Известия Российской академии наук. Серия математическая
RUS  ENG    ЖУРНАЛЫ   ПЕРСОНАЛИИ   ОРГАНИЗАЦИИ   КОНФЕРЕНЦИИ   СЕМИНАРЫ   ВИДЕОТЕКА   ПАКЕТ AMSBIB  
Общая информация
Последний выпуск
Скоро в журнале
Архив
Импакт-фактор
Правила для авторов
Загрузить рукопись

Поиск публикаций
Поиск ссылок

RSS
Последний выпуск
Текущие выпуски
Архивные выпуски
Что такое RSS



Изв. РАН. Сер. матем.:
Год:
Том:
Выпуск:
Страница:
Найти






Персональный вход:
Логин:
Пароль:
Запомнить пароль
Войти
Забыли пароль?
Регистрация


Известия Российской академии наук. Серия математическая, 2023, том 87, выпуск 1, страницы 33–48
DOI: https://doi.org/10.4213/im9244
(Mi im9244)
 

Multiple positive solutions for a Schrödinger–Poisson system with critical and supercritical growths

J. Leia, H. Suob

a School of Mathematics and Statistics, Guizhou University, Guiyang, China
b School of Data Science and Information Engineering, Guizhou Minzu University, Guiyang, China
Список литературы:
Аннотация: In this paper, we are concerned with the following Schrödinger–Poisson system
$$ \begin{cases} -\Delta u+u+\lambda\phi u= Q(x)|u|^{4}u+\mu \dfrac{|x|^\beta}{1+|x|^\beta}|u|^{q-2}u&\text{in }\mathbb{R}^3, \\ -\Delta \phi=u^{2} &\text{in }\mathbb{R}^3, \end{cases} $$
where $0< \beta<3$, $6<q<6+2\beta$, $Q(x)$ is a positive continuous function on $\mathbb{R}^3$, $\lambda,\mu>0$ are real parameters. By the variational method and the Nehari method, we obtain that the system has $k$ positive solutions.
Bibliography: 31 titles.
Ключевые слова: Schrödinger–Poisson system, critical exponent, supercritical growth.
Финансовая поддержка Номер гранта
National Natural Science Foundation of China 11661021
11861021
Supported by National Natural Science Foundation of China (grants № 11661021, № 11861021).
Поступило в редакцию: 19.07.2021
Исправленный вариант: 14.10.2021
Англоязычная версия:
Izvestiya: Mathematics, 2023, Volume 87, Issue 1, Pages 29–44
DOI: https://doi.org/10.4213/im9244e
Реферативные базы данных:
Тип публикации: Статья
УДК: 517.958
MSC: 35B33, 35J50, 35J20
Язык публикации: английский

§ 1. Introduction and main result

We consider the solvability of the Schrödinger–Poisson system

$$ \begin{equation} \begin{cases} -\Delta u+u+\lambda\phi u= Q(x)|u|^{4}u+\mu h(x)|u|^{q-2}u &\text{in }\mathbb{R}^3, \\ -\Delta \phi=u^{2} &\text{in }\mathbb{R}^3, \end{cases} \end{equation} \tag{1.1} $$
where $h(x)=|x|^\beta/(1+|x|^\beta)$, $0< \beta<3$, $6<q<6+2\beta$, $\lambda,\mu> 0$ are real parameters. We assume that the coefficient function $Q(x)\in C(\mathbb{R}^3)$, $Q(x)> 0$ a. e. $x\in \mathbb{R}^3$ and satisfies the following assumption:

$\mathrm{(Q_1)}$ There exist points $a_1,a_2,\dots,a_k\in \mathbb{R}^3$ such that $Q(a_i)$ are strict local maxima satisfying

$$ \begin{equation*} Q(a_i)=Q_M=\max_{x\in\mathbb{R}^3}Q(x) \end{equation*} \notag $$
and
$$ \begin{equation*} Q(x)=Q(a_i)+O(|x-a_i|^2)\quad\text{as}\quad x\to a_i,\qquad 1\leqslant i \leqslant k. \end{equation*} \notag $$
The Schrödinger–Poisson system
$$ \begin{equation} \begin{cases} -\dfrac{\hbar^2}{2m}\, \Delta u+e\omega u+\phi u=f(x,u) &\text{in }\mathbb{R}^3, \\ -\Delta\phi=4\pi eu^2 &\text{in }\mathbb{R}^3, \end{cases} \end{equation} \tag{1.2} $$
was first introduced in [1] describing the behavior of a quantum particle in three- dimensional space interacting with the electromagnetic field. Here $m$ and $e$ are the mass and the charge of the particle, respectively, while $\omega$ denotes the phase, $\hbar$ is Planck’s constant. The unknowns of the system are the field $u$ associated with the particles and the electric potential $\phi$. The presence of the nonlinear term $f(x,u)$ simulates the interaction between many particles or external nonlinear perturbations. The class of nonlinear terms $f(x,u)=\lambda |u|^{p-2}u$, $\lambda> 0$, $p>2$, contains the Slater correction term $C_\mathrm{s}|u|^{2/3}u$, where $C_\mathrm{s}$ is the Slater constant depending on the particles. It is important in various physical frameworks: gravitation, plasma physics, semiconductor theory, quantum chemistry (see [2], [3]). The class of nonlinear terms $f(x,u)=a|u|^{p-2}u+b|u|^{q-2}u$ with $a,b\in \mathbb{R}$ and $2<p<q< \infty$ arises in various regions of mathematical physics. For example, when $a> 0$, $b< 0$, $p=4$ and $q=6$, it models the evolution of a monochromatic wave complex envelope in a medium with weakly saturating nonlinearity. This Schrödinger equation with double power nonlinearities appears in boson gas interaction, nonlinear optics, nuclear hydrodynamics and so on (see [4], [5]). For more physical background of this system, we refer to [1], [6].

In recent years, many papers have been devoted to studying the existence and multiplicity of positive solutions of the system (1.2) under various assumptions on the nonlinear term. We refer the reader to [7]–[11]. The existence of nontrivial solutions for the system (1.2) was obtained in [12]–[14]. The existence of ground state solutions for the system (1.2) has been well studied by various authors, see [15]–[20], [3]. It is worth pointing out that the system (1.2) with critical growth has been studied extensively, for example, in [8], [10], [16], [18]–[22]. In addition, many researchers proved by the variational method that the elliptic equations with critical exponent have at least $k$ positive solutions. We refer the reader to [23]–[28] and references therein.

Specially, Huang [26] studied the following nonhomogeneous equations involving the $p$-Laplacian:

$$ \begin{equation} \begin{cases} -\Delta_p u=Q(x)|u|^{p^*-2}u+\theta h(x),\quad x\in{\mathbb{R}^N}, \\ u\in D^{1,p}(\mathbb{R}^N), \end{cases} \end{equation} \tag{1.3} $$
where $1< p< N$, $p^*=Np/(N-p)$ is the critical Sobolev exponent, $h(x)> 0$ a. e. in $\mathbb{R}^N$, $\theta> 0$ is a real parameter. $Q(x)$ satisfies appropriate assumptions. By the variational method, the author obtained the existence of $k$ positive solutions of (1.3).

Recently, the authors established the multiplicity of positive solutions for a nonlinear Schrödinger–Poisson system

$$ \begin{equation} \begin{cases} -\Delta u+\lambda u+K(x)\phi u= Q(x)|u|^{p-2}u &\text{in }\mathbb{R}^3, \\ -\Delta \phi=K(x)u^{2} &\text{in }\mathbb{R}^3, \end{cases} \end{equation} \tag{1.4} $$
where $\lambda$ is a positive parameter. In [7], Chen, Kuo, and Wu proved that the system (1.4) has $k$ positive solutions with $4\leqslant p<6$, $Q(x)\in C(\mathbb{R}^3)$ and $K(x)\equiv 1$. When $2<p<6$, Sun, Wu, and Feng [11] obtained the number of positive solutions of (1.4) under certain conditions on $Q(x)$ and $K(x)$.

In [14], Li and Gu studied the following Schrödinger–Poisson system with critical and supercritical nonlinearities:

$$ \begin{equation} \begin{cases} -\Delta u+u+\phi u= |u|^{p-2}u+u^5+b(x)|u|^{q-2}u &\text{in } \mathbb{R}^3, \\ -\Delta \phi=u^{2} &\text{in } \mathbb{R}^3, \end{cases} \end{equation} \tag{1.5} $$
where $p\in(4,6)$, $q\in(6,+\infty)$. They proved the existence of a nontrivial solution of the system (1.5) by using the variational method.

Compare our work with [14], where the authors considered the Schrödinger–Poisson system with critical and supercritical nonlinear terms and obtained that the system has a nontrivial solution. Our result is different since we suppose that $Q(x)$ has $k$ strict local maxima and obtain the number of positive solutions.

Since the system (1.1) is of critical growth, the embedding $H_{r}^1(\mathbb{R}^3)\hookrightarrow L^{6}(\mathbb{R}^3)$ need not be compact. We overcome this difficulty by using the Brézis–Lieb lemma. Moreover, we obtain a compactness result for the supercritical term by using the potential function and thus overcome another difficulty.

Throughout this paper, we use the following notation. Let $H^1(\mathbb{R}^3)$, $D^{1,2}(\mathbb{R}^3)$ be the usual Sobolev spaces defined by $H^1(\mathbb{R}^3)=\{u\in L^2(\mathbb{R}^3) \colon |\nabla u|\in L^2(\mathbb{R}^3)\}$, $D^{1,2}(\mathbb{R}^3)=\{u\in L^6(\mathbb{R}^3) \colon |\nabla u|\in L^2(\mathbb{R}^3)\}$, and let $H_r^1(\mathbb{R}^3)$, $D_r^{1,2}(\mathbb{R}^3)$ be the corresponding subspaces of radial functions. The corresponding norms and inner products are defined by

$$ \begin{equation*} \begin{gathered} \, \langle u,v\rangle=\int_{\mathbb{R}^3}(\nabla u\,\nabla v+uv)\, dx, \\ \|u\|_{H^1(\mathbb{R}^3)}=\biggl(\int_{\mathbb{R}^3}(|\nabla u|^2+u^2)\, dx\biggr)^{1/2},\qquad u,v\in H^1(\mathbb{R}^3), \\ \langle u,v\rangle=\int_{\mathbb{R}^3}\nabla u\,\nabla v\, dx,\quad \|u\|_{D^{1,2}(\mathbb{R}^3)}=\biggl(\int_{\mathbb{R}^3}\nabla u\,\nabla v\, dx\biggr)^{1/2},\qquad u,v\in D^{1,2}(\mathbb{R}^3). \end{gathered} \end{equation*} \notag $$
In this paper, we consider the system (1.1) in the subspace $H_r^1(\mathbb{R}^3)$ and write $\|\,{\cdot}\,\|$ for the norm of $H_r^1(\mathbb{R}^3)$. The norm in $L^p(\mathbb{R}^3)$ is denoted by $|\,{\cdot}\,|_p$ for $p\in [1,+\infty)$. We write $C, C_1, C_2,\dots$ for various positive constants, which may vary from line to line. Let $S$ be the best constant for the Sobolev embedding $D^{1,2}(\mathbb{R}^3)\hookrightarrow L^{6}(\mathbb{R}^3)$, namely
$$ \begin{equation*} S=\inf_{u\in D^{1,2}(\mathbb{R}^3)\setminus \{0\}}\frac{\int_{\mathbb{R}^3}|\nabla u|^2\, dx}{\bigl(\int_{\mathbb{R}^3}|u|^{6}\, dx\bigr)^{1/3}}. \end{equation*} \notag $$

We now state our main result.

Theorem 1.1. Assume that $(Q_1)$ holds. Then there exist $\lambda_0,\mu_0> 0$ small enough such that for all $\lambda\in(0,\lambda_0)$ and $\mu\in(0,\mu_0)$, the system (1.1) has at least $k$ positive solutions.

Remark 1.2. A simple example satisfying $(Q_1)$ is

$$ \begin{equation*} Q(x)=\begin{cases} \sin^2x &\text{if }\,x\in\biggl(\dfrac{5\pi}{4}, (k+1)\pi\biggr), \\ \dfrac12 &\text{if }\,x\notin\biggl(\dfrac{5\pi}{4}, (k+1)\pi\biggr), \end{cases} \end{equation*} \notag $$
where $a_i=\pi/2+i\pi$ for all $i$, $1\leqslant i\leqslant k$.

§ 2. Preliminaries

With the help of the Lax–Milgram theorem, we see that for every $u\in H_r^1(\mathbb{R}^3)$ there exists a unique $\phi_u\in D_r^{1,2}(\mathbb{R}^3)$ such that $-\Delta\phi_u=u^2$. Hence the system (1.1) can be transformed into the following semilinear nonlocal elliptic equation:

$$ \begin{equation} -\Delta u+u+\lambda\phi_u u=Q(x)|u|^{4}u+\mu h(x)|u|^{q-2}u \qquad\text{in } \mathbb{R}^3. \end{equation} \tag{2.1} $$
The variational functional associated with (2.1) is of the form
$$ \begin{equation*} I_{\lambda,\mu}(u)=\frac{1}{2}\|u\|^2+\frac{\lambda}{4}\int_{\mathbb{R}^3}\phi_u u^2\, dx -\frac{1}{6}\int_{\mathbb{R}^3}Q(x)|u|^6 \, dx-\frac{\mu}{q}\int_{\mathbb{R}^3}h(x)|u|^{q}\, dx. \end{equation*} \notag $$
We say that $u\in H_{r}^1(\mathbb{R}^3)$ is a weak solution of (2.1) if $u$ satisfies
$$ \begin{equation*} \int_{\mathbb{R}^3}|\nabla u|^2\, dx +\int_{\mathbb{R}^3}u^2\, dx+\lambda\int_{\mathbb{R}^3}\phi_u u^2\, dx-\int_{\mathbb{R}^3}Q(x)|u|^{6}\, dx-\mu\int_{\mathbb{R}^3}h(x)|u|^{q}\, dx=0. \end{equation*} \notag $$
So if a nonzero solution exists, then it must lie in the Nehari manifold $\mathcal{N}_{\lambda,\mu}$, which is defined by
$$ \begin{equation*} \begin{aligned} \, \mathcal{N}_{\lambda,\mu} &=\biggl\{u\in H_{r}^1(\mathbb{R}^3)\setminus\{0\}\colon \int_{\mathbb{R}^3}|\nabla u|^2 \, dx+\int_{\mathbb{R}^3}u^2\, dx+\lambda\int_{\mathbb{R}^3}\phi_u u^2\, dx \\ &\qquad-\int_{\mathbb{R}^3}Q(x)|u|^{6}\, dx-\displaystyle\mu\int_{\mathbb{R}^3}h(x)|u|^{q}\, dx=0\biggr\}. \end{aligned} \end{equation*} \notag $$

By [18], [3], the function $\phi_u$ has the following properties.

Lemma 2.1. For every $u\in H_r^1(\mathbb{R}^3)$ there is a unique solution $\phi_u\in D_r^{1,2}(\mathbb{R}^3)$ of

$$ \begin{equation*} -\Delta\phi=u^{2}\quad\textit{in }\ \mathbb{R}^3. \end{equation*} \notag $$
Moreover, the following assertions hold.

(1) $\phi_u\geqslant 0$ for $x\in \mathbb{R}^3$ and $\phi_{tu}=t^{2}\phi_u$ for all $t> 0$.

(2) If $u$ is a radial function, then $\phi_u$ is radial.

(3)

$$ \begin{equation*} \|\phi_u\|_{D^{1,2}(\mathbb{R}^3)}^{2}=\int_{\mathbb{R}^3}\phi_uu^{2}\, dx\leqslant S^{-1}|u|_{12/5}^{4}\leqslant C\|u\|^4. \end{equation*} \notag $$

(4) Assume that $u_n\rightharpoonup u$ in $H_r^1(\mathbb{R}^3)$. Then $\phi_{u_n}\to \phi_u$ in $D_r^{1,2}(\mathbb{R}^3)$ and

$$ \begin{equation*} \lim_{n\to\infty}\int_{\mathbb{R}^3}\phi_{u_n}u_n^{2}\, dx=\int_{\mathbb{R}^3}\phi_{u}u^{2}\, dx. \end{equation*} \notag $$

Lemma 2.2. Assume that $0< \beta<3$ and $6<q<6+2\beta$. If $\{u_n\}$ is bounded and $u_n\rightharpoonup u$ in $H_{r}^1(\mathbb{R}^3)$, then

$$ \begin{equation*} \lim_{n\to\infty}\int_{\mathbb{R}^3}\frac{|x|^\beta}{1+|x|^\beta}|u_n|^q\, dx=\int_{\mathbb{R}^3}\frac{|x|^\beta}{1+|x|^\beta}|u|^q\, dx. \end{equation*} \notag $$

Proof. By Lemma 2.2 in [14], we have
$$ \begin{equation*} |u_n(x)|\leqslant C\|u_n\|\frac{1}{|x|^{1/2}},\quad\text{for a. e. }\ x\in \mathbb{R}^3. \end{equation*} \notag $$
By the boundedness of $\{u_n\}$, for $6<q<6+2\beta$, one has
$$ \begin{equation*} \begin{aligned} \, \int_{|x|\leqslant1}\frac{|x|^\beta}{1+|x|^\beta}|u_n|^q\, dx &\leqslant C\int_{|x|\leqslant1}\frac{|x|^\beta}{1+|x|^\beta}\, \frac{dx}{|x|^{q/2}} \\ &\leqslant C\int_{|x|\leqslant1} \frac{dx}{|x|^{q/2-\beta}}= C\int_0^1 \frac{dt}{t^{q/2-\beta-2}} < +\infty, \end{aligned} \end{equation*} \notag $$
and
$$ \begin{equation*} \begin{aligned} \, \int_{|x|\geqslant1}\frac{|x|^\beta}{1+|x|^\beta}|u_n|^q\, dx &\leqslant C\int_{|x|\geqslant1}\frac{|x|^\beta}{1+|x|^\beta}\, \frac{dx}{|x|^{q/2}} \\ &\leqslant C\int_{|x|\geqslant1}\, \frac{dx}{|x|^{q/2}}= C\int_1^{+\infty} \frac{dt}{t^{q/2-2}}< +\infty. \end{aligned} \end{equation*} \notag $$
Therefore, by the Lebesgue dominated convergence theorem, we get
$$ \begin{equation*} \int_{\mathbb{R}^3}\frac{|x|^\beta}{1+|x|^\beta}|u_n|^q\, dx\to\int_{\mathbb{R}^3}\frac{|x|^\beta}{1+|x|^\beta}|u|^q\, dx \end{equation*} \notag $$
as $n\to \infty$. The proof is complete.

Lemma 2.3. Suppose that $0< \beta<3$ and $6<q<6+2\beta$, for all $\lambda, \mu> 0$. Then the functional $I_{\lambda, \mu}$ satisfies the $(\mathrm{PS})_{c}$ condition for every $c\in(-\infty, S^{3/2}/(3Q_M^{1/2}))$.

Proof. Let $\{u_n\}\subset H_{r}^1(\mathbb{R}^3)$ be a $(\mathrm{PS})_{c}$ sequence, that is,
$$ \begin{equation} I_{\lambda,\mu}(u_n)\to c\quad\text{and}\quad I'_{\lambda,\mu}(u_n)\to 0\quad\text{as}\quad n\to\infty. \end{equation} \tag{2.2} $$
It follows from (2.2) that
$$ \begin{equation*} \begin{aligned} \, |c|+1+o{(\|u_n\|)} &\geqslant I_{\lambda,\mu}(u_n)-\frac{1}{4}\langle I'_{\lambda,\mu}(u_n),u_n\rangle \\ &=\biggl(\frac{1}{2}-\frac{1}{4}\biggr)\|u_n\|^2 +\biggl(\frac{1}{4}-\frac{1}{6}\biggr)\int_{\mathbb{R}^3}Q(x)|u_n|^6 \, dx \\ &\qquad+\mu\biggl(\frac{1}{4}-\frac{1}{q}\biggr) \int_{\mathbb{R}^3}h(x)|u_n|^{q}\, dx \geqslant\frac{1}{4}\|u_n\|^2. \end{aligned} \end{equation*} \notag $$
Therefore, $\{u_n\}$ is bounded in $H_{r}^1(\mathbb{R}^3)$. Thus, we may assume up to a subsequence, still denoted by $\{u_n\}$, that there exists $u\in H_{r}^1(\mathbb{R}^3)$ such that
$$ \begin{equation} \begin{cases} u_n\rightharpoonup u &\text{weakly in }H_{r}^1(\mathbb{R}^3), \\ u_n\to u &\text{strongly in }L^{p}(\mathbb{R}^3),\ 2\leqslant p<6, \\ u_n(x)\to u(x) &\text{a. e. in }\mathbb{R}^3 \end{cases} \end{equation} \tag{2.3} $$
as $n\to \infty$. Next, we prove that $u_n\to u$ strongly in $H_{r}^1(\mathbb{R}^3)$. Set $w_n=u_n-u$ and $\lim_{n\to\infty}\|w_n\|^2=l$. By (2.2), Brézis–Lieb’s lemma, Lemma 2.1, and Lemma 2.2, we have
$$ \begin{equation} \begin{aligned} \, &\|w_n\|^2+\|u\|^2+\lambda\int_{\mathbb{R}^3}\phi_u u^2\, dx-\int_{\mathbb{R}^3}Q(x)|w_n|^6\, dx-\int_{\mathbb{R}^3}Q(x)|u|^6\, dx \nonumber \\ &\qquad-\mu\int_{\mathbb{R}^3}h(x)|u|^q\, dx=o(1), \end{aligned} \end{equation} \tag{2.4} $$
and
$$ \begin{equation} \|u\|^2+\lambda\int_{\mathbb{R}^3}\phi_u u^2\, dx-\int_{\mathbb{R}^3}Q(x)|u|^6\, dx-\mu\int_{\mathbb{R}^3}h(x)|u|^q\, dx=0. \end{equation} \tag{2.5} $$
It follows from (2.4) and (2.5) that
$$ \begin{equation} \|w_n\|^2-\int_{\mathbb{R}^3}Q(x)|w_n|^6\, dx=o(1). \end{equation} \tag{2.6} $$
Since $\|w_n\|^2\to l$, by (2.6), one has
$$ \begin{equation*} \int_{\mathbb{R}^3}Q(x)|w_n|^6\, dx\to l\quad\text{as}\quad n\to\infty. \end{equation*} \notag $$
Applying the Sobolev inequality, by (2.6), we have
$$ \begin{equation} \|w_n\|^2\leqslant Q_M \int_{\mathbb{R}^3}|w_n|^6\, dx+o(1)\leqslant Q_M S^{-3}\|w_n\|^6+o(1). \end{equation} \tag{2.7} $$
Thus, by (2.7), we can deduce that $l^3\geqslant S^3l/Q_M$ as $n\to \infty$, which implies that either $l= 0$ or $l\geqslant S^{3/2}/Q_M^{1/2}$ as $n\to \infty$. If $l\geqslant S^{3/2}/Q_M^{1/2}$, then, by (2.2),
$$ \begin{equation*} c=\biggl(\frac{1}{2}-\frac{1}{6}\biggr)l+I_{\lambda,\mu}(u)\geqslant \frac{S^{3/2}}{3Q_M^{1/2}}+I_{\lambda,\mu}(u) \end{equation*} \notag $$
as $n\to \infty$. On one hand, by the definition of $c$, we get
$$ \begin{equation*} I_{\lambda,\mu}(u)\leqslant c-\frac{S^{3/2}}{3Q_M^{1/2}}<0. \end{equation*} \notag $$
On the other hand, by (2.5), we have
$$ \begin{equation*} I_{\lambda,\mu}(u) =\frac{1}{4}\, \|u\|^2+\frac{1}{12}\int_{\mathbb{R}^3}Q(x)|u|^6 \, dx +\mu\biggl(\frac{1}{4}-\frac{1}{q}\biggr) \int_{\mathbb{R}^3}h(x)|u|^{q}\, dx \geqslant 0. \end{equation*} \notag $$
This is a contradiction. Hence, $l= 0$ and $u_n\to u$ strongly in $H_{r}^1(\mathbb{R}^3)$. The proof is complete.

Lemma 2.4. Assume that $0< \beta<3$, $6<q<6+2\beta$ and $\mu,\lambda> 0$. Then $\mathcal{N}_{\lambda,\mu}\neq\varnothing$.

Proof. Suppose that $u\in H_{r}^1(\mathbb{R}^3)\setminus\{0\}$. We define a fibering map $J_u\colon t\to I_{\lambda,\mu}(tu)$ by
$$ \begin{equation*} J_u(t)=\frac{t^2}{2}\|u\|^2+\lambda\frac{t^4}{4}\int_{\mathbb{R}^3}\phi_{u}u^2\, dx-\frac{t^6}{6}\int_{\mathbb{R}^3}Q(x)|u|^{6}\, dx -\mu\frac{t^q}{q}\int_{\mathbb{R}^3}h(x)|u|^{q}\, dx \end{equation*} \notag $$
for all $t\geqslant 0$. Then $J_u(0)= 0$ and $J_u(t)\to-\infty$ as $t\to + \infty$. Moreover, we get
$$ \begin{equation*} J'_u(t)= t\biggl[\|u\|^2+\lambda t^2\int_{\mathbb{R}^3}\phi_{u}u^2\, dx-t^{4}\int_{\mathbb{R}^3}Q(x)|u|^{6}\, dx -\mu t^{q-2}\int_{\mathbb{R}^3}h(x)|u|^{q}\, dx\biggr]. \end{equation*} \notag $$
We can see that there exists a unique $t_u> 0$ such that $J'_u(t_u)= 0$, that is,
$$ \begin{equation*} J'_u(t_u)= t_u\|u\|^2+\lambda t_u^3\int_{\mathbb{R}^3}\phi_{u}u^2\, dx-t_u^{5}\int_{\mathbb{R}^3}Q(x)|u|^{6}\, dx -\mu t_u^{q-1} \int_{\mathbb{R}^3}h(x)|u|^{q}\, dx=0. \end{equation*} \notag $$
Therefore, $t_u u\in\mathcal{N}_{\lambda,\mu}$ and $\mathcal{N}_{\lambda,\mu}\neq\varnothing$. The proof is complete.

Lemma 2.5. Suppose that $0< \beta<3$, $6<q<6+2\beta$, and $\lambda,\mu> 0$. Then the functional $I_{\lambda,\mu}(u)$ is coercive and bounded from below on $\mathcal{N_{\lambda,\mu}}$.

Proof. For all $u\in\mathcal{N_{\lambda,\mu}}$, we have
$$ \begin{equation*} \begin{aligned} \, I_{\lambda,\mu}(u)&=\frac{1}{2}\|u\|^2+\frac{\lambda}{4}\int_{\mathbb{R}^3}\phi_u u^2\, dx -\frac{1}{6}\int_{\mathbb{R}^3}Q(x)|u|^6 \, dx -\frac{\mu}{q}\int_{\mathbb{R}^3}h(x)|u|^{q}\, dx \\ &=\biggl(\frac{1}{2}-\frac{1}{4}\biggr)\|u\|^2+\biggl(\frac{1}{4}-\frac{1}{6}\biggr) \int_{\mathbb{R}^3}Q(x)|u|^6 \, dx \\ &\qquad+\mu\biggl(\frac{1}{4}-\frac{1}{q}\biggr) \int_{\mathbb{R}^3}h(x)|u|^{q}\, dx \geqslant\frac{1}{4}\|u\|^2, \end{aligned} \end{equation*} \notag $$
which implies that $I_{\lambda,\mu}(u)$ is coercive and bounded from below on $u\in\mathcal{N_{\lambda,\mu}}$. The proof is complete.

Lemma 2.6. Assume that $u_0$ is a local minimizer of $I_{\lambda,\mu}$ on $\mathcal{N_{\lambda,\mu}}$. Then $I_{\lambda,\mu}'(u_0)= 0$ in $H_r^{-1}(\mathbb{R}^3)$.

Proof. If $u_0$ is a local minimizer for $I_{\lambda,\mu}$ on $\mathcal{N_{\lambda,\mu}}$, then, by the definition of $J_u(t)$ in Lemma 2.4, there exists a neighborhood $U$ of $u_0$ in $H_r^1(\mathbb{R}^3)$ such that
$$ \begin{equation*} I_{\lambda,\mu}(u_0)=\min_{u\in\mathcal{N_{\lambda,\mu}}}I_{\lambda,\mu}(u)=\min_{u\in U\setminus\{0\},\,J'_u(1)=0}I_{\lambda,\mu}(u) \end{equation*} \notag $$
and
$$ \begin{equation*} J'_u(1)=\|u\|^2+\lambda\int_{\mathbb{R}^3}\phi_u u^2\, dx-\int_{\mathbb{R}^3}Q(x)|u|^{6}\, dx -\mu\int_{\mathbb{R}^3}h(x)|u|^{q}\, dx. \end{equation*} \notag $$
Combining this with Lemma 2.4 and the equality $J'_u(1)= 0$, we get
$$ \begin{equation*} \begin{aligned} \, J''_u(1)&=\|u\|^2+3\lambda\int_{\mathbb{R}^3}\phi_u u^2\, dx-5\int_{\mathbb{R}^3}Q(x)|u|^{6}\, dx -(q-1)\mu\int_{\mathbb{R}^3}h(x)|u|^{q}\, dx \\ &=(1-3)\|u\|^2-(5-3)\int_{\mathbb{R}^3}Q(x)|u|^{6}\, dx -(q-1-3)\mu\int_{\mathbb{R}^3}h(x)|u|^{q}\, dx <0. \end{aligned} \end{equation*} \notag $$
It follows from the theory of Lagrange multipliers that there exists a $\theta\in\mathbb{R}$ such that $I'_{\lambda,\mu}(u_0)=\theta J''_{u_0}$. Moreover, since $u_0\in \mathcal{N_{\lambda,\mu}}$, one has
$$ \begin{equation*} 0=\langle I'_{\lambda,\mu}(u_0),u_0\rangle=\theta \langle J''_{u_0}(1),u_0\rangle, \end{equation*} \notag $$
which implies that $\theta= 0$. Therefore, we obtain $I'_{\lambda,\mu}(u_0)= 0$ in $H_r^{-1}(\mathbb{R}^3)$. The proof is complete.

Choose $r> 0$ small enough such that $\bigcup_{i=1}^k\overline{B_{r}(a_i)}\subset \mathbb{R}^3$ and $\overline{B_{r}(a_i)}\cap\overline{B_{r}(a_j)}=\varnothing$ for $i\neq j$, $i,j=1,2,\dots,k$, where $\overline{B_{r}(a_i)}=\{x\in \mathbb{R}^3\colon |x-a_i|\leqslant r\}$. We minimize the functional $I_{\lambda,\mu}$ on some submanifolds of $\mathcal{N_{\lambda,\mu}}$. For this we define a barycenter map $g\colon H_r^{1}(\mathbb{R}^3)\setminus\{0\}\to \mathbb{R}^3$ by

$$ \begin{equation*} g(u)=\frac{\int_{\mathbb{R}^3}x|u|^{6}\, dx}{\int_{\mathbb{R}^3}|u|^{6}\, dx}, \end{equation*} \notag $$
see [23] or [28]. Set
$$ \begin{equation*} \begin{alignedat}{2} \mathcal{N}_{\lambda,\mu}^{i} &=\{u\in\mathcal{N_{\lambda,\mu}}\colon |g(u)-a_i|<r\}, &\qquad &1\leqslant i\leqslant k, \\ \mathcal{O}_{\lambda,\mu}^i &=\{u\in\mathcal{N_{\lambda,\mu}}\colon |g(u)-a_i|=r\}, &\qquad &1\leqslant i\leqslant k. \end{alignedat} \end{equation*} \notag $$
According to Lemma 2.4, we know that $\mathcal{N}_{\lambda,\mu}^{i} \neq \varnothing $ for $1\leqslant i\leqslant k$. By Lemma 2.5, we can define
$$ \begin{equation*} m_{\lambda,\mu}^i=\inf_{u\in \mathcal{N}_{\lambda,\mu}^{i}}I_{\lambda,\mu}(u)\quad\text{and}\quad \overline{m}_{\lambda,\mu}^i=\inf_{u\in \mathcal{O}_{\lambda,\mu}^i}I_{\lambda,\mu}(u), \qquad 1\leqslant i\leqslant k. \end{equation*} \notag $$

As is well known, the function

$$ \begin{equation*} U(x)=\frac{3^{1/4}}{(1+|x|^2)^{1/2}},\qquad x\in \mathbb{R}^3, \end{equation*} \notag $$
satisfies
$$ \begin{equation*} -\Delta U=U^{5}\quad \text{in }\mathbb{R}^3. \end{equation*} \notag $$
Let $\rho> 0$ be small enough such that $0\notin B_\rho(a_i)\subset \mathbb{R}^3$ for $i=1,2,\dots,k$. Choose a cut-off function $\varphi_i(x)\in C_0^\infty(\mathbb{R}^3)$ such that $0\leqslant \varphi_i(x)\leqslant 1$ for $x\in \mathbb{R}^3$ and
$$ \begin{equation*} \varphi_i(x)= \begin{cases} 1, &|x-a_i|\leqslant\dfrac{\rho}{2}, \\ 0, &|x-a_i|\geqslant\rho. \end{cases} \end{equation*} \notag $$
Let
$$ \begin{equation*} u_\varepsilon^i(x)=\varepsilon^{-1/2}\varphi_i(x)U\biggl(\frac{x-a_i}{\varepsilon}\biggr) =\frac{\varphi_i(x)3^{1/4} \varepsilon^{1/2}}{[\varepsilon^2+|x-a_i|^2]^{1/2}},\qquad\varepsilon>0. \end{equation*} \notag $$
Using the estimates obtained by Brézis and Nirenberg [29], we have the following results:
$$ \begin{equation} \begin{cases} |\nabla u_\varepsilon^i|_2^2= S^{3/2}+O(\varepsilon), \\ |u_\varepsilon^i|_6^6= S^{3/2}+O(\varepsilon^{3}) \end{cases} \end{equation} \tag{2.8} $$
and
$$ \begin{equation} |u_\varepsilon^i|_t^t= \begin{cases} O(\varepsilon^{t/2}), &t\in[2,3), \\ O(\varepsilon^{t/2}|{\ln\varepsilon}|), &t=3, \\ O(\varepsilon^{(6-t)/2}), &t\in(3,6). \end{cases} \end{equation} \tag{2.9} $$
Then we have the following lemma.

Lemma 2.7. Suppose that $(Q_1)$ holds, $0< \beta<3$, $6<q<6+2\beta$ and $\lambda, \mu> 0$. Then

$$ \begin{equation*} m_{\lambda,\mu}^i<\frac{S^{3/2}}{3Q_M^{1/2}},\qquad 1\leqslant i\leqslant k. \end{equation*} \notag $$

Proof. By Lemma 2.4, one can find $t_\varepsilon^i> 0$ and $t_\varepsilon^iu_\varepsilon^i\in\mathcal{N}_{\lambda, \mu}$ in such a way that $\sup_{t\geqslant0}I_{\lambda,\mu}(tu_\varepsilon^i)=I_{\lambda,\mu}(t_\varepsilon^i u_\varepsilon^i)$. By Lemma 3.2 in [14], we can assume that there exist positive constants $t_1$, $t_2$ such that $0< t_1<t_\varepsilon^i<t_2< + \infty$. Moreover, by the definition of $g(u)$, we have
$$ \begin{equation*} \begin{aligned} \, g(t_\varepsilon^iu_\varepsilon^i)&=\frac{\int_{\mathbb{R}^3}x|\varepsilon^{-1/2}t_\varepsilon^i \varphi_i(x)U((x-a_i)/\varepsilon)|^6\, dx} {\int_{\mathbb{R}^3}|\varepsilon^{-1/2}t_\varepsilon^i\varphi_i(x)U((x-a_i)/\varepsilon)|^6\, dx} \\ &= \frac{\int_{\mathbb{R}^3}(a_i+\varepsilon y)|t_\varepsilon^i\varphi_i(a_i+\varepsilon y)U(y)|^6\, dy}{\int_{\mathbb{R}^3}|t_\varepsilon^i\varphi_i(a_i+\varepsilon y)U(y)|^6\, dy} \to a_i\quad\text{as}\quad\varepsilon\to0. \end{aligned} \end{equation*} \notag $$
Therefore, there exists an $\varepsilon_0> 0$ such that $g(t_\varepsilon^iu_\varepsilon^i)\in B_{r}(a_i)$ for every $\varepsilon$, $0< \varepsilon<\varepsilon_{0}$. It follows that $t_\varepsilon^i u_\varepsilon^i\in \mathcal{N}_{\lambda,\mu}^{i}$, $1\leqslant i\leqslant k$. Then we have
$$ \begin{equation*} m_{\lambda,\mu}^i\leqslant I_{\lambda,\mu}(t_\varepsilon^i u_\varepsilon^i)=\sup_{t\geqslant0}I_{\lambda,\mu}(tu_\varepsilon^i). \end{equation*} \notag $$
In addition, by $(Q_1)$, for every $\eta> 0$ there exists a $\delta> 0$ such that $|Q(x)-Q(a_i)|<\eta|x-a_i|^2$ for $0< |x-a_i|<\delta$. When $\varepsilon> 0$ is small enough, we have
$$ \begin{equation*} \begin{aligned} \, &\biggl|\int_{\mathbb{R}^3}Q(x)|u_\varepsilon^i|^6\, dx -\int_{\mathbb{R}^3}Q(a_i)|u_\varepsilon^i|^6\, dx\biggr| \leqslant\int_{\mathbb{R}^3}|Q(x)-Q(a_i)||u_\varepsilon^i|^6\, dx \\ &\qquad< \int_{B_{\delta}(a_i)}\eta|x-a_i|^2\frac{(3\varepsilon^2)^{3/2}}{(\varepsilon^2+|x-a_i|^2)^3}\, dx+ C\int_{\mathbb{R}^3\setminus{B_{\delta}(a_i)}} \frac{(3\varepsilon^2)^{3/2}}{(\varepsilon^2+|x-a_i|^2)^3}\, dx \\ &\qquad\leqslant C \eta\varepsilon^3\int_{0}^{\delta}\frac{y^{4}}{(\varepsilon^2+y^2)^{3}}\, dy+C\varepsilon^3\int_{\delta}^{+\infty}\frac{y^2}{(\varepsilon^2+y^2)^{3}}\, dy \\ &\qquad\leqslant C \eta\varepsilon^2\int_{0}^{\delta/\varepsilon}\frac{t^{4}}{(1+t^2)^3}\, dt+C\int_{\delta/\varepsilon}^{+\infty}\frac{t^2}{(1+t^2)^3}\, dt \\ &\qquad\leqslant C \eta\varepsilon^2\int_{0}^{+\infty}\frac{t^{4}}{(1+t^2)^3}\, dt+C\int_{\delta/\varepsilon}^{+\infty}\frac{1}{t^4}\, dt \\ &\qquad\leqslant C_1 \eta\varepsilon^2+C_2\varepsilon^3, \end{aligned} \end{equation*} \notag $$
where $C_1,C_2> 0$ (independent of $\eta$, $\varepsilon$). From this we derive that
$$ \begin{equation} \limsup_{\varepsilon\to 0}\frac{\bigl|\int_{\mathbb{R}^3}Q(x)|u_\varepsilon^i|^6\, dx-\int_{\mathbb{R}^3}Q(a_i)|u_\varepsilon^i|^6\, dx\bigr|}{\varepsilon^2}\leqslant C_1\eta. \end{equation} \tag{2.10} $$
Then from the arbitrariness of $\eta> 0$, by (2.8) and (2.10), one has
$$ \begin{equation} \int_{\mathbb{R}^3}Q(x)|u_\varepsilon^i|^6\, dx=Q(a_i)\int_{\mathbb{R}^3}|u_\varepsilon^i|^6\, dx+O(\varepsilon^2)=Q_MS^{3/2}+O(\varepsilon^2). \end{equation} \tag{2.11} $$
By Lemma 2.2 and (2.9), we have the following estimate:
$$ \begin{equation} \int_{\mathbb{R}^3}\phi_{u_\varepsilon^i}|u_\varepsilon^i|^2\, dx\leqslant C|u_\varepsilon^i|_{12/5}^4\leqslant O(\varepsilon^2). \end{equation} \tag{2.12} $$
To estimate the supercritical term, we use the following inequality:
$$ \begin{equation*} |a+b|^p\leqslant 2^p(|a|+|b|),\qquad a,b\in \mathbb{R},\quad p>0. \end{equation*} \notag $$
According to the definition of $u_\varepsilon^i$, by (2.8) and $0< \beta<3$, there exists $C_3> 0$ such that
$$ \begin{equation} \begin{aligned} \, &\int_{\mathbb{R}^3}\frac{|x|^\beta}{1+|x|^\beta}|u_\varepsilon^i|^q\, dx \geqslant C\int_{|x-a_i|\leqslant \rho/2}\frac{|x|^\beta}{1+|x|^\beta} \frac{\varepsilon^{q/2}}{(\varepsilon^2+|x-a_i|^2)^{q/2}}\, dx \nonumber \\ &\geqslant C\varepsilon^{q/2}\int_{|x-a_i|\leqslant \rho/2}\frac{|x|^\beta}{1+(|a_i|+\rho/2)^\beta}\, \frac{1}{(\varepsilon^2+|x-a_i|^2)^{q/2}}\, dx \nonumber \\ &\geqslant C\varepsilon^{q/2}\int_{|x-a_i|\leqslant \rho/2}\frac{|x|^\beta +|x-a_i|^\beta-|x-a_i|^\beta}{(\varepsilon^2+|x-a_i|^2)^{q/2}}\, dx \nonumber \\ &\geqslant C\varepsilon^{q/2}\int_{|x-a_i|\leqslant \rho/2}\frac{2^{-\beta}|a_i|^\beta}{(\varepsilon^2\,{+}\,|x-a_i|^2)^{q/2}}\, dx- C\varepsilon^{q/2}\int_{|x-a_i|\leqslant \rho/2} \frac{|x-a_i|^\beta}{(\varepsilon^2{+}\,|x-a_i|^2)^{q/2}}\, dx \nonumber \\ &\geqslant C\varepsilon^{q/2}\int_0^{\rho/2}\frac{r^2}{(\varepsilon^2+r^2)^{q/2}}\, dr -C\varepsilon^{q/2}\int_0^{\rho/2}\frac{r^{2+\beta}}{(\varepsilon^2+r^2)^{q/2}}\, dr \nonumber \\ &= C\varepsilon^{3-q/2}\int_0^{\rho/2\varepsilon}\frac{t^2}{(1+t^2)^{q/2}}\, dt -C\varepsilon^{3+\beta-q/2}\int_0^{\rho/2\varepsilon}\frac{t^{2+\beta}}{(1+t^2)^{q/2}}\, dt \nonumber \\ &\geqslant C\varepsilon^{3-q/2}\int_0^1t^2\, dt -C\varepsilon^{3+\beta-q/2}\biggl(\int_0^1t^{2+\beta}\, dt+\int_1^{+\infty}t^{2+\beta-q}\, dt\biggr) \nonumber \\ &\geqslant C\varepsilon^{3-q/2} -C\varepsilon^{3+\beta-q/2}\biggl(\frac{1}{3+\beta} +\frac{1}{3+\beta-q}t^{3+\beta-q}\bigg|_1^{+\infty}\biggr) \geqslant C_3\varepsilon^{3-q/2}. \end{aligned} \end{equation} \tag{2.13} $$
Combining this with (2.11), (2.12) and (2.13), we have
$$ \begin{equation*} \begin{aligned} \, m_{\lambda,\mu}^i&\leqslant I_{\lambda,\mu}(t_\varepsilon^i u_\varepsilon^i)=\sup_{t\geqslant0}I_{\lambda,\mu}(tu_\varepsilon^i) \leqslant\sup_{t\geqslant0}\biggl\{\frac{t^2}{2}\|u_\varepsilon^i\|^2 +\lambda\frac{t^4}{4}\int_{\mathbb{R}^3}\phi_{u_\varepsilon^i}|u_\varepsilon^i|^2\, dx \\ &\qquad\qquad\qquad\qquad-\frac{t^6}{6}\int_{\mathbb{R}^3}Q(x)|u_\varepsilon^i|^{6}\, dx-\mu\frac{t^q}{q}\int_{\mathbb{R}^3}\frac{|x|^\beta}{1+|x|^\beta}|u_\varepsilon^i|^{q}\, dx\biggr\} \\ &\leqslant \sup_{t\geqslant0}\biggl\{\frac{t^2}{2}\int_{\mathbb{R}^3}|\nabla u_\varepsilon^i|^2\, dx-\frac{t^6}{6}\int_{\mathbb{R}^3}Q(x)|u_\varepsilon^i|^{6}\, dx\biggr\}+O(\varepsilon)+\lambda C_4\varepsilon^2-C_3\mu\varepsilon^{3-q/2} \\ &\leqslant\frac{1}{3}\frac{\bigl(\int_{\mathbb{R}^3}|\nabla u_\varepsilon^i|^2\, dx\bigr)^{3/2}}{\bigl(\int_{\mathbb{R}^3}Q(x)|u_\varepsilon^i|^{6}\, dx\bigr)^{1/2}}+O(\varepsilon)+\lambda C_4\varepsilon^2-C_3\mu\varepsilon^{3-q/2} \\ &\leqslant \frac{S^{3/2}}{3Q_M^{1/2}}+O(\varepsilon)+\lambda C_4\varepsilon^2-C_3\mu\varepsilon^{3-q/2}< \frac{S^{3/2}}{3Q_M^{1/2}}, \end{aligned} \end{equation*} \notag $$

where $C_4> 0$. The proof is complete.

Lemma 2.8. There exist $\lambda_0, \mu_0> 0$ small enough such that the following inequalities hold whenever $0< \lambda<\lambda_0$ and $0< \mu<\mu_0$:

$$ \begin{equation*} \overline{m}_{\lambda,\mu}^{\,i}>\frac{S^{3/2}}{3Q_M^{1/2}},\qquad 1\leqslant i\leqslant k. \end{equation*} \notag $$

Proof. Indeed, arguing by contradiction, assume that there are positive sequences $\{\lambda_n\}$ and $\{\mu_n\}$ such that $\lambda_n\to 0$ and $\mu_n\to 0$ as $n\to \infty$ and
$$ \begin{equation*} \overline{m}_{\lambda_n,\mu_n}^{\,i}\to m \leqslant\frac{S^{3/2}}{3Q_M^{1/2}}. \end{equation*} \notag $$
Hence there is a sequence $\{u_n\}\subset \mathcal{O}_{\lambda_n,\mu_n}^i$ such that $I_{\lambda_n,\mu_n}(u_n)\to m$ as $n\to \infty$ and
$$ \begin{equation} \|u_n\|^2+\lambda_n\int_{\mathbb{R}^3}\phi_{u_n} u_n^2\, dx =\int_{\mathbb{R}^3}Q(x)|u_n|^{6}\, dx+\mu_n\int_{\mathbb{R}^3}h(x)|u_n|^{q}\, dx. \end{equation} \tag{2.14} $$
Since $\{u_n\}$ is bounded in $H_r^{1}(\mathbb{R}^3)$, it follows from Lemma 2.1 and Lemma 2.2 that
$$ \begin{equation*} \lim_{n\to\infty}\lambda_n\int_{\mathbb{R}^3}\phi_{u_n} u_n^2\, dx\leqslant\lim_{n\to \infty}\lambda_nC_1\|u_n\|^4=0 \end{equation*} \notag $$
and
$$ \begin{equation*} \lim_{n\to\infty}\mu_n\int_{\mathbb{R}^3}h(x)|u_n|^{q}\, dx\leqslant\lim_{n\to \infty}\mu_nC_2\|u_n\|^q=0, \end{equation*} \notag $$
where $C_1,C_2> 0$ are constants. By (2.14) and the Sobolev inequality, one has
$$ \begin{equation*} \|u_n\|^2\leqslant \|u_n\|^2+\lambda_n\int_{\mathbb{R}^3}\phi_u u^2\, dx\leqslant Q_MS^{-3}\|u_n\|^6+\mu_n C_2\|u_n\|^q, \end{equation*} \notag $$
which implies that there exists a positive constant $C_3> 0$ such that $\|u_n\|\geqslant C_3$ for every $\lambda_n\in(0,\lambda_0)$ and $\mu_n\in(0,\mu_0)$. Thus, we can fix a constant $A> 0$ such that
$$ \begin{equation*} \|u_n\|^2 \geqslant A\quad\text{and}\quad\int_{\mathbb{R}^3}Q(x)|u_n|^{6}\, dx\geqslant A. \end{equation*} \notag $$
Then, up to a subsequence, there exists an $a> 0$ such that
$$ \begin{equation} \lim_{n\to\infty} \|u_n\|^2 =\lim_{n\to\infty}\int_{\mathbb{R}^3}Q(x)|u_n|^6\, dx=a. \end{equation} \tag{2.15} $$
It follows from (2.15) that
$$ \begin{equation} a\leqslant Q_M\lim_{n\to\infty}\int_{\mathbb{R}^3}|u_n|^6\, dx\leqslant Q_MS^{-3}\lim_{n\to\infty}\biggl(\int_{\mathbb{R}^3}|\nabla u_n|^2\, dx\biggr)^3\leqslant Q_MS^{-3}a^3. \end{equation} \tag{2.16} $$
We deduce that $a\geqslant S^{3/2}/Q_M^{1/2}$. On the other hand, we have
$$ \begin{equation} \begin{aligned} \, \frac{1}{3}a &=\frac{1}{2} \|u_n\|^2 +\frac{\lambda_n}{4}\int_{\mathbb{R}^3}\phi_{u_n} u_n^2\, dx -\frac{1}{6}\int_{\mathbb{R}^3}Q(x)|u_n|^{6}\, dx \nonumber \\ &\qquad-\frac{\mu_n}{q}\int_{\mathbb{R}^3}h(x)|u_n|^{q}\, dx+o(1) = I_{\lambda_n,\mu_n}(u_n)+o(1)\leqslant\frac{S^{3/2}}{3Q_M^{1/2}}. \end{aligned} \end{equation} \tag{2.17} $$
Hence, it follows from (2.16) and (2.17) that $a=S^{3/2}/Q_M^{1/2}$ and
$$ \begin{equation*} \lim_{n\to\infty}\int_{\mathbb{R}^3}Q_M|u_n|^6\, dx=\frac{S^{3/2}}{Q_M^{1/2}}. \end{equation*} \notag $$
Then, we have
$$ \begin{equation} \lim_{n\to\infty}\int_{\mathbb{R}^3}(Q_M-Q(x))|u_n|^6\, dx=0. \end{equation} \tag{2.18} $$
Setting $w_n=u_n/|u_n|_{6}$, we have $|w_n|_{6}= 1$ and
$$ \begin{equation*} \lim_{n\to\infty}\int_{\mathbb{R}^3}|\nabla w_n|^2\, dx=\lim_{n\to\infty}\frac{|\nabla u_n|_2^2}{|u_n|_{6}^2}=S. \end{equation*} \notag $$
Therefore, $\{w_n\}$ is a minimizing sequence for $S$. By [30], there exists an $x_0\in\mathbb{R}^3$ such that
$$ \begin{equation*} |\nabla w_n|^2\rightharpoonup d\tilde{\mu}=S\delta_{x_0},\qquad |w_n|^6\rightharpoonup d\tilde{\nu}=\delta_{x_0} \end{equation*} \notag $$
weakly in the sense of measure, where $\tilde{\mu},\tilde{\nu}$ are finite measures and $\delta_{x_0}$ is the Dirac mass at $x_0\in\mathbb{R}^3$. Since $w_n\in \mathcal{O}_{\lambda_n,\mu_n}^i$, we have
$$ \begin{equation*} g(w_n)=\frac{\int_{\mathbb{R}^3}x|w_n|^{6}\, dx}{\int_{\mathbb{R}^3}|w_n|^{6}\, dx}\to x_0, \end{equation*} \notag $$
as $n\to \infty$, which implies that $x_0 \notin \{a_i\colon i=1,2,\dots,k\}$. Consequently, by (2.18), we have
$$ \begin{equation*} Q_M=\lim_{n\to\infty}\int_{\mathbb{R}^3}Q_M|w_n|^6\, dx =\lim_{n\to\infty}\int_{\mathbb{R}^3}Q(x)|w_n|^6\, dx=Q(x_0). \end{equation*} \notag $$
This is a contradiction because $Q(x)$ is not a constant function. The proof is complete.

Lemma 2.9. For every $u\in \mathcal{N}_{\lambda,\mu}^{i}$, $1\leqslant i\leqslant k$, one can find a number $\delta> 0$ and a differentiable function $f\colon B_{\delta}(0)\subset H_r^{1}(\mathbb{R}^3)\to \mathbb{R}$ such that $f(0)= 1$, $f(v)(u-v)\in\mathcal{N}_{\lambda,\mu}^{i}$ for every $v\in B_{\delta}(0)$ and

$$ \begin{equation*} \begin{aligned} \, &\langle f'(0),\varphi\rangle \\ &{=}\,\frac{2\!\int_{\mathbb{R}^3}(\nabla u \nabla\varphi\,{+}\,u\varphi) \, dx\,{+}\,4\lambda\!\int_{\mathbb{R}^{3\!}}\phi_uu\varphi \, dx\,{-}\,6\!\int_{\mathbb{R}^{3\!}}Q(x)|u|^4u\varphi \, dx\,{-}\,\mu q\!\int_{\mathbb{R}^{3\!}}h(x)|u|^{q-2}u\varphi \, dx}{\int_{\mathbb{R}^3}(|\nabla u|^2\,{+}\,u^2) \, dx\,{+}\,3\lambda\!\int_{\mathbb{R}^{3\!}}\phi_u|u|^2\, dx\,{-}\,5\!\int_{\mathbb{R}^{3\!}}Q(x)|u|^{6}\, dx \,{-}\,\mu(q\,{-}\,1)\!\int_{\mathbb{R}^{3\!}}h(x)|u|^{q}\, dx} \end{aligned} \end{equation*} \notag $$
for all $\varphi\in H_r^{1}(\mathbb{R}^3)$.

Proof. It is similar to the proof of Lemma 4.4 in [25]. Given any $u\in\mathcal{N}_{\lambda,\mu}^{i}$, we define a function $G\colon \mathbb{R}^+\times H_r^{1}(\mathbb{R}^3)\to \mathbb{R}$ by
$$ \begin{equation*} \begin{aligned} \, G(t,v)&= t\int_{\mathbb{R}^3}(|\nabla(u-v)|^2+|u-v|^2)\, dx+\lambda t^3\int_{\mathbb{R}^3}\phi_{u-v} |u-v|^2\, dx \\ &\qquad-t^5\int_{\mathbb{R}^3}Q(x)|u-v|^{6}\, dx-\mu t^{q-1}\int_{\mathbb{R}^3}h(x)|u-v|^{q}\, dx. \end{aligned} \end{equation*} \notag $$
Therefore, $G(1,0)=\langle I'_{\lambda,\mu}(u),u\rangle= 0$ and
$$ \begin{equation*} \begin{aligned} \, G_t(1,0)&= G_t(1,0)-3\langle I'_{\lambda,\mu}(u),u\rangle \\ &= (1-3)\int_{\mathbb{R}^3}(|\nabla u|^2+|u|^2)\, dx-(5-3)\int_{\mathbb{R}^3}Q(x)|u|^{6}\, dx \\ &\qquad-\mu(q-1-3)\int_{\mathbb{R}^3}h(x)|u|^{q}\, dx <0. \end{aligned} \end{equation*} \notag $$
Then, according to the implicit function theorem, there exists $\delta> 0$ small enough and a differentiable function $f\colon B_{\delta}(0)\subset H_r^{1}(\mathbb{R}^3)\to \mathbb{R}$ such that $f(0)= 1$ and $G(f(v),v)= 0$ for every $v\in B_{\delta}(0)$. Since $G(f(v),v)= 0$, it is easy to check that $f(v)(u-v)\in\mathcal{N}_{\lambda,\mu}^{i}$ and
$$ \begin{equation*} \begin{aligned} \, &\langle f'(0),\varphi\rangle= -\frac{\langle G_v(1,0),\varphi\rangle}{G_t(1,0)} \\ &{=}\,\frac{2\!\int_{\mathbb{R}^3}(\nabla u\, \nabla\varphi\,{+}\,u\varphi) \, dx\,{+}\,4\lambda\!\int_{\mathbb{R}^3}\!\phi_uu\varphi \, dx\,{-}\,6\!\int_{\mathbb{R}^3}\!Q(x)|u|^4u\varphi \, dx\,{-}\,\mu q\!\int_{\mathbb{R}^3}\!h(x)|u|^{q{-}2}u\varphi \, dx}{\int_{\mathbb{R}^3}(|\nabla u|^2\,{+}\,u^2) \, dx\,{+}\,3\lambda\!\int_{\mathbb{R}^3}\!\phi_u|u|^2\, dx\,{-}\,5\!\int_{\mathbb{R}^3}\!Q(x)|u|^{6}\, dx \,{-}\,\mu(q\,{-}\,1)\!\int_{\mathbb{R}^3}\!h(x)|u|^{q}\, dx}. \end{aligned} \end{equation*} \notag $$
The proof is complete.
Proof of Theorem 1.1. Combining Lemma 2.7 and Lemma 2.8, we infer that
$$ \begin{equation*} m_{\lambda,\mu}^i<\frac{S^{3/2}}{3Q_M^{1/2}}<\overline{m}_{\lambda,\mu}^{\,i}, \end{equation*} \notag $$
for all $\lambda\in(0,\lambda_0)$ and $\mu\in(0,\mu_0)$. Consequently, it follows that
$$ \begin{equation*} m_{\lambda,\mu}^i=\inf\{I_{\lambda,\mu}(u)\colon u\in\mathcal{N}_{\lambda,\mu}^{i}\cup\mathcal{O}_{\lambda,\mu}^i\}. \end{equation*} \notag $$
Let $\{u_n^i\}\subset \mathcal{N}_{\lambda,\mu}^{i}\cup\mathcal{O}_{\lambda,\mu}^i$ be a minimizing sequence for $m_{\lambda,\mu}^i$. Then, by the Ekeland variational principle [31], there exists a subsequence, still denoted by $\{u_n^i\}$ such that
$$ \begin{equation} I_{\lambda,\mu}(u_n^i)\leqslant m_{\lambda,\mu}^i+\frac{1}{n},\qquad I_{\lambda,\mu}(w)\geqslant I_{\lambda,\mu}(u_n^i)-\frac{1}{n}\|w-u_n^i\|,\quad w\in{\mathcal{N}_{\lambda,\mu}^{i}\cup\mathcal{O}_{\lambda,\mu}^i}. \end{equation} \tag{2.19} $$
We now prove that $I'_{\lambda,\mu}(u_n^i)\to 0$ as $n\to \infty$. Indeed, by Lemma 2.9, there exists a $\delta_n^i> 0$ and a differentiable function $f_n^i\colon B_{\delta_n^i}(0)\subset H_r^{1}(\mathbb{R}^3)\to \mathbb{R}$ such that $f_n^i(0)= 1$, $f_n^i(v)(u_n^i-v)\in \mathcal{N}_{\lambda,\mu}^{i}$ for all $v\in B_{\delta_n^i}(0)$. Set $\varphi\in H_r^{1}(\mathbb{R}^3)$ with $\|\varphi\|= 1$ and $0< \delta<\delta_n^i$. Choosing $v=\delta\varphi$, one has $v=\delta\varphi\in B_{\delta_n^i}(0)$ and $f_n^i(\delta \varphi)(u_n^i-\delta \varphi)\in \mathcal{N}_{\lambda,\mu}^{i}$, it follows from (2.19) and the Taylor expansion that
$$ \begin{equation*} \begin{aligned} \, &\frac{\|f_n^i(\delta\varphi)(u_n^i-\delta \varphi)-u_n^i\|}{n} \geqslant I_{\lambda,\mu}(u_n^i)-I_{\lambda,\mu}[f_n^i(\delta\varphi)(u_n^i-\delta\varphi)] \\ &=\langle I'_{\lambda,\mu}(u_n^i),u_n^i-f_n^i(\delta\varphi)(u_n^i-\delta\varphi)\rangle +o\bigl(\|u_n^i-f_n^i(\delta\varphi)(u_n^i-\delta\varphi)\|\bigr) \\ &=\delta f_n^i(\delta\varphi)\langle I'_{\lambda,\mu}(u_n^i),\varphi\rangle+(1-f_n^i(\delta\varphi))\langle I'_{\lambda,\mu}(u_n^i),{u_n^i} \rangle +o\bigl(\|u_n^i-f_n^i(\delta\varphi)(u_n^i-\delta\varphi)\|\bigr) \\ &= \delta f_n^i(\delta\varphi)\langle I'_{\lambda,\mu}(u_n^i),\varphi\rangle +o\bigl(\|u_n^i-f_n^i(\delta\varphi)(u_n^i-\delta\varphi)\|\bigr). \end{aligned} \end{equation*} \notag $$
Hence, letting $\delta\to 0^+$, we get
$$ \begin{equation*} \begin{aligned} \, \langle I'_{\lambda,\mu}(u_n^i),\varphi\rangle &\leqslant\frac{\|u_n^i-f_n^i(\delta \varphi)(u_n^i-\delta\varphi)\|(1/n+|o(1)|)}{\delta |f_n^i(\delta \varphi)|} \\ &\leqslant \frac{\|u_n^i(f_n^i(\delta\varphi)-f_n^i(0))-\delta f_n^i(\delta\varphi)\varphi\|(1/n+|o(1)|)}{\delta |f_n^i(\delta\varphi)|} \\ &\leqslant \frac{\bigl(\|u_n^i\||f_n^i(\delta\varphi)-f_n^i(0)|+\delta |f_n^i(\delta\varphi)|\|\varphi\|\bigr)(1/n+|o(1)|)}{\delta |f_n^i(\delta\varphi)|} \\ &\leqslant\biggl(1+\|u_n^i\|\frac{|f_n^i(\delta\varphi)-f_n^i(0)|}{\delta|f_n^i(\delta\varphi)|}\biggr) \biggl(\frac{1}{n}+|o(1)|\biggr) \\ &\leqslant \bigl(1+\|u_n^i\|\|(f_n^i)'(0)\|\bigr)\biggl(\frac{1}{n}+|o(1)|\biggr). \end{aligned} \end{equation*} \notag $$
Since $\{u_n^i\}$ and $\{(f_n^i)'(0)\}$ are bounded, we deduce that $I'_{\lambda,\mu}(u_n^i)\to 0$ as $n\to \infty$. Therefore, $\{u_n^i\}$ is a $(\mathrm{PS})_{m_{\lambda,\mu}^i}$ sequence for $I_{\lambda,\mu}$ at the level $m_{\lambda,\mu}^i$. Since $I_{\lambda,\mu}(|u_n^i|)=I_{\lambda,\mu}(u_n^i)$, it follows that $u_n^i\geqslant 0$. By Lemma 2.3 and the inequality $m_{\lambda,\mu}^i<S^{3/2}/(3Q_M^{1/2})$, there exists a subsequence of $\{u_n^i\}$, still denoted by $\{u_n^i\}$, and a function $u^i\in H_r^{1}(\mathbb{R}^3)$ with $u^i\geqslant 0$ such that $u_n^i\to u^i$ strongly in $H_r^{1}(\mathbb{R}^3)$ for all $i$, $1\leqslant i\leqslant k$, as $n\to \infty$. It follows from Lemma 2.5 and Lemma 2.8 that $\|u_n^i\|\geqslant A> 0$ and
$$ \begin{equation*} \lim_{n\to\infty}I_{\lambda,\mu}(u_n^i)=I_{\lambda,\mu}(u^i)=m_{\lambda,\mu}^i>0, \end{equation*} \notag $$
whence $u^i\not\equiv 0$. Consequently, applying the strong maximum principle, we see that the system (1.1) has at least $k$ positive solutions $u^i$, $1\leqslant i\leqslant k$, for all $\lambda\in(0,\lambda_0)$ and $\mu\in(0,\mu_0)$. The proof is complete.

Bibliography

1. V. Benci, D. Fortunato, “An eigenvalue problem for the Schrödinger–Maxwell equations”, Topol. Methods Nonlinear Anal., 11:2 (1998), 283–293  crossref  mathscinet  zmath
2. O. Bokanowski, J. L. López, J. Soler, “On an exchange interaction model for quantum transport: the Schrödinger–Poisson–Slater system”, Math. Models Methods Appl. Sci., 13:10 (2003), 1397–1412  crossref  mathscinet  zmath
3. D. Ruiz, “The Schrödinger–Poisson equation under the effect of a nonlinear local term”, J. Funct. Anal., 237:2 (2006), 655–674  crossref  mathscinet  zmath
4. I. V. Barashenkov, A. D. Gocheva, V. G. Makhankov, I. V. Puzynin, “Stability of the soliton-like “bubbles””, Phys. D, 34:1-2 (1989), 240–254  crossref  mathscinet  zmath  adsnasa
5. В. Г. Картавенко, “Решения солитонного типа в ядерной гидродинамике”, Ядерная физика, 40 (1984), 377–388; англ. пер.: V. G. Kartavenko, “Soliton-like solutions in nuclear hydrodynamics”, Soviet J. Nuclear Phys., 40 (1984), 240–246  zmath
6. V. Benci, D. Fortunato, “Solitary waves of the nonlinear Klein–Gordon equation coupled with the Maxwell equations”, Rev. Math. Phys., 14:4 (2002), 409–420  crossref  mathscinet  zmath
7. Ching-yu Chen, Yueh-cheng Kuo, Tsung-fang Wu, “Existence and multiplicity of positive solutions for the nonlinear Schrödinger–Poisson equations”, Proc. Roy. Soc. Edinburgh Sect. A, 143:4 (2013), 745–764  crossref  mathscinet  zmath
8. Fuyi Li, Yuhua Li, Junping Shi, “Existence of positive solutions to Schrödinger–Poisson type systems with critical exponent”, Commun. Contemp. Math., 16:6 (2014), 1450036, 28 pp.  crossref  mathscinet  zmath
9. Gongbao Li, Shuangjie Peng, Shusen Yan, “Infinitely many positive solutions for the nonlinear Schrödinger–Poisson system”, Commun. Contemp. Math., 12:6 (2010), 1069–1092  crossref  mathscinet  zmath
10. Haidong Liu, “Positive solutions of an asymptotically periodic Schrödinger–Poisson system with critical exponent”, Nonlinear Anal. Real World Appl., 32 (2016), 198–212  crossref  mathscinet  zmath
11. Juntao Sun, Tsung-fang Wu, Zhaosheng Feng, “Multiplicity of positive solutions for a nonlinear Schrödinger–Poisson system”, J. Differential equations, 260:1 (2016), 586–627  crossref  mathscinet  zmath  adsnasa
12. A. Azzollini, “Concentration and compactness in nonlinear Schrödinger–Poisson system with a general nonlinearity”, J. Differential Equations, 249:7 (2010), 1746–1763  crossref  mathscinet  zmath  adsnasa
13. A. Ambrosetti, D. Ruiz, “Multiple bound states for the Schrödinger–Poisson problem”, Commun. Contemp. Math., 10:3 (2008), 391–404  crossref  mathscinet  zmath
14. Yuhua Li, Hua Gu, “Existence of solutions to Schrödinger–Poisson systems with critical and supercritical nonlinear terms”, Math. Methods Appl. Sci., 42:7 (2019), 2279–2286  crossref  mathscinet  zmath
15. A. Ambrosetti, “On Schrödinger–Poisson systems”, Milan. J. Math., 76 (2008), 257–274  crossref  mathscinet  zmath
16. A. Azzollini, P. d'Avenia, “On a system involving a critically growing nonlinearity”, J. Math. Anal. Appl., 387:1 (2012), 433–438  crossref  mathscinet  zmath
17. C. O. Alves, M. A. S. Souto, S. H. M. Soares, “Schrödinger–Poisson equations without Ambrosetti–Rabinowitz condition”, J. Math. Anal. Appl., 377:2 (2011), 584–592  crossref  mathscinet  zmath
18. Zhisu Liu, Shangjiang Guo, “On ground state solutions for the Schrödinger–Poisson equations with critical growth”, J. Math. Anal. Appl., 412:1 (2014), 435–448  crossref  mathscinet  zmath
19. Jian Zhang, “On the Schrödinger–Poisson equations with a general nonlinearity in the critical growth”, Nonlinear Anal., 75:18 (2012), 6391–6401  crossref  mathscinet  zmath
20. Jian Zhang, “On ground state and nodal solutions of Schrödinger–Poisson equations with critical growth”, J. Math. Anal. Appl., 428:1 (2015), 387–404  crossref  mathscinet  zmath
21. Lirong Huang, E. M. Rocha, Jianqing Chen, “Positive and sign-changing solutions of a Schrödinger–Poisson system involving a critical nonlinearity”, J. Math. Anal. Appl., 408:1 (2013), 55–69  crossref  mathscinet  zmath
22. Leiga Zhao, Fukun Zhao, “Positive solutions for Schrödinger–Poisson equations with a critical exponent”, Nonlinear Anal., 70:6 (2009), 2150–2164  crossref  mathscinet  zmath
23. Daomin Cao, J. Chabrowski, “Multiple solutions of nonhomogeneous elliptic equation with critical nonlinearity”, Differential Integral Equations, 10:5 (1997), 797–814  mathscinet  zmath
24. Daomin Cao, E. S. Noussair, “Multiple positive and nodal solutions for semilinear elliptic problems with critical exponents”, Indiana Univ. Math. J., 44:4 (1995), 1249–1271  crossref  mathscinet  zmath
25. Pigong Han, “Multiple solutions to singular critical elliptic equations”, Israel J. Math., 156 (2006), 359–380  crossref  mathscinet  zmath
26. Yi Sheng Huang, “Multiple positive solutions of nonhomogeneous equations involving the $p$-Laplacian”, Nonlinear Anal., 43:7 (2001), 905–922  crossref  mathscinet  zmath
27. Huei-li Lin, “Positive solutions for nonhomogeneous elliptic equations involving critical Sobolev exponent”, Nonlinear Anal., 75:4 (2012), 2660–2671  crossref  mathscinet  zmath
28. Jia-feng Liao, Jiu Liu, Peng Zhang, Chun-Lei Tang, “Existence and multiplicity of positive solutions for a class of elliptic equations involving critical Sobolev exponents”, Rev. R. Acad. Cienc. Exactas Fís. Nat. Ser. A Mat. RACSAM, 110:2 (2016), 483–501  crossref  mathscinet  zmath
29. H. Brézis, L. Nirenberg, “Positive solutions of nonlinear elliptic equations involving critical Sobolev exponents”, Comm. Pure. Appl. Math., 36:4 (1983), 437–477  crossref  mathscinet  zmath
30. P.-L. Lions, “The concentration-compactness principle in the calculus of variations. The limit case. I”, Rev. Mat. Iberoamericana, 1:1 (1985), 145–201  crossref  mathscinet  zmath
31. I. Ekeland, “On the variational principle”, J. Math. Anal. Appl., 48:2 (1974), 324–353  crossref  mathscinet  zmath

Образец цитирования: J. Lei, H. Suo, “Multiple positive solutions for a Schrödinger–Poisson system with critical and supercritical growths”, Изв. РАН. Сер. матем., 87:1 (2023), 33–48; Izv. Math., 87:1 (2023), 29–44
Цитирование в формате AMSBIB
\RBibitem{LeiSuo23}
\by J.~Lei, H.~Suo
\paper Multiple positive solutions for a~Schr\"odinger--Poisson system with critical and supercritical growths
\jour Изв. РАН. Сер. матем.
\yr 2023
\vol 87
\issue 1
\pages 33--48
\mathnet{http://mi.mathnet.ru/im9244}
\crossref{https://doi.org/10.4213/im9244}
\mathscinet{http://mathscinet.ams.org/mathscinet-getitem?mr=4634754}
\adsnasa{https://adsabs.harvard.edu/cgi-bin/bib_query?2023IzMat..87...29L}
\transl
\jour Izv. Math.
\yr 2023
\vol 87
\issue 1
\pages 29--44
\crossref{https://doi.org/10.4213/im9244e}
\isi{https://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcApp=Publons&SrcAuth=Publons_CEL&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=001054276700002}
\scopus{https://www.scopus.com/record/display.url?origin=inward&eid=2-s2.0-85171143050}
Образцы ссылок на эту страницу:
  • https://www.mathnet.ru/rus/im9244
  • https://doi.org/10.4213/im9244
  • https://www.mathnet.ru/rus/im/v87/i1/p33
  • Citing articles in Google Scholar: Russian citations, English citations
    Related articles in Google Scholar: Russian articles, English articles
    Известия Российской академии наук. Серия математическая Izvestiya: Mathematics
    Статистика просмотров:
    Страница аннотации:320
    PDF русской версии:25
    PDF английской версии:72
    HTML русской версии:145
    HTML английской версии:118
    Список литературы:50
    Первая страница:10
     
      Обратная связь:
     Пользовательское соглашение  Регистрация посетителей портала  Логотипы © Математический институт им. В. А. Стеклова РАН, 2024