1 Introduction

In the main part of our note we consider splitting-type energies as a particular class of variational integrals

$$\begin{aligned} J[u] := \int _{\Omega }f(\nabla u) \, \mathrm{d}x\end{aligned}$$
(1)

defined for functions u: \({\mathbb {R}}^n \supset \Omega \rightarrow {\mathbb {R}}\). In the splitting case, the energy density f: \({\mathbb {R}}^n \rightarrow {\mathbb {R}}\) admits an additive decomposition into different parts depending on the various first partial derivatives of the admissible functions u, for example, we can consider the case

$$\begin{aligned} f(\nabla u) = \sum _{i=1}^n f_i(\partial _i u) \end{aligned}$$
(2)

with functions \(f_i\): \({\mathbb {R}}\rightarrow {\mathbb {R}}\), \(i=1\), ..., n, of possibly different growth rates.

This also leads us to the more general category of variational problems with non-standard growth, where in the simplest case one studies an energy functional J defined in equation (1) with convex density f being bounded from above and from below by different powers of \(|\nabla u|\).

For an overview of the various aspects of variational problems with non-standard growth including aspects as existence and regularity we refer the interested reader, e.g., to [1,2,3,4,5,6,7,8] and to the recent paper [9] together with the references quoted therein.

Regarding the question of interior regularity of J-minimizers in the non-standard setting, the above references provide rather complete answers, while the problem of boundary regularity for the solution of

$$\begin{aligned} J[u] \rightarrow \min \, ,\qquad u=u_0 \quad \hbox {on}\; \partial \Omega \, , \end{aligned}$$
(3)

with sufficiently smooth boundary datum \(u_0\) seems to be not so well investigated.

We mention the contribution [10], where the higher integrability of the minimizer u up to the boundary is established for densities f being in some sense close to the splitting-class defined in (2). A precise formulation of the assumptions concerning the density f is given in inequality (2.1) of the paper [10], and we also like to stress that a survey of the known global regularity results is provided in the introduction of this reference.

Even more recently the contribution of Koch [11] addresses the global higher integrability of the gradient of solutions of variational problems with (pq)-growth, which to our knowledge is the first result to improve global integrabilty of the gradient allowing full anisotropy with different growth rates with respect to different partial derivatives. The boundary data are supposed to belong to some fractional order spaces (see, e.g., [12] or [13]) and roughly speaking are handled like an additional x-dependence following ideas as outlined in, e.g., [14] for the interior situation. Hence, the admissible energy densities correspond to the examples presented in [14]. On the other hand the case of vectorial functions u: \({\mathbb {R}}^n \supset \Omega \rightarrow {\mathbb {R}}^m\), \(n\ge 2\), \(m \ge 1\), as well as the subquadratic situation are included.

In our note we first discuss the variational problem (3) on a bounded Lipschitz domain \(\Omega \subset {\mathbb {R}}^2\) with energy functional J defined in equation (1) under the following assumptions on the data: for \(i=1\), 2 let \(f_i \in C^2({\mathbb {R}})\) satisfy

$$\begin{aligned} a_i (1+t^2)^{\frac{q_i-2}{2}} \le f_i''(t) \le A_i (1+t^2)^{\frac{q_i -2}{2}}\, , \qquad t\in {\mathbb {R}}\, , \end{aligned}$$
(4)

with constants \(a_i\), \(A_i >0\) and exponents

$$\begin{aligned} q_i > 2 \, . \end{aligned}$$
(5)

We define the energy density f: \({\mathbb {R}}^2 \rightarrow {\mathbb {R}}\) according to (recall the “splitting condition” (2))

$$\begin{aligned} f(Z) := f_1(z_1) + f_2(z_2)\, ,\qquad Z = (z_1,z_2) \in {\mathbb {R}}^2\, . \end{aligned}$$
(6)

Note that (4) yields the anisotropic ellipticity condition

$$\begin{aligned} c_1 (1+|Z|^2)^{\frac{p-2}{2}}|Y|^2 \le D^2f(Z)(Y,Y) \le c_2 (1+|Z|^2)^{\frac{q-2}{2}}|Y|^2 \end{aligned}$$
(7)

valid for Z, \(Y \in {\mathbb {R}}^2\) with positive constants \(c_1\), \(c_2\) and for the choice of exponents \(p=2\) and \(q:= \max \{q_1,q_2\}\). In fact, (7) is a direct consequence of (5) and the formula

$$\begin{aligned} D^2f(Z)(Y,Y) = \sum _{i=1}^2 f_i''(z_i)|y_i|^2\, ,\qquad Z, \, Y \in {\mathbb {R}}^2 \, . \end{aligned}$$
(8)

Finally, we fix a functiom

$$\begin{aligned} u_0\in W^{1,\infty }(\Omega ) \end{aligned}$$
(9)

remarking that actually \(u_0 \in W^{1,t}(\Omega )\) with exponent t being sufficiently large (depending on \(q_1\) and \(q_2\)) is needed in our calculations. For a definition of the Sobolev spaces \(W^{k,r}(\Omega )\) and their local variants we refer to [12].

The function \(u_0\) acts as a prescribed boundary datum in the minimization problem

$$\begin{aligned} J[u] := \int _{\Omega }f(\nabla u) \, \mathrm{d}x\rightarrow \min \qquad \hbox {in}\; u_0+W^{1,1}_0(\Omega )\, . \end{aligned}$$
(10)

Now we state our result in the splitting case by the way establishing global higher integrability without relating \(q_1\) and \(q_2\) and under quite weak assumptions on \(u_0\) which means that we do not have to impose additional hypotheses concerning the regularity of the trace of \(u_0\) in the sense that \({u_0}_{|\partial \Omega }\) belongs to some fractional Sobolev space. The price we have to pay is a weight function which has to be incorporated.

Theorem 1.1

Let the assumptions (4)–(6) and (9) hold. Moreover, let \(u \in u_0 +W^{1,1}_0(\Omega )\) denote the unique solution of problem (10). Then it holds

$$\begin{aligned} |u-u_0|^t \Big (\big |\partial _1 u\big |^{\frac{3}{2}q_1} + \big |\partial _2 u\big |^{\frac{3}{2}q_2}\Big ) \in L^1(\Omega ) \end{aligned}$$
(11)

for any choice of the exponent t such that

$$\begin{aligned} t > T(q_1,q_2) \end{aligned}$$
(12)

with a finite lower bound \(T(q_1,q_2) \ge 1\) depending on the growth rates \(q_1\), \(q_2\).

Remark 1.1

An admissible choice for the quantity \(T(q_1,q_2)\) from (12) is

$$\begin{aligned} T(q_1,q_2) = 6 \max \Bigg \{1, \frac{1}{2} \frac{{q_{\max }}}{{q_{\min }}}\frac{{q_{\min }}-1}{{q_{\min }}-2}, \frac{{q_{\min }}}{{q_{\min }}-2} - \frac{1}{6}\Bigg \} \frac{{q_{\min }}}{{q_{\min }}-2} \, . \end{aligned}$$
(13)

Here and in the following we let

$$\begin{aligned} {q_{\min }}:= \min \{q_1,q_2\}\, , \qquad {q_{\max }}:= \max \{q_1,q_2\}\, . \end{aligned}$$

In the case \(q_i >5\), \(i=1\), 2, we can replace (13) by the expression

$$\begin{aligned} T(q_1,q_2) = 6 {q_{\max }}\max \Bigg \{ \frac{1}{2{q_{\min }}-4},\frac{1}{{q_{\max }}+2}\Bigg \} \end{aligned}$$
(14)

which means that in the case of higher initial integrability we can decrease the exponent t in (11) leading to a stronger result. We also note that for the choice \(q_i > 5\), \(i=1\), 2, the condition (9) can be replaced by the requirement \(u_0 \in W^{1,{q_{\min }}}(\Omega )\) together with \(\partial _i u_0 \in L^{q_i}(\Omega )\), \(i=1\), 2.

Remark 1.2

The existence and the uniqueness of a solution u to problem (10) easily follow from the growth properties of f and its strict convexity combined with our assumption (9) on the boundary datum \(u_0\). Obviously the minimizer satisfies \(\partial _i u\in L^{q_i}(\Omega )\), \(i=1\), 2, and from Theorem 1.1 ii) in [8] we deduce interior Hölder continuity of the first partial derivatives. This in turn implies that u is in the space \(W^{2,2}_{{\text {loc}}}(\Omega )\).

Remark 1.3

Of course we expect that our theorem can be extended to the case \(n>2\), i.e. to densities

$$\begin{aligned} f(Z) = \sum _{i=1}^n f_i(z_i)\, \qquad Z \in {\mathbb {R}}^n \, , \end{aligned}$$

with functions \(f_i\) satisfying (4) and (5) being replaced by \(q_i > n\), \(i=1\), ..., n. We restrict our considerations in the splitting case to the 2D-situation for two reasons: first, a generalization to the case \(n \ge 2\) would be accompanied by a tremendous technical effort without promising a deeper insight. Secondly, the interior regularity results of [8] as the starting point for our arguments would have to be generalized to the full splitting case in n dimensions which goes beyond the aim of this note.

Remark 1.4

Inequality (11) can be seen as weighted global higher integrability result for \(\nabla u\). In fact, on account of (5), u is Hölder continuous up to the boundary, which means \(|u(x) - u_0(x)| \approx {\text {dist}}^\alpha (x,\partial \Omega )\) for points x near \(\partial \Omega\). Thus, in this vague sense, Theorem 1.1 yields \(|\partial _i u|^{t_i} {\text {dist}}^\beta (\cdot , \partial \Omega ) \in L^1(\Omega )\) for a certain range of exponents \(\beta >0\) and \(t_i > q_i\).

Another interpretation will be given in Corollary 1.1 on the behaviour of a suitable Hölder coefficient.

In the second part of our paper we drop the splitting condition (6) and look at densities f with anisotropic (p,q)-growth in the sense of (7) and for a moment we still restrict our considerations to the two-dimensional case. Then the arguments of the first part can be carried over provided that p and q are not too far apart.

Theorem 1.2

Assume that f: \({\mathbb {R}}^2 \rightarrow {\mathbb {R}}\) is of class \(C^2({\mathbb {R}}^2)\) satisfying (7) with exponents \(2 < p \le q\) such that

$$\begin{aligned} q < \min \Big \{p+\frac{1}{2},2p-2\Big \} \, . \end{aligned}$$
(15)

Let (9) hold for the boundary datum \(u_0\) and let \(u \in u_0 + W^{1,1}_0(\Omega )\) denote the unique solution of problem (10). Then we have (\(\Gamma := 1 + |\nabla u|^2\))

$$\begin{aligned} |u-u_0|^{2\kappa } \Gamma ^s \in L^1(\Omega ) \end{aligned}$$

for any exponents \(\kappa\), s such that

$$\begin{aligned} s\in \Big (\frac{q}{2}+\frac{3}{4},p-\frac{1}{4}\Big ) \end{aligned}$$
(16)

and

$$\begin{aligned} \kappa > \max \Bigg \{\frac{p}{p-2},\frac{s}{s- \big (\frac{q}{2}+\frac{3}{4}\Big )}, \frac{2s(p-1)}{(p-2)^2}\Bigg \} \, . \end{aligned}$$
(17)

Remark 1.5

Since f is strictly convex and since we have (9), existence and uniqueness of a solution u to problem (10) are immediate. Note that (15) implies (4.2) in Marcellini’s paper [15], hence \(u \in W^{1,q}_{{\text {loc}}}(\Omega )\) on account of Theorem 4.1 in [15]. Quoting Corollary 2.2 from this reference we deduce \(u \in C^{1,\alpha }(\Omega )\) and \(u \in W^{2,2}_{{\text {loc}}}(\Omega )\).

In our final theorem we incorporate two new features: we modify the idea of the proofs of the previous results by starting with a more subtle inequality being applied to terms on the left-hand sides of our foregoing calculations by the way extending the range of admissible exponents p and q. Moreover, we include the general case of n-dimensions, \(n \ge 2\). This can be done without technical efforts comparable to the ones which are needed to generalize the 2D splitting case.

Now the precise formulation of our last result is as follows:

Theorem 1.3

Consider the variational problem (10) on a bounded Lipschitz domain \(\Omega \subset {\mathbb {R}}^n\) now just assuming (p,q)-growth of the \(C^2\)-density f: \({\mathbb {R}}^n \rightarrow {\mathbb {R}}\) in the sense of (7) without splitting structure. Let the boundary datum \(u_0\) satisfy (9). Suppose that

$$\begin{aligned} 2 \le n< p \le q \qquad \hbox {and}\qquad q < p + \frac{2 (p-n)}{n}\, . \end{aligned}$$
(18)

By assumption (18) we may choose a real number \(\kappa > 1\) sufficiently large such that

$$\begin{aligned} q < p + \frac{2(p-n)}{n} - \frac{2p-1}{n \kappa } \, . \end{aligned}$$
(19)

Moreover, let

$$\begin{aligned} {\overline{s}}< p \frac{\kappa (n+2)-2}{2n\kappa } - \frac{\kappa -1}{2n\kappa } \, . \end{aligned}$$
(20)

Increasing \(\kappa\), if necessary, we suppose in addition that

$$\begin{aligned} {\overline{s}}< (\kappa -1) \frac{p-n}{n} + \frac{p}{2} \end{aligned}$$
(21)

and that we have the inequalities (103) below. Then it holds

$$\begin{aligned} \int _{\Omega }\Gamma ^{{\overline{s}}} |u-u_0|^{2(\kappa -1)} \, \mathrm{d}x\le c \, , \qquad \Gamma := 1+|\nabla u|^2\, , \end{aligned}$$

for a finite constant c depending on n, p, q, \(\kappa\) and \({\overline{s}}\).

Remark 1.6

Passing to the limit \(\kappa \rightarrow \infty\), (20) becomes

$$\begin{aligned} {\overline{s}}\le \frac{p}{2} + \frac{2p-1}{2n}\, . \end{aligned}$$

Remark 1.7

We note that Remark 1.5 extends to the setting studied in Theorem 1.3, since now condition (4.2) from [15] is a consequence of the more restrictive inequality (18).

Our paper is organized as follows: in Section 2 we present the proof of Theorem 1.1 for a given Lipschitz boundary datum \(u_0\) (recall (9)) and just supposing (5). In the case \(q_1\), \(q_2 > 5\) the lower bound \(T(q_1,q_2)\) occurring in (12) looks much nicer and (9) can be considerably weakened. We sketch some arguments in Section 3. The two-dimensional non-splitting case, i.e. the setting described in Theorem 1.2, is discussed in Section 4 and in Section 5 we discuss the higher dimensional situation of Theorem 1.3.

During the proofs of our results we make essential use of Caccioppoli-type inequalities. As a main new feature these inequalities involve “small weights”, for instance weight functions like \((1+|\partial _i u|)^\alpha\), \(i=1\), 2, with a certain range of negative exponents \(\alpha\). A rather general analysis of these tools is presented in Section 6.

Before we get into all these technical details we finish the introduction by adding some extra comments concerning the interpretation of our results. For simplicity we restrict ourselves to the setting described in Theorem 1.1 and denote by u the unique solution of problem (10) under the hypothesis of Theorem 1.1.

We start with a local observation: by adding a distance function if necessary, we find an appropriate representative of the boundary data respecting a lower bound for \(|u-u_0|\) in a neighbourhood of a fixed \(x_0 \in \Omega\).

Proposition 1.1

Fix \(x_0 \in \Omega\). Then there exists a neighbourhood \(U=U(x_0)\Subset \Omega\) of \(x_0\) such that we may assume that for all \(x \in U\)

$$\begin{aligned} |{\hat{u}}|(x) = |u-u_0| (x) \ge {\underline{d}} := \frac{1}{2} d(x_0) \, , \qquad {\hat{u}} := u-u_0\, , \end{aligned}$$

where

$$\begin{aligned} d(x) = {\text {dist}}(x,\partial \Omega ) \, , \end{aligned}$$

denotes the distance function to the boundary \(\partial \Omega\).

Proof of Proposition 1.1

The distance function d is discussed, e.g., in [16], Appendix 14.6. In particular, if \(\Omega\) is a domain in \({\mathbb {R}}^n\) with non-empty boundary, then by (14.91) of [16] d is uniformly Lipschitz continuous.

Let us suppose w.l.o.g. that \({\hat{u}}(x_0) \ge 0\). If \({\hat{u}}(x_0) < d(x_0)\), then we choose \({\tilde{u}}_0 := u_0-d\) as admissible Lipschitz boundary datum with \(u_0 = {\tilde{u}}_0\) on \(\partial \Omega\). Hence both \(u_0\) and \({\tilde{u}}_0\) produce the same minimizer u of the Dirichlet problem (10) and by definition we have

$$\begin{aligned} (u-{\tilde{u}}_0)(x_0) = (u - u_0 + d)(x_0) \ge d(x_0) \, . \end{aligned}$$

The considerations of our note completely remain unchanged if we replace \(u_0\) by \({\tilde{u}}_0\). Hence we have w.l.o.g. that \({\hat{u}}(x_0) \ge d(x_0)\). Proposition 1.1 then follows from the continuity of u and \(u_0\). \(\square\)

Let us fix \(x_0\in \Omega\), and define U, \({\hat{u}}\) as in Proposition 1.1. Moreover we let h: \(\Omega \rightarrow {\mathbb {R}}\), \(h:= |{\hat{u}}|^{\frac{4}{3} \frac{\kappa }{{q_{\min }}}+1}\), hence

$$\begin{aligned} |\nabla h| \le \Bigg [\frac{4}{3} \frac{\kappa }{{q_{\min }}} +1\Bigg ] |{\hat{u}}|^{\frac{4}{3}\frac{\kappa }{{q_{\min }}}}|\nabla {\hat{u}}|\, . \end{aligned}$$

Theorem 1.1 gives

$$\begin{aligned} \int _{\Omega }|\nabla h|^{\frac{3}{2}{q_{\min }}} \, \mathrm{d}x\le c\, ,\qquad {i.e.}\qquad h\in W^{1,\frac{3}{2}{q_{\min }}}(\Omega ) \, . \end{aligned}$$

Letting \(\zeta = \frac{4}{3} \frac{\kappa }{{q_{\min }}}+1\), the imbedding into Hölder spaces yields \(|{\hat{u}}|^{\zeta } \in C^{0,\mu }({\overline{\Omega }})\) with \(\mu = 1- \frac{4}{3{q_{\min }}}\).

Now we consider the Taylor expansion of the function g: \({\mathbb {R}}^+ \rightarrow {\mathbb {R}}\), \(w \mapsto w^{\zeta }\) around a fixed \({\tilde{w}} >0\): for w sufficiently close to \({\tilde{w}}\) we have

$$\begin{aligned} w^{\zeta } = {\tilde{w}}^\zeta + \zeta {\tilde{w}}^{\zeta -1} (w-{\tilde{w}}) + O(|w-{\tilde{w}}|^2) \, . \end{aligned}$$

Suppose w.l.o.g. that \({\hat{u}} > 0\) in U. Inserting \({\tilde{w}} = {\hat{u}}(x_0)\) and \(w = {\hat{u}}(x)\), x sufficiently close to \(x_0\), in the Taylor expansion we obtain using Proposition 1.1

$$\begin{aligned} \frac{|{\hat{u}}^\zeta (x) - {\hat{u}}^\zeta (x_0)|}{\zeta {\underline{d}}^{\zeta -1}}\ge & {} \frac{|{\hat{u}}^\zeta (x) - {\hat{u}}^\zeta (x_0)|}{\zeta {\hat{u}}^{\zeta -1}(x_0)}\nonumber \\\ge & {} u(x) - u(x_0) + u_0(x_0) - u_0(x) + O(|{\hat{u}}(x) - {\hat{u}}(x_0)|^2)\, . \end{aligned}$$
(22)

Since \(|{\hat{u}}|^\zeta\) is of class \(C^{0,\mu }({\overline{\Omega }})\) and since according to (9) \(u_0\) is Lipschitz, by (22) we find a constant \(c >0\) not depending on \(x_0\) such that

$$\begin{aligned} u(x)- u(x_0) \le c \frac{1}{\zeta {\underline{d}}^{\zeta -1}} |x-x_0|^\mu + O(|x-x_0|^{2\mu })\, . \end{aligned}$$
(23)

Changing the roles of x and \(x_0\) and following the arguments leading to (23), we find a constant \(c >0\) not depending on \(x_0\) such that for all x sufficiently close to \(x_0\) we have

$$\begin{aligned} |u(x) - u(x_0)| \le c \frac{1}{\zeta {\underline{d}}^{\zeta -1}}|x-x_0|^\mu \, . \end{aligned}$$
(24)

Summarizing the results we obtain the following corollary to Theorem 1.1.

Corollary 1.1

Under the assumptions of Theorem 1.1 we let

$$\begin{aligned} \zeta : = \frac{4}{3} \frac{\kappa }{{q_{\min }}}+1 \, , \qquad \mu := 1- \frac{4}{3{q_{\min }}} \, . \end{aligned}$$

If \(x_0 \in \Omega\) is fixed, then for every sufficiently small neighbourhood \(U = U(x_0)\Subset \Omega\) of \(x_0\) the Hölder coefficient of u (see [16] for notation) satisfies

$$\begin{aligned}{}[u]_{\mu ;x_0} := \sup _{U} \frac{|u(x)-u(x_0)|}{|x-x_0|^\mu } \le c \frac{1}{\zeta d^{\zeta -1}(x_0)} \, , \end{aligned}$$

where the constant c is not depending on \(x_0\).

2 Proof of Theorem 1.1

We proceed in several steps assuming from now on the validity of the hypotheses (4), (5), (6) and (9). Let u denote the unique solution of problem (10) and recall the (interior) regularity properties of u stated in Remark 1.2. We start with the following elementary observation:

Proposition 2.1

Consider numbers \(\kappa\), \(s_1\), \(s_2 \ge 1\). Then there is a finite constant c such that for \(i=1\), 2 and any \(\eta \in C^1_0(\Omega )\)

$$\begin{aligned} \int _{\Omega }\Gamma _i^{s_i} |\eta |^{2\kappa } \, \mathrm{d}x\le & {} c \Bigg \{ \Bigg [\int _{\Omega }\big | \nabla \Gamma _i^{\frac{s_i}{2}}\big |\, |\eta |^\kappa \, \mathrm{d}x\Bigg ]^2 + \Bigg [\int _{\Omega }\Gamma _i^{\frac{s_i}{2}} |\nabla \eta | |\eta |^{\kappa -1} \, \mathrm{d}x\Bigg ]^2\Bigg \}\nonumber \\ & \quad =: {} c \big \{ (I)_i +(II)_i\big \} \, ,\qquad \Gamma _i := 1+|\partial _i u|^2\, . \end{aligned}$$
(25)

Proof of Proposition 2.1

Let \(v:= \Gamma _i^{s_i/2} |\eta |^{\kappa }\) and observe that due to \(u \in W^{2,2}_{{\text {loc}}} \cap W^{1,\infty }_{{\text {loc}}}(\Omega )\) the function v is in the space \(W^{1,1}_{0}(\Omega )\), hence the Sobolev-Poincaré inequality implies

$$\begin{aligned} \int _{\Omega }v^2 \, \mathrm{d}x\le c \Bigg [ \int _{\Omega }|\nabla v|\, \mathrm{d}x\Bigg ]^2\, , \end{aligned}$$

and inequality (25) directly follows by observing that \(\big |\nabla |\eta |\big | \le |\nabla \eta |\). \(\square\)

Next we replace \(\eta\) by a sequence \(\eta _m\), \(m\in {\mathbb {N}}\), being defined through

$$\begin{aligned} \eta _m := \varphi _m (u-u_0) \end{aligned}$$
(26)

with \(\varphi _m \in C^1_0(\Omega )\), \(0 \le \varphi _m \le 1\),

$$\begin{aligned} \varphi _m = 1 \; \hbox {on}\; \Big \{x\in \Omega : \, {\text {dist}}(x,\partial \Omega ) \ge \frac{1}{m}\Big \}\, ,&\quad\varphi _m = 0\; \hbox {on}\; \Big \{x\in \Omega : \, {\text {dist}}(x,\partial \Omega ) \le \frac{1}{2m}\Big \}\, ,\nonumber \\ |\nabla \varphi _m| \le c\,m \, ,&\quad c=c(\partial \Omega )\, . \end{aligned}$$
(27)

Note that due to (5) and (9) inequality (25) extends to \(\eta _m\), moreover, according to [16], Theorem 7.17, we have

$$\begin{aligned} |u(x) - u_0(x)| \le c \rho ^{1- \frac{2}{{q_{\min }}}}\, , \qquad {q_{\min }}:= \min \{q_1,q_2\}\, , \end{aligned}$$
(28)

for points \(x\in \Omega\) at distance \(\le \rho\) to \(\partial \Omega\), provided \(\rho\) is sufficiently small.

The quantity \((II)_i\) defined in (25) with respect to the choice \(\eta = \eta _m\) behaves as follows:

Proposition 2.2

Let \(\delta := {q_{\min }}-2\) and consider numbers \(s_i\), \(\kappa \ge 1\), \(i=1\), 2, such that

$$\begin{aligned} s_i > \frac{\delta }{2}\, ,\qquad \kappa \ge {\hat{\kappa }} := \frac{{q_{\min }}-1}{{q_{\min }}-2}\, . \end{aligned}$$
(29)

Then it holds for \((II)_i = (II_m)_i\), \(i=1\), 2, defined in (25) with respect to the function \(\eta _m\) from (26)

$$\begin{aligned} (II)_i \le c \int _{\Omega }\Gamma _i^{s_i - \frac{\delta }{2}} |\eta _m|^{2 \kappa - 2 {\hat{\kappa }}} \, \mathrm{d}x=: c (III)_i \, , \qquad i=1 ,\, 2\, , \end{aligned}$$
(30)

c denoting a finite constant being uniform in m.

Proof of Proposition 2.2

By Hölder’s inequality we have

$$\begin{aligned} (II)_i= & {} \Bigg [ \int _{\Omega }\Gamma _i^{\frac{s_i}{2} - \frac{\delta }{4}} |\eta _m|^{\kappa -{{\hat{\kappa }}}} |\eta _m |^{{{\hat{\kappa }}}-1} |\nabla \eta _m| \Gamma _i^{\frac{\delta }{4}}\, \mathrm{d}x\Bigg ]^2\\\le & {} \Bigg [ \int _{\Omega }\Gamma _i^{s_i - \frac{\delta }{2}} |\eta _m|^{2\kappa - 2 {{\hat{\kappa }}}}\, \mathrm{d}x\Bigg ] \Bigg [ \int _{\Omega }\Gamma _i^{\frac{\delta }{2}} |\eta _m|^{2{{\hat{\kappa }}}-2} |\nabla \eta _m|^2 \, \mathrm{d}x\Bigg ] \end{aligned}$$

and for the second integral on the r.h.s. we observe (recall (26) and (27), (28))

$$\begin{aligned} |\eta _m|^{2{{\hat{\kappa }}}-2} |\nabla \eta _m|^2\le \, & {} c \Big [ |\nabla (u-u_0)|^2 + |\nabla \varphi _m |^2 |u-u_0|^{2{{\hat{\kappa }}}} \Big ]\nonumber \\\le & \, {} c \Big [ |\nabla u|^2 + |\nabla u_0|^2 + m^2 \Big (\frac{1}{m}\Big )^{2{{\hat{\kappa }}}(1-2/{q_{\min }})}\Big ] \, . \end{aligned}$$
(31)

Since \(\Omega\) is Lipschitz we observe that \(|{\text {spt}} \nabla \varphi _m| \le c/m\) and obtain

$$\begin{aligned} \int _{{\text {spt}} \nabla \varphi _m} \Gamma _i^{\frac{\delta }{2}} m^{2-2{{\hat{\kappa }}}(1-\frac{2}{{q_{\min }}})} \, \mathrm{d}x\le \int _{\Omega }\Gamma _i^{\frac{{q_{\min }}}{2}} \, \mathrm{d}x+ m^{[2-2{{\hat{\kappa }}}(1-\frac{2}{{q_{\min }}})]\frac{{q_{\min }}}{2}} m^{-1} \end{aligned}$$
(32)

The choice (29) of \({{\hat{\kappa }}}\) shows that the right-hand side of (32) is bounded. Letting \(\Gamma := 1 + |\nabla u|^2\) we arrive at

$$\begin{aligned} (II)_i\le \, & {} c\, (III)_i \Bigg [\int _{\Omega }\Big [ \Gamma ^{\frac{\delta }{2}}\Gamma + \Gamma ^{\frac{\delta }{2}} |\nabla u_0|^2\Big ]\, \mathrm{d}x+1 \Bigg ]\\\le & \, {} c \, (III)_i \Bigg [ \int _{\Omega }\Gamma ^{\frac{{q_{\min }}}{2}}\, \mathrm{d}x+ \int _{\Omega }|\nabla u_0|^{{q_{\min }}}\, \mathrm{d}x+1 \Bigg ] \, , \end{aligned}$$

where in the final step we have used the definition of \(\delta\). Now (30) clearly is a consequence of (9) (in fact, \(u_0 \in W^{1,{q_{\min }}}(\Omega )\) would be sufficient). \(\square\)

In the following we will discuss the quantity \((III)_i\) defined in (30) under the assumptions of Proposition 2.2.

We have by Young’s inequality for any \(\varepsilon > 0\)

$$\begin{aligned} (III)_i= & {} \int _{\Omega }\Gamma _i^{s_i \frac{s_i - \delta /2}{s_i}} |\eta _m|^{2 \kappa - 2 {\hat{\kappa }}}\, \mathrm{d}x= \int _{\Omega }\Big [ \Gamma _i^{s_i} |\eta _m|^{2\kappa }\Big ]^{\frac{s_i - \delta /2}{s_i}} |\eta _m|^{2\kappa - 2{\hat{\kappa }} - 2 \kappa \frac{s_i - \delta /2}{s_i}}\, \mathrm{d}x\\\le \, & {} \varepsilon \int _{\Omega }\Gamma _i^{s_i} |\eta _m|^{2\kappa } \, \mathrm{d}x+ c(\varepsilon ) \int _{\Omega }|\eta _m|^\vartheta \, \mathrm{d}x\end{aligned}$$

with exponent

$$\begin{aligned} \vartheta := \Big [\frac{s_i}{s_i - \delta /2}\Big ]^* \Big (2\kappa - 2 {\hat{\kappa }} - 2 \kappa \frac{s_i - \delta /2}{s_i}\Big ) \, , \end{aligned}$$

\([\dots ]^*\) denoting the exponent conjugate to \(\frac{s_i}{s_i - \delta /2}\). Note that \(\vartheta \ge 0\) provided we additionally assume that \(\kappa \ge 1\) satisfies the inequality

$$\begin{aligned} \kappa \ge \frac{2s_i}{\delta } {\hat{\kappa }} = 2 s_i \frac{{q_{\min }}-1}{({q_{\min }}-2)^2}\, . \end{aligned}$$
(33)

Inserting the above estimate for \((III)_i\) into (30), we find

$$\begin{aligned} (II)_i \le \varepsilon \int _{\Omega }\Gamma _i^{s_i} |\eta _m|^{2\kappa } \, \mathrm{d}x+ c(\varepsilon ) \, , \end{aligned}$$

and if we choose \(\varepsilon\) sufficiently small, we see that (25) yields the following result:

Proposition 2.3

Let \(s_i\), \(\kappa \ge 1\) satisfy (29) and (33). Then there exists a constant c being independent of m such that (\(i=1\), 2)

$$\begin{aligned} \int _{\Omega }\Gamma _i^{s_i} |\eta _m|^{2\kappa } \, \mathrm{d}x\le c \Big [ (I)_i+1\Big ] \, ,\qquad (I)_i := \Bigg [\int _{\Omega }\big | \nabla \Gamma _i^{\frac{s_i}{2}}\big | \, |\eta _m|^\kappa \, \mathrm{d}x\Bigg ]^2 \, . \end{aligned}$$
(34)

In order to prove Theorem 1.1 it remains to discuss the quantity \((I) = (I)_i\) for \(i=1\), 2, i.e. now second derivatives of u have to be handled in an appropriate way. It holds

$$\begin{aligned} (I)_1\le \, & {} c \Bigg [ \int _{\Omega }\Gamma _1^{\frac{s_1}{2}-1} |\nabla \Gamma _1| \, |\eta _m|^{\kappa } \, \mathrm{d}x\Bigg ]^2 \nonumber \\\le & \, {} c \Bigg [ \int _{\Omega }\Gamma _1^{\frac{s_1-1}{2}} |\partial _1 \partial _1 u|\, |\eta _m|^{\kappa } \, \mathrm{d}x\Bigg ]^2 + c \Bigg [\int _{\Omega }\Gamma _1^{\frac{s_1 -1}{2}} |\partial _1 \partial _2 u|\, |\eta _m|^{\kappa }\, \mathrm{d}x\Bigg ]^2 =: T_1 + {\tilde{T}}_1\, , \end{aligned}$$

and in the same manner

$$\begin{aligned} (I)_2\le \, & {} c \Bigg [ \int _{\Omega }\Gamma _2^{\frac{s_2-1}{2}} |\partial _2 \partial _2 u|\, |\eta _m|^{\kappa } \, \mathrm{d}x\Bigg ]^2 + c \Bigg [\int _{\Omega }\Gamma _2^{\frac{s_2 -1}{2}} |\partial _1 \partial _2 u|\, |\eta _m|^{\kappa }\, \mathrm{d}x\Bigg ]^2 =: T_2 + {\tilde{T}}_2\, . \end{aligned}$$

We have

$$\begin{aligned} T_i= \, & {} c \Bigg [ \int _{\Omega }\Gamma _i^{\frac{q_i-2}{4}} |\partial _i \partial _i u|\, |\eta _m|^{\kappa } \Gamma _i^{\frac{\alpha _i}{2}} \Gamma _i^{\frac{s_i -1}{2}- \frac{q_i-2}{4} - \frac{\alpha _i}{2}} \, \mathrm{d}x\Bigg ]^2\\\le \, & {} c \Bigg [ \int _{\Omega }\Gamma _i^{\frac{q_i-2}{2}}|\partial _i \partial _i u|^2 |\eta _m|^{2\kappa } \Gamma _i^{\alpha _i}\, \mathrm{d}x\Bigg ] \Bigg [ \int _{\Omega }\Gamma _i^{s_i -1 - \frac{q_i -2}{2} - \alpha _i} \, \mathrm{d}x\Bigg ]\, , \end{aligned}$$

where the last estimate follows from Hölder’s inequality and \(\alpha _i\), \(i=1\), 2, denote real numbers such that for the moment

$$\begin{aligned} s_i \le q_i + \alpha _i \, , \qquad i=1, \, 2\, . \end{aligned}$$
(35)

Note that the condition (35) guarantees the validity of

$$\begin{aligned} \int _{\Omega }\Gamma _i^{s_i - 1 - \frac{q_i -2}{2} - \alpha _i} \, \mathrm{d}x\le c \, , \qquad i=1,\, 2\, , \end{aligned}$$
(36)

for a finite constant c. Recalling (4) we see that (36) yields the bound

$$\begin{aligned} T_i \le c \int _{\Omega }f_i''(\partial _i u) |\partial _i \partial _i u|^2 |\eta _m|^{2\kappa } \Gamma _i^{\alpha _i} \, \mathrm{d}x\end{aligned}$$
(37)

again for \(i=1\), 2.

Let us look at the quantities \({\tilde{T}}_i\): we have by Hölder’s inequality

$$\begin{aligned} {\tilde{T}}_1= \, & {} c \Bigg [ \int _{\Omega }\Gamma _1^{\frac{q_1-2}{4}} |\partial _1 \partial _2 u| |\eta _m|^{\kappa } \Gamma _1^{\frac{s_1-1}{2}} \Gamma _2^{\frac{\alpha }{2}} \Gamma _1^{-\frac{q_1-2}{4}} \Gamma _2^{- \frac{\alpha _2}{2}}\, \mathrm{d}x\Bigg ]^2\nonumber \\\le \, & {} c \Bigg [\int _{\Omega }\Gamma _1^{\frac{q_1-2}{2}} |\partial _1 \partial _2 u|^2 |\eta _m|^{2\kappa } \Gamma _2^{\alpha _2} \, \mathrm{d}x\Bigg ] \Bigg [ \int _{\Omega }\Gamma _1^{s_1 -1 - \frac{q_1-2}{2}} \Gamma _2^{-\alpha _2}\, \mathrm{d}x\Bigg ]\nonumber \\\le \, & {} c \Bigg [ \int _{\Omega }f_1''(\partial _1 u) |\partial _1 \partial _2 u|^2 |\eta _m|^{2\kappa } \Gamma _2^{\alpha _2}\, \mathrm{d}x\Bigg ] \Bigg [\int _{\Omega }\Gamma _1^{s_1 -1 - \frac{q_1 -2}{2}} \Gamma _2^{-\alpha _2} \, \mathrm{d}x\Bigg ]\nonumber \\=: \, & {} c S_1' \cdot S_1'' \, . \end{aligned}$$
(38)

In order to benefit from the inequality (38), we replace (35) by the stronger bound

$$\begin{aligned} s_i \le \frac{3}{4} q_i \, ,\qquad i=1, \, 2\, , \end{aligned}$$
(39)

together with the requirement

$$\begin{aligned} \alpha _i \in (-1/2,0)\, ,\qquad i=1, \, 2\, . \end{aligned}$$
(40)

Here we note that (39) together with (40) yields (35). We then obtain on account of Young’s inequality

$$\begin{aligned} S_1'' \le c \Bigg [ \int _{\Omega }\Gamma _1^{\frac{q_1}{2}} \, \mathrm{d}x+ \int _{\Omega }\Gamma _2^\beta \, \mathrm{d}x\Bigg ] \end{aligned}$$

with exponent

$$\begin{aligned} \beta := - \alpha _2 \Big [\frac{q_1/2}{s_1 - q_1/2}\Big ]^* = - \alpha _2 \frac{q_1/2}{q_1-s_1} \le - 2 \alpha _2 \le \frac{q_2}{2} \, , \end{aligned}$$

where we used (39), (40) and (40), (5), respectively, for the last inequalities. Thus (38) reduces to

$$\begin{aligned} {\tilde{T}}_1 \le c \int _{\Omega }f_1''(\partial _1 u) |\partial _1 \partial _2 u|^2 |\eta _m|^{2\kappa } \Gamma _2^{\alpha _2}\, \mathrm{d}x\, . \end{aligned}$$
(41)

In the same spirit it follows

$$\begin{aligned} {\tilde{T}}_2 \le c \int _{\Omega }f_2''(\partial _2 u) |\partial _1 \partial _2 u|^2 |\eta _m|^{2\kappa } \Gamma _1^{\alpha _1}\, \mathrm{d}x\, . \end{aligned}$$
(42)

Let us return to (34): we have by (37), (41) and (42) and by (8)

$$\begin{aligned}&{\int _{\Omega }\Big [\Gamma _1^{s_1} + \Gamma _2^{s_2}\Big ] |\eta _m|^{2\kappa } \, \mathrm{d}x\le c \big [1+T_1+{\tilde{T}}_2+{\tilde{T}}_1+T_2]}\\ & \quad = {} c \Bigg [ 1+ \int _{\Omega }\Big (f_1''(\partial _1 u)|\partial _1\partial _1 u|^2 + f_2''(\partial _2 u) |\partial _1\partial _2 u|^2\Big ) |\eta _m|^{2\kappa } \Gamma _1^{\alpha _1} \, \mathrm{d}x\\ & \qquad + \int _{\Omega }\Big (f_1''(\partial _1 u)|\partial _1\partial _2 u|^2 + f_2''(\partial _2 u) |\partial _2\partial _2 u|^2\Big ) |\eta _m|^{2\kappa } \Gamma _2^{\alpha _2} \, \mathrm{d}x\Bigg ]\\ & \quad = {} c \Bigg [ 1 + \int _{\Omega }D^2f(\nabla u) \big (\partial _1 \nabla u,\partial _1\nabla u\big ) |\eta _m|^{2\kappa } \Gamma _1^{\alpha _1}\, \mathrm{d}x\\ & \qquad + \int _{\Omega }D^2f(\nabla u) \big (\partial _2\nabla u,\partial _2 \nabla u\big )|\eta _m|^{2\kappa } \Gamma _2^{\alpha _2}\, \mathrm{d}x\Bigg ] \, . \end{aligned}$$

The remaining integrals are handled with the help of Proposition 6.1 (compare also inequality (4.6) from [17]) replacing l by \(\kappa\) and \(\eta\) by \(|\eta _m|\), respectively. We note that the proof of Proposition 6.1 obviously remains valid with these replacements. We emphasize that (40) is an essential assumption to apply Proposition 6.1. We get

$$\begin{aligned} \int _{\Omega }\Big [\Gamma _1^{s_1} + \Gamma _2^{s_2}\Big ] |\eta _m|^{2\kappa }\mathrm{d}x\le \, & {} c \Bigg [ 1 + \int _{\Omega }D^2f(\nabla u) \big (\nabla |\eta _m|,\nabla |\eta _m|\big ) \Gamma _1^{1+\alpha _1}|\eta _m|^{2\kappa -2}\, \mathrm{d}x\nonumber \\&+ \int _{\Omega }D^2f(\nabla u) \big (\nabla |\eta _m|,\nabla |\eta _m|\big ) \Gamma _2^{1+\alpha _2} |\eta _m|^{2\kappa -2}\, \mathrm{d}x\Bigg ]\, . \end{aligned}$$
(43)

Finally we let \(s_i = \frac{3}{4}q_i\), \(i=1\), 2, which is the optimal choice with respect to (39). We note that with this choice (33) follows from

$$\begin{aligned} \kappa \ge \frac{3}{2} \frac{{q_{\max }}({q_{\min }}-1)}{({q_{\min }}-2)^2}\, , \end{aligned}$$
(44)

and (44) is valid if \(\kappa\) is chosen according to assumption (12) from Theorem 1.1 with \(T(q_1,q_2)\) defined in (13).

From (4) and (8) we deduce

$$\begin{aligned} \hbox {r.h.s.~of~(2.19)}\le \, & {} c \Bigg [ 1+ \sum _{i=1}^2 \int _{\Omega }\Gamma _i^{\frac{q_i}{2}+\alpha _i} |\partial _i \eta _m|^2 |\eta _m|^{2\kappa -2} \, \mathrm{d}x\nonumber \\&+ \int _{\Omega }\Gamma _2^{\frac{q_2 -2}{2}} |\partial _2 \eta _m|^2 |\eta _m|^{2\kappa -2} \Gamma _1^{1+\alpha _1}\, \mathrm{d}x\nonumber \\&+ \int _{\Omega }\Gamma _1^{\frac{q_1-2}{2}} |\partial _1 \eta _m|^2 |\eta _m|^{2\kappa -2} \Gamma _2^{1+\alpha _2} \, \mathrm{d}x\Bigg ]\, . \end{aligned}$$
(45)

We recall the definition (26) of \(\eta _m\) and the gradient bound for \(\varphi _m\) stated in (27). This yields for integrand of the first integral on the right-hand side of (45):

$$\begin{aligned} \Gamma _i^{\frac{q_i}{2}+\alpha _i} |\partial _i \eta _m|^2 |\eta _m|^{2 \kappa -2}\le \, & {} c \Bigg [ \Gamma _i^{\frac{q_i}{2}+\alpha _i} |\partial _i (u-u_0)|^2 |\eta _m|^{2\kappa -2} +\Gamma _i^{\frac{q_i}{2}+\alpha _i} |\partial _i \varphi _m|^2 |u-u_0|^{2\kappa }\Bigg ]\, . \end{aligned}$$

We quote inequality (28) and recall that \(\nabla \varphi _m\) has support in the set \(\{x\in \Omega :\, {\text {dist}}(x,\partial \Omega )\le 1/m\}\) satisfying \(|{\text {spt}}\nabla \varphi _m|\le c/m\), hence

$$\begin{aligned} \int _{\Omega }\Gamma _i^{\frac{q_i}{2}+\alpha _i} |\partial _i \varphi _m|^2 |u-u_0|^{2\kappa } \, \mathrm{d}x\le & {} \int _{{\text {spt}}\nabla \varphi _m} \Gamma _i^{\frac{q_i}{2}+\alpha _i}m^{2-2\kappa \frac{{q_{\min }}-2}{{q_{\min }}}} \, \mathrm{d}x\\\le & {} c \int _{\Omega }\Gamma _i^{\frac{q_i}{2}} \, \mathrm{d}x+ c \int _{{\text {spt}} \nabla \varphi _m} m^{[ 2-2\kappa \frac{{q_{\min }}-2}{{q_{\min }}}] \gamma ^*_i} \, \mathrm{d}x\, , \end{aligned}$$

where (recall (40))

$$\begin{aligned} \gamma _i = \frac{q_i}{q_i+2\alpha _i}\, ,\qquad \gamma _i^* = \Bigg [\frac{q_i}{q_i+2\alpha _i}\Bigg ]^* = - \frac{q_i}{2\alpha _i}\, . \end{aligned}$$

Thus the second integral on the right-hand side is bounded if, with \(\alpha _i\) sufficiently close to \(-1/2\), we have

$$\begin{aligned} \kappa > \frac{2{q_{\max }}-1}{2 {q_{\max }}} \frac{{q_{\min }}}{{q_{\min }}-2} \, . \end{aligned}$$
(46)

Assuming (46) we arrive at

$$\begin{aligned} \int _{\Omega }\Gamma _i^{\frac{q_i}{2}+\alpha _i} |\partial _i \eta _m|^2 |\eta _m|^{2\kappa -2} \, \mathrm{d}x\le c \Bigg [1 + \int _{\Omega }\Gamma _i^{\frac{q_i}{2}+\alpha _i}|\partial _i u - \partial _i u_0|^2 |\eta _m|^{2\kappa -2}\, \mathrm{d}x\Bigg ] \end{aligned}$$
(47)

for any \(\alpha _i\) sufficiently close to \(-1/2\). Assuming this we next let

$$\begin{aligned} \beta _i = \frac{3q_i}{2q_i+4(1+\alpha _i)} \qquad \hbox {with conjugate exponent}\qquad \beta _i^* = \frac{3q_i}{q_i - 4 (1+\alpha _i)} \end{aligned}$$

and apply Young’s inequality in an obvious way to get

$$\begin{aligned} \int _{\Omega }\Gamma _i^{\frac{q_i}{2}+\alpha _i+1} |\eta _m|^{2\kappa -2}\, \mathrm{d}x\le \varepsilon \int _{\Omega }\Gamma _i^{s_i}|\eta _m|^{2\kappa }\, \mathrm{d}x+ c(\varepsilon ) \int _{\Omega }|\eta _m|^{2\kappa - 2 \beta _i^*}\, \mathrm{d}x\, , \end{aligned}$$

where the first term can be absorbed in the left-hand side of (43), while the second one bounded under the assumption

$$\begin{aligned} \kappa > \frac{3{q_{\min }}}{{q_{\min }}-2}\, . \end{aligned}$$
(48)

Here we used the fact that for \(q > 2\) the function \(q/(q-2)\) is a decreasing function. Altogether we have shown that for exponents \(\alpha _i\) close to \(-1/2\) the first integral on the right-hand side of (45) splits into two parts, where the first one can be absorbed in the left-hand side of (45) and the second one stays bounded. During our calculations we evidently used (9), however (9) can be replaced by weaker integrability assumtions concerning \(\partial _i u_0\). We leave the details to the reader.

Let us finally consider the “mixed terms” on the right-hand side of (45). We first observe the inequality

$$\begin{aligned}&{\int _{\Omega }\Gamma _2^{\frac{q_2-2}{2}} |\partial _2 \eta _m|^2 |\eta _m|^{2 \kappa -2}\Gamma _1^{1+\alpha _1}\, \mathrm{d}x}\nonumber \\ & \le {} c \int _{\Omega }\Gamma _2^{\frac{q_2-2}{2}} (\partial _2(u-u_0))^2 |\eta _m|^{2\kappa -2} \Gamma _1^{1+\alpha _1}\, \mathrm{d}x\nonumber \\ & \quad +\int _{\Omega }\Gamma _2^{\frac{q_2-2}{2}} |\partial _2 \varphi _m|^2 |u-u_0|^2 |\eta _m|^{2\kappa -2} \Gamma _1^{1+\alpha _1}\, \mathrm{d}x\, . \end{aligned}$$
(49)

Considering the limit case \(\alpha _1 =-1/2\), the first integral on the right-hand side of (49) basically is of the form

$$\begin{aligned} \int _{\Omega }\Gamma _2^{\frac{q_2}{2}} \Gamma _1^{\frac{1}{2}} |\eta _m|^{2\kappa -2} \, \mathrm{d}x\end{aligned}$$

and this integral directly results from an application of Caccioppoli’s inequality. If we like to show an integrability result for \(\Gamma _2^{t_2}\) with some power \(t_2 > q_2/2\), then the idea is to apply Young’s inequality choosing

$$\begin{aligned} \beta = \frac{2t_2}{q_2}\qquad \hbox {with conjugate exponent}\qquad \beta ^*= \frac{2t_2}{2t_2-q_2} \end{aligned}$$

leading to the quantities \(\Gamma _2^{t_2}\) and \(\Gamma _1^{\frac{t_2}{2t_2 - q_2}}\). If \(t_1 =t_1(q_1) > q_1/2\) denotes the desired integrability exponent for \(\Gamma _1\), then this requires the bound

$$\begin{aligned} \frac{t_2}{2t_2-q_2} \le t_1(q_1) \qquad \hbox {for all}\; q_1 > 2 \end{aligned}$$

and of course we need the same condition changing the roles of \(t_1\) and \(t_2\). With the symmetric Ansatz \(t_i = \theta q_i\) for some \(\theta > 1/2\) we are immediately led to \(t_i = \frac{3}{4} q_i\), \(i=1\), 2, which again motivates our choice of \(s_i\).

More precisely: discussing the first integral on the right-hand side of (49) we choose \(\beta _1 = 3/2\) with conjugate exponent \(\beta _1^* =3\) and obtain

$$\begin{aligned} \int _{\Omega }\Gamma _2^{\frac{q_2}{2}} |\eta _m|^{2\kappa -2} \Gamma _1^{1+\alpha _1} \, \mathrm{d}x= & {} \int _{\Omega }\Gamma _2^{\frac{q_2}{2}} \Gamma _1^{1+\alpha _1} |\eta _m|^{\frac{2\kappa }{\beta _1}} |\eta _m|^{\frac{2\kappa }{\beta _1^*} -2} \, \mathrm{d}x\\\le \, & {} \varepsilon \int _{\Omega }\Gamma _2^{s_2} |\eta _m|^{2\kappa } \, \mathrm{d}x+ c(\varepsilon ) \int _{\Omega }\Gamma _1^{3 (1+\alpha _1)} |\eta _m|^{2\kappa -6}\, \mathrm{d}x\, . \end{aligned}$$

Here the first integral is absorbed in the left-hand side of (43) and since \(\alpha _1\) is chosen sufficiently close to \(-1/2\), the second integral is bounded provided that we suppose in addition

$$\begin{aligned} q_1 > 3 \quad \hbox {and}\quad \kappa \ge 3\, . \end{aligned}$$
(50)

If \(q_1 < 3\), using

$$\begin{aligned} \beta _2 =\frac{q_1}{4 (1+\alpha _1)} \, , \qquad \beta _2^* = \frac{q_1}{q_1-4(1+\alpha _1)}\, , \end{aligned}$$

we are led to (\({\tilde{\varepsilon }} \ll \varepsilon\))

$$\begin{aligned} c(\varepsilon ) \int _{\Omega }\Gamma _1^{3(1+\alpha _1)} \eta _m^{2\kappa -6} \, \mathrm{d}x\le {\tilde{\varepsilon }}\int _{\Omega }\Gamma _1^{s_1} \eta ^{2\kappa } dx + c({\tilde{\varepsilon }}) \int _{\Omega }\eta ^{2\kappa - 6 \beta _2^*}\, \mathrm{d}x\, , \end{aligned}$$

where the first integral is absorbed in the right-hand side of (43) and if \(\alpha _1\) sufficiently close to \(-1/2\) we now suppose

$$\begin{aligned} \kappa > \frac{3q_1}{q_1-2} \, . \end{aligned}$$
(51)

Note that the condition (51) is a consequence of (48).

It remains to discuss the second integral on the right-hand side of (49). Using Young’s inequality with

$$\begin{aligned} \beta _3 = \frac{3}{2} \frac{q_2}{q_2 - 2}\, , \qquad \beta _3^* = \frac{3q_2}{q_2+4}\, , \end{aligned}$$

we obtain

$$\begin{aligned}&{\int _{\Omega }\Gamma _2^{\frac{q_2-2}{2}} |\partial _2 \varphi _m|^2 |u-u_0|^2 |\eta _m|^{2\kappa -2} \Gamma _1^{1+\alpha _1}\, \mathrm{d}x}\nonumber \\ & \le {} \varepsilon \int _{\Omega }\Gamma _2^{s_2} |\eta _m|^{2\kappa } \, \mathrm{d}x\nonumber \\ & \quad + c(\varepsilon ) \int _{\Omega }\Gamma _1^{(1+\alpha _1) \frac{3q_2}{q_2+4}} | \partial _2 \varphi _m|^{2\beta _3^*} |u-u_0|^{2 \beta _3^*}|\eta _m|^{2\kappa -2\beta _3^*}\, \mathrm{d}x\, , \end{aligned}$$
(52)

where again the first integral is absorbed in the left-hand side of (43). Considering the second integral we choose

$$\begin{aligned} \beta _4 = \frac{q_1 (q_2+4)}{4 q_2 (1+ \alpha _1)} \qquad \hbox {with conjugate exponent}\qquad \beta _4^* = \frac{q_1(q_2+4)}{q_2[q_1-4(1+\alpha _1)]+4q_1} \, . \end{aligned}$$

This gives (\({\tilde{\varepsilon }} \ll \varepsilon\))

$$\begin{aligned}&{c(\varepsilon ) \int _{\Omega }\Gamma _1^{(1+\alpha _1) \frac{3q_2}{q_2+4}} | \partial _2 \varphi _m|^{2\beta _3^*} |u-u_0|^{2 \beta _3^*}|\eta _m|^{2\kappa -2\beta _3^*}\, \mathrm{d}x}\nonumber \\ & \le {} {\tilde{\varepsilon }} \int _{\Omega }\Gamma _1^{s_1} |\eta _m|^{2\kappa }\, \mathrm{d}x\nonumber \\ & \quad + c({\tilde{\varepsilon }}) \int _{{\text {spt}}\nabla \varphi _m} |\partial _2 \varphi _m|^{2 \beta _3^*\beta _4^*} |u-u_0|^{2\beta _3^* \beta _4^*} |\eta _m|^{2\kappa -2\beta _3^*\beta _4^*}\, \mathrm{d}x\, . \end{aligned}$$
(53)

As usual the first integral on the right-hand side of (53) is absorbed in the left-hand side of (43) and we calculate

$$\begin{aligned} \beta _3^* \beta _4^* = \frac{3 q_1 q_2}{q_2[q_1 - 4(1+\alpha _1)]+4q_1}\, . \end{aligned}$$
(54)

We obtain recalling (27) and (28)

$$\begin{aligned}&{ \int _{{\text {spt}}\nabla \varphi _m} |\partial _2 \varphi _m|^{2 \beta _3^*\beta _4^*} |u-u_0|^{2\beta _3^* \beta _4^*} |\eta _m|^{2\kappa -2\beta _3^*\beta _4^*}\, \mathrm{d}x}\nonumber \\ & \quad \le {} \int _{{\text {spt}}\nabla \varphi _m} |\partial _2 \varphi _m|^{2 \beta _3^*\beta _4^*} |u-u_0|^{2\beta _3^* \beta _4^*} |u-u_0|^{2\kappa -2\beta _3^*\beta _4^*}\, \mathrm{d}x\nonumber \\ & \quad \le {} c m^{-1} m^{\frac{6 q_1 q_2}{q_2[q_1 - 4(1+\alpha _1)]+4q_1}} m^{- \frac{{q_{\min }}-2}{{q_{\min }}} 2 \kappa } \, . \end{aligned}$$
(55)

For \(\alpha _1\) sufficiently close to \(-1/2\) this leads to the requirement

$$\begin{aligned} \kappa > \frac{{q_{\min }}}{{q_{\min }}-2}\, \frac{1}{2} \Bigg [\frac{6q_1q_2}{q_2(q_1-2)+4q_1} - 1 \Bigg ] \, . \end{aligned}$$
(56)

With \(q_1>2\) fixed and for \(q_2 > 2\) we consider the function \(g(q_2) = \frac{6q_1q_2}{q_2(q_1-2) + 4q_1}\). We have

$$\begin{aligned} g'(q_2)= & {} \frac{6 q_1}{q_2(q_1-2) + 4q_1} - \frac{6q_1q_2}{(q_2(q_1-2)+4q_1)^2} (q_1-2)\\= & {} \frac{24 q_1^2}{(q_2(q_1-2)+4q_1)^2} > 0 \, , \end{aligned}$$

hence g is an increasing function and

$$\begin{aligned} g(q_2) \le \lim _{t \rightarrow \infty } \frac{6 q_1 t}{t(q_1-2)+4q_1} = \frac{6q_1}{q_1-2} \, . \end{aligned}$$

Thus,

$$\begin{aligned} \kappa \ge \frac{{q_{\min }}}{{q_{\min }}-2}\, \frac{1}{2} \Bigg [6 \frac{{q_{\min }}}{{q_{\min }}-2} - 1\Bigg ] \end{aligned}$$
(57)

implies the validity of (56).

Summarizing the conditions imposed on \(\kappa\) during our calculations, i.e. recalling the bounds (29), (44), (46), (48), (57) we are led to the lower bound

$$\begin{aligned} \kappa > 3 \max \Bigg \{1, \frac{1}{2} \frac{{q_{\max }}}{{q_{\min }}}\frac{{q_{\min }}-1}{{q_{\min }}-2}, \frac{{q_{\min }}}{{q_{\min }}-2} - \frac{1}{6}\Bigg \} \frac{{q_{\min }}}{{q_{\min }}-2} \end{aligned}$$
(58)

for the exponent \(\kappa\). Assuming the validity of (58) and returning to (43) we now have shown that for \(\alpha _i\) sufficiently close to \(-1/2\) the right-hand side of (43) can be splitted into terms which either can be absorbed in the left-hand side of (43) or stay uniformly bounded, hence

$$\begin{aligned} \int _{\Omega }\Big [\Gamma _1^{\frac{3}{4}q_1} + \Gamma _2^{\frac{3}{4}q_2}\Big ] |\eta _m|^{2\kappa }\, \mathrm{d}x\le c \end{aligned}$$
(59)

for a finite constant c independent of m. Passing to the limit \(m \rightarrow \infty\) in (59) our claim (11) follows. Obviously (12) is a consequence of (58) and the definition of \(T(q_1,q_2)\) stated in (13), which completes the proof of Theorem 1.1 for arbitrary exponents \(q_1\), \(q_2 >2\). \(\square\)

Remark 2.1

Let us add some comments on the behaviour of \(\kappa\), which means that we look at the lower bound for the exponent \(\kappa\) given by the right-hand side of inequality (58).

  1. i)

    Since \(u-u_0 \in W^{1,{q_{\min }}}(\Omega )\) and thereby \(u-u_0 \in C^{0,\nu }({\overline{\Omega }})\), \(\nu := 1- 2/{q_{\min }}\), the Hölder exponent enters (58), which also corresponds to the natural effect that

  2. ii)

    \(\kappa \rightarrow \infty\) as \({q_{\min }}\rightarrow 2\).

  3. iii)

    The ratio \({q_{\max }}/{q_{\min }}\) determines the growth of \(\kappa\).

  4. iv)

    In the limit \({q_{\max }}= {q_{\min }}\rightarrow \infty\) condition (58) reduces to \(\kappa > 3\).

3 Comments on the case \({q_{\min }}> 5\)

We now choose the sequence \(\eta _m \in C^1_0(\Omega )\) according to

$$\begin{aligned} \partial _i \eta _m \rightarrow \partial _i u \, ,&\hbox {in}\; L^{q_i}(\Omega ) \, ,\quad i= 1 ,\, 2 \, , \nonumber \\ \eta _m \rightarrow u-u_0&\hbox {uniformly as}\; m\rightarrow \infty \, . \end{aligned}$$
(60)

For some elementary properties of anisotropic Sobolev spaces including an appropriate version of this density result we refer, e.g., to [18, 19]). We emphasize that during the following calculations condition (9) can be replaced by the weaker requirement \(\partial _i u_0 \in L^{q_i}(\Omega )\), \(i=1\), 2. Proposition 2.1 obviously holds for \(\eta _m\), and with \(\delta\) as in Proposition 2.2 we now obtain (30) with the choice \({{\hat{\kappa }}}=1\) observing that going through the proof of (30) the quantity \((II)_i\) can be handled as follows: first we note

$$\begin{aligned} (II)_i \le \Bigg [\int _{\Omega }\Gamma ^{s_i - \frac{\delta }{2}} |\eta _m|^{2\kappa -2}\, \mathrm{d}x\Bigg ]\cdot \Bigg [ \int _{\Omega }|\nabla \eta _m|^2 \Gamma _i^{\frac{\delta }{2}}\, \mathrm{d}x\Bigg ]\, , \end{aligned}$$

and then we use

$$\begin{aligned} \int _{\Omega }|\nabla \eta _m|^2 \Gamma _i^{\frac{\delta }{2}} \, \mathrm{d}x\le c \Bigg [1+\int _{\Omega }|\nabla u|^{q_{\min }}\, \mathrm{d}x\Bigg ] \, , \end{aligned}$$

which is a consequence of (60).

We continue with the discussion of \((III)_i\) for the choice \({{\hat{\kappa }}}=1\), and observe that (33) has to be replaced by

$$\begin{aligned} \kappa \ge \frac{2}{\delta } s_i\, . \end{aligned}$$
(61)

Exactly as in Section 2 we obtain the inequalities (43), (45) now being valid for any exponents \(s_i\), \(\alpha _i\), \(\kappa\) such that (\(i=1\), 2)

$$\begin{aligned} \alpha _i \in (-1/2,0)\, ,\qquad s_i \le \frac{3}{4} q_i\, ,\qquad \kappa \ge \frac{2}{\delta } s_i \, . \end{aligned}$$
(62)

Still following the lines of Section 2 we let \(s_i = \frac{3}{4} q_i\), \(i=1\), 2 , and replace (61) by

$$\begin{aligned} \kappa \ge \frac{3}{2 \delta } {q_{\max }}\, . \end{aligned}$$
(63)

We let

$$\begin{aligned} \rho _i := 2 \kappa -2 -2\kappa \Big [\frac{q_i}{2}+\alpha _i\Big ] \frac{1}{s_i} \end{aligned}$$

and consider the terms on the right-hand side of (45). Young’s inequality yields (\(0< \varepsilon < 1\))

$$\begin{aligned}&{\sum _{i=1}^2 \int _{\Omega }\Gamma _i^{\frac{q_i}{2}+\alpha _i} |\partial _i \eta _m|^2 |\eta _m|^{2\kappa -2} \, \mathrm{d}x}\nonumber \\ & \quad = {} \sum _{i=1}^2 \int _{\Omega }\Big [\Gamma _i^{s_i} |\eta _m|^{2\kappa }\Big ]^{\frac{1}{s_i} \big [\frac{q_i}{2}+\alpha _i\big ]} |\partial _i \eta _m|^2 |\eta _m|^{\rho _i}\, \mathrm{d}x\nonumber \\ & \quad \le {} \sum _{i=1}^2 \Bigg [\varepsilon \int _{\Omega }\Gamma _i^{s_i} |\eta _m|^{2 \kappa }\, \mathrm{d}x+ c(\varepsilon ) \int _{\Omega }|\eta _m|^{\rho _i \gamma _i^*} |\partial _i \eta _m|^{2 \gamma _i^*} \, \mathrm{d}x\Bigg ] \end{aligned}$$
(64)

with

$$\begin{aligned} \gamma _i := \frac{s_i}{\frac{q_i}{2}+\alpha _i} = \frac{3 q_i}{2 q_i + 4 \alpha _i}\, ,\qquad \gamma _i^* =\frac{3q_i}{q_i-4\alpha _i}\, . \end{aligned}$$

While the \(\varepsilon\)-part of (64) can be absorbed, the remaining integrals are bounded if for \(i=1\), 2

$$\begin{aligned} \rho _i \ge 0 \qquad \hbox {and}\qquad 2 \gamma _i^* \le q_i \, . \end{aligned}$$

Noting that the function \(t \mapsto t/(t+2)\), \(t \ge 0\), is increasing and by choosing \(\alpha _i\), \(i=1\), 2, sufficiently close to \(-1/2\), we see that these requirements are consequences of the strict inequalities

$$\begin{aligned} {q_{\max }}> 4\, ,\qquad \kappa > \frac{3 {q_{\max }}}{{q_{\max }}+2} \, . \end{aligned}$$
(65)

We thus have assuming (63) and (65)

$$\begin{aligned} \int _{\Omega }\Big [\Gamma _1^{s_1} + \Gamma _2^{s_2}\Big ] |\eta _m|^{2\kappa }\, \mathrm{d}x\le & {} c \Bigg [ 1+ \int _{\Omega }\Gamma _2^{\frac{q_2 -2}{2}} |\partial _2 \eta _m|^2 |\eta _m|^{2\kappa -2} \Gamma _1^{1+\alpha _1}\, \mathrm{d}x\nonumber \\&+ \int _{\Omega }\Gamma _1^{\frac{q_1-2}{2}} |\partial _1 \eta _m|^2 |\eta _m|^{2\kappa -2} \Gamma _2^{1+\alpha _2} \, \mathrm{d}x\Bigg ]\, . \end{aligned}$$
(66)

Let us have closer look at the first integral on the right-hand side of (66). Youngs’s inequality gives

$$\begin{aligned}&{\int _{\Omega }\Gamma _2^{\frac{q_2 -2}{2}} |\partial _2 \eta _m|^2 |\eta _m|^{2\kappa -2} \Gamma _1^{1+\alpha _1}\, \mathrm{d}x}\nonumber \\ & \le {} \int _{\Omega }|\partial _2 \eta _m|^{q_2}\, \mathrm{d}x+ \int _{\Omega }\Gamma _2^{\frac{q_2}{2}} |\eta _m|^{(2\kappa -2) \frac{q_2}{q_2 -2}} \Gamma _1^{(1+\alpha _1)\frac{q_2}{q_2 -2}}\, \mathrm{d}x\nonumber \\ & \le {} c\Bigg [1+ \int _{\Omega }\Big [ \Gamma _2^{s_2}|\eta _m|^{2\kappa }\Big ]^{\frac{2}{3}} |\eta _m|^{(2\kappa -2) \frac{q_2}{q_2-2} - \frac{4}{3}\kappa } \Gamma _1^{(1+\alpha _1) \frac{q_2}{q_2 -2}} \, \mathrm{d}x\Bigg ]\nonumber \\ & \le {} c + \varepsilon \int _{\Omega }\Gamma _2^{s_2} |\eta _m|^{2\kappa } \, \mathrm{d}x\nonumber \\& \quad + c(\varepsilon ) \int _{\Omega }|\eta _m|^{3 \big [(2\kappa -2) \frac{q_2}{q_2 -2} - \frac{4}{3} \kappa \big ]} \Gamma _1^{3(1+\alpha _1)\frac{q_2}{q_2 -2}} \, \mathrm{d}x\, . \end{aligned}$$
(67)

Here the first integral is absorbed and the second is bounded if we have (again \(\alpha _2\) being sufficiently close to \(-1/2\))

$$\begin{aligned} (2\kappa -2) \frac{q_2}{q_2-2} - \frac{4}{3}\kappa \ge 0 \qquad \hbox {and}\qquad \frac{3}{2}\frac{q_2}{q_2-2} < \frac{q_1}{2}\, . \end{aligned}$$

The first condition follows from the requirement (65), the second one holds if we assume in addition that \({q_{\min }}> 5\).

In the same way the last term on the right-hand side of (66) is handled and by combining (63), (65) we have shown Theorem 1.1 together with the formula (14). \(\square\)

4 Proof of Theorem 1.2

We proceed along the lines of Section 2 assuming that all the hypothese of Theorem 1.2 are satisfied. In place of (25) we have (\(\Gamma := 1+|\nabla u|^2\))

$$\begin{aligned} \int _{\Omega }\Gamma ^s |\eta |^{2\kappa } \, \mathrm{d}x\le \, & {} c \Bigg \{\Bigg [\int _{\Omega }|\nabla \Gamma ^{\frac{s}{2}}| \, |\eta |^\kappa \, \mathrm{d}x\Bigg ]^2 +\Bigg [ \int _{\Omega }\Gamma ^{\frac{s}{2}} |\nabla \eta |\,|\eta |^{\kappa -1} \, \mathrm{d}x\Bigg ]^2\Bigg \}\nonumber \\=: & {} c\{(I)+(II)\}\, , \end{aligned}$$
(68)

and (68) holds for any s, \(\kappa \ge 1\) and \(\eta \in C^1_0(\Omega )\). We define the sequence \(\eta _m\) as in (26) observing that (28) holds with \({q_{\min }}\) replaced by p. Letting \(\delta := p-2\), defining \({\hat{\kappa }} := \frac{p-1}{p-2}\) we obtain for

$$\begin{aligned} s > \frac{\delta }{2}\, , \quad \kappa \ge {\hat{\kappa }} \end{aligned}$$
(69)

applying obvious modifications in the proof of (30)

$$\begin{aligned} (II) \le c \int _{\Omega }\Gamma ^{s-\frac{\delta }{2}}|\eta _m|^{2\kappa - 2 {\hat{\kappa }}}\, \mathrm{d}x=: c (III) \, . \end{aligned}$$
(70)

Here (II) just denotes the term occurring on the right-hand side of (68) with \(\eta\) being replaced by \(\eta _m\). The quantity (III) is discussed analogously to the term \((III)_i\): in accordance with the calculations presented after the proof of Proposition 2.2 we get for any \(\varepsilon >0\)

$$\begin{aligned} (III)\le \, & {} \varepsilon \int _{\Omega }\Gamma ^s |\eta _m|^{2\kappa }\, \mathrm{d}x+c(\varepsilon ) \int _{\Omega }|\eta _m|^\vartheta \, \mathrm{d}x\, ,\nonumber \\ \vartheta:= \, & {} \Big [\frac{s}{s-\frac{\delta }{2}}\Big ]^* \Big (2\kappa - 2 {\hat{\kappa }} - 2 \kappa \frac{s-\frac{\delta }{2}}{s}\Big )\, , \end{aligned}$$
(71)

wtith exponent \(\vartheta \ge 0\), if we assume that

$$\begin{aligned} \kappa \ge 2 \frac{s}{\delta }{\hat{\kappa }} \end{aligned}$$
(72)

is satisfied. Choosing \(\varepsilon\) sufficiently small, inserting (71) in (70) and returning to (68) it is shown (compare (35))

$$\begin{aligned} \int _{\Omega }\Gamma ^s |\eta _m|^{2\kappa }\, \mathrm{d}x\le c\big \{(I)+1\big \}\, \quad (I) := \Bigg [\int _{\Omega }|\nabla \Gamma ^{\frac{s}{2}}| \,|\eta _m|^\kappa \, \mathrm{d}x\Bigg ]^2\, . \end{aligned}$$
(73)

Adjusting the calculations presented after inequality (35) to the situation at hand we consider numbers s, \(\alpha\) satisfying

$$\begin{aligned} s \le p + \alpha \end{aligned}$$
(74)

and find recalling (7) and using (74)

$$\begin{aligned} (I)\le \, & {} c \Bigg [\int _{\Omega }\Gamma ^{\frac{s-1}{2}}|\nabla ^2 u| |\eta _m|^\kappa \, \mathrm{d}x\Bigg ]^2\\ \le \, & {} c \Bigg [ \int _{\Omega }\Gamma ^{\frac{p-2}{2}} |\nabla ^2 u|^2 |\eta _m|^{2\kappa } \Gamma ^\alpha \, \mathrm{d}x\Bigg ] \cdot \Bigg [\int _{\Omega }\Gamma ^{s-1-\frac{p-2}{2}-\alpha }\Bigg ]\, \mathrm{d}x\\ \le \, & {} c \int _{\Omega }D^2f(\nabla u) \big (\partial _i \nabla u,\partial _i \nabla u\big ) |\eta _m|^{2\kappa } \Gamma ^\alpha \, \mathrm{d}x\, , \end{aligned}$$

where from now on the sum is taken with respect to the index i. Let us assume in addition to (74) that we have

$$\begin{aligned} \alpha > - \frac{1}{4} \end{aligned}$$
(75)

as lower bound for the parameter \(\alpha\). Quoting (107) from Proposition 6.2 (with \(\eta\) being replaced by \(|\eta _m|^\kappa\)) and returning to (73) we find:

$$\begin{aligned} \int _{\Omega }\Gamma ^s |\eta _m|^{2\kappa } \, \mathrm{d}x\le c \int _{\Omega }D^2f(\nabla u) \big (\nabla |\eta _m|,\nabla |\eta _m|\big ) \Gamma ^{1+\alpha } |\eta _m|^{2\kappa -2}\, \mathrm{d}x\, . \end{aligned}$$
(76)

On the right-hand side of (76) we use the second inequality from the ellipticity condition (7) to obtain (in analogy to the inequalities (43), (45))

$$\begin{aligned} \int _{\Omega }\Gamma ^s |\eta _m|^{2\kappa } \, \mathrm{d}x\le c \int _{\Omega }\Gamma ^{\frac{q}{2}+\alpha }|\nabla \eta _m|^2 |\eta _m|^{2\kappa -2} \, \mathrm{d}x\, . \end{aligned}$$
(77)

The right-hand side of (77) is discussed following the arguments presented after (45): we first have (recall (28))

$$\begin{aligned} \int _{\Omega }\Gamma ^{\frac{q}{2}+\alpha } |\nabla \varphi _m|^2 |u-u_0|^{2\kappa }\, \mathrm{d}x\le & {} \int _{{\text {spt}}\nabla \varphi _m} \Gamma ^{\frac{q}{2}+\alpha } m^{2-2\kappa \frac{p-2}{p}} \, \mathrm{d}x\\ \le \, & {} c \int _{\Omega }\Gamma ^{\frac{p}{2}}\, \mathrm{d}x+ \int _{{\text {spt}}\nabla \varphi _m} m^{\gamma ^*[2-2\kappa \frac{p-2}{p}]}\, \mathrm{d}x\, , \end{aligned}$$

where we have defined

$$\begin{aligned} \gamma := \frac{p/2}{\frac{q}{2} +\alpha }\, , \qquad \gamma ^* := \frac{\gamma }{\gamma -1}\, . \end{aligned}$$

Obviously this requires the bound \(\frac{p}{2} \ge \frac{q}{2}+\alpha\), i.e. the condition

$$\begin{aligned} q < p + \frac{1}{2} \end{aligned}$$
(78)

in combination with (75) yields the maximal range of anisotropy , since then we can choose \(\alpha\) sufficiently close to \(-1/4\) to guarantee \(p > q+2\alpha\). Moreover, we assume the validity of \(-1 + \gamma ^*[2-2\kappa (p-2)/p]\le 0\). This is true again for \(\alpha\) chosen sufficiently close to \(-1/4\), provided that

$$\begin{aligned} \kappa > \frac{p+q-\frac{1}{2}}{2(p-2)}\, . \end{aligned}$$
(79)

Recalling (78) we see that (79) is a consequence of the stronger bound

$$\begin{aligned} \kappa > \frac{p}{p-2} \, . \end{aligned}$$
(80)

In accordance with (47) we therefore arrive at

$$\begin{aligned} \hbox {r.h.s.~of~(4.10)} \le c \Bigg [1+\int _{\Omega }\Gamma ^{\frac{q}{2}+\alpha } |\nabla u - \nabla u_0|^2 |\eta _m|^{2\kappa -2}\, \mathrm{d}x\Bigg ]\, . \end{aligned}$$
(81)

Neglecting the contribution resulting from \(\nabla u_0\) and under the additional hypothesis

$$\begin{aligned} s > \frac{q}{2} + \frac{3}{4}\, , \end{aligned}$$
(82)

which guarantees the validity of \(s > \frac{q}{2}+1+\alpha\) for \(\alpha > -1/4\), we have

$$\begin{aligned} \int _{\Omega }\Gamma ^{\frac{q}{2}+1+\alpha } |\eta _m|^{2\kappa -2}\, \mathrm{d}x\le \, & {} \varepsilon \int _{\Omega }\Gamma ^s |\eta _m|^{2\kappa } \, \mathrm{d}x+ c(\varepsilon ) \int _{\Omega }|\eta _m|^{2\kappa - 2 \beta ^*}\, \mathrm{d}x\, , \end{aligned}$$
(83)
$$\begin{aligned} \beta ^* := \frac{\beta }{\beta -1}\, ,&\quad\beta := \frac{s}{1+\frac{q}{2}+\alpha }\, . \end{aligned}$$
(84)

Let us add a comment concerning the conditions imposed on s: as remarked after (78) and during the subsequent calculations the parameter \(\alpha\) has to be adjusted in an appropriate way and might become very close to the critical value \(-1/4\). For this reason inequality (74) is replaced by the stronger one (recall (16))

$$\begin{aligned} s < p - \frac{1}{4}\, . \end{aligned}$$
(85)

In order to find numbers s satisfying (82) and (85) we need the bound \(q < 2p-2\) being more restrictive for \(p \in (2, 5/2)\) than inequality (78).

Finally we note that the exponent \(2\kappa - 2 \beta ^*\) occurring in the second integral on the right-hand side of (83) is non-negative provided \(\kappa \ge s/(s-[1+\frac{q}{2}+\alpha ])\), and the latter inequality holds for \(\alpha\) sufficiently close to \(-1/4\), if (compare (17))

$$\begin{aligned} \kappa > \frac{s}{s- \big [\frac{q}{2}+\frac{3}{4}\big ]}\, . \end{aligned}$$
(86)

Assuming (86) our claim follows by inserting (83) into (81) and choosing \(\varepsilon\) sufficiently small. \(\square\)

5 Proof of Theorem 1.3

Let all the assumptions of Theorem 1.3 hold. We define the number

$$\begin{aligned} \gamma := \frac{2n}{n+2}\, , \quad 1\le \gamma < 2\, , \qquad \hbox {with Sobolev conjugate}\qquad \frac{n\gamma }{n-\gamma } = 2 \, , \end{aligned}$$

and note that

$$\begin{aligned} \frac{n- \gamma }{n} = \frac{n}{n+2} = \frac{\gamma }{2}\, , \quad \frac{\gamma }{2-\gamma } = \frac{n}{2} \, . \end{aligned}$$

We also remark that for any fixed \(\alpha\) and for any \(\kappa >1\) the inequality

$$\begin{aligned} \frac{p}{2} < p \frac{\kappa (n+2)-2}{2n\kappa } +\alpha \frac{\kappa -1}{\kappa }\, . \end{aligned}$$
(87)

is equivalent to the requirement \(\alpha > - p/n\).

Lemma 5.1

Fix \(\alpha > - 1/(2n)\), \(\kappa > 1\) with (87) and choose a real number \({\overline{s}}_\alpha\) satisfying

$$\begin{aligned} \frac{p}{2}< {\overline{s}}_\alpha < p \frac{\kappa (n+2)-2}{2n\kappa } + \alpha \frac{\kappa -1}{\kappa }\, . \end{aligned}$$
(88)

Then we have for any \(\varepsilon >0\) and for \(\eta \in C^{1}_0(\Omega )\)

$$\begin{aligned}&\int _{\Omega }\Gamma ^{{\overline{s}}_\alpha } |\eta |^{2(\kappa -1)} \, \mathrm{d}x\le \varepsilon \int _{\Omega }\Gamma ^{\frac{q}{2}+\alpha } |\nabla \eta |^{2} |\eta |^{2(\kappa -1)}\, \mathrm{d}x\nonumber \\&\quad +\varepsilon \int _{\Omega }\Gamma ^{-\frac{p}{n} + s} |\nabla \eta |^{2} |\eta |^{2(\kappa -1)}\, \mathrm{d}x+ c \end{aligned}$$
(89)

for a constant being independent on \(\eta\), where s is given by (90).

Proof of Lemma 5.1

We first observe that for

$$\begin{aligned} s = {\overline{s}}_\alpha \frac{\kappa }{\kappa -1} - \frac{p}{2(\kappa -1)} \end{aligned}$$
(90)

we have after an application of Young’s inequality

$$\begin{aligned} \int _{\Omega }\Gamma ^{{\overline{s}}_\alpha } |\eta |^{2(\kappa -1)} \, \mathrm{d}x= & {} \int _{\Omega }\Gamma ^{{\overline{s}}_\alpha -\frac{p}{2\kappa }} |\eta |^{2(\kappa -1)} \Gamma ^{\frac{p}{2\kappa }}\, \mathrm{d}x\le {\tilde{\varepsilon }} \int _{\Omega }\Gamma ^s |\eta |^{2 \kappa }\, \mathrm{d}x+ c({\tilde{\varepsilon }}) \int _{\Omega }\Gamma ^{\frac{p}{2}}\, \mathrm{d}x\end{aligned}$$
(91)

for any \({\tilde{\varepsilon }} > 0\). We then estimate using Sobolev’s inequality

$$\begin{aligned} \int _{\Omega }\Gamma ^s |\eta |^{2\kappa } \, \mathrm{d}x= & {} \int _{\Omega }\big [\Gamma ^{\frac{s}{2}} |\eta |^{\kappa }\big ]^2\, \mathrm{d}x\le c \Bigg [\int _{\Omega }\big | \nabla \big [ \Gamma ^{\frac{s}{2}} |\eta |^\kappa \big ]\big |^\gamma \, \mathrm{d}x\Bigg ]^{\frac{2}{\gamma }}\nonumber \\ \le \, & {} c \Bigg [ \int _{\Omega }\big |\nabla \Gamma ^{\frac{s}{2}}\big |^\gamma |\eta |^{\kappa \gamma }\, \mathrm{d}x\Bigg ]^{\frac{2}{\gamma }} + c(\kappa ) \Bigg [\int _{\Omega }\Gamma ^{\frac{s\gamma }{2}} |\eta |^{(\kappa -1)\gamma } |\nabla \eta |^{\gamma } \, \mathrm{d}x\Bigg ]^{\frac{2}{\gamma }}\nonumber \\ = \, & {} c T_1^{\frac{2}{\gamma }} + c T_2^{\frac{2}{\gamma }}\, . \end{aligned}$$
(92)

Let us first consider \(T_1\). Hölder’s inequality gives

$$\begin{aligned} T_1^\frac{2}{\gamma }\le & {} c \Bigg [ \int _{\Omega }|\nabla ^2 u|^\gamma \Gamma ^{\gamma \frac{s-1}{2}} |\eta |^{\kappa \gamma } \, \mathrm{d}x\Bigg ]^{\frac{2}{\gamma }}\nonumber \\= \, & {} c \Bigg [ \int _{\Omega }|\nabla ^2 u|^\gamma \Gamma ^{\gamma \frac{p-2}{4}} \Gamma ^{\gamma \frac{\alpha }{2}} \Gamma ^{\gamma \frac{2-p}{4}} \Gamma ^{-\gamma \frac{\alpha }{2}}\Gamma ^{\gamma \frac{s-1}{2}} |\eta |^{\kappa \gamma } \, \mathrm{d}x\Bigg ]^{\frac{2}{\gamma }}\nonumber \\\le \, & {} c \Bigg [ \int _{\Omega }|\nabla ^2 u|^2 \Gamma ^{\frac{p-2}{2}}\Gamma ^{\alpha }|\eta |^{2\kappa }\, \mathrm{d}x\Bigg ] \cdot \Bigg [ \int _{\Omega }\Gamma ^{\frac{\gamma }{2-\gamma }( \frac{2-p}{2}-\alpha +s-1)}\, \mathrm{d}x\Bigg ]^{\frac{2-\gamma }{\gamma }} \nonumber \\= \, & {} c T_{1,1} \cdot T_{1,2}^{\frac{2-\gamma }{\gamma }}\, . \end{aligned}$$
(93)

For \(T_{1,2}\) we observe recalling (90)

$$\begin{aligned} \frac{n}{2} \Big [ \frac{2-p}{2} -\alpha + s -1\Big ]< \frac{p}{2}\Leftrightarrow {\overline{s}}_\alpha < p \frac{\kappa (n+2)-2}{2n \kappa } +\alpha \frac{\kappa -1}{\kappa }\, , \end{aligned}$$

which is true by the choice (88) of \({\overline{s}}_\alpha\), hence \(T_{1,2}\) is uniformly bounded.

We handle \(T_{1,1}\) with the help of Proposition 6.2 (replacing \(\eta\) by \(|\eta |^\kappa\)):

$$\begin{aligned} T_{1,1}\le & {} c \int _{\Omega }D^2f(\nabla u) \big (\nabla \partial _\gamma u,\nabla \partial _\gamma u\big ) \Gamma ^{\alpha } |\eta |^{2\kappa } \, \mathrm{d}x\nonumber \\\le \, & {} c \int _{\Omega }D^2f(\nabla u)\big (\nabla |\eta |^\kappa , \nabla |\eta |^\kappa \big )|\nabla u|^2 \Gamma ^{\alpha }\, \mathrm{d}x\nonumber \\\le \, & {} c \int _{\Omega }\Gamma ^{\frac{q-2}{2}} |\eta |^{2\kappa -2}|\nabla \eta |^2 \Gamma ^{1+\alpha } \, \mathrm{d}x= c \int _{\Omega }\Gamma ^{\frac{q}{2} + \alpha } |\nabla \eta |^2 |\eta |^{2\kappa -2}\, \mathrm{d}x\, . \end{aligned}$$
(94)

From (92) - (94) we conclude

$$\begin{aligned} \int _{\Omega }\Gamma ^s |\eta |^{2\kappa } \, \mathrm{d}x\le c \int _{\Omega }\Gamma ^{\frac{q}{2}+\alpha } |\nabla \eta |^2 |\eta |^{2(\kappa -1)}\, \mathrm{d}x+ c T_2^{\frac{2}{\gamma }} \end{aligned}$$
(95)

with constants c being independent of \(\eta\) and it remains to discuss \(T_2\) in (95): we have for \({\hat{\mu }} =2/\gamma\), \(\mu = 2/(2-\gamma )\),

$$\begin{aligned} T_2^{\frac{2}{\gamma }}= & {} \Bigg [\int _{\Omega }\Gamma ^{\frac{s\gamma }{2}}|\nabla \eta |^\gamma |\eta |^{(\kappa -1)\gamma }\, \mathrm{d}x\Bigg ]^{\frac{2}{\gamma }} = \Bigg [\int _{\Omega }\Gamma ^{\frac{p}{2\mu }} \Gamma ^{-\frac{p}{2\mu }} \Gamma ^{\frac{s\gamma }{2}}|\nabla \eta |^\gamma |\eta |^{(\kappa -1)\gamma }\, \mathrm{d}x\Bigg ]^{\frac{2}{\gamma }}\nonumber \\\le \, & {} c \Bigg [\int _{\Omega }\Gamma ^{\frac{p}{2}}\, \mathrm{d}x\Bigg ]^{\frac{2}{n}} \cdot \int _{\Omega }\Gamma ^{-\frac{p}{n}+s}|\nabla \eta |^{2}|\eta |^{2(\kappa -1)}\, \mathrm{d}x\, . \end{aligned}$$
(96)

With (91) for \({\tilde{\varepsilon }} < \varepsilon /c\) sufficiently small, (95) and (96) the proof of Lemma 5.1 is completed. \(\square\)

Now we come to the proof of the theorem: we choose \(\eta _m = (u - u_0)\varphi _m\) with \(\varphi _m\) defined after (26). Then Lemma 5.1 yields

$$\begin{aligned} \int _{\Omega }\Gamma ^{{\overline{s}}_\alpha } |\eta _m|^{2(\kappa -1)} \, \mathrm{d}x\le & {} \varepsilon \int _{\Omega }\Gamma ^{\frac{q}{2}+\alpha } \big [|\nabla u|+|\nabla u_0|\big ]^{2} |\eta _m|^{2(\kappa -1)}\, \mathrm{d}x\nonumber \\ &+\varepsilon \int _{\Omega }\Gamma ^{-\frac{p}{n} + s} \big [|\nabla u| + |\nabla u_0|\big ]^{2} |\eta _m|^{2(\kappa -1)}\, \mathrm{d}x\nonumber \\&+ \varepsilon \int _{\Omega }\Gamma ^{(\frac{q}{2}+\alpha )} |\nabla \varphi _m|^2 |u-u_0|^{2}|\eta _m|^{2(\kappa -1)}\, \mathrm{d}x\nonumber \\&+\varepsilon \int _{\Omega }\Gamma ^{-\frac{p}{n} + s} |\nabla \varphi _m|^2 |u-u_0|^2 |\eta _m|^{2(\kappa -1)}\, \mathrm{d}x+ c \, . \end{aligned}$$
(97)

Here the first two integrals on the right-hand side can be absorbed in the left-hand side provided that we have (recall that \(u_0\) is Lipschitz)

$$\begin{aligned} \max \Big \{\frac{q}{2}+\alpha , -\frac{p}{n} + s \Big \}< {\overline{s}}_\alpha -1 \, . \end{aligned}$$
(98)

Also on account of (98) we can handle the remaining integrals on the right-hand side of (97) with the help of Young’s inequality. We obtain

$$\begin{aligned}&{ \int _{\Omega }\Gamma ^{\frac{q}{2}+\alpha } |\nabla \varphi _m|^2 |u-u_0|^{2}|\eta _m|^{2(\kappa -1)}\, \mathrm{d}x}\nonumber \\ & \quad \le {} \int _{\Omega }\Gamma ^{{\overline{s}}_\alpha } |\eta _m|^{2(\kappa -1)} \, \mathrm{d}x+ c \int _{{\text {spt}} \nabla \varphi _m} |\nabla \varphi _m|^{2\beta _1^*} |u-u_0|^{2\beta _1^*} |\eta _m|^{2(\kappa -1)} \, \mathrm{d}x\, . \end{aligned}$$
(99)

as well as

$$\begin{aligned}&{ \int _{\Omega }\Gamma ^{-\frac{p}{n}+s} |\nabla \varphi _m|^2 |u-u_0|^{2}\eta _m^{2(\kappa -1)}\, \mathrm{d}x}\nonumber \\ & \quad \le {} \int _{\Omega }\Gamma ^{{\overline{s}}_\alpha } \eta _m^{2(\kappa -1)} \, \mathrm{d}x+ c \int _{{\text {spt}} \nabla \varphi _m} |\nabla \varphi _m|^{2\beta _2^*} |u-u_0|^{2\beta _2^*} \eta _m^{2(\kappa -1)} \, \mathrm{d}x\, . \end{aligned}$$
(100)

In (99) we choose \(\beta _1\) such that

$$\begin{aligned} 1< \beta _1 < \Bigg [p \frac{\kappa (n+2)-2}{\kappa } - \frac{\kappa -1}{\kappa }\Bigg ] \frac{1}{qn -1} \end{aligned}$$

with conjugate exponent (recall \(q-p < 2p/n\) on account of (18))

$$\begin{aligned} \beta _1^* > \frac{\kappa \big [p(n+2)-1\big ]-2p+1}{\kappa \big [2p - (q-p)n\big ]-2p+1}\, . \end{aligned}$$
(101)

Note that for \(\alpha\) sufficiently close to \(-1/(2n)\) it holds (recall 88)

$$\begin{aligned} 1< \beta _1 < {\overline{s}}_\alpha \Big [\frac{q}{2}-\frac{1}{2n}\Big ]^{-1} \, . \end{aligned}$$

In (100) we define \(\beta _2\) according to

$$\begin{aligned} 1< \beta _2 < \Bigg [p \frac{\kappa (n+2)-2}{2\kappa } - \frac{\kappa -1}{2\kappa }\Bigg ] \frac{1}{sn -p} \end{aligned}$$

with conjugate exponent

$$\begin{aligned} \beta _2^* > \frac{\kappa \big [pn +4p -1- 2sn\big ]-2p+1}{\kappa \big [2p - (q-p)n\big ]-2p+1} \, . \end{aligned}$$
(102)

and choosing \(\alpha\) close to \(-1/(2n)\) we can arrange that (recall 88)

$$\begin{aligned} 1< \beta _2 < {\overline{s}}_\alpha \Big [-\frac{p}{n}+s\Big ]^{-1}\, . \end{aligned}$$

Returning to (99) and (100), respectively, the first integral again is absorbed in the left-hand side of (97). The remaining integrals stay bounded if we have for \(i=1\), 2

$$\begin{aligned} m^{-1} m^{\beta _i^*\big [2 -2(1-n/p)\big ]-2(\kappa -1)(1-n/p)} \le c\, , \end{aligned}$$

Thus we require the condition

$$\begin{aligned} \beta _i^* < (\kappa -1)\frac{p-n}{n} + \frac{p}{2n} \, , \qquad i=1,\, 2\, , \end{aligned}$$
(103)

which on account of \(p>n\) is satisfied for \(\kappa\) sufficiently large.

It remains to arrange (98) together with the inequality on the right-hand side of (88). Here we first observe

$$\begin{aligned} \frac{q}{2} + \alpha< p \frac{\kappa (n+2)-2}{2n\kappa } + \alpha \frac{\kappa -1}{\kappa } -1\Leftrightarrow & {} q< p \frac{\kappa (n+2)-2}{n \kappa } - 2 - 2 \frac{\alpha }{\kappa } \nonumber \\\Leftrightarrow \, & {} q-p < \frac{2}{n} (p-n) - \frac{2p+2n\alpha }{n\kappa } \end{aligned}$$
(104)

For \(\alpha\) sufficiently close to \(-1/(2n)\), (104) is a consequence of (19). We finally have to discuss

$$\begin{aligned} -\frac{p}{n} + s \le {\overline{s}}_\alpha -1\Leftrightarrow & {} -\frac{p}{n} + {\overline{s}}_\alpha \frac{\kappa }{\kappa -1} - \frac{p}{2(\kappa -1)}< {\overline{s}}_\alpha - 1\nonumber \\\Leftrightarrow & {} {\overline{s}}_\alpha < (\kappa -1) \frac{p-n}{n} + \frac{p}{2}\, . \end{aligned}$$
(105)

Assumption (21) implies (105). Hence, (97), (99) and (100) prove Theorem 1.3 passing to the limit \(m\rightarrow \infty\). \(\square\)

6 Appendix. Caccioppoli-type inequalities

We prove two Caccioppoli-type inequalities with small weights (i.e. involving powers of \(\Gamma = 1+ |\nabla u|^2\) or of \(\Gamma _i =1+|\partial _i u|^2\), \(i=1\), ..., n, with a certain range of negative exponents), where the first one is the appropriate version in the splitting context.

There is no need to restrict the following considerations to the case \(n=2\). Thus, throughout this appendix, we suppose that \(\Omega \subset {\mathbb {R}}^n\) is a bounded Lipschitz domain and that f: \({\mathbb {R}}^n \rightarrow {\mathbb {R}}\) is of class \(C^2\) satisfying \(D^2f(Z)(Y,Y) > 0\) for all Z, \(Y \in {\mathbb {R}}^n\).

Moreover we suppose that \(u \in W^{2,2}_{{\text {loc}}}(\Omega )\cap C^{1}(\Omega )\) solves the differentiated Euler equation

$$\begin{aligned} 0 = \int _{\Omega }D^2 f(\nabla u) \big (\nabla \partial _i u, \nabla \psi \big ) \, \mathrm{d}x\qquad \hbox {for all}\; \psi \in C^\infty _0(\Omega ) \end{aligned}$$
(106)

and for any \(1 \le i \le n\) fixed.

Proposition 6.1

Fix \(l\in {\mathbb {N}}\) and suppose that \(\eta \in C^\infty _0(\Omega )\), \(0 \le \eta \le 1\). Then the inequality

$$\begin{aligned}&{\int _{\Omega }D^2 f(\nabla u)\big (\nabla \partial _i u, \nabla \partial _i u\big ) \eta ^{2l} \Gamma _{i}^{\alpha } \, \mathrm{d}x}\nonumber \\ & \quad \le c \int _{\Omega }D^2f(\nabla u) (\nabla \eta ,\nabla \eta )\eta ^{2l-2} \Gamma _{i}^{\alpha +1} \, \mathrm{d}x\, , \quad \Gamma _i := 1+|\partial _i u|^2\, , \end{aligned}$$
(107)

holds for any \(\alpha > - 1/2\) and for any fixed \(1\le i \le n\).

Proof of Proposition 6.1

Suppose that \(-1/2 < \alpha\) and fix \(1 \le i \le n\) (no summation with respect to i). Using approximation arguments we may insert

$$\begin{aligned} \psi := \eta ^{2l} \partial _i u \Gamma _{i}^{\alpha } \end{aligned}$$

in the equation (106) with the result

$$\begin{aligned} \int _{\Omega }D^2f(\nabla u) \big (\nabla \partial _i u , \nabla \partial _i u\big ) \eta ^{2l} \Gamma _{i}^{\alpha }\, \mathrm{d}x= & {} - \int _{\Omega }D^2f(\nabla u)\big (\nabla \partial _i u, \nabla \Gamma _{i}^{\alpha }\big ) \partial _i u \eta ^{2l}\, \mathrm{d}x\nonumber \\&- \int _{\Omega }D^2f(\nabla u)\big (\nabla \partial _i u, \nabla (\eta ^{2l})\big ) \partial _i u \Gamma _{i}^{\alpha } \, \mathrm{d}x\nonumber \\=: & {} S_1+S_2 \, . \end{aligned}$$
(108)

In (108) we have

$$\begin{aligned} S_1 = - 2 \alpha \int _{\Omega }D^2f (\nabla u) \big (\nabla \partial _i u, \nabla \partial _i u\big ) |\partial _i u|^2\Gamma _{i}^{\alpha -1}\eta ^{2l}\, \mathrm{d}x\end{aligned}$$

which gives \(S_1 \le 0\) if \(\alpha \ge 0\). In this case we will just neglect \(S_1\) in the following. In the case \(-1/2< \alpha < 0\) we estimate

$$\begin{aligned} |S_1|= \, & {} 2 |\alpha | \int _{\Omega }D^2f (\nabla u) \big (\nabla \partial _i u, \nabla \partial _i u\big ) |\partial _i u|^2\Gamma _{i}^{\alpha -1}\eta ^{2l}\, \mathrm{d}x\\ \le \, & {} 2 |\alpha | \int _{\Omega }D^2f (\nabla u) \big (\nabla \partial _i u, \nabla \partial _i u\big )\Gamma _{i}^{\alpha }\eta ^{2l}\, \mathrm{d}x\, . \end{aligned}$$

Since we have \(2 |\alpha | < 1\) we may absorb \(|S_1|\) in the left-hand side of (108), hence

$$\begin{aligned} \int _{\Omega }D^2f (\nabla u) \big (\nabla \partial _i u, \nabla \partial _i u\big ) \eta ^{2l}\Gamma _{i}^{\alpha }\, \mathrm{d}x\le c |S_2| \, . \end{aligned}$$

For \(0 < \varepsilon\) sufficiently small we apply the Cauchy-Schwarz inequality to discuss \(S_2\):

$$\begin{aligned}&{\int _{\Omega }D^2f(\nabla u) (\nabla \partial _i u, \nabla \eta )\eta ^{2l-1} \Gamma _{i}^{\alpha } \partial _iu \, \mathrm{d}x}\\ & \quad \le {} \varepsilon \int _{\Omega }D^2f(\nabla u) (\nabla \partial _i u ,\nabla \partial _i u) \eta ^{2l} \Gamma _{i}^{\alpha }\, \mathrm{d}x\\ & \qquad + c(\varepsilon ) \int _{\Omega }D^2f(\nabla u) (\nabla \eta ,\nabla \eta ) \eta ^{2l-2}\Gamma _{i}^{\alpha } |\partial _i u|^2 \, \mathrm{d}x\, . \end{aligned}$$

After absorbing the first term in the right-hand side of (108) we have established our claim (107). \(\square\)

Instead of the quantities \(\Gamma _i\) our second inequality (compare [20] for the discussion in two dimensions) involves the full derivative, i.e. we incorporate certain negative powers of \(\Gamma = 1+|\nabla u|^2\). As a consequence we do not obtain the range \(-1/2 < \alpha\) and have to replace this condition by the requirement \(-1/(2n) < \alpha\).

Proposition 6.2

Suppose that \(\eta \in C^{\infty }_{0}\big (\Omega \big )\) and fix some real number \(\alpha\) such that \(- 1/(2n) < \alpha\). Then we have (summation with respect to \(i =1\), ..., n)

$$\begin{aligned}&{\big [1 + 2 \alpha n\big ] \int _{\Omega }D^2 f (\nabla u) \big (\nabla \partial _i u, \nabla \partial _i u\big ) \Gamma ^{\alpha }\eta ^2 \, \mathrm{d}x}\nonumber \\ & \quad \le {} c \Bigg [ \int _{{\text {spt}}\nabla \eta } D^2 f (\nabla u) \big (\nabla \partial _i u, \nabla \partial _i u\big ) \Gamma ^{\alpha }\eta ^2 \, \mathrm{d}x\Bigg ]^{\frac{1}{2}} \nonumber \\&\qquad \cdot \Bigg [ \int _{{\text {spt}}\nabla \eta } D^2f(\nabla u)\big (\nabla \eta , \nabla \eta \big ) \big |\nabla u\big |^2 \Gamma ^{\alpha } \, \mathrm{d}x\Bigg ]^{\frac{1}{2}}\, , \end{aligned}$$
(109)

where the constant c is not depending on \(\eta\). In particular it holds

$$\begin{aligned} \int _{\Omega }D^2 f (\nabla u) \big (\nabla \partial _i u, \nabla \partial _i u) \Gamma ^{\alpha } \eta ^2\, \mathrm{d}x\le c \int _{{\text {spt}}\nabla \eta } D^2f(\nabla u)\big (\nabla \eta , \nabla \eta \big ) \big |\nabla u \big |^2\Gamma ^{\alpha } \, \mathrm{d}x\, . \end{aligned}$$
(110)

Proof

For \(i =1\), ..., n and any \(\eta\) as above we deduce from (106)

$$\begin{aligned}&{\int _{\Omega }D^2f(\nabla u) \big (\nabla \partial _i u ,\nabla \partial _i u\big ) \Gamma ^{\alpha } \eta ^2 \, \mathrm{d}x}\nonumber \\ & \quad = {} - \int _{\Omega }D^2f(\nabla u)\big (\nabla \partial _i u, \partial _i u \nabla \Gamma ^{\alpha }\big ) \eta ^2 \, \mathrm{d}x\nonumber \\ & \qquad - 2 \int _{\Omega }D^2 f(\nabla u) \big (\nabla \partial _i u , \nabla \eta \big ) \partial _i u \Gamma ^{\alpha }\eta \, \mathrm{d}x=: I +II\, , \end{aligned}$$
(111)

where at this stage no summation with respect to the index i is performed. Following [20] we denote the bilinear form \(D^2f(\cdot ,\cdot )\) by \(\langle \cdot ,\cdot \rangle\). The first observation is the inequality

$$\begin{aligned} \sum _{i =1}^n \Big \langle \partial _i \nabla u, \partial _i \nabla u\Big \rangle \Gamma ^{\alpha }\ge & {} \sum _{i =1}^n \Big \langle \partial _i \nabla u, \partial _i \nabla u\Big \rangle \sum _{j=1}^n (\partial _j u)^2 \Gamma ^{\alpha -1}\nonumber \\\ge & {} \sum _{i =1}^n \Big \langle \partial _i \nabla u, \partial _i \nabla u\Big \rangle (\partial _i u)^2 \Gamma ^{\alpha -1}\nonumber \\= & {} \sum _{i =1}^n \Big \langle \partial _{i} u \partial _i \nabla u , \partial _i u \partial _i \nabla u \Big \rangle \Gamma ^{\alpha -1} \, . \end{aligned}$$
(112)

For handling the integrand of the term I from (111) we use the identity

$$\begin{aligned} - \sum _{i=1}^n\Big \langle \partial _i \nabla u , \partial _i u \nabla \Gamma ^{\alpha }\Big \rangle= & {} - \alpha \sum _{i=1}^n \Big \langle \partial _i u \partial _i \nabla u , \nabla \sum _{j=1}^n (\partial _j u )^2\Big \rangle \Gamma ^{\alpha -1}\nonumber \\= & {} - 2\alpha \sum _{i =1}^n \Big \langle \partial _i u \partial _i \nabla u, \sum _{j =1}^n \partial _j u \partial _j \nabla u\Big \rangle \Gamma ^{\alpha -1}\nonumber \\= & {} -2 \alpha \sum _{i =1}^n \Big \langle \partial _i u \partial _i \nabla u, \partial _i u \partial _i \nabla u\Big \rangle \Gamma ^{\alpha -1}\nonumber \\&- 2 \alpha \sum _{i =1}^n \Big \langle \partial _i u \partial _i \nabla u, \sum _{j\not =i}\partial _j u \partial _j \nabla u\Big \rangle \Gamma ^{\alpha -1} \, . \end{aligned}$$
(113)

The last term on the right-hand side of (113) is estimated as follows

$$\begin{aligned}&{\sum _{i =1}^n \Big \langle \partial _i u \partial _i \nabla u, \sum _{j\not =i}\partial _j u \partial _j \nabla u\Big \rangle }\\ & \quad \le {} \sum _{i=1}^n \Bigg [\sum _{j\not = i} \frac{1}{2} \Big [ \Big \langle \partial _i u \partial _i \nabla u, \partial _i u \partial _i \nabla u\Big \rangle +\Big \langle \partial _j u \partial _j \nabla u, \partial _j u \partial _j \nabla u\Big \rangle \Big ]\Bigg ]\\ & = {} \frac{1}{2} \sum _{i =1}^n \Bigg [ (n-1) \Big \langle \partial _i u \partial _i \nabla u, \partial _i u \partial _i \nabla u\Big \rangle + \sum _{j\not =i} \Big \langle \partial _j u \partial _j \nabla u, \partial _j u \partial _j \nabla u\Big \rangle \Big ]\Bigg ]\\ & = {} \frac{1}{2} \sum _{i =1}^n \Bigg [ (n-2) \Big \langle \partial _i u \partial _i \nabla u, \partial _i u \partial _i \nabla u\Big \rangle + \sum _{j=1}^n \Big \langle \partial _j u \partial _j \nabla u, \partial _j u \partial _j \nabla u\Big \rangle \Big ]\Bigg ]\\ & = {} \frac{1}{2} \Bigg [ (n-2) \sum _{i =1}^n \Big \langle \partial _i u \partial _i \nabla u, \partial _i u \partial _i \nabla u\Big \rangle + n \sum _{j=1}^n \Big \langle \partial _j u \partial _j \nabla u, \partial _j u \partial _j \nabla u\Big \rangle \Big ]\Bigg ]\\ & = {} (n-1) \sum _{i=1}^n \Big \langle \partial _i u \partial _i \nabla u, \partial _i u \partial _i \nabla u\Big \rangle \, . \end{aligned}$$

This, together with (113) gives (recall \(-2\alpha > 0\))

$$\begin{aligned} - \sum _{i=1}^n\Big \langle \partial _i \nabla u , \partial _i u \nabla \Gamma ^{\alpha }\Big \rangle= & {} -2 \alpha \sum _{i =1}^n \Big \langle \partial _i u \partial _i \nabla u, \partial _i u \partial _i \nabla u\Big \rangle \Gamma ^{\alpha -1}\nonumber \\&- 2 \alpha \sum _{i =1}^n \Big \langle \partial _i u \partial _i \nabla u, \sum _{j\not =i}\partial _j u \partial _j \nabla u\Big \rangle \Gamma ^{\alpha -1} \nonumber \\\le & {} -2 \alpha \sum _{i =1}^n \Big \langle \partial _i u \partial _i \nabla u, \partial _i u \partial _i \nabla u\Big \rangle \Gamma ^{\alpha -1}\nonumber \\&- 2 \alpha (n-1) \sum _{i =1}^n \Big \langle \partial _i u \partial _i \nabla u, \partial _i u \partial _i \nabla u\Big \rangle \Gamma ^{\alpha -1} \nonumber \\= & {} - 2 \alpha n \sum _{i =1}^n \Big \langle \partial _i u \partial _i \nabla u, \partial _i u \partial _i \nabla u\Big \rangle \Gamma ^{\alpha -1} \, . \end{aligned}$$
(114)

Combining (114) and (112) we get

$$\begin{aligned}&{-\sum _{i =1}^n \int _{\Omega }D^2f(\nabla u)\big ( \partial _i \nabla u , \partial _i u \nabla \Gamma ^{\alpha }\big )\eta ^2 \, \mathrm{d}x} \nonumber \\ & \quad \le -2 \alpha n \sum _{i =1}^n \int _{\Omega }D^2f(\nabla u) \big (\partial _i \nabla u, \partial _i \nabla u\big )\Gamma ^{\alpha } \eta ^2 \, \mathrm{d}x\, . \end{aligned}$$
(115)

Returning to (111) and using (115) we get (from now on summation with respect to i)

$$\begin{aligned}&{\big [1+ 2 \alpha n \big ] \int _{\Omega }D^2 f(\nabla u) \big (\nabla \partial _i u , \nabla \partial _i u\big ) \Gamma ^{\alpha } \eta ^2 \, \mathrm{d}x}\nonumber \\& \quad \le -2 \int _{{\text {spt}}\nabla \eta } D^2 f(\nabla u) \big (\eta \nabla \partial _i u, \partial _i u \nabla \eta \big ) \Gamma ^{\alpha } \, \mathrm{d}x\, . \end{aligned}$$
(116)

This finishes the proof by applying the Cauchy-Schwarz inequality. \(\square\)