arXiv:2106.12801v1 [math.AP] 24 Jun 2021 where aaee.Hr euetenotation the use we Here parameter. .Introduction 1. MSC: 2020 Poincar´e inequalities, lemma, Lions’ hypocoercivity. equilibria, local average, t time other by hypocoe obtained derive results Keywords: also similar we to compared consequence, vario are a covers which As which estimates, solutions, equilibria. superexpone the and local t of Maxwellian) exponential estimating norms subexponential, for c of to method averages in corresponding constructive time Mourrat and ine of explicit J.-C. Lions-type rate an Poincar´e and and develop adapted Armstrong we Using weakties, S. on equilibria. by based local Results achieved Maxwellian equilibria. been local recently tail fat have or Maxwellian equ with Vlasov-Fokker-Planck) torus (or Fokker-Planck kinetic consider We Abstract ditions Web: ecnie h pta domain spatial the consider We b iatmnod aeaia“.Csrt” Universit`a deg Casorati”, “F. Matematica di Dipartimento a EEAE NS M 54 nvri´ ai-apie PSL Universit´e Paris-Dauphine, 7534, UMR CNRS, CEREMADE, 1 e scnie the consider us Let mi:[email protected] Email: ieaeae o iei okrPac equations Fokker-Planck kinetic for averages Time https://www.ceremade.dauphine.fr/ n en Ω define and , f safnto ftime of function a is ∂ t iei okrPac qain rsenUlnekequation, Ornstein-Uhlenbeck equation, Fokker-Planck Kinetic f rmr:8C0 eodr:3B0 51,4D6 35K65. 47D06, 35H10, 35B40, Secondary: 82C40; Primary: + v aehld ated asgy 51 ai,France Paris, 75016 Tassigny, de Lattre de Marechal ∇ ⋅ x t f ∶ iei okrPac equation Fokker-Planck kinetic ( = = ∇ ,t t, ⟨ v ivniM Brigati M. Giovanni v = ⟩ ⋅ + ‰ t ∇ 70 ai,Italia Pavia, 27100 τ » ≥ v ) Q f 1 0 × , + + ∶ Q, ( = position ~ α S v brigati/ S ⟨ 0 2 o some for v L , , ⟩ α ) − ∀ d 2 f v x v ∋ velocity , ∈ a,b, x, Ž R τ f , 1 with d > iSuid ai,VaFraa5, Ferrata Via Pavia, di Studi li . 0, t eidcbudr con- boundary periodic ( v, ≥ 0 , n Ω and 0 and ⋅ , nvriy lc du Place University, ⋅ = ) ta (including ntial α f toso the on ations 0 sapositive a is ue2,2021 25, June . sregimes us echniques. = edecay he Ω 0 rcivity norms s of ase quali- . The (1) normalized local equilibrium, that is, the equilibrium of the spatially homoge- neous case, is 1 −⟨v⟩α Rd γα v e , ∀ v , ( ) = Zα ∈ where Zα is a non-negative normalization factor, so that dγα ∶ γα v dv is a measure. We shall distinguish a sublinear regime= if α( ) 0, 1 , a linear regime if α 1 and a superlinear regime if α 1. The superlinear∈ ( ) regime covers the Maxwellian= case α 2. The threshold case≥ α 1 corresponds α to a linear growth of v as v → +∞= . The estimates in the= linear case are similar to the ones of the⟨ ⟩ superlinearS S regime. In the literature, γα is said to be subexponential, exponential or superexponential depending whether the regime is sublinear, linear or superlinear. The mass M f , x, v dx dv ∶ Rd ⋅ = UQ× ( ) is conserved under the evolution according to the kinetic Fokker-Planck equa- tion (1). We are interested in the convergence of the solution to the stationary −d solution ML γα. By linearity, we can assume from now on that M 0 with no loss of generality. The function = f h = γα solves the kinetic-Ornstein-Uhlenbeck equation

∂th + v ⋅ ∇xh ∆αh, h 0, ⋅, ⋅ h0, (2) = ( ) = with α−2 ∆αh ∶ ∆vh − α v v ⋅ ∇vh. = ⟨ ⟩ and zero-average initial datum in the sense that

h0 x, v dx dγα 0. Rd UQ× ( ) = By mass conservation, solutions to (2) are zero-average for any t 0. We also consider the time average defined as > t+τ 1 t+τ g s ds ∶ S g s ds at ( ) = τ t ( ) without specifying the τ dependence when not necessary. Our first result is devoted to the decay rate of h t, ⋅, ⋅ → 0 as t → ∞ using time averages. ( ) Theorem 1. Let α 1. Then, for all L 0 and τ 0, there exists a constant 2 λ 0 such that, for≥ all h0 L dx dγα with> zero-average,> the solution to (2) satisfies> ∈ ( )

t+τ 2 2 −λ t h s, , 2 ds h0 2 e , t 0. (3) a ⋅ ⋅ L (dxdγα) L (dxdγα) ∀ t Y ( )Y ≤ Y Y ≥

2 The expression of λ as a function of τ and L is given in Section 4. To deal with large time asymptotics in kinetic equations, it is by now standard to use hypocoercivity methods. Although not being exactly a hypocoercive method in the usual sense, Theorem 1 provides us with an hypocoercivity-type estimate. Corollary 2. Under the assumptions of Theorem 1, there exists an explicit constant C 1 such that all solutions h to (2) fulfill > 2 2 −λ t h t, , 2 C h 2 , t . ⋅ ⋅ L (dxdγα) 0 L (dxdγα) e ∀ 0 (4) Y ( )Y ≤ Y Y ≥ A typical feature of hypocoercive estimates is the factor C 1 in (4) and we emphasize that there is no such constant in (3). Now, let us turn> our attention to the subexponential case 0 α 1. < < Theorem 3. Let α 0, 1 Then, for all L 0 and τ 0, for all σ 0, there is a constant K 0 such∈ ( that) all solutions to (2)> decay according> to > > t+τ σ 2 − 2 1 α σ 2 h s, , 2 ds K 1 t ( − ) v h0 dx dγα, t 0. a ⋅ ⋅ L (dxdγα) + U d ∀ t Y ( )Y ≤ ( ) Q×R ⟨ ⟩ ≥ (5) Further details will be given in Section 5. The constants λ, C in Corol- lary 2 and K in Theorem 3 depend on L 0 and τ 0 and their values is discussed later. The rate of Theorem 3 is the> same as in> the homogeneous case of [18, Proposition 11]. See Section 5 for a discussion of the limit α → 1−. Equation (2) is used in to describe the distribution function of a system of particles interacting randomly with some background, see for in- stance [10]. The kinetic Fokker-Planck equation is the PDE version of Langevin dynamics ⎧dxt vt dt, ⎪ = ⎨dvt −vt + √2 Wt, ⎪ ⎩⎪ = where Wt is a standard Brownian motion. See [14, Introduction] for further details on connections with probability theory. The kinetic Fokker-Planck equa- tion (1) is a simple kinetic equation which has a long history in mathematics that we will not retrace in details here. Mathematical results go back at least to [31] and are at the basis of the theory of L. H¨ormander (see, e.g., [30]), at least in the case α 2. For the derivation of the kinetic-Fokker-Planck equation from underlying stochastic= ODEs, particularly in the context of astrophysics, we can refer to [23, eq. (328)]. Modern hypoellipticity theory emerged from [29, 25] and built up in a fully developed theory in [38] with important contributions in [28, 34]. We study (1) in this paper, or actually the equivalent equation (2), because it is a useful benchmark for the comparison with other methods. The word hypocoercivty was coined by T. Gallay, in analogy with the already quoted hypoelliptic theory of H¨ormander in [30]. In [38], C. Villani distinguishes the regularity point of view for elliptic and parabolic problems driven by degen- erate elliptic operators from the issue of the long-time behaviour of solutions,

3 which is nowadays attached to the word hypocoercivity. The underlying idea is to twist the direction in which the operator is coercive and use the transport operator to carry its coercivity properties to the directions in the phase space for which the operator is degenerate. Twisting the H1-norm creates equivalent norms, which are coercive along the evolution. So works the H1 framework, see [38, 37, 26]. The H1-framework has been connected to the carr´edu champ method of D. Bakry and M. Emery in [9] by F. Baudoin, who proved decay also w.r.t. the Wasserstein distance, as shown in [11, 12, 13]. The H1 hypocoercivity implies a decay rate for the L2 norm, but the cor- responding estimates turn out to be suboptimal because extra regularity is required on the initial datum. This motivates the development of direct L2 techniques based on a perturbation of the L2 norm. Such an approach can be found in [27] and [18], which is consistent with diffusion limits. In [17], the authors extend the technique to the subexponential case. Another possibility is to perform rotations in the phase space and use a Lyapunov inequality for matrices as in [8]. This approach gives optimal rates, but it is less general as it requires further algebraic properties for the diffusion operator and a detailed knowledge of its spectrum. The core of [8] is a spectral decomposition, that was originally understood via a toy model exposed in [27]. In a domain with periodic boundary conditions, the problem is reduced to an infinite set of ODEs corresponding to spatial modes. See [2, 3, 7] for details and extensions. A H−1 hypocoercivity theory was recently proposed by S. Armstrong and J.-C. Mourrat in [6] using space-time adapted Poincar´einequalities to derive qualitative hypocoercive estimates in the case α 2 and bounded spatial do- mains. An extension to the whole space in presence= of a confining potential can be found in [21]. The method of [6] is our starting point. This paper focuses on decay rates of time averages. Hypocoercivity estimates are obtained as a consequence of these decay rates. We perform an analysis for all positive values of α, which is consistent in the threshold case α 1. The d advantage of the simple setting of a kinetic Fokker-Planck equation on =Q×R is that constants can be computed explicitly in terms of α, τ and L (see Lemma 8), with optimal values in some cases. We make the strategy of [6] quantitative and constructive, for sake of comparison with other hypocoercivity estimates. This paper is organized as follows. In Section 2 we collect some preliminary results: Poincar´eand weighted Poincar´einequalities (Propositions 5 and 6), adapted Lions’ inequality (Lemmas 7 and 8). In Section 3 we introduce an av- eraging lemma (Lemma 12), which is then used to prove the generalized Poincar´e inequality of Proposition 13, at the core of the method. In Section 4 we use Proposition 13 and a Gr¨onwall estimate to prove Theorem 1 and compute an explicit formula for λ (Proposition 15). Section 5 is devoted to the proof of The- orem 3, with additional details, and to the limit α → 1−. Finally, in Section 6, we derive the hypocoercive estimates of Corollary 2. On the benchmark case α 2 in one spatial dimension, we also compare our results with those obtained by= more standard methods.

4 2. Preliminaries

Let us start with some preliminary results.

2.1. Weighted spaces For functions g of the variable v only, that is, of the so-called homogeneous case, we define the weighted Lebesgue and Sobolev spaces

2 2 Rd 1 2 2 d Lα ∶ L , dγα and Hα ∶ g Lα ∶ ∇vg Lα . = ( ) = š ∈ ∈ ‰ Ž Ÿ 2 We equip Lα with the scalar product

g1,g2 g1 v g2 v dγα (6) S d ( ) = R ( ) ( ) 1 and consider on Hα the norm defined by 2 2 2 h 1 h dγα vh dγα Hα ∶ S d + S d ∇ Y Y =‹ R  R S S −1 1 as in [6]. The duality product between Hα z and Hα g is given by ∋ ∋

z,g vwz vg dγα, ∶ S d ∇ ⋅ ∇ ⟨ ⟩ = R 1 where wz is the weak solution in Hα to

∆αwz z z dγα, w dγα 0. − − S d S d = R R = Here we write ∫Rd z dγα for functions which are integrable w.r.t. dγα and, up to a little abuse of notations, this quantity has to be understood in the distri- bution sense for more general measures. As a consequence and with the above notations, we define

2 2 2 z 1 z dγα wz 2 . Hα− ∶ S d + Lα Y Y =‹ R  Y Y

With these notation, the key property of the operator ∆α, is

g1, ∆αg2 − S ∇vg1 ⋅ ∇vg2 dγα ⟨ ⟩ = 1 for any functions g1, g2 Hα. ∈ + d Consider next functions h of t, x, v R × Q × R and define the space ( )∈ 2 R+ 1 2 −1 Hkin ∶ h L × Q; Hα ∶ ∂th + v ⋅ ∇xh L Ωt; Hα ∀ t 0 . = ™ ∈ ‰ Ž ∈ ‰ Ž ≥ ž R+ Rd We recall that Ωt t,t + τ × Q ⊂ t × x and that x-periodic boundary = ( ) 2 1 conditions are assumed. We can equip Hkin ∩ L Ωt × Q; Hα with the norm ‰ Ž 2 2 2 h h 2 2 h kin ∶ L (Ωt;L ) + kin Y Y = Y Y α S S

5 where the kinetic semi-norm is given by

2 2 2 h h 2 2 ∂ h v h 2 1 . kin ∶ ∇v L (Ωt;L ) + t + ⋅ ∇x L (Ωt;H ) S S = Y Y α Y Y α− We refer to [6, Section 7] for the proof of following result. 1 2 1 2 2 Proposition 4. The embedding Hkin ∩ L Ωt × Q; Hα ↪ ∩L Ωt × Q;Lα is continuous and compact for any t 0. ‰ Ž ‰ Ž ≥ 2.2. Poincar´einequalities In this subsection, we consider functions g depending only on the variable v. Let α 1. We can state some Poincar´einequalities. ≥ Proposition 5. If α 1, there exists a constant Pα 0 such that, for all 1 ≥ > functions g Hα, we have ∈ 2 2 g ρg dγα Pα v g dγα with ρg g dγα. (7) S d − S d ∇ ∶ S d R S S ≤ R S S = R 2 With α 1, the operator ∆α admits a compact resolvent on L dγα . Then, (7) holds≥ by the standard results of [19, Chapter 6]. The best constant( ) −1 Pα is such that Pα is the minimal positive eigenvalue of −∆α. See [22] and the references quoted therein for estimates on Pα. In the case of the Gaussian Poincar´einequality, it is shown in [35] that P2 1, although the result was probably known before. =

2.3. Weighted Poincar´einequalities Here we consider again functions depending only on v. For α 0, 1 , in- equality (7) has to be replaced by the following weighted Poincar´einequality.∈ ( )

Proposition 6. If α 0, 1 , there exists a constant Pα 0 such that, for all 1 ∈ ( ) > functions g Hα, we have ∈ 2 (1−α) 2 2 v g ρg dγα Pα v g dγα with ρg g dγα. (8) S d − S d ∇ ∶ S d R ⟨ ⟩ S S ≤ R S S = R For more details, we refer for instance to [17, Appendix A]. Notice that the 2 (1−α) average in the l.h.s. is taken w.r.t. dγα, not w.r.t. v dγα ⟨ ⟩ 2.4. Lions’ Lemma Rd+1 R Rd Let be an open, bounded and Lipschitz-regular subset of t × x. We recallO that ≈

−1 ∗ w w,u C u 1 , C , ∶ H0(O) 0 H (O) = š ∈D (O) S⟨ ⟩O S ≤ Y Y > Ÿ where ∗ denotes the space of distributions over , equipped with the D (O) O −1 1 weak∗ topology, and w,u is the duality product between and 0. The 1 ⟨ ⟩O 1 H H norm on 0 is as usual u ↦ ∇u L2(O). On , we introduce the norm H (O) Y Y H (O) 2 2 2 u H1 S u dt dx + ∇u L2(O). Y Y =V O V Y Y

6 The norm induced on H−1 is then (O) 2 2 2 w 1 w, 1 z 2 , H− (O) + w L (O) Y Y = ⟨ ⟩O Y Y where zw the solution to

− ∂tt + ∆x zw w − w, 1 , S zw dt dx 0. ( ) = ⟨ ⟩O O = Lions’ Lemma gives a sufficient condition for a distribution to be an L2 function. The following statement is taken from [5]. Lemma 7. Let be a bounded, open and Lipschitz-regular subset in Rd+1. Then, for all u O∗ , we have that u L2 if and only if the weak gradient ∈D−1 (O) ∈ (O) ∇u belongs to H . Moreover, there exists a constant CL such that (O) (O) 2 2 u u dx dt CL u H 1 , − S 2 ∇ − (O) \ O \L (O) ≤ Y Y for any u L2 . ∈ (O) According to [20, 16, 24], if is star-shaped w.r.t. a ball, then the con- stant CL has the following structure:O

d D CL 4 S (O), (9) = S S d (O) where D is the diameter of , while d is the diameter of the largest ball one can include in . See in particularO [24,(O) Remark 9.3] and [16, Lemma 1]. As a consequence, weO have the following explicit expression of CL when Ω. O = d Lemma 8. Let L 0, τ 0,L and Ω 0, τ × 0,L . Lemma 7 holds with > ∈ ( ) = ( ) ( ) 2 2 d √dL + τ CL 4 S . (10) = S S τ 2.5. The kinetic Ornstein-Uhlenbeck equation We consider solutions to (2) in the weak sense, i.e., functions h in the space R+ 2 2 C ;L dx dγα with initial datum h0 h 0, ⋅, ⋅ in L dx dγα such that (2) ( ( )) = ( Rd) Rd ( ) holds in the sense of distributions on 0, ∞ × x × v. The following result is taken from [6] if α 2. The extension to( α )2 is straightforward as follows from a careful reading of= the proof in [6, Proposition≠ 7.12].

d Proposition 9. Let L 0 and α 0. With Ω 0, τ × 0,L , for all zero- > 2 > R+ 2 = ( ) ( ) average initial datum h0 L dx dγα ∩C ;L dx dγα , there exists a unique solution h to (2) such that∈ h( Hkin for) all( τ 0.( )) ∈ > Regularity properties for (2) are collected in [6, Section 7]. In the special case α 2, some fractional regularity along all directions of the phase space are known.= Also see [36] for further result on regularity theory for kinetic Fokker- Planck equations.

7 2.6. A priori estimates We state two estimates for solutions to (2).

d Lemma 10. Let L 0, τ 0, Ω 0, τ × 0,L , and α 0. If h is a solution to (2), then we have> > = ( ) ( ) >

∂ v h 2 1 h 2 2 . t + ⋅ ∇x L (Ω;H− ) ∇v L (Ω;L ) (11) Y( ) Y α ≤ Y Y α 2 1 Proof. Take a test function φ L Hα , and write ∈ ( )

S ∂t + v ⋅ ∇x h, φ dx S ∆αh, φ dx − S ∇vh, ∇vφ dx, Q⟨( ) ⟩ = Q⟨ ⟩ = Q( ) from which (11) easily follows, after maximizing over ∇vφ L2 1. Y Y α ≤ For completeness, let us recall the classical L2 decay estimate for solutions to (2).

d Lemma 11. Let L 0, τ 0, Ω 0, τ × 0,L , and α 0. If h is a solution to (2), then we have> > = ( ) ( ) >

d 2 2 h 2 h 2 . L (dxdγα) − 2 ∇v L (dxdγα) dtY Y = Y Y

3. An averaging lemma and a generalized Poincar´einequality

For all functions h Hkin, we define the spatial density ∈

ρh h , , v dγα. ∶ S d ⋅ ⋅ = R ( ) ρ dx h Notice that ∫Q h 0 whenever is a zero-average function. = 3.1. Averaging lemma Inspired by [6, Lemma 3], the following averaging lemma provides a norm of the spatial density, as for instance in [36].

d Lemma 12. Let L τ 0, Ω 0, τ × 0,L , and α 0. For all h Hkin, we have > > = ( ) ( ) > ∈

2 2 2 ρ 1 d h ρ 2 ∂ h v h 2 1 ∇t,x h H (Ω) α − h L (dtdxdγα) + t + ⋅ ∇x L (Ω;H ) (12) Y Y − ≤ ŠY Y Y Y α−  with 2 2 2 2 2 L 2 2 d L 2 dα 2 v1 v L2 + 1 + 4 π2 v L2 + 4 π2 v L2 . (13) = ŠY S S Y α Š  YS S Y α Y Y α  4 Inequality (12) can be extended to any measure dγ such that ∫Rd v dγ ∞ S S < and ∫Rd v dγ 0. The proof of Lemma 12 is technical, but follows in a standard way from the= time-independent case, as it is common in averaging lemmas: see [36]. For sake of simplicity, we detail only the t-independent case below.

8 d Proof of Lemma 12. Assume that h Hkin does not depend on t. Let φ Q be a smooth test-vector field with zero∈ average on each component. We∈ D( write)

− S ρh∇x ⋅ φ dx S ∇xρh ⋅ φ dx, Q = Q with a slight abuse of notation, since the integral of the r.h.s. is in fact a duality −1 2 product. Using Rd vi vj dγα d v 2 δij , we obtain ∫ Lα = Y Y −2 xρh φ dx d v 2 v xρh φ v dx dγα. S ∇ ⋅ Lα U d ⋅ ∇ ⋅ Q = Y Y Q×R

By adding and subtracting ρh, and then integrating by parts, still at formal level, we obtain

v ⋅ ∇xρh φ ⋅ v dx dγα UQ×Rd

v h ρh xφ v dx dγ v xh v φ dx dγα − U d ⋅ − ∇ ⋅ − U d ⋅ ∇ ⋅ = Q×R ( ) Q×R 2 h ρ 2 φ 2 v − h L (dxdγα) ∇x L (dx) L2 ≤ Y Y Y Y Z Z α v h 2 1 φ 2 v 2 + ⋅ ∇x L (H− ) L (dx) Lα Y Y α Y Y Y Y using Cauchy-Schwarz inequalities and duality estimates. By the Poincar´ein- equality, we know that

2 4 π 2 2 φ L2(dx) ∇φ L2(dx) dL2 Y Y ≤ Y Y

Maximizing the r.h.s. on φ such that ∇xφ L2(dx) 1 completes the proof of the t-independent case. When h additionallyY dependsY ≤ on t, the same scheme can be applied with v ⋅ ∇x replaced by ∂t + v ⋅ ∇x.

3.2. A generalized Poincar´einequality The next a priori estimate is at the core of the method. It is a Poincar´e type inequality in t, x and v which relies on Lemma 12 and involves derivatives of various orders.

d Proposition 13. Let L 0, τ 0, Ω 0, τ × 0,L , and α 0. Then, for all h Hkin with zero average,> we have> that= ( ) ( ) > ∈ 2 2 2 h 2 C h ρ 2 ∂ h v h 2 1 L (dtdxdγα) − h L (dtdxdγα) + t + ⋅ ∇x L (Ω;H ) (14) Y Y ≤ ŠY Y Y Y α−  with C 1 + CL dα, where CL and dα are given respectively by (10) and (13). = 2 2 Proof. By orthogonality in L Ω;Lα and because dγα is a probability measure, we have the decomposition ( )

2 2 2 h L2(Ω;L2 ) h − ρh L2(Ω;L2 ) + ρh L2(Ω). Y Y α = Y Y α Y Y

9 The function ρh has zero average on Ω by construction, so that 2 2 2 h L2(Ω;L2 ) h − ρh L2(Ω;L2 ) + CL ∇x,tρ H 1(Ω) Y Y α ≤ Y Y α Y Y − by Lemma 8. Hence 2 2 2 h 2 2 C d h ρ 2 C d ∂ h v h 2 1 L (Ω;L ) 1 + L α − h L (Ω;dγα) + L α t + ⋅ ∇x L (Ω;H ) Y Y α ≤ ( ) Y Y Y Y α− by Lemma 12. This concludes the proof.

4. Linear and superlinear local equilibria: exponential decay rate

In this Section, we consider the case α 1 and the domain Ω t,t + τ × d ≥ = ( ) 0,L , for an arbitrary t 0. Let us define κα ∶ 1 + CL dα Pα + 1 where P( α is) the Poincar´econstant≥ in (7) and where CL and= ( dα are given)( respectively) by (10) and (13). d Lemma 14. Let L 0, τ 0, t 0, Ωt t,t + τ × 0,L , and α 1. Then, for all h Hkin with> zero average,> ≥ we have= (that ) ( ) ≥ ∈ 2 2 h 2 κ h 2 . L (dtdxdγα) α ∇v L (dtdxdγα) (15) Y Y ≤ Y Y Proof. We know that

2 2 2 h 2 C d P h 2 ∂ h v h 2 1 L (dtdxdγα) 1 + L α α ∇v L (dtdxdγα)) + t + ⋅ ∇x L (H ) Y Y ≤ ( ) Š Y Y Y Y α−  as a consequence of (7) and (14). Then (15) follows from Lemma 10. We are ready to prove Theorem 1 with an explicit estimate of the constant λ. Proposition 15. For any α 1, Theorem 1 holds true with ≥ 1 1 2 2 Sd−1 τ + √dL + τ 2 dα Pα + 1 . λ = τ Š  Š S S ( ) Proof. A t integral of (15) on the interval t,t + τ gives ( ) t+τ t+τ 2 2 h s, , 2 ds κα vh s, , 2 ds. S ⋅ ⋅ L (dxdγα) S ∇ ⋅ ⋅ L (dxdγα) t Y ( )Y ≤ t Y ( )Y With λ 2 κα, we deduce from Lemma 11 that = ~ t+τ t+τ d 2 2 h s, , 2 ds 2 vh s, , 2 ds S ⋅ ⋅ L (dxdγα) − S ∇ ⋅ ⋅ L (dxdγα) dt t Y ( )Y = t Y ( )Y t+τ 2 λ h s, , 2 ds. − S ⋅ ⋅ L (dxdγα) ≤ t Y ( )Y 2 Gr¨onwall’s Lemma and the monotonicity of t ↦ h t, ⋅, ⋅ L2(dxdγ ) imply Y ( )Y α t+τ τ 2 2 −λ t h s, , 2 ds h s, , 2 ds e S ⋅ ⋅ L (dxdγα) S ⋅ ⋅ L (dxdγα) t Y ( )Y ≤ 0 Y ( )Y 2 −λ t τ h0 L2(dxdγ ) e ≤ Y Y α for ant t 0, which proves (3), that is, Theorem 1, and gives the claimed expression≥ for λ using the definition of κα.

10 Notice that the r.h.s. vanishes as τ ↓ 0, which is expected because of the 2 degeneracy of ∆α: an exponential decay rate of h t, ⋅, ⋅ L2(dxdγ ) cannot hold. Y ( )Y α 5. Sublinear equilibria: algebraic decay rates 5.1. Proof of Theorem 3 Assume that α 0, 1 . Let us the parameter β 2 1 − α p where p, q 1 ∈ ( ) 1 1 = ( )~ > are H¨older conjugate exponents, i.e., p + q 1 and define = β q 2 Zh t v h ρh dx dγα. (16) ∶ U d − ( ) = Q×R ⟨ ⟩ S S The following estimates replaces Proposition 13. d Proposition 16. Let L 0, τ 0, t 0, Ωt t,t+τ × 0,L , and α 1. With the above notations, for all> h >Hkin with≥ zero= ( average,) we( have) that ≥ ∈ 2 h 2 L (dtdxdγα) Y Y 1 2 1 p p t+τ q 2 C Pα vh 2 Zh s ds C ∂th v xh 2 1 ∇ L (dtdxdγα) ∫t + + ⋅ ∇ L (Ω;Hα− ) ≤ Y Y Š ( )  Y Y where C 1 + CL dα is as in Proposition 13 and Pα denotes the constant in the weighted= Poincar´einequality (8). Proof. Using (14) and H¨older’s inequality w.r.t. the variable v, we find that

1 1 p q 2 −β p 2 β q 2 h ρh 2 v h ρh dγα v h ρh dγα . − Lα S d − S d − Y Y ≤‹ R ⟨ ⟩ S S  ‹ R ⟨ ⟩ S S  The weighted Poincar´einequality (8) with βp 2 1 − α and an additional H¨older inequality w.r.t. the variables t and x allow= us( to complete) the proof.

d Lemma 17. Let L 0, τ 0, t 0, Ωt t,t+τ × 0,L , and α 0, 1 . There is a constant W 0> such> that,≥ for all= solution ( )h ( Hkin) to (2) with∈ ( an) initial datum h0 with zero> average, using the notation (16)∈ as in Proposition 16, we have β q 2 Zh t W v h0 dx dγα, t 0. U d ∀ ( ) ≤ Q×R ⟨ ⟩ ≥ Proof. An elementary computation shows that

β q 2 β q 2 2 v h ρh dx dγα 2 v h ρh dx dγα U d − U d + Q×R ⟨ ⟩ S S ≤ Q×R ⟨ ⟩ ‰ Ž β q β q 2 2 1 Rd v dγα v h dx dγα + ∫ U d ≤ ‰ ⟨ ⟩ Ž Q×R ⟨ ⟩ 2 2 2 β q 2 ρ d h dγ d h dγ d v h dγ . because h ∫R α ∫R α ∫R α According to [17, Proposition 4],= ‰ there is aŽ constant≤ β q 1≤ such⟨ that⟩ K > β q 2 β q 2 v h t, x, v dx dγα β q v h0 dx dγα, t 0. U d U d ∀ Q×R ⟨ ⟩ S ( )S ≤ K Q×R ⟨ ⟩ ≥ β q The result follows with W 2 1 + ∫Rd v dγα β q. = ‰ ⟨ ⟩ Ž K

11 Assume that h Hkin solves (2) with an initial datum h0 with zero average and let us collect∈ our estimates. With Proposition 10, Proposition 16, and Lemma 17, the estimate of Lemma 14 is replaced by

2 2 p 2 h 2 A h 2 C h 2 (17) L (dtdxdγα) ∇v L (dtdxdγ ) + ∇v L (dtdxdγα) Y Y ≤ Y Y α Y Y 1 1~q p 1~q β q 2 A C P τ W d v h dx dγ . with α ∬Q×R 0 α = ( ) Š ⟨ ⟩  The main result of the section is a technical version of Theorem 3. Let t+τ t+τ 2 2 x t h s, , 2 ds and y t vh s, , 2 ds ∶ a ⋅ ⋅ L (dxdγα) ∶ a ∇ ⋅ ⋅ L (dxdγα) ( ) = t Y ( )Y ( ) = t Y ( )Y d where norms are taken on Q × R . We know from Lemma 11 and (17) that

′ 1~p x − 2 y and x ϕ y ∶ A y + C y. = ≤ ( ) = Finally, let us denote by ϕ−1 the inverse of y ↦ ϕ y and consider ( ) x0 dz 2 ψ z with x0 h0 2 . ∶ S −1 L (dxdγα) ( ) = z 2 ϕ z = Y Y ( ) d Theorem 18. Let L 0, τ 0,t 0, Ωt t,t+τ × 0,L , and α 0, 1 . With the above notations, for> all solution> ≥ h H=kin ( to (2)) with( an) initial datum∈ ( )h0 with zero average, we have ∈

t+τ 2 −1 x t h s, , 2 ds ψ t , t 0. a ⋅ ⋅ L (dxdγα) ∀ ( ) = t Y ( )Y ≤ ( ) ≥ Proof. The strategy goes as in [33, 17]. Everything reduces to the differential inequality ′ −1 x − 2 ϕ x ≤ ( ) using the monotonicity of y ↦ ϕ y . From by the elementary Bihari-Lasalle inequality, see [15, 32], which is obtained( ) by a simple integration, we obtain

t+τ 2 −1 x t h s, , 2 ds ψ t ψ x 0 . a ⋅ ⋅ L (dxdγα) + ( ) = t Y ( )Y ≤ ‰ ( ( ))Ž Since, on the one hand

τ 2 2 x 0 h s, , 2 ds h0 2 x0 a ⋅ ⋅ L (dxdγα) L (dxdγα) ( ) = 0 Y ( )Y ≤ Y Y = 2 because s ↦ h s, , 2 is nonincreasing according to Lemma 11, and ψ ⋅ ⋅ L (dxdγα) is nonincreasingY ( on the)Y other hand, then

−1 −1 ψ t + ψ x 0 ψ t , ‰ ( ( ))Ž ≤ ( ) which concludes the proof. Notice that the dependence on h0 enters in A and x0, and henceforth in ϕ and ψ.

12 1~p Proof of Theorem 3. Since limt→+∞ y t 0, we have that ϕ y t A y t as t → +∞, which heuristically explains( ) the = role played by p in( (5).( )) This ∼ can( ) be made rigorous as follows. Notice that

1~p 1~p 1−1~p ϕ y A y + C y A0 y , ∀ y y0, with A0 A + C y0 . ( ) = ≤ ≤ = With A replaced by A0 and C replaced by 0, the computation of the proof −1 of Theorem 18 is now explicit. With the choice y0 ϕ x0 , we know that y t y0 for any t 0 and obtain = ( ) ( ) ≤ ≥ 1 1−p −p − p 1 x t x0 + 2 p − 1 A0 t − , ∀ t 0. (18) ( ) ≤ ‰ ( ) Ž ≥ 1~p −1~p 1~p 1−1~p Using x0 A y0 + C y0 C y0, we know that A0 x0 y0 A + C x0 , which proves= (5) with ≥ = ≤

1~(1−p) 1~p 1−1~p 1~p K max 1, 2 p − 1 C Pα τ W + C . = › ‰ ( )Ž Š ( )  The conclusion holds using σ β q 2 1 − α p − 1 . = = ( )~( ) 5.2. The linear threshold: from algebraic to exponential rates A very natural question arises: is the result Theorem 15 (corresponding to α 0, 1 consistent with the result of Theorem 18 (which covers any α 1) ? A first∈ ( observation) is that we can vary α in the assumptions concerning the≥ initial data.

2 2 σ~2 2 2 Lemma 19. If h0 L Q;Lα0 for some α0 0, 1 , then v h0 L Q;Lα for any α α0 and∈ any(σ 0. ) ∈ ( ) ⟨ ⟩ ∈ ( ) > > The proof is a simple consequence of the fact that v ↦ v σ exp v α−α0 is uniformly bounded. For any α α0, let us denote the corresponding⟨ ⟩ ⟨ solution⟩ (α) of (2) with initial datum h0, of zero> average, by h . If α α0, 1 , then (18) can be rewritten as ∈ ( ) t+τ 1 (α) 2 2 − p 1 h s, , 2 ds h0 2 1 p 1 ℓ α t − . a ⋅ ⋅ L (dxdγα) L (dxdγα) + − t Y ( )Y ≤ Y Y ‰ ( ) ( ) Ž → − By passing to the limit as α 1 , we recover (3) with λ limα→1− ℓ α , where 2(p−1) −p = ( ) ℓ α 2 h0 2 A and A0 A0 α as above. The Poincar´econstant Pα L (dxdγα) 0 in( the) = weightedY Y Poincar´einequality= (8)( ) admits a limit as α → 1−, according to [18, Appendix A].

The limit of limα→1− ℓ α is certainly not optimal. By working directly on the Bihari-Lasalle estimate( ) of Theorem 18, we can recover the value of λ in Theorem 15. Notice here that σ 0 plays essentially no role and can be taken arbitrarily small, even depending> on α, but such that p 2 1 − α σ → 1 as α → 1−. = ( )~

13 6. Hypocoercivity and comparison with some other methods

6.1. A hypocoercivity result 2 Theorem 1 implies an L −hypocoercivity result in the linear and superlinear regimes α 1. ≥ Proof of Corollary 2. For any t 0, we know from Theorem 1 that ≥ t+τ 2 2 2 −λ t h t τ, , 2 h s, , 2 ds h0 2 e , + ⋅ ⋅ L (dxdγα) a ⋅ ⋅ L (dxdγα) L (dxdγα) Y ( )Y ≤ t Y ( )Y ≤ Y Y as a consequence of the monotonicity of the L2 norm, according to Lemma 11. This proves that

2 2 −λ t h t, , 2 C h 2 e ⋅ ⋅ L (dxdγα) 0 L (dxdγα) Y ( )Y ≤ Y Y with C eλτ for any t τ. However, if t 0, τ , it turns out that Ce−λ t 1 so that the= inequality is also≥ true by Lemma∈ [ 11. This) concludes the proof. ≥ The remainder of this section is devoted to a comparison with earlier hypoco- ercivity results in a simple benchmark case: let α 2, d 1 and L 2π. In this case, for the choice τ 2π, Theorem 1 amounts to= = = = 2 π 2 − t 2 8 √3 2 8 √3 h L (dxdγ2) e h0 L (dxdγ2) e , ∀ t 0. Y Y ≤ Y Y ≥ For sake of comparison, notice that λ 1 8 √3 0.0721688. = ~( ) ≈ 6.2. The DMS method The first comparison is with the abstract twisted L2 hypocoercivity method 2 of [27, 18]. Let ⋅ be the norm of L dx dγ2 and ⋅, ⋅ the associated scalar product. We considerY Y the evolution equation( ) ( )

∂th + Th h. (19) = L 2 2 Theorem 20. Let h be a solution of (19) with initial datum h0 L Q;L2 and assume that T and are respectively anti-self-ajoint and self-adjoint∈ ( opera) tors 2 2 on L Q;L such that,L for some positive constants λm, λM and CM , we have ( 2) 2 (A1) − h,h λm 1 − Π h for all h D , ( L 2) ≥ Y( 2 ) Y ∈ (L) (A2) T Π h λM Π h for all h D T Π , Y Y Y Y ( ) (A3) Π T Π h ≥0, ∈ = 2 2 (A4) AT 1 Π h A h CM Id Π h for all admissible h L Q;L Y ( − ) Y+Y Y Y( − ) Y ( 2) L −≤1 ∈ where A Id T Π ∗ T Π T Π ∗ and Π is the projection in L2 onto the ∶ ‰ + Ž 2 kernel of =. Then( we) have ( ) L 2 2 −λ t h t, ⋅, ⋅ C h0 e ∀ t 0 Y ( )Y ≤ Y Y ≥ 1 λm λM 2 δ λM with C 1 + δ 1 − δ , δ 2 min 1, λm, (1+λ ) C2 and λ 3 (1+λ ) . = ( )~( ) = ™ M M ž = M

14 This result is taken from [18, Proposition 4]. According to [18, Corollary 9], we have the estimate λ 1 24 0.041667. A minor improvement is obtained as = ~ ≈ 2 follows. Using Theorem 20 applied with d 1, L 2π, T v ∂x and ∂v −v ∂v, in Fourier variables, we obtain λm λM 1= and C=M 1 =√3 2 accordingL = to [7, ( + )~ Section II.1.3.2], so that λ 1 12=+ 6 √=3 0.0446582.= Using Fourier modes, a slightly better estimate is obtained= ~( from) [7,≈ Section II.1.2] with λ 0.176048. ≈ 6.3. Direct spectral methods In a series of papers, F. Achleitner, A. Arnold, E. Carlen and several other collaborators use direct spectral methods. We refer in particular to [8, 2, 3, 4] and also [7] for an introduction to the method, which can be summarized as follows. Let us consider (19) written after a Fourier transform in x, so that T i ξ ⋅v, 2 2 = and acting on L dx;L2 now considered as a space of complex valued functions. Assume that for( some positive) definite bounded Hermitian operator P and some constant λ 0, +∞ , we have ∈ ( ) ∗ L − T P + P L − T 2 λ P. ( ) ( ) ≥ Let us consider the twisted norm fˆ 2 f,ˆ P fˆ dx where , is the natural Y YP ∶ ∫Q( ) (⋅ ⋅) extension of the scalar product as defined= in (6). From

d ˆ 2 ˆ ∗ ˆ ˆ 2 f P − f, L − T P + P L − T f − 2 λ f P , dtY Y = ⟨ (( ) ( )) ⟩ ≤ Y Y for some C 1, we deduce that > −1 2 2 −2 λ t 2 C f t, , 2 fˆ t, , e fˆ t . ⋅ ⋅ L (dxdγα) ⋅ ⋅ P 0 P ∀ 0 Y ( )Y ≤ Y ( )Y ≤ Y Y ≥ To our knowledge, µ has not yet been computed in the case of (1). The spectral decomposition −i 2π ξ⋅x h t, x, v Q Q aξ,k t Hk v e L ( ) = ξ∈Zd k∈Nd ( ) ( ) provides an easy framework for finite dimensional approximations using the basis of Hermite functions Hk k∈N and the numerical value µ 0.4 has been obtained according to [1]. ( ) ≈

Acknowledgements

The author thanks F. Achleitner and C. Mouhot for stimulating discussions. This research project is funded by the European Union’s Horizon 2020 research and innovation program under the Marie Sklodowska-Curie grant agreement No 754362. Partial support has been obtained from the EFI ANR-17-CE40-0030 Project of the French National Research Agency. © 2021 by the author. Any reproduction for non-commercial purpose is authorized.

15 References

[1] F. Achleitner, A numerical computation of the rates for kinetic fokker- planck equations. Personal communication. [2] F. Achleitner, A. Arnold, and E. A. Carlen, On linear hypocoercive BGK models, in From Particle Systems to Partial Differential Equations III, Springer, 2016, pp. 1–37. [3] F. Achleitner, A. Arnold, and E. A. Carlen, On multi-dimensional hypocoercive BGK models, Kinet. Relat. Models, 11 (2018), pp. 953–1009. [4] F. Achleitner, A. Arnold, and B. Signorello, On optimal de- cay estimates for ODEs and PDEs with modal decomposition, in Inter- national workshop on Stochastic Dynamics out of Equilibrium, Springer, 2017, pp. 241–264. [5] C. Amrouche, P. G. Ciarlet, and C. Mardare, On a lemma of Jacques-Louis Lions and its relation to other fundamental results, Journal de Math´ematiques Pures et Appliqu´ees, 104 (2015), pp. 207–226. [6] S. Armstrong and J.-C. Mourrat, Variational methods for the kinetic Fokker-Planck equation, preprint arXiv:1902.04037, (2019). [7] A. Arnold, J. Dolbeault, C. Schmeiser, and T. Wohrer¨ , Sharpening of decay rates in Fourier based hypocoercivity meth- ods, Preprint https://hal.archives-ouvertes.fr/hal-03078698 and https://arxiv.org/abs/2012.09103, to appear in Recent Advances in Kinetic Equations and Applications (F. Salvarani ed.), Springer INdAM Series, (2020). [8] A. Arnold and J. Erb, Sharp entropy decay for hypocoercive and non-symmetric Fokker-Planck equations with linear drift, preprint arXiv:1409.5425, (2014). [9] D. Bakry and M. Emery´ , Diffusions hypercontractives, in S´eminaire de Probabilit´es XIX 1983/84, Springer, 1985, pp. 177–206. [10] R. Balian, From Microphysics to Macrophysics, Methods and Applica- tions of Statistical Physics, Springer Berlin Heidelberg, 1991. [11] F. Baudoin, Wasserstein contraction properties for hypoelliptic diffusions, preprint arXiv:1602.04177, (2016). [12] , Bakry-Emery´ meet Villani, J. Funct. Anal., 273 (2017), pp. 2275– 2291. [13] F. Baudoin, M. Gordina, and D. P. Herzog, Gamma calculus beyond Villani and explicit convergence estimates for Langevin dynamics with sin- gular potentials, preprint arXiv:1907.03092, (2019).

16 [14] E. Bernard, M. Fathi, A. Levitt, and G. Stoltz, Hypocoercivity with Schur complements, preprint arXiv:2003.00726, (2020). [15] I. Bihari, A generalization of a lemma of Bellman and its applica- tion to uniqueness problems of differential equations, Acta Mathematica Academiae Scientiarum Hungarica, 7 (1956), pp. 81–94. [16] M. E. Bogovskii, Solution of the first boundary value problem for the equation of continuity of an incompressible medium, in Soviet Math. Dokl., vol. 20, 1979, pp. 1094–1098. [17] E. Bouin, J. Dolbeault, L. Lafleche, and C. Schmeiser, Hypoco- ercivity and sub-exponential local equilibria, Monatsh. Math., 194 (2021), pp. 41–65. [18] E. Bouin, J. Dolbeault, S. Mischler, C. Mouhot, and C. Schmeiser, Hypocoercivity without confinement, Pure Appl. Anal., 2 (2020), pp. 203–232. [19] H. Brezis, Op´erateurs maximaux monotones, mathematics studies 5, No- tas de Matematica, 50 (1973). [20] A. P. Calderon´ and A. Zygmund, On singular integrals, American Journal of Mathematics, 78 (1956), pp. 289–309. [21] Y. Cao, J. Lu, and L. Wang, On explicit L2-convergence rate estimate for underdamped Langevin dynamics, preprint arXiv:1908.04746, (2019). [22] P. Cattiaux, A. Guillin, P. Monmarche,´ and C. Zhang, Entropic multipliers method for Langevin diffusion and weighted log-Sobolev inequal- ities, Journal of , 277 (2019), p. 108288. [23] S. Chandrasekhar, Stochastic problems in physics and astronomy, Rev. Mod. Phys., 15 (1943), pp. 1–89. [24] G. Csato,´ B. Dacorogna, and O. Kneuss, The pullback equation for differential forms, vol. 83, Springer Science & Business Media, 2011. [25] L. Desvillettes and C. Villani, On the trend to global equilibrium in spatially inhomogeneous entropy-dissipating systems: the linear Fokker- Planck equation, Comm. Pure Appl. Math., 54 (2001), pp. 1–42. [26] J. Dolbeault and X. Li, ϕ-entropies: convexity, coercivity and hypocoer- civity for Fokker–Planck and kinetic Fokker–Planck equations, Mathemat- ical Models and Methods in Applied Sciences, 28 (2018), pp. 2637–2666. [27] J. Dolbeault, C. Mouhot, and C. Schmeiser, Hypocoercivity for lin- ear kinetic equations conserving mass, Transactions of the American Math- ematical Society, 367 (2015), pp. 3807–3828.

17 [28] F. Herau´ , Hypocoercivity and exponential time decay for the linear in- homogeneous relaxation Boltzmann equation, Asymptot. Anal., 46 (2006), pp. 349–359. [29] F. Herau´ and F. Nier, Isotropic hypoellipticity and trend to equilibrium for the Fokker-Planck equation with a high-degree potential, Arch. Ration. Mech. Anal., 171 (2004), pp. 151–218. [30] L. Hormander¨ et al., Hypoelliptic second order differential equations, Acta Mathematica, 119 (1967), pp. 147–171. [31] A. Kolmogoroff, Zufallige Bewegungen (zur Theorie der Brownschen Bewegung), Annals of Mathematics, (1934), pp. 116–117. [32] J. LaSalle, Uniqueness theorems and successive approximations, Annals of Mathematics, (1949), pp. 722–730. [33] T. M. Liggett, L2 rates of convergence for attractive reversible near- est particle systems: the critical case, The Annals of Probability, (1991), pp. 935–959. [34] C. Mouhot and L. Neumann, Quantitative perturbative study of conver- gence to equilibrium for collisional kinetic models in the torus, Nonlinearity, 19 (2006), pp. 969–998. [35] J. Nash, Continuity of solutions of parabolic and elliptic equations, Amer- ican Journal of Mathematics, 80 (1958), pp. 931–954. [36] B. Perthame et al., Mathematical tools for kinetic equations, Bulletin of the American Mathematical Society, 41 (2004), pp. 205–244. [37] C. Villani, Hypocoercive diffusion operators, in International Congress of Mathematicians, vol. 3, 2006, pp. 473–498. [38] , Hypocoercivity, Memoirs of the American Mathematical Society, 202 (2009).

18