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

Regularity for Double Phase Problems at
Nearly Linear Growth

Cristiana De Filippis & Giuseppe Mingione [email protected]
Abstract

Minima of functionals of the type

wΩ[|Dw|log(1+|Dw|)+a(x)|Dw|q]𝑑x,0a()C0,α,w\mapsto\int_{\Omega}\left[\lvert Dw\rvert\log(1+\lvert Dw\rvert)+a(x)\lvert Dw\rvert^{q}\right]\,dx\,,\quad 0\leq a(\cdot)\in C^{0,\alpha}\,,

with Ωn\Omega\subset\mathbb{R}^{n}, have locally Hölder continuous gradient provided 1<q<1+α/n1<q<1+\alpha/n.

1 Introduction

In this paper we prove the first Schauder type results for minima of nonuniformly elliptic integrals with nearly linear growth, and under optimal bounds on the rate of nonuniform ellipticity. This is a non-trivial task, as in nonuniformly elliptic problems Schauder type estimates change their nature and cannot be achieved via perturbation methods. This was first shown by the counterexamples in sharp ; FMM . A relevant model functional we have in mind is

w(w,Ω):=Ω[𝔠(x)|Dw|log(1+|Dw|)+a(x)|Dw|q]𝑑x.w\mapsto\mathcal{L}(w,\Omega):=\int_{\Omega}\left[\mathfrak{c}(x)\lvert Dw\rvert\log(1+\lvert Dw\rvert)+a(x)\lvert Dw\rvert^{q}\right]\,dx\,. (1.1)

Here, as also in the following, Ωn\Omega\subset\mathbb{R}^{n} denotes an open subset, n2n\geq 2 and q>1q>1; as our results will be local in nature, without loss of generality we assume that Ω\Omega is also bounded. The functions a(),𝔠()a(\cdot),\mathfrak{c}(\cdot) are bounded and non-negative with 𝔠()\mathfrak{c}(\cdot) which is also bounded away from zero; a crucial point here is that a()a(\cdot) is instead allowed to vanish. For the definition of local minimizer see (1.15). We are interested in the case a(),𝔠()a(\cdot),\mathfrak{c}(\cdot) are only Hölder continuous, when proving gradient Hölder regularity of minima in fact corresponds to establish a nonlinear version of Schauder estimates (originally also due to Hopf and Caccioppoli). The functional in (1.1) appears to be a combination of the classical nearly linear growth one

wΩ𝔠(x)|Dw|log(1+|Dw|)𝑑xw\mapsto\int_{\Omega}\mathfrak{c}(x)\lvert Dw\rvert\log(1+\lvert Dw\rvert)\,dx (1.2)

treated when 𝔠()1\mathfrak{c}(\cdot)\equiv 1 for instance in FM and by Marcellini & Papi in MP , with a qq-power growth term weighted by a non-negative coefficient a()a(\cdot). The terminology “nearly linear” accounts for an integrand whose growth in the gradient variable is superlinear, but it is still slower than that of any power type integrand as z|z|pz\mapsto|z|^{p} with p>1p>1. When 𝔠()1\mathfrak{c}(\cdot)\equiv 1 the functional ()\mathcal{L}(\cdot) can be in fact considered as a limiting case (p1p\to 1) of the by now largely studied double phase integral

wΩ(|Dw|p+a(x)|Dw|q)𝑑x,1<p<q.w\mapsto\int_{\Omega}\left(\lvert Dw\rvert^{p}+a(x)\lvert Dw\rvert^{q}\right)\,dx\,,\qquad 1<p<q\,\,. (1.3)

This last one is a basic prototype of a nonautonomous functional featuring nonstandard polynomial growth conditions and a soft kind of nonuniform ellipticity (see discussion below). Originally introduced by Zhikov Z0 ; Z1 ; Zhom in the setting of Homogenization of strongly anisotropic materials, the functional in (1.3) was first studied in CM ; sharp ; FMM and can be thought as a model for composite media with different hardening exponents pp and qq. The geometry of the mixture of the two materials is in fact described by the zero set {a(x)0}\{a(x)\equiv 0\} of the coefficient a()a(\cdot), where the transition from qq-growth to pp-growth takes place. In turn, the functional in (1.2) appears for instance in the theory of plasticity with logarithmic hardening, another borderline case between plasticity with power hardening (p>1p>1) and perfect plasticity (p=1p=1) FrS ; FS1 . Here, we are mostly interested in the theoretical issues raised by functionals as (1.1) in the context of regularity theory. In fact, ()\mathcal{L}(\cdot) provides a basic example of nonuniformly and nonautonomous elliptic functional for which the currently available techniques do not allow to prove sharp Schauder type estimates. In particular, although being a limiting case of (1.3), there is no way to adapt the methods available for (1.3) to treat ()\mathcal{L}(\cdot), already when 𝔠()1\mathfrak{c}(\cdot)\equiv 1. We recall that general nonuniformly elliptic problems have been the object of intensive investigation over the last two decades BS0 ; BS ; BS2 ; bbgi ; bildhauer ; BF ; BF2 ; FM , starting from the more classical case of minimal surface type functionals LU ; Simon and those from the 60s ivanov0 ; ivanov1 ; IO ; serrin ; trudinger . The monograph ivanov2 gives a valuable account of the early age of the theory and we recall that nonuniformly elliptic energies have been considered several times in the setting of nonlinear elasticity BFM ; FMal ; FMa ; M0 ; ma5 . As for linear and nearly linear growth conditions in the vectorial case, we mention the recent extensive work of Kristensen & Gmeineder gme1 ; gme2 ; gmek1 ; gmek2 . But let us first recall the situation for (1.3). The bound

qp1+αn\frac{q}{p}\leq 1+\frac{\alpha}{n} (1.4)

guarantees the local Hölder continuity of the gradient of minima of (1.3) BCM ; CM ; DM . Condition (1.4) is optimal in the sense that its failure generates the existence of minimizers that do not even belong to Wloc1,qW^{1,q}_{{\rm loc}} sharp and develop singularities on fractals with almost maximal dimension FMM . In (1.4) notice the delicate interaction between ambient dimension, the (p,q)(p,q)-growth conditions in the gradient and the regularity of coefficient a()a(\cdot). Despite the simple form of (1.3), all this already happens in the scalar case. In the autonomous case bounds of the type in (1.4), with no appearance of α\alpha, play a key role in the regularity theory of functionals with nonstandard growth as built by Marcellini in his by now classical papers M0 ; M1 ; M2 ; M3 . The key property of the functional (1.3) is that it is uniformly elliptic in the classical (pointwise) sense. This means that, with G(x,z):=|z|p+a(x)|z|qG(x,z):=|z|^{p}+a(x)|z|^{q}, we have

supxΩ,|z|1G(x,)(z)<c(p,q),\sup_{x\in\Omega,\lvert z\rvert\geq 1}\,\mathcal{R}_{G(x,\cdot)}(z)<c(p,q)\,, (1.5)

where G(x,)\mathcal{R}_{G(x,\cdot)} denotes the (pointwise) ellipticity ratio of G(x,)G(x,\cdot), i.e.,

G(x,)(z):=highest eigenvalue ofzzG(x,z)lowest eigenvalue ofzzG(x,z).\mathcal{R}_{G(x,\cdot)}(z):=\frac{\mbox{highest eigenvalue of}\ \partial_{zz}G(x,z)}{\mbox{lowest eigenvalue of}\ \partial_{zz}G(x,z)}\,.

Nevertheless, the functional ()\mathcal{L}(\cdot) exhibits a weaker form of nonuniform ellipticity, detectable via a larger quantity called nonlocal ellipticity ratio ciccio . This is defined by

G(z,B):=supxBhighest eigenvalue ofzzG(x,z)infxB lowest eigenvalue ofzzG(x,z),\mathcal{R}_{G}(z,B):=\frac{\sup_{x\in B}\,\mbox{highest eigenvalue of}\ \partial_{zz}G(x,z)}{\inf_{x\in B}\,\mbox{ lowest eigenvalue of}\ \partial_{zz}G(x,z)}\,,

where BΩB\subset\Omega is any ball. Indeed, note that

G(z,B)1+aL(B)|z|qp\mathcal{R}_{G}(z,B)\approx 1+\|a\|_{L^{\infty}(B)}|z|^{q-p} (1.6)

holds whenever {a(x)0}B\{a(x)\equiv 0\}\cap B is nonempty, and it is therefore unbounded with respect to |z||z|. We refer to ciccio for a discussion on these quantities; we just note that they do coincide in the autonomous case. It is precisely the occurrence of both (1.5) and (1.6) to imply the regularity of minima of the double phase functional in (1.3) when assuming (1.4). Indeed (1.4) and (1.6), allow to implement a delicate perturbation method based on the fact that minimizers of the frozen integral wG(x0,Dw)𝑑xw\mapsto\int G(x_{0},Dw)\,dx enjoy good decay estimates, which is a consequence of (1.5). A similar scheme, based on the occurrence of pointwise uniform ellipticity (1.5), is the starting point for treating larger classes of problems featuring non-polynomial ellipticity conditions, see for instance the recent interesting papers by Hästo & Ok HO1 ; HO2 . These also include integrals of the type

wΩ[Φ1(|Dw|)+a(x)Φ2(|Dw|)]𝑑x,1<p<q,w\mapsto\int_{\Omega}\left[\Phi_{1}(|Dw|)+a(x)\Phi_{2}(|Dw|)\right]\,dx\,,\qquad 1<p<q\,,

where tΦ1(t)t\mapsto\Phi_{1}(t) has superlinear growth (these have been treated in byun0 ; byun1 ). Such schemes completely break down in the case of functionals as ()\mathcal{L}(\cdot) since the boundedness of zG(x,)()z\mapsto\mathcal{R}_{G(x,\cdot)}(\cdot) fails for G(x,z)𝔠(x)|z|log(1+|z|)+a(x)|z|qG(x,z)\equiv\mathfrak{c}(x)|z|\log(1+|z|)+a(x)|z|^{q}. Indeed, already in the case of the functional in (1.2), which is nonuniformly elliptic in the classical sense (i.e., (1.5) fails), no Lipschitz continuity result for minima is available but when considering differentiablity assumptions on 𝔠()\mathfrak{c}(\cdot) (see dm ). In other words, the validity of Hopf-Caccioppoli-Schauder estimates is an open issue. We fix the current situation in the following:

Theorem 1.1 (Nearly linear Hopf-Caccioppoli-Schauder)

Let uWloc1,1(Ω)u\in W^{1,1}_{{\rm loc}}(\Omega) be a local minimizer of the functional ()\mathcal{L}(\cdot) in (1.1), with

{0a()C0,α(Ω),1<q<1+α/n𝔠()Cloc0,α0(Ω),1/Λ𝔠()Λ,\begin{cases}\displaystyle 0\leq a(\cdot)\in C^{0,\alpha}(\Omega),&1<q<1+\alpha/n\\ \mathfrak{c}(\cdot)\in C^{0,\alpha_{0}}_{{\rm loc}}(\Omega),&\displaystyle 1/\Lambda\leq\mathfrak{c}(\cdot)\leq\Lambda\,,\end{cases} (1.7)

where α,α0(0,1)\alpha,\alpha_{0}\in(0,1) and Λ1\Lambda\geq 1. Then DuDu is locally Hölder continuous in Ω\Omega and moreover, for every ball BBrΩB\equiv B_{r}\Subset\Omega, r1r\leq 1, the inequality

DuL(B/2)c(B[|Du|log(1+|Du|)+a(x)|Du|q]dx)ϑ+c\lVert Du\rVert_{L^{\infty}(B/2)}\leq c\left(\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B}\left[\lvert Du\rvert\log(1+\lvert Du\rvert)+a(x)\lvert Du\rvert^{q}\right]\,dx\right)^{\vartheta}+c

holds with cc(n,q,Λ,α,α0,aC0,α,𝔠C0,α0)c\equiv c(n,q,\Lambda,\alpha,\alpha_{0},\lVert a\rVert_{C^{0,\alpha}},\lVert\mathfrak{c}\rVert_{C^{0,\alpha_{0}}}) and ϑϑ(n,q,α,α0)>0\vartheta\equiv\vartheta(n,q,\alpha,\alpha_{0})>0.

Of course the condition q<1+α/nq<1+\alpha/n in (1.7)1 is the sharp borderline version of (1.4) (let p1p\to 1), that guarantees gradient Hölder continuity of minima for the classical double phase functional (1.3). Let us only observe that the equality case in (1.4), in comparison to (1.7)1 assumed here, is typically linked to superlinear growth conditions p>1p>1 in (1.3). The equality in (1.4) is indeed obtained via Gehring type results and it is ultimately again an effect of the uniform ellipticity in the sense of (1.5) (see BCM ; dm0 ). Such effects are missing in the nearly linear growth case. Once local gradient Hölder continuity is at hand, in the non-singular/degenerate case we gain

Corollary 1 (Improved exponents)

Under assumptions (1.7), local minimizers of the functional

wΩ[𝔠(x)|Dw|log(1+|Dw|)+a(x)(|Dw|2+1)q/2]𝑑xw\mapsto\int_{\Omega}\big{[}\mathfrak{c}(x)\lvert Dw\rvert\log(1+\lvert Dw\rvert)+a(x)(\lvert Dw\rvert^{2}+1)^{q/2}\big{]}\,dx (1.8)

are locally C1,α~/2C^{1,\tilde{\alpha}/2}-regular in Ω\Omega, where α~:=min{α,α0}\tilde{\alpha}:=\min\{\alpha,\alpha_{0}\}. In particular, local minimizers of the functional in (1.2) are locally C1,α0/2C^{1,\alpha_{0}/2}-regular in Ω\Omega provided (1.7)2 holds.

We do not expect that, at least with the currently available techniques, the exponent α~/2\tilde{\alpha}/2 in Corollary 1 can be improved manth1 ; manth2 . The loss of 1/21/2 in the exponent is typical when dealing with functionals with subquadratic growth.

1.1 General statements

Theorem 1.1 and Corollary 1 are special cases of a class of results covering general functionals of the type

w𝒩(w,Ω):=Ω[F(x,Dw)+a(x)(|Dw|2+s2)q/2]𝑑x,w\mapsto\mathcal{N}(w,\Omega):=\int_{\Omega}\big{[}F(x,Dw)+a(x)(\lvert Dw\rvert^{2}+\textnormal{{s}}^{2})^{q/2}\big{]}\,dx\,, (1.9)

with F()F(\cdot) being arbitrarily close to have linear growth and s[0,1]\textnormal{{s}}\in[0,1]. This time we are thinking of integrands of the type

{F(x,z)𝔠(x)|z|Lk+1(|z|)for k0Lk+1(|z|)=log(1+Lk(|z|))for k0,L0(|z|)=|z|,\begin{cases}\ F(x,z)\equiv\mathfrak{c}(x)|z|L_{k+1}(|z|)\ \ \mbox{for $k\geq 0$}\\ \ L_{k+1}(|z|)=\log(1+L_{k}(|z|))\ \ \mbox{for $k\geq 0$}\,,\quad L_{0}(|z|)=|z|\,,\end{cases} (1.10)

where 𝔠()\mathfrak{c}(\cdot) is as in (1.7)2. For situations like this, we prepare a more technical theorem. We consider continuous integrands F:Ω×n[0,)F\colon\Omega\times\mathbb{R}^{n}\to[0,\infty) such that zF(x,z)C2(n)z\mapsto F(x,z)\in C^{2}(\mathbb{R}^{n}) for every xΩx\in\Omega and satisfying

{ν|z|g(z)F(x,z)L(|z|g(z)+1)ν|ξ|2(|z|2+1)μ/2zzF(x,z)ξ,ξ,|zzF(x,z)|Lg(|z|)(|z|2+1)1/2|zF(x,z)zF(y,z)|L|xy|α0g(|z|),\begin{cases}\ \nu|z|g(z)\leq F(x,z)\leq L(|z|g(z)+1)\\ \ \displaystyle\frac{\nu|\xi|^{2}}{(|z|^{2}+1)^{\mu/2}}\leq\langle\partial_{zz}F(x,z)\xi,\xi\rangle\,,\quad|\partial_{zz}F(x,z)|\leq\frac{Lg(|z|)}{(|z|^{2}+1)^{1/2}}\\ \ |\partial_{z}F(x,z)-\partial_{z}F(y,z)|\leq L\lvert x-y\rvert^{\alpha_{0}}g(|z|),\end{cases} (1.11)

for every choice of x,yΩx,y\in\Omega, z,ξnz,\xi\in\mathbb{R}^{n}, with α0(0,1)\alpha_{0}\in(0,1), 1μ<3/21\leq\mu<3/2 and 0<ν1L0<\nu\leq 1\leq L being fixed constants. Here g:[0,)[1,)g\colon[0,\infty)\to[1,\infty) is a non-decreasing, concave and unbounded function such that ttg(t)t\mapsto tg(t) is convex. Moreover, we assume that for every ε>0\varepsilon>0 there exists a constant cg(ε)c_{g}(\varepsilon) such that

g(t)cg(ε)tεholds for every t1.g(t)\leq c_{g}(\varepsilon)t^{\varepsilon}\quad\mbox{holds for every $t\geq 1$}\,. (1.12)

The above assumptions are sufficient to get Lipschitz continuity of minimizers. In order to get gradient local Hölder continuity we still need a technical one (see also Remark 1 below):

|zF(x,z)|L|z|holds for every |z|1 .|\partial_{z}F(x,z)|\leq L|z|\quad\mbox{holds for every $|z|\leq 1$\,.} (1.13)

For functionals as in (1.9) we have

Theorem 1.2

Let uWloc1,1(Ω)u\in W^{1,1}_{{\rm loc}}(\Omega) be a local minimizer of the functional in (1.9) under assumptions (1.7)1 and (1.11)-(1.12). There exists μmax(1,3/2)\mu_{\textnormal{max}}\in(1,3/2), depending only n,q,α,α0n,q,\alpha,\alpha_{0}, such that, if 1μ<μmax1\leq\mu<\mu_{\textnormal{max}}, then

DuL(B/2)c(B[F(x,Du)+a(x)(|Du|2+s2)q/2]dx)ϑ+c\lVert Du\rVert_{L^{\infty}(B/2)}\leq c\left(\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B}\left[F(x,Du)+a(x)(\lvert Du\rvert^{2}+\textnormal{{s}}^{2})^{q/2}\right]\,dx\right)^{\vartheta}+c (1.14)

holds whenever BBrΩB\equiv B_{r}\Subset\Omega, r1r\leq 1, where cc(data)c\equiv c(\textnormal{{data}}) and ϑϑ(n,q,α,α0)>0\vartheta\equiv\vartheta(n,q,\alpha,\alpha_{0})\linebreak>0. Moreover, assuming also (1.13) implies that DuDu is locally Hölder continuous is Ω\Omega. Finally, if in addition to all of the above we also assume that s>0\textnormal{{s}}>0 and that zzF()\partial_{zz}F(\cdot) is continuous, then uCloc1,α~/2(Ω)u\in C^{1,\tilde{\alpha}/2}_{{\rm loc}}(\Omega), with α~=min{α,α0}\tilde{\alpha}=\min\{\alpha,\alpha_{0}\}.

The notion of local minimizer we are using in this paper is classical and prescribes that a function uWloc1,1(Ω)u\in W^{1,1}_{{\rm loc}}(\Omega) is a local minimizer of 𝒩()\mathcal{N}(\cdot) if, for every ball BΩB\Subset\Omega, it happens that

{𝒩(u,B) is finite𝒩(u,B)𝒩(w,B) holds whenever wuW01,1(B).\begin{cases}\mbox{$\mathcal{N}(u,B)$ is finite}\\ \mbox{$\mathcal{N}(u,B)\leq\mathcal{N}(w,B)$ holds whenever $w-u\in W^{1,1}_{0}(B)$}\,.\end{cases} (1.15)

This implies, in particular, that 𝒩(u,Ω0)\mathcal{N}(u,\Omega_{0}) is finite whenever Ω0Ω\Omega_{0}\Subset\Omega is an open subset. In Theorem 1.2, we are using the shorthand notation

data:=(n,q,ν,L,α,α0,aC0,α,𝔠C0,α0,cg())\textnormal{{data}}:=(n,q,\nu,L,\alpha,\alpha_{0},\lVert a\rVert_{C^{0,\alpha}},\lVert\mathfrak{c}\rVert_{C^{0,\alpha_{0}}},c_{g}(\cdot)) (1.16)

that we adopt for the rest of the paper. For further notation we refer to Section 2 below. We just remark that, in order to simplify the reading, in the following we shall still denote by cc(data)c\equiv c(\textnormal{{data}}) a constant actually depending only on a subset of the parameters listed in (1.16).

Theorem 1.2 has a technical nature, especially as far as the size of μmax\mu_{\textnormal{max}} is concerned, which is not allowed to be too far from one (a natural limitation in view of the condition q<1+α/nq<1+\alpha/n). This is anyway sufficient to cover the model examples considered here, including (1.10); see Section 5. We note that in the autonomous, nearly linear case, the literature reports several interesting results for functionals wF(Dw)𝑑xw\mapsto\int F(Dw)\,dx, satisfying (1.11)1,2 with restrictions as 1μ<1+o(n)1\leq\mu<1+\texttt{o}(n), o(n)0\texttt{o}(n)\to 0 when nn\to\infty. These bounds are in fact of the type we also need to impose when selecting μmax\mu_{\textnormal{max}} (this can be checked by tracking the constant dependence in the proofs). Starting by FM conditions (1.11)2 have been used in several papers devoted to functionals with linear or nearly linear growth; see for instance bildhauer ; BF ; gmek2 and related references.

Remark 1

The technical assumption (1.13) is automatically implied, for a different constant LL, by zF(x,0n)=0n\partial_{z}F(x,0_{\mathbb{R}^{n}})=0_{\mathbb{R}^{n}}; this follows using the second inequality in (1.11)2. In turn, when F()F(\cdot) is independent of xx we can always assume that zF(0n)=0n\partial_{z}F(0_{\mathbb{R}^{n}})=0_{\mathbb{R}^{n}}. Indeed, we can switch from F(z)F(z) to F(z)zF(0n),zF(z)-\langle\partial_{z}F(0_{\mathbb{R}^{n}}),z\rangle observing that such a replacement in the integrand in (1.9) does not change the set of local minimizers. Assumption (1.13) is anyway verified by the model cases we are considering and in fact it deals with the behaviour of F()F(\cdot) at the origin, while in nonuniformly elliptic equations the main problems usually come from the behaviour at infinity.

1.2 Techniques

The techniques needed for Theorem 1.2 are entirely different from those used in the case of classical superlinear double phase functionals in (1.3) BCM ; CM and for pointwise uniformly elliptic problems with nonstandard growth conditions HO1 ; HO2 . This is essentially due to the failure of uniform ellipticity (1.5) for functionals as ()\mathcal{L}(\cdot) and 𝒩()\mathcal{N}(\cdot). On the other hand, the techniques based on De Giorgi and Moser’s methods used in the literature to deal with nearly linear growth functionals bildhauer ; BF2 ; FrS ; FM are not viable here. In fact, as in our case coefficients are Hölder continuous, it is impossible to differentiate the Euler-Lagrange equation to get gradient regularity. This is in fact the classical obstruction to direct regularity proofs one encounters when proving Schauder estimates, eventually leading to the use of perturbation methods. In turn, perturbation methods are again not working in the nonuniformly elliptic case as the available a priori estimates are not homogeneous and iteration schemes fail. Our approach is new to the context and relies on a number of ingredients, partially already introduced and exploited in piovra . Most notably, we make a delicate use of nonlinear potential theoretic methods allowing to work under the sharp bound

q<1+αn.q<1+\frac{\alpha}{n}\,. (1.17)

We employ a rather complex scheme of renormalized fractional Caccioppoli inequalities on level sets (10). These are aimed at replacing the standard Bernstein technique, and encode the essential data of the problem via a wide range of parameters; see (5.37)-(5.39). Bernstein technique is traditionally based on the fact that functions as |Du|γ|Du|^{\gamma} are subsolutions to linear elliptic equations for certain positive values of γ\gamma. This follows differentiating the original equation, something that is impossible in our setting. We therefore use a fractional and anisotropic version of the method according to which certain nonlinear functions of the gradient, namely

E(x,Du)|Du|2μ+a(x)|Du|q,E(x,Du)\approx|Du|^{2-\mu}+a(x)|Du|^{q}\,, (1.18)

belong to suitable fractional De Giorgi’s classes, i.e., they satisfy inequalities (10). The term renormalized accounts for the fact that inequalities (10) are homogeneous with respect to E(,Du)E(\cdot,Du), despite the functional 𝒩()\mathcal{N}(\cdot) is not. This costs the price of getting multiplicative constants depending on DuL\|Du\|_{L^{\infty}}. Such constants must be carefully kept under control all over the proof and reabsorbed at the very end. Notice that inequalities (10) are obtained via a dyadic/atomic decomposition technique, finding its roots in KM , that resembles the one used for Besov spaces AH ; triebel in the setting of Littlewood-Paley theory; see Section 5.6 and Remark 3. Iterating inequalities (10) via a potential theoretic version of De Giorgi’s iteration (Lemma 5) then leads to establish local Lipschitz bounds in terms of the nonlinear potentials defined in (2.11); see Proposition 2. A crucial point is that the fractional nature of the inequalities (10) allows to sharply quantify how the rate of Hölder continuity of coefficients interacts with the growth of the terms they stick to. Using this last very fact and combining the various parameters in (5.37)-(5.39) leads to determine the optimal nonlinear potentials allowing to prove Lipschitz continuity of local minimizers under the sharp bound in (1.17), but no matter how small the exponent α0\alpha_{0} is. Throughout the whole process, the structure of the functional 𝒩()\mathcal{N}(\cdot) needs to be carefully exploited at every point. This reflects in the peculiar choice of (1.18) as the leading quantity to iterate. Lipschitz regularity is as usual the focal point in nonuniformly elliptic problems, since once DuDu is (locally) bounded condition (1.5) gets automatically satisfied when zDuz\equiv Du. Indeed at this stage gradient Hölder continuity can be achieved by exploiting and extending some hidden facts from more classical regularity theory; see Section 5.10. In particular, we shall use some of the iteration schemes employed in the proof of nonlinear potential estimates for singular parabolic equations presented in KuM . The arguments developed for this point are general and can be used in other settings; see Remark 4. We finally observe that in this paper we preferred to concentrate on model cases in order to highlight the main ideas. Nevertheless we believe that the approach included here can be exported to a variety of different settings where it might yield results that are either sharp or unachievable otherwise.

2 Preliminaries

2.1 Notation

In the following we denote by cc a general constant larger than 11. Different occurrences from line to line will be still denoted by cc. Special occurrences will be denoted by c,c~c_{*},\tilde{c} or likewise. Relevant dependencies on parameters will be as usual emphasized by putting them in parentheses. We denote by Br(x0):={xn:|xx0|<r}B_{r}(x_{0}):=\{x\in\mathbb{R}^{n}:|x-x_{0}|<r\} the open ball with center x0x_{0} and radius r>0r>0; we omit denoting the center when it is not necessary, i.e., BBrBr(x0)B\equiv B_{r}\equiv B_{r}(x_{0}); this especially happens when various balls in the same context share the same center. We also denote

rBr(0):={xn:|x|<r}\mathcal{B}_{r}\equiv B_{r}(0):=\{x\in\mathbb{R}^{n}:|x|<r\}

when the ball in question is centred at the origin. Finally, with BB being a given ball with radius rr and γ\gamma being a positive number, we denote by γB\gamma B the concentric ball with radius γr\gamma r and by B/γ(1/γ)BB/\gamma\equiv(1/\gamma)B. Moreover, QinnQinn(B)Q_{\textnormal{inn}}\equiv Q_{\textnormal{inn}}(B) denotes the inner hypercube of BB, i.e., the largest hypercube, with sides parallel to the coordinate axes and concentric to BB, that is contained in BB, Qinn(B)BQ_{\textnormal{inn}}(B)\subset B. The sidelength of Qinn(B)Q_{\textnormal{inn}}(B) equals 2r/n2r/\sqrt{n}. The vector valued version of function spaces like Lp(Ω),W1,p(Ω)L^{p}(\Omega),W^{1,p}(\Omega), i.e., when the maps considered take values in k\mathbb{R}^{k}, kk\in\mathbb{N}, will be denoted by Lp(Ω;k),W1,p(Ω;k)L^{p}(\Omega;\mathbb{R}^{k}),W^{1,p}(\Omega;\mathbb{R}^{k}); when no ambiguity will arise we shall still denote Lp(Ω)Lp(Ω;k)L^{p}(\Omega)\equiv L^{p}(\Omega;\mathbb{R}^{k}) and so forth. For the rest of the paper we shall keep the notation

ω(z):=|z|2+ω2,\ell_{\omega}(z):=\sqrt{\lvert z\rvert^{2}+\omega^{2}}\,, (2.1)

for zkz\in\mathbb{R}^{k}, k1k\geq 1 and ω[0,1]\omega\in[0,1]. In particular, when k=1k=1, we have a function defined on \mathbb{R}. With 𝒰n\mathcal{U}\subset\mathbb{R}^{n} being a measurable subset with bounded positive measure 0<|𝒰|<0<|\mathcal{U}|<\infty, and with f:𝒰kf\colon\mathcal{U}\to\mathbb{R}^{k}, k1k\geq 1, being an integrable map, we denote

(f)𝒰𝒰f(x)dx:=1|𝒰|𝒰f(x)𝑑x.(f)_{\mathcal{U}}\equiv\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{\mathcal{U}}f(x)\,dx:=\frac{1}{|\mathcal{U}|}\int_{\mathcal{U}}f(x)\,dx\,.

Moreover, given a (scalar) function ff and a number κ\kappa\in\mathbb{R}, we denote

(fκ)+:=max{fκ,0}and(fκ):=max{κf,0}.(f-\kappa)_{+}:=\max\{f-\kappa,0\}\quad\mbox{and}\quad(f-\kappa)_{-}:=\max\{\kappa-f,0\}\,. (2.2)

Finally, whenever f:𝒰kf\colon\mathcal{U}\to\mathbb{R}^{k} and 𝒰\mathcal{U} is any set, we define

osc𝒰f:=supx,y𝒰|f(x)f(y)|.\textnormal{{osc}}_{\mathcal{U}}\,f:=\sup_{x,y\in\mathcal{U}}\,\lvert f(x)-f(y)\rvert\,.

As usual, with β(0,1]\beta\in(0,1] we denote

fC0,β(𝒰):=fL(𝒰)+[f]0,β;𝒰,[f]0,β;𝒰:=supx,y𝒰,xy|f(x)f(y)||xy|β.\|f\|_{C^{0,\beta}(\mathcal{U})}:=\|f\|_{L^{\infty}(\mathcal{U})}+[f]_{0,\beta;\mathcal{U}}\,,\quad[f]_{0,\beta;\mathcal{U}}:=\sup_{x,y\in\mathcal{U},x\not=y}\frac{\lvert f(x)-f(y)\rvert}{\lvert x-y\rvert^{\beta}}\,.

2.2 Auxiliary results

We shall use the vector fields Vω,p:nnV_{\omega,p}\colon\mathbb{R}^{n}\to\mathbb{R}^{n}, defined by

Vω,p:=(|z|2+ω2)(p2)/4z,0<p<andω[0,1]V_{\omega,p}:=(|z|^{2}+\omega^{2})^{(p-2)/4}z,\qquad 0<p<\infty\ \ \mbox{and}\ \ \omega\in[0,1] (2.3)

whenever znz\in\mathbb{R}^{n} (compare (ha, , (2.1)) for such extended range of pp). They satisfy

|Vω,p(z1)Vω,p(z2)|n,p(|z1|2+|z2|2+ω2)(p2)/4|z1z2|\lvert V_{\omega,p}(z_{1})-V_{\omega,p}(z_{2})\rvert\approx_{n,p}(\lvert z_{1}\rvert^{2}+\lvert z_{2}\rvert^{2}+\omega^{2})^{(p-2)/4}\lvert z_{1}-z_{2}\rvert (2.4)

that holds for every z1,z2nz_{1},z_{2}\in\mathbb{R}^{n}, ω[0,1]\omega\in[0,1] and p>0p>0; for this we refer to (ha, , Lemma 2.1). We shall also use the following, trivial

(|z1|2+|z2|2+ω2)γ/201(|z2+τ(z1z2)|2+ω2)γ/2𝑑τ(\lvert z_{1}\rvert^{2}+\lvert z_{2}\rvert^{2}+\omega^{2})^{-\gamma/2}\lesssim\int_{0}^{1}(\lvert z_{2}+\tau(z_{1}-z_{2})\rvert^{2}+\omega^{2})^{-\gamma/2}\,d\tau (2.5)

that holds for every γ0\gamma\geq 0 and whenever |z1|+|z2|>0|z_{1}|+|z_{2}|>0. When restricting from above the range of γ\gamma, the inequality becomes double-sided, i.e.

(|z1|2+|z2|2+ω2)γ/2n,γ01(|z2+τ(z1z2)|2+ω2)γ/2𝑑τ,(\lvert z_{1}\rvert^{2}+\lvert z_{2}\rvert^{2}+\omega^{2})^{-\gamma/2}\approx_{n,\gamma}\int_{0}^{1}(\lvert z_{2}+\tau(z_{1}-z_{2})\rvert^{2}+\omega^{2})^{-\gamma/2}\,d\tau\,, (2.6)

that instead holds for every γ<1\gamma<1; see ha , (giamod, , Section 2). With BΩB\subset\Omega being a ball and σ(0,1)\sigma\in(0,1), we let

{aσ(x):=a(x)+σai(B):=infxBa(x)andaσ,i(B):=ai(B)+σ\begin{cases}a_{\sigma}(x):=a(x)+\sigma\\ a_{\textnormal{i}}(B):=\inf_{x\in B}a(x)\ \ \mbox{and}\ \ a_{\sigma,\textnormal{i}}(B):=a_{\textnormal{i}}(B)+\sigma\end{cases} (2.7)

and define

{𝒱ω,σ2(x,z1,z2):=|V1,2μ(z1)V1,2μ(z2)|2+aσ(x)|Vω,q(z1)Vω,q(z2)|2𝒱ω,σ,i2(z1,z2;B):=|V1,2μ(z1)V1,2μ(z2)|2+aσ,i(B)|Vω,q(z1)Vω,q(z2)|2\begin{cases}\mathcal{V}_{\omega,\sigma}^{2}(x,z_{1},z_{2}):=\lvert V_{1,2-\mu}(z_{1})-V_{1,2-\mu}(z_{2})\rvert^{2}\\ \hskip 84.50467pt+a_{\sigma}(x)\lvert V_{\omega,q}(z_{1})-V_{\omega,q}(z_{2})\rvert^{2}\\ \mathcal{V}_{\omega,\sigma,\textnormal{i}}^{2}(z_{1},z_{2};B):=\lvert V_{1,2-\mu}(z_{1})-V_{1,2-\mu}(z_{2})\rvert^{2}\\ \hskip 85.35826pt+a_{\sigma,\textnormal{i}}(B)\lvert V_{\omega,q}(z_{1})-V_{\omega,q}(z_{2})\rvert^{2}\end{cases} (2.8)

where z1,z2nz_{1},z_{2}\in\mathbb{R}^{n}, xΩx\in\Omega. A consequence of assumptions (1.11) is the following:

Lemma 1

Let F:Ω×n[0,)F\colon\Omega\times\mathbb{R}^{n}\to[0,\infty) be as in (1.11). Then |zF(x,z)|cg(|z|)|\partial_{z}F(x,z)|\leq cg(|z|) holds for every (x,z)Ω(x,z)\in\Omega, where cc depends on LL and the constant cg(1)c_{g}(1) defined in (1.12).

Proof

This is a version of (M1, , Lemma 2.1) and the proof mimics that of Marcellini. Specifically, there we take |h|=1+|z||h|=1+|z| and, after using (1.11)1 as in M1 , we estimate g(|z±hei|)g(|z|+h)g(|z|)+g(h)2g(|z|)+g(1)cg(|z|)g(|z\pm he_{i}|)\leq g(|z|+h)\leq g(|z|)+g(h)\leq 2g(|z|)+g(1)\leq cg(|z|) (here {ek}\{e_{k}\} denotes the standard basis of n\mathbb{R}^{n}). Here we have used that g()1g(\cdot)\geq 1 is non-decreasing and sublinear (since it is concave and non-negative) and finally (1.12) to get g(1)cg(1)g(|z|)g(1)\leq c_{g}(1)\leq g(|z|).

Finally, a classical iteration lemma (giu, , Lemma 6.1).

Lemma 2

Let h:[t,s]h\colon[t,s]\to\mathbb{R} be a non-negative and bounded function, and let a,b,γa,b,\gamma be non-negative numbers. Assume that the inequality h(τ1)(1/2)h(τ2)+(τ2τ1)γa+b,h(\tau_{1})\leq(1/2)h(\tau_{2})+(\tau_{2}-\tau_{1})^{-\gamma}a+b, holds whenever tτ1<τ2st\leq\tau_{1}<\tau_{2}\leq s. Then h(t)c(γ)[a(st)γ+b]h(t)\leq c(\gamma)[a(s-t)^{-\gamma}+b], holds too.

2.3 Fractional spaces

The finite difference operator τh:L1(Ω;k)L1(Ω|h|;k)\tau_{h}\colon L^{1}(\Omega;\mathbb{R}^{k})\to L^{1}(\Omega_{|h|};\mathbb{R}^{k}) is defined as τhw(x):=w(x+h)w(x)\tau_{h}w(x):=w(x+h)-w(x) for xΩ|h|x\in\Omega_{\lvert h\rvert}, for every τL1(Ω)\tau\in L^{1}(\Omega) and where Ω|h|:={xΩ:dist(x,Ω)>|h|}\Omega_{|h|}:=\{x\in\Omega\,:\,\,{\rm dist}(x,\partial\Omega)>|h|\}. With β(0,1)\beta\in(0,1), s[1,)s\in[1,\infty), kk\in\mathbb{N}, n2n\geq 2, and with Ωn\Omega\subset\mathbb{R}^{n} being an open subset, the space Wβ,s(Ω;k)W^{\beta,s}(\Omega;\mathbb{R}^{k}) is defined as the set of maps w:Ωkw\colon\Omega\to\mathbb{R}^{k} such that

wWβ,s(Ω;k)\displaystyle\|w\|_{W^{\beta,s}(\Omega;\mathbb{R}^{k})} :=wLs(Ω;k)+(ΩΩ|w(x)w(y)|s|xy|n+βs𝑑x𝑑y)1/s\displaystyle:=\|w\|_{L^{s}(\Omega;\mathbb{R}^{k})}+\left(\int_{\Omega}\int_{\Omega}\frac{\lvert w(x)-w(y)\rvert^{s}}{\lvert x-y\rvert^{n+\beta s}}\,dx\,dy\right)^{1/s}
=:wLs(Ω;k)+[w]β,s;Ω<.\displaystyle=:\|w\|_{L^{s}(\Omega;\mathbb{R}^{k})}+[w]_{\beta,s;\Omega}<\infty\,.

When considering regular domains, as for instance the ball 1/2\mathcal{B}_{1/2} (this is the only case we are going to consider here in this respect), the embedding inequality

wLnsnsβ(1/2;k)c(n,s,β)wWβ,s(1/2;k)\lVert w\rVert_{L^{\frac{ns}{n-s\beta}}(\mathcal{B}_{1/2};\mathbb{R}^{k})}\leq c(n,s,\beta)\lVert w\rVert_{W^{\beta,s}(\mathcal{B}_{1/2};\mathbb{R}^{k})} (2.9)

holds provided s1,β(0,1)s\geq 1,\beta\in(0,1), kk\in\mathbb{N} and sβ<ns\beta<n. For this and basic results concerning Fractional Sobolev spaces we refer to guide and related references. A characterization of fractional Sobolev spaces via finite difference is contained in the following lemma, that can be retrieved for instance from dm .

Lemma 3

Let BϱBrnB_{\varrho}\Subset B_{r}\subset\mathbb{R}^{n} be concentric balls with r1r\leq 1, wLs(Br;k)w\in L^{s}(B_{r};\mathbb{R}^{k}), s1s\geq 1 and assume that, for α(0,1]\alpha_{*}\in(0,1], S1S\geq 1, there holds

τhwLs(Bϱ;k)S|h|α for every hn with 0<|h|rϱK, where K1.\lVert\tau_{h}w\rVert_{L^{s}(B_{\varrho};\mathbb{R}^{k})}\leq S\lvert h\rvert^{\alpha_{*}}\ \mbox{ for every $h\in\mathbb{R}^{n}$ with $0<\lvert h\rvert\leq\frac{r-\varrho}{K}$, where $K\geq 1$}\,.

Then

wWβ,s(Bϱ;k)\displaystyle\lVert w\rVert_{W^{\beta,s}(B_{\varrho};\mathbb{R}^{k})} c(αβ)1/s(rϱK)αβS\displaystyle\leq\frac{c}{(\alpha_{*}-\beta)^{1/s}}\left(\frac{r-\varrho}{K}\right)^{\alpha_{*}-\beta}S
+c(Krϱ)n/s+βwLs(Br;k)\displaystyle\quad+c\left(\frac{K}{r-\varrho}\right)^{n/s+\beta}\lVert w\rVert_{L^{s}(B_{r};\mathbb{R}^{k})} (2.10)

holds for every β<α\beta<\alpha_{*}, with cc(n,s)c\equiv c(n,s).

2.4 Nonlinear potentials

We shall use a family of nonlinear potentials that, in their essence, goes back to the fundamental work of Havin & Mazya HM . We have recently used such potentials in the setting of nonuniformly elliptic problems in BM ; camel ; camel2 ; ciccio ; piovra and in fact we shall in particular use (piovra, , Section 4) as a main source of tools in this respect. With t,σ>0t,\sigma>0, m,θ0m,\theta\geq 0 being fixed parameters, and fL1(Br(x0))f\in L^{1}(B_{r}(x_{0})) being such that |f|mL1(Br(x0))\lvert f\rvert^{m}\in L^{1}(B_{r}(x_{0})), where Br(x0)nB_{r}(x_{0})\subset\mathbb{R}^{n} is a ball, the nonlinear Havin-Mazya-Wolff type potential 𝐏t,σm,θ(f;){\bf P}_{t,\sigma}^{m,\theta}(f;\cdot) is defined by

𝐏t,σm,θ(f;x0,r):=0rϱσ(Bϱ(x0)|f|mdx)θ/tdϱϱ.{\bf P}_{t,\sigma}^{m,\theta}(f;x_{0},r):=\int_{0}^{r}\varrho^{\sigma}\left(\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B_{\varrho}(x_{0})}\lvert f\rvert^{m}\,dx\right)^{\theta/t}\frac{\,d\varrho}{\varrho}\,. (2.11)

Suitable integrability properties of ff guarantee that 𝐏t,σm,θ(f;){\bf P}_{t,\sigma}^{m,\theta}(f;\cdot) is bounded, as stated in

Lemma 4

Let t,σ,θ>0t,\sigma,\theta>0 be numbers such that

nθ>tσ.n\theta>t\sigma\,. (2.12)

Let BϱBϱ+r0nB_{\varrho}\subset B_{\varrho+r_{0}}\subset\mathbb{R}^{n} be concentric balls with ϱ,r01\varrho,r_{0}\leq 1 and fL1(Bϱ+r0)f\in L^{1}(B_{\varrho+r_{0}}) be a function such that |f|mL1(Bϱ+r0)\lvert f\rvert^{m}\in L^{1}(B_{\varrho+r_{0}}), for some m>0m>0. Then

𝐏t,σm,θ(,r0)L(Bϱ)cfLγ(Bϱ+r0)mθ/tholds for every γ>nmθtσ>0\lVert\mathbf{P}^{m,\theta}_{t,\sigma}(\cdot,r_{0})\rVert_{L^{\infty}(B_{\varrho})}\leq c\|f\|_{L^{\gamma}(B_{\varrho}+r_{0})}^{m\theta/t}\quad\mbox{holds for every $\gamma>\frac{nm\theta}{t\sigma}>0$} (2.13)

and with cc(n,t,σ,m,θ,γ)c\equiv c(n,t,\sigma,m,\theta,\gamma).

Note that in (2.13) we are not requiring that γ1\gamma\geq 1 and any value γ>0\gamma>0 is allowed. Next lemma is a version of (piovra, , Lemma 2.3) and takes information from various results from Nonlinear Potential Theory starting by the seminal paper kilp . It is essentially a nonlinear potential theoretic version of De Giorgi’s iteration. Notice that the very crucial point in the next statement is the tracking, in the various inequalities, of the explicit dependence on the constants M0,MiM_{0},M_{i}.

Lemma 5

Let Br0(x0)nB_{r_{0}}(x_{0})\subset\mathbb{R}^{n} be a ball, n2n\geq 2, and, for j{1,2,3}j\in\{1,2,3\}, consider functions fjf_{j}, |fj|mjL1(B2r0(x0))|f_{j}|^{m_{j}}\in L^{1}(B_{2r_{0}}(x_{0})), and constants χ>1\chi>1, σj,mj,θj>0\sigma_{j},m_{j},\theta_{j}>0 and c,M0>0c_{*},M_{0}>0, κ0,Mj0\kappa_{0},M_{j}\geq 0. Assume that wL2(Br0(x0))w\in L^{2}(B_{r_{0}}(x_{0})) is such that for all κκ0\kappa\geq\kappa_{0}, and for every ball Bϱ(x0)Br0(x0)B_{\varrho}(x_{0})\subseteq B_{r_{0}}(x_{0}), the inequality

(Bϱ/2(x0)(wκ)+2χdx)1/χ\displaystyle\left(\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B_{\varrho/2}(x_{0})}(w-\kappa)_{+}^{2\chi}\,dx\right)^{1/\chi} cM02Bϱ(x0)(wκ)+2dx\displaystyle\leq c_{*}M_{0}^{2}\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B_{\varrho}(x_{0})}(w-\kappa)_{+}^{2}\,dx
+cj=13Mj2ϱ2σj(Bϱ(x0)|fj|mjdx)θj\displaystyle\qquad+c_{*}\sum_{j=1}^{3}M_{j}^{2}\varrho^{2\sigma_{j}}\left(\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B_{\varrho}(x_{0})}\lvert f_{j}\rvert^{m_{j}}\,dx\right)^{\theta_{j}} (2.14)

holds (recall the notation in (2.2)). If x0x_{0} is a Lebesgue point of ww, then

w(x0)\displaystyle w(x_{0}) κ0+cM0χχ1(Br0(x0)(wκ0)+2dx)1/2\displaystyle\leq\kappa_{0}+cM_{0}^{\frac{\chi}{\chi-1}}\left(\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B_{r_{0}}(x_{0})}(w-\kappa_{0})_{+}^{2}\,dx\right)^{1/2}
+cM01χ1j=13Mj𝐏2,σjmj,θj(fj;x0,2r0)\displaystyle\qquad+cM_{0}^{\frac{1}{\chi-1}}\sum_{j=1}^{3}M_{j}\mathbf{P}^{m_{j},\theta_{j}}_{2,\sigma_{j}}(f_{j};x_{0},2r_{0}) (2.15)

holds with cc(n,χ,σj,θj,c)c\equiv c(n,\chi,\sigma_{j},\theta_{j},c_{*}).

3 Auxiliary integrands and their eigenvalues.

With reference to a general functional of the type 𝒩()\mathcal{N}(\cdot) in (1.9), considered under the assumptions of Theorem 1.2 (in particular implying that q<3/2q<3/2), we define the integrand

H(x,z):=F(x,z)+a(x)|z|qH(x,z):=F(x,z)+a(x)\lvert z\rvert^{q} (3.1)

for znz\in\mathbb{R}^{n}, xΩx\in\Omega. We fix an arbitrary ball BΩB\subset\Omega centred at xcx_{\rm c} and, with ω[0,1]\omega\in[0,1] and σ(0,1)\sigma\in(0,1), recalling (2.1) and (2.7), we define

{Hω,σ(x,z):=F(x,z)+aσ(x)[ω(z)]qHω,σ,i(z)Hω,σ,i(z;B):=F(xc,z)+aσ,i(B)[ω(z)]qω,σ,i(t)ω,σ,i(t;B):=tg(t)+aσ,i(B)[t2+ω2]q/2+1,t0.\begin{cases}H_{\omega,\sigma}(x,z):=F(x,z)+a_{\sigma}(x)[\ell_{\omega}(z)]^{q}\\ H_{\omega,\sigma,\textnormal{i}}(z)\equiv H_{\omega,\sigma,\textnormal{i}}(z;B):=F(x_{\rm c},z)+a_{\sigma,\textnormal{i}}(B)[\ell_{\omega}(z)]^{q}\\ \mathbb{H}_{\omega,\sigma,\textnormal{i}}(t)\equiv\mathbb{H}_{\omega,\sigma,\textnormal{i}}(t;B):=tg(t)+a_{\sigma,\textnormal{i}}(B)[t^{2}+\omega^{2}]^{q/2}+1,\ \ t\geq 0\,.\end{cases} (3.2)

It follows that

tω,σ,i(t) is non-decreasing.\mbox{$t\mapsto\mathbb{H}_{\omega,\sigma,\textnormal{i}}(t)$ is non-decreasing}\,. (3.3)

Note that abbreviations as Hω,σ,i(z)Hω,σ,i(z;B)H_{\omega,\sigma,\textnormal{i}}(z)\equiv H_{\omega,\sigma,\textnormal{i}}(z;B), used in (3.2), will take place whenever there will be no ambiguity concerning the identity of the ball BB in question. A similar shortened notation will be adopted for the case of further quantities depending on a single ball BB. We next introduce two functions, λω,σ,Λω,σ:Ω×[0,)[0,)\lambda_{\omega,\sigma},\Lambda_{\omega,\sigma}\colon\Omega\times[0,\infty)\to[0,\infty), bound to describe the behaviour of the eigenvalues of zzHω,σ()\partial_{zz}H_{\omega,\sigma}(\cdot). These are (recall that q,μ<3/2q,\mu<3/2)

{λω,σ(x,|z|):=(|z|2+1)μ/2+(q1)aσ(x)[ω(z)]q2Λω,σ(x,|z|):=(|z|2+1)1/2g(|z|)+aσ(x)[ω(z)]q2.\begin{cases}\lambda_{\omega,\sigma}(x,\lvert z\rvert):=(|z|^{2}+1)^{-\mu/2}+(q-1)a_{\sigma}(x)[\ell_{\omega}(z)]^{q-2}\\ \Lambda_{\omega,\sigma}(x,\lvert z\rvert):=(|z|^{2}+1)^{-1/2}g(|z|)+a_{\sigma}(x)[\ell_{\omega}(z)]^{q-2}\,.\end{cases} (3.4)

It then follows that

{σ[ω(z)]qHω,σ(x,z)c([ω(z)]q+1)λω,σ(x,|z|)|ξ|2czzHω,σ(x,z)ξ,ξ,|zzHω,σ(x,z)|cΛω,σ(x,|z|)𝒱ω,σ2(x,z1,z2)czHω,σ(x,z1)zHω,σ(x,z2),z1z2|zHω,σ(x,z)zHω,σ(y,z)|c|xy|α0g(|z|)+c|xy|α[ω(z)]q1\displaystyle\begin{cases}\ \sigma[\ell_{\omega}(z)]^{q}\leq H_{\omega,\sigma}(x,z)\leq c\left([\ell_{\omega}(z)]^{q}+1\right)\\ \ \lambda_{\omega,\sigma}(x,\lvert z\rvert)\lvert\xi\rvert^{2}\leq c\langle\partial_{zz}H_{\omega,\sigma}(x,z)\xi,\xi\rangle\,,\quad\lvert\partial_{zz}H_{\omega,\sigma}(x,z)\rvert\leq c\Lambda_{\omega,\sigma}(x,\lvert z\rvert)\\ \ \mathcal{V}_{\omega,\sigma}^{2}(x,z_{1},z_{2})\leq c\langle\partial_{z}H_{\omega,\sigma}(x,z_{1})-\partial_{z}H_{\omega,\sigma}(x,z_{2}),z_{1}-z_{2}\rangle\\ \ \lvert\partial_{z}H_{\omega,\sigma}(x,z)-\partial_{z}H_{\omega,\sigma}(y,z)\rvert\leq c\lvert x-y\rvert^{\alpha_{0}}g(|z|)+c\lvert x-y\rvert^{\alpha}[\ell_{\omega}(z)]^{q-1}\end{cases} (3.5)

hold for all x,yΩx,y\in\Omega, z,z1,z2,ξnz,z_{1},z_{2},\xi\in\mathbb{R}^{n}, where cc(data)c\equiv c(\textnormal{{data}}) (recall the definition of 𝒱ω,σ\mathcal{V}_{\omega,\sigma} in (2.8) and (2.3)). Notice that (3.5)1,2 directly follow from (1.11)1,2, respectively, and (1.12) used with ε=q1\varepsilon=q-1. The inequality in (3.5)3\eqref{rege.2}_{3} is a straightforward consequence of (2.4)-(2.5) and (3.5)2 (see Remark 2 below). Finally, (3.5)4 is a consequence of (1.11)3. When ω>0\omega>0, by (1.12) and (3.5)2 the integrand Hω,σ()H_{\omega,\sigma}(\cdot) also satisfies, for all x,yΩx,y\in\Omega, |xy|1\lvert x-y\rvert\leq 1 and z,ξnz,\xi\in\mathbb{R}^{n}

{σ(q1)[1(z)]q2|ξ|2czzHω,σ(x,z)ξ,ξ|zHω,σ(x,z)|1(z)+|zzHω,σ(x,z)|[1(z)]2c[1(z)]q|zHω,σ(x,z)zHω,σ(y,z)|c|xy|α~[1(z)]q1,\displaystyle\begin{cases}\ \sigma(q-1)[\ell_{1}(z)]^{q-2}\lvert\xi\rvert^{2}\leq c\langle\partial_{zz}H_{\omega,\sigma}(x,z)\xi,\xi\rangle\\ \ \lvert\partial_{z}H_{\omega,\sigma}(x,z)\rvert\ell_{1}(z)+\lvert\partial_{zz}H_{\omega,\sigma}(x,z)\rvert[\ell_{1}(z)]^{2}\leq c[\ell_{1}(z)]^{q}\\ \ \lvert\partial_{z}H_{\omega,\sigma}(x,z)-\partial_{z}H_{\omega,\sigma}(y,z)\rvert\leq c\lvert x-y\rvert^{\tilde{\alpha}}[\ell_{1}(z)]^{q-1}\,,\end{cases} (3.6)

with α~=min{α,α0}\tilde{\alpha}=\min\{\alpha,\alpha_{0}\} and cc(data,ω).c\equiv c(\textnormal{{data}},\omega). Note that, on the contrary of the previous displays, the constant cc in (3.6) depends on ω\omega and that for (3.6)2 we also need Lemma 1 and (3.5)2. As for Hω,σ,i()H_{\omega,\sigma,\textnormal{i}}(\cdot), by defining

{λω,σ,i(|z|):=(|z|2+1)μ/2+(q1)aσ,i(B)[ω(z)]q2Λω,σ,i(|z|):=(|z|2+1)1/2g(|z|)+aσ,i(B)[ω(z)]q2\begin{cases}\lambda_{\omega,\sigma,\textnormal{i}}(\lvert z\rvert):=(|z|^{2}+1)^{-\mu/2}+(q-1)a_{\sigma,\textnormal{i}}(B)[\ell_{\omega}(z)]^{q-2}\\ \Lambda_{\omega,\sigma,\textnormal{i}}(\lvert z\rvert):=\ (|z|^{2}+1)^{-1/2}g(|z|)+a_{\sigma,\textnormal{i}}(B)[\ell_{\omega}(z)]^{q-2}\end{cases} (3.7)

we have that

{σ[ω(z)]qHω,σ,i(z)c([ω(z)]q+1)|zHω,σ,i(z)|cg(|z|)+caσ,i(B)[ω(z)]q2|z|λω,σ,i(|z|)|ξ|2czzHω,σ,i(z)ξ,ξ,|zzHω,σ,i(z)|cΛω,σ,i(|z|)𝒱ω,σ,i2(z1,z2;B)czHω,σ,i(z1)zHω,σ,i(z2),z1z2\begin{cases}\ \sigma[\ell_{\omega}(z)]^{q}\leq H_{\omega,\sigma,\textnormal{i}}(z)\leq c([\ell_{\omega}(z)]^{q}+1)\\ \ \lvert\partial_{z}H_{\omega,\sigma,\textnormal{i}}(z)\rvert\leq cg(|z|)+ca_{\sigma,\textnormal{i}}(B)[\ell_{\omega}(z)]^{q-2}|z|\\ \ \lambda_{\omega,\sigma,\textnormal{i}}(\lvert z\rvert)\lvert\xi\rvert^{2}\leq c\langle\partial_{zz}H_{\omega,\sigma,\textnormal{i}}(z)\xi,\xi\rangle,\quad\lvert\partial_{zz}H_{\omega,\sigma,\textnormal{i}}(z)\rvert\leq c\Lambda_{\omega,\sigma,\textnormal{i}}(\lvert z\rvert)\\ \ \mathcal{V}^{2}_{\omega,\sigma,\textnormal{i}}(z_{1},z_{2};B)\leq c\langle\partial_{z}H_{\omega,\sigma,\textnormal{i}}(z_{1})-\partial_{z}H_{\omega,\sigma,\textnormal{i}}(z_{2}),z_{1}-z_{2}\rangle\end{cases} (3.8)

holds whenever z,z1,z2,ξnz,z_{1},z_{2},\xi\in\mathbb{R}^{n}, with cc(data)c\equiv c(\textnormal{{data}}); for (3.8)2 we have used Lemma 1. In particular, it holds that

Λω,σ,i(|z|)λω,σ,i(|z|)(|z|2+1)μ12g(|z|)+1q1c(q)(|z|2+1)μ12g(|z|)\frac{\Lambda_{\omega,\sigma,\textnormal{i}}(\lvert z\rvert)}{\lambda_{\omega,\sigma,\textnormal{i}}(\lvert z\rvert)}\leq(|z|^{2}+1)^{\frac{\mu-1}{2}}g(\lvert z\rvert)+\frac{1}{q-1}\leq c(q)(|z|^{2}+1)^{\frac{\mu-1}{2}}g(\lvert z\rvert) (3.9)

(recall that g()1g(\cdot)\geq 1). As in (3.6), when ω>0\omega>0 the function Hω,σ,i()H_{\omega,\sigma,\textnormal{i}}(\cdot) satisfies

{σ(q1)[1(z)]q2|ξ|2czzHω,σ,i(z)ξ,ξ|zHω,σ,i(z)|1(z)+|zzHω,σ,i(z)|[1(z)]2c[1(z)]q\begin{cases}\ \sigma(q-1)[\ell_{1}(z)]^{q-2}\lvert\xi\rvert^{2}\leq c\langle\partial_{zz}H_{\omega,\sigma,\textnormal{i}}(z)\xi,\xi\rangle\\ \ \lvert\partial_{z}H_{\omega,\sigma,\textnormal{i}}(z)\rvert\ell_{1}(z)+\lvert\partial_{zz}H_{\omega,\sigma,\textnormal{i}}(z)\rvert[\ell_{1}(z)]^{2}\leq c[\ell_{1}(z)]^{q}\end{cases} (3.10)

for all z,ξnz,\xi\in\mathbb{R}^{n}, with cc(data,ω).c\equiv c(\textnormal{{data}},\omega). Two quantities that will have an important role in the forthcoming computations are, for xΩx\in\Omega and t0t\geq 0

{Eω,σ(x,t):=0tλω,σ(x,s)sdsEω,σ,i(t)Eω,σ,i(t;B):=0tλω,σ,i(s;B)sds.\begin{cases}\displaystyle E_{\omega,\sigma}(x,t):=\int_{0}^{t}\lambda_{\omega,\sigma}(x,s)s\,{\rm d}s\\ \displaystyle E_{\omega,\sigma,\textnormal{i}}(t)\equiv E_{\omega,\sigma,\textnormal{i}}(t;B):=\int_{0}^{t}\lambda_{\omega,\sigma,\textnormal{i}}(s;B)s\,{\rm d}s\,.\end{cases} (3.11)

It follows that (recall that μ<2\mu<2)

{Eω,σ(x,t):=12μ([1(t)]2μ1)+(11/q)aσ(x)([ω(t)]qωq)Eω,σ,i(t):=12μ([1(t)]2μ1)+(11/q)aσ,i(B)([ω(t)]qωq)\begin{cases}E_{\omega,\sigma}(x,t):=\frac{1}{2-\mu}\left([\ell_{1}(t)]^{2-\mu}-1\right)+(1-1/q)a_{\sigma}(x)\left([\ell_{\omega}(t)]^{q}-\omega^{q}\right)\\ E_{\omega,\sigma,\textnormal{i}}(t):=\frac{1}{2-\mu}\left([\ell_{1}(t)]^{2-\mu}-1\right)+(1-1/q)a_{\sigma,\textnormal{i}}(B)\left([\ell_{\omega}(t)]^{q}-\omega^{q}\right)\end{cases} (3.12)

and this leads to define

{E~ω,σ(x,t):=12μ[1(t)]2μ+(11/q)aσ(x)[ω(t)]qE~ω,σ,i(t):=12μ[1(t)]2μ+(11/q)aσ,i(B)[ω(t)]q.\begin{cases}\tilde{E}_{\omega,\sigma}(x,t):=\frac{1}{2-\mu}[\ell_{1}(t)]^{2-\mu}+(1-1/q)a_{\sigma}(x)[\ell_{\omega}(t)]^{q}\\ \tilde{E}_{\omega,\sigma,\textnormal{i}}(t):=\frac{1}{2-\mu}[\ell_{1}(t)]^{2-\mu}+(1-1/q)a_{\sigma,\textnormal{i}}(B)[\ell_{\omega}(t)]^{q}\,.\end{cases} (3.13)

We record a few basic properties of these functions in the following:

Lemma 6

The following holds about the functions in (3.11), whenever BΩB\subseteq\Omega is a ball:

  • For every s,t[0,)s,t\in[0,\infty)

    |Eω,σ,i(s)Eω,σ,i(t)|\displaystyle\lvert E_{\omega,\sigma,\textnormal{i}}(s)-E_{\omega,\sigma,\textnormal{i}}(t)\rvert
    [(s2+t2+1)(1μ)/2+aσ,i(B)(s2+t2+ω2)(q1)/2]|st|.\displaystyle\ \leq\left[(s^{2}+t^{2}+1)^{(1-\mu)/2}+a_{\sigma,\textnormal{i}}(B)(s^{2}+t^{2}+\omega^{2})^{(q-1)/2}\right]\lvert s-t\rvert. (3.14)
  • For all xBx\in B

    |Eω,σ(x,t)Eω,σ,i(t)|(11/q)|a(x)ai(B)|([ω(t)]qωq).\lvert E_{\omega,\sigma}(x,t)-E_{\omega,\sigma,\textnormal{i}}(t)\rvert\leq(1-1/q)\lvert a(x)-a_{\textnormal{i}}(B)\rvert\left([\ell_{\omega}(t)]^{q}-\omega^{q}\right). (3.15)
  • There exists constant TT(μ)1T\equiv T(\mu)\geq 1 such that

    t2[Eω,σ(x,t)]12μandt2[Eω,σ,i(t)]12μt\leq 2[E_{\omega,\sigma}(x,t)]^{\frac{1}{2-\mu}}\quad\mbox{and}\quad t\leq 2[E_{\omega,\sigma,\textnormal{i}}(t)]^{\frac{1}{2-\mu}} (3.16)

    hold for all xΩx\in\Omega, tTt\geq T.

  • There exists a constant cc(μ,q,ν)c\equiv c(\mu,q,\nu) such that

    {|z|+E~ω,σ(x,|z|)c[Hω,σ(x,z)+1]|z|+E~ω,σ,i(|z|)c[Hω,σ,i(z)+1]\begin{cases}|z|+\tilde{E}_{\omega,\sigma}(x,|z|)\leq c[H_{\omega,\sigma}(x,z)+1]\\ |z|+\tilde{E}_{\omega,\sigma,\textnormal{i}}(|z|)\leq c[H_{\omega,\sigma,\textnormal{i}}(z)+1]\end{cases} (3.17)

    hold for every xΩx\in\Omega and znz\in\mathbb{R}^{n}.

Proof

The only inequality deserving some comments is (6), the other being direct consequences of the definitions in (3.12)-(3.13). Notice that

|Eω,σ,i(t)|c[1(t)]1μ+caσ(x)[ω(t)]q1|E_{\omega,\sigma,\textnormal{i}}^{\prime}(t)|\leq c[\ell_{1}(t)]^{1-\mu}+ca_{\sigma}(x)[\ell_{\omega}(t)]^{q-1}

with cc(q,μ)c\equiv c(q,\mu). Using this last inequality we have

|Eω,σ,i(s)Eω,σ,i(t)|\displaystyle\lvert E_{\omega,\sigma,\textnormal{i}}(s)-E_{\omega,\sigma,\textnormal{i}}(t)\rvert 01|Eω,σ,i(t+τ(st))|𝑑τ|st|\displaystyle\leq\int_{0}^{1}\lvert E_{\omega,\sigma,\textnormal{i}}^{\prime}(t+\tau(s-t))\rvert\,d\tau|s-t|
c01(|t+τ(st)|2+1)1μ2𝑑τ\displaystyle\leq c\int_{0}^{1}(\lvert t+\tau(s-t)\rvert^{2}+1)^{\frac{1-\mu}{2}}\,d\tau
+caσ,i(B)01(|t+τ(st)|2+ω2)q12𝑑τ\displaystyle\quad+ca_{\sigma,\textnormal{i}}(B)\int_{0}^{1}(\lvert t+\tau(s-t)\rvert^{2}+\omega^{2})^{\frac{q-1}{2}}\,d\tau

so that (6) follows applying (2.6) with γμ1\gamma\equiv\mu-1 and γ1q\gamma\equiv 1-q.

Remark 2

The proof of (3.5)3\eqref{rege.2}_{3} is rather standard, but we report it for completeness as the exponents involved here are not the usual ones. Using, in turn, (2.4) (with p=2μ,qp=2-\mu,q) and (2.5) (with γ=μ,2q\gamma=\mu,2-q), and finally the first inequality in (3.5)2\eqref{rege.2}_{2}, we have

𝒱ω,σ2(x,z1,z2)\displaystyle\mathcal{V}_{\omega,\sigma}^{2}(x,z_{1},z_{2}) c(|z1|2+|z2|2+1)μ/2|z1z2|2\displaystyle\leq c(\lvert z_{1}\rvert^{2}+\lvert z_{2}\rvert^{2}+1)^{-\mu/2}|z_{1}-z_{2}|^{2}
+caσ(x)(|z1|2+|z2|2+ω2)(q2)/2|z1z2|2\displaystyle\qquad+ca_{\sigma}(x)(\lvert z_{1}\rvert^{2}+\lvert z_{2}\rvert^{2}+\omega^{2})^{(q-2)/2}|z_{1}-z_{2}|^{2}
c01(|z2+τ(z1z2)|2+1)μ/2𝑑τ|z1z2|2\displaystyle\leq c\int_{0}^{1}(\lvert z_{2}+\tau(z_{1}-z_{2})\rvert^{2}+1)^{-\mu/2}\,d\tau|z_{1}-z_{2}|^{2}
+caσ(x)01(|z2+τ(z1z2)|2+ω2)(q2)/2𝑑τ|z1z2|2\displaystyle\qquad+ca_{\sigma}(x)\int_{0}^{1}(\lvert z_{2}+\tau(z_{1}-z_{2})\rvert^{2}+\omega^{2})^{(q-2)/2}\,d\tau|z_{1}-z_{2}|^{2}
c01zzHω,σ(x,z2+τ(z1z2))(z1z2),z1z2𝑑τ\displaystyle\leq c\int_{0}^{1}\langle\partial_{zz}H_{\omega,\sigma}(x,z_{2}+\tau(z_{1}-z_{2}))(z_{1}-z_{2}),z_{1}-z_{2}\rangle\,d\tau
=czHω,σ(x,z1)zHω,σ(x,z2),z1z2.\displaystyle=c\langle\partial_{z}H_{\omega,\sigma}(x,z_{1})-\partial_{z}H_{\omega,\sigma}(x,z_{2}),z_{1}-z_{2}\rangle\,.

4 Lipschitz estimates with small anisotropicity

Let BτBτ(xc)ΩB_{\tau}\equiv B_{\tau}(x_{\rm c})\Subset\Omega be a ball with τ1\tau\leq 1 and let u0W1,(Bτ)u_{0}\in W^{1,\infty}(B_{\tau}). With ω(0,1]\omega\in(0,1], we define vω,σu0+W01,q(Bτ)v_{\omega,\sigma}\in u_{0}+W^{1,q}_{0}(B_{\tau}) as the unique solution to the Dirichlet problem

vω,σminwu0+W01,q(Bτ)BτHω,σ,i(Dw)𝑑x,v_{\omega,\sigma}\mapsto\min_{w\in u_{0}+W^{1,q}_{0}(B_{\tau})}\int_{B_{\tau}}H_{\omega,\sigma,\textnormal{i}}(Dw)\,dx, (4.1)

where, according to the notation in (3.2)2, it is Hω,σ,i(z)Hω,σ,i(z;Bτ)H_{\omega,\sigma,\textnormal{i}}(z)\equiv H_{\omega,\sigma,\textnormal{i}}(z;B_{\tau}). In particular, we have

BτHω,σ,i(Dvω,σ)𝑑xBτHω,σ,i(Du0)𝑑x.\int_{B_{\tau}}H_{\omega,\sigma,\textnormal{i}}(Dv_{\omega,\sigma})\,dx\leq\int_{B_{\tau}}H_{\omega,\sigma,\textnormal{i}}(Du_{0})\,dx\,. (4.2)

Existence, and uniqueness, follow by Direct Methods of the Calculus of Variations and the Euler-Lagrange equation

BτzHω,σ,i(Dvω,σ),Dφ𝑑x=0\int_{B_{\tau}}\langle\partial_{z}H_{\omega,\sigma,\textnormal{i}}(Dv_{\omega,\sigma}),D\varphi\rangle\,dx=0 (4.3)

holds for all φW01,q(Bτ)\varphi\in W^{1,q}_{0}(B_{\tau}) by (3.10). Again (3.10) and standard regularity theory (giu, , Chapter 8), manth1 ; manth2 imply

{vω,σWloc1,(Bτ)Wloc2,2(Bτ)zHω,σ,i(Dvω,σ)Wloc1,2(Bτ;n).\begin{cases}v_{\omega,\sigma}\in W^{1,\infty}_{{\rm loc}}(B_{\tau})\cap W^{2,2}_{{\rm loc}}(B_{\tau})\\ \partial_{z}H_{\omega,\sigma,\textnormal{i}}(Dv_{\omega,\sigma})\in W^{1,2}_{{\rm loc}}(B_{\tau};\mathbb{R}^{n})\,.\end{cases} (4.4)
Proposition 1

With vω,σu0+W01,q(Bτ)v_{\omega,\sigma}\in u_{0}+W^{1,q}_{0}(B_{\tau}) as in (4.1), for every integer n2n\geq 2 there exists a continuous and non-decreasing function μn:(0,1)(1,3/2)\mu_{n}\colon(0,1)\to(1,3/2) such that, for every δ(0,1)\delta\in(0,1), if (1.11) holds with 1μ<μn(δ)1\leq\mu<\mu_{n}(\delta), then

Dvω,σL(B3τ/4)cω,σ,iδ(Du0L(Bτ))Du0L(Bτ)+c\lVert Dv_{\omega,\sigma}\rVert_{L^{\infty}(B_{3\tau/4})}\leq c\mathbb{H}_{\omega,\sigma,\textnormal{i}}^{\delta}\left(\lVert Du_{0}\rVert_{L^{\infty}(B_{\tau})}\right)\lVert Du_{0}\rVert_{L^{\infty}(B_{\tau})}+c (4.5)

holds too, with cc(data,δ)1c\equiv c(\textnormal{{data}},\delta)\geq 1, where ω,σ,i()\mathbb{H}_{\omega,\sigma,\textnormal{i}}(\cdot) has been defined in (3.2)3. Moreover, whenever MM is a number such that

Dvω,σL(B)+1M,\lVert Dv_{\omega,\sigma}\rVert_{L^{\infty}(B)}+1\leq M\,, (4.6)

where BBτB\Subset B_{\tau} is a ball (not necessarily concentric to BτB_{\tau}), the Caccioppoli type inequality

3B/4|D(Eω,σ,i(|Dvω,σ|)κ)+|2dxcMδ|B|2/nB(Eω,σ,i(|Dvω,σ|)κ)+2dx\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{3B/4}\lvert D(E_{\omega,\sigma,\textnormal{i}}(\lvert Dv_{\omega,\sigma}\rvert)-\kappa)_{+}\rvert^{2}\,dx\leq\frac{cM^{\delta}}{|B|^{2/n}}\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B}(E_{\omega,\sigma,\textnormal{i}}(\lvert Dv_{\omega,\sigma}\rvert)-\kappa)_{+}^{2}\,dx (4.7)

holds whenever κ0\kappa\geq 0, again with cc(data,δ)c\equiv c(\textnormal{{data}},\delta). Here, according to the notation in (3.11)2, it is Eω,σ,i(|Dvω,σ|)Eω,σ,i(|Dvω,σ|;Bτ)E_{\omega,\sigma,\textnormal{i}}(\lvert Dv_{\omega,\sigma}\rvert)\equiv E_{\omega,\sigma,\textnormal{i}}(\lvert Dv_{\omega,\sigma}\rvert;B_{\tau}).

Proof

The final shape of μn>1\mu_{n}>1 will be determined in the course of the proof via successive choices of μ\mu (to be taken closer and closer to one, depending on δ\delta). All the constants in the forthcoming estimates will be independent of ω,σ\omega,\sigma, so we shall omit indicating the dependence on such parameters, simply denoting vvω,σv\equiv v_{\omega,\sigma}, HiHω,σ,iH_{\textnormal{i}}\equiv H_{\omega,\sigma,\textnormal{i}}, iω,σ,i\mathbb{H}_{\textnormal{i}}\equiv\mathbb{H}_{\omega,\sigma,\textnormal{i}}, EiEω,σ,iE_{\textnormal{i}}\equiv E_{\omega,\sigma,\textnormal{i}}, and so on. Properties (4.4) allow to differentiate (4.3), i.e., replacing φ\varphi by DsφD_{s}\varphi for s{1,,n}s\in\{1,\ldots,n\} and integrate by parts. Summing the resulting equations over s{1,,n}s\in\{1,\ldots,n\}, we get

s=1nBτzzHi(Dv)DDsv,Dφ𝑑x=0,\sum_{s=1}^{n}\int_{B_{\tau}}\langle\partial_{zz}H_{\textnormal{i}}(Dv)DD_{s}v,D\varphi\rangle\,dx=0\,, (4.8)

that, again thanks to (4.4), holds for all φW1,2(Bτ)\varphi\in W^{1,2}(B_{\tau}) such that suppφBτ\,{\rm supp}\,\varphi\Subset B_{\tau}. Let BBτB\Subset B_{\tau} be a ball, κ0\kappa\geq 0 be any non-negative number and ηCc1(Bτ)\eta\in C^{1}_{c}(B_{\tau}) be a cut-off function such that 𝟙3B/4η𝟙5B/6\mathds{1}_{3B/4}\leq\eta\leq\mathds{1}_{5B/6} and |Dη||B|1/n\lvert D\eta\rvert\lesssim|B|^{-1/n}. By (4.4) the functions φφs:=η2(Ei(|Dv|)κ)+Dsv\varphi\equiv\varphi_{s}:=\eta^{2}(E_{\textnormal{i}}(\lvert Dv\rvert)-\kappa)_{+}D_{s}v are admissible in (4.8), therefore via (3.7)-(3.8) and Young’s inequality we obtain

Bτλi(|Dv|)(Ei(|Dv|)κ)+|D2v|2η2𝑑x+Bτ|D(Ei(|Dv|)κ)+|2η2𝑑x\displaystyle\int_{B_{\tau}}\lambda_{\textnormal{i}}(\lvert Dv\rvert)(E_{\textnormal{i}}(\lvert Dv\rvert)-\kappa)_{+}\lvert D^{2}v\rvert^{2}\eta^{2}\,dx+\int_{B_{\tau}}\lvert D(E_{\textnormal{i}}(\lvert Dv\rvert)-\kappa)_{+}\rvert^{2}\eta^{2}\,dx
cBτ[Λi(|Dv|)λi(|Dv|)]2(Ei(|Dv|)κ)+2|Dη|2𝑑x\displaystyle\qquad\qquad\qquad\qquad\leq c\int_{B_{\tau}}\left[\frac{\Lambda_{\textnormal{i}}(\lvert Dv\rvert)}{\lambda_{\textnormal{i}}(\lvert Dv\rvert)}\right]^{2}(E_{\textnormal{i}}(\lvert Dv\rvert)-\kappa)_{+}^{2}\lvert D\eta\rvert^{2}\,dx (4.9)

with cc(data)c\equiv c(\textnormal{{data}}). See (BM, , Lemmas 4.5-4.6) for more details and a similar inequality. Using (3.9) and a few elementary manipulations, we again find

B|D[η2(Ei(|Dv|)κ)+]|2𝑑x\displaystyle\int_{B}\lvert D[\eta^{2}(E_{\textnormal{i}}(\lvert Dv\rvert)-\kappa)_{+}]\rvert^{2}\,dx
cB(|Dv|2+1)μ1[g(|Dv|)]2(Ei(|Dv|)κ)+2|Dη|2𝑑x\displaystyle\qquad\leq c\int_{B}(|Dv|^{2}+1)^{\mu-1}[g(\lvert Dv\rvert)]^{2}(E_{\textnormal{i}}(\lvert Dv\rvert)-\kappa)_{+}^{2}\lvert D\eta\rvert^{2}\,dx

and taking MM as in (4.6) we arrive at

B|D[η2(Ei(|Dv|)κ)+]|2dxcM2(μ1)[g(M)]2|B|2/nB(Ei(|Dv|)κ)+2dx,\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B}\lvert D[\eta^{2}(E_{\textnormal{i}}(\lvert Dv\rvert)-\kappa)_{+}]\rvert^{2}\,dx\leq\frac{cM^{2(\mu-1)}[g(M)]^{2}}{|B|^{2/n}}\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B}(E_{\textnormal{i}}(\lvert Dv\rvert)-\kappa)_{+}^{2}\,dx,

again with cc(data)c\equiv c(\textnormal{{data}}). In turn, using (1.12) with ε>0\varepsilon>0 and recalling that M1M\geq 1, we find

B|D[η2(Ei(|Dv|)κ)+]|2dxcM2(μ1+ε)|B|2/nB(Ei(|Dv|)κ)+2dx,\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B}\lvert D[\eta^{2}(E_{\textnormal{i}}(\lvert Dv\rvert)-\kappa)_{+}]\rvert^{2}\,dx\leq\frac{cM^{2(\mu-1+\varepsilon)}}{|B|^{2/n}}\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B}(E_{\textnormal{i}}(\lvert Dv\rvert)-\kappa)_{+}^{2}\,dx, (4.10)

again with cc(data,ε)c\equiv c(\textnormal{{data}},\varepsilon). This gives (4.7) recalling that η1\eta\geq 1 on 3B/43B/4 and provided μμ(δ)1\mu\equiv\mu(\delta)\geq 1 and εε(δ)>0\varepsilon\equiv\varepsilon(\delta)>0 are such that

μ1+εδ/2.\mu-1+\varepsilon\leq\delta/2\,. (4.11)

We pass to the proof of (4.5). With TT being the constant in (3.16), we can assume without loss of generality that

Dvω,σL(B3τ/4)T1\lVert Dv_{\omega,\sigma}\rVert_{L^{\infty}(B_{3\tau/4})}\geq T\geq 1 (4.12)

holds, otherwise (4.5) follows trivially. By (4.10) and Sobolev embedding theorem we obtain

(B/2(Ei(|Dv|)κ)+2χdx)1/χcM2(μ1+ε)B(Ei(|Dv|)κ)+2dx,\left(\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B/2}(E_{\textnormal{i}}(\lvert Dv\rvert)-\kappa)_{+}^{2\chi}\,dx\right)^{1/\chi}\leq cM^{2(\mu-1+\varepsilon)}\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B}(E_{\textnormal{i}}(\lvert Dv\rvert)-\kappa)_{+}^{2}\,dx, (4.13)

where 2χ2χ(n)>22\chi\equiv 2\chi(n)>2 is the exponent coming from the Sobolev embedding exponent, and cc(data,ε)c\equiv c(\textnormal{{data}},\varepsilon). We now want to apply Lemma 5 to wEi(|Dv|)w\equiv E_{\textnormal{i}}(\lvert Dv\rvert) using (4.13) to satisfy (2.14) (here all the terms involving the functions fjf_{j} in (2.14) are not present). We fix parameters 3τ/4τ1<τ25τ/63\tau/4\leq\tau_{1}<\tau_{2}\leq 5\tau/6 and related balls

B3τ/4(xc)Bτ1(xc)Bτ2(xc)B5τ/6(xc)Bτ(xc)Bτ,B_{3\tau/4}(x_{\rm c})\Subset B_{\tau_{1}}(x_{\rm c})\Subset B_{\tau_{2}}(x_{\rm c})\subset B_{5\tau/6}(x_{\rm c})\Subset B_{\tau}(x_{\rm c})\equiv B_{\tau}\,,

this time all concentric to the initial ball BτBτ(xc)B_{\tau}\equiv B_{\tau}(x_{\rm c}). Take x0Bτ1x_{0}\in B_{\tau_{1}} arbitrary and set r0:=(τ2τ1)/8r_{0}:=(\tau_{2}-\tau_{1})/8, so that Br0(x0)Bτ2B_{r_{0}}(x_{0})\subset B_{\tau_{2}}, choose M2DvL(Bτ2)M\equiv 2\lVert Dv\rVert_{L^{\infty}(B_{\tau_{2}})} and take BBϱ(x0)Br0(x0)B\equiv B_{\varrho}(x_{0})\subset B_{r_{0}}(x_{0}) in (4.13). This yields

(Bϱ/2(x0)(Ei(|Dv|)κ)+2χdx)1/χ\displaystyle\left(\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B_{\varrho/2}(x_{0})}(E_{\textnormal{i}}(\lvert Dv\rvert)-\kappa)_{+}^{2\chi}\,dx\right)^{1/\chi}
cM2(μ1+ε)Bϱ(x0)(Ei(|Dv|)κ)+2dx,\displaystyle\quad\quad\leq cM^{2(\mu-1+\varepsilon)}\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B_{\varrho}(x_{0})}(E_{\textnormal{i}}(\lvert Dv\rvert)-\kappa)_{+}^{2}\,dx\,, (4.14)

where cc(data,ε)c\equiv c(\textnormal{{data}},\varepsilon). Note that the above choice of MM fits (4.6) via (4.12). By (4.14) we are able to apply Lemma 5 with κ0=0\kappa_{0}=0, M0:=Mμ1+εM_{0}:=M^{\mu-1+\varepsilon} and M1=M2=M30M_{1}=M_{2}=M_{3}\equiv 0, to get

Ei(|Dv(x0)|)cM(μ1+ε)χχ1(Br0[Ei(|Dv|)]2dx)1/2,E_{\textnormal{i}}(\lvert Dv(x_{0})\rvert)\leq cM^{\frac{(\mu-1+\varepsilon)\chi}{\chi-1}}\left(\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B_{r_{0}}}[E_{\textnormal{i}}(\lvert Dv\rvert)]^{2}\,dx\right)^{1/2}\,, (4.15)

with cc(n,q)c\equiv c(n,q). Being x0x_{0} arbitrary in Bτ1B_{\tau_{1}}, (4.15) implies

Ei(DvL(Bτ1))cDvL(Bτ2)(μ1+ε)χχ1(τ2τ1)n/2(Bτ2[Ei(|Dv|)]2𝑑x)1/2E_{\textnormal{i}}\left(\lVert Dv\rVert_{L^{\infty}(B_{\tau_{1}})}\right)\leq\frac{c\lVert Dv\rVert_{L^{\infty}(B_{\tau_{2}})}^{\frac{(\mu-1+\varepsilon)\chi}{\chi-1}}}{(\tau_{2}-\tau_{1})^{n/2}}\left(\int_{B_{\tau_{2}}}[E_{\textnormal{i}}(\lvert Dv\rvert)]^{2}\,dx\right)^{1/2}

and we have used the actual definition of MM. Recalling (4.12), we continue to estimate as

Ei(DvL(Bτ1))\displaystyle E_{\textnormal{i}}\left(\lVert Dv\rVert_{L^{\infty}(B_{\tau_{1}})}\right) (3.16)c(τ2τ1)n/2[Ei(DvL(Bτ2))](μ1+ε)χ(χ1)(2μ)+12\displaystyle\stackrel{{\scriptstyle\eqref{0.001}}}{{\leq}}\frac{c}{(\tau_{2}-\tau_{1})^{n/2}}\left[E_{\textnormal{i}}\left(\lVert Dv\rVert_{L^{\infty}(B_{\tau_{2}})}\right)\right]^{\frac{(\mu-1+\varepsilon)\chi}{(\chi-1)(2-\mu)}+\frac{1}{2}}
(B5τ/6[Ei(|Dv|)]𝑑x)1/2\displaystyle\hskip 34.1433pt\cdot\left(\int_{B_{5\tau/6}}[E_{\textnormal{i}}(\lvert Dv\rvert)]\,dx\right)^{1/2}
(3.17)c(τ2τ1)n/2[Ei(DvL(Bτ2))](μ+2ε)χ2+μ2(χ1)(2μ)\displaystyle\stackrel{{\scriptstyle\eqref{0.002}}}{{\leq}}\frac{c}{{(\tau_{2}-\tau_{1})^{n/2}}}\left[E_{\textnormal{i}}\left(\lVert Dv\rVert_{L^{\infty}(B_{\tau_{2}})}\right)\right]^{\frac{(\mu+2\varepsilon)\chi-2+\mu}{2(\chi-1)(2-\mu)}}
(B5τ/6[Hi(Dv)+1]𝑑x)1/2\displaystyle\hskip 34.1433pt\cdot\left(\int_{B_{5\tau/6}}\left[H_{\textnormal{i}}(Dv)+1\right]\,dx\right)^{1/2} (4.16)

for cc(data,ε)c\equiv c(\textnormal{{data}},\varepsilon). By choosing ε>0\varepsilon>0 and μ>1\mu>1 such that

γ(μ,ε):=(μ+2ε)χ2+μ2(χ1)(2μ)<1\gamma_{*}(\mu,\varepsilon):=\frac{(\mu+2\varepsilon)\chi-2+\mu}{2(\chi-1)(2-\mu)}<1 (4.17)

holds (note that this function is increasing in both μ\mu and ε\varepsilon and that γ(1,0)=1/2\gamma_{*}(1,0)=1/2), we can apply Young’s inequality in (4.16), thereby obtaining

Ei(DvL(Bτ1))\displaystyle E_{\textnormal{i}}\left(\lVert Dv\rVert_{L^{\infty}(B_{\tau_{1}})}\right) 12Ei(DvL(Bτ2))\displaystyle\leq\frac{1}{2}E_{\textnormal{i}}\left(\lVert Dv\rVert_{L^{\infty}(B_{\tau_{2}})}\right)
+c(τ2τ1)nγ~(B5τ/6[Hi(Dv)+1]𝑑x)γ~,\displaystyle\qquad+\frac{c}{(\tau_{2}-\tau_{1})^{n\tilde{\gamma}}}\left(\int_{B_{5\tau/6}}\left[H_{\textnormal{i}}(Dv)+1\right]\,dx\right)^{\tilde{\gamma}}\,,

where

γ~(μ,ε):=(χ1)(2μ)(43μ2ε)χ2+μ.\tilde{\gamma}(\mu,\varepsilon):=\frac{(\chi-1)(2-\mu)}{(4-3\mu-2\varepsilon)\chi-2+\mu}\,.

Notice that γ~(1,0)=1\tilde{\gamma}(1,0)=1 and that this is an increasing function of its arguments, therefore we find ε>0\varepsilon>0 and μ>1\mu>1 such that

γ~(μ,ε)1+δ/4\tilde{\gamma}(\mu,\varepsilon)\leq 1+\delta/4 (4.18)

is satisfied in addition to (4.17). Lemma 2, applied with

h(𝔱)Ei(DvL(B𝔱)),t3τ/4<𝔱<5τ/6s,h(\mathfrak{t})\equiv E_{\textnormal{i}}\left(\lVert Dv\rVert_{L^{\infty}(B_{\mathfrak{t}})}\right),\quad t\equiv 3\tau/4<\mathfrak{t}<5\tau/6\equiv s,

which is always finite by (4.4)2, now gives

Ei(DvL(B3τ/4))c(Bτ[Hi(Dv)+1]dx)1+δ/4,E_{\textnormal{i}}\left(\lVert Dv\rVert_{L^{\infty}(B_{3\tau/4})}\right)\leq c\left(\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B_{\tau}}\left[H_{\textnormal{i}}(Dv)+1\right]\,dx\right)^{1+\delta/4}\,,

so that, recalling (4.2), we arrive at

Ei(DvL(B3τ/4))c(Bτ[Hi(Du0)+1]dx)1+δ/4.E_{\textnormal{i}}\left(\lVert Dv\rVert_{L^{\infty}(B_{3\tau/4})}\right)\leq c\left(\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B_{\tau}}\left[H_{\textnormal{i}}(Du_{0})+1\right]\,dx\right)^{1+\delta/4}\,. (4.19)

Setting E~0(t):=t+aσ,i(B)tq\tilde{E}_{0}(t):=t+a_{\sigma,\textnormal{i}}(B)t^{q} for t0t\geq 0, we find, also thanks to (3.16), that

E~0(t)c[Ei(t)]12μ+cholds for every t0,\tilde{E}_{0}(t)\leq c[E_{\textnormal{i}}(t)]^{\frac{1}{2-\mu}}+c\quad\mbox{holds for every $t\geq 0$}\,, (4.20)

where cc(ν,q)c\equiv c(\nu,q). Recalling (1.11)1, the definitions in (3.2) imply

Hi(z)Hω,σ,i(z)Hω,σ,i(z;Bτ)Lω,σ,i(|z|;Bτ)Li(|z|),H_{\textnormal{i}}(z)\equiv H_{\omega,\sigma,\textnormal{i}}(z)\equiv H_{\omega,\sigma,\textnormal{i}}(z;B_{\tau})\leq L\mathbb{H}_{\omega,\sigma,\textnormal{i}}(|z|;B_{\tau})\equiv L\mathbb{H}_{\textnormal{i}}(|z|),

so that inequality (4.19) together with (3.3) and (4.20) now gives

E~0(DvL(B3τ/4))\displaystyle\tilde{E}_{0}\left(\lVert Dv\rVert_{L^{\infty}(B_{3\tau/4})}\right) c[i(Du0L(Bτ))+1]1+δ/42μ\displaystyle\leq c\left[\mathbb{H}_{\textnormal{i}}\left(\lVert Du_{0}\rVert_{L^{\infty}(B_{\tau})}\right)+1\right]^{\frac{1+\delta/4}{2-\mu}}
c[i(Du0L(Bτ))+1]1+δ2\displaystyle\leq c\left[\mathbb{H}_{\textnormal{i}}\left(\lVert Du_{0}\rVert_{L^{\infty}(B_{\tau})}\right)+1\right]^{1+\frac{\delta}{2}}

with cc(data,δ)c\equiv c(\textnormal{{data}},\delta), provided we further take μ>1\mu>1 such that

1+δ/42μ1+δ2.\frac{1+\delta/4}{2-\mu}\leq 1+\frac{\delta}{2}\,. (4.21)

Note that in the above inequality we have incorporated the dependence on ε\varepsilon in the dependence on δ\delta as this last quantity influences the choice of ε\varepsilon via (4.17) and (4.18). Using the very definition (3.2)3 we then continue to estimate

E~0(DvL(B3τ/4))\displaystyle\tilde{E}_{0}\left(\lVert Dv\rVert_{L^{\infty}(B_{3\tau/4})}\right)
ciδ/2(Du0L(Bτ))Du0L(Bτ)g(Du0L(Bτ))\displaystyle\quad\leq c\mathbb{H}_{\textnormal{i}}^{\delta/2}\left(\lVert Du_{0}\rVert_{L^{\infty}(B_{\tau})}\right)\lVert Du_{0}\rVert_{L^{\infty}(B_{\tau})}g\left(\lVert Du_{0}\rVert_{L^{\infty}(B_{\tau})}\right)
+ciδ/2(Du0L(Bτ))aσ,i(Bτ)[ω(Du0L(Bτ))]q+c\displaystyle\qquad+c\mathbb{H}_{\textnormal{i}}^{\delta/2}\left(\lVert Du_{0}\rVert_{L^{\infty}(B_{\tau})}\right)a_{\sigma,\textnormal{i}}(B_{\tau})\left[\ell_{\omega}\left(\lVert Du_{0}\rVert_{L^{\infty}(B_{\tau})}\right)\right]^{q}+c

and, by again using (1.12) with ε=δ/2\varepsilon=\delta/2, we conclude with

E~0(DvL(B3τ/4))ciδ(Du0L(Bτ))E~0(Du0L(Bτ))+c\tilde{E}_{0}\left(\lVert Dv\rVert_{L^{\infty}(B_{3\tau/4})}\right)\leq c\mathbb{H}_{\textnormal{i}}^{\delta}\left(\lVert Du_{0}\rVert_{L^{\infty}(B_{\tau})}\right)\tilde{E}_{0}\left(\lVert Du_{0}\rVert_{L^{\infty}(B_{\tau})}\right)+c (4.22)

with cc(data,δ)c\equiv c(\textnormal{{data}},\delta). Since E~0()\tilde{E}_{0}(\cdot) is monotone increasing, convex and such that E~0(0)=0\tilde{E}_{0}(0)=0, we deduce that its inverse E~01()\tilde{E}_{0}^{-1}(\cdot) is increasing, concave and E~01(0)=0\tilde{E}_{0}^{-1}(0)=0, therefore it is subadditive and, for any given constant c0c_{*}\geq 0 it is E~01(ct)(c+1)E~01(t)\tilde{E}_{0}^{-1}(c_{*}t)\leq(c_{*}+1)\tilde{E}_{0}^{-1}(t), see for instance (DM, , Remark 10 in Section 4). Using this last property, we can apply E~01()\tilde{E}_{0}^{-1}(\cdot) to both sides of (4.22) in order to conclude with (4.5). Finally, the choice of the function μn(δ)\mu_{n}(\delta) mentioned in the statement, that we can always take such that μn(δ)<3/2\mu_{n}(\delta)<3/2, comes by the choices made in (4.11), (4.17), (4.18) and (4.21) and a standard continuity argument (for this recall that both γ()\gamma_{*}(\cdot) and γ~()\tilde{\gamma}(\cdot) are increasing functions of their arguments and that χ\chi depends on nn). Notice that by the above construction it follows that μn(0+)=1\mu_{n}(0_{+})=1.

5 Proof of Theorem 1.2

The proof will take eleven different steps, distributed along Sections 5.1-5.11 below. We shall concentrate on the singular case s=0\textnormal{{s}}=0, then giving remarks on how to deal with the (actually simpler) case s>0\textnormal{{s}}>0 at the very end of Section 5.9. We shall start assuming that (1.11)-(1.12) for some μ[1,3/2)\mu\in[1,3/2). We shall make further restrictions on the size of μ\mu in the course of the proof until we finally come to determine the value μmax\mu_{\textnormal{max}} mentioned in the statement of Theorem 1.2.

5.1 Absence of Lavrentiev phenomenon

We have the following approximation-in-energy result that actually implies the absence of Lavrentiev phenomenon:

Lemma 7

With H()H(\cdot) defined in (3.1), assume (1.7)1 and that F()F(\cdot) satisfies (1.11)1 where g()g(\cdot) is as in Theorem 1.2 (described after (1.11)). Let wWloc1,1(Ω)w\in W^{1,1}_{{\rm loc}}(\Omega) be any function such that H(,Dw)Lloc1(Ω)H(\cdot,Dw)\in L^{1}_{{\rm loc}}(\Omega). For every ball BΩB\Subset\Omega there exists a sequence {wε}W1,(B)\{w_{\varepsilon}\}\subset W^{1,\infty}(B) such that wεww_{\varepsilon}\to w in W1,1(B)W^{1,1}(B) and H(,Dwε)H(,Dw)H(\cdot,Dw_{\varepsilon})\to H(\cdot,Dw) in L1(B)L^{1}(B).

The proof follows (dm, , Section 5) almost verbatim (see also (sharp, , Lemma 13)). Notice that this is essentially the only point where the assumed convexity of ttg(t)t\mapsto tg(t) is used (this implies that z|z|g(|z|)z\mapsto|z|g(|z|) is convex as g()g(\cdot) is also non-decreasing). For more related results on the absence of Lavrentiev phenomenon we refer to the recent papers AFM ; balci ; balci2 ; buli ; koch1 ; koch2 and related references.

5.2 Auxiliary Dirichlet problems and convergence

In the following uWloc1,1(Ω)u\in W^{1,1}_{{\rm loc}}(\Omega) denotes a local minimizer of 𝒩()\mathcal{N}(\cdot), as in Theorem 1.2. By ω,ε{ω},{ε}{ωk}k,{εk}k,\omega,\varepsilon\equiv\{\omega\},\{\varepsilon\}\equiv\{\omega_{k}\}_{k},\{\varepsilon_{k}\}_{k}, we denote two decreasing sequences of positive numbers such that ω,ε0\omega,\varepsilon\to 0, and ε,ω1\varepsilon,\omega\leq 1; we will several times extract subsequences and these will still be denoted by ω,ε\omega,\varepsilon (for this reason we drop the pendice kk). We denote by o(ε)\texttt{o}(\varepsilon) a quantity such that o(ε)0\texttt{o}(\varepsilon)\to 0 as ε0\varepsilon\to 0. Similarly, we denote by oε(ω)\texttt{o}_{\varepsilon}(\omega) a quantity, depending both on ε\varepsilon and ω\omega, such that oε(ω)0\texttt{o}_{\varepsilon}(\omega)\to 0 as ω0\omega\to 0 for each fixed ε\varepsilon. The exact value of such quantities might change on different occurences and only the aforementioned asymptotic properties will matter. Let BrΩB_{r}\Subset\Omega be a ball with 0<r10<r\leq 1; by Lemma 7, there exists a sequence {u~ε}W1,(Br)\{\tilde{u}_{\varepsilon}\}\in W^{1,\infty}(B_{r}) so that

u~εuinW1,1(Br)and𝒩(u~ε,Br)=𝒩(u,Br)+o(ε).\tilde{u}_{\varepsilon}\to u\ \ \mbox{in}\ \ W^{1,1}(B_{r})\quad\ \ \mbox{and}\quad\ \ \mathcal{N}(\tilde{u}_{\varepsilon},B_{r})=\mathcal{N}(u,B_{r})+\texttt{o}(\varepsilon)\,. (5.1)

We define the sequence

σε:=(1+ε1+Du~εLq(Br)2q)1σεBr[ω(Du~ε)]q𝑑x0\sigma_{\varepsilon}:=\left(1+\varepsilon^{-1}+\lVert D\tilde{u}_{\varepsilon}\rVert_{L^{q}(B_{r})}^{2q}\right)^{-1}\ \Longrightarrow\ \sigma_{\varepsilon}\int_{B_{r}}[\ell_{\omega}(D\tilde{u}_{\varepsilon})]^{q}\,dx\to 0 (5.2)

(uniformly with respect to ω(0,1]\omega\in(0,1]). Then we consider uω,εu~ε+W01,q(Br)u_{\omega,\varepsilon}\in\tilde{u}_{\varepsilon}+W^{1,q}_{0}(B_{r}) as the unique solution to the Dirichlet problem

uω,εminwu~ε+W01,q(Br)𝒩ω,ε(w,Br)u_{\omega,\varepsilon}\mapsto\min_{w\in\tilde{u}_{\varepsilon}+W^{1,q}_{0}(B_{r})}\mathcal{N}_{\omega,\varepsilon}(w,B_{r}) (5.3)

where

𝒩ω,ε(w,Br):=BrHω,σε(x,Dw)𝑑x.\mathcal{N}_{\omega,\varepsilon}(w,B_{r}):=\int_{B_{r}}H_{\omega,\sigma_{\varepsilon}}(x,Dw)\,dx\,. (5.4)

Recall that the integrand Hω,σε()H_{\omega,\sigma_{\varepsilon}}(\cdot) has been defined in (3.2), with σσε\sigma\equiv\sigma_{\varepsilon}. The solvability of (5.3) follows by Direct Methods and standard convexity arguments. By (3.6), we can apply the by now classical regularity theory contained in (giu, , Chapter 8) and manth1 ; manth2 , therefore

uω,εCloc1,β(Br)for someββ(n,ν,L,q,ω,ε)(0,1).u_{\omega,\varepsilon}\in C^{1,\beta}_{{\rm loc}}(B_{r})\quad\mbox{for some}\ \ \beta\equiv\beta(n,\nu,L,q,\omega,\varepsilon)\in(0,1)\,. (5.5)

Mean value theorem and (5.2) imply

|𝒩ω,ε(u~ε,Br)𝒩(u~ε,Br)|\displaystyle|\mathcal{N}_{\omega,\varepsilon}(\tilde{u}_{\varepsilon},B_{r})-\mathcal{N}(\tilde{u}_{\varepsilon},B_{r})| cωBr(|Du~ε|q1+1)𝑑x\displaystyle\leq c\omega\int_{B_{r}}(|D\tilde{u}_{\varepsilon}|^{q-1}+1)\,dx
+σεBr[ω(Du~ε)]q𝑑x\displaystyle\qquad+\sigma_{\varepsilon}\int_{B_{r}}[\ell_{\omega}(D\tilde{u}_{\varepsilon})]^{q}\,dx
=oε(ω)+o(ε).\displaystyle=\texttt{o}_{\varepsilon}(\omega)+\texttt{o}(\varepsilon)\,. (5.6)

Similarly, noting that |z|q1|z|+1cF(x,z)+c|z|^{q-1}\leq|z|+1\leq cF(x,z)+c by (1.11)1 and q<3/2q<3/2 (follows from (1.7)1), we find, using also mean value theorem

|𝒩ω,ε(uω,ε,Br)𝒩(uω,ε,Br)σεBr[ω(Duω,ε)]q𝑑x|\displaystyle\left|\mathcal{N}_{\omega,\varepsilon}(u_{\omega,\varepsilon},B_{r})-\mathcal{N}(u_{\omega,\varepsilon},B_{r})-\sigma_{\varepsilon}\int_{B_{r}}[\ell_{\omega}(Du_{\omega,\varepsilon})]^{q}\,dx\right|
cωBr(|Duω,ε|q1+1)𝑑xcω𝒩ω,ε(uω,ε,Br)+cω.\displaystyle\quad\quad\leq c\omega\int_{B_{r}}(|Du_{\omega,\varepsilon}|^{q-1}+1)\,dx\leq c\omega\mathcal{N}_{\omega,\varepsilon}(u_{\omega,\varepsilon},B_{r})+c\omega\,. (5.7)

In turn, using in order: the minimality of uω,εu_{\omega,\varepsilon}, (5.1) and (5.6), we have

𝒩ω,ε(uω,ε,Br)\displaystyle\mathcal{N}_{\omega,\varepsilon}(u_{\omega,\varepsilon},B_{r}) 𝒩ω,ε(u~ε,Br)\displaystyle\leq\mathcal{N}_{\omega,\varepsilon}(\tilde{u}_{\varepsilon},B_{r})
𝒩(u~ε,Br)+|𝒩ω,ε(u~ε,Br)𝒩(u~ε,Br)|\displaystyle\leq\mathcal{N}(\tilde{u}_{\varepsilon},B_{r})+\lvert\mathcal{N}_{\omega,\varepsilon}(\tilde{u}_{\varepsilon},B_{r})-\mathcal{N}(\tilde{u}_{\varepsilon},B_{r})\rvert
𝒩(u,Br)+oε(ω)+o(ε)\displaystyle\leq\mathcal{N}(u,B_{r})+\texttt{o}_{\varepsilon}(\omega)+\texttt{o}(\varepsilon) (5.8)

so that, using the content of the last two displays we gain

𝒩(uω,ε,Br)+σεBr[ω(Duω,ε)]q𝑑x\displaystyle\mathcal{N}(u_{\omega,\varepsilon},B_{r})+\sigma_{\varepsilon}\int_{B_{r}}[\ell_{\omega}(Du_{\omega,\varepsilon})]^{q}\,dx
=𝒩ω,ε(uω,ε,Br)+oε(ω)+o(ε)+cω.\displaystyle\qquad\ \ =\mathcal{N}_{\omega,\varepsilon}(u_{\omega,\varepsilon},B_{r})+\texttt{o}_{\varepsilon}(\omega)+\texttt{o}(\varepsilon)+c\omega\,. (5.9)

Estimate (5.8) and (3.5)1 imply that for every ε(0,1)\varepsilon\in(0,1) the sequence {uω,ε}ω\{u_{\omega,\varepsilon}\}_{\omega} is uniformly bounded in W1,q(Br)W^{1,q}(B_{r}), therefore, up to not relabelled subsequences, we have

uω,εuεweakly inW1,q(Br)anduεu~εW01,q(Br)u_{\omega,\varepsilon}\rightharpoonup u_{\varepsilon}\ \ \mbox{weakly in}\ \ W^{1,q}(B_{r})\quad\mbox{and}\quad u_{\varepsilon}-\tilde{u}_{\varepsilon}\in W^{1,q}_{0}(B_{r}) (5.10)

as ω0\omega\to 0. Letting ω0\omega\to 0 in (5.9) and using standard weak lower semicontinuity theorems, yields

𝒩(uε,Br)lim infω0𝒩(uω,ε,Br)lim infω0𝒩ω,ε(uω,ε,Br)+o(ε)\mathcal{N}(u_{\varepsilon},B_{r})\leq\liminf_{\omega\to 0}\mathcal{N}(u_{\omega,\varepsilon},B_{r})\leq\liminf_{\omega\to 0}\mathcal{N}_{\omega,\varepsilon}(u_{\omega,\varepsilon},B_{r})+\texttt{o}(\varepsilon)

for every fixed ε(0,1)\varepsilon\in(0,1). Using (5.8) we conclude with

𝒩(uε,Br)𝒩(u,Br)+o(ε)\mathcal{N}(u_{\varepsilon},B_{r})\leq\mathcal{N}(u,B_{r})+\texttt{o}(\varepsilon) (5.11)

and again this holds for every ε(0,1)\varepsilon\in(0,1). By (1.11)1 and (5.11) the sequence {|Duε|g(Duε)}\{\lvert Du_{\varepsilon}\rvert g(Du_{\varepsilon})\} is uniformly bounded in L1(Br)L^{1}(B_{r}). Recalling that the assumptions on g()g(\cdot) imply that g(t)g(t)\to\infty as tt\to\infty, by classical results of Dunford & Pettis and de la Vallée Poussin, there exists u^W1,1(Br)\hat{u}\in W^{1,1}(B_{r}) such that uεu^u_{\varepsilon}\rightharpoonup\hat{u} weakly in W1,1(Br)W^{1,1}(B_{r}) and u^uW01,1(Br)\hat{u}-u\in W^{1,1}_{0}(B_{r}). Letting ε0\varepsilon\to 0 in (5.11) weak lower semicontinuity (see (giu, , Theorem 4.3)) and (5.2) yield 𝒩(u^,Br)𝒩(u,Br)\mathcal{N}(\hat{u},B_{r})\leq\mathcal{N}(u,B_{r}), while the opposite inequality follows by the minimality of uu. We conclude with 𝒩(u^,Br)=𝒩(u,Br)\mathcal{N}(\hat{u},B_{r})=\mathcal{N}(u,B_{r}), so that, by strict convexity of the functional w𝒩(w,Br)w\mapsto\mathcal{N}(w,B_{r}) we find that uu^u\equiv\hat{u} in BrB_{r} and we deduce that

uεuweakly inW1,1(Br).u_{\varepsilon}\rightharpoonup u\ \ \mbox{weakly in}\ \ W^{1,1}(B_{r})\,. (5.12)

5.3 Blow-up

We fix ω,ε(0,1]\omega,\varepsilon\in(0,1] and uω,εW1,q(Br)u_{\omega,\varepsilon}\in W^{1,q}(B_{r}) as in (5.3). With Bϱ(x0)BrB_{\varrho}(x_{0})\Subset B_{r} being a ball not necessarily concentric to BrB_{r}, we take 𝔐\mathfrak{M} such that

𝔐E~ω,σε(,|Duω,ε|)L(Bϱ(x0))+1,\mathfrak{M}\geq\lVert\tilde{E}_{\omega,\sigma_{\varepsilon}}(\cdot,\lvert Du_{\omega,\varepsilon}\rvert)\rVert_{L^{\infty}(B_{\varrho}(x_{0}))}+1\,, (5.13)

where E~ω,σε()\tilde{E}_{\omega,\sigma_{\varepsilon}}(\cdot) is defined in (3.13). The above quantities are finite by (5.5). We rescale uu and H(x,z)=F(x,z)+a(x)|z|qH(x,z)=F(x,z)+a(x)|z|^{q} on Bϱ(x0)B_{\varrho}(x_{0}) defining

{uω,ε,ϱ(x):=uω,ε(x0+ϱx)/ϱFϱ(x,z):=F(x0+ϱx,z),𝒶ϱ(x):=a(x0+ϱx)ϱ(x,z):=H(x0+ϱx,z)=Fϱ(x,z)+𝒶ϱ(x)|z|q,\begin{cases}\ \displaystyle u_{\omega,\varepsilon,\varrho}(x):=u_{\omega,\varepsilon}(x_{0}+\varrho x)/\varrho\\ \ F_{\varrho}(x,z):=F(x_{0}+\varrho x,z),\quad\mathcal{a}_{\varrho}(x):=a(x_{0}+\varrho x)\\ \ \mathcal{H}_{\varrho}(x,z):=H(x_{0}+\varrho x,z)=F_{\varrho}(x,z)+\mathcal{a}_{\varrho}(x)|z|^{q}\,,\end{cases} (5.14)

with (x,z)1×n(x,z)\in\mathcal{B}_{1}\times\mathbb{R}^{n}. Note that, obviously, ϱ()\mathcal{H}_{\varrho}(\cdot) is still an integrand of the type in (3.1) (see also (5.15) below) and therefore the content of Section 3 applies to ϱ()\mathcal{H}_{\varrho}(\cdot) as well. Since uω,εu_{\omega,\varepsilon} solves (5.3), recalling the notation fixed in (3.2), it follows that uω,ε,ϱW1,q(1)u_{\omega,\varepsilon,\varrho}\in W^{1,q}(\mathcal{B}_{1}) is a local minimizer on 1\mathcal{B}_{1} of the functional

W1,q(1)w1(ϱ)ω,σε(x,Dw)𝑑xW^{1,q}(\mathcal{B}_{1})\ni w\mapsto\int_{\mathcal{B}_{1}}(\mathcal{H}_{\varrho})_{\omega,\sigma_{\varepsilon}}(x,Dw)\,dx

where, recalling the notation in (2.7)1, with (x,z)1×n(x,z)\in\mathcal{B}_{1}\times\mathbb{R}^{n} it is

{(ϱ)ω,σε(x,z)=Fϱ(x,z)+(𝒶ϱ)σε(x)[ω(z)]q(𝒶ϱ)σε(x)=𝒶ϱ(x)+σε=a(x0+ϱx)+σε.\begin{cases}(\mathcal{H}_{\varrho})_{\omega,\sigma_{\varepsilon}}(x,z)=F_{\varrho}(x,z)+(\mathcal{a}_{\varrho})_{\sigma_{\varepsilon}}(x)[\ell_{\omega}(z)]^{q}\\ (\mathcal{a}_{\varrho})_{\sigma_{\varepsilon}}(x)=\mathcal{a}_{\varrho}(x)+\sigma_{\varepsilon}=a(x_{0}+\varrho x)+\sigma_{\varepsilon}\,.\end{cases} (5.15)

From now on, keeping fixed the choice of ω,ε\omega,\varepsilon made at the beginning, in order to simplify the notation we shall omit to specify dependence on such parameters, simply abbreviating

uϱ(x)uω,ε,ϱ(x),Hϱ(x,z)(ϱ)ω,σε(x,z)u_{\varrho}(x)\equiv u_{\omega,\varepsilon,\varrho}(x),\quad H_{\varrho}(x,z)\equiv(\mathcal{H}_{\varrho})_{\omega,\sigma_{\varepsilon}}(x,z) (5.16)

for (x,z)1×n(x,z)\in\mathcal{B}_{1}\times\mathbb{R}^{n}. The minimality of uϱu_{\varrho} implies the validity of the Euler-Lagrange equation

1zHϱ(x,Duϱ),Dφ𝑑x=0for allφW01,q(1).\int_{\mathcal{B}_{1}}\langle\partial_{z}H_{\varrho}(x,Du_{\varrho}),D\varphi\rangle\,dx=0\quad\mbox{for all}\ \ \varphi\in W^{1,q}_{0}(\mathcal{B}_{1})\,. (5.17)

By (3.5) the integrand Hϱ()H_{\varrho}(\cdot) satisfies

{λϱ(x,|z|)|ξ|2czzHϱ(x,z)ξ,ξ|zzHϱ(x,z)|cΛϱ(x,|z|)|zHϱ(x,z)zHϱ(y,z)|cϱα0|xy|α0g(|z|)+cϱα|xy|α[ω(z)]q1\begin{cases}\ \lambda_{\varrho}(x,\lvert z\rvert)\lvert\xi\rvert^{2}\leq c\langle\partial_{zz}H_{\varrho}(x,z)\xi,\xi\rangle\\ \ \lvert\partial_{zz}H_{\varrho}(x,z)\rvert\leq c\Lambda_{\varrho}(x,\lvert z\rvert)\\ \ \lvert\partial_{z}H_{\varrho}(x,z)-\partial_{z}H_{\varrho}(y,z)\rvert\\ \qquad\leq c\varrho^{\alpha_{0}}\lvert x-y\rvert^{\alpha_{0}}g(|z|)+c\varrho^{\alpha}\lvert x-y\rvert^{\alpha}[\ell_{\omega}(z)]^{q-1}\end{cases} (5.18)

for any x,y1x,y\in\mathcal{B}_{1} and all z,ξnz,\xi\in\mathbb{R}^{n}, where cc(data)c\equiv c(\textnormal{{data}}) and, according to the definitions in (3.4) and the notation in (5.16), we are denoting

λϱ(x,|z|):=λω,σε(x0+ϱx,|z|),Λϱ(x,|z|):=Λω,σε(x0+ϱx,|z|).\lambda_{\varrho}(x,\lvert z\rvert):=\lambda_{\omega,\sigma_{\varepsilon}}(x_{0}+\varrho x,\lvert z\rvert)\,,\quad\Lambda_{\varrho}(x,\lvert z\rvert):=\Lambda_{\omega,\sigma_{\varepsilon}}(x_{0}+\varrho x,\lvert z\rvert)\,. (5.19)

5.4 Minimal integrands

Here we are going to play with auxiliary functionals whose integrands are of the type in (3.2)2, and therefore featuring no explicit dependence on xx (these are usually called “frozen” integrands). The results of Section 4 can be therefore applied. Let us fix a number β0(0,1)\beta_{0}\in(0,1), to be determined in a few lines, and a vector hn{0}h\in\mathbb{R}^{n}\setminus\{0\} such that

0<|h|128/β0.0<\lvert h\rvert\leq\frac{1}{2^{8/\beta_{0}}}\,. (5.20)

We take xc1/2+2|h|β0x_{\rm c}\in\mathcal{B}_{1/2+2|h|^{\beta_{0}}} and fix a ball centered at xcx_{\rm c} with radius |h|β0\lvert h\rvert^{\beta_{0}}, denoted by BhB|h|β0(xc)B_{h}\equiv B_{\lvert h\rvert^{\beta_{0}}}(x_{\rm c}). By (5.20) we have 8Bh18B_{h}\Subset\mathcal{B}_{1}. We set

𝔪𝔪(8Bh):=DuϱL(8Bh)+1.\mathfrak{m}\equiv\mathfrak{m}(8B_{h}):=\lVert Du_{\varrho}\rVert_{L^{\infty}(8B_{h})}+1\,. (5.21)

According to the notation established in (2.7), (3.2) and (5.15)-(5.16), we define

{Hϱ,i(z)(ϱ)ω,σε,i(z;8Bh)Fϱ(xc,z)+a~ϱ,i(8Bh)[ω(z)]qa~ϱ,i(8Bh):=(𝒶ϱ)σε,i(8Bh)=infx8Bh(𝒶ϱ)σε(x)=infx8Bha(x0+ϱx)+σε.\begin{cases}H_{\varrho,\textnormal{i}}(z)\equiv(\mathcal{H}_{\varrho})_{\omega,\sigma_{\varepsilon},\textnormal{i}}(z;8B_{h})\equiv F_{\varrho}(x_{\rm c},z)+\tilde{a}_{\varrho,\textnormal{i}}(8B_{h})[\ell_{\omega}(z)]^{q}\\ \displaystyle\tilde{a}_{\varrho,\textnormal{i}}(8B_{h}):=(\mathcal{a}_{\varrho})_{\sigma_{\varepsilon},\textnormal{i}}(8B_{h})=\inf_{x\in 8B_{h}}(\mathcal{a}_{\varrho})_{\sigma_{\varepsilon}}(x)=\inf_{x\in 8B_{h}}a(x_{0}+\varrho x)+\sigma_{\varepsilon}\,.\end{cases} (5.22)

Note that zHϱ,iC1(n;n)\partial_{z}H_{\varrho,\textnormal{i}}\in C^{1}(\mathbb{R}^{n};\mathbb{R}^{n}). As in (3.7), with

{λϱ,i(|z|)λϱ,i(|z|;8Bh):=(|z|2+1)μ/2+(q1)a~ϱ,i(8Bh)[ω(z)]q2Λϱ,i(|z|)Λϱ,i(|z|;8Bh):=(|z|2+1)1/2g(|z|)+a~ϱ,i(8Bh)[ω(z)]q2,\begin{cases}\lambda_{\varrho,\textnormal{i}}(\lvert z\rvert)\displaystyle\equiv\lambda_{\varrho,\textnormal{i}}(\lvert z\rvert;8B_{h}):=(|z|^{2}+1)^{-\mu/2}+(q-1)\tilde{a}_{\varrho,\textnormal{i}}(8B_{h})[\ell_{\omega}(z)]^{q-2}\\ \Lambda_{\varrho,\textnormal{i}}(\lvert z\rvert)\equiv\Lambda_{\varrho,\textnormal{i}}(\lvert z\rvert;8B_{h}):=(|z|^{2}+1)^{-1/2}g(|z|)+\tilde{a}_{\varrho,\textnormal{i}}(8B_{h})[\ell_{\omega}(z)]^{q-2}\,,\end{cases}

from (1.12) and (3.8) it follows that

{σε[ω(z)]qHϱ,i(z)c([ω(z)]q+1)|zHϱ,i(z)|c([ω(z)]q1+1)λϱ,i(|z|)|ξ|2czzHϱ,i(z)ξ,ξ,|zzHϱ,i(z)|cΛϱ,i(|z|)𝒱ϱ,i2(z1,z2;8Bh)czHϱ,i(z1)zHϱ,i(z2),z1z2\begin{cases}\ \sigma_{\varepsilon}[\ell_{\omega}(z)]^{q}\leq H_{\varrho,\textnormal{i}}(z)\leq c\left([\ell_{\omega}(z)]^{q}+1\right)\\ \ \lvert\partial_{z}H_{\varrho,\textnormal{i}}(z)\rvert\leq c\left([\ell_{\omega}(z)]^{q-1}+1\right)\\ \ \lambda_{\varrho,\textnormal{i}}(\lvert z\rvert)\lvert\xi\rvert^{2}\leq c\langle\partial_{zz}H_{\varrho,\textnormal{i}}(z)\xi,\xi\rangle,\quad\lvert\partial_{zz}H_{\varrho,\textnormal{i}}(z)\rvert\leq c\Lambda_{\varrho,\textnormal{i}}(\lvert z\rvert)\\ \ \mathcal{V}_{\varrho,\textnormal{i}}^{2}(z_{1},z_{2};8B_{h})\leq c\langle\partial_{z}H_{\varrho,\textnormal{i}}(z_{1})-\partial_{z}H_{\varrho,\textnormal{i}}(z_{2}),z_{1}-z_{2}\rangle\end{cases} (5.23)

hold for all z,z1,z2,ξnz,z_{1},z_{2},\xi\in\mathbb{R}^{n}, with cc(data)c\equiv c(\textnormal{{data}}). Consistently with (2.8), here we are denoting

𝒱ϱ,i2(z1,z2;8Bh)\displaystyle\mathcal{V}_{\varrho,\textnormal{i}}^{2}(z_{1},z_{2};8B_{h}) :=|V1,2μ(z1)V1,2μ(z2)|2\displaystyle:=\lvert V_{1,2-\mu}(z_{1})-V_{1,2-\mu}(z_{2})\rvert^{2}
+a~ϱ,i(8Bh)|Vω,q(z1)Vω,q(z2)|2.\displaystyle\qquad+\tilde{a}_{\varrho,\textnormal{i}}(8B_{h})\lvert V_{\omega,q}(z_{1})-V_{\omega,q}(z_{2})\rvert^{2}\,. (5.24)

Recalling (3.11) and (5.19), we define, for t0t\geq 0 and x1x\in\mathcal{B}_{1}

{Eϱ(x,t):=0tλϱ(x,s)sdsEϱ,i(t)Eϱ,i(t;8Bh):=0tλϱ,i(s;8Bh)sds\begin{cases}\displaystyle E_{\varrho}(x,t):=\int_{0}^{t}\lambda_{\varrho}(x,s)s\,{\rm d}s\\ \displaystyle E_{\varrho,\textnormal{i}}(t)\equiv E_{\varrho,\textnormal{i}}(t;8B_{h}):=\int_{0}^{t}\lambda_{\varrho,\textnormal{i}}(s;8B_{h})s\,{\rm d}s\end{cases} (5.25)

with related explicit expressions as in (3.12) and

{E~ϱ(x,t):=12μ[1(t)]2μ+(11/q)(𝒶ϱ)σε(x)[ω(t)]qE~ϱ,i(t):=12μ[1(t)]2μ+(11/q)a~ϱ,i(8Bh)[ω(t)]q.\begin{cases}\ \tilde{E}_{\varrho}(x,t):=\frac{1}{2-\mu}[\ell_{1}(t)]^{2-\mu}+(1-1/q)(\mathcal{a}_{\varrho})_{\sigma_{\varepsilon}}(x)[\ell_{\omega}(t)]^{q}\\ \ \tilde{E}_{\varrho,\textnormal{i}}(t):=\frac{1}{2-\mu}[\ell_{1}(t)]^{2-\mu}+(1-1/q)\tilde{a}_{\varrho,\textnormal{i}}(8B_{h})[\ell_{\omega}(t)]^{q}\,.\end{cases}

By (3.15) we have

|Eϱ(x,t)Eϱ,i(t)|c|h|αβ0ϱα[ω(t)]qfor every x8Bh and t0\lvert E_{\varrho}(x,t)-E_{\varrho,\textnormal{i}}(t)\rvert\leq c|h|^{\alpha\beta_{0}}\varrho^{\alpha}[\ell_{\omega}(t)]^{q}\quad\mbox{for every $x\in 8B_{h}$ and $t\geq 0$} (5.26)

and (5.13) implies

𝔐E~ϱ(,|Duϱ|)L(1)+1.\mathfrak{M}\geq\lVert\tilde{E}_{\varrho}(\cdot,\lvert Du_{\varrho}\rvert)\rVert_{L^{\infty}(\mathcal{B}_{1})}+1\,. (5.27)

Looking at (5.23), Direct Methods and strict convexity provide us with a unique minimizer vuϱ+W01,q(8Bh)v\in u_{\varrho}+W^{1,q}_{0}(8B_{h}) defined by

vv(8Bh)minwuϱ+W01,q(8Bh)8BhHϱ,i(Dw)𝑑xv\equiv v(8B_{h})\mapsto\min_{w\in u_{\varrho}+W^{1,q}_{0}(8B_{h})}\int_{8B_{h}}H_{\varrho,\textnormal{i}}(Dw)\,dx (5.28)

so that, (5.23)1 obviously implies

8BhzHϱ,i(Dv),Dφ𝑑x=0for allφW01,q(8Bh)\int_{8B_{h}}\langle\partial_{z}H_{\varrho,\textnormal{i}}(Dv),D\varphi\rangle\,dx=0\qquad\mbox{for all}\ \ \varphi\in W^{1,q}_{0}(8B_{h}) (5.29)

and minimality gives

8BhHϱ,i(Dv)𝑑x8BhHϱ,i(Duϱ)𝑑x.\int_{8B_{h}}H_{\varrho,\textnormal{i}}(Dv)\,dx\leq\int_{8B_{h}}H_{\varrho,\textnormal{i}}(Du_{\varrho})\,dx\,. (5.30)
Lemma 8

Let δ1(0,1)\delta_{1}\in(0,1) and let μn()\mu_{n}(\cdot) be the (non-decreasing) function introduced in Proposition 1. If 1μ<μn(δ1/2)1\leq\mu<\mu_{n}(\delta_{1}/2), then the inequalities

DvL(6Bh)c𝔪1+δ1\lVert Dv\rVert_{L^{\infty}(6B_{h})}\leq c\mathfrak{m}^{1+\delta_{1}} (5.31)

and

2Bh|D(Eϱ,i(|Dv|)κ)+|2𝑑xc𝔪δ1|h|2β04Bh(Eϱ,i(|Dv|)κ)+2𝑑x\int_{2B_{h}}\lvert D(E_{\varrho,\textnormal{i}}(\lvert Dv\rvert)-\kappa)_{+}\rvert^{2}\,dx\leq\frac{c\,\mathfrak{m}^{\delta_{1}}}{\lvert h\rvert^{2\beta_{0}}}\int_{4B_{h}}(E_{\varrho,\textnormal{i}}(\lvert Dv\rvert)-\kappa)_{+}^{2}\,dx (5.32)

hold with cc(data,δ1)c\equiv c(\textnormal{{data}},\delta_{1}) and 𝔪\mathfrak{m} is defined in (5.21).

Proof

The integrand Hϱ,i()H_{\varrho,\textnormal{i}}(\cdot) is of the type considered in Proposition 1 (compare with (5.22) and (5.23)), and we apply this last result to vv, with Bτ8BhB_{\tau}\equiv 8B_{h}. It follows that for every δ1>0\delta_{1}>0

DvL(6Bh)\displaystyle\lVert Dv\rVert_{L^{\infty}(6B_{h})} (4.5)\displaystyle\stackrel{{\scriptstyle(\ref{8.fz})}}{{\leq}} cϱ,iδ1/q(DuϱL(8Bh))DuϱL(8Bh)+c\displaystyle c\mathbb{H}_{\varrho,\textnormal{i}}^{\delta_{1}/q}\left(\lVert Du_{\varrho}\rVert_{L^{\infty}(8B_{h})}\right)\lVert Du_{\varrho}\rVert_{L^{\infty}(8B_{h})}+c (5.33)
(5.21)\displaystyle\stackrel{{\scriptstyle(\ref{mmm0})}}{{\leq}} cϱ,iδ1/q(𝔪)𝔪+c\displaystyle c\mathbb{H}_{\varrho,\textnormal{i}}^{\delta_{1}/q}(\mathfrak{m})\mathfrak{m}+c

holds provided 1μ<μn(δ1/2)μn(δ1/q)1\leq\mu<\mu_{n}(\delta_{1}/2)\leq\mu_{n}(\delta_{1}/q), where cc(data,δ1)c\equiv c(\textnormal{{data}},\delta_{1}). Here, as in (3.2)3, it is ϱ,i(t):=tg(t)+a~ϱ,i(8Bh)[t2+ω2]q/2+1\mathbb{H}_{\varrho,\textnormal{i}}(t):=tg(t)+\tilde{a}_{\varrho,\textnormal{i}}(8B_{h})[t^{2}+\omega^{2}]^{q/2}+1 for t0t\geq 0 and in fact in (5.33) we have used (3.3). From (5.33) we can derive (5.31) using (1.12) with ε=q1>0\varepsilon=q-1>0. Next, we apply (4.7), that gives

2Bh|D(Eϱ,i(|Dv|)κ)+|2𝑑x\displaystyle\int_{2B_{h}}\lvert D(E_{\varrho,\textnormal{i}}(\lvert Dv\rvert)-\kappa)_{+}\rvert^{2}\,dx
c(DvL(4Bh)+1)δ1/2|h|2β04Bh(Eϱ,i(|Dv|)κ)+2𝑑x\displaystyle\qquad\leq\frac{c\left(\|Dv\|_{L^{\infty}(4B_{h})}+1\right)^{\delta_{1}/2}}{\lvert h\rvert^{2\beta_{0}}}\int_{4B_{h}}(E_{\varrho,\textnormal{i}}(\lvert Dv\rvert)-\kappa)_{+}^{2}\,dx

so that (5.32) follows using (5.31) in the above inequality and observing that (1+δ1)δ1/2δ1(1+\delta_{1})\delta_{1}/2\leq\delta_{1}.

5.5 A comparison estimate

This is in the following:

Lemma 9

Let uϱW1,q(1)u_{\varrho}\in W^{1,q}(\mathcal{B}_{1}) be as in (5.14), vuϱ+W01,q(8Bh)v\in u_{\varrho}+W^{1,q}_{0}(8B_{h}) as in (5.28) and δ2(0,1/2)\delta_{2}\in(0,1/2). The inequality

8Bh𝒱ϱ,i2(Duϱ,Dv;8Bh)𝑑x\displaystyle\int_{8B_{h}}\mathcal{V}_{\varrho,\textnormal{i}}^{2}(Du_{\varrho},Dv;8B_{h})\,dx c|h|β0α~𝔐1δ2/22μϱα8Bh(|Duϱ|+1)q1+δ2𝑑x\displaystyle\leq c\lvert h\rvert^{\beta_{0}\tilde{\alpha}}\mathfrak{M}^{\frac{1-\delta_{2}/2}{2-\mu}}\varrho^{\alpha}\int_{8B_{h}}(|Du_{\varrho}|+1)^{q-1+\delta_{2}}\,dx
+c|h|β0α~𝔐1δ2/22μϱα08Bh(|Duϱ|+1)3δ2𝑑x\displaystyle\quad\ +c\lvert h\rvert^{\beta_{0}\tilde{\alpha}}\mathfrak{M}^{\frac{1-\delta_{2}/2}{2-\mu}}\varrho^{\alpha_{0}}\int_{8B_{h}}(|Du_{\varrho}|+1)^{3\delta_{2}}\,dx (5.34)

holds with cc(data,δ2)c\equiv c(\textnormal{{data}},\delta_{2}), where α~=min{α,α0}\tilde{\alpha}=\min\{\alpha,\alpha_{0}\}, 𝒱ϱ,i()\mathcal{V}_{\varrho,\textnormal{i}}(\cdot) is as in (5.24), and 𝔐\mathfrak{M} is any number satisfying (5.13) and therefore (5.27).

Proof

From (5.21) and (5.27), and yet recalling (3.16), we deduce

{a~ϱ,i(8Bh)[ω(𝔪)]qc(𝒶ϱ)σε()[ω(Duϱ)]qL(8Bh)+cc𝔐𝔪ω(𝔪)c(n,q)𝔐12μ,\begin{cases}\ \tilde{a}_{\varrho,\textnormal{i}}(8B_{h})[\ell_{\omega}(\mathfrak{m})]^{q}\leq c\lVert(\mathcal{a}_{\varrho})_{\sigma_{\varepsilon}}(\cdot)[\ell_{\omega}(Du_{\varrho})]^{q}\rVert_{L^{\infty}(8B_{h})}+c\leq c\mathfrak{M}\\ \ \mathfrak{m}\leq\ell_{\omega}(\mathfrak{m})\leq c(n,q)\mathfrak{M}^{\frac{1}{2-\mu}}\,,\end{cases} (5.35)

where cc(μ,q,aL)c\equiv c(\mu,q,\|a\|_{L^{\infty}}) in (5.35)1. We have

8Bh𝒱ϱ,i2(Duϱ,Dv;8Bh)𝑑x\displaystyle\int_{8B_{h}}\mathcal{V}_{\varrho,\textnormal{i}}^{2}(Du_{\varrho},Dv;8B_{h})\,dx
(5.23)4c8BhzHϱ,i(Duϱ)zHϱ,i(Dv),DuϱDv𝑑x\displaystyle\quad\stackrel{{\scriptstyle\eqref{5}_{4}}}{{\leq}}c\int_{8B_{h}}\langle\partial_{z}H_{\varrho,\textnormal{i}}(Du_{\varrho})-\partial_{z}H_{\varrho,\textnormal{i}}(Dv),Du_{\varrho}-Dv\rangle\,dx
=(5.29)c8BhzHϱ,i(Duϱ),DuϱDv𝑑x\displaystyle\quad\stackrel{{\scriptstyle\eqref{el2}}}{{=}}c\int_{8B_{h}}\langle\partial_{z}H_{\varrho,\textnormal{i}}(Du_{\varrho}),Du_{\varrho}-Dv\rangle\,dx
=(5.17)c8BhzHϱ,i(Duϱ)zHϱ(x,Duϱ),DuϱDv𝑑x\displaystyle\quad\stackrel{{\scriptstyle\eqref{el}}}{{=}}c\int_{8B_{h}}\langle\partial_{z}H_{\varrho,\textnormal{i}}(Du_{\varrho})-\partial_{z}H_{\varrho}(x,Du_{\varrho}),Du_{\varrho}-Dv\rangle\,dx
(5.18)3c|h|β0αϱα8Bh[ω(Duϱ)]q1(|Duϱ|+|Dv|)𝑑x\displaystyle\ \ \ \,\stackrel{{\scriptstyle\eqref{assr}_{3}}}{{\leq}}c\lvert h\rvert^{\beta_{0}\alpha}\varrho^{\alpha}\int_{8B_{h}}[\ell_{\omega}(Du_{\varrho})]^{q-1}(\lvert Du_{\varrho}\rvert+\lvert Dv\rvert)\,dx
+c|h|β0α0ϱα08Bhg(|Duϱ|)(|Duϱ|+|Dv|)𝑑x\displaystyle\quad\qquad\quad+c\lvert h\rvert^{\beta_{0}\alpha_{0}}\varrho^{\alpha_{0}}\int_{8B_{h}}g(|Du_{\varrho}|)(\lvert Du_{\varrho}\rvert+\lvert Dv\rvert)\,dx
(1.12)c|h|β0α~(𝔪q1ϱα+𝔪δ2ϱα0)8Bh(|Duϱ|+|Dv|)𝑑x\displaystyle\quad\stackrel{{\scriptstyle\eqref{0.1}}}{{\leq}}c\lvert h\rvert^{\beta_{0}\tilde{\alpha}}(\mathfrak{m}^{q-1}\varrho^{\alpha}+\mathfrak{m}^{\delta_{2}}\varrho^{\alpha_{0}})\int_{8B_{h}}(\lvert Du_{\varrho}\rvert+\lvert Dv\rvert)\,dx
(1.11)c|h|β0α~(𝔪q1ϱα+𝔪δ2ϱα0)8Bh[Hϱ,i(Duϱ)+Hϱ,i(Dv)]𝑑x\displaystyle\quad\stackrel{{\scriptstyle\eqref{assif}}}{{\leq}}c\lvert h\rvert^{\beta_{0}\tilde{\alpha}}(\mathfrak{m}^{q-1}\varrho^{\alpha}+\mathfrak{m}^{\delta_{2}}\varrho^{\alpha_{0}})\int_{8B_{h}}[H_{\varrho,\textnormal{i}}(Du_{\varrho})+H_{\varrho,\textnormal{i}}(Dv)]\,dx
(5.30)c|h|β0α~(𝔪q1ϱα+𝔪δ2ϱα0)8BhHϱ,i(Duϱ)𝑑x\displaystyle\quad\stackrel{{\scriptstyle\eqref{enes}}}{{\leq}}c\lvert h\rvert^{\beta_{0}\tilde{\alpha}}(\mathfrak{m}^{q-1}\varrho^{\alpha}+\mathfrak{m}^{\delta_{2}}\varrho^{\alpha_{0}})\int_{8B_{h}}H_{\varrho,\textnormal{i}}(Du_{\varrho})\,dx
(1.11)c|h|β0α~(𝔪q1ϱα+𝔪δ2ϱα0)\displaystyle\quad\stackrel{{\scriptstyle\eqref{assif}}}{{\leq}}c\lvert h\rvert^{\beta_{0}\tilde{\alpha}}(\mathfrak{m}^{q-1}\varrho^{\alpha}+\mathfrak{m}^{\delta_{2}}\varrho^{\alpha_{0}})
8Bh(|Duϱ|g(|Duϱ|)+a~ϱ,i(8Bh)|Duϱ|q+1)dx\displaystyle\hskip 36.98857pt\cdot\int_{8B_{h}}(\lvert Du_{\varrho}\rvert g(\lvert Du_{\varrho}\rvert)+\tilde{a}_{\varrho,\textnormal{i}}(8B_{h})\lvert Du_{\varrho}\rvert^{q}+1)\,dx
(1.12)c|h|β0α~(𝔪q1ϱα+𝔪δ2ϱα0)8Bh(|Duϱ|1+δ2/2+a~ϱ,i(8Bh)|Duϱ|q)𝑑x\displaystyle\quad\stackrel{{\scriptstyle\eqref{0.1}}}{{\leq}}c\lvert h\rvert^{\beta_{0}\tilde{\alpha}}(\mathfrak{m}^{q-1}\varrho^{\alpha}+\mathfrak{m}^{\delta_{2}}\varrho^{\alpha_{0}})\int_{8B_{h}}(\lvert Du_{\varrho}\rvert^{1+\delta_{2}/2}+\tilde{a}_{\varrho,\textnormal{i}}(8B_{h})\lvert Du_{\varrho}\rvert^{q})\,dx
+c|h|β0α~(𝔪q1ϱα+𝔪δ2ϱα0)|Bh|,\displaystyle\quad\qquad\quad+c\lvert h\rvert^{\beta_{0}\tilde{\alpha}}(\mathfrak{m}^{q-1}\varrho^{\alpha}+\mathfrak{m}^{\delta_{2}}\varrho^{\alpha_{0}})|B_{h}|\,,

where cc(data,δ2)c\equiv c(\textnormal{{data}},\delta_{2}). We now estimate the four integrals stemming from the second-last line in the above display. As δ2<1/2\delta_{2}<1/2, we have

𝔪δ2ϱα08Bh|Duϱ|1+δ2/2𝑑x\displaystyle\mathfrak{m}^{\delta_{2}}\varrho^{\alpha_{0}}\int_{8B_{h}}\lvert Du_{\varrho}\rvert^{1+\delta_{2}/2}\,dx =𝔪δ2ϱα08Bh|Duϱ|12δ2|Duϱ|5δ2/2𝑑x\displaystyle=\mathfrak{m}^{\delta_{2}}\varrho^{\alpha_{0}}\int_{8B_{h}}\lvert Du_{\varrho}\rvert^{1-2\delta_{2}}\lvert Du_{\varrho}\rvert^{5\delta_{2}/2}\,dx
c𝔪1δ2ϱα08Bh|Duϱ|5δ2/2𝑑x\displaystyle\leq c\mathfrak{m}^{1-\delta_{2}}\varrho^{\alpha_{0}}\int_{8B_{h}}\lvert Du_{\varrho}\rvert^{5\delta_{2}/2}\,dx
c𝔐1δ2/22μϱα08Bh(|Duϱ|+1)3δ2𝑑x\displaystyle\leq c\mathfrak{M}^{\frac{1-\delta_{2}/2}{2-\mu}}\varrho^{\alpha_{0}}\int_{8B_{h}}(|Du_{\varrho}|+1)^{3\delta_{2}}\,dx

and in the last line we have used (5.35)2. Similarly, this time using (5.35)1, we find

𝔪δ2ϱα08Bha~ϱ,i(8Bh)|Duϱ|q𝑑x\displaystyle\mathfrak{m}^{\delta_{2}}\varrho^{\alpha_{0}}\int_{8B_{h}}\tilde{a}_{\varrho,\textnormal{i}}(8B_{h})\lvert Du_{\varrho}\rvert^{q}\,dx
c[a~ϱ,i(8Bh)]1δ2/q𝔪δ2ϱα08Bh|Duϱ|q2δ2|Duϱ|2δ2𝑑x\displaystyle\qquad\leq c[\tilde{a}_{\varrho,\textnormal{i}}(8B_{h})]^{1-\delta_{2}/q}\mathfrak{m}^{\delta_{2}}\varrho^{\alpha_{0}}\int_{8B_{h}}\lvert Du_{\varrho}\rvert^{q-2\delta_{2}}\lvert Du_{\varrho}\rvert^{2\delta_{2}}\,dx
c[a~ϱ,i(8Bh)]1δ2/q𝔪q(1δ2/q)ϱα08Bh|Duϱ|2δ2𝑑x\displaystyle\qquad\leq c[\tilde{a}_{\varrho,\textnormal{i}}(8B_{h})]^{1-\delta_{2}/q}\mathfrak{m}^{q\left(1-\delta_{2}/q\right)}\varrho^{\alpha_{0}}\int_{8B_{h}}\lvert Du_{\varrho}\rvert^{2\delta_{2}}\,dx
c𝔐1δ2/qϱα08Bh|Duϱ|2δ2𝑑x\displaystyle\qquad\leq c\mathfrak{M}^{1-\delta_{2}/q}\varrho^{\alpha_{0}}\int_{8B_{h}}|Du_{\varrho}|^{2\delta_{2}}\,dx
c𝔐1δ2/22μϱα08Bh(|Duϱ|+1)3δ2𝑑x.\displaystyle\qquad\leq c\mathfrak{M}^{\frac{1-\delta_{2}/2}{2-\mu}}\varrho^{\alpha_{0}}\int_{8B_{h}}(|Du_{\varrho}|+1)^{3\delta_{2}}\,dx\,.

Next, note that δ2<1/2\delta_{2}<1/2 and q<3/2q<3/2 (follows from (1.7)1) implies 2qδ2/2>02-q-\delta_{2}/2>0 and therefore we can estimate

𝔪q1ϱα8Bh|Duϱ|1+δ2/2𝑑x\displaystyle\mathfrak{m}^{q-1}\varrho^{\alpha}\int_{8B_{h}}\lvert Du_{\varrho}\rvert^{1+\delta_{2}/2}\,dx =𝔪q1ϱα8Bh|Duϱ|2qδ2/2|Duϱ|q1+δ2𝑑x\displaystyle=\mathfrak{m}^{q-1}\varrho^{\alpha}\int_{8B_{h}}\lvert Du_{\varrho}\rvert^{2-q-\delta_{2}/2}\lvert Du_{\varrho}\rvert^{q-1+\delta_{2}}\,dx
c𝔪1δ2/2ϱα8Bh|Duϱ|q1+δ2𝑑x\displaystyle\leq c\mathfrak{m}^{1-\delta_{2}/2}\varrho^{\alpha}\int_{8B_{h}}\lvert Du_{\varrho}\rvert^{q-1+\delta_{2}}\,dx
c𝔐1δ2/22μϱα8Bh|Duϱ|q1+δ2𝑑x\displaystyle\leq c\mathfrak{M}^{\frac{1-\delta_{2}/2}{2-\mu}}\varrho^{\alpha}\int_{8B_{h}}|Du_{\varrho}|^{q-1+\delta_{2}}\,dx

where we again used (5.35)2. By means of (5.35)1 we find

𝔪q1ϱα8Bha~ϱ,i(8Bh)|Duϱ|q𝑑x\displaystyle\mathfrak{m}^{q-1}\varrho^{\alpha}\int_{8B_{h}}\tilde{a}_{\varrho,\textnormal{i}}(8B_{h})\lvert Du_{\varrho}\rvert^{q}\,dx
c[a~ϱ,i(8Bh)]1δ2/q𝔪q1ϱα8Bh|Duϱ|1δ2|Duϱ|q1+δ2𝑑x\displaystyle\qquad\leq c[\tilde{a}_{\varrho,\textnormal{i}}(8B_{h})]^{1-\delta_{2}/q}\mathfrak{m}^{q-1}\varrho^{\alpha}\int_{8B_{h}}\lvert Du_{\varrho}\rvert^{1-\delta_{2}}\lvert Du_{\varrho}\rvert^{q-1+\delta_{2}}\,dx
c[a~ϱ,i(8Bh)]1δ2/q𝔪q(1δ2/q)ϱα8Bh|Duϱ|q1+δ2𝑑x\displaystyle\qquad\leq c[\tilde{a}_{\varrho,\textnormal{i}}(8B_{h})]^{1-\delta_{2}/q}\mathfrak{m}^{q\left(1-\delta_{2}/q\right)}\varrho^{\alpha}\int_{8B_{h}}\lvert Du_{\varrho}\rvert^{q-1+\delta_{2}}\,dx
c𝔐1δ2/qϱα8Bh|Duϱ|q1+δ2𝑑x\displaystyle\qquad\leq c\mathfrak{M}^{1-\delta_{2}/q}\varrho^{\alpha}\int_{8B_{h}}|Du_{\varrho}|^{q-1+\delta_{2}}\,dx
c𝔐1δ2/22μϱα8Bh|Duϱ|q1+δ2𝑑x.\displaystyle\qquad\leq c\mathfrak{M}^{\frac{1-\delta_{2}/2}{2-\mu}}\varrho^{\alpha}\int_{8B_{h}}|Du_{\varrho}|^{q-1+\delta_{2}}\,dx\,.

Finally, observe that δ2<2q\delta_{2}<2-q as q<3/2q<3/2 so that

𝔪q1ϱα+𝔪δ2ϱα0c𝔐1δ22μ(ϱα+ϱα0)𝔐1δ2/22μ(ϱα+ϱα0).\mathfrak{m}^{q-1}\varrho^{\alpha}+\mathfrak{m}^{\delta_{2}}\varrho^{\alpha_{0}}\leq c\mathfrak{M}^{\frac{1-\delta_{2}}{2-\mu}}(\varrho^{\alpha}+\varrho^{\alpha_{0}})\leq\mathfrak{M}^{\frac{1-\delta_{2}/2}{2-\mu}}(\varrho^{\alpha}+\varrho^{\alpha_{0}})\,.

Merging the content of the last six displays yields (9) with the asserted dependence of the constant cc. Notice that the dependence of the constants on δ2\delta_{2} comes from (1.12) (that has been used with ε=δ2/2\varepsilon=\delta_{2}/2). Notice also that (9) holds whenever we are assuming (1.11) with μ[1,3/2)\mu\in[1,3/2).

5.6 A fractional Caccioppoli inequality via nonlinear
atomic type decompositions

Lemma 10 (Fractional Caccoppoli inequality)

Let uω,εW1,q(Br)u_{\omega,\varepsilon}\in W^{1,q}(B_{r}) be as in (5.3); fix numbers δ1(0,1)\delta_{1}\in(0,1), δ2(0,1/2)\delta_{2}\in(0,1/2) and a ball Bϱ(x0)BrB_{\varrho}(x_{0})\Subset B_{r}. The inequality

(Bϱ/2(x0)(Eω,σε(x,|Duω,ε|)κ)+2χdx)1/χ\displaystyle\left(\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B_{\varrho/2}(x_{0})}(E_{\omega,\sigma_{\varepsilon}}(x,\lvert Du_{\omega,\varepsilon}\rvert)-\kappa)_{+}^{2\chi}\,dx\right)^{1/\chi}
+ϱ2βn[(Eω,σε(,|Duω,ε|)κ)+]β,2;Bϱ/2(x0)2\displaystyle\quad\quad+\varrho^{2\beta-n}[(E_{\omega,\sigma_{\varepsilon}}(\cdot,\lvert Du_{\omega,\varepsilon}\rvert)-\kappa)_{+}]_{\beta,2;B_{\varrho/2}(x_{0})}^{2}
c𝔐2𝔰1Bϱ(x0)(Eω,σε(x,|Duω,ε|)κ)+2dx\displaystyle\quad\leq c\mathfrak{M}^{2\mathfrak{s}_{1}}\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B_{\varrho}(x_{0})}(E_{\omega,\sigma_{\varepsilon}}(x,\lvert Du_{\omega,\varepsilon}\rvert)-\kappa)_{+}^{2}\,dx
+c𝔐2𝔰2ϱ2αBϱ(x0)(|Duω,ε|+1)2(q1+δ2)dx\displaystyle\qquad\quad+c\mathfrak{M}^{2\mathfrak{s}_{2}}\varrho^{2\alpha}\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B_{\varrho}(x_{0})}(|Du_{\omega,\varepsilon}|+1)^{2(q-1+\delta_{2})}\,dx
+c𝔐2𝔰3ϱαBϱ(x0)(|Duω,ε|+1)q1+δ2dx\displaystyle\qquad\quad+c\mathfrak{M}^{2\mathfrak{s}_{3}}\varrho^{\alpha}\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B_{\varrho}(x_{0})}(|Du_{\omega,\varepsilon}|+1)^{q-1+\delta_{2}}\,dx
+c𝔐2𝔰3ϱα0Bϱ(x0)(|Duω,ε|+1)3δ2dx\displaystyle\qquad\quad+c\mathfrak{M}^{2\mathfrak{s}_{3}}\varrho^{\alpha_{0}}\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B_{\varrho}(x_{0})}(|Du_{\omega,\varepsilon}|+1)^{3\delta_{2}}\,dx (5.36)

holds whenever

β(0,α)with χ(β):=nn2β, α:=α~α~+2=min{α,α0}min{α,α0}+2,\displaystyle\beta\in(0,\alpha_{*})\ \ \mbox{with $\chi(\beta):=\frac{n}{n-2\beta},$\ \ \ $\alpha_{*}:=\frac{\tilde{\alpha}}{\tilde{\alpha}+2}=\frac{\min\{\alpha,\alpha_{0}\}}{\min\{\alpha,\alpha_{0}\}+2}$}\,, (5.37)

and provided

{ 1μ<μn(δ1/2)E~ω,σε(,|Duω,ε|)L(Bϱ(x0))+1𝔐,\begin{cases}\,1\leq\mu<\mu_{n}(\delta_{1}/2)\\ \,\lVert\tilde{E}_{\omega,\sigma_{\varepsilon}}(\cdot,\lvert Du_{\omega,\varepsilon}\rvert)\rVert_{L^{\infty}(B_{\varrho}(x_{0}))}+1\leq\mathfrak{M}\,,\end{cases} (5.38)

where cc(data,δ1,δ2,β)c\equiv c(\textnormal{{data}},\delta_{1},\delta_{2},\beta) and

{𝔰1𝔰1(μ,δ1):=δ12(2μ)𝔰2𝔰2(μ,δ2):=1δ22μ𝔰3𝔰3(μ,δ1,δ2):=3μ+(q+1)δ1δ2/22(2μ).\begin{cases}\displaystyle\ \mathfrak{s}_{1}\equiv\mathfrak{s}_{1}(\mu,\delta_{1}):=\frac{\delta_{1}}{2(2-\mu)}\\ \displaystyle\ \mathfrak{s}_{2}\equiv\mathfrak{s}_{2}(\mu,\delta_{2}):=\frac{1-\delta_{2}}{2-\mu}\\ \displaystyle\ \mathfrak{s}_{3}\equiv\mathfrak{s}_{3}(\mu,\delta_{1},\delta_{2}):=\frac{3-\mu+(q+1)\delta_{1}-\delta_{2}/2}{2(2-\mu)}\,.\end{cases} (5.39)

The function μn()\mu_{n}(\cdot) is the one defined in Lemma 8.

Proof

This will be obtained by a technique that works as an analogue of dyadic decomposition in Besov spaces, but using the functions vv(8Bh)v\equiv v(8B_{h}) in (5.28) as “atoms”; see Remark 3 below. We therefore divide the proof in two steps.

Step 1: Estimates on a single ball BhB_{h}. Here we again use the notation and the results in Sections 5.4-5.5. In particular, here we again argue on a fixed ball BhB_{h}. Our goal here is to prove estimate (5.43) below. We recall that basic properties of difference quotients yield

Bh|τh(Eϱ,i(|Dv|)κ)+|2𝑑x|h|22Bh|D(Eϱ,i(|Dv|)κ)+|2𝑑x\int_{B_{h}}\lvert\tau_{h}(E_{\varrho,\textnormal{i}}(\lvert Dv\rvert)-\kappa)_{+}\rvert^{2}\,dx\leq|h|^{2}\int_{2B_{h}}\lvert D(E_{\varrho,\textnormal{i}}(\lvert Dv\rvert)-\kappa)_{+}\rvert^{2}\,dx

where κ0\kappa\geq 0 is any number (recall that |h||h|β0|h|\leq|h|^{\beta_{0}}). Using this in connection with (5.32) we have

Bh|τh(Eϱ,i(|Dv|)κ)+|2𝑑xc|h|2(1β0)𝔪δ14Bh(Eϱ,i(|Dv|)κ)+2𝑑x\int_{B_{h}}\lvert\tau_{h}(E_{\varrho,\textnormal{i}}(\lvert Dv\rvert)-\kappa)_{+}\rvert^{2}\,dx\leq c\lvert h\rvert^{2(1-\beta_{0})}\mathfrak{m}^{\delta_{1}}\int_{4B_{h}}(E_{\varrho,\textnormal{i}}(\lvert Dv\rvert)-\kappa)_{+}^{2}\,dx (5.40)

with cc(data,δ1)c\equiv c(\textnormal{{data}},\delta_{1}) and moreover (5.31) holds by (5.38)1; recall that 𝔪𝔪(8Bh)\mathfrak{m}\equiv\mathfrak{m}(8B_{h}) is defined in (5.21). Use of (5.32) is legitimate here as (5.38)1 is assumed. Let us recall that here it is Eϱ,i(|z|)Eϱ,i(|z|;8Bh)E_{\varrho,\textnormal{i}}(\lvert z\rvert)\equiv E_{\varrho,\textnormal{i}}(\lvert z\rvert;8B_{h}). Let us now record a couple of auxiliary estimates. The first is obtained as follows:

4Bh|Eϱ(x,|Duϱ|)Eϱ,i(|Duϱ|)|2𝑑x\displaystyle\int_{4B_{h}}\lvert E_{\varrho}(x,\lvert Du_{\varrho}\rvert)-E_{\varrho,\textnormal{i}}(\lvert Du_{\varrho}\rvert)\rvert^{2}\,dx
(5.26)c|h|2β0αϱ2α4Bh[ω(Duϱ)]2q𝑑x\displaystyle\qquad\stackrel{{\scriptstyle\eqref{0.00bis}}}{{\leq}}c\lvert h\rvert^{2\beta_{0}\alpha}\varrho^{2\alpha}\int_{4B_{h}}[\ell_{\omega}(Du_{\varrho})]^{2q}\,dx
(5.21)c|h|2β0α𝔪2(1δ2)ϱ2α4Bh(|Duϱ|+1)2(q1+δ2)𝑑x\displaystyle\qquad\stackrel{{\scriptstyle(\ref{mmm0})}}{{\leq}}c\lvert h\rvert^{2\beta_{0}\alpha}\mathfrak{m}^{2(1-\delta_{2})}\varrho^{2\alpha}\int_{4B_{h}}(|Du_{\varrho}|+1)^{2(q-1+\delta_{2})}\,dx
(5.35)2c|h|β0α~𝔐2(1δ2)2μϱ2α4Bh(|Duϱ|+1)2(q1+δ2)𝑑x\displaystyle\qquad\stackrel{{\scriptstyle\eqref{4.1}_{2}}}{{\leq}}c\lvert h\rvert^{\beta_{0}\tilde{\alpha}}\mathfrak{M}^{\frac{2(1-\delta_{2})}{2-\mu}}\varrho^{2\alpha}\int_{4B_{h}}(|Du_{\varrho}|+1)^{2(q-1+\delta_{2})}\,dx (5.41)

for cc(n,q)c\equiv c(n,q). For the second auxiliary inequality, we estimate

4Bh|Eϱ,i(|Duϱ|)Eϱ,i(|Dv|)|2𝑑x\displaystyle\int_{4B_{h}}\lvert E_{\varrho,\textnormal{i}}(\lvert Du_{\varrho}\rvert)-E_{\varrho,\textnormal{i}}(\lvert Dv\rvert)\rvert^{2}\,dx
(6)c4Bh(|Duϱ|2+|Dv|2+1)1μ|DuϱDv|2𝑑x\displaystyle\quad\stackrel{{\scriptstyle(\ref{0.000})}}{{\leq}}c\int_{4B_{h}}(\lvert Du_{\varrho}\rvert^{2}+\lvert Dv\rvert^{2}+1)^{1-\mu}\lvert Du_{\varrho}-Dv\rvert^{2}\,dx
+c[a~ϱ,i(8Bh)]24Bh(|Duϱ|2+|Dv|2+ω2)q1|DuϱDv|2𝑑x\displaystyle\quad\qquad+c[\tilde{a}_{\varrho,\textnormal{i}}(8B_{h})]^{2}\int_{4B_{h}}(\lvert Du_{\varrho}\rvert^{2}+\lvert Dv\rvert^{2}+\omega^{2})^{q-1}\lvert Du_{\varrho}-Dv\rvert^{2}\,dx
(5.31)c𝔪(2μ)(1+δ1)4Bh(|Duϱ|2+|Dv|2+1)μ/2|DuϱDv|2𝑑x\displaystyle\quad\stackrel{{\scriptstyle(\ref{stim3})}}{{\leq}}c\mathfrak{m}^{(2-\mu)(1+\delta_{1})}\int_{4B_{h}}(\lvert Du_{\varrho}\rvert^{2}+\lvert Dv\rvert^{2}+1)^{-\mu/2}\lvert Du_{\varrho}-Dv\rvert^{2}\,dx
+c[a~ϱ,i(8Bh)]2𝔪q(1+δ1)\displaystyle\quad\qquad+c[\tilde{a}_{\varrho,\textnormal{i}}(8B_{h})]^{2}\mathfrak{m}^{q(1+\delta_{1})}
4Bh(|Duϱ|2+|Dv|2+ω2)(q2)/2|DuϱDv|2dx\displaystyle\hskip 48.36967pt\cdot\int_{4B_{h}}(\lvert Du_{\varrho}\rvert^{2}+\lvert Dv\rvert^{2}+\omega^{2})^{(q-2)/2}\lvert Du_{\varrho}-Dv\rvert^{2}\,dx
(2.4)c𝔪(2μ)(1+δ1)4Bh|V1,2μ(Duϱ)V1,2μ(Dv)|2𝑑x\displaystyle\quad\stackrel{{\scriptstyle(\ref{Vm})}}{{\leq}}c\mathfrak{m}^{(2-\mu)(1+\delta_{1})}\int_{4B_{h}}\lvert V_{1,2-\mu}(Du_{\varrho})-V_{1,2-\mu}(Dv)\rvert^{2}\,dx
+c(a~ϱ,i(8Bh)[ω(𝔪)]q)𝔪qδ1\displaystyle\qquad\quad+c\left(\tilde{a}_{\varrho,\textnormal{i}}(8B_{h})[\ell_{\omega}(\mathfrak{m})]^{q}\right)\mathfrak{m}^{q\delta_{1}}
4Bha~ϱ,i(8Bh)|Vω,q(Duϱ)Vω,q(Dv)|2dx\displaystyle\hskip 48.36967pt\cdot\int_{4B_{h}}\tilde{a}_{\varrho,\textnormal{i}}(8B_{h})\lvert V_{\omega,q}(Du_{\varrho})-V_{\omega,q}(Dv)\rvert^{2}\,dx
(5.24)c(𝔪(2μ)(1+δ1)+𝔐𝔪qδ1)4Bh𝒱ϱ,i2(Duϱ,Dv;8Bh)𝑑x,\displaystyle\quad\stackrel{{\scriptstyle\eqref{deffiV}}}{{\leq}}c\left(\mathfrak{m}^{(2-\mu)(1+\delta_{1})}+\mathfrak{M}\mathfrak{m}^{q\delta_{1}}\right)\int_{4B_{h}}\mathcal{V}_{\varrho,\textnormal{i}}^{2}(Du_{\varrho},Dv;8B_{h})\,dx\,,

where cc(data,δ1)c\equiv c(\textnormal{{data}},\delta_{1}) and in the last line we have also used (5.35)1. Using (5.35)2, we gain

4Bh|Eϱ,i(|Duϱ|)Eϱ,i(|Dv|)|2𝑑xc𝔐2μ+qδ12μ4Bh𝒱ϱ,i2(Duϱ,Dv;8Bh)𝑑x.\int_{4B_{h}}\lvert E_{\varrho,\textnormal{i}}(\lvert Du_{\varrho}\rvert)-E_{\varrho,\textnormal{i}}(\lvert Dv\rvert)\rvert^{2}\,dx\leq c\mathfrak{M}^{\frac{2-\mu+q\delta_{1}}{2-\mu}}\int_{4B_{h}}\mathcal{V}_{\varrho,\textnormal{i}}^{2}(Du_{\varrho},Dv;8B_{h})\,dx\,.

Using this last estimate with (9) we conclude with

4Bh|Eϱ,i(|Duϱ|)Eϱ,i(|Dv|)|2𝑑x\displaystyle\int_{4B_{h}}\lvert E_{\varrho,\textnormal{i}}(\lvert Du_{\varrho}\rvert)-E_{\varrho,\textnormal{i}}(\lvert Dv\rvert)\rvert^{2}\,dx
c|h|β0α~𝔐3μ+qδ1δ2/22μϱα8Bh(|Duϱ|+1)q1+δ2𝑑x\displaystyle\qquad\leq c\lvert h\rvert^{\beta_{0}\tilde{\alpha}}\mathfrak{M}^{\frac{3-\mu+q\delta_{1}-\delta_{2}/2}{2-\mu}}\varrho^{\alpha}\int_{8B_{h}}(|Du_{\varrho}|+1)^{q-1+\delta_{2}}\,dx
+c|h|β0α~𝔐3μ+qδ1δ2/22μϱα08Bh(|Duϱ|+1)3δ2𝑑x\displaystyle\qquad\quad+c\lvert h\rvert^{\beta_{0}\tilde{\alpha}}\mathfrak{M}^{\frac{3-\mu+q\delta_{1}-\delta_{2}/2}{2-\mu}}\varrho^{\alpha_{0}}\int_{8B_{h}}(|Du_{\varrho}|+1)^{3\delta_{2}}\,dx (5.42)

where cc(data,δ1,δ2)c\equiv c(\textnormal{{data}},\delta_{1},\delta_{2}), that is the second auxiliary estimate we were aiming at. Triangle inequality now yields

Bh|τh(Eϱ(x,|Duϱ|)κ)+|2𝑑xcBh|τh(Eϱ,i(|Dv|)κ)+|2𝑑x\displaystyle\int_{B_{h}}\lvert\tau_{h}(E_{\varrho}(x,\lvert Du_{\varrho}\rvert)-\kappa)_{+}\rvert^{2}\,dx\leq c\int_{B_{h}}\lvert\tau_{h}(E_{\varrho,\textnormal{i}}(\lvert Dv\rvert)-\kappa)_{+}\rvert^{2}\,dx
+cBh|(Eϱ(x+h,|Duϱ(x+h)|)κ)+(Eϱ,i(|Duϱ(x+h)|)κ)+|2𝑑x\displaystyle\ \ +c\int_{B_{h}}\lvert(E_{\varrho}(x+h,\lvert Du_{\varrho}(x+h)\rvert)-\kappa)_{+}-(E_{\varrho,\textnormal{i}}(\lvert Du_{\varrho}(x+h)\rvert)-\kappa)_{+}\rvert^{2}\,dx
+cBh|(Eϱ,i(|Duϱ(x+h)|)κ)+(Eϱ,i(|Dv(x+h)|)κ)+|2𝑑x\displaystyle\ \ +c\int_{B_{h}}\lvert(E_{\varrho,\textnormal{i}}(\lvert Du_{\varrho}(x+h)\rvert)-\kappa)_{+}-(E_{\varrho,\textnormal{i}}(\lvert Dv(x+h)\rvert)-\kappa)_{+}\rvert^{2}\,dx
+cBh|(Eϱ,i(|Dv|)κ)+(Eϱ,i(|Duϱ|)κ)+|2𝑑x\displaystyle\ \ +c\int_{B_{h}}\lvert(E_{\varrho,\textnormal{i}}(\lvert Dv\rvert)-\kappa)_{+}-(E_{\varrho,\textnormal{i}}(\lvert Du_{\varrho}\rvert)-\kappa)_{+}\rvert^{2}\,dx
+cBh|(Eϱ,i(|Duϱ|)κ)+(Eϱ(x,|Duϱ|)κ)+|2𝑑x.\displaystyle\ \ +c\int_{B_{h}}\lvert(E_{\varrho,\textnormal{i}}(\lvert Du_{\varrho}\rvert)-\kappa)_{+}-(E_{\varrho}(x,\lvert Du_{\varrho}\rvert)-\kappa)_{+}\rvert^{2}\,dx\,.

Using the standard property of translations

Bh|g(x+h)|2𝑑x4Bh|g|2𝑑x\int_{B_{h}}\lvert g(x+h)\rvert^{2}\,dx\leq\int_{4B_{h}}\lvert g\rvert^{2}\,dx

valid for every gL2(4Bh)g\in L^{2}(4B_{h}), (for this note that h+Bh4Bhh+B_{h}\subset 4B_{h} as |h|1|h|\leq 1), and also the Lipschitz continuity of truncations, that is |(sκ)+(tκ)+||st|\lvert(s-\kappa)_{+}-(t-\kappa)_{+}\rvert\leq\lvert s-t\rvert for every s,ts,t\in\mathbb{R}, we continue to estimate as follows

Bh|τh(Eϱ(x,|Duϱ|)κ)+|2𝑑x\displaystyle\int_{B_{h}}\lvert\tau_{h}(E_{\varrho}(x,\lvert Du_{\varrho}\rvert)-\kappa)_{+}\rvert^{2}\,dx
cBh|τh(Eϱ,i(|Dv|)κ)+|2𝑑x\displaystyle\quad\ \ \leq c\int_{B_{h}}\lvert\tau_{h}(E_{\varrho,\textnormal{i}}(\lvert Dv\rvert)-\kappa)_{+}\rvert^{2}\,dx
+c4Bh|Eϱ(x,|Duϱ|)Eϱ,i(|Duϱ|)|2𝑑x\displaystyle\quad\qquad+c\int_{4B_{h}}\lvert E_{\varrho}(x,\lvert Du_{\varrho}\rvert)-E_{\varrho,\textnormal{i}}(\lvert Du_{\varrho}\rvert)\rvert^{2}\,dx
+c4Bh|Eϱ,i(|Dv|)Eϱ,i(|Duϱ|)|2𝑑x\displaystyle\qquad\quad+c\int_{4B_{h}}\lvert E_{\varrho,\textnormal{i}}(\lvert Dv\rvert)-E_{\varrho,\textnormal{i}}(\lvert Du_{\varrho}\rvert)\rvert^{2}\,dx
(5.40)c|h|2(1β0)𝔪δ14Bh(Eϱ(x,|Duϱ|)κ)+2𝑑x\displaystyle\quad\stackrel{{\scriptstyle(\ref{caccia2})}}{{\leq}}c\lvert h\rvert^{2(1-\beta_{0})}\mathfrak{m}^{\delta_{1}}\int_{4B_{h}}(E_{\varrho}(x,\lvert Du_{\varrho}\rvert)-\kappa)_{+}^{2}\,dx
+c4Bh|Eϱ(x,|Duϱ|)Eϱ,i(|Duϱ|)|2𝑑x\displaystyle\quad\qquad+c\int_{4B_{h}}\lvert E_{\varrho}(x,\lvert Du_{\varrho}\rvert)-E_{\varrho,\textnormal{i}}(\lvert Du_{\varrho}\rvert)\rvert^{2}\,dx
+c𝔪δ14Bh|Eϱ,i(|Duϱ|)Eϱ,i(|Dv|)|2𝑑x\displaystyle\quad\qquad+c\mathfrak{m}^{\delta_{1}}\int_{4B_{h}}\lvert E_{\varrho,\textnormal{i}}(\lvert Du_{\varrho}\rvert)-E_{\varrho,\textnormal{i}}(\lvert Dv\rvert)\rvert^{2}\,dx

where cc(data,δ1)c\equiv c(\textnormal{{data}},\delta_{1}). Note that we have used the elementary estimate

(Eϱ,i(|Duϱ|)κ)+(Eϱ(x,|Duϱ|)κ)+on 8Bh,(E_{\varrho,\textnormal{i}}(\lvert Du_{\varrho}\rvert)-\kappa)_{+}\leq(E_{\varrho}(x,\lvert Du_{\varrho}\rvert)-\kappa)_{+}\quad\mbox{on $8B_{h}$}\,,

that follows from the very definition of Eϱ,i()E_{\varrho,\textnormal{i}}(\cdot) in (5.25)2. By then using (5.6) and (5.6) to estimate the last two integrals in the above display, respectively, and again (5.35)2, we come to

Bh|τh(Eϱ(|Duϱ|)κ)+|2𝑑x\displaystyle\int_{B_{h}}\lvert\tau_{h}(E_{\varrho}(\lvert Du_{\varrho}\rvert)-\kappa)_{+}\rvert^{2}\,dx
c~|h|2α𝔐2𝔰14Bh(Eϱ(x,|Duϱ|)κ)+2𝑑x\displaystyle\quad\leq\tilde{c}\lvert h\rvert^{2\alpha_{*}}\mathfrak{M}^{2\mathfrak{s}_{1}}\int_{4B_{h}}(E_{\varrho}(x,\lvert Du_{\varrho}\rvert)-\kappa)_{+}^{2}\,dx
+c~|h|2α𝔐2𝔰2ϱ2α8Bh(|Duϱ|+1)2(q1+δ2)𝑑x\displaystyle\quad\qquad+\tilde{c}\lvert h\rvert^{2\alpha_{*}}\mathfrak{M}^{2\mathfrak{s}_{2}}\varrho^{2\alpha}\int_{8B_{h}}(|Du_{\varrho}|+1)^{2(q-1+\delta_{2})}\,dx
+c~|h|2α𝔐2𝔰3ϱα8Bh(|Duϱ|+1)q1+δ2\displaystyle\quad\qquad+\tilde{c}\lvert h\rvert^{2\alpha_{*}}\mathfrak{M}^{2\mathfrak{s}_{3}}\varrho^{\alpha}\int_{8B_{h}}(|Du_{\varrho}|+1)^{q-1+\delta_{2}}
+c~|h|2α𝔐2𝔰3ϱα08Bh(|Duϱ|+1)3δ2𝑑x\displaystyle\quad\qquad+\tilde{c}\lvert h\rvert^{2\alpha_{*}}\mathfrak{M}^{2\mathfrak{s}_{3}}\varrho^{\alpha_{0}}\int_{8B_{h}}(|Du_{\varrho}|+1)^{3\delta_{2}}\,dx (5.43)

with c~c~(data,δ1,δ2)\tilde{c}\equiv\tilde{c}(\textnormal{{data}},\delta_{1},\delta_{2}), where 𝔰1\mathfrak{s}_{1}, 𝔰2\mathfrak{s}_{2} and 𝔰3\mathfrak{s}_{3} are as in (5.39) and we have taken β0\beta_{0} such that

β0:=2α~+2β0α~=2(1β0)=2α.\beta_{0}:=\frac{2}{\tilde{\alpha}+2}\Longleftrightarrow\beta_{0}\tilde{\alpha}=2(1-\beta_{0})=2\alpha_{*}\,.

Step 2: Patching estimates (5.43) on different balls. In this second and final step we are now going to recover estimates for τh(Eϱ(,|Duϱ|)κ)+\tau_{h}(E_{\varrho}(\cdot,\lvert Du_{\varrho}\rvert)-\kappa)_{+} on 1/2\mathcal{B}_{1/2} by patching up estimates (5.43) via a dyadic covering argument. This goes as follows: we take a lattice of cubes {Qγ}γ𝔫\{Q_{\gamma}\}_{\gamma\leq\mathfrak{n}} with sidelength equal to 2|h|β0/n2\lvert h\rvert^{\beta_{0}}/\sqrt{n}, centered at points {xγ}γ𝔫missingB1/2+2|h|β0\{x_{\gamma}\}_{\gamma\leq\mathfrak{n}}\subset\mathcal{\mathcal{missing}}B_{1/2+2|h|^{\beta_{0}}}, with sides parallel to the coordinate axes, and such that

|1/2γ𝔫Qγ|=0,Qγ1Qγ2=γ1γ2.\displaystyle\left|\ \mathcal{B}_{1/2}\setminus\bigcup_{\gamma\leq\mathfrak{n}}Q_{\gamma}\ \right|=0,\qquad Q_{\gamma_{1}}\cap Q_{\gamma_{2}}=\emptyset\ \Leftrightarrow\ \gamma_{1}\not=\gamma_{2}. (5.44)

This family of cubes corresponds to a family of balls in the sense that QγQinn(Bγ)Q_{\gamma}\equiv Q_{\textnormal{inn}}(B_{\gamma}) and Bγ:=B|h|β0(xγ)B_{\gamma}:=B_{\lvert h\rvert^{\beta_{0}}}(x_{\gamma}), as defined above. By construction, and in particular by (5.20), it is 8Bγ18B_{\gamma}\Subset\mathcal{B}_{1} for all γ𝔫\gamma\leq\mathfrak{n} and 𝔫|h|nβ0\mathfrak{n}\approx\lvert h\rvert^{-n\beta_{0}}, where the implied constant depends on nn. Moreover, by (5.44), each of the dilated balls 8Bγt8B_{\gamma_{t}} intersects the similar ones 8Bγs8B_{\gamma_{s}} fewer than 𝔠n\mathfrak{c}_{n} times, that is a number depending only on nn (uniform finite intersection property). This implies that

γ=1𝔫λ(8Bγ)𝔠nλ(1)\sum_{\gamma=1}^{\mathfrak{n}}\lambda(8B_{\gamma})\leq\mathfrak{c}_{n}\lambda(\mathcal{B}_{1}) (5.45)

holds for every Borel measure λ()\lambda(\cdot) defined on 1\mathcal{B}_{1}. We then write estimates (5.43) on balls BhBγB_{h}\equiv B_{\gamma} and sum up in order to obtain

1/2|τh(Eϱ(x,|Duϱ|)κ)+|2𝑑x\displaystyle\int_{\mathcal{B}_{1/2}}\lvert\tau_{h}(E_{\varrho}(x,\lvert Du_{\varrho}\rvert)-\kappa)_{+}\rvert^{2}\,dx
(5.44)γ=1𝔫Bγ|τh(Eϱ(x,|Duϱ|)κ)+|2𝑑x\displaystyle\quad\stackrel{{\scriptstyle\eqref{11.1}}}{{\leq}}\sum_{\gamma=1}^{\mathfrak{n}}\int_{B_{\gamma}}\lvert\tau_{h}(E_{\varrho}(x,\lvert Du_{\varrho}\rvert)-\kappa)_{+}\rvert^{2}\,dx
(5.43)c~|h|2α𝔐2𝔰1γ=1𝔫8Bγ(Eϱ(x,|Duϱ|)κ)+2𝑑x\displaystyle\quad\stackrel{{\scriptstyle(\ref{patch})}}{{\leq}}\tilde{c}\lvert h\rvert^{2\alpha_{*}}\mathfrak{M}^{2\mathfrak{s}_{1}}\sum_{\gamma=1}^{\mathfrak{n}}\int_{8B_{\gamma}}(E_{\varrho}(x,\lvert Du_{\varrho}\rvert)-\kappa)_{+}^{2}\,dx
+c~|h|2α𝔐2𝔰2ϱ2αγ=1𝔫8Bγ(|Duϱ|+1)2(q1+δ2)𝑑x\displaystyle\qquad\qquad+\tilde{c}\lvert h\rvert^{2\alpha_{*}}\mathfrak{M}^{2\mathfrak{s}_{2}}\varrho^{2\alpha}\sum_{\gamma=1}^{\mathfrak{n}}\int_{8B_{\gamma}}(|Du_{\varrho}|+1)^{2(q-1+\delta_{2})}\,dx
+c~|h|2α𝔐2𝔰3ϱαγ=1𝔫8Bγ(|Duϱ|+1)q1+δ2𝑑x\displaystyle\qquad\qquad+\tilde{c}\lvert h\rvert^{2\alpha_{*}}\mathfrak{M}^{2\mathfrak{s}_{3}}\varrho^{\alpha}\sum_{\gamma=1}^{\mathfrak{n}}\int_{8B_{\gamma}}(|Du_{\varrho}|+1)^{q-1+\delta_{2}}\,dx
+c~|h|2α𝔐2𝔰3ϱα0γ=1𝔫8Bγ(|Duϱ|+1)3δ2𝑑x\displaystyle\qquad\qquad+\tilde{c}\lvert h\rvert^{2\alpha_{*}}\mathfrak{M}^{2\mathfrak{s}_{3}}\varrho^{\alpha_{0}}\sum_{\gamma=1}^{\mathfrak{n}}\int_{8B_{\gamma}}(|Du_{\varrho}|+1)^{3\delta_{2}}\,dx
(5.45)𝔠nc~|h|2α𝔐2𝔰11(Eϱ(x,|Duϱ|)κ)+2𝑑x\displaystyle\quad\stackrel{{\scriptstyle(\ref{sommamis})}}{{\leq}}\mathfrak{c}_{n}\tilde{c}\lvert h\rvert^{2\alpha_{*}}\mathfrak{M}^{2\mathfrak{s}_{1}}\int_{\mathcal{B}_{1}}(E_{\varrho}(x,\lvert Du_{\varrho}\rvert)-\kappa)_{+}^{2}\,dx
+𝔠nc~|h|2α𝔐2𝔰2ϱ2α1(|Duϱ|+1)2(q1+δ2)𝑑x\displaystyle\qquad\qquad+\mathfrak{c}_{n}\tilde{c}\lvert h\rvert^{2\alpha_{*}}\mathfrak{M}^{2\mathfrak{s}_{2}}\varrho^{2\alpha}\int_{\mathcal{B}_{1}}(|Du_{\varrho}|+1)^{2(q-1+\delta_{2})}\,dx
+𝔠nc~|h|2α𝔐2𝔰3ϱα1(|Duϱ|+1)q1+δ2𝑑x\displaystyle\qquad\qquad+\mathfrak{c}_{n}\tilde{c}\lvert h\rvert^{2\alpha_{*}}\mathfrak{M}^{2\mathfrak{s}_{3}}\varrho^{\alpha}\int_{\mathcal{B}_{1}}(|Du_{\varrho}|+1)^{q-1+\delta_{2}}\,dx
+𝔠nc~|h|2α𝔐2𝔰3ϱα01(|Duϱ|+1)3δ2𝑑x.\displaystyle\qquad\qquad+\mathfrak{c}_{n}\tilde{c}\lvert h\rvert^{2\alpha_{*}}\mathfrak{M}^{2\mathfrak{s}_{3}}\varrho^{\alpha_{0}}\int_{\mathcal{B}_{1}}(|Du_{\varrho}|+1)^{3\delta_{2}}\,dx\,. (5.46)

Thanks to the inequality in the last display we apply Lemma 3. This yields (Eϱ(,|Duϱ|)κ)+Wβ,2(1/2)(E_{\varrho}(\cdot,\lvert Du_{\varrho}\rvert)-\kappa)_{+}\in W^{\beta,2}(\mathcal{B}_{1/2}) for all β(0,α)\beta\in(0,\alpha_{*}) with the related a priori bound

(Eϱ(,|Duϱ|)κ)+L2χ(1/2)+[(Eϱ(,|Duϱ|)κ)+]β,2;1/2\displaystyle\lVert(E_{\varrho}(\cdot,\lvert Du_{\varrho}\rvert)-\kappa)_{+}\rVert_{L^{2\chi}(\mathcal{B}_{1/2})}+[(E_{\varrho}(\cdot,\lvert Du_{\varrho}\rvert)-\kappa)_{+}]_{\beta,2;\mathcal{B}_{1/2}}
c𝔐𝔰1Eϱ(,|Duϱ|)κ)+L2(1)\displaystyle\qquad\leq c\mathfrak{M}^{\mathfrak{s}_{1}}\lVert E_{\varrho}(\cdot,\lvert Du_{\varrho}\rvert)-\kappa)_{+}\rVert_{L^{2}(\mathcal{B}_{1})}
+c𝔐𝔰2ϱα|Duϱ|+1L2(q1+δ2)(1)q1+δ2\displaystyle\qquad\quad+c\mathfrak{M}^{\mathfrak{s}_{2}}\varrho^{\alpha}\lVert|Du_{\varrho}|+1\rVert_{L^{2(q-1+\delta_{2})}(\mathcal{B}_{1})}^{q-1+\delta_{2}}
+c𝔐𝔰3ϱα/2|Duϱ|+1Lq1+δ2(1)(q1+δ2)/2\displaystyle\qquad\quad+c\mathfrak{M}^{\mathfrak{s}_{3}}\varrho^{\alpha/2}\lVert|Du_{\varrho}|+1\rVert_{L^{q-1+\delta_{2}}(\mathcal{B}_{1})}^{(q-1+\delta_{2})/2}
+c𝔐𝔰3ϱα0/2|Duϱ|+1L3δ2(1)3δ2/2\displaystyle\qquad\quad+c\mathfrak{M}^{\mathfrak{s}_{3}}\varrho^{\alpha_{0}/2}\lVert|Du_{\varrho}|+1\rVert_{L^{3\delta_{2}}(\mathcal{B}_{1})}^{3\delta_{2}/2} (5.47)

that holds for every χχ(β)\chi\equiv\chi(\beta) as in (5.37), where cc(data,δ1,δ2,β)c\equiv c(\textnormal{{data}},\delta_{1},\delta_{2},\beta). Notice that here we have used, in order, first (2.10), as a consequence of (5.46), and then (2.9). Scaling back from 1\mathcal{B}_{1} to BϱB_{\varrho} in (5.47) via (5.14), (5.19) and (5.25), squaring the resulting inequality, and restoring the original notation, we arrive at (10) and the proof is complete.

Remark 3 (Nonlinear atoms)

Here we briefly expand on the possible analogy between the classic atomic decompositions in fractional spaces (see for instance (AH, , Section 4.6) and (triebel, , Chapter 2)) and the construction made in the proof of Lemma 10. Atomic decompositions of a Besov function ww usually go via decompositions of the space n\mathbb{R}^{n} in dyadic grids with mesh 2k2^{-k}, kk\in\mathbb{N}, corresponding to the annuli (in the frequency space) considered in Littllewood-Paley theory. On each cube Qk,γQ_{k,\gamma} of the grid one considers an atom ak,γ()a_{k,\gamma}(\cdot), i.e., a smooth function with certain control on its derivatives, up to the maximal degree of regularity one is interested in describing for ww. Specifically, one requires that

suppak,γQk,γ,|DSak,γ|2|S|k\textnormal{supp}\,a_{k,\gamma}\subset Q_{k,\gamma}\,,\qquad|D_{S}a_{k,\gamma}|\lesssim 2^{-|S|k} (5.48)

hold for sufficiently large multi-indices SS. Summing up (over γ\gamma) such atoms multiplied by suitable modulating coefficients, and then yet over all possible grids kk\in\mathbb{N}, allows to give a precise description of the smoothness of the function ww. Such “linear” decompositions, although very efficient, are of little use when dealing with nonlinear problems as those considered in this paper. The idea in Lemma 10 is then, given a grid of size |h|β01/2kβ0|h|^{\beta_{0}}\approx 1/2^{k\beta_{0}}, and therefore a certain “height” in the frequency space, to consider atoms vv that are in a sense close to the original solution uu in that they are themselves solutions to nonlinear problems (with frozen coefficients). In other words we attempt a decomposition of the type

uϱ(x)γ𝔫vγ(x)𝟙Bγ(x)+o(|h|α~)u_{\varrho}(x)\approx\sum_{\gamma\leq\mathfrak{n}}v_{\gamma}(x)\mathds{1}_{B_{\gamma}}(x)+\texttt{o}(|h|^{\tilde{\alpha}}) (5.49)

where vγv_{\gamma} is defined as in (5.28), with BγBhB_{\gamma}\equiv B_{h} as in Step 2 from Lemma 10 (𝟙Bγ\mathds{1}_{B_{\gamma}} is the indicator function of the ball BγB_{\gamma}). Notice in fact the analogy with the second information in (5.48), describing the maximal smoothness of a classical atom, with the Caccioppoli inequality (5.32), from which one infers its fractional version for uϱu_{\varrho}, that is (10).

5.7 C0,1C^{0,1}-bounds via nonlinear potentials

Here we deliver

Proposition 2

Let uω,εW1,q(Br)u_{\omega,\varepsilon}\in W^{1,q}(B_{r}) be as in (5.3). The inequality

E~ω,σε(,|Duω,ε|)L(Bt)c(st)nϑHω,σε(,Duω,ε)+1L1(Bs)ϑ+c\lVert\tilde{E}_{\omega,\sigma_{\varepsilon}}(\cdot,\lvert Du_{\omega,\varepsilon}\rvert)\rVert_{L^{\infty}(B_{t})}\leq\frac{c}{(s-t)^{n\vartheta}}\lVert H_{\omega,\sigma_{\varepsilon}}(\cdot,Du_{\omega,\varepsilon})+1\rVert_{L^{1}(B_{s})}^{\vartheta}+c (5.50)

holds whenever BtBsBrB_{t}\Subset B_{s}\subseteq B_{r} are concentric balls such that r1r\leq 1, where cc(data)c\equiv c(\textnormal{{data}}) and ϑϑ(n,q,α,α0)\vartheta\equiv\vartheta(n,q,\alpha,\alpha_{0}). The function E~ω,σε()\tilde{E}_{\omega,\sigma_{\varepsilon}}(\cdot) has been introduced in (3.13).

Proof

In the following we are going to use Lemma 10 with δ1(0,1)\delta_{1}\in(0,1) whose size will determined in a few lines as a function of n,q,α,α0n,q,\alpha,\alpha_{0}, and with δ2\delta_{2} by now fixed by

0<δ2:=12min{1+αnq,α03n}<12.0<\delta_{2}:=\frac{1}{2}\min\left\{1+\frac{\alpha}{n}-q,\frac{\alpha_{0}}{3n}\right\}<\frac{1}{2}\,. (5.51)

Note that δ2>0\delta_{2}>0 follows as a consequence of the assumed bound (1.17) and that the choice in (5.51) makes δ2\delta_{2} depending only on n,q,α,α0n,q,\alpha,\alpha_{0}. Without loss of generality we can assume that

E~ω,σε(,|Duω,ε|)L(Bt)1\lVert\tilde{E}_{\omega,\sigma_{\varepsilon}}(\cdot,\lvert Du_{\omega,\varepsilon}\rvert)\rVert_{L^{\infty}(B_{t})}\geq 1 (5.52)

otherwise (5.50) is obvious. We consider concentric balls BtBτ1Bτ2BsB_{t}\Subset B_{\tau_{1}}\Subset B_{\tau_{2}}\Subset B_{s} and set r0:=(τ2τ1)/8r_{0}:=(\tau_{2}-\tau_{1})/8. It follows that B2r0(x0)Bτ2B_{2r_{0}}(x_{0})\Subset B_{\tau_{2}} holds whenever x0Bτ1x_{0}\in B_{\tau_{1}}. Notice that by (5.5) every point is a Lebesgue point for both |Duω,ε|\lvert Du_{\omega,\varepsilon}\rvert and Eω,σε(,|Duω,ε|)E_{\omega,\sigma_{\varepsilon}}(\cdot,\lvert Du_{\omega,\varepsilon}\rvert). By Lemma 10, and (10) used on Bϱ(x0)Br0(x0)B_{\varrho}(x_{0})\subset B_{r_{0}}(x_{0}), we can apply Lemma 5 on Br0(x0)B_{r_{0}}(x_{0}) verifying (2.14) with

{wEω,σε(,|Duω,ε|)𝔐:=2E~ω,σε(,|Duω,ε|)L(Bτ2)(such a choice of 𝔐 is admissible in (10) by (5.52))M0:=𝔐𝔰1,M1𝔐𝔰2,M2M3:=𝔐𝔰3χ:=n/(n2β)>1,β:=α/2(recall (5.37))σ1:=α,σ2:=α/2,σ3:=α0/2,f1f2f3:=|Duω,ε|+1,θ1θ2θ3:=1,cc(data,δ1,β)c(data,δ1),m1:=2(q1+δ2),m2:=q1+δ2,m3:=3δ2,κ0:=0.\begin{cases}\ w\equiv E_{\omega,\sigma_{\varepsilon}}(\cdot,\lvert Du_{\omega,\varepsilon}\rvert)\\ \ \mathfrak{M}:=2\lVert\tilde{E}_{\omega,\sigma_{\varepsilon}}(\cdot,\lvert Du_{\omega,\varepsilon}\rvert)\rVert_{L^{\infty}(B_{\tau_{2}})}\\ \quad\mbox{(such a choice of $\mathfrak{M}$ is admissible in (\ref{16}) by (\ref{ammissibile}))}\\ \ M_{0}:=\mathfrak{M}^{\mathfrak{s}_{1}},\ M_{1}\equiv\mathfrak{M}^{\mathfrak{s}_{2}},\ M_{2}\equiv M_{3}:=\mathfrak{M}^{\mathfrak{s}_{3}}\\ \ \chi:=n/(n-2\beta)>1,\quad\beta:=\alpha_{*}/2\quad\mbox{(recall (\ref{ilbeta2}))}\\ \ \sigma_{1}:=\alpha,\ \sigma_{2}:=\alpha/2,\ \sigma_{3}:=\alpha_{0}/2,\ \ f_{1}\equiv f_{2}\equiv f_{3}:=\lvert Du_{\omega,\varepsilon}\rvert+1,\\ \ \theta_{1}\equiv\theta_{2}\equiv\theta_{3}:=1,\ c_{*}\equiv c_{*}(\textnormal{{data}},\delta_{1},\beta)\equiv c_{*}(\textnormal{{data}},\delta_{1}),\\ \ m_{1}:=2(q-1+\delta_{2}),\ m_{2}:=q-1+\delta_{2},\ m_{3}:=3\delta_{2},\ \kappa_{0}:=0\,.\end{cases} (5.53)

As x0Bτ1x_{0}\in B_{\tau_{1}} has been chosen arbitrarily, (5) with the choices in (5.53) implies

Eω,σε(,|Duω,ε|)L(Bτ1)\displaystyle\lVert E_{\omega,\sigma_{\varepsilon}}(\cdot,\lvert Du_{\omega,\varepsilon}\rvert)\rVert_{L^{\infty}(B_{\tau_{1}})}
c𝔐𝔰1χχ1(τ2τ1)n/2(Bτ2[Eω,σε(x,|Duω,ε|)]2𝑑x)1/2\displaystyle\qquad\leq\frac{c\mathfrak{M}^{\frac{\mathfrak{s}_{1}\chi}{\chi-1}}}{(\tau_{2}-\tau_{1})^{n/2}}\left(\int_{B_{\tau_{2}}}[E_{\omega,\sigma_{\varepsilon}}(x,\lvert Du_{\omega,\varepsilon}\rvert)]^{2}\,dx\right)^{1/2}
+c𝔐𝔰1χ1+𝔰2𝐏2,α2(q1+δ2),1(|Duω,ε|+1;,(τ2τ1)/4)L(Bτ1)\displaystyle\quad\qquad+c\mathfrak{M}^{\frac{\mathfrak{s}_{1}}{\chi-1}+\mathfrak{s}_{2}}\lVert\mathbf{P}^{2(q-1+\delta_{2}),1}_{2,\alpha}(\lvert Du_{\omega,\varepsilon}\rvert+1;\cdot,(\tau_{2}-\tau_{1})/4)\rVert_{L^{\infty}(B_{\tau_{1}})}
+c𝔐𝔰1χ1+𝔰3𝐏2,α/2q1+δ2,1(|Duω,ε|+1;,(τ2τ1)/4)L(Bτ1)\displaystyle\quad\qquad+c\mathfrak{M}^{\frac{\mathfrak{s}_{1}}{\chi-1}+\mathfrak{s}_{3}}\lVert\mathbf{P}^{q-1+\delta_{2},1}_{2,\alpha/2}(\lvert Du_{\omega,\varepsilon}\rvert+1;\cdot,(\tau_{2}-\tau_{1})/4)\rVert_{L^{\infty}(B_{\tau_{1}})}
+c𝔐𝔰1χ1+𝔰3𝐏2,α0/23δ2,1(|Duω,ε|+1;,(τ2τ1)/4)L(Bτ1),\displaystyle\quad\qquad+c\mathfrak{M}^{\frac{\mathfrak{s}_{1}}{\chi-1}+\mathfrak{s}_{3}}\lVert\mathbf{P}^{3\delta_{2},1}_{2,\alpha_{0}/2}(\lvert Du_{\omega,\varepsilon}\rvert+1;\cdot,(\tau_{2}-\tau_{1})/4)\rVert_{L^{\infty}(B_{\tau_{1}})}, (5.54)

for cc(data,δ1)c\equiv c(\textnormal{{data}},\delta_{1}). Recalling the definitions in (3.12)-(3.13), and the choice of 𝔐\mathfrak{M} in (5.53), after a few elementary manipulations in (5.7) we arrive at

E~ω,σε(,|Duω,ε|)L(Bτ1)\displaystyle\lVert\tilde{E}_{\omega,\sigma_{\varepsilon}}(\cdot,\lvert Du_{\omega,\varepsilon}\rvert)\rVert_{L^{\infty}(B_{\tau_{1}})}
c(τ2τ1)n/2E~ω,σε(,|Duω,ε|)L(Bτ2)𝔰1χχ1+12(BsE~ω,σε(x,|Duω,ε|)𝑑x)12\displaystyle\leq\frac{c}{(\tau_{2}-\tau_{1})^{n/2}}\lVert\tilde{E}_{\omega,\sigma_{\varepsilon}}(\cdot,\lvert Du_{\omega,\varepsilon}\rvert)\rVert_{L^{\infty}(B_{\tau_{2}})}^{\frac{\mathfrak{s}_{1}\chi}{\chi-1}+\frac{1}{2}}\left(\int_{B_{s}}\tilde{E}_{\omega,\sigma_{\varepsilon}}(x,\lvert Du_{\omega,\varepsilon}\rvert)\,dx\right)^{\frac{1}{2}}
+cE~ω,σε(,|Duω,ε|)L(Bτ2)𝔰1χ1+𝔰2\displaystyle\quad+c\lVert\tilde{E}_{\omega,\sigma_{\varepsilon}}(\cdot,\lvert Du_{\omega,\varepsilon}\rvert)\rVert_{L^{\infty}(B_{\tau_{2}})}^{\frac{\mathfrak{s}_{1}}{\chi-1}+\mathfrak{s}_{2}}
𝐏2,α2(q1+δ2),1(|Duω,ε|+1;,(τ2τ1)/4)L(Bτ1)\displaystyle\qquad\quad\cdot\lVert\mathbf{P}^{2(q-1+\delta_{2}),1}_{2,\alpha}(\lvert Du_{\omega,\varepsilon}\rvert+1;\cdot,(\tau_{2}-\tau_{1})/4)\rVert_{L^{\infty}(B_{\tau_{1}})}
+cE~ω,σε(,|Duω,ε|)L(Bτ2)𝔰1χ1+𝔰3\displaystyle\quad+c\lVert\tilde{E}_{\omega,\sigma_{\varepsilon}}(\cdot,\lvert Du_{\omega,\varepsilon}\rvert)\rVert_{L^{\infty}(B_{\tau_{2}})}^{\frac{\mathfrak{s}_{1}}{\chi-1}+\mathfrak{s}_{3}}
𝐏2,α/2q1+δ2,1(|Duω,ε|+1;,(τ2τ1)/4)L(Bτ1)\displaystyle\qquad\quad\cdot\lVert\mathbf{P}^{q-1+\delta_{2},1}_{2,\alpha/2}(\lvert Du_{\omega,\varepsilon}\rvert+1;\cdot,(\tau_{2}-\tau_{1})/4)\rVert_{L^{\infty}(B_{\tau_{1}})}
+cE~ω,σε(,|Duω,ε|)L(Bτ2)𝔰1χ1+𝔰3\displaystyle\quad+c\lVert\tilde{E}_{\omega,\sigma_{\varepsilon}}(\cdot,\lvert Du_{\omega,\varepsilon}\rvert)\rVert_{L^{\infty}(B_{\tau_{2}})}^{\frac{\mathfrak{s}_{1}}{\chi-1}+\mathfrak{s}_{3}}
𝐏2,α0/23δ2,1(|Duω,ε|+1;,(τ2τ1)/4)L(Bτ1)+c\displaystyle\qquad\quad\cdot\lVert\mathbf{P}^{3\delta_{2},1}_{2,\alpha_{0}/2}(\lvert Du_{\omega,\varepsilon}\rvert+1;\cdot,(\tau_{2}-\tau_{1})/4)\rVert_{L^{\infty}(B_{\tau_{1}})}+c (5.55)

with cc(data,δ1)c\equiv c(\textnormal{{data}},\delta_{1}). We now take δ1\delta_{1} to be such that

{𝔰1(1,δ1)χχ1+12=nδ14β+12<1𝔰1(1,δ1)χ1+𝔰2(1,δ2)=(n2β)δ14β+1δ2<1𝔰1(1,δ1)χ1+𝔰3(1,δ1,δ2)=(n2β)δ14β+1+(q+1)δ1δ2/22<1.\begin{cases}\displaystyle\frac{\mathfrak{s}_{1}(1,\delta_{1})\chi}{\chi-1}+\frac{1}{2}=\frac{n\delta_{1}}{4\beta}+\frac{1}{2}<1\\ \displaystyle\frac{\mathfrak{s}_{1}(1,\delta_{1})}{\chi-1}+\mathfrak{s}_{2}(1,\delta_{2})=\frac{(n-2\beta)\delta_{1}}{4\beta}+1-\delta_{2}<1\\ \displaystyle\frac{\mathfrak{s}_{1}(1,\delta_{1})}{\chi-1}+\mathfrak{s}_{3}(1,\delta_{1},\delta_{2})=\frac{(n-2\beta)\delta_{1}}{4\beta}+1+\frac{(q+1)\delta_{1}-\delta_{2}/2}{2}<1\,.\end{cases} (5.56)

Recalling (5.51), the quantity δ1\delta_{1} is now determined as a function of the parameters n,q,α,α0n,q,\alpha,\alpha_{0}. This determines a first restriction on the size of μ\mu via (5.38) and the choice of δ1\delta_{1}. Notice that at this stage the number μ:=μn(δ1/2)\mu_{*}:=\mu_{n}(\delta_{1}/2), determining an upper bound on μ\mu, depends on n,q,α,α0n,q,\alpha,\alpha_{0}; notice also that all the above computation remain valid provided 1μ<μμ(n,q,α,α0)1\leq\mu<\mu_{*}\equiv\mu_{*}(n,q,\alpha,\alpha_{0}). Finally, by (5.56) and a standard continuity argument we can further restrict the value of μ\mu finding μmaxμ\mu_{\textnormal{max}}\leq\mu_{*} such that

1μ<μmax{𝔰1(μ,δ1)χχ1+12<1𝔰1(μ,δ1)χ1+𝔰2(μ,δ2)<1𝔰1(μ,δ1)χ1+𝔰3(μ,δ1,δ2,)<1.1\leq\mu<\mu_{\textnormal{max}}\Longrightarrow\begin{cases}\displaystyle\frac{\mathfrak{s}_{1}(\mu,\delta_{1})\chi}{\chi-1}+\frac{1}{2}<1\\ \displaystyle\frac{\mathfrak{s}_{1}(\mu,\delta_{1})}{\chi-1}+\mathfrak{s}_{2}(\mu,\delta_{2})<1\\ \displaystyle\frac{\mathfrak{s}_{1}(\mu,\delta_{1})}{\chi-1}+\mathfrak{s}_{3}(\mu,\delta_{1},\delta_{2},)<1\,.\end{cases} (5.57)

Notice also that the functions μ𝔰1(μ,),𝔰2(μ,),𝔰3(μ,)\mu\mapsto\mathfrak{s}_{1}(\mu,\cdot),\mathfrak{s}_{2}(\mu,\cdot),\mathfrak{s}_{3}(\mu,\cdot) are increasing. This finally fixes the value of μmax\mu_{\textnormal{max}} from the statement of Theorem 1.2, with the asserted dependence on the constants. We now want to the estimate the potential terms appearing in the right-hand side of (5.55) by means of Lemma 4. In this respect, notice that n>2αn>2\alpha allows to verify (2.12), while we can take γ=1\gamma=1 in (2.13) because

(5.51)1>max{n(q1+δ2)α,3nδ2α0}.(\ref{ildelta})\Longrightarrow 1>\max\left\{\frac{n(q-1+\delta_{2})}{\alpha},\frac{3n\delta_{2}}{\alpha_{0}}\right\}\,.

Applying (2.13) gives

{𝐏2,α/2q1+δ2,1(|Duω,ε|+1;,(τ2τ1)/4)L(Bτ1)c|Duω,ε|+1L1(Bs)(q1+δ2)/2𝐏2,α2(q1+δ2),1(|Duω,ε|+1;,(τ2τ1)/4)L(Bτ1)c|Duω,ε|+1L1(Bs)q1+δ2𝐏2,α0/23δ2,1(|Duω,ε|+1;,(τ2τ1)/4)L(Bτ1)c|Duω,ε|+1L1(Bs)3δ2/2,\displaystyle\begin{cases}\ \lVert\mathbf{P}^{q-1+\delta_{2},1}_{2,\alpha/2}(\lvert Du_{\omega,\varepsilon}\rvert+1;\cdot,(\tau_{2}-\tau_{1})/4)\rVert_{L^{\infty}(B_{\tau_{1}})}\\ \qquad\leq c\lVert\lvert Du_{\omega,\varepsilon}\rvert+1\rVert_{L^{1}(B_{s})}^{(q-1+\delta_{2})/2}\\ \ \lVert\mathbf{P}^{2(q-1+\delta_{2}),1}_{2,\alpha}(\lvert Du_{\omega,\varepsilon}\rvert+1;\cdot,(\tau_{2}-\tau_{1})/4)\rVert_{L^{\infty}(B_{\tau_{1}})}\\ \qquad\leq c\lVert\lvert Du_{\omega,\varepsilon}\rvert+1\rVert_{L^{1}(B_{s})}^{q-1+\delta_{2}}\\ \ \lVert\mathbf{P}^{3\delta_{2},1}_{2,\alpha_{0}/2}(\lvert Du_{\omega,\varepsilon}\rvert+1;\cdot,(\tau_{2}-\tau_{1})/4)\rVert_{L^{\infty}(B_{\tau_{1}})}\\ \qquad\leq c\lVert\lvert Du_{\omega,\varepsilon}\rvert+1\rVert_{L^{1}(B_{s})}^{3\delta_{2}/2},\end{cases} (5.58)

with cc(n,q,α,α0)c\equiv c(n,q,\alpha,\alpha_{0}). Using (5.58) in (5.55), and recalling (5.57), we can use Young’s inequality to finally get

E~ω,σε(,|Duω,ε|)L(Bτ1)12E~ω,σε(,|Duω,ε|)L(Bτ2)\displaystyle\lVert\tilde{E}_{\omega,\sigma_{\varepsilon}}(\cdot,\lvert Du_{\omega,\varepsilon}\rvert)\rVert_{L^{\infty}(B_{\tau_{1}})}\leq\frac{1}{2}\lVert\tilde{E}_{\omega,\sigma_{\varepsilon}}(\cdot,\lvert Du_{\omega,\varepsilon}\rvert)\rVert_{L^{\infty}(B_{\tau_{2}})}
+c(τ2τ1)nϑE~ω,σε(,|Duω,ε|)L1(Bs)ϑ+Duω,εL1(Bs)ϑ+c\displaystyle\qquad\qquad+\frac{c}{(\tau_{2}-\tau_{1})^{n\vartheta}}\lVert\tilde{E}_{\omega,\sigma_{\varepsilon}}(\cdot,\lvert Du_{\omega,\varepsilon}\rvert)\rVert^{\vartheta}_{L^{1}(B_{s})}+\lVert Du_{\omega,\varepsilon}\rVert_{L^{1}(B_{s})}^{\vartheta}+c

with cc(data)c\equiv c(\textnormal{{data}}) and ϑϑ(n,q,α,α0)1\vartheta\equiv\vartheta(n,q,\alpha,\alpha_{0})\geq 1. This allows to apply Lemma 2 that provides

E~ω,σε(,|Duω,ε|)L(Bt)\displaystyle\lVert\tilde{E}_{\omega,\sigma_{\varepsilon}}(\cdot,\lvert Du_{\omega,\varepsilon}\rvert)\rVert_{L^{\infty}(B_{t})}
c(st)nϑE~ω,σε(,|Duω,ε|)L1(Bs)ϑ+cDuω,εL1(Bs)ϑ+c\displaystyle\qquad\leq\frac{c}{(s-t)^{n\vartheta}}\lVert\tilde{E}_{\omega,\sigma_{\varepsilon}}(\cdot,\lvert Du_{\omega,\varepsilon}\rvert)\rVert^{\vartheta}_{L^{1}(B_{s})}+c\lVert Du_{\omega,\varepsilon}\rVert_{L^{1}(B_{s})}^{\vartheta}+c

from which (5.50) follows using (3.17) and a few elementary manipulations.

5.8 Proof of (1.14)

Keeping in mind the notation of Proposition 2, (5.8) and (5.50) give

E~ω,σε(,|Duω,ε|)L(Bt)\displaystyle\lVert\tilde{E}_{\omega,\sigma_{\varepsilon}}(\cdot,\lvert Du_{\omega,\varepsilon}\rvert)\rVert_{L^{\infty}(B_{t})}
c(rt)nϑ[𝒩(u,Br)+oε(ω)+o(ε)+|Br|]ϑ+c,\displaystyle\qquad\leq\frac{c}{(r-t)^{n\vartheta}}\left[\mathcal{N}(u,B_{r})+\texttt{o}_{\varepsilon}(\omega)+\texttt{o}(\varepsilon)+|B_{r}|\right]^{\vartheta}+c\,,

so that, recalling (3.13) we find

Duω,εL(Bt)2μ\displaystyle\lVert Du_{\omega,\varepsilon}\rVert_{L^{\infty}(B_{t})}^{2-\mu} c(rt)nϑ[𝒩(u,Br)+oε(ω)+o(ε)+|Br|]ϑ+c\displaystyle\leq\frac{c}{(r-t)^{n\vartheta}}\left[\mathcal{N}(u,B_{r})+\texttt{o}_{\varepsilon}(\omega)+\texttt{o}(\varepsilon)+|B_{r}|\right]^{\vartheta}+c
=:[𝔪ω,ε(t;Br)]2μ.\displaystyle=:[\mathfrak{m}_{\omega,\varepsilon}(t;B_{r})]^{2-\mu}\,. (5.59)

It follows that, up to not relabelled subsequences, the convergence in (5.10) can be upgraded to uω,εuεu_{\omega,\varepsilon}\rightharpoonup^{*}u_{\varepsilon} in W1,(Bt)W^{1,\infty}(B_{t}) for every ε>0\varepsilon>0. Letting ω0\omega\to 0 in (5.59) yields

DuεL(Bt)2μ\displaystyle\lVert Du_{\varepsilon}\rVert_{L^{\infty}(B_{t})}^{2-\mu} c(rt)nϑ[𝒩(u,Br)+o(ε)+|Br|]ϑ+c\displaystyle\leq\frac{c}{(r-t)^{n\vartheta}}\left[\mathcal{N}(u,B_{r})+\texttt{o}(\varepsilon)+|B_{r}|\right]^{\vartheta}+c
=:[𝔪ε(t;Br)]2μ,\displaystyle=:[\mathfrak{m}_{\varepsilon}(t;B_{r})]^{2-\mu}, (5.60)

that again holds for every ε>0\varepsilon>0, with cc(data)c\equiv c(\textnormal{{data}}) and ϑϑ(n,q,α,α0)\vartheta\equiv\vartheta(n,q,\alpha,\alpha_{0}). Similarly, as after (5.59), by (5.60) the convergence in (5.12) can be upgraded to uεuu_{\varepsilon}\rightharpoonup^{*}u in W1,(Bt)W^{1,\infty}(B_{t}), so that (1.14) (for s=0\textnormal{{s}}=0) follows letting ε0\varepsilon\to 0 in (5.60), taking t=r/2t=r/2, recalling that μ<3/2\mu<3/2 and renaming 2ϑ2\vartheta into ϑ\vartheta.

5.9 Local gradient Hölder continuity

Since the result is local, up passing to smaller open subsets we can assume with no loss of generality that 𝒩(u,Ω)\mathcal{N}(u,\Omega) is finite (see the comment after (1.15)). We fix an open subset Ω0Ω\Omega_{0}\Subset\Omega and the radius r:=min{dist(Ω0,Ω)/8,1}r:=\min\{\,{\rm dist}(\Omega_{0},\partial\Omega)/8,1\}. We take B2rB_{2r} centred in Ω0\Omega_{0} and recall that the quantities 𝔪ω,ε𝔪ω,ε(r;B2r)\mathfrak{m}_{\omega,\varepsilon}\equiv\mathfrak{m}_{\omega,\varepsilon}(r;B_{2r}) and 𝔪ε𝔪ε(r;B2r)\mathfrak{m}_{\varepsilon}\equiv\mathfrak{m}_{\varepsilon}(r;B_{2r}) are defined in (5.59) and (5.60), respectively. With τr\tau\leq r and BτBrB_{\tau}\subset B_{r} being a ball concentric to B2rB_{2r}, we define vuω,ε+W01,q(Bτ)v\in u_{\omega,\varepsilon}+W^{1,q}_{0}(B_{\tau}) as

vminwuω,ε+W01,q(Bτ)BτHω,σε,i(Dw;Bτ)𝑑xv\mapsto\min_{w\in u_{\omega,\varepsilon}+W^{1,q}_{0}(B_{\tau})}\int_{B_{\tau}}H_{\omega,\sigma_{\varepsilon},\textnormal{i}}(Dw;B_{\tau})\,dx

so that

Duω,εL(Bτ)+DvL(Bτ/2)c¯𝔪ω,ε2\lVert Du_{\omega,\varepsilon}\rVert_{L^{\infty}(B_{\tau})}+\lVert Dv\rVert_{L^{\infty}(B_{\tau/2})}\leq\bar{c}\mathfrak{m}_{\omega,\varepsilon}^{2} (5.61)

holds with c¯c¯(data)\bar{c}\equiv\bar{c}(\textnormal{{data}}). The derivation of (5.61) follows using first (5.59) (with BrB_{r} replaced by B2rB_{2r} and t=rt=r) and combining it with (4.5) and easy manipulations as in Proposition 1 and Lemma 8. By Proposition 3 in the subsequent step and (5.61), we find that

Bϱ|Dv(Dv)Bϱ|2dxcω,ε(ϱ/τ)2βω,ε\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B_{\varrho}}\lvert Dv-(Dv)_{B_{\varrho}}\rvert^{2}\,dx\leq c_{\omega,\varepsilon}\left(\varrho/\tau\right)^{2\beta_{\omega,\varepsilon}} (5.62)

holds for every ϱτ/2\varrho\leq\tau/2. Here, in the notation of Proposition 3 below, it is cω,ε:=c(2c¯𝔪ω,ε2)c_{\omega,\varepsilon}:=c_{*}(2\bar{c}\mathfrak{m}_{\omega,\varepsilon}^{2}) and βω,ε:=β(2c¯𝔪ω,ε2)\beta_{\omega,\varepsilon}:=\beta_{*}(2\bar{c}\mathfrak{m}_{\omega,\varepsilon}^{2}); these constants are non-decreasing and non-increasing functions of 𝔪ω,ε\mathfrak{m}_{\omega,\varepsilon}, respectively. Letting first ω0\omega\to 0 and then ε0\varepsilon\to 0, we have that

𝔪ω,ε2μ𝔪ε2μ𝔪2μ:=crnϑ[𝒩(u,B2r)+|B2r|]ϑ+c\mathfrak{m}_{\omega,\varepsilon}^{2-\mu}\to\mathfrak{m}_{\varepsilon}^{2-\mu}\to\mathfrak{m}^{2-\mu}:=cr^{-n\vartheta}[\mathcal{N}(u,B_{2r})+|B_{2r}|]^{\vartheta}+c

(see (5.60)). In particular

𝔪2μcrnϑ[𝒩(u,Ω)+1]+c=:𝔪02μ,\mathfrak{m}^{2-\mu}\leq cr^{-n\vartheta}[\mathcal{N}(u,\Omega)+1]+c=:\mathfrak{m}_{0}^{2-\mu}\,, (5.63)

holds with cc(data)c\equiv c(\textnormal{{data}}) and ϑϑ(n,q,α,α0)\vartheta\equiv\vartheta(n,q,\alpha,\alpha_{0}) so that 𝔪0\mathfrak{m}_{0} only depends on data,dist(Ω0,Ω)\textnormal{{data}},\,{\rm dist}(\Omega_{0},\partial\Omega) and 𝒩(u,Ω)\mathcal{N}(u,\Omega). Again letting ω0\omega\to 0 and then ε0\varepsilon\to 0, and passing to not relabelled subsequences, we can assume that cω,ε:=cε=:c¯c(4c¯𝔪2)<c_{\omega,\varepsilon}\to:=c_{\varepsilon}\to=:\underline{c}\leq c_{*}(4\bar{c}\mathfrak{m}^{2})<\infty. Notice that by (5.63) we have c¯c(4c¯𝔪02)\underline{c}\leq c_{*}(4\bar{c}\mathfrak{m}_{0}^{2}) and this last quantity depends only on data,dist(Ω0,Ω)\textnormal{{data}},\,{\rm dist}(\Omega_{0},\partial\Omega) and 𝒩(u,Ω)\mathcal{N}(u,\Omega) but it is otherwise independent of the chosen ball B2rB_{2r}. In the following, with some abuse of notation, we keep on denoting by cω,ε1c_{\omega,\varepsilon}\geq 1 a double-sequence of constants with the above property, typically being itself released via a non-decreasing function of c(2c¯𝔪ω,ε2)c_{*}(2\bar{c}\mathfrak{m}_{\omega,\varepsilon}^{2}); the exact value of the numbers cω,εc_{\omega,\varepsilon} may vary from line to line. A similar reasoning can be done for the exponents βω,ε\beta_{\omega,\varepsilon}, that is, we have βω,ε:=βε=:ββ(4c¯𝔪2)β(4c¯𝔪02)=:β~(0,1)\beta_{\omega,\varepsilon}\to:=\beta_{\varepsilon}\to=:\beta\geq\beta_{*}(4\bar{c}\mathfrak{m}^{2})\geq\beta_{*}(4\bar{c}\mathfrak{m}_{0}^{2})=:\tilde{\beta}\in(0,1). Similarly, β~\tilde{\beta} depends only data,dist(Ω0,Ω)\textnormal{{data}},\,{\rm dist}(\Omega_{0},\partial\Omega) and 𝒩(u,Ω)\mathcal{N}(u,\Omega) and is independent of the ball B2rB_{2r} considered. Proceeding as for (9), and also using (5.61) repeatedly, it follows that

Bτ|V1,2μ(Duω,ε)V1,2μ(Dv)|2dxcω,ετα~,\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B_{\tau}}\lvert V_{1,2-\mu}(Du_{\omega,\varepsilon})-V_{1,2-\mu}(Dv)\rvert^{2}\,dx\leq c_{\omega,\varepsilon}\tau^{\tilde{\alpha}}\,, (5.64)

where α~=min{α,α0}\tilde{\alpha}=\min\{\alpha,\alpha_{0}\}. Using (2.4) with p=2μp=2-\mu and then (5.61) yields

Bτ/2|Duω,εDv|2dx\displaystyle\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B_{\tau/2}}\lvert Du_{\omega,\varepsilon}-Dv\rvert^{2}\,dx
cBτ/2(|Duω,ε|2+|Dv|2+1)μ/2|V1,2μ(Duω,ε)V1,2μ(Dv)|2dx\displaystyle\quad\leq c\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B_{\tau/2}}(\lvert Du_{\omega,\varepsilon}\rvert^{2}+\lvert Dv\rvert^{2}+1)^{\mu/2}\lvert V_{1,2-\mu}(Du_{\omega,\varepsilon})-V_{1,2-\mu}(Dv)\rvert^{2}\,dx
c𝔪ω,ε2μBτ|V1,2μ(Duω,ε)V1,2μ(Dv)|2dx.\displaystyle\quad\leq c\mathfrak{m}_{\omega,\varepsilon}^{2\mu}\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B_{\tau}}\lvert V_{1,2-\mu}(Du_{\omega,\varepsilon})-V_{1,2-\mu}(Dv)\rvert^{2}\,dx\,.

By again using (5.64) in the above display we conclude with

Bτ/2|Duω,εDv|2dxcω,ετα~.\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B_{\tau/2}}\lvert Du_{\omega,\varepsilon}-Dv\rvert^{2}\,dx\leq c_{\omega,\varepsilon}\tau^{\tilde{\alpha}}\,. (5.65)

We can now complete the proof with more standard arguments (see for instance gg2 ; manth1 ; manth2 ), that we recall for completeness. Combining (5.62) and (5.65) yields

Bϱ|Duω,ε(Duω,ε)Bϱ|2dxcω,ε(ϱ/τ)2βω,ε+cω,ε(τ/ϱ)nτα~\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B_{\varrho}}\lvert Du_{\omega,\varepsilon}-(Du_{\omega,\varepsilon})_{B_{\varrho}}\rvert^{2}\,dx\leq c_{\omega,\varepsilon}(\varrho/\tau)^{2\beta_{\omega,\varepsilon}}+c_{\omega,\varepsilon}(\tau/\varrho)^{n}\tau^{\tilde{\alpha}}

for every ϱτ/2\varrho\leq\tau/2, and taking ϱ=(τ/2)1+α~/(n+2βω,ε)\varrho=(\tau/2)^{1+\tilde{\alpha}/(n+2\beta_{\omega,\varepsilon})} finally yields

Bϱ|Duω,ε(Duω,ε)Bϱ|2dxcω,εϱ2α~βω,εn+2βω,ε+α~\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B_{\varrho}}\lvert Du_{\omega,\varepsilon}-(Du_{\omega,\varepsilon})_{B_{\varrho}}\rvert^{2}\,dx\leq c_{\omega,\varepsilon}\varrho^{\frac{2\tilde{\alpha}\beta_{\omega,\varepsilon}}{n+2\beta_{\omega,\varepsilon}+\tilde{\alpha}}}

for every ϱ(r/2)1+α~/(n+2βω,ε)\varrho\leq(r/2)^{1+\tilde{\alpha}/(n+2\beta_{\omega,\varepsilon})}. In the above display we first let ω0\omega\to 0 and then ε0\varepsilon\to 0, and finally conclude with

Bϱ|Du(Du)Bϱ|2dxcϱ2α~β~n+2β~+α~,\mathop{\int\hskip-10.50005pt-\,\!\!\!}\nolimits_{B_{\varrho}}\lvert Du-(Du)_{B_{\varrho}}\rvert^{2}\,dx\leq c\varrho^{\frac{2\tilde{\alpha}\tilde{\beta}}{n+2\tilde{\beta}+\tilde{\alpha}}}\,, (5.66)

where, by the discussion made after (5.63), both c1c\geq 1 and β~(0,1)\tilde{\beta}\in(0,1) depend only on data, dist(Ω0,Ω)\,{\rm dist}(\Omega_{0},\partial\Omega) and 𝒩(u,Ω)\mathcal{N}(u,\Omega), but are otherwise independent of the ball considered B2rB_{2r}. Since Ω0\Omega_{0} is arbitrary, (5.66) and a standard covering argument and Campanato-Meyers integral characterization of Hölder continuity yield that for every open subset Ω0Ω\Omega_{0}\Subset\Omega it holds that [Du]0,β;Ω0<[Du]_{0,\beta;\Omega_{0}}<\infty, where β:=α~β~/(n+2β~+α~)\beta:=\tilde{\alpha}\tilde{\beta}/(n+2\tilde{\beta}+\tilde{\alpha}). The proof of Theorem 1.2 in the case s=0\textnormal{{s}}=0 is therefore complete up to (5.62), whose proof will be given in the subsequent section. We now briefly comment on how to obtain the local gradient Hölder continuity in the nonsingular case s>0\textnormal{{s}}>0, which is simpler since it requires no approximation via the additional parameter ω\omega. For the proof of (1.14) it is sufficient to use functionals 𝒩s,ε\mathcal{N}_{\textnormal{{s}},\varepsilon} as in (5.4), and all the subsequent estimates remain independent of s and ε\varepsilon; the approximation and the convergence then only take place with respect to the parameter ε\varepsilon and leads to (1.14). The same reasoning applies to the proof in this section by taking ωs\omega\equiv\textnormal{{s}}. It remains to deal with the last issue, that is the local Hölder continuity of DuDu with explicit exponent when s>0\textnormal{{s}}>0. This will be done in Section 5.11.

5.10 A technical decay estimate.

Here we prove (5.62); this relies on certain hidden facts from regularity theory of singular parabolic equations that we take as a starting point to treat the nonuniformly elliptic case considered here. Relevant related methods are also in dibe . Consistently with the notation established in Sections 3 and 4, with a¯>0\bar{a}>0 being a fixed number, in the following we denote

Hi(z):=F(xc,z)+a¯(|z|2+ω2)q/2H_{\textnormal{i}}(z):=F(x_{\rm c},z)+\bar{a}(\lvert z\rvert^{2}+\omega^{2})^{q/2} (5.67)

for every znz\in\mathbb{R}^{n}, where ω(0,1]\omega\in(0,1] and xcΩx_{\rm c}\in\Omega. Here we permanently assume (1.11)-(1.13) and therefore Hi()H_{\textnormal{i}}(\cdot) is an integrand of the type in (3.2)2 with aσ,i(B)a_{\sigma,\textnormal{i}}(B) replaced by a¯\bar{a}. Indeed, conditions (3.8) apply here when accordingly recasted (see also (5.78) below). Moreover, let us define

g~(t):={tif 0t1g(t)if t>1.\tilde{g}(t):=\begin{cases}t&\mbox{if $0\leq t\leq 1$}\\ g(t)&\mbox{if $t>1$}\,.\end{cases} (5.68)

Then (1.13) and Lemma 1 imply that

|zF(x,z)|cg~(|z|)|zHi(z)|cg~(|z|)+ca¯(|z|2+ω2)(q2)/2|z||\partial_{z}F(x,z)|\leq c\tilde{g}(|z|)\Longrightarrow|\partial_{z}H_{\textnormal{i}}(z)|\leq c\tilde{g}(|z|)+c\bar{a}(\lvert z\rvert^{2}+\omega^{2})^{(q-2)/2}|z| (5.69)

holds for every znz\in\mathbb{R}^{n}, where cc(q,L,cg(1))c\equiv c(q,L,c_{g}(1)).

Proposition 3

Under assumptions (1.11)-(1.13), with BnB\subset\mathbb{R}^{n} being a ball, let vW1,(B)v\in W^{1,\infty}(B) such that

{divzHi(Dv)=0DvL(B)+1𝔪.\begin{cases}\,-\,{\rm div}\,\partial_{z}H_{\textnormal{i}}(Dv)=0\\ \,\|Dv\|_{L^{\infty}(B)}+1\leq\mathfrak{m}\,.\end{cases} (5.70)

There exist two functions of 𝔪\mathfrak{m}, c:[1,)[1,)c_{*}\colon[1,\infty)\to[1,\infty) and β:[1,)(0,1)\beta_{*}\colon[1,\infty)\to(0,1), such that

oscτBDvcτβ\textnormal{{osc}}_{\tau B}\,Dv\leq c_{*}\tau^{\beta_{*}} (5.71)

holds for every τ(0,1)\tau\in(0,1). The functions c()c_{*}(\cdot) and β()\beta_{*}(\cdot) are non-decreasing and non-increasing, respectively, and also depend on n,q,μ,ν,L,g()n,q,\mu,\nu,L,g(\cdot). They are independent of a¯\bar{a} and ω(0,1]\omega\in(0,1].

The proof of (5.71) is achieved via two preliminary lemmas. In the rest of section, we denote Dv:=max1sn|Dsv|\|Dv\|:=\max_{1\leq s\leq n}|D_{s}v|.

Lemma 11

With BnB\subset\mathbb{R}^{n} being a ball, assume that vW1,(B)v\in W^{1,\infty}(B) satisfies (5.70)1 and

Dv+ωλin B, where λ𝔪𝔪1.\|Dv\|+\omega\leq\lambda\ \ \mbox{in $B$, where $\lambda\leq\mathfrak{m}$, $\mathfrak{m}\geq 1$}\,. (5.72)

There exists σσ(n,μ,q,ν,L,cg(1),𝔪,g(𝔪))(0,1)\sigma\equiv\sigma(n,\mu,q,\nu,L,c_{g}(1),\mathfrak{m},g(\mathfrak{m}))\in(0,1), which is independent of a¯\bar{a} and is a non-increasing function of 𝔪\mathfrak{m}, such that if either

{|B{Dsv<λ/2}||B|σor|B{Dsv>λ/2}||B|σ=:condition ex(B,λ,s)\begin{cases}\ \displaystyle\frac{|B\cap\{D_{s}v<\lambda/2\}|}{|B|}\leq\sigma\ \quad\mbox{or}\ \quad\frac{|B\cap\{D_{s}v>-\lambda/2\}|}{|B|}\leq\sigma\\ \ \quad=:\mbox{condition $\textnormal{{ex}}(B,\lambda,s)$}\end{cases} (5.73)

holds for some s{1,,n}s\in\{1,\ldots,n\}, then

|Dv||Dsv|λ/4holds a.e. in B/2.\lvert Dv\rvert\geq\lvert D_{s}v\rvert\geq\lambda/4\quad\mbox{holds a.e.\,in $B/2$}\,. (5.74)

In this case

oscτBDvc1τβ1λholds for every τ(0,1),\textnormal{{osc}}_{\tau B}\,Dv\leq c_{1}\tau^{\beta_{1}}\lambda\qquad\mbox{holds for every $\tau\in(0,1)$,} (5.75)

where c11c_{1}\geq 1 and β1(0,1)\beta_{1}\in(0,1) both depend on n,μ,q,ν,L,𝔪,g(𝔪)n,\mu,q,\nu,L,\mathfrak{m},g(\mathfrak{m}), but are otherwise independent of a¯\bar{a}. The constants c1c_{1} and β1\beta_{1} are non-decreasing and non-increasing functions of 𝔪\mathfrak{m}, respectively. On the other hand there exists η~η~(n,μ,q,ν,L,𝔪,g(𝔪))(1/2,1)\tilde{\eta}\equiv\tilde{\eta}(n,\mu,q,\nu,L,\mathfrak{m},g(\mathfrak{m}))\in(1/2,1) such that, if ex(B,λ,s)\textnormal{{ex}}(B,\lambda,s) fails for every s{1,,n}s\in\{1,\ldots,n\}, then

Dvη~λholds a.e. in B/2.\|Dv\|\leq\tilde{\eta}\lambda\quad\mbox{holds a.e.\,in $B/2$}\,. (5.76)

The constant η~\tilde{\eta} is a non-decreasing function of 𝔪\mathfrak{m}.

Proof

For the proof of the first assertion we use and extend some of the arguments in (KuM, , Proposition 3.7); see also dibe for the case of the classical pp-Laplacian operator. We can of course confine ourselves to the case the first inequality in (5.73) occurs (for a σ\sigma to be determined in the course of the proof), the other being similar. By scaling, as in (5.14), we can assume that B1B\equiv\mathcal{B}_{1}. Since conditions (3.10) are satisfied by Hi()H_{\textnormal{i}}(\cdot), it is standard to prove that vWloc2,2(1)v\in W^{2,2}_{{\rm loc}}(\mathcal{B}_{1}). We can then differentiate the equation divzHi(Dv)=0\,{\rm div}\,\partial_{z}H_{\textnormal{i}}(Dv)=0 in the xsx_{s} direction, thereby obtaining div(zzHi(Dv)DDsv)=0\,{\rm div}\,(\partial_{zz}H_{\textnormal{i}}(Dv)DD_{s}v)=0, that is an analogue of (4.8). Recalling that vW1,(1)Wloc2,2(1)v\in W^{1,\infty}(\mathcal{B}_{1})\cap W^{2,2}_{{\rm loc}}(\mathcal{B}_{1}) (as in (4.4) and by (5.72)), we then use as test function (Dsvκ)ϕ2(D_{s}v-\kappa)_{-}\phi^{2}, where ϕC0(1)\phi\in C^{\infty}_{0}(\mathcal{B}_{1}) is non-negative and 0κλ0\leq\kappa\leq\lambda; this is still possible by (3.10)). Integrating by parts the resulting equation we find

1Ds(zHi(Dv)),D(Dsvκ)ϕ2𝑑x\displaystyle\int_{\mathcal{B}_{1}}\langle D_{s}(\partial_{z}H_{\textnormal{i}}(Dv)),D(D_{s}v-\kappa)_{-}\rangle\phi^{2}\,dx
=1zHi(Dv),DsDϕ2(Dsvκ)𝑑x\displaystyle\qquad=\int_{\mathcal{B}_{1}}\langle\partial_{z}H_{\textnormal{i}}(Dv),D_{s}D\phi^{2}\rangle(D_{s}v-\kappa)_{-}\,dx
21{Dsv<κ}zHi(Dv),DϕDssvϕ𝑑x.\displaystyle\qquad\quad-2\int_{\mathcal{B}_{1}\cap\{D_{s}v<\kappa\}}\langle\partial_{z}H_{\textnormal{i}}(Dv),D\phi\rangle D_{ss}v\phi\,dx\,. (5.77)

We define the functions

{G0(t):=g~(t)+a¯(t2+ω2)(q1)/2G(t):=[g~(t)]2(t2+1)μ/2+a¯(t2+ω2)q/2λa¯(t)=(t2+1)μ/2+(q1)a¯(t2+ω2)(q2)/2\begin{cases}G_{0}(t):=\tilde{g}(t)+\bar{a}(t^{2}+\omega^{2})^{(q-1)/2}\\ G(t):=[\tilde{g}(t)]^{2}(t^{2}+1)^{\mu/2}+\bar{a}(t^{2}+\omega^{2})^{q/2}\\ \lambda_{\bar{a}}(t)=(t^{2}+1)^{-\mu/2}+(q-1)\bar{a}(t^{2}+\omega^{2})^{(q-2)/2}\\ \end{cases} (5.78)

for t0t\geq 0, so that

G0(λ)λ+λa¯(λ)λ2+[G0(λ)]2/λa¯(λ)cG(λ)G_{0}(\lambda)\lambda+\lambda_{\bar{a}}(\lambda)\lambda^{2}+[G_{0}(\lambda)]^{2}/\lambda_{\bar{a}}(\lambda)\leq cG(\lambda) (5.79)

holds for cc(q)c\equiv c(q). By also using (3.8), (5.69) and (5.72), and the monotonicity features of λa¯()\lambda_{\bar{a}}(\cdot) and G0()G_{0}(\cdot) in (5.77), we easily find

λa¯(λ)1|D(Dsvκ)|2ϕ2𝑑x\displaystyle\lambda_{\bar{a}}(\lambda)\int_{\mathcal{B}_{1}}|D(D_{s}v-\kappa)_{-}|^{2}\phi^{2}\,dx
cG0(λ)1(ϕ|D2ϕ|+|Dϕ|2)(Dsvκ)𝑑x\displaystyle\qquad\leq cG_{0}(\lambda)\int_{\mathcal{B}_{1}}(\phi|D^{2}\phi|+|D\phi|^{2})(D_{s}v-\kappa)_{-}\,dx
+cG0(λ)1ϕ|Dϕ||D(Dsvκ)|𝑑x.\displaystyle\qquad\quad+cG_{0}(\lambda)\int_{\mathcal{B}_{1}}\phi|D\phi||D(D_{s}v-\kappa)_{-}|\,dx\,.

By then using Young’s inequality and (5.79), and yet recalling that (Dsvκ)2λ(D_{s}v-\kappa)_{-}\leq 2\lambda, we find

λa¯(λ)D(ϕ(Dsvκ))L2(1)2\displaystyle\lambda_{\bar{a}}(\lambda)\|D(\phi(D_{s}v-\kappa)_{-})\|_{L^{2}(\mathcal{B}_{1})}^{2}
cG(λ)|1{Dsv<κ}{ϕ>0}|(DϕL2+D2ϕL).\displaystyle\qquad\leq cG(\lambda)|\mathcal{B}_{1}\cap\{D_{s}v<\kappa\}\cap\{\phi>0\}|(\|D\phi\|_{L^{\infty}}^{2}+\|D^{2}\phi\|_{L^{\infty}})\,. (5.80)

For integers m0m\geq 0 we choose levels {κm=λ(1+1/2m)/4}\{\kappa_{m}=\lambda(1+1/2^{m})/4\}, radii {ϱm=1/2+1/2m+1}\{\varrho_{m}=1/2+1/2^{m+1}\} and cut-off functions ϕmC0(ϱm)\phi_{m}\in C^{\infty}_{0}(\mathcal{B}_{\varrho_{m}}), with ϕm1\phi_{m}\equiv 1 on ϱm+1\mathcal{B}_{\varrho_{m+1}}, DϕmL2+D2ϕmL4m\|D\phi_{m}\|_{L^{\infty}}^{2}+\|D^{2}\phi_{m}\|_{L^{\infty}}\lesssim 4^{m}. With

Am:={Dsv<κm}BϱmA_{m}:=\{D_{s}v<\kappa_{m}\}\cap B_{\varrho_{m}} (5.81)

estimate (5.80) becomes

λa¯(λ)D(ϕm(Dsvκm))L2(1)2c4mG(λ)|Am|.\lambda_{\bar{a}}(\lambda)\|D(\phi_{m}(D_{s}v-\kappa_{m})_{-})\|_{L^{2}(\mathcal{B}_{1})}^{2}\leq c4^{m}G(\lambda)|A_{m}|\,.

We then find, via Sobolev embedding

λa¯(λ)(κmκm+1)2|Am+1|\displaystyle\lambda_{\bar{a}}(\lambda)(\kappa_{m}-\kappa_{m+1})^{2}|A_{m+1}| λa¯(λ)ϕm(Dsvκm)L2(1)2\displaystyle\leq\lambda_{\bar{a}}(\lambda)\|\phi_{m}(D_{s}v-\kappa_{m})_{-}\|_{L^{2}(\mathcal{B}_{1})}^{2}
cλa¯(λ)D(ϕm(Dsvκm))L2(1)2|Am|1/n\displaystyle\leq c\lambda_{\bar{a}}(\lambda)\|D(\phi_{m}(D_{s}v-\kappa_{m})_{-})\|_{L^{2}(\mathcal{B}_{1})}^{2}|A_{m}|^{1/n}
c4mG(λ)|Am|1+1/n.\displaystyle\leq c4^{m}G(\lambda)|A_{m}|^{1+1/n}\,.

Observing that κmκm+12mλ\kappa_{m}-\kappa_{m+1}\approx 2^{-m}\lambda and that (5.72) implies λ2λ2+ω2\lambda^{2}\approx\lambda^{2}+\omega^{2}, the above display gives

|Am+1|\displaystyle|A_{m+1}| \displaystyle\leq c24mG(λ)λa¯(λ)λ2|Am|1+1/n\displaystyle\frac{c2^{4m}G(\lambda)}{\lambda_{\bar{a}}(\lambda)\lambda^{2}}|A_{m}|^{1+1/n} (5.82)
(5.68),(5.78)\displaystyle\stackrel{{\scriptstyle(\ref{ragnar}),(\ref{ragnar2})}}{{\leq}} c[(λ2+1)μ1[g(λ)]2+1]24m|Am|1+1/n\displaystyle c\left[(\lambda^{2}+1)^{\mu-1}[g(\lambda)]^{2}+1\right]2^{4m}|A_{m}|^{1+1/n}
(5.72)\displaystyle\stackrel{{\scriptstyle(\ref{limitatio})}}{{\leq}} c[𝔪2(μ1)[g(𝔪)]2+1]24m|Am|1+1/n\displaystyle c\left[\mathfrak{m}^{2(\mu-1)}[g(\mathfrak{m})]^{2}+1\right]2^{4m}|A_{m}|^{1+1/n}

with cc(n,μ,q,ν,L,cg(1))c\equiv c(n,\mu,q,\nu,L,c_{g}(1)) which is independent of a¯\bar{a}. Inequality (5.82) allows to perform a standard geometric iteration (see for instance (giu, , Lemma 7.1)) leading to Dsvλ/4D_{s}v\geq\lambda/4 once the first inequality in (5.73) is verified with σσ(n,μ,q,ν,L,cg(1),𝔪,g(𝔪))\sigma\equiv\sigma(n,\mu,q,\nu,L,c_{g}(1),\mathfrak{m},g(\mathfrak{m})) small enough (that corresponds to require that |A0||A_{0}| is small, see (5.81)). This allows to prove (5.75) by considering

div(a(x)DDtv)=0,a(x):=[λa¯(λ)]1zzHi(Dv(x))-\,{\rm div}\,(\texttt{a}(x)DD_{t}v)=0\,,\quad\texttt{a}(x):=[\lambda_{\bar{a}}(\lambda)]^{-1}\partial_{zz}H_{\textnormal{i}}(Dv(x)) (5.83)

this time for every t{1,,n}t\in\{1,\ldots,n\}. Recalling (3.8)-(3.9) it follows that a()\texttt{a}(\cdot) satisfies

|ξ|2(5.72)ca(x)ξ,ξ,|a(x)|(5.74)c[𝔪μ1g(𝔪)+1]|\xi|^{2}\stackrel{{\scriptstyle(\ref{limitatio})}}{{\leq}}c\langle\texttt{a}(x)\xi,\xi\rangle\,,\quad\lvert\texttt{a}(x)\rvert\stackrel{{\scriptstyle(\ref{dasotto})}}{{\leq}}c[\mathfrak{m}^{\mu-1}g(\mathfrak{m})+1] (5.84)

a.e. in x1/2x\in\mathcal{B}_{1/2} and for every ξn\xi\in\mathbb{R}^{n}, where cc(n,μ,q,ν,L)c\equiv c(n,\mu,q,\nu,L). Therefore a standard application of De Giorgi-Nash-Moser theory to DtvD_{t}v and (5.72) imply the validity of (5.75) (note that for this we obtain (5.75) for τ1/8\tau\leq 1/8, and then the case 1/8<τ11/8<\tau\leq 1 follows trivially by (5.72)). We now turn to (5.76); the proof is a variant of the one in (KuM, , Proposition 3.11) apart from the parabolic case considered there. The only remark we need to do is that the analogue of equation in (KuM, , (3.53)) is here given by div(b(x)Dz)=0\,{\rm div}\,(\texttt{b}(x)Dz)=0, where b(x):=[λa¯(λ)]1zzHi(Dv(x))=a(x)\texttt{b}(x):=[\lambda_{\bar{a}}(\lambda)]^{-1}\partial_{zz}H_{\textnormal{i}}(Dv(x))=\texttt{a}(x) if xx belongs to the support of v~:=(Dsvλ/2)+\tilde{v}:=(D_{s}v-\lambda/2)_{+}, and b(x)𝕀d\texttt{b}(x)\equiv\mathds{I}_{d} (the identity matrix) otherwise. Then v~\tilde{v} turns out to be a weak subsolution. Moreover, as |Dv|λ/2|Dv|\geq\lambda/2 on the support of v~\tilde{v}, it follows that b(x)\texttt{b}(x) also satisfies (5.84). We can now use the arguments in (KuM, , pp. 784-786) to conclude with (5.76). Finally, as for the monotone dependence on 𝔪\mathfrak{m} of the constants c1,β1,η~,σc_{1},\beta_{1},\tilde{\eta},\sigma, this is classical, and it is a consequence of the fact that all the estimates above feature constants, usually denoted by cc, that are non-decreasing functions of 𝔪\mathfrak{m} and g(𝔪)g(\mathfrak{m}); in turn, g()g(\cdot) is a non-decreasing function too.

Lemma 12

Let vv be as in Lemma 11, assume (5.72) and that mωλm\omega\geq\lambda holds for some integer m1m\geq 1. Then

oscτBDvc2τβ2λholds for every τ(0,1),\textnormal{{osc}}_{\tau B}\,Dv\leq c_{2}\tau^{\beta_{2}}\lambda\qquad\mbox{holds for every $\tau\in(0,1)$,} (5.85)

where c21c_{2}\geq 1 and β2(0,1)\beta_{2}\in(0,1) both depend on n,μ,q,ν,L,𝔪,g(𝔪),mn,\mu,q,\nu,L,\mathfrak{m},g(\mathfrak{m}),m, but are otherwise independent of a¯\bar{a}. The constants c2c_{2} and β2\beta_{2} are non-decreasing and non-increasing functions of 𝔪\mathfrak{m} and mm, respectively.

Proof

As for Lemma 11 we reduce to the case B1B\equiv\mathcal{B}_{1} and consider (5.83). The first inequality in (5.84) follows as Dvλ\|Dv\|\leq\lambda a.e.; using mωλm\omega\geq\lambda and (3.8)3 we instead have

|a(x)|c[𝔪μg(𝔪)+m2q]a.e. in x1,\lvert\texttt{a}(x)\rvert\leq c[\mathfrak{m}^{\mu}g(\mathfrak{m})+m^{2-q}]\quad\mbox{a.e.\,in $x\in\mathcal{B}_{1}$}\,,

where cc(n,μ,q,ν,L)c\equiv c(n,\mu,q,\nu,L). Now (5.85) follows from standard De Giorgi-Nash-Moser theory.

Proof (of Proposition 3)

We take σ,η~σ,η~(n,μ,q,ν,L,𝔪,g(𝔪))(0,1)\sigma,\tilde{\eta}\equiv\sigma,\tilde{\eta}(n,\mu,q,\nu,L,\mathfrak{m},g(\mathfrak{m}))\in(0,1) from Lemma 11 and determine an integer mm(n,μ,q,ν,L,𝔪,g(𝔪))1m\equiv m(n,\mu,q,\nu,L,\mathfrak{m},g(\mathfrak{m}))\geq 1 such that η:=η~+1/m<1\eta:=\tilde{\eta}+1/m<1; notice that mm can be determined as a non-decreasing function of 𝔪\mathfrak{m}. All in all, once 𝔪\mathfrak{m} is fixed so is mm and therefore η\eta. We set λ0:=supBDv+ω\lambda_{0}:=\sup_{B}\,\|Dv\|+\omega, λk:=ηkλ0\lambda_{k}:=\eta^{k}\lambda_{0}, Bk:=B/2kB_{k}:=B/2^{k} for every kk\in\mathbb{N}, and define k¯\bar{k}\in\mathbb{N} as the smallest non-negative integer for which mω>λk¯m\omega>\lambda_{\bar{k}}. This implies that

ωλkm,for 0k<k¯.\omega\leq\frac{\lambda_{k}}{m},\quad\mbox{for \ $0\leq k<\bar{k}$}\,. (5.86)

Note that (5.86) is empty when k¯=0\bar{k}=0. We further define

A:={k:0k<k¯andex(Bk,λk,s) holds for some s{1,,n}}A:=\{k\in\mathbb{N}\,\colon 0\leq k<\bar{k}\ \mbox{and}\ \mbox{$\textnormal{{ex}}(B_{k},\lambda_{k},s)$ holds for some $s\in\{1,\ldots,n\}$}\}

and k~\tilde{k} as k~:=minA\tilde{k}:=\min A if AA\not=\emptyset, and k~=k¯\tilde{k}=\bar{k} if A=A=\emptyset. Using (5.76) and (5.86) iteratively, we find

supBkDv+ωλk=ηkλ0for every 0kk~.\sup_{B_{k}}\|Dv\|+\omega\leq\lambda_{k}=\eta^{k}\lambda_{0}\quad\mbox{for every $0\leq k\leq\tilde{k}$}\,. (5.87)

On the other hand, by the very definition of k~\tilde{k} and Lemmas 11-12 (the latter is only needed when k~=k¯\tilde{k}=\bar{k} and therefore when A=A=\emptyset), we find that

oscτBk~Dvmax{c1,c2}τmin{β1,β2}λk~holds for 0<τ<1.\textnormal{{osc}}_{\tau B_{\tilde{k}}}\,Dv\leq\max\{c_{1},c_{2}\}\tau^{\min\{\beta_{1},\beta_{2}\}}\lambda_{\tilde{k}}\qquad\mbox{holds for \ $0<\tau<1$}\,. (5.88)

At this stage (5.71) follows via a standard interpolation argument combining (5.87)-(5.88).

Remark 4

As a consequence of the local gradient boundedness obtained in Section 5.8, the arguments developed for Proposition 3 do not require any upper bound on μ\mu. Moreover, the specific structure in (5.67) is not indispensable and the methods used here can be used as a starting point to treat general nonautonomous functionals with (p,q)(p,q)-growth of the type considered in M2 .

5.11 Improved Hölder exponent when s>0\textnormal{{s}}>0

The Cloc1,α~/2(Ω)C^{1,\tilde{\alpha}/2}_{{\rm loc}}(\Omega)-regularity proof follows almost verbatim the one in (piovra, , Section 10) since now we already know that DuDu is locally Hölder continuous in Ω\Omega with exponent β\beta. We confine ourselves to give a few remarks on the main modifications and to facilitate the adaptation we use the same notation introduced in piovra . We set M:=DuL(Ω)+[Du]0,β;Ω+1M:=\|Du\|_{L^{\infty}(\Omega)}+[Du]_{0,\beta;\Omega}+1 (we can assume this is finite since our result is local). This time we take Ar(z):=zHrα~,i(z)=zF(xc,z)+qarα~,i(Br)[s(z)]q2zA_{r}(z):=\partial_{z}H_{r^{\tilde{\alpha}},\textnormal{i}}(z)=\partial_{z}F(x_{\rm c},z)+qa_{r^{\tilde{\alpha}},\textnormal{i}}(B_{r})[\ell_{\textnormal{{s}}}(z)]^{q-2}z and recall that |zHrα~,i(z)zH(x,z)|crα~[1(z)]q1|\partial_{z}H_{r^{\tilde{\alpha}},\textnormal{i}}(z)-\partial_{z}H(x,z)|\leq cr^{\tilde{\alpha}}[\ell_{1}(z)]^{q-1} holds whenever xBrx\in B_{r}. With vu+W01,q(Br)v\in u+W^{1,q}_{0}(B_{r}) to solve divAr(Dv)=0\,{\rm div}\,A_{r}(Dv)=0 in BrB_{r}, as in (5.61) we gain DvL(Br/2)cM2\lVert Dv\rVert_{L^{\infty}(B_{r/2})}\leq cM^{2} with cc(data)c\equiv c(\textnormal{{data}}). With these informations estimate DuDvL2(Br/2)2crn+α~\lVert Du-Dv\rVert_{L^{2}(B_{r/2})}^{2}\leq cr^{n+\tilde{\alpha}} follows as in (5.65) (in turn as in (9)), with cc(data,M)c\equiv c(\textnormal{{data}},M). This is the analogue of (piovra, , (10.3)), but for the fact that here we find α~\tilde{\alpha} rather than 2α~2\tilde{\alpha} and the integral is supported in Br/2B_{r/2} instead of BrB_{r}. Next, by defining the matrix [𝔸(x)]ij:=zjAri(Dv(x))[\mathbb{A}(x)]_{ij}:=\partial_{z_{j}}A_{r}^{i}(Dv(x)), to use piovra we need to prove that there exists λλ(data,s,M)>0\lambda\equiv\lambda(\textnormal{{data}},\textnormal{{s}},M)>0, independent of rr, such that λ𝕀d𝔸(x)(1/λ)𝕀d\lambda\mathds{I}_{\rm d}\leq\mathbb{A}(x)\leq(1/\lambda)\mathds{I}_{\rm d} holds for a.e. xBr/2x\in B_{r/2}. As a consequence of (1.11)2 and of |Dv|M2|Dv|\lesssim M^{2} in Br/2B_{r/2}, we have

|ξ|2(M4+1)μ/2c𝔸(x)ξ,ξ,|𝔸(x)|c[g(M2)+sq2]\frac{|\xi|^{2}}{(M^{4}+1)^{\mu/2}}\leq c\langle\mathbb{A}(x)\xi,\xi\rangle\,,\qquad|\mathbb{A}(x)|\leq c[g(M^{2})+\textnormal{{s}}^{q-2}]

for a.e. xBr/2x\in B_{r/2} and z,ξnz,\xi\in\mathbb{R}^{n}, where c(data,M)c\equiv(\textnormal{{data}},M). This is the crucial point where we use After this, the proof follows exactly as (piovra, , (10.6)) and the proof of the entire Theorem 1.2 is finally complete.

6 Theorem 1.1, Corollary 1 and model examples (1.10)

For every integer k0k\geq 0, we consider the integrand

F(x,z):=𝔠(x)|z|Lk+1(|z|)+1,F(x,z):=\mathfrak{c}(x)|z|L_{k+1}(|z|)+1,

with 𝔠()\mathfrak{c}(\cdot) as in (1.7)2 and Lk+1()L_{k+1}(\cdot) as in (1.10)2. Direct computations show that F()F(\cdot) satisfies (1.11) with g(t)Lk+1(t)+1g(t)\equiv L_{k+1}(t)+1, with μ=1\mu=1 when k=0k=0, and with any choice of μ>1\mu>1 when k>0k>0; moreover, it also satisfies (1.13). The constants ν\nu and LL depend on μ,k\mu,k and Λ\Lambda. Theorem 1.2 therefore applies to local minimizers of

wΩ[𝔠(x)|Dw|Lk+1(|Dw|)+a(x)(Dw|2+s2)q/2+1]dx.w\mapsto\int_{\Omega}[\mathfrak{c}(x)|Dw|L_{k+1}(|Dw|)+a(x)(Dw|^{2}+\textnormal{{s}}^{2})^{q/2}+1]\,dx\,.

On the other hand the functional in the above display and

wΩ[𝔠(x)|Dw|Lk+1(|Dw|)+a(x)(Dw|2+s2)q/2]dxw\mapsto\int_{\Omega}[\mathfrak{c}(x)|Dw|L_{k+1}(|Dw|)+a(x)(Dw|^{2}+\textnormal{{s}}^{2})^{q/2}]\,dx

share the same local minimizers, and therefore the regularity results stated in Theorem 1.2 hold for local minimizers of the last functional too. In this way the model case in (1.10) is covered. Theorem 1.1 and Corollary 1 then follow as special cases taking k=0k=0 and s=0\textnormal{{s}}=0 and s=1\textnormal{{s}}=1, respectively. Note that, in the spirit of Corollary 1, local minimizers of the functional

wΩ𝔠(x)|Dw|Lk+1(|Dw|)𝑑xw\mapsto\int_{\Omega}\mathfrak{c}(x)|Dw|L_{k+1}(|Dw|)\,dx

are locally C1,α0/2C^{1,\alpha_{0}/2}-regular in Ω\Omega, for every k0k\geq 0, provided (1.7)2 is assumed.

7 Vectorial cases and further generalizations

When functionals as in (1.1) and (1.9) are considered in the vectorial case, i.e. minima are vector valued u:ΩNu\colon\Omega\to\mathbb{R}^{N} and N>1N>1, we can still obtain Lipschitz continuity results. In this situation it is unavoidable to impose a so-called Uhlenbeck structure uh , that is

F(x,z)=F~(x,|z|),F(x,z)=\tilde{F}(x,|z|)\,, (7.1)

where F~:Ω×[0,)[0,)\tilde{F}\colon\Omega\times[0,\infty)\to[0,\infty) is a continuous function such that tF~(x,t)t\mapsto\tilde{F}(x,t) is C2C^{2}-regular for every choice of xΩx\in\Omega. This is obviously satisfied by the models in (1.1) and (1.10).

Theorem 7.1

Let uWloc1,1(Ω;N)u\in W^{1,1}_{{\rm loc}}(\Omega;\mathbb{R}^{N}) be a local minimizer of the functional in (1.9) under assumptions (1.7)1, (1.11)-(1.12) and (7.1). There exists μmax(1,2)\mu_{\textnormal{max}}\in(1,2), depending only n,q,α,α0n,q,\alpha,\alpha_{0}, such that, if 1μ<μmax1\leq\mu<\mu_{\textnormal{max}}, then (1.14) holds as in Theorem 1.2.

The proof of Theorem 7.1 is essentially the same of the one given for Theorem 1.2, once the content of Proposition 1 is available. This is in fact the only point where (7.1) enters the game. Inspecting the proof of Proposition 1 shows that this works in the vectorial case provided (4.9) holds. In turn, this inequality follows along the lines of the estimates in (BM, , Lemma 5.6). Notice that, without an additional structure assumption as in (7.1), Theorem 7.1 cannot hold and counterexamples to Lipschitz regularity emerge already when considering uniformly elliptic systems SY . We also remark that we expect gradient Hölder continuity in the vectorial case as well; the proof must be different from the one given in Section 5.10, and based on linearization methods as those originally introduced in uh and developed in large parts of the subsequent literature. Another direct generalization, this time in the scalar case, occurs for functionals of the type

w𝒩(w,Ω):=Ω[F(x,Dw)+a(x)G(Dw)]𝑑x,w\mapsto\mathcal{N}(w,\Omega):=\int_{\Omega}\left[F(x,Dw)+a(x)G(Dw)\right]\,dx\,, (7.2)

where F()F(\cdot) is as Theorem 1.2 and G:n[0,)G\colon\mathbb{R}^{n}\to[0,\infty) belongs to C1(n)C2(n{0})C^{1}(\mathbb{R}^{n})\cap C^{2}(\mathbb{R}^{n}\setminus\{0\}) and satisfies

{ν(|z|2+s2)q/2G(z)L(|z|2+s2)q/2ν(|z|2+s2)(q2)/2|ξ|2zzG(z)ξ,ξ|zzG(z)|L(|z|2+s2)(q2)/2\begin{cases}\ \nu(|z|^{2}+\textnormal{{s}}^{2})^{q/2}\leq G(z)\leq L(|z|^{2}+\textnormal{{s}}^{2})^{q/2}\\ \ \displaystyle\nu(|z|^{2}+\textnormal{{s}}^{2})^{(q-2)/2}|\xi|^{2}\leq\langle\partial_{zz}G(z)\xi,\xi\rangle\\ \ |\partial_{zz}G(z)|\leq L(|z|^{2}+\textnormal{{s}}^{2})^{(q-2)/2}\end{cases}

for every z,ξnz,\xi\in\mathbb{R}^{n}, |z|0|z|\not=0. Theorem 1.2 continuous to hold in this last case, with essentially the same proof. The only difference worth pointing out is the new shape of the integrands H(x,z):=F(x,z)+a(x)G(z)H(x,z):=F(x,z)+a(x)G(z) in (3.1) and Hω,σ()H_{\omega,\sigma}(\cdot) (and the related minimal one Hω,σ,i()H_{\omega,\sigma,\textnormal{i}}(\cdot)) in (3.2), accordingly defined as

{Hω,σ(x,z):=F(x,z)+aσ(x)Gω(z)Gω(z):=1G(z+ωλ)ϕ(λ)dλ.\begin{cases}\,H_{\omega,\sigma}(x,z):=F(x,z)+a_{\sigma}(x)G_{\omega}(z)\\ \,\displaystyle G_{\omega}(z):=\int_{\mathcal{B}_{1}}G(z+\omega\lambda)\phi(\lambda)\,{\rm d}\lambda\,.\end{cases}

Here {ϕε}C(n)\{\phi_{\varepsilon}\}\subset C^{\infty}(\mathbb{R}^{n}), denotes a family of radially symmetric mollifiers, defined as ϕε(x):=ϕ(x/ε)/εn\phi_{\varepsilon}(x):=\phi(x/\varepsilon)/\varepsilon^{n}, where ϕCc(1)\phi\in C^{\infty}_{c}(\mathcal{B}_{1}), ϕL1(n)=1\lVert\phi\rVert_{L^{1}(\mathbb{R}^{n})}=1, 3/4suppϕ\mathcal{B}_{3/4}\subset\,{\rm supp}\,\phi. The definition of ω,σ,i()\mathbb{H}_{\omega,\sigma,\textnormal{i}}(\cdot), instead, remains the same. Following for instance (dm, , Section 4.5) and (piovra, , Section 6.7) it is possible to show that the newly defined integrands Hω,σ()H_{\omega,\sigma}(\cdot) still have the properties described in Section 3. In particular (3.5) and (3.6) still hold for a suitable constant cc with the same dependence upon the various parameters described there. Finally, when again considering the vectorial case for functionals as in (7.2) we can still obtain local Lipschitz regularity of minima provided (7.1) holds together with G(z)G~(|z|)G(z)\equiv\tilde{G}(|z|) and G~()\tilde{G}(\cdot) is C2C^{2}-regular outside the origin (see for instance BM ; ciccio for precise assumptions).

Acknowledgments. The first author is supported by INdAM-GNAMPA via the project “Problemi non locali: teoria cinetica e non uniforme ellitticità", and by the University of Parma via the project “Local vs nonlocal: mixed type operators and nonuniform ellipticity". Both the authors are grateful to the referees for the careful reading of the original version of the manuscript and for the many suggestions and comments that eventually led to a better presentation.

Funding Open access funding provided by Università degli Studi di Parma within the CRUI-CARE Agreement.

Data Availibility No data is attached to this paper.

Declarations

Conflict of interests and data. The authors declare to have no conflict of interests. No data are attached to this paper.

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit creativecommons.org/ licenses/by/4.0/.

References

  • (1) E. Acerbi, G. Bouchitté, I. Fonseca, Relaxation of convex functionals: the gap problem, Ann. Inst. Henri Poincaré, Anal. Non Linéaire 20, 359–390 (2003).
  • (2) D.R. Adams, L.I. Hedberg, Function spaces and potential theory, Grundlehren der mathematischen Wissenschaften 314. Springer-Verlag, Berlin, 1996.
  • (3) A. Balci, L. Diening, M. Surnachev, New examples on Lavrentiev gap using fractals, Calc. Var. & PDE 59:180 (2020).
  • (4) A. Balci, C. Ortner, J. Storn, Crouzeix-Raviart finite element method for non-autonomous variational problems with Lavrentiev gap, Numer. Math. 151, 779–805 (2022).
  • (5) P. Baroni, M. Colombo, G. Mingione, Regularity for general functionals with double phase, Calc. Var. & PDE 57:62 (2018).
  • (6) S. Baasandorj, S.S. Byun, Regularity for Orlicz phase problems, Memoirs Amer. Math. Soc., to appear.
  • (7) L. Beck, M. Bulíček, F. Gmeineder, On a Neumann problem for variational functionals of linear growth, Ann. Sc. Norm. Super. Pisa Cl. Sci. (V) 21, 695–737 (2020).
  • (8) L. Beck, G. Mingione, Lipschitz bounds and nonuniform ellipticity, Comm. Pure Appl. Math. 73, 944-1034 (2020).
  • (9) P. Bella, M. Schäffner, Local boundedness and Harnack inequality for solutions of linear nonuniformly elliptic equations, Comm. Pure Appl. Math. 74, 453-477 (2021).
  • (10) P. Bella, M. Schäffner, On the regularity of minimizers for scalar integral functionals with (p,q)(p,q)-growth, Anal. & PDE 13, 2241–2257 (2020).
  • (11) P. Bella, M. Schäffner, Lipschitz bounds for integral functionals with (p,q)(p,q)-growth conditions, Adv. Calc. Var. 10.1515/acv-2022-0016
  • (12) M. Bildhauer, Convex variational problems. Linear, nearly linear and anisotropic growth conditions. Lecture Notes in Mathematics, 1818. Springer-Verlag, Berlin, 2003. x+217.
  • (13) M. Bildhauer, M. Fuchs, Partial regularity for variational integrals with (s,μ,q)(s,\mu,q)-growth, Calc. Var. & PDE 13, 537–560 (2001).
  • (14) M. Bildhauer, M. Fuchs, C1,αC^{1,\alpha}-solutions to non-autonomous anisotropic variational problems, Calc. Var. & PDE 24, 309–340 (2005).
  • (15) G. Bouchitté, I. Fonseca, J. Malý, The effective bulk energy of the relaxed energy of multiple integrals below the growth exponent, Proc. R. Soc. Edinb., Sect. A, Math. 128, 463-479 (1998).
  • (16) M. Bulíček, P. Gwiazda, J. Skrzeczkowski, On a range of exponents for absence of Lavrentiev phenomenon for double phase functionals, Arch. Ration. Mech. Anal. 246, 209–240 (2022).
  • (17) S.S. Byun, J. Oh, Regularity results for generalized double phase functionals, Anal. PDE 13, 1269–1300 (2020).
  • (18) M. Colombo, G. Mingione, Regularity for double phase variational problems, Arch. Ration. Mech. Anal. 215, 443–496 (2015).
  • (19) C. De Filippis, Quasiconvexity and partial regularity via nonlinear potentials, J. Math. Pures Appl. 163, 11–82 (2022).
  • (20) C. De Filippis, G. Mingione, A borderline case of Calderón-Zygmund estimates for nonuniformly elliptic problems, St. Petersburg Math. J. 31, 455–477 (2020).
  • (21) C. De Filippis, G. Mingione, Manifold constrained nonuniformly elliptic problems, J. Geom. Anal. 30:1661-1723 (2020).
  • (22) C. De Filippis, G. Mingione, On the regularity of minima of non-autonomous functionals, J. Geom. Anal. 30:1584-1626 (2020).
  • (23) C. De Filippis, G. Mingione, Lipschitz bounds and nonautonomous integrals, Arch. Ration. Mech. Anal. 242, 973-1057 (2021).
  • (24) C. De Filippis, G. Mingione, Nonuniformly elliptic Schauder theory, arXiv:2201.07369v1.
  • (25) C. De Filippis, B. Stroffolini, Singular multiple integrals and nonlinear potentials, J. Funct. Anal. 285, paper 109952 (2023).
  • (26) E. DiBenedetto, C1+αC^{1+\alpha}-local regularity of weak solutions of degenerate elliptic equations, Nonlinear Anal. 7, 827-850 (1983).
  • (27) E. Di Nezza, G. Palatucci, E. Valdinoci, Hitchhiker’s guide to the fractional Sobolev spaces, Bull. Sci. Math. 136, 521-573 (2012).
  • (28) L. Esposito, F. Leonetti, G. Mingione, Sharp regularity for functionals with (p,q)(p,q) growth, J. Diff. Equ. 204, 5-55 (2004).
  • (29) I. Fonseca, J. Malý, Relaxation of multiple integrals below the growth exponent, Ann. Inst. H. Poincaré Anal. Non Linéaire 14, 309-338 (1997)
  • (30) I. Fonseca, J. Malý, G. Mingione, Scalar minimizers with fractal singular sets, Arch. Ration. Mech. Anal. 172, 295-307 (2004).
  • (31) I. Fonseca, P. Marcellini, Relaxation of multiple integral in subcritical Sobolev spaces, J. Geom. Anal. 7, 57-81 (1997).
  • (32) J. Frehse, G. Seregin, Regularity of solutions to variational problems of the deformation theory of plasticity with logarithmic hardening, Proc. St. Petersburg Math. Soc. V, 127–152, Amer. Math. Soc. Transl. Ser. 2, 193, Amer. Math. Soc., Providence, RI, 1999.
  • (33) M. Fuchs, G. Mingione, Full C1,αC^{1,\alpha}-regularity for free and constrained local minimizers of elliptic variational integrals with nearly linear growth, manuscripta math. 102 (2000), 227–250.
  • (34) M. Fuchs, G. Seregin, Variational methods for problems from plasticity theory and for generalized Newtonian fluids, Lecture Notes in Mathematics, 1749. Springer-Verlag, Berlin, 2000. vi+269 pp.
  • (35) M. Giaquinta, E. Giusti, Differentiability of minima of nondifferentiable functionals, Invent. Math. 72, 285-298 (1983).
  • (36) M. Giaquinta, G. Modica, Remarks on the regularity of the minimizers of certain degenerate functionals, manuscripta math. 57, 55-99 (1986).
  • (37) E. Giusti, Direct Methods in the calculus of variations, World Scientific Publishing Co., Inc., River Edge (2003).
  • (38) F. Gmeineder, Partial regularity for symmetric quasiconvex functionals on BD, J. Math. Pures Appl. (IX) 145 83–129 (2021).
  • (39) F. Gmeineder, The regularity of minima for the Dirichlet problem on BD, Arch. Ration. Mech. Anal. 237, 1099–1171 (2020).
  • (40) F. Gmeineder, J. Kristensen, Partial regularity for BV minimizers, Arch. Ration. Mech. Anal. 232, 1429–1473 (2019).
  • (41) F. Gmeineder, J. Kristensen, Sobolev regularity for convex functionals on BD, Calc. Var. & PDE 58:56 (2019).
  • (42) C. Hamburger, Regularity of differential forms minimizing degenerate elliptic functionals, J. reine angew. Math. (Crelle J.) 431, 7-64 (1992).
  • (43) P. Hästö, J. Ok, Maximal regularity for local minimizers of non-autonomous functionals, J. Eur. Math. Soc. 24, 1285–1334 (2022).
  • (44) P. Hästö, J. Ok, Regularity theory for non-autonomous partial differential equations without Uhlenbeck structure, Arch. Ration. Mech. Anal. 245, 1401–1436 (2022).
  • (45) A.V. Ivanov, The Dirichlet problem for second order quasilinear nonuniformly elliptic equations, Trudy Mat. Inst. Steklov. 116 (1971) 34–54.
  • (46) A.V. Ivanov, Local estimates of the maximum modulus of the first derivatives of the solutions of quasilinear nonuniformly elliptic and nonuniformly parabolic equations and of systems of general form, Proc. Steklov Inst. Math. 110, 48–71 (1970).
  • (47) A.V. Ivanov, Quasilinear degenerate and nonuniformly elliptic and parabolic equations of second order, Proc. Steklov Inst. Math. 1984, no. 1 (160), xi+287 pp.
  • (48) N.M. Ivočkina, A.P. Oskolkov, Nonlocal estimates of the first derivatives of solutions of the Dirichlet problem for nonuniformly elliptic quasilinear equations, Zap. Naučn. Sem. Leningrad. Otdel. Mat. Inst. Steklov. (LOMI) 5, 37–109 (1967).
  • (49) T. Kilpeläinen, J. Malý, The Wiener test and potential estimates for quasilinear elliptic equations, Acta Math. 172, 137–161 (1994).
  • (50) L. Koch, Global higher integrability for minimisers of convex functionals with (p,q)(p,q)-growth, Calc. Var. & PDE 60:63 (2021).
  • (51) L. Koch, On global absence of Lavrentiev gap for functionals with (p,q)(p,q)-growth, arXiv:2210.15454
  • (52) J. Kristensen, G. Mingione, The singular set of ω\omega-minima, Arch. Ration. Mech. Anal. 177, 93–114 (2005).
  • (53) T. Kuusi, G. Mingione, Gradient regularity for nonlinear parabolic equations, Ann. Scu. Norm. Sup. Pisa Cl. Sci. (V) 12, 755–822 (2013).
  • (54) O.A. Ladyzhenskaya, N.N. Ural’tseva, Local estimates for gradients of solutions of nonuniformly elliptic and parabolic equations, Comm. Pure Appl. Math. 23, 677–703 (1970).
  • (55) J. J. Manfredi, Regularity of the gradient for a class of nonlinear possibly degenerate elliptic equations, Ph.D. Thesis, University of Washington, St. Louis.
  • (56) J. J. Manfredi, Regularity for minima of functionals with pp-growth, J. Diff. Equ. 76, 203–212 (1988).
  • (57) P. Marcellini, On the definition and the lower semicontinuity of certain quasiconvex integrals, Ann. Inst. H. Poincaré Anal. Non Linéaire 3, 391–409 (1986).
  • (58) P. Marcellini, The stored-energy for some discontinuous deformations in nonlinear elasticity, Partial Differential Equations and the Calculus of Variations vol. II, Birkhäuser Boston Inc. (1989).
  • (59) P. Marcellini, Regularity of minimizers of integrals of the calculus of variations with non standard growth conditions, Arch. Rat. Mech. Anal. 105, 267–284 (1989).
  • (60) P. Marcellini, Regularity and existence of solutions of elliptic equations with p,qp,q-growth conditions, J. Diff. Equ. 90, 1–30 (1991).
  • (61) P. Marcellini, Regularity for elliptic equations with general growth conditions, J. Diff. Equ. 105, 296–333 (1993).
  • (62) P. Marcellini, G. Papi, Nonlinear elliptic systems with general growth, J. Diff. Equ. 221 412–443 (2006).
  • (63) V. Maz’ya, M. Havin, A nonlinear potential theory, Russ. Math. Surveys, 27, 71–148 (1972).
  • (64) J. Serrin, The problem of Dirichlet for quasilinear elliptic differential equations with many independent variables, Philos. Trans. R. Soc. Lond., Ser. A 264, 413-496 (1969).
  • (65) L. Simon, Interior gradient bounds for nonuniformly elliptic equations, Indiana Univ. Math. J. 25, 821–855(1976).
  • (66) V. Sverák, X. Yan, Non-Lipschitz minimizers of smooth uniformly convex functionals, Proc. Natl. Acad. Sci. USA 99, 15269–15276 (2002).
  • (67) H. Triebel, Theory of function spaces. III. Monographs in Mathematics, 100. Birkhäuser Verlag, Basel, 2006.
  • (68) N. Trudinger, The Dirichlet problem for nonuniformly elliptic equations, Bull. Am. Math. Soc. 73, 410–413 (1967).
  • (69) K. Uhlenbeck, K, Regularity for a class of non-linear elliptic systems, Acta Math. 138, 219–240 (1977).
  • (70) V. V. Zhikov, On Lavrentiev’s Phenomenon, Russian J. Math. Phys. 3, 249–269 (1995).
  • (71) V. V. Zhikov, On some variational problems, Russian J. Math. Phys. 5, 105–116 (1997).
  • (72) V. V. Zhikov, S.M. Kozlov, O. A. Oleĭnik, Homogenization of differential operators and integral functionals. Springer-Verlag, Berlin, 1994. xii+570 pp. ISBN: 3-540-54809-2
\address

Dipartimento SMFI
Università di Parma
Parco Area delle Scienze 53/a, I-43124, Parma, Italy
Dipartimento SMFI
Università di Parma
Parco Area delle Scienze 53/a, I-43124, Parma, Italy
email:[email protected]