35
CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis 1 1.2. Hardy spaces 5 1.3. Harmonic Hardy classes 7 1.4. Fatou’s theorem 9 1.5. The zero sets of functions in H p 13 1.6. Boundary behaviour of functions in Hardy spaces 16 1.7. The Cauchy–Szeg¨ o projection 19 2. Bergman spaces on the unit disc 22 2.1. Function spaces with reproducing kernel 22 2.2. The Bergman spaces 25 2.3. Biholomorphic invariance 28 2.4. L p -boundedness of a family of integral operators 29 2.5. The Bergman kernel and projection on the unit disc 33 3. The Paley–Weiner and Bernstein spaces 35 3.1. The Fourier transform 35 3.2. The Paley–Wiener theorems 40 3.3. The Paley–Wiener spaces 45 3.4. The Bernstein spaces 49 4. Function theory on the upper half plane 51 4.1. Hardy spaces on the upper-half plane 51 4.2. Factorization and boundary behaviour of functions in H p (U). 54 4.3. H 2 and the Paley–Wiener theorem revisited 55 4.4. H 1 , the atomic decomposition and the space of bounded mean oscillations 55 4.5. Weighted Bergman spaces on the upper-half plane 55 4.6. The Paley–Wiener theorem for Bergman spaces 55 5. Further topics 56 5.1. Hardy spaces on tube domains in C n 56 5.2. Weighted Bergman spaces on the unit ball in C n 56 5.3. The Cauchy integral along Lipschitz curves 56 5.4. Real variable Hardy spaces 56 5.5. Hankel and Toeplitz operators 56 References 56 Appunti per il corso Argomenti Avanzati di Analisi Complessa per i Corsi di Laurea in Matematica dell’Universit` a di Milano, a.a. 2011/12. – October 27, 2011.

CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS

MARCO M. PELOSO

Contents

1. Hardy Spaces on the Unit Disc 11.1. Review from complex analysis 11.2. Hardy spaces 51.3. Harmonic Hardy classes 71.4. Fatou’s theorem 91.5. The zero sets of functions in Hp 131.6. Boundary behaviour of functions in Hardy spaces 161.7. The Cauchy–Szego projection 192. Bergman spaces on the unit disc 222.1. Function spaces with reproducing kernel 222.2. The Bergman spaces 252.3. Biholomorphic invariance 282.4. Lp-boundedness of a family of integral operators 292.5. The Bergman kernel and projection on the unit disc 333. The Paley–Weiner and Bernstein spaces 353.1. The Fourier transform 353.2. The Paley–Wiener theorems 403.3. The Paley–Wiener spaces 453.4. The Bernstein spaces 494. Function theory on the upper half plane 514.1. Hardy spaces on the upper-half plane 514.2. Factorization and boundary behaviour of functions in Hp(U). 544.3. H2 and the Paley–Wiener theorem revisited 554.4. H1, the atomic decomposition and the space of bounded mean oscillations 554.5. Weighted Bergman spaces on the upper-half plane 554.6. The Paley–Wiener theorem for Bergman spaces 555. Further topics 565.1. Hardy spaces on tube domains in Cn 565.2. Weighted Bergman spaces on the unit ball in Cn 565.3. The Cauchy integral along Lipschitz curves 565.4. Real variable Hardy spaces 565.5. Hankel and Toeplitz operators 56References 56

Appunti per il corso Argomenti Avanzati di Analisi Complessa per i Corsi di Laurea in Matematicadell’Universita di Milano, a.a. 2011/12.

– October 27, 2011.

Page 2: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

SPACES OF HOLOMORPHIC FUNCTIONS 1

In these notes we consider spaces of holomorphic functions defined on domains of the complexplane or, toward the end of the notes, on domains of the n-dimensional complex space Cn.

We begin with the Hardy spaces on the unit disc.

1. Hardy Spaces on the Unit Disc

1.1. Review from complex analysis. Let C denote the field of complex numbers z = x+ iy,where x, y ∈ R are the real and imaginary part, respectively, of z. We will denote by D the unitdisc, that is,

D =z ∈ C : |z| < 1

.

Here and in what follows, we will denote by D(z0, r) the disc centered at z0 ∈ C and radiusr > 0; we will sometimes abbreviate Dr to denote the disc D(0, r).

If f is holomorphic on D then it can be written as sum of a converging power series

f(z) =+∞∑k=0

akzk .

From Cauchy’s formula we know that, if γ is a simple closed curve contained in D and z liesin the interior of γ (which is a well-defined domain by the Jordan curve theorem), then

f(z) =1

2πi

∫γ

f(ζ)ζ − z

dζ .

If the function f is actually holomorphic in a slightly larger disc DR with R > 1, then wemay take the unit circonference ∂D as curve of integration and obtain

f(z) =1

2πi

∫∂D

f(ζ)ζ − z

dζ ,

for all z ∈ D.One of the main themes of this course is trying to recover the holomorphic function f (in a

given class) from its boundary values, once these values are suitably defined.Setting z = reiη, we rewrite the above identity as

f(reiη) =1

∫ 2π

0f(eiθ)

eiθ

eiθ − reiηdθ

=1

∫ 2π

0f(eiθ)

11− rei(η−θ)

=1

∫ 2π

0f(ei(η−θ))

11− reiθ

dθ , (1.1)

using the periodicity of the exponential.Notice that

11− reiθ

=+∞∑n=0

rneinθ

=: Cr(eiθ) , (1.2)

where Cr(eiθ) is the Cauchy kernel (thought of as a convolution kernel, see (1.4) below).

Page 3: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

2 M. M. PELOSO

Therefore, using the uniform convergence on compact subsets of D of the power series, wehave that (1.1) gives the identity

f(reiη) =+∞∑n=0

rn1

∫ 2π

0f(eiθ)ein(η−θ) dθ

=+∞∑n=0

rneinη1

∫ 2π

0f(eiθ)e−inθ dθ . (1.3)

We recall that the unit circle ∂D = ζ ∈ C : ζ = eiθ inherits the multiplicative structurefrom C and it is homeomorphic (as a topological group) to the interval [−π, π] (or any interval oflength 2π) once we identify the two end points of the interval. This group T = R/2πZ is calledthe torus. Of course, the identification between T and ∂D is given by the mapping t 7→ eit.

Given f, g ∈ L1(T) (where, we recall, T is identified with the unit circle) the convolution onT is defined as

f ∗ g(eiη) =1

∫ 2π

0f(eiθ)g(ei(η−θ)) dθ . (1.4)

Therefore, the identity (1.1) can be expressed as convolution on the group T as

f(reiη) = (f ∗ Cr)(eiη) . (1.5)

Moreover, in (1.3) we recognize the n-th Fourier coefficient of the restriction of f on the unitcircle ∂D (function that, at this stage, is assumed to be continuous)

f(n) =1

∫ 2π

0f(eiθ)e−inθ dθ .

It is clear that in order to define the Fourier coefficients is sufficient to assume that f is anL1-function on the unit circle.

Another integral formula, with its relative integral kernel, that will play a key role in thesenotes is the Poisson reproducing formula for the unit disc D

u(reiη) =1

∫ 2π

0u(eiθ)

1− r2

|1− rei(η−θ)|2dθ

= (Pu)(reiη) , (1.6)

where u is a function harmonic in a domain containing D and 0 < r < 1. The operator P iscalled the Poisson operator and Pu the Poisson integral of the function u defined on the unitcircle.

More generally, the solution of the Dirichlet problem on the unit D (see [CA-notes]) showsthat if we start with a continuous function g on the unit circle ∂D, the function u := Pg isharmonic in D, continuous on the closure of the unit disc and coincides with g on the unit circle.

It is worth to observe that Pg is well defined for all g ∈ L1(T) and that Pg is a harmonicfunction in D.

One of the main questions that we are going to address is if, and in which sense, the functionPg admits boundary values and, in this case, if such values coincide with the function g. In

Page 4: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

SPACES OF HOLOMORPHIC FUNCTIONS 3

other words, a classical question is whether the Dirichlet problem (for the Laplacian on the unitdisc)

∆f = 0 on Df = g on ∂D ,

(1.7)

where g ∈ Lp(∂D) is an assigned function, 1 ≤ p <∞, admits a solution and in what sense thesolution f admits g as values on the boundary.

As in the case of the reproducing formula (1.5) we may think of the Poisson integral as aconvolution integral operator on T by writing

Pr(eiθ) =1− r2

|1− reiθ|2, (1.8)

so that the Poisson integral of a function g ∈ L1(T) is defined by the formula

(Pg)(reiη) =1

∫ 2π

0g(eiθ)

1− r2

|1− rei(η−θ)|2dθ = (Pr ∗ g)(eiη) . (1.9)

In complex analysis we were used to think of the Poisson kernel, that we denote temporarelyas Pz(eiθ) as a function of the variables z = reiη ∈ D and eiθ ∈ ∂D, where

Pz(eiθ) =1− r2

|eiθ − reiη|2=

1− r2

|1− rei(η−θ)|2(1.10)

=1− r2

1− 2r cos(θ − η) + r2. (1.11)

In the current setting it is more convenient to think of the Poisson kernel as a family offunctions Pr defined on the torus T, hence depending on the variable eiη, and of the Poissonintegral of a function g as the convolution of g with Pr, as in (1.9). According to this point ofview, the collection of functions Pr0<r<1 on T is a summability kernel (or a approximationof the identity), A family of functions Φtt>0 is called a summability kernel if they satisfy thefollowing properties:

(1)1

∫ 2π

0Φt(eiθ) dθ = 1 for t > 0;

(2) ‖Φt‖L1 ≤ C with C independent of t;

(3) for every δ > 0, limt→+∞

∫δ≤|θ|≤π

|Φt(eiθ)| dθ = 0.

We recall that if Φtt>0 is a summability kernel on T and g ∈ Lp(T), 1 ≤ p < ∞, thenΦt ∗ g → g in the Lp-norm, and that if g ∈ C(T), then Φt ∗ g → g in the sup-norm.

We collect here the basic properties of the Poisson kernel and integral.

Proposition 1.1. The function P (reiη) = Pr(eiη) is harmonic in D as function of z = reiη

and1

∫ 2π

0Pr(eiη) dη = 1

for all 0 < r < 1.Moreover, as collection Pr of functions on T we have

(i) for 1 ≤ p <∞, ‖f ∗ Pr‖Lp ≤ ‖f‖Lp;(ii) for 1 ≤ p <∞, limr→1− ‖f ∗ Pr − f‖Lp = 0;

Page 5: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

4 M. M. PELOSO

(iii) if g ∈ C(T), then limr→1− ‖g ∗ Pr − g‖L∞ = 0.Finally, for f ∈ Lp(T), 1 ≤ p ≤ ∞, P(f) is harmonic in D.

Proof. It is clear that it suffices to show Pr is a summability kernel and that P (reiη) isharmonic in the variable z = reiη ∈ D.

The latter fact is well known, as well as that 12π

∫ 2π0 Pr(eiη) dη = 1.

In order to prove that Pr is a summability kernel, it remains to show that condition (3) issatisfied. This is easily verified and we leave the details to the reader. 2

Lemma 1.2. For 0 < r < 1 we have that

Pr(eiη) =+∞∑

k=−∞r|k|eikη ,

and the convergence is uniform on T.

Proof. Recall that we always identify the unit circle ∂D with T.We can write

1− r2

|1− reiη|2=

11− reiη

+1

1− re−iη− 1 (1.12)

=+∞∑k=0

rkeikη ++∞∑k=0

rke−ikη − 1

=+∞∑

k=−∞r|k|eikη .

The statement about the uniform convergence is clear, by the Weierstrass M -test. 2

Notice that by (1.12) we obtain a relation between the Poisson and Cauchy kernels, that is,

Pr = 2ReCr − 1 . (1.13)

Lemma 1.3. Let g ∈ L1(∂D). Then for every 0 < r < 1 we have that(Pr ∗ g

)(eiη) =

+∞∑k=−∞

g(k)r|k|eikη .

Proof. Using the uniform convergence of the series development of Pr we see that(Pr ∗ g

)(eiη) =

12π

∫ 2π

0g(eiθ)

+∞∑k=−∞

r|k|eik(η−θ) dθ

=+∞∑

k=−∞

( 12π

∫ 2π

0g(eiθ)e−ikθ dθ

)r|k|eikη

=+∞∑

k=−∞g(k)r|k|eikη ,

and the convergence is uniform on T. 2

Page 6: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

SPACES OF HOLOMORPHIC FUNCTIONS 5

1.2. Hardy spaces. We are finally ready to define the classical Hardy spaces. We will denoteby H(Ω) the space of holomorphic functions on a given domain Ω.

Definition 1.4. For 0 < p <∞ and 0 < r < 1, for a function f defined on D we set

Mp(f, r) =( 1

∫ 2π

0|f(reiθ)|p dθ

)1/p.

We define the Hardy space Hp = Hp(D) as

Hp =f ∈ H(D) : sup

0<r<1Mp(f, r) <∞

,

and for f ∈ Hp we set‖f‖Hp = sup

0<r<1Mp(f, r) .

Furthermore, we define H∞ as the space of holomorphic functions that are bounded on the unitdisc, endowed with the sup-norm.

We mention in passing that, also when 0 < p < 1 we set

‖f‖Lp =( 1

∫ 2π

0|f(eiθ)|p dθ

)1/p

and, with an abuse of language, we call it the Lp-norm. However, setting d(f, g) = ‖f − g‖pLpLp becomes a complete metric space. Notice that ‖f + g‖Lp ≤ 2(1−p)/p(‖f‖Lp + ‖g‖Lp).

Remark 1.5. Aside from the case p = ∞, also the space H2 can be described at once. For,if f is holomorphic on D, then it admits power series expansion f(z) =

∑+∞n=0 anz

n. Using theuniform convergence on compact subsets of D we have

M2(f, r)2 =1

∫ 2π

0|f(reit)|2 dt

=1

∫ 2π

0

+∞∑n,m=0

anamrn+meit(n−m) dt

=+∞∑n=0

r2n|an|2 .

Therefore,

sup0<r<1

M2(f, r) =( +∞∑n=0

|an|2)1/2

,

that is, f ∈ H2 if and only if∑+∞

n=0 |an|2 is finite. In particular, it follows that H2 is a Hilbertspace.1

Observe that, in particular, a function f in H2 can be extended to the boundary ∂D, havingas “boundary values” the function f ∈ L2(∂D) given by

f(eit) =+∞∑n=0

aneint .

1Exercise I.2.

Page 7: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

6 M. M. PELOSO

Moreover, by Lemma 1.3, P(f)(r·) = Pr∗f =∑+∞

n=0 anrnein(·) = f(r·). We can call f “boundary

values” since f(r·) = Pr ∗ f → f in L2(T) as r → 1−.This fundamental relation is a first indication of the deep connection between the theories of

Fourier series and of Hardy spaces.

Remark 1.6. For 0 < p < q ≤ ∞ we have 0 ( Hq ( Hp.For, notice that for 0 < p < q ≤ ∞ we clearly have Hq ⊆ Hp. Moreover, all holomorphic

polynomials are bounded, hence in all the Hp spaces and these are non-trivial. Moreover, usingLemma 2.15 (that we will see later on) one can show that

f(z) =1

(1− z)a∈ Hp if and only if p <

1a.

In these notes we will essentially restrict ourselves to the case 1 ≤ p ≤ ∞, although the Hardyspaces turn out to be of great interest also and, from a certain point of view, especially, in thecase 0 < p ≤ 1. We point out that the Hardy space H1 can be considered as a limit case forboth theories 1 < p ≤ ∞ and 0 < p < 1 and we will say more about it in Section 4.4 when wedeal with the Hardy spaces on the upper half-plane.

The first observation we make is that the “sup” in the definition of the Hp-norm is actuallya “lim”.

Remark 1.7. Given f holomorphic in D, the function Mp(f, r) is increasing in r, 0 < r < 1.This statement holds true for the full range 0 < p ≤ ∞, but its proof is elementary only in

the case p ≥ 1 and we will restrict to this case.For 0 < % < 1 and a function f defined in D we write f% := f(%·) to denote a function on the

unit circle. For 0 < r, % < 1, given a function f holomorphic in D, notice that fr is holomorphicin a ngbh of D so that

f(r%eiη) = (fr ∗ P%)(eiη) .Hence, we have

Mp(f, r%) = ‖fr ∗ P%‖Lp(T)

≤ ‖fr‖Lp(T) ‖P%‖L1(T) = Mp(f, r) ,

that is, Mp(f, r) is increasing in r and

sup0<r<1

Mp(f, r) = limr→1−

Mp(f, r) .

Proposition 1.8. For 1 ≤ p ≤ ∞ Hp is a Banach space.

Proof. It is clear that ‖ · ‖Hp is a norm, so we only need to prove that it is complete.Let 0 < r, % < 1 and notice that for f holomorphic on D we have

|f(r%eit)| =∣∣fr ∗ P%(eit)∣∣ ≤ ‖fr‖Lp(T)‖P%‖Lp′ (T)

≤ C%‖f‖Hp .

This shows thatsup|z|≤%|f(z)| ≤ C%‖f‖Hp ,

Page 8: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

SPACES OF HOLOMORPHIC FUNCTIONS 7

and therefore the convergence in the Hp-norm implies the uniform convergence on compactsubsets.

Thus, let fn be a Cauchy sequence in the Hp-norm, and let f be the function uniform limiton compact subsets of D. Then f is holomorphic and for 0 < r < 1 fixed, using the uniformconvergence, for p <∞ we have

Mp(fn − f, r)p =1

∫ 2π

0|fn(reit)− f(reit)|p dt

= limm→+∞

12π

∫ 2π

0|fn(reit)− fm(reit)|p dt

= limm→+∞

Mp(fn − fm, r)p

≤ limm→+∞

‖fn − fm‖pHp .

Therefore,‖fn − f‖Hp ≤ lim

m→+∞‖fn − fm‖Hp < ε ,

for n sufficiently large. The case p =∞ is similar and we leave the details to the reader.2 2

1.3. Harmonic Hardy classes. We now analyze the action of the Poisson integral on thespaces Lp(T). This analysis has interest on its own right, but also will serve for studying theboundary behaviour of functions in (the holomorphic) Hardy spaces.

We consider now harmonic functions on D that satisfy the Hp-growth condition, 0 < p <∞,that is,

sup0<r<1

Mp(f, r) <∞ , (1.14)

and in an obvious way when p =∞.

Proposition 1.9. Let f be harmonic in D and satisfy (1.14), 1 < p ≤ ∞. Then there existsf ∈ Lp(∂D) such that f = P(f), that is,

f(reit) = (f ∗ Pr)(eit) =1

∫ 2π

0f(eiθ)

1− r2

|1− rei(t−θ)|2dθ .

Moreover, if 1 < p <∞, fr → f in the Lp-norm.

We observe that this statement is false when p = 1. For instance, f(reiη) = Pr(eiη) satisfiesthe hypotheses with p = 1, but there exists no L1-function g on ∂D such that Pr → g andPr ∗ g = Pr. We leave the details as an exercise.3

Proof. Let fr = f(r·), 0 < r < 1. Let rn be a strictly increasing sequence in (0, 1), rn → 1−.Then frn is a bounded set in Lp(T). By the Banach-Alaoglu theorem (see [Ru]), there exists

a subsequence frnj converging in the weak-∗ topology to a funtion f ∈ Lp(T), 1 < p ≤ ∞,

that is, for every g ∈ Lp′(T) we have that∫ 2π

0frnj (e

iθ)g(eiθ) dθ →∫ 2π

0f(eiθ)g(eiθ) dθ as j → +∞ .

2Exercise I.4.3Exercise I.5.

Page 9: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

8 M. M. PELOSO

(Here we need the restriction p > 1, since the Banach-Alaoglu theorem requires that the spacecontaining fr to be the dual space of another Banach space– and L1(T) is not.)

Let 0 < r < 1 and let j0 be such rnj > r for j ≥ j0. Then, frnj is continuous on D andharmonic in D, hence it satisfies the Poisson reproducing formula

f(reit) = frnj((r/rnj )e

it)

= (frnj ∗ Pr/rnj )(eit)

=1

∫ 2π

0frnj (e

iθ)Pr/rnj (ei(t−θ)) dθ

=1

∫ 2π

0frnj (e

iθ)Pr(ei(t−θ)) dθ +1

∫ 2π

0frnj (e

iθ)[Pr/rnj (e

i(t−θ))− Pr(ei(t−θ))]dθ .

(1.15)

Since Pr ∈ C(∂D) ⊂ Lp′(∂D) for all 0 < r < 1, the first term on the right hand side above

converges to1

∫ 2π

0f(eiθ)Pr(ei(t−θ)) dθ .

Now we show that the second term on the right hand side of (1.15) tends to 0 as j → +∞. For,gj := Pr/rnj − Pr → 0 in the Lp

′-norm, since gj ∈ C(T) and tend to 0 uniformly. Therefore,∣∣∣ ∫ 2π

0frnj (e

iθ)gj(ei(t−θ)) dθ∣∣∣ ≤ ‖frnj ‖Lp‖gj‖Lp′ ≤ C‖gj‖Lp′ → 0

as j → +∞. Notice that we have used the assumption that f satisfies (1.14).From (1.15) it follows that

f(reit) =1

∫ 2π

0f(eiθ)Pr(ei(t−θ)) dθ ,

as we wished to show.The last part of the statement is clear: fr = f ∗ Pr as functions on ∂D and since Pr is a

summability kernel, fr = f ∗ Pr → f in the Lp-norm, for 1 < p <∞. 2

Remark 1.10. Notice that, in particular Prop. 1.9 applies to Hp-functions.Given f ∈ Hp, we have constructed the boundary value function f as weak-∗ limit of a

sequence fnj, where nj → 1−. It is easy to see that this construction is independent of thechoice of the sequence fnj. (This is obvious in particular when 1 < p <∞ since f is the limitin the Lp-norm of the whole family fr. )

Moreveor, if fr → g in Lp(∂D), it also easy to see that g = f . In particular, the boundaryvalue function of an H2-function identified in Remark 1.5 coincides with the one constructed inProp. 1.9.

We now prove a partial converse of Prop. 1.9.

Proposition 1.11. Let g ∈ Lp(∂D), 1 ≤ p < ∞. Then the Poisson integral of g, P(g) isharmonic in D, satisfies the Hp-growth condition (1.14), and (in the notation of Prop. 1.9)

P(g) = g .

Page 10: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

SPACES OF HOLOMORPHIC FUNCTIONS 9

Proof. Notice that, as functions on ∂D, fr = f(r·) := g ∗Pr satisfy the norm estimate ‖fr‖Lp ≤‖g‖Lp‖Pr‖L1 = ‖g‖Lp . Therefore,

sup0<r<1

Mp(f, r) ≤ ‖g‖Lp .

This shows that the harmonic function P(g) satisfies the Hp-growth condition (1.14).Next, if 1 < p <∞, by the previous proposition, f := P(g) admits a boundary value function

f such that P(f) = f , that is, P(f − g) = 0. Therefore, Pr ∗ (f − g) = 0 in Lp(T) for all0 < r < 1, since Pr is a summability kernel, 0 = Pr ∗ (f − g)→ f − g in Lp(T). (Notice thathere we need to restrict to the case p <∞.)

Finally, in the case p = 1, we simply notice that fr := Pr ∗ g → g in L1(∂D) as r → 1−, sothat we may define again the “boundary values” of f as the limit of the family fr, that is,f = g. This concludes the proof. 2

Remark 1.12. Recall the Dirichlet problem (1.7). If the boundary datum g ∈ C(∂D) we knowthat the solution f = P(g) is harmonic, continuous in the closure of the unit disc and f restrictedto ∂D coincides with g (e.g. see [CA-notes]).

Prop. 1.11 can be restated by saying that if g ∈ Lp(∂D), 1 ≤ p < ∞, then f = P(g) isharmonic in D and can be extdended to a boundary value function f that coincides with g.Then, we say that f = P(g) is (the) solution of the Dirichlet problem (1.7) with boundarydatum in Lp.

We conclude this part with the following consequence of the last two propositions.

Corollary 1.13. Define

hp(D) =f harmonic on D : ‖f‖hp := sup

0<r<1Mp(f, r) <∞

.

Then, for 1 < p <∞,P : Lp(∂D)→ hp(D)

is an isometric isomorphism.

1.4. Fatou’s theorem. Perhaps, the main result about the Hardy spaces is that if f ∈ Hp with0 < p ≤ ∞, then f , as in the case p = 2, admits “boundary values”, that is,

limr→1−

f(reit) = f(eit) exists for a.e. t ∈ [0, 2π) ,

and f ∈ Lp(∂D). Actually, more is true: the function f converges to f when z approaches apoint eit ∈ ∂D, for a.e. t ∈ [0, 2π), within the non-tangential approach regions.

Definition 1.14. We define the subset of the unit disc

Γα(eit) =z ∈ D : |z − eit| < α(1− |z|)

. (1.16)

The set Γα(eit) is also called the Stoltz region, with vertex in eit and aperture α > 1.

Remark 1.15. We need to make a few comments on the definition of Γα(eit).(i) It is easy to see that the regions Γα(eit) are increasing with α and that they would be

empty for α ≤ 1.

Page 11: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

10 M. M. PELOSO

(ii) The region Γα(eit) is also called a non-tangential approach region, since outside a fixedcompact subset of D, it is contained in a cone with vertex in eit and aperture Cαα, where Cα > 1is a suitable constant.

For, assume that 1 − εα ≤ r < 1, where cα is a (small) positive constant, that we are goingto fix momentarily, and let z = reiη ∈ Γα(eit). Then

|reiη − eit| =∣∣eiη − eit − (1− r)eiη

∣∣ ≥ |eiη − eit| − (1− r) ,

so that|eiη − eit| ≤ (α+ 1)(1− r) .

If cα is chosen sufficientily small, then we must have |η − t| < π/4 (that is, z ∈ Γα(eit) andr ≥ 1− cα implies |η − t| < π/4).

Next, |eiη − eit| = 2 sin(|η − t|/2

)≥ 2

√2

π |η − t|, as long as |η − t| ≤ π/4, which is the case.Therefore, z = reiη ∈ Γα(eit) and r ≥ 1− cα implies

|η − t| ≤ π

2√

2(α+ 1)(1− r) = Cαα(1− r) , (1.17)

as claimed. In particular, we may take Cα = 3.

Theorem 1.16. (Fatou) Let f ∈ Hp, 1 < p ≤ ∞. Then

limΓα(eit)3z→eit

f(z) = f(eit) ,

where f is the function as in Prop. 1.9, and Γα(eit) is as in (1.16).

In order to prove Fatou’s theorem, we need to prove a result concerning the Hardy–Littlewoodmaximal function, defined (in the case of the torus T) as

Mf(eit) = sup0<δ≤π

12δ

∫|θ|<δ

|f(ei(t−θ))| dθ .

We recall that M is weak-type (1, 1), that is, it satisfies the following estimate: there exists aconstant C > 0 such that for every λ > 0,∣∣t : Mf(eit) > λ

∣∣ ≤ C

λ‖f‖L1(T) .

The next result shows the reason for our interest in the Hardy–Littlewood maximal function.

Proposition 1.17. Let g ∈ L1(∂D), α > 1 a fixed parameter. Then there exists Cα > 0 suchthat for every eit ∈ ∂D

supz∈Γα(eit)

|P(g)(z)| ≤ CαMg(eit) .

Proof. The proof is divided in a few steps.

(Step 1.) Assume first that z = reiη varies in a fixed compact subset of D, namely assumethat 1− r ≥ cα, with cα as in Remark 1.15 (ii).

Notice that for all θ

Pr(eiθ) =1− r2

|1− reiθ|2≤ 1− r2

(1− r)2≤ 2

1− r. (1.18)

Page 12: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

SPACES OF HOLOMORPHIC FUNCTIONS 11

Then, for z = reiη as above,

|P(g)(z)| = |(g ∗ Pr)(eiη)| ≤1

∫ 2π

0|g(ei(η−θ))|Pr(eiθ) dθ

≤ 2cα

12π

∫ 2π

0|g(ei(η−θ))| dθ

=2cα

12π

∫ 2π

0|g(ei(t−θ))| dθ

≤ CαMg(eit) ,

where we have used the translation invariance of the integral on T.

(Step 2.) Now we assume that z = reiη ∈ Γα(eit) with 1 − r < cα. By Remark 1.15 (ii) weknow that

|η − t| ≤ 3α(1− r) .

We break the integral on T as follows:

|P(g)(z)| = |(g ∗ Pr)(eiη)| ≤1

∫ 2π

0|g(ei(η−θ))|Pr(eiθ) dθ

≤N+1∑j=0

∫Ej

|g(ei(η−θ))|Pr(eiθ) dθ ,

where

E0 =θ ∈ T : |θ| ≤ 3α(1− r)

;

for j = 1, . . . , N , Ej =θ ∈ T : 2j−13α(1− r) ≤ |θ| < 2j3α(1− r)

and N is chosen so

that ; 2N−13α(1− r) ≤ π/4 < 2N3α(1− r); EN+1 =

θ ∈ T : π/4 < |θ| ≤ π

.

Notice that, for θ ∈ Ej (which implies in particular that |θ| ≤ π/4), arguing as in Remark1.15 (ii) we have that

|1− reiθ| ≥ 2jα(1− r) .

Therefore, for θ ∈ Ej we have that

Pr(eiθ) =1− r2

|1− reiθ|2≤ 1− r2

22jα2(1− r)2

≤ 21

22jα2(1− r).

Finally, for θ ∈ EN+1, |1− reiθ| ≥ 1/√

2 so that

Pr(eiθ) ≤ 2(1− r2) ≤ 2 .

(Step 3.) Having set up the above bounds, we finally come to the estimate in the statement.

Page 13: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

12 M. M. PELOSO

Using the inequalities in Step 2 we have that∫E0

|g(ei(η−θ))|Pr(eiθ) dθ ≤2

1− r

∫|θ|≤3α(1−r)

|g(ei(η−θ))| dθ

≤ Cα1

6α(1− r)

∫|θ|≤3α(1−r)

|g(ei(t−θ))| dθ

≤ CαMg(eit) ,

using the change of variables θ 7→ θ − t+ η.Next, using again the change of variables θ 7→ θ − t+ η,∫

EN+1

|g(ei(η−θ))|Pr(eiθ) dθ ≤ 21

∫|θ|≤π

|g(ei(η−θ))| dθ

= 21

∫|θ|≤π

|g(ei(t−θ))| dθ

≤ cMg(eit) .

Finally, we estimate the central part of the integral:

N∑j=1

∫Ej

|g(ei(η−θ))|Pr(eiθ) dθ ≤ 2N∑j=1

122jα2(1− r)

∫Ej

|g(ei(η−θ))| dθ

≤ 2N∑j=1

122jα2(1− r)

∫|θ|≤2j3α(1−r)

|g(ei(η−θ))| dθ

≤ 2N∑j=1

122jα2(1− r)

∫|θ|≤2j+13α(1−r)

|g(ei(t−θ))| dθ

≤ CαN∑j=1

2−j1

2j+13α(1− r)

∫|θ|≤2j+13α(1−r)

|g(ei(t−θ))| dθ

≤ CαN∑j=1

2−jMg(eit)

≤ CαMg(eit) .

This concludes the proof. 2

We are now ready to prove Thm. 1.16.

Proof of Thm. 1.16. Since H∞ ⊂ Hp for all p <∞, it suffices to deal with the case p <∞.Let f ∈ Lp(∂D) be as in Prop. 1.9 (so that P(f) = f) and for a given γ > 0 let g ∈ C(∂D) be

such that ‖f−g‖Lp ≤ γ; notice that here we need p <∞. By the solution of the Dirichlet problemwe know that P(g)(z)→ g(eit) as D 3 z → eit (even without the restriction z ∈ Γα(eit)).

Page 14: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

SPACES OF HOLOMORPHIC FUNCTIONS 13

We use Prop. 1.17 and argue as in the proof of the Lebesgue differentiation theorem,∣∣∣eit ∈ ∂D : lim supΓα(eit)3z→eit

|f(z)− f(eit)| ≥ ε∣∣∣

≤∣∣∣eit ∈ ∂D : lim sup

Γα(eit)3z→eit|f(z)− P(g)(z)| ≥ ε/3

∣∣∣+∣∣∣eit ∈ ∂D : lim sup

Γα(eit)3z→eit|P(g)(z)− g(eit)| ≥ ε/3

∣∣∣+∣∣∣eit ∈ ∂D : |g(eit)− f(eit)| ≥ ε/3

∣∣∣≤∣∣∣eit ∈ ∂D : CαM(f − g)(eit)| ≥ ε/3

∣∣∣+∫eit∈∂D:|g(eit)−f(eit)|≥ε/3

(3|g(eit)− f(eit)|ε

)pdt

≤ C

ε‖g − f‖L1 +

C ′

εp‖g − f‖pLp ≤

C ′′

εpγp

≤ Cε ,

if we choose γ sufficiently small. This shows that∣∣∣eit ∈ ∂D : lim supΓα(eit)3z→eit

|f(z)− f(eit)| ≥ ε∣∣∣ ≤ Cε ,

which implies thatlim

Γα(eit)3z→eitf(z) = f(eit) ,

a.e., and the statement follows. 2

1.5. The zero sets of functions in Hp. In this section we describe the zero sets of functionsin the Hardy spaces, for the full range 0 < p ≤ ∞. Surprisingly, it turns out that these arethe same for all values of p. The description of such sets will allow us to prove a factorizationtheorem for Hp-functions that will have important consequences.

We begin by recalling a result from complex analysis.

Proposition 1.18. (Jensen’s formula) Let r > 0 and let f be holomorphic in a ngbh ofD(0, r). Let ζ1, . . . , ζN be the zeros of f in D(0, r), counting multiplicity. Assume that f doesnot vanish at the origin and on the circle of radius r. Then

log |f(0)|+ logN∏j=1

r

|ζj |=

12π

∫ 2π

0log |f(reiθ)| dθ .

For a proof we refer the reader to any classical text in complex analysis, as well as for thetheory of infinite products.

A consequence of this result is the following corollary.

Corollary 1.19. Let f ∈ H(D), f(0) 6= 0 and ζ1, ζ2, . . . its zeros counting multiplicity. Then,

log |f(0)|+ log+∞∏j=1

1|ζj |≤ sup

0<r<1

12π

∫ 2π

0log+ |f(reiθ)| dθ ,

where, for t > 0, log+ t = max(log t, 0).

Page 15: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

14 M. M. PELOSO

Proof. Clearly, for every 0 < r < 1,

log |f(0)|+ logN∏j=1

r

|ζj |≤ 1

∫ 2π

0log+ |f(reiθ)| dθ

≤ sup0<r<1

12π

∫ 2π

0log+ |f(reiθ)| dθ .

Letting r → 1− we obtain the result. 2

Corollary 1.20. Let f ∈ Hp, 0 < p ≤ ∞, f 6≡ 0, and let ζ1, ζ2, . . . its zeros countingmultiplicity. Then

+∞∑j=1

(1− |ζj |) <∞ . (1.19)

Proof. If f vanishes of order k at the origin, we consider f(z)/zk, which is still in Hp and it iszero-free at the origin. Notice that

0 ≤ log+ |f(reiθ)| ≤ |f(reiθ)|p .Then we applying Cor. 1.19 to obtain that

0 < log+∞∏j=1

1|ζj |

<∞ ,

which implies that∏+∞j=1

1|ζj | converges; hence

∏+∞j=1 |ζj | converges. But this is equivalent to∑+∞

j=1(1− |ζj |) <∞. 2

We recall that for ζ ∈ D the function

ϕζ(z) =z − ζ1− ζz

is a biholomorphic mapping of D onto itself. In this setting ϕζ is also called a Blaschke factor.

Proposition 1.21. Let ζ1, ζ2, . . . be points in D satisfying condition (1.19). Then the infiniteproduct

+∞∏j=1

−ζj|ζj |

ϕζj

converges uniformly on compact subsets of D.

Proof. It suffices to show that, for r < 1 be fixed and |z| ≤ r, the series+∞∑j=1

∣∣∣1 +ζj|ζj |

ϕζj (z)∣∣∣ <∞

converges uniformly. A simple calculation now shows that∣∣∣1 +ζj|ζj |

ϕζj (z)∣∣∣ =

∣∣|ζj |(1− |ζj |) + ζjz(1− |ζj |)∣∣

|ζj ||1− ζjz|

≤ 1 + r

1− r(1− |ζj |) .

Page 16: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

SPACES OF HOLOMORPHIC FUNCTIONS 15

By the Weierstrass M -test the convergence is uniform in |z| ≤ r, and the conclusion follows. 2

Definition 1.22. Let 0 < p ≤ ∞, f ∈ Hp and let ζ1, ζ2, . . . its zeros different from the origin,counting multiplicity. We set

B(z) = zk+∞∏j=1

−ζj|ζj |

ϕζj (z) ,

where k ≥ 0 is the order of zero of f at the origin. Then, B is a well-defined holomorphicfunction on D, and there exists a holomorphic, zero-free function F on D such that f/B = F .

The function B is called a Blaschke product and f = FB is called the canonical factorizationof f ∈ Hp.

Proposition 1.23. Let 0 < p ≤ ∞, f ∈ Hp and let f = FB be its canonical factorization.Then F ∈ Hp and ‖F‖Hp = ‖f‖Hp.

Proof. Notice that B = zk∏+∞j=1 −

ζj|ζj |ϕζj and every factor satisfies

∣∣ ζj|ζj |ϕζj

∣∣ ≤ 1, so that |B| ≤ 1.Therefore, |f(z)| ≤ |F (z)| for all z ∈ D, so that ‖f‖Hp ≤ ‖F‖Hp .

We set BN =∏Nj=1−

ζj|ζj |ϕζj and FN = f/BN . We recall that ϕζj is defined and holomorphic

in the disc D(0, 1/|ζj |) which is strictly larger than the unit disc D and that |ϕζj (eiη)| = 1 forall η. Therefore, BN converge uniformly on D to a function BN as r → 1−, with |BN | = 1 on∂D.

Hence,

‖FN‖pHp = limr→1−

Mp(FN , r)p

= limr→1−

12π

∫ 2π

0|(f/BN )(reiθ)|p dθ

≤ limr→1−

1infθ |BN (reiθ)|

12π

∫ 2π

0|(f)(reiθ)|p dθ

= ‖f‖pHp .

Moreover,

Mp(F, r)p = limN→+∞

Mp(FN , r)p

≤ limN→+∞

‖FN‖pHp = ‖f‖pHp .

Therefore, ‖F‖Hp ≤ ‖f‖Hp , and the equality follows. 2

Corollary 1.24. Let ζ1, ζ2, . . . be points in D satisfying the condition∑+∞

j=1(1 − |ζj |) < ∞

and let B = zk∏+∞j=1 −

ζj|ζj |ϕζj be the corresponding Blaschke product. Then B ∈ H∞ and |B| = 1

a.e. on ∂D.

Proof. In the proof of Prop. 1.23 we have already noticed that B ∈ H∞ and that its boundaryvalue function B is such that |B| ≤ 1.

Since B ∈ H∞, its canonical factorization is B = 1 ·B. By Prop. 1.23 again it follows that

1 = ‖1‖2H2 = ‖B‖2H2 =1

∫ 2π

0|B(eiθ)|2 dθ .

Page 17: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

16 M. M. PELOSO

Together with the estimate |B| ≤ 1, it implies that |B| = 1 a.e. on ∂D. 2

We conclude this part with the extension of Thm. 1.16 to the range 0 < p ≤ 1.

Theorem 1.25. (Fatou) Let f ∈ Hp, 0 < p ≤ 1. Then

limΓα(eit)3z→eit

f(z) =: f(eit) ,

where Γα(eit) is as in Def. 1.16, exists a.e. The function f ∈ Lp(∂D) and

‖f‖Lp = ‖f‖Hp .

Proof. Notice that, for this range of p’s, the existence of the boundary values f of f is part ofthe statement, as well as the equality of the norms of f and f .

By Prop. 1.23, f admits the canonical factorization f = FB, and the function F ∈ Hp and iszero-free. Therefore, F p/2 is well defined, holomorphic, and in H2. By Thm. 1.16, F p/2 admitsboundary values F p/2 a.e. Therefore, also F admits non-tangential boundary values a.e.

By Cor. 1.24 B admits boundary values B and |B| = 1 a.e. Hence f admits non-tangentialboundary values f =

(F p/2

)2/pB. Finally,

‖f‖pLp =1

∫ 2π

0

∣∣(F p/2)2/pB(eiθ)|p dθ

=1

∫ 2π

0

∣∣F p/2(eiθ)∣∣2 dθ = ‖F p/2‖2H2

= sup0<r<1

12π

∫ 2π

0|F (reiθ)|p dθ

= ‖F‖pHp = ‖f‖Hp ,

by Prop. 1.23 again. This proves the theorem. 2

1.6. Boundary behaviour of functions in Hardy spaces. The goal of this section is todescribe the boundary behaviour of functions in Hp and to characterize them in terms of theFourier coefficients of their boundary values.

In the previous Section 1.3 we studied the classes of functions that are Poisson integrals ofLp-functions on ∂D. In this case, we had to restrict to the case p > 1.

In the section we study the boundary behaviour of functions in Hp, and we will be ableto remove the restriction p > 1. In this analysis, we will consider the action of the Cauchykernel, that is more closely related to Hp-functions, since, in particular, it produces holomorphicfunctions, unlike the Poisson integral.

In the previous section we learned that for 0 < p ≤ ∞, if f ∈ Hp, then f converges pointwisenon-tangentially a.e. to a boundary function that, with a momentary abuse of notation, wedenote by f . We also know that fr → f in Lp(∂D) when 1 < p < ∞. We now show that thisholds true also in the case 0 < p ≤ 1.

Proposition 1.26. Let 0 < p ≤ 1, f ∈ Hp, and f be its boundary value function, as in Thm.1.25. Then fr → f in Lp as r → 1−.

Page 18: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

SPACES OF HOLOMORPHIC FUNCTIONS 17

Proof. We first prove the case p = 1.Let f ∈ H1 and let f = BF its canonical factorization. Set

g = F 1/2 , h = BF 1/2 . (1.20)

Then both g and h are in H2 and gr → g, hr → h in L2(∂D) as r → 1−.Then,

‖fr − f‖L1 = ‖grhr − gh‖L1 ≤ ‖(gr − g)hr‖L1 + ‖g(hr − h)‖L1

≤ ‖gr − g‖L2‖hr‖L2 + ‖g‖L2‖hr − h‖L2 → 0

as r → 1−. This proves the case p = 1.When 0 < p < 1 we proceed similarly, but we have to use an iteration process. Let f ∈ Hp

and define g and h as before. Then, g, h ∈ H2p. If the results holds true in H2p, that is, gr → g,hr → h in L2p(∂D) then the previous arguments applies (with a constant appearing in thetriangular inequality) and therefore the results holds true in Hp. Thus, after k ≥ log2 1/p stepswe can prove the results for Hp. 2

Recall that we denote by Cr(eiθ) the Cauchy kernel. We will denote by C(g) the Cauchyintegral of a given function g on ∂D. Recall that, for g ∈ C(∂D), C(g) is a holomorphic in Dand arguing as in (1.3) (using the uniform convergence of the power series expansion of Cr) wehave that

C(g)(reiη) = (Cr ∗ g)(eiη) =+∞∑n=0

g(n)rneinη .

Hence, we see that if g ∈ C(∂D) and g(n) = 0 when n < 0, then C(g) = P(g).More generally we have

Lemma 1.27. Let g ∈ Lp(∂D), 1 ≤ p ≤ ∞. Then P(g) = C(g) if and only if g(n) for all n < 0.

Proof. It suffices to assume that g ∈ L1(∂D).We use the uniform convergence of the power series expansion of Pr and Cr. We have

P(g)(reiη) = (Pr ∗ g)(eiη) =1

∫ 2π

0g(eiθ)

+∞∑n=−∞

rnei(η−θ) dθ

=+∞∑

n=−∞g(n)rneinη .

Analogously,

C(g)(reiη) = (Cr ∗ g)(eiη) =1

∫ 2π

0g(eiθ)

+∞∑n=0

rnei(η−θ) dθ

=+∞∑n=0

g(n)rneinη .

It follows that P(g) = C(g) if and only if g(n) for all n < 0. 2

Page 19: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

18 M. M. PELOSO

Theorem 1.28. Let f ∈ Hp, 1 ≤ p ≤ ∞ and let f be its boundary value function. Thenˆf(n) = 0 for n < 0 and

P(f) = C(f) = f .

On the other hand, if 1 ≤ p ≤ ∞ and g ∈ Lp(∂D) is such that g(n) = 0, then f := P(g) = C(g)is in Hp and f = g.

Proof. Let f(z) =∑+∞

n=0 anzn be the power series expansion of f in D, f ∈ Hp, 1 ≤ p ≤ ∞.

Notice that it suffices to consider the case p = 1. Then

fr(eiη) =+∞∑n=0

anrneinη

and for all 0 < r < 1, fr(n) = anrn, by the uniform convergence. Since by Prop. 1.26 fr → f

in L1(∂D), fr(n)→ f(n) as r → 1−, so that

f(n) =

an for n ≥ 00 for n < 0 .

This clearly implies that

f(reiη) =+∞∑n=0

( 12π

∫ 2π

0f(eiθ)e−inθ dθ

)rneinη

= (Pr ∗ f)(reiη) ,

that is, f = P(f). The fact that C(f) = P(f) follows from the previous Lemma 1.27.When 1 ≤ p < ∞, he second part of the statement follows immediately from Lemma 1.27,

Prop. 1.11 and the holomorphicity of the Cauchy integrals.If p = ∞, we obtain C(g) = P(g) as before, and clearly f = P(g) = C(g) ∈ H∞ and the

equality f = g follows from the case p <∞. 2

The next result is now obvious, and we state it for sake of completeness.

Corollary 1.29. For 1 ≤ p ≤ ∞ we denote by Hp(∂D) the subspace of Lp(∂D) of the functionsthat are boundary values of functions in Hp.

Then,Hp(∂D) =

g ∈ Lp(∂D) : g(n) = 0 for n < 0

.

We denote by P the set of the (complex) polynomials in z, and by A(D) the space ofholomorphic functions on D that are continuous up to the boundary, that is,

A(D) = H(D) ∩ C(D) ,

endowed with the sup-norm.

Corollary 1.30. For 1 ≤ p < ∞ the polynomials are dense in Hp, while the closure of P inthe sup-norm is A(D) and A(D) ( H∞. Furthermore, if f ∈ A(D), then fr → f uniformly on∂D.

Page 20: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

SPACES OF HOLOMORPHIC FUNCTIONS 19

Proof. We denote by σN (g) the Fejer trigonometric polynomial of an integrable function g onT = ∂D, that is,

σN (g)(eiη) =N∑

k=−N

(1− |k|

N + 1

)g(k)eikη .

It is well known (see [Ka] or your Real Analysis lecture notes) that σN (g) → g in Lp(∂D), asN → +∞, when 1 ≤ p <∞.

Now, let 1 ≤ p <∞, f ∈ Hp, f its boundary function. Then,

σN (f)(eiη) =N∑k=0

(1− |k|

N + 1

) ˆf(k)eikη ,

by Prop. 1.28. Clearly σN (f) is the restriction to the boundary of a polynomial pN and it holdsthat pN = C(σN (f)). Then, we obtain a sequence of polynomials in D such that

‖f − pN‖Hp = ‖f − pN‖Lp(∂D) = ‖f − σN (f)‖Lp(∂D) → 0

as N → +∞. Thus, P is dense in Hp, 1 ≤ p <∞.

Next, it is clear that A(D) ⊂ H∞ and that the closure of P in the sup-norm is contained inA(D) = H(D) ∩ C(D).

Now, if f ∈ A(D), then f ∈ C(∂D), f = P(f) and therefore, fr = Pr ∗ f → f uniformlyon ∂D. If fr(eiη) =

∑+∞n=0 anr

neinη and we set pN (reiη) =∑N

n=0 anrneinη, then pN (r·) → fr

uniformly on ∂D.Therefore, using the maximum modulus principle

supz∈D|f(z)− pN (z)| = sup

eiη∈∂D|f(eiη)− pN (eiη)|

≤ supeiη∈∂D

(|f(eiη)− fr(eiη)|+ |fr(eiη)− pN (reiη)|+ |pN (reiη)− pN (eiη)|

)< ε

choosing r and N large enough.It only remains to show that A(D) is strictly contained in H∞. It suffices to consider an

infinite Blaschke product B, as defined in Def. 1.22. By Cor. 1.24 we know that |B| = 1 a.e.on ∂D. Let ζj be the (infinite) sequence of the zeros of B in D, then they must accumulateat boundary points. Clearly B cannot be continuous at those points. 2

1.7. The Cauchy–Szego projection.

Definition 1.31. We define the Cauchy–Szego projection on L2(∂D) as the Hilbert spaceorthogonal projection as

S(g)(eiη) = S( +∞∑n=−∞

g(n)ein(·))

(eiη) =+∞∑n=0

g(n)einη .

It is clear that S is a projection (that is, S2 = S) and that it is self-adjoint (that is, S∗ = S);hence an orthogonal projection.

The next result follows from the previous lemma.

Page 21: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

20 M. M. PELOSO

Proposition 1.32. For g ∈ L2(∂D) we have that

S(g) = C(g) .

In particular we have obtained a characterization of S as (boundary limit of an) integraloperator. Namely,

S(g)(eiη) = limr→1−

12π

∫ 2π

0g(ei(η−θ))

11− reiθ

dθ . (1.21)

With an abuse of notation, we will simply write

S(g)(eiη) =1

∫ 2π

0g(ei(η−θ))

11− eiθ

dθ . (1.22)

Notice however that the integral is not absolutely convergent since g ∈ L2 and the function1/(1−eiθ) has a non-integrable singularity in θ = 0. The integral in (1.22) has to be interpretedas in (1.21).

We wish to extend the action of S to all the spaces Lp(∂D) and prove its Lp-boundedness,for the appropriate range of p’s.

The following result, of which we will not provide a complete proof, answers the question.

Theorem 1.33. The operator S, initially defined on the dense subset L2 ∩ Lp, extends to anoperator

S : Lp(∂D)→ Lp(∂D)that is of weak-type (1, 1) and is bounded for 1 < p <∞.

Sketch of the proof. Recall the definition S(g) = limr→1− Cr ∗ g and the relation between theCauchy and Poisson kernels (1.13). From the properties of the Poisson kernel we see that

S1(g) = limr→1−

ReCr ∗ g

extends to a bounded linear operator for 1 ≤ p <∞.Thus, in order to prove the theorem, it suffices to prove the statement for (twice) the imaginary

part of Cr, where

2ImCr(eiη) =1i

[ 11− reiη

− 11− re−iη

]=

2r sin η|1− reiη|2

=2r sin η

1− 2r cos η + r2(1.23)

= −i+∞∑

n=−∞sgn(n)r|n|einη := Qr(eiη) , (1.24)

(where the last equality follows from the power series expansions of the right hand side of thefirst line in display).

The kernel Qr is called the conjugate Poisson kernel, the reason being that for every 0 < r < 1,Qr is the harmonic function conjugate to the real harmonic function Pr on the unit disc.

The family of functions on T Qr does not form a summability kernel, and the propertiesof the mapping g 7→ limr→1− Qr ∗ g are substantially different from the ones we obtain when Qris replaced by the Poisson kernel.

Page 22: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

SPACES OF HOLOMORPHIC FUNCTIONS 21

We give a name to the operator g 7→ limr→1− Qr ∗ g, we write

Q(g) = limr→1−

Qr ∗ g , (1.25)

and call Q the Hilbert transform on T. (In fact, this operator is the closely related to theclassical Hilbert transform on the real line. We will return to the latter operator in Section 4.3).

The operator Q belongs to a fundamental class of operators, called singular integrals. Us-ing the fact that the (convolution) kernels Qr are odd and therefore they have built in somecancellation that ultimely is key in proving the theorem.

Notice that the pointwise limit of the kernels

limr→1−

Qr(eiη) =1 sin η

1− cos η= cot(η/2) := Q(eiη)

and that Q 6∈ L1(∂D). It turns out (and it is not hard to see) that,

Q(g)(eiη) = p.v.(Q ∗ g)(eiη) =: limε→0+

∫ε<|θ|≤π

g(ei(η−θ))Q(θ) dθ .

We will not prove that Q is weak-type (1, 1) and bounded on Lp, 1 < p < ∞, but refer to[Ka], Ch. 3, e.g., or also [Gr]. 2

It can be shown that the Cauchy–Szego projection is not bounded from L1(∂D) to itself.4

Thm. 1.33 also gives an answer to the problem of the harmonic conjugate. Suppose u is a realharmonic function on the unit disc. Then we know that u admits a harmonic conjugate v, thatis, a real harmonic function such that u + iv is holomorphic in D. The function v is uniquelydetermined modulo constants, so we require v to vanish at the origin z = 0, i.e. we set v(0) = 0.

In fact, it is not difficult to see that the boundedness of the Cauchy– Szego projection onLp(∂D) is equivalent to the boundedness of the harmonic conjugate operator T .

Corollary 1.34. Let u be a real harmonic function satisfying the Hp-growth condition (1.14),1 < p <∞, and let v be its harmonic conjugate. Then, also v satisfies the Hp-growth condition(1.14) and the mapping T : u 7→ v is bounded on Lp(∂D), 1 < p <∞.

Proof. By Prop. 1.9 u = Pr ∗ u, with u ∈ Lp(∂D). Now, vr = Qr ∗ u and v = Q(u). By theprevious theorem ‖v‖Lp = ‖Q(u)‖Lp ≤ C‖u‖Lp , and we are done. 2

There exists an anologous theory for the Hardy on the upper half-plane. The two theory havemany common aspects, but they also present significant differencies, since the upper half-plane isunbounded, while D is not. Moreover, the symmetry properties of the two domains are different(although analogous) and certain formulas appear simpler in one setting or another.

We will present some further aspects of this theory in the setting of the upper half-plane inSections 4.1 - 4.4.

4Exercise.

Page 23: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

22 M. M. PELOSO

2. Bergman spaces on the unit disc

2.1. Function spaces with reproducing kernel. In this section we briefly describe the theoryof reproducing kernels.

LetH be a Hilbert space of functions defined on a set Ω and suppose that the point evaluationsare bounded (linear) functionals on H, that is, for each z ∈ Ω, there exists a constant Cz > 0such that for all f ∈ H we have

|f(z)| ≤ Cz‖f‖H . (2.1)This implies that the linear functional

Lz(f) = f(z) , f ∈ H ,is bounded. By the Riesz–Fisher theorem, there exists kz ∈ H such that, for all f ∈ H we have

〈f, kz〉H = f(z) .

We define a kernel function K : Ω× Ω→ C by setting

K(w, z) = kz(w) . (2.2)

Such kernel K is called the reproducing kernel for H.

Proposition 2.1. The kernel K satisfies the following properties:(i) for all f ∈ H and z ∈ Ω we have f(z) = 〈f,K(·, z)〉H;(ii) for all z, w ∈ Ω, K(w, z) = K(z, w).

Proof. Property (i) is obvious by construction. For (ii), using the reproducing property of K wehave

K(w, z) = kz(w) = 〈kz,K(·, w)〉H = 〈K(·, w), kz〉H= K(z, w)

and we are done. 2

The following result establishes conditions that characterize the reproducing kernel of H.

Lemma 2.2. Let H(z, w) be a function on Ω× Ω such that(i) H(·, w) ∈ H for all w ∈ Ω fixed;(ii) 〈f,H(·, z)〉H = f(z) for all f ∈ H and z ∈ Ω.

Then, H(z, w) coincides with the reproducing kernel K(z, w) of H.

Proof. This is simple: using the reproducing property of H(·, z) and K(·, z), for every w ∈ Ωfixed we have

H(z, w) = 〈H(·, w),K(·, z)〉H = 〈K(·, z), H(·, w)〉H = K(w, z)

= K(z, w) ,

as we wished to show. 2

Notice that (i) and (ii) together imply the hermitian-symmetric property H(w, z) = H(z, w)for all z, w ∈ Ω, as in the proof of (ii) in Prop. 2.1.

In this notes we will be concerned only with spaces H for which (2.1) holds and that consistof holomorphic functions.

Page 24: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

SPACES OF HOLOMORPHIC FUNCTIONS 23

It will also always be the case that, for each compact subset E of Ω there exists C = CE > 0such that for all f ∈ H we have

supz∈E|f(z)| ≤ C‖f‖H . (2.3)

If (2.3) holds, then clearly the convergence in H implies the uniform convergence on compactsubsets of Ω.

Proposition 2.3. Let H be a Hilbert space of holomorphic functions on Ω for which condition(2.3) holds. Let ϕj be an orthonormal basis for H. Then the series

+∞∑j=1

ϕj(z)ϕj(w)

converges uniformly on compact subsets of Ω× Ω to the reproducing kernel K(z, w) of H.

Proof. It suffices to consider the case of subsets of the form E×E, where E is a compact subsetof Ω. By the Riesz-Fisher theorem for `2, for z ∈ E we have( +∞∑

j=1

|ϕj(z)|2)1/2

=∥∥∥ϕj(z)∥∥∥

`2= sup‖aj‖`2=1

∣∣∣ +∞∑j=1

ajϕj(z)∣∣∣

= sup‖f‖H=1

|f(z)|

≤ sup‖f‖H=1

supz∈E|f(z)|

≤ CE .Therefore,

supz∈E

( +∞∑j=1

|ϕj(z)|2)1/2

≤ CE ,

so that the Cauchy–Schwarz inequality gives that+∞∑j=1

∣∣ϕj(z)ϕj(w)∣∣ ≤ C2

E ,

and the convergence is uniform on E × E.Next, for z ∈ Ω fixed, the previous argument also shows that ϕj(z) ∈ `2, so that

∑+∞j=1 ϕj(z)ϕj

converges in H to a function hz. By the theory of Fourier series in H we have that, for all f ∈ H

〈f, hz〉H =+∞∑j=1

〈f, ϕj〉Hϕj(z) = f(z) .

Therefore, the function H(w, z) = hz(w) =∑+∞

j=1 ϕj(z)ϕj(w) satisfies the conditions ofLemma 2.2 and hence coincides with the reproducing kernel K(w, z).

Notice that the argument above also shows that the series∑+∞

j=1 ϕj(z)ϕj converges in theH-norm, for all z ∈ Ω fixed. 2

An interesting consequence of this proposition is the following result that shows that thereproducing kernel K satisfies an extremal property.

Page 25: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

24 M. M. PELOSO

Corollary 2.4. Let H be a space of holomorphic functions on Ω for which condition (2.3) holdsand let K(z, w) be its reproducing kernel. Then

K(z, z) = supf∈H,‖f‖H=1

|f(z)|2 .

Proof. Let ϕk be an orthonormal basis for H. We have

K(z, z) =∑k

|ϕk(z)|2

=(

supak∈`2, ‖ak‖`2=1

∣∣∣∑k

akϕk(z)∣∣∣)2

= supf∈H,‖f‖H=1

|f(z)|2 ,

as claimed. 2

Notice that the reproducing kernel K satisfies the condition K(z, z) ≥ 0 and in fact K(z, z) >0 unless f(z) = 0 for all f ∈ H.

We conclude this introduction on Hilbert spaces with reproducing kernels with a concreteexample, that we have in fact already encountered.

Proposition 2.5. The Hardy space H2(D) is a Hilbert space with reproducing kernel, with innerproduct 〈f, g〉H2 = 〈f , g〉L2(∂D). Its reproducing kernel is the Cauchy–Szego kernel

K(z, w) =1

1− wz.

Proof. We showed that ‖f‖H2 = ‖f‖L2(∂D). Therefore, H2 is a Hilbert space with inner product〈f, g〉H2 = 〈f , g〉L2(∂D), for f, g ∈ H2. (We also showed that we can describe the inner productin H2 by setting

〈f, g〉H2 = 〈an, bn〉`2 ,

where f(z) =∑+∞

n=0 anzn and g =

∑+∞n=0 bnz

n. However, we will not use this characterization atthis stage.)

We now show that the point evalutions are bounded functionals on H2. Let z ∈ D and let|z| < r < 1. By the Cauchy formula

f(z) =1

2πi

∫|ζ|=r

f(ζ)ζ − z

dζ =1

∫ 2π

0

f(reiθ)reiθ − z

reiθ dθ , (2.4)

and by the Cauchy–Schwartz inequality we have

|f(z)| ≤( 1

∫ 2π

0|f(reiθ)|2 dθ

)1/2( 12π

∫ 2π

0

r2

|reiθ − z|2dθ)1/2

≤ r

r − |z|‖f‖H2 .

This shows that H2 is a Hilbert space with reproducing kernel.

Page 26: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

SPACES OF HOLOMORPHIC FUNCTIONS 25

Consider now the kernel K(z, w) = 11−wz . It is immediate to see that K(·, w) ∈ H2 for every

w ∈ D fixed. Moreover,

〈f,K(·, z)〉H2 = 〈f , K(·, z)〉L2(∂D) =1

∫ 2π

0

f(eiθ)1− e−iθz

= limr→r1−

12π

∫ 2π

0

f(reiθ)reiθ − z

reiθ dθ

= f(z) ,

by (2.4). The conclusion now follows from Lemma 2.2. 2

2.2. The Bergman spaces. Let Ω be a domain in C. We denote by dA the Lebesgue measureon C, which of course is identified with real euclidean space R2, and we write |E| to denotethe Lebesgue measure of a measurable set E. For 0 < p <∞ we denote by Lp(Ω) the space ofthe Lebesgue measurable functions that are p-integrable with respect to dA, and by L∞(Ω) thespace of measurable functions that are essentially bounded.

Definition 2.6. For 0 < p ≤ ∞ we define the Bergman space Ap as Ap(Ω) = Lp(Ω) ∩H(Ω).

Proposition 2.7. Let Ω ⊂ C be a domain and let 0 < p ≤ ∞. Then the space Ap(Ω) is a closedsubspace of Lp(Ω). When p = 2, A2 is a Hilbert space with reproducing kernel.

Proof. We show that the functions in the Bergman space Ap satisfy (2.3). Let E be a compactsubset of Ω and let δ be half the distance of E from ∂Ω, i.e. δ = 1

2dist(E, ∂Ω).For every z ∈ E, D(z, δ) ⊂ Ω. From the mean value property

f(z) =1

∫ 2π

0f(z + reiθ) dθ ,

(valid for every r > 0 such that D(z, r) ⊂ Ω) multiplying by r both sides and integrating inr ∈ [0, δ] we obtain the identity5

f(z) =1

|D(z, δ)|

∫∫D(z,δ)

f(w) dA(w) .

Therefore, for every z ∈ E, using Holder’s inequality, for 1 ≤ p ≤ ∞,

|f(z)|p ≤ 1|D(z, δ)|

∫∫D(z,δ)

|f(w)|p dA(w) (2.5)

≤ 1|D(z, δ)|

‖f‖pLp =1πδ2‖f‖pLp .

We remark that the above inequality (2.5) holds also for 0 < p < 1, but this is a consequenceof the fact that |f |p is subharmonic when f is holomorphic, and to avoid introducing anothernotion, we will not prove this fact here; see [Kr], e.g.

Since E can be covered by a finite number of such discs

supz∈E|f(z)| ≤ CE‖f‖Lp . (2.6)

5 This identity is called the mean value property and it is valid for all holomorphic functions f and discs D(z, δ)contained in its region of holomorphicity.

Page 27: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

26 M. M. PELOSO

This inequality show that if fn is a sequence of holomorphic functions converging in theLp-norm to a function f , then fn converges to f also uniformly on compact subsets. Thisproves that Ap is closed 1 < p <∞. The case p =∞ is obvious.

When p = 2 (2.6) is exactly (2.3) and this concludes the proof.Notice that the constant CE in inequality (2.6), depends on the distance of the set E from

the boundary of Ω, and, more precisely CE ∼ δ−2/p, as δ = 12dist(E, ∂Ω)→ 0. 2

Definition 2.8. Given a domain Ω ⊂ C, let K = KΩ be its Bergman kernel, and for z ∈ Ωwrite kz = K(·, z). We define the Bergman projection on Ω the mapping B defined on L2(Ω) as

(Bf)(z) = 〈f, kz〉 =∫

Ωf(w)kz(w) dA(w)

=∫

Ωf(w)K(z, w) dA(w) .

Proposition 2.9. The Bergman projection is the Hilbert space orthogonal projection of L2(Ω)onto its closed subspace A2(Ω).

Proof. First of all, notice that Bf is well defined for all f ∈ L2, since for all z ∈ Ω, kz ∈ A2 ⊂ L2,and the L2-pairing 〈f, kz〉 makes sense (and converges).

Next, we use the characterization of the Bergman kernel given by Prop. 2.3, that is, kz =∑j ϕj(z)ϕj , where ϕj is an orthonormal basis for A2, where the convergence is in the L2-norm,

for all z ∈ Ω fixed.Therefore,

(Bf)(z) = 〈f, kz〉 =⟨f,∑j

ϕj(z)ϕj⟩

=∑j

ϕj(z)〈f, ϕj〉 =∑j

ajϕj(z) ,

where aj = 〈f, ϕj〉 is the j-th Fourier coefficient of f , w.r.t. the orthonormal system ϕj,(which is not complete in L2). Clearly, aj ∈ `2 and

∑j |aj |2 ≤ ‖f‖2L2 , by Bessel’s inequality.

Hence, the series∑

j〈f, ϕj〉ϕj converges in L2 and therefore

‖Bf‖L2 ≤ ‖f‖L2 .

It is also clear that Bf ∈ A2 since A2 is closed, so that B : L2 → A2 and it is bounded. Itremains to show that B is the orthogonal projection, that is, that B2 = B and B∗ = B.

The former identity follows by the reproducing property of the Bergman kernel when actingon functions in A2, while the latter one follows from the hermitian symmetry property: forf, g ∈ L2

〈Bf, g〉 =⟨∑

j

〈f, ϕj〉ϕj , g⟩

=∑j

〈f, ϕj〉〈ϕj , g〉

=⟨f,∑j

〈g, ϕj〉ϕj⟩

= 〈f,Bg〉 .

This shows that B∗ = B and we are done. 2

Page 28: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

SPACES OF HOLOMORPHIC FUNCTIONS 27

Remark 2.10. We remark that, up to this point, we have not discussed whether, on a givendomain Ω ⊂ C, the Bergman space A2(Ω) is trivial, i.e. reduces to 0, is infinite dimensional,or even if it is not trivial, but finite dimensional. We mention that all these cases can occur. Itis easy to see that A2(C) = 0, and the finite dimensional examples are also fairly elementary;we refer the reader to [Wi].

However, it is important to point out that, an application of the Weierstrass approxima-tion theorem6 shows that, if Ω is a bounded domain, the space of holomorphic polynomials is(contained and) dense in A2(Ω).

We conclude this remark by observing that, as a consequence of the extremal property in Cor.2.4 we have that, it Ω ⊆ Ω′ are domains in C and K,K ′ denote their Bergman kernels resp.,then

K ′(z, z) ≤ K(z, z)

for all z ∈ Ω.

Before proceeding any further, we present an example of Bergman space and kernel.

Proposition 2.11. Let D be the unit disc. Then, the set of functions ϕk, where ϕk =√(k + 1)/πzk, k = 0, 1, 2, . . . , is a complete orthonormal system in A2(D) and the Bergman

kernel has the expression

K(z, w) =1π

1(1− zw)2

.

Proof. It is clear that ϕk is an orthonormal system:

〈ϕj , ϕk〉 =

√(j + 1)(k + 1)

π

∫ 1

0

∫ 2π

0ei(j−k)θ dθ rj+k+1 dr

= δjk2(k + 1)∫ 1

0r2k+1 dr

= δjk ,

where δjk denotes the Kronecker’s delta.By the uniqueness of the Taylor expansion, it is also clear that the holomorphic polinomials

(i.e. the polynomials in z) are dense in A2(D). For, let f ∈ A2(D) be orthogonal to all the zk’sand let f(z) =

∑n anz

n be its power series expansion in D. By integrating over the closed disc|z| ≤ r for some fixed r < 1 and interchanging the summation and integration orders, we obtainthat an = 0 for all n; hence f = 0.

Thus, ϕk is a complete orthonormal system for A2(D) and, by Prop. 2.3 the Bergmankernel is given by the sum

∑k ϕk(z)ϕk(w), that we easily compute:∑k

k + 1π

zkwk =1π

∑k

(k + 1)(zw)k

=1π

1(1− zw)2

,

6True? Recall? Reference?

Page 29: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

28 M. M. PELOSO

as we wished to show.7 2

Notice the Bergman kernel for the unit disc K(z, w) satisfies the condition K(z, w) 6= 0 andK(z, z) > 0 for all z, w ∈ D.

2.3. Biholomorphic invariance. One of the main features of the Bergman kernel is its invari-ance under conformal mappings. This invariance is the content of the next proposition.

Proposition 2.12. Let Ω1,Ω2 ⊂ C be domains, Φ : Ω1 → Ω2 be a biholomorphic mapping, andlet KΩ1 ,B1 and KΩ2 ,B2 be the Bergman kernels and projections of Ω1 and Ω2, resp. Then wehave

KΩ1(z, w) = Φ′(z)KΩ2

(Φ(z),Φ(w)

)Φ′(w) , (2.7)

and, for f ∈ A2(Ω2) we

B1

(Φ′(f Φ)

)= Φ′

((B2f) Φ

). (2.8)

Proof. We begin by observing that if Φ : Ω1 → Ω2 is a biholomorphic mapping, then thedeterminant jacobian of Φ thought as a mapping between domains in R2 equals |Φ′(z)|2. For,if we write Φ = u+ iv, u, v real functions, then Φ is identified with the mapping

Ω 3 z = x+ iy ≡ (x, y) 7→(u(x, y), v(x, y)

).

Then

Jac Φ =(∂xu ∂yu∂xv ∂yv

),

so that, using the Cauchy-Riemann equations,

det(Jac Φ) = (∂xu)2 + (∂xv)2 = |∂xΦ|2 = |Φ′|2 .

This identity implies that the mapping f 7→ Φ′(f Φ) is an isometric isomorphism of A2(Ω2)onto A2(Ω1).

Notice also the analogous statements hold true with Φ−1 in place of Φ and with the roles ofΩ1 and Ω2 switched.

For z, w ∈ Ω1 we set

H(z, w) = Φ′(z)KΩ2

(Φ(z),Φ(w)

)Φ′(w) ,

and easily see that H satisfies the conditions in Lemma 2.2, thus implying that H coincides withthe Bergman kernel on Ω1. 8

7Exercise. Using the residue theorem, show that for every f ∈ A2(D) and z ∈ D we have the identity

f(z) =1

π

ZZD

f(w)

(1− zw)2dA(w) .

Once you have completed your argument, explain where you use the assumption f ∈ A2(D). Can this assumptionbe relaxed?

8Exercise.

Page 30: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

SPACES OF HOLOMORPHIC FUNCTIONS 29

Finally, (2.8) follows easily from (2.7). Let z ∈ Ω1

B1

(Φ′(f Φ)

)(z) =

∫Ω1

Φ′(ζ)(f Φ)(ζ)KΩ1(z, ζ) dA(ζ)

= Φ′(z)∫

Ω1

f(Φ(ζ)

)KΩ2

(Φ(z),Φ(ζ)

)|Φ′(ζ)|2 dA(ζ)

= Φ′(z)∫

Ω2

f(ω)KΩ2

(Φ(z), ω

)dA(ω)

= Φ′(z)(

(B2f) Φ)

(z) ,

as we wanted to prove. 2

2.4. Lp-boundedness of a family of integral operators. In the next section we will provethe boundedness of the Bergman projection B when acting on Lp, 1 < p <∞ as a consequenceof a more general result concerning the boundedness of a family of operators on weighted Lp-spaces. This greater generality does not introduce extra arguments in the proof and at the sametime provides a variety of possible applications.

We begin by recalling a general result concerning the boundedness of integral operators withpositive kernels.

Proposition 2.13. (Schur’s test) Let (X , dµX ), (Y, dµY) be measure spaces. Let T be theintegral operator given by

Tf(x) =∫YK(x, y)f(y) dµY(y) ,

where K is a measurable positive kernel on X × Y. Let 1 < p, p′ < ∞ be conjugate exponents.Suppose there exist positive functions ψ : Y → (0,+∞), ϕ : X → (0,+∞), such that

(i)∫YK(x, y)ψp

′(y) dµY(y) ≤ Cϕ(x)p

′;

(ii)∫XK(x, y)ϕp(x) dµX (x) ≤ Cψ(y)p.

Then T : Lp(Y)→ Lp(X ) is bounded.

The proof is a fairly simple application of Holder’s inequality. We leave the details as anexercise.

For ν > −1 we consider the weights (1− |w|2)ν and the kernels

Kν(z, w) =1π

1|1− zw|2+ν

.

We are interested in the boundedness of the integral operators

T : Lp((1− |w|2)νdA

)→ Lp

((1− |w|2)νdA

),

where

Tf(z) =∫Df(w)Kν(z, w) (1− |w|2)ν dA(w) .

Page 31: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

30 M. M. PELOSO

We begin with the following integral estimate. Given two positive quantities A and B depend-ing on a variable x, we write A ≈ B as x→ x0 if there exists c > 0 such that 1

c ≤ A(x)/B(x) ≤ cas x→ x0.

Lemma 2.14. Let γ > −1, t ∈ R and for z ∈ D set

It(z) =∫D

(1− |w|2)γ

|1− zw|2+γ+tdA(w) .

Then,

It(z) ≈

1 if t < 0log[1/(1− |z|)

]if t = 0

(1− |z|2)−t if t > 0 ,

for |z| → 1−.

Notice that the function It is bounded on every compact subsets of D. Therefore, the be-haviour of It(z) is non-trivial only for z approaching the boundary.

The proof of the lemma is based on another integral estimate.

Lemma 2.15. Let τ ∈ R and for % ∈ (0, 1) set

Jτ (%) =∫ 2π

0

1|1− %eiθ|1+τ

dθ .

Then, as %→ 1−,

Jτ (%) ≈

1 if τ < 0log[1/(1− %)

]if τ = 0

(1− %)−τ if τ > 0 .

Proof. Notice that, for all θ ∈ [0, 2π),

(1− %) ≤ |1− %eiθ| ≤ 1 + % ,

so that the region of integration [0, 2π) is contained in the union of the disjoint sets Ej , j =0, . . . , N , where

Cj =θ : 2j(1− %) ≤ |1− %eiθ| < 2j+1(1− %)

and N = blog2

[(1 + %)/(1− %)

]c.9 We also observe that, if we set

Ej =θ : |1− %eiθ| < 2j(1− %)

,

then Cj = Ej+1 \ Ej , j = 0, . . . , N . Moreover,

|Ej | ≤ c2j(1− %)

for some constant c > 0 (and independent of j and %).10

9We denote by btc the integral part of the real number t.10For, |1− %eiθ|2 < 22(j+1)(1− %)2 if and only if 1− A ≤ cos θ, where A = (22(j+1) − 1)(1− %)2/2%. But this

is equivalent to |θ| ≤ cos−1(1−A) ≈ (1− %), as %→ 1−.

Page 32: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

SPACES OF HOLOMORPHIC FUNCTIONS 31

Then, if τ > 0

Jτ (%) ≤N∑j=0

∫Cj

1|1− %eiθ|1+τ

≤N∑j=0

1(2j(1− %)

)1+τ |Ej |

≤ Cτ (1− %)−τ .

When τ = 0 we need to refine the above estimate. We obtain

Jτ (%) ≤N∑j=0

1(2j(1− %)

) |Cj |≤

N∑j=0

1(2j(1− %)

) |Ej | ≤ cN≤ c log(1/(1− %)) .

When τ < 0 is clear that Jτ is bounded as %→ 1− since |1− eiθ|−1−τ is integrable on [0, 2π).Finally, the reverse inequalities are easy to prove and we leave the details to the reader.11 2

Proof of Lemma 2.14. Integrating in polar coordinates w = reiθ and using the invariance in theintegral over the circle we see that

It(z) =∫ 1

0

∫ 2π

0

1∣∣1− |z|reiθ∣∣2+γ+t dθ(1− r)γr dr .

If t > 0, then τ := 1 + γ + t is also > 0 and Lemma 2.15 gives that

It(z) ≈∫ 1

0

(1− r)γ(1− |z|r

)1+γ+t r dr =∫ 1

0

(1− r)γ((1− r) + (1− |z|)r

)1+γ+t r dr .

We split the latter integral into the regions G1 = r : (1 − r) ≤ (1 − |z|)r and G2 = r :(1− r) > (1− |z|)r, that is,

[0, 1/(2− |z|)

]and

(1/(2− |z|), 1

). Then,

It(z) ≤∫ 1/(2−|z|)

0

1(1− r)1+t

r dr +∫ 1

1/(2−|z|)

(1− r)γ((1− |z|)r

)1+γ+t r dr

≤∫ 1/(2−|z|)

0

1(1− r)1+t

dr +C

(1− |z|)1+γ+t

∫ 1

1/(2−|z|)(1− r)γ dr

≤ Cγ,t(1− |z|)t

.

This proves the estimate from above in the case t > 0. It is not hard to see that the estimatefrom below follows from the same argument.

The case t < 0 is easy, so is the case t = 0, whose details are left to the reader. 2

11Exercise.

Page 33: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

32 M. M. PELOSO

We finally come to the main result of this section. For ν > −1 define the measure on the unitdisc dAν(w) = (1− |w|2)νdA(w).

Theorem 2.16. For α, ν > −1 consider the integral operator Tν defined as

Tνf(z) =∫∫

Df(w)

(1− |w|2)ν

|1− wz|2+νdA(w) .

Then, if α+ 1 < p(ν + 1), Tν : Lp(D, dAα)→ Lp(D, dAα) is bounded.

Proof. We are going to use Schur’s test Prop. 2.13 and the integral estimates in Lemma 2.14.Notice that in this case the pairs (Y, dµY) and (X , dµX ) are both equal to (D, dAα) and that

we write the operator Tν as

Tν(f)(z) =∫∫

Df(w)

(1− |w|2)ν−α

|1− wz|2+νdAα(w) ,

that is, as having integral kernel H(z, w) = (1−|w|2)ν−α

|1−wz|2+ν w.r.t. the measure dAα.Let a > 0 to be chosen later and define ψ(w) = (1 − |w|2)−a. When ν − ap′ > −1, we may

apply Lemma 2.14 with γ = ν − ap′ and t = ap′ and obtain

Tν(ψp′)(z) =

∫∫Dψp′(w)H(z, w) dAα(w)

=∫∫

D

(1− |w|2)ν−ap′

|1− wz|2+νdA(w)

≤ C(1− |z|2)−ap′

= Cψp′(z) .

Notice that this argument is possible if the parameter a satisfies

0 < a <ν + 1p′

. (2.9)

On the other hand, we apply Lemma 2.14 again with γ = α−ap > −1 and t = ν−(α−ap) > 0,and obtain

Tν(ψp)(w) =∫∫

Dψp(z)H(z, w) dAα(z)

= (1− |w|2)ν−α∫∫

D

(1− |z|2)α−ap

|1− wz|2+νdA(z)

≤ C(1− |w|2)−ap = Cψp(w) .

Again, this argument is possible if the parameter a satisfiesα− νp

< a <α− νp

. (2.10)

Therefore, we can apply Prop. 2.13 if we can find a satisfying both (2.9) and (2.10), that is,if the intervals (A,B) = (0, ν+1

p′ ) and (C,D) = (α−νp , α−νp ) have non-empty intersection. SinceC > 0, this happens if and only if C < B, that is,

α− νp

<ν + 1p′

.

Recalling that 1/p′ = 1− 1/p, the statement follows. 2

Page 34: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

SPACES OF HOLOMORPHIC FUNCTIONS 33

2.5. The Bergman kernel and projection on the unit disc. In this section we consider theBergman projection B on the unit disc D. Since C(D) ⊂ L∞(D) ⊂ Lp(D) for all 0 < p < ∞,we have that B is well defined on a dense subset of Lp if 0 < p ≤ 2, and on all Lp if 2 < p ≤ ∞.

Main goal of this section is proving the follwing result.

Theorem 2.17. Let 1 ≤ p ≤ ∞. Then, the Bergman projection B is bounded B : Lp(D) →Ap(D) if and only if 1 < p <∞.

Proof. The sufficiency follows easily from Thm. 2.16. Observe that the bounedness of theoperator T with kernel |K(z, w)| implies the boundedness of P . Thus, it suffices to apply Thm.2.16 with ν = α = 0.

To prove the necessity, observe that if B : L1(D, dAγ) → A1(D, dAγ) were bounded, B :L∞(D) → A∞(D) would also be bounded. Thus, it suffices to show that B is not bounded onL∞(D). Consider the family of functions fz(w) = (1−|w|2)2+ν(1−zw)−(2+ν). Then, ‖fz‖L∞ = 1for all z ∈ D. But,

‖Tνfz‖L∞ ≥ |Tνfz(z)| =∣∣ ∫∫

D

(1− |w|2)2+2ν

|1− wz|2(2+ν)dA(w)

∣∣≥ C log(1− |z|)−1 ,

that clearly is not bounded. Therefore, B cannot be bounded on L∞ and on L1 and we aredone. 2

We now discuss some consequences of the boundedness of the Bergman projection. The firstone is the question of the description of the dual space of Ap(D). It is clear that A2(D) is aHilbert space and can be identified with its dual space, that is,(

A2(D))′ = A2(D) ,

where the identification is w.r.t. to its own inner product.By Holder’s inequality we also immediately have taht any g ∈ Ap′(D) defines a linear func-

tional on Ap(D), again via the L2(D)-inner product

Lg(f) =∫∫

Df(z)g(z) dA(z) .

However, it is not apriori clear that every g ∈ Ap′(D), g 6= 0, gives rise to a non-trivial

functional on Ap(D).

Theorem 2.18. Let 1 < p < ∞, then the dual space of Ap(D) can be identified with Ap′(D),

where p′ is the conjugate exponent.

Proof. Let L be a continuous linear functional on Ap(D), that is, L ∈(Ap(D)

)′. By the Hahn–Banach theorem, L extends to a linear functional on Lp(D), so that there exists g ∈ Lp

′(D)

such that L(f) =∫∫

D f(z)g(z) dA(z) for all f ∈ Ap(D). Using the fact that B is bounded on Lp

we see that〈f, g〉 = 〈B(f), g〉 = 〈f,B(g)〉 =: 〈f, h〉 ,

with h ∈ Ap′(D). Hence,(Ap(D)

)′ ⊆ Ap′(D).It only remains to see that a non-zero element of Ap

′(D) cannot be identically zero on Ap(D).

By symmetry we may assume that p < 2 < p′. Then Ap′ ⊂ A2 ⊂ Ap. If h ∈ Ap′, h 6= 0, and

Page 35: CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS · CLASSICAL SPACES OF HOLOMORPHIC FUNCTIONS MARCO M. PELOSO Contents 1. Hardy Spaces on the Unit Disc 1 1.1. Review from complex analysis

34 M. M. PELOSO

〈f, h〉 = 0 for all f ∈ Ap, then h ∈ A2 and in particular 〈f, h〉 = 0 for all f ∈ A2. Hence h = 0,a contradiction. 2

We show that the problem of the conjugate function has a simpler solution w.r.t. the case ofthe Hardy spaces.

Dipartimento di Matematica, Universita degli Studi di Milano, Via C. Saldini 50, 20133 Milano,Italy

E-mail address: [email protected]

URL: http://wwww.mat.unimi.it/~peloso/