This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Partial Regularity for 𝔸\mathbb{A}-quasiconvex Functionals

Matthias Bärlin  and  Konrad Keßler
Abstract.

We establish partial Hölder regularity for (local) generalised minimisers of variational problems involving strongly quasi-convex integrands of linear growth, where the full gradient is replaced by a first order homogeneous differential operator 𝔸\mathbb{A} with constant coefficients. Working under the assumption of 𝔸\mathbb{A} being \mathbb{C}-elliptic, this is achieved by adapting a method recently introduced in [33, 32].

1. Introduction

1.1. Variational Problems

The analysis of functionals taking the form =Ωf(u)dx\mathcal{F}=\int_{\Omega}f(\nabla u)\,\mathrm{d}x is a major task in the calculus of variations with a long standing tradition. Let us suppose that Ωn\Omega\subset\mathbb{R}^{n} is open and bounded and that uu is a weakly differentiable N\mathbb{R}^{N}-valued map. The growth condition on the integrand fC(N×n)f\in\mathrm{C}(\mathbb{R}^{N\times n}) determines the functional analytic environment in which we analyse \mathcal{F}. A standard growth assumption, which has been studied intensively in the field, constitutes the following: There exist p[1,)p\in[1,\infty) and a constant L>0L>0 such that for all zN×nz\in\mathbb{R}^{N\times n} we have

(1.1) |f(z)|L(1+|z|p).|f(z)|\leq L\left(1+|z|^{p}\right).

The existence of minimisers within a given class of functions (or maps) is a fundamental question. Concretely, we want to minimise the functional \mathcal{F} in g+W01,p(Ω;N)g+\mathrm{W}^{1,p}_{0}(\Omega;\mathbb{R}^{N}) for a prescribed Dirichlet boundary datum gW1,p(Ω;N)g\in\mathrm{W}^{1,p}(\Omega;\mathbb{R}^{N}). In the super-linear growth case p>1p>1, this task can immediately be tackled by means of the direct method, a lower semi-continuity method dating back to Tonelli. Consequently, the question arises under which assumptions on ff is the functional \mathcal{F} sequential weakly lower semi-continuous in W1,p(Ω;N)\mathrm{W}^{1,p}(\Omega;\mathbb{R}^{N}). Towards this question, convexity certainly suffices, but in the vectorial case (N>1)(N>1) it is seen to not be a necessary condition. Ball and Murat [8] have shown that Morrey’s notion of quasi-convexity [43], i. e., for all zN×nz\in\mathbb{R}^{N\times n} and all ζCc1(𝔹;N)\zeta\in\mathrm{C}^{1}_{\mathrm{c}}(\mathbb{B};\mathbb{R}^{N}) we have

f(z)𝔹f(z+ζ)dx,f\left(z\right)\leq\fint_{\mathbb{B}}f\left(z+\nabla\zeta\right)\,\mathrm{d}x,

turns out to be necessary. In fact, for unsigned integrands ff, quasi-convexity is also a sufficient condition, as [1] have shown in [1].

Once having addressed the issue of existence of minimisers, we would like to know what further information about a minimiser we can extract. This question dates back to David Hilbert [35] and is today known by the name of regularity theory in the calculus of variations. There are many different notions of regularity and in our setting, we are interested whether a minimiser is (locally) of the class C1,α\mathrm{C}^{1,\alpha}. In the scalar case (N=1)(N=1) the notions of convexity and quasi-convexity coincide and the regularity theory, at least in the quadratic growth case, reduces to the regularity of solutions of elliptic equations established by De Giorgi [18], Nash [45] and Moser [44]. However, in the vectorial case, the regularity of minimisers can no longer be extracted from the Euler-Lagrange equation only, because, as various counterexamples show [20, 46, 29], there is no such theory for elliptic systems in general. Furthermore, full Hölder regularity can no longer be expected since minimisers may be unbounded within a small set. Adapting ideas from geometric measure theory developed by Almgren [4] and De Giorgi [19], Evans established in the non-parametric setting a fundamental partial regularity result assuming a stronger notion of quasi-convexity [27]. This means that a minimiser enjoys Hölder regularity outside a small set. We stress that partial regularity is a feature of the vectorial case. After the quadratic case, the super-quadratic case (p2p\geq 2) was established by Acerbi and Fusco in [2]. Later on, the sub-quadratic case (1<p21<p\leq 2) was resolved partially by Carozza and Passarelli di Napoli in [14] and then fully by [13] in [13]. Some time later, even the Orlicz growth case has been resolved by [22] in [22]. An overview of related results can be found in [41, 42, 9, 30]. The case of linear growth for quasi-convex integrands, however, had remained an open problem, since the classical methods were bound to fail due to the lack of weak compactness. Only in the recent years, partial local Hölder regularity of the (distributional) gradient of (local) BV\mathrm{BV}-minimisers in the quasi-convex setting has been established by [33] in [33].

Let us try to roughly describe the underlying ideas on how to obtain partial Hölder regularity of the weak gradient of a minimiser uu of \mathcal{F}: The main objective is to prove a decay estimate for the excess of uu. The excess is a quantity that, similar to Campanato’s semi-norm, measures by means of integrals the rate of oscillation of u\nabla u. The goal is to show that if the excess is small enough, it decays with any rate α(0,1)\alpha\in(0,1). By means of a Caccioppoli Inequality, one passes from the excess, which depends on the gradient, to a quantity depending merely on uu. The Caccioppoli Inequality, in turn, builds on the combination of minimality and a stronger notion of quasi-convexity by means of Widman’s hole-filling trick. On the level of order 0, the strategey then is to approximate the minimiser by a \mathcal{B}-harmonic map hh, where \mathcal{B} denotes a strongly Legendre-Hadamard elliptic bilinear form on N×n\mathbb{R}^{N\times n}. The map hh has a good decay since it solves a homogeneous elliptic system. The difficulty is to construct hh in such a way that uhu-h has good decay as well. Classically, the construction of hh follows an indirect approach utilising compactness, for example, of the embedding W1,2(𝔹)L2(𝔹)\mathrm{W}^{1,2}(\mathbb{B})\to\mathrm{L}^{2}(\mathbb{B}). This kind of approximation goes by the name of \mathcal{B}-Harmonic Approximation Lemma [19, 23, 25].

Many generalisations of the functional \mathcal{F} have been studied. Here, we are going to replace the full gradient with a first order homogeneous differential operator 𝔸\mathbb{A}. The two most prominent examples are the symmteric gradient ε\varepsilon and the trace-free symmetric gradient ε~\tilde{\varepsilon}. Let VV and WW be two real and finite dimensional Hilbert spaces. We are going to consider differential operators of the form

𝔸=α=1n𝔸αα,𝔸α(V;W).\mathbb{A}=\sum\limits_{\alpha=1}^{n}\mathbb{A}_{\alpha}\partial_{\alpha},\ \mathbb{A}_{\alpha}\in\mathcal{L}(V;W).

For ξ𝕂n\xi\in\mathbb{K}^{n}, 𝕂=,\mathbb{K}=\mathbb{R},\mathbb{C}, the linear map 𝔸[ξ]v=α=1nξα𝔸αv\mathbb{A}[\xi]v=\sum_{\alpha=1}^{n}\xi_{\alpha}\mathbb{A}_{\alpha}v, modulo a factor of i-i, is called the symbol map associated to the differential operator 𝔸\mathbb{A}. We say that 𝔸\mathbb{A} is 𝕂\mathbb{K}-elliptic if the symbol map 𝔸[ξ]\mathbb{A}[\xi] is one-to-one for all ξ𝕂n{0}\xi\in\mathbb{K}^{n}\setminus\{0\} ([50, 49, 36]). The notion of \mathbb{R}-ellipticity has been characterised by means of Fourier multipliers [40] and singular integrals [11] that the Korn-type Inequality

p(1,)C>0ζCc(n;V):ζLpC𝔸ζLp\forall p\in(1,\infty)\exists C>0\forall\zeta\in\mathrm{C}^{\infty}_{\mathrm{c}}(\mathbb{R}^{n};V)\colon\ \left\lVert\nabla\zeta\right\rVert_{\mathrm{L}^{p}}\leq C\left\lVert\mathbb{A}\zeta\right\rVert_{\mathrm{L}^{p}}

is satisfied by the differential operator 𝔸\mathbb{A}. In the super-linear growth case, Conti and Gmeineder showed that this allows to reduce the question of partial Hölder regularity of a local minimiser of a functional of the form

[u;Ω]=Ωf(𝔸u)dx,\mathcal{F}[u;\Omega]=\int_{\Omega}f\left(\mathbb{A}u\right)\,\mathrm{d}x,

where fC(W)f\in\mathrm{C}(W) is of pp-growth (p>1p>1), to the full gradient case [17]. Ornstein’s Non-Inequality [47, Theorem 1], [16, 37], stating that there is no non-trivial Korn Inequality in the L1\mathrm{L}^{1}-setting, implies that such a reduction is impossible in the linear growth regime. Consequently, it had constituted a highly non-trivial task to adapt the full gradient case [33] to the symmetric gradient case [32] in the linear growth regime.

1.2. Partial Hölder Regularity in the Linear Growth Regime

In the convex case, partial regularity for linear growth functionals has been known in the convex context following the work of [6] [6] (also see [48, 31] for variations of this theme). However, the methods employed therein are confined to the convex case. In the linear growth context, the key difficulty to overcome is the lack of weak compactness. This concerns the existence of minimisers as much as their regularity theory. In particular, this excludes indirect methods like the by now classical \mathcal{B}-Harmonic Approximation Lemma [23, 25, 24]. A direct approach was needed to construct a \mathcal{B}-harmonic approximation. [33] solved this problem by showing that the traces of BV\mathrm{BV}-maps on spheres of radius RR enjoy for 1\mathcal{L}^{1}-almost all sufficiently small radii RR more regularity than the default L1\mathrm{L}^{1}-regularity. This is called a Fubini-type property of BV\mathrm{BV}-maps and it was in effect the key point to construct a \mathcal{B}-harmonic approximation by solving the elliptic system

{div(h)=0inBR(x0)h=uonBR(x0).\begin{cases}\begin{aligned} -\operatorname{div}\left(\mathcal{B}\nabla h\right)&=0&\quad&\text{in}\ B_{R}(x_{0})\\ h&=u&\quad&\text{on}\ \partial B_{R}(x_{0})\end{aligned}\end{cases}.

We note that the solution operator

(C0L1)(BR(x0);N)uhW1,1(BR(x0);N)(\mathrm{C}^{0}\cap\mathrm{L}^{1})\left(\partial B_{R}(x_{0});\mathbb{R}^{N}\right)\ni u\mapsto h\in\mathrm{W}^{1,1}\left(B_{R}(x_{0});\mathbb{R}^{N}\right)

associated to this system cannot be bounded as an operator from L1\mathrm{L}^{1} to W1,1\mathrm{W}^{1,1}. In other words, without more regularity of TrBR(x0)(u)\operatorname{Tr}_{B_{R}(x_{0})}(u) we lack tools to precisely measure how close the \mathcal{B}-harmonic map hh is to uu. This is why the Fubini-type property of BV\mathrm{BV}-maps is essential for a direct approach in the linear growth regime.

[32] was able to adapt the ideas used in [33] to the scenario where the full gradient is replaced by the symmetric gradient [32]. The main difficulties were to prove a Fubini-type property and, building on the latter, to prove precise estimates for uhu-h, where hh denotes a suitable \mathcal{B}-harmonic approximation of uu.

1.3. The Main Theorem

Our scope is to show that the method for the symmetric gradient case extends to an entire class of first order homogeneous differential operators with constant coefficients, namely the class of \mathbb{C}-elliptic operators.

In line with the full [33] and symmetric gradient case [32], we will work from now on under the following assumptions:

  • (H0)

    The differential operator 𝔸\mathbb{A} is \mathbb{C}-elliptic.

  • (H1)

    The integrand fCloc2,1(W)f\in\mathrm{C}^{2,1}_{\mathrm{loc}}(W) is of linear growth.

  • (H2)

    The integrand ff is strongly V1V_{1}-𝔸\mathbb{A}-quasi-convex, where V1V_{1} denotes the reference integrand to be defined in the the upcoming section on preliminaries: There exists ν>0\nu>0 such that F=fνV1||F=f-\nu V_{1}\circ|\cdot| is 𝔸\mathbb{A}-quasi-convex, i. e., for all wWw\in W and all ζCc(𝔹)\zeta\in\mathrm{C}^{\infty}_{\mathrm{c}}(\mathbb{B}) we have

    F(w)𝔹F(w+𝔸ζ)dx.F(w)\leq\fint_{\mathbb{B}}F\left(w+\mathbb{A}\zeta\right)\,\mathrm{d}x.

For ωn\omega\subset\mathbb{R}^{n} open and bounded we associate to the integrand ff the functional [u;ω]=ωF(𝔸u(x))dx\mathcal{F}[u;\omega]=\int_{\omega}F(\mathbb{A}u(x))\,\mathrm{d}x, where uW𝔸,1(ω)u\in\mathrm{W}^{\mathbb{A},1}(\omega).

We recall that to any \mathbb{R}-elliptic potential 𝔸\mathbb{A} exists an annihilator, see [51], 𝒜=|α|=k𝒜αα\mathcal{A}=\sum_{|\alpha|=k}\mathcal{A}_{\alpha}\partial_{\alpha} with 𝒜α(W;V)\mathcal{A}_{\alpha}\in\mathcal{L}(W;V) such that the symbol complex

V𝔸[ξ]W𝒜[ξ]VV\xrightarrow{\mathbb{A}[\xi]}W\xrightarrow{\mathcal{A}[\xi]}V

is exact for all ξn{0}\xi\in\mathbb{R}^{n}\setminus\{0\}. This shows that the notion of 𝔸\mathbb{A}-quasi-convexity is equivalent to the more widely known notion of 𝒜\mathcal{A}-quasi-convexity which is strongly linked to weak sequential lower semi-continuity of the functional \mathcal{F}, which Fonseca and Müller showed in [28].

We fix Ωn\Omega\subset\mathbb{R}^{n} an open and bounded Lipschitz domain and gW𝔸,1(Ω)g\in\mathrm{W}^{\mathbb{A},1}(\Omega) a prescribed boundary datum. Due to the lack of weak compactness, analogously to the full gradient case, the task to minimise \mathcal{F} within a given Dirichlet class g+W0𝔸,1(Ω)g+\mathrm{W}^{\mathbb{A},1}_{0}(\Omega) cannot be tackled by plainly applying the direct method. Hence, we pass from the Sobolev-type space W𝔸,1\mathrm{W}^{\mathbb{A},1} to the BV\mathrm{BV}-type space BV𝔸\mathrm{BV}^{\mathbb{A}}, hoping for better compactness with respect to the weak-topology on the latter space. Already in the full gradient case, this forces us to somehow relax our functional \mathcal{F} to this larger space BV𝔸\mathrm{BV}^{\mathbb{A}}. Without the assumption of \mathbb{C}-ellipticity, this relaxation procedure cannot be implemented analogously: We recall that in the full gradient case, Alberti’s rank-one result [3] for BV\mathrm{BV}-maps has proven to be essential in order to obtain an integral representation of the Lebesgue-Serrin extension [5]. Using that quasi-convexity implies rank-one convexity, Alberti’s result ensures that the strong recession function of a quasi-convex integrand with linear growth is well-defined on the rank-one cone. The requisite result paralleling Alberti’s result in the \mathbb{R}-elliptic case has been established by [21] in [21]. Weak-compactness of closed, norm bounded sets in BV𝔸\mathrm{BV}^{\mathbb{A}} as well as the existence of a strictly continuous and linear trace operator TrΩ:BV𝔸(Ω)L1(Ω)\operatorname{Tr}_{\Omega}\colon\mathrm{BV}^{\mathbb{A}}(\Omega)\to\mathrm{L}^{1}(\partial\Omega) for open and bounded Lipschitz domains are two features exclusive to the \mathbb{C}-elliptic case [10, Theorem 1.1], [34, Theorem 1.1]. Since the trace-operator on BV𝔸\mathrm{BV}^{\mathbb{A}} is discontinuous with respect to weak-convergence, the boundary condition is no longer reflected by the space BV𝔸\mathrm{BV}^{\mathbb{A}} but rather by the relaxed functional itself. The boundary condition then is encoded by a so called penalty term, which already pops up in the full gradient case [38]:

Pf,Ω,g[u]=Ωf(νΩ𝔸TrΩ(ug))dn1,\textnormal{P}_{f,\Omega,g}[u]=\int_{\partial\Omega}f^{\infty}\big{(}\nu_{\partial\Omega}\otimes_{\mathbb{A}}\operatorname{Tr}_{\Omega}(u-g)\big{)}\,\mathrm{d}\mathcal{H}^{n-1},

where f(w)=lim suptf(tw)tf^{\infty}(w)=\limsup_{t\to\infty}\frac{f(tw)}{t} denotes the strong recession function and ξ𝔸v=α=1nξα𝔸αv\xi\otimes_{\mathbb{A}}v=\sum_{\alpha=1}^{n}\xi_{\alpha}\mathbb{A}_{\alpha}v for ξn\xi\in\mathbb{R}^{n} and vVv\in V. We stress that it is necessary to assume that 𝔸\mathbb{A} is \mathbb{C}-elliptic in order for this integral expression to be well-defined. The relaxed functional then takes the form ([10, Section 5], [7])

¯g[u;Ω]=Ωf(𝔸u)+Pf,Ω,g[u].\overline{\mathcal{F}}_{g}[u;\Omega]=\int_{\Omega}f(\mathbb{A}u)+\textnormal{P}_{f,\Omega,g}[u].

We note that (H2) implies that there exist bb\in\mathbb{R} and c>0c>0 such that for all ζg+W0𝔸,1(Ω)\zeta\in g+\mathrm{W}^{\mathbb{A},1}_{0}(\Omega) we have [ζ;Ω]cV1(Ω|𝔸ζ|dx)+b\mathcal{F}[\zeta;\Omega]\geq cV_{1}(\fint_{\Omega}|\mathbb{A}\zeta|\,\mathrm{d}x)+b. This can be inferred from an extension and gluing argument similar to the argument in [15, 218–219]. Identifying ¯g[u;Ω]\overline{\mathcal{F}}_{g}[u;\Omega] as the Lebesgue-Serrin extension [10, Section 5]

¯g[u;Ω]=inf{lim infj[uj;Ω]:(uj)g+W0𝔸,1(Ω),ujBV𝔸u}\overline{\mathcal{F}}_{g}[u;\Omega]=\inf\Big{\{}\liminf_{j\to\infty}\mathcal{F}[u_{j};\Omega]\colon\ (u_{j})\subset g+\mathrm{W}^{\mathbb{A},1}_{0}(\Omega),\ u_{j}\stackrel{{\scriptstyle\mathrm{BV}^{\mathbb{A}}}}{{\rightharpoonup^{*}}}u\Big{\}}

yields coercivity of the relaxation. Hence, under our assumption, generalised minimisers subject to a given Dirichlet boundary condition exist by means of the direct method. Since we are only striving for a local regularity result, it is natural to consider the class of local generalised minimisers:

Definition.

We call a BVloc𝔸(Ω)\mathrm{BV}^{\mathbb{A}}_{\mathrm{loc}}(\Omega)-map uu local generalised minimiser of \mathcal{F} if for any ωΩ\omega\Subset\Omega open and bounded Lipschitz domain and all ζBV𝔸(ω)\zeta\in\mathrm{BV}^{\mathbb{A}}(\omega) we have

¯u[u;ω]¯u[ζ;ω].\overline{\mathcal{F}}_{u}[u;\omega]\leq\overline{\mathcal{F}}_{u}[\zeta;\omega].

At this stage, we are ready to formulate the main theorem:

Theorem 1.1.

Let us assume that (H0), (H1), and (H2) hold. Furthermore, let uBVloc𝔸(Ω)u\in\mathrm{BV}^{\mathbb{A}}_{\mathrm{loc}}(\Omega) be a local generalised minimiser of the to ff associated functional \mathcal{F}. Let α(0,1)\alpha\in(0,1), let M>0M>0 and let B=Br(x0)ΩB=B_{r}(x_{0})\Subset\Omega be a ball. Then there exists ε>0\varepsilon>0 depending on M,α,F,n,dV,dWM,\alpha,F,n,d_{V},d_{W} and Lν\frac{L}{\nu} such that whenever we have

|(𝔸u)B|M,andBV1(𝔸u(𝔸u)B)<ε,|(\mathbb{A}u)_{B}|\leq M,\ \textnormal{and}\ \fint_{B}V_{1}\left(\mathbb{A}u-(\mathbb{A}u)_{B}\right)<\varepsilon,

then u|Bu\bigr{\rvert}_{B} belongs to the class C1,α(B;V)\mathrm{C}^{1,\alpha}(B;V). In particular, the singular set Σu\Sigma_{u} defined by

{xΩ:lim supr0|(𝔸u)Br(x)|=+}{xΩ:lim infr0BV1(𝔸u(𝔸u)B)>0}\left\{x\in\Omega\colon\limsup_{r\to 0}|(\mathbb{A}u)_{B_{r}(x)}|=+\infty\right\}\bigcup\left\{x\in\Omega\colon\liminf_{r\to 0}\fint_{B}V_{1}\left(\mathbb{A}u-(\mathbb{A}u)_{B}\right)>0\right\}

is a relatively closed Lebesgue-null-set and we have uCloc1,α(ΩΣu;V)u\in\mathrm{C}^{1,\alpha}_{\mathrm{loc}}(\Omega\setminus\Sigma_{u};V) for all α(0,1)\alpha\in(0,1).

We wish to point out that for the present paper, the assumption of \mathbb{C}-ellipticity is crucial and visible on several stages (so e. g. in the very definition of the functionals where boundary traces come into play); the elliptic case, however, seems to require refined methods.

2. Preliminaries

2.1. Notation

For a finite dimensional real vector space ZZ we use the shorthand notation dZ=dimZd_{Z}=\dim_{\mathbb{R}}Z. Furthermore, we will suppress the target vector space when dealing with different function spaces. For example, if u:ΩZu\colon\Omega\to Z is a L1\mathrm{L}^{1}-map, we simply write uL1(Ω)u\in\mathrm{L}^{1}(\Omega) instead of uL1(Ω;Z)u\in\mathrm{L}^{1}(\Omega;Z). Within the context, it will be clear which target vector space we are referring to. Furthermore, by |||\cdot| we will denote any norm of a finite dimensional, normed vector space such as (V;W)\mathcal{L}(V;W), VV or (V×V;W)\mathcal{L}(V\times V;W). Since all norms of a finite dimensional normed vector space are equivalent, this is an non-problematic convention. Throughout, we fix an orthonormal basis (vj,,vN)(v_{j},...,v_{N}) of VV, i. e., N=dVN=d_{V}. Furthermore,

(ejk)j=1,k=1j=N,k=n=δjk\left(e_{jk}\right)_{j=1,k=1}^{j=N,k=n}=\delta_{jk}

denotes the standard basis of N×n\mathbb{R}^{N\times n}.

Integration with respect to the (n1)(n-1)-dimensional Hausdorff-measure n1\mathcal{H}^{n-1} will be denoted by dσx\,\mathrm{d}\sigma_{x}, where xx is integration variable.

As usual, Br(x0)nB_{r}(x_{0})\subset\mathbb{R}^{n} denotes the open ball with centre x0x_{0} and radius r>0r>0. Often we will suppress the centre of the ball if it is clear within the context and we will simply write BrB_{r}. Furthermore, we put 𝔹=B1(0)\mathbb{B}=B_{1}(0) and 𝕊=𝔹\mathbb{S}=\partial\mathbb{B}.

By (Ω;Z)\mathcal{M}(\Omega;Z) we denote the space of ZZ-valued finite Radon measures on Ω\Omega. For μ(Ω)\mu\in\mathcal{M}(\Omega) and open, bounded subsets ωΩ\omega\subset\Omega, the total variation-measure of μ\mu will be denoted by |μ||\mu| and the average of μ\mu on ω\omega with respect to the Lebesgue measure will be written as (μ)ω:-μ(ω)n(ω).(\mu)_{\omega}\coloneq\frac{\mu(\omega)}{\mathcal{L}^{n}(\omega)}.

Let V1(t)=1+t21V_{1}(t)=\sqrt{1+t^{2}}-1 denote the reference integrand. We will use the shorthand notation V1(z)=V1(|z|)V_{1}(z)=V_{1}(|z|).

For a C1\mathrm{C}^{1}-function G:ZG\colon Z\to\mathbb{R} and z0Zz_{0}\in Z we denote by

(G)z0:Z,zG(z0+z)(G(z0)+DG(z0)[z])(G)_{z_{0}}\colon Z\to\mathbb{R},\ z\mapsto G(z_{0}+z)-\Big{(}G(z_{0})+DG(z_{0})[z]\Big{)}

the linearisation of GG at z0z_{0}.

For kk\in\mathbb{N} we denote by 𝒫k(Z)\mathcal{P}_{k}(Z) the vector space of all polynomials

p[X1,,Xn]Zp\in\mathbb{R}[X_{1},...,X_{n}]\otimes_{\mathbb{R}}Z

of degree at most kk.

By CC we will denote a generic constant which may vary from line to line. Since it is very important throughout on which parameters a constant depends on, we will write for example C(M,p)C(M,p) if the constant depends on MM and pp.

For ξn\xi\in\mathbb{R}^{n} and vVv\in V we put ξ𝔸v=α=1nξα𝔸αv\xi\otimes_{\mathbb{A}}v=\sum_{\alpha=1}^{n}\xi_{\alpha}\mathbb{A}_{\alpha}v. Furthermore, we denote by

(𝔸):-span{ξ𝔸v:ξn,vV}\mathcal{R}(\mathbb{A})\coloneq\mathrm{span}\{\xi\otimes_{\mathbb{A}}v\colon\xi\in\mathbb{R}^{n},v\in V\}

the effective range of 𝔸\mathbb{A} and we call

𝒩(𝔸):-{u𝒟:𝔸u=0}\mathcal{N}(\mathbb{A})\coloneq\{u\in\mathcal{D}^{*}\colon\mathbb{A}u=0\}

the null-space of 𝔸\mathbb{A}. The formally adjoint operator 𝔸\mathbb{A}^{*} is here defined by the formula

𝔸=α=1n𝔸αα.\mathbb{A}^{*}=-\sum\limits_{\alpha=1}^{n}\mathbb{A}_{\alpha}^{*}\partial_{\alpha}.

Let Ωn\Omega\subset\mathbb{R}^{n} be open and let 𝕄n\mathbb{M}\subset\mathbb{R}^{n} be an embedded (n1)(n-1)-dimensional C1\mathrm{C}^{1}-submanifold of n\mathbb{R}^{n}. For α(0,1)\alpha\in(0,1) and p[1,)p\in[1,\infty), we recall the definition of the fractional Sobolev space (semi)-norms on Ω\Omega and 𝕄\mathbb{M}, respectively:

  • [u]Wα,p(Ω)=(Ω2|u(x)u(y)|p|xy|n+αpdxdy)1p[u]_{\mathrm{W}^{\alpha,p}(\Omega)}=\Big{(}\int_{\Omega^{2}}\frac{|u(x)-u(y)|^{p}}{|x-y|^{n+\alpha p}}\,\mathrm{d}x\,\mathrm{d}y\Big{)}^{\frac{1}{p}}, uWα,p(Ω)=uLp(Ω)+[u]Wα,p(Ω)\left\lVert u\right\rVert_{\mathrm{W}^{\alpha,p}(\Omega)}=\left\lVert u\right\rVert_{\mathrm{L}^{p}(\Omega)}+[u]_{\mathrm{W}^{\alpha,p}(\Omega)},

  • [u]Wα,p(𝕄)=(𝕄2|u(x)u(y)|p|xy|n1+αpdσxdσy)1p[u]_{\mathrm{W}^{\alpha,p}(\mathbb{M})}=\Big{(}\int_{\mathbb{M}^{2}}\frac{|u(x)-u(y)|^{p}}{|x-y|^{n-1+\alpha p}}\,\mathrm{d}\sigma_{x}\,\mathrm{d}\sigma_{y}\Big{)}^{\frac{1}{p}}, uWα,p(𝕄)=uLp(𝕄)+[u]Wα,p(𝕄)\left\lVert u\right\rVert_{\mathrm{W}^{\alpha,p}(\mathbb{M})}=\left\lVert u\right\rVert_{\mathrm{L}^{p}(\mathbb{M})}+[u]_{\mathrm{W}^{\alpha,p}(\mathbb{M})}.

2.2. Space of maps of bounded 𝔸\mathbb{A}-variation

We are going to collect prerequisites on 𝔸\mathbb{A}-weakly differentiable maps. In the spirit of [10], we define Sobolev- and BV\mathrm{BV}-type spaces as follows:

Definition 2.1.

Let Ωn\Omega\subset\mathbb{R}^{n} be open and let p[1,]p\in[1,\infty]. We define:

  • W𝔸,p(Ω){uLp(Ω;V):𝔸uLp(Ω;W)},\mathrm{W}^{\mathbb{A},p}(\Omega)\coloneqq\{u\in\mathrm{L}^{p}(\Omega;V)\colon\ \mathbb{A}u\in\mathrm{L}^{p}(\Omega;W)\}, and

  • BV𝔸(Ω){uLp(Ω;V):𝔸u(Ω;W)}.\mathrm{BV}^{\mathbb{A}}(\Omega)\coloneqq\{u\in\mathrm{L}^{p}(\Omega;V)\colon\ \mathbb{A}u\in\mathcal{M}(\Omega;W)\}.

These spaces can be equipped with the obvious norms making them Banach spaces. Also the spaces W0𝔸,p(Ω)\mathrm{W}^{\mathbb{A},p}_{0}(\Omega) are as usual defined as the closure of Cc(Ω;V)\mathrm{C}^{\infty}_{\mathrm{c}}(\Omega;V) with respect to the according norm.

Let uBVloc𝔸(Ω)u\in\mathrm{BV}^{\mathbb{A}}_{\mathrm{loc}}(\Omega). Then we consider the Radon-Nikodým decomposition of 𝔸u\mathbb{A}u with respect to the Lebesgue measure 𝔸u=𝔸au+𝔸su\mathbb{A}u=\mathbb{A}^{a}u+\mathbb{A}^{s}u, where 𝔸au\mathbb{A}^{a}u denotes the absolutely continuous part and 𝔸su\mathbb{A}^{s}u the singular part. Next, we recall different notions of convergence in BV𝔸\mathrm{BV}^{\mathbb{A}}:

Definition 2.2.

Let uBV𝔸(Ω)u\in\mathrm{BV}^{\mathbb{A}}(\Omega) and (uj)BV𝔸(Ω)(u_{j})\subset\mathrm{BV}^{\mathbb{A}}(\Omega). Then uju_{j} converges to uu in the

  1. (i)

    𝔸\mathbb{A}-weak*-sense (ujuu_{j}\overset{\ast}{\rightharpoonup}u) if ujuu_{j}\to u strongly in L1(Ω)L^{1}(\Omega) and 𝔸uj𝔸u\mathbb{A}u_{j}\overset{\ast}{\rightharpoonup}\mathbb{A}u in the weak*-sense of WW-valued Radon measures on Ω\Omega.

  2. (ii)

    𝔸\mathbb{A}-strict sense (uj𝑠uu_{j}\overset{s}{\to}u) if ujuu_{j}\to u strongly in L1(Ω)L^{1}(\Omega) and |𝔸uj|(Ω)|𝔸u|(Ω)\lvert\mathbb{A}u_{j}\rvert(\Omega)\to\lvert\mathbb{A}u\rvert(\Omega).

  3. (iii)

    𝔸\mathbb{A}-area-strict sense (ujuu_{j}\overset{\langle\cdot\rangle}{\to}u) if ujuu_{j}\to u strongly in L1(Ω)L^{1}(\Omega) and

    Ω1+|d𝔸aujdn|2dx+|𝔸suj|(Ω)Ω1+|d𝔸audn|2dx+|𝔸su|(Ω)\int_{\Omega}\sqrt{1+\left|\frac{\,\mathrm{d}\mathbb{A}^{a}u_{j}}{\,\mathrm{d}\mathcal{L}^{n}}\right|^{2}}\,\mathrm{d}x+\lvert\mathbb{A}^{s}u_{j}\rvert(\Omega)\to\int_{\Omega}\sqrt{1+\left|\frac{\,\mathrm{d}\mathbb{A}^{a}u}{\,\mathrm{d}\mathcal{L}^{n}}\right|^{2}}\,\mathrm{d}x+\lvert\mathbb{A}^{s}u\rvert(\Omega)
Lemma 2.3.

[10, Theorem 2.8, Lemma 4.15] Let Ωn\Omega\subset\mathbb{R}^{n} open. Then (CBV𝔸)(Ω)(\mathrm{C}^{\infty}\cap\mathrm{BV}^{\mathbb{A}})(\Omega) is dense in BV𝔸(Ω)\mathrm{BV}^{\mathbb{A}}(\Omega) with respect to the strict and area-strict topologies. If Ω\Omega is additionally a bounded Lipschitz domain, then C(Ω¯)\mathrm{C}^{\infty}(\overline{\Omega}) is dense in BV𝔸(Ω)\mathrm{BV}^{\mathbb{A}}(\Omega) with respect to the strict and area-strict topologies. Let u0W𝔸,1(Ω)u_{0}\in W^{\mathbb{A},1}(\Omega). For each uBV𝔸(Ω)u\in\mathrm{BV}^{\mathbb{A}}(\Omega) there exists a sequence (uj)u0+Cc(Ω)(u_{j})\subset u_{0}+C_{c}^{\infty}(\Omega) such that ujuL1(Ω)0\lVert u_{j}-u\rVert_{L^{1}(\Omega)}\to 0 and

Ω1+|d𝔸aujdn|2\displaystyle\int_{\Omega}\sqrt{1+\left|\frac{\,\mathrm{d}\mathbb{A}^{a}u_{j}}{\,\mathrm{d}\mathcal{L}^{n}}\right|^{2}} Ω1+|d𝔸audn|2dx+|𝔸su|(Ω)\displaystyle\to\int_{\Omega}\sqrt{1+\left|\frac{\,\mathrm{d}\mathbb{A}^{a}u}{\,\mathrm{d}\mathcal{L}^{n}}\right|^{2}}\,\mathrm{d}x+\lvert\mathbb{A}^{s}u\rvert(\Omega)
+Ω|(Tr(u)Tr(u0))𝔸νΩ|dn1for j.\displaystyle\phantom{\to}+\int_{\partial\Omega}\lvert(\operatorname{Tr}(u)-\operatorname{Tr}(u_{0}))\otimes_{\mathbb{A}}\nu_{\partial\Omega}\rvert\,\mathrm{d}\mathcal{H}^{n-1}\qquad\text{for $j\to\infty$.}

The proof of the last assertion is analogous to the BD\mathrm{BD}-case [32].

Lemma 2.4.

[10, Theorem 3.2] Let BnB\subset\mathbb{R}^{n} be an open ball of radius r>0r>0 and let ΠB\Pi_{B} denote the L2(B)\mathrm{L}^{2}(B)-projection onto 𝒩(𝔸)\mathcal{N}(\mathbb{A}). Then there exists a constant C>0C>0 such that for all uBV𝔸(B)u\in\mathrm{BV}^{\mathbb{A}}(B) we have

uΠBuL1(B)Cr|𝔸u|(B).\left\lVert u-\Pi_{B}u\right\rVert_{\mathrm{L}^{1}(B)}\leq Cr|\mathbb{A}u|(B).
Lemma 2.5.

[10, Theorem 1.2] Let Ωn\Omega\subset\mathbb{R}^{n} be an open and bounded Lipschitz domain. Then there exists a linear and strictly continuous operator TrΩ:BV𝔸(Ω)L1(Ω)\operatorname{Tr}_{\Omega}\colon\mathrm{BV}^{\mathbb{A}}(\Omega)\to\mathrm{L}^{1}(\partial\Omega) such that for all uC1(Ω¯)u\in\mathrm{C}^{1}(\overline{\Omega}) we have TrΩu=u|Ω.\operatorname{Tr}_{\Omega}u=u\bigr{\rvert}_{\partial\Omega}. For an open and bounded Lipschitz subset ΩΩ\Omega^{\prime}\Subset\Omega, we consider so called interior and exterior traces of uu denoted by

TrΩ(u):-TrΩ(u|Ω)andTrΩ+(u):-TrΩΩ(u|ΩΩ).\operatorname{Tr}^{-}_{\Omega^{\prime}}(u)\coloneq\operatorname{Tr}_{\Omega^{\prime}}\left(u\bigr{\rvert}_{\Omega^{\prime}}\right)\ \textnormal{and}\ \operatorname{Tr}^{+}_{\Omega^{\prime}}(u)\coloneq\operatorname{Tr}_{\Omega\setminus\Omega^{\prime}}\left(u\bigr{\rvert}_{\Omega\setminus\Omega^{\prime}}\right).

One can explicitly compute

(2.1) limr0B±(x,r)|u(y)TrB±(u)(x)|dy=0\lim_{r\searrow 0}\fint_{B^{\pm}(x,r)}\lvert u(y)-\operatorname{Tr}^{\pm}_{\partial B}(u)(x)\rvert\,\mathrm{d}y=0

for n1\mathcal{H}^{n-1}-a.e. xBx\in\partial B, where B±(x,r):-{yBr(x)yx,ν(x)0}B^{\pm}(x,r)\coloneq\{y\in B_{r}(x)\mid\langle y-x,\nu(x)\rangle\gtrless 0\}. Here, ν(x)\nu(x) designates the outer unit normal vector to the sphere B\partial B at point xx.

Proposition 2.6.

[10, Proposition 5.1] Let Ωn\Omega\subset\mathbb{R}^{n} be open, bounded and let g:Wg\colon W\to\mathbb{R} be an 𝔸\mathbb{A}-quasi-convex integrand of linear growth. Then the functional

𝒢¯:BV𝔸(Ω),uΩg(𝔸u):-g(𝔸aun)dx+Ωg(d𝔸sud|𝔸su|)d|𝔸su|\overline{\mathcal{G}}\colon\mathrm{BV}^{\mathbb{A}}(\Omega)\to\mathbb{R},u\mapsto\int_{\Omega}g\left(\mathbb{A}u\right)\coloneq\int g\left(\frac{\mathbb{A}^{a}u}{\mathcal{L}^{n}}\right)\,\mathrm{d}x+\int_{\Omega}g^{\infty}\left(\frac{\,\mathrm{d}\mathbb{A}^{s}u}{\,\mathrm{d}|\mathbb{A}^{s}u|}\right)\,\mathrm{d}|\mathbb{A}^{s}u|

is 𝔸\mathbb{A}-area strictly continuous and sequentially lower semi-continuous with respect to weak-convergence.

We will use the shorthand notation Ωg(𝔸u)=Ωg(𝔸u)n(Ω)\fint_{\Omega}g(\mathbb{A}u)=\frac{\int_{\Omega}g(\mathbb{A}u)}{\mathcal{L}^{n}(\Omega)}.

Lemma 2.7.

Let n2n\geq 2, α(0,1)\alpha\in(0,1), let B2rnB_{2r}\subset\mathbb{R}^{n} be a ball of radius 2r>02r>0 and let p:-nn1+αp\coloneq\frac{n}{n-1+\alpha}. Then there exists a constant C>0C>0 independent of the radius rr such that for every ball BrnB_{r}\subset\mathbb{R}^{n} and every uBV𝔸(n)u\in\mathrm{BV}^{\mathbb{A}}(\mathbb{R}^{n}), there exists some b𝒩(𝔸)b\in\mathcal{N}(\mathbb{A}) with

(2.2) (BrBr|ub(x)ub(y)|p|xy|n+αpdxdy)1pCr1αB2r|𝔸u|,\left(\fint_{B_{r}}\int_{B_{r}}\frac{\lvert u_{b}(x)-u_{b}(y)\rvert^{p}}{\lvert x-y\rvert^{n+\alpha p}}\,\mathrm{d}x\,\mathrm{d}y\right)^{\frac{1}{p}}\leq Cr^{1-\alpha}\fint_{B_{2r}}\lvert\mathbb{A}u\rvert,

where ub:-ubu_{b}\coloneq u-b.

Proof.

Let u~:-uφ\widetilde{u}\coloneq u\varphi, where φCc(B2r;[0,1])\varphi\in C_{c}^{\infty}(B_{2r};[0,1]) is a bump function with 𝟏Brφ𝟏B2r\mathbf{1}_{B_{r}}\leq\varphi\leq\mathbf{1}_{B_{2r}} and |φ|C/r\lvert\nabla\varphi\rvert\leq C/r. Then u~BV𝔸(n)\widetilde{u}\in\mathrm{BV}^{\mathbb{A}}(\mathbb{R}^{n}) and

(2.3) 𝔸u~L1(B2r)C(r)uBV𝔸(B2r).\lVert\mathbb{A}\widetilde{u}\rVert_{L^{1}(B_{2r})}\leq C(r)\lVert u\rVert_{\mathrm{BV}^{\mathbb{A}}(B_{2r})}.

Noting that Tr(u~)=0\operatorname{Tr}(\widetilde{u})=0 on B2r\partial B_{2r}, there exists a sequence (u~j)Cc(B2r;N)(\widetilde{u}_{j})\subset C_{c}^{\infty}(B_{2r};\mathbb{R}^{N}) such that u~ju~\widetilde{u}_{j}\to\widetilde{u} strictly. Since 𝔸\mathbb{A} is \mathbb{C}-elliptic it is in particular \mathbb{R}-elliptic and canceling, see [34]. Applying [51, Proposition 8.11] we obtain

u~jWα,p(Br)u~jWα,p(n)C𝔸u~jL1(n)\lVert\widetilde{u}_{j}\rVert_{W^{\alpha,p}(B_{r})}\leq\lVert\widetilde{u}_{j}\rVert_{W^{\alpha,p}(\mathbb{R}^{n})}\leq C\lVert\mathbb{A}\widetilde{u}_{j}\rVert_{L^{1}(\mathbb{R}^{n})}

By passing to a subsequence we may assume that (u~j)(\widetilde{u}_{j}) converges n\mathcal{L}^{n}-almost everywhere. By Fatou’s Lemma and the strict convergence we obtain

(2.4) uWα,p(Br)C(r)uBV𝔸(B2r).\left\lVert u\right\rVert_{W^{\alpha,p}(B_{r})}\leq C(r)\left\lVert u\right\rVert_{\mathrm{BV}^{\mathbb{A}}(B_{2r})}.

We put b:-ΠBrub\coloneq\Pi_{B_{r}}u, (ub)r:-(ub)(rx)(u-b)_{r}\coloneq(u-b)(rx) for x𝔹x\in\mathbb{B} and eventually we obtain by Poincaré’s Inequality 2.4, scaling and applying 2.4 for r=1r=1:

(BrBr|ub(x)ub(y)|p|xy|n+αpdxdy)1pCrα(ub)rWα,p(B1)Cr1αB2r|𝔸u|.\left(\fint_{B_{r}}\int_{B_{r}}\frac{\lvert u_{b}(x)-u_{b}(y)\rvert^{p}}{\lvert x-y\rvert^{n+\alpha p}}\,\mathrm{d}x\,\mathrm{d}y\right)^{\frac{1}{p}}\leq Cr^{-\alpha}\left\lVert(u-b)_{r}\right\rVert_{W^{\alpha,p}(B_{1})}\leq Cr^{1-\alpha}\fint_{B_{2r}}|\mathbb{A}u|.

Lemma 2.8.

Let BΩ\mathrm{B}\Subset\Omega be a ball and uBVloc𝔸(Ω)u\in\mathrm{BV}^{\mathbb{A}}_{\mathrm{loc}}(\Omega). Then there exists a linear map a:nVa\colon\mathbb{R}^{n}\to V such that 𝔸a=(𝔸u)B\mathbb{A}a=(\mathbb{A}u)_{B}.

Proof.

Let ujC(B¯;V)u_{j}\in\mathrm{C}^{\infty}(\overline{B};V) such that uj𝑠uu_{j}\overset{s}{\to}u as jj\to\infty. Clearly, we have supj|(𝔸uj)B|<\sup_{j}|(\mathbb{A}u_{j})_{B}|<\infty and (𝔸uj)B(𝔸)(\mathbb{A}u_{j})_{B}\in\mathcal{R}(\mathbb{A}), since 𝔸uj(B)=α=1n𝔸αBαujdx\mathbb{A}u_{j}(B)=\sum_{\alpha=1}^{n}\mathbb{A}_{\alpha}\fint_{B}\partial_{\alpha}u_{j}\,\mathrm{d}x. After extraction of a non-relabelled sub-sequence we find w(𝔸)w\in\mathcal{R}(\mathbb{A}) such that (𝔸uj)Bw(\mathbb{A}u_{j})_{B}\to w in WW as jj\to\infty. We find vαVv_{\alpha}\in V such that w=α=1n𝔸αvαw=\sum_{\alpha=1}^{n}\mathbb{A}_{\alpha}v_{\alpha} and put a[x]=α=1nxαvαa[x]=\sum_{\alpha=1}^{n}x_{\alpha}v_{\alpha}. This yields the claim. ∎

2.3. Linearisation and the Reference Integrand

Lemma 2.9.

Let f:Wf\colon W\to\mathbb{R} satisfy (H1) and (H2), let m>0m>0. Then there exists a constant C(m)>0C(m)>0 such that for all w0Ww_{0}\in W with |w0|m|w_{0}|\leq m, all ξn\xi\in\mathbb{R}^{n} and all vVv\in V we have

(2.5) D2f(w0)[ξ𝔸v,ξ𝔸v]C(m)ν|ξ𝔸v|2,|D2(f)w0[w,]D(f)w0(w)|C(m)V1(w).\begin{split}D^{2}f(w_{0})[\xi\otimes_{\mathbb{A}}v,\xi\otimes_{\mathbb{A}}v]\geq\frac{C(m)}{\nu}|\xi\otimes_{\mathbb{A}}v|^{2},\\ |D^{2}(f)_{w_{0}}[w,\cdot]-D(f)_{w_{0}}(w)|\leq C(m)V_{1}(w).\end{split}

Also, it is worth noting that we have

(2.6) 0<inft(0,1)V1(t)t2supt(0,1)V1(t)t2<,0<\inf\limits_{t\in(0,1)}\frac{V_{1}(t)}{t^{2}}\leq\sup\limits_{t\in(0,1)}\frac{V_{1}(t)}{t^{2}}<\infty,
(2.7) V1(rt)rV1(t)forr(0,1),V1(rt)r2V1(t)forr(1,),V_{1}(rt)\leq rV_{1}(t)\ \textnormal{for}\ r\in(0,1),\quad V_{1}(rt)\leq r^{2}V_{1}(t)\ \textnormal{for}\ r\in(1,\infty),
(2.8) V1(|s|+|t|)4(V1(|s|)+|V1(t|)).V_{1}(|s|+|t|)\leq 4(V_{1}(|s|)+|V_{1}(t|)).

The proof is analogous to the proof of [33, Lemma 4.2] and [32, Lemma 5.1].

2.4. Caccioppoli Inequality of the second kind

The Caccioppoli Inequality is indispensable for our proof of partial regularity. However, it is line for line analogous to the full and symmetric the gradient case [33, Proposition 4.3], [32, Proposition 5.2], replacing \nabla or ε\varepsilon with 𝔸\mathbb{A}, respectively: by exploiting the strong V1V_{1}-𝔸\mathbb{A}-quasi-convexity and minimality of uu, we apply Widman’s hole-filling trick and iterate the resulting inequality.

Proposition 2.10.

We assume that (H0), (H1), and (H2) hold. Let uBVloc𝔸(Ω)u\in\mathrm{BV}^{\mathbb{A}}_{\mathrm{loc}}(\Omega) be a local generalised minimiser of the functional \mathcal{F} and let a:nNa\colon\mathbb{R}^{n}\to\mathbb{R}^{N} be an affine map with |𝔸a|m|\mathbb{A}a|\leq m for some m>0m>0. Then there exists a constant C=C(dV;dW,Lν)(1,)C=C(d_{V};d_{W},\frac{L}{\nu})\in(1,\infty) such that

Br/2(x0)V(𝔸(ua))CBr(x0)V1(uar)dx\int_{B_{r/2}(x_{0})}V(\mathbb{A}(u-a))\leq C\int_{B_{r}(x_{0})}V_{1}\left(\frac{u-a}{r}\right)\,\mathrm{d}x

for any ball Br(x0)ΩB_{r}(x_{0})\Subset\Omega.

2.5. The Ekeland Variational Principle

Lemma 2.11.

[26, Theorem 1.1] Let (X,d)(X,d) be a complete metric space and let 𝒢:X{+}\mathcal{G}:X\to\mathbb{R}\cup\{+\infty\} be a lower semicontinuous function for the metric topology, bounded from below and taking a finite value at some point. Assume that for some xXx\in X and some ϵ>0\epsilon>0 we have

𝒢(u)infX𝒢+ϵ.\mathcal{G}(u)\leq\inf_{X}\mathcal{G}+\epsilon.

Then, there exists x~X\tilde{x}\in X such that

  1. (i)

    d(x,x~)ϵd(x,\tilde{x})\leq\sqrt{\epsilon},

  2. (ii)

    𝒢(x~)𝒢(x)\mathcal{G}(\tilde{x})\leq\mathcal{G}(x),

  3. (iii)

    𝒢(x~)𝒢(y)+ϵd(x~,y)\mathcal{G}(\tilde{x})\leq\mathcal{G}(y)+\sqrt{\epsilon}d(\tilde{x},y) for all yXy\in X.

2.6. Estimates for Elliptic systems

Lemma 2.12.

We are going to consider a strongly 𝔸\mathbb{A}-Legendre-Hadamard elliptic bilinear form :((𝔸))2\mathcal{B}\colon(\mathcal{R}(\mathbb{A}))^{2}\to\mathbb{R}, i. e., there exist α,β>0\alpha,\beta>0 such that for all ξn\xi\in\mathbb{R}^{n} and vVv\in V we have

[ξ𝔸v,ξ𝔸v]α|ξ𝔸v|2,and||β.\mathcal{B}[\xi\otimes_{\mathbb{A}}v,\xi\otimes_{\mathbb{A}}v]\geq\alpha|\xi\otimes_{\mathbb{A}}v|^{2},\ \textnormal{and}\ |\mathcal{B}|\leq\beta.
  • (i)

    For every gW1n+1,n+1n(𝕊;V)g\in\mathrm{W}^{\frac{1}{n+1},\frac{n+1}{n}}(\mathbb{S};V) there exists a unique weak solution hW1,n+1n(𝔹;V)h\in\mathrm{W}^{1,\frac{n+1}{n}}(\mathbb{B};V) of the elliptic system:

    {𝔸(𝔸h)=0in𝔹h=gon𝕊.\begin{cases}\begin{aligned} \mathbb{A}^{*}(\mathcal{B}\mathbb{A}h)&=0&\quad&\text{in}\ \mathbb{B}\\ h&=g&\quad&\text{on}\ \mathbb{S}.\end{aligned}\end{cases}

    Furthermore, there exists a positive constant C=C(dV,dW,n,βα)C=C(d_{V},d_{W},n,\frac{\beta}{\alpha}) such that we have the estimates

    (2.9) hW1,n+1n(𝔹)CgW1n+1,n+1n(𝕊)andhLn+1n(𝔹)C[g]W1n+1,n+1n(𝕊).\left\lVert h\right\rVert_{\mathrm{W}^{1,\frac{n+1}{n}}(\mathbb{B})}\leq C\left\lVert g\right\rVert_{\mathrm{W}^{\frac{1}{n+1},\frac{n+1}{n}}(\mathbb{S})}\ \textnormal{and}\ \left\lVert\nabla h\right\rVert_{\mathrm{L}^{\frac{n+1}{n}}(\mathbb{B})}\leq C[g]_{\mathrm{W}^{\frac{1}{n+1},\frac{n+1}{n}}(\mathbb{S})}.
  • (ii)

    For every gL(𝔹;W)g\in\mathrm{L}^{\infty}(\mathbb{B};W) and every p>np>n there exists a unique solution u(W1,W01,p)(𝔹;V)u\in(\mathrm{W}^{1,\infty}\cap\mathrm{W}^{1,p}_{0})(\mathbb{B};V) of the elliptic system:

    (2.10) {𝔸(𝔸u)=gin𝔹u=0on𝕊.\begin{cases}\begin{aligned} \mathbb{A}^{*}(\mathcal{B}\mathbb{A}u)&=g&\quad&\text{in}\ \mathbb{B}\\ u&=0&\quad&\text{on}\ \mathbb{S}.\end{aligned}\end{cases}

    Furthermore, there exists a positive constant C=C(p,dV,dW,n,βα)C=C(p,d_{V},d_{W},n,\frac{\beta}{\alpha}) such that uW1,(𝔹)CgLp(𝔹)\left\lVert u\right\rVert_{\mathrm{W}^{1,\infty}(\mathbb{B})}\leq C\left\lVert g\right\rVert_{\mathrm{L}^{p}(\mathbb{B})}.

  • (iii)

    Moreover, if hW𝔸,1(Ω;V)h\in W^{\mathbb{A},1}(\Omega;V) satisfies

    𝔸(𝔸u)=0in 𝒟(Ω;V),\mathbb{A}^{*}(\mathcal{B}\mathbb{A}u)=0\quad\text{in $\mathcal{D}^{\prime}(\Omega;V)$},

    then uC(Br;V)u\in C^{\infty}(B_{r};V) and

    (2.11) supBr/2|u|+rsupBr/2|2u|CBr|u|dx\sup_{B_{r/2}}\lvert\nabla u\rvert+r\sup_{B_{r/2}}\lvert\nabla^{2}u\rvert\leq C\fint_{B_{r}}\lvert\nabla u\rvert\,\mathrm{d}x

    for all balls Br:-Br(x0)ΩB_{r}\coloneq B_{r}(x_{0})\Subset\Omega, where C=C(n,dV,dW,βα)>0C=C(n,d_{V},d_{W},\frac{\beta}{\alpha})>0 is a constant.

Proof.

We define the bilinear form ~:(N×n)2\tilde{\mathcal{B}}\colon(\mathbb{R}^{N\times n})^{2}\to\mathbb{R} by the relations

~[ej1k1,ej2k2]=[𝔸k1vj1,𝔸k2vj2]forj1,j2=1,,N,k1,k2=1,,n.\tilde{\mathcal{B}}[e_{j_{1}k_{1}},e_{j_{2}k_{2}}]=\mathcal{B}[\mathbb{A}_{k_{1}}v_{j_{1}},\mathbb{A}_{k_{2}}v_{j_{2}}]\ \textnormal{for}\ j_{1},j_{2}=1,...,N,\ k_{1},k_{2}=1,...,n.

Note that by construction we have that ~\tilde{\mathcal{B}} is strongly Legendre-Hadamard elliptic, i. e., we have for all z𝒞(N,n)z\in\mathcal{C}(N,n):

~[z,z]αc𝔸|z|2,and|~|βNsupα=1n|𝔸α|2.\tilde{\mathcal{B}}[z,z]\geq\alpha c_{\mathbb{A}}|z|^{2},\ \textnormal{and}\ |\tilde{\mathcal{B}}|\leq\beta N\sup\limits_{\alpha=1}^{n}|\mathbb{A}_{\alpha}|^{2}.

Applying [33, Proposition 2.11] in combination with Morrey’s Inequality yields (i)(i) and (ii)(ii). For the gradient estimate in (i)(i) we note that we have h=(h𝕊gdn1)\nabla h=\nabla(h-\fint_{\mathbb{S}}g\,\mathrm{d}\mathcal{H}^{n-1}) and that we have g𝕊gdn1C[g]\left\lVert g-\fint_{\mathbb{S}}g\,\mathrm{d}\mathcal{H}^{n-1}\right\rVert\leq C[g] for a constant independent of gg. The third item may be derived by means of the difference quotient method as it has been carried out on the level of first order derivatives in the proof of [13, Proposition 2.10].

Corollary 2.13.

Let B=BR(x0)B=B_{R}(x_{0}), r(38R40,39R40)r\in(\frac{38R}{40},\frac{39R}{40}), B~=Br(x0)\tilde{B}=B_{r}(x_{0}) and let gW1n+1,n+1n(B~)g\in\mathrm{W}^{\frac{1}{n+1},\frac{n+1}{n}}(\partial\tilde{B}). We suppose that hW1,n+1n(B~;V)h\in\mathrm{W}^{1,\frac{n+1}{n}}(\tilde{B};V) solves the elliptic system:

(2.12) {𝔸(𝔸u)=0in B~u=gon B~.\begin{cases}\begin{aligned} \mathbb{A}^{*}(\mathcal{B}\mathbb{A}u)&=0&\quad&\text{in $\tilde{B}$}\\ u&=g&\quad&\text{on $\partial\tilde{B}$.}\end{aligned}\end{cases}

Then there exists C(n,dV,dW,βα)>0C(n,d_{V},d_{W},\frac{\beta}{\alpha})>0 such that for all σ(0,110)\sigma\in(0,\frac{1}{10}) we have for Ah[x]=h(x0)+h(x0),xx0A_{h}[x]=h(x_{0})+\langle\nabla h(x_{0}),x-x_{0}\rangle:

(2.13) B2σRV1(hAhσR)dxCσnRnV1(σrn2n+1[g]W1n+1,n+1n(B~))\int_{B_{2\sigma R}}V_{1}\Big{(}\frac{h-A_{h}}{\sigma R}\Big{)}\,\mathrm{d}x\leq C\sigma^{n}R^{n}V_{1}\Big{(}\sigma r^{-\frac{n^{2}}{n+1}}[g]_{\mathrm{W}^{\frac{1}{n+1},\frac{n+1}{n}}(\partial\tilde{B})}\Big{)}

Furthermore, we have:

(2.14) supBr3|h|Crn2n+1[g]W1n+1,n+1n(B~).\sup\limits_{B_{\frac{r}{3}}}|\nabla h|\leq Cr^{-\frac{n^{2}}{n+1}}[g]_{\mathrm{W}^{\frac{1}{n+1},\frac{n+1}{n}}(\partial\tilde{B})}.
Proof.

By Taylor’s formula, we obtain in conjunction with 2.11 and Jensen’s Inequality

supB2σR(x0)|hAhσR|σRsupB15R(x0)|2h|Cσ(B25R(x0)|h|n+1ndx)nn+1.\sup\limits_{B_{2\sigma R}(x_{0})}\Big{|}\frac{h-A_{h}}{\sigma R}\Big{|}\leq\sigma R\sup\limits_{B_{\frac{1}{5}R}(x_{0})}|\nabla^{2}h|\leq C\sigma\Big{(}\fint_{B_{\frac{2}{5}R}(x_{0})}|\nabla h|^{\frac{n+1}{n}}\,\mathrm{d}x\Big{)}^{\frac{n}{n+1}}.

Combining this with the estimate 2.9 yields the corollary. ∎

2.7. Auxiliary Measure Theory

Lemma 2.14.

Let <a<b<-\infty<a<b<\infty and let J(a,b)J\subset(a,b) be a measurable subset with 1((a,b)J)=0\mathcal{L}^{1}((a,b)\setminus J)=0. Then for every gL1((a,b);0)g\in L^{1}((a,b);\mathbb{R}_{\geq 0}), there exists a Lebesgue point ξ0J\xi_{0}\in J for gg such that

g(ξ0)=limr0ξ0rξ0+rgdx2baabgdx,g^{\ast}(\xi_{0})=\lim_{r\searrow 0}\fint_{\xi_{0}-r}^{\xi_{0}+r}g\,\mathrm{d}x\leq\frac{2}{b-a}\int_{a}^{b}g\,\mathrm{d}x,

where gg^{\ast} is the precise representative of gg.

The Lemma can be verified by means of a contraposition argument. One may assume without loss of generality that abg(x)dx>0\int_{a}^{b}g(x)\,\mathrm{d}x>0. Then one integrates the reversed inequality to derive a contradiction to the latter integral being positive.

3. Fubini-type theorem

In this section, we are going to establish a Fubini-type property for BV𝔸\mathrm{BV}^{\mathbb{A}}-maps. Later on, this will prove essential in order to construct a \mathcal{B}-harmonic approximation of a given local generalised minimiser. We have seen that certain semi-norms of a fractional Sobolev space on some ball may be estimated from above by the total 𝔸\mathbb{A}-variation on a larger ball. We will prove that for 1\mathcal{L}^{1}-almost every sufficiently small radii RR, an (n1)(n-1)-dimensional analogous estimate holds on a sphere of radius RR.

Theorem 3.1.

Let n2n\geq 2 and α(0,1)\alpha\in(0,1). Let further be x0nx_{0}\in\mathbb{R}^{n}, R>0R>0 and uBVloc𝔸(n)u\in\mathrm{BV}^{\mathbb{A}}_{\mathrm{loc}}(\mathbb{R}^{n}). Then for 1\mathcal{L}^{1}-a.e. radius r(0,R)r\in(0,R), the restrictions u|Br(x0)u\bigr{\rvert}_{\partial B_{r}(x_{0})} are well-defined and belong to the space Wα,p(Br(x0);V)W^{\alpha,p}(\partial B_{r}(x_{0});V), where p:-nn1+αp\coloneq\frac{n}{n-1+\alpha}.

Moreover, there exists a constant C=C(𝔸,n,α)>0C=C(\mathbb{A},n,\alpha)>0, independent of x0x_{0}, RR and uu, such that for all 0<s<r<R0<s<r<R there exists t(s,r)t\in(s,r) with

(3.1) (Bt(x0)Bt(x0)|ub(x)ub(y)|p|xy|n1+αpdσxdσy)1pCrntn1p(rs)1pB2r(x0)|𝔸u|\left(\fint_{\partial B_{t}(x_{0})}\int_{\partial B_{t}(x_{0})}\frac{\lvert u_{b}(x)-u_{b}(y)\rvert^{p}}{\lvert x-y\rvert^{n-1+\alpha p}}\,\mathrm{d}\sigma_{x}\,\mathrm{d}\sigma_{y}\right)^{\frac{1}{p}}\leq C\frac{r^{n}}{t^{\frac{n-1}{p}}(r-s)^{\frac{1}{p}}}\fint_{B_{2r}(x_{0})}\lvert\mathbb{A}u\rvert

for some suitable b𝒩(𝔸)b\in\mathcal{N}(\mathbb{A}).

Remark.

In the follwoing constellation p=n+1np=\frac{n+1}{n}, α=1n+1\alpha=\frac{1}{n+1} and s>Crs>Cr, the inequality 3.1 then takes the form

(3.2) [ub]W1n+1,n+1n(Br)Crn2n+1B2r|𝔸u|.[u_{b}]_{\mathrm{W}^{\frac{1}{n+1},\frac{n+1}{n}}(\partial B_{r})}\leq Cr^{\frac{n^{2}}{n+1}}\fint_{B_{2r}}|\mathbb{A}u|.
Proof.

The proof is analogous to the one for [32, Theorem 4.1] in the BD\mathrm{BD}-case and only requires minor modifications in the second and the third step.

Let θ(0,1)\theta\in(0,1), q[1,){q\in[1,\infty)} and u(Wθ,qC)(n;V)u\in(W^{\theta,q}\cap C)(\mathbb{R}^{n};V). Then it has been established [32, Theorem 4.1] that there is a constant C=C(n,θ,q)>0C=C(n,\theta,q)>0 such that for all R>0R>0 we have

(3.3) 0RBr×Br|u(x)u(y)|q|xy|n1+θqdσxdσydrCBR×BR|u(x)u(y)|q|xy|n+θqdxdy.\int_{0}^{R}\iint_{\partial B_{r}\times\partial B_{r}}\frac{\lvert u(x)-u(y)\rvert^{q}}{\lvert x-y\rvert^{n-1+\theta q}}\,\mathrm{d}\sigma_{x}\,\mathrm{d}\sigma_{y}\,\mathrm{d}r\leq C\iint_{B_{R}\times B_{R}}\frac{\lvert u(x)-u(y)\rvert^{q}}{\lvert x-y\rvert^{n+\theta q}}\,\mathrm{d}x\,\mathrm{d}y.

The aim now is to establish that uu may be explicitly evaluated n1\mathcal{H}^{n-1}-a.e. point wisely on 1\mathcal{L}^{1}-a.e. sphere centred at the origin. For that matter, let uBV𝔸(n)u\in\mathrm{BV}^{\mathbb{A}}(\mathbb{R}^{n}) and let 0<R1<R2<0<R_{1}<R_{2}<\infty be arbitrary. The set

I:-{t(R1,R2)|𝔸u|(Bt)>0}I\coloneq\{t\in(R_{1},R_{2})\mid\lvert\mathbb{A}u\rvert(\partial B_{t})>0\}

is at most countable and hence a 1\mathcal{L}^{1}-nullset. Now let t(R1,R2)It\in(R_{1},R_{2})\setminus I. Then by [10, Corollary 4.21],

𝔸u  Bt=(TrBt+(u)TrBt(u))𝔸νBtn1  Bt,\mathbb{A}u\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}_{\,\partial}B_{t}=(\operatorname{Tr}_{B_{t}}^{+}(u)-\operatorname{Tr}_{B_{t}}^{-}(u))\otimes_{\mathbb{A}}\nu_{\partial B_{t}}\mathcal{H}^{n-1}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}_{\,\partial}B_{t},

where νBt\nu_{B_{t}} denotes the outer unit normal to Bt\partial B_{t}. Hence, using that t(R1,R2)It\in(R_{1},R_{2})\setminus I in the last step,

Bt|(TrBt+(u)TrBt(u))𝔸νBt|dσ=|𝔸u|(Bt)=0.\int_{\partial B_{t}}\lvert(\operatorname{Tr}_{B_{t}}^{+}(u)-\operatorname{Tr}_{B_{t}}^{-}(u))\otimes_{\mathbb{A}}\nu_{\partial B_{t}}\rvert\,\mathrm{d}\sigma=\lvert\mathbb{A}u\rvert(\partial B_{t})=0.

As a result, we have |TrBt+(u)TrBt(u)|=0\lvert\operatorname{Tr}_{B_{t}}^{+}(u)-\operatorname{Tr}_{B_{t}}^{-}(u)\rvert=0 n1\mathcal{H}^{n-1}-a.e. on Bt\partial B_{t}. Furthermore, writing u~(x):-TrBt+(u)(x)=TrBt(u)(x)\widetilde{u}(x)\coloneq\operatorname{Tr}_{B_{t}}^{+}(u)(x)=\operatorname{Tr}_{B_{t}}^{-}(u)(x) for such xBtx\in\partial B_{t}, (2.1) implies that

(3.4) limr0Br(x)Bt|uu~(x)|dy=limr0Br(x)Bt¯c|uu~(x)|dy=0.\lim_{r\searrow 0}\fint_{B_{r}(x)\cap B_{t}}\lvert u-\widetilde{u}(x)\rvert\,\mathrm{d}y=\lim_{r\searrow 0}\fint_{B_{r}(x)\cap{\mkern 3.0mu\overline{\mkern-3.0muB_{t}\mkern-3.0mu}\mkern 1.0mu}^{c}}\lvert u-\widetilde{u}(x)\rvert\,\mathrm{d}y=0.

In consequence, we have

limr0Br(x)|uu~(x)|dy=0.\lim_{r\searrow 0}\fint_{B_{r}(x)}\lvert u-\widetilde{u}(x)\rvert\,\mathrm{d}y=0.

But this means that n1\mathcal{H}^{n-1}-a.e. xBtx\in\partial B_{t} is a Lebesgue point of uu for 1\mathcal{L}^{1}-a.e. radius t(R1,R2)t\in(R_{1},R_{2}).

Let α(0,1)\alpha\in(0,1) be arbitrary and set p:-n/(n1+α)p\coloneq n/(n-1+\alpha). Let further uBVloc𝔸(n)u\in\mathrm{BV}^{\mathbb{A}}_{\mathrm{loc}}(\mathbb{R}^{n}) and consider for ε>0\varepsilon>0 a family of standard mollifiers uε(x):-(ρεu)(x)u_{\varepsilon}(x)\coloneq(\rho_{\varepsilon}\ast u)(x).

Note that for each Lebesgue point xnx\in\mathbb{R}^{n} of uu, one has uε(x)u(x)u_{\varepsilon}(x)\to u^{\ast}(x) as ε0\varepsilon\searrow 0, where uu^{\ast} is the precise representative of uu.

Now invoking Lemma 2.7 for uεu_{\varepsilon} provides an element bε𝒩(𝔸)b_{\varepsilon}\in\mathcal{N}(\mathbb{A}) such that

(3.5) (BrBr|ub,ε(x)ub,ε(y)|p|xy|n+αpdxdy)1pCr1αB2r|𝔸uε|,\left(\fint_{B_{r}}\int_{B_{r}}\frac{\lvert u_{b,\varepsilon}(x)-u_{b,\varepsilon}(y)\rvert^{p}}{\lvert x-y\rvert^{n+\alpha p}}\,\mathrm{d}x\,\mathrm{d}y\right)^{\frac{1}{p}}\leq Cr^{1-\alpha}\fint_{B_{2r}}\lvert\mathbb{A}u_{\varepsilon}\rvert,

where ub,ε:-uεbεu_{b,\varepsilon}\coloneq u_{\varepsilon}-b_{\varepsilon}. We note also that bεC(n;V)b_{\varepsilon}\in C^{\infty}(\mathbb{R}^{n};V) are in fact the L2L^{2}-orthogonal projections of uεu_{\varepsilon} onto 𝒩(𝔸)\mathcal{N}(\mathbb{A}) and satisfy the L1L^{1}-stability estimate [10, Section 3.1]:

bεL1(Br)CuεL1(Br)uL1(Br).\lVert b_{\varepsilon}\rVert_{L^{1}(B_{r})}\leq C\lVert u_{\varepsilon}\rVert_{L^{1}(B_{r})}\to\lVert u\rVert_{L^{1}(B_{r})}.

Since 𝔸\mathbb{A} is \mathbb{C}-elliptic, the nullspace 𝒩(𝔸)\mathcal{N}(\mathbb{A}) is of finite dimension, so one can find a subsequence (bεj)(bε)(b_{\varepsilon_{j}})\subset(b_{\varepsilon}) and some b𝒩(𝔸)b\in\mathcal{N}(\mathbb{A}) such that bεjbb_{\varepsilon_{j}}\to b in 𝒩(𝔸)\mathcal{N}(\mathbb{A}). Consequently, denoting ubu_{b}^{\ast} to be the precise representative of ubu_{b}, one can estimate

srBt×Bt|ub(x)ub(y)|p|xy|n1+αpdσxdσydt\displaystyle\int_{s}^{r}\iint_{\partial B_{t}\times\partial B_{t}}\frac{\lvert u^{\ast}_{b}(x)-u^{\ast}_{b}(y)\rvert^{p}}{\lvert x-y\rvert^{n-1+\alpha p}}\,\mathrm{d}\sigma_{x}\,\mathrm{d}\sigma_{y}\,\mathrm{d}t
lim infεj0srBt×Bt|ub,εj(x)ub,εj(y)|p|xy|n1+αpdσxdσydt\displaystyle\hskip 42.67912pt\leq\liminf_{\varepsilon_{j}\searrow 0}\int_{s}^{r}\iint_{\partial B_{t}\times\partial B_{t}}\frac{\lvert u_{b,\varepsilon_{j}}(x)-u_{b,\varepsilon_{j}}(y)\rvert^{p}}{\lvert x-y\rvert^{n-1+\alpha p}}\,\mathrm{d}\sigma_{x}\,\mathrm{d}\sigma_{y}\,\mathrm{d}t
(by (3.3)) Clim infεj0BrBr|ub,εj(x)ub,εj(y)|p|xy|n+αpdσxdσy\displaystyle\hskip 42.67912pt\leq C\liminf_{\varepsilon_{j}\searrow 0}\int_{B_{r}}\int_{B_{r}}\frac{\lvert u_{b,\varepsilon_{j}}(x)-u_{b,\varepsilon_{j}}(y)\rvert^{p}}{\lvert x-y\rvert^{n+\alpha p}}\,\mathrm{d}\sigma_{x}\,\mathrm{d}\sigma_{y}
(by (3.5)) Clim infεj0rn(r1αB2r|𝔸uεj|)p\displaystyle\hskip 42.67912pt\leq C\liminf_{\varepsilon_{j}\searrow 0}r^{n}\left(r^{1-\alpha}\fint_{B_{2r}}\lvert\mathbb{A}u_{\varepsilon_{j}}\rvert\right)^{p}
Crn(r1αB2r|𝔸u|)p\displaystyle\hskip 42.67912pt\leq Cr^{n}\left(r^{1-\alpha}\fint_{B_{2r}}\lvert\mathbb{A}u\rvert\right)^{p}

Next, towards employing the auxiliary Lemma 2.14, consider the set

J:-{t(s,r)|𝔸u|(Bt)=0}J\coloneq\{t\in(s,r)\mid\lvert\mathbb{A}u\rvert(\partial B_{t})=0\}

and let g:(s,r)0g\colon(s,r)\to\mathbb{R}_{\geq 0} be defined by

g(t):-{Bt×Bt|ub(x)ub(y)|p|xy|n1+αdσxdσy,tJ,0,else.g(t)\coloneq\begin{cases}\begin{aligned} &\iint_{\partial B_{t}\times\partial B_{t}}\frac{\lvert u^{\ast}_{b}(x)-u^{\ast}_{b}(y)\rvert^{p}}{\lvert x-y\rvert^{n-1+\alpha}}\,\mathrm{d}\sigma_{x}\,\mathrm{d}\sigma_{y},&\quad&t\in J,\\[4.30554pt] &0,&\quad&\text{else.}\end{aligned}\end{cases}

Then, by Step 2 and Lemma 2.14, there is a tJt\in J such that

g(t)2rssrg(t)dtCrnrs(r1αB2r|𝔸u|)p.g^{\ast}(t)\leq\frac{2}{r-s}\int_{s}^{r}g(t)\,\mathrm{d}t\leq C\frac{r^{n}}{r-s}\left(r^{1-\alpha}\fint_{B_{2r}}\lvert\mathbb{A}u\rvert\right)^{p}.

Now plugging the definition of g(t)g(t) in the above inequality finally produces

(BtBt|ub(x)ub(y)|p|xy|n1+αpdσxdσy)1pCrnpr1αtn1p(rs)1pB2r|𝔸u|,\left(\fint_{\partial B_{t}}\int_{\partial B_{t}}\frac{\lvert u^{\ast}_{b}(x)-u^{\ast}_{b}(y)\rvert^{p}}{\lvert x-y\rvert^{n-1+\alpha p}}\,\mathrm{d}\sigma_{x}\,\mathrm{d}\sigma_{y}\right)^{\frac{1}{p}}\leq C\frac{r^{\frac{n}{p}}r^{1-\alpha}}{t^{\frac{n-1}{p}}(r-s)^{\frac{1}{p}}}\fint_{B_{2r}}\lvert\mathbb{A}u\rvert,

which is the desired estimate (3.1). Note that throughout the computations, the generic constant CC did not depend on uu or rr, completing the proof. ∎

4. \mathcal{B}-harmonic Approximation

In this section, we are going to construct a \mathcal{B}-harmonic Approximation hh for a given local generalised minimser uu. We will solve an elliptic system, which later on will be a linearisation of the Euler-Lagrange equation at a given average (𝔸u)B(\mathbb{A}u)_{B}, where BΩB\subset\Omega is a ball. We will assume that uu behaves nicely on the boundary B\partial B and we will give a precise estimation of uhu-h in terms of the excess associated to uu. The Ekeland variational principle allows us to prove a corresponding estimate for all minimisers close to uu and in the limit the estimate is inherited to uu itself. The precise control of uhu-h will play a crucial role in the excess decay estimates.

Theorem 4.1.

Assuming (H0), (H1) and (H2), let uBVloc𝔸(Ω)u\in\mathrm{BV}^{\mathbb{A}}_{\mathrm{loc}}(\Omega) be a generalised local minimiser of \mathcal{F}, let M>0M>0, q(1,nn1)q\in(1,\frac{n}{n-1}), and let B=Br(x0)ΩB=B_{r}(x_{0})\Subset\Omega such that u|B=TrB(u)=Tr+B(u)W1n+1,n+1n(B)u\bigr{\rvert}_{\partial B}=\operatorname{Tr}^{-}_{B}(u)=\operatorname{Tr}^{+}_{B}(u)\in\mathrm{W}^{\frac{1}{n+1},\frac{n+1}{n}}(\partial B). Let a:nVa\colon\mathbb{R}^{n}\to V be an arbitrary affine map with |𝔸a|M|\mathbb{A}a|\leq M. For =D2f(𝔸a)\mathcal{B}=D^{2}f(\mathbb{A}a), let hW1,n+1n(B)h\in\mathrm{W}^{1,\frac{n+1}{n}}(B) be the unique weak solution of the elliptic system

(4.1) {𝔸(𝔸h)=0inBh=u|BaonB.\begin{cases}\begin{aligned} \mathbb{A}^{*}(\mathcal{B}\mathbb{A}h)&=0&\quad&\text{in}\ B\\ h&=u_{|\partial B}-a&\quad&\text{on}\ \partial B\end{aligned}\end{cases}.

Then there exists a constant C=C(M,dV,dW,n,q,Lν)>0C=C(M,d_{V},d_{W},n,q,\frac{L}{\nu})>0 such that

BV(uahr)dxC(BV(𝔸(ua)))q.\fint_{B}V\left(\frac{u-a-h}{r}\right)\,\mathrm{d}x\leq C\left(\fint_{B}V(\mathbb{A}(u-a))\right)^{q}.
Proof.

We confine ourselves to merely sketching the proof, since it follows the lines of the proof [32, Proposition 5.4]. To start with we fix some notation u~=ua\widetilde{u}=u-a, f~=(f)𝔸a\widetilde{f}=(f)_{\mathbb{A}a} and X={vW𝔸,1(B;N)Tr(v)=Tr(u~)}X=\{v\in W^{\mathbb{A},1}(B;\mathbb{R}^{N})\mid\operatorname{Tr}(v)=\operatorname{Tr}(\widetilde{u})\}.

Let ε>0\varepsilon>0. There is a uϵXu_{\epsilon}\in X such that

B|uεu~r|+|V1(𝔸uε)V1(𝔸u~)|ε2,Bf~(𝔸uε)Bf~(𝔸u~)+ε2.\fint_{B}\left|\frac{u_{\varepsilon}-\widetilde{u}}{r}\right|+\left|\fint V_{1}\left(\mathbb{A}u_{\varepsilon}\right)-V_{1}\left(\mathbb{A}\widetilde{u}\right)\right|\leq\varepsilon^{2},\qquad\fint_{B}\widetilde{f}(\mathbb{A}u_{\varepsilon})\leq\fint_{B}\widetilde{f}(\mathbb{A}\widetilde{u})+\varepsilon^{2}.

Noting that u~\widetilde{u} is a local generalised minimiser of the functional a[ζ;ω]=ωf~(𝔸ζ)dx\mathcal{F}_{a}[\zeta;\omega]=\int_{\omega}\widetilde{f}(\mathbb{A}\zeta)\,\mathrm{d}x and taking proposition 2.6 into account allows to apply the Ekeland variational principle 2.11 to (X,d)(X,\,\mathrm{d}), x=uεx=u_{\varepsilon} and 𝒢=a\mathcal{G}=\mathcal{F}_{a}, where d(ζ1,ζ2)=𝔸(ζ1ζ2)L1(B)\,\mathrm{d}(\zeta_{1},\zeta_{2})=\left\lVert\mathbb{A}(\zeta_{1}-\zeta_{2})\right\rVert_{\mathrm{L}^{1}(B)}. This in conjunction with Poincaré’s Inequality, yields u~εX\widetilde{u}_{\varepsilon}\in X such that for all ζW0𝔸,1(B)\zeta\in\mathrm{W}^{\mathbb{A},1}_{0}(B) we have

|BDf~(𝔸u~ε)[𝔸ζ]dx|εB|𝔸ζ|dxanduεu~εW𝔸,1(B)Cε(rn+rn+1)\Big{|}\int_{B}D\widetilde{f}(\mathbb{A}\widetilde{u}_{\varepsilon})[\mathbb{A}\zeta]\,\mathrm{d}x\Big{|}\leq\varepsilon\int_{B}|\mathbb{A}\zeta|\,\mathrm{d}x\ \textnormal{and}\ \left\lVert u_{\varepsilon}-\widetilde{u}_{\varepsilon}\right\rVert_{\mathrm{W}^{\mathbb{A},1}(B)}\leq C\varepsilon(r^{n}+r^{n+1})

for a constant C>0C>0 independent of uε,u~εu_{\varepsilon},\widetilde{u}_{\varepsilon} and rr. Writing [u~ε,]=Λ[𝔸u~ε,]+Df~(𝔸u~ε)\mathcal{B}[\widetilde{u}_{\varepsilon},\cdot]=\Lambda[\mathbb{A}\widetilde{u}_{\varepsilon},\cdot]+D\widetilde{f}(\mathbb{A}\widetilde{u}_{\varepsilon}) and applying the pointwise estimate 2.5 to the first term then yields for some constant C(M,L)>0C(M,L)>0 and all ζW0𝔸,1(B)\zeta\in\mathrm{W}^{\mathbb{A},1}_{0}(B)

(4.2) |B[𝔸u~ε,𝔸ζ]dx|CBV1(𝔸u~ε)|𝔸ζ|dx+εB|𝔸ζ|dx.\Big{|}\int_{B}\mathcal{B}[\mathbb{A}\widetilde{u}_{\varepsilon},\mathbb{A}\zeta]\,\mathrm{d}x\Big{|}\leq C\int_{B}V_{1}(\mathbb{A}\widetilde{u}_{\varepsilon})|\mathbb{A}\zeta|\,\mathrm{d}x+\varepsilon\int_{B}|\mathbb{A}\zeta|\,\mathrm{d}x.

At this stage, we scale things to the unit ball 𝔹\mathbb{B}: For a measurable map pp defined on BB we put S[p](x)=r1p(x0+rx)S[p](x)=r^{-1}p(x_{0}+rx). Furthermore, we put Ψε=S[u~εh]\Psi_{\varepsilon}=S[\widetilde{u}_{\varepsilon}-h] and Uε=S[u~ε]U_{\varepsilon}=S[\widetilde{u}_{\varepsilon}]. We will now truncate the map Ψε\Psi_{\varepsilon}. To this aim, we put

T(w)={w,w1ww,w>1.\textbf{T}(w)=\begin{cases}w,\ &\left\lVert w\right\rVert\leq 1\\ \frac{w}{\left\lVert w\right\rVert},\ &\left\lVert w\right\rVert>1\end{cases}.

We fix p>np>n and let ΦεW1,W01,p)(𝔹;V)\Phi_{\varepsilon}\in\mathrm{W}^{1,\infty}\cap\mathrm{W}^{1,p}_{0})(\mathbb{B};V) be the solution of the elliptic system

(4.3) {𝔸(𝔸Φε)=TΨεin𝔹Φε=0on𝕊.\begin{cases}\begin{aligned} \mathbb{A}^{*}(\mathcal{B}\mathbb{A}\Phi_{\varepsilon})&=\textbf{T}\circ\Psi_{\varepsilon}&\quad&\text{in}\ \mathbb{B}\\ \Phi_{\varepsilon}&=0&\quad&\text{on}\ \mathbb{S}\end{aligned}\end{cases}.

Now testing the system with ΨεW0𝔸,1(𝔹)\Psi_{\varepsilon}\in\mathrm{W}^{\mathbb{A},1}_{0}(\mathbb{B}) and exploiting 4.2 yields the key estimation

𝔹V1(Ψε)dx𝔹TΨε,Ψεdx=𝔹[𝔸Φε,𝔸Ψε]C(𝔹V1(Ψε)dx+ε)TΨεLp(B)C(𝔹V1(Uε)dx+ε)(𝔹V1(Ψε)dx)1p\begin{split}\int_{\mathbb{B}}V_{1}(\Psi_{\varepsilon})\,\mathrm{d}x&\leq\int_{\mathbb{B}}\langle\textbf{T}\circ\Psi_{\varepsilon},\Psi_{\varepsilon}\rangle\,\mathrm{d}x\\ &=\int_{\mathbb{B}}\mathcal{B}[\mathbb{A}\Phi_{\varepsilon},\mathbb{A}\Psi_{\varepsilon}]\\ &\leq C\big{(}\int_{\mathbb{B}}V_{1}(\Psi_{\varepsilon})\,\mathrm{d}x+\varepsilon\big{)}\left\lVert\textbf{T}\circ\Psi_{\varepsilon}\right\rVert_{\mathrm{L}^{p}(B)}\\ &\leq C\big{(}\int_{\mathbb{B}}V_{1}(U_{\varepsilon})\,\mathrm{d}x+\varepsilon\big{)}\big{(}\int_{\mathbb{B}}V_{1}(\Psi_{\varepsilon})\,\mathrm{d}x\big{)}^{\frac{1}{p}}\end{split}

for a constant C=C(M,dV,dW,n,q,Lν)C=C(M,d_{V},d_{W},n,q,\frac{L}{\nu}). Note that in the penultimate step we have exploited the estimate 2.9. Dividing by (𝔹V1(Ψε)dx)1p(\int_{\mathbb{B}}V_{1}(\Psi_{\varepsilon})\,\mathrm{d}x)^{\frac{1}{p}}, sending ε0\varepsilon\to 0, scaling back to the ball BB and setting q=pq=p^{\prime} concludes the proof. ∎

5. Excess Decay

In this section, we will display the most important step towards proving an excess decay for a given local generalised minimiser uu. First, we will invoke Caccioppoli’s Inequality and then, using our Fubini-type theorem for BV𝔸\mathrm{BV}^{\mathbb{A}}-maps, we will construct a \mathcal{B}-harmonic approximation. Then we will show separately that hh and uhu-h have a good decay. Here, the excess of uu is defined as follows:

𝐄(u,x,r)Br(x)V1(𝔸u(𝔸u)Br(x))and𝐄~(u,x,r)𝐄(u,x,r)n(Br(x).\mathbf{E}(u,x,r)\coloneqq\int_{B_{r}(x)}V_{1}\left(\mathbb{A}u-(\mathbb{A}u)_{B_{r}(x)}\right)\ \quad\textnormal{and}\quad\widetilde{\mathbf{E}}(u,x,r)\coloneqq\frac{\mathbf{E}(u,x,r)}{\mathcal{L}^{n}(B_{r}(x)}.
Lemma 5.1.

Assuming (H0), (H1) and (H2), let uBVloc𝔸(Ω)u\in\mathrm{BV}^{\mathbb{A}}_{\mathrm{loc}}(\Omega) be a generalised local minimiser of \mathcal{F}. Furthermore, let M>0M>0 and q(1,nn1)q\in(1,\frac{n}{n-1}). There exists a constant C(M,𝔸,q,dV,dW,Lν)>0C(M,\mathbb{A},q,d_{V},d_{W},\frac{L}{\nu})>0 with the property: If we have B=B(x0,R)ΩB=B(x_{0},R)\Subset\Omega, |(𝔸u)B|<M|(\mathbb{A}u)_{B}|<M and B|𝔸u(𝔸u)B|1\fint_{B}|\mathbb{A}u-(\mathbb{A}u)_{B}|\leq 1, then for all σ(0,1]\sigma\in(0,1] we have

(5.1) 𝐄~(u;x0,σR)c(σ2+σn2(𝐄~(u;x0,R))q1)𝐄~(u;x0,R).\widetilde{\mathbf{E}}(u;x_{0},\sigma R)\leq c\Big{(}\sigma^{2}+\sigma^{-n-2}\big{(}\widetilde{\mathbf{E}}(u;x_{0},R)\big{)}^{q-1}\Big{)}\widetilde{\mathbf{E}}(u;x_{0},R).
Proof.

We are going to set up the proof as follows:

  • Due to Lemma 2.8 we find a(n;V)a\in\mathcal{L}(\mathbb{R}^{n};V) such that 𝔸a=(𝔸u)B\mathbb{A}a=(\mathbb{A}u)_{B}. We put u~=ua\widetilde{u}=u-a, f~=(f)(𝔸u)B\widetilde{f}=(f)_{(\mathbb{A}u)_{B}} and =D2f~(0)\mathcal{B}=D^{2}\widetilde{f}(0).

  • Due to Theorem 3.1 there exists a radius r(38R40,39R40)r\in(\frac{38R}{40},\frac{39R}{40}) such that u~|Br(x0)\widetilde{u}\bigr{\rvert}_{\partial B_{r}(x_{0})} is well-defined and belongs to the space W1n+1,n+1n(B~)\mathrm{W}^{\frac{1}{n+1},\frac{n+1}{n}}(\partial\tilde{B}) with B~=Br\tilde{B}=B_{r}. Furthermore, for some b𝒩(𝔸)b\in\mathcal{N}(\mathbb{A}), we have

    (5.2) [u~b]W1n+1,n+1n(B~)CB2r(x0)|𝔸u~|C𝐄~(u;x0,R)[\widetilde{u}_{b}]_{\mathrm{W}^{\frac{1}{n+1},\frac{n+1}{n}}(\partial\tilde{B})}\leq C\fint_{B_{2r}(x_{0})}|\mathbb{A}\widetilde{u}|\leq C\widetilde{\mathbf{E}}(u;x_{0},R)

    for a constant C(𝔸,n,α)C(\mathbb{A},n,\alpha).

  • Let hh be the solution of the elliptic system

    {𝔸(𝔸h)=0inB~h=u|B~aonB~.\begin{cases}\begin{aligned} \mathbb{A}^{*}(\mathcal{B}\mathbb{A}h)&=0&\quad&\text{in}\ \tilde{B}\\ h&=u_{|\partial\tilde{B}}-a&\quad&\text{on}\ \partial\tilde{B}\end{aligned}\end{cases}.
  • In view of Corollary 2.13, we put a0=a+Ahba_{0}=a+A_{h-b} and let σ(0,110)\sigma\in(0,\frac{1}{10}) be arbitrary but fixed. Since we have BV1(𝔸u~)1\fint_{B}V_{1}(\mathbb{A}\widetilde{u})\leq 1, we may estimate by means of Lemma 2.9 and Jensen’s Inequality

    (5.3) V1(σB𝔸u~)Cσ2(B𝔸u~)2Cσ2V1(B𝔸u~)Cσ2𝐄~(u;x0,R).V_{1}\Big{(}\sigma\fint_{B}\mathbb{A}\widetilde{u}\Big{)}\leq C\sigma^{2}\Big{(}\fint_{B}\mathbb{A}\widetilde{u}\Big{)}^{2}\leq C\sigma^{2}V_{1}\Big{(}\fint_{B}\mathbb{A}\widetilde{u}\Big{)}\leq C\sigma^{2}\widetilde{\mathbf{E}}(u;x_{0},R).

    Combining the estimates 5.3, 2.13 and 5.2 yields

    (5.4) B2σRV1(hbAhb2σR)CRnσn+2𝐄~(u;x0,R).\int_{B_{2\sigma R}}V_{1}\Big{(}\frac{h-b-A_{h-b}}{2\sigma R}\Big{)}\leq CR^{n}\sigma^{n+2}\widetilde{\mathbf{E}}(u;x_{0},R).

We note that we have due to Corollary 2.13 |𝔸a0|M0+CB2r|𝔸u~|M0+Cm|\mathbb{A}a_{0}|\leq M_{0}+C\fint_{B_{2r}}|\mathbb{A}\widetilde{u}|\leq M_{0}+C\coloneqq m for a constant C(m,𝔸,dV,dW,Lν)C(m,\mathbb{A},d_{V},d_{W},\frac{L}{\nu}). Especially, ubu-b is a generalised local minimiser as well. By Caccioppoli’s Inequality, Proposition 2.10, we estimate

𝐄(u;x0,σR)CB2σRV1(uba0σR)C{σ2Br2V1(u~hr)+B2σRV1(hbAhb2σR)}C{σ2Rn(𝐄~(u,x0,R))q+Rnσn+2𝐄(u,x0,2R)}\begin{split}\mathbf{E}(u;x_{0},\sigma R)&\leq C\int_{B_{2\sigma R}}V_{1}\Big{(}\frac{u-b-a_{0}}{\sigma R}\Big{)}\\ &\leq C\Big{\{}\sigma^{-2}\int_{B_{\frac{r}{2}}}V_{1}\Big{(}\frac{\widetilde{u}-h}{r}\Big{)}+\int_{B_{2\sigma R}}V_{1}\Big{(}\frac{h-b-A_{h-b}}{2\sigma R}\Big{)}\Big{\}}\\ &\leq C\Big{\{}\sigma^{-2}R^{n}(\widetilde{\mathbf{E}}(u,x_{0},R))^{q}+R^{n}\sigma^{n+2}\mathbf{E}(u,x_{0},2R)\Big{\}}\end{split}

for a constant C(m,𝔸,dV,dW,Lν)C(m,\mathbb{A},d_{V},d_{W},\frac{L}{\nu}). In the second step, we have exploited Theorem 4.1 and the estimate 5.4. ∎

Proposition 5.2.

Assuming (H0), (H1) and (H2), let uBVloc𝔸(Ω)u\in\mathrm{BV}^{\mathbb{A}}_{\mathrm{loc}}(\Omega) be a generalised local minimiser of \mathcal{F}, where ff satisfies (H1) and (H2). Let α(0,1)\alpha\in(0,1) and M>0M>0. Then there exist constants γ(M,𝔸,α,dV,dW,Lν)\gamma(M,\mathbb{A},\alpha,d_{V},d_{W},\frac{L}{\nu}) and ε(M,𝔸,α,dV,dW,Lν)>0\varepsilon(M,\mathbb{A},\alpha,d_{V},d_{W},\frac{L}{\nu})>0 with the property: If we have B=B(x0,R)ΩB=B(x_{0},R)\Subset\Omega, |(𝔸u)B|<M|(\mathbb{A}u)_{B}|<M and 𝐄~(u;x0;R)<ε\widetilde{\mathbf{E}}(u;x_{0};R)<\varepsilon, then we have for all ϑ(0,1)\vartheta\in(0,1):

(5.5) 𝐄~(u;x0;ϑR)γϑα𝐄~(u;x0;R).\widetilde{\mathbf{E}}(u;x_{0};\vartheta R)\leq\gamma\vartheta^{\alpha}\widetilde{\mathbf{E}}(u;x_{0};R).

The proof is analogous to the proof of [33, Proposition 4.8] or [32, Proposition 5.7] since we already have established Lemma 5.1. At this stage, we sketch the proof of the main theorem. We will pay special attention to the final step, which needs to be modified in regards to [32]:

Proof.

Using [39, 1.6.1, Theorem 1; 1.6.2, Theorem 3], we observe that n(Σu)=0\mathcal{L}^{n}(\Sigma_{u})=0.

Let x0ΩΣux_{0}\in\Omega\setminus\Sigma_{u} and α(0,1)\alpha\in(0,1). For M>0M>0, we denote by γM\gamma_{M} and εM\varepsilon_{M} the constants determined by Proposition 5.2. Then there exists M>0M>0 and a radius R>0R>0 with BR(x0)ΩB_{R}(x_{0})\Subset\Omega such that for all xB~=BR2(x0)x\in\tilde{B}=B_{\frac{R}{2}}(x_{0}) we have

|(𝔸uB~)|Mand𝐄~(u;x;R2)εM.\left|(\mathbb{A}u_{\tilde{B}})\right|\leq M\quad\textnormal{and}\quad\widetilde{\mathbf{E}}\Big{(}u;x;\frac{R}{2}\Big{)}\leq\varepsilon_{M}.

This can be inferred from a similar argument to the one in [32, 32]. In particular, we then have for all xB=BR4(x0)x\in B=B_{\frac{R}{4}}(x_{0}) and all 0<r<R20<r<\frac{R}{2}:

𝐄~(u;x;r)γMεM(2R)αrα.\widetilde{\mathbf{E}}(u;x;r)\leq\gamma_{M}\varepsilon_{M}\Big{(}\frac{2}{R}\Big{)}^{\alpha}r^{\alpha}.

From this decay, it can be deduced by a simple covering argument that the measure |𝔸u  B||\mathbb{A}u\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}_{\,B}| is absolutely continuous with respect to the Lebesgue measure.

Towards the Hölder continuity of the distributional gradient DuDu, we invoke Campanato’s characterisation [12, Theorem 5.1]: Let k=dim𝒩(𝔸)1k=\dim_{\mathbb{R}}\mathcal{N}(\mathbb{A})\geq 1 and qq be a linear VV-valued polynomial such that 𝔸q=(𝔸u)B(x,r)\mathbb{A}q=(\mathbb{A}u)_{B(x,r)}. Furthermore, we put p=q+Π(uq)𝒫k(N)p=q+\Pi(u-q)\in\mathcal{P}_{k}(\mathbb{R}^{N}). Now estimate for all xBx\in B and all 0<r<R20<r<\frac{R}{2}

B(y,r)|up|dxCPoinrB(y,r)|𝔸(uq)|dxCr1+α.\fint_{B(y,r)}|u-p|\,\mathrm{d}x\leq C_{Poin}r\fint_{B(y,r)}|\mathbb{A}(u-q)|\,\mathrm{d}x\leq Cr^{1+\alpha}.

References

  • [1] Emilio Acerbi and Nicola Fusco “Semicontinuity problems in the calculus of variations.” In Arch. Rational Mech. Anal. 86. no. 2, 1984, pp. 125–145
  • [2] Emilio Acerbi and Nicola Fusco “A regularity theorem for minimizers of quasiconvex integrals.” In Arch. Ration. Mech. Anal. 99. no. 3 Berlin/Heidelberg: Springer, 1987, pp. 261–281
  • [3] Giovanni Alberti “Rank one property for derivatives of functions with bounded variation.” In Proc. Roy. Soc. Edinburgh Sect. A 123., 1993, pp. 239–274
  • [4] Frederick J.. Almgren “Existence and regularity almost everywhere of solutions to elliptic variational problems among surfaces of varying topological type and singularity structure.” In Ann. of Math. 87. no. 2 Princeton, NJ: Princeton University, Mathematics Department, 1968, pp. 321–391
  • [5] Luigi Ambrosio and Gianni Dal Maso “On the relaxation in BV(Ω;m)\mathrm{BV}(\Omega;\mathbb{R}^{m}) of quasi-convex integrals.” In J. Funct. Anal. 109. no. 1, 1992, pp. 76–97
  • [6] Gabriele Anzellotti and Mariano Giaquinta “Convex functionals and partial regularity.” In Arch. Ration. Mech. Anal. 102. no. 3 Berlin/Heidelberg: Springer, 1988, pp. 243–272
  • [7] Adolfo Arroyo-Rabasa, Guido De Philippis and Filip Rindler “Lower semicontinuity and relaxation of linear-growth integral functionals under PDE constraints” In Adv. Calc. Var. De Gruyter, Berlin, 2020, pp. 219–255
  • [8] J.. Ball and F. Murat W1,p\mathrm{W}^{1,p}-quasiconvexity and variational problems for multiple integrals.” In J. Funct. Anal. 58. no. 3, 1984, pp. 225–253
  • [9] Lisa Beck “Elliptic regularity theory. A first course.” 19, Lect. Notes Unione Mat. Ital. Cham/Heidelberg/New York/Dordrecht/London: Springer, 2016
  • [10] Dominic Breit, Lars Diening and Franz Gmeineder “On the Trace Operator for Functions of Bounded 𝔸\mathbb{A}-Variation” In Anal. PDE 13. no. 2, 2020, pp. 559–594
  • [11] A.. Caldéron and A. Zygmund “On singular integrals.” In Amer. J. Math. 78, 1956, pp. 289–309
  • [12] Sergio Campanato “Proporietà di una famiglia di spazi funzionali.” In Ann. Scuola Norm. Sup. Pisa Cl. Sci. 3, 1964, pp. 137–160
  • [13] Menita Carozza, Nicola Fusco and Giuseppe Mingione “Partial regularity of minimizers of quasiconvex integrals with subquadratic growth.” In Ann. Mat. Pura Appl. (4) 175 Berlin/Heidelberg: Springer, 1998, pp. 141–164
  • [14] Menita Carozza and Antonia Passarelli di Napoli “A regularity theorem for minimisers of quasiconvex integrals: the case 1 < p < 2.” In Proc. Roy. Soc. Edinburgh Sect. A 126. no. 6, 1996, pp. 1181–1199
  • [15] Chuei Yee Chen and Jan Kristensen “On coercive variational integrals.” In Nonlinear Anal. 153, 2017, pp. 213–229
  • [16] Sergio Conti, Daniel Faraco and Francesco Maggi “A new approach to counterexamples to L1\mathrm{L}^{1} estimates: Korn’s inequality, geometric rigidity, and regularity for gradients of separately convex functions.” In Arch. Ration. Mech. Anal. 175. no. 2, 2005, pp. 287–300
  • [17] Sergio Conti and Franz Gmeineder 𝒜\mathcal{A}-Quasiconvexity and Partial Regularity.”, 2020 arXiv:2009.13820
  • [18] Ennio De Giorgi “Sulla differenziabilità e l’analiticità delle estremali degli integrali multipli regolari.” In Mem. Accad. Sci. Torino. Cl. Sci. Fis. Mat. Nat. (3) 3, 1957, pp. 25–43
  • [19] Ennio De Giorgi “Frontiere orientate di misura minima.”, Sem. Mat. Scuola Norm. Sup. Pisa: Editrice Tecnico Scientifica, 1961
  • [20] Ennio De Giorgi “Un esempio di estremali discontinue per un problema variazionale di tipo ellittico.” In Boll. Unione Mat. Ital., IV. Ser. 1 Bologna: Zanichelli, 1968, pp. 135–137
  • [21] Guido De Philippis and Filip Rindler “On the structure of 𝒜{\mathcal{A}}-free measures and applications.” In Ann. Math. (2) 184. no. 3 Princeton, NJ: Princeton University, Mathematics Department, 2016, pp. 1017–1039
  • [22] Lars Diening, Daniel Lengeler, Bianca Stroffolini and Anna Verde “Partial regularity for minimizers of quasi-convex functionals with general growth.” In SIAM J. Math. Anal. 44. no. 5, 2012, pp. 3594–3616
  • [23] F. Duzaar and J.F. Grotowski “Optimal interior partial regularity for nonlinear elliptic systems: The method of A-harmonic approximation.” In Manuscr. Math. 103, 2000, pp. 267–298
  • [24] Frank Duzaar, Joseph F. Grotowski and Manfred Kronz “Regularity of almost minimizers of quasi-convex variational integrals with subquadratic growth” In Ann. Mat. Pura Appl. 184. no. 4 Springer, Berlin/Heidelberg; Fondazione Annali di Matematica Pura ed Applicata c/o Dipartimento di Matematica “U. Dini”, Firenze, 2005, pp. 421–448
  • [25] Frank Duzaar and Giuseppe Mingione “Harmonic type approximation lemmas.” In J. Math. Anal. Appl. 352. no. 1, 2009, pp. 301–335
  • [26] Ivar Ekeland “On the variational principle.” In J. Math. Anal. Appl. 47, 1974, pp. 324–353
  • [27] Lawrence C. Evans “Quasiconvexity and partial regularity in the calculus of variations.” In Arch. Ration. Mech. Anal. 95. no. 3 Berlin/Heidelberg: Springer, 1986, pp. 227–252
  • [28] Irene Fonseca and Stefan Müller 𝒜\mathcal{A}-quasiconvexity, lower semicontinuity, and Young measures.” In SIAM J. Math. Anal. 30. no. 6 Philadelphia, PA: Society for IndustrialApplied Mathematics (SIAM), 1999, pp. 1355–1390
  • [29] Jens Frehse “Una generalizzazione di un controesempio di De Giorgi nella teoria delle equazioni ellitiche.” In Boll. Un. Mat. Ital. 4. no. 3, 1970, pp. 998–1002
  • [30] Enrico Giusti “Direct methods in the calculus of variations.” Singapore: World Scientific, 2003
  • [31] Franz Gmeineder “The regularity of minima for the Dirichlet problem on BD” In Arch. Ration. Mech. Anal. 237. no. 3 Springer, Berlin/Heidelberg, 2020, pp. 1099–1171
  • [32] Franz Gmeineder “Partial Regularity for Symmetric Quasiconvex Functionals on BD” In J. Math. Pures Appl. 145. no. 9, 2021, pp. 83–129
  • [33] Franz Gmeineder and Jan Kristensen “Partial Regularity for BV Minimizers” In Arch. Ration. Mech. Anal. 232. no. 3 Springer, 2019, pp. 1429–1473
  • [34] Franz Gmeineder and Bogdan Raiţă “Embeddings for 𝔸\mathbb{A}-weakly differentiable functions on domains.” In J. Funct. Anal. 277. no. 12 Amsterdam: Elsevier, 2019
  • [35] David Hilbert “Mathematische Probleme. Vortrag, gehalten auf dem internationalen Mathematiker-Kongreß zu Paris 1900” In Nachrichten von der Königlichen Gesellschaft der Wissenschaften zu Göttingen. Mathematisch-Physikalische Klasse 3, 1900, pp. 253–297
  • [36] Agnieszka Kałamajska “Pointwise multiplicative inequalities and Nirenberg type estimates in weighted Sobolev spaces.” In Stud. Math. 108. no. 3 Warsaw: Polish Academy of Sciences (Polska Akademia Nauk - PAN), Institute of Mathematics (Instytut Matematyczny), 1994, pp. 275–290
  • [37] Bernd Kirchheim and Jan Kristensen “Automatic convexity of rank-1 convex functions.” In C. R. Math. Acad. Sci. Paris 349 no. 7–8, 2011, pp. 407–409
  • [38] J. Kristensen and F. Rindler “Relaxation of signed integral functionals in BV\mathrm{BV}.” In Calc. Var. Partial Differential Equations 37. no. 1–2, 2010, pp. 29–62
  • [39] Evans L.C. and Gariepy R.F. “Measure Theory and Fine Properties of Functions” In Studies in Advanced Mathematics CRC Press, 1992
  • [40] S.. Mihlin “On the multipliers of Fourier integrals.” In Dokl. Akad. Nauk SSSR (N.S.) 109, 1956, pp. 701–703
  • [41] Giuseppe Mingione “Regularity of minima: an invitation to the dark side of the calculus of variations.” In Appl. Math., Praha 51. no. 4 Institute of Mathematics, Prague: Czech Academy of Sciences, 2006, pp. 355–425
  • [42] Giuseppe Mingione “Singularities of minima: a walk on the wild side of the calculus of variations” In J. Glob. Optim. 40. no. 1–3 Springer US, New York, NY, 2008, pp. 209–223
  • [43] Jr. Morrey “Quasi-convexity and the lower semicontinuity of multiple integrals.” In Pacific J. Math. 2, 1952, pp. 25–53
  • [44] Jürgen Moser “A new proof of De Giorgi’s theorem concerning the regularity problem for elliptic differential equations.” In Comm. Pure Appl. Math. 13, 1960, pp. 457–468
  • [45] John F. Nash “Continuity of solutions of parabolic and elliptic equations.” In Amer. J. Math. 80 Baltimore, MD: Johns Hopkins University Press, 1958, pp. 931–954
  • [46] Jindřich Nečas “Example of an irregular solution to a nonlinear elliptic system with analytic coefficients and conditions for regularity.” In Theor. Nonlin. Oper., Constr. Aspects, Proc. 4th Int. Summer School, 1975 Berlin: Akademie-Verlag, 1977, pp. 197–206
  • [47] Donald Ornstein “A non-inequality for differential operators in the L1L_{1} norm.” In Arch. Ration. Mech. Anal. 11 Berlin/Heidelberg: Springer, 1962, pp. 40–49
  • [48] Thomas Schmidt “Partial regularity for degenerate variational problems and image restoration models in BV.” In Indiana Univ. Math. J. 63. no. 1 Bloomington, IN: Indiana University, Department of Mathematics, 2014, pp. 213–279
  • [49] K.. Smith “Formulas to represent functions by their derivatives” In Math. Ann. 188 Springer, Berlin/Heidelberg, 1970, pp. 53–77
  • [50] D.. Spencer “Overdetermined systems of linear partial differential equations” In Bull. Am. Math. Soc. 75 American Mathematical Society (AMS), Providence, RI, 1969, pp. 179–239
  • [51] Jean Van Schaftingen “Limiting Sobolev inequalities for vector fields and canceling linear differential operators.” In J. Eur. Math. Soc. (JEMS) 15. no. 3 Zurich: European Mathematical Society (EMS) Publishing House, 2013, pp. 877–921