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

Global derivation of the 1D Vlasov-Poisson equation from quantum many-body dynamics with screened Coulomb potential

Xuwen Chen Department of Mathematics, University of Rochester, Rochester, NY 14627, USA xuwenmath@gmail.com Shunlin Shen School of Mathematical Sciences, University of Science and Technology of China, Hefei 230026, Anhui Province, China slshen@ustc.edu.cn Ping Zhang Academy of Mathematics & Systems Science and Hua Loo-Keng Center for Mathematical Sciences, Chinese Academy of Sciences, Beijing 100190, China, and School of Mathematical Sciences, University of Chinese Academy of Sciences, Beijing 100049, China zp@amss.ac.cn  and  Zhifei Zhang School of Mathematical Sciences, Peking University, Beijing, 100871, China zfzhang@math.pku.edu.cn
Abstract.

We study the 1D quantum many-body dynamics with a screened Coulomb potential in the mean-field setting. Combining the quantum mean-field, semiclassical, and Debye length limits, we prove the global derivation of the 1D Vlasov-Poisson equation. We tackle the difficulties brought by the pure state data, whose Wigner transforms converge to Wigner measures. We find new weighted uniform estimates around which we build the proof. As a result, we obtain, globally, stronger limits, and hence the global existence of solutions to the 1D Vlasov-Poisson equation subject to such Wigner measure data, which satisfy conservation laws of mass, momentum, and energy, despite being measure solutions. This happens to solve the 1D case of an open problem regarding the conservation law of the Vlasov-Poisson equation raised in [18] by Diperna and Lions.

Key words and phrases:
Quantum Many-body Dynamics, Vlasov-Poisson Equation, Global Weak Solution, Quantum Mean-field Approximation, Semiclassical Limit
2010 Mathematics Subject Classification:
Primary 35Q55, 35Q83, 35D30; Secondary 35A01, 81V70, 82C70.

1. Introduction

Per the superposition principle, the dynamics of NN quantum particles interacting through a two-body interaction potential are governed by the linear NN-body Schrödinger equation

(1.1) {itΨN,,ε=HN,,εΨN,,ε,ΨN,,ε(0)=ΨN,in,\left\{\begin{aligned} i\hbar\partial_{t}\Psi_{N,\hbar,\varepsilon}=&H_{N,\hbar,\varepsilon}\Psi_{N,\hbar,\varepsilon},\\ \Psi_{N,\hbar,\varepsilon}(0)=&\Psi_{N,\hbar}^{\mathrm{in}},\end{aligned}\right.

where ΨN,,ε(t,x1,..,xN)\Psi_{N,\hbar,\varepsilon}(t,x_{1},..,x_{N})\in\mathbb{C} is the NN-particle wave function at time tt and the Hamiltonian operator is

(1.2) HN,,ε=j=1N122Δxj+1N1j<kNVε(xjxk).\displaystyle H_{N,\hbar,\varepsilon}=\sum_{j=1}^{N}-\frac{1}{2}\hbar^{2}\Delta_{x_{j}}+\frac{1}{N}\sum_{1\leq j<k\leq N}V_{\varepsilon}(x_{j}-x_{k}).

In many physical systems dealing with charges in which electro-magnetism is involved, an important physically observable phenomenon is the screening effect, which arises from the collective behaviors of charged particles and modifies the long-range Coulomb potential into an exponentially decaying form at a distance. The concept of a screened Coulomb potential arises in the physics of many-body systems, particularly in plasma physics, condensed matter physics, and certain areas of molecular physics. For example, for an electrically neutral system, the distribution of charges gives rise to an electric potential V(x)V(x) that satisfies Poisson’s equation

2V(x)=j=1Nqjnj(x),\displaystyle\nabla^{2}V(x)=-\sum_{j=1}^{N}q_{j}n_{j}(x),

where qjq_{j} is the charge and nj(x)n_{j}(x) is the concentration at position xx. Under suitable physical assumptions, one often reduces the Poisson’s equation to a simpler one

(2ε2)V(x)=δ(x),\displaystyle(\nabla^{2}-\varepsilon^{2})V(x)=\delta(x),

where the parameter ε\varepsilon denotes the Debye length that characterizes different physical regimes. For more details on the derivation of a screened Coulomb potential, see also the standard monograph [2]. For more physical background on the screened Coulomb potential, see for instance [16, 33, 34, 40, 44, 47].

In the paper, we consider the 1D screened Coulomb potential

(1.3) Vε(x)=±12|x|eε|x|,\displaystyle V_{\varepsilon}(x)=\pm\frac{1}{2}|x|e^{-\varepsilon|x|},

where the sign ±\pm denotes defocusing/focusing. Here, the form (1.3) is a version of approximate solution to the 1D Poisson’s equation.

Totally different from the 3D Coulomb potential 1|x|\frac{1}{|x|} which has a slow decay at the infinity, the 1D Coulomb potential |x||x| tends to infinity as |x||x|\to\infty. Hence, for the 1D interacting systems, it is reasonable to consider the screened Coulomb potential model, as it seems to be counterintuitive that the interaction force grows to be infinitely large with the distance between particles increasing to infinity. From the perspective of physics, the screening effect is widely present in many physical systems. In fact, the Debye length is an experimentally observable parameter of NN-body systems. Some people even use that to define the experimental regimes.

Taking into account the screening effect, the effective interaction range between particles, by which different physical regimes are characterized, is quantified by the Debye length. The most interesting regime might be the Debye length limit ε0\varepsilon\to 0, as the full 1D Coulomb potential is formally recovered in the limit. Thus, not only in the theoretical physics but also in the numerical computation, it is common to take the screened model as an approximation.

Nevertheless, it is a challenge to provide a rigorous proof, as the 1D screened Coulomb potential is far from a perturbation or a regularized model for the Coulomb potential. Moreover, a key goal in mathematical physics is to understand how nonlinear equations of classical physics emerge as descriptions of quantum microscopic linear dynamics in appropriate asymptotic regimes. Staring from the quantum many-body dynamics (1.1), we are concerned with the asymptotic limit of the NN-body wave function as the particle number NN\to\infty, the Planck’s constant 0\hbar\to 0, and the Debye length ε0\varepsilon\to 0, which leads to a kinetic equation, the Vlasov-Poisson equation

(1.4) {tf+ξxf+Eξf=0,xE=±f𝑑ξ,f(0)=f0.\left\{\begin{aligned} &\partial_{t}f+\xi\partial_{x}f+E\partial_{\xi}f=0,\\ &\partial_{x}E=\pm\int_{\mathbb{R}}fd\xi,\\ &f(0)=f_{0}.\end{aligned}\right.

The Vlasov-Poisson systems describe the evolution of the distribution function f(t,x,ξ)f(t,x,\xi) of particle under a self-consistent electric or gravitational field. There have been many developments such as [18, 17, 39, 51] on the global well-posedness problem of weak/measure solutions to the Vlasov-Poisson equation. Moreover, as pointed out in the review [18, p.278], apart from the uniqueness and regularity, the conservation law for the weak solutions in the kinetic theory is an important open question.

The quantum many-body dynamics (1.1) and the kinetic equation (1.4) are linked by the Wigner transform, which takes the form that

(1.5) fN,,ε(1)(t,x,ξ)=W[γN,,ε(1)](t,x,ξ)=12πeiξyγN,,ε(1)(t,x+y2,xy2)𝑑y,\displaystyle f_{N,\hbar,\varepsilon}^{(1)}(t,x,\xi)=W_{\hbar}[\gamma_{N,\hbar,\varepsilon}^{(1)}](t,x,\xi)=\frac{1}{2\pi}\int_{\mathbb{R}}e^{-i\xi y}\gamma_{N,\hbar,\varepsilon}^{(1)}\left(t,x+\frac{\hbar y}{2},x-\frac{\hbar y}{2}\right)dy,

where the first marginal density is

(1.6) γN,,ε(1)(t,x,x)=N1ΨN,,ε(t,x,x2,..,xN)ΨN,,ε(t,x,x2,..,xN)¯dx2dxN.\displaystyle\gamma_{N,\hbar,\varepsilon}^{(1)}(t,x,x^{\prime})=\int_{\mathbb{R}^{N-1}}\Psi_{N,\hbar,\varepsilon}(t,x,x_{2},..,x_{N})\overline{\Psi_{N,\hbar,\varepsilon}(t,x^{\prime},x_{2},..,x_{N})}dx_{2}\cdot\cdot\cdot dx_{N}.

The Wigner function fN,,ε(1)(t,x,ξ)f_{N,\hbar,\varepsilon}^{(1)}(t,x,\xi) turns the spatial marginal density into a real-valued density on the phase space, and satisfies induced basic properties in kinetic theory such as the conservation laws of mass, momentum, and energy.

Our goal is to justify the limit process in which the Wigner function (1.5) from the quantum many-body dynamics (1.1) tends to the Vlasov-Poisson equation (1.4).

Theorem 1.1 (Main theorem).

Let ΨN,,ε(t)\Psi_{N,\hbar,\varepsilon}(t) be the solution to the NN-body dynamics (1.1), and fN,,ε(1)(t)f_{N,\hbar,\varepsilon}^{(1)}(t) be the Wigner transform of γN,,ε(1)(t)\gamma_{N,\hbar,\varepsilon}^{(1)}(t). Assume the initial data ΨN,,ε(0)\Psi_{N,\hbar,\varepsilon}(0) is normalized and factorized in the sense that

ΨN,in=j=1Nψin(xj),ψinLx2=1,\displaystyle\Psi_{N,\hbar}^{\mathrm{in}}=\prod_{j=1}^{N}\psi_{\hbar}^{\mathrm{in}}(x_{j}),\quad\|\psi_{\hbar}^{\mathrm{in}}\|_{L_{x}^{2}}=1,

and ψin\psi_{\hbar}^{\mathrm{in}} satisfies the uniform bounds

(1.7) |x|ψ,εinLx2C,kxkψinLx2Ckkk,k0.\displaystyle\||x|\psi_{\hbar,\varepsilon}^{\mathrm{in}}\|_{L_{x}^{2}}\leq C,\quad\|\hbar^{k}\partial_{x}^{k}\psi_{\hbar}^{\mathrm{in}}\|_{L_{x}^{2}}\leq C^{k}k^{k},\quad k\geq 0.

Then there exist a subsequence of {fN,,ε(1)}\left\{f_{N,\hbar,\varepsilon}^{(1)}\right\}, which we still denote by {fN,,ε(1)}\left\{f_{N,\hbar,\varepsilon}^{(1)}\right\}, and a non-negative bounded Radon measure

f(t,dx,dξ)C([0,);+(2)w),f(t,dx,d\xi)\in C([0,\infty);\mathcal{M}^{+}(\mathbb{R}^{2})-w^{*}),

such that

(1.8) lim(N,,ε)(,0,0)0T2(fN,,ε(1)(t,x,ξ)f(t,x,ξ))ϕ𝑑x𝑑ξ𝑑t=0,\displaystyle\lim_{(N,\hbar,\varepsilon)\to(\infty,0,0)}\int_{0}^{T}\iint_{\mathbb{R}^{2}}\left(f_{N,\hbar,\varepsilon}^{(1)}(t,x,\xi)-f(t,x,\xi)\right)\phi dxd\xi dt=0,\quad

for all T>0T>0 and ϕLt1([0,T];𝒜)\phi\in L_{t}^{1}([0,T];\mathcal{A}), where the space 𝒜\mathcal{A} is defined in (4.1). The Wigner measure f(t,dx,dξ)f(t,dx,d\xi) is a weak solution to the Vlasov-Poisson equation (1.4) with the initial measure datum f(0,dx,dξ)f(0,dx,d\xi) in the sense of Definition A.1. Moreover, the Wigner measure f(t,dx,dξ)f(t,dx,d\xi) satisfies the conservation laws of mass, momentum, and energy

(1.9) 2f(t,dx,dξ)=2f(0,dx,dξ),\displaystyle\iint_{\mathbb{R}^{2}}f(t,dx,d\xi)=\iint_{\mathbb{R}^{2}}f(0,dx,d\xi),
(1.10) 2ξf(t,dx,dξ)=2ξf(0,dx,dξ),\displaystyle\iint_{\mathbb{R}^{2}}\xi f(t,dx,d\xi)=\iint_{\mathbb{R}^{2}}\xi f(0,dx,d\xi),
(1.11) 2ξ2f(t,dx,dξ)±122|xy|ρ(t,dx)ρ(t,dy)\displaystyle\iint_{\mathbb{R}^{2}}\xi^{2}f(t,dx,d\xi)\pm\frac{1}{2}\iint_{\mathbb{R}^{2}}|x-y|\rho(t,dx)\rho(t,dy)
=\displaystyle= 2ξ2f(0,dx,dξ)±122|xy|ρ(0,dx)ρ(0,dy),\displaystyle\iint_{\mathbb{R}^{2}}\xi^{2}f(0,dx,d\xi)\pm\frac{1}{2}\iint_{\mathbb{R}^{2}}|x-y|\rho(0,dx)\rho(0,dy),

where ρ(t,dx)=f(t,dx,ξ)𝑑ξ\rho(t,dx)=\int_{\mathbb{R}}f(t,dx,\xi)d\xi.

Remark 1.2 (Global existence and conservation laws).

One could also consider Theorem 1.1 as proof of global existence of measure solutions to (1.4) subject to such Wigner measure data with conversation of mass, momentum, and energy. Staring from the quantum many-body dynamics, we happen to solve the 1D case of an open problem regarding the conservation law of the Vlasov-Poisson equation raised in [18] by Diperna and Lions.

Remark 1.3 (Existence of initial data).

One can choose the initial data as ψinj\psi_{\hbar}^{\mathrm{in}}*j_{\hbar}, where the mollifier j(x)=1j(x/)j_{\hbar}(x)=\hbar^{-1}j(x/\hbar) with 0j(x)𝒮()0\leq j(x)\in\mathcal{S}(\mathbb{R}), j(x)𝑑x=1\int_{\mathbb{R}}j(x)dx=1, and |xkj|𝑑xCkk\int_{\mathbb{R}}|\partial_{x}^{k}j|dx\leq Ck^{k} for all k0k\geq 0. Then the uniform bounds (1.7) are satisfied. For example, one can choose j(x)=π1ex2j(x)=\pi^{-1}e^{-x^{2}}.

Remark 1.4 (The torus case).

With some modifications, Theorem 1.1 can be extended to the torus case for the Coulomb potential, as the screening effect is more specialized for \mathbb{R} and there is no essential difference between the unscreened and screened cases on 𝕋\mathbb{T} from the mathematical view.

Remark 1.5 (Fixed Debye length).

Our proof also works for any fixed Debye length. The limit equation would then be a Vlasov equation with a screened Coulomb potential characterized by the Debye length.

Currently, there have been many nice developments [6, 8, 15, 20, 28, 29, 30, 31, 37, 38, 41, 45, 48, 50] devoted to the derivation of the Vlasov-type equations from quantum systems. The semiclassical limit of the one-body Schrödinger equation leading to the Vlasov equation was first systematically studied in [38]. For the Coulomb potential case, in [38, 41], this problem was solved for a mixed state initial data

(1.12) j=1λjψj(x)ψj(x)¯,j=1λj=1,\displaystyle\sum_{j=1}^{\infty}\lambda_{j}^{\hbar}\psi_{j}^{\hbar}(x)\overline{\psi_{j}^{\hbar}(x^{\prime})},\quad\sum_{j=1}^{\infty}\lambda_{j}^{\hbar}=1,

under the uniform bound condition

(1.13) 13j=1(λj)2C.\displaystyle\frac{1}{\hbar^{3}}\sum_{j=1}^{\infty}(\lambda_{j}^{\hbar})^{2}\leq C.

However, a pure state in which j=1j=1, λj=1\lambda_{j}^{\hbar}=1 cannot satisfy (1.13). It was then solved in [50] for the 1D case with general initial data including the pure state densities. For the higher dimensional case, apart from the local derivation for the monokinetic case such as [30, 45, 48], it remains an open problem for the global derivation of the 3D Vlasov-Poisson equation from the quantum and classical microscopic systems.

In our setting of justifying the global limit to the weak solution of the Vlasov-Poisson equation from the 1D quantum many-body dynamics, there are also several hard problems which we list below.

  1. (1)

    The problem of the pure state density. The quantum mean-field problem in the NN\to\infty limit is closely related to the Bose-Einstein condensate, a physical phenomenon that all particles take the same quantum state. That is, the NN-body wave function takes the product form that

    ΨN,,ε(t,x1,..,xN)j=1Nψ,ε(t,xj),\displaystyle\Psi_{N,\hbar,\varepsilon}(t,x_{1},..,x_{N})\sim\prod_{j=1}^{N}\psi_{\hbar,\varepsilon}(t,x_{j}),

    which yields a pure state marginal density

    γN,,ε(1)(t,x,x)ψ,ε(t,x)ψ,ε¯(t,x).\gamma_{N,\hbar,\varepsilon}^{(1)}(t,x,x^{\prime})\sim\psi_{\hbar,\varepsilon}(t,x)\overline{\psi_{\hbar,\varepsilon}}(t,x^{\prime}).

    However, the Wigner transform of a pure state density is only known to converge to a Wigner measure as pointed by Lions and Paul in [38]. That is, to obtain the Vlasov-Poisson equation from a pure state density, we have to work in the non-smoothing setting and deal with a not only weak but also measure solution of the Vlasov-Poisson equation. Many exsiting strong-weak stability arguments such as the modulated energy method might not be valid, as the uniqueness of the limiting weak solution to the Vlasov-Poisson equation is unknown.

  2. (2)

    The non-smoothness of the potential at the origin. For the C1,1C^{1,1} interaction potentials, the mean-field and semiclassical approximation to the Vlasov-type equation has been proven in [29]. The singularity at the origin hinders the application of the method in [29]. Hence, new ideas are required to deal with the singularity at the origin.

  3. (3)

    Weak convergence problem in the Debye length limit. To recover the Vlasov-Poisson equation with the full 1D Coulomb potential which is singular at both the origin and the infinity, we need to establish the limit process, which only holds in the weak sense that

    limε0Vε(x)φ(x)𝑑x=12|x|φ(x)𝑑x,φCc().\displaystyle\lim_{\varepsilon\to 0}\int_{\mathbb{R}}V_{\varepsilon}(x)\varphi(x)dx=\int_{\mathbb{R}}\frac{1}{2}|x|\varphi(x)dx,\quad\forall\varphi\in C_{c}(\mathbb{R}).

    Therefore, two weak limits, the semiclassical and Debye length limits, entangle here. Especially for the convergence of the nonlinear term, it requires new uniform estimates and a cancellation structure to deal with these two weak limits at the same time.

  4. (4)

    The conservation laws for the limit measure solution. As we start from basic physics model, the linear NN-body dynamics, it is naturally desired that the limit solution satisfies more physical properties, such as the conservation laws. However, from the view of mathematics, like many other open problems such as the conservation of energy for the renormalized solution of the Boltzmann equation [19] and the Vlasov-Poisson equation [18, 17], it is highly non-trivial to prove these conservation laws for the limit weak measure solution.

1.1. Outline of the Proof of the Main Theorem

We divide the proof into the following five steps.

Step 1. Preliminary reduction to a one-body nonlinear Schrödinger equation.

Since the first wave of work, for example [1, 3, 21, 22, 23, 24, 25, 26, 27] and the references within on deriving the nonlinear Schrödinger equations from the quantum many-body dynamics with the delta-type and Coulomb potentials, there have been a large quantities of work on the study of the quantum mean-field limit using various methods, such as [4, 7, 5, 9, 10, 11, 12, 13, 14, 32, 35, 36, 42, 43]. One of the crucial step of the paper is to take the quantum mean-field limit and reduce the NN-body problem to the one-body nonlinear Schrödinger equation

itψ,ε=\displaystyle i\hbar\partial_{t}\psi_{\hbar,\varepsilon}= 122x2ψ,ε+(Vε|ψ,ε|2)ψ,ε.\displaystyle-\frac{1}{2}\hbar^{2}\partial_{x}^{2}\psi_{\hbar,\varepsilon}+(V_{\varepsilon}*|\psi_{\hbar,\varepsilon}|^{2})\psi_{\hbar,\varepsilon}.

We use directly the result in [4] by Ben Porat and Golse, and obtain

(1.14) fN,,ε(1)(t)f,ε(t)Lx,ξ241Nexp(Ct3ε),\displaystyle\|f_{N,\hbar,\varepsilon}^{(1)}(t)-f_{\hbar,\varepsilon}(t)\|_{L_{x,\xi}^{2}}\leq 4\sqrt{\frac{1}{N\hbar}}\exp\left(\sqrt{\frac{Ct}{\hbar^{3}\varepsilon}}\right),

where f,ε(t)=W[ψ,ε(t)]f_{\hbar,\varepsilon}(t)=W_{\hbar}[\psi_{\hbar,\varepsilon}(t)] is the Wigner transform of the one-body wave function ψ,ε(t)\psi_{\hbar,\varepsilon}(t). With this key observation, it suffices to study the limit problem for f,ε(t)f_{\hbar,\varepsilon}(t).

Step 2. Weighted uniform energy estimates.

In Section 3, we introduce new weighted uniform estimates

(1.15) x12kxkψ,ε(t)Lx2\displaystyle\|\langle x\rangle^{\frac{1}{2}}\hbar^{k}\partial_{x}^{k}\psi_{\hbar,\varepsilon}(t)\|_{L_{x}^{2}}\leq C(k,t),k1,\displaystyle C(k,t),\quad\forall k\geq 1,

based on which we set up

(1.16) xαxαξkf,ε(t,x,ξ)𝑑ξLx1\displaystyle\Big{\|}\langle x\rangle\hbar^{\alpha}\partial_{x}^{\alpha}\int_{\mathbb{R}}\xi^{k}f_{\hbar,\varepsilon}(t,x,\xi)d\xi\Big{\|}_{L_{x}^{1}}\leq C(k,α,t),\displaystyle C(k,\alpha,t),

where x=1+x2\langle x\rangle=\sqrt{1+x^{2}}. The weighted uniform estimates are new, and 1D specific for our subsequent analysis including the compactness, convergence and the conservation laws for the limit solution. The proof and usage of (1.15) and (1.16) are the key.

Step 3. Compactness and convergence. In Section 4, using the uniform estimates in Section 3, we are able to obtain higher moment difference estimates between the Wigner function and the Husimi function which is non-negative, and attain more properties. Then, we prove the compactness of the sequence {f,ε(t,x,ξ)}\left\{f_{\hbar,\varepsilon}(t,x,\xi)\right\} and justify the weak convergence (up to a subsequence) to a non-negative bounded Radon measure

f(t,dx,dξ)C([0,);+(2)w),\displaystyle f(t,dx,d\xi)\in C([0,\infty);\mathcal{M}^{+}(\mathbb{R}^{2})-w^{*}),

in the sense that for T>0\forall T>0, k0k\geq 0, there hold

lim(,ε)(0,0)0T2(ξkf,ε(t,x,ξ)ξkf(t,x,ξ))ϕ𝑑x𝑑ξ𝑑t=0,ϕLt1([0,T];𝒜),\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{0}^{T}\iint_{\mathbb{R}^{2}}\left(\xi^{k}f_{\hbar,\varepsilon}(t,x,\xi)-\xi^{k}f(t,x,\xi)\right)\phi dxd\xi dt=0,\ \forall\phi\in L_{t}^{1}([0,T];\mathcal{A}),

where the test function space 𝒜\mathcal{A} is defined in (4.1). Our method here enables a direct proof that the limit is non-negative. Furthermore, with our method, for the convergence of the moment function ξkf,ε(t,x,ξ)𝑑ξ\int_{\mathbb{R}}\xi^{k}f_{\hbar,\varepsilon}(t,x,\xi)d\xi, we are able to prove the narrow convergence due to (1.15) and (1.16). That is,

lim(,ε)(0,0)0T(ξkf,ε(t,x,ξ)𝑑ξξkf(t,x,ξ)𝑑ξ)φ𝑑x𝑑t=0,φLt1([0,T];Cb()).\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{0}^{T}\int_{\mathbb{R}}\left(\int_{\mathbb{R}}\xi^{k}f_{\hbar,\varepsilon}(t,x,\xi)d\xi-\int_{\mathbb{R}}\xi^{k}f(t,x,\xi)d\xi\right)\varphi dxdt=0,\ \forall\varphi\in L_{t}^{1}([0,T];C_{b}(\mathbb{R})).

The test functions belong to the space of the bounded continuous functions. This is the key to the conservation laws, as we can take the constant 1 as a test function now.

Step 4. Conservation laws for the limit solution.

In Section 5, we prove the conservation laws of mass, momentum, and energy for the limit measure solution as presented in (1.9)–(1.11). The mass, momentum, and kinetic energy parts follow from the narrow convergence. The difficult one is the convergence of the interaction potential energy, as the narrow convergence we obtain in Step 3 remains too weak to deal with the limit problem for the nonlinear term. To circumvent this problem, based on the weighted uniform estimates (1.15) and (1.16), we introduce a weighted transform to obtain the local strong convergence, which is the key to the convergence of the interaction potential energy.

Step 5. Convergence to the Vlasov-Poisson equation

The most intricate part is to verify the limit to the Vlasov-Poisson equation. In Sections 67, we follow the scheme in [50] to prove the moment convergence to the Vlasov-Poisson equation and establish the exponential decay for the limit measure, which is used to obtain the full convergence. More precisely, in Section 6, for the test function φ(t,x)ξk\varphi(t,x)\xi^{k}, we obtain

(1.17) ΩT(tφ+ξxφ)ξkf(t,dx,dξ)𝑑tkΩTφE¯(ξk1f(t,dx,dξ))𝑑t=0,\displaystyle\int_{\Omega_{T}}\int_{\mathbb{R}}\left(\partial_{t}\varphi+\xi\partial_{x}\varphi\right)\xi^{k}f(t,dx,d\xi)dt-k\int_{\Omega_{T}}\varphi\overline{E}\left(\int_{\mathbb{R}}\xi^{k-1}f(t,dx,d\xi)\right)dt=0,

where ΩT=(0,T)×\Omega_{T}=(0,T)\times\mathbb{R} and E¯\overline{E} is the Volpert’s symmetric average defined in (A.1). Then we prove the exponential decay estimate that

ΩTeδ|ξ|f(t,dx,dξ)𝑑tCδ,\displaystyle\iint_{\Omega_{T}}\int_{\mathbb{R}}e^{\delta|\xi|}f(t,dx,d\xi)dt\leq C_{\delta},

based on which we prove

tf+ξxfξ(E¯f)=0,\displaystyle\partial_{t}f+\xi\partial_{x}f-\partial_{\xi}(\overline{E}f)=0,

in the sense of distributions in Section 7. (Notice the difference of test functions in Sections 6 and 7.) Hence, we conclude that the limit measure f(t,dx,dξ)f(t,dx,d\xi) is a weak solution to the Vlasov-Poisson equation.

During the proof of the convergence, the main difficulties lie in dealing with the vanishing problem of the remainder term

R,ε(k)=i2αk(kα)α12k(1(1)α)Dxα(Vερ,ε)ξkαf,ε𝑑ξ,\displaystyle\mathrm{R}_{\hbar,\varepsilon}^{(k)}=i\sum_{2\leq\alpha\leq k}\binom{k}{\alpha}\frac{\hbar^{\alpha-1}}{2^{k}}(1-(-1)^{\alpha})D_{x}^{\alpha}(V_{\varepsilon}*\rho_{\hbar,\varepsilon})\int_{\mathbb{R}}\xi^{k-\alpha}f_{\hbar,\varepsilon}d\xi,

and the convergence problem of the nonlinear term

(xVερ,ε)(ξk1f,ε𝑑ξ),\displaystyle\left(\partial_{x}V_{\varepsilon}*\rho_{\hbar,\varepsilon}\right)\left(\int_{\mathbb{R}}\xi^{k-1}f_{\hbar,\varepsilon}d\xi\right),

both of which yield a entangled product limit problem in the double semiclassical and Debye length limit. Adding more to the difficulty, the limit point is a measure instead of a locally integrable function, which is also known to be a stubborn technical point. Nonetheless, using the new weighted uniform estimates (1.15)–(1.16) and an iteration scheme which reduces ξk+1\xi^{k+1} in (1.17) into ξk\xi^{k} and hence enables an induction, we can fortunately overcome these difficulties.

Finally, we put the definition of the weak solution to the Vlasov-Poisson equation in Appendix A and include some basic properties of the bounded variation functions in Appendix B. Putting together the results in the above steps 1155, we conclude Theorem 1.1.

2. Preliminary Reduction: Quantum Mean-field Limit

In this section, we take the quantum mean-field limit and reduce the quantum NN-body dynamics to the one-body nonlinear Schrödinger equation (NLS)

(2.1) {itψ,ε=122x2ψ,ε+(Vε|ψ,ε|2)ψ,ε,ψ,ε(0)=ψin.\left\{\begin{aligned} i\hbar\partial_{t}\psi_{\hbar,\varepsilon}=&-\frac{1}{2}\hbar^{2}\partial_{x}^{2}\psi_{\hbar,\varepsilon}+(V_{\varepsilon}*|\psi_{\hbar,\varepsilon}|^{2})\psi_{\hbar,\varepsilon},\\ \psi_{\hbar,\varepsilon}(0)=&\psi_{\hbar}^{\mathrm{in}}.\end{aligned}\right.

Certainly, there have been many methods developed to establish a quantitative measurement between the NN-body systems and the one-body NLS. Here, to make our limit problem of the quantum NN-body dynamics concise and clear, we use directly the result in [4], which is inspired by [42] and gives a convergence rate estimate between the NN-body dynamics and the one-body NLS with an explicit \hbar-dependence.

Theorem 2.1 ([4, Corollary 4.2]).

Let ΨN,,ε(t)\Psi_{N,\hbar,\varepsilon}(t) be the solution to the NN-body dynamics (1.1) with the factorized initial data, γN,,ε(1)(t)\gamma_{N,\hbar,\varepsilon}^{(1)}(t) be the first marginal density. Then it holds that

(2.2) Tr|γN,,ε(1)(t,x,x)ψ,ε(t,x)ψ,ε¯(t,x)|41Nexp(30tL,ε(s)𝑑s),\displaystyle\operatorname{Tr}\Big{|}\gamma_{N,\hbar,\varepsilon}^{(1)}(t,x,x^{\prime})-\psi_{\hbar,\varepsilon}(t,x)\overline{\psi_{\hbar,\varepsilon}}(t,x^{\prime})\Big{|}\leq 4\sqrt{\frac{1}{N}}\exp\left(\frac{3}{\hbar}\int_{0}^{t}L_{\hbar,\varepsilon}(s)ds\right),

where

(2.3) L,ε(t):=CVεLxψ,ε(t)H2.\displaystyle L_{\hbar,\varepsilon}(t):=C\|V_{\varepsilon}\|_{L_{x}^{\infty}}\|\psi_{\hbar,\varepsilon}(t)\|_{H^{2}}.

Using Theorem 2.1 and the energy estimate (3.3) in Section 3, we immediately obtain the following corollary.

Corollary 2.2.

Let fN,,ε(1)(t)f_{N,\hbar,\varepsilon}^{(1)}(t), f,ε(t)f_{\hbar,\varepsilon}(t) be the Wigner transform of γN,,ε(1)(t)\gamma_{N,\hbar,\varepsilon}^{(1)}(t), ψ,ε(t)\psi_{\hbar,\varepsilon}(t) respectively. We have

(2.4) fN,,ε(1)(t)f,ε(t)Lx,ξ241Nexp(Ct3ε).\displaystyle\|f_{N,\hbar,\varepsilon}^{(1)}(t)-f_{\hbar,\varepsilon}(t)\|_{L_{x,\xi}^{2}}\leq 4\sqrt{\frac{1}{N\hbar}}\exp\left(\sqrt{\frac{Ct}{\hbar^{3}\varepsilon}}\right).
Proof.

By Plancherel identity and the operator inequality that AHSTr|A|\|A\|_{HS}\leq\operatorname{Tr}|A|, we obtain

(2.5) fN,,ε(1)(t)f,ε(t)Lx,ξ2=\displaystyle\|f_{N,\hbar,\varepsilon}^{(1)}(t)-f_{\hbar,\varepsilon}(t)\|_{L_{x,\xi}^{2}}= W[γN,,ε(1)(t)]W[ψ,ε(t)]Lx,ξ2\displaystyle\|W_{\hbar}[\gamma_{N,\hbar,\varepsilon}^{(1)}(t)]-W_{\hbar}[\psi_{\hbar,\varepsilon}(t)]\|_{L_{x,\xi}^{2}}
=\displaystyle= 12γN,,ε(1)(t,x,x)ψ,ε(t,x)ψ,ε¯(t,x)Lx,x2\displaystyle\hbar^{-\frac{1}{2}}\big{\|}\gamma_{N,\hbar,\varepsilon}^{(1)}(t,x,x^{\prime})-\psi_{\hbar,\varepsilon}(t,x)\overline{\psi_{\hbar,\varepsilon}}(t,x^{\prime})\big{\|}_{L_{x,x^{\prime}}^{2}}
\displaystyle\leq 12Tr|γN,,ε(1)(t,x,x)ψ,ε(t,x)ψ,ε¯(t,x)|.\displaystyle\hbar^{-\frac{1}{2}}\operatorname{Tr}\Big{|}\gamma_{N,\hbar,\varepsilon}^{(1)}(t,x,x^{\prime})-\psi_{\hbar,\varepsilon}(t,x)\overline{\psi_{\hbar,\varepsilon}}(t,x^{\prime})\Big{|}.

Using (2.2), energy estimate (3.3), and VεLxε1\|V_{\varepsilon}\|_{L_{x}^{\infty}}\leq\varepsilon^{-1}, we arrive at (2.4). ∎

The convergence rate estimate (2.4) is enough for the limit problem up to a subsequence. Therefore, in the follow Sections 37, we start from the nonlinear Schrödinger equation (2.1) and justify its limit to the Vlasov-Poisson equation.

3. Weighted Uniform Higher Energy Estimates

In this section, we set up the weighted uniform higher energy estimates on the one-body wave function ψ,ε(t)\psi_{\hbar,\varepsilon}(t) of the nonlinear Schrödinger equation (2.1). Then using the weighted uniform estimates, we provide the higher derivative and weighted uniform estimates for the higher moments of f,ε(t,x,ξ)f_{\hbar,\varepsilon}(t,x,\xi) in Lemma 3.2.

Lemma 3.1 (Weighted uniform estimates).

Let ρ,ε(t)=|ψ,ε(t)|2\rho_{\hbar,\varepsilon}(t)=|\psi_{\hbar,\varepsilon}(t)|^{2}. We have

(3.1) x2ρ,ε(t)Lx1\displaystyle\|\langle x\rangle^{2}\rho_{\hbar,\varepsilon}(t)\|_{L_{x}^{1}}\leq C(t),\displaystyle C(t),
(3.2) x12kxkψ,ε(t)Lx2\displaystyle\|\langle x\rangle^{\frac{1}{2}}\hbar^{k}\partial_{x}^{k}\psi_{\hbar,\varepsilon}(t)\|_{L_{x}^{2}}\leq C(k,t),k1,\displaystyle C(k,t),\quad\forall k\geq 1,

where x=1+x2\langle x\rangle=\sqrt{1+x^{2}}.

Proof.

Estimate (3.1) is usually called a virial estimate, while estimate (3.2) is a new weighted uniform estimate, which might not be true for the higher dimension case.

Before proving the weighted uniform estimates (3.1)–(3.2), we set up the higher energy estimates

(3.3) kxkψ,ε(t)Lx2C(k,t).\displaystyle\big{\|}\hbar^{k}\partial_{x}^{k}\psi_{\hbar,\varepsilon}(t)\big{\|}_{L_{x}^{2}}\leq C(k,t).

For k=0k=0, estimate (3.3) just follows from the mass conservation law of (2.1). For the defousing case, we can also obtain (3.3) with k=1k=1 by using the energy conservation law of (2.1), as the potential energy is positive. However, such an argument is not valid for the focusing case. To provide a unified proof, we take the induction argument to deal with the general case k1k\geq 1 for both defocusing and focusing cases.

We assume that (3.3) holds for nk1n\leq k-1, and we prove it for n=kn=k. Using the nonlinear Schrödinger equation (2.1), we obtain

12ddtkxkψ,εLx22\displaystyle\frac{1}{2}\frac{d}{dt}\|\hbar^{k}\partial_{x}^{k}\psi_{\hbar,\varepsilon}\|_{L_{x}^{2}}^{2}
=\displaystyle= Retkxkψ,ε¯kxkψ,εdx\displaystyle\operatorname{Re}\int\overline{\partial_{t}\hbar^{k}\partial_{x}^{k}\psi_{\hbar,\varepsilon}}\hbar^{k}\partial_{x}^{k}\psi_{\hbar,\varepsilon}dx
=\displaystyle= Imk1xk[(Vερ,ε)ψ,ε]¯kxkψ,εdx\displaystyle-\operatorname{Im}\int\overline{\hbar^{k-1}\partial_{x}^{k}\left[(V_{\varepsilon}*\rho_{\hbar,\varepsilon})\psi_{\hbar,\varepsilon}\right]}\hbar^{k}\partial_{x}^{k}\psi_{\hbar,\varepsilon}dx
=\displaystyle= Im(k1xk[(Vερ,ε)ψ,ε](Vερ,ε)k1xkψ,ε)¯kxkψ,εdx.\displaystyle-\operatorname{Im}\int\overline{(\hbar^{k-1}\partial_{x}^{k}\left[(V_{\varepsilon}*\rho_{\hbar,\varepsilon})\psi_{\hbar,\varepsilon}\right]-(V_{\varepsilon}*\rho_{\hbar,\varepsilon})\hbar^{k-1}\partial_{x}^{k}\psi_{\hbar,\varepsilon})}\hbar^{k}\partial_{x}^{k}\psi_{\hbar,\varepsilon}dx.

Then by Hölder’s inequality, Leibniz rule, and Young’s inequality, we get

(3.4) 12ddtkxkψ,εLx22\displaystyle\frac{1}{2}\frac{d}{dt}\|\hbar^{k}\partial_{x}^{k}\psi_{\hbar,\varepsilon}\|_{L_{x}^{2}}^{2}
\displaystyle\leq k1xk[(Vερ,ε)ψ,ε](Vερ,ε)k1xkψ,εLx2kxkψ,εLx2\displaystyle\big{\|}\hbar^{k-1}\partial_{x}^{k}\left[(V_{\varepsilon}*\rho_{\hbar,\varepsilon})\psi_{\hbar,\varepsilon}\right]-(V_{\varepsilon}*\rho_{\hbar,\varepsilon})\hbar^{k-1}\partial_{x}^{k}\psi_{\hbar,\varepsilon}\big{\|}_{L_{x}^{2}}\|\hbar^{k}\partial_{x}^{k}\psi_{\hbar,\varepsilon}\|_{L_{x}^{2}}
\displaystyle\leq j=1k(kj)j1xjVερ,εLxkjxkjψ,εLx2kxkψ,εLx2\displaystyle\sum_{j=1}^{k}\binom{k}{j}\|\hbar^{j-1}\partial_{x}^{j}V_{\varepsilon}*\rho_{\hbar,\varepsilon}\|_{L_{x}^{\infty}}\|\hbar^{k-j}\partial_{x}^{k-j}\psi_{\hbar,\varepsilon}\|_{L_{x}^{2}}\|\hbar^{k}\partial_{x}^{k}\psi_{\hbar,\varepsilon}\|_{L_{x}^{2}}
\displaystyle\leq j=1k(kj)xVεLxj1xj1ρ,εLx1kjxkjψ,εLx2kxkψ,εLx2\displaystyle\sum_{j=1}^{k}\binom{k}{j}\|\partial_{x}V_{\varepsilon}\|_{L_{x}^{\infty}}\hbar^{j-1}\|\partial_{x}^{j-1}\rho_{\hbar,\varepsilon}\|_{L_{x}^{1}}\|\hbar^{k-j}\partial_{x}^{k-j}\psi_{\hbar,\varepsilon}\|_{L_{x}^{2}}\|\hbar^{k}\partial_{x}^{k}\psi_{\hbar,\varepsilon}\|_{L_{x}^{2}}
\displaystyle\leq Cj=1k(kj)j1xj1ρ,εLx1kjxkjψ,εLx2kxkψ,εLx2,\displaystyle C\sum_{j=1}^{k}\binom{k}{j}\hbar^{j-1}\|\partial_{x}^{j-1}\rho_{\hbar,\varepsilon}\|_{L_{x}^{1}}\|\hbar^{k-j}\partial_{x}^{k-j}\psi_{\hbar,\varepsilon}\|_{L_{x}^{2}}\|\hbar^{k}\partial_{x}^{k}\psi_{\hbar,\varepsilon}\|_{L_{x}^{2}},

where in the last inequality we have used xVεLxC\|\partial_{x}V_{\varepsilon}\|_{L_{x}^{\infty}}\leq C. Using again Leibniz rule and Hölder’s inequality, we have

(3.5) j1xj1ρ,εLx2\displaystyle\hbar^{j-1}\|\partial_{x}^{j-1}\rho_{\hbar,\varepsilon}\|_{L_{x}^{2}}\leq j1+j2=j1(j1j1)j1xj1ψ,εLx2j2xj2ψ,εLx2.\displaystyle\sum_{j_{1}+j_{2}=j-1}\binom{j-1}{j_{1}}\|\hbar^{j_{1}}\partial_{x}^{j_{1}}\psi_{\hbar,\varepsilon}\|_{L_{x}^{2}}\|\hbar^{j_{2}}\partial_{x}^{j_{2}}\psi_{\hbar,\varepsilon}\|_{L_{x}^{2}}.

Plugging (3.5) into (3.4), we use (3.3) for the case n<kn<k to obtain

(3.6) ddtkxkψ,εLx22C(k,t).\displaystyle\frac{d}{dt}\|\hbar^{k}\partial_{x}^{k}\psi_{\hbar,\varepsilon}\|_{L_{x}^{2}}^{2}\leq C(k,t).

Noticing that the initial datum satisfies

kxkψ,ε(0)Lx2Ckkk,\displaystyle\|\hbar^{k}\partial_{x}^{k}\psi_{\hbar,\varepsilon}(0)\|_{L_{x}^{2}}\leq C^{k}k^{k},

by (3.6) we arrive at

kxkψ,ε(t)Lx2C(k,t),\displaystyle\|\hbar^{k}\partial_{x}^{k}\psi_{\hbar,\varepsilon}(t)\|_{L_{x}^{2}}\leq C(k,t),

which completes the proof of (3.3).

Now, we get into the proof of estimate (3.1). For t[0,T]t\in[0,T], we have

ddte2δ|x|2x2ρ,ε(t,x)𝑑x\displaystyle\frac{d}{dt}\int_{\mathbb{R}}e^{-2\delta|x|^{2}}\langle x\rangle^{2}\rho_{\hbar,\varepsilon}(t,x)dx
=\displaystyle= e2δ|x|2x2x(Im(ψ,ε¯xψ,ε))dx\displaystyle-\int_{\mathbb{R}}e^{-2\delta|x|^{2}}\langle x\rangle^{2}\partial_{x}\left(\operatorname{Im}\left(\overline{\psi_{\hbar,\varepsilon}}\hbar\partial_{x}\psi_{\hbar,\varepsilon}\right)\right)dx
=\displaystyle= e2δ|x|22x(12δx2)Im(ψ,ε¯xψ,ε)𝑑x\displaystyle\int_{\mathbb{R}}e^{-2\delta|x|^{2}}2x(1-2\delta\langle x\rangle^{2})\operatorname{Im}\left(\overline{\psi_{\hbar,\varepsilon}}\hbar\partial_{x}\psi_{\hbar,\varepsilon}\right)dx
\displaystyle\lesssim eδ|x|2(δx2+1)Lxeδ|x|2|x|ψ,ε(t,x)Lx2xψ,ε(t,x)Lx2\displaystyle\|e^{-\delta|x|^{2}}(\delta\langle x\rangle^{2}+1)\|_{L_{x}^{\infty}}\|e^{-\delta|x|^{2}}|x|\psi_{\hbar,\varepsilon}(t,x)\|_{L_{x}^{2}}\|\hbar\partial_{x}\psi_{\hbar,\varepsilon}(t,x)\|_{L_{x}^{2}}
\displaystyle\lesssim C(T)e2δ|x|2x2ρ,ε(t,x)𝑑x,\displaystyle C(T)\int_{\mathbb{R}}e^{-2\delta|x|^{2}}\langle x\rangle^{2}\rho_{\hbar,\varepsilon}(t,x)dx,

where in the last inequality we used the energy estimate (3.3) with k=1k=1. By Gronwall’s inequality, we have

e2δ|x|2x2ρ,ε(t,x)Lx1C(t).\displaystyle\|e^{-2\delta|x|^{2}}\langle x\rangle^{2}\rho_{\hbar,\varepsilon}(t,x)\|_{L_{x}^{1}}\leq C(t).

Letting δ0\delta\to 0 and using Fatou’s lemma, we arrive at (3.1).

Next, we can prove the weighted uniform estimate (3.2). By integration by parts, Hölder’s inequality, virial estimate (3.1), and the higher energy estimates (3.3), we have

(χ(xR)x)12kxkψ,εLx22\displaystyle\Big{\|}\left(\chi(\frac{x}{R})\langle x\rangle\right)^{\frac{1}{2}}\hbar^{k}\partial_{x}^{k}\psi_{\hbar,\varepsilon}\Big{\|}_{L_{x}^{2}}^{2}
=\displaystyle= 2k|χ(xR)xxkψ,εxkψ,ε¯dx|\displaystyle\hbar^{2k}\Big{|}\int_{\mathbb{R}}\chi(\frac{x}{R})\langle x\rangle\partial_{x}^{k}\psi_{\hbar,\varepsilon}\overline{\partial_{x}^{k}\psi_{\hbar,\varepsilon}}dx\Big{|}
\displaystyle\leq 2kα=0k(kα)|ψ,εxα(χ(xR)x)x2kαψ,ε¯dx|\displaystyle\hbar^{2k}\sum_{\alpha=0}^{k}\binom{k}{\alpha}\Big{|}\int_{\mathbb{R}}\psi_{\hbar,\varepsilon}\partial_{x}^{\alpha}\left(\chi(\frac{x}{R})\langle x\rangle\right)\overline{\partial_{x}^{2k-\alpha}\psi_{\hbar,\varepsilon}}dx\Big{|}
\displaystyle\leq xψ,εLx2χ(xR)Lx2kx2kψ,εLx2\displaystyle\|\langle x\rangle\psi_{\hbar,\varepsilon}\|_{L_{x}^{2}}\|\chi(\frac{x}{R})\|_{L_{x}^{\infty}}\|\hbar^{2k}\partial_{x}^{2k}\psi_{\hbar,\varepsilon}\|_{L_{x}^{2}}
+α=1k(kα)αψ,εLx2xα(χ(xR)x)Lx2kαx2kαψ,εLx2\displaystyle+\sum_{\alpha=1}^{k}\binom{k}{\alpha}\hbar^{\alpha}\|\psi_{\hbar,\varepsilon}\|_{L_{x}^{2}}\Big{\|}\partial_{x}^{\alpha}\left(\chi(\frac{x}{R})\langle x\rangle\right)\Big{\|}_{L_{x}^{\infty}}\|\hbar^{2k-\alpha}\partial_{x}^{2k-\alpha}\psi_{\hbar,\varepsilon}\|_{L_{x}^{2}}
\displaystyle\leq C(k,t),\displaystyle C(k,t),

where in the last inequality we have used that

α1,xα(χ(xR)x)Lx1,α1.\displaystyle\hbar^{\alpha}\leq 1,\quad\Big{\|}\partial_{x}^{\alpha}\left(\chi(\frac{x}{R})\langle x\rangle\right)\Big{\|}_{L_{x}^{\infty}}\lesssim 1,\quad\forall\alpha\geq 1.

Sending RR\to\infty and using Fatou’s lemma, we arrive at (3.2). ∎

Now, we are able to provide the higher derivative and weighted uniform estimates for the higher moments of f,ε(t,x,ξ)f_{\hbar,\varepsilon}(t,x,\xi). For simplicity, we define the Wigner function

(3.7) W[u1,u2]:=12πeizξu1(x+z2)u2(xz2)¯𝑑z,\displaystyle W_{\hbar}[u_{1},u_{2}]:=\frac{1}{2\pi}\int_{\mathbb{R}}e^{-iz\xi}u_{1}(x+\frac{\hbar z}{2})\overline{u_{2}(x-\frac{\hbar z}{2})}dz,

and use the shorthand W[u]W_{\hbar}[u] to denote W[u,u]W_{\hbar}[u,u]. Next, we give two formulas of the Wigner function. Let Dx:=1ixD_{x}:=\frac{1}{i}\partial_{x}.

  • The Wigner function with a weight function ξk\xi^{k} satisfies

    (3.8) ξkW[u1,u2]=\displaystyle\xi^{k}W_{\hbar}[u_{1},u_{2}]= 12πk2kα=0k(kα)(1)kαeizξDxαu1(x+z2)Dxkαu2(xz2)¯𝑑z\displaystyle\frac{1}{2\pi}\frac{\hbar^{k}}{2^{k}}\sum_{\alpha=0}^{k}\binom{k}{\alpha}(-1)^{k-\alpha}\int e^{-iz\xi}D_{x}^{\alpha}u_{1}(x+\frac{\hbar z}{2})\overline{D_{x}^{k-\alpha}u_{2}(x-\frac{\hbar z}{2})}dz
    =\displaystyle= 12πk2kα=0k(kα)(1)kαW[Dαu1,Dkαu2].\displaystyle\frac{1}{2\pi}\frac{\hbar^{k}}{2^{k}}\sum_{\alpha=0}^{k}\binom{k}{\alpha}(-1)^{k-\alpha}W_{\hbar}[D^{\alpha}u_{1},D^{k-\alpha}u_{2}].
  • The integral of the Wigner function with a weight function ξk\xi^{k} satisfies

    (3.9) ξkW[u1,u2]𝑑ξ=k2kα=0k(kα)(1)kαDxαu1Dxkαu2¯.\displaystyle\int_{\mathbb{R}}\xi^{k}W_{\hbar}[u_{1},u_{2}]d\xi=\frac{\hbar^{k}}{2^{k}}\sum_{\alpha=0}^{k}\binom{k}{\alpha}(-1)^{k-\alpha}D_{x}^{\alpha}u_{1}\overline{D_{x}^{k-\alpha}u_{2}}.
Lemma 3.2.

For t[0,T]t\in[0,T], we have

(3.10) xα+1xαξkf,ε𝑑ξLx\displaystyle\Big{\|}\langle x\rangle\hbar^{\alpha+1}\partial_{x}^{\alpha}\int_{\mathbb{R}}\xi^{k}f_{\hbar,\varepsilon}d\xi\Big{\|}_{L_{x}^{\infty}}\leq C(k,α,t),k0,α0,\displaystyle C(k,\alpha,t),\quad\forall k\geq 0,\ \alpha\geq 0,
(3.11) xαxαξkf,ε𝑑ξLx1\displaystyle\Big{\|}\langle x\rangle\hbar^{\alpha}\partial_{x}^{\alpha}\int_{\mathbb{R}}\xi^{k}f_{\hbar,\varepsilon}d\xi\Big{\|}_{L_{x}^{1}}\leq C(k,α,t),k0,α0,\displaystyle C(k,\alpha,t),\quad\forall k\geq 0,\ \alpha\geq 0,
(3.12) kxkE,εLx\displaystyle\|\hbar^{k}\partial_{x}^{k}E_{\hbar,\varepsilon}\|_{L_{x}^{\infty}}\leq C(k,t),k0,\displaystyle C(k,t),\quad\forall k\geq 0,

where E,ε=xVερ,εE_{\hbar,\varepsilon}=\partial_{x}V_{\varepsilon}*\rho_{\hbar,\varepsilon}.

Proof.

For (3.10), by formula (3.9), Hölder’s inequality, Leibniz rule, we have

xα+1xαξkf,ε𝑑ξLx\displaystyle\Big{\|}\langle x\rangle\hbar^{\alpha+1}\partial_{x}^{\alpha}\int_{\mathbb{R}}\xi^{k}f_{\hbar,\varepsilon}d\xi\Big{\|}_{L_{x}^{\infty}}
=\displaystyle= xα+1xαξkW[ψ,ε]𝑑ξLx\displaystyle\Big{\|}\langle x\rangle\hbar^{\alpha+1}\partial_{x}^{\alpha}\int_{\mathbb{R}}\xi^{k}W_{\hbar}[\psi_{\hbar,\varepsilon}]d\xi\Big{\|}_{L_{x}^{\infty}}
\displaystyle\leq k1+k2=k(kk1)xα+1xα(k1xk1ψ,εk2xk2ψ,ε¯)Lx\displaystyle\sum_{k_{1}+k_{2}=k}\binom{k}{k_{1}}\Big{\|}\langle x\rangle\hbar^{\alpha+1}\partial_{x}^{\alpha}\left(\hbar^{k_{1}}\partial_{x}^{k_{1}}\psi_{\hbar,\varepsilon}\overline{\hbar^{k_{2}}\partial_{x}^{k_{2}}\psi_{\hbar,\varepsilon}}\right)\Big{\|}_{L_{x}^{\infty}}
\displaystyle\leq k1+k2=kα1+α2=α(kk1)(αα1)x12α1+k1+12xα1+k1ψ,εLxx12α2+k2+12xα2+k2ψ,εLx\displaystyle\sum_{k_{1}+k_{2}=k}\sum_{\alpha_{1}+\alpha_{2}=\alpha}\binom{k}{k_{1}}\binom{\alpha}{\alpha_{1}}\|\langle x\rangle^{\frac{1}{2}}\hbar^{\alpha_{1}+k_{1}+\frac{1}{2}}\partial_{x}^{\alpha_{1}+k_{1}}\psi_{\hbar,\varepsilon}\|_{L_{x}^{\infty}}\|\langle x\rangle^{\frac{1}{2}}\hbar^{\alpha_{2}+k_{2}+\frac{1}{2}}\partial_{x}^{\alpha_{2}+k_{2}}\psi_{\hbar,\varepsilon}\|_{L_{x}^{\infty}}
\displaystyle\leq C(k,α,t),\displaystyle C(k,\alpha,t),

where in the last inequality we have used the weighted uniform estimates (3.2) and the interpolation inequality that

x12j+12xjψ,εLxx(x12jxjψ,ε)Lx212x12jxjψ,εLx212C(j,t).\displaystyle\|\langle x\rangle^{\frac{1}{2}}\hbar^{j+\frac{1}{2}}\partial_{x}^{j}\psi_{\hbar,\varepsilon}\|_{L_{x}^{\infty}}\leq\Big{\|}\hbar\partial_{x}\left(\langle x\rangle^{\frac{1}{2}}\hbar^{j}\partial_{x}^{j}\psi_{\hbar,\varepsilon}\right)\Big{\|}_{L_{x}^{2}}^{\frac{1}{2}}\|\langle x\rangle^{\frac{1}{2}}\hbar^{j}\partial_{x}^{j}\psi_{\hbar,\varepsilon}\|_{L_{x}^{2}}^{\frac{1}{2}}\leq C(j,t).

For (3.11), similarly we have

xαxαξkf,ε𝑑ξLx1\displaystyle\Big{\|}\langle x\rangle\hbar^{\alpha}\partial_{x}^{\alpha}\int_{\mathbb{R}}\xi^{k}f_{\hbar,\varepsilon}d\xi\Big{\|}_{L_{x}^{1}}
\displaystyle\leq k1+k2=kα1+α2=α(kk1)(αα1)x12α1+k1xα1+k1ψ,εLx2x12α2+k2xα2+k2ψ,εLx2\displaystyle\sum_{k_{1}+k_{2}=k}\sum_{\alpha_{1}+\alpha_{2}=\alpha}\binom{k}{k_{1}}\binom{\alpha}{\alpha_{1}}\|\langle x\rangle^{\frac{1}{2}}\hbar^{\alpha_{1}+k_{1}}\partial_{x}^{\alpha_{1}+k_{1}}\psi_{\hbar,\varepsilon}\|_{L_{x}^{2}}\|\langle x\rangle^{\frac{1}{2}}\hbar^{\alpha_{2}+k_{2}}\partial_{x}^{\alpha_{2}+k_{2}}\psi_{\hbar,\varepsilon}\|_{L_{x}^{2}}
\displaystyle\leq C(k,α,t).\displaystyle C(k,\alpha,t).

For (3.12), by Young’s inequality and (3.11), we get

kxkE,εLx=xVε(kxkρ,ε)LxxVεLxkxkρ,εLx1C(k,t).\displaystyle\|\hbar^{k}\partial_{x}^{k}E_{\hbar,\varepsilon}\|_{L_{x}^{\infty}}=\big{\|}\partial_{x}V_{\varepsilon}*\left(\hbar^{k}\partial_{x}^{k}\rho_{\hbar,\varepsilon}\right)\big{\|}_{L_{x}^{\infty}}\leq\|\partial_{x}V_{\varepsilon}\|_{L_{x}^{\infty}}\|\hbar^{k}\partial_{x}^{k}\rho_{\hbar,\varepsilon}\|_{L_{x}^{1}}\leq C(k,t).

Hence, we have completed the proof of (3.10)–(3.12).

4. Compactness and Narrow Convergence

In this section, with respect to the weak topology of the dual space of 𝒜\mathcal{A} defined in (4.1), we prove the compactness of the sequence {f,ε(t,x,ξ)}\left\{f_{\hbar,\varepsilon}(t,x,\xi)\right\} justify a weak convergence to a non-negative Radon measure f(t,dx,dξ)f(t,dx,d\xi). In general, the Wigner transform of the wave function is only a real-valued function and may change sign. To ensure the non-negativity of the limit measure, we need to use the Husimi transform of the wave function, which fixes a non-negative sign.

In Section 4.1, we estimate the higher moment differences between the Wigner function and Husimi function, which is used to show that they have the same convergence and limit. Then in Section 4.2, it suffices to prove the convergence of the Husimi function to a non-negative measure, the proof of which relies on the weighted uniform estimates established in Section 3.

More specifically, for the convergence of the sequence {f,ε(t,x,ξ)}\left\{f_{\hbar,\varepsilon}(t,x,\xi)\right\}, we use the test function space introduced in [38]

(4.1) 𝒜={ϕCc(2):(ξϕ)(x,η)L1(η,Cc(x))},\displaystyle\mathcal{A}=\left\{\phi\in C_{c}^{\infty}(\mathbb{R}^{2}):\left(\mathcal{F}_{\xi}\phi\right)(x,\eta)\in L^{1}\left(\mathbb{R}_{\eta},C_{c}\left(\mathbb{R}_{x}\right)\right)\right\},

equipped with the norm

ϕ(x,ξ)𝒜=supx|(ξϕ)(x,η)|dη,\|\phi(x,\xi)\|_{\mathcal{A}}=\int_{\mathbb{R}}\sup_{x}\left|\left(\mathcal{F}_{\xi}\phi\right)(x,\eta)\right|d\eta,

where (ξϕ)(x,η)\left(\mathcal{F}_{\xi}\phi\right)(x,\eta) is the Fourier transform of ϕ(x,ξ)\phi(x,\xi) with respect to ξ\xi.

Furthermore, for the convergence of the moment function ξkf,ε(t,x,ξ)𝑑ξ\int_{\mathbb{R}}\xi^{k}f_{\hbar,\varepsilon}(t,x,\xi)d\xi, we are able to prove the narrow convergence. That is, the test functions belong to the space of the bounded continuous functions, which we denote by Cb()C_{b}(\mathbb{R}). The stronger narrow convergence is the key to the conservation laws for the limit measure in Section 5 and the moment convergence to the Vlasov-Poisson equation in Section 6.

4.1. Higher Moment Estimates between the Wigner and Husimi function

We first define the Husimi transform. For more details, see for instance [38, 49].

Definition 4.1.

Given uL2u\in L^{2}, the Husimi transform of uu is defined by

(4.2) W~[u]=W[u](x,ξ)G,\displaystyle\widetilde{W}_{\hbar}\left[u\right]=W_{\hbar}\left[u\right]*_{(x,\xi)}G_{\hbar},

with

G(x,ξ)=(π)1e|x|2e|ξ|2:=g(x)g(ξ).\displaystyle G_{\hbar}(x,\xi)=(\pi\hbar)^{-1}e^{-\frac{|x|^{2}}{\hbar}}e^{-\frac{|\xi|^{2}}{\hbar}}:=g_{\hbar}(x)g_{\hbar}(\xi).

An important property of the Husimi function is the non-negativity, that is, W~[u]0\widetilde{W}_{\hbar}\left[u\right]\geq 0. To make use of the non-negativity of the Husimi function, we need to provide the higher moment estimates between the Wigner function and Husimi function as follows.

Lemma 4.2.

For ϕ(x,ξ)Cc(2)\phi(x,\xi)\in C_{c}^{\infty}(\mathbb{R}^{2}), there holds that

(4.3) |2ξkW[u1,u2]ϕ𝑑x𝑑ξ|α=0k(kα)αxαu1Lx2kαxkαu2Lx2ϕ𝒜,\displaystyle\Big{|}\iint_{\mathbb{R}^{2}}\xi^{k}W_{\hbar}[u_{1},u_{2}]\phi dxd\xi\Big{|}\leq\sum_{\alpha=0}^{k}\binom{k}{\alpha}\|\hbar^{\alpha}\partial_{x}^{\alpha}u_{1}\|_{L_{x}^{2}}\|\hbar^{k-\alpha}\partial_{x}^{k-\alpha}u_{2}\|_{L_{x}^{2}}\|\phi\|_{\mathcal{A}},

In particular, we obtain

(4.4) |2ξkf,ε(t,x,ξ)ϕ𝑑x𝑑ξ|C(k,t)ϕ𝒜.\displaystyle\Big{|}\iint_{\mathbb{R}^{2}}\xi^{k}f_{\hbar,\varepsilon}(t,x,\xi)\phi dxd\xi\Big{|}\leq C(k,t)\|\phi\|_{\mathcal{A}}.

Moreover, for ϕ(x,ξ)Cc(2)\phi(x,\xi)\in C_{c}^{\infty}(\mathbb{R}^{2}) and φ(x)Cc()\varphi(x)\in C_{c}^{\infty}(\mathbb{R}), we have the estimates on the difference between the Wigner function and Husimi function that

(4.5) |2(ξkW~[u]ξkW[u])ϕ𝑑x𝑑ξ|\displaystyle\Big{|}\iint_{\mathbb{R}^{2}}\left(\xi^{k}\widetilde{W}_{\hbar}[u]-\xi^{k}W_{\hbar}[u]\right)\phi dxd\xi\Big{|}
\displaystyle\leq uLx22ϕ(x,ξ)Gϕ𝒜+C(k)α=0k1α1+α2=αkαα1xα1uLx2α2xα2uLx2ϕ𝒜,\displaystyle\|u\|_{L_{x}^{2}}^{2}\|\phi*_{(x,\xi)}G_{\hbar}-\phi\|_{\mathcal{A}}+C(k)\sum_{\alpha=0}^{k-1}\sum_{\alpha_{1}+\alpha_{2}=\alpha}\hbar^{k-\alpha}\|\hbar^{\alpha_{1}}\partial_{x}^{\alpha_{1}}u\|_{L_{x}^{2}}\|\hbar^{\alpha_{2}}\partial_{x}^{\alpha_{2}}u\|_{L_{x}^{2}}\|\phi\|_{\mathcal{A}},

and

(4.6) (ξkW~[u]𝑑ξξkW[u]𝑑ξ)φ𝑑x\displaystyle\int_{\mathbb{R}}\left(\int_{\mathbb{R}}\xi^{k}\widetilde{W}_{\hbar}[u]d\xi-\int_{\mathbb{R}}\xi^{k}W_{\hbar}[u]d\xi\right)\varphi dx
\displaystyle\leq ξkW[u]𝑑ξLx1φxgφLx+C(k)α=0k1kαξαW[u]𝑑ξLx1φLx.\displaystyle\Big{\|}\int_{\mathbb{R}}\xi^{k}W_{\hbar}[u]d\xi\Big{\|}_{L_{x}^{1}}\|\varphi*_{x}g_{\hbar}-\varphi\|_{L_{x}^{\infty}}+C(k)\sum_{\alpha=0}^{k-1}\hbar^{k-\alpha}\Big{\|}\int_{\mathbb{R}}\xi^{\alpha}W_{\hbar}[u]d\xi\Big{\|}_{L_{x}^{1}}\|\varphi\|_{L_{x}^{\infty}}.
Proof.

For (4.3) with k=0k=0, we have

|2W[u1,u2]ϕ(x,ξ)𝑑x𝑑ξ|\displaystyle\Big{|}\iint_{\mathbb{R}^{2}}W_{\hbar}[u_{1},u_{2}]\phi(x,\xi)dxd\xi\Big{|}
\displaystyle\leq (supx|ξϕ(x,y)|dy)(supy|u1(x+y2)u2(xy2)¯|𝑑x)\displaystyle\left(\int_{\mathbb{R}}\sup_{x}|\mathcal{F}_{\xi}\phi(x,y)|dy\right)\left(\sup_{y}\int_{\mathbb{R}}\Big{|}u_{1}(x+\frac{\hbar y}{2})\overline{u_{2}(x-\frac{\hbar y}{2})}\Big{|}dx\right)
\displaystyle\leq u1Lx2u2Lx2ϕ𝒜.\displaystyle\|u_{1}\|_{L_{x}^{2}}\|u_{2}\|_{L_{x}^{2}}\|\phi\|_{\mathcal{A}}.

For the k1k\geq 1 case, by formula (3.8), we obtain

|2ξkW[u1,u2]ϕ𝑑x𝑑ξ|α=0k(kα)αxαu1Lx2kαxkαu2Lx2ϕ𝒜,\displaystyle\Big{|}\iint_{\mathbb{R}^{2}}\xi^{k}W_{\hbar}[u_{1},u_{2}]\phi dxd\xi\Big{|}\leq\sum_{\alpha=0}^{k}\binom{k}{\alpha}\|\hbar^{\alpha}\partial_{x}^{\alpha}u_{1}\|_{L_{x}^{2}}\|\hbar^{k-\alpha}\partial_{x}^{k-\alpha}u_{2}\|_{L_{x}^{2}}\|\phi\|_{\mathcal{A}},

which completes the proof of (4.3). Then (4.4) follows from (4.3) and the uniform estimate (3.3).

For (4.5), we notice that

(4.7) ξkW~[u]=ξk(W[u](x,ξ)G)=α=0k(kα)(ξαW[u])(x,ξ)(ξkαG),\displaystyle\xi^{k}\widetilde{W}_{\hbar}[u]=\xi^{k}(W_{\hbar}[u]*_{(x,\xi)}G_{\hbar})=\sum_{\alpha=0}^{k}\binom{k}{\alpha}(\xi^{\alpha}W_{\hbar}[u])*_{(x,\xi)}(\xi^{k-\alpha}G_{\hbar}),

and hence get

ξkW~[u]ξkW[u]\displaystyle\xi^{k}\widetilde{W}_{\hbar}[u]-\xi^{k}W_{\hbar}[u]
=\displaystyle= (ξkW[u])(x,ξ)GξkW[u]+α=0k1(kα)(ξαW[u])(x,ξ)(ξkαG).\displaystyle(\xi^{k}W_{\hbar}[u])*_{(x,\xi)}G_{\hbar}-\xi^{k}W_{\hbar}[u]+\sum_{\alpha=0}^{k-1}\binom{k}{\alpha}(\xi^{\alpha}W_{\hbar}[u])*_{(x,\xi)}(\xi^{k-\alpha}G_{\hbar}).

Using (4.3), we obtain

|2(ξkW~[u]ξkW[u])ϕ𝑑x𝑑ξ|\displaystyle\Big{|}\iint_{\mathbb{R}^{2}}\left(\xi^{k}\widetilde{W}_{\hbar}[u]-\xi^{k}W_{\hbar}[u]\right)\phi dxd\xi\Big{|}
\displaystyle\leq |2ξkW[u](ϕ(x,ξ)Gϕ)𝑑x𝑑ξ|+α=0k1(kα)|2(ξαW[u])(ξkαG)ϕ𝑑x𝑑ξ|\displaystyle\Big{|}\iint_{\mathbb{R}^{2}}\xi^{k}W_{\hbar}[u]\left(\phi*_{(x,\xi)}G_{\hbar}-\phi\right)dxd\xi\Big{|}+\sum_{\alpha=0}^{k-1}\binom{k}{\alpha}\Big{|}\iint_{\mathbb{R}^{2}}(\xi^{\alpha}W_{\hbar}[u])*(\xi^{k-\alpha}G_{\hbar})\phi dxd\xi\Big{|}
\displaystyle\leq uLx22ϕ(x,ξ)Gϕ𝒜\displaystyle\|u\|_{L_{x}^{2}}^{2}\|\phi*_{(x,\xi)}G_{\hbar}-\phi\|_{\mathcal{A}}
+α=0k1α1+α2=α(kα)(αα1)α1xα1uLx2α2xα2uLx2(ξkαG)(x,ξ)ϕ𝒜\displaystyle+\sum_{\alpha=0}^{k-1}\sum_{\alpha_{1}+\alpha_{2}=\alpha}\binom{k}{\alpha}\binom{\alpha}{\alpha_{1}}\|\hbar^{\alpha_{1}}\partial_{x}^{\alpha_{1}}u\|_{L_{x}^{2}}\|\hbar^{\alpha_{2}}\partial_{x}^{\alpha_{2}}u\|_{L_{x}^{2}}\|(\xi^{k-\alpha}G_{\hbar})*_{(x,\xi)}\phi\|_{\mathcal{A}}
\displaystyle\leq uLx22ϕ(x,ξ)Gϕ𝒜+C(k)α=0k1α1+α2=αα1xα1uLx2α2xα2uLx2ϕ𝒜ξkαGLx1Lξ1\displaystyle\|u\|_{L_{x}^{2}}^{2}\|\phi*_{(x,\xi)}G_{\hbar}-\phi\|_{\mathcal{A}}+C(k)\sum_{\alpha=0}^{k-1}\sum_{\alpha_{1}+\alpha_{2}=\alpha}\|\hbar^{\alpha_{1}}\partial_{x}^{\alpha_{1}}u\|_{L_{x}^{2}}\|\hbar^{\alpha_{2}}\partial_{x}^{\alpha_{2}}u\|_{L_{x}^{2}}\|\phi\|_{\mathcal{A}}\|\xi^{k-\alpha}G_{\hbar}\|_{L_{x}^{1}L_{\xi}^{1}}
\displaystyle\lesssim uLx22ϕ(x,ξ)Gϕ𝒜+C(k)α=0k1α1+α2=αα1xα1uLx2α2xα2uLx2kαϕ𝒜,\displaystyle\|u\|_{L_{x}^{2}}^{2}\|\phi*_{(x,\xi)}G_{\hbar}-\phi\|_{\mathcal{A}}+C(k)\sum_{\alpha=0}^{k-1}\sum_{\alpha_{1}+\alpha_{2}=\alpha}\|\hbar^{\alpha_{1}}\partial_{x}^{\alpha_{1}}u\|_{L_{x}^{2}}\|\hbar^{\alpha_{2}}\partial_{x}^{\alpha_{2}}u\|_{L_{x}^{2}}\hbar^{k-\alpha}\|\phi\|_{\mathcal{A}},

where in the last two inequalities we have used that

g(x,ξ)ϕ𝒜\displaystyle\|g*_{(x,\xi)}\phi\|_{\mathcal{A}}\leq (ξg)x(ξϕ)Lξ1LxξgLηLx1ξϕLη1LxgLx1Lξ1ϕ𝒜,\displaystyle\|(\mathcal{F}_{\xi}g)*_{x}(\mathcal{F}_{\xi}\phi)\|_{L_{\xi}^{1}L_{x}^{\infty}}\leq\|\mathcal{F}_{\xi}g\|_{L_{\eta}^{\infty}L_{x}^{1}}\|\mathcal{F}_{\xi}\phi\|_{L_{\eta}^{1}L_{x}^{\infty}}\leq\|g\|_{L_{x}^{1}L_{\xi}^{1}}\|\phi\|_{\mathcal{A}},
ξkαGLx1Lξ1\displaystyle\|\xi^{k-\alpha}G_{\hbar}\|_{L_{x}^{1}L_{\xi}^{1}}\lesssim kα.\displaystyle\hbar^{k-\alpha}.

Therefore, we complete the proof of (4.5). Estimate (4.6) follows from a way in which we obtain (4.5).

4.2. Convergence to a Non-negative Radon Measure

As we have established the difference estimates between the Wigner function and the Husimi function, which shows that they have same limit, we can use the non-negativity of the Husimi function to conclude the convergence of {f,ε(t,x,ξ)}\left\{f_{\hbar,\varepsilon}(t,x,\xi)\right\} to a non-negative Radon measure f(t,dx,dξ)f(t,dx,d\xi).

Notation.

Here, for the convenience, we also use the notation f(t,x,ξ)dxdξf(t,x,\xi)dxd\xi to denote the measure f(t,dx,dξ)f(t,dx,d\xi). Hence, one should keep in mind that f(t,x,ξ)f(t,x,\xi) is not an Lloc1L_{loc}^{1} function.

Lemma 4.3.

There exists a subsequence of {f,ε(t,x,ξ)}\left\{f_{\hbar,\varepsilon}(t,x,\xi)\right\}, which we still denote by {f,ε(t,x,ξ)}\left\{f_{\hbar,\varepsilon}(t,x,\xi)\right\}, and a bounded non-negative Radon measure

(4.8) f(t,dx,dξ)C([0,);+(2)w),\displaystyle f(t,dx,d\xi)\in C([0,\infty);\mathcal{M}^{+}(\mathbb{R}^{2})-w^{*}),

such that for T>0\forall T>0, k0k\geq 0, there hold

(4.9) lim(,ε)(0,0)0T2(ξkf,ε(t,x,ξ)ξkf(t,x,ξ))ϕ𝑑x𝑑ξ𝑑t=0,ϕLt1([0,T];𝒜),\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{0}^{T}\iint_{\mathbb{R}^{2}}\left(\xi^{k}f_{\hbar,\varepsilon}(t,x,\xi)-\xi^{k}f(t,x,\xi)\right)\phi dxd\xi dt=0,\ \forall\phi\in L_{t}^{1}([0,T];\mathcal{A}),

and

(4.10) lim(,ε)(0,0)0T(ξkf,ε(t,x,ξ)𝑑ξξkf(t,x,ξ)𝑑ξ)φ𝑑x𝑑t=0,φLt1([0,T];Cb()),\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{0}^{T}\int_{\mathbb{R}}\left(\int_{\mathbb{R}}\xi^{k}f_{\hbar,\varepsilon}(t,x,\xi)d\xi-\int_{\mathbb{R}}\xi^{k}f(t,x,\xi)d\xi\right)\varphi dxdt=0,\ \forall\varphi\in L_{t}^{1}([0,T];C_{b}(\mathbb{R})),

and

(4.11) lim(,ε)(0,0)(ξkf,ε(t,x,ξ)𝑑ξξkf(t,x,ξ)𝑑ξ)φ𝑑x=0,t0,φCb().\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\mathbb{R}}\left(\int_{\mathbb{R}}\xi^{k}f_{\hbar,\varepsilon}(t,x,\xi)d\xi-\int_{\mathbb{R}}\xi^{k}f(t,x,\xi)d\xi\right)\varphi dx=0,\ \forall t\geq 0,\ \varphi\in C_{b}(\mathbb{R}).

Moreover, the limit measure satisfies the weighted estimates that

(4.12) 2x2f(t,dx,dξ)\displaystyle\iint_{\mathbb{R}^{2}}\langle x\rangle^{2}f(t,dx,d\xi)\leq C(t),\displaystyle C(t),
(4.13) 2x|ξ|kf(t,dx,dξ)\displaystyle\iint_{\mathbb{R}^{2}}\langle x\rangle|\xi|^{k}f(t,dx,d\xi)\leq C(k,t).\displaystyle C(k,t).
Proof.

As we need to prove the non-negativity of the limit point, we take the Husimi transform of the wave function ψ,ε\psi_{\hbar,\varepsilon}, which we denote by f~,ε(t,x,ξ):=W~[ψ,ε]0\widetilde{f}_{\hbar,\varepsilon}(t,x,\xi):=\widetilde{W}_{\hbar}[\psi_{\hbar,\varepsilon}]\geq 0.

We first prove (4.8) and (4.9). The proof can be divided into the following three steps.

Step 1. Lx,ξ1L_{x,\xi}^{1} Uniform Bounds.

By estimate (4.5) on the Wigner function and the Husimi function in Lemma 4.3 and the uniform estimate (3.3), we obtain

|0T2(ξkf,ε(t,x,ξ)ξkf~,ε(t,x,ξ))ϕ𝑑x𝑑ξ𝑑t|\displaystyle\Big{|}\int_{0}^{T}\iint_{\mathbb{R}^{2}}\left(\xi^{k}f_{\hbar,\varepsilon}(t,x,\xi)-\xi^{k}\widetilde{f}_{\hbar,\varepsilon}(t,x,\xi)\right)\phi dxd\xi dt\Big{|}
\displaystyle\leq ψ,εLtLx22ϕ(x,ξ)GϕLt1𝒜\displaystyle\|\psi_{\hbar,\varepsilon}\|_{L_{t}^{\infty}L_{x}^{2}}^{2}\|\phi*_{(x,\xi)}G_{\hbar}-\phi\|_{L_{t}^{1}\mathcal{A}}
+C(k)α=0k1α1+α2=αkαα1xα1ψ,εLtLx2α2xα2ψ,εLtLx2ϕLt1𝒜\displaystyle+C(k)\sum_{\alpha=0}^{k-1}\sum_{\alpha_{1}+\alpha_{2}=\alpha}\hbar^{k-\alpha}\|\hbar^{\alpha_{1}}\partial_{x}^{\alpha_{1}}\psi_{\hbar,\varepsilon}\|_{L_{t}^{\infty}L_{x}^{2}}\|\hbar^{\alpha_{2}}\partial_{x}^{\alpha_{2}}\psi_{\hbar,\varepsilon}\|_{L_{t}^{\infty}L_{x}^{2}}\|\phi\|_{L_{t}^{1}\mathcal{A}}
\displaystyle\lesssim ϕ(x,ξ)GϕLt1𝒜+ϕLt1𝒜0.\displaystyle\|\phi*_{(x,\xi)}G_{\hbar}-\phi\|_{L_{t}^{1}\mathcal{A}}+\hbar\|\phi\|_{L_{t}^{1}\mathcal{A}}\to 0.

Thus, the Wigner function f,ε(t,x,ξ)f_{\hbar,\varepsilon}(t,x,\xi) and the Husimi function f~,ε(t,x,ξ)\widetilde{f}_{\hbar,\varepsilon}(t,x,\xi) have the same convergence and limit. We are left to prove the convergence and the limit of the Husimi function f~,ε(t,x,ξ)\widetilde{f}_{\hbar,\varepsilon}(t,x,\xi).

We establish the Lx,ξ1L_{x,\xi}^{1} uniform bound for f~,ε(t,x,ξ)\widetilde{f}_{\hbar,\varepsilon}(t,x,\xi). By formula (4.7), we use the uniform estimate (3.11) to get

ξ2kf~,ε(t,x,ξ)Lx,ξ1=\displaystyle\|\xi^{2k}\widetilde{f}_{\hbar,\varepsilon}(t,x,\xi)\|_{L_{x,\xi}^{1}}= 2ξ2kf~,ε𝑑x𝑑ξ\displaystyle\iint_{\mathbb{R}^{2}}\xi^{2k}\widetilde{f}_{\hbar,\varepsilon}dxd\xi
\displaystyle\leq α=02k(2kα)|2(ξαf,ε)(x,ξ)(ξ2kαG)𝑑x𝑑ξ|\displaystyle\sum_{\alpha=0}^{2k}\binom{2k}{\alpha}\Big{|}\iint_{\mathbb{R}^{2}}(\xi^{\alpha}f_{\hbar,\varepsilon})*_{(x,\xi)}(\xi^{2k-\alpha}G_{\hbar})dxd\xi\Big{|}
=\displaystyle= α=02k(2kα)|(2ξαf,ε𝑑x𝑑ξ)(2ξ2kαG𝑑x𝑑ξ)|\displaystyle\sum_{\alpha=0}^{2k}\binom{2k}{\alpha}\Big{|}\left(\iint_{\mathbb{R}^{2}}\xi^{\alpha}f_{\hbar,\varepsilon}dxd\xi\right)\left(\iint_{\mathbb{R}^{2}}\xi^{2k-\alpha}G_{\hbar}dxd\xi\right)\Big{|}
\displaystyle\leq C(2k,t).\displaystyle C(2k,t).

By Hölder’s inequality, we arrive at

(4.14) ξkf~,ε(t,x,ξ)Lx,ξ1ξ2kf~,ε(t,x,ξ)Lx,ξ112f~,ε(t,x,ξ)Lx,ξ112C(k,t).\displaystyle\|\xi^{k}\widetilde{f}_{\hbar,\varepsilon}(t,x,\xi)\|_{L_{x,\xi}^{1}}\leq\|\xi^{2k}\widetilde{f}_{\hbar,\varepsilon}(t,x,\xi)\|_{L_{x,\xi}^{1}}^{\frac{1}{2}}\|\widetilde{f}_{\hbar,\varepsilon}(t,x,\xi)\|_{L_{x,\xi}^{1}}^{\frac{1}{2}}\leq C(k,t).

Step 2. Equicontinuity.

To obtain the equicontinuity of f~,ε(t,x,ξ)\widetilde{f}_{\hbar,\varepsilon}(t,x,\xi) and apply the compactness argument, we prove the uniform estimates for the time-derivative of f~,ε\widetilde{f}_{\hbar,\varepsilon}. We take the duality argument and notice that

(4.15) 2ξktf~,εϕdxdξ=2ξk(tf,ε(x,ξ)G)ϕ𝑑x𝑑ξ.\displaystyle\iint_{\mathbb{R}^{2}}\xi^{k}\partial_{t}\widetilde{f}_{\hbar,\varepsilon}\phi dxd\xi=\iint_{\mathbb{R}^{2}}\xi^{k}\left(\partial_{t}f_{\hbar,\varepsilon}*_{(x,\xi)}G_{\hbar}\right)\phi dxd\xi.

From the nonlinear Schrödinger equation (2.1), we have

(4.16) tf,ε+ξxf,ε+Θ[Vε,f,ε]=0,\displaystyle\partial_{t}f_{\hbar,\varepsilon}+\xi\partial_{x}f_{\hbar,\varepsilon}+\Theta[V_{\varepsilon},f_{\hbar,\varepsilon}]=0,

where the nonlinear term is

(4.17) Θ[Vε,f,ε]=i2π2Vερ,ε(x+y2)Vερ,ε(xy2)f,ε(t,x,η)ei(ξη)y𝑑η𝑑y.\displaystyle\Theta[V_{\varepsilon},f_{\hbar,\varepsilon}]=\frac{i}{2\pi}\iint_{\mathbb{R}^{2}}\frac{V_{\varepsilon}*\rho_{\hbar,\varepsilon}(x+\frac{\hbar y}{2})-V_{\varepsilon}*\rho_{\hbar,\varepsilon}(x-\frac{\hbar y}{2})}{\hbar}f_{\hbar,\varepsilon}(t,x,\eta)e^{-i(\xi-\eta)y}d\eta dy.

Putting (4.16) into (4.15), we obtain

2ξktf~,εϕdxdξ=\displaystyle\iint_{\mathbb{R}^{2}}\xi^{k}\partial_{t}\widetilde{f}_{\hbar,\varepsilon}\phi dxd\xi= 2ξk((ξxf,ε)(x,ξ)G)ϕ𝑑x𝑑ξ\displaystyle-\iint_{\mathbb{R}^{2}}\xi^{k}\left((\xi\partial_{x}f_{\hbar,\varepsilon})*_{(x,\xi)}G_{\hbar}\right)\phi dxd\xi
2ξk(Θ[Vε,f,ε](x,ξ)G)ϕ𝑑x𝑑ξ\displaystyle-\iint_{\mathbb{R}^{2}}\xi^{k}\left(\Theta[V_{\varepsilon},f_{\hbar,\varepsilon}]*_{(x,\xi)}G_{\hbar}\right)\phi dxd\xi
:=\displaystyle:= I1+I2.\displaystyle I_{1}+I_{2}.

For the linear term I1I_{1}, we rewrite

I1=\displaystyle I_{1}= 2ξk((ξf,ε)(x,ξ)G)(xϕ)𝑑x𝑑ξ\displaystyle\iint_{\mathbb{R}^{2}}\xi^{k}\left((\xi f_{\hbar,\varepsilon})*_{(x,\xi)}G_{\hbar}\right)\left(\partial_{x}\phi\right)dxd\xi
=\displaystyle= α=0k(kα)2((ξα+1f,ε)(x,ξ)(ξkαG))(xϕ)𝑑x𝑑ξ\displaystyle\sum_{\alpha=0}^{k}\binom{k}{\alpha}\iint_{\mathbb{R}^{2}}\left((\xi^{\alpha+1}f_{\hbar,\varepsilon})*_{(x,\xi)}(\xi^{k-\alpha}G_{\hbar})\right)\left(\partial_{x}\phi\right)dxd\xi
=\displaystyle= α=0k(kα)2ξα+1f,ε((ξkαG)(x,ξ)(xϕ))𝑑x𝑑ξ.\displaystyle\sum_{\alpha=0}^{k}\binom{k}{\alpha}\iint_{\mathbb{R}^{2}}\xi^{\alpha+1}f_{\hbar,\varepsilon}\left((\xi^{k-\alpha}G_{\hbar})*_{(x,\xi)}\left(\partial_{x}\phi\right)\right)dxd\xi.

Using estimate (4.4), we get

(4.18) |I1|α=0k(ξkαG)(x,ξ)xϕ𝒜xϕ𝒜ϕHx,ξ2.\displaystyle|I_{1}|\lesssim\sum_{\alpha=0}^{k}\|(\xi^{k-\alpha}G_{\hbar})*_{(x,\xi)}\partial_{x}\phi\|_{\mathcal{A}}\lesssim\|\partial_{x}\phi\|_{\mathcal{A}}\lesssim\|\phi\|_{H_{x,\xi}^{2}}.

For the nonlinear term I2I_{2}, we rewrite

(4.19) I2=\displaystyle I_{2}= α=0k(kα)2((ξαΘ[Vε,f,ε])(x,ξ)(ξkαG))ϕ𝑑x𝑑ξ\displaystyle-\sum_{\alpha=0}^{k}\binom{k}{\alpha}\iint_{\mathbb{R}^{2}}\left(\left(\xi^{\alpha}\Theta[V_{\varepsilon},f_{\hbar,\varepsilon}]\right)*_{(x,\xi)}\left(\xi^{k-\alpha}G_{\hbar}\right)\right)\phi dxd\xi
=\displaystyle= α=0k(kα)2ξαΘ[Vε,f,ε]((ξkαG)(x,ξ)ϕ)𝑑x𝑑ξ.\displaystyle-\sum_{\alpha=0}^{k}\binom{k}{\alpha}\iint_{\mathbb{R}^{2}}\xi^{\alpha}\Theta[V_{\varepsilon},f_{\hbar,\varepsilon}]\left(\left(\xi^{k-\alpha}G_{\hbar}\right)*_{(x,\xi)}\phi\right)dxd\xi.

Using again estimate (4.4), we have

(4.20) 2Θ[Vε,f,ε]ϕ𝑑x𝑑ξ\displaystyle\iint_{\mathbb{R}^{2}}\Theta[V_{\varepsilon},f_{\hbar,\varepsilon}]\phi dxd\xi
=\displaystyle= 2f,ε(t,x,η)η1[y(ϕ)Vερ,ε(x+y2)Vερ,ε(xy2)]𝑑x𝑑η\displaystyle\iint_{\mathbb{R}^{2}}f_{\hbar,\varepsilon}(t,x,\eta)\mathcal{F}_{\eta}^{-1}\left[\mathcal{F}_{y}(\phi)\frac{V_{\varepsilon}*\rho_{\hbar,\varepsilon}(x+\frac{\hbar y}{2})-V_{\varepsilon}*\rho_{\hbar,\varepsilon}(x-\frac{\hbar y}{2})}{\hbar}\right]dxd\eta
\displaystyle\lesssim η1[y(ϕ)Vερ,ε(x+y2)Vερ,ε(xy2)]𝒜\displaystyle\Big{\|}\mathcal{F}_{\eta}^{-1}\left[\mathcal{F}_{y}(\phi)\frac{V_{\varepsilon}*\rho_{\hbar,\varepsilon}(x+\frac{\hbar y}{2})-V_{\varepsilon}*\rho_{\hbar,\varepsilon}(x-\frac{\hbar y}{2})}{\hbar}\right]\Big{\|}_{\mathcal{A}}
\displaystyle\leq y(ϕ)Vερ,ε(x+y2)Vερ,ε(xy2)Ly1Lx\displaystyle\Big{\|}\mathcal{F}_{y}(\phi)\frac{V_{\varepsilon}*\rho_{\hbar,\varepsilon}(x+\frac{\hbar y}{2})-V_{\varepsilon}*\rho_{\hbar,\varepsilon}(x-\frac{\hbar y}{2})}{\hbar}\Big{\|}_{L_{y}^{1}L_{x}^{\infty}}
\displaystyle\leq |y|y(ϕ)Ly1LxxVερ,εLx\displaystyle\||y|\mathcal{F}_{y}(\phi)\|_{L_{y}^{1}L_{x}^{\infty}}\|\partial_{x}V_{\varepsilon}*\rho_{\hbar,\varepsilon}\|_{L_{x}^{\infty}}
\displaystyle\lesssim ϕHx,ξ3.\displaystyle\|\phi\|_{H_{x,\xi}^{3}}.

By the definition of Θ[Vε,f,ε]\Theta[V_{\varepsilon},f_{\hbar,\varepsilon}] in (4.17), we have

(4.21) ξαΘ[Vε,f,ε]=α1+α2=α(αα1)Θ(α1)[Vε,ξα2f,ε],\displaystyle\xi^{\alpha}\Theta[V_{\varepsilon},f_{\hbar,\varepsilon}]=\sum_{\alpha_{1}+\alpha_{2}=\alpha}\binom{\alpha}{\alpha_{1}}\Theta^{(\alpha_{1})}[V_{\varepsilon},\xi^{\alpha_{2}}f_{\hbar,\varepsilon}],

where

Θ(α1)[Vε,ξα2f,ε]\displaystyle\Theta^{(\alpha_{1})}[V_{\varepsilon},\xi^{\alpha_{2}}f_{\hbar,\varepsilon}]
=\displaystyle= i2π2Dyα1(Vερ,ε(x+y2)Vερ,ε(xy2))(ηα2f,ε(t,x,η))ei(ξη)y𝑑η𝑑y.\displaystyle\frac{i}{2\pi}\iint_{\mathbb{R}^{2}}D_{y}^{\alpha_{1}}\left(\frac{V_{\varepsilon}*\rho_{\hbar,\varepsilon}(x+\frac{\hbar y}{2})-V_{\varepsilon}*\rho_{\hbar,\varepsilon}(x-\frac{\hbar y}{2})}{\hbar}\right)\left(\eta^{\alpha_{2}}f_{\hbar,\varepsilon}(t,x,\eta)\right)e^{-i(\xi-\eta)y}d\eta dy.

Putting (4.21) into (4.19), in the same way as (4.20), we obtain

(4.22) |I2|\displaystyle|I_{2}|\leq α=0kα1+α2=α(kα)(αα1)|2Θ(α1)[Vε,ξα2f,ε]((ξkαG)(x,ξ)ϕ)𝑑x𝑑ξ|\displaystyle\sum_{\alpha=0}^{k}\sum_{\alpha_{1}+\alpha_{2}=\alpha}\binom{k}{\alpha}\binom{\alpha}{\alpha_{1}}\Big{|}\iint_{\mathbb{R}^{2}}\Theta^{(\alpha_{1})}[V_{\varepsilon},\xi^{\alpha_{2}}f_{\hbar,\varepsilon}]\left(\left(\xi^{k-\alpha}G_{\hbar}\right)*_{(x,\xi)}\phi\right)dxd\xi\Big{|}
\displaystyle\lesssim α=0kα1+α2=α(ξkαG)(x,ξ)ϕHx,ξ3α1xα1+1Vερ,εLx\displaystyle\sum_{\alpha=0}^{k}\sum_{\alpha_{1}+\alpha_{2}=\alpha}\big{\|}\left(\xi^{k-\alpha}G_{\hbar}\right)*_{(x,\xi)}\phi\big{\|}_{H_{x,\xi}^{3}}\|\hbar^{\alpha_{1}}\partial_{x}^{\alpha_{1}+1}V_{\varepsilon}*\rho_{\hbar,\varepsilon}\|_{L_{x}^{\infty}}
\displaystyle\lesssim C(k,t)ϕHx,ξ3,\displaystyle C(k,t)\|\phi\|_{H_{x,\xi}^{3}},

where in the last inequality we have used Young’s inequality and uniform estimate (3.12).

Combining estimates (4.18) and (4.22) on the terms I1I_{1} and I2I_{2}, we arrive at

tξkf~,εHx,ξ3C(k,t).\displaystyle\|\partial_{t}\xi^{k}\widetilde{f}_{\hbar,\varepsilon}\|_{H_{x,\xi}^{-3}}\leq C(k,t).

Step 3. Compactness Argument.

By Arzela`\grave{\text{a}}-Ascoli compactness lemma and a diagonal argument, for all k1k\geq 1 there exist a subsequence of {f~,ε}\left\{\widetilde{f}_{\hbar,\varepsilon}\right\}, which we still denote by {f~,ε}\left\{\widetilde{f}_{\hbar,\varepsilon}\right\}, and a limit point

fk(t,x,ξ)C([0,T];Hx,ξ3)f_{k}(t,x,\xi)\in C([0,T];H_{x,\xi}^{-3})

such that

(4.23) lim(,ε)(0,0)ξkf~,ε(t,x,ξ)fk(t,x,ξ)C([0,T];Hx,ξ3)=0.\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\|\xi^{k}\widetilde{f}_{\hbar,\varepsilon}(t,x,\xi)-f_{k}(t,x,\xi)\|_{C([0,T];H_{x,\xi}^{-3})}=0.

Actually, we have that fk(t,x,ξ)=ξkf(t,x,ξ)f_{k}(t,x,\xi)=\xi^{k}f(t,x,\xi) due to that

lim(,ε)(0,0)0T2ξkf~,ε(t,x,ξ)ϕ(t,x,ξ)𝑑t𝑑x𝑑ξ\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{0}^{T}\iint_{\mathbb{R}^{2}}\xi^{k}\widetilde{f}_{\hbar,\varepsilon}(t,x,\xi)\phi(t,x,\xi)dtdxd\xi
=\displaystyle= lim(,ε)(0,0)0T2f~,ε(t,x,ξ)(ξkϕ(t,x,ξ))𝑑t𝑑x𝑑ξ\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{0}^{T}\iint_{\mathbb{R}^{2}}\widetilde{f}_{\hbar,\varepsilon}(t,x,\xi)(\xi^{k}\phi(t,x,\xi))dtdxd\xi
=\displaystyle= 0T2f(t,x,ξ)ξkϕ(t,x,ξ)𝑑t𝑑x𝑑ξ.\displaystyle\int_{0}^{T}\iint_{\mathbb{R}^{2}}f(t,x,\xi)\xi^{k}\phi(t,x,\xi)dtdxd\xi.

Moreover, by the non-negativity of the Husimi function and the Lx,ξ1L_{x,\xi}^{1} uniform bound for f~,ε\widetilde{f}_{\hbar,\varepsilon}, we get

f(t,x,ξ)dxdξC([0,T];+(2)w).\displaystyle f(t,x,\xi)dxd\xi\in C([0,T];\mathcal{M}^{+}(\mathbb{R}^{2})-w^{*}).

Therefore, we have completed the proof of (4.8) and (4.9).

Next, we prove estimate (4.10). By (4.6) in Lemma 4.2 and the uniform estimate (3.11), for φLt1([0,T];Cc())\varphi\in L_{t}^{1}([0,T];C_{c}^{\infty}(\mathbb{R})) we have

|0T(ξkf,ε(t,x,ξ)𝑑ξξkf~,ε(t,x,ξ)𝑑ξ)φ𝑑x𝑑t|\displaystyle\Big{|}\int_{0}^{T}\int_{\mathbb{R}}\left(\int_{\mathbb{R}}\xi^{k}f_{\hbar,\varepsilon}(t,x,\xi)d\xi-\int_{\mathbb{R}}\xi^{k}\widetilde{f}_{\hbar,\varepsilon}(t,x,\xi)d\xi\right)\varphi dxdt\Big{|}
\displaystyle\leq ξkf,ε𝑑ξLx1φxgφLx+C(k)α=0k1kαξαf,ε𝑑ξLx1φLx\displaystyle\Big{\|}\int_{\mathbb{R}}\xi^{k}f_{\hbar,\varepsilon}d\xi\Big{\|}_{L_{x}^{1}}\|\varphi*_{x}g_{\hbar}-\varphi\|_{L_{x}^{\infty}}+C(k)\sum_{\alpha=0}^{k-1}\hbar^{k-\alpha}\Big{\|}\int_{\mathbb{R}}\xi^{\alpha}f_{\hbar,\varepsilon}d\xi\Big{\|}_{L_{x}^{1}}\|\varphi\|_{L_{x}^{\infty}}
\displaystyle\leq φxgφLt1Lx+φLt1Lx0.\displaystyle\|\varphi*_{x}g_{\hbar}-\varphi\|_{L_{t}^{1}L_{x}^{\infty}}+\hbar\|\varphi\|_{L_{t}^{1}L_{x}^{\infty}}\to 0.

Thus, it suffices to deal with ξkf~,ε(t,x,ξ)𝑑ξ\int_{\mathbb{R}}\xi^{k}\widetilde{f}_{\hbar,\varepsilon}(t,x,\xi)d\xi. In the same way in which we obtain (4.9), we also have

tξkf~,ε𝑑ξHx3C(k,t),\displaystyle\Big{\|}\partial_{t}\int_{\mathbb{R}}\xi^{k}\widetilde{f}_{\hbar,\varepsilon}d\xi\Big{\|}_{H_{x}^{-3}}\leq C(k,t),

which implies that there exists a limit point Fk(t,x)C([0,T];Hx3)F_{k}(t,x)\in C([0,T];H_{x}^{-3}) such that

lim(,ε)(0,0)ξkf~,ε(t,x,ξ)𝑑ξFk(t,x)C([0,T];Hx3)=0.\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\Big{\|}\int_{\mathbb{R}}\xi^{k}\widetilde{f}_{\hbar,\varepsilon}(t,x,\xi)d\xi-F_{k}(t,x)\Big{\|}_{C([0,T];H_{x}^{-3})}=0.

We claim that Fk(t,x)=ξkf(t,x,dξ)F_{k}(t,x)=\int_{\mathbb{R}}\xi^{k}f(t,x,d\xi). Indeed,

0TFkφ𝑑x𝑑t=\displaystyle\int_{0}^{T}\int_{\mathbb{R}}F_{k}\varphi dxdt= lim(,ε)(0,0)0T(ξkf~,ε(t,x,ξ)𝑑ξ)φ𝑑x𝑑t\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{0}^{T}\int_{\mathbb{R}}\left(\int_{\mathbb{R}}\xi^{k}\widetilde{f}_{\hbar,\varepsilon}(t,x,\xi)d\xi\right)\varphi dxdt
=\displaystyle= lim(,ε)(0,0)0T2(1+ξ2)ξkf~,ε(t,x,ξ)φ1+ξ2𝑑ξ𝑑x𝑑t\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{0}^{T}\int_{\mathbb{R}^{2}}(1+\xi^{2})\xi^{k}\widetilde{f}_{\hbar,\varepsilon}(t,x,\xi)\frac{\varphi}{1+\xi^{2}}d\xi dxdt
=\displaystyle= 0T2(1+ξ2)ξkf(t,x,ξ)φ1+ξ2𝑑ξ𝑑x𝑑t\displaystyle\int_{0}^{T}\int_{\mathbb{R}^{2}}(1+\xi^{2})\xi^{k}f(t,x,\xi)\frac{\varphi}{1+\xi^{2}}d\xi dxdt
=\displaystyle= 0T(ξkf(t,x,ξ)𝑑ξ)φ𝑑x𝑑t,\displaystyle\int_{0}^{T}\int_{\mathbb{R}}\left(\int_{\mathbb{R}}\xi^{k}f(t,x,\xi)d\xi\right)\varphi dxdt,

where in the second-to-last equality we have used convergence (4.23) and the fact that

11+ξ2φ(x)Hx,ξ3.\displaystyle\frac{1}{1+\xi^{2}}\varphi(x)\in H_{x,\xi}^{3}.

Hence, we conclude (4.10) for φLt1([0,T];Cc())\varphi\in L_{t}^{1}([0,T];C_{c}^{\infty}(\mathbb{R})). By the weighted uniform estimate (3.11) and the fact that Cc()C_{c}^{\infty}(\mathbb{R}) is dense in Cc()C_{c}(\mathbb{R}), we arrive at (4.10) for φLt1([0,T];Cc())\varphi\in L_{t}^{1}([0,T];C_{c}(\mathbb{R})). Moreover, for φLt1([0,T];Cb())\varphi\in L_{t}^{1}([0,T];C_{b}(\mathbb{R})), we write

(4.24) 0T(ξkf,ε(t,x,ξ)𝑑ξ)φ𝑑x𝑑t\displaystyle\int_{0}^{T}\int_{\mathbb{R}}\left(\int_{\mathbb{R}}\xi^{k}f_{\hbar,\varepsilon}(t,x,\xi)d\xi\right)\varphi dxdt
=\displaystyle= 0T(ξkf,ε(t,x,ξ)𝑑ξ)φχ(xR)𝑑x𝑑t\displaystyle\int_{0}^{T}\int_{\mathbb{R}}\left(\int_{\mathbb{R}}\xi^{k}f_{\hbar,\varepsilon}(t,x,\xi)d\xi\right)\varphi\chi(\frac{x}{R})dxdt
+0T(ξkf,ε(t,x,ξ)𝑑ξ)φ(1χ(xR))𝑑x𝑑t.\displaystyle+\int_{0}^{T}\int_{\mathbb{R}}\left(\int_{\mathbb{R}}\xi^{k}f_{\hbar,\varepsilon}(t,x,\xi)d\xi\right)\varphi(1-\chi(\frac{x}{R}))dxdt.

By the weighted uniform estimate (3.11), we get

|0T(ξkf,ε(t,x,ξ)𝑑ξ)φ(1χ(xR))𝑑x𝑑t|\displaystyle\Big{|}\int_{0}^{T}\int_{\mathbb{R}}\left(\int_{\mathbb{R}}\xi^{k}f_{\hbar,\varepsilon}(t,x,\xi)d\xi\right)\varphi(1-\chi(\frac{x}{R}))dxdt\Big{|}
\displaystyle\leq 1Rxξkf,ε(t,x,ξ)𝑑ξLtLx1φLt1Lx1R0.\displaystyle\frac{1}{R}\Big{\|}\langle x\rangle\int_{\mathbb{R}}\xi^{k}f_{\hbar,\varepsilon}(t,x,\xi)d\xi\Big{\|}_{L_{t}^{\infty}L_{x}^{1}}\|\varphi\|_{L_{t}^{1}L_{x}^{\infty}}\lesssim\frac{1}{R}\to 0.

Taking (,ε)(0,0)(\hbar,\varepsilon)\to(0,0) and then sending RR\to\infty, (4.24) becomes

lim(,ε)(0,0)0T(ξkf,ε(t,x,ξ)𝑑ξ)φ𝑑x𝑑t\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{0}^{T}\int_{\mathbb{R}}\left(\int_{\mathbb{R}}\xi^{k}f_{\hbar,\varepsilon}(t,x,\xi)d\xi\right)\varphi dxdt
=\displaystyle= limR0T(ξkf(t,x,ξ)𝑑ξ)φχ(xR)𝑑x𝑑t\displaystyle\lim_{R\to\infty}\int_{0}^{T}\int_{\mathbb{R}}\left(\int_{\mathbb{R}}\xi^{k}f(t,x,\xi)d\xi\right)\varphi\chi(\frac{x}{R})dxdt
=\displaystyle= 0T(ξkf(t,x,ξ)𝑑ξ)φ𝑑x𝑑t,\displaystyle\int_{0}^{T}\int_{\mathbb{R}}\left(\int_{\mathbb{R}}\xi^{k}f(t,x,\xi)d\xi\right)\varphi dxdt,

where in the last equality we have used the dominated convergence theorem. Hence, we complete the proof of (4.10).

Next, we handle estimate (4.11). As we have proven that

lim(,ε)(0,0)ξkf~,ε(t,x,ξ)𝑑ξξkf(t,x,ξ)𝑑ξC([0,T];Hx3)=0,\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\Big{\|}\int_{\mathbb{R}}\xi^{k}\widetilde{f}_{\hbar,\varepsilon}(t,x,\xi)d\xi-\int_{\mathbb{R}}\xi^{k}f(t,x,\xi)d\xi\Big{\|}_{C([0,T];H_{x}^{-3})}=0,

by the weighted uniform bound (3.11), we can improve it to the narrow convergence in a similar way in which we obtain (4.10) and hence complete the proof of (4.11).

Finally, we deal with the estimates (4.12)–(4.13). It suffices to prove estimate (4.13), as estimate (4.12) follows similarly. By the non-negativity of f(t,dx,dξ)f(t,dx,d\xi) and Hölder’s inequality, we again only need to prove (4.13) with the weight function ξ2k\xi^{2k}. By the weighted uniform estimate (3.11), we have

2χ(xR)xξ2kf(t,dx,dξ)=\displaystyle\iint_{\mathbb{R}^{2}}\chi(\frac{x}{R})\langle x\rangle\xi^{2k}f(t,dx,d\xi)= lim(,ε)(0,0)χ(xR)x(ξ2kf,ε(t,x,ξ)𝑑ξ)𝑑x\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\mathbb{R}}\chi(\frac{x}{R})\langle x\rangle\left(\int_{\mathbb{R}}\xi^{2k}f_{\hbar,\varepsilon}(t,x,\xi)d\xi\right)dx
\displaystyle\leq sup(,ε)xξ2kf,ε(t,x,ξ)𝑑ξLx1C(2k,t).\displaystyle\sup_{(\hbar,\varepsilon)}\Big{\|}\langle x\rangle\int_{\mathbb{R}}\xi^{2k}f_{\hbar,\varepsilon}(t,x,\xi)d\xi\Big{\|}_{L_{x}^{1}}\leq C(2k,t).

Sending RR\to\infty, by Fatou’s lemma, we arrive at (4.13). ∎

5. Conservation Laws for the Limit Measure

The conservation laws of mass, momentum and energy for the Wigner function f,ε(t,x,ξ)f_{\hbar,\varepsilon}(t,x,\xi) are given by

(5.1) 2f,ε(t,x,ξ)𝑑ξ𝑑x=2f,ε(0,x,ξ)𝑑ξ𝑑x,\displaystyle\iint_{\mathbb{R}^{2}}f_{\hbar,\varepsilon}(t,x,\xi)d\xi dx=\iint_{\mathbb{R}^{2}}f_{\hbar,\varepsilon}(0,x,\xi)d\xi dx,
(5.2) 2ξf,ε(t,x,ξ)𝑑ξ𝑑x=2ξf,ε(0,x,ξ)𝑑ξ𝑑x,\displaystyle\iint_{\mathbb{R}^{2}}\xi f_{\hbar,\varepsilon}(t,x,\xi)d\xi dx=\iint_{\mathbb{R}^{2}}\xi f_{\hbar,\varepsilon}(0,x,\xi)d\xi dx,
(5.3) 2ξ2f,ε(t,x,ξ)𝑑ξ𝑑x+2Vε(xy)ρ,ε(t,x)ρ,ε(t,y)𝑑x𝑑y\displaystyle\iint_{\mathbb{R}^{2}}\xi^{2}f_{\hbar,\varepsilon}(t,x,\xi)d\xi dx+\iint_{\mathbb{R}^{2}}V_{\varepsilon}(x-y)\rho_{\hbar,\varepsilon}(t,x)\rho_{\hbar,\varepsilon}(t,y)dxdy
=\displaystyle= 2ξ2f,ε(0,x,ξ)𝑑ξ𝑑x+2Vε(xy)ρ,ε(0,x)ρ,ε(0,y)𝑑x𝑑y.\displaystyle\iint_{\mathbb{R}^{2}}\xi^{2}f_{\hbar,\varepsilon}(0,x,\xi)d\xi dx+\iint_{\mathbb{R}^{2}}V_{\varepsilon}(x-y)\rho_{\hbar,\varepsilon}(0,x)\rho_{\hbar,\varepsilon}(0,y)dxdy.

In the section, we prove the conservation laws for the limit measure f(t,dx,dξ)f(t,dx,d\xi).

Lemma 5.1.

The limit measure f(t,dx,dξ)f(t,dx,d\xi) satisfies the conservation laws of mass, momentum and energy

(5.4) 2f(t,dx,dξ)=2f(0,dx,dξ),\displaystyle\iint_{\mathbb{R}^{2}}f(t,dx,d\xi)=\iint_{\mathbb{R}^{2}}f(0,dx,d\xi),
(5.5) 2ξf(t,dx,dξ)=2ξf(0,dx,dξ),\displaystyle\iint_{\mathbb{R}^{2}}\xi f(t,dx,d\xi)=\iint_{\mathbb{R}^{2}}\xi f(0,dx,d\xi),
(5.6) 2ξ2f(t,dx,dξ)𝑑x𝑑ξ+122|xy|ρ(t,dx)ρ(t,dy)\displaystyle\iint_{\mathbb{R}^{2}}\xi^{2}f(t,dx,d\xi)dxd\xi+\frac{1}{2}\int_{\mathbb{R}^{2}}|x-y|\rho(t,dx)\rho(t,dy)
=\displaystyle= 2ξ2f(0,dx,dξ)𝑑x𝑑ξ+122|xy|ρ(0,dx)ρ(0,dy),\displaystyle\iint_{\mathbb{R}^{2}}\xi^{2}f(0,dx,d\xi)dxd\xi+\frac{1}{2}\int_{\mathbb{R}^{2}}|x-y|\rho(0,dx)\rho(0,dy),

where ρ(t,dx)=f(t,dx,ξ)𝑑ξ\rho(t,dx)=\int_{\mathbb{R}}f(t,dx,\xi)d\xi.

Proof.

For (5.4), by the narrow convergence (4.11) in Lemma 4.3 and the conservation law of mass (5.1) for f,ε(t,x,ξ)f_{\hbar,\varepsilon}(t,x,\xi), we have

2f(t,dx,dξ)=\displaystyle\iint_{\mathbb{R}^{2}}f(t,dx,d\xi)= lim(,ε)(0,0)2f,ε(t,x,ξ)𝑑x𝑑ξ\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\iint_{\mathbb{R}^{2}}f_{\hbar,\varepsilon}(t,x,\xi)dxd\xi
=\displaystyle= lim(,ε)(0,0)2f,ε(0,x,ξ)𝑑x𝑑ξ\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\iint_{\mathbb{R}^{2}}f_{\hbar,\varepsilon}(0,x,\xi)dxd\xi
=\displaystyle= 2f(0,dx,dξ).\displaystyle\iint_{\mathbb{R}^{2}}f(0,dx,d\xi).

In the same way, we also have the conservation law of momentum (5.5).

For the conservation law of energy (5.6), using again the narrow convergence (4.11) in Lemma 4.3, we obtain the convergence for the kinetic energy part

2ξ2f(t,dx,dξ)=\displaystyle\iint_{\mathbb{R}^{2}}\xi^{2}f(t,dx,d\xi)= lim(,ε)(0,0)2ξ2f,ε(t,x,ξ)𝑑x𝑑ξ,\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\iint_{\mathbb{R}^{2}}\xi^{2}f_{\hbar,\varepsilon}(t,x,\xi)dxd\xi,
2ξ2f(0,dx,dξ)=\displaystyle\iint_{\mathbb{R}^{2}}\xi^{2}f(0,dx,d\xi)= lim(,ε)(0,0)2ξ2f,ε(0,x,ξ)𝑑x𝑑ξ.\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\iint_{\mathbb{R}^{2}}\xi^{2}f_{\hbar,\varepsilon}(0,x,\xi)dxd\xi.

Next, we deal with the potential energy part. For simplicity, we omit the time variable and rewrite

2|xy|eε|xy|ρ,ε(x)ρ,ε(y)𝑑x𝑑y\displaystyle\iint_{\mathbb{R}^{2}}|x-y|e^{-\varepsilon|x-y|}\rho_{\hbar,\varepsilon}(x)\rho_{\hbar,\varepsilon}(y)dxdy
=\displaystyle= 2|xy|ρ,ε(x)ρ,ε(y)𝑑x𝑑y+2|xy|(1eε|xy|)ρ,ε(x)ρ,ε(y)𝑑x𝑑y.\displaystyle\iint_{\mathbb{R}^{2}}|x-y|\rho_{\hbar,\varepsilon}(x)\rho_{\hbar,\varepsilon}(y)dxdy+\iint_{\mathbb{R}^{2}}|x-y|(1-e^{-\varepsilon|x-y|})\rho_{\hbar,\varepsilon}(x)\rho_{\hbar,\varepsilon}(y)dxdy.

By the weighted uniform estimate (3.1), we have

2|xy|(1eε|xy|)ρ,ε(x)ρ,ε(y)𝑑x𝑑y\displaystyle\iint_{\mathbb{R}^{2}}|x-y|(1-e^{-\varepsilon|x-y|})\rho_{\hbar,\varepsilon}(x)\rho_{\hbar,\varepsilon}(y)dxdy
\displaystyle\leq ε2|xy|2ρ,ε(x)ρ,ε(y)𝑑x𝑑y\displaystyle\varepsilon\iint_{\mathbb{R}^{2}}|x-y|^{2}\rho_{\hbar,\varepsilon}(x)\rho_{\hbar,\varepsilon}(y)dxdy
\displaystyle\lesssim ε(|x|2ρ,ε(x)𝑑x)(ρ,ε(y)𝑑y)ε0.\displaystyle\varepsilon\left(\int_{\mathbb{R}}|x|^{2}\rho_{\hbar,\varepsilon}(x)dx\right)\left(\int_{\mathbb{R}}\rho_{\hbar,\varepsilon}(y)dy\right)\lesssim\varepsilon\to 0.

Thus, we are left to consider the convergence of the term

(5.7) 2|xy|ρ,ε(x)ρ,ε(y)𝑑x𝑑y,\displaystyle\iint_{\mathbb{R}^{2}}|x-y|\rho_{\hbar,\varepsilon}(x)\rho_{\hbar,\varepsilon}(y)dxdy,

To do this, we introduce a weighted transform. Set

(5.8) F,ε(x)=\displaystyle F_{\hbar,\varepsilon}(x)= |xy|x1+δρ,ε(y)𝑑y,F,ε(±)=0,\displaystyle\int_{\mathbb{R}}\frac{|x-y|}{\langle x\rangle^{1+\delta}}\rho_{\hbar,\varepsilon}(y)dy,\quad F_{\hbar,\varepsilon}(\pm\infty)=0,
(5.9) G,ε(x)=\displaystyle G_{\hbar,\varepsilon}(x)= xy1+δρ,ε(y)𝑑y,\displaystyle\int_{-\infty}^{x}\langle y\rangle^{1+\delta}\rho_{\hbar,\varepsilon}(y)dy,

where δ(0,1)\delta\in(0,1) is a fixed constant. After the weighted transform, we can use the integration by parts to get

2|xy|ρ,ε(x)ρ,ε(y)𝑑x𝑑y=\displaystyle\iint_{\mathbb{R}^{2}}|x-y|\rho_{\hbar,\varepsilon}(x)\rho_{\hbar,\varepsilon}(y)dxdy= F,ε(x)xG,ε(x)dx\displaystyle\int_{\mathbb{R}}F_{\hbar,\varepsilon}(x)\partial_{x}G_{\hbar,\varepsilon}(x)dx
=\displaystyle= F,ε(x)G,ε(x)|+G,ε(x)xF,ε(x)dx\displaystyle F_{\hbar,\varepsilon}(x)G_{\hbar,\varepsilon}(x)|^{+\infty}_{-\infty}-\int_{\mathbb{R}}G_{\hbar,\varepsilon}(x)\partial_{x}F_{\hbar,\varepsilon}(x)dx
=\displaystyle= G,ε(x)xF,ε(x)dx.\displaystyle-\int_{\mathbb{R}}G_{\hbar,\varepsilon}(x)\partial_{x}F_{\hbar,\varepsilon}(x)dx.

Next, we get into the analysis of G,εG_{\hbar,\varepsilon} and F,εF_{\hbar,\varepsilon}. Using the weighted uniform estimate (3.1), we have

G,εLxC,xG,εLx1C.\displaystyle\|G_{\hbar,\varepsilon}\|_{L_{x}^{\infty}}\leq C,\quad\|\partial_{x}G_{\hbar,\varepsilon}\|_{L_{x}^{1}}\leq C.

Together with the LpL^{p} compactness criteria for 1p<1\leq p<\infty, we conclude that there exist a subsequence of {G,ε}\left\{G_{\hbar,\varepsilon}\right\} and an LlocpL_{loc}^{p} function G(x)G(x) such that

(5.10) G,εLlocpG.\displaystyle G_{\hbar,\varepsilon}\stackrel{{\scriptstyle L_{loc}^{p}}}{{\longrightarrow}}G.

In a similar way, noticing that

(5.11) xF,ε=1x1+δxy|xy|ρ,ε(y)𝑑y(1+δ)xx3+δ|xy|ρ,ε(y)𝑑y,\displaystyle\partial_{x}F_{\hbar,\varepsilon}=\int_{\mathbb{R}}\frac{1}{\langle x\rangle^{1+\delta}}\frac{x-y}{|x-y|}\rho_{\hbar,\varepsilon}(y)dy-(1+\delta)\int_{\mathbb{R}}\frac{x}{\langle x\rangle^{3+\delta}}|x-y|\rho_{\hbar,\varepsilon}(y)dy,

and

x2F,ε=I,ε(1)+I,ε(2)+I,ε(3)+I,ε(4),\displaystyle\partial_{x}^{2}F_{\hbar,\varepsilon}=I_{\hbar,\varepsilon}^{(1)}+I_{\hbar,\varepsilon}^{(2)}+I_{\hbar,\varepsilon}^{(3)}+I_{\hbar,\varepsilon}^{(4)},

where

I,ε(1)=\displaystyle I_{\hbar,\varepsilon}^{(1)}= 2ρ,ε(x)x1+δ,\displaystyle\frac{2\rho_{\hbar,\varepsilon}(x)}{\langle x\rangle^{1+\delta}},
I,ε(2)=\displaystyle I_{\hbar,\varepsilon}^{(2)}= (1+δ)xx3+δxy|xy|ρ,ε(y)𝑑y,\displaystyle-(1+\delta)\int_{\mathbb{R}}\frac{x}{\langle x\rangle^{3+\delta}}\frac{x-y}{|x-y|}\rho_{\hbar,\varepsilon}(y)dy,
I,ε(3)=\displaystyle I_{\hbar,\varepsilon}^{(3)}= (1+δ)xx3+δxy|xy|ρ,ε(y)𝑑y,\displaystyle-(1+\delta)\int_{\mathbb{R}}\frac{x}{\langle x\rangle^{3+\delta}}\frac{x-y}{|x-y|}\rho_{\hbar,\varepsilon}(y)dy,
I,ε(4)=\displaystyle I_{\hbar,\varepsilon}^{(4)}= (1+δ)(1x3+δ(3+δ)x2x5+δ)xy|xy|ρ,ε(y)𝑑y,\displaystyle-(1+\delta)\int_{\mathbb{R}}\left(\frac{1}{\langle x\rangle^{3+\delta}}-(3+\delta)\frac{x^{2}}{\langle x\rangle^{5+\delta}}\right)\frac{x-y}{|x-y|}\rho_{\hbar,\varepsilon}(y)dy,

we use again the weighted uniform estimate (3.1) to get

xF,εLx1LxC,x2F,εLx1C,\displaystyle\|\partial_{x}F_{\hbar,\varepsilon}\|_{L_{x}^{1}\cap L_{x}^{\infty}}\leq C,\quad\|\partial_{x}^{2}F_{\hbar,\varepsilon}\|_{L_{x}^{1}}\leq C,

which implies that there exist a subsequence of {xF,ε}\left\{\partial_{x}F_{\hbar,\varepsilon}\right\} and an LlocpL_{loc}^{p} function which we denote by xF(x)\partial_{x}F(x) such that

(5.12) xF,εLlocpxF.\displaystyle\partial_{x}F_{\hbar,\varepsilon}\stackrel{{\scriptstyle L_{loc}^{p}}}{{\longrightarrow}}\partial_{x}F.

In the following, we identify the limits in (5.10) and (5.12) by

(5.13) G(x)=\displaystyle G(x)= xy1+δρ(dy),a.e.,\displaystyle\int_{-\infty}^{x}\langle y\rangle^{1+\delta}\rho(dy),\quad a.e.,
(5.14) xF(x)=\displaystyle\partial_{x}F(x)= x|xy|x1+δρ(dy),a.e..\displaystyle\partial_{x}\int_{\mathbb{R}}\frac{|x-y|}{\langle x\rangle^{1+\delta}}\rho(dy),\quad a.e..

By the weighted estimate (4.12) and the fact that |xy|x1+δρ(dy)\int_{\mathbb{R}}\frac{|x-y|}{\langle x\rangle^{1+\delta}}\rho(dy) is Lipshitz continuous, (5.13) and (5.14) are indeed well-defined.

Take a test function φ(x)Cc()\varphi(x)\in C_{c}^{\infty}(\mathbb{R}), we have

φ(x)G,ε(x)𝑑x=\displaystyle\int_{\mathbb{R}}\varphi(x)G_{\hbar,\varepsilon}(x)dx= (y+φ(x)𝑑x)y1+δρ,ε(y)𝑑y\displaystyle\int_{\mathbb{R}}\left(\int_{y}^{+\infty}\varphi(x)dx\right)\langle y\rangle^{1+\delta}\rho_{\hbar,\varepsilon}(y)dy
=\displaystyle= A1+A2,\displaystyle A_{1}+A_{2},

where

A1=\displaystyle A_{1}= χ(yR)(y+φ(x)𝑑x)y1+δρ,ε(y)𝑑y,\displaystyle\int_{\mathbb{R}}\chi(\frac{y}{R})\left(\int_{y}^{+\infty}\varphi(x)dx\right)\langle y\rangle^{1+\delta}\rho_{\hbar,\varepsilon}(y)dy,
A2=\displaystyle A_{2}= (1χ(yR))(y+φ(x)𝑑x)y1+δρ,ε(y)𝑑y.\displaystyle\int_{\mathbb{R}}(1-\chi(\frac{y}{R}))\left(\int_{y}^{+\infty}\varphi(x)dx\right)\langle y\rangle^{1+\delta}\rho_{\hbar,\varepsilon}(y)dy.

Using the weighted uniform estimate (3.1), we obtain

|A2|1R1δφLx1y2ρ,εLy10.\displaystyle|A_{2}|\leq\frac{1}{R^{1-\delta}}\|\varphi\|_{L_{x}^{1}}\|\langle y\rangle^{2}\rho_{\hbar,\varepsilon}\|_{L_{y}^{1}}\to 0.

Therefore, letting first (,ε)(0,0)(\hbar,\varepsilon)\to(0,0) and then RR\to\infty, by the convergence (4.11), we get

(5.15) lim(,ε)(0,0)φ(x)G,ε(x)𝑑x=\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\mathbb{R}}\varphi(x)G_{\hbar,\varepsilon}(x)dx= limRχ(yR)(y+φ(x)𝑑x)y1+δρ(dy)\displaystyle\lim_{R\to\infty}\int_{\mathbb{R}}\chi(\frac{y}{R})\left(\int_{y}^{+\infty}\varphi(x)dx\right)\langle y\rangle^{1+\delta}\rho(dy)
=\displaystyle= (y+φ(x)𝑑x)y1+δρ(dy)\displaystyle\int_{\mathbb{R}}\left(\int_{y}^{+\infty}\varphi(x)dx\right)\langle y\rangle^{1+\delta}\rho(dy)
=\displaystyle= φ(x)(xy1+δρ(dy))𝑑x,\displaystyle\int_{\mathbb{R}}\varphi(x)\left(\int_{-\infty}^{x}\langle y\rangle^{1+\delta}\rho(dy)\right)dx,

where in the second and last equalities we have used the dominated convergence theorem and Fubini’s theorem based on the weighted estimate (4.12) that

y2ρ(t,dy)=2y2f(t,dy,dξ)C(t).\displaystyle\int_{\mathbb{R}}\langle y\rangle^{2}\rho(t,dy)=\iint_{\mathbb{R}^{2}}\langle y\rangle^{2}f(t,dy,d\xi)\leq C(t).

Hence, we complete the proof of (5.13) for G(x)G(x).

For (5.14), by (5.11) we have

φ(x)xF,ε(x)dx\displaystyle\int_{\mathbb{R}}\varphi(x)\partial_{x}F_{\hbar,\varepsilon}(x)dx
=\displaystyle= ((xy|xy|φ(x)x1+δ(1+δ)|xy|xφ(x)x3+δ)𝑑x)ρ,ε(y)𝑑y.\displaystyle\int_{\mathbb{R}}\left(\int_{\mathbb{R}}\left(\frac{x-y}{|x-y|}\frac{\varphi(x)}{\langle x\rangle^{1+\delta}}-(1+\delta)|x-y|\frac{x\varphi(x)}{\langle x\rangle^{3+\delta}}\right)dx\right)\rho_{\hbar,\varepsilon}(y)dy.

In a similar fashion as in (5.15), we get

lim(,ε)(0,0)φ(x)xF,ε(x)dx\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\mathbb{R}}\varphi(x)\partial_{x}F_{\hbar,\varepsilon}(x)dx
=\displaystyle= ((xy|xy|φ(x)x1+δ(1+δ)|xy|xφ(x)x3+δ)𝑑x)ρ(dy)\displaystyle\int_{\mathbb{R}}\left(\int_{\mathbb{R}}\left(\frac{x-y}{|x-y|}\frac{\varphi(x)}{\langle x\rangle^{1+\delta}}-(1+\delta)|x-y|\frac{x\varphi(x)}{\langle x\rangle^{3+\delta}}\right)dx\right)\rho(dy)
=\displaystyle= φ(x)(1x1+δxy|xy|ρ(dy)(1+δ)xx3+δ|xy|ρ(dy))𝑑x\displaystyle\int_{\mathbb{R}}\varphi(x)\left(\int_{\mathbb{R}}\frac{1}{\langle x\rangle^{1+\delta}}\frac{x-y}{|x-y|}\rho(dy)-(1+\delta)\int_{\mathbb{R}}\frac{x}{\langle x\rangle^{3+\delta}}|x-y|\rho(dy)\right)dx
=\displaystyle= φ(x)(x|xy|x1+δρ(dy))𝑑x,\displaystyle\int_{\mathbb{R}}\varphi(x)\left(\partial_{x}\int_{\mathbb{R}}\frac{|x-y|}{\langle x\rangle^{1+\delta}}\rho(dy)\right)dx,

where in the last inequality we have used Leibniz rule for the Lipschitz continuous function. This completes the proof of (5.14).

Finally, we prove the convergence of the potential energy part to the desired form. By (5.11) and the weighted uniform estimate (3.1), we notice that

xδ2xF,εLx11x1+δ2𝑑xxρ,εLx1C,\displaystyle\|\langle x\rangle^{\frac{\delta}{2}}\partial_{x}F_{\hbar,\varepsilon}\|_{L_{x}^{1}}\lesssim\int_{\mathbb{R}}\frac{1}{\langle x\rangle^{1+\frac{\delta}{2}}}dx\|\langle x\rangle\rho_{\hbar,\varepsilon}\|_{L_{x}^{1}}\leq C,

and hence obtain

(5.16) ||x|RG,ε(x)xF,ε(x)dx|1Rδ2G,εLxxδ2xF,εLx1CRδ20.\displaystyle\Big{|}\int_{|x|\geq R}G_{\hbar,\varepsilon}(x)\partial_{x}F_{\hbar,\varepsilon}(x)dx\Big{|}\leq\frac{1}{R^{\frac{\delta}{2}}}\|G_{\hbar,\varepsilon}\|_{L_{x}^{\infty}}\|\langle x\rangle^{\frac{\delta}{2}}\partial_{x}F_{\hbar,\varepsilon}\|_{L_{x}^{1}}\leq\frac{C}{R^{\frac{\delta}{2}}}\to 0.

As G,ε(x)Lloc2GG_{\hbar,\varepsilon}(x)\stackrel{{\scriptstyle L_{loc}^{2}}}{{\to}}G, xF,εLloc2xF\partial_{x}F_{\hbar,\varepsilon}\stackrel{{\scriptstyle L_{loc}^{2}}}{{\to}}\partial_{x}F, letting first (,ε)(0,0)(\hbar,\varepsilon)\to(0,0) and then RR\to\infty, we use (5.16) and the dominated convergence theorem to get

lim(,ε)(0,0)G,ε(x)xF,ε(x)dx\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\mathbb{R}}G_{\hbar,\varepsilon}(x)\partial_{x}F_{\hbar,\varepsilon}(x)dx
=\displaystyle= limRlim(,ε)(0,0)χ(xR)G,ε(x)xF,ε(x)dx\displaystyle\lim_{R\to\infty}\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\mathbb{R}}\chi(\frac{x}{R})G_{\hbar,\varepsilon}(x)\partial_{x}F_{\hbar,\varepsilon}(x)dx
+limRlim(,ε)(0,0)(1χ(xR))G,ε(x)xF,ε(x)dx\displaystyle+\lim_{R\to\infty}\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\mathbb{R}}(1-\chi(\frac{x}{R}))G_{\hbar,\varepsilon}(x)\partial_{x}F_{\hbar,\varepsilon}(x)dx
=\displaystyle= limRχ(xR)G(x)xF(x)dx\displaystyle\lim_{R\to\infty}\int_{\mathbb{R}}\chi(\frac{x}{R})G(x)\partial_{x}F(x)dx
=\displaystyle= G(x)xF(x)dx.\displaystyle\int_{\mathbb{R}}G(x)\partial_{x}F(x)dx.

By Fubini’s theorem, we obtain

G(x)xF(x)dx=\displaystyle\int_{\mathbb{R}}G(x)\partial_{x}F(x)dx= (1{yx}xF(x)dx)y1+δρ(dy)\displaystyle\int_{\mathbb{R}}\left(\int_{\mathbb{R}}1_{\left\{y\leq x\right\}}\partial_{x}F(x)dx\right)\langle y\rangle^{1+\delta}\rho(dy)
=\displaystyle= |xy|y1+δρ(dx)y1+δρ(dy)\displaystyle-\int_{\mathbb{R}}\int_{\mathbb{R}}\frac{|x-y|}{\langle y\rangle^{1+\delta}}\rho(dx)\langle y\rangle^{1+\delta}\rho(dy)
=\displaystyle= 2|xy|ρ(dx)ρ(dy).\displaystyle-\iint_{\mathbb{R}^{2}}|x-y|\rho(dx)\rho(dy).

Therefore, together with the convergence of the kinetic energy part, we complete the proof of the conservation law of energy (5.6).

6. Moment Convergence to the Vlasov-Poisson Equation

In the section, our goal is to prove the convergence of some subsequence of f,ε(t,x,ξ)f_{\hbar,\varepsilon}(t,x,\xi) to the Vlasov-Poisson equation for the test functions of the moment form

(6.1) ϕ(t,x,ξ)=φ(t,x)ξk.\displaystyle\phi(t,x,\xi)=\varphi(t,x)\xi^{k}.

That is, the limit point f(t,x,ξ)f(t,x,\xi) satisfies the Vlasov-Poisson equation for the test functions of the form (6.1), based on which we extend it to all test function ϕCc((0,T)×2)\phi\in C_{c}^{\infty}((0,T)\times\mathbb{R}^{2}) in Section 7.

The main result of this section is Lemma 6.1 below.

Lemma 6.1.

Let T>0T>0 and k0k\geq 0. For φCc(ΩT)\varphi\in C_{c}^{\infty}(\Omega_{T}), there holds that

(6.2) ΩT(tφ+ξxφ)ξkf(t,dx,dξ)𝑑tkΩTφE¯(ξk1f(t,dx,dξ))𝑑t=0,\displaystyle\int_{\Omega_{T}}\int_{\mathbb{R}}\left(\partial_{t}\varphi+\xi\partial_{x}\varphi\right)\xi^{k}f(t,dx,d\xi)dt-k\int_{\Omega_{T}}\varphi\overline{E}\left(\int_{\mathbb{R}}\xi^{k-1}f(t,dx,d\xi)\right)dt=0,

where ΩT=(0,T)×\Omega_{T}=(0,T)\times\mathbb{R} and E¯\overline{E} is the Volpert’s symmetric average defined in (A.1).

To motivate the proof of Lemma 6.1, from the equation (4.16) of f,ε(t,x,ξ)f_{\hbar,\varepsilon}(t,x,\xi), we observe that the moment equation

(6.3) tξkf,ε𝑑ξ+xξk+1f,ε𝑑ξ+kE,εξk1f,ε𝑑ξ+R,ε(k)=0,\displaystyle\partial_{t}\int_{\mathbb{R}}\xi^{k}f_{\hbar,\varepsilon}d\xi+\partial_{x}\int_{\mathbb{R}}\xi^{k+1}f_{\hbar,\varepsilon}d\xi+kE_{\hbar,\varepsilon}\int_{\mathbb{R}}\xi^{k-1}f_{\hbar,\varepsilon}d\xi+\mathrm{R}_{\hbar,\varepsilon}^{(k)}=0,

where E,ε=xVερ,εE_{\hbar,\varepsilon}=\partial_{x}V_{\varepsilon}*\rho_{\hbar,\varepsilon} and the remainder term is

(6.4) R,ε(k)={i2αk(kα)α12k(1(1)α)Dxα(Vερ,ε)ξkαf,ε𝑑ξ,k3,0,k=0,1,2.\mathrm{R}_{\hbar,\varepsilon}^{(k)}=\left\{\begin{aligned} &i\sum_{2\leq\alpha\leq k}\binom{k}{\alpha}\frac{\hbar^{\alpha-1}}{2^{k}}(1-(-1)^{\alpha})D_{x}^{\alpha}(V_{\varepsilon}*\rho_{\hbar,\varepsilon})\int_{\mathbb{R}}\xi^{k-\alpha}f_{\hbar,\varepsilon}d\xi,\quad&k\geq 3,\\ &0,\quad&k=0,1,2.\end{aligned}\right.

By the convergence result in Lemma 4.3, we have the convergence for the linear term

lim(,ε)(0,0)ΩT(tφ+ξxφ)ξkf,ε𝑑x𝑑ξ𝑑t=\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\Omega_{T}}\int_{\mathbb{R}}\left(\partial_{t}\varphi+\xi\partial_{x}\varphi\right)\xi^{k}f_{\hbar,\varepsilon}dxd\xi dt= ΩT(tφ+ξxφ)ξkf(t,dx,dξ)𝑑t.\displaystyle\int_{\Omega_{T}}\int_{\mathbb{R}}\left(\partial_{t}\varphi+\xi\partial_{x}\varphi\right)\xi^{k}f(t,dx,d\xi)dt.

Therefore, to conclude Lemma 6.1, we are left to prove the vanishing of the remainder term and the convergence of the nonlinear term, that is,

(6.5) lim(,ε)(0,0)|ΩTφR,ε(k)𝑑x𝑑t|=0,\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\Big{|}\int_{\Omega_{T}}\varphi\mathrm{R}_{\hbar,\varepsilon}^{(k)}dxdt\Big{|}=0,
(6.6) lim(,ε)0ΩTφE,ε(ξk1f,ε𝑑ξ)𝑑x𝑑t=ΩTφE¯(ξk1f(t,dx,dξ))𝑑t.\displaystyle\lim_{(\hbar,\varepsilon)\to 0}\int_{\Omega_{T}}\varphi E_{\hbar,\varepsilon}\left(\int_{\mathbb{R}}\xi^{k-1}f_{\hbar,\varepsilon}d\xi\right)dxdt=\int_{\Omega_{T}}\varphi\overline{E}\left(\int_{\mathbb{R}}\xi^{k-1}f(t,dx,d\xi)\right)dt.

We deal with the remainder term and prove (6.5) in Section 6.1. The convergence of the nonlinear term is usually one of the main difficulties, as it is actually a problem of convergence of the product form in the mixed limit. We prove (6.6) for the k=1,2k=1,2 case in Section 6.2, and the general k3k\geq 3 case in Section 6.3.

6.1. Vanishing Remainder Terms via a Cancellation Structure

In the space of the strong topology, the remainder term R,ε(m)\mathrm{R}_{\hbar,\varepsilon}^{(m)} in (6.4) is only uniformly bounded in L([0,T];Lx1)L^{\infty}([0,T];L_{x}^{1}). We follow the idea of [50] to prove that the remainder term would vanish in the weak sense by an iteration scheme using a cancellation structure.

First, we provide the weighted uniform bound for the remainder term.

Lemma 6.2.

For T>0T>0 and m3m\geq 3, we have

(6.7) xR,ε(m)Lt([0,T];Lx1)C(m,T).\displaystyle\|\langle x\rangle\mathrm{R}_{\hbar,\varepsilon}^{(m)}\|_{L_{t}^{\infty}([0,T];L_{x}^{1})}\leq C(m,T).
Proof.

By the weighted uniform estimate (3.11) and the uniform bound (3.12), we use Hölder’s inequality to get

xR,ε(m)LtLx1\displaystyle\|\langle x\rangle\mathrm{R}_{\hbar,\varepsilon}^{(m)}\|_{L_{t}^{\infty}L_{x}^{1}}\lesssim α=0mxα1xα(Vερ,ε)ξmαf,ε𝑑ξLtLx1\displaystyle\sum_{\alpha=0}^{m}\Big{\|}\langle x\rangle\hbar^{\alpha-1}\partial_{x}^{\alpha}\left(V_{\varepsilon}*\rho_{\hbar,\varepsilon}\right)\int_{\mathbb{R}}\xi^{m-\alpha}f_{\hbar,\varepsilon}d\xi\Big{\|}_{L_{t}^{\infty}L_{x}^{1}}
\displaystyle\leq α1xαVερ,εLtLxxξmαf,ε𝑑ξLtLx1\displaystyle\|\hbar^{\alpha-1}\partial_{x}^{\alpha}V_{\varepsilon}*\rho_{\hbar,\varepsilon}\|_{L_{t}^{\infty}L_{x}^{\infty}}\Big{\|}\langle x\rangle\int_{\mathbb{R}}\xi^{m-\alpha}f_{\hbar,\varepsilon}d\xi\Big{\|}_{L_{t}^{\infty}L_{x}^{1}}
\displaystyle\leq C(m,T).\displaystyle C(m,T).

Next, we get into the analysis of the remainder term.

Lemma 6.3.

Let T>0T>0 and k3k\geq 3. For φ(t,x)Cc1(ΩT)\varphi(t,x)\in C_{c}^{1}(\Omega_{T}) and α=2n+1k\alpha=2n+1\leq k with n1n\geq 1, we have

(6.8) α1ΩTφxα(Vερ,ε)(ξkαf,ε𝑑ξ)dxdt=(,ε),\displaystyle\hbar^{\alpha-1}\int_{\Omega_{T}}\varphi\partial_{x}^{\alpha}(V_{\varepsilon}*\rho_{\hbar,\varepsilon})\left(\int_{\mathbb{R}}\xi^{k-\alpha}f_{\hbar,\varepsilon}d\xi\right)dxdt=\mathcal{E}(\hbar,\varepsilon),

with

(6.9) |(,ε)|C(k,α)(t,xφLt1Lx+φLt1Lx+εφLt1Lx).\displaystyle|\mathcal{E}(\hbar,\varepsilon)|\leq C(k,\alpha)\left(\hbar\|\nabla_{t,x}\varphi\|_{L_{t}^{1}L_{x}^{\infty}}+\hbar\|\varphi\|_{L_{t}^{1}L_{x}^{\infty}}+\varepsilon\|\varphi\|_{L_{t}^{1}L_{x}^{\infty}}\right).

In particular, we have the quantitative estimate that

(6.10) |ΩTφR,ε(k)𝑑x𝑑t|C(k)(t,xφLt1Lx+φLt1Lx+εφLt1Lx),φ(t,x)Cc1(ΩT),\displaystyle\Big{|}\int_{\Omega_{T}}\varphi\mathrm{R}_{\hbar,\varepsilon}^{(k)}dxdt\Big{|}\leq C(k)\left(\hbar\|\nabla_{t,x}\varphi\|_{L_{t}^{1}L_{x}^{\infty}}+\hbar\|\varphi\|_{L_{t}^{1}L_{x}^{\infty}}+\varepsilon\|\varphi\|_{L_{t}^{1}L_{x}^{\infty}}\right),\quad\forall\varphi(t,x)\in C_{c}^{1}(\Omega_{T}),

and have the qualitative convergence that

(6.11) lim(,ε)(0,0)|ΩTφR,ε(k)𝑑x𝑑t|=0,φ(t,x)Lt1([0,T];Cb()).\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\Big{|}\int_{\Omega_{T}}\varphi\mathrm{R}_{\hbar,\varepsilon}^{(k)}dxdt\Big{|}=0,\quad\forall\varphi(t,x)\in L_{t}^{1}([0,T];C_{b}(\mathbb{R})).
Proof.

For convenience, we take up the notation

(6.12) Uε(x):=\displaystyle U_{\varepsilon}(x):= 12|x|Vε(x)=12|x|(1eε|x|),\displaystyle\frac{1}{2}|x|-V_{\varepsilon}(x)=\frac{1}{2}|x|(1-e^{-\varepsilon|x|}),

and hence rewrite

(6.13) α1ΩTφxα(Vερ,ε)(ξkαf,ε𝑑ξ)dxdt=I1I2,\displaystyle\hbar^{\alpha-1}\int_{\Omega_{T}}\varphi\partial_{x}^{\alpha}(V_{\varepsilon}*\rho_{\hbar,\varepsilon})\left(\int_{\mathbb{R}}\xi^{k-\alpha}f_{\hbar,\varepsilon}d\xi\right)dxdt=I_{1}-I_{2},

where

I1=\displaystyle I_{1}= α1ΩTφ(xα|x|2ρ,ε)(ξkαf,ε𝑑ξ)𝑑x𝑑t,\displaystyle\hbar^{\alpha-1}\int_{\Omega_{T}}\varphi\left(\partial_{x}^{\alpha}\frac{|x|}{2}*\rho_{\hbar,\varepsilon}\right)\left(\int_{\mathbb{R}}\xi^{k-\alpha}f_{\hbar,\varepsilon}d\xi\right)dxdt,
I2=\displaystyle I_{2}= α1ΩTφ(xαUερ,ε)(ξkαf,ε𝑑ξ)𝑑x𝑑t.\displaystyle\hbar^{\alpha-1}\int_{\Omega_{T}}\varphi\left(\partial_{x}^{\alpha}U_{\varepsilon}*\rho_{\hbar,\varepsilon}\right)\left(\int_{\mathbb{R}}\xi^{k-\alpha}f_{\hbar,\varepsilon}d\xi\right)dxdt.

We first deal with the term I2I_{2}. Noting that

|xUε(x)|ε|x|a.e.,\displaystyle|\partial_{x}U_{\varepsilon}(x)|\leq\varepsilon|x|\quad a.e.,

we have the pointwise bound

(6.14) α1|xα(Uερ,ε)|=\displaystyle\hbar^{\alpha-1}|\partial_{x}^{\alpha}(U_{\varepsilon}*\rho_{\hbar,\varepsilon})|= α1|xUεxα1ρ,ε|\displaystyle\hbar^{\alpha-1}|\partial_{x}U_{\varepsilon}*\partial_{x}^{\alpha-1}\rho_{\hbar,\varepsilon}|
\displaystyle\leq α1ε|xy||xα1ρ,ε(y)|𝑑y\displaystyle\hbar^{\alpha-1}\varepsilon\int_{\mathbb{R}}|x-y||\partial_{x}^{\alpha-1}\rho_{\hbar,\varepsilon}(y)|dy
\displaystyle\leq εxxα1xα1ρ,εLx1\displaystyle\varepsilon\langle x\rangle\|\langle x\rangle\hbar^{\alpha-1}\partial_{x}^{\alpha-1}\rho_{\hbar,\varepsilon}\|_{L_{x}^{1}}
\displaystyle\lesssim εx,\displaystyle\varepsilon\langle x\rangle,

where in the last inequality we have used the uniform estimate (3.11). By (6.14), we then use Hölder’s inequality and the uniform estimate (3.11) to obtain

(6.15) I2εφLt1Lxxξkαf,ε𝑑ξLtLx1εφLt1Lx.\displaystyle I_{2}\lesssim\varepsilon\|\varphi\|_{L_{t}^{1}L_{x}^{\infty}}\Big{\|}\langle x\rangle\int_{\mathbb{R}}\xi^{k-\alpha}f_{\hbar,\varepsilon}d\xi\Big{\|}_{L_{t}^{\infty}L_{x}^{1}}\lesssim\varepsilon\|\varphi\|_{L_{t}^{1}L_{x}^{\infty}}.

Next, we handle the term I1I_{1} via an iteration scheme. For α=2n+1k\alpha=2n+1\leq k with n1n\geq 1, we set the notation

(6.16) Mφ(k,α,j)=α1ΩTφ(xα2ξjf,ε𝑑ξ)(ξkαjf,ε𝑑ξ)𝑑x𝑑t.\displaystyle M_{\varphi}^{(k,\alpha,j)}=\hbar^{\alpha-1}\int_{\Omega_{T}}\varphi\left(\partial_{x}^{\alpha-2}\int_{\mathbb{R}}\xi^{j}f_{\hbar,\varepsilon}d\xi\right)\left(\int_{\mathbb{R}}\xi^{k-\alpha-j}f_{\hbar,\varepsilon}d\xi\right)dxdt.

In particular, noticing that x2(|x|2)=δ(x)\partial_{x}^{2}(\frac{|x|}{2})=\delta(x), we have

I1=α1ΩTφxα2((x2|x|2)ρ,ε)(ξkαf,ε𝑑ξ)dxdt=Mφ(k,α,0).I_{1}=\hbar^{\alpha-1}\int_{\Omega_{T}}\varphi\partial_{x}^{\alpha-2}((\partial_{x}^{2}\frac{|x|}{2})*\rho_{\hbar,\varepsilon})\left(\int_{\mathbb{R}}\xi^{k-\alpha}f_{\hbar,\varepsilon}d\xi\right)dxdt=M_{\varphi}^{(k,\alpha,0)}.

In the following, we get into the analysis of Mφ(k,α,j)M_{\varphi}^{(k,\alpha,j)}. By integration by parts, we have

Mφ(k,α,j)=\displaystyle M_{\varphi}^{(k,\alpha,j)}= α1ΩTφ(xα3ξjf,ε𝑑ξ)(xξkαjf,ε𝑑ξ)𝑑x𝑑t\displaystyle-\hbar^{\alpha-1}\int_{\Omega_{T}}\varphi\left(\partial_{x}^{\alpha-3}\int_{\mathbb{R}}\xi^{j}f_{\hbar,\varepsilon}d\xi\right)\left(\partial_{x}\int_{\mathbb{R}}\xi^{k-\alpha-j}f_{\hbar,\varepsilon}d\xi\right)dxdt
α1ΩTxφ(xα3ξjf,ε𝑑ξ)(ξkαjf,ε𝑑ξ)dxdt.\displaystyle-\hbar^{\alpha-1}\int_{\Omega_{T}}\partial_{x}\varphi\left(\partial_{x}^{\alpha-3}\int_{\mathbb{R}}\xi^{j}f_{\hbar,\varepsilon}d\xi\right)\left(\int_{\mathbb{R}}\xi^{k-\alpha-j}f_{\hbar,\varepsilon}d\xi\right)dxdt.

Using the moment equation (6.3) in which we take m=kαj1m=k-\alpha-j-1, we have

(6.17) Mφ(k,α,j)=A1+A2+A3+A4,\displaystyle M_{\varphi}^{(k,\alpha,j)}=A_{1}+A_{2}+A_{3}+A_{4},

where

A1=\displaystyle A_{1}= α1ΩTφ(xα3ξjf,ε𝑑ξ)(tξmf,ε𝑑ξ)𝑑x𝑑t,\displaystyle\hbar^{\alpha-1}\int_{\Omega_{T}}\varphi\left(\partial_{x}^{\alpha-3}\int_{\mathbb{R}}\xi^{j}f_{\hbar,\varepsilon}d\xi\right)\left(\partial_{t}\int_{\mathbb{R}}\xi^{m}f_{\hbar,\varepsilon}d\xi\right)dxdt,
A2=\displaystyle A_{2}= α1ΩTφ(xα3ξjf,ε𝑑ξ)(mE,εξm1f,ε𝑑ξ)𝑑x𝑑t,\displaystyle\hbar^{\alpha-1}\int_{\Omega_{T}}\varphi\left(\partial_{x}^{\alpha-3}\int_{\mathbb{R}}\xi^{j}f_{\hbar,\varepsilon}d\xi\right)\left(mE_{\hbar,\varepsilon}\int_{\mathbb{R}}\xi^{m-1}f_{\hbar,\varepsilon}d\xi\right)dxdt,
A3=\displaystyle A_{3}= α1ΩTφ(xα3ξjf,ε𝑑ξ)(R,ε(m))𝑑x𝑑t,\displaystyle\hbar^{\alpha-1}\int_{\Omega_{T}}\varphi\left(\partial_{x}^{\alpha-3}\int_{\mathbb{R}}\xi^{j}f_{\hbar,\varepsilon}d\xi\right)\left(\mathrm{R}_{\hbar,\varepsilon}^{(m)}\right)dxdt,
A4=\displaystyle A_{4}= α1ΩTxφ(xα3ξjf,ε𝑑ξ)(ξkαjf,ε𝑑ξ)dxdt.\displaystyle-\hbar^{\alpha-1}\int_{\Omega_{T}}\partial_{x}\varphi\left(\partial_{x}^{\alpha-3}\int_{\mathbb{R}}\xi^{j}f_{\hbar,\varepsilon}d\xi\right)\left(\int_{\mathbb{R}}\xi^{k-\alpha-j}f_{\hbar,\varepsilon}d\xi\right)dxdt.

We can directly bound the terms A2A_{2}, A3A_{3} and A4A_{4}. For A2A_{2}, by Hölder’s equality, the uniform estimates (3.10) and (3.11), we have

|A2|\displaystyle|A_{2}|\lesssim φLt1Lxα2xα3ξjf,ε𝑑ξLt,xE,εLt,xξm1f,ε𝑑ξLtLx1\displaystyle\hbar\|\varphi\|_{L_{t}^{1}L_{x}^{\infty}}\Big{\|}\hbar^{\alpha-2}\partial_{x}^{\alpha-3}\int_{\mathbb{R}}\xi^{j}f_{\hbar,\varepsilon}d\xi\Big{\|}_{L_{t,x}^{\infty}}\|E_{\hbar,\varepsilon}\|_{L_{t,x}^{\infty}}\Big{\|}\int_{\mathbb{R}}\xi^{m-1}f_{\hbar,\varepsilon}d\xi\Big{\|}_{L_{t}^{\infty}L_{x}^{1}}
\displaystyle\lesssim φLt1Lx.\displaystyle\hbar\|\varphi\|_{L_{t}^{1}L_{x}^{\infty}}.

In a similar way, for A3A_{3} we use the uniform estimate (3.10) and the LtLx1L_{t}^{\infty}L_{x}^{1} bound for R,ε(m)\mathrm{R}_{\hbar,\varepsilon}^{(m)} to obtain

|A3|φLt1Lxα2xα3ξjf,ε𝑑ξLt,xR,ε(m)LtLx1\displaystyle|A_{3}|\leq\hbar\|\varphi\|_{L_{t}^{1}L_{x}^{\infty}}\Big{\|}\hbar^{\alpha-2}\partial_{x}^{\alpha-3}\int_{\mathbb{R}}\xi^{j}f_{\hbar,\varepsilon}d\xi\Big{\|}_{L_{t,x}^{\infty}}\|\mathrm{R}_{\hbar,\varepsilon}^{(m)}\|_{L_{t}^{\infty}L_{x}^{1}}\lesssim φLt1Lx.\displaystyle\hbar\|\varphi\|_{L_{t}^{1}L_{x}^{\infty}}.

In the same manner, we have

|A4|xφLt1Lx.\displaystyle|A_{4}|\lesssim\hbar\|\partial_{x}\varphi\|_{L_{t}^{1}L_{x}^{\infty}}.

Next, we get into the analysis of the term A1A_{1}. Using integration by parts in the time variable, we get

A1=A11+A12,\displaystyle A_{1}=A_{11}+A_{12},

where

A11=\displaystyle A_{11}= α1ΩTφ(xα3tξjf,ε𝑑ξ)(ξmf,ε𝑑ξ)𝑑x𝑑t,\displaystyle-\hbar^{\alpha-1}\int_{\Omega_{T}}\varphi\left(\partial_{x}^{\alpha-3}\partial_{t}\int_{\mathbb{R}}\xi^{j}f_{\hbar,\varepsilon}d\xi\right)\left(\int_{\mathbb{R}}\xi^{m}f_{\hbar,\varepsilon}d\xi\right)dxdt,
A12=\displaystyle A_{12}= α1ΩTtφ(xα3ξjf,ε𝑑ξ)(ξmf,ε𝑑ξ)dxdt.\displaystyle-\hbar^{\alpha-1}\int_{\Omega_{T}}\partial_{t}\varphi\left(\partial_{x}^{\alpha-3}\int_{\mathbb{R}}\xi^{j}f_{\hbar,\varepsilon}d\xi\right)\left(\int_{\mathbb{R}}\xi^{m}f_{\hbar,\varepsilon}d\xi\right)dxdt.

As the term A12A_{12} can be treated in a similar way as A2A_{2}, we have

|A12|tφLt1Lx.\displaystyle|A_{12}|\lesssim\hbar\|\partial_{t}\varphi\|_{L_{t}^{1}L_{x}^{\infty}}.

For the term A11A_{11}, we use again the moment equation (6.3) to get

A11=A111+A112+A113,\displaystyle A_{11}=A_{111}+A_{112}+A_{113},

where

A111=\displaystyle A_{111}= Mφ(k,α,j+1)=α1ΩTφ(xα2ξj+1f,ε𝑑ξ)(ξmf,ε𝑑ξ)𝑑x𝑑t,\displaystyle M_{\varphi}^{(k,\alpha,j+1)}=\hbar^{\alpha-1}\int_{\Omega_{T}}\varphi\left(\partial_{x}^{\alpha-2}\int_{\mathbb{R}}\xi^{j+1}f_{\hbar,\varepsilon}d\xi\right)\left(\int_{\mathbb{R}}\xi^{m}f_{\hbar,\varepsilon}d\xi\right)dxdt,
A112=\displaystyle A_{112}= jα1ΩTφ(xα3(E,εξj+1f,ε𝑑ξ))(ξmf,ε𝑑ξ)𝑑x𝑑t,\displaystyle j\hbar^{\alpha-1}\int_{\Omega_{T}}\varphi\left(\partial_{x}^{\alpha-3}\left(E_{\hbar,\varepsilon}\int_{\mathbb{R}}\xi^{j+1}f_{\hbar,\varepsilon}d\xi\right)\right)\left(\int_{\mathbb{R}}\xi^{m}f_{\hbar,\varepsilon}d\xi\right)dxdt,
A113=\displaystyle A_{113}= α1ΩTφ(xα3R,ε(j))(ξmf,ε𝑑ξ)𝑑x𝑑t.\displaystyle\hbar^{\alpha-1}\int_{\Omega_{T}}\varphi\left(\partial_{x}^{\alpha-3}\mathrm{R}_{\hbar,\varepsilon}^{(j)}\right)\left(\int_{\mathbb{R}}\xi^{m}f_{\hbar,\varepsilon}d\xi\right)dxdt.

For the term A112A_{112}, by the uniform estimates (3.10) and (3.12), we obtain

|A112|φLt1Lx.\displaystyle|A_{112}|\lesssim\hbar\|\varphi\|_{L_{t}^{1}L_{x}^{\infty}}.

For the term A113A_{113}, using Leibniz rule, the uniform estimates (3.10) and (3.12), we have

α3xα3R,ε(j)Lx1C(j,α,t),\displaystyle\|\hbar^{\alpha-3}\partial_{x}^{\alpha-3}\mathrm{R}_{\hbar,\varepsilon}^{(j)}\|_{L_{x}^{1}}\lesssim C(j,\alpha,t),

and hence obtain

|A113|φLt1Lxα3xα3R,ε(j)LtLx1ξmf,ε𝑑ξLt,xφLt1Lx.\displaystyle|A_{113}|\leq\hbar\|\varphi\|_{L_{t}^{1}L_{x}^{\infty}}\|\hbar^{\alpha-3}\partial_{x}^{\alpha-3}\mathrm{R}_{\hbar,\varepsilon}^{(j)}\|_{L_{t}^{\infty}L_{x}^{1}}\Big{\|}\hbar\int_{\mathbb{R}}\xi^{m}f_{\hbar,\varepsilon}d\xi\Big{\|}_{L_{t,x}^{\infty}}\lesssim\hbar\|\varphi\|_{L_{t}^{1}L_{x}^{\infty}}.

To sum up, we finally arrive at

(6.18) Mφ(k,α,j)=Mφ(k,α,j+1)+𝒪(),\displaystyle M_{\varphi}^{(k,\alpha,j)}=M_{\varphi}^{(k,\alpha,j+1)}+\mathcal{O}(\hbar),

with |𝒪()|(t,xφLt1Lx+φLt1Lx)|\mathcal{O}(\hbar)|\lesssim\hbar\left(\|\nabla_{t,x}\varphi\|_{L_{t}^{1}L_{x}^{\infty}}+\|\varphi\|_{L_{t}^{1}L_{x}^{\infty}}\right).

When k=2l+1k=2l+1, α=2n+1\alpha=2n+1, ln1l\geq n\geq 1, iteratively using (6.18) and integration by parts, we have

Mφ(k,α,0)=\displaystyle M_{\varphi}^{(k,\alpha,0)}= Mφ(2l+1,2n+1,ln)+𝒪()\displaystyle M_{\varphi}^{(2l+1,2n+1,l-n)}+\mathcal{O}(\hbar)
=\displaystyle= α1ΩTφ(x2n1ξlnf,ε𝑑ξ)(ξlnf,ε𝑑ξ)𝑑x𝑑t+𝒪()\displaystyle\hbar^{\alpha-1}\int_{\Omega_{T}}\varphi\left(\partial_{x}^{2n-1}\int_{\mathbb{R}}\xi^{l-n}f_{\hbar,\varepsilon}d\xi\right)\left(\int_{\mathbb{R}}\xi^{l-n}f_{\hbar,\varepsilon}d\xi\right)dxdt+\mathcal{O}(\hbar)
=\displaystyle= (1)α1ΩTφ(x2n2ξlnf,ε𝑑ξ)(xξlnf,ε𝑑ξ)𝑑x𝑑t+𝒪()\displaystyle(-1)\hbar^{\alpha-1}\int_{\Omega_{T}}\varphi\left(\partial_{x}^{2n-2}\int_{\mathbb{R}}\xi^{l-n}f_{\hbar,\varepsilon}d\xi\right)\left(\partial_{x}\int_{\mathbb{R}}\xi^{l-n}f_{\hbar,\varepsilon}d\xi\right)dxdt+\mathcal{O}(\hbar)
=\displaystyle= \displaystyle...
=\displaystyle= (1)n1α1ΩTφ(xnξlnf,ε𝑑ξ)(xn1ξlnf,ε𝑑ξ)𝑑x𝑑t+𝒪().\displaystyle(-1)^{n-1}\hbar^{\alpha-1}\int_{\Omega_{T}}\varphi\left(\partial_{x}^{n}\int_{\mathbb{R}}\xi^{l-n}f_{\hbar,\varepsilon}d\xi\right)\left(\partial_{x}^{n-1}\int_{\mathbb{R}}\xi^{l-n}f_{\hbar,\varepsilon}d\xi\right)dxdt+\mathcal{O}(\hbar).

Noticing that ξlnf,ε𝑑ξ\int_{\mathbb{R}}\xi^{l-n}f_{\hbar,\varepsilon}d\xi is real-valued, we hence have

(6.19) Mφ(k,α,0)=\displaystyle M_{\varphi}^{(k,\alpha,0)}= (1)n1α12ΩTφx(xn1ξlnf,ε𝑑ξ)2dxdt+𝒪()\displaystyle(-1)^{n-1}\frac{\hbar^{\alpha-1}}{2}\int_{\Omega_{T}}\varphi\partial_{x}\left(\partial_{x}^{n-1}\int_{\mathbb{R}}\xi^{l-n}f_{\hbar,\varepsilon}d\xi\right)^{2}dxdt+\mathcal{O}(\hbar)
=\displaystyle= (1)n1α12ΩTxφ(xn1ξlnf,ε𝑑ξ)2dxdt+𝒪()\displaystyle(-1)^{n-1}\frac{\hbar^{\alpha-1}}{2}\int_{\Omega_{T}}\partial_{x}\varphi\left(\partial_{x}^{n-1}\int_{\mathbb{R}}\xi^{l-n}f_{\hbar,\varepsilon}d\xi\right)^{2}dxdt+\mathcal{O}(\hbar)
\displaystyle\leq xφLt1Lxnxn1ξlnf,ε𝑑ξLtLx12+𝒪()\displaystyle\|\partial_{x}\varphi\|_{L_{t}^{1}L_{x}^{\infty}}\Big{\|}\hbar^{n}\partial_{x}^{n-1}\int_{\mathbb{R}}\xi^{l-n}f_{\hbar,\varepsilon}d\xi\Big{\|}_{L_{t}^{\infty}L_{x}^{1}}^{2}+\mathcal{O}(\hbar)
\displaystyle\leq 𝒪(),\displaystyle\mathcal{O}(\hbar),

where in the last inequality we have used the uniform estimate (3.10). Putting together the estimates (6.13), (6.15), and (6.19), we thus prove (6.8)–(6.9) for the case k=2l+1k=2l+1.

When k=2lk=2l, α=2n+1\alpha=2n+1, ln1l\geq n\geq 1, repeating the proof of the case k=2l+1k=2l+1, we also have

Mφ(k,α,0)\displaystyle M_{\varphi}^{(k,\alpha,0)}
=\displaystyle= Mφ(2l,2n+1,ln)+𝒪()\displaystyle M_{\varphi}^{(2l,2n+1,l-n)}+\mathcal{O}(\hbar)
=\displaystyle= α1ΩTφ(x2n1ξlnf,ε𝑑ξ)(ξln1f,ε𝑑ξ)𝑑x𝑑t+𝒪()\displaystyle\hbar^{\alpha-1}\int_{\Omega_{T}}\varphi\left(\partial_{x}^{2n-1}\int_{\mathbb{R}}\xi^{l-n}f_{\hbar,\varepsilon}d\xi\right)\left(\int_{\mathbb{R}}\xi^{l-n-1}f_{\hbar,\varepsilon}d\xi\right)dxdt+\mathcal{O}(\hbar)
=\displaystyle= (1)α1ΩTφ(x2n2ξlnf,ε𝑑ξ)(xξln1f,ε𝑑ξ)𝑑x𝑑t+𝒪()\displaystyle(-1)\hbar^{\alpha-1}\int_{\Omega_{T}}\varphi\left(\partial_{x}^{2n-2}\int_{\mathbb{R}}\xi^{l-n}f_{\hbar,\varepsilon}d\xi\right)\left(\partial_{x}\int_{\mathbb{R}}\xi^{l-n-1}f_{\hbar,\varepsilon}d\xi\right)dxdt+\mathcal{O}(\hbar)
=\displaystyle= \displaystyle...
=\displaystyle= (1)n1α1ΩTφ(xnξlnf,ε𝑑ξ)(xn1ξln1f,ε𝑑ξ)𝑑x𝑑t+𝒪().\displaystyle(-1)^{n-1}\hbar^{\alpha-1}\int_{\Omega_{T}}\varphi\left(\partial_{x}^{n}\int_{\mathbb{R}}\xi^{l-n}f_{\hbar,\varepsilon}d\xi\right)\left(\partial_{x}^{n-1}\int_{\mathbb{R}}\xi^{l-n-1}f_{\hbar,\varepsilon}d\xi\right)dxdt+\mathcal{O}(\hbar).

By the moment equation (6.3), in a similar way in which we obtain (6.17), we have

Mφ(k,α,0)=\displaystyle M_{\varphi}^{(k,\alpha,0)}= (1)nα12ΩTφt(xn1ξln1f,ε𝑑ξ)2dxdt+𝒪()\displaystyle(-1)^{n}\frac{\hbar^{\alpha-1}}{2}\int_{\Omega_{T}}\varphi\partial_{t}\left(\partial_{x}^{n-1}\int_{\mathbb{R}}\xi^{l-n-1}f_{\hbar,\varepsilon}d\xi\right)^{2}dxdt+\mathcal{O}(\hbar)
=\displaystyle= (1)n1α12ΩTtφ(xn1ξln1f,ε𝑑ξ)2dxdt+𝒪()\displaystyle(-1)^{n-1}\frac{\hbar^{\alpha-1}}{2}\int_{\Omega_{T}}\partial_{t}\varphi\left(\partial_{x}^{n-1}\int_{\mathbb{R}}\xi^{l-n-1}f_{\hbar,\varepsilon}d\xi\right)^{2}dxdt+\mathcal{O}(\hbar)
\displaystyle\leq tφLt1Lxnxn1ξln1f,ε𝑑ξLtLx12+𝒪()\displaystyle\|\partial_{t}\varphi\|_{L_{t}^{1}L_{x}^{\infty}}\Big{\|}\hbar^{n}\partial_{x}^{n-1}\int_{\mathbb{R}}\xi^{l-n-1}f_{\hbar,\varepsilon}d\xi\Big{\|}_{L_{t}^{\infty}L_{x}^{1}}^{2}+\mathcal{O}(\hbar)
\displaystyle\leq 𝒪().\displaystyle\mathcal{O}(\hbar).

Therefore, putting together the estimates (6.13), (6.15), and (6.19), we have completed the proof of (6.8)–(6.9) for the case k=2lk=2l.

Doing the summation over α\alpha in (6.8), we immediately obtain (6.10). For (6.11), we apply an approximation argument and rewrite

ΩTφR,ε(k)𝑑x𝑑t=B1+B2+B3,\displaystyle\int_{\Omega_{T}}\varphi\mathrm{R}_{\hbar,\varepsilon}^{(k)}dxdt=B_{1}+B_{2}+B_{3},

where

B1=\displaystyle B_{1}= ΩTφ(1χ(xR))R,ε(k)𝑑x𝑑t,\displaystyle\int_{\Omega_{T}}\varphi\left(1-\chi(\frac{x}{R})\right)\mathrm{R}_{\hbar,\varepsilon}^{(k)}dxdt,
B2=\displaystyle B_{2}= ΩT(φχ(xR)φn,R)R,ε(k)𝑑x𝑑t,\displaystyle\int_{\Omega_{T}}\left(\varphi\chi(\frac{x}{R})-\varphi_{n,R}\right)\mathrm{R}_{\hbar,\varepsilon}^{(k)}dxdt,
B3=\displaystyle B_{3}= ΩTφn,RR,ε(k)𝑑x𝑑t,\displaystyle\int_{\Omega_{T}}\varphi_{n,R}\mathrm{R}_{\hbar,\varepsilon}^{(k)}dxdt,

and

φn,RCc(ΩT),limnφχ(xR)φn,RLt1([0,T];Lx)=0.\displaystyle\varphi_{n,R}\in C_{c}^{\infty}(\Omega_{T}),\quad\lim_{n\to\infty}\big{\|}\varphi\chi(\frac{x}{R})-\varphi_{n,R}\big{\|}_{L_{t}^{1}([0,T];L_{x}^{\infty})}=0.

By the weighted uniform estimate (6.7) on the remainder term, we have

|B1|\displaystyle|B_{1}|\leq 1RφLt1([0,T];Lx)xR,ε(k)Lt([0,T];Lx1)1R,\displaystyle\frac{1}{R}\|\varphi\|_{L_{t}^{1}([0,T];L_{x}^{\infty})}\|\langle x\rangle\mathrm{R}_{\hbar,\varepsilon}^{(k)}\|_{L_{t}^{\infty}([0,T];L_{x}^{1})}\lesssim\frac{1}{R},
|B2|\displaystyle|B_{2}|\leq φχ(xR)φn,RLt1([0,T];Lx)R,ε(k)Lt([0,T];Lx1)φχ(xR)φn,RLt1([0,T];Lx).\displaystyle\big{\|}\varphi\chi(\frac{x}{R})-\varphi_{n,R}\big{\|}_{L_{t}^{1}([0,T];L_{x}^{\infty})}\|\mathrm{R}_{\hbar,\varepsilon}^{(k)}\|_{L_{t}^{\infty}([0,T];L_{x}^{1})}\lesssim\big{\|}\varphi\chi(\frac{x}{R})-\varphi_{n,R}\big{\|}_{L_{t}^{1}([0,T];L_{x}^{\infty})}.

Hence, using (6.10) for φn,RCc(ΩT)\varphi_{n,R}\in C_{c}^{\infty}(\Omega_{T}), we obtain

lim(,ε)(0,0)|ΩTφR,ε(k)𝑑x𝑑t|\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\Big{|}\int_{\Omega_{T}}\varphi\mathrm{R}_{\hbar,\varepsilon}^{(k)}dxdt\Big{|}
=\displaystyle= limRlimnlim(,ε)(0,0)(B1+B2+B3)\displaystyle\lim_{R\to\infty}\lim_{n\to\infty}\lim_{(\hbar,\varepsilon)\to(0,0)}\left(B_{1}+B_{2}+B_{3}\right)
\displaystyle\lesssim limRlimn(1R+φχ(xR)φn,RLt1([0,T];Lx))=0,\displaystyle\lim_{R\to\infty}\lim_{n\to\infty}\left(\frac{1}{R}+\big{\|}\varphi\chi(\frac{x}{R})-\varphi_{n,R}\big{\|}_{L_{t}^{1}([0,T];L_{x}^{\infty})}\right)=0,

which completes the proof of (6.11).

6.2. Convergence of the Nonlinear Term for k=1,2k=1,2

As a preliminary part, we prove the convergence of the nonlinear term for the k=1,2k=1,2 case based on the weighted uniform estimates in Section 3. We first provide estimates on E,ε(t,x)E_{\hbar,\varepsilon}(t,x) and study its limit.

Lemma 6.4.

There holds that

(6.20) xE,ε=\displaystyle\partial_{x}E_{\hbar,\varepsilon}= ρ,εx2Uερ,ε,\displaystyle\rho_{\hbar,\varepsilon}-\partial_{x}^{2}U_{\varepsilon}*\rho_{\hbar,\varepsilon},
(6.21) tE,ε=\displaystyle\partial_{t}E_{\hbar,\varepsilon}= ξf,ε𝑑ξ+x2Uεξf,ε𝑑ξ.\displaystyle-\int_{\mathbb{R}}\xi f_{\hbar,\varepsilon}d\xi+\partial_{x}^{2}U_{\varepsilon}*\int_{\mathbb{R}}\xi f_{\hbar,\varepsilon}d\xi.

where Uε(x)=12|x|Vε(x)U_{\varepsilon}(x)=\frac{1}{2}|x|-V_{\varepsilon}(x). We have the uniform estimates on E,ε(t,x)E_{\hbar,\varepsilon}(t,x) that

(6.22) E,εLt([0,T];Lx)\displaystyle\|E_{\hbar,\varepsilon}\|_{L_{t}^{\infty}([0,T];L_{x}^{\infty})}\leq C(T),\displaystyle C(T),
(6.23) xE,εLt([0,T];Lx1)\displaystyle\|\partial_{x}E_{\hbar,\varepsilon}\|_{L_{t}^{\infty}([0,T];L_{x}^{1})}\leq C(T),\displaystyle C(T),
(6.24) tE,εLt([0,T];Lx1)\displaystyle\|\partial_{t}E_{\hbar,\varepsilon}\|_{L_{t}^{\infty}([0,T];L_{x}^{1})}\leq C(T).\displaystyle C(T).

Moreover, for p[1,)p\in[1,\infty) we have the strong convergence that

(6.25) E,ε(t,x)E(t,x):=12xy|xy|(f(t,y,ξ)𝑑ξ)𝑑y,Llocp(ΩT),\displaystyle E_{\hbar,\varepsilon}(t,x)\to E(t,x):=\frac{1}{2}\int_{\mathbb{R}}\frac{x-y}{|x-y|}\left(\int_{\mathbb{R}}f(t,y,\xi)d\xi\right)dy,\quad L_{loc}^{p}(\Omega_{T}),

where

(6.26) E(t,x)BVL(ΩT).\displaystyle E(t,x)\in BV\cap L^{\infty}(\Omega_{T}).

Moreover, the limit function E(t,x)E(t,x) satisfies

(6.27) xE=\displaystyle\partial_{x}E= f𝑑ξ,tE=ξf𝑑ξ,\displaystyle\int_{\mathbb{R}}fd\xi,\quad\partial_{t}E=\int_{\mathbb{R}}\xi fd\xi,

in the sense of measures.

Proof.

Equations (6.20) and (6.21) follow from a direct calculation using the moment equation (6.3). Estimate (6.22) follows from the uniform estimate (3.12). For (6.23), noting that

x2Vε=δ(x)eε|x|2εeε|x|+ε2|x|eε|x|,\displaystyle\partial_{x}^{2}V_{\varepsilon}=\delta(x)e^{-\varepsilon|x|}-2\varepsilon e^{-\varepsilon|x|}+\varepsilon^{2}|x|e^{-\varepsilon|x|},

we use Young’s inequality to get

xE,εLt([0,T];Lx1)=\displaystyle\|\partial_{x}E_{\hbar,\varepsilon}\|_{L_{t}^{\infty}([0,T];L_{x}^{1})}= (x2Vε)ρ,εLt([0,T];Lx1)\displaystyle\|(\partial_{x}^{2}V_{\varepsilon})*\rho_{\hbar,\varepsilon}\|_{L_{t}^{\infty}([0,T];L_{x}^{1})}
\displaystyle\leq ρ,εLt([0,T];Lx1)(1+εeε|x|Lx1+ε2|x|eε|x|Lx1)C(T).\displaystyle\|\rho_{\hbar,\varepsilon}\|_{L_{t}^{\infty}([0,T];L_{x}^{1})}\left(1+\|\varepsilon e^{-\varepsilon|x|}\|_{L_{x}^{1}}+\|\varepsilon^{2}|x|e^{-\varepsilon|x|}\|_{L_{x}^{1}}\right)\leq C(T).

For (6.24), by the moment equation (6.3) we rewrite

tE,ε=Vεtρ,ε=x2Vεξf,ε𝑑ξ.\displaystyle\partial_{t}E_{\hbar,\varepsilon}=V_{\varepsilon}*\partial_{t}\rho_{\hbar,\varepsilon}=-\partial_{x}^{2}V_{\varepsilon}*\int_{\mathbb{R}}\xi f_{\hbar,\varepsilon}d\xi.

Via the same way in which we obtain (6.23), we arrive at (6.24).

By the uniform estimates (6.22)–(6.24) and LpL^{p} compactness criteria, there is a subsequence of {E,ε(t,x)}\left\{E_{\hbar,\varepsilon}(t,x)\right\}, which we still denote by {E,ε(t,x)}\left\{E_{\hbar,\varepsilon}(t,x)\right\}, and some function

E(t,x)BVL(ΩT),E(t,x)\in BV\cap L^{\infty}(\Omega_{T}),

such that

E,ε(t,x)Llocp(ΩT)E(t,x),p[1,).\displaystyle E_{\hbar,\varepsilon}(t,x)\stackrel{{\scriptstyle L_{loc}^{p}(\Omega_{T})}}{{\longrightarrow}}E(t,x),\quad p\in[1,\infty).

To obtain the explicit formula of E(t,x)E(t,x), we consider

ΩTE,εφ𝑑x𝑑t=\displaystyle\int_{\Omega_{T}}E_{\hbar,\varepsilon}\varphi dxdt= ΩT12(x|x|ρ,ε)φ𝑑x𝑑t+ΩT(xUερ,ε)φ𝑑x𝑑t\displaystyle\int_{\Omega_{T}}\frac{1}{2}\left(\frac{x}{|x|}*\rho_{\hbar,\varepsilon}\right)\varphi dxdt+\int_{\Omega_{T}}\left(\partial_{x}U_{\varepsilon}*\rho_{\hbar,\varepsilon}\right)\varphi dxdt
=\displaystyle= ΩT12ρ,ε(x|x|φ)𝑑x𝑑t+ΩT(xUερ,ε)φ𝑑x𝑑t.\displaystyle\int_{\Omega_{T}}\frac{1}{2}\rho_{\hbar,\varepsilon}\left(\frac{x}{|x|}*\varphi\right)dxdt+\int_{\Omega_{T}}\left(\partial_{x}U_{\varepsilon}*\rho_{\hbar,\varepsilon}\right)\varphi dxdt.

On the one hand, by the pointwise estimate (6.14) that |xUερ,ε|εx|\partial_{x}U_{\varepsilon}*\rho_{\hbar,\varepsilon}|\lesssim\varepsilon\langle x\rangle and the weighted uniform estimate (3.1), we have

|ΩT(xUερ,ε)φ𝑑x𝑑t|εxρ,εLtLx1xφLt1Lx10.\displaystyle\Big{|}\int_{\Omega_{T}}\left(\partial_{x}U_{\varepsilon}*\rho_{\hbar,\varepsilon}\right)\varphi dxdt\Big{|}\lesssim\varepsilon\|\langle x\rangle\rho_{\hbar,\varepsilon}\|_{L_{t}^{\infty}L_{x}^{1}}\|\langle x\rangle\varphi\|_{L_{t}^{1}L_{x}^{1}}\to 0.

On the other hand, due to the fact that x|x|φLt1([0,T];Cb())\frac{x}{|x|}*\varphi\in L_{t}^{1}([0,T];C_{b}(\mathbb{R})), we use the narrow convergence (4.10) and hence obtain

lim(,ε)(0,0)ΩTE,εφ𝑑x𝑑t=12ΩT(f(t,x,ξ)𝑑ξ)(x|x|φ)𝑑x𝑑t,\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\Omega_{T}}E_{\hbar,\varepsilon}\varphi dxdt=\frac{1}{2}\int_{\Omega_{T}}\left(\int_{\mathbb{R}}f(t,x,\xi)d\xi\right)\left(\frac{x}{|x|}*\varphi\right)dxdt,

which implies formula (6.25). In the same manner, we also attain (6.27) and hence complete the proof.

Now, we are able to prove the following convergence.

Lemma 6.5.

For φLt1([0,T];Cb())\varphi\in L_{t}^{1}([0,T];C_{b}(\mathbb{R})), we have

(6.28) lim(,ε)(0,0)ΩTφE,ε(f,ε𝑑ξ)𝑑x𝑑t=\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\Omega_{T}}\varphi E_{\hbar,\varepsilon}\left(\int_{\mathbb{R}}f_{\hbar,\varepsilon}d\xi\right)dxdt= ΩTφE¯xEdxdt,\displaystyle\int_{\Omega_{T}}\varphi\overline{E}\partial_{x}Edxdt,
(6.29) lim(,ε)(0,0)ΩTφE,ε(ξf,ε𝑑ξ)𝑑x𝑑t=\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\Omega_{T}}\varphi E_{\hbar,\varepsilon}\left(\int_{\mathbb{R}}\xi f_{\hbar,\varepsilon}d\xi\right)dxdt= ΩTφE¯tEdxdt.\displaystyle\int_{\Omega_{T}}\varphi\overline{E}\partial_{t}Edxdt.
Proof.

It suffices to prove (6.28), as (6.29) follows similarly. First, we prove that (6.28) holds for φCc(ΩT)\varphi\in C_{c}^{\infty}(\Omega_{T}). By (6.20), we rewrite

ΩTφE,ε(f,ε𝑑ξ)𝑑x𝑑t\displaystyle\int_{\Omega_{T}}\varphi E_{\hbar,\varepsilon}\left(\int_{\mathbb{R}}f_{\hbar,\varepsilon}d\xi\right)dxdt
=\displaystyle= ΩTφE,ε(xE,ε)𝑑x𝑑t+ΩTφE,ε(x2Uερ,ε)𝑑x𝑑t\displaystyle\int_{\Omega_{T}}\varphi E_{\hbar,\varepsilon}\left(\partial_{x}E_{\hbar,\varepsilon}\right)dxdt+\int_{\Omega_{T}}\varphi E_{\hbar,\varepsilon}\left(\partial_{x}^{2}U_{\varepsilon}*\rho_{\hbar,\varepsilon}\right)dxdt
:=\displaystyle:= A,ε+B,ε.\displaystyle A_{\hbar,\varepsilon}+B_{\hbar,\varepsilon}.

For the first term A,εA_{\hbar,\varepsilon}, by property (1)(1) of BV functions in Appendix B, we have

lim(,ε)(0,0)A,ε=\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}A_{\hbar,\varepsilon}= lim(,ε)(0,0)12ΩT(xφ)(E,ε)2𝑑x𝑑t\displaystyle-\lim_{(\hbar,\varepsilon)\to(0,0)}\frac{1}{2}\int_{\Omega_{T}}(\partial_{x}\varphi)\left(E_{\hbar,\varepsilon}\right)^{2}dxdt
=\displaystyle= 12ΩT(xφ)(E)2𝑑x𝑑t\displaystyle-\frac{1}{2}\int_{\Omega_{T}}(\partial_{x}\varphi)\left(E\right)^{2}dxdt
=\displaystyle= ΩTφE¯(xE)𝑑x𝑑t.\displaystyle\int_{\Omega_{T}}\varphi\overline{E}\left(\partial_{x}E\right)dxdt.

For the second term B,εB_{\hbar,\varepsilon},

B,ε=\displaystyle B_{\hbar,\varepsilon}= ΩTφE,ε(x2Uερ,ε)𝑑x𝑑t\displaystyle\int_{\Omega_{T}}\varphi E_{\hbar,\varepsilon}\left(\partial_{x}^{2}U_{\varepsilon}*\rho_{\hbar,\varepsilon}\right)dxdt
=\displaystyle= ΩT(xφ)E,ε(xUερ,ε)𝑑x𝑑tΩTφ(xE,ε)(xUερ,ε)𝑑x𝑑t\displaystyle-\int_{\Omega_{T}}\left(\partial_{x}\varphi\right)E_{\hbar,\varepsilon}\left(\partial_{x}U_{\varepsilon}*\rho_{\hbar,\varepsilon}\right)dxdt-\int_{\Omega_{T}}\varphi\left(\partial_{x}E_{\hbar,\varepsilon}\right)\left(\partial_{x}U_{\varepsilon}*\rho_{\hbar,\varepsilon}\right)dxdt
\displaystyle\leq εxxφLx1E,εLx+εxφLxxE,εLx10.\displaystyle\varepsilon\|\langle x\rangle\partial_{x}\varphi\|_{L_{x}^{1}}\|E_{\hbar,\varepsilon}\|_{L_{x}^{\infty}}+\varepsilon\|\langle x\rangle\varphi\|_{L_{x}^{\infty}}\|\partial_{x}E_{\hbar,\varepsilon}\|_{L_{x}^{1}}\to 0.

Hence, we complete the proof of (6.28) for φCc(ΩT)\varphi\in C_{c}^{\infty}(\Omega_{T}). Furthermore, by the uniform bound (6.22) and the weighted uniform estimate (3.11), we get the weighted estimate that

(6.30) xE,ε(f,ε𝑑ξ)LtLx1E,εLtLxxf,ε𝑑ξLtLx1C(T).\displaystyle\Big{\|}\langle x\rangle E_{\hbar,\varepsilon}\left(\int_{\mathbb{R}}f_{\hbar,\varepsilon}d\xi\right)\Big{\|}_{L_{t}^{\infty}L_{x}^{1}}\leq\|E_{\hbar,\varepsilon}\|_{L_{t}^{\infty}L_{x}^{\infty}}\Big{\|}\langle x\rangle\int_{\mathbb{R}}f_{\hbar,\varepsilon}d\xi\Big{\|}_{L_{t}^{\infty}L_{x}^{1}}\leq C(T).

Via the same way in which we obtain (6.11) by an approximation argument, we arrive at (6.28) for φLt1([0,T];Cb())\varphi\in L_{t}^{1}([0,T];C_{b}(\mathbb{R})) and hence complete the proof.

6.3. Convergence of the Nonlinear Term for k3k\geq 3

In this section, we prove the convergence of the nonlinear term for the general k3k\geq 3 case.

Lemma 6.6.

Let T>0T>0 and k3k\geq 3. For φLt1([0,T];Cb())\varphi\in L_{t}^{1}([0,T];C_{b}(\mathbb{R})), there holds that

(6.31) lim(,ε)(0,0)ΩTφE,ε(ξk1f,ε𝑑ξ)𝑑x𝑑t=ΩTφE¯(ξk1f(t,dx,dξ))𝑑t.\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\Omega_{T}}\varphi E_{\hbar,\varepsilon}\left(\int_{\mathbb{R}}\xi^{k-1}f_{\hbar,\varepsilon}d\xi\right)dxdt=\int_{\Omega_{T}}\varphi\overline{E}\left(\int_{\mathbb{R}}\xi^{k-1}f(t,dx,d\xi)\right)dt.

As we have proven the base k=1,2k=1,2 case in Lemma 6.5, we take an induction argument to prove Lemma 6.6, whose proof is postponed to the end of the section.

Induction hypothesis: For lk1l\leq k-1, φLt1([0,T];Cb())\varphi\in L_{t}^{1}([0,T];C_{b}(\mathbb{R})), there holds that

(6.32) lim(,ε)(0,0)ΩTφE,ε(ξl1f,ε𝑑ξ)𝑑x𝑑t=ΩTφE¯(ξl1f𝑑ξ)𝑑x𝑑t.\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\Omega_{T}}\varphi E_{\hbar,\varepsilon}\left(\int_{\mathbb{R}}\xi^{l-1}f_{\hbar,\varepsilon}d\xi\right)dxdt=\int_{\Omega_{T}}\varphi\overline{E}\left(\int_{\mathbb{R}}\xi^{l-1}fd\xi\right)dxdt.

Before getting into the proof, we consider the integral function of the moment function that

(6.33) M,ε(m)(t,x):=xξmf,ε(t,y,ξ)𝑑ξ𝑑y,\displaystyle M_{\hbar,\varepsilon}^{(m)}(t,x):=\int_{-\infty}^{x}\int_{\mathbb{R}}\xi^{m}f_{\hbar,\varepsilon}(t,y,\xi)d\xi dy,

which plays a similar role as E,ε(t,x)E_{\hbar,\varepsilon}(t,x). We set up the uniform estimates for M,ε(m)(t,x)M_{\hbar,\varepsilon}^{(m)}(t,x) and study its limit function, which is important to the convergence of the nonlinear term.

Lemma 6.7.

Let 0mk10\leq m\leq k-1. The function M,ε(m)(t,x)M_{\hbar,\varepsilon}^{(m)}(t,x) satisfies

(6.34) tM,ε(m)+xM,ε(m+1)+mxE,εξm1f,ε𝑑ξ𝑑y+xR,ε(m)𝑑y=0,\displaystyle\partial_{t}M_{\hbar,\varepsilon}^{(m)}+\partial_{x}M_{\hbar,\varepsilon}^{(m+1)}+m\int_{-\infty}^{x}E_{\hbar,\varepsilon}\int_{\mathbb{R}}\xi^{m-1}f_{\hbar,\varepsilon}d\xi dy+\int_{-\infty}^{x}\mathrm{R}_{\hbar,\varepsilon}^{(m)}dy=0,

and enjoys the uniform estimates that

(6.35) M,ε(m)Lt([0,T];Lx)C(T),\displaystyle\|M_{\hbar,\varepsilon}^{(m)}\|_{L_{t}^{\infty}([0,T];L_{x}^{\infty})}\leq C(T),
(6.36) xM,ε(m)Lt([0,T];Lx1)C(m,T),\displaystyle\|\partial_{x}M_{\hbar,\varepsilon}^{(m)}\|_{L_{t}^{\infty}([0,T];L_{x}^{1})}\leq C(m,T),
(6.37) tM,ε(m)Lt([0,T];Lx1)C(m,T).\displaystyle\|\partial_{t}M_{\hbar,\varepsilon}^{(m)}\|_{L_{t}^{\infty}([0,T];L_{x}^{1})}\leq C(m,T).

Moreover, for p[1,)p\in[1,\infty) we have the strong convergence that

(6.38) M,ε(m)(t,x)M(m)(t,x):=xξmf(t,y,ξ)𝑑ξ𝑑y,in Llocp(ΩT).\displaystyle M_{\hbar,\varepsilon}^{(m)}(t,x)\to M^{(m)}(t,x):=\int_{-\infty}^{x}\int_{\mathbb{R}}\xi^{m}f(t,y,\xi)d\xi dy,\quad\text{in $L_{loc}^{p}(\Omega_{T})$}.

Finally, under the induction hypothesis (6.32), the limit function satisfies

(6.39) tM(m)+xM(m+1)+mxE¯ξm1f𝑑ξ𝑑y=0,\displaystyle\partial_{t}M^{(m)}+\partial_{x}M^{(m+1)}+m\int_{-\infty}^{x}\overline{E}\int_{\mathbb{R}}\xi^{m-1}fd\xi dy=0,

in the sense of measures.

Proof.

Equation (6.34) follows from the moment equation (6.3). By the weighted uniform estimates (3.11)–(3.12) and the uniform bound (6.7) on the remainder term, we have (6.35)–(6.37). In the same way in which we obtain (6.25), we get (6.38).

Next, we prove (6.39). For the linear part, we have

lim(,ε)(0,0)ΩT(tM,ε(m)+xM,ε(m+1))φ𝑑x𝑑t=ΩTM(m)tφM(m+1)xφdxdt.\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\Omega_{T}}\left(\partial_{t}M_{\hbar,\varepsilon}^{(m)}+\partial_{x}M_{\hbar,\varepsilon}^{(m+1)}\right)\varphi dxdt=\int_{\Omega_{T}}-M^{(m)}\partial_{t}\varphi-M^{(m+1)}\partial_{x}\varphi dxdt.

For the nonlinear part, we use the induction hypothesis to get

lim(,ε)(0,0)mΩT(xE,εξm1f,ε𝑑ξ𝑑y)φ𝑑x𝑑t\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}m\int_{\Omega_{T}}\left(\int_{-\infty}^{x}E_{\hbar,\varepsilon}\int_{\mathbb{R}}\xi^{m-1}f_{\hbar,\varepsilon}d\xi dy\right)\varphi dxdt
=\displaystyle= lim(,ε)(0,0)mΩT(E,εξm1f,ε𝑑ξ)(yφ𝑑x)𝑑y𝑑t\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}m\int_{\Omega_{T}}\left(E_{\hbar,\varepsilon}\int_{\mathbb{R}}\xi^{m-1}f_{\hbar,\varepsilon}d\xi\right)\left(\int_{y}^{\infty}\varphi dx\right)dydt
=\displaystyle= mΩT(E¯ξm1f𝑑ξ)(yφ𝑑x)𝑑y𝑑t\displaystyle m\int_{\Omega_{T}}\left(\overline{E}\int_{\mathbb{R}}\xi^{m-1}fd\xi\right)\left(\int_{y}^{\infty}\varphi dx\right)dydt
=\displaystyle= ΩT(xE¯ξm1f𝑑ξ𝑑y)φ𝑑x𝑑t.\displaystyle\int_{\Omega_{T}}\left(\int_{-\infty}^{x}\overline{E}\int_{\mathbb{R}}\xi^{m-1}fd\xi dy\right)\varphi dxdt.

For the remainder term, due to the fact that yφ𝑑xLt1([0,T];Cb())\int_{y}^{\infty}\varphi dx\in L_{t}^{1}([0,T];C_{b}(\mathbb{R})), we can use (6.11) in Lemma 6.3 to get

lim(,ε)(0,0)ΩT(xR,ε(m)𝑑y)φ𝑑x𝑑t=\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\Omega_{T}}\left(\int_{-\infty}^{x}\mathrm{R}_{\hbar,\varepsilon}^{(m)}dy\right)\varphi dxdt= lim(,ε)(0,0)ΩTR,ε(m)(yφ𝑑x)𝑑y𝑑t=0.\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\Omega_{T}}\mathrm{R}_{\hbar,\varepsilon}^{(m)}\left(\int_{y}^{\infty}\varphi dx\right)dydt=0.

Hence, by formula (6.34), we complete the proof of (6.39). ∎

The following lemma shows that the limit function M(m)M^{(m)} satisfies an induction equation. This is the key to reduce the order of the weight function ξk\xi^{k} so that one can make use of the induction hypothesis.

Lemma 6.8.

Let 0jk10\leq j\leq k-1, 0mk10\leq m\leq k-1. Under the induction hypothesis (6.32), for φCc(ΩT)\varphi\in C_{c}^{\infty}(\Omega_{T}), we have

(6.40) ΩTφM¯(j)(xM(m+1))𝑑x𝑑t\displaystyle\int_{\Omega_{T}}\varphi\overline{M}^{(j)}\left(\partial_{x}M^{(m+1)}\right)dxdt
=\displaystyle= ΩTφM¯(j+1)(xM(m))𝑑x𝑑t+I1(j,m)+I2(j,m)+I3(j,m)+I4(j,m),\displaystyle\int_{\Omega_{T}}\varphi\overline{M}^{(j+1)}\left(\partial_{x}M^{(m)}\right)dxdt+I_{1}^{(j,m)}+I_{2}^{(j,m)}+I_{3}^{(j,m)}+I_{4}^{(j,m)},

where M¯(j)\overline{M}^{(j)} is the Volpert’s symmetric average defined in (A.1) and

I1(j,m)=\displaystyle I_{1}^{(j,m)}= ΩT(tφ)M(j)M(m)𝑑x𝑑t,\displaystyle\int_{\Omega_{T}}(\partial_{t}\varphi)M^{(j)}M^{(m)}dxdt,
I2(j,m)=\displaystyle I_{2}^{(j,m)}= mΩTφ(xE¯ξj1f𝑑ξ𝑑y)M(m)𝑑x𝑑t,\displaystyle-m\int_{\Omega_{T}}\varphi\left(\int_{-\infty}^{x}\overline{E}\int_{\mathbb{R}}\xi^{j-1}fd\xi dy\right)M^{(m)}dxdt,
I3(j,m)=\displaystyle I_{3}^{(j,m)}= ΩT(xφ)M(j+1)M(m)𝑑x𝑑t,\displaystyle\int_{\Omega_{T}}\left(\partial_{x}\varphi\right)M^{(j+1)}M^{(m)}dxdt,
I4(j,m)=\displaystyle I_{4}^{(j,m)}= jΩTφM(j)(xE¯ξm1f𝑑ξ𝑑y)𝑑x𝑑t.\displaystyle-j\int_{\Omega_{T}}\varphi M^{(j)}\left(\int_{-\infty}^{x}\overline{E}\int_{\mathbb{R}}\xi^{m-1}fd\xi dy\right)dxdt.
Proof.

We consider the test function of the form

(φM(j))ησCc(ΩT),\left(\varphi M^{(j)}\right)*\eta_{\sigma}\in C_{c}^{\infty}(\Omega_{T}),

where φCc(ΩT)\varphi\in C_{c}^{\infty}(\Omega_{T}) and ησ(t,x)=σ2η(t/σ,x/σ)\eta_{\sigma}(t,x)=\sigma^{-2}\eta(t/\sigma,x/\sigma) is a smooth mollifier and approximation of the identity. Putting the test function into the limit equation (6.39), we obtain

ΩT((φM(j))ησ)(xM(m+1))𝑑x𝑑t\displaystyle\int_{\Omega_{T}}\left(\left(\varphi M^{(j)}\right)*\eta_{\sigma}\right)\left(\partial_{x}M^{(m+1)}\right)dxdt
=\displaystyle= ΩT(t(φM(j))ησ)M(m)𝑑x𝑑t\displaystyle\int_{\Omega_{T}}\left(\partial_{t}\left(\varphi M^{(j)}\right)*\eta_{\sigma}\right)M^{(m)}dxdt
mΩT((φM(j))ησ)(xE¯ξm1f𝑑ξ𝑑y)𝑑x𝑑t\displaystyle-m\int_{\Omega_{T}}\left(\left(\varphi M^{(j)}\right)*\eta_{\sigma}\right)\left(\int_{-\infty}^{x}\overline{E}\int_{\mathbb{R}}\xi^{m-1}fd\xi dy\right)dxdt
=\displaystyle= ΩTφ(tM(j))(M(m)ησ)𝑑x𝑑t+ΩT(tφ)M(j)(M(m)ησ)𝑑x𝑑t\displaystyle\int_{\Omega_{T}}\varphi\left(\partial_{t}M^{(j)}\right)(M^{(m)}*\eta_{\sigma})dxdt+\int_{\Omega_{T}}\left(\partial_{t}\varphi\right)M^{(j)}(M^{(m)}*\eta_{\sigma})dxdt
mΩT((φM(j))ησ)(xE¯ξm1f𝑑ξ𝑑y)𝑑x𝑑t\displaystyle-m\int_{\Omega_{T}}\left(\left(\varphi M^{(j)}\right)*\eta_{\sigma}\right)\left(\int_{-\infty}^{x}\overline{E}\int_{\mathbb{R}}\xi^{m-1}fd\xi dy\right)dxdt
:=\displaystyle:= ΩTφ(tM(j))(M(m)ησ)𝑑x𝑑t+I1,σ(j,m)+I2,σ(j,m),\displaystyle\int_{\Omega_{T}}\varphi\left(\partial_{t}M^{(j)}\right)(M^{(m)}*\eta_{\sigma})dxdt+I_{1,\sigma}^{(j,m)}+I_{2,\sigma}^{(j,m)},

where

I1,σ(j,m)=\displaystyle I_{1,\sigma}^{(j,m)}= ΩT(tφ)M(j)(M(m)ησ)𝑑x𝑑t,\displaystyle\int_{\Omega_{T}}\left(\partial_{t}\varphi\right)M^{(j)}(M^{(m)}*\eta_{\sigma})dxdt,
I2,σ(j,m)=\displaystyle I_{2,\sigma}^{(j,m)}= mΩT((φM(j))ησ)(xE¯ξm1f𝑑ξ𝑑y)𝑑x𝑑t.\displaystyle-m\int_{\Omega_{T}}\left(\left(\varphi M^{(j)}\right)*\eta_{\sigma}\right)\left(\int_{-\infty}^{x}\overline{E}\int_{\mathbb{R}}\xi^{m-1}fd\xi dy\right)dxdt.

Using again (6.39) for tM(j)\partial_{t}M^{(j)}, we expand

ΩTφ(tM(j))(M(m)ησ)𝑑x𝑑t\displaystyle\int_{\Omega_{T}}\varphi\left(\partial_{t}M^{(j)}\right)(M^{(m)}*\eta_{\sigma})dxdt
=\displaystyle= ΩTφ(xM(j+1))(M(m)ησ)𝑑x𝑑t\displaystyle-\int_{\Omega_{T}}\varphi\left(\partial_{x}M^{(j+1)}\right)(M^{(m)}*\eta_{\sigma})dxdt
jΩTφ(xE¯ξm1f𝑑ξ𝑑y)(M(m)ησ)𝑑x𝑑t\displaystyle-j\int_{\Omega_{T}}\varphi\left(\int_{-\infty}^{x}\overline{E}\int_{\mathbb{R}}\xi^{m-1}fd\xi dy\right)(M^{(m)}*\eta_{\sigma})dxdt
=\displaystyle= ΩTφM(j+1)(xM(m)ησ)𝑑x𝑑t+ΩT(xφ)M(j+1)(M(m)ησ)𝑑x𝑑t\displaystyle\int_{\Omega_{T}}\varphi M^{(j+1)}(\partial_{x}M^{(m)}*\eta_{\sigma})dxdt+\int_{\Omega_{T}}\left(\partial_{x}\varphi\right)M^{(j+1)}(M^{(m)}*\eta_{\sigma})dxdt
jΩTφ(xE¯ξm1f𝑑ξ𝑑y)(M(m)ησ)𝑑x𝑑t\displaystyle-j\int_{\Omega_{T}}\varphi\left(\int_{-\infty}^{x}\overline{E}\int_{\mathbb{R}}\xi^{m-1}fd\xi dy\right)(M^{(m)}*\eta_{\sigma})dxdt
:=\displaystyle:= ΩT((φM(j+1))ησ)xM(m)dxdt+I3,σ(j,m)+I4,σ(j,m),\displaystyle\int_{\Omega_{T}}\left(\left(\varphi M^{(j+1)}\right)*\eta_{\sigma}\right)\partial_{x}M^{(m)}dxdt+I_{3,\sigma}^{(j,m)}+I_{4,\sigma}^{(j,m)},

where

I3,σ(j,m)=\displaystyle I_{3,\sigma}^{(j,m)}= ΩT(xφ)M(j+1)(M(m)ησ)𝑑x𝑑t,\displaystyle\int_{\Omega_{T}}\left(\partial_{x}\varphi\right)M^{(j+1)}(M^{(m)}*\eta_{\sigma})dxdt,
I4,σ(j,m)=\displaystyle I_{4,\sigma}^{(j,m)}= jΩTφ(xE¯ξm1f𝑑ξ𝑑y)(M(m)ησ)𝑑x𝑑t.\displaystyle-j\int_{\Omega_{T}}\varphi\left(\int_{-\infty}^{x}\overline{E}\int_{\mathbb{R}}\xi^{m-1}fd\xi dy\right)(M^{(m)}*\eta_{\sigma})dxdt.

Therefore, we arrive at

(6.41) ΩT((φM(j))ησ)(xM(m+1))𝑑x𝑑t\displaystyle\int_{\Omega_{T}}\left(\left(\varphi M^{(j)}\right)*\eta_{\sigma}\right)\left(\partial_{x}M^{(m+1)}\right)dxdt
=\displaystyle= ΩT((φM(j+1))ησ)xM(m)dxdt+I1,σ(j,m)+I2,σ(j,m)+I3,σ(j,m)+I4,σ(j,m).\displaystyle\int_{\Omega_{T}}\left(\left(\varphi M^{(j+1)}\right)*\eta_{\sigma}\right)\partial_{x}M^{(m)}dxdt+I_{1,\sigma}^{(j,m)}+I_{2,\sigma}^{(j,m)}+I_{3,\sigma}^{(j,m)}+I_{4,\sigma}^{(j,m)}.

By the dominated convergence theorem, we have

limσ0Ii,σ(j,m)=Ii(j,m),i=1,2,3,4.\displaystyle\lim_{\sigma\to 0}I_{i,\sigma}^{(j,m)}=I_{i}^{(j,m)},\quad i=1,2,3,4.

By the properties (5)-(6) of BV functions at the Appendix B, we have

limσ0ΩT((φM(j))ησ)(xM(m+1))𝑑x𝑑t=\displaystyle\lim_{\sigma\to 0}\int_{\Omega_{T}}\left(\left(\varphi M^{(j)}\right)*\eta_{\sigma}\right)\left(\partial_{x}M^{(m+1)}\right)dxdt= ΩTφM¯(j)(xM(m+1))𝑑x𝑑t,\displaystyle\int_{\Omega_{T}}\varphi\overline{M}^{(j)}\left(\partial_{x}M^{(m+1)}\right)dxdt,
limσ0ΩT((φM(j+1))ησ)(xM(m))𝑑x𝑑t=\displaystyle\lim_{\sigma\to 0}\int_{\Omega_{T}}\left(\left(\varphi M^{(j+1)}\right)*\eta_{\sigma}\right)\left(\partial_{x}M^{(m)}\right)dxdt= ΩTφM¯(j+1)(xM(m))𝑑x𝑑t.\displaystyle\int_{\Omega_{T}}\varphi\overline{M}^{(j+1)}\left(\partial_{x}M^{(m)}\right)dxdt.

Sending σ0\sigma\to 0 in (6.41), we complete the proof of (6.40).

Next, we prove that the function M,ε(m)(t,x)M_{\hbar,\varepsilon}^{(m)}(t,x), which is similar to its limit function M(m)(t,x)M^{(m)}(t,x), also has an induction structure.

Lemma 6.9.

Let 0jk10\leq j\leq k-1, 0mk10\leq m\leq k-1. Under the induction hypothesis (6.32), we have

(6.42) lim(,ε)(0,0)ΩTφM,ε(j)(xM,ε(m+1))𝑑x𝑑t\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\Omega_{T}}\varphi M_{\hbar,\varepsilon}^{(j)}\left(\partial_{x}M_{\hbar,\varepsilon}^{(m+1)}\right)dxdt
=\displaystyle= lim(,ε)(0,0)ΩTφM,ε(j+1)(xM,ε(m))𝑑x𝑑t+I1(j,m)+I2(j,m)+I3(j,m)+I4(j,m).\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\Omega_{T}}\varphi M_{\hbar,\varepsilon}^{(j+1)}\left(\partial_{x}M_{\hbar,\varepsilon}^{(m)}\right)dxdt+I_{1}^{(j,m)}+I_{2}^{(j,m)}+I_{3}^{(j,m)}+I_{4}^{(j,m)}.
Proof.

Using equation (6.34) for xM,ε(m+1)\partial_{x}M_{\hbar,\varepsilon}^{(m+1)}, we get

ΩTφM,ε(j)(xM,ε(m+1))𝑑x𝑑t=\displaystyle\int_{\Omega_{T}}\varphi M_{\hbar,\varepsilon}^{(j)}\left(\partial_{x}M_{\hbar,\varepsilon}^{(m+1)}\right)dxdt= A,ε,1(j,m)+A,ε,2(j,m)+A,ε,3(j,m),\displaystyle A_{\hbar,\varepsilon,1}^{(j,m)}+A_{\hbar,\varepsilon,2}^{(j,m)}+A_{\hbar,\varepsilon,3}^{(j,m)},

where

A,ε,1(j,m)=\displaystyle A_{\hbar,\varepsilon,1}^{(j,m)}= ΩTφM,ε(j)(tM,ε(m)),\displaystyle-\int_{\Omega_{T}}\varphi M_{\hbar,\varepsilon}^{(j)}\left(\partial_{t}M_{\hbar,\varepsilon}^{(m)}\right),
A,ε,2(j,m)=\displaystyle A_{\hbar,\varepsilon,2}^{(j,m)}= mΩTφM,ε(j)(xE,εξm1f,ε𝑑ξ𝑑y)𝑑x𝑑t,\displaystyle-m\int_{\Omega_{T}}\varphi M_{\hbar,\varepsilon}^{(j)}\left(\int_{-\infty}^{x}E_{\hbar,\varepsilon}\int_{\mathbb{R}}\xi^{m-1}f_{\hbar,\varepsilon}d\xi dy\right)dxdt,
A,ε,3(j,m)=\displaystyle A_{\hbar,\varepsilon,3}^{(j,m)}= ΩTφM,ε(j)(xR,ε(m)(y)𝑑y)𝑑x𝑑t.\displaystyle-\int_{\Omega_{T}}\varphi M_{\hbar,\varepsilon}^{(j)}\left(\int_{-\infty}^{x}\mathrm{R}_{\hbar,\varepsilon}^{(m)}(y)dy\right)dxdt.

Using again equation (6.34) for tM,ε(j)\partial_{t}M_{\hbar,\varepsilon}^{(j)}, we expand

A,ε,1(j,m)=\displaystyle A_{\hbar,\varepsilon,1}^{(j,m)}= ΩT(tφ)M,ε(j)M,ε(m)𝑑x𝑑t+ΩTφ(tM,ε(j))M,ε(m)𝑑x𝑑t\displaystyle\int_{\Omega_{T}}(\partial_{t}\varphi)M_{\hbar,\varepsilon}^{(j)}M_{\hbar,\varepsilon}^{(m)}dxdt+\int_{\Omega_{T}}\varphi\left(\partial_{t}M_{\hbar,\varepsilon}^{(j)}\right)M_{\hbar,\varepsilon}^{(m)}dxdt
=\displaystyle= A,ε,10(j,m)+A,ε,11(j,m)+A,ε,12(j,m)+A,ε,13(j,m)+A,ε,14(j,m),\displaystyle A_{\hbar,\varepsilon,10}^{(j,m)}+A_{\hbar,\varepsilon,11}^{(j,m)}+A_{\hbar,\varepsilon,12}^{(j,m)}+A_{\hbar,\varepsilon,13}^{(j,m)}+A_{\hbar,\varepsilon,14}^{(j,m)},

where

A,ε,10(j,m)=\displaystyle A_{\hbar,\varepsilon,10}^{(j,m)}= ΩT(tφ)M,ε(j)M,ε(m)𝑑x𝑑t,\displaystyle\int_{\Omega_{T}}(\partial_{t}\varphi)M_{\hbar,\varepsilon}^{(j)}M_{\hbar,\varepsilon}^{(m)}dxdt,
A,ε,11(j,m)=\displaystyle A_{\hbar,\varepsilon,11}^{(j,m)}= ΩTφM,ε(j+1)(xM,ε(m))𝑑x𝑑t,\displaystyle\int_{\Omega_{T}}\varphi M_{\hbar,\varepsilon}^{(j+1)}\left(\partial_{x}M_{\hbar,\varepsilon}^{(m)}\right)dxdt,
A,ε,12(j,m)=\displaystyle A_{\hbar,\varepsilon,12}^{(j,m)}= ΩT(xφ)M,ε(j+1)M,ε(m)𝑑x𝑑t,\displaystyle\int_{\Omega_{T}}\left(\partial_{x}\varphi\right)M_{\hbar,\varepsilon}^{(j+1)}M_{\hbar,\varepsilon}^{(m)}dxdt,
A,ε,13(j,m)=\displaystyle A_{\hbar,\varepsilon,13}^{(j,m)}= jΩTφ(xE,εξj1f,ε𝑑ξ𝑑y)M,ε(m)𝑑x𝑑t,\displaystyle-j\int_{\Omega_{T}}\varphi\left(\int_{-\infty}^{x}E_{\hbar,\varepsilon}\int_{\mathbb{R}}\xi^{j-1}f_{\hbar,\varepsilon}d\xi dy\right)M_{\hbar,\varepsilon}^{(m)}dxdt,
A,ε,14(j,m)=\displaystyle A_{\hbar,\varepsilon,14}^{(j,m)}= ΩTφ(xR,ε(m)𝑑y)M,ε(m)𝑑x𝑑t.\displaystyle-\int_{\Omega_{T}}\varphi\left(\int_{-\infty}^{x}\mathrm{R}_{\hbar,\varepsilon}^{(m)}dy\right)M_{\hbar,\varepsilon}^{(m)}dxdt.

Therefore, we arrive at

ΩTφM,ε(j)(xM,ε(m+1))𝑑x𝑑t\displaystyle\int_{\Omega_{T}}\varphi M_{\hbar,\varepsilon}^{(j)}\left(\partial_{x}M_{\hbar,\varepsilon}^{(m+1)}\right)dxdt
=\displaystyle= ΩTφM,ε(j+1)(xM,ε(m))𝑑x𝑑t+A,ε,2(j,m)+A,ε,3(j,m)+A,ε,10(j,m)+A,ε,12(j,m)+A,ε,13(j,m)+A,ε,14(j,m).\displaystyle\int_{\Omega_{T}}\varphi M_{\hbar,\varepsilon}^{(j+1)}\left(\partial_{x}M_{\hbar,\varepsilon}^{(m)}\right)dxdt+A_{\hbar,\varepsilon,2}^{(j,m)}+A_{\hbar,\varepsilon,3}^{(j,m)}+A_{\hbar,\varepsilon,10}^{(j,m)}+A_{\hbar,\varepsilon,12}^{(j,m)}+A_{\hbar,\varepsilon,13}^{(j,m)}+A_{\hbar,\varepsilon,14}^{(j,m)}.

We are left to prove that

lim(,ε)(0,0)A,ε,2(j,m)=\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}A_{\hbar,\varepsilon,2}^{(j,m)}= mΩTφM(j)(xE¯ξm1f𝑑ξ𝑑y)𝑑x𝑑t=I2(j,m),\displaystyle-m\int_{\Omega_{T}}\varphi M^{(j)}\left(\int_{-\infty}^{x}\overline{E}\int_{\mathbb{R}}\xi^{m-1}fd\xi dy\right)dxdt=I_{2}^{(j,m)},
lim(,ε)(0,0)A,ε,3(j,m)=\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}A_{\hbar,\varepsilon,3}^{(j,m)}= 0,\displaystyle 0,
lim(,ε)(0,0)A,ε,10(j,m)=\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}A_{\hbar,\varepsilon,10}^{(j,m)}= ΩT(tφ)M(j)M(m)𝑑x𝑑t=I1(j,m),\displaystyle\int_{\Omega_{T}}(\partial_{t}\varphi)M^{(j)}M^{(m)}dxdt=I_{1}^{(j,m)},
lim(,ε)(0,0)A,ε,12(j,m)=\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}A_{\hbar,\varepsilon,12}^{(j,m)}= ΩT(xφ)M(j+1)M(m)𝑑x𝑑t=I3(j,m),\displaystyle\int_{\Omega_{T}}\left(\partial_{x}\varphi\right)M^{(j+1)}M^{(m)}dxdt=I_{3}^{(j,m)},
lim(,ε)(0,0)A,ε,13(j,m)=\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}A_{\hbar,\varepsilon,13}^{(j,m)}= jΩTφM(j)(xE¯ξm1f𝑑ξ𝑑y)𝑑x𝑑t=I4(j,m),\displaystyle-j\int_{\Omega_{T}}\varphi M^{(j)}\left(\int_{-\infty}^{x}\overline{E}\int_{\mathbb{R}}\xi^{m-1}fd\xi dy\right)dxdt=I_{4}^{(j,m)},
lim(,ε)(0,0)A,ε,14(j,m)=\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}A_{\hbar,\varepsilon,14}^{(j,m)}= 0.\displaystyle 0.

It suffices to prove the limits for A,ε,2(j,m)A_{\hbar,\varepsilon,2}^{(j,m)}, A,ε,3(j,m)A_{\hbar,\varepsilon,3}^{(j,m)}, and A,ε,10(j,m)A_{\hbar,\varepsilon,10}^{(j,m)}, as the others can be dealt with in a similar way.

For A,ε,2(j,m)A_{\hbar,\varepsilon,2}^{(j,m)}, we rewrite

A,ε,2(j,m)=\displaystyle A_{\hbar,\varepsilon,2}^{(j,m)}= mΩT(yφM,ε(j)𝑑x)(E,εξm1f,ε𝑑ξ)𝑑y𝑑t.\displaystyle-m\int_{\Omega_{T}}\left(\int_{y}^{\infty}\varphi M_{\hbar,\varepsilon}^{(j)}dx\right)\left(E_{\hbar,\varepsilon}\int_{\mathbb{R}}\xi^{m-1}f_{\hbar,\varepsilon}d\xi\right)dydt.

On the one hand, we have the LxL_{x}^{\infty} convergence that

yφM,ε(j)𝑑xyφM(j)𝑑xLxφLx2M,ε(j)M(j)Lx,loc20.\displaystyle\Big{\|}\int_{y}^{\infty}\varphi M_{\hbar,\varepsilon}^{(j)}dx-\int_{y}^{\infty}\varphi M^{(j)}dx\Big{\|}_{L_{x}^{\infty}}\leq\|\varphi\|_{L_{x}^{2}}\|M_{\hbar,\varepsilon}^{(j)}-M^{(j)}\|_{L_{x,loc}^{2}}\to 0.

On the other hand, by the induction hypothesis for lk1l\leq k-1, we have

lim(,ε)(0,0)ΩTφE,ε(ξm1f,ε𝑑ξ)𝑑x𝑑t=ΩTφE¯(ξm1f𝑑ξ)𝑑x𝑑t,\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\Omega_{T}}\varphi E_{\hbar,\varepsilon}\left(\int_{\mathbb{R}}\xi^{m-1}f_{\hbar,\varepsilon}d\xi\right)dxdt=\int_{\Omega_{T}}\varphi\overline{E}\left(\int_{\mathbb{R}}\xi^{m-1}fd\xi\right)dxdt,

for φLt1([0,T];Cb())\varphi\in L_{t}^{1}([0,T];C_{b}(\mathbb{R})). Therefore, we obtain

lim(,ε)(0,0)A,ε,2(j,m)=\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}A_{\hbar,\varepsilon,2}^{(j,m)}= mΩT(yφM(j)𝑑x)(E¯ξm1f𝑑ξ)𝑑y𝑑t\displaystyle-m\int_{\Omega_{T}}\left(\int_{y}^{\infty}\varphi M^{(j)}dx\right)\left(\overline{E}\int_{\mathbb{R}}\xi^{m-1}fd\xi\right)dydt
=\displaystyle= mΩTφM(j)(xE¯ξm1f𝑑ξ𝑑y)𝑑x𝑑t\displaystyle-m\int_{\Omega_{T}}\varphi M^{(j)}\left(\int_{-\infty}^{x}\overline{E}\int_{\mathbb{R}}\xi^{m-1}fd\xi dy\right)dxdt
=\displaystyle= I2(j,m).\displaystyle I_{2}^{(j,m)}.

For A,ε,3(j,m)A_{\hbar,\varepsilon,3}^{(j,m)}, we rewrite

A,ε,3(j,m)=\displaystyle A_{\hbar,\varepsilon,3}^{(j,m)}= ΩT(y+φM,ε(j)𝑑x)R,ε(m)𝑑y𝑑t\displaystyle-\int_{\Omega_{T}}\left(\int_{y}^{+\infty}\varphi M_{\hbar,\varepsilon}^{(j)}dx\right)\mathrm{R}_{\hbar,\varepsilon}^{(m)}dydt
=\displaystyle= ΩT(1χ(yR)+χ(yR))(y+φM,ε(j)𝑑x)R,ε(m)𝑑y𝑑t.\displaystyle-\int_{\Omega_{T}}\left(1-\chi(\frac{y}{R})+\chi(\frac{y}{R})\right)\left(\int_{y}^{+\infty}\varphi M_{\hbar,\varepsilon}^{(j)}dx\right)\mathrm{R}_{\hbar,\varepsilon}^{(m)}dydt.

By the quantitative estimate (6.10) in Lemma 6.3 and the weighted uniform bound (6.7) on the remainder term, we have

|A,ε,3(j,m)|\displaystyle\big{|}A_{\hbar,\varepsilon,3}^{(j,m)}\big{|}\lesssim t,y(χ(yR)y+φM,ε(j)𝑑x)Lt1Lx+(+ε)χ(yR)y+φM,ε(j)𝑑xLt1Lx\displaystyle\hbar\Big{\|}\nabla_{t,y}\left(\chi(\frac{y}{R})\int_{y}^{+\infty}\varphi M_{\hbar,\varepsilon}^{(j)}dx\right)\Big{\|}_{L_{t}^{1}L_{x}^{\infty}}+(\hbar+\varepsilon)\Big{\|}\chi(\frac{y}{R})\int_{y}^{+\infty}\varphi M_{\hbar,\varepsilon}^{(j)}dx\Big{\|}_{L_{t}^{1}L_{x}^{\infty}}
+1Ry+φM,ε(j)𝑑xLt1LxxR,ε(m)LtLx1\displaystyle+\frac{1}{R}\Big{\|}\int_{y}^{+\infty}\varphi M_{\hbar,\varepsilon}^{(j)}dx\Big{\|}_{L_{t}^{1}L_{x}^{\infty}}\|\langle x\rangle\mathrm{R}_{\hbar,\varepsilon}^{(m)}\|_{L_{t}^{\infty}L_{x}^{1}}
\displaystyle\leq (φLt1Lx+tφLt1Lx1)(M,ε(j)LtLx+tM,ε(j)LtLx1)\displaystyle\hbar\left(\|\varphi\|_{L_{t}^{1}L_{x}^{\infty}}+\|\partial_{t}\varphi\|_{L_{t}^{1}L_{x}^{1}}\right)\left(\|M_{\hbar,\varepsilon}^{(j)}\|_{L_{t}^{\infty}L_{x}^{\infty}}+\|\partial_{t}M_{\hbar,\varepsilon}^{(j)}\|_{L_{t}^{\infty}L_{x}^{1}}\right)
+(R++ε+1R)φLt1LxM,ε(j)LtLx\displaystyle+\left(\frac{\hbar}{R}+\hbar+\varepsilon+\frac{1}{R}\right)\|\varphi\|_{L_{t}^{1}L_{x}^{\infty}}\|M_{\hbar,\varepsilon}^{(j)}\|_{L_{t}^{\infty}L_{x}^{\infty}}
\displaystyle\lesssim +ε+1R0,\displaystyle\hbar+\varepsilon+\frac{1}{R}\to 0,

where in the last inequality we have used the uniform bounds (6.35)–(6.37) on M,ε(j)M_{\hbar,\varepsilon}^{(j)}.

For A,ε,10(j,m)A_{\hbar,\varepsilon,10}^{(j,m)}, noting that

M,ε(j)Lloc2M(j),\displaystyle M_{\hbar,\varepsilon}^{(j)}\stackrel{{\scriptstyle L_{loc}^{2}}}{{\longrightarrow}}M^{(j)},

we immediately get

lim(,ε)(0,0)A,ε,10(j,m)=\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}A_{\hbar,\varepsilon,10}^{(j,m)}= lim(,ε)(0,0)ΩT(tφ)M,ε(j)M,ε(m)𝑑x𝑑t\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\Omega_{T}}(\partial_{t}\varphi)M_{\hbar,\varepsilon}^{(j)}M_{\hbar,\varepsilon}^{(m)}dxdt
=\displaystyle= ΩT(tφ)M(j)M(m)𝑑x𝑑t=I1(j,m).\displaystyle\int_{\Omega_{T}}(\partial_{t}\varphi)M^{(j)}M^{(m)}dxdt=I_{1}^{(j,m)}.

Hence, we complete the proof of Lemma 6.9.

Now, we are able to prove the moment convergence of the nonlinear term, which is Lemma 6.6, the last remaining part of the proof of Lemma 6.1.

Proof of Lemma 6.6.

By the uniform bound (6.22) and the weighted uniform estimate (3.11), we have

xE,ε(ξk1f,ε𝑑ξ)dxdtLtLx1)E,εLtLx1xξk1f,ε𝑑ξLtLx1C(k).\displaystyle\Big{\|}\langle x\rangle E_{\hbar,\varepsilon}\left(\int_{\mathbb{R}}\xi^{k-1}f_{\hbar,\varepsilon}d\xi\right)dxdt\Big{\|}_{L_{t}^{\infty}L_{x}^{1})}\leq\|E_{\hbar,\varepsilon}\|_{L_{t}^{\infty}L_{x}^{1}}\Big{\|}\langle x\rangle\int_{\mathbb{R}}\xi^{k-1}f_{\hbar,\varepsilon}d\xi\Big{\|}_{L_{t}^{\infty}L_{x}^{1}}\leq C(k).

Following the same process as the k=1,2k=1,2 case in Lemma 6.5, it suffices to prove

(6.43) lim(,ε)(0,0)ΩTφE,ε(ξk1f,ε𝑑ξ)𝑑x𝑑t=ΩTφE¯(ξk1f(t,dx,dξ))𝑑t,\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\Omega_{T}}\varphi E_{\hbar,\varepsilon}\left(\int_{\mathbb{R}}\xi^{k-1}f_{\hbar,\varepsilon}d\xi\right)dxdt=\int_{\Omega_{T}}\varphi\overline{E}\left(\int_{\mathbb{R}}\xi^{k-1}f(t,dx,d\xi)\right)dt,

for φCc(ΩT)\varphi\in C_{c}^{\infty}(\Omega_{T}).

First, we get into the analysis of the term on the left hand side of (6.43). Noting that

E,ε=\displaystyle E_{\hbar,\varepsilon}= xVερ,ε=x(|x|2+Uε(x))ρ,ε,\displaystyle\partial_{x}V_{\varepsilon}*\rho_{\hbar,\varepsilon}=\partial_{x}\left(\frac{|x|}{2}+U_{\varepsilon}(x)\right)*\rho_{\hbar,\varepsilon},
x(|x|2ρ,ε)=\displaystyle\partial_{x}\left(\frac{|x|}{2}*\rho_{\hbar,\varepsilon}\right)= (x2|x|ρ,ε)=xρ,ε(y)𝑑y12=M,ε(0)12,\displaystyle\left(\frac{x}{2|x|}*\rho_{\hbar,\varepsilon}\right)=\int_{-\infty}^{x}\rho_{\hbar,\varepsilon}(y)dy-\frac{1}{2}=M_{\hbar,\varepsilon}^{(0)}-\frac{1}{2},

we rewrite

ΩTφE,ε(ξk1f,ε𝑑ξ)𝑑x𝑑t=A1+A2+A3,\displaystyle\int_{\Omega_{T}}\varphi E_{\hbar,\varepsilon}\left(\int_{\mathbb{R}}\xi^{k-1}f_{\hbar,\varepsilon}d\xi\right)dxdt=A_{1}+A_{2}+A_{3},

where

A1=\displaystyle A_{1}= ΩTφM,ε(0)(xM,ε(k1))𝑑x𝑑t,\displaystyle\int_{\Omega_{T}}\varphi M_{\hbar,\varepsilon}^{(0)}\left(\partial_{x}M_{\hbar,\varepsilon}^{(k-1)}\right)dxdt,
A2=\displaystyle A_{2}= 12ΩTφ(ξk1f,ε𝑑ξ)𝑑x𝑑t,\displaystyle-\frac{1}{2}\int_{\Omega_{T}}\varphi\left(\int_{\mathbb{R}}\xi^{k-1}f_{\hbar,\varepsilon}d\xi\right)dxdt,
A3=\displaystyle A_{3}= ΩTφ(xUερ,ε)(ξk1f,ε𝑑ξ)𝑑x𝑑t.\displaystyle\int_{\Omega_{T}}\varphi\left(\partial_{x}U_{\varepsilon}*\rho_{\hbar,\varepsilon}\right)\left(\int_{\mathbb{R}}\xi^{k-1}f_{\hbar,\varepsilon}d\xi\right)dxdt.

For term A3A_{3}, using pointwise estimate (6.14) that |xUερ,ε|εx|\partial_{x}U_{\varepsilon}*\rho_{\hbar,\varepsilon}|\lesssim\varepsilon\langle x\rangle and the weighted estimate (3.11), we get

|A3|=\displaystyle|A_{3}|= |ΩTφ(xUερ,ε)(ξk1f,ε𝑑ξ)𝑑x𝑑t|\displaystyle\Big{|}\int_{\Omega_{T}}\varphi\left(\partial_{x}U_{\varepsilon}*\rho_{\hbar,\varepsilon}\right)\left(\int_{\mathbb{R}}\xi^{k-1}f_{\hbar,\varepsilon}d\xi\right)dxdt\Big{|}
\displaystyle\leq εφLt1Lxxξk1f,ε𝑑ξLtLx10.\displaystyle\varepsilon\|\varphi\|_{L_{t}^{1}L_{x}^{\infty}}\Big{\|}\langle x\rangle\int_{\mathbb{R}}\xi^{k-1}f_{\hbar,\varepsilon}d\xi\Big{\|}_{L_{t}^{\infty}L_{x}^{1}}\to 0.

Therefore, we obtain

(6.44) lim(,ε)(0,0)ΩTφE,ε(ξk1f,ε𝑑ξ)𝑑x𝑑t\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\Omega_{T}}\varphi E_{\hbar,\varepsilon}\left(\int_{\mathbb{R}}\xi^{k-1}f_{\hbar,\varepsilon}d\xi\right)dxdt
=\displaystyle= lim(,ε)(0,0)ΩTφM,ε(0)(xM,ε(k1))𝑑x𝑑t12ΩTφ(ξk1f𝑑ξ)𝑑x𝑑t.\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\Omega_{T}}\varphi M_{\hbar,\varepsilon}^{(0)}\left(\partial_{x}M_{\hbar,\varepsilon}^{(k-1)}\right)dxdt-\frac{1}{2}\int_{\Omega_{T}}\varphi\left(\int_{\mathbb{R}}\xi^{k-1}fd\xi\right)dxdt.

On the other hand, by (6.25) and conservation of mass (5.4) in Lemma 5.1, we have

E(t,x)=\displaystyle E(t,x)= xy2|xy|(f(t,y,ξ)𝑑ξ)𝑑y\displaystyle\int_{\mathbb{R}}\frac{x-y}{2|x-y|}\left(\int_{\mathbb{R}}f(t,y,\xi)d\xi\right)dy
=\displaystyle= xf(t,y,ξ)𝑑ξ𝑑y12=M(0)(t,x)12,\displaystyle\int_{-\infty}^{x}\int_{\mathbb{R}}f(t,y,\xi)d\xi dy-\frac{1}{2}=M^{(0)}(t,x)-\frac{1}{2},

and hence obtain

(6.45) ΩTφE¯(ξk1f(t,dx,dξ))𝑑t\displaystyle\int_{\Omega_{T}}\varphi\overline{E}\left(\int_{\mathbb{R}}\xi^{k-1}f(t,dx,d\xi)\right)dt
=\displaystyle= ΩTφM¯(0)(xM(k1))𝑑x𝑑t12ΩTφ(ξk1f𝑑ξ)𝑑x𝑑t.\displaystyle\int_{\Omega_{T}}\varphi\overline{M}^{(0)}\left(\partial_{x}M^{(k-1)}\right)dxdt-\frac{1}{2}\int_{\Omega_{T}}\varphi\left(\int_{\mathbb{R}}\xi^{k-1}fd\xi\right)dxdt.

Comparing (6.44) with (6.45), to conclude (6.43), we are left to prove

(6.46) lim(,ε)(0,0)ΩTφM,ε(0)(xM,ε(k1))𝑑x𝑑t=ΩTφM¯(0)(xM(k1))𝑑x𝑑t.\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\Omega_{T}}\varphi M_{\hbar,\varepsilon}^{(0)}\left(\partial_{x}M_{\hbar,\varepsilon}^{(k-1)}\right)dxdt=\int_{\Omega_{T}}\varphi\overline{M}^{(0)}\left(\partial_{x}M^{(k-1)}\right)dxdt.

By Lemma 6.8 and Lemma 6.9, the equality (6.46) is equivalent to

(6.47) lim(,ε)(0,0)ΩTφM,ε(1)(xM,ε(k2))𝑑x𝑑t=ΩTφM¯(1)(xM(k2))𝑑x𝑑t.\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\Omega_{T}}\varphi M_{\hbar,\varepsilon}^{(1)}\left(\partial_{x}M_{\hbar,\varepsilon}^{(k-2)}\right)dxdt=\int_{\Omega_{T}}\varphi\overline{M}^{(1)}\left(\partial_{x}M^{(k-2)}\right)dxdt.

Iteratively using Lemma 6.8 and Lemma 6.9, we are left to prove

(6.48) lim(,ε)(0,0)ΩTφM,ε(n)(xM,ε(n))𝑑x𝑑t=\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\Omega_{T}}\varphi M_{\hbar,\varepsilon}^{(n)}\left(\partial_{x}M_{\hbar,\varepsilon}^{(n)}\right)dxdt= ΩTφM¯(n)(xM(n))𝑑x𝑑t,k=2n+1,\displaystyle\int_{\Omega_{T}}\varphi\overline{M}^{(n)}\left(\partial_{x}M^{(n)}\right)dxdt,\quad k=2n+1,
(6.49) lim(,ε)(0,0)ΩTφM,ε(n)(xM,ε(n+1))𝑑x𝑑t=\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\Omega_{T}}\varphi M_{\hbar,\varepsilon}^{(n)}\left(\partial_{x}M_{\hbar,\varepsilon}^{(n+1)}\right)dxdt= ΩTφM¯(n)(xM(n+1))𝑑x𝑑t,k=2n.\displaystyle\int_{\Omega_{T}}\varphi\overline{M}^{(n)}\left(\partial_{x}M^{(n+1)}\right)dxdt,\quad k=2n.

For (6.48), by integration by parts we get

lim(,ε)(0,0)ΩTφM,ε(n)(xM,ε(n))𝑑x𝑑t=\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\Omega_{T}}\varphi M_{\hbar,\varepsilon}^{(n)}\left(\partial_{x}M_{\hbar,\varepsilon}^{(n)}\right)dxdt= 12lim(,ε)(0,0)ΩT(xφ)(M,ε(n))2𝑑x𝑑t\displaystyle-\frac{1}{2}\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\Omega_{T}}\left(\partial_{x}\varphi\right)\left(M_{\hbar,\varepsilon}^{(n)}\right)^{2}dxdt
=\displaystyle= 12ΩT(xφ)(M(n))2𝑑x𝑑t\displaystyle-\frac{1}{2}\int_{\Omega_{T}}\left(\partial_{x}\varphi\right)\left(M^{(n)}\right)^{2}dxdt
=\displaystyle= ΩTφM¯(n)(xM(n))𝑑x𝑑t,\displaystyle\int_{\Omega_{T}}\varphi\overline{M}^{(n)}\left(\partial_{x}M^{(n)}\right)dxdt,

where in the last equality we have used the fact that x(M(n))2=2M¯(n)xM(n)\partial_{x}(M^{(n)})^{2}=2\overline{M}^{(n)}\partial_{x}M^{(n)}.

For (6.49), by the equations (6.34) and (6.39), it suffices to prove

lim(,ε)(0,0)ΩTφM,ε(n)(tM,ε(n))𝑑x𝑑t=\displaystyle\lim_{(\hbar,\varepsilon)\to(0,0)}\int_{\Omega_{T}}\varphi M_{\hbar,\varepsilon}^{(n)}\left(\partial_{t}M_{\hbar,\varepsilon}^{(n)}\right)dxdt= ΩTφM¯(n)(tM(n))𝑑x𝑑t.\displaystyle\int_{\Omega_{T}}\varphi\overline{M}^{(n)}\left(\partial_{t}M^{(n)}\right)dxdt.

This can be done in the same way in which we obtain (6.48). Hence, we complete the proof of Lemma 6.6. ∎

7. Full Convergence to the Vlasov-Poisson Equation

In the section, we prove the limit measure f(t,dx,dξ)f(t,dx,d\xi) satisfies the Vlasov-Poisson equation in the weak sense. Let

(7.1) μ:=tf+ξxfξ(E¯f).\displaystyle\mu:=\partial_{t}f+\xi\partial_{x}f-\partial_{\xi}(\overline{E}f).

We use the following lemma to conclude the full convergence to the Vlasov-Poisson equation, that is, μ(t,x,ξ)=0\mu(t,x,\xi)=0 in the sense of distributions.

Lemma 7.1 ([50, p.620p.620]).

Let ΩT=(0,T)×\Omega_{T}=(0,T)\times\mathbb{R}, δ\delta be an arbitrary positive constant. Assume f(t,dx,dξ)f(t,dx,d\xi) satisfies the following conditions.

  1. (1)(1)

    Exponential decay:

    (7.2) ΩTeδ|ξ|f(t,dx,dξ)𝑑tCδ.\displaystyle\iint_{\Omega_{T}}\int_{\mathbb{R}}e^{\delta|\xi|}f(t,dx,d\xi)dt\leq C_{\delta}.
  2. (2)(2)

    For all test functions of the form ϕ(t,x,ξ)=φ(t,x)ξm\phi(t,x,\xi)=\varphi(t,x)\xi^{m}, φ(t,x)Cc(ΩT)\varphi(t,x)\in C_{c}^{\infty}(\Omega_{T}), there holds that

    (7.3) ΩTϕ𝑑μ(t,x,ξ)=0.\displaystyle\iint_{\Omega_{T}}\int_{\mathbb{R}}\phi d\mu(t,x,\xi)=0.

Then μ(t,x,ξ)=0\mu(t,x,\xi)=0 in the sense of distributions.

By the moment convergence in Lemma 6.1, we have verified the condition (2)(2) in Lemma 7.1. Therefore, we are left to prove the exponential decay condition.

Lemma 7.2.

Let T>0T>0. There holds that

(7.4) 2ξ2kf(t,dx,dξ)C2k(2k)2ket,t[0,T].\displaystyle\iint_{\mathbb{R}^{2}}\xi^{2k}f(t,dx,d\xi)\leq C^{2k}(2k)^{2k}e^{t},\quad\forall t\in[0,T].

In particular, there exists a positive constant δ\delta such that

(7.5) 2eδ|ξ|f(t,dx,dξ)Cδet,t[0,T].\displaystyle\iint_{\mathbb{R}^{2}}e^{\delta|\xi|}f(t,dx,d\xi)\leq C_{\delta}e^{t},\quad\forall t\in[0,T].
Proof.

Recalling the moment equation (6.3) that

tξmf,ε𝑑ξ+xξm+1f,ε𝑑ξ+mE,εξm1f,ε𝑑ξ+R,ε(m)=0,\displaystyle\partial_{t}\int_{\mathbb{R}}\xi^{m}f_{\hbar,\varepsilon}d\xi+\partial_{x}\int_{\mathbb{R}}\xi^{m+1}f_{\hbar,\varepsilon}d\xi+mE_{\hbar,\varepsilon}\int_{\mathbb{R}}\xi^{m-1}f_{\hbar,\varepsilon}d\xi+\mathrm{R}_{\hbar,\varepsilon}^{(m)}=0,

we have

2ξ2kf,ε(t,x,ξ)𝑑ξ𝑑x\displaystyle\iint_{\mathbb{R}^{2}}\xi^{2k}f_{\hbar,\varepsilon}(t,x,\xi)d\xi dx
=\displaystyle= 2ξ2kf,ε(0,x,ξ)𝑑ξ𝑑x+2kΩtE,ε(ξ2k1f,ε𝑑ξ)𝑑x𝑑τ+ΩtR,ε(2k)𝑑x𝑑τ.\displaystyle\iint_{\mathbb{R}^{2}}\xi^{2k}f_{\hbar,\varepsilon}(0,x,\xi)d\xi dx+2k\int_{\Omega_{t}}E_{\hbar,\varepsilon}\left(\int_{\mathbb{R}}\xi^{2k-1}f_{\hbar,\varepsilon}d\xi\right)dxd\tau+\int_{\Omega_{t}}\mathrm{R}_{\hbar,\varepsilon}^{(2k)}dxd\tau.

By the narrow convergence in Lemma 4.3, Lemma 6.3, and Lemma 6.6, taking φ(τ,x)=1\varphi(\tau,x)=1 and letting (,ε)(0,0)(\hbar,\varepsilon)\to(0,0), we obtain

2ξ2kf(t,dx,dξ)=\displaystyle\iint_{\mathbb{R}^{2}}\xi^{2k}f(t,dx,d\xi)= 2ξ2kf(0,dx,dξ)+2kΩtE¯ξ2k1f(τ,dx,dξ)𝑑τ.\displaystyle\iint_{\mathbb{R}^{2}}\xi^{2k}f(0,dx,d\xi)+2k\int_{\Omega_{t}}\overline{E}\int_{\mathbb{R}}\xi^{2k-1}f(\tau,dx,d\xi)d\tau.

For the initial data, we have

2ξ2kf(0,dx,dξ)=\displaystyle\iint_{\mathbb{R}^{2}}\xi^{2k}f(0,dx,d\xi)= lim02ξ2kf(0,x,ξ)𝑑ξ𝑑x\displaystyle\lim_{\hbar\to 0}\iint_{\mathbb{R}^{2}}\xi^{2k}f_{\hbar}(0,x,\xi)d\xi dx
=\displaystyle= lim02k22kα=02k(2kα)(1)2kαDxαψinDx2kαψin¯𝑑x\displaystyle\lim_{\hbar\to 0}\frac{\hbar^{2k}}{2^{2k}}\sum_{\alpha=0}^{2k}\binom{2k}{\alpha}(-1)^{2k-\alpha}\int_{\mathbb{R}}D_{x}^{\alpha}\psi_{\hbar}^{\mathrm{in}}\overline{D_{x}^{2k-\alpha}\psi_{\hbar}^{\mathrm{in}}}dx
\displaystyle\leq sup122kα=02k(2kα)αxαψinLx22kαx2kαψinLx2\displaystyle\sup_{\hbar}\frac{1}{2^{2k}}\sum_{\alpha=0}^{2k}\binom{2k}{\alpha}\|\hbar^{\alpha}\partial_{x}^{\alpha}\psi_{\hbar}^{\mathrm{in}}\|_{L_{x}^{2}}\|\hbar^{2k-\alpha}\partial_{x}^{2k-\alpha}\psi_{\hbar}^{\mathrm{in}}\|_{L_{x}^{2}}
\displaystyle\leq C2k(2k)2k,\displaystyle C^{2k}(2k)^{2k},

where in the last inequality we have used the initial condition (1.7) that

αxαψinLx2Cααα.\|\hbar^{\alpha}\partial_{x}^{\alpha}\psi_{\hbar}^{\mathrm{in}}\|_{L_{x}^{2}}\leq C^{\alpha}\alpha^{\alpha}.

For the nonlinear part, using that 2k|ξ|2k1(2k)2k+ξ2k2k|\xi|^{2k-1}\leq(2k)^{2k}+\xi^{2k}, we get

2k|ΩtE¯ξ2k1f(τ,dx,dξ)𝑑τ|\displaystyle 2k\Big{|}\int_{\Omega_{t}}\overline{E}\int_{\mathbb{R}}\xi^{2k-1}f(\tau,dx,d\xi)d\tau\Big{|}
\displaystyle\leq E¯Lt,x0t2((2k)2k+ξ2k)f(τ,dx,dξ)𝑑τ\displaystyle\|\overline{E}\|_{L_{t,x}^{\infty}}\int_{0}^{t}\iint_{\mathbb{R}^{2}}((2k)^{2k}+\xi^{2k})f(\tau,dx,d\xi)d\tau
\displaystyle\leq (2k)2kT+0t2ξ2kf(τ,dx,dξ)𝑑τ,\displaystyle(2k)^{2k}T+\int_{0}^{t}\iint_{\mathbb{R}^{2}}\xi^{2k}f(\tau,dx,d\xi)d\tau,

where in the last inequality we have used that E¯Lt,x1\|\overline{E}\|_{L_{t,x}^{\infty}}\leq 1 in (6.26). Thus, we arrive at

2ξ2kf(t,dx,dξ)C2k(2k)2k+T(2k)2k+0t2ξ2kf(τ,dx,dξ)𝑑τ.\displaystyle\iint_{\mathbb{R}^{2}}\xi^{2k}f(t,dx,d\xi)\leq C^{2k}(2k)^{2k}+T(2k)^{2k}+\int_{0}^{t}\iint_{\mathbb{R}^{2}}\xi^{2k}f(\tau,dx,d\xi)d\tau.

Then by Gronwall’s inequality, we get

(7.6) 2ξ2kf(t,dx,dξ)(C2k+T)(2k)2ket.\displaystyle\iint_{\mathbb{R}^{2}}\xi^{2k}f(t,dx,d\xi)\leq(C^{2k}+T)(2k)^{2k}e^{t}.

For the exponential decay (7.5), provided that Cδ<1C\delta<1, we have

2eδ|ξ|f(t,dx,dξ)\displaystyle\iint_{\mathbb{R}^{2}}e^{\delta|\xi|}f(t,dx,d\xi)\leq 2(eδ|ξ|+eδ|ξ|)f(t,dx,dξ)\displaystyle\iint_{\mathbb{R}^{2}}\left(e^{\delta|\xi|}+e^{-\delta|\xi|}\right)f(t,dx,d\xi)
=\displaystyle= 2k=0δ2k2k!2ξ2kf(t,dx,dξ)\displaystyle 2\sum_{k=0}^{\infty}\frac{\delta^{2k}}{2k!}\iint_{\mathbb{R}^{2}}\xi^{2k}f(t,dx,d\xi)
\displaystyle\leq 2etk=0δ2k(C2k+T)(2k)2k2k!<.\displaystyle 2e^{t}\sum_{k=0}^{\infty}\frac{\delta^{2k}(C^{2k}+T)(2k)^{2k}}{2k!}<\infty.

Appendix A Measure Solutions to the Vlasov-Poisson Equation

Let us recall the definition of weak measure solutions from [51] by Zheng and Majda.

Definition A.1.

A pair (E(t,x),f(t,x,ξ))(E(t,x),f(t,x,\xi)) of a function and a bounded non-negative Radon measure is called a weak solution to the Vlasov-Poisson equation (1.4) if for any T>0T>0 there hold

  1. (1)

    E(t,x)(BVL)(ΩT)E(t,x)\in(BV\cap L^{\infty})(\Omega_{T}), where ΩT=(0,T)×\Omega_{T}=(0,T)\times\mathbb{R};

  2. (2)

    f(t,x,ξ)L(0,;+(2))f(t,x,\xi)\in L^{\infty}(0,\infty;\mathcal{M}^{+}(\mathbb{R}^{2}));

  3. (3)

    E(t,x)=x|x|f(t,x,ξ)𝑑ξE(t,x)=\frac{x}{|x|}*\int_{\mathbb{R}}f(t,x,\xi)d\xi a.e.;

  4. (4)

    ϕCc((0,T)×2)\forall\phi\in C_{c}^{\infty}((0,T)\times\mathbb{R}^{2}),

    0T2(tϕ)f+(xϕ)ξfdxdξdt0TE¯(ξϕ)f(dξ)𝑑x𝑑t=0.\displaystyle\int_{0}^{T}\iint_{\mathbb{R}^{2}}\left(\partial_{t}\phi\right)f+\left(\partial_{x}\phi\right)\xi fdxd\xi dt-\int_{0}^{T}\int_{\mathbb{R}}\overline{E}\int_{\mathbb{R}}\left(\partial_{\xi}\phi\right)f(d\xi)dxdt=0.
  5. (5)

    fC0,1([0,T);HL(2))f\in C^{0,1}([0,T);H^{-L}(\mathbb{R}^{2})) for some L>0L>0.

The term E¯(t,x)\overline{E}(t,x) in the above definition is the Volpert’s symmetric average:

(A.1) E¯(t,x)={E(t,x)if E(t,x) is approximately continuous at (t,x),El(t,x)+Er(t,x)2if E(t,x) has a jump at (t,x).\displaystyle\overline{E}(t,x)=\left\{\begin{aligned} &E(t,x)&\quad\text{if $E(t,x)$ is approximately continuous at $(t,x)$,}\\ &\frac{E_{l}(t,x)+E_{r}(t,x)}{2}&\quad\text{if $E(t,x)$ has a jump at $(t,x)$.}\end{aligned}\right.

where El(t,x)E_{l}(t,x) and Er(t,x)E_{r}(t,x) denote, respectively, the left and right limits of E(t,x)E(t,x) at a discontinuity line at (t,x)(t,x).

Appendix B Basic Properties of Bounded Variation Functions

We provide some basic properties of BV functions which are used in the paper. For more details, see for instance [46], or [50, 51].

Let Ω\Omega be a Borel measurable subset of 2\mathbb{R}^{2}.

  1. (1)

    If EBV(Ω)L(Ω)E\in BV(\Omega)\cap L^{\infty}(\Omega), then

    E2BV(Ω),E2=2E¯E\displaystyle E^{2}\in BV(\Omega),\quad\nabla E^{2}=2\overline{E}\nabla E

    in the sense of measures.

  2. (2)

    If u,vBV(Ω)u,v\in BV(\Omega), then u¯\overline{u} is almost everywhere defined and measurable with respect to v\nabla v. Furthermore, u¯\overline{u} is integrable with respect to v\nabla v if uu is bounded.

  3. (3)

    If u,vBV(Ω)u,v\in BV(\Omega), u¯\overline{u} is locally integrable with respect to v\nabla v and v¯\overline{v} is locally integrable with respect to u\nabla u. Then uvBV(Ω)uv\in BV(\Omega) and

    (uv)=u¯v+v¯u.\displaystyle\nabla(uv)=\overline{u}\nabla v+\overline{v}\nabla u.
  4. (4)

    φE¯=φE¯\overline{\varphi E}=\varphi\overline{E} if φC1(Ω)\varphi\in C^{1}(\Omega).

  5. (5)

    Let uBV(Ω)L(Ω)u\in BV(\Omega)\cap L^{\infty}(\Omega), and ησ\eta_{\sigma} be an approximation of the identity. Then

    (B.1) uησu¯1a.e.\displaystyle u*\eta_{\sigma}\to\overline{u}\quad\mathcal{H}^{1}-a.e.

    as σ0\sigma\to 0. Here, 1\mathcal{H}^{1} denotes the one-dimensional Hausdorff measure.

  6. (6)

    u\nabla u is absolutely continuous with respect to 1\mathcal{H}^{1} for any uBV(Ω)u\in BV(\Omega).

  7. (7)

    E¯\overline{E} is 1\mathcal{H}^{1}-a.e. defined for any EBVE\in BV.

Acknowledgements X. Chen was supported in part by U.S. NSF grant DMS-2406620. P. Zhang was supported in part by National Key R&\&D Program of China under Grant 2021YFA1000800 and NSF of China under Grants 12288201, 12031006. Z. Zhang was supported in part by National Key R&D Program of China under Grant 2023YFA1008801 and NSF of China under Grant 12288101.

References

  • [1] R. Adami, F. Golse, and A. Teta. Rigorous derivation of the cubic NLS in dimension one. J. Stat. Phys., 127(6):1193–1220, 2007.
  • [2] A. I. Akhiezer, I. Akhiezer, and R. V. Polovin. Plasma Electrodynamics: Linear Theory, volume 1. Elsevier, 2017.
  • [3] C. Bardos, L. Erdős, F. Golse, N. Mauser, and H.-T. Yau. Derivation of the Schrödinger-Poisson equation from the quantum NN-body problem. C. R. Math. Acad. Sci. Paris, 334(6):515–520, 2002.
  • [4] I. Ben Porat and F. Golse. Pickl’s proof of the quantum mean-field limit and quantum Klimontovich solutions. Lett. Math. Phys., 114(2):51, 2024.
  • [5] N. Benedikter, G. de Oliveira, and B. Schlein. Quantitative derivation of the Gross-Pitaevskii equation. Comm. Pure Appl. Math., 68(8):1399–1482, 2015.
  • [6] N. Benedikter, M. Porta, C. Saffirio, and B. Schlein. From the Hartree dynamics to the Vlasov equation. Arch. Ration. Mech. Anal., 221(1):273–334, 2016.
  • [7] C. Brennecke and B. Schlein. Gross-Pitaevskii dynamics for Bose-Einstein condensates. Anal. PDE, 12(6):1513–1596, 2019.
  • [8] R. Carles. Semi-classical analysis for nonlinear Schrödinger equations. World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, 2008.
  • [9] T. Chen and N. Pavlović. The quintic NLS as the mean field limit of a boson gas with three-body interactions. J. Funct. Anal., 260(4):959–997, 2011.
  • [10] T. Chen and N. Pavlović. Derivation of the cubic NLS and Gross-Pitaevskii hierarchy from manybody dynamics in d=3d=3 based on spacetime norms. Ann. Henri Poincaré, 15(3):543–588, 2014.
  • [11] X. Chen and J. Holmer. On the Klainerman-Machedon conjecture for the quantum BBGKY hierarchy with self-interaction. J. Eur. Math. Soc. (JEMS), 18(6):1161–1200, 2016.
  • [12] X. Chen and J. Holmer. The derivation of the 𝕋3\mathbb{T}^{3} energy-critical NLS from quantum many-body dynamics. Invent. Math., 217(2):433–547, 2019.
  • [13] X. Chen and J. Holmer. Quantitative derivation and scattering of the 3D cubic NLS in the energy space. Ann. PDE, 8(2):Paper No. 11, 39, 2022.
  • [14] X. Chen, S. Shen, J. Wu, and Z. Zhang. The derivation of the compressible Euler equation from quantum many-body dynamics. Peking Math. J., 7(1):35–90, 2024.
  • [15] X. Chen, S. Shen, and Z. Zhang. Quantitative derivation of the Euler-Poisson equation from quantum many-body dynamics. Peking Math. J., https://doi.org/10.1007/s42543-023-00065-5, 2023.
  • [16] Y. C. Chen, C. E. Simien, S. Laha, P. Gupta, Y. N. Martinez, P. G. Mickelson, S. B. Nagel, and T. C. Killian. Electron screening and kinetic-energy oscillations in a strongly coupled plasma. Phys. Rev. Lett., 93:265003, Dec 2004.
  • [17] R. DiPerna and P.-L. Lions. Solutions globales d’équations du type Vlasov-Poisson. C. R. Acad. Sci. Paris Sér. I Math., 307(12):655–658, 1988.
  • [18] R. J. DiPerna and P.-L. Lions. Global weak solutions of kinetic equations. Rend. Sem. Mat. Univ. Politec. Torino, 46(3):259–288 (1990), 1988.
  • [19] R. J. DiPerna and P.-L. Lions. On the Cauchy problem for Boltzmann equations: global existence and weak stability. Ann. of Math. (2), 130(2):321–366, 1989.
  • [20] A. Elgart, L. Erdős, B. Schlein, and H.-T. Yau. Nonlinear Hartree equation as the mean field limit of weakly coupled fermions. J. Math. Pures Appl. (9), 83(10):1241–1273, 2004.
  • [21] A. Elgart, L. Erdős, B. Schlein, and H.-T. Yau. Gross-Pitaevskii equation as the mean field limit of weakly coupled bosons. Arch. Ration. Mech. Anal., 179(2):265–283, 2006.
  • [22] A. Elgart and B. Schlein. Mean field dynamics of boson stars. Comm. Pure Appl. Math., 60(4):500–545, 2007.
  • [23] L. Erdős, B. Schlein, and H.-T. Yau. Derivation of the Gross-Pitaevskii hierarchy for the dynamics of Bose-Einstein condensate. Comm. Pure Appl. Math., 59(12):1659–1741, 2006.
  • [24] L. Erdős, B. Schlein, and H.-T. Yau. Derivation of the cubic non-linear Schrödinger equation from quantum dynamics of many-body systems. Invent. Math., 167(3):515–614, 2007.
  • [25] L. Erdős, B. Schlein, and H.-T. Yau. Rigorous derivation of the Gross-Pitaevskii equation with a large interaction potential. J. Amer. Math. Soc., 22(4):1099–1156, 2009.
  • [26] L. Erdős, B. Schlein, and H.-T. Yau. Derivation of the Gross-Pitaevskii equation for the dynamics of Bose-Einstein condensate. Ann. of Math. (2), 172(1):291–370, 2010.
  • [27] L. Erdős and H.-T. Yau. Derivation of the nonlinear Schrödinger equation from a many body Coulomb system. Adv. Theor. Math. Phys., 5(6):1169–1205, 2001.
  • [28] F. Golse, C. Mouhot, and T. Paul. On the mean field and classical limits of quantum mechanics. Comm. Math. Phys., 343(1):165–205, 2016.
  • [29] F. Golse and T. Paul. The Schrödinger equation in the mean-field and semiclassical regime. Arch. Ration. Mech. Anal., 223(1):57–94, 2017.
  • [30] F. Golse and T. Paul. Mean-field and classical limit for the NN-body quantum dynamics with Coulomb interaction. Comm. Pure Appl. Math., 75(6):1332–1376, 2022.
  • [31] S. Graffi, A. Martinez, and M. Pulvirenti. Mean-field approximation of quantum systems and classical limit. Math. Models Methods Appl. Sci., 13(1):59–73, 2003.
  • [32] M. Grillakis and M. Machedon. Pair excitations and the mean field approximation of interacting bosons, I. Comm. Math. Phys., 324(2):601–636, 2013.
  • [33] J. Heyd, G. E. Scuseria, and M. Ernzerhof. Hybrid functionals based on a screened Coulomb potential. J. Chem. Phys, 118(18):8207–8215, 2003.
  • [34] J. Hofierka, B. Cunningham, C. M. Rawlins, C. H. Patterson, and D. G. Green. Many-body theory of positron binding to polyatomic molecules. Nature, 606(7915):688–693, 2022.
  • [35] K. Kirkpatrick, B. Schlein, and G. Staffilani. Derivation of the two-dimensional nonlinear Schrödinger equation from many body quantum dynamics. Amer. J. Math., 133(1):91–130, 2011.
  • [36] S. Klainerman and M. Machedon. On the uniqueness of solutions to the Gross-Pitaevskii hierarchy. Comm. Math. Phys., 279(1):169–185, 2008.
  • [37] L. Lafleche and C. Saffirio. Strong semiclassical limits from Hartree and Hartree-Fock to Vlasov-Poisson equations. Anal. PDE, 16(4):891–926, 2023.
  • [38] P.-L. Lions and T. Paul. Sur les mesures de Wigner. Rev. Mat. Iberoamericana, 9(3):553–618, 1993.
  • [39] P.-L. Lions and B. Perthame. Propagation of moments and regularity for the 33-dimensional Vlasov-Poisson system. Invent. Math., 105(2):415–430, 1991.
  • [40] X. Liu, Z. Wang, K. Watanabe, T. Taniguchi, O. Vafek, and J. Li. Tuning electron correlation in magic-angle twisted bilayer graphene using Coulomb screening. Science, 371(6535):1261–1265, 2021.
  • [41] P. A. Markowich and N. J. Mauser. The classical limit of a self-consistent quantum-Vlasov equation in 33D. Math. Models Methods Appl. Sci., 3(1):109–124, 1993.
  • [42] P. Pickl. A simple derivation of mean field limits for quantum systems. Lett. Math. Phys., 97(2):151–164, 2011.
  • [43] I. Rodnianski and B. Schlein. Quantum fluctuations and rate of convergence towards mean field dynamics. Comm. Math. Phys., 291(1):31–61, 2009.
  • [44] F. J. Rogers, H. C. Graboske, and D. J. Harwood. Bound eigenstates of the static screened Coulomb potential. Phys. Rev. A, 1:1577–1586, Jun 1970.
  • [45] S. Serfaty. Mean field limit for Coulomb-type flows. Duke Math. J., 169(15):2887–2935, 2020. With an appendix by Mitia Duerinckx and Serfaty.
  • [46] A. I. Volpert. Spaces BV{\rm BV} and quasilinear equations. Mat. Sb. (N.S.), 73(115):255–302, 1967.
  • [47] T. O. Wehling, E. Şaşıoğlu, C. Friedrich, A. I. Lichtenstein, M. I. Katsnelson, and S. Blügel. Strength of effective Coulomb interactions in graphene and graphite. Phys. Rev. Lett., 106:236805, Jun 2011.
  • [48] P. Zhang. Wigner measure and the semiclassical limit of Schrödinger-Poisson equations. SIAM J. Math. Anal., 34(3):700–718, 2002.
  • [49] P. Zhang. Wigner measure and semiclassical limits of nonlinear Schrödinger equations, volume 17 of Courant Lecture Notes in Mathematics. Courant Institute of Mathematical Sciences, New York; American Mathematical Society, Providence, RI, 2008.
  • [50] P. Zhang, Y. Zheng, and N. J. Mauser. The limit from the Schrödinger-Poisson to the Vlasov-Poisson equations with general data in one dimension. Comm. Pure Appl. Math., 55(5):582–632, 2002.
  • [51] Y. Zheng and A. Majda. Existence of global weak solutions to one-component Vlasov-Poisson and Fokker-Planck-Poisson systems in one space dimension with measures as initial data. Comm. Pure Appl. Math., 47(10):1365–1401, 1994.