1 Introduction and main results

This work is devoted to the study of the following anisotropic doubly critical problem

$$\begin{aligned} {\left\{ \begin{array}{ll} \displaystyle -\Delta _p^Hu- \frac{\gamma }{H^\circ (x)^p} u^{p-1}= u^{p^*-1} \qquad &{}\text {in } {\mathbb {R}}^N\\ u>0 &{} \text {in } {\mathbb {R}}^N\\ u \in {\mathcal {D}}^{1,p}({\mathbb {R}}^N), \end{array}\right. }\qquad \qquad \qquad ({\mathcal {P}}_H) \end{aligned}$$

where \(1<p<N\), \(p^*:=Np/(N-p)\) is the Sobolev critical exponent, \(\displaystyle 0 \le \gamma < C_H:=\left( (N-p)/p\right) ^p\) is the Hardy constant and

$$\begin{aligned} \Delta _p^Hu:={\text {div}}(H^{p-1}(\nabla u)\nabla H(\nabla u)), \end{aligned}$$
(1.1)

where \(\Delta _p^H\) is the so-called anisotropic p-Laplacian or Finsler p-Laplacian. We point out that H is a Finsler type norm and \(H^\circ \) is its the dual norm (H satisfies assumptions \((h_H)\), see Sect. 2 for further details). In particular, when \(H(\xi )=|\xi |=H^\circ (\xi )\) the Finsler type p-Laplacian coincides with the classical p-Laplacian, and, hence it is singular when \(1<p<2\) and degenerate when \(p>2\). Then, according with standard regularity theory [13, 27] and the regularity results in the anisotropic framework [2, 6], we say that any solution of (\({\mathcal {P}}_H\)) has to be understood in the weak distributional meaning, i.e. \(u \in D^{1,p}({\mathbb {R}}^N)\) satisfies the following integral equality

$$\begin{aligned}{} & {} \int _{{\mathbb {R}}^N}\left( H^{p-1}(\nabla u) \langle \nabla H(\nabla u), \nabla \varphi \rangle - \frac{\gamma }{H^\circ (x)^p} u^{p-1} \varphi \right) \, dx\nonumber \\{} & {} \quad = \int _{{\mathbb {R}}^N} u^{p^*-1} \varphi \, dx \quad \forall \varphi \in C^\infty _c({\mathbb {R}}^N). \end{aligned}$$
(1.2)

The literature about critical problems is really huge. Going back to the Euclidean framework, i.e. when we consider \(H(\xi )=|\xi |=H^\circ (\xi )\) in (\({\mathcal {P}}_H\)), we deal with

$$\begin{aligned} -\Delta _p u - \frac{\gamma }{|x|^p} u^{p-1} = u^{p^*-1} \qquad \text {in } {\mathbb {R}}^N. \end{aligned}$$
(1.3)

In the seminal paper [5], Caffarelli, Gidas and Spruck classified any positive solution to (1.3) with \(p=2\), \(N \ge 3\), and \(\gamma =0\). We point out that a first result, under stronger assumption on the decay of solutions, was obtained by Gidas, Ni and Nirenberg in [14]. Moreover, in this setting a complete answer in the subcritical case was done in the celebrasted work of Gidas and Spruck [15], where the authors proved Liouville-type theorems.

In the quasilinear framework, the situation is much more involved due to the nonlinear nature of the operator. Recently, a classification result of positive solutions to (1.3) with \(p > 2\), \(\gamma =0\) and \(u \in {\mathcal {D}}^{1,p}({\mathbb {R}}^N):=\{u \in L^{p^*}({\mathbb {R}}^N) \ | \ \nabla u \in L^p({\mathbb {R}}^N)\}\) was obtained in [24]. The proof of this result is based on a refined version of the well-known moving plane method of Alexandrov-Serrin [1, 23] and on some a priori estimates of the solutions and their gradients, proved in [29]. To be more precise, we note that the classification result of positive solution to the Sobolev critical quasilinear equation with finite energy started in [12] in the case, and then was extended in [29] for every \(1<p<2\). Subsequently the full case was obtained in [24]. Recently, we refer to the papers [7, 21, 30] for new partial results on the classification of positive solutions without a priori assumption on the energy of solutions. In the anisotropic setting, Ciraolo, Figalli and Roncoroni [9], obtained a complete classification result for positive solution to (\({\mathcal {P}}_H\)) with \(\gamma =0\) using different techniques that do not require the use of the moving plane method, which could not be used in the anisotropic context due to the lack of invariance.

When \(\gamma \ne 0\) the situation is really different. In the seminal paper of Terracini [26], it was proved for the first time the classification result for positive solutions to (1.3) in the case \(p=2\). The author firstly showed the existence of solutions to this problem with a minimization argument based on the concentration and compactness principle. Subsequently, she proved that any solution to this problem is radial and radially decreasing about the origin combining the moving-plane technique and the use of the Kelvin transformation, in the same spirit of [5]. The case \(p \ne 2\) and \(\gamma \ne 0\) is much more involved and it is available in [20], where the techniques used are mainly based on a fine asymptotic analysis at infinity and refined versions of the moving plane procedure, and also on some asymptotic estimates proved in [31, 32].

Our aim is to prove some decay estimates for positive weak solutions to (\({\mathcal {P}}_H\)) in the anisotropic framework \(1<p<N\) and \(\gamma \ne 0\). More precisely, our first main result is the following:

Theorem 1.1

Let \(u \in {\mathcal {D}}^{1,p}({\mathbb {R}}^N)\) be a weak solution of (\({\mathcal {P}}_H\)) with \(1<p<N\), \(0 \le \gamma < C_H\). Then there exist positive constants \(0< R_1< 1 < R_2\) depending on \(N, p, \gamma \) and u, such that

$$\begin{aligned} \frac{c_1}{[H^\circ (x)]^{\mu _1}} \le u(x) \le \frac{C_1}{[H^\circ (x)]^{\mu _1}} \qquad x \in {\mathcal {B}}_{R_1}^{H^\circ }, \end{aligned}$$
(1.4)

and

$$\begin{aligned} \frac{c_2}{[H^\circ (x)]^{\mu _2}} \le u(x) \le \frac{C_2}{[H^\circ (x)]^{\mu _2}} \qquad x \in ({\mathcal {B}}_{R_2}^{H^\circ })^c, \end{aligned}$$
(1.5)

where \(\mu _1\), \(\mu _2\) are the solutions of

$$\begin{aligned} \mu ^{p-2} [(p-1)\mu ^2-(N-p)\mu ] + \gamma =0, \end{aligned}$$
(1.6)

\(C_1, C_2\) are positive constants depending on \(N,p, \gamma , H\) and u, \(c_1\) is a positive constant depending on \(N,p, \gamma , H, R_1, \mu _1\) and u, \(c_2\) is a positive constant depending on \(N,p, \gamma , H, R_2, \mu _2\) and u.

Remark 1.2

In the following we shall assume that \(\mu _1<\mu _2\) and it is easy to see that

$$\begin{aligned} 0 \le \mu _1< \frac{N-p}{p} < \mu _2 \le \frac{N-p}{p-1}. \end{aligned}$$

furthermore \({\mathcal {B}}_R^{H^\circ }\) is the dual anisotropic ball also known as Frank diagram (see Sect. 2 for further details).

In the proof we will also exploit some clever ideas from [31] facing the difficulties of the anisotropic issue. A different approach is in fact needed for the study of the asymptotic behaviour of the gradient. In particular, the fact that the moving plane plane technique cannot be applied, a crucial point is given by the following classification result:

Theorem 1.3

Let \(v\in C^{1,\alpha }_{loc}({\mathbb {R}}^N\setminus \{0\})\) be a positive weak solution of the equation

$$\begin{aligned} -\Delta _p^H v - \frac{\gamma }{[H^\circ (x)]^p} v^{p-1} = 0 \ \text {in } {\mathbb {R}}^N\setminus \{0\}, \end{aligned}$$
(1.7)

where \(0 \le \gamma < C_H\). Assume that there exist two positive constants C and c such that

$$\begin{aligned} \frac{c}{[H^{\circ }(x)]^{\mu _i}}\le v(x)\le \frac{C}{[H^{\circ }(x)]^{\mu _i}} \ \forall x\in {\mathbb {R}}^N\setminus \{0\}, \end{aligned}$$
(1.8)

where \(\mu _i\) (\(i=1,2\)) are the roots of (1.6) and suppose that there exists a positive constant \({{\hat{C}}}\) such that

$$\begin{aligned} |\nabla v(x)|\le \frac{{{\hat{C}}}}{[H^{\circ }(x)]^{\mu _i+1}} \ \ \forall x\in {\mathbb {R}}^N\setminus \{0\}, \end{aligned}$$
(1.9)

then

$$\begin{aligned} v(x)=\frac{{{\overline{c}}}}{[H^{\circ }(x)]^{\mu _i}}, \qquad \qquad \qquad \end{aligned}$$
(1.10)

for some \({{\bar{c}}}>0\).

Theorem 1.3 is new and interesting in itself. The proof is very much different than the ones available in the euclidean case \(H(\xi )=|\xi |\). Here we shall exploit it to deduce the precise asymptotic estimates for the gradient. more precisely we have the following:

Theorem 1.4

Let \(u \in {\mathcal {D}}^{1,p}({\mathbb {R}}^N)\) be a weak solution of (\({\mathcal {P}}_H\)) with \(1<p<N\), and \(0 \le \gamma < C_H\). Then there exist positive constants \({{\tilde{c}}}\), \({{\tilde{C}}}\) depending on \(N,p, \gamma , H\) and u such that

$$\begin{aligned} \frac{{{\tilde{c}}}}{[H^\circ (x)]^{\mu _1+1}} \le |\nabla u(x)| \le \frac{{\tilde{C}}}{[H^\circ (x)]^{\mu _1+1}} \qquad x \in {\mathcal {B}}_{R_1}^{H^\circ }, \end{aligned}$$
(1.11)

and

$$\begin{aligned} \frac{{{\tilde{c}}}}{[H^\circ (x)]^{\mu _2+1}} \le |\nabla u(x)| \le \frac{{{\tilde{C}}}}{[H^\circ (x)]^{\mu _2+1}} \qquad x \in ({\mathcal {B}}_{R_2}^{H^\circ })^c, \end{aligned}$$
(1.12)

where \(\mu _1, \mu _2\) are roots of (1.6) as in Theorem 1.1, and \(0< R_1< 1 < R_2\) are constants depending on \(N, p, \gamma \) and u.

The paper is structured as follows:

  • In Sect. 2 we recall some notions about Finsler type anisotropic geometry, and we prove some technical lemmas that will be crucial in the proof of the main results.

  • In Sect. 3 we prove some preliminary estimates, elliptic estimates and weak comparison principles in bounded and exterior domains that will be essential in the proof of Theorem 1.1.

  • In Sect. 4 we give the proof of decay estimates of solutions to (\({\mathcal {P}}_H\)) near the origin and at infinity, i.e. we prove Theorem 1.1. The, using this result we also prove decay estimates for the gradient of positive weak solutions to (\({\mathcal {P}}_H\)) near the origin and at infinity, i.e. we prove Theorem 1.4.

  • Although the existence of solutions can be easy deduced in the radial-anisotropic setting, in the “Appendix A” we show that problem (\({\mathcal {P}}_H\)) admits at least a positive solution \(u \in D^{1,p}({\mathbb {R}}^N)\) that minimizes the Hardy-Sobolev anisotropic inequality. This result follows using classical arguments (see also [26]) that we decide to add for the readers’ convenience.

2 Preliminaries

Notation. Generic fixed and numerical constants will be denoted by C (with subscript in some case) and they will be allowed to vary within a single line or formula. By |A| we will denote the Lebesgue measure of a measurable set A.

The aim of this section is to recall some properties and geometrical tools about the anisotropic elliptic operator defined above. For \(a, b \in {\mathbb {R}}^N\) we denote by \(a \otimes b\) the matrix whose entries are \((a \otimes b)_{ij}=a_ib_j\). We remark that for any \(v,w \in {\mathbb {R}}^N\) it holds that:

$$\begin{aligned} \langle a \otimes b \ v, w \rangle = \langle b, v \rangle \langle a, w \rangle . \end{aligned}$$

Now, we recall the definition of anisotropic norm.

\((h_H)\):

Let \(H \in C^2({\mathbb {R}}^N \setminus \{0\})\). In all the paper we assume that H is a anisotropic norm if it satisfies the following set of assumptions:

  1. (i)

    \(H(\xi )>0 \quad \forall \xi \in {\mathbb {R}}^N \setminus \{0\}\);

  2. (ii)

    \(H(s \xi ) = |s| H(\xi ) \quad \forall \xi \in {\mathbb {R}}^N \setminus \{0\}, \, \forall s \in {\mathbb {R}}\);

  3. (iii)

    H is uniformly elliptic, that means the set \({\mathcal {B}}_1^H:=\{\xi \in {\mathbb {R}}^N: H(\xi ) < 1\}\) is uniformly convex

    $$\begin{aligned} \exists \Lambda > 0: \quad \langle D^2H(\xi )v, v \rangle \ge \Lambda |v|^2 \quad \forall \xi \in \partial {\mathcal {B}}_1^H, \; \forall v \in \nabla H(\xi )^\bot . \end{aligned}$$
    (2.1)

A set is said uniformly convex if the principal curvatures of its boundary are all strictly positive. Moreover, assumption (iii) is equivalent to assume that \(D^2 (H^2)\) is definite positive.

The dual norm \(H^\circ : {\mathbb {R}}^N \rightarrow [0,+\infty )\) is defined as:

$$\begin{aligned} H^\circ (x) = \sup _{H(\xi ) \le 1} \langle \xi , x \rangle . \end{aligned}$$

It is possible to show that \(H^\circ \) is also a Finsler norm and it has the same regularity properties of H. Moreover, it holds \((H^\circ )^\circ = H\). For \(R > 0\) and \({{\bar{x}}} \in {\mathbb {R}}^N\) we define:

$$\begin{aligned} {\mathcal {B}}_R^H({{\bar{x}}}) = \{ x \in {\mathbb {R}}^N \, \ H(x-{{\bar{x}}}) < R\} \end{aligned}$$

and

$$\begin{aligned} {\mathcal {B}}_R^{H^\circ } ({{\bar{x}}}) = \{ x \in {\mathbb {R}}^N \, \ H^\circ (x-{{\bar{x}}}) < R\}. \end{aligned}$$

For simplicity of exposition, when \({{\bar{x}}} = 0\), we set: \({\mathcal {B}}_R^H={\mathcal {B}}_R^H(0)\), \({\mathcal {B}}_R^{H^\circ }={\mathcal {B}}_R^{H^\circ }(0)\). In literature \({\mathcal {B}}_R^H\) and \({\mathcal {B}}_R^{H^\circ }\) are also called “Wulff shape” and “Frank diagram” respectively. We remark that there holds the following identities:

$$\begin{aligned} H(\nabla H ^\circ (x)) = 1 = H^\circ (\nabla H (x)); \end{aligned}$$
(2.2)

and

$$\begin{aligned} H(x)\nabla H^\circ (\nabla H(x))=x=H^\circ (x)\nabla H(\nabla H^\circ (x)). \end{aligned}$$
(2.3)

We refer the reader to [3, 8] for further details. Observe also that H is a norm equivalent to the euclidean one, i.e. there exist \(\alpha _1, \alpha _2 > 0\) such that:

$$\begin{aligned} \alpha _1 |\xi | \le H(\xi ) \le \alpha _2 |\xi | \qquad \forall \xi \in {\mathbb {R}}^N. \end{aligned}$$
(2.4)

Moreover, recalling that H is 1-homogeneous, by the Euler’s Theorem it follows

$$\begin{aligned} \langle \nabla H(\xi ), \xi \rangle = H(\xi ) \qquad \forall \xi \in {\mathbb {R}}^N \setminus \{0\}. \end{aligned}$$
(2.5)

Since H is 1-homogeneous, we have that \(\nabla H\) is 0-homogeneous and it satisfies

$$\begin{aligned} \nabla H (\xi ) = \nabla H \left( |\xi |\frac{\xi }{|\xi |}\right) = \nabla H \left( \frac{\xi }{|\xi |}\right) \qquad \forall \xi \in {\mathbb {R}}^N\setminus \{0\}. \end{aligned}$$

Hence, by the previous equality, we infer that there exists \(M>0\) such that

$$\begin{aligned} | \nabla H(\xi )| \le M \qquad \forall \xi \in {\mathbb {R}}^N\setminus \{0\}. \end{aligned}$$
(2.6)

For the same reasons there exists a constant \({{\overline{M}}}>0\) such that:

$$\begin{aligned} |D^2H(\xi )|\le \frac{{{\overline{M}}}}{|\xi |}\qquad \forall \,\xi \in {\mathbb {R}}^N\setminus \{0\}, \end{aligned}$$
(2.7)

where \(|\cdot |\) denotes the usual Euclidean norm of a matrix, and

$$\begin{aligned} D^2H(\xi )\xi =0 \qquad \forall \,\xi \in {\mathbb {R}}^N\setminus \{0\}. \end{aligned}$$
(2.8)

We start with some elliptic estimates that can be proved in the same spirit of the Euclidean framework.

Proposition 2.1

For any \(p>1\) and \(\eta , \eta ' \in {\mathbb {R}}^N\) such that \(|\eta |+|\eta '|>0\), it holds

$$\begin{aligned} |H^{p-1}(\eta )\nabla H(\eta )-H^{p-1}(\eta ')\nabla H(\eta ')|\le {{\tilde{C}}}_p(|\eta |+|\eta '|)^{p-2}|\eta -\eta '|. \end{aligned}$$
(2.9)

Moreover, any \(p \ge 2\) it holds the following inequality

$$\begin{aligned} H^p(\eta ) \ge H^p(\eta ') + p H^{p-1}(\eta ') \langle \nabla H(\eta '), \eta -\eta ' \rangle + {{\hat{C}}}(p) H^p(\eta - \eta '), \end{aligned}$$
(2.10)

for any \(\eta , \eta ' \in {\mathbb {R}}^N\). Furthermore, if \(1<p<2\) we have that

$$\begin{aligned} H^p(\eta ) \ge H^p(\eta ') + p H^{p-1}(\eta ') \langle \nabla H (\eta '), \eta -\eta ' \rangle + C_p [H(\eta ) + H(\eta ')]^{p-2}H^2(\eta - \eta '),\nonumber \\ \end{aligned}$$
(2.11)

for any \(\eta , \eta ' \in {\mathbb {R}}^N\) such that \(|\eta |+|\eta '|>0\).

Proof

We start the proof showing (2.9). First of all we note that (2.9) is symmetric in \(\eta \), \(\eta '\). Hence, without loss of generality, we can assume that \(|\eta '| \ge |\eta | > 0\). We note that for \(j=1, \dots ,N\):

$$\begin{aligned}&H^{p-1}(\eta ) \frac{\partial H}{\partial \eta _j}(\eta )-H^{p-1}(\eta ') \frac{\partial H}{\partial \eta _j}(\eta ') \nonumber \\&\quad = \int _0^1 \sum _{i=1}^N \left[ (p-1)H^{p-2}(\eta '+t(\eta -\eta ')) \left( \frac{\partial H}{\partial \eta _i} \cdot \frac{\partial H}{\partial \eta _j}\right) (\eta '+t(\eta -\eta ')) \right. \nonumber \\&\qquad \qquad \left. + H^{p-1}(\eta '+t(\eta -\eta ')) \frac{\partial ^2 H}{\partial \eta _i \partial \eta _j}(\eta '+t(\eta -\eta ')) \right] \cdot (\eta _i-\eta _i') \, dt. \end{aligned}$$
(2.12)

By (2.12), using (2.4), (2.6) and (2.7) we have

$$\begin{aligned}&|H^{p-1}(\eta )\nabla H(\eta )-H^{p-1}(\eta ')\nabla H(\eta ')| \nonumber \\&\quad \le \int _0^1 |(p-1)H^{p-2}(\eta '+t(\eta -\eta ')) \nabla H(\eta '+t(\eta -\eta ')) \otimes \nabla H(\eta '+t(\eta -\eta ')) \nonumber \\&\qquad \ \ \ + H^{p-1}(\eta '+t(\eta -\eta ')))D^2H(\eta '+t(\eta -\eta '))| \cdot |\eta -\eta '| \, dt\nonumber \\&\quad \le ((p-1)\alpha _2^{p-2}M^2+\alpha _2^{p-1}{{\overline{M}}}) |\eta - \eta '| \int _0^1 (|\eta '+t(\eta -\eta ')|)^{p-2} \, dt, \end{aligned}$$
(2.13)

where \(|\cdot |\) denotes the standard matrix Euclidean norm.

Now, we observe that

$$\begin{aligned} |\eta '+t(\eta -\eta ')| \le |\eta |+ |\eta '|, \end{aligned}$$
(2.14)

and, since \(|\eta '| \ge |\eta |\), we have either

$$\begin{aligned} |\eta - \eta '| \le \frac{|\eta '|}{2} \ \ \Longrightarrow \ \ |\eta '+t(\eta -\eta ')| \ge |\eta '|-|\eta -\eta '| \ge \frac{|\eta '|}{2} \ge \frac{|\eta '|+|\eta |}{4}\nonumber \\ \end{aligned}$$
(2.15)

or, putting \(\displaystyle t_0:= \frac{|\eta '|}{|\eta -\eta '|} \in (0,2)\),

$$\begin{aligned} |\eta - \eta '| > \frac{|\eta '|}{2} \ \ \Longrightarrow \ \ |\eta '+t(\eta -\eta ')|&\ge \ \left| |\eta '|-t|\eta -\eta '|\right| = |t_0-t| \cdot | \eta -\eta '| \nonumber \\&\ge |t_0-t| \cdot \frac{|\eta '|}{2} = |t_0-t|\frac{|\eta '|}{2} \nonumber \\&\ge |t_0-t| \frac{|\eta | + |\eta '|}{4}. \end{aligned}$$
(2.16)

If \(p>2\), using (2.14) in (2.13) we have

$$\begin{aligned} |H^{p-1}(\eta )\nabla H(\eta )-H^{p-1}(\eta ')\nabla H(\eta ')| \le {{\tilde{C}}}_p (|\eta '|+|\eta |)^{p-2}|\eta - \eta '| \end{aligned}$$
(2.17)

where \({{\tilde{C}}}_p=(p-1)\alpha _2^{p-2}M^2+\alpha _2^{p-1}{{\overline{M}}}\). Hence (2.9) holds.

If \(p \le 2\) and (2.15) holds, by (2.13) we obtain

$$\begin{aligned}&|H^{p-1}(\eta )\nabla H(\eta )-H^{p-1}(\eta ')\nabla H(\eta ')| \nonumber \\&\quad \le ((p-1)\alpha _2^{p-2}M^2+\alpha _2^{p-1}{{\overline{M}}}) |\eta - \eta '| \int _0^1 ( |\eta '+t(\eta -\eta ')|)^{p-2} \, dt.\nonumber \\&\quad \le ((p-1)\alpha _2^{p-2}M^2+\alpha _2^{p-1}{{\overline{M}}}) |\eta - \eta '| \int _0^1 \left( \frac{|\eta '|+|\eta |}{4}\right) ^{p-2} \, dt\nonumber \\&\quad \le {{\tilde{C}}}_p (|\eta '|+|\eta |)^{p-2}|\eta -\eta '|, \end{aligned}$$
(2.18)

where \({{\tilde{C}}}_p=((p-1)\alpha _2^{p-2}M^2+\alpha _2^{p-1}\overline{M})/4^{p-2}\). Hence (2.9) holds

If \(p \le 2\) and (2.16) holds, by (2.13) we obtain

$$\begin{aligned}&|H^{p-1}(\eta )\nabla H(\eta )-H^{p-1}(\eta ')\nabla H(\eta ')| \nonumber \\&\quad \le ((p-1)\alpha _2^{p-2}M^2+\alpha _2^{p-1}{{\overline{M}}})|\eta - \eta '| \int _0^1 ( |\eta '+t(\eta -\eta ')|)^{p-2} \, dt.\nonumber \\&\quad \le ((p-1)\alpha _2^{p-2}M^2+\alpha _2^{p-1}{{\overline{M}}}) |\eta - \eta '| \int _0^1 |t_0-t|^{p-2}\left( \frac{|\eta '|+|\eta |}{4}\right) ^{p-2} \, dt\nonumber \\&\quad = \frac{(p-1)\alpha _2^{p-2}M^2+\alpha _2^{p-1}{{\overline{M}}}}{4^{p-2}} ( |\eta '|+|\eta |)^{p-2}|\eta -\eta '| \int _0^1 |t_0-t|^{p-2} \, dt\nonumber \\&\quad \le \frac{(p-1)\alpha _2^{p-2}M^2+\alpha _2^{p-1}{{\overline{M}}}}{4^{p-2}} (|\eta '|+|\eta |)^{p-2}|\eta -\eta '| \ 2 \int _0^1 z^{p-2} \, dz\nonumber \\&\quad = {{\tilde{C}}}_p (|\eta '|+|\eta |)^{p-2}|\eta -\eta '|, \end{aligned}$$
(2.19)

where \({{\tilde{C}}}_p=(2(p-1)\alpha _2^{p-2}M^2+2\alpha _2^{p-1}{{\overline{M}}})/4^{p-2}\).

Collecting the estimates above, we deduce that inequality (2.9) holds for every \(p>1\) and for \(\tilde{C}_p=((p-1)\alpha _2^{p-2}M^2+\alpha _2^{p-1}{{\overline{M}}}) \cdot \max \{1, 4^{2-p}, 2 \cdot 4^{2-p}\}\).

Now we will show (2.10) and (2.11). For \(\eta , \eta '\in {\mathbb {R}}^N\), we define

$$\begin{aligned} f(t)=H^p(\eta '+t(\eta -\eta ')) \end{aligned}$$

and Taylor’s formula yields

$$\begin{aligned} H^p(\eta ) = H^p(\eta ') + p H^{p-1}(\eta ') \langle \nabla H(\eta '), \eta -\eta ' \rangle + \int _0^1 (1-t)f''(t) \,dt, \end{aligned}$$
(2.20)

provided \(|\eta '+t(\eta -\eta ')|\ne 0\), for \(0\le t\le 1\). But the case when \(\eta '+t(\eta -\eta ')=0\) can be easily verified. By [10, Theorem 1.5] we obtain that

$$\begin{aligned} f''(t)&=\langle D^2(H^p(\eta '+t(\eta -\eta ')))(\eta -\eta '),\eta -\eta '\rangle \nonumber \\&\ge C(p)|\eta '+t(\eta -\eta ')|^{p-2}|\eta -\eta '|^2, \end{aligned}$$
(2.21)

where C(p) is a constant depending on p. We remark that

$$\begin{aligned} \int _0^1(1-t)f''(t) \,dt \ge \frac{3}{4}\int _0^{\frac{1}{4}}f''(t) \,dt. \end{aligned}$$
(2.22)

If \(1<p<2\), by (2.14) we have

$$\begin{aligned} (H(\eta )+H(\eta '))^{p-2}\le \alpha _2^{p-2}|\eta '+t(\eta -\eta ')|^{p-2} \end{aligned}$$

and using (2.22) we arrive at (2.11).

If \(p\ge 2\), using a similar argument as in the proof of the inequality (2.9), we obtain

$$\begin{aligned} \int _0^{\frac{1}{4}}f''(t)\ge \overline{C}(p)(|\eta |+|\eta '|)|^{p-2}|\eta -\eta '|^2, \end{aligned}$$
(2.23)

with \({{\overline{C}}}(p)\) constant depending on p. Since \(|\eta -\eta '|\le |\eta |+|\eta '|\) and using (2.4), we get (2.10). \(\square \)

We state now the Hardy inequality for the anisotropic operator \(\Delta _p^Hu\), defined in (1.1). We refer to [28, Proposition 7.5].

Theorem 2.2

(Hardy inequality) For any H satisfying the assumption \((h_H)\) and any \(u\in {\mathcal {D}}^{1,p}({\mathbb {R}}^N)\) and \(1<p<N\),

$$\begin{aligned} C_H\int _{{\mathbb {R}}^N}\frac{|u|^p}{{H^\circ (x)^p}}\, dx\le \int _{{\mathbb {R}}^N}H^p(\nabla u)\, dx, \end{aligned}$$
(2.24)

where \(C_H=(({N-p})/{p})^p\) is optimal.

Now we prove a technical lemma that will be very important in the proof of the asymptotic estimates.

Lemma 2.3

Let \(p>1\) and \(a,b \ge 0\). Then, for all \(\delta >0\) there exist \({\mathcal {C}}_\delta >0\) such that

$$\begin{aligned} a^p \ge \frac{1}{1+ 2^{p+1}\delta } (a+b)^p - {\mathcal {C}}_\delta b^p. \end{aligned}$$
(2.25)

Proof

Let us consider \(p>1\) as follows:

$$\begin{aligned} p=\lfloor p \rfloor + \{p\}, \end{aligned}$$

where \(\lfloor \cdot \rfloor \) is the floor function and \(\{\cdot \}\) is the mantissa function. Without loss of generality we assume that \(\{p\} \ne 0\) and, moreover, we set \(m:= \lfloor p \rfloor \). Hence, we have

$$\begin{aligned} (a+b)^p=(a+b)^{\{p\}}(a+b)^m= (a+b)^{\{p\}} \sum _{k=0}^m {m \atopwithdelims ()k} a^{m-k} b^k = (\star ) \end{aligned}$$
(2.26)

Noticing that \(0 \le \{p\} < 1\) it follows that

$$\begin{aligned} (a+b)^{\{p\}} \le a^{\{p\}} + b^{\{p\}}. \end{aligned}$$

Using this inequality in (2.26), we deduce

$$\begin{aligned} \begin{aligned} (\star )&\le a^{\{p\}} \sum _{k=0}^m {m \atopwithdelims ()k} a^{m-k} b^k + b^{\{p\}} \sum _{k=0}^m {m \atopwithdelims ()k} a^{m-k} b^k\\&= a^p+b^p+\sum _{k=1}^m {m \atopwithdelims ()k} a^{p-k} b^k + \sum _{k=0}^{m-1} {m \atopwithdelims ()k} a^{m-k} b^{k+\{p\}}=(*), \end{aligned} \end{aligned}$$
(2.27)

where we used the fact that \(p=m+\{p\}\). Now, we can apply the weighted Young’s inequality to each member of the first sum with conjugate exponents \(\left( p/(p-k),p/k\right) \) and to each member of the second sum with conjugate exponents \(\left( p/(m-k),p/(k+\{p\})\right) \) as follows

$$\begin{aligned} \begin{aligned} a^{p-k}b^k&\le \frac{p-k}{p} \delta a^p + \frac{k}{p} {\mathcal {C}}_\delta b^p \le \delta a^p + {\mathcal {C}}_\delta b^p,\\ a^{m-k} b^{k+\{p\}}&\le \delta a^p + {\mathcal {C}}_\delta b^p. \end{aligned} \end{aligned}$$
(2.28)

Hence, using this estimate we deduce

$$\begin{aligned} \begin{aligned} (*)&\le a^p+b^p+ \left( \delta a^p + {\mathcal {C}}_\delta b^p\right) \sum _{k=1}^m {m \atopwithdelims ()k} + \left( \delta a^p + {\mathcal {C}}_\delta b^p\right) \sum _{k=0}^{m-1} {m \atopwithdelims ()k} \\&\le \left( 1+2^{p+1} \delta \right) a^p + {\mathcal {C}}_\delta b^p, \end{aligned} \end{aligned}$$
(2.29)

where we renamed \({\mathcal {C}}_\delta := \left( 1+2^{p+1} {\mathcal {C}}_\delta \right) \). Collecting (2.26), (2.27) and (2.29), we deduce that

$$\begin{aligned} a^p \ge \frac{1}{1+ 2^{p+1}\delta } (a+b)^p - {\mathcal {C}}_\delta b^p, \end{aligned}$$
(2.30)

with \({\mathcal {C}}_\delta :={\mathcal {C}}_\delta /(1+2^{p+1}\delta )\), and hence the thesis (2.25). \(\square \)

Finally, we recall a lemma (see Lemma 4.19 in [16]) that will be very useful in the proofs of our results.

Lemma 2.4

[16] Let \({\mathcal {L}}\) and g be two nondecreasing functions on the interval \((0, {{\bar{R}}}]\), for some \({{\bar{R}}}>0\). Suppose that it holds

$$\begin{aligned} {\mathcal {L}}(\tau R) \le \sigma {\mathcal {L}}(R) + g(R) \quad \text {for all } R \le {{\bar{R}}}, \end{aligned}$$

for some \(0< \sigma , \tau < 1\). Then, for any \(\mu \in (0,1)\) and \(R \le {{\bar{R}}}\) we have

$$\begin{aligned} {\mathcal {L}}(R) \le \frac{1}{\sigma } \left( \frac{R}{\bar{R}}\right) ^\alpha {\mathcal {L}}({{\bar{R}}}) + \frac{1}{1-\sigma }g(\bar{R}^\mu R^{1-\mu }) \end{aligned}$$

where \(\alpha = \alpha (\sigma , \tau , \mu ) =(1-\mu ) \log \sigma / \log \tau \).

3 Preliminary asymptotic estimates and comparison principles

The aim of this section is to prove some preliminary estimates that will be crucial in the proofs of the main results.

Lemma 3.1

There exists a positive constant \(\tau \) depending only on Np and \(\gamma \) such that for any \({\bar{R}}> 0\) and for any solution u to problem (\({\mathcal {P}}_H\)) satisfying

$$\begin{aligned} \Vert u\Vert _{L^{p^*}({\mathcal {B}}^{H^\circ }_{{{\bar{R}}}})} + \Vert u\Vert _{L^{p^*}({\mathbb {R}}^N \setminus {\mathcal {B}}^{H^\circ }_{1 / {{\bar{R}}}})} \le \tau , \end{aligned}$$
(3.1)

there exists a positive constant \({\mathcal {C}}\) depending only on \(N, p, \gamma \) and \({{\overline{R}}}\) such that

$$\begin{aligned} \Vert u\Vert _{L^{p^*}({\mathcal {B}}^{H^\circ }_R)} \le {\mathcal {C}} R^{\sigma _1} \quad \text {for } R \le {{\bar{R}}}, \end{aligned}$$
(3.2)

and that

$$\begin{aligned} \Vert u\Vert _{L^{p^*}(({\mathcal {B}}^{H^\circ }_R)^c)} \le \frac{{\mathcal {C}}}{R^{\sigma _2}} \quad \text {for } R \ge \frac{1}{{{\bar{R}}}}, \end{aligned}$$
(3.3)

where \(\sigma _1, \sigma _2\) are two positive constants depending on Np and \(\gamma \).

Proof

We start proving (3.2). To this aim let us consider \( R>0\) and a cut-off function \(\eta \in C^\infty _c({\mathbb {R}}^N)\) such that

$$\begin{aligned} \left\{ \begin{array}{llll} 0 \le \eta \le 1 \quad &{}\text {in } {\mathbb {R}}^N\\ \eta \equiv 0 \quad &{}\text {in } ({\mathcal {B}}^{H^\circ }_R)^c\\ \eta \equiv 1 \quad &{}\text {in } {\mathcal {B}}^{H^\circ }_{R / 2}\\ |\nabla \eta | \le \frac{4}{R} &{}\text {in } {\mathcal {B}}^{H^\circ }_R \setminus {\mathcal {B}}^{H^\circ }_{R / 2}. \end{array}\right. \end{aligned}$$
(3.4)

By density argument it is possible to put \(\varphi =\eta ^p u\) as test function in (1.2), so that we obtain

$$\begin{aligned} \begin{aligned}&\int _{{\mathbb {R}}^N} p \eta ^{p-1} u H^{p-1}(\nabla u) \langle \nabla H(\nabla u), \nabla \eta \rangle \, dx + \int _{{\mathbb {R}}^N}\eta ^p H^{p-1}(\nabla u) \langle \nabla H(\nabla u), \nabla u \rangle \, dx\\&\quad = \int _{{\mathbb {R}}^N} \frac{\gamma }{H^\circ (x)^p} u^{p-1} \eta ^p u \, dx + \int _{{\mathbb {R}}^N} u^{p^*-1} \eta ^p u \, dx. \end{aligned}. \end{aligned}$$
(3.5)

First of all, using Euler’s Theorem (2.5), the 0-homogeneity of \(\nabla H\) (2.6) and Schwarz’s inequality, equation (3.5) becomes

$$\begin{aligned} \begin{aligned}&\int _{{\mathbb {R}}^N} H^p (\nabla u) \eta ^p \, dx = \int _{{\mathbb {R}}^N}H^{p-1}(\nabla u) \langle \nabla H(\nabla u), \nabla u \rangle \eta ^p \, dx\\&\quad \le p \ M \int _{{\mathbb {R}}^N} H^{p-1}(\nabla u) |\nabla \eta | \eta ^{p-1} u \, dx + \int _{{\mathbb {R}}^N} \frac{\gamma }{H^\circ (x)^p} u^p \eta ^p \, dx +\int _{{\mathbb {R}}^N} u^{p^*} \eta ^p \, dx. \end{aligned}\nonumber \\ \end{aligned}$$
(3.6)

Recalling that H is 1-homogeneous function, using the weighted Young’s inequality \(ab \le \varepsilon a^{\frac{p}{p-1}} + {\mathcal {C}}_\varepsilon b^p\) on the first term of the right hand side of (3.6), for any \(0<\varepsilon <1\) we have

$$\begin{aligned} \begin{aligned} \int _{{\mathbb {R}}^N} H^p (\eta \nabla u) \, dx&\le \varepsilon \int _{{\mathbb {R}}^N} H^p(\eta \nabla u) \, dx + C(p,M,\varepsilon ) \int _{{\mathbb {R}}^N} |\nabla \eta |^p u^p \, dx \\&\ \ + \int _{{\mathbb {R}}^N} \frac{\gamma }{H^\circ (x)^p} u^{p} \eta ^p \, dx + \int _{{\mathbb {R}}^N} u^{p^*} \eta ^p \, dx,\\ \end{aligned} \end{aligned}$$
(3.7)

where \( C(p,M,\varepsilon ):=(pM)^p {\mathcal {C}}_\varepsilon \). Now, noticing that \(\nabla (\eta u) = u \nabla \eta + \eta \nabla u\), by the triangular inequality, we deduce that for every \(p>1\) it holds

$$\begin{aligned} H^p(\nabla (\eta u)) =H^p(u \nabla \eta + \eta \nabla u)\le \left[ H(u \nabla \eta ) +H(\eta \nabla u) \right] ^p. \end{aligned}$$
(3.8)

Thanks to (3.8) and applying Lemma 2.3 with \(a=H(\eta \nabla u)\) and \(b=H(u\nabla \eta )\), we deduce that

$$\begin{aligned} \int _{{\mathbb {R}}^N} H^p(\eta \nabla u) \,dx \ge \frac{1}{1+2^{p+1}\delta } \int _{{\mathbb {R}}^N} H^p(\nabla (\eta u)) \,dx - {\mathcal {C}}_\delta \int _{{\mathbb {R}}^N} H^p(u \nabla \eta ) \, dx. \end{aligned}$$
(3.9)

Using (3.9) in (3.7) we obtain

$$\begin{aligned} \frac{1-\varepsilon }{1+2^{p+1}\delta }\int _{{\mathbb {R}}^N} H^p (\nabla (\eta u)) \, dx&\le {\mathcal {C}}_\delta \int _{{\mathbb {R}}^N} H^p(u \nabla \eta ) \, dx + C(p,M,\varepsilon ) \int _{{\mathbb {R}}^N} |\nabla \eta |^p u^p \, dx \nonumber \\&\quad + \gamma \int _{{\mathbb {R}}^N} \frac{(u \eta )^p}{H^\circ (x)^p} \, dx + \int _{{\mathbb {R}}^N} u^{p^*} \eta ^p \, dx.\nonumber \\ \end{aligned}$$
(3.10)

Now, applying the anisotropic Hardy inequality (see Theorem 2.2 or [28]) and (2.4) we have

$$\begin{aligned} \frac{1-\varepsilon }{1+2^{p+1}\delta }\int _{{\mathbb {R}}^N} H^p (\nabla (\eta u)) \, dx&\le \left( \alpha _2^p {\mathcal {C}}_\delta + C(p,M,\varepsilon ) \right) \int _{{\mathbb {R}}^N} |\nabla \eta |^p u^p \, dx \nonumber \\&\quad + \frac{\gamma }{C_H}\int _{{\mathbb {R}}^N} H^p (\nabla (\eta u)) \, dx + \int _{{\mathbb {R}}^N} u^{p^*} \eta ^p \, dx. \end{aligned}$$
(3.11)

Let us fix \(\varepsilon , \delta >0\) sufficiently small such that \({\mathcal {C}}_1:=({1-\varepsilon })/({1+2^{p+1}\delta }) - {\gamma }/{C_H} > 0\), so that

$$\begin{aligned} {\mathcal {C}}_1 \int _{{\mathbb {R}}^N} H^p (\nabla (\eta u)) \, dx \le {\mathcal {C}}_2 \int _{{\mathbb {R}}^N} |\nabla \eta |^p u^p \, dx + \int _{{\mathbb {R}}^N} u^{p^*} \eta ^p \, dx, \end{aligned}$$
(3.12)

where \({\mathcal {C}}_2:= C(p,M,\varepsilon ) + \alpha _2^p {\mathcal {C}}_\delta \). By (2.4) we have

$$\begin{aligned} \alpha _1^p{\mathcal {C}}_1 \int _{{\mathbb {R}}^N} |\nabla (\eta u)|^p \, dx \le {\mathcal {C}}_2 \int _{{\mathbb {R}}^N} |\nabla \eta |^p u^p \, dx + \int _{{\mathbb {R}}^N} u^{p^*} \eta ^p \, dx. \end{aligned}$$
(3.13)

Now, using the Sobolev inequality in the left hand side of (3.13), the Hölder inequality and (3.4) in the right hand side, we obtain

$$\begin{aligned} \begin{aligned} \alpha _1^p{\mathcal {C}}_1 {\mathcal {C}}_S^p \left( \int _{{\mathbb {R}}^N} |\eta u|^{p^*} \,dx \right) ^\frac{p}{p^*}&\le \alpha _1^p{\mathcal {C}}_1 \int _{{\mathbb {R}}^N} |\nabla (\eta u)|^p \, dx \\&\le {\mathcal {C}}_2 \int _{{\mathbb {R}}^N} |\nabla \eta |^p u^p \, dx + \int _{{\mathbb {R}}^N} u^{p^*-p} (\eta u)^p \, dx\\&\le {\mathcal {C}}_2 \left( \int _{{\mathcal {B}}^{H^\circ }_R \setminus {\mathcal {B}}^{H^\circ }_{R / 2}} |\nabla \eta |^N \, dx\right) ^{\frac{p}{N}} \left( \int _{{\mathcal {B}}^{H^\circ }_R \setminus {\mathcal {B}}^{H^\circ }_{R / 2}} u^{p^*} \, dx\right) ^{\frac{p}{p^*}} \\&\quad + \left( \int _{{\mathcal {B}}^{H^\circ }_R} u^{p^*} \, dx\right) ^{\frac{p}{N}} \left( \int _{{\mathbb {R}}^N} (\eta u)^{p^*} \, dx\right) ^{\frac{p}{p^*}}\\&\le {\mathcal {C}}_2 C(p,N) \left( \int _{{\mathcal {B}}^{H^\circ }_R \setminus {\mathcal {B}}^{H^\circ }_{R / 2}} u^{p^*} \, dx\right) ^{\frac{p}{p^*}} \\&\quad + \Vert u\Vert _{p^*}^{p^*-p} \left( \int _{{\mathbb {R}}^N} (\eta u)^{p^*} \, dx\right) ^{\frac{p}{p^*}}, \end{aligned} \end{aligned}$$
(3.14)

hence we deduce

$$\begin{aligned} \left( \int _{{\mathbb {R}}^N} |\eta u|^{p^*} \,dx \right) ^\frac{p}{p^*}&\le \frac{{\mathcal {C}}_2 C(p,N)}{\alpha _1^p{\mathcal {C}}_1 {\mathcal {C}}_S^p} \left( \int _{{\mathcal {B}}^{H^\circ }_R \setminus {\mathcal {B}}^{H^\circ }_{R / 2}} u^{p^*} \, dx\right) ^{\frac{p}{p^*}} \nonumber \\&\quad + \frac{\Vert u\Vert _{L^{p^*}({\mathcal {B}}^{H^\circ }_R)}^{p^*-p}}{\alpha _1^p{\mathcal {C}}_1 {\mathcal {C}}_S^p} \left( \int _{{\mathbb {R}}^N} (\eta u)^{p^*} \, dx\right) ^{\frac{p}{p^*}}, \end{aligned}$$
(3.15)

where C(pN) is a positive constant depending on p and N.

Setting

$$\begin{aligned} \tau = \left( \frac{\alpha _1^p{\mathcal {C}}_1 {\mathcal {C}}_S^p}{2}\right) ^{\frac{1}{p^*-p}}, \end{aligned}$$

and choosing \({\overline{R}}>0\) sufficiently small such that (3.1) holds, then \(\Vert u\Vert _{L^{p^*}({\mathcal {B}}^{H^\circ }_R)}^{p^*-p}/(\alpha _1^p{\mathcal {C}}_1 {\mathcal {C}}_S^p) \le 1/2\) for all \(0<R \le {{\overline{R}}}\). Hence we obtain that

$$\begin{aligned} \int _{{\mathcal {B}}^{H^\circ }_{R/2}} u^{p^*} \,dx \le \int _{{\mathbb {R}}^N} |\eta u|^{p^*} \,dx \le {{\bar{C}}} \int _{{\mathcal {B}}^{H^\circ }_R \setminus {\mathcal {B}}^{H^\circ }_{R / 2}} u^{p^*} \, dx \qquad \forall \ 0 < R \le {{\bar{R}}}, \end{aligned}$$

where \({{\bar{C}}}:= \left( (2{\mathcal {C}}_2 C(p,N))/({\alpha _1^p{\mathcal {C}}_1 {\mathcal {C}}_S^p})\right) ^{\frac{p^*}{p}}\) and it depends only on N, p and \(\gamma \). Denoting with \(\displaystyle {\mathcal {L}}(R):=\int _{{\mathcal {B}}^{H^\circ }_R}u^{p^*} \, dx\) for \(0 < R \le {{\bar{R}}}\), we get that

$$\begin{aligned} {\mathcal {L}}(R/2) \le \vartheta {\mathcal {L}}(R) \qquad \forall \ 0 < R \le {{\bar{R}}}, \end{aligned}$$

where \(\vartheta = {{\bar{C}}}/({{\bar{C}}} + 1)\in (0,1)\), depends only on N, p and \(\gamma \). Now, by Lemma 2.4 it follows that

$$\begin{aligned} {\mathcal {L}}(R) \le \frac{1}{\vartheta } {\mathcal {L}}({{\bar{R}}}) \left( \frac{R}{{{\bar{R}}}}\right) ^{\sigma _1 '} \qquad \forall \ 0 < R \le {{\bar{R}}}, \end{aligned}$$

where \(\sigma _1'=\frac{1}{2}\log (1 / \vartheta )/ \log 2\) depends only on \(\vartheta \), Now (3.2) follows by setting \(\sigma _1 = \sigma _1' / p^*\) and \({\mathcal {C}}=(\vartheta ^{-1}{{\overline{R}}}^{-\sigma _1'}{\mathcal {L}}(\bar{R}))^{1/p^*}\). In a similar way, we can deduce (3.3). \(\square \)

Now, we denote by \({\mathcal {A}}_R={\mathcal {B}}^{H^\circ }_{8R} {\setminus } {\mathcal {B}}^{H^\circ }_{R/8}\) and \({\mathcal {D}}_R={\mathcal {B}}^{H^\circ }_{4R} {\setminus } {\mathcal {B}}^{H^\circ }_{R/4}\) for \(R>0\).

Lemma 3.2

Let \(t \in (p^*, N / \mu _1)\). There exists a positive constant \(\sigma = \sigma (N,p,\gamma ,t)\) such that for any solution u to problem (\({\mathcal {P}}_H\)) and for any \({{\bar{R}}}> 0\) satisfying the following inequality

$$\begin{aligned} \Vert u\Vert _{L^{p^*}(B_{{{\bar{R}}}})} + \Vert u\Vert _{L^{p^*}({\mathbb {R}}^N \setminus B_{1 / {{\bar{R}}}})} \le \sigma , \end{aligned}$$
(3.16)

then

(3.17)

where and \({\mathcal {C}}\) is a positive constant depending only on \(N, p, \gamma , {{\overline{R}}}\) and t.

Proof

It is easy to see that, setting \({{\hat{u}}}(x)=u(Rx)\), for \(R>0\),

$$\begin{aligned} -\Delta _p^H {{\hat{u}}} - \gamma \frac{{{\hat{u}}}^{p-1}}{H^\circ (x)^p}=R^p{{\hat{u}}}^{p^*-1} \qquad \text {in } {\mathcal {A}}_1. \end{aligned}$$

Let \(m \ge 1\) and set

$$\begin{aligned} {\mathcal {A}}:= {\mathcal {A}}_1 \cap \{{{\hat{u}}} < m\} \quad \text {and} \quad {\mathcal {B}}:= {\mathcal {A}}_1 \cap \{{{\hat{u}}} \ge m\}. \end{aligned}$$
(3.18)

Hence we can consider the weak formulation of the last equation as follows

$$\begin{aligned} \begin{aligned} I_1+I_2&:= \int _{{\mathcal {A}}} H^{p-1}(\nabla {{\hat{u}}}) \langle \nabla H(\nabla {{\hat{u}}}), \nabla \varphi \rangle \, dx + \int _{{\mathcal {B}}} H^{p-1}(\nabla {{\hat{u}}}) \langle \nabla H(\nabla {{\hat{u}}}), \nabla \varphi \rangle \, dx \\&=\int _{{\mathcal {A}}_1} \gamma \frac{{{\hat{u}}}^{p-1}}{H^\circ (x)^p} \varphi \, dx + R^p\int _{{\mathcal {A}}_1} {{\hat{u}}}^{p^*-1} \varphi \, dx \qquad \forall \varphi \in C^\infty _c({\mathcal {A}}_1). \end{aligned} \end{aligned}$$
(3.19)

Let define \({{\hat{u}}}_m:=\min ({{\hat{u}}}, m)\) for \(m \ge 1\). By density argument, for any \(\eta \in C^\infty _c({\mathcal {A}}_1)\) it is possible to choose \(\varphi =\eta ^p {{\hat{u}}}_m^{p(s-1)}{{\hat{u}}}\), with \(s\ge 1\), as test function in (3.19), so that, using (3.18) and (2.5), we can compute \(I_1\) and \(I_2\)

$$\begin{aligned} \begin{aligned} I_1&= \int _{{\mathcal {A}}} H^{p-1}(\nabla {{\hat{u}}}) \langle \nabla H(\nabla {{\hat{u}}}), p \eta ^{p-1} {{\hat{u}}}^{p(s-1)+1} \nabla \eta \rangle \, dx \\&\quad + [p(s-1)+1] \int _{{\mathcal {A}}} H^{p-1}(\nabla {{\hat{u}}}) \langle \nabla H(\nabla {{\hat{u}}}), \eta ^p {{\hat{u}}}^{p(s-1)} \nabla {{\hat{u}}} \rangle \, dx\\&= p\int _{{\mathcal {A}}} H^{p-1}(\nabla {{\hat{u}}}) \langle \nabla H(\nabla {{\hat{u}}}), \nabla \eta \rangle \eta ^{p-1} {{\hat{u}}}^{p(s-1)+1} \, dx \\&\quad + [p(s-1)+1] \int _{{\mathcal {A}}} H^p(\nabla {{\hat{u}}}) \eta ^p {{\hat{u}}}^{p(s-1)} \, dx.\\ \end{aligned} \end{aligned}$$
(3.20)

In the same way, we obtain

$$\begin{aligned} \begin{aligned} I_2&= \int _{{\mathcal {B}}} H^{p-1}(\nabla {{\hat{u}}}) \langle \nabla H(\nabla {{\hat{u}}}), p \eta ^{p-1} m^{p(s-1)} {{\hat{u}}} \nabla \eta \rangle \, dx \\&\quad + \int _{{\mathcal {B}}} H^{p-1}(\nabla {{\hat{u}}}) \langle \nabla H(\nabla {{\hat{u}}}), \eta ^p m^{p(s-1)} \nabla {{\hat{u}}} \rangle \, dx\\&= p\int _{{\mathcal {B}}} H^{p-1}(\nabla {{\hat{u}}}) \langle \nabla H(\nabla {{\hat{u}}}), \nabla \eta \rangle \eta ^{p-1} m^{p(s-1)} {{\hat{u}}} \, dx + \int _{{\mathcal {B}}} H^p(\nabla {{\hat{u}}}) \eta ^p m^{p(s-1)} \, dx.\\ \end{aligned}\nonumber \\ \end{aligned}$$
(3.21)

Collecting both (3.20), (3.21), using the Scwharz’s inequality and recalling (2.6) we obtain

$$\begin{aligned} \begin{aligned}&{[}p(s-1)+1] \int _{{\mathcal {A}}} H^p(\nabla {{\hat{u}}}) \eta ^p {{\hat{u}}}^{p(s-1)} \, dx +\int _{{\mathcal {B}}} H^p(\nabla {{\hat{u}}}) \eta ^p m^{p(s-1)} \, dx\\&\quad \le p M \left[ \int _{{\mathcal {A}}} H^{p-1}(\nabla {{\hat{u}}}) \eta ^{p-1} {{\hat{u}}}^{p(s-1)+1} |\nabla \eta | \, dx + \int _{{\mathcal {B}}} H^{p-1}(\nabla {{\hat{u}}}) \eta ^{p-1} m^{p(s-1)} |\nabla \eta | {{\hat{u}}} \, dx\right] \\&\quad +\int _{{\mathcal {A}}_1} \gamma \frac{{{\hat{u}}}^{p-1}}{H^\circ (x)^p} \varphi \, dx + R^p\int _{{\mathcal {A}}_1} {{\hat{u}}}^{p^*-1} \varphi \, dx\\&\quad = p M \left[ \int _{{\mathcal {A}}} H^{p-1}(\nabla {{\hat{u}}}) \eta ^{p-1} {{\hat{u}}}^{(p-1)(s-1)} {{\hat{u}}}^s |\nabla \eta | \, dx \right. \\&\quad \left. + \int _{{\mathcal {B}}} H^{p-1}(\nabla {{\hat{u}}}) \eta ^{p-1} m^{(p-1)(s-1)}m^{s-1} |\nabla \eta | {{\hat{u}}} \, dx \right] \\&\quad + \int _{{\mathcal {A}}_1} \gamma \frac{{{\hat{u}}}^{p-1}}{H^\circ (x)^p} \varphi \, dx + R^p\int _{{\mathcal {A}}_1} {{\hat{u}}}^{p^*-1} \varphi \, dx. \end{aligned} \end{aligned}$$
(3.22)

Now we can apply the weighted Young’s inequality to the first two terms in the right hand side of (3.22) with conjugate exponent \(\left( p/(p-1), p\right) \) in order to obtain

$$\begin{aligned} \begin{aligned}&{[}p(s-1)+1] \int _{{\mathcal {A}}} H^p(\nabla {{\hat{u}}}) \eta ^p {{\hat{u}}}^{p(s-1)} \, dx +\int _{{\mathcal {B}}} H^p(\nabla {{\hat{u}}}) \eta ^p m^{p(s-1)} \, dx\\&\quad \le \varepsilon _1 [p(s-1)+1] \int _{{\mathcal {A}}} H^p(\nabla {{\hat{u}}}) \eta ^p {{\hat{u}}}^{p(s-1)} \, dx + {\mathcal {C}}_{\varepsilon _1}(p,s,M) \int _{{\mathcal {A}}} {{\hat{u}}}^{ps} |\nabla \eta |^p \, dx\\&\quad + \varepsilon _2 \int _{{\mathcal {B}}} H^p(\nabla {{\hat{u}}}) \eta ^p m^{p(s-1)} \,dx + {\mathcal {C}}_{\varepsilon _2}(p,M) \int _{{\mathcal {B}}}|\nabla \eta |^p m^{p(s-1)} {{\hat{u}}}^p \, dx \\&\quad + \int _{{\mathcal {A}}_1} \gamma \frac{{{\hat{u}}}^{p-1}}{H^\circ (x)^p} \varphi \, dx \\ {}&\quad + R^p\int _{{\mathcal {A}}_1} {{\hat{u}}}^{p^*-1} \varphi \, dx, \end{aligned} \end{aligned}$$
(3.23)

where \({\mathcal {C}}_{\varepsilon _1}(p,s,M)\) and \({\mathcal {C}}_{\varepsilon _2}(p,M)\) are two positive constants. Hence we obtain

$$\begin{aligned} \begin{aligned}&(1-\varepsilon _1)[p(s-1)+1] \int _{{\mathcal {A}}} H^p(\eta {{\hat{u}}}_m^{s-1}\nabla {{\hat{u}}}) \, dx +(1-\varepsilon _2)\int _{{\mathcal {B}}} H^p(\eta {{\hat{u}}}_m^{s-1}\nabla {{\hat{u}}}) \, dx\\&\quad =(1-\varepsilon _1)[p(s-1)+1] \int _{{\mathcal {A}}} H^p(\nabla {{\hat{u}}}) \eta ^p {{\hat{u}}}^{p(s-1)} \, dx +(1-\varepsilon _2)\int _{{\mathcal {B}}} H^p(\nabla {{\hat{u}}}) \eta ^p m^{p(s-1)} \, dx\\&\quad \le {\mathcal {C}}_{\varepsilon _1}(p,s,M) \int _{{\mathcal {A}}} |\nabla \eta |^p {{\hat{u}}}^{p(s-1)} {{\hat{u}}}^p \, dx + {\mathcal {C}}_{\varepsilon _2}(p,M) \int _{{\mathcal {B}}}|\nabla \eta |^p m^{p(s-1)} {{\hat{u}}}^p \, dx \\&\quad + \int _{{\mathcal {A}}_1} \frac{\gamma }{H^\circ (x)^p} {{\hat{u}}}^{p-1} \eta ^p {{\hat{u}}}_m^{p(s-1)}{{\hat{u}}} \, dx + R^p\int _{{\mathcal {A}}_1} {{\hat{u}}}^{p^*-1} \eta ^p {{\hat{u}}}_m^{p(s-1)}{{\hat{u}}} \, dx. \end{aligned} \end{aligned}$$
(3.24)

Thanks to (3.8) and applying Lemma 2.3, with \(a=H(\eta {{\hat{u}}}^{s-1}\nabla {{\hat{u}}})\) and \(b=H({{\hat{u}}}^s\nabla \eta )\), we deduce that there hold the following inequalities in the sets \({\mathcal {A}}\) and \({\mathcal {B}}\) respectively:

$$\begin{aligned} \begin{aligned}&\int _{{\mathcal {A}}} H^p(\eta {{\hat{u}}}_m^{s-1}\nabla {{\hat{u}}}) \,dx =\int _{{\mathcal {A}}} H^p(\eta {{\hat{u}}}^{s-1}\nabla {{\hat{u}}}) \,dx\\&\ge \frac{1}{1+2^{p+1}\delta _1} \cdot \frac{1}{s^p} \int _{{\mathcal {A}}} H^p(\nabla (\eta {{\hat{u}}}^s)) \,dx - {\mathcal {C}}_{\delta _1} \int _{{\mathcal {A}}} H^p({{\hat{u}}}^s \nabla \eta )\, dx\\&= \frac{1}{1+2^{p+1}\delta _1} \cdot \frac{1}{s^p} \int _{{\mathcal {A}}} H^p(\nabla (\eta {{\hat{u}}}_m^{s-1} {{\hat{u}}})) \,dx - {\mathcal {C}}_{\delta _1} \int _{{\mathcal {A}}} H^p({{\hat{u}}}_m^{s-1} {{\hat{u}}}\nabla \eta ) \, dx \end{aligned} \end{aligned}$$
(3.25)
$$\begin{aligned} \begin{aligned} \int _{{\mathcal {B}}} H^p(\eta {{\hat{u}}}_m^{s-1}\nabla {{\hat{u}}}) \,dx&= \int _{{\mathcal {B}}} H^p(\eta m^{s-1}\nabla {{\hat{u}}}) \,dx\\&\ge \frac{1}{1+2^{p+1}\delta _2} \int _{{\mathcal {B}}} H^p(\nabla (\eta m^{s-1} {{\hat{u}}})) \,dx \\ {}&\quad - {\mathcal {C}}_{\delta _2} \int _{{\mathcal {B}}} H^p(m^{s-1} {{\hat{u}}} \nabla \eta ) \, dx\\&= \frac{1}{1+2^{p+1}\delta _2} \int _{{\mathcal {B}}} H^p(\nabla (\eta {{\hat{u}}}_m^{s-1} {{\hat{u}}})) \,dx \\ {}&\quad - {\mathcal {C}}_{\delta _2} \int _{{\mathcal {B}}} H^p({{\hat{u}}}_m^{s-1} {{\hat{u}}}\nabla \eta ) \, dx \end{aligned} \end{aligned}$$
(3.26)

By (3.25) and (3.26), we obtain

$$\begin{aligned} \begin{aligned}&\frac{1-\varepsilon _1}{1+2^{p+1}\delta _1} \cdot \frac{p(s-1)+1}{s^p} \int _{{\mathcal {A}}} H^p(\nabla (\eta {{\hat{u}}}_m^{s-1} {{\hat{u}}})) \, dx +\frac{1-\varepsilon _2}{1 + 2^{p+1}\delta _2}\int _{{\mathcal {B}}} H^p(\nabla (\eta {{\hat{u}}}_m^{s-1} {{\hat{u}}})) \, dx\\&\quad \le {\mathcal {C}}_{\delta _1}(\varepsilon _1,p,s) \int _{{\mathcal {A}}} H^p({{\hat{u}}}_m^{s-1} {{\hat{u}}}\nabla \eta ) \, dx + {\mathcal {C}}_{\delta _2}(\varepsilon _2) \int _{{\mathcal {B}}} H^p( {{\hat{u}}}_m^{s-1} {{\hat{u}}}\nabla \eta ) \, dx\\&\quad + {\mathcal {C}}_{\varepsilon _1}(p,s,M) \int _{{\mathcal {A}}} |\nabla \eta |^p {{\hat{u}}}_m^{p(s-1)} {{\hat{u}}}^p \, dx + {\mathcal {C}}_{\varepsilon _2}(p,M) \int _{{\mathcal {B}}}|\nabla \eta |^p {{\hat{u}}}_m^{p(s-1)} {{\hat{u}}}^p \, dx \\&\quad + \int _{{\mathcal {A}}_1} \frac{\gamma }{H^\circ (x)^p} {{\hat{u}}}^{p-1} \eta ^p {{\hat{u}}}_m^{p(s-1)}{{\hat{u}}} \, dx + R^p\int _{{\mathcal {A}}_1} {{\hat{u}}}^{p^*-p} \eta ^p {{\hat{u}}}_m^{p(s-1)}{{\hat{u}}}^p \, dx. \end{aligned} \end{aligned}$$
(3.27)

Using (2.4), we deduce that

$$\begin{aligned} \begin{aligned}&\frac{1-\varepsilon _1}{1+2^{p+1}\delta _1} \cdot \frac{p(s-1)+1}{s^p} \int _{{\mathcal {A}}} H^p(\nabla (\eta {{\hat{u}}}_m^{s-1} {{\hat{u}}})) \, dx \\ {}&\quad +\frac{1-\varepsilon _2}{1 + 2^{p+1}\delta _2} \int _{{\mathcal {B}}} H^p(\nabla (\eta {{\hat{u}}}_m^{s-1} {{\hat{u}}})) \, dx\\&\quad \le \int _{{\mathcal {A}}_1} \frac{\gamma }{H^\circ (x)^p} {{\hat{u}}}^{p-1} \eta ^p {{\hat{u}}}_m^{p(s-1)}{{\hat{u}}} \, dx + \hat{{\mathcal {C}}} \int _{{\mathcal {A}}_1}|\nabla \eta |^p {{\hat{u}}}_m^{p(s-1)} {{\hat{u}}}^p \, dx \\ {}&\quad + R^p\int _{{\mathcal {A}}_1} {{\hat{u}}}^{p^*-p} (\eta {{\hat{u}}}_m^{s-1}{{\hat{u}}})^p \, dx, \end{aligned} \end{aligned}$$
(3.28)

where \(\hat{{\mathcal {C}}}\) depends on \(\delta _1, \delta _2, \varepsilon _1, \varepsilon _2, p, s, M\). Using Hardy’s and Hölder’s inequality in the right hand side of (3.28), and the definition of the sets \({\mathcal {A}}\) and \({\mathcal {B}}\), we obtain

$$\begin{aligned} \begin{aligned}&\frac{1-\varepsilon _1}{1+2^{p+1}\delta _1} \cdot \frac{p(s-1)+1}{s^p} \int _{{\mathcal {A}}} H^p(\nabla (\eta {{\hat{u}}}_m^{s-1} {{\hat{u}}})) \, dx +\frac{1-\varepsilon _2}{1 + 2^{p+1}\delta _2}\int _{{\mathcal {B}}} H^p(\nabla (\eta {{\hat{u}}}_m^{s-1} {{\hat{u}}})) \, dx\\&\quad \le \frac{\gamma }{C_H}\int _{{\mathcal {A}}} H^p(\nabla (\eta {{\hat{u}}}_m^{s-1} {{\hat{u}}})) \, dx + \frac{\gamma }{C_H} \int _{{\mathcal {B}}} H^p(\nabla (\eta {{\hat{u}}}_m^{s-1} {{\hat{u}}})) \, dx\\&\quad + \hat{{\mathcal {C}}} \int _{{\mathcal {A}}_1}|\nabla \eta |^p {{\hat{u}}}_m^{p(s-1)} {{\hat{u}}}^p \, dx + \left( \int _{{\mathcal {A}}_1} {{\hat{u}}}^{p^*} R^N \,dx\right) ^{\frac{p^*-p}{p^*}} \left( \int _{{\mathcal {A}}_1} (\eta {{\hat{u}}}_m^{s-1}{{\hat{u}}})^{p \chi }\, dx\right) ^{\frac{1}{\chi }}, \end{aligned} \end{aligned}$$
(3.29)

where \(\chi =p^*/p\). Finally, we deduce

$$\begin{aligned} \begin{aligned}&\left( \frac{1-\varepsilon _1}{1+2^{p+1}\delta _1} \cdot \frac{p(s-1)+1}{s^p} - \frac{\gamma }{C_H}\right) \int _{{\mathcal {A}}} H^p(\nabla (\eta {{\hat{u}}}_m^{s-1} {{\hat{u}}})) \, dx \\&\quad +\left( \frac{1-\varepsilon _2}{1 + 2^{p+1}\delta _2} - \frac{\gamma }{C_H}\right) \int _{{\mathcal {B}}} H^p(\nabla (\eta {{\hat{u}}}_m^{s-1} {{\hat{u}}})) \, dx\\&\quad \le \hat{{\mathcal {C}}} \int _{{\mathcal {A}}_1}|\nabla \eta |^p {{\hat{u}}}_m^{p(s-1)} {{\hat{u}}}^p \, dx + \Vert u \Vert ^{p^*-p}_{L^{p^*}({\mathcal {A}}_R)} \left( \int _{{\mathcal {A}}_1} (\eta {{\hat{u}}}_m^{s-1}{{\hat{u}}})^{p \chi }\, dx\right) ^{\frac{1}{\chi }}. \end{aligned} \end{aligned}$$
(3.30)

Noticing that \((p(s-1)+1)/s^p > \gamma /C_H\) for all \(s \in \left( (N-p)/(p \mu _2), (N-p)(p \mu _1)\right) \), we can fix \(\delta _1, \varepsilon _1>0\) sufficiently small such that

$$\begin{aligned} \frac{1-\varepsilon _1}{1+2^{p+1}\delta _1} \cdot \frac{p(s-1)+1}{s^p} - \frac{\gamma }{C_H}>0, \end{aligned}$$

and \(\delta _2, \varepsilon _2>0\) sufficiently small such that

$$\begin{aligned} \frac{1-\varepsilon _2}{1 + 2^{p+1}\delta _2} - \frac{\gamma }{C_H} > 0. \end{aligned}$$

Hence by (3.30) we get

$$\begin{aligned}{} & {} \bar{{\mathcal {C}}} \int _{{\mathcal {A}}_1} H^p(\nabla (\eta {{\hat{u}}}_m^{s-1} {{\hat{u}}})) \, dx \le \hat{{\mathcal {C}}} \int _{{\mathcal {A}}_1}|\nabla \eta |^p {{\hat{u}}}_m^{p(s-1)} {{\hat{u}}}^p \, dx + \Vert u \Vert ^{p^*-p}_{L^{p^*}({\mathcal {A}}_R)}\nonumber \\ {}{} & {} \left( \int _{{\mathcal {A}}_1} (\eta {{\hat{u}}}_m^{s-1}{{\hat{u}}})^{p \chi }\, dx\right) ^{\frac{1}{\chi }}, \end{aligned}$$
(3.31)

where \(\bar{{\mathcal {C}}}\) is a positive constant depending on \(\varepsilon _1,\varepsilon _2,\delta _1,\delta _2, p,s,\gamma ,C_H\). In conclusion, by (2.4) and Sobolev inequality we have

$$\begin{aligned} \begin{aligned} \alpha _1^p \bar{{\mathcal {C}}} {\mathcal {C}}_S^p\left( \int _{{\mathcal {A}}_1} (\eta {{\hat{u}}}_m^{s-1}{{\hat{u}}})^{p \chi }\, dx \right) ^{\frac{1}{\chi }}&\le \alpha _1^p \bar{{\mathcal {C}}} \int _{{\mathcal {A}}_1} |\nabla (\eta {{\hat{u}}}_m^{s-1} {{\hat{u}}})|^p \, dx \le \bar{{\mathcal {C}}} \int _{{\mathcal {A}}_1} H^p(\nabla (\eta {{\hat{u}}}_m^{s-1} {{\hat{u}}})) \, dx \\&\le \hat{{\mathcal {C}}} \int _{{\mathcal {A}}_1}|\nabla \eta |^p {{\hat{u}}}_m^{p(s-1)} {{\hat{u}}}^p \, dx \\&\quad + \Vert u \Vert ^{p^*-p}_{L^{p^*}({\mathcal {A}}_R)} \left( \int _{{\mathcal {A}}_1} (\eta {{\hat{u}}}_m^{s-1}{{\hat{u}}})^{p \chi }\, dx \right) ^{\frac{1}{\chi }}. \end{aligned} \end{aligned}$$
(3.32)

In order to apply the Moser’s iteration method we need to rewrite the last inequality as follows

$$\begin{aligned}{} & {} \left( \int _{{\mathcal {A}}_1} (\eta {{\hat{u}}}_m^{s-1}{{\hat{u}}})^{p \chi }\, dx \right) ^{\frac{1}{\chi }} \nonumber \\ {}{} & {} \le {\mathcal {C}}_1 \int _{{\mathcal {A}}_1}|\nabla \eta |^p {{\hat{u}}}_m^{p(s-1)} {{\hat{u}}}^p \, dx + {\mathcal {C}}_2 \Vert u \Vert ^{p^*-p}_{L^{p^*}({\mathcal {A}}_R)} \left( \int _{{\mathcal {A}}_1} (\eta {{\hat{u}}}_m^{s-1}{{\hat{u}}})^{p \chi }\, dx \right) ^{\frac{1}{\chi }}, \end{aligned}$$
(3.33)

where \({\mathcal {C}}_1:=\mathcal {{{\hat{C}}}}/(\alpha _1^p \bar{{\mathcal {C}}} {\mathcal {C}}_S^p)\), \({\mathcal {C}}_2:=1/(\alpha _1^p \bar{{\mathcal {C}}} {\mathcal {C}}_S^p)\) and \(\chi ={p^*}/{p}\).

Fix \(t\in (p^*,N/\mu _1)\) and \(k\in {\mathbb {N}}\) so that \(p\chi ^k\le t\le p\chi ^{k+1}\). Then there exist positive constants \({\mathcal {C}}_1\) and \({\mathcal {C}}_2\) such that (3.33) holds for all \(1\le s\le \min \left\{ (N-p)/(p\mu _1),\chi ^k\right\} \). Now, we set \(\sigma =(1 / (2{\mathcal {C}}_2))^\frac{1}{p^*-p}\) and choosing \({{\overline{R}}}\) sufficiently small such that (3.16) holds, we get

$$\begin{aligned} \left( \int _{{\mathcal {A}}_1} (\eta {{\hat{u}}}_m^{s-1}{{\hat{u}}})^{p \chi }\, dx \right) ^{\frac{1}{\chi }} \le {\mathcal {C}} \int _{{\mathcal {A}}_1}|\nabla \eta |^p {{\hat{u}}}_m^{p(s-1)} {{\hat{u}}}^p \, dx \end{aligned}$$
(3.34)

for all \(1\le s\le \min \left\{ (N-p)/(p\mu _1),\chi ^k\right\} \) and \({\mathcal {C}}={\mathcal {C}}_1/2.\) Applying Moser’s iteration method (see e.g. [16, 31] for further details), we conclude, after finitely many times of iteration,

(3.35)

for any \(t\in (p^*,N/\mu _1).\) This proves (3.17). \(\square \)

Let us now prove the following:

Theorem 3.3

Let u be a weak solution of (\({\mathcal {P}}_H\)). Then there exists a positive constant \(C=C(N,p,\gamma ,u)\) such that

$$\begin{aligned} |u(x)|\le \frac{C}{[H^\circ (x)]^{\frac{N-p}{p}-\sigma _1}} \qquad \text {in } {\mathcal {B}}_{R_1}^{H^\circ }, \end{aligned}$$

and that

$$\begin{aligned} |u(x)|\le \frac{C}{{[H^\circ (x)]^{\frac{N-p}{p}+\sigma _2}}} \qquad \text {in } ({\mathcal {B}}_{R_2}^{H^\circ })^c, \end{aligned}$$

where \(\sigma _1, \sigma _2\) are givem in Lemma 3.1 and \( R_1, R_2 >0\) are constants depending on \(N, p, \gamma \) and u.

Proof

Let us fix \(t:= (p^*+ N/\mu _1)/{2} \in \left( p^*, N/\mu _1\right) \) as in Lemma 3.2 and \(\kappa := \min \{\tau , \sigma \}\), where \(\tau \) and \(\sigma \) are respectively as in Lemma 3.1 and Lemma 3.2. Let \({{\bar{R}}} > 0\) such that (3.1) holds for \(\kappa \) and let us consider \({{\hat{u}}}(x)=u(Rx)\), for \(R>0\) fixed. We note that \({{\hat{u}}}\) satisfies the equation

$$\begin{aligned} -\Delta _p^H {{\hat{u}}} + c(x) {{\hat{u}}}^{p-1}=0 \qquad \text {in } {\mathcal {D}}_1, \end{aligned}$$

where

$$\begin{aligned} c(x)=-\frac{\gamma }{H^\circ (x)^p}-R^p{{\hat{u}}} ^{p^*-p}(x). \end{aligned}$$

We note that \(\displaystyle H^\circ (x)^{-p}\) is bounded in \({\mathcal {D}}_1\) and \(V_R(x):=R^p{{\hat{u}}} ^{p^*-p}(x)\in L^q(\mathcal D_1)\) with \(q=t/(p^*-p)>N/p\) due to Lemma 3.2. Hence, as in the proof of [22, Theorem 1] a classical Moser iteration argument yields

(3.36)

for any ball \({\mathcal {B}}^{H^\circ }_{2r}(x) \subset {\mathcal {D}}_1\), where \({\textbf{C}}={\textbf{C}}(N,p,\gamma ,\Vert V_R\Vert _{L^q(\mathcal D_1)})\). We claim that \(\Vert V_R\Vert _{L^q({\mathcal {D}}_1)}\) is uniformly bounded with respect to R. Indeed from Lemma 3.2, since

$$\begin{aligned} p-\frac{N}{q}+N\frac{p^*-p}{t}-N\frac{p^*-p}{p^*}=0, \end{aligned}$$

we get

(3.37)

where \({\mathcal {C}}\) is a positive constant depending on \(N, p, \gamma , q\) and \({{\overline{R}}}\).

Using a covering argument we deduce that

(3.38)

Noticing that \({{\hat{u}}}(x)= u(Rx)\), by (3.38) we obtain that

(3.39)

for each \(0 < R \le {{\bar{R}}}/8\) or \(R \ge 8/{{\bar{R}}}\). By applying the Hölder’s inequality in (3.39), we get

(3.40)

for each \(0< R \le {{\bar{R}}}/8\) or \(R \ge 8/{{\bar{R}}}\) and \({\textbf{C}}\) depends only on \(N, p, \gamma , q, {{\bar{R}}}\) and \(\Vert u \Vert _{L^{p^*} ({\mathbb {R}}^N)}\).

Now we note that, since \({\mathcal {A}}_R \subset {\mathcal {B}}^{H^\circ }_{{{\bar{R}}}}\) for any \(0<R\le {{\bar{R}}}/8\) and \({\mathcal {A}}_R \subset ({\mathcal {B}}^{H^\circ }_{1/{{\bar{R}}}})^c\) for any \(R \ge 8/{{\bar{R}}}\), there exist, by Lemma 3.1, \(\sigma _1, \sigma _2>0\) depending only on \(N, p, \gamma \) such that

$$\begin{aligned} \Vert u\Vert _{L^{p^*}({\mathcal {B}}^{H^\circ }_R)} \le {\mathcal {C}} R^{\sigma _1} \quad \text {for } 0 < R \le \frac{{{\bar{R}}}}{8} \end{aligned}$$

and that

$$\begin{aligned} \Vert u\Vert _{L^{p^*}(({\mathcal {B}}^{H^\circ }_R)^c)} \le \frac{{\mathcal {C}}}{R^{\sigma _2}} \quad \text {for } R \ge 8/\bar{R}. \end{aligned}$$

Now, if we set \(R_1= {{\bar{R}}}/8\) and \(R_2=8/{{\bar{R}}}\), by (3.40) we get the thesis. \(\square \)

The next result is devoted to show the existence of some special supersolutions of our problem, in order to perform a comparison between them and the solutions of the doubly critical equation (\({\mathcal {P}}_H\)).

Proposition 3.4

Given two constants \(A>0\) and \(\alpha < p\), there exist constants \(0<\varepsilon ,\delta <1\), depending on \(N,p,\gamma ,A,\alpha \), such that

$$\begin{aligned} v(H^\circ (x))=\frac{1-\delta [H^\circ (x)]^\varepsilon }{[H^\circ (x)]^{\mu _1}} \in {\mathcal {D}}^{1,p}({\mathcal {B}}^{H^\circ }_{R_1}) \end{aligned}$$
(3.41)

is a positive supersolution to equation

$$\begin{aligned} -\Delta _p^H v - \frac{\gamma }{H^\circ (x)^p} v^{p-1} = g(x) v^{p-1} \quad \text {in } {\mathcal {B}}^{H^\circ }_{R_1},\end{aligned}$$
(3.42)

for some positive constant \(0<R_1<1\) depending only on \(N,p,\gamma , A\) and \(\alpha \), where g(x) is a positive function that belongs to \(L^{\frac{N}{p}}({\mathcal {B}}^{H^\circ }_{R_1})\) such that

$$\begin{aligned} g(x) \ge A[H^\circ (x)]^{-\alpha }\quad \text {in }{\mathcal {B}}^{H^\circ }_{R_1}. \end{aligned}$$
(3.43)

In a similar way, given \(A>0\) and \(\alpha > p\), there exist \(0<\varepsilon , \delta <1\) such that

$$\begin{aligned} v(H^\circ (x))=\frac{1-\delta [H^\circ (x)]^{-\varepsilon }}{[H^\circ (x)]^{\mu _2}} \in {\mathcal {D}}^{1,p}\left( \left( {\mathcal {B}}^{H^\circ }_{R_2}\right) ^c\right) \end{aligned}$$
(3.44)

is a positive supersolution to equation

$$\begin{aligned} -\Delta _p^H v - \frac{\gamma }{H^\circ (x)^p} v^{p-1} = g(x)v^{p-1} \quad \text {in } ({\mathcal {B}}^{H^\circ }_{R_2})^c,\end{aligned}$$
(3.45)

for some positive constant \(R_2>1\) depending only on \(N,p,\gamma \) and \(\alpha \), where g(x) is a positive function that belongs to \(L^{\frac{N}{p}}(({\mathcal {B}}^{H^\circ }_{R_2})^c)\) such that

$$\begin{aligned} g(x) \ge A[H^\circ (x)]^{-\alpha }\quad \text {in } ({\mathcal {B}}^{H^\circ }_{R_2})^c. \end{aligned}$$
(3.46)

Proof

Let us consider \(\mu , \delta , \varepsilon >0\) and let us define the function

$$\begin{aligned} u(x)=v(H^\circ (x))=\frac{1-\delta [H^\circ (x)]^\varepsilon }{[H^\circ (x)]^\mu }. \end{aligned}$$

It is easy to deduce that

$$\begin{aligned} \nabla u (x) = s(H^\circ (x)) \nabla H ^\circ (x), \end{aligned}$$
(3.47)

where \(s(t):=t^{-\mu -1}[-\mu + \delta (\mu -\varepsilon )t^\varepsilon ]\). Using (2.2), we now compute

$$\begin{aligned} \begin{aligned} -\Delta _p^H u =&-{\text {div}} \left( H^{p-1} (\nabla u) \nabla H (\nabla u) \right) = - {\text {div}} \left( |s(H^\circ (x))|^{p-1} {\text {sign}}(s) \nabla H (\nabla H^\circ (x))\right) \\ =&-(p-1) |s(H^\circ (x))|^{p-2} s'(H^\circ (x)) \langle \nabla H^\circ (x), \nabla H(\nabla H^\circ (x)) \rangle \\&- |s(H^\circ (x))|^{p-1} {\text {sign}}(s) {\text {div}} \left( \nabla H (\nabla H^\circ (x))\right) \\ =&-(p-1) |s(H^\circ (x))|^{p-2} s'(H^\circ (x)) - (N-1) \frac{|s(H^\circ (x))|^{p-2}s(H^\circ (x)) }{H^\circ (x)}, \end{aligned} \end{aligned}$$
(3.48)

where in the last line we used the fact that \(\langle \nabla H^\circ (x), \nabla H(\nabla H^\circ (x)) \rangle =1\) and

$$\begin{aligned} \displaystyle {\text {div}}\left( \nabla H (\nabla H^\circ (x))\right) = \frac{N-1}{H^\circ (x)} \end{aligned}$$

due to (2.2) and (2.3).

Making standard computations on the right hand side of (3.48), one can deduce

$$\begin{aligned} \begin{aligned} -\Delta _p^H u -\frac{\gamma }{[H^\circ (x)]^p} |u|^{p-2} u= g(H^\circ (x)) |u|^{p-2} u, \end{aligned} \end{aligned}$$
(3.49)

where

$$\begin{aligned} g(t)= \frac{h(t)}{|1-\delta t^\varepsilon |^{p-2} (1-\delta t^\varepsilon ) t^p}, \end{aligned}$$

and

$$\begin{aligned} \begin{aligned} h(t):=&-\left| \mu - \delta (\mu - \varepsilon ) t^\varepsilon \right| ^{p-2} \left\{ \mu ^2(p-1) - (N-p) \mu - (p-1)(\mu -\varepsilon )^2 \delta t^\varepsilon \right. \\&\qquad \qquad \left. + (N-p) (\mu - \varepsilon )\delta t^{\varepsilon } \right\} \\&- \gamma |1-\delta t^\varepsilon |^{p-2}(1-\delta t^\varepsilon ), \qquad \text {for } t \in [0,+\infty ). \end{aligned} \end{aligned}$$
(3.50)

We note that \(h(0)= - |\mu |^{p-2}[\mu ^2(p-1)-\mu (N-p)]-\gamma \). Hence, using the definition of \(\mu _1\) and \(\mu _2\), we deduce that \(h(0)=0\) when \(\mu =\mu _1\) and \(\mu =\mu _2\). Also we have \(h'(0)>0\) if \(\mu =\mu _1\), \(\varepsilon >0\) or \(\mu =\mu _2\), \(\varepsilon <0\). This implies there exist \(0<\delta _h<1\) such that

$$\begin{aligned} 2h'(0)t\ge h(t)\ge \frac{1}{2}h'(0)t>0 \qquad \forall t\in (0,\delta _h). \end{aligned}$$
(3.51)

We set \(\delta =\min \{\delta _h, 1/2\}\), \(\varepsilon =(p-\alpha )/2\) and

$$\begin{aligned} R_0=\left\{ 1,\left( \frac{\delta }{2A}h'(0)\right) ^{1/(p-\alpha -\varepsilon )}\right\} . \end{aligned}$$

It easy to check that \(v(H^\circ (x))=(1-\delta [H^\circ (x)]^\varepsilon )[H^\circ (x)]^{-\mu _1} \in {\mathcal {D}}^{1,p}({\mathcal {B}}^{H^\circ }_{R_1})\) is positive, which thanks to (3.51), \(g(x)\in L^{\frac{N}{p}}({\mathcal {B}}^{H^\circ }_{R_1}) \) and it satisfies (3.43). The other case is similar. \(\square \)

Now, we consider the following equation

$$\begin{aligned} -\Delta _p^H w - \frac{\gamma }{H^\circ (x)^p}w^{p-1}=f(x) w^{p-1} \qquad \text {in } \Omega , \end{aligned}$$
(3.52)

where \(\Omega \) is an open subset of \({\mathbb {R}}^N\), \(w>0\) and \(w \in {\mathcal {D}}^{1,p}(\Omega )\). Let us start with a comparison principle in bounded domains.

The first result is given by the following pointwise estimate, in the same spirit of [20, 31].

Proposition 3.5

Let uv two weakly differentiable strictly positive functions on a domain \(\Omega \).Footnote 1 We have that:

  • (i) if \(p \ge 2\), then

    $$\begin{aligned} \begin{aligned}&H^{p-1}(\nabla u) \langle \nabla H (\nabla u), \nabla \left( u - \frac{v^p}{u^p} u \right) \rangle + H^{p-1}(\nabla v) \langle \nabla H (\nabla v), \nabla \left( v - \frac{u^p}{v^p} v \right) \rangle \\&\quad \ge C_{p}(u^p+v^p) H^p(\nabla (\ln u -\ln v)), \end{aligned}\nonumber \\ \end{aligned}$$
    (3.53)

    for some positive constant \(C_{p}\) depending only on p;

  • (ii) if \(1<p<2\), then

    $$\begin{aligned} \begin{aligned}&H^{p-1}(\nabla u) \langle \nabla H (\nabla u), \nabla \left( u - \frac{v^p}{u^p} u \right) \rangle + H^{p-1}(\nabla v) \langle \nabla H (\nabla v), \nabla \left( v - \frac{u^p}{v^p} v \right) \rangle \\&\quad \ge C_p (u^p+v^p)\left[ H(\nabla \ln u) + H(\nabla \ln v)\right] ^{p-2}H^2(\nabla (\ln u -\ln v)), \end{aligned}\nonumber \\ \end{aligned}$$
    (3.54)

    for some positive constant \(C_{p}\) depending only on p.

Proof

Let uv two weakly differentiable positive functions and consider the following

$$\begin{aligned} \psi _1:=u-\frac{v^p}{u^p}u \qquad \text {and} \qquad \psi _2:=v-\frac{u^p}{v^p}v. \end{aligned}$$

Then, thanks to the Euler’s theorem for 1-homogeneous functions, and since \(\nabla H\) is 0-homogeneus, we deduce that

$$\begin{aligned} \begin{aligned} H^{p-1}(\nabla u) \langle \nabla&H (\nabla u), \nabla \psi _1 \rangle \\&= H^p(\nabla u) - p \frac{v^{p-1}}{u^{p-1}} H^{p-1}(\nabla u) \langle \nabla H (\nabla u), \nabla v \rangle + (p-1) \frac{v^p}{u^p} H^p(\nabla u)\\&= H^p(\nabla u) - v^p \left[ (1-p)H^p\left( \frac{\nabla u}{u}\right) + p H^{p-1} \left( \frac{\nabla u}{u}\right) \langle \nabla H(\nabla u), \frac{\nabla v}{v} \rangle \right] \\&= H^p(\nabla u) - v^p \left[ H^p(\nabla \ln u) + p H^{p-1} (\nabla \ln u)\langle \nabla H(\nabla \ln u), \nabla \ln v - \nabla \ln u \rangle \right] \end{aligned} \end{aligned}$$
(3.55)

and

$$\begin{aligned} \begin{aligned} H^{p-1}(\nabla v) \langle \nabla&H (\nabla v), \nabla \psi _2 \rangle \\&= H^p(\nabla v) - p \frac{u^{p-1}}{v^{p-1}} H^{p-1}(\nabla v) \langle \nabla H (\nabla v), \nabla u \rangle + (p-1) \frac{u^p}{v^p} H^p(\nabla v )\\&= H^p(\nabla v) - u^p \left[ (1-p)H^p\left( \frac{\nabla v}{v}\right) + p H^{p-1} \left( \frac{\nabla v}{v}\right) \langle \nabla H(\nabla v), \frac{\nabla u}{u} \rangle \right] \\&= H^p(\nabla v) - u^p \left[ H^p (\nabla \ln v) + p H^{p-1} (\nabla \ln v)\langle \nabla H(\nabla \ln v), \nabla \ln u - \nabla \ln v \rangle \right] . \end{aligned} \end{aligned}$$
(3.56)

(i) \(p \ge 2\). We recall that when \(p \ge 2\), it holds (2.10), i.e.

$$\begin{aligned}{} & {} H^p(\eta ) \ge H^p(\eta ') + p H^{p-1}(\eta ') \langle \nabla H(\eta '), \eta -\eta ' \rangle \nonumber \\ {}{} & {} + \frac{1}{2^{p-1}-1} H^p(\eta - \eta '), \qquad \forall \eta , \eta ' \in {\mathbb {R}}^N. \end{aligned}$$
(3.57)

Hence we can apply this inequality, in order to give an estimate from below for (3.55) and (3.56):

$$\begin{aligned}{} & {} \begin{aligned} H^{p-1}(\nabla u) \langle \nabla H (\nabla u), \nabla \psi _1 \rangle&= H^p(\nabla u) - v^p \Big [H^p(\nabla \ln u) + p H^{p-1} (\nabla \ln u)\\&\qquad \langle \nabla H(\nabla \ln u),\nabla \ln v - \nabla \ln u \rangle \Big ]\\&\ge H^p(\nabla u) - v^p \\&\quad \left[ H^p(\nabla \ln v) - C_p H^p(\nabla (\ln v - \ln u))\right] \\&= H^p(\nabla u) - H^p(\nabla v) + C_p v^p H^p(\nabla (\ln v - \ln u)), \end{aligned}\nonumber \\ \end{aligned}$$
(3.58)
$$\begin{aligned}{} & {} \begin{aligned} H^{p-1}(\nabla v) \langle \nabla H (\nabla v), \nabla \psi _2 \rangle&= H^p(\nabla v) - u^p \Big [H^p(\nabla \ln v) + p H^{p-1} (\nabla \ln v)\\&\langle \nabla H(\nabla \ln v), \nabla \ln u - \nabla \ln v \rangle \Big ]\\&\ge H^p(\nabla v) - u^p \left[ H^p(\nabla \ln u) - C_p H^p(\nabla (\ln u - \ln v))\right] \\&= H^p(\nabla v) - H^p(\nabla u) + C_p u^pH^p(\nabla (\ln u - \ln v)). \end{aligned}\nonumber \\ \end{aligned}$$
(3.59)

Adding both these two inequalities, we obtain

$$\begin{aligned} \begin{aligned}&H^{p-1}(\nabla u) \langle \nabla H (\nabla u), \nabla \psi _1 \rangle + H^{p-1}(\nabla v) \langle \nabla H (\nabla v), \nabla \psi _2 \rangle \\ {}&\quad \ge C_p(u^p+v^p) H^p(\nabla (\ln u - \ln v)). \end{aligned} \end{aligned}$$
(3.60)

(ii) \(1< p < 2\). We recall that when \(1< p < 2\), it holds (2.11), i.e.

$$\begin{aligned}{} & {} H^p(\eta ) \ge H^p(\eta ') + p H^{p-1}(\eta ') \langle \nabla H (\eta '), \eta -\eta ' \rangle \nonumber \\ {}{} & {} + C_p [H(\eta ) + H(\eta ')]^{p-2}H^2(\eta - \eta '), \quad \forall \eta , \eta ' \in {\mathbb {R}}^N, \end{aligned}$$
(3.61)

where \(C_p\) is a positive constant depending only on p.

Now we proceed exactly as in the previous case to get an estimate from below for (3.55) and (3.56):

$$\begin{aligned} \begin{aligned}&H^{p-1}(\nabla u) \langle \nabla H (\nabla u), \nabla \psi _1 \rangle \\&= H^p(\nabla u) - v^p \left[ H^p(\nabla \ln u) + p H^{p-1} (\nabla \ln u)\langle \nabla H(\nabla \ln u), \nabla \ln v - \nabla \ln u \rangle \right] \\&\ge H^p(\nabla u) - v^p \left[ H^p(\nabla \ln v) - C_p [H(\nabla \ln u) + H(\nabla \ln v)]^{p-2}H^2(\nabla (\ln v - \ln u))\right] \\&= H^p(\nabla u) - H^p(\nabla v) + C_p v^p \left[ [H(\nabla \ln u) + H(\nabla \ln v)]^{p-2} H^2(\nabla (\ln v - \ln u)\right] , \end{aligned} \end{aligned}$$
(3.62)
$$\begin{aligned} \begin{aligned}&H^{p-1}(\nabla v) \langle \nabla H (\nabla v), \nabla \psi _2 \rangle \\&= H^p(\nabla v) - u^p \left[ H^p(\nabla \ln v) + p H^{p-1} (\nabla \ln v)\langle \nabla H(\nabla \ln v), \nabla \ln u - \nabla \ln v \rangle \right] \\&\ge H^p(\nabla v) - u^p \left[ H^p(\nabla \ln u) - C_p [H(\nabla \ln u) + H(\nabla \ln v)]^{p-2}H^2(\nabla (\ln v - \ln u))\right] \\&= H^p(\nabla v) - H^p(\nabla u) +C_p u^p[H(\nabla \ln u) + H(\nabla \ln v)]^{p-2}H^2(\nabla (\ln v - \ln u)). \end{aligned} \end{aligned}$$
(3.63)

Adding both these two inequalities, we obtain

$$\begin{aligned} \begin{aligned}&H^{p-1}(\nabla u) \langle \nabla H (\nabla u), \nabla \psi _1 \rangle + H^{p-1}(\nabla v) \langle \nabla H (\nabla v), \nabla \psi _2 \rangle \\&\ge C_p (u^p+v^p)[H(\nabla \ln u) + H(\nabla \ln v)]^{p-2}H^2(\nabla (\ln v - \ln u)). \end{aligned} \end{aligned}$$
(3.64)

\(\square \)

Now, we are ready to prove the comparison principles in bounded and exteriors domains.

Proposition 3.6

Let \(\Omega \) be an open bounded smooth domain of \({\mathbb {R}}^N\) and \(f\in L^{\frac{N}{p}}(\Omega )\). Let \(u \in {\mathcal {D}} ^{1,p}(\Omega )\) be a weak positive subsolution to (3.52) and \(v\in {\mathcal {D}}^{1,p}(\Omega )\) be a weak positive supersolution of

$$\begin{aligned} -\Delta _p^H v - \frac{\gamma }{H^\circ (x)^p}v^{p-1}=g(x) v^{p-1} \qquad \text {in } \Omega , \end{aligned}$$
(3.65)

with \(g\in L^{\frac{N}{p}}(\Omega )\). Assume that \(\inf _\Omega v >0\) and \(f\le g\) in \(\Omega .\) If \(u \le v\) on \(\partial \Omega \), then

$$\begin{aligned} u \le v \quad \text {in } \Omega . \end{aligned}$$

Proof

We will give the proof of this result in the case \(p \ge 2\). The case \(1<p<2\) is similar.

Let us define

$$\begin{aligned} \eta _1:=\frac{\min \{(u^p-v^p)^+,m\}}{u^{p-1}} \qquad \text {and} \qquad \eta _2:=\frac{ \min \{(u^p-v^p)^+,m\}}{v^{p-1}} \end{aligned}$$

where \(m>1\). It is quite standard to show that \(\eta _1\) and \(\eta _2\) are good test function that we can use in the weak formulations of (3.52) and (3.65). Taking both these test function and subtracting the two equations, we obtain

$$\begin{aligned} \begin{aligned} \int _\Omega H^{p-1}(\nabla u) \langle \nabla H (\nabla u), \nabla \eta _1 \rangle \, dx&- \int _\Omega H^{p-1}(\nabla v) \langle \nabla H (\nabla v), \nabla \eta _2 \rangle \, dx\\ \le&\int _\Omega \frac{\gamma }{H^\circ (x)^p}u^{p-1} \eta _1 \, dx - \int _\Omega \frac{\gamma }{H^\circ (x)^p}v^{p-1} \eta _2 \, dx \\&+ \int _\Omega f(x)u^{p-1} \eta _1 \,dx - \int _\Omega g(x) v^{p-1} \eta _2 \, dx\\ =&\int _\Omega \frac{\gamma }{H^\circ (x)^p} \left( u^{p-1}\eta _1-v^{p-1}\eta _2\right) \, dx\\&+ \int _\Omega f(x) u^{p-1} \eta _1 -g(x)v^{p-1} \eta _2 \,dx \le 0, \end{aligned} \end{aligned}$$
(3.66)

since \(f\le g\) in \(\Omega \). Hence, setting \(\Omega _1:= \{x \in \Omega \ | \ 0 \le u^p-v^p \le m\}\) and \(\Omega _2:= \{x \in \Omega \ | \ u^p-v^p \ge m\}\), we deduce that

$$\begin{aligned} \begin{aligned}&\int _{\Omega _1} H^{p-1}(\nabla u) \langle \nabla H (\nabla u), \nabla \left( u-\frac{v^p}{u^p}u \right) \rangle \, dx \\ {}&\quad +\int _{\Omega _1} H^{p-1}(\nabla v) \langle \nabla H (\nabla v), \nabla \left( v-\frac{u^p}{v^p}v \right) \rangle \, dx \\&\quad + m(1-p) \left[ \int _{\Omega _2} H^{p-1}(\nabla u) \langle \nabla H (\nabla u), u^{-p}\nabla u \rangle \, dx \right. \\ {}&\quad \left. - \int _{\Omega _2} H^{p-1}(\nabla v) \langle \nabla H (\nabla v), v^{-p}\nabla v\rangle \, dx \right] \le 0. \end{aligned} \end{aligned}$$
(3.67)

Applying (3.53) in (3.67) and making some computations we obtain

$$\begin{aligned} \begin{aligned} C_p&\int _{\Omega _1} (u^p+v^p)H^p(\nabla (\ln u -\ln v)) \, dx \\&+ m(p-1) \left[ \int _{\Omega _2} H^{p}(\nabla \ln v) \, dx - \int _{\Omega _2} H^{p}(\nabla \ln u) \, dx \right] \le 0, \end{aligned} \end{aligned}$$
(3.68)

but this implies that

$$\begin{aligned} \begin{aligned} C_p \int _{\Omega _1} (u^p+v^p)H^p(\nabla (\ln u -\ln v)) \, dx \le m(p-1) \int _{\Omega _2} H^{p}(\nabla \ln u) \, dx. \end{aligned} \end{aligned}$$
(3.69)

For the right hand side of (3.69), we have

$$\begin{aligned} \begin{aligned}&m(p-1) \int _{\Omega _2} H^{p}(\nabla \ln u) \, dx = (p-1) \int _{\Omega _2} mu^{-p} H^p(\nabla u) \, dx \\ {}&\quad \le (p-1)\int _{\Omega _2'} H^p(\nabla u) \, dx, \end{aligned} \end{aligned}$$
(3.70)

where \(\Omega _2':=\{x \in \Omega \ | \ u^p \ge m\}\) and it holds that

$$\begin{aligned} \displaystyle \int _{\Omega _2'} H^p(\nabla u) \, dx \rightarrow 0\qquad \text {for } m \rightarrow + \infty . \end{aligned}$$

Hence, passing to the limit for \(m \rightarrow + \infty \) in (3.69) we obtain that

$$\begin{aligned} \begin{aligned} \int _{\{u^p-v^p \ge 0\}} (u^p+v^p)H^p(\nabla (\ln u -\ln v)) \, dx =0, \end{aligned} \end{aligned}$$
(3.71)

which clearly implies that

$$\begin{aligned} u = K v \qquad \text {in the set } \{x \in \Omega \ | \ u(x) \ge v(x)\}, \end{aligned}$$

for some positive constant K. By our assumptions \(\inf _{x \in \Omega } v > 0\) and \( u \le v\) on \(\partial \Omega \), hence it follows that \(K=1\). But this implies that

$$\begin{aligned} u \le v \qquad \text {in } \Omega \end{aligned}$$

and this complete the proof of this result in the case \(p \ge 2\). The case \(1<p<2\) follows repeating verbatim the proof of the case \(p\ge 2\), but applying inequality (3.54) instead of (3.53). \(\square \)

Now we want to prove the corresponding result of Proposition 3.6 in exterior domains.

Proposition 3.7

Let \(\Omega \) be an exterior domain such that \({\mathbb {R}}^N \setminus \Omega \) is bounded and \(f\in L^{\frac{N}{p}}(\Omega )\). Let \(u \in \mathcal D^{1,p}(\Omega ) \) be a weak positive subsolutions to (3.52) and \(v\in {\mathcal {D}}^{1,p}(\Omega )\) be a positive supersolution of

$$\begin{aligned} -\Delta _p^H v - \frac{\gamma }{H^\circ (x)^p}v^{p-1}=g(x) v^{p-1} \qquad \text {in } \Omega , \end{aligned}$$
(3.72)

with \(g\in L^{\frac{N}{p}}(\Omega )\). Assume that \(\inf _\Omega v >0\) and \(f\le g\) in \(\Omega \). If \(u \le v\) on \(\partial \Omega \) and

$$\begin{aligned} \limsup _{R \rightarrow + \infty } \frac{1}{R} \int _{{B}_{2R} \setminus {B}_R} u^p |\nabla \log v|^{p-1}=0, \end{aligned}$$
(3.73)

then

$$\begin{aligned} u \le v \quad \text {in } \Omega . \end{aligned}$$

Proof

In the same spirit of Proposition 3.6 we prove our result in the case \(p \ge 2\). The other case is similar and it can be shown using the same arguments. To this aim, let \(\varphi _R \in C^\infty _c({B}_{2R})\) be a standard cut-off function such that

$$\begin{aligned} {\left\{ \begin{array}{ll} 0 \le \varphi _R \le 1\\ \varphi _R \equiv 1 \qquad \ \ \text {on } {B}_R\\ |\nabla \varphi _R| \le \frac{2}{R} \quad \text {on } {B}_{2R} \setminus {B}_R. \end{array}\right. } \end{aligned}$$

Let us define

$$\begin{aligned} \eta _1:=\varphi _R\frac{\min \{(u^p-v^p)^+,m\}}{u^{p-1}} \qquad \text {and} \qquad \eta _2:=\varphi _R \frac{\min \{(u^p-v^p)^+,m\}}{v^{p-1}}. \end{aligned}$$

where \(m>1\). As pointed out in the proof of previous proposition, it is possible to show, by standard arguments, that \(\eta _1\) and \(\eta _2\) are good test functions for the weak formulations (3.52) and (3.72). Hence, we obtain

$$\begin{aligned} \begin{aligned} \int _\Omega H^{p-1}(\nabla u) \langle \nabla H (\nabla u), \nabla \eta _1 \rangle \, dx&- \int _\Omega H^{p-1}(\nabla v) \langle \nabla H (\nabla v), \nabla \eta _2 \rangle \, dx\\&\le \int _\Omega \frac{\gamma }{H^\circ (x)^p} \left( u^{p-1}\eta _1-v^{p-1}\eta _2\right) \, dx\\&+ \int _\Omega f(x) u^{p-1} \eta _1 -g(x)v^{p-1} \eta _2 \,dx \le 0, \end{aligned} \end{aligned}$$
(3.74)

since \(f\le g\) in \(\Omega .\) Now we explicitly compute the left hand side of (3.74).

$$\begin{aligned} \begin{aligned} 0&\ge \int _\Omega H^{p-1}(\nabla u) \langle \nabla H (\nabla u), \nabla \eta _1 \rangle \, dx - \int _\Omega H^{p-1}(\nabla v) \langle \nabla H (\nabla v), \nabla \eta _2 \rangle \, dx\\&= \int _{\Omega _1} H^{p-1}(\nabla u) \langle \nabla H (\nabla u), \nabla \left( u-\frac{v^p}{u^p}u\right) \rangle \varphi _R \\ {}&\quad + H^{p-1}(\nabla v) \langle \nabla H (\nabla v), \nabla \left( v-\frac{u^p}{v^p}v\right) \rangle \varphi _R \, dx\\&\quad +\int _{\Omega _1} \left( u-\frac{v^p}{u^p}u\right) H^{p-1}(\nabla u) \langle \nabla H (\nabla u), \nabla \varphi _R \rangle \\ {}&\quad + \left( v-\frac{u^p}{v^p}v\right) H^{p-1}(\nabla v) \langle \nabla H (\nabla v), \nabla \varphi _R \rangle \, dx\\&\quad +\int _{\Omega _2} H^{p-1}(\nabla u) \langle \nabla H (\nabla u), \nabla \left( u^{1-p}\right) \rangle m \varphi _R \,dx \\ {}&\quad +\int _{\Omega _2} H^{p-1}(\nabla u) \langle \nabla H (\nabla u), \nabla \varphi _R\rangle mu^{1-p} \,dx \\&\quad -\int _{\Omega _2} H^{p-1}(\nabla v) \langle \nabla H (\nabla v), \nabla \left( v^{1-p}\right) \rangle m \varphi _R \,dx \\ {}&\quad - \int _{\Omega _2} H^{p-1}(\nabla v) \langle \nabla H (\nabla v), \nabla \varphi _R \rangle mv^{1-p}\, dx\\&\quad =: I_1+I_2+I_3+I_4, \end{aligned} \end{aligned}$$
(3.75)

where \(\Omega _1:= \{x \in \Omega \ | \ 0 \le u^p-v^p \le m\}\) and \(\Omega _2:= \{x \in \Omega \ | \ u^p-v^p \ge m\}\). By Proposition 3.5 and using the definition of \(\varphi _R\), it follows that there exits a positive constant depending only on p such that

$$\begin{aligned} I_1 \ge C_p \int _{\Omega _1 \cap {B}_R}(u^p+v^p) H^p(\nabla (\ln u -\ln v)) \, dx. \end{aligned}$$
(3.76)

Now we are going to give estimates for \(I_2, I_3\) and \(I_4\). We start with \(I_2\). Using (2.6) and the Cauchy–Schwarz inequality, setting \({\tilde{\Omega }}_1:=\{x\in \Omega : v^p\le u^p\},\) we have

$$\begin{aligned} \begin{aligned} |I_2|&\le M\int _{\Omega _1} \left| u-\frac{v^p}{u^p}u\right| H^{p-1}(\nabla u) |\nabla \varphi _R| \,dx + M\int _{\Omega _1} \left| v-\frac{u^p}{v^p}v\right| H^{p-1} (\nabla v) |\nabla \varphi _R|\,dx\\&\le M\int _{{\tilde{\Omega }}_1} \left( 2u H^{p-1}(\nabla u) + v H^{p-1} (\nabla v) \right) |\nabla \varphi _R| \,dx + M\int _{\tilde{\Omega }_1} u^p H^{p-1}(\nabla \ln v) |\nabla \varphi _R| \,dx \\&\le \frac{{\mathcal {C}}}{R}\int _{{B}_{2R} \setminus {B}_R} u \cdot H^{p-1}(\nabla u) \,dx + \frac{{\mathcal {C}}}{R}\int _{{B}_{2R} \setminus {B}_R} v \cdot H^{p-1} (\nabla v) \,dx\\&\quad + \frac{{\mathcal {C}}}{R} \int _{{B}_{2R} \setminus {B}_R} u^p H^{p-1}(\nabla \ln v) \,dx \\&\le {\mathcal {C}} \left( \int _{{B}_{2R} \setminus {B}_R} H^{p}(\nabla u) \,dx \right) ^\frac{p-1}{p}\left( \int _{{B}_{2R} \setminus {B}_R} u^{p^*} \,dx \right) ^\frac{1}{p^*}\\&\quad + {\mathcal {C}} \left( \int _{{B}_{2R} \setminus {B}_R} H^{p}(\nabla v) \,dx \right) ^\frac{p-1}{p}\left( \int _{{B}_{2R} \setminus {B}_R} v^{p^*} \,dx \right) ^\frac{1}{p^*}\\&\quad + \frac{{\mathcal {C}}}{R} \int _{{B}_{2R} \setminus {B}_R} u^p H^{p-1}(\nabla \ln v) \,dx, \end{aligned} \end{aligned}$$
(3.77)

where in the last line we applied the Hölder inequality with conjugate exponent \(\left( N,p/(p-1),p^*\right) \) and \({\mathcal {C}}:=2M\). Passing to the limit for R that goes to \(+\infty \) in the right hand side of (3.77), using also assumption (3.73) and (2.4), we deduce that \(I_2\) goes to zero.

Now we proceed with the estimate of the term \(I_3\). By (2.5), we have

$$\begin{aligned} I_3 = m \int _{\Omega _2} H^{p-1}(\nabla u) \langle \nabla H (\nabla u), \nabla \varphi _R \rangle u^{1-p} \,dx+ m(1-p) \int _{\Omega _2} H^p(\nabla u) \varphi _R u^{-p} \, dx \end{aligned}$$

Therefore

$$\begin{aligned} |I_3| \le m M \int _{{\tilde{\Omega }}_2} H^{p-1}(\nabla u) |\nabla \varphi _R| u^{1-p} \,dx+ m(p-1) \int _{{\tilde{\Omega }}_2} H^p(\nabla u) \varphi _R u^{-p} \, dx, \end{aligned}$$

where \({\tilde{\Omega }}_2:= \{x \in \Omega \ | \ u^p \ge m\}\). Using this definition and also the properties of \(\varphi _R\) we deduce that

$$\begin{aligned} \begin{aligned} |I_3|&\le \frac{2M}{R} \int _{{\tilde{\Omega }}_2} u \cdot H^{p-1}(\nabla u) \,dx+ (p-1) \int _{{\tilde{\Omega }}_2} H^p(\nabla u) \, dx \\&\le {\mathcal {C}} \left( \int _{{B}_{2R} \setminus {B}_R} H^{p}(\nabla u) \,dx\right) ^\frac{p-1}{p} \left( \int _{{B}_{2R} \setminus {B}_R} u^{p^*} \,dx \right) ^\frac{1}{p^*} + (p-1) \int _{{\tilde{\Omega }}_2} H^p(\nabla u) \, dx, \end{aligned} \end{aligned}$$
(3.78)

where \({\mathcal {C}}:=2M\). Passing to the limit for mR that go to \(+\infty \), we deduce that

$$\begin{aligned} \lim _{m,R\rightarrow + \infty } I_3=0. \end{aligned}$$
(3.79)

For the last term \(I_4\), by (2.5), recalling \({\tilde{\Omega }}_2:= \{x\in \Omega : u^p\ge m\}\), we have

$$\begin{aligned} \begin{aligned} I_4&= m(p-1) \int _{\Omega _2} H^{p}(\nabla v) \varphi _R v^{-p} \, dx - \int _{\Omega _2} mH^{p-1}(\nabla v) \langle \nabla H(\nabla v), \nabla \varphi _R \rangle v^{1-p} \, dx\\&\ge - M \int _{\Omega _2} m H^{p-1}(\nabla \ln v) | \nabla \varphi _R |\, dx \ge M \int _{{\tilde{\Omega }}_2} u^p H^{p-1}(\nabla \ln v) | \nabla \varphi _R |\, dx\\&\ge -\frac{2M}{R} \int _{{B}_{2R} \setminus {B}_R} u^p H^{p-1}(\nabla \ln v) \, dx \end{aligned} \end{aligned}$$
(3.80)

Hence, passing to the limit the right hand side of (3.80), by (3.73) we have that \(I_4\) goes to zero when R tends to \(+\infty \).

Finally, if we combine all the estimates (3.76), (3.77), (3.79), (3.80) and we pass to the limit for \(m, R \rightarrow + \infty \) we deduce that

$$\begin{aligned} \begin{aligned} 0\ge&\int _\Omega H^{p-1}(\nabla u) \langle \nabla H (\nabla u), \nabla \eta _1 \rangle \, dx - \int _\Omega H^{p-1}(\nabla v) \langle \nabla H (\nabla v), \nabla \eta _2 \rangle \, dx\\ =&\limsup _{R \rightarrow + \infty } (I_1 + I_2 + I_3 + I_4) \ge \liminf _{R \rightarrow + \infty } (I_1 + I_2 + I_3 + I_4) \\ \ge&\int _{\{u \ge v\}} (u^p+v^p) H^p(\nabla (\ln u - \ln v)) \, dx \ge 0, \end{aligned} \end{aligned}$$

which implies \(u \le v\) in \(\Omega \) as we concluded in the proof of Proposition 3.6. \(\square \)

4 Proof of the main results

This section is dedicated to the proof of our main results: Theorems 1.1, 1.3 and 1.4.

Proof of Theorem 1.1

We start by proving (1.4). To this aim, let us consider u a solution of

$$\begin{aligned} -\Delta _p^H u - \frac{\gamma }{[H^\circ (x)]^p} u^{p-1} = u^{p^*-p}u^{p-1} \qquad \text {in } {\mathbb {R}}^N. \end{aligned}$$
(4.1)

In particular, we have that u is a subsolution of (4.1) in any bounded domain \(\Omega ={\mathcal {B}}_{R_1}^{H^\circ }\). We note that \(f(x):=u^{p^*-p}\in L^{\frac{N}{p}}({\mathcal {B}}_{R_1}^{H^\circ })\) and satisfies \(|f(x)|\le A[H^{\circ }(x)]^{-\alpha }\) with \(\alpha =\left( (N-p)/{p}-\sigma _1\right) (p^*-p)<p\) for \(x\in {\mathcal {B}}_{R_1}^{H^\circ }\) due to Theorem 3.3. By Proposition 3.4 we have that the function

$$\begin{aligned} v(H^\circ (x))=\frac{1-\delta [H^\circ (x)]^\varepsilon }{[H^\circ (x)]^{\mu _1}} \in {\mathcal {D}}^{1,p}({\mathcal {B}}^{H^\circ }_{R_1}) \end{aligned}$$

is a positive supersolution of (3.65) in \({\mathcal {B}}^{H^\circ }_{R_1} \subset \Omega \), with \(g\in L^{\frac{N}{p}}({\mathcal {B}}^{H^\circ }_{R_1})\) satisfying \(g(x)\ge A [H^{\circ }(x)]^{-\alpha }\) and where \(0<\delta , \varepsilon , R_1<1\) are positive constants depending only on \(N, p, \gamma , A\) and \(\alpha .\)

Let us consider \(\Gamma >0\), \({\mathcal {M}}= \sup _{\partial {\mathcal {B}}_{R_1}^{H^\circ }} u\), \({\mathcal {N}}= \sup _{\partial {\mathcal {B}}_{R_1}^{H^\circ }} 1/v\) and define

$$\begin{aligned} w(x):= {\mathcal {N}}({\mathcal {M}}+\Gamma )v(H^{\circ }(x)). \end{aligned}$$

It is easy to chek that w is a positive supersolution of (3.65) in \({\mathcal {B}}^{H^\circ }_{R_1} \subset \Omega \) and \(\inf _{{\mathcal {B}}^{H^\circ }_{R_1}} w = {\mathcal {M}}+\Gamma >0\) and \(u \le w\) on \(\partial {\mathcal {B}}^{H^\circ }_{R_1}\). Hence, by Proposition 3.6 we deduce that

$$\begin{aligned} u \le w \qquad \text {in } {\mathcal {B}}^{H^\circ }_{R_1}. \end{aligned}$$

Passing to the limit for \(\Gamma \rightarrow 0\) we obtain that

$$\begin{aligned} u(x) \le \frac{C}{[H^\circ (x)]^{\mu _1}} \qquad \text {in } {\mathcal {B}}^{H^\circ }_{R_1}, \end{aligned}$$

where \(C={\mathcal {M}} \cdot {\mathcal {N}}\).

Now we have to show the estimate from below. Let u be a weak solution of (\({\mathcal {P}}_H\)), then u is a supersolution of

$$\begin{aligned} -\Delta _p^H u - \frac{\gamma }{[H^\circ (x)]^p} u^{p-1} = 0 \qquad \text {in } {\mathcal {B}}^{H^\circ }_{R_1}. \end{aligned}$$
(4.2)

We set \(c_1:=\inf _{ {\mathcal {B}}_{R_1}^{H^\circ }} u>0\).

Now, we define

$$\begin{aligned} {{\tilde{w}}} (x):= c\frac{ c_1}{[H^\circ (x)]^{\mu _1}} \qquad \text {in } {\mathcal {B}}^{H^\circ }_{R_1}, \end{aligned}$$

where \(c= \inf _{\partial {\mathcal {B}}^{H^\circ }_{R_1}} [H^\circ (x)]^{\mu _1}=R_1^{\mu _1}\). The function \({{\tilde{w}}}\) is a subsolution to (4.2) in \({\mathcal {B}}^{H^\circ }_{R_1}\). Since, it holds that \(u \ge {{\tilde{w}}}\) in \(\partial {\mathcal {B}}^{H^\circ }_{R_1}\), we conclude by using Proposition 3.6 to obtain

$$\begin{aligned} u \ge {{\tilde{w}}} \qquad \text {in } {\mathcal {B}}^{H^\circ }_{R_1}, \end{aligned}$$

and hence combining the estimates from above and below we deduce that (1.4) is proved.

Now, our aim is to prove (1.5). Let us consider u a subsolution of

$$\begin{aligned} -\Delta _p^H u - \frac{\gamma }{[H^\circ (x)]^p} u^{p-1} = u^{p^*-p}u^{p-1} \qquad \text {in } ({\mathcal {B}}_{R_2}^{H^\circ })^c. \end{aligned}$$
(4.3)

We note that \(f(x):=u^{p^*-p}\in L^{\frac{N}{p}}(({\mathcal {B}}_{R_2}^{H^\circ })^c)\) and by Theorem 3.3 we have \(|f(x)|\le A[H^{\circ }(x)]^{-\alpha }\) with \(\alpha =\left( (N-p)/{p}+\sigma _2\right) (p^*-p)>p\) for \(x\in ({\mathcal {B}}_{R_2}^{H^\circ })^c\). By Proposition 3.4 the function

$$\begin{aligned} v(H^\circ (x))=\frac{1-\delta [H^\circ (x)]^{-\varepsilon }}{[H^\circ (x)]^{\mu _2}} \in {\mathcal {D}}^{1,p}(({\mathcal {B}}^{H^\circ }_{R_2})^c) \end{aligned}$$

is a positive supersolution of (3.72) in \(({\mathcal {B}}^{H^\circ }_{R_2})^c \subset \Omega \), with \(g\in L^{\frac{N}{p}}(({\mathcal {B}}^{H^\circ }_{R_2})^c)\) satisfying \(g(x)\ge A [H^{\circ }(x)]^{-\alpha }\) and where \(0<\delta , \varepsilon <1\) and \(R_2>1\) are positive constants depending only on \(N, p, \gamma , A\) and \(\alpha .\)

Let us consider \(\Gamma >0\), \({\mathcal {M}}= \sup _{(\partial {\mathcal {B}}_{R_2}^{H^\circ })^c} u\), \({\mathcal {N}}= \sup _{(\partial {\mathcal {B}}_{R_2}^{H^\circ })^c} 1/v\) and define

$$\begin{aligned} w(x):= {\mathcal {N}}({\mathcal {M}}+\Gamma )v(H^{\circ }(x)). \end{aligned}$$

We note that w is a positive supersolution of (3.72) in \(({\mathcal {B}}^{H^\circ }_{R_2})^c \subset \Omega \) and \(\inf _{({\mathcal {B}}^{H^\circ }_{R_2})^c} w = {\mathcal {M}}+\Gamma >0\) and \(u \le w\) on \((\partial {\mathcal {B}}^{H^\circ }_{R_2})^c\). We verify the condition (3.73). Since \(|\nabla \log v(x)|\le C|x|^{-1}\), by Hölder inequality we have

$$\begin{aligned} \begin{aligned} \limsup _{R \rightarrow + \infty } \frac{1}{R} \int _{{B}_{2R} \setminus {B}_R} u^p |\nabla \log v|^{p-1}\,dx\le \limsup _{R \rightarrow + \infty }\frac{C}{R}\left( \int _{B_{2R}\setminus B_R}|u|^{p^*}\,dx\right) ^{\frac{p}{p^*}}=0, \end{aligned} \end{aligned}$$
(4.4)

where C is a constant independent of R.

Hence, by Proposition 3.7 we deduce that

$$\begin{aligned} u \le w \qquad \text {in } ({\mathcal {B}}^{H^\circ }_{R_2})^c. \end{aligned}$$

Passing to the limit for \(\Gamma \rightarrow 0\) we obtain that

$$\begin{aligned} u(x) \le \frac{C}{[H^\circ (x)]^{\mu _2}} \qquad \text {in } ({\mathcal {B}}^{H^\circ }_{R_2})^c, \end{aligned}$$

where \(C={\mathcal {M}} \cdot {\mathcal {N}}\)

We conclude with the estimate from below. Let u be a weak solution of (\({\mathcal {P}}_H\)). Then u is a supersolution of

$$\begin{aligned} -\Delta _p^H u - \frac{\gamma }{[H^\circ (x)]^p} u^{p-1} = 0 \qquad \text {in } {\mathbb {R}}^N. \end{aligned}$$
(4.5)

We claim that

$$\begin{aligned} \int _{B_{2R}\setminus B_R}|\nabla \log u|^p \le CR^{N-p} \end{aligned}$$
(4.6)

for R sufficiently large and constant C independent of R. Indeed, by (4.5), we have

$$\begin{aligned} -\Delta _p^H u\ge 0\quad \text {in } {\mathbb {R}}^N. \end{aligned}$$
(4.7)

Therefore, considering the test function \(\eta =\zeta ^p u^{1-p}\), with \(\zeta \in C^1_c({\mathbb {R}}^N)\) nonnegative function, and taking in (4.7) we have

$$\begin{aligned} -(p-1)\int _{{\mathbb {R}}^N} H^p(\nabla u) \zeta ^pu^{-p}\,dx+p\int _{{\mathbb {R}}^N} H^{p-1}(\nabla u)\langle \nabla H(\nabla u),\nabla \zeta \rangle \zeta ^{p-1} u^{1-p}\,dx\ge 0. \end{aligned}$$
(4.8)

By Hölder inequality and by 0-homogeneity of \(\nabla H\) we get

$$\begin{aligned} \int _{{\mathbb {R}}^N} H^p(\nabla \log u) \zeta ^p\,dx \le M\frac{p}{p-1}\left( \int _{{\mathbb {R}}^N}H^p(\nabla \log u) \zeta ^p\,dx \right) ^{\frac{p-1}{p}}\left( \int _{{\mathbb {R}}^N}|\nabla \zeta |^p\,dx\right) ^{\frac{1}{p}}. \end{aligned}$$
(4.9)

Taking a standard cutoff function \(\zeta \) in (4.9) we get the claim (4.6).

Now we set \(c_2=\inf _{(\partial {\mathcal {B}}^{H^\circ }_{R_2})^c} u>0\) and \(c=\inf _{(\partial {\mathcal {B}}^{H^\circ }_{R_2})^c}[H^{\circ }(x)]^{\mu _2}\). We note that \(v={c_2}c{[H^{\circ }(x)]^{-\mu _2}}\) is a weak solution of (4.5). Moreover the condition (3.73) is verified. Indeed by Hölder inequality and (4.6) we have

$$\begin{aligned} \begin{aligned} \frac{1}{R} \int _{B_{2R}\setminus B_R} v^p|\nabla \log u|^{p-1}\,dx\le C R^{-1-p\mu _2+\frac{N}{p}}\left( \int _{B_{2R}\setminus B_R}|\nabla \log u|^p\,dx\right) ^{\frac{p-1}{p}} \\ \le C R^{-1-p\mu _2+\frac{p-1}{p}(N-p)+\frac{N}{p}} \rightarrow 0 \end{aligned} \end{aligned}$$
(4.10)

since \(\mu _2> (N-p)/{p}.\) Applying the Proposition 3.7 we conclude that

$$\begin{aligned} u(x)\ge v(x)=c\frac{c_2}{[H^{\circ }(x)]^{\mu _2}}\quad \text {in } ({\mathcal {B}}^{H^\circ }_{R_2})^c \end{aligned}$$

and therefore the thesis. \(\square \)

Now we prove Theorem 1.3 that will be essential to prove the asymptotic behavior of the gradient of solutions to (\({\mathcal {P}}_H\)). For the reader convenience we state a more detailed statement contained in the following:

Theorem 4.1

Let \(v\in C^{1,\alpha }_{loc}({\mathbb {R}}^N\setminus \{0\})\) be a positive weak solution of the equation

$$\begin{aligned} -\Delta _p^H v - \frac{\gamma }{[H^\circ (x)]^p} v^{p-1} = 0 \qquad \text {in } {\mathbb {R}}^N\setminus \{0\}, \end{aligned}$$
(4.11)

where \(0 \le \gamma < C_H\). Assume that there exist two positive constants C and c such that

$$\begin{aligned} \frac{c}{[H^{\circ }(x)]^{\mu _2}}\le v(x)\le \frac{C}{[H^{\circ }(x)]^{\mu _2}} \qquad \forall x\in {\mathbb {R}}^N\setminus \{0\}, \end{aligned}$$
(4.12)

and suppose that there exists a positive constant \({{\hat{C}}}\) such that

$$\begin{aligned} |\nabla v(x)|\le \frac{{{\hat{C}}}}{[H^{\circ }(x)]^{\mu _2+1}} \qquad \ \ \forall x\in {\mathbb {R}}^N\setminus \{0\}, \end{aligned}$$
(4.13)

then

$$\begin{aligned} v(x)=\frac{{{\overline{c}}}_1}{[H^{\circ }(x)]^{\mu _2}}, \qquad \qquad \qquad \qquad \end{aligned}$$
(4.14)

with

$$\begin{aligned} \overline{c_1}:=\limsup _{|x|\rightarrow 0} [H^{\circ }(x)]^{\mu _2}v(x). \qquad \ \ \ \end{aligned}$$

On the other hand, suppose that there exist two positive constants \({{\tilde{C}}}\) and \({{\tilde{c}}}\) such that

$$\begin{aligned} \frac{{{\tilde{c}}}}{[H^{\circ }(x)]^{\mu _1}}\le v(x)\le \frac{\tilde{C}}{[H^{\circ }(x)]^{\mu _1}} \qquad \forall x\in {\mathbb {R}}^N\setminus \{0\}, \end{aligned}$$
(4.15)

and suppose that there exists a positive constant \({{\overline{C}}}\) such that

$$\begin{aligned} |\nabla v(x)|\le \frac{{{\overline{C}}}}{[H^{\circ }(x)]^{\mu _1+1}} \qquad \ \ \forall x\in {\mathbb {R}}^N\setminus \{0\}, \end{aligned}$$
(4.16)

then

$$\begin{aligned} v(x)=\frac{{{\overline{c}}}_2}{[H^{\circ }(x)]^{\mu _1}}, \qquad \qquad \qquad \qquad \end{aligned}$$
(4.17)

with

$$\begin{aligned} \overline{c_2}:=\limsup _{|x|\rightarrow +\infty } [H^{\circ }(x)]^{\mu _1}v(x). \qquad \ \ \ \end{aligned}$$

Proof

We prove this result only for \(p\ge 2.\) The other case is similar. Suppose that (4.12) holds. From definition of \({{\overline{c}}}_1\), there exist r (sufficiently small) such that

$$\begin{aligned} v(x)\le \frac{{{\overline{c}}}_1+a_n}{[H^{\circ }(x)]^{\mu _2}}\quad \text {in } B_r(0)\setminus \{0\}, \end{aligned}$$
(4.18)

given \(a_n\rightarrow 0\).

On the other hand, there exist sequences of radii \(R_n\) and points \(x_n\) with \(R_n\) tending to zero and \(H^{\circ }(x_n)=R_n\), such that

$$\begin{aligned} v(x_n)\ge \frac{{{\overline{c}}}_1-a_n}{[H^{\circ }(x_n)]^{\mu _2}}. \end{aligned}$$
(4.19)

Now we set

$$\begin{aligned} w_n(x):= R_n^{\mu _2}v(R_nx) \qquad \forall x\in {\mathcal {B}}_{2}^{H^\circ }\setminus {\mathcal {B}}_{1/2}^{H^\circ }. \end{aligned}$$
(4.20)

By (4.12), it follows that \(w_n\) is uniformly bounded in \(L^{\infty }({\mathcal {B}}_{2}^{H^\circ }{\setminus } {\mathcal {B}}_{1/2}^{H^\circ })\) and, since \(w_n\) satisfies the equation (4.11), by [2, 17, 18], it is also uniformly bounded in \(C^{1,\alpha }(K),\) for \(0<\alpha <1\) and for any compact set \(K\subset {\mathcal {B}}_{2}^{H^\circ }{\setminus } {\mathcal {B}}_{1/2}^{H^\circ }\). Moreover, from Ascoli-Arzela’s Theorem, we deduce that \(w_n\rightarrow w_{\infty }\) in the norm \(\Vert \cdot \Vert _{C^{1,\alpha }(K)}\), for any compact set \(K\subset {\mathcal {B}}_{2}^{H^\circ }{\setminus } {\mathcal {B}}_{1/2}^{H^\circ }.\) By (4.18), we have

$$\begin{aligned} w_{\infty }\le \frac{\overline{c}_1}{[H^{\circ }(x)]^{\mu _2}} \end{aligned}$$

and by (4.19), there exist a point \({{\overline{x}}}\in \partial {\mathcal {B}}_{1}^{H^\circ }\) such that \(w_{\infty }({{\overline{x}}})={{\overline{c}}}_1\). By the strong comparison principle [11], that holds under our assumption on H (see Sect. 2), we have

$$\begin{aligned} w_{\infty }\equiv \frac{{{\overline{c}}}_1}{[H^{\circ }(x)]^{\mu _2}} \quad \text {in }{\mathcal {B}}_{2}^{H^\circ }\setminus {\mathcal {B}}_{1/2}^{H^\circ }. \end{aligned}$$
(4.21)

Now we set

$$\begin{aligned} {{\hat{v}}}:=\frac{\overline{c}_1}{[H^{\circ }(x)]^{\mu _2}} \end{aligned}$$
(4.22)

and we note that it solves (4.11). Fix \(R>0\) sufficiently large and let \(\varphi _R \in C^\infty _c({B}_{2R})\) be a standard cut-off function such that

$$\begin{aligned} {\left\{ \begin{array}{ll} 0 \le \varphi _R \le 1\\ \varphi _R \equiv 1 \qquad \ \ \text {on } {B}_R\\ |\nabla \varphi _R| \le \frac{2}{R} \quad \text {on } {B}_{2R} \setminus {B}_R. \end{array}\right. } \end{aligned}$$
(4.23)

Let us define

$$\begin{aligned} \varphi _1:=\varphi _R\frac{\min \{(v^p-{{\hat{v}}}^p)^+,m\}}{v^{p-1}} \qquad \text {and} \qquad \varphi _2:=\varphi _R \frac{\min \{(v^p-{{\hat{v}}}^p)^+,m\}}{{{\hat{v}}}^{p-1}},\end{aligned}$$
(4.24)

where \(m>1\).

We remark that \(\varphi _1\) and \(\varphi _2\) are good test function in any domain \(0\not \in {\overline{\Omega }}.\) Testing (4.24) in (4.11) on the domain \({\mathbb {R}}^N{\setminus } {\mathcal {B}}_{R_n}^{H^\circ }\), and, since the stress field \(H(\nabla u)^{p-1}\nabla H(\nabla u)\in W^{1,2}_{loc}({\mathbb {R}}^N{\setminus } {\mathcal {B}}_{R_n}^{H^\circ })\) (see [19]), exploiting the divergence theorem, we have

$$\begin{aligned} \begin{aligned}&\int _{ {\mathbb {R}}^N\setminus {\mathcal {B}}_{R_n}^{H^\circ }} H^{p-1}(\nabla v) \langle \nabla H (\nabla v), \nabla \varphi _1 \rangle dx \, - \int _{ {\mathbb {R}}^N\setminus {\mathcal {B}}_{R_n}^{H^\circ }} H^{p-1}(\nabla {{\hat{v}}}) \langle \nabla H (\nabla {{\hat{v}}}), \nabla \varphi _2 \rangle dx\, \\&\quad - \int _{{\mathbb {R}}^N\setminus {\mathcal {B}}_{R_n}^{H^\circ }}\frac{\gamma }{H^\circ (x)^p} \left( v^{p-1}\varphi _1-v^{p-1}\varphi _2\right) \, dx\\&\quad = \int _{\partial {\mathcal {B}}_{R_n}^{H^\circ }} H^{p-1}(\nabla v) \langle \nabla H (\nabla v), \eta _n \rangle \varphi _1\, - H^{p-1}(\nabla {{\hat{v}}}) \langle \nabla H (\nabla {{\hat{v}}}), \eta _n \rangle \varphi _2 dx, \end{aligned} \end{aligned}$$
(4.25)

where \(\eta _n\) is the inner unite normal vector at \(\partial {\mathcal {B}}_{R_n}^{H^\circ }.\) Now we set \({\mathcal {A}}:= \{ 0\le v^p-{{\hat{v}}}^p\le m\}\) and \({\mathcal {B}}:=\{ v^p-{{\hat{v}}}^p\ge m\}\). Then (4.25) becomes:

$$\begin{aligned} \begin{aligned}&\int _{\partial {\mathcal {B}}_{R_n}^{H^\circ }} -H^{p-1}(\nabla v) \langle \nabla H (\nabla v), \eta _n \rangle \varphi _1\, + H^{p-1}(\nabla {{\hat{v}}}) \langle \nabla H (\nabla {{\hat{v}}}), \eta _n \rangle \varphi _2 \,dx \\ {}&\quad =\int _{({\mathbb {R}}^N\setminus {\mathcal {B}}_{R_n}^{H^\circ })\cap \mathcal A} H^{p-1}(\nabla v) \langle \nabla H (\nabla v), \nabla \left( v-\frac{{{\hat{v}}}^p}{v^p}v\right) \rangle \varphi _R\,dx \\&\quad +\int _{({\mathbb {R}}^N\setminus {\mathcal {B}}_{R_n}^{H^\circ })\cap \mathcal A} H^{p-1}(\nabla {{\hat{v}}}) \langle \nabla H (\nabla {{\hat{v}}}), \nabla \left( {{\hat{v}}}-\frac{v^p}{{{\hat{v}}}^p}{{\hat{v}}}\right) \rangle \varphi _R \, dx\\&\quad +\int _{({\mathbb {R}}^N\setminus {\mathcal {B}}_{R_n}^{H^\circ })\cap {\mathcal {A}}} \left( v-\frac{{{\hat{v}}}^p}{v^p}v\right) H^{p-1}(\nabla v) \langle \nabla H (\nabla v), \nabla \varphi _R \rangle \,dx \\&\quad +\int _{({\mathbb {R}}^N\setminus {\mathcal {B}}_{R_n}^{H^\circ })\cap {\mathcal {A}}} \left( {{\hat{v}}}-\frac{v^p}{{{\hat{v}}}^p}{{\hat{v}}}\right) H^{p-1}(\nabla {{\hat{v}}}) \langle \nabla H (\nabla {{\hat{v}}}), \nabla \varphi _R \rangle \, dx\\&\quad +\int _{({\mathbb {R}}^N\setminus {\mathcal {B}}_{R_n}^{H^\circ })\cap {\mathcal {B}}} H^{p-1}(\nabla v) \langle \nabla H (\nabla v), \nabla \left( v^{1-p}\right) \rangle m \varphi _R \\ {}&\quad + H^{p-1}(\nabla v) \langle \nabla H (\nabla v), \nabla \varphi _R\rangle mv^{1-p} \,dx \\&\quad -\int _{({\mathbb {R}}^N\setminus {\mathcal {B}}_{R_n}^{H^\circ })\cap {\mathcal {B}}} \left( H^{p-1}(\nabla {{\hat{v}}}) \langle \nabla H (\nabla {{\hat{v}}}), \nabla \left( {{\hat{v}}}^{1-p}\right) \rangle m \varphi _R \right. \\ {}&\quad \left. + H^{p-1}(\nabla {{\hat{v}}}) \langle \nabla H (\nabla {{\hat{v}}}), \nabla \varphi _R \rangle m{{\hat{v}}}^{1-p}\right) \, dx\\&\quad =: I_1+I_2+I_3+I_4+I_5+I_6. \end{aligned} \end{aligned}$$
(4.26)

Now we set \(\Omega _1:= ({\mathbb {R}}^N{\setminus } {\mathcal {B}}_{R_n}^{H^\circ })\cap {\mathcal {A}}\) and \(\Omega _2:=({\mathbb {R}}^N{\setminus } {\mathcal {B}}_{R_n}^{H^\circ })\cap \mathcal B\). By Proposition 3.5, it follows that there exits a positive constant depending only on p such that

$$\begin{aligned} I_1 +I_2\ge C_p \int _{\Omega _1\cap {B}_{2R}}(v^p+{{\hat{v}}}^p) H^p(\nabla (\ln v -\ln {{\hat{v}}}))\varphi _R \, dx. \end{aligned}$$
(4.27)

Now we estimate \(I_3+I_4\). Using (2.6) and the Hölder inequality with conjugate exponent \(\left( N,{p}/({p-1}),p^*\right) \) we have

$$\begin{aligned} \begin{aligned} |I_3+I_4|&\le M\int _{\Omega _1} \left| v-\frac{{{\hat{v}}}^p}{v^p}v\right| H^{p-1}(\nabla v) |\nabla \varphi _R| \,dx \\ {}&\quad + M\int _{\Omega _1} \left| {{\hat{v}}}-\frac{v^p}{{{\hat{v}}}^p}{{\hat{v}}}\right| H^{p-1} (\nabla {{\hat{v}}}) |\nabla \varphi _R|\,dx\\&\le M\int _{\Omega _1} \left( 2v H^{p-1}(\nabla v) + {{\hat{v}}} H^{p-1} (\nabla {{\hat{v}}}) \right) |\nabla \varphi _R| \,dx \\ {}&\quad + M\int _{\Omega _1} v^p H^{p-1}(\nabla \ln {{\hat{v}}}) |\nabla \varphi _R| \,dx \\&\le \frac{2M}{R}\int _{{B}_{2R} \setminus {B}_R} v H^{p-1}(\nabla v) \,dx + \frac{2M}{R}\int _{{B}_{2R} \setminus {B}_R} {{\hat{v}}} H^{p-1} (\nabla {{\hat{v}}}) \,dx\\&\quad + \frac{2M}{R} \int _{{B}_{2R} \setminus {B}_R} v^p H^{p-1}(\nabla \ln {{\hat{v}}}) \,dx \\&\le 2M \left( \int _{{B}_{2R} \setminus {B}_R} H^{p}(\nabla v) \,dx \right) ^\frac{p-1}{p}\left( \int _{{B}_{2R} \setminus {B}_R} v^{p^*} \,dx \right) ^\frac{1}{p^*}\\&\quad + 2M \left( \int _{{B}_{2R} \setminus {B}_R} H^{p}(\nabla {{\hat{v}}}) \,dx \right) ^\frac{p-1}{p}\left( \int _{{B}_{2R} \setminus {B}_R} {{\hat{v}}}^{p^*} \,dx \right) ^\frac{1}{p^*}\\&\quad + \frac{2M}{R} \int _{{B}_{2R} \setminus {B}_R} v^p H^{p-1}(\nabla \ln {{\hat{v}}}) \,dx. \end{aligned} \end{aligned}$$
(4.28)

By (4.22) we have \(H(\nabla \ln {{\hat{v}}}(x))\le C[H^{\circ }(x)]^{-1}\), where \(C=\mu _2\), we have that

$$\begin{aligned} \begin{aligned} \frac{1}{R}\int _{B_{2R}\setminus B_R} v^pH^{p-1}(\nabla \ln {{\hat{v}}})&\le \frac{C}{R^p}\int _{B_{2R}\setminus B_R} |v|^p \\ {}&\le C\left( \int _{B_{2R}\setminus B_R}|v|^{p^*}\right) ^{\frac{p}{p^*}}, \end{aligned} \end{aligned}$$
(4.29)

where we used the Hölder inequality. Since v has the right summability at the infinity, for R that goes to infinity in the right hand side of (4.28), we obtain that \(I_3+I_4\) goes to zero.

Now we proceed with the estimate of the term \(I_5\). Recalling that \(v^p \ge m\) in \(\Omega _2\), we get

$$\begin{aligned} |I_5| \le m M \int _{\Omega _2} H^{p-1}(\nabla v) |\nabla \varphi _R| v^{1-p} \,dx+ m(p-1) \int _{ \Omega _2} H^p(\nabla v) \varphi _R v^{-p} \, dx, \end{aligned}$$

Using the properties of \(\varphi _R\) we obtain that

$$\begin{aligned} \begin{aligned} |I_5|&\le \frac{M}{R} \int _{{B}_{2R} \setminus {B}_R} v H^{p-1}(\nabla v) \,dx+ (p-1) \int _{ \Omega _2} H^p(\nabla v) \, dx \\&\le M \left( \int _{{B}_{2R} \setminus {B}_R} H^{p}(\nabla v) \,dx\right) ^\frac{p-1}{p} \left( \int _{{B}_{2R} \setminus {B}_R} v^{p^*} \,dx \right) ^\frac{1}{p^*} + (p-1) \int _{ \Omega _2} H^p(\nabla v) \, dx, \end{aligned} \end{aligned}$$
(4.30)

Passing to the limit for mR that go to \(+\infty \), we deduce that \(I_5\) goes to zero since the set \( \Omega _2\) vanishes as \(m \rightarrow +\infty \).

For the last term \(I_6\) we obtain

$$\begin{aligned} \begin{aligned} I_6&= m(p-1) \int _{\Omega _2} H^{p}(\nabla {{\hat{v}}}) \varphi _R {{\hat{v}}}^{-p} \, dx - \int _{\Omega _2} mH^{p-1}(\nabla {{\hat{v}}}) \langle \nabla H(\nabla {{\hat{v}}}), \nabla \varphi _R \rangle {{\hat{v}}}^{1-p} \, dx\\&\ge - M \int _{\Omega _2} m H^{p-1}(\nabla \ln {{\hat{v}}}) | \nabla \varphi _R |\, dx\\&\ge -\frac{2M}{R} \int _{{B}_{2R} \setminus {B}_R} v^p H^{p-1}(\nabla \ln {{\hat{v}}}) \, dx \end{aligned} \end{aligned}$$
(4.31)

Hence, passing to the limit in the right hand side of (4.31) and using (4.29) we have that the right hand of (4.31) tends to zero when R tends to the infinity.

Now we estimate the left hand of (4.26). By (4.24) and Proposition 2.1 (in particular see (2.11)), we get

$$\begin{aligned} \begin{aligned}&\int _{\partial {\mathcal {B}}_{R_n}^{H^\circ }} -H^{p-1}(\nabla {{\hat{v}}}) \langle \nabla H (\nabla {{\hat{v}}}), \eta _n \rangle \varphi _2\, + H^{p-1}(\nabla v) \langle \nabla H (\nabla v), \eta _n \rangle \varphi _1 \,dx \\ {}&\quad = -\int _{\partial {\mathcal {B}}_{R_n}^{H^\circ }} \varphi _1 \left( H^{p-1}(\nabla {{\hat{v}}}) \langle \nabla H (\nabla {{\hat{v}}}), \eta _n \rangle \, - H^{p-1}(\nabla v) \langle \nabla H (\nabla v), \eta _n \rangle \right) \,dx \\ {}&\quad -\int _{\partial {\mathcal {B}}_{R_n}^{H^\circ }} (\varphi _2-\varphi _1 )H^{p-1}(\nabla {{\hat{v}}}) \langle \nabla H (\nabla {{\hat{v}}}), \eta _n \rangle \,dx \\&\quad \le \int _{\partial {\mathcal {B}}_{R_n}^{H^\circ }}\frac{m}{v^{p-1}}{{\tilde{C}}}_p (|\nabla {{\hat{v}}}|^{p-2}+|\nabla v|^{p-2})|\nabla {{\hat{v}}}-\nabla v| +\frac{2m}{{{\hat{v}}}^{p-1}}M\alpha _2 |\nabla {{\hat{v}}}|^{p-1}\,dx, \\ {}&\quad :=J_1+J_2. \end{aligned} \end{aligned}$$
(4.32)

where, in the last line, we used (2.4) and (2.6).

We set \(x=R_ny\), with \(y\in \partial {\mathcal {B}}_{1}^{H^\circ }\). Using this change of variables and recalling (4.12), (4.20) and (4.21), \(J_1\) can be estimated as

$$\begin{aligned} \begin{aligned} J_1\le {{\tilde{C}}}_1&\int _{\partial {\mathcal {B}}_{1}^{H^\circ }} mR_n^{N-1+\mu _2(p-1)-(\mu _2+1)(p-1)} |-\mu _2{{\overline{c}}}_1\nabla H^{\circ }(y)-\nabla w_n(y)|\,dy \\ {}&={{\tilde{C}}}_1 R_n^{N-p-\epsilon } \end{aligned} \end{aligned}$$
(4.33)

where we choose \(m=R_n^{-\epsilon }\), for \(\epsilon >0\) fixed sufficiently small, and \({{\tilde{C}}}_1\) is a positive constant.

In a similar way \(J_2\) is estimated as

$$\begin{aligned} J_2\le {{\tilde{C}}}_2 R_n^{N-p-\epsilon }, \end{aligned}$$
(4.34)

where \({{\tilde{C}}}_2\) is a positive constant.

Finally, if we combine all the estimates (4.27), (4.28), (4.30), (4.31), (4.33), (4.34) and passing to the limit for \(R_n\rightarrow 0\), and then, exploiting the Fatou’s lemma, for \(R\rightarrow +\infty \), we deduce that

$$\begin{aligned}\begin{aligned}&\int _{\{v \ge {{\hat{v}}}\}} (v^p+{{\hat{v}}}^p) H^p(\nabla (\ln v - \ln {{\hat{v}}})) \, dx=0, \end{aligned}\end{aligned}$$

which implies \(v \le {{\hat{v}}}\) as we concluded in the proof of Proposition 3.6.

To prove that \({{\hat{v}}}\le v\), let us consider

$$\begin{aligned} \varphi _1:=\varphi _R\frac{\min \{({{\hat{v}}}^p-v^p)^+,m\}}{{{\hat{v}}}^{p-1}} \qquad \text {and} \qquad \varphi _2:=\varphi _R \frac{\min \{({{\hat{v}}}^p- v^p)^+,m\}}{ v^{p-1}},\end{aligned}$$
(4.35)

where \(m>1\) and \(\varphi _R\) is the standard cutoff function defined in (4.23). Using (4.35) in the weak formulation of (4.11), proceeding in a similar way as above we obtain (4.14). The only thing to check is that

$$\begin{aligned} \frac{1}{R}\int _{B_{2R}\setminus B_R} {{\hat{v}}}^pH^{p-1}(\nabla \ln v)\,dx\rightarrow 0. \end{aligned}$$
(4.36)

Assumption (4.13) ensures (4.36), and therefore we are done.

Now, assume that (4.15) holds and suppose that \(p\ge 2\), the other case is similar. The proof is similar to the previous one and, for this reason, we omit some details. From definition of \(\overline{c}_2\), there exist r (sufficiently large) such that

$$\begin{aligned} v(x)\le \frac{{{\overline{c}}}_2+a_n}{[H^{\circ }(x)]^{\mu _1}} \quad \text {in } {\mathbb {R}}^N\setminus B_r(0), \end{aligned}$$
(4.37)

given \(a_n\rightarrow 0\).

On the other hand, there exist sequences of radii \(R_n\) and points \(x_n\) with \(R_n\) tending to infinity and \(H^{\circ }(x_n)=R_n\), such that

$$\begin{aligned} v(x_n)\ge \frac{{{\overline{c}}}_2-a_n}{[H^{\circ }(x_n)]^{\mu _1}}. \end{aligned}$$
(4.38)

Now we set

$$\begin{aligned} w_n(x):= R_n^{-\mu _1}v\left( \frac{x}{R_n}\right) \qquad \forall x\in {\mathcal {B}}_{2}^{H^\circ }\setminus {\mathcal {B}}_{1/2}^{H^\circ }. \end{aligned}$$
(4.39)

As in the previous case we obtain

$$\begin{aligned} w_{\infty }\equiv \frac{{{\overline{c}}}_2}{[H^{\circ }(x)]^{\mu _1}} \quad \text {in }{\mathcal {B}}_{2}^{H^\circ }\setminus {\mathcal {B}}_{1/2}^{H^\circ }. \end{aligned}$$
(4.40)

Now we set

$$\begin{aligned} {{\hat{v}}}:=\frac{{{\overline{c}}}_2}{[H^{\circ }(x)]^{\mu _1}} \end{aligned}$$

and we remark that it solves (4.11). We show that \(v={{\hat{v}}}\) in \({\mathbb {R}}^N\). To this aim, fix \(\varepsilon >0\) sufficiently small and let \(\varphi _\varepsilon \in C^\infty ({\mathbb {R}}^N)\) be a function such that

$$\begin{aligned} {\left\{ \begin{array}{ll} 0 \le \varphi _\varepsilon \le 1\\ \varphi _\varepsilon \equiv 0 \qquad \ \ \text {on } {B}_\varepsilon \\ \varphi _\varepsilon \equiv 1 \qquad \ \ \text {on } ({B}_{2\varepsilon })^c\\ |\nabla \varphi _\varepsilon | \le \frac{2}{\varepsilon } \quad \text {on } {B}_{2\varepsilon } \setminus {B}_\varepsilon . \end{array}\right. } \end{aligned}$$
(4.41)

Let us define

$$\begin{aligned} \varphi _1:=\varphi _\varepsilon \frac{\min \{(v^p-{{\hat{v}}}^p)^+,m\}}{v^{p-1}} \qquad \text {and} \qquad \varphi _2:=\varphi _\varepsilon \frac{\min \{(v^p-{{\hat{v}}}^p)^+,m\}}{{{\hat{v}}}^{p-1}},\end{aligned}$$
(4.42)

where \(m>1\).

We note that \(\varphi _1\) and \(\varphi _2\) are good test function in any bounded domain \(0\in \Omega .\) Testing (4.42) in (4.11) on the domain \({\mathcal {B}}_{R_n}^{H^\circ }\) we get

$$\begin{aligned} \begin{aligned}&\int _{ {\mathcal {B}}_{R_n}^{H^\circ }} H^{p-1}(\nabla v) \langle \nabla H (\nabla v), \nabla \varphi _1 \rangle dx \, - \int _{ {\mathcal {B}}_{R_n}^{H^\circ }} H^{p-1}(\nabla {{\hat{v}}}) \langle \nabla H (\nabla {{\hat{v}}}), \nabla \varphi _2 \rangle dx\, \\&\quad - \int _{ {\mathcal {B}}_{R_n}^{H^\circ }}\frac{\gamma }{H^\circ (x)^p} \left( v^{p-1}\varphi _1-v^{p-1}\varphi _2\right) \, dx\\&\quad = \int _{\partial {\mathcal {B}}_{R_n}^{H^\circ }} H^{p-1}(\nabla v) \langle \nabla H (\nabla v), \eta _n \rangle \varphi _1\, - H^{p-1}(\nabla {{\hat{v}}}) \langle \nabla H (\nabla {{\hat{v}}}), \eta _n \rangle \varphi _2 \,dx, \end{aligned} \end{aligned}$$
(4.43)

where \(\eta _n\) is the outward unite normal vector at \(\partial {\mathcal {B}}_{R_n}^{H^{\circ }}.\) By (4.42) and Proposition 2.1, the right hand of (4.43) becomes

$$\begin{aligned} \begin{aligned}&\int _{\partial {\mathcal {B}}_{R_n}^{H^\circ }} -H^{p-1}(\nabla {{\hat{v}}}) \langle \nabla H (\nabla {{\hat{v}}}), \eta _n \rangle \varphi _2\, + H^{p-1}(\nabla v) \langle \nabla H (\nabla v), \eta _n \rangle \varphi _1 \,dx \\ {}&\quad = \int _{\partial {\mathcal {B}}_{R_n}^{H^\circ }} -\varphi _1 \left( H^{p-1}(\nabla {{\hat{v}}}) \langle \nabla H (\nabla {{\hat{v}}}), \eta _n \rangle \, - H^{p-1}(\nabla v) \langle \nabla H (\nabla v), \eta _n \rangle \right) \,dx \\ {}&\quad -\int _{\partial {\mathcal {B}}_{R_n}^{H^\circ }} (\varphi _2-\varphi _1 )H^{p-1}(\nabla {{\hat{v}}}) \langle \nabla H (\nabla {{\hat{v}}}), \eta _n \rangle \,dx \\&\quad \le \int _{\partial {\mathcal {B}}_{R_n}^{H^\circ }}{{\tilde{C}}}_pv (|\nabla {{\hat{v}}}|^{p-2}+|\nabla v|^{p-2})(\nabla {{\hat{v}}}-\nabla v) +\frac{v^p}{{{\hat{v}}}^{p-1}}M\alpha _2 |\nabla {{\hat{v}}}|^{p-1}\,dx, \\ {}&\quad :={{\tilde{J}}}_1+\tilde{J}_2. \end{aligned} \end{aligned}$$
(4.44)

where, in the last line, we used (2.4) and (2.6).

We consider the following change of variables \(x=y/R_n\), with \(y\in \partial {\mathcal {B}}_{1}^{H^\circ }\). Using this fact, by (4.39) and (4.40), we have

$$\begin{aligned} \begin{aligned} {{\tilde{J}}}_1&\le {{\overline{C}}}_1\int _{\partial {\mathcal {B}}_{1}^{H^\circ }}R_n^{1-N-\mu _1+(\mu _1+1)(p-1)}|-\mu _1\overline{c}_2y-\nabla w_n(y)|\,dy \\ {}&={{\overline{C}}}_1R_n^{-N+\mu _1p+p} \end{aligned} \end{aligned}$$
(4.45)

where \( {{\overline{C}}}_1\) is a positive constant.

In a similar way \({{\tilde{J}}}_2\) is estimated as

$$\begin{aligned} {{\tilde{J}}}_2\le {{\overline{C}}}_2 R_n^{-N+\mu _1p+p}, \end{aligned}$$
(4.46)

where \({{\overline{C}}}_2\) is a positive constant. We note that \(-N+\mu _1p+p<0\) since \(0\le \mu _1<(N-p)/p\). Proceeding as in the case (4.12), we prove that

$$\begin{aligned} \int _{\{v \ge {{\hat{v}}}\}} (v^p+{{\hat{v}}}^p) H^p(\nabla (\ln v - \ln {{\hat{v}}})) \, dx=0, \end{aligned}$$
(4.47)

which implies \(v \le {{\hat{v}}}\) in \(\Omega \) as we concluded in the proof of Proposition 3.6.

To prove that \({{\hat{v}}}\le v\), let us consider

$$\begin{aligned} \varphi _1:=\varphi _\varepsilon \frac{\min \{({{\hat{v}}}^p-v^p)^+,m\}}{{{\hat{v}}}^{p-1}} \qquad \text {and} \qquad \varphi _2:=\varphi _\varepsilon \frac{\min \{({{\hat{v}}}^p- v^p)^+,m\}}{ v^{p-1}},\end{aligned}$$
(4.48)

where \(m>1\) and \(\varphi _\varepsilon \) is defined as (4.41). Testing (4.48) in (4.11), using the assumption (4.16) and proceeding in a similar way as above we get (4.17). \(\square \)

At this point we are ready to prove

Proof of Theorem 1.4

We prove (1.12), the other case is similar and it can be proved in a similar way. We start by showing the estimate from above. For \(R_n\) tending to infinity, let us consider

$$\begin{aligned} w_n(x):=R_n^{\mu _2}u(R_nx) \qquad \forall x\in {\mathcal {B}}_{2}^{H^\circ }\setminus {\mathcal {B}}_{1/2}^{H^\circ }. \end{aligned}$$
(4.49)

We remark that \(w_n\) is uniformly bounded in \(L^{\infty }({\mathcal {B}}_{2}^{H^\circ }{\setminus } {\mathcal {B}}_{1/2}^{H^\circ })\) and it weakly solves

$$\begin{aligned} -\Delta _p^Hw_n- \frac{\gamma }{H^\circ (x)^p} w_n^{p-1}= R_n^{\mu _2(p-1)+p-\mu _2(p^*-1)}w_n^{p^*-1} \qquad \text {in } {\mathbb {R}}^N. \end{aligned}$$
(4.50)

Since \(\mu _2(p-1)+p-\mu _2(p^*-1)<0\), by [2, 17, 18], \(w_n\) is also uniformly bounded in \(C^{1,\alpha }(K),\) for \(0<\alpha <1\) and for any compact set \(K\subset {\mathcal {B}}_{2}^{H^\circ }{\setminus } {\mathcal {B}}_{1/2}^{H^\circ }\). For \(R_n\) sufficiently large we get the estimate from above in (1.12).

Now we prove the estimate from below. Suppose by contradiction that there exist sequences of points \(x_n\) such that

$$\begin{aligned} {[}H^{\circ }(x_n)]^{\mu _2+1}|\nabla u(x_n)|\rightarrow 0 \qquad \text {for } |x_n|\rightarrow \infty . \end{aligned}$$
(4.51)

For \(0<a<A\) fixed, let us consider

$$\begin{aligned} w_n(x):=R_n^{\mu _2}u(R_nx). \end{aligned}$$

For n sufficiently large, from Theorem 1.1, relabeling the constants, we have

$$\begin{aligned} \frac{c}{A^{\mu _2}}\le w_n(x)\le \frac{C}{a^{\mu _2}}\qquad \text {in } \overline{{\mathcal {B}}_{A}^{H^\circ }\setminus {\mathcal {B}}_{a}^{H^\circ }}. \end{aligned}$$

Furthermore, recalling the estimate from above of the gradients of the weak solution u, proved previously, we get

$$\begin{aligned} |\nabla w_n(x)|\le \frac{{{\tilde{C}}}}{a^{\mu _2+1}} \quad \text {in } \overline{{\mathcal {B}}_{A}^{H^\circ }\setminus {\mathcal {B}}_{a}^{H^\circ }}. \end{aligned}$$
(4.52)

For aA fixed, by [2, 17, 18], \(w_n\) is also uniformly bounded in \(C^{1,\alpha }(K),\) for \(0<\alpha <1\) and for any compact set \(K\subset {\mathcal {B}}_{A}^{H^\circ }{\setminus } {\mathcal {B}}_{a}^{H^\circ }\). Moreover

$$\begin{aligned} w_n(x)\rightarrow w_{a,A} \qquad \text {in } B_A\setminus B_a, \end{aligned}$$

in the norm \(C^{1,\alpha '}\), for \(0<\alpha ' <\alpha .\) Moreover, since \(w_n\) weakly solves

$$\begin{aligned} -\Delta _p^Hw_n- \frac{\gamma }{H^\circ (x)^p} w_n^{p-1}= R_n^{\mu _2(p-1)+p-\mu _2(p^*-1)}w_n^{p^*-1} \qquad \text {in } {\mathcal {B}}_{A}^{H^\circ }\setminus {\mathcal {B}}_{a}^{H^\circ }, \end{aligned}$$
(4.53)

we deduce that

$$\begin{aligned} -\Delta _p^Hw_{a,A}- \frac{\gamma }{H^\circ (x)^p} w_{a,A}^{p-1}=0 \qquad \text {in } {\mathcal {B}}_{A}^{H^\circ }\setminus {\mathcal {B}}_{a}^{H^\circ }. \end{aligned}$$
(4.54)

Now we take \(a_j=1/j\) and \(A_j=j\), for \(j\in {\mathbb {N}}\) and we construct \(w_{a_j,A_j}\) as above. For j goes to infinity, using a standard diagonal process, we construct a limiting profile \(w_{\infty }\) so that

$$\begin{aligned} -\Delta _p^Hw_{\infty }- \frac{\gamma }{H^\circ (x)^p} w_{\infty }^{p-1}=0 \qquad \text {in } {\mathbb {R}}^N\setminus \{0\}, \end{aligned}$$
(4.55)

with \(w_{\infty }\equiv w_{a_j,A_j}\) in \({\mathcal {B}}_{A_j}^{H^\circ }\setminus {\mathcal {B}}_{a_j}^{H^\circ }\).

Since \(w_{\infty }\) satisfies the assumptions (4.12), (4.13) of the Theorem 4.1, we get

$$\begin{aligned} w_{\infty }(x)= \frac{{{\overline{c}}}_2}{[H^{\circ }(x)]^{\mu _2}}, \end{aligned}$$
(4.56)

where we set \(\overline{c_2}:=\limsup _{|x|\rightarrow 0} [H^{\circ }(x)]^{\mu _2}w_{\infty }(x)\).

Now we set \(y_n=x_n/R_n\), and by (4.51), we deduce that \(|\nabla w_n(y_n)|\) tends to zero as \(R_n\) tends to infinity. This fact and the uniform convergence of the gradients imply that there exist \({{\overline{y}}}\in \partial {\mathcal {B}}_{1}^{H^\circ }\) such that

$$\begin{aligned} |\nabla w_{\infty }({{\overline{y}}})|=0. \end{aligned}$$

This is an absurd since the solution \(w_{\infty }\) has no critical points. \(\square \)