Lecture notes on introduction to Harmonic Analysis
Lecture notes on introduction to Harmonic Analysis
Chengchun Hao
2 Interpolation of Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
2.1 Riesz-Thorin’s and Stein’s interpolation theorems . . . . . . . . . . . . 33
2.2 The distribution function and weak Lp spaces . . . . . . . . . . . . . . . 41
2.3 The decreasing rearrangement and Lorentz spaces . . . . . . . . . . . . 45
2.4 Marcinkiewicz’ interpolation theorem . . . . . . . . . . . . . . . . . . . . . . 51
4 Singular Integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
4.1 Harmonic functions and Poisson equation . . . . . . . . . . . . . . . . . . . 77
4.2 Poisson kernel and Hilbert transform . . . . . . . . . . . . . . . . . . . . . . . 82
4.3 The Calderón-Zygmund theorem . . . . . . . . . . . . . . . . . . . . . . . . . . 94
4.4 Truncated integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
4.5 Singular integral operators commuted with dilations . . . . . . . . . . 101
4.6 The maximal singular integral operator . . . . . . . . . . . . . . . . . . . . . 107
4.7 *Vector-valued analogues . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
L ECTURE N OTES ON c 2016 by Chengchun Hao
Introduction to Harmonic Analysis Email: hcc@amss.ac.cn
Chapter 1
The Fourier Transform and Tempered
Distributions
In this chapter, we introduce the Fourier transform and study its more el-
ementary properties, and extend the definition to the space of tempered dis-
tributions. We also give some characterizations of operators commuting with
translations.
1
-2- 1. The Fourier Transform and Tempered Distributions
Z
h(x) = f (x − y)g(y)dy.
Rn
One can show by an elementary argument that f (x − y)g(y) is a measurable
function of the two variables x and y. It then follows immediately from Fib-
ini’s theorem on the interchange of the order of integration that h ∈ L1 (Rn )
and khk1 6 kf k1 kgk1 . Furthermore, this operation is commutative and as-
sociative. More generally, weRhave, with the help of Minkowski’s integral in-
equality k F (x, y)dykLpx 6 kF (x, y)kLpx dy, the following result:
R
1
Jean Baptiste Joseph Fourier (21 March 1768 – 16 May 1830) was a French mathematician and physi-
cist best known for initiating the investigation of Fourier series and their applications to problems of
heat transfer and vibrations. The Fourier transform and Fourier’s Law are also named in his honor.
Fourier is also generally credited with the discovery of the greenhouse effect.
1.1. The L1 theory of the Fourier transform -3-
Proof. These results are easy to be verified. We only prove (vii). In fact,
Z Z
−ωix·ξ −1
F (f ◦ A)(ξ) = e f (Ax)dx = e−ωiA y·ξ f (y)dy
n n
ZR ZR
>
= e−ωiA y·ξ f (y)dy = e−ωiy·Aξ f (y)dy = fˆ(Aξ),
Rn Rn
where we used the change of variables y = Ax and the fact that A−1 = A>
and | det A| = 1.
Corollary 1.4. The Fourier transform of a radial function is radial.
Proof. Let ξ, η ∈ Rn with |ξ| = |η|. Then there exists some orthogonal matrix
A such that Aξ = η. Since f is radial, we have f = f ◦ A. Then, it holds
F f (η) = F f (Aξ) = F (f ◦ A)(ξ) = F f (ξ),
by (vii) in Proposition 1.3.
It is easy to establish the following results:
Theorem 1.5 (Uniform continuity). (i) The mapping F is a bounded linear
transformation from L1 (Rn ) into L∞ (Rn ). In fact, kF f k∞ 6 kf k1 .
(ii) If f ∈ L1 (Rn ), then F f is uniformly continuous.
Proof. (i) is obvious. We now prove (ii). By
Z
ˆ ˆ
f (ξ + h) − f (ξ) = e−ωix·ξ [e−ωix·h − 1]f (x)dx,
Rn
we have Z
|fˆ(ξ + h) − fˆ(ξ)| 6 |e−ωix·h − 1||f (x)|dx
n
ZR Z
−ωix·h
6 |e − 1||f (x)|dx + 2 |f (x)|dx
|x|6r |x|>r
Z Z
6 |ω|r|h||f (x)|dx + 2 |f (x)|dx
|x|6r |x|>r
=:I1 + I2 ,
since for any θ > 0
iθ
q √
|e − 1| = (cos θ − 1)2 + sin2 θ = 2 − 2 cos θ = 2| sin(θ/2)| 6 |θ|.
Given any ε > 0, we can take r so large that I2 < ε/2. Then, we fix this r
and take |h| small enough such that I1 < ε/2. In other words, for given ε > 0,
there exists a sufficiently small δ > 0 such that |fˆ(ξ + h) − fˆ(ξ)| < ε when
|h| 6 δ, where ε is independent of ξ.
Ex. 1.6. Suppose that a signal consists of a single rectangular pulse of
width 1 and height 1. Let’s say that it gets turned on at x = − 21 and
turned off at x = 12 . The standard name for this “normalized” rectan-
gular pulse is
-4- 1. The Fourier Transform and Tempered Distributions
1
1, if − 21 < x < 12 ,
Π(x) ≡ rect(x) :=
0, otherwise. − 21 1 x
2
It is also called, variously, the normalized boxcar function, the top hat
function, the indicator function, or the characteristic function for the
interval (−1/2, 1/2). The Fourier transform of this signal is
Z 1/2 1/2
e−ωixξ 2 ωξ
Z
−ωixξ
Π(ξ)
b = e Π(x)dx = e−ωixξ dx = = sin
R −1/2 −ωiξ −1/2 ωξ 2
R 1/2
when ξ 6= 0. When ξ = 0, Π(0)b = −1/2 dx = 1. By l’Hôpital’s rule,
sin ωξ
2
ω
cos ωξ
lim Π(ξ)
b = lim 2 = lim 2 2 2
= 1 = Π(0),
b
ξ→0 ξ→0 ωξ ξ→0 ω
so Π(ξ)
b is continuous at ξ = 0. There is a standard function called
“sinc” that is defined by sinc(ξ) = sinξ ξ . In this notation Π(ξ)
2 b = sinc ωξ
2
.
Here is the graph of Π(ξ).
b
− 2π
ω
2π
ω
ξ
Remark 1.7. The above definition of the Fourier transform in (1.1) ex-
tends immediately to finite Borel measures: if µ is such a measure on
Rn , we define F µ by letting
Z
F µ(ξ) = e−ωix·ξ dµ(x).
Rn
Theorem 1.5 is valid for this Fourier transform if we replace the L1
norm by the total variation of µ.
The following theorem plays a central role in Fourier Analysis. It takes its
name from the fact that it holds even for functions that are integrable according
to the definition of Lebesgue. We prove it for functions that are absolutely
integrable in the Riemann sense.3
2
The term “sinc” (English pronunciation:["sINk]) is a contraction, first introduced by Phillip M.
Woodward in 1953, of the function’s full Latin name, the sinus cardinalis (cardinal sine).
3
Let us very briefly recall what this means. A bounded function f on a finite interval [a, b] is
integrable if it can be approximated by Riemann sums from above and below in such a way that the
difference of the integrals of these sums can be made as small as we wish. This definition is then
extended to unbounded functions and infinite intervals by taking limits; these cases are often called
improperR integrals. If I is any interval and f is a function on I such that the (possibly improper)
integral I |f (x)|dx has a finite value, then f is said to be absolutely integrable on I.
1.1. The L1 theory of the Fourier transform -5-
Proof. First, for n = 1, suppose that f (x) = χ(a,b) (x), the characteristic func-
tion of an interval. Then
Z b
ˆ e−ωiaξ − e−ωibξ
f (ξ) = e−ωixξ dx = → 0, as |ξ| → ∞.
a ωiξ
Similarly, the result holds when f is the characteristic function of the n-
dimensional rectangle I = {x ∈ Rn : a1 6 x1 6 b1 , · · · , an 6 xn 6 bn }
since we can calculate F f explicitly as an iterated integral. The same is there-
fore true for a finite linear combination of such characteristic functions (i.e.,
simple functions). Since all such simple functions are dense in L1 , the result
for a general f ∈ L1 (Rn ) follows easily by approximating f in the L1 norm by
such a simple function g, then f = g + (f − g), where F f − F g is uniformly
small by Theorem 1.5, while F g(ξ) → 0 as |ξ| → ∞.
Theorem 1.8 gives a necessary condition for a function to be a Fourier trans-
form. However, that belonging to C0 is not a sufficient condition for being the
Fourier transform of an integrable function. See the following example.
Ex. 1.9. Suppose, for simplicity, that n = 1. Let
1
, ξ > e,
ln ξ
g(ξ) =
ξ
,
0 6 ξ 6 e,
e
g(ξ) = − g(−ξ), ξ < 0.
It is clear that g(ξ) is uniformly continuous on R and g(ξ) → 0 as |ξ| →
∞.
Assume that there exists an f ∈ L1 (R) such that fˆ(ξ) = g(ξ), i.e.,
Z ∞
g(ξ) = e−ωixξ f (x)dx.
−∞
Since g(ξ) is an odd function, we have
Z ∞ Z ∞ Z ∞
−ωixξ
g(ξ) = e f (x)dx = −i sin(ωxξ)f (x)dx = sin(ωxξ)F (x)dx,
−∞ −∞ 0
N
sin t π
Z
lim dt = ,
N →∞ 0 t 2
and by Lebesgue dominated convergence theorem,we get that the inte-
gral of r.h.s. is convergent as N → ∞. That is,
π ∞
Z N
g(ξ)
Z
lim dξ = F (x)dx < ∞,
N →∞ 0 ξ 2 0
R∞ Re
which yields e g(ξ) ξ
dξ < ∞ since 0 g(ξ) ξ
dξ = 1. However,
Z N Z N
g(ξ) dξ
lim dξ = lim = ∞.
N →∞ e ξ N →∞ e ξ ln ξ
This contradiction indicates that the assumption was invalid.
We now turn to the problem of inverting the Fourier transform. That is,
we shall consider the question: Given the Fourier transform fˆ of an integrable
function f , how do we obtain f back again from fˆ ? The reader, who is familiar
with the elementary theory of Fourier series and integrals, would expect f (x)
to be equal to the integral
Z
C eωix·ξ fˆ(ξ)dξ. (1.4)
Rn
Unfortunately, fˆ need not be integrable (for example, let n = 1 and f be the
characteristic function of a finite interval). In order to get around this difficulty,
we shall use certain summability methods for integrals. We first introduce the
Abel method of summability, whose analog for series is very well-known. For
each ε > 0, we define the Abel mean Aε = Aε (f ) to be the integral
Z
Aε (f ) = Aε = e−ε|x| f (x)dx. (1.5)
Rn
The first of these two functions is called the Weierstrass (or Gauss-
Weierstrass) kernel while the second is called the Poisson kernel.
Theorem 1.12 (The multiplication formula). If f, g ∈ L1 (Rn ), then
Z Z
fˆ(ξ)g(ξ)dξ = f (x)ĝ(x)dx.
Rn Rn
Proof. From (iii) and (iv) in Proposition 1.3, it implies (F eωix·ξ Φ(εξ))(y) =
ϕε (y − x). The first result holds immediately with the help of Theorem 1.12.
The last two follow from (1.9), (1.10) and (1.12).
R
Lemma R 1.14. (i) Rn
W (x, ε)dx = 1 for all ε > 0.
(ii) Rn P (x, ε)dx = 1 for all ε > 0.
Letting r = |x|, x0 = x/r (when x 6= 0), S n−1 = {x ∈ Rn : |x| = 1}, dx0 the
element of surface area on S n−1 whose surface area4 is denoted by ωn−1 and,
finally, putting r = tan θ, we have
Z ∞Z
1 1
Z
2 (n+1)/2
dx = 2 (n+1)/2
dx0 rn−1 dr
Rn (1 + |x| ) 0 n−1 (1 + r )
ZS ∞
rn−1
=ωn−1 dr
0 (1 + r2 )(n+1)/2
Z π/2
=ωn−1 sinn−1 θdθ.
0
But ωn−1 sin θ is clearly the surface area of the
n−1 x
n
n+1
S
sphere of radius sin θ obtained by intersecting S n
sin θ
of the upper half of S n is obtained by summing O θ
cos θ
1
x 1
= [f (x − y) − f (x)]ϕε (y)dy,
Rn
the rest of the argument is precisely that used in the last proof.
In particular, we also have
Corollary 1.17. Suppose ϕ ∈ L1 (Rn ) with Rn ϕ(x)dx = 1 and let ϕε (x) =
R
ε−n ϕ(x/ε) for ε > 0. Let f (x) ∈ L∞ (Rn ) be continuous at {0}. Then,
Z
lim f (x)ϕε (x)dx = f (0).
ε→0 Rn
η
|f (x)| < ,
kϕk1
whenever |x| < δ. Noticing that | Rn ϕ(x)dx| 6 kϕk1 , we have
R
η
Z Z Z
f (x)ϕε (x)dx 6 |ϕε (x)|dx + kf k∞ |ϕε (x)|dx
Rn kϕk1 |x|<δ |x|>δ
η
Z
6 kϕk1 + kf k∞ |ϕ(y)|dy
kϕk1 |y|>δ/ε
=η + kf k∞ Iε .
But Iε → 0 as ε → 0. This proves the result.
From Theorems 1.13 and 1.15, we obtain the following solution to the
Fourier inversion problem:
RTheorem 1.18. If both Φ and its Fourier transform ϕ = Φ̂ are integrable and
Rn
ϕ(x)dx = 1, then the Φ means of the integral (|ω|/2π) Rn eωix·ξ fˆ(ξ)dξ
nR
converges to f (x) in the L1 norm. In particular, the Abel and Gauss means of
this integral converge to f (x) in the L1 norm.
We have singled out the Gauss-Weierstrass and the Abel methods of summa-
bility. The former is probably the simplest and is connected with the solution
of the heat equation; the latter is intimately connected with harmonic functions
and provides us with veryn powerful tools in Fourier analysis.
|ω| ωix·ξ −ε|ωξ|2 ˆ
Since s(x, ε) = 2π f (ξ)dξ converges in L1 to f (x) as
R
Rn
e e
ε > 0 tends to 0, we can find a sequence εk → 0 such that s(x, εk ) → f (x)
for a.e. x. If we further assume that fˆ ∈ L1 (Rn ), the Lebesgue dominated
convergence theorem gives us the following pointwise equality:
Theorem 1.19 (Fourier inversion theorem). If both f and fˆ are integrable,
then n Z
|ω|
f (x) = eωix·ξ fˆ(ξ)dξ,
2π Rn
ˆ 1 n
Moreover, we have f ∈ L (R ) and
n Z
|ω|
f (x) = eωix·ξ fˆ(ξ)dξ,
2π Rn
=I1 + I2 .
Since f is continuous at 0, for any given σ > 0, we can choose δ small enough
such that |f (y) − f (0)| 6 σ when |y| < δ. Thus, I1 6 σ by Lemma 1.14. For
the second term, we have, by a change of variables, that
Z
I2 6kf k1 sup P (y, ε) + |f (0)| P (y, ε)dy
|y|>δ |y|>δ
cn ε
Z
=kf k1 2 + |f (0)| P (y, 1)dy → 0,
(ε + δ 2 )(n+1)/2 |y|>δ/ε
n R
as ε → 0. Thus, |ω| 2π Rn
e−ε|ωξ| fˆ(ξ)dξ → f (0) as ε → 0. On the other
hand, by Lebesgue dominated convergence theorem, we obtain
n Z n
|ω| |ω|
Z
ˆ
f (ξ)dξ = lim e−ε|ωξ| fˆ(ξ)dξ = f (0),
2π R n 2π ε→0 Rn
ˆ ˆ
which implies f ∈ L (R ) due to f > 0. Therefore, from Theorem 1.19, it
1 n
The integral defining the Fourier transform is not defined in the Lebesgue
sense for the general function in L2 (Rn ); nevertheless, the Fourier transform
has a natural definition on this space and a particularly elegant theory.
If, in addition to being integrable, we assume f to be square-integrable then
fˆ will also be square-integrable. In fact, we have the following basic result:
Theorem 1.26 (Plancherel theorem). If f ∈ L1 (Rn )∩L2 (Rn ), then kfˆk2 =
−n/2
|ω|
2π
kf k2 .
But this implies that ĝ(x) = 0 for almost every x, contradicting the fact that
−n/2
kĝk2 = |ω|2π
kgk2 6= 0.
- 16 - 1. The Fourier Transform and Tempered Distributions
Theorem 1.27 is a major part of the basic theorem in the L2 theory of the
Fourier transform:
Theorem 1.28. The inverse of the Fourier transform, F −1 , can be obtained
by letting
n
−1 |ω|
(F f )(x) = (F f )(−x)
2π
for all f ∈ L2 (Rn ).
We can also extend the definition of the Fourier transform to other spaces,
such as Schwartz space, tempered distributions and so on.
Proof. Let {ϕk }∞ k=1 ⊂ S be a Cauchy sequence. For any σ > 0 and any
−|γ|
γ ∈ Nn0 , let ε = 21+2σσ , then there exists an N0 (ε) ∈ N such that ρ(ϕk , ϕm ) < ε
when k, m > N0 (ε) since {ϕk }∞ k=1 is a Cauchy sequence. Thus, we have
|ϕk − ϕm |0,γ σ
< ,
1 + |ϕk − ϕm |0,γ 1+σ
and then
sup |∂ γ (ϕk − ϕm )| < σ
x∈K
for any compact set K ⊂ R . It means that {ϕk }∞
n
k=1 is a Cauchy sequence in
the Banach space C (K). Hence, there exists a function ϕ ∈ C |γ| (K) such
|γ|
that
lim ϕk = ϕ, in C |γ| (K).
k→∞
Thus, we can conclude that ϕ ∈ C ∞ (Rn ). It only remains to prove that ϕ ∈ S .
It is clear that for any α, β ∈ Nn0
sup |xα ∂ β ϕ| 6 sup |xα ∂ β (ϕk − ϕ)| + sup |xα ∂ β ϕk |
x∈K x∈K x∈K
Moreover, some easily established properties of S and its topology, are the
following:
Proposition 1.40. i) The mapping ϕ(x) 7→ xα ∂ β ϕ(x) is continuous.
ii) If ϕ ∈ S , then limh→0 τh ϕ = ϕ.
iii) Suppose ϕ ∈ S and h = (0, · · · , hi , · · · , 0) lies on the i-th coordinate
axis of Rn , then the difference quotient [ϕ − τh ϕ]/hi tends to ∂ϕ/∂xi as |h| →
0.
iv) The Fourier transform is a homeomorphism of S onto itself.
v) S is separable.
Finally, we describe and prove a fundamental result of Fourier analysis that
is known as the uncertainty principle. In fact this theorem was "discovered" by
W. Heisenberg in the context of quantum mechanics. Expressed colloquially,
the uncertainty principle says that it is not possible to know both the position
and the momentum of a particle at the same time. Expressed more precisely,
the uncertainty principle says that the position and the momentum cannot be
simultaneously localized.
In the context of harmonic analysis, the uncertainty principle implies that
one cannot at the same time localize the value of a function and its Fourier
transform. The exact statement is as follows.
Theorem 1.41 (The Heisenberg uncertainty principle). Suppose ψ is a
function in S (R). Then
−1/2
|ω| kψk22
kxψk2 kξ ψ̂k2 > ,
2π 2|ω|
2
and equality holds if and only if ψ(x) = Ae−Bx where B > 0 and A ∈ R.
Moreover, we have
−1/2
|ω| kψk22
k(x − x0 )ψk2 k(ξ − ξ0 )ψ̂k2 >
2π 2|ω|
for every x0 , ξ0 ∈ R.
Proof. The last inequality actually follows from the first by replacing ψ(x) by
e−ωixξ0 ψ(x + x0 ) (whose Fourier transform is eωix0 (ξ+ξ0 ) ψ̂(ξ + ξ0 ) by parts (ii)
and (iii) in Proposition 1.3) and changing variables. To prove the first inequal-
ity, we argue as follows.
Since ψ ∈ S , we know that ψ and ψ 0 are rapidly decreasing. Thus, an
integration by parts gives
Z ∞ Z ∞
2 2 d
kψk2 = |ψ(x)| dx = − x |ψ(x)|2 dx
−∞ −∞ dx
Z ∞
=− xψ 0 (x)ψ(x) + xψ 0 (x)ψ(x) dx.
−∞
1.4. The class of tempered distributions - 21 -
Proof. See the previous definition, Theorem 1.49 and its corollary.
Remark 1.53. Since hF −1 F T, ϕi = hF T, F −1 ϕi = hT, F F −1 ϕi =
hT, ϕi, we get F −1 F = F F −1 = I in S 0 .
Ex. 1.54. Since for any ϕ ∈ S ,
Z
hF 1, ϕi =h1, F ϕi = (F ϕ)(ξ)dξ
Rn
−n n Z
|ω| |ω|
= eωi0·ξ (F ϕ)(ξ)dξ
2π 2π Rn
−n −n
|ω| −1 |ω|
= F F ϕ(0) = ϕ(0)
2π 2π
1.4. The class of tempered distributions - 25 -
−n
|ω|
= hδ, ϕi,
2π
we have
−n
|ω|
1̂ = δ, in S 0 .
2π
n
|ω|
Moreover, δ̌ = 2π
· 1.
Ex. 1.55. For ϕ ∈ S , we have
Z
hδ̂, ϕi = hδ, F ϕi = ϕ̂(0) = e−ωix·0 ϕ(x)dx = h1, ϕi.
Rn
0
Thus, δ̂ = 1 in S .
Ex. 1.56. Since
α δ, ϕi =h∂ α δ, ϕ̂i = (−1)|α| hδ, ∂ α ϕ̂i = hδ, F [(ωiξ)α ϕ]i
h∂d
=hδ̂, (ωiξ)α ϕi = h(ωiξ)α , ϕi,
α δ = (ωiξ)α .
we have ∂d
Now, we shall show that the convolution can be defined on the class S 0 .
We first recall a notation we have used: If g is any function on Rn , we define
its reflection, Rg, by letting Rg(x) = g(−x). A direct application of Fubini’s
theorem shows that if u, ϕ and ψ are all in S , then
Z Z
(u ∗ ϕ)(x)ψ(x)dx = u(x)(Rϕ ∗ ψ)(x)dx.
Rn Rn
τx+h Rϕ − τx Rϕ ∂Rϕ
→ −τx ,
hj ∂yj
as |h| → 0, in the topology of S . Thus, since u is continuous, we have
f (x + h) − f (x) τx+h Rϕ − τx Rϕ ∂Rϕ
= hu, i → hu, −τx i
hj hj ∂yj
as hj → 0. This, together with ii) in Proposition 1.40, shows that f has con-
tinuous first-order partial derivatives. Since ∂Rϕ/∂yj ∈ S , we can iterate
this argument and show that ∂ β f exists and is continuous for all multi-index
β ∈ Nn0 . We observe that ∂ β f (x) = hu, (−1)|β| τx ∂ β Rϕi. Consequently, since
∂ β Rϕ ∈ S , if f were slowly increasing, then the same would hold for all the
derivatives of f . In fact, that f is slowly increasing is an easy consequence of
Theorem 1.47: There exist C > 0 and integers ` and m such that
X
|f (x)| = |hu, τx Rϕi| 6 C |τx Rϕ|α,β .
|α|6`,|β|6m
But |τx Rϕ|α,β = supy∈Rn |y ∂ Rϕ(y −x)| = supy∈Rn |(y +x)α ∂ β Rϕ(y)| and
α β
=hu, τx Rϕ(·)ψ(x)dxi
n
Z R Z
= hu, τx Rϕiψ(x)dx = f (x)ψ(x)dx,
Rn Rn
since u is continuous and linear and the fact that the integral
τ Rϕ(y)ψ(x)dx converges in S , which is the desired equality.
R
Rn x
Having set down these facts of distribution theory, we shall now apply them
to the study of the basic class of linear operators that occur in Fourier analysis:
the class of operators that commute with translations.
Definition 1.58. A vector space X of measurable functions on Rn is
called closed under translations if for f ∈ X we have τy f ∈ X for all
y ∈ Rn . Let X and Y be vector spaces of measurable functions on Rn
that are closed under translations. Let also T be an operator from X to
Y . We say that T commutes with translations or is translation invariant if
T (τy f ) = τy (T f )
1.5. Characterization of operators commuting with translations - 27 -
Since (1 + |ξ| )
2 −(n+1)/2
defines anR integrable function on Rn , it follows that
fˆ ∈ L1 (Rn ) and, letting C 000 = C 00 Rn (1 + |ξ|2 )−(n+1)/2 dξ, we get
kfˆk1 6 C 000
X
k∂ α f k1 .
|α|6n+1
Thus, by Theorem 1.19, f equals almost everywhere a continuous function g
and by Theorem 1.5,
n
|ω|
kfˆk1 6 C
X
|g(0)| 6 kf k∞ 6 k∂ α f k1 .
2π
|α|6n+1
- 28 - 1. The Fourier Transform and Tempered Distributions
Suppose now that p > 1. Choose ϕ ∈ D(Rn ) such that ϕ(x) = 1 if |x| 6 1
and ϕ(x) = 0 if |x| > 2. Then, it is clear that f ϕ ∈ L1 (Rn ). Thus, f ϕ equals
almost everywhere a continuous function h such that
X
|h(0)| 6 C k∂ α (f ϕ)k1 .
|α|6n+1
where A > k∂ ϕk∞ , |ν| 6 |α|, and B depends only on p and n. Thus, we can
ν
For further results, one can see [SW71, p.30] and [Gra04, p.137-140].
L ECTURE N OTES ON c 2016 by Chengchun Hao
Introduction to Harmonic Analysis Email: hcc@amss.ac.cn
Chapter 2
Interpolation of Operators
We first present a notion that is central to complex analysis, that is, the
holomorphic or analytic function.
Let Ω be an open set in C and f a complex-valued function on Ω. The
function f is holomorphic at the point z0 ∈ Ω if the quotient
f (z0 + h) − f (z0 )
h
converges to a limit when h → 0. Here h ∈ C and h 6= 0 with z0 + h ∈ Ω,
so that the quotient is well defined. The limit of the quotient, when it exists, is
denoted by f 0 (z0 ), and is called the derivative of f at z0 :
f (z0 + h) − f (z0 )
f 0 (z0 ) = lim . (2.1)
h→0 h
It should be emphasized that in the above limit, h is a complex number that
may approach 0 from any directions.
The function f is said to be holomorphic on Ω if f is holomorphic at every
point of Ω. If C is a closed subset of C, we say that f is holomorphic on C if
f is holomorphic in some open set containing C. Finally, if f is holomorphic
in all of C we say that f is entire.
Every holomorphic function is analytic, in the sense that it has a power
series expansion near every point, and for this reason we also use the term
analytic as a synonym for holomorphic. For more details, one can see [SS03,
pp.8-10].
Ex. 2.1. The function f (z) = z is holomorphic on any open set in C, and
f 0 (z) = 1. The function f (z) = z̄ is not holomorphic. Indeed, we have
f (z0 + h) − f (z0 ) h̄
=
h h
which has no limit as h → 0, as one can see by first taking h real and
then h purely imaginary.
33
- 34 - 2. Interpolation of Operators
Ex. 2.2. The function 1/z is holomorphic on any open set in C that does
not contain the origin, and f 0 (z) = −1/z 2 .
One can prove easily the following properties of holomorphic functions.
Proposition 2.3. If f and g are holomorphic in Ω, then
i) f + g is holomorphic in Ω and (f + g)0 = f 0 + g 0 .
ii) f g is holomorphic in Ω and (f g)0 = f 0 g + f g 0 .
iii) If g(z0 ) 6= 0, then f /g is holomorphic at z0 and
0
f f 0g − f g0
= .
g g2
Moreover, if f : Ω → U and g : U → C are holomorphic, the chain rule
holds
(g ◦ f )0 (z) = g 0 (f (z))f 0 (z), for all z ∈ Ω.
The next result pertains to the size of a holomorphic function.
Theorem 2.4 (Maximum modulus principle). Suppose that Ω is a region
with compact closure Ω̄. If f is holomorphic on Ω and continuous on Ω̄, then
sup |f (z)| 6 sup |f (z)|.
z∈Ω z∈Ω̄\Ω
Proof. Denote
A0 := sup |f (it)|, A1 := sup |f (1 + it)|.
t∈R t∈R
Let λ ∈ R and define
Fλ (z) = eλz f (z).
Then by Theorem 2.5, it follows that
|Fλ (z)| 6 max(A0 , eλ A1 ).
Hence,
|f (θ + it)| 6 e−λθ max(A0 , eλ A1 )
for all t ∈ R. Choosing λ = ln A
A0
1
such that eλ A1 = A0 , we complete the
proof.
In order to state the Riesz-Thorin theorem in a general version, we will state
and prove it in measurable spaces instead of Rn only.
Let (X, µ) be a measure space, µ always being a positive measure. We adopt
the usual convention that two functions are considered equal if they agree ex-
cept on a set of µ-measure zero. Then we denote by Lp (X, dµ) (or simply
Lp (dµ), Lp (X) or even Lp ) the Lebesgue-space of (all equivalence classes of)
scalar-valued µ-measurable functions f on X, such that
- 36 - 2. Interpolation of Operators
Z 1/p
p
kf kp = |f (x)| dµ
X
is finite. Here we have 1 6 p < ∞. In the limiting case, p = ∞, Lp consists
of all µ-measurable and bounded functions. Then we write
kf k∞ = sup |f (x)|.
X
In this section, scalars are supposed to be complex numbers.
Let T be a linear mapping from Lp = Lp (X, dµ) to Lq (Y, dν). This means
that T (αf + βg) = αT (f ) + βT (g). We shall write
T : Lp → Lq
if in addition T is bounded, i.e., if
kT f kq
A = sup
f 6=0 kf kp
is finite. The number A is called the norm of the mapping T .
It will also be necessary to treat operators T defined on several Lp spaces
simultaneously.
Definition 2.7. We define Lp1 + Lp2 to be the space of all functions f ,
such that f = f1 + f2 , with f1 ∈ Lp1 and f2 ∈ Lp2 .
Suppose now p1 < p2 . Then we observe that
Lp ⊂ Lp1 + Lp2 , ∀p ∈ [p1 , p2 ].
In fact, let f ∈ L and let γ be a fixed positive constant. Set
p
f (x), |f (x)| > γ,
f1 (x) =
0, |f (x)| 6 γ,
and f2 (x) = f (x) − f1 (x). Then
Z Z Z
p1 p p1 −p p1 −p
|f1 (x)| dx = |f1 (x)| |f1 (x)| dx 6 γ |f (x)|p dx,
since p1 − p 6 0. Similarly,
Z Z Z
p2 p p2 −p p2 −p
|f2 (x)| dx = |f2 (x)| |f2 (x)| dx 6 γ |f (x)|p dx,
so f1 ∈ Lp1 and f2 ∈ Lp2 , with f = f1 + f2 .
Now, we have the following well-known theorem.
Theorem 2.8 (The Riesz-Thorin interpolation theorem). Let T be a linear
operator with domain (Lp0 + Lp1 )(X, dµ), p0 , p1 , q0 , q1 ∈ [1, ∞]. Assume that
kT f kLq0 (Y,dν) 6 A0 kf kLp0 (X,dµ) , if f ∈ Lp0 (X, dµ),
and
kT f kLq1 (Y,dν) 6 A1 kf kLp1 (X,dµ) , if f ∈ Lp1 (X, dµ),
for some p0 6= p1 and q0 6= q1 . Suppose that for a certain 0 < θ < 1
2.1. Riesz-Thorin’s and Stein’s interpolation theorems - 37 -
1 1−θ θ 1 1−θ θ
= + , = + . (2.3)
p p0 p1 q q0 q1
Then
kT f kLq (Y,dν) 6 Aθ kf kLp (X,dµ) , if f ∈ Lp (X, dµ),
with
Aθ 6 A01−θ Aθ1 . (2.4)
Remark 2.9. 1) (2.4) means that Aθ is logarithmi-
cally convex, i.e., ln Aθ is convex. 1
q (1, 1)
2) The geometrical meaning of (2.3) is that the
( p10 , q10 )
points (1/p, 1/q) are the points on the line seg-
ment between (1/p0 , 1/q0 ) and (1/p1 , 1/q1 ).
3) The original proof of this theorem, published ( p1 , 1q )
supp f
|f (x)| dµ(x) 6 M µ( supp f ) < ∞ for any ` > 0. So g does.
For 0 6 <z 6 1, we put
1 1−z z 1 1−z z
= + , = 0
+ ,
p(z) p0 p1 q 0 (z) q0 q10
and
p f (x)
η(z) =η(x, z) = |f (x)| p(z) , x ∈ X;
|f (x)|
q0 g(y)
ζ(z) =ζ(y, z) = |g(y)| q0 (z) , y ∈ Y.
|g(y)|
Now, we prove η(z), η 0 (z) ∈ Lpj for j = 0, 1. Indeed, we have
1−1/q0
1
Otherwise, it will be p0 = p1 = ∞ if p = ∞, or θ = 1/q1 −1/q0
> 1 if q 0 = ∞.
- 38 - 2. Interpolation of Operators
p
p( 1−z + pz ) p( 1−<z + <z )+ip( =z − =z )
|η(z)| = |f (x)| p(z) = |f (x)| p
0 1 = |f (x)| p 0 p1 p1 p0
p( 1−<z + <z ) p
=|f (x)| p p 0 1 = |f (x)| p(<z) .
Thus, Z Z ppj
kη(z)kppjj = |η(x, z)| dµ(x) =pj
|f (x)| p(<z) dµ(x) < ∞.
X X
We have 0
0 p f (x) p
η (z) =|f (x)| ln |f (x)|
p(z)
p(z) |f (x)|
1 1 p f (x)
=p − |f (x)| p(z) ln |f (x)|.
p1 p0 |f (x)|
On one hand, we have lim|f (x)|→0+ |f (x)|α ln |f (x)| = 0 for any α > 0, that
is, ∀ε > 0, ∃δ > 0 s.t. ||f (x)|α ln |f (x)|| < ε if |f (x)| < δ. On the other
hand, if |f (x)| > δ, then we have ||f (x)|α ln |f (x)|| 6 M α |ln |f (x)|| 6
M α max(| ln M |, | ln δ|) < ∞. Thus, ||f (x)|α ln |f (x)|| 6 C. Hence,
1 1 p
|η 0 (z)| =p − |f (x)| p(z) −α |f (x)|α |ln |f (x)||
p1 p0
p p
6C |f (x)| p(z) −α = C|f (x)| p(<z) −α ,
which yields
Z
p
kη 0
(z)kppjj 6C |f (x)|( p(<z) −α)pj dµ(x) < ∞.
X
0
Therefore, η(z), η (z) ∈ L for j = 0, 1. So ζ(z), ζ 0 (z) ∈ Lqj for j = 0, 1
0 pj
in the same way. By the linearity of T , it holds (T η)0 (z) = T η 0 (z) in view of
(2.1). It follows that T η(z) ∈ Lqj , and (T η)0 (z) ∈ Lqj with 0 < <z < 1, for
j = 0, 1. This implies the existence of
F (z) = hT η(z), ζ(z)i, 0 6 <z 6 1.
Since
dF (z) d d
Z
= hT η(z), ζ(z)i = (T η)(y, z)ζ(y, z)dν(y)
dz dz dz Y
Z Z
= (T η)z (y, z)ζ(y, z)dν(y) + (T η)(y, z)ζz (y, z)dν(y)
Y Y
0 0
=h(T η) (z), ζ(z)i + hT η(z), ζ (z)i,
F (z) is analytic on the open strip 0 < <z < 1. Moreover it is easy to see that
F (z) is bounded and continuous on the closed strip 0 6 <z 6 1.
Next, we note that for j = 0, 1
p
pj
kη(j + it)kpj = kf kp = 1.
Similarly, we also have kζ(j + it)kqj0 = 1 for j = 0, 1. Thus, for j = 0, 1
|F (j + it)| =|hT η(j + it), ζ(j + it)i| 6 kT η(j + it)kqj kζ(j + it)kqj0
2.1. Riesz-Thorin’s and Stein’s interpolation theorems - 39 -
ζ(θ) = g, thus F (θ) = hT f, gi. That is, |hT f, gi| 6 A01−θ Aθ1 . Therefore,
Aθ 6 A01−θ Aθ1 .
Now, we shall give two rather simple applications of the Riesz-Thorin inter-
polation theorem.
Theorem 2.10 (Hausdorff-Young inequality). Let 1 6 p 6 2 and 1/p +
1/p0 = 1. Then the Fourier transform defined as in (1.1) satisfies
−n/p0
|ω|
kF f kp0 6 kf kp .
2π
Proof. It follows by interpolation between the L1 -L∞ result kF f k∞ 6 kf k1
−n/2
(cf. Theorem 1.5) and Plancherel’s theorem kF f k2 = |ω| 2π
kf k2 (cf.
Theorem 1.26).
Proof. We fix f ∈ Lp , p ∈ [1, ∞] and then will apply the Riesz-Thorin in-
terpolation theorem to the mapping g 7→ f ∗ g. Our endpoints are Hölder’s
inequality which gives
|f ∗ g(x)| 6 kf kp kgkp0
0
and thus g 7→ f ∗ g maps Lp (Rn ) to L∞ (Rn ) and the simpler version of
Young’s inequality (proved by Minkowski’s inequality) which tells us that if
g ∈ L1 , then
kf ∗ gkp 6 kf kp kgk1 .
Thus g 7→ f ∗ g also maps L1 to Lp . Thus, this map also takes Lq to Lr where
1 1−θ θ 1 1−θ θ
= + 0 , and = + .
q 1 p r p ∞
Eliminating θ, we have r = p + q − 1.
1 1 1
The condition q > 1 is equivalent with θ > 0 and r > 1 is equivalent with
the condition θ 6 1. Thus, we obtain the stated inequality for precisely the
exponents p, q and r in the hypothesis.
- 40 - 2. Interpolation of Operators
Remark 2.12. The sharp form of Young’s inequality for convolutions can
be found in [Bec75, Theorem 3], we just state it as follows. Under the
assumption of Theorem 2.11, we have
kf ∗ gkr 6 (Ap Aq Ar0 )n kf kp kgkq ,
0
where Am = (m1/m /m01/m )1/2 for m ∈ (1, ∞), A1 = A∞ = 1 and primes
always denote dual exponents, 1/m + 1/m0 = 1.
The Riesz-Thorin interpolation theorem can be extended to the case where
the interpolated operators allowed to vary. In particular, if a family of operators
depends analytically on a parameter z, then the proof of this theorem can be
adapted to work in this setting.
We now describe the setup for this theorem. Suppose that for every z in the
closed strip S there is an associated linear operator Tz defined on the space of
simple functions on X and taking values in the space of measurable functions
on Y such that Z
|Tz (f )g|dν < ∞ (2.5)
Y
whenever f and g are simple functions on X and Y , respectively. The family
{Tz }z is said to be analytic if the function
Z
z→ Tz (f )gdν (2.6)
Y
is analytic in the open strip S ◦ and continuous on its closure S. Finally, the
analytic family is of admissible growth if there is a constant 0 < a < π and a
constant Cf,g such that
Z
−a|=z|
e ln Tz (f )gdν 6 Cf,g < ∞ (2.7)
Y
for all z ∈ S. The extension of the Riesz-Thorin interpolation theorem is now
stated.
Theorem 2.13 (Stein interpolation theorem). Let Tz be an analytic fam-
ily of linear operators of admissible growth. Let 1 6 p0 , p1 , q0 , q1 6 ∞ and
suppose that M0 and M1 are real-valued functions such that
sup e−b|t| ln Mj (t) < ∞ (2.8)
t∈R
for j = 0, 1 and some 0 < b < π. Let 0 < θ < 1 satisfy
1 1−θ θ 1 1−θ θ
= + , and = + . (2.9)
p p0 p1 q q0 q1
Suppose that
kTit (f )kq0 6 M0 (t)kf kp0 , kT1+it (f )kq1 6 M1 (t)kf kp1 (2.10)
for all simple functions f on X. Then
kTθ (f )kq 6 M (θ)kf kp , when 0 < θ < 1 (2.11)
2.2. The distribution function and weak Lp spaces - 41 -
m
{x:|f (x)|>α}
|f (x)|p dx. Thus, ({x : |f (x)| > α}) → 0 as α → +∞ and
R
Rn Z
lim |f (x)|p dx = 0.
α→+∞ {x:|f (x)|>α}
Hence, α f∗ (α) → 0 as α → +∞ since αp f∗ (α) > 0.
p
Proof. In order to prove i), we first prove the following conclusion: If f (x) is
finite and f∗ (α) < ∞ for any α > 0, then
Z Z ∞
p
|f (x)| dx = − αp df∗ (α). (2.12)
Rn 0
Indeed, the r.h.s. of the equality is well-defined from the conditions. For the
integral in the l.h.s., we can split it into Lebesgue integral summation. Let
0 < ε < 2ε < · · · < kε < · · · and
Ej = {x ∈ Rn : (j − 1)ε < |f (x)| 6 jε} , j = 1, 2, · · · ,
m
then, (Ej ) = f∗ ((j − 1)ε) − f∗ (jε), and
∞ ∞
m
Z X X
p p
|f (x)| dx = lim (jε) (Ej ) = − lim (jε)p [f∗ (jε) − f∗ ((j − 1)ε)]
Rn ε→0 ε→0
j=1 j=1
Z ∞
=− αp df∗ (α).
0
Now we return to prove i). If the values of both sides are infinite, then it is
clearly true. If one of the integral is finite, then it is clear that f∗ (α) < +∞
and f (x) is finite almost everywhere.
R ∞ p−1 Thus (2.12) is valid.
If either f ∈ L (R ) or 0 α f∗ (α)dα < ∞ for 1 6 p < ∞ , then we
p n
Thus, i) is true.
For ii), we have
inf {α : f∗ (α) = 0} = inf {α : m({x : |f (x)| > α}) = 0}
= inf {α : |f (x)| 6 α, a.e.}
=ess supx∈Rn |f (x)| = kf kL∞ .
We complete the proofs.
Using the distribution function f∗ , we now introduce the weak Lp -spaces
denoted by Lp∗ .
Definition 2.17. The space Lp∗ , 1 6 p < ∞, consists of all f such that
kf kLp∗ = sup αf∗1/p (α) < ∞.
α
In the limiting case p = ∞, we put L∞ ∞
∗ = L .
By the part (iv) in Proposition 2.15 and the triangle inequality of Lp norms,
we have
kf + gkLp∗ 6 2(kf kLp∗ + kgkLp∗ ).
Thus, one can verify that Lp∗ is a quasi-normed vector space. The weak Lp
spaces are larger than the usual Lp spaces. We have the following:
Theorem 2.18. For any 1 6 p < ∞, and any f ∈ Lp , we have kf kLp∗ 6 kf kp ,
hence Lp ⊂ Lp∗ .
α
m
= sup α( ({x : |x| < α−p/n }))1/p = sup α(α−p Vn )1/p
α
=Vn1/p ,
where Vn = π R/Γ (1 + n/2) is the volume of the unit ball in Rn and Γ -
n/2
∞
function Γ (z) = 0 tz−1 e−t dt for <z > 0.
It is not immediate from their definition that the weak Lp spaces are com-
plete with respect to the quasi-norm k·kLp∗ . For the completeness, we will state
it later as a special case of Lorentz spaces.
2.3. The decreasing rearrangement and Lorentz spaces - 45 -
The spaces Lp∗ are special cases of the more general Lorentz spaces Lp,q . In
their definition, we use yet another concept, i.e., the decreasing rearrangement
of functions.
Definition 2.19. If f is a measurable function on Rn , the decreasing rear-
rangement of f is the function f ∗ : [0, ∞) 7→ [0, ∞] defined by
f ∗ (t) = inf {α > 0 : f∗ (α) 6 t} ,
where we use the convention that inf ∅ = ∞.
Now, we first give some examples of distribution function and decreasing
rearrangement. The first example establish some important relations between
a simple function, its distribution function and decreasing rearrangement.
Ex. 2.20 (Decreasing rearrangement of a simple function). Let f be a
simple function of the following form
k
X
f (x) = aj χAj (x)
j=1
where a1 > a2 > · · · > ak > 0, Aj = {x ∈ R : f (x) = aj } and χA is the
characteristic function of the set A (see Figure (a)). Then
k k
m m
X X
f∗ (α) = ({x : |f (x)| > α}) = ({x : aj χAj (x) > α}) = bj χBj (α),
j=1 j=1
Pj
m
where bj = i=1 (Ai ), Bj = [aj+1 , aj ) for j = 1, 2, · · · , k and ak+1 = 0
which shows that the distribution function of a simple function is a
simple function (see Figure (b)). We can also find the decreasing rear-
rangement (by denoting b0 = 0)
k
X
f ∗ (t) = inf{α > 0 : f∗ (α) 6 t} = inf{α > 0 : bj χBj (α) 6 t}
j=1
k
X
= aj χ[bj−1 ,bj ) (t)
j=1
which is also a simple function (see Figure (c)).
- 46 - 2. Interpolation of Operators
a1 a1
a2 a2
a3 b5 a3
b4
a4 a4
b3
a5 b2 a5
b1
A3 A4 A1 A5 A2 x a5 a4 a3 a2 a1 α b 1 b2 b3 b4 b5 t
(a) (b) (c)
f f∗ f∗
2 2 2
1 1 1
1 2 x 1 2 α 1 2 t
(a) (b) (c)
Observe that the integral over f , f∗ and f ∗ are all the same, i.e.,
Z ∞
√
Z 2 Z 1 Z 2
2
f (x)dx = [1 − (x − 1) ]dx = 2 1 − αdα = (1 − t2 /4)dt = 4/3.
0 0 0 0
2.3. The decreasing rearrangement and Lorentz spaces - 47 -
f∗ f∗
f 5 5
5
4 4 4
3 3 3
2 2 2
1 1 1
1 2 3 x 1 2 3 α 1 2 3 t
(a) (b) (c)
Ex. 2.23. Consider the function f (x) = x for all x ∈ [0, ∞). Then f∗ (α) =
m ({x ∈ [0, ∞) : x > α}) = ∞ for all α > 0, which implies that f ∗ (t) =
inf{α > 0 : ∞ 6 t} = ∞ for all t > 0.
x
Ex. 2.24. Consider f (x) = 1+x for x > 0.
It is clear that f∗ (α) = 0 for α > 1 since
|f (x)| < 1. For α ∈ [0, 1), we have
m
f∗ (α) = ({x ∈ [0, ∞) :
x
1+x
> α})
f∗
m
= ({x ∈ [0, ∞) : x >
1−α
α
}) = ∞.
1
f
That is,
∞, 0 6 α < 1, 1 2
f∗ (α) =
0, α > 1.
Thus, f ∗ (t) = inf{α > 0 : f∗ (α) 6 t} = 1.
- 48 - 2. Interpolation of Operators
Proof. (v) Assume that f ∗ (t1 ) + g ∗ (t2 ) < ∞, otherwise, there is nothing to
prove. Then for α1 = f ∗ (t1 ) and α2 = g ∗ (t2 ), by (x), we have f∗ (α1 ) 6 t1 and
g∗ (α2 ) 6 t2 . From (iv) in Proposition 2.15, it holds
(f + g)∗ (α1 + α2 ) 6 f∗ (α1 ) + g∗ (α2 ) 6 t1 + t2 .
Using the definition of the decreasing rearrangement, we have
(f +g)∗ (t1 +t2 ) = inf{α : (f +g)∗ (α) 6 t1 +t2 } 6 α1 +α2 = f ∗ (t1 )+g ∗ (t2 ).
(vi) Similar to (v), by (v) in Proposition 2.15, it holds that (f g)∗ (α1 α2 ) 6
f∗ (α1 ) + g∗ (α2 ) 6 t1 + t2 . Then, we have
(f g)∗ (t1 + t2 ) = inf{α : (f g)∗ (α) 6 t1 + t2 } 6 α1 α2 = f ∗ (t1 )g ∗ (t2 ).
(xi) If f∗ (α) > t, then by the decreasing of f∗ , we have α < inf{β : f∗ (β) 6
t} = f ∗ (t). Conversely, if f ∗ (t) > α, i.e., inf{β : f∗ (β) 6 t} > α, we get
f∗ (α) > t by the decreasing of f∗ again.
(xii) By the definition and (xi), we have
(f ∗ )∗ (α) = m({t > 0 : f (t) > α}) = m({t > 0 : f (α) > t}) = f (α).
∗
∗ ∗
where σ = α1/p .
(xiv) From Theorem 2.16, we have
2.3. The decreasing rearrangement and Lorentz spaces - 49 -
Z ∞ Z ∞
∗ ∗
kf (t)kpp = p
|f (t)| dt = p αp−1 (f ∗ )∗ (α)dα
Z0 ∞
0
=Akf kp .
2.4. Marcinkiewicz’ interpolation theorem - 55 -
For the case p1 = ∞ the proof is the same except for the use of the estimate
kft k∞ 6 f ∗ (tσ ), we can get
1−1/p
0
1/p−1/q A0 p
1
1/q p −1/p 0
A=2 K |σ| 1 + A1 .
q
p
− p1
0
Thus, we complete the proof.
From the proof given above it is easy to see that the theorem can be extended
to the following situation: The underlying measure space Rn of the Lpi (Rn )
can be replaced by a general measurable space (and the measurable space oc-
curring in the domain of T need not be the same as the one entering in the
range of T ). A less superficial generalization of the theorem can be given in
terms of the notation of Lorentz spaces, which unify and generalize the usual
Lp spaces and the weak-type spaces. For a discussion of this more general
form of the Marcinkiewicz interpolation theorem see [SW71, Chapter V] and
[BL76, Chapter 5].
As an application of this powerful tool, we present a generalization of the
Hausdorff-Young inequality due to Paley. The main difference between the
theorems being that Paley introduced a weight function into his inequality and
resorted to the theorem of Marcinkiewicz. In what follows, we consider the
measure space (Rn , µ) where µ denotes the Lebesgue measure. Let w be a
weihgt function on Rn , i.e., a positive and measurable function on Rn . Then
we denote by Lp (w) the Lp -space with respect to wdx. The norm on Lp (w) is
Z 1/p
p
kf kLp (w) = |f (x)| w(x)dx .
Rn
With this notation we have the following theorem.
Theorem 2.35 (Hardy-Littlewood-Paley theorem on Rn ). Assume that
1 6 p 6 2. Then
kF f kLp (|ξ|−n(2−p) ) 6 Cp kf kp .
which implies that T is of weak type (2, 2). We now work towards showing
that T is of weak type (1, 1). Thus, the Marcinkiewicz interpolation theorem
implies the theorem.
Now, consider the set Eα = {ξ : |ξ|n fˆ(ξ) > α}. For simplicity, we let ν
denote the measure |ξ|−2n dξ and assume that kf k1 = 1. Then, |fˆ(ξ)| 6 1. For
ξ ∈ Eα , we therefore have α 6 |ξ|n . Consequently,
- 56 - 2. Interpolation of Operators
Z Z
−2n
(T f )∗ (α) = ν(Eα ) = |ξ| dξ 6 |ξ|−2n dξ 6 Cα−1 .
Eα |ξ|n >α
Thus, we proves that
α · (T f )∗ (α) 6 Ckf k1 ,
which implies T is of weak type (1, 1). Therefore, we complete the proof.
L ECTURE N OTES ON c 2016 by Chengchun Hao
Introduction to Harmonic Analysis Email: hcc@amss.ac.cn
Chapter 3
The Maximal Function and Calderón-Zygmund
Decomposition
Proof. The argument we give is constructive and relies on the following simple
observation: Suppose B and B 0 are a pair of balls that intersect, with the
radius of B 0 being not greater than that of B. Then B 0 is contained in the ball
B̃ that is concentric with B but with 3 times its radius. (See Fig 3.1.)
As a first step, we pick a ball Bj1 in B with maximal
(i.e., largest) radius, and then delete from B the ball Bj1 as
B0
well as any balls that intersect Bj1 . Thus all the balls that
are deleted are contained in the ball B̃j1 concentric with B
Bj1 , but with 3 times its radius.
The remaining balls yield a new collection B 0 , for which B̃
57
- 58 - 3. The Maximal Function and Calderón-Zygmund Decomposition
In the last step, we have used the fact that in Rn a dilation of a set by δ > 0
results in the multiplication by δ n of the Lebesgue measure of this set.
For the infinite version of Vitali covering lemma, one can see the textbook
[Ste70, the lemma on p.9].
The decomposition of a given set into a disjoint union of cubes (or balls) is
a fundamental tool in the theory described in this chapter. By cubes we mean
closed cubes; by disjoint we mean that their interiors are disjoint. We have in
mind the idea first introduced by Whitney and formulated as follows.
Theorem 3.2 (Whitney covering lemma). Let F be a non-empty closed
set in Rn and Ω be its complement. Then there exists a collection of cubes
F = {Qk } whose sides are parallel to the axes, such that
(i) ∞ c
S
k=1 Qk = Ω = F ,
(ii) Q◦j ∩ Q◦k = ∅ if j 6= k, where Q◦ denotes the interior of Q,
(iii) there exist two constants c1 , c2 > 0 independent of F (In fact we may
take c1 = 1 and c2 = 4.), such that
c1 diam (Qk ) 6 dist (Qk , F ) 6 c2 diam (Qk ).
√ O 1 2 3
ameter n2−k .
F
In addition to the meshes Mk , we Ωk+1
S∞ √ √
Ω = k=−∞ Ωk .
−k −k
c2 − n2 . If we choose c = 2 n we get (4.1).
Then by (4.1) the cubes Q ∈ F are disjoint from F and clearly cover Ω. Therefore, (i) is also
0
Now we make an initial choice of cubes, and denote the resulting collection
proved.
Notice that the collection F has all our required properties, except
0 that the cubes in it are not
necessarily disjoint. To finish the proof of the theorem, we need to refine our choice leading to F ,
by F0 . Our choice is made as follows. Wethose consider the
reallycubes
unnecessary. of the mesh Mk ,
0
eliminating cubes which were
We require the following simple observation. Suppose Q and Q are two cubes (taken respectively 1 2
0
a unique maximal cube in F0 which contains it. By the same taken these maximal cubes are also
disjoint. We let F denote the collection of maximal cubes of F0 . Then obviously
S
(i) Q∈F Q = Ω,
(ii) The cubes of F are disjoint,
(iii) diam (Q) 6 dist (Q, F) 6 4 diam (Q), Q ∈ F .
Therefore, we complete the proof. t
u
3.2. Hardy-Littlewood maximal function - 59 -
[
F0 = {Q ∈ Mk : Q ∩ Ωk 6= ∅} .
k
We then have
[
Q = Ω.
Q∈F 0
For appropriate choice of c, we claim that
diam (Q) 6 dist (Q, F ) 6 4 diam (Q), Q ∈ F0 . (3.1)
√ −k
Let us prove (3.1) first. Suppose Q ∈ Mk ; then diam (Q) = n2 . Since
Q ∈ F0 , there exists an x ∈ Q ∩ Ωk . Thus dist (Q, F ) 6 dist (x, F ) 6
√
c2−k+1 , and dist (Q, F ) > dist (x, F ) − diam (Q) > c2−k − n2−k . If we
√
choose c = 2 n we get (3.1).
Then by (3.1) the cubes Q ∈ F0 are disjoint from F and clearly cover Ω.
Therefore, (i) is also proved.
Notice that the collection F0 has all our required properties, except that the
cubes in it are not necessarily disjoint. To finish the proof of the theorem, we
need to refine our choice leading to F0 , eliminating those cubes which were
really unnecessary.
We require the following simple observation. Suppose Q1 and Q2 are two
cubes (taken respectively from the mesh Mk1 and Mk2 ). Then if Q1 and Q2
are not disjoint, one of the two must be contained in the other. (In particular,
Q1 ⊂ Q2 , if k1 > k2 .)
Start now with any cube Q ∈ F0 , and consider the maximal cube in F0
which contains it. In view of the inequality (3.1), for any cube Q0 ∈ F0 which
contains Q ∈ F0 , we have diam (Q0 ) 6 dist (Q0 , F ) 6 dist (Q, F ) 6
4 diam (Q). Moreover, any two cubes Q0 and Q00 which contain Q have ob-
viously a non-trivial intersection. Thus by the observation made above each
cube Q ∈ F0 has a unique maximal cube in F0 which contains it. By the
same taken these maximal cubes are also disjoint. We let F denote the collec-
tion of maximal cubes of F0 . Then obviously
(i) Q∈F Q = Ω,
S
holds for suitable points x. In higher dimensions we can pose a similar ques-
tion, where the averages of f are taken over appropriate sets that generalize
the intervals in one dimension.
In particular, we can take the sets involved as the ball B(x, r) of radius r,
m
centered at x, and denote its measure by (B(x, r)). It follows
1
Z
lim
r→0 m (B(x, r)) B(x,r)
f (y)dy = f (x), for a.e. x? (3.2)
as desired.
In general, for this “averaging problem” (3.2), we shall have an affirmative
answer. In order to study the limit (3.2), we consider its quantitative analogue,
3.2. Hardy-Littlewood maximal function - 61 -
1
Z
6 sup
r>0 m
(B(x, r)) B(x,r)
|f (y)|dy = M f (x).
1
A measurable function f on Rn is called to be locally integrable, if for every ball B the function
f (x)χB (x) is integrable. We shall denote by L1loc (Rn ) the space of all locally integrable functions.
Loosely speaking, the behavior at infinity does not affact the local integrability of a function. For
example, the functions e|x| and |x|−1/2 are both locally integrable, but not integrable on Rn .
2
The Hardy-Littlewood maximal operator appears in many places but some of its most notable uses
are in the proofs of the Lebesgue differentiation theorem and Fatou’s theorem and in the theory of
singular integral operators.
- 62 - 3. The Maximal Function and Calderón-Zygmund Decomposition
m
00
M f (x) = sup |f (y)|dy, (3.6)
Q3x (Q) Q
where the supremum is taken over all cubes containing x. Again, M 00 is point-
wise equivalent to M . One sometimes distinguishes between M 0 and M 00 by
referring to the former as the centered and the latter as the non-centered max-
imal operator. Alternatively, we could define the non-centered maximal func-
tion with balls instead of cubes:
1
Z
M̃ f (x) = sup
B3x m
(B) B
|f (y)|dy
at each x ∈ Rn . Here, the supremum is taken over balls B in Rn which contain
m
the point x and (B) denotes the measure of B (in this case a multiple of the
radius of the ball raised to the power n).
Ex. 3.6. Let f : R → R, f (x) = χ(0,1) (x). Then
1
2x , x > 1,
0
M f (x) =M f (x) = 1, 0 6 x 6 1,
1
2(1−x)
, x < 0,
1
x , x > 1,
00
M̃ f (x) =M f (x) = 1, 0 6 x 6 1,
1
1−x
, x < 0.
In fact, for x > 1, we get
Z x+h
0 1
M f (x) = M f (x) = sup χ(0,1) (y)dy
h>0 2h x−h
1−x+h
1 1
= max sup , sup = ,
x−h>0 2h x−h60 2h 2x
Z x+h2
00 1
M̃ f (x) = M f (x) = sup χ(0,1) (y)dy
h1 ,h2 >0 h1 + h2 x−h1
1 − x + h1
1 1
= max sup , sup = .
0<x−h1 <1 h1 x−h1 60 h1 x
For 0 6 x 6 1, it follows
Z x+h
0 1
M f (x) = M f (x) = sup χ(0,1) (y)dy
h>0 2h x−h
1−x+h
2h
= max sup , sup ,
0<x−h<x+h<1 2h 0<x−h<16x+h 2h
x+h 1
sup , sup
x−h60<x+h<1 2h x−h60<16x+h 2h
3.2. Hardy-Littlewood maximal function - 63 -
1 1 1
= max 1, 1, 1, min , = 1,
2 x 1−x
Z x+h2
00 1
M̃ f (x) = M f (x) = sup χ(0,1) (y)dy
h1 ,h2 >0 h1 + h2 x−h1
h1 + h2 x + h2
= max sup , sup ,
0<x−h1 <x+h2 <1 h1 + h2 x−h1 <0<x+h2 <1 h1 + h2
1 − x + h1
1
sup , sup
0<x−h1 <1<x+h2 h1 + h2 x−h1 <0<1<x+h2 h1 + h2
=1.
For x < 0, we have
0 x+h 1 1
M f (x) = M f (x) = max sup , sup = ,
0<x+h<1,h>0 2h x+h>1 2h 2(1 − x)
00 x + h2 1
M̃ f (x) = M f (x) = max sup , sup
h1 ,h2 >0,0<x+h2 <1 h1 + h2 h1 >0,x+h2 >1 h1 + h2
1
= .
1−x
Observe that f ∈ L1 (R), but M f, M 0 f, M 00 f, M̃ f ∈
/ L1 (R).
Remark 3.7. (i) M f is defined at every point x ∈ Rn and if f = g a.e.,
then M f (x) = M g(x) at every x ∈ Rn .
(ii) It may be well that M f = ∞ for every x ∈ Rn . For example, let
n = 1 and f (x) = x2 .
(iii) There are several definitions in the literature which are often
equivalent.
Next, we state some immediate properties of the maximal function. The
proofs are left to interested readers.
Proposition 3.8. Let f, g ∈ L1loc (Rn ). Then
(i) Positivity: M f (x) > 0 for all x ∈ Rn .
(ii) Sub-linearity: M (f + g)(x) 6 M f (x) + M g(x).
(iii) Homogeneity: M (αf )(x) = |α|M f (x), α ∈ R.
(iv) Translation invariance: M (τy f ) = (τy M f )(x) = M f (x − y).
With the Vitali covering lemma, we can state and prove the main results for
the maximal function.
Theorem 3.9 (The maximal function theorem). Let f be a given function
defined on Rn .
(i) If f ∈ Lp (Rn ), p ∈ [1, ∞], then the function M f is finite almost every-
where.
(ii) If f ∈ L1 (Rn ), then for every α > 0, M is of weak type (1, 1), i.e.,
- 64 - 3. The Maximal Function and Calderón-Zygmund Decomposition
m
(B(x, r)) B(x,r)
|f (y)|dy > M f (x) − ε > α.
3n 3n
Z Z
= |f (y)|dy 6 |f (y)|dy.
α Ski=1 Bji α Rn
Since this inequality is true for all compact subsets K of Eα , the proof of the
weak type inequality (ii) for the maximal operator is complete.
The above proof also gives the proof of (i) for the case when p = 1. For the
case p = ∞, by Theorem 3.4, (i) and (iii) is true with A∞ = 1.
Now, by using the Marcinkiewicz interpolation theorem between L1 →
L1,∞ and L∞ → L∞ , we can obtain simultaneously (i) and (iii) for the case
p ∈ (1, ∞).
Now, we make some clarifying comments.
3
The Heine-Borel theorem reads as follows: A set K ⊂ Rn is closed and bounded if and only if K is a
compact set (i.e., every open cover of K has a finite subcover). In words, any covering of a compact
set by a collection of open sets contains a finite sub-covering. For the proof, one can see the wiki:
http://en.wikipedia.org/wiki/Heine%E2%80%93Borel_theorem.
3.2. Hardy-Littlewood maximal function - 65 -
Remark 3.10. (1) The weak type estimate (ii) is the best possible for the
distribution function of M f , where f is an arbitrary function in L1 (Rn ).
Indeed, we replace |f (y)|dy in the definition of (3.3) by a Dirac mea-
sure dµ whose
R total measure of one is concentrated at the origin. The
integral B(x,r) dµ = 1 only if the ball B(x, r) contains the origin; other-
wise, it will be zeros. Thus,
1
M (dµ)(x) = sup
r>0, 0∈B(x,r) m
(B(x, r))
= (Vn |x|n )−1 ,
m m
= ( x : Vn |x|n < α−1 ) = (B(0, (Vn α)−1/n ))
−1
=Vn (Vn α) = 1/α.
But we can always find a sequence {fm (x)} of positive integrable func-
tions, whose L1 norm is each 1, and which converges weakly to the
measure dµ. So we cannot expect an estimate essentially stronger than
the estimate (ii) in Theorem 3.9, since, in the limit, a similar stronger
version would have to hold for M (dµ)(x).
(2) It is useful, for certain applications, to observe that
1
Ap = O , as p → 1.
p−1
In contrast with the case p > 1, when p = 1 the mapping f 7→ M f is not
bounded on L1 (Rn ). So the proof of the weak bound (ii) for M f requires a less
elementary arguments of geometric measure theory, like the Vitali covering
lemma. In fact, we have
Theorem 3.11. If f ∈ L1 (Rn ) is not identically zero, then M f is never inte-
grable on the whole of Rn , i.e., M f ∈
/ L1 (Rn ).
1 1
Z Z
M f (x) >
m (B(x, r)) B(x,r)
|f (y)|dy =
Vn (2(|x| + N ))n B(x,r)
|f (y)|dy
1 1
Z
> |f (y)|dy > kf k1
Vn (2(|x| + N ))n B(0,N ) 2Vn (2(|x| + N ))n
1
> kf k1 .
2Vn (4|x|)n
It follows that for sufficiently large |x|, we have
- 66 - 3. The Maximal Function and Calderón-Zygmund Decomposition
m
Z 1 Z ∞
=2 + ({x ∈ E : M f (x) > 2α})dα
0 1
m m
Z ∞
62 (E) + 2 ({x ∈ E : M f (x) > 2α})dα.
1
Decompose f as f1 + f2 , where f1 = f χ{x:|f (x)|>α} and f2 = f − f1 . Then, by
Theorem 3.4, it follows that
M f2 (x) 6 kM f2 k∞ 6 kf2 k∞ 6 α,
which yields
{x ∈ E : M f (x) > 2α} ⊂ {x ∈ E : M f1 (x) > α}.
Hence, by Theorem 3.9, we have
m m
Z ∞ Z ∞
({x ∈ E : M f (x) > 2α})dα 6 ({x ∈ E : M f1 (x) > α})dα
1 1
Z ∞ Z Z max(1,|f (x)|)
1 dα
Z
6C |f (x)|dxdα 6 C |f (x)| dx
1 α {x∈E:|f (x)|>α} E 1 α
Z
=C |f (x)| ln+ |f (x)|dx.
E
This completes the proof.
As a corollary of Theorem 3.9, we have the differentiability almost every-
where of the integral, expressed in (3.2).
Theorem 3.13 (Lebesgue differentiation theorem). If f ∈ Lp (Rn ), p ∈
[1, ∞], or more generally if f is locally integrable (i.e., f ∈ L1loc (Rn )), then
1
Z
lim
r→0 m
(B(x, r)) B(x,r)
f (y)dy = f (x), for a.e. x. (3.9)
3.2. Hardy-Littlewood maximal function - 67 -
Proof. We first consider the case p = 1. It suffices to show that for each α > 0,
the set
1
Z
Eα = x : lim sup
r→0 m
(B(x, r)) B(x,r)
f (y)dy − f (x) > 2α
has measure zero, because this assertion then guarantees that the set E =
S∞
k=1 E1/k has measure zero, and the limit in (3.9) holds at all points of E .
c
m
(B(x, r)) B(x,r)
(f (y) − g(y))dy
1
Z
+
m
(B(x, r)) B(x,r)
g(y)dy − g(x) + g(x) − f (x),
we find that
1
Z
lim sup
r→0 m
(B(x, r)) B(x,r)
f (y)dy − f (x) 6 M (f − g)(x) + |g(x) − f (x)|.
Consequently, if
Fα = {x : M (f − g)(x) > α} and Gα = {x : |f (x) − g(x)| > α} ,
then Eα ⊂ Fα ∪ Gα , because if u1 and u2 are positive, then u1 + u2 > 2α only
if ui > α for at least one ui .
On the one hand, Tchebychev’s inequality4 yields
m 1
(Gα ) 6 kf − gk1 ,
α
and on the other hand, the weak type estimate for the maximal function gives
m 3n
(Fα ) 6 kf − gk1 .
α
Since the function g was selected so that kf − gk1 < ε, we get
m 3n
(Eα ) 6 ε + ε =
α
1
α
3n + 1
α
ε.
m
Since ε is arbitrary, we must have (Eα ) = 0, and the proof for p = 1 is
completed.
4
Tchebychev inequality (also spelled as Chebyshev’s inequality): Suppose f > 0, and f is integrable.
If α > 0 and Eα = {x ∈ Rn : f (x) > α}, then
m 1
Z
(Eα ) 6 f dx.
α Rn
- 68 - 3. The Maximal Function and Calderón-Zygmund Decomposition
Indeed, the limit in the theorem is taken over balls that shrink to the point
x, so the behavior of f far from x is irrelevant. Thus, we expect the result to
remain valid if we simply assume integrability of f on every ball. Clearly, the
conclusion holds under the weaker assumption that f is locally integrable.
For the remained cases p ∈ (1, ∞], we have by Hölder inequality, for any
ball B,
m(B)
Z
1/p0
|f (x)|dx 6 kf kLp (B) k1kLp0 (B) 6 kf kp .
B
Thus, f ∈ L1loc (Rn ) and then the conclusion is valid for p ∈ (1, ∞]. Therefore,
we complete the proof of the theorem.
By the Lebesgue differentiation theorem, we have
Theorem 3.14. Let f ∈ L1loc (Rn ). Then
|f (x)| 6 M f (x), a.e. x ∈ Rn .
Combining with the maximal function theorem (i.e., Theorem 3.9), we get
Corollary 3.15. If f ∈ Lp (Rn ), p ∈ (1, ∞], then we have
kf kp 6 kM f kp 6 Ap kf kp .
As an application, we prove the (Gagliardo-Nirenberg-) Sobolev inequality
by using the maximal function theorem for the case 1 < p < n. We note that
the inequality also holds for the case p = 1 and one can see [Eva98, p.263-264]
for the proof.
Theorem 3.16 ((Gagliardo-Nirenberg-) Sobolev inequality). Let p ∈
(1, n) and its Sobolev conjugate p∗ = np/(n − p). Then for f ∈ D(Rn ),
we have
kf kp∗ 6 Ck∇f kp ,
where C depends only on n and p.
∞
2n−1 X 2n
6 rM (∇f )(x) 2−k = rM (∇f )(x).
n k=0
n
For the second part, by Hölder inequality, we get for 1 < p < n
|∇f (y)|
Z
n−1
dy
Rn \B(x,r) |x − y|
Z 1/p Z 1/p0
p (1−n)p0
6 |∇f (y)| dy |x − y| dy
Rn \B(x,r) Rn \B(x,r)
Z ∞ 1/p0
(1−n)p0 n−1
6 ωn−1 ρ ρ dρ k∇f kp
r
1/p0
(p − 1)ωn−1
= r1−n/p k∇f kp .
n−p
p/n
(p−1)(p−1)/n nk∇f kp
Choose r = 1/n satisfying
M (∇f )(x)
(n−p)(p−1)/n ωn−1 2p
1/p0
(p − 1)ωn−1
n n
2 rM (∇f )(x) = r1−n/p k∇f kp ,
ωn−1 n−p
- 70 - 3. The Maximal Function and Calderón-Zygmund Decomposition
then we get
|f (x)| 6 Ck∇f kp/n
p (M (∇f )(x))
1−p/n
.
Thus, by part (iii) in Theorem 3.9, we obtain for 1 < p < n
1−p/n
kf kp∗ 6Ck∇f kp/n
p kM (∇f )kp∗ (1−p/n)
(0)
Proof. We decompose Rn into a mesh of equal cubes Qk (k = 1, 2, · · · ),
whose interiors are disjoint and edges parallel to the coordinate axes, and
whose common diameter is so large that
1
Z
m (0)
(Qk ) Q(0) k
|f (x)|dx 6 α, (3.11)
since f ∈ L1 .
(0) (1)
Split each Qk into 2n congruent cubes. These we denote by Qk , k =
1, 2, · · · . There are two possibilities:
1 1
Z Z
either |f (x)|dx 6 α, or
m (1)
(Qk ) Q(1)
k m (Q
(1)
k ) Q
(1)
k
|f (x)|dx > α.
(1) (2)
In the first case, we split Qk again into 2n congruent cubes to get Qk (k =
1, 2, · · · ). In the second case, we have
1 1
Z Z
|f (x)|dx 6 2n α
α<
m (1)
(Qk ) Qk (1)
|f (x)|dx 6
2 m
−n (0)
(Qk̃ ) Qk̃
(0)
(j)
/ Ω =: ∞
A repetition of this argument shows that if x ∈ k=1 Qk then x ∈ Qkj
S
(j = 0, 1, 2, · · · ) for which
m 1
Z
(j)
(Qkj ) → 0 as j → ∞, and
m (j)
(Qkj ) Qk(j)j
|f (x)|dx 6 α (j = 0, 1, · · · ).
portional to their distances from F . Let Qk then be one of these cubes, and pk
a point of F such that
dist (F, Qk ) = dist (pk , Qk ).
Let Bk be the smallest ball whose center is pk and which contains the interior
of Qk . Let us set
m
(Bk )
γk =
m
(Qk )
.
- 72 - 3. The Maximal Function and Calderón-Zygmund Decomposition
b = j bj and g = f − b. Consequently,
P
m 1
Z Z Z
Qj
|bj |dx 6
Qj
|f (x)|dx + (Qj )
m (Qj ) Qj
f (x)dx
m
Z
62 |f (x)|dx 6 2n+1 α (Qj ),
Qj
m
which proves kbj k1 6 2n+1 α (Qj ).
Next, we need to obtain the estimates on g. Write Rn = ∪j Qj ∪ F , where
F is the
R closed set obtained by Corollary 3.18. Since b = 0 on F and f − bj =
m(Qj ) Qj f (x)dx, we have
1
f,
on F,
g= 1
Z
(3.12)
m
(Q )
j Qj
f (x)dx, on Qj .
Proof. Except when M ϕ(x) = ∞ a.e., in which case (3.13) holds trivially, M ϕ
is the density of a positive measure µ. Thus, we may assume that M ϕ(x) < ∞
a.e. x ∈ Rn and M ϕ(x) > 0. If we denote
dµ(x) = M ϕ(x)dx and dν(x) = ϕ(x)dx,
- 74 - 3. The Maximal Function and Calderón-Zygmund Decomposition
/ k Q∗k . Then there are two cases for any cube Q with
Proof. Suppose that x ∈
S
1
Z
(Q) Q m
|f (x)|dx 6 α.
m 2 αm(Q )
X
n
6α (Q) + k
Qk ∩Q6=∅
6αm(Q) + 2 αm(3Q)
n
67 αm(Q).
n
Let us return to the proof of weak type (L1 (µ), L1 (ν)). We need to prove
that there exists a constant C such that for any α > 0 and f ∈ L1 (µ)
Z
ϕ(x)dx =ν({x ∈ Rn : M f (x) > α})
n
{x∈R :M f (x)>α}
(3.14)
C
Z
6 |f (x)|M ϕ(x)dx.
α Rn
We may assume that f ∈ L1 (Rn ). In fact, if we take f` = |f |χB(0,`) , then
f` ∈ L1 (Rn ), 0 6 f` (x) 6 f`+1 (x) for x ∈ Rn and ` = 1, 2, · · · . Moreover,
lim`→∞ f` (x) = |f (x)| and
[
{x ∈ Rn : M f (x) > α} = {x ∈ Rn : M f` (x) > α}.
`
By the pointwise equivalence of M and M 0 , there exists cn > 0 such that
M f (x) 6 cn M 0 f (x) for all x ∈ Rn . Applying the Calderón-Zygmund de-
composition on Rn for f and α0 = α/(cn 7n ), we get a sequence {Qk } of
cubes satisfying
1
Z
m
0
α < |f (x)|dx 6 2n α0 .
(Qk ) Qk
By Lemma 3.22 and the pointwise equivalence of M and M 00 , we have that
Z
ϕ(x)dx
{x∈Rn :M f (x)>α}
Z
6 ϕ(x)dx
{x∈Rn :M 0 f (x)>7n α0 }
Z XZ
6 S ϕ(x)dx 6 ϕ(x)dx
k Q∗k k Q∗k
! Z
1 1
Z
m
X
6 ϕ(x)dx 0
|f (y)|dy
k
(Q k ) Q∗k α Q k
!
cn 7n X 2n
Z Z
=
α k Qk
|f (y)|
m (Q∗k ) Q∗k
ϕ(x)dx dy
cn 14n X
Z
6 |f (y)|M 00 ϕ(y)dy
α k Qk
C
Z
6 |f (y)|M ϕ(y)dy.
α Rn
Thus, M is of weak type (L1 (µ), L1 (ν)), and the inequality can be obtained
by applying the Marcinkiewicz interpolation theorem.
L ECTURE N OTES ON c 2016 by Chengchun Hao
Introduction to Harmonic Analysis Email: hcc@amss.ac.cn
Chapter 4
Singular Integrals
Among the most important of all PDEs are undoubtedly Laplace equation
∆u = 0 (4.1)
and Poisson equation
−∆u = f. (4.2)
In both (4.1) and (4.2), x ∈ Ω and the unknown is u : Ω̄ → R, u = u(x),
where Ω ⊂ Rn is a given open set. In (4.2), the function f : Ω → R is also
given. Remember that the Laplacian of u is ∆u = nk=1 ∂x2k u.
P
for k = 1, · · · , n, and so
n−1 0
∆u = v 00 (r) + v (r).
r
Hence ∆u = 0 if and only if
n−1 0
v 00 + v = 0. (4.3)
r
If v 0 6= 0, we deduce
v 00
0 0 1−n
(ln v ) = 0 = ,
v r
and hence v 0 (r) = rn−1
a
for some constant a. Consequently, if r > 0, we have
b ln r + c, n = 2,
v(r) =
b + c, n > 3,
rn−2
where b and c are constants.
These considerations motivate the following
Definition 4.2. The function
1
−
ln |x|, n = 2,
2π
Φ(x) := 1 1 (4.4)
, n > 3,
n(n − 2)Vn |x|n−2
defined for x ∈ Rn , x 6= 0, is the fundamental solution of Laplace equa-
tion.
The reason for the particular choices of the constants in (4.4) will be appar-
ent in a moment.
We will sometimes slightly abuse notation and write Φ(x) = Φ(|x|) to em-
phasize that the fundamental solution is radial. Observe also that we have the
estimates
C C
|∇Φ(x)| 6 n−1 , |∇2 Φ(x)| 6 n , (x 6= 0) (4.5)
|x| |x|
for some constant C > 0.
By construction, the function x 7→ Φ(x) is harmonic for x 6= 0. If we shift
the origin to a new point y, the PDE (4.1) is unchanged; and so x 7→ Φ(x − y)
is also harmonic as a function of x for x 6= y. Let us now take f : Rn → R
and note that the mapping x 7→ Φ(x − y)f (y) (x 6= y) is harmonic for each
point y ∈ Rn , and thus so is the sum of finitely many such expression built for
different points y. This reasoning might suggest that the convolution
1
Z
− 2π 2 (ln |x − y|)f (y)dy, n = 2,
Z
R
u(x) = Φ(x − y)f (y)dy =
1 f (y)
Z
Rn
dy, n > 3
n(n − 2)Vn Rn |x − y|n−2
(4.6)
4.1. Harmonic functions and Poisson equation - 79 -
would solve Laplace equation (4.1). However, this is wrong: we cannot just
compute
Z
∆u(x) = ∆x Φ(x − y)f (y)dy = 0. (4.7)
Rn
Indeed, as intimated by estimate (4.5), ∆Φ(x − y) is not summable near the
singularity at y = x, and so the differentiation under the integral sign above
is unjustified (and incorrect). We must proceed more carefully in calculating
∆u.
Let us for simplicity now assume f ∈ Cc2 (Rn ), that is, f is twice continu-
ously differentiable, with compact support.
Theorem 4.3 (Solving Poisson equation). Let f ∈ Cc2 (Rn ), define u by
(4.6). Then u ∈ C 2 (Rn ) and −∆u = f in Rn .
We consequently see that (4.6) provides us with a formula for a solution of
Poisson’s equation (4.2) in Rn .
Proof. Step 1: To show u ∈ C 2 (Rn ). We have
Z Z
u(x) = Φ(x − y)f (y)dy = Φ(y)f (x − y)dy,
Rn Rn
hence
u(x + hek ) − u(x) f (x + hek − y) − f (x − y)
Z
= Φ(y) dy,
h Rn h
where h 6= 0 and ek = (0, · · · , 1, · · · , 0), the 1 in the k th -slot. But
f (x + hek − y) − f (x − y) ∂f
→ (x − y)
h ∂xk
uniformly on Rn as h → 0, and thus
∂u ∂f
Z
(x) = Φ(y) (x − y)dy, k = 1, · · · , n.
∂xk Rn ∂xk
Similarly,
∂ 2u ∂ 2f
Z
(x) = Φ(y) (x − y)dy, k, j = 1, · · · , n. (4.8)
∂xk ∂xj Rn ∂xk ∂xj
As the expression on the r.h.s. of (4.8) is continuous in the variable x, we see
that u ∈ C 2 (Rn ).
Step 2: To prove the second part. Since Φ blows up at 0, we will need for
subsequent calculations to isolate this singularity inside a small ball. So fix
ε > 0. ThenZ Z
∆u(x) = Φ(y)∆x f (x − y)dy + Φ(y)∆x f (x − y)dy =: Iε + Jε .
B(0,ε) Rn \B(0,ε)
(4.9)
Now
- 80 - 4. Singular Integrals
(
Cε2 (1 + | ln ε|), n = 2,
Z
|Iε | 6 Ck∆f k∞ |Φ(y)|dy 6 (4.10)
B(0,ε) Cε2 , n > 3,
since Z Z ε Z ε
2 ε
| ln |y||dy = − 2π r ln rdr = −π r ln r|0 − rdr
B(0,ε) 0 0
2 2
= − π(ε ln ε − ε /2)
π
=πε2 | ln ε| + ε2 ,
2
for ε ∈ (0, 1] and n = 2 by an integration by parts.
An integration by parts yields
Z
Jε = Φ(y)∆x f (x − y)dy
Rn \B(0,ε)
∂f
Z Z
= Φ(y) (x − y)dσ(y) − ∇Φ(y) · ∇y f (x − y)dy
∂B(0,ε) ∂ν Rn \B(0,ε)
=:Kε + Lε ,
(4.11)
where ν denotes the inward pointing unit normal along ∂B(0, ε). We readily
check Z Z
|Kε | 6k∇f k∞ |Φ(y)|dσ(y) 6 C|Φ(ε)| dσ(y) = C|Φ(ε)|εn−1
∂B(0,ε) ∂B(0,ε)
(
Cε| ln ε|, n = 2,
6
Cε, n > 3,
(4.12)
since Φ(y) = Φ(|y|) = Φ(ε) on ∂B(0, ε) = {y ∈ R : |y| = ε}. n
Proof. Denote
1 1
Z Z
f (r) =
m (∂B(x, r)) ∂B(x,r)
u(y)dσ(y) =
ωn−1 S n−1
u(x + rz)dσ(z).
Obviously,
n
1 1 ∂u
Z X Z
0
f (r) = ∂xj u(x+rz)zj dσ(z) = (x+rz)dσ(z),
ωn−1 S n−1 j=1 ωn−1 S n−1 ∂ν
where ∂ν ∂
denotes the differentiation w.r.t. the outward normal. Thus, by
changes of variable
1 ∂u
Z
0
f (r) = n−1
(y)dσ(y).
ωn−1 r ∂B(x,r) ∂ν
By Stokes theorem, we get
1
Z
0
f (r) = ∆u(y)dy = 0.
ωn−1 rn−1 B(x,r)
Thus f (r) = const. Since limr→0 f (r) = u(x), hence, f (r) = u(x).
Next, observe that our employing polar coordinates gives, by the first iden-
tity proved just now, that
m
Z Z r Z Z r
u(y)dy = u(y)dσ(y) ds = (∂B(x, s))u(x)ds
B(x,r) 0 ∂B(x,s) 0
Z r
=u(x) nVn sn−1 ds = Vn rn u(x).
0
- 82 - 4. Singular Integrals
Proof. If ∆u 6≡ 0, then there exists some ball B(x, r) ⊂ Ω such that, say,
∆u > 0 within B(x, r). But then for f as above,
1
Z
0
0 = f (r) = n−1 ∆u(y)dy > 0,
r ωn−1 B(x,r)
is a contradiction.
because the factor eωiξ·x e−|ωξ|y satisfies this property for each fixed ξ. Thus,
u(x, y) is a harmonic function on Rn+1 + .
By Theorem 1.15, we get that u(x, y) → f (x) in L2 (Rn ) norm, as y →
0. That is, u(x, y) satisfies the boundary condition and so u(x, y) structured
above is a solution for the above Dirichlet problem.
This solution of the problem can also be written without explicit use of the
Fourier transform. For this purpose, we define the Poisson kernel Py (x) :=
P (x, y) by
n Z
|ω|
Py (x) = eωiξ·x e−|ωξ|y dξ = (F −1 e−|ωξ|y )(x), y > 0. (4.16)
2π R n
Since (4.19) is clearly translation invariant w.r.t f and also dilation invariant
w.r.t. ψ and the maximal function, it suffices to show that
(f ∗ ψ)(0) 6 AM f (0). (4.20)
4.2. Poisson kernel and Hilbert transform - 85 -
In proving (4.20), we may clearly assume that M f (0) < ∞. Let us write
λ(r) = S n−1 f (rx )dσ(x ), and Λ(r) = |x|6r f (x)dx, so
0 0
R R
Z rZ Z r
0 0 n−1
Λ(r) = f (tx )dσ(x )t dt = λ(t)tn−1 dt, i.e., Λ0 (r) = λ(r)rn−1 .
0 S n−1 0
We have Z Z ∞ Z
(f ∗ ψ)(0) = f (x)ψ(x)dx = r n−1
f (rx0 )ψ(r)dσ(x0 )dr
Rn 0 S n−1
Z ∞ Z N
= rn−1 λ(r)ψ(r)dr = lim
ε→0
λ(r)ψ(r)rn−1 dr
0 N →∞ ε
Z N Z N
0
= lim
ε→0
Λ (r)ψ(r)dr = lim
ε→0
− [Λ(r)ψ(r)]N
ε Λ(r)dψ(r) .
N →∞ ε N →∞ ε
or r → ∞, we have
0 6 lim Λ(N )ψ(N ) 6 Vn M f (0) lim N n ψ(N ) = 0,
N →∞ N →∞
which implies limN →∞ Λ(N )ψ(N ) = 0 and similarly limε→0 Λ(ε)ψ(ε) = 0.
Thus, by integration by parts, we have
Z ∞ Z ∞
(f ∗ ψ)(0) = Λ(r)d(−ψ(r)) 6 Vn M f (0) rn d(−ψ(r))
0 0
Z ∞ Z
n−1
=nVn M f (0) ψ(r)r dr = M f (0) ψ(x)dx,
0 Rn
since ψ(r) is decreasing which implies ψ (r) 6 0, and nVn = ωn−1 . This
0
Proof. For each ε > 0, the functions ψε (x) = x−1 χ|x|>ε are bounded and
define tempered distributions. It follows at once from the definition that in S 0 ,
1
lim ψε (x) = p.v. .
ε→0 x
Therefore, it will suffice to prove that in S 0
1
lim Qy − ψy = 0.
y→0 π
4.2. Poisson kernel and Hilbert transform - 89 -
|x − a|
1 ε
= lim ln + ln
π ε→0 ε |x − b|
1 |x − a|
= ln ,
π |x − b|
where it is crucial to observe how the cancellation of the odd kernel 1/x
is manifested. Note that H(χ[a,b] )(x) blows up logarithmically for x near
the points a and b and decays like x−1 as x → ±∞. See the following
graph with a = 1 and b = 3:
Proof. Since T commutes with translations and maps L2 (R) to itself, ac-
cording to Theorem 1.62, there is a bounded function m(ξ) such that
Tcf (ξ) = m(ξ)fˆ(ξ). The assumptions (b) and (c) may be written as T δε f =
sgn (ε)δε T f for all f ∈ L2 (R). By part (iv) in Proposition 1.3, we have
F (T δε f )(ξ) =m(ξ)F (δε f )(ξ) = m(ξ)|ε|−1 fˆ(ξ/ε),
sgn (ε)F (δε T f )(ξ) = sgn (ε)|ε|−1 Tcf (ξ/ε) = sgn (ε)|ε|−1 m(ξ/ε)fˆ(ξ/ε),
which means m(εξ) = sgn (ε)m(ξ), if ε 6= 0. This shows that m(ξ) =
c sgn (ξ), and the proposition is proved.
The next theorem shows that the Hilbert transform, now defined for func-
tions in S or L2 , can be extended to functions in Lp , 1 6 p < ∞.
Theorem 4.16. For f ∈ S (R), the following assertions are true:
(i) (Kolmogorov) H is of weak type (1, 1):
m C
({x ∈ R : |Hf (x)| > α}) 6 kf k1 .
α
(ii) (M. Riesz) H is of type (p, p), 1 < p < ∞:
kHf kp 6 Cp kf kp .
Let 2Ij be the interval with the same center as Ij and twice the length, and
m m
let Ω = ∪j Ij and Ω ∗ = ∪j 2Ij . Then (Ω ∗ ) 6 2 (Ω) 6 2α−1 kf k1 .
Since Hf = Hg + Hb, from parts (iv) and (vi) of Proposition 2.15, (4.23)
and (1), we have
(Hf )∗ (α) 6 (Hg)∗ (α/2) + (Hb)∗ (α/2)
m m
Z
−2
6(α/2) |Hg(x)|2 dx + (Ω ∗ ) + ({x ∈ / Ω ∗ : |Hb(x)| > α/2})
R
4
Z Z
2 −1 −1
6 2 |g(x)| dx + 2α kf k1 + 2α |Hb(x)|dx
α R R\Ω ∗
8 2 2
Z Z X
6 |g(x)|dx + kf k1 + |Hbj (x)|dx
α R α α R\Ω ∗ j
8 2 2X
Z
6 kf k1 + kf k1 + |Hbj (x)|dx.
α α α j R\2Ij
4.2. Poisson kernel and Hilbert transform - 93 -
For x ∈
/ 2Ij , we have
1 bj (y) 1 bj (y)
Z Z
Hbj (x) = p.v. dy = dy,
π Ij x − y π Ij x − y
m
since supp bj ⊂ Ij and |x − y| > (Ij )/2 for y ∈ Ij . Denote the center of Ij
by cj , then, since bj is mean zero, we have
1 bj (y)
Z Z Z
|Hbj (x)|dx = dy dx
R\2Ij R\2Ij π Ij x − y
1 1 1
Z Z
= bj (y) − dy dx
π R\2Ij Ij x − y x − cj
!
1 |y − cj |
Z Z
6 |bj (y)| dx dy
π Ij R\2Ij |x − y||x − cj |
m
!
1 (Ij )
Z Z
6 |bj (y)| 2
dx dy.
π Ij R\2Ij |x − cj |
m
The last inequality follows from the fact that |y − cj | < (Ij )/2 and |x − y| >
m
|x − cj |/2. Since |x − cj | > (Ij ), the inner integral equals
m m m
Z ∞
1 1
2 (Ij ) dr = 2 (Ij ) = 2.
m(Ij ) r2 (Ij )
Thus, by (2) and (3),
10 4 X 10 4 X
m
Z
(Hf )∗ (α) 6 kf k1 + |bj (y)|dy 6 kf k1 + 4α (Ij )
α απ j Ij α απ j
10 16 1 10 + 16/π
6 kf k1 + kf k1 = kf k1 .
α π α α
(ii) Since H is of weak type (1, 1) and of type (2, 2), by the Marcinkiewicz
interpolation theorem, we have the strong (p, p) inequality for 1 < p < 2. If
p > 2, we apply the dual estimate with the help of (4.25) and the result for
p0 < 2 (where 1/p + 1/p0 = 1):
kHf kp = sup |hHf, gi| = sup |hf, Hgi|
kgkp0 61 kgkp0 61
The integral
Hχ[1,2]
From this section on, we are going to consider singular integrals whose ker-
nels have the same essential properties as the kernel of the Hilbert transform.
We can generalize Theorem 4.16 to get the following result.
Theorem 4.18 (Calderón-Zygmund Theorem). Let K be a tempered dis-
tribution in Rn which coincides with a locally integrable function on Rn \ {0}
and satisfies
|K(ξ)|
b 6 B, (4.27)
4.3. The Calderón-Zygmund theorem - 95 -
Z
|K(x − y) − K(x)|dx 6 B, y ∈ Rn . (4.28)
|x|>2|y|
Then we have the strong (p, p) estimate for 1 < p < ∞
kK ∗ f kp 6 Cp kf kp , (4.29)
and the weak (1, 1) estimate
C
(K ∗ f )∗ (α) 6
kf k1 . (4.30)
α
We will show that these inequalities are true for f ∈ S , but they can be
extended to arbitrary f ∈ Lp as we did for the Hilbert transform. Condition
(4.28) is usually referred to as the Hörmander condition; in practice it is often
deduced from another stronger condition called the gradient condition (i.e.,
(4.31) as below).
Proposition 4.19. The Hörmander condition (4.28) holds if for every x 6= 0
C
|∇K(x)| 6 n+1 . (4.31)
|x|
Proof. By the integral mean value theorem and (4.31), we have
Z Z Z 1
|K(x − y) − K(x)|dx 6 |∇K(x − θy)||y|dθdx
|x|>2|y| |x|>2|y| 0
1Z 1Z
C|y| C|y|
Z Z
6 n+1
dxdθ 6 n+1
dxdθ
0 |x|>2|y| |x − θy| 0 |x|>2|y| (|x|/2)
Z ∞
n+1 1 n+1 1
62 C|y|ωn−1 2
dr = 2 C|y|ωn−1 = 2n Cωn−1 .
2|y| r 2|y|
This completes the proof.
Proof of Theorem 4.18. Since the proof is (essentially) a repetition of the proof
of Theorem 4.16, we will omit the details.
Let f ∈ S and T f = K ∗ f . From (4.27), it follows that
n/2 n/2
|ω| |ω| b fˆk2
kT f k2 = kT f k2 =
c kK
2π 2π
n/2 n/2
|ω| |ω| (4.32)
6 ˆ
kKk∞ kf k2 6 B
b kfˆk2
2π 2π
=Bkf k2 ,
by the Plancherel theorem (Theorem 1.26) and part (vi) in Proposition 1.3.
It will suffice to prove that T is of weak type (1, 1) since the strong (p, p)
inequality, 1 < p < 2, follows from the interpolation, and for p > 2 it follows
from the duality since the conjugate operator T 0 has kernel K 0 (x) = K(−x)
which also satisfies (4.27) and (4.28). In fact,
Z Z Z
hT f, ϕi = T f (x)ϕ(x)dx = K(x − y)f (y)dyϕ(x)dx
Rn Rn Rn
- 96 - 4. Singular Integrals
Z Z Z Z
= K(−(y − x))ϕ(x)dxf (y)dy = (K 0 ∗ ϕ)(y)f (y)dy
Rn Rn Rn Rn
=hf, T 0 ϕi.
To show that f is of weak type (1, 1), fix α > 0 and from the Calderón-
Zygmund decomposition of f at height α, then as in Theorem 4.16, we can
write f = g + b, where
(i) kgk1 6 kf k1 and kgk∞ 6 2n α.
R (ii) b = j bj , where each bj n+1
is supported in a dyadic cube Qj satisfying
P
Qj j
m
b (x)dx = 0 and kbj k1 6 2 α (Qj ). Furthermore, the cubes Qj and
Qk have disjoint interiors when j 6= k.
m
(iii) j (Qj ) 6 α−1 kf k1 .
P
The argument now proceeds as before, and the proof reduces to showing
that Z Z
|T bj (x)|dx 6 C |bj (x)|dx, (4.33)
Rn \Q∗j Qj
√
where Q∗j is the cube with the same center as Qj and whose sides are 2 n
times longer. Denote their common center by cj . Inequality (4.33) follows from
the Hörmander condition (4.28): since each bj has zero average, if x ∈ / Q∗j
Z Z
T bj (x) = K(x − y)bj (y)dy = [K(x − y) − K(x − cj )]bj (y)dy;
Qj Qj
hence,
Z Z Z !
|T bj (x)|dx 6 |K(x − y) − K(x − cj )|dx |bj (y)|dy.
Rn \Q∗j Qj Rn \Q∗j
1 f (x − y)
Z
lim dy
ε→0 π |y|>ε y
exists in the Lp norm and the resulting operator is bounded in Lp , as
has shown in Theorem 4.16.
For L2 boundedness, we have the following lemma.
Lemma 4.23. Suppose K satisfies the conditions (4.34) and (4.35) of the
above theorem with bound B. Let
K(x), |x| > ε,
Kε (x) =
0, |x| < ε.
Then, we have the estimate
sup |K
b ε (ξ)| 6 CB, ε > 0, (4.38)
ξ
where C depends only on the dimension n.
Proof. First, we prove the inequality (4.38) for the special case ε = 1. Since
K̂1 (0) = 0, thus we can assume ξ 6= 0 and have
Z
K1 (ξ) = lim
b e−ωix·ξ K1 (x)dx
R→∞ |x|6R
Z Z
−ωix·ξ
= e K1 (x)dx + lim e−ωix·ξ K1 (x)dx
|x|<2π/(|ω||ξ|) R→∞ 2π/(|ω||ξ|)<|x|6R
=:I1 + I2 .
By the condition (4.35), K(x)dx = 0 which implies
R
Z1<|x|<2π/(|ω||ξ|)
K1 (x)dx = 0.
|x|<2π/(|ω||ξ|)
Thus, |x|<2π/(|ω||ξ|) e−ωix·ξ K1 (x)dx = |x|<2π/(|ω||ξ|) [e−ωix·ξ − 1]K1 (x)dx.
R R
Hence, from the fact |eiθ − 1| 6 |θ| (see Section 1.1) and the first condition in
(4.34), we get
Z Z
|I1 | 6 |ω||x||ξ||K1 (x)|dx 6 |ω|B|ξ| |x|−n+1 dx
|x|<2π/(|ω||ξ|) |x|<2π/(|ω||ξ|)
Z 2π/(|ω||ξ|)
=ωn−1 B|ω||ξ| dr = 2πωn−1 B.
0
To estimate I2 , choose z = z(ξ) such that e−ωiξ·z = −1. This choice can be
realized if z = πξ/(ω|ξ|2 ), with |z| = π/(|ω||ξ|). Since, by changing variables
x + z = y, we get
Z Z Z
−ωix·ξ −ωi(x+z)·ξ
e K1 (x)dx = − e K1 (x)dx = − e−ωiy·ξ K1 (y − z)dy
Rn n n
ZR R
=− e−ωix·ξ K1 (x − z)dx,
Rn
4.4. Truncated integrals - 99 -
then δε−1 T δε corresponds to the kernel ε−n K(ε−1 x). Notice that if K satisfies
the assumptions of our theorem, then ε−n K(ε−1 x) also satisfies these assump-
tions with the same bounds. (A similar remark holds for the assumptions of all
the theorems in this chapter.) Now, with our K given, let K 0 = εn K(εx). Then
K 0 satisfies the conditions of our lemma with the same bound B, and so if we
denote 0
0 K (x), |x| > 1,
K1 (x) =
0, |x| < 1,
then we know that |K b 10 (ξ)| 6 CB. The Fourier transform of ε−n K10 (ε−1 x)
b 0 (εξ) which is again bounded by CB; however ε−n K 0 (ε−1 x) = Kε (x),
is K 1 1
therefore the lemma is completely proved.
We can now prove Theorem 4.21.
Proof of Theorem 4.21. Since K satisfies the conditions (4.34) and (4.35),
then Kε (x) satisfies the same conditions with bounds not greater than CB.
By Lemma 4.23 and Theorem 4.18, we have that the Lp boundedness of the
operators {Kε }ε>0 , are uniformly bounded.
Next, we prove that {Tε f1 }ε>0 is a Cauchy sequence in Lp provided f1 ∈
Cc (Rn ). In fact, we have
1
Z Z
Tε f1 (x) − Tη f1 (x) = K(y)f1 (x − y)dy − K(y)f1 (x − y)dy
|y|>ε |y|>η
Z
= sgn (η − ε) K(y)[f1 (x − y) − f1 (x)]dy,
min(ε,η)6|y|6max(ε,η)
because of the cancelation condition (4.35). For p ∈ (1, ∞), we get, by the
mean value theorem with some θ ∈ [0, 1], Minkowski’s inequality and (4.34),
that Z
kTε f1 − Tη f1 kp 6 |K(y)||∇f1 (x − θy)||y|dy
min(ε,η)6|y|6max(ε,η) p
Z
6 |K(y)|k∇f1 (x − θy)kp |y|dy
min(ε,η)6|y|6max(ε,η)
Z
6C |K(y)||y|dy
min(ε,η)6|y|6max(ε,η)
4.5. Singular integral operators commuted with dilations - 101 -
Z
6CB |y|−n+1 dy
min(ε,η)6|y|6max(ε,η)
Z max(ε,η)
=CBωn−1 dr
min(ε,η)
=CBωn−1 |η − ε|
which tends to 0 as ε, η → 0. Thus, we obtain Tε f1 converges in Lp as ε → 0
by the completeness of Lp .
Finally, an arbitrary f ∈ Lp can be written as f = f1 + f2 where f1 is
of the type described above and kf2 kp is small. We apply the basic inequality
(4.37) for f2 to get kTε f2 kp 6 Ckf2 kp , then we see that limε→0 Tε f exists in
Lp norm; that the limiting operator T also satisfies the inequality (4.37) is then
obvious. Thus, we complete the proof of the theorem.
In this section, we shall consider those operators which not only commute
with translations but also with dilations. Among these we shall study the class
of singular integral operators, falling under the scope of Theorem 4.21.
If T corresponds to the kernel K(x), then as we have already pointed out,
δε−1 T δε corresponds to the kernel ε−n K(ε−1 x). So if δε−1 T δε = T we are back
to the requirement K(x) = ε−n K(ε−1 x), i.e., K(εx) = ε−n K(x), ε > 0; that
is K is homogeneous of degree −n. Put another way
Ω(x)
K(x) = , (4.39)
|x|n
with Ω homogeneous of degree 0, i.e., Ω(εx) = Ω(x), ε > 0. This condition
on Ω is equivalent with the fact that it is constant on rays emanating from the
origin; in particular, Ω is completely determined by its restriction to the unit
sphere S n−1 .
Let us try to reinterpret the conditions of Theorem 4.21 in terms of Ω.
1) By (4.34), Ω(x) must be bounded and consequently integrable on S n−1 ;
and another condition |x|>2|y| Ω(x−y) − Ω(x) dx 6 C which is not easily re-
R
|x−y|n |x|n
stated precisely in terms of Ω. However, what is evident is that it requires a
certain continuity of Ω. Here we shall content ourselves in treating the case
where Ω satisfies the following “Dini-type” condition suggested by (4.34):
Z 1
0 w(η)
if w(η) := sup |Ω(x) − Ω(x )|, then dη < ∞. (4.40)
|x−x0 |6η 0 η
|x|=|x0 |=1
Z n−1
X
6 |y| |x|−j−1 |x − y|j−n dx
|x|>2|y| j=0
Z n−1
X
6 |y| |x|−j−1 (|x|/2)j−n dx (since |x − y| > |x| − |y| > |x|/2)
|x|>2|y| j=0
Z n−1
X Z
−n−1
= |y| 2n−j
|x| dx = 2(2 − 1)|y| n
|x|−n−1 dx
|x|>2|y| j=0 |x|>2|y|
1
=2(2n − 1)|y|ωn−1 = (2n − 1)ωn−1 .
2|y|
To estimate the first group of terms, we notice that if |x| > 2|y|, the distance
|P Q| between the projections of x − y and x on the unit sphere as in the
following picture.
x P: x x
|x|
x−y
1 y Q: |x−y| 1
y
P P x−y
x−y
Q Q
θ
O O
|y|
Case 1: |x| ≥ |x − y|, sin θ ≤ |x| Case 2: |x| ≤ |x − y|, sin θ ≤ |y|
≤ |y|
|x−y| |x|
sin π−θ
By the sine theorem, we have sin θ
|P Q|
where |OP | = 1. Since |y| 6
= 2
|OP |q √
|x|/2, we have θ 6 π2 and so cos θ > 0. Thus, cos 2θ = 1+cos θ
2
> 1/ 2.
Then, we have
x−y x sin θ sin θ √ |y| |y|
− = |P Q| = π θ
= θ
6 2 62
|x − y| |x| sin( 2 − 2 ) cos 2 |x| |x|
|y|
since sin θ 6 |x| for both cases. So the integral corresponding to the first group
of terms is dominated by
Z ∞
|y| dx
dz dr
Z Z
n n n
2 w 2 n
=2 w(2/|z|) n = 2 ωn−1 w(2/r)
|x|>2|y| |x| |x| |z|>2 |z| 2 r
Z 1
w(η)dη
=2n ωn−1 <∞
0 η
in view of changes of variables x = |y|z and the Dini-type condition (4.40).
Now, we prove (c). Since T is a bounded linear operator on L2 which com-
mutes with translations, we know, by Theorem 1.62 and Proposition 1.3, that
T can be realized in terms of a multiplier m such that Tcf (ξ) = m(ξ)fˆ(ξ). For
such operators, the fact that they commute with dilations is equivalent with the
property that the multiplier is homogeneous of degree 0.
- 104 - 4. Singular Integrals
For our particular operators we have not only the existence of m but also
an explicit expression of the multiplier in terms of the kernel. This formula is
deduced as follows.
Since K(x) is not integrable, we first consider its truncated function. Let
0 < ε < η < ∞, and
Ω(x)
, ε 6 |x| 6 η,
Kε,η (x) = |x|n
0, otherwise.
62 ln(1/|ξ 0 · x0 |).
If 0 < η|ξ 0 · x0 | 6 1, then
|ω|R/|ξ 0 ·x0 |
dt
Z
I1 6 6 2 ln(1/|ξ 0 · x0 |).
|ω|R|ξ 0 ·x0 | t
Thus,
|ω|2 2
0
|<Iε,η (ξ, x )| 6 R + 2 ln(1/|ξ 0 · x0 |),
4
and so the real part converges as ε → 0 and η → ∞. By the fundamental
theorem of calculus, we can write
Z η Z ηZ λ Z λZ η
cos(λr) − cos(µr)
dr = − sin(tr)dtdr = − sin(tr)drdt
ε r ε µ µ ε
Z λZ η Z λ
∂r cos(tr) cos(tη) − cos(tε)
= drdt = dt
µ ε t µ t
- 106 - 4. Singular Integrals
λη λ λη λ
cos(s) cos(tε) sin s sin s cos(tε)
Z Z λη
Z Z
= ds − dt = + ds − dt
µη s µ t s µη µη s2 µ t
λ
1
Z
→0 − dt = − ln(λ/µ) = ln(µ/λ), as η → ∞, ε → 0.
µ t
Take λ = |ω|R|x0 · ξ 0 |, and µ = |ω|R. So
Z ∞
0 0 0 dr
lim <(I ε,η (ξ, x )) = [cos |ω|Rr(x · ξ ) − cos |ω|Rr] = ln(1/|x0 · ξ 0 |).
ε→0
η→∞ 0 r
By the properties of Iε,η just proved, we have
2
|ω|
π
Z
|Kdε,η (ξ)| 6 + 2 0 0
R + 2 ln(1/|ξ · x |) |Ω(x0 )|dσ(x0 )
S n−1 2 4
π |ω|2 2
Z
6C( + R )ωn−1 + 2C ln(1/|ξ 0 · x0 |)dσ(x0 ).
2 4 S n−1
For n = 1, we have S = {−1, 1} and then S n−1 ln(1/|ξ 0 · x0 |)dσ(x0 ) =
0
R
Thus, we have proved the uniform boundedness of K dε,η (ξ), i.e., (i). In view
of the limit of Iε,η (ξ, x ) as ε → 0, η → ∞ just proved, and the dominated
0
if ξ 6= 0, that is (ii).
By the Plancherel theorem, if f ∈ L2 (Rn ), Kε,η ∗ f converges in L2 norm
as ε → 0 and η → ∞, and the Fourier transform of this limit is ˆ
R m(ξ)f (ξ).
However, if we keep ε fixedRand let η → ∞, then clearly Kε,η (y)f (x −
y)dy converges everywhere to |y|>ε K(y)f (x − y)dy, which is Tε f .
Letting now ε → 0, we obtain the conclusion (c) and our theorem is com-
pletely proved.
Remark 4.25. 1) In the theorem, the condition that Ω is mean zero on
S n−1 is necessary and cannot be neglected. Since in the estimate
Z
Ω(y) Ω(y)
Z Z
n
f (x − y)dy = + f (x − y)dy,
Rn |y| |y|61 |y|>1 |y|n
the main difficulty lies in the first integral. For instance, if we assume
Ω(x) ≡ 1, f is a nonzero constant, then this integral is divergent.
2) From the formula of the symbol m(ξ), it is homogeneous of degree
0 in view of the mean zero property of Ω.
3) The proof of part (c) holds under very general conditions on Ω.
Write Ω = Ωe + Ωo where Ωe is the even part of Ω, Ωe (x) = Ωe (−x),
and Ωo (x) is the odd part, Ωo (−x) = −Ωo (x). Then, because of the uni-
form boundedness of the sine integral, i.e., =Iε,η (ξ, x0 ), we required only
|Ωo (x0 )|dσ(x0 ) < ∞, i.e., the integrability of the odd part. For the
R
S n−1
even part, the proof requires the uniform boundedness of
Z
|Ωe (x0 )| ln(1/|ξ 0 · x0 |)dσ(x0 ).
S n−1
This observation is suggestive of certain generalizations of Theorem
4.21, see [Ste70, §6.5, p.49–50].
Proof. The argument for the theorem presents itself in three stages.
The first one is the proof of inequality (c) which can be obtained as a rela-
tively easy consequence of the Lp norm existence of limε→0 Tε , already proved,
and certain general properties of “approximations to the identity”.
Let T f (x) = limε→0 Tε f (x), where the limit is taken in the Lp norm. Its
existence is guaranteed by Theorem 4.24. We shall prove this part by showing
the following Cotlar inequality
T ∗ f (x) 6 M (T f )(x) + CM f (x).
Let ϕ be a smooth non-negative function on Rn , which is supported in the
unit ball, has integral equal to one, and which is also radial and decreasing in
|x|. Consider
(
Ω(x)
, |x| > ε,
Kε (x) = |x|n
0, |x| < ε.
This leads us to another function Φ defined by
Φ = ϕ ∗ K − K1 , (4.45)
where ϕ ∗ K = limε→0 ϕ ∗ Kε = limε→0 |x−y|>ε K(x − y)ϕ(y)dy.
R
The addition of the maximal function to the r.h.s of (6.23) is the main new ele
element of the proof.
To prove (6.23), fix x ∈ Rn \ ∪ j Q∗j , and ε > 0. Now the cubes Q j fall into thre
To prove (4.49), fix x ∈ Rn \ ∪j Q1)∗j for
, and
all yε∈ > |x −Now
Q j ,0. y| < ε;the cubes Qj fall into
three classes: 2) for all y ∈ Q j , |x − y| > ε;
1) for all y ∈ Qj , |x − y| < ε; 3) there is a y ∈ Q j , such that |x − y| = ε.
We now examine
2) for all y ∈ Qj , |x − y| > ε; XZ
3) there is a y ∈ Qj , such that |x − y| = ε. T ε b(x) =
Q
Kε (x − y)b(y)dy.
j j
We now examine
X ZCase 1). Kε (x − y) = 0 if |x − y| < ε, and so the integral over the cube Q j in (6
Tε b(x) = CaseK2).
ε (xKε−(x −y)b(y)dy.
y) = K(x − y), if |x − y| > ε, and(4.50)
therefore this integral over Q
j Qj Z Z
Case 1). Kε (x − y) = 0 if |x − y| < ε, and so theQ K(x − y)b(y)dy [K(x −Q
integral over the cube
= y) − K(x − c j )]b(y)dy.
j Q j j
in (4.50) is zero.
This term is majorized in absolute value by
Case 2). Kε (x − y) = K(x − y), if |x − y| > ε, and therefore this integral
over Qj equals
Z Z
K(x − y)b(y)dy = [K(x − y) − K(x − cj )]b(y)dy.
Qj Qj
This term is majorized in absolute value by
Z
|K(x − y) − K(x − cj )||b(y)|dy,
Qj
which expression appears in the r.h.s. of (4.49).
Case 3). We write simply
Z Z
Kε (x − y)b(y)dy 6 |Kε (x − y)||b(y)|dy
Qj Qj
4.6. The maximal singular integral operator - 111 -
Z
= |Kε (x − y)||b(y)|dy,
Qj ∩B(x,r)
by (ii), with r = γn ε. However, by (iii) and the fact that Ω is bounded, we have
Ω(x − y) C
|Kε (x − y)| = n
6 0 n.
|x − y| (γn ε)
Thus, in this case,
C
Z Z
Kε (x − y)b(y)dy 6 |b(y)|dy.
Qj m(B(x, r)) Qj ∩B(x,r)
If we add over all cubes Qj , we finally obtain, for r = γn ε,
XZ
|Tε b(x)| 6 |K(x − y) − K(x − cj )||b(y)|dy
j Qj
C
Z
+ |b(y)|dy.
m(B(x, r)) B(x,r)
Taking the supremum over ε gives (4.49).
This inequality can be written in the form
|T ∗ b(x)| 6 Σ + CM b(x), x ∈ F ∗,
and so
m({x ∈ R n
\ ∪j Q∗j : |T ∗ b(x)| > α/2})
6m({x ∈ R n
\ ∪j Q∗j : Σ > α/4}) + m({x ∈ R n
\ ∪j Q∗j : CM b(x) > α/4}).
R The first term in the r.h.s. is similar to (4.33), and we can get
Rn \∪j Q∗
j
n
m
Σ(x)dx 6 Ckbk1 which implies ({x ∈ R \ ∪j Q∗j : Σ > α/4}) 6
4C
α
kbk1 .
For the second one, by Theorem 3.9, i.e., the weak type estimate for the
maximal function M , we get m({x ∈ Rn \ ∪j Q∗j : CM b(x) > α/4}) 6
C
α
kbk1 .
The weak type (1, 1) property of T ∗ then follows as in the proof of the same
property for T , in Theorem 4.18 for more details.
The final stage of the proof, the passage from the inequalities of T ∗ to the ex-
istence of the limits almost everywhere, follows the familiar pattern described
in the proof of the Lebesgue differential theorem (i.e., Theorem3.13).
More precisely, for any f ∈ Lp (Rn ), 1 6 p < ∞, let
kf2 kp 6 δ.
We have already proved in the proof of Theorem 4.21 that Tε f1 con-
verges uniformly as ε → 0, so Λf1 (x) ≡ 0. By (4.37), we have kΛf2 kp 6
2Ap kf2 kp 6 2Ap δ if 1 < p < ∞. This shows Λf2 = 0, almost everywhere,
- 112 - 4. Singular Integrals
It is interesting to point out that the results of this chapter, where our func-
tions were assumes to take real or complex values, can be extended to the case
of functions taking their values in a Hilbert space. We present this generaliza-
tion because it can be put to good use in several problems. An indication of
this usefulness is given in the Littlewood-Paley theory.
We begin by reviewing quickly certain aspects of integration theory in this
context.
Let H be a separable Hilbert space. Then a function f (x), from Rn to H
is measurable if the scalar valued functions (f (x), ϕ) are measurable, where
(·, ·) denotes the inner product of H , and ϕ denotes an arbitrary vector of H .
If f (x) is such a measurable function, then |f (x)| is also measurable (as a
function with non-negative values), where | · | denotes the norm of H .
Thus, Lp (Rn , H ) is defined as the equivalent classes of measurable
functions f (x) from Rn to H , with the property that the norm kf kp =
( Rn |f (x)|p dx)1/p is finite, when p < ∞; when p = ∞ there is a similar
R
Suppose that Rf (x) ∈ L1 (Rn , H ). Then we can define its Fourier trans-
form fˆ(ξ) = Rn e−ωix·ξ f (x)dx which is an element of L∞ (Rn , H ). If
f ∈ L1 (Rn , H ) ∩ L2 (Rn , H ), then fˆ(ξ) ∈ L2 (Rn , H ) with kfˆk2 =
−n/2
|ω|
2π
kf k2 . The Fourier transform can then be extended by continuity to
a unitary mapping of the Hilbert space L2 (Rn , H ) to itself, up to a constant
multiplication.
These facts can be obtained easily from the scalar-valued case by introduc-
ing an arbitrary orthonormal basis in H .
Now suppose that H1 and H2 are two given Hilbert spaces. Assume that
f (x) takes values in H1 , and K(x) takes values in L(H1 , H2 ). Then
Z
T f (x) = K(y)f (x − y)dy,
Rn
whenever defined, takes values in H2 .
Theorem 4.27. The results in this chapter, in particular Theorem 4.18, Propo-
sition 4.19, Theorems 4.21, 4.24 and 4.26 are valid in the more general context
where f takes its value in H1 , K takes its values in L(H1 , H2 ) and T f and
Tε f take their value in H2 , and where throughout the absolute value | · | is
replaced by the appropriate norm in H1 , L(H1 , H2 ) or H2 respectively.
This theorem is not a corollary of the scalar-valued case treated in any obvi-
ous way. However, its proof consists of nothing but a identical repetition of the
arguments given for the scalar-valued case, if we take into account the remarks
made in the above paragraphs. So, we leave the proof to the interested reader.
Remark 4.28. 1) The final bounds obtained do not depend on the Hilbert
spaces H1 or H2 , but only on B, p, and n, as in the scalar-valued case.
2) Most of the argument goes through in the even greater generality
of Banach space-valued functions, appropriately defined. The Hilbert
space structure is used only in the L2 theory when applying the variant
of Plancherel’s formula.
The Hilbert space structure also enters in the following corollary.
Corollary 4.29. With the same assumptions as in Theorem 4.27, if in addition
kT f k2 = ckf k2 , c > 0, f ∈ L2 (Rn , H1 ),
then kf kp 6 A0p kT f kp , if f ∈ Lp (Rn , H1 ), if 1 < p < ∞.
Proof. We remark that the L2 (Rn , Hj ) are Hilbert spaces. In fact, let (·, ·)j de-
note the inner product of Hj , j = 1, 2, and let h·, ·ij denote the corresponding
inner product in L2 (Rn , Hj ); that is
Z
hf, gij = (f (x), g(x))j dx.
Rn
- 114 - 4. Singular Integrals
Chapter 5
Riesz Transforms and Spherical Harmonics
We look for the operators in Rn which have the analogous structural charac-
terization as the Hilbert transform. We begin by making a few remarks about
the interaction of rotations with the n-dimensional Fourier transform. We shall
need the following elementary observation.
Let ρ denote any rotation about the origin in Rn . Denote also by ρ its induced
action on functions, ρ(f )(x) = f (ρx). Then
Z Z
−ωix·ξ −1
(F ρ)f (ξ) = e f (ρx)dx = e−ωiρ y·ξ f (y)dy
n Rn
ZR
= e−ωiy·ρξ f (y)dy = F f (ρξ) = ρF f (ξ),
Rn
that is,
F ρ = ρF .
Let `(x) = (`1 (x), `2 (x), ..., `n (x)) be an n-tuple of functions defined on
R . For any rotation ρ about the origin, write ρ = (ρjk ) for its matrix realiza-
n
tion. Suppose that ` transforms like a vector. Symbolically this can be written
as
`(ρx) = ρ(`(x)),
or more explicitly
X
`j (ρx) = ρjk `k (x), for every rotation ρ. (5.1)
k
115
- 116 - 5. Riesz Transforms and Spherical Harmonics
is, the n − 1 dimensional vector (`2 (e1 ), `3 (e1 ), · · · , `n (e1 )) is left fixed by all
the rotations on this n − 1 dimensional vector space. Thus, we have to take
`2 (e1 ) = `3 (e1 ) = · · · = `n (e1 ) = 0.
Inserting again in (5.1) gives `j (ρe1 ) = ρj1 `1 (e1 ) = cρj1 . If we take a
rotation such that ρe1 = x, then we have ρj1 = xj , so `j (x) = cxj , (|x| = 1),
which proves the lemma.
We now define the n Riesz transforms. For f ∈ Lp (Rn ), 1 6 p < ∞, we set
yj
Z
Rj f (x) = lim cn n+1
f (x − y)dy, j = 1, ..., n, (5.3)
ε→0 |y|>ε |y|
Γ ((n+1)/2) π (n+1)/2
with cn = π (n+1)/2
where 1/cn = Γ ((n+1)/2)
is half the surface area of the
Ω (x)
unit sphere S of R . Thus, Rj is defined by the kernel Kj (x) = |x|
n n+1 j
n , and
xj
Ωj (x) = cn |x| .
Next, we derive the multipliers which correspond to the Riesz transforms,
and which in fact justify their definition. Denote
Ω(x) = (Ω1 (x), Ω2 (x), ..., Ωn (x)), and m(ξ) = (m1 (ξ), m2 (ξ), ..., mn (ξ)).
Let us recall the formula (4.42), i.e.,
Z
m(ξ) = Φ(ξ · x)Ω(x)dσ(x), |ξ| = 1, (5.4)
S n−1
with Φ(t) = − πi
2
sgn (ω) sgn (t) + ln |1/t|. For any rotation ρ, since Ω com-
mutes with any rotations, i.e., Ω(ρx) = ρ(Ω(x)), we have, by changes of
variables, Z Z
ρ(m(ξ)) = Φ(ξ · x)ρ(Ω(x))dσ(x) = Φ(ξ · x)Ω(ρx)dσ(x)
n−1 n−1
ZS ZS
= Φ(ξ · ρ−1 y)Ω(y)dσ(y) = Φ(ρξ · y)Ω(y)dσ(y)
S n−1 S n−1
=m(ρξ).
Thus, m commutes with rotations and so m satisfies (5.1). However, the mj
ξ
are each homogeneous of degree 0, so Lemma 5.1 shows that mj (ξ) = c |ξ|j ,
with
5.1. The Riesz transforms - 117 -
Z
c =m1 (e1 ) = Φ(e1 · x)Ω1 (x)dσ(x)
S n−1
πi
Z
= [− sgn (ω) sgn (x1 ) + ln |1/x1 |]cn x1 dσ(x)
S n−1 2
πi
Z
= − sgn (ω) cn |x1 |dσ(x) (the 2nd is 0 since it is odd w.r.t. x1 )
2 S n−1
πi Γ ((n + 1)/2) 2π (n−1)/2
= − sgn (ω) = − sgn (ω)i.
2 π (n+1)/2 Γ ((n + 1)/2)
Here we have used the fact S n−1 |x1 |dσ(x) = 2π (n−1)/2 /Γ ((n + 1)/2). There-
R
fore, we obtain
ξj ˆ
R j f (ξ) = − sgn (ω)i f (ξ), j = 1, ..., n. (5.5)
d
|ξ|
This identity and Plancherel’s theorem also imply the following “unitary”
character of the Riesz transforms
Xn
kRj f k22 = kf k22 .
j=1
By m(ρξ) = ρ(m(ξ)) proved above, we have mj (ρξ) = k ρjk mk (ξ) for
P
any rotation ρ and then mj (ρξ)fˆ(ξ) = k ρjk mk (ξ)fˆ(ξ). Taking the inverse
P
ρjk F −1 mk (ξ)fˆ(ξ) =
X X
= ρjk Rk f.
k k
But by changes of variables, we have
F −1 mj (ρξ)fˆ(ξ)
n Z
|ω|
= eωix·ξ mj (ρξ)fˆ(ξ)dξ
2π n
n ZR
|ω|
= eωiρx·η mj (η)fˆ(ρ−1 η)dη
2π Rn
=ρRj ρ−1 f,
since the Fourier transform commutes with rotations. Therefore, it reaches
X
ρRj ρ−1 f = ρjk Rk f, (5.6)
k
which is the statement that under rotations in Rn , the Riesz operators transform
in the same manner as the components of a vector.
We have the following characterization of Riesz transforms.
- 118 - 5. Riesz Transforms and Spherical Harmonics
Then the Tj is a constant multiple of the Riesz transforms, i.e., there exists a
constant c such that Tj = cRj , j = 1, ..., n.
Proof. All the elements of the proof have already been discussed. We bring
them together.
(i) Since the Tj is bounded linear on L2 (Rn ) and commutes with transla-
tions, by Theorem 1.62 they can be each realized by bounded multipliers mj ,
i.e., F (Tj f ) = mj fˆ.
(ii) Since the Tj commutes with dilations, i.e., Tj δε f = δε Tj f , in view of
Proposition 1.3, we see that F Tj δε f = mj (ξ)F δε f = mj (ξ)ε−n δε−1 fˆ(ξ) =
mj (ξ)ε−n fˆ(ξ/ε) and F δε Tj f = ε−n δε−1 F Tj f = ε−n δε−1 (mj fˆ) =
ε−n mj (ξ/ε)fˆ(ξ/ε), which imply mj (ξ) = mj (ξ/ε) or equivalently mj (εξ) =
mj (ξ), ε > 0; that is, each mj is homogeneous of degree 0.
(iii) Finally, assumption (c) has a consequence by taking the Fourier trans-
form, i.e., the relation (5.1), and so by Lemma 5.1, we can obtain the desired
conclusion.
One of the important applications of the Riesz transforms is that they can
be used to mediate between various combinations of partial derivatives of a
function.
2
Proposition 5.3. Suppose f ∈ Cc2 (Rn ). Let ∆f = nj=1 ∂∂xf2 . Then we have
P
j
the a priori bound
∂ 2f
6 Ap k∆f kp , 1 < p < ∞. (5.7)
∂xj ∂xk p
Proposition 5.4. Suppose f ∈ Cc1 (R2 ). Then we have the a priori bound
5.1. The Riesz transforms - 119 -
∂f ∂f ∂f ∂f
+ 6 Ap +i , 1 < p < ∞.
∂x1 p ∂x2 p ∂x1 ∂x2 p
Remark 5.6. At least locally, the system (5.9) is equivalent with the exis-
∂g
tence of a harmonic function g of the n + 1 variables, such that uj = ∂x j
,
j = 0, 1, 2, ..., n.
Proof. Suppose fj = Rj f , then fbj (ξ) = − sgn (ω) |ξ|j fˆ(ξ), and so by (4.15)
iξ
n Z
|ω| iξj
uj (x, y) = − sgn (ω) fˆ(ξ) eωiξ·x e−|ωξ|y dξ, j = 1, ..., n,
2π Rn |ξ|
and n Z
|ω|
u0 (x, y) = fˆ(ξ)eωiξ·x e−|ωξ|y dξ.
2π Rn
Pk are orthogonal w.r.t. it, since there exists at least one i such that αi > αi0 ,
α0
then ∂xαii xi i = 0. hP, P i = |aα |2 α! where α! = (α1 !) · · · (αn !).
P
and the infinite direct sum is taken in the sense of Hilbert space theory. That
is, if f ∈ L2 (S n−1 ), then f has the development
X∞
f (x) = Yk (x), Yk ∈ Hk , (5.11)
k=0
where the convergence is in the L (S n−1 ) norm, and
2
Z XZ
2
|f (x)| dσ(x) = |Yk (x)|2 dσ(x).
S n−1 k S n−1
1
This is implied by the well-known formula for the Euclidean Laplacian in spherical polar coordi-
nates:
∂ ∂f
∆f = r1−n rn−1 + r−2 ∆S f.
∂r ∂r
2
Sometimes, in order to emphasize the distribution between Hk and Hk , the members of Hk are
referred to as the surface spherical harmonics.
- 122 - 5. Riesz Transforms and Spherical Harmonics
On S , it follows kP =
n−1
where ∂P
∂ν
denotes differentiation w.r.t. the out-
∂
∂ν
ward normal vector. Thus, for P ∈ Hk , and Q ∈ Hj , then by Green’s theorem
∂P ∂ Q̄
Z Z
(k − j) P Q̄dσ(x) = Q̄ −P dσ(x)
S n−1 S n−1 ∂ν ∂ν
Z
= [Q̄∆P − P ∆Q̄]dx = 0,
|x|61
where ∆ is the Laplacean on R . n
(2) Indeed, let |x|2 Pk−2 be the subspace of Pk of all polynomials of the
form |x|2 P2 where P2 ∈ Pk−2 . Then its orthogonal complement w.r.t. h·, ·i
is exactly Hk . In fact, P1 is in this orthogonal complement if and only if
h|x|2 P2 , P1 i = 0 for all P2 . But h|x|2 P2 , P1 i = (P2 (∂x )∆)P1 = hP2 , ∆P1 i,
so ∆P1 = 0 and thus Pk = Hk ⊕ |x|2 Pk−2 , which proves the conclusion. In
addition, we have for P ∈ Pk
|x| P0 (x), k even,
k
2
P (x) = Pk (x) + |x| Pk−2 (x) + · · · +
|x|k−1 P1 (x), k odd,
where Pj ∈ Hj by noticing that Pj = Hj for j = 0, 1.
(3) In fact, by the further result in (2), if |x| = 1, then we have
P0 (x), k even,
P (x) = Pk (x) + Pk−2 (x) + · · · . +
P1 (x), k odd,
with Pj ∈ Hj . That is, the restriction of any polynomial on the unit sphere is a
finite linear combination of spherical harmonics. Since the restriction of poly-
nomials is dense in L2 (S n−1 ) in the norm (see [SW71, Corollary 2.3, p.141])
by the Weierstrass approximation theorem,3 the conclusion is then established.
(4) In fact, for |x| = 1, we have
∆S Yk (x) =∆(|x|−k Yk (x)) = |x|−k ∆Yk + ∆(|x|−k )Yk + 2∇(|x|−k ) · ∇Yk
=(k 2 + (2 − n)k)|x|−k−2 Yk − 2k 2 |x|−k−2 Yk
= − k(k + n − 2)|x|k−2 Yk = −k(k + n − 2)Yk ,
Pn
since j=1 xj ∂xj Yk = kYk for Yk ∈ Pk .
3
If g is continuous on S n−1 , we can approximate it uniformly by polynomials restricted to S n−1 .
5.2. Spherical harmonics and higher Riesz transforms - 123 -
P∞
(5) To prove this, we writeR (5.11) as f (x) = k=0 ak Yk (x), where the
0
Yk0 are normalized such that S n−1 |Yk0 (x)|2 dσ(x) = 1. Our assertion is then
equivalent with ak = O(k −N/2 ), as k → ∞. If f is of class C 2 , then an
application of Green’s theorem shows that
Z Z
0
∆S f Yk dσ = f ∆S Yk0 dσ.
S n−1 S n−1
Thus, if f ∈ C ∞ , then by (4)
Z Z Z ∞
X
r r 0 r
0
∆S f Yk dσ = f ∆S Yk dσ = [−k(k + n − 2)] aj Yj0 Yk0 dσ
S n−1 S n−1 S n−1 j=0
Z
=[−k(k + n − 2)]r ak |Yk0 |2 dσ = ak [−k(k + n − 2)]r .
S n−1
So ak = O(k −2r ) for every r and therefore (5.12) holds.
To prove the converse, from (5.12), we have for any r ∈ N
∞
X ∞
X
k∆rS f k22 =(∆rS f, ∆rS f ) = ( ∆rS Yj (x), ∆rS Yk (x))
j=0 k=0
∞
X ∞
X
=( [−j(j + n − 2)]r Yj (x), [−k(k + n − 2)]r Yk (x))
j=0 k=0
∞
X
= [−k(k + n − 2)]2r (Yk (x), Yk (x))
k=0
∞
X
= [−k(k + n − 2)]2r O(k −N ) 6 C,
k=0
if we take N large enough. Thus, f ∈ C ∞ (S n−1 ).
−n/2
|ω| |ω| 2
=ε−k/2−n/2 (−i sgn (ω))k δε−1/2 (Pk (ξ)e− 2 |ξ| )
2π
−n/2
|ω| |ω| 2
= (−i sgn (ω))k ε−k/2−n/2 Pk (ε−1/2 ξ)e− 2 |ξ| /ε
2π
−n/2
|ω| |ω| 2
= (−i sgn (ω))k ε−k−n/2 Pk (ξ)e− 2 |ξ| /ε .
2π
|ω|
−n/2 |ω| 2
This shows that Tn,k e− 2 εr = |ω|
2
2π
(−i sgn (ω))k ε−k−n/2 e− 2 r /ε , and
so
−k−n/2
− |ω| εr2 |ω| |ω| 2
Tn+2k,0 e 2 = (−i sgn (ω))0 ε−0−(n+2k)/2 e− 2 r /ε
2π
−k−n/2
|ω| |ω| 2
= ε−k−n/2 e− 2 r /ε .
2π
- 126 - 5. Riesz Transforms and Spherical Harmonics
|ω|
k |ω|
Thus, Tn,k e− 2 εr = |ω|
2 2
2π
(−i sgn (ω))k Tn+2k,0 e− 2 εr for ε > 0.
To finish the proof, it suffices to see that the linear combination of
− |ω| 2
{e 2 εr }0<ε<∞ is dense in R. Suppose the contrary, then there exists a (al-
|ω| 2
most everywhere) non-zero g ∈ R, such that g is orthogonal to every e− 2 εr
in the sense of R, i.e.,
Z ∞
|ω| 2
e− 2 εr g(r)r2k+n−1 dr = 0, (5.17)
0
Rs 2
for all ε > 0. Let ψ(s) = 0 e−r g(r)rn+2k−1 dr for s > 0. Then, putting
ε = 2(m + 1)/|ω|, where m is a positive integer, and by integration by parts,
we have Z ∞ Z ∞
2 2
0= e−mr ψ 0 (r)dr = 2m e−mr ψ(r)rdr.
0 0
2
By the change of variable z = e−r , this equality is equivalent to
Z 1 p
0= z m−1 ψ( ln 1/z)dz, m = 1, 2, ....
0
Since the polynomials are uniformly dense in the space of continuous func-
tions on the closed interval [0, 1], this can only be the case when ψ( ln 1/z) =
p
2
0 for all z in [0, 1]. Thus, ψ 0 (r) = e−r g(r)rn+2k−1 = 0 for almost every
r ∈ (0, ∞), contradicting the hypothesis that g(r) is not equal to 0 almost
everywhere.
k
Since the operators Tn,k and |ω| 2π
(−i sgn (ω))k Tn+2k,0 are bounded and
agree on the dense subspace, they must be equal. Thus, we have shown the
desired result.
We come now to what has been our main goal in our discussion of spherical
harmonics.
Theorem 5.12. Let Pk (x) ∈ Hk , k > 1. Then the multiplier corresponding
Pk (x)
to the transform (5.10) with the kernel |x| k+n is
Pk (ξ) Γ (k/2)
γk , with γk = π n/2 (−i sgn (ω))k .
|ξ|k Γ (k/2 + n/2)
Remark 5.13. 1) If k > 1, then Pk (x) is orthogonal to the constants on the
sphere, and so its mean value over any sphere centered at the origin is
zero.
2) The statement of the theorem can be interpreted as
Pk (x) Pk (ξ)
F k+n
= γk . (5.18)
|x| |ξ|k
3) As such it will be derived from the following closely related fact,
Pk (x) Pk (ξ)
F k+n−α
= γk,α k+α , (5.19)
|x| |ξ|
5.2. Spherical harmonics and higher Riesz transforms - 127 -
−α
|ω| Γ (k/2+α/2)
where γk,α = π n/2
2
(−i sgn (ω))k Γ (k/2+n/2−α/2) .
Lemma 5.14. The identity (5.19) holds in the sense that
Pk (x) Pk (ξ)
Z Z
k+n−α
ϕ̂(x)dx = γk,α k+α
ϕ(ξ)dξ, ∀ϕ ∈ S . (5.20)
Rn |x| Rn |ξ|
It is valid for all non-negative integer k and for 0 < α < n.
Remark 5.15. For the complex number α with <α ∈ (0, n), the lemma
and (5.19) are also valid, see [SW71, Theorem 4.1, p.160-163].
Proof. From the proof of Corollary 5.11, we have already known that
−n/2
− |ω| 2 |ω| |ω| 2
F (Pk (x)e 2 ε|x|
)= (−i sgn (ω))k ε−k−n/2 Pk (ξ)e− 2 |ξ| /ε ,
2π
so we have by the multiplication formula,
Z Z
− |ω| 2 |ω| 2
Pk (x)e 2 ε|x|
ϕ̂(x)dx = F (Pk (x)e− 2 ε|x| )(ξ)ϕ(ξ)dξ
Rn Rn
−n/2
|ω|
Z
|ω|
k −k−n/2 |ξ|2 /ε
= (−i sgn (ω)) ε Pk (ξ)e− 2 ϕ(ξ)dξ,
2π Rn
for ε > 0.
We now integrate both sides of the above w.r.t. ε, after having multiplied the
equation by a suitable power of ε, (εβ−1 , β = (k + n − α)/2, to be precise).
ThatZ is
∞ Z
|ω|
ε|x|2
εβ−1 Pk (x)e− 2 ϕ̂(x)dxdε
0 Rn
−n/2 ∞
|ω|
Z Z
|ω|
β−1 −k−n/2 |ξ|2 /ε
= (−i sgn (ω)) k
ε ε Pk (ξ)e− 2 ϕ(ξ)dξdε.
2π 0 Rn
(5.21)
By changing the order of the double integral and a change of variable, we get
Z Z ∞
|ω| 2
l.h.s. of (5.21) = Pk (x)ϕ̂(x) εβ−1 e− 2 ε|x| dεdx
Rn 0
−β Z ∞
|ω| 2
Z
t=|ω|ε|x|2 /2
==== Pk (x)ϕ̂(x) |x| tβ−1 e−t dtdx
Rn 2 0
−β
|ω|
Z
= Γ (β) Pk (x)ϕ̂(x)|x|−2β dx.
2 Rn
Similarly,
−n/2
|ω|
Z
r.h.s. of (5.21) = (−i sgn (ω)) k
Pk (ξ)ϕ(ξ)
2π Rn
Z ∞
|ω| 2
ε−(k/2+α/2+1) e− 2 |ξ| /ε dεdξ
0
- 128 - 5. Riesz Transforms and Spherical Harmonics
−n/2 −(k+α)/2
t= |ω| |ω| |ω| 2
Z
2
|ξ|2 /ε k
==== (−i sgn (ω)) Pk (ξ)ϕ(ξ) |ξ|
2π Rn 2
Z ∞
tk/2+α/2−1 e−t dtdξ
0
−n/2 −(k+α)/2
|ω| |ω|
k
= (−i sgn (ω)) Γ (k/2 + α/2)
2π 2
Z
Pk (ξ)ϕ(ξ)|ξ|−(k+α) dξ.
Rn
Thus, we get
−(k+n−α)/2
|ω|
Z
Γ ((k + n − α)/2) Pk (x)ϕ̂(x)|x|−(k+n−α) dx
2 R n
−n/2 −(k+α)/2
|ω| |ω|
= (−i sgn (ω))k Γ (k/2 + α/2)
2π 2
Z
· Pk (ξ)ϕ(ξ)|ξ|−(k+α) dξ
Rn
which leads to (5.20).
Observe that when 0 < α < n and ϕ ∈ S , then double integrals in the
above converge absolutely. Thus the formal argument just given establishes
the lemma.
Proof of Theorem 5.12. By the assumption that k > 1, we have that the inte-
gral of Pk over any sphere centered at the origin is zero. Thus for ϕ ∈ S , we
get
Pk (x) Pk (x)
Z Z
k+n−α
ϕ̂(x)dx = k+n−α
[ϕ̂(x) − ϕ̂(0)]dx
Rn |x| |x|61 |x|
Pk (x)
Z
+ k+n−α
ϕ̂(x)dx.
|x|>1 |x|
Pk (x)
Obviously, the second term tends to |x|>1 |x| k+n ϕ̂(x)dx as α → 0 by the
R
Pk (x) Pk (x)
Z Z
= k+n
ϕ̂(x)dx = lim ϕ̂(x)dx.
|x|61 |x| ε→0 ε6|x|61 |x|k+n
Thus, we obtain
Pk (x) Pk (x)
Z Z
lim k+n−α
ϕ̂(x)dx = lim ϕ̂(x)dx. (5.22)
α→0+ Rn |x| ε→0 |x|>ε |x|k+n
Similarly,
5.3. Equivalence between two classes of transforms - 129 -
Pk (ξ) Pk (ξ)
Z Z
lim k+α
ϕ(ξ)dξ = lim ϕ(ξ)dξ.
α→0+ Rn |ξ| ε→0 |ξ|>ε |ξ|k
where c is a constant,
R Ω ∈ C (S ) is a homogeneous function of degree
∞ n−1
0, and the integral S n−1 Ω(x)dσ(x) = 0. The second class is given by those
transforms T for which
F (T f )(ξ) = m(ξ)fˆ(ξ) (5.24)
where the multiplier m ∈ C ∞ (S n−1 ) is homogeneous of degree 0.
Theorem 5.16. The two classes of transforms, defined by (5.23) and (5.24)
respectively, are identical.
Proof. First, support that T is of the form (5.23). Then by Theorem 4.24, T is
of the form (5.24) with m homogeneous of degree 0 and
Z
πi
m(ξ) = c + − sgn (ω) sgn (ξ · x) + ln(1/|ξ · x|) Ω(x)dσ(x), |ξ| = 1.
S n−1 2
(5.25)
Now, we need to show m ∈ C ∞ (S n−1 ). Write the spherical harmonic de-
velopments
X∞ ∞
X XN N
X
Ω(x) = Yk (x), m(x) = Ỹk (x), ΩN (x) = Yk (x), mN (x) = Ỹk (x),
k=1 k=0 k=1 k=0
(5.26)
where Yk , Ỹk ∈ Hk in view
R of part (3) in Proposition 5.9. k starts from 1 in the
development of Ω, since S n−1 Ω(x)dx = 0 implies that Ω(x) is orthogonal to
constants, and H0 contains only constants.
Then, by Theorem 5.12, if Ω = ΩN , then m(x) = mN (x), with
Ỹk (x) = γk Yk (x), k > 1.
- 130 - 5. Riesz Transforms and Spherical Harmonics
h i
But mM (x) − mN (x) = S n−1 − πi 1
R
2
sgn (ω) sgn (y · x) + ln |y·x|
[ΩM (y) −
ΩN (y)]dσ(y). Moreover by Hölder’s inequality,
sup |mM (x) − mN (x)|
x∈S n−1
!1/2
2
πi
Z
6 sup − sgn (ω) sgn (y · x) + ln(1/|y · x|) dσ(y)
x S n−1 2
Z 1/2
2
× |ΩM (y) − ΩN (y)| dσ(y) → 0, (5.27)
S n−1
as M , N → ∞, since4 for n = 1, S 0 = {−1, 1},
2
πi π2
Z
− sgn (ω) sgn (y · x) + ln(1/|y · x|) dσ(y) = ,
S0 2 2
and for n > 2, we can pick a orthogonal matrix A satisfying Ae1 = x and
det A = 1 for |x| = 1, and then by a change of variable,
2
πi
Z
sup − sgn (ω) sgn (y · x) + ln(1/|y · x|) dσ(y)
x n−1 2
ZS 2
π 2
= sup + (ln(1/|y · x|)) dσ(y)
x S n−1 4
π2
Z
= ωn−1 + sup (ln |y · Ae1 |)2 dσ(y)
4 x n−1
2 ZS
π
= ωn−1 + sup (ln |A−1 y · e1 |)2 dσ(y)
4 x S n−1
2
z=A y π
Z
−1
==== ωn−1 + (ln |z1 |)2 dσ(z) < ∞.
4 S n−1
Here, we have used the boundedness of the integral in the r.h.s., i.e., (with the
notation z̄ = (z2 , ..., zn ), cf. [Gra04, p.A-20,p.267])
Z 1
dz1
Z Z
2 2
(ln |z1 |) dσ(z) = (ln |z1 |) √ dσ(z̄) p
S n−1 −1 1−z12 S n−2 1 − z12
√ Z 1 Z
y=z̄/ 1−z12
2
====== (ln |z1 |) (1 − z12 )(n−3)/2 dσ(y)dz1
−1 S n−2
Z 1
=ωn−2 (ln |z1 |)2 (1 − z12 )(n−3)/2 dz1
Z−1π
z1 =cos θ
====ω n−2 (ln | cos θ|)2 (sin θ)n−2 dθ = ωn−2 I1 .
0
If n > 3, then, by integration by parts,
Z π Z π Z π
2
I1 6 (ln | cos θ|) sin θdθ = −2 ln | cos θ| sin θdθ = 2 sin θdθ = 4.
0 0 0
4
There the argument is similar with some part of the proof of Theorem 4.24.
5.3. Equivalence between two classes of transforms - 131 -
R π/2
If n = 2, then, by the formula 0 (ln(cos θ))2 dθ = π2 [(ln 2)2 + π 2 /12], cf.
[GR, 4.225.8, p.531], we get
Z π Z π/2
2
I1 = (ln | cos θ|) dθ = 2 (ln(cos θ))2 dθ = π[(ln 2)2 + π 2 /12].
0 0
Thus, (5.27) shows that
∞
X
m(x) = c + γk Yk (x).
k=1
Since Ω ∈ C ∞ , we have, in view of part (5) of Proposition 5.9, that
Z
|Yk (x)|2 dσ(x) = O(k −N )
S n−1
as k → ∞ for every fixed N . However, by the explicit form of γk , we see that
γk ∼ k −n/2 , so m(x) is also indefinitely differentiable on the unit sphere, i.e.,
m ∈ C ∞ (S n−1 ).
Conversely, suppose m(x) ∈ C ∞ (S n−1 ) and let its spherical harmonic de-
velopment be as in (5.26). Set c = Ỹ0 , and Yk (x) = γ1k Ỹk (x). Then Ω(x),
given by (5.26), has mean value zero in the sphere, and is again indefinitely
differentiable there. But as we have just seen the multiplier corresponding to
this transform is m; so the theorem is proved.
As an application of this theorem and a final illustration of the singular
integral transforms we shall give the generalization of the estimates for partial
derivatives given in 5.1.
Let P (x) ∈ Pk (Rn ). We shall say that P is elliptic if P (x) vanishes only
at the origin. For any polynomial P , we consider also its corresponding differ-
ential polynomial. Thus, if P (x) = aα xα , we write P ( ∂x ∂
) as
∂ α
P P
) = aα ( ∂x
in the previous definition.
Corollary 5.17. Suppose P is a homogeneous elliptic polynomial of degree k.
∂ α
Let ( ∂x ) be any differential monomial of degree k. Assume f ∈ Cck , then we
have the a priori estimate
α
∂ ∂
f 6 Ap P f , 1 < p < ∞. (5.28)
∂x p ∂x p
∂ α
Proof. From the Fourier transform of ∂x f and P ∂x∂
f,
∂ ∂
Z
F P f (ξ) = e−ωix·ξ P f (x)dx = (ωi)k P (ξ)fˆ(ξ),
∂x Rn ∂x
and α
∂
F f (ξ) = (ωi)k ξ α fˆ(ξ),
∂x
we have the following relation
- 132 - 5. Riesz Transforms and Spherical Harmonics
α
∂ α ∂
P (ξ)F f (ξ) = ξ F P f (ξ).
∂x ∂x
α
Since P (ξ) is non-vanishing except at the origin, Pξ(ξ) is homogenous of degree
0 and is indefinitely differentiable on the unit sphere. Thus
α
∂ ∂
f =T P f ,
∂x ∂x
where T is one of the transforms of the type given by (5.24). By Theorem 5.16,
T is also given by (5.23) and hence by the result of Theorem 4.24, we get the
estimate (5.28).
L ECTURE N OTES ON c 2016 by Chengchun Hao
Introduction to Harmonic Analysis Email: hcc@amss.ac.cn
Chapter 6
The Littlewood-Paley g-function and Multipliers
133
- 134 - 6. The Littlewood-Paley g-function and Multipliers
Z ∞ 1/2
2
g(f )(x) = |∇u(x, y)| ydy . (6.1)
0
We can also define two partial g-functions, one dealing with the y differ-
entiation and the other with the x differentiations,
!1/2
Z ∞ 2 Z ∞ 1/2
∂u 2
g1 (f )(x) = (x, y) ydy , gx (f )(x) = |∇x u(x, y)| ydy .
0 ∂y 0
(6.2)
2
Obviously, g = g12 + gx2 .
The basic result for g is the following.
Theorem 6.2. Suppose f ∈ Lp (Rn ), 1 < p < ∞. Then g(f )(x) ∈ Lp (Rn ),
and
A0p kf kp 6 kg(f )kp 6 Ap kf kp . (6.3)
2 n 2
∂u X ∂u
= +
∂y L2x j=1
∂xj L2x
" n
#
(−|ωξ|fˆ(ξ)e (ωiξj fˆ(ξ)e
X
−1 −|ωξ|y −1 −|ωξ|y
= kF )k22 + kF )k22
j=1
n " n
#
|ω|
k − |ωξ|fˆ(ξ)e−|ωξ|y k22 + kωiξj fˆ(ξ)e−|ωξ|y k22
X
=
2π j=1
6.1. The Littlewood-Paley g-function - 135 -
n
|ω|
=2 ω 2 k|ξ|fˆ(ξ)e−|ωξ|y k22
2π
n
|ω|
Z
= 2 ω 2 |ξ|2 |fˆ(ξ)|2 e−2|ωξ|y dξ,
Rn 2π
and so Z ∞ Z n
|ω|
2
kg(f )k2 = y 2 ω 2 |ξ|2 |fˆ(ξ)|2 e−2|ωξ|y dξdy
0 Rn 2π
n Z ∞
|ω|
Z
2 ˆ
= 2 2
ω |ξ| |f (ξ)| 2
ye−2|ωξ|y dydξ
n 2π 0
ZR n n
|ω| 2 ˆ 1 1 |ω|
= 2 2
ω |ξ| |f (ξ)| 2
2 |ξ|2
dξ = kfˆk22
R n 2π 4ω 2 2π
1
= kf ||22 .
2
Hence,
kg(f )k2 = 2−1/2 kf k2 . (6.4)
We have also obtained kg1 (f )k2 = kgx (f )k2 = 1
2
kf k2 .
Step 2: We consider the case p 6= 2 and prove kg(f )kp 6 Ap kf kp . We
define the Hilbert spaces H1 and H2 which are to be consider now. H1 is the
one-dimensional Hilbert space of complex numbers. To define H2 , we define
first H20 as the L2 space on (0, ∞) with measure ydy, i.e.,
Z ∞
0 2 2
H2 = f : |f | = |f (y)| ydy < ∞ .
0
Let H2 be the direct sum of n + 1 copies of H20 ; so the elements of H2 can
be represented as (n + 1) component vectors whose entries belong to H20 .
Since H1 is the same as the complex numbers, then L(H1 , H2 ) is of course
identifiable with H2 . Now let ε > 0, and keep it temporarily fixed.
Define
∂Py+ε (x) ∂Py+ε (x) ∂Py+ε (x)
Kε (x) = , ,··· , .
∂y ∂x1 ∂xn
Notice that for each fixed x, Kε (x) ∈ H2 . This is the same as saying that
Z ∞ 2 Z ∞ 2
∂Py+ε (x) ∂Py+ε (x)
ydy < ∞ and ydy < ∞, for j = 1, ..., n.
0 ∂y 0 ∂xj
In fact, since Py (x) = (|x|2 +yc2n)y(n+1)/2 , we have that both ∂P∂y
y
and ∂Py
∂xj
are
bounded by (|x|2 +y2 )(n+1)/2 . So the norm in H2 of Kε (x),
A
Z ∞
2 2 ydy
|Kε (x)| 6A (n + 1)
(|x| + (y + ε)2 )n+1
2
Z0 ∞
dy
6A2 (n + 1) 6 Cε ,
0 (y + ε)2n+1
and in another way
- 136 - 6. The Littlewood-Paley g-function and Multipliers
∞
ydy A2 (n + 1)
Z
2 2
|Kε (x)| 6 A (n+1) = (|x|2 +ε2 )−n 6 C|x|−2n .
ε (|x|2 + y 2 )n+1 2n
Thus,
|Kε (x)| ∈ L1loc (Rn ). (6.5)
Similarly,
2 Z ∞ Z ∞
∂Kε (x) ydy ydy
6C 2 2 n+2
6C 2 2 n+2
6 C|x|−2n−2 .
∂xj 0 (|x| + y ) ε (|x| + y )
Therefore, Kε satisfies the gradient condition, i.e.,
∂Kε (x)
6 C|x|−(n+1) , (6.6)
∂xj
with C independent of ε.
Now we consider the operator Tε defined by
Z
Tε f (x) = Kε (t)f (x − t)dt.
Rn
The function f is complex-valued (take their value in H1 ), but Tε f (x) takes
its value in H2 . Observe that
Z ∞ 21 Z ∞ 12
2 2
|Tε f (x)| = |∇u(x, y + ε)| ydy 6 |∇u(x, y)| ydy 6 g(f )(x).
0 ε
(6.7)
Hence, kTε f (x)k2 6 2−1/2 kf k2 , if f ∈ L2 (Rn ), by (6.4). Therefore,
|K̂ε (x)| 6 2−1/2 . (6.8)
Because of (6.5), (6.6) and (6.8), by Theorem 4.27 (cf. Theorem 4.18), we get
kTε f kp 6 Ap kf kp , 1 < p < ∞ with Ap independent of ε. By (6.7), for each
x, |Tε f (x)| increases to g(f )(x), as ε → 0, so we obtain finally
kg(f )kp 6 Ap kf kp , 1 < p < ∞. (6.9)
Step 3: To derive the converse inequalities,
A0p kf kp 6 kg(f )kp , 1 < p < ∞. (6.10)
In the first step, we have shown that kg1 (f )k2 = 21 kf k2 for f ∈ L2 (Rn ).
Let u1 , u2 are the Poisson integrals of Rf1 , fR2 ∈ L2 , respectively. ThenR we have
∞
kg1 (f1 + f2 )k22 = 41 kf1 + f2 k22 , i.e., Rn 0 | ∂(u1∂y+u2 ) |2 ydydx = 41 Rn |f1 +
f2 |2 dx. It leads to the identity
Z Z ∞
∂u1 ∂u2
Z
4 (x, y) (x, y)ydydx = f1 (x)f2 (x)dx.
Rn 0 ∂y ∂y Rn
This identity, in turn, leads to the inequality, by Hölder’s inequality and the
definition of g1 ,
1
Z Z
f1 (x)f2 (x)dx 6 g1 (f1 )(x)g1 (f2 )(x)dx.
4 Rn Rn
6.1. The Littlewood-Paley g-function - 137 -
0
Suppose now in addition that f1 ∈ Lp (Rn ) and f2 ∈ Lp (Rn ) with kf2 kp0 6
1 and 1/p + 1/p0 = 1. Then by Hölder inequality and the result (6.9).
Z
f1 (x)f2 (x)dx 6 4kg1 (f1 )kp kg1 (f2 )kp0 6 4Ap0 kg1 (f1 )kp . (6.11)
Rn
Now we take the supremum in (6.11) as f2 ranges over all function in
0
L ∩ Lp , with kf2 kp0 6 1. Then, we obtain the desired result (6.10), with
2
∂ k u(x, y)
→ 0 for each x, as y → ∞.
∂y k
Thus,
Z ∞ k+1
∂ k u(x, y) ∂ u(x, s) k ds
k
=− s k.
∂y y ∂sk+1 s
By Schwarz’s inequality, therefore,
- 138 - 6. The Littlewood-Paley g-function and Multipliers
2 2
! Z
∞ ∞
∂ k u(x, y) ∂ k+1 u(x, s) 2k
Z
−2k
6 s ds s ds .
∂y k y ∂sk+1 y
n+1
Lemma 6.6. Suppose F (x, y) ∈ C(R+ ) ∩ C 2 (Rn+1
+ ), and suitably small at
infinity. Then
Z Z
y∆F (x, y)dxdy = F (x, 0)dx. (6.13)
Rn+1
+ Rn
and
∂F ∂y
Z
y −F dσ → 0,
∂D0 ∂N ∂N
as r → ∞. Here ∂D0 is the spherical part of the boundary of D. This will
certainly be the case, if for example ∆F > 0, and |F | 6 O((|x| + y)−n−ε ) and
|∇F | = O((|x| + y)−n−1−ε ), as |x| + y → ∞, for some ε > 0.
Proof. This is the same as the part (a) of Theorem 4.9. It can be proved with a
similar argument as in the proof of part (a) for Theorem 4.10.
Now we use these lemmas to give another proof for the inequality
kg(f )kp 6 Ap kf kp , 1 < p 6 2.
Another proof of kg(f )kp 6 Ap kf kp , 1 < p 6 2. Suppose first 0 6 f ∈
D(Rn ) (and at least f 6= 0Ron a nonzero measurable set). Then the Poisson
integral u of f , u(x, y) = Rn Py (t)f (x − t)dt > 0, since Py > 0 for any
x ∈ Rn and y > 0; and the majorizations u(x, y) = O((|x| + y)−n ) and
|∇u| = O((|x| + y)−n−1 ), as |x| + y → ∞ are valid. We have, by Lemma 6.5,
Lemma 6.7 and the hypothesis 1 < p 6 2,
Z ∞ Z ∞
2 2 1
(g(f )(x)) = y|∇u(x, y)| dy = yu2−p ∆up dy
0 p(p − 1) 0
2−p Z ∞
[M f (x)]
6 y∆up dy.
p(p − 1) 0
We can write this as
g(f )(x) 6 Cp (M f (x))(2−p)/2 (I(x))1/2 , (6.15)
R∞
where I(x) = 0
y∆u dy. However, by Lemma 6.6,
p
- 140 - 6. The Littlewood-Paley g-function and Multipliers
Z Z Z
p
I(x)dx = y∆u dydx = up (x, 0)dx = kf kpp . (6.16)
Rn Rn+1
+ Rn
This immediately gives the desired result for p = 2.
Next, suppose 1 < p < 2. By (6.15), Hölder’s inequality, Theorem 3.9 and
(6.16), we have, for 0 6 f ∈ D(Rn ),
Z Z
p p
(g(f )(x)) dx 6 Cp (M f (x))p(2−p)/2 (I(x))p/2 dx
Rn Rn
Z 1/r0 Z 1/r
0
6Cpp p
(M f (x)) dx I(x)dx 6 Cp0 kf kp/r p/r
p kf kp = Cp0 kf kpp ,
Rn Rn
where r = 2/p ∈ (1, 2) and 1/r + 1/r0 = 1, then r0 = 2/(2 − p).
Thus, kg(f )kp 6 Ap kf kp , 1 < p 6 2, whenever 0 6 f ∈ D(Rn ).
For general f ∈ Lp (Rn ) (which we assume for simplicity to be real-valued),
write f = f + − f − as its decomposition into positive and negative part; then
we need only approximate in norm f + and f − , each by a sequences of positive
functions in D(Rn ). We omit the routine details that are needed to complete
the proof.
Unfortunately, the elegant argument just given is not valid for p > 2. There
is, however, a more intricate variant of the same idea which does work for the
case p > 2, but we do not intend to reproduce it here.
We shall, however, use the ideas above to obtain a significant generalization
of the inequality for the g-functions.
Definition 6.8. Define the positive function
Z ∞Z λn
∗ 2 y
(gλ (f )(x)) = |∇u(x − t, y)|2 y 1−n dtdy. (6.17)
0 Rn |t| + y
Before going any further, we shall make a few comments that will help to
clarify the meaning of the complicated expression (6.17).
First, gλ∗ (f )(x) will turn out to be a pointwise majorant of g(f )(x). To un-
derstand this situation better we have to introduce still another quantity, which
is roughly midway between g and gλ∗ . It is defined as follows.
Definition 6.9. Let Γ be a fixed proper cone in Rn+1 + with vertex at the
origin and which contains (0, 1) in its interior. The exact form of Γ will
not really matter, but for the sake of definiteness let us choose for Γ the
up circular cone:
Γ = (t, y) ∈ Rn+1
+ : |t| < y, y > 0 .
For any x ∈ Rn , let Γ (x) be the cone Γ translated such that its vertex
is at x. Now define the positive Luzin’s S-function S(f )(x) by
Z Z
2 2 1−n
[S(f )(x)] = |∇u(t, y)| y dydt = |∇u(x − t, y)|2 y 1−n dydt.
Γ (x) Γ
(6.18)
y
6.1. The Littlewood-Paley g-function - 141 -
Γ Γ(x)
We assert, as we shall momentarily prove,
that (0, 1)
π
4
Proposition 6.10. O x t
g(f )(x) 6 CS(f )(x) 6 Cλ gλ∗ (f )(x). (6.19) Fig. 6.1 Γ and Γ (x) for n = 1
Figure 1: Γ and Γ(x) for n = 1
What interpretation can we put on the inequalities relating these three quan-
tities? A hint is afforded by considering three corresponding approaches to the
boundary for harmonic functions.
(a) With u(x, y) the Poisson integral of f (x), the simplest approach to the
boundary point x ∈ Rn is obtained by letting y → 0, (with x fixed). This
is the perpendicular approach, and for it the appropriate limit exists almost
everywhere, as we already know.
(b) Wider scope is obtained by allowing the variable point (t, y) to approach
(x, 0) through any cone Γ (x), (where vertex is x). This is the non-tangential
approach which will be so important for us later. As the reader may have al-
ready realized, the relation of the S-function to the g-function is in some sense
analogous to the relation between the non-tangential and the perpendicular ap-
proaches; we should add that the S-function is of decisive significance in its
own right, but we shall not pursue that matter now.
(c) Finally, the widest scope is obtained by allowing the variable point (t, y)
to approach (x, 0) in an arbitrary manner, i.e., the unrestricted approach. The
function gλ∗ has the analogous role: it takes into account the unrestricted ap-
proach for Poisson integrals.
Notice that gλ∗ (x) depends on λ. For each x, the smaller λ the greater gλ∗ (x),
and this behavior is such that that Lp boundedness of gλ∗ depends critically
on the correct relation between p and λ. This last point is probably the main
interest in gλ∗ , and is what makes its study more difficult than g or S.
After these various heuristic and imprecise indications, let us return to firm
ground. The only thing for us to prove here is the assertion (6.19).
Proof of Proposition 6.10. The inequality S(f )(x) 6 Cλ gλ∗ (f )(x) is obvious,
since the integral (6.17) majorizes that part of the integral taken only over Γ ,
and
λn
y 1
> λn
|t| + y 2
since |t| < y there. The non-trivial part of the assertion is:
g(f )(x) 6 CS(f )(x).
y
It suffices to prove this inequality for x = 0. Let By Γ
us denote by By the ball in Rn+1+ centered at (0, y) (0, y)
Γ and1:BΓyand By
Fig. 6.2 Figure
- 142 - 6. The Littlewood-Paley g-function and Multipliers
derivatives ∂u
∂y
and ∂x
∂u
k
are, like u, harmonic func-
tions. Thus, by the mean value theorem of harmonic
functions (i.e., Theorem 4.5 by noticing (0, y) is the
center of By ),
∂u(0, y) 1 ∂u(x, s)
Z
= dxds
∂y m(By ) By ∂s
where m(By ) is the n + 1 dimensional measure of By , i.e., m(By ) = cy n+1
for an appropriate constant c. By Schwarz’s inequality
2 2
∂u(0, y) 1 ∂u(x, s)
Z Z
6 dxds dxds
∂y (m(By ))2 By ∂s By
2
1 ∂u(x, s)
Z
= dxds.
m(By ) By ∂s
If we integrate this inequality, we obtain
!
Z ∞ 2 Z ∞ 2
∂u(0, y) ∂u(x, s)
Z
y dy 6 c−1 y −n dxds dy.
0 ∂y 0 By ∂s
However, (x, s) ∈ By clearly implies that c1 s 6 y 6 c2 s, for two positive
constants c1 and c2 . Thus, apart from a multiplicative factor by changing the
order of the double integrals, the last integral is majorized by
Z Z c2 s 2 2
∂u(x, s) ∂u(x, s) 1−n
Z
−n 0
y dy dxds 6 c s dxds.
Γ c1 s ∂s Γ ∂s
This is another way of saying that,
Z ∞ 2 2
∂u(0, y) ∂u(x, y)
Z
00
y dy 6 c y 1−n dxdy.
0 ∂y Γ ∂y
The same is true for the derivatives ∂xj , j = 1, ..., n, and adding the corre-
∂u
Proof. The part (a) has already been proved in Proposition 6.10. Now, we prove
(b).
For the case p > 2, only the assumption λ > 1 is relevant since 2/λ < 2 6
p.
Let ψ denote a positive function on Rn , we claim that
Z Z
∗ 2
(gλ (f )(x)) ψ(x)dx 6 Aλ (g(f )(x))2 (M ψ)(x)dx. (6.21)
Rn Rn
6.1. The Littlewood-Paley g-function - 143 -
Then M1 f (x) = M f (x), and Mµ f (x) = ((M |f |µ )(x))1/µ . From the theorem
of the maximal function, it immediately follows that, for p > µ,
1/µ 1/µ
kMµ f kp = k((M |f |µ )(x))1/µ kp = k((M |f |µ )(x))kp/µ 6 k|f |µ kp/µ = kf kp .
(6.24)
This inequality fails for p 6 µ, as in the special case µ = 1.
The substitute for Lemma 6.7 is as follows.
- 144 - 6. The Littlewood-Paley g-function and Multipliers
Lemma 6.12. Let f ∈ Lp (Rn ), p > µ > 1; if u(x, y) is the Poisson integral
of f , then
n
|t|
|u(x − t, y)| 6 A 1 + M f (x), (6.25)
y
and more generally
n/µ
|t|
|u(x − t, y)| 6 Aµ 1 + Mµ f (x). (6.26)
y
We shall now complete the proof of the inequality (6.20) for the case 1 <
p < 2, with the restriction p > 2/λ.
Let us observe that we can always find a µ ∈ [1, p) such that if we set
λ0 = λ − 2−pµ
, then one still has λ0 > 1. In fact, if µ = p, then λ − 2−p
µ
>1
since λ > 2/p; this inequality can then be maintained by a small variation of
µ. With this choice of µ, we have by Lemma 6.12
n/µ
y
|u(x − t, y)| 6 Aµ Mµ f (x). (6.27)
y + |t|
We now proceed the argument with which we treated the function g.
(gλ∗ (f )(x))2
λn
1 y
Z
1−n
= y u2−p (x − t, y)|∆up (x − t, y)|dtdy
p(p − 1) Rn+1 +
y + |t|
1 2−p ∗
6 A2−p
µ (Mµ f (x)) I (x), (6.28)
p(p − 1)
where
λ0 n
y
Z
∗ 1−n
I (x) = y ∆up (x − t, y)dtdy.
R+n+1 y + |t|
It is clear that
λ0 n
y
Z Z Z
∗ 1−n
I (x)dx = y ∆up (t, y)dxdtdy
R n n+1
R+ Rxn y + |t − x|
Z
=Cλ0 y∆up (t, y)dtdy.
Rn+1
+
For Qt (x), we have the easy estimates, Qt (x) 6 cn for |x| 6 2t and Qt (x) 6
A0 (1 + |x|2 )−(n+1)/2 , for |x| > 2|t|, from which it is obvious that At 6 A(1 +
|t|)n and hence (6.25) is Rproved.
Since u(x − t, y) = Rn Py (s)f (x − t − s)ds, and Rn Py (s)ds = 1, by
R
(6.32)
Z
−1
kρkM1 =total mass of F ρ = |F −1 ρ(x)|dx
Rn
and
6.2. Fourier multipliers on Lp - 147 -
Proof. It is clear that k · kMp is a norm. Note also that Mp is complete. Indeed,
let {ρk } is a Cauchy sequence in Mp . So does it in L∞ because of Mp ⊂ L∞ .
Thus, it is convergent in L∞ and we denote the limit by ρ. From L∞ ⊂ S 0 ,
we have F −1 ρk F f → F −1 ρF f for any f ∈ S in sense of the strong
topology on S 0 . On the other hand, {F −1 ρk F f } is also a Cauchy sequence
in Lp ⊂ S 0 , and converges to a function g ∈ Lp . By the uniqueness of limit
- 148 - 6. The Littlewood-Paley g-function and Multipliers
multipliers which are functions on Rn . The next theorem says that Mp (Rn ) is
isometrically invariant under affine transforms of Rn .
Theorem 6.16. Let a : Rn → Rm be a surjective affine transform1 with
n > m, and ρ ∈ Mp (Rm ). Then
kρ(a(·))kMp (Rn ) = kρkMp (Rm ) .
If m = n, the mapping a∗ is bijective. In particular, we have
kρ(c·)kMp (Rn ) =kρ(·)kMp (Rn ) , ∀c 6= 0, (6.36)
kρ(hx, ·i)kMp (Rn ) =kρ(·)kMp (R) , ∀x 6= 0, (6.37)
Pn
where hx, ξi = i=1 xi ξi .
Proof. It suffices to consider the case that a : Rn → Rm is a linear transform.
Make the coordinate transform
ηi = ai (ξ), 1 6 i 6 m; ηj = ξj , m + 1 6 j 6 n, (6.38)
which can be written as η = A ξ or ξ = Aη where det A 6= 0. Let A be the
−1 >
1
An affine transform of Rn is a map F : Rn → Rn of the form F (p) = Ap + q for all p ∈ Rn ,
where A is a linear transform of Rn and q ∈ Rn .
6.2. Fourier multipliers on Lp - 149 -
6kρkMp (Rm ) kf kp .
Thus, we have
kρ(a(·))kMp (Rn ) 6 kρkMp (Rm ) . (6.39)
Taking f ((A> )−1 ·) = f1 (x1 , · · · , xm )f2 (xm+1 , · · · , xn ), one can conclude
that the inverse inequality (6.39) also holds.
Now we give a simple but very useful theorem for Fourier multipliers.
Theorem 6.17 (Bernstein multiplier theorem). Assume that k > n/2 is an
integer, and that ∂xαj ρ ∈ L2 (Rn ), j = 1, · · · , n and 0 6 α 6 k. Then we have
ρ ∈ Mp (Rn ), 1 6 p 6 ∞, and
n
!n/2k
1−n/2k
X
kρkMp . kρk2 k∂xkj ρk2 .
j=1
Pn
Proof. Let t > 0 and J(x) = j=1 |xj |k . By the Cauchy-Schwartz inequality
and the Plancherel theorem, we obtain
Z Z n
X
−1 −1 −1 n/2−k
|F ρ(x)|dx = J(x) J(x)|F ρ(x)|dx . t k∂xkj ρk2 .
|x|>t |x|>t j=1
Similarly, we have
Z
|F −1 ρ(x)|dx . tn/2 kρk2 .
|x|6t
Choosing t such that kρk2 = t−k nj=1 k∂xkj ρk2 , we infer, with the help of
P
6Cy −n/2 ,
which yields (6.45).
Now, we return to the identity (6.43), and for each y divide the range of
integration into two parts, |t| 6 y/2 and |t| > y/2. In the first range, use the
estimate (6.44) on M (k) and in the second range, use the estimate (6.45). This
together with Schwarz’ inequality gives immediately
Z
(k+1) 2 −n−2k
|U (x, y)| 6Cy |u(1) (x − t, y/2)|2 dt
|t|6y/2
|u(1) (x − t, y/2)|2 dt
Z
+ Cy −n
|t|>y/2 |t|2k
6.3. The partial sums operators - 153 -
We shall now develop the second main tool in the Littlewood-Paley theory,
(the first being the usage of the functions g and g ∗ ).
Let ρ denote an arbitrary rectangle in Rn . By rectangle we shall mean, in
the rest of this chapter, a possibly infinite rectangle with sides parallel to the
axes, i.e., the Cartesian product of n intervals.
Definition 6.21. For each rectangle ρ denote by Sρ the partial sum opera-
tor, that is the multiplier operator with m = χρ = characteristic function
of the rectangle ρ. So
F (Sρ (f )) = χρ fˆ, f ∈ L2 (Rn ) ∩ Lp (Rn ). (6.48)
For this operator, we immediately have the following theorem.
Theorem 6.22.
kSρ (f )kp 6 Ap kf kp , f ∈ L2 ∩ Lp ,
if 1 < p < ∞. The constant Ap does not depend on the rectangle ρ.
However, we shall need a more extended version of the theorem which
arises when we replace complex-valued functions by functions taking their
value in a Hilbert space.
- 154 - 6. The Littlewood-Paley g-function and Multipliers
Proof. The theorem will be proved in four steps, the first two of which already
contain the essence of the matter.
Step 1: n = 1, and the rectangles ρ1 , ρ2 , · · · , ρj , · · · are the semi-infinite
intervals (−∞, 0).
It is clear that S(−∞,0) f = F −1 χ(−∞,0) F f = F −1 1− sgn 2
(ξ)
F f , so
I − i sgn (ω)H
S(−∞,0) = , (6.51)
2
6.3. The partial sums operators - 155 -
where I is the identity, and S(−∞,0) is the partial sum operator corresponding
to the interval (−∞, 0).
Now if all the rectangles are the intervals (−∞, 0), then by (6.51),
I − i sgn (ω)He
S< =
2
and so by Lemma 6.23, we have the desired result.
Step 2: n = 1, and the rectangles are the intervals (−∞, a1 ), (−∞, a2 ), · · · ,
(−∞, aj ), · · · .
Notice that F (f (x)e−ωix·a ) = fˆ(ξ + a), therefore
F (H(e−ωix·a f (x))) = −i sgn (ω) sgn (ξ)fˆ(ξ + a),
and hence F (eωix·a H(e−ωix·a f (x))) = −i sgn (ω) sgn (ξ − a)fˆ(ξ). From this,
we see that
fj − i sgn (ω)eωix·aj H(e−ωix·aj fj )
(S(−∞,aj ) fj )(x) = . (6.52)
2
If we now write symbolically e−ωix·a f for
(e−ωix·a1 f1 , · · · , e−ωix·aj fj , · · · )
with f = (f1 , · · · , fj , · · · ), then (6.52) may be written as
e −ωix·a f )
f − i sgn (ω)eωix·a H(e
S< f = , (6.53)
2
and so the result again follows in this case by Lemma 6.23.
Step 3: General n, but the rectangles ρj are the half-spaces x1 < aj , i.e.,
ρj = {x : x1 < aj }.
(1)
Let S(−∞,aj ) denote the operator defined on L2 (Rn ), which acts only on the
x1 variable, by the action given by S(−∞,aj ) . We claim that
(1)
Sρj = S(−∞,aj ) . (6.54)
This identity is obvious for L2 functions of the product form
f 0 (x1 )f 00 (x2 , · · · , xn ),
since their linear span is dense in L2 , the identity (6.54) is established.
We now use the Lp inequality, which is the result of the previous step for
each fixed x2 , x3 , · · · , xn . We raise this inequality to the pth power and integrate
w.r.t. x2 , · · · , xn . This gives the desired result for the present case. Notice that
the result holds as well if the half-space {x : x1 < aj }∞ j=1 , is replaced by the
half-space {x : x1 > aj }∞ j=1 , or if the role of the x 1 axis is taken by the x2
axis, etc.
Step 4: Observe that every general finite rectangle of the type considered is
the intersection of 2n half-spaces, each half-space having its boundary hyper-
plane perpendicular to one of the axes of Rn . Thus a 2n-fold application of the
result of the third step proves the theorem, where the family < is made up of
finite rectangles. Since the bounds obtained do not depend on the family <, we
- 156 - 6. The Littlewood-Paley g-function and Multipliers
can pass to the general case where < contains possibly infinite rectangles by
an obvious limiting argument.
We state here the continuous analogue of Theorem 6.24. Let (Γ, dγ) be a σ-
finite measure space,2 and consider the Hilbert space H of square integrable
functions on Γ , i.e., H = L2 (Γ, dγ). The elements
f ∈ Lp (Rn , H )
are the complex-valued functions
R R f (x, γ) =2 fγ (x) on Rn ×Γ , which are jointly
measuable, and for which ( Rn ( Γ |f (x, γ)| dγ)p/2 dx)1/p = kf kp < ∞, if p <
∞. Let < = {ργ }γ∈Γ , and suppose that the mapping γ → ργ is a measurable
function from Γ to rectangles; that is, the numerical-valued functions which
assign to each γ the components of the vertices of ργ are all measurable.
Suppose f ∈ L2 (Rn , H ). Then we define F = S< f by the rule
F (x, γ) = Sργ (fγ )(x), (fγ (x) = f (x, γ)).
Theorem 6.25.
kS< f kp 6 Ap kf kp , 1 < p < ∞, (6.55)
2 n p n
for f ∈ L (R , H ) ∩ L (R , H ), where the bound Ap does not depend on the
measure space (Γ, dγ), or on the function γ → ργ .
Proof. The proof of this theorem is an exact repetition of the argument given
for Theorem 6.24. The reader may also obtain it from Theorem 6.24 by a
limiting argument.
2
If µ is measure on a ring R, a set E is said to have σ-finite measure if there exists a sequence {En }
of sets in R such that E ⊂ ∪∞ n=1 En , and µ(En ) < ∞, n = 1, 2, · · · . If the measure of every set
E in R is σ-finite, the measure µ is called σ-finite on R.
3
Strictly speaking, the origin is left out; but for the sake of simplicity of terminology, we still refer to
it as the decomposition of R.
6.4. The dyadic decomposition - 157 -
r0 (t)
1, 0 6 t 6 1/2,
r0 (t) =
−1, 1/2 < t < 1, Fig. 6.4 r0 (t) and r1 (t)
r0 is extended outside the unit interval by pe- Figure 1: r0 (t) and r1 (t)
riodicity, i.e., r0 (t + 1) = r0 (t). In general, rm (t) = r0 (2m t). The sequences
of Rademacher functions are orthonormal (and in fact mutually independent)
over [0, 1]. In fact, for m < k, the integral
Z 1 Z 1 Z 2m
m k −m
rm (t)rk (t)dt = r0 (2 t)r0 (2 t)dt = 2 r0 (s)r0 (2k−m s)ds
0 0 0
Z 1 Z 1/2 Z 1
= r0 (s)r0 (2k−m s)ds = r0 (2k−m s)ds − r0 (2k−m s)ds
0 0 1/2
"Z k−m−1 Z k−m #
2 2
=2m−k r0 (t)dt − r0 (t)dt
0 2k−m−1
- 158 - 6. The Littlewood-Paley g-function and Multipliers
Z 1 Z 1
−1
=2 r0 (t)dt − r0 (t)dt = 0,
0 0
R1
so, they are orthogonal. It is clear that they are normal since 0 (rm (t))2 dt = 1.
For our purposes, their importance arises from the following fact.
Suppose ∞
P∞
2
and set m=0 am rm (t). Then for every
P
m=0 |am | < ∞ F (t) =
1 < p < ∞, F (t) ∈ L [0, 1] and
p
X∞
Ap kF kp 6 kF k2 = ( |am |2 )1/2 6 Bp kF kp , (6.57)
m=0
for two positive constants Ap and Bp .
Thus, for functions which can be expanded in terms of the Rademacher
functions, all the Lp norms, 1 < p < ∞, are comparable.
We shall also need the n-dimensional form of (6.57). We consider the
unit cube Q ⊂ Rn , Q = {t = (t1 , t2 , · · · , tn ) : 0 6 tj 6 1}. Let
m be an n-tuple of non-negative integers m = (m1 , m2 , · · · , mn ). Define
rm (t) = rm1 (t1 )rm2 (t2 ) · · · rmn (tn ). Write F (t) = am rm (t). With
P
Z 1/p
p
kF kp = |F (t)| dt ,
Q
we also have (6.57), whenever |am |2 < ∞. That is
P
∞
!1/2
X
kF kp ∼ kF k2 = |am |2 , 1 < p < ∞. (6.58)
m=0
since e µ|F | µF
6e +e −µF
, we have
Z 1
2
eµ|F (t)| dt 6 2eµ .
0
Recall the distribution function F∗ (α) = m{t ∈ [0, 1] : |F (t)| > α}. If we
take µ = α/2 in the above inequality, we have
Z Z
2 α2 α2 α2
− α2 α
F∗ (α) = dt 6 e e 2 |F (t)| dt 6 e− 2 2e 22 6 2e− 4 .
|F (t)|>α |F (t)|>α
R∞ 2
From Theorem 2.16, the above and the formula xb e−ax dx = Γ ((b +
√ 0
1)/2)/2 ab+1 , it follows immediately that
Z ∞ 1/p
p−1
kF kp = p α F∗ (α)dα
0
Z ∞ 1/p
2
p−1 − α4
6 2p α e dα = 2(pΓ (p/2))1/p ,
0
for 1 6 p < ∞, and so in general
∞
!1/2
X
kF kp 6 Ap |am |2 , 1 6 p < ∞. (6.59)
m=0
Step 3: We shall now extend the last inequality to several variables. The case
of two variables is entirely of the inductive procedure used in the proof of the
general case.
We can also limit ourselves to the situation when p > 2, since for the
case p < 2 the desired inequality is a simple consequence of Hölder’s in-
equality. (Indeed, for p < 2 and some q > 2, we have kF kLp (0,1) 6
kF kLq (0,1) k1kLqp/(q−p) (0,1) 6 kF kLq (0,1) by Hölder’s inequality.)
We have
- 160 - 6. The Littlewood-Paley g-function and Multipliers
N
X N
X N
X
F (t1 , t2 ) = am1 m2 rm1 (t1 )rm2 (t2 ) = Fm1 (t2 )rm1 (t1 ).
m1 =0 m2 =0 m1 =0
By(6.59), it follows
Z 1 !p/2
X
|F (t1 , t2 )|p dt1 6 App |Fm1 (t2 )|2 .
0 m1
Integrating this w.r.t. t2 , and using Minkowski’s inequlaity with p/2 > 1,
we have
Z 1 X !p/2 p/2 !p/2
X X
|Fm1 (t2 )|2 dt2 = |Fm1 (t2 )|2 6 k|Fm1 (t2 )|2 kp/2
0 m1 m1 m1
p/2
!p/2
X
= kFm1 (t2 )k2p .
m1
However, Fm1 (t2 ) = am1 m2 rm2 (t2 ), and therefore the case already
P
m2
proved shows that
X
kFm1 (t2 )k2p 6 A2p a2m1 m2 .
m2
Inserting this in the above gives
Z 1Z 1 !p/2
XX
|F (t1 , t2 )|p dt1 dt2 6 App a2m1 m2 ,
0 0 m1 m2
which leads to the desired inequality
kF kp 6 Ap kF k2 , 2 6 p < ∞.
Step 4: The converse inequality
kF k2 6 Bp kF kp , p>1
is a simple consequence of the direct inequality.
In fact, for any p > 1, (here we may assume p < 2) by Hölder inequality
1/2
kF k2 6 kF kp1/2 kF kp0 .
We already know that kF kp0 6 A0p0 kF k2 , p0 > 2. We therefore get
kF k2 6 (A0p0 )2 kF kp ,
which is the required converse inequality.
Now, let us return to the proof of the Littlewood-Paley square function the-
orem.
Proof of Theorem 6.26. It will be presented in five steps.
Step 1: We show here that it suffices to prove the inequality
X 1/2
2
|Sρ f (x)| 6 Ap kf kp , 1 < p < ∞, (6.60)
ρ∈∆ p
6.4. The dyadic decomposition - 161 -
0
for f ∈ L2 (Rn ) ∩ Lp (Rn ). To see this sufficiency, let g ∈ L2 (Rn ) ∩ Lp (Rn ),
and consider the identity
XZ Z
Sρ f Sρ gdx = f ḡdx
ρ∈∆ Rn Rn
L2 ∩ Lp such that kfj − f kp → 0; use the inequality (6.60) and (6.61) for fj
and fj − fj 0 ; after a simple limiting argument, we get (6.60) and (6.61) for f
as well.
Step 2: Here we shall prove the inequality (6.60) for n = 1.
We shall need first to introduce a little more notations. We let ∆1 be
the family of dyadic intervals in R, we can enumerate them as I0 , I1 ,
· · · , Im , · · · (the order is here immaterial). For each I ∈ ∆1 , we con-
sider the partial sum operator SI , and a modification of it that we now
define. Let ϕ ∈ C 1 be a fixed function with the following properties:
1, 1 6 ξ 6 2, ϕ(ξ)
ϕ(ξ) = 1
0, ξ 6 1/2, or ξ > 4.
Suppose I is any dyadic interval, and
1 2 3 4 ξ
assume that it is of the form [2k , 2k+1 ].
Define S̃I by Fig. 6.5 ϕ(ξ)
Figure 1: ϕ(ξ)
F (S̃I f )(ξ) = ϕ(2 ξ)fˆ(ξ) = ϕI (ξ)fˆ(ξ).
−k
(6.62)
- 162 - 6. The Littlewood-Paley g-function and Multipliers
That is, S̃I , like SI , is a multiplier transform where the multiplier is equal to
one on the interval I; but unlike SI , the multiplier of S̃I is smooth.
A similar definition is made for S̃I when I = [−2k+1 , −2k ]. We observe that
SI S̃I = SI , (6.63)
since SI has as multiplier the characteristic function of I.
Now for each t ∈ [0, 1], consider the multiplier transform
∞
X
T̃t = rm (t)S̃Im .
m=0
That is, for each t, T̃t is the multiplier transform whose multiplier is mt (ξ),
with
X∞
mt (ξ) = rm (t)ϕIm (ξ). (6.64)
m=0
By the definition of ϕIm , it is clear that for any ξ at most three terms in the
sum (6.64) can be non-zero. Moreover, we also see easily that
dmt B
|mt (ξ)| 6 B, (ξ) 6 , (6.65)
dξ |ξ|
where B is independent of t. Thus, by the Mihlin multiplier theorem (Theorem
6.18)
kT̃t f kp 6 Ap kf kp , for f ∈ L2 ∩ Lp , (6.66)
and with Ap independent of t. From this, it follows obviously that
Z 1 1/p
p
kT̃t f kp dt 6 Ap kf kp .
0
However, by Lemma 6.27 about the Rademacher functions,
Z 1 Z 1Z X p
p
kT̃t f kp dt = rm (t)(S̃Im f )(x) dxdt
0 0 R1
Z !p/2
X
>A0p |S̃Im f (x)|2 dx.
R1 m
Thus, we have
!1/2
X
|S̃Im (f )|2 6 Bp kf kp . (6.67)
m
p
Now using (6.63), applying the general theorem about partial sums,
Theorem 6.24, with < = ∆1 here and (6.67), we get, for F =
(S̃I0 f, S̃I1 f, · · · , S̃Im f, · · · ),
!1/2 !1/2
X X
|SIm (f )|2 = |SIm S̃Im (f )|2 = kS∆1 F kp
m m
p p
6.4. The dyadic decomposition - 163 -
!1/2
X
6Ap kF kp = Ap |S̃Im (f )|2 6 Ap Bp kf kp = Cp kf kp , (6.68)
m
p
which is the one-dimensional case of the inequality (6.60), and this is what we
had set out to prove.
Step 3: We are still in the one-dimensional case, and we write Tt for the
operator
X
Tt = rm (t)SIm .
m
Our claim is that
kTt f kLpt,x 6 Ap kf kp , 1 < p < ∞, (6.69)
with Ap independent of t, and f ∈ L ∩ L . 2 p
PN
Write TtN = m=0 rm (t)SIm , and it suffices to show that (6.69) holds,
with Tt in place of Tt (and Ap independent of N and t). Since each SIm is a
N
We now present another multiplier theorem which is one of the most im-
portant results of the whole theory. For the sake of clarity, we state first the
one-dimensional case.
Theorem 6.28. Let m be a bounded function on R1 , which is of bounded
variation on every finite interval not containing the origin. Suppose
(a) |m(ξ)|
R 6 B, −∞ < ξ < ∞,
(b) I |m(ξ)|dξ 6 B, for every dyadic interval I.
Then m ∈ Mp , 1 < p < ∞; and more precisely, if f ∈ L2 ∩ Lp ,
kTm f kp 6 Ap kf kp ,
where Ap depends only on B and p.
To present general theorem, we consider R as divided into its two half-lines,
R as divided into its four quadrants, and generally Rn as divided into its 2n
2
“octants”. Thus, the first octants in Rn will be the open “rectangle” of those
ξ all of whose coordinates are strictly positive. We shall assume that m(ξ) is
defined on each such octant and is there continuous together with its partial
derivatives up to and including order n. Thus m may be left undefined on the
set of points where one or more coordinate variables vanishes.
For every k 6 n, we regard Rk embedded in Rn in the following obvious
way: Rk is the subspace of all points of the form (ξ1 , ξ2 , · · · , ξk , 0, · · · , 0).
Theorem 6.29 (Marcinkiewicz’ multiplier theorem). Let m be a bounded
function on Rn that is C n in all 2n “octant”. Suppose also
6.5. The Marcinkiewicz multiplier theorem - 165 -
(a) |m(ξ)| 6 B,
(b) for each 0 < k 6 n,
∂km
Z
sup dξ1 · · · dξk 6 B
ξk+1 ,··· ,ξn ρ ∂ξ1 ∂ξ2 · · · ∂ξk
as ρ ranges over dyadic rectangles of Rk . (If k = n, the “sup” sign is omitted.)
(c) The condition analogous to (b) is valid for every one of the n! permuta-
tions of the variables ξ1 , ξ2 , · · · , ξn .
Then m ∈ Mp , 1 < p < ∞; and more precisely, if f ∈ L2 ∩ Lp , kTm f kp 6
Ap kf kp , where Ap depends only on B, p and n.
Proof. It will be best to prove Theorem 6.29 in the case n = 2. This case is
already completely typical of the general situation, and in doing only it we can
avoid some notational complications.
Let f ∈ L2 (R2 ) ∩ Lp (R2 ), and write F = Tm f , that is F (F (x)) =
m(ξ)fˆ(ξ).
Let ∆ denote the dyadic rectangles, and for each ρ ∈ ∆, write fρ = Sρ f ,
Fρ = Sρ F , thus Fρ = Tm fρ .
In view of Theorem 6.26, it suffices to show that
X 1/2 X 1/2
2 2
|Fρ | 6 Cp |fρ | . (6.73)
p p
ρ∈∆ ρ∈∆
The rectangles in ∆ come in four sets, those in the first, the second, the third,
and fourth quadrants, respectively. In estimating the l.h.s. of (6.73), consider
the rectangles of each quadrant separately, and assume from now on that our
rectangles belong to the first quadrant.
We will express Fρ in terms of an integral involving fρ and the partial sum
operators. That this is possible is the essential idea of the proof.
Fix ρ and assume ρ = {(ξ1 , ξ2 ) : 2k 6 ξ1 6 2k+1 , 2l 6 ξ2 6 2l+1 }. Then,
for (ξ1 , ξ2 ) ∈ ρ, it is easy to verify the identity
Z ξ2Z ξ1 2 Z ξ1
∂ m(t1 , t2 ) ∂
m(ξ1 , ξ2 ) = dt1 dt2 + m(t1 , 2l )dt1
2 l 2 k ∂t 1 ∂t 2 2k ∂t 1
Z ξ2
∂
+ m(2k , t2 )dt2 + m(2k , 2l ).
2l ∂t 2
Now let St denote the multiplier transform corresponding to the rectangle
(1)
{(ξ1 , ξ2 ) : 2k+1 > ξ1 > t1 , 2l+1 > ξ2 > t2 }. Similarly, let St1 denote the
(2)
multiplier corresponding to the interval 2k+1 > ξ1 > t1 , similarly for St2 .
(1) (2)
Thus in fact, St = St1 · St2 . Multiplying both sides of the above equation by
the function χρ fˆ and taking inverse Fourier transforms yields, by changing the
order of integrals in view of Fubini’s theorem and the fact that Sρ Tm f = Fρ ,
(1) (1) (2) (2)
and St1 Sρ = St1 , St2 Sρ = St2 , St Sρ = St , we have
- 166 - 6. The Littlewood-Paley g-function and Multipliers
Fρ =Tm Sρ f = F −1 mχρ fˆ
n Z h Z ξ2 Z ξ1 ∂ 2 m(t , t )
|ω| 1 2
i
= e ωix·ξ
dt1 dt2 χρ (ξ)fˆ(ξ) dξ
2π R2 2l 2k ∂t1 ∂t2
n Z Z ξ1
|ω| h ∂ i
+ eωix·ξ m(t1 , 2l )dt1 χρ (ξ)fˆ(ξ) dξ
2π R2 2k ∂t1
n Z Z ξ2
|ω| h ∂ i
+ eωix·ξ m(2k , t2 )dt2 χρ (ξ)fˆ(ξ) dξ
2π R2 2l ∂t2
−1 k l
+ F m(2 , 2 )χρ (ξ)f (ξ) ˆ
n Z Z 2l+1 Z 2k+1 2
|ω| ωix·ξ ∂ m(t1 , t2 )
= e χ[2k ,ξ1 ] (t1 )χ[2l ,ξ2 ] (t2 )dt1 dt2
2π R2 2l 2k ∂t1 ∂t2
· χρ (ξ)fˆ(ξ)dξ
n Z Z 2k+1
|ω| ∂
+ e ωix·ξ
m(t1 , 2l )χ[2k ,ξ1 ] (t1 )dt1 χρ (ξ)fˆ(ξ)dξ
2π R 2 2k ∂t 1
n Z Z 2l+1
|ω| ∂
+ eωix·ξ m(2k , t2 )χ[2l ,ξ2 ] (t2 )dt2 χρ (ξ)fˆ(ξ)dξ
2π R 2 2l ∂t 2
k l
+ m(2 , 2 )fρ
n Z 2l+1 Z 2k+1
|ω|
Z
= eωix·ξ χ[t1 ,2k+1 ] (ξ1 )χ[t2 ,2l+1 ] (ξ2 )χρ (ξ)fˆ(ξ)dξ
2π 2l 2k R2
2
∂ m(t1 , t2 )
· dt1 dt2
∂t1 ∂t2
n Z 2k+1 Z
|ω| ∂
+ eωix·ξ χ[t1 ,2k+1 ] (ξ1 )χρ (ξ)fˆ(ξ)dξ m(t1 , 2l )dt1
2π 2k R2 ∂t 1
n Z 2l+1 Z
|ω| ∂
+ eωix·ξ χ[t2 ,2l+1 ] (ξ2 )χρ (ξ)fˆ(ξ)dξ m(2k , t2 )dt2
2π 2l R 2 ∂t 2
k l
+ m(2 , 2 )fρ
Z 2k+1
∂ 2 m(t1 , t2 ) ∂
Z
(1)
= St fρ dt1 dt2 + St1 fρ m(t1 , 2l )dt1
ρ ∂t 1 ∂t 2 2 k ∂t 1
Z 2l+1
(2) ∂
+ St2 fρ m(2k , t2 )dt2 + m(2k , 2l )fρ .
2 l ∂t 2
We apply the Cauchy-Schwarz inequality in the first three terms of the above
w.r.t. the measures |∂t1 ∂t2 m(t1 , t2 )|dt1 dt2 , |∂t1 m(t1 , 2l )|dt1 , |∂t2 m(2k , t2 )|dt2 ,
respectively, and we use the assumptions of the theorem to deduce
2
Z
2 ∂ 2m Z ∂ 2 m
|Fρ | . |St fρ | dt1 dt2 dt1 dt2
ρ ∂t1 ∂t2 ρ ∂t1 ∂t2
6.5. The Marcinkiewicz multiplier theorem - 167 -
2k+1 Z 2k+1 ∂
Z
(1) ∂
+ |St1 fρ |2 l
m(t1 , 2 ) dt1 m(t1 , 2l ) dt1
2k ∂t1 2k ∂t1
Z 2l+1 ∂ Z 2l+1
∂
(2) 2 k k
+ |St2 fρ | m(2 , t2 ) dt2 m(2 , t2 ) dt2
2l ∂t2 2l ∂t2
+ |m(2k , 2l )|2 |fρ |2
∂ 2m ∂m(t1 , 2l )
Z Z
0 2 (1)
6B |St fρ | dt1 dt2 + |St1 fρ |2 dt1
ρ ∂t1 ∂t2 I1 ∂t1
k
2 ∂m(2 , t2 )
Z
(2) 2
+ |St2 fρ | dt2 + |fρ |
I2 ∂t2
==1ρ + =2ρ + =3ρ + =4ρ , with ρ = I1 × I2 .
To estimate k( ρ |Fρ |2 )1/2 kp , we estimate separately the contributions of each
P
of the four terms on the r.h.s. of the above inequality by the use of Theorem
6.25. To apply that theorem in the case of =1ρ we take for Γ the first quadrant,
2 m(t ,t )
and dγ = | ∂ ∂t 1 2
1 ∂t2
|dt1 dt2 , the functions γ → ργ are constant on the dyadic
rectangles. Since for every rectangle,
∂ 2 m(t1 , t2 )
Z Z
dγ = dt1 dt2 6 B,
ρ ρ ∂t1 ∂t2
then
!1/2 !1/2
X X
1 2
|=ρ | 6 Cp |fρ | .
ρ ρ
p p
Similarly, for =2ρ , =3ρ and =4ρ , which concludes the proof.
L ECTURE N OTES ON c 2016 by Chengchun Hao
Introduction to Harmonic Analysis Email: hcc@amss.ac.cn
Chapter 7
Sobolev Spaces
169
- 170 - 7. Sobolev Spaces
Z Z
Is f (x)g(x)dx = fˆ(ξ)(|ω||ξ|)−s ĝ(ξ)dξ
Rn Rn
whenever f, g ∈ S .
Proof. Part (a) is merely a restatement of Lemma 5.14 since γ(s) = |ω|s γ0,s .
Part (b) follows immediately from part (a) by writing
1
Z Z
Is f (x) = f (x − y)|y|−n+s
dy = (|ω||ξ|)−s f \
(x − ·)dξ
γ(s) Rn Rn
Z Z
−s ˆ
= (|ω||ξ|) f (ξ)e ωiξ·x
dξ = (|ω||ξ|)−s fˆ(ξ)e−ωiξ·x dξ,
Rn Rn
so Z Z Z
Is f (x)g(x)dx = (|ω||ξ|)−s fˆ(ξ)e−ωiξ·x dξg(x)dx
Rn ZR
n Rn
= (|ω||ξ|)−s fˆ(ξ)ĝ(ξ)dξ.
Rn
This completes the proof.
Now, we state two further identities which can be obtained from Lemma 7.2
and which reflect essential properties of the potentials Is .
Is (It f ) = Is+t f, f ∈ S , s, t > 0, s + t < n. (7.6)
∆(Is f ) = Is (∆f ) = −Is−2 f, f ∈ S , n > 3, 2 6 s 6 n. (7.7)
The deduction of these two identities have no real difficulties, and these are
best left to the interested reader to work out.
A simple consequence of (7.6) is the n-dimensional variant of the beta func-
tion,1
γ(s)γ(t) −n+(s+t)
Z
|x − y|−n+s |y|−n+t dy = |x| (7.8)
Rn γ(s + t)
with s, t > 0 and s + t < n. Indeed, for any ϕ ∈ S , we have, by the definition
of Riesz potentials and (7.6), that
ZZ
|x − y|−n+s |y|−n+t dyϕ(z − x)dx
n n
Z R ×R Z
= |y|−n+t |x − y|−n+s ϕ(z − y − (x − y))dxdy
n Rn
ZR
= |y|−n+t γ(s)Is ϕ(z − y)dy = γ(s)γ(t)It (Is ϕ)(z) = γ(s)γ(t)Is+t ϕ(z)
Rn
γ(s)γ(t)
Z
= |x|−n+(s+t) ϕ(z − x)dx.
γ(s + t) Rn
By the arbitrariness of ϕ, we have the desired result.
1
The beta function, also called the Euler integral of the first kind, is a special function defined
by B(x, y) = 01 tx−1 (1 − t)y−1 dt for <x > 0 and <y > 0. It has the relation with Γ -function:
R
We have considered the Riesz potentials formally and the operation for
Schwartz functions. But since the Riesz potentials are integral operators, it
is natural to inquire about their actions on the spaces Lp (Rn ).
For this reason, we formulate the following problem. Given s ∈ (0, n), for
what pairs p and q, is the operator f → Is f bounded from Lp (Rn ) to Lq (Rn )?
That is, when do we have the inequality
kIs f kq 6 Akf kp ? (7.9)
There is a simple necessary condition, which is merely a reflection of the
homogeneity of the kernel (γ(s))−1 |y|−n+s . In fact, we have
Proposition 7.3. If the inequality (7.9) holds for all f ∈ S and a finite
constant A, then 1/q = 1/p − s/n.
Proof. Let us consider the dilation operator δε , defined by δε f (x) = f (εx) for
ε > 0. Then clearly, for ε > 0
1
Z
(δε−1 Is δε f )(x) = |ε−1 x − y|−n+s f (εy)dy
γ(s) Rn
z=εy −n 1
Z
==ε |ε−1 (x − z)|−n+s f (z)dz
γ(s) Rn
−s
=ε Is f (x). (7.10)
Also
kδε f kp = ε−n/p kf kp , kδε−1 Is f kq = εn/q kIs f kq . (7.11)
Thus, by (7.9)
kIs f kq =εs kδε−1 Is δε f kq = εs+n/q kIs δε f kq
6Aεs+n/q kδε f kp = Aεs+n/q−n/p kf kp .
If kIs f kq 6= 0, then the above inequality implies
1/q = 1/p − s/n. (7.12)
If f 6= 0 is non-negative, then Is f > 0 everywhere and hence kIs f kq > 0, and
we can conclude the desired relations.
Next, we observe that the inequality must fail at the endpoints p = 1 (then
q = n/(n − s)) and q = ∞ (then p = n/s).
Let us consider the case p = 1. It is not hard to see that the presumed
inequality
kIs f kn/(n−s) 6 Akf k1 , (7.13)
Rcannot hold. In fact, we can choose a nice positive −n
function ϕ ∈ L1 with
ϕ = 1 and a compact support. Then, with ϕε (x) = ε ϕ(x/ε), we have that
as ε → 0+ ,
Is (ϕε )(x) → (γ(s))−1 |x|−n+s .
- 172 - 7. Sobolev Spaces
2
Fatou’s lamma: If {fk } is a sequence of nonnegative measurable functions, then
Z Z
lim inf fk dµ 6 lim inf fk dµ.
k→∞ k→∞
7.1. Riesz potentials and fractional integrals - 173 -
∞ Z
X
6 (2−(j+1) δ)−n+s |f (y)|dy
j=0 Ej
X∞ Z
6 (2−(j+1) δ)−n+s |f (y)|dy
j=0 B(x,2−j δ)
∞
(2−(j+1) δ)−n+s m(B(x, 2−j δ))
X Z
= |f (y)|dy
j=0
m(B(x, 2−j δ)) B(x,2−j δ)
∞
(2−(j+1) δ)−n+s Vn (2−j δ)n
X Z
= |f (y)|dy
j=0
m(B(x, 2−j δ)) B(x,2−j δ)
∞
X Vn δ s 2n
s n−s
6Vn δ 2 2−sj M f (x) = M f (x).
j=0
2s − 1
Now, we derive an estimate for Hδ (x). By Hölder’s inequality and the con-
dition 1/p > s/n (i.e., q < ∞), we obtain
Z 1/p0
(−n+s)p0
|Hδ (x)| 6kf kp |x − y| dy
Rn \B(x,δ)
Z Z ∞ 1/p0
(−n+s)p0 n−1
=kf kp r r drdσ
S n−1 δ
Z ∞ 1/p0
1/p0 (−n+s)p0 +n−1
=ωn−1 kf kp r dr
δ
1/p0
ωn−1 0
= 0
δ n/p −(n−s) kf kp = C(n, s, p)δ s−n/p kf kp .
(n − s)p − n
By the above two inequalities, we have
|γ(s)Is f (x)| 6 C(n, s)δ s M f (x) + C(n, s, p)δ s−n/p kf kp =: F (δ).
Choose δ = C(n, s, p)[kf kp /M f ]p/n , such that the two terms of the r.h.s. of
the above are equal, i.e., the minimizer of F (δ), to get
|γ(s)Is f (x)| 6 C(M f )1−ps/n kf kps/n
p .
Therefore, by part (i) of Theorem 3.9 for maximal functions, i.e., M f is
finite almost everywhere if f ∈ Lp (1 6 p 6 ∞), it follows that |Is f (x)| is
finite almost everywhere, which proves part (a) of the theorem.
By part (iii) of Theorem 3.9, we know kM f kp 6 Ap kf kp (1 < p 6 ∞),
thus
kIs f kq 6 CkM f k1−ps/n
p kf kps/n
p = Ckf kp .
This gives the proof of part (b).
Finally, we prove (c). Since we also have |Hδ (x)| 6 kf k1 δ −n+s , taking
α = kf k1 δ −n+s , i.e., δ = (kf k1 /α)1/(n−s) , by part (ii) of Theorem 3.9, we get
m{x : |Is f (x)| > 2(γ(s))−1 α}
- 174 - 7. Sobolev Spaces
While the behavior of the kernel (γ(s))−1 |x|−n+s as |x| → 0 is well suited
for their smoothing properties, their decay as |x| → ∞ gets worse as s in-
creases.
We can slightly adjust the Riesz potentials such that we maintain their es-
sential behavior near zero but achieve exponential decay at infinity. The sim-
plest way to achieve this is by replacing the “nonnegative” operator −∆ by
the “strictly positive” operator I − ∆, where I = identity. Here the terms non-
negative and strictly positive, as one may have surmised, refer to the Fourier
transforms of these expressions.
Definition 7.5. Let s > 0. The Bessel potential of order s is the operator
Js = (I − ∆)−s/2
whose action on functions is given by
Js f = F −1 G
cs F f = Gs ∗ f,
where
Gs (x) = F −1 ((1 + ω 2 |ξ|2 )−s/2 )(x).
Now we give some properties of Gs (x) and show why this adjustment yields
exponential decay for Gs at infinity.
Proposition 7.6. Let s > 0.
R∞ |x|2 s−n
(a) Gs (x) = (4π)n/21Γ (s/2) 0 e−t e− 4t t 2 dtt .
(b) Gs (x) > 0, ∀x ∈ Rn ; and Gs (x) ∈ L1 (Rn ), precisely, Rn Gs (x)dx =
R
1.
(c) There exist two constants 0 < C(s, n), c(s, n) < ∞ such that
Gs (x) 6 C(s, n)e−|x|/2 , when |x| > 2,
and such that
1 Gs (x)
6 6 c(s, n), when |x| 6 2,
c(s, n) Hs (x)
where Hs is a function that satisfies
7.2. Bessel potentials - 175 -
s−n
|x| + 1 + O(|x|s−n+2 ), 0 < s < n,
2
Hs (x) = ln |x| + 1 + O(|x|2 ), s = n,
1 + O(|x|s−n ), s > n,
as |x| → 0.
0
(d) Gs (x) ∈ Lp (Rn ) for any 1 6 p 6 ∞ and s > n/p.
2
3
Or use (a) to show it. From part (a), we know Gs (x) > 0. Since Rn e−π|x| /t dx = tn/2 , by
R
∞ 1 ∞ ∞
dt dt dy
Z Z Z Z
− 2t − 2t
1 1
− 2t − 2t
e e 6 e + e dt = e−y + 2e−1/2
0 t 0 t 1 1/2 y
Z ∞
62 e−y dy + 2 6 4.
1/2
Next, suppose that |x| 6 2. Write Gs (x) = G1s (x) + G2s (x) + G3s (x), where
Z |x|2
1 1 −t − |x|
2
s−n dt
Gs (x) = n/2
e e 4t t 2 ,
(4π) Γ (s/2) 0 t
Z 4
2 1 −t − |x|
2
s−n dt
Gs (x) = e e 4t t 2 ,
(4π)n/2 Γ (s/2) |x|2 t
Z ∞
3 1 −t − |x|
2
s−n dt
Gs (x) = n/2
e e 4t t 2 .
(4π) Γ (s/2) 4 t
2
Since t|x|2 6 16 in G1s , we have e−t|x| = 1 + O(t|x|2 ) as |x| → 0; thus after
changing variables, we can write
Z 1
1 2 1 s−n dt
1
Gs (x) =|x| s−n
n/2
e−t|x| e− 4t t 2
(4π) Γ (s/2) 0 t
Z 1 Z 1
1 1 s−n dt
− 4t O(|x|s−n+2 ) 1 s−n
=|x| s−n
n/2
e t 2 + n/2
e− 4t t 2 dt
(4π) Γ (s/2) 0 t (4π) Γ (s/2) 0
s−n Z ∞
2 n−s−2
|x| dy 2 n−s−4
O(|x|s−n+2 ) ∞ −y s−n dy
Z
−y s−n
= e y 2 + e y 2 2
(4π)n/2 Γ (s/2) 1/4 y (4π)n/2 Γ (s/2) 1/4 y
=c1s,n |x|s−n + O(|x|s−n+2 ), as |x| → 0.
|x|2 |x|2
Since 0 6 4t 6 41 and 0 6 t 6 4 in G2s , we have e−17/4 6 e−t− 4t 6 1,
thus as |x| → 0, we obtain
s−n s−n+1
|x|
Z 4 n−s
− 2 n−s , s < n,
dt
2
G2s (x) ∼ t(s−n)/2 = 2 ln |x| , s = n,
|x| 2 t 2
s−n+1
s−n
, s > n.
|x|2
Finally, we have e−1/4 6 e− 4t 6 1 in G3s , which yields that G3s (x) is
bounded above and below by fixed positive constants. Combining the estimates
for Gjs (x), we obtain the desired conclusion.
(d) For p = 1 and so p0 = ∞, by part (c), we have kGs (x)k∞ 6 C for
s > n.
Next, we assume that 1 < p 6 ∞ and so 1 6 p0 < ∞. Again by part (c),
0 0
we have, for |x| > 2, that Gps 6 Ce−p |x|/2 , and then the integration over this
range |x| > 2 is clearly finite.
0
On the range |x| 6 2, it is clear that |x|62 Gps (x)dx 6 C for s > n. For the
R
0
case s = n and n 6= 1, we also have |x|62 Gps (x)dx 6 C by noticing that
R
7.2. Bessel potentials - 177 -
q Z 2 q
2 2
Z
ln dx = C ln rn−1 dr 6 C
|x|62 |x| 0 r
for
R any q 2> q0 since lim
ε
R 2 r→0 r ln(2/r) =R 0. For the case s = n = 1, we have
1
|x|62
(ln |x| ) dx = 2 0 (ln 2/r) dr = 4 0 (ln 1/r)q dr = 4Γ (q + 1) for q > 0
q
R1
by the formula 0 (ln 1/x)p−1 dx = Γ (p) for <p > 0. For the final case s < n,
R2 0
we have 0 r(s−n)p rn−1 dr 6 C if (s − n)p0 + n > 0, i.e., s > n/p.
Thus, we obtain kGs (x)kp0 6 C for any 1 6 p 6 ∞ and s > n/p, which
implies the desired result.
We also have a result analogues to that of Riesz potentials for the operator
Js .
Theorem 7.7. (a) For all 0 < s < ∞, the operator Js maps Lr (Rn ) into itself
with norm 1 for all 1 6 r 6 ∞.
(b) Let 0 < s < n and 1 < p < q < ∞ satisfy 1/q = 1/p − s/n. Then there
exists a constant Cn,s,p < ∞ such that for all f ∈ Lp (Rn ), we have
kJs f kq 6 Cn,s,p kf kp .
(c) If f ∈ L1 (Rn ), then m{x : |Js f (x)| > α} 6 (Cn,s α−1 kf k1 )q , for all
α > 0. That is, the mapping f → Js f is of weak type (1, q), with 1/q =
1 − s/n.
Proof. By Young’s inequality, we have kJs f kr = kGs ∗ f kr 6 kGs k1 kf kr =
kf kr . This proves the result (a).
In the special case 0 < s < n, we have, from the above proposition, that the
kernel Gs of Js satisfies
−n+s
|x| , |x| 6 2,
Gs (x) ∼ −|x|/2
e , |x| > 2.
Then, we can write
Z Z
−n+s −|y|/2
Js f (x) 6Cn,s |f (x − y)||y| dy + |f (x − y)|e dy
|y|62 |y|>2
Z
−|y|/2
6Cn,s Is (|f |)(x) + |f (x − y)|e dy .
Rn
We now use that the function e ∈ Lr for all 1 6 r 6 ∞, Young’s
−|y|/2
inequality and Theorem 7.4 to complete the proofs of (b) and (c).
The affinity between the two potentials is given precisely in the following
lemma.
Lemma 7.8. Let s > 0.
(i) There exists a finite measure µs on Rn such that its Fourier transform µbs
is given by
|ωξ|s
µbs (ξ) = .
(1 + |ωξ|2 )s/2
- 178 - 7. Sobolev Spaces
(d) Suppose 1 < p < ∞ and s > 1. Then f ∈ Hps (Rn ) if and only if
∂f
f ∈ Hps−1 (Rn ) and for each j, ∂xj
∈ Hps−1 (Rn ). Moreover, the two norms are
equivalent:
n
X ∂f
kf kHps ∼ kf kHps−1 + .
j=1
∂xj Hps−1
(e) Hpk (Rn ) = W k,p (Rn ), 1 < p < ∞, ∀k ∈ N.
n
X
(1 + |ωξ| )2 1/2
(1 + χ(ξj )|ξj |)−1 ∈ Mp , χ(ξj )|ξj |ξj−1 ∈ Mp , 1 < p < ∞.
j=1
Thus,
kf kHps =kJ s f kp = kF −1 (1 + |ωξ|2 )1/2 F J s−1 f kp
Xn
−1
6CkF (1 + χ(ξj )|ξj |)F J s−1 f kp
j=1
n
X ∂f
6Ckf kHps−1 + C kF −1 χ(ξj )|ξj |ξj−1 F J s−1 kp
j=1
∂xj
n
X ∂f
6Ckf kHps−1 + .
j=1
∂xj Hps−1
Proof. It is trivial for the case p = p1 since we also have s = s1 in this case.
Now, we assume that p < p1 . Since p11 = p1 − s−sn
1
, by part (b) of Theorem 7.7,
we get
kf kHps11 = kJ s1 f kp1 = kJ s1 −s J s f kp1 = kJs−s1 J s f kp1 6 CkJ s f kp = Ckf kHps .
Similarly, we can show the homogeneous case. Therefore, we complete the
proof.
0
Proof. It follows from the above theorem and the fact that (Lp )0 = Lp , if
1 6 p < ∞.
Finally, we give the connection between the homogeneous and the nonho-
mogeneous spaces.
7.3. Sobolev spaces - 183 -
References
Abe11. Helmut Abels. Short lecture notes: Interpolation theory and function spaces. May 2011.
Bec75. William Beckner. Inequalities in Fourier analysis on Rn . Proc. Nat. Acad. Sci. U.S.A., 72:638–
641, 1975.
BL76. Jöran Bergh and Jörgen Löfström. Interpolation spaces. An introduction. Springer-Verlag, Berlin,
1976. Grundlehren der Mathematischen Wissenschaften, No. 223.
Din07. Guanggui Ding. New talk of functional analysis. Science press, Beijing, 2007.
Duo01. Javier Duoandikoetxea. Fourier analysis, volume 29 of Graduate Studies in Mathematics. Amer-
ican Mathematical Society, Providence, RI, 2001. Translated and revised from the 1995 Spanish
original by David Cruz-Uribe.
Eva98. Lawrence C. Evans. Partial differential equations, volume 19 of Graduate Studies in Mathematics.
American Mathematical Society, Providence, RI, 1998.
GR. I. S. Gradshteyn and I. M. Ryzhik. Table of integrals, series, and products. Seventh edition.
Gra04. Loukas Grafakos. Classical and modern Fourier analysis. Pearson Education, Inc., Upper Saddle
River, NJ, 2004.
HLP88. G. H. Hardy, J. E. Littlewood, and G. Pólya. Inequalities. Cambridge Mathematical Library.
Cambridge University Press, Cambridge, 1988. Reprint of the 1952 edition.
HW64. Richard A. Hunt and Guido Weiss. The Marcinkiewicz interpolation theorem. Proc. Amer. Math.
Soc., 15:996–998, 1964.
Kri02. Erik Kristiansson. Decreasing rearrangement and Lorentz L(p, q) spaces. Master’s thesis, Luleå
University of Technology, 2002. http://bioinformatics.zool.gu.se/~erikkr/
MasterThesis.pdf.
Mia04. Changxing Miao. Harmonic analysis and its application in PDEs, volume 89 of Fundamental
series in mordon Mathematics. Science Press, Beijing, 2nd edition, 2004.
Rud87. Walter Rudin. Real and complex analysis. McGraw-Hill Book Co., New York, third edition, 1987.
SS03. Elias M. Stein and Rami Shakarchi. Complex analysis. Princeton Lectures in Analysis, II. Prince-
ton University Press, Princeton, NJ, 2003.
Ste70. Elias M. Stein. Singular integrals and differentiability properties of functions. Princeton Mathe-
matical Series, No. 30. Princeton University Press, Princeton, N.J., 1970.
Ste93. Elias M. Stein. Harmonic analysis: real-variable methods, orthogonality, and oscillatory inte-
grals, volume 43 of Princeton Mathematical Series. Princeton University Press, Princeton, NJ,
1993. With the assistance of Timothy S. Murphy, Monographs in Harmonic Analysis, III.
SW71. Elias M. Stein and Guido Weiss. Introduction to Fourier analysis on Euclidean spaces. Princeton
University Press, Princeton, N.J., 1971. Princeton Mathematical Series, No. 32.
WHHG11. Baoxiang Wang, Zhaohui Huo, Chengchun Hao, and Zihua Guo. Harmonic analysis method for
nonlinear evolution equations, volume I. World Scientific Publishing Co. Pte. Ltd., 2011.
Index