1 Introduction

Let \(N\ge p>1\), \(q>p-1\) and \(\Omega \subset \mathbb {R}^N\) (\(N>1\)) be a \(C^2\) bounded domain such that \(0\in \partial \Omega \). In this article we study the boundary behavior at 0 of nonnegative functions \(u\in C^1(\Omega )\cap C(\overline{\Omega }{\setminus }\{0\})\) which satisfy

$$\begin{aligned} \left. \begin{array}{ll} -\Delta _p u+|\nabla u|^q=0&\quad \text {in }\Omega , \end{array}\right. \end{aligned}$$
(1.1)

where \(\Delta _pu:=\mathrm{div}(|\nabla u|^{p-2}\nabla u)\).where \(\Delta _pu:=\mathrm{div}(|\nabla u|^{p-2}\nabla u)\). The two main questions we consider are as follows:

Q-1:

Existence of positive solutions of (1.1).

Q-2:

Description of positive solutions with an isolated boundary singularity at 0.

When \(p=2\) a fairly complete description of positive solutions of

$$\begin{aligned} \left. \begin{array}{ll} -\Delta u+|\nabla u|^q=0\\ \end{array} \right. \end{aligned}$$
(1.2)

in \(\Omega \) is provided by Nguyen-Phuoc and Véron [11]. In particular they prove the following series of results in the range of values \(1<q<2\).

  1. 1.

    Any signed solution of (1.3) verifies the estimates

    $$\begin{aligned} |\nabla u(x)|\le c_{N,q}(d(x))^{-\frac{1}{q-1}}\qquad \forall x\in \Omega , \end{aligned}$$
    (1.3)

    where \(d(x)=\text{ dist }\,(x,\partial \Omega )\). As a consequence, if \(u\in C(\overline{\Omega }{\setminus }\{0\})\) is a solution which vanishes on \(\partial \Omega {\setminus }\{0\}\), it satisfies

    $$\begin{aligned} |u(x)|\le c_{q,\Omega }d(x)|x|^{-\frac{1}{q-1}}\quad \forall x\in \Omega . \end{aligned}$$
    (1.4)
  2. 2.

    If \(\frac{N+1}{N}\le q<2\) any positive solution of (1.3) in \(\Omega \) which vanishes on \(\partial \Omega {\setminus }\{0\}\) is identically 0. An isolated boundary point is a removable singularity for (1.2).

  3. 3.

    If \(1<q<\frac{N+1}{N}\) and \(k>0\) there exists a unique positive solution \(u:=u_{k}\) of (1.2) in \(\Omega \) which vanishes on \(\partial \Omega {\setminus }\{0\}\) and satisfies \(u(x)\sim c_N kP^\Omega (x,0)\) as \(x\rightarrow 0\), where \(P^\Omega \) is the Poisson kernel in \(\Omega \times \partial \Omega \).

  4. 4.

    If \(1<q<\frac{N+1}{N}\) there exists a unique positive solution u of (1.2) in the half-space \(\mathbb {R}^N_+:=\{x=(x',x_N):x'\in \mathbb {R}^{N-1}, x_N>0\}\) under the form \(u(x)=|x|^{-\frac{2-q}{q-1}}\omega (|x|^{-1}x)\) which vanishes on \( \partial \mathbb {R}^N_+{\setminus }\{0\}\). The function \(\omega \) is the unique positive solution of

    $$\begin{aligned}&\displaystyle -\Delta ' \omega +\left( \left( \frac{2-q}{q-1}\right) ^2\omega ^2+|\nabla '\omega |^2\right) ^{\frac{q}{2}}-\lambda _{N,q}\omega =0\quad \text {in }S^{N-1}_+,\nonumber \\&\displaystyle \omega =0\quad \text {in }\partial S^{N-1}_+, \end{aligned}$$
    (1.5)

    where \(S^{N-1}\) is the unit sphere of \(\mathbb {R}^N\), \(\partial S^{N-1}_+=\partial \mathbb {R}^N_+\cap S^{N-1}\), \(\Delta '\) the Laplace–Beltrami operator and \(\lambda _{N,q}>0\) an explicit constant.

  5. 5.

    If \(1<q<\frac{N+1}{N}\) and u is a positive solution of (1.3) in \(\Omega \), which is continuous in \(\overline{\Omega }{\setminus }\{0\}\) and vanishes on \(\partial \Omega {\setminus }\{0\}\) the following dichotomy occurs:

  1. (i)

    either \(u(x)\sim |x|^{-\frac{2-q}{q-1}}\omega (|x|^{-1}x)\) as \(x\rightarrow 0\),

  2. (ii)

    or \(u(x)\sim kc_NP^{\Omega }(x,0)\) as \(x\rightarrow 0\) for some \(k\ge 0\).

The aim of this article is to extend to the quasilinear case \(1<p\le N\) the above mentioned results. The following pointwise gradient estimate valid for any signed solution u of (1.1) has been proved in [3]: if \(0<p-1<q\) there exists a constant \(c_{N,p,q}>0\) such that

$$\begin{aligned} \left| \nabla u(x)\right| \le c_{N,p,q}(d(x))^{-\frac{1}{q+1-p}}\quad \forall x\in \Omega . \end{aligned}$$
(1.6)

As a consequence, any solution \(u\in C^1(\overline{\Omega }{\setminus }\{0\}\) satisfies

$$\begin{aligned} \left| u(x)\right| \le c_{p,q,\Omega }d(x)\left| x\right| ^{-\frac{1}{q+1-p}}\quad \forall x\in \Omega . \end{aligned}$$
(1.7)

Concerning boundary singularities, the situation is much more complicated than in the case \(p=2\) and the threshold of critical exponent less explicit. We first consider the problem in \(\mathbb {R}^N_+\). Assuming \(p-1<q\le p\), separable solutions of (1.1) in \(\mathbb {R}^N_+\) vanishing on \(\partial \mathbb {R}^N_+{\setminus }\{0\}\) can be looked for in spherical coordinates \((r,\sigma )\in \mathbb {R}^*_+\times S^{N-1}\) (we denote \(\mathbb {R}^*_+=(0,\infty )\)) under the form

$$\begin{aligned} u(x)=u(r,\sigma )=r^{- \beta _q}\omega (\sigma ),\quad r>0,\;\sigma \in S^{N-1}_+:=\{S^{N-1}\cap \mathbb {R}^N_+\}. \end{aligned}$$
(1.8)

Then \(\omega \) is solution of the following problem

$$\begin{aligned}&\displaystyle -div'\left( \left( \beta _q^2\omega ^2+|\nabla ' \omega |^2\right) ^{\frac{p-2}{2}}\nabla ' \omega \right) -\beta _q\Lambda _{\beta _q}\left( \beta _q^2\omega ^2+|\nabla ' \omega |^2\right) ^{\frac{p-2}{2}}\omega \nonumber \\&\displaystyle +\left( \beta _q^2\omega ^2+|\nabla ' \omega |^2\right) ^{\frac{q}{2}}=0\quad \text{ in } S_+^{N-1}\nonumber \\&\displaystyle \omega =0\quad \text{ on } \partial S_+^{N-1}, \end{aligned}$$
(1.9)

where

$$\begin{aligned} \beta _q=\frac{p-q}{q+1-p}\quad \text {and}\quad \Lambda _{\beta _q}=\beta _q(p-1)+p-N, \end{aligned}$$
(1.10)

and \(\nabla '\) is the covariant derivative on \(S^{N-1}\) identified to the tangential gradient thanks to the canonical isometrical imbedding of \(S^{N-1}\) into \(\mathbb {R}^N\), and \(div'\) the divergence operator acting on vector fields on \(S^{N-1}\).

The existence of a positive solution to this problem cannot be separated from the problem of existence of separable p -harmonic functions which are p-harmonic in \(\mathbb {R}^N_+\) which vanish on \(\partial \mathbb {R}^N_+{\setminus }\{0\}\) and have the form \(\Psi (x)=\Psi (r,\sigma )=r^{-\beta }\psi (\sigma )\) for some real number \(\beta \). Necessarily such a \(\psi \) must satisfy

$$\begin{aligned}&\displaystyle -div'\left( \left( \beta ^2\psi ^2+|\nabla ' \psi |^2\right) ^{\frac{p-2}{2}}\nabla ' \psi \right) -\beta \Lambda _{\beta }\left( \beta ^2\psi ^2+|\nabla ' \psi |^2\right) ^{\frac{p-2}{2}}\psi =0\quad \text{ in } S_+^{N-1}\nonumber \\&\displaystyle \psi =0\quad \text{ on } \partial S_+^{N-1}, \end{aligned}$$
(1.11)

where \(\Lambda _{\beta }=\beta (p-1)+p-N\). We will refer to (1.11) as the spherical p-harmonic eigenvalue problem. The study of this problem has been initiated in the 2-dim case by Krol [8] (\(\beta <0\)) and Kichenassamy and Véron [9] (\(\beta >0\)). In this case \(\omega \) satisfies a completely integrable second order differential equation. In the case where \(S^{N-1}_+\) is replaced by a smooth domain \(S\subset S^{N-1}\) with \(N\ge 3\), Tolksdorf [14] proved the existence of a unique couple \((\tilde{\beta }_s,\tilde{\psi }_s)\) where \(\tilde{\beta }_s<0\) and \(\tilde{\psi }_s\) has constant sign and is defined up to an homothety. Recently Porretta and Véron [12] gave a simpler and more general proof of the existence of two couples \((\tilde{\beta }_s,\tilde{\psi }_s)\) and \((\beta _{*\,s},\psi _{*\,s})\) where \(\beta _{*\,s}>0\) and \(\tilde{\psi }_s\) and \(\psi _{*\,s}\) are positive solutions of (1.11) with \(\beta =\tilde{\beta }_s\) and \(\beta =\beta _{*\,s}\) respectively and are unique up to a multiplication by a real number. When \(p=2\) this problem is an eigenvalue problem for the Laplace–Beltrami operator on a subdomain of \(S^{N-1}\). If \(S=S^{N-1}_+\), \(\tilde{\beta }_s\) and \(\beta _{*\,s}\) are respectively denoted by \(\tilde{\beta }\) and \(\beta _*\) and accordingly \(\tilde{\psi }_s\) and \(\psi _{*\,s}\) by \(\tilde{\psi }\) and \(\psi _*\). Since \(x\mapsto x_N\) is p-harmonic, \(\tilde{\beta }=-1\). Except in the cases \(N=2\) where it is the positive root of some algebraic equation of degree 2, \(p=2\) where it is \(N-1\) and \(p=N\) where it is 1, the value of \(\beta _*\) is unknown besides the straightforward estimate \(\beta _*\ge \max \{1,\frac{N-p}{p-1}\}\). Using the fact that \(\psi _*\) depends only on the azimuthal variable and satisfies a differential equation, we prove in Appendix B the following new estimate:

Theorem A

Let \(1<p\le N\).

  1. (i)

    If \(2\le p\le N\), then \(\beta _*\le \frac{N-1}{p-1}\) with equality only if \(p=2\) or N.

  2. (ii)

    If \(1\le p<2\), then \(\beta _*> \frac{N-1}{p-1}\).

The p-harmonic function \(\Psi _*(x)=\Psi _*(r,\sigma )=r^{-\beta _*}\psi _*(\sigma )\) endows the role of a Poisson kernel. To this exponent \(\beta _*\) is associated the critical value \(q_*\) of q defined by \(\beta _*=\beta _q\), or equivalently

$$\begin{aligned} q_*:=\frac{\beta _*(p-1)+p}{\beta _*+1}=p-\frac{\beta _*}{\beta _*+1}. \end{aligned}$$
(1.12)

The following result characterizes strong singularities.

Theorem B

Let \(0<p-1\le N\), then

  1. (i)

    If \(p-1<q<q_*\) problem (1.9) admits a unique positive solution \(\omega _*\).

  2. (ii)

    If \(q_*\le q <p\) problem (1.9) admits no positive solution.

This critical exponent corresponds to the threshold of criticality for boundary isolated singularities.

Theorem C

Assume \(q_*\le q <p\le N\). If \(u\in C^1(\overline{\Omega }{\setminus }\{0\})\) is a nonnegative solution of (1.1) in \(\Omega \) which vanishes on \(\partial \Omega {\setminus }\{0\}\), it is identical zero.

As in the case \(p=2\), there exist positive solutions (1.1) in \(\Omega \) with weak boundary singularities which are characterized by their blow-up near the singularity. By opposition to the case \(p=2\) where existence is obtained by use of a weak formulation of the boundary value problem, combined with uniform integrability of the absorption term thanks to Poisson kernel estimates (see [11]), this approach cannot be performed in the case \(p\ne 2\); the obtention of solutions with weak singularities necessitates a very long and delicate construction of subsolutions and supersolutions. Furthermore, when \(p\ne N\), the construction is done only if \(\Omega \) is locally an hyperplane near 0.

In the sequel we denote by \(B_R(a)\) the open ball of center a and radius \(R>0\) and \(B_R=B_R(0)\). We also set \(B^+_R(a):=\mathbb {R}^N_+\cap B_R(a)\), \(B^+_R:=\mathbb {R}^N_+\cap B_R\), \(B^-_R(a):=\mathbb {R}^N_-\cap B_R(a)\) and \(B^-_R:=\mathbb {R}^N_-\cap B_R\), where \(\mathbb {R}^N_-:=\{x=(x',x_{N}):x'\in \mathbb {R}^{N-1}, x_N<0\}\). If \(\Omega \) is an open domain and \(R>0\), we put \(\Omega _{R}=\Omega \cap B_{R}\).

Theorem D

Let \(\Omega \subset \mathbb {R}^N_+ \) be a bounded domain such that \(0\in \partial \Omega \). Assume there exists \(\delta >0\) such that \(\Omega _\delta =B^+_\delta \) and \(0<p-1<q<q_* <p\le N\). Then for any \(k>0\) there exists a unique \(u:=u_k\in C^1(\overline{\Omega }{\setminus }\{0\})\), solution of (1.1) in \(\Omega \), vanishing on \(\partial \Omega {\setminus }\{0\}\) and such that

$$\begin{aligned} \lim _{\begin{array}{c} {x\rightarrow 0}\\ \frac{x}{|x|}\rightarrow \sigma \in S^{N-1}_+ \end{array}}\!\!\!\!|x|^{\beta _*}u_k(x)=k\psi _*(\sigma ). \end{aligned}$$
(1.13)

Furthermore \(\lim _{k\rightarrow \infty }u_{k}=u_\infty \) and

$$\begin{aligned} \lim _{\begin{array}{c} {x\rightarrow 0}\\ \frac{x}{|x|}\rightarrow \sigma \in S^{N-1}_+ \end{array}}\!\!\!\!|x|^{\beta _q}u_\infty (x)=\psi _*(\sigma ). \end{aligned}$$
(1.14)

When \(p=N\), then \(q_*=N-\frac{1}{2}\); in such a range of values we use the conformal invariance of \(\Delta _N\) and prove that the previous result holds if \(\Omega \) is any \(C^2\) domain. Finally, the isolated singularities of positive solutions of (1.1) are completely described by the two types of singular solutions obtained in the previous theorem and we prove:

Theorem E

Let \(\Omega \) be a bounded domain such that \(0\in \partial \Omega \). Assume there exists \(\delta >0\) such that \(\Omega _\delta =B^+_\delta \) and \(0<p-1<q<q_* <p\le N\). If \(u\in C^1(\overline{\Omega }{\setminus }\{0\})\) is a positive solution of (1.1) in \(\Omega \) which vanishes on \(\partial \Omega {\setminus }\{0\}\), then

  1. (i)

    either there exists \(k\ge 0\) such that

    $$\begin{aligned} \lim _{\begin{array}{c} {x\rightarrow 0}\\ \frac{x}{|x|}\rightarrow \sigma \in S^{N-1}_+ \end{array}}\!\!\!\!|x|^{\beta _*}u(x)=k\psi _*(\sigma ); \end{aligned}$$
    (1.15)
  2. (ii)

    or

    $$\begin{aligned} \lim _{\begin{array}{c} {x\rightarrow 0}\\ \frac{x}{|x|}\rightarrow \sigma \in S^{N-1}_+ \end{array}}\!\!\!\!|x|^{\beta _q}u(x)=\psi _*(\sigma ). \end{aligned}$$
    (1.16)

2 A priori estimates

2.1 The gradient estimates and its applications

We recall the following estimate and its consequences which are proved in [3].

Proposition 2.1

Assume \(q>p-1\) and u is a \(C^1\) solution of (1.1) in a domain \(\Omega \). Then

$$\begin{aligned} \left| \nabla u(x)\right| \le c_{N,p,q}(d(x))^{-\frac{1}{q+1-p}}\quad \forall x\in \Omega . \end{aligned}$$
(2.1)

The first application is a pointwise upper bound for solutions with isolated singularities.

Corollary 2.2

Assume \(q>p-1>0\), \(R^*>0\) and \(\Omega \) is a domain containing 0 such that \(d(0)\ge 2R^*\). Then for any \(x\in B_{R^*}{\setminus }\{0\}\), and \(0<R\le R^*\), any \(u\in C^1(\Omega {\setminus }\{0\})\) solution of (1.1) in \((\Omega {\setminus }\{0\})\) satisfies

$$\begin{aligned} \left| u(x)\right| \le c_{N,p,q}\left| |x|^{\frac{q-p}{q+1-p}}-R^{\frac{q-p}{q+1-p}}\right| +\max \{\left| u(z)\right| :\left| z\right| =R\}, \end{aligned}$$
(2.2)

if \(p\ne q\), and

$$\begin{aligned} \left| u(x)\right| \le c_{N,p}\left( \ln R-\ln \left| x\right| \right) +\max \{\left| u(z)\right| :\left| z\right| =R\}, \end{aligned}$$
(2.3)

if \(p=q\).

The second application corresponds to solutions with boundary blow-up. For \(\delta >0\) small enough we set \(\Omega _{\delta }:=\{z\in \Omega :d(z)<\delta \}\).

Corollary 2.3

Assume \(q>p-1>0\), \(\Omega \) is a bounded domain with a \(C^2\) boundary. Then there exists \(\delta _1>0\) which depends only on \(\Omega \) such that any \(u\in C^1(\Omega ) \) solution of (1.1) in \(\Omega \) satisfies

$$\begin{aligned} \left| u(x)\right| \le c_{N,p,q}\left| (d(x))^{\frac{q-p}{q+1-p}}-\delta _1^{{\frac{q-p}{q+1-p}}}\right| +\max \{\left| u(z)\right| :d(z)=\delta _1\}\quad \forall x\in \Omega _{\delta _1} \end{aligned}$$
(2.4)

if \(p\ne q\), and

$$\begin{aligned} \left| u(x)\right| \le c_{N,p,q}\left( \ln \delta _1-\ln d(x)\right) +\max \{\left| u(z)\right| :d(z)=\delta _1\}\quad \forall x\in \Omega _{\delta _1} \end{aligned}$$
(2.5)

if \(p=q\).

Remark

As a consequence of (2.4) there holds for \(p>q>p-1\)

$$\begin{aligned} u(x)\le \left( c_{N,p,q}+K\max \{\left| u(z)\right| :d(z)\ge \delta _1\}\right) (d(x))^{\frac{q-p}{q+1-p}}\quad \forall x\in \Omega \end{aligned}$$
(2.6)

where \(K=(\mathrm{diam} (\Omega ))^{\frac{p-q}{q+1-p}}\), with the standard modification if \(p=q\).

As a variant of Corollary 2.3 the following upper estimate of solutions in an exterior domain will be used in the sequel.

Corollary 2.4

Assume \(q>p-1>0\), \(R>0\) and \(u\in C^1(B_{R_0}^c)\) is any solution of (1.1) in \(B_{R_0}^c\). Then for any \(R>R_0\) there holds

$$\begin{aligned} \left| u(x)\right| \le c_{N,p,q}\left| (\left| x\right| -R_0)^{\frac{q-p}{q+1-p}}-(R-R_0)^{\frac{q-p}{q+1-p}}\right| +\max \{\left| u(z)\right| :\left| z\right| =R\}\quad \forall x\in B_{R}^c\nonumber \\ \end{aligned}$$
(2.7)

if \(p\ne q\) and

$$\begin{aligned} \left| u(x)\right| \le c_{N,p,q}\left( \ln (\left| x\right| -R_0)-\ln (R-R_0)\right) +\max \{\left| u(z)\right| :\left| z\right| =R\}\quad \forall x\in B_{R}^c\quad \end{aligned}$$
(2.8)

if \(p=q\).

Proof

The proof is a consequence of the identity

$$\begin{aligned} u(x)=u(z)+{\int _{0}^{1}}{\frac{d}{dt} }u(tx+(1-t)z)dt={\int _{0}^{1}}\langle \nabla u(tx+(1-t)z),x-z\rangle dt \end{aligned}$$

where \(z=\frac{R}{\left| x\right| }x\). Since by (2.1)

$$\begin{aligned} \left| \nabla u(tx+(1-t)z)\right| \le C_{N,p,q}(t\left| x\right| +(1-t)R-R_0)^{-\frac{1}{q+1-p}}, \end{aligned}$$

Equations (2.7) and (2.8) follow by integration. \(\square \)

2.2 Boundary a priori estimates

The next result is the extension to isolated boundary singularities of a previous regularity estimate dealing with singularity in a domain proved in [3, Lemma 3.10].

Lemma 2.5

Assume \(p-1<q<p\), \(\Omega \) is a bounded \(C^2\) domain such that \(0\in \partial \Omega \). Let \(u\in C^1(\overline{\Omega }{\setminus }\{0\})\) be a solution of (1.1) in \(\Omega \) which vanishes on \(\partial \Omega {\setminus }\{0\}\) and satisfies

$$\begin{aligned} \left| u(x)\right| \le \phi (\left| x\right| )\quad \forall x\in \Omega , \end{aligned}$$
(2.9)

where \(\phi :\mathbb {R}^*_+\mapsto \mathbb {R}_+\) is continuous, nonincreasing and satisfies

$$\begin{aligned} \phi (rs)\le \gamma \phi (r)\phi (s)\quad \text {and}\quad r^{\frac{p-q}{q+1-p}}\phi (r)\le c, \end{aligned}$$
(2.10)

for some \(\gamma ,c>0\) and any \(r,s>0\). There exist \(\alpha \in (0,1)\) and \( c_1=c_1(p,q,\Omega )>0\) such that

$$\begin{aligned} \begin{array}{lll} (i)&{}\left| \nabla u(x)\right| \le c_1\phi (\left| x\right| )\left| x\right| ^{-1}&{}\quad \forall x\in \Omega ,\\ (ii)&{}\left| \nabla u(x)-\nabla u(y)\right| \le c_1\phi (\left| x\right| )\left| x\right| ^{-1-\alpha }\left| x-y\right| ^\alpha &{}\quad \forall x,y\in \Omega ,\;\left| x\right| \le \left| y\right| . \end{array} \end{aligned}$$
(2.11)

Furthermore

$$\begin{aligned} u(x)\le c_1\phi (\left| x\right| )\frac{d(x)}{\left| x\right| }\quad \forall x\in \Omega . \end{aligned}$$
(2.12)

Proof

For \(\ell >0\), we set \(\Omega ^\ell :=\frac{1}{\ell }\Omega \). If \(\ell \in (0,1]\) the curvature of \(\partial \Omega ^\ell \) remains uniformly bounded. As in [5, p 622], there exists \(0<\delta _0\le 1\) and an involutive diffeomorphism \(\psi \) from \(\overline{B}_{\delta _0}\cap \overline{\Omega }^{\delta _0}\) into \(\overline{B}_{\delta _0}\cap (\Omega ^{\delta _0})^c\) which is the identity on \(\overline{B}_{\delta _0}\cap \partial \Omega ^{\delta _0}\) and such that \(D\psi (\xi )\) is the symmetry with respect to the tangent plane \(T_\xi \partial \Omega \) for any \(\xi \in \partial \Omega \cap \overline{B}_{\delta _0}\). We extend any function v defined in \(\overline{B}_{\delta _0}\cap \overline{\Omega }^{\delta _0}\) and vanishing on \(\overline{B}_{\delta _0}\cap \partial \Omega ^{\delta _0}\) into a function \(\tilde{v}\) defined in \(\overline{B}_{\delta _0}\) by

$$\begin{aligned} \tilde{v}(x)=\left\{ \begin{array} {ll} v(x)&{}\quad \text {if }x\in \overline{B}_{\delta _0}\cap \overline{\Omega }^{\delta _0}\\ -v\circ \psi (x)&{}\quad \text {if }x\in \overline{B}_{\delta _0}\cap (\Omega ^{\delta _0})^c, \end{array}\right. \end{aligned}$$
(2.13)

If \(v\in C^1(\overline{B}_{\delta _0}\cap \overline{\Omega }^{\delta _0})\) is a solution of (1.1) in \(B_{\delta _0}\cap \Omega ^{\delta _0}\) which vanishes on \(\partial \Omega ^{\delta _0}\cap \overline{B}_{\delta _0}\), \(\tilde{v}\) satisfies

$$\begin{aligned} -\sum _j{\frac{\partial }{\partial x_j} }\tilde{A}_j(x,\nabla \tilde{v})+B(x,\nabla \tilde{v})=0\quad \text {in }B_{\delta _0}. \end{aligned}$$
(2.14)

As in [5, (2.37)] the \(A_j\) and B satisfy the following estimates

$$\begin{aligned} \begin{array}{ll} (i) &{}\quad \tilde{A}_j(x,0)=0\\ (ii)&{}\quad \displaystyle \sum _{i,j}{\frac{\partial }{\partial \eta _i} }\tilde{A}_j(x,\eta )\xi _i\xi _j\ge C_1\left| \eta \right| ^{p-1}\left| \xi \right| ^2\\ (iii)&{}\quad \displaystyle \sum _{i,j}\left| {\frac{\partial }{\partial \eta _j} }\tilde{A}_j(x,\eta )\right| \le C_2\left| \eta \right| ^{p-2}, \end{array} \end{aligned}$$
(2.15)

and

$$\begin{aligned} \left| B(x,\eta )\right| \le C_3(1+\left| \eta \right| )^p, \end{aligned}$$
(2.16)

where the \(C_j\) are positive constants. These estimates are the ones needed to apply Tolksdorf’s result [15, Th1, 2]. There exists a constant C, such that for any ball \(\overline{B}_{3R}\subset \overline{B}_{\delta _0}\), there holds

$$\begin{aligned} \left\| \nabla \tilde{v}\right\| _{L^\infty (B_R)}\le C, \end{aligned}$$
(2.17)

where C depends on the constants \(C_k\) (\(k=1,2,3\)), N, p and \(\left\| \tilde{v}\right\| _{L^\infty (B_{3R})}\). We define

$$\begin{aligned} \Phi _\ell [u](y):=u_\ell =\frac{1}{\phi (\ell )} u(\ell y)\quad \forall y\in \Omega ^\ell . \end{aligned}$$
(2.18)

Then

$$\begin{aligned} |u_\ell (y)|\le \frac{\phi (\ell \left| y\right| )}{\phi (\ell )}\le \gamma \phi (\left| y\right| )\quad \forall y\in \Omega ^\ell \end{aligned}$$
(2.19)

and

$$\begin{aligned} -\Delta _pu_\ell +(\ell ^{\beta _q}\phi (\ell ))^{q+1-p}\left| \nabla u_\ell \right| ^q=0\quad \text {in }\Omega ^\ell . \end{aligned}$$
(2.20)

Using formula (2.13) we extend \(u_\ell \) into a function \(\tilde{u}_\ell \) which satisfies

$$\begin{aligned} -\sum _j{\frac{\partial }{\partial y_j} }\tilde{A}_j(y,\nabla \tilde{u}_\ell )+(\ell ^{\beta _q}\phi (\ell ))^{q+1-p}B(y,\nabla \tilde{u}_\ell )=0\quad \text {in }B_{\delta _0}. \end{aligned}$$
(2.21)

For \(0<\left| x\right| <\delta _0\) there exists \(\ell \in (0,2)\) such that \(\frac{\delta _0\ell }{2}\le \left| x\right| \le \delta _0\ell \). Then \(y\mapsto \tilde{u}_\ell (y)\) with \(y=\frac{x}{\ell }\) satisfies (2.21) in \(B_{\delta _0}\) and \(|\tilde{u}_\ell (y)|\le \gamma _*\phi (\left| y\right| )\) since \(\psi \) is a diffeomorphism and \(D\psi (\xi )\in O(N)\) for any \(\xi \in \partial \Omega \cap B_{\delta _0}\). The function \(\tilde{u}_\ell \) remains bounded on any ball \(B_{3R}(z)\subset \Gamma :=\{y\in \mathbb {R}^N:\frac{\delta _0}{2}<\left| y\right| <\delta _0\}\), therefore \(\left| \nabla \tilde{u}_\ell (y)\right| \le c\) for any \(y\in B_{R}(z)\), for some constant \(c>0\). This implies

$$\begin{aligned} \begin{array}{ll} \left| \nabla u(x)\right| \le c\gamma _*\delta _0\phi \left( \frac{2}{\delta _0}\right) \phi (|x|)|x|^{-1}\quad \forall x\in \Omega \cap B_{\delta _0}, \end{array} \end{aligned}$$
(2.22)

which is (2.11)-(i). Moreover, by standard regularity estimates [10], there exists \(\alpha \in (0,1)\) such that \(\left| \nabla \tilde{u}_\ell (y)-\nabla \tilde{u}_\ell (y')\right| \le c\left| y-y'\right| ^\alpha \) for all y and \(y'\) belonging to \(B_{R}(z)\). This implies (2.11)-(ii).

Next we prove (2.12). Let \(0<\delta _1\le \delta _0\) such that at any boundary point z there exist two closed balls of radius \(\delta _1\) tangent to \(\partial \Omega \) at z and which are included in \(\Omega \cup \{z\}\) and in \(\overline{\Omega }^c\cup \{z\}\) respectively (\(\delta _1\) corresponds to the maximal radius of the interior and exterior sphere condition). Let \(x\in \Omega \) such that \(d(x)\le \delta _1\) (this is not a loss of generality) and \(z_x\) be the projection of x on \(\partial \Omega \). We first assume that x does not belong to the cone \(\Sigma _{\frac{\pi }{4}}\) with vertex 0, axis \(-\mathbf{n}_0\), where \(\mathbf{n}_0\) is the normal outward unit vector at 0, and angle \(\frac{\pi }{4}\). Consider the path \(\zeta \) from \(z_x\) to x defined by \(\zeta (t)=tx+(1-t)z_x\) with \(0\le t\le 1\). Then

$$\begin{aligned} u(x)={\int _{0}^{1}}{\frac{d}{dt} }u\circ \zeta (t)dt={\int _{0}^{1}}\langle \nabla u\circ \zeta (t),x-z_x\rangle dt \end{aligned}$$
(2.23)

Thus, by the Cauchy–Schwarz inequality, using (2.11),

$$\begin{aligned} \left| u(x)\right| \le c_1d(x){\int _{0}^{1}}{\frac{\phi (|\zeta (t)|)}{\left| \zeta (t)\right| } }dt. \end{aligned}$$
(2.24)

Since \(x\notin \Sigma _{\frac{\pi }{4}}\), \(\zeta (t)\notin \Sigma _{\frac{\pi }{4}}\) and there exists \(c_2>0\) depending on \(\Omega \) such that \(c^{-1}_2\left| x\right| \le \left| \zeta (t)\right| \le c_2\left| x\right| \) for all \(0\le t\le 1\). Therefore \(\phi (|\zeta (t)|)\le \phi (c_2\left| x\right| )\le \gamma \phi (c_2)\phi (\left| x\right| )\) by (2.10). This implies

$$\begin{aligned} \left| u(x)\right| \le \gamma c_1c_2\phi (c_2){\frac{d(x)\phi (\left| x\right| )}{\left| x\right| } } \end{aligned}$$
(2.25)

by (2.12) whenever \(x\notin \Sigma _{\frac{\pi }{4}}\). When \(x\in \Sigma _{\frac{\pi }{4}}\) then \(d(x)\le \left| x\right| \le c_3d(x)\) where \(c_3>0\) depends on the curvature of \(\partial \Omega \). Then (2.9) combined with (2.10) implies the claim. \(\square \)

Lemma 2.6

Assume \(p-1<q\le p\), \(\Omega \) is a bounded \(C^2\) domain such that \(0\in \partial \Omega \) and \(R_0=\max \{ \left| z\right| :z\in \Omega \}\). If \(u\in C(\overline{\Omega }{\setminus }\{0\})\cap C^1(\Omega )\) is a positive solution of (1.1) which vanishes on \(\partial \Omega {\setminus }\{0\}\), it satisfies

$$\begin{aligned} u(x)\le \left\{ \begin{array}{ll} c_2\left( \left| x\right| ^{\frac{q-p}{q+1-p}}-R_0^{\frac{q-p}{q+1-p}}\right) &{}\quad \text {if }q< p\\ (p-1)\ln \left( \frac{R_0}{\left| x\right| }\right) &{}\quad \text {if }q= p \end{array}\right. \end{aligned}$$
(2.26)

for all \( x\in \Omega \), where \(c_2=c_2(p,q)>0\).

Proof

For \(\epsilon >0\) we denote by \(P_\epsilon :\mathbb {R}\mapsto \mathbb {R}_+\) the function defined by

$$\begin{aligned} P_\epsilon (r)=\left\{ \begin{array}{ll}0&{}\quad \text {if }0\le r\le \epsilon \\ -\frac{r^4}{2\epsilon ^3}+\frac{3r^3}{\epsilon ^2}-\frac{6r^2}{\epsilon }+5r-\frac{3\epsilon }{2}&{}\quad \text {if }\epsilon <r<2 \epsilon \\ r-\frac{3\epsilon }{2}&{}\quad \text {if }r\ge 2\epsilon , \end{array}\right. \end{aligned}$$
(2.27)

and by \(u_\epsilon \) the extension of \(P_\epsilon (u)\) by zero outside \(\Omega \). There exists \(R_0\) such that \(\Omega \subset B_{R_0}\). Since \(0\le P_\epsilon (r)\le |r|\) and \(P_\epsilon \) is convex, \(u_\epsilon \in C(\mathbb {R}^N{\setminus }\{0\})\cap W^{1,p}_{loc}(\mathbb {R}^N{\setminus }\{0\})\) and

$$\begin{aligned} -\Delta _pu_\epsilon +\left| \nabla u_\epsilon \right| ^q\le 0\qquad \text {in }\mathbb {R}^N. \end{aligned}$$

Let \(R>R_0\). If \(p-1<q<p\)

$$\begin{aligned} U_{\epsilon ,R}(\left| x\right| )=c_2 \left( (|x|-\epsilon )^{\frac{q-p}{q+1-p}}-(R-\epsilon )^{\frac{q-p}{q+1-p}}\right) \quad \text {in }B_R{\setminus } B_\epsilon , \end{aligned}$$
(2.28)

with \(c_2=(p-q)^{-1}(q+p-1)^{\frac{q-p}{q+1-p}}\). Then \(-\Delta _pU_{\epsilon ,R}+\left| \nabla U_{\epsilon ,R}\right| ^q\ge 0\). Since \(u_\epsilon \) vanishes on \(\partial B_R\) and is finite on \(\partial B_\epsilon \), it follows \(u_\epsilon \le U_{\epsilon ,R}\). Letting successively \(\epsilon \rightarrow 0\) and \(R\rightarrow R_0\) yields to (2.26). If \(q=p\) we take

$$\begin{aligned} U_{\epsilon ,R}(\left| x\right| )=(p-1) \ln \left( \frac{R-\epsilon }{\left| x\right| -\epsilon }\right) \quad \text {in }B_R{\setminus } B_\epsilon , \end{aligned}$$
(2.29)

which turns out to be a supersolution of (1.1); the end of the proof is similar.

As a consequence of Lemmas 2.5 and 2.6, we obtain. \(\square \)

Corollary 2.7

Let pq \(\Omega \) and u be as in Lemma 2.6. Then there exists a constant \(c_3=c_3(p,q,\Omega )>0\) such that

$$\begin{aligned} \left| \nabla u(x)\right| \le c_3\left| x\right| ^{-\frac{1}{q+1-p}}\quad \forall x\in \Omega \end{aligned}$$
(2.30)

and

$$\begin{aligned} u(x)\le c_3d(x)\left| x\right| ^{-\frac{1}{q+1-p}}\quad \forall x\in \overline{\Omega }{\setminus }\{0\}. \end{aligned}$$
(2.31)

Remark

If \(\Omega \) is locally flat near 0, then estimates (2.30) and (2.31) are valid without any sign assumption on u. More precisely, if \(\partial \Omega \cap B_{\delta _0}=T_0\partial \Omega \cap B_{\delta _0}\) we can perform the reflection of u through the tangent plane \(T_0\partial \Omega \) to \(\partial \Omega \) at 0 and the new function \(\tilde{u}\) is a solution of (1.1) in \(B_{\delta _0}{\setminus }\{0\}\). By Proposition 2.1, it satisfies

$$\begin{aligned} \left| \nabla \tilde{u}(x)\right| \le c_{N,p,q}\left| x\right| ^{-\frac{1}{q+1-p}}\quad \forall x\in B_{\frac{\delta _0}{2}}{\setminus }\{0\}. \end{aligned}$$
(2.32)

Integrating this relation as in [3], we derive that for any \(x\in B_{\frac{\delta _0}{2}}\cap \Omega \), there holds

$$\begin{aligned} \left| u(x)\right| \le \left\{ \begin{array}{ll} c_{N,p,q}\left( |x|^{-\beta _q}-\left( \frac{\delta _0}{2}\right) ^{-\beta _q}\right) +\max \{\left| u(z)\right| :\left| z\right| =\frac{\delta _0}{2}\}&{}\quad \text {if } p\ne q\\ c_{N,p}\ln \left( \frac{\delta _0}{2\left| x\right| }\right) +\max \{\left| u(z)\right| :\left| z\right| =\frac{\delta _0}{2}\}&{}\quad \text {if } p= q. \end{array}\right. \end{aligned}$$
(2.33)

In the next result we allow the boundary singular set to be a compact set.

Proposition 2.8

Let \(p-1<q<p\) and \(\delta _1\) as above. There exist \( r^*\in (0, \delta _1]\) and \(c_4=c_4(N,p,q)>0\) such that for any nonempty compact set \(K\subset \partial \Omega \), \(K\ne \partial \Omega \) and any positive solution \(u\in C(\overline{\Omega }{\setminus } K)\cap C^1(\Omega )\) of (1.1) which vanishes on \(\partial \Omega {\setminus } K\), there holds

$$\begin{aligned} u(x)\le c_4d(x)(d_K(x))^{-\frac{1}{q+1-p}}\quad \forall x\in \partial \Omega \;\text {s.t.}\;d(x)\le r^*, \end{aligned}$$
(2.34)

where \(d_K(x)=\text{ dist }\,(x,K)\).

Proof

Step 1: tangential estimates Let \(x\in \Omega \) such that \(d(x)\le \delta _1\). We denote by \(\sigma (x)\) the projection of x onto \(\partial \Omega \), unique since \(d(x)\le \delta _1\). Let \(r\,,r',\tau >0\) such that \(\frac{3}{4}r<r'<\frac{7}{8}r\) and \(0<\tau \le \frac{r'}{2}\) and put \(\omega _{\tau ,x}=\sigma (x)+\tau \mathbf{n}_{\sigma (x)}\). Since \(\partial \Omega \) is \(C^2\), there exists \(0< r^*\le \delta _1\) depending on \(\Omega \) such that \(d_K(\omega _{\tau ,x})>\frac{7}{8}r\) whenever \(d(x)\le r^*\). Let \(a>0\) and \(b>0\) to be specified later on; we define \(\tilde{v}(s)=a(r'-s)^{\frac{q-p}{q+1-p}}-b\) and \(v(y)=\tilde{v}(\left| y-\omega _{\tau ,x}\right| )\) in \([0,r')\) and \(B_{r'}(\omega _{\tau ,x})\) respectively. Then

$$\begin{aligned} \left| \tilde{v}'\right| ^{p-2}\left( \left| \tilde{v}'\right| ^{q+2-p}-(p-1)\tilde{v}''-{\frac{N-1}{s} }\tilde{v}'\right) \!=\! a^{p-1}\left( {\frac{p-q}{q+1-p} }\right) ^{p-1} (r'-s)^{-\frac{q}{q+1-p}}X(s) \end{aligned}$$

where

$$\begin{aligned} X(s)=\left( a{\frac{p-q}{q+1-p} }\right) ^{q+1-p}-{\frac{p-1}{q+1-p} }-{\frac{(N-1)(r'-s)}{s} }. \end{aligned}$$

For any \(\tau \in (0,r')\) there exists \(a>0\) such that

$$\begin{aligned} \left( a{\frac{p-q}{q+1-p} }\right) ^{q+1-p}\ge {\frac{p-1}{q+1-p} }+{\frac{(N-1)(r'-s)}{s} }\quad \forall \tau \le s\le r'. \end{aligned}$$

This implies

$$\begin{aligned} -\Delta _p v+\left| \nabla v\right| ^q\ge 0\quad \text {in }B_{r'}(\omega _{\tau ,x}){\setminus } B_{\tau }(\omega _{\tau ,x}). \end{aligned}$$
(2.35)

Next we take \(b=a(r'-\tau )^{\frac{q-p}{q+1-p}}\), thus \(v=0\) on \(\partial B_{\tau }(\omega _{\tau ,x})\). Clearly \(B_{\tau }(\omega _{\tau ,x})\subset \overline{\Omega }^c\) since \(\tau <\delta _1\). Therefore \(v\ge 0=u\) on \(\partial \Omega \cap B_{r'}(\omega _{\tau ,x})\) and \(u\le v=\infty \) on \(\Omega \cap \partial B_{r'}(\omega _{\tau ,x})\). By the comparison principle, \(v\ge u\) in \(\Omega \cap B_{r'}(\omega _{\tau ,x}).\) In particular

$$\begin{aligned} u(x)\le v(x)\le a(r'-\tau -d(x))^{\frac{q-p}{q+1-p}}-a(r'-\tau )^{\frac{q-p}{q+1-p}}. \end{aligned}$$

We take now \(\tau =\frac{r'}{2}\) and \(d(x)\le \frac{r}{4}\) and we derive by the mean value theorem

$$\begin{aligned} u(x)\le c'_4r'^{-\frac{1}{q+1-p}}d(x)=c'_4d(x)(d_K(x))^{-\frac{1}{q+1-p}}, \end{aligned}$$
(2.36)

with \(c'_4=c'_4(p,q)>0\) Letting \(r'\rightarrow \frac{7}{8}r\), we get (2.12).

Step 2: global estimates If \(d(x)\ge \frac{1}{4}d_K(x)\), there holds

$$\begin{aligned} d(x)(d_K(x))^{-\frac{1}{q+1-p}}\ge 2^{-\frac{2}{q+1-p}}(d(x))^{\frac{q-p}{q+1-p}}. \end{aligned}$$

Combining this inequality with (2.6) and obtain (2.34). \(\square \)

Remark

Under the assumption of Proposition 2.8, it follows from the maximum principle that u is upper bounded in the set \(\Omega '_{r^*}:=\{x\in \Omega :d(x)>r^*\}=\Omega {\setminus }\overline{\Omega }_{r^*}\) by the solution w of

$$\begin{aligned} -\Delta _pw+|\nabla w|^q= & {} 0\quad \text {in }\Omega _{r^*}\nonumber \\ w= & {} c_4d(x)(d_K(x))^{-\frac{1}{q+1-p}}\quad \text {in }\partial \Omega _{r^*}, \end{aligned}$$
(2.37)

and w itself is bounded by \(d^*=\max \{cd(x)(d_K(x))^{-\frac{1}{q+1-p}}:d(x)=r^*\}\).

Next we prove a boundary Harnack inequality. We recall that \(\delta _1\) has been introduced at Corollary 2.3, and that the interior and exterior sphere conditions hold in the set \(\{x\in \mathbb {R}^N:d (x)\le \delta _1\}\).

Theorem 2.9

Let \(q>p-1\) and \(0\in \partial \Omega \). Then there exists \(c_5=c_5(N,p,q,\Omega )>0\) such that for any positive solution \(u\in C(\Omega \cup ((\partial \Omega {\setminus }\{0\})\cap B_{2\delta _1})\cap C^1(\Omega )\) of (1.1) in \(\Omega \), vanishing on \(\partial \Omega {\setminus }\{0\})\cap B_{2\delta _1}\), there holds

$$\begin{aligned} {\frac{u(y)}{c_5d(y)} }\le {\frac{u(x)}{d(x)} }\le c_5{\frac{u(y)}{d(y)} } \end{aligned}$$
(2.38)

for all \(x,y\in B_{\frac{2\delta _1}{3}}\cap \Omega \) such that \(\frac{1}{2}\left| x\right| \le \left| y\right| \le 2\left| x\right| \).

For proving Theorem 2.9 we need some intermediate lemmas. First we recall the following result from [1].

Lemma 2.10

Assume that \(a \in \partial \Omega \), \(0<r<\delta _1\) and \(h>1\) is an integer. There exists an integer \(N_0\), depending only on \(\delta _1\), such that for any points x and y in \(\Omega \cap B_{\frac{3r}{2}}(a)\) verifying \(\min \{d(x),d(y)\} \ge r/2^h\), there exists a connected chain of balls \(B_1,\ldots ,B_j\) with \(j\le N_0h\) such that

$$\begin{aligned} x \in B_1, y \in B_j, \quad B_i\cap B_{i+1} \ne \emptyset \quad \text {for } 1\le i \le j-1 \nonumber \\ \quad \text {and}\quad 2B_i \subset B_{2r}(Q) \cap \Omega \quad \text {for } 1\le i \le j. \end{aligned}$$
(2.39)

The next result is a standard Harnack inequality.

Lemma 2.11

Assume \(a \in (\partial \Omega \setminus \{0\}) \cap B_{\frac{2\delta _1}{3}}\) and \(0<r\le \left| a\right| /4\). Let \(u\in C(\Omega \cup ((\partial \Omega \setminus \{0\})\cap B_{2\delta _1}))\cap C^1(\Omega )\) be a positive solution of (1.1) vanishing on \((\partial \Omega \setminus \{0\})\cap B_{2\delta _1}\). Then there exists a positive constant \(c_6>1\) depending on Npq and \(\delta _1\) such that

$$\begin{aligned} u(x) \le c_6^h u(y), \end{aligned}$$
(2.40)

for every \(x,y \in B_{\frac{3r}{2}}(a)\cap \Omega \) such that \(\min \{d(x),d(y)\} \ge r/2^h\) for some \(h \in \mathbb N\).

Proof

For \(\ell >0\), we define \(T_{\ell }[u]\) by

$$\begin{aligned} T_\ell [u](x)=\ell ^{\frac{p-q}{q+1-p}}u(\ell x), \end{aligned}$$
(2.41)

and we notice that if u satisfies (1.1) in \(\Omega \), then \(T_\ell [u]\) satisfies the same equation in \(\Omega ^{\ell }:=\ell ^{-1}\Omega \). If we take in particular \(\ell =|a|\), we can assume \(|a|=1\), thus the curvature of the domain \(\Omega ^{|a|}\) remains bounded. By Proposition 2.8

$$\begin{aligned} u(x) \le c'_6\quad \forall x \in B_{2r}(a)\cap \Omega \end{aligned}$$
(2.42)

where \(c'_6\) depends on N, q, \(\delta _1\). Then we proceed as in [11], using Lemma 2.10 and internal Harnack inequality as quoted in [16, Corollary 10]. \(\square \)

Since the solutions are Hölder continuous, the following statement holds as in [16, Theorem 4.2]:

Lemma 2.12

Let the assumptions on a and u of Lemma 2.11 be fulfilled. If \(b \in \partial \Omega \cap B_r(a)\) and \(0<s\le 2^{-1}r\), there exist two positive constants \(\delta \) and \(c_7\) depending on N, p, q and \(\Omega \) such that

$$\begin{aligned} u(x) \le c_7{\frac{\left| x-b\right| ^\delta }{s^\delta } }\max \{u(z): z\in B_r(b)\cap \Omega \} \end{aligned}$$
(2.43)

for every \(x \in B_s(b) \cap \Omega \).

As a consequence we derive the following Carleson type estimate.

Lemma 2.13

Assume \(a \in (\partial \Omega \setminus \{0\}) \cap B_{\frac{2\delta _1}{3}}\) and \(0<r\le \left| a\right| /8\). Let \(u\in C(\Omega \cup ((\partial \Omega \setminus \{0\})\cap B_{2\delta _1}))\cap C^2(\Omega )\) be a positive solution of (1.1) vanishing on \((\partial \Omega \setminus \{0\})\cap B_{2\delta _1}\). Then there exists a constant \(c_8\) depending only on N, p and q such that

$$\begin{aligned} u(x) \le c_8 u\left( a-\frac{r}{2}\mathbf{n}_{_a}\right) \quad \forall x \in B_r(a) \cap \Omega . \end{aligned}$$
(2.44)

Proof

By Lemma 2.11 it is clear that for any integer h and \(x \in B_r(a) \cap \Omega \) such that \(d(x)\ge 2^{-h}r\), there holds

$$\begin{aligned} u(x) \le c_6^h u\left( a-\frac{r}{2}\mathbf{n}_{_a}\right) . \end{aligned}$$
(2.45)

Therefore u satisfies inequality (2.43) as any Hölder continuous function does. The proof that the constant is independent of r and u is more delicate. It is done in [1, Lemma 2.4] for linear equations, but it is based only on Lemma 2.12 and a geometric construction, thus it is also valid in our case. \(\square \)

Lemma 2.14

Assume \(a \in (\partial \Omega \setminus \{0\}) \cap B_{\frac{2\delta _1}{3}}\) and \(0<r\le \left| a\right| /8\). Let \(u\in C(\Omega \cup ((\partial \Omega \setminus \{0\})\cap B_{2\delta _1}))\cap C^2(\Omega )\) be a positive solution of (1.1) vanishing on \((\partial \Omega \setminus \{0\})\cap B_{2\delta _1}\). Then there exist \(\alpha \in (0,1/2)\) and \(c_9>0\) depending on N, p and q such that

$$\begin{aligned} {\frac{1}{c_9} }{\frac{t}{r} } \le {\frac{u(b-t\mathbf{n}_{_b})}{u(a-\frac{r}{2}\mathbf{n}_{_a})} } \le c_9{\frac{t}{r} } \end{aligned}$$
(2.46)

for any \(b \in B_r(a) \cap \partial \Omega \) and \(0 \le t < \frac{\alpha }{2}r\).

Proof

It is similar to the one of [11, Lemma 3.15]. \(\square \)

Proof of Theorem 2.9

Assume \(x \in B_{\frac{2\delta _1}{3}} \cap \Omega \) and set \(r=\frac{\left| x\right| }{8}\).

Step 1: tangential estimate: we suppose \(d(x) < \frac{\alpha }{2}r\). Let \(a \in \partial \Omega \setminus \{0\}\) such that \(\left| a\right| =\left| x\right| \) and \(x \in B_r(a)\). By Lemma 2.14,

$$\begin{aligned} {\frac{8}{c_9} }{\frac{u(a-\frac{r}{2}\mathbf{n}_{_a})}{\left| x\right| } } \le {\frac{u(x)}{d(x)} } \le 8c_9{\frac{u(a-\frac{r}{2}\mathbf{n}_{_a})}{\left| x\right| } }. \end{aligned}$$
(2.47)

We can connect \(a-\frac{r}{2}\mathbf{n}_{_a}\) with \(-2r\mathbf{n}_{_0}\) by \(m_1\) (depending only on N) connected balls \(B_i=B_\frac{r}{4}(x_i)\) with \(x_i \in \Omega \) and \(d(x_i) \ge \frac{r}{2}\) for every \(1 \le i \le m_1\). It follows from (2.44) that

$$\begin{aligned} c^{-m_1}_6u(-2r\mathbf{n}_{_0}) \le u\left( a-\frac{r}{2}\mathbf{n}_{_a}\right) \le c^{m_1}_6u(-2r\mathbf{n}_{_0}), \end{aligned}$$

which, together with (2.47) leads to

$$\begin{aligned} {\frac{1}{c_{10}} }{\frac{u(-2r\mathbf{n}_{_0})}{\left| x\right| } } \le {\frac{u(x)}{d(x)} } \le c_{10}{\frac{u(-2r\mathbf{n}_{_0})}{\left| x\right| } }, \end{aligned}$$
(2.48)

with \(c_{10}=8c_9c^{m_1}_6\).

Step 2: internal estimate: we suppose \(d(x) \ge \frac{\alpha }{2}r\). We can connect \(-2r\mathbf{n}_{_0}\) with x by \(m_2\) (depending only on N) connected balls \(B'_i=B_\frac{\alpha r}{4}(x'_i)\) with \(x'_i \in \Omega \) and \(d(x'_i) \ge \frac{\alpha }{2}r\) for every \(1 \le i \le m_2\). By Harnack and Carleson inequalities (2.40) and (2.44) and since \(\frac{\alpha }{4}\left| x\right| <d(x)\le \left| x\right| \), we get

$$\begin{aligned} {\frac{\alpha }{4c_{6}'^{m_2}} }{\frac{u(-2r\mathbf{n}_{_0})}{\left| x\right| } } \le {\frac{u(x)}{d(x)} } \le {\frac{4c_{6}'^{m_2}}{\alpha } }{\frac{u(-2r\mathbf{n}_{_0})}{\left| x\right| } }. \end{aligned}$$
(2.49)

Step 3: end of proof Suppose \(\frac{\left| x\right| }{2}\le s \le 2\left| x\right| \), we can connect \(-2r\mathbf{n}_{_Q}\) with \(-s\mathbf{n}_{_Q}\) by \(m_3\) (depending only on N) connected balls \(B''_i=B_\frac{r}{2}(x''_i)\) with \(x''_i \in \Omega \) and \(d(x''_i) \ge r\) for every \(1 \le i \le m_3\). This fact, jointly with (2.48) and (2.49), yields to

$$\begin{aligned} {\frac{1}{c_{11}} }{\frac{u(-s\mathbf{n}_{_0})}{\left| x\right| } } \le {\frac{u(x)}{d(x)} } \le c_{11}{\frac{u(-s\mathbf{n}_{_0})}{\left| x\right| } } \end{aligned}$$
(2.50)

where \(c_{11}=c_{11}(N,q,\Omega )\). Finally, if \(y \in B_{\frac{2r_0}{3}} \cap \Omega \) satisfies \(\frac{\left| x\right| }{2} \le \left| y\right| \le 2\left| x\right| \), then by applying twice (2.50) we get (2.38) with \(c_{5}=c_{11}^2\). \(\square \)

The following inequality is a consequence of Theorem 2.9.

Corollary 2.15

Assume \(q>p-1\) and \(0\in \partial \Omega \). Then there exists \(c_{12}>0\) depending on p, q and \(\Omega \) such that for any positive solutions \(u_1,\,u_2\in C(\Omega \cup ((\partial \Omega {\setminus }\{0\})\cap B_{2\delta _1}))\cap C^1(\Omega )\) of (1.1) in \(\Omega \), vanishing on \((\partial \Omega {\setminus }\{0\})\cap B_{2\delta _1}\), there holds

$$\begin{aligned} \sup \left\{ {\frac{u_1(y)}{u_2(y)} }:y\in B_r{\setminus } B_{\frac{r}{2}}\right\} \le c_{12}\inf \left\{ {\frac{u_1(y)}{u_2(y)} }:y\in B_r{\setminus } B_{\frac{r}{2}}\right\} . \end{aligned}$$
(2.51)

3 Boundary singularities

3.1 Strongly singular solutions

In this section we consider the Eq. (1.1) in \(\mathbb {R}_+^N\). We denote by \((r,\sigma ) \in \mathbb {R}_+ \times S^{N-1}\) the spherical coordinates in \(\mathbb {R}^N \) and

$$\begin{aligned} S_+^{N-1}=\left\{ (\sin \phi \sigma ',\cos \phi ):\sigma '\in S^{N-2},\phi \in [0,\frac{\pi }{2})\right\} . \end{aligned}$$

If \(v(x)=r^{-\beta }\omega (\sigma )\) satisfies (1.1) in \(\mathbb {R}_+^N\) and vanishes on \(\partial {\mathbb {R}^N_+}{\setminus }\{0\}\), then \(\beta =\beta _q\) and \(\omega \) is a solution of

$$\begin{aligned}&-div'\left( \left( \beta _q^2\omega ^2+|\nabla ' \omega |^2\right) ^{\frac{p-2}{2}}\nabla ' \omega \right) -\beta _q\Lambda _{\beta _q}\left( \beta _q^2\omega ^2+|\nabla ' \omega |^2\right) ^{\frac{p-2}{2}}\omega \nonumber \\&\quad +\left( \beta _q^2\omega ^2+|\nabla ' \omega |^2\right) ^{\frac{q}{2}}=0\quad \text{ in } S_+^{N-1}\\&\quad \omega =0\quad \text{ on } \partial S_+^{N-1}\nonumber \end{aligned}$$
(3.1)

where \(\beta _q\) and \(\Lambda _{\beta _q}\) have been defined in (1.10). We denote by \((\beta _*,\psi _*)\in \mathbb {R}_+^*\times C^2(\overline{S}_+^{N-1})\) the unique couple such \(\max \psi _*=1\) with the property that the function \((r,\sigma )\mapsto r^{-\beta _*}\psi _*(\sigma )\) is positive, p-harmonic in \(\mathbb {R}_+^N\) and vanishes on \(\partial {\mathbb {R}^N_+}{\setminus }\{0\}\). Then \(\psi _*=\psi \) satisfies

$$\begin{aligned}&\displaystyle -div'\left( \left( \beta _*^2\psi ^2+|\nabla ' \psi |^2\right) ^{\frac{p-2}{2}}\nabla ' \psi \right) -\beta _*\Lambda _{\beta _*}\left( \beta _*^2\psi ^2+|\nabla ' \psi |^2\right) ^{\frac{p-2}{2}}\psi =0\quad \text{ in } S_+^{N-1} \nonumber \\&\displaystyle \psi =0\quad \text{ on } \partial S_+^{N-1}. \end{aligned}$$
(3.2)

Since the function \(\psi _*\) is unique it depends only on the azimuthal variable \(\theta _{N-1}=\cos ^{-1}(\frac{x_N}{|x|})\) (see Appendix B). Our first result is the following

Theorem 3.1

If \(q\ge q_*\), or equivalently \(\beta _q\le \beta _*\), there exists no positive solution to problem (3.1).

Proof

Suppose such a solution \(\omega \) exists and put \(\theta =\beta _q/\beta _*\), then \(0<\theta \le 1\). Set \(\eta =\psi ^\theta \), where \(\psi \) is a positive solution of (3.2), and define the operator \({\mathcal T}\) by

$$\begin{aligned} {\mathcal T}(\eta )&=-div'\left( \left( \beta _q^2\eta ^2+|\nabla ' \eta |^2\right) ^{\frac{p-2}{2}}\nabla ' \eta \right) -\beta _q\Lambda _{\beta _q}\left( \beta _q^2\eta ^2+|\nabla ' \eta |^2\right) ^{\frac{p-2}{2}}\eta \nonumber \\&\quad +\left( \beta _q^2\eta ^2+|\nabla ' \eta |^2\right) ^{\frac{q}{2}}. \end{aligned}$$
(3.3)

Since \(\nabla \eta =\theta \psi ^{\theta -1}\nabla \psi \),

$$\begin{aligned} \left( \beta _q^2\eta ^2+|\nabla ' \eta |^2\right) ^{\frac{p-2}{2}}&=\theta ^{p-2}\psi ^{(\theta -1)(p-2)}\left( \beta _*^2\psi ^2+|\nabla ' \psi |^2\right) ^{\frac{p-2}{2}},\\ \left( \beta _q^2\eta ^2+|\nabla ' \eta |^2\right) ^{\frac{p-2}{2}}\nabla ' \eta&=\theta ^{p-1}\psi ^{(\theta -1)(p-1)}\left( \beta _*^2\psi ^2+|\nabla ' \psi |^2\right) ^{\frac{p-2}{2}}\nabla '\psi , \end{aligned}$$

therefore

$$\begin{aligned} {\mathcal T}(\eta )&=-\theta ^{p-1}\psi ^{(\theta -1)(p-1)}\mathrm{div'}\left( \left( \beta _*^2\psi ^2+|\nabla ' \psi |^2\right) ^{\frac{p-2}{2}}\nabla '\psi \right) \\&\quad - \theta ^{p-1}(\theta -1)(p-1)\psi ^{(\theta -1)(p-1)-1}\left( \beta _*^2\psi ^2+|\nabla ' \psi |^2\right) ^{\frac{p-2}{2}}|\nabla '\psi |^2\\&\quad -\beta _q\Lambda _{\beta _q}\theta ^{p-2}\psi ^{(\theta -1)(p-1)}\left( \beta _*^2\psi ^2{+}|\nabla ' \psi |^2\right) ^{\frac{p-2}{2}}\psi {+}\theta ^{q}\psi ^{(\theta -1)q}\left( \beta _*^2\psi ^2+|\nabla ' \psi |^2\right) ^{\frac{q}{2}}. \end{aligned}$$

But \(\beta _q\Lambda _{\beta _q}\theta ^{p-2}=\beta _*\Lambda _{\beta _q}\theta ^{p-1}\le \beta _*\Lambda _{\beta _*}\theta ^{p-1}\) since \(\beta _q\le \beta _*\). Using (3.2), we see that \({\mathcal T}(\eta )\ge 0\). Because Hopf Lemma is valid, there holds \(\partial _\mathbf{n}\psi <0\) on \(\partial S^{N-1}_+\). Since \(\omega \) is \(C^1\) in \(\overline{S^{N-1}_+}\) and \(\psi \) is defined up to an homothety, there exists a smallest function \(\psi \) such that \(\eta \ge \omega \), and the graphs of \(\eta \) and \(\omega \) over \(\overline{S^{N-1}_+}\) are tangent, either at some \(\alpha \in S^{N-1}_+\), or only at a point \(\alpha \in \partial S^{N-1}_+\). We put \(w=\eta -\omega \). Then

$$\begin{aligned} {\mathcal T}(\eta )={\mathcal T}(\eta )-{\mathcal T}(\omega )=\Phi (1)-\Phi (0), \end{aligned}$$
(3.4)

where \(\Phi (t)={\mathcal T}(\omega _t)\) with \(\omega _t=\omega +tw\).

We use local coordinates \((\sigma _1,\ldots ,\sigma _{N-1})\) on \(S^{N-1}\) near \(\alpha \). We denote by \(g=(g_{ij})\) the metric tensor on \(S^{N-1}\) and by \(g^{jk}\) its contravariant components. Then, for any \(\varphi \in C^1(S^{N-1})\),

$$\begin{aligned} {\left| \nabla \varphi \right| ^{2}}= \sum _{j,k}g^{jk}\frac{\partial \varphi }{\partial \sigma _{j}}\frac{\partial \varphi }{\partial \sigma _{k}}=\langle \nabla \varphi ,\nabla \varphi \rangle _g. \end{aligned}$$

If \(X=(X^1,\ldots , X^d)\in C^1(TS^{N-1})\) is a vector field, we lower indices by setting \( {X^\ell =\sum _{i}g^{\ell i}X_{i}}\) and define the divergence of X by

$$\begin{aligned} div'_gX=\frac{1}{\sqrt{\left| g\right| }} \sum _{\ell } \frac{\partial }{\partial \sigma _{\ell }}\left( \sqrt{\left| g\right| }X^\ell \right) =\frac{1}{\sqrt{\left| g\right| }} \sum _{\ell ,i}\frac{\partial }{\partial \sigma _{\ell }}\left( \sqrt{\left| g\right| }g^{\ell i}X_{i}\right) . \end{aligned}$$

We write \(\Phi (t)=\Phi _1(t)+\Phi _2(t)+\Phi _3(t)\) where

$$\begin{aligned} \Phi _1(t)=-\beta _q\Lambda _{\beta _q}\left( \beta _q^2\omega _t^2+|\nabla ' \omega _t|^2\right) ^{\frac{p-2}{2}}\omega _t,\quad \Phi _2(t)=\left( \beta _q^2\omega _t^2+|\nabla ' \omega _t|^2\right) ^{\frac{q}{2}} \end{aligned}$$

and

$$\begin{aligned} \Phi _3(t)=-\mathrm {div'}\left( \left( \beta _q^2\omega _t^2+|\nabla ' \omega _t|^2\right) ^{\frac{p-2}{2}}\nabla ' \omega _t\right) . \end{aligned}$$

Then

$$\begin{aligned} \Phi _1(1)-\Phi _1(0)=-\sum _{j}a_{j}\frac{\partial w}{\partial \sigma _{j}}-bw\quad \text {and}\quad \Phi _2(1)-\Phi _2(0)=\sum _{j}c_{j}\frac{\partial w}{\partial \sigma _{j}}+dw, \end{aligned}$$

where

$$\begin{aligned} b&=\beta _q\Lambda _{\beta _q}\left( \beta _q^{2}{\omega _t}^2+\left| \nabla \omega _t\right| ^2\right) ^{\frac{p}{2}-2} \left( (p-1)\beta _q^2\omega _t^2+\left| \nabla \omega _t\right| ^2\right) ,\\ a_{j}&=(p-2)\beta _q\Lambda _{\beta _q}\left( \beta _q^{2}{\omega _t}^2+\left| \nabla \omega _t\right| ^2\right) ^{\frac{p}{2}-2}\omega _t\sum _{k}g^{jk} {\frac{\partial \omega _t}{\partial \sigma _k} },\\ d&=q\beta _q^{2}\left( \beta ^{2}{\omega _t}^2+\left| \nabla \omega _t\right| ^2\right) ^{\frac{q}{2}-1}\omega _t, \end{aligned}$$

and

$$\begin{aligned} c_{j}=q\left( \beta _q^{2}{\omega _t}^2+\left| \nabla \omega _t\right| ^2\right) ^{\frac{q}{2}-1}\sum _{k}g^{jk} {\frac{\partial \omega _t}{\partial \sigma _k} }. \end{aligned}$$

Furthermore

$$\begin{aligned} \Phi _3(1)-\Phi _3(0)&=-(p-2)\mathrm{div'}\left( \left( \beta _q^2\omega _t^2+|\nabla ' \omega _t|^2\right) ^{\frac{p-4}{2}}\left( \beta _q^2\omega _t w+ \langle \nabla '\omega _t,\nabla ' w\rangle _g\right) \nabla '\omega _t\right) \\&\quad -\mathrm{div'}\left( \left( \beta _q^2\omega _t^2+|\nabla ' \omega _t|^2\right) ^{\frac{p-2}{2}}\nabla ' w\right) . \end{aligned}$$

Therefore we can write \(\Phi (1)-\Phi (0)\) under the form

$$\begin{aligned} \Phi (1)-\Phi (0)=-\mathrm{div'}(A\nabla 'w)+\langle B,\nabla ' w\rangle _g+Cw:={\mathcal L}w \end{aligned}$$
(3.5)

where

$$\begin{aligned} \langle AX,X\rangle _g&=\left( \beta _q^2\omega _t^2+|\nabla ' \omega _t|^2\right) ^{\frac{p-4}{2}} \left( p-2)\langle \nabla '\omega _t,X\rangle _g^2+|\nabla ' \omega _t|^2|X|^2\right) \nonumber \\&\ge \left( \beta _q^2\omega _t^2+|\nabla ' \omega _t|^2\right) ^{\frac{p-4}{2}}\min \{1,p-1\}|\nabla ' \omega _t|^2|X|^2. \end{aligned}$$
(3.6)

and B and C can be computed from the previous expressions. It is important to notice that \(\beta _q^2\omega _t^2+|\nabla ' \omega _t|^2\) is bounded between two positive constants \(m_1\) and \(m_2\) in \(\overline{S^{N-1}_+}\). Thus the operator \({\mathcal L}\) is uniformly elliptic with bounded coefficients. Since w is nonnegative and either at some point \(\alpha \), \(\nabla 'w(\alpha )=0\) and \(w(\alpha )>0\), or at some boundary point \(\alpha \) where \(w(\alpha )=0\) and \(\partial _\mathbf{n}w(\alpha )<0\), it follows from the strong maximum principle or Hopf Lemma (see [7]) that \(w=0\), contradiction. \(\square \)

Theorem 3.2

Assume \(q<q_*\) or equivalently \(\beta _q> \beta _*\). There exists a unique positive solution \(\omega _*\) to problem (3.1).

Proof

Existence It will follow from [4]. Indeed problem (3.1) can be written under the form

$$\begin{aligned} \begin{array}{lll} \mathbf{A}(\omega ):=-div'\,\mathbf{a}(\omega ,\nabla '\omega )&{}=\mathbf{B}(\omega ,\nabla '\omega )&{}\quad \text {in }S^{N-1}_+\\ \omega &{}=0&{}\quad \text {on }\partial S^{N-1}_+, \end{array} \end{aligned}$$
(3.7)

where

$$\begin{aligned} \mathbf{a}(r,\xi )&=\left( \beta _q^2r^2+|\xi |^2\right) ^{\frac{p-2}{2}}\xi ,\nonumber \\ \mathbf{B}(r,\xi )&=\beta _q\Lambda _{\beta _q}\left( \beta _q^2r^2+|\xi |^2\right) ^{\frac{p-2}{2}}r -\left( \beta _q^2r^2+|\xi |^2\right) ^{\frac{q}{2}}. \end{aligned}$$
(3.8)

The operator \(\mathbf{A}\) is a Leray–Lions operator which satisfies the assumptions (1.6)–(1.8) of [4, Theorem 2.1], and the term \(\mathbf{B}\) satisfies (1.9), (1.10) in the same article. Therefore the existence of a positive solution \(\omega \in W^{1,p}_0(S^{N-1}_+)\cap L^\infty (S^{N-1}_+)\) is ensured whenever we can find a supersolution \(\overline{\omega }\in W^{1,p}(S^{N-1}_+)\cap L^\infty (S^{N-1}_+)\) and a nontrivial subsolution \(\underline{\omega }\in W^{1,p}(S^{N-1}_+)\) of (3.7) such that

$$\begin{aligned} 0\le \underline{\omega }\le \overline{\omega }\quad&\text {in }S^{N-1}_+. \end{aligned}$$
(3.9)

First we note that \(\eta =\eta _0\) is a supersolution if the positive constant \(\eta _0\) is large enough. In order to find a subsolution, we set again \(\eta =\psi ^\theta \) with \(\theta =\beta _q/\beta _*\) and \(\psi \) as in (3.2). Now \(\theta > 1\), thus \(\eta \in W^{1,p}_0(S^{N-1}_+)\). As above we have

$$\begin{aligned} {\mathcal T}(\eta )&=-\theta ^{p-1}\psi ^{(\theta -1)(p-1)}\mathrm{div'}\left( \left( \beta _*^2\psi ^2+|\nabla ' \psi |^2\right) ^{\frac{p-2}{2}}\nabla '\psi \right) \\&\quad -\theta ^{p-1}(\theta -1)(p-1)\psi ^{(\theta -1)(p-1)-1}\left( \beta _*^2\psi ^2+|\nabla ' \psi |^2\right) ^{\frac{p-2}{2}}|\nabla '\psi |^2\\&\quad -\beta _q\Lambda _{\beta _q}\theta ^{p-2}\psi ^{(\theta -1)(p-1)}\left( \beta _*^2\psi ^2+|\nabla ' \psi |^2\right) ^{\frac{p-2}{2}}\psi {+}\theta ^{q}\psi ^{(\theta -1)q}\left( \beta _*^2\psi ^2{+}|\nabla ' \psi |^2\right) ^{\frac{q}{2}}. \end{aligned}$$

Now \(\beta _q\Lambda _{\beta _q}\theta ^{p-2}=\beta _*\Lambda _{\beta _q}\theta ^{p-1}=\beta _*(\Lambda _{\beta _q}-\Lambda _{\beta _*})\theta ^{p-1}+\beta _*\Lambda _{\beta _*}\theta ^{p-1}\) and \(\Lambda _{\beta _q}-\Lambda _{\beta _*}=(\beta _q-\beta _*)(p-1)=\beta _*(p-1)(\theta -1)\), hence

$$\begin{aligned} {\mathcal T}(\eta )&=- \theta ^{p-1}\psi ^{(\theta -1)(p-1)}\mathrm{div'}\left( \left( \beta _*^2\psi ^2+|\nabla ' \psi |^2\right) ^{\frac{p-2}{2}}\nabla '\psi \right) \\&\quad -\theta ^{p-1}(\theta -1)(p-1)\psi ^{(\theta -1)(p-1)-1}\left( \beta _*^2\psi ^2+|\nabla ' \psi |^2\right) ^{\frac{p-2}{2}}|\nabla '\psi |^2\\&\quad -\beta _*(\Lambda _{\beta _q}-\Lambda _{\beta _*})\theta ^{p-1}\psi ^{(\theta -1)(p-1)}\left( \beta _*^2\psi ^2+|\nabla ' \psi |^2\right) ^{\frac{p-2}{2}}\psi \\&\quad -\beta _*\Lambda _{\beta _*}\theta ^{p-1}\psi ^{(\theta -1)(p-1)}\left( \beta _*^2\psi ^2+|\nabla ' \psi |^2\right) ^{\frac{p-2}{2}}\psi {+}\theta ^{q}\psi ^{(\theta -1)q}\left( \beta _*^2\psi ^2{+}|\nabla ' \psi |^2\right) ^{\frac{q}{2}}. \end{aligned}$$

Using the equation satisfied by \(\psi \) yields to the relation

$$\begin{aligned} {\mathcal T}(\eta )&= -\theta ^{p-1}(\theta -1)(p-1)\psi ^{(\theta -1)(p-1)-1}\left( \beta _*^2\psi ^2+|\nabla ' \psi |^2\right) ^{\frac{p-2}{2}}|\nabla '\psi |^2\\&\quad -\beta _*^2(p-1)(\theta -1)\theta ^{p-1}\psi ^{(\theta -1)(p-1)-1}\left( \beta _*^2\psi ^2+|\nabla ' \psi |^2\right) ^{\frac{p-2}{2}}\psi ^2\\&\quad +\theta ^{q}\psi ^{(\theta -1)q}\left( \beta _*^2\psi ^2+|\nabla ' \psi |^2\right) ^{\frac{q}{2}}\\&=-\theta ^{p-1}(\theta -1)(p-1)\psi ^{(\theta -1)(p-1)-1}\left( \beta _*^2\psi ^2+|\nabla ' \psi |^2\right) ^{\frac{p}{2}}\\&\quad +\theta ^{q}\psi ^{(\theta -1)q}\left( \beta _*^2\psi ^2+|\nabla ' \psi |^2\right) ^{\frac{q}{2}}. \end{aligned}$$

If we replace \(\eta :=\eta _1=\psi ^\theta \) by \(\eta :=\eta _m=(m\psi )^\theta \) in the above computation, the inequality \({\mathcal T}\eta _m)\le 0\) will be true provided

$$\begin{aligned} m^{\theta (q+1-p)}\psi ^{(\theta -1)(q+1-p)+1}\le \theta ^{p-1-q}(\theta -1)(p-1)\left( \beta _*^2\psi ^2+|\nabla ' \psi |^2\right) ^{\frac{p-q}{2}}, \end{aligned}$$

which is satisfied if we choose m small enough so that \((m\psi )^\theta \le \eta _0\) and satisfying

$$\begin{aligned} m^{\theta (q+1-p)}\le \beta _*^{(\theta -1)(q+1-p)+1}\theta ^{p-1-q}(\theta -1)(p-1)\frac{\min _{x\in S_+^{N-1}} \left( \beta _*^2\psi ^2+|\nabla ' \psi |^2\right) ^{\frac{p-q}{2}}}{\max _{x\in S_+^{N-1}} \psi ^{(\theta -1)(q+1-p)+1}}. \end{aligned}$$

Therefore \(0<\eta _m\le \eta _0\) and standard regularity implies that the solution \(\omega \) is \(C^{1}\) in \(\overline{S}^{N-1}_+\). Actually \(\omega \) is \(C^{\infty }\) since the operator is not degenerate.

Uniqueness We use the tangency method developed in the proof of Theorem 3.1. Assume \(\omega _1\) and \(\omega _2\) are two positive solutions of (3.2), then they are positive in \(S^{N-1}_+\) and \(\partial _\mathbf{n}\omega _i<0\) on \(\partial S^{N-1}_+\). Either the \(\omega _i\) are ordered and \(\omega _1\le \omega _2\), or their graphs intersect. In any case we can define

$$\begin{aligned} \tau =\inf \{s>1:s\omega _1\ge \omega _2\}. \end{aligned}$$

We set \(\omega ^*=\tau \omega _1\). Then either the graphs of \(\omega _2\) and \(\omega ^*\) are tangent at some interior point \(\alpha \), or they are not tangent in \(S^{N-1}_+\), \(\partial _\mathbf{n}\omega ^*\le \partial _\mathbf{n}\omega _2<0\) on \(\partial S^{N-1}_+\) and there exists \(\alpha \in \partial S^{N-1}_+\) such that \(\partial _\mathbf{n}\omega ^*(\alpha )= \partial _\mathbf{n}\omega _2(\alpha )<0\). Furthermore \({\mathcal T}(\omega ^*)\ge 0\). If we set \(w=\omega ^*-\omega _2\), then, as in Theorem 3.1,

$$\begin{aligned} -\mathrm{div'}(\tilde{A}\nabla 'w)+\langle \tilde{B},\nabla ' w\rangle _g+\tilde{C}w=\tilde{\mathcal L}w\ge 0 \end{aligned}$$

where

$$\begin{aligned} \langle \tilde{A}X,X\rangle _g&=\left( \beta _q^2\omega _t^2+|\nabla ' \omega _t|^2\right) ^{\frac{p-4}{2}} \left( p-2)\langle \nabla '\omega _t,X\rangle _g^2+|\nabla ' \omega _t|^2|X|^2\right) \nonumber \\&\ge \left( \beta _q^2\omega _t^2+|\nabla ' \omega _t|^2\right) ^{\frac{p-4}{2}}\min \{1,p-1\}|\nabla ' \omega _t|^2|X|^2, \end{aligned}$$
(3.10)

in which \(\omega _t=\omega _2+t(\omega ^*-\omega _2)\) and \(t\in (0,1)\) is obtained by applying the mean value theorem and \(\tilde{B}\) and \(\tilde{C}\) are defined accordingly. Since \(\tilde{\mathcal L}\) is uniformly elliptic and has bounded coefficients, it follows from the strong maximum principle that \(w=0\). Thus \(\omega ^*=\tau \omega _1=\omega _2\) and \(\tau =1\) from the equation. This ends the proof. \(\square \)

3.2 Removable boundary singularities

The following is the basic result for removability of isolated singularities. It is valid in the general case, but with a local geometric constraint.

Theorem 3.3

Assume \(q^*\le q<p\le N\), \(\Omega \) is a \(C^2\) bounded domain with \(0\in \partial \Omega \), such that \(\Omega \cap B_\delta =B_\delta ^+\) for some \(\delta >0\). If \(u\in C^1(\overline{\Omega }{\setminus }\{0\})\) is a nonnegative solution of (1.1) in \(\Omega \) which vanishes on \(\partial \Omega {\setminus }\{0\}\), then it is identically 0.

Proof

Step 1: assume \(\Omega \subset \mathbb {R}^N_+\) For \(\epsilon >0\), we set \(\Omega '_\epsilon =\Omega \cap \overline{B^c_\epsilon }\) and \(H_\epsilon =\mathbb {R}^N_+\cap \overline{B^c_\epsilon }\). For \(k,n\in \mathbb N_*\), \(n\ge \mathrm{diam\,}(\Omega )\), we denote by \(v_{k,n,\epsilon }\) (\(n\in \mathbb N_*\)) the solution of the problem

$$\begin{aligned} -\Delta _pv+\left| \nabla v\right| ^q&=0\qquad \qquad \;\;\;\text {in }H_\epsilon \cap B_n\nonumber \\ v&=k\chi _{_{\mathbb {R}^N_+\cap \partial B_\epsilon }}\quad \text {on }\partial (H_\epsilon \cap B_n). \end{aligned}$$
(3.11)

If \(k>c_2\epsilon ^{\frac{q-p}{q+1-p}}\) for a suitable \(c_2=c_2(p,q)>0\) (see Lemma 2.6), then \(v_{k,n,\epsilon }\ge u\) in \(\Omega '_\epsilon \). Moreover there holds \(v_{k,n,\epsilon }\le v_{k',n',\epsilon }\) for \(n\le n'\) and \(k\le k'\). Furthermore the function

$$\begin{aligned} U_{\epsilon ,n}(x)=c_2\left( (\left| x\right| -\epsilon )^{\frac{q-p}{q+1-p}}-(n-\epsilon )^{\frac{q-p}{q+1-p}}\right) \end{aligned}$$

is a supersolution in \(B_n{\setminus } B_\epsilon \), and there holds \(v_{k,n,\epsilon }\le U_{\epsilon ,n}\). By monotonicity and standard a priori estimate, we obtain that \(v_{k,n,\epsilon }\rightarrow v_\epsilon \) when \(n, k\rightarrow \infty \) and that the function \(v=v_\epsilon \) is solution of

$$\begin{aligned} -\Delta _pv+\left| \nabla v\right| ^q&=0\qquad \text {in }H_\epsilon \nonumber \\ \lim \nolimits _{\left| x\right| \rightarrow \epsilon }v(x)&=\infty \nonumber \\ v&=0\qquad \text {on }\partial \mathbb {R}^N_+\cap \overline{B^c_\epsilon }. \end{aligned}$$
(3.12)

Furthermore

$$\begin{aligned} u(x)\le v_\epsilon (x)\le c_2(\left| x\right| -\epsilon )^{\frac{q-p}{q+1-p}}\quad \text {in }\Omega '_\epsilon . \end{aligned}$$
(3.13)

The function \(v_\epsilon \) may not be unique, however it is the minimal solution of the above problem since the \(v_{k,n,\epsilon }\) is unique, and monotonicity in n and k holds. Actually, \(v_\epsilon \le v_{\epsilon '}\) if \(0\le \epsilon \le \epsilon '\). For \(\ell >0\), we recall that the transformation \(v\mapsto T_\ell [v]\) defined by (2.41) leaves Eq. (1.1) invariant. As a consequence of the uniqueness of the approximations we have \(T_\ell [v_{k,n,\epsilon }]=v_{ \ell ^{\frac{p-q}{q+1-p}}k,\ell ^{-1}n,\ell ^{-1}\epsilon }\), which implies

$$\begin{aligned} T_\ell [v_\epsilon ]=v_{\ell ^{-1}\epsilon }. \end{aligned}$$
(3.14)

Letting \(\epsilon \rightarrow 0\), we derive from the monotonicity with respect to \(\epsilon \) and standard \(C^{1,\alpha }\) estimates, that the following identity holds:

$$\begin{aligned} T_\ell [v_0]=v_0\quad \forall \ell >0. \end{aligned}$$
(3.15)

The function \(v_0\) is a positive and separable solution of (1.1) in \(\mathbb {R}^N_+\) which vanishes on \(\partial \Omega {\setminus }\{0\}\). It follows from Theorem 3.1 that \(v_0=0\), and so is u.

Step 2: the general case We assume that \(\Omega \cap B_\delta \subset \mathbb {R}^N_+\) and we denote by M the maximum of u on \(\partial B_\delta \cap \Omega \). Then the function \((u-M)_+\) is a subsolution of (1.1) in \(\Omega \cap B_\delta \) which vanishes on \(\partial \Omega \cap B_\delta {\setminus }\{0\}\). By Step 1, it is dominated by \(v_0\), which ends the proof. \(\square \)

Remark

The previous result is valid if u is a subsolution with the same regularity. If u is no longer assumed to be nonnegative, only \(u^+\) vanishes. Furthermore, the regularity of the boundary has not been used, but only the fact that \(\Omega \) is locally contained into a half space to the boundary of which 0 belongs.

Remark

If no geometric assumption is made on \(\partial \Omega \), we can prove that \(u(x)=o (\left| x\right| ^{-\beta _q})\) near 0. The next result shows that the removability holds if \(q>q_*\).

Theorem 3.4

Assume \(q^*< q<p\le N\) and \(\Omega \) is a \(C^2\) bounded domain with \(0\in \partial \Omega \). If u is a nonnegative solution of (1.1) in \(\Omega \) which belongs to \(C^1(\overline{\Omega }{\setminus }\{0\})\) and vanishes on \(\partial \Omega {\setminus }\{0\}\), it is identically 0.

Proof

As it is proved in [12], for any smooth subdomain \(S\subset S^{N-1}\), there exists a unique \(\beta _{*\,s}>0\) and \(\psi ^*_s>0\), unique up to an homothety, such that \(x\mapsto \left| x\right| ^{-\beta _{*\,s}}\psi ^*_s(\left| x\right| ^{-1}x)\) is p harmonic in the cone \(C_S=\{x\in \mathbb {R}^N{\setminus }\{0\}:\left| x\right| ^{-1}x\in S\}\) and \(\psi ^*_s\) satisfies

$$\begin{aligned}&\displaystyle -div'\left( \left( \beta _{*\,s}^2\psi ^{*\,2}_s+|\nabla ' \psi ^*_s|^2\right) ^{\frac{p-2}{2}}\nabla ' \psi ^*_s\right) -\beta _{*\,s}\Lambda _{\beta _{*\,s}}\left( \beta _{*\,s}^2\psi ^{*\,2}+|\nabla ' \psi ^*_s|^2\right) ^{\frac{p-2}{2}}\psi ^*_s=0\quad \text{ in } S\nonumber \\&\displaystyle \psi ^*_s=0\quad \text{ on } \partial S, \end{aligned}$$
(3.16)

Furthermore \(S\subset \tilde{S}\subset S^{N-1}\) implies \(\beta _{*\,\tilde{s}}\le \beta _{*\,s}\). Using the system of spherical coordinates defined in (6.5) in Appendix B, for \(\epsilon >0\) we denote by \(S:=S_\epsilon \) the spherical shell with vertex the north pole N and latitude angle \(\theta _{N-1}\in [0,\frac{\pi }{2}+\epsilon ]\). Because of uniqueness of \(\beta _{*\,s}\), \(\beta _{*\,s_\epsilon }\uparrow \beta _*\) as \(\epsilon \rightarrow 0\). Therefore, if \(q>q_*\), or equivalently \(\beta _q<\beta _*\), there exists \(\delta ,\epsilon >0\) such that \(\Omega \cap B_\delta \subset C_{S_\epsilon }\cap B_\delta \) and \(\beta _q<\beta _{*\,s_\epsilon }\). Since 3.1 is valid if \(S^{N-1}_+\) is replaced by \(S_\epsilon \) and \(\beta _q<\beta _{*\,s_\epsilon }\) it follows that \(u=0\) as in the proof of Theorem 3.3, Steps 1 and 2. \(\square \)

The next result, valid in the case \(p=N\), is based upon the conformal invariance of the N-Laplacian. In this case the exponent \(\beta _*\) corresponding to the first spherical N-harmonic eigenvalue is equal to 1 and the corresponding spherical N-harmonic eigenfunction in \(S^{N-1}_+\) is \(x_N/\left| x\right| ^{2}\).

Theorem 3.5

Assume \(N-\frac{1}{2}\le q< N\), \(\Omega \) is a bounded domain and \(0\in \partial \Omega \) is such that there exists a ball \(B\subset \Omega ^c\) to the boundary of which 0 belongs. If u is a nonnegative solution of

$$\begin{aligned} -\Delta _Nu+\left| \nabla u\right| ^q=0\quad \text {in }\Omega , \end{aligned}$$
(3.17)

which belongs to \(C(\overline{\Omega }{\setminus }\{0\})\cap W^{1,N}_0(\Omega {\setminus } \overline{B}_\epsilon (0))\) for any \(\epsilon >0\), it is identically 0.

Proof

We assume that the inward normal unit vector to \(\partial \Omega \) at 0 is \(\mathbf{e}_N=(0,0,\ldots ,1)\) and that the ball \(B=B_{\frac{1}{2}}(a)\) of center \(a=-\frac{1}{2}\mathbf{e}_N\) and radius \(\frac{1}{2}\) touches \(\partial \Omega \) at 0 and is exterior to \(\Omega \) (this can be assumed up to a rotation and a dilation). This is the consequence of the exterior sphere condition at the point 0. It is always valid if \(\partial \Omega \) is \(C^2\). We denote by \({\mathcal I}_\omega \) the inversion of center \(\omega =-\mathbf{e}_N\) and power 1, i.e. \({\mathcal I}_\omega (x)=\omega +\frac{x-\omega }{\left| x-\omega \right| ^2}\). Under this transformation, the complement of the ball \(B_{\frac{1}{2}}(a)\), which contains \(\Omega \), is transformed into the half space \(\mathbb {R}^N_-\) which contains the image \(\tilde{\Omega }\) of \(\Omega \). Since u satisfies (3.17), \(\tilde{u}=u\circ {\mathcal I}_\omega \) satisfies

$$\begin{aligned} -\Delta _N\tilde{u}+\left| x-\omega \right| ^{2(q-N)}\left| \nabla \tilde{u}\right| ^q=0\quad \text {in }\tilde{\Omega }. \end{aligned}$$
(3.18)

Furthermore since \(0={\mathcal I}_\omega (0)\) and \({\mathcal I}_\omega \) is a diffeomorphism, \(\tilde{u}\in C(\overline{\tilde{\Omega }}{\setminus }\{0\})\cap C^1(\tilde{\Omega })\) and it vanishes on \(\partial \tilde{\Omega }{\setminus }\{0\}\). Since \(\left| x-\omega \right| \le 1\) and \(q<N\), \(\tilde{u}\) is a subsolution for (3.17) in \(\tilde{G}\). By Theorem 3.4, \(\tilde{u}=0\). \(\square \)

3.3 Weakly singular solutions

The main result of this section is the following existence and uniqueness result concerning solutions of (1.1) with a boundary weak singularity. We recall that \(\psi _*\) is unique positive solution of (1.11) such that \(\sup \psi _*=1\). Our first result is valid for any \(1<p\le N\) but it needs a geometric constraint on \(\Omega \).

Theorem 3.6

Let \(p-1<q<q_*<p\le N\) and \(\Omega \subset \mathbb {R}^N_+\) be a bounded \(C^2\) domain such that \(0\in \partial \Omega \). Assume that there exists \(\delta >0\) such that \(\Omega _\delta :=\Omega \cap B_\delta = B^+_\delta \). Then for any \(k>0\) there exists a unique positive solution \(u:=u_k\) of (1.1) in \(\Omega \), which belongs to \(C^1(\overline{\Omega }{\setminus }\{0\})\), vanishes on \(\partial \Omega {\setminus }\{0\}\) and satisfies

$$\begin{aligned} \lim _{x\rightarrow 0}\frac{u_k(x)}{\Psi _*(x)}=k \end{aligned}$$
(3.19)

in the \(C^1\)-topology of \(S^{N-1}_+\), where

$$\begin{aligned} \Psi _*(x)=\left| x\right| ^{-\beta _*}\psi _*(|x|^{-1}x). \end{aligned}$$

The proof of this theorem is long and difficult and requires a certain number of intermediate results.

Lemma 3.7

Let the assumptions on p, q and \(\Omega \) of Theorem 3.6 be satisfied. There exists a unique positive p-harmonic function \(\Phi _{*}\) in \(\Omega \), which is continuous in \(\overline{\Omega }{\setminus }\{0\}\), vanishes on \(\partial \Omega {\setminus }\{0\}\) and satisfies

$$\begin{aligned} \lim _{x\rightarrow 0}\frac{\Phi _{*}(x)}{\Psi _*(x)}=1. \end{aligned}$$
(3.20)

Proof

For \(0<\epsilon <\delta \) let \(v_\epsilon \) be the unique nonnegative p-harmonic function in \(\Omega {\setminus } \overline{B^+_{\epsilon }}\) which is continuous in \(\overline{\Omega }{\setminus } {B^+_{\epsilon }}\), vanishes on \(\partial \Omega {\setminus } {B_{\epsilon }}\) and achieves the value \({\Psi _*}\) on \(\partial B_\epsilon \cap \Omega \). Since \(\Omega \subset \mathbb {R}^N_+\), \(v_\epsilon \le \Psi _*\) in \(\Omega {\setminus } {B^+_{\epsilon }}\). Hence inequalities \(0<\epsilon <\epsilon '\le \delta \) imply \(v_\epsilon \le v_{\epsilon '}\) in \(\Omega {\setminus } \overline{B^+_{\epsilon '}}\). Because \(\Psi _*\le \delta ^{-\beta _*}\), there holds

$$\begin{aligned} v_\epsilon +\delta ^{-\beta _*}\ge \Psi _*, \end{aligned}$$
(3.21)

in \(\Omega {\setminus } B_\delta ^+ \). Since \(v_\epsilon \) and \(\Psi _*\) coincide on \(\partial B_\epsilon ^+\) and vanish on \(\partial \mathbb {R}^N_+\cap (B_\delta ^+ {\setminus } B_\epsilon ^+)\), (3.21) holds also in \(B_\delta ^+ {\setminus } B_\epsilon ^+\). Because \(v_\epsilon \ge 0\) there holds

$$\begin{aligned} (\Psi _*-\delta ^{-\beta _*})_+\le v_\epsilon \le \Psi _*\quad \text{ in } \Omega {\setminus } B_\epsilon ^+. \end{aligned}$$
(3.22)

By a standard regularity result \(v_\epsilon \) converges to a function \(\Phi _{*}\) continuous in \(\overline{\Omega }{\setminus }\{0\}\), p-harmonic in \(\Omega \) such that

$$\begin{aligned} (\Psi _*-\delta ^{-\beta _*})_+\le \Phi _{*}\le \Psi _* \end{aligned}$$

in \( \Omega \). Therefore (3.20) holds provided \(\frac{x}{\left| x\right| }\) remains in a compact subset of \(S^{N-1}_+\). Let us define a function \(\tilde{\phi }_{*}\) by \(\tilde{\phi }_*(x)=\left| x\right| ^{\beta _*}\Phi _*(x)\), then \(\tilde{\phi }_*(r,\sigma )\le {\psi _*}(\sigma )\) where \(r=\left| x\right| \) and \(\sigma =\frac{x}{\left| x\right| }\in S^{N-1}_+\). By standard \(C^{1,\alpha }\) estimates, \(\tilde{\phi }_*(r,.)\) is relatively compact in the \(C(\overline{S^{N-1}_+})\)-topology. Therefore the convergence of \(\frac{\Phi _*(x)}{{\Psi _*}(x)}\) to 1 when x to 0 holds not only when \(\frac{x}{\left| x\right| }\) remains in a compact subset of \(S^{N-1}_+\), but uniformly on \(S^{N-1}_+\), which implies (3.20). Uniqueness follows classically by (3.20) and the maximum principle. \(\square \)

Lemma 3.8

Let the assumptions on p, q and \(\Omega \) of Theorem 3.6 be satisfied. If for some \(k>0\) there exists a solution \(u_{k}\) of (1.1) in \(\Omega \), which belongs to \(C^1(\overline{\Omega }{\setminus }\{0\})\), vanishes on \(\partial \Omega {\setminus }\{0\}\) and satisfies (3.19), then for any \(k>0\) there exists such a solution.

Proof

We notice that for any \(c<1\) (resp \(c>1\)), \(cu_{k}\) is a subsolution (resp. supersolution) of (1.1) in \(\Omega \). Let \(\Phi _*\) be as in Lemma 3.7. If \(c<1\), the function \(ck\Phi _*\) is a supersolution of (1.1) which vanishes on \(\partial \Omega {\setminus }\{0\}\). Furthermore

$$\begin{aligned} \lim _{x\rightarrow 0}{\frac{cu_{k}(x)}{\Psi _*(x)} }=ck=\lim _{x\rightarrow 0}{\frac{ck\Phi _*(x)}{\Psi _*(x)} }. \end{aligned}$$

Then there exists a solution \(u_{ck}\) of (1.1) in \(\Omega \) which satisfies \(cu_{k}\le u_{ck}\le ck\Phi _*\). If \(c>1\), we set \(u^*:=T_{c^\theta }[u_k]\), which means \( u^*(x)= c^{\beta _q\theta }u_k(c^\theta \,x)\) with \(\theta =(\beta _q-\beta _*)^{-1}\). Then \(u^*\) is a solution of (1.1) in \(\Omega ^{c^\theta }=\frac{1}{c^\theta }\Omega \). In particular, \(u^*\) satisfies the equation in \(B^+_{\frac{\delta }{c^\theta }}(0)\). Since \(c^\theta >1\), \(B^+_{\frac{\delta }{c^\theta }}(0)\subset B^+_{\delta }(0)\). Put \(m=\max \{u^*:x\in \partial B^+_{\frac{\delta }{c^\theta }}(0)\}\). The function \((u^*-m)_+\), extended by 0 outside \(B^+_{\frac{\delta }{c^\theta }}(0)\), is a subsolution of (1.1) in \(\Omega \). Furthermore it satisfies

$$\begin{aligned} \lim _{x\rightarrow 0}{\frac{(u^*-m)_+(x)}{\Psi _*(x)} }=ck, \end{aligned}$$

uniformly on any compact subset of \(S^{N-1}_+\). Therefore there exists a solution \(u_{ck}\) of (1.1) in \(\Omega \) which satisfies \((u^*-m)_+\le u_{ck}\le ck\Phi _*\), and in particular it vanishes on \(\partial \Omega {\setminus }\{0\}\) and belongs to \(C^1(\overline{\Omega }{\setminus }\{0\})\). By [13], \(u_{ck}\) is positive in \(\Omega \). Thus \(u_{ck}\) belongs to \(C^{1,\alpha }(\overline{B^+_\delta }(0){\setminus }\{0\})\) and satisfies

$$\begin{aligned} \left| x\right| ^{\beta _*}\left| u_{ck}(x)\right| +\left| x\right| ^{1+\beta _*}\left| \nabla u_{ck}(x)\right| +\left| x\right| ^{1+\beta _*+\alpha }\sup _{\begin{array}{c} |y|\le |x|\\ x\ne y \end{array}} \frac{\left| \nabla u_{ck}(x)-\nabla u_{ck}(y)\right| }{|x-y|^\alpha }\le M \end{aligned}$$

by (2.11). Therefore the set of functions \(\{r^{\beta _*+1}\nabla u_{ck}(r,\cdot )\}_{r>0}\) is uniformly relatively compact in the topology of uniform convergence on \(\overline{S}_+^{N-1}\). Since it converges to \(ck\nabla '\psi _* \) uniformly on compact subsets of \(S_+^{N-1}\) as \(r\rightarrow 0\), this convergence holds in \(C(\overline{S_+^{N-1}})\). This implies

$$\begin{aligned} \lim _{x\rightarrow 0}\frac{u_{ck}(x)}{\Psi _*(x)}=ck. \end{aligned}$$
(3.23)

\(\square \)

The next Lemma is the keystone of our construction. Its proof is very delicate and needs several intermediate steps.

Lemma 3.9

Under the assumptions of Theorem 3.6 there exists a real number \(R_0\) such that \(0<R_0\le \delta \) and a positive subsolution \(\tilde{u}\) of (1.1) in \(B^+_{R_0}\) which is Lipschitz continuous in \(\overline{B^+_{R_0}}{\setminus }\{0\}\), vanishes on \(\overline{B^+_{R_0}}\cap \partial \mathbb {R}^N_+{\setminus }\{0\}\), is smaller than \(\Psi _*\) and satisfies

$$\begin{aligned} \lim _{x\rightarrow 0}\frac{\tilde{u}(x)}{\Psi _*(x)}=1. \end{aligned}$$
(3.24)

Proof

The construction of the function \(\tilde{u}\). We look for a subsolution under the form \(\tilde{u}=\Psi _*-w\) for a suitable nonnegative function w.

Step 1: reduction of the problem We use spherical coordinates for a \(C^1\) function \(u:x\mapsto u(x)=u(r,\sigma )\), \(r=|x|\), \(\sigma =\frac{x}{|x|}\). Then \(\nabla u=u_r\mathbf{e}+r^{-1}\nabla ' u\) where \(\mathbf{e}=\left| x\right| ^{-1}x\), \(\left| \nabla u\right| ^2= u^2_r+r^{-2}\left| \nabla ' u\right| ^2\) and \(\left| \nabla u\right| ^q=\left( u^2_r+r^{-2}\left| \nabla ' u\right| ^2\right) ^{\frac{q}{2}} \). The expression of the p-Laplacian in spherical coordinates is

$$\begin{aligned} -\Delta _p u&=-\left( \left( u^2_r+r^{-2}\left| \nabla ' u\right| ^2\right) ^{\frac{p-2}{2}}u_r\right) _r -{\frac{N-1}{r} }\left( u^2_r+r^{-2}\left| \nabla ' u\right| ^2\right) ^{\frac{p-2}{2}}u_r\\&\quad -{\frac{1}{r^2} } div'\left( \left( u^2_r+r^{-2}\left| \nabla ' u\right| ^2\right) ^{\frac{p-2}{2}}\nabla ' u\right) . \end{aligned}$$

Put \(v(t,\sigma )=r^{\beta _*}u(r,\sigma )\) with \(t=\ln r\in (-\infty ,\ln \delta ]\), then v satisfies

$$\begin{aligned}&{\mathcal Q}[v]:=\nonumber \\&\quad -\left( \left( (v_t-\beta _*v)^2+\left| \nabla ' v\right| ^2\right) ^{\frac{p-2}{2}}(v_t-\beta _*v)\right) _t\nonumber \\&\qquad -div'\left( \left( (v_t-\beta _*v)^2+\left| \nabla ' v\right| ^2\right) ^{\frac{p-2}{2}}\nabla 'v\right) \nonumber \\&\qquad +\Lambda _{\beta _*}\left( (v_t-\beta _*v)^2+\left| \nabla ' v\right| ^2\right) ^{\frac{p-2}{2}}(v_t-\beta _*v)+e^{\nu t}\left( (v_t-\beta _*v)^2+\left| \nabla ' v\right| ^2\right) ^{\frac{q}{2}}=0 \end{aligned}$$
(3.25)

in \((-\infty ,\ln \delta )\times S^{N-1}_+\) where \(\nu =1-(q+1-p)(\beta _*+1)=1-\frac{\beta _*+1}{\beta _q+1}>0\) and \(\Lambda _{\beta _*}=\beta _*(p-1)+p-N\). Notice that \(\psi _*\) satisfies

$$\begin{aligned} -div'\left( \left( \beta _*^2\psi _*^2+\left| \nabla ' \psi _*\right| ^2\right) ^{\frac{p-2}{2}}\nabla '\psi _*\right) -\beta _*\Lambda _{\beta _*}\left( \beta _*^2\psi _*^2+\left| \nabla ' \psi _*\right| ^2\right) ^{\frac{p-2}{2}}\psi _*=0,\nonumber \\ \end{aligned}$$
(3.26)

hence it is a supersolution for (3.25). We look for a subsolution under the form

$$\begin{aligned} V(t,\sigma )=\psi _*-a(t)g(\psi _*) \end{aligned}$$

where g is a continuous increasing function defined on \(\mathbb {R}_+\), vanishing at 0 and smooth on \(\mathbb {R}^*_+\) and \(a(t)=e^{\gamma t}\) with \(\gamma >0\) to be chosen. Thus \(a'=\gamma a\), \(a''=\gamma ^2 a\), \(V_t=-\gamma ag(\psi _*)\), \(V_t-\beta _*V=-\beta _*\psi _*+a(\beta _*-\gamma )g(\psi _*)\), \(\nabla ' V=(1-ag'(\psi _*))\nabla '\psi _*\) and

$$\begin{aligned}&(V_t-\beta _*V)^2+\left| \nabla ' V\right| ^2=\left( -\beta _*\psi _*+a(\beta _*-\gamma )g(\psi _*)\right) ^2+(1-ag'(\psi _*))^2\left| \nabla '\psi _*\right| ^2\\&\quad =\left( \beta _*^2\psi _*^2+2a\beta _*(\gamma -\beta _*)g(\psi _*)\psi _*\right) +\left( 1-2ag'(\psi _*)\right) \left| \nabla '\psi _*\right| ^2 +O(a^2\left\| g(\psi )\right\| _{C^1})\\&\quad =\beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2+2a\left( \beta _*(\gamma -\beta _*)\psi _*g(\psi _*)-g'(\psi _*)\left| \nabla \psi _*\right| ^2\right) +O(a^2\left\| g(\psi _*)\right\| _{C^1}). \end{aligned}$$

Therefore

$$\begin{aligned}&\left( (V_t-\beta _*V)^2+\left| \nabla ' V\right| ^2\right) ^{\frac{p-2}{2}}\\&\quad =\left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{p-2}{2}}\left[ 1+(p-2)a{\frac{\beta _*(\gamma -\beta _*)\psi _*g(\psi _*)-g'(\psi _*)\left| \nabla \psi _*\right| ^2 }{\beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2} }\right] \\&\qquad +O(a^2\left\| g(\psi )\right\| _{C^1}), \end{aligned}$$

and

$$\begin{aligned}&e^{\nu t}\left( (V_t-\beta _*V)^2+\left| \nabla ' V\right| ^2\right) ^{\frac{q}{2}} \\&\quad =e^{\nu t}\left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{q}{2}}\left[ 1+qa{\frac{\beta _*(\gamma -\beta _*)\psi _*g(\psi _*)-g'(\psi _*)\left| \nabla \psi _*\right| ^2 }{\beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2} }\right] \\&\qquad +O(e^{\nu t}a^2\left\| g(\psi _*)\right\| _{C^1}), \end{aligned}$$

thus

$$\begin{aligned}&\left( (V_t-\beta _*V)^2+\left| \nabla ' V\right| ^2\right) ^{\frac{p-2}{2}}(V_t-\beta _*V)\\&\quad =-\beta _*\left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{p-2}{2}}\psi _*+ a(\beta _* -\gamma )\left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{p-2}{2}}g(\psi _*)\\&\qquad -a\beta _*(p-2){\frac{\beta _*(\gamma -\beta _*)\psi _*g(\psi _*)-g'(\psi _*)\left| \nabla \psi _*\right| ^2 }{\left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{4-p}{2}}} } \psi _* +O(a^2\left\| g(\psi _*)\right\| _{C^1}). \end{aligned}$$

Finally,

$$\begin{aligned}&-\left( \left( (V_t-\beta _*V)^2+\left| \nabla ' V\right| ^2\right) ^{\frac{p-2}{2}}(V_t-\beta _*V)\right) _t\nonumber \\&\quad =a\left[ (\gamma ^2-\beta _* \gamma )\left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{p-2}{2}}g(\psi _*)\right. \nonumber \\&\qquad \left. +\beta _*(p-2){\frac{\beta _*(\gamma ^2-\beta _*\gamma )\psi _*g(\psi _*)-\gamma g'(\psi _*)\left| \nabla \psi _*\right| ^2 }{\left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{4-p}{2}}} } \psi _*\right] +O(a^2\left\| g(\psi _*)\right\| _{C^2}). \end{aligned}$$
(3.27)

Since

$$\begin{aligned}&\left( (V_t-\beta _*V)^2+\left| \nabla ' V\right| ^2\right) ^{\frac{p-2}{2}}\nabla 'V\\&\quad =\left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{p-2}{2}}(1-ag'(\psi _*))\\&\qquad \times \left[ 1+a(p-2){\frac{\beta _*(\gamma -\beta _*)\psi _*g(\psi _*)-g'(\psi _*)\left| \nabla \psi _*\right| ^2 }{\beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2} }\right] \nabla '\psi _*\\&\qquad +O(a^2\left\| g(\psi _*)\right\| _{C^1})\\&\quad =\left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{p-2}{2}}\nabla '\psi _*\\&\qquad +a\left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{p-2}{2}}\\&\qquad \times \left[ (p-2){\frac{\beta _*(\gamma -\beta _*)\psi _*g(\psi _*)-g'(\psi _*)\left| \nabla \psi _*\right| ^2 }{\beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2} } -g'(\psi _*)\right] \nabla '\psi _*\\&\qquad +O(a^2\left\| g(\psi _*)\right\| _{C^1}), \end{aligned}$$

we get similarly

$$\begin{aligned}&-div'\left( \left( (V_t-\beta _*V)^2+\left| \nabla ' V\right| ^2\right) ^{\frac{p-2}{2}}\nabla 'V\right) = -div'\left( \left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{p-2}{2}}\nabla '\psi _*\right) \nonumber \\&-a\,div'\left( \left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{p-2}{2}}\left[ (p-2){\frac{\beta _*(\gamma -\beta _*)\psi _*g(\psi _*)-g'(\psi _*)\left| \nabla \psi _*\right| ^2 }{\beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2} } -g'(\psi _*)\right] \nabla '\psi _*\right) \nonumber \\&\quad +O(a^2\left\| g(\psi _*)\right\| _{C^2}). \end{aligned}$$
(3.28)

Noting that

$$\begin{aligned} -div'\left( \left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{p-2}{2}}\nabla '\psi _*\right) \psi _* =\beta _*\Lambda _{\beta _*}\left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{p-2}{2}}\psi _*,\nonumber \\ \end{aligned}$$
(3.29)

we obtain

$$\begin{aligned}&e^{-\gamma t}{\mathcal Q}[V]\nonumber \\&\quad =\left[ (\gamma ^2-\beta _* \gamma )\left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{p-2}{2}}g(\psi _*)\right. \nonumber \\&\qquad \left. +\, \beta _*(p-2){\frac{\beta _*(\gamma ^2-\beta _*\gamma )\psi _*g(\psi _*)-\gamma g'(\psi _*)\left| \nabla \psi _*\right| ^2 }{(\beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2)^{\frac{4-p}{2}}} } \psi _*\right] \nonumber \\&\qquad -div'\left( \left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{p-2}{2}}\right. \nonumber \\&\qquad \times \left. \left[ (p-2){\frac{\beta _*(\gamma -\beta _*)\psi _*g(\psi _*)-g'(\psi _*)\left| \nabla \psi _*\right| ^2 }{\beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2} } -g'(\psi _*)\right] \nabla '\psi _*\right) \nonumber \\&\qquad -\Lambda _{\beta _*}\left( (\gamma -\beta _* )\left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{p-2}{2}}g(\psi _*)\right. \nonumber \\&\qquad \left. +\, \beta _*(p-2){\frac{\beta _*(\gamma -\beta _*)\psi _*g(\psi _*)-g'(\psi _*)\left| \nabla \psi _*\right| ^2 }{(\beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2)^{\frac{4-p}{2}}} }\psi _*\right) \nonumber \\&\qquad +e^{(\nu -\gamma ) t}\left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{q}{2}}\left[ 1+qa{\frac{\beta _*(\gamma -\beta _*)\psi _*g(\psi _*)-g'(\psi _*)\left| \nabla \psi _*\right| ^2 }{\beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2} }\right] \nonumber \\&\qquad +O(a\left\| g(\psi _*)\right\| _{C^2}). \end{aligned}$$
(3.30)

In this expression we have in particular

$$\begin{aligned}&-div'\left( \left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{p-2}{2}}\right. \nonumber \\&\qquad \left. \times \left[ (p-2){\frac{\beta _*(\gamma -\beta _*)\psi _*g(\psi _*)-g'(\psi _*)\left| \nabla \psi _*\right| ^2 }{\beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2} } -g'(\psi _*)\right] \nabla '\psi _*\right) \nonumber \\&\quad =(p-1)div'\left[ g'(\psi _*)\left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{p-2}{2}}\nabla \psi _*\right] \nonumber \\&\qquad -\beta _*div'\left( \left( \beta _*^2\psi _*^2\!+\!\left| \nabla '\psi _*\right| ^2\right) ^{\frac{p-4}{2}}\left[ (p-2)\beta _*\psi _*g'(\psi _*)\!+\!(p-2)(\gamma -\beta _*)g(\psi _*)\right] \psi _*\right) \nonumber \\&\quad =(p-1)g''(\psi _*)\left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{p-2}{2}}|\nabla \psi _*|^2\nonumber \\&\qquad +(p-1)g'(\psi _*)div'\left( \left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{p-2}{2}}\nabla \psi _*\right) \nonumber \\&\qquad -(p-2)\beta _*div'\left[ {\frac{\left( (\gamma -\beta _*) g(\psi _*)\psi _*+\beta _*g'(\psi _*)\psi ^2_*\right) }{\left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{4-p}{2}}} } \nabla '\psi _*\right] . \end{aligned}$$
(3.31)

Using the Eq. (3.26) satisfied by \(\psi _*\), it infers that

$$\begin{aligned}&-div'\left( \left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{p-2}{2}}\right. \nonumber \\&\qquad \times \left. \left[ (p-2){\frac{\beta _*(\gamma -\beta _*)\psi _*g(\psi _*)-g'(\psi _*)\left| \nabla \psi _*\right| ^2 }{\beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2} } -g'(\psi _*)\right] \nabla '\psi _*\right) \nonumber \\&\quad =(p-1)\left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{p-2}{2}}\left( g''(\psi _*)|\nabla '\psi _*|^2 -\beta _*\Lambda _{\beta _*}g'(\psi _*)\psi _*\right) \nonumber \\&\quad -(p-2)\beta _*div'\left[ {\frac{\left( (\gamma -\beta _*) g(\psi _*)\psi _*+\beta _*g'(\psi _*)\psi ^2_*\right) }{\left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{4-p}{2}}} } \nabla '\psi _*\right] . \end{aligned}$$
(3.32)

Plugging this identity into the expression (3.30), we obtain after some simplifications

$$\begin{aligned} e^{-\gamma t}{\mathcal Q}[V]=\left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{p-2}{2}}g(\psi _*){\mathcal Q}_1[V] +e^{(\nu -\gamma )t}R[V]+O(a\left\| g(\psi _*)\right\| _{C^2}), \end{aligned}$$
(3.33)

where

$$\begin{aligned} R[V]=e^{\nu t}\left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{q}{2}}\left[ 1+q{\frac{\beta _*(a'-\beta _*a)\psi _*g(\psi _*)-ag'(\psi _*)\left| \nabla \psi _*\right| ^2 }{\beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2} }\right] , \end{aligned}$$
(3.34)

and

$$\begin{aligned} {\mathcal Q}_1[V]&= (\gamma -\Lambda _{\beta _*})(\gamma -\beta _*)\left[ 1+(p-2){\frac{\beta ^2_*\psi ^2_*}{\beta ^2_*\psi ^2_*+|\nabla '\psi _*|^2} }\right] -(p-1)\beta _*\Lambda _{\beta _*}{\frac{\psi _*g'(\psi _*)}{g(\psi _*)} }\nonumber \\&\quad +\left[ (p-4)\beta _*\Lambda _{\beta _*}\psi _*-2\Delta '\psi _*\right] \left( \gamma -\beta _*\left( 1-{\frac{\psi _*g'(\psi _*)}{g(\psi _*)} }\right) \right) {\frac{\beta _*\psi _*}{\beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2} }\nonumber \\&\quad -(p-2)\left[ {\frac{\psi _*g'(\psi _*)}{g(\psi _*)} }\left( (\beta _*+1)\gamma -\beta _*\Lambda _{\beta _*}+\beta _*\right) +\gamma -\beta _*+\beta _*{\frac{\psi _*^2g''(\psi _*)}{g(\psi _*)} }\right] \nonumber \\&\quad \times {\frac{\left| \nabla '\psi _*\right| ^2}{\beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2} }+(p-1){\frac{g''(\psi _*)}{g(\psi _*)} }\left| \nabla '\psi _*\right| ^2. \end{aligned}$$
(3.35)

In this expression the difficult term to deal with is \(\left[ (p-4)\beta _*\Lambda _{\beta _*}\psi _*-2\Delta '\psi _*\right] \) since it has not a prescribed sign. However \(\Delta '\psi _*=O(\psi _*)\) by (6.19) in Appendix B.

Step 2: the perturbation method and the computation with \(g(\psi _*)=\psi _*\) With such a choice of function g

$$\begin{aligned} {\mathcal Q}_1[V]&=(\gamma -\Lambda _{\beta _*})(\gamma -\beta _*)\left[ 1+(p-2){\frac{\beta ^2_*\psi ^2_*}{\beta ^2_*\psi ^2_*+|\nabla '\psi _*|^2} }\right] -(p-1)\beta _*\Lambda _{\beta _*}\nonumber \\&\quad -(p-2)\left[ (\gamma -\Lambda _{\beta _*})\beta _*+2\gamma \right] {\frac{\left| \nabla '\psi _*\right| ^2}{\beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2} }+\gamma \ O(\psi _*^2). \end{aligned}$$
(3.36)

Equivalently

$$\begin{aligned} {\mathcal Q}_1[V]&=\left[ 1+(p-2){\frac{\beta ^2_*\psi ^2_*}{\beta ^2_*\psi ^2_*+|\nabla '\psi _*|^2} }\right] \left( \gamma ^2-(\Lambda _{\beta _*}+\beta _*)\gamma \right) \\&\quad -\gamma \left[ (p-2)(\beta _*+2){\frac{\left| \nabla '\psi _*\right| ^2}{\beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2} }+O(\psi ^2_*)\right] \end{aligned}$$

and finally

$$\begin{aligned} {\mathcal Q}_1[V]=\left[ 1+(p-2){\frac{\beta ^2_*\psi ^2_*}{\beta ^2_*\psi ^2_*+|\nabla '\psi _*|^2} }\right] \gamma [\gamma -(\Lambda _{\beta _*}+\beta _*+(p-2)(\beta _*+2))+O(\psi ^2_*)]. \end{aligned}$$
(3.37)

Using the fact that \(\beta _*>\frac{N-1}{p-1}\) if \(1<p<2\) and \(1<\beta _*<\frac{N-1}{p-1}\) if \(2<p<N\) (see Theorem 6.1 in Appendix B), we have

$$\begin{aligned} \Lambda _{\beta _*}+\beta _*+(p-2)(\beta _*+2)\ge \left\{ \begin{array}{ll} \Lambda _{\beta _*}+\beta _*(p-1)&{}\quad \text {if }p\ge 2\\ N+3(p-2)>N-3&{}\quad \text {if }1<p<2. \end{array}\right. \end{aligned}$$
(3.38)

When \(N=2\), we have explicitly \(\beta _*=\frac{1+2\sqrt{p^2-3p+3}}{3(p-1)}\) (see [9, Th 3.3]). Therefore for all \(N\ge 2\) and \(p>1\), there holds

$$\begin{aligned} \Lambda _{\beta _*}+\beta _*+(p-2)(\beta _*+2)>0. \end{aligned}$$
(3.39)

We fix \(\epsilon _0>0\) such that, whenever \(\psi _*\le \epsilon _0\), there holds

$$\begin{aligned} \Lambda _{\beta _*}+\beta _*+(p-2)(\beta _*+2)+O(\psi ^2_*)>{\frac{1}{2} }\left( \Lambda _{\beta _*}+\beta _*+(p-2)(\beta _*+2)\right) . \end{aligned}$$
(3.40)

If we fix \(\gamma _0>0\) such that

$$\begin{aligned} \gamma _0<\min \left\{ {\frac{1}{2} }\left( \Lambda _{\beta _*}+\beta _*+(p-2)(\beta _*+2)\right) ,\nu ,\beta _*\right\} , \end{aligned}$$
(3.41)

we obtain

$$\begin{aligned} {\mathcal Q}_1[V]\le -\min \{1,p-1\}\gamma m^2\quad \forall \, 0<\gamma \le \gamma _0, \end{aligned}$$
(3.42)

whenever \(\psi _*\le \epsilon _0\), for some m depending only on p, q and N (through \(\psi _*\) and \(\nu \)), which, in the same range of value of \(\psi _*\), yields to

$$\begin{aligned} \left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{p-2}{2}}g(\psi _*){\mathcal Q}_1[V]\le -c_{17} \psi _*\quad \forall \, 0<\gamma \le \gamma _0, \end{aligned}$$
(3.43)

for some \(c_{17}>0\) depending on Npq. This estimate is valid whatever is \(p>1\), but only in a neighborhood of \(\psi _*=0\). If we replace \(g(\psi _*)=\psi _*\) by \(g_k(\psi _*)=\psi _*e^{-k\psi _*}\) for \(0<k<1\), and denote by \({\mathcal Q}_{1,k}[V]\) the corresponding expression of \({\mathcal Q}_1[V]\) which becomes now \({\mathcal Q}_{1,0}[V]\). We define similarly \({\mathcal Q}_k[V]\), and \({\mathcal Q}[V]\) becomes \({\mathcal Q}_0[V]\). Since \(g'_k(\psi _*)=e^{-k\psi _*}-kg_k(\psi _*)\) and \(g''_k=-2ke^{-k\psi _*}+k^2g_k(\psi _*)\), we obtain

$$\begin{aligned} {\mathcal Q}_{1,k}[V]&={\mathcal Q}_{1,0}[V]+k(p-1)\beta _*\Lambda _{\beta _*}\psi _*+(p-1)\left( -{\frac{2k}{\psi _*} }+k^2\right) \left| \nabla '\psi _*\right| ^2\nonumber \\&\quad +(2-p)\beta _*\left( -2k+k^2\right) \psi _*+O(\psi _*^2) \end{aligned}$$
(3.44)

Notice that \(\nabla '\psi _*\) vanishes only at the North pole \(\mathbf{e}_N\), thus there exists \(k_0\in (0,1]\) such that

$$\begin{aligned} k(1-p)\beta _*\Lambda _{\beta _*}\psi _*+(p-1)\left( {\frac{2k}{\psi _*} }-k^2\right) \left| \nabla '\psi _*\right| ^2\ge {\frac{1}{2} }(2-p)_+\beta _*\left( -2k+k^2\right) \psi _*\quad \forall k{\le } k_0 \end{aligned}$$

whenever \(\psi _*\le \epsilon _0\) which yields to

$$\begin{aligned} \left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{p-2}{2}}g_k(\psi _*){\mathcal Q}_{1,k}[V]\le -c_{18}k\quad \forall k\le k_0 \end{aligned}$$
(3.45)

for some \(c_{13}=c_{13}(N,p,q,\epsilon _0)\). There exists \(c_{14}=c_{14}(N,p,q)>0\) such that

$$\begin{aligned} \left( \beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2\right) ^{\frac{q}{2}}\left[ 1+qe^{\gamma t}{\frac{\beta _*(\gamma -\beta _*)\psi _*g_k(\psi _*)-g_k'(\psi _*)\left| \nabla \psi _*\right| ^2 }{\beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2} }\right] \le c_{14}\nonumber \\ \end{aligned}$$
(3.46)

in \(S^{N-1}_+\times (-\infty ,\ln \delta ]\). Moreover

$$\begin{aligned} O(a\left\| g(\psi _*)\right\| _{C^2})\le e^{\gamma t}\tilde{c}_k \end{aligned}$$
(3.47)

for some \(\tilde{c}_k=\tilde{c}_k(N,p,q)>0\). We derive from (3.45)–(3.47)

$$\begin{aligned} \begin{array}{ll} e^{-\gamma t}{\mathcal Q}_k[V]\le -c_{13}k+c_{14}e^{(\nu -\gamma ) t}+e^{\gamma t}\tilde{c}_k\qquad \forall k\le k_0 \end{array} \end{aligned}$$
(3.48)

Thus there exists \(T_k\le \ln \delta \) such that \({\mathcal Q}_k[V]\le 0\), for all \(t\le T_k\) and provided \(\psi _*\le \epsilon _0\). This local estimate will be used in the construction of the subsolution when \(p\ge 2\).

Step 3: the case \(1<p<2\) Since the function \(\psi ^*\) depends only on the azimuthal angle \(\theta \in (0;\frac{\pi }{2}]\) we will write \(\psi _*(\sigma )=\psi _*(\theta )\) and \(\nabla '\psi _*(\sigma )=\psi _{*\theta }(\theta )\mathbf{n}\) where \(\mathbf{n}\) is the downward unit vector tangent to \(S^{N-1}\) in the hyperplane going through \(\sigma \) and the poles. From (6.8),

$$\begin{aligned} (p-4)\beta _*\Lambda _{\beta _*}\psi _*-2\Delta '\psi _*=(p-2)\left( \beta _*\Lambda _{\beta _*}\psi _*+2{\frac{\beta _*^2\psi _*+{\psi _{*}}_{\theta \theta }}{\beta _*^2\psi ^2_{*}+\psi ^2_{*\theta }} }\right) , \end{aligned}$$
(3.49)

since \(\psi ^{\,2}_{*\theta }=\left| \nabla '\psi _*\right| ^2\) and thus

$$\begin{aligned}&\left( (p-4)\beta _*\Lambda _{\beta _*}\psi _*-2\Delta '\psi _*\right) {\frac{\beta _*\gamma \psi _*}{\beta _*^2\psi ^2_{*}+\psi ^2_{*\theta }} }\nonumber \\&\quad =(p-2)\gamma \left( \Lambda _{\beta _*}{\frac{\beta _*^2\psi _*^2}{\beta _*^2\psi ^2_{*}+\psi ^2_{*\theta }} }+2\beta _*{\frac{\beta _*^2\psi _*^2+{\psi _{*}}_ {\theta \theta }\psi _*}{(\beta _*^2\psi ^2_{*}+\psi ^{\,2}_{*\theta })^2} } \right) . \end{aligned}$$
(3.50)

From Theorem 6.1-Step 4 in Appendix B, we know that \(\beta _*^2\psi _*+{\psi _{*}}_{\theta \theta }\ge 0\), thus the contribution of this term to \({\mathcal Q}_1[V]\) is nonpositive. We replace this expression in \({\mathcal Q}_1[V]\) with \(g(\psi _*)=\psi _*\) and obtain

$$\begin{aligned} {\mathcal Q}_1[V]&=(\gamma -\Lambda _{\beta _*})(\gamma -\beta _*)\left( 1+(p-2){\frac{\beta _*^2\psi ^2_{*}}{\beta _*^2\psi ^2_{*}+\psi ^2_{*\theta }} }\right) -\Lambda _{\beta _*}\beta _*(p-1)\nonumber \\&\quad +(p-2)\gamma \Lambda _{\beta _*}{\frac{\beta _*^2\psi ^2_{*}}{\beta _*^2\psi ^2_{*}+\psi ^2_{*\theta }} }-(p-2)\left( (\beta _*+2)\gamma -\Lambda _{\beta _*}\beta _*\right) {\frac{\psi ^2_{*\theta }}{\beta _*^2\psi ^2_{*}+\psi ^2_{*\theta }} }\nonumber \\&\quad +2\beta _*(p-2){\frac{\beta _*^2\psi _*^2+{\psi _{*}}_{\theta \theta }\psi _*}{(\beta _*^2\psi ^2_{*}+\psi ^2_{*\theta })^2} }\gamma \nonumber \\&\le \gamma \left( 1+(p-2){\frac{\beta _*^2\psi ^2_{*}}{\beta _*^2\psi ^2_{*}+\psi ^2_{*\theta }} }\right) \left( \gamma -\Lambda _{\beta _*}-\beta _*\right) \nonumber \\&\quad -(p-2)\gamma {\frac{(\beta _*+2))\psi ^2_{*\theta }-\Lambda _{\beta _*}\beta _*^2\psi _*^2}{\beta _*^2\psi ^2_{*}+\psi ^2_{*\theta }} }\nonumber \\&\le \gamma \left( 1+(p-2){\frac{\beta _*^2\psi ^2_{*}}{\beta _*^2\psi ^2_{*}+\psi ^2_{*\theta }} }\right) \nonumber \\&\quad \times \left( \gamma -\left( \Lambda _{\beta _*}\!+\beta _*+(p-2){\frac{(\beta _*+2)\psi ^2_{*\theta }-\Lambda _{\beta _*}\beta _*^2\psi _*^2}{(p-1)\beta _*^2\psi ^2_{*}+\psi ^2_{*\theta }} }\right) \!\right) . \end{aligned}$$
(3.51)

We can write

$$\begin{aligned}&\Lambda _{\beta _*}+\beta _*+(p-2){\frac{(\beta _*+2)\psi ^2_{*\theta }-\Lambda _{\beta _*}\beta _*^2\psi _*^2}{(p-1)\beta _*^2\psi ^2_{*}+\psi ^2_{*\theta }} }\nonumber \\&\quad ={\frac{\left( \Lambda _{\beta _*}+(p-1)\beta _*\right) \beta _*^2\psi _*^2+\left( \Lambda _{\beta _*}+\beta _*(p-1)+2(p-2)\right) \psi ^2_{*\theta }}{(p-1)\beta _*^2\psi ^2_{*}+\psi ^2_{*\theta }} }\nonumber \\&\quad \ge c_{15}\left( \Lambda _{\beta _*}+\beta _*(p-1)+2(p-2)\right) \end{aligned}$$
(3.52)

for some positive constant \(c_{15}\). This expression \(\Lambda _{\beta _*}+\beta _*(p-1)+2(p-2)\) is always positive: obviously if \(N\ge 3\) and by using the explicit expression of \(\beta _*\) if \(N=2\). Thus there exists \(\gamma _0\) and \(c_{16}>0\) such that \({\mathcal Q}_1[V]<-c_{16}\) for \(0<\gamma \le \gamma _0\). The perturbation method of Step 2, is valid in the whole range of values of \(\psi _*\) and we derive from (3.42)–(3.43) that (3.48) holds for all \(k\le k_0\) and \(t\le T_k\). Therefore \({\mathcal Q}_k[V]\le 0\).

Step 4: the case \(p\ge 2\) For \(c>0\) to be fixed and \(\psi _*\ge \epsilon _0\), \(\gamma \in (0,\gamma _0]\), we take \(g(\psi _*)=c\psi _*^{1-\frac{\gamma }{\beta _*}}\). Then we derive from (3.35):

$$\begin{aligned} {\mathcal Q}_1[V]&= (\gamma -\Lambda _{\beta _*})(\gamma -\beta _*){\frac{(p-1)\beta ^2_*\psi ^2_*+|\nabla '\psi _*|^2}{\beta ^2_*\psi ^2_*+|\nabla '\psi _*|^2} } -(p-1)\beta _*\Lambda _{\beta _*}\left( 1-{\frac{\gamma }{\beta _*} }\right) \nonumber \\&\quad -(p-1){\frac{\gamma (\beta _*-\gamma )}{\beta ^2_*} }\psi _*^{-1-\frac{\gamma }{\beta _*}}\left| \nabla '\psi _*\right| ^2 -(p-2)(\beta _*-\gamma )(\gamma -\Lambda _{\beta _*})\nonumber \\&\quad \times {\frac{\left| \nabla '\psi _*\right| ^2}{\beta _*^2\psi _*^2+\left| \nabla '\psi _*\right| ^2} }\nonumber \\&=(1-p)\left[ \gamma (\beta _*-\gamma )+{\frac{\gamma (\beta _*-\gamma )}{\beta ^2_*} }\psi _*^{-1-\frac{\gamma }{\beta _*}}\left| \nabla '\psi _*\right| ^2\right] . \end{aligned}$$
(3.53)

For \(k\le k_0\) we fix c such that \(c\epsilon _0^{1-\frac{\gamma }{\beta _*}}=\epsilon _0e^{-k\epsilon _0}\Longleftrightarrow c=\epsilon _0^{\frac{\gamma }{\beta _*}}e^{-k\epsilon _0}\) and we define g by

$$\begin{aligned} g(\psi _*)=\min \left\{ \psi _*e^{-k\psi _*},\epsilon _0^{\frac{\gamma }{\beta _*}}e^{-k\epsilon _0}\psi _*^{1-\frac{\gamma }{\beta _*}}\right\} {=}\left\{ \begin{array}{ll} \psi _*e^{-k\psi _*}&{}\quad \text {if }0\le \psi _*\le \epsilon _0\\ \epsilon _0^{\frac{\gamma }{\beta _*}}e^{-k\epsilon _0}\psi _*^{1-\frac{\gamma }{\beta _*}} &{}\quad \text {if }\epsilon _0\le \psi _*\le 1, \end{array}\right. \qquad \end{aligned}$$
(3.54)

and we set \(V(t,\sigma )=\psi ^*(\sigma )-a(t)g(\psi _*(\sigma ))\) with \((t,\sigma )\in (-\infty ,T_k]\times S^{N-1}_+\) and define \(\tilde{u}(r,\sigma )=r^{-\beta _*}(\psi ^*(\sigma )-a(\ln r)g(\psi _*(\sigma )))\) accordingly for \((r,\sigma )\in (-\infty ,e^{T_k}]\times S^{N-1}_+\). Since \(\psi _*\) is a decreasing function the coincidence set \(\{\sigma \in S^{N-1}_+:\psi _*(\sigma )=\epsilon _0\}\) is a circular cone \(\Sigma _{\theta _0}\) with vertex 0, axis \(\mathbf{e}_N\) and angle \(\theta _0\). We set \(R_0=e^{T_k}\)

$$\begin{aligned} \begin{array} {ll}\Gamma _1=\left\{ x=(r,\theta )\in B^+_{R_0}:\theta _0< \theta <\frac{\pi }{2}\right\} =\left\{ (r,\sigma )\in [0,R_0)\times S^{N-1}_+:0<\psi _*(\sigma )<\epsilon _0\right\} ,\\ \Gamma _2=\left\{ x=(r,\theta )\in B^+_{R_0}:0< \theta <\theta _0\right\} =\left\{ (r,\sigma )\in [0,R_0)\times S^{N-1}_+:\epsilon _0<\psi _*(\sigma )<1\right\} , \end{array} \end{aligned}$$

and define

$$\begin{aligned} \tilde{u}(r,\sigma )&=r^{-\beta _*}\left( \psi _*(\sigma )-r^\gamma g(\psi _*(\sigma ))\right) \\&=\left\{ \begin{array}{ll} u_1(r,\sigma )=r^{-\beta _*}(1-r^\gamma e^{-k\psi _*(\sigma )})\psi _*(\sigma )&{}\quad \text {if }(r,\theta )\in \Gamma _1\\ u_2(r,\sigma )=r^{-\beta _*}\left( 1-r^\gamma \epsilon _0^{\frac{\gamma }{\beta _*}}e^{-k\epsilon _0}(\psi _*(\sigma ))^{1-\frac{\gamma }{\beta _*}}\right) \psi _*(\sigma )&{}\quad \text {if }(r,\theta )\in \Gamma _2. \end{array}\right. \end{aligned}$$

The function \(\tilde{u}\) is a subsolution separately on \(\Gamma _1\) and \(\Gamma _2\) and is Lipschitz continuous in \( \overline{\Omega }{\setminus }\{0\}\). If we denote by \(g_1\) and \(g_2\) the restriction of g to \(\Gamma _1\) and \(\Gamma _2\) respectively, that is to \(\Omega _1\) and \(\Omega _2\), then \(g'_1(\epsilon _0)>g'_2(\epsilon _0)>0\). Let \(\zeta \in C^{1}_c(B^+_{R_0})\) which vanishes in neighborhoods of 0 and \(\partial B^+_{R_0}\), \(\zeta \ge 0\), then

$$\begin{aligned} {\int _{\Gamma _i}^{}}\left| \nabla \tilde{u}\right| ^{p-2}\nabla \tilde{u}.\nabla \zeta dx+{\int _{\Omega _i}^{}}\left| \nabla \tilde{u}\right| ^{q}\zeta dx \le {\int _{\Sigma _{\theta _0}}^{}}\left| \nabla u_i\right| ^{p-2}\partial _{\mathbf{n}_i} u_i\zeta dS, \end{aligned}$$
(3.55)

where \(\mathbf{n}_i\) is the normal unit vector on \(\Sigma _{\theta _0}\) outward from \(\Gamma _i\). Actually, \(\mathbf{n}_2=-\mathbf{n}_1=\mathbf{n}\) thus

$$\begin{aligned} \nabla \tilde{u}=\tilde{u}_r\mathbf{e}+r^{-\beta _*-1}(1-r^\gamma g'(\psi _*))\nabla '\psi _*=\tilde{u}_r\mathbf{e}+r^{-\beta _*-1}(1-r^\gamma g'(\psi _*))\psi _{*\theta }\,\mathbf{n}. \end{aligned}$$

and on \(\Sigma _{\theta _0}\),

$$\begin{aligned} \nabla \tilde{u}=\left\{ \begin{array} {ll}\tilde{u}_r\mathbf{e}-r^{-\beta _*-1}(1-r^\gamma g_1'(\epsilon _0))\psi _{*\theta }\,\mathbf{n}&{}\quad \text { in }\Gamma _1\\ \tilde{u}_r\mathbf{e}+r^{-\beta _*-1}(1-r^\gamma g_2'(\epsilon _0))\psi _{*\theta }\,\mathbf{n}&{}\quad \text { in }\Gamma _2 \end{array}\right. \end{aligned}$$

Therefore

$$\begin{aligned}&\left| \nabla u_1\right| ^{p-2}\partial _{\mathbf{n}_1}u_1\nonumber \\&\quad =-r^{-\beta _*-1}(1-r^\gamma g_1'(\epsilon _0))\left( \tilde{u}^2_r+r^{-2\beta _*-2}(1-r^\gamma g_1'(\epsilon _0))^2\psi ^2_{*\theta }\right) ^{\frac{p-2}{2}}\psi _{*\theta }\quad \text { in }\Gamma _1 \end{aligned}$$

and

$$\begin{aligned}&\left| \nabla u_2\right| ^{p-2}\partial _{\mathbf{n}_2}u_2\nonumber \\&\quad =r^{-\beta _*-1}(1-r^\gamma g_2'(\epsilon _0))\left( \tilde{u}^2_r+r^{-2\beta _*-2}(1-r^\gamma g_2'(\epsilon _0))^2\psi ^2_{*\theta }\right) ^{\frac{p-2}{2}}\psi _{*\theta }\quad \text { in }\Gamma _2. \end{aligned}$$

By adding the two inequalities (3.55)

$$\begin{aligned} {\int _{\Omega }^{}}\left| \nabla \tilde{u}\right| ^{p-2}\nabla \tilde{u}.\nabla \zeta dx+{\int _{\Omega }^{}}\left| \nabla \tilde{u}\right| ^{q}\zeta dx \le {\int _{\Sigma _{\theta _0}}^{}}\left( \left| \nabla u_1\right| ^{p-2}\partial _{\mathbf{n_1}} u_1+\left| \nabla u_2\right| ^{p-2}\partial _{\mathbf{n_2}} u_2\right) \zeta dS. \end{aligned}$$
(3.56)

By monotonicity of the function \(X\mapsto \left( \tilde{u}^2_r+X^2\right) ^{\frac{p}{2}}\) and since

$$\begin{aligned} r^{-\beta _*-1}(1-r^\gamma g_2'(\epsilon _0))\ge r^{-\beta _*-1}(1-r^\gamma g_1'(\epsilon _0))\ge 0, \end{aligned}$$

we derive

$$\begin{aligned}&r^{-\beta _*-1}(1-r^\gamma g_2'(\epsilon _0))\left( \tilde{u}^2_r+r^{-2\beta _*-2}(1-r^\gamma g_2'(\epsilon _0))^2\psi ^2_{*\theta }\right) ^{\frac{p-2}{2}}\\&\quad \ge r^{-\beta _*-1}(1-r^\gamma g_1'(\epsilon _0))\left( \tilde{u}^2_r+r^{-2\beta _*-2}(1-r^\gamma g_1'(\epsilon _0))^2\psi ^2_{*\theta }\right) ^{\frac{p-2}{2}} \end{aligned}$$

We derive that the right-hand side of (3.56) is nonpositive because \(\psi _{*\theta }\le 0\), and therefore \(\tilde{u}\) is a positive subsolution of (1.1) in \(B^+_{R_0}\) dominated by \(\Psi _*\) and satisfying (3.24). \(\square \)

.

Proof of Theorem 3.6

Let \(M=\max \{\Psi _*(x):x\in \partial B^+_{R_0}\}\), then \(M=R^{-\beta _*}_0\). The function \(u^{*}\) defined by

$$\begin{aligned} u^{*}(x)=\left\{ \begin{array} {ll} (\tilde{u}(x)-M)_+&{}\quad \text {if } x\in B^+_{R_0}\\ 0&{}\quad \text {if } x\in \Omega {\setminus } B^+_{R_0}, \end{array}\right. \end{aligned}$$

is indeed a subsolution of (1.1) in whole \(\Omega \) where it satisfies \(u^{*}\le \Psi _*\) and it vanishes on \(\partial \Omega {\setminus }\{0\}\). Since \(\Phi _*\) is a positive p-harmonic function in \(\Omega \) which vanishes on \(\partial \Omega {\setminus }\{0\}\) and satisfies (3.20), it is supersolution of (1.1) and therefore it dominates \(u^{*}\). Therefore there exists a solution u of (1.1) in \(\Omega \) which vanishes on \(\partial \Omega {\setminus }\{0\}\) and satisfies \(u^{*}\le u\le \Phi _*\). This implies that (3.19) holds with \(k=1\) and we conclude with Lemma 3.8. This ends the proof of Lemma 3.9. \(\square \)

When \(p=N\) the statement of Theorem 3.6 holds without the flatness assumption on \(\partial \Omega \). The proof of the next theorem is an easy adaptation to the one of Theorem 3.6, provided Lemmas 3.7, 3.8 and 3.9 are modified accordingly.

Theorem 3.10

Assume \(N-1<q<N-\frac{1}{2}\) and \(\Omega \) be a bounded \(C^2\) domain such that \(0\in \partial \Omega \). Then for any \(k>0\) there exists a unique positive solution \(u:=u_k\) of (3.17) in \(\Omega \), which belongs to \(C^1(\overline{\Omega }{\setminus }\{0\})\), vanishes on \(\partial \Omega {\setminus }\{0\}\) and satisfies uniformly with respect to \(\sigma \in S^{N-1}_+\)

$$\begin{aligned} \lim _{\begin{array}{c} x\rightarrow 0\\ x/\left| x\right| \rightarrow \sigma \end{array}}\left| x\right| u_k(x)=k\psi _*(\sigma ). \end{aligned}$$
(3.57)

Since \(p=N\), then \(\beta _*=1\) and \(\psi _*(\sigma )=\frac{x_N}{\left| x\right| }=\cos \theta _{N-1}\) with the identification of \(\sigma \) and \(\theta _{N-1}:=\theta \). In a more intrinsic manner (3.57) can be written under the form

$$\begin{aligned} \lim _{\begin{array}{c} x\rightarrow 0\\ x\in \Omega \end{array}}|x|^2\frac{u_k(x)}{ d(x)}=k. \end{aligned}$$
(3.58)

We recall that if \(\omega \in \mathbb {R}^N\) and \({\mathcal I}_\omega \) denotes the inversion of center \(\omega \) and power 1, i.e. \({\mathcal I}_\omega (x)=\omega +\frac{x-\omega }{|x-\omega |^2}\), then \(\tilde{u}=u\circ {\mathcal I}_\omega \) satisfies (3.18).

Lemma 3.11

Assume \(\Omega \) be a bounded \(C^2\) domain such that \(0\in \partial \Omega \). Then there exists a unique N-harmonic function \(\Phi _*\) in \(\Omega \), which vanishes on \(\partial \Omega {\setminus }\{0\}\) and satisfies

$$\begin{aligned} \lim _{\begin{array}{c} x\rightarrow 0\\ x/\left| x\right| \rightarrow \sigma \end{array}}\left| x\right| \Phi _*(x)=\psi _*(\sigma ), \end{aligned}$$
(3.59)

uniformly with respect to \(\sigma \in S^{N-1}_+\).

Proof

Uniqueness is standard. Let \(\omega =-\mathbf{e}_N\in \overline{\Omega }^c\), with the notations of the proof of Theorem 3.5, \(\omega '=-\omega \), \(a=-\frac{1}{2}\mathbf{e}_N\) and \(a'=-a\). We can assume that the balls \(B_{\frac{1}{2}}(a)\) and \(B_{\frac{1}{2}}(a')\) are tangent to \(\partial \Omega \) at 0 and respectively subset of \(\Omega ^c\) and \(\Omega \). The function \(x\mapsto \Psi (x)=-\frac{x_N}{|x|^2}\) which is N-harmonic in \(\mathbb {R}^N_-\) and vanishes on \(\partial \mathbb {R}^{N}_-{\setminus }\{0\}\) is transformed by the inversion \({\mathcal I}_{\omega '}\) of center \(\omega '\) and power 1 into the function \(\Psi _{\omega '}=\Psi \circ {\mathcal I}_\omega \) which is positive and N-harmonic in \(B_{\frac{1}{2}}(a')\) and vanishes on \(\partial B_{\frac{1}{2}}(a'){\setminus }\{0\}\). The function \(\hat{\Psi }=-\Psi \) which is N-harmonic in \(\mathbb {R}^N_+\) and vanishes on \(\partial \mathbb {R}^{N}_+{\setminus }\{0\}\) is transformed by the inversion \({\mathcal I}_{\omega }\) of center \(\omega \) and power 1 into the function \(\Psi _\omega =\hat{\Psi }\circ {\mathcal I}_{\omega }\) which is positive and N-harmonic in \(B^c_{\frac{1}{2}}(a)\) and vanishes on \(\partial B_{\frac{1}{2}}(a){\setminus }\{0\}\). For \(\epsilon >0\) we denote by \(\Phi _\epsilon \) the solution of

$$\begin{aligned} -\Delta _N\Phi _\epsilon&=0\qquad \;\text {in } \Omega \cap B^c_\epsilon \nonumber \\ \Phi _\epsilon&=0\qquad \;\text {in } (B^c_{\frac{1}{2}}(a')\cap \partial B_\epsilon )\cup (\partial \Omega \cap B^c_\epsilon )\nonumber \\ \Phi _\epsilon&=\Psi _{\omega '}\quad \text {in }B_{\frac{1}{2}}(a')\cap \partial B_\epsilon . \end{aligned}$$
(3.60)

If \(0<\epsilon '<\epsilon \), \(\Phi _{\epsilon '}\ge \Psi _{\omega '}\) in \(B_{\frac{1}{2}}(a')\cap \partial B_\epsilon \), thus \(\Phi _{\epsilon '}\ge \Phi _{\epsilon '}\) in \(\Omega \cap B^c_\epsilon \). We also denote by \(\hat{U}_\epsilon \) the solution of

$$\begin{aligned} -\Delta _N\hat{\Phi }_\epsilon&=0\qquad \;\text {in } \Omega \cap B^c_\epsilon \nonumber \\ \hat{\Phi }_\epsilon&=0\qquad \;\text {in } \partial \Omega \cap B^c_\epsilon \nonumber \\ \hat{\Phi }_\epsilon&=\Psi _{\omega }\quad \;\text {in } \Omega \cap \partial B^c_\epsilon . \end{aligned}$$
(3.61)

In the same way as above

$$\begin{aligned} 0<\epsilon '<\epsilon \Longrightarrow \hat{\Phi }_{\epsilon '}\le \hat{\Phi }_\epsilon \quad \text{ in } \,\Omega \cap \partial B^c_\epsilon \end{aligned}$$

Using the explicit form of \(\Psi \), \({\mathcal I}_\omega :x\mapsto \omega +\frac{x-\omega }{|x-\omega |^2}\) and \({\mathcal I}_{\omega '}:x\mapsto \omega '+\frac{x-\omega '}{|x-\omega '|^2}\) we see that

thus

$$\begin{aligned} \Phi _\epsilon \le {\frac{1+\epsilon }{1-\epsilon } }\hat{\Phi }_\epsilon \quad \text {in }\Omega \cap B^c_\epsilon . \end{aligned}$$

Letting \(\epsilon \rightarrow 0\) we conclude that \(\Phi _\epsilon \) converges uniformly in \(\overline{\Omega }{\setminus } \{0\}\) to \(\Phi _*\) which vanishes on \(\partial \Omega {\setminus }\{0\}\) and satisfies (3.59). \(\square \)

The proof of the next statement is similar to the one of Lemma 3.8 up to some minor modifications, so we omit it.

Lemma 3.12

Let the assumptions on q and \(\Omega \) of Theorem 3.10 be satisfied. If for some \(k>0\) there exists a solution \(u_{k}\) of (3.17) in \(\Omega \), which belongs to \(C^1(\overline{\Omega }{\setminus }\{0\})\), vanishes on \(\partial \Omega {\setminus }\{0\}\) and satisfies (3.57), then for any \(k>0\) there exists such a solution.

Lemma 3.13

Under the assumptions of Theorem 3.10 there exists a Lipschitz continuous nonnegative subsolution \(\tilde{u}\) of (3.17) in \(\Omega \) which vanishes on \(\partial \Omega {\setminus }\{0\}\), is smaller than \(\Phi _*\) and satisfies

$$\begin{aligned} \lim _{\begin{array}{c} x\rightarrow 0\\ x/\left| x\right| \rightarrow \sigma \end{array}}\left| x\right| \tilde{u}(x)=\sigma , \end{aligned}$$
(3.62)

uniformly with respect to \(\sigma \in S^{N-1}_+\).

Proof

Let \(\tau >0\) to be fixed and let w be the solution of

$$\begin{aligned} -\Delta _Nw+|\nabla w|^q=0\quad \text {in }B_2^- \end{aligned}$$
(3.63)

which vanishes on \(\partial B_2^-{\setminus }\{0\}\) and satisfies

$$\begin{aligned} \lim _{\begin{array}{c} x\rightarrow 0\\ {x/|x|}\rightarrow \sigma \end{array}}|x|w(x)=\sigma \end{aligned}$$
(3.64)

in the \(C^1\)-topology of \(S^{N-1}_-\). Its existence follows from Theorem 3.6 and this function is dominated by the N-harmonic function \(\Phi _*\) corresponding to this domain, obtained in Lemma 3.11. By \({\mathcal I}_{\omega '}\), the half-ball \(B_2^-\) is transform into the lunule \(G=B_{\frac{1}{2}}(a'){\setminus } B_{\frac{2}{3}}(\frac{4}{3}\omega ')\) and \(\tilde{w}=w\circ {\mathcal I}_{\omega '}\) satisfies

$$\begin{aligned} -\Delta _N\tilde{w}+|x-\omega '|^{2(q-N)}|\nabla \tilde{w}|^q=0\quad \text {in }G. \end{aligned}$$
(3.65)

Since \(|x-\omega '|\le 1\) in G, \(-\Delta _N\tilde{w}+|\nabla \tilde{w}|^q\le 0\) in G. We extend \(\tilde{w}\) by 0 in \(\Omega {\setminus } G\) and the resulting function \(\tilde{u}\) is a subsolution of (3.17) in \(\Omega \) which vanishes on \(\partial \Omega {\setminus }\{0\})\), is smaller than the N-harmonic function \(\Phi _*\) obtained in Lemma 3.11, and satisfies (3.62). \(\square \)

4 Classification of boundary singularities

We assume that \(\Omega \subset \mathbb {R}^N\) is a \(C^2\) domain and \(0\in \partial \Omega \). Furthermore, in order to avoid extremely technical computations, we shall assume either that \(\partial \Omega \) is flat near 0 or \(p=N\). We suppose that the tangent plane to \(\partial \Omega \) at 0 is \(\partial \mathbb {R}^N_+=\{x=(x',0)\}\) and the normal inward unit vector at 0 is \(\mathbf{e}_N\), therefore \(\mathbf{n}=-\mathbf{e}_N\) in the sequel. We denote by \(\omega _{s^{{N-1}}_+}\) the unique positive solution of (3.1) in \(S^{N-1}_+\) and by \(U_{s^{{N-1}}_+}\) the corresponding singular solution of (1.1) in \(\mathbb {R}^N_+\) defined by

$$\begin{aligned} U_{s^{{N-1}}_+}(x)=\left| x\right| ^{-\beta _q}\omega _{s^{{N-1}}_+}\left( \frac{x}{\left| x\right| }\right) . \end{aligned}$$
(4.1)

We recall that \({\psi _*}\) is the unique positive solution of (3.2) with maximum 1 and \(\Psi _*\) the corresponding p-harmonic function

$$\begin{aligned} \Psi _*(x)=\left| x\right| ^{-\beta _*}\psi _*\left( \frac{x}{\left| x\right| }\right) . \end{aligned}$$
(4.2)

4.1 The case \(1<p<N\)

The first statement points out the link between weak and strong singularities.

Proposition 4.1

Under the assumptions of Theorem 3.6 there exists \(\lim _{k\rightarrow \infty }u_k=u_\infty \) which is the unique element of \(C(\overline{\Omega }{\setminus }\{0\})\cap C^1(\Omega )\) which vanishes on \(\partial \Omega {\setminus }\{0\}\), satisfies (1.1) in \(\Omega \) and

$$\begin{aligned} \lim _{x\rightarrow 0}\frac{u_{\infty }(x)}{U_{s^{{N-1}}_+}(x)}=1. \end{aligned}$$
(4.3)

Proof

Uniqueness follows from (4.3) and the maximum principle. For existence, since the mapping \(k\mapsto u_k\) is increasing and \(u_k\le U_{s^{{N-1}}_+}\), there exists \(\lim _{k\rightarrow \infty }u_k:=u_\infty \le U_{s^{{N-1}}_+}\) and \(u_\infty \in C(\overline{\Omega }{\setminus }\{0\})\cap C^1(\Omega )\). It vanishes on \(\partial B^+_\delta {\setminus }\{0\}\) and satisfies (1.1) in \(B^+_\delta \). In order to take into account the domain \(B^+_\delta \) in the notations, we set \(u_k=u_{k,\delta }\). Since the mapping \(\delta \mapsto u_{k,\delta }\) is also increasing and \(u_{k,\delta }\le k{\Psi _*}\), there also exists \(\lim _{\delta \rightarrow \infty }u_{k,\delta }:=u_{k,\infty }\le k{\Psi _*}\) Then, for all \(\ell >0\),

$$\begin{aligned} T_\ell [u_{k,\delta }](x)=\ell ^{\beta _q}u_{k,\delta }(\ell x)=u_{k\ell ^{\beta _q},\ell ^{-1}\delta }(x). \end{aligned}$$
(4.4)

Letting \(k\rightarrow \infty \), we obtain

$$\begin{aligned} T_\ell [u_{\infty ,\delta }](x)=\ell ^{\beta _q}u_{\infty ,\delta }(\ell x)=u_{\infty ,\ell ^{-1}\delta }(x), \end{aligned}$$
(4.5)

and letting \(\delta \rightarrow \infty \), we obtain

$$\begin{aligned} T_\ell [u_{\infty ,\infty }](x)=\ell ^{\beta _q}u_{\infty ,\infty }(\ell x)=u_{\infty ,\infty }(x). \end{aligned}$$
(4.6)

This implies that

$$\begin{aligned} u_{\infty ,\infty }(r,\sigma )=r^{-\beta _q}\omega '(\sigma ), \end{aligned}$$
(4.7)

and \(\omega '\) is a positive solution of problem (3.1). Therefore \(\omega '=\omega _{s^{{N-1}}_+}\) by Theorem 3.2. If we let \(\ell \rightarrow 0\) in (4.4) and take \(\left| x\right| =1\), \(x=\sigma \), we derive

$$\begin{aligned} \lim _{\ell \rightarrow 0}\ell ^{\beta _q}u_{\infty ,\delta }(\ell , \sigma )=\lim _{\ell \rightarrow 0}u_{\infty ,\ell ^{-1}\delta }(1,\sigma )=u_{\infty ,\infty }(1,\sigma )=\omega _{s^{{N-1}}_+}(\sigma ). \end{aligned}$$
(4.8)

This convergence holds in \(C^1(\overline{S^{N-1}_+})\) because of Lemma 2.5. This implies (4.3). \(\square \)

The main classification result is as follows.

Theorem 4.2

Assume \(1<p<N\), \(p-1<q<q^*\) and \(\partial \Omega \cap B_\delta =\{x=(x',0):\left| x'\right| <\delta \}\), for some \(\delta >0\). If \(u\in C(\overline{\Omega }{\setminus }\{0\})\cap C^1(\Omega )\) is a positive solution of (1.1) in \(\Omega \) which vanishes on \(\partial \Omega {\setminus }\{0\}\), then we have the following alternative:

  1. (i)

    either there exists \(k\ge 0\) such that

    $$\begin{aligned} \lim _{x\rightarrow 0}\frac{u(x)}{{\Psi _*}(x)}=k, \end{aligned}$$
    (4.9)
  2. (ii)

    or

    $$\begin{aligned} \lim _{x\rightarrow 0}\frac{u(x)}{U_{s^{{N-1}}_+}(x)}=1. \end{aligned}$$
    (4.10)

Proof

Step 1. Assume

$$\begin{aligned} \liminf _{x\rightarrow 0}\frac{u(x)}{{\Psi _*}(x)}<\infty , \end{aligned}$$
(4.11)

then we claim that (4.9) holds. We first note that if (4.11) holds, there also holds

$$\begin{aligned} \liminf _{x\rightarrow 0}\frac{u(x)}{u_1(x)}<\infty , \end{aligned}$$
(4.12)

where \(u_1\) is the solution of (1.1) obtained in Theorem 3.6 with \(k=1\). If \(\{x_n\}\) is converging to 0 and such that for some \(k>0\)

$$\begin{aligned} \liminf _{x\rightarrow 0}\frac{u(x)}{u_1(x)}=k=\lim _{n\rightarrow \infty }\frac{u(x_n)}{u_1(x_n)}, \end{aligned}$$

there also holds by the boundary Harnack inequality (2.38) applied to both u and \(u_1\),

$$\begin{aligned} \frac{u(x_n)}{u_1(x_n)}=\frac{u(x_n)}{d(x_n)}\frac{d(x_n)}{u_1(x_n)}\ge c_5^{-2}\frac{u(x)}{u_1(x)}\quad \forall x\text { s.t. }\left| x\right| =\left| x_n\right| . \end{aligned}$$

This implies in particular

$$\begin{aligned} u(x)\le c_5^{2}(k+\epsilon _n)u_1(x)\quad \forall x\text { s.t. }\left| x\right| =\left| x_n\right| \end{aligned}$$

where \(\{\epsilon _n\}\) is converging to \(0_+\), and by the comparison principle

$$\begin{aligned} u(x)\le Ku_1(x)\quad \forall x\in \mathbb {R}^N_+\text { s.t. }\left| x_n\right| \le \left| x\right| \le \frac{\delta }{2}, \end{aligned}$$

for some \(K>0\) and all \(n\in \mathbb N_*\). Therefore

$$\begin{aligned} \limsup _{x\rightarrow 0}\frac{u(x)}{u_1(x)}<\infty . \end{aligned}$$
(4.13)

We can assume that \(k\ne 0\), otherwise (4.9) holds with \(k=0\) and actually u remains bounded near 0. As a consequence of the Hopf Lemma and \(C^1\) regularity, there exists \(K>0\) such that

$$\begin{aligned} u(x)\le K{\Psi _*}(x)\quad \forall x\in B^+_{\frac{\delta }{2}}. \end{aligned}$$
(4.14)

Let \(m=\max \{u(x):\left| x\right| =\delta \}\). For \(0<\tau <\delta \) we denote by \(k_\tau \) the minimum of the \(\kappa >0\) such that \(u(x)\le \kappa {\Psi _*}(x)+m\) for \(\tau \le \left| x\right| \le \delta \). Then \(u(x)\le k_\tau {\Psi _*}(x)+m\), and either the graphs of the mappings \(u(\cdot )\) and \(k_\tau {\Psi _*}(\cdot )+m\) are tangent at some \(x_\tau \in B^+_\delta {\setminus }\overline{B}^+_\tau \), or they are tangent on the boundary of the domain, and the only possibility is that they are tangent on \(\left| x\right| =\tau \). Since

$$\begin{aligned} \left| \nabla {\Psi _*}(x)\right| ^2=\left| x\right| ^{-2(\beta _*+1)}(\beta _*^2\psi _*^2+\left| \nabla {\psi _*}\right| ^2), \end{aligned}$$

it never vanishes. If we set \(w=u-(k_\tau {\Psi _*}(x)+m)\), then

$$\begin{aligned} -{\mathcal L}w+\left| \nabla u\right| ^q=0 \end{aligned}$$
(4.15)

where the operator

$$\begin{aligned} {\mathcal L}=\sum _{i,j}{\frac{\partial }{\partial x_i} }\left( a_{ij}{\frac{\partial }{\partial x_j} }\right) \end{aligned}$$

is uniformly elliptic in a neighborhood of \(x_\tau \) (see [6, Lemma 1.3]). Furthermore \(w\le 0\) and \(w(x_\tau )=0\) by the strong maximum principle \(\nabla u(x_\tau )\) must vanish, which contradicts the fact that \(\nabla u(x_\tau )=\nabla w(x_\tau )\) by the tangency condition, and \(\nabla w(x_\tau )\ne 0\). Therefore \(\left| x_\tau \right| =\tau \) and \(x_\tau \notin \partial \mathbb {R}^N_+\). If \(\tau '<\tau \), \(k_\tau \le k_{\tau '}\), and we set \(k=\lim _{\tau \rightarrow 0}k_{\tau }\), which is finite because of (4.14). There exists \(\{\tau _n\}\) such that \(\sigma _n:=\tau ^{-1}x_{\tau _n}\rightarrow \sigma _0\). Furthermore

$$\begin{aligned} r^{\beta _*}u(r,\sigma )\le k_\tau {\psi _*}(\sigma )+mr^{\beta _*}\quad \text {if }\tau \le r\le \delta \quad \text {and}\quad \tau ^{\beta _*}u(\tau ,\sigma _\tau )= k_\tau {\psi _*}(\sigma _\tau )+m\tau ^{\beta _*}. \end{aligned}$$
(4.16)

Put

$$\begin{aligned} u_\tau (x)=\tau ^{\beta _*}u(\tau x) \end{aligned}$$
(4.17)

Then

$$\begin{aligned} -\Delta _p u_\tau +\tau ^{p-q-\beta _*(p+1-q)}\left| \nabla u_\tau \right| ^q=0\qquad \text {in }B^+_{\frac{\delta }{\tau }}{\setminus } \{0\} \end{aligned}$$

and, by (4.14),

$$\begin{aligned} 0\le u_\tau (x)\le K\left| x\right| ^{-\beta _*}\quad \text {in }B^+_{\frac{\delta }{2\tau }}{\setminus } \{0\}. \end{aligned}$$

By Lemma 2.5, the set of functions \(\{u_\tau (\cdot )\}\) is relatively compact in the \(C_{loc}^1\) topology of \(\overline{\mathbb {R}^N_+}{\setminus } \{0\}\). Therefore, as \(q<q^*\), there exist a sequence \(\{\tau '_n\}\subset \{\tau _n\}\) converging to 0, and a positive p-harmonic function v in \(\mathbb {R}^N_+\), continuous in \(\overline{\mathbb {R}^N_+}{\setminus } \{0\}\) and vanishing on \(\partial \mathbb {R}^N_+{\setminus } \{0\}\), such that \(u_{\tau '_n}\rightarrow v\), and v satisfies (4.14) in \(\overline{\mathbb {R}^N_+}{\setminus } \{0\}\). By Theorem 5.1 in Appendix A, there exists \(k^*\) such that \(v=k^*{\Psi _*}\). In particular,

$$\begin{aligned} \lim _{\tau '_n\rightarrow 0}u_{\tau '_n}(1,\sigma )=k^*{\psi _*}(\sigma ) \end{aligned}$$
(4.18)

in the \(C^1(\overline{S^{N-1}_+})\) topology. Combining (4.16), (4.17) and (4.18) we conclude that \(k^*=k\) and

$$\begin{aligned} \lim _{\tau '_n\rightarrow 0}\tau '^{\beta _*}_nu_{\tau '_n}(1,\sigma )=k{\psi _*}(\sigma ) \end{aligned}$$
(4.19)

Using Theorem 3.6, it is equivalent to

$$\begin{aligned} \lim _{\tau '_n\rightarrow 0}{\frac{u(\tau '_n,\sigma )}{u_k(\tau '_n,\sigma )} }=1 \end{aligned}$$
(4.20)

uniformly on \(S^{N-1}_+\). For any \(\epsilon >0\), there exists \(n_\epsilon >0\) such that \(n\ge n_\epsilon \) implies

$$\begin{aligned} u_{k-\epsilon }(\tau '_n,\sigma )\le u(\tau '_n,\sigma )\le u_{k+\epsilon }(\tau '_n,\sigma ) \end{aligned}$$

By comparison principle,

$$\begin{aligned} u_{k-\epsilon }\le u\le u_{k+\epsilon }+m\quad \text {in }B^+_\delta {\setminus } B^+_{\tau '_n}, \end{aligned}$$
(4.21)

and finally

$$\begin{aligned} u_{k-\epsilon }\le u\le u_{k+\epsilon }+m\quad \text {in }B^+_\delta , \end{aligned}$$
(4.22)

Since \(\epsilon \) is arbitrary and using again Theorem 3.6, it implies

$$\begin{aligned} \lim _{r\rightarrow 0}{\frac{u(r,\sigma )}{{\Psi _*}(r,\sigma )} }=k, \end{aligned}$$
(4.23)

locally uniformly on \(S^{N-1}\). But since the convergence holds in \(C^1(\overline{S^{N-1}_+})\), (4.9) follows.

Step 2. Assume

$$\begin{aligned} \lim _{x\rightarrow 0}\frac{u(x)}{{\Psi _*}(x)}=\infty . \end{aligned}$$
(4.24)

For any \(0<\epsilon <\delta \) and \(k>0\), there holds

$$\begin{aligned} u_k(x)\le u(x)\le v_\epsilon (x)\quad \text {in }B_\delta ^+{\setminus } B_\epsilon ^+ \end{aligned}$$
(4.25)

where \(v_\epsilon \) has been defined in (3.12) and \(u_k\) is given by Theorem 3.6. Letting \(\epsilon \rightarrow 0\), \(k\rightarrow \infty \), and using Proposition 4.1, we derive

$$\begin{aligned} u_\infty (x)\le u(x)\le v_0(x)\quad \text {in }B_\delta ^+{\setminus } \{0\}. \end{aligned}$$
(4.26)

We have seen in Theorem 3.3 that \(v_0\) is a separable solution of (1.1) in \(\mathbb {R}^N_+\) which vanishes on \(\partial \mathbb {R}^N_+{\setminus }\{0\}\), therefore \(v_0(x)=U_{s^{{N-1}}_+}(x)\). This implies

$$\begin{aligned} u_\infty (x)\le u(x)\le \left| x\right| ^{-\beta _q}\omega _{s^{{N-1}}_+}\left( \frac{x}{\left| x\right| }\right) \quad \text {in }B_\delta ^+{\setminus } \{0\}. \end{aligned}$$
(4.27)

We conclude using Proposition 4.1. \(\square \)

4.2 The case \(p=N\)

When \(p=N\), the assumption that \(\partial \Omega \) is an hyperplane near 0 can be removed. The proof of the next results is based upon Theorem 3.10. The following result is the extension to the case \(p=N\) of Proposition 4.1.

Proposition 4.3

Under the assumptions of Theorem 3.10 there exists \(\lim _{k\rightarrow \infty }u_k=u_\infty \) which is the unique element of \(C(\overline{\Omega }{\setminus }\{0\})\cap C^1(\Omega )\) which satisfies (3.17) in \(\Omega \), vanishes on \(\partial \Omega {\setminus }\{0\}\) and such that

$$\begin{aligned} \lim _{x\rightarrow 0}\frac{u_{\infty }(x)}{U_{s^{{N-1}}_+}(x)}=1. \end{aligned}$$
(4.28)

Proof

We denote by \(u_k^{\Omega }\) the unique positive solution of (3.17) satisfying (3.57) obtained in Theorem 3.6. Then

$$\begin{aligned} T_\ell [u_k^\Omega ]=u^{\Omega ^{\ell }}_{\ell ^{\beta _q-\beta _*}k}, \end{aligned}$$
(4.29)

because of uniqueness. We denote by \(B:=B_{\frac{1}{2}}(a)\) and \(B':=B_{\frac{1}{2}}(a')\) the two balls tangent to \(\partial \Omega \) at 0 respectively interior and exterior to \(\Omega \) introduced in the proof of Lemma 3.11. Estimate (3.58) implies

$$\begin{aligned} u_k^{B'^c}\le u_k^\Omega \le u_k^B \end{aligned}$$
(4.30)

the left-hand side inequality holding in \(\Omega \) and the right-hand side one in B. Therefore

$$\begin{aligned} T_\ell [u_k^{B'^c}]:=u^{B'^{c\,\ell }}_{\ell ^{\beta _q-\beta _*}k}\le T_\ell [u_k^\Omega ]\le T_\ell [u_k^B]:=u^{B^{\ell }}_{\ell ^{\beta _q-\beta _*}k}, \end{aligned}$$
(4.31)

the domains of validity of these inequalities being modified accordingly. Using again (3.58) we obtain

$$\begin{aligned} T_{\ell '}[u_{k'}^{B'^c}]\le T_\ell [u_k^{B'^c}]\quad \text {in } B'^{c\,\ell '}, \end{aligned}$$
(4.32)

for any \(0<\ell '\le \ell \) and \(\ell '^{\beta _q-\beta _*}k'\le \ell ^{\beta _q-\beta _*}k\). In the same way

$$\begin{aligned} T_{\ell '}[u_{k'}^{B}]\ge T_\ell [u_k^{B}]\quad \text {in } B^{\ell }, \end{aligned}$$
(4.33)

for any \(0<\ell '\le \ell \) and \(\ell '^{\beta _q-\beta _*}k'\ge \ell ^{\beta _q-\beta _*}k\). Since \(u_k^\Omega \) \(u_k^B\), \(u_k^{B'^c}\) are increasing with respect to k, they converge respectively to \(u_\infty ^\Omega \) \(u_\infty ^B\), \(u_\infty ^{B'^c}\) and there holds for any \(\ell >0\)

$$\begin{aligned} T_\ell [u_\infty ^{B'^c}]\le T_\ell [u_\infty ^\Omega ]\le T_\ell [u_\infty ^{B}], \end{aligned}$$
(4.34)

from (4.31) and

$$\begin{aligned} \begin{array}{ll} (i)\quad T_{\ell '}[u_{\infty }^{B'^c}]\le T_\ell [u_\infty ^{B'^c}]&{}\quad \text {in } B'^{c\,\ell '}\\ (ii)\quad T_{\ell '}[u_{\infty }^{B}]\ge T_\ell [u_\infty ^{B}]&{}\quad \text {in } B^{\ell } \end{array} \end{aligned}$$
(4.35)

for any \(0<\ell '\le \ell \). Notice that , replacing \(\ell \) by \(\ell \ell '\) we can rewrite (4.34) as follows

$$\begin{aligned} T_{\ell '}[T_\ell [u_\infty ^{B'^c}]]\le T_{\ell '}[T_\ell [u_\infty ^\Omega ]]\le T_{\ell '}[T_\ell [u_\infty ^{B}]]. \end{aligned}$$
(4.36)

Because of the monotonicity with respect to \(\ell \) the following limits exist

$$\begin{aligned} U^{B'^c}=\lim _{\ell \rightarrow 0}T_\ell [u_\infty ^{B'^c}]\quad \text {and}\quad U^{B}=\lim _{\ell \rightarrow 0}T_\ell [u_\infty ^{B}]. \end{aligned}$$
(4.37)

By Lemma 2.5 applied with \(\phi (|x|)=|x|^{-\beta _q}\) and since there holds \(u_\infty ^B(x)\le c|x|^{-\beta _q}\) and \(u_\infty ^{B'}(x)\le c|x|^{-\beta _q}\), we derive

$$\begin{aligned} \begin{array}{ll} (i)\quad |\nabla T_\ell [u_\infty ^B](x)|\le c_2|x|^{-\beta _q-1}&{}\quad \forall x\in B^{\ell }\\ (ii)\quad |\nabla T_\ell [u_\infty ^B](x)-\nabla T_\ell [u_\infty ^B](y)|\le c_2|x|^{-\beta _q-1-\alpha }|x-y|^{\alpha }&{}\quad \forall x,y\in B^{\ell },\;|x|\le |y|\\ (iii) \quad T_\ell [u_\infty ^B](x)\le c_2|x|^{-\beta _q-1} (\text{ dist }\,(x,\partial B^\ell ))^{\alpha }&{}\quad \forall x\in B^{\ell }, \end{array} \end{aligned}$$
(4.38)

and

$$\begin{aligned} \begin{array}{lll} (i)\quad |\nabla T_\ell [u_\infty ^{B'^c}](x)|\le c_2|x|^{-\beta _q-1}&{}\quad \forall x\in B'^{c\,\ell }\\ (ii) \quad |\nabla T_\ell [u_\infty ^{B'^c}](x)-\nabla T_\ell [u_\infty ^{B'^c}](y)|\le c_2|x|^{-\beta _q-1-\alpha }|x-y|^{\alpha }&{}\quad \forall x,y\in B'^{c\,\ell },\;|x|\le |y|\\ (iii)\quad T_\ell [u_\infty ^{B'^c}](x)\le c_2|x|^{-\beta _q-1} (\text{ dist }\,(x,\partial B'^{c\,\ell }))^{\alpha }&{}\quad \forall x\in B'^{c\,\ell }. \end{array} \end{aligned}$$
(4.39)

Thus the sets of functions \(\{T_\ell [u_\infty ^B]\}\) and \(\{T_\ell [u_\infty ^{B'}]\}\) are equicontinuous in the \(C^1\)-loc topology and by uniqueness, the limit in (4.37) below holds in this topology. Hence \(U^{B'^c}\) and \(U^{B^c}\) are positive solutions of (3.17) in \(\mathbb {R}^N_+\) which vanish on \(\partial \mathbb {R}^N_+{\setminus }\{0\}\). Furthermore \(U^{B'^c}\le U^{B^c}\) Since for any \(\ell ,\ell '>0\), \(T_{\ell '}[T_\ell [u_\infty ^{B'^c}]]=T_{\ell \ell '}[u_\infty ^{B'^c}]\), it follows \(T_{\ell '}[U^{B'^c}]=U^{B'^c}\) and in the same way \(T_{\ell '}[U^{B}]=U^{B}\). This means that \(U^{B}\) and \(U^{B'^c}\) are self-similar solutions of (3.17) in \(\mathbb {R}^N_+\) and they vanish on \(\partial \mathbb {R}^N_+{\setminus }\{0\}\). Hence

$$\begin{aligned} U^{B}=U^{B'^c}=U_{S^{N-1}_+}. \end{aligned}$$
(4.40)

Applying again Lemma 2.5 to \(u_\infty ^\Omega \) with \(\phi (|x|)=|x|^{-\beta _q}\) we have

$$\begin{aligned} \begin{array}{ll} (i)\quad |\nabla T_\ell [u_\infty ^\Omega ](x)|\le c_2|x|^{-\beta _q-1}&{}\quad \forall x\in \Omega ^{\ell }\\ (ii)\quad |\nabla T_\ell [u_\infty ^\Omega ](x)-\nabla T_\ell [u_k^\Omega ](y)|\le c_2|x|^{-\beta _q-1-\alpha }|x-y|^{\alpha }&{}\quad \forall x,y\in \Omega ^{\ell },\;|x|\le |y|\\ (iii)\quad T_\ell [u_\infty ^\Omega ](x)\le c_2|x|^{-\beta _q-1} (\text{ dist }\,(x,\partial \Omega ^\ell ))^{\alpha }&{}\quad \forall x\in \Omega ^{\ell }. \end{array} \end{aligned}$$
(4.41)

This implies that the set of functions \(\{T_\ell [u_\infty ^\Omega ]\}_\ell \) is equicontinuous in the \(C^1\)-loc topology of \(\mathbb {R}^N_+\) and there exists a sequence \(\{\ell _n\}\rightarrow 0\) and a function U such that \(T_{\ell _n}[u_\infty ^\Omega ]\rightarrow U^\Omega \) in this topology of \(\mathbb {R}^N_+\), and U is a positive solution of (3.17) in \(\mathbb {R}^N_+\) which vanishes on \(\partial \mathbb {R}^N_+{\setminus }\{0\}\). From (4.34) and (4.40) there holds \(U^\Omega = U_{S^{N-1}_+}\) and therefore

$$\begin{aligned} \lim _{\ell \rightarrow 0}T_\ell [u_\infty ^\Omega ]=U_{S^{N-1}_+}. \end{aligned}$$
(4.42)

This implies (4.28) and

$$\begin{aligned} \lim _{r\rightarrow 0}r^{\beta _q}u_\infty ^\Omega (r,\sigma )=\omega _{S^{N-1}_+}(\sigma ) \end{aligned}$$
(4.43)

uniformly on compact subsets of \(S^{N-1}_+\). \(\square \)

Up to minor modifications the proof of the next classification theorem is similar to the one of Theorem 4.2.

Theorem 4.4

Assume \(N-1<q<N-\frac{1}{2}\) If \(u\in C(\overline{\Omega }{\setminus }\{0\})\cap C^1(\Omega )\) is a positive solution of (3.17) in \(\Omega \) which vanishes on \(\partial \Omega {\setminus }\{0\}\), then we have the following alternative:

  1. (i)

    either there exists \(k\ge 0\) such that (4.9) holds,

  2. (ii)

    or (4.10) holds.