In this paper, we will first consider the following system
where $N\geq5$, $p=\frac{N}{N-2}, 2^{\ast}=\frac{2N}{N-2}$, $-\lambda_{1}(\mathbb{B})<\lambda_{1}, \lambda_{2}<0, \mu_{1}, \mu_{2}>0$ and $\beta\neq0$. Here $\mathbb{B}$ is $ [0, 1)\times X$ and $X\subset\mathbb{R}^{N-1}$ is a smooth compact domain, $\lambda_{1}(\mathbb{B})$ is the first eigenvalue of $-\Delta_\mathbb{B}$ with zero Dirichlet condition on $\partial\mathbb{B}$, $ \Delta_{\mathbb{B}}$ =$(x_{1}\partial_{x_{1}})^{2}+\partial^{2}_{x_{2}}+\cdots+\partial^{2}_ {x_{N}}$. We will look for the positive least energy solutions for (1.1) in the cone Sobolev space $\mathcal{H}_{2, 0}^{1, \frac{N}{2}}(\mathbb{B})$, which was introduced in [13]. In [2], Chen-Liu-Wei considered the following problem
and got a positive solution $\varphi $. Recently, the authors in [8] also studied the positive least energy solutions for $p$-Laplacian system. Our study is in fact motivated by the study of Chen-Zou (see [1]), and we investigate the semi-linear equations with critical cone Sobolev exponent terms.
We call a solution $(u, v)\in H$ nontrivial if $u\not\equiv0, v\not\equiv0$, where $H:=\mathcal{H}_{2, 0}^{1, \frac{N}{2}}(\mathbb{B})\times \mathcal{H}_{2, 0}^{1, \frac{N}{2}}(\mathbb{B})$. The weak solutions of (1.1) are the critical points of the functional $J :H\rightarrow\mathbb{R}$, which is given by
We say that a solution $(u, v)$ of (1.1) is a least energy solution if $(u, v)$ is nontrivial and $J(u, v)\leq J(\varphi, \psi)$ for any other nontrivial solution $(\varphi, \psi)$ of (1.1). If we define a " Nehari" manifold (see [1, 4-7, 9])
then any nontrivial solutions of (1.1) belong to $\mathcal{N}$, here $J'(\cdot, \cdot)$ is the Fr${\rm \acute{e}}$chet differentiation of $J$. We define the least energy of (1.3) as
If the equation
has a solution $(d_{0}, g_{0})$ with
then we prove the following theorem.
Theorem 1.1 Let $(d_{0}, g_{0})$ be a solution of (1.4) with $ d_{0}$ in (1.5) and $-\lambda_{1}(\mathbb{B})<\lambda_{1}=\lambda_{2}=\lambda<0$. Then for any $\beta>0$, $(\sqrt{d_{0}}\varphi, \sqrt{g_{0}}\varphi)$ is a positive solution of (1.1). Moreover, if $\beta\geq\frac{2}{N-2}\max\{\mu_{1}, \mu_{2}\}$, then we have $J(\sqrt{d_{0}}\varphi, \sqrt{g_{0}}\varphi)=A$, that is, $(\sqrt{d_{0}}\varphi, \sqrt{g_{0}}\varphi)$ is a positive least energy solution of (1.1).
In the second part of this paper, we consider the existence of the least energy solution of the following problem
where $\mathbb{R}_{+}^{N}=\mathbb{R}_{+}\times\mathbb{R}^{N-1}$ and $D_{2}^{1, \frac{N}{2}}(\mathbb{R}_{+}^{N})=: \{u\in L_{2^{*}}^{\frac{N}{2^{*}}}:|\nabla_{\mathbb{B}}u |\in L_{2}^{\frac{N}{2}}(\mathbb{R}_{+}^{N})\}$ with norm $\| u \|_{D_{2}^{1, \frac{N}{2}}}=(\displaystyle\int_{\mathbb{R}_{+}^{N}}| \nabla_{\mathbb{B}}u|^{2}\frac{dx_{1}}{x_{1}}dx^{'})^{\frac{1}{2}}$. Let $D:=D_{2}^{1, \frac{N}{2}}(\mathbb{R}_{+}^{N})\times D_{2}^{1, \frac{N}{2}}(\mathbb{R}_{+}^{N})$ and the energy functional $E$ for (1.6) is defined as
Analogously, we let
it is easy to see that any nontrivial solutions of (1.6) belong to $\mathcal{M}$. Then we get the following theorem.
Theorem 1.2 If $\beta<0$, then $B$ is not attained.
(2) If $\beta>0$, then there exists a positive least energy solution $(U, V)$ of (1.6) with $E(U, V)=B$, which is partly radially symmetric decreasing. Furthermore, we have
(2-1) Let $(d_{0}, g_{0})$ be as in Theorem 1.1. If $\beta\geq\frac{2}{N-2}\max\{\mu_{1}, \mu_{2}\}$, then
That is, $(\sqrt{d_{0}}U_{\varepsilon}, \sqrt{g_{0}}U_{\varepsilon})$ is a positive least energy solution of (1.6).
(2-2) There exists $0<\beta_{1}\leq\frac{2}{N-2}\max\{\mu_{1}, \mu_{2}\}$ such that for any $0<\beta<\beta_{1}$, we have a solution $(d(\beta), g(\beta))$ of (1.4) with
That is, $(\sqrt{d(\beta)}U_{\varepsilon}, \sqrt{g(\beta)}U_{\varepsilon})$ is a different positive solution of (1.6) with respect to $(U, V)$.
The terminology "partly radially symmetrization decreasing" in Theorem 1.2 will be explained in Section 3. Meanwhile, we will introduce "cone Schwartz symmetrization" in the same section.
The paper is organized as follows. In Section 2, we will give some preliminaries about cone Sobolev spaces and some auxiliary results. In Section 3, we will give the proofs of Theorems 1.1 and 1.2.
Here we first introduce the cone Sobolev spaces. Let $X$ be a closed, compact $C^{\infty}$ manifold of dimension $N-1$, and set $X^{\triangle}=(\overline{\mathbb{R}}_{+}\times X)/(\{0\}\times X)$ which is the local model interpreted as a cone with the base $X$. More details about the manifold with singularities can be found in [10].
Definition 2.1 For $(x_{1}, x^{\prime})\in \mathbb{R}_{+}\times\mathbb{R}^{N-1}$, we say that $u(x_{1}, x^{\prime})\in L_{p}(\mathbb{R}_{+}^{N}, \frac{dx_{1}}{x_{1}}dx^{\prime})$ if
The weighted $L_{p}$-spaces with weight data $\gamma\in \mathbb{R}$ is denoted by $ L_{p}^{\gamma}(\mathbb{R}_{+}^{N}, \frac{dx_{1}}{x_{1}}dx^{\prime})$, and then $x_{1}^{-\gamma}u(x_{1}, x^{\prime})\in L_{p}(\mathbb{R}_{+}^{N}, \frac{dx_{1}}{x_{1}}dx^{\prime})$ with the norm
Definition 2.2 For $m\in\mathbb{N}$, and $\gamma\in\mathbb{R}$, we define the spaces
for arbitrary $\alpha\in\mathbb{N}, \beta\in\mathbb{N}^{N-1}$, and $|\alpha|+|\beta|\leq m$. In other words, if $u(x_{1}, x^{\prime})\in \mathcal{H}_{p}^{m, \gamma}(\mathbb{R}_{+}^{N})$, then $(x_{1}\partial_{x_{1}})^{\alpha}\partial_{x^{\prime}}^{\beta}u\in L^{\gamma}_{p}(\mathbb{R}_{+}^{n}, \frac{dx_{1}}{x_{1}}dx^{\prime})$. It's easy to see that $\mathcal{H}_{p}^{m, \gamma}(\mathbb{R}_{+}^{N})$ is a Banach space with the norm
We will always denote $\omega(x_{1}, x^{\prime})\in C_{0}^{\infty}(\mathbb{B})$ as a real-valued cut-off function which equals $1$ near $\{0\}\times\partial\mathbb{B}$.
Definition 2.3 Let $\mathbb{B}$ be the stretched manifold to a manifold $B$ with conical singularities. Then $\mathcal{H}_{p}^{m, \gamma}(\mathbb{B})$ for $m\in\mathbb{N}, \gamma\in\mathbb{R}$ denotes the subspace of all $u\in W_{{\rm loc}}^{m, p}({\rm int}~\mathbb{B})$ such that
for any cut off function $\omega$, supported by a collar neighbourhood of $[0, 1)\times\partial\mathbb{B}$. Moreover, the subspace $\mathcal{H}_{p, 0}^{m, \gamma}(\mathbb{B})$ of $\mathcal{H}_{p}^{m, \gamma}(\mathbb{B})$ is defined as follows
where $W_{0}^{m, p}({\rm int}~\mathbb{B})$ denotes the closure of $C_{0}^{\infty}({\rm int}~\mathbb{B})$ in Sobolev spaces $W^{m, p}(\tilde{X})$ when $\tilde{X}$ is a closed compact $C^{\infty}$ manifold of dimension of $N$ that containing $\mathbb{B}$ as a submanifold with boundary. More details on the properties of the spaces $\mathcal{H}_{p, 0}^{m, \gamma}(\mathbb{B})$ and $\mathcal{H}_{p}^{m, \gamma}(\mathbb{B})$ can be found in [10].
Next, we will recall the cone Sobolev inequality and Poincaré inequality. For details we refer to [12].
Lemma 2.1 (Cone Sobolev inequality) Assume that $1<p<N, ~\frac{1}{p^{*}}=\frac{1}{p}-\frac{1}{N}$, and $\gamma\in \mathbb{R}$. The following estimate
holds for all $u\in C_{0}^{\infty}(\mathbb{B})$, where $\gamma^{*}=\gamma-1, c_{1}=\frac{(N-1)p}{N(N-p)}, c_{2}=\frac{|N-\frac{(\gamma-1)(N-1)p}{N-p}|^{\frac{1}{N}}}{N}$. Moreover, if $u\in\mathcal{H}_{p, 0}^{1, \gamma}(\mathbb{B})$, we have $\|u\|_{L_{p^{*}}^{\gamma^{*}}(\mathbb{B})}\leq c\|u\|_{\mathcal{H}_{p}^{1, \gamma}(\mathbb{B})}, $ where the constant $c=c_{1}+c_{2}$ and $c_{1}, c_{2}$ are given.
Lemma 2.2 (Poincaré inequality). Let $\mathbb{B}=[0, 1)\times X$ ba a bounded subset in $\mathbb{R}_{+}^{N}$, and $1<p<+\infty, \gamma\in \mathbb{R}$. If $u(x_{1}, x^{\prime})\in\mathcal{H}_{p, 0}^{1, \gamma}(\mathbb{B})$, then $\|u(x_{1}, x^{\prime})\|_{L_{p}^{\gamma}(\mathbb{B})}\leq c\|\nabla_{\mathbb{B}}u(x_{1}, x^{\prime})\|_{L_{p}^{\gamma}(\mathbb{B})}, $ where the positive constant $c$ depending only on $\mathbb{B}$ and $p$.
Lemma 2.3 For $2<p<2^{*}$, the embedding $\mathcal{H}_{2, 0}^{1, N/2}(\mathbb{B})\hookrightarrow\mathcal{H}_{p, 0}^{0, N/p}(\mathbb{B})$ is compact. Then we set
Lemma 2.4 Suppose that $\beta\geq(p-1)\max\{\mu_{1}, \mu_{2}\}$. Then the following system
has a unique solution $(d_{0}, g_{0})$.
Proof See [1, Lemmas 2.1, 2.2, 2.3, 2.4].
Now we consider the solution of (1.2), we will prove that this solution is also a least energy solution.
Lemma 2.5 Assume that $-\lambda_{1}(\mathbb{B})<\lambda<0$, and then (1.2) has a positive least energy solution $\varphi\in \mathcal{H}_{2, 0}^{1, \frac{N}{2}}(\mathbb{B})$ with energy
Proof Let $S_{\lambda}(u;\mathbb{B})= \frac{\parallel \nabla_{\mathbb{B}}u \parallel_{L_{2}^{\frac{N}{2}}}^{2}+\lambda\parallel u\parallel_{L_{2}^{\frac{N}{2}}}^{2}}{\parallel u\parallel_{L_{2^{*}}^{\frac{N}{2^{*}}}}^{2}}$ and $ S_{\lambda}(\mathbb{B})=\inf\limits_{u\in H_{2, 0}^{1, \frac{N}{2}}(\mathbb{B}), u\neq0}S_{\lambda}(u;\mathbb{B})$. Set $C_{0}=\frac{1}{N}[S_{\lambda}(\mathbb{B})]^{\frac{N}{2}}$ and the functional
From the result in [2], we know that (1.2) has a positive solution with energy $C_{0}$. Furthermore, we will show that $C_{0}$ is the least energy of (1.2). We set
If $u$ is the solution of (1.2), then $u\in\overline{\mathcal{N}}$ and $f_{\lambda}(u)=\frac{1}{N}\| u\|_{L_{2^{*}}^{\frac{N}{2^{*}}}}^{2^{*}}=\frac{1}{N}[S_{\lambda}(u;\mathbb{B})]^{\frac{N}{2}}$. If we denote $\inf\limits_{u\in\overline{\mathcal{N}}, u\neq0}f_{\lambda}(u)$ as the least energy of (1.2), then
Therefore $C_{0}=A_{1}=\inf\limits_{u\in\overline{\mathcal{N}}, u\neq0}f_{\lambda}(u)$. Let $\varphi$ be a positive critical point of $f_{\lambda}(u)$ with a critical value $A_{1}$. Then it is easy to get $A_{1}:=\frac{1}{N}\displaystyle\int_{\mathbb{B}}(|\nabla_{\mathbb{B}}\varphi|^{2}+\lambda\varphi^{2}) \frac{dx_{1}}{x_{1}}dx^{'}$.
In this section, we will prove Theorem 1.1 and Theorem 1.2. In particular, we will separate the proof of Theorem 1.2 into several steps.
Proof of Theorem 1.1 For $-\lambda_{1}(\mathbb{B})<\lambda_{1}=\lambda_{2}=\lambda<0$, we can easily get that $A=\inf\limits_{(u, v)\in\mathcal{N}}J(u, v)>0$. $\beta>0$, so (1.3) has a solution $(d_{0}, g_{0})$. By Lemma 2.5, we obtain that $\displaystyle\int_{\mathbb{B}}(|\nabla_{\mathbb{B}}\varphi|^{2}+\lambda\varphi^{2})\frac{dx_{1}}{x_{1}}dx^{'}=\displaystyle\int_{\mathbb{B}} \varphi^{2^{*}}\frac{dx_{1}}{x_{1}}dx^{'}$. For a direct computing, we can get that $(\sqrt{d_{0}}\varphi, \sqrt{g_{0}}\varphi)$ is a positive solution of (1.1). Moreover, we have
Now if $\beta\geq(p-1)\max\{\mu_{1}, \mu_{2}\}$, then we have $A= J(\sqrt{d_{0}}\varphi, \sqrt{g_{0}}\varphi)$. In fact, we can take a minimizing sequence$\{(u_{n}, v_{n})\}_{n\in \mathbb{N}}\subset\mathcal{N}$ for $A$ such that $J(u_{n}, v_{n})\rightarrow A$. Then we get
and
where $c_{n}=(\displaystyle\int_{\mathbb{B}}| u_{n}|^{2p}\frac{dx_{1}}{x_{1}}dx^{'})^{\frac{1}{p}}$ and $k_{n}=(\displaystyle\int_{\mathbb{B}}| v_{n}|^{2p}\frac{dx_{1}}{x_{1}}dx^{'})^{\frac{1}{p}}$. Note that
and then from (3.1), we have
Therefore, the sequences $\{c_{n}\}_{n\in \mathbb{N} }, \{k_{n}\}_{n\in \mathbb{N}}$ are uniformly bounded. Passing to a subsequence, we assume that $c_{n}\rightarrow c$ and $k_{n}\rightarrow k$ as $n\rightarrow \infty$ for some $c\geq0, k\geq0$. By (3.2) and (3.3), we have $\mu_{1}c^{p}+2\beta c^{\frac{p}{2}}k^{\frac{p}{2}}+\mu_{2}k^{p}\geq NA>0$. That means $c$ and $k$ are not necessary to be all vanished. From (3.4)-(3.6), we get
Applying Lemma 2.4, we have
here we get $c_{n}\rightarrow d_{0}(NA_{1})^{1-\frac{2}{N}}$ and $k_{n}\rightarrow g_{0}(NA_{1})^{1-\frac{2}{N}}$ as $n\rightarrow \infty$, and moreover,
That is, $A\geq(d_{0}+g_{0})A_{1}=J(\sqrt{d_{0}}\varphi, \sqrt{g_{0}}\varphi)$, and so $A=J(\sqrt{d_{0}}\varphi, \sqrt{g_{0}}\varphi)=(d_{0}+g_{0})A_{1}$. This tells us that $(\sqrt{d_{0}}\varphi, \sqrt{g_{0}}\varphi)$ is a positive least energy solution of (1.1).
Next we start to prove Theorem 1.2.
Lemma 3.1 For $-\infty<\beta<0$, if $B$ is attained by a couple $(u, v)\in\mathcal{M}$, then this couple is a critical point of $E(u, v)$ in (1.7). The proof is analogous to that in [1, Lemma 2.5]. So we omit it here.
By Lemma 2.1, let $S$ be the sharp constant of $D_{2}^{1, \frac{N}{2}}(\mathbb{R}_{+}^{N})\hookrightarrow L_{2^{*}}^{\frac{N}{2^{*}}}(\mathbb{R}_{+}^{N}), $
For $ \varepsilon>0$, let
Then $U_{\varepsilon}$ satisfies $-\Delta_\mathbb{B}u=| u|^{2^{*}-2}u$ in $\mathbb{R}_{+}^{N}$ (see [2, 5]). Moreover,
Now we give the proof of first part of Theorem 1.2.
Proof of (1) in Theorem 1.2 Let $\varphi_{\mu_{i}}:=\mu_{i}^{\frac{2-N}{4}}U_{1}$ with $U_{1}$ being as in (3.9). Then $\varphi_{\mu_{i}}$ satisfies the equation $-\Delta_\mathbb{B}u=\mu_{i}| u|^{2^{*}-2}u$ in $\mathbb{R}_{+}^{N}$. We set $e_{2}=(0, 1, 0, \cdots, 0)\in\mathbb{R}_{+}^{N}$ and $(u_{r}(x), v_{r}(x))=(\varphi_{\mu_{1}(x)}, \varphi_{\mu_{2}(x+re_{2})}).$ Then $v_{r}\rightharpoonup0$ weakly in $D_{2}^{1, \frac{N}{2}}(\mathbb{R}_{+}^{N})$ and $v^{p}_{r}\rightharpoonup0$ weakly in $L_{2}^{\frac{N}{2}}(\mathbb{R}_{+}^{N})$ as $r\rightarrow\infty$. That is,
To complete this proof, we claim that: for $r>0$ sufficiently large and $\beta<0$, there exists $(t_{r}u_{r}, s_{r}v_{r})\in\mathcal{M}$ with $t_{r}>1, s_{r}>1$.
In fact, note that $u_{r}$ and $v_{r}$ satisfy the equation $-\Delta_\mathbb{B}u=\mu_{i}| u|^{2^{*}-2}u$. If $(tu_{r}, sv_{r})\in\mathcal{M}$, then we have
Since $v_{r}(x)\rightarrow0~~ (r\rightarrow\infty)$, there exists $\delta_{r}>0$ for $r$ sufficiently large such that $v_{r}(x)\leq\delta_{r}$ and $\lim\limits_{r\rightarrow\infty}\delta_{r}=0$. By cone Sobolev inequality, we obtain that for some $C>0$,
For simplicity, we denote
So $D_{1}D_{2}-F_{r}^{2}>0$. Recall that $(tu_{r}, sv_{r})\in\mathcal{M}$, and thus we get
From the first equality of (3.11), we obtain $s^{p}=(t^{2-p}-t^{p})\frac{D_{1}}{F_{r}}>0$, and therefore $t>1$. Similarly, we have $s>1$. Note that (3.11) is equivalent to $w(t)=0$, where
For $1<p<2$, we get $w(1)=-F_{r}>0$, and $\lim\limits_{t\rightarrow\infty}w(t)<0$. So there exists $t_{r}>1$ such that $w(t)=0$.
Note that $(t_{r}u_{r}, s_{r}v_{r})\in\mathcal{M}$, and then we have
Up to a subsequence, if $t_{r}\rightarrow\infty$ as $r\rightarrow\infty$, then by the fact
we also get $t\rightarrow\infty ~~(r\rightarrow\infty)$. As $2-p<p$, for $r$ large enough, we have
Therefore, we obtain
This means that $0<\frac{1}{4}D_{1}D_{2}\leq F^{2}_{r}\rightarrow0, ~~{\rm as}~~r\rightarrow\infty, $ which is a contradiction. Hence $t_{r}$ and $s_{r}$ are uniformly bounded. By (3.12) and $F_{r}\rightarrow0 ~~(r\rightarrow\infty$), we have $\lim\limits_{r\rightarrow\infty}t_{r}=\lim\limits_{r\rightarrow\infty}s_{r}=1.$
For $(t_{r}u_{r}, s_{r}v_{r})\in\mathcal{M}$, from (3.10) we have
Let $r\rightarrow\infty$, we get that $B\leq\frac{1}{N}(\mu_{1}^{-\frac{N-2}{2}}+\mu^{-\frac{N-2}{2}}_{2})S^{\frac{N}{2}}$.
On the other hand, for any $(u, v)\in\mathcal{M}$, by the fact $\beta<0$ and (3.8), we get that
Therefore $\displaystyle\int_{\mathbb{R}_{+}^{N}}|\nabla_{\mathbb{B}} u|^{2}\frac{dx_{1}}{x_{1}}dx^{'}\geq \mu_{1}^{-\frac{N-2}{2}}S^{\frac{N}{2}}$, and similarly, $\displaystyle\int_{\mathbb{R}_{+}^{N}}|\nabla_{\mathbb{B}} v|^{2}\frac{dx_{1}}{x_{1}}dx^{'}\geq \mu_{2}^{-\frac{N-2}{2}}S^{\frac{N}{2}}$. Note that $B=\inf\limits_{(u, \nu)\in\mathcal{M}}\{\frac{1}{N}\displaystyle\int_{\mathbb{R}_{+}^{N}}(| \nabla_{\mathbb{B}}u|^{2}+| \nabla_{\mathbb{B}}v|^{2})\frac{dx_{1}}{x_{1}}dx^{'}\}$, and then we obtain that $B\geq\frac{1}{N}(\mu_{1}^{-\frac{N-2}{2}}+\mu^{-\frac{N-2}{2}}_{2})S^{\frac{N}{2}}$. Hence $B=\frac{1}{N}(\mu_{1}^{-\frac{N-2}{2}}+\mu^{-\frac{N-2}{2}}_{2})S^{\frac{N}{2}}.$
Now if $B$ is attained by some $(u, v)\in\mathcal{M}$, then $(|u|, |v|)\in\mathcal{M}$ and $E(|u|, |v|)=B$. From Lemma 3.1, we know that $(| u|, |v|)$ is a nontrivial solution of (1.6). By the maximum principle, we may assume that $u>0, v>0$, and so $\displaystyle\int_{\mathbb{R}_{+}^{N}} u^{p}v^{p}\frac{dx_{1}}{x_{1}}dx^{'}>0$. Moreover, we get
It is easy to see that
that is a contradiction. We complete the proof.
Now we begin to prove (2-1) in Theorem 1.2.
Proof of (2-1) in Theorem 1.2 For $\beta>0$, $(\sqrt{d_{0}}U_{\varepsilon}, \sqrt{g_{0}}U_{\varepsilon})$ is a nontrivial solution of (1.6) and $B\leq E(\sqrt{d_{0}}U_{\varepsilon}, \sqrt{g_{0}}U_{\varepsilon})=\frac{1}{N}(d_{0}+g_{0})S^{\frac{N}{2}}$.
We let $\beta\geq(p-1)\max\{\mu_{1}, \mu_{2}\}$ and $\{(u_{n}, v_{n})\}_{n\in \mathbb{N}}\subset\mathcal{M}$ be a minimizing sequence for $B$, that is, $E(u_{n}, v_{n})\rightarrow B$. Define $c_{n}=(\displaystyle\int_{\mathbb{R}_{+}^{N}}| u_{n}|^{2p}\frac{dx_{1}}{x_{1}}dx^{'})^{\frac{1}{p}}, k_{n}=(\displaystyle\int_{\mathbb{R}_{+}^{N}}| v_{n}|^{2p}\frac{dx_{1}}{x_{1}}dx^{'})^{\frac{1}{p}}$, and we have
which imply
Similarly as in the proof of Theorem 1.1, we have $c_{n}\rightarrow d_{0}S^{\frac{N}{2}-1}, d_{n}\rightarrow g_{0}S^{\frac{N}{2}-1} ~~(n\rightarrow\infty).$ Moreover, we obtain
Since $B\leq \frac{1}{N}(d_{0}+g_{0})S^{\frac{N}{2}}$, we obtain that
Therefore $(\sqrt{d_{0}}U_{\varepsilon}, \sqrt{g_{0}}U_{\varepsilon})$ is a positive least energy solution of (1.6).
Next we continue the proof of (2-2) in Theorem 1.2. For this purpose we need to show that (1.6) has a positive least energy solution for any $0<\beta<(p-1)\max\{\mu_{1}, \mu_{2}\}$. Therefore, we assume $\beta>0$, and define $B^{'}:=\inf\limits_{(u, v)\in\mathcal{M}^{'}}E(u, v), $ where
It is easy to see that $\mathcal{M}\subset \mathcal{M}^{'}$, and so $B^{'}\leq B$. By cone Sobolev inequality, we have $B^{'}>0$. We set $\Omega_{R}(1, 0):=\{(x_{1}, x^{'})\in \mathbb{R}_{+}^{N};(\ln x_{1})^{2}+| x^{'}|^{2}<R^{2}\}, H(x_{0}, R):=\mathcal{H}_{2, 0}^{1, \frac{N}{2}}(\Omega_{R}(x_{0}))\times \mathcal{H}_{2, 0}^{1, \frac{N}{2}}(\Omega_{R}(x_{0}))$ for $x_{0}=(1, 0, \cdots, 0)\in\mathbb{R}_{+}^{N}$. Consider the system
and define $B^{'}(R):=\inf\limits_{(u, v)\in\mathcal{M}^{'}(R)}E(u, v), $ where
Lemma 3.2 For all $R>0$, we have $B^{'}(R)\equiv B^{'}$.
Proof Let $R_{1}>R_{2}$, since $\mathcal{M}^{'}(R_{2})\subset\mathcal{M}^{'}(R_{1})$, we get $B(R_{1})\leq B^{'}(R_{2})$. For any $(u, v)\in\mathcal{M}^{'}(R_{1})$, we define
It is easy to see that $(u_{1}, v_{1})\in\mathcal{M}^{'}(R_{2})$, and so
That is, $B^{'}(R_{2})\leq B^{'}(R_{1})$. Hence we have $B^{'}(R_{1})= B^{'}(R_{2})$.
Let $\{(u_{n}, v_{n})\}_{n\in\mathbb{N}}\subset\mathcal{M}^{'} $ be a minimizing sequence of $B^{'}$. Moreover, we may assume that $u_{n}, v_{n}\in \mathcal{H}_{2, 0}^{1, \frac{N}{2}}(\Omega_{R_{n}}(x_{0}))$ for some $R_{n}>0$. Then $(u_{n}, v_{n})\in\mathcal{M}^{'}(R_{n}) $ and
Note that $B^{'}\leq B^{'}(R)$ and consequently we have $B^{'}(R)\equiv B^{'}$ for any $ R>0$.
Let $0\leq\varepsilon<p-1$. Consider
and define $B_{\varepsilon}=\inf\limits_{(u, v)\in\mathcal{M}^{'}}E_{\varepsilon}(u, v)$, where
Set $\mathcal{M}^{'}_{\varepsilon}:=\{(u, v)\in H(x_{0}, R)\setminus{(0, 0)}, H_{\varepsilon}(u, v):=E^{'}_{\varepsilon}(u, v)(u, v)=0\}.$
Lemma 3.3 For $0<\varepsilon<p-1$, there holds
The proof is analogous to that in [1, Lemma 2.7]. So we omit it here.
Similarly as in Lemma 3.3, we have
where $\varphi_{\mu_{i}}$ is the same as in the proof of (1) in Theorem 1.2.
Now we introduce the "Cone Schwartz symmetrization". Assume that $\Omega$ is a bounded domain of $\mathbb{R}_{+}^{N}$ and $u$ is a real measurable function defined on $\Omega$. We define the distribution function of $u$ as follows $u_{\sharp}(t)={\rm meas} \{ x\in\Omega:|u(x)|>t\}~~{\rm for}~~t\in\mathbb{R}, $ where "meas" denotes the corresponding measure in cone Sobolev space. Then we can define the decreasing rearrangement of $u$ in the form $\widetilde{u}(s)=\inf\{t\in\mathbb{R}:u_{\sharp}(t)\leq s\}~{\rm for}~s\in[0, | \Omega|].$ We call $u^{*}(x)$ the cone Schwarz symmetrization of $u$ if $u^{*}(x)=\widetilde{u}(o_{n}| x|_{\mathbb{B}}^{N})~{\rm for}~x\in\widetilde{\Omega}, $ where $\widetilde{\Omega}$ is the sphere centred at $x_{0}$ with the same measure of $\Omega$, and $| x-z|_{\mathbb{B}}=(| \ln\frac{x_{1}}{z_{1}}|^{2}+| x^{'}-z^{'}|^{2})^{\frac{1}{2}}$ for $x=(x_{1}, x^{'})$, $z=(z_{1}, z^{'})$, here $o_{n}$ is the measure of the unit ball in $\mathbb{R}_{+}^{N}$. Since $\widetilde{u}$ is decreasing, $u^{*}$ is partly radially symmetric decreasing in relation to $ | x|_{\mathbb{B}}$.
Lemma 3.4 For any $0<\varepsilon<p-1$, (3.14) has a classical least energy solution $(u_{\varepsilon}, v_{\varepsilon})$, and $u_{\varepsilon}, v_{\varepsilon}$ are both partly radially symmetric decreasing.
Proof Fix any $0<\varepsilon<p-1$, and then it is easy to see that $B_{\varepsilon}>0$. Let $(u, v)\in\mathcal{M}^{'}_{\varepsilon}$ with $u\geq0, v\geq0$, and $(u^{*}, v^{*})$ be its cone Schwartz symmetrization. Then we have
Similarly as in Lemma 3.3, there exists $0 < t^{*}\leq1$ such that $(t^{*}u^{*}, t^{*}v^{*})\in\mathcal{M}^{'}_{\varepsilon}$, and then we get
We take a minimizing sequence $\{(u_{n}, v_{n})\}_{n\in \mathbb{N}}\subset \mathcal{M}^{'}_{\varepsilon}$ with $u_{n}\geq0, v_{n}\geq0$ such that $E_{\varepsilon}(u_{n}, v_{n})\rightarrow B_{\varepsilon}$. Let $(u_{n}^{*}, v_{n}^{*})$ be its "cone Schwartz symmetrization". Then there exists $0 < t^{*}_{n}\leq1$ such that $(t^{*}_{n}u^{*}_{n}, t^{*}_{n}v^{*}_{n})\in\mathcal{M}^{'}_{\varepsilon}$. By (3.15), we get
Therefore, we obtain $t^{*}_{n}\rightarrow 1, E_{\varepsilon}(u_{n}^{*}, v_{n}^{*})\rightarrow B_{\varepsilon}$, as $n\rightarrow\infty$, and $u_{n}^{*}, v_{n}^{*}$ are bounded in $\mathcal{H}_{2, 0}^{1, \frac{N}{2}}(\Omega_{1}(x_{0}))$. Passing to a subsequence, we may assume that $u_{n}^{*}\rightharpoonup u_{\varepsilon}, v_{n}^{*}\rightharpoonup v_{\varepsilon}$ weakly in $\mathcal{H}_{2, 0}^{1, \frac{N}{2}}(\Omega_{1}(x_{0}))$. By the compactness of the embedding $\mathcal{H}_{2, 0}^{1, \frac{N}{2}}(\Omega_{1}(x_{0}))\hookrightarrow L^{\frac{N}{2p-2\varepsilon}}_{2p-2\varepsilon}(\Omega_{1}(x_{0})$ and $\mathcal{H}_{2, 0}^{1, \frac{N}{2}}(\Omega_{1}(x_{0}))\hookrightarrow L^{\frac{N}{p-\varepsilon}}_{p-\varepsilon}(\Omega_{1}(x_{0})$, we have
which means $(u_{\varepsilon}, v_{\varepsilon})\neq(0, 0)$. Moreover, $u_{\varepsilon}\geq0, v_{\varepsilon}\geq0$ are partly radially symmetric.
Meanwhile, since
we get
Therefore, there exists $0 < t_{\varepsilon}\leq1$ such that $(t_{\varepsilon}u_{\varepsilon}, t_{\varepsilon}v_{\varepsilon})\in \mathcal{M}^{'}_{\varepsilon}$, and then
That is $t_{\varepsilon}=1$ and $(u_{\varepsilon}, v_{\varepsilon})\in \mathcal{M}^{'}_{\varepsilon}$ with $E_{\varepsilon}(u_{\varepsilon}, v_{\varepsilon})=B_{\varepsilon}$. Therefore, $u_{n}^{*}\rightarrow u_{\varepsilon}, v_{n}^{*}\rightarrow v_{\varepsilon}$ strongly in $\mathcal{H}_{2, 0}^{1, \frac{N}{2}}(\Omega_{1}(x_{0}))$ as $n\rightarrow\infty$.
By Lagrange multiplier theorem, we get that there exists a Lagrange multiplier $\tau\in\mathbb{R}$ such that $E^{'}_{\varepsilon}(u_{\varepsilon}, v_{\varepsilon})-\tau H^{'}_{\varepsilon}(u_{\varepsilon}, v_{\varepsilon})=0.$ Note that $E^{'}_{\varepsilon}(u_{\varepsilon}, v_{\varepsilon})(u_{\varepsilon}, v_{\varepsilon}) =H_{\varepsilon}(u_{\varepsilon}, v_{\varepsilon})=0$ and
We get that $\tau=0$ and $E^{'}_{\varepsilon}(u_{\varepsilon}, \nu_{\varepsilon})=0$. By Lemma 3.3, we see that $u_{\varepsilon}\not\equiv0, v_{\varepsilon}\not\equiv0$. This means that $(u_{\varepsilon}, v_{\varepsilon})$ is a least energy solution of (3.14). By regularity theory and the maximum principle, we see that $u_{\varepsilon}>0, v_{\varepsilon}>0$ in $\Omega_{1}(x_{0}), u_{\varepsilon}, v_{\varepsilon}\in C^{2}(\Omega_{1}(x_{0}))$. This completes the proof.
Completion of the Proof of (2-2) in Theorem 1.2 For any $(u, v)\in\mathcal{M}^{'}(1)$, it is easy to see that there exists $t_{\varepsilon}>0$ such that $(t_{\varepsilon}u, t_{\varepsilon}v)\in\mathcal{M}^{'}_{\varepsilon}$ with $t_{\varepsilon}\rightarrow1 ~~(\varepsilon\rightarrow0)$, then
By Lemma 3.2, we have
By Lemma 3.4, we know that there exists a positive least energy solution $(u_{\varepsilon}, v_{\varepsilon})$ of (3.14), which is partly radically symmetric decreasing. Recall that $E^{'}_{\varepsilon}(u_{\varepsilon}, v_{\varepsilon})(u_{\varepsilon}, v_{\varepsilon})=0$. By cone Sobolev inequality, we have
where $ W_{0}$ is a positive constant independent of $\varepsilon$. Then $u_{\varepsilon}, \nu_{\varepsilon}$ are uniformly bounded in $\mathcal{H}_{2, 0}^{1, \frac{N}{2}}(\Omega_{1}(x_{0}))$. Passing to a subsequence, we may assume that $u_{\varepsilon}\rightharpoonup u_{0}, ~v_{\varepsilon}\rightharpoonup v_{0}$ weakly in $\mathcal{H}_{2, 0}^{1, \frac{N}{2}}(\Omega_{1}(x_{0}))$ as $\varepsilon\rightarrow0$. Then $(u_{0}, v_{0})$ is a solution of the following problem
Note that $u_{\varepsilon}(x_{0})=\max\limits_{\Omega_{1}(x_{0})}u_{\varepsilon}(x), v_{\varepsilon}(x_{0})=\max\limits_{\Omega_{1}(x_{0})}v_{\varepsilon}(x)$ and define $K_{\varepsilon}=\max\{u_{\varepsilon}(x_{0}), v_{\varepsilon}(x_{0})\}$. We claim that $K_{\varepsilon}\rightarrow\infty$ as $\varepsilon\rightarrow0$. Suppose the contrary. If $K_{\varepsilon}$ is uniformly bounded, then by the dominated convergent theorem, we have that
Note that $E^{'}_{\varepsilon}(u_{\varepsilon}, v_{\varepsilon})=E^{'}(u_{0}, v_{0})=0$. It is standard to show that $u_{\varepsilon}^{*}\rightarrow u_{0}, v_{\varepsilon}^{*}\rightarrow v_{0}$ strongly in $\mathcal{H}_{2, 0}^{1, \frac{N}{2}}(\Omega_{1}(x_{0}))$ as $\varepsilon\rightarrow0$. By (3.17), we get that $(u_{0}, v_{0})\neq(0, 0)$. Moreover, $u_{0}\geq0, v_{0}\geq0$. By the strong maximum principle, $u_{0}>0, v_{0}>0$ in $\Omega_{1}(x_{0})$. Combining this with Pohozaev identity (see [11]), we get
which is a contradiction, here $\upsilon$ denotes the outward unit normal vector on $\partial\Omega_{1}(x_{0})$. So $K_{\varepsilon}\rightarrow+\infty$ as $\varepsilon\rightarrow0$. Define
Then we have
and $U_{\varepsilon}, V_{\varepsilon}$ satisfy
Since
we get that $\{(U_{\varepsilon}, V_{\varepsilon})\}$ is bounded in $D_{2}^{1, \frac{N}{2}}(\mathbb{R}_{+}^{N})\times D_{2}^{1, \frac{N}{2}}(\mathbb{R}_{+}^{N})=D$. By elliptic estimates, up to a subsequence, we have $(U_{\varepsilon}, V_{\varepsilon})\rightarrow(U, V)\in D$ uniformly in every compact subset of $\mathbb{R}_{+}^{N}$ as $\varepsilon\rightarrow0$, and $(U, V)$ satisfies (1.6), that is $E^{'}(U, V)=0$. Moreover, $U, V\geq0$ are partly radially symmetric decreasing. Note that (3.18) we get $(U, V)\neq(0, 0)$, and so $(U, V)\in\mathcal{M}^{'}$. Then we deduce from (3.16) that
So $E(U, V)=B^{'}$. Note that $B^{'} < \min\{ \inf\limits_{(u, 0)\in \mathcal{M}^{'}}E(u, 0), \inf\limits_{(0, v)\in \mathcal{M}^{'}}E(0, v)\}$ and we have $U\not\equiv0, V\not\equiv0$. By the strong maximum principle, $U>0$, $V>0$ are partly radially symmetric decreasing. We also have $(U, V)\in\mathcal{M}$, and so $E(U, V)\geq B\geq B^{'}$, that is, $E(U, V)=B=B^{'}.$ Moreover $(U, V)$ is positive least energy solution of (1.6), which is partly radially symmetric decreasing.
Finally, with the help of (2.1) and [1, (2-2) in Theorem 1.6], we get that there exists $d(\beta)$ and $g(\beta)$ on $(-\beta_{2}, \beta_{2})$ for some $\beta_{2}>0$, and $l_{i}(d(\beta), g(\beta))\equiv0$ for $i=1, 2$. This implies that $(\sqrt{d(\beta)}U_{\varepsilon}, \sqrt{g(\beta)}U_{\varepsilon})$ is a positive solution of (1.6). Therefore we have
that is, there exists $0<-\beta_{1} \leq-\beta_{2}$ such that
Recall that
and we have
that is, $(\sqrt{d(\beta)}U_{\varepsilon}, \sqrt{g(\beta)}U_{\varepsilon})$ is a different positive solution of (1.6) with respect to $(U, V)$. We complete the proof of (2-2) in Theorem 1.2.