Loading [MathJax]/extensions/TeX/boldsymbol.js
Research article

Nodal solutions for the Kirchhoff-Schrödinger-Poisson system in R3

  • Received: 23 May 2022 Revised: 05 July 2022 Accepted: 07 July 2022 Published: 13 July 2022
  • MSC : 35J20, 35J65

  • This paper is dedicated to studying the following Kirchhoff-Schrödinger-Poisson system:

    {(a+bR3|u|2dx)Δu+V(|x|)u+λϕu=K(|x|)f(u),xR3,Δϕ=u2,xR3,

    where V,K are radial and bounded away from below by positive numbers. Under some weaker assumptions on the nonlinearity f, we develop a direct approach to establish the existence of infinitely many nodal solutions {ub,λk} with a prescribed number of nodes k, by using the Gersgorin disc's theorem, Miranda theorem and Brouwer degree theory. Moreover, we prove that the energy of {ub,λk} is strictly increasing in k, and give a convergence property of {ub,λk} as b0 and λ0.

    Citation: Kun Cheng, Li Wang. Nodal solutions for the Kirchhoff-Schrödinger-Poisson system in R3[J]. AIMS Mathematics, 2022, 7(9): 16787-16810. doi: 10.3934/math.2022922

    Related Papers:

    [1] Chen Huang, Gao Jia . Schrödinger-Poisson system without growth and the Ambrosetti-Rabinowitz conditions. AIMS Mathematics, 2020, 5(2): 1319-1332. doi: 10.3934/math.2020090
    [2] Jin-Long Zhang, Da-Bin Wang . Existence of least energy nodal solution for Kirchhoff-type system with Hartree-type nonlinearity. AIMS Mathematics, 2020, 5(5): 4494-4511. doi: 10.3934/math.2020289
    [3] Xueqin Peng, Gao Jia, Chen Huang . Multiplicity of solutions for Schrödinger-Poisson system with critical exponent in R3. AIMS Mathematics, 2021, 6(3): 2059-2077. doi: 10.3934/math.2021126
    [4] Ting Xiao, Yaolan Tang, Qiongfen Zhang . The existence of sign-changing solutions for Schrödinger-Kirchhoff problems in R3. AIMS Mathematics, 2021, 6(7): 6726-6733. doi: 10.3934/math.2021395
    [5] Hui Jian, Shenghao Feng, Li Wang . Sign-changing solutions of critical quasilinear Kirchhoff-Schrödinger-Poisson system with logarithmic nonlinearity. AIMS Mathematics, 2023, 8(4): 8580-8609. doi: 10.3934/math.2023431
    [6] Yuan Shan, Baoqing Liu . Existence and multiplicity of solutions for generalized asymptotically linear Schrödinger-Kirchhoff equations. AIMS Mathematics, 2021, 6(6): 6160-6170. doi: 10.3934/math.2021361
    [7] Fumei Ye, Xiaoling Han . Global bifurcation result and nodal solutions for Kirchhoff-type equation. AIMS Mathematics, 2021, 6(8): 8331-8341. doi: 10.3934/math.2021482
    [8] Yang Pu, Hongying Li, Jiafeng Liao . Ground state solutions for the fractional Schrödinger-Poisson system involving doubly critical exponents. AIMS Mathematics, 2022, 7(10): 18311-18322. doi: 10.3934/math.20221008
    [9] Ya-Lei Li, Da-Bin Wang, Jin-Long Zhang . Sign-changing solutions for a class of p-Laplacian Kirchhoff-type problem with logarithmic nonlinearity. AIMS Mathematics, 2020, 5(3): 2100-2112. doi: 10.3934/math.2020139
    [10] Jinfu Yang, Wenmin Li, Wei Guo, Jiafeng Zhang . Existence of infinitely many normalized radial solutions for a class of quasilinear Schrödinger-Poisson equations in R3. AIMS Mathematics, 2022, 7(10): 19292-19305. doi: 10.3934/math.20221059
  • This paper is dedicated to studying the following Kirchhoff-Schrödinger-Poisson system:

    {(a+bR3|u|2dx)Δu+V(|x|)u+λϕu=K(|x|)f(u),xR3,Δϕ=u2,xR3,

    where V,K are radial and bounded away from below by positive numbers. Under some weaker assumptions on the nonlinearity f, we develop a direct approach to establish the existence of infinitely many nodal solutions {ub,λk} with a prescribed number of nodes k, by using the Gersgorin disc's theorem, Miranda theorem and Brouwer degree theory. Moreover, we prove that the energy of {ub,λk} is strictly increasing in k, and give a convergence property of {ub,λk} as b0 and λ0.



    In this paper, we study the existence of infinitely many nodal solutions for the following Kirchhoff-Schrödinger-Poisson problem :

    {(a+bR3|u|2dx)Δu+V(|x|)u+λϕu=K(|x|)f(u),xR3,Δϕ=u2,xR3, (1.1)

    where V,K,F satisfy the following assumptions:

    (V) V:[0,+)R is measurable with V0:=essinf[0,+)V>0;

    (K) KL([0,+)) with essinf[0,+)K>0;

    (F1) fC(R;R),f(t)=f(t), and f(t)=o(t) as t0;

    (F2) there exists a constant C0>0 and p(4,6) such that for all xRN,

    |f(t)|C0(|t|+|t|p1),tR;

    (F3) lim|t||f(t)||t|3=+;

    (F4) there exists θ(0,1) such that for all xRN,t>0 and τR{0},

    K(|x|)[f(τ)τ3f(tτ)(tτ)3]sign(1t)+θV0|1t2|τ20.

    The nonlocal operator (a+bR3|u|2dx)Δu comes from the following Kirchhoff problem

    {(a+bR3|u|2dx)Δu+V(x)u=f(x,u), xRN,uH1(RN), (1.2)

    where a,b>0. Problem (1.2) is related to is related to the stationary analogue of the Kirchhoff equation

    utt(a+bΩ|u|2dx)Δu=f(x,u),

    which was introduced by Kirchhoff [20]. After the pioneer work of Lions [23], lots of interesting results to problem (1.2) or similar problems were obtained in last decade. For example, by category theory, He and Zou [13] studied the existence of multiple positive solutions for (1.2). Moreover, they also studied the concentrated behavior of positive solution. Combining constraint variational methods and the quantitative deformation lemma, Shuai [28] studied the existence and asymptotic behavior of the least energy sign-changing solution to problem (1.2). Soon afterwards, under some more weak assumptions on f (like (F1)(F4)), Tang and Cheng [33] improved and generalized some results obtained in [28]. For the investigation of stationary states of the Kirchhoff type equation and other non-local problems, we refer to [1,9,11,12,13,14,15,16,17,27,33,38,39] and reference therein.

    When a=1,b=0, system (1.1) reduces to the Schrödinger-Poisson system

    {Δu+V(|x|)u+λϕu=K(|x|)f(u),xR3,Δϕ=u2,xR3. (1.3)

    For more details about its physical meaning, we refer the reader to [4,31] and the references therein. Noting that system (1.3) involves a nonlocal term λϕ(x)u, namely, it is not a pointwise identity. In view of this, more mathematical difficulties are waiting to be solved, which makes the study of system (1.3) more meaningful. In the last decade, much effort has been made to investigate system like (1.3) on the existence of positive solutions, ground state solutions, multiple solutions and sign-changing solution; see for examples [2,6,19,24,26,29,32,35,41] and the reference therein. In particular, when K(|x|)f(u)=|u|p1u (3<p<5), Wang and Zhou [35] proved that (1.3) had a least energy sign-changing solution via a constraint variational method combining the Brouwer degree theory. After that, Chen and Tang [6] consider (1.3) under a general nonlinearity f(x,u) like (F1)(F4)). Utilizing the Non-Nehari manifold method, the authors obtained a least energy sign-changing solution.

    On the other hand, system (1.1) includes the well-known nonlinear Schrödinger equation

    {aΔu+V(|x|)u=K(|x|)f(u), xRN,uH1(RN), (1.4)

    Problem of (1.4) has been extensively studied since 1970s. One can refer to the books [31,36] and reference for the details and related results. In particular, Bartsch and Willem [3], Cao and Zhu [5] proved independently that, for every integer k0, there is a pair of solutions u±k of (1.5), having precisely k nodes. They first obtain the solution of (1.1) in each annulus, including every ball and the complement of it, and then glue them by matching the normal derivative at each junction point. After that, many authors pay their attention to find nodal solutions of elliptical partial differential equations with non-local term. When K(x)f(u)=|u|p1u (3<p<5), Kim and Seok [19] construct a sign-changing solution of Schrödinger-Poisson system (1.3) with any prescribed number nodes. More precisely, due to the nonlocal term λϕu, the method used in [3,5] could not apply to (1.3) directly. Hence, by changing (1.3) to another system and take advantage of Nehari method and gluing method in [3,5], Kim and Seok obtained infinitely many sign-changing solutions. It is worth noting that Deng, Peng and Shuai [9] considered the Kirchhoff type problem (1.2), where fC1 and satisfies the Nehari monotonicity condition

    (F5) f(t)|t|3 is increasing on (,0)(0,+).

    By using the Nehari method and gluing method as in [19], they proved (1.2) had a least energy sign-changing radial solution with any prescribed number nodes. It should be mentioned that the assumption (F4) is much weaker than (F5). As stated in [6], there are many functions satisfying (F1)(F4)), but not (F5). For example, f(t)=(|t|3+|t|3/2)t,tR. Very recently, Guo and Wu [12] considered (1.2) with assumption (F1)(F4)), by using Non-Nehari method, matrix theory and Brouwer degree theory, they get a similar conclusion as in [9]. For the reader interested in nodal solutions for the elliptical partial differential equations, we would also like to refer [7,10,18] and references therein.

    For system (1.1) contains bother nonlocal operator and nonlocal nonlinear term, the study of system (1.1) become technically complicated. In recent year, there were some scholars paying attention to system (1.1) or similar problems; see [8,21,37,40] and the reference therein. Especially, when K(x)f(u)=|u|p1u (3<p<5), Deng and Yang [8] studied the nodal solution of (1.1) by using the gluing method. But they did not study the energy property and asymptotic behavior of the nodal solution. Motivated by the above results, it is natural to ask, under condition (F1)(F4)), whether (1.1) had sign-changing solutions with any prescribed number nodes? If there exist such solution, how about its energy property and asymptotic behavior? To the best of our knowledge, this question is open for more general nonlinearities f, especially when fC0.

    Inspired by the work mentioned above, in this paper, we seek the nodal solutions to problem (1.1) under the assumption (V), (K) and (F1)(F4). Before stating our main results, we give several definitions and some notations. Throughout this paper, we define

    H:={uH1r(R3)R3(a|u|2+V(|x|)u2)dx<+},

    with the norm

    u2=R3(a|u|2+V(|x|)u2)dx.

    The embedding HH1r(R3) is continuous due to the condition (V). Then the embedding HLq(R3) is compact for 2<q<6, due to the well known result of Strauss [30]. Besides, as is well known from the Lax-Milgram theorem, for any uH1(R3), there exists a unique

    ϕu=R3u2(y)4π|xy|dyD1,2(R3)

    solving the equation Δϕ=u2. Then the energy functional associated with system (1.1) is well-defined

    Ib,λ(u)=12u2+b(R3|u|2dx)2+λ4R3×R3u2(x)u2(y)4π|xy|dxdyR3K(|x|)F(u)dx, (1.5)

    where F(u)=u0f(τ)dτ. Moreover, Ib,λC1(H,R), due to the growth condition (F1), (F2) and the Hardy-Littlewood-Sobolev inequality, see [22]HY__HY, Theorem 4.3]. Hence, for any u,φH, we have

    Ib,λ(u),φ=R3(auφ+V(|x|)uφ)dx+bR3|u|2dxR3uϕdx+λR3ϕuuφdxR3K(|x|)f(u)φdx. (1.6)

    Clearly, critical points of Ib,λ are the weak solutions for nonlocal problem (1.1). A necessary condition for uH to be a critical point of Ib,λ is that Ib,λ(u),u=0. This necessary condition defines the Nehari manifold

    N={uH{0}:Ib,λ(u),u=0},

    and the corresponding ground state energy is

    c0=infNIb,λ.

    Meanwhile, for each kN+, for each positive integer k, we denote by \boldsymbol{r}_{k} = \left(r_{1}, \ldots, r_{k}\right) \in \mathbb{R}^{k} with 0 = r_{0} < r_{1} < \cdots < r_{k} < r_{k+1} = \infty , and define a Nehari type set

    \begin{equation} \begin{aligned} \mathcal{N}_{k} = \left\{ \right. u \in H: \exists\ \boldsymbol{r}_{k}\ {\rm{such\; that}}&\ (-1)^{i+1} u_{i} > 0\ {\rm{in}}\ B_{i}, u_{i} = 0\ {\rm{on}}\ \partial B_{i}\\ &\left. {\rm{and}}\ \ I_{b,\lambda}^{\prime}(u) u_{i} = 0,\ \forall 1 \leq i \leq k+1 \right\}, \end{aligned} \end{equation} (1.7)

    where B_{1} = \left\{x \in \mathbb{R}^{N}: 0 \leq|x| < r_{1}\right\}, B_{i}: = \left\{x \in \mathbb{R}^{N}: r_{i-1} < |x| < r_{i}\right\}, u_{i} = u_{B_{i}} and \chi_{B_{i}} is the characteristic function on B_{i} . Clearly, u = \sum_{i = 1}^{k+1} u_{i} and \mathcal{N}_{k} consists of nodal functions with precisely k nodes. We consider the level

    \begin{equation} c_{k} = \inf\limits _{\mathcal{N}_{k}} I_{b,\lambda} \end{equation} (1.8)

    Our first main result of the paper can be stated as follows.

    Theorem 1.1. Under the assumptions (\bf{V}) , (\bf{K}) , (\bf{F1}) (\bf{F4}) , for every integer k \geq 0 , there exists a radial solution u_{k} of (1.1), which changes sign exactly k -times.

    The proof of Theorem 1.1 is based on the argument presented in [19] regarding the Nehari method and gluing method (see [3,5]). However, comparing to [19], the contribution of this work is greatly relax the assumptions on f and subsequently deal with the difficulties it brings. The main difficult is when the nonlinearity term f does not satisfy Nehari monotonicity condition (\bf{F5}) , one can not use the method in [19] to prove the set \mathcal{N}_{k} is nonempty. Meanwhile, due to f\notin \mathcal{C}^1 , it is impossible to prove the differentiability of the Nehari set \mathcal{N}_{k} and the Lagrangian multiplier principle dose not work. To overcome this difficulty, we use the Non-Nehari method in [6] and the Gersgorin disc's Theorem (see [12]) to prove \mathcal{N}_{k} is nonempty. Besides, taking advantage of Brouwer degree theory and Miranda theorem, we obtain the least energy sign-changing solution of (1.1) with exactly k nodes for any k\ge 0.

    Similar to [9], our another aim in the present paper is to show that the energy of u_k obtained in Theorem 1.1 is strictly increasing in k .

    Theorem 1.2. If the assumptions of Theorem 1.1 hold, then the energy of u_{k} is strictly increasing in k , i.e.,

    I_{b,\lambda}\left(u_{k+1}\right) > I_{b,\lambda}\left(u_{k}\right), \quad \mathit{{for \;any}}\; k \geq 0.

    Moreover, I_{b, \lambda}\left(u_{k}\right) > (k+1) I_{b, \lambda}\left(u_{0}\right) , where u_0 is the solution u_k with k = 0 in Theorem 1.1.

    Note that u_k obtained in Theorem 1.1 depends on b, \lambda . Thus, we denote it by u_k^{b, \lambda} and analyze the convergence properties of u_k^{b, \lambda} as b \rightarrow 0^{+} and \lambda \rightarrow 0^{+}.

    Theorem 1.3. Under the assumptions of Theorem 1.1, for any sequence \left\{b_{n}\right\} and \left\{\lambda_{n}\right\} with b_{n} \rightarrow 0^{+} and \lambda_{n} \rightarrow 0^{+} as n \rightarrow \infty , there exists a subsequence, still denoted by \left\{b_{n}\right\} and \left\{\lambda_{n}\right\} , such that the solution u_{k}^{b_n, \lambda_{n}} obtained in Theorem 1.1, converge to w_k strongly in H as n \rightarrow \infty , where w_{k} is a least energy radial solution of (1.4) which changes sign exactly k -times.

    The paper is organized as follows. In Section 2, we first present variational framework and and modify the energy functional (1.1) to the functional corresponding to a system of (k+1) -equations. Besides, we give some useful Lemmas in Matrix theory. In Section 3, we devote to find a nontrivial critical point of the modified functional. In Section 4, we obtain a solution of (1.1) with k -nodes by using critical point obtained in Section 2 as a building block, henceforth complete the proof of Theorem 1.1. Finally, in Section 4, we prove Theorems 1.2 and 1.3.

    In this section, we outline the variational framework for problem (1.1) and modify the energy functional to another functional, which corresponds to a system of (k+1) -equations. In addition, we will also give some results from matrix theory, which is important in our proofs.

    For each integer k , we define

    \Gamma_{k} = \left\{\boldsymbol{r}_{k} = \left(r_{1}, \ldots, r_{k}\right) \in \mathbb{R}^{k} \mid \ 0: = r_{0} < r_{1} < \cdots < r_{k+1}: = \infty\right\},

    and for each \boldsymbol{r}_{k} \in \Gamma_{k} , set

    \begin{array}{l} B_{1}^{\boldsymbol{r}_{k}} = \left\{x \in \mathbb{R}^{3}: 0 \leq|x| < r_{1}\right\}, \\ B_{i}^{\boldsymbol{r}_{k}} = \left\{x \in \mathbb{R}^{3}: r_{i-1} < |x| < r_{i}\right\}, \quad {\rm { for }}\; i = 2,3, \cdots, k \\ B_{k+1}^{\boldsymbol{r}_{k}} = \left\{x \in \mathbb{R}^{3}:|x| \geq r_{k}\right\}. \end{array}

    Obviously, B_{1}^{\boldsymbol{r}_{k}} is a ball, B_{2}^{\boldsymbol{r}_{k}}, \cdots, B_{k}^{\boldsymbol{r}_{k}} are annuli and B_{k+1}^{\boldsymbol{r}_{k}} is the complement of a ball. Fix an element \boldsymbol{r}_{k} = \left(r_{1}, \cdots, r_{k}\right) \in \boldsymbol{\Gamma}_{k} and thereby a family of annuli \left\{B^{\boldsymbol{r}_{k}}_{i}\right\}_{i = 1}^{k+1} , we write

    \mathcal{H}^{\boldsymbol{r}_{k}} = H_{1}^{\boldsymbol{r}_{k}} \times \cdots \times H_{k+1}^{\boldsymbol{r}_{k}},

    where

    H^{\boldsymbol{r}_{k}}_{i}: = \left\{u \in H_{0}^{1}\left(B_{i}^{\boldsymbol{r}_{k}}\right) \mid u(x) = u(|x|), u(x) = 0 \;{\rm { if }}\; x \notin B_{i}^{\boldsymbol{r}_{k}}\right\},

    for i = 1, \cdots, k+1. Apparently, H^{\boldsymbol{r}_{k}}_{i} is Hilbert space with the norm

    \|u\|_{i}^{2} = \int_{B_{i}^{\boldsymbol{r}_{k}}}\left(a|\nabla u|^{2}+V(|x|) u^{2}\right) d x .

    Hereafter, we always regard u_{i} \in H_{i}^{\boldsymbol{r}_{k}} as an element in H^{1}\left(\mathbb{R}^{3}\right) by setting u \equiv 0 in \mathbb{R}^{3} \backslash B_{i}^{\boldsymbol{r}_{k}} , and denote by

    \begin{equation} A(u_i,u_j) = \int_{ \mathbb{R}^3} |\nabla u_i|^2|\nabla u_j|^2dx \end{equation} (2.1)

    and

    \begin{equation} B(u_i,u_j) = \int_{ \mathbb{R}^3}\int_{ \mathbb{R}^3} \frac{u^2_{i}(x)u^2_{j}(y)}{4\pi|x-y|}dxdy. \end{equation} (2.2)

    After that, we introduce an auxiliary functional J_{b, \lambda}: \mathcal{H}^{\boldsymbol{r}_{k}}\rightarrow \mathbb{R} related to I_{b, \lambda}: H \rightarrow \mathbb{R} , defined by

    \begin{align*} J_{b,\lambda} & \left(u_{1}, \ldots, u_{k+1}\right): = I_{b,\lambda}\left(\sum\limits_{i = 1}^{k+1} u_{i}\right) \\ & = \sum\limits_{i = 1}^{k+1} \frac12 \left\|u_{i}\right\|_{i}^{2} +\frac{b}{4} \sum\limits_{i = 1}^{k+1} \left(\int_{B^{\boldsymbol{r}_{k}}_{i}} |\nabla u_i|^2dx\right)^2 \\ & +\frac{\lambda}{4} \sum\limits_{i = 1}^{k+1} \int_{B^{\boldsymbol{r}_{k}}_{i}}\int_{B^{\boldsymbol{r}_{k}}_{i}} \frac{u_{i}^{2}(x) u_{i}^{2}(y)}{4 \pi|x-y|}dxdy \\ & \quad +\sum\limits_{i = 1}^{k+1}\sum\limits_{j\neq i} \frac{b}{4} \int_{B_{i}^{\boldsymbol{r}_{k}}} |\nabla u_{i}|^2 dx \int_{B_{j}^{\boldsymbol{r}_{k}}} |\nabla u_{j}|^2 dx+ \\ & \sum\limits_{i = 1}^{k+1}\sum\limits_{j\neq i} \frac{\lambda}{4} \int_{B^{\boldsymbol{r}_{k}}_{j}}\int_{B^{\boldsymbol{r}_{k}}_{i}} \frac{u_{i}^{2}(x) u_{i}^{2}(y)}{4 \pi|x-y|} d x d y \\ &\quad -\sum\limits_{i = 1}^{k+1}\int_{B^{\boldsymbol{r}_{k}}_{i}}K(|x|)F(u)dx \\ & = \sum\limits_{i = 1}^{k+1} \left(\frac12 \|u_i\|_{i}^{2}+\frac{\lambda}{4}A(u_i,u_i) -\frac{1}{p}\int_{B^{\boldsymbol{r}_{k}}_{i}}K(|x|)F(u)dx \right) \\ &\quad +\frac{b}{4}\sum\limits_{i = 1}^{k+1}\sum\limits_{j\neq i} A(u_i,u_j)+\frac{\lambda}{4}\sum\limits_{i = 1}^{k+1}\sum\limits_{j\neq i} B(u_i,u_j). \end{align*}

    Therefore, for each i = 1, \ldots, k+1 ,

    \begin{equation*} \begin{aligned} \partial_{u_i}J_{b,\lambda}\left(u_{1}, \ldots, u_{k+1}\right)u_{i}& = \left\|u_{i}\right\|_{i}^{2}+b A(u_i,u_i)+\lambda B(u_i,u_i) \\ & -\int_{B^{\boldsymbol{r}_{k}}_{i}} K(|x|)F(u) dx\\ &\quad +b\sum\limits_{i = 1}^{k+1}\sum\limits_{j\neq i} A(u_i,u_j)+\lambda\sum\limits_{i = 1}^{k+1}\sum\limits_{j\neq i} B(u_i,u_j). \end{aligned} \end{equation*}

    Moreover, if \left(u_{1}, \ldots, u_{k+1}\right) \in \mathcal{H}^{\boldsymbol{r}_{k}} is a critical point of J_{b, \lambda} , then each component u_i satisfies

    \begin{equation} \left\{\begin{array}{ll} -(a+b \sum\limits_{j = 1}^{k+1}\|\nabla u_j\|_{L^2}^2) \Delta u+V(|x|) u +\lambda \left(\int_{ \mathbb{R}^3}\frac{|\sum\limits_{j = 1}^{k+1}u_j(y)|^2}{4\pi|x-y|} dy \right) u = K(|x|)f(u), & x \in B^{\boldsymbol{r}_{k}}_{i}, \\ u = 0, & x \notin B^{\boldsymbol{r}_{k}}_{i}. \end{array}\right. \end{equation} (2.3)

    Least energy radial solution of (1.1) which change signs exactly k times will constructed by gluing the solutions of the system (2.3). To this end, analogous to Nehari manifold, we define the set \mathcal{N}_{k}^{\boldsymbol{r}_{k}} by

    \begin{equation} \begin{aligned} \mathcal{N}_{k}^{\boldsymbol{r}_{k}}:& = \left\{\left(u_{1}, \cdots, u_{k+1}\right) \in \mathcal{H}^{\boldsymbol{r}_{k}} \mid u_{i} \neq 0,\ \partial_{u_{i}} J_{b,\lambda}\left(u_{1}, \cdots, u_{k+1}\right) u_{i} = 0 \;{\rm { for }}\; i = 1, \cdots, k+1\right\} \\ & = \Big\{\left(u_{1}, \cdots, u_{k+1}\right) \in \mathcal{H}^{\boldsymbol{r}_{k}} \mid u_{i} \neq 0 \;{\rm { and\; for \;each }}\; i = 1, \cdots, k+1,\\ &\qquad\qquad\left\|u_{i}\right\|_{i}^{2}+b \sum\limits_{j = 1}^{k+1} A(u_j,u_i) +\lambda \sum\limits_{j = 1}^{k+1} B(u_j,u_i) -\int_{B^{\boldsymbol{r}_{k}}_{i}} K(|x|)F(u_i)d x = 0\Big\} , \end{aligned} \end{equation} (2.4)

    and consider the least energy level

    \begin{equation} d(\boldsymbol{r}_{k}) = \inf\limits_{\left(u_{1},\cdots,u_{k+1}\right) \in \mathcal{N}_{k}^{\boldsymbol{r}_{k}} } J_{b,\lambda}\left(u_{1},\cdots,u_{k+1}\right) . \end{equation} (2.5)

    We will use constraint minimizer on \mathcal{N}_{k}^{\boldsymbol{r}_{k}} to seek a critical point of J_{b, \lambda} with nonzero component. Until then, we would like to present some properties of F(u) and results from matrix theory to ensure that \mathcal{N}_{k}^{\boldsymbol{r}_{k}} is nonempty.

    Lemma 2.1. Assume that \bf{(F1)} \bf{(F4)} holds, then for any t > 0 and \tau\in \mathbb{R} \backslash\{0\} , we have

    \begin{equation} K(|x|) F(\tau)-K(|x|) F\left(t^{1 / 4} \tau\right)+\frac{t-1}{4} K(|x|) f(\tau) \tau-\frac{\theta V_{0}\left(t^{1 / 2}-1\right)^{2}}{4} \tau^{2}\le 0, \end{equation} (2.6)

    where \theta has been given in \bf{(F4)} .

    Proof. In fact, a straightforward computation shows that

    \begin{equation} \begin{aligned} &K(|x|) F(\tau)-K(|x|) F\left(t^{1 / 4} \tau\right)+\frac{t-1}{4} K(|x|) f(\tau) \tau-\frac{\theta V_{0}\left(t^{1 / 2}-1\right)^{2}}{4} \tau^{2} \\ &\quad = \int_{t}^{1}\left(\frac{1}{4} K(|x|) f\left(s^{1 / 4} \tau\right) s^{-3 / 4} \tau-\frac{1}{4} K(|x|) f(\tau) \tau-\frac{\theta V_{0}}{4}\left(s^{-1 / 2}-1\right) \tau^{2}\right) d s\\ &\quad = \frac{\tau^{4}}{4} \int_{t}^{1}\left(\frac{K(|x|) f\left(s^{1 / 4} \tau\right)}{\left(s^{1 / 4} \tau\right)^{3}}-\frac{K(|x|) f(\tau)}{\tau^{3}}+\theta V_{0} \frac{1-s^{-1 / 2}}{\tau^{2}}\right) d s \\ &\quad = \frac{\tau^{4}}{4} \int_{t}^{1}\left(K(|x|)\left[\frac{f(\tau)}{\tau^{3}}-\frac{f\left(s^{1 / 4} \tau\right)}{\left(s^{1 / 4} \tau\right)^{3}}\right] \operatorname{sign}(1-s)+\theta V_{0} \frac{\left|1-s^{-1 / 2}\right|}{\tau^{2}}\right) \operatorname{sign}(s-1) ds, \end{aligned} \end{equation} (2.7)

    which combines with \bf{(F4)} ensure (2.6).

    As a byproduct of the last lemma is the following corollary.

    Corollary 2.1. Assume that \bf{(F1)} \bf{(F4)} hold, then for any \tau\in \mathbb{R} , we have

    K(|x|) F(\tau)-\frac{1}{4} K(|x|) f(\tau) \tau-\frac{\theta V_{0}}{4} \tau^{2} \leq 0.

    Proof. In fact, the corollary follows by letting t\rightarrow 0 in (2.6).

    Lemma 2.2. (Gersgorin disc Theorem [34,Therome 1.1]) For any matrix A = \left(a_{i, j}\right) \in \mathbb{C}^{n \times n} and any eigenvalue \lambda \in \sigma(A): = \{\mu \in \mathbb{C}: \operatorname{det}(\mu E-A) = 0\} , there is a positive integer l \in\{1, \ldots, n\} such that

    \left|\lambda-a_{l, l}\right| \leq \sum\limits_{j \neq l}\left|a_{l, j}\right| .

    By virtue of this lemma, we have the following results.

    Lemma 2.3. [12] For any b_{i j} = b_{j i} > 0 with i \neq j \in\{1, \ldots, n\} and s_{i} > 0 with i = 1, \ldots, n , define the matrix C: = \left(c_{i j}\right)_{n \times n} b y

    c_{i j} = \left\{\begin{array}{ll} -\sum\limits_{l \neq i} s_{l} b_{i l} / s_{i} & \;{{for}}\; i = j, \\ b_{i j} > 0 & \;{{for}}\; i \neq j. \end{array}\right.

    Then the real symmetric matrix \left(c_{i j}\right)_{n \times n} is non-positive definite.

    Lemma 2.4. Defining map G_{1}:(\mathbb{R}_{\geq 0})^{k+1} \rightarrow \mathbb{R} and G_{2}:(\mathbb{R}_{\geq 0})^{k+1} \rightarrow \mathbb{R} by

    \begin{equation} G_{1}(t_1,\cdots,t_{k+1}) = \sum\limits_{i, j = 1}^{k+1}\left(t_{i}^{1 / 2} t_{j}^{1 / 2}-t_{i}\right) A(u_i,u_j), \end{equation} (2.8)

    and

    \begin{equation} G_{2}(t_1,\cdots,t_{k+1}) = \sum\limits_{i, j = 1}^{k+1}\left(t_{i}^{1 / 2} t_{j}^{1 / 2}-t_{i}\right) B(u_i,u_j), \end{equation} (2.9)

    where A(u_i, u_j) and B(u_i, u_j) are defined in (2.1) and (2.2). Then the map G_1 and G_2 are strictly concave in (\mathbb{R}_{ > 0})^{k+1} . Moreover, for (t_1, \cdots, t_{k+1})\in (\mathbb{R}_{\geq 0})^{k+1} , G_{m}(t_1, \cdots, t_{k+1})\le G_{m}(1, \cdots, 1) = 0 , where m \in \{1, 2\}.

    Proof. We just need to check this result for G_{1} , and then the proof of G_{2} is similar, we omit it. Indeed, direct computations give that

    \frac{\partial G_{1}}{\partial t_i} = \sum\limits_{\substack{l = 1 \\ l \neq i}}^{k+1} \left(\frac{1}{2}t^{-\frac12}_i t^{\frac12}_{l}-1 \right) A(u_i,u_l),
    \frac{\partial^2 G_{1}}{\partial t_{i}^{2}} = \sum\limits_{\substack{l = 1 \\ l \neq i}}^{k+1} \left( -\frac14t^{-\frac32}_i t^{\frac12}_{l}-1 \right)A(u_i,u_l) \quad {\rm{and}}\quad \frac{\partial^2 G_{1}}{\partial t_{i}t_{j}} = \frac{1}{4}t^{-\frac12}_{i} t_{j}^{-\frac12}A(u_i,u_j)\ {\rm{for}}\ i\neq j.

    Defining the matrix \left(a_{i, j}\right) _{(k+1)\times(k+1)} by a_{i, j} = \frac{1}{4}t^{-\frac12}_{i} t_{j}^{-\frac12}A(u_i, u_j), then

    \left(\frac{\partial^2 G_{1}}{\partial t_{i}t_{j}}\right)_{(k+1) \times(k+1)} = \left\{\begin{array}{l} - \sum\limits_{l \neq i} t_{l} a_{i l} / t_{i}\ {\rm{for}}\ i = j ,\\ a_{i j} > 0\ {\rm{for}}\ i \neq j , \end{array}\right.

    which is a negative definite matrix in \left(\mathbb{R}_{ > 0}\right)^{k+1} by Lemma 2.3. Therefore, G_{1} is a strictly concave function in (\mathbb{R}_{ > 0})^{k+1} . Moreover, notice that \nabla G_{1}(1, \cdots, 1) = 0 , we deduce that (1, \cdots, 1) is a unique global maximum point of G_{1} on \left(\mathbb{R}_{ > 0}\right)^{k+1} and

    \max _{\left(\mathbb{R}_{ > 0}\right)^{k+1}} G_{1}(t_1,\cdots,t_{k+1}) = G_{1}(1, \cdots, 1) = 0.

    Consequently, by the continuity of G_{1} ,

    \max _{\left(\mathbb{R}_{\geq 0}\right)^{k+1}} G_{1}(t_1,\cdots,t_{k+1}) = \max _{\left(\mathbb{R}_{ > 0}\right)^{k+1}} G_{1}(t_1,\cdots,t_{k+1}) = g(1, \cdots, 1) = 0.

    As a result, we complete the proof of this lemma.

    In this section, we seek a critical point of the auxiliary functional J_{b, \lambda} . Firstly, we prove the following lemma.

    Lemma 3.1. Assume that (\bf{V}), (\bf{K}), (\bf{F1})-(\bf{F4}) hold and \boldsymbol{r}_{k} \in \Gamma_{k} is fixed. Then for any (u_1, \cdots, u_{k+1}) \in \mathcal{H}^{\boldsymbol{r}_{k}} \backslash\{\boldsymbol{0}\} and \left(t_1, \cdots, t_{k+1} \right) \in \left(\mathbb{R}_{\geq 0}\right)^{k+1}: = [0, \infty)^{k+1} , there holds

    \begin{equation} \begin{aligned} &J_{b,\lambda}\left(t^{\frac14}_{1}u_1, \cdots, t^{\frac14}_{k+1}u_{k+1} \right)-J_{b,\lambda}(u_1,\cdots,u_{k+1})\\ &\qquad\leq \sum\limits_{i = 1}^{k+1}\left[\frac{t_{i}-1}{4} \partial_{u_{i}} J_{b,\lambda}(u_1,\cdots,u_{k+1}) u_{i}-(1-\theta) \frac{\left(t_{i}^{1 / 2}-1\right)^{2}}{4}\left\|u_{i}\right\|_{i}^{2}\right], \end{aligned} \end{equation} (3.1)

    where \theta \in(0, 1) is defined in (\bf{F4}) .

    Proof. For each \left(u_{1}, \cdots, u_{k+1}\right) \in \mathcal{H}^{\boldsymbol{r}_{k}} with u_{i} \neq 0 for i = 1, \cdots, k+1 , we have

    \begin{equation} \begin{aligned} &J_{b,\lambda}\left(t^{\frac14}_{1}u_1, \cdots, t^{\frac14}_{k+1}u_{k+1} \right)-J_{b,\lambda}(u_1,\cdots,u_{k+1})\\ = &\sum\limits_{i = 1}^{k+1}\left(\frac12 t_{i}^{\frac12}\|u_i\|^{2}_{i} +\frac{b}{4}t_{i}^{\frac12} \sum\limits_{j = 1}^{k+1} t_{j}^{\frac12} A(u_j,u_i) \right. \\ & \left.+\frac{\lambda}{4}t_{i}^{\frac12} \sum\limits_{j = 1}^{k+1} t_{j}^{\frac12} B(u_j,u_i) -\int_{B_{i}^{\boldsymbol{r}_{k}}}K(|x|)F(t^{\frac14}_{i}u_i) \right)\\ &-\sum\limits_{i = 1}^{k+1}\left(\frac12 \|u_i\|^{2}_{i} +\frac{b}{4} \sum\limits_{j = 1}^{k+1} A(u_j,u_i) \right. \\ & \left. +\frac{\lambda}{4} \sum\limits_{j = 1}^{k+1} B(u_j,u_i) -\int_{B_{i}^{\boldsymbol{r}_{k}}}K(|x|)F(u_i) \right)\\ = &\sum\limits_{i = 1}^{k+1}\Bigg[ \frac12(t^{\frac12}_{i}-1)\|u_i\|^2_{i} \\ & +\frac{b}{4}\sum\limits_{j = 1}^{k+1}\left(t_{j}^{\frac12} t_{i}^{\frac12}-1 \right) A(u_j,u_i) +\frac{\lambda}{4}\sum\limits_{j = 1}^{k+1}\left(t_{j}^{\frac12} t_{i}^{\frac12}-1 \right) B(u_j,u_i) \\ &\qquad+\int_{B_{i}^{\boldsymbol{r}_{k}}} k(|x|)\left[F(t^{\frac14}_{i}u_i)-F(u_i)\right] dx \Bigg]\\ = &\sum\limits_{i = 1}^{k+1}\frac{t_i-1}{4} \partial_{u_{i}} J_{b,\lambda}(u_1,\cdots,u_{k+1}) u_{i} -\sum\limits_{i = 1}^{k+1}\frac{(t_{i}^{\frac12}-1)^2}{4} \|u_i\|^{2}_{i}\\ &+\frac{b}{4}\sum\limits_{i,j = 1}^{k+1}(t_{j}^{\frac12}t_{i}^{\frac12}-t_{i})A(u_j,u_i) +\frac{\lambda}{4}\sum\limits_{i,j = 1}^{k+1}(t_{j}^{\frac12} t_{i}^{\frac12}-t_{i})B(u_j,u_i)\\ &+\sum\limits_{i = 1}^{k+1} \int_{B_{i}^{\boldsymbol{r}_{k}}} \left( K(|x|) F\left(u_{i}\right)-K(|x|) F\left(t_{i}^{1 / 4} u_{i}\right)+\frac{t_{i}-1}{4} K(|x|) f\left(u_{i}\right) u_{i}\right) dx \\ \leq &\sum\limits_{i = 1}^{k+1}\left[\frac{t_i-1}{4} \partial_{u_{i}} J_{b,\lambda}(u_1,\cdots,u_{k+1}) u_{i} -(1-\theta)\frac{(t_{i}^{\frac12}-1)^2}{4} \|u_i\|^{2}_{i}\right]\\ &+\frac{b}{4}G_{1}(t_1,\cdots,t_{k+1}) +\frac{\lambda}{4}G_{2}(t_1,\cdots,t_{k+1})\\ &+\sum\limits_{i = 1}^{k+1} \int_{B_{i}^{\boldsymbol{r}_{k}}} \left( K(|x|) \left(F\left(u_{i}\right)- F\left(t_{i}^{1 / 4} u_{i}\right)\right)+\frac{t_{i}-1}{4} \right. \\ & \left.K(|x|) f\left(u_{i}\right) u_{i} -\theta V_{0} \frac{(t_{i}^{\frac12}-1)^2}{4} |u_{i}|^2 \right) dx. \end{aligned} \end{equation} (3.2)

    Now, using Lemmas 2.1 and 2.4, we can easily get the conclusion of Lemma 3.1.

    From now on, we always assume that (\bf{V}) , (\bf{K}) , (\bf{F1}) (\bf{F4}) hold. We are devoted to finding the critical points of J_{b, \lambda} . Firstly, for (u_1, \cdots, u_{k+1}) \in \mathcal{H}^{\boldsymbol{r}_{k}} , we define a map \Psi^{{\bf{u}}}:(\mathbb{R}_{\geq 0})^{k+1} \rightarrow \mathbb{R} by

    \begin{equation} \Psi ^{u}(t_1,\cdots,t_{k+1}): = J_{b,\lambda}\left(t_{1}^{\frac{1}{4}}u_1,\cdots,t_{k+1}^{\frac{1}{4}}u_{k+1}\right) . \end{equation} (3.3)

    With the help of Lemma 3.1, we are devoted to prove that the set \mathcal{N}_{k}^{\boldsymbol{r}_{k}} , defined in (2.4), is nonempty in the following Lemma.

    Lemma 3.2. Assume \boldsymbol{r}_{k} \in \Gamma_{k} is fixed. Then for any (u_1, \cdots, u_{k+1}) \in \mathcal{H}^{\boldsymbol{r}_{k}} \backslash\{\boldsymbol{0}\} , there exists a unique (k+1)-tuple (\bar{t}_1, \cdots, \bar{t}_{k+1}) of positive numbers such that

    \begin{equation} \Psi ^{{\bf{u}}}(\bar{t}_1,\cdots, \bar{t}_{k+1}) = \max\limits_{(\mathbb{R}_{\geq 0})^{k+1}}\Psi^{{\bf{u}}}(t_1,\cdots,t_{k+1}), \end{equation} (3.4)

    and

    \begin{equation} \left((\bar{t}_{1})^{\frac{1}{4}}u_1,\cdots,(\bar{t}_{k+1})^{\frac{1}{4}}u_{k+1}\right)\in \mathcal{N}_{k}^{\boldsymbol{r}_{k}} . \end{equation} (3.5)

    Proof. By the definition of \Psi^{{\bf{u}}} and the assumption (\bf{V}) , (\bf{K}) , (\bf{F1}) (\bf{F4}) , we deduce that \Psi^{{\bf{u}}}\left(t_{1}, \ldots, t_{k+1}\right) \rightarrow 0 as |\left(t_{1}, \ldots, t_{k+1}\right)|\rightarrow 0 and \Psi^{{\bf{u}}}\left(t_{1}, \ldots, t_{k+1}\right) \rightarrow -\infty as |\left(t_{1}, \ldots, t_{k+1}\right)|\rightarrow \infty , which yields \Psi^{{\bf{u}}} possess at least one global maximal point \left(\bar{t}_{1}, \ldots, \bar{t}_{k+1}\right) .

    Firstly, we prove that \bar{t}_{i} > 0 for all i = 1, \ldots, k+1 . Otherwise, there exists i_{0} \in\{1, \ldots, k+1\} such that \bar{t}_{i} = 0 . Without loss of generality, we assume \bar{t}_{1} = 0 . Notice that

    \begin{align} \Psi^{{\bf{u}}}\left(\tau,\bar{t}_{2}, \ldots, \bar{t}_{k+1}\right) & = \frac12 \tau^{\frac{1}{2}}\|u_1\|^{2}_1 \\ & +\frac{b}{4} \tau A(u_1,u_1)+\frac{\lambda}{4} \tau B(u_1,u_1)-\int_{B_{1}^{\boldsymbol{r}_{k}}} K(|x|)F(\tau^{\frac14}u_1)dx\\ &+\frac{b}{2} \sum\limits_{j = 2}^{k+1} \tau^{\frac{1}{2}}(\bar{t_j})^{\frac{1}{2}}A(u_1,u_j) +\frac{\lambda}{2} \sum\limits_{j = 2}^{k+1} \tau^{\frac{1}{2}}(\bar{t_j})^{\frac{1}{2}}B(u_1,u_j)\\ &+ \sum\limits_{i = 2}^{k+1}\left[\frac{(\bar{t}_i)^{\frac{1}{2}}}{2} \|u_i\|^{2}_i+ \frac{b}{4} \bar{t}_{i}A(u_i,u_i)+ \frac{\lambda}{4} \bar{t}_{i}B(u_i,u_i)\right. \\ &\left.-\int_{B_{i}^{\boldsymbol{r}_{k}}} K(|x|)F((\bar{t}_{i})^{\frac14}u_i)dx \right]\\ &\quad +\frac{b}{4}\sum\limits_{i = 2}^{k+1}\sum\limits_{j\neq i} (\bar{t_i})^{\frac{1}{2}}(\bar{t_j})^{\frac{1}{2}} A(u_i,u_j) +\frac{\lambda}{4}\sum\limits_{i = 2}^{k+1}\sum\limits_{j\neq i} (\bar{t_i})^{\frac{1}{2}}(\bar{t_j})^{\frac{1}{2}} B(u_i,u_j) \end{align} (3.6)

    is increasing with respect to \tau > 0 for \tau small enough, which leads to a contradiction. Therefore, \bar{t}_{i} > 0 for all i = 1, \ldots, k+1 .

    Since \left(\bar{t}_{1}, \ldots, \bar{t}_{k+1}\right) is a global maximum point of \Psi^u , namely,

    \begin{equation*} \frac{\partial \Psi^{{\bf{u}}}}{\partial t_i}\left(\bar{t}_{1},\bar{t}_{2}, \ldots, \bar{t}_{k+1}\right) = \frac14 \partial_{u_i} J_{b,\lambda}\left((\bar{t}_{1})^{\frac{1}{4}}u_1,\cdots,(\bar{t}_{k+1})^{\frac{1}{4}}u_{k+1}\right) (\bar{t}_{i})^{-\frac{3}{4}}u_{i} = 0, \end{equation*}

    which implies (3.5).

    Thus, our results will be proved if we show the global maximum point of \Psi is unique. Indeed, by Lemma 3.1, if (u_1, \cdots, u_{k+1})\in \mathcal{N}_{k}^{\boldsymbol{r}_{k}} , one has

    \begin{equation} J_{b,\lambda}(u_1,\cdots,u_{k+1})\ge J_{b,\lambda}(t_{1}^{\frac14}u_1,\cdots,t_{k+1}^{\frac14}u_{k+1})+ (1-\theta) \sum\limits_{i = 1}^{k+1}\frac{\left(t_{i}^{1 / 2}-1\right)^{2}}{4}\left\|u_{i}\right\|_{i}^{2}. \end{equation} (3.7)

    Suppose on the contrary that there exists another maximum point (s_1, \cdots, s_{k+1}) of \Psi^{{\bf{u}}} and (s_1, \cdots, s_{k+1})\neq(1, \cdots, 1), Choosing t_i = \frac{1}{s_i} for all 1\le i\le k+1 and by virtue of (2.1), we have

    \begin{equation*} \begin{aligned} J_{b,\lambda}(u_1,\cdots,u_{k+1})& > J_{b,\lambda}(s^{\frac14}_{1}u_1,\cdots,s^{\frac14}_{k+1}u_{k+1})\\ & > J_{b,\lambda}(t^{\frac14}_1 s^{\frac14}_1 u_1,\cdots,t^{\frac14}_{k+1}s^{\frac14}_{k+1}u_{k+1}) = J_{b,\lambda}(u_1,\cdots,&u_{k+1}), \end{aligned} \end{equation*}

    which is impossible. This completes the proof of Lemma 3.2.

    Recall the definition of d(\boldsymbol{r}_{k}) in (2.5), and then we have the following result.

    Lemma 3.3. For each \boldsymbol{r}_{k} \in \Gamma_{k} , there is a minimizer \left({u}^{\boldsymbol{r}_{k}}_{1}, \cdots, {u}^{\boldsymbol{r}_{k}}_{k+1}\right) \in \mathcal{N}_{k}^{\boldsymbol{r}_{k}} with (-1)^{i+1} u_{i}^{\boldsymbol{r}_k} > 0 in B_{i}^{r_{k}} for i = 1, \cdots, k+1 satisfying

    J_{b,\lambda}\left({u}^{\boldsymbol{r}_{k}}_{1}, \cdots, {u}^{\boldsymbol{r}_{k}}_{k+1}\right) = d \left(\boldsymbol{r}_{k}\right).

    Proof. For \left(u_{1}, \cdots, u_{k+1}\right) \in \mathcal{N}_{k}^{\boldsymbol{r}_{k}} , denoted by u = \sum\limits_{i = 1}^{k+1} u_i . From Corollary 2.1, we deduce that

    \begin{align} & J_{b,\lambda}\left(u_{1}, \cdots, u_{k+1}\right) = I_{b,\lambda}\left( \sum\limits_{i = 1}^{k+1} u_i \right) = I_{b,\lambda}(u)-\frac{1}{4}I^{\prime}_{\lambda}(u)u \\ & \qquad\quad = \frac{1}{4}\|u\|^{2}+ \int_{ \mathbb{R}^3}\left(\frac{1}{4}K(|x|f(u)u-K(|x|)F(u) \right) dx \\ & \qquad\quad\ge \frac{1}{4}\|u\|^{2}-\frac{\theta V_0}{4}\int_{ \mathbb{R}^3} u^2 dx \geq \frac{1-\theta}{4} \|u\|^{2}. \end{align} (3.8)

    Suppose that \left\{\left(u_{1}^{n}, \cdots, u_{k+1}^{n}\right)\right\}_{n = 1}^{\infty} \subset \mathcal{N}_{k}^{\boldsymbol{r}_{k}} is a minimizing sequence of \left.J_{b, \lambda}\right|_{\mathcal{N}_{k}} , whose existence and boundedness are guaranteed by (3.8). Then, up to a subsequence, which we still denote by \left(u_{1}^{n}, \cdots, u_{k+1}^{n}\right) , it converges to an element \left(u_{1}^{0}, \cdots, u_{k+1}^{0}\right) weakly in \mathcal{H}^{\boldsymbol{r}_{k}} .

    From now on the proof will be divided into several steps.

    Step 1. We proof that u_{i}^{0}\neq 0 for all i . In fact, we have two cases: Either u^{n}_{i} convergence to u^{0}_{i} strongly in \mathcal{H}_{i}^{\boldsymbol{r}_{k}} or it converges to u^{0}_{i} weakly but not strongly in \mathcal{H}_{i}^{\boldsymbol{r}_{k}} . In the former case, since \left(u^{n}_{1}, \cdots, u^{n}_{k+1}\right) \in \mathcal{N}_{k}^{\boldsymbol{r}_{k}} , then by (\bf{K}), (\bf{F1}), (\bf{F2}) and the Sobolev embedding theorem, it follows that

    \begin{equation} \begin{aligned} \left\|u^{n}_{i}\right\|_{i}^{2} + b \sum\limits_{j = 1}^{k+1} A(u^{n}_j,u^{n}_i) + \lambda \sum\limits_{j = 1}^{k+1} B(u^{n}_j,u^{n}_i) & = \int_{\mathbb{R}^{3}} K(|x|) f\left(u^{n}_{i}\right) u^{n}_{i}dx\\ &\leq\left(\varepsilon \int_{\mathbb{R}^{3}}\left|u^{n}_{i}\right|^{2}dx+ C(\varepsilon) \int_{\mathbb{R}^{N3}}\left|u^{n}_{i}\right|^{p}dx\right) \\ & \leq \frac{1}{2}\left\|u^{n}_{i}\right\|_{i}^{2}+C\left\|u^{n}_{i}\right\|_{i}^{p}, \end{aligned} \end{equation} (3.9)

    which implies \|u^{n}_{i}\|_{i} \ge C_{0} . Consequently, u^{0}_{i} is not zero. In the latter case, we have \|u^{0}_{i}\|_{i} < \liminf_{n\rightarrow \infty} \|u^{n}_{i}\|_{i} . Besides, due to the compactly embedding \mathcal{H}_{i}^{\boldsymbol{r}_{k}} \hookrightarrow L^q for 2 < q < 6 , it follows by Strauss's compactness Lemma [30] that for all i = 1, \cdots, k+1 ,

    \begin{equation} \int_{\mathbb{R}^{3}} K(|x|) F\left(u_{i}^{n}\right)dx \rightarrow \int_{\mathbb{R}^{3}} K(|x|) F\left(u_{i}^{0}\right)dx \end{equation} (3.10)

    and

    \begin{equation} \int_{\mathbb{R}^{3}} K(|x|) f\left(u_{i}^{n}\right) u_{i}^{n}dx \rightarrow \int_{\mathbb{R}^{3}} K(|x|) f\left(u_{i}^{0}\right) u_{i}^{0} dx. \end{equation} (3.11)

    Therefore, we deduce that

    \begin{equation} \left\|u^{0}_{i}\right\|_{i}^{2} < \int_{ \mathbb{R}^3} K(|x|)f(u^{0}_{i})u^{0}_{i} dx\le \frac{1}{2}\left\|u^{0}_{i}\right\|_{i}^{2} +C\left\|u^{0}_{i}\right\|_{i}^{p}, \end{equation} (3.12)

    and we also have u^{0}_{i} is not zero.

    Step 2. We claim that \left(u_{1}^{n}, \ldots, u_{k+1}^{n}\right) \rightarrow \left(u_{1}^{0}, \ldots, u_{k+1}^{0}\right) strongly in \mathcal{H}^{\boldsymbol{r}_{k}}. By contradiction, if this is not the case then, there exists i\in \{1, \cdots, k+1\} such that \|u_i^{0}\|_{i} < \liminf_{n\rightarrow \infty} \|u_i^{n}\|_{i} . Namely, \left(u_{1}^{0}, \ldots, u_{k+1}^{0}\right)\notin \mathcal{N}_{k}^{\boldsymbol{r}_{k}} . Since u_i\neq 0 for all i = 1, \cdots, k+1 , by virtue of Lemma 3.2, we deduce that there exists \left(t_{1}^{0}, \ldots, t_{k+1}^{0}\right) \neq (1, \ldots, 1) satisfying \left((t_{1}^{0})^{\frac14} u_{1}^{0}, \ldots, (t_{k+1}^{0})^{\frac14} u_{k+1}^{0}\right) \in \mathcal{N}_{k}^{\boldsymbol{r}_{k}} and

    \begin{equation*} \begin{aligned} d(\boldsymbol{r}_{k})& = \inf _{\left(u_{1}, \cdots, u_{k+1}\right) \in \mathcal{N}_{k}^{\boldsymbol{r}_{k}} } J_{b,\lambda}\left(u_{1}, \cdots, u_{k+1}\right) \\ &\leq J_{b,\lambda}\left((t_{1}^{0})^{\frac14} u_{1}^{0}, \ldots, (t_{k+1}^{0})^{\frac14} u_{k+1}^{0}\right) \\ & = \frac{1}{2} \sum\limits_{i = 1}^{k+1}\left(t_{i}^{0}\right)^{\frac12} \left\|u_{i}^{0}\right\|_{i}^{2} +\frac{b}{4}\sum\limits_{i = 1}^{k+1}\sum\limits_{j = 1}^{k+1}(t^{0}_i)^{\frac{1}{2}}(t^{0}_j)^{\frac{1}{2}} A(u_i,u_j) \\ & +\frac{\lambda}{4}\sum\limits_{i = 1}^{k+1}\sum\limits_{j = 1}^{k+1}(t^{0}_i)^{\frac{1}{2}}(t^{0}_j)^{\frac{1}{2}} B(u_i,u_j)\\ &\quad-\sum\limits_{i = 1}^{k+1} \int_{B_{i}^{\boldsymbol{r}_{k}}}K(|x|)F((t^{0}_i)^{\frac{1}{4}}u_{i}^{0}) dx \\ & < \frac{1}{2} \sum\limits_{i = 1}^{k+1}\left(\left(t_{i}^{0}\right)^{\frac12} \liminf _{n \rightarrow \infty}\left\|u_{i}^{n}\right\|_{i}^{2}\right) +\frac{b}{4}\sum\limits_{i = 1}^{k+1}\sum\limits_{j = 1}^{k+1}(t^{0}_i)^{\frac{1}{2}}(t^{0}_j)^{\frac{1}{2}} A(u_{i}^{n},u_{j}^{n})\\ &\quad +\frac{\lambda}{4}\sum\limits_{i = 1}^{k+1}\sum\limits_{j = 1}^{k+1}(t^{0}_i)^{\frac{1}{2}}(t^{0}_j)^{\frac{1}{2}} B(u_{i}^{n},u_{j}^{n}) -\sum\limits_{i = 1}^{k+1} \int_{B_{i}^{\boldsymbol{r}_{k}}}K(|x|)F((t^{0}_i)^{\frac{1}{4}}u_{i}^{n}) dx \\ &\le \liminf _{n \rightarrow \infty} J_{b,\lambda}\left(u_{1}^{n}, \cdots, u_{k+1}^{n}\right) \\ & = d(\boldsymbol{r}_{k}) , \end{aligned} \end{equation*}

    which is impossible. Thus, \left(u_{1}^{0}, \cdots, u_{k+1}^{0}\right) is contained in \mathcal{N}_{k}^{\boldsymbol{r}_{k}} and a minimizer of \left.J_{b, \lambda}\right|_{ \mathcal{N}_{k}^{\boldsymbol{r}_{k}} } .

    Step 3. Letting

    \left(u^{\boldsymbol{r}_{k}}_{1}, \cdots, u^{\boldsymbol{r}_{k}}_{k+1}\right): = \left(\left|u_{1}^{0}\right|,-\left|u_{2}^{0}\right|, \cdots,(-1)^{k+2}\left|u_{k+1}^{0}\right|\right),

    we can easily check that \left(u^{\boldsymbol{r}_{k}}_{1}, \cdots, u^{\boldsymbol{r}_{k}}_{k+1}\right) is also a minimizer of \left.J_{b, \lambda}\right|_{ \mathcal{N}_{k}^{\boldsymbol{r}_{k}} } . Using the strong maximum principle, we get (-1)^{i+1} u_{i}^{\boldsymbol{r}_{k}} > 0 in B_{i}^{\boldsymbol{r}_{k}} for all i = 1, \ldots, k+1 . The proof is completed.

    In the following, we will prove the minimizer of \left.J_{b, \lambda}\right|_{ \mathcal{N}_{k}^{\boldsymbol{r}_{k}} } , which obtained by the previous Lemma, is actually a critical point of Functional J_{b, \lambda} . However, we cannot use the Nehari method in [19] directly. More precisely, since the assumption (\bf{F1}) , it seems difficult to prove \mathcal{N}_{k}^{\boldsymbol{r}_{k}} is a manifold as in [19]. Thus, we use the deformation lemma and Brouwer degree theory to achieve this. This idea comes from [12,33].

    In what follows, for convenience, we denote by

    {\bf{u}} = \left(u_{1}, \cdots, u_{k+1}\right) \in \mathcal{H}^{\boldsymbol{r}_{k}}, \ {\bf{t}} = \left(t_{1}, \cdots, t_{k+1}\right) \in \mathbb{R}^{k+1}, \ {\bf{t}}^{1 / 4} \bullet {\bf{u}} = \left(t_{1}^{1 / 4} u_{1}, \cdots, t_{k+1}^{1 / 4} u_{k+1}\right).

    Lemma 3.4. If \boldsymbol{r}_{k} \in \Gamma and \left({u}^{\boldsymbol{r}_{k}}_{1}, \cdots, {u}^{\boldsymbol{r}_{k}}_{k+1}\right) \in \mathcal{N}_{k}^{\boldsymbol{r}_{k}} is a minimizer of J_{b, \lambda}|_{\mathcal{N}_{k}} such that

    J_{b,\lambda} \left({u}^{\boldsymbol{r}_{k} }_{1}, \cdots, {u}^{\boldsymbol{r}_{k} }_{k+1}\right) = d \left(\boldsymbol{r}_{k}\right) ,

    then \left({u}^{\boldsymbol{r}_{k}}_{1}, \cdots, {u}^{\boldsymbol{r}_{k}}_{k+1}\right) is a critical point of J_{b, \lambda} in \mathcal{H}_{k}^{r_{k}} .

    Proof. Suppose on contrary, if {\bf{u}}^{\boldsymbol{r}_{k}} = \left({u}^{\boldsymbol{r}_{k}}_{1}, \cdots, {u}^{\boldsymbol{r}_{k}} _{k+1}\right) is not a critical point of J_{b, \lambda} , then there exist \delta > 0 and \varrho > 0 such that

    {\bf{u}} \in \mathcal{H}^{\boldsymbol{r}_{k}} ,\left\|{\bf{u}}- {\bf{u}}^{\boldsymbol{r}_{k}}\right\| \leq 3 \delta \Rightarrow \left\|(\partial_{u_1}J_{b,\lambda}({\bf{u}}),\cdots,\partial_{u_{k+1}}J_{b,\lambda}({\bf{u}}))\right\| \geq \varrho.

    Let D = \left\{{\bf{t}} \in \mathbb{R}_{\geq 0}^{k+1}:\left|t_{i}^{1 / 4}-1\right| < (\frac12)^{1 / 4}, \ \forall i = 1, \cdots, k+1\right\} . Then D is an open neighborhood of \bf{1}: = (1, \cdots, 1) \in \mathbb{R}^{k+1} . It follows from Lemma 3.2 that

    \begin{equation} \kappa: = \max _{\partial D} J_{b,\lambda}\left({\bf{t}}^{1 / 4} \bullet {\bf{u}}^{\boldsymbol{r}_{k}} \right) < d\left(\boldsymbol{r}_{k}\right). \end{equation} (3.13)

    For \varepsilon: = \min \left\{\left(m_{b}-\kappa\right) / 3, 1, \varrho \delta / 8\right\}, S: = B\left({\bf{u}}^{\boldsymbol{r}_{k}}, \delta\right) , [36,Lemma 2.3] yields a deformation \eta = (\eta_1, \cdots, \eta_{k+1}) \in C\left([0, 1] \times \mathcal{H}^{\boldsymbol{r}_{k}}, \mathcal{H}^{\boldsymbol{r}_{k}} \right) such that

    (ⅰ) \eta (1, {\bf{u}}) = {\bf{u}} if {\bf{u}} \notin J^{-1}_{b, \lambda}\left(\left[d(\boldsymbol{r}_{k})-2 \varepsilon, d(\boldsymbol{r}_{k})+2 \varepsilon\right]\right) \cap S_{2 \delta} ;

    (ⅱ) \eta\left(1, J_{b, \lambda}^{d(\boldsymbol{r}_{k})+\varepsilon} \cap S \right) \subset J_{b, \lambda}^{d(\boldsymbol{r}_{k})-\varepsilon} ;

    (ⅲ) J_{b, \lambda}( \eta (1, {\bf{u}})) \leq J_{b, \lambda}({\bf{u}}), \forall {\bf{u}} \in \mathcal{H}^{\boldsymbol{r}_{k}} .

    By Lemma 3.2,

    \begin{equation*} J_{b,\lambda}\left({\bf{t}}^{1 / 4} \bullet {\bf{u}}^{\boldsymbol{r}_{k}} \right) \leq J_{b,\lambda}\left({\bf{u}}^{\boldsymbol{r}_{k}} \right) = d(\boldsymbol{r}_{k}). \end{equation*}

    Then it follows from (ⅱ) that

    \begin{equation} J_{b,\lambda}\left( \eta \left(1, {\bf{t}}^{1 / 4} \bullet {\bf{u}}^{\boldsymbol{r}_{k}} \right) \right) < d(\boldsymbol{r}_{k})-\varepsilon,\quad \forall {\bf{t}}\in (\mathbb{R}_{\geq 0})^{k+1},\ \sum\limits_{i = 1}^{k+1}|t^{\frac14}_i-1| < \delta /\|{\bf{u}}^{\boldsymbol{r}_{k}}\|. \end{equation} (3.14)

    On the other hand, by (ⅲ) and (2.1), one has

    \begin{equation} J_{b,\lambda}\left( \eta \left(1, {\bf{t}}^{1 / 4} \bullet {\bf{u}}^{\boldsymbol{r}_{k}} \right) \right)\leq J_{b,\lambda}\left({\bf{t}}^{1 / 4} \bullet {\bf{u}}^{\boldsymbol{r}_{k}} \right) < d(\boldsymbol{r}_{k}),\quad \forall {\bf{t}}\in (\mathbb{R}_{\geq 0})^{k+1},\ \sum\limits_{i = 1}^{k+1}|t^{\frac14}_i-1|\geq \delta /\|{\bf{u}}^{\boldsymbol{r}_{k}}\|. \end{equation} (3.15)

    Combining (3.14) with (3.15), we deduce that

    \begin{equation} \max\limits_{{\bf{t}} \in D} J_{b,\lambda}\left( \eta \left(1, {\bf{t}}^{1 / 4} \bullet {\bf{u}}^{\boldsymbol{r}_{k}} \right) \right) < d(\boldsymbol{r}_{k}). \end{equation} (3.16)

    Now we claim that

    \begin{equation} \left\{\eta\left(1, {\bf{t}}^{1 / 4} \bullet {\bf{u}}^{\boldsymbol{r}_{k}} \right): {\bf{t}} \in D\right\} \cap \mathcal{N}_{k}^{\boldsymbol{r}_{k}} \neq \emptyset. \end{equation} (3.17)

    In fact, we denote {\bf{\Phi}} = \left(\Phi_{1}, \cdots, \Phi_{k+1}\right): D \rightarrow \mathbb{R}^{k+1} by

    \begin{equation*} \Phi_{i}({\bf{t}}): = \partial_{u_{i}} E\left(\eta\left(1, {\bf{t}}^{1 / 4} \bullet {\bf{u}}^{\boldsymbol{r}_{k}}\right)\right) \eta_{i}\left(1, {\bf{t}}^{1 / 4} \bullet {\bf{u}}^{\boldsymbol{r}_{k}}\right), \quad {\rm{for}} \ i = 1, \cdots, k+1. \end{equation*}

    Then for all {\bf{t}}\in \partial D , the properties of deformation lemma implies \eta\left(1, {\bf{t}}^{1 / 4} \bullet {\bf{u}}^{\boldsymbol{r}_{k}}\right) = {\bf{t}}^{1 / 4} \bullet {\bf{u}}^{\boldsymbol{r}_{k}} and \Phi_{i}({\bf{t}}): = \partial_{u_{i}} E\left({\bf{t}}^{1 / 4} \bullet {\bf{u}}^{\boldsymbol{r}_{k}}\right) t_{i}^{1 / 4} u^{\boldsymbol{r}_{k} }_{i} for all i = 1, \cdots, k+1 . On the other hand, by virtue of Lemma 3.2 and Brouwer degree theory, it follows that

    \begin{equation*} \begin{aligned} \deg\left(\left(\frac{\partial \Psi^{{\bf{u}}^{\boldsymbol{r}_{k}}}}{\partial t_1 },\cdots,\frac{\partial \Psi^{{\bf{u}}^{\boldsymbol{r}_{k}}}}{\partial t_{k+1} }\right), D,{\bf{0}}\right) = 1, \end{aligned} \end{equation*}

    where \Psi^{{\bf{u}}^{\boldsymbol{r}_{k}}} has been defined in (3.3). Therefore, for all {\bf{t}} \in \partial D we have

    \begin{equation*} \begin{aligned} 1 = &\deg\left(\left(\frac{\partial \Psi^{{\bf{u}}^{\boldsymbol{r}_{k}}}}{\partial t_1 },\cdots,\frac{\partial \Psi^{{\bf{u}}^{\boldsymbol{r}_{k}}}}{\partial t_{k+1} }\right), D,{\bf{0}}\right)\\ & = \deg\left(\left( \partial_{u_{1}} J_{b,\lambda} \left( u^{\boldsymbol{r}_{k} }\right) \frac{1}{4} t_{1}^{-3 / 4} u_{1},\cdots,\partial_{u_{k+1}} J_{b,\lambda} \left( u^{\boldsymbol{r}_{k} }\right) \frac{1}{4} t_{k+1}^{-3 / 4} u_{k+1} \right),D,{\bf{0}}\right)\\ & = \deg\left(\left(\frac{1}{4t_{1}}\Phi_{i}({\bf{t}}),\cdots,\frac{1}{4t_{k+1}}\Phi_{k+1}({\bf{t}})\right),D,0\right), \\ \end{aligned} \end{equation*}

    which shows the correctness of the (3.17). Consequently, we deduce that

    \begin{equation} \sup\limits_{{\bf{t}} \in D} J_{b,\lambda}\left( \eta \left(1, {\bf{t}}^{1 / 4} \bullet {\bf{u}}^{\boldsymbol{r}_{k}} \right) \right) \geq d(\boldsymbol{r}_{k}), \end{equation} (3.18)

    which leads to a contradiction with (3.16). Hence, {\bf{u}}^{\boldsymbol{r}_{k}} = \left({u}^{\boldsymbol{r}_{k}}_{1}, \cdots, {u}^{\boldsymbol{r}_{k}}_{k+1}\right) is a critical point of J_{b, \lambda} and the proof is finished.

    Recall the infimum level d\left(\boldsymbol{r}_{k}\right) defined in (2.5), then we have the following results.

    Lemma 4.1. For any p \in(4, 6) and \boldsymbol{r}_{k} = \left(r_{1}, \ldots, r_{k}\right) \in \Gamma_{k}

    (i) if r_{i}-r_{i-1} \rightarrow 0 for some i \in\{1, \ldots, k\} , then d\left(\boldsymbol{r}_{k}\right) \rightarrow +\infty ;

    (ii) if r_{k} \rightarrow \infty , then d\left(\boldsymbol{r}_{k}\right) \rightarrow +\infty ;

    (iii) d is continuous in \Gamma_{k}. As a consequence, there exists a \overline{\boldsymbol{r}}_{k} \in \Gamma_{k} such that

    d\left(\overline{\boldsymbol{r}}_{k}\right) = \inf\limits _{\boldsymbol{r}_{k} \in \Gamma_{k}} d\left(\boldsymbol{r}_{k}\right).

    Proof. (ⅰ) In view of Lemma 3.3, combining the Hölder inequality and Sobolev inequality, we have

    \begin{equation*} \begin{aligned} \left\|u_{i}^{\boldsymbol{r}_{k}}\right\|_{i}^{2} & = \int_{\mathbb{R}^{3}} K(|x|) f\left(u^{{\boldsymbol{r}_{k}}}_{i}\right) u^{{\boldsymbol{r}_{k}}}_{i}dx-\lambda\sum\limits_{i, j = 1}^{k+1} \int_{\mathbb{R}^{3}}\int_{B_{i}^{\boldsymbol{r}_{k}}} \frac{|u_{i}^{\boldsymbol{r}_{k}}(y)|^2|u_{j}^{\boldsymbol{r}_{k}}(x)|^2}{4\pi|x-y|}dxdy\\ &-b\int_{\mathbb{R}^{3}} \sum\limits_{j = 1}^{k+1}\left|\nabla u_{j}^{\boldsymbol{r}_{k}}(x)\right|^{2} dx \int_{B_{i}^{\boldsymbol{r}_{k}}}\left|\nabla u_{i}^{\boldsymbol{r}_{k}}\right|^{2} d x\\ &\leq\varepsilon \int_{\mathbb{R}^{3}}\left|u^{\boldsymbol{r}_{k}}_{i}\right|^{2}dx+ C(\varepsilon) \int_{\mathbb{R}^{3}}\left|u^{\boldsymbol{r}_{k}}_{i}\right|^{p}dx \\ &\leq C\varepsilon\left\|u_{i}^{\boldsymbol{r}_{k}}\right\|_{i}^{2}+CC(\varepsilon)\left\|u_{i}^{\boldsymbol{r}_{k}}\right\|_{i}^{p}\left|B_{i}^{\boldsymbol{r}_{k}}\right|^{\frac{6-p}{6}}. \end{aligned} \end{equation*}

    If r_{i}-r_{i-1} \rightarrow 0 , then \left\|\sum_{j = 1}^{k+1} u_{j}^{\boldsymbol{r}_{k}}(x)\right\| \rightarrow +\infty due to p \in(4, 6) . Thus,

    \begin{equation} d\left(\boldsymbol{r}_{k}\right) = J_{b,\lambda}\left(u_{1}^{\boldsymbol{r}_{k}}, \ldots, u_{k+1}^{\boldsymbol{r}_{k}}\right) = I_{b,\lambda}\left(\sum\limits_{j = 1}^{k+1} u_{j}^{\boldsymbol{r}_{k}}\right) \geq\frac{1-\theta}{4}\left\|\sum\limits_{i = 1}^{k+1} u_{i}^{\boldsymbol{r}_{k}}\right\|^{2} \rightarrow +\infty, \end{equation} (4.1)

    then, (ⅰ) holds.

    (ⅱ) Recalling the Strauss inequality [26], we can find a_{0} > 0 such that, for all radial function u \in H_{V} ,

    u(x) \leq \frac{a_{0}\|u\|}{|x|},

    for a.e. |x| > 1 . so we have

    \begin{equation} \begin{aligned} \left\|u_{k+1}^{\boldsymbol{r}_{k}}\right\|_{k+1}^{2} &\leqslant \int_{\mathbb{R}^{3}} K(|x|) f\left(u^{\boldsymbol{r}_{k}}_{i}\right) u^{\boldsymbol{r}_{k}}_{i}dx\\ &\leq\varepsilon \int_{\mathbb{R}^{3}}\left|u^{\boldsymbol{r}_{k}}_{i}\right|^{2}dx+ C(\varepsilon) \int_{\mathbb{R}^{3}}\left|u^{\boldsymbol{r}_{k}}_{i}\right|^{p}dx \\ &\leq C\varepsilon \left\|u_{k+1}^{\boldsymbol{r}_{k}}\right\|_{k+1}^{2}+CC(\varepsilon)\|u_{k+1}^{\boldsymbol{r}_{k}}\|^{p-2}_{k+1}\int_{B_{k+1}^{\boldsymbol{r}_{k}}}|x|^{2-p}|u_{k+1}^{\boldsymbol{r}_{k}}|^2 dx\\ &\leq C\varepsilon \left\|u_{k+1}^{\boldsymbol{r}_{k}}\right\|_{k+1}^{2}+CC(\varepsilon) |r_{k}|^{2-p}\|u_{k+1}^{\boldsymbol{r}_{k}}\|^{p-2}_{k+1}\int_{B_{k+1}^{\boldsymbol{r}_{k}}}|u_{k+1}^{\boldsymbol{r}_{k}}|^2 dx\\ &\leq C\varepsilon \left\|u_{k+1}^{\boldsymbol{r}_{k}}\right\|_{k+1}^{2}+CC(\varepsilon) |r_{k}|^{2-p}\|u_{k+1}^{\boldsymbol{r}_{k}}\|^{p}_{k+1} \end{aligned} \end{equation} (4.2)

    which yields that \left\|\sum_{j = 1}^{k+1} u_{j}^{\boldsymbol{r}_{k}}(x)\right\| \rightarrow +\infty as r_{k} \rightarrow \infty due to p\in(4, 6). Therefore,

    \begin{equation*} d\left(\boldsymbol{r}_{k}\right) \geq\frac{1-\theta}{4}\left\|\sum\limits_{i = 1}^{k+1} u_{i}^{\boldsymbol{r}_{k}}\right\|^{2} \rightarrow +\infty \quad {\rm { as }}\ r_{k} \rightarrow \infty, \end{equation*}

    which implies (ⅱ).

    (ⅲ) Take a sequence \left\{\boldsymbol{\boldsymbol{r}}_{k}^{n}\right\}_{n = 1}^{\infty} satisfying \boldsymbol{\boldsymbol{r}}_{k}^{n} \rightarrow \overline{\boldsymbol{\boldsymbol{r}}}_{k} \in \Gamma_{k} . We will prove the conclusion by showing d\left(\overline{\boldsymbol{\boldsymbol{r}}}_{k}\right) \geq \limsup _{n \rightarrow \infty} d\left(\boldsymbol{\boldsymbol{r}}_{k}^{n}\right), d\left(\overline{\boldsymbol{\boldsymbol{r}}}_{k}\right) \leq \liminf _{n \rightarrow \infty} d\left(\boldsymbol{\boldsymbol{r}}_{k}^{n}\right) .

    First, we prove that d\left(\overline{\boldsymbol{\boldsymbol{r}}}_{k}\right) \geq \limsup _{n \rightarrow \infty} d\left(\boldsymbol{r}_{k}^{n}\right) . In order to emphasize that v_{i}^{\boldsymbol{r}_{k}^{n}} is radial in B_{i}^{\boldsymbol{r}_{k}^{n}} , we will rewrite v_{i}^{\boldsymbol{r}_{k}^{n}}(|x|) = v_{i}^{\boldsymbol{r}_{k}^{n}}(r) . Define v_{i}^{\boldsymbol{r}_{k}^{n}}:\left[r_{i-1}^{n}, r_{i}^{n}\right] \rightarrow \mathbb{R} by

    \begin{equation*} v_{i}^{\boldsymbol{r}_{k}^{n}}(r) = \begin{cases}t_{i}^{n} u_{i}^{\bar{\boldsymbol{r}}_{k}} \left(\bar{r}_{i-1}+\frac{\bar{r}_{i}-\bar{r}_{i-1}^{n}}{r_{i}^{n}-r_{i-1}^{n}}\left(r-r_{i-1}^{n}\right)\right), & i = 1, \ldots, k, \\ t_{k+1}^{n} u_{k+1}^{\bar{\boldsymbol{r}}_{k}}\left(\frac{\bar{r}_{k}}{r_{k}^{n}} r\right), & i = k+1,\end{cases} \end{equation*}

    where \left(u_{1}^{\boldsymbol{r}_{k}^{n}}, \ldots, u_{k+1}^{\boldsymbol{r}_{k}^{n}}\right) and \left(u_{1}^{\bar{\boldsymbol{r}}_{k}}, \ldots, u_{k+1}^{\bar{\boldsymbol{r}}_{k}}\right) are minimizers of \left.J_{b, \lambda}\right|_{\mathcal{N}_{k}^{\boldsymbol{r}_{k}^{n}}} and \left.J_{b, \lambda}\right|_{\mathcal{N}_{k}^{\bar{\boldsymbol{r}} k}} respectively. \left(t_{1}^{n}, \ldots, t_{k+1}^{n}\right) is the unique (k+1) tuple of positive numbers such that \left(v_{1}^{\boldsymbol{r}_{k}^{n}}, \ldots, v_{k+1}^{\boldsymbol{r}_{k}^{n}}\right) \in \mathcal{N}_{k}^{\boldsymbol{r}_{k}^{n}} . By the definition of \left(u_{1}^{\boldsymbol{r}_{k}^{n}}, \ldots, u_{k+1}^{\boldsymbol{r}_{k}^{n}}\right) , we know that

    \begin{equation} J_{b,\lambda}(v_{1}^{\boldsymbol{r}_{k}^{n}}, \ldots, v_{k+1}^{\boldsymbol{r}_{k}^{n}})\geq J_{b,\lambda}(u_{1}^{\boldsymbol{r}_{k}^{n}}, \ldots, u_{k+1}^{\boldsymbol{r}_{k}^{n}}) = d(\boldsymbol{r}_{k}^{n}). \end{equation} (4.3)

    Since \boldsymbol{\boldsymbol{r}}_{k}^{n} \rightarrow \overline{\boldsymbol{\boldsymbol{r}}}_{k} \in \Gamma_{k} , by a straightforward computation, one can easily get the following equations,

    \begin{equation*} \begin{aligned} &\int_{B_{i}^{\boldsymbol{r}_{k}^{n}}}\left|v_{i}^{\boldsymbol{r}_{k}^{n}}\right|^{2}dx = \left(t_{i}^{n}\right)^{2} \int_{B_{i}^{\bar{\boldsymbol{r}}_{k}}}\left|u_{i}^{\bar{\boldsymbol{r}}_{k}}\right|^{2} d x+o_{n}(1); \\ &\left\|v_{i}^{\boldsymbol{r}_{k}^{n}}\right\|_{i}^{2} = \left(t_{i}^{n}\right)^{2}\left\|u_{i}^{\bar{\boldsymbol{r}}_{k}}\right\|_{i}^{2}+o_{n}(1);\\ &\int_{B_{i}^{\boldsymbol{r}_{k}^{n}}} f(v_{i}^{\boldsymbol{r}_{k}^{n}})v_{i}^{\boldsymbol{r}_{k}^{n}} dx = \int_{B_{i}^{\bar{\boldsymbol{r}}_{k}}} f(t_{i}^{n} u_{i}^{\bar{\boldsymbol{r}}_{k}}) t_{i}^{n} u_{i}^{\bar{\boldsymbol{r}}_{k}} dx +o_{n}(1);\\ &\int_{B_{i}^{{\boldsymbol{r}}_{k}^{n}}}\int_{B_{i}^{{\boldsymbol{r}}_{k}^{n}}} \frac{v_{i}^{\boldsymbol{r}_{k}^{n}}(x) v_{i}^{\boldsymbol{r}_{k}^{n}}(y)}{4 \pi|x-y|}dxdy = \left(t_{i}^{n}\right)^{4} \int_{B_{i}^{\bar{\boldsymbol{r}}_{k}}}\int_{B_{i}^{\bar{\boldsymbol{r}}_{k}}} \frac{u_{i}^{\bar{\boldsymbol{r}}_{k}}(x) u_{i}^{\bar{\boldsymbol{r}}_{k}}(y)}{4 \pi|x-y|}dxdy+o_{n}(1) \end{aligned} \end{equation*}

    and

    \begin{equation*} \begin{aligned} &\int_{B_{i}^{\boldsymbol{r}_{k}^{n}}}\left|\nabla v_{i}^{\boldsymbol{r}_{k}^{n}}(x)\right|^{2}d x \int_{B_{j}^{\boldsymbol{r}_{k}^{n}}}\left|\nabla v_{j}^{\boldsymbol{r}_{k}^{n}}(x)\right|^{2} d x \\ &\quad = \left(t_{i}^{n}\right)^{2}\left(t_{j}^{n}\right)^{2} \int_{B_{i}^{\bar{\boldsymbol{r}}_{k}}}\left|\nabla u_{i}^{\bar{\boldsymbol{r}}_{k}}(x)\right|^{2} d x \int_{B_{i}^{\bar{\boldsymbol{r}}_{k}}}\left|\nabla u_{j}^{\bar{\boldsymbol{r}}_{k}}(x)\right|^{2} d x+o_{n}(1). \end{aligned} \end{equation*}

    According to \left(v_{1}^{\boldsymbol{r}_{k}^{n}}, \ldots, v_{k+1}^{\boldsymbol{r}_{k}^{n}}\right) \in \mathcal{N}_{k}^{\boldsymbol{r}_{k}^{n}} and \left(u_{1}^{\bar{\boldsymbol{r}}_{k}}, \ldots, u_{k+1}^{\bar{\boldsymbol{r}}_{k}}\right)\in \mathcal{N}_{k}^{\bar{\boldsymbol{r}}_{k}^{n}} , there holds that

    \begin{equation*} \begin{aligned} \left\|u_{i}^{\bar{\boldsymbol{r}}_{k}}\right\|_{i}^{2}&+b \sum\limits_{j = 1}^{k+1} \int_{B_{i}^{\bar{\boldsymbol{r}}_{k}}}\left|\nabla u_{i}^{\bar{\boldsymbol{r}}_{k}}(x)\right|^{2} d x \int_{B_{j}^{\bar{\boldsymbol{r}}_{k}}}\left|\nabla u_{j}^{\bar{\boldsymbol{r}}_{k}}(x)\right|^{2} d x \\ +& \lambda\sum\limits_{j = 1}^{k+1} \int_{B_{i}^{\bar{\boldsymbol{r}}_{k}}}\int_{B_{i}^{\bar{\boldsymbol{r}}_{k}}} \frac{u_{i}^{\bar{\boldsymbol{r}}_{k}}(x) u_{i}^{\bar{\boldsymbol{r}}_{k}}(y)}{4 \pi|x-y|}dxdy -\int_{B_{i}^{\bar{\boldsymbol{r}}_{k}}} K(|x|) f\left(u^{\bar{\boldsymbol{r}}_{k}}_{i}\right) u^{\bar{\boldsymbol{r}}_{k}}_{i} d x = 0, \end{aligned} \end{equation*}

    and

    \begin{equation*} \begin{aligned} &\left (t_{i}^{n})^{2}\|u_{i}^{\bar{\boldsymbol{r}}_{k}}\right\|_{i}^{2}+b(t_{i}^{n})^{2}(t_{j}^{n})^{2} \sum\limits_{j = 1}^{k+1} \int_{B_{i}^{\bar{\boldsymbol{r}}_{k}}}\left|\nabla u_{i}^{\bar{\boldsymbol{r}}_{k}}(x)\right|^{2} d x \int_{B_{j}^{\bar{\boldsymbol{r}}_{k}}}\left|\nabla u_{j}^{\bar{\boldsymbol{r}}_{k}}(x)\right|^{2} d x \\ &\quad+ \lambda(t_{i}^{n})^{4}\sum\limits_{j = 1}^{k+1} \int_{B_{i}^{\bar{\boldsymbol{r}}_{k}}}\int_{B_{i}^{\bar{\boldsymbol{r}}_{k}}} \frac{u_{i}^{\bar{\boldsymbol{r}}_{k}}(x) u_{i}^{\bar{\boldsymbol{r}}_{k}}(y)}{4 \pi|x-y|}dxdy-\int_{B_{i}^{\bar{\boldsymbol{r}}_{k}}} K(|x|) f\left(t_{i}^{n}u^{\bar{\boldsymbol{r}}_{k}}_{i}\right) t_{i}^{n}u^{\bar{\boldsymbol{r}}_{k}}_{i} dx = o_{n}(1). \end{aligned} \end{equation*}

    This combined with Lemma 3.2 we have \lim _{n \rightarrow \infty} t_{i}^{n} = 1 for all i . Hence, from (4.3) we can see that

    \begin{equation} d\left(\overline{\boldsymbol{\boldsymbol{r}}}_{k}\right) = J_{b,\lambda}\left(u_{1}^{\bar{\boldsymbol{r}}_{k}}, \ldots, u_{k+1}^{\overline{\boldsymbol{\boldsymbol{r}}}_{k}}\right) = \limsup\limits _{n \rightarrow \infty} J_{b,\lambda}\left(v_{1}^{\boldsymbol{\boldsymbol{r}}_{k}^{n}}, \ldots, v_{k+1}^{\boldsymbol{\bf{\boldsymbol{r}}}_{k}^{n}}\right) \geq \limsup\limits _{n \rightarrow \infty} d\left(\boldsymbol{\boldsymbol{r}}_{k}^{n}\right). \end{equation} (4.4)

    Next, we prove that d\left(\overline{\boldsymbol{\boldsymbol{r}}}_{k}\right) \leq \liminf _{n \rightarrow \infty} d\left(\boldsymbol{\boldsymbol{r}}_{k}^{n}\right) . By the same argument as former case, let w_{i}^{r_{k}^{n}} = \left[\bar{r}_{i-1}, \bar{r}_{i}\right] \rightarrow \mathbb{R} be defined by

    \begin{equation} w_{i}^{\boldsymbol{r}_{k}^{n}}(r) = \begin{cases}s_{i}^{n} u_{i}^{\boldsymbol{r}_{k}^{n}}\left(r_{i-1}^{n}+\frac{r_{i}^{n}-r_{i-1}^{n}}{\bar{r}_{i}-\bar{r}_{i-1}}\left(r-\bar{r}_{i-1}\right)\right), & {\rm { if }} \quad i = 1, \ldots, k, \\ s_{k+1}^{n} u_{k+1}^{\boldsymbol{r}_{k}^{n}}\left(\frac{r_{k}^{n}}{\bar{r}_{k}} r\right), & {\rm { if }}\; i = k+1,\end{cases} \end{equation} (4.5)

    where \left(s_{1}^{n}, \ldots, s_{k+1}^{n}\right) \in\left(\mathbb{R}_{+}\right)^{k+1} such that \left(w_{1}^{\boldsymbol{r}_{k}^{n}}, \ldots, w_{k+1}^{\boldsymbol{r}_{k}^{n}}\right) \in \mathcal{N}_{k}^{\bar{\boldsymbol{r}}_{k}} . By the same arguments, we can deduce s_{i}^{n} \rightarrow 1 as n \rightarrow \infty for all i = 1, \ldots, k+1 . Hence,

    \begin{equation*} \begin{aligned} d\left(\overline{\boldsymbol{\boldsymbol{r}}}_{k}\right) & = J_{b,\lambda}\left(u_{1}^{\bar{\boldsymbol{r}}_{k}}, \ldots, u_{k+1}^{\bar{\boldsymbol{r}}_{k}}\right) \\ & \leq \liminf _{n \rightarrow \infty} J_{b,\lambda}\left(w_{1}^{\boldsymbol{r}_{k}^{n}}, \ldots, w_{k+1}^{\boldsymbol{r}_{k}^{n}}\right) = \liminf _{n \rightarrow \infty} J_{b,\lambda}\left(u_{1}^{\boldsymbol{r}_{k}^{n}}, \ldots, u_{k+1}^{\boldsymbol{r}_{k}^{n}}\right) = \liminf _{n \rightarrow \infty} d\left(\boldsymbol{r}_{k}^{n}\right). \end{aligned} \end{equation*}

    This combined with (4.4) yields that d is continuous in \Gamma_{k} . Furthermore, taking account into (i), (ii) , we know that there is a \overline{\boldsymbol{\boldsymbol{r}}}_{k} \in \Gamma_{k} such that d\left(\overline{\boldsymbol{\boldsymbol{r}}}_{k}\right) = \inf _{\boldsymbol{\boldsymbol{r}}_{k} \in \Gamma_{k}} d\left(\boldsymbol{\boldsymbol{r}}_{k}\right) . Hence, (iii) holds.

    Proof of Theorem 1.1. By Lemmas 3.3 and 4.1, we deduce that there exists \left(\bar{\boldsymbol{r}}_{k}\right)\in \Gamma_{k} and {\bf{u}}^{\bar{\boldsymbol{r}}_{k}} = \left({u}_{1}^{\bar{\boldsymbol{r}}_{k}}, \cdots, {u}^{\bar{\boldsymbol{r}}_{k}} _{k+1}\right)\in \mathcal{N}_{k}^{\bar{\boldsymbol{r}}_{k}} with (-1)^{i+1} {u}_{i}^{\bar{\boldsymbol{r}}_{k}} > 0 in B_{i}^{\bar{\boldsymbol{r}}_{k}} such that

    \begin{equation*} J_{b,\lambda}\left(u_{1}^{\overline{\boldsymbol{r}}_{k}}, \ldots, u_{k+1}^{\overline{\boldsymbol{r}}_{k}}\right) = d\left(\overline{\boldsymbol{r}}_{k}\right) = \inf _{\boldsymbol{r}_{k} \in \Gamma_{k}} d\left(\boldsymbol{r}_{k}\right), \end{equation*}

    which implies

    \begin{equation*} c_{k} = d\left(\overline{\boldsymbol{r}}_{k}\right) = I_{b,\lambda}\left(\sum\limits_{i = 1}^{k+1} u_{i}^{\overline{\boldsymbol{r}}_{k}}\right), \end{equation*}

    where c_{k} has been defined in (1.8). Now, we claim that u_{k} = \sum_{i = 1}^{k+1} u_{i}^{\overline{\boldsymbol{r}}_{k}} is solution of (1.1). Suppose to the contrary that the claim does not hold, then by density argument, there exists a radial function \psi \in \mathcal{C}_{0}^{\infty}(\mathbb{R}^3) such that

    \begin{equation*} I^{\prime}_{b,\lambda}(u_k)\psi = -2. \end{equation*}

    We denote a function g\in \mathcal{C}\left(\mathbb{R}^{k+1}\times \mathbb{R}; H\right) by

    \begin{equation*} g({\bf{t}},\varepsilon): = \sum\limits_{i = 1}^{k+1} t_i u_{i}^{\overline{\boldsymbol{r}}_{k}}+\varepsilon \psi. \end{equation*}

    Since \sum_{i = 1}^{k+1} u_{i}^{\overline{\boldsymbol{r}}_{k}} is continuous and has exactly k nodes, take into account that g is also continuous, we deduce that there exists \tau > 0 small enough, such that g({\bf{t}}, \varepsilon) also changes sign k times and

    \begin{equation} I^{\prime}_{b,\lambda}(g({\bf{t}},\varepsilon))\psi < -1, \quad \forall ({\bf{t}},\varepsilon))\in D_{\tau}\times [0,\tau], \end{equation} (4.6)

    where D_{\tau}: = \left\{{\bf{t}} = (t_1, \cdots, t_{k+1}) \in \mathbb{R}^{k+1}:|t_i-1| < \tau\ {\rm{for\; all}}\ 1\le i \le k+1 \right\}.

    Now we choose \eta \in \mathcal{C}^{\infty}(\mathbb{R}^3) , 0 \leq \eta \leq 1 with \eta({\bf{t}}) = 1 if {\bf{t}} \in \overline{D_{\frac{\tau}{4}}} and \eta({\bf{t}}) = 0 if {\bf{t}} \notin D_{\frac{\tau}{2}} . Furthermore, we define another continuous function \tilde{g}: \mathbb{R}^{k+1} \rightarrow H by

    \begin{equation*} \tilde{g}({\bf{t}}): = \sum\limits_{i = 1}^{k+1} t_i u_{i}^{\overline{\boldsymbol{r}}_{k}}+\tau \eta({\bf{t}}) \psi. \end{equation*}

    Similarly, for all {\bf{t}}\in D_{\tau} , \tilde{g}({\bf{t}}) also changes sign k times and has k nodes 0 < r_1({\bf{t}}) < \cdots < r_{k}({\bf{t}}) < +\infty .

    We assert that

    \begin{equation} \exists \ \bar{{\bf{t}}}\in D_{\tau}\ {\rm{such\; that}}\ \tilde{g}(\bar{{\bf{t}}}) \in \mathcal{N}_{k}. \end{equation} (4.7)

    If the assertion holds, then

    \begin{equation} I_{b,\lambda}(\tilde{g}(\bar{{\bf{t}}})) \geq c_k. \end{equation} (4.8)

    However, if \bar{{\bf{t}}} \in \overline{D_{\frac{\tau}{2}}} , then \eta(\bar{{\bf{t}}}) > 0 . This combines with (4.1) implies that

    \begin{equation} \begin{aligned} I_{b,\lambda}(\tilde{g}(\overline{{\bf{t}}})) & = I_{b,\lambda}\left(\sum\limits_{i = 1}^{k+1} \bar{t}_{i} u_{i}^{\overline{\boldsymbol{r}}_{k}}\right)+\int_{0}^{1}\langle I_{b,\lambda}^{\prime}\left(\sum\limits_{i = 1}^{k+1} \bar{t}_{i} u_{i}^{\overline{\boldsymbol{r}}_{k}}+\mu \tau \eta(\overline{{\bf{t}}}) \psi\right), \tau \eta(\overline{{\bf{t}}}) \psi\rangle d \mu \\ & \leq I_{b,\lambda}\left(\sum\limits_{i = 1}^{k+1} \bar{t}_{i} u_{i}^{\overline{\boldsymbol{r}}_{k}}\right)-\tau \eta(\overline{{\bf{t}}}) \\ & < c_k. \end{aligned} \end{equation} (4.9)

    On the other hand, if \bar{{\bf{t}}} \notin \overline{D_{\frac{\tau}{2}}} , then \eta(\bar{{\bf{t}}}) = 0 , it follows from Lemma 3.2 that

    \begin{equation} I_{b,\lambda}(\tilde{g}(\overline{{\bf{t}}})) = I_{b,\lambda}\left(\sum\limits_{i = 1}^{k+1} (\bar{t}_{i}) u_{i}^{\overline{\mathrm{r}}_{k}}\right) < I_{b,\lambda}\left(\sum\limits_{i = 1}^{k+1} u_{i}^{\overline{\mathrm{r}}_{k}}\right) = c_{k} . \end{equation} (4.10)

    Consequently, (4.9) and (4.10) lead to a contradiction with (4.8). Therefore, it is enough to prove (4.7). Indeed, we denote by \Omega\left(\sigma_{i}({\bf{t}}), \sigma_{i+1}({\bf{t}})\right) = \left\{x \in \mathbb{R}^{N}:|x| \in\left(\sigma_{i}({\bf{t}}), \sigma_{i+1}({\bf{t}})\right)\right\} and define a map H: \overline{D_{\tau}} \rightarrow \mathbb{R}^{k+1} by

    {\bf{H}}({\bf{t}}) = \left(H_{1}({\bf{t}}), \cdots, H_{k+1}({\bf{t}})\right) \ {\rm { with \;components }} \ H_{i}(\bf{s}): = \langle J_{b,\lambda}^{\prime}(\bar{g}({\bf{t}})),\left.\bar{g}({\bf{t}})\right|_{\Omega\left(\sigma_{i}({\bf{t}}), \sigma_{i+1}({\bf{t}})\right)}\rangle .

    Clearly H \in C\left(\overline{D_{\tau}}, \mathbb{R}^{k+1}\right) and for {\bf{t}} \in \partial D_{\tau} , we have

    \begin{equation} \begin{aligned} H_{i}({\bf{t}}) = J_{b,\lambda}^{\prime}\left(\sum\limits_{j = 1}^{k+1} t_{j} u_{i}^{\overline{\boldsymbol{r}}_{k}}\right) t_{i} u_{i}^{\overline{\boldsymbol{r}}_{k}} & = t_{i}^{2}\left\|u_{i}^{\overline{\boldsymbol{r}}_{k}}\right\|_{j}^{2} +b t_{i}^{2}\sum\limits_{j = 1}^{k+1} t_{j}^{2} A(u_{i}^{\overline{\boldsymbol{r}}_{k}},u_{j}^{\overline{\boldsymbol{r}}_{k}}) +\lambda t_{i}^{2}\sum\limits_{j = 1}^{k+1} t_{j}^{2} B(u_{i}^{\overline{\boldsymbol{r}}_{k}},u_{j}^{\overline{\boldsymbol{r}}_{k}})\\ &-\int_{\mathbb{R}^{N}} K(|x|) f\left(t_{i} u_{i}^{\overline{\boldsymbol{r}}_{k}}\right) t_{i} u_{i}^{\overline{\boldsymbol{r}}_{k}}. \end{aligned} \end{equation} (4.11)

    Since for small \tau > 0 ,

    H_{i}(1-\tau, \cdots, 1-\tau) > 0 \;{\rm { and }}\; H_{i}(1+\tau, \cdots, 1+\tau) < 0.

    Then by a straightforward computations, we deduce that

    \begin{array}{lc} H_{i}\left(t_{1}, \cdots, t_{i-1}, 1-\tau, t_{i+1}, \cdots, t_{k+1}\right) > 0 & {\rm { for \;all }}\; t_{j} \in(1-\tau, 1+\tau), \ j \neq i, \\ H_{i}\left(t_{1}, \cdots, t_{i-1}, 1+\tau, t_{i+1}, \cdots, t_{k+1}\right) < 0 & {\rm { for\; all }}\; t_{j} \in(1-\tau, 1+\tau), \ j \neq i. \end{array}

    Therefore, by the Miranda theorem [25], there exists \overline{{\bf{t}}} \in D_{\tau} such that {\bf{H}}(\overline{{\bf{t}}}) = 0 , which implies that \tilde{g}(\overline{{\bf{t}}}) \in \mathcal{N}_{k} . Thus, the assertion (4.7) is confirmed and the proof of Theorem 1.1 is completed.

    For any integer k \geq 0 , by Theorem 1.1, we know that the problem (1.1) has a radial solution u_{k} which changes sign exactly k -times. Now, we are ready to prove the energy comparison property.

    Proof of Theorem 1.2. By Theorem 1.1, there exists {\overline{\boldsymbol{r}}_{k+1}}\in \Gamma_{k+1} and u_{k+1} = \sum\limits_{i = 1}^{k+1}u_{i}^{\overline{\boldsymbol{r}}_{k+1}} such that u_{k+1} is the solution of (1.1) with k+1 nodes. Moreover, (u_{1}^{\overline{\boldsymbol{r}}_{k+1}}, \cdots, u_{k+1}^{\overline{\boldsymbol{r}}_{k+1}}, u_{k+2}^{\overline{\boldsymbol{r}}_{k+1}}) \in \mathcal{N}_{k+1}^{\overline{\boldsymbol{r}}_{k+1}} and I_{\lambda}(u_{k+1}) = d(\overline{\boldsymbol{r}}_{k+1}) = \inf_{\boldsymbol{r}_{k+1}\in \Gamma_{k+1}}d(\boldsymbol{r}_{k+1}). Meanwhile, Lemma 3.2 implies that there exists (a_1, \cdots, a_{k+1}) such that

    \begin{equation} (a_1u_{1}^{\overline{\boldsymbol{r}}_{k+1}},\cdots,a_{k+1}u_{k+1}^{\overline{\boldsymbol{r}}_{k+1}})\in \mathcal{N}_{k}^{\overline{\boldsymbol{r}}_{k}}. \end{equation} (5.1)

    Thus, using (5.1) and Lemma 3.2 we see that

    \begin{equation*} \begin{aligned} &I_{b,\lambda}(u_k)\le I_{b,\lambda}\left(\sum\limits_{i = 1}^{k+1} a_i u_{i}^{\overline{r}_{k+1}} \right) = J_{b,\lambda} \left( a_1 u_{1}^{\overline{r}_{k+1}},\cdots,a_{k+1} u_{k+1}^{\overline{r}_{k+1}} \right)\\ \quad& = J_{b,\lambda} \left( a_1 u_{1}^{\overline{r}_{k+1}},\cdots,a_{k+1} u_{k+1}^{\overline{r}_{k+1}}, 0\right) < J_{b,\lambda} \left( u_{1}^{\overline{r}_{k+1}},\cdots, u_{k+1}^{\overline{r}_{k+1}}, u_{k+2}^{\overline{r}_{k+2}} \right) = I_{b,\lambda}( u_{k+1}). \end{aligned} \end{equation*}

    On the other hand, by using the Non-Nehari manifold method in [6], we deduce that there exist a ground state solution u_0 of (1.1), such that u_0 \in \mathcal{N} and I_{b, \lambda}(u_0) = c_0 . Besides, for u_{k} = \sum\limits_{i = 1}^{k+1}u_{i}^{\overline{\boldsymbol{r}}_{k}} being a solution of (1.1) with k nodes, by Lemma 3.3 we have known (-1)^{i+1} u_{i}^{\overline{\boldsymbol{r}}_{k}} is positive for each i = 1, \cdots, k+1. Thus, it follows from Lemma 3.2 and Corollary 4.3 in [6] that

    \begin{equation*} \begin{aligned} c_k& = I_{b,\lambda}(u_k) = \max\limits_{{\bf{t}}\in \mathbb{R}_{\geq 0}^{k+1}}J_{b,\lambda}({\bf{t}}\bullet u_k)\\ & = \max\limits_{{\bf{t}}\in \mathbb{R}_{\geq 0}^{k+1}}\left(\sum\limits_{i = 1}^{k+1}I_{b,\lambda}(t_{i}u_{i}^{\overline{\boldsymbol{r}}_{k}})+\sum\limits_{i = 1}^{k+1}\sum\limits_{j\neq i}\frac{b}{4}t_{i}^{2}t_{j}^{2} A(u_{i}^{\overline{\boldsymbol{r}}_{k}},u_{j}^{\overline{\boldsymbol{r}}_{k}}) +\sum\limits_{i = 1}^{k+1}\sum\limits_{j\neq i}\frac{\lambda}{4}t_{i}^{2}t_{j}^{2} B(u_{i}^{\overline{\boldsymbol{r}}_{k}},u_{j}^{\overline{\boldsymbol{r}}_{k}}) \right)\\ & > \max\limits_{{\bf{t}}\in \mathbb{R}_{\geq 0}^{k+1}}\left(\sum\limits_{i = 1}^{k+1}I_{b,\lambda}(t_{i}u_{i}^{\overline{\boldsymbol{r}}_{k}})\right) \ge (k+1)c_0. \end{aligned} \end{equation*}

    This completes the proof.

    Proof of Theorem 1.3. For b, \lambda > 0 , we denote by u_{k}^{b, \lambda} instead of u_{k} to emphasize the dependence of b and \lambda , where u_{k} has been given by Theorem 1.1. For b = \lambda = 0 , by using the similar argument as the proof of Theorem 1.1, we can obtain a solution v^{0}_{k} of equation (1.4) with precisely k nodes. Moreover, I^{\prime}_{0}(v^{0}_{k}) = 0 and I_{0}(v^{0}_{k}) = c_{k}^{0} .

    From now on, we divide the proof into several steps.

    Step 1. We claim that, for any sequence \left\{\lambda_{n}\right\} and \left\{b_{n}\right\} with b_{n} \searrow 0 and \lambda_{n} \searrow 0 as n \rightarrow \infty, \left\{u_{k}^{b_n, \lambda_{n}}\right\} is bounded in H . In fact, choosing nonzero functions \zeta_i\in \mathcal{C}^{\infty}_{0}({B_{i}^{\overline{\boldsymbol{r}}_{k}}}) for i = 1, \cdots, k+1 and define a map h_{i}:\mathcal{H}^{\boldsymbol{r}_{k}}\rightarrow \mathbb{R} by

    \begin{equation*} \begin{aligned} h_{i}({\bf{t}})& = \partial_{u_i}J_{b,\lambda}(t_1^{\frac14}\zeta_1,\cdots,t_{k+1}^{\frac14}\zeta_{k+1})t_{i}^{\frac14}\zeta_{i}\\ & = t_{i}^{\frac12}\|\zeta_{i}\|^{2}_{i}+b t_{i}^{\frac12} \sum\limits_{j = 1}^{k+1} t^{\frac12}_{j}A(\zeta_i,\zeta_j)+\lambda t_{i}^{\frac12} \sum\limits_{j = 1}^{k+1} t^{\frac12}_{j}B(\zeta_i,\zeta_j)- \int_{B_{i}^{\overline{\boldsymbol{r}}_{k}}}K(|x|)f(t_{i}^{\frac14}\zeta_i)t_{i}^{\frac14}\zeta_i dx. \end{aligned} \end{equation*}

    Then, (\bf{F4}) implies that, for any b, \lambda \in[0, 1] , there exists a (k+1) -tuple {\mu} : = \left(\mu_{1}, \cdots, \mu_{k+1}\right) of positive numbers, which does not depend on b and \lambda , such that

    h_{i}( {\mu} ) < 0, {\rm { for }}\; i = 1, \cdots, k+1 .

    Thus, in view of Lemma 3.2, there exists a unique (k+1) tuple \left(t_1(b, \lambda), \cdots, t_{k+1}(b, \lambda)\right)\in (0, +\infty) with t_{i}(b, \lambda) < \mu_i such that

    \left(\bar{\zeta}_{1},\cdots, \bar{\zeta}_{k+1} \right): = \left(t_1^{\frac14}(b,\lambda)\zeta_1,\cdots, t_{k+1}^{\frac14}(b,\lambda)\zeta_{k+1}\right)\in \mathcal{N}_{k}^{\overline{\boldsymbol{r}}_{k}}.

    Besides, by a straightforward computation, there holds

    \begin{equation*} \begin{aligned} I_{b,\lambda}\left(\sum\limits_{i = 1}^{k+1}\bar{\zeta}_i\right)& = I_{b,\lambda}\left(\sum\limits_{i = 1}^{k+1}\bar{\zeta}_i\right) -\frac{1}{4}I^{\prime}_{b,\lambda}\left(\sum\limits_{i = 1}^{k+1}\bar{\zeta}_i\right)\left(\sum\limits_{i = 1}^{k+1}\bar{\zeta}_i\right)\\ & = \frac{1}{4} \sum\limits_{i = 1}^{k+1}\|\bar{\zeta}_i\|^{2}_{i}+\frac{1}{4}\int_{ \mathbb{R}^3}K(|x|)\left(f(\zeta_i)\zeta_i -F(\zeta_i) \right)dx\\ &\le \frac{1}{4} \sum\limits_{i = 1}^{k+1}\|\bar{\zeta}_i\|^{2}_{i}+\frac{1}{4}\|K\|_{L^{\infty}}\int_{ \mathbb{R}^3} \left(C_{1} \bar{\zeta}_{i}^{2}+C_{2} \bar{\zeta}_{i}^{p}\right)dx\\ &\le \frac{1}{4} \sum\limits_{i = 1}^{k+1}\|\mu^{\frac14}_{i}\zeta_i\|^{2}_{i}+\frac{1}{4}\|K\|_{L^{\infty}}\int_{ \mathbb{R}^3} \left(C_{1} \mu_i^{\frac12}\zeta_{i}^{2}+C_{2} \mu_i^{\frac{p}{4}}\zeta_{i}^{p}\right)dx: = C_0, \end{aligned} \end{equation*}

    where we used (\bf{K}) , (\bf{F1}) (\bf{F4}) . Moreover, by Corollary 2.1, we deduce that

    \begin{equation} \begin{aligned} C_{0}& = I_{b,\lambda}(\sum\limits_{i = 1}^{k+1}\bar{\zeta}_i) \geq I_{b,\lambda}\left(u_{k}^{b,\lambda}\right) = I_{b,\lambda}\left(u_{k}^{b,\lambda}\right) -\frac{1}{4} I_{b,\lambda}^{\prime}\left(u_{k}^{b,\lambda}\right) u_{k}^{b,\lambda} \\ & = \frac{1}{4}\left\|u_{k}^{b,\lambda}\right\|^{2} +\frac{1}{4}\int_{\mathbb{R}^{N}}K(|x|)\left( f(u_k^{b,\lambda})u_k^{b,\lambda}-F(u_k^{b,\lambda})\right)dx\\ & \geq \frac{1}{4}(1-\theta) \sum\limits_{i = 1}^{k+1}\left\|u_{k}^{b,\lambda}\right\|^{2}, \end{aligned} \end{equation} (5.2)

    which guarantees the boundedness of \|u_k^{b_n, \lambda_n}\| .

    Step 2. There exists a subsequence of \left\{b_n\right\} and \left\{\lambda_{n}\right\} , still denoted by \left\{b_n\right\} and \left\{\lambda_{n}\right\} , such that u_{k}^{b_n, \lambda_{n}} \rightarrow w_{k} weakly in H . Then, w_{k} is a weak solution of (1.4). Since u_{k}^{b_n, \lambda_{n}} is a sign-changing solution of (1.1) with b = b_{n} and \lambda = \lambda_{n} , then by the Hardy-Littlewood-Sobolev inequality and the compactness of the embedding H \hookrightarrow L^{q}\left(\mathbb{R}^{3}\right) for 2 < q < 6 , we deduce that u_{k}^{b_n, \lambda_{n}} \rightarrow w_{k} strongly in H . Indeed,

    \begin{equation*} \begin{aligned} &\| u_{k}^{b_n,\lambda_{n}} -w_{k} \|^{2} \\ & = \langle I_{b_n,\lambda_{n}}^{\prime}\left(u_{k}^{b_n,\lambda_{n}}\right)-I_{0}^{\prime}\left(w_{k}\right), u_{k}^{b_n,\lambda_{n}}-w_{k}\rangle -b_{n}\int_{ \mathbb{R}^3}|\nabla u^{b_n,\lambda_{n}}_{k}|^2dx \\ & \int_{ \mathbb{R}^3}\nabla u^{b_n,\lambda_{n}}_{k} \left( \nabla u^{b_n,\lambda_{n}}_{k} -\nabla w_{k} \right)dx\\ &-\lambda_{n} \int_{\mathbb{R}^{3}}\int_{\mathbb{R}^{3}} \frac{|u_{k}^{b_n,\lambda_{n}}(x)|^2 u_{k}^{b_n,\lambda_{n}}(y)\left( u_{k}^{b_n,\lambda_{n}}(y)-w_{k}(y)\right)}{|x-y|}dxdy \\ & +\int_{\mathbb{R}^{3}} K(|x|)f\left(u_{k}^{b_n,\lambda_{n}}\right)\left(u_{k}^{b_n,\lambda_{n}}-w_{k}\right) d x \\ &-\int_{\mathbb{R}^{3}} K(|x|)f\left( w_{k}\right)\left(u_{k}^{b_n,\lambda_{n}}-w_{k}\right) d x \rightarrow 0, \quad \;{\rm { as } }\;n \rightarrow \infty , \end{aligned} \end{equation*}

    which implies w_{k}\neq 0 .

    Step 3. Since v^{0}_{k} is a least energy radial solution of (1.4), and we write v^{0}_{k} = v_{k, 1}+\cdots+ v_{k, k+1} , each v_{k, i} is supported on only one annulus B_{i} and vanishes at the complement of it. Thanks to Lemma 3.2, for each \lambda_{n} > 0 , there is a unique (k+1) -tuple \left(a_{1}\left(b_n, \lambda_{n}\right), \cdots, a_{k+1}\left(b_n, \lambda_{n}\right)\right) of positive numbers such that

    \left(a_{1}\left(b_n,\lambda_{n}\right) v_{k, 1}, \cdots, a_{k+1}\left(b_n,\lambda_{n}\right) v_{k, k+1}\right) \in \mathcal{N}_{k} .

    Then, for i = 1, \cdots, k+1 , we have

    \begin{equation*} \begin{aligned} &\left(a_{i}\left(b_n,\lambda_{n}\right)\right)^{2}\left\|v_{k, i}\right\|_{i}^{2}+b_{n}\left(a_{i}\left(b_n,\lambda_{n}\right)\right)^{2}\sum\limits_{j = 1}^{k+1}\left(a_{j}\left(b_n,\lambda_{n}\right)\right)^{2}A(v_{k, i},v_{k, j})\\ &\quad+\lambda_{n}\left(a_{i}\left(b_n,\lambda_{n}\right)\right)^{2}\sum\limits_{j = 1}^{k+1}\left(a_{j}\left(b_n,\lambda_{n}\right)\right)^{2}B(v_{k, i},v_{k, j}) \\ &= \int_{B_{i}} K(|x|)f\left(a_{i}\left(\lambda_{n}\right) v_{k, i}\right) a_{i}\left(\lambda_{n}\right) v_{k, i}dx. \end{aligned} \end{equation*}

    Notice that v_{k, i} satisfies

    \left\|v_{k, i}\right\|_{i}^{2} = \int_{B_{i}}K(|x|) f\left(v_{k, i}\right) v_{k, i} d x .

    Then by (\bf{F1}) (\bf{F4}) and Lemma 3.2, one can easily check that

    \left(a_{1}\left(b_n,\lambda_{n}\right), \cdots, a_{k+1}\left(b_n,\lambda_{n}\right)\right) \rightarrow (1, \cdots, 1), \quad {\rm { as }}\; n \rightarrow \infty.

    Thus,

    \begin{equation*} \begin{aligned} I_{0}(v^{0}_{k})\le I_{0}(w_{k}) = \lim\limits_{n\rightarrow \infty}I_{b_n,\lambda_n}(u_{k}^{b_n,\lambda_n}) &\le \lim\limits_{n\rightarrow \infty} I_{b_n,\lambda_n}\left(\sum\limits_{i = 1}^{k+1}a_{i}(b_n,\lambda_n)v_{k,i} \right)\\ & = I_{0}\left(\sum\limits_{i = 1}^{k+1}v_{k,i}\right) = I_{0}(v^{0}_{k}). \end{aligned} \end{equation*}

    Therefore, w_{k} is a least energy radial solution of (1.1) which changes sign k times.

    Consequently, we complete the proof of Theorem 1.3.

    This manuscript has employed the variational method to study the Kirchhoff-Schrödinger-Poisson system. By using the Gersgorin disc's theorem, Miranda theorem and Brouwer degree theory, we show the existence of infinitely many nodal solutions \{u_k^{b, \lambda}\} with a prescribed number of nodes k for the system. Moreover, we prove that the energy behavior and convergence property of \{u_k^{b, \lambda}\} .

    K. Cheng was supported by Jiangxi Provincial Natural Science Foundation (20202BABL211005), L. Wang was supported by the National Natural Science Foundation of China (No. 12161038) and Science and Technology project of Jiangxi provincial Department of Education (Grant No. GJJ212204).

    The authors declare that there is no conflict of interest.



    [1] P. Agarwal, J. Merker, G. Schuldt, Singular integral Neumann boundary conditions for semilinear elliptic PDEs, Axioms, 10 (2021), 74. https://doi.org/10.3390/axioms10020074 doi: 10.3390/axioms10020074
    [2] A. Azzollini, A. Pomponio, Ground state solutions for the nonlinear Schrödinger-Maxwell equations, J. Math. Anal. Appl., 345 (2008), 90–108. https://doi.org/10.1016/j.jmaa.2008.03.057 doi: 10.1016/j.jmaa.2008.03.057
    [3] T. Bartsch, M. Willem, Infinitely many radial solutions of a semilinear elliptic problem on \mathbb{R}^n, Arch. Rational Mech. Anal., 124 (1993), 261–276. https://doi.org/10.1007/BF00953069 doi: 10.1007/BF00953069
    [4] V. Benci, D. Fortunato, An eigenvalue problem for the Schrödinger-Maxwell equations, Topol. Methods Nonlinear Anal., 11 (1998), 283–293. https://doi.org/10.12775/TMNA.1998.019 doi: 10.12775/TMNA.1998.019
    [5] D. Cao, X. Zhu, On the existence and nodal character of solutions of semilinear elliptic equations, Acta Math. Sci., 8 (1988), 345–359. https://doi.org/10.1016/S0252-9602(18)30312-6 doi: 10.1016/S0252-9602(18)30312-6
    [6] S. Chen, X. Tang, Ground state sign-changing solutions for a class of Schrödinger-Poisson type problems in \mathbb{R}^3, Z. Angew. Math. Phys., 67 (2016). https://doi.org/10.1007/s00033-016-0695-2 doi: 10.1007/s00033-016-0695-2
    [7] W. Chen, T. Zhou, Nodal solutions for Kirchhoff equations with Choquard nonlinearity, J. Fixed Point Theory Appl., 24 (2022), 17. https://doi.org/10.1007/s11784-022-00930-3 doi: 10.1007/s11784-022-00930-3
    [8] J. Deng, J. Yang, Nodal solutions for Schrödinger-Poisson type equations in \mathbb{R}^3, Electron. J. Differ. Eq., 2021 (2021), 277.
    [9] Y. Deng, S. Peng, W. Shuai, Existence and asymptotic behavior of nodal solutions for the Kirchhoff-type problems in \mathbb{R}^3, J. Funct. Anal., 269 (2015), 3500–3527. https://doi.org/10.1016/j.jfa.2015.09.012 doi: 10.1016/j.jfa.2015.09.012
    [10] Y. Deng, S. Peng, W. Shuai, Nodal standing waves for a gauged nonlinear Schrödinger equation in \mathbb{R}^2, J. Differ. Equ., 264 (2018), 4006–4035. https://doi.org/10.1016/j.jde.2017.12.003 doi: 10.1016/j.jde.2017.12.003
    [11] G. Figueiredo, N. Ikoma, J. Júnior, Existence and concentration result for the Kirchhoff type equations with general nonlinearities, Arch. Rational Mech. Anal., 213 (2014), 931–979. https://doi.org/10.1007/s00205-014-0747-8 doi: 10.1007/s00205-014-0747-8
    [12] H. Guo, R. Tang, T. Wang, Infinitely many nodal solutions with a prescribed number of nodes for the Kirchhoff type equations, J. Math. Anal. Appl., 505 (2022), 125519. https://doi.org/10.1016/j.jmaa.2021.125519 doi: 10.1016/j.jmaa.2021.125519
    [13] X. He, W. Zou, Existence and concentration behavior of positive solutions for a Kirchhoff equation in \mathbb{R}^3, J. Differ. Equ., 252 (2012), 1813–1834. https://doi.org/10.1016/j.jde.2011.08.035 doi: 10.1016/j.jde.2011.08.035
    [14] Y. He, G. Li, S. Peng, Concentrating bound states for Kirchhoff type problems in \mathbb{R}^3 involving critical Sobolev exponents, Adv. Nonlinear Stud., 14 (2014), 483–510. https://doi.org/10.1515/ans-2014-0214 doi: 10.1515/ans-2014-0214
    [15] E. Hebey, Stationary Kirchhoff equations with powers, Adv. Calc. Var., 11 (2018), 139–160. https://doi.org/10.1515/acv-2016-0025 doi: 10.1515/acv-2016-0025
    [16] E. Hebey, P. D. Thizy, Stationary Kirchhoff systems in closed 3-dimensional manifolds, Calc. Var. Partial Dif., 54 (2015), 2085–2114. https://doi.org/10.1007/s00526-015-0858-6 doi: 10.1007/s00526-015-0858-6
    [17] E. Hebey, P. D. Thizy, Stationary Kirchhoff systems in closed high dimensional manifolds, Commun. Contemp. Math., 18 (2016), 1550028. https://doi.org/10.1142/S0219199715500285 doi: 10.1142/S0219199715500285
    [18] Z. Huang, J. Yang, W. Yu, Multiple nodal solutions of nonlinear Choquard equations, Electron. J. Differ. Eq., 2017.
    [19] S. Kim, J. Seok, On nodal solutions of the nonlinear Schrödinger-Poisson equations, Commun. Contemp. Math., 14 (2012), 1250041. https://doi.org/10.1142/S0219199712500411 doi: 10.1142/S0219199712500411
    [20] G. Kirchhoff, Mechanik, Teubner, Leipzig, 1883.
    [21] F. Li, Z. Song, Q. Zhang, Existence and uniqueness results for Kirchhoff-Schrödinger-Poisson system with general singularity, Appl. Anal., 96 (2017), 2906–2916. https://doi.org/10.1080/00036811.2016.1253065 doi: 10.1080/00036811.2016.1253065
    [22] E. H. Lieb, M. Loss, Analysis, 2 Eds., Graduate Studies in Mathematics, American Mathematical Society, Providence, RI, 14 (2001).
    [23] J. L. Lions, On some questions in boundary value problems of mathematical physics, North-Holland Math. Stud., North-Holland, New York, 30 (1978), 284–346. https://doi.org/10.1016/S0304-0208(08)70870-3
    [24] Z. Liu, Z. Wang, J. Zhang, Infinitely many sign-changing solutions for the nonlinear Schrödinger-Poisson system, Ann. Mat. Pur. Appl., 195 (2016), 775–794. https://doi.org/10.1007/s10231-015-0489-8 doi: 10.1007/s10231-015-0489-8
    [25] C. Miranda, Un'osservazione su un teorema di brouwer, Boll. UMI, 3 (1940), 5–7.
    [26] D. Ruiz, The Schrödinger-Poisson equation under the effect of a nonlinear local term, J. Funct. Anal., 237 (2006), 655–674. https://doi.org/10.1016/j.jfa.2006.04.005 doi: 10.1016/j.jfa.2006.04.005
    [27] M. Shao, A. Mao, Signed and sign-changing solutions of Kirchhoff type problems, J. Fixed Point Theory Appl., 20 (2018). https://doi.org/10.1007/s11784-018-0486-9 doi: 10.1007/s11784-018-0486-9
    [28] W. Shuai, Sign-changing solutions for a class of Kirchhoff-type problem in bounded domains, J. Differ. Equ., 259 (2015), 1256–1274. https://doi.org/10.1016/j.jde.2015.02.040 doi: 10.1016/j.jde.2015.02.040
    [29] W. Shuai, Q. Wang, Existence and asymptotic behavior of sign-changing solutions for the nonlinear Schrödinger-Poisson system in \mathbb{R}^3, Z. Angew. Math. Phys., 66 (2015), 3267–3282. https://doi.org/10.1007/s00033-015-0571-5 doi: 10.1007/s00033-015-0571-5
    [30] W. A. Strauss, Existence of solitary waves in higher dimensions, Commun. Math. Phys., 55 (1977), 149–162. https://doi.org/10.1007/BF01626517 doi: 10.1007/BF01626517
    [31] M. Struwe, Variational methods, Springer-Verlag, Berlin, 2008. https://doi.org/10.1007/978-3-662-02624-3
    [32] M. Sun, J. Su, L. Zhao, Infinitely many solutions for a Schrödinger-Poisson system with concave and convex nonlinearities, Discrete Contin. Dyn. Syst., 35 (2015), 427. http://doi.org/10.3934/dcds.2015.35.427 doi: 10.3934/dcds.2015.35.427
    [33] X. Tang, B. Cheng, Ground state sign-changing solutions for Kirchhoff type problems in bounded domains, J. Differ. Equ., 261 (2016), 2384–2402. https://doi.org/10.1016/j.jde.2016.04.032 doi: 10.1016/j.jde.2016.04.032
    [34] R. S. Varga, Gersgorin and his circles, Spring Science & Business Media, New York, 2010. https://doi.org/10.1007/978-3-642-17798-9
    [35] Z. Wang, H. Zhou, Sign-changing solutions for the nonlinear schrödinger-poisson system in \mathbb{R}^3, Calc. Var. Partial Dif., 52 (2015), 927–943. https://doi.org/10.1007/s00526-014-0738-5 doi: 10.1007/s00526-014-0738-5
    [36] M. Willem, Minimax theorems, progress in nonlinear differential equations and their applications, Birkhäuser Boston, Inc., Boston, 24 (1996). https://doi.org/10.1007/978-1-4612-4146-1
    [37] M. Xiang, P. Pucci, M. Squassina, B. Zhang, Nonlocal Schrödinger-Kirchhoff equations with external magnetic field, Discrete Contin. Dyn. Syst., 37 (2017), 1631–1649. http://dx.doi.org/10.3934/dcds.2017067 doi: 10.3934/dcds.2017067
    [38] M. Xiang, V. D. Rădulescu, B. Zhang, Nonlocal Kirchhoff diffusion problems: Local existence and blow-up of solutions, Nonlinearity, 31 (2018), 3228–3250. https://doi.org/10.1088/1361-6544/aaba35 doi: 10.1088/1361-6544/aaba35
    [39] Z. Zhang, K. Perera, Sign changing solutions of Kirchhoff type problems via invariant sets of descent flow, J. Math. Anal. Appl., 317 (2006), 456–463. https://doi.org/10.1016/j.jmaa.2005.06.102 doi: 10.1016/j.jmaa.2005.06.102
    [40] G. Zhao, X. Zhu, Y. Li, Existence of infinitely many solutions to a class of Kirchhoff-Schrödinger-Poisson system, Appl. Math. Comput., 256 (2015), 572–581. https://doi.org/10.1016/j.amc.2015.01.038 doi: 10.1016/j.amc.2015.01.038
    [41] L. Zhao, H. Liu, F. Zhao, Existence and concentration of solutions for the Schrödinger-Poisson equations with steep well potential, J. Differ. Equ., 255 (2013), 1–23. https://doi.org/10.1016/j.jde.2013.03.005 doi: 10.1016/j.jde.2013.03.005
  • This article has been cited by:

    1. Yanyan Liu, Leiga Zhao, Least Energy Solutions of the Schrödinger–Kirchhoff Equation with Linearly Bounded Nonlinearities, 2024, 23, 1575-5460, 10.1007/s12346-023-00859-z
  • Reader Comments
  • © 2022 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1794) PDF downloads(86) Cited by(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog