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

On the mean-field and semiclassical limit from quantum NN-body dynamics

Xuwen Chen Department of Mathematics, University of Rochester, Rochester, NY 14627, USA [email protected] Shunlin Shen School of Mathematical Sciences, Peking University, Beijing, 100871, China [email protected]  and  Zhifei Zhang School of Mathematical Sciences, Peking University, Beijing, 100871, China [email protected]
Abstract.

We study the mean-field and semiclassical limit of the quantum many-body dynamics with a repulsive δ\delta-type potential N3βV(Nβx)N^{3\beta}V(N^{\beta}x) and a Coulomb potential, which leads to a macroscopic fluid equation, the Euler-Poisson equation with pressure. We prove quantitative strong convergence of the quantum mass and momentum densities up to the first blow up time of the limiting equation. The main ingredient is a functional inequality on the δ\delta-type potential for the almost optimal case β(0,1)\beta\in(0,1), for which we give an analysis of the singular correlation structure between particles.

Key words and phrases:
Mean-field Limit, Semiclassical Limit, Compressible Euler Equation, Quantum Many-body Dynamics, Modulated Energy.
2010 Mathematics Subject Classification:
Primary 35Q31, 76N10, 81V70; Secondary 35Q55, 81Q05.

1. Introduction

1.1. Background and Problems

The foundations of microscopic physics are Newton’s and Schrödinger equations in the classical and the quantum case respectively. By the first principle of quantum mechanics, a quantum system of NN particles is described by a wave function satisfying a linear NN-body Schrödinger equation. In realistic systems like fluids, the particle number is so large that these NN-body equations are almost impossible to solve. The macroscopic dynamics are therefore modeled by phenomenological equations such as the Euler or the Navier-Stokes equations, which are an important part of many areas of pure and applied mathematics, science, and engineering. These macroscopic equations are usually derived from continuum under ideal assumptions, but they are, in principle, consequences of the microscopic physical laws of Newton or Schrödinger. A key goal of statistical mechanics is to justify these macroscopic equations from microscopic theories in appropriate limit regimes. It is thus of fundamental interest to establish macroscopic equations from the microscopic level.

In the current paper, we start from the bosonic111N2 and O2 molecules are bosons (99.03% of air) and 99.05% H2O molecules are bosons. quantum many-body dynamics with δ\delta-type and Coulomb potentials, and study the mean-field and semiclassical limit which would lead to macroscopic fluid equations as particle number NN tends to infinity and Planck’s constant \hbar tends to zero. The dynamics of NN quantum particles in 3D are governed by, according to the superposition principle, the linear NN-body Schrödinger equation:

(1.1) itψN,=HN,ψN,.\displaystyle i\hbar\partial_{t}\psi_{N,\hbar}=H_{N,\hbar}\psi_{N,\hbar}.

Our Hamiltonian HN,H_{N,\hbar} is

(1.2) HN,=j=1N122Δxj+1N1j<kNVN(xjxk)+κN1j<kNVc(xjxk),\displaystyle H_{N,\hbar}=\sum_{j=1}^{N}-\frac{1}{2}\hbar^{2}\Delta_{x_{j}}+\frac{1}{N}\sum_{1\leq j<k\leq N}V_{N}(x_{j}-x_{k})+\frac{\kappa}{N}\sum_{1\leq j<k\leq N}V_{c}(x_{j}-x_{k}),

where the factor 1/N1/N is the mean-field averaging factor. The δ\delta-type and Coulomb potentials are

(1.3) {VN(x)=N3βV(Nβx),Vc(x)=1|x|,\begin{cases}&V_{N}(x)=N^{3\beta}V(N^{\beta}x),\\ &V_{c}(x)=\frac{1}{|x|},\end{cases}

in which, the parameter β[0,1]\beta\in[0,1] characterises different density regimes which correspond to different physical situations.

To have a fixed number of variables in the NN\to\infty process, we define the marginal densities γN,(k)(t)\gamma_{N,\hbar}^{(k)}(t) associated with ψN,(t)=eitHN,ψN,(0)\psi_{N,\hbar}(t)=e^{itH_{N,\hbar}}\psi_{N,\hbar}(0) in kernel form by

(1.4) γN,(k)(t,Xk,Xk)=ψN,(t,Xk,XNk)ψN,¯(t,Xk,XNk)𝑑XNk,\displaystyle\gamma_{N,\hbar}^{(k)}(t,X_{k},X_{k}^{\prime})=\int\psi_{N,\hbar}(t,X_{k},X_{N-k})\overline{\psi_{N,\hbar}}(t,X_{k}^{\prime},X_{N-k})dX_{N-k},

where Xk=(x1,,xk)3kX_{k}=(x_{1},...,x_{k})\in\mathbb{R}^{3k} and XNk=(xk+1,,xN)3(Nk)X_{N-k}=(x_{k+1},...,x_{N})\in\mathbb{R}^{3(N-k)}. It is believed that nonlinear Schrödinger equations (NLS) is the mean-field limit equation for these quantum NN-body dynamics, that is,

(1.5) γN,(1)(t,x1,x1)|ϕ(t)ϕ(t)|,\displaystyle\gamma_{N,\hbar}^{(1)}(t,x_{1},x_{1}^{\prime})\sim|\phi(t)\rangle\langle\phi(t)|,

where ϕ(t)\phi(t) solves NLS.

There is a large amount of literature devoted to the mean-field theory from quantum many-body dynamics, such as [1, 2, 3, 4, 6, 7, 8, 9, 11, 10, 12, 13, 14, 15, 18, 16, 17, 19, 20, 21, 22, 25, 26, 32, 28, 29, 30, 31, 33, 39, 40, 37, 38, 36, 41, 42, 43, 44, 48, 46, 47, 55, 58, 59, 60, 61, 62]. In particular, for the case of defocusing δ\delta-type potential, it was Erdös, Schlein, and Yau who first rigorously derived the 3D cubic defocusing NLS from quantum many-body dynamics in their groundbreaking papers [28, 29, 30, 31].222Around the same time, see also [1] for 1D case.In their analysis, apart from the uniqueness of the infinite hierarchy which was widely regarded as the most involved part, understanding the singular correlation structure generated by the δ\delta-type potential was one of the main challenges.

In the mean-field limit as the particle number NN tends to infinity, the potential VNV_{N} converges formally to the Dirac-delta interaction (V)δ(\int V)\delta, also called the Fermi potential. For β<13\beta<\frac{1}{3}, the average distance between the particles, which is O(N13)O(N^{-\frac{1}{3}}), is much less than the range of the interaction potential, which is O(Nβ)O(N^{-\beta}), and there are many but weak correlations. For β>13\beta>\frac{1}{3}, then the analysis is much more involved because of the strong correlations between particles. For β\beta close or equal to 11, as the scaling is starting to match the Laplacian operator, it is expected that the δ\delta-type potential generates an interparticle singular correlation structure, closely related to the zero-energy scattering equation

(1.6) {(2Δ+1NVN(x))fN,(x)=0,lim|x|fN,(x)=1.\left\{\begin{aligned} &\left(-\hbar^{2}\Delta+\frac{1}{N}V_{N}(x)\right)f_{N,\hbar}(x)=0,\\ &\lim_{|x|\to\infty}f_{N,\hbar}(x)=1.\end{aligned}\right.

The scattering function (1fN,(x))(1-f_{N,\hbar}(x)) varies effectively on the short scale for |x|Nβ|x|\lesssim N^{-\beta} and has the same singularity as the Coulomb potential at infinity. It is believed that333See for example [49] for the static case and [28, 30, 31] for the time-dependent case., instead of the factorization property, that is,

γN,(2)(t,x1,x2;x1,x2)γN,(1)(t,x1;x1)γN,(1)(t,x2;x2),\displaystyle\gamma_{N,\hbar}^{(2)}(t,x_{1},x_{2};x_{1}^{\prime},x_{2}^{\prime})\sim\gamma_{N,\hbar}^{(1)}(t,x_{1};x_{1}^{\prime})\gamma_{N,\hbar}^{(1)}(t,x_{2};x_{2}^{\prime}),

the marginal densities should be considered as

γN,(2)(t,x1,x2;x1,x2)fN,(x1x2)fN,(x1x2)γN,(1)(t,x1;x1)γN,(1)(t,x2;x2).\displaystyle\gamma_{N,\hbar}^{(2)}(t,x_{1},x_{2};x_{1}^{\prime},x_{2}^{\prime})\sim f_{N,\hbar}(x_{1}-x_{2})f_{N,\hbar}(x_{1}^{\prime}-x_{2}^{\prime})\gamma_{N,\hbar}^{(1)}(t,x_{1};x_{1}^{\prime})\gamma_{N,\hbar}^{(1)}(t,x_{2};x_{2}^{\prime}).

The singular correlation structure is very subtle and plays a crucial role in the mean-field limit from quantum NN-body dynamics, as it gives an O(Nβ)O(N^{\beta}) correction to the NN-body energy.

For the semiclassical limit, the connection between Schrödinger-type equations and the classical fluid mechanics was already noted in 1927 by Madelung [51]. Starting from a single NLS, the asymptotic behavior of the wave function as the Planck’s constant goes to zero is studied by many authors using various approaches based on Madelung’s fluid mechanical formulation. See, for example, [35, 45, 50, 64]. For a more detailed survey related to semiclassical limits, see [5, 65] and references within. There are many deep problems on the study of classical limiting dynamics from quantum equations.

The joint mean-field and semiclassical limit from quantum NN-body dynamics formally gives a direct connection between quantum microscopic systems and classical macroscopic fluid equations. Providing a rigorous proof is certainly a challenging problem. For the repulsive Coulomb potential, Golse and Paul [34], based on Serfaty’s inequality [57, Proposition 1.1], justified the weak convergence to pressureless Euler-Poisson in the mean-field and semiclassical limit. For the case of the δ\delta-type potential, in our previous work [23], we derived the compressible Euler equations with strong and quantitative convergence rate from quantum many-body dynamics by a new strategy of combining the accuracy of the hierarchy method and the flexibility of the modulated energy method. Subsequently, such a scheme was adopted in [24] to obtain the quantitative convergence rate from quantum many-body dynamics to the pressureless Euler-Poisson equation.

Despite a series of progress on the mean-field and semiclassical limit from the quantum NN-body dynamics with singular potentials, a number of challenges remain open:

  1. (1)

    The derivation of the full Euler-Poisson equation with pressure from the quantum many-body dynamics. In [24, 34], the limiting Euler-Poisson equation is pressureless. However, the pressure is a fluid defining feature and essential for the macroscopic fluid equation. It is thus a fundamental question to understand the emergence of pressure from the microscopic level.

  2. (2)

    The large β\beta problem is known to be difficult in the mean-field and semiclassical limit due to the strong correlations between particles. The main challenge lies in the analysis of the singular correlation structure generated by the δ\delta-type potential.

  3. (3)

    To obtain the quantitative strong convergence rate, a double-exponential restriction between NN and \hbar was needed in [23], which is of course not optimal. From the perspective of energy, the restriction should be at least polynomial. To relax the double-exponential restriction, it requires new and finer techniques.

  4. (4)

    The scheme in [23] currently cannot deal with the 𝕋3\mathbb{T}^{3} case, since its proof highly relies on a key collapsing estimate, which fails in the H1H^{1} energy space for the 𝕋3\mathbb{T}^{3} case as proven in [36]. Thus, the 𝕋3\mathbb{T}^{3} case requires new ideas. The torus case is the beginning to understand other related and important problems, such as the microscopic descriptions of the Mach number and Knudsen number and their limit to incompressible fluids.

In this paper, our goal is to settle the above open problems.

1.2. Statement of the Main Theorem

Starting from the quantum NN-body dynamics (1.1), we take the normalization that ψN,(t)LXN2=1\|\psi_{N,\hbar}(t)\|_{L_{X_{N}}^{2}}=1, and define the quantum mass density and momentum density by

(1.7) ρN,(k)(t,Xk)=γN,(k)(t,Xk;Xk),JN,h(1)(t,x)=Im(x1γN,(1))(t,x;x).\displaystyle\rho_{N,\hbar}^{(k)}(t,X_{k})=\gamma_{N,\hbar}^{(k)}(t,X_{k};X_{k}),\quad J_{N,h}^{(1)}(t,x)=\operatorname{Im}\left(\hbar\nabla_{x_{1}}\gamma_{N,\hbar}^{(1)}\right)(t,x;x).

The limiting macroscopic equation would be the compressible Euler-Poisson equation with a pressure term P=b02ρ2P=\frac{b_{0}}{2}\rho^{2}, which is (in velocity form)

(1.8) {tρ+(ρu)=0,tu+(u)u+b0xρ+κx(Vcρ)=0,(ρ,u)|t=0=(ρin,uin),\begin{cases}&\partial_{t}\rho+\nabla\cdot\left(\rho u\right)=0,\\ &\partial_{t}u+(u\cdot\nabla)u+b_{0}\nabla_{x}\rho+\kappa\nabla_{x}(V_{c}*\rho)=0,\\ &(\rho,u)|_{t=0}=(\rho^{in},u^{in}),\end{cases}

or (in momentum form)

(1.9) {tρ+divJ=0,tJ+div(JJρ)+12(b0ρ2)+κρx(Vcρ)=0,(ρ,J)|t=0=(ρin,Jin).\begin{cases}&\partial_{t}\rho+\operatorname{div}J=0,\\ &\partial_{t}J+\operatorname{div}\left(\frac{J\otimes J}{\rho}\right)+\frac{1}{2}\nabla\left(b_{0}\rho^{2}\right)+\kappa\rho\nabla_{x}(V_{c}*\rho)=0,\\ &(\rho,J)|_{t=0}=(\rho^{in},J^{in}).\end{cases}

Here, as usual,

ρ(t,x):×3\displaystyle\rho(t,x):\mathbb{R}\times\mathbb{R}^{3}\mapsto\mathbb{R}
u(t,x)=(u1(t,x),u2(t,x),u3(t,x)):×33\displaystyle u(t,x)=(u^{1}(t,x),u^{2}(t,x),u^{3}(t,x)):\mathbb{R}\times\mathbb{R}^{3}\mapsto\mathbb{R}^{3}
J(t,x)=(ρu)(t,x):×33\displaystyle J(t,x)=\left(\rho u\right)(t,x):\mathbb{R}\times\mathbb{R}^{3}\mapsto\mathbb{R}^{3}

are respectively the mass density, the velocity, and the momentum of the fluid. The coupling constant b0=Vb_{0}=\int V is the macroscopic effect of the microscopic interaction VV. When the coefficient κ=0\kappa=0, the system (1.8)(\ref{equ:euler equation}) is reduced to a compressible Euler equation. Specifically, we consider the initial data satisfying the condition

(1.10) {ρinHs1(3),uinHs(3),ρin(x)0,3ρin(x)𝑑x=1,\left\{\begin{aligned} \rho^{in}\in H^{s-1}(\mathbb{R}^{3}),\quad u^{in}\in H^{s}(\mathbb{R}^{3}),\\ \rho^{in}(x)\geq 0,\quad\int_{\mathbb{R}^{3}}\rho^{in}(x)dx=1,\end{aligned}\right.

with s>92s>\frac{9}{2} and ss\in\mathbb{N}, so that the Euler-Poisson system (1.8)(\ref{equ:euler equation}) has a unique solution (ρ,u)(\rho,u) up to some time T0T_{0} such that444We are not dealing with sharp well-posedness of (1.8) here. The local well-posedness of the Euler system here is known by the standard theory on hyperbolic systems,see [52, 53, 54].

(1.11) {ρC([0,T0];Hs1(3)),uC([0,T0];Hs(3)),ρ(t,x)0,3ρ(t,x)𝑑x=1.\left\{\begin{aligned} &\rho\in C([0,T_{0}];H^{s-1}(\mathbb{R}^{3})),\quad u\in C([0,T_{0}];H^{s}(\mathbb{R}^{3})),\\ &\rho(t,x)\geq 0,\quad\int_{\mathbb{R}^{3}}\rho(t,x)dx=1.\end{aligned}\right.
Theorem 1.1.

Let β(0,1)\beta\in(0,1), κ0\kappa\geq 0 and the marginal densities ΓN,={γN,(k)}\Gamma_{N,\hbar}=\left\{\gamma_{N,\hbar}^{(k)}\right\} associated with ψN,\psi_{N,\hbar} be the solution to the NN-body dynamics (1.1) with a smooth compactly supported, spherically symmetric nonnegative potential VV and a repulsive Coulomb potential VcV_{c}. Assume the initial data satisfy the following conditions:

(a)(a) ψN,(0)\psi_{N,\hbar}(0) is normalized and the NN-body energy bound holds:

(1.12) ψN,(0),(HN,/N+1)2ψN,(0)(E0)2,\displaystyle\langle\psi_{N,\hbar}(0),(H_{N,\hbar}/N+1)^{2}\psi_{N,\hbar}(0)\rangle\leq(E_{0})^{2},

for some E0>0E_{0}>0.

(b)(b) The initial data (ρin,uin)(\rho^{in},u^{in}) to (1.8)(\ref{equ:euler equation}) satisfy condition (1.10)(\ref{equ:initial condition,euler}) with s=5s=5, and the modulated energy between (1.1) and (1.8) at initial time tends to zero, that is, (0)0\mathcal{M}(0)\to 0 where

(0)\displaystyle\mathcal{M}(0)
:=\displaystyle:= 3N|(ix1u(t,x1))ψN,(0,XN)|2𝑑XN\displaystyle\int_{\mathbb{R}^{3N}}|\left(i\hbar\nabla_{x_{1}}-u(t,x_{1})\right)\psi_{N,\hbar}(0,X_{N})|^{2}dX_{N}
+N1NVN(x1x2)ρN,(2)(0,x1,x2)𝑑x1𝑑x2\displaystyle+\frac{N-1}{N}\int V_{N}(x_{1}-x_{2})\rho_{N,\hbar}^{(2)}(0,x_{1},x_{2})dx_{1}dx_{2}
+b03ρin(x1)ρin(x1)𝑑x12b03ρin(x1)ρN,(1)(0,x1)𝑑x1\displaystyle+b_{0}\int_{\mathbb{R}^{3}}\rho^{in}(x_{1})\rho^{in}(x_{1})dx_{1}-2b_{0}\int_{\mathbb{R}^{3}}\rho^{in}(x_{1})\rho_{N,\hbar}^{(1)}(0,x_{1})dx_{1}
+Vc(x1x2)[N1NρN,(2)(0,x1,x2)+ρin(x1)ρin(x2)2ρin(x1)ρN,(1)(0,x2)]𝑑x1𝑑x2.\displaystyle+\int V_{c}(x_{1}-x_{2})\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(0,x_{1},x_{2})+\rho^{in}(x_{1})\rho^{in}(x_{2})-2\rho^{in}(x_{1})\rho_{N,\hbar}^{(1)}(0,x_{2})\right]dx_{1}dx_{2}.

Then under the polynomial restriction555(1.13) is in fact a rational restriction. We say polynomial to avoid confusing “rational” and “reasonable”.

(1.13) r(N,)=C(Nβ16+Nβ34+N1104)0,\displaystyle r(N,\hbar)=C(N^{\beta-1}\hbar^{-6}+N^{-\frac{\beta}{3}}\hbar^{-4}+N^{-\frac{1}{10}}\hbar^{-4})\to 0,

we have the quantitative estimates on the strong convergence of the mass density

(1.14) ρN,(1)(t,x)ρ(t,x)Lt[0,T0]Lx2(3)2(0)+r(N,)+2,\displaystyle\|\rho_{N,\hbar}^{(1)}(t,x)-\rho(t,x)\|_{L_{t}^{\infty}[0,T_{0}]L_{x}^{2}(\mathbb{R}^{3})}^{2}\lesssim\mathcal{M}(0)+r(N,\hbar)+\hbar^{2},

and on the convergence of the momentum density

(1.15) JN,(1)(t,x)(ρu)(t,x)Lt[0,T0]Lx1(3)2(0)+r(N,)+2.\displaystyle\Big{\|}J_{N,\hbar}^{(1)}(t,x)-(\rho u)(t,x)\Big{\|}_{L_{t}^{\infty}[0,T_{0}]L_{x}^{1}(\mathbb{R}^{3})}^{2}\lesssim\mathcal{M}(0)+r(N,\hbar)+\hbar^{2}.

When κ>0\kappa>0, Theorem 1.1 is the first result which simultaneously deals with the δ\delta-type and Coulomb potentials and establishes the quantitative strong convergence to the full Euler-Poisson equation with pressure. Compared to [34, 24], the emergence of the pressure term is the main novelty. We point out that, it is not clear if the scheme in [23, 24] can handle the δ\delta-type and Coulomb potentials simultaneously, since the energy estimates and collapsing estimates are totally different in the δ\delta-type and Coulomb potentials. Therefore, it requires completely new ideas for a simultaneous consideration of δ\delta-type and Coulomb potentials.

When κ=0\kappa=0, it reduces to the sole δ\delta-type potential case. Compared with our previous work [23], we here list the breakthroughs.

  1. (1)

    The parameter β\beta is extended to the full range of (0,1)(0,1), which is almost optimal in the dilute regime.

  2. (2)

    The previous double-exponential restriction between NN and \hbar is relaxed to be polynomial, which is a tremendous improvement.

  3. (3)

    Our new approach also works for the 𝕋3\mathbb{T}^{3} case with slight modifications, as the proof is independent of the hardcore harmonic analysis on 𝕋3\mathbb{T}^{3}.

Additionally, the convergence rate 2\hbar^{2} should be optimal since the convergence rate of the modulated kinetic energy part at initial time is at most the order of 2\hbar^{2}. Besides, this can be achieved with WKB type initial data.

1.3. Outline of the Proof

The proof is based on a modulated energy method.666A closely related method is the relative entropy method, see for example, [63]. The modulated energy we use includes three parts

(1.16) (t)=K(t)+δ(t)+c(t),\displaystyle\mathcal{M}(t)=\mathcal{M}_{K}(t)+\mathcal{F}_{\delta}(t)+\mathcal{F}_{c}(t),

where the kinetic energy part is

(1.17) K(t)=3N|(ix1u(t,x1))ψN,(t,XN)|2𝑑XN,\displaystyle\mathcal{M}_{K}(t)=\int_{\mathbb{R}^{3N}}|\left(i\hbar\nabla_{x_{1}}-u(t,x_{1})\right)\psi_{N,\hbar}(t,X_{N})|^{2}dX_{N},

the δ\delta-type potential part is

(1.18) δ(t)=\displaystyle\mathcal{F}_{\delta}(t)= N1N3×3VN(x1x2)ρN,(2)(t,x1,x2)𝑑x1𝑑x2\displaystyle\frac{N-1}{N}\iint_{\mathbb{R}^{3}\times\mathbb{R}^{3}}V_{N}(x_{1}-x_{2})\rho_{N,\hbar}^{(2)}(t,x_{1},x_{2})dx_{1}dx_{2}
+b03ρ(t,x1)ρ(t,x1)𝑑x12b03ρ(t,x1)ρN,(1)(t,x1)𝑑x1,\displaystyle+b_{0}\int_{\mathbb{R}^{3}}\rho(t,x_{1})\rho(t,x_{1})dx_{1}-2b_{0}\int_{\mathbb{R}^{3}}\rho(t,x_{1})\rho_{N,\hbar}^{(1)}(t,x_{1})dx_{1},

and the Coulomb potential part is

(1.19) c(t)=\displaystyle\mathcal{F}_{c}(t)= 3×3Vc(x1x2)[N1NρN,(2)(t,x1,x2)\displaystyle\iint_{\mathbb{R}^{3}\times\mathbb{R}^{3}}V_{c}(x_{1}-x_{2})\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(t,x_{1},x_{2})\right.
+ρ(t,x1)ρ(t,x2)2ρ(t,x1)ρN,(1)(t,x2)]dx1dx2.\displaystyle\left.\quad\quad\quad+\rho(t,x_{1})\rho(t,x_{2})-2\rho(t,x_{1})\rho_{N,\hbar}^{(1)}(t,x_{2})\right]dx_{1}dx_{2}.

In Section 2, we first derive the time evolution of the modulated energy

(1.20) ddt(t)=~K(t)+~δ(t)+~c(t),\displaystyle\frac{d}{dt}\mathcal{M}(t)=\widetilde{\mathcal{M}}_{K}(t)+\widetilde{\mathcal{F}}_{\delta}(t)+\widetilde{\mathcal{F}}_{c}(t),

where the kinetic energy contribution part is

~K(t)=\displaystyle\widetilde{\mathcal{M}}_{K}(t)= j,k=133N(juk+kuj)(ijψN,ujψN,)(ikψN,ukψN,)¯𝑑XN\displaystyle-\sum_{j,k=1}^{3}\int_{\mathbb{R}^{3N}}\left(\partial_{j}u^{k}+\partial_{k}u^{j}\right)(-i\hbar\partial_{j}\psi_{N,\hbar}-u^{j}\psi_{N,\hbar})\overline{(-i\hbar\partial_{k}\psi_{N,\hbar}-u^{k}\psi_{N,\hbar})}dX_{N}
+223Δ(divu)(t,x1)ρN,(1)(t,x1)𝑑x1,\displaystyle+\frac{\hbar^{2}}{2}\int_{\mathbb{R}^{3}}\Delta(\operatorname{div}u)(t,x_{1})\rho_{N,\hbar}^{(1)}(t,x_{1})dx_{1},

with the notations u=(u1,u2,u3)u=(u^{1},u^{2},u^{3}), x1=(x11,x12,x13)3x_{1}=(x_{1}^{1},x_{1}^{2},x_{1}^{3})\in\mathbb{R}^{3} and j=x1j\partial_{j}=\partial_{x_{1}^{j}}, the δ\delta-type potential contribution part is

(1.21) ~δ(t)=\displaystyle\widetilde{\mathcal{F}}_{\delta}(t)= N1N(u(t,x1)u(t,x2))VN(x1x2)ρN,(2)(t,x1,x2)𝑑x1𝑑x2\displaystyle\frac{N-1}{N}\int(u(t,x_{1})-u(t,x_{2}))\nabla V_{N}(x_{1}-x_{2})\rho_{N,\hbar}^{(2)}(t,x_{1},x_{2})dx_{1}dx_{2}
b0divu(t,x1)ρ(t,x1)[ρ(t,x1)2ρN,(1)(t,x1)]𝑑x1,\displaystyle-b_{0}\int\operatorname{div}u(t,x_{1})\rho(t,x_{1})\left[\rho(t,x_{1})-2\rho_{N,\hbar}^{(1)}(t,x_{1})\right]dx_{1},

and the Coulomb potential contribution part is

(1.22) ~c(t)=\displaystyle\widetilde{\mathcal{F}}_{c}(t)= (u(t,x1)u(t,x2))Vc(x1x2)\displaystyle\int(u(t,x_{1})-u(t,x_{2}))\nabla V_{c}(x_{1}-x_{2})
[N1NρN,(2)(t,x1,x2)+ρ(t,x1)ρ(t,x2)2ρ(t,x1)ρN,(1)(t,x2)]dx1dx2.\displaystyle\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(t,x_{1},x_{2})+\rho(t,x_{1})\rho(t,x_{2})-2\rho(t,x_{1})\rho_{N,\hbar}^{(1)}(t,x_{2})\right]dx_{1}dx_{2}.

It is easy to control the kinetic energy contribution part

(1.23) ~K(t)K(t)+2.\displaystyle\widetilde{\mathcal{M}}_{K}(t)\lesssim\mathcal{M}_{K}(t)+\hbar^{2}.

The toughest part in the modulated energy method is to control the potential contribution part both in the classical and quantum setting. See, for example, [27, 34, 50, 56, 57, 64]. In the classical mean-field limit with Coulomb potential, Serfaty in [57, Proposition 1.1] establishes a crucial functional inequality to solve this challenging problem. Then for the quantum many-body systems with Coulomb potential, based on Serfaty’s inequality, Golse and Paul in [34] managed to control the Coulomb potential contribution part ~c(t)\widetilde{\mathcal{F}}_{c}(t) as follows

(1.24) ~c(t)\displaystyle\widetilde{\mathcal{F}}_{c}(t)\lesssim c(t)+CN13,\displaystyle\mathcal{F}_{c}(t)+CN^{-\frac{1}{3}},
(1.25) 0\displaystyle 0\leq c(t)+CN23.\displaystyle\mathcal{F}_{c}(t)+CN^{-\frac{2}{3}}.

Serfaty’s inequality is a special and impressive tool based on deep observations of the structure of Coulomb potential. It is limited to a special class of singular potentials, as its proof highly relies on the structure and the profile of the potentials, such as the Coulomb characteristic that ΔVc=c0δ-\Delta V_{c}=c_{0}\delta. Therefore, it is quite difficult to establish a Serfaty’s inequality for the δ\delta-type potential case, because of the general profile and sharp singularity of the δ\delta-type potential. In fact, due to the presence of the singular correlation structure caused by the δ\delta-type potential, the analysis would be totally different and is expected to be rather intricate.

In this paper, we develop a new scheme without using Serfaty’s inequality to control the δ\delta-type potential parts δ(t)\mathcal{F}_{\delta}(t) and ~δ(t)\widetilde{\mathcal{F}}_{\delta}(t) and establish

(1.26) ~δ(t)\displaystyle\widetilde{\mathcal{F}}_{\delta}(t)\lesssim δ(t)+r(N,),\displaystyle\mathcal{F}_{\delta}(t)+r(N,\hbar),
(1.27) 0\displaystyle 0\leq δ(t)+r(N,).\displaystyle\mathcal{F}_{\delta}(t)+r(N,\hbar).

The proof is divided in several steps.

Step 1. Preliminary reduction. Applying the approximation of identity to the one-body term of δ(t)\mathcal{F}_{\delta}(t), we have the approximation

(1.28) δ(t)\displaystyle\mathcal{F}_{\delta}(t)\sim VN(xy)[N1NρN,(2)(t,x,y)\displaystyle\int V_{N}(x-y)\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(t,x,y)\right.
ρN,(1)(t,x)ρ(t,y)ρ(t,x)ρN,(1)(t,y)+ρ(t,x)ρ(t,y)]dxdy.\displaystyle\quad\quad\left.-\rho_{N,\hbar}^{(1)}(t,x)\rho(t,y)-\rho(t,x)\rho_{N,\hbar}^{(1)}(t,y)+\rho(t,x)\rho(t,y)\right]dxdy.

To have a closed estimate, namely, letting ~δ(t)\widetilde{\mathcal{F}}_{\delta}(t) match the approximation of δ(t)\mathcal{F}_{\delta}(t), we get by integration by parts for the two-body term that

~δ(t)=\displaystyle\widetilde{\mathcal{F}}_{\delta}(t)= N1Ndivu(t,x1)VN(x1x2)ρN,(2)(t,x1,x2)𝑑x1𝑑x2\displaystyle\frac{N-1}{N}\int\operatorname{div}u(t,x_{1})V_{N}(x_{1}-x_{2})\rho_{N,\hbar}^{(2)}(t,x_{1},x_{2})dx_{1}dx_{2}
(u(t,x)u(t,y))VN(xy)xρN,(2)(t,x,y)𝑑x𝑑y\displaystyle-\int(u(t,x)-u(t,y))V_{N}(x-y)\nabla_{x}\rho_{N,\hbar}^{(2)}(t,x,y)dxdy
b0divu(t,x1)ρ(t,x1)[ρ(t,x1)2ρN,(1)(t,x1)]𝑑x1.\displaystyle-b_{0}\int\operatorname{div}u(t,x_{1})\rho(t,x_{1})\left[\rho(t,x_{1})-2\rho_{N,\hbar}^{(1)}(t,x_{1})\right]dx_{1}.

Using the approximation of identity to the one-body term again, we decompose ~δ(t)\widetilde{\mathcal{F}}_{\delta}(t) into the main part and error part

(1.29) ~δ(t)=MP+EP,\displaystyle\widetilde{\mathcal{F}}_{\delta}(t)=MP+EP,

where

(1.30) MP\displaystyle MP\sim divu(t,x)VN(xy)[N1NρN,(2)(t,x,y)\displaystyle-\int\operatorname{div}u(t,x)V_{N}(x-y)\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(t,x,y)\right.
ρN,(1)(t,x)ρ(t,y)ρ(t,x)ρN,(1)(t,y)+ρ(t,x)ρ(t,y)]dxdy,\displaystyle\left.\quad\quad-\rho_{N,\hbar}^{(1)}(t,x)\rho(t,y)-\rho(t,x)\rho_{N,\hbar}^{(1)}(t,y)+\rho(t,x)\rho(t,y)\right]dxdy,
(1.31) EP=\displaystyle EP= (u(t,x)u(t,y))VN(xy)xρN,(2)(t,x,y)𝑑x𝑑y.\displaystyle-\int(u(t,x)-u(t,y))V_{N}(x-y)\nabla_{x}\rho_{N,\hbar}^{(2)}(t,x,y)dxdy.

Such a decomposition is based on the key observation that the difference coupled with the δ\delta-type potential

(1.32) (u(t,x)u(t,y))VN(xy),\displaystyle(u(t,x)-u(t,y))V_{N}(x-y),

when it is tested against a regular function, would vanish in the NN\to\infty limit. Such a structure is notably special for the δ\delta-type potential, since the difference coupled with a common potential including the Coulomb case cannot provide any smallness.

To prove that the error part (1.31) is indeed a small term, it requires the regularity of the two-body density function. Therefore, we delve into the analysis of two-body energy estimates, then deal with the error part and the main part in the Step 3 and 4 respectively.

Step 2. Two-body energy estimate. As usual, a-priori estimates are one of the toughest parts in the study of many-body dynamics as one must seek a regularity high enough for the limiting argument and at the same time low enough that it is provable. In Section 3, we prove that the wave function with added the singular correlation structure satisfies the two-body H1H^{1} energy bound

(1.33) (12Δx1)(12Δx2)ψN,(t,XN)1wN,(x1x2),ψN,(t,XN)1wN,(x1x2)C,\Big{\langle}(1-\hbar^{2}\Delta_{x_{1}})(1-\hbar^{2}\Delta_{x_{2}})\frac{\psi_{N,\hbar}(t,X_{N})}{1-w_{N,\hbar}(x_{1}-x_{2})},\frac{\psi_{N,\hbar}(t,X_{N})}{1-w_{N,\hbar}(x_{1}-x_{2})}\Big{\rangle}\leq C,

where wN,(x)w_{N,\hbar}(x) satisfies the zero-energy scattering equation

(1.34) {(2Δ+1NVN(x))(1wN,(x))=0,lim|x|wN,(x)=0.\left\{\begin{aligned} &\left(-\hbar^{2}\Delta+\frac{1}{N}V_{N}(x)\right)(1-w_{N,\hbar}(x))=0,\\ &\lim_{|x|\to\infty}w_{N,\hbar}(x)=0.\end{aligned}\right.

The singular correlation function wN,(x)w_{N,\hbar}(x) varies effectively on the short scale for |x|Nβ|x|\lesssim N^{-\beta} and has the same singularity as the Coulomb potential at infinity.

One of the main difficulties here is to understand the interparticle singular correlation structure generated by the δ\delta-type potential. See, for example, [49] for the study of the static case of Bose gas. For the time-dependent systems, Erdös, Schlein, and Yau [28, 30, 31] first introduced the two-body energy estimate which plays a central role in the derivation of Gross-Pitaevskii equation with the nonlinear interaction given by a scattering length. However, instead of showing the emergence of the scattering length, our purpose here is proving the functional inequalities (1.26) and (1.27).

Another difficulty lies in the Coulomb singularity. The Coulomb potential, if taken to high powers, results in singularities which cannot be controlled by derivatives. The (HN,)2(H_{N,\hbar})^{2} energy estimate (1.33) we prove (and require here) is at the borderline case. Indeed, the square of the Coulomb potential is bounded with respect to the kinetic energy in the sense that as operators |Vc(x)|2C(1Δx)|V_{c}(x)|^{2}\leq C(1-\Delta_{x}). However, no such estimates hold for |Vc(x)|3|V_{c}(x)|^{3} due to the singularity of the origin.

Step 3. Analysis of the Error Part. After setting up the energy estimates, we begin to analyze the error part (1.31). Because of the presence of the singular correlation structure, the two-body density function lacks the a-priori energy bound but can be decomposed into the singular and regular (relatively speaking) parts

(1.35) ρN,(2)(t,x,y)=(1wN,(xy))2ρN,(2)(t,x,y)(1wN,(xy))2.\displaystyle\rho_{N,\hbar}^{(2)}(t,x,y)=(1-w_{N,\hbar}(x-y))^{2}\frac{\rho_{N,\hbar}^{(2)}(t,x,y)}{(1-w_{N,\hbar}(x-y))^{2}}.

Hence, we need to rewrite the error part (1.31) as

(u(t,x)u(t,y))VN(xy)x[(1wN,(xy))2ρN,(2)(t,x,y)(1wN,(xy))2]dxdy\displaystyle\int(u(t,x)-u(t,y))\cdot V_{N}(x-y)\nabla_{x}\left[(1-w_{N,\hbar}(x-y))^{2}\frac{\rho_{N,\hbar}^{(2)}(t,x,y)}{(1-w_{N,\hbar}(x-y))^{2}}\right]dxdy
=\displaystyle= (u(t,x)u(t,y))VN(xy)(x(1wN,(xy))2)ρN,(2)(t,x,y)(1wN,(xy))2dxdy\displaystyle\int(u(t,x)-u(t,y))\cdot V_{N}(x-y)\left(\nabla_{x}(1-w_{N,\hbar}(x-y))^{2}\right)\frac{\rho_{N,\hbar}^{(2)}(t,x,y)}{(1-w_{N,\hbar}(x-y))^{2}}dxdy
+(u(t,x)u(t,y))VN(xy)(1wN,(xy))2x[ρN,(2)(t,x,y)(1wN,(xy))2]dxdy.\displaystyle+\int(u(t,x)-u(t,y))\cdot V_{N}(x-y)(1-w_{N,\hbar}(x-y))^{2}\nabla_{x}\left[\frac{\rho_{N,\hbar}^{(2)}(t,x,y)}{(1-w_{N,\hbar}(x-y))^{2}}\right]dxdy.

When the derivative hits the singular correlation function, it produces singularities by the defining feature of the singular correlation function, which would give a rise of O(Nβ)O(N^{\beta}). On the other hand, when the derivative hits the (relatively) regular part, it still requires a careful analysis as we have limited regularity as discussed before on the modified two-body density function.

In Section 4.1, we prove that, the cancellation structure (1.32) indeed dominates the singularity generated by the δ\delta-type potential and singular correlation function, and obtain the error estimate

(1.36) EPNβ16+Nβ24.\displaystyle EP\lesssim N^{\beta-1}\hbar^{-6}+N^{-\frac{\beta}{2}}\hbar^{-4}.

Step 4. Analysis of the Main Part. One difficulty of the analysis of the main part (1.30) is the sharp singularity and the unknown profile of VN(x)V_{N}(x). To overcome it, our strategy is to replace VN(x)V_{N}(x) with a slowly varying potential GN(x)G_{N}(x) which enjoys a number of good properties, but it comes at a price of the integrand’s regularity. Thus, for the main part (1.30), we again need to decompose the two-body density function into the singular part and relatively regular part as follows

(1.37) MP=\displaystyle MP= divu(t,x)VN(xy)[N1NρN,(2)(t,x,y)(1wN,(xy))2(1wN,(xy))2\displaystyle\int\operatorname{div}u(t,x)V_{N}(x-y)\left[\frac{N-1}{N}\frac{\rho_{N,\hbar}^{(2)}(t,x,y)}{(1-w_{N,\hbar}(x-y))^{2}}(1-w_{N,\hbar}(x-y))^{2}\right.
ρN,(1)(t,x)ρ(t,y)ρ(t,x)ρN,(1)(t,y)+ρ(t,x)ρ(t,y)]dxdy.\displaystyle\left.\quad\quad-\rho_{N,\hbar}^{(1)}(t,x)\rho(t,y)-\rho(t,x)\rho_{N,\hbar}^{(1)}(t,y)+\rho(t,x)\rho(t,y)\right]dxdy.

Note that (1wN,(xy))21+O(wN,(xy))(1-w_{N,\hbar}(x-y))^{2}\sim 1+O(w_{N,\hbar}(x-y)). Then by the two-body energy bound and the property for the scattering function wN,(xy)w_{N,\hbar}(x-y), we can prove that

(1.38) MP\displaystyle MP\sim divu(t,x)VN(xy)[N1NρN,(2)(t,x,y)(1wN,(xy))2\displaystyle-\int\operatorname{div}u(t,x)V_{N}(x-y)\left[\frac{N-1}{N}\frac{\rho_{N,\hbar}^{(2)}(t,x,y)}{(1-w_{N,\hbar}(x-y))^{2}}\right.
ρN,(1)(t,x)ρ(t,y)ρ(t,x)ρN,(1)(t,y)+ρ(t,x)ρ(t,y)]dxdy.\displaystyle\left.\quad\quad-\rho_{N,\hbar}^{(1)}(t,x)\rho(t,y)-\rho(t,x)\rho_{N,\hbar}^{(1)}(t,y)+\rho(t,x)\rho(t,y)\right]dxdy.

Since the integrand now enjoys the energy bound, we are able to replace VNV_{N} by GNG_{N} and get

(1.39) MP\displaystyle MP\sim b0divu(t,x)GN(xy)[N1NρN,(2)(t,x,y)(1wN,(xy))2\displaystyle-b_{0}\int\operatorname{div}u(t,x)G_{N}(x-y)\left[\frac{N-1}{N}\frac{\rho_{N,\hbar}^{(2)}(t,x,y)}{(1-w_{N,\hbar}(x-y))^{2}}\right.
ρN,(1)(t,x)ρ(t,y)ρ(t,x)ρN,(1)(t,y)+ρ(t,x)ρ(t,y)]dxdy,\displaystyle\left.\quad\quad-\rho_{N,\hbar}^{(1)}(t,x)\rho(t,y)-\rho(t,x)\rho_{N,\hbar}^{(1)}(t,y)+\rho(t,x)\rho(t,y)\right]dxdy,

where GN(x)=N3ηG(Nηx)G_{N}(x)=N^{3\eta}G(N^{\eta}x) with η<13\eta<\frac{1}{3}.

In Section 4.2, we will give a detailed proof of the above analysis and and arrive at the approximations of δ(t)\mathcal{F}_{\delta}(t) and ~δ(t)\widetilde{\mathcal{F}}_{\delta}(t) given by

(1.40) δ(t)\displaystyle\mathcal{F}_{\delta}(t)\sim b0GN(xy)[N1NρN,(2)(t,x,y)\displaystyle b_{0}\int G_{N}(x-y)\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(t,x,y)\right.
ρN,(1)(t,x)ρ(t,y)ρ(t,x)ρN,(1)(t,y)+ρ(t,x)ρ(t,y)]dxdy,\displaystyle\quad\quad\left.-\rho_{N,\hbar}^{(1)}(t,x)\rho(t,y)-\rho(t,x)\rho_{N,\hbar}^{(1)}(t,y)+\rho(t,x)\rho(t,y)\right]dxdy,

and

(1.41) ~δ(t)\displaystyle\widetilde{\mathcal{F}}_{\delta}(t)\sim b0divu(t,x)GN(xy)[N1NρN,(2)(t,x,y)\displaystyle-b_{0}\int\operatorname{div}u(t,x)G_{N}(x-y)\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(t,x,y)\right.
ρN,(1)(t,x)ρ(t,y)ρ(t,x)ρN,(1)(t,y)+ρ(t,x)ρ(t,y)]dxdy.\displaystyle\left.\quad\quad-\rho_{N,\hbar}^{(1)}(t,x)\rho(t,y)-\rho(t,x)\rho_{N,\hbar}^{(1)}(t,y)+\rho(t,x)\rho(t,y)\right]dxdy.

Now, from the approximations of δ(t)\mathcal{F}_{\delta}(t) and ~δ(t)\widetilde{\mathcal{F}}_{\delta}(t), we are left to prove a reduced form of the functional inequality

divu(x)GN(xy)[N1NρN,(2)(x,y)ρN,(1)(x)ρ(y)ρ(x)ρN,(1)(y)+ρ(x)ρ(y)]𝑑x𝑑y\displaystyle\int\operatorname{div}u(x)G_{N}(x-y)\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(x,y)-\rho_{N,\hbar}^{(1)}(x)\rho(y)-\rho(x)\rho_{N,\hbar}^{(1)}(y)+\rho(x)\rho(y)\right]dxdy
\displaystyle\lesssim GN(xy)[N1NρN,(2)(x,y)ρN,(1)(x)ρ(y)ρ(x)ρN,(1)(y)+ρ(x)ρ(y)]𝑑x𝑑y+o(1),\displaystyle\int G_{N}(x-y)\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(x,y)-\rho_{N,\hbar}^{(1)}(x)\rho(y)-\rho(x)\rho_{N,\hbar}^{(1)}(y)+\rho(x)\rho(y)\right]dxdy+o(1),

which looks more concise and tractable than the original functional inequality (1.26). But, it is unknown if the integrand

(1.42) N1NρN,(2)(x,y)ρN,(1)(x)ρ(y)ρ(x)ρN,(1)(y)+ρ(x)ρ(y)\displaystyle\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(x,y)-\rho_{N,\hbar}^{(1)}(x)\rho(y)-\rho(x)\rho_{N,\hbar}^{(1)}(y)+\rho(x)\rho(y)

is non-negative. We cannot simply rule out the term divu(x)\operatorname{div}u(x) either. Thus, it is still non-trivial to deduce the inequality. In fact, as we will see in Section 4.3, the special structure (1.42) with a slowly varying potential GN(x)G_{N}(x) plays a crucial role in establishing the reduced version of functional inequality. Then, at the end of Section 4.3, we conclude the functional inequalities (1.26) and (1.27).

Finally in Section 5, by using functional inequalities on ~δ(t)\widetilde{\mathcal{F}}_{\delta}(t) and ~c(t)\widetilde{\mathcal{F}}_{c}(t), we prove the Gronwall’s inequality for the positive modulated energy

ddt+(t)+(t)+2,\displaystyle\frac{d}{dt}\mathcal{M}^{+}(t)\lesssim\mathcal{M}^{+}(t)+\hbar^{2},

where +(t)=(t)+2r(N,)\mathcal{M}^{+}(t)=\mathcal{M}(t)+2r(N,\hbar). Subsequently, with the quantitative convergence rate of the positive modulated energy, we further conclude the quantitative strong convergence rate of quantum mass and momentum densities, in which the δ\delta-type potential part plays an indispensable role in upgrading to the quantitative strong convergence.

2. The Time Evolution of the Modulated Energy

We consider the modulated energy in the quantum NN-body dynamics corresponding to the δ\delta-type and Coulomb potentials

(t):=\displaystyle\mathcal{M}(t):= 3N|(ix1u(t,x1))ψN,(t,XN)|2𝑑XN+δ(t)+c(t),\displaystyle\int_{\mathbb{R}^{3N}}|\left(i\hbar\nabla_{x_{1}}-u(t,x_{1})\right)\psi_{N,\hbar}(t,X_{N})|^{2}dX_{N}+\mathcal{F}_{\delta}(t)+\mathcal{F}_{c}(t),

where the δ\delta-type potential part is

(2.1) δ(t)=\displaystyle\mathcal{F}_{\delta}(t)= N1N3×3VN(x1x2)ρN,(2)(t,x1,x2)𝑑x1𝑑x2\displaystyle\frac{N-1}{N}\iint_{\mathbb{R}^{3}\times\mathbb{R}^{3}}V_{N}(x_{1}-x_{2})\rho_{N,\hbar}^{(2)}(t,x_{1},x_{2})dx_{1}dx_{2}
+b03ρ(t,x1)ρ(t,x1)𝑑x12b03ρ(t,x1)ρN,(1)(t,x1)𝑑x1,\displaystyle+b_{0}\int_{\mathbb{R}^{3}}\rho(t,x_{1})\rho(t,x_{1})dx_{1}-2b_{0}\int_{\mathbb{R}^{3}}\rho(t,x_{1})\rho_{N,\hbar}^{(1)}(t,x_{1})dx_{1},

and the Coulomb potential part is

(2.2) c(t)=\displaystyle\mathcal{F}_{c}(t)= 3×3Vc(x1x2)[N1NρN,(2)(t,x1,x2)\displaystyle\iint_{\mathbb{R}^{3}\times\mathbb{R}^{3}}V_{c}(x_{1}-x_{2})\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(t,x_{1},x_{2})\right.
+ρ(t,x1)ρ(t,x2)2ρ(t,x1)ρN,(1)(t,x2)]dx1dx2.\displaystyle\left.\quad\quad\quad+\rho(t,x_{1})\rho(t,x_{2})-2\rho(t,x_{1})\rho_{N,\hbar}^{(1)}(t,x_{2})\right]dx_{1}dx_{2}.

Here, we might as well assume that the coefficient κ=1\kappa=1, as the proof works the same for κ0\kappa\geq 0.

First, we need to derive a time evolution equation for (t)\mathcal{M}(t). The related quantities for ψN,\psi_{N,\hbar} are given as the following.

Lemma 2.1.

We have the following computations regarding ψN,\psi_{N,\hbar}:

(2.3) tρN,(1)+divJN,(1)=0,\displaystyle\partial_{t}\rho_{N,\hbar}^{(1)}+\operatorname{div}J_{N,\hbar}^{(1)}=0,
(2.4) tJN,(1)=22Re((Δx1ψN,¯)x1ψN,+ψN,¯x1Δx1ψN,)𝑑X2,N\displaystyle\partial_{t}J_{N,\hbar}^{(1)}=\frac{\hbar^{2}}{2}\int\operatorname{Re}\left((-\Delta_{x_{1}}\overline{\psi_{N,\hbar}})\nabla_{x_{1}}\psi_{N,\hbar}+\overline{\psi_{N,\hbar}}\nabla_{x_{1}}\Delta_{x_{1}}\psi_{N,\hbar}\right)dX_{2,N}
N1Nx1(VN+Vc)(x1x2)ρN,(2)(t,x1,x2)𝑑x2,\displaystyle\quad\quad\quad\quad-\frac{N-1}{N}\int\nabla_{x_{1}}(V_{N}+V_{c})(x_{1}-x_{2})\rho_{N,\hbar}^{(2)}(t,x_{1},x_{2})dx_{2},
(2.5) EN,(t)=EN,(0)E0,\displaystyle E_{N,\hbar}(t)=E_{N,\hbar}(0)\leq E_{0},

where X2,N=(x2,,xN)X_{2,N}=(x_{2},...,x_{N}) and the momentum density JN,(1)(t,x1)J_{N,\hbar}^{(1)}(t,x_{1}) and the energy EN,(t)E_{N,\hbar}(t) are defined by

(2.6) JN,(1)(t,x1)=\displaystyle J_{N,\hbar}^{(1)}(t,x_{1})= Im(x1γN,(1))(t,x1;x1)=Im(ψN,¯x1ψN,)(t,XN)𝑑X2,N,\displaystyle\operatorname{Im}\left(\hbar\nabla_{x_{1}}\gamma_{N,\hbar}^{(1)}\right)(t,x_{1};x_{1})=\hbar\int\operatorname{Im}(\overline{\psi_{N,\hbar}}\nabla_{x_{1}}\psi_{N,\hbar})(t,X_{N})dX_{2,N},
(2.7) EN,(t)=\displaystyle E_{N,\hbar}(t)= 1N(HN,+N)ψN,(t),ψN,(t).\displaystyle\frac{1}{N}\langle(H_{N,\hbar}+N)\psi_{N,\hbar}(t),\psi_{N,\hbar}(t)\rangle.
Proof.

As the mass and energy conservation laws are well-known, we omit the proof of (2.3) and (2.5). We provide the proof of the evolution (2.4) of the momentum density. From (2.6), we can write out

tJN,(1)=\displaystyle\partial_{t}J_{N,\hbar}^{(1)}= Im(tψN,¯x1ψN,+ψN,¯x1tψN,)𝑑X2,N\displaystyle\hbar\int\operatorname{Im}\left(\overline{\partial_{t}\psi_{N,\hbar}}\nabla_{x_{1}}\psi_{N,\hbar}+\overline{\psi_{N,\hbar}}\nabla_{x_{1}}\partial_{t}\psi_{N,\hbar}\right)dX_{2,N}
=\displaystyle= Im(iHN,ψN,¯x1ψN,iψN,¯x1HN,ψN,)𝑑X2,N\displaystyle\int\operatorname{Im}\left(i\overline{H_{N,\hbar}\psi_{N,\hbar}}\nabla_{x_{1}}\psi_{N,\hbar}-i\overline{\psi_{N,\hbar}}\nabla_{x_{1}}H_{N,\hbar}\psi_{N,\hbar}\right)dX_{2,N}
=\displaystyle= Re(HN,ψN,¯x1ψN,ψN,¯x1HN,ψN,)𝑑X2,N\displaystyle\int\operatorname{Re}\left(\overline{H_{N,\hbar}\psi_{N,\hbar}}\nabla_{x_{1}}\psi_{N,\hbar}-\overline{\psi_{N,\hbar}}\nabla_{x_{1}}H_{N,\hbar}\psi_{N,\hbar}\right)dX_{2,N}
=\displaystyle= IK+IV,\displaystyle I_{K}+I_{V},

where

IK=\displaystyle I_{K}= 22Re(i=1N(ΔxiψN,¯)x1ψN,ψN,¯x1i=1NΔxiψN,)𝑑X2,N,\displaystyle\frac{\hbar^{2}}{2}\int\operatorname{Re}\left(\sum_{i=1}^{N}(-\Delta_{x_{i}}\overline{\psi_{N,\hbar}})\nabla_{x_{1}}\psi_{N,\hbar}-\overline{\psi_{N,\hbar}}\nabla_{x_{1}}\sum_{i=1}^{N}-\Delta_{x_{i}}\psi_{N,\hbar}\right)dX_{2,N},

and

IV=\displaystyle I_{V}= Re(1Ni<jN(VN+Vc)(xixj)ψN,¯x1ψN,)𝑑X2,N\displaystyle\int\operatorname{Re}\left(\frac{1}{N}\sum_{i<j}^{N}(V_{N}+V_{c})(x_{i}-x_{j})\overline{\psi_{N,\hbar}}\nabla_{x_{1}}\psi_{N,\hbar}\right)dX_{2,N}
Re(ψN,¯x11Ni<jN(VN+Vc)(xixj)ψN,)𝑑X2,N.\displaystyle-\int\operatorname{Re}\left(\overline{\psi_{N,\hbar}}\nabla_{x_{1}}\frac{1}{N}\sum_{i<j}^{N}(V_{N}+V_{c})(x_{i}-x_{j})\psi_{N,\hbar}\right)dX_{2,N}.

For IKI_{K}, we use integration by parts with Δxi\Delta_{x_{i}} to obtain

IK=22Re((Δx1ψN,¯)x1ψN,+ψN,¯x1Δx1ψN,)𝑑X2,N,\displaystyle I_{K}=\frac{\hbar^{2}}{2}\int\operatorname{Re}\left((-\Delta_{x_{1}}\overline{\psi_{N,\hbar}})\nabla_{x_{1}}\psi_{N,\hbar}+\overline{\psi_{N,\hbar}}\nabla_{x_{1}}\Delta_{x_{1}}\psi_{N,\hbar}\right)dX_{2,N},

where the other ii-summands vanish when i2i\geq 2.

For IVI_{V}, we note that the ii-summands also vanish when i2i\geq 2 and hence have

IV=\displaystyle I_{V}= Re(1Nj=2N(VN+Vc)(x1xj)ψN,¯x1ψN,)𝑑X2,N\displaystyle\int\operatorname{Re}\left(\frac{1}{N}\sum_{j=2}^{N}(V_{N}+V_{c})(x_{1}-x_{j})\overline{\psi_{N,\hbar}}\nabla_{x_{1}}\psi_{N,\hbar}\right)dX_{2,N}
Re(ψN,¯x11Nj=2N(VN+Vc)(x1xj)ψN,)𝑑X2,N\displaystyle-\int\operatorname{Re}\left(\overline{\psi_{N,\hbar}}\nabla_{x_{1}}\frac{1}{N}\sum_{j=2}^{N}(V_{N}+V_{c})(x_{1}-x_{j})\psi_{N,\hbar}\right)dX_{2,N}
=\displaystyle= N1N|ψN,|2x1(VN+Vc)(x1x2)𝑑X2,N\displaystyle-\frac{N-1}{N}\int|\psi_{N,\hbar}|^{2}\nabla_{x_{1}}(V_{N}+V_{c})(x_{1}-x_{2})dX_{2,N}
=\displaystyle= N1Nx1(VN+Vc)(x1x2)ρN,(2)(t,x1,x2)𝑑x2.\displaystyle-\frac{N-1}{N}\int\nabla_{x_{1}}(V_{N}+V_{c})(x_{1}-x_{2})\rho_{N,\hbar}^{(2)}(t,x_{1},x_{2})dx_{2}.

This completes the proof of (2.4). ∎

Now, we derive the time evolution of (t)\mathcal{M}(t).

Proposition 2.2.

Let (t)\mathcal{M}(t) be defined in (2)\eqref{equ:modulated energy}, there holds

(2.8) ddt(t)\displaystyle\frac{d}{dt}\mathcal{M}(t)
=\displaystyle= j,k=133N(juk+kuj)(ijψN,ujψN,)(ikψN,ukψN,)¯𝑑XN\displaystyle-\sum_{j,k=1}^{3}\int_{\mathbb{R}^{3N}}\left(\partial_{j}u^{k}+\partial_{k}u^{j}\right)(-i\hbar\partial_{j}\psi_{N,\hbar}-u^{j}\psi_{N,\hbar})\overline{(-i\hbar\partial_{k}\psi_{N,\hbar}-u^{k}\psi_{N,\hbar})}dX_{N}
+223Δ(divu)(t,x1)ρN,(1)(t,x1)𝑑x1+~δ(t)+~c(t),\displaystyle+\frac{\hbar^{2}}{2}\int_{\mathbb{R}^{3}}\Delta(\operatorname{div}u)(t,x_{1})\rho_{N,\hbar}^{(1)}(t,x_{1})dx_{1}+\widetilde{\mathcal{F}}_{\delta}(t)+\widetilde{\mathcal{F}}_{c}(t),

where we used the notations u=(u1,u2,u3)u=(u^{1},u^{2},u^{3}), x1=(x11,x12,x13)3x_{1}=(x_{1}^{1},x_{1}^{2},x_{1}^{3})\in\mathbb{R}^{3} and j=x1j\partial_{j}=\partial_{x_{1}^{j}}. Here, the δ\delta-type potential contribution part is

(2.9) ~δ(t)=\displaystyle\widetilde{\mathcal{F}}_{\delta}(t)= N1N(u(t,x1)u(t,x2))VN(x1x2)ρN,(2)(t,x1,x2)𝑑x1𝑑x2\displaystyle\frac{N-1}{N}\int(u(t,x_{1})-u(t,x_{2}))\nabla V_{N}(x_{1}-x_{2})\rho_{N,\hbar}^{(2)}(t,x_{1},x_{2})dx_{1}dx_{2}
b0divu(t,x1)ρ(t,x1)[ρ(t,x1)2ρN,(1)(t,x1)]𝑑x1,\displaystyle-b_{0}\int\operatorname{div}u(t,x_{1})\rho(t,x_{1})\left[\rho(t,x_{1})-2\rho_{N,\hbar}^{(1)}(t,x_{1})\right]dx_{1},

and the Coulomb potential contribution part is

(2.10) ~c(t)=\displaystyle\widetilde{\mathcal{F}}_{c}(t)= (u(t,x1)u(t,x2))Vc(x1x2)\displaystyle\int(u(t,x_{1})-u(t,x_{2}))\nabla V_{c}(x_{1}-x_{2})
[N1NρN,(2)(t,x1,x2)+ρ(t,x1)ρ(t,x2)2ρ(t,x1)ρN,(1)(t,x2)]dx1dx2.\displaystyle\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(t,x_{1},x_{2})+\rho(t,x_{1})\rho(t,x_{2})-2\rho(t,x_{1})\rho_{N,\hbar}^{(1)}(t,x_{2})\right]dx_{1}dx_{2}.
Proof.

We decompose the modulated energy into five parts to do the calculation.

1(t)\displaystyle\mathcal{M}_{1}(t) =3N|ix1ψN,(t,XN)|2𝑑XN\displaystyle=\int_{\mathbb{R}^{3N}}|i\hbar\nabla_{x_{1}}\psi_{N,\hbar}(t,X_{N})|^{2}dX_{N}
+N1N3×3(VN+Vc)(x1x2)ρN,(2)(t,x1,x2)𝑑x1𝑑x2,\displaystyle+\frac{N-1}{N}\iint_{\mathbb{R}^{3}\times\mathbb{R}^{3}}(V_{N}+V_{c})(x_{1}-x_{2})\rho_{N,\hbar}^{(2)}(t,x_{1},x_{2})dx_{1}dx_{2},
2(t)\displaystyle\mathcal{M}_{2}(t) =i3Nu(t,x1)(ψN,¯x1ψN,ψN,x1ψN,¯)(t,XN)𝑑XN,\displaystyle=i\hbar\int_{\mathbb{R}^{3N}}u(t,x_{1})(\overline{\psi_{N,\hbar}}\nabla_{x_{1}}\psi_{N,\hbar}-\psi_{N,\hbar}\nabla_{x_{1}}\overline{\psi_{N,\hbar}})(t,X_{N})dX_{N},
3(t)\displaystyle\mathcal{M}_{3}(t) =3N|u(t,x1)ψN,(t,XN)|2𝑑XN,\displaystyle=\int_{\mathbb{R}^{3N}}|u(t,x_{1})\psi_{N,\hbar}(t,X_{N})|^{2}dX_{N},
4(t)\displaystyle\mathcal{M}_{4}(t) =b03ρ(t,x1)ρ(t,x1)𝑑x1+3ρ(t,x1)(Vcρ)(t,x1)𝑑x1,\displaystyle=b_{0}\int_{\mathbb{R}^{3}}\rho(t,x_{1})\rho(t,x_{1})dx_{1}+\int_{\mathbb{R}^{3}}\rho(t,x_{1})(V_{c}*\rho)(t,x_{1})dx_{1},
5(t)\displaystyle\mathcal{M}_{5}(t) =2b03ρ(t,x1)ρN,(1)(t,x1)𝑑x12(Vcρ)(t,x1)ρN,(1)(t,x1)𝑑x1.\displaystyle=-2b_{0}\int_{\mathbb{R}^{3}}\rho(t,x_{1})\rho_{N,\hbar}^{(1)}(t,x_{1})dx_{1}-2\int(V_{c}*\rho)(t,x_{1})\rho_{N,\hbar}^{(1)}(t,x_{1})dx_{1}.

For 1(t)\mathcal{M}_{1}(t), by the symmetry of the wave function ψN,(t)\psi_{N,\hbar}(t), we obtain

1(t)=\displaystyle\mathcal{M}_{1}(t)= 3N(2Δx1ψN,+N1N(VN+Vc)(x1x2)ψN,)ψN,¯𝑑XN\displaystyle\int_{\mathbb{R}^{3N}}\left(-\hbar^{2}\Delta_{x_{1}}\psi_{N,\hbar}+\frac{N-1}{N}(V_{N}+V_{c})(x_{1}-x_{2})\psi_{N,\hbar}\right)\overline{\psi_{N,\hbar}}dX_{N}
=\displaystyle= 2NHN,ψN,(t),ψN,(t)\displaystyle\frac{2}{N}\langle H_{N,\hbar}\psi_{N,\hbar}(t),\psi_{N,\hbar}(t)\rangle
=\displaystyle= 2NHN,ψN,(0),ψN,(0),\displaystyle\frac{2}{N}\langle H_{N,\hbar}\psi_{N,\hbar}(0),\psi_{N,\hbar}(0)\rangle,

where in the last equality we have used the conservation of energy. Therefore, we have that

ddt1(t)=0.\frac{d}{dt}\mathcal{M}_{1}(t)=0.

For 2(t)\mathcal{M}_{2}(t), from the definition of JN,(1)(t,x)J_{N,\hbar}^{(1)}(t,x) in (2.6), we note that

2(t)=\displaystyle\mathcal{M}_{2}(t)= i3Nu(t,x1)(ψN,¯x1ψN,ψN,x1ψN,¯)(t,XN)𝑑XN\displaystyle i\hbar\int_{\mathbb{R}^{3N}}u(t,x_{1})(\overline{\psi_{N,\hbar}}\nabla_{x_{1}}\psi_{N,\hbar}-\psi_{N,\hbar}\nabla_{x_{1}}\overline{\psi_{N,\hbar}})(t,X_{N})dX_{N}
=\displaystyle= 2u(t,x1)Im(ψN,¯x1ψN,)(t,XN)𝑑X2,N𝑑x1\displaystyle-2\int u(t,x_{1})\hbar\int\operatorname{Im}(\overline{\psi_{N,\hbar}}\nabla_{x_{1}}\psi_{N,\hbar})(t,X_{N})dX_{2,N}dx_{1}
=\displaystyle= 2u(t,x1)JN,(1)(t,x1)𝑑x1.\displaystyle-2\int u(t,x_{1})J_{N,\hbar}^{(1)}(t,x_{1})dx_{1}.

Thus, we have

(2.11) ddt2(t)=\displaystyle\frac{d}{dt}\mathcal{M}_{2}(t)= 2tu(t,x1)JN,(1)(t,x1)dx12u(t,x1)tJN,(1)(t,x1)dx1.\displaystyle-2\int\partial_{t}u(t,x_{1})J_{N,\hbar}^{(1)}(t,x_{1})dx_{1}-2\int u(t,x_{1})\partial_{t}J_{N,\hbar}^{(1)}(t,x_{1})dx_{1}.

For the second term on the r.h.s of (2.11), by (2.4) we obtain

2u(t,x1)tJN,(1)(t,x1)dx1\displaystyle-2\int u(t,x_{1})\partial_{t}J_{N,\hbar}^{(1)}(t,x_{1})dx_{1}
=\displaystyle= 2u(t,x1)Re((Δx1ψN,¯)x1ψN,+ψN,¯x1Δx1ψN,)𝑑XN\displaystyle-\hbar^{2}\int u(t,x_{1})\operatorname{Re}\left((-\Delta_{x_{1}}\overline{\psi_{N,\hbar}})\nabla_{x_{1}}\psi_{N,\hbar}+\overline{\psi_{N,\hbar}}\nabla_{x_{1}}\Delta_{x_{1}}\psi_{N,\hbar}\right)dX_{N}
+2(N1)Nu(t,x1)x1(VN+Vc)(x1x2)ρN,(2)(t,x1,x2)𝑑x1𝑑x2\displaystyle+\frac{2(N-1)}{N}\int u(t,x_{1})\nabla_{x_{1}}(V_{N}+V_{c})(x_{1}-x_{2})\rho_{N,\hbar}^{(2)}(t,x_{1},x_{2})dx_{1}dx_{2}
(2.12) =\displaystyle= 2u(t,x1)Re((Δx1ψN,¯)x1ψN,+ψN,¯x1Δx1ψN,)𝑑XN\displaystyle-\hbar^{2}\int u(t,x_{1})\operatorname{Re}\left((-\Delta_{x_{1}}\overline{\psi_{N,\hbar}})\nabla_{x_{1}}\psi_{N,\hbar}+\overline{\psi_{N,\hbar}}\nabla_{x_{1}}\Delta_{x_{1}}\psi_{N,\hbar}\right)dX_{N}
+(N1)N(u(t,x1)u(t,x2))x1(VN+Vc)(x1x2)ρN,(2)(t,x1,x2)𝑑x1𝑑x2,\displaystyle+\frac{(N-1)}{N}\int(u(t,x_{1})-u(t,x_{2}))\nabla_{x_{1}}(V_{N}+V_{c})(x_{1}-x_{2})\rho_{N,\hbar}^{(2)}(t,x_{1},x_{2})dx_{1}dx_{2},

where in the last equality we used the antisymmetry of (VN+Vc)\nabla(V_{N}+V_{c}).

Next, we deal with (2.12). By integration by parts, we obtain

2u(t,x1)Re((Δx1ψN,¯)x1ψN,+ψN,¯x1Δx1ψN,)𝑑XN\displaystyle-\hbar^{2}\int u(t,x_{1})\operatorname{Re}\left((-\Delta_{x_{1}}\overline{\psi_{N,\hbar}})\nabla_{x_{1}}\psi_{N,\hbar}+\overline{\psi_{N,\hbar}}\nabla_{x_{1}}\Delta_{x_{1}}\psi_{N,\hbar}\right)dX_{N}
=\displaystyle= 2Re2u(t,x1)(Δx1ψN,¯)x1ψN,(divu)ψN,¯Δx1ψN,dXN\displaystyle-\hbar^{2}\operatorname{Re}\int 2u(t,x_{1})(-\Delta_{x_{1}}\overline{\psi_{N,\hbar}})\nabla_{x_{1}}\psi_{N,\hbar}-(\operatorname{div}u)\overline{\psi_{N,\hbar}}\Delta_{x_{1}}\psi_{N,\hbar}dX_{N}
=\displaystyle= 2j,k=13Re2kuj(kψN,¯)jψN,+2uj(kψN,¯)kjψN,dXN\displaystyle-\hbar^{2}\sum_{j,k=1}^{3}\operatorname{Re}\int 2\partial_{k}u^{j}(\partial_{k}\overline{\psi_{N,\hbar}})\partial_{j}\psi_{N,\hbar}+2u^{j}(\partial_{k}\overline{\psi_{N,\hbar}})\partial_{k}\partial_{j}\psi_{N,\hbar}dX_{N}
+2Re(divu)ψN,¯Δx1ψN,𝑑XN\displaystyle+\hbar^{2}\operatorname{Re}\int(\operatorname{div}u)\overline{\psi_{N,\hbar}}\Delta_{x_{1}}\psi_{N,\hbar}dX_{N}
=\displaystyle= 2j,k=13(juk+kuj)jψN,kψN,¯dXN\displaystyle\hbar^{2}\sum_{j,k=1}^{3}\int\left(\partial_{j}u^{k}+\partial_{k}u^{j}\right)\partial_{j}\psi_{N,\hbar}\partial_{k}\overline{\psi_{N,\hbar}}dX_{N}
(2.13) 2j,k=13Re2uj(kψN,¯)kjψN,dXN+2Re(divu)ψN,¯Δx1ψN,𝑑XN,\displaystyle-\hbar^{2}\sum_{j,k=1}^{3}\operatorname{Re}\int 2u^{j}(\partial_{k}\overline{\psi_{N,\hbar}})\partial_{k}\partial_{j}\psi_{N,\hbar}dX_{N}+\hbar^{2}\operatorname{Re}\int(\operatorname{div}u)\overline{\psi_{N,\hbar}}\Delta_{x_{1}}\psi_{N,\hbar}dX_{N},

where we used the notations u=(u1,u2,u3)u=(u^{1},u^{2},u^{3}), x1=(x11,x12,x13)3x_{1}=(x_{1}^{1},x_{1}^{2},x_{1}^{3})\in\mathbb{R}^{3} and j=x1j\partial_{j}=\partial_{x_{1}^{j}}.

Using again integration by parts on the two terms of (2.13) gives

(2.14) 2j,k=13Re2uj(kψN,¯)kjψN,dXN+2Re(divu)ψN,¯Δx1ψN,𝑑XN\displaystyle-\hbar^{2}\sum_{j,k=1}^{3}\operatorname{Re}\int 2u^{j}(\partial_{k}\overline{\psi_{N,\hbar}})\partial_{k}\partial_{j}\psi_{N,\hbar}dX_{N}+\hbar^{2}\operatorname{Re}\int(\operatorname{div}u)\overline{\psi_{N,\hbar}}\Delta_{x_{1}}\psi_{N,\hbar}dX_{N}
=\displaystyle= 2Redivu(|x1ψN,|2+ψN,¯Δx1ψN,)𝑑XN\displaystyle\hbar^{2}\operatorname{Re}\int\operatorname{div}u\left(|\nabla_{x_{1}}\psi_{N,\hbar}|^{2}+\overline{\psi_{N,\hbar}}\Delta_{x_{1}}\psi_{N,\hbar}\right)dX_{N}
=\displaystyle= 22Redivu(Δx1|ψN,|2)𝑑XN\displaystyle\frac{\hbar^{2}}{2}\operatorname{Re}\int\operatorname{div}u(\Delta_{x_{1}}|\psi_{N,\hbar}|^{2})dX_{N}
=\displaystyle= 22(Δdivu)(t,x1)ρN,(1)(t,x1)𝑑x1.\displaystyle\frac{\hbar^{2}}{2}\int(\Delta\operatorname{div}u)(t,x_{1})\rho_{N,\hbar}^{(1)}(t,x_{1})dx_{1}.

Combining estimates (2.11)–(2.14), we provide

ddt2(t)\displaystyle\frac{d}{dt}\mathcal{M}_{2}(t)
=\displaystyle= 2tu,JN,(1)+2j,k=13(juk+kuj)jψN,kψN,¯dXN\displaystyle-2\langle\partial_{t}u,J_{N,\hbar}^{(1)}\rangle+\hbar^{2}\sum_{j,k=1}^{3}\int\left(\partial_{j}u^{k}+\partial_{k}u^{j}\right)\partial_{j}\psi_{N,\hbar}\partial_{k}\overline{\psi_{N,\hbar}}dX_{N}
+22(Δdivu)(t,x1)ρN,(1)(t,x1)𝑑x1\displaystyle+\frac{\hbar^{2}}{2}\int(\Delta\operatorname{div}u)(t,x_{1})\rho_{N,\hbar}^{(1)}(t,x_{1})dx_{1}
+(N1)N(u(t,x1)u(t,x2))x1(VN+Vc)(x1x2)ρN,(2)(t,x1,x2)𝑑x1𝑑x2.\displaystyle+\frac{(N-1)}{N}\int(u(t,x_{1})-u(t,x_{2}))\cdot\nabla_{x_{1}}(V_{N}+V_{c})(x_{1}-x_{2})\rho_{N,\hbar}^{(2)}(t,x_{1},x_{2})dx_{1}dx_{2}.

For 3(t)\mathcal{M}_{3}(t), by the Euler-Poisson equation (1.8) and the mass conservation law (2.3), we obtain

ddt3(t)\displaystyle\frac{d}{dt}\mathcal{M}_{3}(t)
=\displaystyle= ddt|u(t,x1)|2ρN,(1)(t,x1)𝑑x1\displaystyle\frac{d}{dt}\int|u(t,x_{1})|^{2}\rho_{N,\hbar}^{(1)}(t,x_{1})dx_{1}
=\displaystyle= 2u(t,x1)tu(t,x1)ρN,(1)(t,x1)dx1+|u(t,x1)|2tρN,(1)(t,x1)dx1\displaystyle\int 2u(t,x_{1})\cdot\partial_{t}u(t,x_{1})\rho_{N,\hbar}^{(1)}(t,x_{1})dx_{1}+\int|u(t,x_{1})|^{2}\partial_{t}\rho_{N,\hbar}^{(1)}(t,x_{1})dx_{1}
=\displaystyle= 2u(t,x1)(uu+b0ρ+Vcρ)ρN,(1)(t,x1)𝑑x1\displaystyle-2\int u(t,x_{1})\cdot\left(u\cdot\nabla u+b_{0}\nabla\rho+\nabla V_{c}*\rho\right)\rho_{N,\hbar}^{(1)}(t,x_{1})dx_{1}
+(|u(t,x1)|2)JN,(1)(t,x1)𝑑x1.\displaystyle+\int\nabla\left(|u(t,x_{1})|^{2}\right)J_{N,\hbar}^{(1)}(t,x_{1})dx_{1}.

Expanding it gives

ddt3(t)=\displaystyle\frac{d}{dt}\mathcal{M}_{3}(t)= 2j,k=13ukujjukρN,(1)(t,x1)dx12b0ρN,(1),uρ2ρN,(1),uVcρ\displaystyle-2\sum_{j,k=1}^{3}\int u^{k}u^{j}\partial_{j}u^{k}\rho_{N,\hbar}^{(1)}(t,x_{1})dx_{1}-2b_{0}\langle\rho_{N,\hbar}^{(1)},u\cdot\nabla\rho\rangle-2\langle\rho_{N,\hbar}^{(1)},u\cdot\nabla V_{c}*\rho\rangle
+2j,k=13ukjukJN,(1)(t,x1)dx1.\displaystyle+2\sum_{j,k=1}^{3}\int u^{k}\partial_{j}u^{k}J_{N,\hbar}^{(1)}(t,x_{1})dx_{1}.

For 4(t)\mathcal{M}_{4}(t), plugging in the Euler-Poisson equation (1.8), we have

ddt4(t)=\displaystyle\frac{d}{dt}\mathcal{M}_{4}(t)= b0ddt3ρ(t,x1)ρ(t,x1)𝑑x1+ddt3ρ(t,x1)(Vcρ)(t,x1)𝑑x1,\displaystyle b_{0}\frac{d}{dt}\int_{\mathbb{R}^{3}}\rho(t,x_{1})\rho(t,x_{1})dx_{1}+\frac{d}{dt}\int_{\mathbb{R}^{3}}\rho(t,x_{1})(V_{c}*\rho)(t,x_{1})dx_{1},
=\displaystyle= 2b0ρ,uρ+2ρ,uVcρ.\displaystyle 2b_{0}\langle\rho,u\cdot\nabla\rho\rangle+2\langle\rho,u\cdot\nabla V_{c}*\rho\rangle.

For 5(t)\mathcal{M}_{5}(t), similarly we get to

ddt5(t)\displaystyle\frac{d}{dt}\mathcal{M}_{5}(t)
=\displaystyle= 2b0ddt3ρ(t,x1)ρN,(1)(t,x1)𝑑x12ddt(Vcρ)(t,x1)ρN,(1)(t,x1)𝑑x1\displaystyle-2b_{0}\frac{d}{dt}\int_{\mathbb{R}^{3}}\rho(t,x_{1})\rho_{N,\hbar}^{(1)}(t,x_{1})dx_{1}-2\frac{d}{dt}\int(V_{c}*\rho)(t,x_{1})\rho_{N,\hbar}^{(1)}(t,x_{1})dx_{1}
=\displaystyle= 2b0(tρ,ρN,(1)+ρ,tρN,(1))2(tρ,VcρN,(1)+Vcρ,tρN,(1)).\displaystyle-2b_{0}\left(\langle\partial_{t}\rho,\rho_{N,\hbar}^{(1)}\rangle+\langle\rho,\partial_{t}\rho_{N,\hbar}^{(1)}\rangle\right)-2\left(\langle\partial_{t}\rho,V_{c}*\rho_{N,\hbar}^{(1)}\rangle+\langle V_{c}*\rho,\partial_{t}\rho_{N,\hbar}^{(1)}\rangle\right).

Plugging in the Euler-Poisson equation (1.8) and the mass conservation law (2.3), we have

ddt5(t)\displaystyle\frac{d}{dt}\mathcal{M}_{5}(t)
=\displaystyle= 2b0(div(ρu),ρN,(1)+ρ,divJN,(1))+2(div(ρu),VcρN,(1)+Vcρ,divJN,(1))\displaystyle 2b_{0}\left(\langle\operatorname{div}(\rho u),\rho_{N,\hbar}^{(1)}\rangle+\langle\rho,\operatorname{div}J_{N,\hbar}^{(1)}\rangle\right)+2\left(\langle\operatorname{div}(\rho u),V_{c}*\rho_{N,\hbar}^{(1)}\rangle+\langle V_{c}*\rho,\operatorname{div}J_{N,\hbar}^{(1)}\rangle\right)
=\displaystyle= 2b0ρ,uρN,(1)2ρ,uVcρN,(1)2b0ρ,JN,(1)2Vcρ,JN,(1).\displaystyle-2b_{0}\langle\rho,u\cdot\nabla\rho_{N,\hbar}^{(1)}\rangle-2\langle\rho,u\cdot\nabla V_{c}*\rho_{N,\hbar}^{(1)}\rangle-2b_{0}\langle\nabla\rho,J_{N,\hbar}^{(1)}\rangle-2\langle\nabla V_{c}*\rho,J_{N,\hbar}^{(1)}\rangle.

Therefore, putting the five terms together, we reach

ddt(t)=\displaystyle\frac{d}{dt}\mathcal{M}(t)= ddt1(t)+ddt2(t)+ddt3(t)+ddt4(t)+ddt5(t)\displaystyle\frac{d}{dt}\mathcal{M}_{1}(t)+\frac{d}{dt}\mathcal{M}_{2}(t)+\frac{d}{dt}\mathcal{M}_{3}(t)+\frac{d}{dt}\mathcal{M}_{4}(t)+\frac{d}{dt}\mathcal{M}_{5}(t)
(2.15) =\displaystyle= 2tu,JN,(1)2j,k=133N(juk+kuj)jψN,kψN,¯dXN\displaystyle-2\langle\partial_{t}u,J_{N,\hbar}^{(1)}\rangle-\hbar^{2}\sum_{j,k=1}^{3}\int_{\mathbb{R}^{3N}}\left(\partial_{j}u^{k}+\partial_{k}u^{j}\right)\partial_{j}\psi_{N,\hbar}\partial_{k}\overline{\psi_{N,\hbar}}dX_{N}
+223Δ(divu)(t,x1)ρN,(1)(t,x1)𝑑x1\displaystyle+\frac{\hbar^{2}}{2}\int_{\mathbb{R}^{3}}\Delta(\operatorname{div}u)(t,x_{1})\rho_{N,\hbar}^{(1)}(t,x_{1})dx_{1}
+N1N(u(t,x1)u(t,x2))x1(VN+Vc)(x1x2)ρN,(2)(t,x1,x2)𝑑x1𝑑x2\displaystyle+\frac{N-1}{N}\int(u(t,x_{1})-u(t,x_{2}))\cdot\nabla_{x_{1}}(V_{N}+V_{c})(x_{1}-x_{2})\rho_{N,\hbar}^{(2)}(t,x_{1},x_{2})dx_{1}dx_{2}
(2.16) 2j,k=13ukujjukρN,(1)(t,x1)dx12b0ρN,(1),uρ2ρN,(1),uVcρ\displaystyle-2\sum_{j,k=1}^{3}\int u^{k}u^{j}\partial_{j}u^{k}\rho_{N,\hbar}^{(1)}(t,x_{1})dx_{1}-2b_{0}\langle\rho_{N,\hbar}^{(1)},u\cdot\nabla\rho\rangle-2\langle\rho_{N,\hbar}^{(1)},u\cdot\nabla V_{c}*\rho\rangle
+(|u|2),JN,(1)+2b0ρ,uρ+2ρ,uVcρ\displaystyle+\langle\nabla(|u|^{2}),J_{N,\hbar}^{(1)}\rangle+2b_{0}\langle\rho,u\cdot\nabla\rho\rangle+2\langle\rho,u\cdot\nabla V_{c}*\rho\rangle
2b0ρ,uρN,(1)2ρ,uVcρN,(1)2b0ρ,JN,(1)2Vcρ,JN,(1).\displaystyle-2b_{0}\langle\rho,u\cdot\nabla\rho_{N,\hbar}^{(1)}\rangle-2\langle\rho,u\cdot\nabla V_{c}*\rho_{N,\hbar}^{(1)}\rangle-2b_{0}\langle\nabla\rho,J_{N,\hbar}^{(1)}\rangle-2\langle\nabla V_{c}*\rho,J_{N,\hbar}^{(1)}\rangle.

From the above equation, we collect the δ\delta-type potential contribution part ~δ(t)\widetilde{\mathcal{F}}_{\delta}(t) in (2.8) from

N1N(u(t,x1)u(t,x2))x1VN(x1x2)ρN,(2)(t,x1,x2)𝑑x1𝑑x2\displaystyle\frac{N-1}{N}\int(u(t,x_{1})-u(t,x_{2}))\cdot\nabla_{x_{1}}V_{N}(x_{1}-x_{2})\rho_{N,\hbar}^{(2)}(t,x_{1},x_{2})dx_{1}dx_{2}
2b0ρN,(1),uρ2b0ρ,uρN,(1)+2b0ρ,uρ\displaystyle-2b_{0}\langle\rho_{N,\hbar}^{(1)},u\cdot\nabla\rho\rangle-2b_{0}\langle\rho,u\cdot\nabla\rho_{N,\hbar}^{(1)}\rangle+2b_{0}\langle\rho,u\cdot\nabla\rho\rangle
=\displaystyle= N1N(u(t,x1)u(t,x2))x1VN(x1x2)ρN,(2)(t,x1,x2)𝑑x1𝑑x2\displaystyle\frac{N-1}{N}\int(u(t,x_{1})-u(t,x_{2}))\cdot\nabla_{x_{1}}V_{N}(x_{1}-x_{2})\rho_{N,\hbar}^{(2)}(t,x_{1},x_{2})dx_{1}dx_{2}
b0divu(t,x1)ρ(t,x1)[ρ(t,x1)2ρN,(1)(t,x1)]𝑑x1.\displaystyle-b_{0}\int\operatorname{div}u(t,x_{1})\rho(t,x_{1})\left[\rho(t,x_{1})-2\rho_{N,\hbar}^{(1)}(t,x_{1})\right]dx_{1}.

and the Coulomb potential contribution part ~c(t)\widetilde{\mathcal{F}}_{c}(t) in (2.8) from

N1N(u(t,x1)u(t,x2))x1Vc(x1x2)ρN,(2)(t,x1,x2)𝑑x1𝑑x2\displaystyle\frac{N-1}{N}\int(u(t,x_{1})-u(t,x_{2}))\cdot\nabla_{x_{1}}V_{c}(x_{1}-x_{2})\rho_{N,\hbar}^{(2)}(t,x_{1},x_{2})dx_{1}dx_{2}
2ρN,(1),uVcρ2ρ,uVcρN,(1)+2ρ,uVcρ\displaystyle-2\langle\rho_{N,\hbar}^{(1)},u\cdot\nabla V_{c}*\rho\rangle-2\langle\rho,u\cdot\nabla V_{c}*\rho_{N,\hbar}^{(1)}\rangle+2\langle\rho,u\cdot\nabla V_{c}*\rho\rangle
=\displaystyle= (u(t,x1)u(t,x2))Vc(x1x2)\displaystyle\int(u(t,x_{1})-u(t,x_{2}))\nabla V_{c}(x_{1}-x_{2})
[N1NρN,(2)(t,x1,x2)+ρ(t,x1)ρ(t,x2)2ρ(t,x1)ρN,(1)(t,x2)]dx1dx2,\displaystyle\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(t,x_{1},x_{2})+\rho(t,x_{1})\rho(t,x_{2})-2\rho(t,x_{1})\rho_{N,\hbar}^{(1)}(t,x_{2})\right]dx_{1}dx_{2},

where in the last equality we used the antisymmetry of Vc\nabla V_{c}.

As for the first term in (2.8), we use the Euler-Poisson equation (1.8) to combine the terms taking the form of ,JN,(1)\langle\bullet,J_{N,\hbar}^{(1)}\rangle

2tu,JN,(1)2b0ρ,JN,(1)2Vcρ,JN,(1)+(|u|2),JN,(1)\displaystyle-2\langle\partial_{t}u,J_{N,\hbar}^{(1)}\rangle-2b_{0}\langle\nabla\rho,J_{N,\hbar}^{(1)}\rangle-2\langle\nabla V_{c}*\rho,J_{N,\hbar}^{(1)}\rangle+\langle\nabla(|u|^{2}),J_{N,\hbar}^{(1)}\rangle
=\displaystyle= 2uu,JN,(1)+2(uu),JN,(1)\displaystyle 2\langle u\cdot\nabla u,J_{N,\hbar}^{(1)}\rangle+2\langle\nabla\cdot(u\otimes u),J_{N,\hbar}^{(1)}\rangle
(2.17) =\displaystyle= ij,k=133N(juk+kuj)uj(ψN,kψN,¯ψN,¯kψN,)𝑑XN.\displaystyle i\hbar\sum_{j,k=1}^{3}\int_{\mathbb{R}^{3N}}\left(\partial_{j}u^{k}+\partial_{k}u^{j}\right)u^{j}(\psi_{N,\hbar}\partial_{k}\overline{\psi_{N,\hbar}}-\overline{\psi_{N,\hbar}}\partial_{k}\psi_{N,\hbar})dX_{N}.

If we rewrite the first term on the right hand side of (2.8)

j,k=13(juk+kuj)(ijψN,ujψN,)(ikψN,ukψN,)¯𝑑XN\displaystyle-\sum_{j,k=1}^{3}\int\left(\partial_{j}u^{k}+\partial_{k}u^{j}\right)(-i\hbar\partial_{j}\psi_{N,\hbar}-u^{j}\psi_{N,\hbar})\overline{(-i\hbar\partial_{k}\psi_{N,\hbar}-u^{k}\psi_{N,\hbar})}dX_{N}
=\displaystyle= 2j,k=13(juk+kuj)jψN,kψN,¯dXN\displaystyle-\hbar^{2}\sum_{j,k=1}^{3}\int\left(\partial_{j}u^{k}+\partial_{k}u^{j}\right)\partial_{j}\psi_{N,\hbar}\partial_{k}\overline{\psi_{N,\hbar}}dX_{N}
2j,k=13ukujjukρN,(1)(t,x1)dx1\displaystyle-2\sum_{j,k=1}^{3}\int u^{k}u^{j}\partial_{j}u^{k}\rho_{N,\hbar}^{(1)}(t,x_{1})dx_{1}
+ij,k=133N(juk+kuj)uj(ψN,kψN,¯ψN,¯kψN,)𝑑XN,\displaystyle+i\hbar\sum_{j,k=1}^{3}\int_{\mathbb{R}^{3N}}\left(\partial_{j}u^{k}+\partial_{k}u^{j}\right)u^{j}(\psi_{N,\hbar}\partial_{k}\overline{\psi_{N,\hbar}}-\overline{\psi_{N,\hbar}}\partial_{k}\psi_{N,\hbar})dX_{N},

these are the j,k3\sum_{j,k}^{3} terms in (2.15),(2.16) and (2.17). Therefore, we arrive at equation (2.8) and complete the proof.

3. (HN,)2(H_{N,\hbar})^{2} Energy Estimate Using Singular Correlation Structure

As mentioned in the preliminary reduction step in the outline, Section 1.3, the two-body energy estimate is crucial for the analysis of the δ\delta-type potential parts δ(t)\mathcal{F}_{\delta}(t) and ~δ(t)\widetilde{\mathcal{F}}_{\delta}(t). The main difficulty is the singularities simultaneously from the Coulomb potential, from the direct δ\delta-potential in the NN\to\infty limit, and from the interparticle singular correlation structure.

Recall the zero-energy scattering equation

(3.1) {(2Δ+1NVN(x))(1wN,(x))=0,lim|x|wN,(x)=0.\left\{\begin{aligned} &\left(-\hbar^{2}\Delta+\frac{1}{N}V_{N}(x)\right)(1-w_{N,\hbar}(x))=0,\\ &\lim_{|x|\to\infty}w_{N,\hbar}(x)=0.\end{aligned}\right.

and our target estimate

(3.2) (12Δx1)(12Δx2)ψN,(t,XN)1wN,(x1x2),ψN,(t,XN)1wN,(x1x2)C.\displaystyle\Big{\langle}(1-\hbar^{2}\Delta_{x_{1}})(1-\hbar^{2}\Delta_{x_{2}})\frac{\psi_{N,\hbar}(t,X_{N})}{1-w_{N,\hbar}(x_{1}-x_{2})},\frac{\psi_{N,\hbar}(t,X_{N})}{1-w_{N,\hbar}(x_{1}-x_{2})}\Big{\rangle}\leq C.

We first give the properties of the scattering function.

Lemma 3.1.

Suppose that V0V\geq 0 is smooth, spherical symmetric with compact support and 1wN,(x)1-w_{N,\hbar}(x) satisfies the scattering equation (3.1). Then there exists CC, depending on VV, such that

(3.3) 0wN,(x)CN2(|x|+Nβ),\displaystyle 0\leq w_{N,\hbar}(x)\leq\frac{C}{N\hbar^{2}(|x|+N^{-\beta})},
(3.4) |wN,(x)|CN2(|x|2+N2β),\displaystyle|\nabla w_{N,\hbar}(x)|\leq\frac{C}{N\hbar^{2}(|x|^{2}+N^{-2\beta})},

for all x3x\in\mathbb{R}^{3}.

Proof.

The properties of scattering function have been studied by many authors, see, for example, [3, 28, 49]. Here, we include a proof for completeness. First, by the maximum principle, it follows that (1wN,(x))1(1-w_{N,\hbar}(x))\leq 1. From the scattering equation (1.6), we can rewrite

(3.5) wN,(x)=c02N21|xy|VN(y)(1wN,(y))𝑑y,\displaystyle w_{N,\hbar}(x)=\frac{c_{0}}{2N\hbar^{2}}\int\frac{1}{|x-y|}V_{N}(y)(1-w_{N,\hbar}(y))dy,

where c0c_{0} is the renormalized constant.

Then the Hardy-Littlewood-Sobolev inequality implies that

(|x|+Nβ)wN,(x)=\displaystyle(|x|+N^{-\beta})w_{N,\hbar}(x)= c0N2|x|+Nβ|xy|VN(y)(1wN,(y))𝑑y\displaystyle\frac{c_{0}}{N\hbar^{2}}\int\frac{|x|+N^{-\beta}}{|x-y|}V_{N}(y)(1-w_{N,\hbar}(y))dy
\displaystyle\leq c0N2|xy|+|y|+Nβ|xy|VN(y)𝑑y\displaystyle\frac{c_{0}}{N\hbar^{2}}\int\frac{|x-y|+|y|+N^{-\beta}}{|x-y|}V_{N}(y)dy
=\displaystyle= c0VL1N2+c0N1+β21+Nβ|y|VN(y)|xy|𝑑y\displaystyle\frac{c_{0}\|V\|_{L^{1}}}{N\hbar^{2}}+\frac{c_{0}}{N^{1+\beta}\hbar^{2}}\int\frac{1+N^{\beta}|y|V_{N}(y)}{|x-y|}dy
\displaystyle\lesssim 1N2(VL1+yV(y)L32).\displaystyle\frac{1}{N\hbar^{2}}\left(\|V\|_{L^{1}}+\|\langle y\rangle V(y)\|_{L^{\frac{3}{2}}}\right).

For (3.4), by taking the gradient of (3.5), we also have

|(|x|2+N2β)xwN,(x)|\displaystyle|(|x|^{2}+N^{-2\beta})\nabla_{x}w_{N,\hbar}(x)|
\displaystyle\leq 12N2(|x|2+N2β)|x1|xy|VN(y)(1wN,(y))𝑑y|\displaystyle\frac{1}{2N\hbar^{2}}(|x|^{2}+N^{-2\beta})\Big{|}\int\nabla_{x}\frac{1}{|x-y|}V_{N}(y)(1-w_{N,\hbar}(y))dy\Big{|}
\displaystyle\leq 1N2|xy|2+|y|2+N2β|xy|2VN(y)𝑑y\displaystyle\frac{1}{N\hbar^{2}}\int\frac{|x-y|^{2}+|y|^{2}+N^{-2\beta}}{|x-y|^{2}}V_{N}(y)dy
\displaystyle\leq VL1N2+1N21+N2β|y|2VN(y)|xy|2𝑑y\displaystyle\frac{\|V\|_{L^{1}}}{N\hbar^{2}}+\frac{1}{N\hbar^{2}}\int\frac{1+N^{2\beta}|y|^{2}V_{N}(y)}{|x-y|^{2}}dy
\displaystyle\leq 1N2(VL1+y2V(y)L3).\displaystyle\frac{1}{N\hbar^{2}}\left(\|V\|_{L^{1}}+\|\langle y\rangle^{2}V(y)\|_{L^{3}}\right).

For simplicity, we adopt the shorthands

(3.6) w12=wN,(x1x2),w12=(wN,)(x1x2),\displaystyle w_{12}=w_{N,\hbar}(x_{1}-x_{2}),\quad\nabla w_{12}=(\nabla w_{N,\hbar})(x_{1}-x_{2}),

and start the proof of (3.2).

Lemma 3.2.

Let β(0,1)\beta\in(0,1) and Nβ121N^{\beta-1}\hbar^{-2}\ll 1. Then we have

(3.7) ψ,(HN,+N)2ψN(N1)16(12Δx1)(12Δx2)ψ1w12,ψ1w12\displaystyle\langle\psi,(H_{N,\hbar}+N)^{2}\psi\rangle\geq\frac{N(N-1)}{16}\Big{\langle}(1-\hbar^{2}\Delta_{x_{1}})(1-\hbar^{2}\Delta_{x_{2}})\frac{\psi}{1-w_{12}},\frac{\psi}{1-w_{12}}\Big{\rangle}

for ψLs2(3N)\psi\in L_{s}^{2}(\mathbb{R}^{3N}).

Proof.

Let

(3.8) Ti:=122Δi+12Nj:jiVN(xixj)+12Nj:jiVc(xixj),\displaystyle T_{i}:=1-\frac{\hbar^{2}}{2}\Delta_{i}+\frac{1}{2N}\sum_{j:j\neq i}V_{N}(x_{i}-x_{j})+\frac{1}{2N}\sum_{j:j\neq i}V_{c}(x_{i}-x_{j}),

we rewrite the Hamiltonian (1.2)

HN,+N=i=1NTi.\displaystyle H_{N,\hbar}+N=\sum_{i=1}^{N}T_{i}.

By the symmetry of ψ\psi, we have

(3.9) ψ,(HN,+N)2ψ=\displaystyle\langle\psi,(H_{N,\hbar}+N)^{2}\psi\rangle= i,jNψ,TiTjψ\displaystyle\sum_{i,j}^{N}\langle\psi,T_{i}T_{j}\psi\rangle
=\displaystyle= N(N1)ψ,T1T2ψ+Nψ,T12ψ\displaystyle N(N-1)\langle\psi,T_{1}T_{2}\psi\rangle+N\langle\psi,T_{1}^{2}\psi\rangle
\displaystyle\geq N(N1)ψ,T1T2ψ.\displaystyle N(N-1)\langle\psi,T_{1}T_{2}\psi\rangle.

Note that ψ=(1w12)ϕ12\psi=(1-w_{12})\phi_{12} and we have

2Δ1ψ=\displaystyle-\hbar^{2}\Delta_{1}\psi= 2Δ1[(1w12)ϕ12]\displaystyle-\hbar^{2}\Delta_{1}[(1-w_{12})\phi_{12}]
=\displaystyle= (1w12)(2Δ1ϕ12)+21w121ϕ12+2Δ1w12ϕ12.\displaystyle(1-w_{12})(-\hbar^{2}\Delta_{1}\phi_{12})+2\hbar\nabla_{1}w_{12}\hbar\nabla_{1}\phi_{12}+\hbar^{2}\Delta_{1}w_{12}\phi_{12}.

Thus, together with the scattering equation (3.1), we arrive at

(3.10) T1ψ=\displaystyle T_{1}\psi= T1[(1w12)ϕ12]\displaystyle T_{1}[(1-w_{12})\phi_{12}]
=\displaystyle= (1w12)(ϕ1222Δ1ϕ12)+1w121ϕ12+22Δ1w12ϕ12\displaystyle(1-w_{12})(\phi_{12}-\frac{\hbar^{2}}{2}\Delta_{1}\phi_{12})+\hbar\nabla_{1}w_{12}\hbar\nabla_{1}\phi_{12}+\frac{\hbar^{2}}{2}\Delta_{1}w_{12}\phi_{12}
+(1w12)[12Nj2VN(x1xj)ϕ12+12Nj2Vc(x1xj)ϕ12]\displaystyle+(1-w_{12})\left[\frac{1}{2N}\sum_{j\geq 2}V_{N}(x_{1}-x_{j})\phi_{12}+\frac{1}{2N}\sum_{j\geq 2}V_{c}(x_{1}-x_{j})\phi_{12}\right]
=\displaystyle= (1w12)[ϕ1222Δ1ϕ12+1w121w121ϕ12]\displaystyle(1-w_{12})\left[\phi_{12}-\frac{\hbar^{2}}{2}\Delta_{1}\phi_{12}+\frac{\hbar\nabla_{1}w_{12}}{1-w_{12}}\hbar\nabla_{1}\phi_{12}\right]
+(1w12)[12Nj3VN(x1xj)ϕ12+12Nj2Vc(x1xj)ϕ12].\displaystyle+(1-w_{12})\left[\frac{1}{2N}\sum_{j\geq 3}V_{N}(x_{1}-x_{j})\phi_{12}+\frac{1}{2N}\sum_{j\geq 2}V_{c}(x_{1}-x_{j})\phi_{12}\right].

Similarly, we also have

(3.11) T2ψ=\displaystyle T_{2}\psi= (1w12)[ϕ1222Δ2ϕ12+w121w122ϕ12]\displaystyle(1-w_{12})\left[\phi_{12}-\frac{\hbar^{2}}{2}\Delta_{2}\phi_{12}+\frac{\hbar\nabla w_{12}}{1-w_{12}}\hbar\nabla_{2}\phi_{12}\right]
+(1w12)[12Nj3VN(x2xj)ϕ12+12Nj2Vc(x2xj)ϕ12].\displaystyle+(1-w_{12})\left[\frac{1}{2N}\sum_{j\geq 3}V_{N}(x_{2}-x_{j})\phi_{12}+\frac{1}{2N}\sum_{j\neq 2}V_{c}(x_{2}-x_{j})\phi_{12}\right].

Further define the shorthands

(3.12) L1:=122Δ1+1w121w121,\displaystyle L_{1}:=1-\frac{\hbar^{2}}{2}\Delta_{1}+\frac{\hbar\nabla_{1}w_{12}}{1-w_{12}}\hbar\nabla_{1},
(3.13) L2:=122Δ2+2w121w122,\displaystyle L_{2}:=1-\frac{\hbar^{2}}{2}\Delta_{2}+\frac{\hbar\nabla_{2}w_{12}}{1-w_{12}}\hbar\nabla_{2},

which are symmetric with respect to the measure (1w12)2dx(1-w_{12})^{2}dx, that is,

(3.14) (1w12)2f¯(L1g)=\displaystyle\int(1-w_{12})^{2}\overline{f}(L_{1}g)= (1w12)2(L1f¯)g\displaystyle\int(1-w_{12})^{2}(\overline{L_{1}f})g
=\displaystyle= (1w12)2[f¯g+221f¯1g].\displaystyle\int(1-w_{12})^{2}\left[\overline{f}g+\frac{\hbar^{2}}{2}\nabla_{1}\overline{f}\nabla_{1}g\right].

Therefore, from (3.10) and (3.11) we obtain

T1ψ,T2ψ\displaystyle\langle T_{1}\psi,T_{2}\psi\rangle
=\displaystyle= (1w12)2(L1+12Nj3(VN+Vc)(x1xj)+12NVc(x1x2))ϕ¯12\displaystyle\int(1-w_{12})^{2}\left(L_{1}+\frac{1}{2N}\sum_{j\geq 3}(V_{N}+V_{c})(x_{1}-x_{j})+\frac{1}{2N}V_{c}(x_{1}-x_{2})\right)\overline{\phi}_{12}
(L2+12Nj3(VN+Vc)(x2xj)+12NVc(x1x2))ϕ12.\displaystyle\cdot\left(L_{2}+\frac{1}{2N}\sum_{j\geq 3}(V_{N}+V_{c})(x_{2}-x_{j})+\frac{1}{2N}V_{c}(x_{1}-x_{2})\right)\phi_{12}.

Expanding it gives

(3.15) T1ψ,T2ψ\displaystyle\langle T_{1}\psi,T_{2}\psi\rangle
=\displaystyle= (1w12)2L1ϕ¯12L2ϕ12\displaystyle\int(1-w_{12})^{2}L_{1}\overline{\phi}_{12}L_{2}\phi_{12}
+(1w12)2(L1ϕ¯12)[12Nj3(VN+Vc)(x2xj)+12NVc(x1x2)]ϕ12\displaystyle+\int(1-w_{12})^{2}(L_{1}\overline{\phi}_{12})\left[\frac{1}{2N}\sum_{j\geq 3}(V_{N}+V_{c})(x_{2}-x_{j})+\frac{1}{2N}V_{c}(x_{1}-x_{2})\right]\phi_{12}
+(1w12)2[12Nj3(VN+Vc)(x1xj)+12NVc(x1x2)]ϕ¯12L2ϕ12\displaystyle+\int(1-w_{12})^{2}\left[\frac{1}{2N}\sum_{j\geq 3}(V_{N}+V_{c})(x_{1}-x_{j})+\frac{1}{2N}V_{c}(x_{1}-x_{2})\right]\overline{\phi}_{12}L_{2}\phi_{12}
+(1w12)2[12Nj3(VN+Vc)(x1xj)+12NVc(x1x2)]ϕ¯12\displaystyle+\int(1-w_{12})^{2}\left[\frac{1}{2N}\sum_{j\geq 3}(V_{N}+V_{c})(x_{1}-x_{j})+\frac{1}{2N}V_{c}(x_{1}-x_{2})\right]\overline{\phi}_{12}
[12Nj3(VN+Vc)(x1xj)+12NVc(x1x2)]ϕ¯12.\displaystyle\quad\quad\quad\quad\quad\quad\quad\cdot\left[\frac{1}{2N}\sum_{j\geq 3}(V_{N}+V_{c})(x_{1}-x_{j})+\frac{1}{2N}V_{c}(x_{1}-x_{2})\right]\overline{\phi}_{12}.

By the nonnegativity of the potentials, we can discard the last term on the r.h.s of (3.15). The symmetry property (3.14) of the operators L1L_{1} and L2L_{2} then yields

T1ψ,T2ψ\displaystyle\langle T_{1}\psi,T_{2}\psi\rangle
\displaystyle\geq (1w12)2L1ϕ¯12L2ϕ12\displaystyle\int(1-w_{12})^{2}L_{1}\overline{\phi}_{12}L_{2}\phi_{12}
+(1w12)2(|ϕ12|2+22|1ϕ12|2)12Nj3(VN+Vc)(x2xj)\displaystyle+\int(1-w_{12})^{2}\left(|\phi_{12}|^{2}+\frac{\hbar^{2}}{2}|\nabla_{1}\phi_{12}|^{2}\right)\frac{1}{2N}\sum_{j\geq 3}(V_{N}+V_{c})(x_{2}-x_{j})
+(1w12)2|ϕ12|212NVc(x1x2)+22(1w12)21ϕ¯121(12NVc(x1x2)ϕ12)\displaystyle+\int(1-w_{12})^{2}|\phi_{12}|^{2}\frac{1}{2N}V_{c}(x_{1}-x_{2})+\frac{\hbar^{2}}{2}\int(1-w_{12})^{2}\nabla_{1}\overline{\phi}_{12}\nabla_{1}\left(\frac{1}{2N}V_{c}(x_{1}-x_{2})\phi_{12}\right)
+(1w12)2(|ϕ12|2+22|2ϕ12|2)12Nj3(VN+Vc)(x2xj)\displaystyle+\int(1-w_{12})^{2}\left(|\phi_{12}|^{2}+\frac{\hbar^{2}}{2}|\nabla_{2}\phi_{12}|^{2}\right)\frac{1}{2N}\sum_{j\geq 3}(V_{N}+V_{c})(x_{2}-x_{j})
+(1w12)2|ϕ12|212NVc(x1x2)+22(1w12)22(12NVc(x1x2)ϕ¯12)2ϕ12.\displaystyle+\int(1-w_{12})^{2}|\phi_{12}|^{2}\frac{1}{2N}V_{c}(x_{1}-x_{2})+\frac{\hbar^{2}}{2}\int(1-w_{12})^{2}\nabla_{2}\left(\frac{1}{2N}V_{c}(x_{1}-x_{2})\overline{\phi}_{12}\right)\nabla_{2}\phi_{12}.

Again using the nonnegativity of the potentials, we reach

(3.16) T1ψ,T2ψ\displaystyle\langle T_{1}\psi,T_{2}\psi\rangle\geq (1w12)2L1ϕ¯12L2ϕ12\displaystyle\int(1-w_{12})^{2}L_{1}\overline{\phi}_{12}L_{2}\phi_{12}
+22(1w12)21ϕ¯121(12NVc(x1x2)ϕ12)\displaystyle+\frac{\hbar^{2}}{2}\int(1-w_{12})^{2}\nabla_{1}\overline{\phi}_{12}\nabla_{1}\left(\frac{1}{2N}V_{c}(x_{1}-x_{2})\phi_{12}\right)
+22(1w12)22(12NVc(x1x2)ϕ¯12)2ϕ12\displaystyle+\frac{\hbar^{2}}{2}\int(1-w_{12})^{2}\nabla_{2}\left(\frac{1}{2N}V_{c}(x_{1}-x_{2})\overline{\phi}_{12}\right)\nabla_{2}\phi_{12}
=\displaystyle= I+II+III.\displaystyle I+II+III.

For the first term II on the r.h.s of (3.16), by (3.14), we have

(3.17) I=\displaystyle I= (1w12)2[ϕ¯12L2ϕ12+221ϕ¯121L2ϕ12]\displaystyle\int(1-w_{12})^{2}\left[\overline{\phi}_{12}L_{2}\phi_{12}+\frac{\hbar^{2}}{2}\nabla_{1}\overline{\phi}_{12}\nabla_{1}L_{2}\phi_{12}\right]
=\displaystyle= (1w12)2[|ϕ12|2+22|2ϕ12|2+22|1ϕ12|2+44|12ϕ12|2]\displaystyle\int(1-w_{12})^{2}\left[|\phi_{12}|^{2}+\frac{\hbar^{2}}{2}|\nabla_{2}\phi_{12}|^{2}+\frac{\hbar^{2}}{2}|\nabla_{1}\phi_{12}|^{2}+\frac{\hbar^{4}}{4}|\nabla_{1}\nabla_{2}\phi_{12}|^{2}\right]
+22(1w12)21ϕ¯12[1,L2]ϕ12\displaystyle+\frac{\hbar^{2}}{2}\int(1-w_{12})^{2}\nabla_{1}\overline{\phi}_{12}[\nabla_{1},L_{2}]\phi_{12}
\displaystyle\geq 12[|ϕ12|2+22|2ϕ12|2+22|1ϕ12|2+44|12ϕ12|2]\displaystyle\frac{1}{2}\int\left[|\phi_{12}|^{2}+\frac{\hbar^{2}}{2}|\nabla_{2}\phi_{12}|^{2}+\frac{\hbar^{2}}{2}|\nabla_{1}\phi_{12}|^{2}+\frac{\hbar^{4}}{4}|\nabla_{1}\nabla_{2}\phi_{12}|^{2}\right]
+22(1w12)21ϕ¯12[1,L2]ϕ12,\displaystyle+\frac{\hbar^{2}}{2}\int(1-w_{12})^{2}\nabla_{1}\overline{\phi}_{12}[\nabla_{1},L_{2}]\phi_{12},

where in the last inequality we have used Lemma 3.1 that (1w12)212(1-w_{12})^{2}\geq\frac{1}{2}. To control the last term on the r.h.s of (3.17), we note that

|[1,L2]|=2|[1,2w121w12]|2[|2w12|1w12+(w121w12)2].\displaystyle|[\nabla_{1},L_{2}]|=\hbar^{2}\Big{|}\left[\nabla_{1},\frac{\nabla_{2}w_{12}}{1-w_{12}}\right]\Big{|}\leq\hbar^{2}\left[\frac{|\nabla^{2}w_{12}|}{1-w_{12}}+\left(\frac{\nabla w_{12}}{1-w_{12}}\right)^{2}\right].

Therefore, we have

22(1w12)21ϕ¯12[1,L2]ϕ12\displaystyle\frac{\hbar^{2}}{2}\int(1-w_{12})^{2}\nabla_{1}\overline{\phi}_{12}[\nabla_{1},L_{2}]\phi_{12}\leq 42(|2w12|+|w12|2)|1ϕ12||ϕ12|\displaystyle\frac{\hbar^{4}}{2}\int\left(|\nabla^{2}w_{12}|+|\nabla w_{12}|^{2}\right)|\nabla_{1}\phi_{12}||\phi_{12}|
=\displaystyle= I1+I2.\displaystyle I_{1}+I_{2}.

For I1I_{1}, by Hölder and Sobolev inequalities we have

I1\displaystyle I_{1}\leq 42w12Lx2321ϕ12L2Lx26ϕ12L2Lx26\displaystyle\hbar^{4}\|\nabla^{2}w_{12}\|_{L_{x_{2}}^{\frac{3}{2}}}\|\nabla_{1}\phi_{12}\|_{L^{2}L_{x_{2}}^{6}}\|\phi_{12}\|_{L^{2}L_{x_{2}}^{6}}
\displaystyle\lesssim 42w12Lx23212ϕ12L2Lx222ϕ12L2Lx22.\displaystyle\hbar^{4}\|\nabla^{2}w_{12}\|_{L_{x_{2}}^{\frac{3}{2}}}\|\nabla_{1}\nabla_{2}\phi_{12}\|_{L^{2}L_{x_{2}}^{2}}\|\nabla_{2}\phi_{12}\|_{L^{2}L_{x_{2}}^{2}}.

By the Calderón-Zygmund theory which implies that 2fLpΔfLp\|\nabla^{2}f\|_{L^{p}}\lesssim\|\Delta f\|_{L^{p}} for 1<p<1<p<\infty and the scattering equation (3.1), we get

22w12Lx2322Δw12Lx2321NVN(x1x2)Lx232NβN.\displaystyle\hbar^{2}\|\nabla^{2}w_{12}\|_{L_{x_{2}}^{\frac{3}{2}}}\lesssim\hbar^{2}\|\Delta w_{12}\|_{L_{x_{2}}^{\frac{3}{2}}}\leq\frac{1}{N}\|V_{N}(x_{1}-x_{2})\|_{L_{x_{2}}^{\frac{3}{2}}}\lesssim\frac{N^{\beta}}{N}.

Thus, we arrive at

(3.18) I1NβN2(12ϕ12L22+1ϕ12L22).\displaystyle I_{1}\lesssim\frac{N^{\beta}}{N}\hbar^{2}\left(\|\nabla_{1}\nabla_{2}\phi_{12}\|_{L^{2}}^{2}+\|\nabla_{1}\phi_{12}\|_{L^{2}}^{2}\right).

For I2I_{2}, by the properties of the scattering function in Lemma 3.1, we have

|w12|CN2(|x1x2|2+N2β),\displaystyle|\nabla w_{12}|\leq\frac{C}{N\hbar^{2}(|x_{1}-x_{2}|^{2}+N^{-2\beta})},

which implies that

|w12|2N2βN21N2|x1x2|2=N2βN24|x1x2|2.\displaystyle|\nabla w_{12}|^{2}\lesssim\frac{N^{2\beta}}{N\hbar^{2}}\frac{1}{N\hbar^{2}|x_{1}-x_{2}|^{2}}=\frac{N^{2\beta}}{N^{2}\hbar^{4}|x_{1}-x_{2}|^{2}}.

Then by Cauchy-Schwarz and Hardy’s inequalities, we get

(3.19) I2\displaystyle I_{2}\leq 4|w12|2(|1ϕ12|2+|ϕ12|2)\displaystyle\hbar^{4}\int|\nabla w_{12}|^{2}(|\nabla_{1}\phi_{12}|^{2}+|\phi_{12}|^{2})
\displaystyle\lesssim N2βN2(12ϕ12L22+2ϕ12L22).\displaystyle\frac{N^{2\beta}}{N^{2}}\left(\|\nabla_{1}\nabla_{2}\phi_{12}\|_{L^{2}}^{2}+\|\nabla_{2}\phi_{12}\|_{L^{2}}^{2}\right).

Next, we deal with the terms IIII and IIIIII in (3.17). For IIII, we have

II=\displaystyle II= 22(1w12)21ϕ¯121(12NVc(x1x2)ϕ12)\displaystyle\frac{\hbar^{2}}{2}\int(1-w_{12})^{2}\nabla_{1}\overline{\phi}_{12}\nabla_{1}\left(\frac{1}{2N}V_{c}(x_{1}-x_{2})\phi_{12}\right)
=\displaystyle= 22(1w12)2[|1ϕ12|212NVc(x1x2)+(1ϕ¯12)ϕ12112NVc(x1x2)]\displaystyle\frac{\hbar^{2}}{2}\int(1-w_{12})^{2}\left[|\nabla_{1}\phi_{12}|^{2}\frac{1}{2N}V_{c}(x_{1}-x_{2})+(\nabla_{1}\overline{\phi}_{12})\phi_{12}\nabla_{1}\frac{1}{2N}V_{c}(x_{1}-x_{2})\right]
\displaystyle\geq 22(1w12)2(1ϕ¯12)ϕ12112NVc(x1x2),\displaystyle\frac{\hbar^{2}}{2}\int(1-w_{12})^{2}(\nabla_{1}\overline{\phi}_{12})\phi_{12}\nabla_{1}\frac{1}{2N}V_{c}(x_{1}-x_{2}),

where in the last inequality we used the positivity of the Coulomb potential. Noting that |Vc(x)||x|2|\nabla V_{c}(x)|\lesssim|x|^{-2}, we can use Cauchy-Schwarz and Hardy’s inequalities to obtain

(3.20) II\displaystyle II\geq 22N(|1ϕ12|2+|ϕ12|2)1|x1x2|2\displaystyle-\frac{\hbar^{2}}{2N}\int\left(|\nabla_{1}\phi_{12}|^{2}+|\phi_{12}|^{2}\right)\frac{1}{|x_{1}-x_{2}|^{2}}
\displaystyle\gtrsim 2N(12ϕ12L22+2ϕ12L22).\displaystyle-\frac{\hbar^{2}}{N}\left(\|\nabla_{1}\nabla_{2}\phi_{12}\|_{L^{2}}^{2}+\|\nabla_{2}\phi_{12}\|_{L^{2}}^{2}\right).

As the term IIIIII can be estimated in the same way as IIII, we also have

(3.21) III\displaystyle III\gtrsim 2N(12ϕ12L22+2ϕ12L22).\displaystyle-\frac{\hbar^{2}}{N}\left(\|\nabla_{1}\nabla_{2}\phi_{12}\|_{L^{2}}^{2}+\|\nabla_{2}\phi_{12}\|_{L^{2}}^{2}\right).

Together with estimates (3.16)–(3.21), we arrive at

T1ψ,T2ψ\displaystyle\langle T_{1}\psi,T_{2}\psi\rangle\geq 12[|ϕ12|2+22|2ϕ12|2+22|1ϕ12|2+44|12ϕ12|2]\displaystyle\frac{1}{2}\int\left[|\phi_{12}|^{2}+\frac{\hbar^{2}}{2}|\nabla_{2}\phi_{12}|^{2}+\frac{\hbar^{2}}{2}|\nabla_{1}\phi_{12}|^{2}+\frac{\hbar^{4}}{4}|\nabla_{1}\nabla_{2}\phi_{12}|^{2}\right]
C(NβN2+N2βN2)(12ϕ12L22+1ϕ12L22)\displaystyle-C(\frac{N^{\beta}}{N}\hbar^{2}+\frac{N^{2\beta}}{N^{2}})\left(\|\nabla_{1}\nabla_{2}\phi_{12}\|_{L^{2}}^{2}+\|\nabla_{1}\phi_{12}\|_{L^{2}}^{2}\right)
\displaystyle\geq 116|ϕ12|2+2|2ϕ12|2+2|1ϕ12|2+4|12ϕ12|2\displaystyle\frac{1}{16}\int|\phi_{12}|^{2}+\hbar^{2}|\nabla_{2}\phi_{12}|^{2}+\hbar^{2}|\nabla_{1}\phi_{12}|^{2}+\hbar^{4}|\nabla_{1}\nabla_{2}\phi_{12}|^{2}
=\displaystyle= 116(12Δ1)(12Δ2)ϕ12,ϕ12,\displaystyle\frac{1}{16}\langle(1-\hbar^{2}\Delta_{1})(1-\hbar^{2}\Delta_{2})\phi_{12},\phi_{12}\rangle,

where in the second-to-last inequality we have used that Nβ121N^{\beta-1}\hbar^{-2}\ll 1. With (3.9), we complete the proof of the estimate (3.7). ∎

Proposition 3.3.

Let β(0,1)\beta\in(0,1) and Nβ121N^{\beta-1}\hbar^{2}\ll 1. Define

ϕN,,12(t,XN)=ψN,(t,XN)1wN,(x1x2).\displaystyle\phi_{N,\hbar,12}(t,X_{N})=\frac{\psi_{N,\hbar}(t,X_{N})}{1-w_{N,\hbar}(x_{1}-x_{2})}.

There exists a constant C>0C>0 such that

(3.22) (12Δx1)(12Δx2)ϕN,,12(t),ϕN,,12(t)C\displaystyle\langle(1-\hbar^{2}\Delta_{x_{1}})(1-\hbar^{2}\Delta_{x_{2}})\phi_{N,\hbar,12}(t),\phi_{N,\hbar,12}(t)\rangle\leq C

for all tt\in\mathbb{R}.

Proof.

By the (HN,)2(H_{N,\hbar})^{2} energy estimate in Lemma 3.2, we have

(12Δx1)(12Δx2)ϕN,,12(t),ϕN,,12(t)\displaystyle\langle(1-\hbar^{2}\Delta_{x_{1}})(1-\hbar^{2}\Delta_{x_{2}})\phi_{N,\hbar,12}(t),\phi_{N,\hbar,12}(t)\rangle
\displaystyle\leq 16N(N1)ψN,(t),(N+HN,)2ψN,(t)\displaystyle\frac{16}{N(N-1)}\langle\psi_{N,\hbar}(t),(N+H_{N,\hbar})^{2}\psi_{N,\hbar}(t)\rangle
=\displaystyle= 16N(N1)ψN,(0),(N+HN,)2ψN,(0)\displaystyle\frac{16}{N(N-1)}\langle\psi_{N,\hbar}(0),(N+H_{N,\hbar})^{2}\psi_{N,\hbar}(0)\rangle
\displaystyle\leq 16(E0)2,\displaystyle 16(E_{0})^{2},

where we have used the conservation of (HN,)2(H_{N,\hbar})^{2} in the second-to-last equality, and the initial energy condition (1.12) in the last inequality. ∎

4. Functional Inequalities

In the section, with the a-priori energy bound established in Proposition 3.3, we control the δ\delta-type potential parts δ(t)\mathcal{F}_{\delta}(t) and ~δ(t)\widetilde{\mathcal{F}}_{\delta}(t) and establish the functional inequalities (1.26) and (1.27). In Section 4.1, we deal with the error analysis of the two-body term of ~δ(t)\widetilde{\mathcal{F}}_{\delta}(t) and then find the main part of ~δ(t)\widetilde{\mathcal{F}}_{\delta}(t). In Section 4.2, we estimate the main part. By a replacement argument, we find proper approximations of δ(t)\mathcal{F}_{\delta}(t) and ~δ(t)\widetilde{\mathcal{F}}_{\delta}(t), and hence arrive at a reduction of the functional inequality. We then complete the proof of the reduced version of functional inequalities in Section 4.3.

The main goal of the section is the following proposition which is the precise form of (1.26) and (1.27).

Proposition 4.1.

Let β(0,1)\beta\in(0,1), we have the estimate

(4.1) ~δ(t)δ(t)+O(Nβ16+Nβ34+N1104)\widetilde{\mathcal{F}}_{\delta}(t)\lesssim\mathcal{F}_{\delta}(t)+O(N^{\beta-1}\hbar^{-6}+N^{-\frac{\beta}{3}}\hbar^{-4}+N^{-\frac{1}{10}}\hbar^{-4})

and a lower bound of δ(t)\mathcal{F}_{\delta}(t)

(4.2) 0δ(t)+O(Nβ16+Nβ34+N1104).\displaystyle 0\leq\mathcal{F}_{\delta}(t)+O(N^{\beta-1}\hbar^{-6}+N^{-\frac{\beta}{3}}\hbar^{-4}+N^{-\frac{1}{10}}\hbar^{-4}).

Here, the notation O(a+b)O(a+b) is a shorthand for O(a)+O(b)O(a)+O(b) and the notation O(Nα1α2)O(N^{-\alpha_{1}}\hbar^{-\alpha_{2}}) denotes the same order of Nα1α2N^{-\alpha_{1}}\hbar^{-\alpha_{2}} up to an unimportant constant777The constant could depend on the usual Sobolev constants and the fixed parameters such as the time T0T_{0}, the energy bound E0E_{0}, and the Sobolev norms of (ρ,u)(\rho,u) but the constant is independent of (N,)(N,\hbar). CC.

Proof.

We postpone the proof of Proposition 4.1 to the end of the Section 4.3. ∎

4.1. Error Analysis of Two-Body Term

From the expression of ~δ(t)\widetilde{\mathcal{F}}_{\delta}(t)

(4.3) ~δ(t)=\displaystyle\widetilde{\mathcal{F}}_{\delta}(t)= N1N(u(t,x1)u(t,x2))VN(x1x2)ρN,(2)(t,x1,x2)𝑑x1𝑑x2\displaystyle\frac{N-1}{N}\int(u(t,x_{1})-u(t,x_{2}))\nabla V_{N}(x_{1}-x_{2})\rho_{N,\hbar}^{(2)}(t,x_{1},x_{2})dx_{1}dx_{2}
b0divu(t,x1)ρ(t,x1)[ρ(t,x1)2ρN,(1)(t,x1)]𝑑x1,\displaystyle-b_{0}\int\operatorname{div}u(t,x_{1})\rho(t,x_{1})\left[\rho(t,x_{1})-2\rho_{N,\hbar}^{(1)}(t,x_{1})\right]dx_{1},

the difficult part is the two-body term

(u(t,x)u(t,y))VN(xy)ρN,(2)(t,x,y)𝑑x𝑑y.\displaystyle\int(u(t,x)-u(t,y))\cdot\nabla V_{N}(x-y)\rho_{N,\hbar}^{(2)}(t,x,y)dxdy.

At first sight, the lack of a uniform regularity estimate for the two-body density function ρN,(2)(x,y)\rho_{N,\hbar}^{(2)}(x,y) makes further analysis difficult. With the singular correlation structure in mind, we decompose the two-body density function into the singular and regular parts

ρN,(2)(x,y)=(1wN,(xy))2ρN,(2)(x,y)(1wN,(xy))2,\displaystyle\rho_{N,\hbar}^{(2)}(x,y)=(1-w_{N,\hbar}(x-y))^{2}\frac{\rho_{N,\hbar}^{(2)}(x,y)}{(1-w_{N,\hbar}(x-y))^{2}},

and rewrite the two-body term as

(4.4) (u(x)u(y))VN(xy)(1wN,(xy))2ρN,(2)(x,y)(1wN,(xy))2𝑑x𝑑y.\displaystyle\int(u(x)-u(y))\cdot\nabla V_{N}(x-y)(1-w_{N,\hbar}(x-y))^{2}\frac{\rho_{N,\hbar}^{(2)}(x,y)}{(1-w_{N,\hbar}(x-y))^{2}}dxdy.

That is, the singularities come from the potential VN\nabla V_{N} and the singular correlation function (1wN,(xy))2(1-w_{N,\hbar}(x-y))^{2}. As mentioned in (1.32) at the outline, a key observation to beat the singularities is a cancellation structure from the difference coupled with the δ\delta-type potential

(4.5) (u(x)u(y))VN(xy),\displaystyle(u(x)-u(y))V_{N}(x-y),

which would vanish as NN tends to the infinity. Such a structure is special for the δ\delta-type potential. Many common potentials including the Coulomb do not carry such a property.

We will prove that, based on (3.22), the cancellation structure (4.5) dominates the singularities generated by the delta-potential and singular correlation function, which allows us extract the main term from the two-body term.

Lemma 4.2.

Let β(0,1)\beta\in(0,1), we have

(4.6) (u(x)u(y))VN(xy)ρN,(2)(x,y)𝑑x𝑑y\displaystyle\int(u(x)-u(y))\cdot\nabla V_{N}(x-y)\rho_{N,\hbar}^{(2)}(x,y)dxdy
=\displaystyle= divu(x)VN(xy)ρN,(2)(x,y)𝑑x𝑑y±O(Nβ16+Nβ24),\displaystyle-\int\operatorname{div}u(x)V_{N}(x-y)\rho_{N,\hbar}^{(2)}(x,y)dxdy\pm O(N^{\beta-1}\hbar^{-6}+N^{-\frac{\beta}{2}}\hbar^{-4}),

where the notation f=g±O(N,)f=g\pm O(N,\hbar) means |fg|O(N,)|f-g|\leq O(N,\hbar).

Proof.

First, due to the singular correlation structure, we rewrite the two-body term as (4.4). To employ the cancellation structure (4.5), we take the derivative off VNV_{N} by integrating by parts

(u(x)u(y))VN(xy)(1wN,(xy))2ρN,(2)(x,y)(1wN,(xy))2𝑑x𝑑y\displaystyle\int(u(x)-u(y))\cdot\nabla V_{N}(x-y)(1-w_{N,\hbar}(x-y))^{2}\frac{\rho_{N,\hbar}^{(2)}(x,y)}{(1-w_{N,\hbar}(x-y))^{2}}dxdy
=\displaystyle= divu(x)VN(xy)ρN,(2)(x,y)𝑑x𝑑y\displaystyle-\int\operatorname{div}u(x)V_{N}(x-y)\rho_{N,\hbar}^{(2)}(x,y)dxdy
(4.7) (u(x)u(y))VN(xy)x[(1wN,(xy))2ρN,(2)(x,y)(1wN,(xy))2]dxdy.\displaystyle-\int(u(x)-u(y))V_{N}(x-y)\nabla_{x}\left[(1-w_{N,\hbar}(x-y))^{2}\frac{\rho_{N,\hbar}^{(2)}(x,y)}{(1-w_{N,\hbar}(x-y))^{2}}\right]dxdy.

It remains to show the term (4.7) is indeed an error term. For simplicity, we set

w12=wN,(x1x2),w12=(wN,)(x1x2),ϕN,,12=(1w12)ψN,.w_{12}=w_{N,\hbar}(x_{1}-x_{2}),\quad\nabla w_{12}=(\nabla w_{N,\hbar})(x_{1}-x_{2}),\quad\phi_{N,\hbar,12}=(1-w_{12})\psi_{N,\hbar}.

Then we have

(4.8) (u(x)u(y))VN(xy)x[(1wN,(xy))2ρN,(2)(x,y)(1wN,(xy))2]dxdy\displaystyle\int(u(x)-u(y))V_{N}(x-y)\nabla_{x}\left[(1-w_{N,\hbar}(x-y))^{2}\frac{\rho_{N,\hbar}^{(2)}(x,y)}{(1-w_{N,\hbar}(x-y))^{2}}\right]dxdy
=\displaystyle= (u(x1)u(x2))VN(x1x2)x1[(1w12)ϕN,,12]2dXN\displaystyle\int(u(x_{1})-u(x_{2}))V_{N}(x_{1}-x_{2})\nabla_{x_{1}}\left[(1-w_{12})\phi_{N,\hbar,12}\right]^{2}dX_{N}
\displaystyle\leq 2uL|x1x2|VN(x1x2)|w12|(1w12)|ϕN,,12|2𝑑XN\displaystyle 2\|\nabla u\|_{L^{\infty}}\int|x_{1}-x_{2}|V_{N}(x_{1}-x_{2})|\nabla w_{12}|(1-w_{12})|\phi_{N,\hbar,12}|^{2}dX_{N}
+2uL|x1x2|VN(x1x2)(1w12)2|x1ϕN,,12||ϕN,,12|𝑑XN\displaystyle+2\|\nabla u\|_{L^{\infty}}\int|x_{1}-x_{2}|V_{N}(x_{1}-x_{2})(1-w_{12})^{2}|\nabla_{x_{1}}\phi_{N,\hbar,12}||\phi_{N,\hbar,12}|dX_{N}
=:\displaystyle=: 2(A+B).\displaystyle 2(A+B).

We bound AA and BB, using the properties of the scattering function, the two-body energy estimate (3.22) and the operator inequalities in Lemma A.2.

For the term AA, by Lemma 3.1, we have the upper bound estimate

(1w12)1,|w12|N2βN2.(1-w_{12})\leq 1,\quad|\nabla w_{12}|\lesssim\frac{N^{2\beta}}{N\hbar^{2}}.

Therefore, we arrive at

(4.9) A=\displaystyle A= uL|x1x2|VN(x1x2)|w12|(1w12)|ϕN,,12|2𝑑XN\displaystyle\int\|\nabla u\|_{L^{\infty}}|x_{1}-x_{2}|V_{N}(x_{1}-x_{2})|\nabla w_{12}|(1-w_{12})|\phi_{N,\hbar,12}|^{2}dX_{N}
\displaystyle\lesssim NβN2uLNβ|x1x2|VN(x1x2)|ϕN,,12|2𝑑XN\displaystyle\frac{N^{\beta}}{N\hbar^{2}}\|\nabla u\|_{L^{\infty}}\int N^{\beta}|x_{1}-x_{2}|V_{N}(x_{1}-x_{2})|\phi_{N,\hbar,12}|^{2}dX_{N}
\displaystyle\lesssim NβN2uL|x|V(x)L1(1Δx1)(1Δx2)ϕN,,12,ϕN,,12,\displaystyle\frac{N^{\beta}}{N\hbar^{2}}\|\nabla u\|_{L^{\infty}}\||x|V(x)\|_{L^{1}}\langle(1-\Delta_{x_{1}})(1-\Delta_{x_{2}})\phi_{N,\hbar,12},\phi_{N,\hbar,12}\rangle,

where in the last inequality we used the operator inequality (A.2).

For the term BB, we first discard (1w12)2(1-w_{12})^{2} and then use Cauchy-Schwarz to get

B=\displaystyle B= uL|x1x2|VN(x1x2)(1w12)2|x1ϕN,,12||ϕN,,12|𝑑XN\displaystyle\|\nabla u\|_{L^{\infty}}\int|x_{1}-x_{2}|V_{N}(x_{1}-x_{2})(1-w_{12})^{2}|\nabla_{x_{1}}\phi_{N,\hbar,12}||\phi_{N,\hbar,12}|dX_{N}
(4.10) \displaystyle\leq uLNβ[αNβ|x1x2|VN(x1x2)ϕN,,12,ϕN,,12\displaystyle\frac{\|\nabla u\|_{L^{\infty}}}{N^{\beta}}\left[\alpha\langle N^{\beta}|x_{1}-x_{2}|V_{N}(x_{1}-x_{2})\phi_{N,\hbar,12},\phi_{N,\hbar,12}\rangle\right.
+α1Nβ|x1x2|VN(x1x2)x1ϕN,,12,x1ϕN,,12].\displaystyle\left.+\alpha^{-1}\langle N^{\beta}|x_{1}-x_{2}|V_{N}(x_{1}-x_{2})\nabla_{x_{1}}\phi_{N,\hbar,12},\nabla_{x_{1}}\phi_{N,\hbar,12}\rangle\right].

By applying the operator inequality (A.2) to the first term and the operator inequality (A.3) to the second term on the r.h.s of (4.1), we obtain

(4.11) B\displaystyle B\leq uLNβ(α|x|V(x)L1(1Δx1)(1Δx2)ϕN,,12,ϕN,,12\displaystyle\frac{\|\nabla u\|_{L^{\infty}}}{N^{\beta}}\left(\alpha\||x|V(x)\|_{L^{1}}\langle(1-\Delta_{x_{1}})(1-\Delta_{x_{2}})\phi_{N,\hbar,12},\phi_{N,\hbar,12}\rangle\right.
+α1Nβ|x|V(x)L32(1Δx1)(1Δx2)ϕN,,12,ϕN,,12)\displaystyle\left.+\alpha^{-1}N^{\beta}\||x|V(x)\|_{L^{\frac{3}{2}}}\langle(1-\Delta_{x_{1}})(1-\Delta_{x_{2}})\phi_{N,\hbar,12},\phi_{N,\hbar,12}\rangle\right)
\displaystyle\lesssim Nβ2(1Δx1)(1Δx2)ϕN,,12,ϕN,,12,\displaystyle N^{-\frac{\beta}{2}}\langle(1-\Delta_{x_{1}})(1-\Delta_{x_{2}})\phi_{N,\hbar,12},\phi_{N,\hbar,12}\rangle,

where in the last inequality we optimized the choice of α\alpha.

Together with (4.8) and estimates for the terms AA and BB, we reach

(u(x)u(y))VN(xy)x[(1wN,(xy))2ρN,(2)(x,y)(1wN,(xy))2]dxdy\displaystyle\int(u(x)-u(y))V_{N}(x-y)\nabla_{x}\left[(1-w_{N,\hbar}(x-y))^{2}\frac{\rho_{N,\hbar}^{(2)}(x,y)}{(1-w_{N,\hbar}(x-y))^{2}}\right]dxdy
\displaystyle\lesssim (Nβ12+Nβ2)(1Δx1)(1Δx2)ϕN,,12,ϕN,,12\displaystyle\left(N^{\beta-1}\hbar^{-2}+N^{-\frac{\beta}{2}}\right)\langle(1-\Delta_{x_{1}})(1-\Delta_{x_{2}})\phi_{N,\hbar,12},\phi_{N,\hbar,12}\rangle
\displaystyle\lesssim (Nβ12+Nβ2)4(12Δx1)(12Δx2)ϕN,,12,ϕN,,12\displaystyle\left(N^{\beta-1}\hbar^{-2}+N^{-\frac{\beta}{2}}\right)\hbar^{-4}\langle(1-\hbar^{2}\Delta_{x_{1}})(1-\hbar^{2}\Delta_{x_{2}})\phi_{N,\hbar,12},\phi_{N,\hbar,12}\rangle
\displaystyle\lesssim Nβ16+Nβ24,\displaystyle N^{\beta-1}\hbar^{-6}+N^{-\frac{\beta}{2}}\hbar^{-4},

where in the last inequality we used the two-body H1H^{1} energy bound (3.22). This completes the proof of (4.6).

4.2. Tamed Singularities

As a result of the error analysis of the two-body term, we are able to capture the main term

divu(x)VN(xy)ρN,(2)(x,y)𝑑x𝑑y.\displaystyle\int\operatorname{div}u(x)V_{N}(x-y)\rho_{N,\hbar}^{(2)}(x,y)dxdy.

Using the identity approximation to the one-body term of ~δ(t)\widetilde{\mathcal{F}}_{\delta}(t), we arrive at

~δ(t)\displaystyle\widetilde{\mathcal{F}}_{\delta}(t)\sim divu(t,x)VN(xy)[N1NρN,(2)(t,x,y)\displaystyle-\int\operatorname{div}u(t,x)V_{N}(x-y)\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(t,x,y)\right.
ρN,(1)(t,x)ρ(t,y)ρ(t,x)ρN,(1)(t,y)+ρ(t,x)ρ(t,y)]dxdy.\displaystyle\left.\quad\quad-\rho_{N,\hbar}^{(1)}(t,x)\rho(t,y)-\rho(t,x)\rho_{N,\hbar}^{(1)}(t,y)+\rho(t,x)\rho(t,y)\right]dxdy.

By the identity approximation again, we also have the approximation of δ(t)\mathcal{F}_{\delta}(t) that

δ(t)\displaystyle\mathcal{F}_{\delta}(t)\sim VN(xy)[N1NρN,(2)(t,x,y)\displaystyle\int V_{N}(x-y)\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(t,x,y)\right.
ρN,(1)(t,x)ρ(t,y)ρ(t,x)ρN,(1)(t,y)+ρ(t,x)ρ(t,y)]dxdy.\displaystyle\quad\quad\left.-\rho_{N,\hbar}^{(1)}(t,x)\rho(t,y)-\rho(t,x)\rho_{N,\hbar}^{(1)}(t,y)+\rho(t,x)\rho(t,y)\right]dxdy.

We now need to deal with the sharp singularity of VN(x)V_{N}(x). We tame the singularity by replacing VN(x)V_{N}(x) with a slowly varying potential with a number of good properties. However, the replacement relies on the regularity of the integrand. Therefore, we again need to decompose the two-body density function as the singular and relatively regular parts. We obtain proper approximations of δ(t)\mathcal{F}_{\delta}(t) and ~δ(t)\widetilde{\mathcal{F}}_{\delta}(t) and arrive at a reduced version of functional inequalities via a careful analysis.

The following is the main lemma of the section.

Lemma 4.3.

Let

G(x)=(1π)32e|x|2,GN(x)=N3ηG(Nηx).G(x)=\left(\frac{1}{\pi}\right)^{\frac{3}{2}}e^{-|x|^{2}},\quad G_{N}(x)=N^{3\eta}G(N^{\eta}x).

Then for the two-body term we have

(4.12) F(x)VN(xy)ρN,(2)(x,y)𝑑x𝑑y\displaystyle\int F(x)V_{N}(x-y)\rho_{N,\hbar}^{(2)}(x,y)dxdy
=\displaystyle= b0F(x)GN(xy)ρN,(2)(x,y)𝑑x𝑑y±O(Nβ16+Nβ34+Nη34),\displaystyle b_{0}\int F(x)G_{N}(x-y)\rho_{N,\hbar}^{(2)}(x,y)dxdy\pm O(N^{\beta-1}\hbar^{-6}+N^{-\frac{\beta}{3}}\hbar^{-4}+N^{-\frac{\eta}{3}}\hbar^{-4}),

and for the one-body term we have

(4.13) b0F(x)ρ(x)[ρ(x)2ρN,(1)(x)]𝑑x\displaystyle b_{0}\int F(x)\rho(x)\left[\rho(x)-2\rho_{N,\hbar}^{(1)}(x)\right]dx
=\displaystyle= b0F(x)GN(xy)[ρN,(1)(x)ρ(y)ρ(x)ρN,(1)(y)+ρ(x)ρ(y)]𝑑x𝑑y±O(Nη).\displaystyle b_{0}\int F(x)G_{N}(x-y)\left[-\rho_{N,\hbar}^{(1)}(x)\rho(y)-\rho(x)\rho_{N,\hbar}^{(1)}(y)+\rho(x)\rho(y)\right]dxdy\pm O(N^{-\eta}).

In particular, given F(x)=1F(x)=1, we have the approximation of δ(t)\mathcal{F}_{\delta}(t)

(4.14) δ(t)=\displaystyle\mathcal{F}_{\delta}(t)= b0GN(xy)[N1NρN,(2)(x,y)\displaystyle b_{0}\int G_{N}(x-y)\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(x,y)\right.
ρN,(1)(x)ρ(y)ρ(x)ρN,(1)(y)+ρ(x)ρ(y)]dxdy\displaystyle\quad\quad\left.-\rho_{N,\hbar}^{(1)}(x)\rho(y)-\rho(x)\rho_{N,\hbar}^{(1)}(y)+\rho(x)\rho(y)\right]dxdy
±O(Nβ16+Nβ34+Nη34),\displaystyle\pm O(N^{\beta-1}\hbar^{-6}+N^{-\frac{\beta}{3}}\hbar^{-4}+N^{-\frac{\eta}{3}}\hbar^{-4}),

and given F(x)=divu(x)F(x)=\operatorname{div}u(x), we have the approximation of ~δ(t)\widetilde{\mathcal{F}}_{\delta}(t)

(4.15) ~δ(t)=\displaystyle\widetilde{\mathcal{F}}_{\delta}(t)= b0divu(x)GN(xy)[N1NρN,(2)(x,y)\displaystyle-b_{0}\int\operatorname{div}u(x)G_{N}(x-y)\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(x,y)\right.
ρN,(1)(x)ρ(y)ρ(x)ρN,(1)(y)+ρ(x)ρ(y)]dxdy\displaystyle\left.\quad\quad-\rho_{N,\hbar}^{(1)}(x)\rho(y)-\rho(x)\rho_{N,\hbar}^{(1)}(y)+\rho(x)\rho(y)\right]dxdy
±O(Nβ16+Nβ34+Nη34).\displaystyle\pm O(N^{\beta-1}\hbar^{-6}+N^{-\frac{\beta}{3}}\hbar^{-4}+N^{-\frac{\eta}{3}}\hbar^{-4}).
Proof.

For (4.12), we recall ϕN,,12=(1w12)ψN,\phi_{N,\hbar,12}=(1-w_{12})\psi_{N,\hbar} and rewrite

F(x)VN(xy)ρN,(2)(x,y)𝑑x𝑑y\displaystyle\int F(x)V_{N}(x-y)\rho_{N,\hbar}^{(2)}(x,y)dxdy
=\displaystyle= F(x)VN(xy)(1w12)2ϕN,,12,ϕN,,12\displaystyle\langle F(x)V_{N}(x-y)(1-w_{12})^{2}\phi_{N,\hbar,12},\phi_{N,\hbar,12}\rangle
=\displaystyle= F(x)VN(xy)ϕN,,12,ϕN,,12+F(x)VN(xy)(2w12+(w12)2)ϕN,,12,ϕN,,12\displaystyle\langle F(x)V_{N}(x-y)\phi_{N,\hbar,12},\phi_{N,\hbar,12}\rangle+\langle F(x)V_{N}(x-y)(-2w_{12}+(w_{12})^{2})\phi_{N,\hbar,12},\phi_{N,\hbar,12}\rangle
=\displaystyle= I+II.\displaystyle I+II.

For the term II, we use the Poincare´\acute{e} type inequality with θ=13\theta=\frac{1}{3} in Lemma A.3 to obtain

|F(x)(VN(xy)b0δ(xy))ϕN,,12,ϕN,,12|\displaystyle|\langle F(x)(V_{N}(x-y)-b_{0}\delta(x-y))\phi_{N,\hbar,12},\phi_{N,\hbar,12}\rangle|
\displaystyle\lesssim Nβ3x1x2F(x1)ϕN,,12L2x1x2ϕN,,12L2\displaystyle N^{-\frac{\beta}{3}}\|\langle\nabla_{x_{1}}\rangle\langle\nabla_{x_{2}}\rangle F(x_{1})\phi_{N,\hbar,12}\|_{L^{2}}\|\langle\nabla_{x_{1}}\rangle\langle\nabla_{x_{2}}\rangle\phi_{N,\hbar,12}\|_{L^{2}}
\displaystyle\lesssim Nβ3(FL+FL)x1x2ϕN,,12L22Nβ34,\displaystyle N^{-\frac{\beta}{3}}(\|F\|_{L^{\infty}}+\|\nabla F\|_{L^{\infty}})\|\langle\nabla_{x_{1}}\rangle\langle\nabla_{x_{2}}\rangle\phi_{N,\hbar,12}\|_{L^{2}}^{2}\lesssim N^{-\frac{\beta}{3}}\hbar^{-4},

where in the last inequality we used the two-body energy bound (3.22).

For the term IIII, by Lemma 3.1, we have |w12|Nβ12|w_{12}|\lesssim N^{\beta-1}\hbar^{-2}. Therefore, we get

II\displaystyle II\lesssim FLVN(xy)|w12|ϕN,,12,ϕN,,12\displaystyle\|F\|_{L^{\infty}}\langle V_{N}(x-y)|w_{12}|\phi_{N,\hbar,12},\phi_{N,\hbar,12}\rangle
\displaystyle\lesssim Nβ12VN(xy)ϕN,,12,ϕN,,12\displaystyle N^{\beta-1}\hbar^{-2}\langle V_{N}(x-y)\phi_{N,\hbar,12},\phi_{N,\hbar,12}\rangle
\displaystyle\lesssim Nβ12(1Δx1)(1Δx2)ϕN,,12,ϕN,,12Nβ16,\displaystyle N^{\beta-1}\hbar^{-2}\langle(1-\Delta_{x_{1}})(1-\Delta_{x_{2}})\phi_{N,\hbar,12},\phi_{N,\hbar,12}\rangle\lesssim N^{\beta-1}\hbar^{-6},

where we used the operator inequality (A.2) in the second-to-last inequality and the two-body energy bound (3.22) in the last inequality.

In the same way, we also obtain

b0F(x)GN(xy)ρN,(2)(x,y)𝑑x𝑑y\displaystyle b_{0}\int F(x)G_{N}(x-y)\rho_{N,\hbar}^{(2)}(x,y)dxdy
=\displaystyle= b0F(x)δ(xy)ρN,(2)(x,y)𝑑x𝑑y±O(Nη34+Nβ16).\displaystyle b_{0}\int F(x)\delta(x-y)\rho_{N,\hbar}^{(2)}(x,y)dxdy\pm O(N^{-\frac{\eta}{3}}\hbar^{-4}+N^{\beta-1}\hbar^{-6}).

Then by the triangle inequality, we arrive at

|F(x)VN(xy)ρN,(2)(x,y)𝑑x𝑑yb0F(x)GN(xy)ρN,(2)(x,y)𝑑x𝑑y|\displaystyle\Big{|}\int F(x)V_{N}(x-y)\rho_{N,\hbar}^{(2)}(x,y)dxdy-b_{0}\int F(x)G_{N}(x-y)\rho_{N,\hbar}^{(2)}(x,y)dxdy\Big{|}
\displaystyle\lesssim Nβ16+Nβ34+Nη34,\displaystyle N^{\beta-1}\hbar^{-6}+N^{-\frac{\beta}{3}}\hbar^{-4}+N^{-\frac{\eta}{3}}\hbar^{-4},

which completes the proof of (4.12).

For (4.13), we rewrite

F(x)ρ(x)[ρ(x)2ρN,(1)(x)]𝑑x\displaystyle\int F(x)\rho(x)\left[\rho(x)-2\rho_{N,\hbar}^{(1)}(x)\right]dx
=\displaystyle= F(x)δ(xy)[ρN,(1)(x)ρ(y)ρ(x)ρN,(1)(y)+ρ(x)ρ(y)]𝑑x𝑑y\displaystyle\int F(x)\delta(x-y)\left[-\rho_{N,\hbar}^{(1)}(x)\rho(y)-\rho(x)\rho_{N,\hbar}^{(1)}(y)+\rho(x)\rho(y)\right]dxdy
=\displaystyle= F(x)GN(xy)[ρN,(1)(x)ρ(y)ρ(x)ρN,(1)(y)+ρ(x)ρ(y)]𝑑x𝑑y\displaystyle\int F(x)G_{N}(x-y)\left[-\rho_{N,\hbar}^{(1)}(x)\rho(y)-\rho(x)\rho_{N,\hbar}^{(1)}(y)+\rho(x)\rho(y)\right]dxdy
(4.16) FρN,(1),(GNδ)ρ(GNδ)(Fρ),ρN,(1)+Fρ,(GNδ)ρ.\displaystyle-\langle F\rho_{N,\hbar}^{(1)},(G_{N}-\delta)*\rho\rangle-\langle(G_{N}-\delta)*(F\rho),\rho_{N,\hbar}^{(1)}\rangle+\langle F\rho,(G_{N}-\delta)*\rho\rangle.

For the error terms in (4.16), we use Hölder and Sobolev inequalities to get

|FρN,(1),(GNδ)ρ|+|(GNδ)(Fρ),ρN,(1)|+|Fρ,(GNδ)ρ|\displaystyle|\langle F\rho_{N,\hbar}^{(1)},(G_{N}-\delta)*\rho\rangle|+|\langle(G_{N}-\delta)*(F\rho),\rho_{N,\hbar}^{(1)}\rangle|+|\langle F\rho,(G_{N}-\delta)*\rho\rangle|
\displaystyle\leq FL(ρN,(1)L1+ρL1)(GNδ)ρL+ρN,(1)L1(GNδ)(Fρ)L\displaystyle\|F\|_{L^{\infty}}(\|\rho_{N,\hbar}^{(1)}\|_{L^{1}}+\|\rho\|_{L^{1}})\|(G_{N}-\delta)*\rho\|_{L^{\infty}}+\|\rho_{N,\hbar}^{(1)}\|_{L^{1}}\|(G_{N}-\delta)*(F\rho)\|_{L^{\infty}}
\displaystyle\lesssim FL(ρN,(1)L1+ρL1)(GNδ)2ρL2+ρN,(1)L1(GNδ)2(Fρ)L2\displaystyle\|F\|_{L^{\infty}}(\|\rho_{N,\hbar}^{(1)}\|_{L^{1}}+\|\rho\|_{L^{1}})\|(G_{N}-\delta)*\langle\nabla\rangle^{2}\rho\|_{L^{2}}+\|\rho_{N,\hbar}^{(1)}\|_{L^{1}}\|(G_{N}-\delta)*\langle\nabla\rangle^{2}(F\rho)\|_{L^{2}}
\displaystyle\lesssim Nη(FLρH3+FρH3)\displaystyle N^{-\eta}\left(\|F\|_{L^{\infty}}\|\rho\|_{H^{3}}+\|F\rho\|_{H^{3}}\right)
\displaystyle\lesssim NηFH3ρH3,\displaystyle N^{-\eta}\|F\|_{H^{3}}\|\rho\|_{H^{3}},

where in the second-to-last inequality we used Lemma A.1 and the mass conservation, and in the last inequality we used Leibniz rule and Sobolev inequality. Therefore, we complete the proof of (4.13).

For (4.14), by taking F(x)=1F(x)=1 in (4.12) and (4.13), we arrive at the approximation of δ(t)\mathcal{F}_{\delta}(t).

For (4.15), by the error analysis (4.6) in Lemma 4.2 we get

~δ=\displaystyle\widetilde{\mathcal{F}}_{\delta}= b0divu(x)VN(xy)ρN,(2)(x,y)𝑑x𝑑y\displaystyle-b_{0}\int\operatorname{div}u(x)V_{N}(x-y)\rho_{N,\hbar}^{(2)}(x,y)dxdy
b0divu(x)ρ(x)[ρ(x)2ρN,(1)(x)]𝑑x\displaystyle-b_{0}\int\operatorname{div}u(x)\rho(x)\left[\rho(x)-2\rho_{N,\hbar}^{(1)}(x)\right]dx
±O(Nβ16+Nβ24).\displaystyle\pm O(N^{\beta-1}\hbar^{-6}+N^{-\frac{\beta}{2}}\hbar^{-4}).

Then by taking F(x)=divu(x)F(x)=\operatorname{div}u(x) in (4.12) and (4.13), we get the approximation (4.15) of ~δ(t)\widetilde{\mathcal{F}}_{\delta}(t). ∎

4.3. Reduced Version of Functional Inequality

After the analysis of error terms and simplification, we now work with a reduced form of functional inequality

(4.17) divu(x)GN(xy)[N1NρN,(2)(x,y)ρN,(1)(x)ρ(y)ρ(x)ρN,(1)(y)+ρ(x)ρ(y)]𝑑x𝑑y\displaystyle\int\operatorname{div}u(x)G_{N}(x-y)\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(x,y)-\rho_{N,\hbar}^{(1)}(x)\rho(y)-\rho(x)\rho_{N,\hbar}^{(1)}(y)+\rho(x)\rho(y)\right]dxdy
\displaystyle\lesssim GN(xy)[N1NρN,(2)(x,y)ρN,(1)(x)ρ(y)ρ(x)ρN,(1)(y)+ρ(x)ρ(y)]𝑑x𝑑y+o(1),\displaystyle\int G_{N}(x-y)\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(x,y)-\rho_{N,\hbar}^{(1)}(x)\rho(y)-\rho(x)\rho_{N,\hbar}^{(1)}(y)+\rho(x)\rho(y)\right]dxdy+o(1),

which is more concise than the original functional inequality. However, it is unknown whether or not the integrand

(4.18) N1NρN,(2)(x,y)ρN,(1)(x)ρ(y)ρ(x)ρN,(1)(y)+ρ(x)ρ(y)\displaystyle\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(x,y)-\rho_{N,\hbar}^{(1)}(x)\rho(y)-\rho(x)\rho_{N,\hbar}^{(1)}(y)+\rho(x)\rho(y)

is non-negative, so we cannot directly bound the term divu(x)\operatorname{div}u(x) in (4.17). We prove that, if integrated against GN(xy)G_{N}(x-y), (4.18) provides a non-negative contribution up to a small correction and use that to prove the lower bound of δ(t)\mathcal{F}_{\delta}(t). The special structure of a relatively slowly varying and explicit potential GN(x)G_{N}(x) plays a critical role in establishing the reduced version of functional inequality. We then complete the proof of Proposition 4.1.

Lemma 4.4 (Reduced Version of Functional Inequality).

Let η<13\eta<\frac{1}{3} to be determined, we have

(4.19) F(x)GN(xy)[N1NρN,(2)(x,y)ρN,(1)(x)ρ(y)ρ(x)ρN,(1)(y)+ρ(x)ρ(y)]𝑑x𝑑y\displaystyle\int F(x)G_{N}(x-y)\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(x,y)-\rho_{N,\hbar}^{(1)}(x)\rho(y)-\rho(x)\rho_{N,\hbar}^{(1)}(y)+\rho(x)\rho(y)\right]dxdy
\displaystyle\leq FLGN(xy)[N1NρN,(2)(x,y)ρN,(1)(x)ρ(y)ρ(x)ρN,(1)(y)+ρ(x)ρ(y)]𝑑x𝑑y\displaystyle\|F\|_{L^{\infty}}\int G_{N}(x-y)\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(x,y)-\rho_{N,\hbar}^{(1)}(x)\rho(y)-\rho(x)\rho_{N,\hbar}^{(1)}(y)+\rho(x)\rho(y)\right]dxdy
+O(Nη2+N3η1).\displaystyle+O(N^{-\eta}\hbar^{-2}+N^{3\eta-1}).
Proof.

For simplicity, set ρN,(XN)=|ψN,(XN)|2\rho_{N,\hbar}(X_{N})=|\psi_{N,\hbar}(X_{N})|^{2}. By the symmetry of ρN,(XN)\rho_{N,\hbar}(X_{N}), we can write

F(x)GN(xy)[N1NρN,(2)(x,y)ρN,(1)(x)ρ(y)ρ(x)ρN,(1)(y)+ρ(x)ρ(y)]𝑑x𝑑y\displaystyle\int F(x)G_{N}(x-y)\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(x,y)-\rho_{N,\hbar}^{(1)}(x)\rho(y)-\rho(x)\rho_{N,\hbar}^{(1)}(y)+\rho(x)\rho(y)\right]dxdy
=\displaystyle= 1N2ijNF(xi)GN(xixj)ρN,(XN)𝑑XN+F(x)GN(xy)ρ(x)ρ(y)𝑑x𝑑y\displaystyle\frac{1}{N^{2}}\sum_{i\neq j}^{N}\int F(x_{i})G_{N}(x_{i}-x_{j})\rho_{N,\hbar}(X_{N})dX_{N}+\int F(x)G_{N}(x-y)\rho(x)\rho(y)dxdy
1Ni=1NF(xi)GN(xiy)ρ(y)𝑑yρN,(XN)𝑑XN\displaystyle-\frac{1}{N}\sum_{i=1}^{N}\int\int F(x_{i})G_{N}(x_{i}-y)\rho(y)dy\rho_{N,\hbar}(X_{N})dX_{N}
1Nj=1NF(x)GN(xxj)ρ(x)𝑑xρN,(XN)𝑑XN\displaystyle-\frac{1}{N}\sum_{j=1}^{N}\int\int F(x)G_{N}(x-x_{j})\rho(x)dx\rho_{N,\hbar}(X_{N})dX_{N}
=\displaystyle= F(x)GN(xy)[1N2ijNδxi(x)δxj(y)+ρ(x)ρ(y)\displaystyle\int F(x)G_{N}(x-y)\left[\frac{1}{N^{2}}\sum_{i\neq j}^{N}\delta_{x_{i}}(x)\delta_{x_{j}}(y)+\rho(x)\rho(y)\right.
1Ni=1Nδxi(x)ρ(y)ρ(x)1Nj=1Nδxj(y)]dxdyρN,(XN)dXN.\displaystyle\left.-\frac{1}{N}\sum_{i=1}^{N}\delta_{x_{i}}(x)\rho(y)-\rho(x)\frac{1}{N}\sum_{j=1}^{N}\delta_{x_{j}}(y)\right]dxdy\rho_{N,\hbar}(X_{N})dX_{N}.

To simplify, we define the measure

(4.20) νXN(dx)=1Ni=1Nδxi(dx)ρ(x)dx.\displaystyle\nu_{X_{N}}(dx)=\frac{1}{N}\sum_{i=1}^{N}\delta_{x_{i}}(dx)-\rho(x)dx.

We rewrite

(4.21) F(x)GN(xy)[N1NρN,(2)(x,y)ρN,(1)(x)ρ(y)ρ(x)ρN,(1)(y)+ρ(x)ρ(y)]𝑑x𝑑y\displaystyle\int F(x)G_{N}(x-y)\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(x,y)-\rho_{N,\hbar}^{(1)}(x)\rho(y)-\rho(x)\rho_{N,\hbar}^{(1)}(y)+\rho(x)\rho(y)\right]dxdy
=\displaystyle= F(x)GN(xy)νXN(dx)νXN(dy)ρN,(XN)𝑑XNGN(0)NF(x)ρN,(1)(x)𝑑x.\displaystyle\int F(x)G_{N}(x-y)\nu_{X_{N}}(dx)\nu_{X_{N}}(dy)\rho_{N,\hbar}(X_{N})dX_{N}-\frac{G_{N}(0)}{N}\int F(x)\rho_{N,\hbar}^{(1)}(x)dx.

where the last term on the r.h.s of (4.21) comes from the diagonal summation. In particular, if we take F(x)=1F(x)=1, we also have

(4.22) GN(xy)[N1NρN,(2)(x,y)ρN,(1)(x)ρ(y)ρ(x)ρN,(1)(y)+ρ(x)ρ(y)]𝑑x𝑑y\displaystyle\int G_{N}(x-y)\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(x,y)-\rho_{N,\hbar}^{(1)}(x)\rho(y)-\rho(x)\rho_{N,\hbar}^{(1)}(y)+\rho(x)\rho(y)\right]dxdy
=\displaystyle= GN(xy)νXN(dx)νXN(dy)ρN,(XN)𝑑XNGN(0)NρN,(1)(x)𝑑x.\displaystyle\int G_{N}(x-y)\nu_{X_{N}}(dx)\nu_{X_{N}}(dy)\rho_{N,\hbar}(X_{N})dX_{N}-\frac{G_{N}(0)}{N}\int\rho_{N,\hbar}^{(1)}(x)dx.

Note that

(4.23) GN(0)NF(x)ρN,(1)(x)𝑑xN3η1FL,\displaystyle\frac{G_{N}(0)}{N}\int F(x)\rho_{N,\hbar}^{(1)}(x)dx\leq N^{3\eta-1}\|F\|_{L^{\infty}},

which is a smallness term as long as η<13\eta<\frac{1}{3}.

Next, we get into the analysis of the main term. Note that the convolution property of the Gaussian function GG, which is

(4.24) GN(xy)=G0,N(xz)G0,N(zy)𝑑z,\displaystyle G_{N}(x-y)=\int G_{0,N}(x-z)G_{0,N}(z-y)dz,

where G0,N(x)=N3ηG0(Nηx)G_{0,N}(x)=N^{3\eta}G_{0}(N^{\eta}x) and G0(x)=(2π)32e2|x|2G_{0}(x)=\left(\frac{2}{\pi}\right)^{\frac{3}{2}}e^{-2|x|^{2}}. Putting (4.24) into the main term of (4.22) gives

(4.25) F(x)GN(xy)νXN(dx)νXN(dy)ρN,(XN)𝑑XN\displaystyle\int F(x)G_{N}(x-y)\nu_{X_{N}}(dx)\nu_{X_{N}}(dy)\rho_{N,\hbar}(X_{N})dX_{N}
=\displaystyle= F(x)G0,N(xz)G0,N(zy)νXN(dx)νXN(dy)ρN,(XN)𝑑z𝑑XN\displaystyle\int F(x)G_{0,N}(x-z)G_{0,N}(z-y)\nu_{X_{N}}(dx)\nu_{X_{N}}(dy)\rho_{N,\hbar}(X_{N})dzdX_{N}
=\displaystyle= A+B,\displaystyle A+B,

where

(4.26) A=\displaystyle A= (F(x)F(z))G0,N(xz)G0,N(zy)νXN(dx)νXN(dy)ρN,(XN)𝑑z𝑑XN,\displaystyle\int(F(x)-F(z))G_{0,N}(x-z)G_{0,N}(z-y)\nu_{X_{N}}(dx)\nu_{X_{N}}(dy)\rho_{N,\hbar}(X_{N})dzdX_{N},
(4.27) B=\displaystyle B= F(z)G0,N(xz)G0,N(zy)νXN(dx)νXN(dy)ρN,(XN)𝑑z𝑑XN.\displaystyle\int F(z)G_{0,N}(x-z)G_{0,N}(z-y)\nu_{X_{N}}(dx)\nu_{X_{N}}(dy)\rho_{N,\hbar}(X_{N})dzdX_{N}.

Thus, we are left to bound the terms AA and BB.

For the term AA, we use Cauchy-Schwarz inequality to get

A2\displaystyle A^{2}\leq [(F(x)F(z))G0,N(xz)νXN(dx)]2ρN,(XN)𝑑z𝑑XN\displaystyle\int\left[\int(F(x)-F(z))G_{0,N}(x-z)\nu_{X_{N}}(dx)\right]^{2}\rho_{N,\hbar}(X_{N})dzdX_{N}
[G0,N(zy)νXN(dy)]2ρN,(XN)dzdXN\displaystyle\cdot\int\left[\int G_{0,N}(z-y)\nu_{X_{N}}(dy)\right]^{2}\rho_{N,\hbar}(X_{N})dzdX_{N}
\displaystyle\leq 2(A1+A2)[G0,N(zy)νXN(dy)]2ρN,(XN)𝑑z𝑑XN,\displaystyle 2(A_{1}+A_{2})\int\left[\int G_{0,N}(z-y)\nu_{X_{N}}(dy)\right]^{2}\rho_{N,\hbar}(X_{N})dzdX_{N},

where

A1=\displaystyle A_{1}= [(F(x)F(z))G0,N(xz)1Ni=1Nδxi(dx)]2ρN,(XN)𝑑z𝑑XN,\displaystyle\int\left[\int(F(x)-F(z))G_{0,N}(x-z)\frac{1}{N}\sum_{i=1}^{N}\delta_{x_{i}}(dx)\right]^{2}\rho_{N,\hbar}(X_{N})dzdX_{N},
A2=\displaystyle A_{2}= [(F(x)F(z))G0,N(xz)ρ(x)𝑑x]2ρN,(XN)𝑑z𝑑XN.\displaystyle\int\left[\int(F(x)-F(z))G_{0,N}(x-z)\rho(x)dx\right]^{2}\rho_{N,\hbar}(X_{N})dzdX_{N}.

For A1A_{1}, we further decompose it into two parts A1=A11+A12A_{1}=A_{11}+A_{12}, where the diagonal part is

A11=\displaystyle A_{11}= 1N2i=1N(F(xi)F(z))G0,N(xiz)(F(xi)F(z))G0,N(xiz)ρN,(XN)𝑑z𝑑XN,\displaystyle\frac{1}{N^{2}}\sum_{i=1}^{N}\int(F(x_{i})-F(z))G_{0,N}(x_{i}-z)(F(x_{i})-F(z))G_{0,N}(x_{i}-z)\rho_{N,\hbar}(X_{N})dzdX_{N},

and the off-diagonal part is

A12=\displaystyle A_{12}= 1N2ijN(F(xi)F(z))G0,N(xiz)(F(xj)F(z))G0,N(xjz)ρN,(XN)𝑑z𝑑XN.\displaystyle\frac{1}{N^{2}}\sum_{i\neq j}^{N}\int(F(x_{i})-F(z))G_{0,N}(x_{i}-z)(F(x_{j})-F(z))G_{0,N}(x_{j}-z)\rho_{N,\hbar}(X_{N})dzdX_{N}.

For A11A_{11}, by the symmetry of ρN,(XN)\rho_{N,\hbar}(X_{N}), we have

A11\displaystyle A_{11}\leq 1N(F(x1)F(z))G0,N(x1z)(F(x1)F(z))G0,N(x1z)ρN,(XN)𝑑z𝑑XN\displaystyle\frac{1}{N}\int(F(x_{1})-F(z))G_{0,N}(x_{1}-z)(F(x_{1})-F(z))G_{0,N}(x_{1}-z)\rho_{N,\hbar}(X_{N})dzdX_{N}
\displaystyle\leq FL2N(|x1z|G0,N(x1z))2ρN,(XN)𝑑z𝑑XN\displaystyle\frac{\|\nabla F\|_{L^{\infty}}^{2}}{N}\int\left(|x_{1}-z|G_{0,N}(x_{1}-z)\right)^{2}\rho_{N,\hbar}(X_{N})dzdX_{N}
=\displaystyle= FL2N|x|G0,N(x)L22ρN,(XN)𝑑XNNη1,\displaystyle\frac{\|\nabla F\|_{L^{\infty}}^{2}}{N}\||x|G_{0,N}(x)\|_{L^{2}}^{2}\int\rho_{N,\hbar}(X_{N})dX_{N}\lesssim N^{\eta-1},

where in the last inequality we used that |x|G0,N(x)L22Nη\||x|G_{0,N}(x)\|_{L^{2}}^{2}\lesssim N^{\eta} and the mass conservation for ρN,(XN)\rho_{N,\hbar}(X_{N}).

For A12A_{12}, by the symmetry of ρN,(XN)\rho_{N,\hbar}(X_{N}), we also have

A12\displaystyle A_{12}\leq |(F(x1)F(z))G0,N(x1z)(F(x2)F(z))G0,N(x2z)|ρN,(XN)𝑑z𝑑XN\displaystyle\int|(F(x_{1})-F(z))G_{0,N}(x_{1}-z)(F(x_{2})-F(z))G_{0,N}(x_{2}-z)|\rho_{N,\hbar}(X_{N})dzdX_{N}
\displaystyle\leq FL2|x1z|G0,N(x1z)|x2z|G0,N(x2z)|ρN,(XN)dzdXN\displaystyle\|\nabla F\|_{L^{\infty}}^{2}\int|x_{1}-z|G_{0,N}(x_{1}-z)|x_{2}-z|G_{0,N}(x_{2}-z)|\rho_{N,\hbar}(X_{N})dzdX_{N}
\displaystyle\leq FL2N2ηG1,N(x1x2)ρN,(XN)𝑑XN\displaystyle\frac{\|\nabla F\|_{L^{\infty}}^{2}}{N^{2\eta}}\int G_{1,N}(x_{1}-x_{2})\rho_{N,\hbar}(X_{N})dX_{N}
=\displaystyle= FL2N2ηG1,N(x1x2)ψN,,ψN,,\displaystyle\frac{\|\nabla F\|_{L^{\infty}}^{2}}{N^{2\eta}}\langle G_{1,N}(x_{1}-x_{2})\psi_{N,\hbar},\psi_{N,\hbar}\rangle,

where

G1,N(x1x2)=Nη|x1z|G0,N(x1z)Nη|x2z|G0,N(x2z)|dz.\displaystyle G_{1,N}(x_{1}-x_{2})=\int N^{\eta}|x_{1}-z|G_{0,N}(x_{1}-z)N^{\eta}|x_{2}-z|G_{0,N}(x_{2}-z)|dz.

To bound A12A_{12}, we recall ϕN,,12=(1w12)ψN,\phi_{N,\hbar,12}=(1-w_{12})\psi_{N,\hbar} then get

A12\displaystyle A_{12}\lesssim N2ηG1,N(x1x2)(1w12)2ϕN,,12,ϕN,,12\displaystyle N^{-2\eta}\langle G_{1,N}(x_{1}-x_{2})(1-w_{12})^{2}\phi_{N,\hbar,12},\phi_{N,\hbar,12}\rangle
\displaystyle\lesssim N2ηG1,N(x1x2)ϕN,,12,ϕN,,12\displaystyle N^{-2\eta}\langle G_{1,N}(x_{1}-x_{2})\phi_{N,\hbar,12},\phi_{N,\hbar,12}\rangle
\displaystyle\lesssim N2ηG1,NL1(1Δ1)(1Δ2)ϕN,,12,ϕN,,12N2η4,\displaystyle N^{-2\eta}\|G_{1,N}\|_{L^{1}}\langle(1-\Delta_{1})(1-\Delta_{2})\phi_{N,\hbar,12},\phi_{N,\hbar,12}\rangle\lesssim N^{-2\eta}\hbar^{-4},

where we discarded (1w12)2(1-w_{12})^{2} in the second line and used the operator inequality (A.2) in the second-to-last inequality, and the two-body H1H^{1} energy bound (3.22) in the last inequality.

For A2A_{2}, we rewrite

A2=\displaystyle A_{2}= F(G0,Nρ)G0,N(Fρ)L22ρN,(XN)𝑑XN\displaystyle\|F(G_{0,N}*\rho)-G_{0,N}*(F\rho)\|_{L^{2}}^{2}\int\rho_{N,\hbar}(X_{N})dX_{N}
=\displaystyle= F(G0,Nρ)G0,N(Fρ)L22,\displaystyle\|F(G_{0,N}*\rho)-G_{0,N}*(F\rho)\|_{L^{2}}^{2},

where in the last inequality we used the mass conservation for ρN,(XN)\rho_{N,\hbar}(X_{N}). By the triangle, Hölder inequalities and Lemma A.1 we get

A2\displaystyle A_{2}\leq 2F(G0,Nρ)FρL22+2FρG0,N(Fρ)L22\displaystyle 2\|F(G_{0,N}*\rho)-F\rho\|_{L^{2}}^{2}+2\|F\rho-G_{0,N}*(F\rho)\|_{L^{2}}^{2}
\displaystyle\lesssim FL2(G0,Nδ)ρL22+(G0,Nδ)(Fρ)L22\displaystyle\|F\|_{L^{\infty}}^{2}\|(G_{0,N}-\delta)*\rho\|_{L^{2}}^{2}+\|(G_{0,N}-\delta)*(F\rho)\|_{L^{2}}^{2}
\displaystyle\lesssim 1N2ηFL2ρL22+1N2η(Fρ)L22N2η.\displaystyle\frac{1}{N^{2\eta}}\|F\|_{L^{\infty}}^{2}\|\langle\nabla\rangle\rho\|_{L^{2}}^{2}+\frac{1}{N^{2\eta}}\|\langle\nabla\rangle(F\rho)\|_{L^{2}}^{2}\lesssim N^{-2\eta}.

To sum up, we complete the estimates for the term AA and reach

(4.28) AA11+A12+A2Nη12+Nη2Nη2,\displaystyle A\leq\sqrt{A_{11}}+\sqrt{A_{12}}+\sqrt{A_{2}}\lesssim N^{\frac{\eta-1}{2}}+N^{-\eta}\hbar^{-2}\lesssim N^{-\eta}\hbar^{-2},

where in the last inequality we used that Nη12NηN^{\frac{\eta-1}{2}}\leq N^{-\eta} for η<13\eta<\frac{1}{3}.

For the term BB, we rewrite

B=\displaystyle B= F(z)G0,N(xz)G0,N(zy)νXN(dx)νXN(dy)𝑑zρN,(XN)𝑑XN\displaystyle\int F(z)G_{0,N}(x-z)G_{0,N}(z-y)\nu_{X_{N}}(dx)\nu_{X_{N}}(dy)dz\rho_{N,\hbar}(X_{N})dX_{N}
=\displaystyle= F(z)|G0,NνXN(z)|2ρN,(XN)𝑑z𝑑XN.\displaystyle\int F(z)|G_{0,N}*\nu_{X_{N}}(z)|^{2}\rho_{N,\hbar}(X_{N})dzdX_{N}.

Observe that

|G0,NνXN(z)|2ρN,(XN)0\displaystyle|G_{0,N}*\nu_{X_{N}}(z)|^{2}\rho_{N,\hbar}(X_{N})\geq 0

for a.e. (z,XN)3×3N(z,X_{N})\in\mathbb{R}^{3}\times\mathbb{R}^{3N}. Therefore, we can directly bound F(z)F(z) and get

(4.29) B\displaystyle B\leq FL|G0,NνXN(z)|2ρN,(XN)𝑑z𝑑XN\displaystyle\|F\|_{L^{\infty}}\int|G_{0,N}*\nu_{X_{N}}(z)|^{2}\rho_{N,\hbar}(X_{N})dzdX_{N}
=\displaystyle= FLG0,N(xz)G0,N(zy)𝑑zνXN(dx)νXN(dy)ρN,(XN)𝑑XN\displaystyle\|F\|_{L^{\infty}}\int\int G_{0,N}(x-z)G_{0,N}(z-y)dz\nu_{X_{N}}(dx)\nu_{X_{N}}(dy)\rho_{N,\hbar}(X_{N})dX_{N}
=\displaystyle= FLGN(xy)νXN(dx)νXN(dy)ρN,(XN)𝑑XN\displaystyle\|F\|_{L^{\infty}}\int G_{N}(x-y)\nu_{X_{N}}(dx)\nu_{X_{N}}(dy)\rho_{N,\hbar}(X_{N})dX_{N}
=\displaystyle= FLGN(xy)[N1NρN,(2)(x,y)ρN,(1)(x)ρ(y)ρ(x)ρN,(1)(y)+ρ(x)ρ(y)]𝑑x𝑑y\displaystyle\|F\|_{L^{\infty}}\int G_{N}(x-y)\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(x,y)-\rho_{N,\hbar}^{(1)}(x)\rho(y)-\rho(x)\rho_{N,\hbar}^{(1)}(y)+\rho(x)\rho(y)\right]dxdy
+FLGN(0)NρN,(1)(x)𝑑x,\displaystyle+\|F\|_{L^{\infty}}\frac{G_{N}(0)}{N}\int\rho_{N,\hbar}^{(1)}(x)dx,

where in the second-to-last equality we used the property (4.24), and in the last equality we used the equation (4.22).

With the approximation forms (4.21) and (4.25), we use estimates (4.23), (4.28) for AA and (4.29) for BB to arrive at

F(x)GN(xy)[N1NρN,(2)(x,y)ρN,(1)(x)ρ(y)ρ(x)ρN,(1)(y)+ρ(x)ρ(y)]𝑑x𝑑y\displaystyle\int F(x)G_{N}(x-y)\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(x,y)-\rho_{N,\hbar}^{(1)}(x)\rho(y)-\rho(x)\rho_{N,\hbar}^{(1)}(y)+\rho(x)\rho(y)\right]dxdy
=\displaystyle= A+B+O(N3η1)\displaystyle A+B+O(N^{3\eta-1})
\displaystyle\leq FLGN(xy)[N1NρN,(2)(x,y)ρN,(1)(x)ρ(y)ρ(x)ρN,(1)(y)+ρ(x)ρ(y)]𝑑x𝑑y\displaystyle\|F\|_{L^{\infty}}\int G_{N}(x-y)\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(x,y)-\rho_{N,\hbar}^{(1)}(x)\rho(y)-\rho(x)\rho_{N,\hbar}^{(1)}(y)+\rho(x)\rho(y)\right]dxdy
+O(Nη2+N3η1),\displaystyle+O(N^{-\eta}\hbar^{-2}+N^{3\eta-1}),

which is the desired estimate (4.19). ∎

To prove the lower bound estimate (4.2) for δ(t)\mathcal{F}_{\delta}(t), we give the following estimate.

Lemma 4.5.

Let η<13\eta<\frac{1}{3} to be determined, we have

(4.30) GN(xy)[N1NρN,(2)(x,y)ρN,(1)(x)ρ(y)ρ(x)ρN,(1)(y)+ρ(x)ρ(y)]𝑑x𝑑y\displaystyle\int G_{N}(x-y)\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(x,y)-\rho_{N,\hbar}^{(1)}(x)\rho(y)-\rho(x)\rho_{N,\hbar}^{(1)}(y)+\rho(x)\rho(y)\right]dxdy
\displaystyle\geq GN(xy)(ρN,(1)(x)ρ(x))(ρN,(1)(y)ρ(y))𝑑x𝑑yO(N3η1).\displaystyle\int G_{N}(x-y)(\rho_{N,\hbar}^{(1)}(x)-\rho(x))(\rho_{N,\hbar}^{(1)}(y)-\rho(y))dxdy-O(N^{3\eta-1}).
Proof.

We decompose

GN(xy)[N1NρN,(2)(x,y)ρN,(1)(x)ρ(y)ρ(x)ρN,(1)(y)+ρ(x)ρ(y)]𝑑x𝑑y\displaystyle\int G_{N}(x-y)\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(x,y)-\rho_{N,\hbar}^{(1)}(x)\rho(y)-\rho(x)\rho_{N,\hbar}^{(1)}(y)+\rho(x)\rho(y)\right]dxdy
=\displaystyle= I+II,\displaystyle I+II,

where

(4.31) I=\displaystyle I= GN(xy)(ρN,(1)(x)ρ(x))(ρN,(1)(y)ρ(y))𝑑x𝑑y,\displaystyle\int G_{N}(x-y)(\rho_{N,\hbar}^{(1)}(x)-\rho(x))(\rho_{N,\hbar}^{(1)}(y)-\rho(y))dxdy,
(4.32) II=\displaystyle II= GN(xy)[N1NρN,(2)(x,y)ρN,(1)(x)ρN,(1)(y)]𝑑x𝑑y.\displaystyle\int G_{N}(x-y)\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(x,y)-\rho_{N,\hbar}^{(1)}(x)\rho_{N,\hbar}^{(1)}(y)\right]dxdy.

It suffices to prove a lower bound of the term IIII. By the symmetry of the density function ρN,(XN)\rho_{N,\hbar}(X_{N}), we rewrite

II=\displaystyle II= GN(xy)[N1NρN,(2)(x,y)ρN,(1)(x)ρN,(1)(y)]𝑑x𝑑y\displaystyle\int G_{N}(x-y)\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(x,y)-\rho_{N,\hbar}^{(1)}(x)\rho_{N,\hbar}^{(1)}(y)\right]dxdy
=\displaystyle= GN(xy)[1N2ijNδxi(x)δxj(y)+ρN,(1)(x)ρN,(1)(y)\displaystyle\int G_{N}(x-y)\left[\frac{1}{N^{2}}\sum_{i\neq j}^{N}\delta_{x_{i}}(x)\delta_{x_{j}}(y)+\rho_{N,\hbar}^{(1)}(x)\rho_{N,\hbar}^{(1)}(y)\right.
1Ni=1Nδxi(x)ρN,(1)(y)ρN,(1)(x)1Nj=1Nδxj(y)]dxdyρN,(XN)dXN\displaystyle\left.-\frac{1}{N}\sum_{i=1}^{N}\delta_{x_{i}}(x)\rho_{N,\hbar}^{(1)}(y)-\rho_{N,\hbar}^{(1)}(x)\frac{1}{N}\sum_{j=1}^{N}\delta_{x_{j}}(y)\right]dxdy\rho_{N,\hbar}(X_{N})dX_{N}
=\displaystyle= GN(xy)μXN(dx)μXN(dy)ρN,(XN)𝑑XNGN(0)NρN,(1)(x)𝑑x,\displaystyle\int G_{N}(x-y)\mu_{X_{N}}(dx)\mu_{X_{N}}(dy)\rho_{N,\hbar}(X_{N})dX_{N}-\frac{G_{N}(0)}{N}\int\rho_{N,\hbar}^{(1)}(x)dx,

where

μXN(dx)=1Ni=1Nδxi(dx)ρN,(1)(x)dx.\displaystyle\mu_{X_{N}}(dx)=\frac{1}{N}\sum_{i=1}^{N}\delta_{x_{i}}(dx)-\rho_{N,\hbar}^{(1)}(x)dx.

Then by (4.24) and (4.23), we obtain

II=\displaystyle II= |G0,NμXN(z)|2ρN,(XN)𝑑z𝑑XNGN(0)NρN,(1)(x)𝑑xN3η1,\displaystyle\int|G_{0,N}*\mu_{X_{N}}(z)|^{2}\rho_{N,\hbar}(X_{N})dzdX_{N}-\frac{G_{N}(0)}{N}\int\rho_{N,\hbar}^{(1)}(x)dx\gtrsim-N^{3\eta-1},

which completes the proof of estimate (4.30). ∎

To the end, we get into the proof of Proposition 4.1.

Proof of Proposition 4.1.

For estimate (4.1), the approximation (4.15) of ~δ(t)\widetilde{\mathcal{F}}_{\delta}(t) in Lemma 4.3 gives

~δ=\displaystyle\widetilde{\mathcal{F}}_{\delta}= b0divu(x)GN(xy)[N1NρN,(2)(x,y)\displaystyle-b_{0}\int\operatorname{div}u(x)G_{N}(x-y)\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(x,y)\right.
ρN,(1)(x)ρ(y)ρ(x)ρN,(1)(y)+ρ(x)ρ(y)]dxdy\displaystyle\left.\quad\quad-\rho_{N,\hbar}^{(1)}(x)\rho(y)-\rho(x)\rho_{N,\hbar}^{(1)}(y)+\rho(x)\rho(y)\right]dxdy
±O(Nβ16+Nβ34+Nη34).\displaystyle\pm O(N^{\beta-1}\hbar^{-6}+N^{-\frac{\beta}{3}}\hbar^{-4}+N^{-\frac{\eta}{3}}\hbar^{-4}).

Then by the functional inequality (4.19) in Lemma 4.4, we get

~δ\displaystyle\widetilde{\mathcal{F}}_{\delta}\leq divuLGN(xy)[N1NρN,(2)(x,y)\displaystyle\|\operatorname{div}u\|_{L^{\infty}}\int G_{N}(x-y)\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(x,y)\right.
ρN,(1)(x)ρ(y)ρ(x)ρN,(1)(y)+ρ(x)ρ(y)]dxdy\displaystyle\left.-\rho_{N,\hbar}^{(1)}(x)\rho(y)-\rho(x)\rho_{N,\hbar}^{(1)}(y)+\rho(x)\rho(y)\right]dxdy
+O(Nβ16+Nβ34+Nη34+Nη2+N3η1).\displaystyle+O(N^{\beta-1}\hbar^{-6}+N^{-\frac{\beta}{3}}\hbar^{-4}+N^{-\frac{\eta}{3}}\hbar^{-4}+N^{-\eta}\hbar^{-2}+N^{3\eta-1}).

Using the approximation (4.14) of δ\mathcal{F}_{\delta} in Lemma 4.3, we arrive at

~δ\displaystyle\widetilde{\mathcal{F}}_{\delta}\leq divuLδ+O(Nβ16+Nβ34+Nη34+Nη2+N3η1)\displaystyle\|\operatorname{div}u\|_{L^{\infty}}\mathcal{F}_{\delta}+O(N^{\beta-1}\hbar^{-6}+N^{-\frac{\beta}{3}}\hbar^{-4}+N^{-\frac{\eta}{3}}\hbar^{-4}+N^{-\eta}\hbar^{-2}+N^{3\eta-1})
\displaystyle\lesssim δ+O(Nβ16+Nβ34+N1104),\displaystyle\mathcal{F}_{\delta}+O(N^{\beta-1}\hbar^{-6}+N^{-\frac{\beta}{3}}\hbar^{-4}+N^{-\frac{1}{10}}\hbar^{-4}),

where in the last inequality we took η=310\eta=\frac{3}{10}. Therefore, we complete the proof of the estimate (4.1).

For the lower bound estimate (4.2) on δ\mathcal{F}_{\delta}, we use the approximation (4.14) of δ\mathcal{F}_{\delta} and estimate (4.30) to obtain

δ=\displaystyle\mathcal{F}_{\delta}= b0GN(xy)[N1NρN,(2)(x,y)\displaystyle b_{0}\int G_{N}(x-y)\left[\frac{N-1}{N}\rho_{N,\hbar}^{(2)}(x,y)\right.
ρN,(1)(x)ρ(y)ρ(x)ρN,(1)(y)+ρ(x)ρ(y)]dxdy\displaystyle\quad\quad\left.-\rho_{N,\hbar}^{(1)}(x)\rho(y)-\rho(x)\rho_{N,\hbar}^{(1)}(y)+\rho(x)\rho(y)\right]dxdy
±O(Nβ16+Nβ34+Nη34)\displaystyle\pm O(N^{\beta-1}\hbar^{-6}+N^{-\frac{\beta}{3}}\hbar^{-4}+N^{-\frac{\eta}{3}}\hbar^{-4})
(4.33) \displaystyle\geq GN(xy)(ρN,(1)(x)ρ(x))(ρN,(1)(y)ρ(y))𝑑x𝑑y\displaystyle\int G_{N}(x-y)(\rho_{N,\hbar}^{(1)}(x)-\rho(x))(\rho_{N,\hbar}^{(1)}(y)-\rho(y))dxdy
O(Nβ16+Nβ34+Nη34+N3η1).\displaystyle-O(N^{\beta-1}\hbar^{-6}+N^{-\frac{\beta}{3}}\hbar^{-4}+N^{-\frac{\eta}{3}}\hbar^{-4}+N^{3\eta-1}).

By (4.24), we observe that the term on the r.h.s of (4.33)

GN(xy)(ρN,(1)(x)ρ(x))(ρN,(1)(y)ρ(y))𝑑x𝑑y\displaystyle\int G_{N}(x-y)(\rho_{N,\hbar}^{(1)}(x)-\rho(x))(\rho_{N,\hbar}^{(1)}(y)-\rho(y))dxdy
=\displaystyle= |GN,0(ρN,(1)ρ)(z)|2𝑑z0.\displaystyle\int|G_{N,0}*(\rho_{N,\hbar}^{(1)}-\rho)(z)|^{2}dz\geq 0.

Thus, we can discard this positive term and then take η=310\eta=\frac{3}{10} to get

δO(Nβ16+Nβ34+N1104),\displaystyle\mathcal{F}_{\delta}\geq-O(N^{\beta-1}\hbar^{-6}+N^{-\frac{\beta}{3}}\hbar^{-4}+N^{-\frac{1}{10}}\hbar^{-4}),

which is the lower bound estimate (4.2). ∎

5. Quantitative Strong Convergence of Quantum Densities

In the section, using functional inequalities, we prove the Gronwall’s inequality for the modulated energy. Subsequently, with the quantitative convergence rate of the modulated energy, we further conclude the quantitative strong convergence of quantum mass and momentum densities. Notably, the δ\delta-type potential part is crucial in upgrading to the quantitative strong convergence, that is, in the case of only the Coulomb potential, one cannot deduce the strong convergence here.

Recall the modulated energy

(5.1) (t)=K(t)+P(t),\displaystyle\mathcal{M}(t)=\mathcal{M}_{K}(t)+\mathcal{M}_{P}(t),

where the kinetic energy part is

(5.2) K(t)=3N|(ix1u(t,x1))ψN,(t,XN)|2𝑑XN,\displaystyle\mathcal{M}_{K}(t)=\int_{\mathbb{R}^{3N}}|\left(i\hbar\nabla_{x_{1}}-u(t,x_{1})\right)\psi_{N,\hbar}(t,X_{N})|^{2}dX_{N},

and the potential energy part is

(5.3) P(t)=δ(t)+c(t).\displaystyle\mathcal{M}_{P}(t)=\mathcal{F}_{\delta}(t)+\mathcal{F}_{c}(t).

From lower bound estimates (4.2) on δ(t)\mathcal{F}_{\delta}(t) and (1.25) on c(t)\mathcal{F}_{c}(t), we can add a small compensation such that

(5.4) δ(t)+c(t)+r(N,)0,\displaystyle\mathcal{F}_{\delta}(t)+\mathcal{F}_{c}(t)+r(N,\hbar)\geq 0,

where r(N,)=C(Nβ16+Nβ34+N1104).r(N,\hbar)=C(N^{\beta-1}\hbar^{-6}+N^{-\frac{\beta}{3}}\hbar^{-4}+N^{-\frac{1}{10}}\hbar^{-4}). Thus, we introduce the positive modulated energy

(5.5) +(t)=\displaystyle\mathcal{M}^{+}(t)= (t)+2r(N,)r(N,)0.\displaystyle\mathcal{M}(t)+2r(N,\hbar)\geq r(N,\hbar)\geq 0.

We now provide a closed estimate for the positive modulated energy.

Proposition 5.1.

For t[0,T0]t\in[0,T_{0}], we have the differential inequality

(5.6) ddt+(t)+(t)+2.\displaystyle\frac{d}{dt}\mathcal{M}^{+}(t)\lesssim\mathcal{M}^{+}(t)+\hbar^{2}.

Moreover, we conclude

(5.7) 3N|(ix1u(t,x1))ψN,(t,XN)|2𝑑XN+(0)+2,\int_{\mathbb{R}^{3N}}|\left(i\hbar\nabla_{x_{1}}-u(t,x_{1})\right)\psi_{N,\hbar}(t,X_{N})|^{2}dX_{N}\lesssim\mathcal{M}^{+}(0)+\hbar^{2},

and

(5.8) GN(xy)(ρN,(1)(x)ρ(x))(ρN,(1)(y)ρ(y))𝑑x𝑑y+(0)+2.\int G_{N}(x-y)(\rho_{N,\hbar}^{(1)}(x)-\rho(x))(\rho_{N,\hbar}^{(1)}(y)-\rho(y))dxdy\lesssim\mathcal{M}^{+}(0)+\hbar^{2}.
Proof.

From the evolution of the modulated energy (2.8), we find that

ddt+(t)\displaystyle\frac{d}{dt}\mathcal{M}^{+}(t)
=\displaystyle= j,k=133N(juk+kuj)(ijψN,ujψN,)(ikψN,ukψN,)¯𝑑XN\displaystyle-\sum_{j,k=1}^{3}\int_{\mathbb{R}^{3N}}\left(\partial_{j}u^{k}+\partial_{k}u^{j}\right)(-i\hbar\partial_{j}\psi_{N,\hbar}-u^{j}\psi_{N,\hbar})\overline{(-i\hbar\partial_{k}\psi_{N,\hbar}-u^{k}\psi_{N,\hbar})}dX_{N}
+223Δ(divu)(t,x1)ρN,(1)(t,x1)𝑑x1+~δ(t)+~c(t)\displaystyle+\frac{\hbar^{2}}{2}\int_{\mathbb{R}^{3}}\Delta(\operatorname{div}u)(t,x_{1})\rho_{N,\hbar}^{(1)}(t,x_{1})dx_{1}+\widetilde{\mathcal{F}}_{\delta}(t)+\widetilde{\mathcal{F}}_{c}(t)
\displaystyle\lesssim uL3N|(ix1u)ψN,|2𝑑XN+2ψN,L22ΔdivuL+~δ(t)+~c(t).\displaystyle\|\nabla u\|_{L^{\infty}}\int_{\mathbb{R}^{3N}}|(i\hbar\nabla_{x_{1}}-u)\psi_{N,\hbar}|^{2}dX_{N}+\hbar^{2}\|\psi_{N,\hbar}\|_{L^{2}}^{2}\|\Delta\operatorname{div}u\|_{L^{\infty}}+\widetilde{\mathcal{F}}_{\delta}(t)+\widetilde{\mathcal{F}}_{c}(t).

By the functional inequalities (4.1)(\ref{equ:functional inequality,fdelta}) on ~δ(t)\widetilde{\mathcal{F}}_{\delta}(t) and (1.24) on ~c(t)\widetilde{\mathcal{F}}_{c}(t), we get

(5.9) ddt+(t)+(t)+2.\displaystyle\frac{d}{dt}\mathcal{M}^{+}(t)\lesssim\mathcal{M}^{+}(t)+\hbar^{2}.

Then by Gronwall’s inequality, we arrive at

(5.10) +(t)exp(CT0)(+(0)+2t)\displaystyle\mathcal{M}^{+}(t)\leq\exp(CT_{0})\left(\mathcal{M}^{+}(0)+\hbar^{2}t\right)\lesssim +(0)+2\displaystyle\mathcal{M}^{+}(0)+\hbar^{2}

for t[0,T0]t\in[0,T_{0}].

For the kinetic energy estimate (5.7), by (5.5) and (5.10) we have that

3N|(ix1u(t,x1))ψN,(t,XN)|2𝑑XN\displaystyle\int_{\mathbb{R}^{3N}}|\left(i\hbar\nabla_{x_{1}}-u(t,x_{1})\right)\psi_{N,\hbar}(t,X_{N})|^{2}dX_{N}\leq +(t)+(0)+2,\displaystyle\mathcal{M}^{+}(t)\lesssim\mathcal{M}^{+}(0)+\hbar^{2},

which completes the proof of (5.7).

For the potential energy estimate (5.8), by (4.30) in Lemma 4.5 and (4.14) in Lemma 4.3, we have

GN(xy)(ρN,(1)(t,x)ρ(t,x))(ρN,(1)(t,y)ρ(t,y))𝑑x𝑑y\displaystyle\int G_{N}(x-y)(\rho_{N,\hbar}^{(1)}(t,x)-\rho(t,x))(\rho_{N,\hbar}^{(1)}(t,y)-\rho(t,y))dxdy\leq ~c(t)+r(N,)\displaystyle\widetilde{\mathcal{F}}_{c}(t)+r(N,\hbar)
\displaystyle\leq +(t)+r(N,).\displaystyle\mathcal{M}^{+}(t)+r(N,\hbar).

Again by (5.10), we arrive at (5.8). ∎

To the end, we get into the proof of Theorem 1.1.

Proof of Theorem 1.1.

Convergence of the mass density ρN,(1)(t)\rho_{N,\hbar}^{(1)}(t).

We decompose

GN(xy)(ρN,(1)(t,x)ρ(t,x))(ρN,(1)(t,y)ρ(t,y))𝑑x𝑑y\displaystyle\int G_{N}(x-y)(\rho_{N,\hbar}^{(1)}(t,x)-\rho(t,x))(\rho_{N,\hbar}^{(1)}(t,y)-\rho(t,y))dxdy
(5.11) =\displaystyle= (GNδ)(ρN,(1)(t)ρ(t)),ρN,(1)(t)ρ(t)+ρN,(1)(t)ρ(t)L22.\displaystyle\langle(G_{N}-\delta)*(\rho_{N,\hbar}^{(1)}(t)-\rho(t)),\rho_{N,\hbar}^{(1)}(t)-\rho(t)\rangle+\|\rho_{N,\hbar}^{(1)}(t)-\rho(t)\|_{L^{2}}^{2}.

For the first term on the r.h.s of (5.11), we use Hölder inequality and Lemma A.1 to obtain

(GNδ)(ρN,(1)(t)ρ(t)),ρN,(1)(t)ρ(t)\displaystyle\langle(G_{N}-\delta)*(\rho_{N,\hbar}^{(1)}(t)-\rho(t)),\rho_{N,\hbar}^{(1)}(t)-\rho(t)\rangle
\displaystyle\leq (GNδ)(ρN,(1)(t)ρ(t))L32ρN,(1)(t)ρ(t)L3\displaystyle\|(G_{N}-\delta)*(\rho_{N,\hbar}^{(1)}(t)-\rho(t))\|_{L^{\frac{3}{2}}}\|\rho_{N,\hbar}^{(1)}(t)-\rho(t)\|_{L^{3}}
\displaystyle\lesssim Nη(ρN,(1)(t)L32+ρ(t)L32)(ρN,(1)(t)L3+ρ(t)L3).\displaystyle N^{-\eta}\left(\|\langle\nabla\rangle\rho_{N,\hbar}^{(1)}(t)\|_{L^{\frac{3}{2}}}+\|\rho(t)\|_{L^{\frac{3}{2}}}\right)\left(\|\rho_{N,\hbar}^{(1)}(t)\|_{L^{3}}+\|\rho(t)\|_{L^{3}}\right).

Next, we estimate the terms ρN,(1)(t)L32\|\langle\nabla\rangle\rho_{N,\hbar}^{(1)}(t)\|_{L^{\frac{3}{2}}} and ρN,(1)(t)L3\|\rho_{N,\hbar}^{(1)}(t)\|_{L^{3}}. By the Calderón-Zygmund theory which implies that fLpfLp+fLp\|\langle\nabla\rangle f\|_{L^{p}}\lesssim\|\nabla f\|_{L^{p}}+\|f\|_{L^{p}} for 1<p<1<p<\infty, we get

ρN,(1)(t)L32ρN,(1)(t)L32+ρN,(1)(t)L32.\displaystyle\|\langle\nabla\rangle\rho_{N,\hbar}^{(1)}(t)\|_{L^{\frac{3}{2}}}\lesssim\|\nabla\rho_{N,\hbar}^{(1)}(t)\|_{L^{\frac{3}{2}}}+\|\rho_{N,\hbar}^{(1)}(t)\|_{L^{\frac{3}{2}}}.

By the Leibniz rule, Minkowski, Hölder, and Sobolev inequalities, we then obtain

ρN,(1)(t)L32x1ψN,(t)L2ψN,(t)L2Lx16x1ψN,(t)L222,\displaystyle\|\nabla\rho_{N,\hbar}^{(1)}(t)\|_{L^{\frac{3}{2}}}\lesssim\|\nabla_{x_{1}}\psi_{N,\hbar}(t)\|_{L^{2}}\|\psi_{N,\hbar}(t)\|_{L^{2}L_{x_{1}}^{6}}\lesssim\|\langle\nabla_{x_{1}}\rangle\psi_{N,\hbar}(t)\|_{L^{2}}^{2}\lesssim\hbar^{-2},

where in the last inequality we have used the H1H^{1} energy bound (2.5) for ψN,\psi_{N,\hbar}. Similarly, we also have

ρN,(1)(t)L32ψN,(t)L2ψN,(t)L2Lx16ψN,(t)L2x1ψN,(t)L21,\displaystyle\|\rho_{N,\hbar}^{(1)}(t)\|_{L^{\frac{3}{2}}}\lesssim\|\psi_{N,\hbar}(t)\|_{L^{2}}\|\psi_{N,\hbar}(t)\|_{L^{2}L_{x_{1}}^{6}}\lesssim\|\psi_{N,\hbar}(t)\|_{L^{2}}\|\langle\nabla_{x_{1}}\rangle\psi_{N,\hbar}(t)\|_{L^{2}}\lesssim\hbar^{-1},

and

ρN,(1)(t)L3ψN,(t)L2Lx162x1ψN,(t)L222.\displaystyle\|\rho_{N,\hbar}^{(1)}(t)\|_{L^{3}}\leq\|\psi_{N,\hbar}(t)\|_{L^{2}L_{x_{1}}^{6}}^{2}\lesssim\|\langle\nabla_{x_{1}}\rangle\psi_{N,\hbar}(t)\|_{L^{2}}^{2}\lesssim\hbar^{-2}.

With η=310\eta=\frac{3}{10}, these bounds give that

(5.12) (GNδ)(ρN,(1)(t)ρ(t)),ρN,(1)(t)ρ(t)N3104r(N,).\displaystyle\langle(G_{N}-\delta)*(\rho_{N,\hbar}^{(1)}(t)-\rho(t)),\rho_{N,\hbar}^{(1)}(t)-\rho(t)\rangle\lesssim N^{-\frac{3}{10}}\hbar^{-4}\lesssim r(N,\hbar).

Thus, combining (5.11), (5.12) with (5.8), we arrive at

(5.13) ρN,(1)(t)ρ(t)L22\displaystyle\|\rho_{N,\hbar}^{(1)}(t)-\rho(t)\|_{L^{2}}^{2}\lesssim r(N,)++(0)+2+(0)+2,\displaystyle r(N,\hbar)+\mathcal{M}^{+}(0)+\hbar^{2}\lesssim\mathcal{M}^{+}(0)+\hbar^{2},

where in the last inequality we have used that r(N,)+(0)r(N,\hbar)\leq\mathcal{M}^{+}(0).

Convergence of the momentum density JN,(1)(t)J_{N,\hbar}^{(1)}(t).

Recall the momentum density

JN,(1)(t,x1)=Im(ψN,¯x1ψN,)(t,XN)𝑑x2𝑑xN.\displaystyle J_{N,\hbar}^{(1)}(t,x_{1})=\hbar\int\operatorname{Im}(\overline{\psi_{N,\hbar}}\nabla_{x_{1}}\psi_{N,\hbar})(t,X_{N})dx_{2}\cdot\cdot\cdot dx_{N}.

Then by the triangle and Hölder’s inequalities, we have

JN,(1)(t)(ρu)(t)L1\displaystyle\|J_{N,\hbar}^{(1)}(t)-(\rho u)(t)\|_{L^{1}}
\displaystyle\leq JN,(1)(t)(ρN,(1)u)(t)L1+(ρN,(1)u)(t)(ρu)(t)L1\displaystyle\|J_{N,\hbar}^{(1)}(t)-(\rho_{N,\hbar}^{(1)}u)(t)\|_{L^{1}}+\|(\rho_{N,\hbar}^{(1)}u)(t)-(\rho u)(t)\|_{L^{1}}
=\displaystyle= Im(ψN,(t)¯(x1iu(t))ψN,(t))L1+(ρN,(1)u)(t)(ρu)(t)L1\displaystyle\|\operatorname{Im}\left(\overline{\psi_{N,\hbar}(t)}\left(\hbar\nabla_{x_{1}}-iu(t)\right)\psi_{N,\hbar}(t)\right)\|_{L^{1}}+\|(\rho_{N,\hbar}^{(1)}u)(t)-(\rho u)(t)\|_{L^{1}}
\displaystyle\leq ψN,(t)L2(ix1u(t))ψN,(t)L2+u(t)L2ρN,(1)(t)ρ(t)L2\displaystyle\|\psi_{N,\hbar}(t)\|_{L^{2}}\|(i\hbar\nabla_{x_{1}}-u(t))\psi_{N,\hbar}(t)\|_{L^{2}}+\|u(t)\|_{L^{2}}\|\rho_{N,\hbar}^{(1)}(t)-\rho(t)\|_{L^{2}}
\displaystyle\lesssim +(0)+2,\displaystyle\mathcal{M}^{+}(0)+\hbar^{2},

where in the last inequality we used the mass conservation, estimates (5.7)(\ref{equ:convergence of kinetic energy}) and (5.13)(\ref{equ:convergence of mass density,L2}). ∎

Acknowledgements. X. Chen was supported in part by NSF grant DMS-2005469 and a Simons fellowship numbered 916862, S. Shen was supported in part by the Postdoctoral Science Foundation of China under Grant 2022M720263, and Z. Zhang was supported in part by NSF of China under Grant 12171010 and 12288101.

Appendix A Sobolev Type Estimates

Lemma A.1 ([21], Lemma A.5).

Let d=3d=3 and WN(x)=N3βV(Nβx)b0δW_{N}(x)=N^{3\beta}V(N^{\beta}x)-b_{0}\delta, where b0=V(x)𝑑xb_{0}=\int V(x)dx. For any 0s10\leq s\leq 1,

(A.1) WNfLpCxV(x)L1NβssfLp\displaystyle\|W_{N}*f\|_{L^{p}}\leq C\|\langle x\rangle V(x)\|_{L^{1}}N^{-\beta s}\|\langle\nabla\rangle^{s}f\|_{L^{p}}

for any 1<p<1<p<\infty.

Lemma A.2 ([29], Lemma A.3).

Let d=3d=3 and VN(x)=N3βV(Nβx)V_{N}(x)=N^{3\beta}V(N^{\beta}x). Then

(A.2) VN(x1x2)CVL1(1Δx1)(1Δx2),\displaystyle V_{N}(x_{1}-x_{2})\leq C\|V\|_{L^{1}}(1-\Delta_{x_{1}})(1-\Delta_{x_{2}}),
(A.3) VN(x1x2)CNβVL32(1Δx1),\displaystyle V_{N}(x_{1}-x_{2})\leq CN^{\beta}\|V\|_{L^{\frac{3}{2}}}(1-\Delta_{x_{1}}),
(A.4) VN(x1x2)CN3βVL.\displaystyle V_{N}(x_{1}-x_{2})\leq CN^{3\beta}\|V\|_{L^{\infty}}.
Lemma A.3.

Suppose that fL1f\in L^{1} such that

|f(x)||x|12𝑑x<.\int|f(x)||x|^{\frac{1}{2}}dx<\infty.

Let fε(x)=ε3f(εx)f_{\varepsilon}(x)=\varepsilon^{3}f(\varepsilon x) and d0=f𝑑xd_{0}=\int fdx, then we have

|(fε(xy)d0δ(xy))φ,ψ|\displaystyle|\langle(f_{\varepsilon}(x-y)-d_{0}\delta(x-y))\varphi,\psi\rangle|
\displaystyle\lesssim εθ(1Δx)(1Δy)φ,φ12(1Δx)(1Δy)ψ,ψ12\displaystyle\varepsilon^{\theta}\langle(1-\Delta_{x})(1-\Delta_{y})\varphi,\varphi\rangle^{\frac{1}{2}}\langle(1-\Delta_{x})(1-\Delta_{y})\psi,\psi\rangle^{\frac{1}{2}}

for θ(0,12)\theta\in(0,\frac{1}{2}).

Proof.

For the derivation of NLS, this Poincare´\acute{e} type inequality is usually used in the convergence part of the hierarchy method. See, for example, [29, 30, 31, 46]. For completeness, we here include a proof. Without loss of generality, we might as well assume that d0=f𝑑x=1d_{0}=\int fdx=1. Switching to Fourier space, we observe that

φ,(fε(xy)δ(xy))ψ\displaystyle\langle\varphi,(f_{\varepsilon}(x-y)-\delta(x-y))\psi\rangle
=\displaystyle= 𝑑x𝑑p𝑑ξ1𝑑ξ2φ^(ξ1,ξ2)ψ^¯(ξ1+p,ξ2p)f(x)(eiεpx1).\displaystyle\int dxdpd\xi_{1}d\xi_{2}\widehat{\varphi}(\xi_{1},\xi_{2})\overline{\widehat{\psi}}(\xi_{1}+p,\xi_{2}-p)f(x)(e^{i\varepsilon p\cdot x}-1).

By using |eia1|min{a,2}2aθ|e^{ia}-1|\leq\min\left\{a,2\right\}\leq 2a^{\theta} for θ(0,1)\theta\in(0,1) and |p|θξ1θ+ξ1+pθ|p|^{\theta}\leq\langle\xi_{1}\rangle^{\theta}+\langle\xi_{1}+p\rangle^{\theta}, we have

φ,(fε(xy)δ(xy))ψ\displaystyle\langle\varphi,(f_{\varepsilon}(x-y)-\delta(x-y))\psi\rangle
\displaystyle\leq 2εθf(x)|x|θ𝑑x𝑑p𝑑ξ1𝑑ξ2φ^(ξ1,ξ2)ψ^¯(ξ1+p,ξ2p)|p|θ\displaystyle 2\varepsilon^{\theta}\int f(x)|x|^{\theta}dx\int dpd\xi_{1}d\xi_{2}\widehat{\varphi}(\xi_{1},\xi_{2})\overline{\widehat{\psi}}(\xi_{1}+p,\xi_{2}-p)|p|^{\theta}
\displaystyle\leq 2εθf(x)|x|θ𝑑x𝑑p𝑑ξ1𝑑ξ2φ^(ξ1,ξ2)ψ^¯(ξ1+p,ξ2p)(ξ1θ+ξ1+pθ).\displaystyle 2\varepsilon^{\theta}\int f(x)|x|^{\theta}dx\int dpd\xi_{1}d\xi_{2}\widehat{\varphi}(\xi_{1},\xi_{2})\overline{\widehat{\psi}}(\xi_{1}+p,\xi_{2}-p)\left(\langle\xi_{1}\rangle^{\theta}+\langle\xi_{1}+p\rangle^{\theta}\right).

It suffices to bound the term containing ξ1θ\langle\xi_{1}\rangle^{\theta}, as the term containing ξ1pθ\langle\xi_{1}-p\rangle^{\theta} can be estimated similarly. We rewrite

𝑑p𝑑ξ1𝑑ξ2φ^(ξ1,ξ2)ψ^¯(ξ1+p,ξ2p)ξ1θ\displaystyle\int dpd\xi_{1}d\xi_{2}\widehat{\varphi}(\xi_{1},\xi_{2})\overline{\widehat{\psi}}(\xi_{1}+p,\xi_{2}-p)\langle\xi_{1}\rangle^{\theta}
=\displaystyle= 𝑑p𝑑ξ1𝑑ξ2φ^(ξ1,ξ2)ψ^¯(ξ1+p,ξ2p)ξ1ξ2ξ1+pξ2pξ11θξ2ξ1+pξ2p\displaystyle\int dpd\xi_{1}d\xi_{2}\widehat{\varphi}(\xi_{1},\xi_{2})\overline{\widehat{\psi}}(\xi_{1}+p,\xi_{2}-p)\frac{\langle\xi_{1}\rangle\langle\xi_{2}\rangle\langle\xi_{1}+p\rangle\langle\xi_{2}-p\rangle}{\langle\xi_{1}\rangle^{1-\theta}\langle\xi_{2}\rangle\langle\xi_{1}+p\rangle\langle\xi_{2}-p\rangle}

By Cauchy-Schwarz inequality,

𝑑p𝑑ξ1𝑑ξ2φ^(ξ1,ξ2)ψ^¯(ξ1+p,ξ2p)ξ1θ\displaystyle\int dpd\xi_{1}d\xi_{2}\widehat{\varphi}(\xi_{1},\xi_{2})\overline{\widehat{\psi}}(\xi_{1}+p,\xi_{2}-p)\langle\xi_{1}\rangle^{\theta}
\displaystyle\leq [𝑑p𝑑ξ1𝑑ξ2ξ12ξ22ξ1+p2ξ2p2|φ^(ξ1,ξ2)|2]12\displaystyle\left[\int dpd\xi_{1}d\xi_{2}\frac{\langle\xi_{1}\rangle^{2}\langle\xi_{2}\rangle^{2}}{\langle\xi_{1}+p\rangle^{2}\langle\xi_{2}-p\rangle^{2}}|\widehat{\varphi}(\xi_{1},\xi_{2})|^{2}\right]^{\frac{1}{2}}
[𝑑p𝑑ξ1𝑑ξ2ξ1+p2ξ2p2ξ122θξ22|ψ^(ξ1+p,ξ2p)|2]12\displaystyle\cdot\left[\int dpd\xi_{1}d\xi_{2}\frac{\langle\xi_{1}+p\rangle^{2}\langle\xi_{2}-p\rangle^{2}}{\langle\xi_{1}\rangle^{2-2\theta}\langle\xi_{2}\rangle^{2}}|\widehat{\psi}(\xi_{1}+p,\xi_{2}-p)|^{2}\right]^{\frac{1}{2}}
=\displaystyle= [𝑑p𝑑ξ1𝑑ξ2ξ12ξ22ξ1+p2ξ2p2|φ^(ξ1,ξ2)|2]12\displaystyle\left[\int dpd\xi_{1}d\xi_{2}\frac{\langle\xi_{1}\rangle^{2}\langle\xi_{2}\rangle^{2}}{\langle\xi_{1}+p\rangle^{2}\langle\xi_{2}-p\rangle^{2}}|\widehat{\varphi}(\xi_{1},\xi_{2})|^{2}\right]^{\frac{1}{2}}
[𝑑p𝑑ξ1𝑑ξ2ξ12ξ22ξ1+p22θξ2p2|ψ^(ξ1,ξ2)|2]12\displaystyle\cdot\left[\int dpd\xi_{1}d\xi_{2}\frac{\langle\xi_{1}\rangle^{2}\langle\xi_{2}\rangle^{2}}{\langle\xi_{1}+p\rangle^{2-2\theta}\langle\xi_{2}-p\rangle^{2}}|\widehat{\psi}(\xi_{1},\xi_{2})|^{2}\right]^{\frac{1}{2}}
\displaystyle\lesssim (1Δ1)(1Δ2)φ,φ12(1Δ1)(1Δ2)ψ,ψ12\displaystyle\langle(1-\Delta_{1})(1-\Delta_{2})\varphi,\varphi\rangle^{\frac{1}{2}}\langle(1-\Delta_{1})(1-\Delta_{2})\psi,\psi\rangle^{\frac{1}{2}}

where in the last inequality we used that

supξ1,ξ21ξ1p22θξ2p2𝑑p<\displaystyle\sup_{\xi_{1},\xi_{2}}\int\frac{1}{\langle\xi_{1}-p\rangle^{2-2\theta}\langle\xi_{2}-p\rangle^{2}}dp<\infty

for all 0θ<120\leq\theta<\frac{1}{2}. ∎

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] N. Benedikter, G. de Oliveira, and B. Schlein. Quantitative derivation of the Gross-Pitaevskii equation. Comm. Pure Appl. Math., 68(8):1399–1482, 2015.
  • [3] C. Boccato, S. Cenatiempo, and B. Schlein. Quantum many-body fluctuations around nonlinear Schrödinger dynamics. Ann. Henri Poincaré, 18(1):113–191, 2017.
  • [4] C. Brennecke and B. Schlein. Gross-Pitaevskii dynamics for Bose-Einstein condensates. Anal. PDE, 12(6):1513–1596, 2019.
  • [5] R. Carles. Semi-classical analysis for nonlinear Schrödinger equations. World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, 2008.
  • [6] T. Chen, C. Hainzl, N. Pavlović, and R. Seiringer. Unconditional uniqueness for the cubic Gross-Pitaevskii hierarchy via quantum de Finetti. Comm. Pure Appl. Math., 68(10):1845–1884, 2015.
  • [7] T. Chen and N. Pavlović. On the Cauchy problem for focusing and defocusing Gross-Pitaevskii hierarchies. Discrete Contin. Dyn. Syst., 27(2):715–739, 2010.
  • [8] 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.
  • [9] T. Chen and N. Pavlović. A new proof of existence of solutions for focusing and defocusing Gross-Pitaevskii hierarchies. Proc. Amer. Math. Soc., 141(1):279–293, 2013.
  • [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] T. Chen and N. Pavlović. Higher order energy conservation and global well-posedness of solutions for Gross-Pitaevskii hierarchies. Comm. Partial Differential Equations, 39(9):1597–1634, 2014.
  • [12] T. Chen, N. Pavlović, and N. Tzirakis. Energy conservation and blowup of solutions for focusing Gross-Pitaevskii hierarchies. Ann. Inst. H. Poincaré Anal. Non Linéaire, 27(5):1271–1290, 2010.
  • [13] X. Chen. Collapsing estimates and the rigorous derivation of the 2d cubic nonlinear Schrödinger equation with anisotropic switchable quadratic traps. J. Math. Pures Appl. (9), 98(4):450–478, 2012.
  • [14] X. Chen. On the rigorous derivation of the 3D cubic nonlinear Schrödinger equation with a quadratic trap. Arch. Ration. Mech. Anal., 210(2):365–408, 2013.
  • [15] X. Chen and J. Holmer. On the rigorous derivation of the 2D cubic nonlinear Schrödinger equation from 3D quantum many-body dynamics. Arch. Ration. Mech. Anal., 210(3):909–954, 2013.
  • [16] X. Chen and J. Holmer. Correlation structures, many-body scattering processes, and the derivation of the Gross-Pitaevskii hierarchy. Int. Math. Res. Not. IMRN, 2016(10):3051–3110, 2016.
  • [17] X. Chen and J. Holmer. Focusing quantum many-body dynamics: the rigorous derivation of the 1D focusing cubic nonlinear Schrödinger equation. Arch. Ration. Mech. Anal., 221(2):631–676, 2016.
  • [18] 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.
  • [19] X. Chen and J. Holmer. Focusing quantum many-body dynamics, II: The rigorous derivation of the 1D focusing cubic nonlinear Schrödinger equation from 3D. Anal. PDE, 10(3):589–633, 2017.
  • [20] 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.
  • [21] 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.
  • [22] X. Chen and J. Holmer. Unconditional uniqueness for the energy-critical nonlinear Schrödinger equation on 𝕋4\mathbb{T}^{4}. Forum Math. Pi, 10:Paper No. e3, 49, 2022.
  • [23] X. Chen, S. Shen, J. Wu, and Z. Zhang. The derivation of the compressible euler equation from quantum many-body dynamics. To appear in Peking Mathematical Journal.
  • [24] X. Chen, S. Shen, and Z. Zhang. Quantitative derivation of the Euler-Poisson equation from quantum many-body dynamics. To appear in Peking Mathematical Journal.
  • [25] X. Chen, S. Shen, and Z. Zhang. The unconditional uniqueness for the energy-supercritical NLS. Ann. PDE, 8(2):Paper No. 14, 82, 2022.
  • [26] X. Chen and P. Smith. On the unconditional uniqueness of solutions to the infinite radial Chern-Simons-Schrödinger hierarchy. Anal. PDE, 7(7):1683–1712, 2014.
  • [27] M. Duerinckx. Mean-field limits for some Riesz interaction gradient flows. SIAM J. Math. Anal., 48(3):2269–2300, 2016.
  • [28] 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.
  • [29] 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.
  • [30] 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.
  • [31] 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.
  • [32] 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.
  • [33] J. Fröhlich, A. Knowles, and S. Schwarz. On the mean-field limit of bosons with Coulomb two-body interaction. Comm. Math. Phys., 288(3):1023–1059, 2009.
  • [34] F. Golse and T. Paul. Mean-field and classical limit for the N-body quantum dynamics with Coulomb interaction. Comm. Pure Appl. Math., 2021.
  • [35] E. Grenier. Semiclassical limit of the nonlinear Schrödinger equation in small time. Proc. Amer. Math. Soc., 126(2):523–530, 1998.
  • [36] P. Gressman, V. Sohinger, and G. Staffilani. On the uniqueness of solutions to the periodic 3D Gross-Pitaevskii hierarchy. J. Funct. Anal., 266(7):4705–4764, 2014.
  • [37] M. Grillakis and M. Machedon. Pair excitations and the mean field approximation of interacting bosons, I. Comm. Math. Phys., 324(2):601–636, 2013.
  • [38] M. Grillakis and M. Machedon. Pair excitations and the mean field approximation of interacting bosons, II. Comm. Partial Differential Equations, 42(1):24–67, 2017.
  • [39] M. G. Grillakis, M. Machedon, and D. Margetis. Second-order corrections to mean field evolution of weakly interacting bosons. I. Comm. Math. Phys., 294(1):273–301, 2010.
  • [40] M. G. Grillakis, M. Machedon, and D. Margetis. Second-order corrections to mean field evolution of weakly interacting bosons. II. Adv. Math., 228(3):1788–1815, 2011.
  • [41] S. Herr and V. Sohinger. The Gross-Pitaevskii hierarchy on general rectangular tori. Arch. Ration. Mech. Anal., 220(3):1119–1158, 2016.
  • [42] S. Herr and V. Sohinger. Unconditional uniqueness results for the nonlinear Schrödinger equation. Commun. Contemp. Math., 21(7):1850058, 33, 2019.
  • [43] Y. Hong, K. Taliaferro, and Z. Xie. Unconditional uniqueness of the cubic Gross-Pitaevskii hierarchy with low regularity. SIAM J. Math. Anal., 47(5):3314–3341, 2015.
  • [44] Y. Hong, K. Taliaferro, and Z. Xie. Uniqueness of solutions to the 3D quintic Gross-Pitaevskii hierarchy. J. Funct. Anal., 270(1):34–67, 2016.
  • [45] S. Jin, C. D. Levermore, and D. W. McLaughlin. The semiclassical limit of the defocusing NLS hierarchy. Comm. Pure Appl. Math., 52(5):613–654, 1999.
  • [46] 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.
  • [47] S. Klainerman and M. Machedon. On the uniqueness of solutions to the Gross-Pitaevskii hierarchy. Comm. Math. Phys., 279(1):169–185, 2008.
  • [48] A. Knowles and P. Pickl. Mean-field dynamics: singular potentials and rate of convergence. Comm. Math. Phys., 298(1):101–138, 2010.
  • [49] E. H. Lieb, R. Seiringer, J. P. Solovej, and J. Yngvason. The mathematics of the Bose gas and its condensation, volume 34 of Oberwolfach Seminars. Birkhäuser Verlag, Basel, 2005.
  • [50] F. Lin and P. Zhang. Semiclassical limit of the Gross-Pitaevskii equation in an exterior domain. Arch. Ration. Mech. Anal., 179(1):79–107, 2006.
  • [51] E. Madelung. Quantentheorie in hydrodynamischer form. Zeitschrift für Physik, 40(3):322–326, 1927.
  • [52] A. J. Majda. Compressible fluid flow and systems of conservation laws in several space variables, volume 53 of Applied Mathematical Sciences. Springer-Verlag, New York, 1984.
  • [53] T. Makino. On a local existence theorem for the evolution equation of gaseous stars. In Patterns and waves, volume 18 of Stud. Math. Appl., pages 459–479. North-Holland, Amsterdam, 1986.
  • [54] T. Makino, S. Ukai, and S. Kawashima. Sur la solution à support compact de l’équations d’Euler compressible. Japan J. Appl. Math., 3(2):249–257, 1986.
  • [55] I. Rodnianski and B. Schlein. Quantum fluctuations and rate of convergence towards mean field dynamics. Comm. Math. Phys., 291(1):31–61, 2009.
  • [56] S. Serfaty. Mean field limits of the Gross-Pitaevskii and parabolic Ginzburg-Landau equations. J. Amer. Math. Soc., 30(3):713–768, 2017.
  • [57] 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.
  • [58] S. Shen. The rigorous derivation of the 𝕋2\mathbb{T}^{2} focusing cubic NLS from 3D. J. Funct. Anal., 280(8):108934, 72, 2021.
  • [59] V. Sohinger. A rigorous derivation of the defocusing cubic nonlinear Schrödinger equation on 𝕋3\mathbb{T}^{3} from the dynamics of many-body quantum systems. Ann. Inst. H. Poincaré Anal. Non Linéaire, 32(6):1337–1365, 2015.
  • [60] V. Sohinger. Local existence of solutions to randomized Gross-Pitaevskii hierarchies. Trans. Amer. Math. Soc., 368(3):1759–1835, 2016.
  • [61] V. Sohinger and G. Staffilani. Randomization and the Gross-Pitaevskii hierarchy. Arch. Ration. Mech. Anal., 218(1):417–485, 2015.
  • [62] Z. Xie. Derivation of a nonlinear Schrödinger equation with a general power-type nonlinearity in d=1,2d=1,2. Differential Integral Equations, 28(5-6):455–504, 2015.
  • [63] H.-T. Yau. Relative entropy and hydrodynamics of Ginzburg-Landau models. Lett. Math. Phys., 22(1):63–80, 1991.
  • [64] P. Zhang. Wigner measure and the semiclassical limit of Schrödinger-Poisson equations. SIAM J. Math. Anal., 34(3):700–718, 2002.
  • [65] 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.