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

\UseRawInputEncoding

Blow-up Behaviors of Ground States in Ergodic Mean-field Games Systems with Hartree-type Coupling

Fanze Kong , Yonghui Tong, Xiaoyu Zeng and Huan-Song Zhou Department of Applied Mathematics, University of Washington, Seattle, WA 98195, USA; [email protected] for Mathematical Sciences, Wuhan University of Technology, Wuhan 430070, China; [email protected] for Mathematical Sciences, Wuhan University of Technology, Wuhan 430070, China; [email protected] for Mathematical Sciences, Wuhan University of Technology, Wuhan 430070, China; [email protected]
Abstract

In this paper, we investigate the concentration behaviors of ground states to stationary Mean-field Games systems (MFGs) with the nonlocal coupling in n\mathbb{R}^{n}, n2.n\geq 2. With the mass critical exponent imposed on Riesz potentials, we first discuss the existence of ground states to potential-free MFGs, which corresponds to the establishment of Gagliardo-Nirenberg type’s inequality. Next, with the aid of the optimal inequality, we classify the existence of ground states to stationary MFGs with Hartree-type coupling in terms of the L1L^{1}-norm of population density defined by MM. In addition, under certain types of coercive potentials, the asymptotics of ground states to ergodic MFGs with the nonlocal coupling are captured. Moreover, if the local polynomial expansions are imposed on potentials, we study the refined asymptotic behaviors of ground states and show that they concentrate on the flattest minima of potentials.

MSC: 35J47, 35J50, 46N10

Keywords: Mean-field Games, Variational Method, Nonlocal Coupling, Ground States, Blow-up Profiles

1 Introduction

In this paper, we are concerned with the following ergodic stationary Mean-field Games systems

{Δu+H(u)+λ=V(x)Kαm,xn,Δm+(mH(u))=0,xn,nm𝑑x=M>0,\displaystyle\left\{\begin{array}[]{ll}-\Delta u+H(\nabla u)+\lambda=V(x)-K_{\alpha}\ast m,&x\in\mathbb{R}^{n},\\ \Delta m+\nabla\cdot(m\nabla H(\nabla u))=0,&x\in\mathbb{R}^{n},\\ \int_{\mathbb{R}^{n}}m\,dx=M>0,\end{array}\right. (1.4)

where (m,u,λ)(m,u,\lambda) denotes a solution, λ\lambda is a so-called Lagrange multiplier, VV is the potential function and KαK_{\alpha} is defined as the Riesz potential satisfying

K=1|x|nα with 0<α<n.\displaystyle K=\frac{1}{|x|^{n-\alpha}}\text{ with }0<\alpha<n. (1.5)

Here mm represents the population density and uu is the value function of a typical player. In particular, Hamiltonian H:nH:\mathbb{R}^{n}\rightarrow\mathbb{R} is in general assumed to be convex uniformly and the typical form is

H(𝒑):=CH|𝒑|γ,γ>1,CH>0.\displaystyle H(\boldsymbol{p}):={C_{H}}|\boldsymbol{p}|^{\gamma},~{}~{}\exists\gamma>1,~{}C_{H}>0. (1.6)

Correspondingly, the Lagrangian is defined by L(𝒒):=sup𝒑n[𝒒𝒑H(𝒑)]L(\boldsymbol{q}):=\sup_{\boldsymbol{p}\in\mathbb{R}^{n}}[\boldsymbol{q}\cdot\boldsymbol{p}-H(\boldsymbol{p})] and if HH is given by (1.6), LL can be written as

L(𝒒)=CL|𝒒|γ,γ=γγ1>1andCL=1γ(γCH)11γ>0,\displaystyle L(\boldsymbol{q})={C_{L}}|\boldsymbol{q}|^{\gamma^{\prime}},~{}~{}\gamma^{\prime}=\frac{\gamma}{\gamma-1}>1~{}\text{and}~{}C_{L}=\frac{1}{\gamma^{\prime}}(\gamma C_{H})^{\frac{1}{1-\gamma}}>0, (1.7)

where γ\gamma^{\prime} is the conjugate number of γ.\gamma.

Assume HH in system (1.4) is given by (1.6) and V(x)V(x) has polynomial lower and upper bounds when |x||x| is large enough, then Cesaroni and Bernardini [5, 4] studied the existence and concentration of ground states to (1.4) under the subcritical mass exponent case by using the variational method. Motivated by their results and our analysis focused on Mean-field Games systems with the local coupling [14], we shall utilize the variational approach to discuss the existence and asymptotic behaviors of ground states to (1.4) under the critical mass exponent case, i.e. α=α:=nγ\alpha=\alpha^{*}:=n-\gamma^{\prime} in (1.5).

1.1 Mean-field Games Theory and Systems

Motivated by the theories of statistical physics, Huang et al. [20] and Lasry et al. [21] in 2007 developed Mean-field Games theories and proposed a class of coupled PDE systems to describe the differential games among a huge number of players, which have rich applications in the fields of economics, finance and management.

The general form of time-dependent Mean-field Games systems reads as

{ut=Δu+H(u)V(x)f(m),xn,t>0,mt=Δm+(H(u)m),xn,t>0,u|t=T=uT,m|t=0=m0,xn,\left\{\begin{array}[]{ll}u_{t}=-\Delta u+H(\nabla u)-V(x)-f(m),&x\in\mathbb{R}^{n},t>0,\\ m_{t}=\Delta m+\nabla\cdot(\nabla H(\nabla u)m),&x\in\mathbb{R}^{n},t>0,\\ u|_{t=T}=u_{T},m|_{t=0}=m_{0},&x\in\mathbb{R}^{n},\end{array}\right. (1.8)

where mm and uu denote the density and the value function, respectively. Here m0m_{0} represents the initial data of density and uTu_{T} is the terminal data of the value function. Now, we give a brief summary of the derivation of (1.8). Suppose the dynamics of the ii-th player satisfies

dXti=νtidt+2dBti,X0i=xin,i=1,,N,\displaystyle dX_{t}^{i}=-\nu^{i}_{t}dt+\sqrt{2}dB_{t}^{i},\ \ X_{0}^{i}=x^{i}\in\mathbb{R}^{n},~{}i=1,\cdots,N, (1.9)

where xix^{i} is the initial condition, νti\nu^{i}_{t} is the velocity and BtiB_{t}^{i} represents the Brownian motion. Assume BtiB_{t}^{i} for i=1,,Ni=1,\cdots,N are independent and all players are homogeneous, then we have XtiX_{t}^{i} for i=1,,Ni=1,\cdots,N follow the same process and drop “ii" in (1.9). On the other hand, each player aims to minimize the following expected 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}), (1.10)

where LL is the Lagrangian, VV measures the spatial preference and ff is the coupling. Invoking the dynamic programming principle [2, 3], one can formulate the time-dependent system (1.8) by analyzing the minimization of (1.10). We point out that many results are concentrated on the study of global well-posedness to (1.8), see [9, 8, 7, 13, 18, 17, 16].

As stated in [14], the corresponding stationary problem of (1.8) is

{Δu+H(u)+λ=f(m)+V(x),xn,Δm+(mH(u))=0,xn,Nm𝑑x=M>0,\left\{\begin{array}[]{ll}-\Delta u+H(\nabla u)+\lambda=f(m)+V(x),&x\in\mathbb{R}^{n},\\ \Delta m+\nabla\cdot(m\nabla H(\nabla u))=0,&x\in\mathbb{R}^{n},\\ \int_{\mathbb{R}^{N}}mdx=M>0,\end{array}\right. (1.11)

where the triple (m,u,λ)(m,u,\lambda) denotes the solution, VV is the potential function and ff is the cost function. There are also some results concerning the existence and qualitative properties of non-trivial solutions to the stationary problem (1.11), see [10, 15, 19, 23, 12, 14, 5, 4]. We mention that when the cost ff is monotone increasing, as shown in [21], the uniqueness of the solution to (1.11) can be in general guaranteed. Whereas, when the cost ff is monotone decreasing and unbounded, the case is delicate and (1.11) may admit many distinct solutions. In particular, the pioneering work in the study of ground states to stationary Mean-field Games systems with decreasing cost was finished by Cesaroni and Cirant [10].

We also would like to point out the stationary Mean-field Games systems can be trivialized to nonlinear γ\gamma^{\prime}-Laplacian Schrödinger equations when HH is chosen as (1.6). Indeed, Fokker-Planck equation in (1.11) can be reduced into the following form:

m+mCH|u|γ2u=0a.e.,xn.\displaystyle\nabla m+mC_{H}|\nabla u|^{\gamma-2}\nabla u=0~{}~{}\text{a.e.,}~{}~{}x\in\mathbb{R}^{n}. (1.12)

Similarly as shown [11], we define v:=m1γv:=m^{\frac{1}{\gamma^{\prime}}} and obtain from (1.12) and the uu-equation in (1.11) that

{μΔγv+[f(vγ)+V(x)λ]vγ1=0,xn,nvγ𝑑x=M,v>0,μ=(γCH)γ1,\displaystyle\left\{\begin{array}[]{ll}-\mu\Delta_{\gamma^{\prime}}v+[f(v^{\gamma^{\prime}})+V(x)-\lambda]v^{\gamma^{\prime}-1}=0,~{}x\in\mathbb{R}^{n},\\ \int_{\mathbb{R}^{n}}v^{\gamma^{\prime}}\,dx=M,~{}v>0,~{}\mu=\big{(}\frac{\gamma^{\prime}}{C_{H}}\big{)}^{\gamma^{\prime}-1},\end{array}\right. (1.15)

where Δγ\Delta_{\gamma^{\prime}} is the γ\gamma^{\prime}-Laplacian and given by Δγv=(|v|γ2v)\Delta_{\gamma^{\prime}}v=\nabla\cdot(|\nabla v|^{\gamma^{\prime}-2}\nabla v). It is well-known that nonlinear γ\gamma^{\prime}-Laplacian Schrödinger equation (1.15) admits the following variational structures:

(v):=n[μγ|v|γ+F(v)+1γV(x)vγ]𝑑x,\displaystyle\mathcal{F}(v):=\int_{\mathbb{R}^{n}}\bigg{[}\frac{\mu}{\gamma^{\prime}}|\nabla v|^{\gamma^{\prime}}+F(v)+\frac{1}{\gamma^{\prime}}V(x)v^{\gamma^{\prime}}\bigg{]}\,dx, (1.16)

where F(v)F(v) denotes the anti-derivative of f(vγ)vγ1.f(v^{\gamma^{\prime}})v^{\gamma^{\prime}-1}. In particular, when γ=2\gamma^{\prime}=2 and f(v2)=Kαv2f(v^{2})=-K_{\alpha}\ast v^{2} in (1.15), the equation is the standard nonlinear Schrödinger equation with the Hartree-type aggregation term.

Inspired by the relation between Schrödinger equations and Mean-field Games systems discussed above, furthermore, the results of Cirant et al. [14] and Bernardini et al. [4, 5], we focus on the existence and asymptotic behaviors of ground states to (1.4) when α=nγ.\alpha=n-\gamma^{\prime}. In particular, Bernardini and Cesaroni studied the subcritical mass exponent case with α(nγ,n)\alpha\in(n-\gamma^{\prime},n) extensively via the variational method. It is well-known that system (1.4) admits the following variational structure:

(m,w):=n[mL(wm)+V(x)m+F(m)]𝑑x,\displaystyle\mathcal{E}(m,w):=\int_{\mathbb{R}^{n}}\left[mL\bigg{(}-\frac{w}{m}\bigg{)}+V(x)m+F(m)\right]\,dx, (1.17)

where F(m):=12(Kαm)mF(m):=-\frac{1}{2}(K_{\alpha}\ast m)m for m0m\geq 0 and F(m)=0F(m)=0 for m0.m\leq 0. Here Lagrangian LL is defined by

L(wm):={suppn(pwmH(p)),m>0,0,(m,w)=(0,0),+,otherwise.\displaystyle L\bigg{(}-\frac{w}{m}\bigg{)}:=\left\{\begin{array}[]{ll}\sup\limits_{p\in\mathbb{R}^{n}}\big{(}-\frac{p\cdot w}{m}-H(p)\big{)},&m>0,\\ 0,&(m,w)=(0,0),\\ +\infty,&\text{otherwise}.\end{array}\right. (1.21)

To explain the range of exponent α\alpha, we are concerned with the following constrained minimization problem:

eα,M:=inf(m,w)𝒦M(m,w),\displaystyle e_{\alpha,M}:=\inf_{(m,w)\in\mathcal{K}_{M}}\mathcal{E}(m,w), (1.22)

where the admissible set 𝒦M\mathcal{K}_{M} is given by

𝒦M:={\displaystyle\mathcal{K}_{M}:=\Big{\{} (m,w)(L1(n)W1,q^(n))×L1(n)\displaystyle(m,w)\in(L^{1}(\mathbb{R}^{n})\cap W^{1,\hat{q}}(\mathbb{R}^{n}))\times L^{1}(\mathbb{R}^{n})
s. t. nmφdx=nwφdx,φCc(n),\displaystyle\text{s. t. }\int_{\mathbb{R}^{n}}\nabla m\cdot\nabla\varphi\,dx=\int_{\mathbb{R}^{n}}w\cdot\nabla\varphi\,dx,\forall\varphi\in C_{c}^{\infty}(\mathbb{R}^{n}),
nV(x)mdx<+,nmdx=M>0,m0 a.e. },\displaystyle\int_{\mathbb{R}^{n}}V(x)m\,dx<+\infty,~{}\int_{\mathbb{R}^{n}}m\,dx=M>0,~{}m\geq 0\text{~{}a.e.~{}}\Big{\}}, (1.23)

with

q^:=nnγ+1, for each γ<n.\hat{q}:=\frac{n}{n-\gamma^{\prime}+1},\text{ for each }\gamma^{\prime}<n. (1.24)

It is straightforward to show that eα,M<+.e_{\alpha,M}<+\infty. Indeed, by choosing (ms,ws)=(ce|x|,xe|x||x|)(m_{s},w_{s})=\Big{(}ce^{-|x|},-\frac{xe^{-|x|}}{|x|}\Big{)} with cc determined by nm𝑑x=M\int_{\mathbb{R}^{n}}m\,dx=M and ws=msw_{s}=\nabla m_{s}, then one has (ms,ws)𝒦α,M(m_{s},w_{s})\in\mathcal{K}_{\alpha,M} and (ms,ws)<+,\mathcal{E}(m_{s},w_{s})<+\infty, which implies eα,M<+.e_{\alpha,M}<+\infty. Now, we mention that the lower bound α>nγ\alpha>n-\gamma^{\prime} is a necessary condition to guarantee that eα,M>e_{\alpha,M}>-\infty for all M>0M>0. To clarify this, we find if α<nγ,\alpha<n-\gamma^{\prime}, for any (m¯,w¯)𝒦α,M,(\bar{m},\bar{w})\in\mathcal{K}_{\alpha,M},

(m¯δ,w¯δ) as δ0+,\displaystyle\mathcal{E}(\bar{m}_{\delta},\bar{w}_{\delta})\rightarrow-\infty\text{ as }\delta\rightarrow 0^{+},

where (m¯δ,w¯δ)(\bar{m}_{\delta},\bar{w}_{\delta}) is defined as (m¯δ,w¯δ):=(δnm¯(δ1x),δ(n+1)w¯(δ1x))𝒦α,M(\bar{m}_{\delta},\bar{w}_{\delta}):=(\delta^{-n}\bar{m}(\delta^{-1}x),\delta^{-(n+1)}\bar{w}(\delta^{-1}x))\in\mathcal{K}_{\alpha,M} and δ\delta is chosen such that δnnm¯(δ1x)𝑑xM.\delta^{-n}\int_{\mathbb{R}^{n}}\bar{m}(\delta^{-1}x)\,dx\equiv M. Based on the discussion stated above, Bernardini and Cesaroni employed the direct method and the concentration-compactness approach to investigate the ground states to (1.4) with HH given by (1.6) when α\alpha satisfies nγ<α<nn-\gamma^{\prime}<\alpha<n in (1.5). In this paper, similarly as the work finished in [14], we shall study the existence and blow-up behaviors of ground states to (1.4) under the critical mass exponent case. We also would like to mention that there exists the other critical exponent α=n2γ\alpha=n-2\gamma^{\prime} from the restriction of Sobolev embedding Theorem. Next, we state our main results in Subsection 1.2.

Remark 1.1.

We would like to mention that, throughout the paper, we shall only consider the case γ<n\gamma^{\prime}<n. Since when γ<n\gamma^{\prime}<n, the minimization problem (1.22) will be well-posed for any α[nγ,n)\alpha\in[n-\gamma^{\prime},n), where the relevant discussions are shown in [4, 5].

1.2 Main results

We consider Hamiltonian HH satisfies (1.6) and f(m):=Kαmf(m):=-K_{\alpha}*m in (1.11), where KαK_{\alpha} is the Riesz potential of order α[nγ,n)\alpha\in[n-\gamma^{\prime},n) defined by Kα(x)=1|x|nαK_{\alpha}(x)=\frac{1}{|x|^{n-\alpha}}. In particular, we assume potential VV is locally Hölder continuous and satisfies

  • (V1).

    infxnV(x)=0V(x)Lloc(n)\inf\limits_{x\in\mathbb{R}^{n}}V(x)=0\leq V(x)\in L^{\infty}_{\text{loc}}(\mathbb{R}^{n}).

  • (V2).

    there exist positive constants C,C¯,KC,\bar{C},K and b,δb,\delta such that

    C(1+|x|b)V(x)C¯eδ|x|,xn;\displaystyle C(1+|x|^{b})\leq V(x)\leq\bar{C}e^{\delta|x|},\ \ \forall x\in\mathbb{R}^{n}; (1.25a)
    0<CV(x+y)V(x)C¯ for |x|K with |y|<2;\displaystyle 0<C\leq\frac{V(x+y)}{V(x)}\leq\bar{C}\text{~{}for~{}}|x|\geq K\text{~{}with~{}}|y|<2; (1.25b)
    supν[0,1]V(νx)C¯V(x) for |x|K.\displaystyle\sup_{\nu\in[0,1]}V(\nu x)\leq\bar{C}V(x)\text{~{}for~{}}|x|\geq K. (1.25c)
  • (V3).

    |𝒵|=0|\mathcal{Z}|=0 with 𝒵:={xn|V(x)=0}\mathcal{Z}:=\{x\in\mathbb{R}^{n}~{}|~{}V(x)=0\}.

With the assumptions shown above, we shall classify the existence of ground states to (1.4) with α=nγ\alpha=n-\gamma^{\prime} in terms of the total mass of density via the variational method. Compared to the arguments for the existence of ground states to the Mean-field Games system with a local coupling, one has to control mm in some LpL^{p} space with the aid of the nonlinearity in (1.17). Motivated by this, we exploit Hardy-Littlewood-Sobolev inequality stated in Appendix A and establish desired estimates.

One of our main goals is to study the attainability of the constrained minimization problem (1.22)

with the critical mass exponent α=α\alpha=\alpha^{*}, namely,

eα,M:=inf(m,w)𝒦M(m,w),\displaystyle e_{\alpha^{*},M}:=\inf_{(m,w)\in\mathcal{K}_{M}}\mathcal{E}(m,w), (1.26)

where 𝒦M\mathcal{K}_{M} is given in (1.1) and the energy (m,w)\mathcal{E}(m,w) (1.17) is precisely written as

(m,w)=CLn|wm|γm𝑑x+nV(x)m𝑑x12nm(x)(Kαm)(x)𝑑x\displaystyle\mathcal{E}(m,w)=C_{L}\int_{\mathbb{R}^{n}}\Big{|}\frac{w}{m}\Big{|}^{\gamma^{\prime}}m\,dx+\int_{\mathbb{R}^{n}}V(x)m\,dx-\frac{1}{2}\int_{\mathbb{R}^{n}}m(x)(K_{\alpha^{*}}*m)(x)\,dx (1.27)

and

nm(x)(Kαm)(x)𝑑x=nnm(x)m(y)|xy|nα𝑑x𝑑y.\int_{\mathbb{R}^{n}}m(x)(K_{\alpha^{*}}*m)(x)\,dx=\int_{\mathbb{R}^{n}}\int_{\mathbb{R}^{n}}\frac{m(x)m(y)}{|x-y|^{n-\alpha^{*}}}\,dx\,dy.

To this end, similarly as discussed in [14], we have to first investigate the Gagliardo-Nirenberg type’s inequality up to the critical mass exponent, which is

Γα:=inf(m,w)𝒜(CLnm|wm|γ𝑑x)nαγ(nm𝑑x)2γ+αnγnm(x)(Kαm)(x)𝑑x,α[nγ,n),\displaystyle\Gamma_{\alpha}:=\inf_{(m,w)\in\mathcal{A}}\frac{\Big{(}C_{L}\int_{\mathbb{R}^{n}}m\big{|}\frac{w}{m}\big{|}^{\gamma^{\prime}}\,dx\Big{)}^{\frac{n-\alpha}{\gamma^{\prime}}}\Big{(}\int_{\mathbb{R}^{n}}m\,dx\Big{)}^{\frac{2\gamma^{\prime}+\alpha-n}{\gamma^{\prime}}}}{\int_{\mathbb{R}^{n}}m(x)(K_{\alpha}*m)(x)\,dx},\ \ \alpha\in[n-\gamma^{\prime},n), (1.28)

where

𝒜:={\displaystyle\mathcal{A}:=\Big{\{} (m,w)(L1(n)W1,q^(n))×L1(n)\displaystyle(m,w)\in(L^{1}(\mathbb{R}^{n})\cap W^{1,\hat{q}}(\mathbb{R}^{n}))\times L^{1}(\mathbb{R}^{n})
s. t. nmφdx=nwφdx,φCc(n),n|x|bmdx<+,m,0 a.e. },\displaystyle\text{s. t. }\int_{\mathbb{R}^{n}}\nabla m\cdot\nabla\varphi\,dx=\int_{\mathbb{R}^{n}}w\cdot\nabla\varphi\,dx,\forall\varphi\in C_{c}^{\infty}(\mathbb{R}^{n}),~{}\int_{\mathbb{R}^{n}}|x|^{b}m\,dx<+\infty,~{}m\geq,\not\equiv 0\text{~{}a.e.~{}}\Big{\}}, (1.29)

with q^\hat{q} defined as (1.24) and b>0b>0. It is worthy mentioning that problem (1.28) is scaling invariant under the scaling (tβm(tx),tβ+1w(tx))(t^{\beta}m(tx),t^{\beta+1}w(tx)) for any t>0t>0 and β>0.\beta>0.

With the help of the conclusions shown in [4], we can prove the existence of minimizers to (1.28) for any α(nγ,n)\alpha\in(n-\gamma^{\prime},n). Then, we perform the approximation argument to study the case of α=nγ\alpha=n-\gamma^{\prime}. In fact, we have

Theorem 1.1.

Suppose α=α:=nγ\alpha=\alpha^{*}:=n-\gamma^{\prime} in (1.28) and γ(1,n)\gamma^{\prime}\in(1,n), then we have Γα\Gamma_{\alpha^{*}} is finite and attained by some minimizer (mα,wα)𝒜(m_{\alpha^{*}},w_{\alpha^{*}})\in\mathcal{A}. Moreover, we have there exists a classical solution (mα,uα)W1,p(n)×C2(n)\big{(}m_{\alpha^{*}},u_{\alpha^{*}}\big{)}\in W^{1,p}(\mathbb{R}^{n})\times C^{2}(\mathbb{R}^{n}), p>1,\forall p>1, to the following Mean-field Games systems:

{Δu+CH|u|γ1M=Kαm,xn,Δm+CHγ(m|u|γ2u)=0,xn,w=CHγm|u|γ2u,nm𝑑x=M,\displaystyle\left\{\begin{array}[]{ll}-\Delta u+C_{H}|\nabla u|^{\gamma}-\frac{1}{M^{*}}=-K_{\alpha^{*}}*m,&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},\\ w=-C_{H}\gamma m\big{|}\nabla u\big{|}^{\gamma-2}\nabla u,\ \int_{\mathbb{R}^{n}}m\,dx=M^{*},\end{array}\right. (1.33)

where

M:=2Γα.\displaystyle M^{*}:=2\Gamma_{\alpha^{*}}. (1.34)

In particular, there exists constants c1>0c_{1}>0 and c2>0c_{2}>0 such that 0<mα(x)c1ec2|x|0<m_{\alpha^{*}}(x)\leq c_{1}e^{-c_{2}|x|}.

Theorem 1.1 implies the best constant in (1.28) exists even if α=α:=nγ.\alpha=\alpha^{*}:=n-\gamma^{\prime}. Next, with the aid of Theorem 1.1, we are able to study the attainability of eα,Me_{\alpha^{*},M} and classify the existence of minimizers to (1.26), which is

Theorem 1.2.

Suppose VV satisfies assumption (V1)-(V2) and M=2ΓαM^{*}=2\Gamma_{\alpha^{*}}, where Γα\Gamma_{\alpha^{*}} is shown in (1.34), the we have the following alternatives:

  • (i).

    If 0<M<M0<M<M^{*}, problem (1.26) admits at least one minimizer (mM,wM)W1,p(n)×Lp(n)(m_{M},w_{M})\in W^{1,p}(\mathbb{R}^{n})\times L^{p}(\mathbb{R}^{n}),p>1\forall p>1, which satisfies for some λM\lambda_{M}\in\mathbb{R},

    {ΔuM+CH|uM|γ+λM=V(x)K(nγ)mM,ΔmM+CHγ(mM|uM|γ2uM)=0,wM=CHγmM|uM|γ2uM,nmM𝑑x=M<M.\displaystyle\left\{\begin{array}[]{ll}-\Delta u_{M}+C_{H}|\nabla u_{M}|^{\gamma}+\lambda_{M}=V(x)-K_{(n-\gamma^{\prime})}\ast m_{M},\\ \Delta m_{M}+C_{H}\gamma\nabla\cdot(m_{M}|\nabla u_{M}|^{\gamma-2}\nabla u_{M})=0,\\ w_{M}=-C_{H}\gamma m_{M}|\nabla u_{M}|^{\gamma-2}\nabla u_{M},\ \int_{\mathbb{R}^{n}}m_{M}\,dx=M<M^{*}.\end{array}\right. (1.38)
  • (ii).

    If M>MM>M^{*}, problem (1.26) does not admit any minimizer.

  • (iii).

    If M=MM=M^{*} and potential VV satisfies (V3) additionally, then there does not exist any minimizer to problem (1.26).

Remark 1.2.

We remark that in case (i), the LL^{\infty} estimates of mm is crucial due to the maximal regularity properties of Hamilton-Jacobi equations. Following the arguments in [10, 14], we perform the blow-up analysis to obtain the desired estimates.

Theorem 1.2 indicates that the minimizers to (1.26) do not exist when MM is large enough. A natural question is the behaviors of ground states as MMM\nearrow M^{*}, where MM^{*} is the existence threshold defined by (1.34). To explore this, we perform the scaling argument and investigate the convergence to get

Theorem 1.3.

Suppose that V(x)V(x) satisfies (V1)(V3)(V1)-(V3) and let (mM,wM)(m_{M},w_{M}) be the minimizer of eα,Me_{\alpha^{*},M} given in Theorem 1.2 with 0<M<M0<M<M^{*}. Then, we have

  • (i).
    εM=ε:=(CLn|wMmM|γmM𝑑x)1γ0 as MM.\displaystyle{\varepsilon}_{M}={\varepsilon}:=\Big{(}C_{L}\int_{\mathbb{R}^{n}}\bigg{|}\frac{w_{M}}{m_{M}}\bigg{|}^{\gamma^{\prime}}m_{M}\,dx\Big{)}^{-\frac{1}{\gamma^{\prime}}}\rightarrow 0\text{~{}as~{}}M\nearrow M^{*}. (1.39)
  • (ii).

    Let {xε}\{x_{\varepsilon}\} be one of the global minimum points of uMu_{M}, then dist(xε,𝒵)0\text{dist}(x_{\varepsilon},\mathcal{Z})\rightarrow 0 as MMM\nearrow M^{*}, where 𝒵={xn|V(x)=0}\mathcal{Z}=\{x\in\mathbb{R}^{n}~{}|~{}V(x)=0\}. Moreover,

    uε:=ε2γγ1uM(εx+xε),mε:=εnmM(εx+xε),wε:=εn+1wM(εx+xε),\displaystyle u_{\varepsilon}:=\varepsilon^{\frac{2-\gamma}{\gamma-1}}u_{M}(\varepsilon x+x_{\varepsilon}),~{}m_{\varepsilon}:=\varepsilon^{n}m_{M}(\varepsilon x+x_{\varepsilon}),~{}w_{\varepsilon}:=\varepsilon^{n+1}w_{M}(\varepsilon x+x_{\varepsilon}), (1.40)

    satisfies up to a subsequence,

    uεu0 in Cloc2(n),mεm0 in Lp(n)p[1,q^), and wεw0 in Lq^(n),u_{\varepsilon}\rightarrow u_{0}\text{ in }C^{2}_{\rm loc}(\mathbb{R}^{n}),~{}m_{\varepsilon}\rightarrow m_{0}\text{ in }L^{p}(\mathbb{R}^{n})~{}\forall~{}p\in[1,{\hat{q}}^{*}),~{}\text{ and }w_{\varepsilon}\rightharpoonup w_{0}\text{ in }L^{\hat{q}}(\mathbb{R}^{n}), (1.41)

    where (m0,w0)(m_{0},w_{0}) is a minimizer of (1.28), and (u0,m0,w0)(u_{0},m_{0},w_{0}) satisfies (1.33). In particular, when VV satisfies

    CV1(max{|x|CV,0})bV(x)CV(1+|x|)b, for some b,CV>0.\displaystyle C_{V}^{-1}(\max\{|x|-C_{V},0\})^{b}\leq V(x)\leq C_{V}(1+|x|)^{b},~{}~{}\text{ for some }b,C_{V}>0. (1.42)

    and x¯ε\bar{x}_{\varepsilon} denotes any one of global maximum points of mMm_{M}, then

    lim supε0+|x¯εxε|ε<+.\displaystyle\limsup_{\varepsilon\rightarrow 0^{+}}\frac{|\bar{x}_{\varepsilon}-x_{\varepsilon}|}{\varepsilon}<+\infty. (1.43)

Theorem 1.3 implies as MMM\nearrow M^{*}, the ground states to (1.4) concentrate and their basic blow-up behaviors are captured by the least energy solution to potential-free Mean-field Games systems with some mild assumptions imposed on VV. Moreover, by imposing some typical local expansions on potential V(x)V(x), one can obtain the refined asymptotics of ground states, which are summarized as

Theorem 1.4.

Assume that all conditions in Theorem 1.3 hold and suppose that VV has l+l\in\mathbb{N}_{+} distinct zeros denoted by {P1,,Pl}\{P_{1},\cdots,P_{l}\} and there exist ai>0a_{i}>0, qi>0q_{i}>0 and d>0d>0 such that

V(x)=ai|xPi|qi+O(|xPi|qi+1), 0<|xPi|d,i=1,,l.\displaystyle V(x)=a_{i}|x-P_{i}|^{q_{i}}+O\big{(}|x-P_{i}|^{q_{i}+1}\big{)},\ \ 0<|x-P_{i}|\leq d,\ i=1,\cdots,l.

Define

Z:={Pi|qi=q,i=1,,l} and Z0:={Pi|qiZ and μi=μ,i=1,,l},Z:=\{P_{i}~{}|~{}q_{i}=q,\ i=1,\cdots,l\}~{}\text{ and }~{}Z_{0}:=\{P_{i}~{}|~{}q_{i}\in Z\text{~{}and~{}}\mu_{i}=\mu,i=1,\cdots,l\},

where q:=max{q1,,ql}q:=\max\{q_{1},\cdots,q_{l}\} and μ:=min{μi|PiZ,i=1,,l}\mu:=\min\{\mu_{i}~{}|~{}P_{i}\in Z,i=1,\cdots,l\} with

μi:=minynHi(y),Hi(y):=nai|x+y|qim0(x)𝑑x,i=1,,l.\mu_{i}:=\min\limits_{y\in\mathbb{R}^{n}}H_{i}(y),\ H_{i}(y):=\int_{\mathbb{R}^{n}}a_{i}|x+y|^{q_{i}}m_{0}(x)\,dx,~{}i=1,\cdots,l.

Let (mε,wε,uε)(m_{\varepsilon},w_{\varepsilon},u_{\varepsilon}) be the sequence given by (1.40) and (m0,w0,u0)(m_{0},w_{0},u_{0}) be the limiting solution. Then we have xεPiZ0x_{\varepsilon}\rightarrow P_{i}\in Z_{0}. Moreover, as MM,M\nearrow M^{*},

eα,Mq+γq(qμγ)γγ+q[1MM]qγq1,\displaystyle\frac{e_{\alpha^{*},M}}{\frac{q+\gamma^{\prime}}{q}\Big{(}\frac{q\mu}{\gamma^{\prime}}\Big{)}^{\frac{\gamma^{\prime}}{\gamma^{\prime}+q}}\Big{[}1-\frac{M}{M^{*}}\Big{]}^{\frac{q}{\gamma^{\prime}-q}}}\rightarrow 1,

and

ε(γqμ(1MM)]1γ+q1,\displaystyle\frac{\varepsilon}{\left(\frac{\gamma^{\prime}}{q\mu}\left(1-\frac{M}{M^{*}}\right)\right]^{\frac{1}{\gamma^{\prime}+q}}}\rightarrow 1, (1.44)

where eα,Me_{\alpha^{*},M} and ε=εM\varepsilon=\varepsilon_{M} are given by (1.22) and (1.39), respectively. In particular, up to a subsequence,

xεPiεMy0 with PiZ0 and Hi(y0)=infynHi(y)=μ.\displaystyle\frac{x_{\varepsilon}-P_{i}}{\varepsilon_{M}}\rightarrow y_{0}~{}\text{ with }~{}P_{i}\in Z_{0}~{}\text{ and }~{}H_{i}(y_{0})=\inf_{y\in\mathbb{R}^{n}}H_{i}(y)=\mu. (1.45)

Theorem 1.4 demonstrates that under certain types of potentials with the local polynomial expansions, ground states to (1.4) are localized as MMM\nearrow M^{*}, in which the locations converge to the flattest minima of VV.

The rest of this paper is organized as follows. In Section 2, we give some preliminaries for the investigation of ground states to (1.4) with α=nγ\alpha=n-\gamma^{\prime}. Section 3 is dedicated to the formulation of the optimal Gagliardo-Nirenberg type’s inequality and the proof of Theorem 1.1. In Section 4, we prove Theorem 1.2 by using the blow-up analysis and the Gagliardo-Nirenberg inequality shown in Theorem 1.1. Finally, in Section 5, we focus on Theorem 1.3 and 1.4, i.e. discuss the existence and concentration behaviors of ground states in some singular limit of MM given in (1.4). Without confusing readers, C>0C>0 is chosen as a generic constant, which may vary line to line.

2 Preliminaries

This section is devoted to some preliminary results including existence and regularities of the solutions to Hamilton-Jacobi and Fokker-Planck equations.

2.1 Hamilton-Jacobi Equations

Consider the following Hamilton-Jacobi equation:

Δu+CH|u|γ=f,xΩ,\displaystyle-\Delta u+C_{H}|\nabla u|^{\gamma}=f,~{}x\in\Omega, (2.1)

where Ω\Omega is a bounded domain with the smooth boundary, CH>0C_{H}>0 and γ>1\gamma>1. For the local W2,pW^{2,p} estimates of the solutions uu to (2.1), we have

Lemma 2.1 (C.f. Theorem 1.1 in [14]).

Let CH>0C_{H}>0, p>nγp>\frac{n}{\gamma^{\prime}}, γnn1\gamma\geq\frac{n}{n-1} and fLp(Ω)f\in L^{p}(\Omega). Suppose uW2,p(Ω)u\in W^{2,p}(\Omega) solves (2.1) in the strong sense. Then for each M>0M>0 and ΩΩ\Omega^{\prime}\subset\subset\Omega, we have

uLp(Ω)+D2uLp(Ω)C,\|\nabla u\|_{L^{p}(\Omega^{\prime})}+\|D^{2}u\|_{L^{p}(\Omega^{\prime})}\leq C,

where fLp(Ω)M\|f\|_{L^{p}(\Omega)}\leq M and the constant C=C(M,dist(Ω,Ω),n,p,CH,γ)>0C=C(M,\mathrm{dist}(\Omega^{\prime},\partial\Omega),n,p,C_{H},\gamma^{\prime})>0.

Since our arguments in Section 3, 4 and 5 involve some limits of solution sequences, we also focus on the following sequence of 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}, (2.2)

where CH>0C_{H}>0 and γ>1\gamma>1 are fixed. Here (uk,λk)(u_{k},\lambda_{k}) denote the solution pair to (2.2). Concerning the regularities of uku_{k}, we obtain

Lemma 2.2 (C.f. Lemma 3.1 in [14]).

Assume that fkL(n)f_{k}\in L^{\infty}(\mathbb{R}^{n}) satisfies fkLCf\|f_{k}\|_{L^{\infty}}\leq C_{f} and |λk|λ|\lambda_{k}|\leq\lambda. Suppose the potential functions Vk(x)V_{k}(x) are uniformly local Hölder continuous satisfying 0Vk(x)+0\leq V_{k}(x)\rightarrow+\infty as |x|+,|x|\rightarrow+\infty, and there exists R>0R>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, (2.3)

where the positive constants CfC_{f}, λ\lambda, RR, C1C_{1} and C2C_{2} are independent of kk. Define (uk,λk)C2(n)×(u_{k},\lambda_{k})\in C^{2}(\mathbb{R}^{n})\times\mathbb{R} as the solutions to (2.2). Then, we have 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}, (2.4)

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

In particular, if each VkV_{k} satisfies

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}, (2.5)

where b0b\geq 0 and CF>0C_{F}>0 independent of k,k, 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}, (2.6)

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

For the lower bounds of uku_{k}, we have the following results:

Lemma 2.3 (C.f. Lemma 3.2 in [14]).

Suppose all conditions in Lemma 2.2 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)C3Vk1γ(x)C4, xn,for all k.\displaystyle u_{k}(x)\geq C_{3}V^{\frac{1}{\gamma}}_{k}(x)-C_{4},\text{~{}}\forall x\in\mathbb{R}^{n},~{}\text{for all }k. (2.7)

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}, (2.8)

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

uk(x)C3|x|1+bγC4, for allk,xn.\displaystyle u_{k}(x)\geq C_{3}|x|^{1+\frac{b}{\gamma}}-C_{4},\text{~{}for all}k,x\in\mathbb{R}^{n}. (2.9)

If b=0b=0 in (2.8) 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, (2.10)

then (2.9) also holds.

The following results are concerned with the existence of the classical solution to (2.2), which are

Lemma 2.4 (C.f. Lemma 3.3 in [14]).

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

λ¯k:=sup{λ|(2.2) 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})\}. (2.11)

Then we have

  • (i).

    λ¯k\bar{\lambda}_{k} are finite for every kk and (2.2) 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|(2.2) 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 (2.5) 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|bγ+1C,xn.\displaystyle u_{k}(x)\geq C|x|^{\frac{b}{\gamma}+1}-C,\forall x\in\mathbb{R}^{n}. (2.12)

    In particular, if Vk0V_{k}\equiv 0 and b=0b=0 in (1.42) 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}, (2.13)

    where K2>0K_{2}>0 is a large constant independent of kk, then (2.12) also holds.

(iii). If VkV_{k} satisfies (1.25b) with VV replaced by VkV_{k} and positive constants C1C_{1}, C2C_{2} and δ\delta independent of k,k, then there exist uniformly bounded from below classical solutions uku_{k} to problem (2.2) satisfying estimate (2.7).

2.2 Fokker-Planck Equations

Now, we focus on the following Fokker-Planck equations:

Δm+w=0,xn,\displaystyle-\Delta m+\nabla\cdot w=0,\ \ x\in\mathbb{R}^{n}, (2.14)

where ww is given and mm denotes the solution. Firstly, we state the regularity results of solutions to equation (2.14), which are

Lemma 2.5.

Assume that (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}) is a solution to (2.14) and

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 constant 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 [14]. ∎

Lemma 2.6 (C.f. Corollary 1.1 in [14]).

Assume that (m,w)(L1(n)L1+β(n)W1,q(n))×L1(n)(m,w)\in(L^{1}(\mathbb{R}^{n})\cap L^{1+\beta}(\mathbb{R}^{n})\cap W^{1,q}(\mathbb{R}^{n}))\times L^{1}(\mathbb{R}^{n}) is the solution to (2.14) with

1q=1γ+1γ(1+β).\frac{1}{q}=\frac{1}{\gamma^{\prime}}+\frac{1}{\gamma(1+\beta)}.

Then for β(0,γn]\beta\in(0,\frac{\gamma^{\prime}}{n}\big{]}, there exists a positive constant CC depending only on nn and β\beta such that

mLq(n)C(nm|wm|γ𝑑x)1γmL1+β1γ.\displaystyle\|\nabla m\|_{L^{q}(\mathbb{R}^{n})}\leq C\Big{(}\int_{\mathbb{R}^{n}}m\Big{|}\frac{w}{m}\Big{|}^{\gamma^{\prime}}\,dx\Big{)}^{\frac{1}{\gamma^{\prime}}}\|m\|_{L^{1+\beta}}^{\frac{1}{\gamma}}. (2.15)

Moreover, there exists a positive constant CC only depending on γ,\gamma^{\prime}, nn and α\alpha such that

mL1+β(n)1+βC(nm|wm|γ𝑑x)nβγ(nm𝑑x)(β+1)γnβγ.\displaystyle\|m\|^{1+\beta}_{L^{1+\beta}(\mathbb{R}^{n})}\leq C\bigg{(}\int_{\mathbb{R}^{n}}m\Big{|}\frac{w}{m}\Big{|}^{\gamma^{\prime}}\,dx\bigg{)}^{\frac{n\beta}{\gamma^{\prime}}}\bigg{(}\int_{\mathbb{R}^{n}}m\,dx\bigg{)}^{\frac{(\beta+1)\gamma^{\prime}-n\beta}{\gamma^{\prime}}}. (2.16)

Next, we discuss the exponential decay property of the solutions to system (1.11) and obtain

Lemma 2.7.

Assume that (u,m,λ)C2(n)×(W1,p(n)L1(n))×(u,m,\lambda)\in C^{2}(\mathbb{R}^{n})\times\big{(}W^{1,p}(\mathbb{R}^{n})\cap L^{1}(\mathbb{R}^{n})\big{)}\times\mathbb{R} with p>np>n and λ<0\lambda<0 is the solution of the following potential-free problem

{Δu+CH|u|γ+λ=Kαm,xn,Δm+CHγ(m|u|γ2u)=0,xn.\displaystyle\left\{\begin{array}[]{ll}-\Delta u+C_{H}|\nabla u|^{\gamma}+\lambda=-K_{\alpha}\ast m,&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. (2.19)

Suppose uu is bounded from below. 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}. (2.20)
Proof.

Noting that mW1,p(n)m\in W^{1,p}(\mathbb{R}^{n}) with p>np>n, we use Sobolev embedding to get mC0,θ(n)m\in C^{0,\theta}(\mathbb{R}^{n}) for some θ(0,1)\theta\in(0,1), and thus mL(n)m\in L^{\infty}(\mathbb{R}^{n}). Moreover, by using the fact that mL1(n)m\in L^{1}(\mathbb{R}^{n}) and the interpolation inequality, one finds mLq(n)m\in L^{q}(\mathbb{R}^{n}) for every q(1,)q\in(1,\infty). Therefore, invoking Lemma A.1 and Lemma A.2, one can obtain that KαmLβ(n)C0,θ1(n)K_{\alpha}\ast m\in L^{\beta}(\mathbb{R}^{n})\cap C^{0,\theta_{1}}(\mathbb{R}^{n}) for some β>1\beta>1 and θ1(0,1)\theta_{1}\in(0,1), which implies

Kαm0as|x|+.\displaystyle K_{\alpha}\ast m\rightarrow 0\ \ \text{as}\ \ |x|\rightarrow+\infty.

The rest of proof follows from [4, Proposition 4.2] and [14, Lemma 3.6]. ∎

Thanks to Lemma 2.7, we establish Pohozaev identities satisfied by the solution to system (2.19), which are

Lemma 2.8 (C.f. Lemma 3.1 in [5]).

Assume all conditions satisfied by (u,m,λ)(u,m,\lambda) hold in Lemma 2.7 and denote w=CHγm|u|γ2uw=-C_{H}\gamma m|\nabla u|^{\gamma-2}\nabla u. Then the following identities hold:

{λnm𝑑x=α+2γn2γnm(x)(Kαm)(x)𝑑x,CLnm|wm|γ𝑑x=nα2γnm(x)(Kαm)(x)𝑑x=(γ1)CHnm|u|γ𝑑x.\displaystyle\left\{\begin{array}[]{ll}\lambda\int_{\mathbb{R}^{n}}m\,dx=-\frac{\alpha+2\gamma^{\prime}-n}{2\gamma^{\prime}}\int_{\mathbb{R}^{n}}m(x)(K_{\alpha}*m)(x)\,dx,\\ C_{L}\int_{\mathbb{R}^{n}}m\big{|}\frac{w}{m}\big{|}^{\gamma^{\prime}}\,dx=\frac{n-\alpha}{2\gamma^{\prime}}\int_{\mathbb{R}^{n}}m(x)(K_{\alpha}*m)(x)\,dx=(\gamma-1)C_{H}\int_{\mathbb{R}^{n}}m|\nabla u|^{\gamma}\,dx.\end{array}\right. (2.23)
Proof.

Proceeding the similar argument shown in Lemma 3.7 of [14], we can prove this lemma. For the sake of completeness, we exhibit the proof briefly. First all, we multiply the uu-equation and mm-equation in (2.19) by mm and uu, respectively, then integrate them by parts and subtract the two identities to get

(1γ)CHnm|u|γ𝑑x+λnm𝑑x=nm(x)(Kαm)(x)𝑑x,\displaystyle(1-\gamma)C_{H}\int_{\mathbb{R}^{n}}m|\nabla u|^{\gamma}\,dx+\lambda\int_{\mathbb{R}^{n}}m\,dx=-\int_{\mathbb{R}^{n}}m(x)(K_{\alpha}*m)(x)\,dx, (2.24)

where we have used the exponential decay property of mm shown in Lemma 2.7 and the uniformly boundedness of u\nabla u stated in Lemma 2.2.

Next, we focus on the proof of the following identity:

nλnm𝑑xn+α2nm(x)(Kαm)(x)𝑑x+CHnγγ1nm|u|γ𝑑x=0.\displaystyle-n\lambda\int_{\mathbb{R}^{n}}m\,dx-\frac{n+\alpha}{2}\int_{\mathbb{R}^{n}}m(x)(K_{\alpha}*m)(x)\,dx+C_{H}\frac{n-\gamma^{\prime}}{\gamma^{\prime}-1}\int_{\mathbb{R}^{n}}m|\nabla u|^{\gamma}\,dx=0. (2.25)

In fact, by testing the first equation and the second equation in (2.19) against mx\nabla m\cdot x and ux\nabla u\cdot x, we apply the integration by parts to obtain

n((Km)(x)λ)mxdx=\displaystyle\int_{\mathbb{R}^{n}}(-(K*m)(x)-\lambda)\nabla m\cdot x\,dx= nu(mx)dxCHn(|u|γx)m𝑑x,\displaystyle{\int_{\mathbb{R}^{n}}\nabla u\cdot\nabla(\nabla m\cdot x)\,dx}-{C_{H}}\int_{\mathbb{R}^{n}}\nabla\cdot(|\nabla u|^{\gamma}x)m\,dx, (2.26)

and

CHn(|u|γ)xm𝑑x=nm(ux)dx+CHγn|u|γm𝑑x,\displaystyle-{C_{H}}\int_{\mathbb{R}^{n}}\nabla(|\nabla u|^{\gamma})\cdot xm\,dx=\int_{\mathbb{R}^{n}}\nabla m\cdot\nabla(\nabla u\cdot x)\,dx+C_{H}\gamma\int_{\mathbb{R}^{n}}|\nabla u|^{\gamma}m\,dx, (2.27)

where the boundary integrals vanish due to the decay property of mm and the upper bound of u.u. Also, we find

nu(mx)dx=\displaystyle\int_{\mathbb{R}^{n}}\nabla u\cdot\nabla(\nabla m\cdot x)\,dx= i,j=1nnuximxixjxj𝑑x+numdx\displaystyle\sum_{i,j=1}^{n}\int_{\mathbb{R}^{n}}u_{x_{i}}m_{x_{i}x_{j}}x_{j}\,dx+\int_{\mathbb{R}^{n}}\nabla u\cdot\nabla m\,dx
=\displaystyle= i,j=1nnmxiuxixjxj𝑑x+(1n)numdx\displaystyle-\sum_{i,j=1}^{n}\int_{\mathbb{R}^{n}}m_{x_{i}}u_{x_{i}x_{j}}x_{j}\,dx+(1-n)\int_{\mathbb{R}^{n}}\nabla u\cdot\nabla m\,dx
=\displaystyle= nm(ux)dx+(2n)numdx.\displaystyle-\int_{\mathbb{R}^{n}}\nabla m\cdot\nabla(\nabla u\cdot x)\,dx+(2-n)\int_{\mathbb{R}^{n}}\nabla u\cdot\nabla m\,dx. (2.28)

Collecting (2.26), (2.27) and (2.2), we have the following equality holds:

n((Km)(x)λ)mxdx=CH(γn)n|u|γm𝑑x+(2n)numdx.\displaystyle{\int_{\mathbb{R}^{n}}(-(K*m)(x)-\lambda)\nabla m\cdot x\,dx}=C_{H}\big{(}\gamma-{n}\big{)}\int_{\mathbb{R}^{n}}|\nabla u|^{\gamma}m\,dx+(2-n)\int_{\mathbb{R}^{n}}\nabla u\cdot\nabla m\,dx. (2.29)

With the help of the integration by parts, one further gets

nλnm𝑑xn+α2nm(x)(Km)(x)𝑑x+CH(γn)n|u|γm𝑑x+(2n)numdx=0,\displaystyle-n\lambda\int_{\mathbb{R}^{n}}m\,dx-\frac{n+\alpha}{2}\int_{\mathbb{R}^{n}}m(x)(K*m)(x)\,dx+C_{H}\Big{(}\gamma-{n}\Big{)}\int_{\mathbb{R}^{n}}|\nabla u|^{\gamma}m\,dx+(2-n)\int_{\mathbb{R}^{n}}\nabla u\cdot\nabla m\,dx=0,

which indicates (2.25) by using the uu-equation in (2.19). In addition, since w=CHγm|u|γ2uw=-C_{H}\gamma m|\nabla u|^{\gamma-2}\nabla u and CL=1γ(γCH)11γC_{L}=\frac{1}{\gamma^{\prime}}(\gamma C_{H})^{\frac{1}{1-\gamma}}, we obtain

CLnm|wm|γ𝑑x=CL(CHγ)γnm|u|γ𝑑x=(γ1)CHnm|u|γ𝑑x.\displaystyle C_{L}\int_{\mathbb{R}^{n}}m\bigg{|}\frac{w}{m}\bigg{|}^{\gamma^{\prime}}\,dx=C_{L}(C_{H}\gamma)^{\gamma^{\prime}}\int_{\mathbb{R}^{n}}m|\nabla u|^{\gamma}\,dx=(\gamma-1)C_{H}\int_{\mathbb{R}^{n}}m|\nabla u|^{\gamma}\,dx. (2.30)

Finally, by using (2.24), (2.25) and (2.30), we conclude that (2.23) holds.

We mention that the argument shown above hold only when γ2\gamma\geq 2 and in this case, the Fokker-Planck equation can be solved in the strong sense. When 1<γ<21<\gamma<2, one can only solve the Fokker-Planck equation in the weak sense. Whereas, we can replace HH with Hϵ(𝒑):=CH(ϵ+|𝒑|2)γ2H_{\epsilon}(\boldsymbol{p}):=C_{H}(\epsilon+|\boldsymbol{p}|^{2})^{\frac{\gamma}{2}} in (1.6) and proceed the same argument shown above with mϵm_{\epsilon}, then take the limit ϵ0\epsilon\rightarrow 0 to get the desired conclusion.

Now, we are well prepared to prove Theorem 1.1 and the arguments are shown in Section 3.

3 Optimal Gagliardo-Nirenberg Type’s Inequality

In this section, we are going to discuss the existence of minimizers to problem (1.28) and prove Theorem 1.1. As mentioned above, problem (1.28) is scaling invariant under the scaling (tβm(tx),tβ+1w(tx))(t^{\beta}m(tx),t^{\beta+1}w(tx)) for any t>0t>0 and β>0.\beta>0. Therefore, one can verify that (1.28) is equivalent to

Γα:=inf(m,w)missingAM(CLnm|wm|γ𝑑x)nαγ(nm𝑑x)2γ+αnγnm(x)(Kαm)(x)𝑑x,α[nγ,n),\displaystyle\Gamma_{\alpha}:=\inf_{(m,w)\in\mathcal{\mathcal{missing}}A_{M}}\frac{\Big{(}C_{L}\int_{\mathbb{R}^{n}}m\big{|}\frac{w}{m}\big{|}^{\gamma^{\prime}}\,dx\Big{)}^{\frac{n-\alpha}{\gamma^{\prime}}}\Big{(}\int_{\mathbb{R}^{n}}m\,dx\Big{)}^{\frac{2\gamma^{\prime}+\alpha-n}{\gamma^{\prime}}}}{\int_{\mathbb{R}^{n}}m(x)(K_{\alpha}*m)(x)\,dx},\ \ \alpha\in[n-\gamma^{\prime},n), (3.1)

where

𝒜M:={(m,w)𝒜,nm𝑑x=M>0}.\displaystyle\mathcal{A}_{M}:=\Big{\{}(m,w)\in\mathcal{A},\int_{\mathbb{R}^{n}}m\,dx=M>0\Big{\}}. (3.2)

Now, we start by studying the subcritical mass exponent case of problem (3.1), namely, α(nγ,n)\alpha\in(n-\gamma^{\prime},n). For this case, Bernardini [4] proved that there exists (u¯α,M,m¯α,M,λα,M)C2(n)×W1,p(n)×(\bar{u}_{\alpha,M},\bar{m}_{\alpha,M},\lambda_{\alpha,M})\in C^{2}(\mathbb{R}^{n})\times W^{1,p}(\mathbb{R}^{n})\times\mathbb{R} for every p(0,+)p\in(0,+\infty) solving the following system

{Δu+CH|u|γ+λ=Kαm,xn,Δm+CHγ(m|u|γ2u)=0,xn,nm𝑑x=M>0,\displaystyle\left\{\begin{array}[]{ll}-\Delta u+C_{H}|\nabla u|^{\gamma}+\lambda=-K_{\alpha}*m,&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},\\ \int_{\mathbb{R}^{n}}m\,dx=M>0,\end{array}\right. (3.6)

which is the classical solution to system (3.6). Furthermore, the author showed there exist c1,M,c2,M>0c_{1,M},c_{2,M}>0 such that

0<m¯α,M<c1,Mec2,M|x|.0<\bar{m}_{\alpha,M}<c_{1,M}e^{-c_{2,M}|x|}. (3.7)

In particular, Bernardini obtained the following minimization problem

e0,α,M:=inf(m,w)𝒜M0(m,w)\displaystyle e_{0,\alpha,M}:=\inf\limits_{(m,w)\in{\mathcal{A}}_{M}}\mathcal{E}_{0}(m,w) (3.8)

with

0(m,w)=n(CLm|wm|γ)𝑑x12nm(x)(Kαm)(x)𝑑x\displaystyle\mathcal{E}_{0}(m,w)=\int_{\mathbb{R}^{n}}\Bigg{(}C_{L}m\bigg{|}\frac{w}{m}\bigg{|}^{\gamma^{\prime}}\Bigg{)}\,dx-\frac{1}{2}\int_{\mathbb{R}^{n}}m(x)(K_{\alpha}*m)(x)\,dx (3.9)

is attained by the pair (m¯α,M,w¯α,M)(\bar{m}_{\alpha,M},\bar{w}_{\alpha,M}) with w¯α,M=CHγm¯α,M|u¯α,M|γ2u¯α,M\bar{w}_{\alpha,M}=-C_{H}\gamma\bar{m}_{\alpha,M}|\nabla\bar{u}_{\alpha,M}|^{\gamma-2}\nabla\bar{u}_{\alpha,M}. In addition, invoking Lemma 2.8, one finds

{λnm¯α,M𝑑x=2γ+αn2γnm¯α,M(x)(Kαm¯α,M)(x)𝑑x,CLnm¯α,M|w¯α,Mm¯α,M|γ𝑑x=nα2γnm¯α,M(x)(Kαm¯α,M)(x)𝑑x=(γ1)CHnm¯α,M|u|γ𝑑x.\displaystyle\left\{\begin{array}[]{ll}\lambda\int_{\mathbb{R}^{n}}\bar{m}_{\alpha,M}\,dx=-\frac{2\gamma^{\prime}+\alpha-n}{2\gamma^{\prime}}\int_{\mathbb{R}^{n}}\bar{m}_{\alpha,M}(x)(K_{\alpha}*\bar{m}_{\alpha,M})(x)\,dx,\\ C_{L}\int_{\mathbb{R}^{n}}\bar{m}_{\alpha,M}\big{|}\frac{\bar{w}_{\alpha,M}}{\bar{m}_{\alpha,M}}\big{|}^{\gamma^{\prime}}\,dx=\frac{n-\alpha}{2\gamma^{\prime}}\int_{\mathbb{R}^{n}}\bar{m}_{\alpha,M}(x)(K_{\alpha}*\bar{m}_{\alpha,M})(x)\,dx=(\gamma-1)C_{H}\int_{\mathbb{R}^{n}}\bar{m}_{\alpha,M}|\nabla u|^{\gamma}\,dx.\end{array}\right. (3.12)

Collecting the results shown above, we are able to investigate a relationship between (m¯α,M,w¯α,M)(\bar{m}_{\alpha,M},\bar{w}_{\alpha,M}), the minimizer of (3.8) and the minimizer of problem (3.1), which is

Lemma 3.1.

For any fixed α(nγ,n)\alpha\in\big{(}n-\gamma^{\prime},n) and M>0M>0, problem (1.28) is attained by (m¯α,M,w¯α,M)(\bar{m}_{\alpha,M},\bar{w}_{\alpha,M}) with e0,α,M=0(m¯α,M,w¯α,M)e_{0,\alpha,M}=\mathcal{E}_{0}(\bar{m}_{\alpha,M},\bar{w}_{\alpha,M}). More precisely, we have

Γα=(nα)(e0,α,M)nαγγM2γ+αnγ2γ(γn+αnα)γ+αnγ.\displaystyle\Gamma_{\alpha}=\frac{(n-\alpha)(-e_{0,\alpha,M})^{\frac{n-\alpha-\gamma^{\prime}}{\gamma^{\prime}}}M^{\frac{2\gamma^{\prime}+\alpha-n}{\gamma^{\prime}}}}{2\gamma^{\prime}}\Bigg{(}\frac{\gamma^{\prime}-n+\alpha}{n-\alpha}\Bigg{)}^{\frac{\gamma^{\prime}+\alpha-n}{\gamma^{\prime}}}. (3.13)
Proof.

We follow the procedures shown in Lemma 4.1 of [14] to prove this lemma. First of all, we define

Gα(m,w):\displaystyle G_{\alpha}(m,w): =(CLnm|wm|γ𝑑x)nαγ(nm𝑑x)2γ+αnγnm(x)(Kαm)(x)𝑑x.\displaystyle=\frac{\Big{(}C_{L}\int_{\mathbb{R}^{n}}m\big{|}\frac{w}{m}\big{|}^{\gamma^{\prime}}\,dx\Big{)}^{\frac{n-\alpha}{\gamma^{\prime}}}\Big{(}\int_{\mathbb{R}^{n}}m\,dx\Big{)}^{\frac{2\gamma^{\prime}+\alpha-n}{\gamma^{\prime}}}}{\int_{\mathbb{R}^{n}}m(x)(K_{\alpha}*m)(x)\,dx}. (3.14)

With the definition (3.14), the minimization problem (3.1) can be rewritten as

Γα=inf(m,w)𝒜MGα(m,w).\displaystyle\Gamma_{\alpha}=\inf_{(m,w)\in{{\mathcal{A}}}_{M}}G_{\alpha}(m,w). (3.15)

Now, we aim to verify that (3.15) is attained by (m¯α,M,w¯α,M)(\bar{m}_{\alpha,M},\bar{w}_{\alpha,M}), which is the minimizer of (3.8). First of all, we estimate the energy 0\mathcal{E}_{0} defined by (3.8) from below. We remark that Gα(m,w)=+G_{\alpha}(m,w)=+\infty provided with nm|wm|γ𝑑x=+\int_{\mathbb{R}^{n}}m\big{|}\frac{w}{m}\big{|}^{\gamma^{\prime}}\,dx=+\infty. Thus, we only need to consider the case that (m,w)𝒜M(m,w)\in{{\mathcal{A}}}_{M} satisfying nm|wm|γ𝑑x<\int_{\mathbb{R}^{n}}m\big{|}\frac{w}{m}\big{|}^{\gamma^{\prime}}\,dx<\infty. Define (mμ(x),wμ(x))=(μnm(μx),μn+1w(μx))(m_{\mu}(x),w_{\mu}(x))=(\mu^{n}m(\mu x),\mu^{n+1}w(\mu x)) for μ+\{0}\mu\in\mathbb{R}^{+}\backslash\{0\}, then we have

0(mμ,wμ)=\displaystyle\mathcal{E}_{0}(m_{\mu},w_{\mu})= μγnCLm|wm|γ𝑑x12μnαnm(x)(Kαm)(x)𝑑x\displaystyle\mu^{\gamma^{\prime}}\int_{\mathbb{R}^{n}}C_{L}m\Big{|}\frac{w}{m}\Big{|}^{\gamma^{\prime}}\,dx-\frac{1}{2}\mu^{n-\alpha}\int_{\mathbb{R}^{n}}m(x)(K_{\alpha}*m)(x)\,dx
\displaystyle\geq (nα2γ)γγn+α(γn+αnα)(nm(x)(Kαm)(x)𝑑x)γγn+α(CLnm|wm|γ𝑑x)nαγn+α,\displaystyle-\bigg{(}\frac{n-\alpha}{2\gamma^{\prime}}\bigg{)}^{\frac{\gamma^{\prime}}{\gamma^{\prime}-n+\alpha}}\Bigg{(}\frac{\gamma^{\prime}-n+\alpha}{n-\alpha}\Bigg{)}\bigg{(}\int_{\mathbb{R}^{n}}m(x)(K_{\alpha}*m)(x)\,dx\bigg{)}^{\frac{\gamma^{\prime}}{\gamma^{\prime}-n+\alpha}}\bigg{(}C_{L}\int_{\mathbb{R}^{n}}m\Big{|}\frac{w}{m}\Big{|}^{\gamma^{\prime}}\,dx\bigg{)}^{-\frac{n-\alpha}{\gamma^{\prime}-n+\alpha}}, (3.16)

where the equality holds if and only if

μ=[(nα)nm(x)(Kαm)(x)𝑑x2γCLnm|wm|γ𝑑x]1γn+α.\displaystyle\mu=\Bigg{[}\frac{(n-\alpha)\int_{\mathbb{R}^{n}}m(x)(K_{\alpha}*m)(x)\,dx}{2\gamma^{\prime}C_{L}\int_{\mathbb{R}^{n}}m\big{|}\frac{w}{m}\big{|}^{\gamma^{\prime}}\,dx}\Bigg{]}^{\frac{1}{\gamma^{\prime}-n+\alpha}}.

It then follows from the definition of e0,α,M:=inf(m,w)𝒜M0(m,w)e_{0,\alpha,M}:=\inf\limits_{(m,w)\in{{\mathcal{A}}}_{M}}\mathcal{E}_{0}(m,w) and (3) that

(nαγ)γγn+α(γn+αnα)(12nm(x)(Kαm)(x)𝑑x)γγn+α(CLnm|wm|γ𝑑x)nαγn+αe0,α,M,\displaystyle-\bigg{(}\frac{n-\alpha}{\gamma^{\prime}}\bigg{)}^{\frac{\gamma^{\prime}}{\gamma^{\prime}-n+\alpha}}\bigg{(}\frac{\gamma^{\prime}-n+\alpha}{n-\alpha}\bigg{)}\bigg{(}\frac{1}{2}\int_{\mathbb{R}^{n}}m(x)(K_{\alpha}*m)(x)\,dx\bigg{)}^{\frac{\gamma^{\prime}}{\gamma^{\prime}-n+\alpha}}\bigg{(}C_{L}\int_{\mathbb{R}^{n}}m\Big{|}\frac{w}{m}\Big{|}^{\gamma^{\prime}}\,dx\bigg{)}^{-\frac{n-\alpha}{\gamma^{\prime}-n+\alpha}}\geq e_{0,\alpha,M},

which yields

(CLnm|wm|γ𝑑x)nαγn+α(12nm(x)(Kαm)(x)𝑑x)γγn+α(e0,α,M)1(nαγ)γγn+α(γn+αnα).\displaystyle\frac{\bigg{(}C_{L}\int_{\mathbb{R}^{n}}m\Big{|}\frac{w}{m}\Big{|}^{\gamma^{\prime}}\,dx\bigg{)}^{\frac{n-\alpha}{\gamma^{\prime}-n+\alpha}}}{\bigg{(}\frac{1}{2}\int_{\mathbb{R}^{n}}m(x)(K_{\alpha}*m)(x)\,dx\bigg{)}^{\frac{\gamma^{\prime}}{\gamma^{\prime}-n+\alpha}}}\geq(-e_{0,\alpha,M})^{-1}\Big{(}\frac{n-\alpha}{\gamma^{\prime}}\Big{)}^{\frac{\gamma^{\prime}}{\gamma^{\prime}-n+\alpha}}\bigg{(}\frac{\gamma^{\prime}-n+\alpha}{n-\alpha}\bigg{)}. (3.17)

Denote

α,M:=nαγ(e0,α,M)nγαγ(γn+αnα)γn+αγ,\mathcal{H}_{\alpha,M}:=\frac{n-\alpha}{\gamma^{\prime}}(-e_{0,\alpha,M})^{\frac{n-\gamma^{\prime}-\alpha}{\gamma^{\prime}}}\bigg{(}\frac{\gamma^{\prime}-n+\alpha}{n-\alpha}\bigg{)}^{\frac{\gamma^{\prime}-n+\alpha}{\gamma^{\prime}}},

then we invoke (3.17) to obtain

Gα(m,w)=(CLnm|wm|γ𝑑x)nαγ(nm𝑑x)2γ+αnγnm(x)(Kαm)(x)𝑑x12α,MM2γ+αnγ,\displaystyle G_{\alpha}(m,w)=\frac{\Big{(}C_{L}\int_{\mathbb{R}^{n}}m\big{|}\frac{w}{m}\big{|}^{\gamma^{\prime}}\,dx\Big{)}^{\frac{n-\alpha}{\gamma^{\prime}}}\Big{(}\int_{\mathbb{R}^{n}}m\,dx\Big{)}^{\frac{2\gamma^{\prime}+\alpha-n}{\gamma^{\prime}}}}{\int_{\mathbb{R}^{n}}m(x)(K_{\alpha}*m)(x)\,dx}\geq\frac{1}{2}\mathcal{H}_{\alpha,M}M^{\frac{2\gamma^{\prime}+\alpha-n}{\gamma^{\prime}}}, (3.18)

where we have used the definition (3.14) and the fact nm𝑑x=M.\int_{\mathbb{R}^{n}}m\,dx=M.

Next, by using the fact that (m¯α,M,w¯α,M)(\bar{m}_{\alpha,M},\bar{w}_{\alpha,M}) is a minimizer of problem (3.8), we apply (3.12) to get

Gα(m¯α,M,w¯α,M)=12α,MM2γ+αnγ.\displaystyle G_{\alpha}(\bar{m}_{\alpha,M},\bar{w}_{\alpha,M})=\frac{1}{2}\mathcal{H}_{\alpha,M}M^{\frac{2\gamma^{\prime}+\alpha-n}{\gamma^{\prime}}}. (3.19)

Combining (3.18) with (3.19), one can conclude that (3.15) is attained by (m¯α,M,w¯α,M).(\bar{m}_{\alpha,M},\bar{w}_{\alpha,M}). Moreover, we have

Γα=Gα(m¯α,M,w¯α,M)=12α,MM2γ+αnγ=(nα)(e0,α,M)nαγγM2γ+αnγ2γ(γn+αnα)γ+αnγ,\displaystyle\Gamma_{\alpha}=G_{\alpha}(\bar{m}_{\alpha,M},\bar{w}_{\alpha,M})=\frac{1}{2}\mathcal{H}_{\alpha,M}M^{\frac{2\gamma^{\prime}+\alpha-n}{\gamma^{\prime}}}=\frac{(n-\alpha)(-e_{0,\alpha,M})^{\frac{n-\alpha-\gamma^{\prime}}{\gamma^{\prime}}}M^{\frac{2\gamma^{\prime}+\alpha-n}{\gamma^{\prime}}}}{2\gamma^{\prime}}\bigg{(}\frac{\gamma^{\prime}-n+\alpha}{n-\alpha}\bigg{)}^{\frac{\gamma^{\prime}+\alpha-n}{\gamma^{\prime}}},

which shows that (3.13) holds and the proof of this lemma is completed. ∎

We can see from Lemma 3.1 that for all M>0M>0, Gagliardo-Nirenberg type inequalities given by (3.1) can be attained under the subcritical mass exponent case α(nγ,n)\alpha\in(n-\gamma^{\prime},n). In addition, invoking (3.12) and (3.13), we obtain that

e0,α,M=(γ+αn2γ+αn)λM,\displaystyle e_{0,\alpha,M}=\bigg{(}\frac{\gamma^{\prime}+\alpha-n}{2\gamma^{\prime}+\alpha-n}\bigg{)}\lambda M, (3.20)

and

λM=Sα,M(2γ+αnnα)(nαγ)γγ+αn,Sα,M:=[M2γ+αnγ2Γα]γγ+αn.\displaystyle\lambda M=-S_{\alpha,M}\bigg{(}\frac{2\gamma^{\prime}+\alpha-n}{n-\alpha}\bigg{)}\bigg{(}\frac{n-\alpha}{\gamma^{\prime}}\bigg{)}^{\frac{\gamma^{\prime}}{\gamma^{\prime}+\alpha-n}},\ \ S_{\alpha,M}:=\Bigg{[}\frac{M^{\frac{2\gamma^{\prime}+\alpha-n}{\gamma^{\prime}}}}{2\Gamma_{\alpha}}\Bigg{]}^{\frac{\gamma^{\prime}}{\gamma^{\prime}+\alpha-n}}. (3.21)

The next lemma will indicate that Γα\Gamma_{\alpha} defined in (3.1) is uniformly bounded as α(nγ)\alpha\searrow(n-\gamma^{\prime}), which is essential for us to investigate the mass critical exponent case and prove Theorem 1.1.

Lemma 3.2.

There are constants C1>0C_{1}>0 and C2>0C_{2}>0 independent of α\alpha such that for all α[nγ,nγ+ϵ)\alpha\in[n-\gamma^{\prime},n-\gamma^{\prime}+\epsilon) with ϵ>0\epsilon>0 small,

0<C1ΓαC2.\displaystyle 0<C_{1}\leq\Gamma_{\alpha}\leq C_{2}. (3.22)
Proof.

We first estimate Γα\Gamma_{\alpha} from above uniformly in α\alpha. By setting m~=e|x|\tilde{m}=e^{-|x|} with w~=m~\tilde{w}=\nabla\tilde{m}, we have (m~,w~)𝒜(\tilde{m},\tilde{w})\in\mathcal{A} for any α(nγ,n)\alpha\in(n-\gamma^{\prime},n) and

ΓαGα(m~,w~)=(CLnm~|w~m~|γ𝑑x)nαγ(nm~𝑑x)2γnαγnm~(x)(Kαm~)(x)𝑑xC2(CL,γ,n)<+,\displaystyle\Gamma_{\alpha}\leq G_{\alpha}(\tilde{m},\tilde{w})=\frac{\Big{(}C_{L}\int_{\mathbb{R}^{n}}\tilde{m}\Big{|}\frac{\tilde{w}}{\tilde{m}}\Big{|}^{\gamma^{\prime}}\,dx\Big{)}^{\frac{n-\alpha}{\gamma^{\prime}}}\Big{(}\int_{\mathbb{R}^{n}}\tilde{m}\,dx\Big{)}^{\frac{2\gamma^{\prime}-n-\alpha}{\gamma^{\prime}}}}{\int_{\mathbb{R}^{n}}\tilde{m}(x)(K_{\alpha}*\tilde{m})(x)\,dx}\leq C_{2}(C_{L},\gamma^{\prime},n)<+\infty, (3.23)

where we have used the following inequality

nm~(x)(Kαm~)(x)𝑑x\displaystyle\int_{\mathbb{R}^{n}}\tilde{m}(x)(K_{\alpha}*\tilde{m})(x)\,dx =nne|x|e|y||xy|nα𝑑x𝑑y2n{|xy|1}e|x|e|y||xy|nα𝑑x𝑑y\displaystyle=\int_{\mathbb{R}^{n}}\int_{\mathbb{R}^{n}}\frac{e^{-|x|}e^{-|y|}}{|x-y|^{n-\alpha}}\,dx\,dy\geq\iint_{\mathbb{R}^{2n}\cap\{|x-y|\leq 1\}}\frac{e^{-|x|}e^{-|y|}}{|x-y|^{n-\alpha}}\,dx\,dy
2n{|xy|1}{|x|14,|y|14}e|x|e|y|𝑑x𝑑y=C~(n).\displaystyle\geq\iint_{\mathbb{R}^{2n}\cap\{|x-y|\leq 1\}\cap\{|x|\leq\frac{1}{4},|y|\leq\frac{1}{4}\}}e^{-|x|}e^{-|y|}\,dx\,dy=\tilde{C}(n).

We next focus on the positive lower bound satisfied by Γα\Gamma_{\alpha} uniformly in α\alpha. To show this, we argue by contradiction and assume

lim infα(nγ)Γα=0.\displaystyle\liminf_{\alpha\searrow(n-\gamma^{\prime})}\Gamma_{\alpha}=0. (3.24)

Lemma 3.1 implies there is a minimizer (mα,wα)𝒜(m_{\alpha},w_{\alpha})\in\mathcal{A} of problem (1.28). Since (1.28) is invariant under the scaling s(tnm(tx),tn+1w(tx))s(t^{n}m(tx),t^{n+1}w(tx)) for any s>0s>0 and t>0t>0, we normalize mαm_{\alpha} to get

nmα𝑑x=nmα2nn+α𝑑x1.\displaystyle\int_{\mathbb{R}^{n}}m_{\alpha}\,dx=\int_{\mathbb{R}^{n}}m_{\alpha}^{\frac{2n}{n+\alpha}}\,dx\equiv 1. (3.25)

By using this equality and the Hardy-Littlewood-Sobolev inequality given in (A.2), we obtain

0<nmα(x)(Kαmα)(x)𝑑xC(n,α)(nmα2nn+α𝑑x)n+αnC(n,α),\displaystyle 0<\int_{\mathbb{R}^{n}}m_{\alpha}(x)(K_{\alpha}*m_{\alpha})(x)\,dx\leq C(n,\alpha)\bigg{(}\int_{\mathbb{R}^{n}}m_{\alpha}^{\frac{2n}{n+\alpha}}\,dx\bigg{)}^{\frac{n+\alpha}{n}}\leq C(n,\alpha), (3.26)

where C(n,α)>0C(n,\alpha)>0 is the best constant. On the other hand, since α(nγ,n)\alpha\in(n-\gamma^{\prime},n), we follow the argument shown in [22, Theorem 4.3] to get

limα(nγ)C(n,α)=C(n,nγ)<+.\displaystyle\lim_{\alpha\searrow(n-\gamma^{\prime})}C(n,\alpha)=C(n,n-\gamma^{\prime})<+\infty. (3.27)

Hence,

lim infα(nγ)nmα(x)(Kαmα)(x)dx=:C~(n,γ)C(n,nγ)<+.\displaystyle\liminf_{\alpha\searrow(n-\gamma^{\prime})}\int_{\mathbb{R}^{n}}m_{\alpha}(x)(K_{\alpha}*m_{\alpha})(x)\,dx=:\tilde{C}(n,\gamma^{\prime})\leq C(n,n-\gamma^{\prime})<+\infty. (3.28)

Then it follows from (3.24), (3.28) and (1.28) that, as α(nγ),\alpha\searrow(n-\gamma^{\prime}),

nmα|wαmα|γ𝑑x0.\displaystyle\int_{\mathbb{R}^{n}}m_{\alpha}\bigg{|}\frac{w_{\alpha}}{m_{\alpha}}\bigg{|}^{\gamma^{\prime}}\,dx\rightarrow 0. (3.29)

Proceeding the same argument shown in Lemma 3.5 of [14], one finds mαW1,q^(n)0.\|m_{\alpha}\|_{W^{1,\hat{q}}(\mathbb{R}^{n})}\rightarrow 0. By using the Sobolev embedding theorem, we obtain

mαLnnγ(n)0, as α(nγ).\displaystyle\|m_{\alpha}\|_{L^{\frac{n}{n-\gamma^{\prime}}}(\mathbb{R}^{n})}\rightarrow 0,\text{ as }\alpha\searrow(n-\gamma^{\prime}). (3.30)

On the other hand, the following interpolation inequality holds:

mαL2nn+α(n)mαL1(n)1θαmαLnnγ(n)θα,\|m_{\alpha}\|_{L^{\frac{2n}{n+\alpha}}(\mathbb{R}^{n})}\leq\|m_{\alpha}\|^{1-\theta_{\alpha}}_{L^{1}(\mathbb{R}^{n})}\|m_{\alpha}\|^{\theta_{\alpha}}_{L^{\frac{n}{n-\gamma^{\prime}}}(\mathbb{R}^{n})},

where θα:=nαγ(0,1).\theta_{\alpha}:=\frac{n-\alpha}{\gamma^{\prime}}\in(0,1). With the help of (3.30), we further get as α(nγ),\alpha\searrow(n-\gamma^{\prime}), θα1\theta_{\alpha}\nearrow 1 and

mαL2nn+α(n)0,\|m_{\alpha}\|_{L^{\frac{2n}{n+\alpha}}(\mathbb{R}^{n})}\rightarrow 0,

which reaches a contradiction to (3.25). Thus, we have C1>0\exists C_{1}>0 independent of α\alpha such that

0<C1Γα.\displaystyle 0<C_{1}\leq\Gamma_{\alpha}. (3.31)

Finally, combining (3.31) with (3.23), one finds (3.22) holds. This completes the proof of the lemma. ∎

With the aid of the uniform boundedness of Γα\Gamma_{\alpha}, we next establish the uniform LL^{\infty} bound of mαm_{\alpha} as α(nγ)\alpha\searrow(n-\gamma^{\prime}), which is

Lemma 3.3.

Let (uα,mα,λα)C2(n)×W1,p(n)×(u_{\alpha},m_{\alpha},\lambda_{\alpha})\in C^{2}(\mathbb{R}^{n})\times W^{1,p}(\mathbb{R}^{n})\times\mathbb{R}, p>1\forall p>1 be the solution of

{Δu+CH|u|γ+λ=Kαm,xn,ΔmCHγ(m|u|γ2u)=0,xn,nm𝑑x=Mα.\displaystyle\left\{\begin{array}[]{ll}-\Delta u+C_{H}|\nabla u|^{\gamma}+\lambda=-K_{\alpha}*m,&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},\\ \int_{\mathbb{R}^{n}}m\,dx=M_{\alpha}.\end{array}\right. (3.35)

Define wα=CHγmα|uα|γ2uαw_{\alpha}=-C_{H}\gamma m_{\alpha}|\nabla u_{\alpha}|^{\gamma-2}\nabla u_{\alpha}. Assume that each uαu_{\alpha} is bounded from below and there exists a constant C>0C>0 independent of α\alpha such that

lim supα(nγ)nmα|uα|γ𝑑xC,limα(nγ)nmα𝑑x=limα(nγ)MαC,lim supα(nγ)|λα|C,\displaystyle\limsup_{\alpha\searrow(n-\gamma^{\prime})}\int_{\mathbb{R}^{n}}m_{\alpha}|\nabla u_{\alpha}|^{\gamma}\,dx\leq C,\ \ \lim_{\alpha\searrow(n-\gamma^{\prime})}\int_{\mathbb{R}^{n}}m_{\alpha}\,dx=\lim_{\alpha\searrow(n-\gamma^{\prime})}M_{\alpha}\leq C,\ \ \limsup_{\alpha\searrow(n-\gamma^{\prime})}|\lambda_{\alpha}|\leq C, (3.36)

then there is C1>0C_{1}>0 independent of α\alpha such that

lim supα(nγ)mαL(n)C1.\displaystyle\limsup_{\alpha\searrow(n-\gamma^{\prime})}\|m_{\alpha}\|_{L^{\infty}(\mathbb{R}^{n})}\leq C_{1}. (3.37)
Proof.

The proof is similar as shown in [14, Lemma 4.3]. We proceed by contradiction and suppose that up to a subsequence,

μα:=mαL(n)1n0 as α(nγ).\displaystyle\mu_{\alpha}:=\|m_{\alpha}\|_{L^{\infty}(\mathbb{R}^{n})}^{-\frac{1}{n}}\rightarrow 0\ \ \text{~{}as~{}}\alpha\searrow(n-\gamma^{\prime}). (3.38)

Now, we fix 0=uα(0)=infxnuα(x)0=u_{\alpha}(0)=\inf\limits_{x\in\mathbb{R}^{n}}u_{\alpha}(x) without loss of generality, as this is due to the fact that uαu_{\alpha} is bounded from below. Define

u¯α:=μα2γγ1uα(μαx)+1,m¯α:=μαnmα(μαx) and w¯α:=μαn+1wα(μαx),\displaystyle\bar{u}_{\alpha}:=\mu_{\alpha}^{\frac{2-\gamma}{\gamma-1}}u_{\alpha}(\mu_{\alpha}x)+1,\ \bar{m}_{\alpha}:=\mu_{\alpha}^{n}m_{\alpha}(\mu_{\alpha}x)~{}\text{ and }~{}\bar{w}_{\alpha}:=\mu_{\alpha}^{n+1}w_{\alpha}(\mu_{\alpha}x), (3.39)

then, by (3.36) and (3.38), we obtain that up to a subsequence

nm¯α𝑑x=nmα𝑑x=Mα,nm¯α2nn+α𝑑x=μαn(nα)n+αnmα2nn+α𝑑x0as α(nγ),\displaystyle\int_{\mathbb{R}^{n}}\bar{m}_{\alpha}\,dx=\int_{\mathbb{R}^{n}}m_{\alpha}\,dx=M_{\alpha},\ \int_{\mathbb{R}^{n}}\bar{m}_{\alpha}^{\frac{2n}{n+\alpha}}\,dx=\mu_{\alpha}^{\frac{n(n-\alpha)}{n+\alpha}}\int_{\mathbb{R}^{n}}m_{\alpha}^{\frac{2n}{n+\alpha}}\,dx\rightarrow 0\ \ \text{as~{}}\alpha\searrow(n-\gamma^{\prime}), (3.40)

and

nm¯α|u¯α|γ𝑑x=μαγnmα|uα|γ𝑑x0as α(nγ).\displaystyle\int_{\mathbb{R}^{n}}\bar{m}_{\alpha}|\nabla\bar{u}_{\alpha}|^{\gamma}\,dx=\mu_{\alpha}^{\gamma^{\prime}}\int_{\mathbb{R}^{n}}m_{\alpha}|\nabla u_{\alpha}|^{\gamma}\,dx\rightarrow 0\ \ \text{as~{}}\alpha\searrow(n-\gamma^{\prime}). (3.41)

Recall the definition of wαw_{\alpha}, then we deduce from (1.7) and (3.39) that

CLn|w¯αm¯α|γm¯α𝑑x=(γ1)CHnm¯α|u¯α|γ𝑑x0,as α(nγ).\displaystyle C_{L}\int_{\mathbb{R}^{n}}\bigg{|}\frac{\bar{w}_{\alpha}}{\bar{m}_{\alpha}}\bigg{|}^{\gamma^{\prime}}\bar{m}_{\alpha}\,dx=(\gamma-1)C_{H}\int_{\mathbb{R}^{n}}\bar{m}_{\alpha}|\nabla\bar{u}_{\alpha}|^{\gamma}\,dx\rightarrow 0,\ \ \text{as~{}}\alpha\searrow(n-\gamma^{\prime}). (3.42)

In light of (3.38) and (3.39), we infer that

m¯αL1.\displaystyle\|\bar{m}_{\alpha}\|_{L^{\infty}}\equiv 1. (3.43)

This together with (3.40) implies that for any q>2nn+αq>\frac{2n}{n+\alpha},

nm¯αq𝑑x(nm¯α2nn+α𝑑x)m¯αL(n)q2nn+α0asα(nγ).\displaystyle\int_{\mathbb{R}^{n}}\bar{m}_{\alpha}^{q}\,dx\leq\bigg{(}\int_{\mathbb{R}^{n}}\bar{m}_{\alpha}^{\frac{2n}{n+\alpha}}\,dx\bigg{)}\|\bar{m}_{\alpha}\|_{L^{\infty}(\mathbb{{\mathbb{R}}}^{n})}^{q-\frac{2n}{n+\alpha}}\rightarrow 0~{}~{}\text{as}~{}~{}\alpha\searrow(n-\gamma^{\prime}). (3.44)

On the other hand, invoking (3.39) and (3.35), one can obtain that

{Δxu¯α+CH|xu¯α|γ+λαμαγ=μα(γn+α)Kαm¯α,xn,Δxm¯αCHγx(m¯α|xu¯α|γ2xu¯α)=0,xn,nm¯α𝑑x=Mα.\displaystyle\left\{\begin{array}[]{ll}-\Delta_{x}\bar{u}_{\alpha}+C_{H}|\nabla_{x}\bar{u}_{\alpha}|^{\gamma}+\lambda_{\alpha}\mu_{\alpha}^{\gamma^{\prime}}=-\mu_{\alpha}^{(\gamma^{\prime}-n+\alpha)}K_{\alpha}\ast\bar{m}_{\alpha},&x\in\mathbb{R}^{n},\\ -\Delta_{x}\bar{m}_{\alpha}-C_{H}\gamma\nabla_{x}\cdot(\bar{m}_{\alpha}|\nabla_{x}\bar{u}_{\alpha}|^{\gamma-2}\nabla_{x}\bar{u}_{\alpha})=0,&x\in\mathbb{R}^{n},\\ \int_{\mathbb{R}^{n}}\bar{m}_{\alpha}\,dx=M_{\alpha}.\end{array}\right. (3.48)

It follows from (3.36) and (3.38) that λαμαγ0as α(nγ).\lambda_{\alpha}\mu_{\alpha}^{\gamma^{\prime}}\rightarrow 0\ \ \text{as~{}}\alpha\searrow(n-\gamma^{\prime}). In addition, we claim that

0μαγnαKαm¯αL(n)C,\displaystyle 0\leq\mu_{\alpha}^{\gamma^{\prime}-n-\alpha}\|K_{\alpha}*\bar{m}_{\alpha}\|_{L^{\infty}(\mathbb{{\mathbb{R}}}^{n})}\leq C, (3.49)

where C>0C>0 independent of α\alpha. Indeed, by the definition of KαK_{\alpha}, we get

|Kαm¯α|n|m¯α(xy)||y|nα𝑑y=B1|m¯α(xy)||y|nα𝑑yI+B1c|m¯α(xy)||y|nα𝑑yII.\displaystyle|K_{\alpha}*\bar{m}_{\alpha}|\leq\int_{\mathbb{R}^{n}}\frac{|\bar{m}_{\alpha}(x-y)|}{|y|^{n-\alpha}}\,dy=\overbrace{\int_{B_{1}}\frac{|\bar{m}_{\alpha}(x-y)|}{|y|^{n-\alpha}}\,dy}^{I}+\overbrace{\int_{B^{c}_{1}}\frac{|\bar{m}_{\alpha}(x-y)|}{|y|^{n-\alpha}}\,dy}^{II}. (3.50)

For II, we apply (3.43) to get

I=B1|m¯α(xy)||y|nαdym¯αL(n)B11|y|nαdy=:C(n).\displaystyle I=\int_{B_{1}}\frac{|\bar{m}_{\alpha}(x-y)|}{|y|^{n-\alpha}}\,dy\leq\|\bar{m}_{\alpha}\|_{L^{\infty}}(\mathbb{{\mathbb{R}}}^{n})\int_{B_{1}}\frac{1}{|y|^{n-\alpha}}\,dy=:C(n). (3.51)

For IIII, taking into account the condition (3.36), we get from (3.40) that

II=B1c|m¯α(xy)||y|nα𝑑yB1c|m¯α(xy)|𝑑ym¯αL1(n)=MαC.\displaystyle II=\int_{B^{c}_{1}}\frac{|\bar{m}_{\alpha}(x-y)|}{|y|^{n-\alpha}}\,dy\leq\int_{B^{c}_{1}}|\bar{m}_{\alpha}(x-y)|\,dy\leq\|\bar{m}_{\alpha}\|_{L^{1}(\mathbb{{\mathbb{R}}}^{n})}=M_{\alpha}\leq C. (3.52)

Then, we conclude that (3.49) holds by collecting (3.38), (3.50)-(3.52) and the fact γnα\gamma^{\prime}\geq n-\alpha. Hence, applying Lemma 2.2 to the first equation in (3.48), one finds

lim supα(nγ)u¯αL(n)C<.\displaystyle\limsup_{\alpha\searrow(n-\gamma^{\prime})}\|\nabla\bar{u}_{\alpha}\|_{L^{\infty}(\mathbb{{\mathbb{R}}}^{n})}\leq C<\infty. (3.53)

Noting that w¯α=CHγm¯α|u¯α|γ2u¯α\bar{w}_{\alpha}=-C_{H}\gamma\bar{m}_{\alpha}|\nabla\bar{u}_{\alpha}|^{\gamma-2}\nabla\bar{u}_{\alpha}, we deduce from (3.53) that

lim supα(nγ)w¯αL(n)C<.\displaystyle\limsup_{\alpha\searrow(n-\gamma^{\prime})}\|\bar{w}_{\alpha}\|_{L^{\infty}(\mathbb{{\mathbb{R}}}^{n})}\leq C<\infty. (3.54)

Now, we turn our attention to Hölder estimates of m¯α\bar{m}_{\alpha} and the proof of Hölder continuity of m¯α\bar{m}_{\alpha} is the same as shown in Lemma 4.3 of [14]. In fact, we obtain for some θ(0,1),\theta^{\prime}\in(0,1),

m¯αC0,θ(n)0as α(nγ).\displaystyle\|\bar{m}_{\alpha}\|_{C^{0,\theta^{\prime}}(\mathbb{R}^{n})}\rightarrow 0\ \ \text{as~{}}\alpha\searrow(n-\gamma^{\prime}). (3.55)

Assume that xαx_{\alpha} is a maximum point of m¯α\bar{m}_{\alpha}, i.e., m¯α(xα)=m¯αL(n)=1.\bar{m}_{\alpha}(x_{\alpha})=\|\bar{m}_{\alpha}\|_{L^{\infty}(\mathbb{R}^{n})}=1. Then we deduce from (3.55) that there exists R1>0R_{1}>0 independent of α\alpha such that |m¯α(x)|12,xBR1(xα)|\bar{m}_{\alpha}(x)|\geq\frac{1}{2},\forall x\in B_{R_{1}}(x_{\alpha}). Thus,

(12)2n2nγ|BR1|BR1(xα)m¯α2n2nγ𝑑xnm¯α2n2nγ𝑑x,\displaystyle\bigg{(}\frac{1}{2}\bigg{)}^{\frac{2n}{2n-\gamma^{\prime}}}|B_{R_{1}}|\leq\int_{B_{R_{1}}(x_{\alpha})}\bar{m}_{\alpha}^{\frac{2n}{2n-\gamma^{\prime}}}\,dx\leq\int_{\mathbb{R}^{n}}\bar{m}_{\alpha}^{\frac{2n}{2n-\gamma^{\prime}}}\,dx,

which contradicts (3.44), and the proof of the lemma is finished. ∎

With the aid of Lemma 3.1, Lemma 3.2 and Lemma 3.3, we are able to show conclusions stated in Theorem 1.1, which are

Proof of Theorem 1.1:

Proof.

We first recall that, for any M>0M>0 and p(1,+)p\in(1,+\infty), (u¯α,M,m¯α,M,λα,M)C2(n)×W1,p(n)×(\bar{u}_{\alpha,M},\bar{m}_{\alpha,M},\lambda_{\alpha,M})\in C^{2}(\mathbb{R}^{n})\times W^{1,p}(\mathbb{R}^{n})\times\mathbb{R} denotes the solution to system (3.6), and the pair (m¯α,M,w¯α,M)(\bar{m}_{\alpha,M},\bar{w}_{\alpha,M}) with w¯α,M=CHγm¯α,M|u¯α,M|γ2u¯α,M\bar{w}_{\alpha,M}=-C_{H}\gamma\bar{m}_{\alpha,M}|\nabla\bar{u}_{\alpha,M}|^{\gamma-2}\nabla\bar{u}_{\alpha,M} is a minimizer of the minimization problem (3.15), in which m¯α,M\bar{m}_{\alpha,M} satisfies the estimate (3.7). Now, we take

M=Mα:=eγ+αn2γ+αn[2Γα]γ2γ+αnM=M_{\alpha}:=e^{\frac{\gamma^{\prime}+\alpha-n}{2\gamma^{\prime}+\alpha-n}}\big{[}2\Gamma_{\alpha}\big{]}^{\frac{\gamma^{\prime}}{2\gamma^{\prime}+\alpha-n}}

in (3.6), then one can deduce from (3.21) that

Sα,Mα:=[Mα2γ+αnγ2Γα]γγ+αne.S_{\alpha,M_{\alpha}}:=\Bigg{[}\frac{M_{\alpha}^{\frac{2\gamma^{\prime}+\alpha-n}{\gamma^{\prime}}}}{2\Gamma_{\alpha}}\Bigg{]}^{\frac{\gamma^{\prime}}{\gamma^{\prime}+\alpha-n}}\equiv e.

Moreover, we obtain that, up to a subsequence,

Mα2γ+αnγ2Γα1,2γ+αnγ1,asα(nγ).\displaystyle\frac{M_{\alpha}^{\frac{2\gamma^{\prime}+\alpha-n}{\gamma^{\prime}}}}{2\Gamma_{\alpha}}\rightarrow 1,\ \ \frac{2\gamma^{\prime}+\alpha-n}{\gamma^{\prime}}\rightarrow 1,\ \ \text{as}\ \ \alpha\searrow(n-\gamma^{\prime}). (3.56)

Since M=MαM=M_{\alpha} depends on α\alpha, to emphasize the dependence of a solution on α\alpha, we will rewrite (m¯α,M,w¯α,M,λα,M)(\bar{m}_{\alpha,M},\bar{w}_{\alpha,M},\lambda_{\alpha,M}) as (m¯α,Mα,w¯α,Mα,λα,Mα)(\bar{m}_{\alpha,M_{\alpha}},\bar{w}_{\alpha,M_{\alpha}},\lambda_{\alpha,M_{\alpha}}). Hence, we know from (3.6) that (m¯α,Mα,u¯α,Mα,w¯α,Mα,λα,Mα)(\bar{m}_{\alpha,M_{\alpha}},\bar{u}_{\alpha,M_{\alpha}},\bar{w}_{\alpha,M_{\alpha}},\lambda_{\alpha,M_{\alpha}}) satisfies

{Δu+CH|u|γ+λ=Kαm,xn,Δm+CHγ(m|u|γ2u)=0,xn,w=CHγm|u|γ2u,nm𝑑x=Mα.\displaystyle\left\{\begin{array}[]{ll}-\Delta u+C_{H}|\nabla u|^{\gamma}+\lambda=-K_{\alpha}*m,&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},\\ w=-C_{H}\gamma m|\nabla u|^{\gamma-2}\nabla u,\ \int_{\mathbb{R}^{n}}m\,dx=M_{\alpha}.\end{array}\right. (3.60)

We can infer from Lemma 3.2 that, up to a subsequence,

ΓαΓ¯α:=lim infα(nr)Γα>0 asαα.\displaystyle\Gamma_{\alpha}\rightarrow\bar{\Gamma}_{\alpha^{*}}:=\liminf\limits_{\alpha\searrow(n-r)}\Gamma_{\alpha}>0\ \ \text{ as}\ \ \alpha\searrow\alpha^{*}.

In addition, invoking (3.56) we have that MαMα:=MM_{\alpha}\rightarrow M_{\alpha^{*}}:=M^{*}  as  α(nγ)\alpha\searrow(n-\gamma^{\prime}), where

M=2Γ¯α,α=nγ.\displaystyle M^{*}=2\bar{\Gamma}_{\alpha^{*}},\ \alpha^{*}=n-\gamma^{\prime}. (3.61)

Moreover, due to the relation (3.21), we obtain that, up to a subsequence,

λα,Mαλα:=1Masα(nγ)\displaystyle\lambda_{\alpha,M_{\alpha}}\rightarrow\lambda_{\alpha^{*}}:=-\frac{1}{M^{*}}\ as\ \alpha\searrow(n-\gamma^{\prime}) (3.62)

and it follows from (3.12) that

nm¯α,Mα𝑑x=MαM>0,nm¯α,Mα(x)(Kαm¯α,Mα)(x)𝑑x2,CLnm¯α,Mα|w¯α,Mαm¯α,Mα|γ𝑑x1.\displaystyle\int_{\mathbb{R}^{n}}{\bar{m}}_{\alpha,M_{\alpha}}\,dx=M_{\alpha}\rightarrow M^{*}>0,\ \ \int_{\mathbb{R}^{n}}{\bar{m}}_{\alpha,M_{\alpha}}(x)(K_{\alpha}\ast{\bar{m}}_{\alpha,M_{\alpha}})(x)\,dx\rightarrow 2,\ \ C_{L}\int_{\mathbb{R}^{n}}\bar{m}_{\alpha,M_{\alpha}}\bigg{|}\frac{\bar{w}_{\alpha,M_{\alpha}}}{\bar{m}_{\alpha,M_{\alpha}}}\bigg{|}^{\gamma^{\prime}}\,dx\rightarrow 1. (3.63)

Applying Lemma 3.3, we derive from (3.62) and (3.63) that

lim supα(nγ)m¯α,MαL(n)<.\limsup_{\alpha\searrow(n-\gamma^{\prime})}\|\bar{m}_{\alpha,M_{\alpha}}\|_{L^{\infty}(\mathbb{R}^{n})}<\infty. (3.64)

Then, by using the estimate (2.6) with b=0b=0 from Lemma 2.3, we have

lim supα(nγ)u¯α,MαL<,\displaystyle\limsup_{\alpha\searrow(n-\gamma^{\prime})}\|\nabla\bar{u}_{\alpha,M_{\alpha}}\|_{L^{\infty}}<\infty, (3.65)

which, together with the definition of w¯α,Mα\bar{w}_{\alpha,M_{\alpha}}, yields

lim supα(nγ)w¯α,MαL<.\displaystyle\limsup_{\alpha\searrow(n-\gamma^{\prime})}\|\bar{w}_{\alpha,M_{\alpha}}\|_{L^{\infty}}<\infty. (3.66)

Proceeding the arguments similar as those used in the proof of (3.55), we collect (3.63)-(3.66) to obtain that

lim supα(nγ)m¯α,MαW1,q(n)<+,q>n,\displaystyle\limsup_{\alpha\searrow(n-\gamma^{\prime})}\|\bar{m}_{\alpha,M_{\alpha}}\|_{W^{1,q}(\mathbb{R}^{n})}<+\infty,\ \forall\ q>n, (3.67)

and thus

lim supα(nγ)m¯α,MαC0,θ~(n)< for someθ~(0,1).\displaystyle\limsup_{\alpha\searrow(n-\gamma^{\prime})}\|\bar{m}_{\alpha,M_{\alpha}}\|_{C^{0,\tilde{\theta}}(\mathbb{R}^{n})}<\infty~{}\text{ for some}~{}\tilde{\theta}\in(0,1). (3.68)

We may assume that u¯α,Mα(0)=0=infxnu¯α,Mα(x)\bar{u}_{\alpha,M_{\alpha}}(0)=0=\inf_{x\in\mathbb{R}^{n}}\bar{u}_{\alpha,M_{\alpha}}(x) due to the fact that u¯α,MαC2(n)\bar{u}_{\alpha,M_{\alpha}}\in C^{2}(\mathbb{R}^{n}) is bounded from below. Hence, by the first equation of (3.60), one has

(Kαm¯α,Mα)(0)λα,Mα>0,(K_{\alpha}*\bar{m}_{\alpha,M_{\alpha}})(0)\geq-\lambda_{\alpha,M_{\alpha}}>0,

which together with (3.64) and a similar argument as used in [5, Lemma 4.1] to obtain that there are δ1\delta_{1} and a large R>0R>0 independent of α\alpha such that,

|x|Rm¯α,Mα(x)𝑑x>δ12>0.\int_{|x|\leq R}{\bar{m}_{\alpha,M_{\alpha}}(x)}\,dx>\frac{\delta_{1}}{2}>0. (3.69)

Now, we rewrite the first equation of (3.60) as

Δu¯α,Mα=CH|u¯α,Mα|r+hα(x),-\Delta\bar{u}_{\alpha,M_{\alpha}}=-C_{H}|\nabla\bar{u}_{\alpha,M_{\alpha}}|^{r^{\prime}}+h_{\alpha}(x), (3.70)

where hα(x):=λα,MαKαm¯α,Mα,xnh_{\alpha}(x):=-\lambda_{\alpha,M_{\alpha}}-K_{\alpha}*\bar{m}_{\alpha,M_{\alpha}},\ x\in\mathbb{R}^{n}. By performing the same procedure shown in (3.50), one can see that

|(Kαmα,Mα)(x)|C\big{|}(K_{\alpha}*m_{\alpha,M_{\alpha}})(x)\big{|}\leq C

for some C>0C>0 independent of α\alpha. Then, we apply the standard elliptic regularity to (3.70) and obtain

u¯α,MαC2,θ(BR(0))Cθ,R¯< for some θ(0,1),\|\bar{u}_{\alpha,M_{\alpha}}\|_{C^{2,\theta}(B_{R}(0))}\leq C_{\theta,\bar{R}}<\infty\ \text{ for some $\theta\in(0,1)$,} (3.71)

where 0<R<R¯.0<R<\bar{R}. Performing the standard diagonal procedure, we take the limit and apply Arzelà-Ascoli theorem, (3.67) and (3.71) to obtain that there exists (mα,uα)W1,p(n)×C2(n)(m_{\alpha^{*}},u_{\alpha^{*}})\in W^{1,p}(\mathbb{R}^{n})\times C^{2}(\mathbb{R}^{n}) such that

m¯α,Mαmα in W1,p(n), and u¯α,Mαuαin Cloc2(n), as α(nγ).\displaystyle\bar{m}_{\alpha,M_{\alpha}}\rightharpoonup m_{\alpha^{*}}\text{ in }W^{1,p}(\mathbb{R}^{n}),\text{ and }\bar{u}_{\alpha,M_{\alpha}}\rightarrow u_{\alpha^{*}}\ \ \text{in }C^{2}_{\rm loc}(\mathbb{R}^{n}),\text{ as }\alpha\searrow(n-\gamma^{\prime}). (3.72)

Combining (3.60), (3.62) and (3.72), we conclude that (mα,uα)W1,p(n)×C2(n)(m_{\alpha^{*}},u_{\alpha^{*}})\in W^{1,p}(\mathbb{R}^{n})\times C^{2}(\mathbb{R}^{n}) satisfies

{Δu+CH|u|γ1M=Kαm,xn,ΔmγCH(m|u|γ2u)=0,xn,w=CHγm|u|γ2u.\displaystyle\left\{\begin{array}[]{ll}-\Delta u+C_{H}|\nabla u|^{\gamma}-\frac{1}{M^{*}}=-K_{\alpha^{*}}*m,&x\in\mathbb{R}^{n},\\ -\Delta m-\gamma C_{H}\nabla\cdot(m|\nabla u|^{\gamma-2}\nabla u)=0,&x\in\mathbb{R}^{n},\\ w=-C_{H}\gamma m|\nabla u|^{\gamma-2}\nabla u.\end{array}\right. (3.76)

In light of (3.69) and Fatou’s lemma, we have

nmα𝑑x=M~(0,M].\int_{\mathbb{R}^{n}}m_{\alpha^{*}}\,dx=\tilde{M}\in(0,M^{*}]. (3.77)

Moreover, by Lemma 2.7, we obtain that there exists some κ,C>0\kappa,C>0 such that mα(x)<Ceκ|x|m_{\alpha^{*}}(x)<Ce^{-\kappa|x|}. In addition, by using (3.65), we get uαL<\|\nabla u_{\alpha^{*}}\|_{L^{\infty}}<\infty. It then follows from Lemma 2.8 that

CLn|wαmα|γmα𝑑x=12nmα(x)(Kαmα)(x)𝑑x.\displaystyle C_{L}\int_{\mathbb{R}^{n}}\Bigg{|}\frac{w_{\alpha^{*}}}{m_{\alpha^{*}}}\Bigg{|}^{\gamma^{\prime}}m_{\alpha^{*}}\,dx=\frac{1}{2}\int_{\mathbb{R}^{n}}m_{\alpha^{*}}(x)(K_{\alpha^{*}}*m_{\alpha^{*}})(x)\,dx. (3.78)

Next, we discuss the relationship between Γ¯α:=lim infα(nγ)Γα\bar{\Gamma}_{\alpha^{*}}:=\liminf\limits_{\alpha\searrow(n-\gamma^{\prime})}\Gamma_{\alpha} and Γα\Gamma_{\alpha^{*}} with α=(nγ).\alpha^{*}=(n-\gamma^{\prime}). We claim that

Γ¯α=Γ(nγ).\displaystyle\bar{\Gamma}_{\alpha^{*}}=\Gamma_{(n-\gamma^{\prime})}. (3.79)

Indeed, we first utilize Lemma 3.1 and obtain

Γα=Gα(m¯α,Mα,w¯α,Mα)\displaystyle\Gamma_{\alpha}=G_{\alpha}(\bar{m}_{\alpha,M_{\alpha}},\bar{w}_{\alpha,M_{\alpha}}) (3.80)
=G(nγ)(m¯α,Mα,w¯α,Mα)(CLnm¯α,Mα|w¯α,Mαm¯α,Mα|γ𝑑x)nαγ(nm¯α,Mα𝑑x)2γ+αnγ(CLnm¯α,Mα|w¯α,Mαm¯α,Mα|γ𝑑x)(nm¯α,Mα𝑑x)nm¯α,Mα(x)(K(nγ)m¯α,Mα)𝑑xnm¯α,Mα(x)(Kαm¯α,Mα)𝑑x\displaystyle=G_{(n-\gamma^{\prime})}(\bar{m}_{\alpha,M_{\alpha}},\bar{w}_{\alpha,M_{\alpha}})\frac{\Big{(}C_{L}\int_{\mathbb{R}^{n}}\bar{m}_{\alpha,M_{\alpha}}\Big{|}\frac{\bar{w}_{\alpha,M_{\alpha}}}{\bar{m}_{\alpha,M_{\alpha}}}\Big{|}^{\gamma^{\prime}}\,dx\Big{)}^{\frac{n-\alpha}{\gamma^{\prime}}}\Big{(}\int_{\mathbb{R}^{n}}\bar{m}_{\alpha,M_{\alpha}}\,dx\Big{)}^{\frac{2\gamma^{\prime}+\alpha-n}{\gamma^{\prime}}}}{\Big{(}C_{L}\int_{\mathbb{R}^{n}}\bar{m}_{\alpha,M_{\alpha}}\Big{|}\frac{\bar{w}_{\alpha,M_{\alpha}}}{\bar{m}_{\alpha,M_{\alpha}}}\Big{|}^{\gamma^{\prime}}\,dx\Big{)}\Big{(}\int_{\mathbb{R}^{n}}\bar{m}_{\alpha,M_{\alpha}}\,dx\Big{)}}\cdot\frac{\int_{\mathbb{R}^{n}}\bar{m}_{\alpha,M_{\alpha}}(x)(K_{(n-\gamma^{\prime})}*\bar{m}_{\alpha,M_{\alpha}})\,dx}{{\int_{\mathbb{R}^{n}}\bar{m}_{\alpha,M_{\alpha}}(x)(K_{\alpha}*\bar{m}_{\alpha,M_{\alpha}}})\,dx}
Γ(nγ)(CLnm¯α,Mα|w¯α,Mαm¯α,Mα|γ𝑑x)nαγ(nm¯α,Mα𝑑x)2γ+αnγ(CLnm¯α,Mα|w¯α,Mαm¯α,Mα|γ𝑑x)(nm¯α,Mα𝑑x)nm¯α,Mα(x)(K(nγ)m¯α,Mα)𝑑xnm¯α,Mα(x)(Kαm¯α,Mα)𝑑x.\displaystyle\geq\Gamma_{(n-\gamma^{\prime})}\frac{\Big{(}C_{L}\int_{\mathbb{R}^{n}}\bar{m}_{\alpha,M_{\alpha}}\Big{|}\frac{\bar{w}_{\alpha,M_{\alpha}}}{\bar{m}_{\alpha,M_{\alpha}}}\Big{|}^{\gamma^{\prime}}\,dx\Big{)}^{\frac{n-\alpha}{\gamma^{\prime}}}\Big{(}\int_{\mathbb{R}^{n}}\bar{m}_{\alpha,M_{\alpha}}\,dx\Big{)}^{\frac{2\gamma^{\prime}+\alpha-n}{\gamma^{\prime}}}}{\Big{(}C_{L}\int_{\mathbb{R}^{n}}\bar{m}_{\alpha,M_{\alpha}}\Big{|}\frac{\bar{w}_{\alpha,M_{\alpha}}}{\bar{m}_{\alpha,M_{\alpha}}}\Big{|}^{\gamma^{\prime}}\,dx\Big{)}\Big{(}\int_{\mathbb{R}^{n}}\bar{m}_{\alpha,M_{\alpha}}\,dx\Big{)}}\cdot\frac{\int_{\mathbb{R}^{n}}\bar{m}_{\alpha,M_{\alpha}}(x)(K_{(n-\gamma^{\prime})}*\bar{m}_{\alpha,M_{\alpha}})\,dx}{{\int_{\mathbb{R}^{n}}\bar{m}_{\alpha,M_{\alpha}}(x)(K_{\alpha}*\bar{m}_{\alpha,M_{\alpha}}})\,dx}. (3.81)

Then, we derive from (3.63) that, as α(nγ)\alpha\searrow(n-\gamma^{\prime}),

(CLnm¯α,Mα|w¯α,Mαm¯α,Mα|γ𝑑x)nαγ(nm¯α,Mα𝑑x)2γ+αnγ(CLnm¯α,Mα|w¯α,Mαm¯α,Mα|γ𝑑x)(nm¯α,Mα𝑑x)nm¯α,Mα(x)(K(nγ)m¯α,Mα)𝑑xnm¯α,Mα(x)(Kαm¯α,Mα)𝑑x1.\displaystyle\frac{\Big{(}C_{L}\int_{\mathbb{R}^{n}}\bar{m}_{\alpha,M_{\alpha}}\Big{|}\frac{\bar{w}_{\alpha,M_{\alpha}}}{\bar{m}_{\alpha,M_{\alpha}}}\Big{|}^{\gamma^{\prime}}\,dx\Big{)}^{\frac{n-\alpha}{\gamma^{\prime}}}\Big{(}\int_{\mathbb{R}^{n}}\bar{m}_{\alpha,M_{\alpha}}\,dx\Big{)}^{\frac{2\gamma^{\prime}+\alpha-n}{\gamma^{\prime}}}}{\Big{(}C_{L}\int_{\mathbb{R}^{n}}\bar{m}_{\alpha,M_{\alpha}}\Big{|}\frac{\bar{w}_{\alpha,M_{\alpha}}}{\bar{m}_{\alpha,M_{\alpha}}}\Big{|}^{\gamma^{\prime}}\,dx\Big{)}\Big{(}\int_{\mathbb{R}^{n}}\bar{m}_{\alpha,M_{\alpha}}\,dx\Big{)}}\cdot\frac{\int_{\mathbb{R}^{n}}\bar{m}_{\alpha,M_{\alpha}}(x)(K_{(n-\gamma^{\prime})}*\bar{m}_{\alpha,M_{\alpha}})\,dx}{{\int_{\mathbb{R}^{n}}\bar{m}_{\alpha,M_{\alpha}}(x)(K_{\alpha}*\bar{m}_{\alpha,M_{\alpha}}})\,dx}\rightarrow 1.

Moreover, one takes the limit in (3.80) to get

Γ¯α:=lim infα(nγ)ΓαΓ(nγ).\displaystyle\bar{\Gamma}_{\alpha^{*}}:=\liminf_{\alpha\searrow(n-\gamma^{\prime})}\Gamma_{\alpha}\geq\Gamma_{(n-\gamma^{\prime})}. (3.82)

To complete the proof of our claim, it suffices to prove that the "=" holds in (3.82). Suppose the contrary that Γ(nγ)<Γ¯α\Gamma_{(n-\gamma^{\prime})}<\bar{\Gamma}_{\alpha^{*}}, then by the definition of Γ(nγ),\Gamma_{(n-\gamma^{\prime})}, we get that there exists (m^,w^)𝒜(\hat{m},\hat{w})\in\mathcal{A} given in (1.28) such that

G(nγ)(m^,w^)Γ(nγ)+δ<Γ(nγ)+2δ<Γ¯α,\displaystyle G_{(n-\gamma^{\prime})}(\hat{m},\hat{w})\leq\Gamma_{(n-\gamma^{\prime})}+\delta<\Gamma_{(n-\gamma^{\prime})}+2\delta<\bar{\Gamma}_{\alpha^{*}}, (3.83)

where δ>0\delta>0 is sufficiently small. On the other hand, by the definition of Γα,\Gamma_{\alpha}, one finds

G(nγ)(m^,w^)=\displaystyle G_{(n-\gamma^{\prime})}(\hat{m},\hat{w})= Gα(m^,w^)(CLnm^|w^m^|γ𝑑x)(nm^𝑑x)(CLnm^|w^m^|γ𝑑x)nαγ(nm^𝑑x)2γ+αnγnm^(x)(Kαm^)(x)𝑑xnm^(x)(K(nγ)m^)(x)𝑑x\displaystyle G_{\alpha}(\hat{m},\hat{w})\frac{\Big{(}C_{L}\int_{\mathbb{R}^{n}}\hat{m}\Big{|}\frac{\hat{w}}{\hat{m}}\Big{|}^{\gamma^{\prime}}\,dx\Big{)}\Big{(}\int_{\mathbb{R}^{n}}\hat{m}\,dx\Big{)}}{{\Big{(}C_{L}\int_{\mathbb{R}^{n}}\hat{m}\Big{|}\frac{\hat{w}}{\hat{m}}\Big{|}^{\gamma^{\prime}}\,dx\Big{)}^{\frac{n-\alpha}{\gamma^{\prime}}}\Big{(}\int_{\mathbb{R}^{n}}\hat{m}\,dx\Big{)}^{\frac{2\gamma^{\prime}+\alpha-n}{\gamma^{\prime}}}}}\cdot\frac{\int_{\mathbb{R}^{n}}{\hat{m}(x)(K_{\alpha}*\hat{m})(x)}\,dx}{\int_{\mathbb{R}^{n}}\hat{m}(x)(K_{(n-\gamma^{\prime})}*\hat{m})(x)\,dx}
\displaystyle\geq Γα(CLnm^|w^m^|γ𝑑x)(nm^𝑑x)(CLnm^|w^m^|γ𝑑x)nαγ(nm^𝑑x)2γ+αnγnm^(x)(Kαm^)(x)𝑑xnm^(x)(K(nγ)m^)(x)𝑑x.\displaystyle\Gamma_{\alpha}\frac{\Big{(}C_{L}\int_{\mathbb{R}^{n}}\hat{m}\Big{|}\frac{\hat{w}}{\hat{m}}\Big{|}^{\gamma^{\prime}}\,dx\Big{)}\Big{(}\int_{\mathbb{R}^{n}}\hat{m}\,dx\Big{)}}{{\Big{(}C_{L}\int_{\mathbb{R}^{n}}\hat{m}\Big{|}\frac{\hat{w}}{\hat{m}}\Big{|}^{\gamma^{\prime}}\,dx\Big{)}^{\frac{n-\alpha}{\gamma^{\prime}}}\Big{(}\int_{\mathbb{R}^{n}}\hat{m}\,dx\Big{)}^{\frac{2\gamma^{\prime}+\alpha-n}{\gamma^{\prime}}}}}\cdot\frac{\int_{\mathbb{R}^{n}}{\hat{m}(x)(K_{\alpha}*\hat{m})(x)}\,dx}{\int_{\mathbb{R}^{n}}\hat{m}(x)(K_{(n-\gamma^{\prime})}*\hat{m})(x)\,dx}. (3.84)

Since

(CLnm^|w^m^|γ𝑑x)(nm^𝑑x)(CLnm^|w^m^|γ𝑑x)nαγ(nm^𝑑x)2γ+αnrnm^(x)(Kαm^)(x)𝑑xnm^(x)(K(nγ)m^)(x)𝑑x1 as α(nγ),\displaystyle\frac{\Big{(}C_{L}\int_{\mathbb{R}^{n}}\hat{m}\Big{|}\frac{\hat{w}}{\hat{m}}\Big{|}^{\gamma^{\prime}}\,dx\Big{)}\Big{(}\int_{\mathbb{R}^{n}}\hat{m}\,dx\Big{)}}{{\Big{(}C_{L}\int_{\mathbb{R}^{n}}\hat{m}\Big{|}\frac{\hat{w}}{\hat{m}}\Big{|}^{\gamma^{\prime}}\,dx\Big{)}^{\frac{n-\alpha}{\gamma^{\prime}}}\Big{(}\int_{\mathbb{R}^{n}}\hat{m}\,dx\Big{)}^{\frac{2\gamma^{\prime}+\alpha-n}{r}}}}\cdot\frac{\int_{\mathbb{R}^{n}}{\hat{m}(x)(K_{\alpha}\ast\hat{m})(x)}\,dx}{\int_{\mathbb{R}^{n}}\hat{m}(x)(K_{(n-\gamma^{\prime})}\ast\hat{m})(x)\,dx}\rightarrow 1\ \ \text{~{}as~{}}\alpha\searrow(n-\gamma^{\prime}),

then we can pass a limit in (3.83) and (3) to get

Γ¯α=lim infα(nγ)ΓαΓα(nγ)+δ<Γα(nγ)+2δlim infα(nγ)Γα,\displaystyle\bar{\Gamma}_{\alpha^{*}}=\liminf_{\alpha\searrow(n-\gamma^{\prime})}\Gamma_{\alpha}\leq\Gamma_{\alpha\searrow(n-\gamma^{\prime})}+\delta<\Gamma_{\alpha\searrow(n-\gamma^{\prime})}+2\delta\leq\liminf_{\alpha\searrow(n-\gamma^{\prime})}\Gamma_{\alpha},

which reaches a contradiction. Hence, the claim holds, i.e. Γ¯α=Γnγ\bar{\Gamma}_{\alpha^{*}}=\Gamma_{n-\gamma^{\prime}}.

Next, we prove (mα,wα)𝒜.(m_{\alpha^{*}},w_{\alpha^{*}})\in\mathcal{A}. Since (mα,wα)(m_{\alpha^{*}},w_{\alpha^{*}}) solves (3.76) and mαC0,θ(n)m_{\alpha^{*}}\in C^{0,\theta}(\mathbb{R}^{n}) with θ(0,1)\theta\in(0,1), we conclude from (3.72) and Lemma 2.2 that uαC1(n).u_{\alpha^{*}}\in C^{1}(\mathbb{R}^{n}). Then by standard elliptic estimates, the boundedness of uαL\|\nabla u_{\alpha^{*}}\|_{L^{\infty}} and the exponentially decaying property of mαm_{\alpha^{*}}, one can prove that (mα,wα)𝒜(m_{\alpha^{*}},w_{\alpha^{*}})\in\mathcal{A}.

Finally, it follows from (3.63) and (3.82) that

lim infα(nγ)Γα=12M=Γnγ,\displaystyle\liminf_{\alpha\searrow(n-\gamma^{\prime})}\Gamma_{\alpha}=\frac{1}{2}M^{*}=\Gamma_{n-\gamma^{\prime}}, (3.85)

where MM^{*} is given in (3.61). Then, by the fact (mα,wα)𝒜(m_{\alpha^{*}},w_{\alpha^{*}})\in\mathcal{A}, we deduce from (3.77), (3.78) and (3.85) that

Γ(nγ)=12M(CLn|wαmα|γmα𝑑x)(nmα𝑑x)nmα(x)(Kαmα)(x)𝑑x=12M~12M,\displaystyle\Gamma_{(n-\gamma^{\prime})}=\frac{1}{2}M^{*}\leq\frac{\bigg{(}C_{L}\int_{\mathbb{R}^{n}}\Big{|}\frac{w_{\alpha^{*}}}{m_{\alpha^{*}}}\Big{|}^{\gamma^{\prime}}m_{\alpha^{*}}\,dx\bigg{)}\Big{(}\int_{\mathbb{R}^{n}}m_{\alpha^{*}}\,dx\Big{)}}{\int_{\mathbb{R}^{n}}{m_{\alpha^{*}}(x)(K_{\alpha}*m_{\alpha^{*}})(x)}\,dx}=\frac{1}{2}\tilde{M}\leq\frac{1}{2}M^{*}, (3.86)

which shows (mα,wα)𝒜(m_{\alpha^{*}},w_{\alpha^{*}})\in\mathcal{A} is a minimizer of Γ(nγ)\Gamma_{(n-\gamma^{\prime})} and

nmα𝑑x=M and m¯α,Mαmαin L1(n) as α(nγ).\displaystyle\int_{\mathbb{R}^{n}}m_{\alpha^{*}}\,dx=M^{*}~{}\text{ and }~{}\bar{m}_{\alpha,M_{\alpha}}\rightarrow m_{\alpha^{*}}\ \text{in~{}}L^{1}(\mathbb{R}^{n})\text{ as }\alpha\searrow(n-\gamma^{\prime}).

These facts together with (3.76) indicate (1.33) holds. Now, we finish The proof of Theorem 1.1.

As shown in Theorem 1.1, we have obtained the existence of ground states to potential-free MFG systems under the mass critical exponent case, which is the Gagliardo-Nirenberg type’s inequality. In next section, we focus on the proof of Theorem 1.2.

4 Existence of Ground States: Coercive Potential MFGs

In this section, we shall discuss the existence of minimizers to problem (1.26). To this end, we have to perform the regularization procedure on (1.27) since when γ<n\gamma^{\prime}<n, the mm-component enjoys the worse regularity. In detail, we first consider the following auxiliary minimization problem

eϵ,M:=inf(m,w)𝒜Mϵ(m,w),\displaystyle e_{\epsilon,M}:=\inf_{(m,w)\in\mathcal{A}_{M}}\mathcal{E}_{\epsilon}(m,w), (4.1)

where 𝒜M{\mathcal{A}}_{M} is given by (3.2) and

ϵ(m,w):=CLn|wm|γm𝑑x+nV(x)m𝑑x12n{m(x)(K(nγ)m)(x)}ηϵ𝑑x,\displaystyle\mathcal{E}_{\epsilon}(m,w):=C_{L}\int_{\mathbb{R}^{n}}\Big{|}\frac{w}{m}\Big{|}^{\gamma^{\prime}}m\,dx+\int_{\mathbb{R}^{n}}V(x)m\,dx-\frac{1}{2}\int_{\mathbb{R}^{n}}\bigg{\{}m(x)(K_{(n-\gamma^{\prime})}\ast m)(x)\bigg{\}}\ast\eta_{\epsilon}\,dx, (4.2)

and ηϵ0\eta_{\epsilon}\geq 0 is the standard mollifier with

nηϵ𝑑x=1,supp(ηϵ)Bϵ(0),\int_{\mathbb{R}^{n}}\eta_{\epsilon}\,dx=1,\ \ \text{supp}(\eta_{\epsilon})\subset B_{\epsilon}(0),

for ϵ>0\epsilon>0 is sufficiently small. With the regularized energy (4.2), we are able to study the existence of minimizers to (1.26) by taking the limit. The crucial step in this procedure, as discussed in [10], is the uniformly boundedness of mϵm_{\epsilon} in LL^{\infty}, in which (mϵ,wϵ)(m_{\epsilon},w_{\epsilon}) is assumed to be a minimizer of (4.2).

Before proving Theorem 1.2, we collect some vital result shown in Section 3, which is

nm(x)(K(nγ)m)(x)𝑑x2CLM(n|wm|γm𝑑x)(nm𝑑x),(m,w)𝒜,\displaystyle\int_{\mathbb{R}^{n}}m(x)(K_{(n-\gamma^{\prime})}*m)(x)\,dx\leq\frac{2C_{L}}{M^{*}}\bigg{(}\int_{\mathbb{R}^{n}}\big{|}\frac{w}{m}\big{|}^{\gamma^{\prime}}m\,dx\bigg{)}\bigg{(}\int_{\mathbb{R}^{n}}m\,dx\bigg{)},~{}~{}\forall(m,w)\in\mathcal{A}, (4.3)

where 𝒜\mathcal{A} is given by (Blow-up Behaviors of Ground States in Ergodic Mean-field Games Systems with Hartree-type Coupling) and MM^{*} is defined by (1.34).

Then, we shall first prove energy (m,w)\mathcal{E}(m,w) given by (1.27) has a minimizer (m,w)𝒦M(m,w)\in\mathcal{K}_{M} if and only if M<MM<M^{*}, where 𝒦M\mathcal{K}_{M} is defined by (1.1). Next, we show that there exists (u,λ)C2(n)×(u,\lambda)\in C^{2}(\mathbb{R}^{n})\times\mathbb{R} such that (m,u,λ)W1,p(n)×C2(n)×(m,u,\lambda)\in W^{1,p}(\mathbb{R}^{n})\times C^{2}(\mathbb{R}^{n})\times\mathbb{R} is a solution to (1.38) when VV is assumed to satisfy (1.25) when γ>1.\gamma^{\prime}>1. Following the procedures discussed above, we are able to prove conclusions stated in Theorem 1.2. We would like to remark that with (1.25b) in assumption (V2) imposed on potential VV, the condition n|x|bm𝑑x<+\int_{\mathbb{R}^{n}}|x|^{b}m\,dx<+\infty in (3.2) must be satisfied for any minimizer. With this assumption, Gagliardo-Nirenberg type’s inequality (4.3) is valid. Next, we state some crucial propositions and lemmas, which will be used in the proof of Theorem 1.2, as follows:

Lemma 4.1.

Let

𝒲p,V:={m|mW1,p(n)L1(n) and nV(x)|m|𝑑x<}.\mathcal{W}_{p,V}:=\bigg{\{}m\big{|}\ m\in W^{1,p}(\mathbb{R}^{n})\cap L^{1}(\mathbb{R}^{n})\text{ and }\int_{\mathbb{R}^{n}}V(x)|m|\,dx<\infty\bigg{\}}.

Assume that 0V(x)Lloc(n)0\leq V(x)\in L_{\rm loc}^{\infty}(\mathbb{R}^{n}) with lim inf|x|V(x)=\liminf\limits_{|x|\to\infty}V(x)=\infty. Then, the embedding 𝒲p,VLq(n)\mathcal{W}_{p,V}\hookrightarrow L^{q}(\mathbb{R}^{n}) is compact for any 1q<p1\leq q<p^{*}, where p=npnpp^{*}=\frac{np}{n-p} if 1p<n1\leq p<n and p=p^{*}=\infty if pnp\geq n.

Proof.

See [1, Theorem 2.1] or [24, Theorem XIII.67]. ∎

In light of γ<n,\gamma^{\prime}<n, we establish the following lemma for the uniformly boundedness of mϵL:\|m_{\epsilon}\|_{L^{\infty}}:

Lemma 4.2.

Suppose that V(x)V(x) is locally Hölder continuous and satisfies (1.25). Let (uk,λk,mk)C2(n)××(L1(n)L2nn+α(n))(u_{k},\lambda_{k},m_{k})\in C^{2}(\mathbb{R}^{n})\times\mathbb{R}\times(L^{1}(\mathbb{R}^{n})\cap L^{\frac{2n}{n+\alpha^{*}}}(\mathbb{R}^{n})) be solutions to the following systems

{Δuk+CH|uk|γ+λk=Vgk[mk],xn,Δmk+CHγ(mk|uk|γ2uk)=0,xn,nmk𝑑x=M,\displaystyle\left\{\begin{array}[]{ll}-\Delta u_{k}+C_{H}|\nabla u_{k}|^{\gamma}+\lambda_{k}=V-g_{k}[m_{k}],&x\in\mathbb{R}^{n},\\ \Delta m_{k}+C_{H}\gamma\nabla\cdot(m_{k}|\nabla u_{k}|^{\gamma-2}\nabla u_{k})=0,&x\in\mathbb{R}^{n},\\ \int_{\mathbb{R}^{n}}m_{k}\,dx=M,\end{array}\right. (4.7)

where α=nγ\alpha^{*}=n-\gamma^{\prime} with 1<γ<n1<\gamma^{\prime}<n, gk:L1(n)L1(n)g_{k}:L^{1}(\mathbb{{\mathbb{R}}}^{n})\mapsto L^{1}(\mathbb{{\mathbb{R}}}^{n}) with θ(0,1)\theta\in(0,1) satisfies for all mLp(n)m\in L^{p}(\mathbb{R}^{n}), p[1,]p\in[1,\infty] and kk\in\mathbb{N},

gk[m]Lp(n)K(mnαn+αLp(n)+1) for someK>0,\|g_{k}[m]\|_{L^{p}(\mathbb{R}^{n})}\leq\mathrm{K}\bigg{(}\|m^{\frac{n-\alpha^{*}}{n+\alpha^{*}}}\|_{L^{p}(\mathbb{R}^{n})}+1\bigg{)}\ \text{ for some}\ \mathrm{K}>0, (4.8)

and

gk[m]Lp(BR(x0))K(mnαn+αLp(B2R(x0))+1) for any R>0 and x0n.\|g_{k}[m]\|_{L^{p}(B_{R}(x_{0}))}\leq\mathrm{K}\bigg{(}\|m^{\frac{n-\alpha^{*}}{n+\alpha^{*}}}\|_{L^{p}(B_{2R}(x_{0}))}+1\bigg{)}\ \text{ for any }R>0\text{ and }x_{0}\in\mathbb{R}^{n}. (4.9)

Assume that

supkmkL1(n)<,supkmkL2nn+α(n)<,supknVmk𝑑x<,supk|λk|<,\displaystyle\sup_{k}\|m_{k}\|_{L^{1}(\mathbb{R}^{n})}<\infty,~{}~{}\sup_{k}\|m_{k}\|_{L^{\frac{2n}{n+\alpha^{*}}}(\mathbb{R}^{n})}<\infty,~{}~{}\sup_{k}\int_{\mathbb{R}^{n}}Vm_{k}\,dx<\infty,~{}~{}\sup_{k}|\lambda_{k}|<\infty, (4.10)

and for all kk, uku_{k} is bounded from below uniformly. Then we have

lim supkmkL(n)<.\limsup_{k\to\infty}\|m_{k}\|_{L^{\infty}(\mathbb{{\mathbb{R}}}^{n})}<\infty. (4.11)
Proof.

By slightly modifying the argument shown in [14, Lemma 5.2], we finish the proof of this lemma. ∎

With the preliminary results shown above, we now begin the proof of Theorem 1.2.

Proof of Theorem 1.2:

Proof.

We first prove the Conclusion (i) in Theorem 1.2. To this end, we focus on the auxiliary problem (4.1). Invoking the Young’s inequality for convolution and the property of mollifier, one finds

nm(x)(K(nγ)m)(x)𝑑xn([m(K(nγ)m)]ηϵ)(x)𝑑xϵ0+nm(x)(K(nγ)m)(x)𝑑x,\displaystyle\int_{\mathbb{R}^{n}}m(x)(K_{(n-\gamma^{\prime})}\ast m)(x)\,dx\geq\int_{\mathbb{R}^{n}}\bigg{(}\Big{[}m(K_{(n-\gamma^{\prime})}\ast m)\Big{]}\ast\eta_{\epsilon}\bigg{)}(x)\,dx\overset{\epsilon\to 0^{+}}{\longrightarrow}\int_{\mathbb{R}^{n}}m(x)(K_{(n-\gamma^{\prime})}\ast m)(x)\,dx, (4.12)

for any mL2n2nγ(n)m\in L^{\frac{2n}{2n-\gamma^{\prime}}}(\mathbb{R}^{n}). Here, we have used the following Hardy-Littlewood-Sobolev inequality

|nm(x)(K(nγ)m)(x)𝑑x|C(n,γ)mL2n2nγ(n)2,mL2n2nγ(n).\displaystyle\bigg{|}\int_{\mathbb{R}^{n}}m(x)(K_{(n-\gamma^{\prime})}\ast m)(x)\,dx\bigg{|}\leq C(n,\gamma^{\prime})\|m\|^{2}_{L^{\frac{2n}{2n-\gamma^{\prime}}}(\mathbb{R}^{n})},\ \ \forall m\in L^{\frac{2n}{2n-\gamma^{\prime}}}(\mathbb{R}^{n}). (4.13)

As a consequence, in light of (4.3) and (4.12), we get

ϵ(m,w)(m,w)12(MM1)nm(x)(K(nγ)m)(x)𝑑x+nV(x)m𝑑x.\displaystyle\mathcal{E}_{\epsilon}(m,w)\geq\mathcal{E}(m,w)\geq\frac{1}{2}\bigg{(}\frac{M^{*}}{M}-1\bigg{)}\int_{\mathbb{R}^{n}}m(x)(K_{(n-\gamma^{\prime})}\ast m)(x)\,dx+\int_{\mathbb{R}^{n}}V(x)m\,dx. (4.14)

Next, we show that the minimization problem (4.1) is attainable. We first show that there exists C>0C>0 independent of ϵ\epsilon such that

eϵ,M<C<+,e_{\epsilon,M}<C<+\infty, (4.15)

where eϵ,Me_{\epsilon,M} is given by (4.1). Indeed, choosing

(m^,w^):=(e|x|L1(n)Me|x|,e|x|L1(n)Mx|x|e|x|)𝒦M,(\hat{m},\hat{w}):=\left(\frac{\|e^{-|x|}\|_{L^{1}(\mathbb{R}^{n})}}{M}e^{-|x|},\frac{\|e^{-|x|}\|_{L^{1}(\mathbb{R}^{n})}}{M}\frac{x}{|x|}e^{-|x|}\right)\in\mathcal{K}_{M},

one can find

eϵ,MCLn|w^m^|γm^𝑑x+nV(x)m^𝑑x<+,e_{\epsilon,M}\leq C_{L}\int_{\mathbb{R}^{n}}\Big{|}\frac{\hat{w}}{\hat{m}}\Big{|}^{\gamma^{\prime}}\hat{m}\,dx+\int_{\mathbb{R}^{n}}V(x)\hat{m}\,dx<+\infty,

which indicates that (4.15) holds. Let (mϵ,k,wϵ,k)𝒦M(m_{\epsilon,k},w_{\epsilon,k})\in\mathcal{K}_{M} be a minimizing sequence of problem (4.1), then we have from (4.15) that there exists C>0C>0 independent of ϵ\epsilon such that

limkϵ(mϵ,k,wϵ,k)=eϵ,M<C<+.\lim_{k\to\infty}\mathcal{E}_{\epsilon}(m_{\epsilon,k},w_{\epsilon,k})=e_{\epsilon,M}<C<+\infty. (4.16)

Moreover, it follows from (4.3), (4.14), (4.16) and the fact M<MM<M^{*} that

supk+nmϵ,k(x)(K(nγ)mϵ,k)(x)𝑑xC<+,\displaystyle\sup_{k\in\mathbb{N}^{+}}\int_{\mathbb{R}^{n}}m_{\epsilon,k}(x)\big{(}K_{(n-\gamma^{\prime})}\ast m_{\epsilon,k}\big{)}(x)\,dx\leq C<+\infty, (4.17)

and

supk+n(|wϵ,kmϵ,k|γmϵ,k+V(x)mϵ,k)𝑑xC<+,\displaystyle~{}~{}\sup_{k\in\mathbb{N}^{+}}\int_{\mathbb{R}^{n}}\bigg{(}\bigg{|}\frac{w_{\epsilon,k}}{m_{\epsilon,k}}\bigg{|}^{\gamma^{\prime}}m_{\epsilon,k}+V(x)m_{\epsilon,k}\bigg{)}\,dx\leq C<+\infty, (4.18)

where C>0C>0 is independent of ϵ\epsilon. The subsequent argument for proving Conclusion (i) is similar as shown in the proof of Theorem 1.3 in [14]. In fact, with the aid of the key Lemma 2.5, we obtain from (4.17) that

supk+mϵ,kW1,q^(n)C<+ and supk+wϵ,kLp(n)C<+,for any p[1,q^],\displaystyle\sup_{k\in\mathbb{N}^{+}}\|m_{\epsilon,k}\|_{W^{1,\hat{q}}(\mathbb{R}^{n})}\leq C<+\infty\ \text{ and }\ \sup_{k\in\mathbb{N}^{+}}\|w_{\epsilon,k}\|_{L^{p}(\mathbb{R}^{n})}\leq C<+\infty,\ \ \text{for any }p\in[1,\hat{q}], (4.19)

where q^\hat{q} is defined by (1.24) and C>0C>0 is some constant independent of ϵ\epsilon. As a consequence, there exists (mϵ,wϵ)W1,q^(n)×Lq^(n)(m_{\epsilon},w_{\epsilon})\in W^{1,\hat{q}}(\mathbb{R}^{n})\times L^{\hat{q}}(\mathbb{R}^{n}) such that

(mϵ,k,wϵ,k)𝑘(mϵ,wϵ) in W1,q^(n)×Lq^(n).\displaystyle(m_{\epsilon,k},w_{\epsilon,k})\overset{k}{\rightharpoonup}(m_{\epsilon},w_{\epsilon})\text{~{}in~{}}W^{1,\hat{q}}(\mathbb{R}^{n})\times L^{\hat{q}}(\mathbb{R}^{n}). (4.20)

In light of the assumption (V1), lim|x|V(x)=+\lim\limits_{|x|\rightarrow\infty}V(x)=+\infty, given in Subsection 1.2, one can deduce from Lemma 4.1 that

mϵ,k𝑘mϵ in L1(n)L2n2nγ(n).\displaystyle m_{\epsilon,k}\overset{k}{\rightarrow}m_{\epsilon}\text{~{}in~{}}L^{1}(\mathbb{R}^{n})\cap L^{\frac{2n}{2n-\gamma^{\prime}}}(\mathbb{R}^{n}). (4.21)

Therefore, up to a subsequence,

n([mϵ,k(K(nγ)mϵ,k)]ηϵ)(x)𝑑x𝑘n([mϵ(K(nγ)mϵ)]ηϵ)(x)𝑑x.\int_{\mathbb{R}^{n}}\bigg{(}\Big{[}m_{\epsilon,k}\big{(}K_{(n-\gamma^{\prime})}\ast m_{\epsilon,k}\big{)}\Big{]}\ast\eta_{\epsilon}\bigg{)}(x)\,dx\overset{k}{\rightarrow}\int_{\mathbb{R}^{n}}\bigg{(}\Big{[}m_{\epsilon}\big{(}K_{(n-\gamma^{\prime})}\ast m_{\epsilon}\big{)}\Big{]}\ast\eta_{\epsilon}\bigg{)}(x)\,dx. (4.22)

In addition, thanks to the convexity of n|wm|γm𝑑x\int_{\mathbb{R}^{n}}\big{|}\frac{w}{m}\big{|}^{\gamma^{\prime}}m\,dx, by letting kk\to\infty in (4.17), we have there exists C>0C>0 independent of ϵ>0\epsilon>0 such that

n|wϵmϵ|γmϵ𝑑x+nV(x)mϵ𝑑xlim infk+n|wϵ,kmϵ,k|γmϵ,k𝑑x+nV(x)mϵ,k𝑑xC<+.\displaystyle\int_{\mathbb{R}^{n}}\Big{|}\frac{w_{\epsilon}}{m_{\epsilon}}\Big{|}^{\gamma^{\prime}}m_{\epsilon}\,dx+\int_{\mathbb{R}^{n}}V(x)m_{\epsilon}\,dx\leq\liminf_{k\rightarrow+\infty}\int_{\mathbb{R}^{n}}\big{|}\frac{w_{\epsilon,k}}{m_{\epsilon,k}}\big{|}^{\gamma^{\prime}}m_{\epsilon,k}\,dx+\int_{\mathbb{R}^{n}}V(x)m_{\epsilon,k}\,dx\leq C<+\infty. (4.23)

Moreover,

n|wϵ|V1γ𝑑x(n|wϵmϵ|γmϵ𝑑x)γ(nVmϵ𝑑x)γC<.\int_{\mathbb{R}^{n}}|w_{\epsilon}|V^{\frac{1}{\gamma}}\,dx\leq\left(\int_{\mathbb{R}^{n}}\Big{|}\frac{w_{\epsilon}}{m_{\epsilon}}\Big{|}^{\gamma^{\prime}}m_{\epsilon}\,dx\right)^{\gamma^{\prime}}\left(\int_{\mathbb{R}^{n}}Vm_{\epsilon}\,dx\right)^{\gamma}\leq C<\infty. (4.24)

and

n|wϵ|𝑑x(n|wϵmϵ|γmϵ𝑑x)γ(nmϵ𝑑x)γC<.\int_{\mathbb{R}^{n}}|w_{\epsilon}|\,dx\leq\left(\int_{\mathbb{R}^{n}}\Big{|}\frac{w_{\epsilon}}{m_{\epsilon}}\Big{|}^{\gamma^{\prime}}m_{\epsilon}\,dx\right)^{\gamma^{\prime}}\left(\int_{\mathbb{R}^{n}}m_{\epsilon}\,dx\right)^{\gamma}\leq C<\infty. (4.25)

Combining (4.20) and (4.21) with (4.25), we deduce that (mϵ,wϵ)𝒦M(m_{\epsilon},w_{\epsilon})\in\mathcal{K}_{M}. Then, one invokes (4.22) and (4.23) to get

eϵ,M=limkϵ(mϵ,k,wϵ,k)ϵ(mϵ,wϵ)eϵ,M,\displaystyle e_{\epsilon,M}=\lim_{k\rightarrow\infty}\mathcal{E}_{\epsilon}(m_{\epsilon,k},w_{\epsilon,k})\geq\mathcal{E}_{\epsilon}(m_{\epsilon},w_{\epsilon})\geq e_{\epsilon,M},

which indicates (mϵ,wϵ)𝒦M(m_{\epsilon},w_{\epsilon})\in\mathcal{K}_{M} is a minimizer of problem (4.1). Finally, similarly as the proof of Proposition 3.4 in [10] and the arguments shown in Proposition 5.1 and Proposition 5.2 in [14], we apply Lemma 2.4 to obtain that there exists uϵC2(n)u_{\epsilon}\in C^{2}(\mathbb{R}^{n}) bounded from below (depending on ϵ\epsilon) and λϵ\lambda_{\epsilon}\in\mathbb{R} such that

{Δuϵ+CH|uϵ|γ+λϵ=V(x)(K(nγ)mϵ)ηϵ,Δmϵ+CHγ(mϵ|uϵ|γ2uϵ)=0,wϵ=CHγmϵ|uϵ|γ2uϵ,nmϵ𝑑x=M<M.\displaystyle\left\{\begin{array}[]{ll}-\Delta u_{\epsilon}+C_{H}|\nabla u_{\epsilon}|^{\gamma}+\lambda_{\epsilon}=V(x)-\big{(}K_{(n-\gamma^{\prime})}\ast m_{\epsilon}\big{)}\ast\eta_{\epsilon},\\ \Delta m_{\epsilon}+C_{H}\gamma\nabla\cdot(m_{\epsilon}|\nabla u_{\epsilon}|^{\gamma-2}\nabla u_{\epsilon})=0,\\ w_{\epsilon}=-C_{H}\gamma m_{\epsilon}|\nabla u_{\epsilon}|^{\gamma-2}\nabla u_{\epsilon},\ \int_{\mathbb{R}^{n}}m_{\epsilon}\,dx=M<M^{*}.\end{array}\right. (4.29)

For each fixed ϵ>0\epsilon>0, we utilize Lemma 2.2 to obtain that there exists Cϵ>0C_{\epsilon}>0 depends on ϵ\epsilon such that |uϵ(x)|Cϵ(1+V(x))1γ|\nabla u_{\epsilon}(x)|\leq C_{\epsilon}(1+V(x))^{\frac{1}{\gamma}}. Noting that uϵC2(n)u_{\epsilon}\in C^{2}(\mathbb{R}^{n}) and (K(nγ)mϵ)ηϵL(n)\big{(}K_{(n-\gamma^{\prime})}\ast m_{\epsilon}\big{)}\ast\eta_{\epsilon}\in L^{\infty}(\mathbb{R}^{n}), we have from the classical regularity of the uu-equation in (4.29) that |Δuϵ(x)|Cϵ(1+V(x))|\Delta u_{\epsilon}(x)|\leq C_{\epsilon}(1+V(x)). We next prove

|λϵ|C<, with C>0 independent of ϵ>0.|\lambda_{\epsilon}|\leq C<\infty,\text{ with $C>0$ independent of $\epsilon>0$}. (4.30)

To show this, we apply the integration by parts to the mm-equation and the uu-equation in (4.29), then get

nmΔuϵ𝑑x=nwϵuϵdx=CHγnmϵ|uϵ|γ𝑑x,\displaystyle\int_{\mathbb{R}^{n}}m\Delta u_{\epsilon}\,dx=\int_{\mathbb{R}^{n}}w_{\epsilon}\cdot\nabla u_{\epsilon}\,dx=-C_{H}\gamma\int_{\mathbb{R}^{n}}m_{\epsilon}|\nabla u_{\epsilon}|^{\gamma}\,dx,

and

λϵM=(1γ)CHnmϵ|uϵ|γ𝑑x+nVmϵ𝑑xnmϵ(K(nγ)mϵ)ηϵ𝑑x=CLnmϵ|wϵmϵ|γ𝑑x+nVmϵ𝑑xnmϵ(K(nγ)mϵ)ηϵ𝑑x\begin{split}\lambda_{\epsilon}M&=-(1-\gamma)C_{H}\int_{\mathbb{R}^{n}}m_{\epsilon}|\nabla u_{\epsilon}|^{\gamma}\,dx+\int_{\mathbb{R}^{n}}Vm_{\epsilon}\,dx-\int_{\mathbb{R}^{n}}m_{\epsilon}\big{(}K_{(n-\gamma^{\prime})}\ast m_{\epsilon}\big{)}\ast\eta_{\epsilon}\,dx\\ &=C_{L}\int_{\mathbb{R}^{n}}m_{\epsilon}\bigg{|}\frac{w_{\epsilon}}{m_{\epsilon}}\bigg{|}^{\gamma^{\prime}}\,dx+\int_{\mathbb{R}^{n}}Vm_{\epsilon}\,dx-\int_{\mathbb{R}^{n}}m_{\epsilon}\big{(}K_{(n-\gamma^{\prime})}\ast m_{\epsilon}\big{)}\ast\eta_{\epsilon}\,dx\end{split} (4.31)

where we have used the fact that CL=1γ(γCH)11γC_{L}=\frac{1}{\gamma^{\prime}}(\gamma C_{H})^{\frac{1}{1-\gamma}}. Collecting (4.23), (4.24) and (4.31), one finds (4.30) holds.

Next, we let ϵ0\epsilon\rightarrow 0 and show the existence of the minimizer (mM,wM)(m_{M},w_{M}) to problem (1.26). Noting (mϵ,uϵ,λϵ)(m_{\epsilon},u_{\epsilon},\lambda_{\epsilon}) satisfies (4.10) with kk replaced by ϵ.\epsilon. We utilize Young’s inequality for convolution and Hardy-Littlewood-Sobolev inequality (A.1) to get

supk(K(nγ)mk)ηkL1+n+αnα(n)C(n,γ)supkmknαn+αL1+n+αnα(n)=C(n,γ)supkmkL2n2nγ(n)<,\displaystyle\sup_{k}\|\big{(}K_{(n-\gamma^{\prime})}\ast m_{k}\big{)}\ast\eta_{k}\|_{L^{1+\frac{n+\alpha^{*}}{n-\alpha^{*}}}(\mathbb{R}^{n})}\leq C(n,\gamma^{\prime})\sup_{k}\|m_{k}^{\frac{n-\alpha^{*}}{n+\alpha^{*}}}\|_{L^{1+\frac{n+\alpha^{*}}{n-\alpha^{*}}}(\mathbb{R}^{n})}=C(n,\gamma^{\prime})\sup_{k}\|m_{k}\|_{L^{\frac{2n}{2n-\gamma^{\prime}}}(\mathbb{R}^{n})}<\infty,

and

supk(K(nγ)mk)ηkL1+n+αnα(B2R(x0))C(n,γ)supkmkL2n2nγ(B2R(x0))<,\displaystyle\sup_{k}\|\big{(}K_{(n-\gamma^{\prime})}\ast m_{k}\big{)}\ast\eta_{k}\|_{L^{1+\frac{n+\alpha^{*}}{n-\alpha^{*}}}\big{(}B_{2R}(x_{0})\big{)}}\leq C(n,\gamma^{\prime})\sup_{k}\|m_{k}\|_{L^{\frac{2n}{2n-\gamma^{\prime}}}\big{(}B_{2R}(x_{0})\big{)}}<\infty,

Then, collecting (4.23) and (4.30), we invoke Lemma 4.2 to conclude that

lim supϵ0+mϵL(n)<.\limsup_{\epsilon\to 0^{+}}\|m_{\epsilon}\|_{L^{\infty}(\mathbb{R}^{n})}<\infty. (4.32)

Then, by using Lemma 2.2, we obtain

|uϵ(x)|C(1+V(x))1γ, where C>0 is independent of ϵ.|\nabla u_{\epsilon}(x)|\leq C(1+V(x))^{\frac{1}{\gamma}},\text{ where $C>0$ is independent of $\epsilon$.} (4.33)

Since uϵu_{\epsilon} is bounded from below, without loss of generality, we assume that uϵ(0)=0u_{\epsilon}(0)=0. In light of (2.7), one finds that uϵ(x)CϵV1γ(x)Cϵ+u_{\epsilon}(x)\geq C_{\epsilon}V^{\frac{1}{\gamma}}(x)-C_{\epsilon}\to+\infty as |x|+|x|\to+\infty, which indicates each uϵ(x)C2(n)u_{\epsilon}(x)\in C^{2}(\mathbb{R}^{n}) admits its minimum at some finite point xϵx_{\epsilon}. By using (4.30), (4.32) and the coercivity of VV, we obtain from the uu-equation of (4.29) that xϵx_{\epsilon} is uniformly bounded with respect to ϵ\epsilon. The fact uϵ(0)=0u_{\epsilon}(0)=0 together with (4.33) implies that there exists C>0C>0 independent of ϵ\epsilon such that

Cuϵ(x)C|x|(1+V(x))1γ for all xn,-C\leq u_{\epsilon}(x)\leq C|x|(1+V(x))^{\frac{1}{\gamma}}\text{ for all }x\in\mathbb{R}^{n},

where we have used (1.25c) in the second inequality. Since uϵu_{\epsilon} are bounded from below uniformly, one can employ Lemma 2.3 to get that uϵ(x)CV1γ(x)C with C>0 independent of ϵ.u_{\epsilon}(x)\geq CV^{\frac{1}{\gamma}}(x)-C\text{ with $C>0$ independent of $\epsilon$.} Thus, with the assumptions (1.25) imposed on VV, we get

C1V1γ(x)C1uϵC2|x|(1+V(x))1γ,for all xn.C_{1}V^{\frac{1}{\gamma}}(x)-C_{1}\leq u_{\epsilon}\leq C_{2}|x|(1+V(x))^{\frac{1}{\gamma}},\text{for all }x\in\mathbb{R}^{n}. (4.34)

where C1,C2>0C_{1},C_{2}>0 are independent of ϵ\epsilon.

In light of (4.32) and (4.33), one finds for any R>1R>1 and p>1p>1,

wϵLp(B2R(0))=CHγmϵ|uϵ|γ1Lp(B2R(0))Cp,R<,\|w_{\epsilon}\|_{L^{p}(B_{2R}(0))}=C_{H}\gamma\|m_{\epsilon}|\nabla u_{\epsilon}|^{\gamma-1}\|_{L^{p}(B_{2R}(0))}\leq C_{p,R}<\infty, (4.35)

where the constant Cp,R>0C_{p,R}>0 depends only on pp, RR and is independent of ϵ\epsilon. Then, with the help of Lemma 2.5, we obtain from (4.35) that mϵW1,p(B2R(0))Cp,R<\|m_{\epsilon}\|_{W^{1,p}(B_{2R}(0))}\leq C_{p,R}<\infty. Taking p>np>n large enough, we utilize Sobolev embedding theorem to get

mϵC0,θ1(B2R(0))Cθ1,R< for some θ1(0,1).\|m_{\epsilon}\|_{C^{0,\theta_{1}}(B_{2R}(0))}\leq C_{\theta_{1},R}<\infty\text{ for some $\theta_{1}\in(0,1)$.} (4.36)

To estimate uϵ,u_{\epsilon}, we rewrite the uu-equation of (4.29) as

Δuϵ=CH|uϵ|γ+fϵ(x) with fϵ(x):=λϵ+V(x)(K(nγ)mϵ)ηϵ,-\Delta u_{\epsilon}=-C_{H}|\nabla u_{\epsilon}|^{\gamma}+f_{\epsilon}(x)\ \text{ with }f_{\epsilon}(x):=-\lambda_{\epsilon}+V(x)-\big{(}K_{(n-\gamma^{\prime})}\ast m_{\epsilon}\big{)}\ast\eta_{\epsilon}, (4.37)

Since mϵC0,θ1(B2R(0))m_{\epsilon}\in C^{0,\theta_{1}}(B_{2R}(0)), then mϵηϵL1(B2R(0))Lq~(B2R(0))m_{\epsilon}\ast\eta_{\epsilon}\in L^{1}(B_{2R}(0))\cap L^{\tilde{q}}(B_{2R}(0)) with q~>nα\tilde{q}>\frac{n}{\alpha}. Thus, we deduce from Lemma A.2 that (K(nγ)mϵ)ηϵC0,θ2(B2R(0))\big{(}K_{(n-\gamma^{\prime})}\ast m_{\epsilon}\big{)}\ast\eta_{\epsilon}\in C^{0,\theta_{2}}(B_{2R}(0)) for some θ2(0,1)\theta_{2}\in(0,1). Now, by using (4.32), (4.33) and the fact that VV is locally Hölder continuous, we obtain that for any p>1p>1,

fϵLp(B2R(0))+|uϵ|γLp(B2R(0))Cp,R<.\|f_{\epsilon}\|_{L^{p}(B_{2R}(0))}+\||\nabla u_{\epsilon}|^{\gamma}\|_{L^{p}(B_{2R}(0))}\leq C_{p,R}<\infty.

Then we utilize the standard elliptic regularity in (4.37) to get

uϵC2,θ3(BR(0))Cθ3,R<, for some θ3(0,1),\|u_{\epsilon}\|_{C^{2,\theta_{3}}(B_{R}(0))}\leq C_{\theta_{3},R}<\infty,\text{ for some $\theta_{3}\in(0,1)$,} (4.38)

where R>0.R>0. Letting RR\to\infty and proceeding the standard diagonalization procedure, we invoke Arzelà-Ascoli theorem to find there exists uMC2(n)u_{M}\in C^{2}(\mathbb{R}^{n}) such that

uϵϵ0+uM in Cloc2,θ4(n) for some θ4(0,1).u_{\epsilon}\overset{\epsilon\to 0^{+}}{\longrightarrow}u_{M}\text{ in }C^{2,\theta_{4}}_{\rm loc}(\mathbb{R}^{n})\text{ for some $\theta_{4}\in(0,1)$.} (4.39)

In addition, by using Lemma 2.5 and (4.23), we find there exists (mM,wM)W1,q^(n)×(L1(n)Lq^(n))(m_{M},w_{M})\in W^{1,\hat{q}}(\mathbb{R}^{n})\times\big{(}L^{1}(\mathbb{R}^{n})\cap L^{\hat{q}}(\mathbb{R}^{n})\big{)} such that

mϵϵ0+mM a.e. in n,  and (mϵ,wϵ)ϵ0+(mM,wM) in W1,q^(n)×Lq^(n).\displaystyle m_{\epsilon}\overset{\epsilon\to 0^{+}}{\to}m_{M}\text{ a.e. in $\mathbb{R}^{n}$, \ and \ \ }(m_{\epsilon},w_{\epsilon})\overset{\epsilon\to 0^{+}}{\rightharpoonup}(m_{M},w_{M})\text{~{}in~{}}W^{1,\hat{q}}(\mathbb{R}^{n})\times L^{\hat{q}}(\mathbb{R}^{n}). (4.40)

Moreover, invoking Lemma 4.1, one finds

mϵϵ0+mM in L1(n)L2n2nγ(n).\displaystyle m_{\epsilon}\overset{\epsilon\to 0^{+}}{\rightarrow}m_{M}\text{~{}in~{}}L^{1}(\mathbb{R}^{n})\cap L^{\frac{2n}{2n-\gamma^{\prime}}}(\mathbb{R}^{n}). (4.41)

Passing to the limit as ϵ0+\epsilon\to 0^{+} in (4.29), we then obtain from (4.30) and (4.39)-(4.41) that there exists λM\lambda_{M}\in\mathbb{R} such that (mM,uM,wM)(m_{M},u_{M},w_{M}) satisfies (1.38). In addition, we infer from (4.33) and (4.34) that

|uM(x)|C(1+V(x))1γandC1|x|1+bγC1uMC2|x|1+bγ+C2,xn.|\nabla u_{M}(x)|\leq C(1+V(x))^{\frac{1}{\gamma}}\ \text{and}\ C_{1}|x|^{1+\frac{b}{\gamma}}-C_{1}\leq u_{M}\leq C_{2}|x|^{1+\frac{b}{\gamma}}+C_{2},\ \forall x\in\mathbb{R}^{n}. (4.42)

Recall that mϵmMm_{\epsilon}\to m_{M} a.e. as ϵ0+\epsilon\to 0^{+} in n\mathbb{R}^{n}, then we use (4.32) to get that mML(n)m_{M}\in L^{\infty}(\mathbb{R}^{n}). Then, proceeding the same argument as shown in the proof of Proposition 5.2 in [14], one can further find from (1.38) and (4.42) that

wM=CHγmM|uM|γ2uMLp(n)w_{M}=-C_{H}\gamma m_{M}|\nabla u_{M}|^{\gamma-2}\nabla u_{M}\in L^{p}(\mathbb{R}^{n}) and mMW1,p(n)m_{M}\in W^{1,p}(\mathbb{R}^{n}), p>1\forall p>1. (4.43)

Finally, we prove that (mM,wM)𝒦M(m_{M},w_{M})\in\mathcal{K}_{M} is a minimizer of eα,Me_{\alpha^{*},M}. To this end, we claim that for M<MM<M^{*},

limϵ0+eϵ,M=eα,M,\lim_{\epsilon\to 0^{+}}e_{\epsilon,M}=e_{\alpha^{*},M}, (4.44)

where eα,Me_{\alpha^{*},M} is given in (1.26). On one hand, in view of (4.12), it is straightforward to get limϵ0+eϵ,Meα,M\lim\limits_{\epsilon\to 0^{+}}e_{\epsilon,M}\geq e_{\alpha^{*},M}. On the other hand, we aim to show limϵ0+eϵ,Meα,M\lim\limits_{\epsilon\to 0^{+}}e_{\epsilon,M}\leq e_{\alpha^{*},M}. Due to the definition of eα,Me_{\alpha^{*},M}, for any δ>0\delta>0, we choose (m,w)𝒦M(m,w)\in\mathcal{K}_{M} such that (m,w)eα,M+δ2\mathcal{E}(m,w)\leq e_{\alpha^{*},M}+\frac{\delta}{2}. In light of (4.12), we conclude that for ϵ>0\epsilon>0 small enough, ϵ(m,w)(m,w)+δ2\mathcal{E}_{\epsilon}(m,w)\leq\mathcal{E}(m,w)+\frac{\delta}{2}. Thus,

eϵ,Mϵ(m,w)(m,w)+δ2eα,M+δ.e_{\epsilon,M}\leq\mathcal{E}_{\epsilon}(m,w)\leq\mathcal{E}(m,w)+\frac{\delta}{2}\leq e_{\alpha^{*},M}+{\delta}.

Letting ϵ0+\epsilon\to 0^{+} at first and then δ0+\delta\to 0^{+}, one has limϵ0+eϵ,Meα,M\lim\limits_{\epsilon\to 0^{+}}e_{\epsilon,M}\leq e_{\alpha^{*},M}. Combining the two facts, we finish the proof of (4.44).

We collect (4.40), (4.41), (4.44) and the convexity of n|wm|γm𝑑x\int_{\mathbb{R}^{n}}\big{|}\frac{w}{m}\big{|}^{\gamma^{\prime}}m\,dx to get

eα,M=limϵ0+eϵ,M=limϵ0+ϵ(mϵ,wϵ)(mM,wM)eα,M,e_{\alpha^{*},M}=\lim_{\epsilon\to 0^{+}}e_{\epsilon,M}=\lim_{\epsilon\to 0^{+}}\mathcal{E}_{\epsilon}(m_{\epsilon},w_{\epsilon})\geq\mathcal{E}(m_{M},w_{M})\geq e_{\alpha^{*},M},

which implies (mM,wM)𝒦M(m_{M},w_{M})\in\mathcal{K}_{M} is a minimizer of eα,Me_{\alpha^{*},M}. This completes the proof of Conclusion (i).

Now, we focus on Conclusion (ii) of Theorem 1.2. We have the fact that (mα,wα,uα)\big{(}m_{\alpha^{*}},w_{\alpha^{*}},u_{\alpha^{*}}\big{)} given in Theorem 1.1 is a minimizer of problem (1.28) with α=α=(nγ)\alpha=\alpha^{*}=(n-\gamma^{\prime}). To simplify notation, we rewrite (mα,wα,uα)(m_{\alpha^{*}},w_{\alpha^{*}},u_{\alpha^{*}}) as (m,w,u)(m_{*},w_{*},u_{*}), then define

(mt,wt)=(MMtnm(t(xx0)),MMtn+1w(t(xx0)))𝒦M,t>0,x0n.\displaystyle(m_{*}^{t},w_{*}^{t})=\bigg{(}\frac{M}{M^{*}}t^{n}m_{*}(t(x-x_{0})),\frac{M}{M^{*}}t^{n+1}w_{*}(t(x-x_{0}))\bigg{)}\in\mathcal{K}_{M},\ \ \forall t>0,~{}x_{0}\in\mathbb{R}^{n}. (4.45)

where the constraint set 𝒦M\mathcal{K}_{M} and M>0M^{*}>0 are defined by (1.1) and (1.34), respectively. Since uC2(n)u_{*}\in C^{2}(\mathbb{R}^{n}) and mm_{*} decays exponentially as stated in Theorem 1.1, we utilize Lemma 2.8 to find

CLn|wm|γm𝑑x=12nm(x)(K(nγ)m)(x)𝑑x.\displaystyle C_{L}\int_{\mathbb{R}^{n}}\bigg{|}\frac{w_{*}}{m_{*}}\bigg{|}^{\gamma^{\prime}}m_{*}\,dx=\frac{1}{2}\int_{\mathbb{R}^{n}}m_{*}(x)\big{(}K_{(n-\gamma^{\prime})}\ast m_{*}\big{)}(x)\,dx. (4.46)

Thanks to (4.46), we substitute (4.45) into (1.27), then obtain that if M>M,M>M^{*},

eα,M(mt,wt)=\displaystyle e_{\alpha^{*},M}\leq\mathcal{E}(m_{*}^{t},w_{*}^{t})= MM(CLtγn|wm|γm𝑑x+nV(x)m𝑑x)tγ2(MM)2nm(x)(K(nγ)m)(x)𝑑x\displaystyle\frac{M}{M^{*}}\bigg{(}C_{L}t^{\gamma^{\prime}}\int_{\mathbb{R}^{n}}\Big{|}\frac{w_{*}}{m_{*}}\Big{|}^{\gamma^{\prime}}m_{*}\,dx+\int_{\mathbb{R}^{n}}V(x)m_{*}\,dx\bigg{)}-\frac{t^{\gamma^{\prime}}}{2}\bigg{(}\frac{M}{M^{*}}\bigg{)}^{2}\int_{\mathbb{R}^{n}}m_{*}(x)(K_{(n-\gamma^{\prime})}\ast m_{*})(x)\,dx
=\displaystyle= MM[1(MM)]tγ2nm(x)(K(nγ)m)(x)𝑑x+MV(x0)+ot(1)\displaystyle\frac{M}{M^{*}}\Big{[}1-\bigg{(}\frac{M}{M^{*}}\bigg{)}\Big{]}\frac{t^{\gamma^{\prime}}}{2}\int_{\mathbb{R}^{n}}m_{*}(x)\big{(}K_{(n-\gamma^{\prime})}\ast m_{*}\big{)}(x)\,dx+MV(x_{0})+o_{t}(1)
as t+.\displaystyle\rightarrow-\infty\ \ \text{as~{}}t\rightarrow+\infty. (4.47)

Therefore, we have eα,M=e_{\alpha^{*},M}=-\infty for M>M,M>M^{*}, which indicates that problem (1.22) does not admit any minimizer.

Now, we are concentrated at the critical case M=MM=M^{*} and plan to show Conclusion (iii). To begin with, we prove that up to a subsequence,

limMMeα,M=eα,M=0.\displaystyle\lim_{M\nearrow M^{*}}e_{\alpha^{*},M}=e_{\alpha^{*},M^{*}}=0. (4.48)

Indeed, since infxnV(x)=0\inf_{x\in\mathbb{R}^{n}}V(x)=0 as shown in (V1)(V1) and eα,Me_{\alpha^{*},M^{*}} is defined by (1.22), we have for any δ>0\delta>0, (m,w)𝒜M\exists(m,w)\in\mathcal{A}_{M^{*}} such that

eα,M(m,w)eα,M+δ.\displaystyle e_{\alpha^{*},M^{*}}\leq\mathcal{E}(m,w)\leq e_{\alpha^{*},M^{*}}+\delta. (4.49)

Noting that MM(m,w)𝒜M\frac{M}{M^{*}}(m,w)\in\mathcal{A}_{M}, we further obtain

eα,M\displaystyle e_{\alpha^{*},M} (MMm,MMw)\displaystyle\leq\mathcal{E}\bigg{(}\frac{M}{M^{*}}m,\frac{M}{M^{*}}w\bigg{)} (4.50)
=\displaystyle= (m,w)+(MM1)[CLn|wm|γm𝑑x+nV(x)m𝑑x]+12[1(MM)2]nm(x)(K(nγ)m)(x)𝑑x.\displaystyle\mathcal{E}(m,w)+\Big{(}\frac{M}{M^{*}}-1\Big{)}\bigg{[}C_{L}\int_{\mathbb{R}^{n}}\Big{|}\frac{w}{m}\Big{|}^{\gamma^{\prime}}m\,dx+\int_{\mathbb{R}^{n}}V(x)m\,dx\bigg{]}+\frac{1}{2}\bigg{[}1-\bigg{(}\frac{M}{M^{*}}\bigg{)}^{2}\bigg{]}\int_{\mathbb{R}^{n}}m(x)\big{(}K_{(n-\gamma^{\prime})}\ast m\big{)}(x)\,dx.

By a straightforward computation, one has as MM,M\nearrow M^{*},

(MM1)[CLn|wm|γm𝑑x+nV(x)m𝑑x]+12[1(MM)2]nm(x)(K(nγ)m)(x)𝑑x0.\displaystyle\Big{(}\frac{M}{M^{*}}-1\Big{)}\Big{[}C_{L}\int_{\mathbb{R}^{n}}\bigg{|}\frac{w}{m}\bigg{|}^{\gamma^{\prime}}m\,dx+\int_{\mathbb{R}^{n}}V(x)m\,dx\Big{]}+\frac{1}{2}\bigg{[}1-\bigg{(}\frac{M}{M^{*}}\bigg{)}^{2}\bigg{]}\int_{\mathbb{R}^{n}}m(x)\big{(}K_{(n-\gamma^{\prime})}\ast m\big{)}(x)\,dx\rightarrow 0. (4.51)

We collect (4.49), (4.50) and (4.51) to get

lim supMMeα,Meα,M+δ,δ>0.\displaystyle\limsup_{M\nearrow M^{*}}e_{\alpha^{*},M}\leq e_{\alpha^{*},M^{*}}+\delta,\ \ \forall\delta>0. (4.52)

Letting δ0\delta\rightarrow 0 in (4.52), one has from (4.52) that

lim supMMeα,Meα,M.\displaystyle\limsup_{M\nearrow M^{*}}e_{\alpha^{*},M}\leq e_{\alpha^{*},M^{*}}. (4.53)

In addition, define (m¯α,M,w¯α,M)𝒜M(\bar{m}_{\alpha^{*},M},\bar{w}_{\alpha^{*},M})\in\mathcal{A}_{M} as a minimizer of eα,M=inf(m,w)𝒜M(m,w)e_{\alpha^{*},M}=\inf_{(m,w)\in\mathcal{A}_{M}}\mathcal{E}(m,w) for any fixed M(0,M)M\in(0,M^{*}), then we find MM(m¯α,M,w¯α,M)𝒜M\frac{M^{*}}{M}({\bar{m}}_{\alpha^{*},M},{\bar{w}}_{\alpha^{*},M})\in\mathcal{A}_{M^{*}} and

eα,M\displaystyle e_{\alpha^{*},M^{*}}\leq (MM(m¯α,M,w¯α,M))\displaystyle\mathcal{E}\Big{(}\frac{M^{*}}{M}(\bar{m}_{\alpha^{*},M},\bar{w}_{\alpha^{*},M})\Big{)}
=\displaystyle= MM[CLn|w¯α,Mm¯α,M|γm¯α,M𝑑x+nV(x)m¯α,M𝑑x12(MM)nm¯α,M(x)(K(nγ)m¯α,M)(x)𝑑x]\displaystyle\frac{M^{*}}{M}\bigg{[}C_{L}\int_{\mathbb{R}^{n}}\Big{|}\frac{\bar{w}_{\alpha^{*},M}}{\bar{m}_{\alpha^{*},M}}\Big{|}^{\gamma^{\prime}}{\bar{m}}_{\alpha^{*},M}\,dx+\int_{\mathbb{R}^{n}}V(x)\bar{m}_{\alpha^{*},M}\,dx-\frac{1}{2}\left(\frac{M^{*}}{M}\right)\int_{\mathbb{R}^{n}}{\bar{m}}_{\alpha^{*},M}(x)\big{(}K_{(n-\gamma^{\prime})}\ast\bar{m}_{\alpha^{*},M}\big{)}(x)\,dx\bigg{]}
\displaystyle\leq MM(m¯α,M,w¯α,M)=MMeα,M,M<M.\displaystyle\frac{M^{*}}{M}\mathcal{E}(\bar{m}_{\alpha^{*},M},\bar{w}_{\alpha^{*},M})=\frac{M^{*}}{M}e_{\alpha^{*},M},\ \ \forall M<M^{*}.

It follows that

eα,Mlim infMMMMeα,M=limMMeα,M.\displaystyle e_{\alpha^{*},M^{*}}\leq\liminf_{M\nearrow M^{*}}\frac{M^{*}}{M}e_{\alpha^{*},M}=\lim_{M\nearrow M^{*}}e_{\alpha^{*},M}. (4.54)

Combining (4.53) with (4.54), one has

limMMeα,M=eα,M0.\displaystyle\lim_{M\nearrow M^{*}}e_{\alpha^{*},M}=e_{\alpha^{*},M^{*}}\geq 0. (4.55)

In light of assumptions (V1) and (V2) stated in Subsection 1.2 for potential VV, we set M=MM=M^{*} in (4) to get

eα,M(mt,wt)=MV(x0)+ot(1)0, if V(x0)=0 and t+.\displaystyle e_{\alpha^{*},M^{*}}\leq\mathcal{E}(m_{*}^{t},w_{*}^{t})=M^{*}V(x_{0})+o_{t}(1)\rightarrow 0,\text{~{}if~{}}V(x_{0})=0\text{~{}and~{}}t\rightarrow+\infty.

Hence eα,M0e_{\alpha^{*},M^{*}}\leq 0, which together with (4.55) implies (4.48).

Now, we focus on the proof Conclusion (iii). If conclusion (iii) is not true, then we assume that eα,Me_{\alpha^{*},M^{*}} has a minimizer (m^,w^)𝒜M(\hat{m},\hat{w})\in\mathcal{A}_{M^{*}}. By using (4.48), we further obtain

0=(m^,w^)=nCL|w^m^|γm^𝑑x+nV(x)m^𝑑x12nm^(x)(K(nγ)m^)(x)𝑑x0.\displaystyle 0=\mathcal{E}(\hat{m},\hat{w})=\int_{\mathbb{R}^{n}}C_{L}\Big{|}\frac{\hat{w}}{\hat{m}}\Big{|}^{\gamma^{\prime}}\hat{m}\,dx+\int_{\mathbb{R}^{n}}V(x)\hat{m}\,dx-\frac{1}{2}\int_{\mathbb{R}^{n}}\hat{m}(x)\big{(}K_{(n-\gamma^{\prime})}\ast\hat{m}\big{)}(x)\,dx\geq 0.

Combining this with (4.3), one gets

CLn|w^m^|γm^𝑑x=12nm^(x)(K(nγ)m^)(x)𝑑x and nV(x)m^𝑑x=0,\displaystyle C_{L}\int_{\mathbb{R}^{n}}\Big{|}\frac{\hat{w}}{\hat{m}}\Big{|}^{\gamma^{\prime}}\hat{m}\,dx=\frac{1}{2}\int_{\mathbb{R}^{n}}\hat{m}(x)\big{(}K_{(n-\gamma^{\prime})}\ast\hat{m}\big{)}(x)\,dx\text{~{}and~{}}\int_{\mathbb{R}^{n}}V(x)\hat{m}\,dx=0, (4.56)

which implies suppV(x)supp m^=.\text{supp}V(x)\cap\text{supp~{}}\hat{m}=\emptyset. Whereas, with the assumption (1.25c) and the fact γ<n,\gamma^{\prime}<n, we have suppV=n.\text{supp}V=\mathbb{R}^{n}. It follows that m^=0\hat{m}=0 a.e., which is a contradiction. Consequently, we complete the proof of Conclusion (iii). ∎

Theorem 1.2 implies that when the potential VV satisfies some mild assumptions given by (V1), (V2) and (V3) stated in Section 1, system (1.4) admits the ground states only when M<MM<M^{*}, where MM^{*} is explicitly shown in Theorem 1.1 and has a strong connection with the existence of ground states to the potential-free nonlocal Mean-field Games system. In the next section, we shall discuss the asymptotic behaviors of ground states to problem (1.4) as MM.M\nearrow M^{*}.

5 Asymptotics of Ground States as MMM\nearrow M^{*}

This section is devoted to the proof of Theorem 1.3 and Theorem 1.4. More precisely, we shall describe the asymptotic profile of least energy solutions to (1.11) as MM.M\nearrow M^{*}.

5.1 Basic Blow-up Behaviors

In this subsection, we analyze the basic asymptotic behaviors of ground states to (1.11) as MMM\nearrow M^{*} and prove Theorem 1.3.

Proof of Theorem 1.3:

Proof.

To prove Conclusion (i), we perform the blow-up argument and assume

lim supMMn|wMmM|γmM𝑑x<+.\limsup_{M\nearrow M^{*}}\int_{\mathbb{R}^{n}}\bigg{|}\frac{w_{M}}{m_{M}}\bigg{|}^{\gamma^{\prime}}m_{M}\,dx<+\infty.

Then we utilize Lemma 2.5 to get

lim supMMmMW1,q^(n),lim supMMwMLq^(n),lim supMMwML1(n)<+.\displaystyle\limsup_{M\nearrow M^{*}}\|m_{M}\|_{W^{1,\hat{q}}(\mathbb{R}^{n})},~{}\limsup_{M\nearrow M^{*}}\|w_{M}\|_{L^{\hat{q}}(\mathbb{R}^{n})},~{}\limsup_{M\nearrow M^{*}}\|w_{M}\|_{L^{1}(\mathbb{R}^{n})}<+\infty. (5.1)

Consequently, we have there exists (m,w)W1,q^(n)×Lq^(n)(m,w)\in W^{1,\hat{q}}(\mathbb{R}^{n})\times L^{\hat{q}}(\mathbb{R}^{n}) such that

mMm in W1,q^(n) and wMw in Lq^(n) as MM.\displaystyle m_{M}\rightharpoonup m\text{~{}in~{}}W^{1,\hat{q}}(\mathbb{R}^{n})~{}\text{ and }~{}w_{M}\rightharpoonup w\text{~{}in~{}}L^{\hat{q}}(\mathbb{R}^{n})~{}\text{ as }M\nearrow M^{*}. (5.2)

Now, we prove (m,w)𝒦M(m,w)\in\mathcal{K}_{M^{*}} given by (1.1). Indeed, noting (5.1), we have

lim supMMnV(x)mM𝑑x<+.\displaystyle\limsup_{M\nearrow M^{*}}\int_{\mathbb{R}^{n}}V(x)m_{M}\,dx<+\infty. (5.3)

By using the assumptions (V1), (V2) and (V3) satisfied by VV, we conclude from (5.2), (5.3) and Lemma 4.1 that

mMm in L1(n)L2n2nγ(n),as MM,\displaystyle m_{M}\rightarrow m\text{~{}in~{}}L^{1}(\mathbb{R}^{n})\cap L^{\frac{2n}{2n-\gamma^{\prime}}}(\mathbb{R}^{n}),~{}\text{as $M\nearrow M^{*}$}, (5.4)

which implies nm𝑑x=M.\int_{\mathbb{R}^{n}}m\,dx=M^{*}. Moreover, thanks to (5.2), one gets Δm=w\Delta m=\nabla\cdot w weakly. It follows that

n|w|𝑑x=n|w||m|(γ1)γ|m|(γ1)γ𝑑x(n|m||wm|γ𝑑x)1γ(nm𝑑x)γ1γ<+,\int_{\mathbb{R}^{n}}|w|dx=\int_{\mathbb{R}^{n}}|w||m|^{-\frac{(\gamma^{\prime}-1)}{\gamma^{\prime}}}|m|^{\frac{(\gamma^{\prime}-1)}{\gamma^{\prime}}}dx\leq\left(\int_{\mathbb{R}^{n}}|m|\Big{|}\frac{w}{m}\Big{|}^{\gamma^{\prime}}\,dx\right)^{\frac{1}{\gamma^{\prime}}}\left(\int_{\mathbb{R}^{n}}m\,dx\right)^{\frac{\gamma^{\prime}-1}{\gamma^{\prime}}}<+\infty,

which implies wL1(n)w\in L^{1}(\mathbb{R}^{n}). Hence, we obtain (m,w)𝒦M(m,w)\in\mathcal{K}_{M^{*}} and further lim infMM(mM,wM)(m,w)\liminf\limits_{M\nearrow M^{*}}\mathcal{E}(m_{M},w_{M})\geq\mathcal{E}(m,w) due to (5.2) and (5.4). Moreover, one has from (4.48) that

eα,M(m,w)eα,M.e_{\alpha^{*},M^{*}}\geq\mathcal{E}(m,w)\geq e_{\alpha^{*},M^{*}}.

Therefore, (m,w)(m,w) is a minimizer of eα,M,e_{\alpha^{*},M^{*}}, which yields a contradiction to Conclusion (iii) in Theorem 1.2. This finishes the proof of Conclusion (i).

(ii). Note that

εM=ε:=(CLn|wMmM|γmM𝑑x)1γ0 as MM.\displaystyle{\varepsilon}_{M}={\varepsilon}:=\Big{(}C_{L}\int_{\mathbb{R}^{n}}\bigg{|}\frac{w_{M}}{m_{M}}\bigg{|}^{\gamma^{\prime}}m_{M}\,dx\Big{)}^{-\frac{1}{\gamma^{\prime}}}\rightarrow 0\text{~{}as~{}}M\nearrow M^{*}.

As stated in Conclusion (i) of Theorem 1.2, we have each uMC2(n)u_{M}\in C^{2}(\mathbb{R}^{n}) is bounded from below and satisfies lim|x|+uM(x)=+\lim\limits_{|x|\rightarrow+\infty}u_{M}(x)=+\infty. Hence, there exists xεnx_{\varepsilon}\in\mathbb{R}^{n} such that uM(xε)=infxnuM(x)u_{M}(x_{\varepsilon})=\inf\limits_{x\in\mathbb{R}^{n}}u_{M}(x), which indicates 0=uε(0)=infxnuε(x)0=u_{\varepsilon}(0)=\inf\limits_{x\in\mathbb{R}^{n}}u_{\varepsilon}(x) thanks to the definition given in (1.40).

In light of (1.38) and (1.40), we find that (uε,mε,wε)(u_{\varepsilon},m_{\varepsilon},w_{\varepsilon}) satisfies the following system

{Δuε+CH|uε|γ+λMεγ=(K(nγ)mε)(x)+εγV(εx+xε),ΔmεCHγ(mε|uε|γ2uε)=Δmε+wε=0,nmε𝑑x=M.\displaystyle\left\{\begin{array}[]{ll}-\Delta u_{\varepsilon}+C_{H}|\nabla u_{\varepsilon}|^{\gamma}+\lambda_{M}\varepsilon^{\gamma^{\prime}}=-\big{(}K_{(n-\gamma^{\prime})}\ast m_{\varepsilon}\big{)}(x)+\varepsilon^{\gamma^{\prime}}V(\varepsilon x+x_{\varepsilon}),\\ -\Delta m_{\varepsilon}-C_{H}\gamma\nabla\cdot(m_{\varepsilon}|\nabla u_{\varepsilon}|^{\gamma-2}\nabla u_{\varepsilon})=-\Delta m_{\varepsilon}+\nabla\cdot w_{\varepsilon}=0,\\ \int_{\mathbb{R}^{n}}m_{\varepsilon}\,dx=M.\end{array}\right. (5.8)

Collecting (1.39), (4.3) and (4.48), one gets

CLn|wεmε|γmε𝑑x=εγCLn|wMmM|γmM𝑑x1,\displaystyle C_{L}\int_{\mathbb{R}^{n}}\bigg{|}\frac{w_{\varepsilon}}{m_{\varepsilon}}\bigg{|}^{\gamma^{\prime}}m_{\varepsilon}\,dx=\varepsilon^{\gamma^{\prime}}C_{L}\int_{\mathbb{R}^{n}}\bigg{|}\frac{w_{M}}{m_{M}}\bigg{|}^{\gamma^{\prime}}m_{M}\,dx\equiv 1, (5.9)
nmε(x)(K(nγ)mε)(x)𝑑x=εγnmM(x)(K(nγ)mM)(x)𝑑x2,\displaystyle\int_{\mathbb{R}^{n}}m_{\varepsilon}(x)\big{(}K_{(n-\gamma^{\prime})}\ast m_{\varepsilon}\big{)}(x)\,dx=\varepsilon^{\gamma^{\prime}}\int_{\mathbb{R}^{n}}m_{M}(x)\big{(}K_{(n-\gamma^{\prime})}\ast m_{M}\big{)}(x)\,dx\rightarrow 2, (5.10)

and

nV(εx+xε)mε𝑑x=nV(x)mM𝑑x0 as MM.\displaystyle\int_{\mathbb{R}^{n}}V(\varepsilon x+x_{\varepsilon})m_{\varepsilon}\,dx=\int_{\mathbb{R}^{n}}V(x)m_{M}\,dx\rightarrow 0\text{~{}as~{}}M\nearrow M^{*}. (5.11)

Following the similar argument employed in the derivation of (4.31), we utilize (5.8) and (5.9) to obtain

MλM=(mM,wM)12nmM(x)(K(nγ)mM)(x)𝑑x=o(1)12εγnmε(x)(K(nγ)mε)(x)𝑑x,\displaystyle M\lambda_{M}=\mathcal{E}(m_{M},w_{M})-\frac{1}{2}\int_{\mathbb{R}^{n}}m_{M}(x)\big{(}K_{(n-\gamma^{\prime})}\ast m_{M}\big{)}(x)\,dx=o(1)-\frac{1}{2}\varepsilon^{-\gamma^{\prime}}\int_{\mathbb{R}^{n}}m_{\varepsilon}(x)\big{(}K_{(n-\gamma^{\prime})}\ast m_{\varepsilon}\big{)}(x)\,dx,

which implies

λMεγ1M as MM.\displaystyle\lambda_{M}\varepsilon^{\gamma^{\prime}}\rightarrow-\frac{1}{M^{*}}\text{~{}as~{}}M\nearrow M^{*}. (5.12)

We apply the maximum principle to the uu-equation in (5.8), then deduce that

λMεγ(K(nγ)mε)(0)+εγV(xε)(K(nγ)mε)(0),\displaystyle\lambda_{M}\varepsilon^{\gamma^{\prime}}\geq-\big{(}K_{(n-\gamma^{\prime})}\ast m_{\varepsilon}\big{)}(0)+\varepsilon^{\gamma^{\prime}}V(x_{\varepsilon})\geq-\big{(}K_{(n-\gamma^{\prime})}\ast m_{\varepsilon}\big{)}(0), (5.13)

which indicates

(K(nγ)mε)(0)λMεγC>0.\displaystyle\big{(}K_{(n-\gamma^{\prime})}\ast m_{\varepsilon}\big{)}(0)\geq-\lambda_{M}\varepsilon^{\gamma^{\prime}}\geq C>0. (5.14)

Now, we claim that there exists some constant C>0C>0 such that

εγV(xε)C.\displaystyle\varepsilon^{\gamma^{\prime}}V(x_{\varepsilon})\leq C. (5.15)

If this is not the case, one can find some subsequence εl0\varepsilon_{l}\rightarrow 0 such that εlγV(xεl)+\varepsilon_{l}^{\gamma^{\prime}}V(x_{\varepsilon_{l}})\rightarrow+\infty. Then, with the aid of (5.13), one has

(Kαmεl)(0)εlγV(xεl)C,\displaystyle\frac{\big{(}K_{\alpha^{*}}\ast m_{\varepsilon_{l}}\big{)}(0)}{\varepsilon_{l}^{\gamma^{\prime}}V(x_{\varepsilon_{l}})}\geq C, (5.16)

where C>0C>0 is some constant independent of εl.\varepsilon_{l}. Define

vl(x):=alγ2ul(x0+alx),μl(x):=alnml(x0+alx),al:=1εlV(xεl)1γ,\displaystyle v_{l}(x):=a_{l}^{\gamma^{\prime}-2}u_{l}(x_{0}+a_{l}x),~{}~{}\mu_{l}(x):=a_{l}^{n}m_{l}(x_{0}+a_{l}x),~{}~{}a_{l}:=\frac{1}{\varepsilon_{l}V(x_{\varepsilon_{l}})^{\frac{1}{\gamma^{\prime}}}}, (5.17)

then one has

alγ=1εlγV(xεl)0,alγεlγV(xεl)=1.\displaystyle a_{l}^{\gamma^{\prime}}=\frac{1}{\varepsilon_{l}^{\gamma^{\prime}}V(x_{\varepsilon_{l}})}\rightarrow 0,~{}~{}a_{l}^{\gamma^{\prime}}\varepsilon_{l}^{\gamma^{\prime}}V(x_{\varepsilon_{l}})=1.

By substituting (5.17) into (5.8), we find

{Δvl+CH|vl|γ+alγλM=alγV(xl+alx)alγn(Kαμl),xn,Δμl+CHγ(|vl|γ2vlμl)=0,xn.\displaystyle\left\{\begin{array}[]{ll}-\Delta v_{l}+C_{H}|\nabla v_{l}|^{\gamma}+a_{l}^{\gamma^{\prime}}\lambda_{M}=a_{l}^{\gamma^{\prime}}V(x_{l}+a_{l}x)-a^{\gamma^{\prime}-n}_{l}\Big{(}K_{\alpha^{*}}\ast\mu_{l}\Big{)},&x\in\mathbb{R}^{n},\\ \Delta\mu_{l}+C_{H}\gamma\nabla\cdot(|\nabla v_{l}|^{\gamma-2}\nabla v_{l}\mu_{l})=0,&x\in\mathbb{R}^{n}.\end{array}\right. (5.20)

By using the assumption (1.25b), one gets

alγεlγV(alεlx+xεl)=V(aεlεlx+xεl)V(xεl)C,\displaystyle a_{l}^{\gamma^{\prime}}\varepsilon^{\gamma^{\prime}}_{l}V(a_{l}\varepsilon_{l}x+x_{\varepsilon_{l}})=\frac{V(a_{\varepsilon_{l}}\varepsilon_{l}x+x_{\varepsilon_{l}})}{V(x_{\varepsilon_{l}})}\leq C,

where C>0C>0 is some constant independent of l.l. Noting that

μlnαn+αL1+n+αnα(BR(0))1+n+αnα=alγn2nγmεlL2n2nγ(BRal(xl))0 as l+,\displaystyle\|\mu_{l}^{\frac{n-\alpha^{*}}{n+\alpha^{*}}}\|^{1+\frac{n+\alpha^{*}}{n-\alpha^{*}}}_{L^{1+\frac{n+\alpha^{*}}{n-\alpha^{*}}}(B_{R}(0))}=a_{l}^{\frac{\gamma^{\prime}n}{2n-\gamma^{\prime}}}\|m_{\varepsilon_{l}}\|_{L^{\frac{2n}{2n-\gamma^{\prime}}}(B_{Ra_{l}}(x_{l}))}\rightarrow 0\text{~{}as~{}}l\rightarrow+\infty,

we utilize the maximal regularity shown in Lemma 2.1 to obtain

|vl|γL1+n+αnα(BR/2)C,\displaystyle\||\nabla v_{l}|^{\gamma}\|_{L^{1+\frac{n+\alpha^{*}}{n-\alpha^{*}}}(B_{R/2})}\leq C,

where R>0R>0 and C>0C>0 are some constants. Focusing on the mm-equation of (5.20), we similarly apply the standard elliptic regularity estimates (See Theorem 1.6.5 in [6]) to obtain μlC0,θ(BR/4(0))\mu_{l}\in C^{0,\theta}(B_{R/4}(0)) with θ(0,1)\theta\in(0,1) independent of ll. By a direct calculation, we conclude from (5.16) that

(Kαμl)(0)=alr(Kαml)(0)=(Kαμl)(0)εlrV(xεl)C>0.\displaystyle\Big{(}K_{\alpha^{*}}\ast\mu_{l}\Big{)}(0)=a_{l}^{r}\Big{(}K_{\alpha^{*}}\ast m_{l}\Big{)}(0)=\frac{\Big{(}K_{\alpha^{*}}\ast\mu_{l}\Big{)}(0)}{\varepsilon_{l}^{r}V(x_{\varepsilon_{l}})}\geq C>0. (5.21)

This together with the Hölder’s continuity of μl\mu_{l} implies that

BR/4(0)μl(x)𝑑xC>0,\displaystyle\int_{B_{R/4}(0)}\mu_{l}(x)\,dx\geq C>0, (5.22)

where R>0R>0 sufficiently large and independent of ll. In light of εlγV(xεl)+\varepsilon^{\gamma^{\prime}}_{l}V(x_{\varepsilon_{l}})\rightarrow+\infty, we have the fact that |xεl|+.|x_{\varepsilon_{l}}|\rightarrow+\infty. As a consequence, there exists δ>0\delta>0 such that V(xεl)2δ.V(x_{\varepsilon_{l}})\geq 2\delta. Then It follows from (5.22) that

nV(εlx+xεl)mεl(x)𝑑x\displaystyle\int_{\mathbb{R}^{n}}V(\varepsilon_{l}x+x_{\varepsilon_{l}})m_{\varepsilon_{l}}(x)\,dx
=\displaystyle= nV(εlaεlx+xεl)μl𝑑xδBR(0)μl𝑑xCδ>0,\displaystyle\int_{\mathbb{R}^{n}}V(\varepsilon_{l}a_{\varepsilon_{l}}x+x_{\varepsilon_{l}})\mu_{l}\,dx\geq\delta\int_{B_{{R}}(0)}\mu_{l}\,dx\geq C{\delta}>0,

as εl0.\varepsilon_{l}\rightarrow 0. Whereas,

nV(εx+xε)mε(x)𝑑x0asε0,\displaystyle\int_{\mathbb{R}^{n}}V(\varepsilon x+x_{\varepsilon})m_{\varepsilon}(x)\,dx\rightarrow 0~{}\text{as}~{}\varepsilon\rightarrow 0,

which reaches a contradiction. This completes the proof of our claim (5.15).

Moreover, since VV satisfies (1.25b), one further obtains for R>0R>0 large enough,

εγV(εx+xε)CR<+, for all |x|4R,\displaystyle\varepsilon^{\gamma^{\prime}}V(\varepsilon x+x_{\varepsilon})\leq C_{R}<+\infty,\text{ for all }|x|\leq 4R, (5.23)

where constant CR>0C_{R}>0 depends on RR and is independent of ε.\varepsilon.

Similarly as discussed in the proof of Theorem 1.2, we estimate uε\nabla u_{\varepsilon} and rewrite the uu-equation of (5.8) as

Δuε=CH|uε|γ+gε(x) with gε(x):=λMεγ+εγV(xε+εx)(Kαmε)(x).-\Delta u_{\varepsilon}=-C_{H}|\nabla u_{\varepsilon}|^{\gamma}+g_{\varepsilon}(x)\ \text{ with }g_{\varepsilon}(x):=-\lambda_{M}\varepsilon^{\gamma^{\prime}}+\varepsilon^{\gamma^{\prime}}V(x_{\varepsilon}+\varepsilon x)-\big{(}K_{\alpha^{*}}\ast m_{\varepsilon}\big{)}(x). (5.24)

Noting that (Kαmε)L1+n+αnα(n)\big{(}K_{\alpha^{*}}\ast m_{\varepsilon}\big{)}\in L^{1+\frac{n+\alpha^{*}}{n-\alpha^{*}}}(\mathbb{R}^{n}), we utilize Lemma 2.1 to get |uε|γLloc1+n+αnα(n)|\nabla u_{\varepsilon}|^{\gamma}\in L_{\rm loc}^{1+\frac{n+\alpha^{*}}{n-\alpha^{*}}}(\mathbb{R}^{n}), i.e. |uε|γ1Lloc(1+n+αnα)γ(n)|\nabla u_{\varepsilon}|^{\gamma-1}\in L_{\rm loc}^{\big{(}1+\frac{n+\alpha^{*}}{n-\alpha^{*}}\big{)}\gamma^{\prime}}(\mathbb{R}^{n}). By using Lemma 2.6, we further obtain that mCloc0,θ(n)m\in C^{0,\theta}_{\text{loc}}(\mathbb{R}^{n}) for some θ(0,1)\theta\in(0,1) since mm satisfies the second equation in (5.8).

uεC2,θ(BR(0))C<.\displaystyle\|u_{\varepsilon}\|_{C^{2,\theta}(B_{R}(0))}\leq C<\infty. (5.25)

In light of (5.14), we have from (5.25) that there exists a constant R0(0,1)R_{0}\in(0,1) such that

mε(x)C>0,|x|<R0.\displaystyle m_{\varepsilon}(x)\geq C>0,~{}~{}\forall|x|<R_{0}. (5.26)

Now, we claim that up to a subsequence,

limε0xε=x0 with V(x0)=0.\lim_{\varepsilon\to 0}x_{\varepsilon}=x_{0}\ \text{ with }~{}V(x_{0})=0. (5.27)

If not, one has either |xε|+|x_{\varepsilon}|\rightarrow+\infty or xεx0x_{\varepsilon}\rightarrow x_{0} with V(x0)>0.V(x_{0})>0. In the two cases, we both have limxε0V(εx+xε)A\lim\limits_{x_{\varepsilon}\to 0}V(\varepsilon x+x_{\varepsilon})\geq A a.e. in n\mathbb{R}^{n} for some A>0A>0. It then follows from (5.26) that

limε0nV(εx+xε)mε𝑑xA2BR0(0)mε(x)𝑑xAC2|BR0(0)|,\displaystyle\lim_{\varepsilon\to 0}\int_{\mathbb{R}^{n}}V(\varepsilon x+x_{\varepsilon})m_{\varepsilon}\,dx\geq\frac{A}{2}\int_{B_{R_{0}}(0)}m_{\varepsilon}(x)\,dx\geq\frac{AC}{2}|B_{R_{0}}(0)|,

which contradicts (5.11). Therefore, we find (5.27) holds.

By using (5.9), we find there exists (m0,w0)W1,q^(n)×(L1(n)Lq^(n))(m_{0},w_{0})\in W^{1,\hat{q}}(\mathbb{R}^{n})\times\big{(}L^{1}(\mathbb{R}^{n})\cap L^{\hat{q}}(\mathbb{R}^{n})\big{)} such that

(mε,wε)(m0,w0)in W1,q^(n)×Lq^(n) as ε0,\displaystyle(m_{\varepsilon},w_{\varepsilon})\rightharpoonup(m_{0},w_{0})\ \ \text{in~{}}W^{1,\hat{q}}(\mathbb{R}^{n})\times L^{\hat{q}}(\mathbb{R}^{n})\ \text{ as }\varepsilon\to 0, (5.28)

where m00m_{0}\not\equiv 0 thanks to (5.26) and q^\hat{q} is given by (1.24). Furthermore, invoking (5.25), one has uεu0u_{\varepsilon}\rightarrow u_{0} in Cloc2(n)C^{2}_{\rm loc}(\mathbb{R}^{n}). Moreover, combining (5.8) with (5.12), we obtain (m0,u0)(m_{0},u_{0}) satisfies

{Δu0+CH|u0|γ1M=K(nγ)m0,Δm0=CHγ(m0|u0|γ2u0)=w0,0<nm0𝑑xM,w0=CHm0|u0|γ2u0,\displaystyle\left\{\begin{array}[]{ll}-\Delta u_{0}+C_{H}|\nabla u_{0}|^{\gamma}-\frac{1}{M^{*}}=-K_{(n-\gamma^{\prime})}\ast m_{0},\\ -\Delta m_{0}=-C_{H}\gamma\nabla\cdot(m_{0}|\nabla u_{0}|^{\gamma-2}\nabla u_{0})=-\nabla\cdot w_{0},\\ 0<\int_{\mathbb{R}^{n}}m_{0}\,dx\leq M^{*},~{}~{}w_{0}=-C_{H}m_{0}|\nabla u_{0}|^{\gamma-2}\nabla u_{0},\end{array}\right. (5.32)

where we have followed the procedure performed in the proof of (3.86) shown in Section 3. In particular, we have used Lemma 2.8 to obtain that (m0,w0)(m_{0},w_{0}) is a minimizer of (1.28) and nm0𝑑x=M\int_{\mathbb{R}^{n}}m_{0}\,dx=M^{*}. Thus, we have from (5.32) that (u0,m0,w0)(u_{0},m_{0},w_{0}) satisfies (1.33). On the other hand, we obtain mεm0m_{\varepsilon}\to m_{0} in L1(n)L^{1}(\mathbb{R}^{n}), and then with the aid of (5.28), one finds

mεm0in Lp(n),p[1,q^) as ε0,\displaystyle m_{\varepsilon}\to m_{0}\ \ \text{in~{}}L^{p}(\mathbb{R}^{n}),~{}\forall~{}p\in[1,{\hat{q}}^{*})\ \text{ as }\varepsilon\to 0,

which indicates that (1.41) holds.

Finally, we prove that (1.43) holds when (1.42) is imposed on VV. To this end, we argue by contradiction and assume that, then, up to a subsequence,

lim infε0|x¯εxε|ε=+.\displaystyle\liminf_{\varepsilon\rightarrow 0}\frac{|\bar{x}_{\varepsilon}-x_{\varepsilon}|}{\varepsilon}=+\infty. (5.33)

Define

{m¯ε(x):=εnmM(εx+x¯ε)=mε(x+x¯εxεε),w¯ε(x):=εn+1wM(εx+x¯ε)=wε(x+x¯εxεε),u¯ε(x):=ε2γγ1uM(εx+x¯ε)=uε(x+x¯εxεε).\displaystyle\left\{\begin{array}[]{ll}\bar{m}_{\varepsilon}(x):=\varepsilon^{n}m_{M}(\varepsilon x+\bar{x}_{\varepsilon})=m_{\varepsilon}\Big{(}x+\frac{\bar{x}_{\varepsilon}-x_{\varepsilon}}{\varepsilon}\Big{)},\\ \bar{w}_{\varepsilon}(x):=\varepsilon^{n+1}w_{M}(\varepsilon x+\bar{x}_{\varepsilon})=w_{\varepsilon}\Big{(}x+\frac{\bar{x}_{\varepsilon}-x_{\varepsilon}}{\varepsilon}\Big{)},\\ \bar{u}_{\varepsilon}(x):=\varepsilon^{\frac{2-\gamma}{\gamma-1}}u_{M}(\varepsilon x+\bar{x}_{\varepsilon})=u_{\varepsilon}\Big{(}x+\frac{\bar{x}_{\varepsilon}-x_{\varepsilon}}{\varepsilon}\Big{)}.\end{array}\right. (5.37)

Now, we claim that R0>0\exists R_{0}>0 and C>0C>0 independent of ε\varepsilon such that

m¯ε(x)C>0,|x|<R0.\displaystyle\bar{m}_{\varepsilon}(x)\geq C>0,\ \ \forall~{}|x|<R_{0}. (5.38)

Invoking (5.37), we have (5.38) is equivalent to

mε(x)C>0,|xx¯εxεε|<R0.\displaystyle m_{\varepsilon}(x)\geq C>0,\ \ \forall~{}\Big{|}x-\frac{\bar{x}_{\varepsilon}-x_{\varepsilon}}{\varepsilon}\Big{|}<R_{0}. (5.39)

In light of (5.14), we find

m¯ε(0)=m¯εL(n)=mεL(n)>C>0.\bar{m}_{\varepsilon}(0)=\|\bar{m}_{\varepsilon}\|_{L^{\infty}(\mathbb{R}^{n})}=\|m_{\varepsilon}\|_{L^{\infty}(\mathbb{R}^{n})}>C>0. (5.40)

To show (5.39), we have from the first equation in (5.8) that u¯ε\bar{u}_{\varepsilon} satisfies

Δu¯ε+CH|u¯ε|γ=g¯ε(x):=λMεγ(K(nγ)m¯ε)(x)+εγV(εx+x¯ε).\displaystyle-\Delta\bar{u}_{\varepsilon}+C_{H}|\nabla\bar{u}_{\varepsilon}|^{\gamma}=\bar{g}_{\varepsilon}(x):=-\lambda_{M}\varepsilon^{\gamma^{\prime}}-\big{(}K_{(n-\gamma^{\prime})}\ast\bar{m}_{\varepsilon}\big{)}(x)+\varepsilon^{\gamma^{\prime}}V(\varepsilon x+\bar{x}_{\varepsilon}). (5.41)

Following the argument shown in [10, Theorem 4.1], we consider the following two cases:

Case 1: Assume that there exists some constant C>0C>0 independent of ε\varepsilon such that x¯ε\bar{x}_{\varepsilon} satisfies

lim supε0εγV(x¯ε)C<+.\displaystyle\limsup_{\varepsilon\rightarrow 0}\varepsilon^{\gamma^{\prime}}V(\bar{x}_{\varepsilon})\leq C<+\infty.

Then thanks to (5.40), we follow the same argument performed in the derivation of (5.24), (5.25) and (5.26) to obtain the claim (5.38).

Case 2: Suppose that x¯ε\bar{x}_{\varepsilon} satisfies

lim infε0εγV(x¯ε)=+.\displaystyle\liminf_{\varepsilon\rightarrow 0}\varepsilon^{\gamma^{\prime}}V(\bar{x}_{\varepsilon})=+\infty. (5.42)

Define

m~(x)=εnmM(εx)=mε(xxεε),u~(x)=ε2γγ1uM(εx),w~(x)=εn+1wM(εx),\displaystyle\tilde{m}(x)=\varepsilon^{n}m_{M}(\varepsilon x)=m_{\varepsilon}\bigg{(}x-\frac{x_{\varepsilon}}{\varepsilon}\bigg{)},\ \tilde{u}(x)=\varepsilon^{\frac{2-\gamma}{\gamma-1}}u_{M}(\varepsilon x),\tilde{w}(x)=\varepsilon^{n+1}w_{M}(\varepsilon x), (5.43)

then obtain from (5.8) that (m~,u~,w~)(\tilde{m},\tilde{u},\tilde{w}) satisfies

{Δu~+CH|u~|γ+λMεγ=εγV(εx)K(nγ)m~ε,xn,Δm~CHγ(m~|u~|γ2u~)=0,xn,nm~ε𝑑x=MM,w~ε=CHγm~|u~|γ2u~.\displaystyle\left\{\begin{array}[]{ll}-\Delta\tilde{u}+C_{H}|\nabla\tilde{u}|^{\gamma}+\lambda_{M}\varepsilon^{\gamma^{\prime}}=\varepsilon^{\gamma^{\prime}}V(\varepsilon x)-K_{(n-\gamma^{\prime})}\ast\tilde{m}_{\varepsilon},&x\in\mathbb{R}^{n},\\ -\Delta\tilde{m}-C_{H}\gamma\nabla\cdot(\tilde{m}|\nabla\tilde{u}|^{\gamma-2}\nabla\tilde{u})=0,&x\in\mathbb{R}^{n},\\ \int_{\mathbb{R}^{n}}\tilde{m}_{\varepsilon}\,dx=M\nearrow M^{*},\ \ \tilde{w}_{\varepsilon}=-C_{H}\gamma\tilde{m}|\nabla\tilde{u}|^{\gamma-2}\nabla\tilde{u}.\end{array}\right. (5.47)

Since VV satisfies (1.42), we utilize Lemma 2.2 to get

|u~ε|C(1+σε1γ|x|bγ),σε:=εγ+b.\displaystyle|\nabla\tilde{u}_{\varepsilon}|\leq C\big{(}1+\sigma_{\varepsilon}^{\frac{1}{\gamma}}|x|^{\frac{b}{\gamma}}\big{)},\ \ \sigma_{\varepsilon}:=\varepsilon^{\gamma^{\prime}+b}. (5.48)

Denote yε:=xεεy_{\varepsilon}:=\frac{x_{\varepsilon}}{\varepsilon} and y¯ε:=x¯εε\bar{y}_{\varepsilon}:=\frac{\bar{x}_{\varepsilon}}{\varepsilon}, which are the minimum and maximum points of u~ε(x)\tilde{u}_{\varepsilon}(x) and m~ε(x),\tilde{m}_{\varepsilon}(x), respectively. With the aid of (5.27), we obtain |yε|Cε1.|y_{\varepsilon}|\leq C\varepsilon^{-1}. Then, we obtain from (5.48) that

|u~ε(0)||u~ε(yε)|+|yε|sup|y||yε||u~ε(y)|1+Cε1+Cε1σε1γ|yε|bγ1+Cε1.\displaystyle|\tilde{u}_{\varepsilon}(0)|\leq|\tilde{u}_{\varepsilon}(y_{\varepsilon})|+|y_{\varepsilon}|\sup_{|y|\leq|y_{\varepsilon}|}|\nabla\tilde{u}_{\varepsilon}(y)|\leq 1+C\varepsilon^{-1}+C\varepsilon^{-1}\sigma_{\varepsilon}^{\frac{1}{\gamma}}|y_{\varepsilon}|^{\frac{b}{\gamma}}\leq 1+C\varepsilon^{-1}. (5.49)

As a consequence,

u~ε(x)u~ε(0)+|x|sup|uε(x)|1+Cε1+σε1γ|x|bγ+1.\displaystyle\tilde{u}_{\varepsilon}(x)\leq\tilde{u}_{\varepsilon}(0)+|x|\sup|\nabla u_{\varepsilon}(x)|\leq 1+C\varepsilon^{-1}+\sigma_{\varepsilon}^{\frac{1}{\gamma}}|x|^{\frac{b}{\gamma}+1}. (5.50)

Collecting (5.42), (5.49) and (5.50), we proceed the same argument shown in [10, Theorem 4.1] to get m~εC0,θ(BR(y¯ε))\tilde{m}_{\varepsilon}\in C^{0,\theta}(B_{R}(\bar{y}_{\varepsilon})) with θ(0,1)\theta\in(0,1) and R>0R>0 independent of ε.\varepsilon. Since y¯ε\bar{y}_{\varepsilon} is maximum point of m~ε(x)\tilde{m}_{\varepsilon}(x), we combine (5.40) with (5.43) to get m~ε(y¯ε)C>0.\tilde{m}_{\varepsilon}(\bar{y}_{\varepsilon})\geq C>0. Hence, we have there exists some R0>0R_{0}>0 independent of ε\varepsilon such that

m~ε(x)>C2>0,|xy¯ε|<R0.\displaystyle\tilde{m}_{\varepsilon}(x)>\frac{C}{2}>0,\forall~{}|x-\bar{y}_{\varepsilon}|<R_{0}.

Noting y¯ε=x¯εε\bar{y}_{\varepsilon}=\frac{\bar{x}_{\varepsilon}}{\varepsilon}, we find from the above estimate and (5.43) that (5.39) holds.

Thus, if the potential VV satisfies (1.42), then (5.38) and (5.39) hold. Whereas, (5.39) together with (5.33) contradicts the fact that mεm_{\varepsilon} converges strongly to m0m_{0} in L1(n)L^{1}(\mathbb{R}^{n}). As a consequence, (1.43) holds and this completes the proof of Theorem 1.3. ∎

In Theorem 1.3, we see that as MM,M\nearrow M^{*}, the ground states (mM,wM)(m_{M},w_{M}) to problem (1.22) concentrate and become localized patterns, in which the profiles are determined by (m0,w0),(m_{0},w_{0}), the minimizer to problem (3.1). We mention that with some typical expansions imposed on potential VV locally, the detailed asymptotics of ground states can be captured and we shall discuss them in Subsection 5.2.

5.2 Refined Blow-up Behaviors

In this subsection, we shall analyze the refined asymptotic profiles of the rescaled minimizer (mε,wε)(m_{\varepsilon},w_{\varepsilon}) and prove Theorem 1.4. As shown in Theorem 1.4, we assume V(x)V(x) has ll\in\mathbb{N} distinct zeros defined by {P1,,Pl}n\{P_{1},\cdots,P_{l}\}\subset\mathbb{R}^{n}; moreover, ai>0\exists a_{i}>0, qi>0q_{i}>0, d>0d>0 such that

V(x)=ai|xPi|qi+O(|xPi|qi+1), if |xPi|d.\displaystyle V(x)=a_{i}|x-P_{i}|^{q_{i}}+O(|x-P_{i}|^{q_{i}+1}),\ \ \text{~{}if~{}}|x-P_{i}|\leq d. (5.51)

Define q=max{q1,,ql}q=\max\{q_{1},\cdots,q_{l}\}, Z={Pi|qi=q,i=1,,l}Z=\{P_{i}~{}|~{}q_{i}=q,i=1,\cdots,l\} and denote

μ=min{μi|PiZ,i=1,,l} with μi=minynHi(y),Hi(y)=nai|x+y|qim0(x)𝑑x.\mu=\min\{\mu_{i}~{}|~{}P_{i}\in Z,i=1,\cdots,l\}~{}\text{ with }~{}\mu_{i}=\min\limits_{y\in\mathbb{R}^{n}}H_{i}(y),\ H_{i}(y)=\int_{\mathbb{R}^{n}}a_{i}|x+y|^{q_{i}}m_{0}(x)\,dx. (5.52)

Set Z0={Pi|PiZ and μi=μ,i=1,,l}Z_{0}=\{P_{i}~{}|~{}P_{i}\in Z\text{~{}and~{}}\mu_{i}=\mu,i=1,\cdots,l\} consisted of all weighted flattest zeros of V(x).V(x). Collecting the above notations, we first establish the precise upper bound of eα,Me_{\alpha^{*},M} as MMM\nearrow M^{*} stated as follows:

Lemma 5.1.

The eα,Me_{\alpha^{*},M}, defined by (1.26), satisfies

eα,M[1+o(1)]q+γq(qμγ)γγ+q[1MM]qγ+q,asMM.\displaystyle e_{\alpha^{*},M}\leq[1+o(1)]\frac{q+\gamma^{\prime}}{q}\bigg{(}\frac{q\mu}{\gamma^{\prime}}\bigg{)}^{\frac{\gamma^{\prime}}{\gamma^{\prime}+q}}\bigg{[}1-\frac{M}{M^{*}}\bigg{]}^{\frac{q}{\gamma^{\prime}+q}},\ \text{as}\ M\nearrow M^{*}. (5.53)
Proof.

The proof is similar as the argument shown in [14, Lemma 6.1] with slight modifications. We omit the details. ∎

In Section 5, we find (mε,wε,uε)(m_{\varepsilon},w_{\varepsilon},u_{\varepsilon}) converges to (m0,w0,u0)(m_{0},w_{0},u_{0}) in the following sense:

mεm0 in Lp(n)p[1,q^)wεw0 Lq^(n) and uεu0 in Cloc2(n),\text{$m_{\varepsilon}\rightarrow m_{0}$ in $L^{p}(\mathbb{R}^{n})\ \forall~{}p\in[1,{\hat{q}}^{*})$, $w_{\varepsilon}\rightharpoonup w_{0}$ $L^{\hat{q}}(\mathbb{R}^{n})$ and $u_{\varepsilon}\rightarrow u_{0}$ in $C^{2}_{\text{loc}}(\mathbb{R}^{n})$},

where (m0,w0)(m_{0},w_{0}) is the minimizer of Γα\Gamma_{\alpha^{*}} and correspondingly, (u0,m0,w0)(u_{0},m_{0},w_{0}) satisfies (1.33). Moreover, Lemma 2.7 and Lemma 2.8 imply δ1>0\exists\delta_{1}>0 and Cδ1>0C_{\delta_{1}}>0 such that

m0(x)Cδ1Cδ1|x|,m_{0}(x)\leq C_{\delta_{1}}C^{-\delta_{1}|x|}, (5.54)

and

CLn|w0m0|γm0𝑑x=12nm0(x)(K(nγ)m0)(x)𝑑x=1.\displaystyle C_{L}\int_{\mathbb{R}^{n}}\Big{|}\frac{w_{0}}{m_{0}}\Big{|}^{\gamma^{\prime}}m_{0}\,dx=\frac{1}{2}\int_{\mathbb{R}^{n}}m_{0}(x)\big{(}K_{(n-\gamma^{\prime})}\ast m_{0}\big{)}(x)\,dx=1. (5.55)

Next, invoking Lemma 5.1, (5.54) and (5.55), we are going to prove Theorem 1.4, which is

Proof of Theorem 1.4:

Proof.

Thanks to Theorem 1.3, we have xεPix_{\varepsilon}\rightarrow P_{i} for some 1il.1\leq i\leq l. In addition, noting that (mM,wM)(m_{M},w_{M}) is the minimizer of problem (1.26), one gets

eα,M=(mM,wM)=\displaystyle e_{\alpha^{*},M}=\mathcal{E}(m_{M},w_{M})= εγCLn|wεmε|γmε𝑑xεr2nmε(K(nγ)mε)(x)𝑑x+nV(εx+xε)mε(x)𝑑x\displaystyle\varepsilon^{-\gamma^{\prime}}C_{L}\int_{\mathbb{R}^{n}}\bigg{|}\frac{w_{\varepsilon}}{m_{\varepsilon}}\bigg{|}^{\gamma^{\prime}}m_{\varepsilon}\,dx-\frac{\varepsilon^{-r}}{2}\int_{\mathbb{R}^{n}}m_{\varepsilon}\big{(}K_{(n-\gamma^{\prime})}\ast m_{\varepsilon}\big{)}(x)\,dx+\int_{\mathbb{R}^{n}}V(\varepsilon x+x_{\varepsilon})m_{\varepsilon}(x)\,dx
\displaystyle\geq 12εγ[MM1]nmε(K(nγ)mε)(x)𝑑x+nV(εx+xε)mε(x)𝑑x.\displaystyle\frac{1}{2}\varepsilon^{-\gamma^{\prime}}\Big{[}\frac{M^{*}}{M}-1\Big{]}\int_{\mathbb{R}^{n}}m_{\varepsilon}\big{(}K_{(n-\gamma^{\prime})}\ast m_{\varepsilon}\big{)}(x)\,dx+\int_{\mathbb{R}^{n}}V(\varepsilon x+x_{\varepsilon})m_{\varepsilon}(x)\,dx. (5.56)

By the direct calculation, we obtain

nV(εx+xε)mε(x)𝑑x=εqinV(εx+xε)|εx+xεPi|qi|x+xεPiε|qimε(x)𝑑x.\displaystyle\int_{\mathbb{R}^{n}}V(\varepsilon x+x_{\varepsilon})m_{\varepsilon}(x)\,dx=\varepsilon^{q_{i}}\int_{\mathbb{R}^{n}}\frac{V(\varepsilon x+x_{\varepsilon})}{|\varepsilon x+x_{\varepsilon}-P_{i}|^{q_{i}}}\bigg{|}x+\frac{x_{\varepsilon}-P_{i}}{\varepsilon}\bigg{|}^{q_{i}}m_{\varepsilon}(x)\,dx. (5.57)

In light of xεPix_{\varepsilon}\rightarrow P_{i}, then one has

limε0V(εx+xε)|εx+xεPi|qi=limε0ai|εx+xεPi|qi+O(|εx+xεPi|qi+1)|εx+xεPi|qi=ai, a.e. in n.\displaystyle\lim_{\varepsilon\rightarrow 0}\frac{V(\varepsilon x+x_{\varepsilon})}{|\varepsilon x+x_{\varepsilon}-P_{i}|^{q_{i}}}=\lim_{\varepsilon\rightarrow 0}\frac{a_{i}|\varepsilon x+x_{\varepsilon}-P_{i}|^{q_{i}}+O(|\varepsilon x+x_{\varepsilon}-P_{i}|^{q_{i}+1})}{|\varepsilon x+x_{\varepsilon}-P_{i}|^{q_{i}}}=a_{i},\ \text{~{}a.e.~{}in~{}}\mathbb{R}^{n}. (5.58)

Now, we claim that

qi=q=max{q1,,ql} and lim supε0|xεPiε| is uniformly bounded.\displaystyle\text{$q_{i}=q=\max\{q_{1},\cdots,q_{l}\}$ and }~{}\limsup_{\varepsilon\rightarrow 0}\Big{|}\frac{x_{\varepsilon}-P_{i}}{\varepsilon}\Big{|}\text{~{}is uniformly bounded.} (5.59)

Indeed, if (5.59)\eqref{629uniformlyboundxp} is not true, then we have either qi<qq_{i}<q or up to a subsequence, limε0|xεPiε|=+\lim_{\varepsilon\rightarrow 0}\big{|}\frac{x_{\varepsilon}-P_{i}}{\varepsilon}\big{|}=+\infty. Then by using Fatou’s lemma, we conclude from (1.41), (5.57) and (5.58) that

limε0εqnV(εx+xε)mε𝑑x=limε0εqiqnV(εx+xε)|εx+xεPi|qi|x+xεPiε|qimε𝑑xβ1\displaystyle\lim_{\varepsilon\to 0}\varepsilon^{-q}\int_{\mathbb{R}^{n}}V(\varepsilon x+x_{\varepsilon})m_{\varepsilon}\,dx=\lim_{\varepsilon\to 0}\varepsilon^{q_{i}-q}\int_{\mathbb{R}^{n}}\frac{V(\varepsilon x+x_{\varepsilon})}{|\varepsilon x+x_{\varepsilon}-P_{i}|^{q_{i}}}\Big{|}x+\frac{x_{\varepsilon}-P_{i}}{\varepsilon}\Big{|}^{q_{i}}m_{\varepsilon}\,dx\geq\beta\gg 1

for any constant β1\beta\gg 1 large enough. Combining (5.9) with (5.56), one gets

eα,M\displaystyle e_{\alpha^{*},M}\geq 12εγ[MM1]nmε(K(nγ)mε)(x)𝑑x+βεq=[1+oε(1)][MM1]εγ+βεq\displaystyle\frac{1}{2}\varepsilon^{-\gamma^{\prime}}\Big{[}\frac{M^{*}}{M}-1\Big{]}\int_{\mathbb{R}^{n}}m_{\varepsilon}\big{(}K_{(n-\gamma^{\prime})}\ast m_{\varepsilon}\big{)}(x)\,dx+\beta\varepsilon^{q}=[1+o_{\varepsilon}(1)]\Big{[}\frac{M^{*}}{M}-1\Big{]}\varepsilon^{-\gamma^{\prime}}+\beta\varepsilon^{q}
\displaystyle\geq (1+oε(1))q+γq(qβγ)γγ+q[MM1]qγ+q, where β1,\displaystyle(1+o_{\varepsilon}(1))\frac{q+\gamma^{\prime}}{q}\bigg{(}\frac{q\beta}{\gamma^{\prime}}\bigg{)}^{\frac{\gamma^{\prime}}{\gamma^{\prime}+q}}\Bigg{[}\frac{M^{*}}{M}-1\Bigg{]}^{\frac{q}{\gamma^{\prime}+q}},~{}\text{ where }~{}\beta\gg 1,

which contradicts Lemma 5.1. This completes the proof of claim (5.59).

With the help of (5.59), we obtain that y0n\exists y_{0}\in\mathbb{R}^{n} such that, up to a subsequence,

limε0xεPiε=y0.\displaystyle\lim_{\varepsilon\rightarrow 0}\frac{x_{\varepsilon}-P_{i}}{\varepsilon}=y_{0}.

Then we aim to prove that y0y_{0} satisfies (1.45), i.e. Hi(y0)=infynHi(y)=μH_{i}(y_{0})=\inf\limits_{y\in\mathbb{R}^{n}}H_{i}(y)=\mu with PiZ0P_{i}\in Z_{0}. To begin with, noting qi=qq_{i}=q, we apply Fatou’s lemma then conclude from (5.51), (5.52) and (1.41) that

limε0εqnV(εx+xε)mε𝑑x\displaystyle\lim_{\varepsilon\rightarrow 0}\varepsilon^{-q}\int_{\mathbb{R}^{n}}V(\varepsilon x+x_{\varepsilon})m_{\varepsilon}\,dx =limε0nV(ε(x+xεPiε)+Pi)|ε(x+xεPiε)|q|x+xεPiε|qmε𝑑x\displaystyle=\lim_{\varepsilon\rightarrow 0}\int_{\mathbb{R}^{n}}\frac{V\Big{(}\varepsilon\big{(}x+\frac{x_{\varepsilon}-P_{i}}{\varepsilon}\big{)}+P_{i}\Big{)}}{|\varepsilon\big{(}x+\frac{x_{\varepsilon}-P_{i}}{\varepsilon}\big{)}|^{q}}\bigg{|}x+\frac{x_{\varepsilon}-P_{i}}{\varepsilon}\bigg{|}^{q}m_{\varepsilon}\,dx
nai|x+y0|qm0𝑑xμ,\displaystyle\geq\int_{\mathbb{R}^{n}}a_{i}|x+y_{0}|^{q}m_{0}\,dx\geq\mu, (5.60)

where the last two equalities hold if and only if (1.45) holds. Thus, we have

eα,Mεγ[MM1](1+o(1))+εqμ(1+o(1))(1+o(1))q+γq(qμγ)γγ+q[MM1]qγ+q=(1+o(1))q+γq(qμγ)γγ+q[1MM]qγq(MM)qγ+q(1+o(1))q+γq(qμγ)γγ+q[1MM]qγ+q,\begin{split}e_{\alpha^{*},M}\geq&\varepsilon^{-\gamma^{\prime}}\Big{[}\frac{M^{*}}{M}-1\Big{]}(1+o(1))+\varepsilon^{q}\mu(1+o(1))\\ \geq&(1+o(1))\frac{q+\gamma^{\prime}}{q}\Big{(}\frac{q\mu}{\gamma^{\prime}}\Big{)}^{\frac{\gamma^{\prime}}{\gamma^{\prime}+q}}\Big{[}\frac{M^{*}}{M}-1\Big{]}^{\frac{q}{\gamma^{\prime}+q}}\\ =&(1+o(1))\frac{q+\gamma^{\prime}}{q}\Big{(}\frac{q\mu}{\gamma^{\prime}}\Big{)}^{\frac{\gamma^{\prime}}{\gamma^{\prime}+q}}\Big{[}1-\frac{M}{M^{*}}\Big{]}^{\frac{q}{\gamma^{\prime}-q}}\Big{(}\frac{M^{*}}{M}\Big{)}^{\frac{q}{\gamma^{\prime}+q}}\\ \geq&(1+o(1))\frac{q+\gamma^{\prime}}{q}\Big{(}\frac{q\mu}{\gamma^{\prime}}\Big{)}^{\frac{\gamma^{\prime}}{\gamma^{\prime}+q}}\Big{[}1-\frac{M}{M^{*}}\Big{]}^{\frac{q}{\gamma^{\prime}+q}},\end{split} (5.61)

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

ε=[γqμ[1MM]]1γ+q(1+o(1)).\displaystyle\varepsilon=\left[\frac{\gamma^{\prime}}{q\mu}\left[1-\frac{M}{M^{*}}\right]\right]^{\frac{1}{\gamma^{\prime}+q}}(1+o(1)). (5.62)

Thus, combining (5.61) with (5.53), one has all equalities in (5.61) hold. It immediately follows that all "=" in (5.60) also hold. Now, we obtain (1.44) and (1.45), which completes the proof of Theorem. 1.4. ∎

Theorem 1.4 implies that if the local expansion (5.51) is imposed on potential VV, then the minimizers to problem (1.26) will concentrates at the location where VV is weighted flattest as MM.M\nearrow M^{*}. In particular, the asymptotic behavior of scaling factor ε\varepsilon is accurately characterized.

6 Discussion

In this paper, we mainly investigated the existence of ground states to (1.4) with critical mass exponent in the nonlocal coupling. First of all, we analyzed the attainability of the best constant in the Gagliardo-Nirenberg type’s ratio defined by (1.28), which corresponds the existence of ground states to the potential-free Mean-field Games system. Next, with the aid of Gagliardo-Nirenberg type’s inequality, we employ the variational approach to classify the existence of minimizers to the constrained minimization problem (1.22). In particular, while discussing the existence of classical solutions to (1.4) under the subcritical mass, we introduced the mollifier and showed the LL^{\infty} of mm to the mollified minimization problems, in which the Hardy-Littlewood-Sobolev inequality is crucial. Then taking the limit and applying standard elliptic regularities, we obtained the existence of classical solutions to (1.4) under the subcritical mass. Finally, with some assumptions imposed in the potential V,V, we performed the scaling argument and blow-up analysis to derive the asymptotic behaviors of ground states to (1.4) in the singular limit of MM, where the Pohozaev identities have been intensively used for the L1L^{1} convergence of m.m.

There are some interesting problems that deserve the explorations in the future. In Section 3, some technical restriction on mm was imposed, which is the boundedness of nm|x|b𝑑x\int_{\mathbb{R}^{n}}m|x|^{b}\,dx for sufficiently small b>0.b>0. It is an open problem to remove this condition while establishing the Gagliardo-Nirenberg type’s inequality. It is also intriguing to investigate the properties of ground states including uniqueness, symmetries, etc. to potential-free Mean-field Games systems (1.11) with the Hartree coupling and polynomial Hamiltonian. The extension of our results into a general class of potential VV is a challenging problem due to the lower bounds of the value function u.u.

Appendix A Basic proerties of Riesz potential

This Appendix is devoted to some well-known results for the estimates involving Riesz potential, which can be found in [22, Theorem 4.3], [25, Theorem 14.37] and [5, Theorem 2.8].

Lemma A.1 (Hardy Littlewood-Sobolev inequality).

Assume that 0<α<n0<\alpha<n and 1<r<nα1<r<\frac{n}{\alpha}. Then for any fLr(n)f\in L^{r}(\mathbb{R}^{n}), it holds

KαfLnrnαr(n)C(n,α,r)fLr(n),\displaystyle\|K_{\alpha}*f\|_{L^{\frac{nr}{n-\alpha r}}(\mathbb{R}^{n})}\leq C(n,\alpha,r)\|f\|_{L^{r}(\mathbb{R}^{n})}, (A.1)

where constant C>0C>0 depending on nn, α\alpha and r.r.

Moreover, suppose that r,s>1r,s>1 with 1rαn+1s=1\frac{1}{r}-\frac{\alpha}{n}+\frac{1}{s}=1, fLr(n)f\in L^{r}(\mathbb{R}^{n}) and gLs(n)g\in L^{s}(\mathbb{R}^{n}). Then, we have there exists a sharp constant C(n,α,r)C(n,\alpha,r) independent of ff and gg such that

|nnf(x)g(y)|xy|nα𝑑x𝑑y|C(n,α,r)fLr(n)gLs(n).\displaystyle\bigg{|}\int_{\mathbb{R}^{n}}\int_{\mathbb{R}^{n}}\frac{f(x)g(y)}{|x-y|^{n-\alpha}}\,dx\,dy\bigg{|}\leq C(n,\alpha,r)\|f\|_{L^{r}(\mathbb{R}^{n})}\|g\|_{L^{s}(\mathbb{R}^{n})}. (A.2)

In particular, we find from Lemma A.1 that if r=sr=s in (A.2) and fL2nn+α(n)f\in L^{\frac{2n}{n+\alpha}}(\mathbb{R}^{n}), then there exists a sharp constant C(n,α)C(n,\alpha) independent of ff and gg such that

|nnf(x)f(y)|xy|nα𝑑x𝑑y|C(n,α)fL2nn+α(n)2.\displaystyle\bigg{|}\int_{\mathbb{R}^{n}}\int_{\mathbb{R}^{n}}\frac{f(x)f(y)}{|x-y|^{n-\alpha}}\,dx\,dy\bigg{|}\leq C(n,\alpha)\|f\|^{2}_{L^{\frac{2n}{n+\alpha}}(\mathbb{R}^{n})}. (A.3)
Lemma A.2 (C.f. Theorem 2.8 in [5]).

Let 0<α<n0<\alpha<n and 1<r+1<r\leq+\infty be positive constants such that r>nαr>\frac{n}{\alpha} and s[1,nα)s\in\big{[}1,\frac{n}{\alpha}\big{)} . Then for any fLr(n)Ls(n)f\in L^{r}(\mathbb{R}^{n})\cap L^{s}(\mathbb{R}^{n}), we have

KαfL(n)C1fLr(n)+C2fLs(n).\displaystyle\|K_{\alpha}*f\|_{L^{\infty}(\mathbb{R}^{n})}\leq C_{1}\|f\|_{L^{r}(\mathbb{R}^{n})}+C_{2}\|f\|_{L^{s}(\mathbb{R}^{n})}. (A.4)

where C1=C(n,α,r)C_{1}=C(n,\alpha,r) and C2=C(n,α,s)C_{2}=C(n,\alpha,s). Moreover, if 0<αnr<10<\alpha-\frac{n}{r}<1, we have

KαfC0,αnr(n).\displaystyle K_{\alpha}*f\in C^{0,\alpha-\frac{n}{r}}(\mathbb{R}^{n}). (A.5)

In particular, there exists constant C:=C(n,α,r)>0C:=C(n,\alpha,r)>0 such that

|Kαf(x)Kαf(y)||xy|αnrCfLr(n),xy.\displaystyle\frac{\big{|}K_{\alpha}*f(x)-K_{\alpha}\ast f(y)\big{|}}{|x-y|^{\alpha-\frac{n}{r}}}\leq C\|f\|_{L^{r}(\mathbb{R}^{n})},\ \ \ \ \ \forall x\neq y.

Lemma A.2 exhibits the LL^{\infty} and Hölder estimates of KαfK_{\alpha}*f under certain conditions of ff and α\alpha.

Acknowledgments

Xiaoyu Zeng is supported by NSFC (Grant Nos. 12322106, 12171379, 12271417) . Huan-Song Zhou is supported by NSFC (Grant Nos. 11931012, 12371118) .

References

  • [1] T. Bartsch and Z. Wang. Existence and multiplicity results for some superlinear elliptic problems on N{\mathbb{R}}^{N}. Commun. Part. Diff. Eq., 20(9-10):1725–1741, 1995.
  • [2] A. Bensoussan and J. Frehse. Nonlinear elliptic systems in stochastic game theory. J. Reine Angew. Math., 350:23–67, 1984.
  • [3] A. Bensoussan and J. Frehse. Ergodic Bellman systems for stochastic games in arbitrary dimension. Proc. Roy. Soc. London Ser. A, 449(1935):65–77, 1995.
  • [4] C. Bernardini. Mass concentration for Ergodic Choquard Mean-Field Games. ESAIM: COCV, 2023.
  • [5] C. Bernardini and A. Cesaroni. Ergodic mean-field games with aggregation of choquard-type. J. Differential Equations, 364:296–335, 2023.
  • [6] V. Bogachev, N. Krylov, M. Röckner, and S. Shaposhnikov. Fokker–Planck–Kolmogorov Equations, volume 207. American Mathematical Society, 2022.
  • [7] F. Cagnetti, D. Gomes, H. Mitake, and H. Tran. A new method for large time behavior of degenerate viscous Hamilton-Jacobi equations with convex Hamiltonians. Ann. Inst. H. Poincaré, 32(1):183–200, 2015.
  • [8] P. Cardaliaguet. Long time average of first order mean field games and weak KAM theory. Dyn. Games Appl., 3(4):473–488, 2013.
  • [9] P. Cardaliaguet, J. Lasry, P. Lions, and A. Porretta. Long time average of mean field games. Netw. Heterog. Media, 7(2):279–301, 2012.
  • [10] A. Cesaroni and M. Cirant. Concentration of ground states in stationary mean-field games systems. Anal. PDE, 12(3):737–787, 2018.
  • [11] M. Cirant. A generalization of the Hopf–Cole transformation for stationary Mean-Field Games systems. C. R. Math. Acad. Sci. Paris, 353(9):807–811, 2015.
  • [12] M. Cirant. Stationary focusing mean-field games. Commun. Part. Diff. Eq., 41(8):1324–1346, 2016.
  • [13] M. Cirant and A. Goffi. Maximal LqL^{q}-regularity for parabolic Hamilton-Jacobi equations and applications to mean field games. Ann. PDE, 7(2):Paper No. 19, 40, 2021.
  • [14] 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, 2024.
  • [15] D. Gomes and H. Mitake. Existence for stationary mean-field games with congestion and quadratic Hamiltonians. NoDEA Nonlinear Differential Equations Appl., 22(6):1897–1910, 2015.
  • [16] D. Gomes, S. Patrizi, and V. Voskanyan. On the existence of classical solutions for stationary extended mean field games. Nonlinear Anal., 99:49–79, 2014.
  • [17] D. Gomes, E. Pimentel, and H. Sánchez-Morgado. Time-dependent mean-field games in the superquadratic case. ESAIM Control Optim. Calc. Var., 22(2):562–580, 2016.
  • [18] D. Gomes, G. Pires, and H. Sánchez-Morgado. A-priori estimates for stationary mean-field games. Netw. Heterog. Media, 7(2):303–314, 2012.
  • [19] D. A. Gomes, E. Pimentel, and V. Voskanyan. Regularity theory for mean-field game systems. Springer, 2016.
  • [20] 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.
  • [21] J. Lasry and P. Lions. Mean field games. Jpn. J. Math., 2(1):229–260, 2007.
  • [22] E. H. Lieb and M. Loss. Analysis, volume 14 of Grad. Stud. Math. Providence, RI: AMS, American Mathematical Society, 1996.
  • [23] A. Mészáros and F. Silva. A variational approach to second order mean field games with density constraints: the stationary case. J. Math. Pures Appl., 104(6):1135–1159, 2015.
  • [24] M. Reed. Methods of modern mathematical physics: Functional analysis. Elsevier, 2012.
  • [25] R. L. Wheeden and A. Zygmund. Measure and integral. An introduction to real analysis, volume 43 of Pure Appl. Math., Marcel Dekker. Marcel Dekker, Inc., New York, NY, 1977.