Harmonic Analysis
Harmonic Analysis
Harmonic Analysis
Jason Murphy
Missouri University of Science and Technology
Contents
1 Introduction 5
2
CONTENTS 3
5 Wavelet transforms 99
5.1 Continuous wavelet transforms . . . . . . . . . . . . . . . . . 99
5.2 Discrete wavelet transforms . . . . . . . . . . . . . . . . . . . 112
5.3 Multiresolution analysis . . . . . . . . . . . . . . . . . . . . . 123
5.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
Introduction
These notes were written to accompany the courses Math 6461 and Math
6462 (Harmonic Analysis I and II) at Missouri University of Science & Tech-
nology.
The goal of these notes is to provide an introduction into a range of top-
ics and techniques in harmonic analysis, covering material that is interesting
not only to students of pure mathematics, but also to those interested in
applications in computer science, engineering, physics, and so on. We will
focus on giving an overall sense of the available results and the analytic tech-
niques used to prove them; in particular, complete generality or completely
optimal results may not always be pursued. Technical details will sometimes
be left to the reader to work out as exercises; solving these exercises is an
important part of solidifying the reader’s understanding of the material. At
times we will not develop the full theory but rather give a survey of results,
along with citations to references containing full details.
These notes are organized as follows:
5
6 CHAPTER 1. INTRODUCTION
The material from these notes has been drawn from many different
sources. In addition to the references listed in the bibliography, this in-
cludes the author’s personal notes from a harmonic analysis course given by
M. Visan at UCLA.
The author gratefully acknowledges support from the University of Mis-
souri Affordable and Open Educational Resources Initiative Award, as well
as the students of Math 6461-6462 for their useful feedback and corrections.
Chapter 2
u(t, x) = p(t)q(x).
The equation for p is solvable for any λ; indeed, p(t) = e−λt p(0) does the
job. The problem for q is more interesting, since in addition to the ordinary
8
2.1. SEPARATION OF VARIABLES 9
where
inπx
en (x) := e L .
The Fourier coefficients cn are given by
Z L
1
cn = hF, en i = 2L F (x)ēn (x) dx.
−L
We may also write
cn = F̂ (n).
10 CHAPTER 2. FOURIER ANALYSIS, PART I
Using this result, one can recover sine series and cosine series, which
are often used in the solution of PDEs via separation of variables (see the
exercises).
Note that we may identify periodic functions on [−L, L] with functions
on the torus or on the circle, which we may denote by TL .
Remark
P 2.2.2. If one supposes that there is a decomposition of the form
f = k ck φk , one can already see that we should have ck = hf, φk i by taking
the inner product of both sides with φk and using orthonormality.
Remark 2.2.3. The space L2 admits complete orthonormal sets. To see
this, first take a countable dense subset (L2 is separable), and then apply
the Gram–Schmidt algorithm from linear algebra to find an orthonormal
basis for L2 . This means an orthonormal set whose span is dense in L2
(where span means the collection of all finite linear combinations). Using
Cauchy–Schwarz, one can then prove that any orthogonal basis for L2 must
be complete (see the exercises).
The key property of Fourier series is that they give the best L2 approx-
imation using linear combinations of the {φk }.
Lemma 2.2.4. Let {φk } be an orthonormal set in L2 and f ∈ L2 .
(i) Given N , the best L2 approximation to f using the φk is given by the
partial Fourier series.
(ii) (Bessel’s inequality) We have c := {ck } ∈ `2 and
kck`2 ≤ kf kL2 ,
It follows that
N
X
min kf − F (γ)k2 = kf k2 − |ck |2
γ
k=1
and
argminγ kf − F (γ)k2 = (c1 , · · · , cN ).
This proves (i). Furthermore (evaluating at γ = (c1 , . . . , cN )) we can deduce
N
X
|ck |2 = kf k2 − kf − sN k2 ,
k=1
This result does not guarantee uniqueness. However, one does have
uniqueness if the set {φk } is complete. Indeed, if f and g have the same
Fourier coefficients then f − g is perpendicular to each φk .
Finally, we turn to the following:
Proof of Theorem 2.2.1. Theorem 2.2.1 now follows from the combination
of Proposition 2.2.5 and Proposition 2.2.7.
One can consider even more general settings and prove similar results
in the setting of abstract Hilbert spaces. However, at this point we will
return to the more specific setting of Fourier series for periodic functions on
[−L, L].
where (
1 n=m
δnm =
0 otherwise.
14 CHAPTER 2. FOURIER ANALYSIS, PART I
One calls δnm the Kronecker delta. Thus the family { √12π en } is orthonor-
mal.
It remains to prove that this set is complete.
We can rewrite
SN f (x) = f ∗ DN (x),
where DN is the Dirichlet kernel given by
N
X
1
DN (x) = 2π en (x).
n=−N
1 sin([N + 12 ]x)
DN (x) = 2π (2.2)
sin( 12 x)
(see the exercises). One can verify, for example, that the L1 -norm of DN
grows like log N . (Again, see the exercises.)
As we will see, averaging improves the situation. In particular, if we
define the Cesáro means by
N
X −1
1
σN f = N Sn f,
n=0
1 [sin( N2 x)]2
FN (x) = 2πN ,
[sin( 12 x)]2
as well. Finally, we fix δ > 0 and observe that | sin( 21 x)| & δ for |x| > δ.
Thus, using the identity above for example, we find
Z
|FN (x)| dx . N1δ2 → 0 as N → ∞.
|x|>δ
where f (x±) denotes limits from the right/left (see the exercises). In par-
ticular, σN f (x) will converge to f (x) at any point of continuity.
16 CHAPTER 2. FOURIER ANALYSIS, PART I
1 sin([N + 21 ]x)
S N f = f ∗ DN , DN (x) = 2π ,
sin( 12 x)
where we recall that
kDN kL1 & log N.
P −1
We also recall the Fejér kernels FN = N1 Nn=0 Dn .
(i) For each n, we may view Sn as a linear operator from L1 to L1 . We
define the operator norm of Sn by
Recalling that the Fejér kernels FN are uniformly bounded in L1 (by 1), we
have
kSn (FN )kL1 ≤ kSn kL1 →L1 kFN kL1 ≤ kSn kL1 →L1 .
On the other hand, Sn (FN ) = σN (Dn ) (check!), which yields
by showing that
kSn kL∞ →L∞ & kDn kL1 .
We let ψn (x) = signum[Dn (x)], except in small intervals around the 2n
points of discontinuity of signum[Dn (x)]. In particular, we can make ψn be
continuous, with
kψn kL∞ ≤ 1 uniformly in n.
Fixing ε > 0 and choosing the total length of the small intervals {Ik }2n
k=1 to
be smaller than ε/2n, we may derive that
Z
kSn ψn kL∞ ≥ |Sn ψn (0)| & Dn (x)ψn (x) dx & kDn kL1 − ε.
Indeed, we have
Z Z 2n Z
X
Dn ψn (x) dx = |Dn (x)| dx + Dn (x)[ψn (x) − signum[Dn (x)] dx
k=1 Ik
where f (x±) denotes limits from the right/left. In particular, Sn f (x) con-
verges to f (x) at any point of continuity of f .
18 CHAPTER 2. FOURIER ANALYSIS, PART I
with gK (0) ≡ 0 as expected. However we can see that |gK 0 (0)| = K, which
suggests that this series can reach size 1 even over an interval of length 1/K.
In fact, more detailed analysis (or studying this example in Mathemat-
ica) would reveal that (for all large K) gK (ε) reaches size about −.92 for
ε = 0+ and +.92 for ε = 0−, and hence we see that the Fourier series will
undershoot/overshoot the correct value of the function by a fixed amount,
yielding approximate values
a a 4
2 ∓ 2 π · .92)
Proof of Proposition 2.4.2. As we will see, the key is to establish some decay
for the Fourier coefficients fˆ(n) of f .
In fact, we claim that
|fˆ(n)| . 1
|n| kf kBV for |n| ≥ 1. (2.5)
Indeed, this follows from the integration by parts formula for Riemann–
Stieltjes integrals (see Proposition A.1.3):
Z
ˆ −inx
1
|f (n)| = 2π e
f (x) dx
Z
−inx
1 1
= 2πin e df (x) . |n| kf kBV .
ασn+1 f + (1 − α)σm+1 f
will produce a single copy of Sn f plus some ‘error terms’. As any such
combination converges to the desired limit as n, m → ∞, we can complete
the proof if we can find a suitable combination for which we can control the
error terms.
The sums involving |k| ≤ n are the most problematic, as they do not
tend to zero individually; for example, applying (2.5), the best estimate we
have for the term appearing in σn+1 yields
X |k|
ˆ n
n+1 |f (k)| . n+1 ,
|k|≤n
20 CHAPTER 2. FOURIER ANALYSIS, PART I
Our final step is to show that for m chosen suitably depending on n, this
final term can be made arbitrarily small (uniformly as n → ∞). In fact,
using (2.5), we can estimate
|k|
X X
m+1
|(1 − )fˆ(k)| . 1 m
m−n m+1 |k| . log( n ),
n<|k|≤m n<|k|≤m
Given ε > 0, the result now follows by choosing m to be the integer part of
(1 + δ)n for small δ = δ(ε) > 0.
n=−∞
Remark 2.5.1. Recall the viewpoint that the Fourier series of a function
is an expansion in terms of eigenfunctions of −∂x2 (or, in higher dimensions,
the Laplacian). Here we see that the Fourier transform of a function is an
expansion in terms of generalized eigenfunctions of −∂x2 on R (or, in higher
dimensions, the Laplacian). The difference is that the Laplacian has discrete
eigenvalues (or spectrum) on a compact domain, while it has continuous
spectrum on Rd . Spectral theory allows for a unified interpretation of Fourier
series/transform, namely, as a spectral resolution of the Laplacian. We
will also see a unification of Fourier series and Fourier transform through
the perspective of Fourier analysis on locally compact abelian groups in
Chapter 4.
Remark 2.5.2. There are other normalizations for the Fourier transform.
A common one is to define
Z
f (ξ) = e−2πixξ f (x) dx.
ˆ
We will use this normalization in the next chapter and at times below.
This lemma already suggests the connection between the Fourier trans-
form and partial differential equations (PDE): it interchanges taking deriva-
tives and multiplication by x. We leave the following corollary as an exercise:
Corollary 2.5.5. If f ∈ S(Rd ) then fˆ ∈ S(Rd ).
Thus, defining the transformation F which takes f ∈ S(Rd ) and returns
fˆ ∈ S(Rd ), we can see that F (also called the Fourier transform) is a well-
defined linear transformation on S. Linearity is straightforward to check and
is left to the reader. In fact, more is true:
24 CHAPTER 2. FOURIER ANALYSIS, PART I
Proof. Let us verify the third identity and leave the first two as exercises.
This follows from the change of variables formula. Indeed,
Z Z
iyξ
e−ixξ f (λx) dx = λ1d e− λ f (y) dy.
Rd Rd
Proof. Let us prove this identity for the case d = 1. We leave it to the reader
to see why this implies the general case.
We make an ODE argument. Using the fact that f 0 (x) = −xf (x) and
d ˆ 2
Lemma 2.5.4, we find that dξ f = −ξ fˆ. Therefore fˆ(ξ) = e−ξ /2 fˆ(0). How-
ever, Z
− 12 2
ˆ
f (0) = (2π) e−x /2 dx = 1.
R
Thus the result follows.
2.5. THE FOURIER TRANSFORM 25
d 2 /2
In the following, we will define K(x) = (2π)− 2 e−|x| and for ε > 0 set
Kε (x) = ε−d K( xε ).
Proof of Theorem 2.5.6. We begin with the inversion formula for f ∈ S(Rd ).
Using the lemmas above, we compute
Z
f ∗ Kε (x) = f (x − y)Kε (y) dy
Rd
Z
= f (x − y)Ĝε (y) dy
Rd
Z Z
−ixy ˆ
= e f (−y)K(εy) dy = fˆ(y)eixy K(−εy) dy.
Rd Rd
While the Schwartz space is clearly well-suited for the Fourier transform,
it is not the end of the story. Given what we have learned about Fourier
series, it is natural to seek an extension of the Fourier transform to L2 (Rd ).
The key to this is the Plancherel formula. Before we state and prove it,
let us recall the inner product structure on L2 (Rd ), namely,
Z p
hf, gi = f (x)ḡ(x) dx, and kf kL2 = hf, f i.
Rd
26 CHAPTER 2. FOURIER ANALYSIS, PART I
as desired.
In fact, the Fourier transform extends from S(Rd ) to Lp (Rd ) for all
1 ≤ p ≤ 2, with a corresponding bound (known as the Hausdorff–Young
inequality):
In fact, this is the only range for which this is possible. That is, if the
estimate
kfˆkLq . kf kLp
holds for all Schwartz functions, then q = p0 and 1 ≤ p ≤ 2. We will discuss
the Hausdorff–Young inequality later in the setting of interpolation; as for
the second point, we leave it as an exercise.
In particular, this discussion may serve in part to preview some of the topics
to be covered later in these notes.
Recall (from Theorem 2.5.6) that we have the Fourier inversion formula
Z N
f (x) = lim √12π eixξ fˆ(ξ) dξ, x ∈ R, (2.7)
N →∞ −N
for all Schwartz functions f ∈ S(R). Carleson’s theorem [2, 9] states that the
same convergence holds almost everywhere for functions f that are merely
in L2 (R).
The integral appearing on the right-hand side of (2.7) can be written
in terms of the convolution of f with the Dirichlet kernel DN (x) = sinπxN x .
Thus, the convolution is a combination of the singular integral part x1 (which
would fall under the purview of Calderon–Zygmund theory, cf. Section 6.4
below) and the oscillatory part sin N x, which requires some additional tech-
niques.
Following [14], we will work with the (equivalent) one-sided inversion
formula Z N
f (x) = lim √ 1
2π
eixξ fˆ(ξ) dξ, (2.8)
N →∞ −∞
which again holds on R for any Schwartz function f . As Schwartz functions
are dense in L2 (Exercise 2.9.13), the problem reduces to proving that the
set of functions for which almost everywhere convergence holds is closed.
For such purposes, it is common to introduce a suitable maximal func-
tion. In particular, we define the Carleson operator
Z N
ixξ ˆ
Cf (x) = sup e f (ξ) dξ , x ∈ R.
N −∞
We can then show that a weak type (2, 2) bound for C implies the desired
result. (One can compare this to the fact that the weak type (1, 1) bound for
the Hardy–Littlewood maximal function implies the Lebesgue differentiation
theorem, for example; see Proposition 6.3.5.)
Proposition 2.5.14. Suppose that C obeys the weak type (2, 2) bound
Now, setting
N
Z
√1 ixξ ˆ
Lf = lim supf (x) − 2π
e f (ξ) dξ ,
N →∞ −∞
we note that since (2.8) holds for g (for all x), we may write
Lf ≤ C(f − g) + |f − g|.
Thus
|{Lf > ε}| . ε
for any ε > 0. It follows that Lf = 0 almost everywhere, which implies the
desired result.
Actually proving the weak type bound for the Carleson operator is quite
an undertaking. We again refer the interested reader to [14, 13] and end our
discussion of Carleson’s theorem here.
−∆u = f, u : R3 → R,
Applying the Fourier transform and Lemma 2.5.4, we find that the equation
is equivalent to
3
|ξ|2 û = fˆ, so that û = |ξ|−2 fˆ = (2π)− 2 F(K ∗ f )(ξ),
30 CHAPTER 2. FOURIER ANALYSIS, PART I
3
where K = F −1 (|ξ|−2 ). In particular, u(x) = (2π)− 2 K ∗ f (x).
Here we reach a bit of a subtle point: the function |ξ|−2 is not a Schwartz
function, nor is it even an L2 function! Let us ignore this subtlety for
the moment—it can be resolved by the theory of distributions (see below).
Instead, let us see if we can compute a formula for K anyway.
There is an elegant way to compute K(x) exactly using the Gamma
function (see the exercises). Let us instead argue by symmetry to deduce
the form of K(x). First, we observe by Lemma 2.5.8 that
in three dimensions.
Example 2.6.2 (Heat equation). We next consider the heat equation on
(0, ∞) × Rd :
(t, x) ∈ (0, ∞) × Rd ,
ut − ∆u = 0
u(0, x) = f (x) x ∈ Rd .
We apply the Fourier transform in the x variables only. We find
bt (t, ξ) = F(∆u)(t, ξ)
u ⇐⇒ bt (t, ξ) = −|ξ|2 u
u b(t, ξ).
2.7. THE FOURIER TRANSFORM OF DISTRIBUTIONS 31
Thus
2 2
u(t, x) = F −1 [fb e−t|ξ| ](x) = (2π)−d/2 [f ∗ F −1 (e−t|ξ| )](x),
As an exercise, one should check that the map T is injective, and hence we
can identify a function u with the distribution T u. In fact, the mapping T
makes sense for any function that is integrable against Schwartz functions,
32 CHAPTER 2. FOURIER ANALYSIS, PART I
δ0 (f ) = f (0) for f ∈ S.
This definition guarantees that û agrees with the usual definition of the
Fourier transform in the case that u actually arises from a Schwartz function
under the mapping T introduced above. Thus the Fourier transform F
extends to a mapping F : S 0 → S 0 .
Similarly, if we define F ∗ : S 0 → S 0 by F ∗ u(f ) = u(F −1 f ), we can
deduce that FF ∗ = F ∗ F = Id, and hence F ∗ = F −1 and F is a bijection
on S 0 .
To define operations on distributions, one first observes how these op-
erations behave on Schwartz functions. For example, given a multiindex α
and u, f ∈ S, integration by parts yields
Z Z
α |α|
(∂ u)f = (−1) u∂ α f.
(f ∗ u)(x) = u(τg
x f ), τx f (y) = f (y − x). (2.9)
d
Thus δ̂0 = (2π)− 2 .
Kλ (x) = λK(λx),
where Z 1
1 sin(x/2) 2
K(x) = 2π x/2 ) = 1
2π (1 − |ξ|)eixξ dξ
−1
(check!).
We set Z
Gλ (z) = Kλ ∗ F (z) = F (z − w)Kλ (w) dw.
R
Then Gλ is holomorphic in {|y| < a}. We now define
In particular
ĝλ,y (ξ) = K̂λ (ξ)fˆy (ξ) for each λ.
Now observe that each ĝλ,y (ξ) has compact support, specifically, in [−λ, λ].
In particular, (ii) holds for ĝλ,y and hence
2.9 Exercises
Exercise 2.9.1. The 1d Dirichlet Laplacian on (0, 1) is the operator −∂x2
defined on the set of smooth functions f : (0, 1) → R satisfying f (0) =
f (1) = 0. The 1d Neumann Laplacian is also defined to be −∂x2 but on the
set of smooth functions f (0, 1) → R satisfying f 0 (0) = f 0 (1) = 0.
(i) Find the eigenfunctions and eigenvalues for the Dirichlet Laplacian.
(ii) Find the eigenfunctions and eigenvalues for the Neumann Laplacian.
Exercise 2.9.2. Using Fourier series, show the following:
(i) For f : (0, 1) → R smooth and satisfying f (0) = f (1) = 0, we have
∞
X
f (x) = an sin(nπx) for some an ∈ R.
n=1
provided these limits exist. Hint: Use the facts that the Fejér kernels are
good kernels that are positive, even, and decay away from x = 0.
Exercise 2.9.8. Let f be a function on the torus with Fourier coefficients
fˆ(n).
(i) Show that if f is k-times differentiable and f (k) ∈ L1 , then
|fˆ(n)| . |n|−α .
Show that there exists a nonnegative function f ∈ L1 (T) such that the
Fourier coefficients of f are given by fˆ(n) = an .
2.9. EXERCISES 37
Exercise 2.9.11. Show that the Dirichlet kernels satisfy kDn kL1 & log n for
large n.
Exercise 2.9.12. Let f be a Schwartz function. Show that for any multii-
indices α, β and any 1 ≤ p ≤ ∞ we have xα ∂ β f ∈ Lp .
Exercise 2.9.13. Show that Schwartz space is dense in L2 .
Exercise 2.9.14. Show that if an estimate of the form
kfˆkLq . kf kLp
holds for Schwartz functions, then q = p0 and 1 ≤ p ≤ 2. Hint. For the first
2
part, use a scaling argument. For the second, consider f (x) = e−(1+it)|x| /2
and send t → ∞.
Exercise 2.9.15. Let A be a d × d invertible matrix with real entries. Show
that if g(x) = f (Ax), then
This appears in [28, Lemma 1, p.117]. It can be computed using the Gamma
function Z ∞
Γ(z) = e−t tz−1 dt,
0
as we now sketch. First, show that
Z ∞
2 d−α d−α
e−πt|x| t 2 dtt = π −( 2 ) Γ( d−α
2 )|x|
α−d
.
0
Now compute the Fourier transform, using the fact that the Fourier trans-
form of a Gaussian is a Gaussian (which follows from computing the appro-
priate Gaussian integral):
Z Z ∞ Z ∞
2 d−α π|ξ|2 α α
e−2πixξ e−πt|x| t 2 dtt dx = e− t t− 2 dtt = π − 2 |ξ|−α Γ( α2 ),
0 0
In this section, we continue our study of the Fourier transform and con-
sider several applied topics. We will use the following normalization for the
Fourier transform: Z
ˆ
f (ξ) = e−2πitξ f (t) dt,
R
39
40 CHAPTER 3. FOURIER ANALYSIS, PART II
Then
g(kr) = χ[−1,1] (kr) cos(2πk) = χ[−1,1] (kr) = f (kr)
for all k ∈ Z. Thus f and g are indistinguishable by sampling at the points
tk = kr.
The previous example hints at the general principle that in order to
faithfully represent a signal with sampling rate r−1 , the sample should not
contain frequencies higher than r−1 . (Actually, both signals above contain
arbitrarily high frequencies due to the fact that they are compactly sup-
ported in time.)
Let us turn to the positive result. We define the function
sin πx
sinc(x) := πx .
fˆ = χ[− 1 , 1 ] [Π 1 ∗ fˆ]
2r 2r r
and take the inverse Fourier transform. Here ΠN denotes the distribution
X
ΠN (ϕ) = ϕ(kN ),
k∈Z
Computing the inverse Fourier transform of χ[− 1 , 1 ] (see Lemma 3.1.3 be-
2r 2r
low) and Π 1 (see Exercise 3.5.1), one can deduce the formula appearing in
r
Theorem 3.1.1. See Exercise 3.5.2.
We turn to Theorem 3.1.1. Let us begin by recording the key property
of the sinc function.
Lemma 3.1.3. The function sinc belongs to L2 (R). Its Fourier transform
is given by
F[sinc](ξ) = χ[− 1 , 1 ] (ξ).
2 2
Proof. That sinc ∈ L2 follows from the fact that it is bounded near x = 0
and decays like |x|−1 for large x.
Next, we compute
Z 1/2
−1 sin(πx)
F [χ[− 1 , 1 ] ](x) = e2πixξ dξ = = sinc(x),
2 2
−1/2 πx
which yields the desired identity. The final identity then follows from apply-
ing the scaling and translation identities for the Fourier transform, obtained
by a change of variables.
H = {g ∈ L2 : ĝ is supported in 1 1
[− 2r , 2r ]}.
Proof. Let X
Φ(t) = ϕ(t − k).
k∈Z
Note that Φ(0) gives the left-hand side of the Poisson summation formula.
Next, observe that Φ(t) is periodic of period one. Thus it has a Fourier
series expansion X
Φ(t) = Φ̂(m)e2πimt .
m∈Z
We next prove a lemma from which Theorem 3.1.1 will follow directly.
h(x) = f (xr)e−2πiξxr
Thus, if fˆ is supported in [− 2r
1 1
, 2r ], it follows that ĝ = fˆ. This yields the
result.
i.e.
N −1
m X
Fd = f (nr)e−2πimn/N .
Nr
n=0
3.2. DISCRETE FOURIER TRANSFORM 45
fˆ = Ff ∈ CN
with entries
N
X −1
fˆ[m] = e−2πimn/N f [n].
n=0
ω = ωN = e2πi/N ,
Fx = δk for each k = 0, . . . , N − 1,
46 CHAPTER 3. FOURIER ANALYSIS, PART II
Ff · Fg = N (f · g),
P
where f · g = k f [k]ḡ[k] denotes the standard complex inner product on
CN .
Proof. We have
N
X −1 N
X −1
Ff · Fg = f [j]ḡ[k]ω −j · ω −k .
j=0 k=0
as desired.
F f [· − a] = ω −a Ff,
(3.2)
F ω a f = Ff [· − a].
F(f g) = N −1 Ff ∗ Fg,
and the desired identity follows from applying F. Now, an analogous com-
putation shows that
F −1 (a ∗ b) = N F −1 aF −1 b.
Thus
F −1 [N −1 Ff ∗ Fg] = f g,
which implies the second identity upon applying F.
The second term is treated similarly to produce FN/2 fo [m], except there
−m
appears an extra power of e−2πim/N = ωN .
Next consider FN f [m + N/2]. In this case we only need to observe that
e−2πi(m+N/2)(2k)/N = e−2πim(2k)/N e−2πik = e−2πim(2k)/N ,
which leads to the FN/2 fe [m] term again. On the other hand,
1
e−2πi(m+N/2)(2k+1)/N = e−2πim(2k+1)/N e−2πi(k+ 2 ) = −e−2πim(2k+1)/N ,
which accounts for the minus sign in the formula above. This completes the
proof.
For N = 2j , this lemma can be iterated until one is reduced to computing
F1 , which is trivial. Computing the discrete Fourier transform this way is
the fast Fourier transform algorithm.
We will look at the fast Fourier transform in more detail below. First,
let us see what computational advantage it has.
Proposition 3.3.2. Let F (N ) be the number of operations it takes to com-
pute the discrete Fourier transform with the fast Fourier transform algo-
rithm. Then
F (N ) ∼ N log N.
Recalling that computing the discrete Fourier transform using matrix
multiplication takes O(N 2 ) elementary operations (e.g. additions and mul-
tiplications), we see that the fast Fourier transform provides a huge compu-
tational advantage.
Proof. Using Lemma 3.3.1, we find that
F (N ) = 2F (N/2) + cN for some c > 0.
Rearranging, this yields
F (N ) F (N/2)
N = N/2 + c.
F (2j )
Assuming N = 2j and setting aj = 2j
, this simply reads
aj = aj−1 + c, so that aj = jc + a0 .
But a0 = F (1), the number of operations needed to compute the discrete
Fourier transform of a single point. In particular, a0 = 0, so that aj = jc.
Thus
F (N ) = cN log N,
as claimed.
3.3. FAST FOURIER TRANSFORM 51
Let us now look closer at the fast Fourier transform algorithm. As we will
see, this algorithm amounts to a factorization of FN . Recalling Lemma 3.3.1,
we need to define an operation that sorts a vector into its even and odd
indices.
Definition 3.3.3. For f ∈ CN with N even, let
π 01 f = π 1 π 0 f,
and so on. In particular, the input will always be a vector in CN , but the
j
input is a vector in CN/2 , where j is the number of digits appearing in
superscripts.
Next let us write IN for the N ×N identity matrix and ΩN for the N ×N
−m
diagonal matrix with ΩN [m, m] = ω2N . Then Lemma 3.3.1 may be written
as
FN/2 π 0 f
IN/2 ΩN/2
F N f = BN , BN = .
FN/2 π 1 f IN/2 −ΩN/2
Now we repeat the process, yielding
FN/4 π 00 f
FN/4 π 01 f
FN f = BN · diag(BN/2 )
,
FN/4 π 10 f
FN/4 π 11 f
where diag should always be taken with a suitable interpretation, and c(k)
denotes the unique sequence of log N binary digits corresponding to the
52 CHAPTER 3. FOURIER ANALYSIS, PART II
In particular,
p̂ ∗ fˆ = pf
c = 0,
3.4. COMPRESSED SENSING 55
for j = 0, . . . , N − 1.
Now observe (by considering the inverse Fourier transform) that p̂(k) is
the coefficient of p(t) appearing with e2πikt/N . In particular, p̂(0) = 1 and
(since p is a trigonometric polynomial of degree ≤ s) p̂(k) = 0 for k > s.
Rewriting (3.4) for s ≤ j ≤ 2s − 1 leads to
Now, the matrix on the left and the vector on the right are known quan-
tities. In particular, given f ∈ CN and knowledge of its first 2s Fourier
coefficients, we can write down the system (3.5), which must have at least
one solution (namely {p̂[k]}sk=1 ). However, this solution may not be unique.
In the following, we find some solution q̂ to (3.5), which we extend to CN
by setting q̂[0] = 1 and q̂[k] = 0 for k > s.
Then, reversing the steps above, we have q̂ ∗ fˆ[j] = 0 for s ≤ j ≤
2s − 1. In particular qf (component-wise product) has Fourier transform
vanishing on s consecutive indices. We claim that this implies qf ≡ 0. To
see this, we first note that since f = 0 outside of S, to compute the vector
of Fourier coefficients qfc [j] for s ≤ j ≤ 2s − 1 it suffices to multiply the
vector {qf [j]}j∈S by the s × s submatrix A of FN defined by choosing the
rows s through 2s − 1 of FN and the columns defined by indices in S. Thus
it remains to check that A is invertible.
To see this, note that A is of the form A[i, j] = ω −(s+i)nj , where i =
0, . . . , s − 1 and S = {nj : j = 0, . . . , s − 1}. In particular, after factoring
out ω −snj from each column, A is the transpose of a Vandermonde matrix
56 CHAPTER 3. FOURIER ANALYSIS, PART II
• Solve the linear system that produces the first 2s Fourier coefficients
of f from the unknowns f [j], j ∈ S.
The results just discussed suggest that in general, one should expect
that knowledge of a signal’s Fourier coefficients on a set Ω should suffice
to construct signals with support of around the same size as Ω. Strictly
speaking, this is not true.
√
Example 3.4.1. √ Suppose N is a perfect square. Define f by f [j√ N ] = 1 for
j = 0, 1, . . . N − 1 and f [k] = 0 otherwise. Then kf k`0 = N . Let us
compute the Fourier transform. We have
√
N
X −1 N −1
X √
−2πi`k/N
fˆ[k] = e f [`] = e−2πijk/ N
.
`=0 j=0
√ √
Now, if k = p N for some p, then the sum yields N . Otherwise, summing
the geometric series (as we did when computing the inverse Fourier transform
in general) yields zero. Thus
√
fˆ = N f.
3.4. COMPRESSED SENSING 57
In √
particular, we may√ choose Ω to be the set of √
frequencies precisely avoiding
{p N : p = 0, . . . , N − 1}. Then |Ω| = N − N , but knowing the Fourier
coefficients
√ on Ω cannot distinguish f (supported on a set of much smaller
size N ) from the zero signal.
The result that we will present essentially restores the intuition intro-
duced above (which was just shown to be wrong, strictly speaking). The
key is that one must incorporate a probabilistic viewpoint.
Before stating the result, let us first note that the reconstruction problem
under consideration is equivalent to the following `0 minimization problem:
Supposing we also want to avoid the case when |Ω| is comparable to N (so
that |Ω| 12 N ), we can view log(1 − 2|Ω| 2|Ω|
N ) as comparable to − N , whence
√
|Ω| &M N log N ∼M |S| log N.
f [n]
sgn(f )[n] = for n ∈ S,
|f [n]|
• supp P̂ ⊂ Ω,
Then if FS→Ω is injective, then the minimizer to (P1 ) is unique and equals
f . (Conversely, if f is the unique minimizer of (P1 ) then there exists P as
above.)
Proof of Proposition 3.4.8. Let us prove only the forward direction, which
is most directly useful for us.
Suppose such a vector P exists. Suppose that g satisfies ĝ|Ω = fˆ|Ω . Set
h = g − f . Then on S we have
Outside of S, we have
since |P | < 1.
It follows that
N
X −1
kgk`1 ≥ kf k`1 + Re h[n]P̄ [n] .
n=0
where the notation means the following. First, ∗ denotes adjoint (i.e. conju-
gate transpose). In particular, recalling that e is extension by zero, we have
that e∗ : `2 (CN ) → `2 (S) is simply restriction to S.
If we can define such P (given S and Ω), then P automatically has
Fourier support in Ω. Furthermore, we claim that e∗ P = e∗ sgn(f ). In fact,
this follows from
Thus, fixing f and its support S, the proof of Theorem 3.4.5 boils down
to proving that if Ω is chosen uniformly at random from sets of size &M
|S| log N , then
1. The operator FS→Ω is injective with probability 1 − O(N −M ), and
Assuming these two results, let us complete the proof of Theorem 3.4.5.
Proof of Theorem 3.4.5, assuming Propositions 3.4.9 and 3.4.10. Let F (Ω)
be the event that no polynomial P exists as in Proposition 3.4.8 if we choose
the set Ω of Fourier coefficients.
Let Ω be of size Nω drawn uniformly, and Ω0 be drawn according to the
Bernoulli model with τ = NNω . Writing Ωk for a set of frequencies chosen
uniformly at random with |Ωk | = k, we have
N
X N
X
P(F (Ω0 )) = P(F (Ω0 ) : |Ω0 | = k) P(|Ω0 | = k) = P(F (Ωk )) P(|Ω0 | = k).
k=0 k=0
Now note that P(F (Ωk )) is decreasing in k (it only gets easier to reconstruct
using larger sets). We also claim that
P(|Ω0 | ≤ τ N ) > 12 ,
which follows from the fact that τ N is an integer and hence the median of
the random variable |Ω0 |. Thus
Nω
X Nω
X
P(F (Ω0 )) ≥ P(F (Ωk )) P(|Ω0 | = k) ≥ P(F (Ω)) P(|Ω0 | = k) ≥ 1
2 P(F (Ω)).
k=1 k=1
and set H0 = e∗ H. Writing IS for the identity operator on `2 (S) (so that
e∗ e = IS ), we have
1 1 ∗
e− |Ω| H = |Ω| FCN →Ω FS→Ω ,
1 1 ∗
IS − |Ω| H0 = |Ω| FS→Ω FS→Ω .
3.4. COMPRESSED SENSING 63
To prove the upper bound in Proposition 3.4.10, we will also need esti-
mates on H. The crucial estimate will be the following:
uniformly on S.
Assuming these moment bounds for now, let us complete the proof of
Proposition 3.4.9 and (3.4.10). Then, finally, we will prove the moment
bounds and thereby complete the proof of Theorem 3.4.5.
for some 0 < c < 1 (see Exercise 3.5.5). Recall that E{|Ω|} = τ N . We first
deal with the probability that |Ω| is far from its expectation.
We will use a standard large deviation estimate, namely
t 2
P{|Ω| < E(|Ω|) − t} ≤ exp{− 2 E(|Ω|) }
64 CHAPTER 3. FOURIER ANALYSIS, PART II
yields
Next, let AM denote the event {kH0 k ≥ τ√N2 }. We would like to bound
P{AM }. For this we will rely on Lemma 3.4.11 and the fact that (since H0
is self-adjoint)
kH0 k2n = kH0n k2 ≤ tr(H02n ).
τN
See Exercise (3.5.6). In particular, for any n ≤ 4|S|(1−τ ) , we use Tchebychev
and Lemma 3.4.11 to estimate
P{kH0 k ≥ τ√N
2
} = P{kH0 k2n ≥ ( τ√N2 )2n }
≤ P{tr(H02n ) ≥ ( τ√N2 )2n }
2n
≤ (τ N )2n
E{tr(H02n )}
2n n n+1
≤ (τ N )2n
4
2( e(1−τ )) n (τ N )n |S|n+1 .
Recalling the assumed lower bound τ N &M |S| log N , we see that we may
choose n ∼M log N , say n = (M + 1) log N . Choosing constants appropri-
ately, the upper bound above is of the form
8n|S| n −n
2n τ N (1−τ ) |S|e . |τ N |N −(M +1) . N −M .
τ√N |Ω|
kH0 k ≤ 2
≤ √
2(1−εM )
< c|Ω|
for some uniform 0 < c < 1. In fact, this holds with the Frobenius norm
of H0 . This shows the desired invertibility with the desired probability and
completes the proof.
So far we have established that we may define the function (3.6) with high
probability. We turn to the proof of the upper bounds in Proposition 3.4.10,
which will again rely on Lemma 3.4.11 and Lemma 3.4.12.
3.4. COMPRESSED SENSING 65
where we have used the identity (1−M )−1 = (1−M n )−1 (1+M +· · ·+M n−1 ).
In the following, we denote
∞
np
X
1
R= |Ω|np H0
p=1
tr H0 H0∗
p
and regard this as a remainder term. In fact, writing kH0 kF =
for the Frobenius norm, we have the following implication:
αn
kH0 kF ≤ α|Ω| =⇒ kRkF ≤ 1−αn .
where X
kRk∞ := sup |R[i, j]| = sup kRxk`∞ .
i j kxk`∞ ≤1
As we will see, this will deal with the contribution of R in the formulas
above. Thus we will focus on proving estimates for the truncated series
n−1
X
1 m
|Ω|m H0 .
m=0
66 CHAPTER 3. FOURIER ANALYSIS, PART II
P = P0 + P1 off of S,
where
1 ∗
P0 = Dn sgn(f ), P1 = |Ω| HRe (I + Dn−1 ) sgn(f ),
n
X
1 m
Dn = |Ω|m H0 .
m=1
Continuing from above, some rearrangement shows that the claim boils down
to the identity
n−1
X n−1
X
∗ −m ∗ m
e |Ω| (He ) sgn(f ) = |Ω|−m (e∗ H)m sgn(f ).
m=0 m=0
as desired.
Now, choosing any a0 , a1 > 0 with a0 + a1 = 1, we begin with the bound
P sup |P (t)| > 1 ≤ P(kP0 k∞ > a0 ) + P(kP1 k∞ > a1 ).
t∈S
/
Then, by Tchebychev,
j+1 j+1
n
X X 2 X−1
J−1 X 2 X−1
J−1
−2Kj
E{|Ym |2Kj },
P Ym > a0 ≤ P(Ym > βj ) ≤ βj
m=1 j=0 m=2j j=0 m=2j
where Kj := 2J−j . Now, for 2j ≤ m < 2j+1 , we have n ≤ Kj m < 2n. Thus,
recalling |S| . τ N/n (with n ∼M log N ), we can apply Lemma 3.4.12 to get
a bound like
E{|Ym |2Kj } . (1 − εM )−2n ne−n αn
for some α ∈ (0, 1) (one should think of α ∼ 21 , but this parameter will be
−K
specified more precisely below). If we choose βj j ≡ β0−n , then summing
the above gives
With β0 ∼ .42, one has j 2j βj ≤ .91 and hence one can conclude
P
where a0 ∼ .91. In particular, we have a set A(t) with P{A(t)} > 1 − εn and
c . As a consequence,
|P0 (t)| < .91 on A(t) ∩ BM
so that
1
kP1 k∞ ≤ |Ω| kHRk∞ 1 + kQ0 k∞ .
Note the argument just given for P0 applies equally well to Q0 .
Consider the event E = {kH0 kF ≤ α|Ω|}. As we saw in the proof of
Proposition 3.4.9, the probability of E exceeds 1 − O(N −M ). Using the
crude bound kHk∞ ≤ |S||Ω| (since H has |S| columns and each entry is
bounded by |Ω|) and (3.11), we have
3 n
1 α
|Ω| kRk∞ kHk∞ ≤ |S| 2 1−α n on E.
Putting together the pieces and recalling the bound (3.12) (for Q0 ), we see
that
3
αn
kP1 k∞ ≤ 2|S| 2 1−α n ≤ a1 (3.13)
68 CHAPTER 3. FOURIER ANALYSIS, PART II
X X 2n
X 2n
Y
= exp{ 2πi
N ωj (tj − tj+1 )} E Iωj ,
t1 ,...,t2n :tj 6=tj+1 0≤ω1 ,...,ω2n ≤N −1 j=1 j=1
3.4. COMPRESSED SENSING 69
where Iωj is the random variable defined in (3.7) and we have used the
linearity of expectation.
Now, for any ω = {ω1 , . . . , ω2n }, we may define an equivalence relation
R(ω) on A = {1, . . . , 2n} by imposing that
jR(ω)k iff ωj = ωk .
We claim that the expectation above depends only on the equivalence class
of R(ω), denoted A/R(ω). This is because the Iωj are independent and
identically distributed. In particular,
2n
Y
E Iωj = τ |A/R(ω)| .
j=1
(The number of elements in the equivalence class A/R(ω) tells you how
many times you should really multiply the probability τ to compute the
expected value.) With this in mind, we rewrite
X X X 2n
X
|A/R|
E{tr(H02n )} = τ exp N2πi
ωj (tj −tj+1 ) ,
t1 ,...,t2n :tj 6=tj+1 R∈P(A) ω∈Ω(R) j=1
(3.14)
where P(A) is the set of all equivalence relations on A and
where
R0 ≤ R if aRb =⇒ aR0 b
and
Ω≤ (R) = {ω : R(ω) ≤ R} = {ω : aRb =⇒ ωa = ωb }.
70 CHAPTER 3. FOURIER ANALYSIS, PART II
where
2n
X
f (ω) = exp{ 2πi
N ωj (tj − tj+1 )}.
j=1
where
0
X Y
T (R0 ) := τ |A/R| (−1)|A/R|−|A/R | (|A0 /R| − 1)!.
R≥R0 A0 ∈A/R0
We now claim that (by splitting A into equivalence classes of A/R0 and
further splitting the relations R00 on A0 ∈ A/R0 by number of equivalence
classes), we may rewrite this as
|A | 0
0
Y X
0
T (R ) = S(|A0 |, k)τ k (−1)|A |−k (k − 1)!, (3.15)
A0 ∈A/R0 k=1
See Exercise 3.5.9. We denote the sum appearing above by F|A0 | (τ ), i.e.
n
X
Fn (τ ) := (k − 1)!S(n, k)(−1)n−k τ k . (3.16)
k=1
Using this notation, let us continue from above to finally express the desired
expected value in a way that is amenable to estimation. So far, we have
arrived at
X X Y X
E{tr(H02n )} = F|A0 | (τ ) f (ω),
R∈P(A) t1 ,...,t2n :tj 6=tj+1 A0 ∈A/R ω∈Ω≤ (R)
3.4. COMPRESSED SENSING 71
with f as above. Now let us work on this final sum. Note that for any
equivalence class A0 ∈ A/R and ω ∈ Ω≤ (R), we have ωa =Pωb for any a, b ∈
A0 . Denote this common value by ωA0 . Denote also tA0 = a∈A0 (ta − ta+1 ).
Then we can write
X Y
exp{ 2πi
N ω j (t j − t j+1 )} = exp{ 2πi
N ωA tA }.
0 0
j A0 ∈A/R
Using this, X Y X
f (ω) = exp{ 2πi
N ωA tA },
0 0
where the sum is over all possible ωA0 ∈ {0, . . . , N − 1}. But now we observe
that the inner sum equals N when tA0 = 0 and equals zero otherwise. In
conclusion:
X X Y
E{tr(H02n )} = N |A/R| F|A0 | (τ ), (3.17)
R∈P(A) T (R) A0 ∈A/R
where
T (R) = {t1 , . . . , t2n s.t. tj 6= tj+1 and tA0 = 0 for all A0 ∈ A/R}.
This formula implies that we may disregard any R such that some equiv-
alence class in A/R is a singleton. Indeed, if A0 ∈ A/R equals {j}, then
tA0 = tj −tj+1 6= 0 (because of the constraint on the set T (R)). Disregarding
such relations, we can get the bound
#T (R) ≤ |S|2n−|A/R|+1 .
This follows from the fact that there are |A/R| many constraintsPon the tj
coming from the condition tA0 = 0, and one more coming from 2n j=1 (tj −
tj+1 ) = 0. Thus, continuing from (3.17),
n
X X Y
E{tr(H02n )} ≤ N k |S|2n−k+1 F|A0 | (τ ), (3.18)
k=1 R∈P(A,k) A0 ∈A/R
Sketch of proof. Recall the definition of Fn in (3.16). Now, note that the
Stirling numbers satisfy the recurrence relation
x∗ = n−1
log( 1−τ )
.
τ
Continuing from (3.18), we replace F|A0 | (τ ) with G(|A0 |). We are then
faced with estimating
X Y
Q(2n, k) := G(|A0 |).
R∈P(A,k) A0 ∈A/R
We claim
Q(n, k) ≤ G(2)k (2n)n−k , (3.21)
τ
where we note that G(2) = 1−τ . This will complete the proof as follows.
Noting
N G(2)
4n|S| ≥ 1,
we can apply (3.21) to see
n
X
E{tr(H02n )} ≤ N k |S|2n−k+1 G(2)k (4n)2n−k
k=1
n
( N4n|S|
G(2) k
X
≤ |S|2n+1 (4n)2n )
k=1
2n N G(2) n
2n+1
≤ n|S| (4n) ( 4n|S| )
n+1 n n n
≤ n|S| N G(2) (4n) ,
3.5. EXERCISES 73
τ 1
which is the desired estimate (recalling G(2) = 1−τ and τ ≤ 1+e ).I
It remains to verify (3.21), which we only sketch (and leave the details
as an exercise). The key is to establish the recursive estimate
3.5 Exercises
Exercise 3.5.1. Prove that
F −1 [ΠN ] = N −1 ΠN −1 . (3.22)
Hint. Use the Poisson summation formula to treat the case N = 1. Then
compute the general case by scaling.
Exercise 3.5.2. Derive the formula appearing in Theorem 3.1.1 by following
the scheme outlined in Remark 3.1.2.
Exercise 3.5.3. Prove the identities (3.2).
I
Actually, we will miss by the factor e−n if n is not too large. However, in the appli-
cation above, n was of size ∼ log N , which may be expected to be large. Recovering this
factor in general requires an additional argument; see [4].
74 CHAPTER 3. FOURIER ANALYSIS, PART II
and that p
kHk ≤ kHkF := tr(HH ∗ ).
Exercise 3.5.7. Prove exercise Lemma 3.4.13 by completing the following
argument. The details are found in Section IV B of [4].
One can pass from A to A/R to assume that R is simply equality =.
After relabeling A as A = {1, . . . , n}, the formula reduces to
X X Y X
f (ω) = (−1)n−|A/R| (|A| − 1)! f (ω), (3.23)
ω1 ,...ωn distinct R A0 ∈A/R ω∈Ω≤ (R)
where the sum is over all equivalence relations R on A. This formula may
be proved
P by induction. The base case n = 1 follows because both sides
equal f (ω). Suppose the formula has been proven up to level n − 1. Then
rewrite the left-hand side as
X X n−1
X
f (ω 0 , ωn ) − f (ω 0 , ωj ) , (3.24)
ω1 ,...,ωn−1 distinct ωn j=1
and apply the inductive hypothesis to get a new expression for the left-hand
side. Here ω 0 = (ω1 , . . . , ωn−1 ).
Now work on the right-hand side of the formula above that will eventually
lead to the same formula just derived. Note that any equivalence class R on
A can be restricted to an equivalence class R0 on A0 = {1, . . . , n − 1}. Then
R can be formed from R0 either by adding {n} as a new equivalence class
(in which case we write R = {R0 , {n}}), or by having nRj for some j ∈ A0 ,
in which case we write R = {R0 , {n}}/(n = j). In the latter case, there may
3.5. EXERCISES 75
X X X n−1
X
1
F (R) = F ({R0 , {n}}) + |[j]R0 | F ({R0 , {n}/(n = j)}).
R R0 R0 j=1
Apply this identity to the right-hand side of (3.23). This produces two terms
that can be shown to match the two terms arising from (3.24). To make the
second terms match, one must utilize the identity
Y Y
0
1
|[j]R | (|A | − 1)! = (|A0 | − 1)!.
0
A0 ∈A/({R0 ,{n}}/(n=j)) A0 ∈{1,...,n−1}/R0
In this section, we will take a tour through some topics in abstract harmonic
analysis. Our goal will not be to present a thorough theoretical presentation,
but rather to show how many of the preceding topics can be understood as
special cases of a more general theory. In particular, many preliminary
results will simply be quoted as needed; the interested reader is encouraged
to pick up [10] to find complete details. We will also explore related topics
in some new settings (e.g. in the setting of compact Lie groups).
4.1 Preliminaries
Definition 4.1.1. A topological group is a group G with a topology
such that the group operation and inverse operation are continuous (from
G × G → G and from G → G, respectively).
76
4.1. PRELIMINARIES 77
• Every locally compact group possesses a left Haar measure ([10, Theo-
rem 2.10]). Left Haar measure is unique up to a multiplicative constant
([10, Theorem 2.220]).
The two notions of representation above are related, in the sense that any
unitary representation π of G corresponds to a ∗-representation of L1 (G),
which we may still denote by π. In particular, for f ∈ L1 (G) we define the
bounded operator π(f ) ∈ L(Hπ ) by
Z
π(f ) = f (x)π(x) dx,
G
with
hx, ξ −1 i = hx−1 , ξi = hx, ξi.
kξkL2 (G) = 1.
where we have used the translation invariance of Haar measure. This implies
Z
hx, ξihx, ηi dx = 0,
as desired.
In particular,
{f ∈ L∞ (G) : ∫ f > 21 } ∩ Ĝ = {1}.
We claim that this implies that {1} is open in Ĝ. Indeed, because R G is
compact, we have that the constant function 1 is in L1 (G) (so that f may
be viewed as 1(f ) through the identification of (L1 )∗ with L∞ ). This in turn
implies that every singleton set in Ĝ is open, i.e. Ĝ is discrete.
The next result puts the Fourier transform, Fourier series, and the dis-
crete Fourier transform under the same umbrella.
• T̂ = Z with hα, ni = αn .
• Ẑ = T with hn, αi = αn .
Thus φ(t) = ect ; however, since |φ| = 1 we may write c = 2πiξ for some
ξ ∈ R.
Next, since T can be identified with R/Z (via the identification of x ∈
R/Z with α = e2πix ), the characters of T are the characters of R that are
trivial on Z, so the result follows from above.
Now if φ ∈ Ẑ then α := φ(1) ∈ T and φ(n) = [φ(1)]n = αn (by the
homomorphism property).
Finally, the characters of Zk are the characters of Z that are trivial on
kZ. Thus they are of the form φ(n) = αn where α is a k th root of unity.
One can also check that if G1 , . . . , Gn are locally compact Abelian groups,
then
(G1 × · · · × Gn )ˆ = Ĝ1 × · · · × Ĝn .
This allows us to extend the previous result to see R̂n = Rn , T̂n = Zn ,
Ẑn = Tn , and finally Ĝ = G for any finite Abelian group G.
To define the Fourier transform on G, we make use of (4.1). In par-
ticular, the Fourier transform is the map from L1 (G) to C(Ĝ) given by
Z
Ff (ξ) = fˆ(ξ) := hx, ξif (x) dx.
G
Even in this generality, the Fourier transform enjoys many of the familiar
properties that we are used to. For example, it defines a norm-decreasing
∗-homomorphism from L1 (G) to C0 (Ĝ). It can also be extended to complex
Radon measures on G via
Z
µ̂(ξ) = hx, ξidµ(x).
G
where we have used that hx, 1i ≡ 1. This shows that the dual measure on
Ĝ must be counting measure.
On the other hand, if G is discrete then we let g = χ{1} . Then ĝ ≡ 1
and Z
g(x) = hx, ξi dξ,
Ĝ
so that |Ĝ| = 1.
84 CHAPTER 4. ABSTRACT FOURIER ANALYSIS
Example 4.2.2. The groups T and Z are dual. The dual measures can be
dθ
taken to be normalized Lebesgue measure 2π and counting measure. Then
Fourier inversion becomes
Z 2π X
ˆ
f (n) = f (θ)e−inθ 2π
dθ
, f (θ) = fˆ(n)einθ .
0 n∈Z
The general form of the Plancherel theorem is given by the following (see
[10, Theorem 4.26]).
Theorem 4.2.6 (Plancherel). The Fourier transform on L1 (G) ∩ L2 (G)
extends uniquely to a unitary isomorphism from L2 (G) to L2 (Ĝ). Conse-
quently, if G is compact and |G| = 1, then Ĝ is an orthonormal basis for
L2 (G).
We turn to our final main result concerning Fourier analysis on locally
compact Abelian groups, namely, the Pontrjagin duality theorem. Recall
that by definition, elements of Ĝ are characters on G. We can also view
elements of G as characters on Ĝ. Indeed, for x ∈ G we can define a
character Φ(x) on Ĝ via
hξ, Φ(x)i = hx, ξi.
ˆ
It follows that Φ defines a group homomorphism from G to Ĝ. The Pontr-
jagin duality theorem (see [10, Theorem 4.32]) states the following:
Theorem 4.2.7 (Pontrjagin duality). If G is a locally compact Abelian
group, then Φ is an isomorphism of topological groups.
According to this theorem, we may freely write hx, ξi or hξ, xi for the
pairing between G and Ĝ.
One consequence of this theorem is the other form of Fourier inversion:
if f ∈ L1 (G) and fˆ ∈ L1 (Ĝ), then
Z
f (x) = hx, ξifˆ(ξ) dξ
Ĝ
almost everywhere. We also have the dual form of Proposition 4.2.2; that
is, if Ĝ is compact then G is discrete, and if Ĝ is discrete then G is compact.
4.3. COMPACT GROUPS 85
The next result we need is the following theorem (see [10, Theorem 5.11]).
Then E is dense in C(G) in the uniform norm and in Lp (G) for all p < ∞.
We now state the main result (called the Peter–Weyl theorem [10, The-
orem 5.12]). In the following, given an equivalence class [π] we assume we
have chosen one fixed representative π.
and if
πij (x) = hπ(x)ej , ei i,
then the set
p
dim Hπ πij : i, j = 1, . . . dim Hπ , [π] ∈ Ĝ
X dim
X Hπ Z
f= cπij πij , cπij = dim Hπ f (x)πij (x) dx.
G
[π]∈Ĝ i,j=1
i,j i,j
χπ (x) = tr π(x).
4.4 Examples
We work through the details of some special examples, namely SU (2) and
SO(n) for n ∈ {3, 4}.
Example 4.4.1 (SU (2)). Let U (n) denote the group of unitary transforma-
tions of Cn , that is, the set of n × n matrices T satisfying T ∗ T = I. We let
SU (n) be the subgroup consisting of T ∈ U (n) with det T = 1. Note that
T ∈ U (n) if and only if T T ∗ = I, so that the rows of T are an orthonormal
set.
When n = 2, we can write
a b
T := ∈ U (2)
c d
if and only if |a|2 + |b|2 = |c|2 + |d|2 = 1 and ac̄ + bd¯ = 0. In particular,
(a, b) is a unit vector and (c, d) = eiθ (−b̄, ā) for some θ ∈ R. It follows that
det T = eiθ , and so T ∈ SU (2) if and only if eiθ = 1. Writing
a b
Ua,b = ,
−b̄ ā
4.4. EXAMPLES 89
we have
SU (2) = {Ua,b : a, b ∈ C, |a|2 + |b|2 = 1}.
The correspondence Ua,b with (a, b) = (1, 0)Ua,b identifies SU (2) with a
subset of the unit sphere S 3 ⊂ C2 (where the identity element is identified
with (1, 0).
Three one-parameter subgroups of SU (2) are of particular interest:
iθ
e 0
F (θ) = ,
0 e−iθ
cos φ sin φ
G(φ) = ,
− sin φ cos φ
cos ψ i sin ψ
H(ψ) = .
i sin ψ cos ψ
These are three mutually orthogonal great circles in the sphere that intersect
at ±Id.
Proof. Unitary matrices are normal, so by the spectral theorem there exists
V ∈ U (2) with
V T V −1 = diag(α, β).
For T ∈ SU (2) we have β = ᾱ = e−iθ for some θ ∈ [−π, π]. In particular,
1
V T V −1 = F (θ). By replacing V with [det V ]− 2 V , we may assume that
V ∈ SU (2) as well. Furthermore, using
= 2c Γ( m
2 + 2) f (Z 0 ) dσ(Z 0 ).
S3
4.4. EXAMPLES 91
z ∂P
∂w ∈ M and w ∂P
∂z ∈ M.
z ∂P ∂P
∂w + w ∂z =
1 d
i dψ π(H(ψ))P |ψ=0 ∈ M.
Our next goal is to show that the πm ’s give a complete list of the irre-
ducible representations of SU (2):
Proof. First note that none of the πm ’s are equivalent representations, be-
cause they all have different dimensions (and different characters, as we will
see).
Now, let χm be the character of πm , and define
as in the corollary above. Note that the orthogonal basis vectors z j wm−j for
Pm are eigenvectors for πm F (θ); indeed, using (4.2) with a = eiθ and b = 0,
we get
πm (F (θ))(z j wm−j ) = e−i(2j−m)θ z j wm−j .
Thus
m
sin((m+1)θ)
X
χ0m (θ) = e−i(2j−m)θ = sin θ (4.3)
j=0
(see Exercise 4.5.3). It follows that χ00 (θ) ≡ 1, χ01 (θ) = 2 cos θ, and more
generally
χ0m (θ) − χ0m−2 (θ) = 2 cos mθ for m ≥ 2. (4.4)
(see Exercise 4.5.4). Thus the span of {χ0m } equals the span of {cos mθ}.
The latter is uniformly dense in C([0, π]); thus, by Corollary 4.4.2, the span
of {χ0m } is uniformly dense in the space of continuous central functions on
SU (2).
This means that the only function orthogonal to all χm is the zero func-
tion. By Proposition 4.3.6, this shows that χm must include all possible
irreducible characters. This completes the proof.
If we set z = e2πit and w = 1, then the sum becomes a Fourier series and
jk
πm (a, b) are computed like Fourier coefficients:
q Z 1
j!(m−j)!
jk
πm (a, b) = k!(m−k)! (āe2πit − b̄)k (be2πit + a)m−k e−2πijt dt.
0
j0
√ Exercise 4.5.5). In particular {πm : 0 ≤ j ≤ m} span Pm . Here
(see
m + 1 = dim Hπm is needed to normalize the matrix elements.
jk
Let us also discuss the span of {πm : 0 ≤ j ≤ m}. The identities
jk
above show that πm (a, b) is a polynomial in the variables (a, b, ā, b̄) that is
homogeneous of degree m − k in (a, b) and of degree k in (ā, b̄). That is, it
has bidegree (m − k, k).
jk
Furthermore, as a function on C2 , each πm is harmonic:
4
jk
∂ 2 πm
X
∂x2n
= 0, a = x1 + ix2 , b = x3 + ix4
n=1
(check!).
We conclude this example with the following: We identify SU (2) with
the unit sphere in C2 by identifying Ua,b with (a, b̄). Then the Peter–Weyl
decomposition
∞
M
L2 = Eπm
m=0
agrees with the decomposition of functions on the sphere into ‘spherical
harmonics’. We then have the further decomposition
M
Eπ m = Hp,q ,
p+q=m
4.4. EXAMPLES 95
jq
where Hp,q is the span of πp+q for 0 ≤ j ≤ p + q. This is a grouping of the
spherical harmonics of degree m according to their bidegree.
Example 4.4.2 (SO(3)). We write SO(n) for the group of rotations on Rn .
To describe these groups (and their connections to SU (2)) we will make use
of the quaternions, denoted by H.
As a real vector space, we may identify H with R × R3 , with elements
denoted by (a, x). Multiplication in H is given by
where · denotes dot product and × denotes cross product. One can verify
that this product is associative and that |ξη| = |ξ| |η| (where ξ, η ∈ R4 and
| · | denotes euclidean length in R4 ).
The subspace R × {0} is the center of H. We identify this with R and
view it as the ‘real axis’, and we identify {0} × R3 with R3 . Then instead
of writing (a, x), we may write a + x. Denoting the standard basis of R by
i, j, k, we then have
(a, x) = a + x = a + x1 i + x2 j + x3 k.
ξ ξ¯ = ξξ
¯ = a2 + |x|2 = |ξ|2 .
Defining
U (H) = {ξ ∈ H : |ξ| = 1},
we find that U (H) forms a group. For ξ ∈ U (H), the map
η 7→ ξηξ −1
is an isometric linear map from H to H that leaves the center R (and hence
the subspace R3 ) invariant. The restriction of this map to R3 is thus an
element of SO(3) denoted by κ(ξ), i.e.
κ(ξ)x = ξxξ −1 .
This belongs to SO(3) (and not just O(3)) because κ is a continuous map
on the connected set U (H).
96 CHAPTER 4. ABSTRACT FOURIER ANALYSIS
SO(3) ∼
= SU (2)/{±I}.
[SO(3)]ˆ = {[ρk ] : k = 0, 1, 2, . . . },
where
ρk (Ad(U )) = π2k (U ).
Here Ad(U )A := U AU −1 .
SO(4) ∼
= [SU (2) × SU (2)]/{±(1, 1)}.
where
ρmn (L(ξ)R(η)) := πmn (ξ, η).
The restriction that m, n have the same parity comes from the fact that
πm (−1) = (−1)m I.
4.5 Exercises
Exercise 4.5.1. Recover the theory of Fourier series using the Peter–Weyl
theorem and the discussion thereafter.
Exercise 4.5.2. Verify that the maps defined in (4.2) are unitary represen-
tations of SU (2).
Exercise 4.5.3. Compute the sum in (4.3).
Exercise 4.5.4. Prove the identity (4.4).
Exercise 4.5.5. Prove (4.5).
Exercise 4.5.6. Verify (4.6).
Chapter 5
Wavelet transforms
99
100 CHAPTER 5. WAVELET TRANSFORMS
1
Proposition 5.1.1. For ψ ∈ Ḣ − 2 , we have
ZZ
−1
f = Cψ T f (a, b)ψ a,b a−2 da db
Then Z
1 1
fˆ(ξ)|a| 2 e−ibξ ψ̂(aξ) dξ = (2π) 2 F̂a (b),
= Cψ hf, gi,
where we have changed variables in the da integral above and then applied
Plancherel once more.
There are several variations of the situation descibed above. For exam-
ple, if we use a real-valued ψ, then we get
ψ̂(−ξ) = ψ̂(ξ),
Then we have Z ∞Z
f = C̃ψ−1 T f (a, b)ψ a,b a−2 db da.
0 R
One can also investigate using bandlimited functions, wavelets, or using
complex-valued wavelets with real-valued signals, and so on.
One important variation involves using different wavelets for the decom-
position and the reconstruction of the signal f . In particular, we have:
Proposition 5.1.3. If ψ1 , ψ2 satisfy
Z
|ψ̂1 (ξ)| |ψ̂2 (ξ)| |ξ|−1 dξ < ∞,
then ZZ
f = Cψ−1
1 ,ψ2
hf, ψ1a,b iψ2a,b a−2 da db
The function g is called the reproducing kernel for the Hilbert space.
Now let us return to the continuous wavelet transform. For f ∈ L2 (R),
we have ZZ Z
Cψ−1 |T f (a, b)|2 a−2 da db = |f (x)|2 dx,
Let H denote the closed subspace given by the image of L2 under T . Then
H is a reproducing kernel Hilbert space. In fact,
ZZ
−1
a,b
F (a, b) = hf, ψ i = Cψ K(a, b; α, β)F (α, β)α−2 dα dβ,
where
K(a, b; α, β) = hψ α,β , ψ a,b i.
That is, point evaluation is given by integration against a kernel.
The windowed Fourier transform.
We turn to a comparison of the wavelet transform with the so-called
windowed Fourier transform. Given a smooth function g supported
near the origin, we define the windowed Fourier transform by
That is, Fg f (ω, t) represents the Fourier transform of the portion of f (x)
around the point x = t. This is a natural quantity to consider in the setting
of signal processing.
Arguing as in the proof of Proposition 5.1.1, we have the following:
5.1. CONTINUOUS WAVELET TRANSFORMS 103
One can use any g ∈ L2 for the windowing function. Typically, one
normalizes kgkL2 = 1. The windowed Fourier transform also maps L2 to a
reproducing kernel Hilbert space, viz.
ZZ
1
F (ω, t) = 2π K(ω, t; ω 0 , t0 )F (ω 0 , t0 ) dω 0 dt0 ,
where
0 0
F (ω, t) = Fg f (ω, t), K(ω, t; ω 0 , t0 ) = hg ω ,t , g ω,t i.
f 7→ hf, φiφ
For the identity matrix, we give each projection equal weight. However,
by varying the weight given to each projection operator, we can construct a
wide variety of operators.
Let us consider this first in the setting of the windowed Fourier trans-
form. Inspired by applications in quantum mechanics, we switch to the posi-
tion/momentum variables (p, q) (instead of (t, ω)). Given a weight function
w(p, q) and an L2 -normalized window function, we may define the operator
ZZ
W = 2π1
w(p, q)h·, g p,q ig p,q dp dq.
where Z
Cg = ξ 2 |ĝ(ξ)|2 dξ.
The largest value of this ratio corresponds to the largest eigenvalue of the
symmetric operator QT PΩ QT , which is an integral operator with integral
kernel
(QT PΩ QT )(x, y) = 1[−T,T ] (x)1[−T,T ] (y) sin[Ω(x−y)]
π(x−y)
(exercise). The eigenvalues and eigenvectors of this operator are known, due
to the fact that this operator commutes with the operator
2 Ω2 2
A= d
dx (T − x2 ) dx
d
− π2
x ,
which was previously studied in the context of solving the Helmholtz equa-
tion via separation of variables. The eigenfunctions are called ‘prolate
spheroidal wave functions’.
The eigenvalues may be put in decreasing order. One finds (by a rescaling
argument) that the eigenvalues of QT PΩ QT depend only on the product
T Ω. It turns out that (for fixed T Ω) the eigenvalues begin close to 1 before
suddenly plunging down to essentially zero. There are about 2TπΩ eigenvalues
near one before a plunge region of width about log(T Ω).
This gives a rigorous version of something empirically observed long ago.
In particular, in a time and band-limited region, there are 2T Ω/π ‘degrees
of freedom’ (i.e. independent functions that are essentially time and band-
limited in this way). This is proportional to the area of [−T, T ] × [−Ω, Ω];
it also corresponds to the number of sampling times within [−T, T ] dictated
by Shannon’s sampling theorem (Theorem 3.1.1) for a function with Fourier
support in [−Ω, Ω].
Example 5.1.3. Next, returning to the discussion above, let us use the win-
dowed Fourier transform to define an operator that corresponds to time-
frequency localization to a set S ⊂ R2 . We define
ZZ
LS = 2π1
h·, g ω,t ig ω,t dω dt.
(ω,t)∈S
S = SR = {(ω, t) : ω 2 + t2 ≤ R2 }
and
1 2 /2
g(x) = g0 (x) := π − 4 e−x .
In this case, things can be computed fairly explicitly. We only mention some
of the results; more details may be found in [7, pp38–40].
With these choices of g0 and SR , one finds by direct computation that
LR commutes with the harmonic oscillator Hamiltonian, defined by
d 2 2
H = − dx 2 + x − 1.
In fact, one can compute the action of the unitary group e−isH on g0 explic-
itly and prove that
LR e−isH = e−isH LR ,
which implies the result. The proof boils down to showing that the action
of e−isH on g0ω,t simply produces (up to a phase factor) some other g0ωs ,ts ,
where (ωs , ts ) are given by a rotation matrix applied to (ω, t). One then
applies change of variables (given by a rotation, which leaves the domain SR
invariant) to prove the identity above.
5.1. CONTINUOUS WAVELET TRANSFORMS 107
max{n : λn (R) ≥ 21 } ∼ 12 R2 ,
which (writing it as πR2 /2π) is again the area of the time-frequency local-
ization region divided by 2π. The width of the ‘plunge region’ is larger this
time, but still essentially negligible compared to 21 R2 . The eigenfunctions
turn out to be independent of the size of SR ; that is, the R-dependence is
completely represented through the eigenvalues.
So far, we have focused on using the windowed Fourier transform to build
operators. Of course, a similar construction is possible using the wavelet
transform. That is, given a weight w(a, b), we may define
ZZ
−1
W = Cψ w(a, b)h·, ψ a,b iψ a,b a−2 da db.
Once again, one may interested in using a weight w(a, b) that simply cuts off
to a subset S. Once again this leads to operators LS satisfying 0 ≤ LS ≤ 1,
which (provided S is compact and does not contain a = 0) are trace class.
As before, the eigenvalues/eigenvectors can be difficult to analyze except in
some special cases. One such case involves taking ψ̂(ξ) = 2ξe−ξ χ[0,∞) (ξ)
and using the identity
Z ∞Z
−1 a,b a,b a,b a,b −2
1 = Cψ [h·, ψ+ iψ+ + h·, ψ− iψ− ]a db da,
0
108 CHAPTER 5. WAVELET TRANSFORMS
where ψ+ = ψ and ψ̂− (ξ) = ψ̂(ξ). One then considers localization to regions
as claimed.
5.1. CONTINUOUS WAVELET TRANSFORMS 109
can choose ψ2 so that the constant Cψ,ψ2 = 1. Using the resolution of the
identity, ZZ
f (x) = hf, ψ a,b iψ2a,b (x)a−2 da db.
We split the integral into the part where |a| ≤ 1 and |a| > 1. We call
these two terms fS and fL (for small and large scales, respectively).
First consider the large scale piece, which is actually already guaranteed
to be Lipschitz. We begin by showing fL is uniformly bounded. In fact,
ZZ
5
|fL (x)| . kf kL2 kψkL2 a− 2 |ψ2 ( x−b
a )| da db
|a|≥1
Z
3
. kf kL2 kψkL2 kψ2 kL1 |a|− 2 da . 1.
|a|≥1
|fL (x + h) − fL (x)|
ZZZ
a−3 |f (y)||ψ( y−b ψ2 ( x+h−b ) − ψ2 ( x−b ) dy db da
≤ a )| a a
|a|≥1
ZZZ
. |h| a−4 |f (y)| da db dy,
S
We turn to the small scale piece, beginning again with uniform bound-
edness:
ZZ
1 1
|fS (x)| . |a|α+ 2 |a|− 2 |ψ2 ( x−b
a )|a
−2
da db
|a|≤1
Z
. kψ2 kL1 a−1+α da . 1.
|a|≤1
110 CHAPTER 5. WAVELET TRANSFORMS
Next, for |h| ≤ 1, we consider fS (x + h) − fS (x) and split the integral into
two regions. First, we get the contribution
ZZ
x+h−b) −2
|a|α |ψ2 ( x−b
a )| + |ψ2 ( a )| a da db
|a|≤|h|
ZZ Z
. a α−2
da db . |a|−1+α da . |h|α ,
S |a|≤|h|
where
S = {(a, b) : |a| ≤ |h| and |x − b| ≤ |a|R}
for some R > 0. Next, using the fact that ψ2 is Lipschitz, the contribution
of |h| ≤ |a| ≤ 1 is controlled by
ZZ Z
−1 −2
α
|a| |h||a| a da db . |h| |a|−2+α . |h|α ,
S̃ |h|≤|a|≤1
R
Again, the first part follows from a computation: using ψ = 0,
Z
1
|hf, ψ a,b i| ≤ |f (x) − f (0)||a|− 2 |ψ( x−b
a )| dx
Z
1
. |x|α |a|− 2 |ψ( x−b
a )| dx
Z
1 1 1
. |a|α+ 2 |ψ(y)| |y + ab |α dy . |a| 2 +α + |a| 2 |b|α ,
as desired.
We turn to the converse statement. In fact, the argument proceeds
similar to the proof of Theorem 5.1.6. We actually only need to establish a
suitable bound on |fS (h) − fS (0)| for |h| ≤ 1. This time we split into four
pieces:
ZZ
−2
|a|γ |ψ2 ( h−b
a )|a da db, (5.3)
|a|≤|h|α/γ
ZZ
|b|α −2
(|a|α + | log h−b
|b|| )|ψ2 ( a )|a da db, (5.4)
|h| α/γ ≤|a|≤|h|
ZZ
|b|α −2
(|a|α + | log b
|b|| )|ψ2 (− a )|a da db, (5.5)
|a|≤|h|
ZZ
|b|α −b −2
(|a|α + | log h−b
|b|| )|ψ2 ( a ) − ψ2 ( a )|a da db. (5.6)
|h|≤|a|≤1
We first estimate
Z
(5.3) . |a|−1+γ kψ2 kL1 . |h|α ,
|a|≤|h|α/γ
which is acceptable.
Next, supposing the support of ψ2 is contained in [−R, R], we have
Z
(5.4) .kψ2 kL1 |a|−1+α da
|a|≤|h|
(|a|R + |h|)α
Z
+ |a|−1 da
|h|α/γ ≤|a|≤|h| | log(|a|R + |h|)|
Z
. |h|α 1 + 1
| log |h|| |a|−1 da . |h|α .
|h|α/γ ≤|α|≤|h|
Similarly,
(|a|R)α
Z Z
−1+α
(5.5) . |a| da + |a|−1 da
|a|≤|h| |a|≤|h| | log(|a|R)|
α
. |h| .
112 CHAPTER 5. WAVELET TRANSFORMS
For a fixed dilation step a0 > 1, we will restrict to the discrete set of
scales a = am
0 where m ∈ Z. We next fix a translation parameter b0 > 0 and
define the rescaled and translated wavelets
−m −m x−nb0 am
ψm,n (x) = a0 2
ψ(a−m
0 x − nb0 ) = a0
2
ψ( am
0
),
0
where m, n ∈ Z.
The basic questions we are interested in are whether functions are uniquely
determined by their wavelet coefficients hf, ψm,n i, as well as how functions
may be reconstructed using wavelets and wavelet coefficients.
We will need the notion of a frame.
Frames and reconstruction.
which yields X
F ∗c = cj ϕj .
j∈J
1
Note that kF ∗ k = kF k ≤ B 2 and that
X
|hf, ϕj i|2 = kF f k2 = hF ∗ F f, f i.
j
A Id ≤ F ∗ F ≤ B Id.
F̃ = F (F ∗ F )−1 ,
as well as
F̃ ∗ F̃ = (F ∗ F )−1 , F̃ ∗ F = F ∗ F̃ = Id.
Finally, F̃ F ∗ = F F̃ ∗ is the orthogonal projection in `2 onto R(F ) = R(F̃ ).
Proof. Exercise.
One calls {ϕ̃j } the dual frame of ϕj . (The dual frame of ϕ̃j is simply the
ϕj again.) The conclusions of Proposition 5.2.5 may be succinctly written
as X X
hf, ϕj iϕ̃j = f = hf, ϕ̃j iϕj .
j j
Note that typically the ϕj are not even linearly independent, and thus
there may be many linear combinations of the ϕj that yield f . The particular
linear combination given above has a minimality property.
P
Proposition 5.2.6. If f = j cj ϕj but we do not have cj ≡ hf, ϕ̃j i, then
X X
|cj |2 > |hf, ϕ̃j i|2 .
j j
f = F ∗ c = F ∗ F̃ g + F ∗ b =⇒ f = g.
Then X X
|cj |2 = kck2 = kF̃ f k2 + kbk2 = |hf, ϕ̃j i|2 + kbk2 ,
which implies the result.
3
X
2
v= 3 [hv, φj i + α]ϕj for any α ∈ C.
j=1
for all g ∈ H.
116 CHAPTER 5. WAVELET TRANSFORMS
(F ∗ F )−1 ∼ 2
A+B Id, and so ϕ̃j ∼ 2
A+B ϕj .
By construction, we have
− B−A
B+A Id ≤ R ≤
B−A
B+A Id, and so kRk ≤ B−A
B+A = r
2+r .
F ∗F = A+B
2 (Id − R), so that (F ∗ F )−1 = 2
A+B (Id − R)−1 .
B−A
As kRk ≤ B+A < 1, we can write
∞
X
−1
(Id − R) = Rk .
k=0
Then
∞
X
ϕ̃j = 2
A+B Rk ϕj .
k=0
N
X ∞
X
ϕ̃N
j = 2
A+B
k
R ϕj = ϕ̃j − 2
A+B Rk ϕj = [Id − RN +1 ]ϕ̃j .
k=0 k=N +1
5.2. DISCRETE WAVELET TRANSFORMS 117
It follows that
X X
hf, ϕj iϕ̃N hf, ϕj iϕ̃N
f − j
= sup hf − j , gi
j kgk=1 j
X
= sup hf, ϕj ihϕ̃j − ϕ̃N
j , gi
g
j
X
= sup hf, ϕj ihRN +1 ϕ̃j , gi
g
j
r N +1
≤ kRkN +1 kf k ≤
2+r kf k.
As for actually computing the ϕ̃N j , one can use an iterative algorithm
(that actually is relatively practical to implement) and write
−1
ϕ̃N
j =
2
A+B ϕj + Rϕ̃N
j .
We will next discuss necessary and sufficient conditions for this family to
form a frame in L2 (R). The following appear as Theorem 3.3.1 and Propo-
sition 3.3.2 in [7].
Theorem 5.2.7. Suppose ψm,n form a frame for L2 with frame bounds
A, B. Then
Z ∞
b0 log a0
2π A ≤ ξ −1 |ψ̂(ξ)|2 dξ ≤ b0 log a0
2π B,
0
Z 0
b0 log a0
2π A ≤ |ξ|−1 |ψ̂(ξ)|2 dξ ≤ b0 log a0
2π B.
−∞
and X
β(s) := sup |ψ̂(am m
0 ξ)| |ψ̂(a0 ξ + s)|
ξ m
118 CHAPTER 5. WAVELET TRANSFORMS
decays at least as fast as (1 + |s|)−(1+ε) for some ε > 0. Then there exists
b∗0 > 0 such that the ψm,n form a frame for any b0 < b∗0 , with the following
frame bounds:
X X 1
A = 2π |ψ̂(am 2
β( 2π 2π
inf 0 ξ)| − b0 k)β(− b0 k) ,
2
b0
1≤|ξ|≤a0
k6=0
X X 1
2π
|ψ̂(am 2
β( 2π 2π
B= b0 sup 0 ξ)| + b0 k)β(− b0 k)
2
.
1≤|ξ|≤a0 k6=0
imply that m |ψ̂(2m ξ)|2 should be roughly constant (for ξ 6= 0). This is a
P
rather strong condition. One way to remedy this is to use multiple wavelets
ψ 1 , . . . , ψ N , and to consider the frame obtained from {ψm,n
ν } (setting a = 2
0
for each).
The analogue of Theorem 5.2.7 for the windowed Fourier transform is
given by the following. In this case, we consider the functions
for m, n ∈ Z.
Theorem 5.2.8. If gm,n form a frame for L2 with frame bounds A, B, then
2π 2
A≤ ω0 t0 kgkL2 ≤ B.
Conversely, suppose
X X
inf |g(x − nt0 )|2 > 0, sup |g(x − nt0 )|2 < ∞,
0≤x≤t0 0≤x≤t0
n∈Z n∈Z
5.2. DISCRETE WAVELET TRANSFORMS 119
and X
β(s) := sup |g(x − nt0 )| |g(x − nt0 + s)|
0≤x≤t0 n
decays at least as fast as (1 + |s|)−(1+ε) for some ε > 0. Then there exists
ω0∗ > 0 so that the gm,n form a frame whenever ω0 < ω0∗ , with the following
frame bounds:
X X 1
A = ω2π0 inf |g(x − nt0 )|2 − β( ω2π0 k)β(− ω2π0 k) 2 ,
x
n k6=0
X X 1
B = ω2π0 sup |g(x − nt0 )|2 + β( ω2π0 k)β(− ω2π0 k) 2 .
x n k6=0
Example 5.2.3 (The Mexican hat function). The Mexican hat function ψ is
2
the second derivative of the Gaussian e−x /2 . Normalizing the L2 norm and
imposing ψ(0) > 0, one finds
1 2 /2
ψ(x) = √2 π − 4 (1 − x2 )e−x .
3
If one uses at least two voices, this yields an essentially tight frame for
b0 ≤ .75. This wavelet is often used in computer vision applications.
Example 5.2.4. For the windowed Fourier transform, the Gaussian g(x) =
1 2
π − 4 e−x /2 is commonly used. It turns out that the ratio ω0 t0 ÷2π is relevant.
In particular, the gm,n form a frame whenever ω0 t0 < 2π. This is related
to the notion of time-frequency density, which is discussed in detail in [7,
Chapter 4].
for some α > 1, β > 0, and γ > 0. Then for any ε > 0, 0 < Ω0 < Ω1 , and
T > 0, there exists a finite set B such that for any f ∈ L2 ,
X
kf − hf, ψm,n iψ
] m,n kL2
B
q
B
εkf kL2 + kf kL2 (|x|>T ) + kfˆkL2 (|ξ|≤Ω0 ) + kfˆkL2 (|ξ|>Ω1 ) .
≤ A
120 CHAPTER 5. WAVELET TRANSFORMS
Theorem 5.2.10. Suppose gn,m (x) = eimω0 x g(x − nt0 ) form a frame with
bounds A, B. Suppose that
with α > 1. Then for any ε > 0, there exists tε , ωε such that for any f ∈ L2
and T, Ω > 0, we have
X
kf − hf, gm,n ig̃m,n kL2
B
q
B
εkf kL2 + kf kL2 (|x|>T ) + kfˆkL2 (|ξ|>Ω) ,
≤ A
where
B1 = {(m, n) : m ≤ m0 or m > m0 }.
Here we write PΩ to be the Fourier multiplier operator cutting off the fre-
quency support to Ω0 ≤ |ξ| ≤ Ω1 . Next, we have
X
[|hQT f, ψm,n i| + |h(1 − QT )f, ψm,n i|]|hψ
]m,n , hi|,
B2
5.2. DISCRETE WAVELET TRANSFORMS 121
where
where we use Plancherel (on the torus) in the final step and
B10 = {m : m ≤ m0 or m > m0 }.
where
S± = {Ω0 ≤ |ξ|, |ξ ± 2π`a−m −1
0 b0 | ≤ Ω1 }.
122 CHAPTER 5. WAVELET TRANSFORMS
where X
−1
F± = |ψ̂(am
0 ξ)|
2−λ
|ψ̂(am
0 ξ ± 2πb0 `)|
λ
m∈B10
m>m1 m<m0
1 −2γ(1−λ)
. (Ω0 am
0 ) + (Ω1 am0 2β(1−λ)
0 ) .
Redundancy in frames
We close this section by making a brief comment about redundancy in
frames. For frames that are tight or close to tight, this can be measured by
the size of 12 (A + B).
5.3. MULTIRESOLUTION ANALYSIS 123
fapp = F̃ ∗ (F f + α).
However, since F̃ ∗ includes a projection onto the range of F , the part of the
sequence α that is orthogonal to the range of F does not contribute. Thus,
we expect
kf − fapp k = kF̃ ∗ αk
to be smaller than kαk. This effect should become even more pronounced
if the frame is more redundant, since in this case the range of F becomes
even ‘smaller’.
Example 5.2.5. Let u1 = (1, 0) and u2 = (0, 1), giving an orthonormal basis
for C2 . Let
√ √
3 3
e1 = u2 , e2 = − 2 u1 − 12 u2 , e3 = 2 u1 − 21 u2 .
while X
E kf − 32 [hf, ej i + αj ε]ej k2 = 43 ε2 .
T
(iii) j∈Z Vj = {0},
(iv) f ∈ Vj ⇐⇒ f (2j ·) ∈ V0 ,
{φ0,n : n ∈ Z}
Thus, all the spaces are scaled versions of a central space V0 , which is
invariant under integer translations. The final condition (vi) may be relaxed
considerably, but we begin with this simpler setting.
We will write Pj for the orthogonal projection onto Vj .
Example 5.3.1 (Haar multiresolution analysis). Let
Vj−1 = Vj ⊕ Wj .
Thus
J−j−1
M
Vj = VJ WJ−k for all j < J.
k=0
5.3. MULTIRESOLUTION ANALYSIS 125
Next, note that the Wj also inherit the scaling property (iv):
f ∈ Wj ⇐⇒ f (2j ·) ∈ W0 .
We rewrite this as
√ X X
φ(x) = 2 hn φ(2x − n), so that φ̂(ξ) = √1
2
hn e−inξ/2 φ̂(ξ/2).
n n
Defining
X
m0 (ξ) = √1
2
hn e−inξ ,
n
the functions {ψ(· − k)} form an orthonormal basis for W0 (and so {ψj,k }
form a wavelet basis for L2 ).
To this end, we let f ∈ W0 , i.e. f ∈ V−1 and f ⊥ V0 . By assumption,
we have X
f= fn φ−1,n , fn = hf, φ−1,n i,
n
so that X
fˆ(ξ + 2π`)φ̂(ξ + 2π`) ≡ 0.
`
We next claim X
|φ̂(ξ + 2π`)|2 ≡ 1
2π , (5.10)
`∈Z
which yields
|m0 (ξ)|2 + |m0 (ξ + π)|2 ≡ 1.
5.3. MULTIRESOLUTION ANALYSIS 127
In particular, m0 (·) and m0 (·+π) cannot both equal zero on a set of nonzero
measure, and hence (5.11) implies that there exists a 2π-periodic function
λ so that
for suitable νk .
It therefore remains to verify that the {ψ(· − k)} belong to V−1 ∩ V0⊥
and are orthonormal. This we leave as an exercise!
Remark 5.3.3. The proof above provides an example of a wavelet that one
can use. Such examples are not unique—one could modify ψ̂ by multiplying
by any 2π-periodic function that has magnitude one almost everywhere.
Using this freedom, we will take
X
ψ= gn φ−1,n , gn = (−1)n h−n+1 .
n
(in which case they are called a Riesz basis). In many examples, one starts
with the scaling function φ and then defines V0 by taking {φ(· − k)} as an
(orthonormal) basis. The spaces Vj are then the closed subspace spanned
by
φj,k (x) = 2−j/2 φ(2−j x − k).
This construction will lead to a multiresolution analysis provided
X
φ(x) = cn φ(2x − n)
n
f 0 = f 1 + δ 1 ∈ V1 ⊕ W1 = V0 ,
5.3. MULTIRESOLUTION ANALYSIS 129
In fact, we can describe these coefficients c1k and d1k . To see this, first recall
that by construction we may write
X X
ψj,k = gn φj−1,2k+n = gn−2k φj−1,n ,
n n
and similarly X
φj,k = hn−2k φj−1,n .
n
cjk = djk =
X X
hn−2k cj−1
n , gn−2k cj−1
n ,
n n
5.4 Exercises
Exercise 5.4.1. Work out the details in the discussion of Example 5.1.2.
Exercise 5.4.2. Solve the differential equation discussed in Example 5.1.4.
Exercise 5.4.3. Show that if {ϕj } is a tight frame with bound A then
X
f = A−1 hf, ϕj iϕj .
j
Exercise 5.4.7. Show that the functions {ψ(· − k)} defined in Theorem 5.3.2
are orthonormal. [Hint: Use the same trick appearing in the proof of the
theorem. That is, apply Plancherel and split the integral into a sum over
` ∈ Z. Split the sum into even and odd parts and use the properties of m0
and φ̂ that were already established.]
Chapter 6
6.1 Interpolation
In this section we will consider sublinear operators, i.e. those satisfying
|T (cf )| = |c| |T (f )| and |T (f + g)| ≤ |T (f )| + |T (g)|. This includes not only
linear operators, which include many important examples, but also maximal
operators, as well as square function type operators. The latter two may
have the form
X 1
T f = sup |Tn f | or T f = |Tn f |2 2
n n
131
132 CHAPTER 6. CLASSICAL HARMONIC ANALYSIS, PART I
(cf. Section A.1). When q = ∞, we define weak type (p, q) to mean strong
type (p, q).
Note that if T is of strong type (p, q), then T is of weak type (p, q);
indeed, by Tchebychev’s inequality,
kT f kLq,∞ ≤ kT f kLq .
and so on.
In the following, we will always consider operators with the property
that h|T f |, |g|i is finite whenever f, g are taken to be simple functions with
finite measure support.
We will study the problem of interpolating various types of bounds for
operators T . Let us begin with the following application of Hölder’s in-
equality, which is essentially an interpolation-type result for the identity
operator.
whenever 1 ≤ p0 , p1 ≤ ∞ and
1 θ 1−θ
p = p0 + p1 for some θ ∈ [0, 1].
In fact,
kT kLp →Lq . kT kθLp →Lq0 kT k1−θ
Lp →Lq1 .
In fact,
kT kLp →Lq . kT kθLp0 →Lq kT k1−θ
Lp1 →Lq .
Proof. For (i), we apply Hölder’s inequality (in the form of the lemma
above). We find
1 ≤ p0 , q0 , p1 , q1 ≤ ∞, p0 6= p1 , q0 6= q1 .
If T is of weak type (p0 , q0 ) and weak type (p1 , q1 ), then T is of type (p, q)
for all (p, q) such that p ≤ q and
( p1 , 1q ) = ( pθ0 + 1−θ θ
p1 , q0 + 1−θ
q1 ) for some θ ∈ (0, 1). (6.1)
Remark 6.1.5. We have not pursued optimal hypotheses here. For exam-
ple, it suffices to assume ‘restricted weak type’ bounds on T rather than
weak type bounds, and the conclusion of the theorem can be stated in terms
of more general Lorentz spaces (rather than just Lebesgue spaces). We will
not pursue such generality in these notes; however, we would like to point
out that although we will work in the context of Lebesgue measure, this
result applies to operators mapping Lp (dµ1 ) to Lq (dµ2 ) for more general
measures.
Remark 6.1.6. We would also like to point out that to obtain strong type
bounds, the restriction to p ≤ q is necessary. To see this consider for example
1
T f = |x|− 2 f
2p
Finally, to see that T does not map Lp into L p+2 , consider the function
− p1 − p+2
x 7→ |x| [log(|x| + |x|−1 )] 2p .
Lemma 6.1.7. Let f ∈ Lq,∞ with q > 1 and let E be a set of finite measure.
Then
1
|hf, χE i| . kf kLq,∞ |E| q0
so that
Z Z ∞
|f χE | dx = |{x ∈ E : |f | > α}| dα.
0
Setting
− 1q
α0 = kf kLq,∞ |E| ,
we therefore have
Z Z α0 Z ∞
|f χE | dx . |E| dα + α−q kf kqLq,∞ dα
0 α0
1
. kf kLq,∞ |E| , q0
as desired.
for all sets of finite measure. As we may assume q > 1 (cf. q0 6= q1 ), we may
apply the previous lemma to deduce
1 1
|hT χF , χE i| . kT χF kLq,∞ kχE kLq0 . |F | p |E| q0 ,
With the preliminaries in place, we are now ready to prove the Marcinkiewicz
interpolation theorem.
kT f kLq . kf kLp
for (p, q) as defined above. By duality and density, it suffices to prove that
for f and g simple functions with finite measure support, say. Furthermore,
we may assume f, g are nonnegative.
Using Lemma 6.1.8, we may assume that assume that
0
|h|T χF |, |χE |i| . min |F |1/pj |E|1/qj
(6.3)
j=0,1
X X ∞
X
f= f · χ{2n−1 ≤f <2n } = 2 n
2−j χAjn ,
n∈Z n∈Z j=1
where Ajn is the set of x such that 2n−1 ≤ f (x) < 2n and the coefficient of
2−j in the binary expansion of 2−n f (x) is 1. In particular, we may write
∞
X X
f= 2−j 2n χFnj ,
j=1 n∈Z
−j
P
In particular, by applying sublinearity and noting that j≥1 2 is finite,
we find that it also suffices to consider f of the form
X
f= 2n χFn
n∈Z
where we have used p ≤ q and the nesting of sequence spaces in the final
step. It therefore remains to verify that
X X p X X
1 X
n p
2 A . 2np |A| ∼ 2np |Fn |,
A n:|Fn |∼A A n:|Fn |∼A n
q0
and similarly for the `B sum. The desired inequality follows from the fact
that X p X
2n ≤ |2 max 2n |p ≤ 2p 2np
n∈S
n∈S n∈S
for any finite set S ⊂ Z. This completes the proof that T is type (p, q).
6.1. INTERPOLATION 139
( p1 , 1q ) = ( pθ0 + 1−θ θ
p1 , q0 + 1−θ
q1 ) for some θ ∈ (0, 1).
In fact,
1−θ
kT kLp →Lq . kT kθLp0 →Lq0 kT kLp1 →Lq1 .
Aj = kT kLpj →Lqj .
Let us first assume that none of the exponents equal infinity. Writing f =
P
ak χFk , we observe that
θ 1−θ
X p X p
f = sgn(f ) |ak | p0 χFk |ak | p1 χFk ,
f = cf0θ f11−θ ,
One can verify (by linearity of T and the fact that the functions involved
are simple functions) that F is analytic and has at most exponential growth
in z. Furthermore, we claim that by hypothesis we have
(
A0 kf0 kLp0 kg0 kLq00 Re z = 1
|F (z)| .
A1 kf1 kLp1 kg1 kLq10 Re z = 0.
on the support of f , and similarly for the g term. Thus the result follows
from Hölder’s inequality and the assumption that T is type (p0 , q0 ).
Applying the three lines lemma of complex analysis (Lemma A.3.6), we
deduce that
1−θ
|F (θ)| . Aθ0 A11−θ kf0 kθLp0 kf1 kL θ 1−θ
p1 kg0 kLq0 kg1 kLq1
X θ
X p
|ak |χFk = |ak | χFk
p0
A similar argument extends this to the case where T also varies analyt-
ically in z. We leave the following result due to Stein as an exercise.
6.2. SOME CLASSICAL INEQUALITIES 141
z 7→ hTz f, gi
Then
kTθ kLp →Lq ≤ A01−θ Aθ1
for θ ∈ [0, 1] and (p, q) satisfying
( p1 , 1q ) = ( pθ0 + 1−θ θ
p1 , q0 + 1−θ
q1 ).
whenever
1 1
1 < p < r < ∞, 1 < q < ∞, and 1+ r = p + 1q .
142 CHAPTER 6. CLASSICAL HARMONIC ANALYSIS, PART I
In particular,
k|x|−(d−α) ∗ f kLr . kf kLp
whenever
1 α
0 < α < d, 1 < p < r < ∞, and r + d = p1 .
where 1 ≤ p ≤ ∞.
Proof. The first estimate follows from Hölder’s inequality, while the second
0
follows from duality: for h ∈ Lp , we use Fubini’s theorem to write
Proof of Theorem 6.2.2. The first estimate implies the second, cf.
T f = f ∗ g.
We claim that by choosing λ sufficiently small, the first term on the right-
hand side vanishes. To see this, we first need the following bound:
q q
kg · χ|g|≤λ kLa ≤ kgkLa q,∞ λ1− a for any 1 < q < a < ∞. (6.5)
for 0 < c 1.
We are left to consider the contribution of g2 above. In this case, we
need the following:
∂ α = F −1 (iξ)α F
Tm := F −1 m(ξ)F.
ψ ∗ T f = T ∗ ψ ∗ f,
∗ ψ · fˆ.
ψ̂ · Tcf = Td
2
Choosing ψ so that ψ̂ is everywhere nonzero (e.g. by choosing ψ = e−|x| , cf.
Lemma 2.5.9), we find that T is a Fourier multiplier operator, as desired. Let
us finally note here that all Fourier multipliers commute with one another,
as well.
6.2. SOME CLASSICAL INEQUALITIES 145
|∇|s = F −1 |ξ|s F, s ∈ R.
For −d < s < 0 these operators are instead known as ‘fractional integra-
tion’ operators. In this case we can actually compute the inverse Fourier
transform of |ξ|s (in the sense of distributions) and so compute the corre-
sponding convolution kernel. In fact, we already saw a special case of this
in Section 2.6, when we used a symmetry argument to deduce that
Remark 6.2.6. Let us briefly explain the name Sobolev embedding. Just
as one has ‘Lebesgue spaces’, there is also a notion of ‘Sobolev spaces’. In
particular, one defines the spaces
That is, it yields an embedding result between Sobolev and Lebesgue spaces.
Such results have wide application in the field of PDE.
146 CHAPTER 6. CLASSICAL HARMONIC ANALYSIS, PART I
The result just stated only applies to the fractional derivative operators
|∇|s . What about when s ∈ N? For example, do we have
d d
kf kLq . k∇f kLp whenever 1 < p < q < ∞ and q = p − 1.
for any 1 < p < ∞. This question, in turn, boils down to a question about
certain Fourier multiplier operators. Namely, are the operators defined with
multipliers
ξ
mj (ξ) = |ξ|j
Thus we find
N
X
kfN kL∞ ≥ |fN (0)| & cM .
M =1
N
Z X p
s
fN kpLp s
k|∇| = cM M φ(M x) dx
M =1
N
X p Z N
cpM ,
X
= cM M sp |φ(M x)|p dx ∼
M =1 M =1
where we have changed variables and used s = dp . Thus, we may arrange that
kfN kL∞ → ∞ while kfN kLp remains bounded by choosing the coefficients
to obey {cM } ∈ `p \`1 . Recalling M ∈ 2N , we can take cM = [log M ]−1 , say.
Finally, note that we have only stated Theorem 6.2.5 in the setting of
the whole space Rd . Sobolev embedding results for bounded domains are
certainly available (and important), but we do not discuss this topic here.
See e.g. [8] for results of this type.
Theorem 6.3.2. The Hardy–Littlewood maximal operator is weak type (1, 1);
that is,
kM f kL1,∞ . kf kL1 .
As one readily observes that M is strong type (∞, ∞), we may then
apply the Marcinkiewicz interpolation theorem to conclude the following:
Before we begin the proof, let us make a few observations about the
operator M . First, suppose f is a nontrivial function. By considering r ∼
|x| 1, we can observe that
Thus we should not expect M to be strong type (1, 1); instead, mapping into
weak L1 is the natural assertion. Of course, this rate of decay is compatible
with mapping into all higher Lp spaces.
We will need the following lemma.
6.3. HARDY–LITTLEWOOD MAXIMAL FUNCTION 149
1. Choose a ball of largest radius from the remaining collection and add
it to S.
2. Discard from the remaining collection all balls that intersect a ball in
S.
Proof of Theorem 6.3.2. Let α > 0 and consider an arbitrary compact sub-
set K of the set
{x : M f (x) > α}.
We need to show that |K| . α−1 kf kL1 (with implicit constant independent
of K).
Let x ∈ K. Then there exists a radius r(x) > 0 such that
Z
1
|B(x,r(x))| |f (y)| dy ≥ α.
B(x,r(x))
We therefore have [
K⊂ B(x, r(x)),
x∈K
We now write
X X
|K| ≤ |3Bj | ≤ 3d |Bj |.
S S
The details are left as an exercise, but we sketch the idea as follows. First,
one should show that it suffices to treat functions in L1 (not merely locally
integrable). Next, one should observe that the result follows for smooth,
compactly supported functions. Finally, one should extend this to L1 by
approximation; it is here that the maximal function estimate will come into
play.
Let us next discuss a generalization of the result above that we will need
later. For a locally integrable function ω : Rd → [0, ∞), we define a measure
via Z
ω(E) = ω(x) dx.
E
(cf. Exercise A.4.3). In fact, this inequality is a consequence of the fact that
(M ω)(E) = 0 implies |E| = 0.
For the weak type (1, 1) bound, the argument is similar to that appearing
in the proof of Theorem 6.3.2. One constructs the set K, the balls Bj , and
the subcollection S as in that proof. This time we need to estimate ω(3Bj )
(where 3Bj := Bj (xj , 3rj ), which we do by writing
Z Z
ω(3Bj ) = ω(x) dx ≤ ω(x) dx ≤ 4d |Bj |M ω(y)
3Bj |x−y|<4rj
We now have enough tools to prove Theorem 6.3.9 when p ∈ (2, ∞).
Proof of Theorem 6.3.9 for p ∈ (2, ∞). Recall from Theorem 6.3.6 that we
have Z Z
|M fn (x)| ω(x) dx . |fn (x)|2 M ω(x) dx,
2
so that Z Z
2
|M̄ f (x)| ω(x) dx . kf (x)k2`2 M ω(x) dx
for any locally integrable ω. Now let 2 < p < ∞ and set q = ( p2 )0 . Then by
duality we have
Z
2
2 M̄ f 2 ω dx
kM̄ f kLp = k M̄ f k p2 = sup
L
kωkLq =1
Z
. sup kf (x)k2`2 M ω dx
kωkLq =1
It remains to establish the weak type (1, 1) bound in Theorem 6.3.9, for
then all of the remaining cases follow from Marcinkiewicz interpolation. For
this, we introduce the so-called Calderon–Zygmund decomposition.
Lemma 6.3.11 (Calderon–Zygmund decomposition). Let f ∈ L1 (Rd ; `2 (C))
and set α > 0. There exists a decomposition f = g + b such that g, b have
the following properties:
• kg(x)k`2 ≤ α for a.e. x ∈ Rd .
P
• We have g = f (1 − χQk ).
Proof. We begin by decomposing Rd into a mesh of equal-sized nonoverlap-
ping cubes whose common diameter is large enough that
Z
1
|Q| kf (x)k`2 dx ≤ α
Q
154 CHAPTER 6. CLASSICAL HARMONIC ANALYSIS, PART I
stop and select Q0 as one of the cubes Qk . Note that in this case we have
Z Z
2d
1
α < |Q0 | kf (x)k`2 dx ≤ |Q| kf (x)k`2 dx ≤ 2d α.
Q0 Q
If instead Z
1
|Q0 | kf (x)k`2 dx ≤ α,
Q0
then we subdivide further into 2d congruent cubes and repeat the same
selection process for each of the resulting cubes. We continue subdividing
until (if ever) we are forced into the first case.
We repeat this with all of the cubes in the mesh, leading
P to the collection
of nonoverlapping cubes Qk . We then define b = f · χQk , which has the
desired bounds by construction.
It remains to verify that kg(x)k`2 ≤ α for a.e. x ∈ Rd , where g = f − b.
To this end, note that for any x ∈ / ∪Qk , there exists a sequence of cubes
Q 3 x with diameter tending to zero and such that
Z
1
|Q| kf (y)k`2 dy ≤ α.
Q
Remark 6.3.12. The same proof and decomposition works for scalar valued
f . In this case we can extend the result to get a mean zero condition for b,
namely Z
1
|Qk | b dx = 0
Qk
for each k. Indeed, in this case we let
(
f (x) x∈/ ∪Qk ,
g(x) = 1
x ∈ Q◦k
R
|Qk | Qk f (y) dy
6.3. HARDY–LITTLEWOOD MAXIMAL FUNCTION 155
Proof of Theorem 6.3.9. Let α > 0 and f ∈ L1 (Rd ; `2 (C)). We use the
Calderon–Zygmund decomposition to write f = g+b as above. In particular,
We can first estimate by Tchebychev and the strong type (2, 2) bound for
M̄ :
In particular, writing 2Qk for the dilate of Qk with the same center, we have
∪2Qk . 2d α−1 kf kL1 ,
. M b̃n (x),
ξ
mj (ξ) = |ξ|j (known as Riesz transforms) are bounded on Lp spaces.
To answer this question (as well as to understand some other fundamen-
tal operators in harmonic analysis) we will need to develop what is known
as Calderón–Zygmund theory. This theory addresses the case of ‘singular
integral operators’. The precise definition we need is the following.
Definition 6.4.1. A Calderón–Zygmund convolution kernel is a func-
tion K : Rd \{0} → C that obeys
(a) |K(x)| . |x|−d uniformly for x 6= 0,
for some c.
Using this, we readily observe that (a) holds. As for (b), let us con-
sider (without loss of generality) the case j = 1 and d ≥ 2. Define A =
diag(−1, −1, 1, . . . , 1) and consider the change of variables y = Ax. As
det A = 1, |y| = |Ax|, and y1 = −x1 , this shows
Z Z
−(d+1)
xj |x| dx = − xj |x|−(d+1) dx,
R1 ≤|x|≤R2 R1 ≤|x|≤R2
158 CHAPTER 6. CLASSICAL HARMONIC ANALYSIS, PART I
which shows this integral is zero. Finally, for (c) we observe that
|∇Kj | = O(|x|−(d+1) ),
and hence the desired inequality follows from Lemma 6.4.2 below.
Lemma 6.4.2. A kernel K obeys (c) whenever |∇K(x)| . |x|−(d+1) .
Proof. We use the fundamental theorem of calculus to write
Z 1
K(x + y) − K(x) = ∇K(x + θy) · y dθ.
0
Thus
Z Z 1Z
|K(x + y) − K(x)| dx . |y||∇K(x + θy)| dθ dx
|x|≥2|y| 0 |x|≥2|y|
Z 1Z
. |y| |x + θy|−(d+1) dx dθ
0 |x|≥2|y|
Z
. |y| |x|−(d+1) dx . 1.
|x|≥2|y|
This kernel clearly satisfies (a) and (b), while to verify (c) we compute the
derivative of K.
6.4. CALDERÓN–ZYGMUND THEORY 159
f 7→ K ∗ f := lim Kε ∗ f
ε→0
and hence
kK ∗ f kL2 . kf kL2 .
Now let 0 < ε1 < ε2 . Then we may write
The contribution of the first term above is uniformly bounded due to the
assumption that K is a Calderón–Zygmund kernel. Now consider the second
term. The part containing the characteristic function is nonzero if
or if
|x + y| > ε−1 } and ε ≤ |x| ≤ ε−1 .
|x + y| < ε OR (6.8)
6.4. CALDERÓN–ZYGMUND THEORY 161
Consider the scenario (6.7). Suppose |x| < ε. Then the bounds |x+y| ≥ ε
and |x| ≥ 2|y| imply that |x| ≥ 32 ε. Thus we are led to estimate
Z Z
|K(x)| dx . |x|−d dx . 1
2 2
3
ε<|x|<ε 3
ε<|x|<ε
uniformly in ε > 0. If instead |x| > ε−1 , then we similarly deduce that
|x| < 2ε−1 , and we instead estimate
Z
|x|−d dx . 1
ε<|x|<2ε−1
uniformly in ε > 0.
The scenario (6.8) is similar. That is, if |x + y| < ε then we deduce
ε ≤ |x| ≤ 2ε, while if |x + y| > ε−1 then we deduce 23 ε−1 ≤ |x| ≤ ε−1 . In
particular, we obtain a suitable estimate in either case.
This completes the proof that Kε is a Calderon–Zygmund kernel (with
implicit bounds independent of ε).
To complete the the proof, we will show that kK̂ε kL∞ . 1 (uniformly
in ε > 0), which implies that Kε : L2 → L2 boundedly (uniformly in ε).
Indeed,
kKε ∗ f kL2 ∼ kK̂ε fˆkL2 . kK̂ε kL∞ kf kL2
by Plancherel’s theorem.
We fix ξ ∈ Rd \{0}. Then (up to constants depending only on π, d),
Z
K̂ε (ξ) = e−ixξ Kε (x) dx
Z Z
−ixξ
= e Kε (x) dx + e−ixξ Kε (x) dx.
|x|≤|ξ|−1 |x|>|ξ|−1
We begin by writing
Z
2 e−ixξ Kε (x) dx
|x|>|ξ| −1
Z Z
−ixξ
= e Kε (x) dx − e−ixξ eiπ Kε (x) dx
|x|>|ξ|−1 |x|>|ξ|−1
Z Z
−ixξ
= e Kε (x) dx − e−ixξ Kε (x + πξ
|ξ|2
) dx,
|x|>|ξ|−1 |x+ πξ2 |>|ξ|−1
|ξ|
R = {|x| ≤ 2π|ξ|−1 , |x + πξ
|ξ|2
| > |ξ|−1 }.
The term (6.9) is bounded uniformly by property (c). The second term
(6.10) is bounded uniformly by property (a). For the third term, we note
that in the region R we have
|ξ|−1 ≤ |x + πξ
|ξ|2
| ≤ 3π|ξ|−1 ,
and hence term (6.11) is again uniformly bounded by property (a). This
completes the proof.
Our next main result states that Calderón–Zygmund operators are weak
type (1, 1) and strong type (p, p) for all 1 < p < ∞.
Proof. We show the bounds and leave the extension as an exercise. Given
α > 0 and f ∈ L1 , we perform a Calderon–Zygmund decomposition and
write f = g + b, where the support of b is a union of nonoverlapping cubes
Qk , with
Z Z
1 1
|Qk | b = 0 and |Qk | |b| . α for each k,
Qk Qk
|{|Kε ∗ g| > 12 α}| . α−2 kKε ∗ gk2L2 . α−2 kgk2L2 . α−1 kf kL1 ,
where we recall that |g| . α and the definition of g (cf. Remark 6.3.12).
Now consider the contribution of b. We ∗
√ let Qk denote the cube centered
at xk (the center of Qk ) and dilated by 2 d. Then
X X √
| ∪ Q∗k | ≤ |Q∗k | = (2 d)d |Qk | . α−1 kf kL1
Thus
Z XZ
−1
(6.12) . α |Kε (x − y) − Kε (x − xk )| |b(y)| dy dx
∩(Q∗k )c k Qk
X Z Z
−1
. α |Kε (x − y) − Kε (x − xk )| dx |b(y)| dy.
k Qk (Q∗k )c
Now write
Z Z
|Kε (x−y)−Kε (x−xk )| dx = |Kε (x+xk −y)−Kε (x)| dx.
(Q∗k )c −{xk }+(Q∗k )c
Now we claim that |x| ≥ 2|xk − y| for x, y in the appropriate sets above.
Indeed, writing `(Qk ) for the sidelength of Qk , we first have |xk − y| ≤
√
d
√
2 `(Q k for y ∈ Qk , while |x| ≥
) d`(Qk ). Thus condition (c) applies and
this integral is bounded uniformly, leading to
XZ
−1
(6.12) . α |b(y)| dy . α−1 kf kL1 ,
k Qk
as desired.
Marcinkiewicz interpolation now yields (p, p) bounds for 1 < p ≤ 2
(uniformly in ). It remains to treat the case 2 < p < ∞. To this end, we
fix 2 < p < ∞ and write
ZZ
kKε ∗ f kLp = sup Kε (x − y)f (y)ḡ(x) dx dy
kgk 0
Lp =1
Z Z
= sup f (y) Kε (x − y)g(x) dx dy
1
x − a
Hχ[a,b ](x) = π log ,
x−b
which belongs to neither L1 nor L∞ . Indeed, we can write
χ[a,b] (x − y)
Z
1
Hχ[a,b] (x) = π lim dy.
ε→0 ε≤|y|≤ε−1 y
Similar considerations treat the cases a < x < b and x < a (choosing ε small
enough), and so the identity follows. (See the exercises.)
6.5 Exercises
Exercise 6.5.1. Show that that
belongs to L1 (R) if and only if θ > 1. Use this fact to fill in the details in
Remark 6.1.6.
Exercise 6.5.2. Prove the estimates (6.5) and (6.6). Hint: Write the norms
in terms of the integral of the distribution function.
Exercise 6.5.3. Use the Hardy–Littlewood maximal inequality (Theorem 6.3.2)
to prove the Lebesgue differentiation theorem (Proposition 6.3.5).
Exercise 6.5.4. Fill in the details of the proof of Theorem 6.3.6.
Exercise 6.5.5. Show that if fn (x) = χ[2n−1 ,2n ] (x) then f = {fn } ∈ L∞ but
|M̄ f (x)|2 ≡ ∞.
Exercise 6.5.6. Fill in the details in Remark 6.4.6.
Chapter 7
Indeed, X
ψN (x) = ϕ( Nx2 ) − ϕ( N
2x
1
).
N1 ≤N ≤N2
166
7.1. MIHLIN MULTIPLIER THEOREM 167
α1 +α2 =α
X
. N −|α2 | |ξ|1−|α1 | χ{|ξ|∼N } .
α1 +α2 =α
Thus
. N d+1−|α| .
168 CHAPTER 7. CLASSICAL HARMONIC ANALYSIS, PART II
uniformly in x. Thus
X X
|∇m̌(x)| . N d+1 + N −1 |x|−(d+2) . |x|−(d+1) ,
N ≤|x|−1 N >|x|−1
as needed.
Let us now prove condition (c) assuming (7.1) holds only up to |α| ≤
d d+1
2 e. By Plancherel and the computation above,
Z Z
|xα m̌N (x)|2 dx ∼ |∂ξα mN (ξ)|2 dξ
X Z
. |ξ|−2|α1 | N −2|α2 | dξ
α1 +α2 =α |ξ|∼N
. N d−2|α| .
Thus we have
Z XZ
|m̌(x + y) − m̌(x)| dx ≤ |m̌N (x + y) − m̌N (x)| dx
|x|≥2|y| N |x|≥2|y|
X X Z
. N |y| + |m̌N (x)| dx
|x|≥|y|
N ≤|y|−1 N >|y|−1
X d d+1
.1+ (N |y|) 2 −d 2
e
. 1.
N >|y|−1
Remark 7.1.2. This result is sharp in the sense that L1 and L∞ bounds
can fail. To see this, let us re-use the Hilbert transform example, which
essentially corresponds to taking m̌(x) = x1 in d = 1. By using contour
integration (say), one can verify that m(ξ) is a multiple of the signum func-
tion. In particular, it satisfies (7.1). However, as we saw before, the Hilbert
transform is not bounded on L1 or L∞ .
In particular, Z
fN (x) = N d ψ̌(N y)f (x − y) dy.
170 CHAPTER 7. CLASSICAL HARMONIC ANALYSIS, PART II
\ ξ ˆ
fd
≤N (ξ) = P≤N f (ξ) = ϕ( N )f (ξ).
P
Finally, we have P>N = I − P≤N and PN ≤·≤M = N ≤K≤M PK .
2
P̃≤N f := e∆/N f
This viewpoint is useful in settings in which one may not have a nice notion
of a Fourier transform, but one can still solve the heat equation (e.g. on
manifolds).
and d
− dq
kf≤N kLq . N p kf≤N kLp .
Remark 7.2.4. The estimates in (d) and (e) are known as Bernstein esti-
mates. Item (c) may fail in L1 and L∞ . To see this, first observe that for
each N , we have that P≤N f ∈ C ∞ (since P≤N is convolution with a smooth
function), but C ∞ is not dense in L∞ . Alternately, note that f ≡ 1 belongs
to L∞ , while Z
fN (x) = ψ̌(y) dy ≡ 0.
To see why (c) fails in L1 , note that any individual piece fN (and hence any
finite sum of pieces) has mean zero, while mean zero functions are not dense
in L1 . Indeed, Z
fN (x) dx = fˆN (0) = 0.
and similarly for P≤N . This argument also proves (d), since
The following lemma was used in the proof above and may be of inde-
pendent interest.
Lemma 7.2.5. Suppose K ∈ S is nonnegative. For any N > 0, we have
Z
N d K(N (x − y))|f (y)| dy .K M f [x].
7.2. LITTLEWOOD–PALEY THEORY 173
where we sum over dyadic R > 1 and the implicit constants depend only on
khxid+1 KkL∞ .
Then
kSf kLp ∼ kf kLp for all 1 < p < ∞.
for any measurable, bounded f, g. Note that E{Xn } = 0 for the random
variables defined above.
174 CHAPTER 7. CLASSICAL HARMONIC ANALYSIS, PART II
λ2
X − P
P{ cn Xn > λ} ≤ 2e
2 c2
n.
Indeed,
X 2 X
E cn Xn = E cn cm Xn Xm
n,m
X X
= c2n E{Xn2 } + cn cm E{Xn } E{Xm }
n n6=m
X X X
= c2n + cn cm E{Xn } E{Xm } = c2n ,
n n6=m n
Thus
Z
p
kSf kpLp
X
∼E XN fN (x) dx
where X
mX (ξ) := XN ψN (ξ).
N
where we use the fact that for ψ ∈ Cc∞ (Rd \{0}) only finitely many of the
terms above will contribute to the sum.
Remark 7.2.8. The proof shows that the bound kSf kLp . kf kLp holds
for a wider class of ‘square functions’. In particular, instead of using the
multiplier ψ (that defines P1 ), one could use any Cc∞ function supported in
Rd \{0}.
It remains to establish kf kLp . kSf kLp . This will actually depend on
the fact that ψN is a partition of unity. Define
where S̃ denotes the square function defined in terms of the multipliers for
P̃N . This completes the proof.
X 1
2
s 2s 2
k|∇| f kLp ∼s,p
N |fN (x)|
(7.2)
N Lp
1
X 2s 2
s 2
k|∇| f kLp ∼s,p
N |f>N (x)|
(7.3)
N Lp
Proof. Using the (proof of) the square function estimate, we first observe
1
X 2s 2
−s 2
N |P N |∇| g|
. kgkLp .
N Lp
Taking g = |∇|s f now yields RHS(7.2). LHS(7.2). For the reverse in-
equality, we use Plancherel, Cauchy–Schwarz, Hölder, and the estimate just
established to write
Z X
hg, hiL2 = N s |∇|−s PN g · N −s |∇|s P̃N h dx
Z X 1 1
2 2
2s −s 2 −2s s 2
. N |PN |∇| g| N |P̃N |∇| h| dx
N
X 1 X 1
.k N 2s |PN |∇|−s g|2 2
kLp k N −2s |P̃N |∇|s h|2 2
kLp0
X 1
.k N 2s |PN |∇|−s g|2 2
kLp khkLp0 .
0
Taking the supremum over h ∈ Lp and applying this to g = |∇|s f yields
LHS(7.2).RHS(7.2).
We turn to (7.3). We will show RHS(7.2)∼RHS(7.3). Using
fN = f≥N − f≥2N
where in the last step we have used Schur’s test (Lemma A.3.4). This
completes the proof.
Similarly,
X 1 X
k N 2s |M (M f g>N/4 )|2 2 kLp . kM f ( |N s g>N/4 |2 )1/2 kLp
. kM f kLq1 k|∇|s gkLq2
. kf kLq1 k|∇|s gkLq2 .
Then for any 0 < s < 1, 1 < p, p1 < ∞ and 1 < p2 ≤ ∞ with
1 1 1
p = p1 + p2 ,
we have
k|∇|s (F ◦ u)kLp . kG(u)kLp2 k|∇|s ukLp1 .
We will need the following lemma, similar in spirit to Lemma 7.2.5
Lemma 7.2.12. Suppose |h(x)| ≤ g(x), where g is a radial decreasing func-
tion with limr→∞ rd g(r) = 0. Then
Thus
Z Z ∞
|(h ∗ f )(x)| ≤ χ(0,ρ) (|y|)(− ∂g
∂r )(ρ)|f (x − y)| dρ dy
0
Z ∞ Z
≤ |f (x − y)| dy (− ∂g ∂r )(ρ) dρ
0 |y|<ρ
Z ∞
. M f (x) ρd (− ∂g
∂r )(ρ) dρ
0
Z ∞
. M f (x) g(ρ)ρd−1 dρ . kgkL1 M f (x).
0
so that
Z
|[PN (F (u))](x)| ≤ N d |ψ̌(N y)||u(x − y) − u(x)|[G ◦ u(x − y) + G ◦ u(x)] dy.
(7.4)
We decompose
X
|u(x − y) − u(x)| ≤ |u>N (x − y)| + |u>N (x)| + |uK (x − y) − uK (x)|.
K≤N
It suffices to treat the case K|y| < 1. We may apply the fattened projection
P̃K (abusing notation and writing the corresponding convolution kernel as
ψ̌) and write
Z
uK (x − y) − uK (x) = K d ψ̌(Kz)[uK (x − y − z) − uK (x − z)] dz
Z
= K d [ψ̌(K(z − y)) − ψ̌(Kz)]uK (x − z) dz
Z Z 1
d
= K Ky · ∇ψ̌(Kz − θKy) dθ uK (x − z) dz.
0
which is acceptable.
7.2. LITTLEWOOD–PALEY THEORY 181
|PN (F ◦ u)(x)|
≤ M (u>N G ◦ u)(x) + M (u>N )(x)|G ◦ u(x)|
+ |u>N (x)|M (G ◦ u)(x) + |u>N (x)(G ◦ u)(x)|
X Z
+ N d K|y||ψ̌(N y)|[M uK (x − y) + M uK (x)]
K≤N
× {|G ◦ u(x − y)| + |G ◦ u(x)|} dy.
Arguing similarly,
X 1
(7.7) . kM (G ◦ u)kLp2 k( |M (N s u>N )|2 ) 2 kLp1
. kG ◦ ukLp2 k|∇|s ukLp1 .
182 CHAPTER 7. CLASSICAL HARMONIC ANALYSIS, PART II
Using this and arguing as above suffices to treat (7.8) and (7.9). The proof
of (7.10) (and (7.8) and (7.9)) is left as an exercise.
Indeed, this follows from F[e2πixξj ] = δ(ξ − ξj ). This shows that Tm multi-
plies plane waves (adding their frequencies) and modulates their amplitude
by m(ξ1 , ξ2 ).
Our goal will be to prove Lp × Lr → Lq mapping properties for bilin-
ear operators of this type. We will consider multipliers/operators of the
following type.
Definition 7.3.1. We call m : Rd × Rd → C a Coifman–Meyer symbol if
it obeys
|∂ξα11 ∂ξα22 m(ξ1 , ξ2 )| .|α1 |,|α2 |,d (|ξ1 | + |ξ2 |)−|α1 |−|α2 |
for all multiindices α1 , α2 . We call Tm a Coifman–Meyer multiplier.
7.3. COIFMAN–MEYER MULTIPLIERS 183
. N d hy − xi M f (x),
d
7.3. COIFMAN–MEYER MULTIPLIERS 185
Proof. We write
X
πhh (f, g) = πhh (P̃N PN f, P̃M PM g),
N,M
on the support of ψN (ξ1 )ψM (ξ2 ). Now, using the definition of the Fourier co-
efficients and integration by parts, the Coifman–Meyer condition guarantees
that
cn1 ,n2 . (1 + |n1 | + |n2 |)−100d (7.11)
(see Exercise 8.3.2).
186 CHAPTER 7. CLASSICAL HARMONIC ANALYSIS, PART II
so that
|πhh (f, g)(x)|
X X
. (1 + |n1 | + |n2 |)−100d |PN f (x − n1
CN )| |PM g(x − n2
CN )|.
N ∼M n1 ,n2
and hence, using Hölder’s inequality, the vector maximal inequality (Theo-
rem 6.3.9) and the Littlewood–Paley square function estimate (Theorem 7.2.6)
1 X 1
X 2 2
2 2
kπhh (f, g)kLr .
|M PN f (x)| |M PN g(x)|
N N Lr
X 1
X 1
2
2
2 2
.
|M PN f (x)|
|M P N g(x)|
p
L Lq
N N
1
1
X 2
X 2
2 2
.
|PN f (x)|
|PN g(x)|
Lp Lq
N N
. kf kLp kgkLq ,
1 1
provided 1 < p, q, r < ∞ and p + q = 1r .
7.3. COIFMAN–MEYER MULTIPLIERS 187
where
mN (ξ1 , ξ2 ) = m(ξ1 , ξ2 )ϕ̃ N (ξ1 )ψ̃N (ξ2 ).
8
with
|cn1 ,n2 | . h|n1 | + |n2 |i−100d .
0
We estimate the Lr norm by duality. We fix h ∈ Lr . Noting that
P≤ N f PN g = P̃N [P≤ N f PN g],
8 8
Note that without the assumption of compact support inside (a, b), the
best possible decay is λ−1 , which is realized by φ(x) = x and ψ(x) = 1.
eiλφ(x) = 1 d iλφ(x)
iλφ0 (x) dx e
so that
|I(λ)| . λ−1 .
1 d
Df (x) = iλφ0 (x) dx f (x).
Dt f (x) = − dx
d 1
iλφ0 (x) f (x) .
This yields
Z
|I(λ)| . |(Dt )N ψ| dx
N
β
∂ ψ(∂ α1 φ) · · · (∂ αk φ)
X X
−N
.λ
(φ0 )N +k
L1 ([a,b])
k=0 |β|+|α1 |+···+|αk |=N
.N λ−N .
satisfies
1
|I(λ)| .k λ− k ,
Note that if k = 1 then one needs more than just |φ0 | ≥ 1. This can be
seen first by noting that
Z b Z b
iφ(x)
e dx ≥ cos(φ(x)) dx .
a a
where we have used the inductive hypothesis and the fact that
1
|∂ k φ(δ − k y)| ≥ δ −1 δ ≥ 1.
1
as can be seen by choosing δ ∼ λ− k+1 .
Case 2. Suppose φ(k) 6= 0 for x ∈ [a, b]. If φ(k) (a) > 0 then we have
(k)
φ (x) ≥ δ for x ∈ [a + δ, b]. Then we can write
Z b Z a+δ
Z b
iλφ iλφ iλφ
e dx ≤ e dx + e dx
a a a+δ
1
≤ δ + (δλ)− k
1
. λ− k+1
choosing δ as above. If φ(k) (a) < 0 then we have φ(k) (b) < 0 and so φ(k) (x) <
−δ for x ∈ (a, b − δ). Then we can argue similarly. This completes the
proof.
Proof. We write
Z b Z b Z x
iλφ(x)
e ψ(x) dx = eiλφ(y) dy dx
d
ψ(x) dx
a a a
Z b Z b Z x
iλφ(y) 0
= ψ(b) e dy − ψ (x) eiλφ(y) dy dx.
a a a
192 CHAPTER 7. CLASSICAL HARMONIC ANALYSIS, PART II
Thus
Z b Z b
iλφ iλφ
e ψ dx ≤ |ψ(b)| e dy
a
aZ x
e dy kψ 0 kL1
iλφ
+ sup
x∈[a,b] a
− k1 1
. |ψ(b)|λ + λ− k kψ 0 kL1 ,
where in the final step we use the van der Corput lemma.
satisfies
1 1 1 3
I(λ) = (2πi) 2 λ− 2 [φ00 (x0 )]− 2 eiλφ(x0 ) ψ(x0 ) + O(λ− 2 ) as λ → ∞.
On the other hand, using integration by parts (and the fact that φ0 (x) 6= 0
on the support of ψ away from x = x0 ), we can get
for any N .
To get the exact coefficients, we argue as follows. By Taylor’s theorem
and φ0 (x0 ) = 0
φ(x) − φ(x0 ) = 12 φ00 (x0 )(x − x0 )2 1 + η(x) ,
where
2
b(y) := e−y RN (y)ψ2 (y).
Now, Z
|(7.15)| . |y|N |a( yε )| dy . εN +1 .
. λ−m
1 ε
N +1−2m
.
−1
Now choose ε ∼ λ1 2 , so that
εN +1 ∼ λ−m
1 ε
N +1−2m
.
We have
− N2+1 − 32
|(7.16)| . λ1 . λ1
for N ≥ 2. This completes the proof.
we have
i∂t û(t, ξ) = 12 |ξ|2 û(t, ξ),
so that Z
− 12 2 /2
u(t, x) = (2π) eixξ−itξ û0 (ξ) dξ.
utt − uxx + m2 u = 0.
and so √ √
ξ 2 +m2 ξ 2 +m2
û(t, ξ) = A(ξ)eit + B(ξ)e−it
for some A, B defined in terms of the initial data. In particular, to un-
derstand the asymptotic behavior we need to understand the asymptotic
behavior of Z √
2 2
eixξ±it ξ +m ϕ(ξ) dξ.
Consider first p
Φ = xt ξ + ξ 2 + m2 ,
so that
1
m2
Φ0 = x
t + ξ(ξ 2 + m2 )− 2 , Φ00 = 3 6= 0.
(ξ +m2 ) 2
2
−m( xt )
ξ0 = p
1 − ( xt )2
and so
Z √ √ x 2
2 2 1 1 3 3 −m x
eixξ+it ξ +m ϕ(ξ) dξ ∼ (2πi) 2 t− 2 m− 4 [1−( xt )2 ] 4 eimt 1−( t ) ϕ √ tx 2
1−( t )
provided | xt | < 1.
For the other phase, observe
Z √
2 2
eixξ−it ξ +m ψ(ξ) dξ
Z √
2 2
= e−ixξ+it ξ +m ψ̄(ξ) dξ
Z √
2 2
= eixξ+it ξ +m ψ̄(−ξ) dξ
1 1 3 3
√ x 2 mx
∼ (−2πi) 2 t− 2 m− 4 [1 − ( xt )2 ] 4 e−imt 1−( t ) ψ √ t x 2 .
1−( t )
1
for |x| < t. Alternately, writing ρ = (t2 − |x|2 ) 2 , we can write
1 1
u(t, x) ∼m ρ− 2 Re[i 2 eimρ fˆ(mρ)], |x| < t.
obeys
|I(λ)| .N λ−N for all N > 0.
so that
∇φ · ∇eiλφ
eiλφ = .
iλ|∇φ|2
We can then write
∇φ · ∇eiλφ
Z
I(λ) = ψ dx
iλ|∇φ|2
d Z
X ∂j φ
= −∂j ψ eiλφ dx,
iλ|∇φ|2
j=1
and so
−1
∇φ ψ
−1
|I(λ)| . λ
∇ · |∇φ|2
1 . λ .
L
As mentioned above, the case N ≥ 2 follows from iteration.
We skip the analogue of van der Corput’s lemma, which would yield a
− 1
bound of λ |α| whenever one has lower bounds on ∂ α φ.
Instead, we will move to stationary phase in higher dimensions. The
result is the following.
satisfies
d
(2πi) 2 eiλφ(x0 ) ψ(x0 ) − d − d −1
I(λ) =
∂2φ 1 λ 2 + O(λ 2 )
det[ ∂xi ∂xj (x0 )] 2
as λ → ∞.
2
We also write ( ∂x∂i ∂x
φ
j
) = D2 φ.
The proof is similar in spirit to the d = 1 case. The key step there was to
use a change of variables to turn the phase into an exactly quadratic phase.
The necessary result in higher dimensions is the following:
Lemma 7.4.7 (Morse lemma). Let φ : Rd → R be smooth with a nonde-
generate critical point at x0 . Then there exists a smooth local change of
∂y
variables y = y(x) such that y(x0 ) = 0, ∂x |x0 = Id, and
d
X
1 2
φ(x) = φ(x0 ) + 2 λj yj ,
j=1
Proof of the Morse lemma. Noting that D2 φ(x0 ) is symmetric, we may (by
a change of variables) assume that
D2 φ(x0 ) = diag(λ1 , . . . , λd ).
2
Observe that mij is smooth, with mij = mji and mij (x0 ) = 12 ∂x∂i ∂x
φ
j
(x0 ).
We proceed by induction. Suppose we have found a smooth local change
of variables y = y(x) so that
X
φ(x) = φ(x0 ) + 12 λ1 y12 + · · · + 12 λr−1 yr−1
2
+ yi yj m̃ij (y),
i,j≥r
∂y
where y(x0 ) = 0, ∂x |x=x0 = Id, and the m̃ij are smooth and symmetric.
(The computation above gives the base case r = 1 with y(x) = x − x0 .)
Now we compute
In particular, at x = x0 we have
Thus
we deduce
X
D2 yi yj m̃ij (y) = diag{0, . . . , 0, λr , . . . , λd }.
i,j≥r
for some smooth symmetric m̃0ij . This will imply all of the desired properties
for the new variable y 0 . To this end, we write
X X X
yi yj m̃ij (y) = m̃rr (y)yr2 + yr m̃jr (y)yj + yi yj m̃ij (y).
i,j≥r j≥r+1 i,j≥r+1
and modify the matrix m̃ij accordingly. With this change of variables, one
can now verify all of the desired properties and hence complete the induction.
∂ 2 Φ(x,ξ)
det ∂xi ∂ξj 6= 0.
We will prove:
Remark 7.4.9. Note that for fixed λ, we can easily obtain L2 boundedness:
Z
kTλ f kL2 ≤ kψ(x, ξ)kL2 |f (x)| dx ≤ kψkL2 kf kL2 .
ξ ξ x,ξ
Proof of Proposition 7.4.8. Let Tλ∗ denote the adjoint of Tλ . By the method
of T T ∗ , it suffices to prove that the L2 → L2 norm of Tλ Tλ∗ is bounded by
λ−d . (See Exercise 7.5.5.)
We may write
Z
Tλ Tλ∗ f (ξ) = Kλ (ξ, η)f (η) dη,
where Z
Kλ (ξ, η) = eiλ[Φ(x,ξ)−Φ(x,η)] ψ(x, ξ)ψ̄(x, η) dx.
∆ = ∆(x, ξ, η) = ∇a(x)
x [Φ(x, ξ) − Φ(x, η)]
so that
M (ξ, η)a(x) · (ξ − η) = |ξ − η|.
Now, if the support of ψ is sufficiently small, then we can guarantee
This yields
This implies the desired result provided the support of ψ is sufficiently small.
To deal with the more general case, we employ a partition of unity to split
the support of ψ up into a finite number of sufficiently small pieces. This
completes the proof.
7.5 Exercises
Exercise 7.5.1. Complete the proof of (c) in Proposition 7.2.3.
Exercise 7.5.2. Prove (7.10) and complete the proof of (7.8) and (7.9). Hint:
Expand the inner sum and change the order of integration. Performing the
sum in N leads to a bound of
X
1−s s
(K
L) K cK Ls cL ,
K≤L
kT k2 = kT ∗ k2 = kT T ∗ k.
Exercise 7.5.6. In this exercise, you will prove the stationary phase lemma
in the special case of an exactly quadratic phase: Show that for u ∈ Cc∞ (Rd )
and N ≥ 1, we have
N −1 d d π
(2π) 2 hk+ 2 ei 4 sgn Q
Z X
e ix·Qx/2h
u(x) dx = 1 ((Dx ·Q−1 Dx )k u)(0)+SN (u, h),
k=0 (2i)k | det Q| 2 k!
204 CHAPTER 7. CLASSICAL HARMONIC ANALYSIS, PART II
Exercise 7.5.7. Let σ denote the surface measure of the sphere S ⊂ Rd with
d ≥ 2. Show that
d−1
|σ̌(x)| . hxi− 2 ,
where Z p
σ̌(x) = eixξ dσ(ξ) and hxi := 1 + |x|2 .
S
Hint: As dσ is invariant under rotations, we can write
Z Z π
i|x|ξd
σ̌(x) = e dσ(ξ) ∼ ei|x| cos θ [sin θ]d−2 dθ
S 0
where θ is the angle between x and ed . To estimate this integral, use sta-
tionary phase and the van der Corput lemma.
Chapter 8
In this section we discuss some of the basic concepts and results in semiclas-
sical and microlocal analysis, following [20].
205
206 CHAPTER 8. SEMICLASSICAL AND MICROLOCAL ANALYSIS
in x and satisfy
∂xα a(x; h) = O(g)
hj aj .
P
Remark 8.1.2. One calls a a resummation of the formal symbol
Proof. We only sketch the proof; the complete details may be found in [20,
Lemma 2.3.3]. Without loss of generality, take g ≡ 1. One first constructs
a sequence εj → 0 such that if |α| ≤ j,
ε
k(1 − χ( hj ))∂ α aj kL∞
x
≤ h−1
which
Pfor any h > 0 is actually only a finite sum. One can then verify that
a ∼ hj aj .
8.1. SEMICLASSICAL ANALYSIS 207
(ϕ0± )2 = E − V
by setting Z xp
ϕ± (x) = ± E − V (y) dy.
x0
(a± 00
q q
j−1 )
(a±
0 ϕ0± )0 = 0, (a±
j ϕ0± )0 − i √ 0 =0
2 ϕ±
P jj ≥
for 1. We then let a± (x; h) be a resummation of the formal symbol
h a±
j By construction one can check that
.
Indeed, this follows from the fact that the formal series
q ± 00
0 0 − ih (aj )
X
hj a± √
j ϕ± 0
2 ϕ±
j≥0
is identically zero.
These approximate solutions are called WKB solutions (after Wentzel,
Kramers, and Brillouin).
We only considered the case V (x0 ) < 0. In the case V (x0 ) = E (called
a ‘turning point)’, this technique breaks down. In this case one can instead
use a power series expansion for V (x) − E. Solving the ODE to first order
leads to an equation known as the Airy equation, which is solved with spe-
cial functions (the Airy functions). One then needs to patch together the
approximate solutions away from and near the turning points (which will
only be possible for special values of E).
208 CHAPTER 8. SEMICLASSICAL AND MICROLOCAL ANALYSIS
Pseudodifferential operators.
We next define the semiclassical Fourier transform of a Schwartz
function u on Rn by
Z
−n
Fh u(ξ) = û(ξ) = (2πh) 2 e−ixξ/h u(x) dx.
Rn
generalizing the familiar identities obtained for the standard Fourier trans-
form.
Expanding out Fh−1 Fh u = u leads to the identity
ZZ
−n
u(x) = (2πh) ei(x−y)ξ/h u(y) dy dξ.
Our next goal is to make sense of more general operators of the form
ZZ
−n
u(x) 7→ (2πh) ei(x−y)ξ/h a(x, y, ξ)u(y) dy dξ
for some kernel a(x, y, ξ). This requires that we make sense of the integrals
Z
I(a) = I(a; x, y) = ei(x−y)ξ/h a(x, y, ξ) dξ. (8.1)
Suppose a(x, y, ξ) ∈ S3n (hξim ). If m < −n, the integral I(a) converges and
for u ∈ Cc∞ (Rn ) we may define
Z
Aa u(x, h) = ei(x−y)ξ/h a(x, y, ξ)u(y) dy dξ.
for any k > m + n. One can check that Aa defines a continuous linear
operator from Cc∞ to C ∞ ; in particular, (by the ‘Schwartz kernel theorem’)
we may find a distribution K on Rn × Rn (called the distribution kernel
of Aa ) such that
hAa u, vi = hK, v ⊗ ui,
where u, v ∈ Cc∞ , ⊗ denotes tensor product, and h·, ·i denotes the pairing of
distributions and test functions. We denote the distribution kernel by (8.1).
Example 8.1.2. If a = 1, then choosing k > n we may verify that
Aa u(x) = (2πh)n u(x),
so that Z
ei(x−y)ξ/h dξ = (2πh)n δ(y − x)
Proposition 8.1.4. For a ∈ S3n (hξim ), we can extend Oph (a) to a map
from S → S or from S 0 → S 0 .
Proof. Let us show that Oph (a) : S → S. We write
Z
x ∂x Ik u(x) = xβ ∂xα [ei(x−y)ξ/h (t L)k (au)] dy dξ
β α
for any γ > 0 on {|x − y| ≤ 21 |x|}. Choosing γ > |β| + n, this contribution
is integrable with respect to y and ξ.
For the remaining region, we write
1
L̃ = 1+|x−y|2
(1 + h(x − y)Dξ ).
If X
a(x, y, ξ) = bα (y)ξ α
|α|≤m
8.1. SEMICLASSICAL ANALYSIS 211
for all u, v ∈ S.
where
Z
ch (x, z, η) := (2πh)−n ei(x−y)(ξ−η)/h a(x, y, ξ)b(y, z, η) dy dξ.
which satisfies
h|α| α α
X
a#b ∼ ∂ ∂ (a(x, z, η)b(z, y, ξ))η=ξ, z=x
i|α| α! z η
|α|≥0
0
in S3n (hξim+m ).
where
Z
−n
cδ (x, y, ξ) := (2πh) ei(x−z)(η−ξ)/h−δhzi−δhηi a(x, z, η)b(z, y, ξ) dz dη.
0
We will show that cδ = O(hξim+m ) uniformly in δ, so that we may pass to
a limit c0 as δ → 0 (by dominated convergence). We will then show c0 ∈
0
S3n (hξim+m ), which allows us to send ε → 0 (and interpret the resulting
integral in the sense of oscillatory integrals).
We turn to the details. We define
|η−ξ|2 |x−z|2 −1 η−ξ x−z
L1 = 1 + h2
+ h2
[1 − h Dz + h Dη ].
8.1. SEMICLASSICAL ANALYSIS 213
Thus, for example, if m ≥ 0, we can deduce (writing hηim . hξim +hξ −ηim )
that
1 0
|(2πh)n dδ (x, y, ξ)| . hk−n− 2 hξim+m ,
which is acceptable. We leave the remaining case m < 0 as an exercise (hint:
split into regions where |η| ≤ 21 hξi and |η| > 12 hξi and estimate each piece
separately). Similarly one can deduce that
1 0
|(2πh)n eδ (x, y, ξ)| . hk−n− 2 hξim+m
for k ≥ |m| + 2n + 1. We leave this estimate as an exercise. We also note
that one can get the same estimates for any number of derivatives, i.e.
0
|∂ α dδ (x, y, ξ)| + |∂ α eδ (x, y, ξ)| = O(h∞ hξim+m )
uniformly over (x, y, ξ) and δ > 0. In fact, taking derivatives just produces
powers of |x − z| or |η − ξ|, which can always be overcome by choosing k
larger.
It remains to consider fδ , which (undoing the integration by parts) has
the form
fδ (x, y, ξ)
Z
−n
= (2πh) ei(x−z)(η−ξ)/h χ(ξ, η)χ(x, z)e−δhzi−δhηi a(x, z, η)b(z, y, ξ) dz dη.
214 CHAPTER 8. SEMICLASSICAL AND MICROLOCAL ANALYSIS
with appropriate
uδx,y,ξ (z 0 , η 0 ) ∈ Cc∞ (Rnz0 × Rnη0 ).
In particular, by (8.2),
h|α| α α δ
X
fδ (x, y, ξ) = ∂ z ∂η ux,y,ξ (z, η) + SN ,
i|α| α!
z=0,η=0
|α|≤N −1
where
X
|SN | . hN k∂zα ∂ηβ (∂z ∂η )N uδx,y,ξ kL1 (Rnz ×Rnη )
|α+β|≤2n+1
Z
0
. hN hηim hξim dz dη
|η−ξ|≤2, |x−z|≤2
0
. h hξim+m .
N
that are O(h∞ ). (One can compare this to the case of Fourier multiplier
operators; in this case, if |m(ξ)| & 1 then the inverse operator is simply the
1
operator with symbol m(ξ) .)
Oph (a) ◦ Oph (b) = 1 + Oph (r), where r = O(h∞ ) in S3n (1).
Similarly, Oph (b) ◦ Oph (a) = 1 + Oph (r0 ) for some O(h∞ ) symbol r0 .
Proof. LetPus only sketch the first claim. The idea is to construct b in the
form b ∼ hj bj in such a way to guarantee that a#b ∼ 1, where a#b is as
in the composition theorem. To do this, we firstly set b0 = a1 ∈ S3n (hξi−m )
(cf. the chain rule). We can then define bj for j ≥ 1 recursively. For
example, the linear in h terms in the sum will involve cα for |α| = 1 and
b0 , along with cα for |α| = 0 and b1 , which we use to define b1 (so that the
total contribution is zero). Proceeding in this way, we can construct b as
desired.
In particular, we have:
This ultimately leads to the oscillatory integral above; the asymptotic ex-
pansion is a consequence of the stationary phase theorem (after isolating the
relevant part of the integral, as in the proof of the composition theorem).
In the case t = 0, one can write
σt (−h2 ∆ + V ) = ξ 2 + V (x),
which is independent of t.
The next result we state concerns composition:
Theorem 8.1.10 (Symbolic calculus). Let a = a(x, ξ) ∈ S2n (hξim ) and b =
0 0
b(x, ξ) ∈ S2n (hξim ). For all t ∈ [0, 1], there exists unique ct ∈ S2n (hξim+m )
such that
Opth (a) ◦ Opth (b) = Opth (ct ).
8.1. SEMICLASSICAL ANALYSIS 217
In fact, one can write down a formula and asymptotic expansion for
the symbol ct in the preceding theorem. The proof proceeds by applying
the composition theorem to write the composition in the form Oph (c) for
some symbol c, which then satisfies Oph (c) = Opth (ct ) for suitable ct (by the
previous theorem). The asymptotic expansion is again a consequence of the
stationary phase lemma.
If A and B are pseudodifferential operators with symbols in S2n (hξim )
0
and S2n (hξim ), then for all t ∈ [0, 1] one can show
0
σt (A ◦ B) = σt (A)σt (B) + O(h) in S2n (hξim+m ).
hν a0 = σp (A).
L2 boundedness.
Our final topic will be to consider the L2 boundedness of pseudodiffer-
ential operators. To this point, we have only considered such operators as
acting on S or S 0 . We will prove the following result.
218 CHAPTER 8. SEMICLASSICAL AND MICROLOCAL ANALYSIS
• For all µ, ν ∈ Zd ,
P p
• C0 := µ ω(µ) < ∞.
kSk2m = kS ∗ Skm .
Taking the geometric mean of the previous two estimates yields the bound
1
k · k ≤ [ω(0)ω(µ1 − ν1 )ω(ν1 − µ2 ) · · · ω(µm − νm )] 2 .
Thus, continuing from above, we perform one sum at a time (starting from
the νm sum, then µm , then νm−1 , . . . ) to get
X 1
k(S ∗ S)m k ≤ [ω(0)ω(µ1 − ν1 )ω(ν1 − µ2 ) · · · ω(µm − νm )] 2
|µ` |,|ν` |≤M
X p p
≤ ω(0)C02m−1 ≤ (2M + 1)d ω(0)C02m−1 .
|µ1 |≤M
Hence
kSk2m = k(S ∗ S)km ≤ (2M + 1)d
p
ω(0)C02m−1 .
1
Taking the 2m root of both sides and sending m → ∞ yields the desired
result.
Proof of Theorem 8.1.11. Using Proposition 8.1.9, we may find b ∈ S2n (1)
so that A = OpW h (b). Moreover, by integrating by parts in the integral
expression for b, we can control derivatives of b in terms of derivatives of
a. Thus we may take A = OpW h (a) for some a ∈ S2n (1). Furthermore, by
rescaling ξ 7→ hξ, we can further reduce to proving the theorem for the case
h = 1. That is, we may take A = OpW 1 (a).
∞ 2n
Now let χ0 ∈ Cc (R ) yield a partition of unity
P through the translations
χµ (z) = χ0 (z − µ) for µ ∈ Zd . In particular, µ χµ ≡ 1. We set aµ = aχµ
and observe that
Define Aµ = OpW
1 (aµ ), so that
X
Au = Aµ u for all u ∈ Cc∞ (Rn ).
µ
220 CHAPTER 8. SEMICLASSICAL AND MICROLOCAL ANALYSIS
where
Z
−2n y+z
Kµ,ν (x, y) = (2π) ei(xξ−yη−zξ+zη) aµ ( x+z
2 , ξ)āν ( 2 , η) dz dη dξ.
Now, aµ and aν are smooth and compactly supported, so that Kµ,ν is smooth
on R2n . We now use the operator
which obeys
L[ei(xξ−yη−zξ+zη) ] = ei(xξ−yη−zξ+zη) .
Thus for any N ≥ 0, we may integrate by parts N times to get
Z
−2n y+z
Kµ,ν (x, y) = (2π) ei(xξ−yη−zξ+zη) (t L)N [aµ ( x+z
2 , ξ)āν ( 2 , η)] dz dη dξ.
|µ − ν| ∼ |t − s| + |τ − σ|.
Thus, if we write µ = (µ1 , µ2 ) and ν = (ν1 , ν2 ) and define the following set
of (y, z, η, ξ),
then we find
Z Z
|Kµ,ν (x, y)| dy . [1 + |x − z| + |y − z| + |ξ − η|]−N dy dz dη dξ,
Dµ,ν
sup k∂ α ak2L∞ .
|α|≤N
|x − z| + |y − z| ≥ |x − y|
8.1. SEMICLASSICAL ANALYSIS 221
[1 + |µ − ν|]2n+2−N
Z Z
|Kµ,ν (x, y)| dy . dy dz.
(1 + |x − z|)n+1 (1 + |x − y|)n+1
In particular,
Z
sup |Kµ,ν (x, y)| dy . sup k∂ α ak2L∞ · [1 + |µ − ν|]2n+2−N .
x |α|≤N
which then implies kAukL2 ≤ C0 kukL2 for any u ∈ L2 . This completes the
proof.
where
kR1 kL2 →L2 + kR2 kL2 →L2 = O(h∞ ).
222 CHAPTER 8. SEMICLASSICAL AND MICROLOCAL ANALYSIS
OpW
h (a) ≥
1
C1 on L2 (Rn )
hOpW
h (a)u, ui ≥
1 2
C1 kukL2 for all u ∈ L2 .
Write q
B = OpW
h ( a− 1
C2 ),
OpW
h (a −
1
C2 ) = B 2 + hR, where kRkL2 →L2 . 1.
OpW
h (a −
1
C2 ) ≥ −C 0 h,
and hence
OpW
h (a) ≥
1
C2 − C 0h ≥ 1
C1
kT ukL2 (SM,N )
Z
−n −3n/2 0 2 /2h−(x−y 0 )2 /2h
=2 (πh) ei(y −y)ξ/h−(x−y) u(y)ū(y 0 ) dy dy 0 dξ dx,
SM,N
where
SM,N = {|x| ≤ M, |ξ| ≤ N }.
Using that
Z
−n 0
(2πh) ei(y −y)ξ/h dξ → δ(y 0 − y) as N →∞
|ξ|≤N
where we have used the dominated convergence theorem and the fact that
Z
2
e−(x−y) /h dx = (πh)n/2 .
224 CHAPTER 8. SEMICLASSICAL AND MICROLOCAL ANALYSIS
p̃(x, ξ, x∗ , ξ ∗ ) = p(x − ξ ∗ , x∗ ).
where
∞
X
q(x, ξ; h) ∼ hj q̃j (x, ξ) in S2n (1),
j=0
F = hDx − ξ + ∂ξ ψ,
G = hDξ − ∂x ψ,
Qj = Opth (qj ), j = 1, 2.
Using
(hDx − ξ)T = −ihDξ T
and setting Tψ (u) = eψ/h T u, we may also compute
Repeating this argument for q 0 and iterating, we may obtain the expansion
for q.
Next, we let q̃ ∈ S2n (1) be a resummation of the series just obtained,
and we set
R = Πψ [Opth (q) − q̃]Πψ ,
where Πψ is the orthogonal projection onto the range of Tψ (which is a closed
subspace of L2 (R2n ). Using Calderon–Villaincourt and the expansion of q,
we can obtain
Corollary 8.2.9. Let p ∈ S2n (1) and P = Opth (p). Then there exists
p̃(x, ξ; h) ∈ S2n (1) and R(h) : L2 → L2 such that for all u, v ∈ L2 ,
with
X
p̃(x, ξ; h) ∼ hj p̃j (x, ξ) in S2n (1),
j≥0
Theorem 8.2.10 (Sharp Garding Inequality). Let p(x, ξ) ∈ S2n (1) be non-
negative. Then there exists C > 0 such that for all u ∈ L2 and for all h > 0
small enough,
hOpW 2
h (p)u, ui ≥ −ChkukL2 .
Now,
|hR(h)T u, T ui| = O(h∞ )kT uk2L2 = O(h∞ )kuk2L2 .
On the other hand,
where p0 , r ∈ S2n (hξim ) and limh→0 ε(h) = 0. For t ∈ [0, 1] given, let
P = Opth (p). The set
Pu = 0 and kukL2 ≤ 1,
MS(u) ⊂ Char(P ).
(i) ψ = 0 on U c ;
This implies
0 = keψ/h T P uk2L2
= kp(x − 2∂z ψ, ξ + 2i∂z ψ)eψ/h T uk2L2 + O(h)keψ/h T uk2L2 .
This yields
and hence
keψ/h T uk2L2 (U ∩R2n ) = O(h)kT uk2L2 = O(h).
If we now set
then we obtain
kT ukL2 (V ) = O(e−δ/h ),
as desired.
lim E(h) = E0 .
h→0
where the set on the right-hand side is the ‘energy surface’ associated with
E0 .
230 CHAPTER 8. SEMICLASSICAL AND MICROLOCAL ANALYSIS
as a vector field on R2n (the Hamilton field). Given (x0 , ξ0 ) ∈ R2n , we may
solve (locally) the Hamilton-Jacobi system
(
∂t (x(t), ξ(t)) = Hp (x(t), ξ(t))
(x(0), ξ(0)) = (x0 , ξ0 ),
∂p ∂p
i.e. ẋ = ∂ξ and ξ˙ = − ∂x . We denote this solution by
and call
(t, x, ξ) 7→ exp(tHp )(x, ξ)
the Hamilton flow associated to p. We observe that
∂
∂t exp(tHp )(x, ξ) = Hp exp(tHp )(x, ξ).
ẋ = 2ξ and ξ˙ = −2x,
we obtain ẍ + 4x = 0. Thus
exp(tHp )(x0 , ξ0 ) = (x0 cos 2t + ξ0 sin 2t, −x0 sin 2t + ξ0 cos 2t).
Proposition 8.2.14. For any (x0 , ξ0 ) ∈ R2n and for t, s small enough, we
have
exp((t + s)Hp ) = exp(tHp ) exp(sHp )
and
p exp(tHp (x0 , ξ0 )) = p(x0 , ξ0 ).
8.2. MICROLOCAL ANALYSIS 231
∂
Proof. For the second identity, we apply ∂t and observe that the left-hand
side is constant in t.
a null bicharacteristic of p.
We turn to our last main result.
Theorem 8.2.16. Let
hol
p = p(x, ξ, h) = p0 (x, ξ) + ε(h)r(x, ξ, h) ∈ S2n (hξim )
Suppose (x0 , ξ0 ) ∈ R2n and that exp(tHp )(x0 , ξ0 ) exists for t ∈ (T0 , T1 ).
Then
Lemma 8.2.17. For g ∈ S2n (1) (real-valued), there exists c > 0 such that
for all θ > 0 and v ∈ L2 ,
Sketch of the proof of the lemma. Applying Corollary 8.2.8 and Calderón–
Vailliancourt, we firstly obtain
The key to completing the proof is to apply the preceding lemma with
a good choice of function g.
First, recalling the nonvanishing of ∇x,ξ p0 , we may use a local change
of coordinates on a neighborhood V ⊃ W̄δ to coordinates (y1 , . . . , y2n ) for
which
Hp0 = ∂y∂ 1 and (xδ , ξδ ) 7→ (tδ , 0).
We now choose a, b ∈ R and c > 0 so that
{(y1 , y 0 ) : a ≤ y1 ≤ b, |y 0 | ≤ c} ⊂ Wδ ,
• f (0) = 21 αδ , with 0 ≤ f ≤ αδ ;
• χ0 = 1 on [−c/4, c/4];
• χ0 6= 0 on [−c/2, c/2];
8.2. MICROLOCAL ANALYSIS 233
1
• χ≤ 4 outside [−c/2, c/2].
Finally, let g(y) = χ(|y 0 |)f (y1 ) (extended to zero outside of V ), which then
obeys the following:
• Hp0 g = 0 on V c ,
• Hp0 g < 0 on Vδ := {(y1 , y 0 ) : y1 ∈ [−d, a], |y 0 | ≤ c/2},,
• g(x0 , ξ0 ) = 12 αδ , with g ≤ αδ everywhere.
αδ
• g≤ 4 for |y 0 | + |y1 | > c/2.
Now we apply the lemma above to deduce
k(Hp0 g)eθg/h T uk2L2 (Vδ ) ≤ C( h+|ε(h)|
θ2
+ θ)keθg/h T uk2L2 .
Noting that |Hp0 g| & 1 on Vδ , this implies
keθg/h T uk2L2 (Vδ ) . ( h+|ε(h)|
θ2
+ θ)keθg/h T uk2L2 .
For θ small (and h small), this implies
keθg/h T uk2L2 (Vδ ) . keθg/h T uk2L2 (R2n \Vδ ) .
Now we claim that
αδ
−βd)/h
keθg/h T ukL2 (R2n \Vδ ) = O(1 + eθ( 2 ).
Indeed, on V c we have g = 0, and on V \Vδ we may obtain g ≤ 12 αδ − βd
(which is ≥ 41 αδ for β small enough). In particular, restricting to a possibly
smaller neighborhood Vδ0 on which g ≥ 12 αδ − 12 βd, we have
kT ukL2 (Vδ0 ) . O(e−βdθ/2h + e−θ(αδ −βd)/2h . O(e−c̃/h ).
This implies that T u is microlocally exponentially small near (x0 , ξ0 ). (In
fact, the definition used the L∞ -norm, but it is also acceptable to use the L2
norm — see [20, Remark 3.2.4]). In particular, (x0 , ξ0 ) 6∈ MS(u), yielding
the desired contradiction.
We close with one last application.
Example 8.2.3. We return to the eigenvalue problem
Hu = Eu, kukL2 = 1,
with H = −h2 ∆ + V , V ∈ Snhol (1), and E = E(h) → E0 as h → 0. The null
bicharacteristics of P = H − E are given by
˙ = −∇V (x(t)), ξ(0)2 + V (x(0)) = E0 .
ẋ(t) = 2ξ(t), ξ(t)
These are the classical trajectories of energy E0 . In particular, MS(u) is a
union of maximal classical trajectories of E0 .
234 CHAPTER 8. SEMICLASSICAL AND MICROLOCAL ANALYSIS
8.3 Exercises
Exercise 8.3.1. Prove (8.2).
Exercise 8.3.2. Prove (7.11).
Chapter 9
Sharp inequalities
and
kf kL6 ≤ CSob k∇f kL2 for f ∈ Ḣ 1 (R3 ).
Here we use CGN and CSob to denote the best possible constant in these
inequalities. Our goal will be to prove that there exist functions that attain
the best constant. The basic idea is to take an optimizing sequence and
try to prove the existence of a limit, which one then proves is an optimizer.
However, one must contend with a lack of compactness due to the presence
of symmetries that leave the inequalities invariant, namely, translation and
scaling invariance. For example, suppose one already knew that there existed
an optimizer f ∗ to one of these estimates. Then fn := f ∗ (bn x + xn ) would
be an optimizing sequence for any choice of parameters bn ∈ (0, ∞) and
xn ∈ R3 . However, one can readily choose these parameters so that fn
converges weakly to zero (i.e. |xn | → ∞, bn → 0, or bn → ∞; see the
exercises).
Thus, to prove the existence of a limit, we need to restore the loss of com-
pactness. For the case of Gagliardo–Nirenberg, we first perform a rescaling
235
236 CHAPTER 9. SHARP INEQUALITIES
−∆Q + Q − Q3 = 0, Q : R3 → R.
To make sense of this, we only consider functions such that |{|f | > λ}| is
finite for all λ > 0.
This definition guarantees that χ∗S = χS ∗ . Indeed, noting that
(
S λ ∈ (0, 1)
{χS > λ} =
∅ λ ≥ 1,
we find Z 1
χ∗S (x) = χS ∗ (x) dλ = χS ∗ (x).
0
By construction, the rearrangement f ∗ of a function f is a nonnegative,
radial (i.e. spherically symmetric), decreasing function. Furthermore, the
level sets of f ∗ are the rearrangements of the level sets of |f |, that is,
hf, g ∗ hi ≤ hf ∗ , g ∗ ∗ h∗ i.
J1
X J2
X J3
X
f (x) = fj (x − aj ), g(x) = gj (x − bj ), h(x) = hj (x − cj ),
j=1 j=1 j=1
for all k. The key is to prove that χAk → χA∗ in L2 along a subsequence, and
similarly for Bk , Ck . This suffices to complete the proof, since we estimate
along this subsequence
|I(A∗ , B ∗ , C ∗ ) − I(Ak , Bk , Ck )|
≤ kχA∗ − χAk kL2 kχB ∗ kL2 kχC ∗ kL1 + kχA∗ kL2 kχB ∗ − χBk kL2 kχC ∗ kL1
+ kχAk kL2 kχBk kL1 kχC ∗ − χCk kL2
→0 as k → ∞.
We now sketch how to prove the desired convergence. We first note that
each Ak is of the form
2
To see this, one can introduce γ = e−|x| consider the sequence ak =
kγ − χAk kL2 , which converges to a = kγ − χà kL2 . Using symmetries of γ
and the double reflection symmetry of Ã, we can deduce
Z Z Z Z
γRα χà ≤ γXRα χà ≤ γY XRα χà = γRα χà ,
which implies that Rα χà = XY Rα χà . This relies on Exercise 9.3.4. Using
this, we find χà = R2α χà . To complete the argument, one relies on the fact
that any θ may be approximated by irrational multiplies of 2π.
Proof. Recall the heat kernel et∆ introduced in Section 2.6. We compute
Applying the Riesz rearrangement inequality and reversing the steps above,
we find
ZZ
2 1 ∗ 2 − d2 −|x−y|2 /4t ∗ ¯∗
k∇f kL2 ≥ lim t kf kL2 − (4πt) e f (x)f (y) dx dy
t→0
= k∇f ∗ k2L2 .
• Boundedness: there exists M > 0 such that kfn kLp ≤ M for all n.
9.1. SHARP GAGLIARDO–NIRENBERG 241
for all n.
Proof.
R Let φ be a smooth bump function supported on the unit ball with
φ = 1. For R > 0, we define
fnR (x) = φ( R
x
)[Rd φ(R·) ∗ fn ](x)
for all n. We next show that {fnR } are equicontinuous for fixed R > 0. As φ
is smooth, it suffices to show that for any ε > 0, there exists δ > 0 so that
for |y| < δ. As convolution is linear, this follows from the Lp -equicontinuity
of the fn and the estimates above.
By the Arzelá–Ascoli theorem, it follows that for any R > 0, every
subsequence of {fnR } has a subsequence that converges in L∞ ({|x| ≤ R})
and hence in Lp ({|x| ≤ R}).
Now let ε > 0. We claim that there exists R > 0 large enough that
fn − fnR = [1 − φ( R
x x
)]fn + φ( R )[fn − Rd φ(R·) ∗ fn ].
242 CHAPTER 9. SHARP INEQUALITIES
Let us first see that this does the job. We apply this with the sequence
ε = 2−k to find a sequence fn,k satisfying fn,k+1 ∈ B2−k (fn,k ). Such a
subsequence is necessarily Cauchy and hence convergent in Lp .
In fact, by (9.4), it is enough to establish the total boundedness property
(9.5) for any {fnR }. However, this follows from sequential compactness. To
see this, suppose towards a contradiction that for some ε0 , we have
R such that
Then we can inductively build a subsequence fn,k
R R
fn,k+1 ∈
/ Bε0 (fn,k )
Lemma 9.1.8 (Compact embedding). Let Hrad 1 (R3 ) denote the set of radial
Z
. |y|2 |ξ fˆn (ξ)|2 dξ . |y|2 k∇fn k2L2 ,
2 ∞
Z
2 2
r |f (r)| = 2r
f (ρ)∂ρ f (ρ) dρ
Z ∞ r
. ρ|f (ρ)| ρ|∂ρ f (ρ)| dρ
0
Z 1 Z 1
2 2
2 2 2 2
. ρ |f (ρ)| dρ ρ |∂ρ f (ρ)| dρ . kf k2H 1 ,
where in the last line we compute the integral using spherical coordinates.
In particular, we have
Z Z
|fn (x)|4 dx . R−2 |x|2 |fn (x)|4 dx . R−2 kfn k2L2 k|x|fn k2L∞
|x|>R |x|>R
. R−2 kfn k4H 1 ,
With the requisite compactness tools in place, we can now prove exis-
tence of optimizers for Gagliardo–Nirenberg.
244 CHAPTER 9. SHARP INEQUALITIES
−4
for f ∈ H 1 (R3 )\{0}, so that CGN = inf J(f ). We take a sequence fn ∈
1 3
H (R )\{0} satisfying
−4
lim J(fn ) = CGN .
n→∞
1
kfn kL2 kfn kL2 2
bn = and an = 3 ,
k∇fn kL2 k∇fn kL2 2
then
kgn kL2 = k∇gn kL2 = 1.
−4
Note that gn remains an optimizing sequence, i.e. J(gn ) → CGN .
∗
Next, we take radial rearrangements and define hn = gn . By Lemma 9.1.6
and the fact that rearrangements preserve Lp -norms, we have that hn is a
1 (in fact kh k
bounded sequence in Hrad n L2 ≡ 1 and k∇hn kL2 ≤ 1) satisfying
J(hn ) ≤ J(gn ).
N N1 ≤N3
lim sup k∇fn kL2 ≤ A and lim inf kfn kL6 ≥ ε (9.8)
n→∞ n→∞
weakly in Ḣ 1 , with
5
k∇φkL2 & ε( Aε ) 4 . (9.9)
We have the decouplings
and
lim kfn k6L6 − kfn − φn k6L6 − kφn k6L6 = 0,
n→∞
where
−1
φn (x) := λn 2 φ( x−x
λn ).
n
Thus
1 5
kPNn fn kL∞ & εNn2 ( Aε ) 4 ,
and hence there exists xn ∈ R3 such that
−1 5
Nn 2 |PNn fn (xn )| & ε( Aε ) 4 .
The Ḣ 1 decoupling follows from weak convergence and the fact that
(check!).
For the L6 decoupling, we will need the following refined version of Fa-
tou’s lemma due to Brezis and Lieb [1]: if an is a sequence of Lp functions
with lim supn→∞ kan kLp < ∞ and an → a almost everywhere, then
We prove this below. Assuming (9.10) for the moment (see below), we now
claim that 1
λn2 fn (λn x + xn ) → φ
almost everywhere along a subsequence. To see this, we first observe that
H 1 (K) ,→ L2 (K) is a compact embedding for any compact K ⊂ R3 (cf.
Theorem 9.1.7 and the proof of Lemma 9.1.8). Thus, the weak convergence
implies strong L2 convergence along a subsequence for any compact K ⊂
R3 . Using a diagonal argument, we can then deduce almost everywhere
convergence along a subsequence. (See Exercise 9.3.7.)
Thus, by (9.10), we have
1 1
kλn2 fn (λn x + xn )k6L6 − kλn2 fn (λn x + xn ) − φkL6 → kφk6L6 ,
It remains to prove the refined Fatou lemma due to Brézis and Lieb [1].
To see that this holds, consider the cases |y| |x| and |y| & |x| separately;
the first case leads to the first term on the right-hand side, while the second
case leads to the second term.
Continuing from above, we deduce
0 ≤ Wε,n ≤ (1 + Cε )|a|p ∈ L1 .
R
Thus the dominated convergence theorem implies that Wε,n → 0 as n →
∞. Now we observe that
|an |p − |a|p − |an − a|p ≤ Wε,n + ε|an − a|p ,
We can now turn to the main technical tool needed for the proof of
Theorem 9.2.1, namely, a profile decomposition adapted to the Sobolev em-
bedding inequality.
Proposition 9.2.5 (Profile decomposition). Let {fn } be a bounded sequence
∗
in Ḣ 1 (R3 ). There exist J ∗ ∈ {0, 1, 2, . . . } ∪ {∞}, profiles {φj }Jj=1 , positions
∗ ∗
{xjn }Jj=1 , and scales {λjn }Jj=1 such that along a subsequence in n we have
J
X 1 j
fn (x) = [λjn ]− 2 φj ( x−x J
j ) + rn
n
for 0 ≤ J ≤ J ∗.
λn
j=1
• We have Ḣ 1 decoupling:
J
X
2 j 2 J 2
sup lim supk∇fn kL2 −
k∇φ kL2 − k∇rn kL2 = 0.
J n→∞
j=1
Hence, by Ḣ 1 decoupling,
J J
εj−1 2
X X
ε2j−1 A0 ) . k∇φj k2L2 . A20 ,
j=1 j=1
where
J(f ) := kf k6L6 ÷ k∇f k6L2
Let us first normalize the sequence by replacing
fn
fn with k∇fn kL2 ,
Using the nesting `2 ⊂ `6 , this implies that all of the inequalities above are
equalities. In particular,
J ∗ J ∗
X X
k∇φj k6L2 = k∇φj k2L2 = 1.
j=1 j=1
Since the φj are all non-trivial, this implies that there must be only one
profile, say φ, which satisfies k∇φkL2 = 1. We therefore have
−1
fn = λn 2 φ( x−x
λn ) + rn ,
n
1
with λn2 fn (λn x + xn ) * φ and k∇fn kL2 = k∇φkL2 = 1. In particular,
1
λn2 fn (λn x+xn ) converges strongly to φ in H 1 and hence in L6 . In particular,
J(φ) = CSob 6 , i.e. φ optimizes Sobolev embedding.
252 CHAPTER 9. SHARP INEQUALITIES
9.3 Exercises
Exercise 9.3.1. Let f ∈ L2 (Rd ). Show that gn := f (x+xn ) converges weakly
d/2
to zero if |xn | → ∞. Show that hn = λn f (λn x) converges weakly to zero
if λn → 0 or λn → ∞.
Exercise 9.3.2. Prove that the embedding Ḣrad1 (R3 ) ,→ L6 (R3 ) is not com-
pact. Here the subscript ‘rad’ denotes the restriction to radial functions, i.e.
f satisfying f (x) = f (|x|).
Exercise 9.3.3. Investigate the allowed range of exponents in (9.2).
Exercise 9.3.4. Let f, g be nonnegative functions such that |{f > λ}| and
|{g > λ}| are finite for all λ > 0. Then
Z Z
f (x)g(x) dx ≤ f ∗ (x)g ∗ (x) dx.
Rd Rd
253
254 CHAPTER 10. RESTRICTION THEORY AND RELATED TOPICS
Example 10.1.1. There exists a function belonging to Lp for all p > 1 whose
Fourier transform is infinite on an entire hyperplane. Let ψ : Rd−1 → C be
a bump function and set
for ξ ∈ S.
The issue in the previous example is that the hyperplane S has no cur-
vature.
Following [32], we will first focus on the case that S is a quadratic surface;
we return to some more additional cases in Section 10.4. For simplicity, we
will restrict our attention to the case of the paraboloid
ˇ ∗ f k p0 . kf kLp
kdµ (10.2)
L
for the same choice of p. We will look for p such that (10.2) holds. We will
utilize the Stein interpolation theorem, Theorem 6.1.11.
In what follows, we will use the notation
x = (y, xn ) ∈ Rn−1 × R.
10.1. RESTRICTION THEORY 255
lim Kz → c dµ
z→−1
where δ0 is the Dirac delta distribution. Indeed, with (10.4) in place we get
Z Z Z
Kz (x)F (x) dx → c F (−|y|2 , y) dy = c F (x) dµ(x),
as desired.
As γ is assumed to have a simple zero at z = −1, to prove (10.4) it is
enough to establish
In light of the lemma above and (10.2), we have now reduced the restric-
0
tion estimate A. to establishing the Lp → Lp boundedness for the operator
f 7→ T−1 f := Kˇ−1 ∗ f,
with Kz as in (10.3). Writing Tz f = Kz ∗ f , applying Plancherel and Hölder
yields the estimates
kTz f kL2 ≤ kKz kL∞ kf kL2 , i.e. kTz kL2 →L2 ≤ kKz kL∞
kTz f kL∞ ≤ kǨz kL∞ kf kL1 i.e. kTz kL1 →L∞ ≤ kǨz kL∞ .
Then, to apply Stein’s interpolation result (Theorem 6.1.11) we will look for
x0 > 1 and an analytic function γ with a simple zero at z = 1 satisfying the
following on the strip −x0 ≤ Re z ≤ 0:
(i) |γ(x + iy)| has at most exponential growth as |y| → ∞,
(ii) Kz (x) is bounded when Re z = 0.
(iii) Ǩz (x) is bounded when Re z = −x0 .
Then, by Theorem 6.1.11, we can interpolate between the estimates at
Re z = 0 and Re z = −x0 to deduce
2x0
kT−1 kLp →Lp0 . 1, where p= x0 +1 .
Define
γ(z) = Γ(z + 1)−1
and let Kz be as in (10.3). Then γ has a simple zero at z = −1 and satisfies
items (i) and (ii) above. Furthermore, writing x = (y, xn ) as above, we have
2 −[z+ n+1 ]
Ǩz (x) = c exp{− i|y| z
4xn }i xn
2
(10.5)
for some constant c. In particular, item (iii) holds provided
Re z = −x0 := − n+1
2 .
10.1. RESTRICTION THEORY 257
respectively.
One can then check (using Stirling’s formula, say) that |Γ(1 + iσ)| → ∞ as
|σ| → ∞, which yields boundedness of |γ(iσ)|.
Thus, the proof boils down to establishing (10.5). This computation will
also explain the origin of the mysterious choice of γ.
We begin with a change of variables and contour integration. We write
x = (y, xn ) and the dual variable as ξ = (η, ξn ). Then
Z Z Z ∞
z 2
ei[yη+xn ξn ] xn − |y|2 + dx = ei[yη+ξn |y| ] eixn ξn xzn dxn dy
0
Z
z+1 −(z+1) 2
= i ξn Γ(z + 1) ei[yη+ξn |y| ] dy.
by
u(t) = eit∆ φ.
2
Here eit∆ is the Fourier multiplier operator with symbol e−it|ξ| , i.e.
2
eit∆ := F −1 e−it|ξ| F.
where as usual p0 denotes the Hölder dual to p, i.e. the solution to p1 + p10 =
1. This will be an essential ingredient in proving the following Strichartz
estimates.
260 CHAPTER 10. RESTRICTION THEORY AND RELATED TOPICS
Then
keit∆ φkLqt Lrx (R×Rd ) . kφkL2 (Rd ) . (10.11)
which shows Z
∗
T G(x) = e−is∆ G(s, x) ds.
R
In particular, Z
T T ∗ F (t, x) = ei(t−s)∆ F (s, x) ds. (10.12)
R
We now use the dispersive estimate (10.9), the Hardy–Littlewood–Sobolev
inequality (Theorem 6.2.2), and the scaling relation (10.10) to estimate
Z
d d
kT T ∗ F kLq Lr .
|t − s|−( 2 − r ) kF (s)k r0 ds
q
t x Lx Lt
−2
.
|t| q ∗ kF (t)kLr0
Lq
x t
− 2q
. k|t| k q
,∞ kF kLq0 Lr0
Lt2 t x
. kF kLq0 Lr0 ,
t x
10.2. STRICHARTZ ESTIMATES 261
yielding the desired bounds for T T ∗ . Note that the application of Hardy–
Littlewood–Sobolev requires q > 2.
By the method of T T ∗ (see e.g. Exercise 3.5.6), we deduce the desired
L → Lqt Lrx boundedness of eit∆ .
2
Remark 10.2.2. Note that this result covers the special case q = r =
2(d+2)
d covered in the previous section. The endpoint case q = 2 (which is
compatible with (10.10) only for d ≥ 2) is also allowed provided d ≥ 3; it is
valid in d = 2 in the radial setting. However, proving endpoint Strichartz
estimates is a much more challenging problem (see [17]).
Apart from the missing endpoint q = 2, one may ask about the optimality
of the condition (10.10); i.e. are there other exponent pairs for which we
may expect an estimate of the form (10.11)? If we insist on putting the
function φ in L2 , the answer is no. One can check that the scaling relation
is necessary by considering φλ (x) := φ(λx). Then we firstly have
d
kφλ kL2 (Rd ) = λ− 2 kφkL2 .
On the other hand, the solution to the linear Schrödinger equation with data
φλ is
2
eit∆ [φλ (·)](x) = [eitλ ∆ φ](λx),
and so
− 2q − dr
keit∆ [φλ (·)]kLqt Lrx (R×Rd ) = λ keit∆ φkLqt Lrx (R×Rd ) .
In particular, the estimate is only possible if the powers of λ match, which
is exactly (10.10).
One may also ask whether (10.11) could hold with (q, r) satisfying (10.10)
but 1 ≤ q < 2. The answer is also no. To see this, let us give the following
heuristic argument. Fix a nice function φ and consider the linear solution
u(t) = eit∆ φ. Supposing the Strichartz estimate holds in Lqt Lrx , we can
find a sufficiently long time interval around the origin (of length T , say) so
that ‘most’ of the norm is contained this interval. We now consider time-
translates of u by {tj }Nj=1 , which are separated by a length T . Then,
using time-translation symmetry of the equation, we should have
N
X
1
u(t − tj )
&Nq
Lqt Lrx (R×Rd )
j=1
P −itj ∆
the Schrödinger equation with data e φ. Thus, applying Strichartz,
the fact that e−it j ∆ φ and e−it k ∆ φ are almost orthogonal (provided |tj − tk |
is large enough), and Cauchy–Schwarz, we deduce
N
N −it ∆
X
X
1
u(t − tj )
.
e j
φ
. N 2.
Lqt Lrx
j=1 j=1 L2
where u0 denotes the initial condition u|t=0 . In particular, solving the PDE
is equivalent to finding a fixed point of the operator
Z t
it∆
Φu(t) = e u0 − i ei(t−s)∆ F (u(s)) ds.
0
The usual strategy is to utilize the Banach fixed point theorem, which re-
quires proving that Φ is a contraction on a suitable complete metric space.
In particular, this requires mapping properties (i.e. estimates) for the oper-
ators appearing above.
10.2. STRICHARTZ ESTIMATES 263
Defining Z t
T̃ g(t) = K(t, s)g(s) ds,
−∞
Remark 10.2.5. To apply this in the setting of Strichartz estimates for the
Schrödinger equation, we use
Z
0
T f (t) = ei(t−s)∆ f (s) ds, p = q 0 , X = Lr , Y = Lr .
R
Thus we have p < 2 < q, and ei(t−s)∆ ∈ L(X, Y ) by the dispersive estimate.
kf kpLp (R;X) = 1,
264 CHAPTER 10. RESTRICTION THEORY AND RELATED TOPICS
we aim to prove
kT̃ f kLq (R;Y ) . 1.
which implies that f (s) = 0 for almost every s < t with F (s) = F (t). Thus,
in this case we have that both sides of (10.14) are zero (almost surely in s).
10.2. STRICHARTZ ESTIMATES 265
∞ 2X
−1 j
− qj
k(χπ1 Q ◦ F )f kqLp (R;X) ∼ 2 p .
∞ 2Xj
−1 ∞
− qj −j[ pq −1]
kT̃ f kqLq (R;Y )
X X
. 2 p . 2 .1
j=0 k=0 j=0
we have left out the important endpoint estimate of [17], along with many
other extensions and possible refinements. To conclude our discussion, we
will show how to use these estimates to establish a local well-posedness result
for a nonlinear Schrödinger equation. (We also refer the reader to [3] for a
thorough textbook treatment, as well as [18] for an expository introduction
into some more advanced techniques.)
Proof. Let u0 ∈ X and define the sequence {un } ⊂ X inductively via un+1 =
Φ(un ). Observe that by iterating (10.16), we have
Thus there exists u such that un → u in X. Now observe that by the triangle
inequality we have
d(Φ(u), u) ≤ d(Φ(u), Φ(un )) + d(Φ(un ), u)
≤ 12 d(u, un ) + d(un+1 , u) → 0
as n → ∞. Hence Φ(u) = u.
To show uniqueness, we suppose that Φ(u) = u and Φ(v) = v. Then
d(u, v) = d(Φ(u), Φ(v)) ≤ 12 d(u, v),
whence u = v.
to Z t
it∆
u(t) = e u0 + i ei(t−s)∆ |u|2 u(s) ds.
0
Proof. Let T > 0 and C ≥ 1 to be determined below and define
X = {u ∈ L∞ 2 4 ∞
t Lx ∩ Lt Lx ([−T, T ] × R) :
kukL∞ 2 ≤ 2CM
t Lx
and kukL4t L∞
x
≤ 2CM },
with
d(u, v) = ku − vkL∞ 2
t Lx ([−T,T ]×R)
.
Throughout the proof, space-time norms will be taken over [−T, T ] × R.
As (X, d) is a closed subset of the complete metric space L∞ 2
t Lx ([−T, T ]×
R), it follows that (X, d) is a complete metric space.
We let Z t
it∆
Φ(u(t)) = e u0 + i ei(t−s)∆ |u|2 u(s) ds.
0
We will show that (i) Φ : X → X and that (ii) d(Φ(u), Φ(v)) ≤ 21 d(u, v) for
u, v ∈ X.
(i) Let u ∈ X. Then applying Strichartz estimates and Hölder’s inequal-
ity, we have
2
kΦ(u)kL∞ 2 ≤ CM + Ck|u| ukL1 L2
t Lx t x
1
≤ CM + C(2T ) 2 kuk2L4 L∞ kukL∞ 2
t Lx t x
1
≤ CM + C(2T ) (2CM )3 . 2
268 CHAPTER 10. RESTRICTION THEORY AND RELATED TOPICS
Thus if
1
T ≤ ,
128C 6 M 4
where C ≥ 1 is the constant in the Strichartz estimate, we have kΦ(u)kL∞ 2 ≤
t Lx
2CM . Similarly,
kΦ(u)kL4t L∞
x
≤ CM + k|u|2 uk 4
Lt3 L1x
1 1
≤ CM + (2T ) 2 kuk2L∞ L2x kukL4t L∞
x
≤ CM + C(2T ) 2 (2CM )3 ≤ 2CM
t
Thus for
1
T ≤
8192C 6 M 4
we obtain that
d(Φ(u), Φ(v)) ≤ 12 d(u, v).
With (i) and (ii) in place, we conclude that Φ has a unique fixed point,
yielding our desired solution.
We will focus on the case of the sphere. Similar to the paraboloid case,
we will be interested in estimates of the form
kfˆS kLq (S,dσ) . kf kLp (Rd ) ,
(10.17)
fˆ|S = f[
∗ ϕ|S ,
so, by Hölder, Hausdorff–Young, and Young’s inequality, we have
Then Z
ixξ0
(g dσ)ˇ(x) = e ei(ξ−ξ0 )x dσ(ξ).
K
|xd (ξ − ξ0 )d | . 1 if |xd | . R2 ,
while
|x̄(ξ − ξ0 )| . 1 if |x̄| . R.
In particular, (g dσ)ˇ(x) ∼ R−(d−1) throughout the ‘dual tube’ T to D,
centered at the origin, with height R2 and radius R. So
1 d+1
k(g dσ)ˇkLp0 & R−(d−1) |T | p0 & R−(d−1) R p0 .
The restriction conjecture for the sphere states that the necessary
conditions (10.19) and (10.20) are in fact sufficient.
This has been resolved completely in d = 2 [33], while the full result
remains open in higher dimensions.
In the rest of this section, we will discuss a positive result due to Tomas
and Stein.
10.4. MORE RESTRICTION THEORY 271
Proof of Theorem 10.4.2. We will prove the result up to (but not including)
the endpoint p = 2(d+1)
d+3 . We begin by obtaining the smaller range
4d
1<p≤ 3d+1
for any m ≥ 0.
To estimate this integral, we split into two regimes: (i) y ∈ S with
|x − y| . N1 , which has a volume of ∼ N −(d−1) . In this regime the integrand
1
is bounded by N d . (ii) y ∈ S with |x − y| ∼ M for some M N , which
has a volume bound of M −(d−1) (which could of course be replaced by . 1
N
if M ≤ 1). In this case hN (x − y)i ∼ M , and if we choose m = d in the
estimate above we get that the integrand is bounded by M d .
Thus
X
|ψ̂N ∗ dσ(x)| . N d · N −(d−1) + M d M −(d−1) . N.
M ≤N
We now interpolate between the (1, ∞) and (2, 2) estimates and check the
range of p for which we end up with a negative power of N . As p1 = θ + 1−θ
2
2−p
when θ = p , we get
2−p
− d−1 2[ p−1 ]
kf ∗ (ψN σ̌)kLp0 . N 2 p N p kf kLp
2(d+1)
One can check that this power is negative provided p < d+3 . This com-
pletes the proof.
10.5 Exercises
Exercise 10.5.1. Prove (10.6).
Exercise 10.5.2. (Challenge problem.) Develop a profile decomposition adapted
to the Strichartz estimate in order to prove existence of optimizers to this
inequality.
Chapter 11
Additional topics
where
H =∆+V for some V : R3 → R.
To keep things relatively simple, we assume that V is smooth and compactly
supported. We denote the solution to (11.1) by e−itH f ; we will describe this
operator more precisely below.
In this context, ‘scattering’ refers to the statement that as t → +∞
(say), the solution to (11.1) behaves like a solution to the ‘free’ Schrödinger
equation i∂t u = ∆u. That is, for any f ∈ L2 , there exists W+ f ∈ L2 so that
for f ∈ L2 (using the same topology as above). One can consider the same
problem backward in time and construct an analogous operator W− . The
273
274 CHAPTER 11. ADDITIONAL TOPICS
• eitH is unitary on L2 ,
uniformly in t. Using this, we see that we may assume that f, g ∈ Cc∞ (R3 ).
Now, by the fundamental theorem of calculus, we find that the existence of
the limit would be implied by the following estimate:
Z ∞
d heit∆ e−itH f, gi dt < ∞.
dt
1
d it∆ −itH
dt e e f = eit∆ [i∆ − iH]e−itH f = −ieit∆ V e−itH f.
Thus, using unitarity of eit∆ , the dispersive estimate for eit∆ , and Cauchy–
11.1. LINEAR SCATTERING THEORY 275
which is acceptable.
W+ = F −1 FH
such that
2 it|ξ|2
e−itH = FH
∗ −4π
e FH .
Here Φξ is a ‘distorted plane wave’ of frequency ξ that satisfies the general-
ized eigenvalue equation
(H + 4π 2 |ξ|2 )Φξ = 0.
276 CHAPTER 11. ADDITIONAL TOPICS
Note that with this lemma, we already obtain unitarity of e−itH (one of
the facts needed above). Taking this lemma for granted for the moment, let
us prove the second proposition.
Proof of Proposition 11.1.2. We integrate by parts to write
Z ∞ Z ∞
hε e−ε eit∆ e−itH f dt, gi = hf, gi + e−εt h dt
d it∆ −itH
e e f, gi dt.
0 0
Computing as in the proof of the first proposition, we find
Z ∞
hW+ f, gi = hf, gi − lim e−εt hieit∆ V e−itH f, gi dt
ε→0 0
Noting that hf, gi = hFH f, FH gi, we find that the problem reduces to prov-
ing Z ∞
lim e−εt hieit∆ V e−itH f, gi dt = hFH f, FH g − Fgi.
ε→0 0
The left-hand side may be expanded as
Z ∞ ZZ
2 2
lim ci V (x)Φξ (x)e−4π it|ξ| FH f (ξ)eit∆ ḡ(x) dx dξ dt
ε→0 0
ZZ Z ∞
2 2
= lim ci Φξ (x)FH f (ξ)V (x) eit[∆+4π |ξ| +iε] dt ḡ(x) dξ dx
ε→0 0
ZZ
= lim ci Φξ (x)FH f (ξ)V (x)(∆ + 4π 2 |ξ|2 + iε)−1 ḡ(x) dξ dx.
ε→0
We now write
Φξ (x) = e2πixξ + wξ (x),
so that wξ solves
(∆ + 4π 2 |ξ|2 + V )wξ (x) = −V e2πixξ .
(In fact, this is how we will construct Φξ below!) We then rely on another
fact:
11.1. LINEAR SCATTERING THEORY 277
(∆ + 4π 2 k 2 + V )w = h
w = (h − V w) ∗ Gk .
= hFH f, FH g − Fgi,
as desired.
(∆ + 4π 2 k 2 )w = f
The operator on the left-hand side is of the form T + iσ (with T closed and
self-adjoint and σ 6= 0) and hence is guaranteed to be invertible; indeed, one
can verify that
e2πixξ ϕ(ξ/N )
Z
dξ = O(N · (N |x|)−5 ).
(1 + iε)2 − |ξ|2
The main contribution is therefore coming from the region |ξ| ∼ 1, which
requires the most delicate analysis. We sketch the treatment of this term:
Using spherical coordinates ξ = ρω, we write this term as
Z Z 2πiρω1 x1
e ψ(ρ)
dω dρ
1 + iε − ρ
for a suitable cutoff function ψ. We then use a smooth cutoff to split into
three regimes: (i) ω1 > 14 , (ii) |ω1 | < 12 , and (iii) ω1 < − 14 .
The contribution of region (ii) can be seen to be rapidly decaying, which
we see as follows: We first split into regions where |ρ − 1| > |x|−50 and
|ρ − 1| ≤ |x|−50 , say:
For the first term, we see that the ω integral is an oscillatory integral
with non-stationary phase, and hence is O(hxi−100 ). The ρ integral can then
be seen to be O(loghxi), and hence this part is acceptable. For the second
term, we write
where χ, ψ̃ are bump functions and χ = 1 near ρ = 1. Then the ψ̃(ρ) integral
leads to an oscillatory integral with non-stationary phase and hence leads
to rapid decay in x. The χ integral can be dealt with by extending χ into
the complex plane and using contour integration. In the end, by shifting the
contour upward, we end up with a main contribution of
−2πiψ(1)e2πiω1 x1
which (by the method of stationary phase, with the stationary phase point
ξ = e1 ) can be seen to be of the form
(∆ + 4π 2 )G1 = δ0
(H + 4π 2 k 2 )w = f,
(H + 4π 2 k 2 )w = 0
then w ≡ 0.
Brief sketch of proof. Using the radiation condition and Green’s theorem,
one can firstly obtain the identity
w = (−V w) ∗ Gk .
this leads to Z
|V
d w(kω)|2 dω = 0
(H + 4π 2 (k + iε)2 )wε = f,
where (as in the previous section) have that the operator on the left is
guaranteed to be invertible. We will show that wε is bounded in H 1 (BR (0))
for any R > 0. Then, using the compactness of the embedding H 1 (BR (0)) ,→
L2 (BR (0)), we may obtain a subsequential L2 limit of wε as ε → 0. (We can
then verify that the limit obeys w = (f − V w) ∗ Gk as well as the radiation
condition, and hence is the unique solution we are looking for.)
Lemma 11.1.6. The solutions wε are bounded in L2 as ε → 0 and hence
(by elliptic estimates) in H 1 (BR (0)) for any R > 0.
Proof. Suppose not. Then we may find a sequence of solutions wε with
kwε kL2 → ∞ and ε → 0. Normalizing the L2 -norm, we may obtain a
sequence w̃ε with kw̃ε kL2 ≡ 1 but
(H + 4π 2 (k + iε)2 )w̃ε → 0 in L2 .
Using elliptic estimates, we can deduce that w̃ε is compact in L2 (BR (0)) for
any R > 0 and hence we may extract an L2 limit w̃. But then we may derive
(H + 4π 2 k 2 )w̃ = 0 (and the radiation condition) and hence (by uniqueness)
w̃ ≡ 0. However, this contradicts the fact that kw̃ε kL2 ≡ 1.
(H + 4π 2 |ξ|2 )Φξ = 0.
and
∂r GH (x, rω, k) = 2πikCr−1 e2πikr Φkω (x) + O(r−3/2 ).
In fact, an argument using Green’s theorem yields
Z
Im GH (x, y, k) = Ck Im Φkω (x)Φkω (y) dω
S2
(see [26]).
Given the facts above, let us conclude this section by defining the dis-
torted Fourier transform adapted to H and showing how it provides us with
a ‘spectral representation’ of H.
Let us begin by verifying a claim made above, namely:
Lemma 11.1.7. For any λ < 0, the spectral projection χ(−∞,λ) (H) has
finite rank.
Proof. Writing X for the range of χ(−∞,λ) (H), we have
This implies that the unit ball in X with the L2 topology is a bounded
subset of H 1 , and hence compact. This in turn yields that X is finite-
dimensional.
In what follows (as we have done above), we therefore assume that H has
no negative eigenvalues (or that we restrict to the orthogonal complement
of the space of eigenfunctions of H).
To define functions of H, we will use the following Plemelj formula (which
follows from contour integration):
Z ∞
1
φ(x) = 2πi lim φ(λ)[(x + λ − iε)−1 − (x + λ + iε)−1 ] dλ for x > 0,
ε→0 0
284 CHAPTER 11. ADDITIONAL TOPICS
or equivalently
Z ∞
φ(x) = 1
lim
π ε→0 φ(λ) Im[(x + λ − iε)−1 ] dλ.
0
Prerequisite material
285
286 APPENDIX A. PREREQUISITE MATERIAL
where |S| denotes the Lebesgue measure of S. The quantity k · kLp (E) also
defines a norm. We will often drop the underlying set E and simply write
Lp .
To be precise, elements of Lp should be regarded as equivalence classes
of functions that are equal almost everywhere; however, we will typically
ignore this distinction.
The spaces Lp are vector spaces. Furthermore they are complete with
respect to the metric defined by the Lp -norm (namely, d(f, g) = kf − gkLp ).
In particular, they are Banach spaces. Furthermore, for 1 ≤ p < ∞, we
have that the space Lp is separable (i.e. admits a countable dense subset).
On the other hand, L∞ is not separable.
For spaces of sequences c = {ck } we use
X 1
p
p
kck`p = |ck | , kck`∞ = sup |ck |.
where ¯· denotes the complex conjugate. Finiteness of hf, gi follows from the
Cauchy–Schwarz inequality (or Hölder’s inequality):
which yields the desired bound. (This in turn may be used to show that
PN g is Cauchy in H as N → ∞ and hence converges to a limit g.)
As for uniqueness, we use the fact that if hf, gi = hf, hi for all f ∈ H
then g = h.
Using the Riesz representation theorem, we can identify H with its dual
space via the pairing hf, gi. We say that a sequence fn converges weakly
to f if
hfn , gi → hf, gi for all g ∈ H.
We write fn * f .
Lemma A.2.3. The following properties hold concerning weak convergence:
(a) If fn * f , then kf k ≤ lim supn→∞ kfn k.
hf, gi > kf k − ε.
to a limit (say ck ) for all k. We claim that fn converges weakly along this
subsequence.
To see this, we need to define the limit f . By duality, it will suffice to
specify the values hf, gi for all g ∈ H. To this end, we fix g ∈ H and take
a sequence of gk ∈ S converging to g. We will show that {ck } is Cauchy,
so that it has a limit c; we will then define hf, gi = c and check that this
defines an L2 function to which fn converge weakly. Let ε > 0 and choose
K large enough that kg − gk k < ε for k > K. Now fix k, ` > K. By choosing
n large enough, we may guarantee that
hfn , gk i − hfn , g` i − (ck − c` ) < ε.
where M is the uniform bound for the {fn }. Thus {ck } is Cauchy and hence
converges to c. An intertwining argument shows that we may uniquely define
f as a (linear) functional on H via hf, gi = c. To see that f ∈ H, we observe
that
|hf, gi| = | lim lim hfn , gk i| ≤ M kgk,
k→∞ n→∞
Remark A.2.4. Item (b) in the previous lemma is a special case of Alaoglu’s
theorem.
kT k = sup kT f k.
kf k≤1
T fnn − T fm
m
= (T − TK )fnn + TK (fnn − fm
m m
) + (TK − T )fm .
Choosing K large enough, the first and third terms may be made arbitrarily
small small due to the fact that {fnn }, {fm
m } are bounded sequences. For
as desired. In the final step we have used the dual formulation to compute
the norm of xn , i.e. kxn k = sup{|hxn , yi|} where the supremum is taken
over all y ∈ X.
Sketch of proof. We first show that A has an eigenvalue α0 with |α0 | = kAk.
We let α = kAk ≥ 0. Using symmetry, one observes
Let A1 = A|H (1) . Then A1 is symmetric and compact, and hence we can find
a largest eigenvalue α1 with normalized eigenvector u1 ... we now continue
in this way to construct a sequence of normalized eigenvectors {uj } that are
mutually orthogonal by construction with eigenvalues αj .
We claim that αj → 0. If not, then (passing to a subsequence) we get
αj 6= 0 for all j. Now consider the bounded sequence α1j uj . Then {A α1j uj }
has no convergent subsequence, since
In fact, in this case the trace tr(T ) is independent of the basis chosen. Trace
class operators are compact. We can prove this provided we take a few
Fourier analysis type results for granted (see Section 2.2 for details).
We first claim that if T is trace class, then in fact
X
kT ϕn k2 < ∞. (A.2)
n
We use this both when the functions are defined on all of Rd , as well as
when the functions are periodic on some torus. In this section we will focus
on the case of Rd .
Definition A.3.1. We call a family of functions {Kn } on Rd a family of
good kernels if the following three conditions hold:
R
(i) Rd Kn (y) dy = 1 for all n,
R
(ii) there exists M such that for all n we have Rd |Kn (y)| dy ≤ M ,
R
(iii) for each δ > 0, we have |y|>δ |Kn (y)| dy → 0 as n → ∞.
Example A.3.1. Let K : Rd → R satisfy
Z Z
|K(x)| dx . 1 and K(x) dx = 1.
Thus (writing M as the uniform upper bound for kKn kL1 and applying
Minkowski’s integral inequality)
Z Z p
K n (y)[f (x) − f (x − y)] dy dx
|y|<δ
Z Z p
p
. sup |f (x) − f (x − y)| dx · Kn (y) dy dx
|y|<δ
. εp M p
A.3. ANALYSIS TOOLS 297
then
sup kT kX→Y < ∞. (A.5)
T ∈F
Proof. The standard proof relies on the Baire category theorem. Here we
present a completely elementary proof appearing in [27]. Let us omit the
subscripts from the norms below; the meaning should be clear from context.
First, for any linear operator T and any x, y,
kT yk ≤ kT ( 12 (x + y) − 12 (x − y))k
≤ 21 kT (x + y)k + kT (x − y)k
≤ max kT (x ± y)k.
Now suppose (A.5) fails; we will show (A.4) fails as well. We choose
a sequence Tn ∈ F such that kTn k ≥ 4n . Define x0 = 0. Proceeding
inductively, we may find a sequence xn such that xn ∈ B(xn−1 , 3−n ) and
9 −n 9 4 n
kTn xn k ≥ 10 3 kTn k ≥ 10 ( 3 ) .
one finds kx − xn k ≤ 1
2 · 3−n . Now
so that
kTn xk ≥ 52 3−n kTn k ≥ 52 ( 43 )n → ∞,
showing the failure of (A.4).
Proof. Let us show the proof in the simplest case p = 2 and leave the
remaining cases as an exercise. Using Cauchy–Schwarz and exchanging the
order of summation,
2
XX
kT f k2`2
=
Tjk fk
j k
XX X
. |Tjk | |fk |2 |Tj` |
j k `
X XX
. sup |Tj` | · |Tjk ||fk |2
j j
` k
X X
2
.C· |fk | · sup |Tjk | . C 2 kf k2`2 ,
k k j
A.4 Exercises
Exercise A.4.1. Let 1 ≤ p1 ≤ p2 ≤ ∞. Show that
kak`p2 ≤ kak`p1
for all f, g.
Exercise A.4.7. Show that if fn * f weakly in a Hilbert space H and
kfn k → kf k, then fn → f strongly in H.
Bibliography
301
302 BIBLIOGRAPHY
[22] B. Osgood, Lecture Notes for EE251: The Fourier Transform and its
Applications.
[23] T. Tao, Lecture notes for Math 247A and Math 247B. Available at
BIBLIOGRAPHY 303
http://www.ucla.edu/~tao.
[24] T. Tao, Spherically averaged endpoint Strichartz estimates for the two-
dimensional Schrödinger equation, Comm. Partial Differential Equa-
tions 25 (2000), 1471–1485.