and also highlighted the shortcomings of implicit-solvent models in numerous reviews [3, 20, 60, 349, 350]. Along similar lines to PDLD, but in a standard PDE framework, Sandberg and Edholm have developed the Langevin–Debye (LD) model for water's polarizability and applied it to molecular solvation [351, 352]. In the LD model, the polarization field density P(r) is defined in terms of not only the usual electric field E(r) but also a locally reduced electric field F(r), so that the overall constitutive relation is given by the equations

Equation (36)

Equation (37)

Equation (38)

where the Langevin function is

Equation (39)

with β = (kBT)−1,

Equation (40)

α0 is the molecular electric polarizability, μ the permanent dipole moment of water, n the number density of water molecules and Cμ and g correction factors from Onsager and Kirkwood (for more details, see [353]). One solves equations (36)–(38) numerically to determine the actual dielectric relationship D(r) = fLD(E(r)). We note at this point that although the LD model is a nonlinear theory, it is a local one: the displacement field at a point r is determined purely by the electric field at r. Freed and co-workers expanded the LD model to study charge burial in proteins, showing that the new theory gives markedly reduced penalties [353356]. Thus, nonlinear dielectric behavior may help explain the long-running controversy over charge burial [302, 354].

The nonlocal continuum models of Dogonadze [357] and Kornyshev et al [358] represent the other axis, extending the Poisson theory to nonlocal dielectric response while preserving the appealing mathematical properties of linear response, e.g. reciprocity. The basic idea of the nonlocal model is that the displacement field at r is no longer purely a function of the electric field there; instead, we have the nonlocal relationship

Equation (41)

where the integral is taken over all space and ${\hbox{\fontencoding{OT1}\fontfamily{pzc}\fontsize{13.5}{14}\selectfont E}}(\mathbf {r},\mathbf {r}')$ is the nonlocal dielectric function. The essential idea of nonlocal models is that we are interested in fields on length scales that are not 'well separated' from the length scale of the constituent material, i.e. the usual continuum assumption. As a result, nonlocal models are appearing with increasing frequency throughout molecular and nano-science, with applications in optical properties of nanostructures [359], liquid crystals [360] and the elastic properties of carbon nanotubes [361].

Combining the constitutive relationship equation (41) with the displacement-field form of Gauss's law (∇·D = ρ), we find that in nonlocal electrostatics the Poisson equation is an integrodifferential equation

Equation (42)

Note that equation (41) has absolutely nothing in common with the widely used notion of a 'distance-dependent dielectric' wherein the potential at rj due to a unit point charge at ri is given by

Equation (43)

with rij = ∥rj − ri∥. Nonlocal dielectric response represents a well-founded continuum theory, whereas the latter is merely a useful and very fast computational heuristic.

Equation (42) is much more complex to solve than the local PDE Poisson equation, and for many years the only studies using nonlocal dielectric theory relied on analytically tractable model geometries. Kornyshev et al [358] applied the theory widely to study problems at the interface between planar electrodes and surrounding electrolytes, Basilevsky and Parsons [362] studied the solvation of spherical ions, Scott et al [363] addressed ions in carbon nanotubes, and Rubinstein and Sherman modeled charge burial and molecular binding using planar interfaces [364366]. Early models tended to use a complicated form of the nonlocal dielectric function, but more recently, a simpler version known as the Lorentz model has become favored (see the discussion in [362] and criticisms of the model in [367]), so that the dielectric function for both r, r' in solvent is

Equation (44)

where ε is the short-range (optical) dielectric constant, usually taken as ∼2 and λ is a length scale parameter associated with the nonlocality of response. It is important to note that this model exactly recovers the standard local Poisson model in the limit λ → 0.

However, even this simplest nonlocal model poses severe challenges due to the integrodifferential Poisson equation. Hildebrandt et al [368] made a substantial breakthrough in recognizing that the nonlocal component of the Lorentz response (the second term of equation (44)) is a Green's function for the Yukawa (LPB) equation. This observation allows an important reformulation of the problem into a set of coupled local PDEs, and this reformulated problem can be solved very efficiently using standard numerical techniques; analytical studies are also simplified by the local reformulation [369]. Hildebrandt et al illustrated a boundary-integral method for nonlocal electrostatics [370] as well as an immersed interface method [371]. Bardhan and Hildebrandt have recently shown that standard fast BEM solvers (based on the work of Altman et al [127, 174]) can solve the nonlocal problem in about the same time as required for standard local-response LPB problems [372]. Applying the nonlocal model and BEM to the problem of charge burial, Bardhan illustrated that nonlocal dielectric behavior substantially reduces burial penalties, which may have implications for protein pKa calculations or binding free energies [373]. However, the model is still immature given the very recent development of theory to solve it efficiently, and parameterization remains a critical subject for future work in the area.

There are many models that address both nonlinearity and nonlocality of solvent response. Particularly noteworthy are the recent papers of Orland et al on a dipolar Poisson model [234237, 374] that captures some aspects of the molecular nature of the solvent, yet can be addressed with widely available, and massively scalable, solvers for continuum PDE-based models (the authors have implemented their model using the multigrid finite-difference package underlying APBS [32]). Additional models that are both nonlinear and nonlocal may be found in recent energy functional theories [375], sometimes known as density-functional theories (DFT) [376381], variational models [72] and statistical–mechanical integral equations [84, 333335, 382386]. Some advanced numerical algorithms have recently been presented for the latter [387, 388].

8. Advanced models for electrolyte behavior

Theories of electrolyte solutions represent a large body of literature (reviewed in [336, 337, 339, 389]), and here we merely survey the basic nature of proposed extensions to the PB model. An early extension, due in part to Kirkwood [390] and others, included ion–ion correlations using a set of nonlinearly coupled PDE problems. Outhwaite and Bhuiyan have developed this theory extensively for model systems [391393], and Tomac and Gräslund [394] have presented a multigrid solver for the modified PB model. Much more recently, Borukhov et al [338] proposed a modified PB to account for finite ion sizes; for a symmetric monovalent electrolyte with both cation and anion diameters equal to a, they obtain the PB-like equation

Equation (45)

for the potential in the solvent, where cb is the bulk concentration and ϕ0 = 2a3cb. This model, later extended by Lu and Zhou [395] for more realistic models with unequal ion radii, exactly simplifies to the usual NLPB for low salt concentrations, because then ϕ0 → 0; however, it is still a local theory. Doniach and co-workers have extended this model to unequal ion sizes [396], as have Li and co-workers from [397, 398]. Particularly noteworthy are the computational approaches employed by these groups: Doniach et al adapted the APBS solver [32], highlighting the value of releasing open-source software with sufficient documentation for others to read the code and make substantial modifications. Li et al adopt a constrained-optimization approach [399] that is advantageous for their variational models.

Antypov et al [400] used a density-functional framework to compare both local and nonlocal methods for adding finite-size effects to PB theory. A principal finding was that the local functionals (one of which was a lattice-gas approach similar to Borukhov's [338]) could not match Monte Carlo data as accurately as the nonlocal functionals. Gillespie et al [401] have also demonstrated the importance of nonlocality in predicting charge inversion in their comparison of Rosenfeld-type DFT to standard PB and the Bikerman equation (a local modified PB theory). Important biological applications of DFT have been shown in predicting the selectivity of ion channels [402404] and lipid bilayer systems [405, 406]. A review of classical DFT based on Rosenfeld's work [378380, 407409] is outside the scope of this paper, particularly because the computational treatment involves several challenges unlike those in standard PB models [410412]. For details, see the several recent reviews on Rosenfeld-type DFT [413416].

9. Challenges and promising developments

9.1. Semi-implicit solvent models

Implicit-solvent models are often used to avoid the large cost of explicit-solvent calculations, which are dominated by thousands of solvent molecules. Relatively few of these solvent molecules actually interact directly with the solute, i.e. the water and ions in the hydration layer immediately surrounding it. As one of the noted weaknesses of implicit-solvent models has been its lack of molecular realism, it is of interest to explore a 'best of both worlds' approach in which a small number of explicit-solvent molecules are modeled near the solute, with an implicit-solvent model employed outside the explicit region [417, 418]. Another advantage is that these models also eliminate the need for periodic boundary conditions in explicit-solvent calculations. Beglov and Roux [418] derived a solvent boundary potential model for a sphere whose radius varies with time to include all of the explicit-solvent molecules. The strong fundamentals in their approach allowed application to a variety of problems, including the computationally expensive parameterization of implicit-solvent models [266]. Later extensions allowed much more general boundary shapes [419, 420] and application to quantum-mechanical simulations [421]. A more recent approach was presented by Wagoner and Pande [422], who developed a semi-implicit model in which the solvent molecules are smoothly decoupled from the system as they approach a spherical boundary; other methods have used GB electrostatics [423] and BIE methods [424]. Recently, Li et al [425] introduced the 'elastic bag' model which allows a complex boundary to vary dynamically. These models and other studies suggest that in many cases, fewer than ten 'solvent shells' are needed to recover the results of fully explicit-solvent calculations [332, 424].

Semi-explicit models for electrolyte solutions have also been an active area [54, 426430]. In these models, the water is treated implicitly (as a high-dielectric continuum) but the ions in solution are treated explicitly as charged hard spheres. This approach eliminates important weaknesses of the PB theories, namely that the ions' excluded volume is handled properly and ion–ion correlations are included naturally. However, these advantages are balanced by the need to compute statistical averages over the ions' degrees of freedom, often using Monte Carlo sampling [54, 426], but sometimes with Langevin dynamics [427]. These types of electrolyte models have been studied for decades in the physical chemistry community, where they are known as primitive-model electrolytes [431, 432]. These primitive-model calculations have been very successful in predicting the selectivity of some ion-channel proteins [429, 430].

9.2. High-performance parallel computing

Perhaps the two most widely known successes for massive parallelism in molecular simulations are MD [30, 433] and scalable FEMs [32]. However, promising results have been demonstrated for many other theories, including statistical–mechanical integral equations [388], DFT [411], BEM simulations [187] and GB-type models [434]. Significant effort is being made to port these algorithms to an emerging computing platform called graphics processing units (GPUs), and in this section we address how GPU computing may influence the evolution of implicit-solvent models.

For years, the data-processing capabilities of high-end graphics cards have grown significantly faster than the capabilities of standard microprocessors (CPU). Most graphics applications require fairly simple computations applied independently to a large number of data elements, for instance applying the same rotation to millions of triangles. This independence means that the overall computation is massively parallel, and so GPUs combine a large number of relatively simple compute cores, each of which applies the same computational kernel to a single element. Many problems in computational biology seem to map naturally to this hardware, and demonstrated applications include bioinformatics, estimating molecular binding affinities [435] and MD in explicit and implicit solvent [436438]. In some problems, a GPU code can be dozens or even hundreds of times faster than the corresponding CPU code. However, before the exciting potential of these computing platforms can be reached for arbitrary applications, a GPU implementation must resolve several practical complications.

The main price for this dramatic performance improvement is code complexity, especially for scalability to clusters of GPU-enabled computers. Consider how many software packages are designed to run on a single CPU or at most a small cluster of CPUs, and how few applications can make effective use of massively parallel computers (due to the need for complicated analysis of what limits scalability). Single GPU computing is considerably more complex than single CPU computing, so very few algorithms can make effective use of even two GPUs attached to the same CPU (node)—much less exploit hundreds of nodes which each have one or more GPUs. Computational scientists specializing in parallel computing should therefore be integral team members for any serious GPU implementation.

Even when implementing an algorithm for a single GPU on a single node, there are significant complications that argue for collaboration with GPU experts and computer scientists. First, modern GPUs are fast enough that data movement is usually a bottleneck, i.e. it is hard to keep the compute cores busy all the time because data cannot be transferred fast enough to them. Algorithm experts can often improve performance without sacrificing calculation accuracy by minimizing data movement. A second complication arises because most compute cores are substantially faster at single-precision floating point arithmetic than they are at double precision. In pursuit of performance, many GPU implementations of existing software have moved from double to single precisions; the user should be careful to check that the applications of interest have been tested thoroughly for subtle issues that can arise (see, e.g., the thorough analysis by Elber and co-workers [439]).

To be sure, at some point, we may arrive at new mathematical models due to novel insights into algorithms that run efficiently on GPUs. In the short run, however, we must resist the temptation to introduce new, modified implicit-solvent models simply because they can be optimized to run extremely quickly on a single GPU. Firstly, unless the algorithm is implemented within an existing code, e.g. [440], coupling codes is a time-consuming and error-prone task. Secondly, GPU hardware continues to evolve rapidly, which risks an early obsolescence for proposed models. Thirdly, non-scalable GPU algorithms will not be suitable for the increasingly large biological systems of current study. Thus, a new model that runs on one GPU but does not scale to multiple GPUs (many workstations hold two or even four), or clusters of GPU-enabled computers, is of limited value. For these reasons, it seems prudent to focus GPU development on solidly understood and accepted models (and algorithms).

9.3. High-level mathematical abstractions for testing new solvation models

Why is defining a new mathematical model of solvation so much easier than solving it numerically? This state of affairs is as lamentable as it is obvious—yet the disparity is not in any sense 'fundamental'. In this final section, I would like to point out an emerging paradigm of automated code generation that may ultimately make testing new implicit-solvent models as simple as writing down the governing equations.

Admittedly, at first glance it appears that each new mathematical model would require a new implementation—usually a slow and complicated business, even if one is only modifying an existing code. Compounding the difficulty, the new implementation must be fast enough to run on problems of current scientific interest, and as always, parallelism significantly increases complexity. Thus, the computational chemist/biophysicist developing more realistic models seems to face an unavoidable reality that the vast majority of her efforts will address mundane but complicated problems of software engineering and numerical analysis. Similar problems are faced by scientists across many disciplines and length scales that range from the molecular to the global, including cardiac electrophysiology and geological dynamics (to name just two examples).

Computational methods built around high-level mathematical interfaces offer a powerful way to address these challenges. Just as high-level programming interfaces streamline the design of sophisticated new algorithms, high-level mathematical interfaces streamline the design of sophisticated new models. Many scientists in the computational biophysics and biochemistry communities are familiar with an excellent example without realizing it: the APBS software [32] is built on top of the extremely general FETk software toolkit developed by M Holst [441]. The FEniCS [442446] and Sundance [447] projects also address the problem very generally, by automating the generation of finite-element simulation codes for continuum-theory models. In FEniCS, the user needs only to specify the so-called weak form of the PDE, along with basic aspects of the numerical method, e.g. how the finite-element mesh is to be used as input, and what basis functions should be used. These are specified at a high level of mathematical abstraction in either C + + or Python. The 'model compiler' then automatically generates a parallel FEM code for computing the model [448, 449]. Numerous applications show that these solvers already scale to several hundreds of nodes and improved scalability is a priority in ongoing work (J Hake, personal communication); such a parallelism should be more than adequate for many basic modeling studies in the near future. A demonstration calculation of turbulence in computational fluid dynamics illustrates the capabilities of high-level modeling flexibility [450], and recent work extends these tools from traditional FEM to newer discontinuous Galerkin FEM [451].

The main mechanism for the success of FEniCS and related projects such as Sundance is that the developers were willing to circumscribe a class of models (specifically, those based on weak forms) and then developed automatic methods to solve that whole class. Importantly, they did so while retaining a programming interface that allows users to solve closely related problems. The point is not that the present version of any of these libraries will solve all of our implicit-solvent models, of course. Instead, consider these ideas as important steps to let the scientist explore an important part of 'model space' efficiently. Sundance and FEniCS address models based on standard PDEs, but one could imagine similar tools for variational models, which have become increasingly popular recently [72, 375], classical DFT [378380, 403, 404, 408, 412, 415, 416], or statistical–mechanical integral equations [84, 382, 384, 386]. With such a software framework, the scientist can stay focused on the application science itself, rather than spending time implementing complex numerical algorithms on sophisticated hardware. Many chemists are already familiar with the MADNESS software [452], which uses wavelets for electronic structure methods; it has been constructed to be very general, and could be a suitable starting point.

10. Summary

Biological solvents—aqueous electrolytes with tightly controlled ionic composition—exhibit extremely complex behavior, especially in their physiologically critical influence on the function of biological molecules. However, the substantial computational cost of modeling solvent molecules explicitly, even with classical, nonpolarizable all-atom MD simulations, continues to motivate the biophysics community to develop faster (and generally, much more approximate) models. Most of these new models and approximations derive from novel physical insights, whose largely empirical nature poses severe challenges for transferring and combining these insights between models. Optimizing the rate of progress demands a robust 'model ecosystem' that readily permits model testing, re-combination and improvement. Obviously, any such ecosystem must rest on a unifying foundation of rigorous mathematics. Of course, this is not to say that only one such foundation exists for the task—three present themselves immediately. Classical continuum Poisson-based models represent one, encompassing many dielectric modifications, PB theory and its modifications, and most of the GB models. Variational/density-functional approaches represent another foundation, and liquid-state integral equations a third.

New computing technologies, both hardware and software, promise dramatic improvements to modeling capabilities. Capitalizing on these opportunities will require us to meet two fundamentally interdisciplinary challenges: (i) translating chemical and physical insights into the relevant mathematical formalisms for these (and future) mathematical foundations; (ii) collaborating with computational specialists in high-performance parallel computing, to develop massively scalable, general software for solving wide classes of these models.

Acknowledgments

This work was supported in part by a Rush University New Investigator award. It is a pleasure to acknowledge valuable comments and insights from numerous members of the Computational Science Graduate Fellowship (CSGF) community, including A E Ismail, C Rinderspacher, M Radhakrishnan, M Reuter and especially G Oxberry. The CSGF program is supported by the US Department of Energy and administered by the Krell Institute. The author is indebted to R S Eisenberg for creating an environment conducive to bridging the biophysics and computational mathematics communities.

Footnotes

  • See [1].

Please wait… references are loading.