• Aucun résultat trouvé

The canonical injection of the Hardy-Orlicz space $H^\Psi$ into the Bergman-Orlicz space ${\mathfrak B}^\Psi$

N/A
N/A
Protected

Academic year: 2021

Partager "The canonical injection of the Hardy-Orlicz space $H^\Psi$ into the Bergman-Orlicz space ${\mathfrak B}^\Psi$"

Copied!
23
0
0

Texte intégral

(1)

HAL Id: hal-00482846

https://hal.archives-ouvertes.fr/hal-00482846

Preprint submitted on 11 May 2010

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

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

The canonical injection of the Hardy-Orlicz space H Ψ into the Bergman-Orlicz space B Ψ

Pascal Lefèvre, Daniel Li, Hervé Queffélec, Luis Rodriguez-Piazza

To cite this version:

Pascal Lefèvre, Daniel Li, Hervé Queffélec, Luis Rodriguez-Piazza. The canonical injection of the

Hardy-Orlicz space

HΨ

into the Bergman-Orlicz space

BΨ

. 2010. �hal-00482846�

(2)

The canonical injection of the Hardy-Orlicz space H Ψ into the

Bergman-Orlicz space B Ψ

Pascal Lef` evre, Daniel Li,

Herv´ e Queff´ elec, Luis Rodr´ıguez-Piazza

May 11, 2010

Abstract. We study the canonical injection from the Hardy-Orlicz space HΨ into the Bergman-Orlicz spaceBΨ.

Mathematics Subject Classification. Primary: 46E30 – Secondary: 30D55;

30H05; 32A35; 32A36; 42B30

Key-words. absolutely summing operator – Bergman-Orlicz space – compact- ness – Dunford-Pettis operator – Hardy-Orlicz space – weak compactness

1 Introduction and notation

1.1 Introduction

There are two natural Orlicz spaces of analytic functions on the unit disk Dof the complex plane: the Hardy-Orlicz space HΨ and the Bergman-Orlicz spaceBΨ. It is well-known that in the classical case Ψ(x) =xp,Hp⊆Bp and the canonical injection Jp from Hp to Bp is bounded, and even compact. In fact, for any Orlicz function Ψ, one hasHΨ ⊆BΨ and the canonical injection JΨ:HΨ→BΨ is bounded, but we shall see in this paper that its compactness requires that Ψ does not grow too fast. We actually characterize in Section 2 the compactness: JΨ is compact if and only if limx→+∞Ψ(Ax)/[Ψ(x)]2= 0 for everyA >1, and the weak compactness: JΨ is weakly compact if and only if lim supx→+∞Ψ(Ax)/[Ψ(x)]2 <+∞ for every A > 1 . We show that, if these two properties are “often” equivalent (this happens for example if Ψ(x)/x is non-decreasing forxlarge enough), it is not always the case. We actually show a stronger result in Section 4: there is an Orlicz function Ψ such thatJΨ is weakly compact and is Dunford-Pettis, but such thatJΨ is not compact.

1.2 Notation

An Orlicz function is a non-decreasing convex function Ψ : [0,+∞[→[0,+∞[ such that Ψ(0) = 0 and Ψ(∞) = ∞. One says that the Orlicz function Ψ has

(3)

property ∆2 (Ψ∈∆2) if Ψ(2x)≤CΨ(x) for some constantC >0 andxlarge enough. It is equivalent to say that, for every β > 1, Ψ(βx) ≤ CβΨ(x). It is known that if Ψ ∈ ∆2, then Ψ(x) = O(xp) for some 1 ≤ p < +∞. One says (see [6], [7]) that Ψ satisfies the condition ∆0 if, for some β > 1, one has lim

x→∞Ψ(βx)/Ψ(x) = +∞. If Ψ ∈ ∆0, then Ψ(x)/xpx→∞−→ +∞ for every 1≤p <∞. Indeed, let 1≤p <∞. For everyβ >1 one can findx0>0 such that Ψ(βx)/Ψ(x)≥βp forx≥x0; then Ψ(βnx0)≥βnpΨ(x0) for everyn≥1.

That implies that Ψ(x) ≥Cpxp for every x >0 large enough. Since p≥1 is arbitrary, we getxp=o[Ψ(x)].

We say that Ψ∈ ∇0(1) if, for every A >1, Ψ(Ax)/Ψ(x) is non-decreasing for x large enough. This is equivalent to say (see [7], Proposition 4.7) that log Ψ(ex) is convex. When Ψ∈ ∇0(1), one has either Ψ∈∆2, or Ψ∈∆0.

If (S,S, µ) is a finite measure space, one defines the Orlicz space LΨ(µ) as the set of all (classes of) measurable functions f:S →C for which there is a C >0 such thatR

SΨ(|f|/C)dµis finite. The normkfkΨ is the infimum of all C >0 for which the above integral is≤1. The Morse-Transue spaceMΨ(µ) is the subspace off ∈LΨ(µ) for whichR

SΨ(|f|/C)dµis finite for allC >0; it is the closure ofL(µ) inLΨ(µ). One hasMΨ(µ) =LΨ(µ) if and only if Ψ∈∆2.

If Ψ(x)/x −→

x→+∞+∞, the conjugate function Φ of Ψ is defined by Φ(y) = supx>0 xy−Ψ(x)

. It is an Orlicz function and [MΨ(µ)] =LΦ(µ), isomorphi- cally.

We may note that if Ψ(x)/x does not converges to infinity, we must have Ψ(x)≤axfor somea≥1 andxlarge enough. ThenLΨ(µ) =L1(µ) isomorphi- cally and then Φ(y) = +∞for y > a (givingLΦ(µ) = L(µ) isomorphically).

We denote byDthe open unit disk ofCand byT=∂Dthe unit circle. The normalized area-measure on D is denoted by A and the normalized Lebesgue measure onTis denoted bym.

The Hardy-Orlicz space HΨ is defined as {f ∈ H1; f ∈ LΨ(m)}, where f is the boundary values function of f, and HMΨ = HΨ ∩MΨ(m) is the closure ofHinHΨ. The Bergman-Orlicz spaceBΨis the subspace of analytic f ∈LΨ(A), andBMΨ =BΨ∩MΨ(A) is the closure ofHin BΨ. Since, for f ∈ HΨ, kfkHΨ = sup0<r<1kfrkHΨ (see [7], Proposition 3.1), where fr(z) = f(rz), one has:

Z 0

Ψ

|f(reit)| kfkHΨ

dt 2π ≤

Z 0

Ψ

|f(reit)| kfrkHΨ

dt 2π ≤1 ; hence:

Z

D

Ψ

|f(reit)| kfkHΨ

dA=

Z 1 0

Z 0

Ψ

|f(reit)| kfkHΨ

dt 2π

2r dr≤1,

sof ∈BΨ andkfkBΨ ≤ kfkHΨ. It follows thatHΨ ⊆BΨ and the canonical injection JΨ:HΨ → BΨ is bounded, and has norm 1. Let us point out that

(4)

the boundedness also follows from [7], Theorem 4.10, 2), sinceJΨ is a Carleson embeddingJΨ:HΨ→BΨ⊆LΨ(A).

This injection is not onto, since there are functions f ∈BΨ with no radial limit on a subset ofTof positive measure (the proof is the same as in Bp: see [4],§3.2, Lemma 2, page 81). Note that JΨ is not an into-isomorphism: take fn(z) =zn, for everyn∈N; it is easy to see that{fn}n tends to 0 inBΨ, but not inHΨ.

Acknowledgment. This work is partially supported by a Spanish research project MTM 2009-08934. Part of this paper was made during an invitation of the second-named author by the Departamento de An´alisis Matem´atico of the Universidad de Sevilla. It is a pleasure to thanks the members of this department for their warm hospitality.

2 Compactness and weak-compactness

In order to characterize the compactness and the weak-compactness ofJΨ, we introduce the following quantityQA,A >1:

(2.1) QA= lim sup

x→+∞

Ψ(Ax) [Ψ(x)]2, which will turn out to be essential.

We are going to start with the compactness.

Theorem 2.1 The canonical injection JΨ: HΨ → BΨ is compact if and only if

(2.2) lim

x→+∞

Ψ(Ax)

[Ψ(x)]2 = 0 for every A >1.

Remarks. 1) Condition (2.2) means that QA = 0 for every A > 1. It is equivalent to say that:

(2.3) sup

A>1

QA<+∞.

Indeed, assume thatM := supA>1QA <+∞. Let 0 < ε≤1 and A > 1; we can find xA = xA(ε) > 0 such that Ψ(Ax/ε)/[Ψ(x)]2 ≤ 2M for x ≥xA. By convexity, one has Ψ(Ax)≤ εΨ(Ax/ε), and hence Ψ(Ax)/[Ψ(x)]2 ≤2εM for x≥xA. We getQA= 0.

2) It is clear that condition (2.2) is satisfied whenever Ψ∈∆2, but Ψ(x) = e[log(x+1)]2 −1 satisfies (2.2) without being in ∆2. However, condition (2.2) implies that Ψ cannot grow too fast. More precisely, we must have

Ψ(x) =o(exα) for everyα >0.

(5)

Indeed, one has Ψ(At) ≤ [Ψ(t)]2 for t ≥ tA, and, by iteration, Ψ(AntA) ≤ [Ψ(tA)]2n for every n ≥1. For every x > 0 large enough, taking n≥ 1 such thatAntA≤x < An+1tA, we get Ψ(x)≤C1eC2xα, withα= log 2/logA. Since A > 1 is arbitrary, α may be any positive number. The little-oh condition follows from the fact that the inequality is true for allα >0.

Proof of Theorem 2.1. By definition, BΨ is a subspace of LΨ(D,A); hence we can seeJΨas a Carleson embeddingJΨ:HΨ→LΨ(D,A). IfS(ξ, h) ={z∈ D; |z−ξ|< h}, the compactness ofJΨ implies, by [7], Theorem 4.11, that, for everyA >1, everyε >0, andh >0 small enough:

h2≤4A[S(ξ, h)]≤ 4ε Ψ[AΨ−1(1/h)],

that is, settingx= Ψ−1(1/h), Ψ(Ax)≤4ε[Ψ(x)]2, and (2.2) is satisfied.

Conversely, one has:

sup

0<t≤h

sup

|ξ|=1

A[S(ξ, t)]

t ≤ sup

0<t≤h

t2 t =h , which iso (1/h)/Ψ[AΨ−1(1/h)]

for everyA >1, if (2.2) holds; hence, by [7],

Theorem 4.11, again,JΨ is compact.

We now turn ourself to the weak compactness.

Theorem 2.2 The following assertions are equivalent:

(a) JΨ:HΨ→BΨ is weakly compact;

(b)JΨ fixes no copy of c0; (c) JΨ fixes no copy of; (d)QA<+∞, for every A >1;

(e) HΨ⊆BMΨ;

(f) JΨ is strictly singular.

Recall that an operatorT:X →Y between two Banach spaces is said to be strictly singular if there is no infinite-dimensional subspaceX0 of X on which T is an into-isomorphism.

The proof will be somewhat long, and before beginning it, we shall remark that if Ψ∈∆0, then condition

(2.4) QA<+∞ for everyA >1 implies condition (2.2). Indeed, if lim

x→+∞

Ψ(βx)

Ψ(x) = +∞, we get, for everyA >1:

lim sup

x→+∞

Ψ(Ax)

[Ψ(x)]2 = lim sup

x→+∞

Ψ(Ax) Ψ(βAx)

Ψ(βAx)

[Ψ(x)]2 ≤lim sup

x→+∞

Ψ(Ax)

Ψ(βAx)QβA= 0. Now, if, for someA > 1, Ψ(Ax)/Ψ(x) is non-decreasing forxlarge enough (in particular if Ψ∈ ∇0(1)), one has the dichotomy: either Ψ∈∆2, and then JΨ

is compact; or Ψ∈∆0 and hence the weak compactness ofJΨ implies, by the two above theorems, its compactness. Hence:

(6)

Proposition 2.3 If, for some A > 1, Ψ(Ax)/Ψ(x) is non-decreasing, for x large enough, then the weak compactness ofJΨ is equivalent to its compactness.

However, it is easy to construct an Orlicz function Ψ which satisfies condi- tion (2.4), but not condition (2.2). We do not give an axample here because we have a stronger result in Section 4.

In order to prove Theorem 2.2, we shall need several lemmas.

Lemma 2.4 Let Ψbe any Orlicz function. If we defineΨ1(t) = [Ψ(t)]2,t≥0, thenΨ1is an Orlicz function for which HΨ⊆BΨ1 and the canonical injection ofHΨ intoBΨ1 is continuous.

Proof. It is enough to see that HΨ continuously embeds intoLΨ1(A), and for this we can use Theorem 4.10 in [7]. Following the notation of that theorem for the measure µ = A, it is easy to see that, as h → 0+, ρA(h) ≈ h2, and KA(h) ≈ h. Observe that, for t > 1, we have Ψ1−1(t)] = t2, and so, for h∈(0,1),

1/h

Ψ1−1(1/h)] = 1/h

1/h2 =hKA(h).

Using part 2) of Theorem 4.10 in [7], the lemma follows.

Lemma 2.5 LetM > δ >0 and{fn}nbe a sequence inHΨ∩BMΨ such that:

(a){fn}n tends to 0uniformly on compact subsets of D; (b)kfnkBΨ ≥δ, for everyn≥1;

(c) kfnkHΨ≤M, for everyn≥1.

Then there exists a subsequence {fnk}k such that P

k|fnk(z)| < +∞, for every z ∈ D, and for every α = (αk)k ∈ ℓ, one has, writing T α(z) = P

k=1αkfnk(z):

(2.5) T α∈BΨ and (δ/2)kαk≤ kT αkBΨ ≤2Mkαk.

Remark. It is clear that, by (2.5), we are defining an operatorT from ℓ

intoBΨ which is an isomorphism betweenℓand its image. In particular, the subsequence{fnk}k is equivalent, inBΨ, to the canonical basis of c0.

Proof. First we are going to construct, inductively, a subsequence {fnk}k of {fn}, and an increasing sequence{rk}k in (0,1), such that limk→∞rk = 1 and, setting

Dk={z∈D; |z| ≤rk}, fork≥1, and

C1=D1, Ck =Dk\Dk−1={z∈D; rk−1<|z| ≤rk}, k≥2, we have:

(2.6) |fnk(z)| ≤2−k, for everyz∈Dk−1, and everyk≥2 ;

(7)

and

(2.7) kfnk1ID\CkkLΨ< δ2−k−2, for everyk≥1.

Start the construction by taking n1= 1. It is a known fact that, for every functionf in the Morse-Transue spaceMΨ(A), we have

(2.8) lim

A(A)→0kf1IAkLΨ= 0.

Now, using (2.8), with f = fn1 and considering sets A of the form A= {z ∈ D; r <|z|<1}, we getr1∈(0,1) so that, for C1 =D1={z ∈D; |z| ≤r1}, we have

kf11ID\C1kLΨ< δ2−3.

By the uniform convergence of{fn}nto 0 onD1, we can findn2> n1such that

|fn2(z)| ≤1/4, for everyz∈D1, and kfn21ID1kLΨ< δ2−5. Using this last inequality and (2.8) again (for f = fn2), we get r2 ∈ (r1,1), r2>1−1/2, such that, settingC2={z∈D; r1<|z| ≤r2}, we have

kfn21ID\C2kLΨ< δ2−4.

Now that we have (2.6) and (2.7) fork= 1 andk= 2, it is clear how we must iterate the inductive construction. At the time of choosing rk ∈ (rk−1,1), we also impose the conditionrk>1−1/kin order to get limk→∞rk = 1.

Once the construction is achieved, let us see why the subsequence {fnk}k works. The condition (2.6) and the fact that limk→∞rk = 1 imply that, for every compact setK inDandz∈D, there existslK ∈Nsuch that:

|fnk(z)| ≤2−k, for everyz∈K, and everyk≥lK. This yields two facts. First,P

k|fnk(z)|<+∞, for everyz∈D, and secondly:

for every bounded complex sequence α = (αk)k ∈ ℓ, the series P

kαkfnk

converges uniformly on compact subsets ofD, and its sum, the functionT α, is analytic onD.

It remains to prove the estimates in (2.5) about the norm ofT αinLΨ(A).

By homogeneity, we may assume thatkαk= 1. Let us writegk =fnk1ICkand hk =fnk1ID\Ck, for everyk≥1,

g= X

k=1

αkgk and h= X

k=1

αkhk.

We have T α = g+h. By (2.7) and the fact that |αk| ≤ 1, we have that h∈LΨ(A) andkhkLΨ≤δ/4.

By the condition (c) in the statement and the definition of the norm inHΨ we have, for everynand everyr∈(0,1):

(2.9) 1

2π Z

0

Ψ |fn(reit)|/M dt≤1.

(8)

The functiongk is 0 outside ofCk, and the sequence{Ck}k is a partition ofD. Therefore:

Z

D

Ψ(|g|/M)dA= X k=1

Z

Ck

Ψ(|g|/M)dA= X k=1

Z

Ck

Ψ(|αk| |fnk|/M)dA

≤ X

k=1

Z

Ck

Ψ(|fnk|/M)dA.

Integrating in polar coordinates, settingr0= 0, and using (2.9), we get:

Z

D

Ψ(|g|/M)dA ≤ X

k=1

Z rk

rk−1

2r 1 2π

Z 0

Ψ(|fnk(reit)|/M)dt dr

≤ X

k=1

Z rk

rk−1

2r dr= 1,

and thereforekgkLΨ≤M, andkT αkLΨ≤δ/4 +M ≤2M. On the other hand, for everyk, we have:

kgkLΨ≥ kg1ICkkLΨ=|αk|kfnk−hkkLΨ≥ |αk|(δ−δ/22+k)≥ 3δ 4|αk|. Taking the supremum onk, we getkgkLΨ≥(3δ/4)kαk= 3δ/4. Consequently,

kT αkLΨ≥ kgkLΨ− khkLΨ≥(3δ/4)−δ/4≥δ/2,

and Lemma 2.5 is fully proved.

In the following lemma we isolate the proof of the implication (c) =⇒ (d) in the statement of Theorem 2.2.

Lemma 2.6 Assume that the Orlicz functionΨ is such that, for someA >1,

(2.10) lim sup

x→+∞

Ψ(Ax) [Ψ(x)]2 = +∞ Then the injectionJΨ:HΨ→BΨ fixes a copy of.

Proof.Let us take a sequence of positive numbers{dn}n, and a sequence{ξn}n

inT, such that the disks{D(ξn, dn)}n are pairwise disjoint inD. In particular, we should have limn→∞dn = 0.

The convexity of Ψ implies the existence of somec >0 such that Ψ(x)≥cx for everyx≥1. Given a sequence {βn}n in (4,+∞) to be fixed later, we can find, thanks to (2.10), an increasing sequence{xn}satisfying:

(2.11) xn>1, Ψ(xn)>1, Ψ(Axn)> βn[Ψ(xn)]2, for everyn∈N.

(9)

Defineyn as the point in the interval (xn, Axn) such that

(2.12) [Ψ(yn)]2= Ψ(Axn).

Put now hn = 1/Ψ(yn) and rn = 1−hn. By (2.11) and (2.12), we have [Ψ(yn)]2> βn>4, and thereforehn∈(0,1/2). Define

un(z) = hn

1−rnξnz 2

, and fn(z) =ynun(z). It is easy to see thatkunk= 1, and thatkunkH1 ≤hn.

The first condition imposed to βn is βn > 16/d2n. That gives [Ψ(yn)]2 >

16/d2n andhn < dn/4. Let us write Dn for the disk D(ξn, dn). Observe that, forz∈D\Dn, we have

|1−rnξnz|=|1−rn+rnξnξn−rnξnz| ≥rnn−z|−hn ≥(1/2)dn−hn≥dn/4, and therefore, since [Ψ(xn)]2≥Ψ(xn)≥c xn,

|fn(z)| ≤yn

4hn

dn

2

= 16yn

d2n[Ψ(yn)]2 ≤ 16Axn

d2nβn[Ψ(xn)]2 ≤ 16A c d2nβn ·

We also impose the conditionβn >16An2/cd2n, and so we have:

(2.13) |fn(z)| ≤ 1

n2, forz∈D\Dn.

From (2.13) we deduce that {fn}n converges to 0 uniformly on compact subsets ofD. Moreover (2.13) yields that, for every bounded sequence{αn}n of complex numbers, the seriesP

n≥1αnfn is uniformly convergent on compact subsets ofD. Let us writefnfor the boundary value (onT=∂D) of the function fn. We claim that :

(2.14) S=

X n=1

|fn| ∈LΨ(T, m).

From this, it is not difficult to deduce that, for every bounded sequence{αn}n

of complex numbers, the functionP

n=1αnfnis inHΨand, forM =kSkLΨ(T),

(2.15)

X n=1

αnfn

HΨ≤Mk{αn}nk.

On the other hand, taking An = {z ∈ D; |z−ξn| ≤ hn}, there exists a constantγ∈(0,1) such thatA(An)≥γh2n, and, for everyz∈An, we have:

|1−rnξnz| ≤ |1−rn|+|rnξnξn−rnξnz|=hn+rn|z−ξn| ≤2hn,

(10)

and consequently|un(z)| ≥1/4. If δ=γ/4A, we have, for everyn, Z

D

Ψ|fn| δ

dA ≥ Z

An

Ψyn

dA ≥γh2nΨ1 γAyn

≥h2nΨ(Ayn)> h2nΨ(Axn) = 1.

ThuskfnkBΨ ≥δ, for everyn∈N. We can apply Lemma 2.5. Using this lemma and (2.15), we get a subsequence{fnk}k such that, for everyα= (αk)k ∈ℓ, we have:

(δ/2)k{αk}kk≤ X

k=1

αkfnk

BΨ

X

k=1

αkfnk

HΨ ≤Mk{αk}kk. This clearly says thatJΨ fixes a copy ofℓ.

It remains to prove (2.14). For obtaining this we impose the last condition to the sequence{βn}n. We shall need:

(2.16)

X n=1

1/p βn ≤1.

Let us set gn = |fn|1IDn. Thanks to (2.13), S −P

n=1gn is a bounded function. Thus we just need to prove thatG=P

n=1gn is inLΨ(T). We have kGkLΨ(T)≤A. Indeed, recalling that the Dn’s are pairwise disjoint, and that eachgn is 0 out ofDn, we have:

Z

T

ΨG A

dm= X n=1

Z

DnT

ΨG A

dm= X n=1

Z

DnT

Ψ|fn| A

dm

≤ X

n=1

Z

T

Ψyn|un| A

dm

and by the convexity of Ψ, and the fact that|un| ≤1,

≤ X n=1

Z

T

|un|Ψyn

A

dm= X n=1

kunkH1Ψyn

A

≤ X n=1

Ψ(yn/A) Ψ(yn) ≤

X n=1

Ψ(xn) Ψ(yn) =

X n=1

Ψ(xn) pΨ(Axn) ≤

X n=1

√1

βn ≤1, by the required condition (2.16), and that ends the proof of Lemma 2.6.

We are now in position to prove Theorem 2.2.

Proof of Theorem 2.2. We shall prove that:

(a) =⇒ (b) =⇒ (c) =⇒ (d) =⇒ (e) =⇒ (a), and that (b)⇐⇒(f).

(11)

The implications (a) =⇒(b) =⇒(c) and (f) =⇒(b) are trivial, and we have seen in Lemma 2.6 that (c) =⇒(d).

(d) =⇒ (e). By Lemma 2.4, there exists a constant C > 0 such that, for everyf in the unit ball ofHΨ, we have:

(2.17)

Z

D

[Ψ(|f|/C)]2dA ≤1.

For everyA >0, there existxA, such that Ψ(Ax)≤(QA+ 1)[Ψ(x)]2, for every x≥xA. Thus for every x≥0 we have Ψ(Ax)≤(QA+ 1)[Ψ(x)]2+ Ψ(AxA).

Then, by (2.17), we have Z

D

Ψ(A|f|/C)dA<+∞, for everyA >0.

Thereforef ∈BMΨ, for everyf in the unit ball of HΨ, and thus for everyf inHΨ.

(e) =⇒ (a). Let {fn}n be in the unit ball of HΨ. We have to prove that {fn}n has a subsequence which converges in the weak topology of BΨ. By Montel’s Theorem{fn}n has a subsequence converging uniformly on compact subsets ofD, to a functiongwhich, by Fatou’s lemma, also belongs to the unit ball ofHΨ. If this subsequence converges togin the norm ofBΨ we are done.

If not, after perhaps a new extraction of subsequence, there existδ >0 and a subsequence{fnk}k, such that

kfnk−gkBΨ≥δ, and kfnk−gkHΨ ≤2.

Since moreover{fnk−g}k converges to 0 uniformly on compact subsets of D and, by condition (e),fnk−g ∈BMΨ, we may apply Lemma 2.5 and we get that {fnk−g}k has a subsequence equivalent to the canonical basis of c0 in BΨ, and is therefore weakly null. This yields that {fn}n has a subsequence converging tog in the weak topology ofBΨ.

(b) =⇒(f). Suppose there exists an infinite-dimensional subspaceX ofHΨ on which the normsk · kBΨ andk · kHΨ are equivalent. We shall have finished if we prove thatX contains a subspace isomorphic toc0 because thenJΨ will fix a copy ofc0.

We can assume thatX is contained inBMΨ because we already know that (b) implies (e). X being infinite-dimensional, there exists, for every n ∈ N, fn ∈ X, such that kfnkHΨ = 1, and cfn(k) = 0, for k = 0,1, . . . , n. By the equivalence of the norms in X, there exists δ > 0 such thatkfnkBΨ ≥ δ, for everyn. The unit ball ofHΨ is compact in the topology ofH(D). Since

n→∞lim cfn(k) = 0, for everyk≥0,

the only possible limit of a subsequence of {fn}n is the function 0. So {fn}n converges to 0 uniformly on compact subsets of D. As fn ∈ X ⊆ BMΨ, for

(12)

everyn, we can apply Lemma 2.5, and we get that {fn}n has a subsequence generating an spaceY isomorphic toc0inBΨ. This spaceY is contained inX, where the norms are equivalent, soY is also isomorphic toc0 for the norm of

HΨ.

3 Other properties

3.1 Dunford-Pettis

Recall that an operator T: X →Y between two Banach spaces X and Y is said to beDunford-Pettis if{T xn}n converges in norm whenever{xn}n con- verges weakly. Every compact operator is Dunford-Pettis. The next proposition shows that, in “most” of the cases, these two properties are equivalent forJΨ. Proposition 3.1 If the conjugate function of Ψ satisfies condition2, then JΨ:HΨ→BΨ is Dunford-Pettis if and only if it is compact.

We shall see in Section 4 that without condition ∆2for the conjugate func- tion,Jψ may be Dunford-Pettis without being compact.

Proof. Remark first that speaking of the conjugate function of Ψ implicitly assume that Ψ(x)/x tends to +∞as xgoes to +∞.

Assume that JΨ is not compact. By Theorem 2.1, there are some A > 1 and a sequence{xj}j going to +∞such that Ψ(Axj)≥[Ψ(xj)]2. Settingrj = 1−1/Ψ(xj), this is equivalent to say thatAΨ−1 1/(1−rj)

≥Ψ−1 1/(1−rj)2 . Define:

fj(z) =xj

1−rj

1−rjz 2

·

One has fj ∈ HMΨ and kfjkHΨ ≤ 1 (see [7], Corollary 3.10). Since {fj}j

converges to 0 uniformly on compact subsets ofD, {fj}j converges to 0 in the weak-star topology ofHΨ ([7], Proposition 3.7). But, since the conjugate func- tion of Ψ satisfies condition ∆2,HΨ is the bidual of HMΨ([7], Corollary 3.3);

hence{fj}j converges weakly to 0 in HMΨ.

On the other hand, ifSj =D(1,1−rj)∩D, one has|1−rjz| ≤2(1−rj) forz∈Sj; hence, writingK=kfjkBΨ, one has:

1 = Z

D

Ψ(|fj|/K)dA ≥ Z

Sj

Ψ(|fj|/K)dA ≥ A(Sj)Ψ(xj/4K).

Since A(Sj) ≥ α(1 −rj)2, with 0 < α < 1, we get (since Ψ(αxj/4K) ≤ αΨ(xj/4K), by convexity):

kfjkBΨ≥(α/4) xj

Ψ−1 1/(1−rj)2 = (α/4) Ψ−1 1/(1−rj) Ψ−1 1/(1−rj)2 ≥ α

4A·

ThereforeJΨ is not Dunford-Pettis.

On the other hand, one has:

(13)

Proposition 3.2 If JΨ is Dunford-Pettis, thenJΨ is weakly compact.

Proof. By Theorem 2.2, if JΨ is not weakly compact, there is a subspaceX0

ofHΨ isomorphic to c0 on whichJΨ is an into-isomorphism; hence JΨ cannot

be Dunford-Pettis.

We shall see in the next section that JΨ may be weakly compact without being Dunford-Pettis.

3.2 Absolutely summing

Everyp-summing operator is weakly compact and Dunford-Pettis; so it may be expected thatJΨ isp-summing for somep <∞. The next results show that this is never the case as soon as Ψ grows faster than all the power functions.

Recall that an operatorT:X →Y between two Banach spacesX andY is called (p, q)-summing if there is a constantC >0 such that

Xn

k=1

kT xkkp1/p

≤C sup

kxkX∗≤1

Xn

k=1

|x(xk)|q1/q

,

for every finite sequence (x1, . . . , xn) inX. Ifq=p, it is saidp-summing. Every p-summing operator is (p, q)-summing forq≤p.

Theorem 3.3 If JΨ:HΨ→BΨ isp-summing, then, for everyq > p,Ψ(x) = O(xq)for xlarge enough. Moreover, ifp <2, thenJΨ is compact.

In order to prove this, we need two lemmas.

Lemma 3.4 If the canonical injection IΨ:A→ BΨ is (p,1)-summing, where A=A(D)is the disk algebra, thenΨ(x) =O(x2p)for xlarge enough.

In particular, Jr:Hr →Br is (p,1)-summing for no p < r/2, and, if Ψ∈

0, thenJΨ is(p,1)-summing for nop <∞.

Recall that the disk algebra is the space of continuous functions onDwhich are analytic inD.

We refer to [9] for a detailed study ofr-summing Carleson embeddingsHr→ Lr(µ). In particular, it follows from these results that Jr: Hr → Br is 1- summing for 1≤r <2. On the other hand, it is easy to see thatJ2: H2→B2 is not Hilbert-Schmidt (i.e. not 2-summing): for the canonical orthonormal basis {zn}n and {√

n+ 1zn}n of H2 and B2, J2 is the diagonal operator of multiplication by{1/√

n+ 1}n. It also follows from [9] that, for r ≥2, Jr is p-summing for no finitep.

Proof. Assume that we do not have Ψ(x) =O(x2p) forxlarge enough. Then lim supx→+∞Ψ(x)/x2p = +∞. Given any K > 0, take y > 0 such that Ψ(y)/y2p ≥ K and such that h = 1/p

Ψ(y) ≤ 1/2. Let N be the integer part of (1/h) + 1. Writingξj = e2πij/N, we set:

uj(z) = h2

[1−(1−h)ξjz]2·

(14)

We haveuj ∈A(D). By [7], Lemma 5.6, one has, sinceh≥1/N:

NX−1

j=0

|uj(eit)| ≤N h2 1−(1−h)2N

[1−(1−h)2][1−(1−h)N]2 ≤ e2

(1−e)2 :=C . Hence:

sup

kxkA≤1 NX−1

j=0

|x(uj)| ≤C .

On the other hand, it is easy to see that|uj(z)| ≥1/9 when|z−(1−h)ξj|< h;

hence, ifSj ={z∈D; |z−(1−h)ξj|< h}, one has, sinceA(Sj) =h2: 1 =

Z

D

Ψ |uj(z)| kujkBΨ

dA(z)≥ Z

Sj

Ψ 1/9 kujkBΨ

dA ≥h2Ψ 1/9 kujkBΨ

,

sokujkBΨ≥1/9Ψ−1(1/h2). Sincey= Ψ−1(1/h2), one gets:

N−1X

j=0

kujkpBΨ≥(1/9)pN

yp ≥(1/9)p Ψ(y)

y2p 1/2

≥ K1/2 9p ·

This yields that the (p,1)-summing norm of IΨ should be greater than K1/2p/9C, and, asKis arbitrary, thatIΨ is not (p,1)-summing.

Remark. WhenIΨ: A ֒→BΨ is p-summing, we have this shorter argument.

By Pietsch’s factorization theorem, thisIΨ factors throughHp. It follows from [7], Theorem 4.10, that α h2 ≤ ρA(h) ≤ 1/Ψ−1(A/h1/p), for some constants 0< α <1 andA >0, andhsmall enough. That means that Ψ(x)≤C x2p for xlarge enough.

Lemma 3.5 If the canonical injection IΨ:A→BΨ is1-summing, thenJΨ is compact.

Proof. The canonical injectionJ1:H1→B1(as well asJΨwhenever Ψ∈∆2) is compact. Hence we may assume thatHΨ is not H1 and hence that Ψ(x)/x tends to +∞as xtends to +∞.

Assume thatJΨ is not compact. Then, as in the proof of Proposition 3.1, there are someA >1 and a sequence{xk}k going to +∞such that Ψ(Axk)≥ [Ψ(xk)]2. Settinghk= 1/Ψ(xk), we define, as in the proof of Proposition 3.4:

uk,j(z) = h2k

[1−(1−hkk,jz]2,

whereξk,j = e2πij/Nk, withNkthe integer part of (1/hk) + 1. One hasuk,j ∈A and (see the proofs of the two quoted propositions):

NXk−1 j=0

|uk,j(eit)| ≤C and kuk,jkBΨ≥ δα A

1 Ψ−1(1/hk

(15)

It follows that:

Nk−1

X

j=0

kuk,jkBΨ≥δα A

Nk

Ψ−1(1/hk)≥ δα A

1/hk

Ψ−1(1/hk) = δα A

Ψ(xk) xk −→

k→∞+∞.

HenceIΨ is not 1-summing.

Proof of Theorem 3.3. Since JΨ: HΨ →BΨ is p-summing and the canon- ical injection IΨ: A → BΨ factors as IΨ:A → HΨ → BΨ, this injection is p-summing. By Lemma 3.4, Ψ(x) = O(x2p) for x large enough. Hence we have the factorizationA →H2p →HΨ →BΨ. Since the first injection is 2p- summing and the last one isp-summing, the composition is max(1, p1)-summing, with p11 = 2p1 +1p (see [2], Theorem 2.22),i. e. p1= 23p. If p1>1, we can use again Lemma 3.4 withp1instead of 2p; we get that Ψ(x) =O(x2p1), forxlarge enough, and that the factorizationIΨ:A→H2p1 →HΨ →BΨ is max(1, p2)- summing, with p1

2 = 2p1

1 + 1p. Going on the same way, we get a decreas- ing sequence {pn}n such that the canonical injection A → BΨ is max(1, pn)- summing and p1

n+1 =2p1

n +1p·Writingpnnp, we getαn+1= n

n+1; hence pnn→∞−→ p/2. In particular, Ψ(x) =O(xq) for everyq > p.

Ifp <2, one has max(1, pn) = 1 fornlarge enough, and Lemma 3.4 implies

thatJΨ is compact.

Remark 1. It is not clear whetherJΨ p-summing, withp≥2, implies thatJΨ

is compact. However, whenr ≥2,Jr:Hr →Br isp-summing for nop <∞ (see [9]).

Remark 2. An operator T:X →Y between two Banach spaces is said to be finitely strictly singular (orsuperstrictly singular) if for everyε >0, there is an integerNε≥1 such that, for every subspaceX0 ofX of dimension≥Nε, there is anx∈X0such thatkT xk ≤εkxk. Every finitely strictly singular operator is strictly singular. It is not difficult to see that every compact operator is finitely strictly singular and it is shown in [10] (see also [5], Corollary 2.3) that every p-summing operator is finitely strictly singular. We do not know when JΨ is finitely strictly singular.

3.3 Order boundedness

Recall that an operator T: X →Y from a Banach spaceX into a Banach latticeY is said to beorder bounded if there isy ∈Y+ such that|T x| ≤y for everyxin the unit ball ofX. Since the Bergman-Orlicz spaceBΨis a subspace of the Banach lattice LΨ(D,A), we may study the order boundedness of JΨ. Actually, we are going to see that the natural space for the order boundedness ofJΨ is notLΨ(D,A), but theweak Orlicz space LΨ,∞(D,A), the definition of which we are recalling below (see [7], Definition 3.16).

(16)

Definition 3.6 Let (S,S, µ) be a measure space; the weak-LΨ space LΨ,∞ is the set of the (classes of ) measurable functions f: S →C such that, for some constantc >0, one has, for every t >0:

µ(|f|> t)≤ 1 Ψ(ct)·

One has LΨ ⊆ LΨ,∞ and ([7], Proposition 3.18) the equality LΨ = LΨ,∞

implies that Ψ ∈ ∆0. On the other hand, this equality holds when Ψ grows sufficiently; for example, if Ψ satisfies the condition ∆1: xΨ(x) ≤ Ψ(αx), for some constantα >1 andxlarge enough.

Proposition 3.7 JΨ:HΨ→BΨ is always order bounded intoLΨ,∞(D,A).

Proof. Since (see [7], Lemma 3.11):

(3.1) 1

−1 1 1− |z|

≤ sup

kfkHΨ≤1|f(z)| ≤4Ψ−1 1 1− |z|

,

one has, denoting byS(z) the supremum in (3.1), fortlarge enough:

A(|S|> t)≤ A {z∈D; |z|>1−1/Ψ(t/4)}

≤ 2

Ψ(t/4) ≤ 1 Ψ(t/8),

and the result follows.

Since we also have, fort large enough:

A(|S|> t)≥ A {z∈D; |z|>1−1/Ψ(4t)}

≥ 1 Ψ(4t), we get:

Corollary 3.8 JΨ is order bounded into LΨ(D,A)if and only if LΨ =LΨ,∞. This is the case if Ψ∈∆1.

Remark. Contrary to the compactness, or the weak compactness, which re- quires that Ψ does not grow too fast, the order boundedness ofJΨintoLΨ(D,A) holds when Ψ grows fast enough. Nevertheless, for Ψ(x) = exp[ log(x+1)2

]−1, JΨ is compact and order bounded intoLΨ(D,A).

WhenJΨis weakly compact,JΨmapsHΨintoBMΨ(Theorem 2.2); hence, we may ask whether JΨ may be order bounded into MΨ(D,A); however, we have:

Proposition 3.9 JΨ is never order bounded intoMΨ(D,A).

Proof. If it were the case, we should haveS∈MΨ(D,A), and hence Z

D

Ψ

4×1

−1 1 1− |z|

dA(z)<+∞,

which is false.

(17)

4 An example

Theorem 4.1 There exists an Orlicz functionΨsuch thatJΨis weakly compact and Dunford-Pettis, but which is not compact.

Note that such an Orlicz function is very irregular: Ψ ∈/ ∆2, Ψ∈/ ∆0, so, for everyA >1, Ψ(Ax)/Ψ(x) is not non-decreasing forxlarge enough, and the conjugate function of Ψ does not satisfies condition ∆2.

The following lemma is undoubtedly well-known, but we have found no ref- erence, so we shall give a proof. Recall that a sublatticeX of L0(µ) is solid if

|f| ≤ |g|andg∈X impliesf ∈X andkfk ≤ kgk.

Lemma 4.2 Let (S,S, µ) be a measure space, and let X be a solid Banach sublattice ofL0(µ), the space of all measurable functions. Then, for every weakly null sequence{fn}ninX and every sequence{An}nof disjoint measurables sets, the sequence{fn1IAn}n converges weakly to0 inX.

Proof. If the conclusion does not hold, there are a continuous linear functional σ: X→Cand someδ >0 such that, up to taking a subsequence,|σ(fn1IAn)| ≥ δ. Set, for every measurable setA∈ S:

µn(A) =σ(fn1IA).

Thenµn is a finitely additive measure with bounded variation. By Rosenthal’s lemma (see [3], Lemma I.4.1, page 18, or [1], Chapter VII, page 82), there is an increasing sequence of integers{nk}k such that:

µnk

[

l6=k

Anl≤ |µnk| [

l6=k

Anl

≤δ/2.

Now, ifA=S

l≥1Anl,{fnk1IA}k is weakly null, but:

|σ(fnk1IA)| ≥ |σ(fnk1IAnk)| − |µnk| [

l6=k

Anl

≥δ−δ 2 = δ

2,

so we get a contradiction.

Proof of Theorem 4.1. We begin by defining a sequence{xn}n of positive numbers in the following way: setx1 = 4 and, for everyn≥1, xn+1 >2xn is the abscissa of the second intersection point of the parabolay =x2 with the straight line containing (xn, x2n) and (2xn, x4n); we have xn+1 =x3n−2xn (for example,x2= 56). Define Ψ : [0,+∞)→[0,+∞) by Ψ(x) = 4xfor 0≤x≤4, and, forn≥1:

(4.1) Ψ(xn) =x2n, Ψ(2xn) =x4n, Ψ affine between xn andxn+1. Then Ψ is an Orlicz function and

(4.2) x2≤Ψ(x)≤x4 for x≥4.

(18)

For this Orlicz function Ψ,JΨ is not compact, since Ψ(2x)/[Ψ(x)]2does not tend to 0. However, JΨ is weakly compact, because one has the factorization HΨ֒→H2֒→B4֒→BΨ (by (4.2) and Lemma 2.4).

Assume that JΨ is not Dunford-Pettis: there exists a weakly null sequence {fn}n in the unit ball of HΨ which does not converges for the norm in BΨ. Then{fn}n converges uniformly to 0 on the compact subsets of D(since it is weakly null) and we may assume thatkfnkBΨ ≥ δ for some δ >0. We may also assume thatkfnkn→∞−→ +∞because if{fn}nwere uniformly bounded, we should havekfnkBΨn→∞−→ 0, by dominated convergence.

We are going to show that there exist a subsequence{fnk}kand pairwise dis- joint measurable setsAk ⊆Tsuch that the sequence{fnk1IAk}k ⊆LΨ(T, m) is equivalent to the canonical basis ofℓ1, whence a contradiction with Lemma 4.2.

It is worth to note from now that the Poisson integral P maps boundedly L2(T) into L4(D). Indeed, L2(T) = H2 ⊕H02 and the canonical injection is bounded fromH2into B4, by Lemma 2.4.

We have seen in the proof of Lemma 2.5 that there exist a subsequence {fnk}k and disjoint measurable annuli C1 = {z ∈ D; |z| ≤ r1} and Ck = {z ∈ D; rk−1 < |z| ≤ rk}, k ≥2, with 0 < r1 < r2 <· · · < rn < · · · < 1, such thatkfnk1ICkkLΨ(D) ≥δ/2. The assumptions of that lemma are satisfied here:kfnkHΨ≤1,kfnkBΨ≥δ,{fn}nconverges uniformly to 0 on the compact subsets ofD, andfn∈BMΨbecauseHΨ⊆BMΨ, sinceJΨis weakly compact.

Then:

Fact 1. There exist two sequencesk}k andk}k, with βn > αnn→∞−→ +∞ such that, ifgk =fnk1Ik≤|fnk |≤βk}, then:

kP(gk)kLΨ(D)≥δ/3, wherefnk is the boundary value offnk onT. Proof. 1) Letαk = 12δ Ψ−1 1/A(Ck)

and vk =P fnk1I{|fnk |<αk}

1ICk. One has:

Z

D

Ψ |vk|/(δ/12) dA=

Z

Ck

Ψ |vk|/(δ/12)

dA ≤Ψ αk/(δ/12)

A(Ck) = 1, so kvkkLΨ(D) ≤ δ/12. Since P(fnk) = fnk, we have kP(fnk) 1ICkkLΨ(D) = kfnk1ICkkLΨ(D)≥δ/2, and we get:

kP(fnk1I{|fnk ≥αk}) 1ICkkLΨ(D)≥ kfnk1ICkkLΨ(D)− kvkkLΨ(D)≥δ 2 − δ

12 =5δ 12· 2) Let wk =fnk1I{|fnk |≥αk}. Since P(wk1I{|wk|>β}) tends to 0 uniformly on Ck whenβ goes to infinity, Lebesgue’s dominated convergence theorem gives:

kP(wk1I{|wk|>β}) 1ICkkLΨ(D)≤ kP(wk1I{|wk|>β}) 1ICkkL4(D) −→

β→+∞0,

(19)

so there is someβk > αk such thatkP(wk1I{|wk|>β}) 1ICkkLΨ(D)≤δ/12.

We then have, withgk=fnk1Ik≤|f

nk|≤βk}: kP(gk)kLΨ(D)≥ kP(gk) 1ICkkLΨ(D)≥ 5δ

12− δ 12 =δ

3,

and that ends the proof of Fact 1.

Fact 2. There are a further subsequence, denoted yet by {fnk}k, and pairwise disjoint measurable subsetsEk ⊆ {αk≤ |fnk| ≤βk}, such that, ifhk=fnk1IEk, then:

kP(hk)kLΨ(D)≥δ/4.

Proof. First, since gk ∈ L(T) ⊆ MΨ(T), there exists εk > 0 such that m(A)≤εk implieskgk1IAkLΨ(T)≤δ/(12kPk) (wherekPkstands for the norm ofP: L2(T)→L4(D)). Now, P: LΨ(T)→LΨ(D) is bounded and its norm is

≤ kPk, thanks to the factorizationLΨ(T)֒→L2(T)֒→L4(D)֒→LΨ(D). Hence kP(gk1IA)kLΨ(D)≤δ/12 form(A)≤εk.

LetBk ={αk ≤ |fnk| ≤βk}. Up to taking a subsequence, we may assume thatP

l>km(Bl)≤εk. Let

Ek =Bk\[

l>k

Bl.

The setsEk,k≥1, are pairwise disjoint, and

kP(gk1IEk)kLΨ(D)≥ kP(gk1IBk)kLΨ(D)− kP gk1ISl>kBl

kLΨ(D)≥ δ 3− δ

12 =δ 4; so we get the Fact 2 withhk =gk1IEk=fnk1IEk.

Set

Fk={z∈Ek; Ψ |fnk(z)

| ≤M|fnk(z)|2}. Forz∈Ek\Fk, one has:

Z

Ek\Fk|fnk|2dm≤ 1 M

Z

T

Ψ(|fnk)|dm≤ 1 M , sokfnk1IEk\FkkL2(T)≤1/√

M and:

kP(fnk1IEk\Fk)kLΨ(D)≤ kP(fnk1IEk\Fk)kL4(D)

≤ kPk k(fnk1IEk\Fk)kL2(T)≤ kPk

√M ≤δ 8,

forM large enough. It follows that, for M large enough,kP(fnk1IFk)kLΨ(D)≥ δ/8 and

(4.3) kfnk1IFkkLΨ(D)≥δ/(8kPk).

Références

Documents relatifs

its scope is the smallest textually enclosing routine body or result block and its initial Rvalue is a bit pattern representing the program position of the command

(b) The previous question shows that it is enough to consider the case k, l 6 n/2, for otherwise we can replace the representations by isomorphic ones so that these inequalities

When the vector field Z is holomorphic the real vector fields X and Y (cf. This follows from a straightforward computation in local coordinates.. Restriction to the

In the first part, by combining Bochner’s formula and a smooth maximum principle argument, Kr¨ oger in [15] obtained a comparison theorem for the gradient of the eigenfunctions,

Kobayashi, S., Ochiai, T.: Three-dimensional compact K~ihler manifolds with positive holo- morphic bisectional curvature... Kobayashi, S., Ochiai, T.: Characterizations

A natural question about a closed negatively curved manifold M is the following: Is the space MET sec&lt;0 (M) of negatively curved met- rics on M path connected?. This problem has

Suppose R is a right noetherian, left P-injective, and left min-CS ring such that every nonzero complement left ideal is not small (or not singular).. Then R

Characterization of removable sets in strongly pseudoconvex boundaries 459 The known proof of Theorem 0 is function-theoretic in nature; on the contrary the proofs of Theorem