• Aucun résultat trouvé

A limiting case for the divergence equation

N/A
N/A
Protected

Academic year: 2021

Partager "A limiting case for the divergence equation"

Copied!
22
0
0

Texte intégral

(1)

HAL Id: hal-00639326

https://hal.archives-ouvertes.fr/hal-00639326v2

Submitted on 9 Nov 2011

HAL is a multi-disciplinary open access

archive for the deposit and dissemination of

sci-entific research documents, whether they are

pub-lished or not. The documents may come from

teaching and research institutions in France or

abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est

destinée au dépôt et à la diffusion de documents

scientifiques de niveau recherche, publiés ou non,

émanant des établissements d’enseignement et de

recherche français ou étrangers, des laboratoires

publics ou privés.

A limiting case for the divergence equation

Pierre Bousquet, Petru Mironescu, Emmanuel Russ

To cite this version:

(2)

(will be inserted by the editor)

A limiting case for the divergence equation

Pierre Bousquet · Petru Mironescu · Emmanuel Russ

November 7, 2011

Abstract We consider the equation divY = f , with f a zero average function on the torus Td. In their seminal paper [5], Bourgain and Brezis proved the existence of a solutionY ∈ W1,d

∩L∞for a datum f ∈ Ld. We extend their result to the critical

Sobolev spaces Ws,pwith (s + 1)p = d and p ≥ 2. More generally, we prove a similar result in the scale of Triebel-Lizorkin spaces. We also consider the equation divY = f in a bounded domainΩ subject to zero Dirichlet boundary condition.

Keywords Divergence equation · Triebel-Lizorkin spaces · critical embedding Mathematics Subject Classification (2000) 35F05 · 35F15 · 42B05 · 46E35

1 Introduction

This paper is devoted to the regularity of solutions of the equation

divY = f , where Z

f = 0. (1.1)

Here, the scalar function f and the vector fieldY are defined on the d-dimensional torus1Tdor in a smooth bounded domain

inRd.

Standard regularity theory asserts thatY gains one derivative with respect to f . For example, when f ∈ Lp, 1 < p < ∞, one can pickY ∈ W1,p.2In the limiting cases where f ∈ L1, respectively f ∈ L, it is not always possible to pickY ∈ W1,1[5],

respectivelyY ∈ W1,∞[18]; see also [8, 13].

In their seminal paper, Bourgain and Brezis [5] discovered another limiting situation: the case where f ∈ Ld. For such f,

(1.1) has a solutionY ∈ W1,d, so thatY ”almost” belongs to L∞. It turns out that (1.1) does have a solution in L∞. This was noted in [5, Proposition 1]. In principle, there is no reason to have a solution in both L∞and W1,d. Bourgain and Brezis [5] obtained the existence of a solution of (1.1) in L∞∩W1,d

for a datum f ∈ Ld. The proof is a real tour de force: the construction ofY is highly nontrivial and the proof of the fact that Y has the desired regularity is extremely involved. In the special case p = d = 2, existence of Y ∈ L∞∩ W1,2can be established via a simpler duality argument [5, Section 4]. More generally, when

f ∈ Wd/2−1,2, existence of a solutionY ∈ L∞∩Wd/2,2of (1.1) can be obtained by a similar strategy [12, 14]. However, no simple proof of existence is known when p 6= 2.

In this paper, we investigate the regularity properties of the Bourgain-Brezis fieldY in other function spaces. We con-sider regularity in Sobolev-Slobodeskii spaces Ws,p and more generally, in Triebel-Lizorkin spaces Fqs,p, focusing on the

limiting situation where (s + 1)p = d. In this case, if f ∈ Fqs,p, we expect a solutionY ∈ F s+1,p

q , and the latter space is ”close”

to L∞, thanks to the condition (s + 1)p = d.

Our main result is

Pierre Bousquet

Aix-Marseille Université; Université de Provence; CNRS, UMR 6632, LATP, Centre de Mathématiques et Informatique, 39 rue Frédéric Joliot-Curie, F-13453 Marseille Cedex 13, France

E-mail: bousquet@cmi.univ-mrs.fr Petru Mironescu

Université de Lyon; Université Lyon 1; INSA de Lyon, F-69621; Ecole Centrale de Lyon; CNRS, UMR 5208, Institut Camille Jordan, 43 blvd du 11 novembre 1918, F-69622 Villeurbanne-Cedex, France

E-mail: mironescu@math.univ-lyon1.fr Emmanuel Russ

Université Joseph Fourier; Institut Fourier; CNRS, UMR 5582; 100 rue des maths, BP 74, F-38402 Saint-Martin d’Hères Cedex, France Tel: (33)4 76 51 43 26

Fax: (33)4 76 51 44 78

E-mail: emmanuel.russ@ujf-grenoble.fr 1 We identifyTdwithRd/(2πZ)d.

2 This follows from elliptic regularity theory: it suffices to letY = ∇u, where u is an appropriate solution of∆ u

(3)

Theorem 1.1 Let p ≥ 2, q ∈ [2, p], d ≥ 2 and s > −1

2. We assume that

(s + 1)p = d. (1.2)

Let f ∈ Fqs,p(Td) satisfy the compatibility condition

Z

Tdf = 0. Then (1.1) has a solution Y ∈ L

(Td ) ∩ Fqs+1,p(Td) such that kYkL∞(Td)+ kYk Fs+1,pq (Td)≤ Ck f kF s,p q (Td). (1.3)

The result of Bourgain and Brezis corresponds to s = 0, p = d, q = 2. In particular, Theorem 1.1 contains the following regularity result in the scale of Sobolev-Slobodeskii spaces:

Corollary 1.2 Let p ≥ 2 and d ≥ 2. We assume that (1.2) holds and that s > −1

2. Let f ∈ W

s,p(Td) be such thatZ

Tdf = 0. Then (1.1) has a solutionY ∈ L∞(Td) ∩ Ws+1,p(Td) such that

kYkL∞(Td)+ kYkWs+1,p(Td)≤ Ck f kWs,p(Td).

Our proof follows the main lines of the one of Bourgain and Brezis in [5]. In particular, as in [5], we make use of a one-sided inequality, due to Rubio de Francia [20]; see Theorem 5.1 below. This inequality requires p ≥ 2; this is why the condition p ≥ 2 appears in both Theorem 1.1 and Corollary 1.2.3Whether the assumption p ≥ 2 can be removed is a challenging question. Observe also that we do not obtain the conclusions of Theorem 1.1 and Corollary 1.2 when s ≤ −1

2. This is probably due to our method.

The regularity result concerning the divergence equation was subsequently extended by Bourgain and Brezis [6] to more general Hodge systems. It is plausible that a version of Theorem 1.1 still holds for these systems. We will return to this question in a subsequent work.

Our paper is organized as follows. In Section 2, we recall the definition of the Triebel-Lizorkin spaces. In Section 3, we describe the Bourgain-Brezis construction, give the main steps of the proof of Theorem 1.1 and state the main estimates. In Sections 4 and 5 we collect the harmonic analysis background used in the proof of Theorem 1.1; the estimates we prove in Section 5 are crucial in the proof of Theorem 1.1. In Section 6, we establish the validity of the main estimates stated in Section 3. In Section 7, we discuss the solvability of (1.1) in bounded domains. More specifically, we prove the following

Theorem 1.3 Let p ≥ 2, q ∈ [2, p], d ≥ 2 and s > −1

2 satisfy (1.2). LetΩ be a smooth bounded domain in R

d. Let f ∈ Fs,p q (Ω)

satisfy the compatibility condition Z

Ωf = 0. 4

Then there existsY ∈ Fsq+1,p(Ω)∩ L∞(Ω) such that

(

divY = f in

trY = 0 onΩ. Moreover, we can chooseY such that kYkL∞(Ω)+ kYkFs+1,p

q (Ω)≤ C k f kF s,p q (Ω).

A final appendix gathers the proofs of some elementary estimates involving trigonometric polynomials used in the proof of Theorem 1.1.

Contents

1 Introduction . . . 1

2 Projections and Triebel-Lizorkin spaces . . . 2

3 Proof of Theorem 1.1 . . . 3

4 Kernels and multipliers . . . 6

5 An inequality of Rubio de Francia and applications . . . 9

6 Proof of Proposition 3.1 . . . 11

7 Inversion of the divergence in domains . . . 15

8 Appendix. Some estimates involving trigonometric polynomials . . . 20

2 Projections and Triebel-Lizorkin spaces

We denote byP (Td) the space of trigonometric polynomials.

If∆ is a subset of Zd, then we denote byPthe projection on the Fourier coefficients in∆: P∆( f )(x) :=X

n∈∆

b f (n)eın·x.

This projection makes sense if either∆ is finite and f ∈ D0(Td), or∆ is arbitrary and f ∈ P (Td).

Of special importance in the theory of function spaces are the projections on dyadic sets. These sets are finite unions of intervals5and are defined as follows:

∆d 0= {0},∆ d j:= [2 j−1 ≤ |n| < 2j], ∀ j ≥ 1; here, we work with the norm |n| = max©|nj|; j ∈J1, dKª.

6

3 The condition p2 amounts to sd/21. 4 When s

<0, this condition has to be suitably interpreted, see Section 7 below. 5 An interval inZdis a Cartesian product of intervals inZ.

6 The notation

(4)

One of the possible equivalent definitions of the Triebel-Lizorkin spaces is [22, Theorem 3.5.3] Fqs,p(Td) = ( f ∈ D0(Td) : kf kFqs,p(Td)= ° ° ° ° ° ° ° ° ° µ 2jsPd j ( f )(x) ¶ j∈N ° ° ° ° lq(N) ° ° ° ° °Lp(Td) < ∞ ) . (2.1)

The scale of Triebel-Lizorkin spaces contains in particular most of the Sobolev-Slobodeskii spaces: when 1 < p < ∞, we have Ws,p(Td) =

(

F2s,p(Td), if s ∈ N Fps,p(Td), if s 6∈ N.

In particular, when s = 0, the above and (2.1) amount to the square function theorem k f kLp ° ° ° ° ° ° ° ° µ Pd j ( f )(x) ¶° ° ° °l2(N) ° ° ° ° Lp(Td) , 1 < p < ∞. (2.2)

A word about the condition Z

Tdf = 0, which appears in (1.1): when s in Theorem 1.1 is positive, we have f ∈ L

1(Td), so

that the condition Z

Tdf = 0 makes sense. For arbitrary s, this condition has to be understood as b

f (0) = 0. (2.3)

3 Proof of Theorem 1.1

It suffices to establish (1.3) when f ∈ P (Td). The general case is obtained by density. We use the notations which will be introduced in (3.6)-(3.8) below. Since bf (0) = 0, we have f = f1+ · · · + fd, with fk

= PBk( f ). Theorem 1.1 will be a consequence of

Proposition 3.1 There existδ0> 0 andα> 0 such that, for every f ∈ P (Td) satisfying kf kFqs,p(Td)≤δ0, there exist functions Y1, . . . ,Yd:Td→ C such that kYkkL∞(Td)≤ 1, kYkkFs+1,p q (Td)≤ 1, for k ∈J1, dK, (3.1) and kkYk− fkkFs,pq (Td)≤ k f k 1 Fs,pq (Td), for k ∈J1, dK. (3.2)

From Proposition 3.1, we get at once

Corollary 3.2 There existδ0> 0 andα> 0 such that, for every 0 <δδ0and everyg ∈ Fqs,p(Td) satisfying (2.3), there exists

X = (X1, . . . ,Xd) ∈ F s+1,p q (Td) such that kXkkL∞(Td)≤ kgkFqs,p(Td) δ , kXkkFs+1,pq (Td)≤ kgkFs,pq (Td) δ , for k ∈J1, dK, (3.3) and kdiv X − gkFs,pq (Td)≤ dδ αkgk Fqs,p(Td). (3.4)

Proof of Theorem 1.1 using Corollary 3.2. Let f ∈ Fqs,p(Td). We set g0:= f andδ:= min¡(2d)−1/α,δ0¢. By Corollary 3.2, there

existsX0 ∈ Fqs+1,p(Td) such that kX0kkL∞(Td)Fs+1,p q (Td)≤ kg0kFs,pq (Td) δ , for k ∈J1, dK, kdiv X 0 − g0kFqs,p(Td)≤ dδ αkg 0kFqs,p(Td).

More generally, assume that g0, . . . , gl have been defined, as well as the corresponding vector fieldsX0, . . . ,Xl provided by

Corollary 3.2. Thus we have

kXlkkL∞(Td)Fs+1,p q (Td)≤ kglkFs,pq (Td) δ , for k ∈J1, dK, kdiv X l − glkFs,pq (Td)≤ dδ αkg lkFs,pq (Td). (3.5) We then set gl+1:= gl− div Xland letXl+1∈ Fqs+1,p(Td) satisfy (3.3) and (3.4) with g replaced by gl+1.

Combining (3.5) with our choice ofδ, we find that kgl+1kFqs,p(Td)≤ dδ αkg lkFs,pq (Td)≤ . . . ≤ (dδ α)l+1 kg0kFqs,p(Td)≤ 1 2l+1kg0kFqs,p(Td). It then follows that

kXlkkL∞(Td)Fs+1,p q (Td)≤

kg0kFqs,p(Td)

2lδ , for k ∈J1, dK. We can thus defineX := X

l

Xl

, which satisfies kXkL∞(Td)Fs+1,p

q (Td)≤ Ck f kF s,p

q (Td). Moreover, the fact that kglkF s,p

(5)

It remains to explain the

Proof of Proposition 3.1. We write

Zd \ {0} =[ j≥0 ³ B1 j∪ . . . B d j ´ , (3.6) with B1 j:= {n ∈∆ d j : 2 j−1 ≤ |n1| < 2j},B2j:= {n ∈∆ d j : 2 j−1 ≤ |n2| < 2j} \B1j, . . . ,B d j :=∆ d j\ Ãd −1 [ k=1 Bk j ! . (3.7)

Hence, for each j ≥ 0,∆dj = d

[

k=1

Bk

j, the union being disjoint. We also let

Bk =[ j≥0 Bk j, for k ∈J1, dK, (3.8) and B = B1 ∩ (N × Zd−1),Bj= B1j∩ (N × Z d−1).

We describe the construction ofY1which only involvesB1,B1jand f1= PB1( f ). The componentYk, with k ∈J2, dK, is built in the same way from fk= PBk( f ). We note that f1is a trigonometric polynomial, and that supp cf1⊂ B1. We also note that

kPBf1kFs,pq (Td)= ° ° ° ° ° ° ° ° ° µ 2s jPd jPB f1(x) ¶ j∈N ° ° ° ° lq(N) ° ° ° ° °Lp(Td) = ° ° ° ° ° ° ° ° ° µ 2s jPBjPd j f (x) ¶ j∈N ° ° ° ° lq(N) ° ° ° ° °Lp(Td) ≤ C ° ° ° ° ° ° ° ° ° µ 2s jPd j f (x) ¶ j∈N ° ° ° ° lq(N) ° ° ° ° °Lp(Td) = Ck f kFs,pq (Td),

the above inequality being a consequence of Theorem 4.9. It follows that it suffices to prove Proposition 3.1 when f is replaced byPBf1. Therefore, we assume, in the sequel, that f is a trigonometric polynomial such that supp bf ⊂ B.

Following the strategy in [5], we divide the proof of Proposition 3.1 into five steps.

Step 1. Estimates for the naive solution of (1.1). When supp bf ⊂ B, an exact solution of divX = f is given by X = (F,0,...,0), where F =X n∈B 1 ın1 b f (n)eın·x. (3.9)

The fieldY1will be constructed by modifyingX.

The purpose of Step 1 is to collect some estimates involving F and the projections of f and F, defined by

fj:= X n∈Bj b f (n)eın·x= PBjf = P∆d j f , Fj:= X n∈Bj 1 ın1 b f (n)eın·x= PBjF = P∆d j F. (3.10)

More specifically, we will establish the following

Lemma 3.3 We have kFkFs+1,pq (Td)≤ C(s, p, q)k f kF s,p q (Td), (3.11) ° °Fj°°L(Td)≤ C(s, p, q) ° °fj°°Fs,p q (Td), (3.12) ° °∇Fj ° °L∞(Td)≤ C(s, p, q)2j ° °fj ° °Fs,p q (Td), (3.13) and k fjkFs,pq (Td)≤ C(s, p, q)k f kF s,p q (Td). (3.14)

Step 2. Construction of a good function Hj dominating |Fj|. Letε= 2−`with`∈ N to be determined later. For j >`we

define k( j) := 1/ε= 2`and we decomposeBjinto disjoint vertical stripsBj,r, r ∈J1, k( j)K. More specifically, each Bj,ris of the form Bj,r= ³ Jaj,r, bj,r− 1K × Z d−1´ ∩ Bj, with bj,r− aj,r=ε2j−1:= l( j). Following [5], we set Gj(x) = X 1≤r≤1/ε ¯ ¯ ¯ ¯ ¯ X n∈Bj,r 1 n1 b f (n)eın·x ¯ ¯ ¯ ¯ ¯ (3.15) and Hj(x) = 3dGj∗ (Fε2j⊗ F2j⊗ · · · ⊗ F2j). (3.16) Here and after, we denote byFN the Fejér kernel given by

(6)

We set, for all (i1, ..., id) ∈ Nd,

Fi1⊗ . . . ⊗ Fid(x1, . . . , xd) := Fi1(x1). . .Fid(xd). We clearly have, for all x ∈ Td,

|Fj(x)| ≤ Gj(x) (3.18)

and also

suppHbj⊂ [|n| ≤ 2j− 1]. (3.19)

A slightly more involved estimate, taken from [5], is that, for all x ∈ Td,

Gj(x) ≤ Hj(x); (3.20)

this follows from Corollary 8.5 in the appendix.

Up to now, the functions Gjand Hjhave been defined only for j >`. When j ≤`(that is, whenε2j−1< 1), we do not split Bj.

Instead, we let

aj,1= 2j−1, bj,1= 2j, Gj= |Fj| (3.21)

and define Hjvia

Hj= 3dGj∗ (F2j⊗ F2j⊗ · · · ⊗ F2j). (3.22) In this case, we set k( j) = 1 and l( j) = bj,1− aj,1= 2j−1. It will follow form the proof of (3.20) that estimates (3.18), (3.21) and

(3.22) remain valid.

The main purpose of Step 2 is to establish the following estimates satisfied by Gjand Hj.

Lemma 3.4 We have X 1≤r≤k( j) ° ° ° ° ° x 7→ X n∈Bj,r b f (n)eın·x ° ° ° ° ° p Lp(Td) ≤ C(p)k fjkLpp(Td)≤ C(p)2− s j p k f kpFs,p q (Td) , (3.23) kGjkL∞(Td)≤ C(s, p, q)k( j)1−2/p ° °fj°°Fs,p q (Td), (3.24) and k∇GjkL∞(Td)≤ C(s, p, q)2jk( j)1−2/pk fjkFqs,p(Td). (3.25) An immediate consequence of (3.14), (3.16), (3.24) and of kFNkL1(T)= 1 is

° °Hj ° °L∞(Td)≤ C(s, p, q)k( j)1−2/pk fjkFs,pq (Td)≤ C(s, p, q)k( j) 1−2/p k f kFs,pq (Td). (3.26) Similarly, we have k∇HjkL∞(Td)≤ 3dk∇GjkL∞(Td)≤ C(s, p, q)2jk( j)1−2/pk fjkFs,pq (Td)≤ C(s, p, q)2 j k( j)1−2/pk f kFs,p q (Td). (3.27)

Step 3. Construction ofY1. Following [5], we let

Y1:= X j≥0 Fj Y k>j (1 − Hk). Set Kj:= X k<j Fk Y k<m<j (1 − Hm) if j > 0, K0:= 0.

The next result will be an immediate consequence of Lemma 3.4.

Lemma 3.5 One has

Y1= X j≥0 Fj− X j≥0 HjKj (3.28) and 1Y1= X j≥0 fj− X j≥0 1(HjKj). (3.29)

Assume furthermore that k( j)1−2/p

k f kFqs,p(Td)≤η0for all j ∈ N, whereη0> 0 is a constant which will be determined in the proof. Then, for allx ∈ Td,

|Y1(x)| ≤ 1, (3.30) ¯ ¯Kj(x) ¯ ¯≤ 1, ∀ j ∈ N (3.31) and 0 ≤¯ ¯Fj(x)¯¯≤ Gj(x) ≤ Hj(x) ≤ 1, ∀ j ∈ N. (3.32)

In particular, Lemma 3.5 implies the L∞estimate in (3.1).

(7)

Lemma 3.6 There existδ0> 0,α> 0 and C = C(s, p, q) > 0 such that, if k f kFqs,p(Td)≤δ0, then ° ° ° ° ° X j≥0 1(HjKj) ° ° ° ° ° Fs,pq (Td) ≤ C k f k1Fs,p q (Td) . (3.33)

Step 5. Proof of (3.1) completed: estimate of k∇Y1kFs,pq (Td). This is achieved via the following

Lemma 3.7 Letη0be as in Lemma 3.6. Then there exists a constantδ0> such that, if

k f kFqs,p(Td)≤ min ³ ε1−2/pη 0,δ0 ´ , (3.34) then k∇Y1kFqs,p(Td)≤ 1.

It is easy to see that Proposition 3.1 is a consequence of the five above steps. ä

4 Kernels and multipliers

We will make use of the following classical kernels defined onT: 1. The Fejèr kernelFN, given by (3.17).

2. The de la Vallée Poussin kernelVN= 2F2N− FN.

Ifϕ:Zd→ C and f is a trigonometric polynomial on Td, we let Tϕ( f )(x) := X

n∈Zd

ϕ(n) bf (n)eın·x, x ∈ Td.

Similarly, given, for any j ∈ N, a functionϕj:Zd→ C and a trigonometric polynomial fj, we setϕ= (ϕj)j∈N, respectively

f = (fj)j∈N, and define Tϕby the formula

f 7→  x 7→ Ã X n∈Zd ϕj(n) bfj(n)eın·x ! j∈N  . (4.1)

More generally, the above formulae make sense if the trigonometric polynomial f has coefficients in a vector space overC. We start by recalling a classical estimate on Fourier multipliers.

Theorem 4.1 [23, Section I.6.2] Let k ∈ D0(Td). Assume that

1. ϕ:= bk ∈ l∞(Zd).

2. k ∈ L1l oc

¡

Td\ {0}¢.

Then for each 1 < p < ∞ there exists Cp> 0 such that

¯ ¯ ¯ ¯Tϕf¯ ¯ ¯ ¯Lp(Td)≤ CpB(ϕ) || f ||Lp(Td) , ∀ f ∈ P (Td). Here, B(ϕ) := max ( ¯ ¯ ¯ ¯ϕ ¯ ¯ ¯ ¯l∞(Zd), sup x∈Td Z |x−y|≥2|x||k(y − x) − k(y)| d y ) .

The next result that we quote is a vector-valued version of Theorem 4.1; see [1, Theorem 2]. In [1], Theorem 4.2 is stated inRd, but the proof is easily adapted toTd.

Theorem 4.2 [1, Theorem 2] Let A, B be Banach spaces and let k ∈ L1(Td,L (A,B)). Define T : Lp(Td, A) → Lp(Td, B) through the formula

T f (x) = k ∗ f (x) = Z

k(x − y)f (y) d y, ∀ f ∈ Lp(Td, A). (4.2) Fix 1 < p0< ∞ and let

M ≥ Z

|y−x|≥2|x|kk(y − x) − k(y)kL(A,B)d y, ∀ x ∈ T

d. (4.3)

Then, for 1 < p < ∞, the norm kTkLpLpofT as a linear continuous operator from Lp(Td, A) into Lp(Td, B) is controlled by a quantity depending solely onM, on p and on kTkLp0Lp0, but not on kkkL1(Td,L(A,B)).

We next derive some consequences of Theorem 4.2. To start with, we consider the case where i.d = 1.

ii.A = B = lq(N), with 1 < q < +∞.

iii. The operator T f acts, on a sequence f = (fj)j∈Nof functions in L1(T), through the formula

T f (x) =³FNj∗ fj(x) ´

j∈N. (4.4)

More specifically, we will be interested in the case where

(8)

Lemma 4.3 Let 1 < p < ∞ and 1 < q < ∞. Then T defined by (4.4) and (4.5) is continuous from Lp(T; lq(N)) into Lp(T; lq(N)). In addition, we have ° ° ° ° ° ° ° ° ³ FNj∗ fj ´ j∈N ° ° ° °lq(N) ° ° ° ° Lp(T)≤ C(p, q) ° ° ° ° ° °¡ fj ¢ j∈N ° ° °lq(N) ° ° ° Lp(T), (4.6) withC(p, q) independent of kj.

Proof By a standard limiting procedure, it suffices to prove (4.6) (with bounds independent of J ∈ N) for the truncated operator, still denoted T, defined on Lp¡

T, lq¡

J0, J K¢¢ with values into L

p¡ T, lq¡

J0, J K¢¢ through the formula T f (x) =³FNj∗ fj(x)

´

j∈J0, J K

, ∀fj∈ Lp(T,C), j ∈J0, J K.

For this purpose, it suffices to check that the assumptions of Theorem 4.2 are satisfied with p0= q, and that the norm

kTkLqLq as well as the right-hand side of (4.3) have upper bounds independent of kjand J. Let us start by computing kTkLqLq. For each j, we have

Z T|FNj∗ fj| q (x) dx ≤ Z T| fj| q (x) dx, since°° °FNj ° °

°L1(T)= 1. By taking the sum over j in the above inequality, we find that kTkLq→Lq≤ 1. We next obtain a uniform estimate for the right-hand side of (4.3). We have

kk(y − x) − k(y)kL(lq(N),lq(N))= sup

j |F

Nj( y − x) − FNj( y)| ≤ X

k

|F2k( y − x) − F2k( y)|. (4.7)

In view of (4.7), it suffices to establish the estimate

X

k

Z

|y−x|≥2|x||F2

k( y − x) − F2k( y)| d y ≤ C, ∀ x ∈ T. (4.8) In order to prove (4.8), we rely, on the one hand, on Bernstein’s integral inequality [9, Theorem D.2.1], which yields

Z

T|FN( y − x) − FN( y)| d y ≤ CN|x|. (4.9)

On the other hand, we haveFN( y) ≤

C N y2, and therefore Z |x−y|≥2|x||F N( y − x) − FN( y)| d y ≤ C N Z |y|≥|x| d y y2 ≤ C N|x|. (4.10)

Let x ∈ T \ {0}. If |x| ≥ 1, then (4.10) implies that X k Z |x−y|≥2|x||F2 k( y − x) − F2k( y)| d y ≤ C X k 1 2k |x|≤ C, (4.11)

guaranteeing the validity of (4.8) for such x.

Assume next that |x| < 1. Let k0∈ N be such that 2−k0−1≤ |x| < 2−k0. Thanks to (4.9) and (4.10), we have

X k Z |x−y|≥2|x||F2 k( y − x) − F2k( y)| d y = X k≤k0 . . . + X k>k0 . . . ≤ C|x| X k≤k0 2k+ C |x| X k>k0 1 2k≤ C,

which proves (4.8). This completes the proof of the lemma. ä

Lemma 4.3 combined with a Fubini type argument7leads to

Lemma 4.4 Let 1 < p < ∞ and 1 < q < ∞. Let

T³¡ fj¢j∈N ´ =³³F2k1, j⊗ . . . ⊗ F2kd, j ´ ∗ fj ´ j∈N, ∀ fj∈ P (T d ), ∀ j ∈ N. Then ° ° ° ° ° °T ³ ¡ fj¢j∈N ´° ° °lq(N) ° ° ° Lp(Td)≤ C(p, q) ° ° ° ° ° °¡ fj ¢ j∈N ° ° °lq(N) ° ° ° Lp(Td), (4.12) withC(p, q) independent of k1, j, . . . , kd, j.

Let us also recall the scalar and the vector-valued Marcinkiewicz Theorem on Fourier multipliers:

Theorem 4.5 Letϕ:Z → C and 1 < p < ∞.

(9)

1. [9, Theorem 8.2.1] The operator Tϕ, initially defined for trigonometric polynomials f ∈ P (Td), can be extended to an

Lp-bounded operator satisfying

kTϕf kLp(T)≤ C(p)A(ϕ)kf kLp(T), (4.13) provided A(ϕ) := max  kϕkl∞(Z), sup j X n∈∆1j |ϕ(n + 1) −ϕ(n)|  < ∞.

2. [25, Proposition 2] LetX be a Banach space satisfying the UMD property. Then the conclusion of item 1. holds for X -valued maps. More specifically, ifA(ϕ) < ∞, then the operator

f 7→ Tϕ( f ) =X

n∈Z

ϕ(n) bf (n)eınx,

initially defined for trigonometric polynomialsf with coefficients in X , f ∈ P (T, X ), can be extended to a linear operator onLp(T, X) satisfying (4.13).

For a discussion on the UMD property, see [11]. Some results related to item 2. above and the UMD property can be found, e.g., in [10]. Of importance for us is that the space lq(N), with 1 < q < ∞, has the UMD property.

We will also need the following d-dimensional version of Theorem 4.5:

Corollary 4.6 Let 1 < p < ∞. Let X be a Banach space with the UMD property, ϕj:Z → C,1 ≤ j ≤ d andϕ=ϕ1⊗ . . . ⊗ϕd.

Then kTϕf kLp(Td,X )≤ C(p, X )A(ϕ1). . . A(ϕd)kf kLp(Td,X ), ∀ f ∈ P (Td), where A(ϕl) := max  kϕlkl∞(Z), sup j X n∈∆1 j |ϕl(n + 1) −ϕl(n)|  . Proof If g :Td

→ X , we denote by g∗x0the map x17→ g(x1, x0). Observe that

Tϕf (x) = Tϕ1((Tψf )∗x0)(x1),

whereψ(n1, n2, . . . , nd) =ϕ2(n2). . .ϕd(nd). Theorem 4.5 and the Fubini theorem imply

kTϕf kLpp(Td,X )= Z Td−1kTϕ1((Tψf )∗x0)k p Lp(T,X )dx0. A(ϕ1)p Z Td−1k(Tψf )∗x0k p Lp(T,X )dx0 = A(ϕ1)pkTψf kpLp(Td,X ). ... . A(ϕ1) p. . . A(ϕ d)pk f kLpp(Td,X ). ä

Here and in what follows, . stands for ≤ C, with appropriate C. We will often rely on the following special case of Corollary 4.6.

Corollary 4.7 Let X be a Banach space with the UMD property and let 1 < p < ∞.

Letf ∈ P (Td, X ) be such that supp bf ⊂

d

Y

j=1

Jaj, bjK.

Letϕj:Z → R, j ∈J1, dK, be monotonic in the sets [nj> 0] and [nj< 0]. Setϕ=ϕ1⊗ . . . ⊗ϕd. Then

kTϕf kLp(Td,X )≤ C(p, X )kϕ1kl∞(Ja1,b1K). . . kϕdkl∞(Jad,bdK)k f kLp(Td,X ). (4.14)

Proof Apply Corollary 4.6 to the functionsϕj1Jaj,bjK. ä

Vector-valued Fourier multiplier are valid beyond multipliersϕof the formϕ1⊗ ... ⊗ϕd. Here is a special case we will

rely on in the sequel. The next result is a straightforward consequence of [25, Proposition 2].

Lemma 4.8 Let 1 < p < ∞ and 1 < q < ∞. Let (Pj)j∈Nbe a sequence of trigonometric polynomials. Then

° ° ° ° ° ° ° ° ° µ 1Pd j (Pj) ¶ j∈N ° ° ° ° lq(N) ° ° ° ° °Lp(Td) ≤ C(p, q) ° ° ° ° ° ° ° ° µ 2jPd j (Pj) ¶° ° ° °lq(N) ° ° ° ° Lp(Td) . (4.15)

We will also make use of the following vector-valued version of the Riesz inequality.

Theorem 4.9 [19, 3] Let 1 < p < ∞ and 1 < q < ∞. Let (Ij)j be an arbitrary sequence of intervals inZd. For all j ∈ N, let

(10)

5 An inequality of Rubio de Francia and applications

We start by recalling the following one-sided estimate due to Rubio de Francia [20].

Theorem 5.1 [20] Let 2 ≤ p < ∞. Let (Ij) be an arbitrary sequence of pairwise disjoint intervals inZ. Then

° ° ° ° ° ° ° ° ³ PIjf ´ j∈N ° ° ° °l2(N) ° ° ° ° Lp(T) ≤ C(p) k f kLp(T), ∀ f ∈ P (T).

Here,C(p) is independent of the choice of the disjoint intervals.

In the remaining part of this section, we establish two consequences of Theorem 5.1 (Corollaries 5.3 and 5.5). These results will play a crucial role in the proof of Theorem 1.1.

Lemma 5.2 Let 2 ≤ q ≤ p < ∞. Consider, in each dyadic set∆1j, a family ofk( j) pairwise disjoint intervals Ij,r,r ∈J1, k( j)K. Then there existsC(p, q) such that

° ° ° ° ° ° ° ° ° ° µ X r ¯ ¯ ¯PIj,rf ¯ ¯ ¯ ¶ j∈N ° ° ° ° °lq(N) ° ° ° ° ° Lp(T) ≤ C(p, q) ° ° ° ° ° ° ° ° ³ (k( j))1/2P1 jf ´ j∈N ° ° ° °lq(N) ° ° ° ° Lp(T) , ∀ f ∈ P (T). (5.1) Though we will apply the above lemma with k( j) as in Step 2 of the proof of Proposition 3.1 and to the equal length intervals considered there, we emphasize the fact that here k( j) and the intervals are arbitrary.

Proof By the Cauchy-Schwarz inequality, we have, for all j ∈ N, X r ¯ ¯ ¯PIj,rf ¯ ¯ ¯ ≤(k( j)) 1/2µX r ¯ ¯ ¯PIj,rf ¯ ¯ ¯ 2¶1/2 . (5.2)

In view of (5.2), (5.1) will follow from the estimate

° ° ° ° ° ° ° ° ° ° ° õ X r ¯ ¯ ¯PIj,rf ¯ ¯ ¯ 2¶1/2! j∈N ° ° ° ° ° lq(N) ° ° ° ° ° °Lp(T) . ° ° ° ° ° ° ° ° ³ P1 jf ´ j∈N ° ° ° °lq(N) ° ° ° ° Lp(T) (5.3)

applied to the functionX(k( j))1/2P1 jf .

In turn, estimate (5.3) will be obtained by interpolation, starting from the limiting cases q = 2 and q = p. i. Proof of (5.3) when q = 2. When q = 2, (5.3) amounts to

° ° ° ° ° ° ° ° ³ PIj,rf ´ j∈N,r∈J1,k( j)K ° ° ° °l2 ° ° ° °Lp(T). ° ° ° ° ° ° ° ° ³ P1 jf ´ j∈N ° ° ° °l2(N) ° ° ° ° Lp(T) . (5.4)

By Theorem 5.1, the left hand side of (5.4) can be estimated by kf kLp(T). By the square function theorem, the right hand side of (5.4) is equivalent to kf kLp(T). This proves (5.3) for q = 2.

ii. Proof of (5.3) when q = p. In this case, (5.3) is equivalent to X j≥0 Z T µ X r |P Ij,rf | 2¶p/2 .X j≥0 Z T|P∆1jf | p.

The above estimate follows from

Z T µ X r |P Ij,rf | 2¶p/2 . kP1 jf k p Lp(T), ∀ j ≥ 0, which in turn is a consequence of Theorem 5.1 applied toP1

jf . iii. Proof of (5.3) when 2 < q < p. Let us consider the Banach space

Xp,q:=

n

( fj)j∈N∈ Lp(T, lq(N)); suppfbj⊂∆1j, ∀ j ∈ N

o .

Define the linear map

Xp,q3 ( fj)j∈N T 7−→³¡uj,r¢r∈J1,k( j)K ´ j∈N, where uj,r=(P Ij,rfj, if r ∈J1, k( j)K 0, otherwise .

Up to now, we have proved (5.3) for q = 2 and q = p. Equivalently, we have proved that T is continuous from Xp,2 into

Lp(T, l2(N, l2(N))) and from Xp,pinto Lp(T, lp(N, l2(N))), with norms controlled by a constant.

(11)

We continue as follows: we have (l2(N, l2(N)), lp(N, l2(N)))θ

= lq(N, l2(N)), which implies

(Lp(T, l2(N, l2(N))),Lp(T, lp(N, l2(N))))θ= Lp(T, lq(N, l2(N))).

Note that the interpolation results used in this part of the proof can be found in [2, Theorem 5.1.2]. By complex interpolation, T is continuous from (Xp,2, Xp,p)θinto Lp(T, lq(N, l2(N)). In view of (5.5), T maps continuously Xp,qinto Lp(T, lq(N, l2(N))).

This is precisely (4.8) and the conclusion of Lemma 5.2. Let us now prove (5.5). By Theorem 4.9, the projection map

Lp(T, lq(N)) 3 (fj)j∈N7−→π ³ P1 jfj ´ j∈N∈ Xp,q

is well-defined and continuous. Clearly,πis onto and the continuous embedding J : Xp,q→ Lp(lq) is a right inverse ofπ.

By complex interpolation,πis continuous from the space (Lp(T, l2(N)),Lp(T, lp(N)))θ into the space (Xp,2, Xp,p)θ, and the continuous embedding J : (Xp,2, Xp,p)θ→ (Lp(T, l2(N)),Lp(T, lp(N)))θis a right inverse ofπ. Using the equality

(Lp(T, l2(N)),Lp(T, lp(N)))θ= Lp(T, lq(N)),

we find that (5.5) holds. ä

The Fubini theorem combined with Lemma 5.2 leads to the following

Corollary 5.3 Let 2 ≤ q ≤ p < ∞. Consider, in each dyadic set∆1j, a family ofk( j) pairwise disjoint intervals Ij,r,r ∈J1, k( j)K. Then there existsC(p, q) such that, for each f ∈ P (Td), we have

° ° ° ° ° ° ° ° ° ° µ X r ¯ ¯ ¯PIj,r×Zd−1f ¯ ¯ ¯ ¶ j∈N ° ° ° ° °lq(N) ° ° ° ° ° Lp(Td) ≤ C(p, q) ° ° ° ° ° ° ° ° ³ (k( j))1/2P1 j×Zd−1f ´ j∈N ° ° ° °lq(N) ° ° ° ° Lp(Td) . (5.6)

In contrast with the hypotheses of Lemma 5.2, in the next result the intervals Ij,rare not arbitrary.

Lemma 5.4 Assume that 2 ≤ q ≤ p < ∞. Consider, in each dyadic set∆1j, a family ofk( j) pairwise disjoint intervals Ij,r,

r ∈J1, k( j)K, all of same cardinal l ( j) = 2

m( j), with

m( j) ∈ N. Then there exists C(p, q) such that ° ° ° ° ° ° ° ° ° ° µ X r ¯ ¯ ¯PIj,rf ¯ ¯ ¯ 0¶ j∈N ° ° ° ° °lq(N) ° ° ° ° ° Lp(T) ≤ C(p, q) ° ° ° ° ° ° ° ° ³ (k( j))1/2l( j)P1 j f´ j∈N ° ° ° °lq(N) ° ° ° ° Lp(T) , ∀ f ∈ P (T). (5.7) Proof Arguing as at the beginning of the proof of Lemma 5.2, it suffices to establish the estimate

° ° ° ° ° ° ° ° ° ° ° à µ X r ³¯ ¯ ¯PIj,rf ¯ ¯ ¯ 0´2¶1/2 ! j∈N ° ° ° ° ° lq(N) ° ° ° ° ° °Lp(T) . ° ° ° ° ° ° ° ° ³ l( j)P1 jf ´ j∈N ° ° ° °lq(N) ° ° ° ° Lp(T) , ∀ f ∈ P (T). (5.8) We write Ij,r=Jaj,r, bj,r− 1K, where bj,r− aj,r= l( j) = 2

m( j). We rely on the key inequality

¯ ¯ ¯ ¯ ¯ ¯PIj,rf ¯ ¯ ¯ 0¯ ¯ ¯ ≤l( j) ¯ ¯ ¯Fl( j)∗ (e −ıaj,rxP Ij,rf ) ¯ ¯ ¯, (5.9)

whose proof is postponed to the appendix; see Lemma 8.6. Taking (5.9) for granted, we obtain (5.8) provided the following inequality holds: ° ° ° ° ° ° ° ° ° ° ° Ã µ X r ¯ ¯ ¯Fl( j)∗ (e −ıaj,rxP Ij,rf ) ¯ ¯ ¯ 2¶1/2! j∈N ° ° ° ° ° lq(N) ° ° ° ° ° °Lp(T) . ° ° ° ° ° ° ° ° ³ P1 jf ´ j∈N ° ° ° °lq(N) ° ° ° ° Lp(T) . (5.10)

Indeed, (5.8) follows by applying (5.10) toX l( j)P1

jf and using (5.9).

We now turn to (5.10). It suffices to establish the validity of this estimate in the limiting cases q = 2 and q = p. Indeed, if this is proved then we are in position to repeat the interpolation argument used in the proof of Lemma 5.2.8

i. Proof of (5.10) when q = 2. We have to prove the inequality ° ° ° ° ° ° ° ° ³ Fl( j)∗ (e−ıaj,rxPIj,rf ) ´ j∈N,r∈J1,k( j)K ° ° ° °l2(N) ° ° ° ° Lp(T) . ° ° ° ° ° ° ° ° ³ P1 jf ´ j∈N ° ° ° °l2(N) ° ° ° ° Lp(T) .

This is obtained as follows: we have ° ° ° ° ° ° ° ° ³ Fl( j)∗ (e−ıaj,rxPIj,rf ) ´ j∈N,r∈J1,k( j)K ° ° ° °l2(N) ° ° ° ° Lp(T) (a) . ° ° ° ° ° ° ° ° ³ e−ıaj,rxP Ij,rf ´ j∈N,r∈J1,k( j)K ° ° ° °l2(N) ° ° ° ° Lp(T) (b) . ° ° ° ° ° ° ° ° ³ P1 j f´ j∈N ° ° ° °l2(N) ° ° ° ° Lp(T) . Here,

(a) is a consequence of Lemma 4.3.

(b) follows from Theorem 5.1 and the square function theorem (2.2).

8 Interpolation is applied to the operator f

(12)

ii. Proof of (5.10) when q = p. In this case, the left-hand side of (5.10), denoted by A, is estimated as follows A = ° ° ° ° ° ° ð ° ° ° ° µ X r ¯ ¯ ¯Fl( j)∗ (e −ıaj,rxP Ij,rf ) ¯ ¯ ¯ 2¶1/2° ° ° ° °Lp(T) ! j∈N ° ° ° ° ° °lp(N) (a) . ° ° ° ° ° ° ð ° ° ° ° µ X r ¯ ¯ ¯e −ıaj,rxP Ij,rf ¯ ¯ ¯ 2¶1/2° ° ° ° °Lp(T) ! j∈N ° ° ° ° ° °lp(N) = ° ° ° ° ° ° ð ° ° ° ° µ X r ¯ ¯ ¯PIj,rf ¯ ¯ ¯ 2¶1/2° ° ° ° °Lp(T) ! j∈N ° ° ° ° ° °lp(N) (b) . ° ° ° ° ° ° ° ° ³ P1 jf ´ j∈N ° ° ° °Lp(T) ° ° ° ° lp(N) = ° ° ° ° ° ° ° ° ³ P1 jf ´ j∈N ° ° ° °lp(N) ° ° ° ° Lp(T) ;

items (a) and (b) are justified as above. ä

Corollary 5.5 Assume that 2 ≤ q ≤ p < ∞. Consider, in each dyadic set∆1j, a family ofk( j) pairwise disjoint intervals Ij,r,

r ∈J1, k( j)K, all of same cardinal l ( j) = 2

m( j), with

m( j) ∈ N. Then there exists C(p, q) such that, for each f ∈ P (Td), we have ° ° ° ° ° ° ° ° ° ° µ X r 1 ¯ ¯ ¯PIj,r×Zd−1f ¯ ¯ ¯ ¶ j∈N ° ° ° ° °lq(N) ° ° ° ° ° Lp(Td) ≤ C(p, q) ° ° ° ° ° ° ° ° ³ (k( j))1/2l( j)P1 j×Zd−1f ´ j∈N ° ° ° °lq(N) ° ° ° ° Lp(Td) . (5.11)

Remark 5.6 The proof of Lemma 5.4 also yields the following more general result. Assume that 2 ≤ q ≤ p < ∞. Consider,

in each dyadic set ∆1j, a family of k( j) pairwise disjoint intervals Ij,r, r ∈J1, k( j)K, each of cardinal l ( j, r) = 2

m( j,r), with

m( j, r) ∈ N. Then there exists C(p, q) such that ° ° ° ° ° ° ° ° ° ° µ X r ¯ ¯ ¯PIj,rf ¯ ¯ ¯ 0¶ j∈N ° ° ° ° °lq(N) ° ° ° ° ° Lp(T) ≤ C(p, q) ° ° ° ° ° ° ° ° ° ° µ (k( j))1/2X r l( j, r)PIj,rf ¶ j∈N ° ° ° ° °lq(N) ° ° ° ° ° Lp(T) , ∀ f ∈ P (T). (5.12) 6 Proof of Proposition 3.1 Proof of Lemma 3.3.

i. Proof of (3.11). Recall that supp ˆf ⊂ B. Since in Bjwe have 2j−1≤ n1< 2j, we find that

³ 2(s+1) jFj ´ j∈N= Tϕ µ ³ 2s jfj ´ j∈N ¶ . (6.1)

Here,ϕis the scalar function (acting on vector-valued functions)ϕ=ϕ1⊗ . . . ⊗ϕd, whereϕ1(n1) =

2j

ın1

if 2j−1

≤ n1< 2j,

0 otherwise andϕk≡ 1 for k ∈J2, dK. By combining (6.1) with Corollary 4.6, we find that

kFkFs+1,p q = ° ° ° ° ° ° ° ° ³ 2(s+1) jFj ´ j∈N ° ° ° °lq(N) ° ° ° ° Lp(Td) = ° ° ° ° ° ° ° ° Tϕ³2s jfj ´ j∈N ° ° ° °lq(N) ° ° ° ° Lp(Td) . ° ° ° ° ° ° ° ° ³ 2s jfj ´ j∈N ° ° ° °lq(N) ° ° ° ° Lp(Td) = k f kFs,pq .

ii. Proof of (3.12). SincePd k (Fj) =δjkFjandPd k ( fj) =δjkfj, we have kFjkFqs,p(Td)= 2 s j kFjkLp(Td) and kfjkFs,pq (Td)= 2 s j k fjkLp(Td). (6.2) Corollary 4.7 applied toϕ1(n1) = 1 ın1

if n1≥ 1 andϕ1(n1) = 0 otherwise andϕk≡ 1 for k ∈J2, dK implies that

kFjkLp(Td). 2−jk fjkLp(Td). (6.3) Lemma 8.2 in the appendix combined with (1.2) and (6.3) yields

° °Fj ° °L(Td). 2jd/p ° °Fj ° °Lp(Td). 2j(d/p−1) ° °fj ° °Lp(Td)= ° °fj ° °Fs,p q (Td). (6.4)

iii. Proof of (3.13). Lemma 8.2 combined with (6.4) gives

k∇FjkL∞(Td). 2jkFjkL∞(Td). 2j ° °fj°°Fs,p

q (Td). iv. Proof of (3.14). Since fj= Pd

(13)

i. Proof of (3.23). Using the fact thatBj,r=¡Jaj,r, bj,r− 1K × Z

d−1¢ ∩ B

jfor all 1 ≤ r ≤ k( j), we obtain that

I := X 1≤r≤k( j) ° ° ° ° ° x 7→ X n∈Bj,r b fj(n)eın·x ° ° ° ° ° p Lp(Td) satisfies I = X 1≤r≤k( j) Z Td−1 Z T ¯ ¯ ¯ ¯ ¯ X n1∈Jaj,r,bj,r−1K à ( fj)∗x0(n1)eın1x1 ¯ ¯ ¯ ¯ ¯ p dx1dx0 (a) ≤ Z Td−1 Z T à X 1≤r≤k( j) ¯ ¯ ¯ ¯ ¯ X n1∈Jaj,r,bj,r−1K à ( fj)∗x0(n1)eın1x1 ¯ ¯ ¯ ¯ ¯ 2!p/2 dx1dx0 (b) . Z Td−1 Z T|( fj)∗x0| pdx 1dx0= k fjkLpp(Td). Here, (a) is a consequence ofX |ar|p≤¡X |ar|2¢ p/2 , valid when p ≥ 2. (b) follows from Theorem 5.1 applied to I1

r=Jaj,r, bj,r− 1K and f = ( fj)∗x0. This completes the proof of (3.23).

ii. Proof of (3.24). When j ≤ l, this is just (3.12). If j > l, by Hölder’s inequality we have

° °Gj ° °L∞(Td)≤ k( j)1/p 0   X 1≤r≤k( j) ° ° ° ° ° x 7→ X n∈Bj,r 1 n1 b fj(n)eın·x ° ° ° ° ° p L∞(Td)   1/p . (6.5) Corollary 4.7 applied toϕ1(n1) = 1 n1

when n1≥ 1, 0 otherwise, andϕj≡ 1 for j ∈J2, dK implies that ° ° ° ° ° x 7→ X n∈Bj,r 1 n1 b fj(n)eın·x ° ° ° ° ° Lp(Td) . 2−j ° ° ° ° ° x 7→ X n∈Bj,r b fj(n)eın·x ° ° ° ° ° Lp(Td) . (6.6) Corollary 8.3 applied to x 7→ X n∈Bj,r 1 n1 b

fj(n)eın·xand combined with (1.2), (6.2), (6.5), (6.6) and (3.23) yields

kGjkL∞(Td). k( j)1/p 0   X 1≤r≤k( j) ³ 2j(d−1)/p(l( j))1/p´p ° ° ° ° ° x 7→ X n∈Bj,r 1 n1 b fj(n)eın·x ° ° ° ° ° p Lp(Td)   1/p . k( j)1/p0l( j)1/p2j(d−1)/p2−j   X 1≤r≤k( j) ° ° ° ° ° x 7→ X n∈Bj,r b fj(n)eın·x ° ° ° ° ° p Lp(Td)   1/p . k( j)1−2/p2s j   X 1≤r≤k( j) ° ° ° ° ° x 7→ X n∈Bj,r b fj(n)eın·x ° ° ° ° ° p Lp(Td)   1/p . k( j)1−2/p2s jk fjkLp(Td)= k( j)1−2/pk fjkFqs,p(Td).

In the third inequality, we used the fact that k( j)l( j) = 2j−1.

iii. Proof of (3.25). The proof is similar to the one of (3.24). When j ≤ l, (3.25) is a consequence of (3.13). When j > l, we start from the inequality

|kGj(x)| ≤ X 1≤r≤k( j) ¯ ¯ ¯ ¯ ¯ X n∈Bj,r nk n1 b f (n)eın·x ¯ ¯ ¯ ¯ ¯ .

We continue as in the proof of (3.24). The only noticeable difference is that, in the proof, estimate (6.6) is replaced with

° ° ° ° ° x 7→ X n∈Bj,r nk n1 b fj(n)eın·x ° ° ° ° ° Lp(Td) . ° ° ° ° ° x 7→ X n∈Bj,r b fj(n)eın·x ° ° ° ° ° Lp(Td) . (6.7)

In turn, (6.7) with k 6= 1 follows from Corollary 4.7 applied toϕ1(n1) =

1 n1

,ϕk(nk) = nk,ϕm(nm) ≡ 1 for m 6= 1, k. ä

Proof of Lemma 3.5.

i. Proof of (3.28) and (3.29). Clearly, (3.29) is a consequence of (3.28). By straightforward induction on m ≥ 0, we find that

m X j=0 Fj Y j<k≤m (1 − Hk) = m X j=0 Fj− m X j=0 HjKj. (6.8)

(14)

ii. Proof of (3.30), (3.31) and (3.32). Inequality (3.32) follows immediately from (3.20) and (3.26) providedη0is sufficiently

small.

In order to establish (3.30), we start from (3.20), which implies that |Fj| ≤ Hj. This yields

|Y1| ≤ X j Hj Y k>j (1 − Hk), (6.9)

with 0 ≤ Hk≤ 1. Estimate (3.30) is a consequence of (6.9) and of the inequality

X j aj Y k>j (1 − ak) ≤ 1, valid if aj∈ [0, 1], ∀ j ≥ 0.

The proof of (3.31) is similar to the one of (3.30). ä

Proof of Lemma 3.6. In what follows, we use the convention Hm= Km= 0 when m < 0.

In view of (3.19), we have suppKcj⊂ [|n| ≤ 2j− 2]. Using again (3.19), we have supp ƒHjKj⊂ [|n| ≤ 2j+1− 3]. Thus

X j≥0 1(HjKj) = X j≥0 X 0≤m≤j+1 Pd m(1(HjKj)) = X m≥0 X j≥0 Pd m(1(Hm+j−1Km+j−1)). Hence, ° ° ° ° ° X j≥0 1(HjKj) ° ° ° ° ° Fqs,p(Td) ≤X j≥0 ° ° ° ° ° X m≥0 Pd m(1(Hm+j−1Km+j−1)) ° ° ° ° °Fs,p q (Td) ≡X j≥0 Sj. (6.10) Note that Sj= ° ° ° ° ° ° ° ³ 2smPd m(1(Hm+j−1Km+j−1)) ´ m∈N ° ° °lq(N) ° ° ° °Lp(Td) . (6.11)

We are going to estimate the sum in (6.10) by an interpolation technique: we will establish two estimates involving Sj

(estimates (6.13) and (6.20) below) and use the most convenient one according to the values of j. These two estimates will rely on:

° ° ° ° ° ° ° ³ 2(s+1)mHm ´ m∈N ° ° °lq(N) ° ° ° °Lp(Td).ε −1/2 k f kFqs,p. (6.12) Let us prove (6.12): ° ° ° ° ° ° ° ³ 2(s+1)mHm ´ m∈N ° ° °lq(N) ° ° ° °Lp(Td) (a) . ° ° ° ° ° ° ° ³ 2(s+1)mGm ´ m∈N ° ° °lq(N) ° ° ° °Lp(Td) (b) . ° ° ° ° ° ° ° ³ 2(s+1)mk(m)1/2Fm ´ m∈N ° ° °lq(N) ° ° ° °Lp(Td)ε−1/2 ° ° ° ° ° ° ° ³ 2(s+1)mFm ´ m∈N ° ° °lq(N) ° ° ° °Lp(Td)ε −1/2 kFkFs+1,p q (c) .ε−1/2k f kFqs,p. The above estimates are obtained as follows:

(a) is obtained by combining Lemma 4.4 with (3.16) and (3.22). (b) is a consequence of Corollary 5.3 applied to the functionX

m

2(s+1)mFm+j−1.

(c) relies on (3.11).

This ends the proof of (6.12).

i. First estimate ofSj. Starting from (6.11), we obtain successively

Sj (a) . ° ° ° ° ° ° ° ³ 2(s+1)mPd m(Hm+j−1Km+j−1) ´ m∈N ° ° °lq(N) ° ° ° °Lp(Td) (b) . ° ° ° ° ° ° ° ³ 2(s+1)mHm+j−1Km+j−1 ´ m∈N ° ° °lq(N) ° ° ° °Lp(Td) (c) ≤ ° ° ° ° ° ° ° ³ 2(s+1)mHm+j−1 ´ m∈N ° ° °lq(N) ° ° ° °Lp(Td) (d) .ε−1/22−(s+1) jk f kFqs,p. The above estimates are obtained as follows:

(a) follows from Lemma 4.8. (b) is a consequence of Theorem 4.9.

(c) relies on (3.31). (d) relies on (6.12).

We established our first estimate:

Sj.ε−1/22−(s+1) jk f kFqs,p. (6.13) ii. Second estimate ofSj. Using (6.11) and Theorem 4.9, we find that

(15)

We proceed by estimating Tjand Ujseparately.

iii. Estimate ofTj. Let l ∈ N be such thatε= 2−l. Using (3.21) and the definition of Kjas well as (3.31) and (3.18), we find

Tj≤ ° ° ° ° °¡2smK m1Hm ¢ m≤l ° °lq ° ° °Lp(Td)+ ° ° ° ° °¡2smK m1Hm ¢ m>l ° °lq ° ° °Lp(Td) ≤ ° ° ° ° ° ° ° ° ° ° Ã 2sm ¯ ¯ ¯ ¯ ¯ m X k=1 Gk ¯ ¯ ¯ ¯ ¯ 1Hm ! m≤l ° ° ° ° °lq ° ° ° ° °Lp(Td) + ° ° ° ° °¡2sm 1Hm ¢ m>l ° °lq ° ° °Lp(Td) (a) . lk f kFs,pq (Td) ° ° ° ° °¡2sm 1Hm¢ml ° °lq ° ° °Lp(Td)+ ° ° ° ° °¡2sm 1Hm¢m>l ° °lq ° ° °Lp(Td). (6.15)

Inequality (a) follows from the fact that, when m ≤ l, we have ¯ ¯ ¯ ¯ ¯ m−1 X k=1 Gk ¯ ¯ ¯ ¯ ¯ . m−1 X k=1 k fkkFs,pq (Td). lk f kF s,p q (Td); (6.16)

here, we rely on (3.24) and (3.14). For I =J0, l K or I = Jl + 1, ∞J, we have ° ° ° ° °¡2sm 1Hm¢mI°° lq(N) ° ° °Lp(Td) (a) .°° ° ° °¡2sm 1Gm¢mI°° lq(N) ° ° °Lp(Td) (b) . ° ° ° ° ° ° ° ³ 2smk(m)1/2l(m)Fm ´ m∈I ° ° °lq(N) ° ° ° °Lp(Td) . (6.17)

The above estimates are obtained as follows:

(a) follows from Lemma 4.4.

(b) is a consequence of Corollary 5.5 applied toP

m2smFm.

Let us recall that, when I =J0, l K, we have k(m) = 1 and l (m) = 2

m−1

for any m ∈ I. On the other hand, in the case where I =Jl + 1, ∞J, we have k(m) =ε

−1

= 2land l(m) =ε2m−1

= 2m−l−1. By combining this with (6.15), (6.17), (3.11) and with the definition of Fm, we obtain Tj. ³ − lnεk f kFs,pq (Td)+ε 1/2´ k f kFs,pq (Td). (6.18) iv. Estimate ofUj. We start by estimating

|1Km| (a) . X n<m¡k∇F nkL∞(Td)+ k∇HnkL∞(Td)¢ (b) . X n<m ³ 2n(1+d/p)kFnkLp(Td)+ 2nk(n)1−2/pk fnkFs,pq (Td) ´ (c) . X n<m k(n)1−2/p2nk fnkFs,pq (Td) (d) .ε2/p−12mk f kFs,p q (Td). Here are the explanations:

(a) follows from the definition of Kmand (3.32).

(b) is a consequence of Lemma 8.2 and (3.27). (c) relies on (6.2) and (6.3).

(d) is a consequence of the fact that k(m) ≤ε−1

for any m ≥ 1. In view of the definition of Uj, this implies that

Uj.ε2/p−1k f kFs,pq (Td) ° ° ° ° ° ° ° ³ 2(s+1)mHm ´ m∈N ° ° °lq(N) ° ° ° °Lp(Td).ε 2/p−3/2 k f k2Fs,p q (Td) , (6.19)

where the second inequality follows from (6.12). By combining (6.18) and (6.19), we obtain the second estimate of Sj:

Sj. 2−s j

³

(−lnε+ε2/p−3/2) kf kFqs,p(Td)+ε

1/2´

k f kFqs,p(Td). (6.20) v. Optimization. In view of (6.10), we have for everyτ≥ 1

° ° ° ° ° X j≥0 1(HjKj) ° ° ° ° °Fs,p q (Td) ≤X j≥τ Sj+ X j Sj. (6.21)

In the first sum, we use the first estimate of Sj, while in the second sum we use the second estimate. This gives

(16)

withδ:= kf kFs,pq (Td). We now choose l (and thusε): we let l be the integer part of −p lnδ 2(p + 1)ln2. Hence 1 2δ p/(2p−2) <ε= 2−l≤ δp/(2p−2) .

The next step consists in choosing the integerτ. Assume first that s ≥ 0. Then we define andτas the integer part of l s + 1, and an elementary computation gives

° ° ° ° ° X j≥0 1(HjKj) ° ° ° ° °Fs,p q (Td) . k f k1F+s,pp/(4p−4) q (Td) ³ 1 − lnkf kFqs,p(Td) ´ .

Assume next that −1

2< s < 0. We then letτ= l and obtain ° ° ° ° ° X j≥0 1(HjKj) ° ° ° ° ° Fs,pq (Td) . k f k1F+s,p(2s+1)p/(4p−4) q (Td) .

Takingδ0sufficiently small, we can thus ensure that (3.33) holds for any

0 <α< min µ p 4p − 4, (2s + 1)p 4p − 4 ¶ and any kf kFs,pq (Td)≤δ0.

Moreover, the condition k( j)1−2/pk f kFs,p

q (Td)≤η0is satisfied provided that

ε1−2/p

k f kFs,pq (Td)≤η0. (6.23) In turn, (6.23) is satisfied for smallδ0; this follows from the inequality

ε1−2/p k f kFs,pq (Td)≤ k f k (3p−4)/(2p−2) Fqs,p(Td) ≤δ (3p−4)/(2p−2) 0 .

This completes the proof of Lemma 3.6. ä

Proof of Lemma 3.7. We have

k∇Y1kFs,pq (Td)≤ ° ° ° ° ° X j ∇Fj ° ° ° ° ° Fqs,p(Td) + ° ° ° ° ° X j ∇(HjKj) ° ° ° ° ° Fs,pq (Td) := A + B. By Lemma 3.6, we have that

° ° ° ° ° X j 1(HjKj) ° ° ° ° ° Fs,pq (Td) . k f kFqs,p(Td)≤ 1 2d

provided that kf kFs,pq ≤δ0andδ0is chosen sufficiently small. A proof similar to the one of (6.13) combined with the choice ofεleads to ° ° ° ° ° X j m(HjKj) ° ° ° ° °Fs,p q (Td) .ε−1/2k f kFqs,p(Td).δ (3p−4)/(4p−4) for m ∈J2, dK. Hence B ≤ 1 2 for smallδ0. In order to estimate A, we prove that

° ° ° ° ° X j ∇Fj ° ° ° ° °Fs,p q (Td) . k f kFs,pq (Td). (6.24) This is obtained as follows: for m ∈J1, dK, we have

° ° ° ° ° X j mFj ° ° ° ° ° Fqs,p(Td) = ° ° ° ° ° ° à X k 2ksq ¯ ¯ ¯ ¯ ¯ Pd k à X j mFj !¯ ¯ ¯ ¯ ¯ q!1/q°° ° ° ° °Lp(Td) = ° ° ° ° ° à X k 2ksq|mFk|q !1/q° ° ° ° °Lp(Td) . ° ° ° ° ° à X k 2ksq2qk|Fk|q !1/q° ° ° ° °Lp(Td) = kFkFs+1,pq (Td). k f kF s,p q (Td), by Lemma 4.8. This implies (6.24).

Finally, we have k∇Y1kFs,pq (Td)≤ 1, provided k f kFqs,p(Td)≤δ0withδ0sufficiently small. ä

The proof of Proposition 3.1 is complete. ä

7 Inversion of the divergence in domains

Throughout this section,Ω is a smooth bounded domain in Rd. With no loss of generality, we may also assume thatΩ is

connected.

For the definition of the space Fs,pq (Rd) and its basic properties, we refer the reader to [21, Chapter 4]. Let us recall that

Fqs,p(Ω) is the space of restrictions to Ω of elements in F s,p q (Rd).9

9 When s

>0, functions in Fs,pq (Rd) are locally integrable, and we deal with restrictions of functions. When s

(17)

We next give a meaning to the condition Z

Ωf = 0, when f ∈ F s,p

q (Ω). When s > 0, any element f ∈ F s,p

q (Ω) belongs to

Lp(Ω) ⊂ L1(Ω) so that the condition Z

Ωf = 0 has an obvious meaning. When s ≤ 0, we have −s < 1/p

0(since (s + 1)p = d).

Therefore, we are in position to apply [21, Section 4.6.3, Theorem 1] and obtain that the characteristic functionχofΩ is a multiplier of F−q0s,p0(Rd). Equivalently, the extension operator by 0 of maps in F−q0s,p0(Ω), which will be denoted by u 7→ Eu in the sequel, is bounded from Fq−0s,p0(Ω) into F

−s,p0

q0 (Rd). In particular, there exists C > 0 such that for everyϕ∈ F

−s,p0

q0 (Rd) vanishing outsideΩ we have

° °ϕ ° ° F−s,p0q0 (Rd)≤ C ° °ϕ|Ω ° ° F−s,p0q0 (Ω). (7.1)

A consequence of this fact is that

Fqs,p(Ω) = (F− s,p0

q0 (Ω))0. (7.2)

Indeed, whenΩ = Rdthis corresponds to [21, Section 2.1.5]. In order to prove (7.2) for a generalΩ, let f ∈ Fs,pq (Ω). For any

extension ef ∈ Fqs,p(Rd) of f and anyϕ∈ C∞c (Ω), define 〈f ,ϕ〉 := 〈 ef , Eϕ〉. This definition does not depend on the choice of ef ,

and the linear functional thus defined is continuous on C∞

c (Ω) with respect to the F− s0,p0

q0 (Ω) norm. The density of C∞c (Ω)

in Fq−0s,p0(Ω) [21, Theorem 2.4.4.3] allows to extend this linear functional to a continuous linear functional on the whole F−q0s,p0(Ω) space. This proves the inclusion Fqs,p(Ω) ⊂ (F−

s,p0

q0 (Ω))0. Conversely, let g ∈ (F−q0s,p0(Ω))0. Thanks to the boundedness of E, F

−s,p0

q0 (Ω) can be identified with a subspace of F

−s,p0

q0 (R

d).

The Hahn-Banach theorem yields an extension e

g of g as a bounded linear functional on F−q0s,p0(Rd) with the same norm as

g. Thusg ∈ Fe s,pq (Rd) [21, Proposition 2.1.5], which shows that g ∈ F s,p q (Ω).

The claim (7.2) is therefore proved. In addition, a closer look to the above arguments shows that the norms of Fqs,p(Ω) and

(Fq0s,p0(Ω))0are equivalent.

By the above discussion, we may interpret, when s ≤ 0, the condition Z

Ωf = 0 as 〈f ,1〉 = 0, where 〈, 〉 denotes the pairing

between Fqs,p(Ω) and F− s,p0

q0 (Ω).

Proof of Theorem 1.3. The proof follows the one of [5, Theorem 2], with several modifications due to the possibly high regularity of functions in Fqs,p(whereas, in [5, Theorem 2], we only have f ∈ Lp). We first introduce the following notation:

Q0r:= (−r, r)d−1, Qr:= Q0r× (0, r) (r > 0) and k := max(0, [s]). Here, [ ] denotes the integer part. When we deal with traces, we

identify Q0

rwith Q0r× {0}.

i. Extension by reflection. Let f ∈ Fs,pq (Q1). We introduce the invertible Vandermonde matrix A := ((−1/ j)i−1)1≤i, j≤k+1and let

α=t(α1, . . . ,αk+1) := A−1[t(1, . . . , 1)] ∈ Rk+1. (7.3)

If s > 0, we then define, for x = (x0, xd) ∈ Q01× (−1, 1):

˜ f (x0, xd) :=      f (x0, x d), if xd> 0 k+1 X i=1 αif ³ x0, −xd i ´ , if xd< 0 . Then ˜f ∈ Fqs,p(Q01× (−1, 1)) and ° ° ˜f ° °Fs,p q (Q01×(−1,1)). k f kF s,p

q (Q1). Indeed, this is easily checked when s ∈ N and q = 2 (that is, when Fqs,pis the classical Sobolev space Ws,p). This special case combined with interpolation implies that these assertions

still hold when s > 0, p > 1 and q > 1. When s ≤ 0, define ˜f in the following way:

〈 ˜f,ϕ〉 = 〈 f ,ϕe〉, ∀ϕ∈ Fq−0s,p0(Q01× (−1, 1)),

whereϕe(x0, x

d) =ϕ(x0, xd) +ϕ(x0, −xd). Since the map e: F−s,p0

q0 (Q01× (−1, 1)) → F

−s,p0

q0 (Q1),ϕ7→ϕe,

is continuous, it follows that the map

˜: Fqs,p(Q1) → Fqs,p(Q01× (−1, 1)), f 7→ ˜f

is also continuous.

ii. Inversion of the divergence in the case of a flat boundary. This is performed in Lemma 7.2.

Lemma 7.1 Let f ∈ Fqs,p(Q1). Then there exists X ∈ Fsq+1,p(Q2) ∩ L∞(Q2) such that divX = f in Q1and trXd= 0 on Q01.

Moreover, there existsC > 0 (independent of f ) such that kXkFs+1,p

(18)

Proof Let g be a compactly supported Fs,pq extension of ˜f to (−π,π)d. Extending g by periodicity, we may identify g with

an element of Fs,pq (Td). We may further assume that

Z

Tdg = 0 and kgkF s,p

q (Td). k f kFqs,p(Q1). By Theorem 1.1, there exists Y ∈ Fs+1,p

q (Td) ∩ L∞(Td) such that divY = g and

kYkFs+1,p q (Td)+ kYkL∞(T d)≤ C kgkFs,p q (Td). Let B :=            1 2 ··· k + 1 −1 1 α1 1 0 α2 . .. ... 0 ... 1 αk+1            ,

whereα1, ...,αk+1are given by (7.3). In order to prove that B is non singular, we note that

µ 1 0 0 A ¶ B =       1 2 ··· k + 1 −1 1 A ... 1       . (7.4)

Therefore, it suffices to prove that the matrix in the right hand side of (7.4) is non singular. Its determinant equals

Λ = − ¯ ¯ ¯ ¯ ¯ ¯ ¯ ¯ ¯ ¯ p−11p−21· · · p−k+111 p01 p02 · · · p0k +11 .. . ... pk1 pk2 · · · pkk +11 ¯ ¯ ¯ ¯ ¯ ¯ ¯ ¯ ¯ ¯ = − 1 p1. . . pk+1 V (p1, . . . , pk+1, 1),

where pj = −1/ j, 1 ≤ j ≤ k + 1 and V (p1, . . . , pk+1, 1) is the Vandermonde determinant associated with the parameters

{p1, . . . , pk+1, 1}. This implies thatΛ 6= 0 and B is non singular.

We can thus definet(βj)1≤j≤k+2:= B−1[t(0, 1, 0, . . . , 0)]. We then consider for any 1 ≤ i ≤ d − 1,

Xi(x1, . . . , xd) := k+1 X j=1 βjYi µ x0,xd j ¶ +βk+2Yi(x0, −xd), and for i = d, Xd(x1, . . . , xd) := k+1 X j=1 jβjYd µ x0,xd j ¶ −βk+2Yd(x0, −xd).

It is easy to check that divX = f in Q1, that tr Xd= 0 on Q01and that kXkFs+1,p

q (Q2)+ kXkL∞(Q2). k f kF s,p

q (Q1). ä

Lemma 7.2 Let f ∈ Fqs,p(Q1). Then there exists X ∈ Fsq+1,p(Q2) ∩ L∞(Q2) such that divX = f in Q1and trX = 0 on Q01.

Moreover, there existsC > 0 (independent of f ) such that kXkFs+1,p

q (Q2)+ kXkL∞(Q2)≤ C k f kF s,p q (Q1).

Proof By Lemma 7.1, there existsY ∈ Fqs+1,p(Q4) ∩ L∞(Q4) such that divY = f on Q1and trYd= 0 on Q01(with the

cor-responding estimates). For 1 ≤ j ≤ d, we set Cj:= (−1)j+dYj. Observe that trCd= 0 on Q10. We claim that there exists

Dj∈ Fsq+2,p(Q2) ∩ W1,∞(Q2) such that trDj= 0 on Q02and

tr µD j t ¶ = Cjon Q02. Indeed, letEj:= tr|Q0 4Cj∈ F s+1−1/p,p

p (Q04) ∩ L∞(Q04). Fix someρ∈ C∞c (Rd−1) such that

Z Rd−1ρ= 1 and suppρ⊂ Q 0 1. With the notationρt(x0) = 1 td−1ρ µx0 t ¶ , set, for x0∈ Q0 2, Dj(x0, t) := ( tEj∗ρt(x0), if 0 < t < 2 0, if t = 0 . We then have for 1 ≤ k ≤ d − 1

Dj x0 k (x0, t) = Ej∗ Ã ∂ρ x0 k ! t and also Dj t (x 0, t) = E j∗ (ρ+η)t(x0),

(19)

Lemma 7.3 Let L > 0 and R = Rd−1× (0, L). Letρ∈ C∞c (Rd−1) andσ> 1/p. For anyγ∈ Fσ− 1/p,p

p (Rd−1), let

fγ: R → C, R 3 (x0, t) 7→γρt(x0) ∈ C.

Then the mapFσ−p 1/p,p(Rd−1) 3γ7→ fγ∈ Fqσ,p(R) is continuous for any q ∈ [2, p].

Proof of Lemma 7.3. Let K (x0, xd) :=ρxd(x

0). A straightforward computation shows that

b

K (ξ0, xd) =ρb(xdξ

0), where

b

K denotes the d − 1 dimensional Fourier transform with respect to x0= (x1, ..., x

d−1). Sinceρb∈ S (R

d−1), the kernel

b

K belongs to the class S0introduced in [21, Section 3.2.1, Definition 2]. The conclusion of Lemma 7.3 follows therefore from [21, Theorem

3.2.2.3]. ä

Proof of Lemma 7.2 continued. By Lemma 7.3 we have Dj

x0

k

,Dj

t ∈ F

s+1,p

q (Q2), so thatDj∈ Fqs+2,p(Q2). We observe that

tr µD j xk ¶ = 0 if k ≤ d − 1, tr µD j xd ¶ = Ej.

For 1 ≤ j ≤ d, we then define

Zj:= (−1)d+jD j xd−δjd d X i=1 (−1)i+dDi xi ∈ F s+1,p q (Q2) ∩ L∞(Q2).

Finally, letZ := (Z1, . . . ,Zd). The identities divZ = 0 and tr Z = tr Y are straightforward. The vector field X = Y − Z has all

the required properties. ä

iii. Inversion of the divergence in epigraphs. This is achieved in Lemma 7.5. Letψ∈ C∞(Rd−1) and for r > 0,

Ur:= {(x0, xd) ∈ Q0r× R :ψ(x0) < xd<ψ(x0) + 1}.

Lemma 7.4 Let k = [s]. There existsε0> 0 such that ifPki=+12

° °Diψ°

°L∞(Rd−1)≤ε0, then given any f ∈ Fqs,p(U1), we may find

someX ∈ Fsq+1,p(U1) ∩ L∞(U1) satisfying divX = f in U1and trX = 0 on {(x0,ψ(x0)) : x0∈ Q01}. Moreover, we may chooseX such

that kXkFs+1,p q (U1)+ kXkL∞(U1)≤ C k f kF s,p q (U1), withC > 0 independent of f .

Proof As in the proof of [5, Lemma 6], for x0∈ Q0

1and 0 < y < 1, set ˜f(x0, y) = f (x0, y +ψ(x0)). When s ≤ 0, this definition has

to be understood as 〈 ˜f,ϕ〉 := 〈 f ,ϕ◦Ψ−1〉, whereΨ(x0, y) := (x0, y +ψ(x0)). We claim that

° ° ˜f ° °Fs,p q (Q1)≤ c k f kF s,p q (U1). (7.5)

Indeed, when s is an integer (with s ≥ −1) and q = 2, we have F2s,p(Ω) = Ws,p(Ω), and in this case (7.5) is easily checked using

the definition of Sobolev spaces. By complex interpolation [24, Theorem 2.13] one obtains (7.5) for all s ≥ −1 and q = 2. Then real interpolation yields°

° ˜f ° °Bs,p

q (Q1)≤ c k f kB s,p

q (U1)for all s > −1 and all q ∈ R. In particular, (7.5) holds for all s whenever q = p. Since (7.5) also holds when q = 2, complex interpolation yields (7.5) for all s, p and q ∈ [2, p].

By Step 2, there exist C0> 0 and ˜X ∈ Fqs+1,p(Q2) ∩ L∞(Q2) such that div ˜X = ˜f in Q1, tr ˜X = 0 on Q01and

° ° ˜X°°Fs+1,p q (Q2)+ ° ° ˜X°°L∞(Q2)≤ C0 ° ° ˜f ° °Fs,p q (Q1). Now, setX(x0, y) := ˜X(x0, y −ψ(x0)) and write the components of ˜X as ˜X = ( ˜X0, ˜Xd). We obtain

divX(x0, xd) − f (x0, xd) = d−1 X i=1 X˜d x0 i (x0, xd−ψ(x0))∂ψ x0 i (x0).

Clearly, thisX satisfies

kdivX − f kFs,pq (U1)≤ C0 d−1 X i=1 ° ° ° ° ° X˜d x0 i ∂ψ x0 i ° ° ° ° °Fs,p q (Q1) .

By [21, Theorem 4.6.2.2] (this requires s > −1/2), we obtain kdivX − f kFs,pq (U1)≤ C1 ³° °∇ψ ° °L(Rd−1)+ ° °∇ψ ° ° Fs+1,p(Rd−1) ´ k f kFs,pq (U1)≤ C2ε0k f kFqs,p(U1), (7.6) kXkFs+1,pq (U1)+ kXkL∞(U1)≤ C3k f kF s,p q (U1), (7.7)

where the constants Ci, 0 ≤ i ≤ 3, do not depend onψ. If we chooseε0<

1 2C2

, then Lemma 7.5 follows by iterating the above construction ofX (as in the proof of Theorem 1.1) and using (7.6) and (7.7). ä

Références

Documents relatifs

In 1965, the author showed that there is an unique integral representation for the subclass of (nS^) consisting of posi- tive n-harmonic functions with the aid of Radon measures on

Many different methods have been used to estimate the solution of Hallen’s integral equation such as, finite element methods [4], a general method for solving dual integral equation

In this section we first recover the result of (3) on the characterisation of dilation analytic vectors in a very simple way and then give a dilation analyticity

In fact, using our results, Ramalho’s argument for the existence of precisely two positive solutions could

Our main contribution consists in establishing a concentration inequality whose variance matches asymptotically the carr´ e du champ, see (1.4) and Theorem 1 in what we called

We take both a seman- tic and a syntactic approach in order to define a suitable notion of theory dilation in the context of approximate reasoning on the one hand, and a generalized

There are some beautiful intersections between two described types of polynomial Poisson structures: when we are restricting ourselves to the class of quadratic Poisson brackets

In particular, approximating an invariant Gaussian Field with a Kriging model relying on a symmetrized kernel obtained by projection over a fundamental domain may give very