0% found this document useful (0 votes)
65 views107 pages

Uncertainty Principles in Framed Hilbert Spaces

This dissertation by Haodong Li from Clemson University examines uncertainty principles in framed Hilbert spaces. Chapter 1 introduces key concepts like multiplication and differentiation operators, the classical uncertainty principle, and frames. Chapter 2 develops the theory of continuous frames and Toeplitz operators in framed Hilbert spaces. Chapter 3 defines d-approximate (weak) interpolation sets in framed Hilbert spaces and proves a necessary density condition for such sets. Chapter 4 studies sampling problems in multiband-limited function spaces. Chapter 5 generalizes the Balian-Low theorem to non-commuting operator pairs in a weak sense and examines joint eigenbasis properties. The dissertation establishes fundamental results about uncertainty principles in the general framework of framed Hilbert spaces.

Uploaded by

Rosana Radović
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
65 views107 pages

Uncertainty Principles in Framed Hilbert Spaces

This dissertation by Haodong Li from Clemson University examines uncertainty principles in framed Hilbert spaces. Chapter 1 introduces key concepts like multiplication and differentiation operators, the classical uncertainty principle, and frames. Chapter 2 develops the theory of continuous frames and Toeplitz operators in framed Hilbert spaces. Chapter 3 defines d-approximate (weak) interpolation sets in framed Hilbert spaces and proves a necessary density condition for such sets. Chapter 4 studies sampling problems in multiband-limited function spaces. Chapter 5 generalizes the Balian-Low theorem to non-commuting operator pairs in a weak sense and examines joint eigenbasis properties. The dissertation establishes fundamental results about uncertainty principles in the general framework of framed Hilbert spaces.

Uploaded by

Rosana Radović
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 107

Clemson University

TigerPrints

All Dissertations Dissertations

December 2020

Uncertainty Principles in Framed Hilbert Spaces


Haodong Li
Clemson University, [email protected]

Follow this and additional works at: https://tigerprints.clemson.edu/all_dissertations

Recommended Citation
Li, Haodong, "Uncertainty Principles in Framed Hilbert Spaces" (2020). All Dissertations. 2748.
https://tigerprints.clemson.edu/all_dissertations/2748

This Dissertation is brought to you for free and open access by the Dissertations at TigerPrints. It has been
accepted for inclusion in All Dissertations by an authorized administrator of TigerPrints. For more information,
please contact [email protected].
Uncertainty Principles in Framed Hilbert Spaces

A Dissertation
Presented to
the Graduate School of
Clemson University

In Partial Fulfillment
of the Requirements for the Degree
Doctor of Philosophy
Mathematical Science

by
Haodong Li
December 2020

Accepted by:
Mishko Mitkovski, Committee Chair
Shitao Liu
Martin Schmoll
Jeong-Rock Yoon
Introduction

The uncertainty principle is one of the most fundamental concepts in harmonic analysis. It

has many facets and appears in many different (non-equivalent) forms. However, a common theme

of all uncertainty principles can be vaguely summarized into two points of view. From Fourier

analysis perspective: In 1925, Wiener first formulate the uncertainty principle [19] in his Göttingen

lecture. It says a function and its Fourier transform cannot both be sharply localized. From operator

theoretical perspective: In 1927, two years after the Wiener’s lecture, Heisenberg formulated his

famous uncertainty principle [7], saying that the position and the momentum of a quantum particle

cannot be measure simultaneously. Mathematically, this can be expressed as the multiplication and

differentiation operators X and D satisfy the canonical commutation relation:

1
[D, X] := DX − XD = I.
2πi

Which shows a strong non-commutativity of X and D. These two points of view are related with

each other through the fact that the Fourier transform F conjugates the operators X and D. Namely,

X = FDF ∗ .

The thesis mainly consists of three projects whose common theme is the uncertainty prin-

ciple. Two of them concern interpolation and sampling problems, while the third one is about the

so-called generalized Balian-Low theorem. We now briefly describe each of them, and give an outline

of how the thesis is organized.

The main results of the thesis are included in the last three chapters. The main purpose

of Chapter 1 is to introduce the well-known concepts that will be used throughout the thesis, and

set up the notations. Besides this, we state the classical uncertainty principle and the classical

Balian-Low theorem that will be generalized in the later chapters. We also include proofs of these

ii
results which are based on the similar ideas with our proofs of more general results. The reason for

including them is to illustrate those ideas in the simplest possible setting.

In Chapter 2, we introduce the concept of framed Hilbert spaces as a general setting for

studying problems of uncertainty principle type. These spaces could be viewed as a special type of

reproducing kernel Hilbert spaces which include many function spaces that play an important role

in harmonic analysis. In the chapter, we prove some basic results about framed Hilbert spaces that

will be used in the rest of the thesis. Many of those are new results in this level of generality.

Sampling and interpolation problems could be viewed as a manifestation of the uncertainty

principle in the following way. Take the classical Paley-Wiener space as an example. This space

consists of functions (band-limited functions) in L2 (R) whose Fourier transform is localized in a

fixed interval. As a consequence of the classical uncertainty principle, these functions cannot be

localized on any finite intervals. However, we have more than that. It turns out that the oscillation

of the functions get controlled by their band limit. Therefore, it is difficult to solve interpolation

problems for such functions. More precisely, to be able to solve interpolation problems, the density

of interpolating points must be sparse enough relative to the band limit. However, at the same time,

it is easy to solve sampling problems. In other words, such functions could be completely determined

on the set (sampling set) which is sufficiently dense relative to the band limit.

In Chapter 3, we define a very general interpolation set called d-approximate (weak) in-

terpolation set in framed Hilbert spaces. This type of sets were relatively recently introduced by

Olevskii and Ulanovskii in the classical Paley-Wiener space. And we will prove a necessary density

condition for those sets similar to the one for usual interpolation sets.

Like Chapter 3, sampling problems could be also studied in the general framework of framed

Hilbert spaces. However, we will concentrate on a more precise problem of estimating a sampling

constant in the space of multiband-limited functions. A sharp estimate of this constant in classi-

cal Paley-Wiener spaces is given by Kovrijkine, who also provides a quantitative estimate for this

constant in multiband cases. In both of these cases, the dependence of this sampling constant on

the length of the (multi)band is exponential. In Chapter 4, we will impose additional (suboptimal)

conditions on the sampling set which allow us to obtain a much-improved sampling constant which

depends linearly on the length of the multiband.

The classical uncertainty principle gives a bound on how close a function could become a

joint eigenvector of X and D. This classical version of uncertainty principles also gives us the optimal

iii
approximate joint eigenvectors, which are the time-frequency shifts of the Gaussian. However, it

turns out that the time-frequency shifts of the Gaussian, unfortunately, cannot form an orthonormal

basis (even a Riesz basis). Moreover, if we use an integer lattice of time-frequency shifts of some

generating function to form an orthonormal basis, it turns out this generating function must have

a very poor time-frequency localization. This result of uncertainty principle type is the so-called

Balian-Low theorem.

In Chapter 5, we will give a very general version of Balian-Low theorem, which only requires

the operator pair to satisfy a weak non-commutativity property and does not necessarily require the

orthonormal (or Riesz) basis to be the time-frequency shifts of a single function. Furthermore, we

examine the relationship between diagonalization and possibility of joint orthonormal (or Riesz)

eigenbasis of the operator pair both in a very weak sense.

iv
Table of Contents

Title Page . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . i

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ii

1 Preliminaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Multiplication and Differentiation Operators . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Classical Uncertainty Principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 Classical Balian–Low Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.4 Frames and Riesz Sequences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

2 Continuous Frames and Toeplitz Operators in Framed Hilbert Spaces . . . . . . 12


2.1 Continuous Frames and Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.2 Framed Hilbert Spaces and Examples . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.3 Toeplitz Operators on Framed Hilbert Spaces . . . . . . . . . . . . . . . . . . . . . . 20
2.4 Additional Conditions on Framed Hilbert Spaces . . . . . . . . . . . . . . . . . . . . 27
2.5 Boundedness, Compactness and Trace Class Membership of Toeplitz Operators . . . 34
2.6 Beurling Densities of Sampling and Interpolation Sets . . . . . . . . . . . . . . . . . 36

3 Approximate Interpolation in Framed Hilbert Spaces . . . . . . . . . . . . . . . . 40


3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.2 d-Approximate Interpolation in Framed Hilbert Spaces . . . . . . . . . . . . . . . . . 42
3.3 d-Approximate Weak Interpolation in Classical Bargmann-Fock Spaces . . . . . . . . 50

4 Frame Bound Estimates for Continuous Frames of Exponentials . . . . . . . . . 58


4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
4.2 Lower Frame Bound Estimates for Continuous Frames of Exponentials . . . . . . . . 59

5 Uncertainty Principles in Framed Hilbert Spaces . . . . . . . . . . . . . . . . . . . 75


5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
5.2 Generalized Balian-Low Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
5.3 Balian-Low Theorem in Framed Hilbert Spaces . . . . . . . . . . . . . . . . . . . . . 84

Acknowledgment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .100

Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .101

v
Chapter 1

Preliminaries

1.1 Multiplication and Differentiation Operators

Arguably, the two most important operators in analysis are the multiplication and differen-

tiation operators:

The multiplication X : D(X) → L2 (R) is defined by

Xf (x) := xf (x), ∀x ∈ R.

And the differentiation D : D(D) → L2 (R) is defined by

1 0
Df (x) := f (x), ∀x ∈ R.
2πi

Where the domains of X and D are defined in the usual way:

D(X) := {f ∈ L2 (R) : Xf ∈ L2 (R)}; D(D) := {f ∈ L2 (R) : Df ∈ L2 (R)}.

By definitions, we could show X and D are both self-adjoint operators, i.e.,

X ∗ = X; D∗ = D.

Using X and D, we could also define the following two groups of operators:

1
The group of translations Ta : L2 (R) → L2 (R) is given by

Ta := e−2πiaD , ∀a ∈ R.

And the group of modulations Mb : L2 (R) → L2 (R) is given by

Mb := e2πibX , ∀b ∈ R.

By Taylor expansions, we have

Ta f (x) = f (x − a); Mb f (x) = e2πibx f (x).

That’s why we call Ta translation and Mb modulation.

Moreover, we could show a → Ta and b → Mb are two unitary representations from (R, +)

to U(H), which means Ta and Mb are the unitary operators such that

Ta Tb = Ta+b , Ta∗ = T−a ;

Ma Mb = Ma+b , Mb∗ = M−a .

Since Mb is generated by X and Ta is generated by D, we have two pairs of commuting operators:

XMb = Mb X; DTa = Ta D.

However, simple calculation shows that X and D do not commute but satisfy the so-called canonical

commutation relation:
1
[D, X] := DX − XD = I, (1.1)
2πi

where I : D(X) ∩ D(D) → D(X) ∩ D(D) is the identity map. Rewrite (1.1) in terms of Ta and Mb ,

it is equivalent to

Ta Mb = e−2πiba Mb Ta .

2
As a consequence of the canonical commutation relation, we obtain

[X, Ta ] := XTa − Ta X = aTa ;

[D, Mb ] := DMb − Mb D = bMb . (1.2)

In next section, we will see the canonical commutation relation plays an important role in the

classical uncertainty principle.

1.2 Classical Uncertainty Principle

Before we state the classical uncertainty principle, we need to introduce a very important

concept in analysis which is the Fourier transform:

For f ∈ L1 (R) ∩ L2 (R), the Fourier transform fb : R → C is defined by

Z
fb(ξ) := f (x)e−2πixξ dx, ∀ξ ∈ R.
R

And we will also denote fb by Ff . This notation indicates that we would consider the Fourier

transform F as an operator as well. By Plancherel theorem, F could be extended as an isometry

from L2 (R) onto L2 (R). Throughout the thesis, we would view the Fourier transform F as an

unitary operator on L2 (R). Since F is unitary on L2 (R), its inverse Fourier transform F −1

equals its adjoint F ∗ given by

Z
f (x) = fb(ξ)e2πixξ dξ, ∀x ∈ R.
R

For us, the importance of the Fourier transform F mainly comes from the fact that it connects

multiplication X and differentiation D in the following way:

FX = −DF; FD = XF.

As a consequence, F could also transform Ta and Mb as the following:

FMb = Tb F; FTa = M−a F.

3
We now could state the classical uncertainty principle (which is often referred to as the

Heisenberg-Pauli-Weyl inequality) as the following:

Theorem 1.2.1 ([5], Theorem 1.1). For any f ∈ L2 (R) and any a, b ∈ R,

Z Z
1 4 2
kf k2 ≤ |(x − a)f (x)| dx |(ξ − b)fb(ξ)|2 dξ. (1.3)
16π 2 R R

2
Equality holds if and only if f (x) = Ce2πibx e−c(x−a) for some C ∈ C and c > 0.

Note the inequality (1.3) can be written as the following equivalent form in terms of X and

D. Namely, for any f ∈ D(X) ∩ D(D) and any a, b ∈ R,

1 4 2 2
kf k2 ≤ k(X − a)f k2 k(D − b)f k2 . (1.4)
16π 2

Proof. Here is a simple proof for Schwartz class S(R) using the canonical commutation relation.

Assume f ∈ S(R). For every a, b ∈ R, by (1.1) we have

 
1 2 1
kf k2 = f, f
2πi 2πi

= h[D, X]f, f i

= h(DX − XD)f, f i

= h((D − b)(X − a) − (X − a)(D − b))f, f i

= h(X − a)f, (D − b)f i − h(D − b)f, (X − a)f i

= 2iIm h(X − a)f, (D − b)f i .

Apply Cauchy-Schwartz, we get the desired inequality,

1 4 2
kf k2 = |Im h(X − a)f, (D − b)f i|
16π 2
2
≤ |h(X − a)f, (D − b)f i|
2 2
≤ k(X − a)f k2 k(D − b)f k2 . (1.5)

Notice equality in (1.5) holds if and only if (D − b)f = ic(X − a)f for some c ∈ R. This is the

4
differential equation:

f 0 − 2πibf = −2πc(x − a)f.

2
Its solutions are Ce2πibx e−c(x−a) for every C ∈ C. Finally, f ∈ L2 (R) requires c > 0.

The proof for f ∈ L2 (R) is similar. Instead of using (1.1), we need the generalized

canonical commutation relation. That is, for any f ∈ D(X) ∩ D(D),

1 2
kf k2 = hXf, Df i − hDf, Xf i . (1.6)
2πi

Since (1.6), X and D cannot have a common eigenvector in L2 (R) (otherwise we would get

a contradiction). And (1.4) could be viewed as a quantitive version of this argument. It tell us there

is no sequence of L2 (R) functions which approximates to and performs like a common eigenvector

for X and D. However, we could find some functions in L2 (R) which minimize the right hand side in

(1.4). Naturally, we might call these functions the “best approximate common eigenvectors”. And
2
it turns out these approximants are the multiples of Mb Ta g, where g(x) = e−cx is the Gaussian

function for some c > 0.

1.3 Classical Balian–Low Theorem

Recall the spectral theorem in functional analysis, if there are two compact self-adjoint

operators on a Hilbert space commuting with each other, then we could find a sequence of common

eigenvectors which forms an orthonormal basis for the Hilbert space. Go back to operators X and

D, the classical uncertainty principle shows that instead of common eigenvectors, we could find some

“best approximate common eigenvectors” for X and D. The natural question is: can we use these

“best approximants” to form an orthonormal basis just like in the spectral theorem. Unfortunately,

the answer is No!

If we set g(x) := 1[0,1] (x), the sequence {Mm Tn g}m,n∈Z which does not consist of the “best

approximants” does form an orthonormal basis of L2 (R). However, in this case, the Fourier transform

of g has a very poor localization property. The question is: can we find some g ∈ L2 (R) such that

{Mm Tn g}m,n∈Z forms an orthonormal basis, in addition, g and gb are both well localized. Again,

the answer is No! It is given by the so-called Balian-Low theorem. In order to state the theorem,

5
we need to introduce the Gabor systems.

Definition 1.3.1. A Gabor system is a sequence in L2 (R) of the form {Mmb Tna g}m,n∈Z , where

a, b > 0 and g ∈ L2 (R) is a fixed non-zero function.

The function g is called the window function or the generator. Explicitly,

Mmb Tna g(x) = e2πimbx g(x − na), x ∈ R.

The set Λ = {(na, mb)}m,n∈Z is called the lattice with volume ab. Actually, we could generalize the

definition for high dimensional cases, i.e., we could define the Gabor system {Mmb Tna g}m,n∈Zn for

L2 (Rn ) in a similar way.

Now we state the classical Balian-Low theorem:

Theorem 1.3.2 ([2], Theorem 1.1). Let g ∈ L2 (R) and let a, b > 0 satisfy ab = 1. If the Gabor

system {Mmb Tna g}m,n∈Z is an orthonormal basis for L2 (R), then

Z Z
2
|xg(x)| dx g (ξ)|2 dξ = ∞.
|ξb (1.7)
R R

Note the expression (1.7) can be replaced by: for any x0 , ξ0 ∈ R

Z Z
|(x − x0 )g(x)|2 dx g (ξ)|2 dξ = ∞.
|(ξ − ξ0 )b
R R

Proof. We prove the theorem by contradiction. Suppose

Z Z
|xg(x)|2 dx g (ξ)|2 dξ < ∞.
|ξb
R R

6
Which implies Xg, Dg ∈ L2 (R), then by (1.2) we have

hXg, Dgi
X
= hXg, Mmb Tna gi hMmb Tna g, Dgi
m,n
X
= hT−na M−mb Xg, gi hg, T−na M−mb Dgi
m,n
X
= h(XM−mb T−na + naM−mb T−na )g, gi hg, (DM−mb T−na + mbM−mb T−na )gi
m,n
X
= hXM−mb T−na g, gi hg, DM−mb T−na gi
m,n
X
= hM−mb T−na g, Xgi hDg, M−mb T−na gi
m,n
X
= hDg, M−mb T−na gi hM−mb T−na g, Xgi
m,n

= hDg, Xgi .

Combining (1.6), we have


1 2
kgk2 = hXg, Dgi − hDg, Xgi = 0.
2πi

It contradicts with the fact that {Mmb Tna g}m,n∈Z forms an orthonormal basis.

The classical Balian-Low theorem tells us that there is no well-localized window function

g in both time and frequency, which generates an orthonormal basis. Furthermore, in some sense

(1.7) is sharp. Actually, Benedetto et al. [1] shows that for any ε > 0, there exists an orthonormal

basis {Mm Tn g}m,n∈Z such that

2 2
1 + |x| 1 + |ξ|
Z Z
|g(x)|2 dx g (ξ)|2
|b dξ < ∞.
R log1+ε (2 + |x|) R
2+ε
log (2 + |ξ|)

1.4 Frames and Riesz Sequences

If we ask for the Gabor system to be a weaker type of basis such that the Balian-Low theorem

still holds. The answer would be the Riesz basis which is slightly weaker than the orthonormal basis.

Here is the definition of such basis:

Definition 1.4.1. A sequence {fk }∞


k=1 in a complex and separable Hilbert space H is called a Riesz

7
basis, if there exists an invertible operator U : H → H such that

f k = U ek , ∀k ∈ N,

where {ek }∞
k=1 is an orthonormal basis of H.

Throughout the thesis, we always view H as a complex and separable Hilbert space. Just

as any basis, the Riesz basis has two dual features: spanning and independence. The sequence in

H which is usually viewed as a representative of spanning property is the so-called frame. And the

sequence in H which is usually viewed as a representative of independence property is the so-called

Riesz sequence. In this section, for completeness we will give their definitions and state their main

properties. And we will not include proofs which can be found in [4, 20].

Definition 1.4.2. A sequence {fk }∞


k=1 in H is a (discrete) frame of H, if there exist constants

A, B > 0 such that



X
2 2 2
A kf k ≤ |hf, fk i| ≤ B kf k , ∀f ∈ H. (1.8)
k=1

The (optimal: maximal for A and minimal for B) constants are called the (optimal) lower

and upper frame bounds. A frame is said to be tight if we can pick A = B as frame bounds. And

a frame is called a Parseval frame if A = B = 1, i.e.,


X
2 2
kf k = |hf, fk i| , ∀f ∈ H.
k=1

Definition 1.4.3. A sequence {fk }∞


k=1 in H is said to be complete if

span{fk }∞
k=1 = H.

It is not hard to see that if a sequence satisfies the inequality to the left in (1.8), then it

has to be complete. As a consequence, any frame is a complete sequence. If a sequence satisfies the

inequality to the right in (1.8), such sequence is said to be Bessel:

Definition 1.4.4. A sequence {fk }∞


k=1 in H is called a Bessel sequence if there exists a constant

B > 0 such that



X 2 2
|hf, fk i| ≤ B kf k , ∀f ∈ H.
k=1

8
The (optimal) constant B is called the (optimal) Bessel bound. By the definition, we can

prove every Bessel sequence is a bounded sequence. Another equivalent definition of Bessel sequences

is as the following.

Proposition 1.4.5. A sequence {fk }∞


k=1 in H is a Bessel sequence if and only if there exist constant

B > 0 such that ∞ 2


X ∞
X 2
c f ≤ B |ck | ,

k k

k=1 k=1

for any finite complex sequence {ck }.

Given a Bessel sequence, we could well define the following operators:

Definition 1.4.6. Let {fk }∞


k=1 be a Bessel sequence of H.

1. The synthesis operator T is defined by


X
T : l2 (N) → H, T {ck }∞
k=1 := ck fk .
k=1

2. Its adjoint operator T ∗ called the analysis operator is defined by

T ∗ : H → l2 (N), T ∗ f := {hf, fk i}∞


k=1 .

3. The frame operator S is the composition of T and T ∗ ,


X
S : H → H, Sf := T T ∗ f = hf, fk i fk .
k=1

Actually, {fk }∞
k=1 being a Bessel sequence not only guarantees the operators are well-defined,

but also make them being bounded. A very interesting fact is that if the operator T or T ∗ is well

defined, then the sequence {fk }∞


k=1 has to be Bessel.

From the definition, it is easy to see that the frame operator S is positive and self-adjoint.

In addition, if {fk }∞
k=1 is a frame of H, then the frame operator S is invertible, i.e., S
−1
exists. It

9
follows from the definition that


X
f, S −1 fk fk ,


f= ∀f ∈ H;
k=1
X∞
f= hf, fk i S −1 fk , ∀f ∈ H.
k=1

Notice there are some elements fek := S −1 fk (for any k ∈ N) appearing in the last two series. The

sequence {fek }∞ ∞
k=1 is called the canonical dual frame of {fk }k=1 . And one could show the canonical

dual frame is also a frame of H.

Furthermore, if {fk }∞
k=1 is a Parseval frame, then S = I. So the canonical dual frame is

itself, i.e., fek = fk for any k ∈ N. In this case


X
f= hf, fk i fk , ∀f ∈ H.
k=1

Definition 1.4.7. A sequence {fk }∞


k=1 in H is a Riesz sequence, if there exist constants A, B > 0

such that 2

X X∞ ∞
X
2 2
A |ck | ≤ ck fk ≤ B |ck | , (1.9)


k=1 k=1 k=1

for any finite complex sequence {ck }.

The (optimal) constants A and B are called the (optimal) lower and upper Riesz bounds.

By Proposition 1.4.5, we have already known a sequence satisfies the inequality to the right

in (1.9) is a Bessel sequence. Now we introduce another important sequence which satisfies the

inequality to the left in (1.9), such sequence is the so-called Riesz-Fischer sequence.

Definition 1.4.8. A sequence {fk }∞


k=1 in H is called a Riesz-Fischer sequence if there exists a

constant A > 0 such that 2



X X∞
2
A |ck | ≤ ck fk , (1.10)


k=1 k=1

for any finite complex sequence {ck }.

By previous discussions, we have an equivalent definition of Riesz sequences.

Proposition 1.4.9. {fk }∞ ∞


k=1 is a Riesz sequence of H if and only if {fk }k=1 is not only a Bessel

sequence but also a Riesz-Fischer sequence of H.

10
About Riesz-Fischer sequences, we also have the following two useful equivalent definitions.

And they will be applied in Chapter 3.

Proposition 1.4.10. {fk }∞


k=1 is a Riesz-Fischer sequence of H if and only if the moment problem

of sequence {fk }∞
k=1 in H is solvable, i.e., there exists an element f ∈ H such that

hf, fk i = ck , ∀k ∈ N,

for any sequence {ck }∞ 2


k=1 ∈ l (N).

Proposition 1.4.11. {fk }∞


k=1 is a Riesz-Fischer sequence of H if and only if there exists a Bessel

sequence {gk }∞
k=1 of H such that

hfk , gj i = δkj , ∀k, j ∈ N,

where δkj is the Kronecker delta symbol.

Definition 1.4.12. Two sequences {fk }∞ ∞


k=1 and {gk }k=1 in H are said to be biorthogonal if

hfk , gj i = δkj , ∀k, j ∈ N.

And a sequence that admits a biorthogonal sequence will be called minimal.

By Proposition 1.4.11, every Riesz-Fischer sequence is a biorghogonal sequence (or a minimal

sequence). Go back to the Riesz basis, we have the following two important propositions:

Proposition 1.4.13. A sequence {fk }∞


k=1 in H is a Riesz basis if and only if it is a complete Riesz

sequence.

Proposition 1.4.14. A Riesz basis {fk }∞ e ∞


k=1 of H is a frame. And its canonical dual frame {fk }k=1

is also a Riesz basis. In addition, {fk }∞ e ∞


k=1 and {fk }k=1 are biorthogonal sequences.

11
Chapter 2

Continuous Frames and Toeplitz

Operators in Framed Hilbert

Spaces

2.1 Continuous Frames and Examples

In Chapter 2, we will introduce a very important concept for the thesis which is the so-called

“continuous frame”, and make some preparations for the rest of the chapters.

2.1.1 Continuous Frames

Let H be a Hilbert space.

Definition 2.1.1. A continuous frame of H is a family of elements {kx }x∈(X,σ) indexed on a

measure space (X, σ), such that

1. σ is a σ-finite positive measure;

2. x → hf, kx i is a measurable function on X for any f ∈ H;

12
3. there exist constants A, B > 0 such that

Z
2 2 2
A kf k ≤ |hf, kx i| dσ(x) ≤ B kf k , ∀f ∈ H.
X

Throughout the thesis, we may omit the measure σ for some continuous frame {kx }x∈X ,

when doing this will not cause any doubt.

A continuous frame {kx }x∈(X,σ) is said to be a continuous Parseval frame if A = B = 1.

Namely,
Z
2 2
kf k = |hf, kx i| σ(x), ∀f ∈ H.
X

And a continuous frame {kx }x∈X is said to be a normalized continuous frame if

kkx k = 1, ∀x ∈ X.

Note that if the measure space X = N with σ being the counting measure, then the continu-

ous frame {kx }x∈N becomes a discrete frame (see 1.4.2). So continuous frames are the generalization

of discrete frames, and some mathematicians use the phrase the generalized frame instead of the

continuous frame.

Proposition 2.1.2. Let {kx }x∈(X,σ) be a continuous frame of H, then {kx }x∈(X,σ) is complete, i.e.,

span{kx }x∈X = H.

Definition 2.1.3. Let {kx }x∈(X,σ) be a continuous frame of H, the frame operator S : H → H is

given by
Z
Sf := hf, kx i kx dσ(x),
X

where the right-hand side is defined in the weak sense, i.e., as the unique element in H such that

Z
hSf, gi = hf, kx i hkx , gi dσ(x), ∀g ∈ H.
X

Just like the discrete case, the frame operator is invertible for continuous frames. Let

kx := S −1 kx for any x ∈ X. The family of elements {e


e kx }x∈X is called the canonical dual frame

of {kx }x∈X . In addition, the canonical dual frame is also a continuous frame of H.

13
Again, if {kx }x∈X is a continuous Parseval frame, the corresponding frame operator is the

identity map I : H → H. Then the canonical dual frame is itself, i.e., e


kx = kx for any x ∈ X. In

this case
Z
f= hf, kx i kx dσ(x), ∀f ∈ H.
X

2.1.2 Continuous Gabor Frames

One example of continuous frames is the continuous Gabor frame.

Definition 2.1.4. A continuous Gabor frame is a family in L2 (Rn ) of the form {Mb Ta g}(a,b)∈R2n ,

where g ∈ L2 (Rn ) is a fixed non-zero function.

Proposition 2.1.5. The continuous Gabor frame {Mb Ta g}(a,b)∈(R2n ,m) is a continuous frame of

L2 (Rn ), where m is the Lebesgue measure on R2n . If kgk2 = 1, then {Mb Ta g}(a,b)∈R2n is a normalized

continuous Parseval frame of L2 (Rn ).

Proof. We give the following proof for n = 1. In high dimensional case, the proof is similar. For any

f ∈ L2 (R),

Z Z
2
|hf, Mb Ta gi| dbda
R R
Z Z Z 2

= 2πibx g(x − a)dx dbda
f (x)e

R R R
Z Z Z 2
f (x)g(x − a)e−2πibx dx dbda

=
ZR ZR R  2
= F f (·)g(· − a) (b) dbda

R R
Z Z 2
= f (b)g(b − a) dbda

ZR ZR
2
= |f (b)g(b − a)| dadb
ZR R Z
2 2
= |f (b)| |g(b − a)| dadb
ZR ZR
2 2
= |f (b)| |g(a)| dadb
R R
2 2
= kgk2 kf k2 . (2.1)

14
2
Denote kgk2 by A, then (2.1) shows that

Z Z
2
A kf k2 = | hf, Mb Ta gi |2 dbda, ∀f ∈ L2 (R).
R R

So {Mb Ta g}(a,b)∈(R2 ,m) is a continuous frame of L2 (R).

Note if kgk2 = 1, we have kMb Ta gk2 = 1 for every (a, b) ∈ R2 . In addition,

Z Z
2 2
kf k2 = |hf, Mb Ta gi| dbda, ∀f ∈ L2 (R).
R R

In this case, {Mb Ta g}(a,b)∈R2 is a normalized continuous Parseval frame of L2 (R).

2.1.3 Continuous Exponential Frames

Another example of continuous frames is the exponential functions, which is very familiar

to us but easy to ignore as a continuous frame.

Proposition 2.1.6. Let E be a subset of R with finite Lebesgue measure, then the exponential func-

tions {e2πiξx }ξ∈(R,m) is a continuous Parseval frame of L2 (E). If m(E) = 1, then {e2πiξx }ξ∈(R,m)

is a normalized continuous Parseval frame.

Proof. Let f be any function in L2 (E), and we view f as a L2 (R) function supported on E. Notice

Z D E 2
f, e2πiξ(·)


2
L (E)

R
Z Z 2
= f (x)e−2πiξx dx dξ

R E
Z Z 2
−2πiξx

= f (x)e
dx dξ
ZR R 2
= fb(ξ) dξ

ZR
2
= |f (x)| dx
ZR
2
= |f (x)| dx
E
2
= kf kL2 (E) .

15
It follows that {e2πiξx }ξ∈(R,m) is a continuous Parseval frame of L2 (E). Furthermore, if m(E) = 1,

Z
2πiξ(·) 2
2πiξx 2
e 2 = e dx = m(E) = 1.
L (E) E

Then, {e2πiξx }ξ∈(R,m) is a normalized continuous Parseval frame.

2.2 Framed Hilbert Spaces and Examples

2.2.1 Framed Hilbert Spaces

Definition 2.2.1. A framed Hilbert space (FHS) is a pair (H, {kx }x∈(X,σ) ), where H is a complex

Hilbert space, and {kx }x∈(X,σ) is a normalized continuous Parsevel frame of H.

Actually, one can view a FHS as another more familiar space, which is the so-called repro-

ducing kernel Hilbert space.

Definition 2.2.2. Let X be an arbitrary set. A Hilbert space H consisting of functions f : X → C

is said to be a reproducing kernel Hilbert space (RKHS) if for any x ∈ X, the evaluation

functional δx given by

δx : H → C, δx (f ) := f (x),

is bounded.

By Riesz representation theorem, for any x ∈ X, there exists an unique element Kx ∈ H

such that

f (x) = hf, Kx i , ∀f ∈ H.

The collection of elements {Kx }x∈X is called the reproducing kernel of H; the element Kx is

called the reproducing kernel of H at point x. And the collection {kx := Kx / kKx k}x∈X is called

the normalized reproducing kernel of H; the element kx is called the normalized reproducing

kernel of H at point x.

The following two propositions show that every RKHS which is embedded in a L2 (X, γ) for

some positive measure γ is a FHS and vice versa.

Proposition 2.2.3. Let H ⊆ L2 (X, γ) be a RKHS, and {Kx }x∈X be its reproducing kernel. Then

16
2
(H, {kx }x∈(X,σ) ) forms a FHS, where dσ(x) := kKx k dγ(x), and {kx }x∈X is the normalized repro-

ducing kernel.

Proof. Obviously, kkx k = 1 for any x ∈ X. In addition,

Z
2 2
kf k = |f (x)| dγ(x)
ZX
2
= |hf, Kx i| dγ(x)
ZX
2 2
= |hf, kx i| kKx k dγ(x)
X
Z
2
= |hf, kx i| dσ(x),
X

for any f ∈ H. It implies {kx }x∈(X,σ) is a normalized continuous Parsevel frame, and (H, {kx }x∈(X,σ) )

forms a FHS.

Proposition 2.2.4. Let (H, {kx }x∈(X,σ) ) be a framed Hilbert space. Then H can be viewed as a

RKHS embedded in L2 (X, σ).

Proof. Define the space H


e by

H
e := {fe : X → C| fe(x) := hf, kx i , f ∈ H}.

Notice for any f ∈ H,

Z
e 2

kfek22 : = f (x) dσ(x)
ZX
2
= |hf, kx i| dσ(x)
X
2
= kf k .

e ⊆ L2 (X, σ). Then the map f → fe is an


So the linear map f → fe is an isometry from H onto H

unitary operator from H to H.


e Notice for any x ∈ X, the evaluation functional δx on H
e satisfies the

17
boundedness property:


δx (fe) : = fe(x)

= |hf, kx i|

≤ kf k kkx k

= kf k

= kfek2 ,

for any fe ∈ H.
e By δx is bounded, H
e is a RKHS. Since f → fe is unitary, for any x ∈ X we have

D E
fe(x) = hf, kx i = fe, e
kx , ∀fe ∈ H.
e
2

So {e
kx }x∈X is the reproducing kernel of H.
e

Now we are going to list some classical examples of FHS.

2.2.2 L2 (Rn ) Space with Continuous Gabor Frame

Define L2 (Rn ) as the Hilbert space as usual. And let {Mb Ta g}(a,b)∈R2n be the continuous

Gabor frame with kgk2 = 1, i.e., for any (a, b) ∈ R2n

Mb Ta g(x) = e2πib·x g(x − a), ∀x ∈ Rn .

By Proposition 2.1.5, (L2 (Rn ), {Mb Ta g}(a,b)∈(R2n ,m) ) is a FHS.

2.2.3 Classical Bargmann-Fock Spaces

The classical Bargmann-Fock space Fα (Cn ) with the parameter α > 0 is the space of

all entire functions f : Cn → C satisfying the following integrability condition

αn
Z
2 2 2
kf kα := |f (z)| e−α|z| dm(z) < ∞,
πn Cn

18
where m is the Lebesgue measure on Cn . It is known that Fα (Cn ) is a reproducing kernel Hilbert

space. Its reproducing kernel at point z ∈ Cn equals

Kzα (w) = eαhw,zi , ∀w ∈ Cn ;

and the normalized reproducing kernel at point z equals

α 2
kzα (w) = eαhw,zi− 2 |z| , ∀w ∈ Cn .

Define measure mα by

2 αn −α|z|2 αn
dmα (z) := kKz kα n
e dm(z) = n dm(z),
π π

then the classical Bargmann-Fock space (Fα , {kzα }z∈(Cn ,mα ) ) forms a framed Hilbert space.

When α = π, by convention, we will drop the sub(super)scripts α in the above notations.

We will use k·k , Kz (w), and kz (w) to denote the norm, the reproducing kernel, and the normalized

reproducing kernel of F(Cn ) at z in this case.

Because there is an unitary operator B (the Bargmann transform) from L2 (Rn ) onto F(Cn ),

which maps the continuous Gabor frame {Mb Ta g}(a,b)∈R2n generated by the normalized Gaussian

window function g onto the normalized reproducing kernel {kz }z∈Cn of F(Cn ) (up to some complex

numbers with modulus 1). One could view the classical Bargmann-Fock space (Fα , {kzα }z∈(Cn ,mα ) ) as

a special case of the FHS (L2 (Rn ), {Mb Ta g}(a,b)∈(R2n ,m) ), that is when g is the normalized Gaussian

function.

2.2.4 General Paley-Wiener Spaces

Let E be a subset of R with finite Lebesgue measure, we define the general Paley-Wiener

space as the following

L2 (E) := {f ∈ L2 (R) : suppf ⊆ E}.

2πiξx
e 1
let { √ }ξ∈R be the √ multiple of exponential functions. As shown in Proposition 2.1.6,
m(E) m(E)
2πiξx
e
we obtain { √ }ξ∈(R,mE ) is a normalized continuous Parseval frame of L2 (E), where dmE :=
m(E)

m(E)dm is the m(E) multiple of the Lebesgue measure.

19
2πiξx
e
The general Paley-Wiener space (L2 (E), { √ }ξ∈(R,mE ) ) forms a FHS.
m(E)

2.2.5 Classical Paley-Wiener Spaces

The classical Paley-Wiener space PWα with the parameter α > 0 is given by

PWα := {f ∈ L2 (R, k·k2 ) : suppfb ⊆ [−α, α]}.

PWα is a RKHS as well. Its reproducing kernel at point x ∈ R equals

sin α(y − x)
Kxα (y) = , ∀y ∈ R;
π(y − x)

and the normalized reproducing kernel at x equals


α sin α(y − x)
kxα (y) = √ , ∀y ∈ R.
π α(y − x)

Define measure mα by
2 α
dmα (x) := kKxα k2 dm(x) = dm(x),
π

Then the classical Paley-Wiener space (PWα , {kxα }x∈(R,mα ) ) forms a FHS.

Because the Fourier transform F is an unitary operator from PWα onto L2 [−α, α], which
2πiξx
maps the normalized reproducing kernel {kxα }x∈R of PWα onto the continuous frame { e√2α }ξ∈R
2πiξx
of L2 [−α, α]. One could view the space (L2 [−α, α], { e√2α }ξ∈(R,m[−α,α] ) ) as the classical Paley-

Wiener space (PWα , {kxα }x∈(R,mα ) ). In addition, for any finite interval I ⊆ R with length 2α, there
2πiξx
exists a translation operator mapping from L2 [−α, α] onto L2 (I), which also maps { e√2α }ξ∈R onto
2πiξx
e
{√ }ξ∈R up to some constants with modulus 1. Based on these arguments, we could also view
m(I)
2πiξx
e
the space (L2 (I), { √ }ξ∈(R,|·|,mI ) ) as a classical Paley-Wiener space.
m(I)

2.3 Toeplitz Operators on Framed Hilbert Spaces

2.3.1 Toeplitz Operators

Let (H, {kx }x∈(X,σ) ) be a FHS.

Definition 2.3.1. Let µ be a σ-finite positive measure on the same measurable sets with σ, such

20
that
Z
|hkx , ky i| dµ(y) < ∞, ∀x ∈ X.
X

We densely define the Toeplitz operator Tµ of {kx }x∈X with symbol µ by

Z
Tµ : span{kx }x∈X → H, Tµ f := hf, kx i kx dµ(x),
X

where the right-hand side is defined in the weak sense.

Proposition 2.3.2. The Toeplitz operator Tµ with symbol µ is a positive symmetric linear operator.

Proof. Obviously, Tµ is linear.

1. Tµ is positive:

Z
hTµ f, f i = hf, kx i hkx , f i dµ(x)
X
Z
2
= |hf, kx i| dµ(x)
X

≥ 0,

for any f ∈ span{kx }x∈X .

2. Tµ is symmetric:

Z
hTµ f, gi = hf, kx i hkx , gi dµ(x)
X
Z
= hg, kx i hkx , f i dµ(x)
X

= hTµ g, f i

= hf, Tµ gi ,

for any f, g ∈ span{kx }x∈X .

Proposition 2.3.3. Let Tµ be a Toeplitz operator with a finite measure µ. Then Tµ can be extend

to a positive bounded self-adjoint linear operator on H.

21
Proof. By Proposition 2.3.2, we only need to show the boundedness of Tµ on span{kx }x∈X . For any

f ∈ span{kx }x∈X , we have

2 2
kTµ f k = sup |hTµ , gi|
kgk=1
Z 2

= sup hf, kx i hkx , gi dµ(x)

kgk=1 X
Z Z
2 2
≤ sup |hf, kx i| dµ(x) |hkx , gi| dµ(x)
kgk=1 X X
Z Z
2 2 2 2
≤ sup kf k kkx k dµ(x) kkx k kgk dµ(x)
kgk=1 X X

2 2
= sup µ(X)2 kf k kgk
kgk=1

2
= µ(X)2 kf k .

Now we could view the Toeplitz operator Tµ as a bounded operator on the whole H, if we

impose the assumption that µ is a finite measure.

Proposition 2.3.4. Let Tµ be a Toeplitz operator with a finite measure µ. Then Tµ is in the trace

class, and its trace and Hilbert-Schmidt norm satisfy the following identities:

Z Z
2
kTµ kT r = µ(X) = |hkx , ky i| dσ(y)dµ(x); (2.2)
X X

Z Z
2 2
kTµ kHS = |hkx , ky i| dµ(y)dµ(x). (2.3)
X X

22
Proof. Let {en }∞
n=1 be an orthonormal basis of H. On one hand,

kTµ kT r = T r(|Tµ |)

= T r(Tµ )

X
= hTµ en , en i
n=1
X∞ Z
2
= |hen , kx i| dµ(x)
n=1 X

Z X
2
= |hen , kx i| dµ(x)
X n=1
Z
2
= kkx k dµ(x)
X

= µ(X) < ∞.

Which implies Tµ is in the trace class, and kTµ kT r = µ(X). Then (2.2) comes from the following

Z Z
2
|hkx , ky i| dσ(y)dµ(x)
X X
Z
2
= kkx k dµ(x)
X

=µ(X).

23
On the other hand,

2
kTµ kHS = T r(Tµ∗ Tµ )

X
Tµ∗ Tµ en , en


=
n=1
X∞
= hTµ en , Tµ en i
n=1
X∞ Z Z 
= hen , kx i kx dµ(x), hen , ky i ky dµ(y)
n=1 X X

X∞ Z  Z 
= hen , kx i kx , hen , ky i ky dµ(y) dµ(x)
n=1 X X

X∞ Z Z
= hen , kx i hky , en i hkx , ky i dµ(y)dσ(x)
n=1 X X

X∞ Z Z
= hky , en i hen , kx i hkx , ky i dµ(y)dµ(x)
n=1 X X

Z Z X
= hky , en i hen , kx i hkx , ky i dµ(y)dµ(x)
X X n=1
Z Z
= hky , kx i hkx , ky i dµ(y)dµ(x)
X X
Z Z
2
= |hkx , ky i| dµ(y)dµ(x).
X X

In general, it is not that easy to characterize the boundedness, compactness and the trace

class membership for Toeplitz operators without the assumption µ(X) < ∞. We will continue to do

the characterizations later.

2.3.2 Concentration Operators

Let (H, {kx }x∈(X,σ) ) be a FHS, we now introduce a very special Toeplitz operator by letting

the measure dµ := 1E dσ for some measurable subset E of X, such operator is called the concentration

operator.

Definition 2.3.5. Let E be a measurable subset of X. Define the concentration operator TE

over E by
Z
TE : H → H, TE f := hf, kx i kx dσ(x),
E

24
where the right-hand side is defined in the weak sense.

Proposition 2.3.6. Let TE be a concentration operator over the set E, then TE is bounded with

kTE k ≤ 1.

Proof. For any f ∈ H, we have

2 2
kTE f k = sup |hTE , gi|
kgk=1
Z 2

= sup hf, kx i hkx , gi dσ(x)

kgk=1 E
Z Z
2 2
≤ sup |hf, kx i| dσ(x) |hkx , gi| dσ(x)
kgk=1 E E
Z Z
2 2
≤ sup |hf, kx i| dσ(x) |hkx , gi| dσ(x)
kgk=1 X X

2 2
= sup kf k kgk
kgk=1

2
= kf k .

So kTE k ≤ 1.

Proposition 2.3.6 shows that as a special Toeplitz operator, the concentration operator TE

(equals Tµ , dµ := 1E dσ) will always being bounded. In addition, if we assume σ(E) is finite, we

have µ(X) = σ(E) < ∞. By Proposition 2.3.3 and 2.3.4, we obtain the following corollary.

Corollary 2.3.7. Let TE be the concentration operator over the set E with σ(E) < ∞. Then TE is

a positive compact self-adjoint operator, and its trace and Hilbert-Schmidt norm satisfy the following

identities:
Z Z
2
kTE kT r = σ(E) = |hkx , ky i| dσ(y)dσ(x); (2.4)
E X

Z Z
2 2
kTE kHS = |hkx , ky i| dσ(y)dσ(x). (2.5)
E E

Let E be a finite measure set, Corollary 2.3.7 tell us the concentration operator TE is a

positive compact self-adjoint operator. Combine Proposition 2.3.6, we know all the eigenvalues of

TE are belonged to the interval [0, 1].

Definition 2.3.8. Let E be a measurable subset of X with σ(E) < ∞. Given a number 0 < c < 1,

25
we say that a subspace G of H is c-concentrated on set E if

Z
2 2
c kf k ≤ |hf, kx i| dσ(x),
E

for every f ∈ G.

Lemma 2.3.9. Let E be a Borel subset of X with σ(E) < ∞. Given a number 0 < c < 1, if G is a

subspace of H which is c-concentrated on E, then

σ(E)
dimG ≤ .
c

Proof. Let TE be the concentration operator over E. Denote all the eigenvalues of TE in the

decreasing order by 1 ≥ l1 ≥ l2 ≥ · · · ≥ 0, where entries are repeated with multiplicity.

Let G 0 be an arbitrary finite-dimensional subspace of G and k = dimG 0 . By the min-max

principle,

lk = max min hTE f, f i


dimH=k f ∈H,kf k=1

≥ min hTE f, f i
f ∈G 0 ,kf k=1
Z
2
= min |hf, kx i| dσ(x)
f ∈G 0 ,kf k=1 E
2
≥ min c kf k
f ∈G 0 ,kf k=1

= c.

Combine (2.4) we have


X
σ(E) = kTE kT r = T r(TE ) = lj ≥ klk ≥ kc.
j=1

So we obtain
σ(E)
dimG 0 = k ≤ .
c

Since G 0 was an arbitrary finite-dimensional subspace of G, we obtain that G is finite-dimensional

and the same estimate of the dimension holds for G.

26
Lemma 2.3.9 shows that, given a finite measure set E, we could find a “inverse” relation

between the dimension of a concentrated subspace G on E and its concentration level c. Notice,

from the proof, if we knew the distribution of eigenvalues of TE , we could get a more precise relation

between dimG and c. We will see it later.

2.4 Additional Conditions on Framed Hilbert Spaces

In this section, we will impose additional conditions on FHS to be able to obtain additional

properties. All these conditions are natural, and almost all of them are satisfied in the previous

examples.

2.4.1 Framed Hilbert Spaces with Metric

Definition 2.4.1. We say (H, {kx }x∈(X,d,σ) ) is a framed Hilbert space with metric d, if

(H, {kx }x∈(X,σ) ) is a FHS and σ is a Borel measure with respect to the imposed metric d on X.

Recall the examples of FHS in section 2.2,

1. The space (L2 (Rn ), {Mb Ta g}(a,b)∈(R2n ,m) ),

2. the classical Bargmann-Fock space (Fα , {kzα }z∈(Cn ,mα ) ),

2πiξx
e
3. the general Paley-Wiener space (L2 (E), { √ }ξ∈(R,mE ) ),
m(E)

4. the classical Paley-Wiener space (PWα , {kxα }x∈(R,mα ) )

Because all of the measures in these FHS are the Borel measure with respect to the Euclidean metric

|·|. If we impose the Euclidean metric |·| on their index space respectively, then all of them are the

framed Hilbert space with the Euclidean metric |·|.

For convenience, we will continue using the phrase “framed Hilbert space” or “FHS ” to call

a framed Hilbert space (H, {kx }x∈(X,d,σ) ) with the metric d.

2.4.2 Additional Conditions

Let (H, {kx }x∈(X,d,σ) ) be a FHS, and we continue to impose conditions on it.

27
We say the metric measure space (X, d, σ) satisfies the doubling measure property

(DMP ), if there exists a constant C > 0 such that

0 < µ(B(a, 2r)) ≤ Cµ(B(a, r)) < ∞,

for all a ∈ X and r > 0.

Simple calculation shows that the Euclidean metric space with the Lebesgue measure (Rn , |·| , m)

satisfies DMP with C = 2n .

The metric measure space (X, d, σ) is said to satisfy the annular decay property (ADP ),

if for any ρ > 0 we have


σ(B(a, r + ρ))
lim sup sup = 1. (2.6)
r→∞ a∈X σ(B(a, r))

Notice (2.6) is equivalent to

σ(B(a, r + ρ)) \ σ(B(a, r))


lim sup sup = 0.
r→∞ a∈X σ(B(a, r))

Easily see (Rn , |·| , m) satisfies ADP as well. For other metric measure spaces, we provide a

method to verify ADP by the following proposition.

Proposition 2.4.2 ([3]). Let (X, d, σ) be a metric measure space. If (X, d, σ) satisfies DMP and

(X, d) is also a length space, then (X, d, σ) satisfies ADP.

We say (X, d, σ) satisfies the uniform measure distribution (UMD), if for every r > 0

there exist constants D(r) ≥ c(r) > 0 such that

c(r) ≤ σ(B(x, r)) ≤ D(r),

for any ball B(x, r) in X with the radius r, where c(r) → ∞ as r → ∞.

By translation invariance of the Lebesgue measure, (Rn , |·| , m) satisfies UMD.

We say H satisfies the mean value property (MVP ), if for every r > 0 there exists Cr > 0

such that
Z
2 2
|hf, ka i| ≤ Cr |hf, kx i| dσ(x),
B(a,r)

for any f ∈ H and any a ∈ X.

28
Combining MVP with UMD, we have

Z
2 1 2
|hf, ka i| . |hf, kx i| dσ(x).
σ(B(a, r)) B(a,r)

It tell us that the averaging value of the function x → hf, kx i over a ball is greater or equal to

the value of the center. That’s the reason we borrow the terminology “mean value property” from

complex analysis here. Actually, the MVP usually holds when the Hilbert space is an analytic

function space.

We say {kx }x∈(X,d,σ) satisfies the approximate orthogonality property (AOP ), if the

followings hold:

1. there exists δ > 0 and c > 0 such that for any x, y ∈ X with d(x, y) < δ,

c ≤ |hkx , ky i| ;

2.

|hkx , ky i| → 0, as d(x, y) → ∞.

Notice AOP tell us that the angle between continuous frame kx and ky is small when x is

closed to y, and the angle is big when x is far from y.

The continuous frame {kx }x∈(X,d,σ) is said to satisfy the uniform localization property

(ULP ), if
Z
2
lim sup |hka , kx i| dσ(x) = 0.
r→∞ a∈X B(a,r)c

In the rest part of the thesis, for convenience we just say (H, {kx }x∈(X,d,σ) ) satisfies DMP,

ADP, UMD, MVP, AOP and ULP, instead of mentioning (X, d, σ), H or {kx }x∈(X,d,σ) respectively.

In the previous examples of FHS, we have the following properties hold:

1. The space (L2 (Rn ), {Mb Ta g}(a,b)∈(R2n ,|·|,m) ) satisfies DMP, ADP, UMD and AOP.

2. The classical Bargmann-Fock space (Fα , {kzα }z∈(Cn ,|·|,mα ) ) satisfies DMP, ADP, UMD, MVP,

AOP and ULP.

2πiξx
e
3. The general Paley-Wiener space (L2 (E), { √ }ξ∈(R,|·|,mE ) ) satisfies DMP, ADP, UMD and
m(E)

AOP.

29
4. The classical Paley-Wiener space (PWα , {kxα }x∈(R,|·|,mα ) ) satisfies DMP, ADP, UMD, MVP, AOP

and ULP.

2.4.3 Additional Properties

Proposition 2.4.3. Let (H, {kx }x∈(X,d,σ) ) be a FHS, which satisfies ADP, UMD and ULP. Then

Z Z
1 2
lim sup |hkx , ky i| dσ(y)dσ(x) = 0. (2.7)
r→∞ a∈X σ(B(a, r)) B(a,r) B(a,r)c

Proof. Let ρ > 0. We break the double integral above as follows

Z Z Z Z
+ .
B(a,r) B(a,r+ρ)c B(a,r) B(a,r+ρ)\B(a,r)

Estimate each term separately, and in both estimates we will divide by σ(B(a, r)) ( σ(B(a, r)) < ∞,

by UMD). About the first term, we have for any a ∈ X and r > 0,

Z Z
1 2
|hkx , ky i| dσ(y)dσ(x)
σ(B(a, r)) B(a,r) B(a,r+ρ)c
Z Z
1 2
≤ |hkx , ky i| dσ(y)dσ(x)
σ(B(a, r)) B(a,r) B(x,ρ)c
Z Z !
1 2
≤ sup |hkx , ky i| dσ(y) dσ(x)
σ(B(a, r)) B(a,r) x∈X B(x,ρ)c
Z
1 2
= σ(B(a, r)) sup |hkx , ky i| dσ(y)
σ(B(a, r)) x∈X B(x,ρ)c
Z
2
= sup |hkx , ky i| dσ(y)
x∈X B(x,ρ)c

For any ε > 0, by ULP we can find a positive ρ such that

Z
2
sup |hkx , ky i| dσ(y) < ε.
x∈X B(x,ρ)c

It follows that for such ρ

Z Z
1 2
lim sup sup |hkx , ky i| dσ(y)dσ(x)
r→∞ a∈X σ(B(a, r)) B(a,r) B(a,r+ρ)c
Z
2
≤ sup |hkx , ky i| dσ(y) < ε. (2.8)
x∈X B(x,ρ)c

30
Now we estimate the second term, using the positive ρ > 0 from above,

Z Z
1 2
|hkx , ky i| dσ(y)dσ(x)
σ(B(a, r)) B(a,r) B(a,r+ρ)\B(a,r)
Z Z
1 2
= |hkx , ky i| dσ(x)dσ(y)
σ(B(a, r)) B(a,r+ρ)\B(a,r) B(a,r)
Z Z
1 2
≤ |hkx , ky i| dσ(x)dσ(y)
σ(B(a, r)) B(a,r+ρ)\B(a,r) X
Z
1 2
= kky k dσ(y)
σ(B(a, r)) B(a,r+ρ)\B(a,r)
σ(B(a, r + ρ) \ B(a, r))
= .
σ(B(a, r))

By ADP, we obtain for such ρ

Z Z
1 2
lim sup sup |hkx , ky i| dσ(y)dσ(x)
r→∞ a∈X σ(B(a, r)) B(a,r) B(a,r+ρ)\B(a,r)
σ(B(a, r + ρ) \ B(a, r))
≤ lim sup sup = 0. (2.9)
r→∞ a∈X σ(B(a, r))

Combining (2.8) and (2.9), we obtain

Z Z
1 2
lim sup sup |hkx , ky i| dσ(y)dσ(x) < ε.
r→∞ a∈X σ(B(a, r)) B(a,r) B(a,r)c

Since ε is arbitrary, we get the desired equality.

Notice (2.7) is very similar with ULP, for some particular spaces, one implies another. Out

of the thesis, we might sometimes name (2.7) as ULP. As a direct consequence of Proposition 2.4.3,

we have the following corollary.

Corollary 2.4.4. Let (H, {kx }x∈(X,d,σ) ) be a FHS satisfying ADP, UMD and ULP, and TB(a,r) be

the concentration operator over the ball B(a, r). Denote all its eigenvalues in the decreasing order

by 1 ≥ l1 (TB(a,r) ) ≥ l2 (TB(a,r) ) ≥ · · · ≥ 0, where entries are repeated with multiplicity. Then for any

 > 0, there exists R > 0 such that


X ∞
X
(1 − ) li (TB(a,r) ) ≤ li (TB(a,r) )2 ,
i=1 i=1

for any a ∈ X and all r > R.

31
Proof. By UMD, we have σ(B(a, r)) < ∞ for any a ∈ X and all r > 0. So we can apply Corol-

lary 2.3.7 and obtain



X
li (TB(a,r) ) = TB(a,r) T r = σ(B(a, r)).
i=1

In order to prove the corollary, it is sufficient to show

∞ ∞
!
1 X X
2
lim sup sup li (TB(a,r) ) − li (TB(a,r) ) = 0.
r→∞ a∈X σ(B(a, r))
i=1 i=1

Again by Corollary 2.3.7


X ∞
X
li (TB(a,r) ) − li (TB(a,r) )2
i=1 i=1
2
= TB(a,r) T r − TB(a,r) HS
Z Z Z Z
2 2
= |hkx , ky i| dσ(y)dσ(x) − |hkx , ky i| dσ(y)dσ(x)
B(a,r) X B(a,r) B(a,r)
Z Z
2
= |hkx , ky i| dσ(y)dσ(x).
B(a,r) B(a,r)c

Then by Proposition 2.4.3, we get the desired equality.

Because ADP and UMD are common properties, we usually view Corollary 2.4.4 as a

consequence of ULP. Study Corollary 2.4.4, it tell us every eigenvalue li (TB(a,r) ) of TB(a,r) should

be closed to either 1 or 0, and the closeness depends on how large of r. It somehow tell us the

information about the distribution of eigenvalues of TB(a,r) when r is large. Based on this concern,

we have the following Lemma which improves the inequality in Lemma 2.3.9 when r is large.

Lemma 2.4.5. Given a number 0 < c < 1, then for every 0 < θ < 1, there exists R > 0 such that

for every ball B(a, r) ⊆ X with r > R,

σ(B(a, r))
dimG < ,
θ

where G denote any subspace of H which is c-concentrated on B(a, r).

Proof. Let TB(a,r) be the concentration operator over an arbitrary ball B(a, r). Denote all the eigen-

values of TB(a,r) by {li (TB(a,r) )}∞


i=1 indexed in decreasing order. And Let G be any c-concentrated

subspace of H on B(a, r) and denote dimG by k (k < ∞, by Lemma 2.3.9). By the min-max

32
principle,

D E
lk (TB(a,r) ) = max min TB(a,r+ δ ) g, g
dimF =k g∈F ,kgk=1 2

D E
≥ min TB(a,r+ δ ) g, g
g∈G,kgk=1 2

≥ c.

For any 0 < ε < 1 − c, we have

dimG = k

≤ # i : li (TB(a,r) ≥ c
 
= # i : li (TB(a,r) ) > 1 − ε + # i : c ≤ li (TB(a,r) ) ≤ 1 − ε
X li (TB(a,r) ) X li (TB(a,r) )
≤ +
1−ε c
li >1−ε c≤li ≤1−ε

1 X 1 X
≤ li (TB(a,r) ) + li (TB(a,r) ). (2.10)
1−ε c
i=1 li ≤1−ε

Let  = cε2 . By Corollary 2.4.4, there exists R > 0 such that for any a ∈ X and all r > R, we have


X
(1 − ) li (TB(a,r) )
i=1

X
≤ li (TB(a,r) )2
i=1
X X
= li (TB(a,r) )2 + li (TB(a,r) )2
li ≤1−ε li >1−ε
X X
≤(1 − ε) li (TB(a,r) ) + li (TB(a,r) )
li ≤1−ε li >1−ε

X X
= li (TB(a,r) ) − ε li (TB(a,r) ).
i=1 li ≤1−ε

It follows that for any a ∈ X and all r > R,


1 X X
li (TB(a,r) ) ≤ ε li (TB(a,r) ). (2.11)
c
li ≤1−ε i=1

33
Combining (2.10) and (2.11), we obtain that for any a ∈ X and all r > R,


1−ε X
dimG ≤ li (TB(a,r) ). (2.12)
1 + ε − ε2 i=1

Notice as ε → 0+ ,
1−ε
↑ 1.
1 + ε − ε2

So for any 0 < θ < 1, there exists ε > 0 such that

1−ε
> θ.
1 + ε − ε2

Then by (2.12), we have

P∞ P∞
i=1 li (TB(a,r) ) i=1 li (TB(a,r) ) σ(B(a, r))
dimG ≤ 1−ε < = ,
1+ε−ε2
θ θ

for any a ∈ X and all r > R. Where R only dependents on c and θ.

2.5 Boundedness, Compactness and Trace Class Member-

ship of Toeplitz Operators

In this section, based on the additional conditions we impose on FHS, we could give some

criteria to characterize the boundedness, compactness and trace class membership of Toeplitz oper-

ators without assuming µ(X) is finite. All of the results in this section are based on my master’s

thesis [11].

Definition 2.5.1. Let {kx }x∈(X,σ) be a continuous frame of H, and µ be another positive measure

for the same measurable sets with measure σ. We define the Berezin transform µ
e of measure µ

by
Z
2
e : X → R,
µ µ
e(x) := |hkx , ky i| dµ(y).
X
R
It is easy to check that if X
|hkx , ky i| dµ(y) < ∞ for every x ∈ X, then

e(x) = hTµ kx , kx i ,
µ ∀x ∈ X.

34
Therefore, the Berezin transform µ
e(x) of measure µ can be viewed as the “continuous diagonal

terms” of the matrix of the Toeplitz operator Tµ .

Definition 2.5.2. Let µ be a positive Borel measure on a metric space (X, d). For any r > 0, we

define the volume function µ


br of measure µ by

br : X → R,
µ µ
br (x) := µ(B(x, r)).

We now list the following theorems to characterize the boundedness, compactness and trace

class membership of Toeplitz operators Tµ .

Theorem 2.5.3 ([11], Theorem 5). Let (H, {kx }x∈(X,d,σ) ) be a FHS which satisfies DMP, UMD,

MVP and AOP. And Let µ be a σ-finite positive Borel measure on X that satisfies

Z
|hkx , ky i| dµ(y) < ∞, ∀x ∈ X.
X

Then the followings are equivalent.

a. Tµ : span{kx : x ∈ X} → H can be extended to a bounded operator on H.

b. The Berezin transform µ̃ : X → R is bounded.

c. The volume function µ̂r : X → R is bounded for some r > 0.

Theorem 2.5.4 ([11], Theorem 6). Let (H, {kx }x∈(X,d,σ) ) be a FHS which satisfies DMP, UMD,

MVP and AOP. And Let µ be a σ-finite positive Borel measure on X that satisfies

Z
|hkx , ky i| dµ(y) < ∞, ∀x ∈ X.
X

Then the followings are equivalent.

a. Tµ : H → H is a compact operator.

b. µ̃(x) → 0 as x → ∞.

c. µ̂r (x) → 0 as x → ∞ for some r > 0.

35
Theorem 2.5.5 ([11], Theorem 8). Given 1 ≤ p < ∞, Let (H, {kx }x∈(X,d,σ) ) be a FHS which

satisfies DMP, UMD, MVP and AOP. And Let µ be a σ-finite positive Borel measure on X that

satisfies
Z
|hkx , ky i| dµ(y) < ∞, ∀x ∈ X.
X

Then the followings are equivalent.

a. Tµ belongs to the Schatten class Sp .

b. µ̃ belongs to Lp (X, dσ).

c. µ̂r belongs to Lp (X, dσ) for some r > 0.

2.6 Beurling Densities of Sampling and Interpolation Sets

2.6.1 Sampling and Interpolation Sets

For RKHS, one could define the sampling and interpolation set. Similarly, we could give

the definitions of those for FHS.

Definition 2.6.1. Let (H, {kx }x∈(X,σ) ) be a FHS. A countable subset Λ := {λ} of X is called a

sampling set for H, if there exist constants A, B > 0 such that

2
X 2 2
A kf k ≤ |hf, kλ i| ≤ B kf k , ∀f ∈ H.
λ∈Λ

From the definition, Λ is a sampling set for H if and only if the corresponding sequence of

continuous frame {kλ }λ∈Λ is a frame of H.

Definition 2.6.2. Let (H, {kx }x∈(X,σ) ) be a FHS. A countable subset Λ := {λ} of X is called an

interpolation set for H, if for any complex sequence (cλ ) ∈ l2 (Λ), there exists f ∈ H such that

hf, kλ i = cλ , ∀λ ∈ Λ.

By Propositions 1.4.10, Λ is an interpolation set if and only if the corresponding sequence

of continuous frame {kλ }λ∈Λ is a Riesz-Fischer sequence of H.

36
2.6.2 Density Theorems in Particular Framed Hilbert Spaces

Now we are going to list some important results about sampling and interpolation sets which

will be applied in the rest of the chapters. All of them are related to the so-called Beurling densities.

Definition 2.6.3. Let Λ be a countable subset of a metric measure space (X, d, σ). The lower

Beurling density of Λ is defined as

#{Λ ∩ B(a, r)}


Dσ− (Λ) := lim inf inf .
r→∞ a∈X σ(B(a, r))

Similarly, the upper Beurling density of Λ is defined as

#{Λ ∩ B(a, r)}


Dσ+ (Λ) := lim sup sup .
r→∞ a∈X σ(B(a, r))

If σ is the Lebesgue measure, by convention, we will drop the subscript σ in the above notations. We

will use D− (Λ) and D+ (Λ) to denote the lower and upper Beurling densities in this case.

For example, all the integers Z is a countable subset of (R, |·| , m). It is not hard to see

D− (Z) = D+ (Z) = 1. If we pick all the natural numbers N as the countable subset, then D− (N) = 0

and D+ (N) = 1.

The following density theorem is based on the general Paley-Wiener space which is proved

by Landua in 1967.

Theorem 2.6.4 ([10]). Let E be a bounded measurable set on R, and Λ be a countable subset of R.
2πiξx
e
For the general Paley-Wiener space (L2 (E), { √ }ξ∈(R,|·|,mE ) ),
m(E)

1. if Λ is a sampling set for L2 (E), then


Dm E
(Λ) ≥ 1;

2. if Λ is an interpolation set for L2 (E), then

+
Dm E
(Λ) ≤ 1.

Note Theorem 2.6.4 gives us a necessary density condition for sampling and interpolation

sets in the general Paley-Wiener space. Besides this, Beurling and Kahane proved a sufficient density

37
condition for sampling and interpolation sets in the classical Paley-Wiener space by the following

theorem:

Theorem 2.6.5 ([8]). Let I be a finite interval of R, and Λ be countable subset of R. For the
2πiξx
e
classical Paley-Wiener space (L2 (I), { √ }ξ∈(R,|·|,mI ) ),
m(I)


1. if Λ is relatively separated with Dm I
(Λ) > 1, then Λ is a sampling set for L2 (I), i.e. there exist

constants A, B > 0 such that

X D E 2
2 ≤ B kf k2 2 ,
f, e2πiλ(·) ∀f ∈ L2 (I);

A kf kL2 (I) ≤ L (I)
L2 (I)

λ∈Λ

+
2. if Λ is separated with Dm I
(Λ) < 1, then Λ is an interpolation set for L2 (I), i.e. there exists a

constant C > 0 such that 2


X X
2 2πiλ(·)
C |cλ | ≤ cλ e ,



λ∈Λ λ∈Λ L2 (I)

for any finite sequence {cλ }λ∈Λ .

Similar results for the one-dimensional classical Bargmann-Fock space are proved by Seip

and Wallstén as the following two theorems:

Theorem 2.6.6 ([17, 18]). Let Λ be a countable subset in C. For the one-dimensional classical

Bargmann-Fock space (Fα (C), {kzα }z∈(C,|·|,mα ) ), Λ is a sampling set for Fα (C) if and only if Λ is

relatively separated and contains a separated subset Λ0 for which Dm


+
α
(Λ0 ) > 1.

Theorem 2.6.7 ([17, 18]). Let Λ be a countable subset in C. For the one-dimensional classical

Bargmann-Fock space (Fα (C), {kzα }z∈(C,|·|,mα ) ), Λ is an interpolation set for Fα (C) if and only if Λ
+
is separated and Dm α
(Λ) < 1.

For high dimensional Bargmann-Fock spaces, we don’t have the perfect if and only if density

conditions for sampling and interpolation sets. But still, Lindholm gives a very similar necessary

density condition for high dimensional weighted Bargmann-Fock spaces which implies the following

theorem:

Theorem 2.6.8 ([12]). Let Λ be a countable subset of Cn . For the classical Bargmann-Fock space

(Fα (Cn ), {kzα }z∈(Cn ,|·|,mα ) ),

38
1. if Λ is a sampling set for Fα (Cn ), then Λ is relatively separated and contains a separated subset

Λ0 which is also sampling and satisfies


Dm α
(Λ0 ) ≥ 1;

2. if Λ is an interpolation set for Fα (Cn ), then Λ is separated with

+
Dm α
(Λ) ≤ 1.

In general cases, we have the following theorem proved by Mitkovski and Ramirez which

gives a necessary density condition for the general FHS.

Theorem 2.6.9 ([13], Theorem 4.5). Let (H, {kx }x∈(X,d,σ) ) be a FHS which satisfies the ADP,

UMD, MVP and ULP, and Λ be a separated subset of X,

1. If Λ is a sampling set for H, then

Dσ+ (Λ) ≥ 1;

2. If Λ is an interpolation set for H, then

Dσ+ (Λ) ≤ 1.

39
Chapter 3

Approximate Interpolation in

Framed Hilbert Spaces

3.1 Introduction

3.1.1 (Weak) Interpolation Sets

Let (H, {kx }x∈(X,σ) ) be a FHS. A countable subset Λ := {λ} of X is called an interpolation

set for H, if the corresponding sequence of continuous frame {kλ }λ∈Λ is a Riesz-Fischer sequence of

H (see 2.6.2 and 1.4.10).

By Proposition 1.4.11, Λ is an interpolation set if and only if there exists a Bessel sequence

{fλ }λ∈Λ of H such that

hfλ , kν i = δλν , ∀λ, ν ∈ Λ.

If we replace the Bessell sequence condition by the condition of bounded sequence, we would

get a weak type of interpolation sets.

Definition 3.1.1. Let (H, {kx }x∈(X,σ) ) be a FHS. A countable subset Λ := {λ} of X is called a

weak interpolation set for H, if there exists a bounded sequence {fλ }λ∈Λ such that

hfλ , kν i = δλν , ∀λ, ν ∈ Λ.

40
Because any Bessel sequence is bounded, we have every interpolation set is a weak in-

terpolation set. And for particular FHS, the converse is also true. For example, in the classical

one-dimensional Bargmann-Fock space F(C), these two classes of sets coincide.

3.1.2 d-Approximate (Weak) Interpolation Sets

Recall Theorem 2.6.9: let (H, {kx }x∈(X,d,σ) ) be a FHS which satisfies the ADP, UMD, MVP

and ULP , if a separated set Λ (will be defined in next section) is an interpolation set for H, then

its upper Beurling density satisfies

Dσ+ (Λ) ≤ 1.

The goal of this chapter is to provide a similar type necessary density condition for an even

larger class of “interpolation sets”. This type of sets (to be defined momentarily) were relatively

recently introduced by Olevskii and Ulanovskii (see [15]) in the classical Paley-Wiener space, where

it was shown that all such sets have to satisfy a Beurling density condition similar to the one for

usual interpolation sets. Our results can be viewed as the generalization of their results.

Definition 3.1.2. Given 0 ≤ d < 1, we will say that a countable subset Λ := {λ} of X is a

d-approximate interpolation set for H, if there exists a Bessel sequence {hλ }λ∈Λ of H such that

X 2
|hhλ , kν i − δλν | ≤ d2 , ∀λ ∈ Λ.
ν∈Λ

Namely, the l2 (Λ) distance between the sequences (hhλ , kν i) and (δλν ) is no greater than d for any

λ.

Note that 0-approximate interpolation sets coincide with interpolation sets. Again, using a

bounded sequence instead of a Bessel sequence in Definition 3.1.2, we could define a weak type of

d-approximate interpolation sets.

Definition 3.1.3. Given 0 ≤ d < 1, we will say that a countable subset Λ = {λ} of X is a d-

approximate weak interpolation set for H, if there exists a bounded sequence {fλ }λ∈Λ in H

such that
X 2
|hfλ , kν i − δλν | ≤ d2 , ∀λ ∈ Λ.
ν∈Λ

Again, 0-approximate weak interpolation sets coincide with weak interpolation sets in H.

41
And every d-approximate interpolation set is a d-approximate weak interpolation set. But we don’t

know whether the converse is true in general.

3.2 d-Approximate Interpolation in Framed Hilbert Spaces

3.2.1 Main Theorem

Our first result is based on FHS which gives a necessary density condition for d-approximate

interpolation sets.

Theorem 3.2.1. Let (H, {kx }x∈(X,d,σ) ) be a FHS which satisfy ADP, UMD, MVP and ULP. Given

0 ≤ d < 1, suppose a relatively separated subset Λ of X is a d-approximate interpolation set for H.

Then
1
Dσ+ (Λ) ≤ .
1 − d2

Before we give a proof of the theorem, we need to collect some basics.

3.2.2 Relatively Separated Sets

Definition 3.2.2. A subset Λ in a metric space (X, d) is said to be separated or uniformly discrete

if

δ := inf{d(λ, ν) : λ 6= ν ∈ Λ} > 0.

And the constant δ > 0 is called a separation constant of Λ.

Definition 3.2.3. A subset Λ in a metric space (X, d) is said to be relatively separated if it is a

finite union of separated sets. Namely,

Λ = ∪K
k=1 Λk ,

where Λk is a separated set for any k = 1, · · · , K.

The following two propositions will show some simple but important properties for Λ being

relatively separated.

42
Proposition 3.2.4. Let (X, d, σ) be a metric measure space satisfying UMD, and Λ be a relatively

separated set in X. Then for any ball B(a, r) in X with center a ∈ X and radius r > 0,

#{Λ ∩ B(a, r)} < ∞.

If (X, d, σ) also has ADP, then

Dσ+ (Λ) < ∞.

Proof. Since Λ is relatively separated, we have

Λ = ∪K
k=1 Λk ,

where Λk is a separated set with the separation constant δk for any k = 1, · · · , K. Let δ :=

min1≤k≤K δk , and {λki }i∈Ik := {Λk ∩ B(a, r)} for any k = 1, · · · , K. For every 1 ≤ k ≤ K, by Λk is

separated,
δ δ
B(λki , ) ∩ B(λkj , ) = ∅, ∀i 6= j.
2 2

In addition, for every 1 ≤ k ≤ K,

δ δ
∪i∈Ik B(λki , ) ⊆ B(a, r + ).
2 2

It follows that for every 1 ≤ k ≤ K,

X δ δ
σ(B(λki , )) ≤ σ(B(a, r + )).
2 2
i∈Ik

By UMD, there exists c( 2δ ) > 0 such that

δ X δ δ
#{Λk ∩ B(a, r)}c( ) ≤ σ(B(λki , )) ≤ σ(B(a, r + )).
2 2 2
i∈Ik

Again by UMD, there exists D(r + 2δ ) > 0 such that

K K
X X σ(B(a, r + 2δ )) mD(r + 2δ )
#{Λ ∩ B(a, r)} ≤ #{Λk ∩ B(a, r)} ≤ ≤ < ∞.
k=1 k=1
c( 2δ ) c( 2δ )

43
Combine ADP,

#{Λ ∩ B(a, r)}


Dσ+ (Λ) : = lim sup sup
r→∞ a∈X σ(B(a, r))
PK
#{Λk ∩ B(a, r)}
≤ lim sup sup k=1
r→∞ a∈X σ(B(a, r))
K
X σ(B(a, r + 2δ ))
≤ lim sup sup
r→∞ a∈X
k=1
c( 2δ )σ(B(a, r))
K σ(B(a, r + 2δ ))
= δ
lim sup sup
c( 2 ) r→∞ a∈X σ(B(a, r))
K
= δ < ∞. (3.1)
c( 2 )

Actually, Theorem 3.2.1 tell us under additional interpolation assumptions on Λ, this trivial

density upper bound (3.1) could be significantly improved (especially when δ is close to 0).

The second consequence of Λ being relatively separated is that it will generate a Bessel

sequence of continuous frames {kλ }λ∈Λ of H.

Proposition 3.2.5. Let (H, {kx }x∈(X,d,σ) ) be a FHS satisfying ADP, UMD, MVP and ULP. If Λ

is a relatively separated set in X, then there exists a constant C > 0 such that

X 2 2
|hf, kλ i| ≤ C kf k , ∀f ∈ H.
λ∈Λ

Proof. Since Λ is relatively separated, we have

Λ = ∪K
k=1 Λk ,

where Λk is a separated set with the separation constant δk for any k = 1, · · · , K. Let δ :=

min1≤k≤K δk , and {λki }i∈Ik := {Λk ∩ B(a, r)} for any k = 1, · · · , K. Then by MVP, there exists

44
Cδ > 0 such that

X K X
X
2 2
|hf, kλ i| ≤ |hf, kλ i|
λ∈Λ k=1 λ∈Λk
K X
X Z
2
≤ Cδ |hf, kx i| dσ(x)
k=1 λ∈Λk B(λ, δ2 )

K X Z
X 2
= Cδ |hf, kx i| dσ(x)
k=1 λ∈Λk B(λ, δ2 )

K Z
X 2
≤ Cδ |hf, kx i| dσ(x)
k=1 X

2
= KCδ kf k .

3.2.3 Some Lemmas

In order to prove Theorem 3.2.1, we will adopt the proof strategy of Olevskii and Ulanovskii [15].

The argument has two crucial ingredients. The first one (essentially going back to Landau [10]) says

that any subspace of H which is c-concentrated on a fixed finite measure set cannot have arbitrar-

ily large dimension. We have these results as Lemma 2.3.9 and Lemma 2.4.5 in Chapter 2, for

convenience we restate them here.

Lemma. 2.3.9 Let E be a Borel subset of X with σ(E) < ∞. Given a number 0 < c < 1, if G is a

subspace of H which is c-concentrated on E, then

σ(E)
dimG ≤ .
c

Lemma. 2.4.5 Given a number 0 < c < 1, then for every 0 < θ < 1, there exists R > 0 such that

for every ball B(a, r) ⊆ X with r > R,

σ(B(a, r))
dimG < ,
θ

where G denote any subspace of H which is c-concentrated on B(a, r).

The following lemma is the second important ingredient in our proofs. It allows us to

45
generate a fairly high-dimensional subspace which is concentrated on some ball so that we can apply

the above two lemmas. Note the following result is a finite-dimensional result, which can be applied

in our proofs only when we restrict the set Λ to a ball.

Lemma 3.2.6 ([15], Lemma 2). Let {uj }1≤j≤N be an orthonormal basis in an N-dimensional

complex Euclidean space U. Given 0 < d < 1, suppose that {vj }1≤j≤N is a set of vectors in U

satisfying

kvj − uj k2 ≤ d2 , 1 ≤ j ≤ N.

Then for any b with 1 < b < 1/d, there is a subspace X of CN , such that

1. (1 − b2 d2 )N − 1 < dimX;

2. the estimate
N N
1 X X
(1 − )2 |cj |2 ≤ k cj vj k2 ,
b j=1 j=1

holds for any vector c = (c1 , c2 , ..., cN ) ∈ X.

3.2.4 Proof of Theorem 3.2.1

Proof. By Λ = {λ} ⊆ X is a d-approximate interpolation set for H, there exists a Bessel sequence

{hλ }λ∈Λ for H such that


X 2
|hhλ , kν i − δλν | ≤ d2 , ∀λ ∈ Λ.
ν∈Λ

Let B(a, r) be an arbitrary open ball in X. Since Λ is relatively separated, Λ ∩ B(a, r) is finite set.

Let Λ ∩ B(a, r) = {λ1 , λ2 , ..., λN }. Consider the following vectors in CN ,




vj := ( hλj , kλ1 , · · · , hλj , kλN ), 1 ≤ j ≤ N,

and the standard basis of CN ,

uj := (δλj λ1 , · · · , δλj λN ), 1 ≤ j ≤ N.

46
Notice that

N
hλj , kλi − δλj λi 2
2
X

kvj − uj k =
i=1

hλ , kν − δλ ν 2
X

≤ j j
ν∈Λ

≤ d2 , 1 ≤ j ≤ N.

By Lemma 3.2.6, for any 1 < b < 1/d, there exists a subspace U of CN , such that

1.

(1 − b2 d2 )N − 1 < dimU ; (3.2)

2. the inequality
N N
1 X X
(1 − )2 |cj |2 ≤ k cj vj k2 , (3.3)
b j=1 j=1

holds for any vector c = (c1 , ..., cN ) ∈ U.

PN
Let Y := { j=1 cj vj |c = (c1 , ..., cN ) ∈ U } be the subspace of CN . we want to show dimU ≤ dimY .

The idea is that using a basis of U to create some linearly independent vectors of Y .

Let M = dimU and let {ci := (ci1 , ..., ciN )}M N


i=1 be a basis of U ⊆ C . And let {wi :=
PN i M
j=1 cj vj }i=1
be the vectors of Y . Assume a1 , a2 , · · · , aM are the scalars such that the linear
PM
combination i=1 ai wi = 0, then

M
X
0= a i wi
i=1
M
X N
X
= ai cij vj
i=1 j=1
N M
!
X X
= ai cij vj
j=1 i=1

47
P 
M PM
Notice the scalars of coefficients i=1 ai cij = i=1 ai ci ∈ U , by (3.3) we have
1≤j≤N

2
!
N X M
X
ai cij vj

0=
j=1 i=1
N
M 2
1 2 X X i
≥ (1 − ) ai cj ,
b j=1 i=1

PM
which implies the linear combination i=1 ai ci = 0. By {ci }M
i=1 is a basis, ai = 0, 1 ≤ i ≤ M. So

{wi }M
i=1 are the linearly independent vectors of Y . Then

dimU ≤ dimY. (3.4)

PN
Let G := { j=1 cj hλj |c = (c1 , ..., cN ) ∈ U } be the subspace of H, now we want to prove dimY ≤

dimG.

Let T ∗ be the analysis operator of the finite sequence {kλ1 , kλ2 , ..., kλN } on H. By T ∗ hλj =

vj , 1 ≤ j ≤ N , we have T ∗ (G) = Y . It follows that

dimY ≤ dimG. (3.5)

Combine (3.2), (3.4) and (3.5), we obtain

(1 − b2 d2 )N − 1 < dimG. (3.6)

PN
Let g = j=1 cj hλj ∈ G for some (c1 , ..., cN ) ∈ U . Using that {hλ }λ∈Λ is a Bessel sequence and

48
(3.3), we get

N
*
N X N
+ 2
X 2
X
|hg, kλi i| =
cj hλj , kλi
i=1 i=1 j=1
N
X
=k cj vj k2
j=1
N
1 X
≥ (1 − )2 |cj |2
b j=1
N
1 X
≥ (1 − )2 Ck cj hλj k2
b j=1

2
= C1 kgk , (3.7)

for any g ∈ G.

Let δ be the separation constant of Λ. Then B(λ, δ/2) ∩ B(ν, δ/2) = ∅ for any λ 6= ν ∈ Λ.

It follows from the MVP that

N
X N
X Z
2 2
|hg, kλi i| ≤ Cδ |hg, kx i| dσ(x)
i=1 i=1 B(λi , δ2 )
Z
2
≤ Cδ |hg, kx i| dσ(x), (3.8)
B(a,r+ δ2 )

for any g ∈ G. Combining (3.7) and (3.8), we obtain

Z D E
2 2
c kgk ≤ |hg, kx i| dσ(x) = TB(a,r+ δ ) g, g , (3.9)
2
B(a,r+ δ2 )

for any g ∈ G, where 0 < c := C1 /Cδ < 1 is independent of a and r. It follows that G is a

c-concentrated subspace of H on B(a, r + 2δ ).

For every 0 < θ < 1, apply Lemma 2.4.5, there exists R > 0 such that for any ball B(a, r+ 2δ )

with r > R,
σ(B(a, r + 2δ ))
dimG < . (3.10)
θ

Combine (3.6) and (3.10), we have for every a ∈ X and r > R,

σ(B(a, r + 2δ ))
(1 − b2 d2 )#{Λ ∩ B(a, r)} − 1 < .
θ

49
Then for every a ∈ X and r > R,

σ(B(a, r + 2δ )) 1
#{Λ ∩ B(a, r)} < + .
θ(1 − b2 d2 ) (1 − b2 d2 )

Finally, using ADP and UMD, we obtain

#{Λ ∩ B(a, r)}


Dσ+ (Λ) = lim sup sup
r→∞ a∈X σ(B(a, r))
!
σ(B(a, r + 2δ )) 1
≤ lim sup sup +
r→∞ a∈X θ(1 − b2 d2 )σ(B(a, r)) (1 − b2 d2 )σ(B(a, r))
σ(B(a, r + 2δ )) 1
≤ lim sup sup 2 2
+ lim sup sup 2 2
r→∞ a∈X θ(1 − b d )σ(B(a, r)) r→∞ a∈X (1 − b d )σ(B(a, r))

σ(B(a, r + 2δ ))
= lim sup sup
r→∞ a∈X θ(1 − b2 d2 )σ(B(a, r))
1 σ(B(a, r + 2δ ))
= lim sup sup
θ(1 − b2 d2 ) r→∞ a∈X σ(B(a, r))
1
= .
θ(1 − b2 d2 )

Since θ < 1 and b > 1 are arbitrary,


1
D+ (Λ) ≤ .
1 − d2

3.3 d-Approximate Weak Interpolation in Classical Bargmann-

Fock Spaces

Our next result is based on the classical Bargmann-Fock space (F(Cn ), |·| , m). In the

classical one-dimensional Bargmann-Fock space F(C), it was shown by Seip and Schuster [16] that

the class of weak interpolation sets coincides with the class of interpolation sets.

Another important result (Theorem 2.6.7) of Seip and Wallstén shows that, in the classical

one-dimensional Bargmann-Fock space F(C), interpolation sets Λ can be completely characterized

in terms of the upper Beurling density D+ (Λ). Namely, Λ is an interpolation set (or equivalently

weak interpolation set) if and only if Λ is separated, and D+ (Λ) < 1.

50
3.3.1 Main Theorem

For the classical Bargmann-Fock space (F(Cn ), |·| , m), as a special FHS, we could improve

Theorem 3.2.1 by applying a weaker condition that Λ is a d-approximate weak interpolation set.

Theorem 3.3.1. Let (F(Cn ), |·| , m) be the classical Bargmann-Fock space. Given 0 ≤ d < 1,

suppose a relatively separated subset Λ of Cn is a d-approximate weak interpolation set for F(Cn ).

Then
1
D+ (Λ) ≤ .
1 − d2

This result can be easily extended to all of the classical Bargmann-Fock spaces Fα (Cn ).

3.3.2 Basics in Classical Bargmann-Fock Spaces

Recall in Chapter 2, the classical Bargmann-Fock space (Fα (Cn ), {kz }z∈(Cn ,|·|,mα ) ) is a FHS

satisfying DMP, ADP, UMD, MVP, AOP and ULP. In addition, Fα (Cn ) is a RKHS. Its reproducing

kernel at point z ∈ Cn equals

Kzα (w) = eαhw,zi , ∀w ∈ Cn ,

and its normalized reproducing kernel at point z ∈ Cn is

α 2
kzα (w) = eαhw,zi− 2 |z| , ∀w ∈ Cn .

About the reproducing kernels, we have the following important equalities hold:

2 2
kKzα kα = hKzα , Kzα iα = eα|z| , ∀z ∈ Cn ,

α 2
|hkzα , kw
α
iα | = e− 2 |z−w| , ∀z, w ∈ Cn .

3.3.3 Proof of Theorem 3.3.1

Proof. By Λ = {λ} ⊆ Cn is a d-approximate weak interpolation set for F(Cn ), there exists a

bounded sequence {fλ }λ∈Λ such that

X 2
|hfλ , kν i − δλν | ≤ d2 , ∀λ ∈ Λ.
ν∈Λ

51
Let ε > 0. For any λ ∈ Λ define gλ (z) := fλ (z)kλε (z). Clearly gλ : Cn → C is entire as a product of

two entire functions. Also, since

2 2
|kλε (z)| = |hkλε , Kzε i|
2 2
≤ kkλε kε kKzε kε
2
= eε|z| ,

for all z, λ ∈ Cn , we have

Z
2
|gλ (z)|2 e−(π+ε)|z| dm(z)
Cn
Z
2
= |fλ (z)kλε (z)|2 e−(π+ε)|z| dm(z)
n
ZC
2
≤ |fλ (z)|2 e−π|z| dm(z) < ∞.
Cn

Therefore, gλ ∈ Fπ+ε (Cn ) for all λ ∈ Λ. Moreover,

X
2
gλ , kνπ+ε π+ε − δλν

ν∈Λ
X
−1 2
gλ , Kνπ+ε π+ε Kνπ+ε π+ε − δλν

=

ν∈Λ
X −1 2
gλ (ν) Kνπ+ε π+ε − δλν

=

ν∈Λ
X −1 2
fλ (ν)kλε (ν) Kνπ+ε π+ε − δλν

=

ν∈Λ
X −1 2
hfλ , Kν i hkλε , Kνε iε Kνπ+ε π+ε − δλν

=

ν∈Λ
X −1 2
hfλ , kν i kKν k hkλε , kνε iε kKνε kε Kνπ+ε π+ε − δλν

=

ν∈Λ
X 2
= |hfλ , kν i hkλε , kνε iε − δλν |
ν∈Λ
X 2
≤ |hfλ , kν i − δλν | ≤ d2 , (3.11)
ν∈Λ

for any λ ∈ Λ. Note that in this simple computation we used that

π+ε

π+ε
= kKν k kKνε kε .

52
Let B(a, r) be any ball in Cn . Since Λ is relatively separated, Λ ∩ B(a, r) is a finite set. Let

Λ ∩ B(a, r) = {λ1 , λ2 , ..., λN }. Consider the following vectors in CN

vj := ( gλj , kλπ+ε , ..., gλj , kλπ+ε





1 π+ε N π+ε
), 1 ≤ j ≤ N,

and the standard basis

uj := (δλj λ1 , ..., δλj λN ), 1 ≤ j ≤ N.

By the above inequality (3.11), we have

N
X 2
2
gλj , kλπ+ε


kvj − uj k = − δ

i π+ε λj λi
i=1
X
2
gλj , kνπ+ε π+ε − δλj ν



ν∈Λ

≤ d2 , 1 ≤ j ≤ N.

By Lemma 3.2.6, for any 1 < b < 1/d, there exists a subspace U of CN , such that

1. (1 − b2 d2 )N − 1 < dimU ;

2. the inequality
N N
1 X X
(1 − )2 |cj |2 ≤ k cj vj k2 , (3.12)
b j=1 j=1

holds for any vector c = (c1 , ..., cN ) ∈ U.

PN
Let G := { j=1 cj gλj |c = (c1 , ..., cN ) ∈ U }. By the same argument of (3.6) in the proof of

Theorem 3.2.1,

(1 − b2 d2 )N − 1 ≤ dimG. (3.13)

53
PN
Let g = j=1 cj gλj ∈ G for some (c1 , ..., cN ) ∈ U . Using that {kλπ+ε }λ∈Λ is a Bessel sequence (due

to the relative separateness of Λ) and (3.12), we have

X
2
2
g, kλπ+ε π+ε

kgkπ+ε ≥ C

λ∈Λ
* + 2
N X
N
X
π+ε

≥C
cj gλj , kλi

i=1 j=1 π+ε

2
N X
N

X
π+ε


≥C
cj gλj , kλi π+ε
i=1 j=1
2
N
X

=C cj vj

j=1
N
X
≥ C(1 − 1/b)2 |cj |2
j=1
N
X
= C1 |cj |2 , (3.14)
j=1

for any g ∈ G, where C1 is independent of the radius r.

Fix a small ρ > 0. A simple application of the Cauchy-Schwarz inequality and the already

mentioned identity kKzπ+ε kπ+ε = kKz k kKzε kε yields

Z
2 2 π
|g(z)| e−(π+ε)|z| −nlog( π+ε )
dm(z)
B(a,r+ρr)c
Z
2 2
= |g(z)| e−(π+ε)|z| dmπ+ε (z)
B(a,r+ρr)c
2
N
Z X
ε
−(π+ε)|z|2
=
c f
j λj (z)kλj (z) e
dmπ+ε (z)
B(a,r+ρr)c j=1
N Z N 2
X 2
X 2
≤ |cj | fλj (z)kλε j (z) e−(π+ε)|z| dmπ+ε (z)

j=1 B(a,r+ρr)c j=1

N Z N E 2
X 2
X D −2
fλj , Kz kλε j , Kzε Kzπ+ε π+ε dmπ+ε (z)

= |cj |

B(a,r+ρr)c j=1 ε
j=1
N Z N E 2
X 2
X D
fλj , kz kλε j , kzε dmπ+ε (z),

= |cj | (3.15)

B(a,r+ρr)c j=1 ε
j=1

for any g ∈ G. We now estimate the integral term in (3.15). Applying {fλ }λ∈Λ is bounded and

54
doing a simple change of variables, we obtain

Z N E 2
X D
fλj , kz kλε j , kzε dmπ+ε (z)


B(a,r+ρr)c j=1 ε

Z N
X D E 2
fλj kkz k kλε j , kzε dmπ+ε (z)



B(a,r+ρr)c j=1 ε

Z N D
ε ε 2
X E
≤C kλj , kz dmπ+ε (z)
B(λj ,ρr)c j=1 ε

N Z
ε ε 2
X D E
=C kλj , kz dmπ+ε (z)
B(λj ,ρr)c ε
j=1
N Z
X 2
=C e−ε|z−λj | dmπ+ε (z)
j=1 B(λj ,ρr)c
Z
2
=CN e−ε|z| dmπ+ε (z). (3.16)
B(0,ρr)c

Since Λ is relatively separated, we have

Λ = ∪K
k=1 Λk ,

where Λk is a separated set with the separation constant δk for any k = 1, · · · , K. Let δ :=

min1≤k≤K δk , and {λki }i∈Ik := {Λk ∩ B(a, r)} for any k = 1, · · · , K. The simple counting argument

shows, for r > 1

N = #{Λ ∩ B(a, r)}


K
X
≤ #{Λk ∩ B(a, r)}
k=1
K
X m(B(a, r + 2δ ))

k=1
m(B(0, 2δ ))
n
π
(r + δ )2n
= K n!πn δ 22n
n! ( 2 )
1 2
= K( + )2n r2n
r δ
2 2n 2n
< K(1 + ) r .
δ

55
And doing change of variables by surface coordinates, we get (3.16) is bounded by

Z
2
CN e−ε|z| dmπ+ε (z)
B(0,ρr)c
Z
2 2n 2n 2
<CK(1 + ) r e−ε|z| dmπ+ε (z)
δ B(0,ρr)c
Z ∞
2
=C 0 r2n e−εt t2n−1 dt
ρr

=C2 (r), (3.17)

where C 0 does not depend on r (which depends on n, δ and K). Denote the last expression by C2 (r).

Observe that C2 (r) → 0 as r → ∞ (to be used in a moment). Combine (3.15), (3.16) and (3.17), we

obtain for every g ∈ G

Z N
2 2 π X 2
|g(z)| e−(π+ε)|z| −nlog( π+ε )
dm(z) ≤ C2 (r) |cj | . (3.18)
B(a,r+ρr)c j=1

Combining (4.2.4) and (4.2.5), we have for every g ∈ G

Z
C2 (r) 2 2 2 π
(1 − ) kgkπ+ε ≤ |g(z)| e−(π+ε)|z| −nlog( π+ε )
dm(z)
C1 B(a,r+ρr)
Z
2
g, kzπ+ε π+ε dmπ+ε (z)

=

B(a,r+ρr)


= TB(a,r+ρr) g, g π+ε .

Let 0 <  < 1. Since C2 (r)/C1 → 0 as r → ∞, there exists R > 0, such that

2

(1 − ) kgkπ+ε ≤ TB(a,r+ρr) g, g π+ε ,

for every g ∈ G when r > R. In other words, the subspace G is (1 − )-concentrated on B(a, r + ρr)

whenever r > R. By Lemma 2.3.9, we obtain

mπ+ε (B(a, r + ρr))


dimG ≤ . (3.19)
1−

56
Combining (3.13) and (3.19), we obtain for every a ∈ Cn and r > R,

mπ+ε (B(a, r + ρr))


(1 − b2 d2 )N − 1 < .
1−

Then for every a ∈ Cn and r > R,

mπ+ε (B(a, r + ρr)) 1


#{Λ ∩ B(a, r)} < 2 2
+ .
(1 − )(1 − b d ) (1 − b2 d2 )

Therefore,

#{Λ ∩ B(a, r)}


D+ (Λ) = lim sup sup
r→∞ a∈Cn m(B(a, r))
 
mπ+ε (B(a, r + ρr)) 1
≤ lim sup sup +
r→∞ a∈Cn (1 − )(1 − b2 d2 )m(B(a, r)) (1 − b2 d2 )m(B(a, r))
(π + ε)n (1 + ρ)2n
= .
π n (1 − )(1 − b2 d2 )

Since ε > 0, ρ > 0,  > 0, b > 1 are arbitrary,

1
D+ (Λ) ≤ .
1 − d2

57
Chapter 4

Frame Bound Estimates for

Continuous Frames of Exponentials

4.1 Introduction

Recall Beurling and Kahane’s density theorem for sampling sets in the classical Paley-Wiener

space.

Theorem. 2.6.5 Let I be a finite interval of R, and Λ be countable subset of R. For the classical
2πix(·)

Paley-Wiener space (L2 (I), { e√ }x∈(R,|·|,mI ) ), if Λ is relatively separated with Dm I
(Λ) > 1, then
m(I)

Λ is a sampling set for L2 (I). Namely, there exist constants A, B > 0 such that

X D E 2
2 ≤ B kf k2 2 ,
f, e2πiλ(·) ∀f ∈ L2 (I).

A kf kL2 (I) ≤ 2 L (I)
L (I)

λ∈Λ

The following theorem could be viewed as an analogue for “continuous sampling sets” in

the classical Paley-Wiener space.

Theorem 4.1.1 ([9], Theorem 1 Logvinenko-Sereda). Let E be an interval with the length b and let

S be a measurable subset of R. If there exist a > 0 and γ > 0 such that |S ∩ I| := m(S ∩ I) ≥ γa

for every interval I with length a, then

 −C(ab+1) Z
C b 2 b 2

f 2 ≤ f (x) dx, (4.1)
γ L (R) S

58
for any function f ∈ L2 (E).

Note that the above inequality implies

 −C(ab+1) Z D 2
C 2 f, e2πix(·)
E
dx ≤ kf k2 2

kf kL2 (E) ≤ L (E) .
γ
S L2 (E)

2πix(·)
e
It shows us the exponentials { √ }x∈(S,mE ) forms a continuous frame of L2 (E) with lower frame
m(E)
 −C(ab+1)
bound Cγ and upper frame bound 1. By this observation, Theorem 4.1.1 tell us that

when the subset S is pretty “dense” in R, S could be viewed as a “continuous sampling set” of the
2πix(·)
e
classical Paley-Wiener space (L2 (E), { √ }x∈(R,|·|,mE ) ).
m(E)

In 1973, Logvinenko and Sereda first proved inequality (4.1) for some constant of lower

frame bound. Then Kovrijkine improved it by giving a formula for that lower frame bound. In

addition, Kovrijkine developed and generalized Theorem 4.1.1 even for multiband-limited functions.

Theorem 4.1.2 ([9], Theorem 2 Kovrijkine). Let E := ∪ni=1 Ii be the union of n intervals with the

same length b. And let S be a measurable subset of R. If there exist a > 0 and γ > 0 such that

|S ∩ I| ≥ γa for every interval I of length a, then

 −2ab( Cγ )n −2n+1 Z
C b 2 b 2

f ≤ f (x) dx, (4.2)
γ L2 (R) S

for any function f ∈ L2 (E).

Notice the inverses of lower frame bound appearing in (4.1) and (4.2) both grow exponen-

tially with the length of E. The goal of this chapter is to provide a better lower frame bound whose

inverse would grow linearly with the length of E under certain conditions.

4.2 Lower Frame Bound Estimates for Continuous Frames of

Exponentials

4.2.1 Main Theorems

In order to state our theorems in an easier way, it is better to impose the conditions on set

Σ, which plays the role of the complement of set S in Theorem 4.1.1 and Theorem 4.1.2. In those

59
theorems, we impose conditions to make set S “dense” in R. For our theorems, we would like to

make set Σ kind of “sparse”.

When Σ is a countable union of bounded intervals, we find two different ways to make set

Σ “sparse”. One way is to require that the length of intervals and the density of centers of intervals

are small; another way is to require the density of Σ itself is small. Based on these, we formulate

the following two theorems.

Theorem 4.2.1. Let E := ∪ni=1 Ii be the union of n bounded intervals. And let Λ := {λk }k∈Z be

a sequence, and Σ := ∪k∈Z (λk − 2b , λk + 2b ) be the countable union of intervals with center λk and
1
length b. If [b]D+ (Λ) < n, then there exists a constant C such that

2 Z
C fb 2 ≤ |fb(x)|2 dx,

L (R) Σc

for any function f ∈ L2 (E). And C −1 grows linearly with |E| when |E| is large. Note: [b] denote

the biggest integer less or equal to b.

Theorem 4.2.2. Let E := ∪ni=1 Ii be the union of n bounded intervals. And let Σ = ∪k∈Z (λk −
bk bk
2 , λk + 2 ) be the countable disjoint union of intervals with the center λk and the length bk . If
1
inf k {bk } ≥ b > 0 for some constant b, and D+ (Σ) < n, then there exists a constant C such that

2 Z
C fb ≤ |fb(x)|2 dx,

L2 (R) Σc

for any function f ∈ L2 (E). And C −1 grows linearly with |E| when |E| is large. Note: D+ (Σ) =
|Σ∩[a−r,a+r]|
lim supr→∞ supa∈R 2r as usual.

4.2.2 Random Periodization

In order to prove Theorem 4.2.1 and Theorem 4.2.2, we will adopt the proof strategy of

Nazarov in [14], where the random periodization was introduced.

Definition 4.2.3. Given any function f ∈ L1 (R) and any positive number ε. We define the random

periodization g εν of f by
1 X k+t
g εν (t) := √ f( ), t ∈ [0, 1)
εν εν
k∈Z

where ν is a random variable equidistributed on the interval (1, 2).

60
The series in the definition of g εν converges in L1 [0, 1), and for every ν this series represents

a 1-periodic function on R. The random periodization g εν allows some useful properties which will

be stated in the following two propositions.

Proposition 4.2.4. Given any function f ∈ L1 (R), let g εν be the random periodization of f . Then

the Fourier coefficients of g εν satisfy

εν (k) =

gc εν fb(ενk),

for any k ∈ Z.

Proof. By Fubini’s theorem, we have

Z 1
εν (k) =
gc g εν (t)e−2πikt dt
0
Z 1
1 X n + t −2πikt
= √ f( )e dt
0 εν εν
n∈Z
XZ 1 1 n + t −2πikt
= √ f( )e dt
εν εν
n∈Z 0
X Z n+1 1 t
= √ f ( )e−2πikt dt
εν εν
n∈Z n
Z ∞
1 t
= √ f ( )e−2πikt dt
−∞ εν εν

= εν fb(ενk).

Let Σ be a measurable subset of R and α be a positive number. Define the following subset

of Z

Λ(α, Σ) := {k ∈ Z : kα ∈ Σ},

then we have the following property:

Proposition 4.2.5 ([14], Proposition 2.2). Let g εν be the random periodization of the function f ,

and Λ := Λ(εν, Σ) be the subset of Z. Then

X Z
E εν (k)|2 ≤ 2
|gc |fb(x)|2 dx.
06=k∈Λ
/ Σc

61
Note: E denote the expectation of random variable.

To prove Proposition 4.2.5, we need the random lattice averaging lemma:

Lemma 4.2.6 ([14], Lemma 2.1). Let ϕ : R → [0, ∞) be a positive function, and let ε > 0 be a fixed

number. Then
Z 2 Z
X 1
ϕ(kεν)dν ≤ ϕ(x)dx.
1 k6=0 ε R

Proof. Split the left hand side into two terms,

Z 2 X Z 2 X Z 2 X
ϕ(kεν)dν = ϕ(kεν)dν + ϕ(kεν)dν.
1 k6=0 1 k>0 1 k<0

Estimating the first term, by Tonelli’s theorem we have

Z 2 X XZ 2
ϕ(kεν)dν = ϕ(kεν)dν
1 k>0 k>0 1

X 1 Z 2εk
= ϕ(x)dx (x = kεν)
εk εk
k>0

1 X 2εk ϕ(x)
Z
= dx
ε k
k>0 εk
1 ∞ X ϕ(x)
Z
= dx
ε ε x x k
2ε <k< ε

1 ∞
Z X 1
= ϕ(x) dx
ε 0 x x k
2ε <k< ε

1 ∞
Z
≤ ϕ(x)dx, (4.3)
ε 0

1
P
where the last inequality comes from x x ≤ 1 for any x > 0. Similarly, for the second term,
2ε <k< ε k

we have
Z 2 Z 0
X 1
ϕ(kεν)dν ≤ ϕ(x)dx. (4.4)
1 k<0 ε −∞

Adding (4.3) and (4.4) together, we obtain

Z 2 Z
X 1
ϕ(kεν)dν ≤ ϕ(x)dx.
1 k6=0 ε R

62
Proof of Proposition 4.2.5. By Proposition 4.2.4 and Lemma 4.2.6, we have

X X
E εν (k)|2 = E
|gc εν|fb(kεν)|2
06=k∈Λ
/ 06=k∈Λ
/
X
=E εν|fb(kεν)|2 1Σc (kεν)
k6=0
X
≤ 2εE |fb(kεν)|2 1Σc (kεν)
k6=0
Z 2 X
= 2ε |fb(kεν)|2 1Σc (kεν)dν
1 k6=0
Z
≤2 |fb(x)|2 dx.
Σc

4.2.3 Proof of Theorem 4.2.1

To prove Theorem 4.2.1, we need Beurling and Kahane’s interpolation theorem (see Theo-

rem 2.6.5). For convenience, we modify the statement and restate as the following:

Theorem. 2.6.5 Let Λ be a sequence of real numbers, and I be an interval of torus T := R/Z. If Λ

is separated with D+ (Λ) < |I|, then there exists a constant C such that

X X
|cλ |2 ≤ Ck cλ e2πiλ(·) k2L2 (I)
λ∈Λ λ∈Λ

for any {cλ }λ∈Λ ∈ l2 (Λ). Where the constant C only depends on D+ (Λ) and |I|, i.e., C =

C(D+ (Λ), |I|).

Besides of Beurling and Kahane’s interpolation theorem, we also need the following lemmas.

Consider any function f ∈ L2 (R) supported on the δ-neighbourhood Eδ of E for some δ > 0. Let

fy := M−y f be the modulation of f , and gyεν be the random periodization of fy .

Let Eyεν := {t ∈ T : gyεν (t) 6= 0} be the support of gyεν , and Fyεν := {t ∈ T : gyεν (t) = 0} be

the zero set of gyεν which is the complement of Eyεν in T.

And denote Λεν


y := {k ∈ Z : kεν ∈ Σ − y}, where Σ − y := {t − y : t ∈ Σ}.

Lemma 4.2.7.

D+ (Λεν +
y ) ≤ ([b] + 2ε)D (Λ).

63
Proof. Without loss of generality, consider the case when y = 0. Let ενΛεν
0 := {kεν|kεν ∈ Σ, k ∈ Z}

be the εν multiple of Λεν


0 . It can be viewed as the intersection of random lattice ενZ and Σ. For

any r > 0 and a ∈ R

#{ενΛεν
0 ∩ (a − r, a + r)}

=#{kεν : kεν ∈ Σ ∩ (a − r, a + r)}


b b
≤#{kεν : kεν ∈ ∪a−r− b <λk <a+r+ b (λk − , λk + )}
2 2 2 2
X b b
≤ #{kεν : kεν ∈ (λk − , λk + )}
b b
2 2
a−r− 2 <λk <a+r+ 2
X b
≤ d e
εν
a−r− 2b <λk <a+r+ 2b

b b b
=d e#{Λ ∩ (a − r − , a + r + )},
εν 2 2

b b
where d εν e denote the smallest integer bigger or equal to εν . Then

D+ (ενΛεν
0 )

#{ενΛεν 0 ∩ (a − r, a + r)}
= lim sup
r→∞ a 2r
b
d εν e#{Λ ∩ (a − r − 2b , a + r + 2b )}
≤ lim sup
r→∞ a 2r
b #{Λ ∩ (a − r − 2b , a + r + 2b )} 2r + b
=d e lim sup
εν r→∞ a 2r + b 2r
b
=d eD+ (Λ).
εν

Simple computation shows that

D+ (Λεν + εν
0 ) = ενD (ενΛ0 )
b
≤ ενd eD+ (Λ)
εν
≤ ([b] + 2ε)D+ (Λ).

1
Lemma 4.2.8. For ε < 2|Eδ | , Fyεν contains an interval I εν with length |I εν | > (1 − 2ε|Eδ |) n1 .

64
Proof. By gyεν is the random periodization of fy ,

1 X k + t −2πi( k+t )y
gyεν (t) = √ f( )e εν .
εν εν
k∈Z

By Eyεν := {t ∈ T : gyεν (t) 6= 0} is the support of gyεν ,

∀t ∈ Eyεν ⇒ gyεν (t) 6= 0,


k0 + t
⇒ ∃k0 ∈ Z, s.t.f ( ) 6= 0,
εν
k0 + t
⇒ ∈ Eδ ,
εν
⇒ t ∈ ενEδ − k0 ,

⇒ t ∈ ενEδ (mod1),

where ενEδ := {ενt : t ∈ Eδ }, and ενEδ (mod1) := {t + Z : t ∈ ενEδ } ⊆ T.

So Eyεν ⊆ ενEδ (mod1) ⊆ T. Since Eδ is at most n union of intervals, so is ενEδ (mod1).

Then T \ ενEδ (mod1) is at most n union of intervals as well. Notice

|T \ ενEδ (mod1)| = 1 − |ενEδ (mod1)|

> 1 − |ενEδ |

> 1 − 2ε|Eδ |.

1−2ε|Eδ |
So T \ ενEδ (mod1) contains an interval I with length |I| > n . By Fyεν = T \ Eyεν ⊃ T \

ενEδ (mod1), Fyεν contains interval I as well.

1
Lemma 4.2.9. If [b]D+ (Λ) < n, then there exists a constant C such that

Z
sup |fb(y)|2 ≤ C |fb(x)|2 dx,
y∈Σ Σc

for any function f ∈ L2 (R) supported on Eδ .

Proof. Let gyεν be the random periodization of f ∈ L2 (R) supported on Eδ . Rewrite gyεν as the sum

65
of functions pεν εν
y and qy ,

X
gyεν (t) = εν
gc
y (k)e
2πikt

k∈Z

= pεν εν
y (t) + qy (t),

where pεν 2πikt


and qyεν (t) := 2πikt
P εν
P εν
y (t) := k∈Λεν
y
gc
y (k)e / εν
k∈Λy
gc
y (k)e .

When y ∈ Σ, it implies 0 ∈ Λεν


y . Then for any y ∈ Σ,

|fb(y)|2 = |fby (0)|2


1 εν
= |gc (0)|2
εν y
1 εν
≤ |gc (0)|2
ε y
1 X εν 2
≤ |gc
y (k)| . (4.5)
ε εν
k∈Λy

On one hand, by lemma 4.2.7, D+ (Λεν + +


y ) ≤ ([b] + 2ε)D (Λ). Let ε → 0 ,

D+ (Λεν + +
y ) ≤ ([b] + 2ε)D (Λ) ↓ [b]D (Λ).

On the other hand, by lemma 4.2.8, Fyεν contains an interval I εν with length |I εν | > (1 − 2ε|Eδ |) n1 .

Let ε → 0+ ,
1 1
|I εν | > (1 − 2ε|Eδ |) ↑ .
n n
1 1
Since [b]D+ (Λ) < n, let d := n − [b]D+ (Λ) > 0, and ε0 := min{ 6D+d(Λ) , 6|E
nd
δ|
}. Then

d 1 d
D+ (Λεy0 ν ) ≤ [b]D+ (Λ) + < − < |I ε0 ν |.
3 n 3

By Theorem 2.6.5, there exists a constant C0 = C0 ([b]D+ (Λ) + d3 , n1 − d3 ) such that

2
X 2 X
ε0 ν ε0 ν 2πik(·)

gy (k) ≤ C0 gy (k)e .
d d

ε ν k∈Λεy0 ν
k∈Λy0

L2 (I ε0 ν )

Notice gyε0 ν (t) = pεy0 ν (t) + qyε0 ν (t) = 0 on its zero set Fyε0 ν which contains interval I ε0 ν . Combin-

66
ing (4.5), we have for any y ∈ Σ,

1 X d
|fb(y)|2 ≤ |gyε0 ν (k)|2
ε0 ε0 ν
k∈Λy

C0 X d
≤ k gyε0 ν (k)e2πik(·) k2L2 (I ε0 ν )
ε0 ε0 ν
k∈Λy

C0 ε0 ν 2
= kp k 2 ε ν
ε0 y L (I 0 )
C0 ε0 ν 2
= kq k 2 ε ν
ε0 y L (I 0 )
C0 ε0 ν 2
≤ kq k (4.6)
ε0 y
C0 X d
= |gyε0 ν (k)|2 .
ε0 ε0 ν
k∈Λ
/ y

By Proposition 4.2.5,

X Z
Ekqyε0 ν k2 =E ε0 ν (k)|2 ≤ 2
|gd |fb(x)|2 dx.
ε ν Σc
/ y0
k∈Λ

By Chebyshev’s inequality,

Z
P{kqyε0 ν k2 > 4 |fb|2 }
Σc

≤P{kqyε0 ν k2 > 2Ekqyε0 ν k2 }


Ekqyε0 ν k2
<
2Ekqyε0 ν k2
1
= .
2

Then,
Z
1
P{kqyε0 ν k2 ≤ 4 |fb|2 } > .
Σc 2

So there exists a ν0 ∈ (1, 2) such that

Z
kqyε0 ν0 k2 ≤4 |fb(x)|2 dx.
Σc

Combine (4.6), we have


Z
C0 ε0 ν0 2 4C0
|fb(y)|2 ≤ kqy k ≤ |fb(x)|2 dx.
ε0 ε0 Σc

67
Since y ∈ Σ is arbitrary,
Z
4C0
sup |fb(y)|2 ≤ |fb(x)|2 dx,
y∈Σ ε0 Σc

nd
where ε0 = min{ 6D+d(Λ) , 6|E δ|
} with d = 1
n − [b]D+ (Λ), and C0 does not depend on |E|.

Proof of theorem 4.2.1. For every function f ∈ L2 (E) and every y ∈ Σ, define

e2πiyt
hy (t) := 1(δ,δ) (t), t ∈ R,

and

Fy (t) := f ∗ hy (t), t ∈ R,

where f ∗ hy denote the convolution of f and hy . Then

cy (x) = fb(x) sin(2πδ(x − y)) ,


F (4.7)
2πδ(x − y)

and suppFy ⊆ Eδ . By Lemma 4.2.9,

Z
cy (x)|2 ≤ 4C0
sup |F cy (x)|2 dx.
|F (4.8)
x∈Σ ε0 Σc

Let x = y ∈ Σ. Combining (4.7) and (4.8), we obtain

Z
4C0
|fb(y)|2 = |F
cy (y)|2 ≤ sup |F
cy (x)|2 ≤ cy (x)|2 dx.
|F
x∈Σ ε0 Σc

Integrating this inequality over Σ on both sides, we have

Z Z Z
4C0
|fb(y)|2 dy ≤ cy (x)|2 dxdy
|F
Σ Σ ε 0 c
Z ZΣ
4C0 sin(2πδ(x − y)) 2
= |fb(x)|2 | | dxdy
ε0 Σ Σc 2πδ(x − y)
sin(2πδ(x − y)) 2
Z Z
4C0
= |fb(x)|2 | | dydx
ε0 Σc Σ 2πδ(x − y)
sin(2πδ(x − y)) 2
Z Z
4C0
= |fb(x)|2 | | dydx
ε0 Σc Σ 2πδ(x − y)
sin(2πδ(x − y)) 2
Z Z
4C0
≤ |fb(x)|2 | | dydx
ε0 Σc R 2πδ(x − y)
Z
2C0
= |fb(x)|2 dx.
ε0 δ Σc

68
Then we obtain
2 Z
ε0 δ
f 2 ≤ |fb(x)|2 dx.
b
ε0 δ + 2C0 L (R) Σc

nd
Since ε0 = min{ 6D+d(Λ) , 6|E δ|
} with d = 1 +
n −[b]D (Λ), and C0 does not rely on |E| (see Lemma 4.2.9).
1
Let δ = 2n , for large enough |E| we have

d ε0 δ
C := ≤ .
(24C0 + 1)(|E| + 1) ε0 δ + 2C0

It follows that
2 Z
C fb 2 ≤ |fb(x)|2 dx,

L (R) Σc

where C −1 grows linearly with |E| when |E| is large.

4.2.4 Proof of Theorem 4.2.2

In order to prove Theorem 4.2.2, we need the following new lemmas which play the same

role with Lemmas 4.2.7 and 4.2.9 for the proof of Theorem 4.2.1.

Again, consider any function f ∈ L2 (R) supported on the δ-neighbourhood Eδ of E for some

δ > 0. Let fy := M−y f be the modulation of f , and gyεν be the random periodization of fy .

Let Eyεν := {t ∈ T : gyεν (t) 6= 0} be the support of gyεν , and Fyεν := {t ∈ T : gyεν (t) = 0} be

the zero set of gyεν .

And denote Λεν


y := {k ∈ Z : kεν ∈ Σ − y}.

Lemma 4.2.10. If inf k {bk } > b for some positive number b, then

2ε +
D+ (Λεν
y ) ≤ (1 + )D (Σ).
b

Proof. Without loss of generality, consider the case when y = 0. Let ενΛεν
0 := {kεν : kεν ∈ Σ, k ∈ Z}

be the εν multiple of Λεν


0 . For any r > 0 and a ∈ R, by inf k {bk } > b > 0, we have finite number of
bk bk
intervals (λk − 2 , λk + 2 ) ⊆ Σ which intersect with interval (a − r, a + r). Denote the length of

69
those intersections by {l1 , l2 , · · · , ln }, then

#{ενΛεν
0 ∩ (a − r, a + r)}

=#{kεν : kεν ∈ Σ ∩ (a − r, a + r)}


l1 l2 ln
≤d e + d e + ··· + d e
εν εν εν
l1 l2 ln
<( + 1) + ( + 1) + · · · + ( + 1)
εν εν εν
l1 + l2 + · · · + ln
= +n
εν
|Σ ∩ (a − r, a + r)|
= + n.
εν

Furthermore, it’s easy to see (n − 2)b ≤ |Σ ∩ (a − r, a + r)|, so

|Σ ∩ (a − r, a + r)|
n≤ + 2.
b

Then

#{ενΛεν
0 ∩ (a − r, a + r)}
D+ (ενΛεν
0 ) = lim sup sup
r→∞ a 2r
|Σ∩(a−r,a+r)|
εν +n
≤ lim sup sup
r→∞ a 2r
|Σ∩(a−r,a+r)|
εν+ |Σ∩(a−r,a+r)|
b +2
≤ lim sup sup
r→∞ a 2r
 
1 1 |Σ ∩ (a − r, a + r)| 2
= lim sup sup ( + ) +
r→∞ a εν b 2r 2r
1 1 |Σ ∩ (a − r, a + r)|
= ( + ) lim sup sup
εν b r→∞ a 2r
1 1 +
= ( + )D (Σ).
εν b

By simple computations,

D+ (Λεν + εν
0 ) = ενD (ενΛ0 )
1 1
≤ εν( + )D+ (Σ)
εν b
2ε +
≤ (1 + )D (Σ).
b

70
1
Lemma 4.2.11. If D+ (Σ) < n, then there exists a constant C such that

Z
sup |fb(y)|2 ≤ C |fb(x)|2 dx,
y∈Σ Σc

for any function f ∈ L2 (R) supported on Eδ .

Proof. Let gyεν be the random periodization of f ∈ L2 (R) supported on Eδ . We can rewrite gyεν as

the sum of two functions.

X
gyεν (t) = εν
gc
y (k)e
2πikt

k∈Z

= pεν εν
y (t) + qy (t),

where pεν 2πikt


, qyεν (t) = 2πikt
P εν
P εν
y (t) = k∈Λεν
y
gc
y (k)e / εν
k∈Λy
gc
y (k)e .

When y ∈ Σ, 0 ∈ Λεν
y . Then for y ∈ Σ,

|fb(y)|2 = |fby (0)|2


1 εν
= |gc (0)|2
εν y
1 εν
≤ |gc (0)|2
ε y
1 X εν 2
≤ |gc
y (k)| . (4.9)
ε ενk∈Λy


By lemma 4.2.10, D+ (Λεν
y ) ≤ (1 +
+
b )D (Σ). Let ε → 0+ ,

2ε +
D+ (Λεν
y ) ≤ (1 + )D (Σ) ↓ D+ (Σ).
b

By lemma 4.2.8, Fyεν contains an interval I εν with length |I εν | > (1 − 2ε|Eδ |) n1 . Let ε → 0+ ,

1 1
|I εν | > (1 − 2ε|Eδ |) ↑ .
n n

71
1 1
Since D+ (Σ) < n, let d := n − D+ (Σ) > 0, and ε0 := min{ 6Dbd nd
+ (Σ) , 6|E | }. Then
δ

d 1 d
D+ (Λεy0 ν ) ≤ D+ (Σ) + < − < |I ε0 ν |.
3 n 3

Again, by Theorem 2.6.5, there exists a constant C0 = C0 (D+ (Σ) + d3 , n1 − d3 ) such that

2
2
X X
ε0 ν ε0 ν 2πik(·)

gy (k) ≤ C0 gy (k)e .
d d

ε ν k∈Λεy0 ν
k∈Λy0

L2 (I ε0 ν )

By gyε0 ν (t) = pεy0 ν (t) + qyε0 ν (t) = 0 on its zero set Fyε0 ν which contains the interval I ε0 ν . Combin-

ing (4.9), we have for every y ∈ Σ,

1 X d
|fb(y)|2 ≤ |gyε0 ν (k)|2
ε0 ε0 ν
k∈Λy

C0 X d
≤ k gyε0 ν (k)e2πik(·) k2L2 (I ε0 ν )
ε0 ε0 ν
k∈Λy

C0 ε0 ν 2
= kp k 2 ε ν
ε0 y L (I 0 )
C0 ε0 ν 2
= kq k 2 ε ν
ε0 y L (I 0 )
C0 ε0 ν 2
≤ kq k (4.10)
ε0 y
C0 X d
= |gyε0 ν (k)|2 .
ε0 ε0 ν
k∈Λ
/ y

By Proposition 4.2.5,

X Z
Ekqyε0 ν k2 =E ε0 ν (k)|2 ≤ 2
|gd |fb(x)|2 dx.
ε ν Σc
/ y0
k∈Λ

72
By Chebyshev’s inequality,

Z
P{kqyε0 ν k2 >4 |fb|2 }
Σc

≤P{kqyε0 ν k2 > 2Ekqyε0 ν k2 }


Ekqyε0 ν k2
<
2Ekqyε0 ν k2
1
= .
2

Then,
Z
1
P{kqyε0 ν k2 ≤4 |fb|2 } > .
Σc 2

So there exists a ν0 ∈ (1, 2) such that

Z
kqyε0 ν0 k2 ≤ 4 |fb(x)|2 dx.
Σc

Combining (4.10), we obtain

Z
C0 ε0 ν0 2 4C0
|fb(y)|2 ≤ kq k ≤ |fb(x)|2 dx.
ε0 y ε0 Σc

Since y ∈ Σ is arbitrary,
Z
4C0
sup |fb(y)|2 ≤ |fb(x)|2 dx,
y∈Σ ε0 Σc

where ε0 = min{ 6Dbd nd


+ (Σ) , 6|E | } with d =
δ
1
n − D+ (Σ) > 0, and C0 does not depend on |E|.

Proof of theorem 4.2.2. For every function f ∈ L2 (E) and every y ∈ Σ, define

e2πiyt
hy (t) := 1(δ,δ) (t), t ∈ R,

and

Fy (t) := f ∗ hy (t), t ∈ R.

Then
cy (x) = fb(x) sin(2πδ(x − y)) ,
F
2πδ(x − y)

73
and suppFy ⊆ Eδ . By Lemma 4.2.11

Z
cy (x)|2 ≤ 4C0
sup |F cy (x)|2 dx.
|F
x∈Σ ε0 Σc

Let x = y ∈ Σ, we obtain

Z
|fb(y)|2 = |F cy (x)|2 ≤ 4C0
cy (y)|2 ≤ sup |F cy (x)|2 dx.
|F
x∈Σ ε0 Σc

Integrating the inequality over Σ on both sides, we get:

Z Z Z
4C0
|fb(y)|2 dy ≤ cy (x)|2 dxdy
|F
Σ Σ ε0 Σc
sin(2πδ(x − y)) 2
Z Z
4C0
= |fb(x)|2 | | dxdy
ε0 Σ Σc 2πδ(x − y)
sin(2πδ(x − y)) 2
Z Z
4C0
= |fb(x)|2 | | dydx
ε0 Σc Σ 2πδ(x − y)
sin(2πδ(x − y)) 2
Z Z
4C0
= |fb(x)|2 | | dydx
ε0 Σc 2πδ(x − y)

sin(2πδ(x − y)) 2
Z
4C0
≤ |fb(x)|2 | | dydx
ε0 Σc R 2πδ(x − y)
Z
2C0
= |fb(x)|2 dx.
ε0 δ Σc

Then we obtain
2 Z
ε0 δ
f 2 ≤ |fb(x)|2 dx.
b
ε0 δ + 2C0 L (R) Σc

Since ε0 = min{ 6Dbd nd


+ (Σ) , 6|E | } with d =
δ
1
n − D+ (Σ) > 0, and C0 does not depend on |E| (see
1
Lemma 4.2.11). Let δ = 2n , for large enough |E| we have

d ε0 δ
C := ≤ .
(24C0 + 1)(|E| + 1) ε0 δ + 2C0

It follows that
2 Z
C fb ≤ |fb(x)|2 dx,

L2 (R) Σc

where C −1 grows linearly with |E| when |E| is large.

74
Chapter 5

Uncertainty Principles in Framed

Hilbert Spaces

5.1 Introduction

5.1.1 Uncertainty Principle for Gabor Riesz Bases

Let X and D be the multiplication and differentiation operators on their domain D(X) and

D(D) (see Section 1.1). For any f ∈ L2 (R) and any a, b ∈ R, we define the notations:

Z
2 2
k(X − a)f k2 := |(x − a)f (x)| dx = ∞, when f ∈
/ D(X),
R

Z 2
2
k(D − a)f k2 := (ξ − b)fb(ξ) dξ = ∞, when f ∈
/ D(D).

R

Recall the classical uncertainty principle (Theorem 1.2.1): for any f ∈ L2 (R) and any a, b ∈ R,

2 2 1 4
k(X − a)f k2 k(D − b)f k2 ≥ kf k2 . (5.1)
16π 2

And the classical Balian-Low theorem (Theorem 1.3.2) could be restated as: Let g ∈ L2 (R) and

α, β > 0 with αβ = 1. If the Gabor system {Mmβ Tnα g}m,n∈Z forms a Riesz basis for L2 (R), then

for any a, b ∈ R,
2 2
k(X − a)f k2 k(D − b)f k2 = ∞, (5.2)

75
or equivalently
2 2
k(X − a)f k2 + k(D − b)f k2 = ∞.

Comparing the similarity of (5.1) and (5.2), the Balian-Low theorem could be view as the uncertainty

principle type of results for Gabor Riesz Bases.

5.1.2 Uncertainty Principle for General Riesz Bases

It is natural for mathematicians to release the Gabor system structures and work on more

general bases in next step. So Meyer asks the following question: is it true that for any orthonormal

basis {fn }∞ 2
n=1 ⊆ L (R) (which may not be a Gabor system) and any sequences of real numbers

{an }∞ ∞
n=1 and {bn }n=1 , we have

 
2 2
sup k(X − an )fn k2 + k(D − bn )fn k2 = ∞. (5.3)
n∈N

If this is true, then it would be a “generalized Balian-Low theorem” for general orthonormal bases.

However the answer is No! Bourgain constructed an orthonormal basis {fn }∞ 2


n=1 of L (R) such

that (5.3) fails.

Based on this fact, the next question will be: can we impose some extra conditions to

make (5.3) still hold. In 2011, Gröchenig and Malinnikova proved the following very similar equality

holds for general Riesz bases.

Theorem 5.1.1 ([6], Theorem 1). If {fn }∞ 2


n=1 is a Riesz basis of L (R) and s > 1, then

Z Z 
sup |x − an |2s |fn (x)|2 dx + |ξ − bn |2s |fbn (ξ)|2 dξ = ∞,
n∈N R R

for any sequences of real numbers {an }∞ ∞


n=1 , {bn }n=1 .

By studying Theorem 5.1.1, we get some hints to formulate the conditions for our “gener-

alized Balian-Low theorem”.

76
5.2 Generalized Balian-Low Theorem

5.2.1 Main Results

In order to introduce our main theorem, we set up the following definition and notation:

Definition 5.2.1. We say a countable subset Λ of a metric space (X, d) decays annularly around

point a ∈ X, if for any ρ > 0

#{Λ ∩ (B(a, r + ρ) − B(a, r))}


lim = 0.
r→∞ #{Λ ∩ B(a, r)}

Let A and B be two operators on a Hilbert space H, denote their domain by D(A) := {f ∈

H : Af ∈ H} and D(B) := {f ∈ H : Bf ∈ H}. When f, g ∈ D(A) ∩ D(B), we define the notation

hi[B, A]f, gi := i hAf, Bgi − i hBf, Agi .

We now state our generalized Balian-Low theorem:

Theorem 5.2.2. Let {fn }∞ ∞ ∞


n=1 and {gn }n=1 be dual Riesz bases of H, and Λ := {λn }n=1 be a

countable subset of some metric space (X, d) which decays annularly around point a ∈ X and satisfies

#{Λ ∩ B(a, r)} < ∞ for any r > 0. And let A : D(A) → H and B : D(B) → H be two symmetric

operators such that for every n ∈ N, hi[B, A]gn , fn i ≥ c for some c > 0. If
P 2
1. limr→∞ supn∈N d(λm −λn )>r |hAgn , fm i| = 0,

P 2
2. limr→∞ supn∈N d(λm −λn )>r |hBgn , fm i| = 0.

Then for any sequences of complex numbers {an }∞ ∞


n=1 , {bn }n=1 ,

 
2 2
sup k(A − an )fn k + k(B − bn )fn k = ∞.
n∈N

Notice if we let Λ := Zd embedded in the Euclidean metric space (Rd , |·|), then Λ would

satisfy all the requirements in Theorem 5.2.2. Based on this fact, we have the following corollary:

Corollary 5.2.3. Let {fn }n∈Zd be an orthonormal basis of H. And let A : D(A) → H and B :

D(B) → H be two symmetric operators such that for every n ∈ Zd , hi[B, A]fn , fn i ≥ c for some

c > 0. If

77
P 2
1. limr→∞ supn∈Zd |m−n|>r |hAfn , fm i| = 0,

P 2
2. limr→∞ supn∈Zd |m−n|>r |hBfn , fm i| = 0.

Then for any sequences of complex numbers {an }n∈Zd , {bn }n∈Zd ,

 
2 2
sup k(A − an )fn k + k(B − bn )fn k = ∞.
n∈Zd

5.2.2 Proof of Theorem 5.2.2

Proof. Suppose there exists C > 0 such that

 
2 2
sup k(A − an )fn k + k(B − bn )fn k ≤ C < ∞.
n∈N

We claim that a contradiction will follow from this assumption.

Since {fn }n∈N and {gn }n∈N are dual Riesz bases, for any n ∈ N we have

c ≤ hi[A, B]gn , fn i

:= i hBgn , Afn i − i hAgn , Bfn i


X X
=i hBgn , fm i hgm , Afn i − i hAgn , fm i hgm , Bfn i
m∈N m∈N
X
=i hBgn , fm i hgm , Afn i − hAgn , fm i hgm , Bfn i .
m∈N

Let B := B(a, r) be the ball with the fixed center a (for which Λ decays annularly around)

and radius r > 0, then

X X
c#{Λ ∩ B} ≤ i hBgn , fm i hgm , Afn i − hAgn , fm i hgm , Bfn i
λn ∈Λ∩B m∈N
X X
=i hBgn , fm i hgm , Afn i − hAgn , fm i hgm , Bfn i
λn ∈Λ∩B λm ∈Λ∩B
X X
+i hBgn , fm i hgm , Afn i − hAgn , fm i hgm , Bfn i .
λn ∈Λ∩B λm ∈Λ∩B c

78
By A, B are symmetric, the first term vanishes. So we have

X X
c#{Λ ∩ B} ≤ i hBgn , fm i hgm , Afn i − hAgn , fm i hgm , Bfn i .
λn ∈Λ∩B λm ∈Λ∩B c

Put absolute value sign inside, we obtain

X X
c#{Λ ∩ B} ≤ |hBgn , fm i hgm , Afn i − hAgn , fm i hgm , Bfn i|
λn ∈Λ∩B λm ∈Λ∩B c
X X
≤ |hBgn , fm i hgm , Afn i| + |hAgn , fm i hgm , Bfn i|
λn ∈Λ∩B λm ∈Λ∩B c
X X
≤ |hBgn , fm i hgm , Afn i|
λn ∈Λ∩B λm ∈Λ∩B c
X X
+ |hAgn , fm i hgm , Bfn i| . (5.4)
λn ∈Λ∩B λm ∈Λ∩B c

Apply Cauchy-Schwarz twice for the first term, we have

X X
|hBgn , fm i hgm , Afn i|
λn ∈Λ∩B λm ∈Λ∩B c
! 12 ! 21
X X 2
X X 2
≤ |hBgn , fm i| |hgm , Afn i| .
λn ∈Λ∩B λm ∈Λ∩B c λn ∈Λ∩B λm ∈Λ∩B c

Let ρ > 0 and denote B(a, r + ρ) by Bρ which is the ρ-neighbourhood of B. By the

79
assumption, the first factor satisfies

X X 2
|hBgn , fm i|
λn ∈Λ∩B λm ∈Λ∩B c
X X 2
X X 2
= |hBgn , fm i| + |hBgn , fm i|
λn ∈Λ∩B λm ∈Λ∩Bρc λn ∈Λ∩B λm ∈Λ∩(Bρ \B)
X X 2
X X 2
≤ |hBgn , fm i| + |hgn , Bfm i|
λn ∈Λ∩B λm ∈Λ∩B(λn ,ρ)c λm ∈Λ∩(Bρ \B) λn ∈Λ∩B
X X 2
X X 2
= |hBgn , fm i| + |hgn , (B − bm )fm i|
λn ∈Λ∩B λm ∈Λ∩B(λn ,ρ)c λm ∈Λ∩(Bρ \B) λn ∈Λ∩B
X X 2
X 2
≤ sup |hBgn , fm i| + C k(B − bm )fm k
n∈N
λn ∈Λ∩B λm ∈Λ∩B(λn ,ρ)c λm ∈Λ∩(Bρ \B)
X 2
X
≤ #{Λ ∩ B} sup |hBgn , fm i| + C
n∈N
λm ∈Λ∩B(λn ,ρ)c λm ∈Λ∩(Bρ \B)
X 2
= #{Λ ∩ B} sup |hBgn , fm i| + C#{Λ ∩ (Bρ \ B)}.
n∈N
λm ∈Λ∩B(λn ,ρ)c

Again by the assumption, the second factor satisfies

X X 2
X X 2
|hgm , Afn i| = |hgm , (A − an )fn i|
λn ∈Λ∩B λm ∈Λ∩B c λn ∈Λ∩B λm ∈Λ∩B c
X 2
X
≤ C k(A − an )fn k ≤ C ≤ C#{Λ ∩ B}.
λn ∈Λ∩B λn ∈Λ∩B

Then for some C > 0, we get an estimate for the first term,

X X
|hBgn , fm i hgm , Afn i|
λn ∈Λ∩B λm ∈Λ∩B c
  21
1
X 2
≤C#{Λ ∩ B} 2 #{Λ ∩ B} |hBgn , fm i| + #{Λ ∩ (Bρ \ B)} . (5.5)
λm ∈Λ∩B(λn ,ρ)c

Similarly, for some C > 0, we get an estimate for the second term,

X X
|hAgn , fm i hgm , Bfn i|
λn ∈Λ∩B λm ∈Λ∩B c
  12
1
X 2
≤C#{Λ ∩ B} #{Λ ∩ B}
2 |hAgn , fm i| + #{Λ ∩ (Bρ \ B)} . (5.6)
λm ∈Λ∩B(λn ,ρ)c

80
So for any r > 0 and ρ > 0, combine (5.4) (5.5) (5.6), we have

1 X X
c≤ |hBgn , fm i hgm , Afn i|
#{Λ ∩ B}
λn ∈Λ∩B λm ∈Λ∩B c
1 X X
+ |hAgn , fm i hgm , Bfn i|
#{Λ ∩ B}
λn ∈Λ∩B λm ∈Λ∩B c
  21
X 2 #{Λ ∩ (Bρ \ B)}
≤ C sup |hBgn , fm i| + 
n∈N c
#{Λ ∩ B}
λm ∈Λ∩B(λn ,ρ)
  12
X 2 #{Λ ∩ (B ρ \ B)}
+ C sup |hAgn , fm i| +  . (5.7)
n∈N c
#{Λ ∩ B}
λm ∈Λ∩B(λn ,ρ)

Let ε > 0, by conditions 1 and 2, we can find a ρ > 0 such that

X 2
X 2
sup |hBgn , fm i| < ε2 , sup |hAgn , fm i| < ε2 . (5.8)
n∈N n∈N
λm ∈Λ∩B(λn ,ρ)c λm ∈Λ∩B(λn ,ρ)c

So for such ρ, combine (5.7) and (5.8), we have

 1  1
c 2 #{Λ ∩ (Bρ \ B)} 2 2 #{Λ ∩ (Bρ \ B)} 2
≤ ε + + ε + ,
C #{Λ ∩ B} #{Λ ∩ B}

for any r > 0. Let r → ∞, by Λ decays annularly around point a,

 1  1
c 2 #{Λ ∩ (Bρ \ B)} 2 2 #{Λ ∩ (Bρ \ B)} 2
≤ lim ε + + ε +
C r→∞ #{Λ ∩ B} #{Λ ∩ B}
1 1
= ε2 + 0 2 + ε2 + 0 2 = 2ε.
 

By ε is arbitrarily small, contradiction completes.

5.2.3 Applications

Apply Corollary 5.2.3, we could give another proof of the classical Balian-Low theorem.

Proof of Theorem 1.3.2. For convenience, we let a = b = 1. The multiplication operator X and

differentiation operator D are two symmetric operators which will play the role of operators A and

B in Corollary 5.2.3 respectively.

81
Again, we will prove the theorem by contradiction. Suppose

Z Z
2 2 2
|xg(x)| dx g (ξ)|2 dξ = kXgk2 kDgk2 < ∞.
|ξb (5.9)
R R

Denote {Mm Tn g}m,n∈Z by {gm,n }m,n∈Z , then for any m, n ∈ Z

Z
2 2
kXgm,n k2 := |Xgm,n (x)| dx
R
Z
2πimx 2
= xe g(x − n) dx
ZR
2 2
= |x + n| |g(x)| dx
R
Z Z
2 2 2 2
≤2 |x| |g(x)| dx + 2 |n| |g(x)| dx
R R
2 2 2
= 2 kXgk2 + 2 |n| kgk2 < ∞,

which implies {gm,n }m,n∈Z ⊆ D(X). Similarly, we could obtain {gm,n }m,n∈Z ⊆ D(D). By general-

ized canonical commutation relation (see (1.6)), we have for any m, n ∈ Z

hi[D, X]gm,n , gm,n i

:=i hXgm,n , Dgm,n i − i hDgm,n , Xgm,n i


1 2
= kgm,n k2

1
= > 0.

82
By (1.2), we obtain

X 2
|hXgm,n , gk,l i|
k(k,l)−(m,n)k>r
X 2
= |hXMm Tn g, gk,l i|
k(k,l)−(m,n)k>r
X 2
= |hMm XTn g, gk,l i|
k(k,l)−(m,n)k>r
X 2
= |hMm (Tn X + nTn )g, gk,l i|
k(k,l)−(m,n)k>r
X 2
= |hMm Tn Xg, gk,l i|
k(k,l)−(m,n)k>r
X 2
= |hMm Tn Xg, Mk Tl gi|
k(k,l)−(m,n)k>r

Xg, M(k−m) T(l−n) g 2


X

=
k(k,l)−(m,n)k>r
X 2
= |hXg, Mk Tl gi| .
k(k,l)k>r
X 2
= |hXg, gk,l i| .
k(k,l)k>r

By {gm,n }m,n∈Z forms an orthonormal basis,

X 2
lim sup |hXgm,n , gk,l i|
r→∞ m,n∈Z
k(k,l)−(m,n)k>r
X 2
= lim sup |hXg, gk,l i|
r→∞ m,n∈Z
k(k,l)k>r
X 2
= lim |hXg, gk,l i| = 0.
r→∞
k(k,l)k>r

Similarly,
X 2
lim sup |hDgm,n , gk,l i| = 0.
r→∞ m,n∈Z
k(k,l)−(m,n)k>r

Apply Corollary 5.2.3, we have for any sequences {am,n }m,n∈Z and {bm,n }m,n∈Z ,

 
2 2
sup k(X − am,n )gm,n k2 + k(D − bm,n )gm,n k2 = ∞ (5.10)
m,n∈Z

83
On the other hand, let am,n = n and bm,n = m for any m, n ∈ Z. Again by (1.2), we have

2 2
k(X − n)gm,n k2 + k(D − m)gm,n k2
2 2
= k(X − n)Mm Tn gk2 + k(D − m)Mm Tn gk2
2 2
= kMm (X − n)Tn gk2 + kMm DTn gk2
2 2
= kMm Tn Xgk2 + kMm Tn Dgk2
2 2
= kXgk2 + kDgk2 ,

for any m, n ∈ Z. By assumption (5.9),

 
2 2
sup k(X − n)gm,n k2 + k(D − m)gm,n k2
m,n∈Z
 
2 2
= sup kXgk2 + kDgk2
m,n∈Z

2 2
= kXgk2 + kDgk2 < ∞.

Which contradicts with (5.10).

5.3 Balian-Low Theorem in Framed Hilbert Spaces

5.3.1 Uniformly Localized Sequences in Framed Hilbert Spaces

Our next result is based on FHS. In order to state the theorem, we introduce the following

definitions. Let (H, {kx }x∈(X,d,σ) ) be a FHS, and Λ := {λn }∞


n=1 be a countable subset of X.

Definition 5.3.1. A sequence {fn }∞


n=1 of H is said to be uniformly localized on Λ if

Z
2
lim sup |hfn , kx i| dσ(x) = 0.
r→∞ n∈N B(λn ,r)c

Recall the ULP (in Section 2.4.2), it is not hard to see if {kx }x∈(X,d,σ) satisfies the ULP, then

the sequence {kλn }∞ ∞


n=1 in {kx }x∈(X,d,σ) is always uniformly localized on its index set Λ = {λn }n=1 .

Definition 5.3.2. Let w : [0, ∞) → [0, ∞) be a weight function satisfies w(d) ↑ ∞ as d → ∞. A

84
sequence {fn }∞
n=1 of H is said to be uniformly w-localized on Λ if

Z
2
sup |hfn , kx i| w(d(λn , x))dσ(x) < ∞.
n∈N X

The following proposition proves that uniform w-localization implies uniform localization.

Proposition 5.3.3. Let {fn }∞ ∞


n=1 be a sequence of H. If {fn }n=1 is uniformly w-localized on Λ,

then it is uniformly localized on Λ.

Proof. By {fn }∞
n=1 is uniformly w-localized on Λ,

Z
2
sup |hfn , kx i| w(d(λn , x))dσ(x) < ∞,
n∈N X

for some weight function w. It follows that

Z
2
sup |hfn , kx i| dσ(x)
n∈N B(λn ,r)c
Z
2 w(d(λn , x))
= sup |hfn , kx i| dσ(x)
n∈N B(λn ,r)c w(d(λn , x))
Z
2 w(d(λn , x))
≤ sup |hfn , kx i| dσ(x)
n∈N B(λn ,r)c w(r)
Z
1 2
≤ sup |hfn , kx i| w(d(λn , x))dσ(x)
w(r) n∈N X
1
. .
w(r)

By the property of w,

Z
2 1
lim sup |hfn , kx i| dσ(x) . lim = 0.
r→∞ n∈N B(λn ,r)c r→∞ w(r)

Definition 5.3.4. Let w : [0, ∞) → [0, ∞) be a weight function satisfies w(d) ↑ ∞ as d → ∞. We

say w is localized on Λ if

X 1
lim sup = 0.
r→∞ n∈N w(d(λn , λm ))
d(λn ,λm )≥r

One example of localized weights is the power weight on the lattice of integers. Let Λ := Zn

85
be the subset of the Euclidean metric space (Rn , |·|), and let the weight function w(d) := ds be the

power function with s > n. Easily see w(d) ↑ ∞ as d → ∞.

By Euclidean metric is equivalent to maximum metric, i.e., |x| ' maxi |xi |,

X 1 X 1
s .
|m| max{|m1 | , |m2 | , · · · , |mn |}s
m∈Zn \{0} m∈Zn \{0}

X 1
= ((l + 1)n − ln )
ls
l=1

X (2n − 1)ln−1

ls
l=1

X 1
.
ls−n+1
l=1

< ∞. (5.11)

It implies

X 1
lim sup s
r→∞ n∈Zn |n − m|
|n−m|≥r
X 1
= lim s
r→∞ |m|
|m|≥r

=0.

So in the Euclidean metric space (Rn , |·|), the weight function w(d) = ds with s > n is localized on

Λ = Zn .

5.3.2 Main Results

Let (H, {kx }x∈(X,d,σ) ) be a FHS which satisfies ADP and UMD, then our next result is the

following:

Theorem 5.3.5. Let {fn }∞ ∞


n=1 and {gn }n=1 be dual Riesz bases of H where the index set Λ :=

{λn }∞ +
n=1 is relatively separated with 0 < D (Λ). Let A : D(A) → H and B : D(B) → H be two

symmetric operators such that for every n ∈ N, hi[B, A]gn , fn i ≥ c for some c > 0. If there exist

sequences of complex numbers {an }∞ ∞


n=1 , {bn }n=1 and a weight function w such that the following

conditions hold:

86
 
2 2
1. supn∈N k(A − an )gn k + k(B − bn )gn k < ∞,

2. {(A − an )gn }∞ ∞
n=1 and {(B − bn )gn }n=1 are uniformly localized on Λ,

3. {fn }∞
n=1 is uniformly w-localized on Λ,

4. w is localized on Λ, and w(2d) . w(d) for every d ≥ 0.

Then
 
2 2
sup k(A − an )fn k + k(B − bn )fn k = ∞.
n∈N

5.3.3 Prerequisites

Before we give a proof of Theorem 5.3.5, we need to do some preparations.

Proposition 5.3.6. Let (X, d, σ) be a metric measure space satisfying ADP and UMD, and Λ be a

relatively separated set in X. Then for every ρ > 0,

#{Λ ∩ (B(a, r + ρ) \ B(a, r))}


lim sup sup = 0.
r→∞ a∈X σ(B(a, r))

Proof. Since Λ is relatively separated, we have

Λ = ∪K
k=1 Λk ,

where Λk is a separated set with the separation constant δk for any k = 1, · · · , K. Let δ :=

min1≤k≤K δk , and {λki }i∈Ik := {Λk ∩ B(a, r)} for any k = 1, · · · , K.

By UMD, there exist c( 2δ ) > 0 and D(r + 2δ ) > 0 such that

#{Λ ∩ (B(a, r + ρ) \ B(a, r))}


K
X
≤ #{Λk ∩ (B(a, r + ρ) \ B(a, r))}
k=1
K
X σ(B(a, r + ρ + 2δ ) \ B(a, r − 2δ ))

k=1
c( 2δ )
K δ δ
= δ
σ(B(a, r + ρ + ) \ B(a, r − )).
c( 2 ) 2 2

87
Combine ADP,

#{Λ ∩ (B(a, r + ρ) \ B(a, r))}


lim sup sup
r→∞ a∈X σ(B(a, r))
K σ(B(a, r + ρ + 2δ ) \ B(a, r − 2δ ))
≤ lim sup sup
r→∞ a∈X c( 2δ ) σ(B(a, r))
K σ(B(a, r + ρ + 2δ ) \ B(a, r − 2δ ))
= lim sup sup
c( 2δ ) r→∞ a∈X σ(B(a, r))
K σ(B(a, r + ρ + 2δ ) \ B(a, r − 2δ ))
≤ lim sup sup
c( 2δ ) r→∞ a∈X σ(B(a, r − 2δ ))
K σ(B(a, r + ρ + δ) \ B(a, r))
= δ lim sup sup
c( 2 ) r→∞ a∈X σ(B(a, r))

=0.

The next proposition is very important for us. We will use it to verify

X 2
lim sup |hAgn , fm i| = 0,
r→∞ n∈N
d(λm ,λn )>r

and
X 2
lim sup |hBgn , fm i| = 0,
r→∞ n∈N
d(λm ,λn )>r

hold in Theorem 5.3.5.

Proposition 5.3.7. Let w : [0, ∞) → [0, ∞) be a weight function satisfies w(d) ↑ ∞ as d → ∞. Let

{hn }∞ ∞
n=1 be a sequence in H, and {fn }n=1 be a frame of H. If the following conditions hold

2
1. supn∈N khn k < ∞,

2. {hn }∞
n=1 is uniformly localized on Λ,

3. {fn }∞
n=1 is uniformly w-localized on Λ,

4. w is localized on Λ, and w(2d) . w(d) for every d,

then
X 2
lim sup |hhn , fm i| = 0.
r→∞ n∈N
d(λm ,λn )>r

88
Proof. Fix any λn ∈ Λ, and define the ball B := B(λn , 2r ) for some r > 0. And let TB , TB c be two

concentration operators on H given by

Z Z
TB f := hf, kx i kx dσ(x), T Bc f := hf, kx i kx dσ(x).
B Bc

Then for any f ∈ H,

Z
f= hf, kx i kx dσ(x)
ZX Z
= hf, kx i kx dσ(x) + hf, kx i kx dσ(x)
B Bc

= TB f + TB c f.

So the identity map I = TB + TB c . It follows that

2 2
|hhn , fm i| = |h(TB + TB c )hn , fm i|
2
= |hTB hn , fm i + hTB c hn , fm i|
2 2
≤ 2 |hTB hn , fm i| + 2 |hTB c hn , fm i| . (5.12)

Estimate the first term to the right in (5.12). By condition 1 and 3,

Z

|hTB hn , fm i| = hhn , kx i hkx , fm i dσ(x)
ZB Z
2 2
≤ |hhn , kx i| dσ(x) |hkx , fm i| dσ(x)
X B
Z
2 2 w(d(λm , x))
= khn k |hfm , kx i| dσ(x)
B w(d(λm , x))
Z
2 1 2
≤ sup khn k sup |hfm , kx i| w(d(λm , x))dσ(x)
n∈N x∈B w(d(λm , x)) X
1
. sup . (5.13)
x∈B w(d(λm , x))

89
Notice when x ∈ B(λn , r/2) and d(λm , λn ) > r, we have

d(λm , λn ) ≤ d(λm , x) + d(x, λn )

≤ d(λm , x) + r/2

≤ 2d(λm , x). (5.14)

Combine (5.13) and (5.14) , by condition 4 we obtain

X 2
lim sup |hTB hn , fm i|
r→∞ n∈N
d(λm ,λn )>r
X 1
. lim sup sup
r→∞ n∈N x∈B(λn , r2 ) w(d(λm , x))
d(λm ,λn )>r
X 1
≤ lim sup d(λm ,λn )
r→∞ n∈N
d(λm ,λn )>r w( 2 )
X 1
. lim sup
r→∞ n∈N w(d(λm , λn ))
d(λm ,λn )>r

=0. (5.15)

Estimate the second term to the right in (5.12). By {fn }∞


n=1 is a frame and {kx }x∈X is a

90
continuous Parseval frame,

X 2
lim sup |hTB c hn , fm i|
r→∞ n∈N
d(λm ,λn )>r
X 2
≤ lim sup |hTB c hn , fm i|
r→∞ n∈N
m
2
. lim sup kTB c hn k
r→∞ n∈N
Z 2

= lim sup hhn , kx i kx dσ(x)
r→∞ n∈N Bc

Z  2

= lim sup sup hhn , kx i kx dσ(x), g
r→∞ n∈N kgk=1 B c

Z 2

= lim sup sup hhn , kx i hkx , gi dσ(x)
r→∞ n∈N kgk=1 Bc
Z Z
2 2
≤ lim sup sup |hhn , kx i| dσ(x) |hkx , gi| dσ(x)
r→∞ n∈N kgk=1 Bc X
Z
2 2
= lim sup sup |hhn , kx i| dσ(x) kgk
r→∞ n∈N kgk=1 Bc
Z
2
= lim sup |hhn , kx i| dσ(x)
r→∞ n∈N Bc
Z
2
= lim sup |hhn , kx i| dσ(x)
r→∞ n∈N B(λn , r2 )c

=0, (5.16)

where the last equality comes from condition 2. Combining (5.12), (5.15) and (5.16), we have

X 2
lim sup |hhn , fm i|
r→∞ n
d(λm ,λn )>r
X 2 2
. lim sup |hTB hn , fm i| + |hTB c hn , fm i|
r→∞ n
d(λm ,λn )>r

=0.

91
5.3.4 Proof of Theorem 5.3.5

Proof. By conditions 1-4 of Theorem 5.3.5, applying Proposition 5.3.7, we have

X 2
lim sup |hAgn , fm i|
r→∞ n∈N
d(λm ,λn )>r
X 2
= lim sup |h(A − an )gn , fm i| = 0; (5.17)
r→∞ n∈N
d(λm ,λn )>r

and

X 2
lim sup |hBgn , fm i|
r→∞ n∈N
d(λm ,λn )>r
X 2
= lim sup |h(B − bn )gn , fm i| = 0. (5.18)
r→∞ n∈N
d(λm ,λn )>r

Notice we have different assumptions on Λ from Theorem 5.2.2, we cannot use Theorem 5.2.2 to

prove Theorem 5.3.5 directly. However, we could borrow the first half of the proof of Theorem 5.2.2.

Again, we assume
 
2 2
sup k(A − an )fn k + k(B − bn )fn k < ∞,
n∈N

and expect to derive a contradiction.

Let B := B(a, r) be any ball with the center a ∈ X and the radius r > 0, as in (5.4) we

have

X X
c#{Λ ∩ B} ≤ |hBgn , fm i hgm , Afn i|
λn ∈Λ∩B λm ∈Λ∩B c
X X
+ |hAgn , fm i hgm , Bfn i| . (5.19)
λn ∈Λ∩B λm ∈Λ∩B c

Let ρ > 0 and denote B(a, r + ρ) by Bρ which is the ρ-neighbourhood of B. As in (5.5),

there exists some C > 0 such that

X X
|hBgn , fm i hgm , Afn i|
λn ∈Λ∩B λm ∈Λ∩B c
  12
1
X 2
≤C#{Λ ∩ B} 2 #{Λ ∩ B} |hBgn , fm i| + #{Λ ∩ (Bρ \ B)} . (5.20)
λm ∈Λ∩B(λn ,ρ)c

92
Similarly, as in (5.6), for some C > 0 we have

X X
|hAgn , fm i hgm , Bfn i|
λn ∈Λ∩B λm ∈Λ∩B c
  12
1
X 2
≤C#{Λ ∩ B} #{Λ ∩ B} 2 |hAgn , fm i| + #{Λ ∩ (Bρ \ B)} . (5.21)
λm ∈Λ∩B(λn ,ρ)c

Divide (5.19) by σ(B) on both sides. Combine (5.20) and (5.21), we have for every a ∈ X, r > 0

and ρ > 0,

c#{Λ ∩ B}
σ(B)
1 X X
≤ |hBgn , fm i hgm , Afn i|
σ(B)
λn ∈Λ∩B λm ∈Λ∩B c
1 X X
+ |hAgn , fm i hgm , Bfn i|
σ(B)
λn ∈Λ∩B λm ∈Λ∩B c
  21
2
#{Λ ∩ B} X 2 #{Λ ∩ B} #{Λ ∩ (Bρ \ B)} 
. sup |hBgn , fm i| +
σ(B)2 n∈N σ(B) σ(B)
λm ∈Λ∩B(λn ,ρ)c
  21
2
#{Λ ∩ B} X 2 #{Λ ∩ B} #{Λ ∩ (Bρ \ B)} 
+ sup |hAgn , fm i| + . (5.22)
σ(B)2 n∈N σ(B) σ(B)
λm ∈Λ∩B(λn ,ρ)c

Let ε > 0, by (5.17) and (5.18), we can find a ρ > 0 such that

X 2
X 2
sup |hBgn , fm i| < ε2 , sup |hAgn , fm i| < ε2 . (5.23)
n∈N n∈N
λm ∈Λ∩B(λn ,ρ)c λm ∈Λ∩B(λn ,ρ)c

So for such ρ, combine (5.22) and (5.23), we have

 21
#{Λ ∩ B}2 2 #{Λ ∩ B} #{Λ ∩ (Bρ \ B)}

c#{Λ ∩ B(a, r)}
. ε + ,
σ(B(a, r)) σ(B)2 σ(B) σ(B)

93
for any a ∈ X and r > 0. Then by Proposition 5.3.6, we obtain for such ρ

#{Λ ∩ B(a, r)}


cD+ (Λ) = c lim sup sup
r→∞ a∈X σ(B(a, r))
1
#{Λ ∩ B}2 2 #{Λ ∩ B} #{Λ ∩ (Bρ \ B)} 2

. lim sup sup ε +
r→∞ a∈X σ(B)2 σ(B) σ(B)
1
#{Λ ∩ B}2 2

#{Λ ∩ B} #{Λ ∩ (Bρ \ B)} 2
≤ lim sup sup ε + lim sup sup
r→∞ a∈X σ(B)2 r→∞ a∈X σ(B) σ(B)
  12
#{Λ ∩ (Bρ \ B)}
≤ D+ (Λ)2 ε2 + D+ (Λ) lim sup sup
r→∞ a∈X σ(B)
1
= D+ (Λ)2 ε2 + 0 2


= εD+ (Λ).

Since Λ is relatively separated with D+ (Λ) > 0, by Proposition 3.1, we also have D+ (Λ) < ∞. By

ε is arbitrarily small, contradiction completes.

5.3.5 Applications

Now we could give another proof of the main result in [6] applying Theorem 5.3.5.

Theorem 5.3.8 ([6], Lemma 3). Assume that {fn }∞ 2


n=1 is a Riesz basis for L (R) with the biorthog-

onal basis {gn }∞ ∞ ∞


n=1 . If there exist sequences of real numbers {an }n=1 , {bn }n=1 such that

2
2s 2 2s b
dξ ≤ S 2 < ∞ for every n,
R R
(a) |x − an | |fn (x)| dx + |ξ − b | f (ξ)

R R n n

2s 2 2s 2
gn (ξ)| dξ ≤ T 2 < ∞ for every n,
R R
(b) R
|x − an | |gn (x)| dx + R
|ξ − bn | |b

(c) Λ := {(an , bn )}∞ 2 +


n=1 ⊆ R is relatively separated with 0 < D (Λ) < ∞.

Then s ≤ 1.

Proof. Let (L2 (R), {Mb Ta g}(a,b)∈(R2 ,|·|,m) ) be our FHS which satisfies ADP and UMD, where g(x) :=
π 2
e− 2 x is normalized Gaussian function. And multiplication X and differentiation D will play the

role of operators A and B in Theorem 5.3.5 respectively.

Again, we will prove the theorem by contradiction. Suppose all the above assumptions

(a),(b),(c) hold and s > 1. Our proof strategy is the following: We will verify the following conditions

1-4 of Theorem 5.3.5 hold. Namely, there exists a weight function w such that

94
 
2 2
1. supn∈N k(X − an )gn k + k(D − bn )gn k < ∞,

2. {(X − an )gn }∞ ∞
n=1 and {(D − bn )gn }n=1 are uniformly localized on Λ,

3. {fn }∞
n=1 is uniformly w-localized on Λ,

4. w is localized on Λ, and w(2d) . w(d) for every d.

Meanwhile, we will show

 
2 2
sup k(X − an )fn k + k(D − bn )fn k < ∞.
n∈N

Then Theorem 5.3.5 will give us the desired contradiction.

Let w(d) := d2s be the power weight function with exponent 2s. Obviously, w(2d) . w(d)

for every d. Since Λ = {(an , bn )}∞ 2


n=1 ⊆ R is relatively separated and s > 1, by the same argument

of (5.11) in Section 5.3.1, the weight function w is localized on Λ. So condition 4 holds.

By assumption (a),

Z
2 2 2
k(X − an )fn k = |x − an | |fn (x)| dx
R
Z
2s 2
≤ (|x − an |
+ 1) |fn (x)| dx
R
Z Z
2s 2 2
= |x − an | |fn (x)| dx + |fn (x)| dx
R R
2
≤ S + C,

where C is the Riesz upper bound of {fn }∞


n=1 . And

2 2
k(D − bn )fn k = kF(D − bn )fn k
2
= k(X − bn )Ffn k
Z 2
2
= |ξ − bn | fbn (ξ) dξ

R
Z 2
2s
≤ (|ξ − bn | + 1) fbn (ξ) dξ

R

≤ (S 2 + C).

95
So we have
 
2 2
sup k(X − an )fn k + k(D − bn )fn k < ∞. (5.24)
n∈N

Similarly by assumption (b), condition 1 holds, i.e.,

 
2 2
sup k(X − an )gn k + k(D − bn )gn k < ∞.
n∈N

Denote Mb Ta g by gb,a for any a, b ∈ R. Notice

Z
2 2s
|hfn , gb,a i| |a − an | dadb
R2
Z
2 2s
= |F(fn Ta g)(b)| |a − an | dbda
R2
Z
2 2s
= |(fn Ta g)(b)| db |a − an | da
R2
Z
2 2 2s
= |fn (b)| |g(b − a)| |a − b + b − an | dbda
R2
Z  
2 2 2s 2s
. |fn (b)| |g(b − a)| |a − b| + |b − an | dadb
R2
Z Z
2 2s 2
= |g(x)| |x| dx |fn (x)| dx
R R
Z Z
2 2 2s
+ |g(x)| dx |fn (x)| |x − an | dx. (5.25)
R R

Denote M−a Tb gb by gb−a,b for any a, b ∈ R, similarly,

Z
2 2s
|hfn , gb,a i| |b − bn | dadb
R2
Z D E 2
2s
= fn , gb−a,b |b − bn | dadb
b
R2
Z Z 2
2 2s
. |b g (ξ)| |ξ| dξ fbn (ξ) dξ

ZR Z R 2
2 2s
+ |b g (ξ)| dξ fbn (ξ) |ξ − bn | dξ. (5.26)

R R

96
Combine (5.25) and (5.26), by g ∈ S(R) and assumption (a), we have

Z
2
sup |hfn , gb,a i| d2s ((a, b), (an , bn ))dadb
n∈N R2
Z  s
2 2 2
= sup |hfn , gb,a i| |a − an | + |b − bn | dadb
n∈N R2
Z  
2 2s 2s
. sup |hfn , gb,a i| |a − an | + |b − bn | dadb
n∈N R2

<∞. (5.27)

Which implies condition 3 holds. Similarly by assumption (b), we have

Z
2
sup |hgn , gb,a i| d2s ((a, b), (an , bn ))dadb < ∞. (5.28)
n∈N R2

In order to verify condition 2, let kb,a := Mb Ta Xg be the continuous Gabor frame generated by the

new window function Xg. Notice

Z
2
|h(X − an )gn , gb,a i| dadb
B((an ,bn ),r)c
Z
2
= |hgn , (X − an )gb,a i| dadb
B((an ,bn ),r)c
Z
2
= |hgn , (X − an )Mb Ta gi| dadb
B((an ,bn ),r)c
Z
2
= |hgn , (a − an )Mb Ta g + Mb Ta Xgi| dadb
B((an ,bn ),r)c
Z
2
≤2 |hgn , (a − an )Mb Ta gi| dadb
B((an ,bn ),r)c
Z
2
+2 |hgn , Mb Ta Xgi| dadb
B((an ,bn ),r)c
Z
2 2
=2 |hgn , gb,a i| |a − an | dadb
B((an ,bn ),r)c
Z
2
+2 |hgn , kb,a i| dadb. (5.29)
B((an ,bn ),r)c

97
Estimate the first term in (5.29),

Z
2 2
|hgn , gb,a i| |a − an | dadb
B((an ,bn ),r)c
Z  
2 2 2
≤ |hgn , gb,a i| |a − an | + |b − bn | dadb
B((an ,bn ),r)c
Z
2
= |hgn , gb,a i| d2 ((a, b), (an , bn ))dadb
B((an ,bn ),r)c
d2s ((a, b), (an , bn ))
Z
2
= |hgn , gb,a i| dadb
B((an ,bn ),r)c d2s−2 ((a, b), (an , bn ))
Z
2−2s 2
≤r |hgn , gb,a i| d2s ((a, b), (an , bn ))dadb
B((an ,bn ),r)c
Z
2
≤r2−2s |hgn , gb,a i| d2s ((a, b), (an , bn ))dadb.
R2

By (5.28), we have

Z
2 2
lim sup |hgn , gb,a i| |a − an | dadb = 0. (5.30)
r→∞ n∈N B((an ,bn ),r)c

Estimate the second term in (5.29),

Z
2
|hgn , kb,a i| dadb
B((an ,bn ),r)c
d2s ((a, b), (an , bn ))
Z
2
= |hgn , kb,a i| dadb
B((an ,bn ),r)c d2s ((a, b), (an , bn ))
Z
2
≤r−2s |hgn , kb,a i| d2s ((a, b), (an , bn ))dadb
B((an ,bn ),r)c
Z
2
≤r−2s |hgn , kb,a i| d2s ((a, b), (an , bn ))dadb.
R2

Repeat the same argument of (5.25), (5.26) and (5.27), by Xg ∈ S(R) and assumption (b), we

obtain
Z
2
sup |hgn , kb,a i| d2s ((a, b), (an , bn ))dadb < ∞.
n∈N R2

So we have
Z
2
lim sup |hgn , kb,a i| dadb = 0. (5.31)
r→∞ n∈N B((an ,bn ),r)c

98
Combine (5.29), (5.30) and (5.31),

Z
2
lim sup |h(X − an )gn , gb,a i| dadb = 0.
r→∞ n∈N B((an ,bn ),r)c

Similarly,
Z
2
lim sup |h(D − bn )gn , gb,a i| dadb = 0.
r→∞ n∈N B((an ,bn ),r)c

Which imply condition 2 holds as well. Since we collect all the conditions 1-4, we could apply

Theorem 5.3.5, and obtain

 
2 2
sup k(X − an )fn k + k(D − bn )fn k = ∞.
n∈N

It contradicts with (5.24)!

99
Acknowledgment

First, I would like to thank my advisor, Dr. Mishko Mitkovski. Without his guidance and

knowledge, I couldn’t have done it. Thank you for leading me to begin my journey of analysis, and

also thank you for providing suggestions and helps when I met a problem in daily life.

Besides, I want to say thank you to all of my professors at Clemson, especially to: Dr.

Shitao Liu, Dr. Martin Schmoll, Dr. Jeong-Rock Yoon, and Dr. Brian Fralix. Thank all of you

for teaching me with great patience. Thank you again to Dr. Liu for sharing your experience of

studying abroad and treating me more like a friend. It was a huge support and encouragement to

me.

Finally, I would like to express my gratitude to my family members. Thank you to my

sister Xiaoming Wang for taking care of me all the time. Thank you to my parents. You give me

everything you can give. I feel sorry about not being at home for a long time. Because of you, I

could play math for fun.

100
Bibliography

[1] John J Benedetto, Wojciech Czaja, Przemystaw Gadziński, and Alexander M Powell, The balian-low theorem
and regularity of gabor systems, The Journal of Geometric Analysis 13 (2003), no. 2, 239. ↑7
[2] John J Benedetto, Christopher Heil, and David F Walnut, Differentiation and the balian-low theorem, Journal
of Fourier Analysis and Applications 1 (1994), no. 4, 355–402. ↑6
[3] Stephen M Buckley, Is the maximal function of a lipschitz function continous?, Annales academiæ scientiarum
fennicæ mathematica, 1999, pp. 519–528. ↑28
[4] Ole Christensen et al., An introduction to frames and riesz bases, Springer, 2016. ↑8
[5] Gerald B Folland and Alladi Sitaram, The uncertainty principle: a mathematical survey, Journal of Fourier
analysis and applications 3 (1997), no. 3, 207–238. ↑4
[6] Karlheinz Gröchenig and Eugenia Malinnikova, Phase space localization of riesz bases for lˆ 2 (rˆ d), arXiv
preprint arXiv:1102.3097 (2011). ↑76, 94
[7] Werner Heisenberg, Über den anschaulichen inhalt der quantentheoretischen kinematik und mechanik, Original
scientific papers wissenschaftliche originalarbeiten, 1985, pp. 478–504. ↑ii
[8] Jean-Pierre Kahane, Sur les fonctions moyenne-périodiques bornées, Annales de l’institut fourier, 1957, pp. 293–
314. ↑38
[9] Oleg Kovrijkine, Some results related to the logvinenko-sereda theorem, Proceedings of the American Mathemat-
ical Society 129 (2001), no. 10, 3037–3047. ↑58, 59
[10] HJ Landau et al., Necessary density conditions for sampling and interpolation of certain entire functions, Acta
Mathematica 117 (1967), 37–52. ↑37, 45
[11] Haodong Li, Multiplier operator on framed hilbert spaces (2015). ↑34, 35, 36
[12] Niklas Lindholm, Sampling in weighted lp spaces of entire functions in cn and estimates of the bergman kernel,
Journal of Functional Analysis 182 (2001), no. 2, 390–426. ↑38
[13] Mishko Mitkovski and Aaron Ramirez, Density results for continuous frames, Journal of Fourier Analysis and
Applications 26 (2020), no. 4, 1–26. ↑39
[14] Fedor L’vovich Nazarov, Local estimates for exponential polynomials and their applications to inequalities of the
uncertainty principle type, Algebra i analiz 5 (1993), no. 4, 3–66. ↑60, 61, 62
[15] Alexander Olevskii and Alexander Ulanovskii, Approximation of discrete functions and size of spectrum, St.
Petersburg Mathematical Journal 21 (2010), no. 6, 1015–1025. ↑41, 45, 46
[16] Alexander P Schuster and Kristian Seip, Weak conditions for interpolation in holomorphic spaces, Publicacions
matematiques (2000), 277–293. ↑50
[17] Kristian Seip, Density theorems for sampling and interpolation in the bargmann-fock space, Bulletin of the
American Mathematical Society 26 (1992), no. 2, 322–328. ↑38
[18] Kristian Seip and Robert Wallstén, Density theorems for sampling and interpolation in the bargmann-fock space.
ii, J. reine angew. Math 429 (1992), 107–113. ↑38
[19] Norbert Wiener, I am a mathematician, the later life of a prodigy: An autobiographical account of the mature
years and career of norbert wiener and a continuation of the account of his childhood in ex-prodigy, Vol. 20,
Doubleday, 1956. ↑ii
[20] Robert M Young, An introduction to nonharmonic fourier series, Academic press, 1981. ↑8

101

You might also like