-Moment Estimates for the Singular -Laplace Equation and Applications
Abstract.
We provide -moment estimates on annuli for weak solutions of the singular -Laplace equation where and are conjugates.
We derive -uniform integrability for some critical parameter range.
As a application, we derive a mass conservation as well as a weak convergence result for a larger critical parameter range.
Concerning the latter point, we further provide a rate of convergence of order of the solution in the -Wasserstein distance.
Keywords: Singular -Laplace Equation; Moment Estimates; Mass Conservation; Convergence Rate in Wasserstein Distance.
1. Introduction
We consider the Cauchy problem for the parabolic -Laplace equation posed in the whole Euclidean space
(1.1) |
with a positive Radon measure as initial data and where the critical value is defined as
Solutions are understood in the weak sense on the whole space and finite time horizon where .
For , such Cauchy problem is know as the fast diffusion or singular equation. Existence and uniqueness and regularity of weak solutions with various conditions for initial data are studied by Di Benedetto and Herrero [6], Junning [10], Chen and Di Benedetto [5]. For further insights about parabolic -Laplace equation, we refer to [3, 15]. While for a Dirac measure as initial data, this Cauchy problem does not admit a positive solution for any , see [10], it is known that for , there exists a fundamental solution called Barenblatt source-type solution, see [15].
In the context of optimal matching problem, Ambrosio et al. [1] draw a link between -Wasserstein distance and Poisson equation through linearization of Monge-Ampère equation and regularize empirical measures with heat semigroup. They suggest that such a linearization for the -Wasserstein distance leads to -Laplacian where and are conjugates. Such an approach has been used in a different context by Evans and Gangbo [7].
Based on these ideas, it seems natural to regularize empirical measures with -Laplacian semigroup. And further, to connect the Cauchy problem (1.1) to flows in the -Wasserstein space, as done in Kell [12]. A key point in this connection concerns the behavior of -moment of weak solutions of Cauchy problem (1.1), namely,
For the degenerate case where , that is, , such -moments are easily estimated, at least for initial data with finite mass and compact support, due to the finite propagation property, see [11]. However, for the singular case , solutions diffuse everywhere even with a Dirac measure as initial data. Hence, the focus on the parameter range .
Following similar methods as in [6, 10], our main result are the following -moment estimate for weak solutions of (1.1) on annuli
(1.2) |
for any , where is a constant depending only on and and is the polynomial
This estimate shows in particular that with a finite -moment Radon measure as initial data, there exists a weak solution which has uniformly bounded -moments on all compact time interval for any , where the critical value is the largest root of the polynomial . This critical range is sharp in the sense that the Barenblatt source-type solution has finite -moment if and only if .
As applications, we show that for such a weak solution with a finite -moment Radon measure as initial data, the mass conservation as well as weak convergence at holds for any . Finally, we obtain a -Wasserstein convergence rate for weak solutions with a finite -moment Radon measure as initial data for any , namely,
(1.3) |
where is a constant depending only on , , and . In particular, when converges to , the convergence rate coincides with the well-known heat flow case in [1].
To the best of our knowledge, there does not exists -moment estimates for singular -Laplacian equation. However, similar results in terms of mass conservation and connection to -Wasserstein space has been done. The mass conservation of singular parabolic -Laplace equation for is proved by Fino et al. [8] for any weak solution where the initial data are Radon measures with compact support. Our result only requires finite -moment instead of compact support for initial data, but for any weak solution constructed through mollification. As for the connection to the flow in the -Wasserstein space, the work of Kell [12] is the closest to the present one. There, the author shows that a smooth solution of Cauchy problem (1.1) solves a generalized gradient flow problem of Rényi entropy functional in the -Wasserstein space based on gradient flow and functional analysis methods. The author further provides a condition for the mass conservation of the flow with an integrable function as initial data for in a general metric space. In contrast, our -moment estimate (1.2) is a local estimate based on pure PDE and analysis method, and, while holding for , our result shows that mass conservation holds for a different range .
The paper is organized as follows: in Section 2, we introduce notations and definitions and auxiliary Lemmas as well as the critical parameters , and . Section 3, is dedicated to the main theorem for the local -moment estimate (1.2). Section 4 addresses the mass conservation and the weak convergence. Finally, in Section 5 we prove the Wasserstein convergence rate.
2. Notation and preliminary results
On equipped with the Euclidean norm , we denote by the closed ball centered at with radius , and set . We denote by be the closed annulus centered at with radius and . Let further be a measurable set. For , we denote by the space of -integrable functions with respect to the Lebesgue measure and set . We denote by the space of first-order Sobolev space and set . For a function in , we denote by the weak derivative of . For and , we denote by the space of functions which belong to for all intervals and compact subsets of . For or , we denote by and in a Bochner sense. We further denote by the set of positive Radon measures on .
Definition 2.1.
We say that a measurable function is a weak solution of the Cauchy problem (1.1) with -initial data, that is
(2.1) |
if for every bounded open set , it holds
(2.2) |
and for all and such that for all for some ,
(2.3) |
and for all ,
(2.4) |
For the existence and uniqueness of weak solution of (1.1) with initial data in , we refer readers to [6]. In particular, if the initial data is in , then the weak solution is unique and satisfies that , and and . In this case, the condition (2.3) is equivalent to the following condition:
(2.7) |
Furthermore, is locally -Hölder continuous on for all bounded subsets and , with depending only on and supremum norm on , see [5]. Moreover, if initial data , then the weak solution for all .
For the existence of weak solution of (1.1) with a positive Radon measure as initial data, we refer readers to [10, Theorem 1] and [6, Theorem III.8.1]. However, to our knowledge, in this case the uniqueness of weak solution is unknown and depends on the choice of the definition of the solution (viscosity, entropy, distributional, etc.).
Recall that a sequence of positive Radon measures converges vaguely, or weakly, to in if for every in , or for every in , respectively.
We briefly recall two important Lemmas for the proofs that are formulated in our notations.
Lemma 2.2.
Lemma 2.3.
([9, Lemma 3.1]) Let be a bounded function with . Suppose that for any with , we have
where are nonnegative constants with . Then for any with , it holds
for some positive constant depending only on and .
Finally, recall the following critical values
(2.10) |
Note that since , it holds
3. Moments estimates
In this section, we derive -moment estimates for a weak solution of (1.1) with finite positive Radon measure as initial data. We denote by the Hölder conjugate number of , that is, , and denote by the constant . Given a , we say that has finite -moment if . We begin to state our main result in this section.
Theorem 3.1.
(-moment estimates.) Let and be in . Then for any and , there exists a weak solution of (1.1) with as initial data satisfying
(3.1) |
for some constant depending only on and . Furthermore, if in has finite total mass and finite -moment, and , then for any , the family has uniformly bounded -moment on and that
(3.2) |
Remark 3.2.
Let be a family of -valued random variables such that the law of is . Then equality (3.2) is equivalent to say that is uniformly integrable for any .
Remark 3.3.
Note that is a sharp range for in the following sense: let be the Barenblatt source-type solution, that is,
(3.3) |
where for and some positive constant . It is well-known that is a weak solution of (1.1) with initial data where . Computations show that has finite -moment if and only if .
To show Theorem 3.1, we first establish the following auxiliary lemma.
Lemma 3.4.
Let and be a smooth function such that for some bounded domain . Let be the positive weak solution of (1.1) with initial data . Then there exists a positive constant such that for any ,
where and if and if .
Proof.
Let be fixed and and . Note that is bounded since .
On the one hand, by the regularity of on it follows that
Since , with in and in and boundedness of , we can integrate over on both sides and get
(3.4) |
By the regularity of it follows that
(3.5) |
For the first term on the right hand side of (3.5), applying Young’s inequality for and
Equation (3.5) yields
(3.6) |
Since is in and is bounded and , it follows that
Hence, integrating (3.6) over yields
(3.7) |
We address now the proof of Theorem 3.1. In the following proof, we denote by a generic positive constant depending only on and .
Proof of Theorem 3.1.
In a first step, we show inequality (3.1) in the case where is the positive weak solution of (1.1) with initial data in and . Let and . Let denote closed annulus and respectively. Let be a smooth cut-off function such that in and in and in and in . Let be the test function where is indicator function. Then for any , by (2.7), it follows that
(3.8) |
For the second term on the right hand side of (3.8), since in and and in and respectively, and note that , it follows that
(3.9) |
Plugging (3.9) into (3.8) and use the fact that and yields
(3.10) |
Let for . We consider the second term on the right hand side of (3.10) for different cases of :
-
Case 1:
If , then by Hölder’s inequality, it follows that
(3.11) where and .
For the first term on the right hand of (3.11), by Lemma 3.4 and upper bounds of in and , it follows that
(3.12) For the first term on the right hand side of (3.12), since , Hölder’s inequality yields
(3.13) The second term on the right hand side of (3.12) is the same as the first term, replacing by . As for the third term on the right hand side of (3.12), Hölder’s inequality yields
(3.14) We now turn to the second term on right hand side of (3.11) for which holds
(3.15) Plugging (3.13), (3.14) into (3.12), and then applying inequality for to (3.12), and plugging it into (3.11) together with (3.15) and taking , it follows that
(3.16) where . Applying Young’s inequaliy and to the first and second terms of the right hand side of (3.16), respectively, whereby , and using identity , it follows that
(3.17) -
Case 2:
If , by Hölder’s inequality, it follows that
(3.18) where and .
For the first term on the right hand side of (3.19), note that we have
Hence, applying inequality and upper bounds of in and and , it follows that
(3.20) By the fact that when and and , it follows from (3.20) that
(3.21) As for the second term on the right hand side of (3.19), note that , which by Hölder’s inequality yields
(3.22) Note that if , it is obvious that . If , since , then it holds that
Taking on the right hand side of (3.22), yields
(3.23) Plugging (3.21) and (3.23) into (3.19) and applying inequality yields
(3.24) As for the second term on the right hand side of (3.18), Hölder’s inequality and then letting , yields
(3.25) Plugging (3.24) and (3.25) into (3.18), it follows that
(3.26) Applying Young’s inequality as well as to the first and second terms on the right hand side of (3.26) respectively, where , it follows that
(3.27)
Together with (3.17) and (3.27), we deduce from (3.10) that
(3.28) |
Inequality (3.28) holds by replacing by any in implying that
(3.29) |
By Lemma 2.2, it holds that
Hence has bounded value on . Applying Lemma 2.3 for and , it follows that
(3.30) |
for some . Thus we prove inequality (3.1) for the case where the initial data is in .
Let us finally address the inequality (3.1) in the general case where is taken as initial data. Let be a mollifier function with and and . Let be a smooth cut-off function such that in and in . Let be defined as follows:
Obviously, . Moreover, for any , since , it follows that
(3.31) |
Furthermore, for any , take large enough such that , where . Then it follows that
where . By the uniform convergence of to on compact sets, see [4, Proposition 4.21], it follows that
(3.32) |
Let be the unique positive weak solution of (1.1) and (2.1) with initial data for . Then by inequality (2.9) of Lemma 2.2 and (3.31), it follows that is locally equibounded in , that is, for any and bounded domain , is equibounded in . Then by [5, Theorem 1], it follows that is uniformly equicontinuous in for all and all bounded domains . By diagonalization procedure, we can find a subsequence of , which is relabelled by , such that
(3.33) |
and is a positive weak solution of (1.1) and (2.5) with initial data , see [6, Theorem III.8.1].
Now for any and , by inequality (3.30), it follows that for any ,
(3.34) |
For the left hand side of (3.34), by dominated convergence and (3.33), it follows that
(3.35) |
For the first term on the right hand side of (3.34), let and such that on and on . Then taking and by (3.32), it follows that
(3.36) |
For together with (3.35) and (3.36), it follows that for any ,
Taking the supremum over on the left hand side yields the desired result.
We are left to show uniform boundedness of -moment and inequality (3.2). Assume that is in with finite total mass and finite -moment, and is such that . By inequality (2.8) in Lemma 2.2 and similar arguments, see also [10, Theorem 1], it follows that for any and ,
(3.37) |
Taking and together with (3.1), it follows that for any
Since , taking and it follows that
From this we obtain that has uniformly bounded -moment for any . For inequality (3.2), taking for both side of (3.1) and it follows that
Since has finite -moment, taking and it follows that
∎
4. Mass conservation and weak convergence
In [8], the authors show that mass conservation of any positive weak solution of (1.1) and (2.5) with initial data in with compact support. In this section, we will show that the weak solution constructed in Theorem 3.1 with initial data in with finite total mass and finite -moment, preserves mass. As a by-product, measure converges to weakly.
Theorem 4.1 (Mass Conservation).
Let and be a positive finite Radon measure in and is the positive weak solution in Theorem 3.1. If has -finite moment, then preserves mass, that is,
(4.1) |
Corollary 4.2 (Weak Convergence).
Let be a positive finite Radon measure in with finite -moment and be the positive weak solution in Theorem 3.1. Then the measure converges weakly to the initial data as goes to .
We start with a lemma providing gradient estimate of the weak solution of (1.1) with respect to -moment of smooth and compact-supported initial data on the annulus.
Lemma 4.3.
Let be the positive weak solution of (1.1) with initial data in and . Then there exists a positive constant such that for any and , it holds that
(4.2) |
where .
Proof.
Let and denote by the annulus for . Let be a smooth cut-off functions such that on and on , and on and on . Since on , it follows that
(4.3) |
We consider the right hand side of (4.3) for different cases of :
- Case 1:
- Case 2:
Hence, from (4.5) and (4.7), we obtain that for any , it holds that
(4.8) |
Let and and and denote by the terms for . By applying inequality (3.1) in Theorem 3.1 to and inequality for , it follows that
(4.9) |
By using inequality for and respectively, we have and . Then for the first two terms on the right hand side in (4.9), it follows that
Plugging it into (4.9), we obtain
which together with (4.3) yields the result. ∎
Proof of Theorem 4.1.
Let with finite total mass be fixed and be the sequence of functions constructed in the proof of Theorem 3.1, and be the positive weak solution of (1.1) with as initial data, and be the positive weak solution constructed in the proof of Theorem 3.1. Let and be a smooth cut-off function such that on and on and on . By using as the test function in (2.7), it follows that for any and ,
(4.10) |
By Lemma 4.3 with and , it follows that
(4.11) |
By (3.33) in the proof of Theorem 3.1, it holds that converges to , up to a subsequence, uniformly on each compact subset of . Together with equality (3.32), taking on both side of (4.11) and it follows that
(4.12) |
For the left hand side of (4.12), by letting on both side of (3.37) in the proof of Theorem 3.1, it follows that . Together with , by dominated convergence theorem, it follows that
(4.13) |
For the right hand side of (4.12), it is easy to check that for ,
Since has finite -moment, the term on the right hand side of (4.12) converges to as . So we obtain that for any ,
∎
5. Convergence rate of Wasserstein distance
We address now the convergence rate of the constructed weak solution to in the -Wasserstein distance. We recall that the -Wasserstein distance between finite Borel measures in with equal mass is defined as
(5.1) |
where is the family of all Borel measures on having and as their first and second marginal measures respectively.
Theorem 5.1.
To address the proof of Theorem 5.1, we first establish an auxiliary lemma, prove the theorem in the case where the initial data is in ), and then show the general case by an approximation procedure. The key ingredients of the proof are to use Brenier-Benamou formulation for -Wasserstein distance and write the parabolic -Laplace equation as a law of mass conservation where .
Lemma 5.2.
Let and be the positive weak solution of Cauchy problem (1.1) with and as initial data. Then for any , it holds that
(5.3) |
for some .
Proof.
Let and . Let be a smooth cut-off function such that on and on and on . Let and and . Multiplying on (1.1) and integrating it over , by integral by part and similar argument in Lemma 3.4, it follows that
(5.4) |
Note that implies that . Hence, for the first term on the right hand side of (5), by using the similar argument as in the proof of Theorem 3.1, it follows that
(5.5) |
As for the second term on the right hand side of (5), we take and . By Hölder’s inequality and for , it follows that
(5.6) |
Plugging (5) and (5) into (5), taking and using monotone convergence theorem as well as identity yields
(5.7) |
where . Using Inequality (3.1) from Theorem 3.1 and (2.8) from 2.2 in the right hand side of (5), it follows that
(5.8) |
where . Since , taking on both side of (5.8), monotone convergence theorem and -finite moment of yields the result. ∎
Proof of Theorem 5.1.
We first show that inequality (5.2) holds in the case where is the positive weak solution of (1.1) with in and as initial data. Let and . By Benamou-Brenier formulation for -Wasserstein, see [14, Theorem 5.28] and [2], it follows that
Rescaling as and and changing variable yields
Note that by the property of singular -Laplace equation, for all . So choosing and for , it follows that
(5.9) |
We are left to show that inequality (5.2) holds in the case where is the positive weak solution of the Cauchy problem (1.1) with finite Radon measure with finite -moment as initial data, constructed in the proof of Theorem 3.1. Let be the sequence of smooth initial data in defined in the proof of Theorem 3.1 and be the corresponding positive weak solution, and be the subsequence of such that and converges to and uniformly on all compact subset of . Let be fixed. For , by triangle inequality, it follows that
(5.10) |
where and .
As for the first term on the right hand side of (5.10), we claim that converges to weakly and that as . Indeed, for all and , by construction of and Fubini’s theorem, it follows that
(5.11) |
Hence, as . Choose large enough such that and and let in such that on and on . Since converges to vaguely by (3.32), choose large enough such that
(5.12) |
Hence, as . By classical result [13, Theorem 13.16], converges to weakly. Together with (5), by [16, Theorem 7.12], it follows that as .
For the third term on the right hand side of (5.10), we claim that converges to weakly and that as . Indeed, by inequality (3.1) of Theorem 3.1, it follows that
(5.13) |
Applied to (5), it follows that as . Furthermore, using inequality (5.13) as well as (3.2), choose large enough such that and . Then since converges to uniformly on all compact subsets on , choose large enough, such that
By the same argument as previously, it follows that as .
References
- Ambrosio et al. [2019] Luigi Ambrosio, Federico Stra, and Dario Trevisan. A pde approach to a 2-dimensional matching problem. Probability Theory and Related Fields, 173(1):433–477, 2019.
- Benamou and Brenier [2000] Jean-David Benamou and Yann Brenier. A computational fluid mechanics solution to the monge-kantorovich mass transfer problem. Numerische Mathematik, 84(3):375–393, 2000.
- Benedetto [1993] Emmanuele Di Benedetto. Degenerate Parabolic Equations. Universitext. Springer, 1993.
- Brezis [2010] Haim Brezis. Functional Analysis, Sobolev Spaces and Partial Differential Equations. Universitext. Springer-Verlag New York, 2010.
- Chen and Di Benedetto [1988] Ya-Zhe Chen and Emmanuele Di Benedetto. On the local behavior of solutions of singular parabolic equations. Archive for Rational Mechanics and Analysis, 103(4):319–345, 1988.
- Di Benedetto and Herrero [1990] Emmanuele Di Benedetto and Miguel A. Herrero. Non-negative solutions of the evolution p-Laplacian equation. Initial traces and Cauchy-problem when . Archive for Rational Mechanics and Analysis, 111(3):225–290, 1990.
- Evans and Gangbo [1999] Lawrence C Evans and Wilfrid Gangbo. Differential equations methods for the Monge-Kantorovich mass transfer problem. Number 653. American Mathematical Soc., 1999.
- Fino et al. [2014] Ahmad Z. Fino, Fatma Gamze Düzgün, and Vincenzo Vespri. Conservation of the mass for solutions to a class of singular parabolic equations. Kodai Math. J., 37(3):519–531, 10 2014.
- Giaquinta [1983] Mariano Giaquinta. Multiple Integrals in the Calculus of Variations and Nonlinear Elliptic Systems. Annals of Mathematics Studies 105. Princeton University Press, 1983.
- Junning [1995] Zhao Junning. The cauchy problem for when . Nonlinear Analysis: Theory, Methods & Applications, 24(5):615 – 630, 1995.
- Kamin and Vázquez [1988] Shoshana Kamin and Juan Luis Vázquez. Fundamental solutions and asymptotic behaviour for the -laplacian equation. Revista matemática iberoamericana, 4(2):339–354, 1988.
- Kell [2016] Martin Kell. q-Heat flow and the gradient flow of the renyi entropy in the p-wasserstein space. Journal of Functional Analysis, 271(8):2045–2089, 2016.
- Klenke [2014] Achim Klenke. Probability Theory: a Comprehensive Course. Universitext. Springer-Verlag London, 2nd ed. edition, 2014.
- Santambrogio [2015] Filippo Santambrogio. Optimal Transport for Applied Mathematicians: Calculus of Variations, PDEs, and Modeling. Progress in nonlinear differential equations and their applications 67. Birkhäuser, 1st ed. edition, 2015.
- Vázquez [2006] Juan Luis Vázquez. Smoothing and Decay Estimates for Nonlinear Diffusion Equations: Equations of Porous Medium Type. Oxford Lecture Series in Mathematics and Its Applications. Oxford University Press, USA, 2006.
- Villani [2003] Cedric Villani. Topics in Optimal Transportation. Graduate Studies in Mathematics 58. American Mathematical Society, 2003.