This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Existence and Blow-up Profiles of Ground States in Second Order Multi-population Mean-field Games Systems

Fanze Kong , Juncheng Wei and Xiaoyu Zeng Department of Applied Mathematics, University of Washington, Seattle, WA 98195, USA; [email protected] of Mathematics, Chinese University of Hong Kong, Shatin, NT, Hong Kong; [email protected] of Mathematics, Wuhan University of Technology, Wuhan 430070, China; [email protected].
Abstract

In this paper, we utilize the variational structure to study the existence and asymptotic profiles of ground states in multi-population ergodic Mean-field Games systems subject to some local couplings with mass critical exponents. Of concern the attractive and repulsive interactions, we impose some mild conditions on trapping potentials and firstly classify the existence of ground states in terms of intra-population and interaction coefficients. Next, as the intra-population and inter-population coefficients approach some critical values, we show the ground states blow up at one of global minima of potential functions and the corresponding profiles are captured by ground states to potential-free Mean-field Games systems for single population up to translations and rescalings. Moreover, under certain types of potential functions, we establish the refined blow-up profiles of corresponding ground states. In particular, we show that the ground states concentrate at the flattest global minima of potentials.

2020 MSC: 35Q89 (35A15)
Keywords: Multi-population Mean-field Games Systems; Variational Approaches; Constrained Minimization; Blow-up Solutions

1 Introduction

Mean-field Games systems are proposed to describe decision-making among a huge number of indistinguishable rational agents. In real world, various problems involve numerous interacting players, which causes theoretical analysis and even numerical study become impractical. To overcome this issue, Huang et al. [17] and Lasry et al. [18] borrowed the ideas arising from particle physics and introduced Mean-field Games theories and systems independently. For their rich applications in economics, finance, management, etc, we refer the readers to [14].

Focusing on the derivation of Mean-field Games systems, we assume that the ii-th agent with i=1,,ni=1,\cdots,n satisfies the following controlled stochastic differential equation (SDE):

dXti=γtidt+2dBti,X0i=xiN,\displaystyle dX_{t}^{i}=-\gamma^{i}_{t}dt+\sqrt{2}dB_{t}^{i},\ \ X_{0}^{i}=x^{i}\in\mathbb{R}^{N},

where xix^{i} is the initial state, γti\gamma^{i}_{t} denotes the controlled velocity and BtiB_{t}^{i} represent the independent Brownian motion. Suppose all agents are indistinguishable and minimize the following average cost:

J(γt):=𝔼0T[L(γt)+V(Xt)+f(m(Xt))]𝑑t+uT(XT),\displaystyle J(\gamma_{t}):=\mathbb{E}\int_{0}^{T}[L(\gamma_{t})+V(X_{t})+f(m(X_{t}))]dt+u_{T}(X_{T}), (1.1)

where LL is the Lagrangian, VV describes the spatial preference and function ff depends on the population density. By applying the standard dynamic programming principle, the coupled PDE system consisting of Hamilton–Jacobi–Bellman equation and Fokker-Planck equation is formulated, in which the second equation characterize the distribution of the population. The crucial assumption here is all agents are homogeneous and minimize the same cost (1.1). Whereas, in some scenarios, the game processes involve several classes of players with distinct objectives and constraints. Correspondingly, the distributions of games can not be modelled by classical Mean-field Games systems. Motivated by this, multi-population Mean-field Games systems were proposed and the derivations of multi-population stationary problems used to describe Nash equilibria are shown in [13]. For some relevant results of the study of multi-population Mean-field Games systems, we refer the readers to [16, 21, 9, 5, 7]. We also mention that some general theories for the study of Mean-field Games systems can be found in [2, 4, 15, 1]. Recently, with the consideration of a common noise, some researchers established the extended mean-field games systems and discussed some properties such as well-posedness [19, 6].

The objective of this paper is to study the following stationary two-population second order Mean-field Games system:

{Δu1+H(u1)+λ1=V1(x)+f1(m1,m2),xN,Δm1+(m1H(u1))=0,xN,Δu2+H(u2)+λ2=V2(x)+f2(m1,m2),xN,Δm2+(m2H(u2))=0,xN,Nm1𝑑x=Nm2𝑑x=1,\displaystyle\left\{\begin{array}[]{ll}-\Delta u_{1}+H(\nabla u_{1})+\lambda_{1}=V_{1}(x)+f_{1}(m_{1},m_{2}),&x\in\mathbb{R}^{N},\\ \Delta m_{1}+\nabla\cdot(m_{1}\nabla H(\nabla u_{1}))=0,&x\in\mathbb{R}^{N},\\ -\Delta u_{2}+H(\nabla u_{2})+\lambda_{2}=V_{2}(x)+f_{2}(m_{1},m_{2}),&x\in\mathbb{R}^{N},\\ \Delta m_{2}+\nabla\cdot(m_{2}\nabla H(\nabla u_{2}))=0,&x\in\mathbb{R}^{N},\\ \int_{\mathbb{R}^{N}}m_{1}\,dx=\int_{\mathbb{R}^{N}}m_{2}\,dx=1,\end{array}\right. (1.7)

where H:NH:\mathbb{R}^{N}\rightarrow\mathbb{R} is a Hamiltonian, (m1,m2)(m_{1},m_{2}) represents the population density, (u1,u2)(u_{1},u_{2}) denotes the value function and (f1,f2)(f_{1},f_{2}) is the coupling. Here Vi(x)V_{i}(x), i=1,2i=1,2 are potential functions and (λ1,λ2)×(\lambda_{1},\lambda_{2})\in\mathbb{R}\times\mathbb{R} denotes the Lagrange multiplier. In particular, Hamiltonian HH is in general chosen as

H(p)=CH|p|γ with CH>0 and γ>1.\displaystyle H(p)=C_{H}|p|^{\gamma}\text{ with }C_{H}>0\text{ and }\gamma>1. (1.8)

In light of the definition, the corresponding Lagrangian is given by

L=CL|γ|γ,γ=γγ1>1,CL=1γ(γCH)11γ>0.\displaystyle L={C_{L}}|\gamma|^{\gamma^{\prime}},~{}~{}\gamma^{\prime}=\frac{\gamma}{\gamma-1}>1,~{}C_{L}=\frac{1}{\gamma^{\prime}}(\gamma C_{H})^{\frac{1}{1-\gamma}}>0.

From the viewpoint of variational methods, the single population counterpart of (1.7) has been studied intensively when the coupling ff is local and satisfies f=emαf=-em^{\alpha} with constant e>0e>0, see [8, 11, 12]. In detail, there exists a mass critical exponent α=α:=γN\alpha=\alpha^{*}:=\frac{\gamma^{\prime}}{N} such that only when α<α\alpha<\alpha^{*}, the stationary problem admits ground states for any e>0e>0. Moreover, when α=α,\alpha=\alpha^{*}, one can find e>0e^{*}>0 such that the stationary Mean-field Games system has ground states only for e<ee<e^{*} [12]. In this paper, we shall extend the above results into two-species stationary Mean-field Games system (1.7). Similarly as in [12], we consider the mass critical exponent case and define

f1=α1m1γNβm1γ2N12m212+γN,f2=α2m2γNβm2γ2N12m112+γN,\displaystyle f_{1}=-\alpha_{1}m_{1}^{\frac{\gamma^{\prime}}{N}}-\beta m_{1}^{\frac{\gamma^{\prime}}{2N}-\frac{1}{2}}m_{2}^{\frac{1}{2}+\frac{\gamma^{\prime}}{N}},~{}~{}f_{2}=-\alpha_{2}m_{2}^{\frac{\gamma^{\prime}}{N}}-\beta m_{2}^{\frac{\gamma^{\prime}}{2N}-\frac{1}{2}}m_{1}^{\frac{1}{2}+\frac{\gamma^{\prime}}{N}}, (1.9)

where αi>0\alpha_{i}>0, i=1,2i=1,2 and β\beta measure the strengths of intra-population and inter-population interactions, respectively. We shall employ the variational approach to classify the existence of ground states and analyze their asymptotic profiles to (1.7) in terms of αi\alpha_{i}, i=1,2i=1,2 and β.\beta. Noting the forms of nonlinearities shown in (1.9), we assume γ>N\gamma^{\prime}>N here and in the sequel for our analysis; otherwise the strong singularities might cause difficulties for finding ground states to (1.7) while taking limits. It is an intriguing but challenging problem to explore the existence of global minimizers in the case of 1<γN.1<\gamma^{\prime}\leq N.

By employing the variational methods, the existence of ground states to (1.7) is associated with the following constrained minimization problem:

eα1,α2,β=inf(m1,w1,m2,w2)𝒦(m1,w1,m2,w2),\displaystyle e_{\alpha_{1},\alpha_{2},\beta}=\inf_{(m_{1},w_{1},m_{2},w_{2})\in\mathcal{K}}\mathcal{E}(m_{1},w_{1},m_{2},w_{2}), (1.10)

where

α1,α2,β(m1,w1,m2,w2):=\displaystyle\mathcal{E}_{\alpha_{1},\alpha_{2},\beta}(m_{1},w_{1},m_{2},w_{2}):= i=1,2(CLN|wimi|γmi𝑑x+NVimi𝑑xNN+γαiNmi1+γN𝑑x)\displaystyle\sum_{i=1,2}\bigg{(}C_{L}\int_{\mathbb{R}^{N}}\bigg{|}\frac{w_{i}}{m_{i}}\bigg{|}^{\gamma^{\prime}}m_{i}\,dx+\int_{\mathbb{R}^{N}}V_{i}m_{i}\,dx-\frac{N}{N+\gamma^{\prime}}\alpha_{i}\int_{\mathbb{R}^{N}}m_{i}^{1+\frac{\gamma^{\prime}}{N}}\,dx\bigg{)}
2βNN+γNm112+γ2Nm212+γ2N𝑑x,\displaystyle-\frac{2\beta N}{N+\gamma^{\prime}}\int_{\mathbb{R}^{N}}m_{1}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}m_{2}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\,dx, (1.11)

and 𝒦=𝒦1×𝒦2\mathcal{K}=\mathcal{K}_{1}\times\mathcal{K}_{2} with

𝒦i={\displaystyle\mathcal{K}_{i}=\bigg{\{} (mi,wi)|Nmiφdx+Nwiφdx=0,φCc(N),\displaystyle(m_{i},w_{i})\bigg{|}-\int_{\mathbb{R}^{N}}\nabla m_{i}\cdot\nabla\varphi\,dx+\int_{\mathbb{R}^{N}}w_{i}\cdot\nabla\varphi\,dx=0,~{}\forall\varphi\in C_{c}^{\infty}(\mathbb{R}^{N}),
miW1,γ(N),wiL1(N),Nmidx=1,NVimidx<+,mi0 a.e. }\displaystyle m_{i}\in W^{1,\gamma^{\prime}}(\mathbb{R}^{N}),~{}w_{i}\in L^{1}(\mathbb{R}^{N}),~{}\int_{\mathbb{R}^{N}}m_{i}\,dx=1,~{}\int_{\mathbb{R}^{N}}V_{i}m_{i}\,dx<+\infty,~{}m_{i}\geq 0\text{~{}a.e.~{}}\bigg{\}} (1.12)

for i=1,2i=1,2. Due to the technical restriction of our analysis, we impose the following assumptions on potential functions Vi(x)V_{i}(x) with i=1,2i=1,2:

  • (H1).
    infxNVi(x)=0,ViC1(N) and lim|x|+Vi(x)=+;\displaystyle\inf_{x\in\mathbb{R}^{N}}V_{i}(x)=0,~{}V_{i}\in C^{1}(\mathbb{R}^{N})\text{ and }\lim_{|x|\rightarrow+\infty}V_{i}(x)=+\infty; (1.13)
  • (H2).
    lim inf|x|+Vi(x)|x|b>0,lim sup|x|+Vi(x)eδ|x|<+ with constants b>0,δ>0.\displaystyle\liminf_{|x|\rightarrow+\infty}\frac{V_{i}(x)}{|x|^{b}}>0,~{}~{}\limsup_{|x|\rightarrow+\infty}\frac{V_{i}(x)}{e^{\delta|x|}}<+\infty\text{ with constants }b>0,~{}\delta>0. (1.14)

Similarly as shown in [12], the existence of ground states to (1.7) has a strong connection with the following minimization problem for the single species potential-free Mean-field Games System:

(M)γN=inf(m,w)𝒜(NCL|wm|γm𝑑x)(Nm𝑑x)γN11+γNNm1+γN𝑑x,\displaystyle(M^{*})^{\frac{\gamma^{\prime}}{N}}=\inf_{(m,w)\in\mathcal{A}}\frac{\bigg{(}\int_{\mathbb{R}^{N}}C_{L}\big{|}\frac{w}{m}\big{|}^{\gamma^{\prime}}m\,dx\bigg{)}\bigg{(}\int_{\mathbb{R}^{N}}m\,dx\bigg{)}^{\frac{\gamma^{\prime}}{N}}}{\frac{1}{1+\frac{\gamma^{\prime}}{N}}\int_{\mathbb{R}^{N}}m^{1+\frac{\gamma^{\prime}}{N}}\,dx}, (1.15)

where

𝒜:={\displaystyle\mathcal{A}:=\bigg{\{} (m,w)W1,γ(N)L1(N)|Nmφdx+Nwφdx=0,φCc(N),\displaystyle(m,w)\in W^{1,\gamma^{\prime}}(\mathbb{R}^{N})\cap L^{1}(\mathbb{R}^{N})\bigg{|}-\int_{\mathbb{R}^{N}}\nabla m\cdot\nabla\varphi\,dx+\int_{\mathbb{R}^{N}}w\cdot\nabla\varphi\,dx=0,~{}\forall\varphi\in C_{c}^{\infty}(\mathbb{R}^{N}),
0m0,Nm|x|bdx<+ with b>0 given by (1.14}.\displaystyle 0\leq m\not\equiv 0,~{}\int_{\mathbb{R}^{N}}m|x|^{b}\,dx<+\infty\text{~{}with~{}}b>0\text{~{}given by \eqref{Vicondition2}~{}}\bigg{\}}.

We would like to point out that it was shown in Theorem 1.2 [12] that problem (1.15) is attainable and admits at least a minimizer satisfying

{Δu+CH|u|γγNM=mγN,Δm+CHγ(m|u|γ2u)=0,w=CHγm|u|γ2u,Nm𝑑x=M,0<m<Ceδ0|x|,\displaystyle\left\{\begin{array}[]{ll}-\Delta u+C_{H}|\nabla u|^{\gamma}-\frac{\gamma^{\prime}}{NM^{*}}=-m^{\frac{\gamma^{\prime}}{N}},\\ \Delta m+C_{H}\gamma\nabla\cdot(m|\nabla u|^{\gamma-2}\nabla u)=0,~{}w=-C_{H}\gamma m|\nabla u|^{\gamma-2}\nabla u,\\ \int_{\mathbb{R}^{N}}m\,dx=M^{*},~{}0<m<Ce^{-\delta_{0}|x|},\end{array}\right. (1.19)

where δ0>0\delta_{0}>0 is some constant. As a consequence, the following Gagliardo-Nirenberg type’s inequality holds:

NN+γNm1+γN𝑑x1a(CLN|wm|γm𝑑x)(Nm𝑑x)γN(m,w)𝒜,\displaystyle\frac{N}{N+\gamma^{\prime}}\int_{\mathbb{R}^{N}}m^{1+\frac{\gamma^{\prime}}{N}}\,dx\leq\frac{1}{a^{*}}\bigg{(}C_{L}\int_{\mathbb{R}^{N}}\bigg{|}\frac{w}{m}\bigg{|}^{\gamma^{\prime}}m\,dx\bigg{)}\bigg{(}\int_{\mathbb{R}^{N}}m\,dx\bigg{)}^{\frac{\gamma^{\prime}}{N}}\ \forall\ (m,w)\in\mathcal{A}, (1.20)

where a:=(M)γNa^{*}:=(M^{*})^{\frac{\gamma^{\prime}}{N}}. With the aid of (1.20), we shall establish several results for the existence and non-existence of global minimizers to (1.7) and further study the blow-up behaviors of ground states in terms of αi\alpha_{i}, i=1,2i=1,2 and β\beta defined in (1.9). We emphasize that αi>0\alpha_{i}>0, i=1,2i=1,2 represent the self-focusing of the ii-th component and β>0\beta>0 denotes the attractive interaction, while β<0\beta<0 represents the repulsive interaction.

In the next subsection, we shall first state our existence results for attractive and repulsive interactions then discuss the corresponding blow-up profiles results.

1.1 Main Results

Theorem 1.1.

Suppose that Vi(x)V_{i}(x) with i=1,2i=1,2 satisfy (H1) and (H2) given by (1.13) and (1.14), respectively. Define a:=(M)γNa^{*}:=(M^{*})^{\frac{\gamma^{\prime}}{N}} with MM^{*} given in (1.19), then we have

  • (i).

    if 0<α1,α2<a0<\alpha_{1},\alpha_{2}<a^{*} and <β<β:=(aα1)(aα2)-\infty<\beta<\beta_{*}:=\sqrt{(a^{*}-\alpha_{1})(a^{*}-\alpha_{2})}, problem (1.10) has at least one global minimizer (m1,a,w1,a,m2,a,w2,a)𝒦(m_{1,\textbf{a}},w_{1,\textbf{a}},m_{2,\textbf{a}},w_{2,\textbf{a}})\in\mathcal{K}. Correspondingly, there exists a solution (m1,a,m2,a,u1,a,u2,a)W1,p(N)×W1,p(N)×C2(N)×C2(N)(m_{1,\textbf{a}},m_{2,\textbf{a}},u_{1,\textbf{a}},u_{2,\textbf{a}})\in W^{1,p}(\mathbb{R}^{N})\times W^{1,p}(\mathbb{R}^{N})\times C^{2}(\mathbb{R}^{N})\times C^{2}(\mathbb{R}^{N}) with any p>1p>1 and (λ1,a,λ2,a)×(\lambda_{1,\textbf{a}},\lambda_{2,\textbf{a}})\in\mathbb{R}\times\mathbb{R} such that

    {Δu1+CH|u1|γ+λ1=V1(x)α1m1γNβm1γ2N12m212+γN,xN,Δm1+CHγ(m1|u1|γ2u1)=0,xN,Δu2+CH|u2|γ+λ2=V2(x)α2m2γNβm2γ2N12m112+γN,xN,Δm2+CHγ(m2|u2|γ2u2)=0,xN,Nm1𝑑x=Nm2𝑑x=1;\displaystyle\left\{\begin{array}[]{ll}-\Delta u_{1}+C_{H}|\nabla u_{1}|^{\gamma}+\lambda_{1}=V_{1}(x)-\alpha_{1}m_{1}^{\frac{\gamma^{\prime}}{N}}-\beta m_{1}^{\frac{\gamma^{\prime}}{2N}-\frac{1}{2}}m_{2}^{\frac{1}{2}+\frac{\gamma^{\prime}}{N}},&x\in\mathbb{R}^{N},\\ \Delta m_{1}+C_{H}\gamma\nabla\cdot(m_{1}|\nabla u_{1}|^{\gamma-2}\nabla u_{1})=0,&x\in\mathbb{R}^{N},\\ -\Delta u_{2}+C_{H}|\nabla u_{2}|^{\gamma}+\lambda_{2}=V_{2}(x)-\alpha_{2}m_{2}^{\frac{\gamma^{\prime}}{N}}-\beta m_{2}^{\frac{\gamma^{\prime}}{2N}-\frac{1}{2}}m_{1}^{\frac{1}{2}+\frac{\gamma^{\prime}}{N}},&x\in\mathbb{R}^{N},\\ \Delta m_{2}+C_{H}\gamma\nabla\cdot(m_{2}|\nabla u_{2}|^{\gamma-2}\nabla u_{2})=0,&x\in\mathbb{R}^{N},\\ \int_{\mathbb{R}^{N}}m_{1}\,dx=\int_{\mathbb{R}^{N}}m_{2}\,dx=1;\end{array}\right. (1.26)
  • (ii).

    either α1>a\alpha_{1}>a^{*} or α2>a\alpha_{2}>a^{*} or β>β:=2aα1α22,\beta>\beta^{*}:=\frac{2a^{*}-\alpha_{1}-\alpha_{2}}{2}, problem (1.10) has no minimizer.

Theorem 1.1 indicates that when the self-focusing cofficients αi\alpha_{i}, i=1,2i=1,2 are small and the interaction is repulsive, or attractive but with the weak effect, problem (1.10) admits minimizers and correspondingly, there exist classical solutions to (1.26). Whereas, if the self-focusing effects and the attractive interaction are strong, problem (1.10) does not have any minimizer. In fact, there are some gap regions for the existence results shown in Theorem 1.1 since we have ββ\beta^{*}\geq\beta_{*} and the equality holds only when α1=α2\alpha_{1}=\alpha_{2}. It is also an interesting problem to explore the case of αi<a\alpha_{i}<a^{*}, i=1,2i=1,2 and β<β<β\beta_{*}<\beta<\beta^{*}.

Of concern one borderline case β=β=β\beta=\beta_{*}=\beta^{*} with α1=α2<a\alpha_{1}=\alpha_{2}<a^{*} shown in Theorem 1.1, we further obtain

Theorem 1.2.

Assume all conditions in Theorem 1.1 hold and suppose Vi(x)V_{i}(x), i=1,2i=1,2 satisfy

infxN(V1(x)+V2(x))=0.\displaystyle\inf_{x\in\mathbb{R}^{N}}(V_{1}(x)+V_{2}(x))=0. (1.27)

Then if α:=α1=α2<a\alpha:=\alpha_{1}=\alpha_{2}<a^{*} and 0<β=β=β=aα<a,0<\beta=\beta^{*}=\beta_{*}=a^{*}-\alpha<a^{*}, we have problem (1.10) has no minimizer.

Theorem 1.2 demonstrates that when the self-focusing effects are subcritical but the attractive interaction is strong and under critical case, there is no minimizer to problem (1.10). Besides the borderline case discussed in Theorem 1.2, we also study the case of αi=a\alpha_{i}=a^{*} for i=1i=1 or 22 and obtain

Theorem 1.3.

Assume all conditions in Theorem 1.1 hold. If one of the following conditions holds:

  • (i).

    α1=α2=a\alpha_{1}=\alpha_{2}=a^{*} and <β0;-\infty<\beta\leq 0;

  • (ii).

    α1=a\alpha_{1}=a^{*}, 0<α2<a0<\alpha_{2}<a^{*} and 0ββ=aα22,0\leq\beta\leq\beta^{*}=\frac{a^{*}-\alpha_{2}}{2},

then we have problem (1.10) does not admit any minimizer.

Remark 1.1.

We remark that when α2=a\alpha_{2}=a^{*}, 0<α1<a0<\alpha_{1}<a^{*} and 0ββ0\leq\beta\leq\beta^{*}, (1.10) also does not have any minimizer since m1m_{1}-population and m2m_{2}-population are symmetric in (1.7),

Theorem 1.3 shows that if one of self-focusing coefficients are critical, system (1.7) does not admit the ground state. We next summarize results for the study of blow-up profiles of ground states in some singular limits, in which two cases are concerned: attractive interactions with β>0\beta>0 and repulsive ones with β<0\beta<0. Before stating our results, we give some preliminary notations. Define

Zi:={x|Vi(x)=0},i=1,2.\displaystyle Z_{i}:=\{x|V_{i}(x)=0\},i=1,2. (1.28)

For any p>0p>0, we denote

Hm¯,p(y):=N|x+y|pm¯(x)𝑑x, and ν¯p:=inf(m¯,w¯)infyNHm¯,p(y),\displaystyle H_{\bar{m},p}(y):=\int_{\mathbb{R}^{N}}|x+y|^{p}\bar{m}(x)\,dx,\text{ and }\bar{\nu}_{p}:=\inf_{(\bar{m},\bar{w})\in\mathcal{M}}\inf_{y\in\mathbb{R}^{N}}H_{\bar{m},p}(y), (1.29)

with

:={(m¯,w¯)|u such that (m¯,w¯,u) satisfies (1.19) and (m¯,w¯) is a minimizer of (1.15)}.\displaystyle\mathcal{M}:=\{(\bar{m},\bar{w})|\exists u\text{ such that }(\bar{m},\bar{w},u)\text{ satisfies }(\ref{equmpotentialfree})\text{ and }(\bar{m},\bar{w})\text{ is a minimizer of }(\ref{GNinequalitybest})\}. (1.30)

The following two theorems address the attractive case with (α1,α2)(aβ,aβ)(\alpha_{1},\alpha_{2})\nearrow(a^{*}-\beta,a^{*}-\beta) and Z1Z2Z_{1}\cap Z_{2}\not=\emptyset, which are

Theorem 1.4.

Assume that Vi(x)V_{i}(x) satisfies (1.13), (1.14) and Z1Z2Z_{1}\cap Z_{2}\not=\emptyset. Let 0<β<a0<\beta<a^{*}, 0<α1,α2<aβ:=αβ0<\alpha_{1},\alpha_{2}<a^{*}-\beta:=\alpha^{*}_{\beta}, (m1,𝐚,w1,𝐚,m2,𝐚,w2,𝐚)(m_{1,\bf{a}},w_{1,\bf{a}},m_{2,\bf{a}},w_{2,\bf{a}}) be a minimizer of eα1,α2,βe_{\alpha_{1},\alpha_{2},\beta} with 𝐚:=(α𝟏,α𝟐)\bf{a}:=(\alpha_{1},\alpha_{2}) and (m1,𝐚,u1,𝐚,m2,𝐚,u2,𝐚)(m_{1,\bf{a}},u_{1,\bf{a}},m_{2,\bf{a}},u_{2,\bf{a}}) be a solution of (1.26). Define 𝐚β:=(αβ,αβ)=(aβ,aβ),{\bf{a}}^{*}_{\beta}:=(\alpha_{\beta}^{*},\alpha_{\beta}^{*})=(a^{*}-\beta,a^{*}-\beta), then as 𝐚𝐚β\bf{a}\nearrow\bf{a}_{\beta}^{*}, we have for i=1,2,i=1,2,

lim𝐚𝐚β(NCL|wi,𝐚mi,𝐚|γmi,𝐚𝑑xN(αi+β)N+γNmi,𝐚1+γN𝑑x)=0,\displaystyle\lim_{\bf{a}\nearrow\bf{a}_{\beta}^{*}}\bigg{(}\int_{\mathbb{R}^{N}}C_{L}\bigg{|}\frac{w_{i,\bf{a}}}{m_{i,\bf{a}}}\bigg{|}^{\gamma^{\prime}}m_{i,\bf{a}}\,dx-\frac{N(\alpha_{i}+\beta)}{N+\gamma^{\prime}}\int_{\mathbb{R}^{N}}m_{i,\bf{a}}^{1+\frac{\gamma^{\prime}}{N}}\,dx\bigg{)}=0, (1.31)
lim𝐚𝐚βNV1(x)m1,𝐚+V2(x)m2,𝐚dx=0,lim𝐚𝐚βN(m1,𝐚12+γ2Nm2,𝐚12+γ2N)2𝑑x=0,\displaystyle\lim_{\bf{a}\nearrow\bf{a}^{*}_{\beta}}\int_{\mathbb{R}^{N}}V_{1}(x)m_{1,\bf{a}}+V_{2}(x)m_{2,\bf{a}}\,dx=0,~{}\lim_{\bf{a}\nearrow\bf{a}^{*}_{\beta}}\int_{\mathbb{R}^{N}}\bigg{(}m_{1,\bf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}-m_{2,\bf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\bigg{)}^{2}\,dx=0, (1.32)
lim𝐚𝐚βCLN|wi,𝐚mi,𝐚|γmi,𝐚𝑑x+ for both i=1,2\displaystyle~{}\lim_{\bf{a}\nearrow\bf{a}^{*}_{\beta}}C_{L}\int_{\mathbb{R}^{N}}\bigg{|}\frac{w_{i,\bf{a}}}{m_{i,\bf{a}}}\bigg{|}^{\gamma^{\prime}}m_{i,\bf{a}}\,dx\rightarrow+\infty\text{~{}for both~{}}i=1,2 (1.33)

and

lim𝐚𝐚βN|w1,𝐚m1,𝐚|γm1,𝐚𝑑xN|w2,𝐚m2,𝐚|γm2,𝐚𝑑x=1,lim𝐚𝐚βNm1,𝐚1+γN𝑑xNm2,𝐚1+γN𝑑x=1.\displaystyle\lim_{\bf{a}\nearrow\bf{a}^{*}_{\beta}}\frac{\int_{\mathbb{R}^{N}}\big{|}\frac{w_{1,\bf{a}}}{m_{1,\bf{a}}}\big{|}^{\gamma^{\prime}}m_{1,\bf{a}}\,dx}{\int_{\mathbb{R}^{N}}\big{|}\frac{w_{2,\bf{a}}}{m_{2,\bf{a}}}\big{|}^{\gamma^{\prime}}m_{2,\bf{a}}\,dx}=1,~{}~{}\lim_{\bf{a}\nearrow\bf{a}_{\beta}^{*}}\frac{\int_{\mathbb{R}^{N}}m_{1,\bf{a}}^{1+\frac{\gamma^{\prime}}{N}}\,dx}{\int_{\mathbb{R}^{N}}m_{2,\bf{a}}^{1+\frac{\gamma^{\prime}}{N}}\,dx}=1. (1.34)

Moreover, define

ε:=εa:=(CLN|w1,am1,a|γm1,a)1γ0.\displaystyle\varepsilon:=\varepsilon_{\textbf{a}}:=\bigg{(}C_{L}\int_{\mathbb{R}^{N}}\bigg{|}\frac{w_{1,\textbf{a}}}{m_{1,\textbf{a}}}\bigg{|}^{\gamma^{\prime}}m_{1,\textbf{a}}\bigg{)}^{-\frac{1}{\gamma^{\prime}}}\rightarrow 0. (1.35)

Let xi,εx_{i,\varepsilon}, i=1,2i=1,2 be one global minimal point of ui,au_{i,\textbf{a}} and yi,εy_{i,\varepsilon}, i=1,2i=1,2 be one global maximal point of mi,am_{i,\textbf{a}}. Then we have up to a subsequence x0\exists x_{0} s.t. V1(x0)=V2(x0)=0,V_{1}(x_{0})=V_{2}(x_{0})=0, and

xi,ε,yi,εx0, as aaβ;\displaystyle x_{i,\varepsilon},y_{i,\varepsilon}\rightarrow x_{0},\text{ as }\textbf{a}\nearrow\textbf{a}_{\beta}^{*};

moreover, we find

lim supε0+|x1,εx2,ε|ε<+,\displaystyle\limsup_{\varepsilon\rightarrow 0^{+}}\frac{|x_{1,\varepsilon}-x_{2,\varepsilon}|}{\varepsilon}<+\infty, (1.36)

and

lim supε0+|xi,εyj,ε|ε<+,i,j=1,2.\displaystyle\limsup_{\varepsilon\rightarrow 0^{+}}\frac{|x_{i,\varepsilon}-y_{j,\varepsilon}|}{\varepsilon}<+\infty,~{}~{}i,j=1,2. (1.37)

In addition, let

ui,ε:=ε2γγ1ui,a(εx+x1,ε),mi,ε:=εNmi,a(εx+x1,ε),wi,ε:=εN+1wi,a(εx+x1,ε),\displaystyle u_{i,\varepsilon}:=\varepsilon^{\frac{2-\gamma}{\gamma-1}}u_{i,\textbf{a}}(\varepsilon x+x_{1,\varepsilon}),~{}m_{i,\varepsilon}:=\varepsilon^{N}m_{i,\textbf{a}}(\varepsilon x+x_{1,\varepsilon}),~{}{{w}}_{i,\varepsilon}:=\varepsilon^{N+1}w_{i,\textbf{a}}(\varepsilon x+x_{1,\varepsilon}), (1.38)

then there exist uC2(N)u\in C^{2}(\mathbb{R}^{N}), 0mW1,γ(N)0\leq m\in W^{1,\gamma^{\prime}}(\mathbb{R}^{N}), and wLγ(N){{w}}\in L^{\gamma^{\prime}}(\mathbb{R}^{N}) such that

ui,εu in Cloc2(N),mi,εm in Lp(N),p1,wi,εwin Lγ(N),i=1,2.\displaystyle u_{i,\varepsilon}\rightarrow u\text{~{}in~{}}C_{\text{loc}}^{2}(\mathbb{R}^{N}),~{}~{}m_{i,\varepsilon}\rightarrow m\text{~{}in~{}}L^{p}(\mathbb{R}^{N}),~{}\forall p\geq 1,~{}~{}{{w}}_{i,\varepsilon}\rightharpoonup{{w}}~{}\text{in~{}}L^{\gamma^{\prime}}(\mathbb{R}^{N}),~{}i=1,2. (1.39)

In particular, (m,w)(m,{{w}}) is a minimizer of problem (1.15) and (u,m,w)(u,m,{{w}}) solves

{Δu+CH|u|γγN=amγN,xN,Δm+CHγ(m|u|γ2u)=0,w=CHγ|u|γ2u,xN,Nm𝑑x=1.\displaystyle\left\{\begin{array}[]{ll}-\Delta u+C_{H}|\nabla u|^{\gamma}-\frac{\gamma^{\prime}}{N}=-a^{*}m^{\frac{\gamma^{\prime}}{N}},&x\in\mathbb{R}^{N},\\ \Delta m+C_{H}\gamma\nabla\cdot(m|\nabla u|^{\gamma-2}\nabla u)=0,~{}{w}=-C_{H}\gamma|\nabla u|^{\gamma-2}\nabla u,&x\in\mathbb{R}^{N},\\ \int_{\mathbb{R}^{N}}m\,dx=1.\end{array}\right. (1.43)

Theorem 1.4 implies that as (α1,α2)(aβ,aβ)(\alpha_{1},\alpha_{2})\nearrow(a^{*}-\beta,a^{*}-\beta), there are concentration phenomena in the multi-population Mean-field Games system (1.7) with attractive interactions under the mass critical exponent case. In addition, the basic blow-up profiles of ground states are given in Theorem 1.4. Moreover, by imposing the local polynomial expansion on potential functions, we obtain the following results of refined blow-up profiles:

Theorem 1.5.

Assume all conditions in Theorem 1.4 hold. Suppose that V1(x)V_{1}(x) and V2(x)V_{2}(x) have ll common global minimum points, i.e., Z1Z2={x1,,xlN}Z_{1}\cap Z_{2}=\{x_{1},\cdots,x_{l}\in\mathbb{R}^{N}\}, and there exist d>0d>0, aij>0a_{ij}>0, pij>0p_{ij}>0 with i=1,2i=1,2, j=1,,lj=1,\cdots,l such that

Vi=aij|xxj|pij+O(|xxj|pij+1) for 0<|xxj|<d.\displaystyle V_{i}=a_{ij}|x-x_{j}|^{p_{ij}}+O(|x-x_{j}|^{p_{ij}+1})\text{ for }0<|x-x_{j}|<d. (1.44)

Let pj:=min{p1j,p2j}p_{j}:=\min\{p_{1j},p_{2j}\}, p0:=max1jlpjp_{0}:=\max_{1\leq j\leq l}p_{j} and

μj=limxxjV1(x)+V2(x)|xxj|pj=\displaystyle\mu_{j}=\lim_{x\rightarrow x_{j}}\frac{V_{1}(x)+V_{2}(x)}{|x-x_{j}|^{p_{j}}}= {a1j, if p1j<p2j,a1j+a2j, if p1j=p2j,a2j, if p1j>p2j.\displaystyle\left\{\begin{array}[]{ll}a_{1j},\text{ if }p_{1j}<p_{2j},\\ a_{1j}+a_{2j},\text{ if }p_{1j}=p_{2j},\\ a_{2j},\text{ if }p_{1j}>p_{2j}.\end{array}\right. (1.48)

Define

Z0={xj|xjZ¯ and μj=μ} with Z¯:={xj|pj=p0,j=1,,l} and μ:=min{μj|xjZ¯}.\displaystyle Z_{0}=\{x_{j}|x_{j}\in\bar{Z}\text{ and }\mu_{j}=\mu\}\text{ with $\bar{Z}:=\{x_{j}|p_{j}=p_{0},~{}j=1,\cdots,l\}$ and $\mu:=\min\{\mu_{j}|x_{j}\in\bar{Z}\}$.} (1.49)

Let (ui,ε,mi,ε,wi,ε)(u_{i,\varepsilon},m_{i,\varepsilon},w_{i,\varepsilon}), i=1,2i=1,2 be given as (1.38). Then we have

limε0+ε(2γp0μν¯p0a)1γ+p0(aα1+α2+2β2)1γ+p0=1,\displaystyle\lim_{\varepsilon\rightarrow 0^{+}}\frac{\varepsilon}{\big{(}\frac{2\gamma^{\prime}}{p_{0}\mu\bar{\nu}_{p_{0}}a^{*}}\big{)}^{\frac{1}{\gamma^{\prime}+p_{0}}}\big{(}a^{*}-\frac{\alpha_{1}+\alpha_{2}+2\beta}{2}\big{)}^{\frac{1}{\gamma^{\prime}+p_{0}}}}=1, (1.50)

and

x1,εx0εy0 with x0Z0 and y0NsatisfyingHm,p0(y0)=ν¯p0,\displaystyle\frac{x_{1,\varepsilon}-x_{0}}{\varepsilon}\rightarrow y_{0}\text{ with }x_{0}\in Z_{0}\text{ and $y_{0}\in\mathbb{R}^{N}$}satisfyingH_{m,p_{0}}(y_{0})=\bar{\nu}_{p_{0}}, (1.51)

where mm and ν¯p0\bar{\nu}_{p_{0}} are given in (1.39) and (1.29), respectively.

Next, we discuss the blow-up profiles of ground states to (1.7) under repulsive interactions. We remark that on one hand, one has shown in Theorem 1.1 that (1.7) admits ground states when 0<α1,α2<a0<\alpha_{1},\alpha_{2}<a^{*} and β0;\beta\leq 0; on the other hand, Theorem 1.3 indicates that (1.10) does not have any minimizer when α1=α2=a\alpha_{1}=\alpha_{2}=a^{*} and β0\beta\leq 0. Similarly as discussed in the proof of Theorem 1.4, we investigate the concentration phenomena in (1.7) with repulsive interactions and obtain

Theorem 1.6.

Assume that Vi(x)V_{i}(x) with i=1,2i=1,2 satisfy (H1) and (H2) given by (1.13) and (1.14), respectively. Suppose

Z1Z2=,\displaystyle Z_{1}\cap Z_{2}=\emptyset, (1.52)

where Z1Z_{1} and Z2Z_{2} are given by (1.28). Let β<0\beta<0, 0<α1,α2<a0<\alpha_{1},\alpha_{2}<a^{*}, (m1,𝐚,w1,𝐚,m2,𝐚,w2,𝐚)(m_{1,\bf{a}},w_{1,\bf{a}},m_{2,\bf{a}},w_{2,\bf{a}}) be a minimizer of eα1,α2,βe_{\alpha_{1},\alpha_{2},\beta} with 𝐚:=(α𝟏,α𝟐)\bf{a}:=(\alpha_{1},\alpha_{2}) and (m1,𝐚,u1,𝐚,m2,𝐚,u2,𝐚)(m_{1,\bf{a}},u_{1,\bf{a}},m_{2,\bf{a}},u_{2,\bf{a}}) be a solution of (1.26). Define 𝐚:=(a,a),{\bf{a}}^{*}:=(a^{*},a^{*}), then we have as 𝐚𝐚\bf{a}\nearrow\bf{a}^{*},

limaa(NCL|wi,ami,a|γmi,a𝑑x+NVimi,a𝑑xNαiN+γNmi,a1+γN𝑑x)=0;\displaystyle\lim_{\textbf{a}\nearrow\textbf{a}^{*}}\bigg{(}\int_{\mathbb{R}^{N}}C_{L}\bigg{|}\frac{w_{i,\textbf{a}}}{m_{i,\textbf{a}}}\bigg{|}^{\gamma^{\prime}}m_{i,\textbf{a}}\,dx+\int_{\mathbb{R}^{N}}V_{i}m_{i,\textbf{a}}\,dx-\frac{N\alpha_{i}}{N+\gamma^{\prime}}\int_{\mathbb{R}^{N}}m_{i,\textbf{a}}^{1+\frac{\gamma^{\prime}}{N}}\,dx\bigg{)}=0; (1.53)
limaaNm1,a12+γ2Nm2,a12+γ2N𝑑x=0,\displaystyle\lim_{\textbf{a}\nearrow\textbf{a}^{*}}\int_{\mathbb{R}^{N}}m_{1,\textbf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}m_{2,\textbf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\,dx=0, (1.54)

and

NV1m1,a+V2m2,adx0;\displaystyle\int_{\mathbb{R}^{N}}V_{1}m_{1,\textbf{a}}+V_{2}m_{2,\textbf{a}}\,dx\rightarrow 0; (1.55)
CLN|wi,ami,a|γmi,a𝑑x+,Nmi,a1+γN𝑑x+,i=1,2.\displaystyle C_{L}\int_{\mathbb{R}^{N}}\big{|}\frac{w_{i,\textbf{a}}}{m_{i,\textbf{a}}}\big{|}^{\gamma^{\prime}}m_{i,\textbf{a}}\,dx\rightarrow+\infty,~{}~{}\int_{\mathbb{R}^{N}}m_{i,\textbf{a}}^{1+\frac{\gamma^{\prime}}{N}}\,dx\rightarrow+\infty,~{}~{}i=1,2. (1.56)

Moreover, define

ε^i:=(CLN|wi,ami,a|γmi,a𝑑x)1γ0 as aa,i=1,2.\displaystyle\hat{\varepsilon}_{i}:=\bigg{(}C_{L}\int_{\mathbb{R}^{N}}\bigg{|}\frac{w_{i,\textbf{a}}}{m_{i,\textbf{a}}}\bigg{|}^{\gamma^{\prime}}m_{i,\textbf{a}}\,dx\bigg{)}^{-\frac{1}{\gamma^{\prime}}}\rightarrow 0\text{~{}as~{}}\textbf{a}\nearrow\textbf{a}^{*},~{}i=1,2.

Let xi,ε^x_{i,\hat{\varepsilon}}, i=1,2i=1,2 be a global minimum point of ui,𝐚u_{i,\bf{a}} and

mi,ε^=ε^iNmi,a(ε^ix+xi,ε^),wi,ε^=ε^iN+1wi,a(ε^ix+xi,ε^),ui,ε^=ε^i2γγ1ui,a(ε^ix+xi,ε^),\displaystyle m_{i,\hat{\varepsilon}}=\hat{\varepsilon}^{N}_{i}m_{i,\textbf{a}}(\hat{\varepsilon}_{i}x+x_{i,\hat{\varepsilon}}),~{}~{}w_{i,\hat{\varepsilon}}=\hat{\varepsilon}_{i}^{N+1}w_{i,\textbf{a}}(\hat{\varepsilon}_{i}x+x_{i,\hat{\varepsilon}}),~{}u_{i,\hat{\varepsilon}}=\hat{\varepsilon}_{i}^{\frac{2-\gamma}{\gamma-1}}u_{i,\textbf{a}}(\hat{\varepsilon}_{i}x+x_{i,\hat{\varepsilon}}), (1.57)

then there exist (ui,mi,wi)C2(N)×W1,γ(N)×Lγ(N)(u_{i},m_{i},w_{i})\in C^{2}(\mathbb{R}^{N})\times W^{1,\gamma^{\prime}}(\mathbb{R}^{N})\times L^{\gamma^{\prime}}(\mathbb{R}^{N}) with i=1,2i=1,2 such that

ui,εui in Cloc2(N),mi,εmi in Lp(N),p1,wi,εwiin Lγ(N),i=1,2.\displaystyle u_{i,\varepsilon}\rightarrow u_{i}\text{~{}in~{}}C_{\text{loc}}^{2}(\mathbb{R}^{N}),~{}~{}m_{i,\varepsilon}\rightarrow m_{i}\text{~{}in~{}}L^{p}(\mathbb{R}^{N}),~{}\forall p\geq 1,~{}~{}{{w}}_{i,\varepsilon}\rightharpoonup{{w}}_{i}~{}\text{in~{}}L^{\gamma^{\prime}}(\mathbb{R}^{N}),~{}i=1,2. (1.58)

In particular, (mi,ui,wi)(m_{i},u_{i},w_{i}), i=1,2i=1,2 both solve system (1.43).

Remark 1.2.

We point out that unlike the attractive case discussed in Theorem 1.4 and Theorem 1.5, ε^1\hat{\varepsilon}_{1} and ε^2\hat{\varepsilon}_{2} given in Theorem 1.6 both converge to zero but might not be in the same order since β<0\beta<0 and the behaviors of V1V_{1} and V2V_{2} might be distinct around global minimum points locally.

Theorem 1.6 indicates that when the interaction is repulsive, there are concentration phenomena within system (1.7) in some singular limit of parameters α1\alpha_{1}, α2\alpha_{2} and β\beta. Moreover, similarly as the conclusion shown in Theorem 1.5, we explore the refined blow-up profiles and obtain

Theorem 1.7.

Assume all conditions in Theorem 1.6 hold. Suppose that each ViV_{i}, i=1,2i=1,2 has only one global minimum point xix_{i} with x1x2x_{1}\not=x_{2} and there exist d>0d>0, bi>0b_{i}>0 and qi>0q_{i}>0 such that

Vi(x)=bi|xxi|qi+O(|xxi|qi+1) for 0<|xxi|<d.\displaystyle V_{i}(x)=b_{i}|x-x_{i}|^{q_{i}}+O(|x-x_{i}|^{q_{i}+1})\text{ for }0<|x-x_{i}|<d. (1.59)

Define for i=1,2,i=1,2,

ϵ~i:=(aαi)1γ+qi and assume s(0,1] such that ϵ~1=O(ϵ~2s).\displaystyle\tilde{\epsilon}_{i}:=(a^{*}-\alpha_{i})^{\frac{1}{\gamma^{\prime}+q_{i}}}\ \text{ and assume $\exists\ s\in(0,1]$ such that }\tilde{\epsilon}_{1}=O(\tilde{\epsilon}^{s}_{2}). (1.60)

Let (m1,a,w1,a,m2,a,w2,a)(m_{1,\textbf{a}},w_{1,\textbf{a}},m_{2,\textbf{a}},w_{2,\textbf{a}}) be a minimizer of (1.10) and (mi,ε^,wi,ε^,ui,ε^)(m_{i,\hat{\varepsilon}},w_{i,\hat{\varepsilon}},u_{i,\hat{\varepsilon}}) be defined as (1.57). Then we have

xi,ε^xiε^iyi0 such that Hmi,qi(yi0)=ν¯qi,\displaystyle\frac{x_{i,\hat{\varepsilon}}-x_{i}}{\hat{\varepsilon}_{i}}\rightarrow y_{i0}\text{ such that }H_{m_{i},q_{i}}(y_{i0})=\bar{\nu}_{q_{i}}, (1.61)

where mim_{i} and ν¯qi\bar{\nu}_{q_{i}}, i=1,2i=1,2 are given by (1.58) and (1.29), respectively. Moreover, the following asymptotics hold as 𝐚𝐚\bf{a}\nearrow\bf{a}^{*},

ε^iγ=(1+o(1))(γ(aαi)abiν¯qiqi)1γ+qi,i=1,2.\displaystyle\hat{\varepsilon}_{i}^{\gamma^{\prime}}=(1+o(1))\bigg{(}\frac{\gamma^{\prime}(a^{*}-\alpha_{i})}{a^{*}b_{i}\bar{\nu}_{q_{i}}q_{i}}\bigg{)}^{\frac{1}{\gamma^{\prime}+q_{i}}},~{}i=1,2.
Remark 1.3.

In Theorem 1.7, we discuss the refined blow-up profiles of ground states when the interaction coefficient is non-positive under some technical assumption (1.60). We would like to remark that this condition is technical and could be improved if the refined decay estimate of population density mm is given. In fact, the improved condition will be exhibited in Section 5.

The rest of this paper is organized as follows: In Section 2, we give some preliminary results for the existence and properties of the solutions to Hamilton-Jacobi equations and Fokker-Planck equations, which are used to investigate the existence and blow-up behaviors of minimizers to problem (1.10) . Section 3 is devoted to the exploration of the effect of the potentials Vi(x),V_{i}(x), i=1,2i=1,2 and coefficients α1,α2,β\alpha_{1},\alpha_{2},\beta on the existence of minimizers. Correspondingly, the proof of Theorems 1.1-1.3 will be finished. In Section 4, we perform the blow-up analysis of minimizers under the case of attractive interactions β>0\beta>0, and show the conclusions of Theorem 1.4 and Theorem 1.5. Finally, in Section 5, we focus on the asymptotic profiles of ground states with β<0\beta<0 and complete the proof of Theorem 1.6 and Theorem 1.7.

2 Preliminary Results

In this section, we collect some preliminaries for the existence and regularities of solutions to Hamilton-Jacobi equations and Fokker-Planck equations, respectively. Furthermore, some useful equalities and estimates satisfied by the solution to the single population Mean-field Games system will be listed.

2.1 Hamilton-Jacobi Equations

Consider the following second order Hamilton-Jacobi equations:

Δuk+CH|uk|γ+λk=Vk(x)+fk(x),xN,\displaystyle-\Delta u_{k}+C_{H}|\nabla u_{k}|^{\gamma}+\lambda_{k}=V_{k}(x)+f_{k}(x),\ \ x\in\mathbb{R}^{N}, (2.1)

where γ>1\gamma>1 is fixed, CHC_{H} is a given positive constant independent of kk and (uk,λk)(u_{k},\lambda_{k}) denote the solutions to (2.1). For the gradient estimates of uku_{k}, we find

Lemma 2.1.

Suppose that fkL(N)f_{k}\in L^{\infty}(\mathbb{R}^{N}) satisfies fkLCf\|f_{k}\|_{L^{\infty}}\leq C_{f}, |λk|λ|\lambda_{k}|\leq\lambda, and the potential functions Vk(x)Cloc0,θ(N)V_{k}(x)\in C^{0,\theta}_{\rm loc}(\mathbb{R}^{N}) with θ(0,1)\theta\in(0,1) satisfy 0Vk(x)+0\leq V_{k}(x)\rightarrow+\infty as |x|+,|x|\rightarrow+\infty, and R>0\exists~{}R>0 sufficiently large such that

0<C1Vk(x+y)Vk(x)C2, for all k and all |x|R with |y|<2,\displaystyle 0<C_{1}\leq\frac{V_{k}(x+y)}{V_{k}(x)}\leq C_{2},\text{~{}for~{}all~{}}k\text{~{}and~{}all~{}}|x|\geq R\text{~{}with~{}}|y|<2,

where the positive constants CfC_{f}, λ\lambda, RR, C1C_{1} and C2C_{2} are independent of kk. Let (uk,λk)C2(N)×(u_{k},\lambda_{k})\in C^{2}(\mathbb{R}^{N})\times\mathbb{R} be a sequence of solutions to (2.1). Then, for all kk,

|uk(x)|C(1+Vk(x))1γ, for all xN,\displaystyle|\nabla u_{k}(x)|\leq C(1+V_{k}(x))^{\frac{1}{\gamma}},\text{ for all }x\in\mathbb{R}^{N},

where constant CC depends on CHC_{H}, C1C_{1}, C2C_{2}, λ\lambda, γ\gamma, NN and Cf.C_{f}.

In particular, if there exist b0b\geq 0 and CF>0C_{F}>0 independent of k,k, such that following conditions hold on VkV_{k}

CF1(max{|x|CF,0})bVk(x)CF(1+|x|)b,for all k and xN,\displaystyle C_{F}^{-1}(\max\{|x|-C_{F},0\})^{b}\leq V_{k}(x)\leq C_{F}(1+|x|)^{b},~{}~{}\text{for all }k\text{ and }x\in\mathbb{R}^{N}, (2.2)

then we have

|uk|C(1+|x|)bγ,for all k and xN,\displaystyle|\nabla u_{k}|\leq C(1+|x|)^{\frac{b}{\gamma}},~{}\text{for all }k\text{ and }x\in\mathbb{R}^{N},

where constant CC depends on CHC_{H}, CFC_{F}, bb, λ\lambda, γ\gamma, NN and Cf.C_{f}.

Proof.

See Lemma 3.1 in [12] and the argument is the slight modification of the proof of Theorem 2.5 in [8]. ∎

For the lower bound of uku_{k}, we have

Lemma 2.2 (C.f. Lemma 3.2 in [12]).

Suppose all conditions in Lemma 2.1 hold. Let uku_{k} be a family of C2C^{2} solutions and assume that uk(x)u_{k}(x) are bounded from below uniformly. Then there exist positive constants C3C_{3} and C4C_{4} independent of kk such that

uk(x)C3Vk1r(x)C4, xn,for all k.\displaystyle u_{k}(x)\geq C_{3}V^{\frac{1}{r^{\prime}}}_{k}(x)-C_{4},\text{~{}}\forall x\in\mathbb{R}^{n},~{}\text{for all }k. (2.3)

In particular, if the following conditions hold on VkV_{k}

CF1(max{|x|CF,0})bVk(x)CF(1+|x|)b,for all k and xn,\displaystyle C_{F}^{-1}(\max\{|x|-C_{F},0\})^{b}\leq V_{k}(x)\leq C_{F}(1+|x|)^{b},~{}~{}\text{for all }k\text{ and }x\in\mathbb{R}^{n}, (2.4)

where constants b>0b>0 and CFC_{F} are independent of k,k, then we have

uk(x)C3|x|1+brC4, for all k,xn.\displaystyle u_{k}(x)\geq C_{3}|x|^{1+\frac{b}{r^{\prime}}}-C_{4},\text{~{}for all }k,x\in\mathbb{R}^{n}. (2.5)

If b=0b=0 in (2.4) and there exist R>0R>0 and δ^>0\hat{\delta}>0 independent of kk such that

fk+Vkλk>δ^>0 for all |x|>R,\displaystyle f_{k}+V_{k}-\lambda_{k}>\hat{\delta}>0\text{~{}for~{}all }|x|>R,

then (2.5) also holds.

The existence result of the classical solution to (2.1) is summarized as

Lemma 2.3 (C.f. Lemma 3.3 in [12]).

Suppose Vk+fkV_{k}+f_{k} are locally Hölder continuous and bounded from below uniformly in kk. Define

λ¯k:=sup{λ|(2.1) has a solution ukC2(n)}.\displaystyle\bar{\lambda}_{k}:=\sup\{\lambda\in\mathbb{R}~{}|~{}(\ref{HJB-regularity})\text{ has a solution }u_{k}\in C^{2}(\mathbb{R}^{n})\}.

Then

  • (i).

    λ¯k\bar{\lambda}_{k} are finite for every kk and (2.1) admits a solution (uk,λk)C2(n)×(u_{k},\lambda_{k})\in C^{2}(\mathbb{R}^{n})\times\mathbb{R} with λk=λ¯k\lambda_{k}=\bar{\lambda}_{k} and uk(x)u_{k}(x) being bounded from below (may not uniform in kk). Moreover,

    λ¯k=sup{λ|(2.1) has a subsolution ukC2(n)}.\bar{\lambda}_{k}=\sup\{\lambda\in\mathbb{R}~{}|~{}(\ref{HJB-regularity})\text{ has a subsolution }u_{k}\in C^{2}(\mathbb{R}^{n})\}.
  • (ii).

    If VkV_{k} satisfies (2.2) with b>0b>0, then uku_{k} is unique up to constants for fixed kk and there exists a positive constant CC independent of kk such that

    uk(x)C|x|br+1C,xn.\displaystyle u_{k}(x)\geq C|x|^{\frac{b}{r^{\prime}}+1}-C,\forall x\in\mathbb{R}^{n}. (2.6)

    In particular, if Vk0V_{k}\equiv 0 in (2.1) and there exists σ>0\sigma>0 independent of kk such that

    fkλkσ>0,for |x|>K2,\displaystyle f_{k}-\lambda_{k}\geq\sigma>0,\ \ \text{for~{}}|x|>K_{2},

    where K2>0K_{2}>0 is a large constant independent of kk, then (2.6) also holds.

(iii). If VkV_{k} satisfies (1.14), then there exist uniformly bounded from below classical solutions uku_{k} to problem (2.1) satisfying estimate (2.3).

2.2 Fokker-Planck Equations

Of concern the second order Fokker-Planck equation

Δm+w=0,xN,\displaystyle-\Delta m+\nabla\cdot w=0,\ \ x\in\mathbb{R}^{N}, (2.7)

where ww is given and mm denotes the solution, we have the following results for the regularity:

Lemma 2.4.

Let (m,w)(L1(N)W1,q^(N))×L1(N)(m,w)\in\left(L^{1}(\mathbb{R}^{N})\cap W^{1,\hat{q}}(\mathbb{R}^{N})\right)\times L^{1}(\mathbb{R}^{N}) be a solution to (2.7) with

q^:={NNγ+1 if γ<N,(2NN+2,N) if γ=N,γ if γ>N.\hat{q}:=\begin{cases}\frac{N}{N-\gamma^{\prime}+1}&\text{ if }\gamma^{\prime}<N,\\ \in\left(\frac{2N}{N+2},N\right)&\text{ if }\gamma^{\prime}=N,\\ \gamma^{\prime}&\text{ if }\gamma^{\prime}>N.\end{cases}

Assume that

Λγ:=n|m||wm|γ𝑑x<,\Lambda_{\gamma^{\prime}}:=\int_{\mathbb{R}^{n}}|m|\Big{|}\frac{w}{m}\Big{|}^{\gamma^{\prime}}\,dx<\infty,

then we have wL1(N)Lq^(N)w\in L^{1}(\mathbb{R}^{N})\cap L^{\hat{q}}(\mathbb{R}^{N}) and there exists 𝒞=𝒞(Λγ,mL1(N))>0\mathcal{C}=\mathcal{C}(\Lambda_{\gamma^{\prime}},\|m\|_{L^{1}(\mathbb{R}^{N})})>0 such that

mW1,q^(N),wL1(N),wLq^(N)𝒞.\|m\|_{W^{1,\hat{q}}(\mathbb{R}^{N})},\|w\|_{L^{1}(\mathbb{R}^{N})},\|w\|_{L^{\hat{q}}(\mathbb{R}^{N})}\leq\mathcal{C}.
Proof.

See the proof of Lemma 3.5 in [12]. ∎

Next, we state some useful identities satisfied by the single population Mean-field Games system. First of all, we have the exponential decay estimates of mm when some condition is imposed on the Lagrange multiplier, which is

Lemma 2.5 ( C.f. Proposition 5.3 in [8] ).

Assume γ>N\gamma^{\prime}>N. Let (u,λ,m)C2(n)××(W1,γ(n)L1(n))(u,\lambda,m)\in C^{2}(\mathbb{R}^{n})\times\mathbb{R}\times\big{(}W^{1,\gamma^{\prime}}(\mathbb{R}^{n})\cap L^{1}(\mathbb{R}^{n})\big{)} with uu bounded from below, and λ<0\lambda<0 be the solution of the following Mean-field Games system

{Δu+CH|u|γ+λ=mν,xN,Δm+CHγ(m|u|γ2u)=0,xN,\displaystyle\left\{\begin{array}[]{ll}-\Delta u+C_{H}|\nabla u|^{\gamma}+\lambda=-m^{\nu},&x\in\mathbb{R}^{N},\\ \Delta m+C_{H}\gamma\nabla\cdot(m|\nabla u|^{\gamma-2}\nabla u)=0,&x\in\mathbb{R}^{N},\end{array}\right. (2.10)

where ν(0,γN].\nu\in(0,\frac{\gamma^{\prime}}{N}]. Then, we have there exist κ1,κ2>0\kappa_{1},\kappa_{2}>0 such that

m(x)κ1eκ2|x| for all xN.\displaystyle m(x)\leq\kappa_{1}e^{-\kappa_{2}|x|}~{}\text{ for all }x\in\mathbb{R}^{N}.

With the aid of Lemma 2.5, we have the following results for the Pohozaev identities satisfied by the solution to system (2.10):

Lemma 2.6 (C.f. Proposition 3.1 in [10]).

Assume all conditions in Lemma 2.5 hold and denote w=CHγm|u|γ2uw=-C_{H}\gamma m|\nabla u|^{\gamma-2}\nabla u. Then we have the following Pohozaev type identities hold:

{λNm𝑑x=(ν+1)γNν(α+1)γNmν+1𝑑x,CLNm|wm|γ𝑑x=Nν(ν+1)γNmν+1𝑑x=(γ1)CHNm|u|γ𝑑x.\displaystyle\left\{\begin{array}[]{ll}\lambda\int_{\mathbb{R}^{N}}m\,dx=-\frac{(\nu+1)\gamma^{\prime}-N\nu}{(\alpha+1)\gamma^{\prime}}\int_{\mathbb{R}^{N}}m^{\nu+1}\,dx,\\ C_{L}\int_{\mathbb{R}^{N}}m\big{|}\frac{w}{m}\big{|}^{\gamma^{\prime}}\,dx=\frac{N\nu}{(\nu+1)\gamma^{\prime}}\int_{\mathbb{R}^{N}}m^{\nu+1}\,dx=(\gamma-1)C_{H}\int_{\mathbb{R}^{N}}m|\nabla u|^{\gamma}\,dx.\end{array}\right.

3 Existence of ground states

In this section, we shall discuss the existence of ground states to system (1.7) under some conditions of coefficients αi\alpha_{i} with i=1,2i=1,2 and β\beta. To this end, we first estimate the energy α1,α2,β(m1,w1,m2,w2)\mathcal{E}_{\alpha_{1},\alpha_{2},\beta}(m_{1},w_{1},m_{2},w_{2}) from below. Then, if the energy is shown to have some finite lower bound and the minimizers is proved to exist, we will find the existence of ground states to (1.7) by the standard duality argument. Before stating our main results for the existence of minimizers, we give some preliminary definitions, which are

eαii:=inf(m,w)𝒦iαii(m,w),i=1,2,\displaystyle e^{i}_{\alpha_{i}}:=\inf_{(m,w)\in\mathcal{K}_{i}}\mathcal{E}^{i}_{\alpha_{i}}(m,w),~{}i=1,2, (3.1)

where 𝒦i\mathcal{K}_{i} is given by (1) and

αii(m,w)=CLN|wm|γm𝑑x+NVim𝑑xαi1+γNNm1+γN𝑑x.\displaystyle\mathcal{E}_{\alpha_{i}}^{i}(m,w)=C_{L}\int_{\mathbb{R}^{N}}\bigg{|}\frac{w}{m}\bigg{|}^{\gamma^{\prime}}m\,dx+\int_{\mathbb{R}^{N}}V_{i}m\,dx-\frac{\alpha_{i}}{1+\frac{\gamma^{\prime}}{N}}\int_{\mathbb{R}^{N}}m^{1+\frac{\gamma^{\prime}}{N}}\,dx. (3.2)

Concerning the existence of ground states in (1.7), we have

Lemma 3.1.

Assume all conditions in Theorem 1.1 hold, then we have

  • (i).

    if 0<α1<a0<\alpha_{1}<a^{*}, 0<α2<a0<\alpha_{2}<a^{*} and <β<β:=(aα1)(aα2)-\infty<\beta<\beta_{*}:=\sqrt{(a^{*}-\alpha_{1})(a^{*}-\alpha_{2})}, then problem (1.10) has a global minimizer (m1,a,w1,a,m2,a,w2,a)𝒦(m_{1,\textbf{a}},w_{1,\textbf{a}},m_{2,\textbf{a}},w_{2,\textbf{a}})\in\mathcal{K};

  • (ii).

    either α1>a\alpha_{1}>a^{*} or α2>a\alpha_{2}>a^{*} or β>β:=2aα1α22,\beta>\beta^{*}:=\frac{2a^{*}-\alpha_{1}-\alpha_{2}}{2}, then problem (1.10) has no minimizer.

Proof.

(i). Invoking inequality (1.20) and condition (1.13) satisfied by ViV_{i} with i=1,2i=1,2, we have for any (m1,w1,m2,w2)𝒦(m_{1},w_{1},m_{2},w_{2})\in\mathcal{K},

α1,α2,β(m1,w1,m2,w2)\displaystyle\mathcal{E}_{\alpha_{1},\alpha_{2},\beta}(m_{1},w_{1},m_{2},w_{2})
\displaystyle\geq i=12NVimi𝑑x+NN+γ[i=12(aαi)Nmi1+γN𝑑x2βNm112+γ2Nm212+γ2N𝑑x]\displaystyle\sum_{i=1}^{2}\int_{\mathbb{R}^{N}}V_{i}m_{i}\,dx+\frac{N}{N+\gamma^{\prime}}\bigg{[}\sum_{i=1}^{2}(a^{*}-\alpha_{i})\int_{\mathbb{R}^{N}}m_{i}^{1+\frac{\gamma^{\prime}}{N}}\,dx-2\beta\int_{\mathbb{R}^{N}}m_{1}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}m_{2}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\,dx\bigg{]}
\displaystyle\geq 2(ββ)NN+γNm112+γ2Nm212+γ2N𝑑x,\displaystyle\frac{2(\beta_{*}-\beta)N}{N+\gamma^{\prime}}\int_{\mathbb{R}^{N}}m_{1}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}m_{2}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\,dx, (3.3)

where α1,α2,β\mathcal{E}_{\alpha_{1},\alpha_{2},\beta} is given by (1). Then, letting {(m1,k,w1,k,m2,k,w2,k}𝒦\{(m_{1,k},w_{1,k},m_{2,k},w_{2,k}\}\subset\mathcal{K} with k+k\in\mathbb{Z}^{+} being a minimizing sequence of eα1,α2,βe_{\alpha_{1},\alpha_{2},\beta}, one has from (3) and <β<β-\infty<\beta<\beta_{*} that

supki=12NVimi,k𝑑x<+,supkNm1,k12+γ2Nm2,k12+γ2N𝑑x<+,\displaystyle\sup_{k}\sum_{i=1}^{2}\int_{\mathbb{R}^{N}}V_{i}m_{i,k}\,dx<+\infty,~{}~{}\sup_{k}\int_{\mathbb{R}^{N}}m_{1,k}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}m_{2,k}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\,dx<+\infty, (3.4)

and then

supki=12N|wi,kmi,k|γmi,k𝑑x<+.\displaystyle\sup_{k}\sum_{i=1}^{2}\int_{\mathbb{R}^{N}}\bigg{|}\frac{w_{i,k}}{m_{i,k}}\bigg{|}^{\gamma^{\prime}}m_{i,k}\,dx<+\infty. (3.5)

Thanks to Lemma 2.4 and (3.5), one obtains as k+,k\rightarrow+\infty, for i=1,2i=1,2,

(mi,k,wi,k)(mi,a,wi,a) in W1,γ(N)×Lγ(N).\displaystyle(m_{i,k},w_{i,k})\rightharpoonup(m_{i,\textbf{a}},w_{i,\textbf{a}})\text{~{}in~{}}W^{1,\gamma^{\prime}}(\mathbb{R}^{N})\times L^{\gamma^{\prime}}(\mathbb{R}^{N}).

Moreover, by the compactly Sobolev embedding (C.f. Lemma 5.1 in [12]) and Fatou’s lemma, we find from (3.4) that

mi,kmi,a in L1+γN(N)L1(N).\displaystyle m_{i,k}\rightarrow m_{i,\textbf{a}}\text{~{}in~{}}L^{1+\frac{\gamma^{\prime}}{N}}(\mathbb{R}^{N})\cap L^{1}(\mathbb{R}^{N}).

Then it follows that (m1,a,w1,a,m2,a,w2,a)𝒦(m_{1,\textbf{a}},w_{1,\textbf{a}},m_{2,\textbf{a}},w_{2,\textbf{a}})\in\mathcal{K} is a minimizer.

(ii). Let \mathcal{M} be given by (1.30). Since γ>N\gamma^{\prime}>N, by using Morrey’s embedding, the standard elliptic regularity and the maximum principle, one follows the idea shown in [3] then obtain for any (m,w)(m,w)\in\mathcal{M}, m(x)>0m(x)>0 for all xN.x\in\mathbb{R}^{N}. Next, we utilize some rescaled pair of (m0,w0)(m_{0},w_{0})\in\mathcal{M} to analyze the bound of α1,α2,β\mathcal{E}_{\alpha_{1},\alpha_{2},\beta} from below.

Let (m0,w0)(m_{0},w_{0})\in\mathcal{M} and define

(mt,wt)=(tNMm0(t(xx0)),tN+1Mw0(t(xx0))),for t>0 and x0N.\displaystyle(m_{t},w_{t})=\bigg{(}\frac{t^{N}}{M^{*}}m_{0}(t(x-x_{0})),\frac{t^{N+1}}{M^{*}}w_{0}(t(x-x_{0}))\bigg{)},~{}\text{for~{}}t>0\text{~{}and~{}}x_{0}\in\mathbb{R}^{N}. (3.6)

From Lemma 2.5 and Lemma 2.6, we have that

CLN|w0m0|γm0𝑑x=1,Nm01+γN𝑑x=N+γN,Nm0𝑑x=M.\displaystyle C_{L}\int_{\mathbb{R}^{N}}\bigg{|}\frac{w_{0}}{m_{0}}\bigg{|}^{\gamma^{\prime}}m_{0}\,dx=1,~{}~{}\int_{\mathbb{R}^{N}}m_{0}^{1+\frac{\gamma^{\prime}}{N}}\,dx=\frac{N+\gamma^{\prime}}{N},~{}~{}\int_{\mathbb{R}^{N}}m_{0}\,dx=M^{*}. (3.7)

Combining (3.6) with (3.7), one finds

CLN|wtmt|γmt𝑑x=CLtγMN|w0m0|γm0𝑑x=tγM,\displaystyle C_{L}\int_{\mathbb{R}^{N}}\bigg{|}\frac{w_{t}}{m_{t}}\bigg{|}^{\gamma^{\prime}}m_{t}\,dx=C_{L}\frac{t^{\gamma^{\prime}}}{M^{*}}\int_{\mathbb{R}^{N}}\bigg{|}\frac{w_{0}}{m_{0}}\bigg{|}^{\gamma^{\prime}}m_{0}\,dx=\frac{t^{\gamma^{\prime}}}{M^{*}}, (3.8)

and

Nmt1+γN𝑑x=tN(1+γN)(M)1+γNNm01+γN(tx)𝑑x=N+γNtγ(M)1+γN.\displaystyle\int_{\mathbb{R}^{N}}m_{t}^{1+\frac{\gamma^{\prime}}{N}}\,dx=\frac{t^{N(1+\frac{\gamma^{\prime}}{N})}}{(M^{*})^{1+\frac{\gamma^{\prime}}{N}}}\int_{\mathbb{R}^{N}}m^{1+\frac{\gamma^{\prime}}{N}}_{0}(tx)\,dx=\frac{N+\gamma^{\prime}}{N}\frac{t^{\gamma^{\prime}}}{(M^{*})^{1+\frac{\gamma^{\prime}}{N}}}. (3.9)

Then it follows from (3.2), (3.6), (3.8) and (3.9) that

α11(mt,wt)=\displaystyle\mathcal{E}^{1}_{\alpha_{1}}(m_{t},w_{t})= CLN|wtmt|γmt𝑑xNα1N+γNmt1+γN𝑑x+NV1mt𝑑x\displaystyle C_{L}\int_{\mathbb{R}^{N}}\bigg{|}\frac{w_{t}}{m_{t}}\bigg{|}^{\gamma^{\prime}}m_{t}\,dx-\frac{N\alpha_{1}}{N+\gamma^{\prime}}\int_{\mathbb{R}^{N}}m_{t}^{1+\frac{\gamma^{\prime}}{N}}\,dx+\int_{\mathbb{R}^{N}}V_{1}m_{t}\,dx
=\displaystyle= tγM(1α1a)+1MNV(xt+x0)m0𝑑x.\displaystyle\frac{t^{\gamma^{\prime}}}{M^{*}}\bigg{(}1-\frac{\alpha_{1}}{a^{*}}\bigg{)}+\frac{1}{M^{*}}\int_{\mathbb{R}^{N}}V\bigg{(}\frac{x}{t}+x_{0}\bigg{)}m_{0}\,dx. (3.10)

On the other hand, we choose

m¯=eδ1|x|eδ1|x|L1,w¯=m¯ with (m¯,w¯)𝒦2,\displaystyle\bar{m}=\frac{e^{-\delta_{1}|x|}}{\|e^{-\delta_{1}|x|}\|_{L^{1}}},~{}~{}\bar{w}=\nabla\bar{m}\text{ with }(\bar{m},\bar{w})\in\mathcal{K}_{2},

and apply Hölder’s inequality to get

Nmt12+γ2Nm¯12+γ2N𝑑x\displaystyle\int_{\mathbb{R}^{N}}m_{t}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\bar{m}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\,dx\leq (Nmt1+γN𝑑x)12(Nm¯1+γN𝑑x)12Ctγ2,\displaystyle\bigg{(}\int_{\mathbb{R}^{N}}m_{t}^{1+\frac{\gamma^{\prime}}{N}}\,dx\bigg{)}^{\frac{1}{2}}\bigg{(}\int_{\mathbb{R}^{N}}\bar{m}^{1+\frac{\gamma^{\prime}}{N}}\,dx\bigg{)}^{\frac{1}{2}}\leq Ct^{\frac{\gamma^{\prime}}{2}}, (3.11)

where C>0C>0 is some constant. Upon collecting (3) and (3.11), we obtain if α1>a\alpha_{1}>a^{*},

α1,α2,β(mt,wt,m¯,w¯)tγM(1α1a)Ctγ2C, as t+.\displaystyle\mathcal{E}_{\alpha_{1},\alpha_{2},\beta}(m_{t},w_{t},\bar{m},\bar{w})\geq\frac{t^{\gamma^{\prime}}}{M^{*}}\bigg{(}1-\frac{\alpha_{1}}{a^{*}}\bigg{)}-Ct^{\frac{\gamma^{\prime}}{2}}-C\rightarrow-\infty,\text{~{}as }t\rightarrow+\infty.

Thus, eα1,α2,β= when α1>a.e_{\alpha_{1},\alpha_{2},\beta}=-\infty\text{~{}when~{}}\alpha_{1}>a^{*}. Similarly, we find if α2>a,\alpha_{2}>a^{*}, then eα1,α2,β=e_{\alpha_{1},\alpha_{2},\beta}=-\infty. Consequently, we have if any αi>a\alpha_{i}>a^{*} or α2>a\alpha_{2}>a^{*}, problem (1.10) does not have a minimizer.

It is left to study the case of β>β\beta>\beta^{*}. To this end, we compute and obtain

α1,α2,β(mt,wt,mt,wt)=\displaystyle\mathcal{E}_{\alpha_{1},\alpha_{2},\beta}(m_{t},w_{t},m_{t},w_{t})= tγM(2α1aα2a2βa)+O(1), as t+,\displaystyle\frac{t^{\gamma^{\prime}}}{M^{*}}\bigg{(}2-\frac{\alpha_{1}}{a^{*}}-\frac{\alpha_{2}}{a^{*}}-\frac{2\beta}{a^{*}}\bigg{)}+O(1)\rightarrow-\infty,\text{ as }t\rightarrow+\infty,

when β>β:=2aα1α22.\beta>\beta^{*}:=\frac{2a^{*}-\alpha_{1}-\alpha_{2}}{2}. This completes the proof.

Lemma 3.1 states some existence results for the global minimizers (m1,w1,m2,w2)(m_{1},w_{1},m_{2},w_{2}) to (1.10) under some conditions of α1\alpha_{1}, α2\alpha_{2} and β\beta. In particular, when intra-population and inter-population coefficients are all small, Lemma 3.1 implies there exists a minimizer to (1.10). Whereas, the existence of ground states to (1.7) can not be shown unless (u1,u2)(u_{1},u_{2}) and (λ1,λ2)(\lambda_{1},\lambda_{2}) are obtained. Hence, to finish the proof of Theorem 1.1, we establish the following lemma for the existence of the value function pair (u1,u2)(u_{1},u_{2}) and Lagrange multipliers (λ1,λ2)(\lambda_{1},\lambda_{2}):

Lemma 3.2.

Let (m1,a,w1,a,m2,a,w2,a)𝒦(m_{1,\textbf{a}},{{w}}_{1,\textbf{a}},m_{2,\textbf{a}},{{w}}_{2,\textbf{a}})\in\mathcal{K} be a minimizer of eα1,α2,βe_{\alpha_{1},\alpha_{2},\beta} with 𝒦=𝒦1×𝒦2\mathcal{K}=\mathcal{K}_{1}\times\mathcal{K}_{2} defined by (1), then there exist (u1,a,u2,a)(C2(N))2(u_{1,\textbf{a}},u_{2,\textbf{a}})\in\big{(}C^{2}(\mathbb{R}^{N})\big{)}^{2} and (λ1,a,λ2,a)2(\lambda_{1,\textbf{a}},\lambda_{2,\textbf{a}})\in\mathbb{R}^{2} such that (m1,a,u1,a,m2,a,u2,a,λ1,a,λ2,a)(m_{1,\textbf{a}},u_{1,\textbf{a}},m_{2,\textbf{a}},u_{2,\textbf{a}},\lambda_{1,\textbf{a}},\lambda_{2,\textbf{a}}) solves

{Δu1+CH|u1|γ+λ1=V1(x)α1m1γNβm1γ2N12m212+γN,xN,Δm1+w1=0,w1=γCHm1|u1|γ2u1xN,Δu2+CH|u2|γ+λ2=V2(x)α2m2γNβm2γ2N12m112+γN,xN,Δm2+w2=0,w2=γCHm2|u2|γ2u2,xN.\displaystyle\left\{\begin{array}[]{ll}-\Delta u_{1}+C_{H}|\nabla u_{1}|^{\gamma}+\lambda_{1}=V_{1}(x)-\alpha_{1}m_{1}^{\frac{\gamma^{\prime}}{N}}-\beta m_{1}^{\frac{\gamma^{\prime}}{2N}-\frac{1}{2}}m_{2}^{\frac{1}{2}+\frac{\gamma^{\prime}}{N}},&x\in\mathbb{R}^{N},\\ \Delta m_{1}+\nabla\cdot{{w}}_{1}=0,~{}{{w}}_{1}=-\gamma C_{H}m_{1}|\nabla u_{1}|^{\gamma-2}\nabla u_{1}&x\in\mathbb{R}^{N},\\ -\Delta u_{2}+C_{H}|\nabla u_{2}|^{\gamma}+\lambda_{2}=V_{2}(x)-\alpha_{2}m_{2}^{\frac{\gamma^{\prime}}{N}}-\beta m_{2}^{\frac{\gamma^{\prime}}{2N}-\frac{1}{2}}m_{1}^{\frac{1}{2}+\frac{\gamma^{\prime}}{N}},&x\in\mathbb{R}^{N},\\ \Delta m_{2}+\nabla\cdot{{w}}_{2}=0,~{}{{w}}_{2}=-\gamma C_{H}m_{2}|\nabla u_{2}|^{\gamma-2}\nabla u_{2},&x\in\mathbb{R}^{N}.\end{array}\right. (3.16)

Moreover, we have the following identities and estimates hold:

λi,a=CLN|wi,ami,a|γmi,a𝑑x+NVimi,a𝑑xαiNmi,a1+γN𝑑xβNm1,a12+γ2Nm2,a12+γ2N𝑑x,i=1,2,\displaystyle\lambda_{i,\textbf{a}}=C_{L}\int_{\mathbb{R}^{N}}\bigg{|}\frac{w_{i,\textbf{a}}}{m_{i,\textbf{a}}}\bigg{|}^{\gamma^{\prime}}m_{i,\textbf{a}}\,dx+\int_{\mathbb{R}^{N}}V_{i}m_{i,\textbf{a}}\,dx-\alpha_{i}\int_{\mathbb{R}^{N}}m_{i,\textbf{a}}^{1+\frac{\gamma^{\prime}}{N}}\,dx-\beta\int_{\mathbb{R}^{N}}m_{1,\textbf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}m_{2,\textbf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\,dx,~{}i=1,2, (3.17)

and there exists a constant C>0C>0 such that

|ui,a(x)|C(1+Vi1γ(x)),ui,a(x)CVi1γC, for all xN,i=1,2.\displaystyle|\nabla u_{i,\textbf{a}}(x)|\leq C\Big{(}1+V_{i}^{\frac{1}{\gamma}}(x)\Big{)},~{}~{}u_{i,\textbf{a}}(x)\geq CV_{i}^{\frac{1}{\gamma}}-C,\text{ for all }x\in\mathbb{R}^{N},~{}i=1,2. (3.18)
Proof.

To prove this lemma, we follow the approaches employed to show Proposition 3.4 in [8] and make slight modifications. Define admissible sets 𝒜i\mathcal{A}_{i} as

𝒜i={ψC2(N)|lim sup|x||ψ|Vi1γ<+,lim sup|x||Δψ|Vi<+},i=1,2,\displaystyle\mathcal{A}_{i}=\bigg{\{}\psi\in C^{2}(\mathbb{R}^{N})\bigg{|}\limsup_{|x|\rightarrow\infty}\frac{|\nabla\psi|}{V_{i}^{\frac{1}{\gamma}}}<+\infty,~{}\limsup_{|x|\rightarrow\infty}\frac{|\Delta\psi|}{V_{i}}<+\infty~{}\bigg{\}},~{}~{}i=1,2,

then we proceed the similar argument shown in the proof of Proposition 5.1 in [12] and obtain

Nmi,aΔψ𝑑x=Nwi,aψdx,ψ𝒜i,i=1,2.\displaystyle-\int_{\mathbb{R}^{N}}m_{i,\textbf{a}}\Delta\psi\,dx=\int_{\mathbb{R}^{N}}{w}_{i,\textbf{a}}\cdot\nabla\psi\,dx,~{}~{}\forall\psi\in\mathcal{A}_{i},~{}i=1,2. (3.19)

Next, we define

J~1(m,w):=N[CL|wm|γm+[V1(x)+f1(m1,a,m2,a)]m]𝑑x,\displaystyle\tilde{J}_{1}(m,{w}):=\int_{\mathbb{R}^{N}}\bigg{[}C_{L}\bigg{|}\frac{w}{m}\bigg{|}^{\gamma^{\prime}}m+[V_{1}(x)+f_{1}(m_{1,\textbf{a}},m_{2,\textbf{a}})]m\bigg{]}\,dx, (3.20)

where

f1(m1,a,m2,a):=α1m1,aγNβm2,aγ2N+12m1,aγ2N12,\displaystyle f_{1}(m_{1,\textbf{a}},m_{2,\textbf{a}}):=-\alpha_{1}m_{1,\textbf{a}}^{\frac{\gamma^{\prime}}{N}}-\beta m_{2,\textbf{a}}^{\frac{\gamma^{\prime}}{2N}+\frac{1}{2}}m_{1,\textbf{a}}^{\frac{\gamma^{\prime}}{2N}-\frac{1}{2}},

and set

i:={\displaystyle\mathcal{B}_{i}:=\bigg{\{} (m,w)(L1(N)W1,γ(N))×Lγ(N)|NmΔψ𝑑x=Nwψdx,ψ𝒜i,\displaystyle(m,w)\in(L^{1}(\mathbb{R}^{N})\cap W^{1,\gamma^{\prime}}(\mathbb{R}^{N}))\times L^{\gamma^{\prime}}(\mathbb{R}^{N})\bigg{|}-\int_{\mathbb{R}^{N}}m\Delta\psi\,dx=\int_{\mathbb{R}^{N}}{w}\cdot\nabla\psi\,dx~{},\forall\psi\in\mathcal{A}_{i},
m0 a.e. in N,Nmdx=1,NVimdx<+,N|w|Vi1γdx<+},i=1,2.\displaystyle m\geq 0\text{~{}a.e.~{}in~{}}\mathbb{R}^{N},~{}\int_{\mathbb{R}^{N}}m\,dx=1,~{}\int_{\mathbb{R}^{N}}V_{i}m\,dx<+\infty,~{}\int_{\mathbb{R}^{N}}|{w}|V_{i}^{\frac{1}{\gamma^{\prime}}}\,dx<+\infty\bigg{\}},~{}i=1,2.

We have the fact that (m1,a,w1,a,m2,a,w2,a)(m_{1,\textbf{a}},{w}_{1,\textbf{a}},m_{2,\textbf{a}},{w}_{2,\textbf{a}}) is a minimizer of α1,α2,β\mathcal{E}_{\alpha_{1},\alpha_{2},\beta} in 1×2,\mathcal{B}_{1}\times\mathcal{B}_{2}, i.e.

eα1,α2,β:=inf(m1,w1,m2,w2)𝒦1×𝒦2α1,α2,β(m1,w1,m2,w2)=inf(m1,w1,m2,w2)1×2α1,α2,β(m1,w1,m2,w2).\displaystyle e_{\alpha_{1},\alpha_{2},\beta}:=\inf_{(m_{1},{w}_{1},m_{2},{w}_{2})\in\mathcal{K}_{1}\times\mathcal{K}_{2}}\mathcal{E}_{\alpha_{1},\alpha_{2},\beta}(m_{1},{w}_{1},m_{2},{w}_{2})=\inf_{(m_{1},{w}_{1},m_{2},{w}_{2})\in\mathcal{B}_{1}\times\mathcal{B}_{2}}\mathcal{E}_{\alpha_{1},\alpha_{2},\beta}(m_{1},{w}_{1},m_{2},{w}_{2}). (3.21)

Now, we claim

J~1(m1,a,w1,a)=min(m,w)1J~1(m,w),\displaystyle\tilde{J}_{1}(m_{1,\textbf{a}},{w}_{1,\textbf{a}})=\min_{(m,{w})\in\mathcal{B}_{1}}\tilde{J}_{1}(m,w), (3.22)

where J~1\tilde{J}_{1} is defined by (3.20). Indeed, we set

J1(m,w):=α1,α2,β(m,w,m2,a,w2,a):=φ(m,w)+Λ(m)+G~,\displaystyle J_{1}(m,{w}):=\mathcal{E}_{\alpha_{1},\alpha_{2},\beta}(m,{w},m_{2,\textbf{a}},{w}_{2,\textbf{a}}):=\varphi(m,{w})+\Lambda(m)+\tilde{G}, (3.23)

where

φ(m,w):=CLNm|wm|γ𝑑x,\displaystyle\varphi(m,{w}):=C_{L}\int_{\mathbb{R}^{N}}m\bigg{|}\frac{w}{m}\bigg{|}^{\gamma^{\prime}}\,dx,
Λ(m):=Nα1N+γNm1+γN𝑑x+NV1m𝑑x2βNN+γNm12+γ2Nm2,a12+γ2N𝑑x,\displaystyle\Lambda(m):=-\frac{N\alpha_{1}}{N+\gamma^{\prime}}\int_{\mathbb{R}^{N}}m^{1+\frac{\gamma^{\prime}}{N}}\,dx+\int_{\mathbb{R}^{N}}V_{1}m\,dx-\frac{2\beta N}{N+\gamma^{\prime}}\int_{\mathbb{R}^{N}}m^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}m_{2,\textbf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\,dx,

and

G~:=CLN|w2,am2,a|γm2,a𝑑xNα1γ+NNm2,a1+γN𝑑x+NV2m2,a𝑑x.\displaystyle\tilde{G}:=C_{L}\int_{\mathbb{R}^{N}}\bigg{|}\frac{w_{2,\textbf{a}}}{m_{2,\textbf{a}}}\bigg{|}^{\gamma^{\prime}}m_{2,\textbf{a}}\,dx-\frac{N\alpha_{1}}{\gamma^{\prime}+N}\int_{\mathbb{R}^{N}}m_{2,\textbf{a}}^{1+\frac{\gamma^{\prime}}{N}}\,dx+\int_{\mathbb{R}^{N}}V_{2}m_{2,\textbf{a}}\,dx.

For any (m,w)1(m,{w})\in\mathcal{B}_{1}, we define

mλ=λm+(1λ)m1,a,wλ=λw+(1λ)w1,a,0<λ<1,\displaystyle m_{\lambda}=\lambda m+(1-\lambda)m_{1,\textbf{a}},~{}w_{\lambda}=\lambda w+(1-\lambda)w_{1,\textbf{a}},~{}0<\lambda<1,

and have the fact that (mλ,wλ)1(m_{\lambda},w_{\lambda})\in\mathcal{B}_{1}. Thus, by using (3.21) and (3.23), we obtain

J1(mλ,wλ)=α1,α2,β(mλ,wλ,m2,a,w2,a)α1,α2,β(m1,a,w1,a,m2,a,w2,a)=J1(m1,a,w1,a),\displaystyle J_{1}(m_{\lambda},{w}_{\lambda})=\mathcal{E}_{\alpha_{1},\alpha_{2},\beta}(m_{\lambda},{w}_{\lambda},m_{2,\textbf{a}},{w}_{2,\textbf{a}})\geq\mathcal{E}_{\alpha_{1},\alpha_{2},\beta}(m_{1,\textbf{a}},{w}_{1,\textbf{a}},m_{2,\textbf{a}},{w}_{2,\textbf{a}})=J_{1}(m_{1,\textbf{a}},{w}_{1,\textbf{a}}),

which implies

φ(mλ,wλ)+Λ(mλ)φ(m1,a,w1,a)+Λ(m1,a),\displaystyle\varphi(m_{\lambda},{w}_{\lambda})+\Lambda(m_{\lambda})\geq\varphi(m_{1,\textbf{a}},{w}_{1,\textbf{a}})+\Lambda(m_{1,\textbf{a}}),

i.e.

φ(mλ,wλ)φ(m1,a,w1,a)Λ(m1,a)Λ(mλ).\displaystyle\varphi(m_{\lambda},{w}_{\lambda})-\varphi(m_{1,\textbf{a}},{w}_{1,\textbf{a}})\geq\Lambda(m_{1,\textbf{a}})-\Lambda(m_{\lambda}). (3.24)

Next, we simplitfy (3.24). On one hand, by the convexity of φ\varphi in (m,w)(m,w), we have

φ(mλ,wλ)λφ(m,w)+(1λ)φ(m1,a,w1,a),\displaystyle\varphi(m_{\lambda},{w}_{\lambda})\leq\lambda\varphi(m,{w})+(1-\lambda)\varphi(m_{1,\textbf{a}},{w}_{1,\textbf{a}}),

i.e.

φ(mλ,wλ)φ(m1,a,w1,a)λ[φ(m,w)φ(m1,a,w1,a)].\displaystyle\varphi(m_{\lambda},{w}_{\lambda})-\varphi(m_{1,\textbf{a}},{w}_{1,\textbf{a}})\leq\lambda[\varphi(m,{w})-\varphi(m_{1,\textbf{a}},{w}_{1,\textbf{a}})]. (3.25)

On the other hand, for λ>0\lambda>0 sufficiently small, we have

Λ(mλ)=Λ(m1,a)+λΛ(m1,a),(mm1,a)+O(λ).\displaystyle\Lambda(m_{\lambda})=\Lambda(m_{1,\textbf{a}})+\lambda\langle\nabla\Lambda(m_{1,\textbf{a}}),(m-m_{1,\textbf{a}})\rangle+O(\lambda). (3.26)

In addition, invoking (3.20) and (3.23), one can obtain

Λ(m1,a)=V1+f1(m1,a,m2,a).\displaystyle\nabla\Lambda(m_{1,\textbf{a}})=V_{1}+f_{1}(m_{1,\textbf{a}},m_{2,\textbf{a}}).

Upon substituting (3.25) and (3.26) into (3.24), we get

φ(m,w)φ(m1,a,w1,a)Λ(m1,a),mm1,a.\displaystyle\varphi(m,{w})-\varphi(m_{1,\textbf{a}},{w}_{1,\textbf{a}})\geq-\langle\nabla\Lambda(m_{1,\textbf{a}}),m-m_{1,\textbf{a}}\rangle.

Hence,

J~1(m,w)=φ(m,w)+Λ(m1,a),mφ(m1,a,w1,a)+Λ(m1,a),m1,a=J~1(m1,a,w1,a),\displaystyle\tilde{J}_{1}(m,{w})=\varphi(m,{w})+\langle\nabla\Lambda(m_{1,\textbf{a}}),m\rangle\geq\varphi(m_{1,\textbf{a}},{w}_{1,\textbf{a}})+\langle\nabla\Lambda(m_{1,\textbf{a}}),m_{1,\textbf{a}}\rangle=\tilde{J}_{1}(m_{1,\textbf{a}},{w}_{1,\textbf{a}}),

which indicates that claim (3.22) holds.

Now, we prove

sup{λ:Δψ+CH|ψ|γ+λV1+f1(m1,a,m2,a) in N for some ψ1}=min(m,w)1J~1(m,w).\displaystyle\sup\{\lambda:-\Delta\psi+C_{H}|\nabla\psi|^{\gamma}+\lambda\leq V_{1}+f_{1}(m_{1,\textbf{a}},m_{2,\textbf{a}})\text{~{}in~{}}\mathbb{R}^{N}\text{~{}for some~{}}\psi\in\mathcal{B}_{1}\}=\min_{(m,{w})\in\mathcal{B}_{1}}\tilde{J}_{1}(m,{w}). (3.27)

In fact, by following the similar argument shown in the proof of Proposition 3.4 in [8], we define

1(m,w,λ,ψ):=J~1(m,w)+N(mΔψ+wψλm)𝑑x+λ,\displaystyle\mathcal{L}_{1}(m,w,\lambda,\psi):=\tilde{J}_{1}(m,w)+\int_{\mathbb{R}^{N}}\big{(}m\Delta\psi+{w}\cdot\nabla\psi-\lambda m\big{)}\,dx+\lambda,

and obtain

min(m,w)1J~1(m,w)=min(m,w)Γsup(λ,ψ)×𝒜11(m,w,λ,ψ),\displaystyle\min_{(m,{w})\in\mathcal{B}_{1}}\tilde{J}_{1}(m,{w})=\min_{(m,{w})\in\Gamma}\sup_{(\lambda,\psi)\in\mathbb{R}\times\mathcal{A}_{1}}\mathcal{L}_{1}(m,{w},\lambda,\psi),

where Γ:=(L1(N)W1,γ(N))×Lγ(N)\Gamma:=(L^{1}(\mathbb{R}^{N})\cap W^{1,\gamma^{\prime}}(\mathbb{R}^{N}))\times L^{\gamma^{\prime}}(\mathbb{R}^{N}). Invoking the convexity of 1(,,λ,ψ)\mathcal{L}_{1}(\cdot,\cdot,\lambda,\psi) and the linearity of 1(m,w,,)\mathcal{L}_{1}(m,{w},\cdot,\cdot), one has

min(m,w)Γsup(λ,ψ)×𝒜11(m,w,λ,ψ)=sup(λ,ψ)×𝒜1min(m,w)Γ1(m,w,λ,ψ)\displaystyle\min_{(m,{w})\in\Gamma}\sup_{(\lambda,\psi)\in\mathbb{R}\times\mathcal{A}_{1}}\mathcal{L}_{1}(m,{w},\lambda,\psi)=\sup_{(\lambda,\psi)\in\mathbb{R}\times\mathcal{A}_{1}}\min_{(m,{w})\in\Gamma}\mathcal{L}_{1}(m,{w},\lambda,\psi)
=\displaystyle= sup(λ,ψ)×𝒜1Nmin(m,w)×N[CL|wm|γm+[V1+f1(m1,a,m2,a)]m+mΔψ+wψλm]𝑑x+λ\displaystyle\sup_{(\lambda,\psi)\in\mathbb{R}\times\mathcal{A}_{1}}\int_{\mathbb{R}^{N}}\min_{(m,{w})\in\mathbb{R}\times\mathbb{R}^{N}}\bigg{[}C_{L}\bigg{|}\frac{w}{m}\bigg{|}^{\gamma^{\prime}}m+[V_{1}+f_{1}(m_{1,\textbf{a}},m_{2,\textbf{a}})]m+m\Delta\psi+{w}\cdot\nabla\psi-\lambda m\bigg{]}\,dx+\lambda
=\displaystyle= {0,V1+f1(m1,a,m2,a)[Δψ+CH|ψ|γ]0,,V1+f1(m1,a,m2,a)[Δψ+CH|ψ|γ]<0\displaystyle\left\{\begin{array}[]{ll}0,&V_{1}+f_{1}(m_{1,\textbf{a}},m_{2,\textbf{a}})-[-\Delta\psi+C_{H}|\nabla\psi|^{\gamma}]\geq 0,\\ -\infty,&V_{1}+f_{1}(m_{1,\textbf{a}},m_{2,\textbf{a}})-[-\Delta\psi+C_{H}|\nabla\psi|^{\gamma}]<0\end{array}\right.
=\displaystyle= sup{λ|V1+f1(m1,a,m2,a)[Δψ+CH|ψ|γ]0 for some ψ𝒜1},\displaystyle\sup\{\lambda|V_{1}+f_{1}(m_{1,\textbf{a}},m_{2,\textbf{a}})-[-\Delta\psi+C_{H}|\nabla\psi|^{\gamma}]\geq 0\text{ for some }\psi\in\mathcal{A}_{1}\},

which shows (3.27). Moreover, with the aid of Lemma 2.3, we have

λ1,a:=\displaystyle\lambda_{1,\textbf{a}}:= sup{λ|V1+f1(m1,a,m2,a)[Δψ+CH|ψ|γ]0 for some ψ𝒜1}\displaystyle\sup\{\lambda|V_{1}+f_{1}(m_{1,\textbf{a}},m_{2,\textbf{a}})-[-\Delta\psi+C_{H}|\nabla\psi|^{\gamma}]\geq 0\text{ for some }\psi\in\mathcal{A}_{1}\}
=\displaystyle= min(m,w)1J~1(m,w)<+,\displaystyle\min_{(m,{w})\in\mathcal{B}_{1}}\tilde{J}_{1}(m,{w})<+\infty, (3.28)

and there exists u1,aC2(N)u_{1,\textbf{a}}\in C^{2}(\mathbb{R}^{N}) such that

Δu1,a+CH|u1,a|γ+λ1,a=V1+f1(m1,a,m2,a) in N.\displaystyle-\Delta u_{1,\textbf{a}}+C_{H}|\nabla u_{1,\textbf{a}}|^{\gamma}+\lambda_{1,\textbf{a}}=V_{1}+f_{1}(m_{1,\textbf{a}},m_{2,\textbf{a}})\text{~{}in~{}}\mathbb{R}^{N}. (3.29)

In particular, we have from Lemma 2.1 and Lemma 2.2 that (3.18) holds for u1,au_{1,\textbf{a}}.

Since m1,am_{1,\textbf{a}}, m2,aL(N)m_{2,\textbf{a}}\in L^{\infty}(\mathbb{R}^{N}) by Sobolev embedding, one obtains f1(m1,a,m2,a)L(N).f_{1}(m_{1,\textbf{a}},m_{2,\textbf{a}})\in L^{\infty}(\mathbb{R}^{N}). Then it follows from (3.18) and (3.29) that

|Δu1,a(x)|C(1+V1(x)).\displaystyle|-\Delta u_{1,\textbf{a}}(x)|\leq C(1+V_{1}(x)).

Thus, u1,a𝒜1u_{1,\textbf{a}}\in\mathcal{A}_{1}. Combining (3.22) with (3), one finds (3.17) holds for i=1,i=1, i.e.

λ1,a=J~1(m1,a,w1,a)=N[CL|w1,am1,a|γm1,a+[V1+f1(m1,a,m2,a)]m1,a]𝑑x,\displaystyle\lambda_{1,\textbf{a}}=\tilde{J}_{1}(m_{1,\textbf{a}},{w}_{1,\textbf{a}})=\int_{\mathbb{R}^{N}}\bigg{[}C_{L}\bigg{|}\frac{w_{1,\textbf{a}}}{m_{1,\textbf{a}}}\bigg{|}^{\gamma^{\prime}}m_{1,\textbf{a}}\,+[V_{1}+f_{1}(m_{1,\textbf{a}},m_{2,\textbf{a}})]m_{1,\textbf{a}}\bigg{]}\,dx,

where we have used (3.20). Next, we shall show

w1,a=CHγm1,a|u1,a|γ2u1,a.{w}_{1,\textbf{a}}=-C_{H}\gamma m_{1,\textbf{a}}|\nabla u_{1,\textbf{a}}|^{\gamma-2}\nabla u_{1,\textbf{a}}.

First of all, (3.17) and (3.29) imply that

0=\displaystyle 0= N[CL|w1,am1,a|γ+V1+f1(m1,a,m2,a)λ1,a]m1,a𝑑x\displaystyle\int_{\mathbb{R}^{N}}\bigg{[}C_{L}\bigg{|}\frac{{w}_{1,\textbf{a}}}{m_{1,\textbf{a}}}\bigg{|}^{\gamma^{\prime}}+V_{1}+f_{1}(m_{1,\textbf{a}},m_{2,\textbf{a}})-\lambda_{1,\textbf{a}}\bigg{]}m_{1,\textbf{a}}\,dx
=\displaystyle= N[CL|w1m1,a|γΔu1,a+CH|u1,a|γ]m1,a𝑑x.\displaystyle\int_{\mathbb{R}^{N}}\bigg{[}C_{L}\bigg{|}\frac{{w}_{1}}{m_{1,\textbf{a}}}\bigg{|}^{\gamma^{\prime}}-\Delta u_{1,\textbf{a}}+C_{H}|\nabla u_{1,\textbf{a}}|^{\gamma^{\prime}}\bigg{]}m_{1,\textbf{a}}\,dx.

Then we take ψ=u1,a\psi=u_{1,\textbf{a}} in (3.19) to get

0=\displaystyle 0= {x|m1,a>0}[CL|w1,am1,a|γ+CH|u1,a|γ+u1,aw1,am1,a]m1,a𝑑x.\displaystyle\int_{\{x|m_{1,\textbf{a}}>0\}}\bigg{[}C_{L}\bigg{|}\frac{{w}_{1,\textbf{a}}}{m_{1,\textbf{a}}}\bigg{|}^{\gamma^{\prime}}+C_{H}|\nabla u_{1,\textbf{a}}|^{\gamma^{\prime}}+\nabla u_{1,\textbf{a}}\cdot\frac{{w}_{1,\textbf{a}}}{m_{1,\textbf{a}}}\bigg{]}m_{1,\textbf{a}}\,dx. (3.30)

By using the definition of HH that

L(w1,am1,a)=CL|w1,am1,a|γ=\displaystyle L\bigg{(}-\frac{{w}_{1,\textbf{a}}}{m_{1,\textbf{a}}}\bigg{)}=C_{L}\bigg{|}\frac{{w}_{1,\textbf{a}}}{m_{1,\textbf{a}}}\bigg{|}^{\gamma^{\prime}}= suppN(pw1,am1,aH(p))CH|u1,a|γu1,aw1,am1,a,\displaystyle\sup_{p\in\mathbb{R}^{N}}\bigg{(}-p\frac{{w}_{1,\textbf{a}}}{m_{1,\textbf{a}}}-H(p)\bigg{)}\geq-C_{H}|\nabla u_{1,\textbf{a}}|^{\gamma}-\nabla u_{1,\textbf{a}}\cdot\frac{{w}_{1,\textbf{a}}}{m_{1,\textbf{a}}},

where H(p)=CH|p|γ.H(p)=C_{H}|p|^{\gamma}. Therefore, (3.30) indicates that

CL|w1,am1,a|γ+CH|u1,a|γ+u1,aw1,am1,a0 a.e. in {xN|m1,a>0}.\displaystyle C_{L}\bigg{|}\frac{{w}_{1,\textbf{a}}}{m_{1,\textbf{a}}}\bigg{|}^{\gamma^{\prime}}+C_{H}|\nabla u_{1,\textbf{a}}|^{\gamma}+\nabla u_{1,\textbf{a}}\cdot\frac{{w}_{1,\textbf{a}}}{m_{1,\textbf{a}}}\geq 0\text{~{}a.e.~{}in~{}}\{x\in\mathbb{R}^{N}|m_{1,\textbf{a}}>0\}. (3.31)

Since suppN(pw1,am1,aH(p))\sup\limits_{p\in\mathbb{R}^{N}}(-p\frac{{w}_{1,\textbf{a}}}{m_{1,\textbf{a}}}-H(p)) is attained by p=u1,ap=\nabla u_{1,\textbf{a}} when m1,a>0,m_{1,\textbf{a}}>0, one has from (3.31) that

w1,am1,a=H(u1,a) in {xN|m1,a>0}.\displaystyle\frac{{w}_{1,\textbf{a}}}{m_{1,\textbf{a}}}=-\nabla H(\nabla u_{1,\textbf{a}})\text{~{}in~{}}\{x\in\mathbb{R}^{N}|m_{1,\textbf{a}}>0\}.

Thus, we obtain

Δm1,aCH(m1,a|u1,a|γ2u1,a)=0 in a weak sense.\displaystyle-\Delta m_{1,\textbf{a}}-C_{H}\nabla\cdot(m_{1,\textbf{a}}|\nabla u_{1,\textbf{a}}|^{\gamma-2}\nabla u_{1,\textbf{a}})=0\text{~{}in a weak sense}.

Proceeding the similar argument shown above, we have (3.17) holds for i=2i=2 and there exists u2C2(N)u_{2}\in C^{2}(\mathbb{R}^{N}) such that

w2,a=CHγm2,a|u2,a|γ2u2,a,xN in a weak sense.{w}_{2,\textbf{a}}=-C_{H}\gamma m_{2,\textbf{a}}|\nabla u_{2,\textbf{a}}|^{\gamma-2}\nabla u_{2,\textbf{a}},~{}x\in\mathbb{R}^{N}\text{~{}in a weak sense}.

Finally, by the standard elliptic regularity, we find (3.16) holds in a classical sense. This completes the proof of this lemma. ∎

By summarizing Lemma 3.1 and Lemma 3.2, we are able to show conclusions stated in Theorem 1.1, which are

Proof of Theorem 1.1:

Proof.

For Conclusion (i), we invoke Lemma 3.1 to get there exists a minimizer (m1,a,w1,a,m2,a,m2,a)𝒦(m_{1,\textbf{a}},w_{1,\textbf{a}},m_{2,\textbf{a}},m_{2,\textbf{a}})\in\mathcal{K} to (1.10). Moreover, Lemma 3.2 implies there exist (u1,a,u2,a)C2(N)×C2(N)(u_{1,\textbf{a}},u_{2,\textbf{a}})\in C^{2}(\mathbb{R}^{N})\times C^{2}(\mathbb{R}^{N}) and (λ1,a,λ2,a)2(\lambda_{1,\textbf{a}},\lambda_{2,\textbf{a}})\in\mathbb{R}^{2} such that (m1,a,m2,a,u1,a,u2,a,λ1,a,λ2,a)(m_{1,\textbf{a}},m_{2,\textbf{a}},u_{1,\textbf{a}},u_{2,\textbf{a}},\lambda_{1,\textbf{a}},\lambda_{2,\textbf{a}}) solves (1.26). By standard regularity arguments, we have from Lemma 2.4 that

(m1,a,m2,a,u1,a,u2,a)W1,p(N)×W1,p(N)×C2(N)×C2(N),(m_{1,\textbf{a}},m_{2,\textbf{a}},u_{1,\textbf{a}},u_{2,\textbf{a}})\in W^{1,p}(\mathbb{R}^{N})\times W^{1,p}(\mathbb{R}^{N})\times C^{2}(\mathbb{R}^{N})\times C^{2}(\mathbb{R}^{N}),

which completes the proof of this conclusion. Conclusion (ii) is the straightforward corollary of Lemma 3.1. ∎

We next focus on the borderline case when α1=α2\alpha_{1}=\alpha_{2} shown in Theorem 1.1. In detail, we impose the extra assumption (1.27) on the potentials and investigate the conclusions shown in Theorem 1.2, which are

Proof of Theorem 1.2:

Proof.

In light of the assumption (1.27), we let (mt,wt)(m_{t},w_{t}) be (3.6) with x0Nx_{0}\in\mathbb{R}^{N} satisfying

V1(x0)=V2(x0)=0.\displaystyle V_{1}(x_{0})=V_{2}(x_{0})=0.

Then for i=1,2,i=1,2, we compute to get

NVi(x)mt𝑑x=\displaystyle\int_{\mathbb{R}^{N}}V_{i}(x)m_{t}\,dx= 1MNVi(x)tNm0(t(xx0))𝑑x=1MNVi(yt+x0)m0(y)𝑑y.\displaystyle\frac{1}{M^{*}}\int_{\mathbb{R}^{N}}V_{i}(x)t^{N}m_{0}(t(x-x_{0}))\,dx=\frac{1}{M^{*}}\int_{\mathbb{R}^{N}}V_{i}\bigg{(}\frac{y}{t}+x_{0}\bigg{)}m_{0}(y)\,dy.

By invoking Lebesgue dominated theorem, we further obtain as t+,t\rightarrow+\infty,

NVi(x)mt𝑑xVi(x0)=0, for i=1,2.\displaystyle\int_{\mathbb{R}^{N}}V_{i}(x)m_{t}\,dx\rightarrow V_{i}(x_{0})=0,\text{~{}for~{}}i=1,2.

Proceeding the similar argument shown in the proof of Lemma 3.1, we get

aβ,aβ,β(mt,wt,mt,wt)=[V1(x0)+V2(x0)]+ot(1),\displaystyle\mathcal{E}_{a^{*}-\beta,a^{*}-\beta,\beta}(m_{t},w_{t},m_{t},w_{t})=[V_{1}(x_{0})+V_{2}(x_{0})]+o_{t}(1), (3.32)

where ot(1)0o_{t}(1)\rightarrow 0 as t+.t\rightarrow+\infty. We take t+t\rightarrow+\infty in (3.32) to obtain

eαβ,αβ,β0.\displaystyle e_{\alpha^{*}-\beta,\alpha^{*}-\beta,\beta}\leq 0. (3.33)

On the other hand, we rewrite (1) as

α1,α2,β(m1,w1,m2,w2)=\displaystyle\mathcal{E}_{\alpha_{1},\alpha_{2},\beta}(m_{1},w_{1},m_{2},w_{2})= i=12(NCL|wimi|γmi+VimiN(αi+β)N+γNmi1+γN𝑑x)\displaystyle\sum_{i=1}^{2}\bigg{(}\int_{\mathbb{R}^{N}}C_{L}\big{|}\frac{w_{i}}{m_{i}}\bigg{|}^{\gamma^{\prime}}m_{i}+V_{i}m_{i}-\frac{N(\alpha_{i}+\beta)}{N+\gamma^{\prime}}\int_{\mathbb{R}^{N}}m_{i}^{1+\frac{\gamma^{\prime}}{N}}\,dx\bigg{)}
+NβN+γN(m112+γ2Nm212+γ2N)2𝑑x.\displaystyle+\frac{N\beta}{N+\gamma^{\prime}}\int_{\mathbb{R}^{N}}\bigg{(}m_{1}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}-m_{2}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\bigg{)}^{2}\,dx. (3.34)

Upon substituting α1=α2=aβ\alpha_{1}=\alpha_{2}=a^{*}-\beta and β=aα\beta=a^{*}-\alpha into (3), we deduce that

eaβ,aβ,β0.\displaystyle e_{a^{*}-\beta,a^{*}-\beta,\beta}\geq 0. (3.35)

Combining (3.33) with (3.35), one has

eaβ,aβ,β=0.\displaystyle e_{a^{*}-\beta,a^{*}-\beta,\beta}=0. (3.36)

Now, we argue by contradiction and assume that (m1,w1,m2,w2)(m_{1},w_{1},m_{2},w_{2}) is a minimizer of (1.10) with α1=α2=aβ\alpha_{1}=\alpha_{2}=a^{*}-\beta and β=aα\beta=a^{*}-\alpha. Then we have

aβ,aβ,β(m1,w1,m2,w2)=\displaystyle\mathcal{E}_{a^{*}-\beta,a^{*}-\beta,\beta}(m_{1},w_{1},m_{2},w_{2})= i=12CLN|wimi|γmi𝑑xNaN+γNmi1+γN𝑑x\displaystyle\sum_{i=1}^{2}C_{L}\int_{\mathbb{R}^{N}}\bigg{|}\frac{w_{i}}{m_{i}}\bigg{|}^{\gamma^{\prime}}m_{i}\,dx-\frac{Na^{*}}{N+\gamma^{\prime}}\int_{\mathbb{R}^{N}}m_{i}^{1+\frac{\gamma^{\prime}}{N}}\,dx
+NβN+γN(m112+γ2Nm212+γ2N)2𝑑x\displaystyle+\frac{N\beta}{N+\gamma^{\prime}}\int_{\mathbb{R}^{N}}\bigg{(}m_{1}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}-m_{2}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\bigg{)}^{2}\,dx
+NV1(x)m1+V2(x)m2dx\displaystyle+\int_{\mathbb{R}^{N}}V_{1}(x)m_{1}+V_{2}(x)m_{2}\,dx
:=\displaystyle:= I1+I2+I3.\displaystyle I_{1}+I_{2}+I_{3}. (3.37)

In light of (3.36), one finds from (3) that I1=I2=I3=0,I_{1}=I_{2}=I_{3}=0, in which I1=0I_{1}=0 implies each (mi,wi)(m_{i},w_{i}), i=1,2i=1,2 is a minimizer of problem (1.15). In addition, I2=0I_{2}=0 indicates that m1=m2m_{1}=m_{2} in N.\mathbb{R}^{N}. Morever, one gets from I3=0I_{3}=0 that

NV1(x)m1+V2(x)m2dx=0,\displaystyle\int_{\mathbb{R}^{N}}V_{1}(x)m_{1}+V_{2}(x)m_{2}\,dx=0,

which leads to a contradiction since mi>0m_{i}>0 for i=1,2i=1,2 by using the compactly Sobolev embedding and the maximum principle as shown in [3]. ∎

For the existence of minimizers, we next consider the case of α1=a\alpha_{1}=a^{*} and show Theorem 1.3, which is

Proof of Theorem 1.3:

Proof.

We define the test solution-pair as

mi,τ(x)=τNMm0(τ(xx¯i+(1)iιlnττν)),wi,τ=τN+1Mw0(τ(xx¯i+(1)iιlnττν)),\displaystyle m_{i,\tau}(x)=\frac{\tau^{N}}{M^{*}}m_{0}\bigg{(}\tau\bigg{(}x-\bar{x}_{i}+(-1)^{i}\iota\frac{\ln\tau}{\tau}\nu\bigg{)}\bigg{)},~{}w_{i,\tau}=\frac{\tau^{N+1}}{M^{*}}w_{0}\bigg{(}\tau\bigg{(}x-\bar{x}_{i}+(-1)^{i}\iota\frac{\ln\tau}{\tau}\nu\bigg{)}\bigg{)}, (3.38)

where (m0,w0)(m_{0},w_{0}) denotes a minimizer of (1.15) satisfying (1.19), ν𝕊N1\nu\in\mathbb{S}^{N-1}, x¯iN\bar{x}_{i}\in\mathbb{R}^{N} and constant ι\iota will be determined later.

By using Lemma 2.6, we have

CLN|wi,τmi,τ|γmi,τ𝑑x=τγM,Nmi,τ1+γN𝑑x=N+γNτγ(M)1+γN,\displaystyle C_{L}\int_{\mathbb{R}^{N}}\bigg{|}\frac{w_{i,\tau}}{m_{i,\tau}}\bigg{|}^{\gamma^{\prime}}m_{i,\tau}\,dx=\frac{\tau^{\gamma^{\prime}}}{M^{*}},~{}~{}\int_{\mathbb{R}^{N}}m_{i,\tau}^{1+\frac{\gamma^{\prime}}{N}}\,dx=\frac{N+\gamma^{\prime}}{N}\frac{\tau^{\gamma^{\prime}}}{(M^{*})^{1+\frac{\gamma^{\prime}}{N}}},

and

Nm1,τ12+γ2Nm2,τ12+γ2N𝑑x=τγ(M)1+γNNm012+γ2N(x)m012+γ2N(x+τ(x¯1x¯2)+2ιlnτν)𝑑x.\displaystyle\int_{\mathbb{R}^{N}}m_{1,\tau}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}m_{2,\tau}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\,dx=\frac{\tau^{\gamma^{\prime}}}{(M^{*})^{1+\frac{\gamma^{\prime}}{N}}}\int_{\mathbb{R}^{N}}m_{0}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}(x)m_{0}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}(x+\tau(\bar{x}_{1}-\bar{x}_{2})+2\iota\ln\tau\nu)\,dx. (3.39)

We have the fact that

|τ(x¯1x¯2)+2ιlnτν|2ιlnτwhen τ1.\displaystyle|\tau(\bar{x}_{1}-\bar{x}_{2})+2\iota\ln\tau\nu|\geq 2\iota\ln\tau~{}~{}\text{when }\tau\gg 1.

Hence, for τ\tau large, if xBιlnτ={x||x|<ιlnτ}x\in B_{\iota\ln\tau}=\{x||x|<\iota\ln\tau\}, one gets from Lemma 2.5 that

m0(x+τ(x¯1x¯2)+2ιlnτν)Ceδ0ιlnτ,\displaystyle m_{0}(x+\tau(\bar{x}_{1}-\bar{x}_{2})+2\iota\ln\tau\nu)\leq Ce^{-\delta_{0}\iota\ln\tau}, (3.40)

where C>0C>0 and δ0>0\delta_{0}>0 are some constants. And if xBιlnτcx\in B_{\iota\ln\tau}^{c}, then

m0(x)Ceδ0ιlnτ.\displaystyle m_{0}(x)\leq Ce^{-\delta_{0}\iota\ln\tau}. (3.41)

Combining (3.40) and (3.41), one finds from (3.39) that as τ+\tau\rightarrow+\infty,

Nm1,τ12+γ2Nm2,τ12+γ2Ndx=τγ(M)1+γN[\displaystyle\int_{\mathbb{R}^{N}}m_{1,\tau}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}m_{2,\tau}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\,dx=\frac{\tau^{\gamma^{\prime}}}{(M^{*})^{1+\frac{\gamma^{\prime}}{N}}}\bigg{[} Bιlnτm012+γ2N(x)m012+γ2N(x+τ(x¯1x¯2)+2ιlnτν)𝑑x\displaystyle\int_{B_{\iota\ln\tau}}m_{0}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}(x)m_{0}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}(x+\tau(\bar{x}_{1}-\bar{x}_{2})+2\iota\ln\tau\nu)\,dx
+Bιlnτcm012+γ2N(x)m012+γ2N(x+τ(x¯1x¯2)+2ιlnτν)dx]\displaystyle+\int_{B_{\iota\ln\tau}^{c}}m_{0}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}(x)m_{0}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}(x+\tau(\bar{x}_{1}-\bar{x}_{2})+2\iota\ln\tau\nu)\,dx\bigg{]}
\displaystyle\leq CN,γτγe(12+γ2N)δ0ιlnτ=CN,γτγ(12+γ2N)δ0ι0,\displaystyle C_{N,\gamma^{\prime}}\tau^{\gamma^{\prime}}e^{-(\frac{1}{2}+\frac{\gamma^{\prime}}{2N})\delta_{0}\iota\ln\tau}=C_{N,\gamma^{\prime}}\tau^{\gamma^{\prime}-\big{(}\frac{1}{2}+\frac{\gamma^{\prime}}{2N}\big{)}\delta_{0}\iota}\rightarrow 0, (3.42)

where constant ι\iota is chosen as ι>2γN(N+γ)δ0.\iota>\frac{2\gamma^{\prime}N}{(N+\gamma^{\prime})\delta_{0}}. In addition,

NVi(x)mi,τ(x)𝑑x=1MNVi(xτ+x¯i(1)iιlnττν)m0𝑑x:=1MNgτ(x)𝑑x.\displaystyle\int_{\mathbb{R}^{N}}V_{i}(x)m_{i,\tau}(x)\,dx=\frac{1}{M^{*}}\int_{\mathbb{R}^{N}}V_{i}\bigg{(}\frac{x}{\tau}+\bar{x}_{i}-(-1)^{i}\iota\frac{\ln\tau}{\tau}\nu\bigg{)}m_{0}\,dx:=\frac{1}{M^{*}}\int_{\mathbb{R}^{N}}g_{\tau}(x)\,dx.

Noting that gτ(x)Vi(x¯i)m0(x)g_{\tau}(x)\rightarrow V_{i}(\bar{x}_{i})m_{0}(x) a.e. in N\mathbb{R}^{N}, we obtain from (1.14) and Lemma 2.5 that when τ\tau is large,

|gτ(x)|Ceδ|xτ+x¯i(1)iιlnττν|eδ0|x|Ceδ02|x|L1(N).\displaystyle|g_{\tau}(x)|\leq Ce^{\delta|\frac{x}{\tau}+\bar{x}_{i}-(-1)^{i}\iota\frac{\ln\tau}{\tau}\nu|}e^{-\delta_{0}|x|}\leq Ce^{-\frac{\delta_{0}}{2}|x|}\in L^{1}(\mathbb{R}^{N}).

Thus, by Lebesgue dominated theorem, we further get

NVimi,τ𝑑xVi(x¯i) as τ+.\displaystyle\int_{\mathbb{R}^{N}}V_{i}m_{i,\tau}\,dx\rightarrow V_{i}(\bar{x}_{i})\text{~{}as~{}}\tau\rightarrow+\infty. (3.43)

Collecting (3.39), (3) and (3.43), one finds if α1=α2=a\alpha_{1}=\alpha_{2}=a^{*} and β0\beta\leq 0, then

α1,α2,β(m1,τ,w1,τ,m2,τ,w2,τ)=V1(x¯1)+V2(x¯2)+oτ(1),\displaystyle\mathcal{E}_{\alpha_{1},\alpha_{2},\beta}(m_{1,\tau},w_{1,\tau},m_{2,\tau},w_{2,\tau})=V_{1}(\bar{x}_{1})+V_{2}(\bar{x}_{2})+o_{\tau}(1),

where oτ(1)0o_{\tau}(1)\rightarrow 0 as τ+.\tau\rightarrow+\infty. Choose x¯i\bar{x}_{i} such that Vi(x¯i)=0V_{i}(\bar{x}_{i})=0 for i=1,2,i=1,2,, it then follows that

eα1,α2,βV1(x¯1)+V2(x¯2)=0.\displaystyle e_{\alpha_{1},\alpha_{2},\beta}\leq V_{1}(\bar{x}_{1})+V_{2}(\bar{x}_{2})=0.

Moreover, by using (1.20) and β0\beta\leq 0, one has eα1,α2,β0.e_{\alpha_{1},\alpha_{2},\beta}\geq 0. Therefore, we summarize to get eα1,α2,β=0.e_{\alpha_{1},\alpha_{2},\beta}=0. Proceeding the same argument as shown in the proof of Theorem 1.2, we show there is no minimizer in case (i).

For case (ii), if β=0,\beta=0, one finds

eα1,α2,0=ea1+eα22,\displaystyle e_{\alpha_{1},\alpha_{2},0}=e^{1}_{a^{*}}+e^{2}_{\alpha_{2}},

where ea1e^{1}_{a^{*}} and eα22e^{2}_{\alpha_{2}} are given by (3.1). Noting that this is the decoupled case, we have the fact that there is no minimizer as shown in [12].

If 0<βaα220<\beta\leq\frac{a^{*}-\alpha_{2}}{2}, taking ι=0\iota=0 in (3.38), we compute to get

Nm1,τ12+γ2Nm012+γ2N𝑑x=τ12(γN)(M)12+γ2NNm0(xτ+x¯1)m012+γ2N𝑑x:=τ12(γN)(M)12+γ2NIτ.\displaystyle\int_{\mathbb{R}^{N}}m_{1,\tau}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}m_{0}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\,dx=\frac{\tau^{\frac{1}{2}(\gamma^{\prime}-N)}}{(M^{*})^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}}\int_{\mathbb{R}^{N}}m_{0}\bigg{(}\frac{x}{\tau}+\bar{x}_{1}\bigg{)}m_{0}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\,dx:=\frac{\tau^{\frac{1}{2}(\gamma^{\prime}-N)}}{(M^{*})^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}}I_{\tau}. (3.44)

We choose x¯1N\bar{x}_{1}\in\mathbb{R}^{N} such that m0(x¯1)>C0>0m_{0}(\bar{x}_{1})>C_{0}>0 then obtain

limτ+IτC0Nm012+γ2N𝑑xC1>0 as τ+.\displaystyle\lim_{\tau\rightarrow+\infty}I_{\tau}\geq C_{0}\int_{\mathbb{R}^{N}}m_{0}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\,dx\geq C_{1}>0\text{ as }\tau\rightarrow+\infty.

Thus, (3.44) implies

Nm1,τ12+γ2Nm012+γ2N𝑑xC1τ12(γN)+,\displaystyle\int_{\mathbb{R}^{N}}m_{1,\tau}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}m_{0}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\,dx\geq C_{1}\tau^{\frac{1}{2}(\gamma^{\prime}-N)}\rightarrow+\infty,

It follows that

a.α2,β(m1,τ,w1,τ,m0,w0)oτ(1)+CCγβτ12(γN) for β>0.\displaystyle\mathcal{E}_{a^{*}.\alpha_{2},\beta}(m_{1,\tau},w_{1,\tau},m_{0},w_{0})\leq o_{\tau}(1)+C-C_{\gamma^{\prime}}\beta\tau^{\frac{1}{2}(\gamma^{\prime}-N)}\rightarrow-\infty\text{ for }\beta>0.

Hence ea,α2,β=e_{a^{*},\alpha_{2},\beta}=-\infty if β>0\beta>0, which indicates (1.10) has no minimizer.

As shown in Theorem 1.1 and Theorem 1.2, we have obtained when all coefficients α1,\alpha_{1}, α2\alpha_{2} and β\beta are subcritical, (1.7) admits classical ground states; whereas, if α1=α2\alpha_{1}=\alpha_{2} are subcritical and β\beta is critical, then (1.10) has no minimizer. A natural question is the behaviors of ground states as (α1,α2)(aβ,aβ)(\alpha_{1},\alpha_{2})\nearrow(a^{*}-\beta,a^{*}-\beta). In fact, we can show there are concentration phenomena as coefficients approach critical ones. In the next section, we shall discuss the asymptotic profiles of ground states in the singular limits mentioned above.

4 Asymptotic Profiles of Ground States with β>0\beta>0

This section is devoted to the blow-up behaviors of ground states to (1.7) in some singular limits under the attractive interaction case. We proceed the proof of Theorem 1.4 as follows.

Proof of Theorem 1.4:

Proof.

First of all, we have from (3) that

α1,α2,β(m1,a,w1,a,m2,a,w2,a)=\displaystyle\mathcal{E}_{\alpha_{1},\alpha_{2},\beta}(m_{1,\textbf{a}},w_{1,\textbf{a}},m_{2,\textbf{a}},w_{2,\textbf{a}})= i=12(NCL|wi,ami,a|γmi,aN(αi+β)N+γNmi,a1+γN𝑑x)\displaystyle\sum_{i=1}^{2}\bigg{(}\int_{\mathbb{R}^{N}}C_{L}\big{|}\frac{w_{i,\textbf{a}}}{m_{i,\textbf{a}}}\bigg{|}^{\gamma^{\prime}}m_{i,\textbf{a}}-\frac{N(\alpha_{i}+\beta)}{N+\gamma^{\prime}}\int_{\mathbb{R}^{N}}m_{i,\textbf{a}}^{1+\frac{\gamma^{\prime}}{N}}\,dx\bigg{)}
+NβN+γN(m1,a12+γ2Nm2,a12+γ2N)2𝑑x\displaystyle+\frac{N\beta}{N+\gamma^{\prime}}\int_{\mathbb{R}^{N}}\bigg{(}m_{1,\textbf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}-m_{2,\textbf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\bigg{)}^{2}\,dx
+NV1(x)m1,a+V2(x)m2,adx\displaystyle+\int_{\mathbb{R}^{N}}V_{1}(x)m_{1,\textbf{a}}+V_{2}(x)m_{2,\textbf{a}}\,dx
:=\displaystyle:= II1+II2+II3.\displaystyle II_{1}+II_{2}+II_{3}. (4.1)

In light of (1.13) and (1.20), one finds IIj0II_{j}\geq 0, j=1,2,3.j=1,2,3. Moreover, assumption (1.13) implies II30.II_{3}\geq 0. Proceeding the same argument shown in the proof of Lemma 3.1, we use the test pair (3.6) and compute from (4) that

limaaβeα1,α2,β=eaβ,aβ,β=0.\displaystyle\lim_{\textbf{a}\nearrow\textbf{a}^{*}_{\beta}}e_{\alpha_{1},\alpha_{2},\beta}=e_{a^{*}-\beta,a^{*}-\beta,\beta}=0. (4.2)

Combining (4) with (4.2), we obtain (1.31) and (1.32).

We next prove (1.33) and argue by contradiction. Without loss of generality, we assume that

lim supaaβCLN|w1,am1,a|γm1,a𝑑x<+.\displaystyle\limsup_{\textbf{a}\nearrow\textbf{a}^{*}_{\beta}}C_{L}\int_{\mathbb{R}^{N}}\bigg{|}\frac{w_{1,\textbf{a}}}{m_{1,\textbf{a}}}\bigg{|}^{\gamma^{\prime}}m_{1,\textbf{a}}\,dx<+\infty.

Then, it follows from (1.31), (1.32) and Lemma 2.4 that (m1,a,w1,a,m2,a,w2,a)(m_{1,\textbf{a}},w_{1,\textbf{a}},m_{2,\textbf{a}},w_{2,\textbf{a}}) is uniformly bounded in (W1,γ(N)×Lγ(N))2.(W^{1,\gamma^{\prime}}(\mathbb{R}^{N})\times L^{\gamma^{\prime}}(\mathbb{R}^{N}))^{2}. Moreover, by compactly Sobolev embedding (C.f. Lemma 5.1 in [12]), one finds mi,ami,0m_{i,\textbf{a}}\rightarrow m_{i,0} strongly in L1(N)L1+γN(N)L^{1}(\mathbb{R}^{N})\cap L^{1+\frac{\gamma^{\prime}}{N}}(\mathbb{R}^{N}) for i=1,2.i=1,2. By using the convexity of N|wm|γm𝑑x\int_{\mathbb{R}^{N}}\big{|}\frac{w}{m}\big{|}^{\gamma^{\prime}}m\,dx, we have

eaβ,aβ,β=limaaβeα1,α2,β=\displaystyle e_{a^{*}-\beta,a^{*}-\beta,\beta}=\lim_{\textbf{a}\nearrow\textbf{a}_{\beta}^{*}}e_{\alpha_{1},\alpha_{2},\beta}= limaaβα1,α2,β(m1,a,w1,a,m2,a,w2,a)\displaystyle\lim_{\textbf{a}\nearrow\textbf{a}^{*}_{\beta}}\mathcal{E}_{\alpha_{1},\alpha_{2},\beta}(m_{1,\textbf{a}},w_{1,\textbf{a}},m_{2,\textbf{a}},w_{2,\textbf{a}})
\displaystyle\geq aβ,aβ,β(m1,0,w1,0,m2,0,w2,0)eaβ,aβ,β,\displaystyle\mathcal{E}_{a^{*}-\beta,a^{*}-\beta,\beta}(m_{1,0},w_{1,0},m_{2,0},w_{2,0})\geq e_{a^{*}-\beta,a^{*}-\beta,\beta},

which implies (m1,0,w1,0,m2,0,w2,0)(m_{1,0},w_{1,0},m_{2,0},w_{2,0}) is a minimizer of eaβ,aβ,βe_{a^{*}-\beta,a^{*}-\beta,\beta} and it is a contradiction since we have showed that eaβ,aβ,βe_{a^{*}-\beta,a^{*}-\beta,\beta} has no minimizer in Theorem 1.2.

Now, we find (1.33) holds and further obtain from (1.31) that for i=1,2i=1,2

limaaβNmi,a1+γN𝑑x=+ and limaaβCLN|wi,ami,a|γmi,a𝑑xNmi,a1+γN𝑑x=N+γNa.\displaystyle\lim_{\textbf{a}\nearrow\textbf{a}_{\beta}^{*}}\int_{\mathbb{R}^{N}}m_{i,\textbf{a}}^{1+\frac{\gamma^{\prime}}{N}}\,dx=+\infty\text{ and }~{}\lim_{\textbf{a}\nearrow\textbf{a}_{\beta}^{*}}\frac{C_{L}\int_{\mathbb{R}^{N}}\big{|}\frac{w_{i,\textbf{a}}}{m_{i,\textbf{a}}}\big{|}^{\gamma^{\prime}}m_{i,\textbf{a}}\,dx}{\int_{\mathbb{R}^{N}}m_{i,\textbf{a}}^{1+\frac{\gamma^{\prime}}{N}}\,dx}=\frac{N+\gamma^{\prime}}{N}a^{*}.

Noting that as aaβ,\textbf{a}\nearrow\textbf{a}_{\beta}^{*},

[(Nm1,a1+γN𝑑x)12(Nm2,a1+γN𝑑x)12]2N(m1,a12+γ2Nm2,a12+γ2N)2𝑑x0,\displaystyle\bigg{[}\bigg{(}\int_{\mathbb{R}^{N}}m_{1,\textbf{a}}^{1+\frac{\gamma^{\prime}}{N}}\,dx\bigg{)}^{\frac{1}{2}}-\bigg{(}\int_{\mathbb{R}^{N}}m_{2,\textbf{a}}^{1+\frac{\gamma^{\prime}}{N}}\,dx\bigg{)}^{\frac{1}{2}}\bigg{]}^{2}\leq\int_{\mathbb{R}^{N}}\bigg{(}m_{1,\textbf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}-m_{2,\textbf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\bigg{)}^{2}\,dx\rightarrow 0,

one gets (1.34) holds.

Noting that (m1,a,w1,a,m2,a,w2,a)(m_{1,\textbf{a}},{{w}}_{1,\textbf{a}},m_{2,\textbf{a}},{{w}}_{2,\textbf{a}}) satisfy (3.16), we have from the integration by parts that

Nu1,am1,adx+CHN|u1,a|γm1,a𝑑x+λ1\displaystyle\int_{\mathbb{R}^{N}}\nabla u_{1,\textbf{a}}\cdot\nabla m_{1,\textbf{a}}\,dx+C_{H}\int_{\mathbb{R}^{N}}|\nabla u_{1,\textbf{a}}|^{\gamma}m_{1,\textbf{a}}\,dx+\lambda_{1}
=\displaystyle= NV1m1,a𝑑xα1Nm1,a1+γN𝑑xβNm1,a12+γ2Nm2,a12+γ2N𝑑x,\displaystyle\int_{\mathbb{R}^{N}}V_{1}m_{1,\textbf{a}}\,dx-\alpha_{1}\int_{\mathbb{R}^{N}}m_{1,\textbf{a}}^{1+\frac{\gamma^{\prime}}{N}}\,dx-\beta\int_{\mathbb{R}^{N}}m_{1,\textbf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}m_{2,\textbf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\,dx, (4.3)

and

Nu1,am1,adx=CHγNm1,a|u1,a|γ𝑑x.\displaystyle\int_{\mathbb{R}^{N}}\nabla u_{1,\textbf{a}}\cdot\nabla m_{1,\textbf{a}}\,dx=-C_{H}\gamma\int_{\mathbb{R}^{N}}m_{1,\textbf{a}}|\nabla u_{1,\textbf{a}}|^{\gamma}\,dx. (4.4)

Combining (4.3) with (4.4), one finds

λ1=\displaystyle\lambda_{1}= CLN|w1,am1,a|γm1,a𝑑x+NV1m1,a𝑑xα1Nm1,a1+γN𝑑xβNm1,a12+γ2Nm2,a12+γ2N𝑑x\displaystyle C_{L}\int_{\mathbb{R}^{N}}\bigg{|}\frac{w_{1,\textbf{a}}}{m_{1,\textbf{a}}}\bigg{|}^{\gamma^{\prime}}m_{1,\textbf{a}}\,dx+\int_{\mathbb{R}^{N}}V_{1}m_{1,\textbf{a}}\,dx-\alpha_{1}\int_{\mathbb{R}^{N}}m_{1,\textbf{a}}^{1+\frac{\gamma^{\prime}}{N}}\,dx-\beta\int_{\mathbb{R}^{N}}m_{1,\textbf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}m_{2,\textbf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\,dx
=\displaystyle= (CLN|w1,am1,a|γm1,a𝑑xN(α1+β)N+γNm1,a1+γN𝑑x)γ(α1+β)γ+NNm1,a1+γN𝑑x\displaystyle\bigg{(}C_{L}\int_{\mathbb{R}^{N}}\bigg{|}\frac{w_{1,\textbf{a}}}{m_{1,\textbf{a}}}\bigg{|}^{\gamma^{\prime}}m_{1,\textbf{a}}\,dx-\frac{N(\alpha_{1}+\beta)}{N+\gamma^{\prime}}\int_{\mathbb{R}^{N}}m_{1,\textbf{a}}^{1+\frac{\gamma^{\prime}}{N}}\,dx\bigg{)}-\frac{\gamma^{\prime}(\alpha_{1}+\beta)}{\gamma^{\prime}+N}\int_{\mathbb{R}^{N}}m_{1,\textbf{a}}^{1+\frac{\gamma^{\prime}}{N}}\,dx
+βNm1,a12+γ2N(m1,a12+γ2Nm2,a12+γ2N)𝑑x+NV1m1,a𝑑x\displaystyle+\beta\int_{\mathbb{R}^{N}}m_{1,\textbf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\bigg{(}m_{1,\textbf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}-m_{2,\textbf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\bigg{)}\,dx+\int_{\mathbb{R}^{N}}V_{1}m_{1,\textbf{a}}\,dx
=\displaystyle= oε(1)γ(α1+β)γ+NNm1,a1+γN𝑑x+βNm1,a12+γ2N(m1,a12+γ2Nm2,a12+γ2N)𝑑x,\displaystyle o_{\varepsilon}(1)-\frac{\gamma^{\prime}(\alpha_{1}+\beta)}{\gamma^{\prime}+N}\int_{\mathbb{R}^{N}}m_{1,\textbf{a}}^{1+\frac{\gamma^{\prime}}{N}}\,dx+\beta\int_{\mathbb{R}^{N}}m_{1,\textbf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\bigg{(}m_{1,\textbf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}-m_{2,\textbf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\bigg{)}\,dx, (4.5)

where we have used (1.31) and (1.32) as aaβ.\textbf{a}\nearrow\textbf{a}_{\beta}^{*}. To further simplify (4), we use (1.32) to get

|Nm1,a12+γ2N(m1,a12+γ2Nm2,a12+γ2N)𝑑x|\displaystyle\bigg{|}\int_{\mathbb{R}^{N}}m_{1,\textbf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\bigg{(}m_{1,\textbf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}-m_{2,\textbf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\bigg{)}\,dx\bigg{|}\leq (Nm1,a1+γN𝑑x)12[N(m1,a12+γ2Nm2,a12+γ2N)2𝑑x]12\displaystyle\bigg{(}\int_{\mathbb{R}^{N}}m_{1,\textbf{a}}^{1+\frac{\gamma^{\prime}}{N}}\,dx\bigg{)}^{\frac{1}{2}}\bigg{[}\int_{\mathbb{R}^{N}}\bigg{(}m_{1,\textbf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}-m_{2,\textbf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\bigg{)}^{2}\,dx\bigg{]}^{\frac{1}{2}}
=\displaystyle= oε(1)(Nm1,a1+γN𝑑x)12.\displaystyle o_{\varepsilon}(1)\bigg{(}\int_{\mathbb{R}^{N}}m_{1,\textbf{a}}^{1+\frac{\gamma^{\prime}}{N}}\,dx\bigg{)}^{\frac{1}{2}}. (4.6)

By utilizing (1.31) and (1.35), one finds

limaaβNεaγ(α1+β)N+γNm1,a1+γN𝑑x=1.\displaystyle\lim_{\textbf{a}\nearrow\textbf{a}_{\beta}^{*}}{\frac{N\varepsilon_{\textbf{a}}^{\gamma^{\prime}}(\alpha_{1}+\beta)}{N+\gamma^{\prime}}\int_{\mathbb{R}^{N}}m_{1,\textbf{a}}^{1+\frac{\gamma^{\prime}}{N}}\,dx}=1. (4.7)

Collecting (4), (4) and (4.7), we have

λ1=(α1+β)γN+γNm1,a1+γN𝑑x+oε(1)(Nm1,a1+γN𝑑x)12+oε(1)=γNεγ+oε(1)(1+εγ2),\displaystyle\lambda_{1}=-\frac{(\alpha_{1}+\beta)\gamma^{\prime}}{N+\gamma^{\prime}}\int_{\mathbb{R}^{N}}m_{1,\textbf{a}}^{1+\frac{\gamma^{\prime}}{N}}\,dx+o_{\varepsilon}(1)\bigg{(}\int_{\mathbb{R}^{N}}m_{1,\textbf{a}}^{1+\frac{\gamma^{\prime}}{N}}\,dx\bigg{)}^{\frac{1}{2}}+o_{\varepsilon}(1)=\frac{\gamma^{\prime}}{N}\varepsilon^{-\gamma^{\prime}}+o_{\varepsilon}(1)(1+\varepsilon^{-\frac{\gamma^{\prime}}{2}}), (4.8)

where ε0\varepsilon\rightarrow 0 given by (1.35). This implies that limε0λ1εγ=γN\lim_{\varepsilon\rightarrow 0}\lambda_{1}\varepsilon^{\gamma^{\prime}}=-\frac{\gamma^{\prime}}{N}. Proceeding the similar argument shown above, one obtains from (1.34) that limε0λ2εγ=γN\lim_{\varepsilon\rightarrow 0}\lambda_{2}\varepsilon^{\gamma^{\prime}}=-\frac{\gamma^{\prime}}{N}. Now, we substitute (1.38) into (3.16) and obtain

{Δu1,ε+CH|u1,ε|γ+λ1εγ=εγV1(εx+x1,ε)α1m1,εγNβm1,εγ2N12m2,ε12+γN,xN,Δm1,ε+w1,ε=0,w1,ε=γCHm1,ε|u1,ε|γ2u1,εxN,Δu2,ε+CH|u2,ε|γ+λ2εγ=εγV2(εx+x1,ε)α2m2,εγNβm2,εγ2N12m1,ε12+γN,xN,Δm2,ε+w2,ε=0,w2,ε=γCHm2,ε|u2,ε|γ2u2,ε,xN,Nm1,ε𝑑x=Nm2,ε𝑑x=1.\displaystyle\left\{\begin{array}[]{ll}-\Delta u_{1,\varepsilon}+C_{H}|\nabla u_{1,\varepsilon}|^{\gamma}+\lambda_{1}\varepsilon^{\gamma^{\prime}}=\varepsilon^{\gamma^{\prime}}V_{1}(\varepsilon x+x_{1,\varepsilon})-\alpha_{1}m_{1,\varepsilon}^{\frac{\gamma^{\prime}}{N}}-\beta m_{1,\varepsilon}^{\frac{\gamma^{\prime}}{2N}-\frac{1}{2}}m_{2,\varepsilon}^{\frac{1}{2}+\frac{\gamma^{\prime}}{N}},&x\in\mathbb{R}^{N},\\ \Delta m_{1,\varepsilon}+\nabla\cdot{{w}}_{1,\varepsilon}=0,~{}{{w}}_{1,\varepsilon}=-\gamma C_{H}m_{1,\varepsilon}|\nabla u_{1,\varepsilon}|^{\gamma-2}\nabla u_{1,\varepsilon}&x\in\mathbb{R}^{N},\\ -\Delta u_{2,\varepsilon}+C_{H}|\nabla u_{2,\varepsilon}|^{\gamma}+\lambda_{2}\varepsilon^{\gamma^{\prime}}=\varepsilon^{\gamma^{\prime}}V_{2}(\varepsilon x+x_{1,\varepsilon})-\alpha_{2}m_{2,\varepsilon}^{\frac{\gamma^{\prime}}{N}}-\beta m_{2,\varepsilon}^{\frac{\gamma^{\prime}}{2N}-\frac{1}{2}}m_{1,\varepsilon}^{\frac{1}{2}+\frac{\gamma^{\prime}}{N}},&x\in\mathbb{R}^{N},\\ \Delta m_{2,\varepsilon}+\nabla\cdot{{w}}_{2,\varepsilon}=0,~{}{{w}}_{2,\varepsilon}=-\gamma C_{H}m_{2,\varepsilon}|\nabla u_{2,\varepsilon}|^{\gamma-2}\nabla u_{2,\varepsilon},&x\in\mathbb{R}^{N},\\ \int_{\mathbb{R}^{N}}m_{1,\varepsilon}\,dx=\int_{\mathbb{R}^{N}}m_{2,\varepsilon}\,dx=1.\end{array}\right. (4.14)

Without loss of the generality, we assume

infxNu1,a(x)=infxNu2,a(x)=0.\displaystyle\inf_{x\in\mathbb{R}^{N}}u_{1,\textbf{a}}(x)=\inf_{x\in\mathbb{R}^{N}}u_{2,\textbf{a}}(x)=0.

In light of (1.34), (1.35) and (1.38), one finds

supε0+CLN|wi,εmi,ε|γmi,ε𝑑x<+,i=1,2.\displaystyle\sup_{\varepsilon\rightarrow 0^{+}}C_{L}\int_{\mathbb{R}^{N}}\bigg{|}\frac{w_{i,\varepsilon}}{m_{i,\varepsilon}}\bigg{|}^{\gamma^{\prime}}m_{i,\varepsilon}\,dx<+\infty,~{}i=1,2.

Then it follows from Lemma 2.4 that for i=1,2i=1,2

supε0+mi,εW1,γ(N)<+,supε0+wi,εL1(N)<+,supε0+wi,εLγ(N)<+.\displaystyle\sup_{\varepsilon\rightarrow 0^{+}}\|m_{i,\varepsilon}\|_{W^{1,\gamma^{\prime}}(\mathbb{R}^{N})}<+\infty,~{}~{}\sup_{\varepsilon\rightarrow 0^{+}}\|w_{i,\varepsilon}\|_{L^{1}(\mathbb{R}^{N})}<+\infty,~{}~{}\sup_{\varepsilon\rightarrow 0^{+}}\|w_{i,\varepsilon}\|_{L^{\gamma^{\prime}}(\mathbb{R}^{N})}<+\infty. (4.15)

Invoking (1.32) and (1.35) , one finds for i=1,2i=1,2,

limε0+NVi(εx+x1,ε)mi,ε(x)𝑑x=0,\displaystyle\lim_{\varepsilon\rightarrow 0^{+}}\int_{\mathbb{R}^{N}}V_{i}(\varepsilon x+x_{1,\varepsilon})m_{i,\varepsilon}(x)\,dx=0, (4.16)

and

limε0+N(m1,ε12+γ2Nm2,ε12+γ2N)2𝑑x=0.\displaystyle\lim_{\varepsilon\rightarrow 0^{+}}\int_{\mathbb{R}^{N}}\bigg{(}m_{1,\varepsilon}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}-m_{2,\varepsilon}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\bigg{)}^{2}\,dx=0. (4.17)

By using the standard Sobolev embedding, we have from (4.15) and (4.17) that

mi,εmin W1,γ(N),mi,εm0, a.e. in N.\displaystyle m_{i,\varepsilon}\rightharpoonup m~{}\text{in~{}}W^{1,\gamma^{\prime}}(\mathbb{R}^{N}),~{}~{}m_{i,\varepsilon}\rightarrow m\geq 0,\text{~{}a.e.~{}in~{}}\mathbb{R}^{N}. (4.18)

Moreover, by using the Morrey’s embedding W1,γ(N)C0,θ(N)W^{1,\gamma^{\prime}}(\mathbb{R}^{N})\hookrightarrow C^{0,\theta}(\mathbb{R}^{N}) with θ(0,1γN)\theta\in(0,1-\frac{\gamma^{\prime}}{N}), one finds

mi,εm in C0,θ(N),supε0+mi,εC0,θ(N)<+,i=1,2.\displaystyle m_{i,\varepsilon}\rightarrow m\text{~{}in~{}}C^{0,\theta}(\mathbb{R}^{N}),~{}~{}\sup_{\varepsilon\rightarrow 0^{+}}\|m_{i,\varepsilon}\|_{C^{0,\theta}(\mathbb{R}^{N})}<+\infty,~{}~{}i=1,2. (4.19)

Recall that u1,a(x1,ε)=infxNu1,a(x)=0,u_{1,\textbf{a}}(x_{1,\varepsilon})=\inf_{x\in\mathbb{R}^{N}}u_{1,\textbf{a}}(x)=0, then we have u1,ε(0)=infxNu1,εu_{1,\varepsilon}(0)=\inf_{x\in\mathbb{R}^{N}}u_{1,\varepsilon}. Moreover, by applying the maximum principle, one gets from the first equation of (4.14) and (4.19) that

λ1εγ\displaystyle\lambda_{1}\varepsilon^{\gamma^{\prime}}\geq α1m1,εγN(0)βm2,ε12+γ2N(0)m1,εγ2N12(0)\displaystyle-\alpha_{1}m_{1,\varepsilon}^{\frac{\gamma^{\prime}}{N}}(0)-\beta m_{2,\varepsilon}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}(0)m^{\frac{\gamma^{\prime}}{2N}-\frac{1}{2}}_{1,\varepsilon}(0)
=\displaystyle= (α1+β)m1,εγN(0)+oε(1),\displaystyle-(\alpha_{1}+\beta)m_{1,\varepsilon}^{\frac{\gamma^{\prime}}{N}}(0)+o_{\varepsilon}(1),

In addition, noting (4.8) and α1+βa\alpha_{1}+\beta\nearrow a^{*}, we have

limε0m1,εγN(0)γNa.\displaystyle\lim_{\varepsilon\rightarrow 0}m_{1,\varepsilon}^{\frac{\gamma^{\prime}}{N}}(0)\geq\frac{\gamma^{\prime}}{Na^{*}}.

Recall again W1,γ(N)C0,θ(N),W^{1,\gamma^{\prime}}(\mathbb{R}^{N})\hookrightarrow C^{0,\theta}(\mathbb{R}^{N}), we have from (4.17) that for i=1,2,i=1,2, there exists R0>0R_{0}>0 and C>0C>0 such that

mi,ε(x)C>0, |x|<R0, for i=1,2.\displaystyle m_{i,\varepsilon}(x)\geq C>0,\text{~{}~{}}\forall|x|<R_{0},\text{ for }i=1,2. (4.20)

Moreover, we utilize (4.16) and (4.20) to get up to a subsequence,

limε0x1,ε=x0, s.t. V1(x0)=0=V2(x0).\displaystyle{\lim_{\varepsilon\rightarrow 0}x_{1,\varepsilon}=x_{0},\text{~{}s.t.~{}}V_{1}(x_{0})=0=V_{2}(x_{0}).}

Combining (4.19) with (4.20), one also has

m(x)C>0, |x|<R0.\displaystyle m(x)\geq C>0,\text{~{}}\forall|x|<R_{0}. (4.21)

Next, we study the regularity of the value function uu. To this end, we rewrite the u1u_{1}-equation in (4.14) as

Δu1,ε+CH|u1,ε|γ=\displaystyle-\Delta u_{1,\varepsilon}+C_{H}|\nabla u_{1,\varepsilon}|^{\gamma}= λ1εγ+εγV1(εx+x1,ε)α1m1,εγNβm1,εγ2N12m2,ε12+γ2N\displaystyle-\lambda_{1}\varepsilon^{\gamma^{\prime}}+\varepsilon^{\gamma^{\prime}}V_{1}(\varepsilon x+x_{1,\varepsilon})-\alpha_{1}m_{1,\varepsilon}^{\frac{\gamma^{\prime}}{N}}-\beta m_{1,\varepsilon}^{\frac{\gamma^{\prime}}{2N}-\frac{1}{2}}m_{2,\varepsilon}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}
:=\displaystyle:= gε(x)Lloc(N)Cloc0,θ(N).\displaystyle g_{\varepsilon}(x)\in L^{\infty}_{\text{loc}}(\mathbb{R}^{N})\cap C_{\text{loc}}^{0,\theta}(\mathbb{R}^{N}). (4.22)

For R>0R>0 large enough, we have

gε(x)L(BR(0))<CR<+, |x|<2R,\displaystyle\|g_{\varepsilon}(x)\|_{L^{\infty}(B_{R}(0))}<C_{R}<+\infty,\text{~{}}\forall|x|<2R,

where CR>0C_{R}>0 is independent of ε.\varepsilon. Then it follows from (4) and Sobolev embedding that

|u1,ε(x)|CR,|x|<2R.\displaystyle|\nabla u_{1,\varepsilon}(x)|\leq C_{R},~{}\forall|x|<2R.

Since u1,ε(0)=0u_{1,\varepsilon}(0)=0, we further have

|u1,ε(x)|CR, |x|<2R.\displaystyle|u_{1,\varepsilon}(x)|\leq C_{R},\text{~{}}\forall|x|<2R.

By using the W2,pW^{2,p}- estimate, one gets

u1,εW2,p(BR+1(0))Cp,R(u1,εLp(B2R(0))+gεLp(B2R(0))+|u1,ε|γLp(B2R(0))),p>1,\displaystyle\|u_{1,\varepsilon}\|_{W^{2,p}(B_{R+1}(0))}\leq C_{p,R}\big{(}\|u_{1,\varepsilon}\|_{L^{p}(B_{2R}(0))}+\|g_{\varepsilon}\|_{L^{p}(B_{2R}(0))}+\||\nabla u_{1,\varepsilon}|^{\gamma}\|_{L^{p}(B_{2R}(0))}\big{)},~{}\forall p>1,

where Cp,R>0C_{p,R}>0 is a constant depending on pp and R.R. Let p>N,p>N, then we obtain

u1,εC1,θ1(BR+1(0))Cθ1,R<+,\|u_{1,\varepsilon}\|_{C^{1,\theta_{1}}(B_{R+1}(0))}\leq C_{\theta_{1},R}<+\infty,

for some θ1(0,1).\theta_{1}\in(0,1). Moreover, we rewrite (4) as

Δu1,ε=CH|u1,ε|γ+gεC1,θ2(BR+1(0)).\displaystyle-\Delta u_{1,\varepsilon}=-C_{H}|\nabla u_{1,\varepsilon}|^{\gamma}+g_{\varepsilon}\in C^{1,\theta_{2}}(B_{R+1}(0)).

One further deduces from the standard W2,pW^{2,p} estimate that

u1,εC2,θ3(BR(0))Cθ3,R<+.\displaystyle\|u_{1,\varepsilon}\|_{C^{2,\theta_{3}}(B_{R}(0))}\leq C_{\theta_{3},R}<+\infty.

Then by the standard diagonal procedure and Arzelà-Ascoli theorem, we have from (4.14), (4.15), (4.18) and (4.19) that there exist u1C2(N)u_{1}\in C^{2}(\mathbb{R}^{N}) and w1Lγ(N){{w}}_{1}\in L^{\gamma^{\prime}}(\mathbb{R}^{N}) such that

u1,εu1 in Cloc2(N),w1,εw1 in Lγ(N),\displaystyle u_{1,\varepsilon}\rightarrow u_{1}\text{~{}in~{}}C_{\text{loc}}^{2}(\mathbb{R}^{N}),~{}~{}{{w}}_{1,\varepsilon}\rightharpoonup{{w}}_{1}\text{~{}in~{}}L^{\gamma^{\prime}}(\mathbb{R}^{N}), (4.23)

and (u1,m,w1)(u_{1},m,{{w}}_{1}) satisfies

{Δu1+CH|u1|γγN=amγN,xN,Δm=γCH(m|u1|γ2u1)=w1,xN,0<Nm𝑑x1,\displaystyle\left\{\begin{array}[]{ll}-\Delta u_{1}+C_{H}|\nabla u_{1}|^{\gamma}-\frac{\gamma^{\prime}}{N}=-a^{*}m^{\frac{\gamma^{\prime}}{N}},&x\in\mathbb{R}^{N},\\ -\Delta m=\gamma C_{H}\nabla\cdot(m|\nabla u_{1}|^{\gamma-2}\nabla u_{1})=-\nabla\cdot{w}_{1},&x\in\mathbb{R}^{N},\\ 0<\int_{\mathbb{R}^{N}}m\,dx\leq 1,\end{array}\right.

where we have used (4.8) and (4.21). In addition, by Lemma 2.6 and (1.20), one finds

Nm𝑑x=1.\displaystyle\int_{\mathbb{R}^{N}}m\,dx=1. (4.24)

Thus, with the aid of (4.18), we obtain for i=1,2i=1,2, mi,εm in L1(N).m_{i,\varepsilon}\rightarrow m\text{~{}in~{}}L^{1}(\mathbb{R}^{N}). Moreover, (4.19) indicates

mi,εm in Lp(N),p1.\displaystyle m_{i,\varepsilon}\rightarrow m\text{~{}in~{}}L^{p}(\mathbb{R}^{N}),~{}\forall p\geq 1. (4.25)

This combine with (4.23) show that (1.39) holds for i=1i=1.

Next, to prove (1.36), we first recall that u2,a(x2,ε)=0=infxNu2,a(x)u_{2,\textbf{a}}(x_{2,\varepsilon})=0=\inf_{x\in\mathbb{R}^{N}}u_{2,\textbf{a}}(x). Then, we have from (4.14) and (1.38) that

λ2\displaystyle\lambda_{2}\geq V2(x2,ε)α2m2,aγN(x2,ε)βm2,aγ2N12(x2,ε)m1,a12+γN(x2,ε)\displaystyle V_{2}(x_{2,\varepsilon})-\alpha_{2}m_{2,\textbf{a}}^{\frac{\gamma^{\prime}}{N}}(x_{2,\varepsilon})-\beta m_{2,\textbf{a}}^{\frac{\gamma^{\prime}}{2N}-\frac{1}{2}}(x_{2,\varepsilon})m_{1,\textbf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{N}}(x_{2,\varepsilon})
\displaystyle\geq εγ[α2m2,εγN(x2,εx1,εε)βm2,εγ2N12(x2,εx1,εε)m1,ε12+γN(x2,εx1,εε)],\displaystyle\varepsilon^{-\gamma^{\prime}}\bigg{[}-\alpha_{2}m_{2,\varepsilon}^{\frac{\gamma^{\prime}}{N}}\bigg{(}\frac{x_{2,\varepsilon}-x_{1,\varepsilon}}{\varepsilon}\bigg{)}-\beta m_{2,\varepsilon}^{\frac{\gamma^{\prime}}{2N}-\frac{1}{2}}\bigg{(}\frac{x_{2,\varepsilon}-x_{1,\varepsilon}}{\varepsilon}\bigg{)}m_{1,\varepsilon}^{\frac{1}{2}+\frac{\gamma^{\prime}}{N}}\bigg{(}\frac{x_{2,\varepsilon}-x_{1,\varepsilon}}{\varepsilon}\bigg{)}\bigg{]},

which implies

α2m2,εγN(x2,εx1,εε)+βm2,εγ2N12(x2,εx1,εε)m1,ε12+γN(x2,εx1,εε)εγλ2γ2N as ε0.\displaystyle\alpha_{2}m_{2,\varepsilon}^{\frac{\gamma^{\prime}}{N}}\bigg{(}\frac{x_{2,\varepsilon}-x_{1,\varepsilon}}{\varepsilon}\bigg{)}+\beta m_{2,\varepsilon}^{\frac{\gamma^{\prime}}{2N}-\frac{1}{2}}\bigg{(}\frac{x_{2,\varepsilon}-x_{1,\varepsilon}}{\varepsilon}\bigg{)}m_{1,\varepsilon}^{\frac{1}{2}+\frac{\gamma^{\prime}}{N}}\bigg{(}\frac{x_{2,\varepsilon}-x_{1,\varepsilon}}{\varepsilon}\bigg{)}\geq\varepsilon^{\gamma^{\prime}}\lambda_{2}\geq\frac{\gamma^{\prime}}{2N}\text{~{}as~{}}\varepsilon\rightarrow 0. (4.26)

Combining (4.19) with (4.25), one can easily check that for i=1,2i=1,2

lim|x|+mi,ε(x)=0 uniformly in ε.\displaystyle\lim_{|x|\rightarrow+\infty}m_{i,\varepsilon}(x)=0\text{~{}uniformly~{}in~{}}\varepsilon.

Combining this with (4.26), one has (1.36) holds.

We next similarly show that there exist u2C2(N)u_{2}\in C^{2}(\mathbb{R}^{N}) and w2Lγ(N){w}_{2}\in L^{\gamma^{\prime}}(\mathbb{R}^{N}) such that

u2,εu2 in Cloc2(N), and w2,εw2 in Lγ(N),\displaystyle u_{2,\varepsilon}\rightarrow u_{2}\text{~{}in~{}}C_{\text{loc}}^{2}(\mathbb{R}^{N}),\text{~{}and~{}}{\textbf{w}}_{2,\varepsilon}\rightharpoonup{\textbf{w}}_{2}\text{~{}in~{}}L^{\gamma^{\prime}}(\mathbb{R}^{N}),

and (u2,m,w2)(u_{2},m,{w}_{2}) satisfies (1.43), in which (m,w2)(m,{w}_{2}) is a minimizer of (1.15). Indeed, we rewrite the u2u_{2}-equation in (4.14) as

Δu2,ε+CH|u2,ε|γ\displaystyle-\Delta u_{2,\varepsilon}+C_{H}|\nabla u_{2,\varepsilon}|^{\gamma} =λ2εγ+εγV2(εx+x1,ε)α2m2,εγNβm2,εγ2N12m1,ε12+γ2N\displaystyle=-\lambda_{2}\varepsilon^{\gamma^{\prime}}+\varepsilon^{\gamma^{\prime}}V_{2}(\varepsilon x+x_{1,\varepsilon})-\alpha_{2}m_{2,\varepsilon}^{\frac{\gamma^{\prime}}{N}}-\beta m_{2,\varepsilon}^{\frac{\gamma^{\prime}}{2N}-\frac{1}{2}}m_{1,\varepsilon}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}
:=hεLloc(N)Cloc0,θ4(N).\displaystyle:=h_{\varepsilon}\in L^{\infty}_{\text{loc}}(\mathbb{R}^{N})\cap C_{\text{loc}}^{0,\theta_{4}}(\mathbb{R}^{N}). (4.27)

Moreover, by Lemma 2.1, one has for any R>0R>0 large enough,

|u2,ε(x)|CR<+,|x|<2R.\displaystyle|\nabla u_{2,\varepsilon}(x)|\leq C_{R}<+\infty,~{}\forall|x|<2R. (4.28)

In light of u2,ε(x2,εx1,εε)=0=infxNu2,ε(x),u_{2,\varepsilon}\bigg{(}\frac{x_{2,\varepsilon}-x_{1,\varepsilon}}{\varepsilon}\bigg{)}=0=\inf_{x\in\mathbb{R}^{N}}u_{2,\varepsilon}(x), we use (1.36) and (4.28) to get

|u2,ε(0)|CR|x2,εx1,εε|+|u2,ε(x2,εx1,εε)|C~R<+.\displaystyle|u_{2,\varepsilon}(0)|\leq C_{R}\bigg{|}\frac{x_{2,\varepsilon}-x_{1,\varepsilon}}{\varepsilon}\bigg{|}+\bigg{|}u_{2,\varepsilon}\bigg{(}\frac{x_{2,\varepsilon}-x_{1,\varepsilon}}{\varepsilon}\bigg{)}\bigg{|}\leq\tilde{C}_{R}<+\infty.

Thus, thanks to (4.28), we find

|u2,ε(x)|CR, |x|<2R.\displaystyle|u_{2,\varepsilon}(x)|\leq C_{R},\text{~{}~{}}\forall|x|<2R. (4.29)

Upon collecting (4), (4.28) and (4.29), one obtains

u2,εC2,θ5(BR(0))Cθ5,R<+.\displaystyle\|u_{2,\varepsilon}\|_{C^{2,\theta_{5}}(B_{R}(0))}\leq C_{\theta_{5},R}<+\infty.

Moreover, we similarly get (u2,m,w2)(u_{2},m,{w}_{2}) satisfies (1.43), in which (m,w2)(m,{w}_{2}) is a minimizer of (1.15). To finish the proof of (1.39), it remains to show that u1=u2u_{1}=u_{2} and w1=w2w_{1}=w_{2}, which can be obtained by following the argument shown in the proof of Theorem 2.4 in [18], Indeed, since (m,u1,λ)(m,u_{1},\lambda) and (m,u2,λ)(m,u_{2},\lambda) solve (1.43) with wi:=γm|ui|γ2uiw_{i}:=\gamma m|\nabla u_{i}|^{\gamma-2}\nabla u_{i}, we test the u1u2u_{1}-u_{2} equation and m1m2m_{1}-m_{2} equation against m1m2m_{1}-m_{2} and u1u2u_{1}-u_{2} and integrate them by parts, then subtract them to get a useful identity. With the aid of the strict convexity of |p|γ|p|^{\gamma}, γ>1\gamma>1, one has the conclusion u1=u2\nabla u_{1}=\nabla u_{2} and then w1=w2.w_{1}=w_{2}. By fixing the same minimum points of u1u_{1} and u2u_{2}, we obtain u1=u2.u_{1}=u_{2}.

Finally, proceeding the similar argument as shown in the proof of Case (ii), Theorem 1.4 in [12], we have (1.37) holds. ∎

Theorem 1.4 demonstrates that under mild assumptions (1.13) and (1.14), ground states are localized as 𝐚𝐚β\bf{a}\nearrow\bf{a}_{\beta}^{*}. We next present the proof of Theorem 1.5, which is for the refined asymptotic profiles of ground states. First of all, we establish the following upper bound of eα1,α2,βe_{\alpha_{1},\alpha_{2},\beta} given by (1.10):

Lemma 4.1.

Under the assumptions of Theorem 1.5, we have as (α1,α2)(aβ,aβ)(\alpha_{1},\alpha_{2})\nearrow(a^{*}-\beta,a^{*}-\beta),

0eα1,α2,β(γ+p0p0)(μν¯p0p0γ)γγ+p0(2a)p0γ+p0(aα1+α2+2β2)p0γ+p0(1+o(1)).\displaystyle 0\leq e_{\alpha_{1},\alpha_{2},\beta}\leq\bigg{(}\frac{\gamma^{\prime}+p_{0}}{p_{0}}\bigg{)}\bigg{(}\frac{\mu\bar{\nu}_{p_{0}}p_{0}}{\gamma^{\prime}}\bigg{)}^{\frac{\gamma^{\prime}}{\gamma^{\prime}+p_{0}}}\bigg{(}\frac{2}{a^{*}}\bigg{)}^{\frac{p_{0}}{\gamma^{\prime}+p_{0}}}\bigg{(}a^{*}-\frac{\alpha_{1}+\alpha_{2}+2\beta}{2}\bigg{)}^{\frac{p_{0}}{\gamma^{\prime}+p_{0}}}(1+o(1)). (4.30)
Proof.

From the definition of ν¯p0\bar{\nu}_{p_{0}} in (1.29), one can easily derive that, for any ν>ν¯p0\nu>\bar{\nu}_{p_{0}}, there exist (m0,w0)(m_{0},w_{0})\in\mathcal{M} and yNy\in\mathbb{R}^{N} such that

ν¯p0Hm0,p0(y)=N|x+y|p0m0(x)𝑑xν.\displaystyle\bar{\nu}_{p_{0}}\leq H_{m_{0},p_{0}}(y)=\int_{\mathbb{R}^{N}}|x+y|^{p_{0}}m_{0}(x)dx\leq\nu. (4.31)

Since (m0,w0)(m_{0},w_{0})\in\mathcal{M} is a minimizer of (1.15), we have from (1.15) and Lemma 2.6 that

Nm0𝑑x=1,CLN|w0m0|γm0𝑑x=1, and NN+γNm01+γN𝑑x=1a.\displaystyle\int_{\mathbb{R}^{N}}m_{0}\,dx=1,~{}C_{L}\int_{\mathbb{R}^{N}}\bigg{|}\frac{w_{0}}{m_{0}}\bigg{|}^{\gamma^{\prime}}m_{0}\,dx=1,\text{ and }\frac{N}{N+\gamma^{\prime}}\int_{\mathbb{R}^{N}}m^{1+\frac{\gamma^{\prime}}{N}}_{0}\,dx=\frac{1}{a^{*}}. (4.32)

Let xjZ0x_{j}\in Z_{0} with Z0Z_{0} given by (1.49), and define

mτ(x)=τNm0(τ(xxj)y),wτ(x)=τN+1w0(τ(xxj)y),\displaystyle m_{\tau}(x)=\tau^{N}m_{0}(\tau(x-x_{j})-y),~{}~{}w_{\tau}(x)=\tau^{N+1}w_{0}(\tau(x-x_{j})-y), (4.33)

then one finds from (4.32) and (4.33) that

CLN|wτmτ|γmτ𝑑x=τγCLN|w0m0|γm0𝑑x=τγ,\displaystyle C_{L}\int_{\mathbb{R}^{N}}\bigg{|}\frac{w_{\tau}}{m_{\tau}}\bigg{|}^{\gamma^{\prime}}m_{\tau}\,dx=\tau^{\gamma^{\prime}}C_{L}\int_{\mathbb{R}^{N}}\bigg{|}\frac{w_{0}}{m_{0}}\bigg{|}^{\gamma^{\prime}}m_{0}\,dx=\tau^{\gamma^{\prime}},
NN+γNmτ1+γN𝑑x=NN+γτγNm01+γN𝑑x=τγa,\displaystyle\frac{N}{N+\gamma^{\prime}}\int_{\mathbb{R}^{N}}m^{1+\frac{\gamma^{\prime}}{N}}_{\tau}\,dx=\frac{N}{N+\gamma^{\prime}}\tau^{\gamma^{\prime}}\int_{\mathbb{R}^{N}}m_{0}^{1+\frac{\gamma^{\prime}}{N}}\,dx=\frac{\tau^{\gamma^{\prime}}}{a^{*}},

and

N(V1+V2)mτ𝑑x=\displaystyle\int_{\mathbb{R}^{N}}(V_{1}+V_{2})m_{\tau}\,dx= N(V1+V2)(x+yτ+xj)m0(x)𝑑x\displaystyle\int_{\mathbb{R}^{N}}(V_{1}+V_{2})\bigg{(}\frac{x+y}{\tau}+x_{j}\bigg{)}m_{0}(x)\,dx
=\displaystyle= 1τp0N(V1+V2)(x+yτ+xj)|x+yτ|p0|x+y|pm0𝑑x.\displaystyle\frac{1}{\tau^{p_{0}}}\int_{\mathbb{R}^{N}}\frac{(V_{1}+V_{2})\big{(}\frac{x+y}{\tau}+x_{j}\big{)}}{\big{|}\frac{x+y}{\tau}\big{|}^{p_{0}}}|x+y|^{p}m_{0}\,dx. (4.34)

Note from (1.49) that

limτ+(V1+V2)(x+yτ+xj)|x+yτ|p0=μ.\displaystyle\lim_{\tau\rightarrow+\infty}\frac{(V_{1}+V_{2})\big{(}\frac{x+y}{\tau}+x_{j}\big{)}}{\big{|}\frac{x+y}{\tau}\big{|}^{p_{0}}}=\mu. (4.35)

Combining (4.31), (4) with (4.35), one can get as τ+\tau\to+\infty,

N(V1+V2)mτ(x)𝑑x=μντp0+O(1τp0).\displaystyle\int_{\mathbb{R}^{N}}(V_{1}+V_{2})m_{\tau}(x)\,dx=\frac{\mu\nu}{\tau^{p_{0}}}+O\bigg{(}\frac{1}{\tau^{p_{0}}}\bigg{)}.

Finally, by taking τ=(μνp0a2γ(aα1+α2+2β2))1γ+p0\tau=\bigg{(}\frac{\mu\nu p_{0}a^{*}}{2\gamma^{\prime}\big{(}a^{*}-\frac{\alpha_{1}+\alpha_{2}+2\beta}{2}\big{)}}\bigg{)}^{\frac{1}{\gamma^{\prime}+p_{0}}} in (1), we obtain

0eα1,α2,β\displaystyle 0\leq e_{\alpha_{1},\alpha_{2},\beta} (mτ,wτ,mτ,wτ)=τγ[2α1+α2+2βa]+μντp0+O(1τp0)\displaystyle\leq\mathcal{E}(m_{\tau},w_{\tau},m_{\tau},w_{\tau})=\tau^{\gamma^{\prime}}\bigg{[}2-\frac{\alpha_{1}+\alpha_{2}+2\beta}{a^{*}}\bigg{]}+\frac{\mu\nu}{\tau^{p_{0}}}+O\bigg{(}\frac{1}{\tau^{p_{0}}}\bigg{)}
=\displaystyle= γ+p0p0(μνp0γ)γγ+p0(2a)p0γ+p0(aα1+α2+2β2)p0γ+p0(1+oτ(1)),\displaystyle\frac{\gamma^{\prime}+p_{0}}{p_{0}}\bigg{(}\frac{\mu\nu p_{0}}{\gamma^{\prime}}\bigg{)}^{\frac{\gamma^{\prime}}{\gamma^{\prime}+p_{0}}}\bigg{(}\frac{2}{a^{*}}\bigg{)}^{\frac{p_{0}}{\gamma^{\prime}+p_{0}}}\bigg{(}a^{*}-\frac{\alpha_{1}+\alpha_{2}+2\beta}{2}\bigg{)}^{\frac{p_{0}}{\gamma^{\prime}+p_{0}}}(1+o_{\tau}(1)),

which indicates (4.30) since ν>ν¯p0\nu>\bar{\nu}_{p_{0}} is arbitrary. ∎

Now, we are ready to prove Theorem 1.5, which is

Proof of Theorem 1.5:

Proof.

In light of (1.38), we compute

eα1,α2,β=\displaystyle e_{\alpha_{1},\alpha_{2},\beta}= (m1,ε,w1,ε,m2,ε,w2,ε)\displaystyle\mathcal{E}(m_{1,\varepsilon},w_{1,\varepsilon},m_{2,\varepsilon},w_{2,\varepsilon})
=\displaystyle= i=12[εγCLN|wi,εmi,ε|γmi,ε𝑑xαiεγ1+γNNmi,ε1+γN𝑑x+NV(εx+xε)mi,ε𝑑x]\displaystyle\sum_{i=1}^{2}\bigg{[}\varepsilon^{-\gamma^{\prime}}C_{L}\int_{\mathbb{R}^{N}}\bigg{|}\frac{w_{i,\varepsilon}}{m_{i,\varepsilon}}\bigg{|}^{\gamma^{\prime}}m_{i,\varepsilon}\,dx-\frac{\alpha_{i}\varepsilon^{-\gamma^{\prime}}}{1+\frac{\gamma^{\prime}}{N}}\int_{\mathbb{R}^{N}}m_{i,\varepsilon}^{1+\frac{\gamma^{\prime}}{N}}\,dx+\int_{\mathbb{R}^{N}}V(\varepsilon x+x_{\varepsilon})m_{i,\varepsilon}\,dx\bigg{]}
2βεγ1+γNNm1,ε12+γ2Nm2,ε12+γ2N𝑑x,\displaystyle-\frac{2\beta\varepsilon^{-\gamma^{\prime}}}{1+\frac{\gamma^{\prime}}{N}}\int_{\mathbb{R}^{N}}m^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}_{1,\varepsilon}m^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}_{2,\varepsilon}\,dx, (4.36)

where we redefine x1,εx_{1,\varepsilon} as xεx_{\varepsilon} here and in the sequel for simplicity. Noting that V1(x0)=V2(x0)V_{1}(x_{0})=V_{2}(x_{0}) shown in Theorem 1.4, we find there exists some 1jl1\leq j\leq l such that x0=xjx_{0}=x_{j}. Then, we rewrite the potential energy as

NVi(εx+xε)mi,ε(x)𝑑x=εpjNVi(εx+xε)|εx+xεxj|pj|x+xεxjε|pjmi,ε𝑑x,\displaystyle\int_{\mathbb{R}^{N}}V_{i}(\varepsilon x+x_{\varepsilon})m_{i,\varepsilon}(x)\,dx=\varepsilon^{p_{j}}\int_{\mathbb{R}^{N}}\frac{V_{i}(\varepsilon x+x_{\varepsilon})}{|\varepsilon x+x_{\varepsilon}-x_{j}|^{p_{j}}}\bigg{|}x+\frac{x_{\varepsilon}-x_{j}}{\varepsilon}\bigg{|}^{p_{j}}m_{i,\varepsilon}\,dx, (4.37)

where i=1,2.i=1,2. In addition, since xεxjx_{\varepsilon}\rightarrow x_{j}, we obtain

limε0+i=12Vi(εx+xε)|εx+xεxj|pj=μj a.e. in N,\displaystyle\lim_{\varepsilon\rightarrow 0^{+}}\sum_{i=1}^{2}\frac{V_{i}(\varepsilon x+x_{\varepsilon})}{|\varepsilon x+x_{\varepsilon}-x_{j}|^{p_{j}}}=\mu_{j}\text{ a.e. in }\mathbb{R}^{N},

where μj\mu_{j} is defined in (1.48)\eqref{mujthm1point5}. Without loss of generality, we assume p2jp1j=pjp_{2j}\geq p_{1j}=p_{j} with p1jp_{1j} and p2jp_{2j} defined by (1.44).

Now, we claim that

pj=p0=max{p1,,pl}, and |xεxjε| is uniformly bounded as ε0+.\displaystyle p_{j}=p_{0}=\max\{p_{1},\cdots,p_{l}\},\text{ and }\big{|}\frac{x_{\varepsilon}-x_{j}}{\varepsilon}\big{|}\text{ is uniformly bounded as }\varepsilon\rightarrow 0^{+}. (4.38)

To show (4.38), we argue by contradiction and obtain either pj<p0p_{j}<p_{0} or up to a subsequence,

limε0+|xεxjε|=+.\displaystyle\lim_{\varepsilon\rightarrow 0^{+}}\bigg{|}\frac{x_{\varepsilon}-x_{j}}{\varepsilon}\bigg{|}=+\infty.

By using (4.37) and mi,εm0 in L1Lm_{i,\varepsilon}\rightarrow m_{0}\text{ in }L^{1}\cap L^{\infty} shown in Theorem 1.4, one deduces that for any Γ>0\Gamma>0 large enough,

limε0εp0NV1(εx+xε)m1,ε𝑑x\displaystyle\lim_{\varepsilon\rightarrow 0}\varepsilon^{-p_{0}}\int_{\mathbb{R}^{N}}V_{1}(\varepsilon x+x_{\varepsilon})m_{1,\varepsilon}\,dx
=\displaystyle= limε0εpjp0NV1(εx+xε)|εx+xεxj|pj|x+xεxjε|pjm1,ε𝑑xΓ.\displaystyle\lim_{\varepsilon\rightarrow 0}\varepsilon^{p_{j}-p_{0}}\int_{\mathbb{R}^{N}}\frac{V_{1}(\varepsilon x+x_{\varepsilon})}{|\varepsilon x+x_{\varepsilon}-x_{j}|^{p_{j}}}\big{|}x+\frac{x_{\varepsilon}-x_{j}}{\varepsilon}\big{|}^{p_{j}}m_{1,\varepsilon}\,dx\geq\Gamma. (4.39)

Recall the definition of ε\varepsilon shown in (1.35) and the estimate of (1.32), then we find

Nm1,ε12+γ2Nm2,ε12+γ2N𝑑x\displaystyle\int_{\mathbb{R}^{N}}m_{1,\varepsilon}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}m^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}_{2,\varepsilon}\,dx =Nm1,ε1+γN𝑑x+N(m2,ε12+γ2Nm1,ε12+γ2N)m1,ε12+γ2N𝑑x\displaystyle=\int_{\mathbb{R}^{N}}m^{1+\frac{\gamma^{\prime}}{N}}_{1,\varepsilon}\,dx+\int_{\mathbb{R}^{N}}\bigg{(}m^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}_{2,\varepsilon}-m^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}_{1,\varepsilon}\bigg{)}m^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}_{1,\varepsilon}\,dx
=Nm1,ε1+γN𝑑x+oε(1).\displaystyle=\int_{\mathbb{R}^{N}}m^{1+\frac{\gamma^{\prime}}{N}}_{1,\varepsilon}\,dx+o_{\varepsilon}(1). (4.40)

Thus, one finds

CLN|w1,εm1,ε|γm1,ε𝑑xα11+γNNm1,ε1+γN𝑑x2β1+γNNm1,ε12+γ2Nm2,ε12+γ2N𝑑x\displaystyle C_{L}\int_{\mathbb{R}^{N}}\bigg{|}\frac{w_{1,\varepsilon}}{m_{1,\varepsilon}}\bigg{|}^{\gamma^{\prime}}m_{1,\varepsilon}\,dx-\frac{\alpha_{1}}{1+\frac{\gamma^{\prime}}{N}}\int_{\mathbb{R}^{N}}m_{1,\varepsilon}^{1+\frac{\gamma^{\prime}}{N}}\,dx-\frac{2\beta}{1+\frac{\gamma^{\prime}}{N}}\int_{\mathbb{R}^{N}}m_{1,\varepsilon}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}m_{2,\varepsilon}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\,dx
=\displaystyle= CLN|w1,εm1,ε|γm1,ε𝑑xα1+2β1+γNNm1,ε1+γN𝑑x+oε(1)\displaystyle C_{L}\int_{\mathbb{R}^{N}}\big{|}\frac{w_{1,\varepsilon}}{m_{1,\varepsilon}}\big{|}^{\gamma^{\prime}}m_{1,\varepsilon}\,dx-\frac{\alpha_{1}+2\beta}{1+\frac{\gamma^{\prime}}{N}}\int_{\mathbb{R}^{N}}m_{1,\varepsilon}^{1+\frac{\gamma^{\prime}}{N}}\,dx+o_{\varepsilon}(1)
\displaystyle\geq (1α1+2βa)N|w1,εm1,ε|γm1,ε𝑑x+oε(1).\displaystyle\bigg{(}1-\frac{\alpha_{1}+2\beta}{a^{*}}\bigg{)}\int_{\mathbb{R}^{N}}\bigg{|}\frac{w_{1,\varepsilon}}{m_{1,\varepsilon}}\bigg{|}^{\gamma^{\prime}}m_{1,\varepsilon}\,dx+o_{\varepsilon}(1). (4.41)

In addition, in light of (1.34) and (1.35), one has

N|wi,εmi,ε|γmi,ε𝑑x=1+oε(1),i=1,2,\displaystyle\int_{\mathbb{R}^{N}}\bigg{|}\frac{w_{i,\varepsilon}}{m_{i,\varepsilon}}\bigg{|}^{\gamma^{\prime}}m_{i,\varepsilon}\,dx=1+o_{\varepsilon}(1),~{}i=1,2, (4.42)

and obtain from (4) that

N|w2,εm2,ε|γm2,ε𝑑xα21+γNNm2,ε1+γN𝑑x(1α2a)N|w2,εm2,ε|γm2,ε𝑑x.\displaystyle\int_{\mathbb{R}^{N}}\bigg{|}\frac{w_{2,\varepsilon}}{m_{2,\varepsilon}}\bigg{|}^{\gamma^{\prime}}m_{2,\varepsilon}\,dx-\frac{\alpha_{2}}{1+\frac{\gamma^{\prime}}{N}}\int_{\mathbb{R}^{N}}m^{1+\frac{\gamma^{\prime}}{N}}_{2,\varepsilon}\,dx\geq\bigg{(}1-\frac{\alpha_{2}}{a^{*}}\bigg{)}\int_{\mathbb{R}^{N}}\bigg{|}\frac{w_{2,\varepsilon}}{m_{2,\varepsilon}}\bigg{|}^{\gamma^{\prime}}m_{2,\varepsilon}\,dx. (4.43)

Upon substituting (4), (4) and (4.43), (4.42), one finds from (4) that

(m1,ε,w1,ε,m2,ε,w2,ε)\displaystyle\mathcal{E}(m_{1,\varepsilon},w_{1,\varepsilon},m_{2,\varepsilon},w_{2,\varepsilon})\geq εγ[1α1+α2+2βa](1+o(1))+Γεp0\displaystyle\varepsilon^{-\gamma^{\prime}}\bigg{[}1-\frac{\alpha_{1}+\alpha_{2}+2\beta}{a^{*}}\bigg{]}(1+o(1))+\Gamma\varepsilon^{-p_{0}}
\displaystyle\geq (1+oε(1))p0+γp0(p0Γγ)γγ+p0(2a)γγ+p0(aα1+α2+2β2)p0γ+p0,\displaystyle(1+o_{\varepsilon}(1))\frac{p_{0}+\gamma^{\prime}}{p_{0}}\bigg{(}\frac{p_{0}\Gamma}{\gamma^{\prime}}\bigg{)}^{\frac{\gamma^{\prime}}{\gamma^{\prime}+p_{0}}}\bigg{(}\frac{2}{a^{*}}\bigg{)}^{\frac{\gamma^{\prime}}{\gamma^{\prime}+p_{0}}}\bigg{(}a^{*}-\frac{\alpha_{1}+\alpha_{2}+2\beta}{2}\bigg{)}^{\frac{p_{0}}{\gamma^{\prime}+p_{0}}},

which is contradicted to Lemma 4.1. This completes the proof of claim (4.38). Hence, we obtain y0N\exists y_{0}\in\mathbb{R}^{N} such that

limε0xεxjε=y0.\displaystyle\lim_{\varepsilon\rightarrow 0}\frac{x_{\varepsilon}-x_{j}}{\varepsilon}=y_{0}.

We next show that y0y_{0} satisfies (1.51). Since pi=p0p_{i}=p_{0}, it follows from Theorem 1.4 that

limε0+εp0Ni=12Vi(εx+xε)mi,ε(x)dx\displaystyle\lim_{\varepsilon\rightarrow 0^{+}}\varepsilon^{-p_{0}}\int_{\mathbb{R}^{N}}\sum_{i=1}^{2}V_{i}(\varepsilon x+x_{\varepsilon})m_{i,\varepsilon}(x)\,dx
=\displaystyle= limε0+Ni=12Vi(ε(x+xxjε)+xj)|ε(x+xxjε)|p0|x+xxεε|p0mi,ε𝑑x\displaystyle\lim_{\varepsilon\rightarrow 0^{+}}\int_{\mathbb{R}^{N}}\frac{\sum_{i=1}^{2}V_{i}\bigg{(}\varepsilon(x+\frac{x-x_{j}}{\varepsilon})+x_{j}\bigg{)}}{\big{|}\varepsilon(x+\frac{x-x_{j}}{\varepsilon})\big{|}^{p_{0}}}\bigg{|}x+\frac{x-x_{\varepsilon}}{\varepsilon}\bigg{|}^{p_{0}}m_{i,\varepsilon}\,dx
\displaystyle\geq μjN|x+y0|p0m0𝑑xμν¯p0,\displaystyle\mu_{j}\int_{\mathbb{R}^{N}}|x+y_{0}|^{p_{0}}m_{0}\,dx\geq\mu\bar{\nu}_{p_{0}}, (4.44)

where the last two inequalities hold if and only if one has (1.51). As a consequence, we deduce from (4) and (4.43) that

eα1,α2,β\displaystyle e_{\alpha_{1},\alpha_{2},\beta}\geq εγ[2α1+α2+2βa][1+o(1)]+εp0μν¯p0[1+o(1)]\displaystyle\varepsilon^{-\gamma^{\prime}}\bigg{[}2-\frac{\alpha_{1}+\alpha_{2}+2\beta}{a^{*}}\bigg{]}[1+o(1)]+\varepsilon^{p_{0}}\mu\bar{\nu}_{p_{0}}[1+o(1)]
\displaystyle\geq [1+o(1)][γ+p0p0(μν¯p0p0γ)γγ+p0(2a)p0γ+p0(aα1+α2+2β2)p0γ+p0],\displaystyle[1+o(1)]\bigg{[}\frac{\gamma^{\prime}+p_{0}}{p_{0}}\bigg{(}\frac{\mu\bar{\nu}_{p_{0}}p_{0}}{\gamma^{\prime}}\bigg{)}^{\frac{\gamma^{\prime}}{\gamma^{\prime}+p_{0}}}\bigg{(}\frac{2}{a^{*}}\bigg{)}^{\frac{p_{0}}{\gamma^{\prime}+p_{0}}}\bigg{(}a^{*}-\frac{\alpha_{1}+\alpha_{2}+2\beta}{2}\bigg{)}^{\frac{p_{0}}{\gamma^{\prime}+p_{0}}}\bigg{]}, (4.45)

where the equality in the second inequality holds if and only if

ε=(2γp0μν¯p0a)1γ+p0(aα1+α2+2β2)1γ+p0(1+oε(1)).\displaystyle\varepsilon=\bigg{(}\frac{2\gamma^{\prime}}{p_{0}\mu\bar{\nu}_{p_{0}}a^{*}}\bigg{)}^{\frac{1}{\gamma^{\prime}+p_{0}}}\bigg{(}a^{*}-\frac{\alpha_{1}+\alpha_{2}+2\beta}{2}\bigg{)}^{\frac{1}{\gamma^{\prime}+p_{0}}}\big{(}1+o_{\varepsilon}(1)\big{)}.

Combining the lower bound (4) with the upper bound (4.30), we find the equalities in (4) and (4) hold. As a consequence, we obtain (1.50) and (1.51) and finish the proof of this theorem. ∎

5 Asymptotic Profiles of Ground States with β<0\beta<0

In this section, we shall discuss the concentration phenomena within (1.7) under the repulsive case with β<0.\beta<0. Similarly as shown in Section 4, we first investigate the basic blow-up profiles of ground states with some assumptions imposed on the potentials, which is summarized as Theorem 1.6. Then, we investigate the refined blow-up profiles shown in Theorem 1.7 when potentials satisfy local polynomial expansions.

Proof of Theorem 1.6:

Proof.

As shown in the proof of Theorem 1.1, we have proved that when β<0\beta<0,

limaaeα1,α2,β=0.\displaystyle\lim_{\textbf{a}\nearrow{\textbf{a}}^{*}}e_{\alpha_{1},\alpha_{2},\beta}=0. (5.1)

In addition, one obtains from (1.20) that

αii(mi,wi)0 if αi<a,\displaystyle\mathcal{E}_{\alpha_{i}}^{i}(m_{i},w_{i})\geq 0\text{~{}if~{}}\alpha_{i}<a^{*},

where αii(mi,wi),\mathcal{E}_{\alpha_{i}}^{i}(m_{i},w_{i}), i=1,2i=1,2 are given by (3.2). Moreover, noting that α1,α2,β(m1,w1,m2,w2)\mathcal{E}_{\alpha_{1},\alpha_{2},\beta}(m_{1},w_{1},m_{2},w_{2}) defined by (1) can be written as

α1,α2,β(m1,w1,m2,w2)=i=12αii(mi,wi)2βNN+γNm112+γ2Nm212+γ2N𝑑x,\displaystyle\mathcal{E}_{\alpha_{1},\alpha_{2},\beta}(m_{1},w_{1},m_{2},w_{2})=\sum_{i=1}^{2}\mathcal{E}_{\alpha_{i}}^{i}(m_{i},w_{i})-\frac{2\beta N}{N+\gamma^{\prime}}\int_{\mathbb{R}^{N}}m_{1}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}m_{2}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\,dx,

we find from (5.1) that (1.53), (1.54) and (1.55) hold.

Next, we shall prove (1.56) and argue by contradiction. Assume that

lim supaaNCL|wi,ami,a|γmi,a𝑑x<+, for i=1 or 2,\displaystyle\limsup_{\textbf{a}\nearrow\textbf{a}^{*}}\int_{\mathbb{R}^{N}}C_{L}\bigg{|}\frac{w_{i,\textbf{a}}}{m_{i,\textbf{a}}}\bigg{|}^{\gamma^{\prime}}m_{i,\textbf{a}}\,dx<+\infty,\text{ for }i=1\text{ or }2,

then it follows from (1.20) that

lim supaaNmi,a1+γN𝑑x<+.\displaystyle\limsup_{\textbf{a}\nearrow\textbf{a}^{*}}\int_{\mathbb{R}^{N}}m_{i,\textbf{a}}^{1+\frac{\gamma^{\prime}}{N}}\,dx<+\infty.

Therefore, we deduce from (1.53) that

limaaai(m1,a,w1,a,m2,a,w2,a)=limaaαii(m1,a,w1,a,m2,a,w2,a)=0=eai.\displaystyle\lim_{\textbf{a}\nearrow\textbf{a}^{*}}\mathcal{E}_{\textbf{a}^{*}}^{i}(m_{1,\textbf{a}},w_{1,\textbf{a}},m_{2,\textbf{a}},w_{2,\textbf{a}})=\lim_{\textbf{a}\nearrow\textbf{a}^{*}}\mathcal{E}_{\alpha_{i}}^{i}(m_{1,\textbf{a}},w_{1,\textbf{a}},m_{2,\textbf{a}},w_{2,\textbf{a}})=0=e^{i}_{a^{*}}.

This implies that {(mi,a,wi,a)}\{(m_{i,\textbf{a}},w_{i,\textbf{a}})\} is a is a bounded minimizing sequence of eaie^{i}_{a^{*}} given by (3.1) and its limit is a minimizer of eaie^{i}_{a^{*}}. This is a contradiction to the fact that eaie^{i}_{a^{*}} does not admit any minimizer as shown in [12]. Hence, one finds (1.56) holds.

Let

ε^i:=(CLN|wi,ami,a|γmi,a𝑑x)1γ0 as aa.\displaystyle\hat{\varepsilon}_{i}:=\bigg{(}C_{L}\int_{\mathbb{R}^{N}}\bigg{|}\frac{w_{i,\textbf{a}}}{m_{i,\textbf{a}}}\bigg{|}^{\gamma^{\prime}}m_{i,\textbf{a}}\,dx\bigg{)}^{-\frac{1}{\gamma^{\prime}}}\rightarrow 0\text{ as }\textbf{a}\nearrow\textbf{a}^{*}.

Recall that (m1,a,w1,a,m2,a,w2,a)𝒦(m_{1,\textbf{a}},w_{1,\textbf{a}},m_{2,\textbf{a}},w_{2,\textbf{a}})\in\mathcal{K} is a minimizer and by using Lemma 3.2, one has for i=1,2i=1,2,

λi,a=\displaystyle\lambda_{i,\textbf{a}}= CLN|wi,ami,a|γmi,a𝑑x+NVimi,a𝑑xαiNmi,a1+γN𝑑xβNm1,a12+γ2Nm2,a12+γ2N𝑑x\displaystyle C_{L}\int_{\mathbb{R}^{N}}\bigg{|}\frac{w_{i,\textbf{a}}}{m_{i,\textbf{a}}}\bigg{|}^{\gamma^{\prime}}m_{i,\textbf{a}}\,dx+\int_{\mathbb{R}^{N}}V_{i}m_{i,\textbf{a}}\,dx-\alpha_{i}\int_{\mathbb{R}^{N}}m_{i,\textbf{a}}^{1+\frac{\gamma^{\prime}}{N}}\,dx-\beta\int_{\mathbb{R}^{N}}m_{1,\textbf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}m_{2,\textbf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\,dx
=\displaystyle= αii(mi,a,wi,a)NαiN+γNmi,a1+γN𝑑xβNm1,a12+γ2Nm2,a12+γ2N𝑑x\displaystyle\mathcal{E}_{\alpha_{i}}^{i}(m_{i,\textbf{a}},w_{i,\textbf{a}})-\frac{N\alpha_{i}}{N+\gamma^{\prime}}\int_{\mathbb{R}^{N}}m_{i,\textbf{a}}^{1+\frac{\gamma^{\prime}}{N}}\,dx-\beta\int_{\mathbb{R}^{N}}m_{1,\textbf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}m_{2,\textbf{a}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\,dx
=\displaystyle= γNε^iγ+oεi(1),\displaystyle-\frac{\gamma^{\prime}}{N}\hat{\varepsilon}_{i}^{-\gamma^{\prime}}+o_{\varepsilon_{i}}(1),

which implies

λi,aε^iγγN as ε^i0+,i=1,2.\displaystyle\lambda_{i,\textbf{a}}\hat{\varepsilon}_{i}^{\gamma^{\prime}}\rightarrow-\frac{\gamma^{\prime}}{N}\text{~{}as~{}}\hat{\varepsilon}_{i}\rightarrow 0^{+},~{}i=1,2. (5.2)

Since (u1,a,u2,a)(u_{1,\textbf{a}},u_{2,\textbf{a}}) is bounded from below, we have ui,a(x)+u_{i,\textbf{a}}(x)\rightarrow+\infty as |x|+.|x|\rightarrow+\infty. Thus, there exist xi,ε^x_{i,\hat{\varepsilon}}, i=1,2i=1,2 such that

ui,ε^(0)=ui,a(xi,ε^)=infxNui,a(x).\displaystyle u_{i,\hat{\varepsilon}}(0)=u_{i,\textbf{a}}(x_{i,\hat{\varepsilon}})=\inf_{x\in\mathbb{R}^{N}}u_{i,\textbf{a}}(x).

By using (1.57) and (3.16), we find (m1,ε^,u1,ε^,m2,ε^,u2,ε^)(m_{1,\hat{\varepsilon}},u_{1,\hat{\varepsilon}},m_{2,\hat{\varepsilon}},u_{2,\hat{\varepsilon}}) satisfies

{Δu1,ε^+CH|u1,ε^|γ+λ1,aε^1γ=ε^1γV1(ε^1x+x1,ε^)α1m1,ε^γNβ(ε^1ε^2)γ2+N2m1,ε^γ2N12m2,ε^γ2N+12(ε^1x+x1,ε^x2,ε^ε^2),Δm1,ε^=CHγ(m1,ε^|u1,ε^|γ2u1,ε^)=w1,ε^,Δu2,ε^+CH|u2,ε^|γ+λ2,aε^2γ=ε^2γV2(ε^2x+x2,ε^)α2m2,ε^γNβ(ε^2ε^1)γ2+N2m2,ε^γ2N12m1,ε^γ2N+12(ε^2x+x2,ε^x1,ε^ε^1),Δm2,ε^=CHγ(m2,ε^|u2,ε^|γ2u2,ε^)=w2,ε^.\displaystyle\left\{\begin{array}[]{ll}-\Delta u_{1,\hat{\varepsilon}}+C_{H}|\nabla u_{1,\hat{\varepsilon}}|^{\gamma}+\lambda_{1,\textbf{a}}\hat{\varepsilon}_{1}^{\gamma^{\prime}}=\hat{\varepsilon}_{1}^{\gamma^{\prime}}V_{1}(\hat{\varepsilon}_{1}x+x_{1,\hat{\varepsilon}})-\alpha_{1}m_{1,\hat{\varepsilon}}^{\frac{\gamma^{\prime}}{N}}-\beta\big{(}\frac{\hat{\varepsilon}_{1}}{\hat{\varepsilon}_{2}}\big{)}^{\frac{\gamma^{\prime}}{2}+\frac{N}{2}}m_{1,\hat{\varepsilon}}^{\frac{\gamma^{\prime}}{2N}-\frac{1}{2}}m_{2,\hat{\varepsilon}}^{\frac{\gamma^{\prime}}{2N}+\frac{1}{2}}\big{(}\frac{\hat{\varepsilon}_{1}x+x_{1,\hat{\varepsilon}}-x_{2,\hat{\varepsilon}}}{\hat{\varepsilon}_{2}}\big{)},\\ -\Delta m_{1,\hat{\varepsilon}}=C_{H}\gamma\nabla\cdot(m_{1,\hat{\varepsilon}}|\nabla u_{1,\hat{\varepsilon}}|^{\gamma-2}\nabla u_{1,\hat{\varepsilon}})=-\nabla\cdot w_{1,\hat{\varepsilon}},\\ -\Delta u_{2,\hat{\varepsilon}}+C_{H}|\nabla u_{2,\hat{\varepsilon}}|^{\gamma}+\lambda_{2,\textbf{a}}\hat{\varepsilon}_{2}^{\gamma^{\prime}}=\hat{\varepsilon}_{2}^{\gamma^{\prime}}V_{2}(\hat{\varepsilon}_{2}x+x_{2,\hat{\varepsilon}})-\alpha_{2}m_{2,\hat{\varepsilon}}^{\frac{\gamma^{\prime}}{N}}-\beta\big{(}\frac{\hat{\varepsilon}_{2}}{\hat{\varepsilon}_{1}}\big{)}^{\frac{\gamma^{\prime}}{2}+\frac{N}{2}}m_{2,\hat{\varepsilon}}^{\frac{\gamma^{\prime}}{2N}-\frac{1}{2}}m_{1,\hat{\varepsilon}}^{\frac{\gamma^{\prime}}{2N}+\frac{1}{2}}\big{(}\frac{\hat{\varepsilon}_{2}x+x_{2,\hat{\varepsilon}}-x_{1,\hat{\varepsilon}}}{\hat{\varepsilon}_{1}}\big{)},\\ -\Delta m_{2,\hat{\varepsilon}}=C_{H}\gamma\nabla\cdot(m_{2,\hat{\varepsilon}}|\nabla u_{2,\hat{\varepsilon}}|^{\gamma-2}\nabla u_{2,\hat{\varepsilon}})=-\nabla\cdot w_{2,\hat{\varepsilon}}.\end{array}\right. (5.7)

Then by applying the maximum principle on (5.7), one finds for i,j=1,2i,j=1,2 and iji\not=j that

λi,aε^iγαimi,ε^γN(0)+ε^iγVi(ε^ix+xi,ε^)β(ε^iε^j)γ2+N2mi,ε^γ2N12(0)mj,ε^γ2N+12(xi,ε^xj,ε^ε^j),\displaystyle\lambda_{i,\textbf{a}}\hat{\varepsilon}_{i}^{\gamma^{\prime}}\geq-\alpha_{i}m_{i,\hat{\varepsilon}}^{\frac{\gamma^{\prime}}{N}}(0)+\hat{\varepsilon}_{i}^{\gamma^{\prime}}V_{i}(\hat{\varepsilon}_{i}x+x_{i,\hat{\varepsilon}})-\beta\bigg{(}\frac{\hat{\varepsilon}_{i}}{\hat{\varepsilon}_{j}}\bigg{)}^{\frac{\gamma^{\prime}}{2}+\frac{N}{2}}m_{i,\hat{\varepsilon}}^{\frac{\gamma^{\prime}}{2N}-\frac{1}{2}}(0)m_{j,\hat{\varepsilon}}^{\frac{\gamma^{\prime}}{2N}+\frac{1}{2}}\bigg{(}\frac{x_{i,\hat{\varepsilon}}-x_{j,\hat{\varepsilon}}}{\hat{\varepsilon}_{j}}\bigg{)},

Noting that αi>0\alpha_{i}>0, β<0\beta<0 and Vi0V_{i}\geq 0 with i=1,2i=1,2, we further have from (5.2) when αia,\alpha_{i}\nearrow a^{*},

Cmi,εγN(0)>γ2aN>0,\displaystyle C\geq m_{i,\varepsilon}^{\frac{\gamma^{\prime}}{N}}(0)>\frac{\gamma^{\prime}}{2a^{*}N}>0, (5.8)

where C>0C>0 is a constant. Invoking (1.55) and (1.56), we obtain

NVi(ε^ix+xi,ε^)mi,ε^(x)𝑑x0 as aa,\displaystyle\int_{\mathbb{R}^{N}}V_{i}(\hat{\varepsilon}_{i}x+x_{i,\hat{\varepsilon}})m_{i,\hat{\varepsilon}}(x)\,dx\rightarrow 0\text{~{}as~{}}\textbf{a}\nearrow\textbf{a}^{*}, (5.9)

and

NCL|wi,ε^mi,ε^|γmi,ε^𝑑x=1,Nmi,ε^1+γN𝑑xN+γNa.\displaystyle\int_{\mathbb{R}^{N}}C_{L}\bigg{|}\frac{w_{i,\hat{\varepsilon}}}{m_{i,\hat{\varepsilon}}}\bigg{|}^{\gamma^{\prime}}m_{i,\hat{\varepsilon}}\,dx=1,~{}~{}\int_{\mathbb{R}^{N}}m_{i,\hat{\varepsilon}}^{1+\frac{\gamma^{\prime}}{N}}\,dx\rightarrow\frac{N+\gamma^{\prime}}{Na^{*}}. (5.10)

Now, we claim up to a subsequence,

xi,ε^xi with Vi(xi)=0,i=1,2.\displaystyle x_{i,\hat{\varepsilon}}\rightarrow x_{i}\text{ with }V_{i}(x_{i})=0,~{}i=1,2. (5.11)

Indeed, we have from (5.10) and Lemma 2.4 that

lim supε^1,ε^20+mi,ε^W1,γ(N)<+.\displaystyle\limsup_{\hat{\varepsilon}_{1},\hat{\varepsilon}_{2}\rightarrow 0^{+}}\|m_{i,\hat{\varepsilon}}\|_{W^{1,\gamma^{\prime}}(\mathbb{R}^{N})}<+\infty. (5.12)

Moreover, since γ>N\gamma^{\prime}>N, one gets from Morrey’s estimate that

lim supε^1,ε^20+mi,ε^C0,1Nγ(N)<+.\displaystyle\limsup_{\hat{\varepsilon}_{1},\hat{\varepsilon}_{2}\rightarrow 0^{+}}\|m_{i,\hat{\varepsilon}}\|_{C^{0,1-\frac{N}{\gamma^{\prime}}}(\mathbb{R}^{N})}<+\infty. (5.13)

(5.13) together with (5.8) gives that there exists R>0R>0 such that

mi,ε^(x)C2>0,|x|<R,i=1,2,\displaystyle m_{i,\hat{\varepsilon}}(x)\geq\frac{C}{2}>0,~{}\forall|x|<R,~{}i=1,2, (5.14)

where C>0C>0 is a constant independent of ε^i\hat{\varepsilon}_{i}. As a consequence, we obtain claim (5.11) thanks to (5.9) and (5.14). In light of (1.52) and (5.11), one finds

limε^1,ε^20+|x1,ε^x2,ε^|ε^i=+,i=1,2.\displaystyle\lim_{\hat{\varepsilon}_{1},\hat{\varepsilon}_{2}\rightarrow 0^{+}}\frac{|x_{1,\hat{\varepsilon}}-x_{2,\hat{\varepsilon}}|}{\hat{\varepsilon}_{i}}=+\infty,~{}i=1,2.

Next, we study the convergence of (m1,ε^,u1,ε^,m2,ε^,u2,ε^)(m_{1,\hat{\varepsilon}},u_{1,\hat{\varepsilon}},m_{2,\hat{\varepsilon}},u_{2,\hat{\varepsilon}}) as ε^i0\hat{\varepsilon}_{i}\rightarrow 0 with i=1,2.i=1,2. First of all, we have from (5.12) and (5.14) that there exist 0,miW1,γ(N)0\not\equiv,\leq m_{i}\in W^{1,\gamma^{\prime}}(\mathbb{R}^{N}) with i=1,2i=1,2 such that

mi,ε^mi in W1,γ(N).\displaystyle m_{i,\hat{\varepsilon}}\rightharpoonup m_{i}\text{ in }W^{1,\gamma^{\prime}}(\mathbb{R}^{N}).

Without loss of the generality, we assume

ε^1ε^2.\displaystyle\hat{\varepsilon}_{1}\geq\hat{\varepsilon}_{2}. (5.15)

In light of (5.13) and (5.15), one has there exists C>0C>0 independent of ε^1\hat{\varepsilon}_{1} and ε^2\hat{\varepsilon}_{2} such that

β(ε^2ε^1)N2+γ2m1,ε^12+γ2N(ε^2x+x2,ε^x1,ε^ε^1)m2,ε^γ2N12(x)C for all xN.\displaystyle\beta\bigg{(}\frac{\hat{\varepsilon}_{2}}{\hat{\varepsilon}_{1}}\bigg{)}^{\frac{N}{2}+\frac{\gamma^{\prime}}{2}}m_{1,\hat{\varepsilon}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\bigg{(}\frac{\hat{\varepsilon}_{2}x+x_{2,\hat{\varepsilon}}-x_{1,\hat{\varepsilon}}}{\hat{\varepsilon}_{1}}\bigg{)}m_{2,\hat{\varepsilon}}^{\frac{\gamma^{\prime}}{2N}-\frac{1}{2}}(x)\leq C\text{ for all }x\in\mathbb{R}^{N}.

In addition, by using Lemma 2.1, one obtains for any xBR(0),x\in B_{R}(0),

|u2,ε^(x)|CR,\displaystyle|\nabla u_{2,\hat{\varepsilon}}(x)|\leq C_{R}, (5.16)

where CR>0C_{R}>0 is a constant. Moreover, the u2,ε^u_{2,\hat{\varepsilon}}-equation in (5.7) becomes

Δu2,ε^=CH|u2,ε^|γ+hε^(x),\displaystyle-\Delta u_{2,\hat{\varepsilon}}=-C_{H}|\nabla u_{2,\hat{\varepsilon}}|^{\gamma}+h_{\hat{\varepsilon}}(x),

where

hε^(x):=λ2,aε^2γ+ε^2γV2(ε^2x+x2,ε^)α2m2,ε^γNβ(ε^2ε^1)γ2+N2m2,ε^γ2N12m1,ε^γ2N+12(ε^2x+x2,ε^x1,ε^ε^1).h_{\hat{\varepsilon}}(x):=-\lambda_{2,\textbf{a}}\hat{\varepsilon}_{2}^{\gamma^{\prime}}+\hat{\varepsilon}_{2}^{\gamma^{\prime}}V_{2}(\hat{\varepsilon}_{2}x+x_{2,\hat{\varepsilon}})-\alpha_{2}m_{2,\hat{\varepsilon}}^{\frac{\gamma^{\prime}}{N}}-\beta\big{(}\frac{\hat{\varepsilon}_{2}}{\hat{\varepsilon}_{1}}\big{)}^{\frac{\gamma^{\prime}}{2}+\frac{N}{2}}m_{2,\hat{\varepsilon}}^{\frac{\gamma^{\prime}}{2N}-\frac{1}{2}}m_{1,\hat{\varepsilon}}^{\frac{\gamma^{\prime}}{2N}+\frac{1}{2}}\big{(}\frac{\hat{\varepsilon}_{2}x+x_{2,\hat{\varepsilon}}-x_{1,\hat{\varepsilon}}}{\hat{\varepsilon}_{1}}\big{)}.

is given by (4) with ε\varepsilon replaced by ε^\hat{\varepsilon}. We further find from (5.16) that |CH|u2,ε^|γ+gε^|C~R|-C_{H}|\nabla u_{2,\hat{\varepsilon}}|^{\gamma}+g_{\hat{\varepsilon}}|\leq\tilde{C}_{R} with C~R>0\tilde{C}_{R}>0. Then we apply the standard elliptic regularity to get u2,ε^C2,θ(BR)CR,\|u_{2,\hat{\varepsilon}}\|_{C^{2,\theta}(B_{R})}\leq C_{R}, where CR>0C_{R}>0 is a constant and θ(0,1).\theta\in(0,1). Thus, we take the limit in the u2,ε^u_{2,\hat{\varepsilon}}-equation and m2,ε^m_{2,\hat{\varepsilon}}-equation of (5.7), use the diagonalization procedure and Arzelà-Ascoli theorem to deduce that u2,ε^u2 in C2,θ^(N)u_{2,\hat{\varepsilon}}\rightarrow u_{2}\text{ in }C^{2,\hat{\theta}}(\mathbb{R}^{N}) as ε^1\hat{\varepsilon}_{1}, ε^20+,\hat{\varepsilon}_{2}\rightarrow 0^{+}, and (m2,u2)(m_{2},u_{2}) satisfies

{Δu2+CH|u2|γγN=am2,Δm2=CHγ(m2|u2|γ2u2)=w2,0<Nm2𝑑x1.\displaystyle\left\{\begin{array}[]{ll}-\Delta u_{2}+C_{H}|\nabla u_{2}|^{\gamma}-\frac{\gamma^{\prime}}{N}=a^{*}m_{2},\\ -\Delta m_{2}=C_{H}\gamma\nabla\cdot(m_{2}|\nabla u_{2}|^{\gamma-2}\nabla u_{2})=-\nabla\cdot w_{2},\\ 0<\int_{\mathbb{R}^{N}}m_{2}\,dx\leq 1.\end{array}\right.

Similar as the derivation of (4.24), one uses Lemma 2.6 to get Nm2𝑑x=1.\int_{\mathbb{R}^{N}}m_{2}\,dx=1. It follows that m2,ε^m2m_{2,\hat{\varepsilon}}\rightarrow m_{2} in L1(N)L^{1}(\mathbb{R}^{N}). Combining this with (5.13) , we deduce

m2,ε^m2 in Lq(N),q1.\displaystyle m_{2,\hat{\varepsilon}}\rightarrow m_{2}\text{ in }L^{q}(\mathbb{R}^{N}),\forall q\geq 1. (5.17)

Invoking Lemma 2.2, (5.2) and (5.13), one has

u2,ε^(x)Cmax{|x|,(ε2γV2(ε^2x+x2,ε^))1γ}, if |x|>R,\displaystyle u_{2,\hat{\varepsilon}}(x)\geq C\max\big{\{}|x|,\big{(}\varepsilon_{2}^{\gamma^{\prime}}V_{2}(\hat{\varepsilon}_{2}x+x_{2,\hat{\varepsilon}})\big{)}^{\frac{1}{\gamma}}\big{\}},\text{ if }|x|>R, (5.18)

where C>0C>0 and R>0R>0 are constants independent of ε^1\hat{\varepsilon}_{1} and ε^2\hat{\varepsilon}_{2}. Indeed, it suffices to prove u2,ε^(x)C|x|u_{2,\hat{\varepsilon}}(x)\geq C|x| for some constant C>0C>0 when |x|>R.|x|>R. To this end, we find from (5.7) that when ε^i\hat{\varepsilon}_{i}, i=1,2i=1,2 are small,

Δu2,ε^+CH|u2,ε^|γ+λ0γ3Nα2m2,ε^γNβ(ε^2ε^1)γ2+N2m2,ε^γ2N12m1,ε^γ2N+12(ε^2x+x2,ε^x1,ε^ε^1),\displaystyle-\Delta u_{2,\hat{\varepsilon}}+C_{H}|\nabla u_{2,\hat{\varepsilon}}|^{\gamma}+\lambda_{0}\geq\frac{\gamma^{\prime}}{3N}-\alpha_{2}m_{2,\hat{\varepsilon}}^{\frac{\gamma^{\prime}}{N}}-\beta\bigg{(}\frac{\hat{\varepsilon}_{2}}{\hat{\varepsilon}_{1}}\bigg{)}^{\frac{\gamma^{\prime}}{2}+\frac{N}{2}}m_{2,\hat{\varepsilon}}^{\frac{\gamma^{\prime}}{2N}-\frac{1}{2}}m_{1,\hat{\varepsilon}}^{\frac{\gamma^{\prime}}{2N}+\frac{1}{2}}\bigg{(}\frac{\hat{\varepsilon}_{2}x+x_{2,\hat{\varepsilon}}-x_{1,\hat{\varepsilon}}}{\hat{\varepsilon}_{1}}\bigg{)}, (5.19)

where λ0:=γ2N\lambda_{0}:=-\frac{\gamma^{\prime}}{2N} and we have used (5.2) and the positivity of V2V_{2}. In addition, (5.15) and (5.17) indicate that as |x|+,|x|\rightarrow+\infty,

α2m2,ε^γNβ(ε^2ε^1)γ2+N2m2,ε^γ2N12m1,ε^γ2N+12(ε^2x+x2,ε^x1,ε^ε^1)0 uniformly in ε^1 and ε^2.\displaystyle-\alpha_{2}m_{2,\hat{\varepsilon}}^{\frac{\gamma^{\prime}}{N}}-\beta\bigg{(}\frac{\hat{\varepsilon}_{2}}{\hat{\varepsilon}_{1}}\bigg{)}^{\frac{\gamma^{\prime}}{2}+\frac{N}{2}}m_{2,\hat{\varepsilon}}^{\frac{\gamma^{\prime}}{2N}-\frac{1}{2}}m_{1,\hat{\varepsilon}}^{\frac{\gamma^{\prime}}{2N}+\frac{1}{2}}\bigg{(}\frac{\hat{\varepsilon}_{2}x+x_{2,\hat{\varepsilon}}-x_{1,\hat{\varepsilon}}}{\hat{\varepsilon}_{1}}\bigg{)}\rightarrow 0\text{ uniformly in }\hat{\varepsilon}_{1}\text{ and }\hat{\varepsilon}_{2}. (5.20)

Thus, one further obtains from (5.19) and (5.20) that

Δu2,ε^+CH|u2,ε^|γ+λ0>0 when |x|1.\displaystyle-\Delta u_{2,\hat{\varepsilon}}+C_{H}|\nabla u_{2,\hat{\varepsilon}}|^{\gamma}+\lambda_{0}>0\text{ when }|x|\gg 1. (5.21)

Now, we fix any |x~||\tilde{x}| large enough and define

h(x):=K1|x~|χ(x|x~|),h(x):=K_{1}|\tilde{x}|\chi\bigg{(}\frac{x}{|\tilde{x}|}\bigg{)},

where constant K1>0K_{1}>0 will be chosen later and χ()0\chi(\cdot)\geq 0 denotes the smooth cut-off function satisfying χ0\chi\equiv 0 when x(0,12)(32,+).x\in(0,\frac{1}{2})\cup(\frac{3}{2},+\infty). We compute to get

Δh+CH|h|γ+λ0K1|x~|+CHK1γ+λ0<0,\displaystyle-\Delta h+C_{H}|\nabla h|^{\gamma}+\lambda_{0}\leq\frac{K_{1}}{|\tilde{x}|}+C_{H}K_{1}^{\gamma}+\lambda_{0}<0, (5.22)

if we choose K1K_{1} small enough. Applying the comparison principle into (5.21) and (5.22), one has

u2,ε^(x)h(x) for 12|x~|<|x|<32|x~|,\displaystyle u_{2,\hat{\varepsilon}}(x)\geq h(x)\text{ for }\frac{1}{2}|\tilde{x}|<|x|<\frac{3}{2}|\tilde{x}|,

which finishes the proof of (5.18).

Next, we claim that for any p>1p>1, there exist R>0R>0 and C>0C>0 such that

m2,ε^(x)C|x|p,|x|>R.m_{2,\hat{\varepsilon}}(x)\leq C|x|^{-p},~{}\forall|x|>R.

Indeed, let ϕ=u2,ε^p\phi=u_{2,\hat{\varepsilon}}^{p}, then we have

Δϕ+CHγ|u2,ε^|γ2u2,ε^ϕ\displaystyle-\Delta\phi+C_{H}\gamma|\nabla u_{2,\hat{\varepsilon}}|^{\gamma-2}\nabla u_{2,\hat{\varepsilon}}\cdot\nabla\phi
=\displaystyle= pu2,ε^p1[Δu2,ε^(p1)|u2,ε^|2u2,ε^+CHγ|u2,ε^|γ]\displaystyle pu^{p-1}_{2,\hat{\varepsilon}}[-\Delta u_{2,\hat{\varepsilon}}-(p-1)\frac{|\nabla u_{2,\hat{\varepsilon}}|^{2}}{u_{2,\hat{\varepsilon}}}+C_{H}\gamma|\nabla u_{2,\hat{\varepsilon}}|^{\gamma}]
=\displaystyle= pu2,ε^p1[CH(γ1)|u2,ε^|γλ2ε^2γ(p1)|u2,ε^|2u2,ε^\displaystyle pu_{2,\hat{\varepsilon}}^{p-1}\bigg{[}C_{H}(\gamma-1)|\nabla u_{2,\hat{\varepsilon}}|^{\gamma}-\lambda_{2}\hat{\varepsilon}_{2}^{\gamma^{\prime}}-(p-1)\frac{|\nabla u_{2,\hat{\varepsilon}}|^{2}}{u_{2,\hat{\varepsilon}}}
+ε^2γV2(ε^2x+x2,ε^)α2m2,ε^γNβ(ε^2ε^1)γ2+N2m2,ε^γ2N12m1,ε^γ2N+12(ε^2x+x2,ε^x1,ε^ε^1)]\displaystyle+\hat{\varepsilon}_{2}^{\gamma^{\prime}}V_{2}(\hat{\varepsilon}_{2}x+x_{2,\hat{\varepsilon}})-\alpha_{2}m_{2,\hat{\varepsilon}}^{\frac{\gamma^{\prime}}{N}}-\beta\bigg{(}\frac{\hat{\varepsilon}_{2}}{\hat{\varepsilon}_{1}}\bigg{)}^{\frac{\gamma^{\prime}}{2}+\frac{N}{2}}m_{2,\hat{\varepsilon}}^{\frac{\gamma^{\prime}}{2N}-\frac{1}{2}}m_{1,\hat{\varepsilon}}^{\frac{\gamma^{\prime}}{2N}+\frac{1}{2}}\bigg{(}\frac{\hat{\varepsilon}_{2}x+x_{2,\hat{\varepsilon}}-x_{1,\hat{\varepsilon}}}{\hat{\varepsilon}_{1}}\bigg{)}\bigg{]}
:=\displaystyle:= pu2,ε^p1Gε^(x).\displaystyle pu_{2,\hat{\varepsilon}}^{p-1}G_{\hat{\varepsilon}}(x). (5.23)

Lemma 2.1 implies

|u2,ε^|C[1+ε^2γV2(ε^2x+x2,ε^)]1γ.\displaystyle|\nabla u_{2,\hat{\varepsilon}}|\leq C\big{[}1+\hat{\varepsilon}^{\gamma^{\prime}}_{2}V_{2}(\hat{\varepsilon}_{2}x+x_{2,\hat{\varepsilon}})\big{]}^{\frac{1}{\gamma}}. (5.24)

Hence, we deduce from (5.18) that

|u2,ε^|2γu2,ε^C[1+ε^2γV2(ε^2x+x2,ε^)2γγ]max{|x|,[ε^2γV2(ε^2x+x2,ε^)]1γ}CH(γ1)2(p1),for |x|>R.\displaystyle\frac{|\nabla u_{2,\hat{\varepsilon}}|^{2-\gamma}}{u_{2,\hat{\varepsilon}}}\leq C\frac{\bigg{[}1+\hat{\varepsilon}_{2}^{\gamma^{\prime}}V_{2}(\hat{\varepsilon}_{2}x+x_{2,\hat{\varepsilon}})^{\frac{2-\gamma}{\gamma}}\bigg{]}}{\max\{|x|,[\hat{\varepsilon}_{2}^{\gamma^{\prime}}V_{2}(\hat{\varepsilon}_{2}x+x_{2,\hat{\varepsilon}})]^{\frac{1}{\gamma}}\}}\leq\frac{C_{H}(\gamma-1)}{2(p-1)},~{}\text{for~{}}|x|>R.

Thus,

CH(γ1)|u2,ε^|γ(p1)|u2,ε^|2u2,ε^\displaystyle C_{H}(\gamma-1)|\nabla u_{2,\hat{\varepsilon}}|^{\gamma}-(p-1)\frac{|\nabla u_{2,\hat{\varepsilon}}|^{2}}{u_{2,\hat{\varepsilon}}}
=\displaystyle= |u2,ε^|γ[CH(γ1)(p1)|u2,ε^|2γu2,ε^]>0for |x|>R.\displaystyle|\nabla u_{2,\hat{\varepsilon}}|^{\gamma}\bigg{[}C_{H}(\gamma-1)-(p-1)\frac{|\nabla u_{2,\hat{\varepsilon}}|^{2-\gamma}}{u_{2,\hat{\varepsilon}}}\bigg{]}>0~{}\text{for }|x|>R.

In light of (5), we further find

Δϕ+CHγ|u2,ε^|γ2u2,ε^ϕCpu2,ε^p1,for |x|>R.\displaystyle-\Delta\phi+C_{H}\gamma|\nabla u_{2,\hat{\varepsilon}}|^{\gamma-2}\nabla u_{2,\hat{\varepsilon}}\cdot\nabla\phi\geq Cpu^{p-1}_{2,\hat{\varepsilon}},~{}\text{for }|x|>R. (5.25)

By using Theorem 3.1 in [20], one gets Nm2,ε^u2,ε^p1𝑑x<+\int_{\mathbb{R}^{N}}m_{2,\hat{\varepsilon}}u^{p-1}_{2,\hat{\varepsilon}}\,dx<+\infty. We next show that

lim supε^1,ε^20+Nm2,ε^u2,ε^p1𝑑x<+.\displaystyle\limsup_{\hat{\varepsilon}_{1},\hat{\varepsilon}_{2}\rightarrow 0^{+}}\int_{\mathbb{R}^{N}}m_{2,\hat{\varepsilon}}u^{p-1}_{2,\hat{\varepsilon}}\,dx<+\infty. (5.26)

Indeed, we test the m2,ε^m_{2,\hat{\varepsilon}}-equation in (5.7) against ϕ\phi and integrate it by parts to obtain

0=Nm2,ε^[Δϕ+CHγ|u2,ε^|γ2u2,ε^ϕ]𝑑x=pNm2,ε^Gε^u2,ε^p1𝑑x.\displaystyle 0=\int_{\mathbb{R}^{N}}m_{2,\hat{\varepsilon}}[-\Delta\phi+C_{H}\gamma|\nabla u_{2,\hat{\varepsilon}}|^{\gamma-2}\nabla u_{2,\hat{\varepsilon}}\cdot\nabla\phi]\,dx=p\int_{\mathbb{R}^{N}}m_{2,\hat{\varepsilon}}G_{\hat{\varepsilon}}u_{2,\hat{\varepsilon}}^{p-1}\,dx.

It follows that for some large R1>0R_{1}>0 independent of ε^i\hat{\varepsilon}_{i}, i=1,2i=1,2,

{x||x|>R1}m2,ε^Gε^u2,ε^p1𝑑x={x||x|R1}m2,ε^Gε^u2,ε^p1𝑑x.\displaystyle\int_{\{x||x|>R_{1}\}}m_{2,\hat{\varepsilon}}G_{\hat{\varepsilon}}u_{2,\hat{\varepsilon}}^{p-1}\,dx=-\int_{\{x||x|\leq R_{1}\}}m_{2,\hat{\varepsilon}}G_{\hat{\varepsilon}}u_{2,\hat{\varepsilon}}^{p-1}\,dx. (5.27)

On one hand, in light of (5.25), one has

{x||x|>R1}m2,ε^u2,ε^p1𝑑xC{x||x|>R1}m2,ε^Gε^u2,ε^p1𝑑x,\displaystyle\int_{\{x||x|>R_{1}\}}m_{2,\hat{\varepsilon}}u_{2,\hat{\varepsilon}}^{p-1}\,dx\leq C\int_{\{x||x|>R_{1}\}}m_{2,\hat{\varepsilon}}G_{\hat{\varepsilon}}u_{2,\hat{\varepsilon}}^{p-1}\,dx, (5.28)

where C>0C>0 is some constant independent of ε^i\hat{\varepsilon}_{i}, i=1,2i=1,2. On the other hand, by fixing infxNu2,ε^=1\inf\limits_{x\in\mathbb{R}^{N}}u_{2,\hat{\varepsilon}}=1 in (5), we get |Gε^|C|G_{\hat{\varepsilon}}|\leq C for some constant C>0C>0 independent of ε^i,i=1,2.\hat{\varepsilon}_{i},~{}i=1,2. Combining this with (5.24), one has from the boundedness of |x2,ε^||x_{2,\hat{\varepsilon}}| that

|{x||x|R1}m2,ε^Gε^u2,ε^p1𝑑x|CNm2,ε^𝑑xC~,\displaystyle\bigg{|}\int_{\{x||x|\leq R_{1}\}}m_{2,\hat{\varepsilon}}G_{\hat{\varepsilon}}u_{2,\hat{\varepsilon}}^{p-1}\,dx\bigg{|}\leq C\int_{\mathbb{R}^{N}}m_{2,\hat{\varepsilon}}\,dx\leq\tilde{C}, (5.29)

where C~>0\tilde{C}>0 is independent of ε^i,i=1,2.\hat{\varepsilon}_{i},~{}i=1,2. Collecting (5.27), (5.28) and (5.29), one finds (5.26) holds. Moreover, (5.26) indicates

m2,ε^(x)C|x|1p,p>1,\displaystyle m_{2,\hat{\varepsilon}}(x)\leq C|x|^{1-p},~{}\forall p>1,

As a consequence, for any fixed xNx\in\mathbb{R}^{N}, we have

|ε^1x+x1,ε^x2,ε^ε^2|ε^1|x|ε^2+12|x1,ε^x2,ε^|ε^2Cε^2.\displaystyle\bigg{|}\frac{\hat{\varepsilon}_{1}x+x_{1,\hat{\varepsilon}}-x_{2,\hat{\varepsilon}}}{\hat{\varepsilon}_{2}}\bigg{|}\geq\frac{\hat{\varepsilon}_{1}|x|}{\hat{\varepsilon}_{2}}+\frac{1}{2}\frac{|x_{1,\hat{\varepsilon}}-x_{2,\hat{\varepsilon}}|}{\hat{\varepsilon}_{2}}\geq\frac{C}{\hat{\varepsilon}_{2}}.

It follows that

(ε^1ε^2)γ2+N2m2,ε^γ2N+12(ε^1x+x1,ε^x2,ε^ε^2)(ε^1ε^2)γ2+N2ε^2pε^1γ2N2 by choosing p>γ2+N2.\displaystyle\bigg{(}\frac{\hat{\varepsilon}_{1}}{\hat{\varepsilon}_{2}}\bigg{)}^{\frac{\gamma^{\prime}}{2}+\frac{N}{2}}m^{\frac{\gamma^{\prime}}{2N}+\frac{1}{2}}_{2,\hat{\varepsilon}}\bigg{(}\frac{\hat{\varepsilon}_{1}x+x_{1,\hat{\varepsilon}}-x_{2,\hat{\varepsilon}}}{\hat{\varepsilon}_{2}}\bigg{)}\leq\bigg{(}\frac{\hat{\varepsilon}_{1}}{\hat{\varepsilon}_{2}}\bigg{)}^{\frac{\gamma^{\prime}}{2}+\frac{N}{2}}\hat{\varepsilon}_{2}^{p}\leq\hat{\varepsilon}_{1}^{\frac{\gamma^{\prime}}{2}-\frac{N}{2}}\text{ by choosing }p>\frac{\gamma^{\prime}}{2}+\frac{N}{2}. (5.30)

We rewrite the u1,ε^u_{1,\hat{\varepsilon}}-equation in (5.7) as

Δu1,ε^+CH|u1,ε^|γ+λ1,aε1γ\displaystyle-\Delta u_{1,\hat{\varepsilon}}+C_{H}|\nabla u_{1,\hat{\varepsilon}}|^{\gamma}+\lambda_{1,\textbf{a}}\varepsilon_{1}^{\gamma^{\prime}}
=\displaystyle= ε^1V1(ε^1x+x1,ε^)α1m1γNβ(ε^1ε^2)γ2+N2m1,ε^m2,ε^(ε^1x+x1,ε^x2,ε^ε^2):=IVε^.\displaystyle\hat{\varepsilon}_{1}V_{1}(\hat{\varepsilon}_{1}x+x_{1,\hat{\varepsilon}})-\alpha_{1}m_{1}^{\frac{\gamma^{\prime}}{N}}-\beta\bigg{(}\frac{\hat{\varepsilon}_{1}}{\hat{\varepsilon}_{2}}\bigg{)}^{\frac{\gamma^{\prime}}{2}+\frac{N}{2}}m_{1,\hat{\varepsilon}}m_{2,\hat{\varepsilon}}\bigg{(}\frac{\hat{\varepsilon}_{1}x+x_{1,\hat{\varepsilon}}-x_{2,\hat{\varepsilon}}}{\hat{\varepsilon}_{2}}\bigg{)}:=IV_{\hat{\varepsilon}}. (5.31)

Since (5.30) indicates for any R^>0,\hat{R}>0,

|IVε^|CR^,for |x|<R^,|IV_{\hat{\varepsilon}}|\leq C_{\hat{R}},~{}\text{for }|x|<\hat{R},

we have from Lemma 2.1 that

|u1,ε^|CR^, for |x|<R^.\displaystyle|\nabla u_{1,\hat{\varepsilon}}|\leq C_{\hat{R}},\text{ for }|x|<\hat{R}.

Thus, we find by the standard diagonal procedure that

u1,εu1 in Cloc1,α(N) with α(0,1),\displaystyle u_{1,\varepsilon}\rightarrow u_{1}\text{ in }C^{1,\alpha}_{\text{loc}}(\mathbb{R}^{N})\text{ with }\alpha\in(0,1),

then take the limit in (5) to obtain u1u_{1} satisfies

{Δu1+CH|u1|γγN=am1γN,Δm1=CHγ(m1|u1|γ2u1),0<Nm1𝑑x1.\displaystyle\left\{\begin{array}[]{ll}-\Delta u_{1}+C_{H}|\nabla u_{1}|^{\gamma}-\frac{\gamma^{\prime}}{N}=a^{*}m_{1}^{\frac{\gamma^{\prime}}{N}},\\ -\Delta m_{1}=C_{H}\gamma\nabla\cdot(m_{1}|\nabla u_{1}|^{\gamma-2}\nabla u_{1}),\\ 0<\int_{\mathbb{R}^{N}}m_{1}\,dx\leq 1.\end{array}\right.

Moreover, we deduce from Lemma 2.6 and (1.20) that Nm1𝑑x=1.\int_{\mathbb{R}^{N}}m_{1}\,dx=1. It then follows from (5.12) that

m1,ε^m1 in Lp(N),p1,\displaystyle m_{1,\hat{\varepsilon}}\rightarrow m_{1}\text{ in }L^{p}(\mathbb{R}^{N}),~{}\forall p\geq 1,

which finishes the proof of this theorem. ∎

Next, we focus on the refined blow-up rate of minimizers under the case β0\beta\leq 0 and proceed to complete the proof of Theorem 1.7. Before proving Theorem 1.7, we collect the results of the existence of minimizers to (3.1) and the corresponding asymptotic profiles as follows

Lemma 5.1.

Problem (3.1) admits a minimizer (mi,wi)W1,p(N)×Lp(N)(m_{i},w_{i})\in W^{1,p}(\mathbb{R}^{N})\times L^{p}(\mathbb{R}^{N}) with p>1p>1, where wi=CHγmi|ui|γ2uiw_{i}=-C_{H}\gamma m_{i}|\nabla u_{i}|^{\gamma-2}\nabla u_{i} with uiC2(N)u_{i}\in C^{2}(\mathbb{R}^{N}) for i=1,2i=1,2 . Moreover, the following conclusions hold:

  • (i).

    ϵi:=(CLN|wimi|γmi𝑑x)1γ0\epsilon_{i}:=\bigg{(}C_{L}\int_{\mathbb{R}^{N}}\big{|}\frac{w_{i}}{m_{i}}\big{|}^{\gamma^{\prime}}m_{i}\,dx\bigg{)}^{-\frac{1}{\gamma^{\prime}}}\rightarrow 0 as αia\alpha_{i}\nearrow a^{*};

  • (ii).

    Let xi,ϵix_{i,\epsilon_{i}} be a global minimum point of uiu_{i}, then

    ui,ϵ:=ϵi2γγ1ui(ϵix+xi,ϵi),mi,ϵ:=ϵiNmi(ϵix+xi,ϵi),wi,ϵ:=ϵiN+1w(ϵix+xi,ϵi)\displaystyle u_{i,\epsilon}:=\epsilon_{i}^{\frac{2-\gamma}{\gamma-1}}u_{i}(\epsilon_{i}x+x_{i,\epsilon_{i}}),~{}m_{i,\epsilon}:=\epsilon_{i}^{N}m_{i}(\epsilon_{i}x+x_{i,\epsilon_{i}}),~{}w_{i,\epsilon}:=\epsilon_{i}^{N+1}w(\epsilon_{i}x+x_{i,\epsilon_{i}}) (5.32)

    satisfies up to a subsequence,

    ui,ϵiu¯i in Cloc2(N),mi,ϵim¯i in Lp(N),p[1,+],wi,ϵiw¯i in Lγ(N),\displaystyle u_{i,\epsilon_{i}}\rightarrow\bar{u}_{i}\text{ in }C^{2}_{\text{loc}}(\mathbb{R}^{N}),~{}m_{i,\epsilon_{i}}\rightarrow\bar{m}_{i}\text{ in }L^{p}(\mathbb{R}^{N}),\forall p\in[1,+\infty],~{}w_{i,\epsilon_{i}}\rightarrow\bar{w}_{i}\text{ in }L^{\gamma^{\prime}}(\mathbb{R}^{N}), (5.33)

    where (m¯i,w¯i)(\bar{m}_{i},\bar{w}_{i}) is a minimizer of (1.15) and (m¯i,u¯i)(\bar{m}_{i},\bar{u}_{i}) satisfies (1.19);

  • (iii).

    if ViV_{i} satisfies (1.59) and set

    νqi:=infyNN|x+y|qim¯i𝑑x,\displaystyle\nu_{q_{i}}:=\inf_{y\in\mathbb{R}^{N}}\int_{\mathbb{R}^{N}}|x+y|^{q_{i}}\bar{m}_{i}\,dx, (5.34)

    then νqi=ν¯qi\nu_{q_{i}}=\bar{\nu}_{q_{i}} with ν¯qi\bar{\nu}_{q_{i}} given in (1.29) and

    eαii:=(1+o(1))qi+γqi(qiν¯qibiγ)γγ+1(aαia)qiγ+qi,ϵi=(1+o(1))(γ(aαi)abiν¯qiqi)1γ+qi.\displaystyle e^{i}_{\alpha_{i}}:=(1+o(1))\frac{q_{i}+\gamma^{\prime}}{q_{i}}\bigg{(}\frac{q_{i}\bar{\nu}_{q_{i}}b_{i}}{\gamma^{\prime}}\bigg{)}^{\frac{\gamma^{\prime}}{\gamma^{\prime}+1}}\bigg{(}\frac{a^{*}-\alpha_{i}}{a^{*}}\bigg{)}^{\frac{q_{i}}{\gamma^{\prime}+q_{i}}},~{}\epsilon_{i}=(1+o(1))\bigg{(}\frac{\gamma^{\prime}(a^{*}-\alpha_{i})}{a^{*}b_{i}\bar{\nu}_{q_{i}}q_{i}}\bigg{)}^{\frac{1}{\gamma^{\prime}+q_{i}}}. (5.35)

    Moreover, we have

    xi,ϵixiϵiyi,\displaystyle\frac{x_{i,\epsilon_{i}}-x_{i}}{\epsilon_{i}}\rightarrow y_{i},

    where yiNy_{i}\in\mathbb{R}^{N} satisfies

    Hm¯i,qi(yi)=infyNHm¯i,qi(y)=ν¯qi.\displaystyle H_{\bar{m}_{i},q_{i}}(y_{i})=\inf_{y\in\mathbb{R}^{N}}H_{\bar{m}_{i},q_{i}}(y)=\bar{\nu}_{q_{i}}.

    In particular, there exist R>0R>0, C>0C>0 and κ1,\kappa_{1}, δ0>0\delta_{0}>0 small such that

    0<mi,ϵiCeκ12|x|δ0 for |x|>R,\displaystyle 0<m_{i,\epsilon_{i}}\leq Ce^{-\frac{\kappa_{1}}{2}|x|^{\delta_{0}}}\text{ for }|x|>R, (5.36)
Proof.

Proceeding the similar arguments shown in [12], we are able to show Conclusion (i), (ii) and (iii) with ν¯pi\bar{\nu}_{p_{i}} replaced by νpi\nu_{p_{i}}. (5.36) follows directly from Proposition A.1 shown in Appendix A. It is left to show ν¯pi=νpi\bar{\nu}_{p_{i}}=\nu_{p_{i}} with νpi\nu_{p_{i}} defined by (5.34). First of all, it is straightforward to see that ν¯qiνqi\bar{\nu}_{q_{i}}\leq\nu_{q_{i}}. Then, we argue by contradiction and assume

ν¯qi<νqi, for i=1 or 2.\displaystyle\bar{\nu}_{q_{i}}<\nu_{q_{i}},~{}\text{ for }i=1\text{ or }2.

In light of the definition of ν¯qi\bar{\nu}_{q_{i}} given in (1.29), we find that there exists (m,w)(m,w)\in\mathcal{M} with \mathcal{M} defined by (1.30) and yiNy_{i}\in\mathbb{R}^{N} such that

νi0:=infyNN|x+y|qim(x)𝑑x=N|x+yi|qim(x)𝑑x<νqi.\displaystyle\nu_{i0}:=\inf_{y\in\mathbb{R}^{N}}\int_{\mathbb{R}^{N}}|x+y|^{q_{i}}m(x)\,dx=\int_{\mathbb{R}^{N}}|x+y_{i}|^{q_{i}}m(x)\,dx<\nu_{q_{i}}.

Let

mτ:=τNm(τ(xxi)yi),wτ:=τN+1w(τ(xxi)yi),\displaystyle m_{\tau}:=\tau^{N}m(\tau(x-x_{i})-y_{i}),~{}w_{\tau}:=\tau^{N+1}w(\tau(x-x_{i})-y_{i}),

where τ=(biνi0qia2γ(aα1+α2+2β2))1γ+qi\tau=\bigg{(}\frac{b_{i}\nu_{i0}q_{i}a^{*}}{2\gamma^{\prime}\big{(}a^{*}-\frac{\alpha_{1}+\alpha_{2}+2\beta}{2}\big{)}}\bigg{)}^{\frac{1}{\gamma^{\prime}+q_{i}}}. Then one can obtain

eαii(1+o(1))qi+γpi(qiνi0biγ)γγ+1(aαia)qiγ+qi.\displaystyle e^{i}_{\alpha_{i}}\leq(1+o(1))\frac{q_{i}+\gamma^{\prime}}{p_{i}}\bigg{(}\frac{q_{i}\nu_{i0}b_{i}}{\gamma^{\prime}}\bigg{)}^{\frac{\gamma^{\prime}}{\gamma^{\prime}+1}}\bigg{(}\frac{a^{*}-\alpha_{i}}{a^{*}}\bigg{)}^{\frac{q_{i}}{\gamma^{\prime}+q_{i}}}. (5.37)

Whereas, (5.35) gives that

eαii=(1+o(1))qi+γqi(qiνqibiγ)γγ+1(aαia)qiγ+qi,\displaystyle e^{i}_{\alpha_{i}}=(1+o(1))\frac{q_{i}+\gamma^{\prime}}{q_{i}}\bigg{(}\frac{q_{i}\nu_{q_{i}}b_{i}}{\gamma^{\prime}}\bigg{)}^{\frac{\gamma^{\prime}}{\gamma^{\prime}+1}}\bigg{(}\frac{a^{*}-\alpha_{i}}{a^{*}}\bigg{)}^{\frac{q_{i}}{\gamma^{\prime}+q_{i}}},

which reaches a contradiction to (5.37). ∎

Now, we establish the lower and upper bounds of eα1,α2,βe_{\alpha_{1},\alpha_{2},\beta} in the following lemma.

Lemma 5.2.

Assume that each ViV_{i} satisfies (1.59) with x1x2x_{1}\not=x_{2} and (1.60) holds. Let (m1,a,w1,a,m2,a,w2,a)(m_{1,\textbf{a}},w_{1,\textbf{a}},m_{2,\textbf{a}},w_{2,\textbf{a}}) be a minimizer of eα1,α2,βe_{\alpha_{1},\alpha_{2},\beta} defined by (1.10) with β<0.\beta<0. Then for any q>max{q1,q2}q>\max\{q_{1},q_{2}\}, we have there exists Cq>0C_{q}>0 such that

eα11+eα22\displaystyle e^{1}_{\alpha_{1}}+e^{2}_{\alpha_{2}}\leq eα1,α2,β=(m1,a,w1,a,m2,a,w2,a)\displaystyle e_{\alpha_{1},\alpha_{2},\beta}=\mathcal{E}(m_{1,\textbf{a}},w_{1,\textbf{a}},m_{2,\textbf{a}},w_{2,\textbf{a}})
\displaystyle\leq eα11+eα22+Cqϵ~2q, as (α1,α2)(a,a),\displaystyle e^{1}_{\alpha_{1}}+e^{2}_{\alpha_{2}}+C_{q}\tilde{\epsilon}_{2}^{q},\text{ as }(\alpha_{1},\alpha_{2})\nearrow(a^{*},a^{*}), (5.38)

where eαiie_{\alpha_{i}}^{i}, i=1,2i=1,2 are defined by (3.1). In particular, the following estimates hold:

eαiiαii(mi,a.wi,a)eαii+Cqϵ~2q,i=1,2.\displaystyle e^{i}_{\alpha_{i}}\leq\mathcal{E}^{i}_{\alpha_{i}}(m_{i,\textbf{a}}.w_{i,\textbf{a}})\leq e^{i}_{\alpha_{i}}+C_{q}\tilde{\epsilon}^{q}_{2},~{}i=1,2. (5.39)
Proof.

Noting that β<0\beta<0, we deduce from (3.1) and (3.2) that

(m1,a,w1,a,m2,a,w2,a)i=12αii(mi,a,wi,a)eα11+eα22.\displaystyle\mathcal{E}(m_{1,\textbf{a}},w_{1,\textbf{a}},m_{2,\textbf{a}},w_{2,\textbf{a}})\geq\sum_{i=1}^{2}\mathcal{E}_{\alpha_{i}}^{i}(m_{i,\textbf{a}},w_{i,\textbf{a}})\geq e^{1}_{\alpha_{1}}+e^{2}_{\alpha_{2}}. (5.40)

Moreover, let (mi,wi)(m_{i},w_{i}) be the minimizers of eαiie^{i}_{\alpha_{i}}, i=1,2i=1,2 obtained in Lemma 5.1, then one has

eα1,α2,β(m1,w1,m2,w2)=\displaystyle e_{\alpha_{1},\alpha_{2},\beta}\leq\mathcal{E}(m_{1},w_{1},m_{2},w_{2})= i=12αii(mi,wi)2β1+γNNm112+γ2Nm212+γ2N𝑑x\displaystyle\sum_{i=1}^{2}\mathcal{E}_{\alpha_{i}}^{i}(m_{i},w_{i})-\frac{2\beta}{1+\frac{\gamma^{\prime}}{N}}\int_{\mathbb{R}^{N}}m_{1}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}m_{2}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\,dx
=\displaystyle= eα11+eα222β1+γNNm112+γ2Nm212+γ2N𝑑x.\displaystyle e^{1}_{\alpha_{1}}+e^{2}_{\alpha_{2}}-\frac{2\beta}{1+\frac{\gamma^{\prime}}{N}}\int_{\mathbb{R}^{N}}m_{1}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}m_{2}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\,dx. (5.41)

By using (5.32), one finds

Nm112+γ2Nm212+γ2N𝑑x=(ϵ1ϵ2)N(12+γ2N)ϵ1NNm1,ϵ12+γ2N(x)m2,ϵ12+γ2N(ϵ1ϵ2x+x1,ϵ1x2,ϵ2ϵ2)𝑑x.\displaystyle\int_{\mathbb{R}^{N}}m_{1}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}m_{2}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\,dx=(\epsilon_{1}\epsilon_{2})^{-N(\frac{1}{2}+\frac{\gamma^{\prime}}{2N})}\epsilon_{1}^{N}\int_{\mathbb{R}^{N}}m^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}_{1,\epsilon}(x)m^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}_{2,\epsilon}\bigg{(}\frac{\epsilon_{1}}{\epsilon_{2}}x+\frac{x_{1,\epsilon_{1}}-x_{2,\epsilon_{2}}}{\epsilon_{2}}\bigg{)}\,dx. (5.42)

Since xi,ϵixix_{i,\epsilon_{i}}\rightarrow x_{i}, i=1,2i=1,2 and x1x2x_{1}\not=x_{2}, we take R:=14|x1x2|R:=\frac{1}{4}|x_{1}-x_{2}| and obtain for any xBR/ϵ1(0)x\in B_{R/\epsilon_{1}}(0),

|x1,ϵ1x2,ϵ2ϵ2+ϵ1ϵ2x||x1,ϵ1x2,ϵ2|ϵ2ϵ1ϵ2|x|\displaystyle\bigg{|}\frac{x_{1,\epsilon_{1}}-x_{2,\epsilon_{2}}}{\epsilon_{2}}+\frac{\epsilon_{1}}{\epsilon_{2}}x\bigg{|}\geq\frac{|x_{1,\epsilon_{1}}-x_{2,\epsilon_{2}}|}{\epsilon_{2}}-\frac{\epsilon_{1}}{\epsilon_{2}}|x|\geq 34|x1x2|ϵ2Rϵ2\displaystyle\frac{3}{4}\frac{|x_{1}-x_{2}|}{\epsilon_{2}}-\frac{R}{\epsilon_{2}}
=\displaystyle= 12|x1x2|ϵ2=O(1ϵ2)+.\displaystyle\frac{1}{2}\frac{|x_{1}-x_{2}|}{\epsilon_{2}}=O\bigg{(}\frac{1}{\epsilon_{2}}\bigg{)}\rightarrow+\infty.

It then follows from (5.36) that there exists constant C>0C>0 such that

m2,ϵ(ϵ1ϵ2x+x1,ϵ1x2,ϵ2ϵ2)C|x1x2ϵ2|q^,xBR/ϵ1(0),\displaystyle m_{2,\epsilon}\bigg{(}\frac{\epsilon_{1}}{\epsilon_{2}}x+\frac{x_{1,\epsilon_{1}}-x_{2,\epsilon_{2}}}{\epsilon_{2}}\bigg{)}\leq C\bigg{|}\frac{x_{1}-x_{2}}{\epsilon_{2}}\bigg{|}^{-\hat{q}},~{}\forall x\in B_{R/\epsilon_{1}}(0),

which implies

|x|<Rϵ1m1,ϵ12+γ2N(x)m2,ϵ12+γ2N(ϵ1ϵ2x+x1,ϵ1x2,ϵ2ϵ2)𝑑xCϵ2q^(12+γ2N)Nm1,ϵ12+γ2N𝑑xCϵ2q^(12+γ2N),\displaystyle\int_{|x|<\frac{R}{\epsilon_{1}}}m^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}_{1,\epsilon}(x)m_{2,\epsilon}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\bigg{(}\frac{\epsilon_{1}}{\epsilon_{2}}x+\frac{x_{1,\epsilon_{1}}-x_{2,\epsilon_{2}}}{\epsilon_{2}}\bigg{)}\,dx\leq C\epsilon_{2}^{\hat{q}\big{(}\frac{1}{2}+\frac{\gamma^{\prime}}{2N}\big{)}}\int_{\mathbb{R}^{N}}m_{1,\epsilon}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\,dx\leq C\epsilon_{2}^{\hat{q}\big{(}\frac{1}{2}+\frac{\gamma^{\prime}}{2N}\big{)}}, (5.43)

where we have used (5.33). In addition, invoking (5.36), we obtain m1,ϵm_{1,\epsilon} satisfies for any q^>0,\hat{q}>0,

m1,ϵ(x)Cq^|x|q^ for |x|>R,\displaystyle m_{1,\epsilon}(x)\leq C_{\hat{q}}|x|^{-\hat{q}}\text{ for }|x|>{R}, (5.44)

where Cq^>0C_{\hat{q}}>0 is some constant. On the other hand, (5.33) indicates that

lim supϵ20+m2,ϵL<+.\displaystyle\limsup_{\epsilon_{2}\rightarrow 0^{+}}\|m_{2,\epsilon}\|_{L^{\infty}}<+\infty. (5.45)

Combining (5.44) with (5.45), we deduce that

|x|>Rϵ1m1,ϵ12+γ2Nm2,ϵ12+γ2N(ϵ1ϵ2x+x1,ϵ1x2,ϵ2ϵ2)𝑑xC|x|>Rϵ1m1,ϵ112+γ2N𝑑x\displaystyle\int_{|x|>\frac{R}{\epsilon_{1}}}m^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}_{1,\epsilon}m^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}_{2,\epsilon}\bigg{(}\frac{\epsilon_{1}}{\epsilon_{2}}x+\frac{x_{1,\epsilon_{1}}-x_{2,\epsilon_{2}}}{\epsilon_{2}}\bigg{)}\,dx\leq C\int_{|x|>\frac{R}{\epsilon_{1}}}m_{1,\epsilon_{1}}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\,dx
Cq^Rϵ1+rq^(12+γ2N)rN1𝑑rCq^ϵ1q^(12+γ2N)N,\displaystyle\leq C_{\hat{q}}\int_{\frac{R}{\epsilon_{1}}}^{+\infty}r^{-\hat{q}\big{(}\frac{1}{2}+\frac{\gamma^{\prime}}{2N}\big{)}}r^{N-1}\,dr\leq C_{\hat{q}}\epsilon_{1}^{\hat{q}\big{(}\frac{1}{2}+\frac{\gamma^{\prime}}{2N}\big{)}-N}, (5.46)

where we have used (5.44). Upon collecting (5.42), (5.43) and (5), we deduce that

Nm112+γ2Nm212+γ2N𝑑xCq^(ϵ1ϵ2)N(12+γ2N)ϵ1N(ϵ2q^(12+γ2N)+ϵ1q^(12+γ2N)N).\displaystyle\int_{\mathbb{R}^{N}}m_{1}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}m_{2}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\,dx\leq C_{\hat{q}}(\epsilon_{1}\epsilon_{2})^{-N(\frac{1}{2}+\frac{\gamma^{\prime}}{2N})}\epsilon_{1}^{N}\bigg{(}\epsilon_{2}^{\hat{q}(\frac{1}{2}+\frac{\gamma^{\prime}}{2N})}+\epsilon_{1}^{\hat{q}(\frac{1}{2}+\frac{\gamma^{\prime}}{2N})-N}\bigg{)}. (5.47)

Since (1.60) and (5.35) imply up to a subsequence,

limαiaϵiϵ~i=Ci,i=1,2,Ci>0 are constants,\displaystyle\lim_{\alpha_{i}\nearrow a^{*}}\frac{\epsilon_{i}}{\tilde{\epsilon}_{i}}=C_{i},~{}i=1,2,C_{i}>0\text{ are constants},

we have from (5.47) that

Nm112+γ2Nm212+γ2N𝑑x\displaystyle\int_{\mathbb{R}^{N}}m_{1}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}m_{2}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\,dx\leq Cq^ϵ~1N(γ2N12)ϵ~2(q^N)(12+γ2N)+Cq^ϵ~1(q^N)(12+γ2N)ϵ~2N(12+γ2N)\displaystyle C_{\hat{q}}\tilde{\epsilon}_{1}^{-N(\frac{\gamma^{\prime}}{2N}-\frac{1}{2})}\tilde{\epsilon}_{2}^{(\hat{q}-N)\big{(}\frac{1}{2}+\frac{\gamma^{\prime}}{2N}\big{)}}+C_{\hat{q}}\tilde{\epsilon}_{1}^{(\hat{q}-N)\big{(}\frac{1}{2}+\frac{\gamma^{\prime}}{2N}\big{)}}\tilde{\epsilon}_{2}^{-N\big{(}\frac{1}{2}+\frac{\gamma^{\prime}}{2N}\big{)}}
=\displaystyle= Cq^ϵ~2(q^N)(12+γ2N)sN(γ2N12)+ϵ~2s(q^N)(12+γ2N)N(12+γ2N).\displaystyle C_{\hat{q}}\tilde{\epsilon}^{(\hat{q}-N)\big{(}\frac{1}{2}+\frac{\gamma^{\prime}}{2N}\big{)}-sN\big{(}\frac{\gamma^{\prime}}{2N}-\frac{1}{2}\big{)}}_{2}+\tilde{\epsilon}_{2}^{s(\hat{q}-N)\big{(}\frac{1}{2}+\frac{\gamma^{\prime}}{2N}\big{)}-N\big{(}\frac{1}{2}+\frac{\gamma^{\prime}}{2N}\big{)}}. (5.48)

By choosing q^>0\hat{q}>0 large enough, one finds for any q>max{p1,p2}q>\max\{p_{1},p_{2}\},

(q^N)(12+γ2N)sN(γ2N12)>q,s(q^N)(12+γ2N)N(12+γ2N)>q.(\hat{q}-N)\bigg{(}\frac{1}{2}+\frac{\gamma^{\prime}}{2N}\bigg{)}-sN\bigg{(}\frac{\gamma^{\prime}}{2N}-\frac{1}{2}\bigg{)}>q,~{}s(\hat{q}-N)\bigg{(}\frac{1}{2}+\frac{\gamma^{\prime}}{2N}\bigg{)}-N\bigg{(}\frac{1}{2}+\frac{\gamma^{\prime}}{2N}\bigg{)}>q.

Thus, we obtain from (5) that

int\displaystyle\int_{\mathbb{R}^{N}}m_{1}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}m_{2}^{\frac{1}{2}+\frac{\gamma^{\prime}}{2N}}\,dx\leq C_{q}\tilde{\epsilon}_{2}^{q},~{}\forall q>\max\{q_{1},q_{2}\}, (5.49)

where Cq>0C_{q}>0 is a constant. Finally, (5.49) together with (5.40) and (5) implies (5.2).

We next show estimate (5.39). First of all, it is straightforward to obtain from the definitions of eiαie^{i}_{\alpha_{i}} that

eiαiiαi(mi,a,wi,a),i=1,2.\displaystyle e^{i}_{\alpha_{i}}\leq\mathcal{E}^{i}_{\alpha_{i}}(m_{i,\textbf{a}},w_{i,\textbf{a}}),~{}i=1,2. (5.50)

Then, we argue by contradiction to establish the estimate shown in the right hand side of (5.39). Without loss of generality, we assume for i=1,i=1,

1α1(m1,a,w1,a)e1αi+Γϵ~2q,\displaystyle\mathcal{E}^{1}_{\alpha_{1}}(m_{1,\textbf{a}},w_{1,\textbf{a}})\geq e^{1}_{\alpha_{i}}+\Gamma\tilde{\epsilon}_{2}^{q},

where Γ>0\Gamma>0 is large enough and note that ϵ~2qmin{e1α1,e2α2}\tilde{\epsilon}_{2}^{q}\ll\min\{e^{1}_{\alpha_{1}},e^{2}_{\alpha_{2}}\} thanks to (5.35). Whereas, by using (5.50), one has

e1α1+e2α2+Γϵ~2qi=12αi(mi,a,wi,a)eα1,α2,β,\displaystyle e^{1}_{\alpha_{1}}+e^{2}_{\alpha_{2}}+\Gamma\tilde{\epsilon}_{2}^{q}\leq\sum_{i=1}^{2}\mathcal{E}_{\alpha_{i}}(m_{i,\textbf{a}},w_{i,\textbf{a}})\leq e_{\alpha_{1},\alpha_{2},\beta},

which is contradicted to (5.2). Therefore, we find (5.39) holds for i=1i=1. Proceeding the similar argument, we can show (5.39) holds for i=2.i=2.

Remark 5.1.

We remark that by using the exponential decay properties of mm shown in Proposition A.1, the conclusion in Lemma 5.2 holds when (1.60) is replaced by the following condition

lim𝐚𝐚eϵ~1δ^ϵ~2q2=0,\displaystyle\lim_{\bf{a}\nearrow\bf{a}^{*}}\frac{e^{{-\tilde{\epsilon}_{1}}^{-\hat{\delta}}}}{\tilde{\epsilon}_{2}^{q_{2}}}=0, (5.51)

where q2q_{2} is given in (1.59) and constant δ^>0\hat{\delta}>0 depends on δ0\delta_{0} and κ1\kappa_{1}, which are defined in Proposition A.1. Moreover, with the aid of (5.51), one can show all conclusions of Theorem 1.7. In other words, assumption (1.60) can be relaxed as (5.51) if the exponential decay properties of m1m_{1} and m2m_{2} are established.

Now, we are ready to prove Theorem 1.7, which is

Proof of Theorem 1.7:

Proof.

First of all, we have the fact that

iαi(mi,a,wi,a)ε^iγ(1αia)+NVi(ε^ix+xi,ε^)mi,ε^dx.\displaystyle\mathcal{E}^{i}_{\alpha_{i}}(m_{i,\textbf{a}},w_{i,\textbf{a}})\geq\hat{\varepsilon}_{i}^{\gamma^{\prime}}\bigg{(}1-\frac{\alpha_{i}}{a^{*}}\bigg{)}+\int_{\mathbb{R}^{N}}V_{i}(\hat{\varepsilon}_{i}x+x_{i,\hat{\varepsilon}})m_{i,\hat{\varepsilon}}\,dx.

We compute to get

NVi(ε^ix+xi,ε^)mi,ε^dx\displaystyle\int_{\mathbb{R}^{N}}V_{i}(\hat{\varepsilon}_{i}x+x_{i,\hat{\varepsilon}})m_{i,\hat{\varepsilon}}\,dx =ε^iqiNVi(ε^ix+xi,ε^)|ε^ix+xi,ε^xi|pi|x+xi,ε^xiε^i|mi,ε^dx=ε^qiiIε^.\displaystyle=\hat{\varepsilon}_{i}^{q_{i}}\int_{\mathbb{R}^{N}}\frac{V_{i}(\hat{\varepsilon}_{i}x+x_{i,\hat{\varepsilon}})}{|\hat{\varepsilon}_{i}x+x_{i,\hat{\varepsilon}}-x_{i}|^{p_{i}}}\bigg{|}x+\frac{x_{i,\hat{\varepsilon}}-x_{i}}{\hat{\varepsilon}_{i}}\bigg{|}m_{i,\hat{\varepsilon}}\,dx=\hat{\varepsilon}^{q_{i}}_{i}I_{\hat{\varepsilon}}. (5.52)

By using (3.1), (5.39) and (5.35), we proceed the similar argument shown in Theorem 1.5, then obtain up to a subsequence,

xi,ε^xiε^iyi0 for some yi0N.\displaystyle\frac{x_{i,\hat{\varepsilon}}-x_{i}}{\hat{\varepsilon}_{i}}\rightarrow y_{i0}\text{ for some }y_{i0}\in\mathbb{R}^{N}.

Hence, one has Iε^I_{\hat{\varepsilon}} defined in (5.52) satisfies

limε^0Iε^biN|x+yi0|qimi(x)dxν¯qibi,\displaystyle\lim_{\hat{\varepsilon}\rightarrow 0}I_{\hat{\varepsilon}}\geq b_{i}\int_{\mathbb{R}^{N}}|x+y_{i0}|^{q_{i}}m_{i}(x)\,dx\geq\bar{\nu}_{q_{i}}b_{i},

where the “=” in the second inequality holds if and only if Hm¯i,qi(yi0)=infyNHm¯i,qi(y)=ν¯qi.H_{\bar{m}_{i},q_{i}}(y_{i0})=\inf_{y\in\mathbb{R}^{N}}H_{\bar{m}_{i},q_{i}}(y)=\bar{\nu}_{q_{i}}. It then follows that

iαi(mi,a,wi,a)\displaystyle\mathcal{E}^{i}_{\alpha_{i}}(m_{i,\textbf{a}},w_{i,\textbf{a}})\geq ε^iγaαia+ε^iqibiν¯qi(1+o(1))\displaystyle\hat{\varepsilon}_{i}^{\gamma^{\prime}}\frac{a^{*}-\alpha_{i}}{a^{*}}+\hat{\varepsilon}_{i}^{q_{i}}b_{i}\bar{\nu}_{q_{i}}(1+o(1))
\displaystyle\geq (1+o(1))qi+γpi(qiν¯qibiγ)γγ+1(aαia)qiγ+qi,\displaystyle(1+o(1))\frac{q_{i}+\gamma^{\prime}}{p_{i}}\bigg{(}\frac{q_{i}\bar{\nu}_{q_{i}}b_{i}}{\gamma^{\prime}}\bigg{)}^{\frac{\gamma^{\prime}}{\gamma^{\prime}+1}}\bigg{(}\frac{a^{*}-\alpha_{i}}{a^{*}}\bigg{)}^{\frac{q_{i}}{\gamma^{\prime}+q_{i}}},

and the equality holds if and only if

ε^iγ=(1+o(1))(γ(aαi)abiν¯qiqi)1γ+qi.\displaystyle\hat{\varepsilon}_{i}^{\gamma^{\prime}}=(1+o(1))\bigg{(}\frac{\gamma^{\prime}(a^{*}-\alpha_{i})}{a^{*}b_{i}\bar{\nu}_{q_{i}}q_{i}}\bigg{)}^{\frac{1}{\gamma^{\prime}+q_{i}}}.

Comparing the lower bound and the upper bound of αii(mi,a,wi,a)\mathcal{E}_{\alpha_{i}}^{i}(m_{i,\textbf{a}},w_{i,\textbf{a}}) with i=1,2i=1,2 shown in Lemma 5.2, we see that (1.61) and (1.61) hold. The proof of this theorem is finished. ∎

Theorem 1.7 exhibits the refined blow-up profiles of ground states when interaction coefficient β<0\beta<0 under some technical assumptions (1.59) and (1.60). It is worthy mentioning that with the aid of Proposition A.1, we are able to improve the condition (1.60) such that the conclusion shown in Theorem 1.7 still holds.

6 Conclusions

In this paper, we have studied the stationary multi-population Mean-field Games system (1.7) with decreasing cost self-couplings and interactive couplings under critical mass exponents via variational methods. Concerning the existence of ground states, we classified the existence of minimizers to constraint minimization problem (1.10) in terms of self-focusing coefficients and interaction coefficients, in which the attractive and repulsive interactions were discussed, respectively. In particular, when all coefficients are subcritical, we showed the existence of ground states to (1.7) by the duality argument. Then, the basic and refined blow-up profiles of ground states were studied under some mild assumptions of potential functions Vi,V_{i}, i=1,2.i=1,2.

We would like to mention that there are also some open problems deserve explorations in the future. In this paper, we focus on the existence and asymptotic profiles of ground states to (1.7) with mass critical local couplings under the case of γ<N\gamma<N^{\prime} with γ\gamma given in (1.8) since population density mm can be shown in some Hölder space by using Morrey’s estimate and system (1.7) enjoys the better regularity. Whereas, if γN\gamma\geq N^{\prime}, nonlinear terms (1.9) in (1.7) become singular and one can only show mLp(N)m\in L^{p}(\mathbb{R}^{N}) for some p>1p>1 by standard Sobolev embedding. Correspondingly, the positivities of m1m_{1} and m2m_{2} given in (1.7) can not be shown due to the worse regularities. Hence, when γN\gamma\geq N^{\prime}, it seems a challenge but interesting to prove the existence of ground states even under the mass subcritical local couplings. On the other hand, while discussing the concentration phenomena in (1.7), we impose some assumptions on potential functions ViV_{i}, i=1,2.i=1,2. In detail, when the interaction coefficient β\beta satisfies β>0\beta>0, (1.27) is assumed for the convenience of analysis. However, when V1V_{1} and V2V_{2} satisfy infxN(V1(x)+V2(x))>0\inf_{x\in\mathbb{R}^{N}}(V_{1}(x)+V_{2}(x))>0, the classification of the existence of minimizers is more intriguing and the corresponding blow-up profiles analysis might be more complicated. Similarly, if the interaction is repulsive, the investigation of the concentration property of global minimizers is also challenging when V1V_{1} and V2V_{2} have common global minima.

Acknowledgments

We thank Professor M. Cirant for stimulating discussions and many insightful suggestions. Xiaoyu Zeng is supported by NSFC (Grant Nos. 12322106, 11931012, 12171379, 12271417).

Appendix A Exponential Decay Estimates of Population Densities

In this appendix, we investigate the exponential decay property of population density mm. More precisely, we consider the following system:

{Δuε+CH|uε|γ+λε=εγV(εx+xε)+gε(x),xN,Δmε+CHγ(mε|uε|γ2uε)=0,xN,\displaystyle\left\{\begin{array}[]{ll}-\Delta u_{\varepsilon}+C_{H}|\nabla u_{\varepsilon}|^{\gamma}+\lambda_{\varepsilon}=\varepsilon^{\gamma}V(\varepsilon x+x_{\varepsilon})+g_{\varepsilon}(x),&x\in\mathbb{R}^{N},\\ -\Delta m_{\varepsilon}+C_{H}\gamma\nabla\cdot(m_{\varepsilon}|\nabla u_{\varepsilon}|^{\gamma-2}\nabla u_{\varepsilon})=0,&x\in\mathbb{R}^{N},\end{array}\right.

where γ>1\gamma>1, VV and gεg_{\varepsilon} are given. Under some assumptions of gεg_{\varepsilon} and λε\lambda_{\varepsilon}, one can show mεm_{\varepsilon} satisfies the exponential decay property, which is

Proposition A.1.

Denote (mε,uε,λε)W1,p(N)×C2(N)×(m_{\varepsilon},u_{\varepsilon},\lambda_{\varepsilon})\in W^{1,p}(\mathbb{R}^{N})\times C^{2}(\mathbb{R}^{N})\times\mathbb{R} as the solution to

{Δuε+CH|uε|γ+λε=εγV(εx+xε)+gε(x),xN,Δmε+CHγ(mε|uε|γ2uε)=0,xN,\displaystyle\left\{\begin{array}[]{ll}-\Delta u_{\varepsilon}+C_{H}|\nabla u_{\varepsilon}|^{\gamma}+\lambda_{\varepsilon}=\varepsilon^{\gamma}V(\varepsilon x+x_{\varepsilon})+g_{\varepsilon}(x),&x\in\mathbb{R}^{N},\\ -\Delta m_{\varepsilon}+C_{H}\gamma\nabla\cdot(m_{\varepsilon}|\nabla u_{\varepsilon}|^{\gamma-2}\nabla u_{\varepsilon})=0,&x\in\mathbb{R}^{N},\end{array}\right.

where mε0m_{\varepsilon}\geq 0 in N\mathbb{R}^{N}, uεu_{\varepsilon} is uniformly bounded from below and Hölder continuous function VV satisfies (1.13) and (1.14), and gεC0,θ(N)g_{\varepsilon}\in C^{0,\theta}(\mathbb{R}^{N}) with θ(0,1)\theta\in(0,1) independent of ε\varepsilon. Suppose that

  • (i).

    λελ0\lambda_{\varepsilon}\rightarrow\lambda_{0} up to a subsequence with λ0<0\lambda_{0}<0;

  • (ii).

    gε(x)0g_{\varepsilon}(x)\rightarrow 0 uniformly as |x|+|x|\rightarrow+\infty,

then we have there exist constants C>0C>0 and R>0R>0 independent of ε\varepsilon such that

0<mεCeκ12|x|δ0 when |x|>R,\displaystyle 0<m_{\varepsilon}\leq Ce^{-\frac{\kappa_{1}}{2}|x|^{\delta_{0}}}\text{ when }|x|>R, (A.1)

where constant δ0(0,min{γ1,1})\delta_{0}\in(0,\min\{\gamma-1,1\}), constant κ1>0\kappa_{1}>0 and they are independent of ε.\varepsilon.

Proof.

By following the same argument shown in the proof of Theorem 1.6, one has

uε(x)C|x| for |x|>R^,\displaystyle u_{\varepsilon}(x)\geq C|x|\text{ for }|x|>\hat{R}, (A.2)

where R^>0\hat{R}>0 is some constant. Then we define the Lyapunov function Φ=eκuεδ0\Phi=e^{\kappa u_{\varepsilon}^{\delta_{0}}} with 0<κ<10<\kappa<1 and 0<δ0<10<\delta_{0}<1 will be determined later. We compute to get

ΔΦ+CHγ|uε|γ2uεΦ\displaystyle-\Delta\Phi+C_{H}\gamma|\nabla u_{\varepsilon}|^{\gamma-2}\nabla u_{\varepsilon}\cdot\nabla\Phi
=\displaystyle= κδ0Φuεδ01[Δuε(κδ0uεδ01+(δ01)uε1)|uε|2+CHγ|uε|γ]\displaystyle\kappa\delta_{0}\Phi u_{\varepsilon}^{\delta_{0}-1}[-\Delta u_{\varepsilon}-(\kappa\delta_{0}u_{\varepsilon}^{\delta_{0}-1}+(\delta_{0}-1)u_{\varepsilon}^{-1})|\nabla u_{\varepsilon}|^{2}+C_{H}\gamma|\nabla u_{\varepsilon}|^{\gamma}]
=\displaystyle= κδ0Φuεδ01[CH(γ1)|uε|γλε+εγV(εx+xε)+gε(x)(κδ0uεδ01+(δ01)uε1)|uε|2].\displaystyle\kappa\delta_{0}\Phi u_{\varepsilon}^{\delta_{0}-1}[C_{H}(\gamma-1)|\nabla u_{\varepsilon}|^{\gamma}-\lambda_{\varepsilon}+\varepsilon^{\gamma}V(\varepsilon x+x_{\varepsilon})+g_{\varepsilon}(x)-(\kappa\delta_{0}u_{\varepsilon}^{\delta_{0}-1}+(\delta_{0}-1)u_{\varepsilon}^{-1})|\nabla u_{\varepsilon}|^{2}].

Without loss of generality, we assume uε1u_{\varepsilon}\geq 1 by fixing uε(0)u_{\varepsilon}(0). Then it is straightforward to show that

(κδ0uεδ01+(δ01)uε1)|uε|22κδ0uεδ01|uε|2,|x|>R,\displaystyle(\kappa\delta_{0}u_{\varepsilon}^{\delta_{0}-1}+(\delta_{0}-1)u_{\varepsilon}^{-1})|\nabla u_{\varepsilon}|^{2}\leq 2\kappa\delta_{0}u_{\varepsilon}^{\delta_{0}-1}|\nabla u_{\varepsilon}|^{2},~{}|x|>R,

where R>0R>0 is a large constant and we have used uα1u1u^{\alpha-1}\geq u^{-1}. In addition, by using Lemma 2.1 and Lemma 2.2, we have the facts that

|uε|2γC(1+εγV)2γγ, and uε1δ0C(1+εγV)1δ0γ for |x|>R,\displaystyle|\nabla u_{\varepsilon}|^{2-\gamma}\leq C(1+\varepsilon^{\gamma}V)^{\frac{2-\gamma}{\gamma}},\text{ and }u_{\varepsilon}^{1-\delta_{0}}\geq C(1+\varepsilon^{\gamma}V)^{\frac{1-\delta_{0}}{\gamma}}\text{ for }|x|>R, (A.3)

where C>0C>0 is a constant and 0<δ0<10<\delta_{0}<1.

Next, we would like to prove there exists R>0R>0 independent of ε\varepsilon such that

CH(γ1)2|uε|γ2κδ0uεδ01|uε|2,|x|>R.\displaystyle\frac{C_{H}(\gamma-1)}{2}|\nabla u_{\varepsilon}|^{\gamma}\geq 2\kappa\delta_{0}u_{\varepsilon}^{\delta_{0}-1}|\nabla u_{\varepsilon}|^{2},~{}\forall|x|>R. (A.4)

Actually, when γ2\gamma\geq 2, it is easy to show (A.4) holds by (A.3) and choosing κ\kappa small enough. When 1<γ<21<\gamma<2, by taking δ0\delta_{0} and κ\kappa such that 2γ1δ02-\gamma\leq 1-\delta_{0} and κ\kappa small, one finds (A.4) holds. In summary, upon choosing δ0(0,min{γ1,1})\delta_{0}\in(0,\min\{\gamma-1,1\}) and κ\kappa small enough, we apply Condition (i) and (ii) to get

ΔΦ+CHγ|uε|γ2uεΦCκδ0uεδ01Φ, if |x|>R.\displaystyle-\Delta\Phi+C_{H}\gamma|\nabla u_{\varepsilon}|^{\gamma-2}\nabla u_{\varepsilon}\cdot\nabla\Phi\geq C\kappa\delta_{0}u_{\varepsilon}^{\delta_{0}-1}\Phi,\text{ if }|x|>R.

Proceeding the similar argument shown in the proof of (5.26), one finds

supεNeκuεδ0uεδ01mεdx<+.\displaystyle\sup_{\varepsilon}\int_{\mathbb{R}^{N}}e^{\kappa u_{\varepsilon}^{\delta_{0}}}u_{\varepsilon}^{\delta_{0}-1}m_{\varepsilon}\,dx<+\infty.

Therefore, by using the uniformly Hölder continuity of mεm_{\varepsilon} and the fact that uε1u_{\varepsilon}\geq 1, we obtain for |x|>R|x|>R with constant R>0R>0 independent of ε,\varepsilon,

0<mε(x)Ceκ2uεδ0,δ0(0,γ1),\displaystyle 0<m_{\varepsilon}(x)\leq Ce^{-\frac{\kappa}{2}u_{\varepsilon}^{\delta_{0}}},~{}\delta_{0}\in(0,\gamma-1), (A.5)

where C>0C>0, κ>0\kappa>0 is small and δ0(0,min{γ1,1})\delta_{0}\in(0,\min\{\gamma-1,1\}), which are all independent of ε\varepsilon. Moreover, in light of (A.2), one has from (A.5) that (A.1) holds. ∎

References

  • Achdou et al. [2020] Y. Achdou, P. Cardaliaguet, F. Delarue, A. Porretta, and F. Santambrogio. Mean field games, volume 2281 of Lecture Notes in Math. Springer, Cham; Centro Internazionale Matematico Estivo (C.I.M.E.), Florence, 2020.
  • Bensoussan et al. [2013] A. Bensoussan, J. Frehse, and P. Yam. Mean field games and mean field type control theory, volume 101. Springer, 2013.
  • Bernardini and Cesaroni [2023] C. Bernardini and A. Cesaroni. Ergodic mean-field games with aggregation of choquard-type. J. Differential Equations, 364:296–335, 2023.
  • Cardaliaguet [2010] P. Cardaliaguet. Notes on mean field games. Technical report, Paris Dauphine University, 2010.
  • Cardaliaguet and Rainer [2020] P. Cardaliaguet and C. Rainer. An example of multiple mean field limits in ergodic differential games. NoDEA Nonlinear Differential Equations Appl., 27(3):Paper No. 25, 19, 2020.
  • Cardaliaguet and Souganidis [2022] P. Cardaliaguet and P. Souganidis. On first order mean field game systems with a common noise. Ann. Appl. Probab., 32(3):2289–2326, 2022.
  • Cardaliaguet et al. [2017] P. Cardaliaguet, A. Porretta, and D. Tonon. A segregation problem in multi-population mean field games. In Ann. Internat. Soc. Dynam. Games, volume 15, pages 49–70. Birkhäuser/Springer, Cham, 2017.
  • Cesaroni and Cirant [2018] A. Cesaroni and M. Cirant. Concentration of ground states in stationary mean-field games systems. Anal. PDE, 12(3):737–787, 2018.
  • Cirant [2015] M. Cirant. Multi-population mean field games systems with neumann boundary conditions. J. Math. Pures Appl., 103(5):1294–1315, 2015.
  • Cirant [2016] M. Cirant. Stationary focusing mean-field games. Comm. Partial Differential Equations, 41(8):1324–1346, 2016.
  • Cirant et al. [2024a] M. Cirant, A. Cosenza, and G. Verzini. Ergodic mean field games: existence of local minimizers up to the sobolev critical case. Calc. Var. Partial Differential Equations, 63(5):134, 2024a.
  • Cirant et al. [2024b] M. Cirant, F. Kong, J. Wei, and X. Zeng. Critical mass phenomena and blow-up behavior of ground states in stationary second order mean-field games systems with decreasing cost. preprint, 2024b.
  • Feleqi [2013] E. Feleqi. The derivation of ergodic mean field game equations for several populations of players. Dyn. Games Appl., 3:523–536, 2013.
  • Guéant [2009] O. Guéant. Mean field games and applications to economics. PhD Thesis (Univ. Paris-Dauphine), 2009.
  • Guéant et al. [2011] O. Guéant, J. Lasry, and P. Lions. Mean field games and applications. In Paris-Princeton Lectures on Mathematical Finance 2010, volume 2003 of Lecture Notes in Math., pages 205–266. Springer, Berlin, 2011.
  • Huang [2010] M. Huang. Large-population LQG games involving a major player: the nash certainty equivalence principle. SIAM J. Control Optim., 48(5):3318–3353, 2010.
  • Huang et al. [2006] M. Huang, R. Malhamé, and P. Caines. Large population stochastic dynamic games: closed-loop McKean-Vlasov systems and the Nash certainty equivalence principle. Commun. Inf. Syst., 6(3):221–251, 2006.
  • Lasry and Lions [2007] J. Lasry and P. Lions. Mean field games. Japan. J. Math., 2(1):229–260, 2007.
  • Lions and Souganidis [2020] P. Lions and P. Souganidis. Extended mean-field games. Rend. Lincei, 31(3):611–625, 2020.
  • Metafune et al. [2005] G. Metafune, D. Pallara, and A. Rhandi. Global properties of invariant measures. J. Funct. Anal., 223(2):396–424, 2005.
  • Nourian and Caines [2013] M. Nourian and P. Caines. ϵ\epsilon-nash mean field game theory for nonlinear stochastic dynamical systems with major and minor agents. SIAM J. Control Optim., 51(4):3302–3331, 2013.