The Kernels of The Hankel and Toeplitz Operator
The Kernels of The Hankel and Toeplitz Operator
Jim-Felix Lobsien
Date: August 4, 2011
1
CONTENTS 2
Contents
1 List of Symbols 3
Introduction 4
1.1 Foreword . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2 Preliminaries 5
2.1 Preliminary Introduction into the Topic . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.2 An Introduction to H
p
Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.3 The Poisson kernel and its mollier properties . . . . . . . . . . . . . . . . . . . . . . . . 9
2.4 A digression on harmonic functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.5 Boundary Behaviour . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.6 Back to Analytic Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.7 Projections from L
p
to H
p
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3 Necessary Tools 18
3.1 Jensens Inequality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.2 Factorization for functions in H
p
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.3 The Nevanlinna class . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.4 The forward and backward shift on H
p
, for p [1, ) . . . . . . . . . . . . . . . . . . . 23
4 Kernels of Hankel operators 24
4.1 The symbols of the Hankel operator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.2 Description of the kernels of Hankel operators . . . . . . . . . . . . . . . . . . . . . . . . 25
5 Kernels of Toeplitz operators 26
5.1 The symbols of Toeplitz operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
5.2 Description of the kernels of Toeplitz 0perators . . . . . . . . . . . . . . . . . . . . . . . 28
6 Appendix 36
6.1 Factorization for functions in H
p
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
6.2 Theorems used in this paper . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
1 LIST OF SYMBOLS 3
1 List of Symbols
1. T denotes the unit circle T := z C : [z[ = 1.
2. D denotes the unit disk D := z C : [z[ < 1.
3. N
0
are the natural numbers including the zero.
4. N are the natural numbers without the zero.
5. L
p
(T) is the set of all measurable functions from T to C whose absolute value raised to the p-th
power has nite integral.
6. w
f(e
i
)
2
dm()
1
2
< .
2. The corresponding Fourier series of f is given by
f(e
i
) =
nZ
f(n)e
in
with
f(n) =
T
f(e
i
)e
in
dm().
Remark
Due to the Riesz-Fischer theorem the series converges in L
2
-norm if and only if the function is square
integrable and Lennart Carleson proved that the Fourier expansion of any function in L
2
converges
almost everywhere.
Therefore we have a new but equivalent denition of L
2
.
L
2
:=
f(e
i
) =
nZ
a
n
e
in
: a
n
C,
nZ
[a
n
[
2
<
.
With the aid of this denition we are able to dene a closed subspace of L
2
.
Denition 2.0.2 The space
H
2
:=
f L
2
:
f(n) = 0 for n < 0
nN
0
, where
the coecients are square summable. Hence it is a Hilbert space.
We are going to use now the most useful feature of a Hilbert space; the fact that each closed subspace
of L
2
has a natural complementary closed subspace.
Denition 2.0.3 The orthogonal complement of H
2
is dened as
H
2
= f L
2
: (f, g) = 0 g H
2
0
nZ
a
n
b
n
with f(e
i
) =
nZ
a
n
e
in
and g(e
i
) =
nZ
b
n
e
in
.
2 PRELIMINARIES 6
Consequently the space L
2
is the direct sum of the two subspaces (L
2
= H
2
H
).
Later on we will concentrate on the space L
p
:= L
p
(T, dm) on the circle with respect to normalized
Lebesgue measure which is unfortunately not a Hilbert space anymore for p = 2. Therefore the
orthogonal compliment does not exist and we will deal with the complex conjugate of H
2
, where the
zero Fourier coecient vanishes.
H
2
0
:=
f : f H
2
,
f(0) = 0
.
To illustrate that both denitions are equivalent, we state the following
Lemma 2.1
H
2
= H
2
0
.
Proof "" Let f
H
2
with f(e
i
) =
nZ
a
n
e
in
, then
(f, g) =
nZ
a
n
b
n
= 0 f H
2
with g(e
i
) =
nN
0
b
n
e
in
.
But since b
n
= 0n N
0
it follows that a
n
= 0n N
0
and hence f(e
i
) =
nN
a
n
e
in
. Let
c
n
:= a
n
we have
f(e
i
) =
nN
c
n
e
in
=
nN
c
n
e
in
= g with g H
2
0
.
"" Let f H
2
0
with f =
nN
c
n
e
in
. If we dene a
n
:= c
n
it is obvious that (g, f) = 0 g H
2
,
since f has no coecients for n 0.
Remark The proof above illustrates that another denition holds for the complex conjugate of H
2
,
where the zero Fourier coecient vanish.
H
2
0
:=
f L
2
:
f(n) = 0 for n 0
.
As is customary, we are going to use H
2
0
with the above denition instead of
H
2
.
The direct sum decomposition gives rise to natural projections, the so-called Riesz projections.
Denition 2.1.1 Let f be in L
2
with the Fourier series f(e
i
) =
nZ
f(n)e
in
, then
1. the operator P
: L
2
H
2
0
P
(f) :=
nN
f(n)e
in
is called the Riesz projection onto H
2
0
.
2. the operator P
+
: L
2
H
2
P
+
(f) :=
nN
0
f(n)e
in
is called the Riesz projection onto H
2
.
Remark The operators are well-dened because cutting o the positive or negative part still guarantees
the convergence of the Fourier series, and since the coecients are still square summable, the image of
the projections is still in L
2
and hence due to the denitions in H
2
or H
2
0
.
Both projections are bounded since L
2
is the sum of two closed subspaces. Therefore it is quite natural
to dene the following two operators.
2 PRELIMINARIES 7
Denition 2.1.2 Let f H
2
and L
.
1. The operator H
: H
2
H
2
0
H
f := P
(f)
is called the Hankel operator with symbol .
2. The operator T
: H
2
H
2
T
f := P
+
(f)
is called the Toeplitz operator with symbol .
Remark The Hankel and Toeplitz operator are bounded due to the boundedness of the Riesz projec-
tions.
2.2 An Introduction to H
p
Spaces
The Hankel and Toeplitz operators in Section 4 and 5 are dened for H
p
with p [1, ]. Thus our
rst step is to dene them, but the question arises is how? Introducing them using the same route like
we did for H
2
fails due to the following problems:
1. Although Carlesons theorem which he published in 1966 was later generalized by Richard Hunt
for p > 1, in 1922 already Andrey Kolmogorov constructed an example of a function in L
1
whose
Fourier series diverges everywhere.
2. Leaving aside the convergence of the Fourier series, cutting o the negative parts of a function in
L
1
or L
nN
0
f(n)e
in
with
f(n) =
T
f(e
i
)e
in
dm().
Let us have a look at the functions
f(z) =
nN
0
f(n)z
n
with
f(n) =
T
f(e
i
)e
in
dm() and [z[ < 1.
Since [z[ < 1 the pointwise convergence of f(z) follows from the pointwise convergence for [z[ = 1 and
hence f is analytic in D. We arrive at the following denition.
Denition 2.1.3 The space
H
2
:= f analytic in D : the coecients are square summable
is called the Hardy-Hilbert space on D with norm
|f|
H
2
=
nN
0
[a
n
[
2
1
2
.
2 PRELIMINARIES 8
There is a natural identication between H
2
and
H
2
. Namely we identify the function f H
2
as
having Fourier series
nN
0
a
n
e
in
with the analytic function f(z) =
nN
0
a
n
z
n
. This identication
is clearly an isomorphism between H
2
and
H
2
.
Of course, this identication, although natural, does not describe the relationship between f H
2
and
f
H
2
as functions (We come back to this later in the next subsection).
Theorem 1.1.12 in [2] (Pg. 7) provides an insight into the denition of H
p
functions since one just has
to replace 2 with p. It states the following:
Theorem 2.2 Let f be analytic on D. Then f
H
2
if and only if
sup
0<r<1
T
[f(re
i
)[
2
dm() < .
Moreover for f
H
2
|f|
2
H
2
= sup
0<r<1
T
[f(re
i
)[
2
dm().
Of course this was not the reason why Frigyes Riesz introduced the H
p
spaces in [4] (1923) which he
named after Godfrey Harold Hardy because of his paper [3] published in 1915.
The maximum modulus principle says that for f analytic in D the maximum modulus function
M
(r, f) := max
[f(re
i
)[ : [0, 2]
, r (0, 1)
is an increasing function of r. For a xed p (0, ) G.H. Hardy proved in [3] the same result for the
integral means
M
p
(r, f) :=
T
[f(re
i
)[
p
dm()
1
p
, r (0, 1)
and from here it is quite natural to consider the following class of analytic functions.
Denition 2.2.1 The space
H
p
(D) :=
f analytic in D : sup
0<r<1
M
p
(r, f) <
M
p
(r, f).
Denition 2.2.2 When p = , we dene H
(r, f) < .
These functions are precisely the bounded analytic functions on D.
From the basic inequality
(a +b)
p
2
p
(a
p
+b
p
) a 0, b 0
it follows that H
p
is a complex linear space and with
1 p < : |f|
H
p := sup
0<r<1
M
p
(r, f)
p = : |f|
H
:= sup
0<r<1
M
(r, f)
2 PRELIMINARIES 9
it is a normed space. One easily checks, that H
p
for 1 p is a Banach space (Remark 17.8 in [8]).
Furthermore a simple application of Hoelders inequality shows that for p, q [1, ], p < q, the
inequality
M
p
(r, f) M
q
(r, f)
holds for all r (0, 1). Therefore H
q
H
p
.
2.3 The Poisson kernel and its mollier properties
Before we continue our introduction let us have a look at the Poisson kernel, since it plays a major part
in this theory.
Denition 2.2.3 The function
{
r
() =
1 r
2
1 +r
2
2r cos
with (, ] is called the Poisson kernel.
It also has the following form
{
r
() = {
r
( ) = Re
1 +re
i
1 re
i
= Re
e
i
+re
i
e
i
re
i
= Re
e
i
+z
e
i
z
= {
z
()
with = and z = re
i
.
Theorem 2.3 The Poisson kernel has the following properties also referred to as mollier properties
1. {
z
() > 0 for z D and (, ].
2. {
z
( + 2) = {
z
().
3. for each z D
T
{
z
()dm() = 1.
4. Given a > 0
{
z
() 0 uniformly for [[ as r 1.
2.4 A digression on harmonic functions
We arrived at a point where we dened a class of analytic functions in the unit disc, but the connection
to the space L
p
is still missing. The connection does not seem to be natural since H
p
is living on the
unit disk and L
p
on the unit circle. In fact we need to talk in terms of harmonic functions, since the
theorems we need were proved for this class of functions. Therefore we dene the class of harmonic
functions with bounded integral means.
Denition 2.3.1 1. Let p [1, ), then the space
h
p
(D) := U harmonic in D : sup
0<r<1
M
p
(r, U) <
is called the harmonic Hardy space h
p
in the disc.
2 PRELIMINARIES 10
2. If p = , we dene h
(r, U) < .
These functions are precisely the bounded harmonic functions in D.
The following theorem gives us the connection between harmonic functions in the disk and integrable
functions on the unit circle. It was published by Pierre Fatou in his famous paper [9] in 1906 and is
the starting point of the whole subject dealt with here.
Theorem 2.4 Let 1 < p and U(z) h
p
. Then there is a F L
p
such that
U(z) =
T
{
z
()F()dm().
The proof of this theorem uses strongly the fact, that L
q
is the dual of L
p
with
1
q
+
1
p
= 1. But the
problem is that L
1
itself is not the dual of any Banach space. A way out of this dilemma is to use
the space of nite complex measures on T, because it is the dual of C(T) ,i.e. the space of continuous
functions on T.
If p L
1
, we can associate to p a complex measure
p
, by putting
T
G()d
p
() =
T
G()p()d.
With this argument one can obtain by using the same route of the proof of Theorem 2.4 the following
Theorem 2.5 Let U(z) h
1
. Then there is a nite complex measure on T such that
U(z) =
T
{
z
()
d()
2
.
Conversely there is a theorem which for a given function in L
p
provides the corresponding harmonic
function in D. It states the following
Theorem 2.6 If 1 p and f L
p
and
U(z) =
T
{
z
()F()dm()
then U(z) h
p
.
The same result holds for complex measures.
Theorem 2.7 Let be a nite complex measure on T. Then
U(z) =
T
{
z
()
d()
2
is in h
1
.
2.5 Boundary Behaviour
We are now able to assign an harmonic function in the unit disc with an integrable function on the unit
circle and conversely. It would be convenient if the integrable function was the limit of the harmonic
function, if z tended to the boundary, since consequently a natural isomorphism would arise between
the functions in D and on T. Indeed with a few restrictions this is the case but let us rst examine the
useful L
p
and w
convergence.
2 PRELIMINARIES 11
Theorem 2.8 Let F L
p
, 1 p < and let
U(z) =
T
{
z
()F()dm()
then
T
U(re
i
) F()
p
d 0 as r 1
,i.e. U(re
i
) approaches F() in L
p
-norm, as r 1.
If p = or is a complex measure all we have is weak-star convergence (w
).
Theorem 2.9 1. If F L
and
U(z) =
T
{
z
()F()dm()
then U(re
i
) F() in w
, as r 1.
2. Let
U(z) =
T
{
z
()
d()
2
with a nite complex measure on T. Then U(re
i
)d d() in w
, as r 1.
Remark The proof of the above theorem needs in particular the mollier properties of {
z
()
1
2
.
Therefore all kinds of other kernels besides the Poisson kernel would work to yield similar results.
Since we now have the radial convergence in L
p
-norm and w
T
{
z
()F()dm() with F L
p
can be subsumed to the case
U(z) =
T
{
z
()
d()
2
since d() := F()d is a nite complex measure on T. The theorem of Fatou, which deals with the
non-tangential limits of the Poisson integral says the following.
Theorem 2.10 Let
0
T and suppose that the derivative
(
0
) exists and is nite. Then
U(z) =
T
{
z
()
d()
2
tends to
(
0
) for z = re
i
tending to e
i
0
, from within any region of the form [
0
[ c(1 r), the
so-called Stolz angle.
Remark Strictly speaking it is the Stolz angle not a sector with straight sides and vertex at e
i
0
.
However it becomes asymptotic to such a sector near the point e
i
0
.
2 PRELIMINARIES 12
Let F L
p
, p 1 and let
U(z) =
T
{
z
()F()dm().
A classical theorem of Lebesgue says that
d
d
0
F(t)dt exists a.e. and equals F().
In conjunction with Fatous theorem (Theorem 2.10) we see, that a.e. in T
U(re
i
) F(
0
) as re
i
e
i
0
where
means within the Stolz angle. Combined with Theorem 2.4 we obtain the following
Theorem 2.11 Let 1 < p and let U(z) h
p
. Then for almost all
0
, U(z) tends to a nite limit,
say U(e
i
0
), as z = re
i
e
i
0
. The function
0
U(e
i
0
) L
p
and for all z D we have
U(z) =
T
{
z
()U(e
i
)dm().
Remark U(e
i
) is called the (non-tangential) limit of U(z). We frequently write
U(e
i
) = lim
z
e
i
U(z) a.e..
In case p = 1 the theorem is not completely true, since we are dealing in that case with a measure
d(). The decomposition theorem of Lebesgue says that then the derivative
() L
1
but that d() is not in general
()d. Instead
d() =
()d +d
s
()
where
s
is a singular measure.
Thus if we barely know that U(z) L
1
, we still have a.e. existence of the nite non-tangential limit
lim
z
e
i
U(z) =
()
but we cannot recover U(z) from this boundary value function. Instead we have
U(z) =
T
{
z
()
()dm() +
T
{
z
()
d
s
()
2
with some singular
s
.
Example The ordinary Poisson kernel shows that a representation with nonzero
s
can actually occur.
U(re
i
) =
1 r
2
1 +r
2
2r cos
.
It follows from this representation that
lim
z
e
i
U(z) = 0
save for = 0 and
U(z) =
1
2
T
{
z
()2d
s
()
where
s
is the unit point mass at 0.
This distinction between the cases p = 1 and p > 1 is one of the fundamental complications of the
theory and will be seen to have deep and far-reaching implications in its further developments.
2 PRELIMINARIES 13
2.6 Back to Analytic Functions
It is a well known fact that analytic functions are harmonic and since the only dierence between f h
p
and f H
p
is the analyticity, all the theorems in the last subsection apply for f H
p
.
Therefore we have for f H
p
, p > 1
f(z) =
T
{
z
()f(e
i
0
)dm()
and
f(e
i
) = lim
z
e
i
f(z) a.e..
Hence we have a natural isomorphism between the function f H
p
(D) and f H
p
(T). Since f on the
circle is in L
p
, we can view H
p
as a subspace of L
p
.
If p = 1, we saw in the previous subsection that we utterly failed to construct an isomorphism be-
tween the function on T and the function in D.
But since f is now analytic the situation changed after the proceedings of the fourth scandinavian
mathematical congress in 1917. There Frigyes and Marcel Riesz published the celebrated
Theorem 2.12 (F. and M.Riesz) If
T
e
in
d() = 0 for n = 1, 2, 3, . . .
then is absolutely continuous with respect to Lebesgue measure.
This theorem is the key in the following context.
If f H
1
, then we have since f is in particular harmonic for [z[ < 1 by Theorem 2.5
f(z) =
T
{
z
()
d()
2
for some complex measure (here complex valued) on T. By Theorem 2.7 we obtain
f(re
i
)d d() in w
(2.6.1)
as r 1.
Since f H
1
, the following holds.
Lemma 2.13 If f H
1
, then
T
e
in
d() = 0 for n = 1, 2, 3, . . . .
Proof Let = e
i
, then
F
z
() :=
+z
z
=
1 +z
1 z
= 1 + 2
z
1 z
= 1 + 2
kN
z
k
k
since [z
k
k
[ = [z[
k
< 1.
2 PRELIMINARIES 14
Therefore
{
z
() = Re [F
z
()] =
1
2
F
z
() +F
z
()
=
1
2
+
kN
z
k
k
+
1
2
+
kN
z
k
k
=
kN
0
z
k
k
+
kN
z
k
k
=
1
1 z
+
kN
z
k
k
.
By Theorem 2.5 we obtain
f(z) =
T
{
z
()
d()
2
=
T
1
1 z
d()
2
+
kN
z
k
k
d()
2
.
With [z[ < r it follows that [z
k
k
[ < r
k
and therefore the sum uniformly converges on compact subsets
which gives us the possibility to arrive at the following
f(z) =
T
1
1 z
d()
2
+
1
2
kN
z
k
k
d().
Since f and the left part of the sum is due to the Cauchy integral formula analytic, the anti-analytic
part on the right has to vanish. Therefore
0 =
n
d() =
T
e
in
d() for n = 1, 2, 3, . . . .
Now the theorem of the Riesz brothers (Theorem 2.12) guarantees that is absolutely continuous and
therefore due to the Radon-Nikodym theorem
d() = h()d for some h L
1
.
So in this case we really have
f(z) =
T
{
z
()h()dm()
with a function h.
The distinction between this case and the more general case (f merely harmonic) is very important for
the whole development of the theory.
By Fatous theorem (Theorem 2.10), we now have f(z) h() a.e. for z
e
i
, so if we call
f(e
i
) = lim
z
e
i
f(z)
we have
f(z) =
T
{
z
()f(e
i
)dm()
for f H
1
. Due to our observation before H
1
L
1
and consequently the case p = 1 is not special
anymore for analytic functions and we can state the following theorem.
2 PRELIMINARIES 15
Theorem 2.14 Let 1 p and let f(z) H
p
. Then for almost all
0
, f(z) tends to a nite limit,
say f(e
i
0
), as z = re
i
e
i
0
. The function
0
f(e
i
0
) L
p
and for all z D we have
f(z) =
T
{
z
()f(e
i
)dm().
Due to the following theorem the boundary functions are unique since H
1
contains all other H
p
spaces
for p > 1.
Theorem 2.15 Let f H
1
and suppose, for a set of E of positive measure, that f(e
i
) = 0 for E.
Then f 0.
2.7 Projections from L
p
to H
p
If X is a Banach space and A is a closed subspace of X, then there is not usually any closed subspace
B such that X = AB.
If nevertheless X = AB, we have a unique projection from X onto A with null space B. The aim is
in our case to compute a bounded projection from L
p
to H
p
, if possible. We have so far that H
p
is a
closed subspace of L
p
. The question now arises is, how do we dene a projection onto H
p
? In our rst
chapter we cut the negative Fourier coecients and due to our denition of H
2
and the fact that the
function without the negative part is still in L
2
, this was a projection. It was bounded, since L
2
has a
natural decomposition in two closed subspaces.
The core of the denition of the projection was our preliminary denition of H
2
, since we dened it as
functions in L
2
, where the Fourier coecients vanish for n < 0.
Our situation now is completely dierent. For p [1, ] we are dealing with special classes of analytic
functions in D which are isomorphic to functions on T which belong to a much larger space L
p
on T.
For the last class, there exists a Fourier representation (excluding p = 1) but where is the connection
to H
p
(T)? The answer is contained in the following two theorems proved by the Riesz brothers in [5].
Theorem 2.16 (F. and M. Riesz) 1. For p [1, ] a function f L
p
belongs to H
p
(T) if and
only if the Fourier coecients
f(n) =
T
f(e
i
)e
in
dm()
vanish for n < 0.
2. If f H
p
(T) has Fourier series f(e
i
) =
nN
0
f(n)e
in
, then the Taylor series of the corre-
sponding H
p
(D) is f(z) =
nN
0
f(n)z
n
. Conversely, if f H
p
(D) has Taylor series f(z) =
nN
0
a
n
z
n
, then f H
p
(T) has Fourier series f(e
i
) =
nN
0
a
n
e
in
.
Consequently we have a new denition of H
p
(T).
Denition 2.16.1 The space
H
p
(T) := f L
p
:
f(n) = 0 for n < 0
is called the Hardy space H
p
on T.
From here we have a similarly to the case p = 2 a natural candidate for a closed subspace complementary
to H
p
,
B := f L
p
:
f(n) = 0 for n 0.
The functions in B live on the unit circle and the following lemma gives the connection to anti-analytic
functions in D.
2 PRELIMINARIES 16
Lemma 2.17
B = H
p
0
(T) := f : f H
p
,
f(0) = 0.
Proof "" Let f B. It follows from the denition of B that
f(0) = 0. Then
f(e
i
) =
nN
f(n)e
in
=
nN
a
n
e
in
with a
n
:=
f(n).
Therefore f = g with g H
p
0
.
"" similarly
The extension of H
p
0
(T) to the disk makes use of the Theorem 2.16 (part 2) so we have, along with H
p
the space H
p
0
(D) of anti-analytic functions in D.
Therefore it is quite natural to introduce the following denition.
Denition 2.17.1 If f L
p
has Fourier series
f
nZ
f(n)e
in
then the operator
1.
P
+
(f) :=
nN
0
f(n)e
in
is called the analytic Riesz projection.
2.
P
(f) :=
nN
f(n)e
in
is called the anti-analytic Riesz projection.
Remark We are now using the symbol , because we are including the case p = 1, where Fourier
series must not converge. Therefore we are just associating f with its Fourier series.
Some care must be taken with our previous result in order for it not to be misunderstood. One might
be tempted to say that the image of the analytic Riesz projection P
+
belongs to H
p
for p [1, ].
Unfortunately this is not always the case as the following well known examples show.
Example 1. f() =
nN
sin(n)
n
= arg(1e
i
) is the Fourier series of a bounded function. However
with
nN
sin(n)
n
=
1
2i
nN
z
n
z
n
n
=
1
2i
nZ\{0}
z
n
n
with z = e
i
we obtain with the power series representation of the logarithm
P
+
(f) =
1
2i
nN
z
n
n
=
1
2i
log(1 z)
and the image does not belong to H
n=2
cos(n)
log n
is the Fourier series of an L
1
function ([10] pg. 64) but
P
+
(f) =
1
2
n=2
z
n
log n
does not belong to H
1
due to Hardys inequality.
The last example shows that the image under P
+
of a function in L
1
must not be in L
1
anymore.
2 PRELIMINARIES 17
Therefore the hope that for an f L
p
it holds
f = g +h with g H
p
and h H
p
0
is utterly lost for the case p = 1 and p = .
For p (1, ) we do have the following positive result due to Marcel Riesz which he published in [6].
Theorem 2.18 (M.Riesz) For 1 < p < the operator P
+
maps L
p
boundedly onto H
p
.
Remark Due to this theorem, the operators P
+
and P
L
p
which implies
L
p
= H
p
H
p
0
.
Therefore we have for a given f L
p
, p (1, ) a g H
p
, and an h H
p
0
such that
f = g +h. (2.7.1)
For p = 1 we gave an example such that
L
1
= H
1
H
1
0
but it holds that the functions of the from 2.7.1 are dense in L
1
.
Our example showed that the projection P
+
is unbounded but even more holds due to a theorem of
Donald J. Newman which he published in [7].
Theorem 2.19 (J.D. Newman) There is no closed subspace of L
1
complementary to H
1
. Equiva-
lently there is no bounded projection from L
1
to H
1
.
For p = the situation is (in a certain sense) even worse. As we noted
L
= H
0
but furthermore, the functions of the form 2.7.1 are not even dense.
Theorem 2.20 The closed linear span of the functions in H
.
Remark The theorem is proved in [14], page 151.
It turns out that P
+
L
1
H
p
p < 1 and P
+
L
and it holds
H
p
BMOA H
T
log [f(e
i
)[dm() > . (3.1)
If f(0) = 0, then
log [f(0)[
T
log [f(e
i
)[dm() (Jensens inequality) (3.2)
and more generally, if f(z
0
) = 0
log [f(z
0
)[
T
{
z
0
() log [f(e
i
)[dm().
Remark A fundamental result of H
p
theory is that the condition of (3.1) characterizes the moduli of
H
p
functions [f(t)[ among the positive L
p
functions, i.e.
= [f[, for some f H
p
, f 0
L
p
(T)
log L
1
(T).
Jensens inequality shows the direction "" but the other is contained in Section 3.2 where I compute
a more detailed factorization of H
p
functions.
A useful result which follows from Jensens inequality is the following
Corollary 3.2 Let 0 < p, r . If f(z) H
p
and if the boundary function f(e
i
) L
r
, then
f(z) H
r
. This is often written as
H
p
L
r
= H
r
.
3.2 Factorization for functions in H
p
The next result plays an important role in H
p
theory and especially with respect to our observations
regarding the kernels of the Hankel and Toeplitz operator. It is known as the "factorization theorem"
and is owed to F.Riesz [4] and Vladimir Ivanovich Smirnov [13].
Theorem 3.3 (factorization theorem) 1. Let f H
p
, f 0 with 1 p , then f has a
unique factorization of the form
f = BSF
where B is the Blaschke product
B(z) = z
m
|
k
| =0
k
[
k
[
k
z
1
k
z
,
n
: n N D
whose zeros, repeated according to multiplicity, satisfy
nN
(1 [
n
[) < .
3 NECESSARY TOOLS 19
S is a singular function of the form
S(z) = exp
T
e
i
+z
e
i
z
d()
with positive singular measure (with respect to dm) , and F is an outer function (later dened
as the associated outer function of f)
F(z) = exp
T
e
i
+z
e
i
z
log [f(e
i
)[dm()
T
e
i
+z
e
i
z
log [g(e
i
)[dm()
= BSF (3.1)
where B is a Blaschke product , S is a singular function and g is an L
p
function with log[g[ L
1
,
belongs to the space H
p
. Furthermore [F(e
i
)[ = [g(e
i
)[ almost everywhere.
The whole derivation is quite interesting which is why it is included in this paper. However since it is
not directly related to our topic it is stated in the back in Section 6.1.
An important property of outer and inner functions is the following
Lemma 3.4 Outer and inner functions are multiplicative.
Proof For inner functions this follows directly from the denition.
Let f, g be outer functions (The general denition of outer functions is stated in (3.2), where the only
dierence is that log [f(e
i
) is just a real valued integrable function k on T). Therefore
fg =
1
exp
T
e
i
+z
e
i
z
k
1
()dm()
2
exp
T
e
i
+z
e
i
z
k
2
()dm()
= exp
T
e
i
+z
e
i
z
k()dm()
where :=
2
a complex number of modulus 1 and k := k
1
+ k
2
is a real valued integrable function
on the unit circle.
Remark For f, g the quotient
f
g
is as well an outer function, since k := k
1
k
2
is as well a real valued
integrable function on T and :=
1
2
a complex number of modulus 1. For inner functions this is not
true, since they contain a Blaschke product with zeros.
The converse is also true for outer functions in H
p
.
Lemma 3.5 Let F be an outer function in H
1
, then F =
g
h
, with g, h H
T
e
i
+z
e
i
z
log [F(e
i
)[dm()
(3.2)
3 NECESSARY TOOLS 20
with log [F[ L
1
(T). Let us dene
(e
i
) :=
[F(e
i
)[, [F(e
i
)[ 1
1, [F(e
i
)[ > 1
(e
i
) :=
1, [F(e
i
)[ 1
1
|F(e
i
)|
, [F(e
i
)[ > 1.
Then , are bounded functions in D, hence O
, O
are in H
and
O
(z)
O
(z)
= exp
T
e
i
+z
e
i
z
log
(e
i
)
(e
i
)
dm()
= exp
T
e
i
+z
e
i
z
log [F(e
i
)[dm()
= F.
Another representation of outer functions we are going to use later is the following theorem.
Theorem 3.6 Let F be an outer function with the well known form for z D
F(z) = exp
T
e
i
+z
e
i
z
k()dm()
.
Then F has on the boundary T the form
F(e
i
) = exp [k() +iH(k())] a.e. (3.3)
where H is the harmonic conjugation operator.
Proof Without loss of generality we assume = 1. Then
log F(z) =
T
e
i
+z
e
i
z
k()dm()
=
T
Re
e
i
+z
e
i
z
k()dm() +
T
Im
e
i
+z
e
i
z
k()dm()
=
T
{
z
()k()dm() +
T
O
z
()k()dm().
with {
z
() the Poisson kernel and O
z
() the conjugate Poisson kernel. Then with Theorem 2.14 we
know that the non-tangential limits of the Poisson integral exists a.e. and is in this case equal to k. On
the other hand Theorem 3.1.1. in [11] states, that the same is true for the conjugate Poisson integral
and that the limit is equal to the conjugate function in this case the conjugate function of k. Therefore
log F(e
i
) = k() +iH(k()) almost everywhere.
Another factorization of functions in H
1
which is proved in [14], pg. 52 is the following
Theorem 3.7 Every function in H
1
is the product of two functions in H
2
.
In our observation of the kernels of Hankel and Toeplitz operators, we will often deal with a space
IH
p
:= If : f H
p
T
log
+
f(re
i
)
dm() < .
It is clear that H
p
N for all p > 0, since log
+
[f(z)[
1
p
[f(z)[
p
.
Before we mention the factorization of functions in N and the resulting connection to functions in H
p
,
let us rst introduce an important subclass of N.
If f N, f 0 then it can be shown (stated in [11] Theorem 2.5.3.) that
T
log
f(e
i
)
dm() > .
However the sharper inequality
log [f(z
0
)[
1
2
{
z
0
() log
f(e
i
)
d (3.1)
which was proven for H
p
functions (Jensens inequality Theorem 3.1), can fail for f(z) N which
shows the following example.
Example Let
g(z) = exp
1 +z
1 z
1+z
1z
= {
z
(1). The function g has the harmonic majorant {
z
(1) is therefore in N
(It can be shown, that this is equivalent to the denition of functions in N) and the measure determined
by log [g(z)[ is the unit charge at e
i
= 1. Since
log [g(0)[ = 1 > 0 =
log [g(e
i
)[d
,(3.1) fails for g(z).
This counterexample contains the only thing that can go wrong with (3.1) for a function in N. The
functions in N for which (3.1) holds form a subclass of the Nevanlinna class the so called Smirnov class
which is denoted by N
+
.
In this paper we abandon the original denition of functions in N
+
, because it is not simple to see the
connection to N and H
p
and requires a more detailed examination of the topic. Hence we are going to
work with an equivalent denition of N
+
, which was proven in [11].
From the factorization theorem for H
p
functions, it holds, that any H
p
function can be written in
the form BS
F
f
, where B is the Blaschke product, S
2
as the dierence of two positive measures, one obtains the factorization
for functions in N, which states the following
Theorem 3.9 Let f(z) N, f 0. Then
f(z) = CB(z)
S
1
(z)
S
2
(z)
F(z) [C[ = 1 (3.2)
where B(z) is a Blaschke product, F(z) is an outer function and S
1
(z) and S
2
(z) are singular functions.
Except for the choice of the constant C, [C[ = 1, the factorization (3.2) is unique. Every function of
the form (3.2) is in N.
3 NECESSARY TOOLS 22
Remark The main dierence here, is that we have a quotient of two singular functions and that the
outer function is not necessarily in H
p
.
The factorization of functions in N
+
is similar to the factorization of H
p
functions with one important
dierence.
Theorem 3.10 (and Denition) Let f(z) N, f 0. Then in (3.2) the singular factor S
2
1 if
and only if f(z) N
+
. Therefore we have the unique factorization
f(z) = CB(z)S(z)F(z) [C[ = 1.
From here it is easy to see the following connection.
H
p
N
+
N
for p > 1. It also follows from a theorem in [11], that f(z) H
p
if and only if f(z) N
+
and
f(e
i
) L
p
. This fact can be written as
N
+
L
p
= H
p
.
The example we gave above shows that
N L
p
= H
p
.
The denition of functions in N
+
and the representation in Subsection 3.2 equation (3.1) of functions
in H
p
, 0 p is quite similar. The only dierence is that the outer part of functions in N
+
must
not be integrable which illustrates the following example.
Example The function h(z) = 1 z is outer since Corollary 3.4.8. in [11] which says, that every
function in H
p
with non-negative real values is outer. Therefore the function
f(z) =
1
1 z
is due to (3.4) in N
+
but is not integrable.
An equivalent denition of functions in N and N
+
is the following corollary
Corollary 3.11
N =
f
g
: f, g H
, g zero-free (3.3)
N
+
=
f
g
: f, g H
.
"" Let f be of the form (3.3), then since g, h H
nN
0
a
n
z
n
H
p
the operator
1.
(Sf)(z) := zf(z) =
nN
0
a
n
z
n+1
= a
0
z +a
1
z
2
+a
2
z
3
. . .
is called the forward shift operator.
(Bf)(z) :=
f(z) f(0)
z
= a
1
+a
2
z +a
3
z
2
. . .
is called the backward shift operator.
Remark It is clear from the denition of H
p
, that S and B are linear transformations from H
p
to H
p
A theorem we will use in conjunction with the kernels of Hankel operators is the following
Theorem 3.12 (Beurlings theorem) Let p (0, ) and K be a non-zero S-invariant subspace of
H
p
. Then the following statements are true.
1. K = IH
p
for some inner function I.
2. I
1
and I
2
are inner functions, then I
1
H
p
I
2
H
p
if and only if I
1
/I
2
H
, i.e. I
2
divides I
1
.
Remark 1. Beurlings theorem was originally proven by Beurling for the Hilbert space H
2
and then
generalized by others to H
p
, (0 < p < ).
2. From the Nevanlinna factorization theory, one can show that for any family M of inner functions,
there is an inner function I
M
with the property, that
(a) I
M
divides every I M, specically, I/I
M
H
, and
(b) if J is an inner function which divides every I M, then J divides I
M
.
The inner function I
M
is called the greatest common divisor of M. If K is an S-invariant subspace
of H
p
and I is the greatest common divisor of the inner factors of non-zero functions from K,
then K = IH
p
.
Conversely we will use the B-invariant subspace for our examination of the kernels of Toeplitz operators.
The following theorem explains the structure of this set.
Theorem 3.13 Let p (1, ) then the B-invariant subspace is
K
p
I
:= H
p
IH
p
0
where I ranges over the inner functions.
Remark 1. The operator B is often called S
and let P
: H
p
H
p
0
,
H
f := P
(f)
is called Hankel operator with symbol .
2. The kernels of the Hankel operator are denoted by
Ker
p
H
:= f H
p
: H
f = 0.
Main questions:
1. For which symbols are the kernels non trivial?
2. What are the kernels?
4.1 The symbols of the Hankel operator
In this subsection we are going to compute the set of symbols for which the kernels are non trivial. The
set is precisely the following:
Denition 4.0.2 The space
N
mer
:=
u
v
: u, v H
, v 0
, then H
(f) = 0.
Since this answer is not quite sucient let us assume L
` H
.
Now if f H
p
and H
f = P
and u
2
and
v
2
zero free. Let us dene u := v
1
u
2
and v := v
2
u
1
to conclude
=
g
f
=
v
1
u
2
v
2
u
1
=
u
v
with u, v H
then with N
+
L
= H
this is equivalent to
N
mer
` N
+
.
Recalling the denition of N
+
=
f
g
: f, g H
, g outer we obtain
N
mer
` N
+
:=
u
v
: u, v H
.
4 KERNELS OF HANKEL OPERATORS 25
When v is not outer and in H
= IH
p
, with I the inner function we obtained
in (4.1).
Proof "" Let f = Ig for some g H
p
. Then
H
f = P
(f) = P
(Ig)
4.1
= P
(Fg) = 0
since Fg H
p
.
"" If f Ker
p
H
, then P
(f) = 0. Therefore
f = g H
p
f =
g
yield
g
F
H
p
which proves this direction.
This proof is very straight and needs the results we obtained before. Conversely we could obtain the
same result by using Beurlings theorem (Theorem 3.12), but in this situation just for some inner
function not as specialized as we did it before.
Proof Let M = Ker
p
H
and let P
+
be the Riesz projection from L
p
onto H
p
.
Then the operator T
: H
p
H
p
,
T
f := P
+
(f)
is called Toeplitz operator with symbol .
2. The kernels of the Toeplitz operator are denoted by
Ker
p
T
:= f H
p
: T
f = 0.
Before we start with our examination of the kernels let us prove a useful lemma.
Lemma 5.1 Let L
= Ker
p
T
. (5.1)
Proof Let f H
p
, then (5.1) reduces to
T
f = 0 T
f = 0.
Let us dene F := O
||
and :=
F
, then [[ = 1, since
[[ =
[[
exp[Re(log [[ +iH(log h))]
=
= 1 almost everywhere.
"" If T
.
Since g H
p
N
+
, we have the decomposition
g =
g
1
g
2
with g
1
, g
2
H
g
F
=
[[[f[
[F[
= [f[ L
p
and consequently
g
F
N
+
L
p
= H
p
.
"" If T
f = 0, this is equivalent to
f
F
= g with g H
p
0
, hence
f = Fg. (5.2)
Since f L
p
it holds with (5.2) that Fg L
p
. Conversely by use of Lemma 6.3 is F in H
1
N
+
and due to the multiplicativity of functions in N
+
it holds Fg N
+
and following the same route like
before we proved that f H
p
0
.
The Main questions are still the same:
1. For which symbols are the kernels non trivial?
2. What are the kernels?
5 KERNELS OF TOEPLITZ OPERATORS 27
5.1 The symbols of Toeplitz operators
Let us start with the rst question. We have seen for the Hankel operator, that the space of symbols
for which the kernels are non trivial is easy to nd and quite sucient. For the Toeplitz operator the
situation is completely dierent. The problem is reducible to the following. Let f H
p
, then
T
f
= 0 f = g with g H
p
0
=
g
f
where the diculties are to describe the space of functions
g
f
. One description is the following
Lemma 5.2 If Ker
p
T
f = 0 =
g
f
with g H
p
0
.
Since g, f H
p
, we have the following factorization
g = I
g
F
g
f = I
f
F
f
with I an inner function and F an outer function.
Due to Lemma 5.1, we can say that [[ = 1 without loss of generality. Therefore
1 = [[ =
F
g
I
g
I
f
F
f
F
g
F
f
F
g
= F
f
=: F
with J := I
g
I
f
the the greatest common devisor of I
f
H
p
. We arrive at our result
=
F
JF
. (5.1)
Remark The question now becomes, given an L
= 0 or Ker
2
T
= 0.
Proof Suppose 0 f Ker
2
T
and 0 g Ker
2
T
and
T
f = P
+
zf
f
f
= P
+
zf
= 0.
Remark The question for which symbols are the kernels non trivial is still unsolved and it does not
seem to get solved suciently in the nearby future.
5.2 Description of the kernels of Toeplitz 0perators
Let us start with some preliminary observations. If you choose H
,then f H
p
for a given
f H
p
and hence there is no need of a projection onto H
p
anymore which leads to a trivial kernel.
Conversely it holds the following:
Lemma 5.4 If you choose = I with I an inner function then
Ker
p
T
I
= K
p
I
where K
p
I
is the backward shift invariant subspace.
Proof Let f H
p
. Then
T
I
f = 0 fI H
p
0
f IH
p
0
and hence f H
p
IH
p
0
= K
p
I
.
Due to Theorem 2.20, we know that not only L
= H
0
but the closure of H
0
is not
all of L
nor in H
0
are the functions
which complicate matters. That this is already the case for p = 2 shows the following theorem by Eric
Hayashi which he published in [16].
Theorem 5.5 (Hayashi) Let L
, 0 and Ker
2
T
= G K
2
(5.1)
where G H
2
is such that G
2
is exposed in H
1
and is an inner function with (0) = 0.
Remark We are not giving a denition of exposed points, since we are not dealing with them later.
Furthermore, they make up a chapter on their own thus are not of interest. On top of that there does
not seem to be a comfortable characterization of them which is the reason Konstantin Dyakonov (my
supervisor) felt it would be wise to devise a set without them.
Before we concentrate on the work of Konstantin Dyakonov, it is useful to state a few denitions and
useful lemmas.
Denition 5.5.1 1. Given an integer n 1, we denote P(n) the set of (analytic) polynomials of
degree < n.
2. (H
)
1
:= f H
:
1
f
H
.
3. Elements of the form B B (H
)
1
, where B is the set of all Blaschke products are called
triple.
Lemma 5.6 Let F H
.
2. Let f H
p
0
N
+
0
and the factorization functions in H
p
yields
f
F
=
O
|f|
I
f
F
.
O
|f|
F
is with Lemma 3.4 outer and therefore
f
F
N
+
0
follows from the denition of N
+
0
.
Corollary 5.7 Let f L
p
, F H
. Then f H
p
0
f F
1
H
p
0
.
Proof Follows by the previous lemma with f N
+
0
L
p
= H
p
0
.
A more detailed description of the kernels is the following lemma.
Lemma 5.8 Let T
f H
p
0
.
Proof The case 1 < p < is obviously true since L
p
= H
p
H
p
0
.
If p = we have f H
H
2
and it holds
T
f = 0 f H
2
0
and since f L
this is equivalent to
f L
H
2
0
= H
0
.
If p = 1 then f L
1
and since T
f = P
+
(f) = 0, f has to be in H
1
0
due to the denition (recall:
H
1
0
= g L
1
: P
+
(g) = 0).
Remark The Subsection 2.7 deals with the question; what happens if you take a function in L
1
and
you truncate o the negative parts? It is possible that the new function is not even in L
1
anymore.
A necessary description of the symbols, for which the kernels are G H
p
multiplied by a special case
of a backward shift invariant subspace P(n) = H
p
z
n
H
p
0
= K
p
z
n.
Proposition 5.9 Let 1 p and n N. Suppose that G H
p
,
1
G
H
and set := z
n G
G
. Then
Ker
p
T
= G P(n).
Proof Fix f H
p
. In order that f Ker
p
T
u
, Lemma 5.8 says it is necessary and sucient that
fu H
p
0
. (5.2)
Recalling the denition of and setting h :=
f
G
, we rewrite (5.2) as
fz
n
G
G
= Ghz
n
H
p
0
. (5.3)
5 KERNELS OF TOEPLITZ OPERATORS 30
Since we have
Ghz
n
= [fz
n
[ = [f[ [z
n
[ = [f[ L
p
since [z[ = 1 we are able to use Corollary 5.7 and
thus (5.3) is equivalent to
hz
n
H
p
. (5.4)
Finally noting that h H
p
and therefore h H
p
z
n
H
p
0
, we see when keeping the power series
representation in mind that (5.4) reduces to the inclusion h P(n) , i.e.
f G P(n). (5.5)
The resulting equivalence relation (5.2) (5.5) proves the proposition.
A naive attempt to extend (5.1) to the L
p
-scale might be to replace it with something like
Ker
p
T
= G K
p
(5.6)
where is an inner and G is a suitable weight function. Unfortunately this plan fails due to the following
example.
Example Let us take the function h(z) := (1z)
p
0
with p
0
(1, ) and =
zh
h
. Then L
and
if z = e
it
then for t being small z is reducible to the rst two parts of its power series representation
and therefore e
it
1 +it and hence
[dz[
[1 z[
p
p
0
dt
[t[
p
p
0
.
Now it follows from basic calculus that h H
p
for 0 < p < p
0
. It holds h Ker
p
T
for 1 p p
0
since P
+
(h) = P
+
(
z
hh
h
) = P
+
( zh) = 0.
Furthermore since Lemma 5.9 the subspace Ker
p
T
Ker
1
T
and log[f[ L
1
, then f = gh
for some g, h H
.
When restricted to unimodular functions, Bourgains result amounts to the following:
Theorem 5.11 Given a function L
)
1
.
If I
g
and I
h
are constant this would be the proof since the constant inner function 1 is a Blaschke
product and G multiplied by a constant is still in H
)
1
and are outer functions,
since the group structure of outer functions yield to g := EFG an outer function. It holds
1 aI
g
1 +
aI
g
E H
[aI
g
[ [a[ < 1
1
E
H
.
Following the same route for F yields E, F (H
)
1
and by using Corollary 6.13 we see that E, F
are outer functions.
Remark The conjugate part of the factorization does not have to be the numerator. By looking at
the proof closely you can change the denitions of E, F and G to
1
E
,
1
F
and
1
G
and one obtains an
equivalent factorization just with the conjugate part as a divisor.
The following lemma is the reason why we can use Burgains factorization theorem in our next theorem
Lemma 5.12 Let L
0. If log [[ L
1
then Ker
p
T
= 0.
Proof Let us assume Ker
p
T
f H
p
0
.
Therefore it exists g H
p
0
with
f = g. (5.7)
Because L
=
g
b
(K
p
B
bH
p
) p [1, ]. (5.8)
2. Conversely, given a triple (B, b, g) one can nd a L
. (5.9)
Lemma 5.8 says that it is necessary and sucient that f H
p
0
. This can be rewritten as
f
1
F
H
p
0
. (5.10)
With [f[ = [f[ L
p
and the result of Corollary 5.7, (5.10) reduces to
f H
p
0
. (5.11)
Next we invoke the factorization to restate (5.11) as
fb
g
BH
p
0
(5.12)
recalling that g (H
)
1
. Since
fb
g
belongs to bH
p
and hence to H
p
(Lemma 3.8), we may as
well write (5.12) in the form
fb
g
BH
p
0
H
p
bH
p
= K
p
B
bH
p
. (5.13)
Finally (5.13) amounts to the inclusion
f
g
b
(K
p
B
bH
p
) (5.14)
and the resulting equivalence relation (5.9) (5.14) proves (5.8).
2. Given triple (B, b, g) dene =
bg
Bg
and set = , F = O
||
= 1. This done, the above reasoning
shows that conditions (5.9) (5.14) are all equivalent, whenever f H
p
(1 p ). A
juxtaposition of the endpoint statements (5.9) and (5.14) yields to the equality, just as before.
5 KERNELS OF TOEPLITZ OPERATORS 33
Five years later Stephan Ramon Garcia extended the work of Dyakonov (which was entirely dierent
as well) but just for 0 < p < . He dened an easy malleable space which is equal to the kernels of
the Toeplitz operators. We start with some necessary denitions:
Denition 5.13.1 1. A function f belonging to the Smirnov class N
+
is called a real Smirnov
function, f R
+
if its boundary function is real valued a.e. on T.
2. Let C
+
denote the subspace R
+
+iR
+
of N
+
.
3. We dene A
p
to be C
+
H
p
with an outer function.
4. Let f = h A
p
f H
p
and since h C
+
it follows that
f A
p
. We refer to f and
f as a conjugate pair in A
p
.
A set like R
+
for functions in H
p
would not make any sense due to the following lemma.
Lemma 5.14 If f H
1
and f(e
i
) R T, then f constant.
Proof With the Poisson kernel representation
f(z) =
T
{
z
()f(e
i
)dm()
and the fact that {
z
() R it follows that f(z) R z D which implies that f has to be constant.
Before we state the results of Stephan Ramon Garcia, let us obtain some preliminary results.
Lemma 5.15 If f and
f are conjugate functions in A
p
, then
f =
f
.
Proof If f A
p
, then h C
+
such that f = h. Consequently
f = h =
h
=
f
.
Proposition 5.16 Two functions f and g in H
p
satisfy g =
f
and
combining it with
I
f
I
g
= I
f
I
g
I
f
+I
g
I
f
+I
g
=
I
f
+I
g
1
I
f
+
1
I
g
=
I
f
+I
g
I
f
+I
g
=
1
2
(I
f
+I
g
)
1
2
(I
f
+I
g
)
=
1
2i
(I
f
I
g
)
1
2i
(I
f
I
g
)
(5.15)
on T yields
1
2
(I
f
+I
g
)F
=
1
2
(I
f
+I
g
)F
and
1
2i
(I
f
I
g
)F
=
1
2i
(I
f
I
g
)F
. (5.16)
This means, that the functions
1
2
(I
f
+I
g
)F
and
1
2i
(I
f
I
g
)F
and are
a conjugate pair.
"" Since f, g are conjugate pairs in A
p
, g =
f
.
5 KERNELS OF TOEPLITZ OPERATORS 34
Denition 5.16.1 Let f = I
f
F and
f = I
f
F are a conjugate pair in A
p
f
F =
F
. Therefore it exists a
conjugate function
F of F with
F =
F
which leads to
I
f
I
f
F =
F. (5.17)
Moreover the inner function J
F
:= I
f
I
f
=
F
F
depends only upon F and . We call this inner function
J
F
the associated inner function of F in A
p
.
Remark The functions f = I
f
F in A
p
be a Toeplitz operator on H
p
for some p (1, ). If Ker
p
T
is
non trivial, then there exists an outer function in H
p
such that Ker
p
T
= Ker
p
T
z
.
Proposition 5.18 If p (1, ) and H
p
is outer, then Ker
p
T
z
= A
p
.
Proof "" Since f H
p
, f = I
f
F with I
f
inner and F outer. Let f belongs to Ker
p
T
z
which is
equivalent to I
f
F
z
H
p
0
. Therefore g H
p
such that l :=
I
f
Fz
= zIF
for some inner function I. Consequently
IF =
I
f
F
.
"" If f belongs to A
p
f) = P
+
(zh) it follows that
T
z
=
b
B
g
g
. (5.18)
Our nal proposition relates A
p
) =
b
B
g
g
where (B, b, g) is a triple, then the following subspaces of H
p
are identical:
1. A
p
.
2. Ker
p
T
z
.
5 KERNELS OF TOEPLITZ OPERATORS 35
3.
g
b
(bH
p
BH
p
).
4. f g(H
p
BH
p
) : bI
f
[J
F/g
.
Remark The notation J
F/g
in (4) refers to the associated inner function for the outer function F/g
in H
p
bH
p
.
Proof The equality of (1) and (2) follows from Proposition 5.18.
"(1)(4)" If f = I
f
F belongs to A
p
, then
f = I
f
F =
I
f
F
= I
f
F
B
b
g
g
.
Hence
bI
f
I
f
F/g = (F/g)B. (5.19)
Since bI
f
I
f
(F/g) H
p
, it follows with (5.19) that bI
f
I
f
(F/g) BH
p
. Therefore F/g belongs to
H
p
BH
p
and since this is a B-invariant subspace F/g has the associated inner function J
F/g
:= bI
f
I
f
.
"(1)(4)" Suppose f = I
f
F belongs to g(H
p
BH
p
). Then (F/g)B H
p
and consequently (F/g)B
H
p
. Since
(F/g)b H
p
= [F/g[, O
[(F/g)B[
= F, the canonical function of Fb is
FB = FI
for some inner function I. The function I is the associated inner function to (F/g) and therefore
I = J
F/g
. With bI
f
[J
F/g
we obtain J
F/g
= bI
f
J for some inner function J. If we merge everything
together we obtain
J
F/g
= bI
f
JF/g = (F/g)B.
This is equivalent to
F := I
f
JF = F
B
b
g
g
=
F
.
Hence from the denition of the associated inner function, f belongs to A
p
1
2
T
e
i
+z
e
i
z
k()d
Re
T
e
i
+z
e
i
z
k()dm()
= exp
T
{
z
()k()dm()
.
Then by Theorem 2.14 we obtain
log
F(e
i
)
T
[F(z)[ d =
T
exp
T
{
z
()k()dm()
d
and let d() := {
z
()dm(), then d() is a probability measure. Since k is real-valued and the
exponential function is convex we can apply Jensens inequality for convex functions (Theorem 6.12)
to obtain
T
exp
T
k()d()
T
e
k()
d()d.
Now we can use Fubini-Tonelli to switch the integrals and since
T
{
z
()
d
2
= 1, we arrive at the nal
estimate
T
[F(z)[ d
T
e
k()
d < .
One can state the connection between an outer function and a function in H
1
in the following way.
Lemma 6.3 Let F be outer, f H
1
and k() = log [f(e
i
)[, then F H
1
and [F[ = [f[ a.e. on T.
6 APPENDIX 37
Proof By using Lemma 6.1 we have log [F(e
i
)[ = k() = log [f(e
i
)[ almost everywhere. The rest
follows from Lemma 6.2.
We now provide another useful denition we employ.
Denition 6.3.1 Let f be in H
1
. The function
F(z) = exp
T
e
i
+z
e
i
z
log [f(e
i
)[dm()
(6.1)
where is a complex number of modulus 1 is called the associated outer function. This is well dened
since log [f(e
i
)[ is real-valued on T and with Jensens inequality (Theorem 3.1) Lebesgue integrable.
A deeper result then Lemma 6.3 can be obtained.
Theorem 6.4 Let F H
1
and F 0. The following are equivalent:
1. F is an outer function.
2. If f is any function in H
1
such that [F[ = [f[ a.e. on T, then [F(z)[ [f(z)[ z D.
Proof "" We use Jensens inequality (Theorem 3.1) and get
log [f(z)[
T
{
z
(e
i
) log
f(e
i
)
T
e
i
+z
e
i
z
log
F(e
i
)
dm()
. (6.3)
Then we use Lemma 6.3 and nd that [F[ = [G[ a.e. on T. Now we can use step (1)(2) and obtain
[F(z)[ [G(z)[. On the other hand one easily checks that [G(z)[ [F(z)[.
Thus
F
G
is analytic and always absolute value 1. So F = G, where [[ = 1, which proves that F is an
outer function.
Another useful characterization is the following theorem.
Theorem 6.5 Let F be a non zero function in H
1
. The following are equivalent:
1. F is an outer function.
2. log [F(0)[ =
T
log
F(e
i
)
dm().
Proof "" This is obviously true for all outer functions since {
0
() = 1.
"" We dene G as in the theorem above and following the same route to obtain F(z) G(z) z D.
Thus
F(z)
G(z)
1 z D but since
F(0)
G(0)
=
exp
T
log
F(e
i
)
dm()
exp
log [F(e
i
)[ dm()
= 1
the maximum principle yields
F(z)
G(z)
= z D with [[ = 1.
Let us now have a look at the function g(z) :=
f(z)
F(z)
z D with f H
1
and F the associated outer
function. Since [F[ = [f[ a.e. on T it follows that g is a bounded function with
[g(z)[ =
f(z)
F(z)
1 z D.
It will be convenient for us to make the following denition.
6 APPENDIX 38
Denition 6.5.1 (inner function) An inner function is an analytic function g on D such that
[g(z)[ 1 z D and [g[ = 1 a.e. on T.
With this result we are able to do a preliminary factorization:
Theorem 6.6 Let f be a non zero function in H
1
. Then f can be written in the form f = gF, where
g is an inner function and F is the associated outer function in H
1
. This factorization is unique up to
a constant of modulus 1.
Proof By dening
F(z) = exp
T
e
i
+z
e
i
z
log
f(e
i
)
dm()
kN
(1 [
k
[) .
Proof Without loss of generality, we assume [f[ 1. Since for a nite number of zeros the case is
trivial, we assume that f has a countable number of zeros. Let B
n
(z) be the nite product:
B
n
(z) =
n
k=1
k
[
k
[
k
z
1
k
z
(6.4)
Now B
n
(z) is a rational complex valued function and since the poles are lying outside of D it is analytic
inside. It holds for all [0, 2]
k
[
k
[
k
e
i
1
k
e
i
k
e
i
1
k
e
i
k
e
i
e
i
k
e
i
e
i
= 1
and therefore
B
n
(e
i
)
= 1 n N. Furthermore
f
B
n
is a bounded analytic function in D. Since n N
f(e
i
)
[B
n
(e
i
)[
=
f(e
i
)
1 almost everywhere.
we have [f(z)[ [B
n
(z)[ on D. In particular
0 < [f(0)[ [B(0)[ =
n
k=1
[
k
[ .
Since [
k
[ < 1 k N and since each of the partial products
n
k=1
[
k
[ is not less then [f(0)[, the
innite product converges.
6 APPENDIX 39
Remark Let us assume that the
k
are not nite then
1. Since f 0, the identity theorem for holomorphic functions states that the accumulation point
cannot lie in D and lies therefore either on T or outside of D. Consequently the modulus of the
k
increases when k goes to innity.
2. It is impossible to have an with f() = 0 and innity multiplicity since with [[ < 1 it would
lead to
kN
[[
k
= 0.
3. The zeros of f are not allowed to lie too dense in D. This can be illustrated with the following
example:
Let
k
:= 1
1
k
k N. Since the convergency of
kN
[
k
[ is equivalent to
kN
(1[
k
[) <
the
k
cannot be the zeros of an bounded analytic function, since
kN
(1 [
k
[) =
kN
(1
1 +
1
k
) =
kN
(
1
k
) is the harmonic series.
4. If we drop the boundedness of the analytic function Remark (3) becomes false since the function
f(z) = sin(
1z
) vanishes at
k
k N and is analytic in D.
Theorem 6.8 (and Denition of the Blaschke product) Let
k
be a sequence of points in D
such that
kN
(1 [
k
[) < . Let m be the number of
k
equal to zero. Then the Blaschke product
B(z) = z
m
|
k
| =0
k
[
k
[
k
z
1
k
z
converges on D. The function B(z) is in H
k
z
1
k
z
. (6.5)
With Theorem 15.6. in [8] the product
kN
b
k
converges on D to an analytic function having the
k
for zeros if and only if
kN
[1 b
k
(z)[ converges uniformly on each compact subset of D. By
calculation it follows
[1 b
k
(z)[ =
[
k
[(1
k
z)
k
(
k
z)
[
k
[(1
k
z)
[
k
[ [
k
[
k
z [
k
[
2
k
+z
k
[
k
[(1
k
z)
[
k
[ +z
k
[
k
[(1
k
z)
(1 [
k
[)
[
k
[ +[z[[
k
[
[
k
[[1
k
z[
(1 [
k
[)
=
1 +[z[
1 [
k
[[z[
(1 [
k
[)
1 +[z[
1 [z[
(1 [
k
[)
and the convergence follows from
kN
[1 b
k
(z)[ < . Since [b
k
(z)[ 1 z D, it is clear that
B(z) H
and [B(z)[ 1. The bounded harmonic function B(z) has non tangential limits [B(e
i
)[ 1
almost everywhere. To see that [B(e
i
)[ = 1 a.e., set B
n
(z) :=
n
k=1
b
k
(z). Then
B
B
n
is another Blaschke
product and
B(0)
B
n
(0)
B(e
i
)
[B
n
(e
i
)[
dm() =
B(e
i
)
dm().
6 APPENDIX 40
Letting n now yields
T
B(e
i
)
dm() = 1
so that
B(e
i
)
= 1 almost everywhere.
Remark We are now able to factorize every inner function into a Blaschke product B and a function
g without zeros since g =
f
B
is analytic and bounded in D. The factorization f = gB is unique since a
Blaschke product is uniquely determined by its zeros.
Theorem 6.9 Let g be an inner function without zeros and suppose that g(0) is positive, then there is
a unique singular positive measure on T such that
g(z) = exp
T
e
i
+z
e
i
z
d()
. (6.6)
Proof Since g is analytic in the disc and has no zeros, it follows from a well-known theorem that
g = e
h
, where h is an analytic function in the disc. Since g is bounded by 1, the real part of h must
be non negative in D. Let h = u + iv so that u 0. The non-negative harmonic function is uniquely
expressible in the form
u(z) =
T
{
z
()d()
where is a positive measure on T and z = re
i
. Since g(0) > 0, we have e
uiv
and consequently
v(0) = 0 or with e
2ni
= 1 n N, v(0) = 2ni. Without loss of generality, v(0) = 0 and thus
h(z) =
T
e
i
+z
e
i
z
d()
where is a positive measure on T. This is due to Herglotzs theorem (Theorem 6.14) since h takes
values in the right half plane. Now we have as an inner function [g[ = 1 a.e. on T. Since [g[ = e
u
,
this just means that the non-tangential limits of u must vanish almost everywhere on T. But by using
Fatous theorem these non-tangential limits are equal to 2
d
d
= 2f with f the Radon-Nikodym
derivative. By using Lebesgues decomposition theorem (Theorem 6.16) we obtain
d = fd +d
s
where d
s
is a singular measure with respect to d. Since the Radon-Nikodym derivative is zero a.e.
d is singular and this completes the proof.
Denition 6.9.1 (singular function) We dene an inner function without zeros which is positive at
the origin a singular function.
Theorem 6.10 Let f 0 be an H
1
function on D. Then f is uniquely expressible in the form
f = BSF, where B is a Blaschke product, S is a singular function and F is the associated outer
function (in H
1
).
Proof We know from our previous results (Theorem 6.6) that f = gF, where g is an inner function
and F is the associated outer function and that this factorization is unique up to a constant multiple
of modulus 1. If B is the Blaschke product formed from the zeros of g (i.e. the zeros of f), then
g = BS, where S is an inner function without zeros. By multiplying g by a constant of modulus 1, we
can arrange that S(0) > 0 ,i.e. that S is a singular function. We then absorb that constant into the
associated outer function F and we are done!
6 APPENDIX 41
6.2 Theorems used in this paper
Theorem 6.11 (Frostmans theorem) Let f(z) be a non-constant inner function on D. Then for
all , [[ < 1, except possibly for a set of capacity zero, the function
f
(z) =
f(z)
1 f(z)
(6.1)
is a Blaschke product.
Remark This is the key part of the proof that the set of Blaschke products is uniformly dense in the
set of inner functions.
Theorem 6.12 (Jensens inequality for convex functions) Let (, ) be a measure space such
that is a probability measure ,i.e. () = 1. Let L
1
() be a real-valued function and let (t) be
a convex function on R. Then
(
d)
()d. (6.2)
Corollary 6.13 If f(z) H
p
and if for some r > 0,
1
f(z)
H
r
, then f(z) is an outer function.
Theorem 6.14 (Herglotzs theorem) Every analytic function in the unit disc with values in the
right half-plane such that f(0) > 0 has the form
f(z) =
T
e
i
+z
e
i
z
d() (6.3)
where is a nite positive measure on T.
Theorem 6.15 (Radon-Nikodym theorem) Given a measurable space (, ), if a -nite on
(, ) is absolutely continuous with respect to a -nite measure on (, ), then there is a measurable
function f on and taking values in [0, ) such that
(A) =
A
fd (6.4)
for any measurable set A.
Remark The function f satisfying the above equality is uniquely dened up to -null set. The function
f is commonly written as
d
d
and is called the Radon-Nikodym derivative.
Theorem 6.16 (Lebesgues decomposition theorem) Given and , two -nite measures on a
measurable space (, ), there exist two -nite signed measures
a
and
s
such that
1. =
a
+
s
.
2.
a
<.
3.
s
.
6.3 Acknowledgments
I thank Prof. Konstantin Dyakonov for accepting my request of beeing my supervisor for this bachelor
thesis since it is not part of the Erasmus program, for his numerous private lectures which gave me
a more detailed understanding of the topic and for his support especially at the end of the project.
Also, it is my pleasure to thank Prof. A.I. Bobenko who gave me the possibility to write this paper in
Barcelona during my Erasmus year.
REFERENCES 42
References
[1] J.R. PARTINGTON An Introduction to Hankel Operators - London Mathematical Society Student
Texts
[2] R.A. MARTINEZ-AVENADO and P.ROSENTHAL An Introduction to Operators on the Hard-
Hilbert Space - Graduate Texts in Mathematics
[3] G.H. HARDY On the mean value of the modulus of an analytic function - Proceedings of the
London mathematical society series 2 14:269-277
[4] F. RIESZ Ueber die Randwerte einer analytischen Funktion - Math.Z. 18:87-95
[5] F. and M. RIESZ Ueber die Randwerte einer analytischen Funktion - Quatrime Congrs des
Math. Scand. Stockholm (1916) 27-44
[6] M. RIESZ Les fonctions conjugeres et les sries de Fourier - C.R.Acad. Sci. Paris, Sr. A-B 178
(1924), 1464-1467
[7] J.D. NEWMAN The non.existence of projections from L
1
to H
1
- Proc. Amer. Math.Soc. 12
(1961), 98-99
[8] W. RUDIN Real and Complex Analysis - International Series in Pure and Applied Mathematics
[9] P. FATOU Sries trigonomtriques et seris de Taylor - Acta Math. 30 (1906), no.1, 335-400
[10] P.L. DUREN Theory of H
p
spaces - Pure and Applied Math., Vol. 38, Academic Press, New
York-London, 1970
[11] JOHN B. GARNETT, Bounded Analytic Functions - Graduate Texts in Mathematics
[12] JOSEPH A. CIMA and WILLIAM T. ROSS, The Backward Shift on the Hardy Space - Mathe-
matical Surveys and Monographs
[13] V.I. SMIRNOV, Sur les valeurs limites de fonctions rgulires linterrieur dun cerle - Journal
de la Socit Phys.-Math. de Lningrade, 2(1929), no. 2, 22-37
[14] K. HOFFMAN, Banach Spaces of Analytic Functions - Dover Publications, Inc., New York, 1988.
[15] E.HAYASHI, The solution sets of extremal problems in H
1
- Proc. Amer. Math. Soc. 93(1985),
690-696.
[16] E.HAYASHI, Classication of the nearly invariant Subspaces of the Backward Shift - Proc. Amer.
Math. Soc. 110(1990), 441-448.
[17] K. DYAKONOV, Kernels of Toeplitz Operators via Bourgains factorization theorem - J. Funct.
Anal. 170 (2000), 93-106
[18] S.R. GARCIA, Conjugation, the backward shift, and Toeplitz kernels - J. Operator Theory 54
(2005), no. 2, 239-250.