Abstract
We show that charge-sign-dependent asymmetric hydration can be modeled accurately using linear Poisson theory after replacing the standard electric-displacement boundary condition with a simple nonlinear boundary condition. Using a single multiplicative scaling factor to determine atomic radii from molecular dynamics Lennard-Jones parameters, the new model accurately reproduces MD free-energy calculations of hydration asymmetries for: (i) monatomic ions, (ii) titratable amino acids in both their protonated and unprotonated states, and (iii) the Mobley “bracelet” and “rod” test problems [D. L. Mobley, A. E. Barber II, C. J. Fennell, and K. A. Dill, “Charge asymmetries in hydration of polar solutes,” J. Phys. Chem. B 112, 2405–2414 (2008)]. Remarkably, the model also justifies the use of linear response expressions for charging free energies. Our boundary-element method implementation demonstrates the ease with which other continuum-electrostatic solvers can be extended to include asymmetry.
I. INTRODUCTION
Implicit-solvent models represent an intuitive and fast approach to understand molecular solvation,1–4 and have a rigorous statistical-mechanical interpretation as an approximation to the potential of mean force (PMF) experienced by a molecular solute due to the surrounding solvent molecules.1 The PMF is usually decomposed into non-polar and electrostatic terms, the latter of which are often modeled using macroscopic continuum models based on the Poisson–Boltzmann partial-differential equation (PDE). Continuum models approximate the free energy required to grow the solute charge distribution into the solute cavity.1,5–13 Although implicit-solvent models can be orders of magnitude faster than explicit-solvent molecular-dynamics (MD) simulations, most popular continuum theories ignore numerous potentially important effects, including solvent molecules' finite size and specific molecular interactions such as hydrogen bonding (the AGBNP2 model, which addresses the latter, is a notable exception14).
One of the Poisson model's most perplexing and long-standing shortcomings is the difficulty of extending it to model charge-sign asymmetric solvation: for example, given two monatomic ions of equal radius, one of +q charge and the other of −q, the negative charge experiences stronger interactions with the solvent (more negative solvation free energy).6,15–21 However, standard Poisson models are charge-sign symmetric; that is, they predict the same solvation free energy for ±q. The need to include asymmetric effects is difficult to exaggerate, particularly in biological contexts. Consider that the protein avidin binds its ligand biotin with a binding free energy of approximately −20 kcal/mol, one of the most favorable in biology;22 solvent-exposed +1e and −1e charges can experience as much as 40 kcal/mol difference in their solvation free energies.21 Dominant factors in charge-sign asymmetric response include the liquid-vapor interface potential16,18 and the fact that water hydrogens can approach a negative solute charge closer than water oxygens can approach a positive one.6,15,20,21 Spherical solutes with central charges provide a useful data set for developing an understanding of size- and charge-sign dependent hydration, including the characterization of interface potentials, solvent packing, and dielectric saturation.15,16,18,19,23,24 These analyses and the continuum macroscopic-dielectric framework suggest that improvements require a more detailed, accurate representation of the solvent dipole field P(r),25,26 or, equivalently, the solvent charge density ρinduced(r) = ∇ · P(r). Because P and ρinduced do not respond linearly to the solute charge distribution,27 particularly in the first solvent shell,28 many groups have developed solvent models in which the solvent potential obeys a nonlinear PDE.29–33 Unfortunately, most of these models are still charge-sign symmetric.
However, in 1939 Latimer et al. proposed an approach to increase or decrease an ion's radius based on the charge,6 and recent developments in high-performance computing and explicit-solvent MD free-energy calculations provide important new data to extend this approach. Mobley et al. constructed a challenging test set and conducted extensive MD simulations on charge-sign asymmetry,34 enabling important new developments in modeling asymmetry28,35,36 that extend Latimer's6 work to Generalized-Born (GB) models of complex solutes. GB theory was a natural setting for these developments because Latimer's6 work and GB theory share the conceptual picture of an effective atomic radius. These early studies provided an important insight: a buried charge still affects the electric field at the boundary, so merely parameterizing charge-dependent radii cannot (indeed, should not) provide a satisfactory explanation. The accuracy of asymmetric GB models suggests that a simple Poisson-based model exists, but finding one has proven to be surprisingly difficult.
In this paper, we propose a simple Poisson continuum model that includes charge-sign asymmetry and show that it is remarkably accurate even without parameterization on an atom-by-atom basis. The key feature of our theory is a nonlinear boundary condition (NLBC) for the normal displacement field; in contrast, the displacement boundary condition for the standard (symmetric) Poisson theory is linear.37 Importantly, even though our proposed displacement boundary condition is nonlinear, the electrostatic potential in the solvent and solute volumes still satisfy linear Poisson/Laplace equations. Two phenomena motivated us to propose a nonlinear boundary condition instead of a nonlinear governing equation. First, numerous results illustrate that the solute reaction potential obeys nearly linear response even though the solvent charge distribution does not.21,28,35,36,38,39 For example, the new asymmetric GB models use the charge distribution only to modify the Born radii, with the overall energy still computed using superposition (independent sum of individual charge responses).28,35,36 Furthermore, we found in our previous work that the solute reaction potential is essentially a piecewise-linear function of charge,21 i.e., the proportionality coefficient depends on whether one is charging an ion from zero to +q or to −q. In fact, we began this work seeking primarily to reproduce this curiously simple nonlinearity.
The second phenomenon motivating our NLBC approach is the fact that the solute reaction potential is a harmonic field—that is, it satisfies the Laplace equation. This property is useful for numerical computations40,41 and also provides a path to improve models via boundary-integral methods:42 harmonicity means that regardless of the solvent model of interest, there exists some surface charge density that reproduces the reaction potential inside. For a given solvent model, the surface charge density might satisfy a nonlinear boundary-integral equation, but the very fact that such a density always exists suggests that one might improve continuum models by adding nonlinear terms to widely used BIE formalisms.43–46
II. CONTINUUM MODEL AND EXTENSION TO NONLINEAR BOUNDARY CONDITIONS
We first present the standard (charge-sign symmetric) Poisson electrostatic model and then describe the difference between it and our proposed NLBC model. In both theories, the molecular solute is treated as a macroscopic linear dielectric continuum obeying the Poisson equation where r is a point in space, φ1(r) is the potential in the solute, ε1 is the relative permittivity, and the molecular charge distribution ρ(r) is a set of Nq point charges, i.e., . The solute and solvent are separated by the interface Γ, and the solvent exterior is a linear dielectric with permittivity ε2 ≫ ε1, so the electric potential obeys ∇2φ2(r) = 0; note that modeling realistic biological solutions requires inclusion of screening effects due to mobile ions using, e.g., some form of the Poisson–Boltzmann equation for φ2(r).2,4 From macroscopic dielectric theory and Gauss's law, we obtain the standard Maxwell boundary conditions for :
| (1) |
| (2) |
where denotes the normal derivative (the normal at is defined pointing outward into solvent). Assuming that φ2(r) decays sufficiently quickly as |r| → ∞, this mixed-dielectric Poisson problem is well posed and the unknown potential φ1 can be rewritten as a linear boundary-integral equation for an unknown surface charge distribution on Γ. In particular, the apparent-surface charge (ASC) model44,46,47 (also known as the polarizable continuum model4,48) can be interpreted as finding an equivalent surface charge σ(r) in a homogeneous medium with permittivity ε1 everywhere. In this equivalent problem, the analogous boundary condition to Eq. (2) is simpler due to homogeneity, but adds a term for the surface charge
| (3) |
and we use to emphasize our use of an equivalent problem. Defining , one obtains
| (4) |
where and K is the normal electric field operator.46 The reaction potential in the solute is then , and φ1(r) = φREAC(r) + φCoulomb(r), with the latter term representing the Coulomb potential due to ρ(r).
The standard Maxwell displacement boundary condition (2) is obtained using Gauss's law in integral form and the fact that the divergence of the polarization field P(r) represents a volume charge density. However, near the solute–solvent boundary, the assumption that P(r) is pointwise proportional to the local electric field breaks down due to water structure at the interface; that is, it is no longer necessarily true that P(r) = (ε(r) − 1)E(r).
To model nonlinear solvent response at the boundary, we propose to replace the linear boundary condition, Eq. (2), with the phenomenological nonlinear boundary condition
| (5) |
where En is the electric field just inside Γ, i.e., , and
| (6) |
| (7) |
with α, β, and γ representing model parameters and μ = −αtanh ( − γ). The specification of μ ensures that h(En = 0) = 0, so that in the limit of weak electric fields, such as induced at the surface by a deeply buried charge, the boundary condition reduces to the familiar Poisson model. The NLBC leads to the modified, nonlinear BIE:
| (8) |
with the nonlinearity arising in the dependence of h on En (see the supplementary material72 for details on the numerical implementation).
One challenge in developing more accurate solvent models is the fact that nonlinear response49 generally requires a charging process,50 i.e., the expression no longer holds (L denotes the reaction-potential operator1,42). However, our previous work showed the remarkable fact that the solute reaction potential is piecewise linear, with the breakpoint at q = 0,21 so that φREAC = L+q for q > 0 and φREAC = L−q for q < 0, with L+ ≠ L−. The proposed NLBC in Eqs. (5) and (7) immediately explains this curious phenomenon: consider the limit β → ∞, so that tanh is constant everywhere, but discontinuous at q = 0. The Debye charging process51 scales all charges uniformly, i.e., , so the Coulomb field has the same sign for all finite λ. The Coulomb-field approximation (CFA) shows that the reaction field is nearly proportional to the direct Coulomb field, but slightly smaller in magnitude,42,52,53 so for finite λ, at almost all , the total field has the same sign as . This implies that almost everywhere on the surface, the tanh boundary condition takes its limiting (λ → 1) value for any finite λ, which means that the boundary condition is essentially linear: . With this justification, in this work we compute solvation free energies as . Note that a more precise definition of the charging free energy would be piecewise affine, because the charging free energy also includes a linear term that results from the liquid-vapor interface potential;21,54,55 as noted above, however, in the present work its influence is approximated via the offset parameter γ.
III. RESULTS AND DISCUSSION
We parameterized the NLBC model using the Mobley et al. MD free-energy calculations, who studied asymmetry using fictitious bracelet and rod molecules34 constructed from AMBER Cα atoms with Rmin/2 = 1.908 Å. We obtained optimal results with α = 0.5, β = −60, γ = −0.5, and a continuum-model Cα radius of 1.75 Å (a scale factor of approximately 0.92). Note that in this first exploration of the NLBC, we have parameterized against the overall solvation free energies computed by Mobley et al.34 rather than the more correct charging free energy.
Figure 1 plots NLBC and MD free-energy calculations for ion charging free energies; the MD charging simulations used the same protocol as our previous work21 (see the supplementary material72) and CHARMM Lennard-Jones parameters. We remind the reader that no additional parameters were fit in obtaining these NLBC results, i.e., ion radii were assigned Rion = 0.92Rmin/2. For additional data, ions were charged to both +1e and −1e, regardless of the charge on the real ion, and the NLBC accurately predicts these charging free energies as well. The largest deviations occur for radii less than 1.4 Å, where discrete packing effects and actual dielectric saturation are likely.
FIG. 1.
Asymmetric polarization free energies for a monovalent central charge in a sphere, as a function of sphere radius. The labeled symbols denote results from MD free-energy calculations charging CHARMM monatomic ions from zero to +1e or −1e, with Rion = 0.92Rmin/2. The dashed black curve in the middle is the (charge-sign symmetric) Born polarization free energy.
Calculations for the Mobley test set are summarized in Table I; Figures 1–8 in the supplementary material72 plot the NLBC and Mobley MD solvation free energies and asymmetry energies, and include illustrations of the test problems. The rod molecules are composed of 5 or 6 atoms along a line, with one atom possessing +1e charge, one with −1e, and the rest neutral. The asymmetry errors in Table I represent the difference in solvation energies when reversing the charged atoms' signs. The bracelet molecules are regular polygons with between 3 and 8 sides; atoms are at the vertices (1.4 Å apart). Bracelets were simulated with three charge distributions: the “opposing” case had a +1e charge neutralized by two −0.5e charges positioned symmetrically on the opposite side. The “distributed” case has one +1e charge and a neutralizing −1e distributed equally on all the other atoms; the “dipole” case is similar to “opposing” but fixes the dipole moment.34 Solvent charge-densities from the MD calculations34 suggest that solvent packing may be responsible for size-dependent deviations; parameterizing radii for actual atoms should significantly reduce these errors.
Table I.
Comparison of NLBC model to MD free-energy calculations of Mobley et al.34 for rod and bracelet molecule test set. All energies are in kcal/mol. See the supplementary material72 for detailed results.
| Solvation errors | Asymmetry errors | |||
|---|---|---|---|---|
| Problem | RMSE | Max. | RMS | Max. |
| Rods | 5.57 | 9.63 | 0.88 | 1.49 |
| Bracelets (opposing) | 2.88 | 6.10 | 2.04 | 3.08 |
| Bracelets (distributed) | 2.20 | 2.72 | 0.29 | 0.59 |
| Bracelets (dipole) | 2.67 | 3.52 | 0.85 | 1.09 |
To test the model on real but nonspherical molecules, we compared NLBC and MD charging free energies for isolated titratable amino acids in both protonated and unprotonated states (see the supplementary material72 for details on structure preparation). Parameters were from the CHARMM force field56 when available, with other protonation states defined so that the protonated and unprotonated states had the same number of atoms. The MD free-energy-perturbation (FEP) calculations used the same protocol as the ions,21 holding the solute rigid so that ε1 = 1 unambiguously.1 The deviations between our MD results and the MD calculations of Nina et al.57 are small compared to the energies of interest, and likely due to our use of: (i) periodic boundary conditions, (ii) a larger solvent box (1959 waters vs. 150), and (iii) slightly different backbone angles.
As in the ion and Mobley examples, the NLBC radii were defined by the scaling R = 0.92Rmin/2. The results in Figure 2 illustrate that the NLBC model correctly captures solvation free energies in both charge states, despite the fact that radii were not adjusted individually or even for the atomic charges. In contrast, standard Poisson model results computed using the Nina et al.57 or PARSE58 radii exhibit larger deviations, particularly for arginine, aspartic acid, cysteine, glutamic acid, and tyrosine. These data suggest that the differences between symmetric and asymmetric electrostatic models are robust with respect to radii (the PARSE calculations are merely suggestive because these calculations used the CHARMM charges; for consistent comparison to experiment, one should use PARSE charges with PARSE radii).
FIG. 2.
Comparison of NLBC model to explicit-solvent MD FEP calculations for titratable residues with neutral blocking groups. MD results from Nina et al.57 are shown where available; standard continuum model results are shown for Nina et al.57 radii and PARSE radii58 (using CHARMM charges).
IV. CONCLUSION
We have proposed a Poisson-based theory that models charge-sign-dependent asymmetries in electrostatic solvation free energies using a NLBC, while still using linear continuum theory in the solute and solvent volumes. The NLBC model accurately reproduces MD free-energy results for monatomic ions, the Mobley et al.34 bracelet and rod problems, and titratable residues, even though we have used charge-independent radii that were fixed by a single scaling factor applied to MD radii. Furthermore, the NLBC reduces smoothly to the standard Poisson model as the parameter α approaches zero. Finally, our boundary-element method implementation for non-trivial molecules demonstrates that the new model is easily implemented in numerical Poisson and Poisson–Boltzmann solvers.
Our introduction of a modified boundary condition to account for solvation-shell response follows a long history in continuum mechanics, where phenomenological techniques find applications in many areas of science and engineering to capture a particular physical behavior in continuum theory rather than modeling or deriving it from first principles.59–61 Non-equilibrium micro-scale gas flows offer a well-developed example: velocity-slip and temperature-jump boundary conditions are simplified phenomenological approaches to represent both non-equilibrium and gas-surface interaction effects occurring near solid walls. Such boundary conditions were first suggested in the 19th century by Maxwell62 and von Smoluchowski,63 respectively. More recent examples include the partitioning of minerals at phase boundaries in geophysics,64 tumor growth,65 the deformation of biological membranes,66 and thin electric double layers in electro-osmotic flow.67
Much as Beglov and Roux showed that solvent response approaches the linear Poisson model in the limit as the solvent molecule approaches zero size,68 our model emphasizes that the nonlinear response is generally localized in the first solvent shell. Conceptually, the nonlinear boundary condition penalizes negative surface charge because the larger water oxygen cannot approach a solute charge as closely as the water hydrogens can. From a boundary-integral point of view, this has the same effect as adjusting the atomic radii, an approach pioneered by Latimer et al.,6 and extended recently to GB models.20,28,35,36 Purisima's work is particularly relevant due to their use of surface-charge boundary-integral approach, adjusting GB radii using σ(r).28,35 Our work differs substantially from these approaches because we have included asymmetry directly in the underlying Poisson model.
The present theory can be extended in several important ways. First, the proposed NLBC model has only three parameters whose particular dependencies on solvent model have not yet been established theoretically. Second, it seems straightforward to include ionic screening via the Poisson–Boltzmann equation. Third, the proposed NLBC depends exclusively on the normal electric field; improved models might include local curvature or higher-order moments of the potential. Importantly, the latter could distinguish between small-magnitude charges near the surface, and larger charges further away.20 Fourth, water's length-scale-dependent dielectric behavior might be included using nonlocal electrostatics.8,24,69–71 The new model also does not necessarily capture specific hydrogen-bonding effects like AGBNP2 does,14 which motivates future work comparing the two approaches.
ACKNOWLEDGMENTS
The authors thank David Mobley for sharing detailed calculation results, Jed Brown and David Green for valuable discussions, and Matt Reuter for a critical reading of the paper. M.G.K. was partially supported by the (U.S.) Department of Energy (DOE), Office of Science, Advanced Scientific Computing Research, under Contract No. DE-AC02-06CH11357, and also National Science Foundation (NSF) Grant No. OCI-1147680. J.P.B. has been supported in part by the National Institute of General Medical Sciences (NIGMS) of the National Institutes of Health (NIH) under Award No. R21GM102642. The content is solely the responsibility of the authors, and does not necessarily represent the official views of the National Institutes of Health.
REFERENCES
- 1. Roux B. and Simonson T., “Implicit solvent models,” Biophys. Chem. 78, 1–20 (1999). 10.1016/S0301-4622(98)00226-9 [DOI] [PubMed] [Google Scholar]
- 2. Sharp K. A. and Honig B., “Electrostatic interactions in macromolecules: Theory and applications,” Annu. Rev. Biophys. Biophys. Chem. 19, 301–332 (1990). 10.1146/annurev.bb.19.060190.001505 [DOI] [PubMed] [Google Scholar]
- 3. Davis M. E. and McCammon J. A., “Electrostatics in biomolecular structure and dynamics,” Chem. Rev. 90, 509–521 (1990). 10.1021/cr00101a005 [DOI] [Google Scholar]
- 4. Tomasi J. and Persico M., “Molecular interactions in solution: An overview of methods based on continuous descriptions of the solvent,” Chem. Rev. 94, 2027–2094 (1994). 10.1021/cr00031a013 [DOI] [Google Scholar]
- 5. Kirkwood J. G., “Theory of solutions of molecules containing widely separated charges with special application to zwitterions,” J. Chem. Phys. 2, 351 (1934). 10.1063/1.1749489 [DOI] [Google Scholar]
- 6. Latimer W. M., Pitzer K. S., and Slansky C. M., “The free energy of hydration of gaseous ions, and the absolute potential of the normal calomel electrode,” J. Chem. Phys. 7, 108–112 (1939). 10.1063/1.1750387 [DOI] [Google Scholar]
- 7. Kornyshev A. A. and Sutmann G., “Nonlocal dielectric saturation in liquid water,” Phys. Rev. Lett. 79, 3435–3438 (1997). 10.1103/PhysRevLett.79.3435 [DOI] [PubMed] [Google Scholar]
- 8. Hildebrandt A., Blossey R., Rjasanow S., Kohlbacher O., and Lenhof H.-P., “Novel formulation of nonlocal electrostatics,” Phys. Rev. Lett. 93, 108104 (2004). 10.1103/PhysRevLett.93.108104 [DOI] [PubMed] [Google Scholar]
- 9. Rizzo R. C., Aynechi T., Case D. A., and Kuntz I. D., “Estimation of absolute free energies of hydration using continuum methods: Accuracy of partial charge models and optimization of nonpolar contributions,” J. Chem. Theory Comput. 2, 128–139 (2006). 10.1021/ct050097l [DOI] [PubMed] [Google Scholar]
- 10. Abrashkin A., Andelman D., and Orland H., “Dipolar Poisson–Boltzmann equation: Ions and dipoles close to charge interfaces,” Phys. Rev. Lett. 99, 077801 (2007). 10.1103/PhysRevLett.99.077801 [DOI] [PubMed] [Google Scholar]
- 11. Gong H. and Freed K. F., “Langevin–Debye model for nonlinear electrostatic screening of solvated ions,” Phys. Rev. Lett. 102, 057603 (2009). 10.1103/PhysRevLett.102.057603 [DOI] [PubMed] [Google Scholar]
- 12. Guo Z., Li B., Dzubiella J., Cheng L.-T., McCammon J. A., and Che J., “Evaluation of hydration free energy by level-set variational implicit-solvent model with Coulomb-field approximation,” J. Chem. Theory Comput. 9, 1778–1787 (2013). 10.1021/ct301087w [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13. Zhou S., Cheng L.-T., Dzubiella J., Li B., and McCammon J. A., “Variational implicit solvation with Poisson–Boltzmann theory,” J. Chem. Theory Comput. 10, 1454–1467 (2014). 10.1021/ct401058w [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14. Gallicchio E., Paris K., and Levy R. M., “The AGBNP2 implicit solvation model,” J. Chem. Theory Comput. 5, 2544–2564 (2009). 10.1021/ct900234u [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15. Rashin A. A. and Honig B., “Reevaluation of the Born model of ion hydration,” J. Phys. Chem. 89, 5588–5593 (1985). 10.1021/j100272a006 [DOI] [Google Scholar]
- 16. Ashbaugh H. S., “Convergence of molecular and macroscopic continuum descriptions of ion hydration,” J. Phys. Chem. B 104(31), 7235–7238 (2000). 10.1021/jp0015067 [DOI] [Google Scholar]
- 17. Lynden-Bell R. M., Rasaiah J. C., and Noworyta J. P., “Using simulation to study solvation in water,” Pure Appl. Chem. 73, 1721–1731 (2001). 10.1351/pac200173111721 [DOI] [Google Scholar]
- 18. Rajamani S., Ghosh T., and Garde S., “Size dependent ion hydration, its asymmetry, and convergence to macroscopic behavior,” J. Chem. Phys. 120, 4457 (2004). 10.1063/1.1644536 [DOI] [PubMed] [Google Scholar]
- 19. Grossfield A., “Dependence of ion hydration on the sign of the ion's charge,” J. Chem. Phys. 122, 024506 (2005). 10.1063/1.1829036 [DOI] [PubMed] [Google Scholar]
- 20. Mukhopadhyay A., Fenley A. T., Tolokh I. S., and Onufriev A. V., “Charge hydration asymmetry: The basic principle and how to use it to test and improve water models,” J. Phys. Chem. B 116, 9776–9783 (2012). 10.1021/jp305226j [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21. Bardhan J. P., Jungwirth P., and Makowski L., “Affine-response model of molecular solvation of ions: Accurate predictions of asymmetric charging free energies,” J. Chem. Phys. 137, 124101 (2012). 10.1063/1.4752735 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22. Green N. M., “Avidin,” Adv. Prot. Chem. 29, 85–133 (1975). 10.1016/S0065-3233(08)60411-8 [DOI] [PubMed] [Google Scholar]
- 23. Hummer G., Pratt L. R., and García A. E., “Free energy of ionic hydration,” J. Phys. Chem. 100, 1206–1215 (1996). 10.1021/jp951011v [DOI] [Google Scholar]
- 24. Fedorov M. V. and Kornyshev A. A., “Unravelling the solvent response to neutral and charged solutes,” Mol. Phys. 105, 1–16 (2007). 10.1080/00268970601110316 [DOI] [Google Scholar]
- 25. Warshel A. and Levitt M., “Theoretical studies of enzymic reactions: Dielectric, electrostatic and steric stabilization of the carbonium ion in the reaction of lysozyme,” J. Mol. Biol. 103, 227–249 (1976). 10.1016/0022-2836(76)90311-9 [DOI] [PubMed] [Google Scholar]
- 26. Azuara C., Orland H., Bon M., Koehl P., and Delarue M., “Incorporating dipolar solvents with variable density in Poisson–Boltzmann electrostatics,” Biophys. J. 95, 5587–5605 (2008). 10.1529/biophysj.108.131649 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27. Alper H. E. and Levy R. M., “Field strength dependence of dielectric saturation in liquid water,” J. Phys. Chem. 94, 8401–8403 (1990). 10.1021/j100385a008 [DOI] [Google Scholar]
- 28. Purisima E. O. and Sulea T., “Restoring charge asymmetry in continuum electrostatic calculations of hydration free energies,” J. Phys. Chem. B 113, 8206–8209 (2009). 10.1021/jp9020799 [DOI] [PubMed] [Google Scholar]
- 29. Kornyshev A. A. and Sutmann G., “Nonlocal nonlinear static dielectric response of polar liquids,” J. Electroanal. Chem. 450, 143–156 (1998). 10.1016/S0022-0728(97)00622-0 [DOI] [Google Scholar]
- 30. Sandberg L. and Edholm O., “Nonlinear response effects in continuum models of the hydration of ions,” J. Chem. Phys. 116, 2936–2944 (2002). 10.1063/1.1435566 [DOI] [Google Scholar]
- 31. Jha A. K. and Freed K. F., “Solvation effect on conformations of 1,2:dimethoxyethane: Charge-dependent nonlinear response in implicit solvent models,” J. Chem. Phys. 128, 034501 (2008). 10.1063/1.2815764 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32. Gong H., Hocky G., and Freed K. F., “Influence of nonlinear electrostatics on transfer energies between liquid phases: Charge burial is far less expensive than Born model,” Proc. Natl. Acad. Sci. U.S.A. 105, 11146–11151 (2008). 10.1073/pnas.0804506105 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33. Hu L. and Wei G.-W., “Nonlinear Poisson equation for heterogeneous media,” Biophys. J. 103, 758–766 (2012). 10.1016/j.bpj.2012.07.006 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34. Mobley D. L., Barber A. E. II, Fennell C. J., and Dill K. A., “Charge asymmetries in hydration of polar solutes,” J. Phys. Chem. B 112, 2405–2414 (2008). 10.1021/jp709958f [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35. Corbeil C. R., Sulea T., and Purisima E. O., “Rapid prediction of solvation free energy. 2. The first-shell hydration (FiSH) continuum model,” J. Chem. Theory Comput. 6, 1622–1637 (2010). 10.1021/ct9006037 [DOI] [PubMed] [Google Scholar]
- 36. Mukhopadhyay A., Aguilar B. H., Tolokh I. S., and Onufriev A. V., “Introducing charge hydration asymmetry into the Generalized Born model,” J. Chem. Theory Comput. 10, 1788–1794 (2014). 10.1021/ct4010917 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37. Jackson J. D., Classical Electrodynamics, 3rd ed. (Wiley, 1998). [Google Scholar]
- 38. Lin B., Wong K.-Y., Hu C., Kokubo H., and Pettitt B. M., “Fast calculations of electrostatic solvation free energy from reconstructed solvent density using proximal radial distribution functions,” J. Phys. Chem. Lett. 2, 1626–1632 (2011). 10.1021/jz200609v [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39. Boda D., Valiskó M., Henderson D., Gillespie D., Eisenberg B., and Gilson M. K., “Ions and inhibitors in the binding site of HIV protease: Comparison of Monte Carlo simulations and the linearized Poisson–Boltzmann theory,” Biophys. J. 96, 1293–1306 (2009). 10.1016/j.bpj.2008.10.059 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40. Chern I.-L., Liu J.-G., and Wang W.-C., “Accurate evaluation of electrostatics for macromolecules in solution,” Methods Appl. Anal. 10, 309–328 (2003). 10.4310/MAA.2003.v10.n2.a9 [DOI] [Google Scholar]
- 41. Holst M., McCammon J. A., Yu Z., Zhou Y. C., and Zhu Y., “Adaptive finite element modeling techniques for the Poisson–Boltzmann equation,” Commun. Comput. Phys. 11, 179–214 (2012). 10.4208/cicp.081009.130611a [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42. Bardhan J. P., “Biomolecular electrostatics—I want your solvation (model),” Comput. Sci. Discovery 5, 013001 (2012). 10.1088/1749-4699/5/1/013001 [DOI] [Google Scholar]
- 43. Rizzo F. J., “An integral equation approach to boundary value problems of classical elastostatics,” Q. Appl. Math. 25, 83–95 (1967). [Google Scholar]
- 44. Shaw P. B., “Theory of the Poisson Green's-function for discontinuous dielectric media with an application to protein biophysics,” Phys. Rev. A 32(4), 2476–2487 (1985). 10.1103/PhysRevA.32.2476 [DOI] [PubMed] [Google Scholar]
- 45. Juffer A. H., Botta E. F. F., van Keulen B. A. M., van der Ploeg A., and Berendsen H. J. C., “The electric potential of a macromolecule in a solvent: A fundamental approach,” J. Comput. Phys. 97(1), 144–171 (1991). 10.1016/0021-9991(91)90043-K [DOI] [Google Scholar]
- 46. Bardhan J. P., “Numerical solution of boundary-integral equations for molecular electrostatics,” J. Chem. Phys. 130, 094102 (2009). 10.1063/1.3080769 [DOI] [PubMed] [Google Scholar]
- 47. Altman M. D., Bardhan J. P., White J. K., and Tidor B., “An efficient and accurate surface formulation for biomolecule electrostatics in non-ionic solution,” in Proceedings of the Engineering in Medicine and Biology Conference (EMBC; ), 2005 10.1109/IEMBS.2005.1616269 [DOI] [PubMed] [Google Scholar]
- 48. Miertus S., Scrocco E., and Tomasi J., “Electrostatic interactions of a solute with a continuum – A direct utilization of ab initio molecular potentials for the prevision of solvent effects,” Chem. Phys. 55(1), 117–129 (1981). 10.1016/0301-0104(81)85090-2 [DOI] [Google Scholar]
- 49. Sharp K. A. and Honig B., “Calculating total electrostatic energies with the nonlinear Poisson–Boltzmann equation,” J. Phys. Chem. 94(19), 7684–7692 (1990). 10.1021/j100382a068 [DOI] [Google Scholar]
- 50. Zhou H.-X., “Macromolecular electrostatic energy within the nonlinear Poisson–Boltzmann equation,” J. Chem. Phys. 100, 3152–3162 (1994). 10.1063/1.466406 [DOI] [Google Scholar]
- 51. Kirkwood J. G., “On the theory of strong electrolyte solutions,” J. Chem. Phys. 2, 767–772 (1934). 10.1063/1.1749393 [DOI] [Google Scholar]
- 52. Kharkats Yu. I., Kornyshev A. A., and Vorotyntsev M. A., “Electrostatic models in the theory of solutions,” J. Chem. Soc. Faraday Trans. 2 72, 361–371 (1976). 10.1039/f29767200361 [DOI] [Google Scholar]
- 53. Bardhan J. P., “Interpreting the Coulomb-field approximation for Generalized-Born electrostatics using boundary-integral equation theory,” J. Chem. Phys. 129, 144105 (2008). 10.1063/1.2987409 [DOI] [PubMed] [Google Scholar]
- 54. Harder E. and Roux B., “On the origin of the electrostatic potential difference at the liquid-vapor interface,” J. Chem. Phys. 129, 234706 (2008). 10.1063/1.3027513 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55. Kathmann S. M., Kuo I.-F. W., Mundy C. J., and Schenter G. K., “Understanding the surface potential of water,” J. Phys. Chem. B 115, 4369–4377 (2011). 10.1021/jp1116036 [DOI] [PubMed] [Google Scholar]
- 56. Brooks B. R., Bruccoleri R. E., Olafson B. D., States D. J., Swaminathan S., and Karplus M., “CHARMM: A program for macromolecular energy, minimization, and dynamics calculations,” J. Comput. Chem. 4, 187–217 (1983). 10.1002/jcc.540040211 [DOI] [Google Scholar]
- 57. Nina M., Beglov D., and Roux B., “Atomic radii for continuum electrostatics calculations based on molecular dynamics free energy simulations,” J. Phys. Chem. B 101, 5239–5248 (1997). 10.1021/jp970736r [DOI] [Google Scholar]
- 58. Sitkoff D., Sharp K. A., and Honig B., “Accurate calculation of hydration free energies using macroscopic solvent models,” J. Phys. Chem. B 98, 1978–1988 (1994). 10.1021/j100058a043 [DOI] [Google Scholar]
- 59. Mizzi S., Barber R. W., Emerson D. R., Reese J. M., and Stefanov S. K., “A phenomenological and extended continuum approach for modelling non-equilibrium flows,” Contin. Mech. Thermodyn. 19, 273–283 (2007). 10.1007/s00161-007-0054-9 [DOI] [Google Scholar]
- 60. Brenner H., “Beyond the no-slip boundary condition,” Phys. Rev. E 84(4), 046309 (2011). 10.1103/PhysRevE.84.046309 [DOI] [PubMed] [Google Scholar]
- 61. Bardhan J. P. and Knepley M. G., “Mathematical analysis of the boundary-integral based electrostatics estimation approximation for molecular solvation: Exact results for spherical inclusions,” J. Chem. Phys. 135, 124107 (2011). 10.1063/1.3641485 [DOI] [PubMed] [Google Scholar]
- 62. Maxwell J. C., “On stresses in rarefied gases arising from inequalities of temperature,” Proc. R. Soc. London 27, 304–308 (1878). 10.1098/rspl.1878.0052 [DOI] [Google Scholar]
- 63. von Smolan Smoluchowski M., “Über wärmeleitung in verdünnten gasen,” Ann. Phys. 300(1), 101–130 (1898). 10.1002/andp.18983000110 [DOI] [Google Scholar]
- 64. L'Heureux I. and Fowler A. D., “Dynamical model of oscillatory zoning in plagioclase with nonlinear partition relation,” Geophys. Res. Lett. 23, 17–20, doi: 10.1029/95GL03327 (1996). [DOI] [Google Scholar]
- 65. Macklin P. and Lowengrub J., “Evolving interfaces via gradients of geometry-dependent interior Poisson problems: Application to tumor growth,” J. Comput. Phys. 203, 191–220 (2005). 10.1016/j.jcp.2004.08.010 [DOI] [Google Scholar]
- 66. Fan T.-H. and Fedorov A. G., “Electrohydrodynamics and surface force analysis in AFM imaging of a charged, deformable biological membrane in a dilute electrolyte solution,” Langmuir 19, 10930–10939 (2003). 10.1021/la035663j [DOI] [Google Scholar]
- 67. Yossifon G., Frankel I., and Miloh T., “Symmetry breaking in induced-charge electro-osmosis over polarizable spheroids,” Phys. Fluids 19, 068105 (2007). 10.1063/1.2746847 [DOI] [Google Scholar]
- 68. Beglov D. and Roux B., “Solvation of complex molecules in a polar liquid: An integral equation theory,” J. Chem. Phys. 104(21), 8678–8689 (1996). 10.1063/1.471557 [DOI] [Google Scholar]
- 69. Bardhan J. P., “Nonlocal continuum electrostatic theory predicts surprisingly small energetic penalties for charge burial in proteins,” J. Chem. Phys. 135, 104113 (2011). 10.1063/1.3632995 [DOI] [PubMed] [Google Scholar]
- 70. Bardhan J. P. and Hildebrandt A., “A fast solver for nonlocal electrostatic theory in biomolecular science and engineering,” in Proceedings of the IEEE/ACM Design Automation Conference (DAC), 2011. [Google Scholar]
- 71. Bardhan J. P., “Gradient models in molecular biophysics: Progress, challenges, opportunities,” J. Mech. Behavior Mater. 22, 169–184 (2013). 10.1515/jmbm-2013-0024 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72. See supplementary material at http://dx.doi.org/10.1063/1.4897324E-JCPSA6-141-016439 for figures comparing NLBC and MD calculations for the Mobley test set.34 The source code (MATLAB) and surface discretizations for running the nonlinear boundary-condition calculations, data files, parameters, and scripts for preparing and running the MD calculations of titratable residues, as well as source code to generate the figures, are freely and publicly available online at https://bitbucket.org/jbardhan/si-nlbc.
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Data Citations
- See supplementary material at http://dx.doi.org/10.1063/1.4897324E-JCPSA6-141-016439 for figures comparing NLBC and MD calculations for the Mobley test set.34 The source code (MATLAB) and surface discretizations for running the nonlinear boundary-condition calculations, data files, parameters, and scripts for preparing and running the MD calculations of titratable residues, as well as source code to generate the figures, are freely and publicly available online at https://bitbucket.org/jbardhan/si-nlbc.


