On the mean-field and semiclassical limit from quantum -body dynamics
Abstract.
We study the mean-field and semiclassical limit of the quantum many-body dynamics with a repulsive -type potential 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 -type potential for the almost optimal case , 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 particles is described by a wave function satisfying a linear -body Schrödinger equation. In realistic systems like fluids, the particle number is so large that these -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 -type and Coulomb potentials, and study the mean-field and semiclassical limit which would lead to macroscopic fluid equations as particle number tends to infinity and Planck’s constant tends to zero. The dynamics of quantum particles in 3D are governed by, according to the superposition principle, the linear -body Schrödinger equation:
(1.1) |
Our Hamiltonian is
(1.2) |
where the factor is the mean-field averaging factor. The -type and Coulomb potentials are
(1.3) |
in which, the parameter characterises different density regimes which correspond to different physical situations.
To have a fixed number of variables in the process, we define the marginal densities associated with in kernel form by
(1.4) |
where and . It is believed that nonlinear Schrödinger equations (NLS) is the mean-field limit equation for these quantum -body dynamics, that is,
(1.5) |
where 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 -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 -type potential was one of the main challenges.
In the mean-field limit as the particle number tends to infinity, the potential converges formally to the Dirac-delta interaction , also called the Fermi potential. For , the average distance between the particles, which is , is much less than the range of the interaction potential, which is , and there are many but weak correlations. For , then the analysis is much more involved because of the strong correlations between particles. For close or equal to , as the scaling is starting to match the Laplacian operator, it is expected that the -type potential generates an interparticle singular correlation structure, closely related to the zero-energy scattering equation
(1.6) |
The scattering function varies effectively on the short scale for 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,
the marginal densities should be considered as
The singular correlation structure is very subtle and plays a crucial role in the mean-field limit from quantum -body dynamics, as it gives an correction to the -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 -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 -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 -body dynamics with singular potentials, a number of challenges remain open:
-
(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)
The large 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 -type potential.
-
(3)
To obtain the quantitative strong convergence rate, a double-exponential restriction between and 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)
The scheme in [23] currently cannot deal with the case, since its proof highly relies on a key collapsing estimate, which fails in the energy space for the case as proven in [36]. Thus, the 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 -body dynamics (1.1), we take the normalization that , and define the quantum mass density and momentum density by
(1.7) |
The limiting macroscopic equation would be the compressible Euler-Poisson equation with a pressure term , which is (in velocity form)
(1.8) |
or (in momentum form)
(1.9) |
Here, as usual,
are respectively the mass density, the velocity, and the momentum of the fluid. The coupling constant is the macroscopic effect of the microscopic interaction . When the coefficient , the system is reduced to a compressible Euler equation. Specifically, we consider the initial data satisfying the condition
(1.10) |
with and , so that the Euler-Poisson system has a unique solution up to some time 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) |
Theorem 1.1.
Let , and the marginal densities associated with be the solution to the -body dynamics (1.1) with a smooth compactly supported, spherically symmetric nonnegative potential and a repulsive Coulomb potential . Assume the initial data satisfy the following conditions:
is normalized and the -body energy bound holds:
(1.12) |
for some .
The initial data to satisfy condition with , and the modulated energy between (1.1) and (1.8) at initial time tends to zero, that is, where
Then under the polynomial restriction555(1.13) is in fact a rational restriction. We say polynomial to avoid confusing “rational” and “reasonable”.
(1.13) |
we have the quantitative estimates on the strong convergence of the mass density
(1.14) |
and on the convergence of the momentum density
(1.15) |
When , Theorem 1.1 is the first result which simultaneously deals with the -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 -type and Coulomb potentials simultaneously, since the energy estimates and collapsing estimates are totally different in the -type and Coulomb potentials. Therefore, it requires completely new ideas for a simultaneous consideration of -type and Coulomb potentials.
When , it reduces to the sole -type potential case. Compared with our previous work [23], we here list the breakthroughs.
-
(1)
The parameter is extended to the full range of , which is almost optimal in the dilute regime.
-
(2)
The previous double-exponential restriction between and is relaxed to be polynomial, which is a tremendous improvement.
-
(3)
Our new approach also works for the case with slight modifications, as the proof is independent of the hardcore harmonic analysis on .
Additionally, the convergence rate should be optimal since the convergence rate of the modulated kinetic energy part at initial time is at most the order of . 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) |
where the kinetic energy part is
(1.17) |
the -type potential part is
(1.18) | ||||
and the Coulomb potential part is
(1.19) | ||||
In Section 2, we first derive the time evolution of the modulated energy
(1.20) |
where the kinetic energy contribution part is
with the notations , and , the -type potential contribution part is
(1.21) | ||||
and the Coulomb potential contribution part is
(1.22) | ||||
It is easy to control the kinetic energy contribution part
(1.23) |
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 as follows
(1.24) | ||||
(1.25) |
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 . Therefore, it is quite difficult to establish a Serfaty’s inequality for the -type potential case, because of the general profile and sharp singularity of the -type potential. In fact, due to the presence of the singular correlation structure caused by the -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 -type potential parts and and establish
(1.26) | ||||
(1.27) |
The proof is divided in several steps.
Step 1. Preliminary reduction. Applying the approximation of identity to the one-body term of , we have the approximation
(1.28) | ||||
To have a closed estimate, namely, letting match the approximation of , we get by integration by parts for the two-body term that
Using the approximation of identity to the one-body term again, we decompose into the main part and error part
(1.29) |
where
(1.30) | ||||
(1.31) |
Such a decomposition is based on the key observation that the difference coupled with the -type potential
(1.32) |
when it is tested against a regular function, would vanish in the limit. Such a structure is notably special for the -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 energy bound
(1.33) |
where satisfies the zero-energy scattering equation
(1.34) |
The singular correlation function varies effectively on the short scale for 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 -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 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 . However, no such estimates hold for 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) |
Hence, we need to rewrite the error part (1.31) as
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 . 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 -type potential and singular correlation function, and obtain the error estimate
(1.36) |
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 . To overcome it, our strategy is to replace with a slowly varying potential 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) | ||||
Note that . Then by the two-body energy bound and the property for the scattering function , we can prove that
(1.38) | ||||
Since the integrand now enjoys the energy bound, we are able to replace by and get
(1.39) | ||||
where with .
In Section 4.2, we will give a detailed proof of the above analysis and and arrive at the approximations of and given by
(1.40) | ||||
and
(1.41) | ||||
Now, from the approximations of and , we are left to prove a reduced form of the functional inequality
which looks more concise and tractable than the original functional inequality (1.26). But, it is unknown if the integrand
(1.42) |
is non-negative. We cannot simply rule out the term 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 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 and , we prove the Gronwall’s inequality for the positive modulated energy
where . 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 -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 -body dynamics corresponding to the -type and Coulomb potentials
where the -type potential part is
(2.1) | ||||
and the Coulomb potential part is
(2.2) | ||||
Here, we might as well assume that the coefficient , as the proof works the same for .
First, we need to derive a time evolution equation for . The related quantities for are given as the following.
Lemma 2.1.
We have the following computations regarding :
(2.3) | |||
(2.4) | |||
(2.5) |
where and the momentum density and the energy are defined by
(2.6) | ||||
(2.7) |
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
where
and
For , we use integration by parts with to obtain
where the other -summands vanish when .
For , we note that the -summands also vanish when and hence have
This completes the proof of (2.4). ∎
Now, we derive the time evolution of .
Proposition 2.2.
Let be defined in , there holds
(2.8) | ||||
where we used the notations , and . Here, the -type potential contribution part is
(2.9) | ||||
and the Coulomb potential contribution part is
(2.10) | ||||
Proof.
We decompose the modulated energy into five parts to do the calculation.
For , by the symmetry of the wave function , we obtain
where in the last equality we have used the conservation of energy. Therefore, we have that
For , from the definition of in (2.6), we note that
Thus, we have
(2.11) |
For the second term on the r.h.s of (2.11), by (2.4) we obtain
(2.12) | ||||
where in the last equality we used the antisymmetry of .
Next, we deal with (2.12). By integration by parts, we obtain
(2.13) |
where we used the notations , and .
Using again integration by parts on the two terms of (2.13) gives
(2.14) | ||||
Combining estimates (2.11)–(2.14), we provide
For , by the Euler-Poisson equation (1.8) and the mass conservation law (2.3), we obtain
Expanding it gives
For , plugging in the Euler-Poisson equation (1.8), we have
For , similarly we get to
Plugging in the Euler-Poisson equation (1.8) and the mass conservation law (2.3), we have
Therefore, putting the five terms together, we reach
(2.15) | ||||
(2.16) | ||||
From the above equation, we collect the -type potential contribution part in (2.8) from
and the Coulomb potential contribution part in (2.8) from
where in the last equality we used the antisymmetry of .
As for the first term in (2.8), we use the Euler-Poisson equation (1.8) to combine the terms taking the form of
(2.17) |
If we rewrite the first term on the right hand side of (2.8)
these are the terms in (2.15),(2.16) and (2.17). Therefore, we arrive at equation (2.8) and complete the proof.
∎
3. 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 -type potential parts and . The main difficulty is the singularities simultaneously from the Coulomb potential, from the direct -potential in the limit, and from the interparticle singular correlation structure.
Recall the zero-energy scattering equation
(3.1) |
and our target estimate
(3.2) |
We first give the properties of the scattering function.
Lemma 3.1.
Suppose that is smooth, spherical symmetric with compact support and satisfies the scattering equation (3.1). Then there exists , depending on , such that
(3.3) | |||
(3.4) |
for all .
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 . From the scattering equation (1.6), we can rewrite
(3.5) |
where is the renormalized constant.
Then the Hardy-Littlewood-Sobolev inequality implies that
Lemma 3.2.
Let and . Then we have
(3.7) |
for .
Proof.
Let
(3.8) |
we rewrite the Hamiltonian (1.2)
By the symmetry of , we have
(3.9) | ||||
Note that and we have
Thus, together with the scattering equation (3.1), we arrive at
(3.10) | ||||
Similarly, we also have
(3.11) | ||||
Further define the shorthands
(3.12) | |||
(3.13) |
which are symmetric with respect to the measure , that is,
(3.14) | ||||
Therefore, from (3.10) and (3.11) we obtain
Expanding it gives
(3.15) | ||||
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 and then yields
Again using the nonnegativity of the potentials, we reach
(3.16) | ||||
For the first term on the r.h.s of (3.16), by (3.14), we have
(3.17) | ||||
where in the last inequality we have used Lemma 3.1 that . To control the last term on the r.h.s of (3.17), we note that
Therefore, we have
For , by Hölder and Sobolev inequalities we have
By the Calderón-Zygmund theory which implies that for and the scattering equation (3.1), we get
Thus, we arrive at
(3.18) |
For , by the properties of the scattering function in Lemma 3.1, we have
which implies that
Then by Cauchy-Schwarz and Hardy’s inequalities, we get
(3.19) | ||||
Next, we deal with the terms and in (3.17). For , we have
where in the last inequality we used the positivity of the Coulomb potential. Noting that , we can use Cauchy-Schwarz and Hardy’s inequalities to obtain
(3.20) | ||||
As the term can be estimated in the same way as , we also have
(3.21) |
Proposition 3.3.
Let and . Define
There exists a constant such that
(3.22) |
for all .
4. Functional Inequalities
In the section, with the a-priori energy bound established in Proposition 3.3, we control the -type potential parts and 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 and then find the main part of . In Section 4.2, we estimate the main part. By a replacement argument, we find proper approximations of and , 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 , we have the estimate
(4.1) |
and a lower bound of
(4.2) |
Here, the notation is a shorthand for and the notation denotes the same order of up to an unimportant constant777The constant could depend on the usual Sobolev constants and the fixed parameters such as the time , the energy bound , and the Sobolev norms of but the constant is independent of . .
4.1. Error Analysis of Two-Body Term
From the expression of
(4.3) | ||||
the difficult part is the two-body term
At first sight, the lack of a uniform regularity estimate for the two-body density function makes further analysis difficult. With the singular correlation structure in mind, we decompose the two-body density function into the singular and regular parts
and rewrite the two-body term as
(4.4) |
That is, the singularities come from the potential and the singular correlation function . 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 -type potential
(4.5) |
which would vanish as tends to the infinity. Such a structure is special for the -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 , we have
(4.6) | ||||
where the notation means .
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 by integrating by parts
(4.7) |
It remains to show the term (4.7) is indeed an error term. For simplicity, we set
Then we have
(4.8) | ||||
We bound and , 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 , by Lemma 3.1, we have the upper bound estimate
Therefore, we arrive at
(4.9) | ||||
where in the last inequality we used the operator inequality (A.2).
For the term , we first discard and then use Cauchy-Schwarz to get
(4.10) | ||||
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) | ||||
where in the last inequality we optimized the choice of .
Together with (4.8) and estimates for the terms and , we reach
where in the last inequality we used the two-body 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
Using the identity approximation to the one-body term of , we arrive at
By the identity approximation again, we also have the approximation of that
We now need to deal with the sharp singularity of . We tame the singularity by replacing 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 and 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
Then for the two-body term we have
(4.12) | ||||
and for the one-body term we have
(4.13) | ||||
In particular, given , we have the approximation of
(4.14) | ||||
and given , we have the approximation of
(4.15) | ||||
Proof.
For (4.12), we recall and rewrite
For the term , we use the Poincar type inequality with in Lemma A.3 to obtain
where in the last inequality we used the two-body energy bound (3.22).
For the term , by Lemma 3.1, we have . Therefore, we get
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
Then by the triangle inequality, we arrive at
which completes the proof of (4.12).
For (4.13), we rewrite
(4.16) |
For the error terms in (4.16), we use Hölder and Sobolev inequalities to get
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).
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) | ||||
which is more concise than the original functional inequality. However, it is unknown whether or not the integrand
(4.18) |
is non-negative, so we cannot directly bound the term in (4.17). We prove that, if integrated against , (4.18) provides a non-negative contribution up to a small correction and use that to prove the lower bound of . The special structure of a relatively slowly varying and explicit potential 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 to be determined, we have
(4.19) | ||||
Proof.
For simplicity, set . By the symmetry of , we can write
To simplify, we define the measure
(4.20) |
We rewrite
(4.21) | ||||
where the last term on the r.h.s of (4.21) comes from the diagonal summation. In particular, if we take , we also have
(4.22) | ||||
Note that
(4.23) |
which is a smallness term as long as .
Next, we get into the analysis of the main term. Note that the convolution property of the Gaussian function , which is
(4.24) |
where and . Putting (4.24) into the main term of (4.22) gives
(4.25) | ||||
where
(4.26) | ||||
(4.27) |
Thus, we are left to bound the terms and .
For the term , we use Cauchy-Schwarz inequality to get
where
For , we further decompose it into two parts , where the diagonal part is
and the off-diagonal part is
For , by the symmetry of , we have
where in the last inequality we used that and the mass conservation for .
For , by the symmetry of , we also have
where
To bound , we recall then get
where we discarded in the second line and used the operator inequality (A.2) in the second-to-last inequality, and the two-body energy bound (3.22) in the last inequality.
For , we rewrite
where in the last inequality we used the mass conservation for . By the triangle, Hölder inequalities and Lemma A.1 we get
To sum up, we complete the estimates for the term and reach
(4.28) |
where in the last inequality we used that for .
To prove the lower bound estimate (4.2) for , we give the following estimate.
Lemma 4.5.
Let to be determined, we have
(4.30) | ||||
Proof.
To the end, we get into the proof of Proposition 4.1.
Proof of Proposition 4.1.
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 -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) |
where the kinetic energy part is
(5.2) |
and the potential energy part is
(5.3) |
From lower bound estimates (4.2) on and (1.25) on , we can add a small compensation such that
(5.4) |
where Thus, we introduce the positive modulated energy
(5.5) |
We now provide a closed estimate for the positive modulated energy.
Proposition 5.1.
For , we have the differential inequality
(5.6) |
Moreover, we conclude
(5.7) |
and
(5.8) |
Proof.
From the evolution of the modulated energy (2.8), we find that
By the functional inequalities on and (1.24) on , we get
(5.9) |
Then by Gronwall’s inequality, we arrive at
(5.10) |
for .
To the end, we get into the proof of Theorem 1.1.
Proof of Theorem 1.1.
Convergence of the mass density .
We decompose
(5.11) |
For the first term on the r.h.s of (5.11), we use Hölder inequality and Lemma A.1 to obtain
Next, we estimate the terms and . By the Calderón-Zygmund theory which implies that for , we get
By the Leibniz rule, Minkowski, Hölder, and Sobolev inequalities, we then obtain
where in the last inequality we have used the energy bound (2.5) for . Similarly, we also have
and
With , these bounds give that
(5.12) |
Thus, combining (5.11), (5.12) with (5.8), we arrive at
(5.13) |
where in the last inequality we have used that .
Convergence of the momentum density .
Recall the momentum density
Then by the triangle and Hölder’s inequalities, we have
where in the last inequality we used the mass conservation, estimates and . ∎
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 and , where . For any ,
(A.1) |
for any .
Lemma A.2 ([29], Lemma A.3).
Let and . Then
(A.2) | |||
(A.3) | |||
(A.4) |
Lemma A.3.
Suppose that such that
Let and , then we have
for .
Proof.
For the derivation of NLS, this Poincar 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 . Switching to Fourier space, we observe that
By using for and , we have
It suffices to bound the term containing , as the term containing can be estimated similarly. We rewrite
By Cauchy-Schwarz inequality,
where in the last inequality we used that
for all . ∎
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 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 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 . 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 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 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 . 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.