Reliable Numerical Solution of a Class of Nonlinear Elliptic Problems Generated by the Poisson–Boltzmann Equation Skip to content
Publicly Available Published by De Gruyter June 13, 2019

Reliable Numerical Solution of a Class of Nonlinear Elliptic Problems Generated by the Poisson–Boltzmann Equation

  • Johannes Kraus EMAIL logo , Svetoslav Nakov and Sergey I. Repin

Abstract

We consider a class of nonlinear elliptic problems associated with models in biophysics, which are described by the Poisson–Boltzmann equation (PBE). We prove mathematical correctness of the problem, study a suitable class of approximations, and deduce guaranteed and fully computable bounds of approximation errors. The latter goal is achieved by means of the approach suggested in [19] for convex variational problems. Moreover, we establish the error identity, which defines the error measure natural for the considered class of problems and show that it yields computable majorants and minorants of the global error as well as indicators of local errors that provide efficient adaptation of meshes. Theoretical results are confirmed by a collection of numerical tests that includes problems on 2D and 3D Lipschitz domains.

1 Introduction

1.1 Classical Statement of the Problem

Let Ωd,d=2,3 be a bounded domain with Lipschitz boundary Ω. We assume that Ω contains an interior subdomain Ω1 with Lipschitz boundary Γ. In general, Ω1 may consist of several disconnected parts (in this case, all of them are assumed to have Lipschitz continuous boundaries). We consider a class of nonlinear elliptic equations motivated by the Poisson–Boltzmann equation (PBE), which is widely used for computation of electrostatic interactions in a system of biomolecules in ionic solution [23, 10, 11],

(1.1)

(1.1a)-(ϵu)+k2sinh(u+w)=linΩ1Ω2,
(1.1b)[u]Γ=0,
(1.1c)[ϵun]Γ=0,
(1.1d)u=0onΩ,

where Ω2:-Ω(Ω1Γ), the coefficients ϵ,kL(Ω), ϵmaxϵϵmin>0, w is measurable, and lL2(Ω). Typically, in biophysical applications, Ω1 is occupied by one or more macromolecules, and Ω2 is occupied by a solution of water and moving ions. The coefficients ϵ and k represent the dielectric constant and the modified Debye–Hückel parameter, and u is the dimensionless electrostatic potential. Concerning the given functions k and w, we can identify three main cases:

  1. kmaxk(x)kmin>0 in Ω and wL(Ω),

  2. k(x)0 in Ω1, kmaxk(x)kmin>0 in Ω2 and wL(Ω2),

  3. k(x)0 in Ω2, kmaxk(x)kmin>0 in Ω1 and wL(Ω1).

Throughout the paper, the major attention is paid to case b, which arises when solving the PBE and which is the most interesting from the practical point of view. Cases a and c can be studied analogously (with some rather obvious modifications). The case with nonhomogeneous Dirichlet boundary condition u=g on Ω can also be treated in this framework provided that the boundary condition is defined as the trace of a function g such that gH1(Ω)L(Ω) and gLs(Ω) with s>max{2,d}.

The ability to find reliable and efficient solutions of the nonlinear Poisson–Boltzmann equation (PBE) for complex geometries of the interior domain Ω1 (with Lipschitz boundary) and piecewise constant dielectrics is important for applications in biophysics and biochemistry, e.g., in modeling the effects of water and ion screening on the potentials in and around soluble proteins, nucleic acids, membranes, and polar molecules and ions; see [23] and the references therein. Although the solution of the linearized PBE (as in the linear Debye–Hückel theory) often yields accurate approximations [22], certain mathematical models are valid only if they are based on the nonlinear PBE.

Over the recent years, adaptive finite element methods have proved to be an adequate technique in the numerical solution of elliptic problems with sharp local features such as point sources, heterogeneous coefficients or nonsmooth boundaries or interfaces (e.g., see [4, 9] and also have been successfully used to solve the nonlinear PBE [5, 14]. Adaptivity heavily relies on reliable and efficient error indicators, which are typically developed in the framework of a posteriori error control methods. While the theory of a posteriori error estimates for linear elliptic partial differential equations is already well established and understood, it is far less developed for nonlinear problems. A posteriori error analysis based on functional estimates has already been successfully applied to variational nonlinear problems including obstacle problems in [20, 21]. The accuracy verification approach taken in this work is also based on arguments that are commonly used in duality theory and convex analysis and can be found, e.g., in [8, 17]. Fast solution methods for systems of nonlinear differential equations is another important issue that affects efficiency of computer simulation methods. Multigrid methods may provide optimal or nearly optimal algorithms (in terms of complexity) to perform this task (e.g., see [18]). However, a systematic discussion of this topic is beyond the framework of this paper.

The main questions studied in the paper are related to the well-posedness of problem (1.1) and a posteriori error estimation of its numerical solution. We use a suitable weak formulation (Definition 2.1), where the nonlinearity does not satisfy any polynomial growth condition, and consequently it does not induce a bounded mapping from H01(Ω) to its dual H-1(Ω). For this (more general) weak formulation, we can guarantee existence of a solution and prove its uniqueness using a result of Brezis and Browder [3]. Additionally, in Proposition 2.1, we show that the solution is bounded (here [3] is used, again together with special test functions suggested in [26, 16]). Boundedness of the solution is important and later used in the derivation of functional a posteriori error estimates. Applying the general approach from [19, 17], we derive guaranteed and computable bounds of the difference between the exact solution and any function from the respective energy class in terms of the energy and combined energy norms (equation (3.20)). Moreover, we obtain an error identity (3.19) with respect to a certain measure for the error which is the sum of the usual combined energy norm |||(v-u)|||2+|||y*-p*|||*2 and a nonlinear measure. In the case of a linear elliptic equation of the form -div(ϵu)+u=l, this nonlinear measure reduces to v-uL2(Ω)2+div(y*-p*)L2(Ω)2, where v and y* are approximations to the exact solution u and the exact flux p*=ϵu. One advantage of the presented error estimate is that it is valid for any conforming approximations of u and ϵu and that it does not rely on Galerkin orthogonality or properties specific to the used numerical method. Another advantage is that only the mathematical structure of the problem is exploited, and therefore no mesh dependent constants are present in the estimate. Majorants of the error not only give guaranteed bounds of global (energy) error norms but also generate efficient error indicators (cf. (1.1a), Figures 12 and 13). Also, we derive a simple, but efficient lower bound for the error in the combined energy norm. Using only the error majorant, we obtain an analog of Cea’s lemma which forms a basis for the a priori convergence analysis of finite element approximations for this class of semilinear problems. Finally, we present three numerical examples that verify the accuracy of error majorants and minorants and confirm efficiency of the error indicator in mesh adaptive procedures.

The outline of the paper is as follows. In Section 2, we discuss correctness of problem (1.1) and prove an a priori L(Ω) estimate for the solution u. In Section 3, first we recall some facts from the duality theory and general a posteriori error estimation method for convex variational problems. Then we apply this abstract framework and derive explicit forms of all the respective terms. A special attention is paid to the general error identity that defines a combined error measure natural for the considered class of problems. At the end of Section 3, we prove convergence of the conforming finite element method based on P1 Lagrange elements. In Section 4, we consider numerical examples associated with 2D and 3D problems and make a systematic comparison of numerical solutions computed by adaptive mesh refinements based on different indicators. The last section includes a summary of the presented results.

2 Variational Form of the Problem

From now on, we assume that the functions k and w fall in case b of Section 1.1.

Definition 2.1.

A function uH01(Ω) is called a weak solution of (1.1) if u is such that b(x,u+w)vL1(Ω) for any vH01(Ω)L(Ω) and

(2.1)a(u,v)+Ωb(x,u+w)vdx=Ωlvdxfor allvH01(Ω)L(Ω),

where a(u,v)=Ωϵuvdx and b(x,z):-k2(x)sinh(z).

Define the functional J:H01(Ω){+} by the relation

(2.2)J(v):-{Ω[ϵ(x)2|v|2+k2cosh(v+w)-lv]dxifk2cosh(v+w)L1(Ω),+ifk2cosh(v+w)L1(Ω),

and consider the variational problem:

(2.3)finduH01(Ω)such thatJ(u)=minvH01(Ω)J(v).

2.1 Existence of a Minimizer

We begin with proving that the variational problem is well-posed. First it is necessary to make some comments on specific features of the above defined variational problem associated with the term k2cosh(v+w). Notice that, for d2, the function evL2(Ω) for all vH01(Ω) (e.g., see [27, 15]) and, therefore, the set dom(J):-{vH01(Ω):J(v)<} is a linear subspace of H01(Ω). However, if d=3, then dom(J) is a convex set but not a linear subspace (Remark 2.1). Since dom(J) is convex and obviously J is convex over domJ, it follows that J is convex over H01(Ω) (e.g., see [8]). Next we note that J is a proper functional, i.e., J is not identically equal to + (e.g., J(0)=Ωk2cosh(w)dx<) and does not take the value - (J(v) is nonnegative). Therefore, existence of the minimizer u is guaranteed by known theorems of the calculus of variations (e.g., see [8]) if we will show that

  1. J is sequentially weakly lower semicontinuous (s.w.l.s.c.), i.e., J(v)lim infnJ(vn) for any sequence {vn}n=1H01(Ω) weakly converging to v in H01(Ω)) (vnv),

  2. J is coercive, i.e., limnJ(vn)=+ whenever vnH1(Ω).

To prove that (1) is fulfilled, notice that J is the sum of the functionals

Ω(ϵ2|v|2-lv)dxandΩk2(x)cosh(v+w)dx.

The first functional is convex and continuous in H01(Ω) and, therefore, it is s.w.l.s.c. (sequentially weakly lower semicontinuous). The second functional is convex and, for d=2, it is Gateaux differentiable, which implies that it is also s.w.l.s.c. (the proof of this implication can be found in [24, Corollary 2.4]). However, for d=3, the functional Ωk2(x)cosh(v+w)dx is not Gateaux differentiable (Remark 2.2). Nevertheless, one can show that it is also s.w.l.s.c. For this purpose, we use Fatou’s lemma and compact embedding of H01(Ω) into L2(Ω).

Let {vn}n=1H01(Ω) be a sequence weakly converging in H01(Ω) to a vH01(Ω), i.e., vnv. Since the embedding H01(Ω)L2(Ω) is compact, it follows that vnv (strongly) in L2(Ω). Therefore, we can extract a subsequence vnm(x)v(x), which converges almost everywhere in the pointwise sense. Recall that k2(x)cosh(z(x)+w(x))0 for all zH01(Ω) and that k2(x)cosh() is a continuous function for a.e. xΩ. Hence, by Fatou’s lemma, we obtain

(2.4)lim infmΩk2(x)cosh(vnm(x)+w(x))dxΩlim infmk2(x)cosh(vnm(x)+w(x))dx=Ωk2(x)cosh(v(x)+w(x))dx.

Now it is clear that if {vnm}m=1 is an arbitrary subsequence of {vn}n=1, then there exists a further subsequence {vnms}s=1 for which (2.4) is satisfied. This means that (2.4) is also satisfied for the whole sequence {vn}n=1, and hence Ωk2cosh(v+w)dx is s.l.w.s.c.

The coercivity of J follows from the estimate

J(v)12a(v,v)-lL2(Ω)vL2(Ω)ϵminvL2(Ω)2-lL2(Ω)vH1(Ω)ϵmin1+CF2vH1(Ω)2-lL2(Ω)vH1(Ω),

where CF is the constant in the Friedrichs inequality vL2(Ω)CFvL2(Ω) for all vH01(Ω).

Thus conditions (1) and (2) are fulfilled, and the existence of a minimizer uH01(Ω) is guaranteed. Moreover, noting that J is strictly convex, we arrive at the following result.

Theorem 2.1.

There exists a unique minimizer uH01(Ω) of problem (2.3).

Remark 2.1.

It is worth noting that dom(J):-{vH01(Ω):k2(x)cosh(v+w)L1(Ω)} is a linear subspace of H01(Ω) for d2 and not a linear subspace of H01(Ω) if d3. In dimension d2, from [27, 15], we know that evL2(Ω) for any vH01(Ω), and thus eλv1+μv2L2(Ω) for any λ,μ and any v1,v2H01(Ω).

On the other hand, if d3, let Ω=B(0,1), and let the inner domain Ω2 be given by Ω2=B(0,r¯) for some r¯<1, where B(0,r) denotes the ball in d with radius r centered at zero. Consider the function v=ln1|x|H01(B(0,1)). Since ev=1|x|L1(Ω2) and eλv=1|x|λL1(Ω2) for any λd,[1] we find on the one hand that

Ωk2cosh(v+w)dx=Ω2k2(ev+w+e-v-w)2dx12kmax2ewL(Ω2)Ω2(ev+e-v)dx12kmax2ewL(Ω2)(Ω2evdx+|Ω2|)<.

On the other hand,

Ωk2cosh(λv+w)dx12Ω2k2eλv+wdx12kmin2e-wL(Ω2)Ω2eλvdx=for anyλ>d.

Hence vdom(J), but λvdom(J) for any λd and, therefore, dom(J) is not a linear subspace. However, dom(J)H01(Ω) is a convex set. Indeed, let v1,v2dom(J), i.e., k2cosh(v1+w),k2cosh(v2+w)L1(Ω). Since k2cosh() is convex for almost all xΩ and any λ[0,1], we have

Ωk2cosh(λv1+(1-λ)v2+w)dxλΩk2cosh(v1+w)dx+(1-λ)Ωk2cosh(v2+w)dx<+.

Hence dom(J) is a convex set.

Remark 2.2.

The functional Ωk2cosh(v+w)dx is not Gateaux differentiable at any uH01(Ω)L(Ω) if d=3 (therefore, J is also not Gateaux differentiable). In fact, Ωk2cosh(v+w)dx is discontinuous at every uH01(Ω)L(Ω). This fact is easy to see by the following example. Let Ω2=B(0,2)Ω be the ball centered at 0 with radius 2. There exists a function vH01(Ω) such that Ω2eλvdx=+ for any λ>0. In particular, we can set v=ϕ|x|-1/3, where ϕ is a smooth function equal to 1 in B(0,1) and 0 in 3B(0,2). Then vH01(Ω), but eλvL1(Ω2) for any λ>0 since eλv>|x|-3 for small enough |x|. In this case, for any uH01(Ω)L(Ω) and any λ>0, we have

Ωk2cosh(u+λv+w)dx12Ω2k2eu+λv+wdxkmin2e-u+wL(Ω2)2Ω2eλvdx=+.

Now our goal is to show that the minimizer u is a solution of (2.1). To prove this, we use the Lebesgue dominated convergence theorem and the fact that, at the unique minimizer u of J, we have k2cosh(u+w)L1(Ω). Since J(u+λv)-J(u)0 for all vH01(Ω)L(Ω) and any λ0, we have

(2.5)12a(u+λv,u+λv)+Ωk2cosh(u+λv+w)dx-Ωl(u+λv)dx-12a(u,u)-Ωk2cosh(u+w)dx+Ωludx0.

Making equivalent transformations of (2.5) and dividing by λ>0, we obtain

(2.6)a(u,v)+limλ0+Ωk2(cosh(u+λv+w)-cosh(u+w))λdx-Ωlvdx0.

To compute the limit in the second term of (2.6), we will apply the Lebesgue dominated convergence theorem. We have

(2.7)fλ(x):-k2(x)(cosh(u(x)+w(x)+λv(x))-cosh(u(x)+w(x)))λλ0+k2(x)sinh(u(x)+w(x))v(x)for a.e.xΩ.

By the mean value theorem, we obtain

fλ(x)=k2(x)sinh(ς(x))v(x),

where ς(x):-u(x)+w(x)+Θ(x)λv(x) and Θ(x)(0,1), a.e. xΩ. Then

|fλ(x)|k2(x)(eς(x)+e-ς(x)2)|v(x)|,

from which it follows that

(2.8)|fλ|vL(Ω)k2eu+w+e-u-w2e|Θ(x)|λ|v(x)|vL(Ω)evL(Ω)k2cosh(u+w)L1(Ω)for allλ1.

From the Lebesgue dominated convergence theorem, (2.7) and (2.8), it follows that the limit in (2.6) is equal to Ωk2sinh(u+w)vdx, and therefore we obtain

(2.9)a(u,v)+Ωb(x,u+w)vdx-Ωlvdx0for allvH01(Ω)L(Ω).

Since the test functions belong to a linear manifold, (2.9) is equivalent to the weak formulation (2.1).

2.2 Uniqueness of the Solution to (2.1)

Uniqueness of the solution of (2.1) follows from the monotonicity of b(x,):

(2.10)Ω(b(x,v+w)-b(x,z+w))(v-z)dx0for allv,zH01(Ω)L(Ω).

If u1,u2H01(Ω) are two different solutions of (2.1), then

(2.11)a(u1-u2,v)+Ω(b(x,u1+w)-b(x,u2+w))vdx=0for allvH01(Ω)L(Ω).

Note that the difference u1-u2 is not necessarily in H01(Ω)L(Ω). To show that we can test with u1-u2 in (2.11), we apply a property of Sobolev spaces proved in [3].

Theorem 2.2.

Let Ω be an open set in Rd, TH-1(Ω)Lloc1(Ω), and vH01(Ω). If there exists a function fL1(Ω) such that T(x)v(x)f(x) a.e in Ω, then TvL1(Ω) and the duality product T,v in H-1(Ω)×H01(Ω) coincides with ΩTvdx.

We have the following situation: a locally summable function gLloc1(Ω) defines a bounded linear functional Tg over the dense subspace C0(Ω) of H01(Ω) through the integral formula Tg,φ=Ωgφdx. It is clear that the functional Tg is uniquely extendable by continuity to a bounded linear functional T¯g over the whole space H01(Ω). The question is whether this extension is still representable by the same integral formula for any vH01(Ω) (if the integral makes sense at all). If the function vH01(Ω) is fixed, then Theorem 2.2 gives a sufficient condition for gv to be summable and for the extension T¯g evaluated at v to be representable with the same integral formula as above, i.e., T¯g,v=Ωgvdx.

Now, applying Theorem 2.2 to the functional Tg defined by

Tg,v:-b(x,u1+w)-b(x,u2+w),vfor allvH01(Ω)L(Ω)

and the function v=u1-u2H01(Ω), using (2.11), we conclude that

a(u1-u2,u1-u2)+Ω(b(x,u1+w)-b(x,u2+w))(u1-u2)dx=0.

Using the monotonicity (2.10) of b(x,) and the coercivity of a(,), we obtain u1=u2.

2.3 Boundedness of the Minimizer

Next we show that the solution to problem (2.1) is essentially bounded. To prove this, we need the following lemma [16].

Lemma 2.1.

Let φ(t) be a nonnegative function, which is nonincreasing for s0t< and such that

φ(h)Cφ(s)β(h-s)αfor allh>s>s0,

where C and α are positive constants and β>1. If jR is defined by jα:-Cφ(s0)β-12αββ-1, then φ(s0+j)=0.

Now we present a main result of this section.

Proposition 2.1.

The unique weak solution u to problem (2.1) belongs to L(Ω). Moreover, there exists a positive constant j¯>0, depending only on d, Ω, lL2(Ω), ϵmin, such that uL(Ω)wL(Ω2)+j¯. If l=0, then the constant j¯ is equal to zero.

Proof.

To prove that u is bounded, we apply Theorem 2.2 once again.

The first step is to show that (2.1) holds for the test function

(2.12)v=Gs(u):-sgn(u)max{|u|-s,0},

where swL(Ω2) (we notice that similar test functions Gs have been used in [16, Theorem B.2] in the context of linear elliptic problems).

It is easy to see that Gs(0)=0, this function is Lipschitz continuous and, therefore, Gs(u)H01(Ω) (e.g., see [16, 12]). Next the functional Tb defined by

Tb,v:-Ωb(x,u+w)vdxfor allvH01(Ω)L(Ω)

is bounded and linear and b(x,u+w)Lloc1(Ω). This follows from (2.1) and from the fact that the functionals a(u,) and (l,) belong to H-1(Ω). In view of Theorem 2.2, to show that Tb,Gs(u)=Ωb(x,u+w)Gs(u)dx, it suffices to verify the inequality

(2.13)b(x,u+w)Gs(u)fa.e. for somefL1(Ω).

Choosing swL(Ω2), using the monotonicity of b(x,), and the fact that b(x,0)=0, we obtain

(2.14)b(x,u+w)Gs(u)={b(x,u+w)(u-s)0foru>s,0foru[-s,s],b(x,u+w)(u+s)0foru<-s,

which shows that assumption (2.13) holds for f=0.

Now we are ready to prove that uL(Ω). First we consider the case l=0. From (2.14), it follows that

(2.15)Ωb(x,u+w)Gs(u)dx0.

Moreover, using the definition of a(,) and the definition (2.12) of Gs(u), we obtain

(2.16)a(u,Gs(u))=ΩϵuGs(u)dx=ΩϵGs(u)Gs(u)dxϵminGs(u)L2(Ω)2ϵminCF2Gs(u)L2(Ω)2,

where CF is the constant in Friedrichs’ inequality vL2(Ω)CFvL2(Ω) that holds for all vH01(Ω). Finally, using (2.1), (2.15), and (2.16), we get Gs(u)L2(Ω)20 for all swL(Ω2). Consequently, |u|s almost everywhere and for all swL(Ω2). In the case where l is not identically zero in Ω, we further estimate a(u,Gs(u)) from below and ΩlGs(u)dx from above using the Sobolev embedding H1(Ω)Lq(Ω), where q= for d=1, q< for d=2, and q=2dd-2 for d3. Let q* denote the Hölder conjugate to q. Then q*=1 for d=1, q*=qq-1>1 for d=2, and q*=2dd+2 for d>2. In order to treat both cases in which we are interested simultaneously, namely, d=2,3, we can take q=6 and q*=6/5. By CE we denote the embedding constant in the inequality vL6(Ω)CEvH1(Ω) for all vH1(Ω), which depends only on the domain Ω and d. For a(u,Gs(u)), we have

(2.17)a(u,Gs(u))=ΩϵGs(u)Gs(u)dxϵmin1+CF2Gs(u)H1(Ω)2

and, for ΩlGs(u)dx, we obtain

(2.18)ΩlGs(u)dx=A(s)lGs(u)dxlLq*(A(s))Gs(u)Lq(Ω)CElLq*(A(s))Gs(u)H1(Ω),

where A(s):-{xΩ:|u(x)|>s}. Combining (2.17), (2.18), (2.15), and (2.1), we arrive at the estimate

(2.19)ϵmin1+CF2Gs(u)H1(Ω)CElLq*(A(s)).

The final step before applying Lemma 2.1 is to estimate the left-hand side of (2.19) from below in terms of |A(h)| for h>swL(Ω2) and the right-hand side of (2.19) from above in terms of |A(s)|. Again, using the Sobolev embedding H1(Ω)Lq(Ω) and Hölder’s inequality yields

(2.20)Gs(u)H1(Ω)1CE(Ω|Gs(u)|qdx)1q=1CE(A(s)||u|-s|qdx)1q1CE(A(h)(h-s)qdx)1q=1CE(h-s)|A(h)|1q

and

(2.21)lLq*(A(s))lL2(Ω)|A(s)|2-q*2q*.

Combining (2.20), (2.21), and (2.19), we obtain the following inequality for the nonnegative and non-increasing function φ(t):-|A(t)|:

|A(h)|(CE2(1+CF2)ϵminlL2(Ω))q|A(s)|q-22(h-s)qfor allh>swL(Ω2).

Since q-22=2>1, applying Lemma 2.1, we conclude that there is some j>0 such that

0<jq=(CE2(1+CF2)ϵminlL2(Ω))q|A(wL(Ω2))|q-422q(q-2)q-4(CE2(1+CF2)ϵminlL2(Ω))q|Ω|q-422q(q-2)q-4-:j¯q

and |A(wL(Ω2)+j¯)|=0. Hence uL(Ω)wL(Ω2)+j¯. ∎

The results of this section are summarized in the following theorem.

Theorem 2.3.

Problem (2.1) has a unique solution uH01(Ω)L(Ω), which is the unique minimizer of variational problem (2.3).

Remark 2.3.

Since k=0 in Ω1, wL(Ω2) and uL(Ω), we conclude that (2.1) holds for all vH01(Ω) resulting in a standard weak formulation. If k2 is uniformly positive in the whole domain Ω and wL(Ω), then uL(Ω)wL(Ω)+j¯. On the other hand, if k=0 in Ω2, k2 is uniformly positive in Ω1, and wL(Ω1), we have uL(Ω)wL(Ω1)+j¯.

3 A Posteriori Error Estimates

3.1 Abstract Framework

First we briefly recall some results from the duality theory [17, 8]. Consider a class of variational problems having the following common form:

(P)finduVsuch thatJ(u)=infvVJ(v),whereJ(v)=G(Λv)+F(v).

Here V,Y are reflexive Banach spaces with the norms V and Y, respectively, F:V¯, G:Y¯ are convex and proper functionals, and Λ:VY is a bounded linear operator. By 0V we denote the zero element in V. It is assumed that J is coercive and lower semicontinuous. In this case, problem (P) has a solution u, which is unique if J is strictly convex.

The spaces topologically dual to V and Y are denoted by V* and Y*, respectively. They are endowed with the norms V* and Y*. Henceforth, v*,v denotes the duality product of v*V* and vV. Analogously, (y*,y) is the duality product of y*Y* and yY, and Λ*:Y*V* is the operator adjoint to Λ. It is defined by the relation

Λ*y*,v=(y*,Λv)for allvV,y*Y*.

The functional J*:V*¯ defined by the relation

J*(v*):-supvV{v*,v-J(v)}

is called dual (or Fenchel conjugate) to J (see, e.g., [8]). In accordance with the general duality theory of the calculus of variations, the primal problem (P) has a dual counterpart:

(P*)findp*Y*such thatI*(p*)=supy*Y*I*(y*),whereI*(y*):--G*(y*)-F*(-Λ*y*),

where G* and F* are the functionals conjugate to G and F, respectively. Problems (P) and (P*) are generated by the Lagrangian L:V×Y*¯ defined by the relation L(v,y*)=(y*,Λv)-G*(y*)+F(v). If we additionally assume that G* is coercive and that F(0V) is finite, then it is well known that problems (P) and (P*) have unique solutions uV and p*Y* and that strong duality relations hold (see [17], or the book [8, Proposition 2.3, Remark 2.3, and Proposition 1.2 in Chapter VI]):

J(u)=infvVJ(v)=infvVsupy*Y*L(v,y*)=supy*Y*infvVL(v,y*)=supy*Y*I*(y*)=I*(p*).

Furthermore, the pair (u,p*) is a saddle point of the Lagrangian L, i.e.,

L(u,y*)L(u,p*)L(v,p*)for allvV,y*Y*,

and u and p* satisfy the relations

ΛuG*(p*),p*G(Λu).

We have

(3.1)J(v)-I*(y*)=G(Λv)+F(v)+G*(y*)+F*(-Λ*y*)=DG(Λv,y*)+DF(v,-Λ*y*)-:M2(v,y*),

where

DG(Λv,y*):-G(Λv)+G*(y*)-(y*,Λv),
DF(v,-Λ*y*):-F(v)+F*(-Λ*y*)+Λ*y*,v

are the compound functionals for G and F, respectively [17]. A compound functional is nonnegative by the definition. Equality (3.1) shows that DG and DF can vanish simultaneously if and only if v=u and y*=p*. Moreover, setting v:-u and y*:-p* in (3.1), we obtain analogous identities for the primal and dual parts of the error:

(3.2)

(3.2a)J(u)-I^*(y^*)=M2(u,y*)=DG(Λu,y*)+DF(u,-Λ*y*),
(3.2b)J(v)-I*(p*)=M2(v,p*)=DG(Λv,p*)+DF(v,-Λ*p*).

Using the fact that J(u)=I*(p*) and that the above equalities (3.2) hold, we obtain another important relation (see [17])

(3.3)M2(v,y*)=J(v)-I*(y*)=J(v)-I*(p*)+J(u)-I*(y*)=M2(v,p*)+M2(u,y*).

Notice that M2(v,y*) depends on the approximations v and y* only and, therefore, is fully computable. The right-hand side of (3.3) can be viewed as a certain measure of the distance between (u,p*) and (v,y*), which vanishes if and only if v=u and y*=p*. Hence the relation

(3.4)DG(Λv,p*)+DF(v,-Λ*p*)+DG(Λu,y*)+DF(u,-Λ*y*)=M2(v,y*)

establishes the equality of the computable term M2(v,y*) and an error measure natural for this class of variational problems.

It is worth noting that identity (3.4) can be represented in terms of norms if G and F are quadratic functionals. For example, if V=H01(Ω), V*=H-1(Ω), Y=[L2(Ω)]d=Y*, G(Λv)=G(v)=Ω12Avvdx and F(v)=Ω(12v2-lv)dx (where A is a symmetric positive definite matrix with bounded entries), then

(3.5)

DG(Λv,p*)=12ΩA(v-u)(v-u)dx,DG(Λu,y*)=12ΩA-1(y*-p*)(y*-p*)dx,
DF(v,-Λ*p*)=12v-uL2(Ω)2,DF(u,-Λ*y*)=12div(y*-p*)L2(Ω)2.

In this case, the minimizer of (P) solves the linear elliptic problem -div(Au)+u=l in Ω, and (3.4) is reduced to the error identity

(3.6)ΩA(v-u)(v-u)dx+ΩA-1(y*-p*)(y*-p*)dx+v-uL2(Ω)2+div(y*-p*)L2(Ω)2=|||Av-y*|||*2+v-divy*-lL2(Ω)2=2M2(v,y*).

The sum of the first and the third term in (3.6) represents the primal, the sum of the second and fourth term the dual error.

We set V:-H01(Ω), Y:-[L2(Ω)]d, where d=2,3, and Λ the gradient operator :H01(Ω)[L2(Ω)]d. We further denote g:Ω×3, g(x,ξ):-ϵ(x)2|ξ|2, and B:Ω×, B(x,ξ):-k2(x)cosh(ξ). With this notation, we have

G(Λv):-Ωg(x,v(x))dx=Ωϵ2|v|2dx,
F(v):-ΩB(x,v(x)+w(x))dx-Ωlvdx=Ωk2cosh(v+w)dx-Ωlvdx,

and the functional J, defined by (2.2), can be written in the form J(v)=G(Λv)+F(v). For any vV the functional G(Λv) is finite, while F:V{+} may take the value + for some vV if d3 (e.g., v=log1|x|α, αd on the unit ball in d). However, if d2, then exp(v)L1(Ω) for all vH01(Ω) and F:V (see [15]). Also, F(0V) is obviously finite since wL(Ω2). We set V*=H-1(Ω) and Y*=Y=[L2(Ω)]d. In this case, Λ* coincides with -div considered as an operator from [L2(Ω)]d to H-1(Ω). We will present the particular form of error equality (3.4) where the error is measured in a special “nonlinear norm”. This measure contains the usual combined energy norm terms, i.e., the sum of the energy norms of the errors for the primal and dual problem, but also two additional nonnegative terms due to the nonlinearity B(x,ξ) (or equivalently b(x,ξ)) which in some cases may dominate the usual energy norm terms. We start by deriving explicit expressions for G*,F*, and then we will use these expressions to get an explicit form of the abstract error equality (3.4).

3.2 Fenchel Conjugates of the Functionals G and F

It is easy to find that G*(y*)=Ω12ϵ(x)|y*(x)|2dx. For y*H(div;Ω) and an arbitrary function z:Ω2, we introduce the functional

Iy*(z):-Ω2[(divy*+l)z-B(x,z+w)]dx.

Recalling that the nonlinearity B is supported on Ω2, we have

(3.7)F*(-Λ*y*)=supzH01(Ω)[-Λ*y*,z-F(z)]=supzH01(Ω)[(-y*,Λz)-F(z)]=supzH01(Ω)Ω[-y*z-B(x,z+w)+lz]dx(ify*H(div;Ω))=supzH01(Ω)Ω[divy*z-B(x,z+w)+lz]dx(finite ifdivy*+l=0inΩ1)=supzH01(Ω)Iy*(z)Ω2supξ[(divy*(x)+l(x))ξ-B(x,ξ+w(x))]dx=Ω2[(divy*(x)+l(x))ξ0(x)-B(x,ξ0(x)+w(x))]dx=Iy*(ξ0).

Here ξ0:Ω2 is computed by the condition

(3.8)ddξ[(divy*(x)+l(x))ξ-B(x,ξ+w(x))]=0for a.e.xΩ2,

which is equivalent to

divy*(x)+l(x)-k2(x)sinh(ξ+w(x))=0for a.e.xΩ2.

We notice that (3.8) is a necessary condition for a maximum which is also sufficient since B(x,) is convex. The solution of the last equation exists, is unique, and is given by

ξ0(x)=arsinh(ρk(y*))-w(x)=ln(ρk(y*)+ρk2(y*)+1)-w(x)=ln(Θ(ρk(y*)))-w(x),

where ρk(y*):-divy*(x)+l(x)k2(x) and Θ(s):-s+s2+1 for s. Note that the exact solution p*=ϵu of dual problem (P*) also satisfies the relation div(ϵu)+l=0 because, for any xΩ1, it holds k(x)=0. Moreover, since uL(Ω), wL(Ω2), and lL2(Ω), we see that divp*=k2sinh(u+w)+lL2(Ω) and thus p*H(div;Ω). In Proposition 3.1, we will later prove that we have not overestimated the supremum over zH01(Ω) in (3.7) and that we actually have equalities everywhere. Denoting S:-arsinh(ρk(y*)) and using the expression for ξ0(x) and the formula

cosh(arsinh(x))=x2+1for allx,

for any y*H(div;Ω)[L2(Ω)]d=Y* with divy*+l=0 in Ω1, we obtain an explicit formula for F*(-Λ*y*):

(3.9)F*(-Λ*y*)=Ω2[k2ρk(y*)(ln(Θ(ρk(y*)))-w)-k2ρk2(y*)+1]dx=Ω2[k2sinh(S)(S-w)-k2cosh(S)]dx.

Remark 3.1.

Since |ln(t+t2+1)||t| for all t, the function ln(Θ(f(x)))-w(x) belongs to L2(Ω2) for any fL2(Ω2), and we conclude that ξ0(x)L2(Ω2) if y*H(div;Ω). Therefore, the integral in (3.9) is well defined.

Now our goal is to prove that the inequality supzH01(Ω)Iy*(z)Iy*(ξ0) holds as the equality. In other words, we want to prove that the error estimate remains sharp and that the computed majorant M2(v,y*) will be indeed zero if approximations (v,y*) coincide with the exact solution (u,p*).

Proposition 3.1.

For any y*H(div;Ω) with divy*+l=0 in Ω1, it holds

supzH01(Ω)Iy*(z)=Iy*(ξ0)<.

Proof.

The idea is to approximate f=divy*+lk2L2(Ω2) and wΩ2L(Ω2) by C0(Ω2) functions (in the a.e. sense) and use the Lebesgue dominated convergence theorem. Let fnC0(Ω2) and wnC0(Ω2) be such that fn(x)f(x) a.e. in Ω2, |fn(x)|h(x)L2(Ω2) (see [2, Theorem 4.9]), wn(x)w(x) a.e. in Ω2, |wn(x)|m+2, where m:-wL(Ω2). Then

zn(x):-ln(Θ(fn(x)))-wn(x)ξ0(x)a.e. inΩ2

and znC0(Ω2)H01(Ω2)H01(Ω) (by extending the functions by zero in Ω1). Since B(x,) is continuous, we have the pointwise a.e. in Ω2 convergence

(divy*(x)+l(x))zn(x)-B(x,zn+w(x))(divy*(x)+l(x))ξ0(x)-B(x,ξ0(x)+w(x))

Now we search for a function in L1(Ω2) that majorates the function |(divy*(x)+l(x))zn(x)-B(x,zn+w(x))|:

(3.10)|(divy*(x)+l(x))zn(x)-k2(x)cosh(zn(x)+w(x))||divy*(x)+l(x)||zn(x)|+k2(x)ewL(Ω2)e|zn(x)|.

Our next goal is to bound |zn(x)| in (3.10). For the first summand, we have

|zn(x)|=|ln(Θ(fn(x)))-wn(x)||fn(x)|+m+2h(x)+m+2L2(Ω2),

where Remark 3.1 has been used. However, this bound cannot be used in the second term because eh might not belong even to L1(Ω2). In order to find an L1-majorant for the second summand in (3.10), we distinguish the following two cases: In the first case, fn(x)>0. Then |ln(Θ(fn(x)))||ln(Θ(h(x)))|.

In the second case (fn(x)0), we have Θ(fn(x))1. Therefore, 0fn(x)-h(x). Since Θ(s) is a monotonically increasing function, Θ(0)=1Θ(fn(x))Θ(-h(x))>0. From here, we obtain

ln(1)=0ln(Θ(fn(x)))ln(Θ(-h(x))),

and using the relation Θ(-h)=1Θ(h), we conclude that

|ln(Θ(fn(x)))||ln(Θ(-h(x)))|=|ln(Θ(h(x)))|.

Finally, for almost all xΩ2, we have

|zn(x)|=|ln(Θ(fn(x)))-wn(x)||ln(Θ(h(x)))|+m+2=ln(Θ(h(x)))+m+2becauseh(x)0for a.e.xΩ2.

Therefore,

|(divy*(x)+l(x))zn(x)-k2(x)cosh(zn(x)+w(x))||divy*(x)+l(x)|(h(x)+wL(Ω2)+2)+k2(x)e2wL(Ω2)+2Θ(h(x)):-H(x)L2(Ω2),

where, in the last line, we used the fact that Θ(h(x))L2(Ω2). All the conditions of the Lebesgue dominated convergence theorem are satisfied, and we see that Iy*(zn)Iy*(ξ0) and, consequently,

supzH01(Ω)Iy*(z)=Iy*(ξ0).

3.3 Error Measures

In this section, we apply the abstract framework from Section 3.1 and derive an explicit form of relation (3.4) adapted to our problem. For any y*H(div;Ω) with divy*+l=0 in Ω1, the quantity M2(v,y*) is fully computable and is given by the relation

(3.11)M2(v,y*)=DG(Λv,y*)+DF(v,-Λ*y*)=G(Λv)+G*(y*)-(y*,Λv)+F(v)+F*(-Λ*y*)+Λ*y*,v=Ωη2(x)dx=12|||ϵv-y*|||*2+DF(v,-Λ*y*),

where

(3.12)η2(x)={12ϵ|ϵv-y*|2,xΩ112ϵ|ϵv-y*|2+k2cosh(v+w)-lv+k2ρk(y*)(ln(Θ(ρk(y*)))-w)-k2ρk2(y*)+1-divy*v,xΩ2

and we have used the expressions for G* and F*. It is clear that η2(x)0 since it is the sum of the compound functionals generated by g~x(s):-g(x,s) and B~x(s)-l(x)s=B(x,s+w(x))-l(x)s evaluated at (v(x),y*(x)) and (v(x),divy*(x)), respectively. It therefore qualifies as an error indicator, provided that y* is chosen appropriately, which we demonstrate with numerical experiments in the next section. Using the expression for G*, we obtain

(3.13) D_G(v,p^*)=12Ωϵ|(v-u)|2dx-:12|||(v-u)|||2,
DG(Λu,y*)=12Ω1ϵ|y*-p*|2dx-:12|||y*-p*|||*2.

Now we find explicit expressions for the nonlinear measures DF(v,-Λ*p*) and DF(u,-Λ*y*) similar to the ones for the case of quadratic F in (3.5) for the linear elliptic equation -div(Au)+u=l. We will need the following assertion, which is easy to prove.

Proposition 3.2.

For all s,tR, it holds

(3.14)(t-s)22A(s,t)(sinh(t)-sinh(s))22,

where A(s,t)=cosh(t)-cosh(s)+ssinh(s)-tsinh(s).

Since, for the exact solution u, we have ρk(p*)=sinh(u+w) and u=arsinh(ρk(p*))-w for a.e. xΩ2, we find that

DF(v,-Λ*p*)=Ω2(k2cosh(v+w)-lv+k2sinh(u+w)u-k2cosh(u+w)-divp*v)dx=Ω2k2(cosh(v+w)-cosh(u+w)+usinh(u+w)-vsinh(u+w))dx.

Similarly, DF(u,-Λ*y*)=Ω2k2(cosh(T)-cosh(S)+Ssinh(S)-Tsinh(S))dx, where T:-arsinh(ρk(p*)). The nonlinear quantities DF(v,-Λ*p*) and DF(u,-Λ*y*) measure the error in v and in divy*, respectively. Using inequality (3.14), we can represent these two measures in a form which resembles the corresponding estimates in the case (3.5) of a quadratic functional F, namely,

(3.15)Ω2k22(v-u)2dxDF(v,-Λ*p*)Ω2k22(sinh(v+w)-sinh(u+w))2dx,
(3.16)Ω2k22(T-S)2dxDF(u,-Λ*y*)Ω212k2(divp*-divy*)2dx.

Note that, for kkmin>0 in Ω, the following equivalences hold:

Ωk22(v-u)2dxv-uL2(Ω)2andΩ12k2(divp*-divy*)2dxdivy*-divp*L2(Ω)2.

Moreover, replacing the nonlinear term k2sinh(u+w) with u, inequalities (3.15) and (3.16) reduce to the equalities for DF(v,-Λ*p*) and DF(u,-Λ*y*) in (3.5) because, in this case, the inverse function of f(x)=x is again f(x). The functions on the left-hand side, in the middle, and on the right-hand side in inequality (3.14) are depicted on Figure 1.

Figure 1 Functions in inequality (3.14).
Figure 1

Functions in inequality (3.14).

Further, if v is in a δ1-neighborhood of u in L(Ω) norm, then we can find a constant C1(δ1,uL(Ω))>1 such that

(3.17)Ω2k22(sinh(v+w)-sinh(u+w))2dxC1(δ1,uL(Ω))Ω2k22(v-u)2dx.

Analogously, if lL(Ω2) and div(y*-p*)L(Ω2)δ2 (recall that when lL(Ω2), divp* is in L(Ω2)), then we can find a constant C2(δ2,divp*L(Ω2))<1 such that

(3.18)C2(δ2,divp*L(Ω2))Ω212k2(divp*-divy*)2dxΩ2k22(T-S)2dx.

The constants C1 and C2 are just Lipschitz constants for the locally Lipschitz function sinh. Notice that if k2kmin>0 in Ω, then everywhere in (3.15), (3.16), (3.17), and (3.18), the integrals are taken over the entire domain Ω. Now the abstract error identity (3.4) takes the form

(3.19)12|||(u-v)|||2+12|||p*-y*|||*2+Ω2k22(v-u)2dx+C2(δ2,divp*L(Ω))Ω212k2(divp*-divy*)2dx12|||(u-v)|||2+12|||p*-y*|||*2+DF(v,-Λ*p*)+DF(u,-Λ*y*)=M2(v,y*)¯12|||(u-v)|||2+12|||p*-y*|||*2+C1(δ1,uL(Ω))Ω2k22(v-u)2dx+Ω212k2(divy*-divp*)2dx,

where we have used p*=ϵΛu=ϵu. Relation (3.19) shows that the computable majorant M2(v,y*) is bounded from below and above by a multiple of one and the same error norm. Since DF(v,-Λ*p*)0 and DF(u,-Λ*y*)0, we also obtain a guaranteed bound on the error in the combined energy norm,

(3.20)|||(u-v)|||2+|||p*-y*|||*22M2(v,y*).

From the pointwise equality

(3.21)1ϵ|ϵv-y*|2=1ϵ|ϵ(v-u)-(y*-p*)|2=ϵ|(v-u)|2+1ϵ|y*-p*|2-2(y*-p*)(v-u),

after applying Young’s inequality and integrating over Ω, we obtain a lower bound for the error in combined energy norm,

(3.22)12|||ϵv-y*|||*2|||(v-u)|||2+|||y*-p*|||*2

Remark 3.2.

Integrating (3.21) over Ω, we obtain the algebraic identity

(3.23)|||ϵv-y*|||*2=|||(v-u)|||2+|||y*-p*|||*2-2Ω(y*-p*)(v-u)dx,

from which the Prager–Synge identity is derived. Comparing the last relation with (3.19), using the fact that M(v,y*)2=12|||ϵv-y*|||*2+DF(v,-Λ*y*), we arrive at the relation

(3.24)DF(v,-Λ*y*)=DF(v,-Λ*p*)+DF(u,-Λ*y*)+Ω(y*-p*)(v-u)dx.

From here, it is seen that if the integral on the right-hand side is small compared to the other terms, then the error in v and divy* measured with DF(v,-Λ*p*)+DF(u,-Λ*y*) is controlled mainly by the computable term DF(v,-Λ*y*) in the majorant M2(v,y*). Moreover, (3.23) enables us to give a practical estimation of the error in combined energy norm, which is very close to the real error in all of the experiments that we have conducted.

We conclude the section by presenting a near best approximation result. Contrary to the result in [5, Theorem 6.2], we do not make any restrictive assumptions on the meshes to ensure that the finite element approximations uh are uniformly bounded in L norm. In our analysis, VhL is a finite-dimensional subspace of H01, and uh is the minimizer of J over Vh, which is the unique solution of the Galerkin problem:

(3.25)finduhVhsuch thata(uh,v)+Ωb(x,uh+w)vdx=(l,v)for allvVh.

Then, using (3.2b) and expression (3.13) for DG(Λv,p*), for any vVh, we can write

|||(uh-u)|||2+2DF(uh,-Λ*p*)=2(J(uh)-J(u))2(J(v)-J(u))=|||(v-u)|||2+2DF(v,-Λ*p*).

Since 2DF(uh,-Λ*p*)0, using (3.15), we obtain the following generalization of Cea’s lemma to the case of our nonlinear problem.

Proposition 3.3.

Let VhL(Ω) be a closed subspace of H01(Ω) and uhVh the Galerkin approximation of u defined by (3.25). Then

(3.26)|||(uh-u)|||2infvVh{|||(v-u)|||2+Ω2k2(sinh(v+w)-sinh(u+w))2dx}.

Since we use the finite element method with P1 Lagrange elements, let Vh be the corresponding space, where h refers to the maximum element size. By Ih(φ), we denote the Lagrange finite element interpolant of φC0(Ω). Using (3.26), we can show unqualified convergence of the finite element approximations uh to u as h0. Let ε>0, and u¯C0(Ω) is such that (u¯-u)L2(Ω)ε and u¯L(Ω)uL(Ω)+2. Also, let L be the Lipschitz constant in the inequality

|sinh(s)-sinh(t)|L|s-t|for alls,t[-2wL(Ω2)-j¯-2,2wL(Ω2)+j¯+2],

where j¯ is the constant from Proposition 2.1. Then, applying the triangle inequality together with Young’s inequality, we obtain

(3.27)|||(uh-u)|||22(|||(Ih(u¯)-u¯)|||2+|||(u¯-u)|||2)+2(Ωk2(sinh(Ih(u¯)+w)-sinh(u¯+w))2dx+Ωk2(sinh(u¯+w)-sinh(u+w))2dx).

For the first term in (3.27), assuming mesh regularity, we have

|||(Ih(u¯)-u¯)|||2+|||(u¯-u)|||2ϵmax(C|u¯|22h2+ε2),

where |u¯|2 denotes the H2 seminorm of u¯ and C>0 is a constant depending on the mesh regularity. Using the fact that Ih(u¯)L(Ω)u¯L(Ω)uL(Ω)+2, for the second term in (3.27), we obtain the upper bound

2kmax2L2(Ih(u¯)-u¯L2(Ω)2+u¯-uL2(Ω)2)2kmax2L2CF2((Ih(u¯)-u¯)L2(Ω)2+(u¯-u)L2(Ω)2)2kmax2L2CF2(C|u¯|22h2+ε2).

This inequality shows that the right-hand side of (3.27) can be made as small as desired provided that we choose ε and h small enough, and therefore |||(uh-u)|||0 when h0. Moreover, (3.26) can be also used to obtain qualified convergence of uh in the energy norm under additional assumptions on the interface Γ, the meshes, and the regularity of u.

4 Numerical Results

In this section, we present numerical examples illustrating error identity (3.19) and performance of functional a posteriori error estimates. All numerical experiments are carried out in FreeFem++ developed and maintained by Frederich Hecht [13], and all pictures are generated in VisIt [6]. We solve adaptively the homogeneous nonlinear problem (1.1) with w:-whref=g-zhref, where zhref is a good Galerkin finite element approximation of the solution z of

-(ϵz)=-k2sinh(g)+linΩ1Ω2,
[z]Γ=0,
[ϵzn]Γ=0,
z=0onΩ,

for given functions g and l. We compare the accuracy of the adaptively computed solution uh of (1.1) for w=whref to the reference solution zhref. The adaptive mesh refinement is based on the error indicator 2ηL2(Oi) on subdomains Oi, where the function η is defined in (3.12) and η2 is the integrand of the majorant M2(v,y*). The factor 2 accounts for the factor 2 in (3.20). More precisely, we find approximations uh to the exact solution uH01(Ω) of the problem

Ωϵuvdx+Ωb(x,u+whref)vdx=Ωlvdx=0for allvH01(Ω).

In all examples, we use piecewise constant parameters ϵ and k, and for y*H(div;Ω), we used a patchwise equilibrated reconstruction of the numerical flux ϵuh based on [1]. More precisely, we find y* in the Raviart–Thomas space RT0 over the same mesh such that its divergence is equal to the L2 orthogonal projection of k2sinh(uh+w)+l onto the space of piecewise constants.

Recall that

M2(v,y*)=M2(v,p*)+M2(u,y*),

where M2(v,y*)=12|||ϵv-y*|||*2+DF(v,-Λ*y*) and M2(v,p*)=J(v)-J(u)=12|||(v-u)|||2+DF(v,-Λ*p*) is the primal error, whereas M2(u,y*)=I*(p*)-I*(y*)=12|||y*-p*|||*2+DF(u,-Λ*y*) is the dual error. Further, we use v for the approximate solution uh and u for the reference solution zhref and define the efficiency index of the lower bound for the error in combined energy norm (3.22) by

IEffCEN,Low:-22|||ϵv-y*|||*|||(v-u)|||2+|||y*-p*|||*2.

Similarly,

IEffCEN,Up:-2M2(v,y*)|||(v-u)|||2+|||y*-p*|||*2

defines the efficiency index of the upper bound (3.20) for the error in combined energy norm. Finally,

IEffE:-2M2(v,y*)|||(v-u)|||andPrelCEN:-|||ϵv-y*|||*|||v|||2+|||y*|||*2

define the efficiency index of the upper bound for the error in energy norm and the practical estimate of the relative error in combined energy norm, respectively.

4.1 Example 1 (2D Problem)

In the first example, the domain Ω is a square with a side 20 with Ω1 being a regular 15-sided polygon with a radius of its circumscribed circle equal to 2. The coefficients ϵ and k are defined by the relations

ϵ(x)={ϵ1=1,xΩ1,ϵ2=100,xΩ2,  k(x)={k1=0.15,xΩ1,k2=0.4,xΩ2,

and

g=L(exp(-b1((x1-c1)2σ12-1))-exp(-b2((x2-c2)2σ22-1))),

l=0, where b1=2=b2=2, c1=-1, c2=6, σ1=σ2=1.5, L=0.8. The reference solution zhref is computed on a multiply refined mesh with 50 086 142 triangles. Note that k2=0.0225 in Ω1 and k2=0.16 in Ω2. The mesh adaptation is done with the built-in function “adaptmesh” of FreeFem++. The localized error indicator 2ηL2(Oi), computed on each vertex patch Oi of the mesh, is compared to its average value over all patches, and the local mesh size is divided by two if this average is smaller than the local value.

Table 1 illustrates the main error identity (3.3) and the convergence of its constituent parts. Further, it is seen that the dual error 2M2(u,y*) dominates the primal error in this example. This is due to the fact that the term 2DF(u,-Λ*y*), measuring the error in divy* (cf. (3.16) and (3.18)), is much larger than |||(v-u)|||2+DF(v,-Λ*p*), where DF(v,-Λ*p*) behaves like v-uL2(Ω2)2 (cf. (3.15) and (3.17)). As we mentioned earlier, for y*, we use a partially equilibrated reconstruction of the numerical flux ϵv, which is the reason why the integral term in (3.23) is negligible compared to the combined energy norm of the error. This fact is confirmed by the values of the efficiency index of the lower bound (3.22).

Table 1

Constituent parts of main error identity (3.3) for Example 1 (2D).

Example 1 (2D): k1=0.15, k2=0.4, ϵ1=1, ϵ2=100
# eltsv-u0u0 [%]|||(v-u)||||||u||| [%]|||y*-p*|||*|||p*|||* [%]2M2(v,y*)2M2(v,p*)2M2(u,y*)
19615.007751.558286.10211778.1466.59801711.54
3475.6933930.853441.7241703.59420.7780682.816
6304.2038421.771531.4858217.71910.2201207.498
13152.3955215.853223.124476.80185.3757471.4261
28651.8707511.735317.165533.93102.9441430.9869
59380.646117.9300111.469216.08121.3387414.7425
120060.369855.647868.235447.752320.678727.07360
245710.160233.942415.760543.852680.330393.52229
484830.089092.802654.093661.900430.166821.73361
974230.039611.978752.884550.962750.083040.87970
1929050.022301.398322.032000.475240.041360.43388
3861850.010150.994711.446160.241340.020820.22052
Table 2

Constituent parts of error identity (3.6) for example Example 1 (2D).

Example 1 (2D): k1=0.15, k2=0.4, ϵ1=1, ϵ2=100
# elts|||(v-u)|||2|||y*-p*|||*22DF(v,-Λ*p*)2DF(u,-Λ*y*)
19656.5057157.58810.09231553.95
34720.235037.00580.54296645.811
63010.075621.07290.14450186.425
13155.3423511.36680.0333860.0593
28652.927426.263380.0167124.7235
59381.336732.796190.0020011.9462
120060.678051.441690.000675.63191
245710.330380.705380.000012.81691
484830.166960.356220.000001.37739
974230.083230.176870.000000.70283
1929050.041560.087770.000000.34611
3861850.021030.044450.000000.17606

In Table 3, we can see that IEffCEN,Low is approximately equal to 0.707122. The value of the efficiency index with respect to the combined energy norm and the value of the ratio DF(v,-Λ*y*)/M2(v,y*) are also coupled in the sense that if we have only one of these two quantities, we can estimate the other one by using the main error equality (3.19). This estimation is accurate because the integral term in (3.24) is very close to zero, and therefore DF(v,-Λ*y*)DF(v-Λ*p*)+DF(u-Λ*y*). One more consequence of using a partially equilibrated flux is that we obtain a very accurate practical estimate of the absolute and relative error in combined energy norm as illustrated in the last two columns of Table 3.

Table 3

Efficiency indices for Example 1 (2D).

Example 1 (2D): k1=0.15, k2=0.4, ϵ1=1, ϵ2=100
# eltsDF(v,-Λ*y*)M2(v,y*) [%]IEffCEN,LowIEffCEN,UpIEffE,UpPrelCEN [%]True rel. error in CEN [%]
19689.07010.673712.881915.6096674.697370.9641
34792.49420.679193.505975.8967136.263836.6935
63085.95250.700662.643804.6484827.157427.0680
131578.26160.706812.143923.7915819.938319.8250
286572.89920.707291.921423.4045214.752314.7032
593874.30090.707081.972563.468469.874199.85973
1200672.64730.707221.912383.381307.067627.06119
2457173.11760.707081.928643.414854.937534.93591
4848372.48260.706941.905883.373713.507893.50805
9742373.00840.706781.923923.401082.472562.47347
19290572.84860.706291.916923.381451.742261.74418
38618572.99120.705461.919723.387481.238291.24114

Figure 2 depicts a mesh that is a part of a sequence of meshes obtained by mesh adaptation using the localized functional error indicator 2ηL2(Oi). Figure 3 depicts a mesh with approximately the same number of elements but obtained by mesh adaptation using the error indicator |||ϵv-y*|||*(Oi). The mesh in Figure 2 is refined mainly where the error in divy* is the dominant part of the error M2(v,-Λ*p*)+M2(u,-Λ*y*). On the other hand, the mesh in Figure 3 is refined most around the extrema of the solution. Figure 5 depicts the minimal set of elements K of a mesh Th that contains at least 30 % of the total indicated error KTh|||ϵv-y*|||*(K) (greedy algorithm with a bulk factor of 0.3), where Th is part of the same sequence as the mesh illustrated in Figure 3.

Figure 2 Mesh on the 9th level of AMR (97 423 elements) based on the error indicator ∥2⁢η∥L2⁢(Oi){\lVert\sqrt{2}\eta\rVert_{L^{2}(O_{i})}} with flux equilibration for y*{y^{*}}.
Figure 2

Mesh on the 9th level of AMR (97 423 elements) based on the error indicator 2ηL2(Oi) with flux equilibration for y*.

Figure 3 Mesh on the 9th level of AMR (97 353 elements) based on the error indicator |||ϵ⁢∇⁡v-y*|||*(Oi){\lvert\kern-1.2pt\lvert\kern-1.2pt\lvert\epsilon\nabla v-y^{*}\rvert\kern-1.2%
pt\rvert\kern-1.2pt\rvert_{*(O_{i})}} with flux equilibration for y*{y^{*}}.
Figure 3

Mesh on the 9th level of AMR (97 353 elements) based on the error indicator |||ϵv-y*|||*(Oi) with flux equilibration for y*.

Figure 4 Reference solution for Example 1 (2D).
Figure 4

Reference solution for Example 1 (2D).

Figure 5 Mesh on the 7th level of AMR (24 122 elements) based on the error indicator |||ϵ⁢∇⁡v-y*|||*(Oi){\lvert\kern-1.2pt\lvert\kern-1.2pt\lvert\epsilon\nabla v-y^{*}\rvert\kern-1.2%
pt\rvert\kern-1.2pt\rvert_{*(O_{i})}} with flux equilibration for y*{y^{*}}.
The elements are marked by applying the error indicator ∥2⁢η∥L2⁢(K){\lVert\sqrt{2}\eta\rVert_{L^{2}(K)}} and using the greedy algorithm with bulk factor 0.3.
Figure 5

Mesh on the 7th level of AMR (24 122 elements) based on the error indicator |||ϵv-y*|||*(Oi) with flux equilibration for y*. The elements are marked by applying the error indicator 2ηL2(K) and using the greedy algorithm with bulk factor 0.3.

Figure 6 depicts the elements marked by the greedy algorithm using a bulk factor of 0.5 and employing the true error

2M2(v,p*)+2M2(u,y*)

as indicator. Figure 7 depicts elements which are marked additionally or fail to be marked by the same greedy algorithm when employing the functional error indicator 2ηL2(Oi) for the same bulk factor. The ratio of the number of these differently marked elements, that is, elements which are marked by one of the two methods but not by the other one, and the total number of elements is 0.022, and the ratio of the number of differently marked elements to the number of marked elements using the true error is 0.048 (Table 4). Comparing the indicated error and the true error elementwise, one finds that the error indicator generated by the majorant M2(v,y*) reproduces the local distribution of the error with a very high accuracy. This is also confirmed by Figure 8, where it can be seen that all error measures are almost identical in both cases of adaptive mesh refinement. Mesh adaptation based on the functional error indicator 2ηL2(Oi) instead of the error indicator |||ϵv-y*|||*(Oi) (Figure 9) yields approximately twice smaller efficiency indices in energy and combined energy norms and approximately twice smaller values for the full error M2(v,p*)+M2(u,y*) on meshes with a comparable number of elements. The reason for the higher efficiency indices is that no adaptive control is applied on the nonlinear part of the error measure in (3.19), and consequently, the ratio DF(v,-Λ*y*)/M2(v,y*) is increasing, reaching values close to 100 % on fine meshes. However, the error in |||(v-u)||| and |||y*-p*|||* might be a little higher in the case of the functional error indicator 2ηL2(Oi). For example, on the mesh from Figure 5 with 24 122 elements, M2(v,p*)+M2(u,y*)=3.8314, |||(v-u)|||=0.4674, |||y*-p*|||*=0.6540, whereas on a mesh with 24 571 elements from the sequence adapted with the indicator 2ηL2(Oi), we obtained a value of 1.9263 for M2(v,y*), and 0.574791 and 0.8399 for |||(v-u)||| and |||y*-p*|||*, respectively. This shows that, reducing the error in divy*, the functional error indicator 2ηL2(Oi) provides a better approximation for the primal and dual problem together.

Figure 6 Mesh on the 2nd level of AMR (630 elements) based on the error indicator
∥2⁢η∥L2⁢(Oi){\lVert\sqrt{2}\eta\rVert_{L^{2}(O_{i})}} with flux equilibration for y*{y^{*}}.
The elements are marked by applying the true error 2⁢M⊕2⁢(v,p*)+2⁢M⊕2⁢(u,y*){\sqrt{2M_{\oplus}^{2}(v,p^{*})+2M_{\oplus}^{2}(u,y^{*})}} as an indicator using greedy algorithm with bulk factor 0.5.
Figure 6

Mesh on the 2nd level of AMR (630 elements) based on the error indicator 2ηL2(Oi) with flux equilibration for y*. The elements are marked by applying the true error 2M2(v,p*)+2M2(u,y*) as an indicator using greedy algorithm with bulk factor 0.5.

Figure 7 Mesh on the 2nd level of AMR (630 elements) based on the error indicator ∥2⁢η∥L2⁢(Oi){\lVert\sqrt{2}\eta\rVert_{L^{2}(O_{i})}} with flux equilibration for y*{y^{*}}.
Here we mark red those elements, which differ in the markings based on the indicator ∥2⁢η∥L2⁢(K){\lVert\sqrt{2}\eta\rVert_{L^{2}(K)}} and on the true error 2⁢M⊕2⁢(v,p*)+2⁢M⊕2⁢(u,y*){\sqrt{2M_{\oplus}^{2}(v,p^{*})+2M_{\oplus}^{2}(u,y^{*})}}.
Marking is done by greedy algorithm with bulk factor 0.5.
Figure 7

Mesh on the 2nd level of AMR (630 elements) based on the error indicator 2ηL2(Oi) with flux equilibration for y*. Here we mark red those elements, which differ in the markings based on the indicator 2ηL2(K) and on the true error 2M2(v,p*)+2M2(u,y*). Marking is done by greedy algorithm with bulk factor 0.5.

Figure 8 Comparison of errors for AMR based on the functional error indicator ∥2⁢η∥L2⁢(Oi){\lVert\sqrt{2}\eta\rVert_{L^{2}(O_{i})}} versus AMR based on the indicator generated by the true error 2⁢M⊕2⁢(v,p*)+2⁢M⊕2⁢(u,y*){\sqrt{2M_{\oplus}^{2}(v,p^{*})+2M_{\oplus}^{2}(u,y^{*})}}.
Figure 8

Comparison of errors for AMR based on the functional error indicator 2ηL2(Oi) versus AMR based on the indicator generated by the true error 2M2(v,p*)+2M2(u,y*).

Figure 9 Comparison of errors for AMR based on the functional error indicator ∥2⁢η∥L2⁢(Oi){\lVert\sqrt{2}\eta\rVert_{L^{2}(O_{i})}} versus AMR based on the indicator |||ϵ⁢∇⁡v-y*|||*(Oi){\lvert\kern-1.2pt\lvert\kern-1.2pt\lvert\epsilon\nabla v-y^{*}\rvert\kern-1.2%
pt\rvert\kern-1.2pt\rvert_{*(O_{i})}}.
Figure 9

Comparison of errors for AMR based on the functional error indicator 2ηL2(Oi) versus AMR based on the indicator |||ϵv-y*|||*(Oi).

Table 4

Marking based on true error and functional error indicator in Example 1 (2D).

Example 1 (2D): k1=0.15, k2=0.4, ϵ1=1, ϵ2=100
# elts# marked elts with true error# differently marked eltsdifferently marked elts in % of all mesh elts
1966263.06122
347150102.88184
630288142.22222
1315632392.96578
286514391133.94415
593829492163.63759
1200659815344.44778
24571120999613.91111
484832419422334.60574
974234778440124.11812

Now we want to demonstrate that flux equilibration is indeed an important subtask to make the proposed error bounds reliable and efficient. For this purpose, we use a simple global gradient averaging procedure, i.e., project the numerical flux ϵvL2(Ω) onto the subspace [Vh]2, where Vh is the finite element space of continuous piecewise linear functions. Then the problem in Example 1 is solved adaptively once applying the functional error indicator 2ηL2(Oi) and next applying the error indicator |||ϵv-y*|||*(Oi). Figure 10 shows the adapted mesh with 563 965 elements, which is a part of a sequence of meshes obtained applying the functional error indicator with gradient averaging for y*. Figure 11 shows a mesh with 444 092 elements, which is part of a sequence of meshes adapted using the second indicator with gradient averaging for y*. Comparing with the results based on flux equilibration for y*, it can be seen that the mesh in Ω2 close to the interface Γ is refined too much for both error indicators. Apart from that, the meshes on Figures 11 and 3 look quite similar, unlike the meshes on Figures 10 and 2. For meshes with a comparable number of elements, applying the indicator |||ϵv-y*|||*(Oi) using gradient averaging instead of flux equilibration, we obtained  30 % larger values for the error |||(v-u)||| and 60 % larger values for the error |||y*-p*|||*. The difference in the errors when applying the functional indicator 2ηL2(Oi) with gradient averaging for y* instead of flux equilibration resulted in an even more drastic increase of the error, namely, between 40 % and 180 % for |||(v-u)||| and between 64 % and 66 % for |||y*-p*|||*, where the meshes had between 21 528 and 563 965 elements. In both cases, we obtained an increasing sequence of efficiency indices with respect to energy and combined energy norms reaching values of 133 and 107 with the functional error indicator on a mesh with 2 089 022 elements, and 570 and 269 with the error indicator |||ϵv-y*|||*(Oi) on a mesh with 2 954 218 elements. This is due to the fact that the nonlinear term DF(u,-Λ*y*), which measures the error in divy* (see (3.16) and (3.18)), dominates the other terms in the nonlinear measure M2(v,p*)+M2(u,y*) for the error, reaching more than 99.99 % of it in both cases. In both experiments with gradient averaging for y*, increasing values of DF(u,-Λ*y*) are in correspondence with increasing error divy*-divp*L2(Ω) and increasing efficiency indices.

Figure 10 Mesh with 563 965 elements, adapted using the error indicator ∥2⁢η∥L2⁢(Oi){\lVert\sqrt{2}\eta\rVert_{L^{2}(O_{i})}} with gradient averaging for y*{y^{*}}.
Figure 10

Mesh with 563 965 elements, adapted using the error indicator 2ηL2(Oi) with gradient averaging for y*.

Figure 11 Mesh with 444 092 elements, adapted using the error indicator |||ϵ⁢∇⁡v-y*|||*(Oi){\lvert\kern-1.2pt\lvert\kern-1.2pt\lvert\epsilon\nabla v-y^{*}\rvert\kern-1.2%
pt\rvert\kern-1.2pt\rvert_{*(O_{i})}} with gradient averaging for y*{y^{*}}.
Figure 11

Mesh with 444 092 elements, adapted using the error indicator |||ϵv-y*|||*(Oi) with gradient averaging for y*.

4.2 Example 2 (2D Problem)

Figures 13 and 15 show how meshes depend on the indicator in another example, where ϵ1=1ϵ2=100, k1=0.2, k2=0.3. The functions

g=exp(-b1(|x-c1|2σ12-1))-exp(-b2(|x-c2|2σ22-1))andl=exp(-b3(|x|2σ32-1))sin(x1x24),

where b1=2.2, b2=2.5, b3=6, c1=(-1,0), c2=(5,5), σ1=σ2=2, σ3=10. The indicator |||ϵv-y*|||*(Oi) correctly approximates elementwise errors in the combined energy norm but does not capture the rest of the error, which results from the nonlinearity k2sinh(u+w) and the right-hand side l in (1.1). On the other hand, the term DF(v,-Λ*y*) controls the error DF(v,-Λ*p*)+DF(u,-Λ*y*), and this is the reason why the mesh on Figure 13 resembles the wavy features of the function f=-k2sinh(u+w)+l.

The isolines of the reference solution and of the function f are depicted on Figures 14 and 12.

Figure 12 Function f=-k2⁢sinh⁡(u+w)+l{f=-k^{2}\sinh(u+w)+l}.
Figure 12

Function f=-k2sinh(u+w)+l.

Figure 13 Mesh with 395 935 elements, obtained by AMR using the error indicator ∥2⁢η∥L2⁢(Oi){\lVert\sqrt{2}\eta\rVert_{L^{2}(O_{i})}} with
flux equilibration for y*{y^{*}}.
Figure 13

Mesh with 395 935 elements, obtained by AMR using the error indicator 2ηL2(Oi) with flux equilibration for y*.

Figure 14 Reference solution.
Figure 14

Reference solution.

Figure 15 Mesh with 555 489 elements, obtained by AMR using the error indicator |||ϵ⁢∇⁡v-y*|||*(Oi){\lvert\kern-1.2pt\lvert\kern-1.2pt\lvert\epsilon\nabla v-y^{*}\rvert\kern-1.2%
pt\rvert\kern-1.2pt\rvert_{*(O_{i})}} with flux equilibration for y*{y^{*}}.
Figure 15

Mesh with 555 489 elements, obtained by AMR using the error indicator |||ϵv-y*|||*(Oi) with flux equilibration for y*.

4.3 Example 3 (3D Problem)

Here we consider an example close to a real physical problem. The computational domain Ω is a cube of side length 20 angstrom with a triangulated water molecule Ω1 in it. The diameter of the water molecule, which is positioned in the center of the cube, is about 2.75 angstrom. Its shape is not changed during the mesh adaptation process. The surface mesh of the water molecule is taken from [28]. Figure 16 illustrates the initial tetrahedral mesh, which consists of 60 222 elements. It is generated using TetGen [25] and adaptively refined with the help of mmg3d [7]. Using the localized error indicator 2ηL2(Oi) computed on each vertex patch Oi of the mesh, a new local mesh size at each vertex is defined by the formula

hinew=hiold(max{min{AM{2ηL2(Oj)}2ηL2(Oi),1},0.35})

and supplied to mmg3d, where AM{2ηL2(Oj)} is the arithmetic mean of {2ηL2(Oj)} over all vertex patches Oj. The coefficients ϵ and k for this example are typical for electrostatic computations in biophysics using the PBE and are given by

ϵ(x)={ϵ1=2,xΩ1,ϵ2=80,xΩ2,  k(x)={k1=0,xΩ1,k2=0.84,xΩ2.

We consider the homogeneous problem, i.e., l=0, and

g=exp(-b1(|x-c1|2σ12-1))-exp(-b2(|x-c2|2σ22-1))+exp(-b3(|x-c3|2σ32-1))+exp(-b4(|x-c4|2σ42-1)),

where b1=b2=b3=b4=2.3, c1=(1,1,0), c2=(4,4,0), c2=(0,6,0), c2=(-5,0,0), σ1=σ2=σ3=σ4=2. The reference solution zhref is computed on a very fine mesh, obtained after a sequence of adaptive mesh refinements, that contains 79 917 007 tetrahedrons.

Figure 16 Initial mesh in Example 3 consisting of 60 222 tetrahedrons.
Figure 16

Initial mesh in Example 3 consisting of 60 222 tetrahedrons.

Figure 17 Ratio of the error indicator |||ϵ⁢∇⁡v-y*|||*{\lvert\kern-1.2pt\lvert\kern-1.2pt\lvert\epsilon\nabla v-y^{*}\rvert\kern-1.2%
pt\rvert\kern-1.2pt\rvert_{*}} and combined energy norm of the error, elementwise.
Mesh on the 4th level of AMR (1.1736e+06 elements) in Example 3 using the error indicator ∥2⁢η∥L2⁢(Oi){\lVert\sqrt{2}\eta\rVert_{L^{2}(O_{i})}} with flux equilibration for y*{y^{*}}.
Figure 17

Ratio of the error indicator |||ϵv-y*|||* and combined energy norm of the error, elementwise. Mesh on the 4th level of AMR (1.1736e+06 elements) in Example 3 using the error indicator 2ηL2(Oi) with flux equilibration for y*.

Since l=0 in Ω1 is a constant function, the patchwise reconstruction from [1] produces a flux y* with zero divergence in Ω1 and, therefore, the reliability of our majorant is guaranteed. In this example, we achieved a very tight guaranteed bound on the error in combined energy norm, as well as in energy norm. The efficiency index IEffCEN,Up settles at around 1.05 and the efficiency index IEffE,Up decreases to 1.30 (Table 6). This is in a good agreement with the fact that, in this example, the ratio DF(v,-Λ*y*)/M2(v,y*) is well controlled and decreases to around 10 % (column 2 in Table 7). We also note that, in this example, we obtained very similar results with the error indicator |||ϵv-y*|||*(Oi). For the efficiency index IEffCEN,Low of the lower bound on the combined energy norm of the error, we obtain values converging to approximately 0.7071, which is the approximate value of 22 (column 3 in Table 6). This means that the combined energy norm of the error

|||(v-u)|||2+|||y*-p*|||*2

is practically equal to |||ϵv-y*|||*.

Table 5

Constituent parts of main error identity (3.3) for Example 3 (3D).

Example 3 (3D): k1=0, k2=0.84, ϵ1=2, ϵ2=80
# eltsv-u0u0 [%]|||(v-u)||||||u||| [%]|||y*-p*|||*|||p*|||* [%]2M2(v,y*)2M2(v,p*)2M2(u,y*)
6022276.8320108.015167.589425569117373308196
10323611.925746.330655.121047104.517845.029259.5
2221181.0923317.735314.95784484.442224.692259.75
5529360.498208.672227.09062965.067513.706451.361
1.17360e+060.256096.580755.33661539.734295.254244.481
2.05668e+060.170945.376254.18207350.648197.016153.631
2.97315e+060.123174.734663.53852265.167152.783112.385
3.90692e+060.100714.328863.12966216.336127.70388.6336
Table 6

Constituent parts of error identity (3.6) for Example 3 (3D).

Example 3 (3D): k1=0, k2=0.84, ϵ1=2, ϵ2=80
# elts|||(v-u)|||2|||y*-p*|||*22DF(v,-Λ*p*)2DF(u,-Λ*y*)REUp [%]
6022279487.019134637886.0116850
10323614623.920699.73221.128559.78310.049
2221182142.921524.2881.7757735.47433.9219
552936512.376342.5281.32980108.83313.4714
1.17360e+06295.039194.0260.2145850.4559.75647
2.05668e+06196.919119.1550.0974334.47627.72193
2.97315e+06152.72485.30440.0585727.08056.64970
3.90692e+06127.66666.73030.0366321.90335.96873
Table 7

Efficiency indices for Example 3 (3D).

Example 3 (3D): k1=0, k2=0.84, ϵ1=2, ϵ2=80
# eltsDF(v,-Λ*y*)M2(v,y*) [%]IEffCEN,LowIEffCEN,UpIEffE,UpPrelCEN [%]True rel. error in CEN [%]
6022240.05410.686271.253532.3138692.8434140.985
10323620.45000.728281.154781.7947347.687050.9159
22211816.11720.716151.105831.4466116.404016.4054
55293611.22490.707861.062481.372417.909667.92099
1.17360e+069.334770.707311.050531.352545.985055.99106
2.05668e+069.822890.707251.053271.334424.813434.81632
2.97315e+0610.20570.707221.055471.317674.177844.17960
3.90692e+0610.11940.707191.054921.301753.775923.77716

Another consequence of this fact is the good accuracy of the practical estimation PrelCEN of the relative error in combined energy norm (columns 6 and 7 in Table 6). The tight bounds on the error also enable us to compute tight and guaranteed upper bounds on the relative error in energy norm:

(4.1)|||(v-u)||||||u|||2M2(v,y*)|||v|||-2M2(v,y*)-:REUp,

where (4.1) is valid if |||v|||-2M2(v,y*)>0.

As a remark, we note that the efficiency indices with respect to the energy and combined energy norms of the error can be improved if we use a flux reconstruction in a bigger space, say, RT1, which has better approximation properties. In this way, the error in divy* will decrease and, as a result, the term DF(v,-Λ*y*) and consequently the dual part of the error 2M2(u,y*)=|||y*-p*|||*2+DF(u,-Λ*y*) will constitute a smaller part of the whole majorant and the error, respectively. Even better, we can minimize the majorant with respect to y* in a subspace of H(div;Ω) like RT0, possibly on another finer mesh. Note that, in the limit case, we have infy*H(div;Ω)M2(v,y*)=M2(v,p*)=12|||(v-u)|||2+DF(v,-Λ*p*), and the dual error completely vanishes. In this case,

IEffCEN,Up=IEffE=2M2(v,p*)|||(v-u)|||=|||(v-u)|||2+2DF(v,-Λ*p*)|||(v-u)|||,

where the last ratio tends to 1 because, by (3.15) and (3.17), the term DF(v,-Λ*p*)v-uL2(Ω2)2 and has a higher order of convergence than |||(v-u)|||2. In practice, we can minimize the majorant with respect to y* only once on a sufficiently big subspace of H(div;Ω) to find some good approximation y¯* of p* and then reuse this y¯* and obtain guaranteed and tight bounds on the error in energy and combined energy norm.

A key factor that determines the efficiency index is the ratio DF(v,-Λ*y*)M2(v,y*). Assuming that

DF(v,-Λ*y*)DF(v,-Λ*p*)+DF(u,-Λ*y*),

which means that the last term in (3.24) is close to zero, we obtain from (3.19) the estimate

IEffCEN,Up11-DF(v,-Λ*y*)M2(v,y*).

From what we have observed, the efficiency index IEffE,Up with respect to the energy norm usually is no more than twice bigger than IEffCEN,Up (assuming we have a good approximation y* to p*). Therefore, if, during the computations, we detect that this ratio is increasing, we can apply the so-called estimation with one step delay, i.e., compute the value of the majorant M2(v,y*) for the current mesh level with the reconstructed y* from the next level.

5 Conclusions

We proved the existence and uniqueness of a solution u of the nonlinear elliptic problem (1.1), which appears in the context of solving the nonlinear PBE numerically by two- or three-term regularization. Also, we established the L(Ω) regularity of the solution u and an analog of Cea’s lemma (cf. (3.26)). These results are used to prove convergence of P1 FEM approximations under natural conditions on regularity of the meshes used in the constructions of Galerkin approximations.

The main result is the error identity (3.19). We deduced it by finding explicit forms of the terms in general error relations (3.1) and (3.4). The identity defines a natural error measure for the considered class of problems and forms a basis for fully computable and guaranteed tight bounds on the global errors (Table 6).

An advantage of the suggested approach is that it can be used for any conforming approximation (e.g., P1 or P2 finite element, IGA, or spectral approximations) and that the estimates do not contain local (mesh dependent) constants or unknown global constants.

As we have confirmed by our theoretical findings and numerical experiments, flux equilibration is an important paradigm to obtain an accurate error indicator (cf. Figures 10 and 11) as well as to guarantee that the last term in (3.11) does not dominate the majorant.

References

[1] D. Braess and J. Schöberl, Equilibrated residual error estimator for Maxwell’s equations, RICAM report 2006-19, Austrian Academy of Sciences, 2006. Search in Google Scholar

[2] H. Brézis, Functional Analysis, Sobolev Spaces and Partial Differential Equations, Springer, New York, 2011. 10.1007/978-0-387-70914-7Search in Google Scholar

[3] H. Brézis and F. E. Browder, Sur une propriété des espaces de Sobolev, C. R. Acad. Sci. Paris Sér. A-B 287 (1978), no. 3, A113–A115. Search in Google Scholar

[4] C. Carstensen, M. Feischl, M. Page and D. Praetorius, Axioms of adaptivity, Comput. Math. Appl. 67 (2014), no. 6, 1195–1253. 10.1016/j.camwa.2013.12.003Search in Google Scholar

[5] L. Chen, M. J. Holst and J. Xu, The finite element approximation of the nonlinear Poisson–Boltzmann equation, SIAM J. Numer. Anal. 45 (2007), no. 6, 2298–2320. 10.1137/060675514Search in Google Scholar

[6] H. Childs, E. Brugger, B. Whitlock, J. Meredith, S. Ahern, D. Pugmire, K. Biagas, M. Miller, C. Harrison, G. H. Weber, H. Krishnan, T. Fogal, A. Sanderson, C. Garth, E. Wes Bethel, D. Camp, O. Rübel, M. Durant, J. M. Favre and P. Navrátil, VisIt: An end-user tool for visualizing and analyzing very large data, High Performance Visualization–Enabling Extreme-Scale Scientific Insight, CRC Press, Boca Raton (2012), 357–372. Search in Google Scholar

[7] C. Dobrzynski, MMG3D: User guide, Technical Report RT-0422, INRIA, 2012. Search in Google Scholar

[8] I. Ekeland and R. Temam, Convex Analysis and Variational Problems, North-Holland, Amsterdam, 1976. Search in Google Scholar

[9] M. Feischl, D. Praetorius and K. G. van der Zee, An abstract analysis of optimal goal-oriented adaptivity, SIAM J. Numer. Anal. 54 (2016), no. 3, 1423–1448. 10.1137/15M1021982Search in Google Scholar

[10] F. Fogolari, A. Brigo and H. Molinari, The Poisson–Boltzmann equation for biomolecular electrostatics: A tool for structural biology, J. Mol. Recognit. 15 (2002), 377–392. 10.1002/jmr.577Search in Google Scholar

[11] F. Fogolari, P. Zuccato, G. Esposito and P. Viglino, Biomolecular electrostatics with the linearized Poisson–Boltzmann equation, Biophys. J. 76 (1999), 1–16. 10.1016/S0006-3495(99)77173-0Search in Google Scholar

[12] D. Gilbarg and N. S. Trudinger, Elliptic Partial Differential Equations of Second Order, Classics Math., Springer, Berlin, 2001. 10.1007/978-3-642-61798-0Search in Google Scholar

[13] F. Hecht, New development in freefem++, J. Numer. Math. 20 (2012), no. 3–4, 251–265. 10.1515/jnum-2012-0013Search in Google Scholar

[14] M. Holst, J. A. McCammon, Z. Yu, Y. C. Zhou and Y. Zhu, Adaptive finite element modeling techniques for the Poisson–Boltzmann equation, Commun. Comput. Phys. 11 (2012), no. 1, 179–214. 10.4208/cicp.081009.130611aSearch in Google Scholar PubMed PubMed Central

[15] B. Kawohl and M. Lucia, Best constants in some exponential Sobolev inequalities, Indiana Univ. Math. J. 57 (2008), no. 4, 1907–1927. 10.1512/iumj.2008.57.3307Search in Google Scholar

[16] D. Kinderlehrer and G. Stampacchia, An Introduction to Variational Inequalities and Their Applications, Class. Appl. Math. 31, Society for Industrial and Applied Mathematics (SIAM), Philadelphia, 2000. 10.1137/1.9780898719451Search in Google Scholar

[17] P. Neittaanmäki and S. Repin, Reliable Methods for Computer Simulation. Error Control and a Posteriori Estimates, Stud. Math. Appl. 33, Elsevier Science, Amsterdam, 2004. Search in Google Scholar

[18] H. Oberoi and N. Allewell, Multigrid solution of the nonlinear Poisson–Boltzmann equation and calculation of titration curves, Biophys. J. 65 (1993), 48–55. 10.1016/S0006-3495(93)81032-4Search in Google Scholar

[19] S. I. Repin, A posteriori error estimation for variational problems with uniformly convex functionals, Math. Comp. 69 (2000), no. 230, 481–500. 10.1090/S0025-5718-99-01190-4Search in Google Scholar

[20] S. I. Repin, On measures of errors for nonlinear variational problems, Russian J. Numer. Anal. Math. Modelling 27 (2012), no. 6, 577–584. 10.1515/rnam-2012-0033Search in Google Scholar

[21] S. Repin and J. Valdman, Error identities for variational problems with obstacles, ZAMM Z. Angew. Math. Mech. 98 (2018), no. 4, 635–658. 10.1002/zamm.201700105Search in Google Scholar

[22] I. Sakalli, J. Schöberl and E. W. Knapp, mfes: A robust molecular finite element solver for electrostatic energy computations, J. Chem. Theory Comput. 10 (2014), 5095–5112. 10.1021/ct5005092Search in Google Scholar PubMed

[23] K. Sharp and B. Honig, Calculating total electrostatic energies with the nonlinear Poisson–Boltzmann equation, J. Phys. Chem 94 (1990), 7684–7692. 10.1021/j100382a068Search in Google Scholar

[24] R. E. Showalter, Hilbert Space Methods for Partial Differential Equations, Pitman, London, 1977. Search in Google Scholar

[25] H. Si, TetGen, a Delaunay-based quality tetrahedral mesh generator, ACM Trans. Math. Software 41 (2015), no. 2, Article ID 11. 10.1145/2629697Search in Google Scholar

[26] G. Stampacchia, Le problème de Dirichlet pour les équations elliptiques du second ordre à coefficients discontinus, Ann. Inst. Fourier (Grenoble) 15 (1965), no. 1, 189–258. 10.5802/aif.204Search in Google Scholar

[27] N. S. Trudinger, On imbeddings into Orlicz spaces and some applications, J. Math. Mech. 17 (1967), 473–483. 10.1512/iumj.1968.17.17028Search in Google Scholar

[28] A collection of molecular surface meshes, https://www.rocq.inria.fr/gamma/gamma/download/affichage.php?dir=MOLECULE&name=water_mol&last_page=6, Accessed: 2017-08-18. Search in Google Scholar

Received: 2018-09-27
Revised: 2019-04-07
Accepted: 2019-05-03
Published Online: 2019-06-13
Published in Print: 2020-04-01

© 2020 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 15.12.2024 from https://www.degruyter.com/document/doi/10.1515/cmam-2018-0252/html
Scroll to top button