Hardy Spaces Lecture Notes 2
Hardy Spaces Lecture Notes 2
Hardy Spaces Lecture Notes 2
1
CHAPTER 1
Preliminaries
and assume that there is z0 ∈ U such that (Fn (z0 )) has a nonzero limit.
Then (Fn ) converges uniformly on compacts to a nonzero holomorphic
function F on U and each zero of F is the zero of some fn , n ∈ N.
Analogously, one shows that |H| ≤ |G| and we obtain that |G| = |H|.
This implies that G = αH for some α ∈ C with |α| = 1. But, by
hypothesis, we have also G(z0 ) = H(z0 ) 6= 0 which implies α = 1
and G = H. To see the assertion about the zeros, note first that if
K ⊂ U is compact then K ∩ (∪n fn−1 ({0})) is finite, otherwise F would
be identically zero. Then if z is a zero of F that is not a zero of any fn
there is an open neighborhood of z that contains no zeros of fn , n ≥ 1
and this leads to a contradiction by Theorem 1.1.
Let us denote from now on by D the (open) unit disc (i.e. the disc
centered at the origin and of radius one). Our next result will be fre-
quently used in what follows and is a direct consequence of Proposition
1.4 combined with the following simple but important identity:
If z, w ∈ C with wz 6= 1 then
z − w 2 (1 − |z|2 )(1 − |w|2 )
(1.2) 1− = .
1 − wz |1 − wz|2
Indeed, this can be verified by a direct calculation (exercise!).
P
Corollary 1.1. Let (an ) be a sequence in D\{0}. If n (1−|an |) = ∞
then the infinite product
∞
Y −an z − an
B(z) =
n=1
|an | 1 − an z
1. SEQUENCES AND FAMILIES OF HOLOMORPHIC FUNCTIONS 5
P
converges uniformly to zero on compact subsets of D. If n (1−|an |) <
∞ then the product converges uniformly on compacts to a nonzero holo-
morphic function in D. Moreover, in this case, the zeros of B are pre-
cisely the points an , n ≥ 1 and for fixed m, the multiplicity of the zero
am equals the number of occurences of am in the sequence.
P
Proof. Assume that n (1 − |an |)
= ∞ and let w ∈ D. Note that
N N
! N !
Y |w − an |2 X |w − an |2 X |w − an |2
2
= exp log 2
≤ exp 2
−1 .
n=1
|1 − a n w| n=1
|1 − a n w| n=1
|1 − a n w|
By (1.2) we have
|w − an |2 (1 − |w|2 )(1 − |an |2 ) (1 − |w|2 )(1 − |an |2 )
1− = ≥
|1 − an w|2 |1 − an w|2 (1 + |w|)2
which implies that
N N
|w − an |2 1 − |w|2 X
X
2
−1 ≤− 2
(1 − |an |2 ) → −∞
n=1
|1 − an w| (1 + |w|) n=1
when N → ∞, and hence, the product converges to zero uniformly on
compacts.
P
Now assume that n (1 − |an |) < ∞. From (1.2) we deduce that the
factors fn (z) = −a n z−an
|an | 1−an z
satisfy
(ii) For every angle A with vertex at 1, symmetric w.r.t. the segment
[0, 1) and with opening less than π we have
lim f (z) = 0,
z→1
z∈A∩D
1+z
To prove these assertions, let g(z) = 1−z , z ∈ D. Then clearly, |g(z)| →
∞ if and only if z → 1. Let us show first that g(D) equals the right
half plane. Indeed, if g(z) = w and |z| < 1 then
1+z Re(1 + z)(1 − z) 1 − |z|2
Rew = Re = = > 0.
1−z |1 − z|2 |1 − z|2
Conversely, if Re w > 0, a similar computation shows that |z| < 1.
(i) Clearly, |f | = exp(−Re g) ≤ 1 if Re g > 0. The second assertion is
1+r
obvious since Re g(r) = 1−r → ∞ when r → 1.
(ii) Every z = reit ∈ A ∩ D can be written in the form z = 1 + ρeiθ ,
where ρ > 0 and |θ − π| ≤ α < π/2. Then
1 + ρeiθ
r = (1 + ρ2 + 2ρ cos θ)1/2 , eit =
(1 + ρ2 + 2ρ cos θ)1/2
and
eit − 1 1 + ρeiθ − (1 + ρ2 + 2ρ cos θ)1/2
= −
1 − r2 (1 + ρ2 + 2ρ cos θ)1/2 ρ(ρ + 2 cos θ)
Now the inequality satisfied by the angle θ implies that this quantity
stays bounded when ρ → 0, or equivalently, z → 1, z ∈ A ∩ D. This is
further equivalent to the fact that
|t|
(2.1) lim sup < ∞.
reit →1 1−r
reit ∈A∩D
With this fact at hand, the proof of (ii) is immediate. Indeed, from
above we have that
1 − r2 1 − r2 1 − r2
Reg(reit ) = = =
|1 − reit |2 1 + r2 − 2r cos t (1 − r)2 + 4r sin2 ( 2t )
8 1. PRELIMINARIES
satisfy
∞
!
X µ
|fN (z)| < exp µrn = exp( ).
n=0
1−r
Thus, the family {fN , N ≥ 1} is uniformly bounded on compacts
and hence normal, by Montel’s theorem. Moreover, exactly the same
argument as above shows that (fN (r)) converges for every 0 < r < 1
and by Vitali’s theorem the convergence assertion follows. Finally,
using the inequality 1 + x ≥ e−x we see that for 0 < r < 1
∞
!
X µ
f (r) ≥ exp − µrn = exp(− ),
n=0
1 − r
i.e. f is not identically zero.
On the other hand, every point
(2k+1)πi
znk = µ−1/n e n , n, k ∈ N
n
is a zero of the function f because znk = −1/µ. We claim that for
every t ∈ [0, 2π] there exists a subsequence of (znk )n,k≥1 that converges
nontangentially to eit . Indeed, given n ≥ 1 we can find 0 ≤ kn < n
(2kn +1)πi
such that |t − 2knn+1 π| < πn . Then clearly, znkn = µ−1/n e n → eit
and since
|t − 2knn+1 π| π π
−1/n
< −1/n
→
1−µ n(1 − µ ) log µ
it follows by Exercise 1 that (znkn ) converges nontangentially to eit .
This shows that whenever f has a nontangential limit at some point eit
the limit must be zero. But by Privalov’s theorem the set of these points
must have measure zero (recall that the set in question is measurable!).
Example 2.3. There exist analytic functions f in D that have a radial
limit but no nontangential limit at almost every boundary point!
2. BOUNDARY BEHAVIOR OF POWER SERIES 11
It will turn out that such an erratic behavior near the boundary points
as described in the previous examples, becomes impossible if functions
in question satisfy an appropriate growth restriction. This is an impor-
tant point of view because, as we shall see in the sequel, these growth
restrictions are very easy to define and provide large classes (spaces) of
analytic functions that behave nicely near the boundary.
Theorem 2.2. Let f be analytic and bounded in D. If f has a radial
limit at a boundary point eit then f has a nontangential limit at that
point and these limits coincide.
To this end, consider the sequence (fn ) with fn (z) = f ( nz ), z ∈ Γa,b and
apply Vitali’s theorem. The family {fn , n ≥ 1} is uniformly bounded
in Γa,b and if 0 < r < a then
r
lim fn (r) = lim f ( ) = L.
n→∞ n→∞ n
Since the set (0, a) is not discrete in Γa,b Vitali’s theorem implies that
(fn ) converges uniformly on compacts to the constant function L. Now
let (zk ) be a sequence in Γa,c , 0 < c < b that converges to 0. We
want to show that f (zk ) → L. Choose a sequence (nk ) of integers such
that nk |zk | ∈ (a/2, a) for sufficiently large k (for example, the choice
nk + 1 = integer part of |zak | will do). Then the points wk = nk zk satisfy
a/2 < |wk | < a, | arg wk | < c for sufficiently large k which shows that
these points wk lie in a compact subset of Γa,b . Then by the above
reasoning we have
wk
L = lim fnk (wk ) = lim f ( ) = lim f (zk )
k→∞ k→∞ nk k→∞
Of course, at the first sight the condition in Abel’s theorem seems diffi-
cult to test and has little to do with a growth restriction. Nevertheless
there is a deep result in analysis that yields a global existence the-
orem for nontangential limits of power series with square summable
coefficients. The following famous theorem was proved by L. Carleson.
n
P
Theorem 2.4. (Carleson) Let f (z) = n≥0 an z be a power series
with square summable coefficients, that is,
X
|an |2 < ∞.
n≥0
Proof. Write
∞
! ∞ ! ∞
X X X
it 2 n int n −int
|f (re )| = an r e an r e = an am rm+n ei(n−m)t
n=0 n=0 m,n=0
and note that the last sum converges uniformly on [0, 2π]. But then
we can interchange limit and integration to obtain
Z 2π ∞ Z 2π ∞
it 2 dt dt
X X
m+n
|f (re )| = an am r ei(n−m)t = |an |2 r2n ,
0 2π m,n=0 0 2π n=0
R 2π ikt dt
because 0 e 2π = δk0 , that is, it equals 0 if k 6= 0 and 1 if k = 0.
The rest of the proof is a routine exercise based on the fact that the
integrals involved here increase with r as the last identity shows.
The above result is classical, and, as we shall see in the sequel, its proof
does not need the ”heavy artillery” provided by Carleson’s theorem.
We close this section with a result which shows that, in terms of Taylor
coefficients, the square summability condition in Corollary 2.6 cannot
be improved. The theorem below is a special case of a much stronger
result of Khinchin and Kolmogorov (see [Du],p.).
Theorem 2.6. Let f (z) = n≥0 an z n be a convergent power series in
P
D such that X
|an |2 = ∞.
n≥0
Then there exists a sequence (εn ) with εn = ±1, n = 0, 1, 2 . . . such
that the power series X
g(z) = ε n an z n
n≥0
has no radial limit almost everywhere on the unit circle.
CHAPTER 2
Poisson integrals
1. Harmonic functions
∂ ∂u
Ref = Ref 0 =
∂x ∂x
and
∂ ∂u
Ref = −Imf 0 = .
∂y ∂y
Thus, Ref − u is constant in G and the result follows.
N
X
u(z) = akn z k z n , z∈C
k,n=0
z∈D
Z
1 f (ζ)dζ
(2.1) u(z) = Ref (z) = Re
2πi |ζ|=1 ζ −z
Z 2π it it
1 f (e )e dt
= Re .
2π 0 eit − z
This is an integral representation of u which, unfortunately, involves
one of its harmonic conjugates and therefore needs to be improved.
Lemma 2.1. If u is harmonic in a disc containing D then for all z ∈ D
we have Z 2π
1 1 − |z|2
u(z) = it 2
u(eit )dt.
2π 0 |e − z|
Proof. The trick is to insert in the last integral in (2.1) the harm-
less term zf (eit )/(e−it −z), where f is again an analytic function whose
real part equals u. Note that for every integer n ≥ 0 we have
Z 2π int Z 2π i(n+1)t
ζ n dζ
Z
e dt e dt
= = −i = 0,
0 e−it − z 0 1 − eit z |ζ|=1 1 − ζz
2π
eit
Z
1 it z
u(z) = Re f (e ) it
+ −it dt .
2π 0 e −z e −z
Now note also that
eit z 1 − |z|2
(2.2) + = , z 6= eit ,
eit − z e−it − z |eit − z|2
and that the right hand side of this equality is always real (it becomes
also positive if |z| < 1). Then
Z 2π
1 1 − |z|2
u(z) = Ref (eit )dt
2π 0 |eit − z|2
18 2. POISSON INTEGRALS
Proof. For 0 < r < 1 fixed but arbitrary, consider the dilation ur
of u defined in D by ur (z) = u(rz). Clearly, ur is harmonic in the disc
of radius 1/r centered at the origin, so that Lemma 2.1 gives for every
z∈D Z 2π
1 1 − |z|2
ur (z) = ur (eit )dt.
2π 0 |eit − z|2
Now let r → 1 and note that ur (z) → u(z), z ∈ D and also, that the
functions t 7→ ur (eit ), t ∈ [0, 2π] converge uniformly to t 7→ u(eit ), t ∈
[0, 2π], by the continuity assumption. This implies that for fixed z ∈ D
Z 2π Z 2π
1 − |z|2 it 1 − |z|2
it 2
u r (e )dt → it 2
u(eit )dt, r → 1
0 |e − z| 0 |e − z|
and we are done.
if p = ∞. Then:
(i) If p > 1, there exists a unique ũ ∈ Lp (m) such that
Z 2π
1 − |z|2 dt
u(z) = it 2
ũ(eit ) .
0 |e − z| 2π
2. REPRESENTATION BY POISSON INTEGRALS 19
The kernel
it 1 − |z|2
Pz (e ) = it
|e − z|2
is called the Poisson kernel, and a harmonic function of the form de-
scribed in the above theorem is called the Poisson integral of ũ, or µ.
The simplest example that (i) cannot hold for p = 1 are the Poisson
kernels themselves. For example, Pz (1) satisfies the theorem with p = 1
and is the Poisson integral of the Dirac measure at 1 which is uniquely
determined by this assertion.
Corollary 2.2. Let u be harmonic and nonnegative in D. Then there
exists a unique nonnegative Borel measure µ on T such that µ(T) =
20 2. POISSON INTEGRALS
u(0) and
1 − |z|2
Z
u(z) = dµ(ζ).
T |ζ − z|2
Proof. A nonnegative harmonic function u in D satisfies the as-
sumption in part (ii) of Theorem 2.1 since
Z 2π Z 2π
it dt dt
|u(re )| = u(reit ) = u(0).
0 2π 0 2π
Since the measure µ goiven by the theorem is the weak-star limit of
nonnegative measures, it will be nonnegative as well.
Corollary 2.3. Let f be analytic with nonnegative real part in D.
Then there exists a unique nonnegative Borel measure µ on T such
that µ(T) = Ref (0) and
Z
ζ +z
f (z) = dµ(ζ).
T ζ −z
Proof. (i), (ii), and (iv) are obvious. The first integral in (iii)
equals 1 by an application of Lemma 2.1u ≡ 1, the second equality
follows from Preiθ (t) = Preit (θ). To see (v), write
1 − r2 1 − r2 |λ − rζ|2 |λ − rζ|2
Pz (λ) = = = P rζ (λ) .
|λ − z|2 |λ − rζ|2 |λ − z|2 |λ − z|2
Moreover,
|λ − rζ| |λ − z| + |z − rz| + r|z − ζ| |z − ζ|
≤ ≤1+r+r ,
|λ − z| λ − z| 1−r
and the right hand side is bounded within Γσ (ζ) by a constant that
depends only on σ. Finally, the first part of (vi) is immediate since
∂ 2r(1 − r2 )t sin t
t Pt (eit ) = − ,
∂t |eit − r|2
while the second part follows from this together with integration by
parts:
Z π Z π
∂
t Pr (e ) dt = − ∂ dt
it
∂t 2π t Pr (eit )
−π −π ∂t 2π
Z π
dt
= −Pr (−1) + Pr (eit )
−π 2π
1−r
=− + 1.
1+r
Proof. Let z ∈ Γσ (eiθ ), with |z| = r ∈ [0, 1), and use Proposition
3.1 (v) to obtain the estimate
Z Z
(3.5) |u(z)| ≤ Pz (λ)|h(λ)|dm(λ) ≤ C Preiθ (λ)|h(λ)|dm(λ).
T T
4. BOUNDARY BEHAVIOR OF POISSON INTEGRALS 23
Proof. Let ε > 0, and let g ∈ C(T) with kh − gk1 < ε. By the
Hardy-Littlewood theorem it follows that
√ √
m({ζ ∈ T : (h − g)† (ζ) > ε}) < C1 ε.
Let v denote the Poisson integral of g. By Theorem 3.2 we have that
√ √
m({ζ ∈ T : (u − v)∗ (ζ) > C ε}) < C1 ε,
where C > 0 is the constant given in that theorem. For a fixed integer
n > 0 consider the set An ⊂ T consisting of those points ζ for which
the nontangential ”limsup” of |u − h(ζ)| at ζ exceeds n1 . Since v is
24 2. POISSON INTEGRALS
dµ
where h = dm ∈ L1 (m) is the Radon-Nykodim derivative of µ w.r.t.
m, and ν is singular w.r.t. m. By Theorem 4.1 and the above lemma
we obtain:
Corollary 4.1. If u is the Poisson integral of the finite Borel measure
dµ
µ on T then u has the nontangential limit dm (ζ) at almost every ζ ∈ T.
5. hp -spaces
and if p = 1, then
kuk1 = kµu k.
For 1 ≤ p ≤ ∞, hp is a Banach space with the norm (5.6).
(iii) If 1 < p < ∞, and u ∈ hp , ur (z) = u(rz), 0 ≤ r < 1, then
lim kur − ukp = 0.
r→1−
Exercise 1 Show that part (iii) of Theorem 5.1 fails for p = ∞ when-
ever u has a discontinuous boundary function, or when p = 1, and u is
the Poisson integral of a measure with a nonzero singular part.
It turns out that the necessary condition above is also sufficient for our
embedding. Let us rewrite this condition in a more transparent way.
Proposition 6.1. Given 0 < h < 1, and θ ∈ [0, 2π], let
Sh (eiθ ) = {reit : 1 − h < r < 1, |t − θ| < h}.
A positive Borel measure ν on D satisfies (6.9) if and only if
ν(Sh (eiθ )) ν({|z − ζ| < r})
sup < ∞, or sup < ∞.
0<h<1 h ζ∈T r
θ∈[0,2π] r>0
which shows that if (6.9) holds, then the second supremum above is
finite. Conversely, fix λ ∈ D with |λ| ≥ 1/2, and let
En = {z ∈ D : 2n (1 − |λ|2 ) ≤ |1 − λ| < 2n+1 (1 − |λ|2 )}.
Then Z X ν(En )
(1 − |λ| )2
|1 − λz|−2 dν(z) ≤ .
n
22n (1 − |λ|2 )
Moreover, if z ∈ En , then
1 λ
|z − | < |1 − λz| + 1 − |λ| < 2n+2 (1 − |λ|2 ),
2 |λ|
hence,
ν({|z − ζ| < r})
ν(En ) ≤ 2n+2 (1 − |λ|2 ) sup ,
ζ∈T r
r>0
which gives
ν({|z − ζ| < r})
Z
2
(1 − |λ| ) |1 − λz|−2 dν(z) ≤ C sup ,
ζ∈T r
r>0
when λ ∈ D with |λ| ≥ 1/2. Since the integrals on the left are obviously
bounded or |λ| < 1/2, the result follows.
The sets Sh have been used by Carleson who was the first to study
embeddings of this type. They will be called Carleson boxes, and the
measures satisfying one of the conditions in Proposition 6.1 are called
Carleson measures. The next theorem is essentially due to Carleson.
Theorem 6.1. If ν is a finite positive Borel measure on D, and 1 <
p < ∞ then hp ⊂ Lp (ν) if and only if ν is a Carleson measure.
which implies the claim and completes the proof of the theorem.
Example 6.1. As pointed out in Exercise 1, the one-dimensional Lebesgue
measure on the segment (−1, 1) is a Carleson measure, that is, there
exists C > 0 such that
Z 1
|u(x)|p dx ≤ Ckukpp ,
−1
6. EMBEDDINGS OF hp INTO Lp (ν) 31
and by Theorem 5.1 (iii) it follows that ∇ur − ∇uρ converges to zero in
1
L2 (D, log |z| dA), when r, ρ → 1− . This easily shows that when λ = 0,
(6.11) holds for every u ∈ h2 .
z−λ
If λ ∈ D is arbitrary, let φλ (z) = 1−λz . Clearly, φλ is an analytic
bijection from D onto itself. By what we have proved above, we have
for every u ∈ h2 ,
Z Z
2 1
|u ◦ φλ − u(λ)| dm = |∇u ◦ φλ |2 (z) log dA(z).
T D |z|
A direct computation similar to the above yields
|∇u ◦ φλ |2 (z) = |∇u|2 (φλ (z))|φ0λ (z)|2 ,
i.e.
Z Z
1
2
|u ◦ φλ − u(λ)| dm = |∇u|2 (φλ (z))|φ0λ (z)|2 log dA(z).
T D |z|
Then (6.11) follows directly from the change of variables ξ = φλ (ζ), w =
φλ (z).
1
Corollary 6.1. Let u ∈ h2 . Then the measure |∇u|2 (z) log |z| dA, is
a Carleson measure if and only if
Z
sup |u(ζ) − u(λ)|2 Pλ (ζ)dm(ζ) < ∞.
λ∈D T
Hardy spaces
1. Outer functions.
and Z
ζ +z
F1 (z) = exp log max{h(ζ), 1}dm(ζ) ,
T ζ −z
are analytic and bounded in D, hence both functions have nontangential
limits almost everywhere on the unit circle. By the above computation
we also know that these limits are nonzero m− a.e.. Thus F = F1 /F2
F has nontangential limits almost everywhere on the unit circle as
well.
Exercise 1. Show that the product and quotient of two outer functions
is outer. Moreover, prove that if F is outer with |F (eit )| ≥ a > 0
a.e. then |F (z)| ≥ a for all z ∈ D. Finally, show that any function
F that is analytic in a larger disc (than D) and has no zeros in D is
an outer function, but the statement is no longer true if we replace the
assumption ”F analytic in a larger disc” by ”F continuous in D.
Proposition 1.2. Let h be a nonnegative function on the unit circle
such that Z
| log h|dm < ∞
T
and let Fh be the outer function whose modulus equals h a.e. on the
boundary. Then
Z
|Fh (z)| ≤ Pz (ζ)h(ζ)dm(ζ).
T
2. Zeros of H p -functions
so that we can consider the outer function Fr with |Fr (ζ)| = |f (rζ)|,
m-a.e. on T . Moreover, if f has no zeros on rT, then |Fr | is bounded
below, hence fr /Fr is bounded in D, where, as usual fr (z) = f (rz), z ∈
D. Since the nontangential limits of fr /Fr have modulus one a.e. on
the circle, it follows by the Poisson representation that |fr /Fr | ≤ 1,
that is,
|Fr (z)| ≥ |f (rz)|, z ∈ D.
Then for 0 < ρ < 1 we have by Corollary 1.1
Mp (rρ, f ) ≤ kfr kpp ≤ kFr kpp = Mp (r, f ).
Thus we have shown the inequality Mp (rρ, f ) ≤ Mp (r, f ) for all ρ ∈
(0, 1) and all r in a dense subset of [0, 1]). The result follows.
Theorem 2.1. Let 0 < p ≤ ∞, and assume that f ∈ H p is not
identically zero. Let a1 , a2 , . . . be the zeros of f in D, repeated according
to their multiplicity. Then
X
(1 − |an |) < ∞.
n≥1
valid for all r ∈ (0, 1). Thus the Blaschke condition in the statement
holds true. Letting N → inf ty, we obtain from the last estimates
Mp (r, f /B) ≤ kf kpp , max |f /B(z)| ≤ kf k∞ ,
|z|=r
z−a
with equality if and only if f has the form f (z) = α 1−az for some fixed
a ∈ D and α ∈ ∂D. In particular, for all z ∈ D we have also
1
|f 0 (z)| ≤ .
1 − |z|2
3. Applications
The norm convergence of dilations also extends for all values of 0 <
p < ∞ as the following result shows.
Theorem 3.2. If f ∈ H p , 0 < p < ∞, and for 0 ≤ r < 1, fr (z) =
f (rz), z ∈ D, then
lim− kf − fr kp = 0.
r→1
The next theorem is a famous result due to the brothers Riesz with wide
applications in function theory. The original proof is quite different
from the one below and is of interest in its own right.
Theorem 3.4. Let µ be a finite Borel measure on T with the property
that Z
ζ n dµ(ζ) = 0,
T
for all nonnegative integers n. Then µ is absolutely continuous w.r.t.
dµ
m and there exists f ∈ H 1 with f (0) = 0 such that dm = f.
The outer function F in Proposition 3.2 will be called the outer factor
of f .
Definition 3.1. A bounded analytic function I in D is called inner if
its nontangential limits satisfy |I(ζ)| = 1, m− a.e. on T.
At its turn, by Theorem 2.1 the inner factor can be further decomposed
as I = BS, where B is a Blaschke product and a function S which is
zero-free in D. Here we consider B = 1 if I has no zeros to begin with.
Definition 3.2. An inner function without zeros in D is called sin-
gular inner.
Then
Z Z Z
p
|f (ϕ(rz))| dm(z) ≤ Pϕ(rz) (ζ)|f (ζ)|p dm(ζ)dm(z)
T
ZT ZT
= |f (ζ)|p Pϕ(rz) (ζ)dm(z)dm(ζ),
T T
The dual of H p
since
∞
X z̄ n X ∞
ζ̄
= n
= z̄ n ζ n , ζ ∈ T, z ∈ D,
ζ̄ − z̄ n=0
ζ̄ n=0
with [g] ∈ Lq (m)/(H p )⊥ and kgkLq (m) ≤ 2k[g]k = klk. Moreover, since
fr → f in H p , when r → 1− we obtain easily that l(fr ) → l(f ), i.e.
Z
l(f ) = lim− f (rζ)g(ζ)dm(ζ).
r→1 T
Using the Cauchy formula as in Corollary 3.4 we have for ζ ∈ T
Z
zf (z)
f (rζ) = dm(z), ζ ∈ T, 0 < r < 1,
T z − rζ
and by an application of Fubini’s theorem we obtain
Z Z Z
g(ζ)
f (rζ)g(ζ)dm(ζ) = zf (z) dm(ζ)dm(z)
T T T z − rζ
Z Z
ζ̄g(ζ)
= f (z) dm(ζ)dm(z).
T T ζ̄ − rz̄
Then the result follows with
Z
ζg(ζ)
h(z) = dm(ζ), z ∈ D,
T ζ −z
using the considerations preceeding the proposition.
This formula gives 2 analytic functions, one in D, and the other in C\D,
but we will be mostly concerned with the first, since the properties of
the second are very similar. In fact their boundary values are related
by the so-called ”jump theorem”.
Theorem 1.1. If h ∈ L1 (m)
1
h(rz) − b
lim− b h( z) = z̄h(z)
r→1 r
m−a.e. on T.
satisfies
Ref (z) = u(z), z ∈ D, , f (0) ∈ R.
Thus for such functions, we can reformulate the above in terms of real
parts:
If f is analytic in D and Ref ∈ hp , 1 ≤ p ≤ ∞, does f ∈ H p (or
Imf ∈ hp )?
Obviously the original question reduces to real-valued functions, so that
the two are equivalent.
It turns out that the answer is negative for the ”endpoints” p = 1 and
p = ∞. Indeed, if
1+z
f1 (z) = f2 (z) = i log(1 − z), z ∈ D,
1−z
then Ref1 (z) = Pz (1), so that Ref1 ∈ h1 , and Ref2 ∈ h∞ . However,
/ h1 , because its boundary function is not integrable, and f2 is ob-
f1 ∈
viously unbounded.
kb
hkp ≤ Cp khkLp (m) .
Equivalently, if f is analytic in D, f (0) ∈ R and Ref ∈ hp , then
f ∈ H p and there exists Cp0 > 0 such that
kf kp ≤ Cp0 kRef kp .
2. THE M. RIESZ THEOREM 57
where g1 (ζ) = ζg(ζ). Using the first part of the proof, the fact that
1 < q < 2, and the equivalence of the two statements, it follows that
G(z) = z gb1 (z) belongs to H q with
q
kGkqq ≤ kgkqLq (m) .
q−1
Thus by Hölder’s inequality
Z 1/q
q
h(rζ)g(ζ)dm(ζ) ≤ khkLp (m) kGkq ≤ q − 1 khkLp (m) kgkLq (m) ,
b
T
58 4. THE DUAL OF H p
which gives
1/q
q
k(b
h)r kp ≤ khkLp (m) ,
q−1
and the result follows letting r → 1− .
3. The dual of H 1
Moreover,
ζ − āζz − a + z 1 − āζ
ζ − φa (z) = = (−φa (ζ) + z).
1 − āz 1 − āz
From these three computations we obtain
1 1 (1 − |a|2 )z
− = , z ∈ D , ζ ∈ T,
ζ − φa (z) ζ − a (φa (ζ) − z)(1 − āζ)(ζ − a)
and since ζ ∈ T, we have ζ = ζ̄1 , i.e.
1 1 z ζ̄ (1 − |a|2 ) z ζ̄
− = = |φ0 (ζ)|.
ζ − φa (z) ζ − a (φa (ζ) − z) |1 − āζ|2 (φa (ζ) − z) a
Thus,
Z
ζ̄
h(φa (z)) − b
b h(a) = z h(ζ)|φ0a (ζ)|dm(ζ).
T (φa (ζ) − z)
Now use the fact that φ−1
a = φa , together with the change of variable
Z Z
0
g ◦ φa |φa |dm = gdm,
T T
to obtain
Z
φa (ζ)h(φa (ζ)
h(φa (z)) − b
b h(a) = z dm(ζ),
T ζ −z
which is the formula in the statement.
sup kb
h−b
h(a)k2 ≤ CkhkL∞ (m) .
a∈D
for some C > 0 and all f ∈ H 1 , 0 < r < 1. Indeed, if (3.16) holds then
for r, ρ → 1− , r > ρ, we have fr − fρ = (f − fρ/r )r , hence
Z
(fr − f ρ)gdm ≤ Ckf − fρ/r k1 → 0,
T
From the discussion at the beginning of the chapter we now obtain the
following description of the dual of H 1 . Until now, for the definition of
BM OA we have used the seminorm
sup kg ◦ φa − g(a)k2 .
a∈D
Proof. Let u = Ref . Since |∇u|2 = 2|f 0 |2 , the measure |∇u|2 log |z|
1
dA
0 2 1
is Carleson if and only if |f | log |z| dA is.
One of the basic applications of the Cauchy formula is the fact that
the set of holomorphic functions in a given region in the complex plane
is closed with respect to uniform convergence on compacts. That is,
given any sequence (fn ) of holomorphic functions in the open set U ⊂ C
which converges uniformly on every compact subset of U to a function
f : U 7→ C, it follows that f is also holomorphic on U . This is an
important result which, in particular, enables us to construct highly
nontrivial examples of holomorphic functions. Some of these are listed
below as exercises.
∞
X 1
(i) ζ(z) = z
, Re z > 1, (Zeta function),
n=1
n
Z ∞
dt
(ii) Γ(z) = e−t tz , Re z > 0, (Gamma function),
0 t
∞
X 1
(iii) f (z) = , z ∈ C \ N, (no special name),
n=1
(z − n)2
∞
1 X 1 1
(iv) ℘(z) = 2 + − ,
z m,n∈Z
(z − m − in)2 (m + in)2
m2 +n2 6=0
(fn1k ) such that (fn1k (z1 )) converges. This sequence has at its turn a
subsequence (fn2k ) such that (fn2k (z2 )) converges. If we continue this
process indefinitely we obtain subsequences (fnj ) such that (fnj (zl ))
k k
converges for 1 ≤ l ≤ j. If we set nk = nkk then clearly, the subsequence
(fnk ) has the property that (fnk (zj )) converges for all j = 1, 2, . . .. This
is called a diagonalization process.
Let us now show that (fnk ) converges uniformly on compacts. Note
first that if w ∈ U and ε > 0 then by the claim (equicontinuity) there
is δ > 0 such that
|fnk (z) − fnk (w)| < ε
for all k ≥ 1 and z ∈ U with |z − w| < δ. Then, if z, ζ ∈ U satisfy
|z − w| < δ and |ζ − w| < δ, we have
|fnk (z) − fnk (ζ)| ≤ |fnk (z) − fnk (w)| + |fnk (ζ) − fnk (w)| < 2ε.
Fix a point zj (j depends on w) in the above set, such that |zj −w| < δ.
Then there exists k0 depending only on w such that for k, p ≥ k0
|fnk (zj ) − fnp (zj )| < ε.
Putting this together we obtain for all z ∈ U with |z − w| < δ that
|fnk (z)−fnp (z)| ≤ |fnk (z)−fnk (zj )|+|fnk (zj )−fnp (zj )|+|fnp (zj )−fnp (z)| < 3ε
for all k, p ≥ k0 .
Finally, let K ⊂ U be compact ε, δ > 0 as above and cover K by finitely
many sets of the form {z, |z − w| < δ} ∩ U, w ∈ K. Then the above
reasoning shows that there exists k1 ≥ 1 such that for all z ∈ K and
k, p ≥ k1 we have
|fnk (z) − fnp (z)| < 3ε.
Since K was arbitrary we obtain that (fnk ) converges pointwise to some
function f and from the above inequality, letting p → ∞ we get that
for every compact set K ⊂ U there exists k1 ≥ 1 such that for all
z ∈ K and k ≥ k1 we have
|fnk (z) − f (z)| < 3ε,
i.e. (fnk ) converges uniformly on compacts to f .
It remains to prove the claim from the beginning of the proof, i.e. the
equicontinuity of the family F. For z0 ∈ U let ∆0 be a disc of radius
r0 > 0 with ∆0 ⊂ U . By hypothesis, we know that there exists a
constant C0 such that
sup |f (z)| ≤ C0
z∈∆0
68 A. SEQUENCES AND FAMILIES OF HOLOMORPHIC FUNCTIONS