Fatou’s theorem, Bergman spaces, and Hardy spaces on the circle
1 Introduction
In this note I am writing out proofs of some facts about Fourier series, Bergman spaces, and Hardy spaces. §§1–3 follow the presentation in Stein and Shakarchi’s Real Analysis and Fourier Analysis. The questions in Halmos’s Hilbert Space Problem Book that deal with Hardy spaces are: §§24–35, 67, 116–117, 124–125, 127, 193–199, and I present solutions to some of these in §§4–5, on Bergman spaces, and §6, on Hardy spaces.
2 Poisson kernel
Let . Let
If , let
Define
One checks that is an approximation to the identity, which implies that for , for almost all we have as .11 1 If is continuous, then converges to uniformly on as . This is proved for example in Lang’s Complex Analysis, fourth ed., chapter VIII, §5.
For , for any we have
hence by the dominated convergence theorem we have
and so
3 Harmonic functions
For , define on by
In polar coordinates, the Laplacian is
Then
Hence is harmonic on the open unit disc.
4 Fatou’s theorem
Let . If is holomorphic, let it have the power series
By the Cauchy integral formula, for and for any , , we have
Hence, for and for ,
On the other hand, for , we have
and because is holomorphic on for , by the residue theorem the right-hand side of the above equation is equal to . Hence, for and for ,
Let be holomorphic, and suppose there is some such that for all . For , define by . From our above work, we have
For , note that , so, by Parseval’s identity,
On the other hand,
It follows that
Define by
this defines an element of if and only if , and indeed
As , . Then by our work in §2, for almost all we have
which means here that for almost all ,
Thus, for almost all ,
In words, we have proved that if is a bounded holomorphic function on the unit disc, then it has radial limits at almost every angle. This is Fatou’s theorem.
5 Bergman spaces
This section somewhat follows Problem 24 of Halmos. Let be Lebesgue measure on . .
If is a nonempty bounded open subset of and , let denote the set of functions that are holomorphic and that satisfy
and let denote the set of functions that are holomorphic and that satisfy
It is apparent that is a vector space over . By Minkowski’s inequality, is a norm, and thus is a normed space. If then by Jensen’s inequality we have
and so
is called a Bergman space. It is not apparent that it is a complete metric space. We show this using the following lemmas. We use the following lemma to prove the lemma after it, and use that lemma to prove the theorem.
Lemma 1.
If , , and , then
Proof.
Put , with . For , define
We have as . Then,
which tends to as . Thus
For , using polar coordinates we have
Therefore
That is, for each we have
(1) |
Because ,
Thus, taking the limit as of (1), we obtain
∎
If and , denote
and for , let
This is the radius of the largest open disc centered at that is contained in (it is equal to the union of all open discs centered at that are contained in , and thus makes sense). As is open, , and as is bounded, .
Lemma 2.
If , , and , then
Proof.
Now we prove that is a complete metric space, showing that it is a Banach space.
Theorem 3.
If , then is a Banach space.
Proof.
Suppose that is a Cauchy sequence. We have to show that there is some such that in . The space of holomorphic functions on is a Fréchet space: there is an increasing sequence of compact sets whose union is , and the seminorms on are the supremum of a function on . (See Henri Cartan, Elementary Theory of Analytic Functions of One or Several Complex Variables, §V.1.3.) For each of these compact sets , let be the distance between and , which are both compact sets. If then . Thus if and , using Lemma 2 we get
From this and the fact that as , we get that
That is, is a Cauchy sequence in each of the seminorms , and as is a Fréchet space it follows that there is some such that in . In particular, for all we have as (because each is included in one of the compact sets , on which the converge uniformly to and hence pointwise to ).
On the other hand, is a Banach space, and hence there is some such that as . This implies that there is some subsequence such that for almost all , . Thus, for almost all we have . Therefore, in we have and so
∎
6 Inner products
In this section we follow Problem 25 of Halmos. In this section we restrict our attention to the Bergman space , where is the open unit disc, on which we define the inner product
As , it follows that is a Hilbert space with this inner product. If we have a Hilbert space we would like to find an explicit orthonormal basis.
Theorem 4.
If and , define by
Then are an orthonormal basis for .
Proof.
If is a subset of a Hilbert space and , we write if for all . If is an orthonormal set in , is an orthonormal basis if and only if implies that . This is proved in John B. Conway, A Course in Functional Analysis, second ed., p. 16, Theorem 4.13. For ,
while
Therefore is an orthonormal set. Hence, to show that it is an orthonormal basis for we have to show that if for all then .
For , let be the open disc centered at of radius , and let . Let , and for each this power series converges uniformly in . Then
One checks that , and hence
Therefore
As for each , this gives us that for all and hence . This shows that is an orthonormal basis for . ∎
Steven G. Krantz, Geometric Function Theory: Explorations in Complex Analysis, p. 9, §1.2, writes about the Bergman space , where is a connected open subset of , not necessarily bounded.
7 Hardy spaces
In a Hilbert space , if are subsets of , let denote the closure in of . Thus, to say that a set is an orthonormal basis for a Hilbert space is to say that is orthonormal and that .
Let , and let be normalized arc length, so that . Define by , for . It is a fact that are an orthonormal basis for the Hilbert space , with inner product
We define the Hardy space to be . As it is a closed subspace of the Hilbert space , it is itself a Hilbert space. For , we denote .
The following is Problem 26 of Halmos. Note .
Theorem 5.
If and , then is constant.
Proof.
If and , then
Thus is continuous .
If , then, as is an orthonormal basis for , we have , and so, as ,
Therefore if then
(2) |
For ,
the first equality is because , the second equality is by what we showed for any element of , and the third equality is because . It follows that , and thus that is constant. ∎
If , define by
and by
and, by (2), , so
and
(3) |
, and we have and ; that is, both and are real valued, like how the real and imaginary parts of a complex number are both real numbers.
The following is Problem 35 of Halmos. In words, it states that a real valued function has a corresponding real valued function (made unique by demanding that have constant term) such that the sum is an element of the Hardy space . This is called the Hilbert transform of . This is analogous to how if is harmonic on an open subset of , then satisfies the Cauchy-Riemann equations at every point in and hence is holomorphic on . Since is holomorphic on , for every there is some open neighborhood of on which has a primitive ( might not have a primitive defined on , e.g. on ), and there is a constant such that for all in this neighborhood. and are called harmonic conjugates.
Theorem 6.
If and , then there is a unique such that , , and .
Proof.
Define by
As , and using Parseval’s identity,
This is finite, hence .
For any and , by (2) we have . As , if then . Using this, we check that .
Put , hence . gives , and applying this and (3) we get . Thus satisfies the conditions , , and . We are not obliged to do so, but let’s write out the Fourier coefficients of . If then, using ,
Thus .
If , then, as ,
Thus
∎