1 Introduction

We are interested in Euler-Lagrange equations, existence and uniqueness of minimizers for some problems in the vectorial calculus of variations emanating from elasticity theory. These variational problems are related to an open problem in Partial Differential Equations that we describe as follows: let \(T>0\) and let \(\varOmega \) and \(\varLambda \) be two open subsets of \(\mathbb{R}^{d}\); suppose that \(\mathbf{u}_{0}\) is a diffeomorphism between \(\varOmega \) and \(\varLambda \); we seek \(\mathbf{u} : \varOmega \times (0,T)\longrightarrow \mathbb{R}^{d} \) such that \(\mathbf{u}(\cdot ,t) (\varOmega )=\varLambda \) for each \(t\) and

$$ \textstyle\begin{cases} \mathbf{u}_{t}=\operatorname{div}_{x} D_{\xi }L(\nabla \mathbf{u}) & \text{on }\varOmega \times (0,T),\\ \mathbf{u}(0, \cdot )=\mathbf{u}_{0} &\text{on }\varOmega , \end{cases} $$
(1.1)

in the sense of distributions. In (1.1), we assume that the map \(\mathbb{R} ^{d\times d}\ni \xi \mapsto L(\xi )\) is quasiconvex. We refer the reader to [2], [7], [5], [11], and [12] for further details on these gradient flows. Understanding variational problems associated to the time-discretization of (1.1) is arguably an important step toward the construction of a solution. In that regard, several partial results are available in the literature (see for instance [7] and [5]).

In [2], the authors have focused on a class of Lagrangians that arises in elastic materials. More precisely, they have considered polyconvex Lagrangians of the form \(\xi \mapsto L(\xi )=f(\xi )+H( \det \xi )\). Here \(f\) is a \(C^{1}(\mathbb{R}^{d} )\) strictly convex function with \(p\)-th order growth, and the map \(H\) is a \(C^{1}(0, \infty )\) convex function that satisfies

$$ \lim_{t\rightarrow 0^{+}} H(t)= \lim_{t\rightarrow \infty }\frac{H(t)}{t}=+\infty . $$
(1.2)

As a result, a variational problem emerges from the time discretization and has a relaxation that takes the general form:

$$ \min \biggl\{ \int _{\varOmega } \bigl( f(\nabla u)+H(\beta )-F\cdot u \bigr)\,dx ;\; (u,\beta )\in \mathscr{U} \biggr\} $$
(1.3)

where \(F\in L^{1}(\varOmega ,\mathbb{R} ^{d}) \) and

$$\begin{aligned} \mathscr{U}= \biggl\{ &(u,\beta ): u\in W^{1,p}(\varOmega ,\bar{\varLambda }),\; \beta :\varOmega \to [0, \infty ); \\ &\int _{\varOmega }l(u)\beta \,dx= \int _{\varLambda }l(y)\,dy;\forall l\in C_{c}\bigl(\mathbb{R}^{d} \bigr) \biggr\} . \end{aligned}$$
(1.4)

Although the existence of minimizers in (1.3) follows from the direct methods in the calculus of variations, the uniqueness is a rather challenging problem. Indeed, because of (1.2) and the non-convexity of the integrand, standard techniques in calculus of variations do not apply.

To bypass these difficulties, the authors of [2] have introduced a pseudo-projected gradient operator \(\mathscr{U}_{\mathcal{S}}\ni u\mapsto \nabla _{\mathcal{S}}u\) defined as follows: for a given \(u\in \mathscr{U}_{\mathcal{S}}\), the map \(\nabla _{ \mathcal{S}}u\) is the unique minimizer of

$$ \int _{\varOmega }f(G)\,dx $$

over

$$ \mathcal{G}_{\mathcal{S}}(u):= \biggl\{ G\in L^{p}\bigl(\varOmega ,\mathbb{R} ^{d\times d}\bigr) : \int _{\varOmega }u\operatorname{div}\varphi \; dx =- \int _{\varOmega } \langle {G},{\varphi } \rangle \,dx\; \forall \varphi \in \mathcal{S} \biggr\} . $$

Here, \(\mathcal{S}\) is a finite-dimensional subspace of \(W_{0}^{1,q}( \varOmega ,\mathbb{R} ^{d\times d})\), \(q\) is the conjugate of \(p\), \(\mathscr{U}_{\mathcal{S}}\) is the set of all \(u:\varOmega \to \bar{ \varLambda }\) measurable such that there exists a \(c=c(u,\varOmega ,\varLambda ) >0\) satisfying:

$$ \biggl\vert \int _{\varOmega }u\cdot \operatorname{div}\, \varphi \;dx \biggr\vert \leq c \|\varphi \| _{ L^{q}( {\varOmega , \mathbb{R}^{d\times d}})}, \; \quad \forall \varphi \in \mathcal{S}. $$
(1.5)

We point out that the pseudo-projected gradient operator depends also on \(f\), though the dependence is not exhibited in its notation. As a first step to approaching (1.3), they have considered the following perturbed problem:

$$ \inf \biggl\{ \int _{\varOmega } \bigl( f(\nabla _{S} u)+H(\beta )-F \cdot u \bigr)\,dx ;\; (u,\beta )\in \mathscr{U} \biggr\} . $$
(1.6)

The choice of problem (1.6) is justified by the construction of a family of finite dimensional subspaces \(\lbrace \mathcal{S} _{n} \rbrace _{n} \) dense in \(W_{0}^{1,q}(\varOmega ,\mathbb{R} ^{d\times d}) \) such that for \(u\in W^{1,p}(\varOmega ,\mathbb{R} ^{d})\), one has

$$ \lim_{{n} \rightarrow \infty } \int _{\varOmega }f(\nabla _{\mathcal{S} _{n}} u)\;dx= \int _{\varOmega }f(\nabla u )\;dx. $$
(1.7)

We note that a \(L^{p}(\varOmega ,\; \mathbb{R}^{d})\)-bounded subset of \(\mathscr{U}_{\mathcal{S}}\) whose image by the operator \(\nabla _{\mathcal{S}}\) is bounded in \(L^{p}(\varOmega ,\; \mathbb{R}^{d \times d})\) is not in general strongly pre-compact with respect to the \(L^{p}(\varOmega ,\; \mathbb{R}^{d})\) topology. As a result, compactness of level subsets of the functional in (1.6) cannot be guaranteed. Nevertheless, the authors of [2] have successfully shown existence and, more importantly, uniqueness in (1.6) under the assumption that \(F\) is non-degenerate (see definition below). This condition of non-degeneracy for uniqueness is crucial in a similar problem, the so-called Brenier polar factorization, and more generally, in optimal transport problems. Confer [1], [3], [9], [8], [10] and [15].

In this paper, we investigate the respective roles played by the strict convexity of \(f\), the convexity and smoothness of \(H\), and the non-degeneracy of \(F\) in problem (1.6). More precisely, we impose less stringent conditions so that either the map \(F\) is allowed to be degenerate or \(f\) is allowed to be merely convex or \(H\) is neither smooth nor strictly-convex. These considerations are not just technicalities. Indeed we note that a prominent case of mere convexity, \(f(\xi )=|\xi | \), is typical for the study of minimal surfaces as well as for the study of functionals involving the total variation (see for instance [4]). Furthermore, we observe that cases where \(H\) is taken to be the characteristic function of a singleton of ℝ arise in the study of incompressible deformations in Elasticity theory (see for instance [12] and [15]). Finally, the non-degeneracy condition tests the extent to which one can hope for uniqueness in the variational problem we considered. To deal with these weaker assumptions, we introduce a family of operators \(\lbrace V_{\mathcal{S}}^{f} : \mathcal{S}\subset W_{0}^{1,q}(\varOmega ,\mathbb{R} ^{d}), \;f \text{ convex} \rbrace \) defined by

$$ W^{1,p}\bigl(\varOmega ,\mathbb{R} ^{d} \bigr) \ni u\mapsto V^{f}_{\mathcal{S}}[u]:= \sup _{\varphi \in \mathcal{S}} \int _{\varOmega } \bigl(- u \operatorname{div}\varphi -f^{*}(\varphi ) \bigr)\;dx. $$
(1.8)

We note that the operator \(V_{\mathcal{S}}^{f}\) is actually well defined on the set of measurable functions \(u\) defined from \(\varOmega \) to \(\bar{\varLambda }\) when the set \(\mathcal{S}\) is a finite dimensional nonempty set and the function \(f\) satisfies appropriate growth conditions. As a family, these operators extend the pseudo-projected gradient operators and the distributional gradient. Indeed, \(V_{\mathcal{S}}^{f}[u]=\int _{\varOmega}f(\nabla _{\mathcal{S}}u)\;dx\) if \(\mathcal{S}\) is a finite dimensional subspace of \(W_{0}^{1,q}(\varOmega ,\mathbb{R} ^{d\times d})\) and \(u\in \mathscr{U}_{\mathcal{S}}\) and furthermore \(V_{\mathcal{S}}^{f}[u]=\int _{\varOmega }f(\nabla u)\;dx\) if \(\mathcal{S}=W_{0}^{1,q}(\varOmega ,\mathbb{R} ^{d\times d})\) and \(u\in W^{1,p}(\varOmega ,\mathbb{R} ^{d})\). These extensions are only valid under appropriate conditions on \(f\). It is worth pointing out that if \(f(\xi )=|\xi |\) and \(\mathcal{S}=W_{0}^{1,q}(\varOmega,\mathbb{R} ^{d \times d})\) then \(V_{\mathcal{S}}^{f}(u)\) is nothing but the total variation of \(u\) on the set \(\varOmega \). We show that for a collection of sets \(\lbrace \mathcal{S}_{n} \rbrace _{n=1}^{\infty } \) of \(W_{0}^{1,q}(\varOmega ,\mathbb{R} ^{d})\) satisfying Hypothesis (H1) or Hypothesis (H2) (see Sect. 2), we have a convergence result in the same spirit as (1.7):

$$ \lim_{n \rightarrow \infty }V^{f}_{\mathcal{S}_{n}}[u]=V^{f}_{W _{0}^{1,q}(\varOmega ,\mathbb{R} ^{d})}[u] \biggl( = \int _{\varOmega }f( \nabla u )\;dx \biggr) $$
(1.9)

for any \(u\in W^{1,p}(\varOmega ,\mathbb{R} ^{d\times d})\) and appropriate conditions on \(f\). We thus proceed to study a more general problem:

$$ \inf_{(u,\beta )\in \mathscr{U}^{*}_{\mathcal{S}}} \biggl\{ V_{\mathcal{S}}^{f}[u]+ \int _{\varOmega }\left(H(\beta )-F\cdot u\right)\; dx \biggr\} $$
(1.10)

where \(\mathcal{S}\) is an element of a collection of sets satisfying Hypothesis (H1) or Hypothesis (H2), and

$$\begin{aligned} \mathscr{U}_{\mathcal{S}}^{*}= \biggl\{ &(u,\beta ): u \in \mathscr{U} _{\mathcal{S}};\; \beta :\varOmega \to [0,\infty ); \\ &\int _{\varOmega }l\bigl(u(x)\bigr)\beta (x)\;dx= \int _{\varOmega }l(y)\;dy \; \forall l \in C_{c}\bigl( \mathbb{R} ^{d}\bigr) \biggr\} . \end{aligned}$$
(1.11)

Sublevel sets of the integrand in (1.10) are not compact. Nor is \(f\) necessarily strictly convex. However, we show existence and uniqueness in problem (1.10). In fact, this result holds for \(F\) non-degenerate as well as for a class of degenerate \(F\) provided that the set \(\mathcal{S}\) is chosen accordingly (see Corollaries 3.6 and 3.7). Unlike optimal transport theory, this analysis suggests that the non-degeneracy condition is not essential for a uniqueness result in (1.3).

Existence and uniqueness results for problem (1.10) are established thanks to the discovery of suitable dual problems. Indeed, call \(\mathcal{C}\) the set of all functions \((k,l)\) with \(k,\;l:\mathbb{R}^{d} \rightarrow \mathbb{R} \cup \{\infty \}\) Borel measurable, finite at least at one point, and satisfying the relation \(l\equiv \infty \) on \(\mathbb{R}^{d} \setminus \bar{\varLambda }\) and such that

$$ k(v) +tl( u) +H(t)\geq u\cdot v\quad \forall u,v\in \mathbb{R}^{d} , t>0. $$

Let \(\mathscr{A}\) be the set of \((k,l,\varphi )\) such that \((k,l) \in \mathcal{C}\) and \(\varphi \in \mathcal{S}\). Define the following functional over the set \(\mathscr{A}\):

$$ J(k,l,\varphi ):= \int _{\varOmega }k(F+\operatorname{div} \varphi )\;dx+ \int _{\varLambda }l\;dy+ \int _{\varOmega }f^{*}(\varphi )\;dx. $$

Next, assume that the map \(F \) and the set \(\mathcal{S}\) are such that for all \(\varphi \in \mathcal{S}\),

$$ F+\operatorname{div} \varphi \text{ is non-degenerate.} $$
(1.12)

Then \(-J\) admits a maximizer \((k_{0},l_{0},\varphi _{0}) \) with \(k_{0} \) convex and \(\operatorname{diam}(\varLambda )\)-Lipschitz. As a consequence, problem (1.10) admits a unique minimizer \((u_{0},\beta _{0}) \) and \(u_{0} \) satisfies

$$ \textstyle\begin{cases} u_{0}= \nabla k_{0}(F+\operatorname{div} \varphi _{0})\\ \varphi _{0}\in \varPhi _{\mathcal{S}}(u_{0}). \end{cases} $$
(1.13)

Here, we have denoted by \(\varPhi _{\mathcal{S}}(u_{0})\), the non-empty set of maximizers of problem (1.8) (see Proposition 2.8). In order to obtain condition (1.12), we consider two distinct situations.

First, we assume that \(F \) has a countable range, thus degenerate. If \(\mathcal{S}\) is an element of a collection of sets satisfying hypothesis (H2) then it holds that \(F+\operatorname{div} \varphi \) is non-degenerate.

Second, we assume \(F \) non-degenerate and \(\mathcal{S}\) is a finite dimensional vector space, as in [2]. It holds again that \(F+\operatorname{div} \varphi \) is non-degenerate. However, unlike the hypotheses in [2], we have allowed the map \(f\) to be as singular as the map \(\mathbb{R} ^{d\times d}\ni \xi \mapsto |\xi |\).

We have also studied (1.10) when \(H\) is replaced by \(H_{0}:(0,\infty )\to \mathbb{R} \cup \{\infty \}\) defined by \(H_{0}(1)=0\) and \(H_{0}(t)=\infty \) if \(t\neq 1\). This case corresponds to the case of measure preserving maps. Note that \(H_{0}\) is not even continuous. However, it may be obtained as a limit of functions \(H_{n}\) which are \(C^{1}(0,\infty )\) convex functions and satisfy (1.2). We show that for such singular \(H_{0}\), the corresponding problem

$$ \inf_{u\in \mathscr{U}^{1}_{\mathcal{S}}} \biggl\{ V_{\mathcal{S}}^{f}[u]- \int _{\varOmega }F\cdot u \;dx \biggr\} $$
(1.14)

with

$$ \mathscr{U}_{\mathcal{S}}^{1}= \biggl\{ u\in \mathscr{U}_{\mathcal{S}}: \int_{\varOmega }l\bigl(u(x)\bigr)\;dx= \int _{\varOmega }l(y)\;dy \; \forall l \in C _{c}\bigl( \mathbb{R} ^{d}\bigr) \biggr\} $$
(1.15)

admits a unique minimizer. (See Theorem 4.3.)

To obtain existence and uniqueness results in problem (1.14), we exploit a dual formulation and maximize \(-J\) over the set that consists of \((k,l,\varphi )\) such that \(\varphi \in \mathcal{S}\) and \(k,\;l:\mathbb{R}^{d} \rightarrow \mathbb{R} \cup \{\infty \}\) are Borel measurable, finite at least at one point, and satisfy the relations \(l\equiv \infty \) on \(\mathbb{R} ^{d} \setminus \bar{\varLambda }\) and

$$ k(v) +l( u) \geq u\cdot v\quad \forall u,v\in \mathbb{R}^{d}. $$

One shows that \(-J\) admits a maximizer \((k_{0},l_{0},\varphi _{0}) \) with \(k_{0} \) convex and Lipschitz and the unique minimizer of problem (1.14) is \(u_{0} \) given by

$$ u_{0}=\nabla k_{0}(F+\operatorname{div} \varphi _{0}). $$

Finally, we show convergence of a sequence of problems of the form (1.10) to (1.14). More precisely, we show that the minimizer of problem (1.14) may be obtained as limit of minimizers of problems of the form (1.10) provided that the dual problems admit regular enough maximizers. In fact, suppose the map \(F \) and the set \(\mathcal{S}\) are such that for all \(\varphi \in \mathcal{S}\), the map \(F+\operatorname{div} \varphi \) is non-degenerate. For \((u,\beta ) \in \mathscr{U}_{\mathcal{S}}\), define

$$ I_{n}(u,\beta )=V_{\mathcal{S}}^{f}[u]+ \int _{\varOmega }\left(H_{n}(\beta )-u \cdot F\right)\;dx $$

and set

$$ I_{0}(u )=V_{\mathcal{S}}^{f}[u]- \int _{\varOmega }u\cdot F\;dx. $$

Thanks to Theorem 3.5, the problem

$$ \inf_{(u,\beta )\in \mathscr{U}_{\mathcal{S}}^{*} }I_{n}(u,\beta ) $$
(1.16)

admits a unique minimizer that we denote \((u_{n},\beta _{n})\) with \(u_{n}=\nabla k_{n}(F+\operatorname{div}\varphi _{n})\) for some \(k_{n}:\mathbb{R} ^{d}\to \mathbb{R} \) convex and \(\varphi _{n}\in \mathcal{S}\). Denote \(u_{0}\) the unique minimizer of (1.14). If for all \(n\in \mathbb{N} ^{*}\) the map \(k_{n}\) is differentiable then the sequence \(\{ u_{n} \}_{n\in \mathbb{N} ^{*}}\) converges almost everywhere to \(u_{0}\) and in addition, the minima \(\{ I_{n}(u_{n},\beta _{n})\}_{n\in \mathbb{N} ^{*}}\) converge to \(I_{0}(u_{0})\) (cf. Theorem 4.7).

2 Preliminaries

2.1 Notation and Definitions

  • Throughout this manuscript, \(\varOmega \) and \(\varLambda \subset \mathbb{R}^{d}\) are two bounded open convex sets; \(r^{*}>1\) is such that \(B(0,1/r^{*})\subset \varLambda \subset B(0,r^{*}/2)\); \(p\in (1,\infty )\) and \(q\) is its conjugate, that is, \(p^{-1}+q^{-1}=1\).

  • Given \(A\subset \mathbb{R} ^{d}\), the indicator function of \(A\) is defined as

    $$ \chi _{A}(x)= \textstyle\begin{cases} 0&\text{if } x\in A,\\ \infty&\text{otherwise.} \end{cases} $$
  • For any subset \(\mathcal{S}\) of \(W_{0}^{1, q}( {\varOmega ,\mathbb{R} ^{d\times d}})\), we denote by \(\operatorname{span}(\mathcal{S})\) the linear subspace of \(W_{0}^{1, q}( {\varOmega , \mathbb{R} ^{d\times d}})\) generated by \(\mathcal{S}\).

  • We denote by \(f^{*}\) the Legendre transform of a map \(f:\mathbb{R} ^{d\times d}\longrightarrow \mathbb{R} \) so that

    $$ f^{*}\bigl(\xi ^{*}\bigr)=\sup_{\xi \in \mathbb{R} ^{d\times d}} \bigl\{ \xi \cdot \xi ^{*}-f(\xi ) \bigr\} . $$
  • If \(h : \mathbb{R} ^{d}\longrightarrow \mathbb{R} \cup \{\infty \}\) is convex then the subdifferential \(\partial h (x)\) of \(h\) at \(x\in \operatorname{Dom} (h)\) is closed and convex. If \(\partial h (x)\) is non-empty we denote by \(\operatorname{grad}[h](x)\) the element of \(\partial h (x)\) with minimum norm:

    $$ \bigl|\operatorname{grad}[h](x)\bigr|= \min \bigl\lbrace |y|: y\in \partial h (x) \bigr\rbrace ; \quad x\in \operatorname{Dom}(h). $$
  • Let \(\mathcal{S}\subset W_{0}^{1, q}( {\varOmega , \mathbb{R} ^{d\times d}})\). We denote by \(\mathscr{S}_{f}\) the set

    $$ \mathscr{S}_{f} := \biggl\{ \varphi \in \mathcal{S}: \int _{\varOmega }f ^{*}(\varphi ) \text{ is finite} \biggr\} . $$
    (2.1)
  • Let \(F :\mathbb{R}^{d} \longrightarrow \mathbb{R} ^{d} \) be measurable. We say that \(F\) is non-degenerate if for any \(N\subset \mathbb{R}^{d} \) such that \(\mathcal{L}^{d}(N) =0 \) we have \(\mathcal{L}^{d}(F^{-1}(N)) =0\).

2.2 Assumptions

(A0) :

We additionally assume that there exists a strictly convex function that is \(C^{1}(\bar{\varOmega }) \) and vanishes on the boundary of \(\varOmega \).

(A1) :

The set \(\mathcal{S}\) is a subset of \(W_{0}^{1, q}( {\varOmega ,\mathbb{R} ^{d\times d}})\). In addition, the map \(f:\mathbb{R} ^{d\times d}\to \mathbb{R} \) is convex and satisfies the following three properties:

  1. (i)

    There exist \(a,b,c >0\) such that for all \(\xi \in \mathbb{R} ^{d\times d}\),

    $$ c\frac{|\xi |^{p}}{p}+b\geq f(\xi )\geq a|\xi |-b $$
    (2.2)

    and for all \(\xi ^{*}\in \partial f(\xi ) \),

    $$ \bigl|\xi ^{*}\bigr|^{q}\leq c|\xi |^{p} + b. $$
    (2.3)
  2. (ii)

    The set \(\mathscr{S}_{f}\) is non-empty.

  3. (iii)

    One of the following two conditions holds:

    1. (a)

      The map \(f\) is such that \(\partial f^{*}(x^{*})\) is non-empty and \(\operatorname{grad}[f^{*}](x^{*})=0\) for each \(x^{*}\in \operatorname{Dom} f^{*}\).

    2. (b)

      The map \(f\) is strictly convex and there exist \(\bar{a}, \bar{b}>0\) such that for all \(\xi ^{*}\in \mathbb{R}^{d\times d}\), one has

      $$ f^{*}\bigl(\xi ^{*}\bigr)\leq \bar{a}+\bar{b}\bigl|\xi ^{*}\bigr|^{q} \quad \mbox{and} \quad \bigl| \nabla f^{*}\bigl(\xi ^{*}\bigr)\bigr|\leq \bar{a} +\bar{b}\bigl| \xi ^{*}\bigr|^{q-1}. $$
      (2.4)
(A2) :

The map \(H\) is \(C^{1}(0,\infty )\), strictly convex, and such that

$$ \lim_{t\rightarrow 0^{+}} H(t)= \lim_{t\rightarrow \infty } \frac{H(t)}{t}=+\infty . $$
(A3) :

The function \(F\) is measurable and belongs to \(L^{1}(\varOmega )\).

Let \(\mathcal{S}\) be a subset of \(W_{0}^{1,q }({\varOmega ,\mathbb{R} ^{d\times d}})\). We say that \(F\) satisfies the condition \(\textbf{(ND)}_{\mathcal{S}}\) if

$$ \operatorname{div}(\varphi ) + F \text{ is non-degenerate} $$

for all \(\varphi \in \mathcal{S}\).

Remark 2.1

  1. (i)

    As\(f\)satisfies (2.2), we have

    $$ -b+c^{p}\frac{|\xi ^{*}|^{q}}{q}\leq f^{*}\bigl(\xi ^{*}\bigr)\leq \chi _{\bar{B}(0,a)} \bigl( \xi ^{*}\bigr)+b $$
    (2.5)

    for all\(\xi ^{*}\in \mathbb{R} ^{d\times d}\).

  2. (ii)

    If\(f\)satisfies case (b) in (iii) of Assumption\(\mathbf{(A1)}\), then\(f^{*}\)is continuously differentiable. In that case, \(\operatorname{grad}[f^{*}]=\nabla f^{*}\).

  3. (iii)

    If\(f\)satisfies case (a) of Assumption\(\mathbf{(A1)}\)(iii) then\(0\in \partial f^{*}(x^{*}) \)for every element\(x^{*} \in \operatorname{Dom}(f^{*} )\). Consequently, the map\(f^{*} \)is constant on\(\operatorname{Dom}(f^{*} )\)and the following equalities are satisfied for all\(x^{*} \)and\(y^{*} \)in\(\operatorname{Dom}(f^{*} ) \):

    $$ f^{*}\bigl(x^{*}\bigr)-f^{*} \bigl(y^{*}\bigr)=\operatorname{grad}\bigl[f^{*}\bigr] \bigl(x^{*}\bigr)= \operatorname{grad}\bigl[f^{*}\bigr] \bigl(y^{*}\bigr)=0. $$
    (2.6)
  4. (iv)

    Assumption\(\mathbf{(A0)}\)is satisfied by\(\varOmega =B(0,1) \subset \mathbb{R}^{d}\)with the strictly convex function being the map\(\mathbb{R}^{d}\ni x\mapsto |x|^{2}-1 \).

  5. (v)

    The map\(f:\mathbb{R} ^{d\times d}\to \mathbb{R} \)defined by\(f(\xi )=|\xi |\)satisfies case (a) in (iii) of Assumption\(\mathbf{(A1)}\). The map\(f:\mathbb{R} ^{d\times d}\to \mathbb{R} \)defined by\(f(\xi )=|\xi |^{p}\)satisfies case (b) in (iii) of Assumption\(\mathbf{(A1)}\).

The following lemma summarizes some elementary properties of \(H \). We refer the reader to Remark 2.1 in [2].

Lemma 2.2

Assume\(\mathbf{(A2)}\)holds. Then,

  1. (i)

    The map\(H':(0,\infty )\to \mathbb{R} \)is a strictly increasing bijection.

  2. (ii)

    The Legendre transform\(H^{*}\)of\(H\)is a strictly increasing bijection fromto ℝ.

  3. (iii)

    Let\(g:\mathbb{R} \to \bar{\mathbb{R} } \)be defined by\(g(s)=\alpha s-\beta H^{*}(s)\), with\(\alpha ,\beta >0\). Then

    $$ \lim_{s\to -\infty }g(s)=\lim_{s\to \infty }g(s)=-\infty . $$

Define \(H_{0}\) by

$$ H_{0}(t)= \textstyle\begin{cases} 0& t=1\\ \infty& t\neq 1 \end{cases} $$
(2.7)

and, for \(n\geq 1\),

$$ H_{n}(t)= H(t)- H(1)+ n(t-1)^{2}. $$
(2.8)

The following lemma is straightforward.

Lemma 2.3

Assume\(\mathbf{(A2)}\)holds. Then,

  1. (i)

    There exists\(\bar{H}\in \mathbb{R} \)such that

    $$ \bar{H}= \min_{t\in [0,\infty )} H(t). $$
  2. (ii)

    The collection\(\lbrace H_{n} \rbrace _{n=1} ^{\infty } \)is a non-decreasing sequence of functions that converges pointwise to\(H_{0}\). In addition, for all\(n\in \mathbb{N} ^{*}\), the map\(H_{n}\)is a\(C^{1}(0,\infty )\)strictly convex function that satisfies

    $$ \lim_{t\rightarrow 0^{+}} H_{n}(t)= \lim_{t\rightarrow \infty } \frac{H _{n}(t)}{t}=+\infty . $$
  3. (iii)

    Let\(t>0\). If\(\lbrace H_{n}(t) \rbrace _{n=1} ^{\infty } \)is uniformly bounded above by a constant\(c_{0}\)then

    $$ n(t-1)^{2} \leq c_{0} +H(1)-\bar{H} $$

    and\(t=1\).

2.3 Hypothesis on the Underlying Sets of Pseudo-Gradients

We recall that in [2], the construction of \(\nabla_{\mathcal{S}^{\tau }} u\) has relied on hypothesis on the underlying sets \(\mathcal{S}^{\tau } \) that we summarize in Hypothesis (H1) below.

Hypothesis (H1).

A collection \(\{ \mathfrak{A}_{n }\}^{\infty }_{ n=1} \) of subsets of \(W_{0}^{1,q }({\varOmega ,\mathbb{R}^{d\times d}}) \) satisfies Hypothesis (H1) if

  1. (i)

    \(\mathfrak{A}_{n }\) of a finite dimensional subspace of \(W_{0}^{1,q }({\varOmega ,\mathbb{R}^{d\times d}}) \) for each \(n\in \mathbb{N^{*}}\).

  2. (ii)

    The map \(\nabla \varphi \) has a countable range whenever \(\varphi \in \mathfrak{A}_{n} \), for any \(n\in \mathbb{N} ^{*} \).

  3. (iii)

    The set \(\bigcup _{n\in \mathbb{N} ^{*}}\mathfrak{A} _{n} \) is dense in \(W_{0}^{1,q }({\varOmega ,\mathbb{R} ^{d\times d}} )\).

  4. (iv)

    For \(i\leq j\), we have the inclusion \(\mathfrak{A}_{i } \subset \mathfrak{A}_{j}\).

An explicit construction of sets satisfying Hypothesis (H1) is provided in [2]. Here, we build on the conditions of Hypothesis (H1) and we relax conditions on the underlying sets:

Hypothesis (H2).

A collection \(\{ \mathfrak{Q}_{n }\}^{\infty }_{ n=1} \) of subsets of \(W_{0}^{1,q }({\varOmega ,\mathbb{R}^{d\times d}}) \) satisfies Hypothesis (H2) if

  1. (i)

    \(\text{Span}(\mathfrak{Q}_{n})\) is of finite dimension and \(\mathfrak{Q}_{n }\) is a non-empty closed and convex subset of \(W_{0}^{1,q }({\varOmega ,\mathbb{R}^{d\times d}})\).

  2. (ii)

    The map \(\operatorname{div} \varphi \) is non-degenerate whenever \(\varphi \in \mathfrak{Q}_{n} \), for any \(n\in \mathbb{N} ^{*} \).

  3. (iii)

    The set \(\bigcup _{n\in \mathbb{N} ^{*}}\mathfrak{Q} _{n} \) is dense in \(W_{0}^{1,q }({\varOmega ,\mathbb{R} ^{d\times d}} )\).

  4. (iv)

    For \(i\leq j\), the inclusion \(\mathfrak{Q}_{i }\subset \mathfrak{Q}_{j} \) holds.

The next lemma asserts that a collection of sets can be constructed to satisfy Hypothesis (H2).

Lemma 2.4

Assume\(\mathbf{(A0)}\)holds. Then, there exists a collection of sets\(\{\mathfrak{Q}_{n }\}^{\infty } _{n=1}\)satisfying the requirements of Hypothesis (H2).

Remark 2.5

The condition\(\mathbf{(A0)}\)in Lemma 2.4 is only needed for requirement (ii) of Hypothesis (H2).

Proof

Suppose that \(\psi \) is a strictly convex function that is \(C^{1}(\bar{\varOmega }) \) and vanishes on the boundary of \(\varOmega \) as given by Assumption (A0). Let \(\varphi _{0}:\varOmega \to \mathbb{R} ^{d\times d} \) be defined by

$$ \varphi _{0}= \begin{pmatrix} \psi &0&\cdots &0 \\ 0&\psi &\cdots &0 \\ \vdots &\vdots &\ddots &\vdots \\ 0&0&\cdots &\psi \end{pmatrix} . $$

As \(\psi \) is \(C^{1}(\bar{\varOmega }) \), we have \(\varphi _{0}\in W_{0} ^{1,q }({\varOmega ,\mathbb{R}^{d\times d}}) \) and it follows that \(\operatorname{div}\varphi _{0}=\nabla \psi \). Thus, for almost every \(x \) in \(\varOmega \), we have

$$ \det \bigl(\nabla ( \operatorname{div}\varphi _{0}) (x)\bigr)=\det \bigl(\nabla ^{2} \psi (x)\bigr)>0. $$

Thanks to Lemma 5.5.3 in [1], the map \(\operatorname{div}\varphi _{0} \) is non-degenerate. Let \(\{ \mathfrak{A}_{n }\}^{\infty }_{ n=1} \) be a collection of sets satisfying Hypothesis (H1). One readily checks that the family of sets defined by

$$ \mathfrak{Q} _{n}= \biggl\{ \varphi +\epsilon \varphi _{0}: \varphi \in \mathfrak{A}_{n}; \epsilon \geq \frac{1}{n} \biggr\} $$

for \(n\in \mathbb{N} ^{*}\), satisfies hypothesis (H2). □

2.4 Special Displacements

To \(\mathcal{S}\subset W_{0}^{1,q}(\varOmega ,\mathbb{R} ^{d\times d})\) we associate \(\mathscr{U}_{\mathcal{S}}\), the set of all \(u:\varOmega \to \bar{\varLambda }\) measurable such that there exists \(\bar{c}= \bar{c}(u,\varOmega ,\varLambda ) >0 \) satisfying:

$$ \biggl\vert \int _{\varOmega }u\cdot \operatorname{div}\, \varphi \;dx \biggr\vert \leq \bar{c} \|\varphi \| _{ L^{q}( {\varOmega , \mathbb{R}^{d\times d}})} \quad \forall \varphi \in \mathcal{S}. $$
(2.9)

Remark that if \(u \in \mathscr{U}_{\mathcal{S}}\), then \(u \) belongs to \(L^{\infty }(\varOmega ,\mathbb{R}^{d}) \) since \(u\) has values in \(\bar{\varLambda }\) which is bounded. If \(\operatorname{span}(\mathcal{S})\) is of finite dimension then \(\mathscr{U}_{\mathcal{S}}\) is the set of all measurable maps \(u:\varOmega \to \bar{\varLambda }\). In fact, the linear map \(\operatorname{span}(\mathcal{S})\ni \varphi \mapsto \int _{\varOmega }u \operatorname{div}\varphi \) is continuous with respect to the \(L^{q} \)-norm as in finite dimension, all norms are equivalent. Therefore, we may find \(c\) for which inequality (2.9) holds for all \(\varphi \in \operatorname{span}(\mathcal{S})\) and in particular for all \(\varphi \in \mathcal{S}\).

At any rate, \(\mathscr{U}_{\mathcal{S}}\) contains \(W^{1,p}(\varOmega , \mathbb{R} ^{d})\). Indeed, notice that for a fixed \(u\in W^{1,p}( \varOmega ,\mathbb{R} ^{d})\), we have, for all \(\varphi \in \mathcal{S}\):

$$ \biggl\vert \int _{\varOmega }u\cdot \operatorname{div}\varphi \; dx \biggr\vert = \biggl\vert - \int _{\varOmega }\langle \nabla u,\varphi \rangle \;dx \biggr\vert \leq \Vert \nabla u \Vert _{ L^{p}( {\varOmega , \mathbb{R}^{d\times d}})} \Vert \varphi \Vert _{ L^{q}( {\varOmega , \mathbb{R}^{d \times d}})}. $$

We introduce the following set

$$ \mathscr{U}_{\mathcal{S}}^{1}= \biggl\{ u\in \mathscr{U}_{\mathcal{S}}: \int _{\varOmega }l\bigl(u(x)\bigr)\;dx= \int _{\varLambda }l(y)\;dy \; \forall l \in C _{c}\bigl( \mathbb{R} ^{d}\bigr) \biggr\} $$

and

$$\begin{aligned} \mathscr{U}_{\mathcal{S}}^{*}= \biggl\{ &(u,\beta ): u \in \mathscr{U} _{\mathcal{S}};\; \beta :\varOmega \to [0,\infty );\\ & \int _{\varOmega }l\bigl(u(x)\bigr)\beta (x)\;dx= \int _{\varLambda }l(y)\;dy \; \forall l \in C_{c}\bigl( \mathbb{R} ^{d}\bigr) \biggr\} . \end{aligned}$$

Notice that \(\mathscr{U}_{\mathcal{S}}^{1}=\{u\in \mathscr{U}_{ \mathcal{S}}: (u,1)\in \mathscr{U}_{\mathcal{S}}^{*}\}\). This corresponds to measure preserving displacements.

2.5 Extended Pseudo-Projected Gradient

Let \(\mathcal{S}\subset W_{0}^{1,q}(\varOmega ,\mathbb{R} ^{d\times d})\) and \(u \in \mathscr{U}_{\mathcal{S}}\). Define

$$ \mathcal{G}_{\mathcal{S}}( u ):= \biggl\{ G\in L^{p}\bigl( \varOmega , \mathbb{R} ^{d\times d}\bigr) : \int _{\varOmega }u \operatorname{div}\varphi \;dx =- \int _{\varOmega } \langle {G},{\varphi } \rangle \;dx \; \forall \varphi \in \mathcal{S} \biggr\} . $$

Consider the operator

$$ V^{f}_{\mathcal{S}}( u ):=\sup _{\varphi \in \mathcal{S}} \int _{\varOmega } \bigl(- u \operatorname{div}\varphi -f^{*}(\varphi ) \bigr)\;dx = \sup_{\varphi \in \mathscr{S}_{f}} \int _{\varOmega } \bigl(- u \operatorname{div}\varphi -f^{*}(\varphi ) \bigr)\;dx. $$
(2.10)

We denote by \(\varPhi _{S}( u )\) the set of maximizers of problem (2.10).

Lemma 2.6

Let\(\mathcal{S}\subset W_{0}^{1,q}(\varOmega ,\mathbb{R} ^{d\times d})\)and\(u \in \mathscr{U}_{\mathcal{S}}\).

  1. 1.

    We have

    $$ \mathcal{G}_{\mathcal{S}}( u )= \biggl\{ G\in L^{p}\bigl( \varOmega , \mathbb{R} ^{d\times d}\bigr) : \int _{\varOmega }u \operatorname{div}\varphi \,dx=- \int _{\varOmega } \langle {G},{\varphi } \rangle\, dx;\ \forall \varphi \in \operatorname{span}(\mathcal{S}) \biggr\} . $$
  2. 2.

    If\(\operatorname{span}(\mathcal{S})\)is finite dimensional, then\(\mathcal{G}_{\mathcal{S}}( u )\)is nonempty.

Proof

Set

$$\begin{aligned} \bar{\mathcal{G}}_{\mathcal{S}}(u) &= \biggl\{ G\in L^{p}\bigl({ \varOmega , \mathbb{R}^{d\times d}}\bigr): \int _{\varOmega }u \operatorname{div}\varphi \,dx=- \int _{\varOmega }\langle G,\varphi \rangle \,dx \; \forall \varphi \in \operatorname{span}(\mathcal{S}) \biggr\} . \end{aligned}$$

As \(\mathcal{S}\subset \operatorname{span}(\mathcal{S})\), we have \(\bar{\mathcal{G}}_{\mathcal{S}}(u) \subset \mathcal{ G}_{\mathcal{S}}(u) \). Next, let \(G\in \mathcal{ G}_{\mathcal{S}}(u)\). Assume that \(\varphi \in \operatorname{span}(\mathcal{S})\). We may find \(n\in \mathbb{N}\), \(\lambda _{1},\dots ,\lambda _{n}\in \mathbb{R} \) and \(\varphi _{1},\dots ,\varphi _{n}\in \mathcal{S}\) such that \(\varphi = \sum_{i=1}^{n}\lambda _{i}\varphi _{i}\). Then

$$ \int _{\varOmega }u\operatorname{div}\varphi \,dx = \int _{\varOmega }u \operatorname{div}\sum _{i=1}^{n}\lambda _{i}\varphi _{i} \,dx =\sum_{i=1} ^{n} \lambda _{i} \int _{\varOmega }u\operatorname{div}\varphi _{i} \,dx = \sum_{i=1}^{n} -\lambda _{i} \int _{\varOmega }\langle G,\varphi _{i}\rangle \,dx $$

and

$$ - \int _{\varOmega }\langle G, \varphi \rangle \,dx =- \int _{\varOmega }\Biggl\langle G,\sum_{i=1}^{n} \lambda _{i}\varphi _{i}\Biggr\rangle \,dx =\sum _{i=1}^{n} - \lambda _{i} \int _{\varOmega }\langle G,\varphi _{i}\rangle \,dx. $$

Thus \(G\in \bar{\mathcal{G}}_{\mathcal{S}}(u) \). We deduce that \(\mathcal{G}_{\mathcal{S}}(u) \subset \bar{\mathcal{G}}_{\mathcal{S}}(u) \). It follows that part (1.) holds. To obtain part (2.), we use part (1.) and the Riesz Representation Theorem. □

The following results are essentially found in Proposition 3.1 in [2].

Proposition 2.7

Suppose that the set\(\mathcal{S}\)is a finite dimensional subspace of\(W_{0}^{1,q }({\varOmega ,\mathbb{R}^{d\times d}})\)and\(f\)is\(C^{1}\)and strictly convex. Suppose, in addition that there exist constants\(c_{1},c_{2},c_{3}>0\)such that

$$\begin{aligned} -c_{3}+c_{2}|\xi |^{p}\leq f(\xi )&\leq c_{3}+c_{1}|\xi |^{p}\\ \bigl|Df(\xi )\bigr|&\leq c_{3}+c_{1}|\xi |^{p-1}\\ \bigl|Df^{*}(\xi )\bigr|&\leq c_{3}+c_{1}|\xi|^{q-1} \end{aligned}$$

for all\(\xi \in \mathbb{R} ^{d\times d}\). Then, there exists a unique map denoted\(\nabla _{\mathcal{S}}u \)that minimizes

$$ \inf_{G\in \mathcal{G}_{\mathcal{S}}( u )} \int _{\varOmega }f(G)\;dx. $$

Moreover, \(\nabla _{\mathcal{S}}u \)is the unique map\(G\in \mathcal{G}_{\mathcal{S}}( u )\)that satisfies\(Df(G)\in \mathcal{S}\).

In the next proposition, we establish similar results as in Proposition 2.7 but under weaker assumptions on \(\mathcal{S}\) and \(f\) (except in part 4).

Proposition 2.8

Assume\(\mathbf{(A1)}\)holds. Assume\(\mathcal{S}\)is a finite dimensional non-empty closed and convex subset of\(W_{0}^{1,q }( {\varOmega ,\mathbb{R}^{d\times d}})\)and let\(u \in \mathscr{U}_{ \mathcal{S}}\).

  1. 1.

    For all\(G\in \mathcal{G}_{\mathcal{S}}( u )\), \(\varphi \in \mathcal{S}\), we have

    $$ \int _{\varOmega }f(G) \;dx\geq \int _{\varOmega } \bigl(- u \operatorname{div}\varphi -f^{*}(\varphi ) \bigr)\;dx. $$
  2. 2.

    The supremum in problem (2.10) is attained.

  3. 3.

    A map\(\bar{\varphi }\)belongs to\(\varPhi _{\mathcal{S}}( u ) \)if and only if\(\bar{\varphi }\)belongs to\(\mathscr{S}_{f} \)and

    $$ \int _{\varOmega } \bigl( \operatorname{grad}\bigl[f^{*} \bigr](\bar{\varphi }) \cdot (\varphi -\bar{\varphi } ) + u \cdot ( \operatorname{div} \varphi - \operatorname{div}\bar{\varphi } ) \bigr)\;dx\geq 0 $$

    for all\(\varphi \in \mathscr{S}_{f} \).

  4. 4.

    Suppose that the hypotheses of Proposition 2.7are satisfied. Then we have

    $$ \int _{\varOmega }f(\nabla _{\mathcal{S}}u )\;dx=V_{\mathcal{S}}^{f}( u ) $$

    and\(\varPhi _{\mathcal{S}}( u )=\{Df(\nabla _{\mathcal{S}}u )\}\).

Proof

(1.) Let \(\varphi \in \mathcal{S}\) and \(G\in \mathcal{G}_{ \mathcal{S}}( u)\). By using the Legendre transformation,

$$ \int _{\varOmega }f(G)\;dx\geq \int _{\varOmega }\left( G\cdot \varphi -f^{*}(\varphi )\right)\;dx= \int _{\varOmega }\left(- u \cdot \operatorname{div}\varphi -f^{*}( \varphi )\right)\;dx. $$

(2.) Let \(\varphi \in \mathcal{S}\). We use (2.9) and (2.5) to get

$$\begin{aligned} \int _{\varOmega } \bigl( u \operatorname{div}\varphi +f^{*}(\varphi ) \bigr) \;dx&\geq -\bar{c} \|\varphi \|_{ L^{q}(\varOmega ,\mathbb{R} ^{d\times d}) }+ \int _{\varOmega }f^{*}(\varphi )\;dx \\ &\geq -\bar{c} \|\varphi \|_{ L^{q}(\varOmega ,\mathbb{R} ^{d\times d}) } +q^{-1}c^{-q} \|\varphi \|^{q}_{ L^{q}(\varOmega ,\mathbb{R} ^{d\times d}) }. \end{aligned}$$
(2.11)

In light of (2.11), \(q>1\) implies that the map

$$ \mathscr{S}_{f}\ni \varphi \mapsto T(\varphi ):= \int _{\varOmega } \bigl( u \operatorname{div}\varphi +f^{*}(\varphi ) \bigr)\;dx $$

is \(L^{q}\)-coercive. Moreover, the convexity of \(f^{*}\) guarantees that \(T\) is lower semi-continuous. The direct methods of the calculus of variations thus yield the existence of a maximizer in problem (2.10).

(3.) Let \(\bar{\varphi }\in \varPhi _{\mathcal{S}}( u ) \) so that \(\bar{\varphi }\in \mathscr{S}_{f}\). Let \(\varphi \in \mathscr{S}_{f}\) and \(\epsilon \in (0,1)\). The convexity of \(f^{*}\) ensures that \(\bar{\varphi }+\epsilon (\varphi -\bar{\varphi })\in \mathscr{S}_{f}\) and the maximality property of \(\bar{\varphi }\) implies that

$$ \int _{\varOmega }u \cdot \operatorname{div}\bar{\varphi }+f^{*}(\bar{ \varphi })\;dx\leq \int _{\varOmega }u \cdot \bigl(\operatorname{div}\bar{ \varphi }+ \epsilon \operatorname{div}(\varphi -\bar{\varphi }) \bigr) \;dx + f^{*}\bigl(\bar{\varphi }+\epsilon (\varphi -\bar{\varphi })\bigr) \;dx. $$
(2.12)

We rewrite (2.12), in turn, as

$$ \int _{\varOmega }\frac{ f^{*}(\bar{\varphi }+\epsilon (\varphi -\bar{ \varphi }))-f^{*}(\bar{\varphi })}{\epsilon }+ u \cdot \operatorname{div}( \varphi -\bar{\varphi })\; dx \geq 0. $$
(2.13)

Note that \(\operatorname{grad}[f^{*}](\bar{\varphi }+\epsilon ( \varphi -\bar{\varphi })) \) belongs to the set \(\partial f^{*}((\bar{ \varphi }+\epsilon (\varphi -\bar{\varphi }))) \) whenever \((\bar{ \varphi }+\epsilon (\varphi -\bar{\varphi })) \) is in the domain of \(f^{*} \). It follows that

$$ \int _{\varOmega } \bigl(\operatorname{grad}\bigl[f^{*} \bigr]\bigl(\bigl(\bar{\varphi }+ \epsilon (\varphi -\bar{\varphi })\bigr)\bigr) \cdot \bigl(-\epsilon (\varphi -\bar{ \varphi })\bigr) \bigr)\;dx\leq \int _{\varOmega } \bigl(f^{*}(\bar{\varphi })-f ^{*}\bigl(\bar{\varphi }+\epsilon (\varphi -\bar{\varphi })\bigr) \bigr)\;dx $$

that is,

$$ \int _{\varOmega } \bigl(\operatorname{grad}\bigl[f^{*} \bigr]\bigl(\bar{\varphi }+\epsilon (\varphi -\bar{\varphi })\bigr)\cdot (\varphi -\bar{\varphi }) \bigr)\,dx \geq \int _{\varOmega }\frac{f^{*}(\bar{\varphi }+\epsilon (\varphi -\bar{ \varphi }))-f^{*}(\bar{\varphi })}{\epsilon }\;dx. $$
(2.14)

We combine (2.13) and (2.14) to get

$$ \int _{\varOmega } \bigl(\operatorname{grad}\bigl[f^{*} \bigr]\bigl(\bar{\varphi }+\epsilon (\varphi -\bar{\varphi })\bigr)\cdot ( \varphi -\bar{\varphi })+u \operatorname{div}(\varphi -\bar{\varphi }) \bigr)\,dx\geq 0. $$
(2.15)

First, we assume that \(\mathbf{(A1)}\)(iii)(a) holds. In light of (2.6), we have \(\operatorname{grad}[f^{*}](\bar{\varphi }+\epsilon (\varphi -\bar{\varphi }))=\operatorname{grad}[f^{*}](\bar{ \varphi })\). Equation (2.15) becomes

$$ \int _{\varOmega } \bigl(\operatorname{grad}\bigl[f^{*} \bigr](\bar{\varphi })\cdot ( \varphi -\bar{\varphi })+u\operatorname{div}(\varphi -\bar{\varphi }) \bigr)\,dx\geq 0 . $$

Second, we assume that \(\mathbf{(A1)}\)(iii)(b) holds. In light of Remark 2.1(ii), we use the growth condition on \(\nabla f^{*} \) in (2.4), the Lebesgue dominated convergence theorem and let \(\epsilon \) go to 0 in (2.15) to obtain that:

$$ \int _{\varOmega } \bigl(\operatorname{grad}\bigl[f^{*} \bigr](\bar{\varphi })\cdot ( \varphi -\bar{\varphi })+u\operatorname{div}(\varphi -\bar{\varphi }) \bigr)\,dx\geq 0. $$

We next show the converse implication. Let \(\varphi \in \mathscr{S} _{f} \) such that

$$ 0\leq \int _{\varOmega } \bigl(u\operatorname{div}(\varphi -\bar{\varphi })+\operatorname{grad}\bigl[f^{*}\bigr](\bar{\varphi })\cdot ( \varphi -\bar{ \varphi }) \bigr)\,dx, $$
(2.16)

for all \(\varphi \in \mathscr{S}_{f}\). We notice that, as \(f^{*}\) is convex, the range of the map \(\operatorname{grad}[f^{*}](\bar{\varphi })\) lies in the sub-differential of \(f^{*}\) so that \(f^{*}(\varphi )-f ^{*}(\bar{\varphi })\geq \operatorname{grad}[f^{*}](\bar{\varphi })(\varphi -\bar{\varphi } )\) for all \(\varphi \in \mathscr{S}_{f}\). Then, the inequality (2.16) implies that

$$\begin{aligned} 0&\leq \int _{\varOmega } \bigl(u\operatorname{div}(\varphi -\bar{\varphi })+ \bigl(f^{*}(\varphi )-f^{*}(\bar{\varphi })\bigr) \bigr)\,dx \end{aligned}$$

for all \(\varphi \in \mathscr{S}_{f}\), that is,

$$\begin{aligned} \int _{\varOmega } \bigl(u\operatorname{div}\bar{\varphi }+ f^{*}(\bar{ \varphi }) \bigr)\,dx\leq \int _{\varOmega } \bigl(u\operatorname{div} \varphi + f^{*}(\varphi ) \bigr)\,dx \end{aligned}$$

for all \(\varphi \in \mathscr{S}_{f}\). We conclude that \(\bar{\varphi }\in \varPhi _{\mathcal{S}}( u ) \).

(4.) Thanks to Proposition 2.7, \(D f( \nabla _{\mathcal{S}}u)\in \mathcal{S}\). Next, we set \(\varphi _{0}:= D f(\nabla _{\mathcal{S}}u)\). By definition of \(f^{*}\),

$$ f(\nabla _{\mathcal{S}}u ) + f^{*}(\varphi )\geq \varphi \cdot \nabla _{\mathcal{S}}u $$

for all \(\varphi \in \mathcal{S}\). As \(f\) is convex and \(\varphi _{0} = D f(\nabla _{\mathcal{S}}u)\), we have

$$ f(\nabla _{\mathcal{S}}u )+ f^{*}(\varphi _{0}) = \varphi _{0}\cdot \nabla _{\mathcal{S}}u. $$

Thus,

$$ \int _{\varOmega }f(\nabla _{\mathcal{S}}u )\;dx\geq \int _{\varOmega }\varphi \cdot \nabla _{\mathcal{S}}u \;dx- \int _{\varOmega }f^{*}(\varphi )\;dx= \int _{\varOmega }- u \operatorname{div}\varphi \;dx - \int _{\varOmega }f^{*}( \varphi )\;dx $$

and

$$ \int _{\varOmega }f(\nabla _{\mathcal{S}}u )\;dx= \int _{\varOmega }\varphi _{0} \cdot \nabla _{\mathcal{S}}u\;dx - \int _{\varOmega }f^{*}(\varphi _{0})\;dx= \int _{\varOmega }- u \operatorname{div}\varphi _{0} \;dx- \int _{\varOmega }f ^{*}(\varphi _{0})\;dx. $$

We deduce that \(\varphi _{0}\in \varPhi _{\mathcal{S}}( u )\). Since \(f^{*}\) is strictly convex, we conclude that \(\varPhi _{\mathcal{S}}( u )= \{D f(\nabla _{\mathcal{S}}u)\}\) and moreover, \(\int _{\varOmega }f( \nabla _{\mathcal{S}}u )=V^{f}_{\mathcal{S}}( u )\), see (2.10). □

In the next proposition, we establish a convergence result in the spirit of (1.7). We also connect the operator \(V_{ \mathcal{S}}^{f}\) with the usual notions of gradient and total variation.

Proposition 2.9

Assume\(\mathbf{(A1)}\)holds. Assume that\(\mathcal{S}_{n}\)is a finite dimensional non-empty closed and convex subset of\(W_{0}^{1,q }( {\varOmega ,\mathbb{R}^{d\times d}})\)for each\(n\geq 1\). The following holds.

  1. 1.

    If\(\lbrace \mathcal{S}_{n} \rbrace _{n=1}^{\infty } \)is a monotonically increasing family of subsets of some set\(\mathcal{S}_{0} \)and\(\bigcup _{n\in \mathbb{N} ^{*}}\mathcal{S}_{n} \)is dense in\(\mathcal{S}_{0} \)with respect to the\(W_{0}^{1,q}(\varOmega ,\mathbb{R} ^{d\times d}) \)norm then

    $$ \lim_{n\rightarrow \infty } V_{\mathcal{S}_{n}}^{f}[u]=V_{\mathcal{S} _{0}}^{f}[u] $$

    for any\(u\in \mathscr{U}_{\mathcal{S}_{0}}\).

  2. 2.

    If\(\mathcal{S}=W_{0}^{1,q}(\varOmega ,\mathbb{R} ^{d\times d} )\)and\(u\in W^{1,p}(\varOmega ,\mathbb{R} ^{d} )\)then\(V_{\mathcal{S}}^{f}[u]= \int _{\varOmega }f(\nabla u)\,dx \).

  3. 3.

    Assume\(u\in BV(\varOmega ,\mathbb{R} ^{d\times d}) \)and\(f(\xi )=|\xi | \)for all\(\xi \in \mathbb{R} ^{d\times d}\). If\(\mathcal{S}=W_{0}^{1,q}(\varOmega ,\mathbb{R} ^{d\times d}) \)then\(V_{\mathcal{S}}^{f}[u] \)is the total variation of\(u \).

Remark 2.10

A consequence of Proposition 2.9is the following: If the sequence of sets\(\{\mathcal{S}_{n} \}_{n\in \mathbb{N} ^{*}}\)is monotonically increasing to\(W_{0}^{1,q}(\varOmega ,\mathbb{R} ^{d\times d})\)and\(u\in W^{1,p}(\varOmega ,\mathbb{R} ^{d} )\)we have

$$ \lim_{n\rightarrow \infty } V_{\mathcal{S}_{n}}^{f}[u]= \int _{\varOmega }f( \nabla u)\;dx. $$

Proof

(1.) Recall that

$$ V_{\mathcal{S}_{n}}^{f}[u]=\sup_{\varphi \in \mathcal{S}_{n}} \biggl\{ \int _{\varOmega } \bigl(-u\cdot \operatorname{div}\varphi -f^{*}(\varphi ) \bigr)\,dx \biggr\} . $$

As \(\lbrace \mathcal{S}_{n} \rbrace _{n=1}^{\infty } \) is a monotonically increasing, \(\lim_{n\rightarrow \infty } V_{S_{n}} ^{f}[u]\) exists. Moreover, since \(\mathcal{S}_{n}\subset \mathcal{S} _{0} \) for all \(n\geq 1\),

$$ \lim_{n\rightarrow \infty } V_{\mathcal{S}_{n}}^{f}[u] \leq V_{ \mathcal{S}_{0}}^{f}[u]. $$
(2.17)

Let \(\epsilon >0 \) and choose \(\varphi ^{\epsilon }\in \mathcal{S} _{0}\) such that

$$ V_{\mathcal{S}_{0}}^{f}[u]\leq \epsilon + \int _{\varOmega } \bigl(-u \cdot \operatorname{div}\varphi ^{\epsilon }-f^{*}\bigl(\varphi ^{\epsilon }\bigr) \bigr)\,dx. $$

Let \(\{\varphi ^{\epsilon }_{n}\} _{n\in \mathbb{N} ^{*}} \) be a sequence converging to \(\varphi ^{\epsilon }\) in \(W_{0}^{1,q}(\varOmega ,\mathbb{R} ^{d\times d}) \) and such that \(\varphi _{n}^{\epsilon } \in \mathcal{S}_{n} \) for all \(n\in \mathbb{N} ^{*} \). Then, we use the growth conditions on \(f^{*} \) in (2.4) and (2.5), the continuity of \(f^{*}\) on its domain and the Lebesgue dominated convergence theorem to obtain that

$$ \int _{\varOmega }-f^{*}\bigl(\varphi ^{\epsilon } \bigr)\,dx = \lim_{n\rightarrow \infty } \int _{\varOmega }-f^{*}\bigl(\varphi ^{\epsilon }_{n}\bigr)\,dx. $$

It follows that

$$\begin{aligned} V_{\mathcal{S}_{0}}^{f}[u]&\leq \epsilon + \int _{\varOmega } \bigl(-u \cdot \operatorname{div}\varphi ^{\epsilon }-f^{*}\bigl(\varphi ^{\epsilon }\bigr) \bigr)\,dx \\ &= \epsilon + \lim_{n\rightarrow \infty } \int _{\varOmega } \bigl(-u \cdot \operatorname{div}\varphi ^{\epsilon }_{n}-f^{*}\bigl(\varphi ^{\epsilon }_{n}\bigr) \bigr)\,dx \\ &\leq \epsilon + \limsup_{n\to \infty }V_{\mathcal{S}_{n}}^{f}[u] \\ &= \epsilon + \lim_{n\to \infty }V_{\mathcal{S}_{n}}^{f}[u]. \end{aligned}$$

As \(\epsilon \) is arbitrary, we have

$$ \lim_{n\rightarrow \infty } V_{\mathcal{S}_{n}}^{f}[u] \geq V_{ \mathcal{S}_{0}}^{f}[u]. $$
(2.18)

From (2.17) and (2.18), we conclude that \(\lim_{n\rightarrow \infty } V_{\mathcal{S}_{n}}^{f}[u]=V_{\mathcal{S}_{0}}^{f}[u]\).

(2.) One has

$$\begin{aligned} V_{\mathcal{S}}^{f}[u]&= \sup_{\varphi \in W_{0}^{1,q}(\varOmega ,\mathbb{R} ^{d\times d} ) } \biggl\{ \int _{\varOmega } \bigl(-u\cdot \operatorname{div}\varphi -f^{*}( \varphi ) \bigr)\,dx \biggr\} \\ &= \sup_{\varphi \in W_{0}^{1,q}(\varOmega ,\mathbb{R} ^{d\times d} ) } \biggl\{ \int _{\varOmega } \bigl(\nabla u\cdot \varphi -f^{*}( \varphi ) \bigr)\,dx \biggr\} \\ &\leq \int _{\varOmega }f(\nabla u)\;dx. \end{aligned}$$

The inequality above is obtained by using the definition of the Legendre transform \(f^{*}\) of \(f\). Let \(\bar{\varphi }\in \partial f (\nabla u) \). Then \(f^{*}(\bar{\varphi })+f(\nabla u)=\nabla u\cdot \bar{ \varphi }\). Thanks to the growth conditions (2.2) and (2.3) on \(f \), it holds that \(\bar{\varphi }\in L ^{q}(\varOmega ,\mathbb{R} ^{d\times d}) \). Since \(W_{0}^{1,q}(\varOmega , \mathbb{R} ^{d\times d})\) is dense in \(L^{q}(\varOmega ,\mathbb{R} ^{d \times d}) \) for the \(L^{q}(\varOmega ,\mathbb{R} ^{d\times d}) \) norm, we get

$$\begin{aligned} \int _{\varOmega }f(\nabla u)\;dx&= \int _{\varOmega } \bigl( \nabla u\cdot \bar{ \varphi }-f^{*}(\bar{\varphi }) \bigr)\;dx \\ &\leq \sup_{\varphi \in W_{0}^{1,q}(\varOmega ,\mathbb{R} ^{d\times d} ) } \biggl\{ \int _{\varOmega } \bigl(\nabla u\cdot \varphi -f^{*}( \varphi ) \bigr)\;dx \biggr\} \\ &= V^{f}_{\mathcal{S}}[u]. \end{aligned}$$

We conclude that \(V^{f}_{S}[u]= \int _{\varOmega }f(\nabla u)\;dx\).

(3.) The total variation of \(u\in BV(\varOmega ,\mathbb{R} ^{d\times d}) \) is

$$ \|Du\|(\varOmega ) =\sup \biggl\{ \int _{\varOmega }u\cdot \operatorname{div}\varphi \;dx: \varphi \in C_{c}^{1}\bigl({\varOmega , \mathbb{R} ^{d\times d}} \bigr); |\varphi |\leq 1 \biggr\} $$
(2.19)

while, using the Legendre transform of \(f(\xi )= |\xi |\), we obtain for every \(q>1\)

$$ V^{f}_{S}(u)=\sup \biggl\{ \int _{\varOmega }u\cdot \operatorname{div} \varphi \;dx: \varphi \in W_{0}^{1,q}\bigl(\varOmega ,\mathbb{R} ^{d\times d} \bigr); |\varphi |\leq 1 \biggr\} . $$
(2.20)

It follows directly from (2.19) and (2.20) that \(\|Du\|(\varOmega )\leq V^{f}_{\mathcal{S}}[u]\). The converse inequality \(\|Du\|(\varOmega )\geq V^{f}_{\mathcal{S}}[u]\) follows from the density of \(C_{c}^{1}({\varOmega ,\mathbb{R} ^{d\times d}})\) in \(W_{0}^{1,q}(\varOmega ,\mathbb{R} ^{d\times d})\) and an argument similar to the one made in the proof of (2) in the proposition. □

3 Minimization with General Displacements

We consider the following:

$$ \inf_{(u,\beta )\in \mathscr{U}^{*}_{\mathcal{S}}} \biggl\{ I(u,\beta )= V_{\mathcal{S}}^{f}(u)+ \int _{\varOmega } \bigl( H(\beta )-F\cdot u \bigr)\,dx \biggr\} . $$
(3.1)

This problem will be studied via a dual problem that we will formulate next. We assume in this section that Assumption (A2) holds.

3.1 An Auxiliary Problem

For \(l, k:\mathbb{R}^{d} \rightarrow (-\infty ,\infty ]\), define for \(u,v\in \mathbb{R} ^{d}\)

$$\begin{aligned} l^{\#}(v):=\sup_{u\in \bar{\varLambda },t>0} \bigl\{ u \cdot v-l( u) t-H(t) \bigr\} \end{aligned}$$
(3.2)

and

$$\begin{aligned} k_{\#}(u):=\sup_{v\in \mathbb{R}^{d} ,t>0} \bigl\{ (1/t) \bigl( u \cdot v-k( v) -H(t) \bigr) \bigr\} . \end{aligned}$$
(3.3)

Under Assumption (A2), it is known that \(((l^{\#})_{\#})^{\#}=l ^{\#}\) and \(((k_{\#})^{\#})_{\#}=k_{\#}\) (see for instance Lemma A1 of [11]). Call \(\mathcal{C}\) the set of all functions \((k,l)\) with \(k,\;l:\mathbb{R}^{d} \rightarrow \mathbb{R} \cup \{ \infty \}\) Borel measurable, finite at least at one point, and satisfying \(l\equiv \infty \) on \(\mathbb{R}^{d} \setminus \bar{\varLambda }\) and such that

$$ k(v) +t\cdot l( u) +H(t)\geq u\cdot v\quad \forall u,v\in \mathbb{R}^{d} , t>0. $$
(3.4)

Call \(\mathcal{C}'\) the set of all functions \((k,l)\in \mathcal{C}\) such that \(l=k_{\#} \) and \(k=l ^{\#} \). The set \(\mathcal{C}'\) is nonempty. Indeed, \((\chi _{\bar{\varLambda }}^{\#},(\chi _{\bar{\varLambda }}^{\#})_{\#}) \in \mathcal{C}'\) as \(((\chi _{\bar{\varLambda }}^{\#})_{\#})^{\#}= \chi _{\bar{\varLambda }}^{\#}\).

Let \(\mathscr{A}\) be the set of \((k,l,\varphi )\) such that \((k,l) \in \mathcal{C}\) and \(\varphi \in \mathcal{S}\). Consider the following functional defined on \(\mathscr{A}\):

$$ J(k,l,\varphi ):= \int _{\varOmega }k(F+\operatorname{div} \varphi )\;dx+ \int _{\varLambda }l\;dy+ \int _{\varOmega }f^{*}(\varphi )\;dx. $$

The following problem will play an important role in this section:

$$ \inf \bigl\{ J(k,l,\varphi ): (k,l,\varphi )\in \mathscr{A} \bigr\} . $$
(3.5)

The value of the expression (3.5) is the opposite of the value of the following expression:

$$ \sup \bigl\{ -J(k,l,\varphi ): (k,l,\varphi )\in \mathscr{A} \bigr\} . $$
(3.6)

Let \(\mathscr{A}'\) denote the subset of \(\mathscr{A}\) consisting of all \((k,l,\varphi )\in \mathscr{A}\) that satisfy \((k,l)\in \mathcal{C}' \). It holds that

$$ \inf \bigl\{ J(k,l,\varphi ): (k,l,\varphi )\in \mathscr{A} \bigr\} = \inf \bigl\{ J(k,l,\varphi ): (k,l,\varphi )\in \mathscr{A}' \bigr\} . $$
(3.7)

Indeed, the key observation to this end is that for \((k,l,\varphi ) \in \mathscr{A}\), one has \(l\geq k_{\#}\) and \(k\geq (k_{\#})^{\#}\) so that

$$ J(k,l,\varphi )\geq J\bigl((k_{\#})^{\#}, k_{\#},\varphi \bigr)\quad \mbox{and} \quad \bigl((k_{\#})^{\#}, k_{\#},\varphi \bigr)\in \mathscr{A}'. $$

For \(R>0\), we set

$$ \mathscr{A}_{R}= \bigl\{ (k,l,\varphi )\in \mathscr{A}':J(k,l, \varphi )\leq R \bigr\} . $$

Lemma 3.1

Assume that\(\mathbf{(A1)}\), \(\mathbf{(A2)}\)and\(\mathbf{(A3)}\)hold. Let\((k,l,\varphi )\in \mathscr{A}_{R}\). Set\(s_{l}:=- \inf_{u\in \bar{\varLambda}}l(u)\). Then,

$$ \int _{ \varOmega }k(F+ \operatorname{div}\varphi )\;dx\geq \mathcal{L} ^{d}{(\varOmega )}H^{*}(s_{l}) -r^{*} \|F\|_{ L^{1}( \varOmega ) }. $$

Moreover, there exists\(M:= M(R,F,f,\varOmega ,\varLambda )>0 \)such that

$$ |s_{l}|\leq M. $$
(3.8)

Proof

As \(\varLambda \) is bounded and \(l\) is convex, we choose \(u_{l}\in \overline{\varLambda }\) such that \(-l(u_{l})=s_{l}\). Since \(k:=l ^{\#}\), in view of (3.2), we have

$$ -tl(u_{l}) -H(t)+u_{l}\cdot v= ts_{l} -H(t)+u_{l}\cdot v \leq H^{*}(s _{l})+u_{l}\cdot v \leq k(v). $$
(3.9)

Using the last inequality in (3.9), one gets

$$\begin{aligned} \int _{ \varOmega }k(F+ \operatorname{div}\varphi )\;dx &\geq \int _{ \varOmega } \bigl(H^{*}(s_{l})+u_{l} \cdot (F+\operatorname{div}\varphi ) \bigr) \;dx \end{aligned}$$
(3.10)
$$\begin{aligned} &= H^{*}(s_{l})\mathcal{L}^{d}{(\varOmega )}+ \int _{\varOmega }u_{l}\cdot F \;dx. \end{aligned}$$
(3.11)

We have used the fact that \(u_{l}\) is a constant vector and \(\varphi \in W_{0}^{1,q}(\varOmega ,\mathbb{R} ^{d\times d} )\) to obtain the equality in (3.11). Hence,

$$ \int _{ \varOmega }k(F+ \operatorname{div}\varphi )\;dx\geq \mathcal{L} ^{d}{(\varOmega )}H^{*}(s_{l}) -r^{*} \|F\|_{L^{1}(\varOmega )}. $$

Thus,

$$ R\geq J(k,l,\varphi )\geq -s_{l} \mathcal{L}^{d}( \varLambda )+ \mathcal{L}^{d}{(\varOmega )}H^{*}(s_{l}) -r^{*} \|F\|_{ L^{1}( \varOmega ) }+ \inf f^{*}. $$

Thanks to Lemma 2.2(iii), \(s_{l} \) is bounded uniformly in \(l \). □

Lemma 3.2

Assume that\(\mathbf{(A1)}\), \(\mathbf{(A2)}\)and\(\mathbf{(A3)}\)hold.

  1. 1.

    There exists\(M>0 \)such that for all\((k,l,\varphi )\in \mathscr{A}_{R}\)one has

    $$ \int _{\varLambda }\bigl|l(y)\bigr|\,dy \leq M. $$
    (3.12)
  2. 2.

    There exist\(a_{0},b_{0},c_{0}>0 \)such that for all\((k,l,\varphi )\in \mathscr{A}_{R}\), the map\(k \)is\(r^{*} \)-Lipschitz, and one has for all\(v\in \mathbb{R}^{d} \)

    $$ -c_{0}+a_{0}|v|\leq k(v)\leq b_{0}+r^{*}|v|. $$
    (3.13)

Proof

(1.) Recall that for \((k,l,\varphi )\in \mathcal{A}_{R}\), one has

$$ { J(k,l,\varphi )= \int _{\varOmega }k(F+\operatorname{div} \varphi )\;dx+ \int _{\varLambda }l\;dy+ \int _{\varOmega }f^{*}(\varphi )\;dx. } $$

By Lemma 3.1, for all \((k,l,\varphi )\in \mathcal{A}_{R}\), if we define \(s_{l}:=-\inf_{u\in \bar{\varLambda }}l(u)\), we get

$$ R\geq J(k,l,\varphi )\geq \mathcal{L}^{d}{(\varOmega )}H^{*}(s_{l})-r ^{*} \|F \|_{ L^{1}( \varOmega ) } + \int _{\varLambda } l(y) \,dy + \mathcal{L} ^{d}{(\varOmega )}\inf f^{*}. $$

Rearranging the terms, we get:

$$ \int _{\varLambda } l(y) \,dy\leq R-\mathcal{L}^{d}{( \varOmega )}H^{*}(s_{l})+r ^{*} \|F \|_{ L^{1}( \varOmega ) } - \inf f^{*}\mathcal{L}^{d}{( \varOmega )}. $$

By definition of \(s_{l}\) we also have \(-s_{l}\mathcal{L}^{d}{(\varOmega )} \leq \int _{\varLambda } l(y) \,dy\) and thus

$$ -s_{l}\mathcal{L}^{d}{(\varOmega )} \leq \int _{\varLambda } l(y) \,dy\leq R- \mathcal{L}^{d}{( \varOmega )}H^{*}(s_{l})+r^{*} \|F \|_{ L^{1}( \varOmega ) } - \inf f^{*}\mathcal{L}^{d}{( \varOmega )}. $$
(3.14)

We consider the negative part of \(l\) defined by \(l^{-}:=\max \{-l,0\}\) and note that

$$ \int _{\varLambda } \bigl|l(y)\bigr| \,dy = \int _{\varLambda } l(y) \,dy+2 \int _{\varLambda } l ^{-}(y) \,dy. $$
(3.15)

Observe that, by the definition of \(s_{l}\), we have \(l^{-}\leq |s_{l}|\). This, combined with (3.14), (3.15) and (3.8) yields (3.12).

(2.) Let \((k,l,\varphi )\in \mathscr{A}_{R}\). Since \(k=l^{\#}\), by Eq. (3.2), \(k\) is a \(r^{*}\)-Lipschitz as \(\varLambda \) has diameter less or equal to \(r^{*}\). Next, we have

$$\begin{aligned} k(0)&= \sup_{u\in \bar{\varLambda }, t>0} \bigl\{ -tl(u)-H(t) \bigr\} \\ &= \sup_{ t>0} \bigl\{ -ts_{l}-H(t) \bigr\} . \end{aligned}$$

As \(s_{l}\) is uniformly bounded, the growth condition on \(H\) ensures that \(|k(0)|\) is uniformly bounded say by some \(b_{0}>0\). We get then the inequality \(k(v)\leq b_{0}+r^{*}|v|\) for all \(v\in \mathbb{R} ^{d}\).

Because of the hypothesis on the domain \(\varLambda \), we take \(a_{0}>0\) such that \(B(0,a_{0})\subset \varLambda \). As \((k,l,\varphi ) \in \mathcal{A}_{R}\), we use relation (3.4) to obtain for \(v\neq 0\)

$$\begin{aligned} k(v)\geq v\cdot \biggl(a_{0}\frac{v}{|v|} \biggr)-l \biggl(a_{0} \frac{v}{|v|} \biggr)-H(1). \end{aligned}$$
(3.16)

Thanks to inequality (3.12), \(\int _{\varLambda }|l|\,dy\) is uniformly bounded in \(l \). We use in addition the fact that \(l \) is bounded to deduce that \(\sup_{y\in \bar{B}(0,a_{0})}|l|(y) \) is bounded by a constant independent of \(l\) (see for instance Theorem 1, p. 236 in [6]). Thus Eq. (3.16) implies that there exists \(c_{0}>0\) such that \(k(v)\geq a_{0}|v|-c_{0}\) for all \(v\in \mathbb{R} ^{d}\). □

Proposition 3.3

Assume that\(\mathbf{(A1)}\), \(\mathbf{(A2)}\), and\(\mathbf{(A3)}\)hold. Assume\(\mathcal{S}\)is a finite dimensional non-empty closed and convex subset of\(W_{0}^{1,q }({\varOmega ,\mathbb{R}^{d\times d}})\). Then, the functional\(J \)admits a minimizer\(( k_{0}, l_{0},\varphi _{0})\)in\(\mathscr{A}' \).

Proof

Let \((\bar{k},\bar{l},\bar{\varphi })\in \mathscr{A} \). Set \(R=J(\bar{k},\bar{l},\bar{\varphi })\). Take a minimizing sequence \(\{(k_{n},l_{n},\varphi _{n})\}_{n\in \mathbb{N} ^{*}} \) of problem (3.5) that is in \(\mathscr{A}_{R} \). By Lemma 3.1 and the growth condition on \(f^{*}\) we may assume without loss of generality that \(\{ \varphi _{ n}\}^{\infty } _{ n=1} \) converges to some \(\varphi _{0}\in \mathcal{S}\) weakly in \(L^{q}(\varOmega ,\mathbb{R} ^{d\times d}) \). Since \(\text{Span}( \mathcal{S})\) is finite dimensional, \(\{ \varphi _{ n}\}^{\infty } _{ n=1} \) converges to some \(\varphi _{0}\in \mathcal{S}\) strongly in the \(L^{q}(\varOmega ,\mathbb{R} ^{d\times d}) \) norm. We deduce

$$ \int _{\varOmega }f^{*}(\varphi _{0})\;dx \leq \liminf_{n\to \infty } \int _{\varOmega }f^{*}(\varphi _{n})\;dx. $$
(3.17)

From Lemma 3.2, as \(l_{n}\) is convex, we use Ascoli-Arzelà Theorem together with Theorem 1, p. 236 in [6] to deduce that up to a subsequence, we may assume that \((k_{n},l_{n}) \) converges locally uniformly \(\mathbb{R} ^{d} \times \varLambda \) to \((k_{0},l_{0}) \in \mathcal{C}' \). The Lebesgue dominated convergence together with inequality (3.13) yield

$$ \int _{\varOmega }k(F+\operatorname{div}\varphi _{0}) \;dx\leq \liminf_{n\to \infty } \int _{\varOmega }k_{n}(F+\operatorname{div}\varphi _{n})\;dx. $$
(3.18)

Since \(\{ l_{n}\}^{\infty } _{n=1} \) is uniformly bounded below (thanks to Lemma 3.1), by Fatou’s Lemma we get

$$ \int _{\varLambda }l_{0}\;dy\leq \liminf _{n\to \infty } \int _{\varLambda }l _{n}\;dy. $$
(3.19)

By inequalities (3.17), (3.18) and (3.19), we get

$$ J(k_{0},l_{0},\varphi _{0})\leq \liminf _{n\to \infty }J(k_{n},l_{n}, \varphi _{n}) $$

and \((k_{0},l_{0},\varphi _{0})\) is a minimizer of \(J\) over \(\mathscr{A}'\). □

3.2 A Uniqueness Result

Here, we prove the main result of this section. We will need the following lemma which is in the spirit of Lemma 4.3 and Lemma 4.4 in [2]. A proof of Lemma 3.4 is given in Sect. A.1.

Lemma 3.4

Assume that(A2)holds. Consider a lower semicontinuous function\(l_{0}:\mathbb{R}^{d} \to \bar{\mathbb{R}} \)such that\(\inf_{\bar{\varLambda }}l_{0}>-\infty \); \(l_{0}\)is finite on\(\varLambda \)and\(l_{0}\equiv +\infty \)on\(\mathbb{R}^{d} \setminus \bar{ \varLambda }\). Set\(k_{0}= ({l_{0}})^{\#}\). Let\(v\in \mathbb{R}^{d} \)be such that\(k_{0}\)is differentiable at\(v\).

  1. 1.

    There exist unique\(u_{0}\in \bar{\varLambda }\)and\(t_{0}>0\)such that\(k_{0}( v ) =-t_{0}l_{0}(u_{0})-H(t_{0})-u_{0}\cdot v\). In addition, \(u_{0} \)and\(t_{0} \)are characterized by\(u_{0}=\nabla k_{0}( v ) \)and\(H'(t_{0})+l(u_{0})=0\).

  2. 2.

    Let\(\hat{l}\in C_{b}(\mathbb{R}^{d} )\)and let\(1\geq \epsilon >0\). Define\(l_{\epsilon }=l_{0}+\epsilon \hat{l}\)and\(k_{\epsilon }= { (l_{\epsilon } )}^{\#}\).

    1. (a)

      There exists a constant\(M\)independent of\(v\)and\(\epsilon \)such that

      $$ \biggl\vert \frac{k_{\epsilon }(v)-k_{0}( v ) }{\epsilon } \biggr\vert \leq M. $$
    2. (b)

      We have

      $$ \lim_{\epsilon \rightarrow 0 } \frac{k_{\epsilon }(v)-k_{0}( v ) }{ \epsilon }=- t_{0} \hat{l}(u_{0}). $$

Next, we give the main result of this section.

Theorem 3.5

Assume that\(\mathbf{(A1)}\), \(\mathbf{(A2)}\), and\(\mathbf{(A3)}\)hold. Assume\(\mathcal{S}\)is a finite dimensional non-empty closed and convex subset of\(W_{0}^{1,q }({\varOmega ,\mathbb{R}^{d\times d}})\). Assume\(F\)satisfies the condition\(\textbf{(ND)}_{\mathcal{S}}\). Then, problems (3.1) and (3.6) are dual. Problem (3.6) admits a maximizer\((k_{0},l_{0}, \varphi _{0}) \)with\(k_{0}=l_{0}^{\#} \)and\(l_{0}=(k_{0})_{\#} \). Problem (3.1) admits a unique minimizer\((u_{0},\beta _{0}) \). Moreover\(u_{0} \)satisfies

$$\begin{aligned} \textstyle\begin{cases} u_{0}= \nabla k_{0}(F+\operatorname{div} \varphi _{0})\\ \varphi _{0}\in \varPhi _{S}(u_{0}). \end{cases}\displaystyle \end{aligned}$$

Proof

Step 1. For \((u,\beta )\in \mathscr{U}^{*}_{ \mathcal{S}}\) and \((k,l,\varphi )\in \mathscr{A}\), one has

$$\begin{aligned} I(u, \beta ) &= V_{\mathcal{S}}^{f}(u)+ \int _{\varOmega } \bigl( H(\beta )-F\cdot u \bigr)\;dx \\ &\geq \int _{\varOmega } \bigl( -u\cdot (\mathop{div} \varphi +F) \bigr)\,dx- \int _{\varOmega }f^{*}(\varphi )\;dx \\ &\quad{}+ \int _{\varOmega }H(\beta )\,dx+ \int _{\varOmega }\beta l (u)\;dx - \int _{\varLambda }l(y)\;dy \\ &\geq \int _{\varOmega }-k (\mathop{div} \varphi +F)\;dx - \int _{\varOmega }f^{*}(\varphi )\;dx - \int _{\varLambda }l(y)\;dy. \end{aligned}$$

Thus \(I(u,\beta )\geq -J(k,l,\varphi )\) with equality if and only if \(\varphi \in \varPhi _{\mathcal{S}}(u)\) and

$$ k(F+\operatorname{div} \varphi )+\beta l(u)+H(\beta )=u\cdot (F+ \operatorname{div} \varphi ). $$

Note that if \(k\) is convex, the map \(\nabla k (F+\operatorname{div} \varphi )\) is well defined as the map \(F+\operatorname{div} \varphi \) is non-degenerate. Using Lemma 3.4(i), it follows that if \(k\) is convex, then \(I(u,\beta )=-J(k,l,\varphi )\) if and only if

$$ \textstyle\begin{cases} \varphi \in \varPhi _{\mathcal{S}}(u)\\ u =\nabla k(F+\operatorname{div} \varphi )\\ \beta =(H')^{-1}(-l(u)). \end{cases} $$
(3.20)

Step 2. Thanks to Eq. (3.7), we may find a maximizer \((k_{0},l_{0},\varphi _{0}) \) of problem (3.5) satisfying \(k_{0}=l_{0}^{\#} \) and \(l_{0}=(k_{0})_{ \#} \). The function \(u_{0}=\nabla k_{0}(F+\operatorname{div} \varphi _{0})\) is well defined as \(k_{0}\) is convex and we set \(\beta _{0} =(H')^{-1}(-l(u _{0}))\). We have to show that \((u_{0},\beta _{0})\ \in \mathscr{U}^{*} _{\mathcal{S}}\) and \(\varphi _{0} \in \varPhi _{\mathcal{S}}(u_{0})\).

Step 3. Let \(\bar{l}\in C_{c}(\mathbb{R}^{d} ) \). For \(\epsilon \in (0,1) \), define \(l_{\epsilon }=l_{0}+\epsilon \bar{l} \) and \(k_{\epsilon }=(l_{\epsilon })^{\#} \). Using Lemma 3.4, one has

$$\begin{aligned} &\lim_{\epsilon \rightarrow 0^{+}} \int _{\varOmega }(1/\epsilon ) \bigl( k_{0}(F+ \operatorname{div} \varphi _{0})-k_{\epsilon }(F+ \operatorname{div} \varphi _{0}) \bigr) \,dx \\ &\quad= \int _{\varOmega }\beta _{0}\bar{l}\bigl(\nabla k_{0}(F+\operatorname{div} \varphi _{0})\bigr)\;dx= \int _{\varOmega }\beta _{0}\bar{l}(u_{0}) \;dx. \end{aligned}$$
(3.21)

Since \(J(k_{0},l_{0},\varphi _{0})\leq J(k_{\epsilon },l_{\epsilon }, \varphi _{0}) \), we deduce that \(-\int _{\varLambda }\bar{l}\;dy + \int _{\varOmega }\beta _{0}\bar{l}(u_{0})\;dx\leq 0\). As we can replace \(\bar{l} \) by \(-\bar{l} \), one deduces that \(\int _{\varLambda } \bar{l} \;dy=\int _{\varOmega }\beta _{0}\bar{l}(u_{0})\;dx\). Therefore \((u_{0},\beta _{0})\in \mathscr{U}_{S}^{*} \).

Step 4. Let \(\varphi \in \mathcal{S}\). For \(\epsilon \in (0,1) \), set \(\varphi _{\epsilon }=\epsilon \varphi +(1-\epsilon ) \varphi _{0} \). By the convexity of \(\mathcal{S}\), the map \(\varphi _{\epsilon }\) belongs to \(\mathcal{S}\). As \(J(k_{0},l_{0}, \varphi _{0})\leq J(k_{0},l_{0},\varphi _{\epsilon }) \), we have

$$\begin{aligned} 0&\leq \int _{\varOmega }(1/\epsilon ) \bigl( k_{0}\bigl(F+ \operatorname{div} \varphi _{0}+\epsilon \operatorname{div}(\varphi - \varphi _{0})\bigr) - k _{0}(F+\operatorname{div} \varphi _{0}) \bigr)\;dx \\ &\quad{} +(1/\epsilon ) \int _{\varOmega } \bigl(f^{*}\bigl(\varphi _{0}+\epsilon (\varphi -\varphi _{0})\bigr) -f^{*}(\varphi _{0}) \bigr)\;dx \end{aligned}$$
(3.22)

Thanks to Lemma 3.4, Inequality (3.22) implies

$$\begin{aligned} &\int _{\varOmega } \bigl( u_{0}\cdot \operatorname{div}( \varphi - \varphi _{0}) +\operatorname{grad}\bigl[ f^{*}\bigr](\varphi _{0})\cdot (\varphi - \varphi _{0}) \bigr) \;dx \\ &\quad= \int _{\varOmega } \bigl( \nabla k_{0}(F+ \operatorname{div} \varphi _{0})\cdot \operatorname{div}(\varphi - \varphi _{0}) +\operatorname{grad}\bigl[ f^{*}\bigr]( \varphi _{0})\cdot (\varphi - \varphi _{0}) \bigr)\;dx \\ &\quad\geq 0. \end{aligned}$$

It follows from Proposition 2.8 that \(\varphi _{0}\in \varPhi _{\mathcal{S}}(u_{0}) \).

Step 5. Since \((u_{0},\beta _{0}) \in \mathscr{U}^{*}_{ \mathcal{S}}\), \(\varphi _{0} \in \varPhi _{\mathcal{S}}(u_{0})\), \(u_{0}=\nabla k_{0}(F+\operatorname{div} \varphi _{0})\), and \(\beta _{0} = (H')^{-1}(-l(u_{0}))\), we deduce that \(I(u_{0},\beta _{0})=J(k _{0},l_{0},\varphi _{0})\) and \(u_{0}\) is a minimizer of problem (3.1) thanks to relation (3.20). Suppose \((u_{1},\beta _{1})\ \in \mathscr{U}^{*}_{S}\) is another minimizer of problem (3.1). Then we have \(I(u_{1},\beta _{1})=J(k_{0},l _{0},\varphi _{0})\) and by relation (3.20), we get \(u_{1}= \nabla k_{0}(F+\operatorname{div} \varphi _{0})\) which implies \(u_{1}=u_{0}\). Next the strict convexity of \(H\) yields that \(\beta _{0}=\beta _{1}\). We conclude that \((u_{0},\beta _{0})\) is the unique minimizer of problem (3.1) and \(u_{0}\) is characterized by

$$\begin{aligned} \textstyle\begin{cases} u_{0}= \nabla k_{0}(F+\operatorname{div} \varphi _{0})\\ \varphi _{0}\in \varPhi _{\mathcal{S}}(u_{0}). \end{cases}\displaystyle \end{aligned}$$

 □

Corollary 3.6

Assume that\(\mathbf{(A0)}\), \(\mathbf{(A1)}\), \(\mathbf{(A2)}\), and\(\mathbf{(A3)}\)hold. Assume\(\mathcal{S}\)is a finite dimensional non-empty closed and convex subset of\(W_{0}^{1,q }({\varOmega , \mathbb{R}^{d\times d}})\)and\(\nabla \varphi \)is non-degenerate whenever\(\varphi \in \mathcal{S}\). Suppose\(F \)has a countable range (thus degenerate). Then, \(F\)satisfies the condition\(\textbf{(ND)} _{\mathcal{S}}\)and problem (3.1) admits a unique solution.

Corollary 3.7

Assume that\(\mathbf{(A1)}\), \(\mathbf{(A2)}\), and\(\mathbf{(A3)}\)hold. Assume\(\mathcal{S}\)is a finite dimensional subspace of\(W_{0}^{1,q }({\varOmega ,\mathbb{R}^{d\times d}})\)and\(\nabla \varphi \)has a countable range whenever\(\varphi \in \mathcal{S}\). Suppose\(F \)is non-degenerate. Then, \(F\)satisfies the condition\(\textbf{(ND)}_{ \mathcal{S}}\)and problem (3.1) admits a unique solution.

4 The Incompressible Case

Throughout this section, we assume that \(\mathcal{S}\) is a subset of \(W_{0}^{1,q}(\varOmega ,\mathbb{R} ^{d\times d})\). We consider the following problem:

$$ \inf_{u\in \mathcal{U}_{\mathcal{S}}^{1}} \biggl\{ I_{0}(u):=V^{f}_{ \mathcal{S}}( u)- \int _{\varOmega }F\cdot u\; dx \biggr\} $$
(4.1)

and we recall that the set \(\mathcal{U}_{\mathcal{S}}^{1}\) is defined as

$$ \mathscr{U}_{\mathcal{S}}^{1}= \biggl\{ u\in \mathscr{U}_{\mathcal{S}}: \int _{\varOmega }l\bigl(u(x)\bigr)\;dx= \int _{\varLambda }l(y)\;dy \; \forall l \in C _{c}\bigl( \mathbb{R} ^{d}\bigr) \biggr\} . $$

We assume \(\mathcal{L}^{d}(\varOmega )=\mathcal{L}^{d}(\varLambda )\) so that \(\mathscr{U}_{\mathcal{S}}^{1}\) is non-empty.

4.1 Existence and Uniqueness via Duality

We study problem (4.1) via duality. Let \(u\in \mathscr{U}^{1}_{\mathcal{S}}\), \(\varphi \in \mathcal{S}\), \(l\in C( \varLambda )\) and \(k:\mathbb{R}^{d} \to \mathbb{R} \) satisfy \(k(v)+l(u)\geq u\cdot v\) for all \(u\in \varLambda \) and all \(v\in \mathbb{R}^{d} \). One has

$$\begin{aligned} &V^{f}_{\mathcal{S}}(u)-\int _{\varOmega }F\cdot u\; dx \end{aligned}$$
(4.2)
$$\begin{aligned} &\quad= - \int _{\varOmega }u\cdot (F+\operatorname{div} \varphi )\;dx+ \int _{\varOmega }l(u)\;dx- \int _{\varLambda }l(y)\;dy- \int _{\varOmega }f^{*}( \varphi )\;dx \end{aligned}$$
(4.3)
$$\begin{aligned} &\quad\geq - \int _{\varOmega }k(F+\operatorname{div} \varphi )\;dx- \int _{\varLambda }l(y)\,dy - \int _{\varOmega }f^{*}(\varphi )\;dx. \end{aligned}$$
(4.4)

This suggests that we consider the dual problem

$$ M_{0}:=\inf_{( k,l,\varphi )\in A_{0}} \biggl\{ J(k,l,\varphi ):= \int _{\varOmega }k(F+\operatorname{div} \varphi )\;dx+ \int _{\varLambda }l(y) \;dy + \int _{\varOmega }f^{*}(\varphi )\;dx \biggr\} $$
(4.5)

with \(A_{0} \) being the set of all \((k,l,\varphi ) \) such that \(\varphi \in \mathcal{S}\), \(l\in C( \varLambda )\), \(\inf_{\varLambda }l=0 \) and \(k:\mathbb{R}^{d} \to \mathbb{R} \) satisfies \(k(v)+l(u)\geq u \cdot v\) for all \(u\in \bar{\varLambda }\), and all \(v\in \mathbb{R}^{d} \). Remark that we have

$$ -M_{0}=\sup_{( k,l,\varphi )\in A_{0}} \bigl\{ -J(k,l,\varphi ) \bigr\} . $$
(4.6)

4.1.1 Existence and Regularity of Minimizers of Problem (4.5)

Denote by \(\mathcal{C}\) the set of all \((k,l) \) such that \(k:\mathbb{R} ^{d}\to \mathbb{R} \) and \(l:\mathbb{R} ^{d}\to \mathbb{R} \cup \{\infty \} \) satisfy

$$ k(v)+l(u)\geq u\cdot v;\quad \forall u\in \bar{\varLambda };\ \forall v\in \mathbb{R} ^{d} $$
(4.7)

and \(l\equiv \infty \) on \(\mathbb{R} ^{d}\setminus \bar{\varLambda }\). Consider the subset \(\mathcal{C}_{0}\) of \(\mathcal{C}\) consisting of \((k,l)\in \mathcal{C}\) such that \(l\in C( \varLambda )\) and \(\inf_{\varLambda }l=0\). The following lemma is standard:

Lemma 4.1

Let\((k,l)\in \mathcal{C}\). It holds that\((l^{*}, l^{**})\in \mathcal{C}\), \(l^{*}\leq k\), \(0\leq l^{**}\leq l\)and\(l^{***}=l^{*}\). If\((k,l)\in \mathcal{C}_{0}\)then\(l^{*}(0)=0\).

Let us denote by \(C_{0}' \) the set of all \((k,l)\in C_{0}\) such that \(l^{*}=k\), \(k^{*}=l\), \(k(0)=0\), and \(l\geq 0\), and by \(A'_{0}\) the set of all \((k,l,\varphi )\) with \((k,l)\in C_{0}'\) and \(\varphi \in \mathcal{S}\). Remark that an element in \(C_{0}' \) is the couple \((\chi _{\bar{\varLambda }},(\chi _{\bar{\varLambda }})^{*})\). Hence \(A'_{0}\) is nonempty when \(\mathcal{S}\) is nonempty. One readily checks that, in light of Lemma 4.1, problem (4.5) has the same infimum value as

$$ \inf_{(k,l,\varphi )\in A_{0}'} \biggl\{ J(k,l,\varphi ):= \int _{ \varOmega }k(F+\operatorname{div} \varphi )\;dx+ \int _{\varOmega }f^{*}( \varphi )\;dx+ \int _{\varLambda }l(y)\;dy \biggr\} . $$
(4.8)

We recall that \(r^{*}\) is such that \(B(0,1/r^{*})\subset \varLambda \subset B(0,r^{*}/2)\).

Lemma 4.2

Assume that\(\mathbf{(A1)}\)and\(\mathbf{(A3)}\)hold. Assume that the set\(\mathcal{S}\)is a finite dimensional non-empty closed and convex subset of\(W_{0}^{1,q }({\varOmega ,\mathbb{R}^{d\times d}})\). Then, problem (4.8) admits a minimizer\((k_{0},l_{0}, \varphi _{0})\in A_{0}' \)with\(k_{0} \)convex and\(r^{*} \)-Lipschitz and\(k_{0}(0)=0 \).

Proof

Consider a minimizing sequence \(\lbrace (k_{n},l_{n}, \varphi _{n}) \rbrace _{n=1}^{\infty }\) of problem (4.8). Since \(k_{n}=l_{n}^{*}\) and \(l_{n}=(k_{n})^{*}\), \(k_{n}\) is \(r^{*}\)-Lipschitz. As \(k_{n}(0)=0\), we use Ascoli-Arzelà theorem to deduce that a subsequence of \(\lbrace k_{n} \rbrace _{n=1}^{\infty }\) converges locally uniformly to some \(k_{0}\). Next, using the growth condition (2.5) on \(f^{*}\) as well as the facts that \(k_{n}\) is \(r^{*}\)-Lipschitz, \(k_{n}(0)=0\), we establish the following estimate:

$$ J(k_{n},l_{n},\varphi _{n})\geq \int _{\varOmega } \biggl(-r^{*}|F+\operatorname{div} \varphi _{n}|+c^{p}\frac{|\varphi _{n}|^{q}}{q} -b \biggr) \;dx+ \int _{\varLambda }l_{n}(y)\;dy. $$
(4.9)

As the left hand side of (4.9) is bounded, \(l_{n}\geq 0\) and \(\mathcal{S}\) is finite dimensional, we deduce from (4.9) that a subsequence of \(\lbrace \varphi _{n} \rbrace _{n=1}^{\infty }\) converges strongly to some \(\varphi _{0}\) in \(W_{0}^{1,q}(\varOmega ,\mathbb{R} ^{d\times d})\). Invoking (4.9) again, we show that \(\lbrace \int _{\varLambda }l_{n}(y)\,dy \rbrace _{n=1}^{\infty } \) is bounded. This, combined with the fact that \(l_{n}\) is non-negative and convex, yields the existence of a subsequence of \(\lbrace l_{n} \rbrace _{n=1}^{\infty }\) that converges locally uniformly to some \(l_{0}\) (see for instance Theorem 1, p. 236 in [6]). One readily checks that \((k_{0}, l_{0}, \varphi _{0})\in A_{0}'\). We next exploit lower semi-continuity properties of the functional \(J\) to conclude that \((k_{0}, l_{0}, \varphi _{0})\) is a minimizer of \(J\) over \(A_{0}'\). □

4.1.2 A Duality Result

We have the following theorem.

Theorem 4.3

Assume that\(\mathbf{(A1)}\)and\(\mathbf{(A3)}\)hold. Assume\(\mathcal{S}\)is a finite dimensional non-empty closed and convex subset of\(W_{0}^{1,q }({\varOmega ,\mathbb{R}^{d\times d}})\). Suppose that the map\(F \)satisfies the condition\(\textbf{(ND)}_{\mathcal{S}}\). Then problems (4.1) and (4.6) are dual. Problem (4.6) admits a maximizer\((k_{0},l_{0}, \varphi _{0}) \)with\(k_{0}=l_{0}^{*} \)and\(l_{0}=(k_{0})^{*} \). Problem (4.1) admits a unique minimizer\(u_{0} \). Moreover\(u_{0} \)satisfies

$$\begin{aligned} \textstyle\begin{cases} u_{0}=\nabla k_{0}(F+\operatorname{div} \varphi )\\ \varphi _{0}\in\varPhi _{S}(u_{0}). \end{cases}\displaystyle \end{aligned}$$

Proof

Suppose \(u\in \mathscr{U}_{S}^{1} \) and \(( k,l,\varphi ) \in A_{0} \). Using (4.3) and (4.4), we see that \(I_{0}(u)\geq -J(k,l,\varphi )\) with equality if and only if \(\varphi \in \varPhi _{S}(u) \) and \(l(u)+k(F+\operatorname{div}\varphi )=u\cdot (F+\operatorname{div} \varphi ) \) for almost every \(x\in \varOmega \). The latter condition reduces to \(u(x)=\nabla k(F(x)+\operatorname{div} \varphi (x)) \) if \(k\) is convex, under the assumption \(F+\operatorname{div} \varphi \) is non-degenerate. Now, let \((k_{0},l_{0},\varphi _{0}) \in A_{0}'\) be a minimizer of \(J \) over \(A_{0} \). Since \(F+\operatorname{div} \varphi _{0} \) is non-degenerate and \(k_{0} \) is convex, the map \(u_{0}=\nabla k_{0}(F+\operatorname{div}\varphi _{0}) \) is well defined.

Variation around\(l_{0}\). Let \(\bar{l}\in C_{c}( \mathbb{R}^{d} ) \). For \(\epsilon \in (0,1) \), set \(l_{\epsilon }=l _{0}+\epsilon \bar{l} \) and \(k_{\epsilon }=(l_{\epsilon })^{*}\). Let \(v\in \mathbb{R} ^{d} \) be a point where \(k_{0} \) is differentiable. Using the measurable selection theorem, one deduces that there exists \(T_{\epsilon }:\mathbb{R} ^{d}\to \mathbb{R}^{d} \) measurable such that for all \(\epsilon \in [0,1) \)

$$ k_{\epsilon }(v)=T_{\epsilon }(v)\cdot v-l_{\epsilon } \bigl(T_{\epsilon }(v)\bigr). $$

Then, for \(\epsilon \in (0,1)\), we have

$$\begin{aligned} \bar{l}\bigl(T_{\epsilon }(v)\bigr)\leq -(1/\epsilon ) \bigl( k_{\epsilon }(v)-k _{0}( v ) \bigr) \leq \bar{l} \bigl(T_{0}(v)\bigr) \end{aligned}$$
(4.10)

and

$$\begin{aligned} \bigl\vert (1/\epsilon ) \bigl( k_{\epsilon }(v)-k_{0}( v ) \bigr) \bigr\vert \leq \|\bar{l}\|_{ L^{\infty }(\mathbb{R}^{d} ) }. \end{aligned}$$
(4.11)

Moreover,

$$\begin{aligned} \lim_{\epsilon \rightarrow {0^{+}} } -(1/\epsilon ) \bigl( k_{\epsilon }(v)-k_{0}( v ) \bigr) =\bar{l} \bigl(T_{0}(v)\bigr). \end{aligned}$$
(4.12)

We refer the reader to Lemma A.3 for (4.10)–(4.12). Hence, as

$$ T_{0}(F+\operatorname{div} \psi _{0})=\nabla k_{0}(F+ \operatorname{div} \psi _{0})=u_{0} \quad \mbox{a.e.} $$

using again (4.12), one has

$$\begin{aligned} &\lim_{\epsilon \rightarrow 0^{+}} \int _{\varOmega }(1/{\epsilon }) \bigl( k_{0}(F+ \operatorname{div} \psi _{0})-k_{\epsilon }(F+ \operatorname{div} \psi _{0}) \bigr) \,dx \\ &\quad= \int _{\varOmega }\bar{l}\bigl(T _{0}(F+ \operatorname{div} \psi _{0})\bigr)\;dx= \int _{\varOmega }\bar{l}(u_{0}) \;dx. \end{aligned}$$
(4.13)

Since \(J(k_{0},l_{0},\varphi _{0})\leq J(k_{\epsilon },l_{\epsilon }, \varphi _{0}) \), we deduce from (4.13) that \(-\int _{\varLambda }\bar{l} +\int _{\varOmega }\bar{l}(u_{0})\leq 0\). By replacing \(l \) by \(-l \) in the above argument, one deduces that \(\int _{\varLambda }\bar{l} =\int _{\varOmega }\bar{l}(u_{0})\). As a result, \(u_{0}\in \mathscr{U}_{S}^{1} \).

Variation around\(\varphi _{0}\). Let \(\varphi \in \mathcal{S}\). For \(\epsilon \in (0,1) \), by convexity of \(\mathcal{S}\), we have \(\varphi _{\epsilon }:=\epsilon \varphi +(1- \epsilon ) \varphi _{0}\in \mathcal{S}\). Then \(J(k_{0},l_{0},\varphi _{0})\leq J(k_{0},l_{0},\varphi _{\epsilon }) \). This implies that

$$\begin{aligned} 0&\geq \int _{\varOmega }(1/\epsilon ) \bigl( k_{0}(F+ \operatorname{div} \varphi _{0})-k_{0}\bigl(F+ \operatorname{div} \varphi _{0}+\epsilon \operatorname{div}(\varphi - \varphi _{0})\bigr) \\ &\quad{}+ f^{*}(\varphi _{0})-f^{*}(\varphi _{0}+\epsilon (\varphi -\varphi _{0})\bigr) \;dx . \end{aligned}$$

As \(\epsilon \) tends to \(0^{+}\), the above equation yields

$$\begin{aligned} 0&\geq - \int _{\varOmega }\nabla k_{0}(F+\operatorname{div} \varphi _{0}) \cdot \operatorname{div}(\varphi - \varphi _{0}) -\operatorname{grad}\bigl[ f^{*}\bigr](\varphi _{0})\cdot (\varphi - \varphi _{0})\;dx \end{aligned}$$

It follows from Proposition 2.8 that \(\varphi _{0}\in \varPhi _{\mathcal{S}}(u_{0}) \). □

Corollary 4.4

Assume that\(\mathbf{(A0)}\), \(\mathbf{(A1)}\), and\(\mathbf{(A3)}\)hold. Assume that\(\mathcal{S}\)is a finite dimensional non-empty closed and convex subset of\(W_{0}^{1,q }({\varOmega ,\mathbb{R}^{d\times d}})\)and\(\nabla \varphi \)is non-degenerate whenever\(\varphi \in \mathcal{S}\). Suppose\(F \)has a countable range (thus degenerate). Then, \(F \)satisfies the condition\(\textbf{(ND)}_{\mathcal{S}}\)and problem (4.1) admits a unique solution.

Corollary 4.5

Assume that\(\mathbf{(A1)}\)and\(\mathbf{(A3)}\)hold. Assume that\(\mathcal{S}\)is a finite dimensional subspace of\(W_{0}^{1,q }({\varOmega , \mathbb{R}^{d\times d}})\)and\(\nabla \varphi \)has a countable range whenever\(\varphi \in \mathcal{S}\). Suppose\(F \)is non-degenerate. Then, \(F \)satisfies the condition\(\textbf{(ND)}_{\mathcal{S}}\)and problem (4.1) admits a unique solution.

4.2 A Link Between Problem (3.1) and Problem (4.1)

Here, we explore the relationships between problem (3.1) and problem (4.1). For this purpose, we make a further assumption of the domains \(\varOmega \) and \(\varLambda \) by requiring that \(\varOmega =\varLambda \). Assume \(\mathbf{(A1)}\) holds and recall \(\lbrace H_{n} \rbrace _{n=0}^{\infty }\) as defined in (2.7) and (2.8). Then, Lemma 2.3 ensures that \(\mathbf{(A2)}\) holds for \(H_{n}\) for all \(n\geq 1\). Define

$$ I_{n}(u,\beta ):=V_{\mathcal{S}}^{f}( u)+ \int _{\varOmega }\left( H_{n}(\beta )-u \cdot F\right)\;dx \quad {n \geq 1} $$

and

$$ I_{0}(u ):=V_{\mathcal{S}}^{f}( u)- \int _{\varOmega }u\cdot F\;dx. $$

Recall that \(C_{0}\) is the set of all \((k,l)\) such that \(l\in C(\bar{ \varLambda })\), \(\inf l=0\) and \(k:\mathbb{R} ^{d}\to \mathbb{R} \) satisfies for all \(u\in \varLambda \) and all \(v\in \mathbb{R} ^{d}\):

$$ k(v)+l(u)\geq u\cdot v. $$
(4.14)

Let \(C_{n}\) be the set of all \((k,l)\) such that \(l\in C(\bar{\varLambda})\) and \(k:\mathbb{R} ^{d}\to \mathbb{R} \) satisfy:

$$ k(v)+tl(u)+H_{n}(t)\geq u\cdot v;\quad \forall u \in \varLambda ;\ \forall v \in \mathbb{R} ^{d}. $$
(4.15)

We denote by \(\mathcal{A}_{0}\) the set of all \((k,l,\varphi )\) satisfying \((k,l)\in C_{0}\) and \(\varphi \in S\). Similarly \(\mathcal{A}_{n}\) denotes the set of all \((k,l,\varphi )\) satisfying \((k,l)\in C_{n}\) and \(\varphi \in S\). If \((k,l,\varphi )\in \mathcal{A}_{0}\cup \mathcal{A}_{n}\), we still set

$$ J(k,l,\varphi )= \int _{\varOmega }k(F+\operatorname{div} \varphi )\;dx+ \int _{\varLambda }l(y)\;dy + \int _{\varOmega }f^{*}(\varphi )\;dx. $$

Lemma 4.6

Assume that\(\mathbf{(A1)}\), \(\mathbf{(A2)}\), and\(\mathbf{(A3)}\)hold. Assume that\(\mathcal{S}\)is a finite dimensional non-empty closed and convex subset of\(W_{0}^{1,q }({\varOmega ,\mathbb{R}^{d\times d}})\). For each\(n\in \mathbb{N} \), let\((u_{n},\beta _{n})\)be the unique minimizer of\(I_{n}\)over\(\mathscr{U}_{\mathcal{S}}^{*}\)as given by Theorem 3.5and let\((k_{n},l_{n},\varphi _{n})\)be a minimizer of\(J\)over\(\mathcal{A}_{n}\)with\(k_{n}\)convex and\(r^{*}\)-Lipschitz as ensured by Proposition 3.3and Lemma 4.2. Then,

  1. 1.

    The sequence\(\{I_{n}(u_{n},\beta _{n})\}_{n\in \mathbb{N} ^{*}}\)is bounded.

  2. 2.

    The sequence\(\{ \beta _{n} \}_{n\in \mathbb{N} ^{*}}\)converges to 1 in\(L^{2}(\varOmega ) \).

  3. 3.

    The sequence\(\{ \varphi _{n} \}_{n\in \mathbb{N} ^{*}}\)admits a subsequence that converges to some\(\bar{\varphi }\)in\(S\)with respect to the\(W_{0}^{1,q} (\varOmega ,\mathbb{R}^{d\times d} )\)-norm.

Proof

Step 1. Let \(\bar{u}\in \mathscr{U}_{\mathcal{S}}^{1}\). We have \((\bar{u},1)\in \mathscr{U}_{\mathcal{S}}^{*}\) and thus \(I_{n}(u_{n},\beta _{n})\leq I_{n}(\bar{u},1)\) for all \(n\geq 1\). As \(H_{n}(1)=0\), it holds that \(I_{n}(\bar{u},1)=V_{\mathcal{S}}^{f}( \bar{u})-\int _{\varOmega }\bar{u}\cdot F\;dx\) which is finite. Hence

$$ R_{0}:=V_{\mathcal{S}}^{f}( \bar{u})- \int _{\varOmega }\bar{u}\cdot F\;dx \geq I_{n}(u_{n}, \beta _{n}). $$
(4.16)

On the other hand, we use growth condition (2.5) to get

$$ I_{n}(u_{n},\beta _{n}) \geq \int _{\varOmega }(-b+u_{n}\cdot F)\; dx\geq -b \mathcal{L}^{d}(\varOmega )-r^{*}\|F\|_{ L^{1}(\varOmega ,\mathbb{R} ^{d}) } := - R_{1}. $$
(4.17)

Finally, we use (4.16) and (4.17) to prove (1).

Step 2. Let \(\varphi _{0} \in \mathcal{S}\). As \(u_{n}\) has values in \(\varLambda \), it holds that

$$ V_{\mathcal{S}}^{f}(u_{n})=\sup _{\varphi \in S} \int _{\varOmega }\bigl(-u_{n} \operatorname{div}\varphi -f^{*}(\varphi )\bigr)\;dx\geq \int _{\varOmega }\bigl(-r ^{*}|\operatorname{div} \varphi _{0}|-f^{*}(\varphi _{0})\bigr) \;dx=: R_{2} $$
(4.18)

and

$$ \int _{\varOmega }-u_{n}\cdot F\;dx\geq -r^{*}\|F\|_{ L^{1}(\varOmega , \mathbb{R} ^{d}) }. $$
(4.19)

We combine (4.16), (4.17), (4.18), (4.19) to get

$$ R_{2} -r^{*}\|F\|_{ L^{1}(\varOmega ,\mathbb{R} ^{d}) }+ \int _{\varOmega }H _{n}(\beta _{n})\;dx \leq I_{n}(u_{n},\beta _{n})\leq R_{0}. $$
(4.20)

Setting \(c_{0} \mathcal{L}^{d}(\varOmega ):= R_{0}- R_{2}+r^{*}\|F\|_{ L ^{1}(\varOmega ,\mathbb{R} ^{d}) }\), we use Lemma 2.3 and (4.20) to obtain

$$\begin{aligned} \int _{\varOmega }n\bigl(\beta _{n}(x)-1 \bigr)^{2} \,dx\leq \bigl(c_{0}+\bar{H}-H(1)\bigr) \mathcal{L}^{d}(\varOmega ). \end{aligned}$$

This establishes (2).

Step 3. As \(\lbrace H_{n} \rbrace _{n=1}^{ \infty } \) is a non-decreasing sequence that converges to \(H_{0}\), it holds that \(C_{n+1}\subset C_{n}\subset C_{0}\) for all \(n\in \mathbb{N} \). Thus, as \((k_{n},l_{n})\in C_{n}\), we have \((k_{n},l _{n})\in C_{0}\) so that

$$\begin{aligned} k_{n}(F+\operatorname{div}\varphi _{n}) + l_{n}(x)\geq x\cdot (F+ \operatorname{div} \varphi _{n}). \end{aligned}$$
(4.21)

Since \(-J(k_{n},l_{n},\varphi _{n})=I_{n}(u_{n},\beta _{n})\), we have \(J(k_{n},l_{n},\varphi _{n})\leq R_{1}\) for all \(n\in \mathbb{N} ^{*}\). This, combined with \(\varOmega =\varLambda \), and (4.21) yields

$$\begin{aligned} R_{1}&\geq \int _{\varOmega } \bigl(k_{n}(F+\operatorname{div} \varphi _{n})+l_{n}(x)+f^{*}(\varphi _{n}) \bigr)\,dx\\ \end{aligned}$$
(4.22)
$$\begin{aligned} &\geq \int _{\varOmega } \bigl(x\cdot (F+\operatorname{div}\varphi _{n})+f ^{*}(\varphi _{n}) \bigr)\,dx. \end{aligned}$$
(4.23)

In view of the growth condition (2.5) and boundedness of \(\varOmega \), (4.22) implies

$$\begin{aligned} R_{1}\geq \int _{\varOmega } \biggl(r^{*}|F+\operatorname{div} \varphi _{n}|-b+c ^{p}\frac{|\varphi _{n}|^{q}}{q} \biggr)\,dx. \end{aligned}$$
(4.24)

As the space \(\mathcal{S}\) is of finite dimension and the \(\operatorname{div}\) operator is continuous on \(\mathcal{S}\), we conclude that \(\lbrace \varphi _{n} \rbrace _{n=1}^{\infty } \) is convergent up to a subsequence in \(W_{0}^{1,q} (\varOmega ,\mathbb{R}^{d\times d} ) \) which allows us to conclude (3). □

Theorem 4.7

Assume that\(\mathbf{(A1)}\), \(\mathbf{(A2)}\), and\(\mathbf{(A3)}\)hold. Assume that\(\mathcal{S}\)is a finite dimensional non-empty closed and convex subset of\(W_{0}^{1,q }({\varOmega ,\mathbb{R}^{d\times d}})\). Assume\(F\)satisfies the condition\(\mathbf{(ND)_{\mathcal{S}}}\). For each\(n\in \mathbb{N} \), let\((u_{n},\beta _{n})\)be the unique minimizer of\(I_{n}\)over\(\mathscr{U}_{\mathcal{S}}^{*}\)as given by Theorem 3.5and let\((k_{n},l_{n},\varphi _{n})\)be a minimizer of\(J\)over\(\mathcal{A}_{n}\)with\(k_{n}\)convex and\(r^{*}\)-Lipschitz as ensured by Proposition 3.3and Lemma 4.2. Suppose that\(k_{n}\)is differentiable for all\(n\in \mathbb{N} ^{*}\). Then, the sequence\(\{ u_{n} \}_{n\in \mathbb{N} ^{*}}\)converges almost everywhere to the unique minimizer\(u_{0}\)of\(I_{0}\)over \(\mathscr{U}_{S}^{1}\). In addition, the minima\(\lbrace I_{n}(u_{n},\beta _{n}) \rbrace _{n=1}^{\infty } \)converge to\(I_{0}(u_{0})\).

Proof

Step 1. For \(n\in \mathbb{N} ^{*}\), set \(\bar{k}_{n}=k _{n}-k_{n}(0)\). Note that we have \(\bar{k}_{n}(0)=0\). Since the functions \(k_{n}\) are \(r^{*}\)-Lipschitz, so are the functions \(\bar{k}_{n}\) and we obtain that, up to a subsequence, the sequence \(\lbrace \bar{k}_{n} \rbrace _{n=1}^{\infty }\) converges locally uniformly to a certain function \(\bar{k}\). Since \(F+ \operatorname{div}\varphi _{n}\) is non-degenerate, we have that \(\nabla \bar{k}_{n}(F+\operatorname{div}\varphi _{n})\) is well-defined. Furthermore, Lemma 4.6 ensures that \(\lbrace \varphi _{n} \rbrace _{n=1}^{\infty } \) converges up to a subsequence to some \(\bar{\varphi }\in S\) with respect to the \(W^{1,q} (\varOmega , \mathbb{R}^{d} )\)-norm. As a result, \(\lbrace \operatorname{div}\varphi _{n} \rbrace _{n=1}^{ \infty }\) converges to \(\operatorname{div}\bar{\varphi }\) in \(L^{q}(\varOmega ,\mathbb{R}^{d})\). Since \(\mathcal{S}\) is of finite dimension, the \(L^{q}\) convergence of \(\lbrace \operatorname{div} \varphi _{n} \rbrace _{n=1}^{\infty }\) reduces to a pointwise convergence. Next, using the convexity of the \(\bar{k}_{n}\) and the pointwise convergence of \(\lbrace \operatorname{div}\varphi _{n} \rbrace _{n=1}^{\infty }\) to \(\operatorname{div}\varphi \), we deduce that up to a subsequence \(\lbrace \nabla \bar{k}_{n}(F+ \operatorname{div}\varphi _{n}) \rbrace _{n=1}^{\infty } \) converges a.e to \(\nabla \bar{k}(F+\operatorname{div}\bar{\varphi })\) (cf. [13] Theorem 25.7).

As a duality result, Theorem 3.5 ensures that \(\nabla \bar{k}_{n}(F+\operatorname{div}\varphi _{n})=u_{n}\). If we denote \(\bar{u}:= \nabla \bar{k}(F+\operatorname{div}\bar{\varphi })\), then, up to a subsequence, the sequence \(\{u_{n}\}_{n\in \mathbb{N} }\) converges a.e to \(\bar{u}\).

Step 2. Let \(l\in C_{b}(\mathbb{R}^{d} )\). The strong convergence in \(L^{2}(\varOmega )\) of \(\lbrace \beta _{n} \rbrace _{n=1}^{\infty } \) to 1 established in Lemma 4.6 and the almost everywhere convergence of \(\{u_{n}\}_{n\in \mathbb{N} }\) to \(\bar{u}\) obtained in Step 1 ensure that \(\lim_{n\rightarrow \infty }\int _{\varOmega }\beta _{n} l(u_{n})\,dx= \int _{\varOmega }l(\bar{u} (x))\,dx\). As \((u_{n},\beta _{n})\in \mathscr{U} _{S}^{*}\), \(\int _{\varOmega }\beta _{n}(x) l(u_{n})\,dx=\int _{\varOmega }l(y)\,dy\) for all \(l\in C_{b}(\mathbb{R}^{d} )\). It follows that in the limit \(\int _{\varOmega }l(\bar{u})\,dx=\int _{\varOmega }l(y)\,dy\) for all \(l\in C_{b}( \mathbb{R}^{d} )\) and thus \(\bar{u}\in \mathscr{U}_{S}^{1}\).

Step 3. We recall that

$$ I_{ n}(u,\beta ) =V_{\mathcal{S}}^{f}(u)+ \int _{\varOmega } \bigl( H_{ n}( \beta )-u\cdot F \bigr)\;dx. $$

Since \(u\mapsto V_{S}^{f}(u)\) is lower-semicontinuous as a supremum of affine functions, by applying the Fatou’s Lemma, we have

$$ \liminf_{n} I_{ n}(u_{n},\beta _{n}) \geq V_{S}^{f}(\bar{u})+ \int _{ \varOmega }-\bar{u}\cdot F\;dx =I_{0}(\bar{u}). $$

Let \(u_{0}\) be the unique minimizer of \(I_{0}\) over \(\mathscr{U}_{S} ^{1}\) as given by Theorem 4.3. Then,

$$ \liminf_{n} I_{ n}(u_{n}, \beta _{n}) \geq I_{0}(\bar{u})\geq I_{0}(u _{0}). $$
(4.25)

Meanwhile, as \(C_{n}\subset C_{0}\) and \((k_{0},l_{0},\varphi _{0})\) is a minimizer of \(J\) over \(C_{0}\), we have

$$ J(k_{0},l_{0},\varphi _{0})\leq J(k_{n},l_{n},\varphi _{n}). $$

This, along with the duality established in Theorem 3.5 imply that

$$ \limsup_{n} I_{ n}(u_{n},\beta _{n})\leq \limsup _{n} \bigl(-J(k_{n},l _{n},\varphi _{n}) \bigr) \leq -J(k_{0},l_{0},\varphi _{0})=I_{0}(u _{0}). $$
(4.26)

We combine (4.25) and (4.26) to obtain \(I_{0}(\bar{u})=I_{0}(u_{0})\). As \(u_{0}\) is the unique minimizer of \(I_{0}\) over \(\mathscr{U}_{S}^{1}\) we have \(u_{0}=\bar{u}\). We note that the limit \(\bar{u}\) does not depend on the subsequence of \(\{u_{n}\} _{n}\) chosen. Thus, the whole sequence \(\{u_{n}\}_{n}\) converges a.e. to \(u_{0}\). In addition, \(\{I_{n}(u_{n},\beta _{n})\}_{n}\) converges to \(I_{0}(u_{0})\). □