Processing math: 87%
Research article Special Issues

Quasiconvex bulk and surface energies with subquadratic growth

  • We establish partial Hölder continuity of the gradient for equilibrium configurations of vectorial multidimensional variational problems, involving bulk and surface energies. The bulk energy densities are uniformly strictly quasiconvex functions with p-growth, 1<p<2, without any further structure conditions. The anisotropic surface energy is defined by means of an elliptic integrand Φ not necessarily regular.

    Citation: Menita Carozza, Luca Esposito, Lorenzo Lamberti. Quasiconvex bulk and surface energies with subquadratic growth[J]. Mathematics in Engineering, 2025, 7(3): 228-263. doi: 10.3934/mine.2025011

    Related Papers:

    [1] Mirco Piccinini . A limiting case in partial regularity for quasiconvex functionals. Mathematics in Engineering, 2024, 6(1): 1-27. doi: 10.3934/mine.2024001
    [2] Donatella Danielli, Rohit Jain . Regularity results for a penalized boundary obstacle problem. Mathematics in Engineering, 2021, 3(1): 1-23. doi: 10.3934/mine.2021007
    [3] Daniela De Silva, Ovidiu Savin . Uniform density estimates and Γ-convergence for the Alt-Phillips functional of negative powers. Mathematics in Engineering, 2023, 5(5): 1-27. doi: 10.3934/mine.2023086
    [4] Miyuki Koiso . Stable anisotropic capillary hypersurfaces in a wedge. Mathematics in Engineering, 2023, 5(2): 1-22. doi: 10.3934/mine.2023029
    [5] Aleksandr Dzhugan, Fausto Ferrari . Domain variation solutions for degenerate two phase free boundary problems. Mathematics in Engineering, 2021, 3(6): 1-29. doi: 10.3934/mine.2021043
    [6] Claudia Lederman, Noemi Wolanski . Lipschitz continuity of minimizers in a problem with nonstandard growth. Mathematics in Engineering, 2021, 3(1): 1-39. doi: 10.3934/mine.2021009
    [7] Daniela De Silva, Giorgio Tortone . Improvement of flatness for vector valued free boundary problems. Mathematics in Engineering, 2020, 2(4): 598-613. doi: 10.3934/mine.2020027
    [8] Morteza Fotouhi, Andreas Minne, Henrik Shahgholian, Georg S. Weiss . Remarks on the decay/growth rate of solutions to elliptic free boundary problems of obstacle type. Mathematics in Engineering, 2020, 2(4): 698-708. doi: 10.3934/mine.2020032
    [9] Giovanni Scilla, Bianca Stroffolini . Partial regularity for steady double phase fluids. Mathematics in Engineering, 2023, 5(5): 1-47. doi: 10.3934/mine.2023088
    [10] Manuel Friedrich . Griffith energies as small strain limit of nonlinear models for nonsimple brittle materials. Mathematics in Engineering, 2020, 2(1): 75-100. doi: 10.3934/mine.2020005
  • We establish partial Hölder continuity of the gradient for equilibrium configurations of vectorial multidimensional variational problems, involving bulk and surface energies. The bulk energy densities are uniformly strictly quasiconvex functions with p-growth, 1<p<2, without any further structure conditions. The anisotropic surface energy is defined by means of an elliptic integrand Φ not necessarily regular.



    Let us consider a functional F with density energy discontinuous through an interface A, inside an open bounded subset Ω of Rn, of the form

    F(v,A):=Ω(F(Dv)+1AG(Dv))dx+P(A,Ω), (1.1)

    where vW1,ploc(Ω;RN), F,G:Rn×NR are C2-integrands, AΩ and P(A,Ω) stands for the perimeter of the set A in Ω. Assume that these integrands satisfy the following growth and uniformly strict p-quasiconvexity conditions, for p>1 and positive constants 1,2,L1,L2:

    0F(ξ)L1(1+|ξ|2)p2, (F1)
    ΩF(ξ+Dφ)dxΩ(F(ξ)+1|Dφ|2(1+|Dφ|2)p22)dx, (F2)
    0G(ξ)L2(1+|ξ|2)p2, (G1)
    ΩG(ξ+Dφ)dxΩ(G(ξ)+2|Dφ|2(1+|Dφ|2)p22)dx, (G2)

    for every ξRn×N and φC10(Ω;RN).

    Existence and regularity results have been obtained initially in the scalar case (N=1) in [4,5,10,17,22,23,24,25,26,29,34,35,36]. In the vectorial case (N>1), the authors in [11] proved the existence of local minimizers of (1.1), for any p>1 under the quasiconvexity assumption quoted above. In the same paper, the C1,α partial regularity is proved for minimal configurations outside a negligible set, in the quadratic case p=2.

    In [9] the same regularity result has been established in the general case p2, also addressing anisotropic surface energies. Almgren was the first to study such surface energies in his celebrated paper [3] (see also [8,21,27,39,40] for subsequent results). This kind of energies arises in many physical contexts such as the formation of crystals (see [6,7]), liquid drops (see [16,28]), capillary surfaces (see [18,19]) and phase transitions (see [33]).

    In this paper, we consider the same functional as in [9], given by

    I(v,A):=Ω(F(Dv)+1AG(Dv))dx+ΩAΦ(x,νA(x))dHn1(x), (1.2)

    in the case of sub-quadratic growth, 1<p<2. We achieve analogous regularity results as those established in [9], thereby completing the answer to the problem for all p>1.

    In this setting AΩ is a set of finite perimeter, uW1,ploc(Ω;RN), 1A is the characteristic function of the set A, A denotes the reduced boundary of A in Ω and νA is the measure-theoretic outer unit normal to A. Moreover, Φ is an elliptic integrand on Ω (see Definition 2.8), i.e., Φ:¯Ω×Rn[0,] is lower semicontinuous, Φ(x,) is convex and positively one-homogeneous, Φ(x,tν)=tΦ(x,ν) for every t0, and the anisotropic surface energy of a set A of finite perimeter in Ω is defined as follows

    Φ(A;B):=BAΦ(x,νA(x))dHn1(x), (1.3)

    for every Borel set BΩ. The further assumption

    1ΛΦ(x,ν)Λ, (1.4)

    with Λ>1, allows to compare the surface energy introduced in (1.3) with the usual perimeter. Let us recall that in the vectorial setting, as in the previously cited papers, the regularity we can expect for the gradient of the minimal deformation u:ΩRN, (N>1), even in absence of a surface term, is limited to a partial regularity result.

    Definition 1.1. We say that a pair (u,E) is a local minimizer of I in Ω, if for every open set UΩ and every pair (v,A), where vuW1,p0(U;RN) and A is a set of finite perimeter with AΔEU, we have

    U(F(Du)+1EG(Du))dx+Φ(E;U)U(F(Dv)+1AG(Dv))dx+Φ(A;U).

    Existence and regularity results for local minimizers of integral functionals with uniformly strict p-quasiconvex integrand, also in the non autonomous case, have been widely investigated (see [1,2,12,13,14,15,30,31,32,38]).

    Regarding the functional (1.2), the existence of local minimizers is guaranteed by the following theorem, proved in [9].

    Theorem 1.2. Let p>1 and assume that (F1), (F2), (G1), and (G2) hold. Then, if vW1,ploc(Ω;RN) and AΩ is a set of finite perimeter in Ω, for every sequence {(vk,Ak)}kN such that {vk} weakly converges to v in W1,ploc(Ω;RN) and 1Ak strongly converges to 1A in L1loc(Ω), we have

    I(v,A)lim infkI(vk,Ak).

    In particular, I admits a minimal configuration (u,1E)W1,ploc(Ω;RN)×BVloc(Ω;[0,1]).

    We emphasize that, in particular, the previous theorem implies the semicontinuity of the anisotropic perimeter functional (1.3) (see [9] Proposition 3.2 for the proof).

    In this paper, we obtain a C1,α regularity result for local minimizers of (1.2) in the case of sub-quadratic growth, 1<p<2. If we further assume a closeness condition on F and G (see assumption (H) in Theorem 1.3), we prove that uC1,γ(Ω1) for every γ(0,1p) on a full measure set Ω1Ω. Furthermore, we do not assume any regularity on Φ in order to get the regularity of u.

    Our main theorem is the following:

    Theorem 1.3. Let (u,E) be a local minimizer of I. Let the bulk density energies F and G satisfy (F1), (F2), (G1), and (G2), with 1<p<2, and let the surface energy Φ be of general type (1.3) with Φ satisfying (1.4). Assume in addition that

    L21+2<1, (H)

    then there exists an open set Ω1Ω of full measure such that uC1,γ(Ω1;RN) for every γ(0,1p).

    In the case where hypothesis (H) does not hold, it is still possible to establish a partial C1,β regularity result. To avoid redundancy and overlap, we have chosen to present this result in the form of a remark. Nevertheless, throughout the paper, we will provide some sketches and insights into the proof in this case as well.

    Remark 1.4. We remark that if (u,E) is a local minimizer of I with the bulk density energies F and G satisfying (F1), (F2), (G1), (G2), 1<p<2, and the surface energy Φ of general type (1.3) satisfying (1.4), then there exist an exponent β(0,1) and an open set Ω0Ω with full measure such that uC1,β(Ω0;RN).

    The proof of the Theorem 1.3 is based on a blow-up argument aimed to establish a decay estimate for the excess function

    U(x0,r):=Br(x0)|V(Du)V((Du)x0,r)|2dx+P(E,Br(x0))rn1+r,

    where

    V(ξ)=(1+|ξ|2)(p2)/4ξ,ξRk.

    To this aim, we use a comparison argument between the blow-up sequence vh at small scale in the balls Brh(xh) and the solution v of a suitable linearized system. The challenging part of the argument, as usual, is to prove that the 'good' decay estimates available for the function v (see Proposition 2.1), are inherited by the vh as h.

    To achieve this result, the main tool is a Caccioppoli type inequality that we prove for minimizers of perturbed rescaled functionals (see (3.16)) involving the function V(Dvh) and the perimeter of the rescaled minimal set Eh. The Caccioppoli inequality combined with the Sobolev-Poincaré inequality will lead us to a contradiction (see Step 6 of Proposition 3.1). In this final step, the issue to deal with the function V(Du) in the sub-quadratic case, is overcome by using a suitable Sobolev Poincaré inequality involving V(Du) (see Theorem 2.6), whose proof is due to [12].

    Let Ω be a bounded open set in Rn, n2, u:ΩRN, N>1. We denote by Br(x):={yRn:|yx|<r} the open ball centered at xRn of radius r>0, Sn1 represents the unit sphere of Rn, c a generic constant that may vary.

    For Br(x0)Rn and uL1(Br(x0);RN) we denote

    (u)x0,r:=Br(x0)u(x)dx

    and we will omit the dependence on the center when it is clear from the context. We denote by || the standard Euclidean norm, defined as

    |ξ|=(Nα=1ni=1(ξαi)2)1/2,

    for every ξRn×N.

    If F:Rn×NR is sufficiently differentiable, we write

    DF(ξ)η:=Nα=1ni=1Fξαi(ξ)ηαi and D2F(ξ)ηη:=Nα,β=1ni,j=1Fξαiξβj(ξ)ηαiηβj,

    for ξ, ηRn×N.

    It is well known that for quasiconvex C1 integrands the assumptions (F1) and (G1) yield the upper bounds

    |DF(ξ)|c1L1(1+|ξ|2)p12and|DG(ξ)|c2L2(1+|ξ|2)p12 (2.1)

    for all ξRn×N, with c1 and c2 constants depending only on p (see [32, Lemma 5.2] or [38]).

    Furthermore, if F and G are C2, then (F2) and (G2) imply the following strong Legendre-Hadamard conditions

    Nα,β=1ni,j=1Fξαiξβj(Q)λiλjμαμβc3|λ|2|μ|2andNα,β=1ni,j=1Gξαiξβj(Q)λiλjμαμβc4|λ|2|μ|2,

    for all QRn×N, λRn, μRN, where c3=c3(p,1) and c4=c4(p,2) are positive constants (see [32, Proposition 5.2]). Throughout the paper, we frequently employ the Einstein summation convention. We will need the following quite standard regularity result (see [12] for its proof).

    Proposition 2.1. Let vW1,1(Ω;RN) be such that

    ΩQijαβDivαDjφβdx=0,

    for every φCc(Ω;RN), where Q={Qijαβ} is a constant matrix satisfying |Qijαβ|L and the strong Legendre-Hadamard condition

    Qijαβλiλjμαμβ|λ|2|μ|2,

    for all λRn, μRN and for some positive constants ,L>0. Then vC and, for any BR(x0)Ω, the following estimate holds

    supBR/2|Dv|cRnBR|Dv|dx,

    where c=c(n,N,,L)>0.

    We assume that 1<p<2 and we refer to the auxiliary function

    V(ξ)=(1+|ξ|2)(p2)/4ξ,ξRk, (2.2)

    whose useful properties are listed in the following lemma (see [12] for the proof).

    Lemma 2.2. Let 1<p<2 and let V:RkRk be the function defined in (2.2), then for any ξ,ηRk and t>0 the following inequalities hold:

    (i) 2(p2)/4min{|ξ|,|ξ|p/2}|V(ξ)|min{|ξ|,|ξ|p/2},

    (ii) |V(tξ)|max{t,tp/2}|V(ξ)|,

    (iii) |V(ξ+η)|c(p)[|V(ξ)|+|V(η)|],

    (iv) p2|ξη|(1+|ξ|2+|η|2)(2p)/4|V(ξ)V(η)|c(k,p)|ξη|,

    (v) |V(ξ)V(η)|c(k,p)|V(ξη)|,

    (vi) |V(ξη)|c(p,M)|V(ξ)V(η)|, if |η|M.

    We will also use the following iteration lemma (see [32, Lemma 6.1]).

    Lemma 2.3. Let 0<ρ<R and let ψ:[ρ,R]R be a bounded non negative function. Assume that for all ρs<tR we have

    ψ(s)ϑψ(t)+A+B(st)α+C(st)β

    where ϑ[0,1), α>β>0 and A,B,C0 are constants. Then there exists a constant c=c(ϑ,α)>0 such that

    ψ(ρ)c(A+B(Rρ)α+C(Rρ)β).

    An easy extension of this result can be obtained by replacing homogeneity with condition (ⅱ) of Lemma 2.2.

    Lemma 2.4. Let R>0 and let ψ:[R/2,R][0,+) be a bounded function. Assume that for all R/2s<tR we have

    ψ(s)ϑψ(t)+ABR|V(h(x)ts)|2dx+B,

    where hLp(Br), A,B>0, and 0<ϑ<1. Then there exists a constant c(ϑ)>0 such that

    ψ(R2)c(ϑ)(ABR|V(h(x)R)|2dx+B).

    Given a C1 function f:RkR, QRk and λ>0, we set

    fQ,λ(ξ):=f(Q+λξ)f(Q)Df(Q)λξλ2,ξRk.

    In the next sections we will use the following lemma about the growth of fQ,λ and DfQ,λ.

    Lemma 2.5. Let 1<p<, and let f be a C2(Rk) function such that

    |f(ξ)|L(1+|ξ|p)and|Df(ξ)|L(1+|ξ|2)(p1)/2, (2.3)

    for any ξRk and for some L>0. Then for every M>0 there exists a constant c=c(p,L,M)>0 such that, for every QRk, |Q|M and λ>0, it holds

    |fQ,λ(ξ)|c(1+|λξ|2)(p2)/2|ξ|2and|DfQ,λ(ξ)|c(1+|λξ|2)(p2)/2|ξ|, (2.4)

    for all ξRk.

    Proof. Applying Taylor's formula for every ξRk, there exists θ[0,1] such that,

    fQ,λ(ξ)=12D2f(Q+θλξ)ξξ,
    DfQ,λ(ξ)=1λ(Df(Q+λξ)Df(Q))=10D2f(Q+sλξ)ξds.

    If we denote KM:=max{|D2f(ξ)|:|ξ|M+1}, we have

    |fQ,λ(ξ)|12KM|ξ|2,|DfQ,λ(ξ)|KM|ξ|,if |λξ|1. (2.5)

    On the other hand, using growth condition (2.3) and the definitions of fQ,λ and DfQ,λ, we get

    |fQ,λ(ξ)|c(p,L,M)λp2|ξ|p,|DfQ,λ(ξ)|c(L,M)λp2|ξ|p1,whereas |λξ|>1. (2.6)

    We get the result by combining (2.5) and (2.6).

    A fundamental tool in order to handle the subquadratic case is the following Sobolev-Poincaré inequality related to the function V, as established in Theorem 2.4 of [12].

    Theorem 2.6. If 1<p<2, there exist 2/p<α<2 and σ>0 such that if uW1,p(B3R(x0),RN), then

    (BR(x0)|V(uuxo,RR)|2(1+σ)dx)12(1+σ)C(B3R(x0)|V(Du)|αdx)1α, (2.7)

    where the positive constant C=C(n,N,p) is independent of R and u.

    We remark that a sharper version of Theorem 2.6 can be found in [20].

    In the remaining part of this section, we recall some elementary definitions and well-known properties of sets of finite perimeter. We introduce the notion of anisotropic perimeter as well.

    Given a set ERn and t[0,1], we define the set of points of E of density t as

    E(t)={xRn: |EBr(x)|=t|Br(x)|+o(rn) as r0+}.

    Let U be an open subset U of Rn. A Lebesgue measurable set ERn is said to be a set of locally finite perimeter in U if there exists a Rn-valued Radon measure μE on U (called the Gauss-Green measure of E) such that

    Eϕ dx=UϕdμE,ϕC1c(U).

    Moreover, we denote the perimeter of E relative to GU by P(E,G)=|μE|(G).

    It is well known that the support of μE can be characterized by

    sptμE={xU:0<|EBr(x)|<ωnrn,r>0}UE, (2.8)

    (see [37, Proposition 12.19]). If E is of finite perimeter in U, the reduced boundary EU of E is the set of those xU such that

    νE(x):=limr0+μE(Br(x))|μE|(Br(x))

    exists and belongs to Sn1. The essential boundary of E is defined as eE:=Rn(E(0)E(1)). It is well-understood that

    EUeEsptμEUE,U¯E=sptμE.

    Furthermore, Federer's criterion (see for instance [37, Theorem 16.2]) ensures that

    Hn1((UeE)E)=0.

    By De Giorgi's rectifiability theorem (see [37, Theorem 15.9]), the Gauss-Green measure μE is completely characterized as follows:

    μE=νEHn1E,|μE|=Hn1E.

    The equality holds in the class of Borel sets compactly contained in U. Here, we have denoted μE(F)=μ(EF), for any subset F of Rn.

    Remark 2.7 (Minimal topological boundary). If ERn is a set of locally finite perimeter in U and FRn is such that |(EΔF)U|=0, then F is a set of locally finite perimeter in U and μE=μF. In the rest of the paper, the topological boundary E must be understood by considering the suitable representative of E in order to have that ¯E=EU. We will choose E(1) as representative of E. With such a choice it can be easily verified that

    UE={xU:0<|EBr(x)|<ωnrn,r>0}.

    Therefore, by (2.8),

    ¯E=sptμE=EU.

    In what follows, we give the definition of anisotropic surface energies and we recall some properties.

    Definition 2.8 (Elliptic integrands). Given an open subset Ω of Rn, Φ:¯Ω×Rn[0,] is said to be an elliptic integrand on Ω if it is lower semicontinuous, with Φ(x,) convex and positively one-homogeneous for any x¯Ω, i.e., Φ(x,tν)=tΦ(x,ν) for every t0. Accordingly, the anisotropic surface energy of a set E of finite perimeter in Ω is defined as

    Φ(E;B):=BEΦ(x,νE(x))dHn1(x), (2.9)

    for every Borel set BΩ.

    In order to prove the regularity of minimizers of anisotropic surface energies, it is well known that a Ck-dependence of the integrand Φ on the variable ν, and a continuity condition with respect to the variable x, must be assumed (see the seminal paper [3]). In fact, one more condition is essential, that is a non-degeneracy type condition for the integrand Φ. More precisely, we have to assume that there exists a constant Λ>1 such that

    1ΛΦ(x,ν)Λ, (2.10)

    for any xΩ and νSn1. We emphasize that (2.10) is the only assumption we make for the elliptic integrand Φ. We observe that, if the elliptic integrand Φ satisfies the previous condition, then the anisotropic surface energy (2.9) satisfies the following comparability condition to the perimeter:

    1ΛHn1(BE)Φ(E;B)ΛHn1(BE),

    for any set E of finite perimeter in Ω and any Borel set BΩ.

    A useful relation is given by proposition below proved in [9].

    Proposition 2.9. Let URn be an open set and let E,FU be two sets of finite perimeter in U. It holds that

    Φ(EF;U)=Φ(E;F(0))+Φ(F;E(0))+Φ(E;{νE=νF}).

    In this section we prove decay estimates for local minimizers u of the functionals (1.2), see Definition 1.1, by using a well-known blow-up technique involving a suitable excess function. We consider the bulk excess function defined as

    U(x0,r):=Br(x0)|V(Du)V((Du)x0,r)|2dx, (3.1)

    for Br(x0)Ω.

    When the assumption (H) is in force, we refer to the following "hybrid" excess:

    U(x0,r):=U(x0,r)+P(E,Br(x0))rn1+r.

    Proposition 3.1. Let (u,E) be a local minimizer of the functional I in (1.2) and let the assumptions (F1), (F2), (G1), (G2), and (H) hold. For every M>0 and every 0<τ<14, there exist two constants ε0=ε0(τ,M)>0 and C=C(n,p,1,2,L1,L2,Λ,M)>0 such that if for some ball Br(x0)Ω the following condition hold: |(Du)x0,r|MandU(x0,r)ε0, then

    U(x0,τr)CτU(x0,r). (3.2)

    Proof. In order to prove (3.2), we argue by contradiction. Let M>0 and τ(0,1/4) be such that for every hN, C>0, there exists a ball Brh(xh)Ω such that

    |(Du)xh,rh|M,U(xh,rh)0 (3.3)

    and

    U(xh,τrh)CτU(xh,rh). (3.4)

    The constant C will be determined later. We remark that we can confine ourselves to the case in which EBrh(xh), since the case in which Brh(xh)ΩE is well known, being U=U+r.

    Step 1. Blow-up. We set λ2h:=U(xh,rh), Ah:=(Du)xh,rh, ah:=(u)xh,rh, and we define

    vh(y):=u(xh+rhy)ahrhAhyλhrh,yB1. (3.5)

    One can easily check that (Dvh)0,1=0 and (vh)0,1=0. We set

    Eh:=Exhrh,Eh:=ExhrhB1.

    By using (ii) and (vi) of Lemma 2.2, we deduce

    B1|V(Dvh(y))|2dy=Brh(xh)|V(Du(x)(Du)xh,rhλh)|2dxc(M)λ2hBrh(xh)|V(Du(x))V((Du)xh,rh)|2dx=c(M)λ2hB1|V(Du(xh+rhy))V(Ah))|2dy.

    Then, since the integral in the last expression appear in the definition of the excess U(xh,rh),

    λ2h=U(xh,rh)=B1|V(Du(xh+rhy))V(Ah)|2dy+P(E,Brh(xh))rn1h+rh,

    it follows that rh0, P(Eh,B1)0, and

    rhλ2h1,B1|V(Dvh(y))|2dyc(M),P(Eh,B1)λ2h1. (3.6)

    Therefore, by (3.3) and (3.6), using also (ⅰ) of Lemma 2.2 and Poincaré inequality, we deduce that there exist a (not relabeled) subsequence of {vh}hN, ARn×N and vW1,p(B1;RN), such that

    vhvweakly in W1,p(B1;RN),vhvstrongly in Lp(B1;RN),AhA,λhDvh0in Lp(B1;Rn×N) and pointwise a.e. in B1, (3.7)

    where we have used the fact that (vh)0,1=0. Moreover, by (3.3) and (3.6), we have that for every 0ϵ<1n1

    limh(P(Eh,B1))nn1λ2(1+ϵ)hlimhP(Eh,B1)1n1ϵlim suphP(Eh,B1)1+ϵλ2(1+ϵ)h=0, (3.8)

    where we have used (3.6) and the choice of ϵ<1n1 in the last inequalities. Therefore, by the relative isoperimetric inequality,

    limhmin{|Eh|λ2(1+ϵ)h,|B1Eh|λ2(1+ϵ)h}c(n)limh(P(Eh,B1))nn1λ2(1+ϵ)h=0. (3.9)

    In the sequel the proof will proceed differently depending on

    min{|Eh|,|B1Eh|}=|Eh| or min{|Eh|,|B1Eh|}=|B1Eh|.

    The first case is easier to handle. To understand the reason, let us introduce the expansions of F and G around Ah as follows:

    Fh(ξ):=F(Ah+λhξ)F(Ah)DF(Ah)λhξλ2h,Gh(ξ):=G(Ah+λhξ)G(Ah)DG(Ah)λhξλ2h, (3.10)

    for any ξRn×N. In the first case the suitable rescaled functional to consider in the blow-up procedure is the following:

    Ih(w):=B1[Fh(Dw)dy+1EhGh(Dw)]dy. (3.11)

    We claim that vh satisfies the minimality inequality

    Ih(vh)Ih(vh+ψ)+1λhB11EhDG(Ah)Dψ(y)dy, (3.12)

    for any ψW1,p0(B1;RN). Indeed, using the minimality of (u,E) with respect to (u+φ,E), for φW1,p0(Brh(xh);RN), where φ is defined by the change of variable y=xxhrh, setting φ(x):=λhrhψ(xxhrh), it holds that

    B1[(Fh(Dvh(y))+1EhGh(Dvh(y))]dyB1[Fh(Dvh(y)+Dψ(y))+1EhGh(Dvh(y)+Dψ(y))]dy+1λhB11EhDG(Ah)Dψ(y)dy,

    and (3.12) follows by the definition of Ih in (3.11).

    In the second case, the suitable rescaled functional to consider in the blow-up procedure is

    Hh(w):=B1[Fh(Dw)+Gh(Dw)]dy.

    We claim that

    Hh(vh)Hh(vh+ψ)+L2λ2h(B1Eh)suppψ(1+|Ah+λhDvh|2)p2dy, (3.13)

    for all ψW1,p0(B1;RN). Indeed, the minimality of (u,E) with respect to (u+φ,E), for φW1,p0(Brh(xh);RN), implies that

    Brh(xh)(F+G)(Du)dx=Brh(xh)[F(Du)+1EG(Du)]dx+Brh(xh)EG(Du)dxBrh(xh)[F(Du+Dφ)+1EG(Du+Dφ)]dx+Brh(xh)EG(Du)dx=Brh(xh)(F+G)(Du+Dφ)dx+Brh(xh)E[G(Du)G(Du+Dφ)]dxBrh(xh)(F+G)(Du+Dφ)dx+(Brh(xh)E)suppφG(Du)dx, (3.14)

    where we used that the last integral vanishes outside the support of φ and that G0. Using the change of variable x=xh+rhy in the previous formula, we get

    B1(F+G)(Du(xh+rhy))dyB1(F+G)(Du(xh+rhy)+Dφ(xh+rhy))dy+(B1Eh)suppψG(Du(xh+rhy))dy,

    or, equivalently, using the definitions of vh,

    B1(F+G)(Ah+λhDvh)dyB1(F+G)(Ah+λh(Dvh+Dψ))dy+(B1Eh)suppψG(Ah+λhDvh)dy,

    where ψ(y):=φ(xh+rhy)λhrh, for yB1. Therefore, setting

    Hh:=Fh+Gh,

    by the definitions of Fh and Gh in (3.10) and using the assumption (G1), we have that

    B1Hh(Dvh)dyB1Hh(Dvh+Dψ)dy+1λ2h(B1Eh)suppψG(Ah+λhDvh)dyB1Hh(Dvh+Dψ)dy+L2λ2h(B1Eh)suppψ(1+|Ah+λhDvh|2)p2dy, (3.15)

    i.e., (3.13).

    Step 2. A Caccioppoli type inequality. The key ingredient in our proof is the following Caccioppoli-type inequality. The version presented here, which involves the auxiliary function V, was used in [12] to address the subquadratic case 1<p<2. In our setting, there is also a perimeter term, which is a distinctive feature of our problem. We also draw attention to [20], where a suitable variant of the Caccioppoli-type inequality involving a modified auxiliary function V|A| was established to handle potential degeneracy of the strict quasiconvexity.

    We claim that there exists a constant c=c(n,p,1,2,L1,L2,M)>0 such that for every 0<ρ<1 there exists h0=h0(n,p,M,ρ)N such that

    Bρ2|V(λh(Dvh(Dvh)ρ2)|2dyc[Bρ|V(λh(vh(vh)ρ(Dvh)ρ2y)ρ)|2dy+P(Eh,B1)nn1], (3.16)

    for all h>h0. We divide the proof into two steps.

    Substep 2.a The case min{|Eh|,|B1Eh|}=|Eh|. We consider 0<ρ2<s<t<ρ<1 and let ηC0(Bt) be a cut off function between Bs and Bt, i.e., 0η1, η1 on Bs and |η|cts. Set ph:=(vh)Bρ, Ph:=(Dvh)Bρ2, and set

    wh(y):=vh(y)phPhy, (3.17)

    for any yB1. Proceeding similarly as in (3.10), we rescale F and G around Ah+λhPh,

    ˜Fh(ξ):=F(Ah+λhPh+λhξ)F(Ah+λhPh)DF(Ah+λhPh)λhξλ2h,˜Gh(ξ):=G(Ah+λhPh+λhξ)G(Ah+λhPh)DG(Ah+λhPh)λhξλ2h, (3.18)

    for any ξRn×N. By Lemma 2.5, two growth estimates on ˜Fh, ˜Gh and their gradients hold with some constants that depend on p,L1,L2,M (see (3.3)) and could also depend on ρ through |λhPh|. However, given ρ, we may choose h0=h0(n,p,M,ρ) large enough to have

    |λhPh|<c(n,p,M)λhρnp<1,

    for any hh0. Indeed, by (3.6) the sequence {Dvh}h is equibounded in Lp(B1), then we have

    |Ph|2nωnρn[Bρ2{|Dvh|1}|Dvh|dy+Bρ2{|Dvh|>1}|Dvh|dy]
    1+2nω1pnρnp(B1|V(Dvh)|2dy)1pc(n,p,M)ρnp,

    and so the constant in (2.4) can be taken independently of ρ.

    Set

    ψ1,h:=ηwhandψ2,h:=(1η)wh.

    By the uniformly strict quasiconvexity of ˜Fh, we have

    1λ2hBs|V(λhDwh)|2dy1Bt(1+|λhDψ1,h|2)p22|Dψ1,h|2dyBt˜Fh(Dψ1,h)dy=Bt˜Fh(Dwh)dy+Bt˜Fh(DwhDψ2,h)dyBt˜Fh(Dwh)dy=Bt˜Fh(Dwh)dyBt10D˜Fh(DwhθDψ2,h)Dψ2,hdθdy. (3.19)

    We estimate separately the two addends in the right-hand side of the previous chain of inequalities. We deal with the first addend by means of a rescaling of the minimality condition of (u,E). Using the change of variable x=xh+rhy, the fact that G0 and the minimality of (u,E) with respect to (u+φ,E) for φW1,p0(Brh(xh);RN), we have

    B1F(Du(xh+rhy))dyB1[F(Du(xh+rhy))+1EhG(Du(xh+rhy))]dyB1[F(Du(xh+rhy)+Dφ(xh+rhy))+1EhG(Du(xh+rhy)+Dφ(xh+rhy))]dy,

    i.e., by the definitions (3.5) and (3.17) of vh and wh, respectively,

    B1F(Ah+λhPh+λhDwh)dyB1[F(Ah+λhPh+λh(Dwh+Dψ))+1EhG(Ah+λhPh+λh(Dwh+Dψ))dy,

    for ψ:=φ(xh+rhy)λhrhW1,p0(B1;RN). Therefore, recalling the definitions of ˜Fh and ˜Gh in (3.18), we have that

    B1˜Fh(Dwh)dyB1[˜Fh(Dwh+Dψ)+1Eh˜Gh(Dwh+Dψ)]dy+1λ2hB11Eh[G(Ah+λhPh)+DG(Ah+λhPh)λh(Dwh+Dψ)]dy.

    Choosing φsuchthatψ=ψ1,h, the previous inequality becomes

    Bt˜Fh(Dwh)dyBt[˜Fh(DwhDψ1,h)+1Eh˜Gh(DwhDψ1,h)]dy+1λ2hB11Eh[G(Ah+λhPh)+DG(Ah+λhPh)λh(DwhDψ1,h)]dy=BtBs[˜Fh(Dψ2,h)+1Eh˜Gh(Dψ2,h)]dy+1λ2hB11Eh[G(Ah+λhPh)+DG(Ah+λhPh)λhDψ2,h]dyc(p,L1,L2,M)λ2hBtBs|V(λhDψ2,h)|2dy+c(n,p,L2,M)[|Eh|λ2h+1λhEh|Dψ2,h|dy], (3.20)

    where we have used Lemma 2.5, the second estimate in (2.1), and the fact that |Ah+λhPh|M+1. By applying Hölder's and Young's inequalities, we get

    1λhEh|Dψ2,h|dy|Eh|p1pλ2h(Eh(BtBs)|λhDψ2,h|pdy)1p1λ2h[|Eh|+Eh(BtBs)|λhDψ2,h|pdy]1λ2h[2|Eh|+Eh(BtBs){|λhDψ2,h|>1}|λDψ2,h|pdy]1λ2h[2|Eh|+BtBs|V(λhDψ2,h))|2dy].

    The previous chain of inequalities combined with (3.20) yields

    B1˜Fh(Dwh)dyc(n,p,L1,L2,M)λ2h[BtBs|V(λhDψ2,h)|2dy+|Eh|]. (3.21)

    Now we estimate the second addend in the right-hand side of (3.19). Using the upper bound on D˜Fh in Lemma 2.5,

    Bt10D˜Fh(DwhθDψ2,h)Dψ2,hdθdyc(p,L1,M)BtBs10(1+λ2h|DwhθDψ2,h|2)p22|DwhθDψ2,h||Dψ2,h|dθdy. (3.22)

    Regarding the integrand in the latest estimate, we distinguish two cases:

    Case 1: |Dψ2,h||DwhθDψ2,h|. By the definition of V, we have

    (1+λ2h|DwhθDψ2,h|2)p22|DwhθDψ2,h||Dψ2,h|λ2h|V(λh(DwhθDψ2,h)|2.

    Case 2: |DwhθDψ2,h|<|Dψ2,h|. If |Dψ2,h|<1/λh, using (i) of Lemma 2.2 we get

    (1+λ2h|DwhθDψ2,h|2)p22|DwhθDψ2,h||Dψ2,h||Dψ2,h|2λ2h|V(λhDψ2,h)|2.

    If |Dψ2,h|1/λh, using again (ⅰ) of Lemma 2.2, we deduce that

    (1+λ2h|DwhθDψ2,h|2)p22|DwhθDψ2,h||Dψ2,h|
    λp2h|DwhθDψ2,h|p1|Dψ2,h|λ2h|λhDψ2,h|pλ2h|V(λhDψ2,h)|2.

    By combining the two previous cases, we can proceed in the estimate (3.22) as follows:

    Bt10D˜Fh(DwhθDψ2,h)Dψ2,hdθdyc(p,L1,M)λ2hBtBs10D(|V(λh(DwhθDψ2,h)|2+|V(λhDψ2,h)|2)dθdyc(p,L1,M)λ2hBtBs(|V(λhDwh)|2+|V(λhDψ2,h)|2)dy. (3.23)

    Hence, combining (3.19) with (3.21) and (3.23), we obtain

    1λ2hBs|V(λhDwh)|2dyc(n,p,L1,L2,M)λ2h[BtBs(|V(λhDwh)|2+|V(λhDψ2,h)|2) dy+|Eh|].

    By the definition of ψ2,h and (ii) and (iii) of Lemma 2.2, we infer that

    1Bs|V(λhDwh)|2dy
    ˜C[BtBs(|V(λhDwh)|2+|V(λhwhts)|2)dy+|Eh|],

    for some ˜C=˜C(n,p,L1,L2,M)

    By adding ˜CBs|V(λhDwh)|2dy to both sides of the previous estimate, dividing by 1+˜C and thanks to Lemma 2.4, we deduce that

    Bρ2|V(λhDwh)|2dyc(n,p,1,L1,L2,M)(Bρ|V(λhwhρ)|2dy+|Eh|).

    Therefore, by the definition of wh, we conclude that

    Bρ2|V(λh(Dvh(Dvh)ρ2)|2dyc(n,p,1,L1,L2,M)[Bρ|V(λh(vh(vh)ρ(Dvh)ρ2y)ρ)|2dy+|Eh|]

    which, by the relative isoperimetric inequality and the hypothesis of this substep, i.e.,

    min{|Eh|,|B1Eh|}=|Eh|,

    yields the estimate (3.16).

    Substep 2.b The case min{|Eh|,|B1Eh|}=|B1Eh|.

    Let us fix 0<ρ2<s<t<ρ<1 and let ηC0(Bt), ph, Ph as in Substep 2.a and define

    wh(y):=vh(y)phPhy,yB1,

    and

    ˜Hh:=˜Fh+˜Gh.

    We remark that Lemma 2.5 can be applied to ˜Hh, that is

    |˜Hh(ξ)|c(p,L1,L2,M)(1+|λhξ|2)p22|ξ|2,ξRn×N,

    and, by the uniformly strict quasiconvexity conditions (F2) and (G2),

    B1˜Hh(ξ+Dψ)dxBt[˜Hh(ξ)+˜(1+|λhDψ|2)p22|Dψ|2]dy,ψW1,p0(B1;RN), (3.24)

    where we have set

    ˜=1+2.

    We set again

    ψ1,h:=ηwhandψ2,h:=(1η)wh.

    By the quasiconvexity condition (3.24) and since ˜Hh(0)=0, we have

    ˜λ2hBs|V(λhDwh)|2dy=˜Bs(1+|λhDwh|2)p22|Dwh|2dy˜Bt(1+|λhDψ1,h|2)p22|Dψ1,h|2dyBt˜Hh(Dψ1,h)dy=Bt˜Hh(DwhDψ2,h)dy=Bt˜Hh(Dwh)dy+Bt˜Hh(DwhDψ2,h)dyBt˜Hh(Dwh)dy=Bt˜Hh(Dwh)dyBt10D˜Hh(DwhθDψ2,h)Dψ2,hdθdy. (3.25)

    Similarly to the previous case, we estimate separately the two addends in the right-hand side of the previous chain of inequalities. Using the minimality condition (3.15) for the rescaled functions vh and recalling the definition of ˜Hh, since Dvh=Dwh+Ph, we get

    B1˜Hh(Dwh)dyB1˜Hh(Dwh+Dψ)dy+L2λ2h(B1Eh)suppψ(1+|Ah+λhPh+λhDwh|2)p2dy. (3.26)

    Choosing ψ=ψ1,h as test function in (3.26) and using the fact that ˜Hh(0)=0, we estimate

    Bt˜Hh(Dwh)dyBt˜Hh(DwhDψ1,h)dy+L2λ2hBtEh(1+|Ah+λhPh+λhDwh|2)p2dy=BtBs˜Hh(Dψ2,h)dy+L2λ2hBtEh(1+|Ah+λhPh+λhDwh|2)p2dy
    c(p,L1,L2,M)λ2hBtBs|V(λhDψ2,h)|2dy+L2λ2hBtEh(1+|Ah+λhPh+λhDwh|2)p2dy.

    We note that, since |Ah+λhPh|c(M), for every fixed ε>0 there exists a constant C=C(p,ε) such that

    (1+|Ah+λhPh+λhDwh|2)p2C(p,ε)c(M)p+(1+ε)λph|Dwh|p.

    Summarizing, we get

    Bt˜Hh(Dwh)dyc(p,L1,L2,M)λ2hBtBs|V(λhDψ2,h)|2dy+(1+ε)L2λ2hBt1{|λhDwh|1}|λhDwh|pdy+c(p,L2,M,ε)|B1Eh|λ2h. (3.27)

    Now we estimate the second addend in the right-hand side of (3.25). Using the upper bound on D˜Hh in Lemma 2.5, we obtain

    Bt10D˜Hh(DwhθDψ2,h)Dψ2,hdθdy
    c(p,L1,L2,M)BtBs10(1+λ2h|DwhθDψ2,h|2)p22|DwhθDψ2,h||Dψ2,h|dθdy.

    Proceeding exactly as in the estimate (3.23) of the step 2.a, we obtain

    Bt10D˜Hh(DwhθDψ2,h)Dψ2,hdθdyc(p,L1,L2,M)λ2hBtBs(|V(λhDwh)|2+|V(λhDψ2,h)|2)dy. (3.28)

    Inserting (3.27) and (3.28) in (3.25), we infer that

    ˜λ2hBs|V(λhDwh)|2dyc(p,L1,L2,M)λ2hBtBs(|V(λhDwh)|2+|V(λhDψ2,h)|2) dy+(1+ε)L2λ2hBt1{|λhDwh|1}|λhDwh|pdy+c(p,L2,M,ε)|B1Eh|λ2hc(p,L1,L2,M)λ2hBtBs|V(λhDwh)|2dy+c(p,M,L1,L2)λ2hBtBs|V(λhwhts)|2dy+(1+ε)L2λ2hBt|V(λhDwh)|2dy+c(p,L2,M,ε)|B1Eh|λ2h.

    Taking advantage of the hole filling technique as in the previous case, we obtain

    Bs|V(λhDwh)|2 dy(c(p,L1,L2,M)+(1+ε)L2)(c(p,M,L1,L2)+˜)Bt|V(λhDwh)|2 dy+c(p,M,L1,L2)BtBs|V(λhwhts)|2dy+c(p,L2,M,ε)|B1Eh|λ2h.

    The assumption (H) implies that there exists ε=ε(p,1,2,L2)>0 such that (1+ε)L21+2<1. Therefore we have

    c+(1+ε)L2c+˜=c+(1+ε)L2c+1+2<1.

    So, by virtue of Lemma 2.4, from the previous estimate we deduce that

    Bρ2|V(λhDwh)|2dyc(n,p,1,2,L1L2,M)(Bρ|V(λhwhρ)|2dy+|B1Eh|).

    By definition of wh and the relative isoperimetric inequality, since |B1Eh|=min{|Eh|,|B1Eh|}, we get the estimate (3.16).

    Step 3. v solves a linear system in B1.

    Let us divide the proof into two cases, depending on which one is the smallest between |Eh| and |B1Eh|.

    We divide the proof in two substeps.

    Substep 3.a The case min{|Eh|,|B1Eh|}=|Eh|.

    We claim that v solves the linear system

    B1D2F(A)DvDψdy=0,

    for all ψC10(B1;RN). Since vh satisfies (3.12), we have that

    0Ih(vh+sψ)Ih(vh)+1λhB11EhDG(Ah)sDψdy,

    for every ψC10(B1;RN) and s(0,1). Dividing by s and passing to the limit as s0, by the definition of Ih, we get (see [9] or [11, Substep 3.a])

    01λhB1(DF(Ah+λhDvh)DF(Ah))Dψdy+1λhB11EhDG(Ah+λhDvh)Dψdy. (3.29)

    We partition the unit ball as follows:

    B1=B+hBh={yB1:λh|Dvh|>1}{yB1:λh|Dvh|1}.

    By (3.6), we get

    |B+h|B+hλph|Dvh|pdyλphB1|Dvh|pdyc(n,p,M)λph. (3.30)

    We rewrite (3.29) as follows:

    01λhB+h(DF(Ah+λhDvh)DF(Ah))Dψdy+Bh10(D2F(Ah+tλhDvh)D2F(A))dtDvhDψdy+BhD2F(A)DvhDψdy+1λhB11EhDG(Ah+λhDvh)Dψdy. (3.31)

    By growth condition in (2.1) and Hölder's inequality, we get

    1λh|B+h(DF(Ah+λhDvh)DF(Ah))Dψdy|
    c(p,L1,M,Dψ)[|B+h|λh+λp2hB+h|Dvh|p1dy]
    c(n,p,L1,M,Dψ)[λp1h+λp1h(B1|Dvh|pdy)p1p(|B+h|λph)1p]
    c(n,p,L1,M,Dψ)λp1h,

    thanks to (3.3), (3.6) and (3.30). Thus

    limh1λh|B+h(DF(Ah+λhDvh)DF(Ah))Dψdy|=0. (3.32)

    By (3.3) and the definition of Bh we have that |Ah+λhDvh|M+1 on Bh. Hence we estimate

    |Bh10(D2F(Ah+tλhDvh)D2F(A))dtDvhDψdy|Bh|10(D2F(Ah+tλhDvh)D2F(A))dt||Dvh||Dψ|dy(Bh|10(D2F(Ah+tλhDvh)D2F(A))dt|pp1dy)p1pDvhLp(B1)DψL(B1)c(n,p,M,Dψ)(Bh|10(D2F(Ah+tλhDvh)D2F(A))dt|pp1dy)p1p,

    where we have used (3.6). Since, by (3.7), λhDvh0 a.e. in B1, the uniform continuity of D2F on bounded sets and the Severini-Egorov's Theorem implies that

    limh|Bh10(D2F(Ah+tλhDvh)D2F(A))dtDvhDψdy|=0. (3.33)

    Note that (3.30) yields that 1Bh1B1 in Lr(B1), for every r<. Therefore, by the weak convergence of Dvh to Dv in Lp(B1), it follows that

    limhBhD2F(A)DvhDψdy=B1D2F(A)DvDψdy. (3.34)

    By growth condition (2.1), we deduce

    1λh|B11Eh[DξG(Ah+λhDvh)Dψdy|c(p,L2)λhB11Eh(1+|Ah+λhDvh|2)p12|Dψ|dyc(p,L2,M,||Dψ||)[1λh|Eh|+λp2hEh|Dvh|p1dy]c(p,L2,M,||Dψ||)[1λh|Eh|+λp2+2ph(B1|Dvh|pdy)p1p(|Eh|λ2h)1p]c(n,p,L2,M,||Dψ||)[1λh|Eh|+λp2+2ph(|Eh|λ2h)1p],

    where we have used (3.3) and (3.6). Since min{|Eh|,|B1Eh|}=|Eh|, by (3.9), we have

    limh|Eh|λ2h=0,

    and so

    limh1λhB11EhDG(Ah+λhDvh)Dψdy=0. (3.35)

    By (3.32)–(3.35), passing to the limit as h in (3.31), we get

    B1DF(A)DvDψdy0.

    Furthermore, plugging ψ in place of ψ, we get

    B1DF(A)DvDψdy=0,

    i.e., v solves a linear system with constant coefficients.

    Substep 3.b The case min{|Eh|,|B1Eh|}=|B1Eh|.

    We claim that v solves the linear system

    B1D2(F+G)(A)DvDψdy=0,

    for all ψC10(B1;RN).

    Arguing as in (3.14) and rescaling, we have that

    B1Hh(Dvh)dyB1Hh(Dvh+sDψ)+1λ2hB1Eh[G(Ah+λhDvh)G(Ah+λhDvh+sλhDψ)]dy=B1Hh(Dvh+sDψ)dy+1λhB1Eh10DG(Ah+λhDvh+tsλhDψ)sDψdtdyB1Hh(Dvh+sDψ)dy+c(p,L2)λhB1Eh10(1+|Ah+λhDvh+tsλhDψ|2)p12s|Dψ|dtdyB1Hh(Dvh+sDψ)dy+c(p,L2,M)[1λhB1Ehs|Dψ|dy+B1Eh10λp2h|Dvh+tsDψ|p1s|Dψ|dtdy],

    for every ψC10(B1;RN) and for every s(0,1). Therefore

    0B110DHh(Dvh+sθDψ)dθsDψdy+c(p,L2,M)[1λhB1Ehs|Dψ|dy+B1Eh10λp2h|Dvh+tsDψ|p1s|Dψ|dtdy].

    Dividing by s and passing to the limit as s0, by the definition of Hh we get

    01λhB1[D(F+G)(Ah+λhDvh)DψD(F+G)(Ah)Dψ]dy+c(p,L2,M)[1λhB1Eh|Dψ|dy+B1Ehλp2h|Dvh|p1|Dψ|dy]. (3.36)

    As before, we partition B1 as follows:

    B1=B+hBh={yB1:λh|Dvh|>1}{yB1:λh|Dvh|1}.

    We rewrite (3.36) as

    01λhB+h(D(F+G)(Ah+λhDvh)D(F+G)(Ah))Dψdy+1λhBh(D(F+G)(Ah+λhDvh)D(F+G)(Ah))Dψdy+c(p,L2,M)[1λhB1Eh|Dψ|dy+B1Ehλp2h|Dvh|p1|Dψ|dy]. (3.37)

    Arguing as in (3.32), we obtain that

    limh1λh|B+h(D(F+G)(Ah+λhDvh)D(F+G)(Ah))Dψdy|=0, (3.38)

    and, as in (3.33) and (3.34),

    limh1λhBh[D(F+G)(Ah+λhDvh)D(F+G)(Ah)]Dψdy=B1D(F+G)(A)DvDψdy.

    Moreover, we have that

    1λhB1Eh|Dψ|dy+B1Ehλp2h|Dvh|p1|Dψ|dyc(p,Dψ)[|B1Eh|λh+λp2+2ph(B1|Dvh|pdy)p1p(|B1Eh|λ2h)1p]c(n,p,Dψ)[|B1Eh|λh+λp2+2ph(|B1Eh|λ2h)1p],

    where we used (3.6). Since min{|Eh|,|B1Eh|}=|B1Eh|, by (3.12), we have

    limh|B1Eh|λ2h=0,

    and we obtain

    limh[1λhB1Eh|Dψ|dy+B1Ehλp2h|Dvh|p1|Dψ|dy]=0. (3.39)

    By (3.38) and (3.39), passing to the limit as h in (3.37) we conclude that

    B1D2(F+G)(A)DvDψdy0

    and, with ψ in place of ψ, we finally get

    B1D2(F+G)(A)DvDψdy=0,

    as claimed.

    Substep 3.c. A decay estimate for Dv.

    By Proposition 2.1 and the theory of linear systems (see [30, Theorem 2.1 and Chapter 3]), we deduce in both cases that vC and there exists a constant ˜c=˜c(n,N,p,1,2,L1,L2)>0 such that

    Bτ|Dv(Dv)τ|2˜cτ2B12|Dv(Dv)12|2dx,

    for any τ(0,12). Moreover, by Proposition 2.1 again,

    B12|Dv(Dv)12|2dxsupB12|Dv|2˜c(B1|Dv|pdx)2/p.

    Observing that

    DvLp(B1)lim suphDvhLp(B1)c(n,p),

    it follows that

    Bτ|Dv(Dv)τ|2¯Cτ2, (3.40)

    for some fixed ¯C=¯C(n,N,p,1,2,L1,L2).

    Step 4. An estimate for the perimeters.

    Our aim is to show that there exists a constant c=c(n,p,L2,Λ,M)>0 such that

    P(Eh,Bτ)c[1τP(Eh,B1)nn1+rhτn+rhλph]. (3.41)

    By the minimality of (u,E) with respect to (u,˜E), where ˜E is a set of finite perimeter such that ˜EΔEBrh(xh) and Brh(xh) are the balls of the contradiction argument, we get

    Brh(xh)1EG(Du)+Φ(E;Brh(xh))Brh(xh)1˜EG(Du)+Φ(˜E;Brh(xh)).

    Using the change of variable x=xh+rhy and dividing by rn1h, we have

    rhB11EhG(Ah+λhDvh)dy+Φh(Eh;B1)rhB11˜EhG(Ah+λhDvh)dy+Φh(˜Eh;B1),

    where

    Φh(Eh;V):=VEhΦ(xh+rhy,νEh(y))dHn1(y),

    for every Borel set VΩ. Assume first that min{|B1Eh|,|Eh|}=|B1Eh|. Choosing ˜Eh:=EhBρ, we get

    Φh(Eh;B1)rhB11BρG(Ah+λhDvh)dy+Φh(˜Eh;B1). (3.42)

    By the coarea formula, the relative isoperimetric inequality, the choice of the representative E(1)h of Eh, which is a Borel set, we get

    2ττHn1(BρEh)dρ|B1Eh|c(n)P(Eh,B1)nn1.

    Therefore, thanks to Chebyshev's inequality, we may choose ρ(τ,2τ), independent of h, such that, up to subsequences, it holds

    Hn1(EhBρ)=0andHn1(BρEh)c(n)τP(Eh,B1)nn1. (3.43)

    We remark that Proposition 2.9 holds also for Φh. Thus, thanks to the choice of ρ, being Hn1(EhBρ)=0, we have that

    Φh(˜Eh;B1)=Φh(Eh;B(0)ρ)+Φh(Bρ;E(0)h)+Φh(Eh;{νEh=νBρ})=Φh(Eh;B1¯Bρ)+Φh(Bρ;E(0)h).

    By the choice of the representative of Eh (see Remark 2.7), taking into account (2.10) and the inequality in (3.43), it follows that

    Φh(˜Eh;B1)Φh(Eh;B1¯Bρ)+ΛHn1(BρE(0)h)Φh(Eh;B1¯Bρ)+ΛHn1(BρEh).Φh(Eh;B1¯Bρ)+Λc(n)τP(Eh,B1)nn1. (3.44)

    On the other hand, by (2.10) and the additivity of the measure Φh(Eh,) it holds that

    1ΛP(Eh,Bτ)Φh(Eh;Bτ)Φh(Eh;B1)Φh(Eh;B1¯Bρ), (3.45)

    since ρ>τ. Combining (3.42), (3.44) and (3.45), we obtain

    1ΛP(Eh,Bτ)Φh(Eh;B1)Φh(Eh;B1¯Bρ)Φh(˜Eh;B1)+rhB11BρG(Ah+λhDvh)dyΦh(Eh;B1¯Bρ)Λc(n)τP(Eh,B1)nn1+rhB11BρG(Ah+λhDvh)dyΛc(n)τP(Eh,B1)nn1+c(p,L2)rhB2τ(1+|Ah+λhDvh|2)p2dyΛc(n)τP(Eh,B1)nn1+c(n,p,L2,M)rhτn+c(p,L2)rhλphB2τ|Dvh|pdyΛc(n)τP(Eh,B1)nn1+c(n,p,L2,M)rhτn+c(n,p,L2)rhλph, (3.46)

    where we used (3.6). The previous estimate leads to (3.41). We reach the same conclusion if

    min{|B1Eh|,|Eh|}=|Eh|,

    choosing ˜Eh=EhBρ as a competitor set.

    Step 5. Higher integrability of vh.

    We will prove that there exist two positive constants C and δ depending on n,p,1,2,L1,L2 such that for every BrB1 it holds

    Br2|V(λhDvh)|2(1+δ)dyC[(B1|V(λhDvh)|2dy)1+δ+min{|B1Eh|,|Eh|}]. (3.47)

    We remark that, using (2.4) in Lemma 2.5 and (ⅳ) of Lemma 2.2,

    |Fh(ξ)|+|Gh(ξ)|c(p,L1,L2,M)λ2h|V(λhξ)|2,ξRn×N, (3.48)

    and

    B1Fh(Dϕ)dy1λ2hB1|V(λhDϕ)|2dy,ϕC1c(B1,RN).

    Let r>0 be such that B3rB1, r2<s<t<r and ηC1c(Bt) be such that 0η1, η=1 on Bs, |Dη|cts, for some positive constant c. We define

    ϕ1:=[vh(vh)r]η,ϕ2:=[vh(vh)r](1η).

    We deal with the case min{|Eh|,|B1Eh|}=|Eh|, the other one is similar. Using the fact that Gh0 and the minimality relation (3.12) we deduce

    1λ2hBt|V(λhDϕ1)|2dyBtFh(Dϕ1)dy=BtFh(Dvh)dy+BtBs[Fh(DvhDϕ2)Fh(Dvh)]dyIh(vh)+BtBs[Fh(DvhDϕ2)Fh(Dvh)]dyIh(ϕ2+(vh)r)+BtBs[Fh(DvhDϕ2)Fh(Dvh)]dy+1λhBtEhDG(Ah)|Dϕ1|dy.

    Then, using growth condition (3.48) and the fact that Ah is controlled by M, we conclude that

    1λ2hBt|V(λhDϕ1)|2dyc(p,L1,L2,M)λ2h[BtBs[|V(λhDϕ2)|2+|V(λhDϕ1)|2+|V(λhDvh)|2]dy+λhBtEh|Dϕ1|dy].

    By the properties of V, it holds that

    |ξ|C(p)(1+|V(ξ)|2p),ξRn×N.

    Thus, using Young's inequality, it follows that

    1λ2hBtEh|λhDϕ1|dyc(p)λ2h[|EhBt|+BtEhV(|λhDϕ1|)2pdy]c(p)λ2h[c(ε)|EhBt|+εBtEh|V(λhDϕ1)|2dy],

    for some ε>0 to be chosen. Combining the previous two chains of inequalities, we deduce that

    1λ2hBt|V(λhDϕ1)|2dyc(p,L1,L2,M)λ2h[BtBs[|V(λhDϕ2)|2+|V(λhDϕ1)|2+|V(λhDvh)|2]dy+c(ε)|EhBt|+εBtEh|V(λhDϕ1)|2dy].

    Choosing ε sufficiently small, we absorb the last integral to the left-hand side

    1λ2hBt|V(λhDϕ1)|2dyc(p,1,L1,L2,M)λ2h[BtBs[|V(λhDϕ2)|2+|V(λhDϕ1)|2+|V(λhDvh)|2]dy+|EhBt|].

    By (ii) and (iii) of Lemma 2.2, it follows

    Bs|V(λhDvh)|2dyc(p,1,L1,L2,M)[BtBs|V(λhDvh)|2dy+BtBs|V(λhvh(vh)rts)|2dy+|EhBt|].

    By applying the hole-filling technique, we add c(p,1,L1,L2,M)Bs|V(λhDvh)|2dy, and we get

    Bs|V(λhDvh)|2dyc(p,1,L1,L2,M)c(p,1,L1,L2,M)+1[Bt|V(λhDvh)|2dy+BtBs|V(λhvh(vh)rts)|2dy+|EhBt|].

    Now we can apply Lemma 2.4 and derive

    Br/2|V(λhDvh)|2dyc(p,1,L1,L2,M)[Br|V(λhvh(vh)rr)|2dy+Br1Eh dy].

    Finally, by Hölder's inequality and Theorem 2.7 we gain

    Br/2|V(λhDvh)|2dyc(p,1,L1,L2,M){[Br|V(λhvh(vh)rr)|2(1+σ)dy]11+σ+Br1Eh dy}c(p,1,L1,L2,M){[B3r|V(λhDvh)|αdy]12α+Br1Eh dy}.

    We conclude the proof by applying Gehring's lemma (see [32, Theorem 6.6]).

    Step 6. Conclusion.

    By the change of variable x=xh+rhy, inequalities (3.6), (3.7) and (v) of Lemma 2.2, for every 0<τ<14, we have

    lim suphU(xh,τrh)λ2hlim suphBτrh(x0)|V(Du)V((Du)x0,τrh)|2dx+lim suphP(E,Bτrh(xh))λ2hτn1rn1h+lim suphτrhλ2hlim suph1λ2hBτ|V(λhDvh+Ah)V(Ah+λh(Dvh)τ)|2dy+lim suphP(Eh,Bτ)λ2hτn1+τlim suphc(M,n,p)λ2hBτ|V(λh(Dvh(Dvh)τ)|2dy+lim suphP(Eh,Bτ)λ2hτn1+τ.

    Then, using Caccioppoli inequality in (3.16) and estimate of the perimeter (3.46), we get

    lim suphU(xh,τrh)λ2hc(n,p,1,2,L1,L2,Λ,M){lim suph1λ2hB2τ|V(λh(vh(vh)2τ(Dvh)τy)2τ)|2dy+1τnlim suphP(Eh,B1)nn1λ2h+1τn1lim suph(rhτnλ2h+rhλ2hλph)+τ}c(n,p,1,2,L1,L2,Λ,M){lim suph1λ2hB2τ|V(λh(vh(vh)2τ(Dvh)τy)2τ)|2dy+τ},

    where we have used (3.6), (3.8) and estimate (3.46).

    Now we want to prove the following extimate:

    lim suph1λ2hB2τ|V(λh(vh(vh)2τ(Dvh)τy)2τ)|2dy=lim suph1λ2hB2τ|V(λh(v(v)2τ(Dv)τy)2τ)|2dyB2τ|v(v)2τ(Dv)τy|2τ2dy.

    The last inequality is obtained by using that v and Dv are bounded, λh0 and |V(ξ)||ξ| for |ξ|1.

    We observe that proving the equality is equivalent to show

    I:=limh1λ2hB2τ|V(λh((vhv)(vhv)2τ(DvhDv)τy)2τ)|2dy=0.

    In the sequel σ will denote the exponent given in the Sobolev-Poincaré type inequality of the Theorem 2.7. We can assume that the higher integrability exponent δ given in the Step 5 is such that δ<σ.

    Let us choose θ(0,1) such that 2θ+1θ1+σ=1. Applying Hölder's inequality, it holds that

    0Ilim suph1λ2h(B2τ|V(λh((vhv)(vhv)2τ(DvhDv)τy)2τ)|dy)2θ×(B2τ|V(λh((vhv)(vhv)2τ(DvhDv)τy)2τ)|2(1+σ)dy)1θ1+σ.

    Using the fact that |V(ξ)||ξ| and (iii) of Lemma 2.2, for the first factor in the previous product, and using also Theorem 2.7 applied to (vhv)(vhv)2τ(DvhDv)τy, we deduce

    0Ilim suphcλ2h(λhB2τ(|vhvτ|+|(DvhDv)ττ|)dy)2θ×(B6τ|V(λh(DvhDv)λh(DvhDv)τ)|αdy)2(1θ)α,

    with 2/p<α<2 given in Theorem 2.7.

    In the last term we can increase choosing α=2, moreover, using again (iii) of Lemma 2.2 we deduce

    0Ilim suphcλ2h(λhB2τ(|vhvτ|+|(DvhDv)ττ|)dy)2θ×(B6τ|V(λh(DvhDv)|2+|V(λh((Dvh)τ(Dv)τ))|2dy)1θ.

    In the last term, we observe that the second addend can be estimated by making use of (i) of Lemma 2.2, the fact that DvhDv weakly in Lp(B1,RnN) and λh0. In particular, we obtain

    |V(λh((Dvh)τ(Dv)τ))|2cλ2h.

    Regarding the term

    B6τ|V(λh(DvhDv)|2dy,

    using (3.47) and the definition of vh, we deduce

    B12|V(λhDvh)|2(1+δ)dyC[(B1|V(λhDvh)|2dy)1+δ+min{|B1Eh|,|Eh|}]=C[(Brh(xh)|V(Du(x)(Du)xh,rh)|2dx)1+δ+min{|B1Eh|,|Eh|}]C[(Brh(xh)|V(Du(x))V((Du)xh,rh))|2dx)1+δ+min{|B1Eh|,|Eh|}]C[λ2(1+δ)h+λ2(1+ϵ)h]Cλ2(1+δ)h,

    where 0ϵ<1n1. Therefore, by Hölder's inequality, we have

    B12|V(λhDvh)|2dyC(M)λ2h.

    We conclude that

    0Ilimhcλ2hλ2θh(B2τ(|vhvτ|+|(DvhDv)ττ|)dy)2θλ2(1θ)h=limhC(B2τ(|vhv|+|(DvhDv)τ|)dy)2θ=0.

    By virtue of (3.6), (3.8), (3.9), the Poincaré-Wirtinger inequality and (3.40), we get

    lim suphU(xh,τrh)λ2hc(n,p,1,2,L2,Λ,M){B2τ|v(v)2τ(Dv)τy|2τ2dy+τ}c(n,p,1,2,L2,Λ,M){B2τ|Dv(Dv)τ|2dy+τ}c(n,N,p,1,2,L1,L2,Λ,M)[τ2+τ]C(n,N,p,1,2,L1,L2,Λ,M)τ.

    The contradiction follows, by choosing C_* such that C_* > C , since, by (3.5),

    \begin{equation} \liminf\limits_h\frac{U_*(x_h, \tau r_h)}{\lambda^2_h}\ge C_*\tau. \end{equation}

    If assumption (H) is not taken into account, it is still possible to establish a decay result for the excess, analogous to the previous one. However, this requires employing a modified ''hybrid" excess, defined as:

    \begin{equation} U_{**}(x_0, r): = U(x_0, r)+ \left(\frac{P(E, B_r(x_0))}{r^{n-1}}\right)^{\frac{\delta}{1+\delta}}+r^\beta, \end{equation}

    where U(x_0, r) is defined in (3.1), \delta is the higher integrability exponent given in the Step 5 of Proposition 3.1 and 0 < \beta < \frac{\delta}{1+\delta} . The following result still holds true.

    Proposition 3.2. Let (u, E) be a local minimizer of \mathcal{I} in (1.2) under the assumptions (F1), (F2), (G1), and (G2). For every M > 0 and 0 < \tau < \frac{1}{4} , there exist two positive constants \varepsilon_0 = \varepsilon_0(\tau, M) and c_{**} = c_{**}(n, p, \ell_1, \ell_2, L_1, L_2, \Lambda, \delta, M) for which, whenever B_r(x_0)\Subset{\Omega} verifies

    \begin{equation*} |(Du)_{x_0, r}|\leq M\quad\mathrm{and}\quad U_{**}(x_{0}, r)\leq \epsilon_0, \end{equation*}

    then

    \begin{equation} U_{**}(x_{0}, \tau r)\leq c_{**}\, \tau^\beta\, U_{**}(x_{0}, r). \end{equation}

    In order to avoid unnecessary repetition we do not include the proof here, as it is almost identical to the proof of the Proposition 3.1, with the obvious adjustments, see [9].

    Here we give the proof of Theorem 1.3 through a suitable iteration procedure. It is easy to show the validity of the following lemma by arguing exactly in the same way as in [11, Lemma 6.1].

    Lemma 4.1. Let (u, E) be a local minimizer of the functional \mathcal{I} and let c_* the constant introduced in Proposition 3.1. For every \alpha\in (0, 1) and M > 0 there exists \vartheta_0 = \vartheta_0(c_*, \alpha) < 1 such that for \vartheta\in (0, \vartheta_0) there exists a positive constant \varepsilon_1 = \varepsilon_1(n, p, \ell_1, \ell_2, L_1, L_2, M, \vartheta) such that, if B_r(x_0)\Subset \Omega ,

    \begin{equation*} |Du|_{x_0, r} < M\quad \mathit{\text{and}}\quad U_*(x_0, r) < \varepsilon_1, \end{equation*}

    then

    \begin{equation} |D u|_{x_0, \vartheta^h r} < 2M\quad\mathit{\text{and}}\quad U_*(x_0, \vartheta^{h}r) \leq\vartheta^{h\alpha} U_*(x_0, r), \quad \forall h\in \mathbb{N}_0. \end{equation} (4.1)

    Proof. Let M > 0 , \alpha\in (0, 1) and \vartheta\in (0, \vartheta_0) , where \vartheta_0 < 1 . Let \varepsilon_1 < \varepsilon_0 , where \varepsilon_0 is the constant appearing in Proposition 3.1. We first prove by induction that

    \begin{equation} |D u|_{x_0, \vartheta^h r} < 2M, \quad\forall h\in{\mathbb N}_0. \end{equation} (4.2)

    If h = 0 , the statement holds. Assuming that (4.1) holds for h > 0 , applying properties (i) and (vi) of Lemma 2.2, we compute

    \begin{align} |Du|_{x_0, \vartheta^{h+1} r} & \leq |Du|_{x_0, r}+\sum\limits_{j = 1}^{h+1}||Du|_{x_0, \vartheta^{j} r}-|Du|_{x_0, \vartheta^{j-1}r}| \\ & \leq M+\sum\limits_{j = 1}^{h+1} \rlap{-} \displaystyle {\int }_{B_{\vartheta^j r}}|Du-(Du)_{x_0, \vartheta^{j-1}r}|\, dx\\ & \leq M +\vartheta^{-n}\sum\limits_{j = 1}^{h+1}\Bigg[ \frac{1}{|B_{\vartheta^{j-1} r}|}\int_{B_{\vartheta^{j-1} r}\cap\{|Du-(Du)_{x_0, \vartheta^{j-1}r}|\leq 1\}}\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!|Du-(Du)_{x_0, \vartheta^{j-1}r}|\, dx\\ & + \frac{1}{|B_{\vartheta^{j-1} r}|}\int_{B_{\vartheta^{j-1} r}\cap\{|Du-(Du)_{x_0, \vartheta^{j-1}r}| > 1\}}|Du-(Du)_{x_0, \vartheta^{j-1}r}|\, dx\\ & \leq M+\vartheta^{-n}\sum\limits_{j = 1}^{h+1} \Bigg[\bigg(\rlap{-} \displaystyle {\int }_{B_{\vartheta^{j-1} r}}|V(Du-(Du)_{x_0, \vartheta^{j-1}r})|^2\, dx\bigg)^{\frac{1}{2}} \\ & +\bigg(\rlap{-} \displaystyle {\int }_{B_{\vartheta^{j-1} r}}|V(Du-(Du)_{x_0, \vartheta^{j-1}r})|^2\, dx\bigg)^{\frac{1}{p}}\Bigg] \\ & \leq M+c(p, M)\vartheta^{-n}\sum\limits_{j = 1}^{h+1}\big[U_*(x_0, \vartheta^{j-1}r)^{\frac{1}{2}}+U_*(x_0, \vartheta^{j-1}r)^{\frac{1}{p}}\big] \\ & \leq M+c(p, c_*, M)\varepsilon_1^{\frac{1}{2}}\vartheta^{-n}\sum\limits_{j = 1}^{h+1} \vartheta^{\frac{j-1}{2}}\leq M+c(p, c_*, M)\varepsilon_1^{\frac{1}{2}}\frac{\vartheta^{-n}}{1-\vartheta^{\frac{1}{2}}}\leq 2M, \end{align}

    where we have chosen \varepsilon_1 = \varepsilon_1(p, c_*, M, \vartheta) > 0 sufficiently small. Now we prove the second inequality in (4.1). The statement is obvious for h = 0 . If h > 0 and (4.1) holds, we have that

    \begin{equation} U_*(x_0, \vartheta^{h}r)\leq \vartheta^{h\alpha} U_*(x_0, r) < \varepsilon_1, \end{equation} (4.3)

    by our choice of \vartheta and \varepsilon_1 . Thus thanks to (4.2) we can apply Proposition 3.1 with \vartheta^h r in place of r to deduce that

    \begin{equation*} U_*(x_0, \vartheta^{h+1}r)\leq \vartheta^\alpha U_*(x_0, \vartheta ^h r)\leq \vartheta^{(h+1)\alpha}U_*(x_0, r), \end{equation*}

    where we have chosen \vartheta_0 = \vartheta_0(c_*, \alpha) sufficiently small and we have used (4.3). Therefore, the second inequality in (4.1) is also true for every k\in \mathbb{N} .

    Analogously, it is possible to prove an iteration lemma for U_{**} .

    Lemma 4.2. Let (u, E) be a local minimizer of the functional \mathcal{I} and let \beta be the exponent of Proposition 3.2. For every M > 0 and \vartheta\in (0, \vartheta_0) , with \vartheta_0 < \min\left\{ c_{**}, \frac{1}{4}\right\} , there exist \varepsilon_1 > 0 and R > 0 such that, if r < R and x_0\in\Omega satisfy

    \begin{equation*} B_r(x_0)\Subset \Omega, \quad |Du|_{x_0, r} < M\quad and\quad U_{**}(x_0, r) < \varepsilon_1, \end{equation*}

    where c_{**} is the constant introduced in Proposition 3.2, then

    \begin{equation} |D u|_{x_0, \vartheta^h r} < 2M\quad\mathit{\text{and}}\quad U_{**}(x_0, \vartheta^{k}r)\leq \vartheta^{k\beta} U_{**}(x_0, r), \quad\forall k\in{\mathbb N}. \end{equation}

    Proof of Theorem 1.3. We consider the set

    \begin{equation} \Omega_1: = \bigg\{x\in \Omega:\, \, \limsup\limits_{\rho\to 0}|(Du)_{x, \rho}| < \infty \, \, \mathrm{and}\, \, \limsup\limits_{\rho\to 0} U_*(x, \rho) = 0\bigg\} \end{equation}

    and let x_0\in \Omega_1 . For every M > 0 and for \varepsilon_1 determined in Lemma 4.1 there exists a radius R_{M, \varepsilon_1} > 0 such that

    \begin{equation} |Du|_{x_0, r} < M\quad \text{and}\quad U_{*}(x_0, r) < \varepsilon_1, \end{equation}

    for every 0 < r < R_{M, \varepsilon_1} . Let 0 < \rho < \vartheta r < R and h\in\mathbb{N} be such that \vartheta^{h+1}r < \rho < \vartheta^h r , where \vartheta = \frac{\vartheta_0}{2} and \vartheta_0 is the same constant appearing in Lemma 4.1. By Lemma 4.1, we obtain

    \begin{equation*} |D u|_{x_0, \rho}\leq \frac{1}{\vartheta^n}|D u|_{x_0, \vartheta^h r}\leq c(M, c_*, \alpha). \end{equation*}

    Using the properties of Lemma 2.2 and reasoning as in the proof of Lemma 4.1, we estimate

    \begin{align} &\quad |V((Du)_{x_0, \vartheta^h r})-V((Du)_{x_0, \rho})|^2 \\ & \leq c(n, p)|(Du)_{x_0, \vartheta^h r}-(Du)_{x_0, \rho}|^2 \\ & \leq c(n, p)\bigg(\rlap{-} \displaystyle {\int }_{B_{\rho}(x_0)}|Du-(Du)_{x_0, \vartheta^{h}r}|\, dx\bigg)^2 \\ & \leq c(n, p)\vartheta_0^{-2n}\Bigg[ \frac{1}{|B_{\vartheta^{h} r}|}\int_{B_{\vartheta^{h} r}\cap\{|Du-(Du)_{x_0, \vartheta^{h}r}|\leq 1\}}\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!|Du-(Du)_{x_0, \vartheta^{h}r}|\, dx \\ & + \frac{1}{|B_{\vartheta^{h} r}|}\int_{B_{\vartheta^{h} r}\cap\{|Du-(Du)_{x_0, \vartheta^{h}r}| > 1\}}|Du-(Du)_{x_0, \vartheta^{h}r}|\, dx\Bigg]^2 \\ & \leq c(n, p)\vartheta_0^{-2n} \Bigg[\bigg(\rlap{-} \displaystyle {\int }_{B_{\vartheta^{h} r}}|V(Du-(Du)_{x_0, \vartheta^{h}r})|^2\, dx\bigg)^{\frac{1}{2}} \\ & +\bigg(\rlap{-} \displaystyle {\int }_{B_{\vartheta^{h} r}}|V(Du-(Du)_{x_0, \vartheta^{h}r})|^2\, dx\bigg)^{\frac{1}{p}}\Bigg]^2 \\ & \leq c(n, p, M)\vartheta_0^{-2n}\big[U_*(x_0, \vartheta^{h}r)+U_*(x_0, \vartheta^{h}r)^{\frac{2}{p}}\big] \\ & \leq c(n, p, c_*, M)\vartheta_0^{-2n}\vartheta^{h\alpha}U_*(x_0, r). \end{align}

    Thus, taking the previous chain of inequalities into account, applying again Lemma 4.1, we estimate

    \begin{align} U_*(x_0, \rho) &\leq 2\rlap{-} \displaystyle {\int }_{B_\rho(x_0)}|V(Du)-V((Du)_{x_0, \vartheta^h r})|^2\, dx+2|V((Du)_{x_0, \vartheta^h r})-V((Du)_{x_0, \rho})|^2 \\ & + \frac{P(E, B_\rho(x_0))}{\rho^{n-1}}+\rho \\ & \leq c(n, p, M, c_*\vartheta_0)\bigg[\rlap{-} \displaystyle {\int }_{B_{\vartheta^h r}(x_0)}|V(Du)-V((Du)_{x_0, \vartheta^h r})|^2\, dx+\vartheta^{h\alpha}U_*(x_0, r)\\ & +\frac{P(E, B_{\vartheta^h r}(x_0))}{(\vartheta^h r)^{n-1}}+\vartheta^h r\bigg] \\ & \leq c(n, p, c_*, M, \vartheta_0) \big[U_*(x_0, \vartheta^h r)+\vartheta^{h\alpha} U_*(x_0, r)\big] \\ &\leq c(n, p, c_*, M, \vartheta_0)\left(\frac{\rho}{r}\right)^\alpha U_*(x_0, r). \end{align}

    The previous estimate implies that

    \begin{equation} {U(x_0, \rho)}\leq C_*\left( \frac{\rho}{r}\right)^{\alpha}U_*(x_0, r), \end{equation}

    where C_{*} = C_{*}(n, p, c_*, M, \vartheta_0) . Since U_*(y, r) is continuous in y , we have that U_*(y, r) < \varepsilon_1 for every y in a suitable neighborhood I of x_0 . Therefore, for every y\in I we have that

    \begin{equation*} U(y, \rho)\leq C_* \left(\frac{\rho}{r}\right)^\alpha U_*(y, r). \end{equation*}

    The last inequality implies, by the Campanato characterization of Hölder continuous functions (see [32, Theorem 2.9]), that u is C^{1, \alpha} in I for every 0 < \alpha < \frac{1}{2} , and we can conclude that the set \Omega_1 is open and the function u has Hölder continuous derivatives in \Omega_1 .

    When the assumption (H) is not enforced, the proof goes exactly in the same way provided we use Lemma 4.2 in place of Lemma 4.1, with

    \begin{equation} \Omega_0: = \bigg\{x\in \Omega:\, \, \limsup\limits_{\rho\to 0}|(Du)_{x_0, \rho}| < \infty \, \, \mathrm{and}\, \, \limsup\limits_{\rho\to 0} U_{**}(x_0, \rho) = 0\bigg\}. \end{equation}

    In this paper, we studied the C^{1, \alpha} partial regularity for a wide class of multidimensional vectorial variational problems involving both bulk and surface energies. The bulk energy densities are uniformly strictly quasiconvex functions with subquadratic growth p\in (1, 2) . Since the case p \geq 2 had been addressed in a previous work by the authors, the present paper completes the analysis by covering the entire range p > 1 . The overall strategy of the proof is to establish an excess decay property for a suitably chosen excess function. The core of the argument - and the main contribution of the paper - is Proposition 3.1, where a one-step improvement of the excess is established. The proof proceeds via a contradiction and blow-up argument. The proof of Proposition 3.1 is rather long; nevertheless, we would like to highlight two fundamental estimates that are pivotal in the proof strategy. These are the Caccioppoli estimate (3.16) and the higher integrability estimate (3.47) for the blow-up sequences, in which the influence of the set E appears explicitly. These estimates, together with the Sobolev–Poincaré inequality (2.7), which is specific to the subquadratic case, constitute the main tools used to establish the result.

    Finally, we would like to mention two possible directions for future research, kindly suggested by one of the referees. The first concerns the potential extension of the same type of regularity to the non-uniformly elliptic case. Another intriguing question concerns the double-phase case, which may be more challenging, but should still be manageable - at least in the situation where the two phases are separated in the sets E and \Omega \setminus E .

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    The authors are members of the Gruppo Nazionale per l'Analisi Matematica, la Probabilità e le loro Applicazioni (GNAMPA) of the Istituto Nazionale di Alta Matematica (INdAM) and wish to acknowledge financial support from INdAM GNAMPA Project 2024 "Regolarità per problemi a frontiera libera e disuguaglianze funzionali in contesto finsleriano" (CUP E53C23001670001).

    Prof. Menita Carozza is a Guest Editor of special issue ''Multi-Rate Processes and Hysteresis" for Mathematics in Engineering. Prof. Menita Carozza was not involved in the editorial review and the decision to publish this article.

    The author declares no conflicts of interest.



    [1] E. Acerbi, N. Fusco, An approximation lemma for W^{1;p} functions, Material Instabilities in Continuum Mechanics (Edinburgh, 1985-1986), Oxford Science Publications, Oxford University Press,, 1988, 1–5.
    [2] E. Acerbi, N. Fusco, A regularity theorem for minimizers of quasiconvex integrals, Arch. Ration. Mech. Anal., 99 (1987), 261–281. https://doi.org/10.1007/BF00284509 doi: 10.1007/BF00284509
    [3] F. J. Almgren, Existence and regularity almost everywhere of solutions to elliptic variational problems among surfaces of varying topological type and singularity structure, Ann. Math., 87 (1968), 321–391. https://doi.org/10.2307/1970587 doi: 10.2307/1970587
    [4] L. Ambrosio, G. Buttazzo, An optimal design problem with perimeter penalization, Calc. Var. Partial Differential Equations, 1 (1993), 55–69. https://doi.org/10.1007/BF02163264 doi: 10.1007/BF02163264
    [5] A. Arroyo-Rabasa, Regularity for free interface variational problems in a general class of gradients, Calc. Var. Partial Differential Equations, 55 (2016), 154. https://doi.org/10.1007/s00526-016-1085-5 doi: 10.1007/s00526-016-1085-5
    [6] G. Bellettini, M. Novaga, M. Paolini, On a Crystalline Variational Problem, Part Ⅰ: first variation and global L^{\infty} regularity, Arch. Rational Mech. Anal., 157 (2001), 165–191. https://doi.org/10.1007/s002050010127 doi: 10.1007/s002050010127
    [7] G. Bellettini, M. Novaga, M. Paolini, On a Crystalline Variational Problem, Part Ⅱ: BV regularity and structure of minimizers on facets, Arch. Rational Mech. Anal., 157 (2001), 193–217. https://doi.org/10.1007/s002050100126 doi: 10.1007/s002050100126
    [8] E. Bombieri, Regularity theory for almost minimal currents, Arch. Ration. Mech. Anal., 78 (1982), 99–130. https://doi.org/10.1007/BF00250836 doi: 10.1007/BF00250836
    [9] M. Carozza, L. Esposito, L. Lamberti, Quasiconvex bulk and surface energies: C^{1, \alpha} regularity, Adv. Nonlinear Anal., 13 (2024), 20240021. https://doi.org/10.1515/anona-2024-0021 doi: 10.1515/anona-2024-0021
    [10] M. Carozza, I. Fonseca, A. P. Di Napoli, Regularity results for an optimal design problem with a volume constraint, ESAIM: COCV, 20 (2014), 460–487. https://doi.org/10.1051/cocv/2013071 doi: 10.1051/cocv/2013071
    [11] M. Carozza, I. Fonseca, A. P. Di Napoli, Regularity results for an optimal design problem with quasiconvex bulk energies, Calc. Var. Partial Differential Equations, 57 (2018), 68. https://doi.org/10.1007/s00526-018-1343-9 doi: 10.1007/s00526-018-1343-9
    [12] M. Carozza, N. Fusco, G. Mingione, Partial regularity of minimizers of quasiconvex integrals with subquadratic growth, Ann. Mat. Pura Appl., 175 (1998), 141–164. https://doi.org/10.1007/BF01783679 doi: 10.1007/BF01783679
    [13] M. Carozza, G. Mingione, Partial regularity of minimizers of quasiconvex integrals with subquadratic growth: the general case, Ann. Pol. Math., 77 (2001), 219–243. https://doi.org/10.4064/ap77-3-3 doi: 10.4064/ap77-3-3
    [14] M. Carozza, A. P. Di Napoli, A regularity theorem for minimisers of quasiconvex integrals: The case 1 < p < 2, Proc. Roy. Soc. Edinb. Sec. A Math., 126 (1996), 1181–1200. https://doi.org/10.1017/S0308210500023350 doi: 10.1017/S0308210500023350
    [15] M. Carozza, A. P. Di Napoli, Partial regularity of local minimizers of quasiconvex integrals with sub-quadratic growth, Proc. Roy. Soc. Edinb. Sec. A Math., 133 (2003), 1249–1262. https://doi.org/10.1017/S0308210500002900 doi: 10.1017/S0308210500002900
    [16] R. Choksi, R. Neumayer, I. Topaloglu, Anisotropic liquid drop models, Adv. Calc. Var., 15 (2022), 109–131. https://doi.org/10.1515/acv-2019-0088 doi: 10.1515/acv-2019-0088
    [17] G. De Philippis, A. Figalli, A note on the dimension of the singular set in free interface problems, Differ. Integral Equ., 28 (2015), 523–536. https://doi.org/10.57262/die/1427744099 doi: 10.57262/die/1427744099
    [18] G. De Philippis, N. Fusco, M. Morini, Regularity of capillarity droplets with obstacle, Trans. Amer. Math. Soc., 377 (2024), 5787–5835. https://doi.org/10.1090/tran/9152 doi: 10.1090/tran/9152
    [19] G. De Philippis, F. Maggi, Dimensional estimates for singular sets in geometric variational problems with free boundaries, J. Reine Angew. Math., 725 (2017), 217–234. https://doi.org/10.1515/crelle-2014-0100 doi: 10.1515/crelle-2014-0100
    [20] F. Duzaar, G. Mingione, Regularity for degenerate elliptic problems via p-harmonic approximation, Ann. Inst. H. Poincaré Anal. Non Linéaire, 21 (2004), 735–766. https://doi.org/10.1016/j.anihpc.2003.09.003 doi: 10.1016/j.anihpc.2003.09.003
    [21] F. Duzaar, K. Steffen, Optimal interior and boundary regularity for almost minimizers to elliptic variational integrals, J. Reine Angew. Math., 546 (2002), 73–138. https://doi.org/10.1515/crll.2002.046 doi: 10.1515/crll.2002.046
    [22] L. Esposito, Density lower bound estimate for local minimizer of free interface problem with volume constraint, Ricerche Mat., 68 (2019), 359–373. https://doi.org/10.1007/s11587-018-0407-7 doi: 10.1007/s11587-018-0407-7
    [23] L. Esposito, N. Fusco, A remark on a free interface problem with volume constraint, J. Convex Anal., 18 (2011), 417–426.
    [24] L. Esposito, L. Lamberti, Regularity Results for an Optimal Design Problem with lower order terms, Adv. Calc. Var., 16 (2023) 1093–1122. https://doi.org/10.1515/acv-2021-0080 doi: 10.1515/acv-2021-0080
    [25] L. Esposito, L. Lamberti, Regularity results for a free interface problem with Hölder coefficients, Calc. Var. Partial Differential Equations, 62 (2023), 156. https://doi.org/10.1007/s00526-023-02490-x doi: 10.1007/s00526-023-02490-x
    [26] L. Esposito, L. Lamberti, G. Pisante, Epsilon-regularity for almost-minimizers of anisotropic free interface problem with Hölder dependence on the position, Interfaces Free Bound., 2024. https://doi.org/10.4171/ifb/535 doi: 10.4171/ifb/535
    [27] A. Figalli, Regularity of codimension-1 minimizing currents under minimal assumptions on the integrand, J. Differential Geom., 106 (2017), 371–391. https://doi.org/10.4310/jdg/1500084021 doi: 10.4310/jdg/1500084021
    [28] A. Figalli, F. Maggi, On the shape of liquid drops and crystals in the small mass regime, Arch. Rational Mech. Anal., 201 (2011), 143–207. https://doi.org/10.1007/s00205-010-0383-x doi: 10.1007/s00205-010-0383-x
    [29] N. Fusco, V. Julin, On the regularity of critical and minimal sets of a free interface problem, Interfaces Free Bound., 17 (2015), 117–142. https://doi.org/10.4171/IFB/336 doi: 10.4171/IFB/336
    [30] M. Giaquinta, Multiple integrals in the calculus of variations and nonlinear elliptic systems, Vol. 105, Princeton: Princeton University Press, 1984. https://doi.org/10.1515/9781400881628
    [31] M. Giaquinta, G. Modica, Partial regularity of minimizers of quasiconvex integrals, Ann. Inst. H. Poincaré, Anal. Non Linéaire, 3 (1986), 185–208. https://doi.org/10.1016/S0294-1449(16)30385-7 doi: 10.1016/S0294-1449(16)30385-7
    [32] E. Giusti, Direct methods in the calculus of variations, World Scientific, 2003. https://doi.org/10.1142/5002
    [33] M. Gurtin, On phase transitions with bulk, interfacial, and boundary energy, Arch. Ration. Mech. Anal., 96 (1986), 243–264. https://doi.org/10.1007/BF00251908 doi: 10.1007/BF00251908
    [34] L. Lamberti, A regularity result for minimal configurations of a free interface problem, Boll. Unione Mat. Ital., 14 (2021), 521–539. https://doi.org/10.1007/s40574-021-00285-6 doi: 10.1007/s40574-021-00285-6
    [35] F. H. Lin, Variational problems with free interfaces, Calc. Var. Partial Differential Equations, 1 (1993), 149–168. https://doi.org/10.1007/BF01191615 doi: 10.1007/BF01191615
    [36] F. H. Lin, R. V. Kohn, Partial regularity for optimal design problems involving both bulk and surface energies, Chin. Ann. Math., 20 (1999), 137–158. https://doi.org/10.1142/S0252959999000175 doi: 10.1142/S0252959999000175
    [37] F. Maggi, Sets of finite perimeter and geometric variational problems: an introduction to geometric measure theory, Cambridge University Press, 2012. https://doi.org/10.1017/CBO9781139108133
    [38] P. Marcellini, Approximation of quasiconvex functions, and lower semicontinuity of multiple integrals, Manuscripta Math., 51 (1985), 1–28. https://doi.org/10.1007/BF01168345 doi: 10.1007/BF01168345
    [39] R. Schoen, L. Simon, A new proof of the regularity theorem for rectifiable currents which minimize parametric elliptic functionals, Indiana Univ. Math. J., 31 (1982), 415–434. https://doi.org/10.1512/iumj.1982.31.31035 doi: 10.1512/iumj.1982.31.31035
    [40] D. A. Simmons, Regularity of almost-minimizers of Hölder-coefficient surface energies, Discrete. Contin. Dyn. Syst., 42 (2022), 3233–3299. https://doi.org/10.3934/dcds.2022015 doi: 10.3934/dcds.2022015
  • Reader Comments
  • © 2025 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(154) PDF downloads(43) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog