Loading [MathJax]/jax/element/mml/optable/BasicLatin.js
Research article

Equity market integration of China and Southeast Asian countries: further evidence from MGARCH-ADCC and wavelet coherence analysis

  • This paper investigates the short-term and long-term dynamics between China and four Southeast Asian countries (Vietnam, Thailand, Singapore and Malaysia) during period 2008–2018. Our empirical research is based on the Generalized Autoregression Conditional Heteroscedasticity-Asymmetric Dynamic Conditional Covariance (MGARCH-ADCC) model and the wavelet coherence technique which allow us to estimate the time-varying correlation and the co-movement in both time-frequency spaces of stock markets of China and its neighboring countries. The results of the study reveal that stock markets of China and its trading partners are relatively integrated after the global financial crisis of 2008, frequency changes in the pattern of the co-movements and a positive linkage throughout the sample period. Specifically, the conditional correlation of stock returns between China and Singapore is more significantly influenced by negative innovations than by positive shocks to return. Furthermore, the study provides evidence of significant coherence between both the variables for almost the entire studied period in long scale. Therefore, these findings are positive signs for the Chinese and international investors to diversify their portfolio among the stock markets of China and its trading partners.

    Citation: Ngo Thai Hung. Equity market integration of China and Southeast Asian countries: further evidence from MGARCH-ADCC and wavelet coherence analysis[J]. Quantitative Finance and Economics, 2019, 3(2): 201-220. doi: 10.3934/QFE.2019.2.201

    Related Papers:

    [1] Lirong Huang, Jianqing Chen . Existence and asymptotic behavior of bound states for a class of nonautonomous Schrödinger-Poisson system. Electronic Research Archive, 2020, 28(1): 383-404. doi: 10.3934/era.2020022
    [2] Senli Liu, Haibo Chen . Existence and asymptotic behaviour of positive ground state solution for critical Schrödinger-Bopp-Podolsky system. Electronic Research Archive, 2022, 30(6): 2138-2164. doi: 10.3934/era.2022108
    [3] Xiaoyong Qian, Jun Wang, Maochun Zhu . Existence of solutions for a coupled Schrödinger equations with critical exponent. Electronic Research Archive, 2022, 30(7): 2730-2747. doi: 10.3934/era.2022140
    [4] Jun Wang, Yanni Zhu, Kun Wang . Existence and asymptotical behavior of the ground state solution for the Choquard equation on lattice graphs. Electronic Research Archive, 2023, 31(2): 812-839. doi: 10.3934/era.2023041
    [5] Yixuan Wang, Xianjiu Huang . Ground states of Nehari-Pohožaev type for a quasilinear Schrödinger system with superlinear reaction. Electronic Research Archive, 2023, 31(4): 2071-2094. doi: 10.3934/era.2023106
    [6] Xiaoguang Li . Normalized ground states for a doubly nonlinear Schrödinger equation on periodic metric graphs. Electronic Research Archive, 2024, 32(7): 4199-4217. doi: 10.3934/era.2024189
    [7] Zhiyan Ding, Hichem Hajaiej . On a fractional Schrödinger equation in the presence of harmonic potential. Electronic Research Archive, 2021, 29(5): 3449-3469. doi: 10.3934/era.2021047
    [8] Li Cai, Fubao Zhang . The Brezis-Nirenberg type double critical problem for a class of Schrödinger-Poisson equations. Electronic Research Archive, 2021, 29(3): 2475-2488. doi: 10.3934/era.2020125
    [9] Hui Guo, Tao Wang . A note on sign-changing solutions for the Schrödinger Poisson system. Electronic Research Archive, 2020, 28(1): 195-203. doi: 10.3934/era.2020013
    [10] Qihong Shi, Yaqian Jia, Xunyang Wang . Global solution in a weak energy class for Klein-Gordon-Schrödinger system. Electronic Research Archive, 2022, 30(2): 633-643. doi: 10.3934/era.2022033
  • This paper investigates the short-term and long-term dynamics between China and four Southeast Asian countries (Vietnam, Thailand, Singapore and Malaysia) during period 2008–2018. Our empirical research is based on the Generalized Autoregression Conditional Heteroscedasticity-Asymmetric Dynamic Conditional Covariance (MGARCH-ADCC) model and the wavelet coherence technique which allow us to estimate the time-varying correlation and the co-movement in both time-frequency spaces of stock markets of China and its neighboring countries. The results of the study reveal that stock markets of China and its trading partners are relatively integrated after the global financial crisis of 2008, frequency changes in the pattern of the co-movements and a positive linkage throughout the sample period. Specifically, the conditional correlation of stock returns between China and Singapore is more significantly influenced by negative innovations than by positive shocks to return. Furthermore, the study provides evidence of significant coherence between both the variables for almost the entire studied period in long scale. Therefore, these findings are positive signs for the Chinese and international investors to diversify their portfolio among the stock markets of China and its trading partners.


    In this paper, we study a class of Schrödinger-Poisson system with the following version

    {Δu+u+K(x)ϕu=|u|p1u+μh(x)u in R3,Δϕ=K(x)u2 in R3, (1)

    where p(3,5), μ>0, K(x) and h(x) are nonnegative functions. System (1) can be looked on as a non-autonomous version of the system

    {Δu+u+ϕu=f(u) in R3,Δϕ=u2 in R3, (2)

    which has been derived from finding standing waves of the Schrödinger-Poisson system

    {iψtΔψ+ϕψ=f(ψ) in R3,Δϕ=|ψ|2 in R3.

    A starting point of studying system (1) is the following fact. For any uH1(R3) and KL(R3), there is a unique ϕuD1,2(R3) with

    ϕu(x)=14πR3K(y)|u(y)|2|xy|dy

    such that Δϕu=K(x)u2, see e.g. [11,20]. Inserting this ϕu into the first equation of the system (1), we get that

    Δu+u+K(x)ϕuu=|u|p1u+μh(x)u,uH1(R3). (3)

    Problem (3) can be also looked on as a usual semilinear elliptic equation with an additional nonlocal perturbation K(x)ϕuu. Our aim here is to prove some new phenomenon of (3) due to the presence of the term K(x)ϕuu. Before giving the main results, we state the following assumptions.

    (A1) h(x)0, h(x)0 in R3 and h(x)L32(R3)L(R3).

    (A2) There exist b>0 and H0>0 such that h(x)H0eb|x| for all xR3.

    (A3) K(x)0 and K(x)L2(R3)L(R3).

    (A4) There exist a>0 and K0>0 such that K(x)K0ea|x| for all xR3.

    From Lemma 2.1, we know that under the condition (A1), the following eigenvalue problem

    Δu+u=μh(x)u,u H1(R3)

    has a first eigenvalue μ1>0 and μ1 is simple. Denote

    F(u):=R3K(x)ϕu(x)|u(x)|2dx

    and introduce the energy functional Iμ:H1(R3)R associated with (3)

    Iμ(u)=12u2+14F(u)R3(1p+1|u|p+1+μ2h(x)u2)dx,

    where u2=R3(|u|2+u2)dx. From [11] and the Sobolev inequality, Iμ is well defined and IμC1(H1(R3),R). Moreover, for any vH1(R3),

    Iμ(u),v=R3(uv+uv+K(x)ϕuuv|u|p1uv+μh(x)uv)dx.

    It is known that there is a one to one correspondence between solutions of (3) and critical points of Iμ in H1(R3). Note that if uH1(R3) is a solution of (3), then (u,ϕu) is a solution of the system (1). If u0 and u is a solution of (3), then (u,ϕu) is a nonnegative solution of (1) since ϕu is always nonnegative. We call uH1(R3){0} a bound state of (3) if Iμ(u)=0. At this time (u,ϕu) is called a bound state of (1). A bound state u is called a ground state of (3) if Iμ(u)=0 and Iμ(u)Iμ(w) for any bound state w. In this case, we call (u,ϕu) a ground state of (1). The first result is about μ less than μ1.

    Theorem 1.1. Suppose that the assumptions of (A1) - (A4) hold and 0<b<a<2. If 0<μμ1, then problem (3) has at least one nonnegative bound state.

    The second result is about μ in a small right neighborhood of μ1.

    Theorem 1.2. Under the assumptions of (A1) - (A4), if 0<b<a<1, then there exists δ>0 such that, for any μ(μ1,μ1+δ),

    (1) problem (3) has at least one nonnegative ground state u0,μ with Iμ(u0,μ)<0. Moreover, u0,μ(n)0 strongly in H1(R3) for any sequence μ(n)>μ1 and μ(n)μ1;

    (2) problem (3) has another nonnegative bound state u2,μ with Iμ(u2,μ)>0. Moreover, u2,μ(n)uμ1 strongly in H1(R3) for any sequence μ(n)>μ1 and μ(n)μ1, where uμ1 satisfies Iμ1(uμ1)=0 and Iμ1(uμ1)>0.

    The proofs of Theorem 1.1 and Theorem 1.2 are based on critical point theory. There are several difficulties in the road of getting critical points of Iμ in H1(R3) since we are dealing with the problem in the whole space R3, the embedding from H1(R3) into Lq(R3) (2<q<6) is not compact, the appearance of a nonlocal term K(x)ϕuu and the non coercive linear part. To explain our strategy, we review some related known results. For the system (2), under various conditions of f, there are a lot of papers dealing with the existence and nonexistence of positive solutions (u,ϕu)H1(R3)×D1,2(R3), see for example [2,23] and the references therein. The lack of compactness from H1(R3)Lq(R3) (2<q<6) was overcome by restricting the problem in H1r(R3) which is a subspace of H1(R3) containing only radial functions. The existence of multiple radial solutions and non-radial solutions have been obtained in [2,13]. See also [6,15,16,17,18,19,24,29,30] for some other results related to the system (2).

    While for nonautonomous version of Schrödinger-Poisson system, only a few results are known in the literature. Jiang et.al.[21] have studied the following Schrödinger-Poisson system with non constant coefficient

    {Δu+(1+λg(x))u+θϕ(x)u=|u|p2u in R3,Δϕ=u2in R3,lim

    in which the authors prove the existence of ground state solution and its asymptotic behavior depending on \theta and \lambda . The lack of compactness was overcome by suitable assumptions on g(x) and \lambda large enough. The Schrödinger-Poisson system with critical nonlinearity of the form

    \left\{ \begin{array}{ll} -\Delta u +u + \phi u = V(x)|u|^4u + \mu P(x)|u|^{q-2}u \quad \hbox{in}\ \mathbb{R}^3,& \\ -\Delta \phi = u^2 \quad \hbox{in} \ \mathbb{R}^3, \quad 2 < q < 6, \mu > 0& \end{array} \right.

    has been studied by Zhao et al. [31]. Besides some other conditions, Zhao et. al. [31] assume that V(x)\in C(\mathbb{R}^3, \mathbb{R}) , \lim_{|x|\to\infty}V(x) = V_\infty \in (0, \infty) and V(x)\geq V_\infty for x\in \mathbb{R}^3 and prove the existence of one positive solution for 4 < q < 6 and each \mu > 0 . It is also proven the existence of one positive solution for q = 4 and \mu large enough. Cerami et. al. [11] study the following type of Schrödinger-Poisson system

    \begin{equation} \left\{ \begin{array}{ll} -\Delta u +u+L(x)\phi u = g(x, u) \quad & \ \hbox{in}\ \mathbb{R}^3,\\ -\Delta \phi = L(x)u^2\quad & \ \hbox{in} \ \mathbb{R}^3. \end{array} \right. \end{equation} (4)

    Besides some other conditions and the assumption L(x) \in L^2(\mathbb{R}^3) , they prove the existence and nonexistence of ground state solutions. We emphasize that L(x) \in L^2(\mathbb{R}^3) will imply suitable compactness property of the coupled term L(x)\phi u . Huang et. al. [20] have used this property to prove the existence of multiple solutions of (4) when g(x,u) = a(x)|u|^{p-2}u + \mu h(x)u and \mu stays in a right neighborhood of \mu_1 . The lack of compactness was overcome by suitable assumptions on the sign changing function a(x) . While for (3), none of the aboved mentioned properties can be used. We have to analyze the energy level of the functional I_\mu such that the Palais-Smale ( (PS) for short) condition may hold at suitable interval. Also for (3), another difficulty is to find mountain pass geometry for the functional I_\mu in the case of \mu \geq \mu_1 . We point out that for the semilinear elliptic equation

    \begin{equation} -\Delta u = a(x)|u|^{p-2}u + \tilde{\mu}k(x)u, \quad \ \hbox{in}\ \mathbb{R}^N, \end{equation} (5)

    Costa et.al.[14] have proven the mountain pass geometry for the related functional of (5) when \tilde{\mu} \geq \tilde{\mu}_1 , where \tilde{\mu}_1 is the first eigenvalue of -\Delta u = \tilde{\mu}k(x)u in D^{1,2}(\mathbb{R}^N) . Costa et. al. have managed to do these with the help of the condition \int_{\mathbb{R}^N} a(x)\tilde{e}_1^p dx < 0 , where \tilde{e}_1 is a positive eigenfunction corresponding to \tilde{\mu}_1 . In the present paper, it is not possible to use such kind of condition. We will develop further the techniques in [20] to prove the mountain pass geometry. A third difficulty is to look for a ground state of (3). A usual method of getting a ground state is by minimizing the functional I_\mu over the Nehari set \{u\in H^1(\mathbb{R}^3)\backslash\{0\}\ :\ \langle I'_\mu(u), u\rangle = 0\} . But in the case of \mu > \mu_1 , one can not do like this because we do not know if 0 belongs to the boundary of this Nehari set. To overcome this trouble, we will minimize the functional over the set \{u\in H^1(\mathbb{R}^3)\backslash\{0\}\ :\ I'_\mu(u) = 0\} .

    This paper is organized as follows. In Section 2, we give some preliminaries. Special attentions are focused on several lemmas analyzing the Palais-Smale conditions of the functional I_\mu , which will play an important role in the proofs of Theorem 1.1 and Theorem 1.2. In Section 3, we prove Theorem 1.1. And Section 4 is devoted to the proof of Theorem 1.2.

    Notations. Throughout this paper, o(1) is a generic infinitesimal. The H^{-1}(\mathbb{R}^3) denotes dual space of H^1(\mathbb{R}^3) . L^q(\mathbb{R}^3) (1\leq q\leq +\infty) is a Lebesgue space with the norm denoted by \|u\|_{L^q} . The S_{p+1} is defined by

    S_{p+1} = {\inf\limits_{u\in H^1(\mathbb{R}^3)\setminus\{0\}}}\frac{\int_{\mathbb{R}^3}\left(|\nabla u|^2 + |u|^2\right)dx}{\left(\int_{\mathbb{R}^3}| u|^{p+1}dx\right)^{\frac2{p+1}}}.

    For any \rho>0 and x\in \mathbb{R}^3 , B_\rho(x) denotes the ball of radius \rho centered at x . C or C_j ( j = 1,\ 2,\ \cdots ) denotes various positive constants, whose exact value is not important.

    In this section, we give some preliminary lemmas, which will be helpful to analyze the (PS) conditions for the functional I_\mu . Firstly, for any u\in H^1(\mathbb{R}^3) and K\in L^{\infty}(\mathbb{R}^3) , defining the linear functional

    L_u(v) = \int_{{\mathbb{R}}^3} K(x)u^2vdx,\qquad v\in D^{1, 2}(\mathbb{R}^3),

    one may deduce from the Hölder and the Sobolev inequalities that

    \begin{equation} |L_u(v)|\leq C\|u\|_{L^{\frac{12}{5}}}^2\|v\|_{L^6}\leq C \|u\|_{L^{\frac{12}{5}}}^2\|v\|_{D^{1, 2}}. \end{equation} (6)

    Hence, for any u\in H^1(\mathbb{R}^3) , the Lax-Milgram theorem implies that there exists a unique \phi_u\in D^{1, 2}(\mathbb{R}^3) such that -\Delta \phi = K(x) u^2 in D^{1, 2}(\mathbb{R}^3) . Moreover it holds that

    \phi_u(x) = \frac{1}{4\pi}\int_{\mathbb{R}^3}\frac{K(y)u^2(y)}{|x-y|}dy.

    Clearly \phi_u(x)\geq 0 for any x\in \mathbb{R}^3 . We also have that

    \begin{equation} \|\phi_u\|_{D^{1, 2}}^2 = \int_{\mathbb{R}^3}|\nabla \phi_u|^2dx = \int_{\mathbb{R}^3}K(x)\phi_uu^2dx. \end{equation} (7)

    Using (6) and (7), we obtain that

    \begin{equation} \|\phi_{u}\|_{L^6} \leq C\|\phi_u\|_{D^{1, 2}}\leq C\|u\|_{L^{\frac{12}{5}}}^2\leq C\|u\|^2. \end{equation} (8)

    Then we deduce that

    \begin{equation} \int_{\mathbb{R}^3} K(x)\phi_u(x)u^2(x)dx\leq C \|u\|^4. \end{equation} (9)

    Hence on H^1(\mathbb{R}^3) , both the functional

    \begin{equation} F(u) = \int_{\mathbb{R}^3} K(x)\phi_u(x)u^2(x)dx \end{equation} (10)

    and

    \begin{equation} I_\mu(u) = \frac{1}{2}\|u\|^2+\frac{1}{4}F(u)-\int_{\mathbb{R}^3}\left({\frac{1}{p+1}} |u|^{p+1} + \frac{\mu}{2} h(x)u^2\right)dx \end{equation} (11)

    are well defined and C^1 . Moreover, for any v\in H^1(\mathbb{R}^3) ,

    \langle I'_\mu(u), v\rangle = \int_{\mathbb{R}^3}\left( \nabla u\nabla v+uv + K(x) \phi_uuv - |u|^{p-1}uv - \mu h(x)uv\right)dx.

    The following Lemma 2.1 is a direct consequence of [28,Lemma 2.13].

    Lemma 2.1. Assume that the hypothesis (A1) holds. Then the functional u\in H^1(\mathbb{R}^3) \mapsto \int_{\mathbb{R}^3} h(x)u^2dx is weakly continuous and for each v\in H^1(\mathbb{R}^3) , the functional u\in H^1(\mathbb{R}^3) \mapsto \int_{\mathbb{R}^3} h(x)uv dx is weakly continuous.

    Using the spectral theory of compact symmetric operators on Hilbert space, the above lemma implies the existence of a sequence of eigenvalues (\mu_n)_{n\in \mathbb{N}} of

    -\Delta u+u = \mu h(x)u, \; \hbox{in}\; H^1(\mathbb{R}^3)

    with \mu_1 < \mu_2 \leq \cdots and each eigenvalue being of finite multiplicity. The associated normalized eigenfunctions are denoted by e_1, e_2, \cdots with \|e_i\| = 1, i = 1, 2, \cdots . Moreover, one has \mu_1>0 with an eigenfunction e_1>0 in \mathbb{R}^3 . In addition, we have the following variational characterization of \mu_n :

    \mu_1 = {\inf\limits_{u\in {H^1(\mathbb{R}^3)}\setminus \{0\}}}\frac{\|u\|^2}{ \int_{\mathbb{R}^3} h(x)u^2dx}, \quad \mu_n = {\inf\limits_{u\in {S_{n-1}^\perp}\setminus \{0\}}}\frac{\|u\|^2}{ \int_{\mathbb{R}^3} h(x)u^2dx},

    where S_{n-1}^\perp = \{span \{e_1, e_2, \cdots, e_{n-1}\}\}^\perp .

    Next we analyze the (PS) condition of the functional I_\mu in H^1(\mathbb{R}^3) . The following definition is standard.

    Definition 2.2. For d\in\mathbb{R} , the functional I_\mu is said to satisfy (PS)_d condition if for any (u_n)_{n\in\mathbb{N}} \subset H^1(\mathbb{R}^3) with I_\mu(u_n)\to d and I'_\mu(u_n)\to 0 , the (u_n)_{n\in\mathbb{N}} contains a convergent subsequence in H^1(\mathbb{R}^3) . The functional I_\mu is said to satisfy (PS) conditions if I_\mu satisfies (PS)_d condition for any d\in \mathbb{R} .

    Lemma 2.3. Let (u_n)_{n\in\mathbb{N}} \subset H^1(\mathbb{R}^3) be such that I_\mu(u_n) \to d \in \mathbb{R} and I'_\mu(u_n) \to 0 , then (u_n)_{n\in\mathbb{N}} is bounded in H^1(\mathbb{R}^3) .

    Proof. For n large enough, we have that

    \begin{eqnarray} &d& + 1 + o(1)\|u_n\| = I_\mu(u_n) -\frac14 \langle I_\mu'(u_n),u_n \rangle \\ & = & \frac{1}{4}\|u_n\|^2 - \frac{\mu}{4}\int_{\mathbb{R}^3}h(x)u_n^2dx + \frac{p-3}{4(p+1)}\int_{\mathbb{R}^3}|u_n|^{p+1} dx. \end{eqnarray} (12)

    Note that \frac{p+1}{p-1}>\frac{3}{2} for p\in (3,5) . Then for any \vartheta > 0 , we obtain from h\in L^{\frac32}(\mathbb{R}^3)\cap L^{\infty}(\mathbb{R}^3) that

    \begin{array}{rl} \int_{\mathbb{R}^3}h(x)u_n^2dx &\leq \left(\int_{\mathbb{R}^3}|u_n|^{p+1}dx\right)^{\frac2{p+1}} \left(\int_{\mathbb{R}^3}|h(x)|^{\frac{p+1}{p-1}}dx\right)^{\frac{p-1}{p+1}} \\ & \leq \frac{2\vartheta}{p+1} \int_{\mathbb{R}^3}|u_n|^{p+1}dx + \frac{p-1}{p+1} \vartheta^{-\frac2{p-1}}\int_{\mathbb{R}^3}|h(x)|^{\frac{p+1}{p-1}}dx. \end{array}

    Choosing \vartheta = \frac{p-3}{2\mu} , we get

    \begin{equation} d + 1 + o(1)\|u_n\| \geq \frac{1}{4}\|u_n\|^2 - D(p, h) \mu^{\frac{p+1}{p-1}}, \end{equation} (13)

    where D(p,h) = \frac{p-1}{4(p+1)} \left(\frac{p-3}{2}\right)^{-\frac{2}{p-1}} \int_{\mathbb{R}^3}|h(x)|^{\frac{p+1}{p-1}}dx. Hence (u_n)_{n\in\mathbb{N}} is bounded in H^1(\mathbb{R}^3) .

    The following lemma is a variant of Brezis-Lieb lemma. One may find the proof in [20].

    Lemma 2.4. [20] If a sequence (u_n)_{n\in\mathbb{N}} \subset H^1(\mathbb{R}^3) and u_n \rightharpoonup u_0 weakly in H^1(\mathbb{R}^3) , then

    \lim\limits_{n\to\infty}F(u_n) = F(u_0) + \lim\limits_{n\to\infty}F(u_n-u_0).

    Lemma 2.5. There is a \delta_1 > 0 such that for any \mu \in [\mu_1, \mu_1 + \delta_1) , any solution u of (3) satisfies

    I_\mu(u) > - \frac{p-1}{2(p+1)} S_{p+1}^{\frac{p+1}{p-1}}.

    Proof. Since u is a solution of (3), we get that

    \begin{array}{rl} I_\mu(u) & = \frac12 \left(\|u\|^2 - \mu\int_{\mathbb{R}^3} h(x)u^2 dx\right) + \frac14 F(u) - \frac1{p+1}\int_{\mathbb{R}^3} |u|^{p+1} dx\\ & = \frac{p-1}{2(p+1)} \left(\|u\|^2 - \mu\int_{\mathbb{R}^3} h(x)u^2 dx\right) +\frac{p-3}{4(p+1)}F(u). \end{array}

    Noticing that \|u\|^2 \geq \mu_1\int_{\mathbb{R}^3} h(x)u^2 dx for any u\in H^1(\mathbb{R}^3) , we deduce that for any u\neq 0 ,

    I_{\mu_1}(u) \geq \frac{p-3}{4(p+1)}F(u) > 0.

    Next, we claim: there is a \delta_1 > 0 such that for any \mu \in [\mu_1, \mu_1 + \delta_1) , any solution u of (3) satisfies

    I_\mu(u) > - \frac{p-1}{2(p+1)}S_{p+1}^{\frac{p+1}{p-1}}.

    Suppose this claim is not true, then there is a sequence \mu^{(n)} > \mu_1 with \mu^{(n)}\to \mu_1 and solutions u_{\mu^{(n)}} of (3) such that

    I_{\mu^{(n)}}(u_{\mu^{(n)}}) \leq - \frac{p-1}{2(p+1)}S_{p+1}^{\frac{p+1}{p-1}}.

    Note that I'_{\mu^{(n)}}(u_{\mu^{(n)}}) = 0. Then we deduce that for n large enough,

    \begin{eqnarray} I_{\mu^{(n)}}(u_{\mu^{(n)}}) + o(1)\|u_{\mu^{(n)}}\|&\geq & I_{\mu^{(n)}}(u_{\mu^{(n)}}) -\frac14 \langle I'_{\mu^{(n)}}(u_{\mu^{(n)}}), u_{\mu^{(n)}} \rangle \\ &\geq & \frac{1}{4}\|u_{\mu^{(n)}}\|^2 - D(p,h) \left(\mu^{(n)}\right)^{\frac{p+1}{p-1}}. \end{eqnarray}

    This implies that (u_{\mu^{(n)}})_{n\in\mathbb{N}} is bounded in H^1(\mathbb{R}^3) . Since for any n\in \mathbb{N} , \|u_{\mu^{(n)}}\|^2 \geq \mu_1\int_{\mathbb{R}^3} h(x)(u_{\mu^{(n)}})^2 dx , we obtain that as \mu^{(n)}\to \mu_1

    \|u_{\mu^{(n)}}\|^2 - \mu^{(n)} \int_{\mathbb{R}^3} h(x)(u_{\mu^{(n)}})^2 dx \geq \left(1-\frac{\mu^{(n)}}{\mu_1}\right)\|u_{\mu^{(n)}}\|^2\to 0

    because (u_{\mu^{(n)}})_{n\in\mathbb{N}} is bounded in H^1(\mathbb{R}^3) . Noting that

    \begin{array}{rl} I_{\mu^{(n)}}(u_{\mu^{(n)}}) & = \frac{p-1}{2(p+1)} \left(\|u_{\mu^{(n)}}\|^2 - \mu^{(n)} \int_{\mathbb{R}^3} h(x)(u_{\mu^{(n)}})^2 dx\right) \\ &\qquad + \frac{p-3}{4(p+1)}F(u_{\mu^{(n)}}), \end{array}

    we deduce that

    \liminf\limits_{n\to\infty}I_{\mu^{(n)}}(u_{\mu^{(n)}}) \geq \frac{p-3}{4(p+1)}\liminf\limits_{n\to\infty} F(u_{\mu^{(n)}}) \geq 0,

    which contradicts to the

    I_{\mu^{(n)}}(u_{\mu^{(n)}}) \leq - \frac{p-1}{2(p+1)} S_{p+1}^{\frac{p+1}{p-1}}.

    This proves the claim and the proof of Lemma 2.5 is complete.

    Lemma 2.6. If \mu\in [\mu_1, \mu_1 + \delta_1) , then I_\mu satisfies (PS)_d condition for any d < 0 .

    Proof. Let (u_n)_{n\in\mathbb{N}}\subset H^1(\mathbb{R}^3) be a (PS)_d sequence of I_\mu with d < 0 . Then for n large enough,

    d + o(1) = \frac{1}{2}\|u_n\|^2 - \frac{\mu}{2}\int_{\mathbb{R}^3}h(x)u_n^2dx + \frac{1}{4}F(u_n) -\frac{1}{p+1}\int_{\mathbb{R}^3}|u_n|^{p+1}dx

    and

    \langle I'_\mu (u_n), u_n\rangle = \|u_n\|^2 - \mu\int_{\mathbb{R}^3}h(x)u_n^2dx + F(u_n) - \int_{\mathbb{R}^3}|u_n|^{p+1}dx.

    Then we can prove that (u_n)_{n\in\mathbb{N}} is bounded in H^1(\mathbb{R}^3) . Without loss of generality, we may assume that u_n \rightharpoonup u_0 weakly in H^1(\mathbb{R}^3) and u_n \to u_0 a. e. in \mathbb{R}^3 . Denoting w_n : = u_n - u_0 , we obtain from Brezis-Lieb lemma and Lemma 2.4 that for n large enough,

    \|u_n\|^2 = \|u_0\|^2 + \|w_n\|^2 +o(1),
    F(u_n) = F(u_0) + F(w_n) +o(1)

    and

    \|u_n\|^{p+1}_{L^{p+1}} = \|u_0\|^{p+1}_{L^{p+1}} + \|w_n\|^{p+1}_{L^{p+1}} +o(1).

    Using Lemma 2.1, we also have that \int_{\mathbb{R}^3}h(x)u_n^2dx\to \int_{\mathbb{R}^3}h(x)u_0^2dx as n\to\infty . Therefore

    \begin{eqnarray} d + o(1)& = & I_\mu(u_n) = I_\mu(u_0) + \frac{1}{2}\| w_n\|^2\\ &+& \frac{1}{4}F(w_n) -\frac{1}{p+1}\int_{\mathbb{R}^3}|w_n|^{p+1}dx. \end{eqnarray} (14)

    Noticing \langle I'_\mu (u_n), \psi \rangle \to 0 for any \psi\in H^1(\mathbb{R}^3) , we obtain that I'_\mu (u_0) = 0 . From which we deduce that

    \begin{equation} \|u_0\|^2 - \mu\int_{\mathbb{R}^3}h(x)u_0^2dx + F(u_0) = \int_{\mathbb{R}^3}|u_0|^{p+1}dx. \end{equation} (15)

    Since (u_n)_{n\in\mathbb{N}} is bounded in H^1(\mathbb{R}^3) , we obtain from I'_\mu(u_n) \to 0 that

    o(1) = \|u_n\|^2 - \mu\int_{\mathbb{R}^3}h(x)u_n^2dx + F(u_n) - \int_{\mathbb{R}^3}|u_n|^{p+1}dx.

    Combining this with (15) as well as Lemma 2.1, we obtain that

    \begin{equation} o(1) = \|w_n\|^2 + F(w_n) -\int_{\mathbb{R}^3}|w_n|^{p+1}dx. \end{equation} (16)

    Recalling the definition of S_{p+1} , we have that \| u\|^2 \geq S_{p+1} \|u\|_{L^{p+1}}^2 for any u\in\ H^1(\mathbb{R}^3) . Now we distinguish two cases:

    \bf (i) \int_{\mathbb{R}^3}|w_n|^{p+1}dx\not\to 0 as n\to\infty ;

    \bf (ii) \int_{\mathbb{R}^3}|w_n|^{p+1}dx \to 0 as n\to\infty .

    Suppose that the case (ⅰ) occurs. We may obtain from (16) that

    \|w_n\|^2 \geq S_{p+1} \left(\|w_n\|^2 + F(w_n) - o(1)\right)^{\frac2{p+1}}.

    Hence we get that for n large enough,

    \begin{equation} \|w_n\|^2 \geq S_{p+1}^{\frac{p+1}{p-1}} + o(1). \end{equation} (17)

    Therefore using (14), (16) and (17), we deduce that for n large enough,

    \begin{eqnarray} & d& + o(1) = I_\mu(u_n) \\ & = & I_\mu(u_0) + \frac{1}{2}\|w_n\|^2 + \frac{1}{4}F(w_n) -\frac{1}{p+1}\int_{\mathbb{R}^3}|w_n|^{p+1}dx. \\ & = & I_\mu(u_0) + \frac{p-1}{2(p+1)}\|w_n\|^2 + \frac{p-3}{4(p+1)}F(w_n) \\ & > & - \frac{p-1}{2(p+1)}S_{p+1}^{\frac{p+1}{p-1}} + \frac{p-1}{2(p+1)}\|w_n\|^2 + \frac{p-3}{4(p+1)}F(w_n) \\ & > & 0, \end{eqnarray} (18)

    which contradicts to the condition d < 0 . This means that the case (ⅰ) does not occur. Therefore the case (ⅱ) occurs. Using (16), we deduce that \|w_n\|^2 \to 0 as n\to\infty . Hence we have proven that u_n\to u_0 strongly in H^1(\mathbb{R}^3) .

    Next we give a mountain pass geometry for the functional I_\mu .

    Lemma 2.7. There exist \delta_2 > 0 with \delta_2 \leq \delta_1 , \rho > 0 and \alpha > 0 , such that for any \mu\in [\mu_1, \mu_1 + \delta_2) , I_{\mu}|_{\partial B_{\rho}}\geq \alpha > 0.

    Proof. For any u\in H^1(\mathbb{R}^3) , there exist t\in \mathbb{R} and v\in S_1^\perp such that

    \begin{equation} u = te_1+v, \; \hbox{where}\; \int_{\mathbb{R}^3} \left(\nabla v \nabla e_1 + ve_1\right)dx = 0. \end{equation} (19)

    Hence we deduce that

    \begin{equation} \|u\|^2 = \|\nabla (te_1+v)\|_{L^2}^2 + \|te_1+v\|_{L^2}^2 = t^2+\|v\|^2, \end{equation} (20)
    \begin{equation} \mu_2\int_{\mathbb{R}^3} h(x)v^2dx\leq \|v\|^2, \; \; \mu_1\int_{\mathbb{R}^3} h(x)e_1^2dx = \|e_1\|^2 = 1 \end{equation} (21)

    and

    \begin{equation} \mu_1\int_{\mathbb{R}^3} h(x)e_1 v dx = \int_{\mathbb{R}^3} \left(\nabla v \nabla e_1+ve_1\right)dx = 0. \end{equation} (22)

    We first consider the case of \mu = \mu_1 . Denoting \theta_1: = (\mu_2-\mu_1)/2\mu_2 > 0 , then by the relations from (19) to (22), we obtain that

    \begin{array}{rl} I_{\mu_1}(u)& = \frac{1}{2} \|u\|^2+\frac{1}{4}F(u)- \frac{\mu_1}{2}\int_{\mathbb{R}^3}h(x)u^2dx - \frac{1}{p+1}\int_{\mathbb{R}^3} |u|^{p+1}dx\\ & = \frac{1}{2} \|t e_1 + v\|^2 +\frac{1}{4} F(te_1+v)\\ & \quad -\frac{\mu_1}{2}\int_{\mathbb{R}^3}h(x)(te_1+v)^2dx-\frac{1}{p+1}\int_{\mathbb{R}^3} |te_1 + v|^{p+1}dx\\ & \geq\frac{1}{2}\left(1-\frac{\mu_1}{\mu_2}\right) \|v\|^2 +\frac{1}{4} F(te_1+v)-\frac{1}{p+1}\int_{\mathbb{R}^3} |te_1 + v|^{p+1}dx\\ & \geq \theta_1\|v\|^2 +\frac{1}{4} F(te_1+v) - C_1|t|^{p+1} - C_2\|v\|^{p+1}. \end{array}

    Next we estimate the term F(te_1+v) . Using the expression of F(u) , we have that

    F(te_1+v) = \frac{1}{4\pi}\int_{\mathbb{R}^3\times \mathbb{R}^3}\frac{K(x)K(y)(te_1(y)+v(y))^2(te_1(x)+v(x))^2}{|x-y|}dydx.

    Since

    \begin{array}{rl} & \left(te_1(y)+v(y)\right)^2\left(te_1(x)+v(x)\right)^2 = t^4 (e_1(y))^2(e_1(x))^2 + (v(y))^2(v(x))^2\\ & \qquad + 2t^3\left(e_1(y)(e_1(x))^2v(y) + e_1(x)(e_1(y))^2v(x)\right)\\ & \qquad + 2t\left(e_1(x)v(x)(v(y))^2 + e_1(y)v(y)(v(x))^2\right)\\ & \qquad +t^2\left((e_1(x))^2(v(y))^2 + 4e_1(y)e_1(x)v(y)v(x) + (e_1(y))^2(v(x))^2\right), \end{array}

    we know that

    \begin{equation} \left|\int_{\mathbb{R}^3\times \mathbb{R}^3}\frac{K(x)K(y)\left(e_1(y)(e_1(x))^2v(y) + e_1(x)(e_1(y))^2v(x)\right)}{|x-y|}dydx\right|\leq C\|v\|; \end{equation} (23)
    \begin{equation} \left|\int_{\mathbb{R}^3\times \mathbb{R}^3}\frac{K(x)K(y)\left(2(e_1(x))^2(v(y))^2 + 4e_1(y)e_1(x)v(y)v(x)\right)}{|x-y|}dydx\right|\leq C\|v\|^2 \end{equation} (24)

    and

    \begin{equation} \left|\int_{\mathbb{R}^3\times \mathbb{R}^3}\frac{K(x)K(y)\left(e_1(x)v(x)(v(y))^2 + e_1(y)v(y)(v(x))^2\right)}{|x-y|}dydx\right|\leq C\|v\|^3. \end{equation} (25)

    Hence

    \begin{array}{rl} I_{\mu_1}(u)& \geq \theta_1\|v\|^2 + \theta_2|t|^4 - C_1|t|^{p+1} - C_2\|v\|^{p+1} \\ & \qquad -C_3|t|^3\|v\| - C_4|t|^2\|v\|^2 - C_5|t|\|v\|^3+\frac{1}{4}F(v), \end{array}

    where \theta_2 = \frac14 \int_{\mathbb{R}^3}K(x)\phi_{e_1} e_1^2 dx . Note that

    t^2\|v\|^2 \leq \frac2{p+1} |t|^{p+1} + \frac{p-1}{p+1} \|v\|^{\frac{2(p+1)}{p-1}},
    |t|\|v\|^3 \leq \frac1{p+1} |t|^{p+1} + \frac{p}{p+1}\|v\|^{\frac{3(p+1)}{p}}

    and for some q_0 with 2 < q_0 < 4 , we also have that

    |t|^3\|v\| \leq \frac{1}{q_0} \|v\|^{q_0} + \frac{q_0 - 1}{q_0} |t|^{\frac{3q_0}{q_0 - 1}}.

    Therefore we deduce that

    \begin{equation} \begin{array}{rl} & I_{\mu_1}(u) \geq \theta_1 \|v\|^2 + \theta_2 |t|^4 - \frac{C_3}{q_0} \|v\|^{q_0} - \frac{C_3(q_0 - 1)}{q_0} |t|^{\frac{3q_0}{q_0 - 1}}\\ & \quad -\frac{2C_4}{p+1} |t|^{p+1} - \frac{(p-1)C_4}{p+1}\|v\|^{\frac{2(p+1)}{p-1}} - \frac{C_5}{p+1} |t|^{p+1} \\ & \quad - \frac{pC_5}{p+1}\|v\|^{\frac{3(p+1)}{p}} - C|t|^{p+1} - C\|v\|^{p+1}. \end{array} \end{equation} (26)

    From q_0 > 2 and \frac{3q_0}{q_0 - 1} > 4 (since q_0 < 4 ), we know that there are positive constants \theta_3 , \theta_4 and \tilde{\theta}_3 , \tilde{\theta}_4 such that

    I_{\mu_1}(u) \geq \theta_3 \|v\|^2 + \theta_4 |t|^4

    provided that \|v\| \leq \tilde{\theta}_3 and |t|\leq \tilde{\theta}_4 . Hence there are positive constants \theta_5 and \tilde{\theta}_5 such that

    \begin{equation} I_{\mu_1}(u) \geq \theta_5 \|u\|^4\qquad \hbox{for}\quad \|u\|^2 \leq \tilde{\theta}_5^2. \end{equation} (27)

    Set \bar{\delta} : = \min\{ \frac{\mu_1}{2}\theta_5\tilde{\theta}^2_5,\ \ \mu_2 - \mu_1\} > 0 and \delta_2 : = \min \{\bar{\delta}, \delta_1\} . Then for any \mu\in [\mu_1, \mu_1 + \delta_2) , we deduce from (27) that

    \begin{array}{rl} I_\mu (u) & = I_{\mu_1}(u) + \frac{1}{2}(\mu_1 - \mu)\int_{\mathbb{R}}h(x)u^2dx\\ & \geq \theta_5 \|u\|^4 - \frac{\mu - \mu_1}{2\mu_1}\|u\|^2\\ & = \|u\|^2 \left(\theta_5 \|u\|^2 - \frac{\mu - \mu_1}{2\mu_1}\right)\\ & \geq \|u\|^2\left(\frac{1}{2} \theta_5 \tilde{\theta}^2_5 - \frac{1}{4}\theta_5\tilde{\theta}^2_5\right) = \frac{1}{4}\theta_5\tilde{\theta}^2_5\|u\|^2 \end{array}

    for \frac{1}{2}\tilde{\theta}^2_5 \leq \|u\|^2 \leq \tilde{\theta}^2_5 . Choosing \rho^2 = \frac{1}{2}\tilde{\theta}^2_5 and \alpha = \frac{1}{4}\theta_5\tilde{\theta}^2_5\rho^2, we finish the proof of Lemma 2.7.

    In this section, our aim is to prove Theorem 1.1. For 0 < \mu < \mu_1 , it is standard to prove that the functional I_{\mu} contains mountain pass geometry. For \mu = \mu_1 , as we have seen in Lemma 2.7, with the help of the competing between the Poisson term K(x)\phi_u u and the nonlinear term, the 0 is a local minimizer of the functional I_{\mu_1} and I_{\mu_1} contains mountain pass geometry. To get a mountain pass type critical point of the functional I_\mu , it suffices to prove the (PS)_{d} condition by the mountain pass theorem of [3]. In the following we will focus our attention to the case of \mu = \mu_1 , since the case of 0 < \mu < \mu_1 is similar.

    Proposition 3.1. Let the assumptions (A1)-(A4) hold and 0 < b <a < 2 . Define

    d_{\mu_1} = \inf\limits_{\gamma \in \Gamma_1}\sup\limits_{t\in [0,1]}I_{\mu_1}(\gamma(t))

    with

    \Gamma_1 = \left\{ \gamma \in C([0,1], H^1(\mathbb{R}^3))\ : \ \gamma(0) = 0,\ I_{\mu_1}(\gamma(1)) < 0 \right\}.

    Then d_{\mu_1} is a critical value of I_{\mu_1} .

    Before proving Proposition 3.1, we analyze the (PS)_{d_{\mu_1}} condition of I_{\mu_1} . Let U(x) be the unique positive solution of -\Delta u + u = |u|^{p-1}u in H^1(\mathbb{R}^3) . We know that for any \varepsilon \in (0,1) , there is a C \equiv C(\varepsilon)>0 such that U(x) \leq C e^{-(1-\varepsilon)|x|}.

    Lemma 3.2. If the assumptions (A1)-(A4) hold and 0 < b <a < 2 , then the d_{\mu_1} defined in Proposition 3.1 satisfies d_{\mu_1} < \frac{p-1}{2(p+1)} S_{p+1}^{\frac{p+1}{p-1}}.

    Proof. It suffices to find a path \gamma(t) starting from 0 such that

    \sup\limits_{t\in [0,1]}I_{\mu_1}(\gamma(t)) < \frac{p-1}{2(p+1)} S_{p+1}^{\frac{p+1}{p-1}}.

    Define U_R(x) = U(x-R\theta) with \theta = (0,0,1) . Note that for the U_R defined as above, the I_{\mu_1}(tU_R) \to -\infty as t\to +\infty and I_{\mu_1}(tU_R) \to 0 as t\to 0 . We know that there is a unique T_R > 0 such that \frac{\partial}{\partial t}I_{\mu_1}(t U_R)|_{t = T_R} = 0 , which is

    \|U_R\|^2 -\mu_1\int h(x)U_R^2 dx+ T_R^2F(U_R) - T_R^{p-1}\int U_R^{p+1} dx = 0.

    If T_R\to 0 as R\to\infty , then \|U_R\|^2 -\mu_1\int h(x)U_R^2dx \to 0 as R\to\infty , which is impossible. If T_R\to\infty as R\to\infty , then as R\to\infty ,

    \frac{1}{T_R^2}\left(\|U_R\|^2 -\mu_1\int h(x)U_R^2dx\right)+F(U_R) = T_R^{p-3}\int U_R^{p+1}dx \to \infty,

    which is impossible either. Hence we only need to estimate I_{\mu_1}(tU_R) for t in a finite interval and we may write

    I_{\mu_1}(tU_R) \leq g(t) + CF(U_R),

    where

    g(t) = \frac{t^2}{2}\left(\|U_R\|^2 - \mu_1 \int_{\mathbb{R}^3} h(x) U_R^2 dx\right) - \frac{|t|^{p+1}}{p+1}\int_{\mathbb{R}^3} U_R^{p+1} dx.

    Noting that under the assumptions (A1)-(A4) , we obtain that for R large enough,

    \begin{equation} \begin{array}{rl} F(U_R) & \leq \left(\int_{\mathbb{R}^3}K(x)^{\frac65}U_R^{\frac{12}{5}}dx\right)^{\frac56} \left(\int_{\mathbb{R}^3} \phi_{U_R}^6dx\right)^{\frac16}\\ & \leq C\left(\int_{\mathbb{R}^3} e^{-\frac65 a|x+R\theta|}(U(x))^{\frac{12}{5}} dx\right)^{\frac56}\\ & \leq C\left(\int_{\mathbb{R}^3} e^{-{\frac65}aR}e^{({\frac65}a-{\frac{12}5}(1-\varepsilon))|x|} dx\right)^{\frac56} \leq C e^{-aR} \end{array} \end{equation} (28)

    since 0 < a < 2 . We can also prove that

    \begin{equation} \begin{array}{rl} & \int_{\mathbb{R}^3}h(x)U_R^2dx = \int_{\mathbb{R}^3}h(x+R\theta)U^2(x)dx\\ & \geq C\int_{\mathbb{R}^3} e^{-b|x+R\theta|}U^2(x) dx \geq C\int_{\mathbb{R}^3} e^{-b|x| -bR } U^2(x) dx\\ & \geq C e^{-bR} \int_{\mathbb{R}^3} e^{-b|x|} U^2(x)dx \geq C e^{-bR}. \end{array} \end{equation} (29)

    It is now deduced from (28) and (29) that

    \begin{array}{rl} & \sup\limits_{t > 0} I_{\mu_1}(tU_R) \leq \sup\limits_{t > 0} g(t) + C e^{-aR} \\ &\leq \frac{p-1}{2(p+1)} \left(\|U_R\|^2 - \mu_1 \int_{\mathbb{R}^3} h(x) U_R^2 dx\right)^{\frac{p+1}{p-1}}(\|U_R\|_{L^{p+1}}^{-2})^{\frac{p+1}{p-1}} + C e^{-aR} \\ &\leq \frac{p-1}{2(p+1)} S_{p+1}^{\frac{p+1}{p-1}} - Ce^{-bR} + o(e^{-bR}) + C e^{-aR} \\ & < \frac{p-1}{2(p+1)} S_{p+1}^{\frac{p+1}{p-1}} \end{array}

    for R large enough since 0 < b < a . The proof is complete.

    Lemma 3.3. Under the assumptions (A1)-(A4) , I_{\mu_1} satisfies (PS)_d condition for any d < \frac{p-1}{2(p+1)} S_{p+1}^{\frac{p+1}{p-1}} .

    Proof. Let (u_n)_{n\in\mathbb{N}}\subset H^1(\mathbb{R}^3) be a (PS)_d sequence of I_{\mu_1} with d < \frac{p-1}{2(p+1)} S_{p+1}^{\frac{p+1}{p-1}} . Then we have that for n large enough,

    d + o(1) = \frac{1}{2}\|u_n\|^2 - \frac{\mu_1}{2}\int_{\mathbb{R}^3}h(x)u_n^2dx + \frac{1}{4}F(u_n) -\frac{1}{p+1}\int_{\mathbb{R}^3}|u_n|^{p+1}dx

    and

    \langle I'_{\mu_1} (u_n), u_n\rangle = \|u_n\|^2 - \mu_1\int_{\mathbb{R}^3}h(x)u_n^2dx + F(u_n) - \int_{\mathbb{R}^3}|u_n|^{p+1}dx.

    Hence we can deduce that (u_n)_{n\in\mathbb{N}} is bounded in H^1(\mathbb{R}^3) . Going if necessary to a subsequence, we may assume that u_n \rightharpoonup u_0 weakly in H^1(\mathbb{R}^3) and u_n \to u_0 a. e. in \mathbb{R}^3 . Denote w_n : = u_n - u_0 . We then obtain from Brezis-Lieb lemma and Lemma 2.4 that for n large enough,

    \|u_n\|^2 = \|u_0\|^2 + \|w_n\|^2 +o(1),\quad F(u_n) = F(u_0) + F(w_n) +o(1)

    and

    \|u_n\|^{p+1}_{L^{p+1}} = \|u_0\|^{p+1}_{L^{p+1}} + \|w_n\|^{p+1}_{L^{p+1}} +o(1).

    Since \int_{\mathbb{R}^3}h(x)u_n^2dx\to \int_{\mathbb{R}^3}h(x)u_0^2dx as n\to\infty , we deduce that

    \begin{eqnarray} & & d + o(1) = I_{\mu_1}(u_n) = I_{\mu_1}(u_0) + \frac{1}{2}\| w_n\|^2\\ & & \qquad \quad + \frac{1}{4}F(w_n) -\frac{1}{p+1}\int_{\mathbb{R}^3}|w_n|^{p+1}dx. \end{eqnarray} (30)

    From \langle I'_{\mu_1} (u_n), \psi\rangle \to 0 for any \psi\in H^1(\mathbb{R}^3) , one may deduce that I'_{\mu_1} (u_0) = 0 . Therefore

    \|u_0\|^2 - \mu_1\int_{\mathbb{R}^3}h(x)u_0^2dx + F(u_0) = \int_{\mathbb{R}^3}|u_0|^{p+1}dx

    and then

    I_{\mu_1}(u_0) \geq \frac{p-1}{2(p+1)}\left(\|u_0\|^2 - \mu_1 \int_{\mathbb{R}^3}h(x)u_0^2dx\right) + \frac{p-3}{4(p+1)}F(u_0) \geq 0.

    Now using an argument similar to the proof of (16), we obtain that

    \begin{equation} o(1) = \|w_n\|^2 + F(w_n) - \int_{\mathbb{R}^3}|w_n|^{P+1}dx. \end{equation} (31)

    By the relation \|u\|^2 \geq S_{p+1} \|u\|_{L^{p+1}}^2 for any u\in H^1(\mathbb{R}^3) , we proceed our discussion according to the following two cases:

    \bf (I) \int_{\mathbb{R}^3}|w_n|^{p+1}dx\not\to 0 as n\to\infty ;

    \bf (II) \int_{\mathbb{R}^3}|w_n|^{p+1}dx \to 0 as n\to\infty .

    Suppose that the case (I) occurs. Then up to a sbusequence, we may obtain from (31) that

    \|w_n\|^2 \geq S_{p+1} \left(\| w_n\|^2 + F(w_n) - o(1)\right)^{\frac2{p+1}},

    which implies that for n large enough,

    \|w_n\|^2 \geq S_{p+1}^{\frac{p+1}{p-1}} + o(1).

    It is deduced from this and (30) that d \geq \frac{p-1}{2(p+1)} S_{p+1}^{\frac{p+1}{p-1}}, which is a contradiction. Therefore the case (II) must occur. This and (31) imply that \|w_n\| \to 0 . Hence we have proven that I_{\mu_1} satisfies (PS)_d condition for any d < \frac{p-1}{2(p+1)} S_{p+1}^{\frac{p+1}{p-1}} .

    Proof of Proposition 3.1. Since 0 is a local minimizer of I_{\mu_1} and for v\neq 0 , I_{\mu_1}(s v) \to -\infty as s\to +\infty , Lemma 3.2, Lemma 3.3 and the mountain pass theorem [3] imply that d_{\mu_1} is a critical value of I_{\mu_1} .

    Proof of Theorem 1.1. By Proposition 3.1, the d_{\mu_1} is a critical value of I_{\mu_1} and d_{\mu_1} > 0 . The proof of nonnegativity for at least one of the corresponding critical point is inspired by the idea of [1]. In fact, since I_{\mu_1}(u) = I_{\mu_1}(|u|) for any u \in H^1(\mathbb{R}^3) , for every n\in \mathbb{N} , there exists \gamma_n\in \Gamma_1 with \gamma_n(t)\geq 0 (a.e. in \mathbb{R}^3 ) for all t\in [0, 1] such that

    \begin{equation} d_{\mu_1} \leq \max\limits_{t\in [0, 1]} I_{\mu_1}(\gamma_n(t)) < d_{\mu_1} + \frac{1}{n}. \end{equation} (32)

    By Ekeland's variational principle [5], there exists \gamma_n^*\in \Gamma_1 satisfying

    \begin{equation} \left\{ \begin{array}{ll} d_{\mu_1} \leq \max\limits_{t\in [0, 1]} I_{\mu_1} (\gamma_n^*(t))\leq \max\limits_{t\in [0, 1]} I_{\mu_1} (\gamma_n(t)) < d_{\mu_1} +\frac{1}{n};\\ \max\limits_{t\in [0, 1]}\|\gamma_n(t)-\gamma_n^*(t)\| < \frac{1}{\sqrt{n}}; \\ \hbox{ there exists} \; t_n\in [0, 1]\; \hbox{such that}\; z_n = \gamma_n^*(t_n) \hbox{ satisfies}: \\ I_{\mu_1}(z_n) = \max\limits_{t\in [0, 1]} I_{\mu_1}(\gamma_n^*(t)), \;\hbox{and}\; \|I'_{\mu_1}(z_n)\|\leq\frac{1}{\sqrt{n}}. \end{array} \right. \end{equation} (33)

    By Lemma 3.2 and Lemma 3.3 we get a convergent subsequence (still denoted by (z_n)_{n\in \mathbb{N}} ). We may assume that z_n\rightarrow z in H^1(\mathbb{R}^3) as n\rightarrow \infty . On the other hand, by (33), we also arrive at \gamma_n(t_n)\rightarrow z in H^1(\mathbb{R}^3) as n\rightarrow \infty . Since \gamma_n(t)\geq 0 , we conclude that z\geq 0 , z\not\equiv 0 in \mathbb{R}^3 with I_{\mu_1}(z)>0 and it is a nonnegative bound state of (3) in the case of \mu = \mu_1 .

    In this section, we always assume the conditions (A1)-(A4) . We will prove the existence of ground state and bound states of (3) as well as their asymptotical behavior with respect to \mu . We emphasize that if 0 < \mu < \mu_1 , then one may consider a minimization problem like

    \inf\{ I_\mu(u)\ : \ u\in \mathcal{M}\},\quad \mathcal{M} = \{u\in H^1(\mathbb{R}^3)\ :\ \langle I'_\mu(u), u\rangle = 0\}

    to get a ground state solution. But for \mu \geq \mu_1 , we can not do like this because for \mu > \mu_1 , we do not know if 0 \not\in \partial\mathcal{M} . To overcome this difficulty, we define the set of all nontrivial critical points of I_\mu in H^1(\mathbb{R}^3) :

    \mathcal{N} = \{u\in H^1(\mathbb{R}^3)\backslash\{0\}:I_\mu'(u) = 0\}.

    And then we consider the following minimization problem

    \begin{equation} c_{0,\mu} = \inf\{I_\mu(u):u\in\mathcal{N}\}. \end{equation} (34)

    Lemma 4.1. Let \delta_2 and \rho be as in Lemma 2.7 and \mu\in (\mu_1, \mu_1 + \delta_2) . Define the following minimization problem

    d_{0, \mu} = \inf\limits_{\|u\| < \rho} I_\mu (u).

    Then the d_{0, \mu} is achieved by a nonnegative function w_{0,\mu}\in H^1(\mathbb{R}^3) . Moreover this w_{0,\mu} is a nonnegative solution of (3).

    Proof. Firstly, we prove that -\infty < d_{0, \mu} < 0 for \mu\in (\mu_1, \mu_1 + \delta_2) . Keeping the expression of I_\mu(u) in mind, we obtain from the Sobolev inequality that

    \begin{eqnarray} I_\mu(u) & = &\frac12 \|u\|^2-\frac{\mu}{2}\int_{\mathbb{R}^3} h(x)u^2dx + \frac14 F(u) -\frac{1}{p+1}\int_{\mathbb{R}^3}|u|^{p+1}dx \\ &\geq& \frac12 \|u\|^2-\frac{\mu}{2\mu_1}\|u\|^2 - C\|u\|^{p+1} > -\infty \end{eqnarray}

    as \|u\| < \rho . Next, for any t > 0 , we have that

    I_\mu(te_1) = \frac{t^2}2 \|e_1\|^2-\frac{\mu t^2}{2}\int_{\mathbb{R}^3} h(x)e_1^2dx + \frac{t^4}4 F(e_1) - \frac{t^{p+1}}{p+1}\int_{\mathbb{R}^3}|e_1|^{p+1}dx.

    It is now deduced from \mu_1\int_{\mathbb{R}^3} h(x)e_1^2dx = \|e_1\|^2 that

    I_\mu(te_1) = \frac{t^2}2\left(1-\frac{\mu}{\mu_1}\right)\|e_1\|^2 + \frac{t^4}4 F(e_1) -\frac{t^{p+1}}{p+1}\int_{\mathbb{R}^3}|e_1|^{p+1}dx.

    Since \mu > \mu_1 , we obtain that for t small enough, the I_\mu(t e_1) < 0 . Thus we have proven that -\infty < d_{0, \mu} < 0 for \mu\in (\mu_1, \mu_1 + \delta_2) .

    Secondly, let (v_n)_{n\in\mathbb{N}} be a minimizing sequence, that is, \|v_n\| < \rho and I_\mu(v_n)\to d_{0, \mu} as n\to\infty . By the Ekeland's variational principle, we can obtain that there is a sequence (u_n)_{n\in\mathbb{N}}\subset H^1(\mathbb{R}^3) with \|u_n\| < \rho such that as n\to\infty ,

    I_\mu(u_n) \to d_{0, \mu}\qquad \hbox{and}\qquad I'_\mu(u_n) \to 0.

    Then we can prove that (u_n)_{n\in\mathbb{N}} is bounded in H^1(\mathbb{R}^3) . Using Lemma 2.6, we obtain that (u_n)_{n\in\mathbb{N}} contains a convergent subsequence, still denoted by (u_n)_{n\in\mathbb{N}} , such that u_n\to u_0 strongly in H^1(\mathbb{R}^3) . Noticing the fact that if (v_n)_{n\in\mathbb{N}} is a minimizing sequence, then (|v_n|)_{n\in\mathbb{N}} is also a minimizing sequence, we may assume that for each n\in \mathbb{N} , the u_n\geq 0 in \mathbb{R}^3 . Therefore we may assume that u_0 \geq 0 in \mathbb{R}^3 . The I'_\mu(u_n) \to 0 and u_n\to u_0 strongly in H^1(\mathbb{R}^3) imply that I_\mu'(u_0) = 0 . Hence choosing w_{0,\mu} \equiv u_0 , we know that w_{0,\mu} is a nonnegative solution of the (3).

    We emphasize that the above lemma does NOT mean that w_{0,\mu} is a ground state of (3). But it does imply that \mathcal{N} \neq\emptyset for any \mu\in (\mu_1, \mu_1 + \delta_2) . Now we are in a position to prove that the c_{0,\mu} defined in (34) can be achieved.

    Lemma 4.2. For \mu\in (\mu_1, \mu_1 + \delta_2) , the c_{0,\mu} is achieved by a nontrivial v_{0, \mu}\in H^1(\mathbb{R}^3) , which is a nontrivial critical point of I_\mu and hence a solution of the (3).

    Proof. By Lemma 4.1, we know that \mathcal{N} \neq\emptyset for \mu\in (\mu_1, \mu_1 + \delta_2) . Hence we have that c_{0,\mu} < 0 . Next we prove that the c_{0,\mu} > -\infty .

    For any u\in \mathcal{N} , since I'_\mu(u) = 0 , then \langle I'_\mu(u), u\rangle = 0 . Then we can deduce that

    I_\mu(u) = I_\mu(u) - \frac{1}{4}\langle I'_\mu(u), u\rangle \geq \frac{1}{4}\|u\|^2 - D(p,h) \mu^{\frac{p+1}{p-1}}.

    Therefore the c_{0,\mu} > -\infty.

    Now let (u_n)_{n\in\mathbb{N}}\subset \mathcal{N} be a sequence such that

    I_\mu(u_n) \to c_{0,\mu}\qquad \hbox{and}\qquad I'_\mu(u_n) = 0.

    Since -\infty < c_{0,\mu} < 0 , we know from Lemma 2.6 that (u_n)_{n\in\mathbb{N}} contains a convergent subsequence in H^1(\mathbb{R}^3) and then we may assume without loss of generality that u_n \to v_0 strongly in H^1(\mathbb{R}^3) . Therefore we have that I_\mu(v_0) = c_{0,\mu} and I'_\mu(v_0) = 0 . Choosing v_{0,\mu} \equiv v_0 and we finish the proof of the Lemma 4.2.

    Next, to analyze further the (PS)_d condition of the functional I_\mu , we have to prove a relation between the minimizer w_{0,\mu} obtained in Lemma 4.1 and the minimizer v_{0,\mu} obtained in Lemma 4.2.

    Lemma 4.3. There exists \delta_3 \in (0, \delta_2] such that for any \mu\in (\mu_1, \mu_1 + \delta_3) , the v_{0,\mu} obtained in Lemma 4.2 can be chosen to coincide the w_{0,\mu} obtained in Lemma 4.1.

    Proof. The proof is divided into two steps. In the first place, for u\neq 0 and I_{\mu_1}'(u) = 0 , we have that

    \|u\|^2 - \mu_1\int_{\mathbb{R}^3} h(x)u^2dx + F(u) = \int_{\mathbb{R}^3} |u|^{p+1}dx

    and hence

    I_{\mu_1}(u) = \frac{p-1}{2(p+1)} \left(\|u\|^2 - \mu_1\int_{\mathbb{R}^3} h(x)u^2dx\right) + \frac{p-3}{4(p+1)}F(u).

    Since \|u\|^2 \geq \mu_1\int_{\mathbb{R}^3} h(x)u^2dx for any u\in H^1(\mathbb{R}^3) , we obtain that

    I_{\mu_1}(u) \geq \frac{p-3}{4(p+1)}F(u) > 0.

    In the second place, denoted by u_{0,\mu} a ground state obtained in Lemma 4.2. For any sequence \mu^{(n)} > \mu_1 and \mu^{(n)} \to \mu_1 as n\to\infty , we have that u_{0,\mu^{(n)}} satisfies

    I'_{\mu^{(n)}}(u_{0,\mu^{(n)}}) = 0

    and we also have that

    c_{0, \mu^{(n)}} = I_{\mu^{(n)}}(u_{0,\mu^{(n)}}) < 0.

    Hence we deduce that (u_{0,\mu^{(n)}})_{n\in \mathbb{N}} is bounded in H^1(\mathbb{R}^3) . Since I'_{\mu^{(n)}}(u_{0,\mu^{(n)}}) = 0 , one also has that

    \begin{array}{rl} I_{\mu^{(n)}}(u_{0,\mu^{(n)}}) & = \frac{p-1}{2(p+1)} \left(\|u_{0,\mu^{(n)}}\|^2 - \mu^{(n)}\int_{\mathbb{R}^3} h(x)(u_{0,\mu^{(n)}})^2dx\right)\\ & \qquad \quad + \frac{p-3}{4(p+1)}F(u_{0,\mu^{(n)}}). \end{array}

    Using the definition of \mu_1 , we obtain that, as n\to \infty ,

    \|u_{0,\mu^{(n)}}\|^2 - \mu^{(n)}\int_{\mathbb{R}^3} h(x)(u_{0,\mu^{(n)}})^2dx \geq \left(1-\frac{\mu^{(n)}}{\mu_1}\right)\|u_{0,\mu^{(n)}}\|^2\to 0

    because (u_{0,\mu^{(n)}})_{n\in \mathbb{N}} is bounded in H^1(\mathbb{R}^3) . Next since (u_{0,\mu^{(n)}})_{n\in\mathbb{N}} is bounded in H^1(\mathbb{R}^3) , we may assume without loss of generality that u_{0,\mu^{(n)}}\rightharpoonup \tilde{u}_0 weakly in H^1(\mathbb{R}^3) .

    Claim. As n\to \infty , the u_{0,\mu^{(n)}} \to \tilde{u}_0 strongly in H^1(\mathbb{R}^3) and \tilde{u}_0 = 0 .

    Proof of the Claim. From u_{0,\mu^{(n)}}\rightharpoonup \tilde{u}_0 weakly in H^1(\mathbb{R}^3) , we may assume that u_{0,\mu^{(n)}}\to \tilde{u}_0 a. e. in \mathbb{R}^3 . Using these and the fact of I'_{\mu^{(n)}}(u_{0,\mu^{(n)}}) = 0 , we deduce that I'_{\mu_1}(\tilde{u}_0) = 0 . Then similar to the proof in Lemma 2.6, we obtain that

    \begin{eqnarray} o(1) + I_{\mu^{(n)}}(u_{0,\mu^{(n)}}) & = & I_{\mu^{(n)}}(\tilde{u}_0) + \frac{1}{2}\|\tilde{w}_n\|^2 \\ & & + \frac{1}{4}F(\tilde{w}_n) - \frac{1}{p+1}\int_{\mathbb{R}^3}|\tilde{w}_n|^{p+1}dx, \end{eqnarray} (35)

    where \tilde{w}_n : = u_{0,\mu^{(n)}} - \tilde{u}_0 .

    Now we distinguish two cases:

    \bf (i) \int_{\mathbb{R}^3}|\tilde{w}_n|^{p+1}dx\not\to 0 as n\to\infty ;

    \bf (ii) \int_{\mathbb{R}^3}|\tilde{w}_n|^{p+1}dx \to 0 as n\to\infty .

    Suppose that the case (ⅰ) occurs. We may deduce from a proof similar to Lemma 2.6 that

    I_{\mu^{(n)}}(u_{0,\mu^{(n)}}) + o(1) \geq I_{\mu_1}(\tilde{u}_0) + \frac{p-1}{2(p+1)} S_{p+1}^{\frac{p+1}{p-1}},

    which is a contradiction because I_{\mu_1}(\tilde{u}_0) > - \frac{p-1}{2(p+1)} S_{p+1}^{\frac{p+1}{p-1}} by Lemma 2.5 and the fact of I_{\mu^{(n)}}(u_{0,\mu^{(n)}}) < 0 . Therefore the case (ii) occurs, which implies that u_{0,\mu^{(n)}} \to \tilde{u}_0 strongly in H^1(\mathbb{R}^3) (the proof is similar to those in Lemma 2.6). From this we also deduce that F(\tilde{w}_n) \to F(\tilde{u}_0).

    Next we prove that \tilde{u}_0 = 0 . Arguing by a contradiction, if \tilde{u}_0 \neq 0 , then we know from I'_{\mu^{(n)}}(u_{0,\mu^{(n)}}) = 0 that

    \liminf\limits_{n\to\infty} I_{\mu^{(n)}}(u_{0,\mu^{(n)}}) \geq \frac{p-3}{4(p+1)} F(\tilde{u}_0) > 0,

    which is also a contradiction since I_{\mu^{(n)}}(u_{0,\mu^{(n)}}) < 0 . Therefore \tilde{u}_0 = 0 .

    Hence there is \delta_3 \in (0, \delta_2] such that for any \mu\in (\mu_1, \mu_1 + \delta_3) , \|u_{0,\mu}\| < \rho , which implies that c_{0,\mu} = d_{0,\mu} . Using Lemma 4.1, we can get a nonnegative ground state of (3), called w_{0,\mu} and c_{0,\mu} = d_{0,\mu} = I_\mu(w_{0,\mu}) . The proof is complete.

    Remark 4.4. The proof of Lemma 4.3 implies that (1) of Theorem 1.2 holds.

    In the following, we are going to prove the existence of another nonnegative bound state solution of (3). To obtain this goal, we have to analyze further the (PS)_d condition of the functional I_\mu .

    Lemma 4.5. Under the assumptions of (A1)-(A4) , if \mu\in (\mu_1, \mu_1 + \delta_3) , then I_\mu satisfies (PS)_d condition for any d < c_{0,\mu} + \frac{p-1}{2(p+1)} S_{p+1}^{\frac{p+1}{p-1}} .

    Proof. Let (u_n)_{n\in\mathbb{N}}\subset H^1(\mathbb{R}^3) be a (PS)_d sequence of I_\mu with d < c_{0,\mu} + \frac{p-1}{2(p+1)}S_{p+1}^{\frac{p+1}{p-1}} . Then we have that for n large enough,

    d + o(1) = \frac{1}{2}\|u_n\|^2 - \frac{\mu}{2}\int_{\mathbb{R}^3}h(x)u_n^2dx + \frac{1}{4}F(u_n) -\frac{1}{p+1}\int_{\mathbb{R}^3}|u_n|^{p+1}dx

    and

    \langle I'_\mu (u_n), u_n\rangle = \|u_n\|^2 - \mu\int_{\mathbb{R}^3}h(x)u_n^2dx + F(u_n) - \int_{\mathbb{R}^3}|u_n|^{p+1}dx.

    Similar to the proof in Lemma 2.3, we can deduce that (u_n)_{n\in\mathbb{N}} is bounded in H^1(\mathbb{R}^3) . Going if necessary to a subsequence, we may assume that u_n \rightharpoonup u_0 weakly in H^1(\mathbb{R}^3) and u_n \to u_0 a. e. in \mathbb{R}^3 . Denote w_n : = u_n - u_0 . We then obtain from Brezis-Lieb lemma and Lemma 2.4 that for n large enough,

    \|u_n\|^2 = \|u_0\|^2 + \|w_n\|^2 +o(1),
    F(u_n) = F(u_0) + F(w_n) +o(1)

    and

    \|u_n\|^{p+1}_{L^{p+1}} = \|u_0\|^{p+1}_{L^{p+1}} + \|w_n\|^{p+1}_{L^{p+1}} +o(1).

    Using Lemma 2.1, we also have that \int_{\mathbb{R}^3}h(x)u_n^2dx\to \int_{\mathbb{R}^3}h(x)u_0^2dx as n\to\infty . Therefore we deduce that

    \begin{eqnarray} & & d + o(1) = I_\mu(u_n) = I_\mu(u_0) + \frac{1}{2}\| w_n\|^2\\ & & \qquad \quad + \frac{1}{4}F(w_n) -\frac{1}{p+1}\int_{\mathbb{R}^3}|w_n|^{p+1}dx. \end{eqnarray} (36)

    Since \langle I'_\mu (u_n), \psi\rangle \to 0 for any \psi\in H^1(\mathbb{R}^3) , we know that I'_\mu (u_0) = 0 . Moreover we have that

    I_\mu(u_0) \geq c_{0,\mu}

    and

    \|u_0\|^2 - \mu\int_{\mathbb{R}^3}h(x)u_0^2dx + \int_{\mathbb{R}^3}\phi_{u_0}u_0^2 = \int_{\mathbb{R}^3}|u_0|^{p+1}dx.

    Note that (u_n)_{n\in\mathbb{N}} is bounded in H^1(\mathbb{R}^3) . The Brezis-Lieb lemma, Lemma 2.4 and

    o(1) = \|u_n\|^2 - \mu\int_{\mathbb{R}^3}h(x)u_n^2dx + F(u_n) - \int_{\mathbb{R}^3}|u_n|^{p+1}dx

    imply that

    \begin{equation} o(1) = \| w_n\|^2 + F(w_n) -\int_{\mathbb{R}^3}|w_n|^{p+1}dx. \end{equation} (37)

    Using \|u\|^2 \geq S_{p+1} \|u\|^2_{L^{p+1}} for any u\in H^1(\mathbb{R}^3) , we distinguish two cases:

    \bf (I) \int_{\mathbb{R}^3}|w_n|^{p+1}dx\not\to 0 as n\to\infty ;

    \bf (II) \int_{\mathbb{R}^3}|w_n|^{p+1}dx \to 0 as n\to\infty .

    Suppose (I) occurs. Up to a subsequence, we may obtain from (37) that

    \|w_n\|^2 \geq S_{p+1} \left(\|w_n\|^2 + F(w_n) - o(1)\right)^{\frac2{p+1}}.

    Hence we get that for n large enough,

    \begin{equation} \|w_n\|^2 \geq S_{p+1}^{\frac{p+1}{p-1}} + o(1). \end{equation} (38)

    Therefore using (36) and (38), we deduce that for n large enough,

    \begin{eqnarray} & d & + o(1) = I_\mu(u_n) \\ & = & I_\mu(u_0) + \frac{1}{2}\|w_n\|^2 + \frac{1}{4}F(w_n) -\frac{1}{p+1}\int_{\mathbb{R}^3}|w_n|^{p+1}dx \\ & = & I_\mu(u_0) + \frac{p-1}{2(p+1)}\|w_n\|^2 + \frac{p-3}{4(p+1)}F(w_n) \\ & \geq & c_{0,\mu} + \frac{p-1}{2(p+1)}\|w_n\|^2 + \frac{p-3}{4(p+1)}F(w_n) \\ & > & c_{0,\mu} + \frac{p-1}{2(p+1)} S_{p+1}^{\frac{p+1}{p-1}}, \end{eqnarray} (39)

    which contradicts to the assumption d < c_{0,\mu} + \frac{p-1}{2(p+1)} S_{p+1}^{\frac{p+1}{p-1}} . Therefore the case (Ⅱ) must occur, i.e., \int_{\mathbb{R}^3}|w_n|^{p+1}dx\to 0 as n\to\infty . This and (37) imply that \|w_n\| \to 0 . Hence we have proven that I_\mu satisfies (PS)_d condition for any d < c_{0,\mu} + \frac{p-1}{2(p+1)} S_{p+1}^{\frac{p+1}{p-1}} .

    Next, for the w_{0,\mu} obtained in Lemma 4.3, we define

    d_{2, \mu} = \inf\limits_{\gamma \in \Gamma_2}\sup\limits_{t\in [0,1]}I_\mu(\gamma(t))

    with

    \Gamma_2 = \{ \gamma \in C([0,1], H^1(\mathbb{R}^3))\ : \ \gamma(0) = w_{0, \mu},\ I_\mu(\gamma(1)) < c_{0,\mu} \}.

    Lemma 4.6. Suppose that the conditions (A1)-(A4) hold and 0<b<a<1 . If \mu\in (\mu_1, \mu_1 + \delta_3) , then

    d_{2, \mu} < c_{0,\mu} + \frac{p-1}{2(p+1)} S_{p+1}^{\frac{p+1}{p-1}}.

    Proof. It suffices to find a path starting from w_{0, \mu} and the maximum of the energy functional over this path is strictly less than c_{0,\mu} + \frac{p-1}{2(p+1)} S_{p+1}^{\frac{p+1}{p-1}}. To simplify the notation, we denote w_0 : = w_{0,\mu} , which corresponds to the critical value c_{0,\mu} . We will prove that there is a T_0 such that the path \gamma(t) = w_0 + t T_0U_R is what we need, here U_R(x) \equiv U(x-R\theta) is defined as before. Similar to the discussion in the proof of Lemma 3.2, we only need to estimate I_\mu(w_0 + tU_R) for positive t in a finite interval. By direct calculation, we have that

    \begin{array}{rl} I_\mu(w_0 + tU_R) & = \frac{1}{2}\left(\left\|w_0 + tU_R\right\|^2 - \mu \int_{\mathbb{R}^3} h(x)|w_0 + t U_R|^2dx\right) \\ & \quad + \frac{1}{4}F(w_0 + tU_R) - \frac1{p+1} \int_{\mathbb{R}^3}|w_0 + t U_R|^{p+1}dx \\ & = I_\mu(w_0) + A_1 + A_2 + A_3 +\frac{t^2}{2}\|U_R\|^2 - \frac{\mu}{2}\int_{\mathbb{R}^3}h(x)U_R^2dx, \end{array}

    where

    A_1 = \langle w_0, tU_R\rangle - \mu t\int_{\mathbb{R}^3} h(x)w_0 U_R dx,
    A_2 = \frac{1}{4}\left(F(w_0 + tU_R) - F(w_0)\right)

    and

    A_3 = \frac1{p+1} \int_{\mathbb{R}^3}\left(|w_0|^{p+1} - |w_0 + t U_R|^{p+1}\right)dx.

    Since w_0 is a solution of (3), we have that

    A_1 = \int_{\mathbb{R}^3}(w_0)^p t U_R dx - \int_{\mathbb{R}^3} K(x)\phi_{w_0}w_0 t U_R dx.

    From an elementary inequality:

    (a+b)^q - a^q \geq b^q + q a^{q-1}b,\qquad q > 1,\ \ a > 0, b > 0,

    we deduce that

    |A_3| \leq - \frac1{p+1}\int_{\mathbb{R}^3} |t U_R|^{p+1}dx - \int_{\mathbb{R}^3} |w_0|^p tU_Rdx.

    For the estimate of A_2 , using the expression of F(u) = \int_{\mathbb{R}^3} K(x)\phi_u u^2 dx and the symmetry property of the integral with respect to x and y , we can obtain that

    \begin{array}{rl} |A_2| &\leq t\int_{\mathbb{R}^3}K(x)\phi_{w_0}w_0 U_R dx + \frac{t^2}{2}\int_{\mathbb{R}^3}K(x)\phi_{w_0} (U_R)^2 dx \\ & \qquad \quad + \frac{t^4}{4}\int_{\mathbb{R}^3} K(x)\phi_{U_R}(U_R)^2 dx + t^3\int_{\mathbb{R}^3} K(x)\phi_{U_R}w_0 U_R dx \\ & \qquad \qquad \quad + t^2\int_{\mathbb{R}^3\times\mathbb{R}^3}\frac{K(x)K(y)w_0(x)w_0(y)U_R(x)U_R(y)}{|x-y|}dxdy. \end{array}

    Since w_0 is a nonnegative solution of (3) and w_0\in L^\infty(\mathbb{R}^3) , we obtain from the assumption on K(x) that

    \begin{array}{rl} & \int_{\mathbb{R}^3\times\mathbb{R}^3}\frac{K(x)K(y)w_0(x)w_0(y)U_R(x)U_R(y)}{|x-y|}dxdy \\ & = \int_{\mathbb{R}^3}K(x)\phi_{\sqrt{w_0 U_R}}w_0 U_R dx\\ & \leq \|\phi_{\sqrt{w_0 U_R}}\|_{L^6} \left(\int_{\mathbb{R}^3} K(x)^{\frac65}(w_0U_R)^{\frac65} dx\right)^{\frac56} \\ & \leq C \left(\int_{\mathbb{R}^3} e^{-\frac65 aR} e^{\left(\frac65 a - \frac65(1-\delta)\right)|x|} dx\right)^{\frac56}\\ & \leq C e^{-a R}\quad \hbox{since}\quad 0 < a < 1. \end{array}

    Similarly we can deduce that for R large enough,

    \int_{\mathbb{R}^3}K(x)\phi_{w_0}w_0 U_R dx \leq C e^{-a R}, \ \ \int_{\mathbb{R}^3}K(x)\phi_{w_0} (U_R)^2 dx \leq C e^{-a R},
    \int_{\mathbb{R}^3} K(x)\phi_{U_R}(U_R)^2 dx \leq C e^{-a R}\ \ \hbox{and}\ \ \int_{\mathbb{R}^3} K(x)\phi_{U_R}w_0 U_R dx \leq C e^{-a R}.

    Since \int_{\mathbb{R}^3} h(x) (U_R)^2 dx \geq C e^{-b R} for R large enough, we obtain that

    \begin{array}{rl} & I_\mu(w_0 + tU_R) \leq I_\mu(w_0) +\frac{t^2}{2}\|U_R\|^2dx - \frac{\mu}{2}\int_{\mathbb{R}^3}h(x)U_R^2dx\\ & \qquad \qquad - \frac1{p+1}\int_{\mathbb{R}^3} |t U_R|^{p+1}dx+C e^{-a R}\\ & \leq I_\mu(w_0) + \frac{p-1}{2(p+1)}S_{p+1}^{\frac{p+1}{p-1}} +C e^{-a R} - C e^{-b R} + o(e^{-b R})\\ & < c_{0,\mu} + \frac{p-1}{2(p+1)}S_{p+1}^{\frac{p+1}{p-1}} \end{array}

    for R large enough since 0 < b < a < 1 . The proof is complete.

    Proposition 4.7. Under the conditions (A1)-(A4), if \mu\in (\mu_1, \mu_1 + \delta_3) and w_{0,\mu} be the minimizer obtained in Lemma 4.3, then the d_{2, \mu} is a critical value of I_\mu .

    Proof. Since for \mu\in (\mu_1, \mu_1 + \delta_3) , we know from Lemma 4.1 and Lemma 4.3 that the w_{0, \mu} is a local minimizer of I_\mu . Moreover, one has that I_\mu(w_{0, \mu} + s U_R) \to -\infty as s\to +\infty . Therefore Lemma 4.5, Lemma 4.7 and the mountain pass theorem of [3] imply that d_{2, \mu} is a critical value of I_\mu .

    Proof of Theorem 1.2. The conclusion (1) of Theorem 1.2 follows from Lemma 4.3 and Remark 4.4. It remains to prove (2) of Theorem 1.2. By Proposition 4.7, the d_{2, \mu} is a critical value of I_{\mu} and d_{2, \mu} > 0 . The proof of nonnegativity for at least one of the corresponding critical point is inspired by the idea of [1]. In fact, since I_{\mu}(u) = I_{\mu}(|u|) for any u\in H^1(\mathbb{R}^3) , for every n\in \mathbb{N} , there exists \gamma_n\in \Gamma_2 with \gamma_n(t)\geq 0 (a.e. in \mathbb{R}^3 ) for all t\in [0, 1] such that

    \begin{equation} d_{2, \mu} \leq \max\limits_{t\in [0, 1]} I_{\mu} (\gamma_n(t)) < d_{2, \mu} + \frac{1}{n}. \end{equation} (40)

    By Ekeland's variational principle, there exists \gamma_n^*\in \Gamma_2 satisfying

    \begin{equation} \left\{ \begin{array}{ll} d_{2, \mu} \leq \max\limits_{t\in [0, 1]} I_{\mu} (\gamma_n^*(t))\leq \max\limits_{t\in [0, 1]} I_{\mu} (\gamma_n(t)) < d_{2, \mu} +\frac{1}{n};\\ \max\limits_{t\in [0, 1]}\|\gamma_n(t)-\gamma_n^*(t)\| < \frac{1}{\sqrt{n}}; \\ \hbox{ there exists} \; t_n\in [0, 1]\; \hbox{such that}\; z_n : = \gamma_n^*(t_n) \hbox{ satisfies}: \\ I_{\mu}(z_n) = \max\limits_{t\in [0, 1]} I_{\mu} (\gamma_n^*(t)), \;\hbox{and}\; \|I'_{\mu}(z_n)\|\leq\frac{1}{\sqrt{n}}. \end{array} \right. \end{equation} (41)

    By Lemma 4.6 we get a convergent subsequence (still denoted by (z_n)_{n\in \mathbb{N}} ). We may assume that z_n\rightarrow z strongly in H^1(\mathbb{R}^3) as n\rightarrow \infty . On the other hand, by (41), we also arrive at \gamma_n(t_n)\rightarrow z strongly in H^1(\mathbb{R}^3) as n\rightarrow \infty . Since \gamma_n(t)\geq 0 , we conclude that z\geq 0 , z\not\equiv 0 in \mathbb{R}^3 with I_{\mu}(z)>0 and it is a nonnegative solution of problem (3).

    Next, let u_{2,\mu} be the nonnegative solution given by the above proof, that is, I_{\mu}'(u_{2,\mu}) = 0 and I_{\mu}(u_{2,\mu}) = d_{2, \mu} . We claim that for any sequence \mu^{(n)}> \mu_1 and \mu^{(n)}\to \mu_1 , there exist a sequence of solution u_{2,\mu^{(n)}} of (3) with \mu = \mu^{(n)} and a u_{\mu_1} with I_{\mu_1}'(u_{\mu_1}) = 0 such that u_{2,\mu^{(n)}}\rightarrow u_{\mu_1} strongly in H^1(\mathbb{R}^3) . In fact, denoted by w_{0,\mu^{(n)}} the minimizer corresponding to d_{0, \mu^{(n)}} , according to the definition of d_{2, \mu} and the proof of Lemma 4.6, we deduce that for n large enough,

    0 < \alpha \leq d_{2, \mu^{(n)}}\leq \max\limits_{s > 0}I_{\mu^{(n)}}(w_{0,\mu^{(n)}} + sU_R)

    and

    I_{\mu^{(n)}}(w_{0,\mu^{(n)}} + sU_R) \leq \frac{p-1}{2(p+1)}S_{p+1}^{\frac{p+1}{p-1}} +C e^{-a R} - C e^{-b R} + o(e^{-b R}),
    \begin{equation} \limsup\limits_{n\to\infty} d_{2, \mu^{(n)}} \leq \frac{p-1}{2(p+1)}S_{p+1}^{\frac{p+1}{p-1}}. \end{equation} (42)

    Next, similar to the proof in Lemma 2.3, we can deduce that (u_{2,\mu^{(n)}})_{n\in \mathbb{N}} is bounded in H^1(\mathbb{R}^3) . Going if necessary to a subsequence, we may assume that u_{2,\mu^{(n)}} \rightharpoonup \tilde{u}_2 weakly in H^1(\mathbb{R}^3) and u_{2,\mu^{(n)}} \to \tilde{u}_2 a. e. in \mathbb{R}^3 . Then we have that I'_{\mu_1}(\tilde{u}_2) = 0 . Moreover I_{\mu_1}(\tilde{u}_2) \geq 0 . If (u_{2,\mu^{(n)}})_{n\in \mathbb{N}} does not converge strongly to \tilde{u}_2 in H^1(\mathbb{R}^3) , then using an argument similar to the proof of Lemma 4.5, we may deduce that

    I_{\mu^{(n)}}(u_{2,\mu^{(n)}}) \geq I_{\mu_1}(\tilde{u}_2) + \frac{p-1}{2(p+1)}S_{p+1}^{\frac{p+1}{p-1}},

    which contradicts to (42). Hence u_{2,\mu^{(n)}}\to \tilde{u}_2 strongly in H^1(\mathbb{R}^3) and hence I_{\mu_1}(\tilde{u}_2) > 0 . The proof is complete by choosing u_{\mu_1} = \tilde{u}_2 .

    The author thanks the unknown referee for helpful comments.



    [1] Aloui C, Hkiri B (2014) Co-movements of GCC emerging stock markets: New evidence from wavelet coherence analysis. Econ Model 36: 421–431. doi: 10.1016/j.econmod.2013.09.043
    [2] Aggarwal S, Raja A (2019) Stock market interlinkages among the BRIC economies. Int J Ethics Syst 35: 59–74. doi: 10.1108/IJOES-04-2018-0064
    [3] Bissoondoyal-Bheenick E, Brooks R, et al. (2018) Volatility spillover between the US, Chinese and Australian stock markets. Aus J Manage 43: 263–285. doi: 10.1177/0312896217717305
    [4] Bai S, Cui W, Zhang L (2018) The Granger causality analysis of stocks based on clustering. Cluster Comput, 1–6.
    [5] Cai XJ, Tian S, Yuan N, et al. (2017) Interdependence between oil and East Asian stock markets: Evidence from wavelet coherence analysis. J Int Financ Mark Inst Money 48: 206–223. doi: 10.1016/j.intfin.2017.02.001
    [6] Cappiello L, Engle RF, Sheppard K (2006) Asymmetric dynamics in the correlations of global equity and bond returns. J Financ Econ 4: 537–572.
    [7] Chien M, Lee C, Hu T, et al. (2015) Dynamic Asian stock market convergence: Evidence from dynamic cointegration analysis among China and ASEAN-5. Econ Model 51: 84–98.
    [8] Caporale G, Ali F, Spagnolo N (2015) Oil price uncertainty and sectoral stock returns in China: A time-varying approach. China Econ Rev 34: 311–321.
    [9] Chiang T, Chen X (2016) Empirical analysis of dynamic linkages between China and international stock markets. J Math Financ 6: 189.
    [10] Caporin M, McAleer M (2013) Ten things you should know about the dynamic conditional correlation representation. Econometrics 1: 115–126. doi: 10.3390/econometrics1010115
    [11] Drometer-Oesingmann K (2015) Household Debt in Central and Eastern Europe CEE. In CESifo Forum, München: ifo Institut–Leibniz-Institut für Wirtschaftsforschung an der Universität München. 16: 82–84.
    [12] Engle R (2002) Dynamic conditional correlation: A simple class of multivariate generalized autoregressive conditional heteroskedasticity models. J Bus Econ Stat 20: 339–350 doi: 10.1198/073500102288618487
    [13] Fang L, Bessler D (2018) Is it China that leads the Asian stock market contagion in 2015? Appl Econ Lett 25: 752–757.
    [14] Grinsted A, Moore J, Jevrejeva S(2004) Application of the cross wavelet transform and wavelet coherence to geophysical time series. Nonlinear Process Geophys 11: 561–566.
    [15] Hung N (2019) Return and volatility spillover across equity markets between China and Southeast Asian countries. J Econ Financ Administrative Sci.
    [16] Hung N (2018) Volatility Behaviour of the Foreign Exchange Rate and Transmission Among Central and Eastern European Countries: Evidence from the EGARCH Model. Global Bus Rev.
    [17] Huyghebaert N, Wang L (2010) The co-movement of stock markets in East Asia: Did the 1997–1998 Asian financial crisis really strengthen stock market integration? China Econ Rev 21: 98–112.
    [18] Jain A, Biswal P (2016) Dynamic linkages among oil price, gold price, exchange rate, and stock market in India. Resour Policy 49: 179–185.
    [19] Joyo A, Lefen L (2019) Stock Market Integration of Pakistan with Its Trading Partners: A Multivariate DCC-GARCH Model Approach. Sustainability 11: 303. doi: 10.3390/su11020303
    [20] Kumar D (2014) Return and volatility transmission between gold and stock sectors: Application of portfolio management and hedging effectiveness. IIMB Manage Rev 26: 5–16.
    [21] Kao W, Kao T, Changchien C, et al. (2018) Contagion in international stock markets after the subprime mortgage crisis. Chin Econ 51: 130–153. doi: 10.1080/10971475.2018.1447822
    [22] Khan F, Bangash R, Khan M (2018) Dynamic Linkage among Pakistan, Emerging and Developed Equity Market. FWU J Social Sci 112.
    [23] Li Y, Giles D (2015) Modelling volatility spillover effects between developed stock markets and Asian emerging stock markets. Int J Financ Econ 20: 155–177. doi: 10.1002/ijfe.1506
    [24] Li B, Zeng Z (2018) Time-varying dependence structures of equity markets of China, ASEAN and the USA. Appl Econ Lett 25: 87–91. doi: 10.1080/13504851.2017.1296545
    [25] Lee B (2019) Asian financial market integration and the role of Chinese financial market. Int Rev Econ Financ 59: 490–499. doi: 10.1016/j.iref.2018.10.012
    [26] Lau C, Sheng X (2018) Inter-and intra-regional analysis on spillover effects across international stock markets. Res Int Bus Financ 46: 420–429. doi: 10.1016/j.ribaf.2018.04.013
    [27] Majdoub J, Sassi S (2017) Volatility spillover and hedging effectiveness among China and emerging Asian Islamic equity indexes. Emerg Mark Rev 31: 16–31. doi: 10.1016/j.ememar.2016.12.003
    [28] Nagayev R, Disli M, Inghelbrecht K, et al. (2016) On the dynamic links between commodities and Islamic equity. Energy Econ 58: 125–140. doi: 10.1016/j.eneco.2016.06.011
    [29] Najeeb S, Bacha O, Masih M (2015) Does heterogeneity in investment horizons affect portfolio diversification? Some insights using M-GARCH-DCC and wavelet correlation analysis. Emerg Mark Financ Trade 51: 188–208.
    [30] Prasad N, Grant A, Kim S (2018) Time varying volatility indices and their determinants: Evidence from developed and emerging stock markets. Int Rev Financ Anal 60: 115–126. doi: 10.1016/j.irfa.2018.09.006
    [31] Paramati S, Zakari A, Jalle M, et al. (2018) The dynamic impact of bilateral trade linkages on stock market correlations of Australia and China. Appl Econ Lett 25: 141–145. doi: 10.1080/13504851.2017.1305067
    [32] Raza S, Shah N, Sharif A (2019) Time frequency relationship between energy consumption, economic growth and environmental degradation in the United States: Evidence from transportation sector. Energy.
    [33] Raza S, Sharif A, Wong WK, et al. (2017) Tourism development and environmental degradation in the United States: Evidence from wavelet-based analysis. Curr Issues Tour 20: 1768–1790. doi: 10.1080/13683500.2016.1192587
    [34] Ramsey J, Zhang Z (1996) The application of wave form dictionaries to stock market index data. In Predictability of complex dynamical systems, Springer, 189–205.
    [35] Rizvi S, Arshad S (2014) An empirical study of Islamic equity as a better alternative during crisis using multivariate GARCH DCC. Islamic Econ Stud 130: 1–27.
    [36] Roni B, Abbas G, Wang S (2018) Return and Volatility Spillovers Effects: Study of Asian Emerging Stock Markets. J Syst Sci Inf 6: 97–119.
    [37] Kirkulak Uludag B, Khurshid M (2019) Volatility spillover from the Chinese stock market to E7 and G7 stock markets. J Econ Stud 46: 90–105. doi: 10.1108/JES-01-2017-0014
    [38] Sehgal S, Pandey P, Deisting F (2018) Stock market integration dynamics and its determinants in the East Asian Economic Community Region. J Quant Econ 16: 389–425. doi: 10.1007/s40953-017-0090-7
    [39] Sznajderska A (2019) The role of China in the world economy: evidence from a global VAR model. Appl Econ 51: 1574–1587. doi: 10.1080/00036846.2018.1527464
    [40] Torrence C, Webster P (1999) Interdecadal changes in the ENSO–monsoon system. J clim 12: 2679–2690. doi: 10.1175/1520-0442(1999)012<2679:ICITEM>2.0.CO;2
    [41] Tamakoshi G, Hamori S (2013) An asymmetric dynamic conditional correlation analysis of linkages of European financial institutions during the Greek sovereign debt crisis. Eur J Financ 19: 939–950. doi: 10.1080/1351847X.2012.712921
    [42] Toyoshima Y, Tamakoshi G, Hamori S (2012) Asymmetric dynamics in correlations of treasury and swap markets: Evidence from the US market. J Int Financ Mark Inst Money 22: 381–394. doi: 10.1016/j.intfin.2011.12.002
    [43] Yarovaya L, Brzeszczyński J, Lau C (2016) Volatility spillovers across stock index futures in Asian markets: Evidence from range volatility estimators. Financ Res Lett 17: 158–166. doi: 10.1016/j.frl.2016.03.005
    [44] Yilmaz K (2010) Return and volatility spillovers among the East Asian equity markets. J Asian Econ 21: 304–313. doi: 10.1016/j.asieco.2009.09.001
    [45] Yang L, Cai XJ, Zhang H, et al. (2016) Interdependence of foreign exchange markets: A wavelet coherence analysis. Econ Model 55: 6–14. doi: 10.1016/j.econmod.2016.01.022
    [46] Yang L, Cai XJ, Hamori S (2017). Does the crude oil price influence the exchange rates of oil-importing and oil-exporting countries differently? A wavelet coherence analysis. Int Rev Econ Financ 49: 536–547.
    [47] Zhou X, Zhang W, Zhang J (2012) Volatility spillovers between the Chinese and world equity markets. Pac-Basin Financ J 20: 247–270. doi: 10.1016/j.pacfin.2011.08.002
  • This article has been cited by:

    1. Chao Yang, Sharp condition of global well-posedness for inhomogeneous nonlinear Schrödinger equation, 2021, 14, 1937-1632, 4631, 10.3934/dcdss.2021136
  • Reader Comments
  • © 2019 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(5182) PDF downloads(1049) Cited by(26)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog