1 Introduction

In this note, we study regularity properties of local minimizers of integral functionals

$$\begin{aligned} {\mathcal {F}}[u]:=\int _\varOmega f(D u)\,dx, \end{aligned}$$
(1)

where \(\varOmega \subset {\mathbb {R}}^n\), \(n\ge 3\), is a bounded domain, \(u:\varOmega \rightarrow {\mathbb {R}}^N\), \(N\ge 1\) and \(f:{\mathbb {R}}^{N\times n}\rightarrow {\mathbb {R}}\) is a sufficiently smooth integrand satisfying (pq)-growth of the form

Assumption 1

There exist \(0<\nu \le L<\infty \) such that \(f\in C^2({\mathbb {R}}^{N\times n})\) satisfies for all \(z,\xi \in {\mathbb {R}}^{N\times n}\)

$$\begin{aligned} {\left\{ \begin{array}{ll} \nu |z|^p\le f(z)\le L(1+|z|^q),&{}\\ \nu |z|^{p-2}|\xi |^2\le \langle \partial ^2 f(z)\xi ,\xi \rangle \le L(1+|z|^2)^\frac{q-2}{2}|\xi |^2. \end{array}\right. } \end{aligned}$$
(2)

Regularity properties of local minimizers of (1) in the case \(p=q\) are classical, see, e.g., [24]. A systematic regularity theory in the case \(p<q\) was initiated by Marcellini in [27, 28], see [31] for an overview (for a more up-to-date overview see the introduction in [30]). In particular, Marcellini [29] proves (among other things):

  1. (A)

    If \(N=1\), \(2\le p<q\) and \(\frac{q}{p}<1+\frac{2}{n}\), then every local minimizer \(u\in W_\mathrm{loc}^{1,p}(\varOmega )\) of (1) satisfies \(u\in W_\mathrm{loc}^{1,\infty }(\varOmega )\).

Local boundedness of the gradient implies that the non-standard growth of f and \(\partial ^2f\) in (1) becomes irrelevant and higher regularity (depending on the smoothness of f) follows by standard arguments, see e.g. [27, Chapter 7].

Only very recently, Bella and the author improved in [6] the result (A) in the sense that ’n’ in the assumption on the ratio \(\frac{q}{p}\) can be replaced by ’\(n-1\)’ for \(n\ge 3\) (to be precise, [6] considers the non-degenerate version (4) of (2)). The argument in [6] relies on scalar techniques, e.g., Moser-iteration type arguments, and thus cannot be extended to the vectorial case \(N>1\).

For the vectorial case \(N>1\), Esposito, Leonetti and Mingione showed in [18] that

  1. (B)

    If \(2\le p<q\) and \(\frac{q}{p}<1+\frac{2}{n}\), then every local minimizer \(u\in W_\mathrm{loc}^{1,p}(\varOmega ,{\mathbb {R}}^N)\) of (1) satisfies \(u\in W_\mathrm{loc}^{1,q}(\varOmega ,{\mathbb {R}}^N)\).

To the best of our knowledge, there is no improvement of (B) with respect to the relation between the exponents pq and the dimension n available in the literature. Here we provide such an improvement for \(n\ge 3\).

Before we state the results, we recall a standard notion of local minimizer in the context of functionals with (pq)-growth

Definition 1

We call \(u\in W_\mathrm{loc}^{1,1}(\varOmega )\) a local minimizer of \({\mathcal {F}}\) given in (1) iff

$$\begin{aligned} f(Du)\in L^1_\mathrm{loc}(\varOmega ) \end{aligned}$$

and

$$\begin{aligned} \int _{\mathrm{supp}\,\varphi }f(D u)\,dx\le \int _{\mathrm{supp}\,\varphi }f(D u+D \varphi )\,dx \end{aligned}$$

for any \(\varphi \in W^{1,1}(\varOmega ,{\mathbb {R}}^N)\) satisfying \(\mathrm{supp}\;\varphi \Subset \varOmega \).

The main result of the present paper is

Theorem 2

Let \(\varOmega \subset {\mathbb {R}}^n\), \(n\ge 3\), and suppose Assumption 1 is satisfied with \(2\le p<q<\infty \) such that

$$\begin{aligned} \frac{q}{p}<1+\frac{2}{n-1}. \end{aligned}$$
(3)

Let \(u\in W_\mathrm{loc}^{1,1}(\varOmega ,{\mathbb {R}}^N)\) be a local minimizer of the functional \({\mathcal {F}}\) given in (1). Then, \(u\in W_\mathrm{loc}^{1,q}(\varOmega ,{\mathbb {R}}^N)\).

Higher gradient integrability is a first step in the regularity theory for integral functionals with (pq)-growth, see [7, 11, 19, 20] for further higher integrability results under (pq)-conditions. Clearly, we cannot expect to improve from \(W_\mathrm{loc}^{1,q}\) to \(W_\mathrm{loc}^{1,\infty }\) for \(N>1\), since this even fails in the classic setting \(p=q\), see [34]. Direct consequences of Theorem 2 are higher differentiability and a further improvement in gradient integrability in the form:

  1. (i)

    (Higher differentiability). In the situation of Theorem 2 it holds \(|\nabla u|^\frac{p-2}{2} \nabla u\in W^{1,2}_\mathrm{loc}(\varOmega )\), see Theorem 5.

  2. (ii)

    (Higher integrability). Sobolev inequality and (i) imply \(\nabla u\in L^{\kappa p}_\mathrm{loc}(\varOmega ,{\mathbb {R}}^{N\times n})\) with \(\kappa =\frac{n}{n-2}\). Note that \(\kappa p>q\) provided \(\frac{q}{p}<1+\frac{2}{n-2}\).

A further, on first glance less direct, consequence of Theorem 2 is partial regularity of minimizers of (1), see, e.g.,  [1, 7, 10, 32], for partial regularity results under (pq)-conditons. For this, we slightly strengthen the assumptions on the integrand and suppose

Assumption 3

There exist \(0<\nu \le L<\infty \) such that \(f\in C^2({\mathbb {R}}^{N\times n})\) satisfies for all \(z,\xi \in {\mathbb {R}}^{N\times n}\)

$$\begin{aligned} {\left\{ \begin{array}{ll} \nu |z|^p\le f(z)\le L(1+|z|^q),&{}\\ \nu (1+|z|^2)^\frac{p-2}{2}|\xi |^2\le \langle \partial ^2 f(z)\xi ,\xi \rangle \le L(1+|z|^2)^\frac{q-2}{2}|\xi |^2. \end{array}\right. } \end{aligned}$$
(4)

In [7], Bildhauer and Fuchs prove partial regularity under Assumption 3 with \(\frac{q}{p}<1+\frac{2}{n}\) ( [7] contains also more general conditions including, e.g., the subquadratic case). Here we show

Theorem 4

Let \(\varOmega \subset {\mathbb {R}}^n\), \(n\ge 3\), and suppose Assumption 3 is satisfied with \(2\le p<q<\infty \) such that (3). Let \(u\in W_\mathrm{loc}^{1,1}(\varOmega ,{\mathbb {R}}^N)\) be a local minimizer of the functional \({\mathcal {F}}\) given in (1). Then, there exists an open set \(\varOmega _0\subset \varOmega \) with \(|\varOmega \setminus \varOmega _0|=0\) such that \(\nabla u\in C^{0,\alpha }(\varOmega _0,{\mathbb {R}}^{N\times n})\) for each \(0<\alpha <1\).

We do not know if (3) in Theorems 2 and 4 is optimal. Classic counterexamples in the scalar case \(N=1\), see, e.g., [23, 28], show that local boundedness of minimizers can fail if \(\frac{q}{p}\) is to large depending on the dimension n. In fact, [28, Theorem 6.1] and the recent boundedness result [26] show that \(\frac{1}{p}-\frac{1}{q}\le \frac{1}{n-1}\) is the sharp condition ensuring local boundedness in the scalar case \(N=1\) (for sharp results under additional structure assumptions, see, e.g., [14, 22]).

For non-autonomous functionals, i.e., \(\int _\varOmega f(x,Du)\,dx\), rather precise sufficiently & necessary conditions are established in [20], where the conditions on pq and n has to be balanced with the (Hölder)-regularity in space of the integrand. However, if the integrand is sufficiently smooth in space, the regularity theory in the non-autonomous case essentially coincides with the autonomous case, see [10]. Currently, regularity theory for non-autonomous integrands with non-standard growth, e.g. p(x)-Laplacian or double phase functionals are a very active field of research, see, e.g., [2, 12, 13, 15,16,17, 25, 33].

Coming back to autonomous integral functionals: In [11] higher gradient integrability is proven assuming so-called ’natural’ growth conditions, i.e., no upper bound assumption on \(\partial ^2f\), under the relation \(\frac{q}{p}<1+\frac{1}{n-1}\). Moreover, in two dimensions we cannot improve the previous results on higher differentiability and partial regularity of, e.g., [7, 18], see [8] for a full regularity result under Assumption 3 with \(n=2\) and \(\frac{q}{p}<2\). Finally, we mention the recent paper [3] where optimal Lipschitz-estimates with respect to a right-hand side are proven for functionals with (pq)-growth.

Let us briefly describe the main idea in the proof of Theorem 2 and from where our improvement compared to earlier results comes from. The main point is to obtain suitable a priori estimates for minimizers that may already be in \(W_\mathrm{loc}^{1,q}(\varOmega ,{\mathbb {R}}^N)\). The claim then follows by a known regularization and approximation procedure, see, e.g., [18]. For minimizers \(v\in W_\mathrm{loc}^{1,q}(\varOmega ,{\mathbb {R}}^N)\) a Caccioppoli-type inequality

$$\begin{aligned} \int \eta ^2 |D(|Dv|^\frac{p-2}{2}Dv)|^2\lesssim \int |\nabla \eta |^2(1+|Dv|^q) \end{aligned}$$
(5)

is valid for all sufficiently smooth cut-off functions \(\eta \), see Lemma 1. Very formally, the Caccioppoli inequality (5) can be combined with Sobolev inequality and a simple interpolation inequality to obtain

$$\begin{aligned} \Vert Dv\Vert _{L^{\kappa p}}^p\lesssim \Vert D(|Dv|^\frac{p-2}{2}Dv)\Vert _{L^2}^2\lesssim \Vert Dv\Vert _{L^q}^q\lesssim \Vert Dv\Vert _{L^{\kappa p}}^{q\theta }\Vert Dv\Vert _{L^{p}}^{(1-\theta )q}, \end{aligned}$$

where \(\theta =\frac{\frac{1}{p}-\frac{1}{q}}{\frac{1}{p}-\frac{1}{\kappa p}}\in (0,1)\) and \(\kappa =\frac{n}{n-2}\). The \(\Vert Dv\Vert _{L^{\kappa p}}\)-factor on the right-hand side can be absorbed provided we have \(\frac{q\theta }{p}<1\), but this is precisely the ’old’ (pq)-condition \(\frac{q}{p}<1+\frac{2}{n}\), this type of argument was previously rigorously implemented in, e.g., [7, 19]. Our improvement comes from choosing a cut-of function \(\eta \) in (5) that is optimized with respect to v, which enables us to use Sobolev inequality on \(n-1\)-dimensional spheres wich gives the desired improvement, see Sect. 3. This idea has its origin in joint works with Bella [4, 5] on linear non-uniformly elliptic equations.

With Theorem 2 at hand, we can follows the arguments of [7] almost verbatim to prove Theorem 4. In Sect. 4, we sketch (following [7]) a corresponding \(\varepsilon \)-regularity result from which Theorem 4 follows by standard methods.

2 Preliminary results

In this section, we gather some known facts. We begin with a well-known higher differentiability result for minimizers of (1) under the assumption that \(u\in W^{1,q}_\mathrm{loc}(\varOmega ,{\mathbb {R}}^N)\):

Lemma 1

Let \(\varOmega \subset {\mathbb {R}}^n\), \(n\ge 2\), and suppose Assumption 1 is satisfied with \(2\le p<q<\infty \). Let \(v\in W_\mathrm{loc}^{1,q}(\varOmega ,{\mathbb {R}}^N)\) be a local minimizer of the functional \({\mathcal {F}}\) given in (1). Then, \(|Dv|^\frac{p-2}{2}Dv\in W_\mathrm{loc}^{1,2}(\varOmega ,{\mathbb {R}}^{N\times n})\) and there exists \(c=c(\frac{L}{\nu },n,N,p,q)\in [1,\infty )\) such that for every \(Q\in {\mathbb {R}}^{N\times n}\) and every \(\eta \in C_c^1(\varOmega )\)

$$\begin{aligned} \int _\varOmega \eta ^2|D(|Dv|^\frac{p-2}{2}Dv)|^2\,dx \le c\int _{\varOmega }(1+|Dv|^2)^\frac{q-2}{2} |D v-Q|^2|\nabla \eta |^2\,dx. \end{aligned}$$
(6)

The Lemma 1 is known, see e.g. [7, 18, 28]. Since we did not find a precise reference for estimate (6), we included a prove here following essentially the argument of [18].

Proof of Lemma 1

Without loss of generality, we suppose \(\nu =1\) the general case \(\nu >0\) follows by replacing f with \(f/\nu \) (and thus L with \(L/\nu \)). Throughout the proof, we write \(\lesssim \) if \(\le \) holds up to a multiplicative constant depending only on nNp and q.

Thanks to the assumption \(v\in W^{1,q}_\mathrm{loc}(\varOmega ,{\mathbb {R}}^N)\), the minimizer v satisfies the Euler-Largrange equation

$$\begin{aligned} \int _\varOmega \langle \partial f(D v), D \varphi \rangle \,dx=0\qquad \text{ for } \text{ all } \varphi \in W_0^{1,q}(\varOmega ,{\mathbb {R}}^N) \end{aligned}$$
(7)

(for this we use that the convexity and growth conditions of f imply \(|\partial f(z)|\le c(1+|z|^{q-1})\) for some \(c=c(L,n,N,q,)<\infty \)). Next, we use the difference quotient method, to differentiate the above equation: For \(s\in \{1,\dots ,n\}\), we consider the difference quotient operator

$$\begin{aligned} \tau _{s,h}v:=\tfrac{1}{h}(v(\cdot +he_s)-v)\qquad \text{ where } v\in L_\mathrm{loc}^1({\mathbb {R}}^n,{\mathbb {R}}^N). \end{aligned}$$

Fix \(\eta \in C_c^1(\varOmega )\). Testing (7) with \(\varphi :=\tau _{s,-h}(\eta ^2(\tau _{s,h}(v-\ell _Q)))\in W_0^{1,q}(\varOmega )\), where \(\ell _Q(x)=Qx\), we obtain

$$\begin{aligned} (I):=&\int _\varOmega \eta ^2 \langle \tau _{s,h}\partial f(D v),\tau _{s,h}D v\rangle \,dx\\ =&-2\int _\varOmega \eta \langle \tau _{s,h}\partial f(D v), \tau _{s,h}(v-\ell _Q)\otimes \nabla \eta \rangle \,dx=:(II). \end{aligned}$$

Writing \(\tau _{s,h}\partial f(D v)=\frac{1}{h} \partial f(D v+th\tau _{s,h}D v)\big |_{t=0}^{t=1}\), the fundamental theorem of calculus yields

$$\begin{aligned}&\int _\varOmega \int _0^1 \eta ^2\langle \partial ^2f(D v+th\tau _{s,h}D v))\tau _{s,h}D v,\tau _{s,h}D v\rangle \,dt\,dx=(I)\nonumber \\ =&(II)=-2\int _\varOmega \int _0^1 \eta \langle \partial ^2f(D v+th\tau _{s,h}D v)\tau _{s,h}D v, (\tau _{s,h}v-Qe_s) \otimes \nabla \eta \rangle \,dt\,dx, \end{aligned}$$
(8)

where we use \(\tau _{h,s}\ell _Q=Qe_s\). Youngs inequality yields

$$\begin{aligned} |(II)|\le \tfrac{1}{2}(I)+2(III), \end{aligned}$$
(9)

where

$$\begin{aligned} (III):=\int _\varOmega \int _0^1\langle \partial ^2f(D u+th\tau _{s,h}D u) (\tau _{s,h}v-Qe_s)\otimes \nabla \eta ,(\tau _{s,h}v-Qe_s)\otimes \nabla \eta \rangle \,dt\,dx. \end{aligned}$$

Combining (8), (9) with the assumptions on \(\partial ^2f\), see (2), with the elementary estimate

$$\begin{aligned} |\tau _{s,h}(|Dv|^{\frac{p-2}{2}}Dv)|^2\lesssim \int _0^1 |D v+th\tau _{s,h}D v|^\frac{p-2}{2} |\tau _{s,h}D v|^2\,dt\ \end{aligned}$$

for \(h>0\) sufficiently small (see e.g. [18, Lemma 3.4]), we obtain

$$\begin{aligned}&\int _\varOmega \eta ^2|\tau _{s,h}(|Dv|^{\frac{p-2}{2}}Dv)|^2\,dx\nonumber \\ \lesssim&\int _\varOmega \int _0^1 \eta ^2|D v+th\tau _{s,h}D v|^\frac{p-2}{2} |\tau _{s,h}D v|^2\,dt\,dx\le (I)\nonumber \\ \le&4(III)\le 4L\int _\varOmega \int _0^1 (1+|Dv+th\tau _{s,h}D v|^{q-2})|\nabla \eta |^2|\tau _{s,h}v-Qe_s|^2\,dt\,dx. \end{aligned}$$
(10)

Estimate (10), the fact \(v\in W_\mathrm{loc}^{1,q}(\varOmega )\) and the arbitrariness of \(\eta \in C_c^1(\varOmega )\) and \(s\in \{1,\dots ,n\}\) yield \(|Dv|^{\frac{p-2}{2}}Dv\in W_\mathrm{loc}^{1,2}(\varOmega )\). Sending h to zero in (10), we obtain

$$\begin{aligned} \int _\varOmega \eta ^2|\partial _s(|Dv|^{\frac{p-2}{2}}Dv)|^2\,dx\lesssim L\int _\varOmega (1+|Dv|^{q-2})|\nabla \eta |^2|\partial _sv-Qe_s|^2\,dx \end{aligned}$$

the desired estimate (6) follows by summing over s. \(\square \)

Next, we state a higher differentiability result under the more restrictive Assumption 3 which will be used in the proof of Theorem 4.

Lemma 2

Let \(\varOmega \subset {\mathbb {R}}^n\), \(n\ge 2\), and suppose Assumption 3 is satisfied with \(2\le p<q<\infty \). Let \(v\in W_\mathrm{loc}^{1,q}(\varOmega ,{\mathbb {R}}^N)\) be a local minimizer of the functional \({\mathcal {F}}\) given in (1). Then, \(h:=(1+|Dv|^2)^\frac{p}{4}\in W_\mathrm{loc}^{1,2}(\varOmega )\) and there exists \(c=c(\frac{L}{\nu },n,N,p,q)\in [1,\infty )\) such that for every \(Q\in {\mathbb {R}}^{N\times n}\)

$$\begin{aligned} \int _\varOmega \eta ^2|\nabla h|^2\,dx \le c\int _{\varOmega }(1+|Dv|^2)^\frac{q-2}{2} |D v-Q|^2|\nabla \eta |^2\,dx\quad \text{ for } \text{ all } \eta \in C_c^1(\varOmega ). \end{aligned}$$
(11)

A variation of Lemma 2 can be found in [7] and we only sketch the proof.

Proof of Lemma 2

With the same argument as in the proof of Lemma 1 but using (4) instead of (2), we obtain \(v\in W^{2,2}_\mathrm{loc}(\varOmega ,{\mathbb {R}}^N)\) and the Caccioppoli inequality

$$\begin{aligned} \int _\varOmega \eta ^2(1+|Dv|^2)^\frac{p-2}{2}|D^2v|^2\,dx \le c\int _{\varOmega }(1+|Dv|^2)^\frac{q-2}{2} |D v-Q|^2|\nabla \eta |^2\,dx \end{aligned}$$
(12)

for all \(\eta \in C_c^1(\varOmega )\), where \(c=c(\frac{L}{\nu },n,N,p,q)<\infty \). Formally, the chain-rule implies

$$\begin{aligned} |\nabla h|^2\le c(1+|Dv|^2)^\frac{p-2}{2}|D^2v|^2, \end{aligned}$$
(13)

where \(c=c(n,p)<\infty \), and the claimed estimate (11) follows from (12) and (13). In general, we are not allowed to use the chain rule, but the above reasoning can be made rigorous: Consider a truncated version \(h_m\) of h, where \(h_m:=\varTheta _m(|Dv|)\) with

$$\begin{aligned} \varTheta _m(t):={\left\{ \begin{array}{ll}(1+t^2)^\frac{p}{4}&{}\text{ if } 0\le t\le m\\ (1+m^2)^\frac{p}{4}&{}\text{ if } t\ge m\end{array}\right. }. \end{aligned}$$

For \(h_m\) we are allowed to use the chain-rule and (12) together with (13) with h replaced by \(h_m\) imply (11) with h replaced by \(h_m\). The claimed estimate follows by taking the limit \(m\rightarrow \infty \), see [7, Proposition 3.2] for details. \(\square \)

The following technical lemma is contained in [6] (see also [4, proof of Lemma 2.1, Step 1]) and plays a key role in the proof of Theorem 2

Lemma 3

( [6], Lemma 3) Fix \(n\ge 2\). For given \(0<\rho<\sigma <\infty \) and \(v\in L^1(B_{\sigma })\), consider

$$\begin{aligned} J(\rho ,\sigma ,v):=\inf \left\{ \int _{B_{\sigma }}|v||\nabla \eta |^2\,dx \;|\;\eta \in C_0^1(B_{\sigma }),\,\eta \ge 0,\,\eta =1 \text{ in } B_\rho \right\} . \end{aligned}$$

Then for every \(\delta \in (0,1]\)

$$\begin{aligned} J(\rho ,\sigma ,v)\le (\sigma -\rho )^{-(1+\frac{1}{\delta })} \biggl (\int _{\rho }^\sigma \left( \int _{\partial B_r} |v|\,d{\mathcal {H}}^{n-1}\right) ^\delta \,dr\biggr )^\frac{1}{\delta }. \end{aligned}$$
(14)

For convenience of the reader we include a short proof of Lemma 3

Proof of Lemma 3

Estimate (14) follows directly by minimizing among radial symmetric cut-off functions. Indeed, we obviously have for every \(\varepsilon \ge 0\)

$$\begin{aligned}&J(\rho ,\sigma ,v)\\ \le&\inf \left\{ \int _{\rho }^\sigma \eta '(r)^2\left( \int _{\partial B_r}|v|\,d{\mathcal {H}}^{n-1}+\varepsilon \right) \,dr \;|\;\eta \in C^1(\rho ,\sigma ),\,\eta (\rho )=1,\,\eta (\sigma )=0\right\} \\ =:&J_{\mathrm{1d},\varepsilon }. \end{aligned}$$

For \(\varepsilon >0\), the one-dimensional minimization problem \(J_{\mathrm{1d},\varepsilon }\) can be solved explicitly and we obtain

$$\begin{aligned} J_{\mathrm{1d},\varepsilon }=\biggl (\int _{\rho }^\sigma \biggl (\int _{\partial B_r}|v|\,d{\mathcal {H}}^{n-1}+\varepsilon \biggr )^{-1}\,dr\biggr )^{-1}. \end{aligned}$$
(15)

To see (15), we observe that using the assumption \(v\in L^1(B_\sigma )\) and a simple approximation argument we can replace \(\eta \in C^1(\rho ,\sigma )\) with \(\eta \in W^{1,\infty }(\rho ,\sigma )\) in the definition of \(J_{\mathrm{1d},\varepsilon }\). Let \({{\widetilde{\eta }}}:[\rho ,\sigma ]\rightarrow [0,\infty )\) be given by

$$\begin{aligned} {{\widetilde{\eta }}}(r):=1-\biggl (\int _\rho ^\sigma b(r)^{-1}\,dr\biggr )^{-1}\int _{\rho }^rb(r)^{-1}\,dr,\quad \text{ where } b(r):=\int _{\partial B_r}|v|+\varepsilon . \end{aligned}$$

Clearly, \({{\widetilde{\eta }}}\in W^{1,\infty }(\rho ,\sigma )\) (since \(b\ge \varepsilon >0\)), \({{\widetilde{\eta }}}(\rho )=1\), \({{\widetilde{\eta }}}(\sigma )=0\), and thus

$$\begin{aligned} J_{\mathrm{1d},\varepsilon }\le \int _{\rho }^\sigma {{\widetilde{\eta }}}'(r)^2b(r)\,dr=\biggl (\int _{\rho }^\sigma b(r)^{-1}\,dr\biggr )^{-1}. \end{aligned}$$

The reverse inequality follows by Hölder’s inequality. Next, we deduce (14) from (15): For every \(s>1\), we obtain by Hölder inequality \(\sigma -\rho =\int _\rho ^\sigma (\frac{b}{b})^\frac{s-1}{s}\le \left( \int _\rho ^\sigma b^{s-1}\right) ^\frac{1}{s}\left( \int _\rho ^\sigma \frac{1}{b}\right) ^\frac{s-1}{s}\) with b as above, and by (15) that

$$\begin{aligned} J_{\mathrm{1d},\varepsilon }\le (\sigma -\rho )^{-\frac{s}{s-1}}\biggl (\int _{\rho }^\sigma \left( \int _{\partial B_r}|v|+\varepsilon \right) ^{s-1}\,dr\biggr )^{\frac{1}{s-1}}. \end{aligned}$$

Sending \(\varepsilon \) to zero, we obtain (14) with \(\delta =s-1>0\). \(\square \)

3 Higher integrability - Proof of Theorem 2

In this section, we prove the following higher integrability and differentiability result which clearly contains Theorem 2

Theorem 5

Let \(\varOmega \subset {\mathbb {R}}^n\), \(n\ge 2\), and suppose Assumption 1 is satisfied with \(2\le p<q<\infty \) such that \(\frac{q}{p}<1+\min \{\frac{2}{n-1},1\}\). Let \(u\in W_\mathrm{loc}^{1,1}(\varOmega ,{\mathbb {R}}^N)\) be a local minimizer of the functional \({\mathcal {F}}\) given in (1). Then, \(u\in W_\mathrm{loc}^{1,q}(\varOmega ,{\mathbb {R}}^N)\) and \(|Du|^\frac{p-2}{2}Du\in W_\mathrm{loc}^{1,2}(\varOmega ,{\mathbb {R}}^{N\times n})\). Moreover, for

$$\begin{aligned} \chi =\frac{n-1}{n-3}\quad \text{ if } n\ge 4\quad \chi \in (\frac{1}{2-\frac{q}{p}},\infty )\quad \text{ if } n=3 \text{ and } \quad \chi :=\infty \quad \text{ if } n=2. \end{aligned}$$
(16)

there exists \(c=c(\frac{L}{\nu },n,N,p,q,\chi )\in [1,\infty )\) such that for every \(B_{R}(x_0)\Subset \varOmega \)

(17)

where

$$\begin{aligned} \alpha :=\frac{1-\frac{q}{\chi p}}{2-\frac{q}{p}-\frac{1}{\chi }}. \end{aligned}$$
(18)

Proof of Theorem 5

Without loss of generality, we suppose \(\nu =1\) the general case \(\nu >0\) follows by replacing f with \(f/\nu \). Throughout the proof, we write \(\lesssim \) if \(\le \) holds up to a multiplicative constant depending only on LnNp and q.

Following, e.g., [7, 18, 19], we consider the perturbed integral functionals

$$\begin{aligned} {\mathcal {F}}_\lambda (w):=\int _\varOmega f_\lambda (Dw)\,dx,\qquad \text{ where }\quad f_\lambda (z):=f(z)+\lambda |z|^q\quad \text{ with } \lambda \in (0,1). \end{aligned}$$
(19)

We then derive suitable a priori higher differentiability and integrability estimates for local minimizers of \({\mathcal {F}}_\lambda \) that are independent of \(\lambda \in (0,1)\). The claim then follows with help of a by now standard double approximation procedure in spirit of [18].

Step 1. One-step improvement.

Let \(v\in W^{1,1}_\mathrm{loc}(\varOmega ,{\mathbb {R}}^N)\) be a local minimizer of the functional \({\mathcal {F}}_\lambda \) defined in (19), \(B_1\Subset \varOmega \), and let \(\chi >1\) be defined in (16). We claim that there exists \(c=c(L,n,N,p,q,\chi )\in [1,\infty )\) such that for all \(\frac{1}{2}\le \rho <\sigma \le 1\) and every \(\lambda \in (0,1]\)

$$\begin{aligned}&\int _{B_1}1+f_\lambda (Dv)+\int _{B_\rho }|D(|Dv|^\frac{p-2}{2}Dv)|^2\,\,dx\nonumber \\ \le&\frac{c\biggl (\int _{B_1}1+f_\lambda (Dv)\biggr )^{\frac{\chi }{\chi -1}(1-\frac{q}{\chi p})}}{(\sigma -\rho )^{1+\frac{q}{p}}}\nonumber \\&\times \biggl (\int _{B_1}1+f_\lambda (Dv)+\int _{B_{\sigma }}|D(|Dv|^\frac{p-2}{2}Dv)|^2\,\,dx\biggr )^{\frac{\chi }{\chi -1}(\frac{q}{p}-1)} \end{aligned}$$
(20)

with the understanding \(\frac{\infty }{\infty -1}=1\) and

$$\begin{aligned} \int _{B_\rho }|D(|Dv|^\frac{p-2}{2}Dv)|^2\,\,dx \lesssim \frac{1}{(\sigma -\rho )^2}\frac{1}{\lambda }\int _{B_{\sigma }}1+ f_\lambda (Dv)\,dx. \end{aligned}$$
(21)

The growth conditions of \(f_\lambda \) and the minimality of v imply \(v\in W^{1,q}_\mathrm{loc}(\varOmega ,{\mathbb {R}}^N)\) and thus by Lemma 1

$$\begin{aligned} \int _\varOmega |D(|Dv|^\frac{p-2}{2}Dv)|^2\eta ^2\,\,dx \lesssim \int _{\varOmega }(1+|Dv|^2)^\frac{q-2}{2} |D v|^2|\nabla \eta |^2\,dx \end{aligned}$$
(22)

for all \(\eta \in C_c^1(\varOmega )\). Estimate (21) follows directly from (22) for \(\eta \in C_c^1(B_\sigma )\) with \(0\le \eta \le 1\), \(\eta \equiv 1\) on \(B_\rho \) and \(|\nabla \eta |\le \frac{2}{\sigma -\rho }\), combined with \(|z|^q\le \frac{1}{\lambda }f_\lambda (z)\) and \(\lambda \in (0,1]\).

Hence, it is left to show (20). For this, we use a technical estimate which follows from Lemma 3 and Hölders inequality: For given \(0<\rho<\sigma <\infty \) and \(w\in L^q(B_{\sigma })\) it holds

$$\begin{aligned} J(\rho ,\sigma ,|w|^q)\le \frac{\biggl (\int _{B_{\sigma }\setminus B_\rho } |w|^p\biggr )^{\frac{\chi }{\chi -1}(1-\frac{q}{\chi p})}}{(\sigma -\rho )^{1+\frac{q}{p}}} \biggl (\int _{\rho }^{\sigma }\Vert w\Vert _{L^{\chi p}(\partial B_r)}^p\,dr\biggr )^{\frac{\chi }{\chi -1}(\frac{q}{p}-1)}, \end{aligned}$$
(23)

where J is defined as in Lemma 3. We postpone the derivation of (23) to the end of this step.

Combining (22) with \((1+|Dv|^2)^\frac{q-2}{2} |D v|^2\le (1+|Dv|)^q\) and estimate (23) with \(w=1+|Dv|\), we obtain

$$\begin{aligned}&\int _{B_\rho }|D(|Dv|^\frac{p-2}{2}Dv)|^2\,\,dx\nonumber \\ \lesssim&\frac{\biggl (\int _{B_{\sigma }\setminus B_\rho }(1+|D v|)^p\,dx\biggr )^{\frac{\chi }{\chi -1}(1-\frac{q}{\chi p})}}{(\sigma -\rho )^{1+\frac{q}{p}}}\biggl (\int _{\rho }^{\sigma }\Vert 1+|Dv|\Vert _{L^{\chi p}(\partial B_r)}^p\,dr\biggr )^{\frac{\chi }{\chi -1}(\frac{q}{p}-1)}. \end{aligned}$$
(24)

Next, we use the Sobolev inequality on spheres to estimate the second factor on the right-hand side in (24): For \(n\ge 2\) there exists \(c=c(n,N,\chi )\in [1,\infty )\) such that for all \(r>0\)

$$\begin{aligned}&\Vert D v\Vert _{L^{\chi p}(\partial B_r)}^p\nonumber \\ \le&c r^{(n-1)(\frac{1}{\chi }-1)}\biggl (\int _{\partial B_r}|D v|^p\,d{\mathcal {H}}^{n-1}+r^2\int _{\partial B_r}|D(|Dv|^{\frac{p-2}{2}}Dv)|^2\,d{\mathcal {H}}^{n-1}\biggr ). \end{aligned}$$
(25)

Combining (25) with elementary estimates and assumption \(\frac{1}{2}\le \rho <\sigma \le 1\), we obtain

$$\begin{aligned} \int _{\rho }^{\sigma }\Vert 1+|Dv|\Vert _{L^{\chi p}(\partial B_r)}^p\,dr\lesssim&\int _{\rho }^{\sigma }1+\Vert Dv\Vert _{L^{\chi p}(\partial B_r)}^p\,dr\nonumber \\ \lesssim&\int _{\rho }^{\sigma }1+\biggl (\int _{\partial B_r}|D v|^p+|D(|Dv|^{\frac{p-2}{2}}Dv)|^2\,d{\mathcal {H}}^{n-1}\biggr )\,dr\nonumber \\ \lesssim&\int _{B_{\sigma }\setminus B_\rho }1+|D v|^p+|D(|Dv|^{\frac{p-2}{2}}Dv)|^2\,dx. \end{aligned}$$
(26)

Combining (24) and estimate (26), we obtain

$$\begin{aligned}&\int _{B_\rho }|D(|Dv|^\frac{p-2}{2}Dv)|^2\,\,dx\\ \le&\frac{c\biggl (\int _{B_{1}}(1+|D v|)^p\,dx\biggr )^{\frac{\chi }{\chi -1}(1-\frac{q}{\chi p})}}{(\sigma -\rho )^{1+\frac{q}{p}}}\biggl (\int _{B_{\sigma }}1+|D v|^p+|D(|Dv|^{\frac{p-2}{2}}Dv)|^2\,dx\biggr )^{\frac{\chi }{\chi -1}(\frac{q}{p}-1)}, \end{aligned}$$

The claimed estimate (20) now follows since \(|z|^p\le f(z)\le f_\lambda (z)\), \(\frac{\chi }{\chi -1}(1-\frac{q}{\chi p}+\frac{q}{p}-1)=\frac{q}{p}\ge 1\) and \(\int _{B_1}1+f_\lambda (Dv)\,dx\ge |B_1|\).

Finally, we present the computations regarding (23): Lemma 3 yields

$$\begin{aligned} J(\sigma ,\rho ,|w|^q)\le \frac{\biggl (\int _{\rho }^{\sigma }\Vert w\Vert _{L^q(\partial B_r)}^{q\delta }\,dr\biggr )^\frac{1}{\delta }}{(\sigma -\rho )^{1+\frac{1}{\delta }}}\qquad \text{ for } \text{ every } \delta >0. \end{aligned}$$

Using two times the Hölder inequality, we estimate

$$\begin{aligned} \biggl (\int _{\rho }^{\sigma }\Vert w\Vert _{L^q(\partial B_r)}^{q\delta }\,dr\biggr )^\frac{1}{\delta }\le&\biggl (\int _{\rho }^{\sigma }\Vert w\Vert _{L^p(\partial B_r)}^{\theta q\delta }\Vert w\Vert _{L^{\chi p}(\partial B_r)}^{(1-\theta ) q\delta }\,dr\biggr )^\frac{1}{\delta }\quad \text{ where } \frac{\theta }{p}+\frac{1-\theta }{\chi p}=\frac{1}{q}\\ \le&\biggl (\int _{\rho }^{\sigma }\Vert w\Vert _{L^p(\partial B_r)}^{\theta q\delta \frac{s}{s-1}}\,dr\biggr )^\frac{s-1}{s\delta }\biggl (\int _{\rho }^{\sigma }\Vert w\Vert _{L^{\chi p}(\partial B_r)}^{(1-\theta ) q\delta s}\,dr\biggr )^\frac{1}{\delta s}\quad \text{ for } \text{ every } s>1. \end{aligned}$$

Inequality (23) follows with the admissible choice

$$\begin{aligned} \delta =\frac{p}{q}\quad \text{ and }\quad s=\frac{1}{1-\theta }\qquad \biggl (\text{ recall } 1-\theta =\frac{\frac{1}{p}-\frac{1}{q}}{\frac{1}{p}-\frac{1}{\chi p}}\quad \text{ and } p<q\biggr ) \end{aligned}$$

which ensures \(\theta q\delta \frac{s}{s-1}=(1-\theta ) q\delta s=p\).

Step 2. Iteration.

We claim that there exists \(c=c(L,n,N,p,q,\chi )\in [1,\infty )\) such that

$$\begin{aligned} \int _{B_\frac{1}{2}}|Dv|^p+|D(|Dv|^\frac{p-2}{2}Dv)|^2\,\,dx\le&c\biggl (\int _{B_1}1+f_\lambda (Dv)\,dx\biggr )^\alpha , \end{aligned}$$
(27)

where \(\alpha \) is defined in (18). For \(k\in {\mathbb {N}}\cup \{0\}\), we set

$$\begin{aligned} \rho _k=\frac{3}{4}- \frac{1}{4^{1+k}}\quad \text{ and }\quad J_k:=\int _{B_1}1+f_\lambda (Dv)+\int _{B_{\rho _k}}|D(|Dv|^\frac{p-2}{2}Dv)|^2\,\,dx. \end{aligned}$$

Estimate (21) and the choice of \(\rho _k\) imply for \(\lambda \in (0,1]\)

$$\begin{aligned} \sup _{k\in {\mathbb {N}}}J_k\le \int _{B_1}1+f_\lambda (Dv)+\int _{B_{\frac{3}{4}}}|D(|Dv|^\frac{p-2}{2}Dv)|^2\,\,dx\lesssim \frac{1}{\lambda }\int _{B_1}1+f_\lambda (Dv)\,dx<\infty .\nonumber \\ \end{aligned}$$
(28)

From (20) we deduce the existence of \(c=c(L,n,N,p,q,\chi )\in [1,\infty )\) such that for every \(k\in {\mathbb {N}}\)

$$\begin{aligned} J_{k-1}\le c 4^{(1+\frac{q}{p}) k}\biggl (\int _{B_1}1+f_\lambda (Dv)\biggr )^{\frac{\chi }{\chi -1}(1-\frac{q}{\chi p})}J_k^{\frac{\chi }{\chi -1}\frac{q-p}{p}}. \end{aligned}$$
(29)

Assumption \(\frac{q}{p}<1+\min \{1,\frac{2}{n-1}\}\) and the choice of \(\chi \) yield

$$\begin{aligned} {\frac{\chi }{\chi -1}\frac{q-p}{p}}{\mathop {=}\limits ^{(16)}}{\left\{ \begin{array}{ll}\frac{q}{p}-1&{}\text{ if } n=2\\ {\frac{\chi }{\chi -1}\frac{q-p}{p}}&{}\text{ if } n=3\\ \frac{n-1}{2}(\frac{q}{p}-1)&{}\text{ if } n\ge 4\end{array}\right. }<1, \end{aligned}$$

where we use for \(n=3\) that \(\chi {\mathop {>}\limits ^{(16)}}\frac{1}{2-\frac{q}{p}}>0\) and

$$\begin{aligned} \frac{\chi }{\chi -1}\frac{q-p}{p}<1\quad \Leftrightarrow \quad \frac{q-p}{p}<1-\frac{1}{\chi }\quad \Leftrightarrow \quad \frac{1}{\chi }<2-\frac{q}{p}. \end{aligned}$$

Hence, iterating (29) we obtain (using the uniform bound (28) on \(J_k\) and \({\frac{\chi }{\chi -1}\frac{q-p}{p}}<1\))

$$\begin{aligned}&\int _{B_{\frac{1}{2}}}|Dv|^p+|D(|Dv|^\frac{p-2}{2}Dv)|^2\,\,dx\le J_0\lesssim \biggl (\int _{B_1}1+f_\lambda (Dv)\biggr )^{\frac{\chi }{\chi -1}(1-\frac{q}{\chi p})\sum _{k=0}^\infty ({\frac{\chi }{\chi -1}\frac{q-p}{p}})^k}\nonumber \\ \end{aligned}$$
(30)

and the claimed estimate (27) follow from

$$\begin{aligned} \alpha = \frac{\chi }{\chi -1}(1-\frac{q}{\chi p})\sum _{k=0}^\infty ({\frac{\chi }{\chi -1}\frac{q-p}{p}})^k. \end{aligned}$$

Step 3. Conclusion.

We assume \(B_1\Subset \varOmega \) and show that there exists \(c=c(L,n,N,p,q,\chi )\in [1,\infty )\)

$$\begin{aligned} \int _{B_\frac{1}{8}}|D u|^{q}\,dx\le c\left( \int _{B_1}1+f(D u)\,dx\right) ^{\frac{\alpha q}{p}}, \end{aligned}$$
(31)

where \(\alpha \) is given as in (18) above. Clearly, standard scaling, translation and covering arguments yield

for all \(B_R(x_0)\Subset \varOmega \) and \(c=c(L,n,N,p,q,\chi )\in [1,\infty )\). The claimed estimate (17) then follows from Lemma 1.

Following [18], we introduce in addition to \(\lambda \in (0,1)\) a second small parameter \(\varepsilon >0\) which is related to a suitable regularization of u. For \(\varepsilon \in (0,\varepsilon _0)\), where \(0<\varepsilon _0\le 1\) is such that \(B_{1+\varepsilon _0}\Subset \varOmega \), we set \(u_\varepsilon :=u*\varphi _\varepsilon \) with \(\varphi _\varepsilon :=\varepsilon ^{-n}\varphi (\frac{\cdot }{\varepsilon })\) and \(\varphi \) being a non-negative, radially symmetric mollifier, i.e. it satisfies

$$\begin{aligned} \varphi \ge 0,\quad \mathrm{supp}\; \varphi \subset B_1,\quad \int _{{\mathbb {R}}^n}\varphi (x)\,dx=1,\quad \varphi (\cdot )={{\widetilde{\varphi }}}(|\cdot |)\quad \text{ for } \text{ some } {{\widetilde{\varphi }}}\in C^\infty ({\mathbb {R}}). \end{aligned}$$

Given \(\varepsilon ,\lambda \in (0,\varepsilon _0)\), we denote by \(v_{\varepsilon ,\lambda }\in u_\varepsilon +W_0^{1,q}(B_1)\) the unique function satisfying

$$\begin{aligned} \int _{B_1}f_\lambda (D v_{\varepsilon ,\lambda })\,dx\le \int _{B_1}f_\lambda (D v)\,dx\qquad \text{ for } \text{ all } v\in u_\varepsilon +W_0^{1,q}(B_1). \end{aligned}$$
(32)

Combining Sobolev inequality with the assumption \(\frac{q}{p}<1+\frac{2}{n-2}\) and estimate (27), we have

$$\begin{aligned} \biggl (\int _{B_\frac{1}{8}}|D v_{\varepsilon ,\lambda }|^{q}\,dx\biggr )^\frac{p}{q}&\lesssim \int _{B_\frac{1}{8}}|Dv_{\varepsilon ,\lambda }|^p+|D(|Dv_{\varepsilon ,\lambda }|^\frac{p-2}{2}Dv_{\varepsilon ,\lambda })|^2\,\,dx\nonumber \\&{\mathop {\lesssim }\limits ^{(27)}}\biggl (\int _{B_1}1+f_\lambda (Dv_{\varepsilon ,\lambda })\,dx\biggr )^\alpha \nonumber \\&{\mathop {\le }\limits ^{(19),(32)}}\left( \int _{B_{1}}1+f(D u_\varepsilon )+\lambda |D u_\varepsilon |^q\,dx\right) ^\alpha \nonumber \\&\le \left( |B_1|+\int _{B_{1+\varepsilon }}f(D u)\,dx+\lambda \int _{B_1}|D u_\varepsilon |^q\,dx\right) ^\alpha , \end{aligned}$$
(33)

where we used Jensen’s inequality and the convexity of f in the last step. Similarly,

$$\begin{aligned} \int _{B_1}|D v_{\varepsilon ,\lambda }|^p\,dx{\mathop {\le }\limits ^{(2)}}&\int _{B_1}f(D v_{\varepsilon ,\lambda })\,dx{\mathop {\le }\limits ^{(19)(32)}} \int _{B_1}f(D u_\varepsilon )+\lambda |D u_\varepsilon |^q\,dx\nonumber \\ \le&\int _{B_{1+\varepsilon }}f(D u)\,dx+\lambda \int _{B_1}|D u_\varepsilon |^q\,dx. \end{aligned}$$
(34)

Fix \(\varepsilon \in (0,\varepsilon _0)\). In view of (33) and (34), we find \(w_\varepsilon \in u_\varepsilon +W_0^{1,p}(B_1)\) such that as \(\lambda \rightarrow 0\), up to subsequence,

$$\begin{aligned} v_{\varepsilon ,\lambda }\rightharpoonup w_\varepsilon \qquad \text{ weakly } \text{ in } W^{1,p}(B_1),\\ D v_{\varepsilon ,\lambda }\rightharpoonup Dw_\varepsilon \qquad \text{ weakly } \text{ in } L^{q}(B_\frac{1}{8}). \end{aligned}$$

Hence, a combination of (33), (34) with the weak lower-semicontinuity of convex functionals yield

$$\begin{aligned} \Vert D w_\varepsilon \Vert _{L^{q}(B_\frac{1}{8})}\le&\liminf _{\lambda \rightarrow 0}\Vert D v_{\varepsilon ,\lambda }\Vert _{L^{\kappa p}(B_\frac{1}{8})}\lesssim \left( \int _{B_{1+\varepsilon }}f(D u)\,dx+1\right) ^\frac{\alpha }{p} \end{aligned}$$
(35)
$$\begin{aligned} \int _{B_1}|D w_\varepsilon |^p\,dx\le&\int _{B_{1}}f(D w_\varepsilon )\,dx\le \int _{B_{1+\varepsilon }}f(D u)\,dx. \end{aligned}$$
(36)

Since \(w_\varepsilon \in u_\varepsilon +W_0^{1,q}(B_1)\) and \(u_\varepsilon \rightarrow u\) in \(W^{1,p}(B_1)\), we find by (36) a function \(w\in u+W_0^{1,p}(B_1)\) such that, up to subsequence,

$$\begin{aligned} D w_{\varepsilon }\rightharpoonup D w\quad \text{ weakly } \text{ in } L^p(B_1). \end{aligned}$$

Appealing to the bounds (35), (36) and lower semicontinuity, we obtain

$$\begin{aligned} \Vert D w\Vert _{L^{q}(B_\frac{1}{8})}\lesssim&\left( \int _{B_{1}}f(D u)\,dx+1\right) ^\frac{\alpha }{p} \end{aligned}$$
(37)
$$\begin{aligned} \int _{B_{1}}f(D w)\,dx\le&\int _{B_{1}}f(D u)\,dx. \end{aligned}$$
(38)

Inequality (38), strict convexity of f and the fact \(w\in u+W_0^{1,p}(B_1)\) imply \(w=u\) and thus the claimed estimate (31) is a consequence of (37). \(\square \)

4 Partial regularity - Proof of Theorem 4

Theorem 4 follows from, the higher integrability statement Theorem 2, the \(\varepsilon \)-regularity statement of Lemma 4 below and a well-known iteration argument.

Lemma 4

Let \(\varOmega \subset {\mathbb {R}}^n\), \(n\ge 3\), and suppose Assumption 3 is satisfied with \(2\le p<q<\infty \) such that \(\frac{q}{p}<1+\frac{2}{n-1}\). Fix \(M>0\). There exists \(C^*=C^*(n,N,p,q,\frac{L}{\nu },M)\in [1,\infty )\) such that for every \(\tau \in (0,\frac{1}{4})\) there exists \(\varepsilon =\varepsilon (M,\tau )>0\) such that the following is true: Let \(u\in W_\mathrm{loc}^{1,1}(\varOmega ,{\mathbb {R}}^N)\) be a local minimizer of the functional \({\mathcal {F}}\) given in (1). Suppose for some ball \(B_r(x)\Subset \varOmega \)

$$\begin{aligned} |(Du)_{x,r}|\le M, \end{aligned}$$

where we use the shorthand , and

then

$$\begin{aligned} E(x,\tau r)\le C^*\tau ^2E(x,r). \end{aligned}$$

With the higher integrability of Theorem 5 and the Caccioppoli inequality of Lemma 2 at hand, we can prove Lemma 4 following almost verbatim the proof of the corresponding result [7, Lemma 4.1], which contain the statement of Lemma 4 under the assumption \(\frac{q}{p}<1+\frac{2}{n}\) (note that in [7] somewhat more general growth conditions including also the case \(1<p<q\) are considered). Thus, we only sketch the argument.

Proof of Lemma 4

Fix \(M>0\). Suppose that Lemma 4 is wrong. Then there exists \(\tau \in (0,\frac{1}{4})\), a local minimizer \(u\in W^{1,1}_\mathrm{loc}(\varOmega ,{\mathbb {R}}^N)\), which in view of Theorem 2 satisfies \(u\in W^{1,q}_\mathrm{loc}(\varOmega ,{\mathbb {R}}^N)\), and a sequence of balls \(B_{r_m}(x_m)\Subset B_R\) satisfying

$$\begin{aligned}&|(Du)_{x_m,r_m}|\le M,\quad E(x_m,r_m)=:\lambda _m\quad \text{ with }\quad \lim _{m\rightarrow \infty }\lambda _m=0, \end{aligned}$$
(39)
$$\begin{aligned}&E(x_m,\tau r_m)>C^*\tau ^2\lambda _m^2, \end{aligned}$$
(40)

where \(C^*\) is chosen below. We consider the sequence of rescaled functions given by

$$\begin{aligned}&v_m(z):=\frac{1}{\lambda _m r_m}(u(x_m+r_mz)-a_m-r_mA_mz), \end{aligned}$$

where \(a_m:=(u)_{x_m,r_m}\) and \(A_m:=(Du)_{x_m,r_m}\). Assumption (39) implies \(\sup _m|A_m|\le M\) and thus, up to subsequence,

$$\begin{aligned} A_m\rightarrow A\in {\mathbb {R}}^{N\times n}. \end{aligned}$$

The definition of \(v_m\) yields

$$\begin{aligned} Dv_m(z)=\lambda _m^{-1}(Du(x_m+r_mz)-A_m),\quad (v_m)_{0,1}=0,\quad (Dv_m)_{0,1}=0 \end{aligned}$$
(41)

Assumptions (39) and (40) imply

(42)
(43)

The bound (42) together with (41) imply the existence of \(v\in W^{1,2}(B_1,{\mathbb {R}}^N)\) such that, up to extracting a further subsequence,

$$\begin{aligned}&v_m\rightharpoonup v&\text{ in } W^{1,2}(B_1,{\mathbb {R}}^{ N}),\\&\lambda _m Dv_m\rightarrow 0&\text{ in } L^2(B_1,{\mathbb {R}}^{N\times n}) \text{ and } \text{ almost } \text{ everywhere }\\&\lambda _m^{1-\frac{2}{q}}v_m\rightharpoonup 0\quad&\text{ in } W^{1,q}(B_1,{\mathbb {R}}^N). \end{aligned}$$

The function v satisfies the linear equation with constant coefficients

$$\begin{aligned} \int _{B_1}\langle \partial ^2 f(A) Dv,D\varphi \rangle \,dz=0\qquad \text{ for } \text{ all } \varphi \in C^1_0(B_1), \end{aligned}$$

see, e.g., [21] or [7, Proposition 4.2]. Standard estimates for linear elliptic systems with constant coefficients imply \(v\in C_\mathrm{loc}^\infty (B_1,{\mathbb {R}}^N)\) and existence of \(C^{**}<\infty \) depending only on nN and the ellipticity contrast of \(\partial ^2 f(A)\) (and thus on \(\frac{L}{\nu },p,q,\) and M) such that

(44)

Choosing \(C^*=2C^{**}\) we obtain a contradiction between (43) and (44) provided we have as \(m\rightarrow \infty \)

$$\begin{aligned}&Dv_m\rightarrow Dv\quad&\text{ in } L^2_\mathrm{loc}(B_1),\end{aligned}$$
(45)
$$\begin{aligned}&\lambda _m^{1-\frac{2}{q}}Dv_m\rightarrow 0\quad&\text{ in } L^q_\mathrm{loc}(B_1). \end{aligned}$$
(46)

Exanctly as in [7, Proposition 4.3] (with \(\mu =2-p\), see also [9, Section 3.4.3.2] for a more detailed presentation of the proof), we have for all \(\rho \in (0,1)\),

$$\begin{aligned} \lim _{m\rightarrow \infty }\int _{B_\rho }\int _0^1(1-s)\biggl (1+|A_m+\lambda _m(Dv+sDw_m)|^2\biggr )^\frac{p-2}{2}|Dw_m|^2\,dz=0, \end{aligned}$$
(47)

where \(w:=v_m-v\), and thus the local \(L^2\)-convergence (45) follows. It is left to prove (46). For this, we introduce for \(\rho \in (0,1)\) and \(T>0\) the sequence of subsets

$$\begin{aligned} U_m:=U_m(\rho ,T):=\{\,z\in B_\rho \,:\,\lambda _m|Dv_m|\le T\,\}. \end{aligned}$$

The local Lipschitz regularity of v, \(q>2\) and (45) imply for all \(\rho \in (0,1)\) and \(T>0\)

$$\begin{aligned} \limsup _{m\rightarrow \infty }\int _{U_m(\rho ,T)}\lambda _m^{q-2}|Dv_m|^q\,dz\lesssim&\limsup _{m\rightarrow \infty }\int _{U_m(\rho ,T)}\lambda _m^{q-2}|Dw_m|^q\,dz\\ \lesssim&\limsup _{m\rightarrow \infty }\int _{B_\rho }(M^{q-2}+\lambda _m^{q-2}|Dv|^{q-2})|Dw_m|^2\,dz\\ =&0, \end{aligned}$$

where here and for the rest of the proof \(\lesssim \) means \(\le \) up to a multiplicative constant depending only on LnNp and q. Hence, it is left to show that there exists \(T>0\) such that

$$\begin{aligned} \limsup _{m\rightarrow \infty }\int _{B_\rho \setminus U_m(\rho ,T)}\lambda _m^{q-2}|Dv_m|^q\,dz\le 0\quad \text{ for } \text{ all } \rho \in (0,1). \end{aligned}$$

As in [7], we introduce a sequence of auxiliary functions

$$\begin{aligned} \psi _m:=\lambda _m^{-1}\biggl [(1+|A_m+\lambda _m Dv_m|^2)^\frac{p}{4}-(1+|A_m|^2)^{\frac{p}{4}}\biggr ], \end{aligned}$$

which satisfy

$$\begin{aligned} \limsup _{m\rightarrow \infty }\Vert \psi _m\Vert _{W^{1,2}(B_\rho )}\lesssim c(\rho )\in [1,\infty )\qquad \text{ for } \text{ all } \rho \in (0,1). \end{aligned}$$
(48)

Indeed, by Theorem 2 and Lemma 2, we have for every \(\rho \in (0,1)\) and every \(Q\in {\mathbb {R}}^{N\times n}\)

$$\begin{aligned} \int _{B_{\rho r_m}(x_m)}|\nabla (1+|D u(x)|^2)^\frac{p}{4}|^2\,dx\lesssim r_m^{-2}c(\rho )\int _{B_{r_m}(x_m)}(1+|\nabla u(x)|)^{q-2}|Du(x)-Q|^2\,dx \end{aligned}$$

and thus by rescaling and setting \(Q=A_m\)

$$\begin{aligned} \int _{B_\rho }|\nabla \psi _m|^2\,dz\lesssim c(\rho )\int _{B_1}(1+|A|^{q-2}+|\lambda _m Dv_m|^{q-2}))|Dv_m|^2\,dz{\mathop {\lesssim }\limits ^{(42)}} c(\rho )(1+M^{q-2}). \end{aligned}$$

The identity \(\psi _m=\lambda _m^{-1}\int _0^1\frac{d}{dt}\varTheta (A_m+t\lambda _mv_m)\,dt\) with \(\varTheta (F):=(1+|F|^2)^\frac{p}{4}\) implies

$$\begin{aligned} |\psi _m|\le c(|Dv_m|+\lambda _m^{\frac{p-2}{2}}|Dv_m|^{\frac{p}{2}}) \end{aligned}$$

(see [7, p. 555] for details) and thus with help of (47), we obtain

$$\begin{aligned} \limsup _{m\rightarrow \infty }\int _{B_\rho }|\psi _m|^2\,dz\lesssim c(\rho ). \end{aligned}$$

For T sufficiently large (depending on M) there exists \(c>0\) such that for all \(z\in B_\rho \setminus U_m(\rho ,T)\)

$$\begin{aligned} \psi _m(z)\ge c\lambda _m^{-1}\lambda _m^{\frac{p}{2}}|Dv_m(z)|^\frac{p}{2}\quad \text{ and } \text{ thus } \quad \lambda _m^{2(1+\frac{q}{p})}\psi _m^{\frac{2q}{p}}(z)\ge c^{\frac{2q}{p}}\lambda _m^{q-2}|Dv_m(z)|^q \end{aligned}$$

Estimate (48) and Sobolev embedding imply \(\limsup _{m\rightarrow \infty }\Vert \psi _m\Vert _{L^{\frac{2n}{n-2}}(B_\rho )}\lesssim c(\rho )\in [1,\infty )\). Hence, using assumption \(\frac{q}{p}<1+\frac{2}{n-1}\) (and thus \(\frac{2q}{p}<\frac{2n}{n-2}\)), we obtain for every \(\rho \in (0,1)\)

$$\begin{aligned} \limsup _{m\rightarrow \infty }\int _{B_\rho \setminus U_m(\rho ,T)}\lambda _m^{q-2}|Dv_m|^q\,dz\lesssim \lambda _m^{2(1+\frac{q}{p})}\int _{B_\rho }\psi _m^{\frac{2q}{p}}(z)\,dz\lesssim c(\rho )\limsup _{m\rightarrow \infty }\lambda _m^{2(1+\frac{q}{p})}=0, \end{aligned}$$

which finishes the proof. \(\square \)