1 Introduction and main results

In this paper, we consider the existence of sign-changing bubble tower solutions for the following supercritical Hénon-type equation

$$\begin{aligned} {\left\{ \begin{array}{ll} -\Delta u=|x|^\alpha |u|^{p_\alpha -1-\varepsilon }u~~&{}\text {in}~\Omega ,\\ u=0~~&{}\text {on}~\partial \Omega , \end{array}\right. } \end{aligned}$$
(1.1)

where \(\alpha >-2\), \(\varepsilon >0\), \(p_\alpha =\frac{N+2+2\alpha }{N-2}\), \(N\ge 3\), \(\Omega \) is a smooth bounded domain in \(\mathbb {R}^N\) containing the origin.

When \(\Omega \) is the unit ball \(B_1(0)\) of \(\mathbb {R}^N\), problem (1.1) becomes the well-known Hénon equation, i.e.,

$$\begin{aligned} {\left\{ \begin{array}{ll} -\Delta u=|x|^\alpha u^{p}, \; u>0, &{} \text {in}\; B_1(0),\\ u=0, &{}\text {on}\; \partial B_1(0). \end{array}\right. } \end{aligned}$$
(1.2)

Problem (1.2) was proposed by Hénon in [15] when he studied rotating stellar structures, which has attracted a lot of interest in recent years. Ni [21] first considered (1.2) and proved that it possesses a positive radial solution when \(p\in (1,p_\alpha )\). Due to the appearance of the weighted term \(|x|^\alpha \), the classical moving plane method in [13] cannot be applied to problem (1.2) when \(\alpha >0\). Therefore it is quite natural to ask whether problem (1.2) with \(\alpha >0\) has non-radial solutions. Based on numerical results in [4], Smets, Su and Willem [27] obtained the existence of non-radial solutions for \(1<p<\frac{N+2}{N-2}\), when \(\alpha \) is large enough. For \(p=\frac{N+2}{N-2}-\varepsilon \), Cao and Peng [5] showed that the ground state solution is non-radial and blows up near the boundary of \(B_1(0)\) as \(\varepsilon \rightarrow 0\). Later on, Peng [22] constructed multiple boundary peak solutions for problem (1.2). When \(p=\frac{N+2}{N-2}\), Serra [26] proved that problem (1.2) has a non-radial solution provided \(\alpha \) is large enough. More recently, Wei and Yan [28] showed that there are infinitely many non-radial positive solutions for problem (1.2) with \(\alpha >0\). For other results related to the Hénon-type problems, see [1, 2, 6, 16, 24] and the references therein.

On the other hand, using the Pohozaev-type identity [25], we know that for \(p\ge p_\alpha \) there are no nontrivial solutions to problem (1.2). So it seems more interesting whether there are solutions for \(p\in (\frac{N+2}{N-2},p_\alpha )\). When \(p=p_\alpha -\varepsilon \) with small \(\varepsilon >0\), Gladiali and Grossi in [11] showed that there exists a solution concentrating at origin provided \(0<\alpha \le 1\). By the results in [12], the same results still hold when \(\alpha \) is not an even integer. In [17], the asymptotic behavior of the radial solutions obtained by Ni in [21] was analyzed as \(\varepsilon \rightarrow 0^+\). More recently, Liu and Peng [19] constructed large number of peak solutions for (1.2) with \(p=\frac{N+2}{N-2}+\varepsilon \). However, as far as we know, it seems that there are no results on existence of sign-changing solutions for (1.2) when \(p\in (\frac{N+2}{N-2},p_\alpha )\)

Our purpose in the present paper is to construct sign-changing bubble tower solutions for (1.1) which blow up at the origin. It seems that this is the first existence result of sign-changing bubble tower solutions for (1.1).

The main result of this paper is as follows.

Theorem 1.1

Assume that \(N\ge 3\), \(\alpha >-2\) is not an even integer, then for any \(k\in \mathbb {N}^+\), there exists \(\varepsilon _k>0\) such that for \(\varepsilon \in (0,\varepsilon _k)\), problem (1.1) has a sign-changing bubble tower solution \(u_\varepsilon \) with exactly k nodal sets in \(\Omega \).

Remark 1.2

When \(\alpha =0\) and \(\Omega \) has some symmetry property, problem (1.1) has been studied in [3] and [23]. Our results do not need any symmetry property of \(\Omega \). Further more, compared with the classical Hénon equation where \(\alpha >0\), our result covers the more general case \(\alpha >-2\).

Remark 1.3

If \(\Omega \) is the unit ball of \(\mathbb {R}^N\) , then Theorem 1.1 holds for all \(\alpha >-2\). Actually, we can construct a sign-changing radial bubble tower solution \(u_\varepsilon \). Considering the transformation \(\texttt {w}(s)=u(r),~~r=s^{\frac{2}{\alpha +2}}\), problem (1.1) can be changed into the following problem

$$\begin{aligned} {\left\{ \begin{array}{ll} -\texttt {w}^{\prime \prime }-\frac{M-1}{s}{} \texttt {w}^\prime =\frac{4}{(2+\alpha )^2}|\texttt {w}|^{\frac{M+2}{M-2}-\varepsilon }{} \texttt {w} , &{} \text {in}\; (0,1),\\ \texttt {w}^\prime (0)=\texttt {w}(1)=0, &{} \end{array}\right. } \end{aligned}$$
(1.3)

where \(M=\frac{2(N+\alpha )}{2+\alpha }\). When M is an integer, problem (1.3) was studied in [3]. However, problem (1.3) can be dealt with in a similar way if M is not an integer.

Let us outline the main idea to prove Theorem 1.1. To do this, we introduce a few notations first. For \(x\in \mathbb {R}^N\) and \(\mu >0\), set

$$\begin{aligned} U_{\mu }(x)=C_{\alpha ,N}\left( \frac{\mu ^{\frac{2+\alpha }{2}}}{\mu ^{2+\alpha }+|x|^{2+\alpha }}\right) ^{\frac{N-2}{2+\alpha }},\,\, C_{\alpha ,N}=((N+\alpha )(N-2))^{\frac{N-2}{4+2\alpha }}. \end{aligned}$$

It is well known from [12, 14] that \(U_{\mu }(x)\) are the only radial solutions of

$$\begin{aligned} -\Delta u=|x|^\alpha u^{p_\alpha }, u>0~\text {in}~\mathbb {R}^N. \end{aligned}$$
(1.4)

We define the following Emden–Fowler-type transformation

$$\begin{aligned} v(y,\Theta )=\mathcal {T}(u)(y,\Theta )=\left( \frac{p_\alpha -1}{2}\right) ^{\frac{2}{p_\alpha -1}}r^{\frac{N-2}{2}}u(r,\Theta ), \end{aligned}$$
(1.5)

where

$$\begin{aligned} r=e^{-\frac{p_\alpha -1}{2}y}, \Theta \in \mathbb {S}^{N-1}. \end{aligned}$$

Define

$$\begin{aligned} D=\{(y,\Theta )\in \mathbb {R}\times \mathbb {S}^{N-1}:(e^{-\frac{p_\alpha -1}{2}y},\Theta )\in \Omega \}. \end{aligned}$$

After these changes of variables, problem (1.1) becomes

$$\begin{aligned} {\left\{ \begin{array}{ll} L(v)=\sigma _\varepsilon e^{-\frac{2+\alpha }{2}\varepsilon y}|v|^{p_\alpha -1-\varepsilon }v~~&{}\text {in}~D,\\ v=0~~&{}\text {on}~\partial D, \end{array}\right. } \end{aligned}$$
(1.6)

where

$$\begin{aligned} \sigma _\varepsilon =\left( \frac{p_\alpha -1}{2}\right) ^{\frac{2\varepsilon }{p_\alpha -1}} \end{aligned}$$

and

$$\begin{aligned} L(v)=-v^{\prime \prime }+\frac{(2+\alpha )^2}{4}v-\left( \frac{p_\alpha -1}{2}\right) ^2 \Delta _{\mathbb {S}^{N-1}}v. \end{aligned}$$

The energy functional corresponding to problem (1.6) is

$$\begin{aligned} \begin{aligned} I_\varepsilon (v)&=\frac{1}{2}\int _D\left( |v^\prime |^2+\frac{(2+\alpha )^2}{4} |v|^2\right) \mathrm{d}y\mathrm{d}\Theta +\frac{1}{2}\left( \frac{p_\alpha -1}{2}\right) ^2\int _D|\nabla _{\mathbb {S}^{N-1}}v|^2\mathrm{d}y\mathrm{d}\Theta \\&\quad -\,\frac{\sigma _\varepsilon }{p_\alpha +1-\varepsilon }\int _D e^{-\frac{2+\alpha }{2}\varepsilon y}|v|^{p_\alpha +1-\varepsilon }\mathrm{d}y\mathrm{d}\Theta . \end{aligned} \end{aligned}$$

It is easy to see that

$$\begin{aligned} I_\varepsilon (v)=\left( \frac{p_\alpha -1}{2}\right) ^{\frac{p_\alpha +3}{p_\alpha -1}}J_\varepsilon (u), \end{aligned}$$
(1.7)

where

$$\begin{aligned} J_\varepsilon (u)=\frac{1}{2}\int _\Omega |\nabla u|^2dx-\frac{1}{p_\alpha +1-\varepsilon }\int _\Omega |x|^\alpha |u|^{p_\alpha +1-\varepsilon }dx. \end{aligned}$$

We observe that W(y) is the unique solution of the problem

$$\begin{aligned} {\left\{ \begin{array}{ll} W^{\prime \prime }-\frac{(2+\alpha )^2}{4}W+W^{p_\alpha }=0~~\text {in}~\mathbb {R},\\ W^\prime (0)=0,\quad W(y)>0, \\ W(y)\rightarrow 0~~~\text {as}~~y\rightarrow \pm \infty , \end{array}\right. } \end{aligned}$$
(1.8)

where

$$\begin{aligned} W(y)=\gamma _{\alpha ,N} \frac{e^{-\frac{2+\alpha }{2}y}}{\left( 1+e^{-\frac{(2+\alpha )^2}{N-2}y}\right) ^{\frac{N-2}{2+\alpha }}}, ~~\gamma _{\alpha ,N}=\left( \frac{(2+\alpha )^2(N+\alpha )}{N-2}\right) ^{\frac{N-2}{4+2\alpha }}. \end{aligned}$$

We denote the function \(\mathrm{PU}_\mu :=U_\mu +R_\mu \), which is the projection onto \(H_0^1(\Omega )\) of the function \(U_\mu \), that is,

$$\begin{aligned} {\left\{ \begin{array}{ll} -\Delta \mathrm{PU}_\mu =|x|^\alpha U_\mu ^{p_\alpha }~~&{}\text {in}~\Omega ,\\ \mathrm{PU}_\mu =0~~&{}\text {on}~\partial \Omega . \end{array}\right. } \end{aligned}$$

Then, we have

$$\begin{aligned} R_\mu =-C_{\alpha ,N}\mu ^{\frac{N-2}{2}}H(x,0)+O\left( \mu ^{\frac{N+2+2\alpha }{2}}\right) , \end{aligned}$$

where H(x, 0) is the Robin function.

For given \(\Lambda _i>0,~i=1,2,\ldots ,k,\) set

$$\begin{aligned} \begin{aligned} \xi _1&= -\frac{1}{2+\alpha }\log \varepsilon +\frac{2}{2+\alpha }\log \Lambda _1, \\ \xi _{i+1}-\xi _i&=-\frac{2}{2+\alpha } \log \varepsilon -\frac{2}{2+\alpha }\log \Lambda _{i+1},~~i=1,2,\ldots ,k-1. \end{aligned} \end{aligned}$$
(1.9)

Let us write

$$\begin{aligned} W_i(y)=W(y-\xi _i),~~~V_i(y)=W_i(y)+\Pi _i(y),~~~V(y)=\sum _{i=1}^k(-1)^iV_i(y), \end{aligned}$$
(1.10)

where

$$\begin{aligned} \Pi _i(y)=\mathcal {T}(R_{\mu _i}),~\mu _i=e^{-\frac{p_\alpha -1}{2}\xi _i}. \end{aligned}$$

We will prove Theorem 1.1 by verifying the following result.

Theorem 1.4

Suppose that \(\alpha >-2\) is not an even integer. Then for any integer \(k\ge 1\), there exists \(\varepsilon _k>0\) such that for \(\varepsilon \in (0,\varepsilon _k)\), problem (1.6) has a pair of solutions \(v_\varepsilon \) and \(-v_\varepsilon \) of the form

$$\begin{aligned} v_\varepsilon =V+\phi _\varepsilon , \end{aligned}$$

where \(\Vert \phi _\varepsilon \Vert _{L^\infty }\rightarrow 0\), as \(\varepsilon \rightarrow 0\).

Remark 1.5

Using the Emden–Fowler-type transformation (1.5), we can give the explicit expression of solution to problem (1.1), that is,

$$\begin{aligned} u_\varepsilon (x)=C_{\alpha ,N}\sum _{i=1}^k(-1)^i \left( \frac{M_i^{\frac{2+\alpha }{2}}\varepsilon ^{\frac{(2+\alpha )(2i-1)}{2(N-2)}}}{M_i^{2+\alpha }\varepsilon ^{\frac{(2+\alpha )(2i-1)}{N-2}}+|x|^{2+\alpha }}\right) ^{\frac{N-2}{2+\alpha }}(1+o(1)), \end{aligned}$$

where \(M_i, i=1,2,\ldots ,k\) are some certain positive constants (see (4.3)) and \(o(1)\rightarrow 0\) uniformly on compact subsets of \(\Omega \) as \(\varepsilon \rightarrow 0\) .

The proof of Theorem 1.4 is motivated by [7, 23]. More precisely, we will use the Lyapunov–Schmidt reduction argument to prove Theorem 1.4, which reduces the construction of the solutions to a finite-dimensional variational problem. As a final remark, we point out that bubble tower concentration phenomena have been observed in [3, 7, 8, 10, 18, 20, 23] near the critical Sobolev exponent, i.e., \(\alpha =0\). However, as far as we know, there are no such results for \(\alpha \ne 0\).

This paper is organized as follows. In Sect. 2, we give some basic estimates and asymptotic expansion. In Sect. 3, we will carry out the finite-dimensional reduction argument and the main results will be proved in Sect. 4.

2 Energy expansion

In this section, we give some estimates and asymptotic expansion used in the later sections.

Lemma 2.1

For fixed \(\delta >0\) and \(\delta<\Lambda _i<\delta ^{-1}, i=1,2,\ldots ,k\), we have the following estimates:

$$\begin{aligned}&\displaystyle \int _D |V|^{p_\alpha +1}\mathrm{d}y\mathrm{d}\Theta = k\omega _{N-1}\int _{\mathbb {R}}W^{p_\alpha +1}+o(1), \end{aligned}$$
(2.1)
$$\begin{aligned}&\displaystyle \int _D\left( |V|^{p_\alpha +1}-|V|^{p_\alpha +1-\varepsilon }\right) \mathrm{d}y\mathrm{d}\Theta = k\omega _{N-1}\varepsilon \int _{\mathbb {R}}W^{p_\alpha +1}\log W+o(\varepsilon ), \end{aligned}$$
(2.2)
$$\begin{aligned}&\displaystyle \int _D y|V|^{p_\alpha +1}\mathrm{d}y\mathrm{d}\Theta =\left( \sum _{\ell =1}^k\xi _\ell \right) \omega _{N-1}\int _{\mathbb {R}}W^{p_\alpha +1}+o(1), \end{aligned}$$
(2.3)
$$\begin{aligned}&\displaystyle \int _{D_\ell }W_i^{p_\alpha }W_j\mathrm{d}y\mathrm{d}\Theta =o\left( \varepsilon \right) ,~~i\ne \ell , \end{aligned}$$
(2.4)
$$\begin{aligned}&\displaystyle \int _{D_\ell }W_\ell ^{p_\alpha }W_j\mathrm{d}y\mathrm{d}\Theta =a_3 e^{-\frac{2+\alpha }{2}|\xi _\ell -\xi _j|}+o\left( \varepsilon \right) ,~~j\ne \ell , \end{aligned}$$
(2.5)
$$\begin{aligned}&\displaystyle \int _{D_\ell }\left( |V_\ell |^{p_\alpha +1}-|V|^{p_\alpha +1}+(p_\alpha +1)V_\ell ^{p_\alpha }\sum _{j\ne \ell }(-1)^{\ell +j}V_j\right) \mathrm{d}y\mathrm{d}\Theta =o\left( \varepsilon \right) , \end{aligned}$$
(2.6)
$$\begin{aligned}&\displaystyle \int _{D_\ell }\left( W_\ell ^{p_\alpha }-V_\ell ^{p_\alpha }\right) V_j\mathrm{d}y\mathrm{d}\Theta = o\left( \varepsilon \right) ,~~~j\ne \ell , \end{aligned}$$
(2.7)

where \(a_3=\gamma _{\alpha ,N}\omega _{N-1}\int _{\mathbb {R}}e^{-\frac{2+\alpha }{2}y}W^{p_\alpha }\), \(D_\ell =\{(y,\Theta )\in D: \eta _\ell \le y<\eta _{\ell +1}\}\), \( \eta _1=0, \eta _\ell =\frac{\xi _{\ell -1}+\xi _\ell }{2},~\ell =2,\ldots ,k, \eta _{k+1}=+\infty \).

Proof

The results are similar to Lemma 4.4 in [23], we omit the details. \(\square \)

Next, we will calculate the asymptotic expansions of the energy functional \(I_\varepsilon (V)\).

Proposition 2.2

For any \(\delta >0\), there exists \(\varepsilon _0>0\) such that for \(\varepsilon \in (0,\varepsilon _0)\), we have the following asymptotic expansion

$$\begin{aligned} \begin{aligned} I_\varepsilon (V)=k a_0+ka_1\varepsilon -\frac{k^2}{2}a_4\varepsilon \log \varepsilon +\varepsilon \Psi _k(\Lambda )+\varepsilon R_\varepsilon (\Lambda ), \end{aligned} \end{aligned}$$
(2.8)

where

$$\begin{aligned} \Psi _k(\Lambda )=ka_4\log \Lambda _1+\frac{a_2H(0,0)}{\Lambda _1^2}+\sum _{\ell =2}^k\left( a_3\Lambda _\ell -(k-\ell +1)a_4\log \Lambda _\ell \right) \end{aligned}$$

and \(R_\varepsilon \rightarrow 0\), as \(\varepsilon \rightarrow 0\) uniformly in \(C^1\)-norm on the set of \(\Lambda _i\)’s with \(\delta<\Lambda _i<\delta ^{-1}\), \(i=1,2,\ldots ,k\). Here \(a_i, i=0,1,\ldots ,4\), are given by

$$\begin{aligned} {\left\{ \begin{array}{ll}\displaystyle a_0=\frac{2+\alpha }{2(N+\alpha )}\left( \frac{p_\alpha -1}{2}\right) ^{\frac{p_\alpha +3}{p_\alpha -1}}\int _{\mathbb {R}^N}|x|^\alpha U_1^{p_\alpha +1},\\ \displaystyle a_1=\frac{\omega _{N-1}}{p_\alpha +1}\left( \int _{\mathbb {R}}W^{p_\alpha +1}\log W-\frac{1}{p_\alpha +1}\int _{\mathbb {R}}W^{p_\alpha +1} -\frac{2}{p_\alpha -1}\log \frac{p_\alpha -1}{2}\int _{\mathbb {R}}W^{p_\alpha +1}\right) ,\\ \displaystyle a_2=\frac{2+\alpha }{2(N+\alpha )}\left( \frac{p_\alpha -1}{2}\right) ^{\frac{p_\alpha +3}{p_\alpha -1}}C_{\alpha ,N}\int _{\mathbb {R}^N}|x|^\alpha U_1^{p_\alpha },\\ \displaystyle a_3=\gamma _{\alpha ,N}\omega _{N-1}\int _{\mathbb {R}}e^{-\frac{2+\alpha }{2}y}W^{p_\alpha },\\ \displaystyle a_4=\frac{\omega _{N-1}}{p_\alpha +1}\int _{\mathbb {R}}W^{p_\alpha +1}. \end{array}\right. } \end{aligned}$$

Proof

The proof is standard, and we only give a sketch here.

Note that

$$\begin{aligned} \begin{aligned} I_{\varepsilon }(V)=&I_0(V)+\frac{1}{p_\alpha +1}\int _D|V|^{p_\alpha +1}-\frac{\sigma _\varepsilon }{p_\alpha +1-\varepsilon }\int _De^{-\frac{2+\alpha }{2}\varepsilon y}|V|^{p_\alpha +1-\varepsilon }\\ =&I_0(V)-\frac{1}{p_\alpha +1}\int _D\left( e^{-\frac{2+\alpha }{2}\varepsilon y}-1\right) |V|^{p_\alpha +1}+k a_1\varepsilon +o(\varepsilon ). \end{aligned} \end{aligned}$$

It follows from Lemma 2.1 that

$$\begin{aligned} \begin{aligned} \frac{1}{p_\alpha +1}\int _D\left( e^{-\frac{2+\alpha }{2}\varepsilon y}-1\right) |V|^{p_\alpha +1}&=-\frac{(2+\alpha )\varepsilon }{2(p_\alpha +1)}\int _Dy|V|^{p_\alpha +1}+o(\varepsilon )\\&=-\varepsilon \frac{a_4(2+\alpha )}{2}\sum _{j=1}^k\xi _j+o(\varepsilon ). \end{aligned} \end{aligned}$$

It is easy to check that

$$\begin{aligned} I_0(V)-\sum _{i=1}^kI_0(V_i)= \frac{1}{p_\alpha +1}\int _D\left( \sum _{i=1}^kV_i^{p_\alpha +1}-|V|^{p_\alpha +1}\right) +\sum _{i,j=1,i>j}^k(-1)^{i+j}\int _D W_i^{p_\alpha }V_j. \end{aligned}$$

Since \(\Pi _j=O(e^{-\frac{2+\alpha }{2}\xi _j})=O(\varepsilon ^{\frac{3}{2}})\), \(j\ge 2\), from Lemma 2.1, we have

$$\begin{aligned} \begin{aligned}&I_0(V)-\sum _{i=1}^kI_0(V_i)\\&\quad =\frac{1}{p_\alpha +1} \sum _{\ell =1}^k\int _{D_\ell }\left( V_\ell ^{p_\alpha +1}-|V|^{p_\alpha +1}+(p_\alpha +1)\sum _{j<\ell }(-1)^{\ell +j}W_\ell ^{p_\alpha }V_j\right) +o(\varepsilon )\\&\quad =-\sum _{\ell =1}^k\sum _{j>\ell }(-1)^{\ell +j}\int _{D_\ell }W_\ell ^{p_\alpha }W_j+o(\varepsilon )\\&\quad =a_3\sum _{\ell =1}^{k-1}e^{-\frac{2+\alpha }{2}|\xi _{\ell +1}-\xi _\ell |}+o(\varepsilon ). \end{aligned} \end{aligned}$$

Next, we estimate \(I_0(V_i)\), \(i=1,2,\ldots ,k\).

Recall that

$$\begin{aligned} V_i=W_i+\Pi _i, ~~~\Pi _i(y)=\mathcal {T}(R_{\mu _i}),~\mu _i=e^{-\frac{p_\alpha -1}{2}\xi _i} \end{aligned}$$

and

$$\begin{aligned} I_0(V_i)=\left( \frac{p_\alpha -1}{2}\right) ^{\frac{p_\alpha +3}{p_\alpha -1}}J_0(PU_{\mu _i}). \end{aligned}$$

Thus, we find

$$\begin{aligned} J_0(PU_{\mu _i})= & {} \frac{2+\alpha }{2(N+\alpha )}\int _{\mathbb {R}^N}|x|^\alpha U_1^{p_\alpha +1}\\&+\frac{2+\alpha }{2(N+\alpha )}C_{\alpha ,N}H(0,0)\mu _i^{N-2}\int _{\mathbb {R}^N}|x|^\alpha U_1^{p_\alpha }+O(\mu _i^N). \end{aligned}$$

Since \(\mu _i=e^{-\frac{p_\alpha -1}{2}\xi _i}\), we find

$$\begin{aligned} \sum _{i=1}^kI_0(V_i)=ka_0+a_2H(0,0)e^{-(2+\alpha )\xi _1}+o(\varepsilon ). \end{aligned}$$

Hence, we can deduce

$$\begin{aligned} \begin{aligned} I_\varepsilon (V)&=ka_0+ka_1\varepsilon +a_2H(0,0)e^{-(2+\alpha )\xi _1} +a_3\sum _{\ell =1}^{k-1}e^{-\frac{2+\alpha }{2}|\xi _{\ell +1}-\xi _\ell |}\\&\quad +\,a_4\varepsilon \frac{2+\alpha }{2}\sum _{\ell =1}^k\xi _\ell +o(\varepsilon ). \end{aligned} \end{aligned}$$

By the definition of \(\xi _i, i=1,2,\ldots ,k\), we can obtain (2.8) immediately and the proof of Proposition 2.2 is concluded. \(\square \)

3 The finite-dimensional reduction

In this section, we perform the finite-dimensional procedure, which reduces problem (1.6) to a finite-dimensional problem on \(\mathbb {R}_+\).

For given \(\xi _i, ~i=1,2,\ldots ,k\), let

$$\begin{aligned} \Vert \phi \Vert _*=\sup \limits _{(y,\Theta )\in D}\left( \sum _{i=1}^ke^{-\frac{2+\alpha }{2}\sigma |y-\xi _i|}\right) ^{-1}|\phi (y,\Theta )|, \end{aligned}$$

where \(\sigma >0\) is a small constant. We denote \(\mathcal {C}_*\) by the continuous function space defined on D with finite norm defined as above.

Define

$$\begin{aligned} \tilde{Z}_i(x)=\mu _i\frac{\partial U_{\mu _i}}{\partial \mu _i}, \quad \quad \mu _i=e^{-\frac{p_\alpha -1}{2}\xi _i},~i=1,2,\ldots ,k. \end{aligned}$$

Then, \(\tilde{Z}_i(x)\) solves

$$\begin{aligned} -\Delta \tilde{Z}_i(x)=p_\alpha U_{\mu _i}^{p_\alpha -1}\tilde{Z}_i(x) ~~\text {in}~\mathbb {R}^N. \end{aligned}$$

Let \(P\tilde{Z}_i\) be the projection onto \(H_0^1(\Omega )\) of the function \(\tilde{Z}_i(x)\), that is,

$$\begin{aligned} {\left\{ \begin{array}{ll} -\Delta P\tilde{Z}_i=p_\alpha U_{\mu _i}^{p_\alpha -1}\tilde{Z}_i(x) ~~&{}\text {in}~\Omega ,\\ P\tilde{Z}_i=0&{}\text {on} ~\partial \Omega . \end{array}\right. } \end{aligned}$$

Set

$$\begin{aligned} Z_i(y,\Theta )=\mathcal {T}(P\tilde{Z}_i)(y,\Theta ). \end{aligned}$$

Then, \(Z_i\) satisfies

$$\begin{aligned} {\left\{ \begin{array}{ll} L(Z_i)=p_\alpha W_i^{p_\alpha -1}W^\prime _i ~~&{}\text {in}~D,\\ Z_i=0&{}\text {on}~\partial D. \end{array}\right. } \end{aligned}$$

First, we consider the following linear problem

$$\begin{aligned} {\left\{ \begin{array}{ll} \mathbb {L}_\varepsilon (\phi )=h+\sum _{j=1}^kc_jZ_j~~~&{}\text {in}~~D,\\ \phi =0&{}\text {on}~~\partial D,\\ \int _D Z_i\phi \mathrm{d}y\mathrm{d}\Theta =0,~~i=1,2,\ldots ,k, \end{array}\right. } \end{aligned}$$
(3.1)

where \(c_i\), \(i=1,2,\ldots ,k\), are some constants and

$$\begin{aligned} \mathbb {L}_\varepsilon (\phi )=L(\phi ) -(p_\alpha -\varepsilon )\sigma _\varepsilon e^{-\frac{2+\alpha }{2}\varepsilon y}|V|^{p_\alpha -1-\varepsilon }\phi . \end{aligned}$$

Lemma 3.1

Assume that there are sequences \(\varepsilon _n\rightarrow 0\) and points \(0<\xi _1^n<\xi ^n_2<\cdots <\xi _k^n\) with

$$\begin{aligned} \xi _1^n\rightarrow \infty ,~~\min \limits _{1\le i\le k-1}(\xi _{i+1}^n-\xi _{i}^n)\rightarrow +\infty ,~~ \xi _k^n=o(\varepsilon _n^{-1}), \end{aligned}$$

such that \(\phi _n\) solves (3.1) for scalars \(c_i^n\) and \(h_n\) with \(\Vert h_n\Vert _*\rightarrow 0\), then \(\lim \limits _{n\rightarrow \infty }\Vert \phi _n\Vert _*=0\).

Proof

We will first show that

$$\begin{aligned} \lim \limits _{n\rightarrow \infty }\Vert \phi _n\Vert _{L^\infty }=0. \end{aligned}$$

Arguing by contradiction, we may assume that \(\Vert \phi _n\Vert _{L^\infty }=1\). Multiplying (3.1) by \(Z_\ell ^n\) and integrating by parts, we find

$$\begin{aligned} \sum _{i=1}^kc_i^n\int _D Z_i^nZ_\ell ^n\mathrm{d}y\mathrm{d}\Theta =\int _D\mathbb {L}_{\varepsilon _n}(Z_\ell ^n)\phi _n\mathrm{d}y\mathrm{d}\Theta -\int _D h_nZ_\ell ^n\mathrm{d}y\mathrm{d}\Theta . \end{aligned}$$

Note that

$$\begin{aligned} \int _D Z_i^nZ_\ell ^n\mathrm{d}y\mathrm{d}\Theta =C\delta _{i\ell }+o(1) \end{aligned}$$

where \(\delta _{i\ell }\) is the Kronecker’s delta function. This defines an almost diagonal system in the \(c_i^n\)’s as \(n\rightarrow \infty \).

Thus, we have

$$\begin{aligned} \begin{aligned} \sum _{i=1}^kc_i^n\int _D Z_i^nZ_\ell ^n=&\int _D\left[ L(Z_\ell ^n) -(p_\alpha -\varepsilon _n)\sigma _{\varepsilon _n}e^{-\frac{2+\alpha }{2}\varepsilon _n y}|V|^{p_\alpha -1-\varepsilon _n}Z_\ell ^n\right] \phi _n -\int _D h_nZ_\ell ^n. \end{aligned} \end{aligned}$$
(3.2)

But

$$\begin{aligned} L(Z_\ell ^n)=p_\alpha W^{p_\alpha -1}(y-\xi _\ell ^n)W^\prime (y-\xi _\ell ^n), \end{aligned}$$

by the dominated convergence theorem, we know that \(\lim \limits _{n\rightarrow \infty }c_i^n=0\). Assume that \((y_n,\Theta _n)\in D\) is such that \(|\phi _n(y_n,\Theta _n)|=1\), we claim that there is an \(\ell \in \{1,\ldots ,k\}\) and a fixed \(R>0\), such that \(|\xi ^n_\ell -y_n|\le R\) for n large enough. Otherwise, we can suppose that \(|\xi ^n_\ell -y_n|\rightarrow +\infty \) as \(n\rightarrow +\infty \) for any \(\ell =1,2,\ldots ,k\). Then either \(|y_n|\rightarrow +\infty \) or \(|y_n|\) is bounded. Assume first that \(|y_n|\rightarrow +\infty \).

Define

$$\begin{aligned} \tilde{\phi }_n(y,\Theta )=\phi _n(y+y_n,\Theta ). \end{aligned}$$

By the standard elliptic regularity theory, we may assume that \(\tilde{\phi }_n\) converges uniformly over compact sets to a function \(\tilde{\phi }\). Set \(\tilde{\psi }=\mathcal {T}^{-1}(\tilde{\phi })\), then we have

$$\begin{aligned} \Delta \tilde{\psi }=0~~~\text {in}~~\mathbb {R}^N{\setminus }\{0\}. \end{aligned}$$

Due to \(\Vert \tilde{\phi }_n\Vert _{L^\infty }=1\), we see that \(|\tilde{\psi }(x)|\le |x|^{-\frac{N-2}{2}}\). Hence, \(\tilde{\psi }\) can extend smoothly to 0 to be a harmonic function in \(\mathbb {R}^N\) with this decay condition. So, \(\tilde{\phi }=0\) gives a contradiction. The fact that \(|y_n|\) cannot be bounded can be handled in similar way. Thus, there exists an integer \(\ell \in \{1,\ldots ,k\}\) and a positive number \(R>0\) such that for n large enough, \(|y_n-\xi _\ell ^n|\le R.\)

Define again

$$\begin{aligned} \tilde{\phi }_n(y,\Theta )=\phi _n(y+\xi _\ell ^n,\Theta ). \end{aligned}$$

Thus, \(\tilde{\phi }_n\) converges uniformly over compact sets to a function \(\tilde{\phi }\). Set again that \(\tilde{\psi }=\mathcal {T}^{-1}(\tilde{\phi })\). Hence, \(\tilde{\psi }\) is a nontrivial solution of

$$\begin{aligned} \Delta \tilde{\psi }+p_\alpha |x|^\alpha U_1^{p_\alpha -1}\tilde{\psi }=0~~\text {in}~~\mathbb {R}^N{\setminus }\{0\}. \end{aligned}$$

Moreover, \(|\tilde{\psi }(x)|\le C|x|^{-\frac{N-2}{2}}\). Therefore, we obtain a classical solution in \(\mathbb {R}^N{\setminus }\{0\}\) decaying at infinity. It follows from [12] that it equals a linear combination of the \(\{\tilde{Z}_i\}\) provided that \(\alpha \) is not an even integer. However, the orthogonality conditions imply \(\tilde{\phi }=0\). This is again a contradiction. Thus, we can deduce that \(\lim \limits _{n\rightarrow \infty }\Vert \phi _n\Vert _{L^\infty }=0.\)

Next we shall establish that

$$\begin{aligned} \lim \limits _{n\rightarrow \infty }\Vert \phi _n\Vert _*\rightarrow 0. \end{aligned}$$

Now we see that (3.1) possesses the following form

$$\begin{aligned} -\phi _n^{\prime \prime }+\frac{(2+\alpha )^2}{4}\phi _n-\left( \frac{p_\alpha -1}{2}\right) ^2 \Delta _{\mathbb {S}^{N-1}}\phi _n=g_n, \end{aligned}$$
(3.3)

where

$$\begin{aligned} g_n=h_n+(p_\alpha -\varepsilon _n)\sigma _\varepsilon e^{-\frac{2+\alpha }{2}\varepsilon _n y}|V|^{p_\alpha -1-\varepsilon _n}\phi _n+\sum _{i=1}^n c_i^nZ_i^n. \end{aligned}$$

If \(0<\sigma <\min \{p_\alpha -1,1\}\), we find

$$\begin{aligned} |g_n(y)|\le \theta _n\sum _{i=1}^ke^{-\frac{2+\alpha }{2}\sigma |y-\xi _i^n|}~~\text {with}~~\theta _n\rightarrow 0. \end{aligned}$$

Choosing \(C>0\) large enough, we see that

$$\begin{aligned} \varphi _n(y)=C\theta _n\sum _{i=1}^ke^{-\frac{2+\alpha }{2}\sigma |y-\xi _i^n|} \end{aligned}$$

is a supersolution of (3.3), and \(-\varphi _n(y)\) will be a subsolution of (3.3). Thus,

$$\begin{aligned} |\phi _n|\le C\theta _n\sum _{i=1}^ke^{-\frac{2+\alpha }{2}\sigma |y-\xi _i^n|}. \end{aligned}$$

The following proposition is a direct consequence of Proposition 1 in [9] combining with Lemma 3.1.

Proposition 3.2

There exist positive numbers \(\varepsilon _0, \delta _0, R_0\), such that if

$$\begin{aligned} R_0<\xi _1,~~R_0<\min \limits _{i=1,\ldots ,k-1}(\xi _{i+1}-\xi _{i}),~~\xi _k<\frac{\delta _0}{\varepsilon }, \end{aligned}$$
(3.4)

then for all \(0<\varepsilon <\varepsilon _0\) and \(h\in \mathcal {C}_*\), problem (3.1) has a unique solution \(\phi =T_\varepsilon (h)\). Moreover, there exists \(C>0\) such that

$$\begin{aligned} \Vert T_\varepsilon (h)\Vert _*\le C\Vert h\Vert _*,~~~|c_i|\le C\Vert h\Vert _*. \end{aligned}$$

For later purposes, we need to understand the differentiability of the operator \(T_\varepsilon \) on the variables \(\xi _i\). We will use the notation \(\xi =(\xi _1,\xi _2,\ldots ,\xi _k)\). We also consider the space \(L(\mathcal {C}_*)\) of the linear operator of \(\mathcal {C}_*\). We have the following result.

Proposition 3.3

Under the same assumptions of Proposition 3.2, the map \(\xi \rightarrow T_\varepsilon \) with values in \(L(\mathcal {C}_*)\) is of class \(C^1\). Besides, there is a constant \(C>0\) such that

$$\begin{aligned} \Vert D_\xi T_\varepsilon \Vert _{L({\mathcal {C}_*})}\le C \end{aligned}$$

uniformly on the vectors \(\xi \) satisfying (3.4).

Proof

Fix \(h\in \mathcal {C}_*\), and let \(\phi =T_\varepsilon (h)\). We are interested in studying the differentiability of \(\phi \) with respect to \(\xi _\ell \) for \(\ell =1,2,\ldots ,k\). Recall that \(\phi \) satisfies

$$\begin{aligned} {\left\{ \begin{array}{ll} \mathbb {L}_\varepsilon (\phi )=h+\sum _{j=1}^kc_jZ_j~~~&{}\text {in}~~D,\\ \phi =0&{}\text {on}~~\partial D,\\ \int _D Z_i\phi \mathrm{d}y\mathrm{d}\Theta =0,~~i=1,2,\ldots ,k, \end{array}\right. } \end{aligned}$$

for certain constants \(c_i\). Differentiating the above equation with respect to \(\xi _\ell ,~\ell =1,\ldots ,k.\) Define \(Y=\partial _{\xi _\ell }\phi \) and \(d_i=\partial _{\xi _\ell }c_i\), we find

$$\begin{aligned} {\left\{ \begin{array}{ll} \mathbb {L}_\varepsilon (Y)=(p_\alpha -\varepsilon )\sigma _\varepsilon e^{-\frac{2+\alpha }{2}\varepsilon y}(\partial _{\xi _\ell }|V|^{p_\alpha -1-\varepsilon })\phi +c_\ell \partial _{\xi _\ell }Z_\ell +\sum _{j=1}^kd_jZ_j~~~&{}\text {in}~~D,\\ Y=0&{}\text {on}~~\partial D,\\ \int _D (YZ_i+\phi \partial _{\xi _\ell }Z_i)\mathrm{d}y\mathrm{d}\Theta =0,~~i=1,2,\ldots ,k. \end{array}\right. } \end{aligned}$$

Set \(\chi =Y-\sum _{i=1}^kb_iZ_i\), where the constants \(b_i\) satisfy

$$\begin{aligned} \begin{aligned}&\sum _{i=1}^k b_i\int _D Z_iZ_j\mathrm{d}y\mathrm{d}\Theta =0,~~~j\ne \ell ,\\&\sum _{i=1}^k b_i\int _D Z_iZ_\ell \mathrm{d}y\mathrm{d}\Theta =-\int _D\phi \partial _{\xi _\ell }Z_\ell \mathrm{d}y\mathrm{d}\Theta . \end{aligned} \end{aligned}$$

This is also an almost diagonal system and \(Y=\chi +\sum _{j=1}^kb_jZ_j,\) where\(\int _D\chi Z_j \mathrm{d}y\mathrm{d}\Theta =0,~j=1,2,\ldots ,k.\) Moreover, it is easy to see that \(\chi \) satisfies

$$\begin{aligned} {\left\{ \begin{array}{ll} \mathbb {L}_\varepsilon (\chi )=g+\sum _{j=1}^kd_jZ_j~~~&{}\text {in}~~D,\\ \chi =0&{}\text {on}~~\partial D,\\ \int _D\chi Z_j \mathrm{d}y\mathrm{d}\Theta =0,~j=1,2,\ldots ,k, \end{array}\right. } \end{aligned}$$

where

$$\begin{aligned} g=(p_\alpha -\varepsilon )\sigma _\varepsilon e^{-\frac{2+\alpha }{2}\varepsilon y}(\partial _{\xi _\ell }|V|^{p_\alpha -1-\varepsilon })\phi +c_\ell \partial _{\xi _\ell }Z_\ell -\sum _{j=1}^k b_j\mathbb {L}_\varepsilon (Z_j). \end{aligned}$$

Then, we find

$$\begin{aligned} \chi =T_\varepsilon (g) \end{aligned}$$

and

$$\begin{aligned} \partial _{\xi _\ell }\phi =T_\varepsilon (g)+\sum _{j=1}^k b_jZ_j. \end{aligned}$$

By Proposition 3.2, we find

$$\begin{aligned} \Vert T_\varepsilon (g)\Vert _*\le C\Vert g\Vert _*. \end{aligned}$$

Since

$$\begin{aligned} \Vert g\Vert _*\le C\left( \Vert \phi \Vert _*+|c_\ell |+\sum _{j=1}^k|b_j|\right) \end{aligned}$$

and

$$\begin{aligned} |b_i|\le C\Vert \phi \Vert _*,~~|c_i|\le C\Vert h\Vert _*,~~\Vert \phi \Vert _*\le C\Vert h\Vert _*. \end{aligned}$$

Thus, we can obtain that \(\Vert \partial _{\xi _\ell }\phi \Vert _*\le C\Vert h\Vert _*\), and \(\partial _{\xi _\ell }\phi \) depends continuously on \(\xi \) for this norm. \(\square \)

Now we consider

$$\begin{aligned} {\left\{ \begin{array}{ll} L(V+\phi )-\sigma _\varepsilon e^{-\frac{2+\alpha }{2}\varepsilon y}|V+\phi |^{p_\alpha -1-\varepsilon }(V+\phi ) =\sum \limits _{j=1}^kc_jZ_j~~~\text {in}~D,\\ \phi =0~~~\text {on}~\partial D,\\ \int _D Z_i\phi \mathrm{d}y\mathrm{d}\Theta =0,\quad i=1,2,\ldots ,k. \end{array}\right. } \end{aligned}$$
(3.5)

In order to solve problem (3.5), we rewrite it as

$$\begin{aligned} {\left\{ \begin{array}{ll} \mathbb {L}_\varepsilon (\phi )=N_\varepsilon (\phi )+R_\varepsilon +\sum \limits _{j=1}^kc_jZ_j~~~\text {in}~~D,\\ \phi =0\quad \text {on}~~\partial D,\\ \int _D Z_i\phi \mathrm{d}y\mathrm{d}\Theta =0,\quad i=1,2,\ldots ,k, \end{array}\right. } \end{aligned}$$
(3.6)

where

$$\begin{aligned} N_\varepsilon (\phi )=\sigma _\varepsilon e^{-\frac{2+\alpha }{2}\varepsilon y} \left( |V+\phi |^{p_\alpha -1-\varepsilon }(V+\phi )-|V|^{p_\alpha -1-\varepsilon }V-(p_\alpha -\varepsilon )|V|^{p_\alpha -1-\varepsilon }\phi \right) \end{aligned}$$

and

$$\begin{aligned} R_\varepsilon= & {} \sigma _\varepsilon e^{-\frac{2+\alpha }{2}\varepsilon y}|V|^{p_\alpha -1-\varepsilon }V-\sum _{i=1}^k(-1)^iW_i^{p_\alpha }. \end{aligned}$$

Let us fix a large number \(M>0\), \(\xi \) satisfies the following conditions

$$\begin{aligned} \xi _1>\frac{1}{2}\log \frac{1}{M\varepsilon },~~ \min \limits _{1\le i\le k-1}(\xi _{i+1}-\xi _i)>\log \frac{1}{M\varepsilon },~~ \xi _k<k\log \frac{1}{M\varepsilon }. \end{aligned}$$
(3.7)

In order to prove that (3.6) is uniquely solvable in the set that \(\Vert \phi \Vert _*\) is small, we need to estimate \(R_\varepsilon \) and \(N_\varepsilon (\phi )\).

Lemma 3.4

If \(N\ge 3\), then

$$\begin{aligned} \begin{aligned}&\Vert N_\varepsilon (\phi )\Vert _*\le C\Vert \phi \Vert _*^{\min \{p_\alpha -\varepsilon ,2\}},\\&\left\| \frac{\partial N_\varepsilon (\phi )}{\partial \phi }\right\| _*\le C\Vert \phi \Vert _*^{\min \{p_\alpha -1-\varepsilon ,1\}}. \end{aligned} \end{aligned}$$
(3.8)

Proof

Since

$$\begin{aligned} |N_\varepsilon (\phi )|\le {\left\{ \begin{array}{ll} C|\phi |^{p_\alpha -\varepsilon },~~&{}p_\alpha -1\le 1,\\ C|V|^{p_\alpha -2-\varepsilon }\phi ^2+C|\phi |^{p_\alpha -\varepsilon },~&{}p_\alpha -1>1. \end{array}\right. } \end{aligned}$$

First, we consider the case \(p_\alpha -1\le 1\).

$$\begin{aligned} \begin{aligned} |N_\varepsilon (\phi )|&\le C\Vert \phi \Vert _*^{p_\alpha -\varepsilon }\left( \sum _{i=1}^ke^{-\frac{2+\alpha }{2}\sigma |y-\xi _i|}\right) ^{p_\alpha -\varepsilon }\\&\le C\Vert \phi \Vert _*^{p_\alpha -\varepsilon }\left( \sum _{i=1}^ke^{-\frac{2+\alpha }{2}\sigma |y-\xi _i|}\right) . \end{aligned} \end{aligned}$$

where we have used the fact that

$$\begin{aligned} \sum _{i=1}^ke^{-\frac{2+\alpha }{2}\sigma |y-\xi _i|}\le C. \end{aligned}$$

Thus, the result follows.

Now we show the result holds for \(p_\alpha -1>1\).

$$\begin{aligned} \begin{aligned} |N_\varepsilon (\phi )|\le&C\Vert \phi \Vert _*^2|V|^{p_\alpha -2-\varepsilon }\left( \sum _{i=1}^ke^{-\frac{2+\alpha }{2}\sigma |y-\xi _i|}\right) ^2 +C\Vert \phi \Vert _*^{p_\alpha -\varepsilon }\left( \sum _{i=1}^ke^{-\frac{2+\alpha }{2}\sigma |y-\xi _i|}\right) ^{p_\alpha -\varepsilon }\\ \le&C\left( \Vert \phi \Vert _*^{p_\alpha -\varepsilon }+\Vert \phi \Vert _*^2\right) \left( \sum _{i=1}^ke^{-\frac{2+\alpha }{2}\sigma |y-\xi _i|}\right) . \end{aligned} \end{aligned}$$

Thus,

$$\begin{aligned} \Vert N_\varepsilon (\phi )\Vert _*\le C\Vert \phi \Vert _*^{\min \{{p_\alpha -\varepsilon ,2}\}}. \end{aligned}$$

The other terms can be estimated similarly, and the proof of the lemma is completed. \(\square \)

Lemma 3.5

If \(N\ge 3\), then

$$\begin{aligned} \Vert R_\varepsilon \Vert _*\le C\varepsilon ^{\frac{1+\tau }{2}},~~\Vert \partial _\xi R_\varepsilon \Vert _*\le C\varepsilon ^{\frac{1+\tau }{2}}, \end{aligned}$$
(3.9)

where \(\tau >0\) is a small constant.

Proof

We give here the proof of the first one only. The second one can be obtained similarly.

Note that

$$\begin{aligned} R_\varepsilon= & {} (\sigma _\varepsilon -1)e^{-\frac{2+\alpha }{2}\varepsilon y}|V|^{p_\alpha -1-\varepsilon }V+e^{-\frac{2+\alpha }{2}\varepsilon y}\left( |V|^{p_\alpha -1-\varepsilon }V-|V|^{p_\alpha -1}V\right) \\&+\,|V|^{p_\alpha -1}V\left( e^{-\frac{2+\alpha }{2}\varepsilon y}-1\right) +|V|^{p_\alpha -1}V-\sum _{i=1}^k(-1)^iV_i^{p_\alpha }\\&+\sum _{i=1}^k(-1)^iV_i^{p_\alpha }-\sum _{i=1}^k(-1)^iW_i^{p_\alpha }\\= & {} :J_1+J_2+J_3+J_4+J_5. \end{aligned}$$

Recalling that

$$\begin{aligned} V=\sum _{i=1}^k(-1)^iV_i,~~0\le V_i\le W_i. \end{aligned}$$

Thus, we find

$$\begin{aligned} \begin{aligned}&|J_1|\le C\varepsilon e^{-\frac{2+\alpha }{2}\varepsilon y}|V|^{p_\alpha -\varepsilon }\le C\varepsilon \sum _{i=1}^ke^{-\frac{2+\alpha }{2}\sigma |y-\xi _i|}, \\&|J_2|\le C\varepsilon |\log V||V|^{p_\alpha -1}\le C\varepsilon \sum _{i=1}^ke^{-\frac{2+\alpha }{2}\sigma |y-\xi _i|},\\&|J_3|=\left| \left( e^{-\frac{2+\alpha }{2}\varepsilon y}-1\right) |V|^{p_\alpha -1}V\right| \le C\varepsilon y|V|^{p_\alpha }\le C\varepsilon \sum _{i=1}^ke^{-\frac{2+\alpha }{2}\sigma |y-\xi _i|}. \end{aligned} \end{aligned}$$

Next we estimate \(J_4\) and \(J_5\).

Define

$$\begin{aligned} \chi _\ell =\frac{\xi _{\ell -1}+\xi _\ell }{2},\,\,\,\,\ell =1,2,\ldots ,k+1,~\text {where}~~\xi _0=\inf \limits _{(y,\Theta )\in D}|y|,~~\xi _{k+1}=+\infty . \end{aligned}$$

Thus, for \(\chi _\ell \le y<\chi _{\ell +1}\), we have

$$\begin{aligned} \begin{aligned} |J_4|&=\Big ||V|^{p_\alpha -1}V-\sum _{i=1}^k(-1)^iV_i^{p_\alpha }\Big |\le CV_\ell ^{p_\alpha -1}\left( \sum _{j\ne \ell }V_j\right) \\&\le \,C\sum _{j\ne \ell }e^{-\frac{2+\alpha }{2}(p_\alpha -1)|y-\xi _\ell |}e^{-\frac{2+\alpha }{2}|y-\xi _j|}\\&\le \,Ce^{-\frac{2+\alpha }{2}\sigma |y-\xi _\ell |}\sum _{j\ne \ell }e^{-\frac{2+\alpha }{2}(p_\alpha -\sigma -1)|y-\xi _\ell |}e^{-\frac{2+\alpha }{2}|y-\xi _j|}\\&\le \,Ce^{-\frac{2+\alpha }{2}\sigma |y-\xi _\ell |}\sum _{j\ne \ell }e^{-\frac{(2+\alpha )(1+\tau )}{4}|\xi _\ell -\xi _{\ell -1}|}\\&\le \,C\varepsilon ^{\frac{1+\tau }{2}}\sum _{i=1}^ke^{-\frac{2+\alpha }{2}\sigma |y-\xi _i|} \end{aligned} \end{aligned}$$

and

$$\begin{aligned} \begin{aligned} |J_5|&=\left| \sum _{i=1}^k\left( V_i^{p_\alpha }-W_i^{p_\alpha }\right) \right| \le C\sum _{i=1}^kW_i^{p_\alpha -1}|\Pi _i|\\&\le \,CR_{\mu _1}\left( e^{-\frac{p_\alpha -1}{2}y},\Theta \right) \sum _{i=1}^ke^{-\frac{2+\alpha }{2}(p_\alpha -1)|y-\xi _i|}e^{-\frac{2+\alpha }{2}y}, ~\mu _1=e^{-\frac{p_\alpha -1}{2}\xi _1}\\&\le \,C\varepsilon ^{\frac{1+\tau }{2}}\sum _{i=1}^ke^{-\frac{2+\alpha }{2}\sigma |y-\xi _i|}. \end{aligned} \end{aligned}$$

Therefore, \(\Vert R_\varepsilon \Vert _*\le C\varepsilon ^{\frac{1+\tau }{2}}\) and the results follow. \(\square \)

The next proposition enables us to reduce the problem of finding a solution for (1.6) to a finite-dimensional problem.

Proposition 3.6

Suppose that condition (3.7) holds. Then there exists a positive constant C such that, for \(\varepsilon >0\) small enough, problem (3.6) admits a unique solution \(\phi =\phi (\xi )\), which satisfies

$$\begin{aligned} \Vert \phi \Vert _*\le C\varepsilon ^{\frac{1+\tau }{2}}. \end{aligned}$$

Moreover, \(\phi (\xi )\) is of class \(C^1\) on \(\xi \) with the \(\Vert \cdot \Vert _*\)-norm, and

$$\begin{aligned} \Vert D_\xi \phi \Vert _*\le C\varepsilon ^{\frac{1+\tau }{2}}, \end{aligned}$$

where \(\tau >0\) is a small constant.

Proof

Define

$$\begin{aligned} A_\varepsilon (\phi ):=T_\varepsilon (N_\varepsilon (\phi )+R_\varepsilon ), \end{aligned}$$

then we know that problem (3.6) is equivalent to the fixed point problem \(\phi =A_\varepsilon (\phi )\). We will use the contraction mapping theorem to solve it.

Set

$$\begin{aligned} E_\rho =\{\phi \in \mathcal {C}_*:\Vert \phi \Vert _*\le \rho \varepsilon ^{\frac{1+\tau }{2}}\}, \end{aligned}$$

where \(\rho >0\) will be fixed later.

We will show that \(A_\varepsilon \) is a contraction map from \(E_\rho \) to \(E_\rho \).

In fact, for \(\varepsilon >0\) small enough, we find

$$\begin{aligned} \begin{aligned} \Vert A_\varepsilon (\phi )\Vert _*\le C\Vert N_\varepsilon (\phi )+R_\varepsilon \Vert _* \le C\left( (\rho \varepsilon )^{\min \{p_\alpha -\varepsilon ,2\}}+\varepsilon ^{\frac{1+\tau }{2}}\right) \le \rho \varepsilon ^{\frac{1+\tau }{2}}, \end{aligned} \end{aligned}$$

provided \(\rho \) is chosen large enough, but independent of \(\varepsilon \).

Thus, \(A_\varepsilon \) maps \(E_\rho \) into itself. Moreover,

$$\begin{aligned} \begin{aligned} |N_\varepsilon (\phi _1)-N_\varepsilon (\phi _2)|\le&|\partial _\phi N_\varepsilon (t\phi _1+(1-t)\phi _2)||\phi _1-\phi _2|\\ \le&C\left( \rho \varepsilon ^{\frac{1+\tau }{2}}\right) ^{\min \{p_\alpha -1-\varepsilon ,1\}}|\phi _1-\phi _2|. \end{aligned} \end{aligned}$$

Hence,

$$\begin{aligned} \begin{aligned} \Vert A_\varepsilon (\phi _1)-A_\varepsilon (\phi _2)\Vert _*\le&C\left( \rho \varepsilon ^{\frac{1+\tau }{2}}\right) ^{\min \{p_\alpha -1-\varepsilon ,1\}}\Vert \phi _1-\phi _2\Vert _*\\ \le&\frac{1}{2}\Vert \phi _1-\phi _2\Vert _*. \end{aligned} \end{aligned}$$

Thus, there is a unique \(\phi \in E_\rho \), such that \(\phi =A_\varepsilon (\phi )\).

Now we consider the differentiability of \(\xi \rightarrow \phi (\xi )\).

Let

$$\begin{aligned} B(\xi ,\phi )=\phi -T_\varepsilon (N_\varepsilon (\phi )+R_\varepsilon ). \end{aligned}$$

First, we have \(B(\xi ,\phi (\xi ))=0\). Let us write

$$\begin{aligned} D_\phi B(\xi ,\phi )[\psi ]=\psi -T_\varepsilon (\psi D_\phi N_\varepsilon (\phi ))=\psi +M(\psi ), \end{aligned}$$

where

$$\begin{aligned} M(\psi )=-T_\varepsilon (\psi D_\phi N_\varepsilon (\phi )). \end{aligned}$$

From (3.8), we find

$$\begin{aligned} \Vert M(\psi )\Vert _*\le C\varepsilon ^{\frac{1+\tau }{2}\min \{p_\alpha -1-\varepsilon ,1\}}\Vert \psi \Vert _*. \end{aligned}$$

Thus, the linear operator \(D_\phi B(\varepsilon ,\phi )\) is invertible in \(\mathcal {C}_*\) with uniformly bounded inverse depending continuously on its parameters. Differentiating with respect to \(\xi \), we deduce

$$\begin{aligned} D_\xi B(\xi ,\phi )=-D_\xi T_\varepsilon [N_\varepsilon (\phi )+R_\varepsilon ]-T_\varepsilon [D_\xi N_\varepsilon (\xi ,\phi )+D_\xi R_\varepsilon ], \end{aligned}$$

where all these expressions depend continuously on their parameters.

By the implicit function theorem, we see that \(\phi (\xi )\) is of class \(C^1\) and

$$\begin{aligned} D_\xi \phi =-\left( D_\phi B(\xi ,\phi )\right) ^{-1}[D_\xi B(\xi ,\phi )]. \end{aligned}$$

Thus,

$$\begin{aligned} \Vert D_\xi (\phi )\Vert _*\le C\left( \Vert N_\varepsilon (\phi )+R_\varepsilon \Vert _*+\Vert D_\xi N_\varepsilon (\xi ,\phi )\Vert _*+\Vert D_\xi R_\varepsilon \Vert _*\right) \le C\varepsilon ^{\frac{1+\tau }{2}}. \end{aligned}$$

The proof of Proposition 3.6 is concluded. \(\square \)

4 Proof of the main result

In this section, we will prove Theorem 1.1. As deduced in the introduction, we need to verify Theorem 1.4. To do this, we will choose \(\xi \) such that \(V+\phi \) is a solution of (1.6), where \(\phi \) is the map obtained in Proposition 3.6.

Recall that

$$\begin{aligned} \begin{aligned} I_\varepsilon (v)=\,&\frac{1}{2}\int _D\left( |v^\prime |^2+\frac{(2+\alpha )^2}{4} |v|^2\right) \mathrm{d}y\mathrm{d}\Theta +\frac{1}{2}\left( \frac{p_\alpha -1}{2}\right) ^2\int _D|\nabla _{\mathbb {S}^{N-1}}v|^2\mathrm{d}y\mathrm{d}\Theta \\&-\frac{\sigma _\varepsilon }{p_\alpha +1-\varepsilon }\int _D e^{-\frac{2+\alpha }{2}\varepsilon y}|v|^{p_\alpha +1-\varepsilon }\mathrm{d}y\mathrm{d}\Theta . \end{aligned} \end{aligned}$$
(4.1)

Define

$$\begin{aligned} K_\varepsilon (\xi )=I_\varepsilon (V+\phi ). \end{aligned}$$

It is now well known that if \(\xi \) is a critical point of \(K_\varepsilon (\xi )\), then \(V+\phi \) is a solution of (1.6). Next, we will prove that \(K_\varepsilon (\xi )\) has a critical point. To this end, we need the next lemma, which is important in finding the critical point of \(K_\varepsilon \).

Lemma 4.1

The following expansion holds

$$\begin{aligned} K_\varepsilon (\xi )=I_\varepsilon (V)+O(\varepsilon ^{1+\tau }), \end{aligned}$$
(4.2)

where \(O(\varepsilon ^{1+\tau })\) is uniformly in the \(C^1\)-sense on the vectors \(\xi \) satisfying (3.4).

Proof

Using the Taylor expansion

$$\begin{aligned} F(u+v)=F(u)+dF(u)[v]+\int _0^1(1-t)d^2F(u+tv)[v,v]dt \end{aligned}$$

and the fact that \(\nabla I_\varepsilon (V+\phi )[\phi ]=0\), we have

$$\begin{aligned} \begin{aligned}&I_\varepsilon (V+\phi )-I_\varepsilon (V)=\int _0^1\nabla ^2I_\varepsilon (V+t\phi )[\phi ,\phi ]tdt\\&\quad =\int _0^1\left( \int _D(N_\varepsilon (\phi )+R_\varepsilon )\phi +(p_\alpha -\varepsilon )\sigma _\varepsilon \int _D e^{-\frac{2+\alpha }{2}\varepsilon y}\right. \nonumber \\&\qquad \left. \times \left( |V|^{p_\alpha -1-\varepsilon }-|V+t\phi |^{p_\alpha -1-\varepsilon }\right) \phi ^2\right) tdt. \end{aligned} \end{aligned}$$

Since \(\Vert \phi \Vert _*\le C\varepsilon ^{\frac{1+\tau }{2}}\), we find that

$$\begin{aligned} \begin{aligned} \int _D|(N_\varepsilon (\phi )+R_\varepsilon )\phi |\le&C(\Vert N_\varepsilon (\phi )\Vert _*+\Vert R_\varepsilon \Vert _*)\Vert \phi \Vert _* \int _D\left( \sum _{i=1}^ke^{-\frac{2+\alpha }{2}\sigma |y-\xi _i|}\right) ^2\\ \le&C(\Vert N_\varepsilon (\phi )\Vert _*+\Vert R_\varepsilon \Vert _*)\Vert \phi \Vert _*={O(\varepsilon ^{{1+\tau }})}, \end{aligned} \end{aligned}$$

and

$$\begin{aligned} \begin{aligned}&\int _D\left| |V|^{p_\alpha -1-\varepsilon }-|V+t\phi |^{p_\alpha -1-\varepsilon }\right| \phi ^2\\&\quad \le C\Vert \phi \Vert _*^2\int _D \left( \sum _{i=1}^ke^{-\frac{2+\alpha }{2}\sigma |y-\xi _i|}\right) ^2\\&\quad \le C\Vert \phi \Vert _*^2. \end{aligned} \end{aligned}$$

Thus,

$$\begin{aligned} I_\varepsilon (V+\phi )=I_\varepsilon (V)+O(\varepsilon ^{1+\tau }). \end{aligned}$$

Differentiating with respect to \(\xi _\ell \), we see that

$$\begin{aligned} \begin{aligned}&\partial _{\xi _\ell }\left( I_\varepsilon (V+\phi )-I_\varepsilon (V)\right) \\&\quad =\int _0^1\int _D\partial _{\xi _\ell }\left[ (N_\varepsilon (\phi )+R_\varepsilon )\phi \right] tdt\\&\qquad +(p_\alpha -\varepsilon )\sigma _\varepsilon \int _0^1\int _D e^{-\frac{2+\alpha }{2}\varepsilon y}\partial _{\xi _\ell } \left[ \left( |V|^{p_\alpha -1-\varepsilon }-|V+t\phi |^{p_\alpha -1-\varepsilon }\right) \phi ^2\right] tdt. \end{aligned} \end{aligned}$$

In a similar way, we have that

$$\begin{aligned} \partial _{\xi _\ell }I_\varepsilon (V+\phi )=\partial _{\xi _\ell }I_\varepsilon (V)+O(\varepsilon ^{1+\tau }). \end{aligned}$$

Thus, the result follows. \(\square \)

Proof of Theorem 1.4

Recalling that

$$\begin{aligned} \begin{aligned} \xi _1 =&-\frac{1}{2+\alpha }\log \varepsilon +\frac{2}{2+\alpha }\log \Lambda _1, \\ \xi _{i+1}-\xi _i=&-\frac{2}{2+\alpha } \log \varepsilon -\frac{2}{2+\alpha }\log \Lambda _{i+1},~~i=1,2,\ldots ,k-1, \end{aligned} \end{aligned}$$

where \(\delta<\Lambda _i<\frac{1}{\delta }\), \(\delta >0\) is a fixed constant. To simplify the notation, we denote \(\Lambda =(\Lambda _1,\Lambda _2,\ldots ,\Lambda _k)\). Thus, it is sufficient to find a critical point of the function

$$\begin{aligned} \widetilde{K}_\varepsilon (\Lambda )=\varepsilon ^{-1}\left( K_\varepsilon (\xi (\Lambda ))-ka_0\right) . \end{aligned}$$

From Lemma 4.1 and Proposition 2.2, we have

$$\begin{aligned} \widetilde{K}_\varepsilon (\Lambda )=\Psi _k(\Lambda )+ka_1-\frac{k^2}{2}a_4\log \varepsilon +o(1), \end{aligned}$$

where the term o(1) goes to 0 uniformly as \(\varepsilon \rightarrow 0\).

It is easy to see that the function

$$\begin{aligned} \Lambda _1\rightarrow ka_4\log \Lambda _1+\frac{a_2H(0,0)}{\Lambda _1^2} \end{aligned}$$

has a stable minimum point \(\Lambda _1^*=\left( \frac{2a_2H(0,0)}{ka_4}\right) ^{\frac{1}{2}}\) on \((0,+\infty )\), and for \(i=2,\ldots ,k\), the function

$$\begin{aligned} \Lambda _i\rightarrow a_3\Lambda _i-(k-i+1)a_4\log \Lambda _i \end{aligned}$$

also has a stable minimum point \(\Lambda _i^*=\frac{(k-i+1)a_4}{a_3}\) on \((0,+\infty )\). Thus, the function \(\Psi _k(\Lambda )\) has a stable minimum point \(\Lambda ^*=(\Lambda _1^*,\ldots ,\Lambda _k^*)\). Therefore, for \(\varepsilon \) small enough, there exists a critical point \(\Lambda ^\varepsilon =(\Lambda _1^\varepsilon ,\ldots ,\Lambda _k^\varepsilon )\) of the function \(\widetilde{K}_\varepsilon (\Lambda )\), such that \(\Lambda _i^\varepsilon \rightarrow \Lambda _i^*\) as \(\varepsilon \rightarrow 0\) for \(i=1,2,\ldots ,k\).

For the \(\Lambda _i^\varepsilon ~~(i=1,\ldots ,k)\) obtained above, let

$$\begin{aligned} \xi _1^\varepsilon =\frac{2}{2+\alpha }\log \frac{\Lambda _1^\varepsilon }{\varepsilon ^{\frac{1}{2}}}, ~~\xi _i^\varepsilon =\frac{2}{2+\alpha }\log \frac{\Lambda _1^\varepsilon }{\Lambda _2^\varepsilon \ldots \Lambda _i^\varepsilon \varepsilon ^{\frac{2i-1}{2}}}, ~~i=2,3,\ldots ,k. \end{aligned}$$

Hence, \(\xi ^\varepsilon =(\xi _1^\varepsilon ,\ldots ,\xi _k^\varepsilon )\) is a critical point of \(K_\varepsilon (\xi )\) and \(V+\phi (\xi ^\varepsilon )\) is a solution of (1.6). \(\square \)

Proof of Theorem 1.1

Note that \(\Lambda _i^\varepsilon =\Lambda _i^*+o(1), i=1,2,\ldots ,k\) as \(\varepsilon \rightarrow 0\). Then

$$\begin{aligned} \begin{aligned} \xi _1^\varepsilon&=\frac{2}{2+\alpha }\log \frac{\Lambda _1^*}{\varepsilon ^{\frac{1}{2}}}+o(1),\\ \xi _i^\varepsilon&=\frac{2}{2+\alpha }\log \frac{\Lambda _1^*}{\Lambda _2^*\ldots \Lambda _i^*\varepsilon ^{\frac{2i-1}{2}}}+o(1), ~~i=2,3,\ldots ,k. \end{aligned} \end{aligned}$$

Using the fact that \(e^{-\frac{p_\alpha -1}{2}\xi ^\varepsilon _i}=M_i\varepsilon ^{\frac{2i-1}{N-2}}(1+o(1))\), \(i=1,\ldots ,k\), where

$$\begin{aligned} \begin{aligned} M_1=\left( \frac{1}{\Lambda _1^*}\right) ^{\frac{2}{N-2}},\,\,\, M_i=\left( \frac{\Lambda _2^*\ldots \Lambda _i^*}{\Lambda _1^*}\right) ^{\frac{2}{N-2}},~i=2,\ldots ,k. \end{aligned} \end{aligned}$$
(4.3)

Thus, by the transformation (1.5), we find

$$\begin{aligned} u_\varepsilon (x)= C_{\alpha ,N}\sum _{i=1}^k(-1)^i \left( \frac{M_i^{\frac{2+\alpha }{2}}\varepsilon ^{\frac{(2+\alpha )(2i-1)}{2(N-2)}}}{M_i^{2+\alpha }\varepsilon ^{\frac{(2+\alpha )(2i-1)}{N-2}}+|x|^{2+\alpha }}\right) ^{\frac{N-2}{2+\alpha }}(1+o(1)), \end{aligned}$$

where \(o(1)\rightarrow 0\) uniformly on compact subsets of \(\Omega \) as \(\varepsilon \rightarrow 0\).

Let

$$\begin{aligned} \begin{aligned} \hat{u}_\varepsilon (x)&=\sum _{i=1}^k(-1)^i \left( \frac{M_i^{\frac{2+\alpha }{2}}\varepsilon ^{\frac{(2+\alpha )(2i-1)}{2(N-2)}}}{M_i^{2+\alpha }\varepsilon ^{\frac{(2+\alpha )(2i-1)}{N-2}}+|x|^{2+\alpha }}\right) ^{\frac{N-2}{2+\alpha }}\\&=\sum _{i=1}^k(-1)^i\left( \frac{1}{M_i^{\frac{2+\alpha }{2}}\varepsilon ^{\frac{(2+\alpha )(2i-1)}{2(N-2)}} +M_i^{-\frac{2+\alpha }{2}}\varepsilon ^{-\frac{(2+\alpha )(2i-1)}{2(N-2)}}|x|^{2+\alpha }}\right) ^{\frac{N-2}{2+\alpha }}. \end{aligned} \end{aligned}$$

Hence,

$$\begin{aligned} u_\varepsilon (x)=C_{\alpha ,N}\hat{u}_\varepsilon (x)(1+o(1)). \end{aligned}$$
(4.4)

Set \(S_\varepsilon ^j=\{x\in \mathbb {R}^N: |x|=\varepsilon ^{\frac{2j-1}{N-2}}\}, j=1,2,\ldots ,k\), and choose a compact subset \(K\subset \Omega \) such that, for \(\varepsilon \) small enough, \(S_\varepsilon ^j\subset K\) for \(j=1,2,\ldots ,k\).

Then, for \(x\in S_\varepsilon ^j\), we have

$$\begin{aligned} \begin{aligned} \hat{u}_\varepsilon (x)&=\sum _{i=1}^k(-1)^i\left( \frac{1}{M_i^{\frac{2+\alpha }{2}}\varepsilon ^{\frac{(2+\alpha )(2i-1)}{2(N-2)}} +M_i^{-\frac{2+\alpha }{2}}\varepsilon ^{\frac{(2+\alpha )(4j-2i-1)}{2(N-2)}}}\right) ^{\frac{N-2}{2+\alpha }}\\&=\varepsilon ^{-\frac{2j-1}{2}}\sum _{i=1}^k(-1)^i\left( \frac{1}{M_i^{\frac{2+\alpha }{2}}\varepsilon ^{\frac{(2+\alpha )(i-j)}{(N-2)}} +M_i^{-\frac{2+\alpha }{2}}\varepsilon ^{\frac{(2+\alpha )(j-i)}{(N-2)}}}\right) ^{\frac{N-2}{2+\alpha }}\\&=(-1)^j\varepsilon ^{-\frac{2j-1}{2}}\left( \frac{1}{\big (M_j^{\frac{2+\alpha }{2}}+M_j^{-\frac{2+\alpha }{2}}\big )^{\frac{N-2}{2+\alpha }}}+o(1)\right) . \end{aligned} \end{aligned}$$

Thus, for \(\varepsilon >0\) small enough, \((-1)^j\hat{u}_\varepsilon >0\) on \(S_\varepsilon ^j\), \(j=1,2,\ldots ,k\), which implies that \((-1)^ju_\varepsilon >0\) on \(S_\varepsilon ^j\). Therefore, \(u_\varepsilon \) has at least k nodal domains \(\Omega _1,\ldots ,\Omega _k\) such that \(\Omega _i\) contains the sphere \(S_\varepsilon ^i\).

Next we show that, for \(\varepsilon \) small enough, \(u_\varepsilon \) has at most k nodal sets. Thanks to Proposition 2.2, Lemma 4.1, (1.7) and (1.10), we have

$$\begin{aligned} J_\varepsilon (PU_{\mu _i})\rightarrow \frac{(2+\alpha )}{2(N+\alpha )}\int _{\mathbb {R}^N}|x|^\alpha U_1^{p_\alpha +1},\quad i=1,2,\ldots ,k,~~\text {as}~\varepsilon \rightarrow 0 \end{aligned}$$
(4.5)

and

$$\begin{aligned} J_\varepsilon (u_\varepsilon )\rightarrow \frac{(2+\alpha )k}{2(N+\alpha )}\int _{\mathbb {R}^N}|x|^\alpha U_1^{p_\alpha +1},~~\text {as}~\varepsilon \rightarrow 0. \end{aligned}$$
(4.6)

Argue by contradiction, we can assume that there exists another nodal domain denoted by \(\Omega _{k+1}\). If \(\alpha >0\), we find that

$$\begin{aligned} \left( \int _{\Omega _{k+1}}|u_\varepsilon |^{\frac{2N}{N-2}}\right) ^{\frac{N-2}{N}}\le C\int _{\Omega _{k+1}}|x|^\alpha |u_\varepsilon |^{p_\alpha +1-\varepsilon }. \end{aligned}$$
(4.7)

Hence,

$$\begin{aligned} \left( \int _{\Omega _{k+1}}|u_\varepsilon |^{\frac{2N}{N-2}}\right) ^{\frac{N-2}{N}}\le C\Vert u_\varepsilon \Vert ^{\frac{2\alpha }{N-2}-\varepsilon }_{L^\infty (\Omega _{k+1})}\int _{\Omega _{k+1}} |u_\varepsilon |^{\frac{2N}{N-2}}. \end{aligned}$$

By (4.4), we see that \(\Vert u_\varepsilon \Vert _{L^\infty (\Omega _{k+1})}\le C\). Thus, \(\int _{\Omega _{k+1}}|u_\varepsilon |^{\frac{2N}{N-2}}\ge C>0,\) which implies \(J_\varepsilon (u_\varepsilon )>\frac{(2+\alpha )k}{2(N+\alpha )}\int _{\mathbb {R}^N}|x|^\alpha U_1^{p_\alpha +1}\). This is a contradiction with (4.6). If \(-2<\alpha <0\), by Hardy inequality, we obtain that \(\int _{\Omega }|x|^\alpha |u|^{p_\alpha +1}\le C\big (\int _{\Omega }|\nabla u|^2\big )^{\frac{p_\alpha +1}{2}}\). Similar to the case \(\alpha =0\) in [23], we still have that \(J_\varepsilon (u_\varepsilon )>\frac{(2+\alpha )k}{2(N+\alpha )}\int _{\mathbb {R}^N}|x|^\alpha U_1^{p_\alpha +1}\) and the proof of Theorem 1.1 is finished. \(\square \)