• Nenhum resultado encontrado

On approximation numbers of composition operators Daniel Li, Hervé Queffélec, Luis Rodriguez-Piazza

N/A
N/A
Protected

Academic year: 2024

Share "On approximation numbers of composition operators Daniel Li, Hervé Queffélec, Luis Rodriguez-Piazza"

Copied!
35
0
0

Texto

(1)

HAL Id: hal-00588256

https://hal-univ-artois.archives-ouvertes.fr/hal-00588256

Preprint submitted on 22 Apr 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

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

On approximation numbers of composition operators

Daniel Li, Hervé Queffélec, Luis Rodriguez-Piazza

To cite this version:

Daniel Li, Hervé Queffélec, Luis Rodriguez-Piazza. On approximation numbers of composition oper-

ators. 2011. �hal-00588256�

(2)

On approximation numbers of composition operators

Daniel Li, Hervé Queffélec, Luis Rodríguez-Piazza

April 22, 2011

Abstract. We show that the approximation numbers of a compact composi- tion operator on the weighted Bergman spacesBα of the unit disk can tend to 0 arbitrarily slowly, but that they never tend quickly to 0: they grow at least exponentially, and this speed of convergence is only obtained for symbols which do not approach the unit circle. We also give an upper bounds and explicit an example.

Mathematics Subject Classification.Primary: 47B06 – Secondary: 47B33;

47B10

Key-words. approximation number – Bergman space – Carleson measure – composition operator – Hardy space – interpolation sequence – reproducing kernel – weighted Bergman space – weighted shift

1 Introduction

LetDbe the open unit disk of the complex plane, equipped with its normal- ized area measuredA(z) = dxdyπ . Forα >−1, letBαbe the weighted Bergman space of analytic functionsf(z) =P

n=0anzn onDsuch that kfk2α= (α+ 1)

Z

D

|f(z)|2(1− |z|2)αdA(z) = X n=0

n!Γ(2 +α)

Γ(n+ 2 +α)|an|2<∞. The limiting case, as α−→ −> 1, of those spaces is the usual Hardy space H2 (indeed, iff is a polynomial, we havelimα >

−→ −1kfk2α=P

n=0|an|2=kfk2H2), which we shall treat asB1. Note that kfk2α≈P

n=0 |an|2

(n+1)α+1 and that dAα(z) = (α+ 1)(1− |z|2)αdA(z)

is a probability measure onD.

Bergman spaces ([46] page 75, page 78) are Hilbert spaces of analytic func- tions on D with reproducing kernel Ka ∈ Bα, given by Ka(z) = (11az)α+2, namely, for everya∈D:

(1.1) f(a) =hf, Kai, ∀f ∈Bα; and kKak2=Ka(a) = 1 1− |a|2

α+2

.

(3)

An important common feature of those spaces is that the multipliers ofBαcan be (isometrically) identified with the space H of bounded analytic functions onD, that is:

(1.2) ∀g∈H, kgk= sup

fBα,kfkα1kf gkα.

Indeed,kf gkα≤ kgkkfkαis obvious, and if kf gkα≤Ckfkαfor all f ∈Bα, testing this inequality successively on f = 1, g, . . . , gn, . . . easily givesg ∈H andkgk≤C.

Let nowϕbe anon-constant analytic self-map (a so-calledSchur function) ofDand letCϕ:Bα→H(D)the associated composition operator:

Cϕ(f) =f◦ϕ.

It is well-known ([9] page 30) that such an operator is always bounded fromBα into itself, and we are interested in its approximation numbers.

Also recall that theapproximation (or singular)numbers an(T)of an oper- atorT ∈ L(H1, H2), between two Hilbert spaces H1 and H2, are defined, for n= 1,2, . . ., by:

an(T) = inf{kT−Rk; rank (R)< n}. We have:

an(T) =cn(T) =dn(T),

where the numberscn (resp. dn) are theGelfand (resp. Kolmogorov)numbers ofT ([6], page 59 and page 51 respectively).

In the sequel we shall need the following quantity:

(1.3) β(T) = lim inf

n→∞

an(T)1/n

.

Those approximation numbers form a non-increasing sequence such that a1(T) =kTk, an(T) =an(T) =p

an(TT)

and verify the so-called “ideal” and “subadditivity” properties ([17] page 57 and page 68):

(1.4) an(AT B)≤ kAkan(T)kBk; an+m1(S+T)≤an(S) +am(T).

Moreover, the sequence(an(T))tends to0iffT is compact. If(an(T))∈ℓp, we say thatT belongs to the Schatten class Sp of indexp,0< p <∞. Taking for T a compact diagonal operator, we see that this sequence is non-increasing with limit0, but otherwise arbitrary. But if we restrict ourselves to aspecified class of operators, the answer is far from being so simple, although in some cases the situation is completely elucidated. For example, for the class of Hankel operators onH2(those operatorsHφwhose matrix(ai,j)on the canonical basis

(4)

of H2 is of the formai,j = φ(ib +j) for some function φ ∈ L), it is known that Hφ is compact if and only if the conjugate φ¯ of the symbol φ belongs to H+C, where C denotes the space of continuous, 2π-periodic functions (Hartman’s theorem, [32] page 214). For those Hankel operators, the following theorem, due to A. V. Megretskii, V. V. Peller, and S. R. Treil ([31] and [37], Theorem 0.1, page 490), shows that the approximation numbers are absolutely arbitrary, under the following form.

Theorem 1.1 (Megretskii-Peller-Treil) Letn)n1be a non-increasing se- quence of positive numbers. Then there exists a Hankel operatorHφ satisfying:

an(Hφ) =εn, ∀n≥1.

Indeed, if we take a positive self-adjoint operatorAwhose eigenvaluessn coin- cide with theεn’s and whose kernel is infinite-dimensional, it is easily checked that this operator A verifies the three necessary and sufficient conditions of Theorem 0.1, page 490 in [37] and is therefore unitarily equivalent to a Hankel operatorHφwhich will verify, in view of (1.4):

an(Hφ) =an(A) =εn, n= 1,2, . . .

In particular, ifεn→0, the above Hankel operator will be compact, and in no Schatten class if εn = 1/log(n+ 1) for example. We also refer to [16] for the following slightly weaker form due to S. V. Khruscëv and V. Peller, but with a more elementary proof based on interpolation sequences in the Carleson sense:

for anyδ >0, there exists a Hankel operatorHφ such that 1

1 +δεn≤an(Hφ)≤(1 +δ)εn, n= 1,2, . . .

Now, the aim of this work is to prove analogous theorems for the class of composition operators (whose compactness was characterized in [29] and [42]).

But if we are able to obtain the Khruscëv-Peller analogue for thelower bounds, we will only obtain subexponential estimates for the upper bounds, a fact which is explained by our second result: the speed of convergence to0 of the approx- imation numbers of a composition operator cannot be greater than geometric (andis geometric for symbolsϕverifyingkϕk<1). Our first result involves a constant<1 and is not as precise as the result of Megretskii-Peller-Treil or even that of Khruscëv-Peller; this is apparently due to the non-linearity of the dependence with respect to the symbol for the class of composition operators, contrary to the case of the Hankel class.This latter lower bound improves several previously known results on “non-Schattenness” of those operators (see Corol- lary 4.2 below) and also answers in the positive to a question which was first asked to us by C. Le Merdy ([26]) in the OT Conference 2008 of Timisoara, con- cerning the bad rate of approximation of compact composition operators. Those theorems are, to our knowledge, the first individual results on approximation numbersan of composition operators (in the work of Parfenov [35], some good estimates are given for the approximation numbers of the Carleson embedding

(5)

operator in the case of the spaceH2 = B1, but they remain fairly implicit, and are not connected with composition operators), whereas all previous results where in terms of symmetric norms of the sequence(an), not on the behaviour of eachan.

Before describing our results, let us recall two definitions. For everyξwith

|ξ|= 1and0< h <1, the Carleson windowW(ξ, h)centered atξand of sizeh is the set

W(ξ, h) ={z∈D; |z| ≥1−hand|arg(zξ)| ≤πh}.

Letµbe a positive, finite, measure onD; the associated maximal functionρµis defined by:

(1.5) ρµ(h) = sup

|ξ|=1

µ W(ξ, h) .

The measureµ is called aCarleson measure for the Bergman space Bα, or an (α+ 2)-Carleson measure (including the caseB1 =H2), ifρµ(h) =O(h2+α) ash→0. For any Schur functionϕ, we shall denote bymϕthe imageϕ(m)of the Haar measuremof the unit circle under the radial limits functionϕ(u) = limr1ϕ(ru)ofϕ,|u|= 1, and byAϕ,α+2the image of the probability measure (α+ 1)(1− |z|2)αdA(z) underϕ. The corresponding maximal function will be denoted byρϕ,α+2. This notation is justified by the fact thatmϕ

def= Aϕ,1is a1- Carleson measure andAϕ,α an(α+ 2)-Carleson measure forα >−1, in view of the famous Carleson embedding theorem which, expressed under a quantitative and generalized form, states the following, implicit as concerns kjk and with different notations, but fully proved in [44], Theorem 1.2, for the caseα >−1 (see [32], page 153).

Theorem 1.2 (Carleson’s theorem) For any (α+ 2)-Carleson measure µ, the canonical inclusion mappingj:Bα→L2(µ)is defined and continuous, and its norm satisfies

(1.6) C1 sup

0<h<1

µ(h)

h2+α ≤ kjk ≤C sup

0<h<1

µ(h) h2+α·

The paper is organized as follows. Section 1 is this introduction. In Sec- tion 2, we prove some preliminary lemmas. Our first theorems concern lower bounds. In Section 3, we prove (Theorem 3.1) that the convergence of the ap- proximation numbersan(Cϕ)of a composition operatorCϕ:Bα→Bαcannot exceed an exponential speed: for somer∈(0,1) and some constantc >0, one hasan(Cϕ)≥c rn. More precisely, with the notations (1.3) and (3.1), one has β(Cϕ)≥[ϕ]2. Moreover, this speed of convergence is only attained if the values ofϕdo not approach the boundary of the unit disk: kϕk<1(Theorem 3.4).

On the other hand, the speed of convergence to0 ofan(Cϕ)can be arbitrarily slow; this is proved in Section 4. The proof is mainly an adaptation of the

(6)

one in [7], but is fairly technical at some points, and will require several addi- tional explanations. In Section 5, we prove an upper estimate (Theorem 5.1), and give three applications of this theorem. In the final Section 6, we test our general results against the example of lens maps, which are known to generate composition operators belonging to all Schatten classes.

2 Preliminary lemmas

In this Section, we shall state several lemmas, which are either already known or quite elementary, but turn out to be necessary for the proofs of our Theo- rem 3.1 and Theorem 4.1.

For the proof of Theorem 3.1, we shall need the Weyl lemma ([6] Proposi- tion 4.4.2, page 157).

Lemma 2.1 (Weyl lemma) Let T:H →H be a compact operator. Suppose thatn)n1 is the sequence of eigenvalues of T rearranged in non-increasing order. Then, we have:

Yn k=1

ak(T)≥ Yn k=1

k|.

We recall ([3], [13] pages 194–195, [33] pages 302–303) that aninterpolation sequence(zn)with (best) interpolation constantCis a sequence(zn)(necessarily Blaschke, i.e. P

n=1(1− |zn|)<∞) in the unit disk such that, for any bounded sequence(wn) of scalars, there exists a bounded analytic functionf (i.e. f ∈ H) such that:

f(zn) =wn, ∀n≥1, and kfk≤Csupn1|wn|.

TheCarleson constant δof a Blaschke sequence(zn)is defined as follows:

(2.1) δn=Y

j6=n

ρ(zn, zj) ; δ= infδn= inf

n1(1− |zn|2)|B(zn)|,

where B is the Blaschke product with zeroes zn, n ≥ 1. The interpolation constant C is related to the Carleson constant δ by the following inequality ([10] page 278), in whichλis a positive numerical constant:

(2.2) 1

δ ≤C ≤λ δ

1 + log1 δ

·

This latter inequality can be viewed as a quantitative form of the Carleson interpolation theorem. Interpolation sequences and reproducing kernels ofBα are related as follows ([33] pages 302–303).

Lemma 2.2 Let (zn)n1 be an H-interpolation sequence of the unit disk, with interpolation constant C. Then, the sequence (fn) = (Kzn/kKznk) of

(7)

normalized reproducing kernels atzn isC-equivalent to an orthonormal basis in Bα, namely we have for any finite sequencen)of scalars:

(2.3) C1 X

nn|21/2

≤X

nλnfn

α≤C X

nn|21/2

.

The proof in [33] is only forH2, therefore we indicate a simple proof valid for Bergman spacesBα as well. LetS =P

λnKzn be a finite linear combination of the kernelsKzn,ω= (ωn)be a sequence of complex signs,Sω=P

ωnλnKzn

andg ∈H an interpolating function for the sequence (ωn), i.e. g(zn) =ωn

andkgk≤C. Iff ∈Bαandkfkα≤1, we see that:

hSω, fi=X

ωnλnf(zn) =X

λn(f g)(zn) =X

λnhKzn, f gi=hS, f gi, so that using (1.2):

|hSω, fi| ≤ kSkαkf gkα≤ kSkαkgkkfkα≤CkSkα

and passing to the supremum onf, we get kSωkα ≤CkSkα. Since the coeffi- cientsλn are arbitrary, this implies that(fn)isC-unconditional, namely:

C1 X

ωnλnfn

α

X

λnfn

α≤C

X

ωnλnfn

α.

Now, squaring and integrating with respect to random, independent, choices of

signsωn’s, we get (2.3).

We also recall ([13] pages 203–204) that an increasing sequence(rn)of num- bers such that 0 < rn < 1 and 11rn+1rn ≤ ρ < 1 (i.e. verifying the so-called Hayman-Newman condition) is an interpolation sequence (see also [32]). In the following, let(rn)be such a sequence verifying moreover the backward induction relation:

(2.4) ϕ(rn+1) =rn.

Setfn = Krn/kKrnk and W = span(fn). Let (en)n1 be the canonical basis of ℓ2, ϕa Schur function and h∈H a function vanishing at r1. Denote by Mh:Bα→Bαthe operator of multiplication byh. Then, we have the following basic lemma, which shows that some compression ofCϕ is a backward shift with controlled weights ([7]).

Lemma 2.3 Let J:ℓ2 →W be the isomorphism given byJ(en) =fn. Then, the operatorB=J1CϕMhJ:ℓ2→ℓ2 is the weighted backward shift given by:

(2.5) B(en+1) =wnen and B(e1) = 0, where wn=h(rn+1) kKrnk kKrn+1

(8)

To exploit Lemma 2.3, we shall need the following simple fact on approxi- mation numbers of weighted backward shifts.

Lemma 2.4 Let (en)n1 be an orthonormal basis of the Hilbert space H and B∈ L(H)the weighted backward shift defined by

B(e1) = 0 and B(en+1) =wnen, wherewn→0.

Assume that |wn| ≥εn for all n≥1, wheren) is a non-increasing sequence of positive numbers. Then Bis compact, and satisfies:

(2.6) an(B)≥εn, ∀n≥1.

Proof. The compactness of Bis obvious. Let R be an operator of rank< n.

Then kerR is of codimension < n, and therefore intersects the n-dimensional space generated bye2, . . . , en+1 in a vector x=Pn

j=1xjej+1 of norm one. We then have:

kB−Rk2≥ kBx−Rxk2=kBxk2=Xn

j=1|wj|2|xj|2

≥Xn

j=1ε2j|xj|2≥ε2n

Xn

j=1|xj|22n.

This ends the proof of Lemma 2.4.

Now, in view of (1.1) and (2.5), the weightwnroughly behaves asq

1rn+1

1rn , so we shall need good estimates on that quotient, before defining the sequence (rn)explicitly.

We first connect this estimate with the hyperbolic distancedinD. We denote (see [12] or [15] for the definition) byd(z, w;U)the hyperbolic distance of two points z, w of a simply connected domainU. It follows from the generalized Schwarz-Pick lemma ([15] Theorem 7.3.1, page 130) applied to the canonical injectionU →V that the bigger the domain the smaller the hyperbolic distance, namely:

(2.7) U ⊂V andz, w∈U =⇒ d(z, w;V)≤d(z, w;U).

Moreover, as is well-known,

0≤r <1 =⇒ d(0, r;D) = 1

2log1 +r 1−r·

Recall that the pseudo-hyperbolic and hyperbolic distances ρ and d onD are defined by:

ρ(a, b) = a−b

1−ab

, d(a, b) =1

2log1 +ρ(a, b)

1−ρ(a, b), a, b∈D,

In the sequel, we shall omit the symbolDas far as the open unit disk is con- cerned. For this unit disk, we have the following simple inequality ([7]) . Lemma 2.5 Let a, b∈Dwith 0< a < b <1. Then:

(2.8) e2d(a,b)≤ 1−b

1−a ≤2 e2d(a,b).

(9)

Finally, before proceeding to the construction of our Schur function ϕ in Section 4, it will be useful to note the following simple technical lemma.

Lemma 2.6 Letn)be a non-increasing sequence of positive numbers of limit 0. Then there exists a decreasing and logarithmically convex sequencen) of positive numbers, with limit0, such thatδn≥εn for all n≥1.

Proof. Provided that we replace εn by εn+n1, we may assume that (εn) is decreasing. Let us define our new sequence by the inductive relation:

δ11; δ22; δn+1= max εn+1, δn2n1

.

This sequence is log-convex by definition, i.e. δn2 ≤ δn+1δn1. By induction, it is seen to be decreasing. Therefore, it has a limit l ≥ 0. If δn = εn for infinitely many indices, l = 0. Otherwise, for n large enough, we have the inductive relation δn+1 = δn2n1, which implies that δn = exp(λn+µ) for some constantsλ, µ. Since(δn)is decreasing, we must haveλ <0and again we

getl= 0.

In the sequel, we may and will thus assume, without loss of generality, that (εn)is decreasing and logarithmically convex.

3 Lower bounds

We first introduce a notation. If ϕ#(z) = lim

wz

ρ(ϕ(w), ϕ(z))

ρ(w, z) = |ϕ(z)|(1− |z|2) 1− |ϕ(z)|2 is the pseudo-hyperbolic derivative ofϕ, we set:

(3.1) [ϕ] = sup

z∈D

ϕ#(z) =kϕ#k.

In our first theorem, we get that the approximation numbers cannot super- sede a geometric speed.

Theorem 3.1 For any Schur function ϕ, there exist positive constantsc > 0 and0< r <1 such that, forCϕ:Bα→Bα, we have:

(3.2) an(Cϕ)≥c rn, n= 1,2, . . .

More precisely, one hasβ(Cϕ)≥[ϕ]2 and hence, for eachκ <[ϕ], there exists a constantcκ>0 such that:

(3.3) an(Cφ)≥cκκ2n.

(10)

For the proof, we need the following lemma.

Lemma 3.2 Let T:H →H be a compact operator. Suppose thatn)n1, the sequence of eigenvalues of T rearranged in non-increasing order, satisfies, for someδ >0 andr∈(0,1):

n| ≥δrn, n= 1,2, . . . Then there existsδ1>0 such that

an(T)≥δ1r2n, n= 1,2, . . . In particular β(T)≥r2.

Proof. By Weyl’s inequality (Lemma 2.1), we have Yn

k=1

ak(T)≥ Yn k=1

k| ≥δnrn(n+1)/2.

Sinceak(T)is non-increasing andak(T)≤ kTkfor everyk, changingninto2n, we get:

kTknan(T)n ≥ Y2n

k=1

ak(T)≥δ2nrn(2n+1)≥δ2nr2n2

and thereforean(T)≥ kδT2kr2n1r2n, as claimed.

By applying this lemma to composition operators, we get the following result, which ends the proof of Theorem 3.1.

Proposition 3.3 For every composition operator Cϕ: Bα → Bα of symbol ϕ:D→D, we haveβ(Cϕ)≥[ϕ]2.

Proof. For everya∈ D, letΦa be the (involutive) automorphism of the unit disk defined by

Φa(z) = a−z

1−az, z∈D. Observe that we have

Φa(a) = 0, Φa(0) =a, Φa(a) = 1

|a|2−1, Φa(0) =|a|2−1.

Define nowψ = Φϕ(a)◦ϕ◦Φa. We have that 0 is a fixed point of ψ, whose derivative is, by the chain rule:

(3.4) ψ(0) = Φϕ(a)(φ(a))ϕ(a)Φa(0) = ϕ(a)(1− |a|2) 1− |ϕ(a)|2

def= ϕ#(a).

By Schwarz’s lemma, we know that |ψ(0)| ≤ 1 and so |ϕ1(a)−||ϕ(a)(1−||a2|2) ≤ 1 (Schwarz-Pick’s inequality).

(11)

Let us first assume that the composition operatorCϕ is compact. Then, so isCψ, since we have

(3.5) Cψ=CΦa◦Cϕ◦CΦϕ(a).

Ifψ(0)6= 0, the sequence of eigenvalues ofCψ is [ψ(0)]n

n0 ([41], page 96;

the result given for the space H2 holds forBα ⊂H2, and would also hold for any space of analytic functions inDon whichCψ is compact). Lemma 3.2 then gives us:

β(Cψ)≥ |ψ(0)|2= [ϕ#(a)]2≥0.

This trivially still holds ifψ(0) = 0.

Now, sinceCΦaandCΦϕ(a) are invertible operators, (3.5) clearly implies that β(Cϕ) =β(Cψ), and therefore, with the notation of (3.4):

β(Cϕ)≥[ϕ#(a)]2, for alla∈D.

By passing to the supremum ona∈D, we end the proof of Proposition 3.3, and that of Theorem 3.1 in the compact case. IfCϕis not compact, the proposition trivially holds. Indeed, in this case, we haveβ(Cϕ) = 1≥[ϕ]2. Remark. It is easy to see that the composition operatorCϕis always of infinite rank, contrary to the case of a Hankel operator, so that in some sense it refuses to be approached by finite-rank operators. Theorem 3.1 quantifies things: it is a well-known and easy fact (see for example [41], page 25 and see Theorem 5.1 to come) that, in the case kϕk < 1, we have an(Cϕ) ≤ ckϕkn (and hence β(Cϕ) ≤ kϕk < 1), showing that the approximation numbers can decrease at an exponential speed. Theorem 3.1 shows that this speed is the maximal possible one. The next theorem says that this maximal speed isonly obtained whenkϕk<1.

Theorem 3.4 For everyα≥ −1, there exists, for any0< r <1,s=s(r)<1, satisfying limr1s(r) = 1, such that, for Cϕ: Bα → Bα, one has, with the notation coined in (1.3):

(3.6) kϕk> r =⇒ β(Cϕ)≥s2.

In particular, the exponential speed of convergence to 0 of the approximation numbers of a composition operator Cφ of symbol ϕ takes place if and only if kϕk<1; in other words, we have:

(3.7) kϕk= 1 ⇐⇒ β(Cϕ) = 1.

The proof will proceed through a series of lemmas. Throughout that proof, we assume, without loss of generality, thatϕ(0) = 0.

(12)

Lemma 3.5 LetKbe a compact subset ofϕ(D)andµbe a probability supported byK. Then, there exists a constantδ >0such that, ifRµ:Bα→L2(µ)denotes the restriction operator, we have:

an(Cϕ)≥δ an(Rµ).

In particular:

β(Cϕ)≥β(Rµ).

Proof. Sinceϕis an open map, there exists a compact setL⊂Dand a Borel subset A ⊂ L such that ϕ(A) = K and ϕ: A → K is a bijection (see [36], Chapter I, Theorem 4.2). Then µ= ϕ(ν), whereν =ϕ1(µ) is a probability measure supported by L, and we have automatically kRνk < ∞. Then, for everyf ∈Bα:

kfk2L2(µ)= Z

K|f|2dµ= Z

L|f◦ϕ|2dν=kCϕfk2L2(ν).

This yieldskRµfk=k(Rν◦Cϕ)fk, soCϕ acts as an isometry fromL2(µ)into L2(ν), and the lemma follows, since we have then:

an(Rµ) =an(Rν◦Cϕ)≤ kRνkan(Cϕ)

for everyn≥1.

Observe that this provides a new proof of Theorem 3.1. Indeed, ifK⊂ϕ(D) is a small ball of center0 and radius r, we can take forµ the normalized area measure on K; then Parseval’s formula easily shows that β(Rµ) ≥ r in that case.

The strategy of the proof of Theorem 3.4 will consist of refining this obser- vation. More precisely, we shall show that the situation can be reduced to the caseK= [0, r], and that an appropriate choice ofµ can be made in that case, giving a sharp lower bound forβ(Rµ). We begin with explaining that choice in the next two lemmas.

Lemma 3.6 For everyr ∈(0,1) there exists s=s(r)<1 and f =fr ∈H with the following properties:

1)limr1s(r) = 1;

2)kfk≤1;

3)f((0, r]) =s ∂Din a one-to-one way.

Proof. Letρ=1r1r2. Thenr= 1+ρ2 and the automorphismϕρ(z) = 1ρρzz maps [0, r] onto [−ρ, ρ]. We define ε = ε(r) and s = s(r) by the following relations:

(3.8) ε(r) = π

log1+ρ1ρ , and s= eεπ/2.

(13)

Let now

(3.9) χ(z) =εlog1 +ϕρ(z)

1−ϕρ(z) and

(3.10) f(z) =seiχ(z).

Note thatf = eh, where

h(z) =iεlog1 +ϕρ(z) 1−ϕρ(z)−επ

2

is a conformal mapping fromDonto a small vertical strip of the left-half plane.

This function f fulfills all the requirements of the lemma. Indeed, we have

|f(z)| ≤1for allz∈Dand h([0, r]) =

−iεlog1 +ρ

1−ρ,iεlog1 +ρ 1−ρ

−επ

2 = [−iπ, iπ]−επ 2 , so thatf((0, r]) ={w=se; −π≤θ≤π}, in a one-to-one way.

Lemma 3.6 allows a good choice of the measureµas follows.

Lemma 3.7 Letf be as in Lemma 3.6. Then, there exists a probability measure µ=µrsupported by[0, r]and a constantδr>0such that, for any integern≥1 and any choice of scalarsc0, c1, . . . , cn1, we have:

n1

X

j=0

cjRµ(fj)

L2(µ)

≥ sn

√n

nX1 j=0

cjfj

H2

≥ sn

√n

n1

X

j=0

cjfj

Bα

.

As a consequence, we can claim that, forCϕ:Bα→Bα: (3.11) ϕ(D)⊃[0, r] =⇒ β(Cϕ)≥s=s(r).

Proof. With our previous notations, we know thatχ is a bijective map from ]0, r]onto the unit circle∂D. Letµ=χ1(m)be the image of the Haar measure mof∂Dbyχ1. We have by definition ofµ:

nX1 j=0

cjRµ(fj)

2

L2(µ)

= Z r

0

n1

X

j=0

cjfj(x)

2

dµ(x) = Z r

0

n1

X

j=0

cjsjeijχ(x))

2

dµ(x)

= Z π

π

n1

X

j=0

cjsjeijθ

2 dθ 2π =

n1

X

j=0

|cj|2s2j

≥s2n

n1

X

j=0

|cj|2.

(14)

Now,kfjkH2≤ kfjk≤1, so that we have, using the Cauchy-Schwarz inequal- ity:

n1

X

j=0

cjfj

H2

nX1 j=0

|cj| kfjkH2

nX1 j=0

|cj| ≤√ n

nX1 j=0

|cj|2 1/2

,

giving the first inequality, since k kH2 ≥ k kBα. Finally, let R: Bα → L2(µ) be an operator of rank< n. We can find a functiong =Pn1

j=0cjfj such that kgkBα= 1 andR(g) = 0. The first part of the proof gives:

kRµ−Rk ≥ kRµ(g)−R(g)k=kRµ(g)k=

nX1 j=0

cjfj

L2(µ)

≥ sn

√n

nX1 j=0

cjfj

Bα

= sn

√n· Thereforean(Rµ)≥sn/√

n and, in view of Lemma 3.5, the last conclusion of

Lemma 3.7 follows.

The next lemma explains how to reduce the situation to the caseK= [0, r]

when we only know that kϕk > r. It was inspired to us by the proof of the Lindelöf theorem that convergence along a curve implies non-tangential convergence for functions in Hardy spaces ([39] page 300).

Lemma 3.8 Suppose that 0 andr belong to ϕ(D), with0< r <1. Letµ be a probability measure carried by[0, r]. Then, there exists a probability measure ν carried by a compact setK⊂ϕ(D)such that, for any f ∈ H(D):

(3.12)

Z

[0,r]|f(x)|2dµ(x)≤ 1 2 Z

K |f(z)|2+|f(¯z)|2 dν(z).

Proof. Sinceϕ(D)is open and connected and0, r∈ϕ(D), there is a curve with imageK⊂ϕ(D)connecting0 andr. PutK˜ ={z¯; z∈K}. Then, there exists a compact setLsuch that[0, r]⊂Land whose boundary∂L⊂(K∪K). Now,˜ the existence ofν carried byK will be provided by an appropriate application of the Pietsch factorization Theorem. To that effect, letX be the real subspace ofC(L)formed by the real functions which are harmonic in the interior ofL. By the maximum principle for harmonic functions,X can be viewed as a subspace ofC(K∪K). Now, the inclusion map˜ j of X intoL2(µ)has2-summing norm less than one ([1] page 208, or [24], Chapitre 5, Proposition I.3). Therefore, the Pietsch factorization Theorem ([1] page 209, or [24], Chapitre 5, Théorème I.5) implies the existence of a probabilityσonK∪K˜ such that, for everyu∈X:

(3.13) kuk2L2(µ)= Z

[0,r]

u2dµ≤ Z

KK˜

u2dσ.

(15)

For any harmonic functionuonD, we can apply (3.13) tou(z)andu(¯z)to get:

2 Z

[0,r]

u2dµ≤ Z

KK˜

u2(z) +u2(¯z)

dσ(z) = Z

KK˜

u2(z) +u2(¯z) d˜σ(z), where σ˜ is the symmetric measure ofσ, defined by σ(E) =˜ σ( ¯E). There is a probabilityν on K such thatν + ˜ν = σ+ ˜σ. For this probability ν, we thus have, for any real harmonic functionuonD:

(3.14) kuk2L2(µ)≤ Z

K

u2(z) +u2(¯z) dν(z).

Now, givenf ∈ H(D), we use (3.14) withuthe real and imaginary parts off,

and sum up to get (3.12).

We can now finish the proof of Theorem 3.4 as follows.

Suppose that kϕk > r. Then, making a rotation if necessary, we may assume that0, r∈ϕ(D)(recall that ϕ(0) = 0). Let µas in Lemma 3.7. Using Lemma 3.8, we find a probability measure ν, compactly supported by ϕ(D), such that (3.12) holds. This inequality shows that:

kRµfk2≤1

2 kRνfk2+kRν˜fk2 , so thatRµ =A(Rν ⊕Rν˜)withkAk ≤1/√

2 ≤1. Therefore, by the ideal and sub-additivity properties (1.4):

a2n(Rµ)≤a2n(Rν⊕R˜ν)≤an(Rν) +an(R˜ν) = 2an(Rν), implyingβ(Rν)≥β(Rµ)2. Finally, Lemma 3.5 and Lemma 3.7 give:

β(Cϕ)≥β(Rν)≥β(Rµ)2≥s(r)2,

and this ends the proof of Theorem 3.4.

Remark. The proof of Theorem 3.4 is strongly influenced by the papers [8]

and [45]. In the first one, it is proved that, ifK is a continuum of a connected open setΩand if the doubly connected region Ω\Kis conformally equivalent to the annulus1 <|z| < R, then there exists a linearly independent sequence (fn)inH(Ω) satisfying, for all scalarscj:

Xn j=1

cjfj

H(Ω)

≤Rn

Xn j=1

cjfj

C(K)

.

As a consequence, the author proves thatlimn→∞d1/nn = 1/R, where the num- bersdn are the Kolmogorov numbers (see [6] page 49 for the definition) of the restriction mapH(Ω)→ C(K). This statement led us to Lemma 3.7. In the second paper, it is proved that, for the same operator, one haslimn→∞d1/nn =

(16)

e1/C(K,Ω), where C(K,Ω)is the Green capacity of K relative to Ω. So that one has1/R= e1/C(K,Ω). In the case we were interested in, namelyΩ =Dand Kr= [−r, r], it seemed to us, for topological and analytic reasons, thatRshould tend to1asr→1, in other terms that we should havelimr1C(Kr,D) =∞. This is indeed the case ([40], Example II.1), but the proof is fairly involved, and the desire to get a reasonably simple and self-contained proof of Theorem 3.4 led us to the previous series of lemmas, once we were sure that the result was true.

4 Slow speed

In this section, we shall see that the speed of convergence to 0 of the ap- proximation numbers of a compact composition operator can be as slow as one wants. This answers in the positive to a question which was first asked to us by C. Le Merdy ([26]) in the OT Conference 2008 of Timisoara.

Theorem 4.1 Letn)n1 be a non-increasing sequence of positive real num- bers of limit zero. Then, there exists an injective Schur function ϕ such that ϕ(0) = 0andCϕ:Bα→Bα is compact, i.e. an(Cϕ)→0, but:

(4.1) lim inf

n→∞

an(Cϕ) εn >0.

Equivalently, we have for some positive numberδ >0, independent ofn:

an(Cϕ)≥δ εn for alln≥1.

As in the case of Hankel operators, an immediate consequence of Theorem 4.1 is the following:

Corollary 4.2 There exists a composition operator Cϕ: H2 → H2 which is compact, but in no Schatten class.

This corollary, which Theorem 4.1 reinforces and precises, was an answer to a question of Sarason, and has been first proved in [7]. Other proofs appeared in [2], [14], [19], [20], [47] (for a positive result on Schatten-ness, we refer to [28]).

The construction of the symbolϕin Theorem 4.1 follows that given in [7], but we have to proceed to some necessary adjustments. In order to exploit (2.8), we shall use, as in [7], the following two results due to Hayman ([12]) concerning the hyperbolic distanced(z, w;U)of two pointsz, w of a simply connected domain U (see also [15]), whose proof uses in particular the comparison principle (2.7):

Proposition 4.3 Suppose thatU contains the rectangle

R={z∈C; a1−b <Rez < a2+b,|Imz|< b}, wherea1< a2 andb >0. Then, we have the upper estimate:

(4.2) d(a1, a2;U)≤ π

4b(a2−a1) +π 2·

(17)

Proposition 4.4 Suppose thatU contains the rectangle

R={z∈C; a1−c <Rez < a2+c,|Imz|< c}, wherea1< a2 andc >0, but that the horizontal sides

{z∈C; a1−c≤Rez≤a2+c,|Imz|=c}

of that rectangle are disjoint fromU. Then, we have the lower estimate:

(4.3) d(a1, a2;U)≥ π

4c(a2−a1)−π 2·

We now proceed to the construction of our Schur functionϕ.

We first define a continuous mapψ:R→Ras follows:

ψ(t) =

(K(1 +|t|) if |t| ≤1

|t|/A(|t|) if |t|>1,

whereKis a positive constant adjusted below andA: [0,∞[→[0,∞[an increas- ing piecewise linear function on the intervals(0,1)and(en1,en)such that

A(0)def= A0= 0, A(en1)def=An forn≥1, and 2K= 1/A(1), the increasing sequence(An)being positive and concave forn≥1, and tending to∞. It then follows that the sequence of slopes AennAen−1n1 is decreasing, since An+1−An ≤An −An1 ≤e (An−An1), that the function A is increasing and concave on(0,∞)and vanishing at0, implying thatA(t)/tis decreasing on (0,∞), and that in particularψis increasing on(1,∞).

We then define a domainΩof the complex plane by:

(4.4) Ω ={w∈C; |Imw|< ψ(|Rew|)}.

Letσ: D→Ωbe the unique Riemann map such that σ(0) = 0 andσ(0)>0.

This map exists in view of the following simple fact.

Lemma 4.5 The domaindefined by (4.4)is star-shaped with respect to the origin and σ: (−1,1) → R is an increasing bijection such that σ(−1) = −∞

andσ(1) =∞.

Proof. The star-shaped character ofΩwill follow from the implication:

|Imw|< ψ(|Rew|) and 0< λ <1 =⇒ |Im (λw)|< ψ(|Re (λw)|).

We may assume that bothRew,Imw are positive, and it is enough to prove:

(4.5) λψ(x)≤ψ(λx), ∀λ∈[0,1], ∀x >0.

This is easy to check separating three cases:

(18)

1)x≤1; thenλψ(x) =λK(1 +x)≤K(1 +λx) =ψ(λx);

2)λx≤1< x; then, sinceA(x)> A(1), λψ(x) =λ x

A(x) <2Kλx≤K(1 +λx) =ψ(λx);

3)λx >1; we then have, sinceψ increases, λψ(x) =λ x

A(x) ≤ λx

A(λx) =ψ(λx)

and this ends the proof of (4.5). Now, sinceσis determined by the value ofσ(0) and the sign ofσ(0), we haveσ(z) =σ(z)for allz∈D, so thatσ[(−1,1)]⊂R. And since the derivative of an injective analytic function does not vanish and σ(0)>0, we get thatσis increasing on(−1,1). Finally, ifw∈Randw=σ(z), we havew=w, so thatσ(z) =σ(z)andz=z, which proves the surjectivity of

σ: (−1,1)→R.

We now chooseAn as follows,η >0denoting a positive numerical constant to be specified later.

(4.6) An =ηlog 1

εn

, n≥1.

Observe that this is an increasing, concave sequence tending to ∞ since we assumed that(εn)is log-convex and decreasing to0.

Finally, we define our Schur functionϕand our sequence(rn)under the form of the following lemma, in which the increasing character ofψis important.

Lemma 4.6 Let ϕbe defined by

ϕ(z) =σ1(e1σ(z)), and letrn1(en). Then we have:

1. ϕis univalent and maps DtoD,(rn)increases, andϕ(0) = 0;

2. ϕ(rn+1) =rn;

3. 11rn+1rn →0 and therefore (rn)is an interpolation sequence;

4. Cϕ: Bα→Bα is compact.

Proof. 1. SinceΩis star-shaped,e1σ(z)∈Ωwhenz∈D, soϕis well-defined and mapsDto itself in a univalent way. Moreover,ϕ(0) =σ1(0) = 0, and(rn) increases sinceσ1 increases onR.

2. We haveϕ(rn+1) =σ1 1eσ(rn+1)

1 1een+1

1(en) =rn. 3. This assertion is more delicate and relies on Proposition 4.4 as follows.

Setdn =ψ(en). We have clearlyen+1+dn+2<en+2 for largen(recall that ψ(t) = o(t) as t → ∞), so that ψ(en+1+dn+2)< ψ(en+2) =dn+2 sinceψ is increasing. By the intermediate value theorem for the functionψ(en+1+x)−x, we can therefore find a positive numbercn< dn+2such thatψ(en+1+cn) =cn.

(19)

Now, consider the open sets:

Rn ={z∈C; en−cn<Rez <en+1+cn and|Imz|< cn}, Un=Rn∪Ω.

Those setsUnsatisfy the assumptions of Proposition 4.4 in view of (4.4). Indeed, ifz belongs to the horizontal sides ofRn, we havez /∈Un since

en−cn ≤Rez≤en+1+cn =⇒ ψ(Rez)≤ψ(en+1+cn) =cn=|Imz|. This proposition then gives, sinceΩ⊂Un andcn < dn+2, and since the hyper- bolic metric is conformally invariant,

d(rn, rn+1) =d(en,en+1; Ω)≥d(en,en+1;Un)≥ π 4cn

(en+1−en)−π 2

≥c en+2

ψ(en+2) =cA(en+2)≥cAn,

wherecis a positive constant. Now, we use Lemma 2.5 to obtain:

1−rn+1

1−rn ≤2 e2d(rn,rn+1)≤2 e2cAn, which proves that 11rn+1r

n → 0, and implies that (rn) is an interpolation se- quence.

4. Sinceϕis univalent, the compactness ofCϕ:Bα→Bαamounts to prov- ing thatlim|z|→11−|ϕ(z)|

1−|z| =∞. Forα >−1, this follows from [30], Theorem 3.5 and forα=−1 from [41], page 39. By the Julia-Carathéodory Theorem ([41], page 57), this in turn is equivalent to proving that for anyu, von the unit circle, the quotient ϕ(z)zuv has no finite limit asz tends touradially. This latter fact requires some precise justification.

First, we notice that σ extends continuously to an injective map of the open upper half of the unit circle onto the upper part of the boundary of Ω (and similarly for lower parts). This follows from the Carathéodory extension theorem ([39], page 290), applied to the restriction ofσ1 to the Jordan region limited by ∂Ω and two vertical lines Rew = ±R where R > 0 is arbitrarily large. Now, let u∈∂Dwith u6=±1. Then, σ(ru) →w∈ ∂Ωas r →1, so that e1σ(ru) →e1w =w ∈ Ωand that ϕ(ru)→ σ1(w) ∈D. Therefore the image ofϕ touches the unit circle only at ±1, and the assumption of the Julia-Carathéodory Theorem is fulfilled ifu6=±1. By symmetry, it remains to test the pointu= 1 for which we have:

lim sup

r−→< 1

1−ϕ(r)

1−r ≥lim sup

n→∞

1−ϕ(rn+1) 1−rn+1

= lim sup

n→∞

1−rn

1−rn+1

=∞ by the preceding point 3. Since |v−ϕ(r)| ≥ 1−ϕ(r), this ends the proof of

Lemma 4.6.

(20)

We now want a good lower bound for the weightswn appearing in (2.5). To that effect, we apply Proposition 4.3 with

U = Ω, a1= en, a2= en+1 and bn =ψ(en1), as well as

Rn={z∈C; en−bn <Rez <en+1+bn and|Imz|< bn}. We haveen−bn>en1for largen, since this amounts to

en−en1> bn= en1

A(en1), or e−1> 1 A(en1),

which holds for large n since A(t) tends to ∞ with t. We then observe that Rn⊂Ω. Indeed,z∈Rn=⇒Rez >en−bn>en1 and, sinceψis increasing, we haveψ(Rez)> ψ(en1) =bn >|Imz|. Therefore, we can apply (4.2) and get, for alln≥1:

d(en,en+1; Ω)≤ π

4ψ(en1)(en+1−en) +π

2 ≤C0A(en1) =C0An, where C0 is a numerical constant. By conformal invariance, we have as well d(rn, rn+1)≤C0An. It then follows from (2.8) that:

(4.7) 1−rn+1

1−rn ≥exp −2d(rn, rn+1)

≥exp(−2C0An).

Now, we takeh(z) =z−r1in Lemma 2.4 and use the ideal property (1.4) of the approximation numbers. We get, denoting byCthe interpolation constant of the sequence(rn), and using the fact thatkMhk=khk≤2:

(4.8) an(B)≤ kJ1kan(Cϕ)kMhk kJk ≤2C2an(Cϕ).

Next, we chooseη= 1/C0in (4.6) and we setd= (r2−r1)/√

2. Using Lemma 2.3 and relations (1.1), (2.5) and (4.7), we see that the weightswn associated with Bverify:

|wn|=h(rn+1) kKrnk

kKrn+1k =h(rn+1) s

1−r2n+1

1−rn2 ≥ r2−r1

√2

r1−rn+1

1−rn

≥dexp(−C0An)≥dεn for alln≥1.

(4.9)

Finally, using Lemma 2.4, (4.8) and (4.9):

an(Cϕ)≥ 1

2C2an(B)≥ 1 2C2d εn

def=δεn for alln≥1.

We thus get the desired conclusion (4.1) of Theorem 4.1.

(21)

5 An upper bound

We do not obtain a fairly good upper bound, and we shall content ourselves with the following result, whose proof is quite simple and, for the caseα=−1, partly contained in [35], but under a very cryptic form which is not easy to decipher.

Theorem 5.1 Let ϕ be a Schur function and α ≥ −1.Then, we have for the approximation numbers ofCϕ:Bα→Bα the upper bound:

(5.1) an(Cϕ)≤C inf

0<h<1

nα+12 (1−h)n+

ϕ,α+2(h) h2+α

, n= 1,2, . . . whereCis a constant. In particular, if ρϕ,α+2h2+α(h)≤eh/A(h), where the function A: [0,1] →[0,1] is increasing, with A(0) = 0 and with inverse function A1, we have:

(5.2) an(Cϕ)≤Cnα+12 enA−1(1/2n), n= 1,2, . . . .

The proof of (5.1) uses a contraction principle which was first proved for α=−1([18]) andα= 0([23]), but is also valid for anyα≥ −1, as follows from the forthcoming work [25].

To prove Theorem 5.1, it will be convenient to prove first the following simple lemma.

Lemma 5.2 Let nbe a positive integer, g∈Bα andf(z) =zng(z). Then, we have:

(5.3) kgkα≤Cnα+12 kfkα. Proof. Letwn= Γ(n+2+α)n!Γ(2+α) andf(z) =P

n=0anzn. We first observe that

(5.4) wk

wk+n ≤Cnα+1, ∀k≥0, ∀n≥1.

Indeed, we have:

wk

wk+n = k!

(k+n)!

Γ(k+α+ 2 +n) Γ(k+α+ 2) =

Yn j=1

(k+j+α+ 1) (k+j) ≤

Yn j=1

j+α+ 1 j

= Yn j=1

1 + α+ 1 j

≤exp

(α+ 1) Xn j=1

1 j

≤Cnα+1, which proves (5.4).

Now, if f(z) = P

k=nakzk, we have g(z) = P

k=0ak+nzk so that, using (5.4):

kgk2α= X k=0

|ak+n|2wk = X l=n

|al|2wln≤Cnα+1 X l=n

|al|2wl=Cnα+1kfk2α,

Referências

Documentos relacionados