Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2016 Feb 12.
Published in final edited form as: SIAM J Appl Math. 2015 Sep 15;75(5):2072–2092. doi: 10.1137/15M100701X

DIFFUSED SOLUTE-SOLVENT INTERFACE WITH POISSON–BOLTZMANN ELECTROSTATICS: FREE-ENERGY VARIATION AND SHARP-INTERFACE LIMIT

BO LI *, YUAN LIU
PMCID: PMC4752183  NIHMSID: NIHMS733536  PMID: 26877556

Abstract

A phase-field free-energy functional for the solvation of charged molecules (e.g., proteins) in aqueous solvent (i.e., water or salted water) is constructed. The functional consists of the solute volumetric and solute-solvent interfacial energies, the solute-solvent van der Waals interaction energy, and the continuum electrostatic free energy described by the Poisson–Boltzmann theory. All these are expressed in terms of phase fields that, for low free-energy conformations, are close to one value in the solute phase and another in the solvent phase. A key property of the model is that the phase-field interpolation of dielectric coefficient has the vanishing derivative at both solute and solvent phases. The first variation of such an effective free-energy functional is derived. Matched asymptotic analysis is carried out for the resulting relaxation dynamics of the diffused solute-solvent interface. It is shown that the sharp-interface limit is exactly the variational implicit-solvent model that has successfully captured capillary evaporation in hydrophobic confinement and corresponding multiple equilibrium states of underlying biomolecular systems as found in experiment and molecular dynamics simulations. Our phase-field approach and analysis can be used to possibly couple the description of interfacial fluctuations for efficient numerical computations of biomolecular interactions.

Keywords and phrases: variational implicit-solvent model, diffused solute-solvent interface, Poisson–Boltzmann theory, matched asymptotic analysis, sharp-interface limit

1. Introduction

The solvation of charged molecules such as DNAs and proteins in aqueous solvent (i.e., water or salted water) is a fundamental biological process. Implicit-solvent models provide efficient predictions of the solvation free energies and equilibrium conformations of underlying molecular systems [35]. In such a model, the solvent molecules and ions are treated implicitly and their effects are coarse-grained. In a large class of implicit-solvent models, the solute-solvent interface that separates the solvent region from solutes is used to calculate the surface energy and electrostatic energy. The later is often done by solving Poisson’s or the Poisson–Boltzmann equation with the dielectric coefficient close to 1 and 80 in the solute and solvent regions, respectively [7,16,27,38]. In a variational implicit-solvent model (VISM) [19,20,44], one determines the solvation free energies and stable equilibrium conformations by minimizing a macroscopic free-energy functional of all possible solute-solvent interfaces, i.e., dielectric boundaries.

Let us denote by Ω ⊂ ℝ3 the region of an underlying solvation system; cf. Figure 1.1. Assume that there are total N solute atoms located at x1, …, xN inside Ω and carrying partial charges Q1, …, QN, respectively. Assume also that there are M ionic species in the solvent, and denote by cj and qj = zje the bulk concentration and charge for the jth ionic species, respectively, where zj is the valence and e elementary charge. Let Γ be a closed and smooth surface, a possible solute-solvent interface, that encloses all x1, …, xN and that divides Ω into two regions: the solute region Ωp (p stands for protein) and the solvent region Ωw (w stands for water). The VISM solvation free-energy functional is then defined for all such solute-solvent interface Γ by

Fig. 1.1.

Fig. 1.1

A schematic view of a solvation system with an implicit solvent. A solute-solvent interface Γ separates the solute region ΩP (p stands for protein) from the solvent region ΩW (w stands for water). Dots represent solute atoms located at xi carrying partial charges Qi(i = 1, …, N).

F[Γ]=PVol(Ωp)+γ0Area(Γ)+ρwΩwU(x)dx+Ω[-εΓ2ψ2+ρfψ-χwV(ψ)]dx. (1.1)

The first term here is the energy of creating a solute cavity in the solvent against the pressure difference P between the solvent liquid and solute vapor. The second term is the surface energy with γ0 being the constant surface tension. For simplicity, we neglect the Tolman correction to the surface tension [40]. The third term is the energy of the van der Waals (vdW) type interaction between the solute atoms xi (1 ≤ iN) and continuum solvent, where ρw is the constant bulk solvent density. The interaction potential U is often defined by

U(x)=i=1NULJ(i)(x-xi),

where each

ULJ(i)(r)=4εi[(σir)12-(σir)6]

is a Lennard-Jones potential, and all the parameters εi of energy and σi of length are given.

The last term is the electrostatic free energy, where ψ is the electrostatic potential, εΓ is the coefficient of permittivity, ρf is a fixed charge density approximating the solute point charges i=1NQiδxi with δxi being the Dirac delta function at xi, χw is the characteristic function of the solvent region Ωw, and V(ψ) describes the ionic contribution to the electrostatic interaction. Here, ions are assumed to be in an equilibrium macroscopically; and the equilibrium ionic concentrations are given by the Boltzmann distributions through the electrostatic potential ψ [7, 27]. The potential ψ is the solution to a boundary-value problem of the dielectric-boundary Poisson–Boltzmann equation

·εΓψ-χwV(ψ)=-ρfinΩ, (1.2)
ψ=ψonΩ, (1.3)

where ψ is a given function on the boundary Ω. The coefficient εΓ is defined by

εΓ(x)={εpε0ifxΩp,εwε0ifxΩw, (1.4)

where εp and εw are the dielectric coefficients (relative permitivities) of the solute and solvent regions, respectively, and ε0 is the vacuum permittivity. In general, εp ≈ 1 and εw ≈ 80. In the classical Poisson–Boltzmann theory,

V(ψ)=β-1j=1Mcj(e-βqjψ-1),

where β = 1/(kBT) is the inverse thermal energy with kB the Boltzmann constant and T absolute temperature. Different forms of V (ψ) can be used for the linearized or size-modified Poisson–Boltzmann equation [26, 45]. More complicated forms V = V(ψ, ∇ψ) involving both ψ and ∇ψ can be used to model effects such as the ionic concentration dependent dielectric response but can also be more complicated for implementation [2,23,29].

Cheng et al. [9,10,12,41,44] developed a highly accurate, compact coupling interface method for solving the dielectric-boundary Poisson–Boltzmann equation and a robust level-set method to minimize the VISM functional (1.1). The normal velocity in the level-set formulation is the (normal component of) effective boundary force defined to be the negative variational derivative −δΓF [Γ] : Γ → ℝ of the free-energy functional F[Γ] with respect to the location change of interface Γ. Let n denote the unit normal to the interface Γ pointing from the solvent region Ωw to solute region Ωp. We have [5,911,28,44]

-δΓF[Γ]=P+2γ0H-ρwU+12(1εpε0-1εwε0)(εΓψn)2+12(εwε0-εpε0)Γψ2+V(ψ)onΓ, (1.5)

where H is the mean curvature that is positive if Ωp is a sphere and ∇Γ = (Inn) ∇, with I the identity matrix, is the surface gradient along Γ.

While successful initially [9, 12, 22, 41, 44], the sharp-interface VISM needs to be improved in several aspects. The most critical one is to include the description of fluctuations, both around the solute-solvent interface and in the bulk solvent. Such fluctuations are particularly crucial in the transition of one equilibrium conformation to another and in sampling different states to accurately predict the free energies of underlying biomolecular systems. It is certainly possible to describe interfacial fluctuations within a sharp-interface framework. But several implementational difficulties can arise. For instance, the extension of normal velocity in the level-set method can be hard for a moving fluctuating interface. Moreover, fluctuations can nucleate and coalesce small bubbles (water regions) inside solutes, making it hard to solve the dielectric boundary Poisson–Boltzmann equation in a sharp-interface formulation. In seeking for an alternative approach, we notice that initial theoretical and computational studies of interfacial fluctuations using a diffused-interface approach seem promising [3,17,24]. We therefore propose in this work a diffused-interface approach to the solvation of charged molecules. In a slight different language, this is a phase-field approach as it is well appreciated that the solvent-solute interface in a biomolecular system resembles a liquid-vapor interface, and the solvent and solute can be regarded as two different phases [6,42]. The phase-field approach has been widely used in studying interface problems arising in many scientific areas, such as materials physics, complex fluids, biomembranes, and cell motility, cf. e.g., [1, 13, 18, 25, 31, 37] and the references therein.

Our phase-field model is governed by the free-energy functional

Fξ[ϕ]=PΩϕ2dx+γ0Ω[ξ2ϕ2+1ξW(ϕ)]dx+ρwΩ(ϕ-1)2Udx+Ω[-ε(ϕ)2ψϕ2+ρfψϕ-(ϕ-1)2V(ψϕ)]dx (1.6)

for all phase fields ϕ : Ω → ℝ, where the electrostatic potential ψϕ is the solution to the boundary-value problem of Poisson–Boltzmann equation with a diffused dielectric boundary

·ε(ϕ)ψϕ-(ϕ-1)2V(ψϕ)=-ρfinΩ, (1.7)
ψϕ=ψonΩ. (1.8)

Here ξ > 0 is a small number. With a low free energy, a phase field ϕ should be close to two values, say, 0 and 1, respectively, in the solvation region Ω, except a thin transition layer that represents the dielectric boundary. The sets {ϕ ≈ 0} and {ϕ ≈ 1} represent the solvent and solute regions, respectively. With such a convention, the first term in (1.6) corresponds to that in (1.1), describing the volumetric energy. The second term in (1.6), in which,

W(ϕ)=18ϕ2(1-ϕ)2,

approximates the surface energy [30, 33, 39]. The third term in (1.6) corresponds to that in (1.1), describing the energy of solute-solvent interaction. Finally, the last term in (1.6) corresponds to that in (1.1), describing the electrostatic free energy. In this term, ε(ϕ) is defined to be a smooth function of ϕ that interpolates the piecewise constant dielectric coefficient: ε(0) = εwε0 and ε(1) = εpε0. We shall require that ε′(0) = ε′(1) = 0. Note that the electrostatic potential ψϕ depends on ϕ.

We study our proposed phase-field solvation free-energy functional (1.6), and the corresponding relaxation dynamics. Our main results are as follows:

  1. We prove the well-posedness of the boundary-value problem of Poisson–Boltzmann equation (1.7) and (1.8), and obtain some bounds for the electrostatic potential; cf. Theorem 2.1;

  2. We derive, first intuitively and then rigorously, the first variation of the phase-field free-energy functional (1.6)
    δϕF[ϕ]=2Pϕ-γ0[ξΔϕ-1ξW(ϕ)]+2ρw(ϕ-1)U-12ε(ϕ)ψ2-2(ϕ-1)V(ψ);

    cf. Theorem 3.1;

  3. We consider the relaxation dynamics tϕ = −δϕF[ϕ] for ϕ = ϕ (x, t) with t denoting the time variable. Using the method of matched asymptotic analysis [4, 14, 15, 34, 36], we show that, in the sharp-interface limit, the normal velocity of the solute-solvent interface is exactly the boundary force (1.5) in the original sharp-interface variational approach; cf. Theorem 5.1.

Several remarks are in order. First, our diffuse-interface model (1.6) differs from our previous one [30,43] significantly in that we use the Poisson–Boltzmann equation, rather than the Coulomb-field approximation [8, 41] which requires no solution of partial differential equations, to describe the electrostatic interaction. Second, our relaxation dynamics tϕ = −δϕF[ϕ] is a non-conservative dynamics with respect to the phase-field function ϕ. In fact, the relaxation process can change the volume of solute region {ϕ ≈ 1}. The amount of ions, however, are controlled through their bulk concentrations cj(j=1,,M) that are input parameters. Third, the surface energy term described by the gradient-square and double-well terms in our phase-field energy functional (1.6) coincides with the part in an effective Hamiltonian for the large-scale solvent density in the Lum–Chandler–Weeks theory of solvation [32].

The rest of the paper is organized as follows: In Section 2, we prove the existence and uniqueness of, and provide bounds for, the solution to the Poisson–Boltzmann equation with a diffused dielectric boundary that is given by a phase field. In Section 3, we derive the first variation of the phase-field free-energy functional. In Section 4, we use the matched asymptotic analysis to study the relaxation dynamics in a fast and the regular time scales. In Section 5, we continue our matched asymptotic analysis for a slow time scale to show that the sharp-interface limit of the relaxation dynamics is exactly the same as in the original sharp-interface variational model.

2. Poisson–Boltzmann Electrostatics with a Diffused Dielectric Boundary

We make the following assumptions:

  • (A1)

    The solvation region Ω is a bounded, connected, open set with a sufficiently smooth boundary Ω. All x1, …, xN are given points in Ω. All P, γ0, ρw, εp, and εw are positive constants with εp < εw. The function ρfL(Ω). The function ψW1,∞ (Ω);

  • (A2)

    The function U : ℝ3 \ {x1, …, xN} → ℝ is continuously differentiable and bounded below. Moreover, limxxiU(x) = +∞ for each i (1 ≤ iN) and limx→∞ U(x) = 0;

  • (A3)

    The function εC1(ℝ). There exist positive numbers εmin, εmax, and εmax such that εminε(ϕ) ≤ εmax and ε(ϕ)εmax for all ϕ ∈ ℝ. Moreover, ε(0) = εwε0, ε(1) = εpε0, ε′(0) = ε′(1) = 0, and ε′(ϕ) ≠ 0 if 0 < ϕ < 1;

  • (A4)

    The function VC2(ℝ). It is strictly convex. Moreover, V (0) = mins∈ℝ V(s) < V(s) for any s ≠ 0, V(±∞) = +∞, and V′(±∞) = ±∞.

Here are two examples of the function ε = ε(ϕ) with all the desired properties.

Example 1.

ε(ϕ)={εwε0ifϕ0,εpε0etan(π(ϕ-1/2))+εwε0e-tan(π(ϕ-1/2))etan(π(ϕ-1/2))+e-tan(π(ϕ-1/2))if0<ϕ<1,εpε0ifϕ1.

In fact, this is a C-function. Moreover, for any integer m ≥ 1, ε(m)(ϕ) = 0 for all ϕ ∉ (0, 1) and ε(m)(ϕ) ≠ 0 for all ϕ ∈ (0, 1).

Example 2.

ε(ϕ)={εwε0ifϕ0,12(εwε0-εpε0)cosπϕ+12(εwε0+εpε0)if0<ϕ<1,εpε0ifϕ1.

In what follows, we shall denote by C a generic constant that may depend on Ω, ρf, ψ, and ε but is independent of ϕH01(Ω). We shall also denote by C(ϕ) a generic constant that can depend on ϕ in addition to Ω, ρf, ψ, and ε. The following theorem is our main result in this section:

Theorem 2.1

For any ϕH01(Ω) there exists a unique weak solution ψϕ ∈ H1(Ω) to the boundary-value problem (1.7) and (1.8), i.e., ψϕ = ψ on ∂Ω in the sense of trace and

Ω[ε(ϕ)ψϕ·η+(ϕ-1)2V(ψϕ)η]dx=ΩρfηdxηH01(Ω). (2.1)

Moreover, ψϕL(Ω) and

ψϕH1(Ω)C(1+ϕL2(Ω)), (2.2)
χ{ϕ1}ψϕL(Ω)C, (2.3)
ψϕL(Ω)C(1+ϕH1(Ω)2). (2.4)

Proof

We divide our proof in three steps.

Step 1

Let ψ̂ϕH1(Ω) be the unique weak solution to the boundary-value problem −∇ · ε(ϕ)∇ψ̂ϕ = ρf in Ω and ψ̂ϕ = ψ on Ω. This means that ψ̂ϕ is the unique function in H1(Ω) such that ψ̂ϕ = ψ on Ω and

Ωε(ϕ)ψ^ϕ·ηdx=ΩρfηdxηH01(Ω). (2.5)

Since 0 < εminε(ϕ) ≤ εmax for all ϕ ∈ ℝ, we have by the standard elliptic theory (cf. Theorem 8.3 and Theorem 8.16 in [21]) that ψ̂ϕH1(Ω) ∩ L(Ω), and

ψ^ϕH1(Ω)+ψ^ϕL(Ω)C. (2.6)
Step 2

We define I: H01(Ω){+} by

I[u]=Ω[ε(ϕ)2u2+(ϕ-1)2V(u+ψ^ϕ)]dxuH01(Ω).

It is easy to see that infuH01(Ω)I[u] is finite. By the direct method in the calculus of variations, there exists a unique minimizer uH01(Ω) of I over H01(Ω).

By our assumption (A4) on the function V and by (2.6), there exists λ > 0, independent of ϕ, such that V′ (λ + ψ̂ϕ) > 1 and V′(−λ + ψ̂ϕ) < −1 a.e. on Ω. We prove that

uλa.e.on{ϕ1}. (2.7)

Define uλ : Ω → ℝ by

uλ(x)={-λifu(x)<-λ,u(x)if-λu(x)λ,λifu(x)>λ.

Since I[uλ] ≥ I[u] and |∇uλ| ≤ |∇u| a.e. Ω, we have

Ω(ϕ-1)2V(u+ψ^0)dx=I[u]-Ωε(ϕ)2u2dxI[uλ]-Ωε(ϕ)2uλ2dx=Ω(ϕ-1)2V(uλ+ψ^ϕ)dx.

Consequently, it follows from the convexity of V and our choice of λ that

0Ω(ϕ-1)2V(u+ψ^ϕ)dx-Ω(ϕ-1)2V(uλ+ψ^ϕ)dx=Ωχ{u>λ}(ϕ-1)2[V(u+ψ^ϕ)-V(λ+ψ^ϕ)]dx+Ωχ{u<-λ}(ϕ-1)2[V(u+ψ^ϕ)-V(-λ+ψ^ϕ)]dxΩχ{u>λ}(ϕ-1)2V(λ+ψ^ϕ)(u-λ)dx+Ωχ{u<-λ}(ϕ-1)2V(-λ+ψ^ϕ)(u+λ)dxΩχ{u>λ}(ϕ-1)2(u-λ)dx+Ωχ{u<-λ}(ϕ-1)2(-u-λ)dx=Ωχ{u>λ}(ϕ-1)2u-λdx0.

This leads to (2.7).

The minimizer u of I: H01(Ω){+} is the weak solution to the Euler–Lagrange equation

·ε(ϕ)u-(ϕ-1)2V(u+ψ^ϕ)=0inΩ,

i.e.,

Ω[ε(ϕ)u·η+(ϕ-1)2V(u+ψ^ϕ)η]dx=0ηH01(Ω). (2.8)

This is true by (2.7) and the Lebesgue Dominated Convergence Theorem if ηH01(Ω)L(Ω). For a general ηH01(Ω), this is also true, since (ϕ − 1)2V′(u+ψ̂ϕ) ∈ L4(Ω) by (2.7) and since H01(Ω)L(Ω) is dense in H01(Ω). The convexity of V implies that the weak solution uH01(Ω) defined by (2.8) is unique.

If we choose η = u in the equation of (2.8), we obtain by (2.6), (2.7), and Poincaré’s inequality that

uH1(Ω)2C{ϕ1}|(ϕ-1)2V(u+ψ^ϕ)u|dxCΩ(1+ϕ2)dx,

leading to

uH1(Ω)C(1+ϕL2(Ω)). (2.9)

It now follows from the regularity theory (cf. Theorem 8.16 in [21]), (2.6), (2.7), and the imbedding H1(Ω) ↪ L4(Ω) that

uL(Ω)C(ϕ-1)2V(u+ψ^ϕ)L2(Ω)C(1+ϕL4(Ω)2)C(1+ϕH1(Ω)2). (2.10)
Step 3

Let ψϕ = u + ψ̂ϕ. Then ψϕ = ψ on Ω. Moreover, (2.5) and (2.8) imply (2.1). The uniqueness of ψϕ follows from the convexity of V and a usual argument. The estimates (2.2)–(2.4) follow from (2.6), (2.9), (2.7), and (2.10).

3. First Variation of Free-Energy Functional

We calculate the first variation of the free-energy functional Fξ defined by (1.6). The first variation of each of the first three parts in the functional Fξ can be obtained by routine calculations. So, we focus on the electrostatic part

Fele,ξ[ϕ]:=Ω[-ε(ϕ)2ψϕ2+ρfψϕ-(ϕ-1)2V(ψϕ)]dx, (3.1)

where ψϕ is the unique solution to the boundary-value problem (1.7) and (1.8). Heuristically, if we denote by δϕ the “derivative” with respect to ϕ and δϕ the variation of ϕ, and “differentiate” both sides of the above equation, then we obtain by the chain rule and product rule for differentiation that

δϕFele,ξ[ϕ]δϕ=Ω[-ε(ϕ)2δϕψϕ2-ε(ϕ)ψϕ·δϕψϕ+ρfδϕψϕ-2(ϕ-1)δϕV(ψϕ)-(ϕ-1)2V(ψϕ)δϕψϕ]dx=Ω[-ε(ϕ)2ψϕ2-2(ϕ-1)V(ψϕ)]δϕdx+Ωρfδϕψϕdx-Ω[ε(ϕ)ψϕ·δϕψϕ+(ϕ-1)2V(ψϕ)δϕψϕ]dx=Ω[-ε(ϕ)2ψϕ2-2(ϕ-1)V(ψϕ)]δϕdx,

where in the last step we used the weak form (2.1) of the Poisson–Boltzmann equation with η = δϕψϕ. We then identify δϕFele,ξ[ϕ] as

δϕFele,ξ[ϕ]=-ε(ϕ)2ψϕ2-2(ϕ-1)V(ψϕ).

We remark that −δϕFele [ϕ] is an approximation of the electrostatic part of the boundary force −δΓF [Γ] in (1.5) if ϕ ≈ 0 in the solvent region and ϕ ≈ 1 in the solute region. In fact, the V-terms are similar, since the perturbation δϕ is localized at the interface. With our notation, the unit normal n ≈ −∇ϕ/|∇ϕ|. Moreover, ε(ϕ) ≈ εΓ. Therefore,

ε(ϕ)2ψϕ2=ε(ϕ)2ψϕ·n2+ε(ϕ)2Γψϕ2=-12ddϕ[1ε(ϕ)]ε(ϕ)ψϕ·n2+ε(ϕ)2Γψϕ2-12(1εpε0-1εwε0)εΓψϕ·n2+12(εpε0-εwε0)Γψ02.

The final result is exactly the corresponding part in (1.5).

We now give a rigorous justification of our derivation. We recall that if G:H01(Ω) is a functional and if ϕ, ηH01(Ω), then the first variation of G at ϕ in the “direction” η is defined to be

δϕG[ϕ][η]=limt0G[ϕ+tη]-G[ϕ]t=ddt|t=0G[ϕ+tη],

if the limit exists.

Theorem 3.1

Let ϕH01(Ω). Let ψϕ ∈ H1(Ω) be the weak solution of the corresponding boundary-value problem (1.7) and (1.8). Let ηL(Ω)H01(Ω). Then

δϕFele,ξ[ϕ][η]=Ω[-12ε(ϕ)ψϕ2-2(ϕ-1)V(ψϕ)]ηdx. (3.2)

To prove the theorem, we need the following lemma:

Lemma 3.2

Let ϕH01(Ω) and ηL(Ω)H01(Ω). Let t ∈ ℝ with |t| ≤ 1. Let ψϕ and ψϕ+ be the weak solutions to the boundary-value problems of Poisson–Boltzmann equation (1.7) and (1.8) corresponding to ϕ and ϕ + tη, respectively. Then there exists a constant C = C(ϕ, η) > 0 that may depend on both ϕ and η such that

ψϕ+tη-ψϕH1(Ω)C(ϕ,η)tift1.

Proof

By the weak form for ψϕ and that for ψϕ+, cf. (2.1), we have for any ζH01(Ω) that

Ω[ε(ϕ)ψϕ·ζ+(ϕ-1)2V(ψϕ)ζ]dx=Ωρfζdx,Ω[ε(ϕ+tη)ψϕ+tη·ζ+(ϕ+tη-1)2V(ψϕ+tη)ζ]dx=Ωρfζdx.

It follows from these two equations with ζ = ψϕ+ψϕ, the property of ε, and the convexity of V that

εmin(ψϕ+tη-ψϕ)L2(Ω)2Ωε(ϕ)(ψϕ+tη-ψϕ)·(ψϕ+tη-ψϕ)dx=Ωε(ϕ)ψϕ+tη·(ψϕ+tη-ψϕ)dx-Ωε(ϕ)ψϕ·(ψϕ+tη-ψϕ)dx=-Ω[(ε(ϕ+tη)-ε(ϕ))]ψϕ+tη·(ψϕ+tη-ψϕ)dx+Ωε(ϕ+tη)ψϕ+tη·(ψϕ+tη-ψϕ)dx-Ωε(ϕ)ψϕ·(ψϕ+tη-ψϕ)dx=-Ω[(ε(ϕ+tη)-ε(ϕ)]ψϕ+tη·(ψϕ+tη-ψϕ)dx-Ω[(ϕ+tη-1)2-(ϕ-1)2]V(ψϕ+tη)(ψϕ+tη-ψϕ)dx-Ω(ϕ-1)2[V(ψϕ+tη)-V(ψϕ)](ψϕ+tη-ψϕ)dx-Ω[(ε(ϕ+tη)-ε(ϕ))]ψϕ+tη·(ψϕ+tη-ψϕ)dx-Ω[(ϕ+tη-1)2-(ϕ-1)2]V(ψϕ+tη)(ψϕ+tη-ψϕ)dxεmaxtηL(Ω)ψϕ+tηL2(Ω)(ψϕ+tη-ψϕ)L2(Ω)+(2tη(ϕ-1)L2(Ω)+t2η2L2(Ω))V(ψϕ+tη)L(Ω)ψϕ+tη-ψϕL2(Ω).

Since |t| ≤ 1, we have by (2.2) and (2.4) that both ||ψϕ+||H1(Ω) and ||V′(ψϕ + )||L(Ω) are bounded by some constants that can depend on ϕ and η. An application of Poincaré’s inequality then implies the desired inequality.

Proof of Theorem 3.1

For any t with |t| ≪ 1, we denote by ψϕ+ the weak solution to the boundary-value problem (1.7) and (1.8) with ϕ + replacing ϕ. Setting ψϕ+tη-ψϕH01(Ω) as the test function in the corresponding weak formulation for ψϕ+, we obtain

Ωρf(ψϕ+tη-ψϕ)dx=Ωε(ϕ+tη)ψϕ+tη·(ψϕ+tη-ψϕ)dx+Ω(ϕ+tη-1)2V(ψϕ+tη)(ψϕ+tη-ψϕ)dx=12Ωε(ϕ+tη)(ψϕ+tη2+ψϕ+tη-ψϕ2-ψϕ2)dx+Ω(ϕ+tη-1)2V(ψϕ+tη)(ψϕ+tη-ψϕ)dx.

Consequently, by the definition of the functional Fele (cf. (3.1)), we have

Fele,ξ[ϕ+tη]-Fele,ξ[ϕ]=Ω[-ε(ϕ+tη)2ψϕ+tη2+ε(ϕ)2ψϕ2+ρf(ψϕ+tη-ψϕ)-(ϕ+tη-1)2V(ψϕ+tη)+(ϕ-1)2V(ψϕ)]dx=Ω12ε(ϕ+tη)ψϕ+tη-ψϕ2dx-Ω12[ε(ϕ+tη)-ε(ϕ)]ψϕ2dx-Ω[(ϕ+tη-1)2-(ϕ-1)2]V(ψϕ)dx-Ω(ϕ+tη-1)2[V(ψϕ+tη)-V(ψϕ)-V(ψϕ+tη)(ψϕ+tη-ψϕ)]dx=:Λ1(t)-Λ2(t)-Λ3(t)-Λ4(t), (3.3)

where Λi(t) (i = 1,, 4) denote the corresponding integrals.

By Lemma 3.2, we have

|Λ1(t)t|εmax2tψϕ+tη-ψϕH1(Ω)2C(ϕ,η)t0ast0. (3.4)

Since ηL(Ω), we have for a.e. x ∈ Ω that

1t[ε(ϕ+tη)-ε(ϕ)]=ε(ϕ+θtη)ηε(ϕ)ηast0,

where θ = θ(x) ∈ [0, 1]. The Lebesgue Dominated Convergence Theorem then implies that

limt0Λ2(t)t=Ω12ε(ϕ)ψϕ2ηdx. (3.5)

Similarly, since ψϕL(Ω) by Theorem 2.1,

limt0Λ3(t)t=Ω2(ϕ-1)V(ψϕ)ηdx. (3.6)

Let a ∈ ℝ and define g(b) = V (b) − V (a) − V′(b)(ba) for all b ∈ ℝ. Note that g(a) = 0 and g′(b) = −V″(b)(ba). By Taylor’s expansion,

g(b)=g(a)+g(a+μ(b-a))(b-a)=-μV(a+μ(b-a))(b-a)2,

where μ = μ(b) ∈ [0, 1]. With a = ψϕ and b = ψϕ+, we obtain by (2.4), the Cauchy–Schwarz inequality, the imbedding H1(Ω) ↪ L4(Ω), and Lemma 3.2 that

|Λ4(t)t|C(ϕ,η)tΩ(ϕ+tη-1)2ψϕ+tη-ψϕ2dxC(ϕ,η)t(ϕ+tη-1)2L2(Ω)ψϕ+tη-ψϕL4(Ω)2C(ϕ,η)t(ϕ+tη-1)2L2(Ω)ψϕ+tη-ψϕH1(Ω)20ast0. (3.7)

Now (3.2) follows from (3.3)–(3.7).

The next theorem provides the formula of first variation of the solvation free-energy functional Fξ defined in (1.6). It follows from Theorem 3.1 and routine calculations. We thus omit the proof.

Theorem 3.3

Let ϕH01(Ω) and ηL(Ω)H01(Ω) be such that

Ω(ϕ-1)2Udx<andΩη2Udx<.

Let ψϕH1(Ω) be the unique weak solution to the boundary-value problem (1.7) and (1.8) corresponding to ϕ. Then

δϕFξ[ϕ][η]=Ω{γ0ξϕ·η+[2Pϕ+γ0ξW(ϕ)+2ρw(ϕ-1)U-ε(ϕ)2ψϕ2-2(ϕ-1)V(ψϕ)]η}dx.

If in addition ϕH2(Ω) then the integral of ∇ϕ · ∇η becomes that of −Δϕη. Therefore, we shall denote

δϕFξ[ϕ]=2Pϕ-γ0[ξΔϕ-1ξW(ϕ)]+2ρw(ϕ-1)U-ε(ϕ)2ψϕ2-2(ϕ-1)V(ψϕ), (3.8)

and call it the first variation of Fξ at ϕ.

4. Relaxation Dynamics: Fast and Regular Time Scales

We now consider the relaxation dynamics tϕ = −δϕFξ[ϕ]. By Theorem 3.3, this dynamics is described by the following initial-boundary-value problem for the phase field ϕ = ϕ(x, t) and the corresponding electrostatic potential ψ = ψ(x, t):

tϕ=-2Pϕ+γ0[ξΔϕ-1ξW(ϕ)]-2ρw(ϕ-1)U+12ε(ϕ)ψ2+2(ϕ-1)V(ψ), (4.1)
·ε(ϕ)ψ-(ϕ-1)2V(ψ)=-ρf, (4.2)
ϕ=0onΩ, (4.3)
ψ=ψonΩ, (4.4)

together with some initial condition for ϕ. If the solution ϕ = ϕ(x, t) and ψ = ψ(x, t) are smooth, then we have by (1.6), (3.8), the chain rule, and integration by parts that

ddtFξ[ϕ(·,t)]=ΩtϕδϕFξ[ϕ]dx-Ω[ε(ϕ)ψ·tψ-ρftψ+(ϕ-1)2V(ψ)tψ]dx=-Ω(tϕ)2dx0,

where the vanish of the second integral follows from the weak formulation for (4.2) with the test function tψ. Therefore, the energy decays, explaining the terminology “relaxation dynamics”.

Assume 0 < ξ ≪ 1. We expect that, after a short time period, the solution ϕ will take the value close to 0 or 1 in most of the region Ω, due to the fast reaction described by the term −(1)W′(ϕ). Moreover, in the regions {ϕ ≈ 0} and {ϕ ≈ 1}, the leading-order term of the electrostatic potential should solve the boundary-value problem (4.2) and (4.4), as the electrostatic relaxation is instantaneous: there is no tψ term. In the subsequent long time period, the relaxation dynamics is mainly the motion of the interface that separates the two regions {ϕ ≈ 0} and {ϕ ≈ 1}. It is reasonable to assume that the early fast time scale is determined by the basic reaction-diffusion equation

tϕ=ξΔϕ-1ξW(ϕ),

as again the electrostatic relaxation is instantaneous. Perturbing the unstable constant equilibrium solution ϕ0(x) = 1/2 by δ exp (ik · x + ωt) with |δ| ≪ 1, we find the most unstable mode kc = 0 and the corresponding fastest growth rate ω(kc) = O(1). Therefore, the fast time scale is t/ξ. The next time scales are the regular time scale t, a slow time scale ξt, and so on.

Let us first consider the fast time scale T = t/ξ. We assume that

ϕ(x,t)=ϕ0(x,T)+ξϕ1(x,T)+ξ2ϕ2(x,T)+, (4.5)
ψ(x,t)=ψ0(x,T)+ξψ1(x,T)+ξ2ψ2(x,T)+, (4.6)

where ϕi = ϕi(x, T) and ψi = ψi(x, T) (i = 0, 1, …) are smooth and bounded functions satisfying the boundary conditions

ϕi=0(i0),ψ0=ψ,andψi=0(i1)onΩ. (4.7)

Substitute ϕ and ψ in (4.1) by these expansions and match terms with the same orders of ξ. At the leading order O(ξ−1), we obtain T ϕ0 = −γ0W′(ϕ0). This equation has two stable fixed points 0 and 1. So,

foranyxΩ,ϕ0(x,T)0or1exponentiallyasT. (4.8)

Similarly, placing our expansions into (4.2), we obtain the leading-order O(1)-equation

·ε(ϕ0)ψ0-(ϕ0-1)2V(ψ0)=-ρf. (4.9)

This is exactly (4.2), as expected. More equations corresponding to other higher powers of ξ can be obtained but they provide less useful information. We therefore need to examine the next time scale, the regular time scale.

We assume that, at the regular time scale, the solution ϕ and ψ have the expansions

ϕ(x,t)=ϕ0(x,t)+ξϕ1(x,t)+ξ2ϕ2(x,t)+,ψ(x,t)=ψ0(x,t)+ξψ1(x,t)+ξ2ψ2(x,t)+,

where ϕi = ϕi(x, t) and ψi = ψi(x, t) (i = 0, 1, …) are smooth and bounded functions that satisfy the boundary conditions (4.7). Note that these are functions of (x, t), different from those of (x, T) in (4.5) and (4.6). Substituting ϕ and ψ in (4.1) by these expansions and matching the terms with the same orders of ξ, we obtain the first two equations

O(ξ-1):W(ϕ0)=0,O(1):tϕ0=-2Pϕ0-γ0W(ϕ0)ϕ1-2ρw(ϕ0-1)U (4.10)
+12ε(ϕ0)ψ02+2(ϕ0-1)V(ψ0). (4.11)

Note that Eq. (4.2) does not involve any time derivatives. Thus, substituting ϕ and ψ in (4.2) by their expansions, we obtain the same equation (4.9) in the leading order.

Eq. (4.10) implies that ϕ0 = 0, 1, or 1/2. By (4.8), we must have ϕ0 = 0 or 1. Since ϕ0 = 0 on Ω, we have ϕ0 = 0 near Ω. By the assumption (A2) on the function U in Section 2 and (4.11), ϕ0 cannot be identically 0 in Ω. Thus, both {ϕ0 = 0} and {ϕ0 = 1} are nonempty. But ϕ0 is a continuous function. This means that our above expansions for ϕ and ψ are valid only in the outer region, i.e., the union of Ωξ+(t):={xΩ:ϕ(x,t)1} and Ωξ-(t):={xΩ:ϕ(x,t)0}. The region Ωξ+(t) contains all the points xi (1 ≤ iN). The boundary of the region Ωξ-(t) contains the boundary of Ω.

We assume the inner region, a thin transition layer of width O(ξ) that is complement to the outer region, centers around the closed and smooth interface Γξ(t) = {x ∈ Ω : ϕ(x, t) = 1/2} enclosing Ωξ+(t). This interface is the perturbation of a closed and smooth interface Γ(t) that is also inside the inner region and that does not depend on ξ. We assume that all the principal curvatures of these interfaces are O(1) with respect to ξ → 0. For a given point x in the inner region, we denote by s(x, t) the signed distance from x to Γξ(t), positive interior and negative exterior of the interface. This is a smooth function depending on ξ such that |∇s(x, t)| = 1. Let Ptx ∈ Γξ(t) be the projection of x onto Γξ(t), defined by |s(x, t)| = ||xPtx||. (We use the Euclidean norm and distance.) The projection Pt depends on ξ. For ξ > 0 small enough, one can expect that the projection Ptx ∈ Γξ(t) is unique. We shall assume so. Consequently, the vector xPt(x) is normal to Γξ(t) at Ptx.

We shall show that, to the leading order with respect to ξ → 0, the interface Γ(t), or equivalently the interface Γξ(t), does not move at this time scale. To this end, let us introduce a local coordinate system for the inner region [14, 15]. Consider an open neighborhood D(t) in ℝ3 that covers an O(ξ)-neighborhood of a part of Γξ(t). Let Q(t) ⊂ ℝ2 be a smooth domain and Φ(·, t) : Q(t) → D(t) ∩ Γξ(t) a smooth parameterization of the patch of interface D(t) ∩ Γξ(t) such that the coordinate lines yi = 0 (i = 1, 2) are exactly the lines of principal curvature and the coordinates y1 and y2 of y = (y1, y2) ∈ Q(t) are the corresponding arc lengths of these lines on Γξ(t). Then any point xD(t) of the inner region can be expressed uniquely as

x=Φ(y,t)+ξz(t)n(y,t), (4.12)

where Φ(y, t) is exactly the projection of x onto Γξ(t), z(t) = s(x, t)/ξ, and n(y, t) = ∇xs(x, t) is the unit normal to the interface Γξ(t) at the projection Φ(y, t), pointing from the exterior to interior of the interface Γξ(t). We shall denote by J = J (y, t) the Jacobian matrix of the mapping Φ(·, t) : Q(t) → D(t) ∩ Γξ(t). This is a 3 × 2 matrix at each point y = (y1, y2) with the jth column being the vector yj Φ(y, t) (j = 1, 2). These columns are orthonormal vectors and are tangent to the surface Γξ(t). Note that all D(t), Φ(y, t), J (y, t), z(t), and n(y, t) can depend on ξ. Both Φ(y, t) and J(y, t) are O(1) with respect to ξ → 0. Note also that the z component of the (z, y) coordinate is a global variable, independent of the parameterization Φ(·, t) of the patch of interface D(t) ∩ Γξ(t).

The relation (4.12) determines locally y = y(x, t) and z = z(x, t) as smooth functions of x and t. In particular, z(x, t) = s(x, t). Let v = v(y, t) and H = H(y, t) be the normal velocity and mean curvature of the point Φ(y, t) ∈ Γξ(t) defined by v(y, t) = tΦ(y, t) · n(y, t) and H(y, t) = ∇y · n(y, t)/2, respectively. We have for any x related to (z, y) by (4.12) that

xz(x,t)=ξ-1n(y,t), (4.13)
Δxz(x,t)=2ξ-1H(y,t)+O(1), (4.14)
tz(x,t)=-ξ-1v(y,t), (4.15)
xy(x,t)=J(y,t)+O(ξ). (4.16)

Eq. (4.13) follows from n(y, t) = ∇xs(x, t) and ξz(x, t) = s(x, t). Eq. (4.14) follows from the fact that ∂yjn is the jth column of J multiplied by the jth principal curvature; cf. (2.3) in [15]. Fixing x and replacing z(t) by s(x, t) in (4.12), differentiating both sides of this equation with respect to t, and then dotting with the unit vector n = n(y, t), we then obtain (4.15). Finally, Eq. (4.16) follows from a formula for ∇xy in Lemma 2.1 of [15]. With these, we further obtain by a series of calculations that, for any smooth functions f = f(x, t) and = (z, y, t) that satisfy f(x, t) = (z, y, t) with x and (z, y) related by (4.12),

xf=(Jy+ξ-1nz)f+O(ξ), (4.17)
Δxf=(ξ-12Hz+ξ-2zz2)f+O(1), (4.18)
tf=(t-ξ-1vz)f. (4.19)

We now assume that the solutions ϕ = ϕ(x, t) and ψ = ψ(x, t) have the following expansions in the intersection of D(t) and the inner region [4, 34, 36]:

ϕ(x,t)=ϕ0(z,y,t)+ξϕ1(z,y,t)+ξ2ϕ2(z,y,t)+,ψ(x,t)=ψ0(z,y,t)+ξψ1(z,y,t)+ξ2ψ2(z,y,t)+,

where x and (z, y) are related by (4.12), and all ϕ̃i = ϕ̃i(z, y, t) and ψ̃i = ψ̃i(z, y, t) (i = 0, 1, …) are smooth and bounded functions. Substitute ϕ and ψ in (4.1) with these inner expansions, apply (4.17)–(4.19), and group terms of the same orders in terms of powers of ξ.

Equating the terms of the order O(ξ−2), we get ε′(ϕ̃0)(∂z ψ̃0)2 = 0. Since ϕ0 = 0 or 1 in the outer region, we can assume that the set of points in the inner region at which ϕ̃ = 0 or 1 has the Lebesgue measure zero. Thus, by our assumption (A3) in Section 2, ε′(ϕ̃0) ≠ 0 almost everywhere in the inner region. Consequently, ∂z ψ̃0 = 0 in the inner region. With this, we obtain at the next order O(ξ−1) that

-vzϕ0=γ0[zz2ϕ0-W(ϕ0)]. (4.20)

By matching the inner and outer solutions, we have ϕ̃0(z, y, t) → 0 if z → −∞ and ϕ̃0(z, y, t) → 1 if z → +∞. Moreover, since ϕ̃0 is smooth and bounded, we have also that ∂z ϕ̃0 → 0 as z → −∞ or ∞. Therefore, multiplying both sides of (4.20) by ∂z ϕ̃0 and integrating over z from −∞ and ∞, we obtain

-v-(zϕ0)2dz=γ0[12(zϕ0)2-W(ϕ0)]|z=-z==0.

Hence v = v(y, t) = 0. Therefore, at the leading order of ξ, the interface does not move at the regular time scale. However, in the next long time period measured by a slow time scale, the interface still moves. Note that all these results depend only on the global variable z but not on the local variable y, i.e., not on the parametrization Φ(·, t) : Q(t) → D(t) ∩ Γξ(t).

5. Relaxation Dynamics: Slow Time Scale and Sharp-Interface Limit

We now consider the slow time scale τ = ξt and obtain the leading order of the normal velocity for this interface motion. Based on the previous analysis, we assume at each τ the system region Ω is divided by a closed and smooth interface Γ(τ) into two regions Ω(τ) and Ω+(τ), outside and inside of Γ(τ), respectively. We assume the principal curvatures at any point of Γ(τ) are of order O(1) with respect to ξ → 0. We have ϕ ≈ 0 in Ω(τ) and ϕ ≈ 1 in Ω+(τ). Moreover, all xi ∈ Ω+(τ) (1 ≤ iN). Hence Ω+(τ) approximates the solute region Ωp and Ω(τ) approximates the solvent region Ωw.

We assume that in the outer region—the region outside an O(ξ) neighborhood of Γ(τ)—the solutions ϕ and ψ have the expansions

ϕ(x,t)=ϕ0(x,τ)+ξϕ1(x,τ)+ξ2ϕ2(x,τ)+,ψ(x,t)=ψ0(x,τ)+ξψ1(x,τ)+ξ2ψ2(x,τ)+,

where ϕi = ϕi(x, τ) and ψi = ψi(x, τ) with τ = ξt are smooth and bounded functions that satisfy the boundary conditions (4.7). Plugging these into (4.1) and comparing the terms of the same orders of ξ, we obtain

O(ξ-1):0=W(ϕ0), (5.1)

Similarly, we have by (4.2) that

O(1):·ε(ϕ0)ψ0-(ϕ0-1)2V(ψ0)=-ρf. (5.2)

By (5.1) and the boundary condition for ϕ0, we get ϕ0 = 0 in the Ω(τ) part of the outer region and ϕ0 = 1 in the Ω+(τ) part of the outer region. In particular, as ξ → 0, ϕ0 converges to 1 and 0, inside and outside Γ(τ), respectively.

We now consider the solutions ϕ and ψ in the inner region, an O(ξ)-neighborhood of Γ(τ). As before, we introduce a local coordinate system for the inner region. Let s(x, τ), Pτ(x), D(τ), Q(τ), Φ = Φ(y, τ), J = J(y, τ), z = z(τ), n = n(y, τ), H = H(y, τ), and v = v(y, τ) be all the same as those in Section 4 with τ replacing t and Γ(τ) replacing Γξ(t), respectively. We relate a point xD(τ) in the inner region to the local coordinates (z, y) by

x=Φ(y,τ)+ξz(τ)n(y,τ). (5.3)

We still have (4.13), (4.14), and (4.16) with t replaced by τ. Eq. (4.15) is changed now to ∂tz(x, τ) = −v(x, τ). If f(x, τ) and (z, y, τ) are two smooth functions and f(x, τ) = (z, y, τ) with x and (z, y) related by (5.3), then (4.17) and (4.18) still hold true; but (4.19) is changed to ∂tf = (ξτvz) .

We assume now the following local expansions in the inner region:

ϕ(x,t)=ϕ0(z,y,τ)+ξϕ1(z,y,τ)+ξ2ϕ2(z,y,τ)+,ψ(x,t)=ψ0(z,y,τ)+ξψ1(z,y,τ)+ξ2ψ2(z,y,τ)+,

where ϕ̃i = ϕ̃i(z, y, τ) and ψ̃i = ψ̃i(z, y, τ) (i = 0, 1, …) are smooth and bounded functions, and x and (z, y) are related by (5.3). We also have by (5.3) that

U(x)=U(Φ(y,τ)+ξz(τ)n(y,τ))=U0(y,τ)+O(ξ).

where Ũ0(y, τ) = U(Φ(y, τ)). If x = Φ(y, τ) ∈ Γ(τ) then Ũ0(y, τ) = U(x).

Plug these expansions into (4.1) and equate the terms with the same power of ξ to obtain

O(ξ-2):0=zψ0, (5.4)
O(ξ-1):0=zz2ϕ0-W(ϕ0).O(1):-vzϕ0=-2Pϕ0+γ0[2Hϕ0+zz2ϕ1-W(ϕ0)ϕ1]-2ρw(ϕ0-1)U0 (5.5)
+12ε(ϕ0)[Γ(τ)ψ02+(zψ1)2]+2(ϕ0-1)V(ψ0). (5.6)

In obtaining (5.4), we use the fact that ϕ0 = 0 or 1 in the outer region to assume that ϕ̃0 ∈ (0, 1) and hence ε′(ϕ̃0) ≠ 0 almost everywhere in the inner region. Eq. (5.4) is used to obtain (5.5) and (5.6). By (5.4), ψ̃0(z, y, τ) = ψ̃0(0, y, τ) is the value of ψ̃0 at the point Φ(y, τ) ∈ Γ(τ). Note that the surface graident term ∇Γ(τ)ψ̃0 in (5.6) is an O(ξ)-approximation of Jyψ̃0 by (4.16) and the fact that columns of J are orthogonal to n:

Γ(τ)ψ0=(I-nn)xψ0=(I-nn)Dxyyψ0=(I-nn)Jyψ0+O(ξ)=Jyψ0+O(ξ).

Similarly, plugging the inner expansions of ϕ and ψ into (4.2), we obtain by (5.4) and the fact that columns of J = J(y, τ) are orthogonal to the normal n = n(y, τ) that

O(ξ-1):z[ε(ϕ0)zψ1]=0. (5.7)

Fix x = Φ(y, τ) ∈ Γ(τ). Let us employ the following matching conditions [4, 34, 36]:

limz±ϕ0(z,y,τ)=ϕ0±(x,τ), (5.8)
limz±ψ0(z,y,τ)=ψ0±(x,τ), (5.9)
ψ1(z,y,τ)=(ψ1±+zψ0±·s)(x,τ)+o(1)asz±, (5.10)

where ϕ0±(x,τ)=limα0±ϕ0(x+αs(x,τ)) and ψi±(x,τ)=limα0±ψi(x+αs(x,τ)). Since ϕ0 = 0 in Ω(τ) and ϕ0 = 1 in Ω+(τ), we have by (5.5) and (5.8) that

ϕ0(z,y,τ)=ϕ0(z)=12+e3z-e-3z2(e3z+e-3z). (5.11)

We now derive the following matching condition that will be used later:

zψ1(z,y,τ)=ψ0±(x,τ)·s(x,τ)+o(1)asz±. (5.12)

Let us rewrite the matching condition (5.10) as

ψ1(z,y,τ)=ψ1±(x,τ)+zψ0±(x,τ)·s(x,τ)+g±(z,y,τ)

for some smooth functions g+ = g+(z, y, τ) (z > 0) and g = g(z, y, τ) (z < 0) such that g±(z, y, τ) = o(1) as z → ±∞, respectively. Thus

zψ1(z,y,τ)=ψ0±(x,τ)·s(x,τ)+zg±(z,y,τ). (5.13)

By (5.7), we have ∂zψ̃1(z, y, τ) = h(y, τ)/ε(ϕ̃0(z)) for some smooth and bounded function h = h(y, τ) independent of z. This and (5.13) imply that ∂zg± are bounded as z → ±∞. Moreover, with these two equations and (5.11), we obtain

zz2g±(z,y,τ)=|zz2ψ1(z,y,τ)|=|z[h(y,τ)ε(ψ0(z))]|6εmaxh(y,τ)εmax2e-6z.

Consequently, for any A2 > A1 > 0,

|[zg+(A2)]2-[zg+(A1)]2|=|A1A2z[(zg+)2]dz|2A1A2zg+zzg+dz0

as A1, A2 → ∞. Therefore, (∂zg+)2a+ as z → ∞ for some a+ ≥ 0. We must have a+ = 0, since g+ → 0 as z → ∞. Therefore ∂zg+ → 0 as z → ∞. Similarly ∂zg → 0 as z → −∞. Hence (5.12) follows from (5.13).

We now show that ψ0 solves the Poisson–Boltzmann equation with the dielectric boundary Γ(τ). By (5.4), ψ̃0 is independent of z. Therefore, (5.2), the matching condition (5.9), and the boundary condition ψ0 = ψ on ∂Ω allow us to determine ψ0 in the outer region. Moreover, the matching condition (5.9) implies that ψ0=ψ0+=ψ0- on Γ(τ). Integrating both sides of Eq. (5.7) over z from −∞ to ∞, we obtain by (5.12) that εwε0ψ0-·n=εpε0ψ0+·n on Γ(τ). These, together with (5.2) and (4.7), imply that ψ0 is the solution to the elliptic interface problem

{εwε0Δψ0-V(ψ0)=-ρfinΩ-(τ),εpε0Δψ0=-ρfinΩ+(τ),ψ0-=ψ0+onΓ(τ),εwε0ψ0-·n=εpε0ψ0+·nonΓ(τ),ψ0=ψonΩ. (5.14)

This is equivalent to the boundary-value problem of the Poisson–Boltzmann equation (1.2) and (1.3) with ψ replaced by ψ0 [27].

We finally derive the motion law for the interface Γ(τ). For notational simplicity, let us skip writing the variables y and τ for a moment. Multiplying both sides of (5.6) by ϕ0 and integrating over z from −∞ to ∞, we obtain that

-vS=-R1(z)ϕ0(z)dz+-R2(z)ϕ0(z)dz+-R3(z)ϕ0(z)dz, (5.15)

where by (5.11)

S:=-[ϕ0(z)]2dz=-36e12z(e6z+1)4dz=06x(x+1)4dx=1,R1(z):=-2Pϕ0-2ρw(ϕ0-1)U0+2(ϕ0-1)V(ψ0),R2(z):=γ0[2Hzϕ0+zz2ϕ1-W(ϕ0)ϕ1],R3(z):=12ε(ϕ0)[Γ(τ)ψ02+(zψ1)2]. (5.16)

Since Ũ0 and V(ψ̃0) are independent of z, we have

-R1(z)ϕ0(z)dz=-2P-ϕ0ϕ0dz+2[V(ψ0)-ρwU0]-(ϕ0-1)ϕ0dz=-Pϕ02|z=-z=+[V(ψ0)-ρwU0](ϕ0-1)2|z=-z==-P+ρwU-V(ψ0), (5.17)

where we have on Γ(τ) that Ũ0 = U, and that ψ̃0 = ψ0 by (5.4) and (5.9).

By integration by parts and the fact that ϕ0(±)=ϕ0(±)=0, we have

-zz2ϕ1ϕ0dz=-ϕ0ϕ1dz.

Therefore, by (5.16) and (5.5), we have

-R2(z)ϕ0(z)dz=2γ0H+0-z[ϕ0-W(ϕ0)]ϕ1dz=2γ0H. (5.18)

Now, since ψ0=ψ0-=ψ0+ on Γ(τ), we have Γ(τ)ψ0=Γ(τ)ψ0-=Γ(τ)ψ0+=Γ(τ)ψ0. Consequently,

-R3(z)ϕ0(z)dz=12-ε(ϕ0)[Γ(τ)ψ02+(zψ1)2]ϕ0dz=12(εpε0-εwε0)Γ(τ)ψ02+12-ε(ϕ0)(zψ1)2ϕ0dz. (5.19)

Rewrite (5.7) as ε(ϕ0)zz2ψ1=-ε(ϕ0)ϕ0zψ1. With this and (5.12), we have

12-ε(ϕ0)(zψ1)2ϕ0dz=12ε(ϕ0)(zψ1)2|z=-z=--ε(ϕ0)zψ1zz2ψ1dz=12εpε0ψ0+·n2-12εwε0ψ0-·n2+-ε(ϕ0)(zψ1)2ϕ0dz.

Note that the last term on the right-hand side doubles the term on the left-hand side. This and the third equation in (5.14) then lead to

12-ε(ϕ0)(zψ1)2ϕ0dz=12εwε0ψ0-·n2-12εpε0ψ0+·n2=12(1εwε0-1εpε0)εΓ(τ)ψ0·n2, (5.20)

where εΓ(τ) is defined in (1.4) with Γ(τ) replacing Γ.

Finally, it follows from (5.15)–(5.20) that

v=P-2γ0H-ρwU+12(1εpε0-1εwε0)εΓ(τ)ψ0·n2+12(εwε0-εpε0)Γ(γ)ψ02+V(ψ0).

Since our normal n points from the interior to exterior and mean curvature H = ∇ · n/2, opposite to that in the sharp-interface model (cf. the description above Eq. (1.5)), this is exactly the normal velocity vn = −δΓF[Γ] in the sharp-interface model (cf. (1.5)).

We summarize our analysis in the following:

Theorem 5.1

Assume as above the existence of a closed and smooth interface Γ(τ), the outer and inner expansions for the solutions ϕ and ψ to the initial-boundary-value problem (4.1)–(4.4), and the corresponding matching conditions. Then the following are true:

  1. As ξ → 0, the leading order of the phase-field function ϕ converges to 1 in Ω+(τ) and 0 in Ω(τ), respectively;

  2. The leading-order of the electrostatic potential is the unique solution to the boundary-value problem of the Poisson–Boltzmann equation (1.2) and (1.3);

  3. As ξ → 0, the normal velocity of the interface Γ(τ) converges to the boundary forceδΓF[Γ] in (1.5), with Γ(τ) replacing Γ, as in the sharp-interface model.

Acknowledgments

This work was supported by the US National Institutes of Health (NIH) through the grant R01GM096188 (B.L.), the US National Science Foundation (NSF) through the grant DMS-1319731 (B.L.), and the China Scholar Council (Y.L.). The authors thank Professor John Lowengrub, Dr. Jiayi Wen, and Professor Yanxiang Zhao for many helpful discussions.

Contributor Information

BO LI, Email: bli@math.ucsd.edu.

YUAN LIU, Email: yul400@ucsd.edu.

References

  • 1.Anderson DM, McFadden GB, Wheeler AA. Diffuse-interface methods in fluid mechanics. Ann Rev Fluid Mech. 1998;30:139–165. [Google Scholar]
  • 2.Ben-Yaakov D, Andelman D, Podgornik R. Dielectric decrement as a source of ion-specific effects. J Chem Phys. 2011;134:074705. doi: 10.1063/1.3549915. [DOI] [PubMed] [Google Scholar]
  • 3.Benítez R, Ramírez-Piscina L. Sharp-interface projection of a fluctuating phase-field model. Phys Rev E. 2005;71:061603. doi: 10.1103/PhysRevE.71.061603. [DOI] [PubMed] [Google Scholar]
  • 4.Caginalp G, Fife P. Dynamics of layered interfaces arising from phase boundaries. SIAM J Applied Math. 1988;48:506–518. [Google Scholar]
  • 5.Cai Q, Ye X, Wang J, Luo R. Dielectric boundary forces in numerical Poisson–Boltzmann methods: Theory and numerical strategies. Chem Phys Lett. 2011;514:368–373. doi: 10.1016/j.cplett.2011.08.067. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Chandler D. Interfaces and the driving force of hydrophobic assembly. Nature. 2005;437:640–647. doi: 10.1038/nature04162. [DOI] [PubMed] [Google Scholar]
  • 7.Che J, Dzubiella J, Li B, McCammon JA. Electrostatic free energy and its variations in implicit solvent models. J Phys Chem B. 2008;112:3058–3069. doi: 10.1021/jp7101012. [DOI] [PubMed] [Google Scholar]
  • 8.Cheng HB, Cheng LT, Li B. Yukawa-field approximation of electrostatic free energy and dielectric boundary force. Nonlinearity. 2011;24:3215–3236. doi: 10.1088/0951-7715/24/11/011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Cheng LT, Dzubiella J, McCammon JA, Li B. Application of the level-set method to the implicit solvation of nonpolar molecules. J Chem Phys. 2007;127:084503. doi: 10.1063/1.2757169. [DOI] [PubMed] [Google Scholar]
  • 10.Cheng LT, Li B, Wang Z. Level-set minimization of potential controlled Hadwiger valuations for molecular solvation. J Comput Phys. 2010;229:8497–8510. doi: 10.1016/j.jcp.2010.07.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Cheng LT, Li B, White M, Zhou S. Motion of a cylindrical dielectric boundary. SIAM J Appl Math. 2013;73:594–616. doi: 10.1137/120867986. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Cheng LT, Xie Y, Dzubiella J, McCammon JA, Che J, Li B. Coupling the level-set method with molecular mechanics for variational implicit solvation of nonpolar molecules. J Chem Theory Comput. 2009;5:257–266. doi: 10.1021/ct800297d. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Collins JB, Levine H. Diffuse interface model of diffusion-limited crystal growth. Phys Rev B. 1985;31:6119–6122. doi: 10.1103/physrevb.31.6119. [DOI] [PubMed] [Google Scholar]
  • 14.Dai S, Du Q. Motion of interfaces governed by the Cahn–Hilliard equation with highly disparate diffusion mobility. SIAM J Appl Math. 2012;72(6):1818–1841. [Google Scholar]
  • 15.Dai S, Promislow K. Geometric evolution of bilayers under the functionalized Cahn–Hilliard equation. Proc Royal Soc A. 2013;469:20120505. [Google Scholar]
  • 16.Davis ME, McCammon JA. Electrostatics in biomolecular structure and dynamics. Chem Rev. 1990;90:509–521. [Google Scholar]
  • 17.Dembo A, Funaki T. Lectures on Probability Theory and Statistics. Vol. 1869. Springer–Verlag; 2005. [Google Scholar]
  • 18.Du Q, Liu C, Wang X. A phase field approach in the numerical study of the elastic bending energy for vesicle memberanes. J Comput Phys. 2004;198:450–468. [Google Scholar]
  • 19.Dzubiella J, Swanson JMJ, McCammon JA. Coupling hydrophobicity, dispersion, and electrostatics in continuum solvent models. Phys Rev Lett. 2006;96:087802. doi: 10.1103/PhysRevLett.96.087802. [DOI] [PubMed] [Google Scholar]
  • 20.Dzubiella J, Swanson JMJ, McCammon JA. Coupling nonpolar and polar solvation free energies in implicit solvent models. J Chem Phys. 2006;124:084905. doi: 10.1063/1.2171192. [DOI] [PubMed] [Google Scholar]
  • 21.Gilbarg D, Trudinger NS. Elliptic Partial Differential Equations of Second Order. 2. Springer; 1983. [Google Scholar]
  • 22.Guo Z, Li B, Dzubiella J, Cheng LT, McCammon JA, Che J. Heterogeneous hydration of p53/MDM2 complex. J Chem Theory Comput. 2014;10:1302–1313. doi: 10.1021/ct400967m. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Hatlo MM, van Roij R, Lue L. The electric double layer at high surface potentials: The influence of excess ion polarizability. Europhys Letters. 2012;97(2):28010. [Google Scholar]
  • 24.Karma A, Rappel WJ. Phase-field model of dendritic sidebranching with thermal noise. Phys Rev E. 1999;60:3614–3625. doi: 10.1103/physreve.60.3614. [DOI] [PubMed] [Google Scholar]
  • 25.Langer JS. Models of pattern formation in first-order phase transitions. In: Grinstein G, Mazenko G, editors. Directions in Condensed Matter Physics. World Scientic; 1986. [Google Scholar]
  • 26.Li B. Continuum electrostatics for ionic solutions with nonuniform ionic sizes. Nonlinearity. 2009;22:811–833. [Google Scholar]
  • 27.Li B. Minimization of electrostatic free energy and the Poisson–Boltzmann equation for molecular solvation with implicit solvent. SIAM J Math Anal. 2009;40:2536–2566. [Google Scholar]
  • 28.Li B, Cheng X-L, Zhang Z-F. Dielectric boundary force in molecular solvation with the Poisson–Boltzmann free energy: A shape derivative approach. SIAM J Applied Math. 2011;71:2093–2111. doi: 10.1137/110826436. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Li B, Wen J, Zhou S. Mean-field theory and computation of electrostatics with ionic concentration dependent dielectrics. Commun Math Sci. 2015 doi: 10.4310/CMS.2016.v14.n1.a10. accepted. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Li B, Zhao Y. Variational implicit solvation with solute molecular mechanics: From diffuse-interface to sharp-interface models. SIAM J Applied Math. 2013;73(1):1–23. doi: 10.1137/120883426. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Lowengrub JS, Rätz A, Voigt A. Phase-field modeling of the dynamics of multicomponent vesicles: Spinodal decomposition, coarsening, budding, and fission. Phys Rev E. 2009;79:031926. doi: 10.1103/PhysRevE.79.031926. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Lum K, Chandler D, Weeks JD. Hydrophobicity at small and large length scales. J Phys Chem B. 1999;103:4570–4577. [Google Scholar]
  • 33.Modica L. The gradient theory of phase transitions and the minimal interface criterion. Arch Rational Mech Anal. 1987;98:123–142. [Google Scholar]
  • 34.Pego RL. Front migration in the nonlinear Cahn–Hilliard equation. Proc Roy Soc London Ser A. 1989;422:261–278. [Google Scholar]
  • 35.Roux B, Simonson T. Implicit solvent models. Biophys Chem. 1999;78:1–20. doi: 10.1016/s0301-4622(98)00226-9. [DOI] [PubMed] [Google Scholar]
  • 36.Rubinstein J, Sternberg P, Keller JB. Fast reaction, slow diffusion, and curve shortening. SIAM J Appl Math. 1989;49(1):116–133. [Google Scholar]
  • 37.Shao D, Rappel WJ, Levine H. Computational model for cell morphodynamics. Phys Rev Lett. 2010;105:108104. doi: 10.1103/PhysRevLett.105.108104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Sharp KA, Honig B. Electrostatic interactions in macromolecules: Theory and applications. Annu Rev Biophys Chem. 1990;19:301–332. doi: 10.1146/annurev.bb.19.060190.001505. [DOI] [PubMed] [Google Scholar]
  • 39.Sternberg P. The effect of a singular perturbation on nonconvex variational problems. Arch Rational Mech Anal. 1988;101:209–260. [Google Scholar]
  • 40.Tolman RC. The effect of droplet size on surface tension. J Chem Phys. 1949;17:333–337. [Google Scholar]
  • 41.Wang Z, Che J, Cheng LT, Dzubiella J, Li B, McCammon JA. Level-set variational implicit solvation with the Coulomb-field approximation. J Chem Theory Comput. 2012;8:386–397. doi: 10.1021/ct200647j. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Willard AP, Chandler D. The molecular structure of the interface between water and a hydrophobic substrate is liquid-vapor like. J Chem Phys. 2014;141:18C519. doi: 10.1063/1.4897249. [DOI] [PubMed] [Google Scholar]
  • 43.Zhao Y, Kwan Y, Che J, Li B, McCammon JA. Phase-field approach to implicit solvation of biomolecules with Coulomb-field approximation. J Chem Phys. 2013;139:024111. doi: 10.1063/1.4812839. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Zhou S, Cheng LT, Dzubiella J, Li B, McCammon JA. Variational implicit solvation with Poisson–Boltzmann theory. J Chem Theory Comput. 2014;10(4):1454–1467. doi: 10.1021/ct401058w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Zhou S, Wang Z, Li B. Mean-field description of ionic size effects with non-uniform ionic sizes: A numerical approach. Phys Rev E. 2011;84:021901. doi: 10.1103/PhysRevE.84.021901. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES