Three-Field Modelling of Nonlinear Nonsmooth Boundary Value Problems and Stability of Differential Mixed Variational Inequalities
Abstract
The purpose of this paper is twofold. Firstly we consider nonlinear nonsmooth elliptic boundary value problems, and also related parabolic initial boundary value problems that model in a simplified way steady-state unilateral contact with Tresca friction in solid mechanics, respectively, stem from nonlinear transient heat conduction with unilateral boundary conditions. Here a recent duality approach, that augments the classical Babuška-Brezzi saddle point formulation for mixed variational problems to twofold saddle point formulations, is extended to the nonsmooth problems under consideration. This approach leads to variational inequalities of mixed form for three coupled fields as unknowns and to related differential mixed variational inequalities in the time-dependent case. Secondly we are concerned with the stability of the solution set of a general class of differential mixed variational inequalities. Here we present a novel upper set convergence result with respect to perturbations in the data, including perturbations of the associated nonlinear maps, the nonsmooth convex functionals, and the convex constraint set. We employ epiconvergence for the convergence of the functionals and Mosco convergence for set convergence. We impose weak convergence assumptions on the perturbed maps using the monotonicity method of Browder and Minty.
1. Introduction
The classical Babuška-Brezzi theory for mixed variational problems has been extended by Gatica [1, 2] to some classes of variational problems and nonlinear operator equations. This extension leads to three-field variational models that can be understood as dual-dual mixed variational models or as twofold saddle point formulations. Such augmented variational models are well adapted for multiphysics problems with different coupled unknown quantities and in particular for engineering problems, where speaking in terms of solid mechanics, strains and stresses are often of more interest then the displacements.
In a series of papers Gatica with coauthors applied his duality approach to the numerical treatment of various linear/nonlinear interior/exterior elliptic boundary value problems by the finite element method (FEM), the boundary element method (BEM), or by the coupling of these discretization methods. Here we refer to the paper of Gatica et al. [3] that presents a numerical analysis of nonlinear two-fold saddle point problems involving a nonlinear operator equation with a uniformly monotone operator.
This novel duality approach to nonlinear nonsmooth boundary value problems has to be distinguished from the standard duality approach which hinges on the Lagrange duality theory of convex analysis in calculus of variations (see [4] for a systematic study) and which is employed in the numerical FEM analysis of various unilateral boundary and obstacle problems as pioneered by Haslinger and Lovíšek [5, 6]; see also the monograph [7].
In this paper, we address a simplified scalar model of steady-state unilateral contact problems with Tresca friction and nonlinear transient heat conduction problems with unilateral boundary conditions. We extend the duality approach of Gatica to such problems. This approach leads to variational inequalities of mixed form for three coupled fields as unknowns and to related differential mixed variational inequalities (DMVI) in the time-dependent case.
Differential variational inequalities have recently been introduced and studied in depth by Pang and Stewart [8] in finite dimensions as a new modeling paradigm of variational analysis. In their seminal paper the authors have already shown that this new class of differential inclusions contains ordinary differential equations with possibly discontinuous right-hand sides, differential algebraic systems, dynamic complementarity systems, and evolutionary variational systems. More recently, some results of [8] have been extended to DMVI by Li et al. [9] in finite dimensions.
Furthermore in this paper, we are concerned with stability of the solution set to DMVI. In this connection, let us refer to [10], where a Lyapunov approach is developed for strong solutions of evolution variational inequalities and to [11], where first several sensitivity results are established for initial value problems of ordinary differential equations with nonsmooth right hand sides and then applied to treat differential variational inequalities. Related stability results for more general evolution inclusions by Papageorgiou [12, 13] and by Hu and Papageorgiou in the memoir [14] are not applicable here since these results are limited to finite dimensions, respectively, and need more stringent compactness assumptions.
Here we present a novel upper set convergence result for DMVI with respect to perturbations in the data. In particular, we admit perturbations of the nonlinear maps, of the non-smooth convex functionals, and the convex constraint set that describe the DMVI. We employ epiconvergence for the convergence of the functionals and Mosco convergence for set convergence. Since our analysis of the underlying mixed nonlinear variational inequality relies on the monotonicity method of Browder and Minty (see e.g., [15, 16]), we need to impose comparably weak convergence assumptions on the perturbations in the nonlinear maps. Thus we extend the stability results of [17] (without considering slow solutions here) and of [18] to this new more general class of mixed differential variational inequalities.
The outline of this paper is as follows. In Sections 2 and 3, we show how the duality approach of Gatica can be extended to a nonlinear transient heat problem with unilateral boundary conditions and to a scalar nonsmooth boundary value problem modelling the steady-state unilateral contact of an elastic body with Tresca friction. Thus we obtain variational inequalities of mixed form involving three unknown fields and related differential mixed variational inequalities (DMVI) in the time-dependent case. Then in Section 4, we turn to the stability analysis of a general class of DMVI. After a discussion of some preliminaries including Mosco set convergence and epiconvergence (Section 4.2), we establish our novel stability result in Section 4.3 based on the Browder-Minty monotonicity method. Section 5 gives an outlook to some open directions of research.
2. A Nonlinear Nonsmooth Boundary Value Problem from Heat Conduction
In this section we consider a nonlinear boundary value problem with Signorini boundary conditions that arises from nonlinear heat conduction [19] with semipermeable walls [20]. We first show how the steady-state problem can be variationally formulated as a variational inequality in mixed form. Then we turn to the transient problem and derive the associated differential mixed variational inequality.
3. A Simplified Scalar Nonsmooth Boundary Value Problem from Frictional Contact
In this section we treat a nonsmooth boundary value problem that can be considered as a simplified scalar model of a nonsmooth contact mechanics problem involving Tresca friction and a unilateral constraint of an elastic body with a rigid foundation. Instead of the vector Navier-Lamé pde system or a nonlinear extension of it to model nonlinear elastic material, we are concerned with a nonlinear Helmholtz-like pde. This pde will be complemented by nonclassical boundary conditions involving the non-smooth modulus function. We show how a three-field modelling transforms this non-smooth boundary value problem to a variational inequality of mixed type.
These implications reflect Tresca′s law of friction (given friction model) and the more general Coulomb′s law of friction [20]. Namely, with σ · ν denoting (the tangential component of) the traction in the general elasticity problem, the first implication means that when the body is not in contact with the obstacle, there is no tangential stress due to friction. If on the other hand, the body is in contact with the rigid obstacle, then—this is the meaning of the second implication—there arises a tangential stress proportional and opposite to the tangential displacement u.
Remark 1. The obtained convex functional j and the constraint |u | ≤ g are related by Fenchel duality of convex analysis, see for example [4]. Namely, introduce the convex, lower semicontinuous, proper function φ(y)∶ = g | y| with g > 0. Then the Fenchel dual function
Finally in virtue of Green′s formula (5), the proven claim provides the variational inequality of the second kind as follows:
4. Differential Mixed Variational Inequalities and Their Stability
Motivated by the non-smooth boundary value problems and their variational formulation in the previous sections, we deal in this section with a general class of differential mixed variational inequalities. As we will see that, with some changes of notation, all the concrete variational inequalities of mixed form of the previous sections can be subsumed in this class, when introducing some appropriate product spaces.
Since in our stability analysis, we permit perturbations in the non-smooth convex functionals and in the convex constraint set, we provide auxiliary results on epiconvergence and Mosco convergence. Using the monotonicity method of Browder and Minty, we can establish a general stability result under weak convergence assumptions.
4.1. The Setting of Differential Mixed Variational Inequalities Considered
In what follows we study stability of differential mixed variational inequalities formulated as DMVI and admit perturbations x0,n of x0 in the initial condition x(0) = x0, fn, gn of the maps f : X × V → X, g : X × V → V, Kn of the convex closed subset K ⊂ V, and ϕn of the convex, lower semicontinuous proper functional ϕ : V → ℝ ∪ +∞. Suppose that (xn, un) solves (DMVI)(fn, gn, Kn, ϕn; x0,n), and assume that (xn, un)→(x, u) with respect to an appropriate convergence for X-valued, respectively, V-valued functions on [0, T]. Then we seek conditions on fn → f, gn → g, Kn → K, ϕn → ϕ, and x0,n → x0 that guarantee that (x, u) solves the limit problem (DMVI)(f, g, K, ϕ; x0). Such a stability result can be understood as a result of upper set convergence for the solution set of (DMVI)(f, g, K, ϕ; x0).
4.2. Preliminaries; Mosco Convergence of Sets; Epiconvergence of Functions
Here the prefix σ and mean sequentially weak convergence in contrast to strong convergence denoted by the prefix s and by . Further, limsup , liminf respectively are in the sense of Kuratowski upper, lower limits respectively of sequences of sets (see [25] for more information on Mosco convergence). Here we note that for the nonempty set K the second inclusion provides gn ∈ Kn, such that for some g ∈ K. Clearly, , if and only if , this simple translation argument shows that there is no loss of generality to assume when needed that 0 ∈ Kn, K.
As a preliminary result we next show that Mosco convergence of convex closed sets Kn inherits to Mosco convergence of the polars and to Mosco convergence of the associated sets 𝒦n = L2(0, T; Kn), derived from Kn similar to (34).
Lemma 2. Let . Then (a) ; (b) in L2(0, T; V).
Proof. To show (a) we verify that
(1) .
Let ζ = σ − lim n→∞ζn with . Choose z ∈ K arbitrarily. Then by assumption, there exists (eventually for a subsequence) zn ∈ Kn with z = s − lim n→∞zn. By definition of polar, (ζn, zn) ≤ 1, for all n, hence in the limit (ζ, z) ≤ 1, for all z ∈ K which gives ζ ∈ K0.
(2) .
By [25, Proposition 3.23], , where
Now let ζ ∈ K0. Then s(K)(ζ) ≤ 1, and hence ζ ∈ dom s(K). For the previous sequences {ζn}, {sn : = s(Kn)(ζn)}, we obtain limsup n sn ≤ 1. If for a subsequence holds, then and the argument is complete. Otherwise, for almost all n we have sn > 1. But then , and ; the proof of part (a) terminates.
To show (b) we verify that
(1) σ − limsup n→∞L2(0, T; Kn) ⊂ L2(0, T; K).
Let w = σ − lim n→∞wn with wn ∈ L2(0, T; Kn). By the bipolar theorem (K00 = K, assume here without loss of generality that 0 ∈ K) it is enough to show that for all ζ ∈ K0, for a.a. t ∈ (0, T) there holds (w(t), ζ) ≤ 1.
Assume not. Then there exist with measure |A | > 0, such that on A. Define (with χA denoting the characteristic function of A). By part (a), there exist , such that ; hence for , we have s − lim nvn = v in L2(0, T; V). By construction,
(2) 𝒦 ⊂ s − liminf n→∞𝒦n.
It is enough to verify the claim for a dense subset of 𝒦; this follows from a diagonal sequence argument; see also [26, Lemma 2.6] for a similar reasoning. Here we use the well-known fact from Bochner-Lebesgue integration theory that the set 𝒮(0, T; V) of simple V-valued functions on (0, T) is dense in L2(0, T; V). This extends to density of 𝒮(0, T; V)∩𝒦, the K-valued simple functions on (0, T), in 𝒦. This can be seen by taking averages or mean value approximations; see [26] for approximations on a multidimensional integration domain instead of the interval (0, T).
Thus let w be a K-valued simple function on (0, T); that is, , where J is finite, zj ∈ K, and Aj ⊂ (0, T) are pairwise disjoint with measure |Aj | > 0 and ∪j∈JAj = (0, T). Since , there exist zj,n ∈ Kn, such that for all j ∈ J, zj,n → zj (n → ∞). Hence lies in 𝒦n, and w = s − lim n→∞wn in L2(0, T; V) follows.
Let us note that part (b) of the preceding lemma is of intrinsic interest for time-dependent variational inequalities. An analogous implication (b) (1). was already shown in [26] in the more general context of probability spaces instead of the interval (0, T), however for the restricted class of translated convex closed cones.
- (1)
,
- (2)
for all w with ϕ(w)<+∞ ∃ {wn} n∈ℕ,
()
As a further preliminary result we next show that epiconvergence of convex lsc functions ϕn inherits to epiconvergence of the associated functionals Φn, derived from ϕn similar to (33).
Lemma 3. Let the convex lsc proper functionals ϕn epiconverge to a convex lsc proper functional ϕ on V. Suppose that the functionals ϕn are equi-lower bounded in the sense that there exist co ∈ ℝ, w0 ∈ V, such that
Proof. To show (1) let v = σ − lim n→∞vn in L2(0, T; V). Hence v(t) = σ − lim n→∞vn(t) in V for almost all t ∈ (0, T). Now by equi-lower boundedness, conclude from the lemma of Fatou using the nonnegativity of the integrand that
To show (2) let u ∈ L2(0, T; V) with Φ(u) < ∞. Hence ρ(t): = ϕ(u(t)) is real valued for almost all t ∈ (0, T) and [ρ, u] ∈ L2(0, T; epi ϕ). Now in virtue of the previous Lemma 2, part (b), there exists a sequence {[ρn, un]} strongly convergent to [ρ, u] ∈ L2(0, T; ℝ × V). Hence un → u ∈ L2(0, T; V), un(t) → u(t) ∈ V for almost all t, and ρn → ρ ∈ L2(0, T). Therefore conclude from the lemma of Fatou using the nonpositivity of the integrand that
By combination of the previous lemmata we obtain the following auxiliary result.
Lemma 4. Let in V. Let the convex lsc proper functionals ϕn : Kn → ℝ epiconverge to ϕ : K → ℝ on V. Suppose that the functionals ϕn are equi-lower bounded in the sense that there exist co ∈ ℝ, w0 ∈ V, such that
The details of the proof are omitted.
As a further tool in our stability analysis we recall from [17] the following technical result.
Lemma 5. Let H be a separable Hilbert space, and let T > 0 be fixed. Then for any sequence {zn} n∈ℕ converging to some z in L1(0, T; H) there exists a subsequence , such that for some set N of zero measure, for all t ∈ [0, T]∖N.
4.3. The Stability Result
Before stating the result, some remarks are in order. In view of the existence theory of variational inequalities in infinite dimensional spaces (see, e.g., [15]) the best one can hope for is weak convergence of the perturbations un in the general case of nonunique solutions of the underlying variational inequalities in (DMVI). Weak convergence can, namely, be readily derived from a posteriori estimates. However, continuity of a nonlinear map (here g, f) with respect to weak convergence is a hard requirement. To circumvent these weak convergence difficulties we apply the monotonicity method of Browder and Minty. Then as we shall see below, a stability condition on the maps gn with respect to the basic Hilbert space norm suffices.
These weak convergence difficulties also affect fn. Therefore we have to impose a generally strong stability condition on the nonlinear maps fn. In the situation of linear operators this condition can be drastically simplified to a stability condition with respect to convergence in the operator norm, see [18, Theorem 4.1] in the case ϕ = ϕn = 0.
On the other hand, stronger assumptions on gn, like uniform monotonicity, imply that the solution sets Σ(Kn, ϕn, G(xn, ·)) are single valued. Uniform monotonicity with respect to n moreover entails that the sequence un strongly converges. Then the stability assumption for fn can be relaxed.
Since our stability assumptions pertain the given maps gn, g, not the derived maps Gn, G, we have a delicate interplay between the pointwise almost everywhere formulation and the integrated formulation of the variational inequality in the perturbed DMVI and in the limit DMVI.
- (H1)
Let in X and in V. Moreover, let in X. Then p = f(z, w).
- (H2)
All maps gn(z, ·) for any z ∈ X are monotone. If in X, and V respectively, then in V. g is hemicontinuous in the sense that for any z ∈ X; v, w ∈ V, the real-valued function r ∈ ℝ ↦ (g(z, v + rw), rw) is lower semicontinuous.
Now we can state the following stability result.
Theorem 6. Let (xn, un) solve (DMVI)(fn, gn, Kn, ϕn; x0,n). Suppose that fn, f, gn, and g, respectively, satisfy (H1), (H2), respectively. Let . Let the convex closed nonvoid sets Kn Mosco-converge to K in V, and let the convex lsc proper functions ϕn : Kn → ℝ epiconverge to ϕ : K → ℝ on V. Suppose that the functions ϕn are equi-lower bounded in the sense of (47). Assume that in 𝒳(0, T; X) and that un ∈ L2(0, T; V) converges weakly to u pointwise in V for a.a. t ∈ (0, T) with , for all a.a. t ∈ (0, T) for some m ∈ L2(0, T). Then (x, u) is a solution to (DMVI) (f, g, K, ϕ; x0).
Proof. The proof consists of three parts.
(1) Feasibility: u ∈ 𝒦, x(0) = x0.
First we observe that, for any w ∈ L2(0, T; V), in virtue of Lebesgue′s theorem of dominated convergence,
Since by continuous embedding in C[0, T; X], we conclude that .
(2) u solves the variational inequality in (DMVI)(f, g, K, ϕ; x0):
Fix an arbitrary w ∈ 𝒦. Then by Lemma 4, there exist wn ∈ 𝒦n, such that in L2(0, T; V) and limsup nΦn(wn) ≤ Φ(w). Moreover, by extracting eventually a subsequence, we have by Lemma 5 that also wn(t) strongly converges to w(t) for a.a. t ∈ (0, T). For any measureable set A ⊂ (0, T) we can define by on A, on (0, T)∖A. Hence and by construction it follows:
(3) (x, u) solves the limit (DMVI)(f, g, K, ϕ; x0).
By Lemma 5 applied to {xn} and , we can extract a subsequence, such that xn(t) → x(t) and strongly in X pointwise for all t ∈ (0, T)∖N0, where N0 is a null set. Fix t ∈ (0, T)∖N0. Then by assumption, for all n ∈ ℕ we have . Then in virtue of (H1), follows, and (x, u) solves the (DMVI)(f, g, K, ϕ; x0).
We refrain from deriving a stability result for stationary problems from Theorem 6. Instead we can refer to [27, Theorem 3] for a much stronger result.
5. Some Concluding Remarks
Let us first shortly remark on possible extensions and limitations of the previous stability result. The monotonicity method extends easily to set-valued operators. Also, monotonicity can be replaced by the more general, however more abstract notion of (order-) pseudomonotonicity.
Here motivated by the considered non-smooth boundary value problems we discussed differential variational inequalities in a Hilbert space framework. Let us point out that differential variational inequalities and their stability using Mosco convergence can be investigated in more general reflexive Banach spaces, but not beyond [28].
Finally let us give an outlook of some open directions of research. In this paper we confined the three-field modelling to a simplified scalar model of nonsmooth contact in continuum mechanics. An extension of the three-field duality approach to such non-smooth boundary value problems will involve the tensor fields of continuum mechanics. In particular, the basic Green’s formula (5) has then to be replaced by the more general “integration by parts” formula [29, (3.46)].
In this paper, we did not touch the issue of existence of solutions to DMVI. As shown by Pang and Stewart [8] and Li et al. respectively, [9], existence of solutions to DMVI can be obtained from the theory of multivalued differential equations [30] and the theory of the more general differential inclusions [31]. However, compactness assumptions as used in [8, 9] are too strong in infinite dimensions; moreover the feasible set K defined by the unilateral boundary conditions is a convex cone, but not compact. For an existence result for linear differential variational inequalities based on maximal monotone operator theory, we can refer to [18].
When it comes to numerical approximation, note that Galerkin approximation with respect to space discretization leads in the case of higher than piecewise linear approximation; to a nonconforming approximation, that is, the approximating convex set Kh is not a subset of the given convex K. In this situation Mosco convergence (and more refined Glowinski convergence) is instrumental to arrive at convergence of Galerkin approximation; see [7, 32]. Thus our stability result can be seen as an important step towards convergence of semidiscretization methods that provide finite dimensional differential variational inequalities. Clearly a full space time discretization needs an additional analysis (see, e.g., [33]) and is beyond the scope of the present paper.