Global well-posedness of the quantum Boltzmann equation for bosons interacting via inverse power law potentials
Abstract.
We consider the spatially inhomogeneous quantum Boltzmann equation for bosons with a singular collision kernel, the weak-coupling limit of a large system of Bose-Einstein particles interacting through inverse power law. Global well-posedness of the corresponding Cauchy problem is proved in a periodic box near equilibrium for initial data satisfying high temperature condition.
AMS Subject Classification (2020): 35Q20, 82C40.
1. Introduction
Quantum Boltzmann equations are proposed to describe the time evolution of a dilute system of weakly interacting bosons or fermions. The derivation of such equations dates back to as early as 1920s by Nordheim [43] and 1933 by Uehling-Uhlenbeck [47]. As a result, the quantum Boltzmann equations are also called Boltzmann-Nordheim equations or Uehling-Uhlenbeck equations. Later on, further developments were made by Erdős-Salmhofer-Yau [17], Benedetto-Castella-Esposito-Pulvirenti [6], [10], [7], [9] and [8], Lukkarinen-Spohn [41]. One can refer the classical book [13] for physical backgrounds.
In this article, we consider the Cauchy problem of the quantum Boltzmann equation for bosons
(1.1) |
Here is the density function of particles with velocity at time in position . Here is the periodic box with unit volume . The quantum Boltzmann operator acting only on velocity variable is defined by
(1.2) |
where according to [17] and [10] the Boltzmann kernel has the following form
(1.3) |
Here is the Plank constant. In (1.3), the radial function is the Fourier transform of a radial potential function .
In (1.2) and the rest of the article, we use the convenient shorthand , , , where , are given by
(1.4) |
Now it remains to see the notation in (1.2). For or , we denote
(1.5) |
If , we write . The term is interpreted as “difference” before and after collision. We will consider a singular kernel and always need some function difference to remove the singularity. The notation or indicates which function is offering help to cancel the singularity.
By the following scaling
(1.6) |
we can normalize the Plank constant . Indeed, it is easy to check is a solution to (1.1) if and only if is a solution of the following normalized equation
(1.7) |
where the operator is defined by
(1.8) |
with the kernel given by
(1.9) |
In this article, we will take the inverse power law potential . Before that, we give a short review on some relevant mathematical research of the quantum Boltzmann equation.
1.1. Existing results
We recall some existing mathematical results on the quantum Boltzmann equation in this subsection. The quantum Boltzmann equations include two models: one for bosons or Bose-Einstein(B-E) particles, the other for fermions or Fermi-Dirac(F-D) particles. The quantum Boltzmann equation for B-E particles is written by (1.7), (1.8) and (1.9) and usually referred as “BBE equation”. If the four “” in (1.8) and the one “” in (1.9) are replaced by minus sign “”, we will get the quantum Boltzmann equation for F-D particles which is usually referred as “BFD equation”.
We begin with the mathematical results on BFD equations. In spatially homogeneous case , we refer the monograph Escobedo-Mischler-Valle [18] for global existence and weak convergence to equilibrium. We also refer Lu [31] and Lu-Wennberg [40] for weak and strong convergence to equilibrium of mild solutions. On the whole space , Dolbeault [14] established global existence and uniqueness of mild solutions for globally integrable kernels. Lions [29] proved global existence of distributional solutions for locally integrable kernels. Alexandre [1] proved global existence of H-solutions for non-cutoff kernels that allow angular singularity given by inverse power law potentials. Based on the averaging compactness result in [34], on the torus , Lu [35] proved global existence of weak solutions for kernels with very soft potentials. Recently in a perturbation framework, on the whole space , Jiang-Xiong-Zhou [27] went further to consider the incompressible Navier-Stokes-Fourier limit from the BFD equation with hard sphere collisions.
Unlike Fermi-Dirac particles whose density has a natural bound due to Pauli’s exclusion principle, density of Bose-Einstein particles may blow up in finite time, which corresponds to the intriguing phenomenon: Bose-Einstein condensation(BEC) in low temperature. As a consequence, the mathematical study of BBE equations is much more difficult than BFD equations. As a result, most of existing results on BBE equations are concerned with isotropic solutions in the homogeneous case with non-singular kernel, for instance hard sphere model or hard potentials with some cutoff. Lu [30] proved global existence of solutions for some kernels with strong cutoff assumption (not satisfied by the hard sphere model). Based on the results in [30], Lu [32] further established global existence of conservative distributional (measure-valued) isotropic solutions for some kernels including the hard sphere model. Also see Escobedo-Mischler-Velázquez [19] and Escobedo-Mischler-Velázquez [20] for some singular (near ) solutions.
In terms of long time behavior of the conservative measure-valued isotropic solutions, we refer Lu [33] for strong convergence to equilibrium and single point concentration. With some local condition on the initial datum, long time strong convergence to equilibrium was proved in [38] and [39]. Remarkably for any temperature and without any local condition on the initial datum, Cai-Lu [12] provided algebraic rate of strong convergence to equilibrium.
BEC is an interesting physical phenomenon that deserves deep mathematical understanding. It is fortunate to see many excellent mathematical results (for instance, Spohn [45], Lu [36, 37], Escobedo-Velázquez [21, 22]) on this topic.
1.2. Potential function and Boltzmann kernel
In weak-coupling regime for bosons, the Boltzmann kernel depends on the potential function via (1.9). As mentioned before, existing results on BBE equations are mostly concerned with the hard sphere model. This amounts to taking the Dirac delta function as the potential function and thus for some universal constant . Observe that the kernel in the quantum case is the same as that in the classical case. Another physically relevant potential is the inverse power law which has been extensively studied in the research of classical Boltzmann equations but has rarely been considered in the quantum case. To our best knowledge, this article seems to be the first to study BBE equation with inverse power law.
Note that corresponds to the famous Coulomb potential which is the critical case for Boltzmann equation to be meaningful. In this article, we work in dimension 3 where is the critical value such that is locally integrable near . Fix , then the Fourier transform of is Let be the angle between and , then . As a result, the Boltzmann kernel given by (1.9) is
Here and in the rest of the article denotes a constant that depends only on fixed parameters and could change from line to line. Due the symmetry structure of (1.9), we can always assume . Then , since , we get
Since , the following integral over is bounded,
(1.10) |
As in (1.10), in the rest of the article, when taking integration, the range of some frequently used variables will be omitted if there is no ambiguity. For instance, the usual ranges will be consistently used unless otherwise specified. Whenever a new variable appears, we will specify its range once and then omit it thereafter.
By (1.10), for the kernel has finite momentum transfer which is a basic condition for the classical Boltzmann equation to be well-posed. The constant in (1.10) blows up as since for . If , the angular function satisfies Grad’s angular cutoff assumption
Note that is the critical value such that Grad’s cutoff assumption fails. To summarize, can be seen as angular cutoff while can be seen as angular non-cutoff. We consider the much harder case in this article. By taking , it suffices to consider the more general kernel,
(1.11) |
The parameter pair is commonly used in the study of classical Boltzmann equations with inverse power law. We find the resulting kernel (1.11) has close relation (see (2.18) for details) to the Boltzmann kernel defined in (2.16). The condition is referred as (very) soft potentials. From now on, the notation or stands for the kernel in (1.11) unless otherwise specified.
1.3. Temperature and initial datum
Temperature plays an important role in the study of quantum Boltzmann equation. For example, for B-E particles, BEC will happen in low temperature. We now introduce some basic knowledge about temperature in the quantum context. Let us consider a homogeneous density with zero mean . For , we recall the moment function
Let for simplicity. Let be the mass of a particle, then and are the total mass and kinetic energy per unit space volume. Referring [32], the kinetic temperature and the critical temperature of the particle system are defined by
(1.13) |
where is the Boltzmann constant and is the Riemann zeta function.
We now recall some known facts about equilibrium distribution. The equilibrium of the classical Boltzmann equation is the Maxwellian distribution with density function defined by
Here is density, is mean velocity and is temperature. The famous Bose-Einstein distribution has density function
(1.14) |
Now and do not represent density, mean velocity and temperature anymore, but only three parameters. The ratio quantifies high and low temperature. In high temperature , the equilibrium of BBE equation is the Bose-Einstein distribution (1.14) with . In low temperature , the equilibrium of BBE equation is the Bose-Einstein distribution (1.14) with plus some Dirac delta function. That is, the equilibrium contains a Dirac measure. In the critical case , the equilibrium is (1.14) with . One can refer to [33] for the classification of equilibria.
In this article, we work with high temperature and thus the equilibrium of (1.1) is defined in (1.14). For perturbation around equilibrium, we define
(1.15) |
We remark that the function serves as the multiplier in the expansion .
Recall that the solution of (1.1) conserves mass, momentum and energy. That is, for any ,
(1.16) |
Once is appropriately given, the constants in (1.14) are uniquely determined through
(1.17) |
Without any loss of generality, we assume that has zero mean and thus gives from now on. Also without any loss of generality, we assume that gives in this article. Indeed, we can make the transform to reduce the general case to the special case . As a result, we only keep as a parameter. That is, we only consider those initial data with according to (1.17). Taking the equilibrium and the multiplier function reduce to
(1.18) |
When is small, it is easy to see and so by recalling (1.13),
In our main result Theorem 1.1, we will assume for some small constant which means
(1.19) |
That is, we need high temperature assumption. Note that high temperature assumption is also imposed in [28] to prove global well-posedness of homogeneous BBE equation with (slightly general than) hard sphere collisions.
1.4. Perturbation around equilibrium and main result
For simplicity, let . With the expansion , the linearized quantum Boltzmann equation corresponding to (1.12) reads
(1.20) |
Here the linearized quantum Boltzmann operator is define by
(1.21) |
where is defined by
(1.22) |
The bilinear term and the trilinear term are defined by
(1.23) | |||||
(1.24) |
The notation in (1.23) is defined by
(1.28) | |||||
Remark that the three operators , and depends on through .
Our goal is to prove global well-posedness of (1.20) in some weighted Sobolev space. More precisely, we use the following energy and dissipation functional
(1.29) |
where . See subsection 1.8 for the definition of and . See (1.51) for the definition of . For the moment, just keep in mind is the dissipation corresponding to the energy . Here is a polynomial weight on the velocity variable . The weight order depends on the derivative order and the sequence verifies
(1.30) | |||
(1.31) |
The condition (1.30) is used to deal with linear streaming term as (see (1.44) below). The two conditions (1.30) and (1.31) together ensure that increases as decreases (see (5.49)).
Now we are ready to give global well-posedness of (1.20) in the following theorem.
Theorem 1.1.
Let . There exist two universal constants such that the following global well-posedness is valid. Let . If
(1.32) |
then the Cauchy problem (1.20) has a unique global solution satisfying for some universal constant ,
(1.33) |
and for all ,
We give some explanations and comments on Theorem 1.1 in the rest of this subsection and the next two subsections.
We emphasize that the constants in Theorem 1.1 could depend on . In Theorem 1.1 and the rest of the article, if a constant depends only on the given parameters , we say it is a universal constant. In particular, the two constants are independent of .
The condition owes to Theorem 5.8. Understandably, a certain high order is required to deal with the singular kernel (1.11) and the trilinear operator . Recall that the hard sphere model in [44] needs .
Note that Theorem 1.1 gives global well-posedness under the condition
Roughly speaking, these conditions mean
Note that this -related smallness did not appear in previous works where the parameter was regarded as a fixed constant. Other effects of viewing as a variable parameter will be pointed out later.
By (1.32) and (1.33), we have for any . This estimate is consistent with the trivial case where is the solution to problem (1.12) starting with a zero initial datum .
In terms of physical relevance of Theorem 5.8, we give the following two remarks.
Remark 1.1.
Note that and thus for any . Theorem 1.1 shows that BEC will not happen if the initial datum is close enough to an equilibrium with high temperature, which is reasonable and consistent with physical observation.
Remark 1.2.
1.5. Feature of our result
In terms of condition and conclusion, Theorem 1.1 is the closest to the main result (Theorem 1.4) of [28]. More precisely, both of these two results validate global existence of anisotropic solutions under very high temperature condition, see (1.19) and Remark 1.6 of [28] for more details. The main differences of these two results are also obvious. To reiterate, [28] considers spatially homogeneous case and (slightly general than) hard sphere model, while we work with the inhomogeneous case and non-cutoff kernels.
In terms of mathematical methods, this article is closer to [5], [44] and [27] since all of these works fall into the close-to-equilibrium framework well established for the classical Boltzmann equation. There are many great works that contribute to this mathematically satisfactory theory for global well-posedness of the classical Boltzmann equation. For reference, we mention [48, 24] for angular cutoff kernels and [23, 4] for non-cutoff kernels.
Each of [5], [44] and [27] has its own features and focuses. The work [5] is the first to investigate both relativistic and quantum effects. The article [27] studies the hydro dynamic limit from BFD (but not BBE) to incompressible Navier-Stokes-Fourier equation. The work [44] includes three cases: torus near equilibrium, whole space near equilibrium and whole space near vacuum. These novel works contribute to the literature of quantum Boltzmann equation from different aspects.
Like the above works, our main result Theorem 1.1 also has some unique features that may better our understanding of quantum Boltzmann equation. In particular, this article may be the first in the spatially inhomogeneous case to
-
•
study the quantum Boltzmann equation with inverse power law potentials;
-
•
incorporate high temperature condition into global well-posedness;
-
•
rule out BEC globally in time.
Different from the three works ([5], [44] and [27]), we keep the parameter along our derivation throughout the article. This intentional choice enables us to relate the high temperature condition to the smallness of quantitatively in (1.19). As a result, BEC is rigorously ruled out globally.
1.6. Possible improvements
Compared to the well-established global well-posedness theory for classical Boltzmann equations with inverse power law potentials, our Theorem 1.1 has much room to improve. In this subsection, some possible improvements of Theorem 1.1 are given based on the author’s limited knowledge. To keep the present article in a reasonable length, we leave these improvements in future works.
Recall is defined in (1.29) with the weight order satisfying (1.30) and (1.31). Let be the space with the minimal weight order where and verifies the two conditions (1.30) and (1.31) with identities. To illustrate, we give an example of the weight order . Recall for the inverse power law , and thus . Let us take , the corresponding with is shown in Table 1. In Table 1, the column index represents derivative order of space variable and the row index represents derivative order of velocity variable . For example .
9 | 0 | ||||||||||
8 | 9 | 1 | |||||||||
7 | 17 | 10 | 2 | ||||||||
6 | 24 | 18 | 11 | 3 | |||||||
5 | 30 | 25 | 19 | 12 | 4 | ||||||
4 | 35 | 31 | 26 | 20 | 13 | 5 | |||||
3 | 39 | 36 | 32 | 27 | 21 | 14 | 6 | ||||
2 | 42 | 40 | 37 | 33 | 28 | 22 | 15 | 7 | |||
1 | 44 | 43 | 41 | 38 | 34 | 29 | 23 | 16 | 8 | ||
0 | 45 | 45 | 44 | 42 | 39 | 35 | 30 | 24 | 17 | 9 | |
0 | 1 | 2 | 3 | 4 | 5 | 6 | 7 | 8 | 9 | 10 |
Note that in Theorem 1.1, smallness of is imposed to prove well-posedness in . We may try to prove well-posedness in under the condition . That is, smallness assumption is only imposed on . Such kind better result was derived by Guo in [25] for the Vlasov-Possion-Landau system.
Theorem 1.1 does not consider relaxation to equilibrium. Using the methods in [46] and [16] for classical Boltzmann equations, it is very promising to derive similar long time behaviors for the solution obtained in Theorem 1.1. For example, we can derive almost exponential decay like in [46] by recalling (1.51) and using some interpolation inequality to deal with soft potential . In this regard, a possible result may be as follows. Let and assume . Try to prove for any ,
The first upper bound estimate on is used together with interpolation method to derive the polynomial decay of . By interpolation, we should have .
We may try to derive sub-exponential decay rate under more assumption on initial data like in [16]. Roughly speaking, we could get something as follows. Let be small enough and assume . Try to prove for any ,
where .
Another topic is to prove global well-posedness in a larger space than . Note that for classical Boltzmann equation with inverse power law potential, see [16] for global well-posedness in the up-to-date largest space (containing for any ). One may try to establish global well-posedness in such kind low regularity space.
1.7. Strategy of proof
The proof of Theorem 1.1 will be given in subsection 7.3 by a rigorous continuity argument based on the local well-posedness result in Theorem 6.1 and the a priori estimate in Theorem 7.1. We spend this subsection to outline the procedure and give the key points in deriving Theorem 6.1 and Theorem 7.1. The best way is to glance over this subsection first and come back to read it carefully when appropriate.
1.7.1. An auxiliary Cauchy problem
In order to prove local well-posedness of (1.12) or (1.20), we study the following linear problem
(1.34) |
Here is a given function and is unknown. The operator is defined by
(1.35) |
Since is a given function, the operator is linear in . Using the expansion , the equation (1.34) is equivalent to
(1.36) |
where the linear operators are defined by
(1.37) | |||
(1.38) |
and the operator is defined in (3.2). In this article, the subscripts “” and “” are referred to “main” and “remaining” respectively. For example, is the main part of . Corresponding to the remaining part of , we also have the main part defined in (5.10). Note that .
Local well-posedness of (1.20) is proved by iterating on equation (1.36). In order to implement energy method on equations (1.20) and (1.36), we first give necessary estimates on the linear operators , the bilinear operators and the trilinear operator in Section 2, 3 and 4 respectively. We now give some keys ideas of these estimates.
1.7.2. Linear operators
In this sequel, we consider the four linear operators: .
Coercivity estimate of . Coercivity estimate of plays a central role in the close-to-equilibrium framework. Note that is a self-joint operator. Indeed,
(1.39) |
The null space of is
(1.40) |
Let (see (2.11) for its precise definition) be the projection operator on the null space . Then enjoys the following estimate (see Theorem 2.2 for the precise statement)
(1.41) |
where the norm is defined in (1.51). We use to denote the identity operator. Note that vanishes with order-1 as . There are two key points in deriving (1.41).
-
•
The first one is to reduce quantum case to classical case. Let us recall the classical (non-quantum) linearized Boltzmann operator with kernel in (1.11),
(1.42) where we use defined in (1.22). In this article, the subscript “” is referred to “classical”. When is small, we reduce to with some small correction term. See Lemma 2.2 for details.
- •
With these two key observations, using the coercivity estimate of in Theorem 2.1, we get Theorem 2.2 for the coercivity estimate of .
Some preliminary formulas. We collect some preliminary formulas for Boltzmann type integrals in subsection 2.2. We first give two commonly used changes of variable in (2.27) and (2.28). When estimating integrals, we have to deal with singularity in the kernel . To cancel angular singularity, we give Lemma 2.4 (based on Taylor expansion) and Lemma 2.5(based on the cancelation lemma in [2]) for integrals that have particular structures. To cancel velocity singularity, we prepare Lemma 2.6. In order to retain negative exponential weight, we give Lemma 2.7 and Remark 2.2.
Upper bound of . We make several rearrangements to and then use the formulas in subsection 2.2 to get in Proposition 2.1. Note that we succeed in obtaining an upper bound only involving -norm with negative exponential weight. This result is comparable to the classical case, see Lemma 2.15 in [4].
Upper bound of . Using the upper bound for functionals and involving classical Boltzmann kernel defined in (2.16), we prove in Theorem 2.3 that the two functionals and are bounded from up by . See (2.48) for the definition of these functionals. We make several rearrangements to such that it can be controlled by the two functionals, which enables us to apply Theorem 2.3 to get . See Proposition 2.2 for details.
1.7.3. Nonlinear operators
To analyze the nonlinear operators , we will make use of the classical (non-quantum) Boltzmann operator with kernel in (1.11),
(1.43) |
By some rearrangement, we find
where is defined in (3.3) and is defined in (3.4). Here . In order to estimate , we need to consider . We now give some keys points.
Upper bound of . Observing (2.18), we first get an upper bound of in Corollary 3.1 by using some known estimates of classical Boltzmann operator with defined in (2.16). Then noting that and using the elementary Lemma 3.1, we get the upper bound of in Proposition 3.2.
Upper bound of . See from (3.24), (3.25) and (3.26) that contains differences and . To deal with functionals containing such differences, we introduce Lemma 3.2 and Remark 3.1. We then make suitable rearrangements and use some basic tools (such as Cauchy-Schwartz inequality, changes of variable in (2.27) and (2.28)) to bound from up by the functionals in Lemma 3.2 and Theorem 2.3. See Proposition 3.3 for more details.
Upper bounds of . Lemma 3.3 and Remark 3.4 are introduced to deal with various functionals containing . For the upper bound of and , we make a variety of rearrangements and use some basic tools (such as Cauchy-Schwartz inequality, changes of variable in (2.27) and (2.28), weight retainment (2.37), the imbedding , the usual change of variable or ) to bound and from up by and the functionals in Lemma 3.2(Remark 3.1), Lemma 3.3(Remark 3.4) and Theorem 2.3. keep in mind that the ending upper bounds will be used in later energy estimate of the equations (1.20) and (1.36) and so they must be sharp enough. We illustrate this point by looking at . We use to denote -derivative. In later energy estimate of , we will encounter
where . The most dangerous term appears when all the derivatives fall on a single function. There are three cases
To deal with these terms, the upper bound of must only involve or for at least one of . To achieve such flexibility, we try hard to figure out suitable rearrangements. See Theorem 4.1 for the three final estimates of . Applicable estimates of are given in Propositions 3.4, 3.5 and 3.6 corresponding to the three terms in (3.4).
1.7.4. Commutator estimates
To implement energy method in weighted Sobolev space, we need to consider commutators between the weight function and the above operators. Such commutators always contain the difference . To deal with functionals containing such difference, we introduce Lemma 5.1 and Remark 5.1. Then estimates of commutators like are derived by using Lemma 5.1(Remark 5.1) and some other known results. With these commutator estimates and the operator estimates in Section 2, 3 and 4, we give weighted inner product estimates first in space and then in space. Finally, some useful energy estimates are derived in Theorem 5.2, 5.5 and 5.8. See Section 5 for details.
1.7.5. Local well-posedness
We first prove well-posedness of the equation (1.36) in Proposition 6.2 under suitable smallness assumption on the given function . In addition, some continuity (see (6.4)) to initial datum is provided. Based on these results on equation (1.36), we construct a function sequence through iteration. More concretely, we start with and take in (1.36) to construct a function sequence . Then using the continuity result (6.4), the sequence is proved to be a Cauchy sequence generating a local solution to the nonlinear equation (1.20). See Theorem 6.1 and its proof for more details.
1.7.6. A priori estimate
The standard macro-micro decomposition method is used in this step. The decomposition reads . Recall that and are referred as “macroscopic” and “microscopic” parts respectively. For the “macroscopic” part, we first derive a system of macroscopic equations (7.6) and some local conservation laws (7.9). With these equations and some other elementary estimates, the dissipation on can be derived as in [15]. See Lemma 7.3 and 7.4 for the precise results. Full dissipation functional in (1.29) is derived in Proposition 7.1 for the equation where is a general source term.
Let us see a key point in the proof of Proposition 7.1. The most difficult term is the free streaming term when taking -derivative. More precisely, we need to deal with the commutator . By the condition (1.30), in Lemma 6.2 for any we get
(1.44) |
Here . Observe that there is a factor before the dissipation (see (1.41) or Theorem 2.2) for the corresponding energy. Here we have
For this reason, in (1.44) we should take for some sufficiently small . The resulting latter term needs the dissipation in the energy estimate of for some large constant . Such kind of treatment results in the following combination
(1.45) |
for some constants . Note that the power of depends on the -derivative order . See the proof of Proposition 7.1 for a detailed and rigourous derivation. A comparison of (1.45) and the energy functional in (1.29) explains the factor in Theorem 1.1.
1.7.7. Global well-posedness and smallness of parameter
Global well-posedness (Theorem 1.1) is proved in subsection 7.3 by a continuity argument based on Theorem 6.1 and Theorem 7.1 for . Let us see the places where smallness condition appears.
-
(1)
In Lemma 2.1, we need in order to bound from below the denominator .
-
(2)
In Lemma 2.3, we need to estimate the operator difference .
-
(3)
In Theorem 2.2, we need to get the coercivity estimate of .
- (4)
- (5)
As , we finally take in Theorem 1.1.
1.8. Notations
In this subsection, we give a list of notations.
Given a set , is the characteristic function of .
Given two operator , their commutator is denoted by .
The notation means that there is a universal constant such that . The constant could depend on the kernel parameters and the energy space index .
If both and , we write .
We denote or by a constant depending on .
The bracket is defined by . The weight function .
For , and .
For , and .
For , and .
For a multi-index , define .
For , denote .
We now introduce some norm.
For and a function on , define
(1.46) |
Note that .
For and a function on , define
Note that .
For and a function on , define
(1.47) | |||
(1.48) |
For and a function on , define
(1.49) |
where is the Fourier transform.
For and a function on , define
(1.50) |
where . Here are the real spherical harmonics verifying that where is the Laplacian operator on the unit sphere . Note that is an orthonormal basis of . Here the notation refers to “anisotropic regularity”.
For , we define
(1.51) |
Recalling (1.46), (1.49) and (1.50), the three norms on the right-hand of (1.51) share a common weight function for .
For and a function on , define
(1.52) |
Note that .
For and a function on , define
(1.53) |
1.9. Plan of the article
Section 2 contains estimates of linear operators, including coercivity estimate of and upper bounds of . Section 3 and 4 are devoted to upper bounds of the bilinear operator and the trilinear operator respectively. In Section 5, various functionals that will appear in later energy method are estimated after necessary commutator estimates. In Section 6, we derive local well-posedness. In Section 7, we first prove a priori estimate and then establish global well-posedness. Section 8 is an appendix in which we put some elementary proof for the sake of completeness.
2. Linear operator estimate
In the rest of the article, in the various functional estimates, the involved functions are assumed to be functions on or such that the corresponding norms of them are well-defined. For simplicity, we use the notation .
2.1. Coercivity estimate of
Recall (1.18), (1.21) and (1.42). The operator will vanish as . However, formally it is easy to see as . We define
(2.1) |
Recalling (1.18), we can see that
(2.2) |
When is small and close to 0, it is obvious that . More precisely, we have
Lemma 2.1.
If , then
(2.3) |
As a direct result, there holds
(2.4) |
Proof.
In the rest of the article, we always assume which enables us to use the results in Lemma 2.1. Other smallness condition on will be specified as we go further.
We relate and in the following lemma.
Lemma 2.2.
It holds that
Proof.
Recalling (1.40), since , for , we can also write
(2.7) |
Note that based on (2.7), is well-defined by
(2.8) |
Observe that is the null space of .
We construct an orthogonal basis for for as follows
(2.9) |
Note that is the projection of on . We denote the coefficient which depends only on . Then . By normalizing , an orthonormal basis of can be obtained as
(2.10) |
With this orthonormal basis, the projection operator on the null space is defined by
(2.11) |
Let us see more clearly. Let us define
(2.12) |
For simplicity, we set for and . Then by direct derivation and rearrangement, we have
(2.13) | |||||
where
Note that is a vector of length 3. Let us define
(2.14) |
For simplicity, let . Then there holds
(2.15) |
The next lemma shows that is of order when is small.
Lemma 2.3.
The proof of Lemma 2.3 will be given in the appendix.
We now give the coercivity estimate of based on some known result on the classical linearized Boltzmann operator with inverse power law potential. For inverse power law potential, it suffices to consider the following Boltzmann kernel
(2.16) |
Note that the superscript “” is short for “inverse power law potential”. Let be the associated classical linearized Boltzmann operator, i.e., is defined through (1.42) by replacing with . By [26], it turns out that
(2.17) |
where the norm is defined in (1.51). We remind that the norm is equivalent to in [23] and in [4].
Since for , recalling (2.16) and (1.11), there holds
(2.18) |
Recalling (2.6) and using (2.18), we get
By using (2.17), we get the following theorem.
Theorem 2.1.
It holds that
For later reference, let be the two optimal universal constants such that
(2.19) |
We now prove the following the following coercivity estimate of by using Theorem 2.1, Lemma 2.2 and Lemma 2.3.
Theorem 2.2.
There are three universal constants such that for any verifying , it holds that
The constants are explicitly given in (2.26).
Proof.
Since , it suffices to prove for verifying . From now on, we assume . By Lemma 2.2, for , we have
(2.20) |
By (2.19) in Theorem 2.1, we have
(2.21) | |||
(2.22) |
Here is the optimal constant such that for any and is the optimal constant such for any . The existence of is ensured by (3.12). We remark that are universal constants independent of .
For the upper bound estimate in Theorem 2.2 we only need . Indeed, based on the proof, we only need Lemma 2.2 and Theorem 2.1 to get (2.23). With the constant defined in (2.26), we have
Remark 2.1.
For , we have .
2.2. Some preliminary formulas
In this subsection, we recall some useful formulas for the computation of integrals involving defined in (1.11). It is obvious that the change of variable has unit Jacobian and thus
(2.27) |
where is a general function such that the integral exists. Thanks to the symmetry of elastic collision formula (1.4), the change of variable has unit Jacobian and thus
(2.28) |
From now on, we will frequently use the notation (1.5). By (1.5) and the shorthand , it is easy to see
Similar to (1.5), we introduce
(2.29) |
The term is interpreted as “addition” before and after collision.
In upper bound estimate, we frequently encounter quantities like and . Thanks to (2.18), it suffices to consider the kernel defined (2.16) for upper bound estimates involving . To cancel the angular singularity of near , one usually relies on the order-2 factor and the following fact for ,
(2.30) |
Thanks to the symmetry of -integral, the factor will appear if we appropriately apply Taylor expansion to . By Taylor expansion, we have the following two candidates.
(2.31) | |||
(2.32) |
Here . Since , the second order contains . So it remains to deal with the first order term. Fortunately we have the following two identities,
(2.33) | |||
(2.34) |
We remark that (2.34) holds for any fixed .
Note that the right-hand side of (2.33) contains and so we can use (2.31) and (2.33) to cancel the angular singularity in . We can use (2.32) and (2.34) to cancel the angular singularity in since the first order term vanishes by (2.34). As a result, we have
Lemma 2.4.
The following two estimates are valid.
(2.36) |
Lemma 2.5.
There holds
The following Lemma 2.6 is used to cancel singularity near when .
Lemma 2.6.
For , there holds
Note that the points for connecting and appear in the formulas (2.31) and (2.32) . If we apply Taylor expansion to , the points for will appear. Thanks to the symmetry of elastic collision formula (1.4), we have
Lemma 2.7.
For any , there holds
Remark 2.2.
Since
for simplicity in the rest of the article we will use
As a result, recalling , for any , there holds
(2.37) |
We will frequently use (2.37) to retain the good negative exponential (-type) weight.
2.3. Estimate of
After the preliminary preparation in subsection 2.2, we are ready to give upper bound estimate of the operator .
Proposition 2.1.
There holds
Proof.
In what follows, we derive the estimates of and with full details because the tricks used here will be used later. Recalling (2.1), it is easy to see
(2.40) |
Using (2.36) and (2.40), we have
By (2.37), we have and thus
(2.41) | |||||
where from the first to the second line we use the following change of variable (see Lemma 1 in [2] for details)
(2.42) |
From the second to the third line we use the fact . The last inequality in (2.41) follows Cauchy-Schwartz inequality and Lemma 2.6. More precisely, we deduce that
where is used to in the last inequality.
We now go to see . Recalling by (2.3), recalling (1.5) and (2.29), using , we get
(2.44) |
Here the role of is to produce to cancel angular singularity later. The factor is kept to retain -type weight. Such a treatment will be used frequently in the rest of the article. Similar to (2.44), we have
(2.45) |
Patching together (2.44) and (2.45), using (2.37), we have
Note that we get the factor to cancel singularity and also the good -type weight factor . As a result, by (2.30), using (2.3), we get
Applying (2.4) to , using (2.40), (2.37) and (2.30), we get
Using (2.27), we have which is the same as if we exchange and . Therefore we also have
Patching together the above estimates of for , we get
(2.46) |
We next go to see . It is further decomposed into two terms where
Comparing and , we can use the same arguments for and (2.42) to get Note that is the same as if we exchange and . Therefore we also have Patching together the above estimates of and , we get
(2.47) |
2.4. Estimate of
Theorem 2.3.
For , there holds
Proof.
Recalling (2.3) and following the proof of Theorem 1.1 and Theorem 2.1 in [26], we give the following remark.
Remark 2.3.
If is replaced by or in Theorem 2.3, the result is still valid.
We now derive an upper bound estimate for the linear operator defined in (1.37).
Proposition 2.2.
There holds
Proof.
Recalling (2.1) and (2.2), we rewrite the operator in (1.37) as
Taking inner product with , we have
where we use (2.28) in the second line. Then by Cauchy-Schwartz inequality and (2.4), we get
where we use Theorem 2.3 in the second line. Noting that by (2.1), we get
which gives where
By Theorem 2.3, we have Since , we get
where we use (2.27), (2.28) and Theorem 2.3. Patching together the above estimates, we finish the proof. ∎
2.5. Estimate of
We now derive a result about which will be used in Proposition 6.1 to derive non-negativity of solutions to the linear equation (1.34).
Proposition 2.3.
If , then
Proof.
Recalling (1.35), using and , we have
where
By Lemma 2.5, the imbedding and Lemma 2.6, we get
By (2.28), we have and . By (2.27), Lemma 2.5, the imbedding and Lemma 2.6, we get
By (2.4), we get
Noting the following fact
and similarly , using (2.37), (2.30) and Lemma 2.6, we get
Patching together the above estimates, we finish the proof. ∎
3. Bilinear operator estimate
Recall (2.1). For simplicity we continue to write . Recalling the relation (2.2) between and , the definition of in (1.23) and in (1.28), we have
(3.1) |
Here stands for the main term (“” is referred to “main”) defined by
(3.2) |
where is defined in (1.43) and is defined by
(3.3) |
Here represents the remaining term (“” is referred to “remaining”). This term consists of three parts corresponding to (1.28), (1.28), (1.28) respectively,
(3.4) | |||||
(3.5) | |||||
(3.6) | |||||
(3.7) |
We call the main term for two reasons. First, the factor before is , while the factor before is . Second, when , the term corresponds to the nonlinear term in the classical linearized Boltzmann equation.
3.1. The main operator
Proposition 3.1.
It holds that
(3.8) | |||||
(3.9) |
For the convenience of later reference, we write Proposition 3.1 as
Corollary 3.1.
There holds
The following lemma is used to deal with the norm of the product of two functions.
Lemma 3.1.
Let . Then
(3.10) |
Let . There exists a constant such that
(3.11) |
Recall . There holds
(3.12) |
Proof.
Proposition 3.2.
There holds
Note that Proposition 3.2 has the flexibility to balance the regularity between and . Such flexibility allows us to close energy estimate in high order Sobolev spaces.
Now we set to consider We first prepare some intermediate estimates.
Lemma 3.2.
Let . Let , then the following four estimates are valid.
(3.13) | |||||
(3.14) | |||||
(3.15) | |||||
(3.16) |
Let , then the following two estimates are valid.
(3.17) | |||||
(3.18) |
Let (which means the two sets are equal) and , then the following four estimates are valid.
(3.19) | |||||
(3.20) | |||||
(3.21) | |||||
(3.22) |
Remark 3.1.
If we replace with or in Lemma 3.2, all the results are still valid. Let be two polynomials on . If we replace and with and respectively in Lemma 3.2, all the results are still valid. Moreover, if we replace and with (or ) and (or ) respectively in Lemma 3.2, all the results are still valid. Since , if we replace with in (3.17), the result is still valid.
Remark 3.1 says that Lemma 3.2 have a more general version. This flexibility allows us to deal with various similar integrals. For example, we will use Remark 3.1 to deal with integrals involving derivatives of in subsection 3.2.
Now we are ready to prove the following upper bound estimate for the operator .
Proposition 3.3.
Let . There holds
Proof.
Note that by recalling (2.1). From which together with , we have
(3.23) |
Plugging (3.23) into (3.3), we have
(3.24) | |||||
(3.25) | |||||
(3.26) |
Now it suffices to estimate and .
Estimate of . Recalling (3.25) and using (2.28), we have
Using and we get where
(3.27) | |||||
(3.28) |
Since , there holds . By Cauchy-Schwartz inequality, (2.27) and (2.28), recalling (2.48), using (3.17) and Theorem 2.3, we get
(3.29) |
Recalling (3.27), by Cauchy-Schwartz inequality, recalling (2.48), using (3.17) and Theorem 2.3, we have
(3.30) |
Recalling (3.28), applying (3.13) and Lemma 3.1, we get
Patching together the estimates of for , we get
(3.31) |
Theorem 3.1.
It holds that
3.2. Some remark on derivative before operator
In this subsection, we address the issue of taking derivative w.r.t. variable of various Boltzmann type operators appeared in this article. The conclusion is that we can focus on derivatives of the involved functions and safely ignore the derivatives on those good functions like appearing in the operators. We begin with the following fact about Boltzmann type operator
(3.33) |
That is, when taking derivative w.r.t. , we can use the binomial formula for the product first and then take integral.
Recalling (3.2) and (3.24), we have Therefore
(3.34) |
Thanks to (3.33), we have
Note that . Then by Corollary 3.1 and Lemma 3.1, we have
(3.35) |
Note that if we use Proposition 3.2 to estimate , we will get the same result as (3.35). Therefore we conclude that to estimate , it suffices to estimate using Proposition 3.2 for all such that . That is, we can safely regard as .
Now we visit . Recalling (3.25) and using (3.33), we have
Note that when . For index , there is a polynomial such that . Observe
(3.36) |
Using (2.28) and (3.36), setting for simplicity, we have
Recall . Therefore for any index , there exists some constant such that
(3.37) |
Then by Cauchy-Schwartz inequality, using (2.27) and (2.28), we have
which can be handled like in (3.29) thanks to Remark 3.1. With the identity , both and can be split into two terms like (3.27) and (3.28). Then by following the estimate of and in Proposition 3.3, thanks to Remark 3.1, using (3.37), we will get the same upper bound. In a word, shares the same upper bound as (3.31) for . Therefore we conclude that to estimate , it suffices to consider for all such that . That is, we can safely regard general as . By nearly the same analysis, the same conclusion holds for . Recalling (3.34), we arrive at the following remark.
Remark 3.2.
In order to estimate , it suffices to consider for all such that .
Following the proof of upper bound estimates in Section 2, 3 and 4, we can go further to conclude that
Remark 3.3.
To estimate upper bound concerning , it suffices to consider for all such that . To estimate upper bound concerning , it suffices to consider for all such that . To estimate upper bound concerning and , it suffices to consider and for all such that .
3.3. The remaining operator
We first prepare some intermediate estimates.
Lemma 3.3.
Let . The following four estimates are valid.
(3.38) | |||||
(3.39) | |||||
(3.40) | |||||
(3.41) |
We will only use (3.41) in this subsection. The proof of Lemma 3.3 is given in the Appendix 8. Appropriately revising the proof of Lemma 3.3, we have
Remark 3.4.
In the rest of the article, are two constants verifying unless otherwise specified.
Recalling (3.4), there holds . We will give estimates of in Propositions 3.4, 3.5, 3.6 respectively.
Proposition 3.4.
It holds that
Proof.
Recalling (3.5), we have
(3.42) |
We need to make rearrangement in order to use previous results. By the following identity
(3.43) |
we have
(3.44) | |||
Estimate of . Using the identity , we have where
Recalling (3.3), using Proposition 3.3 and Lemma 3.1, we have
(3.46) |
By Cauchy-Schwartz inequality, have
(3.47) |
By (2.27) and (2.28), we observe
(3.48) |
where we use Theorem 2.3 and Remark 2.3. Similar to (3.23), we have
which yields and thus
where we use (2.27) for the latter integral and then get the final estimate by using (3.17). Plugging (3.48) and (3.3) into (3.47), we get
(3.50) |
Patching together (3.46) and (3.50), recalling , we have
(3.51) |
Estimate of . Using the identity , we have where
Observe . Then Corollary 3.1 and Lemma 3.1 yield
(3.52) | |||||
By Cauchy-Schwartz inequality, (2.3) and (2.27), have
(3.53) |
where we use (3.18), Remark 3.1 and Theorem 2.3. Patching together (3.52) and (3.53), we have
(3.54) |
Proposition 3.5.
It holds that
Proof.
Recalling the definition of in (3.6), we have
(3.58) |
Using the identity (3.43), we get
(3.59) | |||||
(3.60) | |||||
(3.61) | |||||
(3.62) | |||||
(3.63) |
Estimate of . Using (2.28) to (3.60), we have
(3.64) |
Recalling (2.1), there holds Similar to (3.23), we have
(3.65) |
Plugging (3.65) into (3.64), using and , we have
(3.66) | |||||
Estimate of . By Cauchy-Schwartz inequality and (2.3), we have
By (3.17) and Remark 3.1, we have
By (2.27) and Theorem 2.3, we get
By (2.27) and the estimate (3.18), we get
Patching together the previous three estimates, we have
(3.67) |
Estimate of . By Cauchy-Schwartz inequality and (2.3), we have
(3.68) |
By (3.17), we have
(3.69) |
By (2.27) and the estimate (3.18), we get
(3.70) |
Observe and thus
(3.71) |
which gives
(3.72) |
where we use (2.27), Theorem 2.3 and the estimate (3.18). Plugging (3.69), (3.70) and (3.72) into (3.68), we get
(3.73) |
Another estimate of . We now give another estimate of to put more regularity on . By Cauchy-Schwartz inequality and (2.3), we have
(3.74) |
By (3.17), we have By (2.27) and the estimate (3.18), we get
By (3.71) and (2.27), using (3.41) and (3.18), we have
Plugging the previous three estimates into (3.74), we get
(3.75) |
Estimate of . Now we set to estimate . By (2.27), using (3.13) and (3.15), we have
(3.76) | |||||
Patching together (3.67), (3.73) and (3.76), recalling (3.66), we get
(3.77) |
Patching together (3.67), (3.75) and (3.76), recalling (3.66), we get
(3.78) |
Estimate of . Recalling (3.61) and , using (2.28), we have
(3.79) | |||
By Cauchy-Schwartz inequality, (2.3) and (2.27), we have
(3.80) | |||||
where we use Theorem 2.3, the estimate (3.18) and Remark 3.1.
By (2.27), the estimate (3.15) and Remark 3.1, we have
(3.82) |
Patching together (3.80) and (3.82), recalling (3.79), we get
(3.83) |
Patching together (3.81) and (3.82), recalling (3.79), we get
(3.84) |
Estimate of . Recalling (3.62), by Cauchy-Schwartz inequality, the imbedding , (2.3), (2.28) and Theorem 2.3, we have
(3.85) | |||||
By Cauchy-Schwartz inequality, (2.3), (2.27), (2.28), the estimate (3.41) and Theorem 2.3, we have
(3.86) | |||||
Proposition 3.6.
It holds that
Proof.
Recalling the definition of in (3.7), we have
(3.90) | |||||
(3.91) | |||||
(3.92) |
Then it suffices to consider and .
Estimate of . Recalling (3.91), using (2.28), we have
(3.93) | |||
(3.94) |
By Cauchy-Schwartz inequality, the imbedding , (2.28), (2.3) and (2.37), we get
(3.95) | |||||
where we use Theorem 2.3, the estimate (3.18) and Remark 3.1 in the last line. By the same derivation, we have
(3.96) |
By (3.16)(in which play the same role) and Remark 3.1, we have
(3.97) |
Here and in the rest of the article, or unless otherwise specified.
Another estimate of . Recalling (3.94) for , we have
(3.99) | |||
By Cauchy-Schwartz inequality, (2.3), (2.28), (2.27), the estimate (2.37), we get
(3.100) | |||||
where we use the imbedding , the estimate (3.18) and Remark 3.1 in the last line. By Cauchy-Schwartz inequality, (2.28), the estimate (2.37), Theorem 2.3, the estimate (3.18) and Remark 3.1, we get
(3.101) | |||||
where we use (3.12) in the last inequality. Patching together (3.100), (3.101) and (3.97), recalling (3.99), by taking , we get
(3.102) |
Estimate of . Recalling (3.92), taking inner product, we have
(3.104) |
Recalling , we have which gives
Plugging which into (3.104), we get
(3.105) | |||||
(3.106) | |||||
(3.107) |
Since the two quantities have a similar structure, we only consider . Indeed, we will not use the factor before in and so the estimate of is similar. Using (2.28), note that
(3.108) | |||
By Cauchy-Schwartz inequality, (2.3), (2.28), (2.27), the estimate (2.37), the imbedding , the estimate (3.18), Remark 3.1 and Theorem 2.3, we have
We will deal with in two ways. On one hand, by Cauchy-Schwartz inequality, (2.3), (2.28), (2.27), the estimate (2.37), Theorem 2.3 and the estimate (3.18), we have
On the other hand, by Cauchy-Schwartz inequality, (2.3), (2.27), the estimate (2.37), the estimate (3.18) and the estimate (3.41), we have
Now we set to estimate . By Cauchy-Schwartz inequality, (2.3), (2.27), the estimate (2.37), the imbedding , the estimate (3.18), Theorem 2.3 and Remark 2.3, we have
Recalling (3.1) and (3.4), patching together Theorem 3.1, Propositions 3.4, 3.5 and 3.6, since , we get
Theorem 3.2.
It holds that
4. Trilinear operator estimate
In this section, we estimate the trilinear operator defined in (1.24). Recalling the relation (2.2) between and , we have
(4.1) | |||||
(4.2) | |||||
(4.3) |
The whole section is devoted to prove
Theorem 4.1.
The following functional estimates are valid.
(4.5) | |||||
(4.6) |
Proof of (4.1)..
Estimate of . Using (2.28), and , we get
(4.10) | |||||
By Cauchy-Schwartz inequality and (2.3), using (3.39), Remark 3.4 and Theorem 2.3, we have
(4.11) | |||||
By Cauchy-Schwartz inequality, the imbedding , (2.3), (2.27), the estimate (3.41) and Theorem 2.3, we have
(4.12) |
Note that . Then by Corollary 3.1 and Lemma 3.1, we have
(4.13) |
Recalling (4.10), patching together (4.11), (4.12) and (4.13), we get
(4.14) |
Estimate of . Using (2.28), we get
(4.15) |
Plugging the following two identities
(4.16) |
into (4.15), we have
(4.17) |
where
(4.18) | |||||
(4.19) |
By Cauchy-Schwartz inequality and (2.3), we have
By the change of variable and the estimate (3.17), we get
Recalling (3.71) and (2.27), using (3.41) and (3.18) we get
Patching together the previous two estimates, we get
By nearly the same argument as that for , thanks to the factor , we can also get
By Cauchy-Schwartz inequality and (2.3), we have
By the estimate (3.17), we get
(4.20) |
Recalling (3.71), we have
(4.21) |
By (4.21), using (3.38) and (3.17), we get
(4.22) |
Patching together (4.20) and (4.22), we get By nearly the same argument as that for , we can also get Using (3.19), for , we get
(4.23) |
(4.24) |
Recalling (4.17), patching together the above estimates of , taking , we get
(4.25) |
Recalling (4.7), patching together (4.14) and (4.25), we have
(4.26) |
We now divide into two terms. Using , we get where
(4.30) |
By Cauchy-Schwartz inequality, (2.3), the imbedding and Theorem 2.3, we have
Using Corollary 3.1 and Lemma 3.1, we have
We now divide into two terms. Using and , we get where
By Cauchy-Schwartz inequality, (2.3), (2.27), the estimate (2.37), the imbedding , Theorem 2.3, the estimate (3.18) and Remark 3.1, we have
By (2.27), using (3.22)(in which play the same role), we have
Recalling (4.28), patching together the above estimates of , we have
(4.31) |
Proof of (4.5) and (4.6)..
We first estimate . Recalling (4.8), using , we get
(4.32) |
By Cauchy-Schwartz inequality, (2.3), (3.39) and Theorem 2.3, we have
(4.33) | |||||
Observe . Then by Corollary 3.1 and Lemma 3.1, we have
(4.34) | |||||
Recalling (4.32), patching together (4.33) and (4.34), we get
(4.35) |
We then estimate . Recalling (4.9), using (4.16) and we have
(4.36) | |||||
By Cauchy-Schwartz inequality, (3.17), (3.38) and Remark 3.4, we have
By nearly the same argument as that for , thanks to the factor , we can get
By Cauchy-Schwartz inequality, (2.3), the imbedding , the estimate (3.17) and Theorem 2.3, we have
By nearly the same argument as that for , thanks to the factor , we get
Note that where is defined in (4.18). Then by taking or in (4.23), we get
Note that where is defined in (4.19). Then by taking or in (4.24), we get
Recalling (4.36), patching together the above estimates of , we get
(4.37) |
We first deal with . Recall where and are defined in (4.30). We now give new estimates to and . By Cauchy-Schwartz inequality, (2.3), (2.27), (3.40) and Theorem 2.3, we have
Observe . Then by Corollary 3.1 and Lemma 3.1, we have
We then consider . Recalling (4.29), using and the identity (4.16), we have
(4.39) | |||||
By Cauchy-Schwartz inequality, (2.3), (2.27), (2.37), using (3.18) and (3.40)(by regarding ), we have
By nearly the same argument as that for , using the factor , we get
By Cauchy-Schwartz inequality, (2.3), (2.27), (2.37), the imbedding for , Theorem 2.3, the estimate (3.18) and Remark 3.1, we have
By nearly the same argument as that for , using the factor , we get
By (3.21), we have Using (2.27) and (3.22), we get Recalling . Patching together the above estimates of , we have
(4.40) |
5. Commutator estimate and weighted energy estimate
In this section, we derive some estimates for the commutators between the weight function and the various linear, bilinear and trilinear operators whose un-weighted estimates are already given in Section 2, 3 and 4. These estimates together produce some weighted energy estimates that will be used in the next two sections.
Direct computation shows that the various commutators share a common term . We collect estimates of various functionals involving in the following lemma. In all the estimates for functionals involving in the article, the “” could bring a constant depending on on the righthand side of “”. We do not specify this dependence for brevity.
Lemma 5.1.
Let . Let and . The following estimates are valid.
(5.1) | |||||
(5.2) | |||||
(5.3) | |||||
(5.4) | |||||
(5.5) | |||||
(5.6) |
The proof of Lemma 5.1 is given in the Appendix 8. Based on the proof, we give a remark to make Lemma 5.1 more applicable.
Remark 5.1.
5.1. Commutator estimate between and
In the following proposition, we give an estimate of the commutator between and .
Proposition 5.1.
Let . The following two estimates are valid.
(5.7) | |||
(5.8) |
Proof.
Recalling (1.21), (2.2) and (2.38), we have
(5.9) | |||||
(5.10) |
By (5.9), we need to consider and . For the operator , using (2.28), we derive
where we use (5.1)(writing ). By Proposition 2.1 and (3.11), we have
(5.11) |
Patching together the previous two estimates, we get (5.7).
We now set to prove (5.8). Recalling (5.9), we need to consider and . For the operator , we already have (5.11). For the operator , recalling (5.10), using (2.28), we deduce
Note that by using (5.21). Then by Proposition 5.2(which will be proved soon), we have By Cauchy-Schwartz inequality, (2.3), (2.28) and (2.27), we have
where we use (5.1), Theorem 2.3 and Remark 2.3. Patching together the estimates of and (5.11), we get (5.8). ∎
Theorem 5.1.
Let . If , there holds
(5.12) |
where is the constant appearing in Theorem 2.2. As a direct consequence, in the full space, there holds
(5.13) |
Proof.
In the space, we have the following lower bound estimates which yield the dissipation functional in later energy estimates.
Theorem 5.2.
Recall . If , the following statements are valid. When there is only derivative, it holds that
(5.14) |
If , it holds that
Proof.
Note that where
We use (5.13) to get
By binomial formula and Remark 3.3, to estimate , it suffices to consider
for all .
Recalling (2.13) and (2.15), it is easy to see . Then by the upper bound in Theorem 2.2 and Remark 2.1, for , we have
(5.16) |
Recalling (1.39), using Cauchy-Schwartz inequality and (5.16), for , we have
(5.17) |
which gives
(5.18) |
By (5.8), we have
(5.19) |
Patching together (5.18) and (5.19), using the basic inequality , we have
Taking sum over and taking suitably small, we finish the proof. ∎
5.2. Estimate of the commutator
In this subsection, we derive an estimate of the commutator between and the operator . That is, we want to estimate Recalling (3.1), (3.2) and (3.4), we have
(5.20) |
We will estimate the five terms on the right-hand side of (5.20) in the following five propositions. First, we give an estimate for the commutator in the following proposition.
Proposition 5.2.
Let and . It holds that
Proof.
The following proposition gives an estimate for the commutator .
Proposition 5.3.
Let and . It holds that
Proof.
Recalling (3.24), (3.25), (3.26), we have
(5.22) | |||||
(5.23) | |||||
By Cauchy-Schwartz inequality, (2.27) and (2.28), using (5.1), (5.2) and Theorem 2.3, we have
(5.24) | |||||
By Cauchy-Schwartz inequality, (2.27) and (2.28), using (5.1) and (3.18), we have
(5.25) | |||||
Recalling (5.22), patching together (5.24) and (5.25), we finish the proof. ∎
The next proposition gives an estimate for the commutator .
Proposition 5.4.
Let and . It holds that
Proof.
Recalling (3.42), we have where
Recalling and (5.21), we observe
Then by Proposition 5.2, we get
(5.26) |
Note that and , we get
which gives where
(5.27) | |||||
(5.28) | |||||
(5.29) | |||||
(5.30) |
Note that the two lines (5.27) and (5.29) resemble the line (5.23). Then by (5.24) we get
(5.31) |
Note that the two lines (5.28) and (5.30) resemble (5.21). Then by Proposition 5.2, we get
(5.32) |
Recalling , patching together (5.26), (5.31) and (5.32), we finish the proof. ∎
The next proposition gives an estimate for the commutator .
Proposition 5.5.
Let and . It holds that
Proof.
Recalling (3.58), we have where
We first visit . By (2.28) and the identity , we get
(5.33) | |||||
(5.34) | |||||
By Cauchy-Schwartz inequality and (2.3), using (5.1), Remark 5.1 and Theorem 2.3, we get
(5.35) | |||||
By Cauchy-Schwartz inequality, (2.3), (2.27) and (2.28), we have
(5.36) | |||||
where in the last line we use (5.1), Remark 5.1, Theorem 2.3 and Remark 2.3. By (5.4) and Remark 5.1, we have
(5.37) |
The next proposition gives an estimate for the commutator .
Proposition 5.6.
Let . It holds that
Proof.
Recalling (3.90), we have . Recalling (3.91), it is easy to see . For the term involving , it suffices to consider for . By (3.105), (3.106) and (3.107), we have where
Since the two quantities have a similar structure, it suffices to consider . Note that only differs from in (3.107) by the two weight functions and . We can follow the derivation of the estimate (3.109) for . Observing that in the derivation there is a factor . By using the fact , we can get the same upper bound as that in (3.109). ∎
By the above commutator estimates and the upper bound estimates in Section 3, weighted upper bounds of and are given in the following theorem.
Theorem 5.3.
Let or . The following two estimates are valid.
(5.41) | |||||
(5.42) |
Proof.
In the 3-dimensional space , by imbedding or with , based on Theorem 5.3, estimates of inner product in the full space are given in the following theorem.
Theorem 5.4.
Let or . The following two estimates are valid.
(5.44) | |||||
(5.45) |
Based on Theorem 5.4, by making a suitable choice of parameters to deal with different distribution of derivative order, we get
Theorem 5.5.
Let . The following two estimates are valid.
(5.46) | |||
(5.47) |
Proof.
By binomial formula and Remark 3.2, we need to consider for all combinations of with . By (5.44), it suffices to prove that the following inequality
holds for some verifying or .
The following is divided into two cases: and for .
Case 1: . In this case, there holds and we take and use to get
Here we recall (1.48) for the definition of .
Case 2: for . We consider two subcases: and for . In the first subcase , there holds . As the same as (5.2), we get
5.3. Estimate of the commutator
In this subsection, we derive a commutator estimate between and the operator . More precisely, we will give some upper bound for the quantity Recalling (4.1), it suffices to estimate and . We will estimate in Proposition 5.7 and in Proposition 5.8.
Proposition 5.7.
The following two estimates are valid.
(5.50) | |||||
(5.51) |
Proof.
We first prove (5.50). Recalling (4.7), using and (2.27), we have
By rearrangement, we have where
(5.53) | |||||
Recalling (2.3), we have . By Cauchy-Schwartz inequality, using (4.21), (3.38), (3.17), (5.1) and Remark 5.1, we have
(5.54) |
By Cauchy-Schwartz inequality, use the imbedding for , the estimate (5.1), Remark 5.1, Theorem 2.3 and (3.12), we have
(5.55) |
By (5.4) and Remark 5.1, using (3.10), we have
(5.56) |
Recalling , patching together (5.54), (5.55) and (5.56), we arrive at (5.50).
We now prove (5.51). Recalling the line (5.3) and , we have where
Recalling (2.3), we have . By Cauchy-Schwartz inequality and (2.28), using (5.2) and (3.38), we have
(5.57) |
By Cauchy-Schwartz inequality, using (5.1), (3.17) and Remark 3.1. we have
(5.58) |
By Cauchy-Schwartz inequality, we have
(5.59) |
Here we use the imbedding for and the estimate (5.1) to deal with the former bracket. The latter bracket is estimated by using Theorem 2.3 and (3.12). By (5.6) and (3.10), we get
(5.60) |
Recalling , patching together (5.57), (5.58), (5.59) and (5.60), we arrive at (5.51). ∎
Proposition 5.8.
The following two estimates are valid.
Proof.
Recalling (4.28), we have
(5.61) | |||
Recalling (2.3), we have . We first consider . By Cauchy-Schwartz inequality, we have
The latter bracket is bounded by according to Theorem 2.3. On one hand, using the imbedding for , the estimate (5.1) and Remark 5.1, we estimate the former bracket as
On the other hand, since , using the imbedding for and the estimate (5.1), we get
Therefore we have two estimates
(5.62) |
We next consider . By Cauchy-Schwartz inequality, on one hand, we have
The latter bracket is bounded by according to Theorem 2.3. Using the imbedding for and the estimate (5.1), the former bracket is bounded by Patching together the two estimates, we get
(5.63) |
By Cauchy-Schwartz inequality, on the other hand, we have
The latter bracket is bounded by according to the estimate (5.1) and (3.11). Using (2.27) and (3.40), the former bracket is bounded by Patching together the two estimates, we get
(5.64) |
Patching together Proposition 5.7 and Proposition 5.8, recalling (4.1), we get the following proposition.
Proposition 5.9.
The following functional estimates are valid.
Patching together Proposition 5.9 and Theorem 4.1, using (3.11), we arrive at the following theorem for the weighted upper bound of the trilinear term.
Theorem 5.6.
The following functional estimates are valid.
By Sobolev imbedding in the 3-dimensional space , based on Theorem 5.6, we have the following result in the full space .
Theorem 5.7.
The following functional estimates are valid.
(5.66) | |||||
(5.67) | |||||
(5.68) |
Theorem 5.7 allows us to derive the following weighted energy estimate.
Theorem 5.8.
Let , then
Proof.
By binomial formula and Remark 3.3, it suffices to establish the following estimate for all combinations of with ,
(5.69) |
The following is divided into three cases: .
Case 1: . In this case, We use (5.67) and the condition to get
Case 2: . In this case, . We consider two subcases: and . In the first subcase, . As the same as (5.3), we use (5.67) and the condition to get (5.69).
Case 3: for . We consider three subcases: ; ; . In the first subcase, . As the same as (5.3), we use (5.67) and the condition to get (5.69).
In the second subcase, implies and so . Since , then and so by (5.49). Therefore we use (5.68) to get
In the third subcase, note that gives and so . Since , then and so by (5.49). Therefore we can use (5.66) to get
Patching together the three cases, we finish the proof. ∎
6. Local well-posedness
In this section, we will prove local existence of (1.12) or (1.20). To this end, we need local well-posedness of (1.34)(or (1.36)). We first apply Proposition 2.3 to show that the solution to (1.34) is non-negative.
Proposition 6.1.
Let . Let . If and . Let be a solution to (1.34), then .
Proof.
Let . Taking inner product with , we have
where we use in the inequality. Since , by Proposition 2.3 and the imbedding , we have
which yields
Therefore the initial condition yields for any . ∎
We next prepare two lemmas. By Proposition 2.1, Proposition 2.2 and Remark 3.3, using Cauchy-Schwartz inequality, (3.11) and (3.12), we have the following lemma for energy estimates involving and .
Lemma 6.1.
Let and , then
The following lemma gives two standard results for energy estimate involving . One is controlled by dissipation, while the other is directly controlled by energy.
Lemma 6.2.
Let . For any , it holds that
(6.1) |
Here . In addition,
(6.2) |
Proof.
Now we are ready to prove well-posedness of the equation (1.36).
Proposition 6.2.
Let and . There are universal constants such that for the following statement is valid. Suppose the initial data and the given function verifies
(6.3) |
then (1.36) has a unique solution verifying for any and
(6.4) |
for some universal constant (independent of ).
Proof.
Based on the operator estimates in Section 2, 3 and 4, we can use Hahn-Banach Theorem like in [3, 42] to prove existence.
Positivity is a direct consequence of Proposition 6.1. Indeed, note that is the solution to (1.34) with the given function and initial data . Recall that . Recall (1.29) and (1.48) for the definition of and , we have
which gives
The second inequality in (6.3) gives . By Proposition 6.1, we get for any .
As for uniqueness, it is standard to take difference and do energy estimate. Indeed, one can revise the following proof to get uniqueness naturally.
Now it remains to prove the a priori estimate (6.4). For simplicity, let be the solution to (1.36). Fix . Take two indexes and such that and . Set . Applying to both sides of (1.36), taking inner product with over , using periodic condition, we get
(6.5) |
where for simplicity in this proof
By (6.2), we have By Lemma 6.1, we have
By (5.46) and the basic inequality , we have
By Theorem 5.8 and the basic inequality , we have
Patching together the above estimates, back to (6.5), we get
We now apply Theorem 5.2 to deal with the term involving . From now on in this proof, we assume . When , by (5.14), since , we have
(6.6) |
If , by (5.2), since , we have
(6.7) |
Note that the last term in (6.7) only involves the following term with index . Since , there holds
Therefore, there are universal constants with for and a large universal constant such that
Let us take such that . Suppose verify
(6.8) |
then by (6.3) there holds By these smallness conditions on and the choice of , we have
Note that
(6.9) |
where . We arrive at
By Grönwall’s lemma and the assumption (6.3), for any , we have
Recalling (6.9), since and are universal constants, we get the desired result (6.4) with a different constant .
Let be the largest constant appearing in the above proof. According to (6.8), we set
Then if and , the above proof is valid. ∎
By iteration on the equation (1.36), we derive local well-posedness of the Cauchy problem (1.20). The local well-posedness result can be concluded as follows:
Theorem 6.1.
Let . There are universal constants such that for , if
then the Cauchy problem (1.20) admits a unique solution verifying
(6.10) |
for some universal constant .
Proof.
Let and for , is the solution to the following problem
with the initial condition
Let . Here and in the rest of this paragraph are the constants appearing in Proposition 6.2. We now set to prove the following statement. If and , then the sequence is well-defined on the time interval and has uniform(in ) estimate
(6.11) |
Obviously (6.11) is valid when since . We now use mathematical induction over . Let us assume that the partial sequence is well-defined on the time interval and satisfies the uniform estimate (6.11). In particular, for , it holds that
Since , by the definition of , we have . Now applying Proposition 6.2(in which is replaced with ), there is a unique solution verifying
where we use and in the last inequality. Now the statement about the sequence is proved.
Moreover by Proposition 6.2, we have for any and ,
(6.12) |
We now prove that is a Cauchy sequence in for some . Let for . Then for , the function solves
where
By basic energy estimate(similar to (6.5)), we have
(6.14) |
where for simplicity in this proof
Let us first deal with the inner product involving on the right-hand side of (6.14). By (6.2), we have
By Lemma 6.1, we have
Recalling (6), there are three terms involving and five terms involving . For the three terms involving , by (5.46), we have
For the five terms involving , by Theorem 5.8, we have
Now let us deal with the inner product involving on the left-hand side of (6.14). Since , we can apply Theorem 5.2. As the same as (6.6) and (6.7), it holds that
Therefore as in the proof of Proposition 6.2 making a combination of the energy inequality of order , using the fact , we get
where is a universal constant. Let us take such that . Suppose verify
(6.15) |
then by (6.11) there holds By these smallness conditions on and the choice of , we have
where we use and . Recalling (6.9), we have
For simplicity, let us define , then the above inequality is
Let , then and thus
When , recall that and thus
Note that by (6.11), then for any . Then with a different constant , we get for any ,
With a different , we get for any ,
If is small enough such that
(6.16) |
then by (6.11) there holds which gives
Suppose are small enough such that
(6.17) |
then by (6.11) there holds
Let , we get
Note that the above inequality is valid for all . As a result, we get for all ,
Note that . By (6.11), it holds that
Recalling (6.9) and , we arrive at
and so is a Cauchy sequence in . The sequence has a limit which is the solution of the Cauchy problem (1.20) verifying (6.10) thanks to (6.11) and (6.12). As for uniqueness, see the proof of Theorem 1.1 at the end of Section 7. Let be the largest constant in (6.15), (6.16) and (6.17), we set
Finally, set . Then if and , all the conditions in the proof are fulfilled. ∎
7. A priori estimate and global well-posedness
This section is devoted to the proof to Theorem 1.1. We first provide the a priori estimate for the equation (1.20) in Theorem 7.1. From which together with the local existence result in Theorem 6.1, the global well-posedness result Theorem 1.1 is constructively established by a standard continuity argument at the end of this section.
7.1. A priori estimate of a general equation
This subsection is devoted to some a priori estimate for the following equation
(7.1) |
where is a given function.
Let be a solution to (7.1). Recalling the formula (2.13) of projection operator , we denote
(7.2) |
where and
(7.3) |
Here we recall (2.15) for the definition of for fixed . Note that in (7.3) we omit . However, we should always keep in mind that are functions of originating from the solution to (7.1) for fixed .
We set and . We first recall some basics of macro-micro decomposition. Plugging the macro-micro decomposition into (7.1) and using the fact , we get
(7.4) |
Recalling (7.2), the left-hand of (7.4) is
(7.5) |
Here for and . We order the 13 functions of on the right-hand side of (7.5) as
We emphasize that depends on through . We also order the 13 functions of on the right-hand side of (7.5) as
Use to denote vector transpose. For simplicity, we define two column vectors
With these two column vectors, (7.5) becomes and thus (7.4) can be written as
Taking inner product with in the space , since depends on but not on , we get
The matrix is invertible for small and so
For simplicity, let us define
Then we get the following system that consists of 13 macroscopic equations
(7.6) |
Observing and using (5.17), the functions can be controlled as:
Lemma 7.1.
Let . It holds that
We now estimate the dynamics of in the following lemma.
Lemma 7.2.
Let . It holds that
(7.7) | |||
(7.8) |
Proof.
Recall (2.15). In , taking inner products between equation (7.1) and the five functions respectively, using we get
Since for . Recalling the definition of in (2.14), it is straightforward to see
which gives the local conservation laws
(7.9) |
By (8.6), for . Recalling , we get (7.7) and (7.8) directly from (7.9). ∎
Let us recall the temporal energy functional in [15] as
(7.10) | |||||
With Lemma 7.1 and Lemma 7.2, based on the macroscopic system (7.6), using integration by parts to balance derivative it is standard to derive
Lemma 7.3.
For a rigorous proof of Lemma 7.3, one can refer to [15]. As a result of the Poincaré inequality and Lemma 7.3, we have
Lemma 7.4.
Let . Let be a solution to (7.1) verifying
(7.13) |
then there exist two universal constant such that
(7.14) |
Proof.
We now give an interpolation result.
Lemma 7.5.
Let . For any , there exists a constant such that
Proof.
We derive the following a priori estimate for equation (7.1).
Proposition 7.1.
Proof.
Note that is a combination of several functionals. We already have from Lemma 7.4. That is, the solution verifies (7.14). We add the other functionals step by step.
Step 1: Pure -derivative without weight . Applying to equation (7.1), taking inner product with , we have
Recalling and . By the lower bound in Theorem 2.2, taking sum over , we have
(7.18) |
Then gives
Thanks to (7.12), we can choose large enough such that
and then
Step 2: Pure -derivative with weight . Applying to equation (7.1), taking inner product with , taking sum over , we have
By (5.14), we have
(7.20) |
Note that for some universal constant . Taking large enough such that and the combination gives
Step 3: Weighted mixed derivatives. We prove by mathematical induction that for any , there exist some constants , such that
(7.22) | |||
Recalling (1.29), our final goal (7.1) is a result of (7.22) by taking .
We prove (7.22) by induction on . Suppose (7.22) is true when for some , we prove it is also valid when .
Take two indexes and such that and . Set . Applying to both sides of (7.1), we have
Taking inner product with , one has
(7.23) |
By (5.2), since if , we get
By (6.1), we have
Plugging the previous two inequalities into (7.23), by taking , taking sum over and , we get
By the interpolation Lemma 7.5, we have
By taking small, we have
By our induction assumption, (7.22) is true when . That is,
(7.25) | |||
Let be large enough such that and , then gives
where for , . Thus (7.22) is proved when and so we finish the proof. ∎
7.2. A priori estimate of the quantum Boltzmann equation.
In this subsection, we derive the following a priori estimate for solutions to the Cauchy problem (1.20).
Theorem 7.1.
Let . Let . Fix . There exists a constant which is independent of and , such that if a solution to the Cauchy problem (1.20) satisfies
then verifies
(7.26) |
for some universal constant .
Proof.
Observe that solves (7.1) with . Note that (1.16) and (1.17) give
(7.27) |
Thanks to (7.27) and recalling , verifies (7.13). Then we can apply Proposition 7.1 and the three terms on the righthand of (7.1) are
It is not necessary to estimate , since the upper bound of controls naturally. By (5.47) and Theorem 5.8, we have
In view of the proof of (5.47) and Theorem 5.8, it is much easier to check
Therefore by (7.1), we get
We take small enough such that If , then
and thus
(7.28) |
Integrating (7.28) w.r.t. time, we finish the proof by recalling (7.17). ∎
7.3. Global well-posedness.
With Theorem 6.1(local well-posedness) and Theorem 7.1(a priori estimate) in hand, we are ready to prove Theorem 1.1 for global well-posedness.
Proof of Theorem 1.1.
Recall the constants in Theorem 6.1. Recall the constant in Theorem 7.1. Take . Note that and so both Theorems are valid if . Denote still by the larger one of the two universal constants in (6.10) and (7.26). Take
Now we assume and set to establish global existence and the estimate (1.33). Note that . First since and , we can apply Theorem 6.1(taking ) to conclude that the Cauchy problem (1.20) admits a unique solution verifying
Then by Theorem 7.1(taking ), the solution verifies
(7.29) |
Now we go to establish the following result. For any , the Cauchy problem (1.20) admits a solution verifying
(7.30) |
We will prove (7.30) by mathematical induction on . First, is given by (7.29). Suppose (7.30) is valid for , we now prove it is also valid for . By the assumption, the Cauchy problem (1.20) admits a solution verifying
(7.31) |
In particular, and , we can apply Theorem 6.1(taking ) to conclude that the Cauchy problem (1.20) admits a unique solution verifying
(7.32) |
By (7.31) and (7.32), the solution verifying
Then by Theorem 7.1(taking ), the solution verifies
That is, (7.30) is valid for . Sending , we get a global solution satisfying (1.33).
Note that positivity follows from Theorem 6.1.
In the last, we prove uniqueness. Uniqueness can be easily proved by using the arguments in the proof of Theorem 6.1. Indeed, suppose are two solutions to the Cauchy problem (1.20) satisfying (1.33). Let , then solves
with the initial condition . By energy estimates like in the proof of Theorem 6.1, one will get
Since are two solutions satisfying (1.33), the quantity is integrable over for any . Then by Grönwall’s inequality and using the initial condition , one has in for any and so . Now the proof is complete. ∎
8. Appendix
Recalling (2.9) and (2.10) for the definition of and . Recalling (2.9), one has
For , we recall
(8.1) |
From which, for it is easy to see
(8.2) |
Recalling (2.10), the orthonormal basis for is
Intuitively, when is small, we expect the orthogonal basis is close to , so is to . The following lemma reflects this expectation mathematically.
Lemma 8.1.
Proof.
If , one has
(8.7) |
Recall (2.1) and (2.9), we define
Taking derivative w.r.t. , using (8.7) and , recalling (8.2), we get
(8.8) |
Note that by (8.2). If , by mean value theorem,
(8.9) |
From which we get
(8.10) |
Recalling (2.9) and (2.10) for the definition of and , we have
Using (8.7), since , it is easy to see . From which together with (8.10), we get
(8.11) |
Fix , let for . Note that is independent of and by (8.2). Recalling the definition of in (2.9), there holds
Taking derivative w.r.t. , using (8.7) and , we get
(8.12) |
Recalling . When , by mean value theorem, we have
(8.13) |
From which we get
(8.14) |
Recalling (2.9) and (2.10) for the definition of and , for , we have
It is easy to check . From which together with (8.14), we have
(8.15) |
Let . Recalling (8.8), (8.9), (8.12) and (8.13), we have
(8.16) |
Taking derivative w.r.t. and using (8.16), we get
Note that by recalling . When , by mean value theorem, we have
(8.17) |
Next, recalling (2.9) and using , we have
Taking derivative w.r.t. , using (8.7) and to get recalling (8.16), we get
Note that by (8.2). If , by mean value theorem, we have
(8.18) |
From which we get
(8.19) |
By (8.18) and (8.19), the first two inequalities in (8.5) are proved. Recalling (2.9) and (2.10) for the definition of and , we have
and thus
(8.20) |
Now it remains to consider . Recall
Then
Recalling (8.7) and , using and the results in (8.17), we have
By recalling (8.20) and , we get the last inequality in (8.5).
We can revise the proof of Lemma 8.1 to get
Lemma 8.2.
Let . For , there holds
for some universal constant and a constant depending only on .
Proof of Lemma 2.3..
When and are close, we can exchange negative exponential weight freely. For example, if , then and thus . More general, we have
Lemma 8.3.
Recall . For , if , then
Lemma 8.3 is obvious since .
We estimate various integrals over involving the difference in the following lemma.
Lemma 8.4.
The following estimates are valid for any .
(8.22) | |||||
(8.23) | |||||
(8.24) | |||||
(8.25) |
Proof.
Note that when , all the results contain a type weight. This is given by Lemma 8.3. We will give a detailed proof to (8.22) and then only sketch the proof of the other three results.
Applying Taylor expansion, we have
(8.26) |
where . We consider two cases: and .
Case 1: . By (8.26), where
Noting that , using (2.33), the estimate (2.30), the fact and Lemma 8.3, we get
By the fact , Lemma 8.3 and the estimate (2.30), we have
Patching together the previous two estimates, since , we get
(8.27) |
Case 2: . We divide the integral into two parts according to and as where
It is obvious where
Note that
(8.28) |
which gives The analysis of is a little bit technical. We will make use of the nice weight on . Since and , we have
Plugging which into and using (8.28), we get
Note that and share the same upper bound and so Plugging (8.26) into , we get where
Using (2.33) and , we have
where we use
(8.29) |
We will make use of the nice weight on in . Since and , we have
From which together with (8.29), we get
Patching together the estimates of and , we get Therefore when , we get
(8.30) |
Patching together (8.27) and (8.30), we finish the proof of (8.22).
Thanks to the additional factor on the left-hand side of (8.23), we can use (2.37) to retain the weight for the case and get the desired result (8.23).
Now we explain how to prove (8.24). Note that contains order 2 cancelation by first order Taylor expansion. For the case , applying Taylor expansion to up to first order, using the fact and Lemma 8.3, we can get the result. For the case , like in the proof of (8.22), we divide the integral into two parts according to and . For the part with , we can use the same estimate for . For the part with , applying Taylor expansion to up to first order, we can use the same estimate for
The exponents of in Lemma 8.4 can be relaxed as long as they have a lower bound. Since is enough for our purpose, we give
Lemma 8.5.
Let , then the following estimates are valid.
(8.31) | |||||
(8.32) | |||||
(8.33) | |||||
(8.34) |
Proof.
Since , we can follow the proof of Lemma 8.4 to get the desired results. ∎
Remark 8.1.
By Hardy-Sobolev-Littlewood inequality, we can derive
Lemma 8.6.
Let . There holds
(8.35) |
Let . There holds
(8.36) |
Proof of Lemma 3.2.
We first prove (3.13). By (8.31) and (8.36), we have
By the same argument as the above, using the imbedding (for or or ), we can get (3.19). By (8.32) and (8.36), we can similarly prove (3.14) and (3.15). By (8.32), using the imbedding and (8.36), we can similarly prove (3.20). By (8.32), using the imbedding for and (8.36), we can similarly get (3.21). By the same argument, we can also get (3.22). By (8.32), using the imbedding for and Lemma 2.6, we can get (3.16).
Proof of Lemma 3.3.
We will give a detailed proof to (3.38) and sketch the proof of the other three results by pointing out the differences.
According to and , we write where
By Taylor expansion, we have
(8.37) |
Estimate of . By (8.37), we get
(8.38) |
Since , by Lemma 8.3, we have . For any fixed , using the following change of variable
(8.39) |
the fact and the estimate (2.30), we get
where we use (8.35) in the last inequality.
Estimate of . According to and , we write where
(8.40) | |||
(8.41) |
It is obvious where
Recalling (8.28), we have
where we use
(8.42) |
The analysis of is a little bit technical. We will make use of the change of variable and the nice weight on . Since and , we have
(8.43) |
Plugging which into , by the change of variable (take in (8.39) and let ) and using (8.28), we get
Patching together the estimates of and , we arrive at We now go to see . Plugging (8.37) into , we get
(8.44) |
We will use the change of variable according to (8.39). Recall and , then
(8.45) | |||
which yield
where we use the change of variable in the second line and (8.29) in the last line. Patching together the estimates of and , we get Patching together the estimates of and , we finish the proof of (3.38).
Now let us see how to get (3.39) by revising the proof of (3.38). Keep in mind the additional term . For where , from (8.38) we just take out and thus we do not need the change of variable . Finally we will get . Let us turn to where . Let us first see in (8.41). We simply use and then just use the same argument as that for . Finally we will get . Now it remains to see defined in (8.40). From (8.44) we just take out and thus we do not need the change of variable . The line (8.45) is revised to
and we will get . Patching together the three parts, we get (3.39).
Comparing to (3.39), there is an additional factor in (3.40) and so we can also get weight for and by using (2.37). We omit the details here.
Now let us see how to get (3.41) by revising the proof of (3.38). Now we have the factor instead of . In where , by using Lemma 8.3 to get , we can get exactly the same upper bound. As for where , we indicate the differences. Firstly, (8.42) is revised to
(8.46) |
Secondly, (8.43) is revised to
(8.47) |
Thirdly, the line (8.45) is revised to
(8.48) |
Note that in all the three estimates (8.46), (8.47) and (8.48), we have polynomial weight for and negative exponential weight for . Therefore, we can get (3.41). ∎
The following lemma gives upper bounds of various integrals over involving the difference .
Lemma 8.7.
Let . The following estimates are valid.
(8.49) | |||||
(8.50) | |||||
(8.51) |
Proof.
Applying (2.31) to , we get where
Using (2.33) and (2.30), since we get
If , then . If , then by Lemma 8.3 and since . Patching together these two cases, we get
(8.52) |
Using to get
(8.53) |
If , then and . As a result, we have
If , then by Lemma 8.3 and since . Using , we get . As a result, we have
Patching these two cases together, using (2.30), we get
(8.54) |
Proof of Lemma 5.1..
Acknowledgments. Yu-Long Zhou was supported by National Key R&D Program of China under the grant 2021YFA1002100 and NSF of China under the grant 12001552. The author is indebted to Prof. Ling-Bing He for his continuous encouragement and supervision. Great gratitude goes to Prof. Xuguang Lu for his insightful comments, especially in the scope and focus of the article.
References
- [1] R. Alexandre, On some related non homogeneous 3D Boltzmann models in the non cutoff case, Journal of Mathematics of Kyoto University, 40 (2000), pp. 493–524.
- [2] R. Alexandre, L. Desvillettes, C. Villani, and B. Wennberg, Entropy dissipation and long-range interactions, Archive for Rational Mechanics and Analysis, 152 (2000), pp. 327–355.
- [3] R. Alexandre, Y. Morimoto, S. Ukai, C.-J. Xu, and T. Yang, Global existence and full regularity of the Boltzmann equation without angular cutoff, Communications in Mathematical Physics, 304 (2011), pp. 513–581.
- [4] , The Boltzmann equation without angular cutoff in the whole space: I, Global existence for soft potential, Journal of Functional Analysis, 262 (2012), pp. 915–1010.
- [5] G.-C. Bae, J. W. Jang, and S.-B. Yun, The relativistic quantum Boltzmann equation near equilibrium, Archive for Rational Mechanics and Analysis, 240 (2021), pp. 1593–1644.
- [6] D. Benedetto, F. Castella, R. Esposito, and M. Pulvirenti, Some considerations on the derivation of the nonlinear quantum Boltzmann equation, Journal of Statistical Physics, 116 (2004), pp. 381–410.
- [7] , Some considerations on the derivation of the nonlinear quantum Boltzmann equation II: the low density regime, Journal of Statistical Physics, 124 (2006), pp. 951–996.
- [8] , A short review on the derivation of the nonlinear quantum Boltzmann equations, Communications in Mathematical Sciences, 5 (2007), pp. 55–71.
- [9] , From the N-body Schrödinger equation to the quantum Boltzmann equation: a term-by-term convergence result in the weak coupling regime, Communications in Mathematical Physics, 277 (2008), pp. 1–44.
- [10] D. Benedetto, M. Pulvirenti, F. Castella, and R. Esposito, On the weak-coupling limit for bosons and fermions, Mathematical Models and Methods in Applied Sciences, 15 (2005), pp. 1811–1843.
- [11] M. Briant and A. Einav, On the Cauchy problem for the homogeneous Boltzmann–Nordheim equation for bosons: local existence, uniqueness and creation of moments, Journal of Statistical Physics, 163 (2016), pp. 1108–1156.
- [12] S. Cai and X. Lu, The spatially homogeneous Boltzmann equation for Bose–Einstein particles: Rate of strong convergence to equilibrium, Journal of Statistical Physics, 175 (2019), pp. 289–350.
- [13] S. Chapman and T. G. Cowling, The mathematical theory of non-uniform gases: an account of the kinetic theory of viscosity, thermal conduction and diffusion in gases, Cambridge university press, 1990.
- [14] J. Dolbeault, Kinetic models and quantum effects: A modified Boltzmann equation for Fermi-Dirac Particles, Archive for Rational Mechanics and Analysis, 127 (1994), pp. 101–131.
- [15] R. Duan, On the Cauchy problem for the Boltzmann equation in the whole space: Global existence and uniform stability in , Journal of Differential Equations, 244 (2008), pp. 3204–3234.
- [16] R. Duan, S. Liu, S. Sakamoto, and R. M. Strain, Global mild solutions of the Landau and non-cutoff Boltzmann equations, Communications on Pure and Applied Mathematics, 74 (2021), pp. 932–1020.
- [17] L. Erdős, M. Salmhofer, and H.-T. Yau, On the quantum Boltzmann equation, Journal of Statistical Physics, 116 (2004), pp. 367–380.
- [18] M. Escobedo, S. Mischler, and M. A. Valle, Homogeneous Boltzmann equation in quantum relativistic kinetic theory, Electronic Journal of Differential Equations, 4 (2003), pp. 1–85.
- [19] M. Escobedo, S. Mischler, and J. J. L. Velázquez, On the fundamental solution of a linearized Uehling–Uhlenbeck equation, Archive for Rational Mechanics and Analysis, 186 (2007), pp. 309–349.
- [20] , Singular solutions for the Uehling–Uhlenbeck equation, Proceedings of the Royal Society of Edinburgh Section A: Mathematics, 138 (2008), pp. 67–107.
- [21] M. Escobedo and J. Velázquez, On the blow up and condensation of supercritical solutions of the Nordheim equation for bosons, Communications in Mathematical Physics, 330 (2014), pp. 331–365.
- [22] , Finite time blow-up and condensation for the bosonic Nordheim equation, Inventiones Mathematicae, 200 (2015), pp. 761–847.
- [23] P. Gressman and R. M. Strain, Global classical solutions of the Boltzmann equation without angular cut-off, Journal of the American Mathematical Society, 24 (2011), pp. 771–847.
- [24] Y. Guo, Classical solutions to the Boltzmann equation for molecules with an angular cutoff, Archive for Rational Mechanics and Analysis, 169 (2003), pp. 305–353.
- [25] , The Vlasov-Poisson-Landau system in a periodic box, Journal of the American Mathematical Society, 25 (2012), pp. 759–812.
- [26] L.-B. He and Y.-L. Zhou, Asymptotic analysis of the linearized boltzmann collision operator from angular cutoff to non-cutoff, Annales de l’Institut Henri Poincaré C, (2022).
- [27] N. Jiang, L. Xiong, and K. Zhou, The incompressible Navier-Stokes-Fourier limit from Boltzmann-Fermi-Dirac equation, arXiv preprint arXiv:2102.02656, (2021).
- [28] W. Li and X. Lu, Global existence of solutions of the Boltzmann equation for Bose–Einstein particles with anisotropic initial data, Journal of Functional Analysis, 276 (2019), pp. 231–283.
- [29] P. L. Lions, Compactness in Boltzmann’s equation via Fourier integral operators and applications. III, Journal of Mathematics of Kyoto University, 34 (1994), pp. 539–584.
- [30] X. Lu, A modified Boltzmann equation for Bose–Einstein particles: isotropic solutions and long-time behavior, Journal of Statistical Physics, 98 (2000), pp. 1335–1394.
- [31] , On spatially homogeneous solutions of a modified Boltzmann equation for Fermi–Dirac particles, Journal of Statistical Physics, 105 (2001), pp. 353–388.
- [32] , On isotropic distributional solutions to the Boltzmann equation for Bose-Einstein particles, Journal of Statistical Physics, 116 (2004), pp. 1597–1649.
- [33] , The Boltzmann equation for Bose–Einstein particles: velocity concentration and convergence to equilibrium, Journal of Statistical Physics, 119 (2005), pp. 1027–1067.
- [34] , On the Boltzmann equation for Fermi–Dirac particles with very soft potentials: Averaging compactness of weak solutions, Journal of Statistical Physics, 124 (2006), pp. 517–547.
- [35] , On the Boltzmann equation for Fermi–Dirac particles with very soft potentials: Global existence of weak solutions, Journal of Differential equations, 245 (2008), pp. 1705–1761.
- [36] , The Boltzmann equation for Bose-Einstein particles: condensation in finite time, Journal of Statistical Physics, 150 (2013), pp. 1138–1176.
- [37] , The Boltzmann equation for Bose–Einstein particles: regularity and condensation, Journal of Statistical Physics, 156 (2014), pp. 493–545.
- [38] , Long time convergence of the Bose–Einstein condensation, Journal of Statistical Physics, 162 (2016), pp. 652–670.
- [39] , Long time strong convergence to Bose-Einstein distribution for low temperature, Kinetic and Related Models, 11 (2018), pp. 715–734.
- [40] X. Lu and B. Wennberg, On stability and strong convergence for the spatially homogeneous Boltzmann equation for Fermi-Dirac particles, Archive for Rational Mechanics and Analysis, 168 (2003), pp. 1–34.
- [41] J. Lukkarinen and H. Spohn, Not to normal order–notes on the kinetic limit for weakly interacting quantum fluids, Journal of Statistical Physics, 134 (2009), pp. 1133–1172.
- [42] Y. Morimoto and S. Sakamoto, Global solutions in the critical Besov space for the non-cutoff Boltzmann equation, Journal of Differential Equations, 261 (2016), pp. 4073–4134.
- [43] L. Nordhiem, On the kinetic method in the new statistics and application in the electron theory of conductivity, Proceedings of the Royal Society of London. Series A, Containing Papers of a Mathematical and Physical Character, 119 (1928), pp. 689–698.
- [44] Z. Ouyang and L. Wu, On the quantum Boltzmann equation near Maxwellian and vacuum, arXiv preprint arXiv:2102.00657, (2021).
- [45] H. Spohn, Kinetics of the Bose–Einstein condensation, Physica D: Nonlinear Phenomena, 239 (2010), pp. 627–634.
- [46] R. M. Strain and Y. Guo, Almost exponential decay near Maxwellian, Communications in Partial Differential Equations, 31 (2006), pp. 417–429.
- [47] E. A. Uehling and G. Uhlenbeck, Transport phenomena in Einstein-Bose and Fermi-Dirac gases. i, Physical Review, 43 (1933), p. 552.
- [48] S. Ukai, On the existence of global solutions of mixed problem for non-linear Boltzmann equation, Proceedings of the Japan Academy, 50 (1974), pp. 179–184.