Morrey Spaces - 25 - 05 - 01 - 02 - 32 - 41
Morrey Spaces - 25 - 05 - 01 - 02 - 32 - 41
This series will maintain the highest editorial standards, publishing well-developed monographs
as well as research notes on new topics that are final, but not yet refined into a formal
monograph. The notes are meant to be a rapid means of publica!on for current material where
the style of exposi!on reflects a developing topic.
Glider Representaons
Frederik Caenepeel, Fred Van Oystaeyen
Summable Spaces and Their Duals, Matrix Transformaons and Geometric Properes
Feyzi Basar, Hemen Du!a
Morrey Spaces: Introducon and Applicaons to Integral Operators and PDE’s, Volume I
Yoshihiro Sawano, Giuseppe Di Fazio, Denny Ivanal Hakim
Morrey Spaces: Introducon and Applicaons to Integral Operators and PDE’s, Volume II
Yoshihiro Sawano, Giuseppe Di Fazio, Denny Ivanal Hakim
For more informa!on about this series please visit: h$ps://www.crcpress.com/Chapman--HallCRC-Monographs-and-Research-Notes-in-Mathema!cs/book-
series/CRCMONRESNOT
Morrey Spaces
Introduction and Applications to
Integral Operators and PDE’s,
Volume I
Yoshihiro Sawano
Chuo University
Giuseppe Di Fazio
University of Catania
Denny Ivanal Hakim
Bandung Institute of Technology
First edition published 2020
by CRC Press
6000 Broken Sound Parkway NW, Suite 300, Boca Raton, FL 33487-2742
Reasonable efforts have been made to publish reliable data and information, but the author
and publisher cannot assume responsibility for the validity of all materials or the conse-
quences of their use. The authors and publishers have attempted to trace the copyright
holders of all material reproduced in this publication and apologize to copyright holders if
permission to publish in this form has not been obtained. If any copyright material has not
been acknowledged please write and let us know so we may rectify in any future reprint.
Except as permitted under U.S. Copyright Law, no part of this book may be reprinted,
reproduced, transmitted, or utilized in any form by any electronic, mechanical, or other
means, now known or hereafter invented, including photocopying, microfilming, and record-
ing, or in any information storage or retrieval system, without written permission from the
publishers.
For permission to photocopy or use material electronically from this work, access
www.copyright.com or contact the Copyright Clearance Center, Inc. (CCC), 222 Rose-
wood Drive, Danvers, MA 01923, 978-750-8400. For works that are not available on CCC
please contact mpkbookspermissions@tandf.co.uk
Typeset in CMR
by Nova Techset Private Limited, Bengaluru & Chennai, India
Contents
Preface xi
Acknowledgement xv
1.1 Lp spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1.1 Measure space . . . . . . . . . . . . . . . . . . . . . . 2
1.1.2 Integration theorems . . . . . . . . . . . . . . . . . . 3
1.1.3 Fubini theorem and Lebesgue spaces . . . . . . . . . 4
1.1.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.2 Morrey spaces . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.2.1 Morrey norms . . . . . . . . . . . . . . . . . . . . . . 13
1.2.2 Examples of functions in Morrey spaces . . . . . . . 16
1.2.3 The role of the parameters . . . . . . . . . . . . . . . 23
1.2.4 Inclusions in Morrey spaces . . . . . . . . . . . . . . 24
1.2.5 Weak Morrey spaces . . . . . . . . . . . . . . . . . . 25
1.2.6 Morrey spaces and ball Banach function spaces . . . 27
1.2.7 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . 30
1.3 Local Morrey spaces, Bσ -spaces, Herz spaces and Herz–
Morrey spaces . . . . . . . . . . . . . . . . . . . . . . . . . . 31
1.3.1 Local Morrey spaces . . . . . . . . . . . . . . . . . . 32
1.3.2 Herz spaces and Herz–Morrey spaces . . . . . . . . . 34
1.3.3 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . 36
1.4 Distributions and Lorentz spaces . . . . . . . . . . . . . . . 37
1.4.1 Distribution function . . . . . . . . . . . . . . . . . . 37
1.4.2 Lorentz spaces . . . . . . . . . . . . . . . . . . . . . . 42
1.4.3 Hardy operators and Hardy’s inequality . . . . . . . 43
1.4.4 Inequalities for monotone functions and their
applications to Lorentz norms . . . . . . . . . . . . . 46
1.4.5 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . 48
1.5 Young functions and Orlicz spaces . . . . . . . . . . . . . . 49
1.5.1 Young functions . . . . . . . . . . . . . . . . . . . . . 50
1.5.2 Orlicz spaces . . . . . . . . . . . . . . . . . . . . . . . 57
v
vi Contents
1.5.3 Orlicz-averages . . . . . . . . . . . . . . . . . . . . . 60
1.5.4 Lebesgue spaces with a variable exponent . . . . . . 64
1.5.5 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . 64
1.6 Smoothness function spaces . . . . . . . . . . . . . . . . . . 64
1.6.1 Sobolev spaces . . . . . . . . . . . . . . . . . . . . . . 65
1.6.2 Hölder–Zygmund spaces . . . . . . . . . . . . . . . . 73
1.7 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
Bibliography 441
Index 475
Preface
xi
xii Preface
up in these chapters will play such a role. We can say that this note expands
the editorial [386].
Chapters 2 and 3 are of preliminary nature, which is interesting in its own
right. Chapter 2 collects some auxiliary observations in functional analysis.
Chapter 3 takes up polynomials. The facts that we prove in Chapters 2 and 3
will appear not so many times. The facts on functional analysis in this book
are quite fundamental and can be found in some textbooks such as [310].
Chapters 4, 5, 6 and 7 are also of preliminary nature. However, unlike
Chapters 2 and 3, we will frequently use what we prove in Chapters 4, 5, 6
and 7. This part largely overlaps the existing textbooks such as [146, 176, 156,
157, 415, 416, 382]. Although a thorough explanation of fundamental facts in
harmonic analysis makes the book thicker, we decided to include it to the
minimum so as to be self-contained.
When we consider some problems in mathematics, it is not sufficient to
be able to handle functions themselves. Usually, functions we will consider
are transformed by some mappings described by means of integrals. So, we
introduce some important operators here. Chapter 4 deals with the funda-
mental operators in many branches of mathematics. We consider the Hardy–
Littlewood maximal operator, singular integral operators fractional integral
operators and fractional maximal operators. Chapter 5 includes commutators
generated by these operators.
The theory of Morrey spaces can be staged on metric measure spaces; it is
enough to have a distance function and a Borel measure to develop a theory
of Morrey spaces. So, we work in general metric measure spaces in Chapter
6. As a special case of metric measure spaces, Chapter 7 deals with weighted
function spaces. Chapter 7 contains the solution of the A2 conjecture.
We embark on the study of Morrey spaces, especially operators acting on
Morrey spaces in Chapters 8, 9 and 10. In some sense, Morrey spaces are
introduced so as to cover what Lebesgue spaces cannot do. We seek to have a
better understanding of mathematical transforms by the use of Morrey spaces.
For example, the fractional integral operator is given by f 7→ | · |−n+α ∗ f .
Here 0 < α < n. The well-known theorem by Hardy–Littlewood–Sobolev
says that this transform maps Lp to Lq under a certain condition. Since the
fractional integral operator is the convolution operator, the Young inequality
will be a useful tool. However, the kernel | · |−n+α never belongs to Lp for
any 0 < p ≤ ∞. We expect Morrey spaces to explain why this convolution
mapping is bounded. In many cases we resort to the density argument in func-
tional analysis when we define operators acting on function spaces. Chapter
8 precisely discuss whether this is possible in Morrey spaces. For example,
L∞ n
c (R ), the space of all compactly supported essentially bounded functions,
is not dense in Morrey spaces. Many other plausible candidates fail to be dense
too. Another way of investigating the boundedness of operators in function
spaces is to employ duality. Chapter 9 invesitigates duality. Unfortunately, we
do not have any description of the dual space of Morrey spaces. However, Mor-
rey spaces are realized as the dual spaces of certain function spaces. Based
xiv Preface
xv
Notation in this book
B(x, r) ≡ {y ∈ Rn : kx − yk < r}
xvii
xviii Notation in this book
Numbers
(1) Let a ∈ R. Then write a+ ≡ max(a, 0) and a− = a ∧ 0 ≡ min(a, 0).
Correspondingly, given an R-valued function f , f+ and f− are functions
given by f+ (x) ≡ max(f (x), 0) and f− (x) ≡ min(f (x), 0), respectively.
(2) Let a, b ∈ R. Then write a ∧ b ≡ min(a, b). Correspondingly, given
R-valued functions f, g, f ∧ g are functions given by f ∧ g(x) ≡
min(f (x), g(x)).
(3) The constants C and c denote positive constants that may change from
one occurrence to another. Because the two constants c can be different,
the inequality 0 < 2c < c is by no means a contradiction. When we add
a subscript, for example, this means that the constant c depends upon
the parameter. It can happen that the constants with subscript differ
according to the above rule. In particular, we prefer to use cn for various
constants that depend on n, when we do not want to specify its precise
value.
(4) Let A, B ≥ 0. Then A . B and B & A mean that there exists a constant
C > 0 such that A ≤ CB, where C depends only on the parameters of
importance. The symbol A ∼ B means that A . B and B . A happen
simultaneously, while A ' B means that there exists a constant C > 0
such that A = CB.
(5) When we need to emphasize or keep in mind that the constant C depends
on the parameters α, β, γ etc:
(1) Instead of A . B, we write A .α,β,γ,... B.
Notation in this book xix
Function spaces
(1) We use · for functions; f = f (·).
(2) The function spaces are tacitly on Rn unless otherwise stated.
(3) Let X be a Banach space. We denote its norm by k · kX .
(4) Let Ω be an open set in Rn . Then Cc∞ (Ω) denotes the set of smooth
functions with compact support in Ω.
(5) Let 1 ≤ j ≤ n. The symbol xj denotes not only the j-th coordinate but
also the function x = (x1 , x2 , . . . , xn ) 7→ xj .
(6) The space L2 (Rn ) is the Hilbert space of square integrable functions on
Rn whose inner product is given by
Z
hf, gi = f (x)g(x)dx. (0.2)
Rn
(10) The space C(Rn ) denotes the set of all continuous functions on Rn .
(11) The space BC(Rn ) denotes the set of all bounded continuous functions
on Rn .
xx Notation in this book
(12) Occasionally we identify the value of functions with functions. For exam-
ple sin x denotes the function on R defined by x 7→ sin x.
(17) We denote the Lp (Rn )-norm by k·kLp . For other function spaces such as
Hölder’s continuous function space C γ (Rn ) of order γ > 0, we use k · kC γ
to stress the function spaces.
(18) When we consider function spaces on a domain Ω, we denote by C γ (Ω)
the Hölder continuous function space of order γ.
(19) A quasi-norm over a linear space X enjoys positivity, homogeneity,
and quasi-triangle inequality: for some α ≥ 1, kf + gkX ≤ α( kf kX +
kgkX ) (f, g ∈ X ). However, to simplify, we frequently omit the word
“quasi”. Likewise, we abbreviate the word “quasi-Banach space” to
Banach space.
The Kronecker delta function defined on a set X is given by δjk ≡
(20) (
1 (j = k),
for j, k ∈ X.
0 (j 6= k).
v + A ≡ {v + a : a ∈ A}, A + B ≡ { a + b : a ∈ A, b ∈ B}.
(24) When A and B are sets, A ⊂ B stands for the inclusion of sets. If,
in addition, both A and B are topological spaces, and if the natural
embedding mapping A → B is continuous, we write A ,→ B in the sense
of continuous embedding. If A and B are quasi-normed spaces with the
k
embedding constant k > 0, then A ,→ B.
Chapter 1
Banach function lattices
1.1 Lp spaces
Here we recall some fundamental properties on integration theory includ-
ing Lebesgue spaces. This will be useful when we consider Morrey spaces since
Lebesgue spaces are realized as a special case of Morrey spaces. Conversely
in this book it will be important what happens when Morrey spaces differ
from Lebesgue spaces. Section 1.1.1 deals with measure spaces. Section 1.1.2
handles integration theorems. Section 1.1.3 considers Fubini’s theorem. The-
orem 5 is a fundamental tool to express the Lebesgue norm of functions. The
Lebesgue space Lp initially appears in Section 1.1.3. Among other theorems,
1
2 Morrey Spaces
we postpone the proof of Theorem 12 till Section 4.1. Many of the results in
Section 1.1 are standard; see some textbooks on integration theory.
Let L0 (µ) be the space of all measurable functions. When we consider the
case of Euclidean space, L0 (Rn ) denotes the set of all Lebesgue measurable
functions. It may happen that the value of f ∈ L0 (Rn ) lies in C or [0, ∞]. We
need other terminologies for later consideration.
Definition 3. A regular measure µ means that µ has the inner and outer
regularities. To be precise, if µ is a regular Borel measure (Radon measure),
then for each Borel set E of finite measure, we have the inner regularity
µ(E) = inf{µ(U ) : U runs over all open sets containing E}. (1.1)
p−1
= p t λf,µ (t)dt.
0
When 0 < p < 1, the triangle inequality fails to be true. Instead, we use
the following p-convexity or p-triangle inequality:
Proposition 8 (p-convexity). Let (X, B, µ) be a measure space. Let 0 < p ≤ 1.
Then kf + gkLp (µ) p ≤ kf kLp (µ) p + kgkLp (µ) p for all f, g ∈ L0 (µ).
The proof, which uses Exercise 6, is left for the readers.
Duality is an important tool to estimate the size of functions. Let 1 ≤ p <
∞ and (X, B, µ) be a σ-finite measure space. Then we can describe the dual
0
of Lp (µ) by way of Lp (µ) as follows:
0
Theorem 9 (Duality Lp (µ)-Lp (µ)). Let 1 ≤ p < ∞ and (X, B, µ) be a
0
σ-finite measure space. Then Lp (µ)∗ = Lp (µ) with coincidence of norms.
0
Speaking precisely, let f ∈ Lp (µ) and define Ff ∈ Lp (µ)∗ by
Z
p
g ∈ L (µ) 7→ f (x)g(x)dµ(x) ∈ C.
X
p0 p ∗
Then f ∈ L (µ) 7→ Ff ∈ L (µ) is well defined and kf kLp0 (µ) = kFf kLp (µ)∗ .
Furthermore if F : Lp (µ) → K is a continuous linear functional, i.e., there
exists M > 0 such that |F (h)| ≤ M khkLp (µ) for all h ∈ Lp (µ), then F is
0
realized as F = Ff with some f ∈ Lp (µ).
Proof We do not recall the proof of this fundamental theorem. For the
proof, see [81, Example 3.5.4] and [319, Theorem 12.7], for example.
In addition to Theorem 9, it is important to write the Lebesgue norm in
the dual form.
Theorem 10. Let 1 ≤ p ≤ ∞ and (X, B, µ) be a measure space. Then for all
f ∈ L0 (µ), kf kLp (µ) = sup{kf · gkL1 (µ) : kgkLp0 (µ) = 1}.
Proof Again we do not recall the proof of this fundamental theorem. For
the proof see [258, Theorem 3.1.6], for example.
Local integrability will be important in this book. This will be made clearer
after we define Morrey spaces.
Definition 6 (Lploc (Rn )). Let 1 ≤ p ≤ ∞. The space Lploc (Rn ) collects all
f ∈ L0 (Rn ) such that f ∈ Lp (K) for each compact set K, or equivalently,
χB ·f ∈ Lp (Rn ) for each ball B inR Rn . The topology of Lploc (Rn ) is generated by
the functional f ∈ Lploc (Rn ) 7→ |f (y)|p dy, where K moves over all compact
K
sets.
Example 2. Let γ ∈ R and 0 < p < ∞. We define fγ (x) ≡ |x|γ for x ∈ Rn .
n
Since fγ ∈ Lp (Rn ) if and only if γ > − , it follows that f ∈ Lploc (Rn ) if and
p
n n
only if γ > − . To compensate for the failure in the case of γ = − , we
p p
will define the weak Lebesgue space WLp (Rn ) as the set of all g ∈ L0 (Rn ) for
which kgkWLp ≡ sup λkχ(λ,∞] (|g|)kLp is finite. See Definition 7 below.
λ>0
8 Morrey Spaces
for almost all x ∈ Rn . A point x for which (1.7) holds is called a Lebesgue
point of f .
We will prove Theorem 12 after we prove the boundedness of the Hardy-
Littlewood maximal operator:
n
We postpone its proof.
The function | · |− p is not a member in Lp (Rn ). Hence,
n
we are interested
in function spaces close to Lp (Rn ) which contain | · |− p . We will see that the
weak Lp (Rn )-space (the weak Lebesgue spaces, the Marcinkiewicz space) will
serve this purpose.
Definition 7 (Marcinkiewicz space, weak Lebesgue space).
(1) Let 0 < p < ∞. Then the weak Lp (µ)-space (or the Marcinkiewicz
space) WLp (µ) is the space of all functions f : X → C which satisfies
1 1
kf kWLp (µ) ≡ sup tλf (t) p = sup t p f ∗ (t) < ∞ if p < ∞.
t>0 t>0
∞ ∞
(2) Define WL (µ) ≡ L (µ).
(3) As usual if (X, B, µ) is the Lebesgue measure, then we write Lp (Rn ) =
Lp (µ) and k · kWLp = k · kWLp (µ) .
Example 3. Let γ ∈ R and 0 < p < ∞. Define fγ and gγ as in Example 1.
n
Then fγ ∈ WLp (Rn ) if and only if γ ≥ − and gγ ∈ WLp (Rn ) if and only if
p
Banach function lattices 9
n n
γ ≤ − , while fγ ∈ Lp (Rn ) if and only if γ > − and gγ ∈ Lp (Rn ) if and
p p
n n
only if γ < − . Note f− np + g− np = | · |− p ∈ WLp (Rn ).
p
We will investigate an embedding relation for decreasing functions. Here,
recall that a continuous (Borel) measure is a measure µ which does not charge
any point in the space; µ({a}) = 0 for any point a. Here and below w ∈ B(I)
means that w is Borel measurable. Here, we state our result in full generality
using continuous measures. We use Lemma 13 to prove Theorem 315. Here
and below denote by M↓ (0, ∞), the set of all non-negative decreasing functions
on (0, ∞).
Lemma 13. Let 0 < p ≤ p q < ∞, and let µ, ν be continuous Borel measures
q
µ(0, s)
on (0, ∞). If B ≡ sup p
p
< ∞, then kf kLq (µ) ≤ Bkf kLp (ν) for all
s>0 ν(0, s)
f ∈ M↓ (0, ∞).
Proof We may approximate f with simple decreasing functions. Let g ≡
K
X
f p . Then for the function g of the form g = ak χ(0,bk ) , with a1 , a2 , . . . , aK >
k=1
a0 = 0 and b0 = 0 < b1 < b2 < · · · < bK , we have
K Z ∞ pq
X q
kgk q ≤ ak p χ(0,bk ) (t)dµ(t)
Lp (µ) 0
k=1
for all f ∈ M↓ (0, ∞) and for all 0 < p ≤ 1, as long as v, w ∈ M+ (0, ∞) satisfy
! !
1 1
sup sup w(t) = sup sup w(t) < ∞,
a>0 v(0, a) t∈(0,a) a>0 kvkL1 (0,a) t∈(0,a)
Lemma 14. Let −∞ < a < b < ∞ and θ ∈ (0, 1). For all f ∈ M↑ (0, ∞), we
Z b !θ Z b
have g(t)dt ≤θ (t − a)θ−1 g(t)θ dt.
a a
Proof Assume for the time being that g is a function of the form:
N
X
g= λj χ(aj−1 ,aj ]
j=1
and
Z b N
X
θ (t − a)θ−1 g(t)θ dθ = λj θ ((aj − a)θ − (aj−1 − a)θ ).
a j=1
1.1.4 Exercises
Exercise 1. Let f : X → C be a complex-valued measurable function. Denote
h+ = max(h, 0) and h− = max(−h, 0) for a real-valued measurable function
h. Show that the following are equivalent:
(1) f ∈ L1 (µ).
(2) Re(f ) ∈ L1 (µ) and Im(f ) ∈ L1 (µ).
(3) Re(f )± ∈ L1 (µ) and Im(f )± ∈ L1 (µ).
Exercise 2.
(1) Choose functions integrable on [0, ∞) with respect to the Lebesgue mea-
sure among the functions listed below.
(2) Do the same thing replacing [0, ∞) with [1, ∞).
(3) Do the same thing replacing [0, ∞) with [0, 1).
(A) e−x (B) xα (α < −1) (C) x−1 (D) xα (−1 < α < 0) (E) 1
√ −x
e
(F ) xα (α > 0) (G) xe−x (H) x4 sin x (I) x e k − x
(J)
x−1
1 2
log(x + 1) 2 e−x
(K) 4 (L) p .
x3 |x − 2|
Exercise 3. Find the necessary and sufficient condition for a, b ∈ R to satisfy
|x|a
Z
b
dx < ∞.
Rn 1 + |x|
Exercise 4. Let (X, B, µ) be a finite measure space, and let f ∈ L0 (µ). Show
that lim kf kLp (µ) = kf kL∞ (µ) and that
p→∞
Z
1 1
lim µ(X)− p kf kLp = exp log |f (x)|dµ(x) .
p↓0 µ(µ) X
12 Morrey Spaces
2
Re log x
Exercise 5. Calculate the Lebesgue integral x dx.
e
Exercise 6.
(1) Show that (a + b)p ≤ ap + bp for all a, b > 0 and 0 < p ≤ 1.
(2) Prove Proposition 8.
Exercise 7. Show the following scaling nlaw in Lebesgue spaces Lp (Rn ) : For
all f ∈ Lp (Rn ) and t > 0 kf (t·)kLp = t− p kf kLp .
Exercise 8. Show that the Lebesgue space Lp ([0, 1]) with 0 < p < 1 is not a
normed space. What is the property of the normed space that fails?
Exercise 9. Suppose that h : R → C is a continuous function such that
lim h(t) = A exists.
|t|→∞
Z T
1
(1) Show that lim h(t)dt = A.
T →∞ 2T −T
Z T
1 T
(2) Show that lim log h(t)dt = A. Hint : Let M+ (R) denote
T →∞ 2T −T |t|
the set of all non-negative measurable functions defined on R. Justify
that we can assume that h ∈ M+ (R) is an even function. Then use
Z T
1 T
Z
1 T T
log h(t)dt = log h(t)dt
2T −T |t| T 0 t
!
1 T
Z Z T
ds
= h(t)dt
T 0 t s
1 T 1 s
Z Z
= h(t)dt ds,
T 0 s 0
where for the last equality we have used Fubini’s theorem.
Exercise 10. Let M+ (µ) stand for the set of all non-negative µ-measurable
functions. Let (X, B, µ) be a measure space, and let f ∈ M+ (µ). Define induc-
tively
1
f0 ≡ f, fj+1 ≡ fj − χ 1 (fj ).
j + 1 ( j+1 ,∞)
Then show that lim fj = 0. Consequently, we have a decomposition f =
j→∞
∞
1
P
j+1 χEj for some measurable set Ej .
j=0
expected to enjoy nicer properties than Lebesgue spaces. Here, we define Mor-
rey spaces in Section 1.2.1 and give some functions in Morrey spaces in Section
1.2.2.
Section 1.2.3 also considers some examples of functions in Morrey spaces
by paying attention to the parameters in Morrey spaces. Since Morrey spaces
have two parameters, we expect that Morrey spaces with different parameters
are connected with each other. We check that Morrey spaces are nested in
Section 1.2.4. Section 1.2.5 is a different nature from the previous sections.
We consider weak Morrey spaces as an analogue of Morrey spaces. We handle
weak function spaces. Weak Lebesgue spaces appeared in Example 2. We
expand the idea there. As shown, weak Morrey spaces are used to compensate
for the failure of the boundedness of operators. We return to Morrey spaces
in Section 1.2.6. There are many function spaces other than Morrey spaces.
Some of them are Banach lattices. The notion of Banach lattices is used to
cover many (solid) function spaces. One of the other useful tools to classify
function spaces is the use of Banach function spaces. However, as it turns out,
Morrey spaces are not Banach function spaces in general. Hence, we propose a
notion which substitutes for Banach function spaces. We actually define ball
Banach function spaces in Section 1.2.6. Once again a different function in
Morrey spaces appears there.
If there is need to stress the dimension we work in, we write kf kMpq (Rn ) instead
of kf kMpq . The (classical) Morrey space Mpq (Rn ) is the set of all f ∈ Lqloc (Rn )
for which the norm kf kMpq is finite.
If 0 < q ≤ p = ∞, then the Lebesgue differentiation theorem shows that
Mpq (Rn ) = L∞ (Rn ), so we exclude this case.
In other words, f ∈ Mpq (Rn ) if and only if there exists D ≥ 0 such that
1 1
kf kLq (B(x,r)) ≤ D|B(x, r)| p − q for all x ∈ Rn and r > 0. The minimal value
of D in this inequality is kf kMpq .
Since f · g ∈ Mpq (Rn ) for any f ∈ Mpq (Rn ) and g ∈ L∞ (Rn ), Mpq (Rn ) is
not a space of functions possessing any kind of common smoothness of any
order.
14 Morrey Spaces
Remark 1.
(1) Define (temporarily) the Morrey norms kf kB
Mp
0
q
and kf kQ
Mpq
of f ∈
0 n
L (R ) by
Z ! q1
n n
kf kB0
≡ sup r p−q q
|f (y)| dy (1.2)
Mp q
(x,r)∈Rn+1
+
B(x,r)
and
Z ! q1
1 1
kf kQ ≡ sup |Q(x, r)| p−q q
|f (y)| dy . (1.3)
Mpq
(x,r)∈Rn+1
+
Q(x,r)
Z q1
1 1
kf kD
Mpq
≡ sup |Q| p − q |f (y)|q dy .
Q∈D(Rn ) Q
Then kf kMpq ∼ kf kD
Mpq
for all f ∈ L0 (Rn ). Then kf kMpq ∼n,p,q
kf kB
Mp
0
q
∼n,p,q kf kQ
Mpq
∼n,p,q kf kD
Mpq
for all f ∈ L0 (Rn ). This means that
we can define the Morrey space Mpq (Rn ) by using the norms k · kB0
Mp q
,
k · kQ
Mpq
k · kD
Mpq
: The resulting space will be the same. We sometimes
identify all of them; the superindexes B, B0 , D are sometimes omitted.
See Exercise 13.
(2) In this book, all norms k · kMpq , k · kB
Mp
0
q
and k · kQ
Mpq
are used. Usually
there will be no confusion.
(3) There are some traditions of how to express Morrey norms. Some prefer
to use the notation:
Z ! q1
1
kf kLqλ ≡ sup |f (y)|q dy .
(x,r)∈Rn+1 rλ B(x,r)
+
Banach function lattices 15
Here, 0 < q < ∞ and 0 < λ < n. See the classical but important work
of Peetre [353]. By letting
q
λ≡ 1− n (1.4)
p
for all f, g ∈ L0 (Rn ). Note that (1.5) follows from the triangle inequality
immediately when 1 ≤ q < ∞. Meanwhile, when 0 < q < 1, (1.5) follows from
the inequality (a + b)q ≤ 2max(q−1,0) (aq + bq ) for all a, b ≥ 0.
As the following theorem shows, we can say that Lebesgue spaces are
realized as a special case of Morrey spaces.
16 Morrey Spaces
Theorem 16. For 0 < p < ∞, Mpp (Rn ) = Lp (Rn ) with coincidence of norms.
Proof Let f ∈ L0 (Rn ). We write down the definition of the norm kf kMpp ;
1 1
kf kMpp ≡ sup |B(x, r)| p − p kf kLp (B(x,r)) = sup kf kLp (B(x,r)) .
(x,r)∈Rn+1
+ (x,r)∈Rn+1
+
Then, we have
Z p1
p
kf kMpp ≤ |f (y)| dy = kf kLp . (1.6)
Rn
Meanwhile, by the monotone convergence theorem, for any x ∈ Rn , we have
kf kLp = lim kf kLp (B(x,m)) ≤ sup kf kLp (B(x,r)) = kf kMpp . (1.7)
m→∞ r>0
Z ! q1
1 1
kχB kMpq = sup |B(x, r)| p−q |χB (y)|q dy .
(x,r)∈Rn+1
+
B(x,r)
We can calculate and evaluate the integral precisely. The result is:
1 1 1 1
kχB kMpq = sup |B(x, r)| p − q |B(x, r) ∩ B| q ≥ |B| p = kχB kLp . (1.10)
(x,r)∈Rn+1
+
Z ! q1
1 1
kf kMpq ≡ kf kQ = sup |Q(x, r)| p−q q
|f (y)| dy .
Mpq
(x,r)∈Rn+1
+
Q(x,r)
Since f is supported on [0, 1]n , we may assume that Q(x, r) runs over
all cubes contained in [0, 1]n . We first note that the supremum is almost
attained by letting r = 2α: If we choose x ∈ Rn suitably, then we have
1 1 1
|Q(x, r)| p − q kf kLq (Q(x,r)) . |Q(r)| p ' kχαQj kLp = kf kLq . If 2r < α instead,
1 1 1
then we have |Q(x, r)| p − q kf kLq (Q(x,r)) ≤ |Q(x, r)| p < kf kLq no matter where
x is. Thus, to show the result, we have only to consider the cubes Q(x, r)
with α ≤ 2r ≤ 1. If Q(x, r) intersects QN N
j and Qk with 1 ≤ j < k ≤ N ,
n
1 R
Tε (x) ≡ x+ ε (x ∈ Rn ).
1+R 1+R
Let E0 ≡ [0, 1]n and Ej,0 ≡ [0, (1 + R)−j ]n .SSuppose that we have defined
E0 , E1 , E2 , . . . , Ej , j ∈ N. Define Ej+1 ≡ Tε (Ej ). Then we will show
ε∈{0,1}n
that
n
kχEj kMpq ∼ (1 + R)−j p = kχEj,0 kMpq = kχEj,0 kLp = kχEj kLq , (1.13)
Banach function lattices 19
kχEj kMpq
1 1 1
∼ sup{|Q| p − q |Q ∩ Ej | q : Q contains a connected component of Ej }.
where co(A) stands for the smallest convex set containing a set A. Then a
geometric observation shows that S ∗ engulfs k n connected component of Ej
for some 1 ≤ k ≤ 2j . Take an integer l so that 2l−1 ≤ k ≤ 2l . Then
|S ∗ ∩ Ej | = k n (1 + R)−jn , |S ∗ | ∼ (1 + R)−jn+ln .
20 Morrey Spaces
= 2 q (1 + R)l( p − q ) (1 + R)−j p
ln n n n
n
= (1 + R)−j p .
if q < r ≤ p.
The following is a sort of normalization: It can be arranged that Ej ⊂
[0, 2−j )n for each j ∈ N.
Hence, we can say that the Morrey norm k · kMpq reflects local regularity
of functions more precisely than the Lebesgue norm k · kLp .
A chain of equalities in (1.13) is the motivation of choosing R in (1.12).
Based on Example 9, we present another example. Although this is just a
matter of scaling, we sometimes find it convenient to use increasing sequences.
Example 11 (Self-similar increasing sequence for Mpq (Rn )). Let R > 1 solve
n n n
(1 + R) p − q 2 q = 1. In Example 9, if one defines Fj ≡ {x ∈ Rn : (1 + R)−j x ∈
Banach function lattices 21
Ej }, then {Fj }∞
j1 is an increasing sequence of sets and each Fj is made up of
a disjoint union of cubes of length 1. It is noteworthy that
j
( )
X
Fj ≡ y + Rk ak : {ak }jk=1 ∈ {0, 1}j , y ∈ [0, 1]n .
k=1
∞
X 1 a
(2) We set f0N ≡ N − p − q fN (· − N !) and fN ≡ ⊗n fN
0 0
= fN 0
⊗ fN ⊗···⊗
N =1
0
fN . Then f ∈ Mpq (Rn ).
We end this section with other functions defined on the real line.
N
X
Example 16. Let N ∈ N and ε ∈ (0, 1). Define f0N ≡ χ(l,l+l−ε ) . We
l=1
Z N +1
1 1 1 ε
have |[0, N + 1]| p − q kf0N kLq ([0,N +1]) ∼ N p − q , since f0N (t)dt ∼ N 1−ε .
0
1 ε
Consequently, kf0N kMpq (R) ∼ max(1, N p − q ).
Z ! q1
1 1 q
kf (t·)kMpq = sup |B(x, r)| p |f (y)| dy .
(x,r)∈Rn+1 |B(tx, tr)| B(tx,tr)
+
Since tn |B(x, r)| = |B(tx, tr)| for any balls B(x, r), we have the desired result.
As we discussed before, the parameter q seems to serve to describe the
local integrability of functions. We will clarify here the role of the parameter
q. Denote by L0c (Rn ) the set of all compactly supported measurable functions.
24 Morrey Spaces
Theorem 18 (The local integrability of q). Let 0 < q < r < p < ∞. Then
Mpq (Rn ) ∩ L0c (Rn ) \ Lrloc (Rn ) 6= ∅.
Proof We use the function f ≡ fN in Example 8. With α fixed, we have
!
1 kfN kLr 1 kfN kLr
lim inf log = lim inf log > 1.
N →∞ N kfN kLq N →∞ N kfN kMpq
∞
X 1
Thus, f ≡ fN ∈ Mpq (Rn ) ∩ L0c (Rn ) \ Lrloc (Rn ).
N 2 kfN kMpq
N =1
Thus, by inserting inequality (1.20) into (1.18) and (1.19), we obtain the
desired result.
The following theorem indicates that the embedding Mpq (Rn ) ,→ Mpq̃ (Rn )
is not dense unlike Lebesgue spaces over a set of finite measure.
Theorem 20. Let 1 < q < q̃ < p. Then Mpq̃ (Rn ) is not dense in Mpq (Rn ).
This result in particular shows that L∞ n p n
c (R ) is not dense in Mq (R ).
Proof Let F be the set defined in Example 11. The function χF belongs
to Mpq (Rn ) thanks to Example 11.
Banach function lattices 25
(2) Let fj ≡ χB(2j ) f , j ∈ N. Then kfj − fk kMpq = kf1 − f0 kMpq > 0 for all
k, j ∈ N with k < j.
The following example explains the gap between Mpq (Rn ) and WMpq (Rn ):
Example 17. Let 0 < q < p < ∞. Let F be the set in Example 11. Define
χF ((1 + R)−k ·)
f ≡ sup −k ·)k p
.
k∈N kχF ((1 + R) Mq
= 1,
so that f ∈ WMpq (Rn ).
1
We can refine the embedding Mpr (Rn ) ,→ Mpq (Rn ) for 0 < q < r ≤ p < ∞.
Lemma 23. Let 0 < q < r ≤ p < ∞. Then WMpr (Rn ) ,→ Mpq (Rn ).
Proof Let f ∈ WMpr (Rn ). Also let Q ∈ Q. We need to estimate
Z q1
1 1
|Q| p − q |f (x)|q dx .
Q
Theorem 24. Let 0 < q < p < ∞. Then WLp (Rn ) ,→ Mpq (Rn ).
Proof Let f ∈ WLp (Rn ). Then for any cube Q,
Z Z ∞
q q
|Q| p −1 |f (x)|q dx = |Q| p −1 qλq−1 |{x ∈ Q : |f (x)| > λ}|dλ
Q 0
q
Z ∞
1
≤ |Q| p qλq−1 min(1, (|Q|− p λ−1 kf kWLp )p )dλ
0
' (kf kWLp )q ,
as required.
We distinguish two cases 2a < b and 2a > b. For the latter case, there is at
most one j such that (a, b) ∩ (10j , 10j + aj ) 6= ∅. This implies
∞
X q
aj − 2 |(a, b) ∩ (10j , 10j + aj )| ≤ 1.
j=1
Banach function lattices 29
Consequently
q1
∞
1 1 X 1
kf kMpq ≤ 1 + sup (b − a) 2 − q j j
q |(a, b) ∩ (10 , 10 + aj )|
0<2a<b<∞ j=1
aj 2
q1
∞
1 1 1 1 X 1
≤ 1 + 2 q − 2 sup b 2 − q j j
q |(0, b) ∩ (10 , 10 + aj )|
0<b<∞ a
j=1 j
2
q1
∞
1 1 1 1 X
≤1+2 q−2 sup b 2−q χ(0,b+3) (10j + aj )
1≤b<∞ j=1
1 1 1 1 1
≤1+2 q−2 sup b 2−q (log10 (b + 3)) q < ∞.
1≤b<∞
Thus, we have
f ∈ Mpq (R) (1.22)
Meanwhile, a direct calculation shows that:
Z ∞
1
X
f (t)dt = (4−j ) 2 · 2j = ∞. (1.23)
E j=1
Example 22. When 1 ≤ q ≤ p < ∞, then Mpq (Rn ) is a ball Banach function
1 1
space as is seen from kχB kMpq = |B| p and kf kL1 (B) ≤ |B|1− p kf kMpq .
It is noteworthy that mixed Lebesgue spaces are not Banach function
spaces but that they are ball Banach function spaces [470]. See Exercise 19.
1.2.7 Exercises
Exercise 11. Let 0 < q < u ≤ p < ∞.
(1) [393, Proposition 2.3] Let Fj be the set in Example 11. Then show that
jn − jn
1+R q u
kχFj kMpu ≥ by considering the convex hull of Fj .
2
Mpr (Rn ) is not closed in Mpq (Rn )
S
(2) [393, Proposition 2.5] Show that
q<r≤p
by mimicking the proof of Theorem 20.
Mpr (Rn ) is not dense in Mpq (Rn )
S
(2) [393, Proposition 3.1] Show that
q<r≤p
by mimicking the proof of Theorem 20.
Exercise 12. [265, Lemma 2] Let 0 < q ≤ p < ∞. Let Ej be the set con-
q
structed in Example 10, so that |Ej | = 2−jn [2jn(1− p ) ] and that each Ej con-
tain [0, 2−j )n . Then for a positive sequence {aj }∞
j=1 , the following are equiv-
alent:
∞
aj χEj ∈ Lq (Rn ).
P
(1)
j=1
∞ jn
(2− aj )q < ∞.
P
(2) p
j=1
Exercise 13. Let f ∈ L0 (Rn ), and let 0 < q ≤ p < ∞. Recall that the
Morrey norm kf kMpq = kf kB
Mpq
is defined by Definition 8. Define
Z ! q1
1 1
kf kQ ≡ sup |Q(x, r)| p−q q
|f (y)| dy . (1.24)
Mpq
(x,r)∈Rn+1
+
Q(x,r)
for f ∈ L0 (Rn ). The local Morrey space LMpq (Rn ) is the set of all f ∈ L0 (Rn )
for which kf kLMpq is finite. Likewise, define the weak local Morrey space
WLMpq (Rn ) as the set of all f ∈ L0 (Rn ) for which the norm kf kWLMpq =
sup λkχ(λ,∞] (|f |)kMpq is finite.
λ>0
(3) If 0 < p < q ≤ ∞ or −∞ < p < 0 < q ≤ ∞, then LMpq (Rn ) = {0},
which is similar to Mpq (Rn ) = {0}.
We collect some elementary inclusion properties of local Morrey spaces.
Proposition 25.
(1) Let 0 < p < ∞. Then LMpp (Rn ) = Lp (Rn ) with coincidence of norms.
1
(2) If 0 < q1 ≤ q2 ≤ p < ∞, then LMpq2 (Rn ) ,→ LMpq1 (Rn ).
(3) Let 1 ≤ q ≤ p < ∞. Then WLMpq (Rn ) and LMpq (Rn ) are complete. We
have the following inclusions:
1 1 1
WMpq (Rn ) ,→ WLMpq (Rn ), Mpq (Rn ) ,→ LMpq (Rn ) ,→ WLMpq (Rn ).
Banach function lattices 33
n
(4) Let 1 ≤ q < p < ∞. Then | · |− p ∈ LMpq (Rn )
Proof Overall, the proofs are careful reexaminations of the case of Morrey
spaces.
(1) Reexamine the proof of Theorem 16.
(2) Reexamine the proof of Theorem 19.
(3) Completeness of these spaces is clear because these spaces are continu-
ously embedded into WLqloc (Rn ). Embeddings follow from the definition
of the norms.
1
(4) This is a simple consequence of the embedding Mpq (Rn ) ,→ LMpq (Rn )
and Example 6.
Some people call local Morrey spaces Bσ spaces. This simply reformulates
the definition.
Definition 16. Let E be a ball quasi-Banach function space. Then, for σ ∈
[0, ∞), we define the E-based nonhomogeneous Bσ space Bσ (E)(Rn ) and the
E-based homogeneous Bσ space Ḃσ (E)(Rn ) as the sets of f ∈ L0 (Rn ) for which
kf kBσ (E) ≡ sup r−σ kf χQ(r) kE < ∞ and kf kḂσ (E) ≡ sup r−σ kf χQ(r) kE < ∞,
r≥1 r>0
respectively.
Example 24. As the candidate of E(Rn ), we can list the following function
spaces:
(1) E = Lp (Rn ) with 1 ≤ p ≤ ∞. In this case the corresponding func-
tion spaces Bσ (E)(Rn ) and Ḃσ (E)(Rn ) are written Bσ (Lp )(Rn ) and
Ḃσ (Lp )(Rn ) and are called a homogeneous Bσ -Lp space and a nonho-
mogeneous Bσ -Lp space, respectively. It is noteworthy that
Ḃ nq − np (Lq )(Rn ) ≈ LMpq (Rn )
with equivalence of norms for any 1 < q ≤ p < ∞. We define
Z ! p1
1
kf kḂ p,λ ≡ sup r−λ |f (x)|p dx
r>0 |Q(r)| Q(r)
and ! p1
Z
1
kf kB p,λ ≡ sup r−λ |f (x)|p dx .
r≥1 |Q(r)| Q(r)
∼ kf kK̇ α .
p∞
is finite.
Example 27. Let 0 < p, q ≤ ∞ and α, λ ∈ R. Let f (x) ≡ |x|β χRn \B(1) (x),
n
x ∈ Rn with β ∈ Rn . Then 2jα kχCj f kLp ∼ 2j(α+β+ p ) .
n
(1) If α + β + p
α,λ
> 0, then f ∈ M Kpq (Rn ) if and only if λ ≥ α + β + np .
n α,λ
(2) If α + β + p = 0, then f ∈ M Kpq (Rn ) if and only if λ > 0.
n α,λ
(3) If α + β + p ≤ 0, then f ∈ M Kpq (Rn ) if and only if λ ≥ 0.
1.3.3 Exercises
Exercise 20. Let 1 ≤ p ≤ ∞ and σ ≥ 0.
(1) Show that χQ(1) ∈ Bσ (Lp )(Rn ) for all σ ≥ 0 and 1 ≤ p ≤ ∞.
n
(2) Show that χQ(1) ∈ Ḃσ (Lp )(Rn ) if and only if σ ≤ .
p
(3) Show that χ[1,2]n ∈ Bσ (Lp )(Rn ) for all σ ≥ 0 and 1 ≤ p ≤ ∞.
(4) Show that χ[1,2]n ∈ Ḃσ (Lp )(Rn ) for all σ ≥ 0 and 1 ≤ p ≤ ∞.
Exercise 21. Let n
0 < q ≤ p < ∞ and f ∈ L0 (Rn ). Then show that
−
kf (t·)kLMpq = t kf (·)kLMpq for all t > 0.
p
Exercise 22 (Diversity of local Morrey spaces). Let 0 < q < r ≤ p < ∞. Let
F be the set defined in Example 11. Then using χF , show that LMpq (Rn ) and
LMpr (Rn ) are different.
n
Exercise 23. Let 0 < q < p < ∞, and let f (x) = |x|− p .
(1) Show that kf kLMpq = kf kMpq .
(2) Show that dist(f, L∞ n
c (R )) = kf kLMq .
p
α,λ
Exercise 24. Let 0 < p, q ≤ ∞ and α, λ ∈ R. Show that M Kpq (Rn ) is a
quasi-ball Banach function space.
n
Exercise 25. Let 0 < q < p < ∞, and let f (x) = |x|− p .
Banach function lattices 37
(1) Show that there exists an increasing sequence {Rj }∞ j=1 of positive real
numbers such that R1 = 1, kf χB(Rj+1 )\B(Rj ) kLMpq ≥ (1 − 2−j )kf kLMpq .
(2) Define
∞
X
Φ : {aj }∞ ∞
j=1 ∈ ` (N) 7→ aj f χB(Rj+1 )\B(Rj ) ∈ LMpq (Rn ).
j=1
Example 28. Consider the Euclidean space Rn here. Let γ 6= 0. Let fγ and
gγ be as in Example 1. For each t > 0, we have:
N
X −1
f ∗ = αN χ(0,µ(EN )) + αk χ[µ(EN ∪EN −1 ∪···∪Ek+1 ),µ(EN ∪EN −1 ∪···∪Ek )) .
k=1
∞
[
{x ∈ Rn : |f (x)| > t0 } = {x ∈ Rn : |f (x)| > tk }
k=1
and
λf (t1 ) < a1 and λf (t2 ) < a2 . This means that µ{|f + g| > t1 + t2 } > a1 + a2
and that µ{|f | > t1 } ≤ a1 , µ{|f | > t2 } ≤ a2 . This is a contradiction.
The distribution function is used to define Lorentz spaces. The definition
is as follows:
It is also clear from the definition of distribution functions that f ∗ (t) ≤ g ∗ (t)
if f, g ∈ L0 (Rn ) satisfy |f | ≤ |g|.
We consider a couple of examples.
Example 30. Let γ < 0. Let fγ and gγ be as in Example 1. Then
v γ/n
n
fγ∗ (t) = χ[0,vn ) (t) (0 < t < ∞)
t
and γ/n
t
gγ∗ (t) = +1 (0 < t < ∞).
vn
Example 31. Since λχ[0,1] = χ[0,1) ,
(
[1, ∞) 0 < t < 1,
{s ∈ [0, ∞) : λχ[0,1] (s) ≤ t} =
[0, ∞) t ≥ 1.
Lemma 29. Let f ∈ L0 (µ). Then for any t > 0, λf (f ∗ (t)) ≤ t, namely,
µ{x ∈ X : |f (x)| > f ∗ (t)} ≤ t.
40 Morrey Spaces
Proof We may assume that f ∗ (t) < ∞; otherwise the conclusion follows
readily from 0 = λf (∞) ≤ t. Suppose f ∗ (t) < ∞. Then, since f ∗ (t) = inf{s ∈
[0, ∞) : λf (s) ≤ t}, we have λf (s) ≤ t for all s > f ∗ (t). Therefore, by
Theorem 27, λf (f ∗ (t)) = lim
∗
λf (s) ≤ t. Thus, the proof is complete.
s↓f (t)
Proof It is by the monotonicity clear that f ∗ (t) ≥ lim (fj )∗ (t). So, we
j→∞
need to establish that f ∗ (t) ≤ lim (fj )∗ (t). Assume otherwise: a = f ∗ (t) −
j→∞
lim (fj )∗ (t) > 0. Then f ∗ (t) − a ∈
/ {s ∈ [0, ∞) : λf (s) ≤ t}, or equivalently,
j→∞
λf (f ∗ (t) − a) > t. Consequently, there exists b > 0 such that λf (f ∗ (t) − a) >
b + t. Thus, λfj (f ∗ (t) − a) > b + t by the monotone convergence theorem
as long as j 1. Consequently, λfj ((fj )∗ (t)) > b + t for such j. This is a
contradiction to Lemma 29.
Example 33. Let f ∈ M↓ (0, ∞). Then we can approximate f almost every-
where with simple decreasing functions. Since fj is a simple decreasing func-
tion, then (fj )∗ = fj as is seen from Example 32. Thanks to Proposition 31,
f ∗ = f almost everywhere.
In a special case like Example 30, we can regard λf as the inverse of f ∗ .
Proposition 32. Let f ∈ L0 (µ). Assume that λf is positive, strictly decreas-
ing and continuous. Then;
(1) λ−1 ∗
f (t) = f (t) for all t > 0,
Proof Simply consider the case where f is a simple function. In this case,
the proof is direct.
As an application of Theorem 33, we justify the name of WLp (µ):
1
Lemma 34. Let 0 < p < ∞. Then Lp (µ) ,→ WLp (µ).
Proof Let f ∈ Lp (µ). According to Theorem 33, we have
Z ∞ p1 Z t p1
∗ p ∗ p
kf kLp = f (s) ds ≥ f (s) ds .
0 0
1
Hence, kf kLp ≥ sup t f ∗ (t) = kf kWLp . Thus, the proof is complete.
p
t>0
which implies f ∗ (t1 ) + g ∗ (t2 ) ≤ (f + g)∗ (t1 + t2 ). Thus, the proof is complete.
0 t
is finite.
(2) The space Lq,∞ (µ) stands for L∞ (µ) for any 0 < q ≤ ∞.
(3) If 0 < p < ∞, then the Lorentz space Lp,∞ (µ) denotes the weak Lp (µ)-
space: Lp,∞ (µ) = WLp (µ).
As usual if (X, B, µ) is the Lebesgue measure, then we write Lp,q (Rn ) =
Lp,q (µ) and k · kWLp,q = k · kWLp,q (µ) .
Theorem 33 shows that Lp,p (µ) = Lp (µ) with coincidence of norms. Indeed,
Z ∞ p1 Z ∞ p1
1
∗ p dt ∗ p
kf kLp,p = (t f (t))
p = f (t) dt = kf kLp .
0 t 0
When we consider integral inequalities in (0, ∞), we frequently face the fol-
lowing norm:
Banach function lattices 43
Proof We calculate
r1
k ·−λ0 r−1 |ϕ|r kΦθ(λ −λ )r−1, q (0,∞)
1 0 r
Z ∞ q1
−θ(λ1 −λ0 )r−λ0 r rq q dη
= (η ) ϕ(η)
0 η
Z ∞ q1
dη
= η −θ(λ1 −λ0 )q−λ0 q ϕ(η)q
0 η
Z ∞ q1
dη
= η −λq ϕ(η)q
0 η
= kϕkΦλ,q (0,∞) .
We now will consider the space Φ↑λ,q (0, ∞) ≡ Φλ,q (0, ∞) ∩ M↑ (0, ∞). We
define kf kΦλ,q (0,∞) ≡ kf kΦλ,q (0,∞) for f ∈ Φ↑λ,q (0, ∞). One of the promi-
nent properties in the scale {Φλ,q (0, ∞)}q>0 is the monotonicity in the second
parameter q.
Lemma 37. Let 0 < q0 < q1 ≤ ∞, and let 0 < λ < ∞. Then Φ↑λ,q0 (0, ∞) ⊂
1 1
Φ↑λ,q1 (0, ∞). More quantitatively, kϕkΦλ,q1 (0,∞) ≤ (λq0 ) q0 − q1 kϕkΦλ,q0 (0,∞) for
all ϕ ∈ Φ↑λ,q0 (0, ∞).
Proof First let q1 = ∞. Then
Z ∞ q1
−λ
1 dτ 0
kϕkΦλ,∞ (0,∞) = sup t ϕ(t) = (λq0 ) q0
sup ϕ(t).
t>0 t>0 t τ 1+λq0
as long as the integral makes sense. Both operators are called Hardy operators.
min(1, t) max(1 − t, 0)
Example 39. Since Hχ(0,1) (t) = and Hχ(0,1) (t) =
t t
for t > 0. So H and H fail to be L1 (0, ∞)-bounded.
Since µ > −1, we obtain the desired result once we integrate the right-
hand side against τ .
46 Morrey Spaces
We estimate
Z ρ 1 ∞ Z µ(ρ)θ qj
µ(t) q X 1
dt ≤ µ(ρ)θj ds. (1.2)
0 t j=0 µ(ρ)θ q(j+1) ν(s)µ0 (ν(s))
1.4.5 Exercises
Exercise 26. Let 0 < p < ∞ and 0 < q ≤ ∞.
(1) Argue similarly to Example 28(3) to show that χ∗B(R) = χ[0,|B(R)|) .
Exercise 29. Let p and q be as in Definition 20. Find the necessary and
sufficient condition on the parameter γ ∈ R ensuring that fγ ∈ Lp,q (Rn ) and
gγ ∈ Lp,q (Rn ).
Exercise 30. By using the set E defined by (6.25), prove that L4,∞ (R) is
strictly wider than M42 (R). Here, the norm of L4,∞ (R) is given by
q
kf kL4,∞ ≡ sup λ 4 λf (λ)
λ>0
for f ∈ L0 (R).
Exercise 31.
(1) For g ∈ Φµ,∞ (0, ∞) and µ > −1, show that kHgkΦµ,∞ (0,∞) ≤
Z t
kgkΦµ,∞ (0,∞) 1
using sµ ds = .
µ+1 0 µ+1
Z ∞
1
(2) For g ∈ Φ−µ,∞ (0, ∞) and µ > 1, using s−µ ds = , show that
t µ−1
kgkΦ−µ,∞ (0,∞)
kHgkΦ−µ,∞ (0,∞) ≤ .
µ−1
kgkΦµ,1 (0,∞)
(3) For g ∈ Φµ,1 (0, ∞) and µ > −1, show that kHgkΦµ,1 (0,∞) ≤
µ+1
using Fubini’s theorem.
Banach function lattices 49
kgkΦ−µ,1 (0,∞)
(4) Show that kHgkΦ−µ,1 (0,∞) ≤ for g ∈ Φ−µ,∞ (0, ∞) and
µ−1
µ > 1 using Fubini’s theorem.
(4) Show that there is no measurable function f on (0, ∞) such that 0 <
kHf kLp (0,∞) = p0 kf kLp (0,∞) < ∞.
Exercise 33. [156, Exercise 1.4.13] Let 0 < p < ∞, and let f ∈ L0 (Rn ). Use
Lemma 29 to show that f ∈ Lp,1 (Rn ) if and only if there exist a decreas-
ing sequence {aj }∞ ∞ ∞ n
j=−∞ ⊂ [0, ∞), {ϕj }j=−∞ ⊂ L (R ) and a sequence
{Ej }∞ j
j=−∞ of measurable sets safisfying |ϕj | ≤ aj χEj and |Ej | ≤ 2 for each
∞
P ∞
P j
j ∈ Z such that f = ϕj and 2 p aj < ∞.
j=−∞ j=−∞
(7) The function Φ(t) = t exp(−t−1 ), t > 0 has the limit Φ(+0) = lim Φ(t) =
t↓0
0. Consequently, if we extend Φ continuously to [0, ∞), then Φ is a Young
function since Φ00 (t) = t−4 Φ(t) for all t > 0.
Let us start with the following simple estimates for Young functions.
Proposition 41. For a Young function Φ : [0, ∞) → [0, ∞) and t ≥ 0,
θΦ(t) ≥ Φ(θt) if 0 < θ < 1, or equivalently, θΦ(t) ≤ Φ(θt) if 1 < θ < ∞.
Proof Since Φ is convex and Φ(0) = 0, for 0 < θ < 1,
As a corollary, we can show that the tangent of the line connecting (0, 0)
and (a, Φ(a)) is increasing in a ≥ 0.
Lemma 42. Let Φ : [0, ∞) → [0, ∞) be a Young function. Then mapping
t ∈ (0, ∞) 7→ t−1 Φ(t) ∈ [0, ∞) is increasing.
Proof This can be derived easily from Proposition 41.
Since Young functions are locally absolutely continuous, they have density:
Any Young function can be expressed for some ϕ ∈ M↑ (0, ∞) as
Z t
Φ(t) = ϕ(s)ds (t ≥ 0).
0
We call this equality the canonical representation of Φ. Next, we will treat the
information on the density function ϕ of a Young function.
Lemma 43. Let Φ : [0, ∞) → [0, ∞) be a Young function that is expressed as
Z t
Φ(t) = ϕ(s)ds (t ≥ 0).
0
1 t 1 t
Z Z
Φ(t)
= ϕ(s)ds ≤ ϕ(t)ds = ϕ(t)
t t 0 t 0
and
Z 2t Z 2t Z 2t
1 1 1 Φ(2t)
ϕ(t) = ϕ(t)dt ≤ ϕ(s)ds ≤ ϕ(s)ds = ,
t t t t t 0 t
52 Morrey Spaces
(2) Note that [0, x] ⊃ ϕ−1 ([0, v]). Thus, we can argue as in (1).
We have the following expression of Φ∗ :
Banach function lattices 53
for all s, t ≥ 0 and that for each s ≥ 0, there exists t ≥ 0 for which equality
in (1.5) holds.
If st = 0, then (1.5) is trivial. Therefore, we may assume that s, t > 0.
Inserting (1.4) into the canonical representation, we have
Z s Z ∞
Φ∗ (s) = χϕ−1 ([0,v]) (x)dx dv
Z0Z 0
Observe that
To show that equality holds for some t, we set t ≡ ϕ∗ (s). We obtain a kind
of reverse inclusion in (1.6). We aim to show
and that
{(x, v) ∈ (0, ∞) × (0, ∞) : 0 < x < t, 0 < v < ϕ(x)} ⊂ [0, t] × [0, s].
Let 0 < v < s satisfy ϕ(x) < v. Then t = ϕ∗ (s) ≥ ϕ∗ (v) ≥ x ≥ 0 thanks to
Lemma 44. Meanwhile, let 0 ≤ x < t and 0 ≤ v < ϕ(x). Since x < t = ϕ∗ (s),
we have ϕ(x) ≤ s thanks to Lemma 44 once again. Therefore v ≤ s.
54 Morrey Spaces
st ≤ Φ∗ (2s) (1.9)
λ ≤ ts ≤ 2λ. (1.10)
Banach function lattices 55
Proof
(1) Recall that t ∈ (0, ∞) 7→ t−1 Φ(t) ∈ (0, ∞) is increasing in t thanks to
Lemma 42. Thus,
∗ Φ(t) Φ(t) Φ(s) Φ(t) Φ(s)
Φ = sup s − ≤ t sup − ≤ Φ(t).
t 0<s<t t s 0<s<t t s
The right inequality is easier to prove. We have only to use Theorem 45:
2Φ(t) 2Φ(t) Φ(s)
Φ∗ = sup s − ≥ Φ(t).
t s>0 t s
Thus, (1.7) is proven.
(2) Let st ≤ Φ(t). Since t−1 Φ(t) is increasing in t > 0 again thanks to
Lemma 42, we have
Φ∗ (s)
Φ(u)
= sup u −
s u>0 s
u Φ(u)
≤ t sup −
u>0 t Φ(t)
u Φ(u)
= t sup −
0<u<t t Φ(t)
≤ t.
Definition 24 (∆2 , ∇2 ).
(1) We claim that θ ≡ τ + 1 − q̃ does the job. For all 0 < t ≤ s < ∞, we
choose a number k ∈ N0 such that 2−k−1 s ≤ t ≤ 2−k s. By using the ∇2
condition, we obtain
s 1+log2 D Φ(t)
Φ(s) ≥ Φ 2k t ≥ (2D)k Φ(t) = 2(1+log2 D)k Φ(t) ≥
,
t 21+log2 D
or equivalently,
Φ(t) Φ(s)
θ
≤ 2θ θ .
t s
(2) We claim that θ† ≡ τ does the job. Since q̃ ∈ (0, 1), we have
Z r Z r
Φ(r) r 1
Z
Φ(t) Φ(t) dt Φ(r)
τ +1
dt = θ q̃
. θ q̃
dt ∼ τ
0 t 0 t t r 0 t r
thanks to (1). Thus, the proof is complete.
After writing the norm in this way, we generalize the Lp (µ)-norm. We define
the Orlicz space LΦ (µ) as follows:
58 Morrey Spaces
for f ∈ L0 (µ). The Orlicz space LΦ (µ) over X is the set of all f ∈ L0 (µ) for
which kf kLΦ (µ) is finite.
Here and below for the sake of simplicity, we assume that Φ is nice.
Example 45.
(1) If Φ(r) = rp , 1 ≤ p < ∞, then LΦ (µ) = Lp (µ) with coincidence of
norms.
(2) Let p > 1, q > 0 be parameters. The most important example of Orlicz
spaces is generated by a function satisfying Φ(t) ∼ tp [log(3 + t)]q for all
t ≥ 0. Denote by Lp logq L(µ) such a function space. The space L log L(µ)
is occasionally called the Hardy–Littlewood space. If p = 1, then write
L logq L(µ) instead of Lp logq L(µ).
(3) If Φ(t) = et − 1 for t ≥ 0, then we write exp(µ) instead of LΦ (µ).
(4) Let E be a measurable set, and let Φ : [0, ∞) → [0, ∞) be a Young
1
function. Then kχE kLΦ (µ) = −1 . In fact, assuming that
Φ (µ(E)−1 )
1
µ(E) < ∞, we see that λ ≡ kχE kLΦ (µ) solves µ(E)Φ = 1.
λ
As the following theorem shows, k · kLΦ (µ) is a complete norm.
From the definition of the set defining the norm kf kLΦ (µ) , we deduce
|f (x)|
Z
1
kf kLΦ (µ) + ε ∈ λ>0 : Φ dµ(x) ≤ 1 .
2 X λ
Banach function lattices 59
Namely, we have
|f (x)|
Z
Φ dµ(x) ≤ 1.
X kf kLΦ (µ) + 2−1 ε
1 1 β
Set α ≡ kf kLΦ (µ) + ε, β ≡ kgkLΦ (µ) + ε and θ ≡ .
2 2 α+β
Since Φ is convex, we have a pointwise estimate: for all x ∈ X,
|f (x) + g(x)| |f (x)| |g(x)|
Φ = Φ (1 − θ) +θ
α+β α β
|f (x)| |g(x)|
≤ (1 − θ) Φ +θΦ . (1.13)
α β
Proof By the Fatou lemma, we may assume that f ∈ L∞ (µ) and that
µ{f 6= 0} < ∞. We may assume that kf kLΦ = 1 by normalization. By con-
|f (x)|
Z
sidering two cases; λ ≤ 1 and λ ≥ 1, we obtain 1 ≤ λ + λ Φ dx
X λ
for all λ > 0. The opposite inequality can be obtained by taking λ ≡ 1.
We next prove the dual inequality for Orlicz spaces.
Theorem 52. Let Φ : [0, ∞) → [0, ∞) be a Young function, and let Φ∗
∗
be its conjugate. Then for all f ∈ LΦ (µ) and g ∈ LΦ (µ), kf · gkL1 (µ) ≤
2kf kLΦ (µ) kgkLΦ∗ (µ) .
60 Morrey Spaces
Proof We may assume that kf kLΦ (µ) = kgkLΦ∗ (µ) = 1. In this case, by
integrating |f (x)g(x)| ≤ Φ(|f (x)|) + Φ∗ (|g(x)|) over X we obtain the desired
result.
We end Section 1.5.2 with a duality result on Orlicz spaces.
Theorem 53. Let Φ ∈ ∆2 . Then the dual space of LΦ (µ) is isomorphic to
∗
LΦ (µ) in the following sense:
∗ ∗
(1) For all f ∈ LΦ (µ) and g ∈ LΦ (µ), f ·g ∈ L1 (µ), Φ
R so that any g ∈ L (µ)
Φ
defines a continuous functional h ∈ L (µ) 7→ h(x)g(x)dµ(x) ∈ C.
X
Φ
(2) Conversely any continuous functional on L (µ) is realized in this way.
Proof From Theorem 52, (1) follows. We prove (2). Let L : LΦ (µ) → C
be continuous. Since X is σ-finite, we may assume µ(X) < ∞ by a routine
truncation procedure.
Let f ∈ LΦ (µ) with kf kLΦ (µ) = 1. We remark that g ≡ χ{f 6=0} |f |−1 Φ(|f |)
satisfies
Z Z
kg · f kL1 (µ) = 1, Φ∗ (g(x))dµ(x) ≤ Φ(|f (x)|)dµ(x) = 1
X X
thanks to Lemma 47(1). Thus, for any f ∈ LΦ (µ) with kf kLΦ (µ) = 1, there
exists a simple function g satisfying kgkLΦ∗ (µ) ≤ 1 such that kf · gkL1 (µ) = 1.
Since Φ ∈ ∆2 , we can find a number p > 0 such that Φ(t) . tp for t ≥ 1.
Thus, since µ(X) < ∞, the mapping f ∈ Lp (µ) 7→ L(f ) ∈ C is a continuous
0
linear functional, so that there exists g ∈ Lp (µ) such that
Z
L(f ) = f (x)g(x)dµ(x)
X
p
for all f ∈ L (µ). Consequently, by a simple truncation procedure once again,
we have
kf · gkL1 (µ) ≤ kLkLΦ (µ)→C kf kLΦ (µ)
∗
for all f ∈ LΦ (µ). Once we show that g ∈ LΦ (µ) with the estimate
kgkLΦ∗ (µ) . kLkLΦ (µ)→C , we see that g realizes L. Again by a simple trunca-
∗
tion procedure, we may assume that g is a simple function, so that g ∈ LΦ (µ).
This means that we can concentrate on the norm estimate. It remains to test
this estimate on f = χ{g6=0} |g|−1 Φ∗ (|g|).
1.5.3 Orlicz-averages
(p)
On many occasions it is important to consider the powered average mQ (f )
of functions f over a cube Q for 0 < p < ∞. We will apply the idea of
generalizing p to a Young function Φ : [0, ∞) → [0, ∞) to generalize the
above notion of the powered average. Here, we are concerned with the case
Banach function lattices 61
mQ (|f |)
Z Z
1 1
≤ χ[0,1] (|f (x)|)|f (x)|dx + χ(1,∞] (|f (x)|)Φ(|f (x)|)dx
|Q| Q Φ(1)|Q| Q
1
≤1+ .
Φ(1)
In Lemma 55, we used the fact that Φ(1)t ≤ Φ(t) for t ≥ 1. Note that
Φ(1)t ≥ Φ(t) for 0 ≤ t ≤ 1. As is guessed from this fact, the value Φ(t) with
0 ≤ t ≤ 1 does not affect very much. This fact can be summarized in the
following proposition:
Proposition 56. Let Φ1 , Φ2 : [0, ∞) → [0, ∞) be Young functions. Assume
that Φ1 (t) ≤ Φ2 (t) for all t > t0 , where t0 > 0 is a fixed number. Then for all
cubes Q and all f ∈ L0 (Q), kf kΦ1 ;Q . kf kΦ2 ;Q .
Proof Assume that kf kΦ2 ;Q = 1. Then
Z Z
1 1
Φ1 (|f (x)|χ(t0 ,∞] (|f (x)|))dx = Φ2 (|f (x)|χ(t0 ,∞] (|f (x)|))dx
|Q| Q |Q| Q
≤ 1,
Proof We may assume that the right-hand side is finite or more strongly
equals 1. Then Z
1
Φ(|Q2 | · |f (x)|)dx ≤ 1.
|Q2 | Q2
Since Φ is a Young function and Q1 ⊂ Q2 , we have
1 1
Φ(|Q1 | · |f (x)|) ≤ Φ(|Q2 | · |f (x)|) (x ∈ Rn ).
|Q1 | |Q2 |
It remains to integrate this pointwise estimate over Q2 using Lemma 42 and
discard the contribution of Q2 \ Q1 in the left-hand side.
Banach function lattices 63
Let Q ∈ Q, and let U be any partition of Q into cubes. Also let f ∈ L1 (Q).
Then by the pigeon-hole principle, we have mR (|f |) ≥ mQ (|f |) for some R ∈
U. We will generalize this fact to the Φ-average as follows:
Lemma 58. Let f ∈ L0 (Rn ). Then kf kΦ;Q ≤ 2n sup kf kΦ;Q0 for
Q0 ∈Q(Q) : `(Q0 )=t0
any cube Q ∈ Q and any positive number t0 ≤ `(Q).
Proof By the monotone convergence theorem, we may assume f ∈
L∞ (Rn ). By the normalization, we may also assume kf kΦ;Q = 1. Let t0 ∈
(0, `(Q)] be fixed. Let
`(Q)
N = 1+ 0 ∈ [1, ∞).
t
n
Y
We write Q = [aj , aj +`(Q)], so that (a1 , a2 , . . . , an ) is the “bottom” corner
j=1
of Q. Define (
0 (mj < N ),
δ(mj ) ≡ 0
N t − `(Q) (mj = N )
for mj ∈ {1, 2, . . . , N } and
n
Y
Qm ≡ [aj + (mj − 1)t0 − δ(mj ), aj + mj t0 − δ(mj )]
j=1
t0n 1
X Z
= Φ(|f (x)|)dx.
n
|Q| |Qm | Qm
m∈{1,2,...,N }
Since
t0 N t0 t0 t0
`(Q) `(Q)
= 1+ 0 ≤ 1+ 0 ≤ + 1 ≤ 2,
`(Q) `(Q) t `(Q) t `(Q)
it follows that
Z Z
1 n 1
Φ(|f (x)|)dx ≤ 2 max Φ(|f (x)|)dx.
|Q| Q m∈{1,2,...,N }n |Qm | Qm
Thus, the right-hand side equals ∞. If the left-hand side is finite, then we
argue similarly.
64 Morrey Spaces
where
ρp (f ) ≡ k |f |p(·) χ(0,∞) (p)kL1 (µ) + kf χ{∞} (p)kL∞ (µ) .
Moreover, for f ∈ Lp(·) (µ) one defines the variable Lebesgue norm by
1.5.5 Exercises
Exercise 34. Show that any homeomorphism from [0, ∞) to [0, ∞) leaves 0
invariant. Hint: Count the number of connected components of the set [0, ∞)\
{a} for a ∈ [0, ∞).
Exercise 35.
(1) Let p > 1 and t > 0. Calculate sup(st − sp ).
s≥0
s
(2) Calculate sup(st − e + 1) for t > 1.
s≥0
Exercise 36. [192, Theorem 1.3] Let Φ : [0, ∞) → [0, ∞) be a Young function,
and let f ∈ L0 (Rn ).
kf χB(x,r) kLΦ
(1) Show that lim inf ≥ |f (x)| for almost all x ∈ Rn .
r↓0 kχB(x,r) kLΦ
kf χB(x,r) kLΦ
(2) Assume that Φ ∈ ∆2 . Then prove that lim inf = |f (x)| for
r↓0 kχB(x,r) kLΦ
almost all x ∈ Rn .
Lp (Rn )-norm. The space L∞ (Rn ) is a natural extension of the Lp (Rn )-norm.
The failure of completeness is the case, if we endow Cc∞ (Rn ) with a norm
given by X
kf kW m,p ≡ k∂ α f kLp .
α∈N0 n
|α|≤m
We can say that W m,p (Rn ), whose precise definition we are about to present,
overcomes this defect, if p < ∞.
Definition 29 (The Sobolev spaces in Lp (Rn ) with m derivatives). Let m ∈
N0 , and let 1 ≤ p ≤ ∞. Then define the Sobolev space W m,p (Rn ) by:
W m,p (Rn )
\
≡ {f ∈ Lp (Rn ) : ∂ α f exists in weak sense and belongs to Lp (Rn ) }.
|α|≤m
X
The norm of f ∈ W m,p (Rn ) is defined by: kf kW m,p ≡ k∂ α f kLp .
n
α∈N0
|α|≤m
Banach function lattices 67
Theorem 60. Let 1 ≤ p ≤ ∞ and m ∈ N. Then the Sobolev space W m,p (Rn )
is a Banach space.
Proof The proof is left as an exercise; see Exercise 38.
Next, we are going to discuss the mollification and consider the dense
subspace of W m,p (Rn ).
Lemma 61. Let 1 ≤ p < ∞. Let ϕ ∈ Cc∞ (Rn ) be a positive function with
integral 1. Then for given f ∈ W m,p (Rn ) and t > 0, we set ft ≡ ϕt ∗ f , where
1 ·
ϕt = n ϕ . Then we have lim ft = f in the topology of W m,p (Rn ).
t t t↓0
Proof The proof uses the weak compactness of Lp (Rn ) and is left as an
exercise; see Exercise 39.
Now we consider the chain rule. What is totally different from the classical
analysis is that if f ∈ W 1,p (Rn ), then we have |f | ∈ W 1,p (Rn ). This property
fails for the space C 1 (Rn ) because the limit of the difference quotient comes
into play. The next theorem is the first step for this purpose.
68 Morrey Spaces
Theorem 64. Let 1 < p < ∞. Assume that η ∈ BC1 (C) with η(0) = 0. Also
let f ∈ W 1,p (Rn ). Then η ◦ f ∈ W 1,p (Rn ).
Here if f (x) = ±∞, then it will be understood that η ◦ f (x) = 0 and below
we shall disregard such a point. We shall not allude to this point later.
Proof The proof is left as an exercise; see Exercise 40.
We next consider the Sobolev embedding theorem. Suppose that f : R → R
is
Z x smooth function with the first derivative integrable over R. Then f (x) =
a
f 0 (t) dt + f (0) and hence f is bounded. In this way if we have some infor-
0
mation on the partial derivatives, then we can say more about the function.
The Sobolev embedding theorem quantifies such a situation.
We write Du ≡ (∂x1 u, ∂x2 u, . . . , ∂xn u) for a function u. For a Banach
lattice X, we define kDukX ≡ k |Du| kX .
Theorem 65 (Gagliardo-Nirenberg-Sobolev inequality). Assume 1 ≤ p < n.
1 1 1
Define q by = − . Then kukLq .p,n kDukLp for all u ∈ Cc1 (Rn ).
q p n
Proof The proof is left as an exercise; see Exercise 41.
Function spaces are tools with which to describe the size and the smooth-
ness of functions. Let p, q, m ∈ [1, ∞] and suppose that m is an integer.
For example, Lq (Rn ) measures the size. Since the pointwise multiplication
of L∞ (Rn )-functions is closed in Lq (Rn ), Lq (Rn ) never measures the smooth-
ness of functions. However, W m,p (Rn ) can control somehow the smoothness
of functions. What counts about size and smoothness is that the control of
smoothness can be transformed into that of size. This aspect of these two
quantities can be illustrated by the next theorem.
n n n
Theorem 66. Let m ∈ N and 1 ≤ p < . Then define q by m − = − .
m p q
Then W m,p (Rn ) ,→ Lq (Rn ).
Proof The proof is left as an exercise; see Exercise 41.
n
The quantity m − in the space W m,p (Rn ) has a special name.
p
Definition 31 (Differential/Sobolev index). In W m,p (Rn ), the quantity m −
n
is referred to as the differential/Sobolev index of functions.
p
We move on to the extension property. The extension problem is a classical
problem in the theory of function spaces with important applications in many
fields of mathematical analysis, in particular harmonic analysis and the theory
of partial differential equations. Broadly speaking, the problem consists in
extending the elements of a space of functions defined on a given subset of
Rn to the whole of Rn , preserving certain differentiability and summability
properties.
Banach function lattices 69
Proof Let τ ∈ Cc∞ (B(r)) ∩ M+ (Rn ) for some small r > 0 and kτ kL1 = 1.
Set ψ̃k ≡ 2kn τ (2k ·) ∗ χρ−1
n (2
n
−k−1 ,2−k ) for each k ∈ Z. Since ϕ ∈ Lip(R ), if we
Lemma 69. Let a > 0 be fixed. Then (x0 − 2−k z 0 , xn − A2−k zn ) ∈ Ω for all
k ∈ Z, x ∈ G and z ∈ (−a, a)n−1 × (2a, 4a).
Proof This is clear from the definition of the Lipschitz norm.
We are now ready for the construction of Burenkov’s extension operator.
Choose ω ∈ Cc∞ ((−a, a)n−1 ×(2a, 4a))∩M+ (Rn ) so that kωkL1 = 1. For every
f ∈ L1loc (Ω) and x ∈ G, we set
Z
fk (x) ≡ f (x0 − 2−k z 0 , xn − A2−k zn )ω(z)dz
(−a,a)n−1 ×(2a,4a)
Z
= A−1 2kn ω(2k (x0 − y 0 ), A−1 2k (xn − yn ))f (y)dy
Ω
and
f (x), if x ∈ Ω,
∞
T f (x) ≡ X (1.3)
ψk (x)fk (x), if x ∈ G ≡ Rn \ Ω.
k=−∞
Note that ∂Ω has Lebesgue measure zero, so that T f is defined almost every-
where on Rn . We note that the operator T in (1.3) is defined by means of a
sequence of mollifiers with discrete steps and has a local nature in the sense
that the values of the extended function T f around a point in Rn \ Ω depend
only on the values of f localized around certain “reflected” points inside Ω.
Theorem 70. Let l ∈ N. Let f ∈ C l (Ω) satisfy ∂ α f ∈ L∞ (Ω) for any α ∈ N0 n
with |α| ≤ l, so that its derivative up to any order extends continuously to the
boundary. Then T f ∈ C l (Rn ).
Proof Since T f ∈ C ∞ (Rn \ ∂Ω), it suffices to prove lim ∂ α T f (x) =
x∈G,x→x0
∂ α f (x0 ) for all x0 ∈ ∂Ω and |α| ≤ l. We let x = (x0 , xn ) ∈ G. We set
Z(x)
≡ (∂ α f (x0 − 2−k z 0 , xn − A2−k zn ) − ∂ α f (x0 , ϕ(x0 )))ω(z),
Z 0 (x)
Z
≡ (∂ β f (x0 − 2−k z 0 , xn − A2−k zn ) − ∂ β f (x0 , ϕ(x0 )))ω(z)dz
(−a,a)n−1 ×(2a,4a)
ζ(s) = ζ(s; k, x, z, A, β)
≡ ∂ β f (x0 − 2−k sz 0 , s(xn − A2−k zn ) + (1 − s)ϕ(x0 )) − ∂ β f (x0 , ϕ(x0 )).
∞
∂ γ ψk (x) = 0 for any γ ∈ N0 n \ {0}, we have
P
Since
k=−∞
|fk (x)|
Z
. 2kn χ(−a,a)n−1 ×(2a,4a) (2k (x0 − y 0 ), A−1 2k (xn − yn ))|f (y)|dy
Ω
Z
kn
=2 |f (x − w)|dw
(−2−k a,2−k a)n−1 ×(2−k+1 aA,2−k+2 aA)
Z
≤ 2kn |f (x − w)|dw
(−2axn +2aϕ(x0 ),2axn −2aϕ(x0 )))n−1 ×(A(xn −ϕ(x0 )),8A(xn −ϕ(x0 )))
Z
∼ |f (x0 − (xn − ϕ(x0 ))z 0 , xn − (xn − ϕ(x0 ))zn )|dz.
(−2a,2a)n−1 ×(aA,8aA)
result.
Proof of Theorem 73 Write Σ ≡ {z ∈ Rn−1 : |z| = 1, zn < −M |z 0 |}.
Equip Σ with the (hyper)surface measure σ. We may suppose that f ∈
C ∞ (Ωϕ ) thanks to Lemma 67. Then we have
Z
|v · Df (x)|p dσ(v) ∼ |Df (x)|p (x ∈ Ω).
Σ
or equivalently,
Z 1 Z 1
|v · Df (x)| . |f (x + tσ)|dt + |D2 f (x + tσ)|dt.
0 0
Consequently,
Z Z 1 Z Z 1
|Df (x)|p . |f (x + tσ)|p dtdσ(v) + |D2 f (x + tσ)|p dtdσ(v).
Σ 0 Σ 0
Banach function lattices 73
for all g ∈ W 2,p (Ωϕε ). For f ∈ W 2,p (Ωϕ ), the function g = f (ε·) belongs to
W 2,p (Ωϕε ). Thus,
The homogeneous Lipschitz space Lipγ (Rn ) is the set of all continuous func-
tions f for which the norm kf kLipγ is finite.
74 Morrey Spaces
This definition of Lipγ (Rn ) is reasonable since the space Lipγ (Rn ) gen-
eralizes the familiar space Lip(Rn ) made up of all functions f satisfying
kf kLip ≡ sup |x − y|−1 |f (x) − f (y)| < ∞. The trouble is that we cannot
x,y∈Rn ,x6=y
extend this definition naturally to γ > 1; the space Lipγ (Rn ) is nothing but
P0 (Rn ) once we use the same formula using the first order difference.
So we consider another expression of Lipγ (Rn ). We define C˙γ (Rn ) with
0 < γ < 2 using the difference operator of order 2.
Definition 33 (Homogeneous Hölder–Zygmund space). For 0 < γ < 2 and a
continuous function f : Rn → C, define
The homogeneous Hölder–Zygmund space C˙γ (Rn ) is the set of all continuous
functions f for which the norm kf kĊ γ is finite.
Remark that if f ∈ C(Rn ) satisfies kf kĊ γ = 0 in Definition 33, then
f ∈ P1 (Rn ), so that we regard C˙γ (Rn ) as a normed space modulo P1 (Rn ). See
Exercise 42.
We will show that Lipγ (Rn ) and C˙γ (Rn ) are isomorphic for 0 < γ < 1 in
Theorem 82.
We employ Definition 33 to investigate the property of Hölder–Zygmund
spaces. Let us now choose ψ ∈ Cc∞ (Rn ) so that
For j ∈ Z, define
We show how to use the difference operator of order 2 in the next lemma.
Lemma 76. Let f ∈ C˙γ (Rn ) with 0 < γ < 2. Then kϕj ∗ f kL∞ . 2−jγ kf kĊ γ
for all j ∈ Z.
Z
1
Proof We note that ϕj ∗f (x) = ϕj (y)(f (x−y)−2f (x)+f (x+y))dy
2 Rn
since ϕj has zero integral and ϕj is an even function. It remains to use the
triangle inequality and the definition of the norm in Definition 33.
Before we go further, we will develop the argument in Lemma 76.
Let β be a multi-index with length 2. By the use of the Taylor expan-
sion we see that ϕj ∗ ϕk and ∂ β ϕj are even functions having zero inte-
gral and satisfying |ϕj ∗ ϕk | . 2min(j,k)n−2|j−k| χB(2min(j,k)+2 ) , |∂ β ϕj | .
2j(2+n) χB(2j+1 ) and |∂ α ϕj | . 2j(n+|α|) χB(2j+1 ) , which yields kϕj ∗ϕk ∗f kL∞ .
2min(j,k)n−2|j−k| kf kĊ γ , k∂ β ψ j ∗ f kL∞ . 2j(2−γ) kf kĊ γ , and k∂ α (ϕj ∗ ϕk ∗
f )kL∞ . min(2j|α|−kγ , 2k|α|−jγ )kf kĊ γ for all f ∈ C˙γ (Rn ). So we are led to
the following estimate:
Banach function lattices 75
Lemma 77. Let f ∈ C˙γ (Rn ) with 0 < γ < 2. Then kϕk ∗ ϕj ∗ f kL∞ .
∞
2−2|j−k|−min(j,k)γ kf kĊ γ for all j, k ∈ Z. In particular, ϕj ∗ϕk ∗f = ϕk ∗f
P
j=−∞
and k∂ α (ϕk ∗ f )kL∞ . 2k|α|−kγ kf kĊ γ for all k ∈ Z, α ∈ N0 n and f ∈ C˙γ (Rn )
with 0 < γ < 2 and for all multi-indexes β with |β| = 2, lim ∂ β [ψ j ∗ f ] = 0
j→−∞
in L∞ (Rn ).
∞ ∞
kϕj ∗ f kL∞ . 2−jγ kf kĊ γ ∼ kf kĊ γ from Lemma 76, which
P P
Hence
j=0 j=0
leads us to the following definition:
∞
Definition 34. Let f ∈ C˙γ (Rn ) with 0 < γ < 2. The function H ≡ ϕj ∗ f
P
j=0
∞
ϕ−j ∗f ,
P
is called the high frequency part of f . Meanwhile, the function G̃ ≡
j=1
which is defined formally, is called the low frequency part of f .
The trouble is that there is no guarantee that the right-hand side defining
G̃ is convergent in some suitable topology. To circumvent this problem, we
consider its derivative as follows:
Lemma 78. Let f ∈ C˙γ (Rn ) with 0 < γ < 2. Let ψ ∈ Cc∞ (Rn ) satisfy (1.4).
0
Define ϕj by (1.5) for j ∈ Z. Then Gα ≡ ∂ α [ϕj ∗ f ] is convergent in
P
j=−∞
L∞ (Rn ) whenever |α| ≥ 2. Hence, Gα is smooth.
Proof Simply use Lemma 77.
To construct a substitute of the lower part, we need to depend on a geo-
metric property of Rn ; HDR
1
(Rn ) = 0. Applying this fact, we can prove the
following:
Lemma 79. Let N ∈ N. Suppose that we have {fα }α∈N0 n ,|α|=N ⊂ C ∞ (Rn )
satisfying the curl-free condition: ∂β fα = ∂β 0 fα0 for all α, α0 , β, β 0 ∈ N0 n with
|α| = |α0 | = N and α + β = α0 + β 0 . Then there exists f ∈ C ∞ (Rn ) such that
∂ α f = fα .
Proof We prove this lemma by induction on N . If N = 1, then the result
1
is immediate from the Poincaré lemma, or equivalently, HDR (Rn ) = 0. Let
N ≥ 2. Define gjγ ≡ fej +γ if j = 1, 2, . . . , n and |γ| = N − 1. Let ej ≡
(0, . . . , 0, 1, 0, . . . , 0), where 1 is in the j-th lot. Then ∂ δ gjγ = ∂ δ fej +γ =
0 0
∂ δ fej0 +γ = ∂ δ gj 0 γ for all j, j 0 = 1, 2, . . . , n, γ, δ, δ 0 ∈ N0 n with |γ| = N − 1
and ej + δ = ej 0 + δ 0 . Thus, we are in the position of applying the Poincaré
lemma to have gγ ∈ C ∞ (Rn ) satisfying gjγ = ∂ j gγ for all γ with |γ| = N − 1
and j = 1, 2, . . . , n.
Let α, α0 , γ, γ 0 ∈ N satisfy |γ| = |γ 0 | = N − 1 and |α| = |α| > 0 and α + γ =
α + γ 0 . Suppose α − ej , α0 − ej 0 ≥ 0. Then ∂ α gγ = ∂ α−ej ∂ j gγ = ∂ α−ej gjγ =
0
76 Morrey Spaces
0 0 0 0 0
∂ α−ej fej +γ and ∂ α gγ 0 = ∂ α −ej0 ∂ j gγ 0 = ∂ α −ej0 gj 0 γ 0 = ∂ α −ej0 fej0 +γ 0 . Since
0
(α − ej ) + (ej + γ) = (α0 − ej 0 ) + (ej 0 + γ 0 ), we obtain ∂ α gγ = ∂ α gγ 0 , thus,
∞ n
by the induction assumption to have a function f ∈ C (R ) such that gγ =
∂ γ f . Consequently, if |α| = N , then by choosing j so that ej ≤ α we have
fα = ∂j gα−ej = ∂j ∂ α−ej f = ∂ α f .
We go back to the analysis of the low frequency part. As before choose
ψ ∈ Cc∞ (Rn ) so that χB(1) ≤ ψ = ψ(−·) ≤ χB(2) . Define ϕj ≡ 2jn ψ(2j ·) −
2(j−1)n ψ(2j−1 ·) for j ∈ Z. By using Lemma 79, we have the following control
of the low frequency part:
Corollary 80. Let f ∈ C˙γ (Rn ) with 0 < γ < 2. Let ψ ∈ Cc∞ (Rn ) satisfy
(1.4). Define ϕj by (1.5) for j ∈ Z. There exists a function G ∈ C ∞ (Rn ) such
0
that ∂ α G = ∂ α ϕj ∗ f for all α ∈ N0 with |α| ≥ 2.
P
j=−∞
and
|Hj0 (x + y) − Hj0 (x)| . |y|γ kf kĊ γ , (1.9)
then we have (1.6).
To prove (1.8), by Lemma 77 we estimate
as was to be shown.
From now on, we can identify Lipγ (Rn ) with C˙γ (Rn ) for 0 < γ < 1.
So far, we do not consider the meaning of the parameter γ. If γ ∈ (1, 2),
we can differentiate functions in C˙γ (Rn ).
Theorem 83. Let f ∈ C˙γ (Rn ), and let 1 < γ < 2. Then f ∈ C 1 (Rn ) and
∂j f ∈ C˙γ−1 (Rn ) for all j = 1, 2, . . . , n.
Proof We use the functions G and H in the proof of Theorem 82. Arguing
∞
∂j [ϕl ∗ f ], j = 1, 2, . . . , n, converges uniformly,
P
as before, we can show that
l=0
so that H ∈ C 1 (Rn ). Since f differs from G + H by a linear function, we see
that f ∈ C 1 (Rn ). We can show that ∂j (G + H) ∈ C˙γ−1 (Rn ) by an argument
similar to Theorem 82. Hence, ∂j f ∈ C˙γ−1 (Rn ).
We consider the converse of Theorem 83.
Proposition 84. Let f ∈ C 1 (Rn ), and let 1 < γ < 2. If ∂j f ∈ C˙γ−1 (Rn ) for
each j = 1, 2, . . . , n, then f ∈ C˙γ (Rn ).
Proof Let x, y ∈ Rn . We may assume that f is real-valued. By the mean-
value theorem,
for some 0 < t < 1. If we use ∂j f ∈ C˙γ−1 (Rn ) for each j = 1, 2, . . . , n, then
n
we obtain |f (x + y) − 2f (x) + f (x − y)| . |y|γ
P
k∂j f kC γ−1 .
j=1
Exercises
Exercise 37.
(1) Prove Lemma 59(1) by the use of the Lebesgue differentiation theorem.
(2) Prove Lemma 59(2) by the use of the Stokes theorem.
Exercise 38. Let 1 ≤ p ≤ ∞ and m ∈ N.
(1) Show that W m,p (Rn ) is a normed space using the fact that Lp (Rn ) is a
normed space.
(2) Prove Theorem 60.
Exercise 39. Using the Banach–Alaoglu theorem (see Theorem 89), prove
Theorem 63.
Exercise 40. Prove Theorem 64 for f ∈ Cc∞ (Rn ) and then complete the
proof of Theorem 64.
Exercise 41.
(1) Prove Theorem 65 by using Theorems 182 and 193 to follow.
(2) Prove Theorem 66 by inducting on m. Hint: The base case m = 1 is
Theorem 65.
Exercise 42. Let f : Rn → C be a continuous function satisfying f (x + y) +
f (x − y) = 2f (x) for all x, y ∈ Rn .
Banach function lattices 79
1.7 Notes
Section 1.1
General remarks and textbooks in Section 1.1
There are many books on Lebesgue spaces; see [156, 258, 319] for example.
We refer to the exhaustive textbooks [29] and [296] for Banach function spaces.
We refer to the textbooks [189, 319] for an introduction of integration theory.
We refer to the recent textbook [203] for mixed Morrey spaces. The authors
hit upon Exercise 9 after the lecture of Kaoru Yoneda.
80 Morrey Spaces
Section 1.1.3
See [302] for Marcinkiewicz spaces.
In 1961, Benedek and Panzone introduced Lebesgue spaces with mixed
norm [27]. See [24, 128, 195, 196, 211] for the studies of mixed Lebesgue
spaces such as the boundedness of the Hardy–Littlewood and strong maximal
operators, multivariate rearrangements, the theory of variable exponents, and
interpolation theory.
Section 1.2
General remarks and textbooks in Section 1.2
Morrey spaces go back to the work of Morrey [323], which handled elliptic
differential equations. Actually Morrey pointed out that the function is con-
tinuous once the derivative satisfies a weak integrability condition. Peetre and
Stampaccia considered Morrey and Campanato spaces [353, 413]. We refer to
the textbooks [235, Chapter 13], [258, Chapter 5], [438] and [382, §6.1.5] for
Morrey spaces. We refer to [29], [258, Chapter 6] for Banach function spaces.
Section 1.2.1
n
There are many papers dealing with | · |− p in the context of Morrey
spaces. Zorko considered this function to explain that smooth functions cannot
approximate this function [472, p. 587]. See [304, Proposition 4.1] for Example
6, for example.
The space Mp1 (Rn ) was defined by Adams, Xiao, Giga and Miyakawa.
See [1, (4)], [149, p. 582, §2] as well as [8, p.1631] for Mp1 (Rn ), Definition 9.
Among others, Adams used Mp1 (Rn ) to characterize the boundedness property
of the convolution operator [1, Theorems 1 and 2]. We can find Proposition
15, asserting that Mpq (Rn ) is trivial for 0 < p < q < ∞, in [42, p. 13].
Section 1.2.2
We calculated kχB(x,r) kMpq ; see Example 6 [252, Lemma 1], for example.
The dyadic rearrangement (Example 10) is due to Olsen [342]. Olsen worked
in R3 and constructed an example in [342, Theorem 10]. Example 10 is a
passage to Rn of [342, Theorem 10].
We considered the power function in Example 7. Example 7(1) can be
found in [472], while Example 7(2) and (3) can be found in [42, Examples 1
and 2].
In [390], Sawano, Sugano and Tanaka considered the self-similar increas-
ing/decreasing sequence for Morrey spaces. Examples 9 and 11 come from [390,
Banach function lattices 81
Proposition 4.1], based on the proof of [64, Theorem 2.1]. See also [304, Lemma
4.6] and [391, Proposition 2.1] for Examples 9 and 11. The term “wasteful”
comes from the proof of the survey paper [391, Proposition 2.1].
Section 1.2.3
Giaquinta pointed out the role of the parameter q; see [148, p. 67] for
Theorem 18.
As we saw in this book, we have an approach using a self-similar decreasing
sequence for Morrey spaces. See [390, Proposition 4.1] together with detailed
calculation [304, p.120] and [393, Lemma 2.1] for Example 9.
As we saw in Theorem 20, Mpq̃ (Rn ) is not dense in Mpq (Rn ) whenever
1 < q < q̃ < p; see the paper [380].
Section 1.2.4
Triebel pointed out that the Morrey space Mpq (Rn ) with 0 < q < p < ∞
does not have L∞ n
c (R ) and the Schwartz space as dense subspaces; see [438,
Proposition 2.16] for Proposition 22.
As a modification of Morrey spaces, we can consider the Stummel class
considered by Ragusa and Zamboni [365]. See [442] for the recent development.
Eridani and Gunawan, pointed out the inclusion between Mp1 (Rn ) and the
Stummel class [115]. See [116] for a further generalization to various directions.
As we saw in Proposition 21, Morrey spaces are not rearrangment invari-
ant, that is, it can happen that kf kMpq > kgkMpq if λf = λg . Furusho and Ono
proposed to restrict the class of Morrey functions to a special case by using a
finite subdivision of cubes; see [137, Definition 1] and [345, 346, 347]. See also
[347] for the inclusionship similar to Theorem 19.
Section 1.2.5
The definition of weak Morrey spaces can be found in [405]; see also [2,
Section 5]. Applications to PDE can be found in [314, 360]. See [304, Lemma
4.7(ii)] for Example 17. See also [173] for Example 17, a more quantitative
approach for the difference between WMpq (Rn ) and Mpq (Rn ).
In [172], we can find Lemma 23, which compares WMpr (Rn ) and Mpq (Rn )
for 0 < q < r ≤ p.
As we saw in Theorem 24, the weak Morrey space WMpq (Rn ) is smaller
than Mpq (Rn ); see [139, p. 2143], [149, Proposition 2.2], [440, p. 4] and [471,
p.86] as well as [172]. We refer to [186] for a weighted version.
Section 1.2.6
The terminology of ball Banach function spaces can be tracked back to
[178, Definition 2]. See [398, Example 3.3] for an account of why Morrey
spaces are not Banach function spaces.
82 Morrey Spaces
Section 1.3
General remarks and textbooks in Section 1.3
Theory of Herz spaces goes back to the work of Herz [193]. See the textbook
[458] for Herz spaces.
Section 1.3.1
Although Wiener did not consider local Morrey spaces when his paper [447]
was published, the idea of defining local Morrey spaces can be found in [447]. In
fact, Wiener considered the case of p = 1, 2 in the context of Fourier analysis.
Beurling [31] introduced the spaces B p (Rn ) and Ḃ p (Rn ), whose norm is given
(p) (p)
by kf kB p ≡ sup mQ(r) (f ) and kf kḂ p ≡ sup mQ(r) (f ) together with the pred-
r≥1 r>0
ual Ap (Rn ) so-called a Beurling algebra. In this case B p (Rn ) = B np (Lp )(Rn )
and Ḃ p (Rn ) = Ḃ np (Lp )(Rn ). Later, to extend Wiener’s ideas [447, 448] which
describe the behavior of functions at infinity, Feichtinger [126] gave an equiv-
alent norm on B p (Rn ), which is a special case of norms to describe non-
α
homogeneous Herz spaces Kp,r (Rn ) introduced in [193]. Garcı́a-Cuerva and
Herrero [142] and Alvarez, Guzmán-Partida and Lakey [17] introduced the
non-homogeneous central Morrey space B p,λ (Rn ) (the non-homogeneous Bσ -
Lebesgue space/ the non-homogeneous central Morrey space) and the central
Morrey space Ḃ p,λ (Rn ) (the homogeneous Bσ -Lebesgue space/ the homoge-
neous central Morrey space).
p1
−λ− n p
kf kWB p,λ ≡ sup r p sup t |{|f | > t} ∩ Q(r)|
r≥1 t>0
and
p1
−λ− n p
kf kWḂ p,λ ≡ sup r p sup t |{|f | > t} ∩ Q(r)| .
r>0 t>0
Section 1.3.2
Herz–Morrey spaces are defined by replacing the Lp (Rn )-norm with the
Morrey norm k · kMpq in some sense. Lu and Xu defined Herz–Morrey spaces
[295, Definition 1.1]. Wang and Shu considered the Littlewood–Paley gλµ -
function in Herz–Morrey spaces [445]. Herz–Morrey spaces are investigated
by Izuki in the setting of variable exponents. For details see [214, 215].
Wu investigated the boundedness property of Hardy operators as well as
the Riesz transform in [452, 453, 454]. We refer to [431] for the multilinear
case.
Local Morrey spaces and Herz spaces overlap largely [167, Lemma 8.1]; see
Theorem 26. Mizuta and Ohno took an approach similar to Theorem 26 [321].
Gao, Tang, Xue and Zhou investigated the boundedness properties of com-
mutators generated by Hardy operators and BMO functions [140, 429].
Herz–Morrey spaces can be defined over locally compact groups. More
precisely, Wu and Liu worked in locally compact Vilenkin groups in [451, 455].
Dzhabrailov and Khaligova worked in the anisotropic setting in [110, 111].
Fan and Gao considered the boundedness property of the bilinear fractional
integral operator acting on Herz–Morrey spaces [120].
Kuang considered Hilbert/Hausdorff operators acting on Herz–Morrey
spaces; see [256, 257].
Section 1.4
General remarks and textbooks in Section 1.4
Edmunds, Kokilashvili and Meskhi dealt with Hardy operators. Especially,
[112, Chapter 1] deals intensively with this topic [112].
See [373, Chapter 2] for Hardy’s inequality.
See [30, §1.3], [29, Chapter 2 §1], [156, Chapter 1], [147, §1.4], [258, Chapter
8] and [223, Chapter 0, §II] for the distribution of functions and Lorentz spaces.
In [147, Chapter 2], we can find some information on the Hardy–Littlewood
maximal operator. See [30], [156, §1.4], [417, Chapter V, §3] for a detailed
explanation on Lorentz spaces.
Section 1.4.3
We can find inclusion Φ↑λ,q0 (0, ∞) ⊂ Φ↑λ,q1 (0, ∞); see [50, Lemma 1] for
Lemma 37. We considered the boundedness of K0 ; see [77, Theorem 2.5] for
Example 40.
84 Morrey Spaces
Section 1.4.4
See [364] Lemma 40, Chiarenza’s lemma.
We polished [32, §2] to state Example 41.
Section 1.5
General remarks and textbooks in Section 1.5
See the exhaustive textbook and seminar note [29, 299, 366] as well as
[147, pp. 22–23], [258, Chapter 4] and [382, §6.1.6] for Orlicz spaces and Young
functions. Trudinger pointed out that Orlicz spaces can be used to compensate
for the failure of the endpoint case for integral operators. See [439]. We can
find some maximal estimates in [147, §5.2] (including [147, Chapter 6]) and
[382] and some singular integral estimates in [147, §5.6]. We refer to [258,
Chapter 13] for Lebesgue spaces with variable exponents. See also the survey
[216].
Section 1.5.1
Young proved an inequality for the Young function and its conjugate [466];
see Theorem 45. We obtained an estimate for ∇2 -functions. See [232, Lemma
1.3.2] for Lemma 49(1) and [387, Lemma 4.3] for Lemma 49(2). We borrowed
[297, Example 2.3] for Example 43.
Section 1.5.2
Koshi and Shimogaki obtained an equivalent expression of Orlicz norms
[238]; see Theorem 51.
Kita established similar results to Proposition 156 for Ψ(t) ≡ t on LΦ (Rn ).
See [228, 229].
Section 1.5.3
Carro, Pérez, F. Soria and J. Soria defined the Φ-average in [64, §3]; see
Definition 26. See [64, (3.1)] for Definition 26 including Example 46.
Section 1.5.4
Orlicz initially considered Lebesgue spaces with a variable exponent [348,
§2]. Later on Nakano considered variable Lebesgue norms [336, 337].
In particular, function spaces with variable exponents are necessary in
the field of electronic fluid mechanics [374] and the applications to image
restoration [69, 187, 276]. Kovacik and Rakosnik [243] gave an application of
generalized Lebesgue spaces with variable exponents to Dirichlet boundary
value problems for nonlinear partial differential equations with coefficients
of a variable growth. Another simple example of the application to differen-
tial equations can be found in [121, p. 438, Example], where Fan and Zhao
Banach function lattices 85
implicitly showed that the variable Lebesgue spaces can be used to control
the non-linear term of differential equations.
See the survey paper [216] for a detailed explanation of the old books
[336, 337].
Section 1.6
General remarks and textbooks in Section 1.6
There are many textbooks dealing with Sobolev spaces; see the exhaustive
textbooks [12, 13, 309] as well as the textbooks [5], [117, §5.2], [118, Chapter
4], [415, §V.2] and [410, §6.3] for example.
See [117, §5.1], [415, Chapter V, §4] and [416, Chapter VI, §5] for a char-
acterization of Hölder–Zygmund spaces. For the extension, we refer to the
textbook [41].
Section 1.6.1
We followed the idea of Burenkov [41]. See [40] for Lemma 68. Theorem
70 is due to Burenkov [41]; see also the textbook [41, Lemma 18]. See [411]
for Theorem 66.
Section 1.6.2
See [415] for Definition 33. We followed [384] overall in this section.
Chapter 2
Fundamental facts in functional
analysis
This chapter collects some preliminary facts. The readers may choose to skip
Chapter 2. We indicate where we use each theorem in the other parts of this
book. We start with normed spaces in Section 2.1, while we concentrate on
Hilbert spaces in Section 2.2. Section 2.3 is used for the last chapter of the
second book. We will consider integration theory for measurable functions
assuming their values in a Banach space.
87
88 Morrey Spaces
We next recall the Banach–Alaoglu theorem. Recall that the metric space
is separable if it is realized as the closure of a countable subset.
Theorem 89 (Banach and Alaoglu). Let X be a separable normed space.
Assume that {xj ∗ }∞ ∗
j=1 belongs to the closed unit ball B in X . Then there
∗
exists a subsequence {xjk }k∈N convergent with respect to weak-∗ topology.
That is,
lim xjk ∗ (x) = x∗ (x)
k→∞
for all x ∈ X .
We do not recall the proof of Theorems 85, 86, 87, 88 and 89. See [310] for
example.
X0 , X1 ,→ X Y0 , Y1 ,→ Y,
X0 + X1 ≡ {x ∈ X : x = x0 + x1 , x0 ∈ X0 , x1 ∈ X1 }.
| hx∗ , xi |
kx∗ kX0 ∗ ∩X1 ∗ = sup for all x∗ ∈ X0 ∗ ∩ X1 ∗ .
x∈X0 +X1 \{0} kxkX0 +X1
and that
| hx∗ , xi |
kx∗ kX0 ∗ +X1 ∗ ≥ sup , (2.2)
x∈X0 ∩X1 kxkX0 ∩X1
| hx∗ , xi |
kx∗ kX0 ∗ ∩X1 ∗ = sup . (2.3)
x∈X0 +X1 kxkX0 +X1
Fundamental facts in functional analysis 91
kx0 ∗ kX0 ∗ + kx1 ∗ kX1 ∗ = kLk(X0 ⊕X1 )∗ ≤ kx∗ k(X0 ∩X1 )∗ (2.4)
2.1.4 Exercises
Exercise 45. [225] Let X and Y be Banach spaces and T : X → Y a linear
operator. Suppose that there exist constants M > 0 and 0 < ε < 1 with the
following property: For every y ∈ Y with unit norm, there exists x ∈ X with
kxkX ≤ M and kT x − ykY < ε. Then show that T is onto.
Exercise 46. Let X , Y be normed spaces, and let T : X → Y be a lin-
ear operator. Define kT kX →Y ≡ sup{kT xkY : kxkX = 1}. Then show that
kT kX →Y = sup{kT xkY : kxkX < 1}.
Exercise 47 (Clarkson’s inequality). [319, Theorem 12.6] Let (X, B, µ) be a
measure space, and let 1 < p ≤ 2.
p0 p0 p 1
a+b a−b |a| + |b|p p−1
(1) Show that + ≤ for all a, b ∈ R.
2 2 2
(Hint: We may assume b = 1 ≥ a.)
!p0 !p0
f +g f −g
(2) Show that +
2 Lp (µ) 2 Lp (µ)
0
p p p −1
(kf kLp (µ) ) + (kgkLp (µ) )
≤ for all f, g ∈ L0 (µ). (Hint: Prove
2
0 0 0
first that kf kLp0 (µ) p + kgkLp0 (µ) p ≤ kf + gkLp0 (µ) p for f, g ∈ M+ (µ).)
92 Morrey Spaces
p0 p0 0 0
a+b a−b |a|p + |b|p
(3) Show that + ≤ for all a, b ∈ R.
2 2 2
!p0 !p0
f +g f −g
(4) Show that +
2 Lp0 (µ) 2 Lp0 (µ)
0 0
(kf kLp0 (µ) )p + (kgkLp0 (µ) )p
≤ for all f, g ∈ L0 (µ).
2
Exercise 48. In Example 50, we considered the norm of f ∈ L1 (Rn ) +
L∞ (Rn ). Let L ≥ 0. Then find the minimum of kf − hkL1 , where h ∈ L0 (Rn )
moves over all functions satisfying khkL∞ = L.
Exercise 49. Let p(·) : Rn → [1, ∞) be a variable exponent. Let p+ ≡
ess.supx∈Rn p(x) and p− ≡ ess.inf x∈Rn p(x). Then establish that Lp+ (Rn ) ∩
Lp− (Rn ) ⊂ Lp(·) (Rn ) ⊂ Lp+ (Rn ) + Lp− (Rn ).
Proof
(1) This is a standard result in functional analysis. See [409, Theorem 3.2.1],
for example.
(2) Write vk ≡ min{kgkH ∈ co({fk+1 , fk+2 , . . .})}. Note that kgk −gk0 kH 2 =
2kgk kH 2 + 2kgk0 kH 2 − kgk + gk0 kH 2 ≤ 2vk 2 + 2vk0 2 − 4vmin(k,k0 ) 2 , since
gk + gk0
∈ co({fmin(k,k0 )+1 , fmin(k,k0 )+2 , . . .})}.
2
Note that {vk }∞
k=1 is a bounded increasing sequence. Hence, it converges
and we obtain the desired result.
and that
We will estimate this sum by majorizing the norms of the individual sum-
mands.
First, associating the factors in each summand as
m
Y
Tj∗1 Tj2 Tj∗3 · · · Tj2m = (Tj∗2l−1 Tj2l ),
l=1
We take the geometric mean of (2.5) and (2.6) and insert this into (2.4). The
result is
X 2m−1
Y
k (TJ∗ TJ )m kH→H ≤ A γjl −jl+1 .
j1 ,j2 ,...,j2m ∈J l=1
P
In the above, we first sum in j1 and use the fact that γj1 −j2 ≤ A. Next, we
P j1 ∈J
sum in j2 , using γj2 −j3 ≤ A. Continuing in this way for j1 , j2 , . . . , j2m−1
j2 ∈J
yields k (TJ∗ TJ )m kH→H ≤ A2m−1 ]J. Recall that we assumed J finite. Thus,
1
kTJ kH→H ≤ A(]J) 2m . Finally, we let m → ∞, proving the theorem.
2.2.3 Exercises
Exercise 50. Let 1 < p < ∞. By using Clarkson’s inequality (see Exercise
47), prove an analogue in Lp ([0, 1]) to Proposition 93.
Exercise 51. Suppose that we have {Tj }j∈Z ⊂ B(H). and a sequence of
positive constants {γj }j∈Z satisfying (2.3) with J = Z and
∞
X
A≡ γj < ∞. (2.7)
j=−∞
∞
Tj Tk∗ x converges for all x ∈ H and k ∈ Z.
P
(1) Show that
j=−∞
Fundamental facts in functional analysis 95
∞
P S
(2) Show that Tj z converges for all z ∈ Im(Tj ).
j=−∞ j∈Z
∞
P
(3) Construct the limit of Tj in an appropriate sense. Hint: What
j=−∞
∞
T
happens on ker(Tj )?
j=−∞
The integral appearing in the right-hand side of this formula is strictly speak-
ing the Bochner integral.
For the sake of simplicity, we assume that (X, B, µ) is a complete σ-finite
measure space in Section 2.3.
Section 2.3 is interested mainly in Bochner integrals. Section 2.3.1 extends
the notion of measurability of functions to Banach spaces-valued functions.
Section 2.3.2 considers some convergence theorems in analogy with the ones
in the theory of the Lebesgue integral. Section 2.3.3 is devoted to the theory
of repeated integrals.
where we have used the Lebesgue convergence theorem for the second equality.
Uniqueness of Φ follows from the Hahn–Banach theorem, Theorem 87.
We generalize the triangle inequality so that it is adapted to our current
setting.
Lemma 98. Let ϕ : X → X be an integrable countably simple function. Then
Z Z
ϕ(x)dµ(x) ≤ kϕkX (x)dµ(x). (2.1)
X X X
∞
P
In particular, if ϕ has an integrable representation ϕ ' aj χEj , then
j=1
Z ∞
X
ϕ(x)dµ(x) ≤ kaj kX µ(Ej ). (2.2)
X X j=1
98 Morrey Spaces
∞
P
showing (2.1). Since kϕkX (x) ≤ kaj kX χEj (x) for µ-almost every x ∈ X,
j=1
(2.2) follows. Therefore, the proof is now complete.
The next lemma shows that the choice of the integrable representation can
be made not so wasteful.
Lemma 99. Suppose that ϕ : X → X is a countably simple measurable
∞
P
function and ε > 0. Then ϕ has an integrable representation ϕ ' aj χEj
j=1
such that
∞
X Z
kaj kX µ(Ej ) < kϕkX (x)dµ(x) + ε. (2.3)
j=1 X
∞
P
Proof Suppose that ϕ ' bj χFj is an integrable representation. Then
j=1
there exists an increasing sequence of integers N1 < N2 < · · · such that
∞
X ε
kbj kX µ(Fj ) < . (2.4)
2k+1
j=Nk
k −1
Then the disjointness of {Ej }M
j=Mk−1 yields
Z k −1
NX Z Mk −1
X
bj χFj (x) dµ(x) = al χEl (x) dµ(x)
X j=Nk−1 X l=Mk−1
X X
Mk −1
X
= kal kX µ(El ).
l=Mk−1
Fundamental facts in functional analysis 99
We also have
∞
X ∞ Z
X k −1
NX
kal kX µ(El ) = bj χFj (x) dµ(x).
l=1 k=1 X j=Nk−1
X
∞
X Z 1 −1
NX ∞ X
X ∞
kal kX µ(El ) ≤ bj χFj (x) dµ(x) + kbj kX µ(Fj ).
l=1 X j=1 k=1 j=Nk
X
∞
X Z ∞
X X ∞
∞ X
kal kX µ(El ) ≤ f (x) − bj χFj (x) dµ(x) + kbj kX µ(Fj ).
l=1 X j=N1 k=1 j=Nk
X
Therefore, we obtain
∞
X Z X ∞
∞ X
kal kX µ(El ) ≤ kf (x)kX dµ(x) + 2 kbj kX µ(Fj )
l=1 X k=1 j=Nk
Z
≤ kf (x)kX dµ(x) + ε.
X
The symbol L1 (X; X ) denotes the set of all Bochner integrable functions.
100 Morrey Spaces
R
Lemma 100. The definition of the integral ϕ(x)dµ(x) makes sense. That
R X
is, in that the limit (2.8) defining ϕ(x)dµ(x) does exist and the element
X
ϕ(x)dµ(x) ∈ X is independent of the choice of {ϕj }∞
R
j=1 .
X
Proof We will show that the limit (2.8) exists. To see this, we observe, if
j, k ≥ J
Z Z Z
ϕj (x)dµ(x) − ϕk (x)dµ(x) ≤ 2 sup kϕ(x) − ϕl (x)kX dµ(x).
X X X l≥J X
Z ∞
Therefore ϕj (x)dµ(x) is a Cauchy sequence in X and the limit there-
X j=1
fore does exist.
We will show that the integral is independent of the admissible representa-
tion of ϕ. Suppose {ψj }∞
j=1 is another sequenceZof countably simple integrable
functions such that lim ψj = ϕ and that lim kϕ(x) − ψj (x)kX dµ(x) = 0.
j→∞ j→∞ X
Then by the triangle inequality we obtain
Z Z
ϕj (x)dµ(x) − ψj (x)dµ(x)
X X X
Z
≤ (kϕ(x) − ϕj (x)kX + kϕ(x) − ψj (x)kX ) dµ(x),
X
Z ∞
which tends to 0 as j → ∞. Thus, it follows that both ϕj (x)dµ(x)
X j=1
Z ∞
and ψj (x)dµ(x) tend to the same limit. Therefore, the integral is
X j=1
independent of the admissible representations of ϕ.
Lemma 101. In addition to the same setting as above, we let ϕ : X → X be
a Bochner integrable function. Then
Z Z
ϕ(x)dµ(x) ≤ kϕkX (x)dµ(x).
X X X
Proof It is just a simple matter to pass to the limit of the countably simple
functions. However, it is worth considering the case when X = C. Writing out
the statement in full with X = C, we can easily imagine what is the heart of
the matter in the proof of Lemma 101.
We characterize Bochner integrable functions.
Theorem 102 (Bochner). The function ϕ : X → X is Bochner integrable, if
and only if ϕ is an integrable countably simple function.
Fundamental facts in functional analysis 101
Z Z ∞
X
kϕ(x) − ϕj (x)kX dµ(x) = ak χEk (x) dµ(x)
X X k=j+1
X
∞
X
≤ kak kX µ(Ek ).
k=j+1
If we invoke Lemma 99, then for each j ∈ N, we can find an integrable repre-
sentation
∞
X ∞
X
ϕj+1 − ϕj ' aj+1,k χEj+1,k with kaj,k kX µ(Ej,k ) < 2−j . (2.9)
k=1 k=1
∞
P
We also have an integrable representation of ϕ1 : ϕ1 ' a1,k χE1,k . Using
k=1
(2.9), we conclude that ϕ has the following integrable representation
∞
X
ϕ' aj,k χEj,k .
j,k=1
∞
S
Proof Since Yj is separable, whenever each Yj is separable, ϕ is
j=1
strongly measurable.
We claim that ϕ ∈ L1 (X; X ) and (2.10) holds. Since ϕj ∈ L1 (X; X ), by
Theorem 102 we conclude that ϕj is an integrable countably simple function.
Furthermore, kϕ(x)−ϕj (x)kX ≤ 2ψ(x) for µ-almost everywhere x ∈ X. Thus,
if we invoke Lebesgue’s convergence theorem, we obtain
Z
lim kϕ(x) − ϕj (x)kX dµ(x) = 0.
j→∞ X
∞
P
Then for x ∈ X, we define ψ(x) ≡ aj,k χEj,k (x), provided the series
j,k=1
converges and ψ(x) = 0 otherwise. Then ψ has an integrable representation
Fundamental facts in functional analysis 103
∞ l
aj,k χEj,k . Therefore ψ ∈ L1 (X; X ). Since ψ −
P P
ψ ' ϕj has an inte-
j,k=1 j=1
l
P ∞
P ∞
P ∞
P
grable representation ψ− ϕj ' aj,k χEj,k , we conclude ψ = ϕj .
j=1 j=l+1 k=1 j=1
Finally let us characterize the Bochner integral in terms of the dual spaces.
Proposition
Z 105. Let ϕ : X → X be a Bochner integrable function. Then
b≡ ϕ(x)dµ(x) is a unique element in X satisfying
X
Z
b∗ (b) = b∗ (ϕ(x))dµ(x)
X
∗ ∗
for all b ∈ X .
Proof Now that “Bochner integrable function” and “integrable countably
simple function” are synonymous, this proposition is Lemma 97 itself.
(2) We have
ZZ Z Z
ϕ(x, y)dµ ⊗ ν(x, y) = ϕ(x, y)dν(y) dµ(x)
X×Y X Y
Z Z
= ϕ(x, y)dµ(y) dν(y).
Y X
Proof
ZZ
(1) Since kϕ(x, y)kX dµ ⊗ ν(x, y) < ∞, assertion (1) follows imme-
X×Y
diately.
∞
P
(2) Pick an integrable representation of ϕ: ϕ ' aj χGj , where each Gj is
j=1
a µ ⊗ ν-measurable set. By disregarding a set of measure zero we may
104 Morrey Spaces
∞
P R
assume kaj kX · Y
χGj (x, y)dν(y) < ∞ for all x ∈ X. Then for all
j=1
Z ∞
X Z
x ∈ X, ϕ(x, y)dµ(y) = aj χGj (x, y)dν(y). Since
Y j=1 Y
∞ Z
X Z ∞
X Z
aj χGj (x, y)dν(y) dµ(y) = kaj kX χGj (x, y)dν(y)
j=1 X Y X j=1 Y
< ∞,
we are in the position of using the Lebesgue convergence theorem (The-
orem 103) for the Bochner integral to obtain
Z Z X∞ Z
ϕ(x, y)dµ(y) dν(x) = aj χGj (x, y)dµ(x)dν(y)
X Y j=1 X×Y
ZZ
= ϕ(x, y)dµ ⊗ ν(x, y).
X×Y
≤ kϕk∗ · |t − s|.
Let us define
Φj (t) ≡ 2j (Ψ(t + 2−j ) − Ψ(t)) (t ∈ R).
Then
kΦj (t)kX ∗ ≤ kϕk∗ (t ∈ R)
by virtue of (2.11). Therefore, by the Banach Alaoglu theorem, Theorem 89,
there exists a subsequence {Φjk }k∈N such that the weak-∗ limit
Φ(t) ≡ lim Φjk (t)
k→∞
Since s ∈ Q is arbitrary, this implies lim Φjk (t) = Φ(t) for all t ∈ R. As a
k→∞
result, for each g ∈ L∞ d
c (R ; X ),
Z Z
hΦ(t), g(t)idt = lim hg(t), Φjk (t)iX ,X ∗ dt = ϕ(g).
R k→∞ R
Since L∞ d 1 d
c (R ; X ) is dense in L (R ; X ), we see that ϕ is realized by Φ.
surjective; see [310, 1.11.6]. The proof is similar to Theorem 107 and we omit
the proof.
Theorem 108. Let (X, B, µ) be a σ-finite measure space, and let 1 ≤ p < ∞.
0
Suppose that X is a reflexive Banach space. Then Lp (X; X )∗ = Lp (X; X ∗ ).
Let X be a Banach space, and let Lip(R; X ) be the set of all Lipschitz
continuous functions f : R → X for which the quantity kf kLip(R;X ) ≡
sup |t−s|−1 kf (t)−f (s)kX is finite. If X = C, then abbreviate Lip(R; X )
−∞<s<t<∞
to Lip(R).
Lemma 109. Let X be a Banach space.
(1) Let f ∈ L1 (R; X ) ∩ C(R; X ) and g ∈ Lip(R; X ∗ ). Then the limit
∞
X
Lg (f ) ≡ lim hg(N −1 (j + 1)) − g(N −1 j), f (N −1 j)i
N →∞
j=−∞
2.3.4 Exercises
Exercise 52. Let (X, B, µ) be a measure space, and let X be a Banach space.
Suppose that {ϕj }∞ 1
j=1 ∈ L (X; X ) satisfies
∞ Z
X
kϕj (x)kX dµ(x) < ∞.
j=1 X
∞
P
Then show that ϕj (x) converges for µ-almost every x ∈ X. Modifying
j=1
∞ ∞
ϕj ∈ L1 (X; X ) and
P P
ϕj (x) over a set of measure zero, show that
j=1 j=1
Z X ∞ ∞ Z
X
ϕj dµ =
ϕj (x)dµ(x).
X j=1 j=1 X
Fundamental facts in functional analysis 107
Exercise 53. Let (X, B, µ) be a measure space. Also let X be a Banach space.
2.4 Notes
Section 2.1
General remarks and textbooks in Section 2.1
See the textbooks [241, Chapter 14], [310] and [409, Chapter 5] for func-
tional analysis, for example.
Section 2.1.1
We can consult [310, 409]. We refer to [310, Theorem 1.9.8] and [310,
Proposition 1.9.15] for Theorems 87 and 88, respectively or see [409, Theo-
rem 5.5.5], [409, Theorem 5.5.8] and [409, Theorem 5.5.9] for Hahn–Banach
extensions, Theorems 85, 86 and 87, respectively.
See [239] as well.
See [310, 2.6.18 and 2.6.19] for Theorem 89 for the Banach Alaoglu
theorem.
Section 2.1.2
Many people strengthened the triangle inequality. We employed one of
them in Theorem 90; see [300, Theorem 1].
108 Morrey Spaces
Section 2.1.3
Section 2.1.3 heavily depends on Lions and Peetre [280]. See [280, Chapter
premier] for the definition of sum spaces and intersection spaces. We refer to
[280, 3. I] for Theorem 92.
Section 2.2
General remarks and textbooks in Section 2.2
There are many textbooks on Hilbert spaces. Any one of them will suffice
for this book. Here we content ourselves with the book by Yoshida [465].
Section 2.2.1
Beigböck, Schachermayer and Veliyev considered a variant of Komlós’ the-
orem. See [26, §2.1] for Proposition 93, while the original Komlós theorem is
due to Komlós [244].
Section 2.2.2
Cotlar proved the “so called” Cotlar’s lemma, Theorem 94, in [88].
Section 2.3
General remarks and textbooks in Section 2.3
The Bochner integral is named after Salomon Bochner [38]. We refer to
[81, Appendix E] and [465] for Bochner integrals. See also the lecture note
[223, Chapter 0, §I].
Section 2.3.1
Pettis characterized the strong measurability [357, Theorem 1.1]; see The-
orem 96. We followed Pettis [357, Definition 2.1] when we defined the integral
of functions assuming their values in a Banach space; see Definition 37.
Section 2.3.2
See the paper by Bochner [38, §3] for Theorem 103. The completeness of
the Bochner integrable spaces can be found in [357, Theorem 4.2]; see Theorem
104. Pettis obtained a dense subspace; see [357, Theorem 4.3].
Section 2.3.3
See the paper by Bochner [38, §4] for Theorem 106. Pettis considered the
duality with respect to Bochner integral [357, Theorem 3.3]; see Theorem 107.
Calderón specified the dual space of L1 (R; X ) [52, §32.1]; see Lemma 109.
Chapter 3
Polynomials and harmonic functions
This chapter collects some preliminary facts. The readers may skip Chapter
3. We indicate where we use each theorem in the other parts of this book.
As a preliminary step, in Section 3.1 we collect some fundamental facts
on polynomials. Section 3.2 is the main part of Chapter 3, where we consider
harmonic functions.
109
110 Morrey Spaces
where Ω ∈ (M+ (Rn ) ∩ Cc∞ (Rn )) \ {0} is a fixed function supported on V . Set
U ≡ {P ∈ Pl (Rn ) : P (0) = 0} and
Z
U ⊥ ≡ R ∈ L1 (V ) : R(x)P (x)Ω(x)dx = 0, P ∈ U .
Rn
Polynomials and harmonic functions 111
Z
Then any function R ∈ U ⊥ ∩ Pl (Rn ) \ {0} satisfies R(x)Ω(x)dx 6= 0;
V
n
otherwise R is orthogonal to any element in Pl (R ) with respect to this
inner
Z product, that is, R = 0. Choosing R suitably, we can assume that
R(x)Ω(x)dx = 1. If we set ω ≡ R · Ω, then we have the desired function.
V
A passage to higher dimension can be done with the tensor product. Thus, it
∞ n ⊥ n
R there exists ψ ∈ C n (R ) ∩ Pk (R ⊥) supported
follows that in a cube (−1, 1)n
n
such that ψ(x)dx = 1 and 2 ψ(2·) − ψ ∈ Pk (R ).
Rn
We start with the inequality without the moment condition. Among the
convolutions of functions, we consider those generated by non-negative func-
tions with polynomial decay. When we handle integral operators, we occasion-
ally want to regard h·i−n−1 and χ[0,1]n as the same functions. Although the
former spreads more than the latter, the effect away from the origin of the
former is not so strong. This idea can be quantified in the following theorem:
112 Morrey Spaces
to complete the proof of (3.1). When |x| ≤ 10, we use kf ∗gkL1 ≤ kf kL∞ kgkL1
for f ∈ L∞ (Rn ) and g ∈ L1 (Rn ) and the result follows.
|x|
Hence, it can be assumed that |x| ≥ 10. Write M (x; t) ≡ −1 + log2 and
t
[M (x;t)]
X 1 Z hx − yi−λ dy
I ≡ n
,
j=1
t B(2j t) 2λj
∞
hx − yi−λ dy
Z
X 1
II ≡ .
tn B(2j t) 2λj
j=[1+M (x;t)]
Then
Z D y E−λ ∞ Z D y E−λ
1 −λ
X 1
hx − yi dy = hx − yi−λ dy
tn Rn \B(t) t j=1
t n
B(2 j t)\B(t) t
∞ Z
X 1 1
. n λj
hx − yi−λ dy = I + II.
j=1
t j
B(2 t) 2
|x|
Note that |x| ≥ max(10t, 2j+1 t) as long as 1 ≤ j ≤ M (x; t) = −1 + log2 .
t
j
Thus, |x| ≥ 10t, y ∈ B(2 t) and 1 ≤ j ≤ M (x; t), provided 2|x − y| ≥ |x|.
Consequently,
[M (x;t)] Z D x E−λ 1 [M (x;t)]
X 1 X 1 1
I≤ dy . hxi−λ j(λ−n) . .
j=1
tn B(2j t) 2 2λj
j=1
2 hxiλ
Polynomials and harmonic functions 113
Meanwhile,
∞ ∞
tλ−n
Z
X 1 dy X 1
II ≤ ∼ . .
tn Rn 2λj hx − yiλ 2λj tn hxiλ
j=[1+M (x;t)] j=[1+M (x;t)]
X 1 α
a(x − t y) − ∂ a(x)(−ty)α . |t y|N hxi−λ (3.6)
α!
α∈N0 n ,|α|≤N −1
114 Morrey Spaces
Changing the variables: r 7→ r−1 on the integral on the right-hand side yields
Z t−1 Z ∞
rn+N −1 (1 ∧ r−λ )dr = r−n−N −1 (1 ∧ rλ )dr.
0 t
Consequently,
Z X 1 α
a(x − t y) − ∂ a(x)(−ty)α η(y)dy
B(t−1 ) α!
α∈N0 n ,|α|≤N −1
Z 2
dr
. .
t rN +n+1−λ
X 1 α
a(x − t y) − ∂ a(x)(−ty)α
α!
α∈N0 n ,|α|≤N −1
X 1 α
≤ |a(x − t y)| + ∂ a(x)(−ty)α
α!
α∈N0 n ,|α|≤N −1
−λ
+ |t y|N −1 hxi−λ
. hx − t yi (3.7)
√
by Peetre’s inequality, hx + yi ≤ 2hxihyi and (3.3). Hence
Z
a(y)η(x − y; t)dy
Rn \B(t−1 )
Z
hx − t yi−λ + |t y|N −1 hxi−λ hyi−λ dy.
.
Rn \B(t−1 )
Recall also that λ > N + n − 1. With this in mind we change the variables:
y 7→ z = ty to obtain
Z Z
N −1−λ λ−N −n+1
|y| dy = t |z|N −1−λ dz ∼ tλ−N −n+1 .
Rn \B(t−1 ) Rn \B(t−1 )
Consequently,
Z Z
|t y|N −1 hyi−λ dy < tN −1 |y|N −1−λ dy
Rn \B(t−1 ) Rn \B(t−1 )
' tλ−n
Z 2
. tN rλ−N −n−1 dr.
t
Hence
Z X 1 α
a(x − t y) − ∂ a(x)(−ty)α η(y)dy
Rn \B(t−1 ) α!
α∈N0 n ,|α|≤N −1
Z 2
. tN hxi−λ rλ−N −n−1 dr.
t
12
X
kf kL2 . |ajν |2 . (3.8)
j∈Z,ν∈Zn
116 Morrey Spaces
. 1.
Thus,
X X Z
|ajν aj̃ ν̃ | ϕjν (x)ψj̃ ν̃ (x)dx
j∈Z,ν∈Zn j̃∈Z,ν̃∈Zn Rn
X X Z
≤2 |aj̃ ν̃ |2 ϕjν (x)ψj̃ ν̃ (x)dx
j∈Z,ν∈Zn j̃∈Z,ν̃∈Zn Rn
X
. |ajν |2 .
j∈Z,ν∈Zn
Lemma 115. Let k ∈ N0 ∪ {−1}, q ≥ 1 and A ∈ (0, 1). Also let x0 ∈ Ω and
0 < ρ < diam(Ω). If E is a measurable subset in Ω(x0 ; ρ) such that |E| ≥
Z q1
β 1 q
A|B(ρ)|, then |∂ P (x0 )| .β,A |P (x)| dx for all P ∈ Pk (Rn )
ρn+|β|q E
and β ∈ N0 n .
It should be noted that if E = B(x0 , ρ), then Lemma 115 is trivial from
the dilation and translation.
Proof By the translation and the dilation we may assume that B(x0 , ρ) =
B(1). The symbol Tk denotes the set of all polynomials in Pk (Rn ) of the form
X
P (x) = aα xα
|α|≤k
with X
|aα |2 = 1. (3.10)
|α|≤k
Let us define Z
γ(A) ≡ inf |P (x)|q f (x)dx.
P ∈Tk ,f ∈F B(1)
Let us prove that the infimum is attained;
Z
γ(A) = min |P (x)|q f (x)dx.
P ∈Tk ,f ∈F B(1)
This means that f −Pk (·; x0 , ρ, f ) and f −Qk (·; x0 , ρ, f ) are linearly dependent
thanks to Theorem 90 and the constant K satisfying f −Pk (·; x0 , ρ, f ) = K(f −
Qk (·; x0 , ρ, f )) is positive. Since f − Pk (·; x0 , ρ, f ) and f − Qk (·; x0 , ρ, f ) have
the same Lq (Ω(x0 , ρ))-norm, we see that K = 1, implying that Pk (·; x0 , ρ, f )
is unique.
We prove the existence. Set
X aα
P (x) ≡ (x − x0 )α (x ∈ Rn ).
α!
|α|≤k
Observe that 21
X
|aα |2 ∼ kP kLq (x0 ;ρ)
|α|≤k
again thanks to the fact that the norms in a finite dimensional normed space
Pk (Rn ) are equivalent. Thus, when a sequence {Pl }∞ n
l=0 ∈ Pk (R ) satisfies
Z ! q1
q
inf kf − P kLq (Ω(x0 ;ρ)) = lim |f (x) − Pl (x)| dx ,
P ∈Pk (Rn ) l→∞ Ω(x0 ;ρ)
Polynomials and harmonic functions 119
X
Pl (x) = al,α xα (x ∈ Rn ).
|α|≤k
3.1.5 Exercises
Exercise 56. Prove dimC (Ṗk (Rn )) = k+n−1 Cn−1 by considering the number
of the integer solutions of x1 + x2 + · · · + xn = k.
Exercise 57. Prove Proposition 117 as follows:
(4) Use (1) to show that |kB (x, y)| . |B|−1 for all x, y ∈ Rn .
Example 57. We have Ḣ0 (Rn ) = Ṗ0 (Rn ), H1 (Rn ) = P1 (Rn ) and Ḣ1 (Rn ) =
Ṗ1 (Rn ).
We are interested in the structure of these polynomial linear spaces. For a
start, we investigate the image of ∆ : Pk (Rn ) → Pk−2 (Rn ).
Lemma 118. The Laplacian ∆ maps Pk (Rn ) onto Pk−2 (Rn ) for all k ∈ N.
Proof Since the polynomials of the form (a1 x1 + a2 x2 + · · · + an xn )k ,
a1 , a2 , . . . , an ∈ R span Pk (Rn ), we have only to show that such a polynomial
is realized as the image of ∆ : Pk (Rn ) → Pk−2 (Rn ). Since
∆(a1 x1 + a2 x2 + · · · + an xn )k+2
= (k + 2)(k + 1)(a1 2 + · · · + an 2 )(a1 x1 + a2 x2 + · · · + an xn )k ,
this is clear.
Since Pk (Rn ) is a finite dimensional linear space, so are Hk (Rn ) and
Ḣk (Rn ). In higher dimensions, it is difficult to take a basis. Consequently,
we do not go in this direction here. However, we will need to know how large
the spaces Hk (Rn ) and Ḣk (Rn ) are.
Corollary 119. Let k ∈ N0 . Then dimC (Hk (Rn )) = k+n Cn − k+n−2 Cn =
O(k n−1 ) and dimC (Ḣk (Rn )) = k+n−1 Cn−1 − k+n−3 Cn−1 = O(k n−2 ).
Proof We will deal with the nonhomogeneous spaces; the homogeneous
spaces can be handled similarly. See Exercise 59. Using the definition of
Hk (Rn ) and Lemma 118, we calculate
Consequently, Ḣk (Rn ) and Im(M : Pk−2 (Rn ) → Pk (Rn )) are orthogonal.
To check that Ḣk (Rn ) and Im(M : Pk−2 (Rn ) → Pk (Rn )) span Pk (Rn ),
we can compare the dimension of these spaces using Corollaries 110 and 119.
122 Morrey Spaces
while
m
∂ ∂
hf, f iP = +i (x − iy)m = 2m m!.
∂x ∂y
for all x ∈ Rn .
In the proof, we use the orthogonal group of degree n; O(n) denotes the
set of all orthogonal matrices.
Polynomials and harmonic functions 123
which is called zonal harmonic of degree k with pole at x, then Y (z) = hY, Zz iH
for all Y ∈ Ḣk . By the uniqueness of the Riesz representation theorem for
Hilbert spaces, we see that
Ḣk (Rn ))
dimC (P
for all A ∈ O(n). Thus, |Yjk (z)|2 is a constant. To determine the
j=1
precise value of this constant, we integrate over S n−1 and use the fact that
dim (Ḣ (Rn ))
{Yjk }j=1C k is an orthonormal system.
dim (Ḣk (Rn ))
The family {Yjk }j=1C has the following growth in L∞ (S n−1 )-norm.
s
dimC (Ḣk (Rn )) n
Corollary 122. |Yjk (x)| ≤ = O(k 2 −1 ) for all x ∈ S n−1 ,
ωn−1
k ∈ N0 and j = 1, 2, . . . , dimC (Ḣk (Rn )).
Proof Simply use Proposition 121:
In particular,
Z Z
k(n + 2k − 2) |P (x)|2 dσ(x) = |grad(P )(x)|2 dσ(x) (3.2)
S n−1 S n−1
for k ≥ 1.
124 Morrey Spaces
Ḣk (Rn ))
dimC (P
Proof Let x ∈ Rn be fixed. Since P = hP, Yjk iH Yjk , we obtain
j=1
Ḣk (Rn ))
dimC (P
grad(P )(x) = hP, Yjk iH grad(Yjk )(x). Notice that there exists a
j=1
constant A = Ak > 0 such that
dimC (Ḣk (Rn ))
X
|grad(Yjk )(x)|2 = A|x|2k−2 . (3.3)
j=1
Ḣk (Rn ))
dimC (P Ḣk (Rn ))
dimC (P
Indeed, |grad(Yjk )(M x)|2 = |grad(Yjk )(x)|2 for all
j=1 j=1
M ∈ O(n). To calculate the precise value of A, we integrate (3.3) over S n−1 :
By (3.2), we obtain
x4 + y 4 − 6x2 y 2 x4 + y 4 − 6x2 y 2
∆ 2 2 2
= −16 .
(x + y ) (x2 + y 2 )3
If we write this by the spherical coordinate, then we obtain the desired result.
3.2.4 Exercises
Exercise 58. Show that P1 (Rn ) = Ḣ0 (Rn ) ⊕ Ḣ1 (Rn ). Hint: Can we write
these spaces explicitly ?
Exercise 59. Let k, n ∈ N.
(1) Fix n ∈ N. Prove dimC (Ḣk (Rn )) = k+n−1 Cn−1 − k+n−3 Cn−1 = O(k n−2 )
as k → ∞ by reexaming the proof of Corollary 119.
(2) Prove Ṗk (Rn ) = Ḣk (Rn ) ⊕ Im(M : Ṗk−2 (Rn ) → Ṗk (Rn )) from the
decomposition Pk (Rn ) = Hk (Rn )⊕Im(M : Pk−2 (Rn ) → Pk (Rn )). Hint:
How do M in Proposition 120 and ∆ act on Pk (Rn ) ?
Exercise 60 (Euler’s formula). Let α ∈ R and f ∈ C ∞ (Rn \ {0}) be a
function such that f (tx1 , tx2 , . . . , txn ) = tα f (x1 , x2 , . . . , xn ) for all t > 0 and
(x1 , x2 , . . . , xn ) ∈ Rn \ {0}. Then, by differentiating the above equation in t
n
xj fxj (x) = αf (x) for all x = (x1 , x2 , . . . , xn ) ∈ Rn \ {0}.
P
show that
j=1
3.3 Notes
Section 3.1
General remarks and textbooks in Section 3.1
See [85] for example.
Section 3.1.1
See [85, Theorem 3.8] for the expression of P (D)Q, Lemma 111. Janson,
Taibleson and Weiss used the Gram–Schmidt polynomial. See [220, p.111] for
Proposition 117, for example.
Section 3.1.2
See [129, Lemma 3.3].
Polynomials and harmonic functions 127
Section 3.1.3
We can find some quantitative information on the relation between the
Lq (Rn )-norm and the L∞ (Rn )-norm of the derivative; see [59, Lemma 2.I] for
Lemma 115.
Section 3.1.4
We followed [59, (3.1)] to find the best approximation; see Lemma 116.
Section 3.2
General remarks and textbooks in Section 3.2
The properties and uses of spherical harmonics can be found in the books
by [415, 417] and [410, §3.5].
Section 3.2.1
Corollary 119, which calculates dimC (Hk (Rn )), can be found in [127,
(2.55), (2.56)]. Proposition 120, which decomposes Pk (Rn ) can be found in
[127, (2.49)].
Section 3.2.2
Proposition 121 can be found in [127, (2.57)].
Section 3.2.3
The functional equation in Proposition 125 can be found in [127, (2.62)].
The self-adjointness of the spherical Laplacian can be found in [76, Lemma
2.9]; see Proposition 126.
Chapter 4
Various operators in Lebesgue spaces
One of the primary aims in this book is the boundedness properties of the
integral operators. Here, we collect the properties of the linear operators acting
on Lebesgue spaces. The boundedness obtained here will be used for analysis
of operators acting on Morrey spaces. We take up the following operators.
(1) Hardy–Littlewood maximal operators, including the related maximal
operators (Sections 4.1 and 4.2).
(2) Fractional maximal operators (Section 4.3).
(3) Singular integral operators (Section 4.4).
(4) Fractional integral operators (Section 4.5).
The first three operators are used for the purpose of controlling the remaining
two operators and these two operators are mainly used to express the solution
of the differential equations.
129
130 Morrey Spaces
for all R > 0. In particular, M χ[−1,1] (t) = min 1, 2(1 + |t|)−1 for t ∈ R.
Proof Since we are considering the average of the function of χ(−R,R) , we
have M χ(−R,R) (t) ≤ 1 for all t ∈ R. Let |t| < R. Then
Z t+r Z t+r
1 1
M χ(−R,R) (t) ≥ lim χ(−R,R) (y)dy = lim dy = 1.
r↓0 2r t−r r↓0 2r t−r
1 |x − c(R)|
≤ M Q χR ≤ 1, 1≤1+ ≤ 1 + 3n.
3n `(R)
132 Morrey Spaces
(2) Let x ∈ 3l+1 R \ 3l R for some l ∈ N. In this case, we can show that
1 4n 3l |x − c(R)|
≤ M Q χR (x) ≤ , ≤1+ ≤ 1 + n · 3l+1 .
3(l+1)n 3ln 2 `(R)
Remark that cubes can be replaced by balls and that 5Q can be readily
replaced by κQ for any κ > 1 once the implicit constant can be allowed to
depend on κ. The right inequality is clear but we frequently use it. However,
it will turn out that we need to pay attention to the difference between the
summation and the supremum.
Proof The right inequality is clear, so we prove the left equivalence. One
inequality is clear from the definition of M . We write out M [χRn \5Q f ](x) in
full: Z
χR (x)
M [χRn \5Q f ](x) = sup |f (y)|dy.
R∈Q |R| R\5Q
Z
In order that χR (x) |f (y)|dy be not zero, we need to have x ∈ R and
R\5Q
R \ 5Q 6= ∅. Thus, R meets both Q and Rn \ 5Q. If R ∈ Q is a cube that meets
both Q and Rn \ 5Q, then `(R) ≥ 2`(Q) and 2R ⊃ Q. Thus, the desired result
follows.
Various operators in Lebesgue spaces 133
k Z
X 1
|Φ ∗ f (x)| ≤ aj |B(x, rj )| |f (y)|dy ≤ kΦkL1 M f (x).
j=1
|B(x, rj )| B(x,rj )
A passage to the limit, makes this inequality valid for any integrable, contin-
uous and radial decreasing function Φ.
Example 64. We list some examples of functions to which Theorem 131 is
applicable.
(1) Φ = χB(1) .
|x|2
1
(2) Gaussian: Φ(x) = p exp − for x ∈ Rn .
(4π)n 4
(3) Let j ∈ N0 and m > n. Define ηj,m (x) ≡ 2jn (1 + 2j |x|)−m for x ∈ Rn .
The function ηj,m is called the η-function.
Thus, the balls B1 , B2 , . . . , BK satisfy the condition of the theorem and the
proof is complete.
Usually, we mean by the “5r-covering lemma” the following theorem (due
to Wiener), whose proof we omit due to similarity:
Theorem 133 (Infinite Vitali’s 5r-covering lemma). Suppose that we have a
family {Bλ }λ∈Λ of balls of general cardinality. If the radii of the balls in the
family is bounded from above by a positive constant, then we can find Λ0 ⊂ Λ
such that
(1) {Bλ }λ∈Λ0 is disjoint,
Various operators in Lebesgue spaces 135
L
X
χBj ≤ 1, (4.4)
j=1
We can modify the argument above to have an estimate for the maximal
function of the measure.
Example 66. Let µ be a non-negative finite Radon measure and define (tem-
porarily) the Hardy–Littlewood maximal operator of measures by
µ(Q)
M µ(x) ≡ sup : Q ∈ Q, c(Q) = x (x ∈ Rn ). (4.7)
|Q|
for λ > 0. Let δ0 be the measure massed at the origin. It is also noteworthy
that M δ0 (x) ∼ |x|−n for x ∈ Rn . The proof of (4.8) is left as an exercise; see
Exercise 63.
Although in Proposition 128 we explained that the function M f is never
integrable, we still substitute this inequality for the weak variant.
Theorem 138. Let 0 < p < 1, E ⊂ Rn be a measurable set with finite
1
measure, and let f ∈ L1 (Rn ). Then kM f kLp (E) . |E| p −1 kf kL1 .
The idea that lies behind it is the use of the finiteness of the measure
and the weak-(1, 1) boundedness of operators. This estimate is called the Kol-
mogorov inequality. Note that this type of estimate is valid for any weak-(1, 1)
bounded linear operator.
Proof This is a consequence of
Z Z ∞
M f (x)p dx = pλp−1 |{x ∈ E : M f (x) > λ}|dλ,
E 0
We also call the above inequality the weighted weak (1, 1)-inequality. In
Proposition 139, by letting w ≡ 1, we can recover Theorem 133.
Proof By the monotone convergence theorem, it can be assumed that w
is bounded. For λ > 0, set Eλ ≡ {M f > λ}. In view of the inner regularity of
the measure w(x)dx, it suffices to show
3n
Z
w(K) ≤ |f (x)|M w(x)dx
λ Rn
w(3Bj )
≤ 3n inf M w(y). (4.12)
|Bj | y∈Bj
Various operators in Lebesgue spaces 139
M
P
By (4.11), we have w(K) ≤ w(3 Bj ). If we use (4.10), (4.12), and (4.9),
j=1
in that order, then
M Z
1X w(3 Bj )
w(K) ≤ |f (x)|dx ·
λ j=1 Bj |Bj |
M Z
n X
3
≤ |f (x)|M w(x)dx
λ j=1 Bj
3n
Z
≤ |f (x)|M w(x)dx.
λ Rn
We use (1.4) for the function M f and the measure w(x)dx. If we change
variables: σ = 2λ, then
!
Z Z Z M f (x)
M f (x)p w(x)dx = p σ p−1 dσ w(x)dx
Rn Rn 0
2−1 M f (x)
Z Z !
= 2p p λp−1 dλ w(x)dx. (4.13)
Rn 0
3n
Z
|{M f > λ}| ≤ |f (x)|dx.
λ {M f >λ}
Going through the same argument using the Layer-Cake formula, we have
Z Z
M f (x)p dx ≤ 3n p0 M f (x)p−1 |f (x)|dx
Rn Rn
Z p−1
p
Z p1
≤ 3n p0 M f (x)p dx |f (x)|p dx .
Rn Rn
Since f ∈ L∞ n
c (R ), we can arrange this inequality to obtain the desired con-
clusion.
About the constant obtained in Theorem 141, we have the following clar-
ifying remark:
Remark 5. Theorem 141 is very important and we use it many many times
in this book. Usually, it is not necessary to learn the constant 3n p0 by heart.
However, it is not so difficult to keep track of the constants in the proof.
Example 67. Let p ∈ (1, ∞) and r ∈ [1, ∞). Let T : L∞ n 0 n
c (R ) → L (R ) be a
sublinear operator such that T can be extended to a bounded operator both
Various operators in Lebesgue spaces 141
and let c∗ (r) be the nminimal value of c(r) in the inequality above. Then we
claim c∗ (r) = c∗ (1)r p , which redues matters to the case where r = 1. Indeed,
we calculate
Z Z
n
M f (x)p dx rp M f (ry)p dx
B(r) B(1)
c∗ (r)p = sup Z = sup Z
f |f (x)|p f |f (ry)|p
n
dx n
dy
Rn (|x| + r) Rn (|y| + 1)
Z Z
n
rp M (f (r·))(y)p dx M g(x)p dx
B(1) n B(1)
= sup = r p sup Z
|f (ry)|p |g(x)|p dx
Z
f g
n
dy n
Rn (|y| + 1) Rn (|x| + 1)
∗ p n
= c (1) r .
q1 q1
∞
X X∞
M fj q .p,q |fj |q .
j=1 j=1
Lp Lp
The space `q (Lp , Rn ) is the set of all collections {fj }j∈J for which the
quantity k{fj }j∈J k`q (Lp ) is finite.
(2) For a sequence {fj }j∈J of functions we define a function k{fj }j∈J k`q :
Rn → [0, ∞] by
q1
X
k{fj }j∈J k`q ≡ |fj |q
j∈J
The space Lp (`q , Rn ) is the set of all collections {fj }j∈J for which the
quantity k{fj }j∈J k`q (Lp ) is finite.
(4) For a measure space (X, B, µ), define Lp (`q , µ) and `q (Lp , µ) similarly.
These two norms are called vector-valued norms. A natural modification is
made in the above when q = ∞.
Example 70. We extend Hölder’s inequality. Let 1 ≤ p ≤ ∞ and 1 ≤
q ≤ ∞. Let (X, B, µ) be a measure space. By original Hölder’s inequality
k{aj bj }∞ ∞ ∞ ∞ ∞
j=1 k`1 ≤ k{aj }j=1 k`q k{bj }j=1 k`q0 for sequences {aj }j=1 and {bj }j=1
0
and by the one kF · GkL1 (µ) ≤ kF kLp (µ) kGkLp0 (µ) for F, G ∈ L (µ),
k{Fj Gj }∞ ∞ ∞
j=1 kL1 (`1 ,µ) ≤ k{Fj }j=1 kLp (`q ,µ) k{Gj }j=1 kLp0 (`q0 ,µ) .
{fj }∞
j=1
+ n
⊂ M (R ), then we can arrange that {gj }∞
j=1 ⊂ M (Rn ).
+
∞ Z
X
k{M fj θ }∞
j=1 kLs (`t ) = M fj (x)θ gj (x)dx, k{gj }∞
j=1 kLs0 (`t0 ) = 1.
j=1 Rn
The relation for s, t is reversed if we pass to the conjugate index; s0 > t0 . Thus,
by Step 2, k{M gj }∞ j=1 kLs0 (`t0 ) . 1. Consequently, we have
Z Z
k{M fj (x)}∞ j=1 k`
p
q dx . k{fj (x)}∞ p
j=1 k`q dx.
Rn Rn
Vλ ≡ x ∈ Rn : k{M fj }∞
j=1 k`u (x) > λ (4.16)
Z
. λ−θ k{fj }∞ θ ∞
j=1 k`u (x) χ(0,λ) (k{fj }j=1 k`u (x))dx
Rn
Z
†
θ†
+ λ−θ k{fj }∞ ∞
j=1 k`u (x) χ[λ,∞] (k{fj }j=1 k`u (x))dx.
Rn
Corollary 147. For any Young function Φ : [0, ∞) → [0, ∞) and f ∈ L0 (Rn )
M f . MΦ f .
Proof Simply compare the definition of M f and MΦ f using Lemma 55.
As before, the maximal operator MΦ is bigger than the identity operator.
Corollary 148. Suppose that f ∈ L1loc (Rn ) and that Φ is a Young function.
Then |f (x)| . MΦ f (x) for almost all x ∈ Rn .
Proof This is clear since |f (x)| ≤ M f (x) . MΦ f (x) for any Lebesgue
point x of f ; see Lemma 142 and Corollary 147.
We investigate the boundedness property of the Orlicz maximal operators.
We first establish the following weak-type inequality:
Proposition 149. Suppose that f ∈ L1loc (Rn ) and that Φ : [0, ∞) → [0, ∞)
is a Young function. Then for all λ > 0,
3n
|f (x)|
Z
|{y ∈ Rn : MΦ f (y) > λ}| ≤ Φ dx.
λ Rn λ
Proof By the inner regularity (1.1) of the Lebesgue measure, it suffices
to show
|f (x)|
Z
n
|K| ≤ 3 Φ dx (4.19)
Rn λ
for any compact set K contained in {MΦ f > λ}.
Let x ∈ K be fixed. By the definition of M , there exists Qx ∈ Q containing
x such that
kf kΦ;Qx > λ. (4.20)
By expanding Qx slightly, we may assume x ∈ Int(Qx ).
By the compactness of K, there exists a finite collection x1 , x2 , . . . , xJ of
J
S
points such that K ⊂ Qxj . To simplify the notation, write Qj ≡ Qxj . By
j=1
Theorem 132, if we relabel the Qj ’s, for some L ≤ J, the collection {Qj }L
j=1
is disjoint, that is,
L
X
χQj ≤ 1, (4.21)
j=1
and
J
[ L
[
K⊂ Qj ⊂ 3 Qj . (4.22)
j=1 j=1
So far, we have seen how the operator MΦ arises. Here, we are interested
in showing that MΦ is bounded on Lp (Rn ) under some suitable conditions.We
propose the following condition:
Definition 49. Let 1 ≤ Zp < ∞. A Young function Φ : [0, ∞) → [0, ∞)
∞
Φ(t)
belongs to the class Bp if dt < ∞. In this case write Φ ∈ Bp .
1 tp+1
Example 73. Let 1 < p < ∞.
(1) Since Φ(t) ≥ Φ(1)t for any Young function Φ : [0, ∞) → [0, ∞) and
t ≥ 0, we see that B1 = ∅.
(2) Let Φ1 (t) = tq for 1 ≤ q < ∞. Then Φ1 ∈ Bp if and only if q < p.
(3) Let 1 ≤ q < ∞ and r ∈ R. Let Φ2 be a Young function such that
Φ2 (t) ∼ tq log(3 + t) for t 1. Then Φ2 ∈ Bp if and only if q < p.
The class Bp is important because this class characterizes the boundedness
of MΦ on Lp (Rn ) for 1 < p < ∞.
Theorem 152. Let 1 < p < ∞, and let Φ : [0, ∞) → [0, ∞) be a bijective
Young function. Then MΦ is bounded on Lp (Rn ) if and only if Φ ∈ Bp .
Proof Suppose Φ ∈ Bp . Let C0 ≡ 1 + Φ−1 (1). Then by Theorem 5 and
Proposition 150,
Z Z ∞
p
MΦ f (y) dy ∼ λp−1 |{y ∈ Rn : MΦ f (y) > C0 λ}|dλ
Rn 0
Z ∞ !
|f (x)|
Z
p−2
. λ Φ dx dλ
0 {y∈Rn : |f (y)|>λ} λ
Z ∞ Z
Φ(t)
= dt · |f (y)|p dy.
1 tp+1 Rn
implying Φ ∈ Bp .
Various operators in Lebesgue spaces 151
Let Q̃j ∈ D(Q) be the dyadic parent of Qj , the unique dyadic cube contain-
ing Qj with side-length twice that of Qj . Thus, from Example 47, we have
kf kΦ, Qj ≤ 2n kf kΦ, Q̃j and by the maximality of Qj kf kΦ, Qj ≤ 2n kf kΦ, Q̃j ≤
2n t. It follows from the definition of the
mean
Luxemberg
normand the
fact
↑ f f
that Φ ∈ M [0, ∞) satisfies 1 ≥ mQj Φ ≥ mQj Φ .
kf kΦ, Qj 2n t
Z
f (x)
This yields that |Qj | ≥ Φ dx. Thus, we obtain the left inequality
Qj 2n t
S
of (4.25) by observing further that Qj ⊃ {x ∈ Q : f (x) > t}, which holds
j
by the Lebesgue differential theorem.
We use the following criteria to check that MΨ is bounded.
Lemma 154. Let Ψ : [0, ∞) → [0, ∞) be a normalized Young function. Sup-
pose that normalized Young functions Φ, Φ1 , Φ2 : [0, ∞) → [0, ∞) fulfill, for
some positive constants C1 and C2 ,
Z t
t
Φ1 (C1 t) − 1 ≤ Ψ Φ0 (s)ds ≤ Φ2 (C2 t) (t > 1).
1 s
This yields, by the definition of the mean Luxemburg norm, for some C > 1
Z
1 MΨ [f χQ ](x)
Φ dx ≤ 1,
|Q| Q Ckf kΦ2 , Q
We now claim that then kf kΨ, Q ≤ 1. If it were not, then we must have
that the above integral mean is bigger than one, which contradicts our
normalization above, by virtue of the fact that for almost every x ∈ Q
kf kΨ, Q ≤ MΨ [f χQ ](x).
We claim
kf kΦ1 , Q . 1. (4.26)
In fact, since kf kΨ, Q ≤ 1,
Z
|Q| ≥ Φ (MΨ [f χQ ](x)) dx
Q
Z ∞
= |{x ∈ Q : MΨ [f χQ ](x) > t}|Φ0 (t)dt.
0
154 Morrey Spaces
By our assumption,
f (x)/2n
Z Z !
f (x) 0
|Q| ≥ Ψ Φ (t)dt dx
Q∩{f >2n } 1 2n t
Z
≥ Φ1 (2−n C1 f (x))dx − |Q ∩ {f > 2n }|
Q∩{f >2n }
Z
≥ Φ1 (2−n C1 f (x))dx − (Φ1 (C1 ) + 1)|Q|.
Q
Z
Φ1 (C1 ) + 2 1 f (x)
Let C ≡ . Then the above estimate reads Φ1 dx ≤
2−n C1 |Q| Q C
1 and, hence, (4.26) follows. The proof is now complete.
Example 75. Let Q ∈ Q, and let f ∈ L0 (Q). Then mQ (M [χ3Q f ]) .
kf kL log L,3Q from Lemma 75 with Φ(t) = Ψ(t) = t, t ≥ 0.
As a special case of Φ1 (t) = Φ2 (t) = t log(e + t) and Φ(t) = Ψ(t) = t,
t ≥ 0, we have the equivalent expression of the composition of the maximal
operators.
Corollary 155. One has M ◦ M f ' ML log L f for any f ∈ L0 (Rn ).
Proof Simply combine Example 65 and Lemma 75 with Φ(t) = Ψ(t) = t
and Φ1 (t) = Φ2 (t) = t log(e + t), t ≥ 0.
Example 76. For any f ∈ L0 (Rn ) and λ > 0,
|f (x)| |f (x)|
Z
n
|{x ∈ R : M ◦ M f (x) > λ}| . log 3 + dx
Rn λ λ
Proof We have already proven that (4.28) implies (4.27); see Lemma 154.
First, we verify (4.27) implies (4.28). Suppose
Z that (4.27), namely,
assume
MΨ [f χQ ](x)
that there exists a constant C0 ≥ 1 such that Φ dx ≤ |Q|.
Q C0 kf kΦ;Q
We now claim that then kf kΨ, Q . C0 kf kΦ;Q . If it were not, then we would
have that the above integral mean is bigger than one, which contradicts our
normalization above, by virtue of the fact that kf kΨ, Q ≤ MΨ [f χQ ](x) for
almost every x ∈ Q. Thus we have by Lemma 154
Z
MΨ [f χQ ](x)
|Q| ≥ Φ dx
Q C0 kf kΦ;Q
Z ∞
= |{x ∈ Q : MΨ [f χQ ](x) > C0 kf kΦ;Q · s}| Φ0 (s)ds
0
Z ∞Z
|f (x)|
≥ Ψ dx Φ0 (s)ds
1 Q∩{|f |>C0 kf kΦ;Q ·s} 2n C0 kf kΦ;Q · s
Z 2n C|fkf
(x)|
|f (x)|
Z kΦ;Q
0
≥ Ψ Φ0 (s)dsdx.
Q∩{|f |>2n C0 kf kΦ;Q } 1 2n C0 kf kΦ;Q · s
and that Z Z
ϕ(x) −β −β
β
dx ≥ 2 r ϕ(x)dx.
Rn (|x| + r) B(r)
If we decompose the integral dyadically in the second term in the right-hand
side and use Lemma 161 and
Z ∞ Z ! Z
d` −β
β ϕ(x)dx β+1 = r ϕ(x)dx,
r B(r) ` B(r)
Xm Z
. (2k ρ)γ dy
k=0 B(r)∩B(x,2k+1 ρ)
Xm
. (2k ρ)γ min(rn , 2(k+1)n ρn )
k=0
m
X
. ργ min(rn , ρn ) 2k(γ+n) ,
k=0
and the desired result follows since for any t ∈ R,
1
(t < 0),
1 − 2t
m
X
kt
2 ≤ m+1 (t = 0),
(m+1)t
2
k=0
(t > 0).
2t − 1
We also collect another type of estimate.
160 Morrey Spaces
Proof
(1) We decompose
Z
I≡ |x − y|γ (|y| + r)β dy
Rn \B(x,ρ)
∞ Z
X
= |x − y|γ (|y| + r)β dy
k=0 B(x,2k+1 ρ)\B(x,2k ρ)
Note that
∞
X Z
γ kγ
I≤ρ 2 (|y| + r)β dy
k=0 B(x,2k+1 ρ)\B(x,2k ρ)
∞
X n β
. ργ 2kγ 2k+1 ρ |x| + 2k+1 ρ + r
k=0
∞
X
β
= 2n+β ργ+n (|x| + ρ + r) 2k(γ+β+n)
k=0
γ+n β
∼ρ (|x| + ρ + r) .
4.1.9 Exercises
Exercise 61. Let f : R → R be a measurable function. Also let a > 0 and
g : R → [0, ∞) is a function such that g|[a, ∞) is decreasing and g|(−∞, a] is
increasing. Then show that kf ·gkL1 ≤ kgkL1 M f (a) by mimicking the proof of
Proposition 159. Here M denotes the uncentered Hardy–Littlewood maximal
operator.
Exercise 62. Let 0 < a < b < 1, and let Q be a cube. Then show that
M Q χbQ\aQ ≥ (bn − an )χQ by considering mQ (χbQ\aQ ).
Various operators in Lebesgue spaces 161
k
(1) Prove that |PQ(x,r) f (x)| . M f (x) using Proposition 117.
(2) Prove Lemma 143 by mimicking the proof of Lemma 142.
=: Ik + IIk .
Choose k ∈ Z ∩ (−∞, log2 ε). Note again that |Gk (x)| ≤ 2k+n almost every-
where by the maximality of the cubes Qkj and the Lebesgue differentiation
theorem. Since
Ik ≤ 2k+n kf kL∞ |{f > ε}| + εkgkL1 ,
Various operators in Lebesgue spaces 163
∞ X Z
X
|IIk | ≤ |f (x) − mQlj (f )| · |B l (x) − B l+1 (x)|dx
l=k j∈J l Qlj
∞
X XZ
n+l
≤3 2 |f (x) − mQlj (f )|dx.
l=k Qlj
j∈J l
∞
X Z
We deduce |IIk | ≤ 3 2 n+l
M ] f (x)dx ≤ 6kM ] f · M D gkL1 from
l=−∞ {M D g>2l }
(4.1). The proof is therefore complete.
We transform Theorem 165 into the form which we actually use.
Theorem 166 (Sharp maximal inequality). Let 1 < p < ∞. Then kf kLp .p
kM ] f kLp for f ∈ L0 (Rn ) with min(M f, 1) ∈ Lp (Rn ).
We really have to assume that min(1, M f ) ∈ Lp (Rn ) as the example of
the function f ≡ 1 shows.
Proof By truncation, we may suppose that f ∈ L∞ (Rn ). In this case,
M f ∈ Lp (Rn ). Hence by Corollary 148, we have f ∈ Lp (Rn ), so that (4.2) is
0
satisfied. Let w ∈ Lp (Rn ) ∩ M+ (Rn ) be a function with norm 1 satisfying
Z
kf kLp ≤ 2 f (x)w(x)dx.
Rn
WL1 (Rn ) = L1,∞ (Rn ), so that it is not always locally integrable. Hence,
we are interested in the maximal operator which yields something for any
measurable function. Recall that the decreasing rearrangement of f ∈ L0 (Rn )
is given as follows:
See Definition 19. We will consider a maximal operator which can be useful
for any measurable function.
For Q ∈ Q, the median of f ∈ L0 (Q), which is close to (χQ f )∗ 2−1 |Q| ,
(3) Assume otherwise; sup MED(f ; Q) > sup MED(g; Q). We choose β ∈
(sup MED(g; Q), sup MED(f ; Q)). Then
1
|{x ∈ Q : g(x) > β}| ≤ |{x ∈ Q : g(x) > Med(g; Q)}| ≤ |Q|.
2
Meanwhile,
1
|{x ∈ Q : g(x) < β}| > |Q| ≥ |{x ∈ Q : f (x) < β}|
2
≥ |{x ∈ Q : g(x) < β}|,
which is a contradiction.
(4) Assume otherwise, assume λ > f ∗ (2−1 a|Q|) or λ < −f ∗ (2−1 a|Q|). Then
if λ > f ∗ (2−1 a|Q|), then
1
|{x ∈ Q : f (x) ≤ f ∗ (2−1 a|Q|)}| ≤ |{x ∈ Q : f (x) < λ}| ≤ |Q|
2
since λ ∈ MED(f ; Q). Meanwhile, thanks to Lemma 29,
|{x ∈ Q : f (x) > f ∗ (2−1 a|Q|)}| ≤ |{x ∈ Q : |f (x)| > f ∗ (2−1 a|Q|)}|
≤ λf (f ∗ (2−1 a|Q|))
a
≤ |Q|,
2
which is a contradiction since |Q| > 21 |Q| + a2 |Q|.
If λ < −f ∗ (2−1 a|Q|), then we consider −f instead of f . Thanks to the
above argument, we obtain a contradiction.
where
Ej,k ≡ x ∈ Rn : 2−j (k − 1) ≤ f (x) < 2−j k .
Then
0 ≤ f (x) − sj (x) ≤ 2−j , Inf(MED(sj ; Q)) ≤ Med(f ; Q)
for every Q ∈ Q(Rn ). Let
∞ [∞
\ |Ej,k ∩ Q|
F ≡ x ∈ Ej,k : lim =1 .
j=1
|Q|→0,x∈Q |Q|
k=1
(2) The family E ≡ {Ekj }j∈N0 ,k∈Kj is disjoint, decomposes Q0 and satisfies
τ |Ekj | 1
|Qjk | ≤ ≤ |Ekj | for all j ∈ N0 and k ∈ Kj .
τ − 2n 2
Proof
(1) If j = 0, then the assertion is clear from the maximality. Let us suppose
j ≥ 1. The left inequality is a consequence of the fact that we chose Qjk
from Zj . If we consider the dyadic parent R of Qjk , then mR (h) ≤ γ0 τ j .
Otherwise, instead of Qjk R would have been chosen as an element of
Zj . Since mQj,k (h) ≤ 2n mR (h), we obtain the right inequality.
(2) A geometric observation shows that two dyadic cubes never intersect
unless one is not included in the other. We freeze j ∈ N and k ∈ Kj .
From the observation above, we have
[ n j+1 o
Dj+1 ∩ Qjk = Qk∗ : k ∗ ∈ Kj+1 , Qj+1
k ∗ ⊂ Qj
k
because Qj+1 j
k∗ ) Qk never happens thanks to the maximality
Z of Qjk and
the fact that τ ≥ 2n+1 . Note that |Qj+1
k∗ |γ0 τ
j+1
≤ h(x)dx, since
Qj+1
k∗
Qj+1
k∗
∗
∈ Zj+1 . If we add this estimate over Zall k ∈ Kj+1 such that
j+1
Qk∗ ⊂ Qjk , we obtain |Dj+1 ∩ Qjk |γ0 τ j+1 ≤ h(x)dx ≤ |Qjk |2n γ0 τ j .
Qjk
Consequently, if we consider the complement of Dj+1 , then we obtain
the desired result.
168 Morrey Spaces
Although the set of dyadic cubes has a good geometric property, there
are many dyadic cubes that dyadic cubes cannot cover: there does not exist a
dyadic cube R such that Q(1) cannot be covered by R. However, if we consider
2n translated families, we can have this property.
Definition 53 (Da ). Let a = (a1 , a2 , . . . , an ) ∈ {0, 1/3}n . The generalized
dyadic grid Da is defined by Da ≡ {Q : Q − (−1)log2 `(Q) `(Q)a ∈ D}. If
n = 1, then we write D+ ≡ D(1) .
and
1 1 1 5 5 4 4 11 11 7
..., − , , , , , , , , , , . . . ∈ D+ .
6 3 3 6 6 3 3 6 6 3
Furthermore, [−1, 1] ⊂ − 38 , 43 ∈ D+ .
The family Da , a ∈ {0, 1/3}n is useful because we can control any cubes
in the following sense:
Lemma 171. Let Q be any cube. Then there exist a = (a1 , a2 , . . . , an ) ∈
{0, 1/3}n and R ∈ Da ∩ Q] (Q) such that |R| ≤ 6n |Q|.
Proof We work in R; n = 1. By translation, we may assume Q = (a, b),
where 0 < a < b. Choose k0 ∈ Z such that 2−k0 −1 ≤ 3(b−a) < 2−k0 . If (a, b)∩
2−k0 Z = ∅, then (a, b) is contained in an interval of the form 2−k0 [m0 , m0 +1).
Thus, we may assume (a, b) ∩ 2−k0 Z 6= ∅. Let m0 2−k0 ∈ (a, b). Then m2−k0 ∈ /
(a, b); otherwise b − a ≥ 2−k0 and this is a contradiction. Then
Thus, if we choose R = (m0 2−k0 − 21−k0 /3, m0 2−k0 + 2−k0 /3) ∈ D+ , then we
have `(R) = 2−k0 ≤ 6(b − a) = 6`(Q).
Motivated by the generalized dyadic grid Da , we introduce the notion of
generalized dyadic grids.
Definition 54 (Generalized dyadic grid). A generalized dyadic grid D is the
set of all cubes of the form
x + [0, 2l )n (x ∈ Rn , l ∈ Z)
Let Ekj≡ Qjk\ Dj+1 . Then the family {Ekj }j∈N0 ,k∈Kj is pairwise disjoint and
decomposes Q0 . Consequently, we need only verify that 2|Ekj | ≥ |Qjk |. Notice
that, if k 0 ∈ Kj+1 satisfies Qj+1 j
k0 ⊂ Qk , then
α α α
a0 ak+1 ≤ |Qj+1
k0 | m3Qj+1
n
0
(f ) ≤ |Qjk | n m3Qj+1
0
(f ) ≤ |Qjk | n M [f χ3Qj ](x)
k k k
Here, the inequality is understood as the one for almost every point.
For the proof we use the following notation: Recall that given a cube Q,
D1 (Q) denotes the set of all dyadic children. We rely on a somewhat distorted
stopping time argument. For any fixed measurable function f on a fixed cube
Q0 and Q ∈ D(Q0 ), let S0 (Q) ≡ {Q}, and define S1 (Q) ⊂ D(Q0 ) to be the
collection of all maximal dyadic subcubes R of Q such that
max |Med(f, Q0 ) − Med(f ; Q0 )| > ω2−n−2 (f ; Q). (4.4)
Q0 ∈D1 (R)
∞
X X
χQ0 |f − Med(f ; Q0 )| ≤ ω(f ; Q)χQ
j=0 Q∈Sj (Q)
almost everywhere. Thus, the proof of Theorem 174 will be complete once we
prove Lemma 175.
Proof of Lemma 175 We may assume that Med(f ; Q0 ) = 0 by con-
sidering f − Med(f ; Q0 ) instead of f itself. In this case ω2−n−2 (f ; Q0 ) =
(χQ0 f )∗ (2−n−2 |Q0 |).
(1) We distinguish two cases. We first assume x ∈ Ω(Q0 ) and then assume
x ∈ Q0 \ Ω(Q0 ).
Fix x ∈ Ω(Q0 ). For the disjoint subcubes S1 (Q0 ) of Q0 , we write
X
A(x) ≡ Med(f ; Q)χQ (x),
Q∈S1 (Q0 )
X
B(x) ≡ χQ (x)(f (x) − Med(f ; Q)),
Q∈S1 (Q0 )
so that χΩ(Q0 ) (x)f (x) = A(x) + B(x). Note that the quantity B(x)
matches the second term in the right-hand side of the conclusion; we
have only to handle A(x). From the maximality it follows that any Q ∈
S1 (Q0 ) satisfies the opposite estimate in place of Q0 ∈ D(Q). That is,
|Med(f ; Q)| ≤ ω2−n−2 (f ; Q0 ). Thus, since S1 (Q0 ) is disjoint, |A(x)| ≤
ω(f ; Q0 )χΩ(Q0 ) (x).
Let x ∈ Q0 \ Ω(Q0 ) instead. Then we have |Med(f ; R)| ≤ ω2−n−2 (f ; Q0 )
for any cube R ∈ D(Q0 ) containing x. It should be noted that, for any
x ∈ Q0 \ Ω(Q0 ), such a cube R can be chosen as small as we wish. Thus,
|f (x)| ≤ ω2−n−2 (f ; Q0 ) for almost all x ∈ Q0 \Ω(Q0 ) by the Fujii lemma;
see Corollary 169.
(2) Let Q ∈ S1 (Q0 ) be arbitrary, so that some Q0 ∈ D1 (Q) satisfies
4.2.4 Exercises
Exercise 65. We let
D2k ≡ {[j · 4k − 2 · 4k /3, j · 4k + 4k /3)}j∈Z
D2k+1 ≡ {[2j · 4k − 2 · 4k /3, 2j · 4k + 2 · 4k /3)}j∈Z
for k ∈ Z. Let 0 < q ≤ p < ∞.
(1) [273] Show that D ≡ {Dk }∞
k=−∞ is a generalized dyadic grid.
(2) [273] Show that for any compact set K there exist k ∈ Z and Q ∈ Dk
such that K ⊂ Q.
Z q1
1
− q1 q
(3) Show that kf kMq ∼ sup sup |Q|
p p |f (y)| dy for f ∈ L0 (Rn ).
k∈Z Q∈Dk Q
0
Exercise 66. Let f ∈ L (Q0 ) with Q0 ∈ Q, and let λ ∈ (0, 1).
(1) Show that ωλ (f ; Q0 ) = 0 if and only if f is constant.
],D
(2) Show that Mλ;Q 0
f = 0 if and only if f is constant.
174 Morrey Spaces
0 n
for f ∈ L (R ), so that M = M0 is the Hardy–Littlewood maximal operator.
In connection with the Hardy–Littlewood maximal operator, we will consider
this generalization in Section 4.3.1.
Example 80. It seems that Morrey spaces are useful when we consider the
boundedness property of the fractional maximal operator Mα , 0 < α < n. In
n
fact, Mα maps M1α (Rn ) boundedly to L∞ (Rn ).
As is the case with the Hardy–Littlewood maximal operator, we have the
endpoint weak estimate.
Theorem 176. Let 0 < α < n. Then kMα f kWL n−α
n . kf kL1 for all f ∈
1 n
L (R ).
Proof Simply modify the case of α = 0; see Exercise 67.
The following theorem, called the Hardy–Littlewood–Sobolev theorem, is
fundamental.
1 1
Theorem 177. Let 1 < p < q < ∞ and 0 < α < n satisfy q = p −α
n . Then
kMα f kLq . kf kLp for all f ∈ Lp (Rn ).
Before the proof we remark that arithmetic shows
0
n (n − α)p
q− × = q.
n−α nu
Proof By the monotone convergence theorem we may assume that f ∈
L∞ n
c (R ); see Exercise 68. For λ > 0, we let Ωλ ≡ {Mα f > λ}. We observe
that
Z ∞ Z ∞
n n
(kMα f kLq )q = q λq−1 |Ωλ |dλ . λq−1− n−α (kf χΩλ kL1 ) n−α dλ
0 0
Various operators in Lebesgue spaces 175
or equivalently
(Z ! ) p1
∞ Z 1
n
p1 dt
kMα f kLp2 (Q(r)) . r p2
|f (x)| dx (4.4)
r B(t) tn−αp1 +1
. sup rα mpQ(r)
1
(f )
r≥1
. kf kLp1 (W ) ,
n
which forces −n + α − γ + ≤ 0. If p1 6= p2 , use the truncated power function
p01
n n
| · |β χB(1) to have α ≥ p1 − p2 .
Example 81. Let 1 < p < ∞. Fix r > 0 and x ∈ Rn . We consider the
following “partial maximal function” M r f (x) ≡ sup mB(x,t) (|f |) together with
t≥r
(p) (p)
its powered version: M r f (x) ≡ sup mB(x,t) (f ). Then one can prove that
t≥r
n
kM f kLp (B(x,r)) ' kM (f χB(x,2r) )kLp (B(x,r)) + r p M 2r f (x).
Various operators in Lebesgue spaces 177
n n (p)
Hence r p M r f (x) . kM f kLp (B(x,r)) . r p M 2r f (x). Note that
! ! p1
Z ∞ Z
(p) n
p dt
M 2r f (x) .r p |f (y)| dy
r B(x,t) tn+1
thanks to (4.4).
Since the family {Ekj }j∈N0 ,k∈Kj is pairwise disjoint and decomposes Q0 , we
fα f (x) . LSα f (x) for all x ∈ Q0 . This completes the proof.
conclude that M
We have the following variant of Theorem 179:
Remark 7. Let Da be the dyadic grid with a ∈ {0, 1/3}n . Denote by D]a (Q0 )
the set of all cubes in Da containing a cube Q0 . Then define
D] (Q0 )
Mα a f≡ sup `(Q)α mQ (|f |)χQ
Q∈D]a (Q0 )
for f ∈ L0 (Rn ). For a sparse family S ⊂ D]a (Q0 ) with the nutshell {EQ }Q∈S ,
write
α D]a (Q0 ) α
X
LSα f ≡ |Q| n mQ (f )χEQ and c∞,α (f ) ≡ sup |Q| n mQ (f ).
Q∈S Q∈D]a (Q0 )
178 Morrey Spaces
Then, for any fixed cube Q0 ∈ Q, there exists a sparse family S = Sf ⊂ D(Q0 )
D] (Q0 )
such that, for all x ∈ Q0 , MαDa f (x) .α,n LSα f (x) + c∞,α
a
(f ).
4.3.4 Exercises
n−α n−α n−α
Exercise 67. Prove Theorem 176 by the use of (a + b) n ≤a n +b n
Here, 0 < α < n. Section 4.4.1 examines the close connection between Mα
and Iα .
Example 82. Let 0 < α < n and 0 < β < n. If f (x) ≡ |x|−β χB(1) (x) for
x ∈ Rn , then a simple calculation shows Iα f (x) ∼ |x|−α−β as x → 0.
Various operators in Lebesgue spaces 179
Example 83. Let 0 < α < n. Observe that Iα χB(1) (x) > 0 for all x ∈
Rn and Iα χB(1) ∈ L∞ (Rn ). Furthermore, Iα χB(1) (x) ∼ |x|α−n for |x| ≥ 2.
Consequently, Iα χB(1) (x) ∼ (|x| + 1)α−n for x ∈ Rn . Thus a dilation and
translation argument yields Iα χB(x0 ,r) (x) ∼ rα (|x − x0 | + r)α−n . A similar
conclusion Iα χQ (x) ∼ `(Q)n (|x − c(Q)| + `(Q))α−n holds for cubes.
Example 84. Let 0 < α < n and f ∈ M+ (Rn ). Using Lemma 161, we write
Z ∞ Z !
1
Iα f (x) = (n − α) f (y)dy dl.
0 ln+1−α B(x,l)
Proof Suppose first that (4.2) holds. Let us check that the integral con-
verges for almost all y ∈ B(x, 1), where x ∈ Rn is chosen arbitrarily. It is not
so hard to see that
Z
Iα [χB(x,1) |f |](z)dz < ∞,
B(x,1)
which implies Iα [χB(x,1) |f |](y) < ∞ for almost all y ∈ B(x, 1) thanks to Theo-
rem 182(1) to follow. Therefore, it remains to show that Iα [χRn \B(x,1) |f |](y) <
∞ for almost all y ∈ B(x, 1).
Note that (4.2) implies
|f (y)|
Z
dy < ∞. (4.3)
R n 1 + |x − y|n−α
Thus
|f (y)|
Z
Iα [χRn \B(x,1) |f |](y) ≤ dy
Rn \B(x,1) |x − y|n−α
|f (y)|
Z
≤ dy
Rn \B(x,1) 1 + |x − y|n−α
< ∞.
Assume instead that Iα [|f |](x) < ∞ for almost all x ∈ Rn . Then choose a
point x0 ∈ Rn such that Iα [|f |](x0 ) < ∞. Then, we have
|f (y)|dy |f (y)|dy
Z Z
n−α
n−α
≤ 2 n−α
Rn \B(x0 ,2|x0 |) 1 + |y| Rn \B(x0 ,2|x0 |) |x0 − y|
If S = D(Rn ), then IαS f is simply called the discrete fractional integral oper-
ator. In particular, for S = D(Q0 ), we have
0 X
IαD(Q ) f ≡ `(Q)α m3Q (f )χQ .
Q∈D(Q0 )
X X
χQ (x)`(Q)α m3Q (f ) ≤ a0 aj+1 χQ (x)`(Q)α
j j
Q∈Dk Q∈Dk
. a0 aj+1 `(Qjk )α
. `(Qjk )α m3Qj (f )
k
0
D(Q )
for all Qjk ∈ S. Consequently, Iα χQ (x)`(Q)α m3Q (f ) for all
P
f (x) .
Q∈S
x ∈ Q0 . This completes the proof.
Following the procedure of Theorem 170 with h replaced by f , we can find
a sparse family S such that the following estimate holds:
Theorem 191. For f ∈ M+X (Rn ) and 0 < α < n, there exists a sparse family
S = Sf such that Iα f (x) ∼ `(Q)α m3Q (f ) · χQ (x) for x ∈ Rn .
Q∈S
Proof Since the proof is similar to Lemma 190, we omit the proof.
1− n2
n
1 X
Γ(x) ≡ √ aij xi xj (x ∈ Rn ).
(n − 2)ωn−1 det A i,j=1
Thus, we obtain
Z
1
Γ ∗ [Lϕ](0) = |y|2−n (−∆ϕA )(y)dy.
(n − 2)ωn−1 Rn
Note that
|y| 1
∆ τ |y| 2−n
. ε−n χB(ε) (y).
ε (n − 2)ωn−1
so that by the triangle inequality for integrals, we have the desired result.
s
4.4.5 The Bessel potential operator (1 − ∆)− 2 , s > 0
The operator dealt with in Section 4.4.5 lies outside the main stream of
Section 4.4. However due to the similarity and the strong connection with
fractional integral operartors, we consider the Bessel potential operator (1 −
s
∆)− 2 , s > 0. We will show the multiplicative law (1 − ∆)−s1 (1 − ∆)−s2 = (1 −
∆)−s1 −s2 and the inverse formula (1−∆)(1−∆)−1 f = (1−∆)−1 (1−∆)f = f
for suitable functions f .
We start with the definition of the Bessel kernel.
Definition 56. Let s > 0. Then we define the Bessel kernel of order s by
exp(−|εξ|2 )eix·ξ
Z
1
Gs (x) ≡ lim s dξ (x ∈ Rn ).
ε↓0 (2π)n Rn (1 + |ξ|2 ) 2
eix·ξ
Z
Roughly speaking, Gs (x) is the integral 2 s
dξ. However, if
Z Rn (1 + |ξ| ) 2
1
0 < s ≤ n, then 2 s
dξ = ∞, so that we need to take an indirect
Rn (1 + |ξ| ) 2
way to define Gs . We need to justify:
Various operators in Lebesgue spaces 189
Lemma 194. Let s > 0. Then the limit defining Gs (x) exists for each x ∈ Rn .
Proof We use integration by parts to the expression. To this end we trans-
form
exp(−|εξ|2 )eix·ξ exp(−|εξ|2 )(−∆ξ )n eix·ξ
Z Z
1
s dξ = s dξ.
Rn (1 + |ξ|2 ) 2 |x|2n Rn (1 + |ξ|2 ) 2
Since sup tm exp(−t) < ∞ for each m ∈ N,
t>0
η(ξ)eix·ξ
Z
1
gs,0 (x) ≡ n s dξ (x ∈ Rn )
(2π) Rn (1 + |ξ|2 ) 2
and for l ∈ N
(η(2−l ξ) − η(21−l ξ))eix·ξ
Z
1
gs,l (x) ≡ s dξ (x ∈ Rn ).
(2π)n Rn (1 + |ξ|2 ) 2
190 Morrey Spaces
∞
P
We decompose Gs (x) = gs,l (x). It is easy to see that gs,0 is a bounded
l=0
function, so that we have only to take care of gs,l . If we perform integration
by parts, |gs,l (x)| . (2l |x|)−N 2l(n−s) for any N ∈ N and x ∈ Rn . If we add
this estimate over l ∈ N, we obtain
∞
X ∞
X
|gs,l (x)| . min(2l(n−s) , 2−s |x|−n ) = O(|x|−n+s )
l=1 l=1
as x → 0.
As a result, although Gs is defined in a tricky manner, this is integrable.
Corollary 197. Let s > 0. Then Gs ∈ L1 (Rn ).
If we use Corollary 197, then we see that the operator f ∈ L1 (Rn ) 7→
Gs ∗ f ∈ L1 (Rn ) is a bounded linear operator.
As we will prove, we have the following multiplicative law:
Theorem 198. Let s1 , s2 > 0. Then Gs1 +s2 = Gs1 ∗ Gs2 .
Proof We remark that in the course of the proof of Corollary 197, we
have showed that
exp(−|εξ|2 )eix·ξ
Z
1
sup n 2 ) 2s
dξ . min(|x|−n+s , |x|−n−1 ).
ε>0 (2π) R n (1 + |ξ|
We calculate
thanks to Corollary 195, Proposition 196 and Fubini’s theorem. Using the heat
kernel, we calculate
exp(−|εξ|2 )eiy·ξ exp(−|εζ|2 )ei(x−y)·ζ
Z Z Z
s1 dξ s2 dζ dy
Rn Rn (1 + |ξ|2 ) 2 Rn (1 + |ζ|2 ) 2
π n2 Z Z exp(−ε (|ξ| + |ζ| ))eix·ζ
2 2 2
1 2
= lim s s exp − |ξ − ζ| dξ dζ
δ↓0 δ 2 1 2 2 4δ
Rn ×Rn (1 + |ξ| ) 2 (1 + |ζ| ) 2
exp(−|εξ|2 )2 eix·ζ
Z
= (2π)n s1 +s2 dξ.
Rn (1 + |ξ|2 ) 2
4.4.7 Exercises
0
θp θ−p
Exercise 69. Let f (θ) ≡ A + 0 B with A, B > 0. Find the minimum of
p p
f : [0, ∞) → [0, ∞).
Exercise 70. Let f (x) ≡ (1 + |x|)θ , x ∈ Rn with θ ∈ R. Let p0 = 1 < p1 < ∞.
(1) Find the necessary and sufficient condition on θ for f ∈ L∞ (Rn ).
(2) Find an analogue to (4.7) to show α ≤ n + γ if (4.6) holds.
Exercise 71. Prove Corollary 186. Hint: Use Theorem 184(1).
Exercise 72. [424, p. 252] Let Q be a dyadic cube, and let 0 < α < n. Then
for any f ∈ M+ (Rn ), show that
X Z Z
`(R)α f (x)dx . `(Q)α f (x)dx
R∈D(Q) R Q
which is initially defined for f ∈ Cc∞ (Rn ), extends to a bounded linear operator
on L2 (Rn ).
We will prove Theorem 201 as the special case of Example 92. It is not so
hard to show that the limit defining Rj f exists for f ∈ Cc∞ (Rn ); see Exercise
74.
Example 85. The most prominent case is the case where n = 1. In this case,
Z −ε Z ∞
f (t − s) f (t − s)
Hf (t) ≡ lim ds + ds (t ∈ R).
ε↓0 −∞ s ε s
194 Morrey Spaces
As a result,
Z
yj η(ε|y|)
∂j I1 f (x) = (1 − n) lim f (x − y)dy
ε↓0 Rn |y|n+1
Various operators in Lebesgue spaces 195
As a result,
Z
yj η(ε|y|)
n+1
f (x − y)dy
B(ε)\B ( 2ε ) |y|
|yj |η(ε|y|)
Z
≤ |f (x − y) − f (x)|dy
B(ε)\B ( 2 )
ε |y|n+1
|yj |
Z
≤ sup |f (x − y) − f (x)| n+1
dy
y∈B(ε)\B ( 2 )ε B(ε)\B ( 2 )
ε |y|
|yj |
Z
= sup |f (x − y) − f (x)| n+1
dy.
y∈B(ε)\B ( 2 )ε B(1)\B ( 2 )
1 |y|
xj − yj
Z
Consequently, ∂j I1 f (x) = −(n − 1) lim f (y)dy.
ε↓0 Rn \B(x,ε) |x − y|n+1
(2) There exist constants D1 and D2 such that K satisfies the following
estimates.
(i) Size condition:
|K(x, y)| ≤ D1 |x − y|−n , (4.2)
if x 6= y.
(ii) Hörmander’s condition:
|x − y|
|K(x, z) − K(y, z)| + |K(z, x) − K(z, y)| ≤ D2 , (4.3)
|x − z|n+1
if 0 < 2|x − y| < |z − x|.
If K(x, y) = K0 (x − y) for some function K0 , then T is called a singular
integral operator of convolution type.
Example 87. Let {ajν }j∈Z,ν∈Zn ∈ `2 (Z × Zn ). Also let ψ ∈ Cc∞ (Rn ) be
jn
supported on Q(3). For j ∈ Z and ν ∈ Zn , we set ϕjν (x) ≡ 2 2 ∆ψ(2j x −
ν), x ∈ Rn . Since
X
`(Q)−a−n χ3Q (x)χ3Q (y) ∼ |x − y|−n−a
Q∈D(Rn )
xj − yj
Z
Rj f (x) ≡ p.v. n+1
f (y)dy (x ∈ Rn ).
R n |x − y|
Various operators in Lebesgue spaces 197
When there is a singularity in the integral kernel, we use p.v. in this way to
approximate the integral.
Example 89. Surprisingly enough, the identity operator idL2 is a generalized
singular integral operator with the kernel K = 0.
∞ n
Example 90. Let n ≥ 3 and f ∈ C Zc (R ). Also let I2 be the fractional
f (y)
integral operator of order 2; I2 f (x) ≡ dy (x ∈ Rn ). A similar
R n |x − y|n−2
technique to Example 86 shows that
(xj − yj )(xk − yk )f (y)
Z
∂j ∂k I2 f (x) = n(n − 2) lim dy (4.5)
ε↓0 Rn \B(x,ε) |x − y|n+2
if 1 ≤ j < k ≤ n and
(n(xj − yj )2 − |x − y|2 )f (y)
Z
∂j 2 I2 f (x) = (n − 2) lim dy (4.6)
ε↓0 Rn \B(x,ε) |x − y|n+2
if 1 ≤ j ≤ n. See Exercise 73. Since ∂j ∂k I2 is L2 (Rn )-bounded thanks to
Example 92 to follow, we see that ∂j ∂k I2 is a singular integral operator of
convolution type.
In what follows we abbreviate this type of operator to CZ-operator. The
next proposition explains why we want to pose symmetric condition (4.3).
Lemma 202. For any CZ-operator T and f ∈ L2c (Rn ), its dual T ∗ on the
complex Hilbert space L2 (Rn ) satisfies
Z
∗
T f (x) = K(y, x)f (y)dy for a.e. x ∈
/ supp(f ). (4.7)
Rn
Proof We take a function g ∈ L2c (Rn ) whose support does not meet that
of f . Then by (4.1)
Z Z
∗
f (x)T g(x)dx = T f (x)g(x)dx
Rn n
ZRZ
= K(x, y)f (y)g(x)dxdy
Rn ×Rn
Z Z
= f (y) K(x, y)g(x)dx dy.
Rn Rn
Thus, since g is an arbitrary L2c (Rn )-function with its support contained in
Rn \ supp(f ), we have
Z
T ∗ f (x) = K(y, x)f (y)dy for almost everywhere x ∈ / supp(f ).
Rn
C ≡ {y = (y 0 , yn ) ≡ (y1 , y2 , . . . , yn ) ∈ Rn : yn > a |y 0 |} .
If n = 1 then we assume that K(x, y) & |x − y|−1 for all x ∈ R and for all
y > x or for all x ∈ R and for all y < x.
A function F defined on Rn \ {0} is said to be homogeneous of degree k if
F (tx) = tk F (x) for t > 0 and x ∈ Rn \ {0}.
Example 91.
1
(1) The Hilbert transform H, in which case K(x, y) ≡ x−y , is a genuine
−1
Calderon–Zygmund operator because K(x, y) ≥ |x − y| for all x ∈ R
and for all y < x.
n−1
R Ω ∈ C(S
(2) Suppose that we have )\{0}, which is homogeneous of degree
zero and satisfies S n−1 Ω(η)dη = 0. We let T be a Calderon–Zygmund
1 x−y
operator whose kernel K takes the form K(x, y) = nΩ .
|x − y| |x − y|
n−1
The properties of Ω imply that there exist η0 ∈ S and δ > 0 such
that Ω(η) & 1 for all η ∈ S n−1 ∩ B (η0 , δ). So, T is a genuine singular
integral operator.
We will show that there are Calderón–Zygmund operators amply. Here,
we present an application of Theorem 114 to create an example to which the
results in Section 4.5.1 are available.
Example 92. Let K ∈ C 1 (S n−1 ), and let ϕ ∈ Cc∞ (Rn \ {0}) be such that
Z ∞
X
K(x)dσ(x) = 0, ϕ(2l x) = χRn \{0} (x) (x ∈ Rn ).
S n−1 l=−∞
for all k1 , k2 ∈ Z. Meanwhile for f ∈ Cc∞ (Rn ), we know that the limit T f (x) =
k2
Tk f (x) exists for all x ∈ Rn . By the Fatou lemma, we have
P
lim
k1 ,k2 →∞ k=−k1
kT f kL2 . kf kL2 . Since Cc∞ (Rn ) is dense in L2 (Rn ), T extends to a bounded
linear operator on L2 (Rn ) with the bound kT kL2 →L2 . k∂ α KkL∞ (S n−1 ) ,
P
|α|≤1
in particular proving Theorem 201.
Recall that Ḣk (Rn ) denotes the homogeneous harmonic polynomial space
dim (Ḣk (Rn ))
of order k ∈ N0 . Fix an orthonormal basis {Yjk }j=1C in Ḣk (Rn ).
Example 93. Let k ∈ N0 and j = 1, 2, . . . , dimC (Ḣk (Rn )). Fix x ∈ Rn . Then
s
dimC (Ḣk (Rn )) k n
|Yjk (x)| ≤ |x| . (k + 1) 2 −1 |x|k
ωn−1
thanks to Corollary 122 and
s
k(n + 2k − 2) dimC (Ḣk (Rn )) k−1 n
|grad(Yjk )(x)| ≤ |x| . (k + 1) 2 |x|k−1
ωn−1
thanks to Corollary 124. Consequently, if we consider the integral operator
Tjk given by
Yjk (x − y)
Z
Tjk f (x) ≡ n+k
f (y)dy (x ∈ Rn )
Rn |x − y|
n
for k ≥ 1 and j = 1, 2, . . . , dimC (Ḣk (Rn )), then kTjk f kL2 . k 2 kf kL2 for all
f ∈ L2 (Rn ) with the implicit constants independent of such j and k.
200 Morrey Spaces
The next theorem is the heart of the matter. To formulate this theorem,
we define the quadrant as a set of the form
quadrants, the set of j satisfying x ∈ {Ej [|f |] > λ} is bounded from below for
each x ∈ Eλ . Hence, there exists j ∗ such that Ej [ |f | ](x) ≤ λ for all j < j ∗ .
Hence, we can partition Eλ :
[ \
Eλ = {x ∈ Rn : Ej [ |f | ](x) > λ, Ek [ |f | ](x) ≤ λ}.
j∈Z k∈Z∩(−∞,j)
(2) (The L1 (Rn )-condition) The L1 (Rn )-norm of g is less than that of f :
kgkL∞ ≤ 2n λ. (4.11)
(6) (Partition of thePset {M D f > λ}) The family {Qj }j∈J is disjoint and
D
{M f > λ} = Qj .
j∈J
The first term is easy to estimate. In fact, using the L2 (Rn )-boundedness
as usual, we have
Z Z
λ 1 1 1 1
|T g| ≥ . 2 |T g(x)|2 dx . 2 |g(x)|2 dx . kgkL1 . kf kL1 .
2 λ Rn λ Rn λ λ
Hence
λ 1
|T g| ≥ . kf kL1 . (4.13)
2 λ
To estimate the second term we need the following observation:
Z
Claim 208. For any j ∈ J, √
|T bj (x)|dx . kbj kL1 .
Rn \2 n Qj
204 Morrey Spaces
S Admitting
√ Claim 208, let us finish the proof of the theorem. Write Z ≡
2 nQj . Firstly we proceed as follows:
j∈J
X λ X λ
T bj ≥ ≤ (Rn \ Z) ∩ T bj ≥ + |Z|
2 2
j∈J j∈J
1 X X
. T bj + |Qk |.
λ
j∈J k∈J
L1 (Rn \Z)
X [ 1
Recall that |Qk | = Qk = |{M D f > λ}| . kf kL1 . By the trian-
λ
k∈J k∈J
gle inequality we obtain
X X XZ
T bj ≤ kT bj kL1 (Rn \Z) ≤ √
|T bj (x)|dx.
j∈J j∈J j∈J Rn \2 nQj
L1 (Rn \Z)
P
We can estimate kbj kL1 (Qj ) because the bj have support in Qj and the Qj
j∈J
are disjoint. If we invoke Claim 208, then we obtain
X X Z
T bj . kbj kL1 (Qk ) . (|f (x)| + |g(x)|)dx . kf kL1 .
j∈J j∈J Rn
L1 (Rn \Z)
Thus, we have finished the proof of the claim and hence the theorem.
Although it is impossible to have a pointwise control of singular integral
operators in terms of given functions, we can control their integral in terms
of the original functions in Lp (Rn )-norm.
Theorem 209. Let 1 < p < ∞ and T be a singular integral operator, which
is initially defined on L2 (Rn ). Let D1 and D2 be constants in (4.2) and (4.3),
respectively. Then T can be extended to Lp (Rn ) for all 1 < p < ∞ so that
kT f kLp .p (kT kL2 →L2 + D1 + D2 )kf kLp for all f ∈ Lp (Rn ) ∩ L2 (Rn ).
We still write T defined on Lp (Rn ).
Proof We may suppose that kT kL2 →L2 + D1 + D2 ≤ 1 as before. We
distinguish two cases keeping in mind that the case p = 2 is contained in the
assumption.
Case 1 : 1 < p < 2 In this case we will use interpolation of L2 (Rn )-
boundedness and weak-(1, 1) boundedness, the same technique that we used
for the maximal operator.
0
Case 2 : p > 2 In this case we use the duality Lp (Rn )-Lp (Rn ). Since
0
Cc∞ (Rn ) is dense in Lp (Rn ), for all f ∈ L2 (Rn ) ∩ Lp (Rn ),
Z
kT f kLp = sup T f (x)g(x)dx : g ∈ Cc∞ (Rn ), kgkLp0 ≤ 1
n
ZR
= sup f (x)T ∗ g(x)dx : g ∈ Cc∞ (Rn ), kgkLp0 ≤ 1
Rn
∗
≤ kf kLp kT gkLp0 .
Recall that T ∗ is again a singular integral operator and that 1 < p0 < 2. Thus
0
for all g ∈ Lp (Rn ) ∩ L2 (Rn ), kT ∗ gkLp0 . kgkLp0 .
Combining these estimates allows us to conclude that T can be extended
to a bounded operator on Lp (Rn ).
Lp (Rn ).
S
for all f ∈
1≤p<∞
Proof We may suppose that kT R kL2 →L2 +D1 +D2 ≤ 1 as before. Fix ε > 0
and x ∈ Rn . We will estimate K(x, y)f (y)dy independently of ε.
Rn \B(x,ε)
1
Let z ∈ B x, ε . We will decompose f according to B(x, 2ε): Let f0 ≡
2
χB(x,ε) · f and f1 ≡ f − f0 . As for the estimate of f0 , we will employ the size
condition |K(x, y)| . |x − y|−n . It follows from the definition of f0 that
Z Z
K(x, y)f (y)dy = K(x, y)f0 (y)dy . M f (x).
Rn \B(x,ε) B(x,2ε)\B(x,ε)
By the Hörmander condition (4.3), the first term is readily estimated by the
Hardy–Littlewood maximal function. Thus, we have
Z
K(x, y)f1 (y)dy . M f (x) + |T f1 (z)|. (4.15)
Rn \B(x,ε)
Various operators in Lebesgue spaces 207
Combining Cotlar’s inequality with Theorem 211, we can prove that the
truncated singular integral operator converges if and only if K enjoys a sort
of cancellation condition.
Theorem 212. Suppose that T is a singular integral kernel and that K is an
associated kernel. Then the limit
Z
T0 f (x) ≡ lim K(x, y)f (y)dy
ε↓0 Rn \B(x,ε)
exists for almost every x ∈ Rn . If this is the case, then for all 1 < p < ∞
there exists a constant cp such that
We define Z Z
p.v. ∂ij Γ(x)dx ≡ lim ∂ij Γ(x)dx.
Rn ε↓0 Rn \B(ε)
∞
Then for all ϕ ∈ Cc (Rn ),
Z Z
Γ(x)∂ij ϕ(x)dx = p.v. ∂ij Γ(x)dx + cij ϕ(0).
Rn Rn
It remains to use k + n − 2 ≥ k.
With these preliminary observations in mind, we present an example of sin-
gular integral operators. Roughly speaking, we will see that we can handle the
operator of the form T f (x) = k(x)T0 f (x), where T0 is a singular integral oper-
ator and k ∈ L∞ (Rn ). We consider the kernel (x, z) ∈ Rn × Rn 7→ K(x, z) ∈ C
below.
Theorem 215. Let K ∈ L0 (Rn × Rn ) satisfy the following conditions:
(1) K(x, ·) ∈ C ∞ (Rn \ {0}) for all x ∈ Rn ;
(2) K(x, tz) = t−n K(x, z) for every t > 0 and (x, z) ∈ Rn × S n−1 ;
Z
(3) K(x, z)dσ(z) = 0;
S n−1
∂αK
(4) max < ∞.
|α|≤2n ∂z α L∞ (Rn ×S n−1 )
Define Z
T f (x) ≡ K(x, x − y)f (y)dy (x ∈ Rn )
Rn
We claim that the series converges absolutely thanks to the strong decay in
each summand. Due to Example 93,
Z
Yjk (z) n
k+n
f (· − z)dz . (1 + k) 2 kf kLp .
Rn |z| Lp
due to Lemma 214, we conclude from Corollary 124 that (4.22) takes place in
the topology of the operator norm k · kLp →Lp . Thus, T is Lp (Rn )-bounded.
We can also discuss the pointwise convergence for this class of operators.
Theorem 216. Let K ∈ L0 (Rn × Rn ) satisfy the following conditions:
(1) K(x, ·) ∈ C ∞ (Rn \ {0}) for all x ∈ Rn ;
(2) K(x, tz) = t−n K(x, z) for every t > 0 and (x, z) ∈ Rn × S n−1 ;
Z
(3) K(x, z)dσ(z) = 0 for all x ∈ Rn ;
S n−1
∂αK
(4) max < ∞.
|α|≤2n ∂z α L∞ (Rn ×S n−1 )
Then there exists a function T f ∈ Lp (Rn ) such that lim kTε f − T f kLp = 0.
ε↓0
Lemma 217. Let T be a singular integral operator, and let Q be a cube. Then
∞
X
ω(Re(T f ); Q) + ω(Im(T f ); Q) . 2−l m2l Q (|f |)
l=0
for all f ∈ L∞ n
c (R ).
Proof We consider the real part solely. We set c ≡ Re(T [f ·χRn \3Q ](c(Q))).
Then there exist D1 , D2 > 0 such that
Z
`(Q)|f (y)|
|Re(T f (x)) − c| ≤ D1 |T [f χ3Q ](x)| + D1 dy
n
R \Q |y − c(Q)|n+1
∞
X
≤ D2 |T [f χ3Q ](x)| + D2 2−l m2l Q (|f |).
l=0
The idea of the proof is similar. The only difference from the case of the
maximal operator lies in the point where we handle the integral kernel with
the size condition.
Proof We split f = f1 + f2 , where f1 ≡ χB(x,2r) f and f2 ≡ f − f1 . For
f1 , apply the Lq (Rn )-boundedness of T and, for f2 , apply the size condition
together with the triangle inequality.
We transform Theorem 219 into the form which we use.
n
Corollary 220. Fix x ∈ Rn . Let 1 < q < ∞ and 0 < δ < q. Then
! ! q1
Z ∞ Z
n
q −δ
dt
kT f kLq (B(x,r)) . r |f (z)|q dz
r B(x,t) tn−δq+1
4.5.9 Exercises
Exercise 73.
(1) Prove (4.5) and (4.6) using the Stokes theorem. Alternatively, we may
argue as follows: Let ρ ∈ C ∞ (Rn ) satisfy χ(1,∞) ≤ ρ ≤ χ(2,∞) . For
ρ(ε|x − y|)f (y)
Z
ε > 0, we set I2,ε f (x) ≡ dy (x ∈ Rn ). Prove first
Rn |x − y|n−2
that I2,ε f converges in some suitable topology.
214 Morrey Spaces
4.6 Notes
Section 4.1
General remarks and textbooks in Section 4.1
See the fundamental textbooks and lecture notes [104, Chapter 2], [146,
Chapter II, §2], [156, §2.1], [223, Chapter 1] [293, §1.1], [319, §1.10], [417,
Chapter 2, §3], [416, Chapter I, §3 and Chapter II, §1], [382, §1.4] and [410,
§2.3] for the maximal operator considered in this book.
For the Fefferman–Stein vector-valued inequality, we refer to [146, Chapter
V, §4], where a different approach can be found.
See [118, §1.5] for various covering lemmas such as Vitali’s covering lemma
and Besicovitch’s covering lemma.
We viewed the Hedberg inequality in the context of Hedberg’s inequality;
see Lemma 181 for more about Example 61.
Section 4.1.1
The history of the Hardy–Littlewood maximal operator dates back to 1930
[182]. The original aim was to apply it to the functions on the unit disk ∆(1) on
the complex plane. Consequently, originally the Hardy–Littlewood maximal
operator was used for the analysis over the torus T = R/2πZ. Here, to study
the function spaces on R, we work in R. A natural passage to higher dimensions
is due to Wiener [449]. His purpose was to apply it to the ergodic theory.
Section 4.1.2
In retrospect, the proof of Theorem 141 essentially relies upon the interpo-
lation theorem due to Marcinkiewicz [302] whose proof Zygmund gave [473].
Wiener proved the sunrise lemma in [449]; see Theorem 137.
Section 4.1.3
We showed how to calculate the maximal operator on a cube Q for func-
tions supported outside 2Q. Lemma 130 is well known. See [387, Lemma 4.2]
for example. Fefferman and Stein considered Example 68.
Various operators in Lebesgue spaces 215
See [43] for the relation between the Kantorovitch operator and the Hardy–
Littlewood maximal operator on [0, 1].
Section 4.1.4
See [124] for the Fefferman–Stein vector-valued inequality, Theorem 145.
Our proof depends on the method of [378, Theorem 1.3].
Section 4.1.5
Carro, Pérez, F. Soria and J. Soria used the maximal operator ML log L to
investigate the dual inequality of Stein-type; see [64, Theorem 3.1].
Section 4.1.6
Gogatishvili, Mustafayev and Agcayazi considered the composition of the
maximal operator and fractional maximal operators. See [152, Theorem 4.1]
with α = 0 for Corollary 155.
Section 4.1.7
Sawano, Sugano and Tanaka characterized the condition on the bounded-
ness for the maximal operator MΨ to be locally bounded in the norm deter-
mined by Φ including the powered maximal operator M (η) in [392, Proposition
5.7], [392, Claim 2.18] and [392, Corollary 2.19]; see Proposition 156, Theorem
157 and Corollary 158, respectively.
Section 4.1.8
Recall that we considered the integral expression of the integral kernel
| · |α−n . See the survey paper [391, Lemma 3.2] for Lemma 160. We recorded
a general estimate on convolution by Burenkov and Guliyev; see [46, Lemmas
5,6 and 7] for Propositions 163, 164(1) and 164(2), respectively.
Section 4.2
General remarks and textbooks in Section 4.2
See the fundamental textbooks [29, Chapter 5 §8], [104, Chapter 6], [146,
Chapter II, §3], [157, §3.4], [293], [416, Chapter IV, §2] and [410, §6.2].
Section 4.2.1
Lerner proved the sharp maximal inequality, Theorem 165, [267, Theorem
2.1]. The sharp maximal inequality, Theorem 166, is due to Fefferman and
Stein; see [125, Theorem 5] as well as [125, (4.1)].
216 Morrey Spaces
Section 4.2.2
Fujii refined the Lebesgue differentiation theorem as in Lemma 168 [136].
Section 4.2.3
Lerner considered the local mean oscillation and the local sharp maximal
function in [269, Section 4] see also [270, Section 2.3]. Theorem 174 has a
history. Lerner prove Theorem 174 with an extra term in [270]. Later on
Hytönen proved Theorem 174 as it stands. We followed [270, Theorem 4.5].
We followed observations on Da made by Lerner [270, Proposition 2.1]. See
[269, Lemma 3.1] for Lemma 171.
We constructed a sparse family in Theorem 170; see the work by Tanaka
[424].
Section 4.3
General remarks and textbooks in Section 4.3
We refer to the fundamental textbooks [156, Exercise 1.2.6] and [293, §3.2].
See also [104, p. 89].
The fractional maximal operator can control the operator of the form
V α (−∆ + V )−β with 0 ≤ α ≤ β ≤ 1 [421, Theorem 1].
Section 4.3.1
Adams and Xiao calculated Iα [| · |−β ] in [8, Example 5.5]; see Exam-
ple 82. Theoren 182 is known as the Hardy–Littlewood–Sobolev inequality
[184, 185, 411]. We employed the method of Hedberg [190]. Lemma 181,
known as the Hedberg inequality, is nowadays a standard method of prov-
ing the Hardy–Littlewood–Sobolev inequality. It is nowadays common to use
the maximal operator and fractional maximal operators to control fractional
integral operators. See [446, p.146] for Lemma 183 and [7, Proposition 3.1.2]
for its variants.
Next, we consider fractional integral operators with rough kernel.
Definition 62 (Fractional integral operators with rough kernel). Let s > 1,
and let Ω ∈ L0 (Rn ) be a function homogeneous of degree zero such that
Ω|S n−1 ∈ Ls (S n−1 ). Let 0 < α < n.
(1) One defines the fractional maximal operator MΩ,α with rough kernel Ω
by Z
1
MΩ,α f (x) ≡ sup n−α |Ω(y)f (x − y)|dy (x ∈ Rn ).
r>0 r B(r)
(2) One defines the fractional integral operator IΩ,α with rough kernel Ω by
Ω(y)f (x − y)
Z
IΩ,α f (x) ≡ n−α
dy (x ∈ Rn ).
Rn |x − y|
Various operators in Lebesgue spaces 217
These operators go back to [101, p. 29]. See also the textbook [293, §3.6].
See [293, Chapter 2] for a counterpart to singular integrals together with a
detailed motivation. See [311, Theorem 1.1] for the Morrey–BMO boundedness
and [?, 375] for the Morrey boundedness.
Section 4.3.2
See [164, Theorem 4.1] for Theorem 178 with α = 0 and p2 = 1.
Section 4.3.3
See [91, Theorem 5.1] for the sparse estimate of the fractional maximal
operator Mα . Theorem 179, a sparse estimate for fractional maximal opera-
tors, is [334, Lemma 2.3].
Section 4.4
See [90, Proposition 2.3] for the sparse estimate of fractional integral oper-
ators Iα .
Section 4.4.1
Example 84 can be found in [320, 394] for example. See [16] for the Sobolev
embedding in terms of Morrey spaces over metric measure spaces. We followed
[319] for Lemma 180.
Section 4.4.4
We have some explicit formulas of the solution of the variable coefficient
elliptic differential equations; see [60] for Theorem 192.
We know that functions are controlled by the Riesz potential I1 of their
gradient; see [439, Lemma 1] as well as [415, Section V.2.3] for Theorem 193.
Section 4.4.2
We followed the works by Burenkov, Gogatishvili, Guliyev and Mustafayev
for the local estimate of fractional integral operators; see [46, Lemma 8], [44,
Lemma 3.3] and [44, Lemma 3.6] for Theorem 184.
218 Morrey Spaces
Section 4.4.3
The idea of decomposing fractional integral operators, as we did in Propo-
sition 188, dates back to 1995; see the paper [354, Section 3] by Pérez. See [199]
for a non-linear version of the discrete fractional integral operator appearing
in Proposition 188. The sparse estimate used here is [334, Lemma 3.1]; see
Theorem 191.
Section 4.4.5
The Bessel potential is a potential named after Friedrich Wilhelm Bessel,
a German astronomer, mathematician, physicist and geodesist.
Section 4.4.6
See the original paper [323] by Morrey for Morrey’s lemma. Cianchi and
Pick used Morrey’s lemma to show that embedding relation between Sobolev
(type) spaces and generalized Campanato spaces [79].
Section 4.5
General remarks and textbooks in Section 4.5
See the fundametal papers [28, 55] for the theory of singular integral oper-
ators. A general overview can be found in [53].
See [104, Theorem 2.11], [293, §1.2], [416, Chapter I, §4] (for general dou-
bling measures), [416, Chapter IV, §3] (using the dyadic cubes) [382] for the
Calderón–Zygmund decomposition.
For further results on singular integral operators considered in this book,
we refer to the textbooks [104, Chapters 3,4 and 5], [146, Chapter II, §5],
[147, Chapter 5], [223, Chapter 4], [293, §2.1], [415, Chpaters II and III], [417,
Chapter VI], [416, Chapter 1, §5], [382, §1.5] and [410, §6.4].
Although we did not consider the boundedness of singular integral oper-
ators from L∞ (Rn ) to BMO(Rn ), this is due to Spanne, Peetre and Stein
[352, 405, 414].
Section 4.5.1
The weak-(1, 1) boundedness of the Riesz transform is due to Kolmogorov
[237]. Its strong-(p, p) boundedness is due to M. Riesz [367, 368]. See also
[51, 287].
Section 4.5.2
We refer to [83] for Definition 57. We borrowed the term “genuine singular
integral operator” from [48]; see also [47, p. 34].
Various operators in Lebesgue spaces 219
Section 4.5.3
We refer to [54, Lemmas 1 and 2] for Theorem 205.
Section 4.5.4
See [54, Lemma 2] for the weak estimate of singular integral operators,
Theorem 207.
Section 4.5.5
See [54, Chapter 4] for Theorem 211.
Section 4.5.6
Theorems 215 and 216 (see [76, Theorem 2.5]) including Lemma 214 (see
[76, (2.6)]) are some of the important results in the theory of elliptic differen-
tial operators; see [463] for an application of Theorem 215 together with the
technique of the proof.
Section 4.5.7
Lerner obtained a control of ω(Re(T f ); Q) and ω(Im(T f ); Q) as in Lemma
217; see [271, Proposition 2.3].
Section 4.5.8
We followed Guliyev’s works [162, 163] for the local estimates; see Theorem
219. Corollary 220, the local estimate for singular integral operators, is [47,
Corollary 1].
Chapter 5
BMO spaces and
Morrey–Campanato spaces
Let us look at Morrey’s original observation in 1938 [323]. For 0 < θ < 1, recall
that the Lipschitz space of order θ denotes the set of all continuous functions
F for which kF kLipθ ≡ sup |x − y|−θ |F (x) − F (y)| is finite. Roughly,
x,y∈Rn ,x6=y
he showed that if f ∈ L1loc (Rn ) has a weak derivative in Mp1 (Rn ) with p > n,
n
then f ∈ Lip1− p (Rn ).
However, it is not easy to guarantee that a function in L1loc (Rn ) has a weak
derivative, let alone that it belongs to Mp1 (Rn ) for some p ∈ [1, ∞). Often,
it is taken for granted that the functions are locally integrable. Consequently,
there is a gap between what is assumed and what needs to be proven.
It is desirable to obtain some smoothness of locally integrable functions
from inequalities derived from integration. Actually, this chapter is interested
in how to replace the condition for a derivative belonging to Mp1 (Rn ). Morrey–
Campanato spaces can be used for this purpose. As the limiting case, the
bounded mean oscillation (BMO) space, arises naturally.
Section 5.1 investigates the BMO space since it is a fundamental func-
tion space. Section 5.2 handles operators made up of BMO functions. We
define Morrey–Campanato spaces in Section 5.3. Section 5.3, which includes
an auxiliary observation to Morrey–Campanato spaces, shows that Morrey–
Campanato spaces are isomorphic to known function spaces.
221
222 Morrey Spaces
is finite. This mapping makes sense despite ambiguity caused by the additive
constant in BMO(Rn ) as long as we consider the commutator. This type of
operator will be used when we consider the elliptic differential operators. We
aim in Section 5.1 to define the commutator described above more precisely.
One of the prominent properties of BMO(Rn ) is the distributional inequality
which is called the John–Nirenberg inequality and is investigated in Section
5.1.2. The John–Nirenberg inequality will be a key tool when we consider
operators involoved with BMO.
Example 97.
(1) In the sense of Remark 10, L∞ (Rn ) ,→ BMO(Rn ), since mQ (|f |) ≤
kf kL∞ for all f ∈ L∞ (Rn ).
(2) Let f (x) ≡ log |x| for x ∈ Rn . Then f ∈ BMO(Rn ). In fact, we need to
check mQ (|f −mQ (f )|) . 1 for all cubes Q. By the scaling f 7→ f (a·), we
may assume that |Q| = 1. If |Q| = 1 and |c(Q)| ≤ 2n, then mQ (|f |) . 1,
so that it is clear that mQ (|f − mQ (f )|) . 1. Meanwhile if |Q| = 1 and
|c(Q)| > 2n, then we can use | log |y| − log |z|| . 1 for all y, z ∈ Q.
BMO spaces and Morrey–Campanato spaces 223
As a result,
|mQ (f ) − m2j Q (f )| ≤ 2n jkf k∗ (5.1)
for j ∈ N. Therefore,
|f (x) − mQ(1) (f )|
Z
dx
Rn 1 + |x|n+1
∞ Z
|f (x) − mQ(1) (f )| |f (x) − mQ(1) (f )|
Z X
≤ n+1
dx + dx
Q(1) 1 + |x| j=1 Q(2 )
j 1 + |x|n+1
∞ Z
X 1
≤ |f (x) − mQ(1) (f )|dx
j=1
2(j−1)(d+1) Q(2j )
∞ Z !
X j 1
. · kf k∗ + |f (x) − mQ(2j ) (f )|dx .
j=1
2j |Q(2j )| Q(2j )
∞
fj = F in BMO(Rn ).
P
Thus, we have proven that
j=1
As a result, BMO(Rn ) is a Banach space.
Although we have remarked that we have to identify functions modulo
additive constants, we are still able to embed BMO(Rn ) to the set of all mea-
surable functions. To do this, we choose a special representative with integral
over Q(1) zero. Nevertheless, we usually regard BMO(Rn ) as a function space
modulo C.
Hence,
Corollary 226. There exists a constant θ > 0 depending on n such that the
following inequality:
|b(x) − mQ (b)|
Z
1
exp θ dx . 1
|Q| Q kbkBMO
j j!(j + 1)j+1
Proof By the Cauchy–Hadamard test, lim √ = lim =
j→∞ j j! j→∞ j j (j + 1)!
1
e < 3. Use a constant D in Corollary 225. Set θ ≡ . Then,
3D
∞
!
|b(x) − mQ (b)|j
|b(x) − mQ (b)|
Z Z
1 X 1
exp θ dx = θ dx
|Q| Q kbkBMO j=0
j!|Q| Q kbkjBMO
∞
X (Dθ)j j j
≤
j=0
j!
. 1.
5.1.3 Exercises
Exercise 75. Let Γ be the Gamma function given by
Z ∞
Γ(p) ≡ tp−1 e−t dt (p > 0).
0
Exercise 77. [344, Lemma 2.1] Let 1 ≤ p < ∞ and Q ∈ Q. Then show that
Z Z p p1 Z Z 1
p p
|u(x) − u(y)|dy ≤2 (u(x) − u(y))dy
Q Q Q Q
for all u ∈ Lp (Q). Hint: Use the decomposition u(x) − u(y) = u(x) − mQ (u) +
mQ (u) − u(y) to the left-hand side.
Exercise 78. [331, Lemma 2.2] Let F : C → C be a Lipschitz mapping.
(1) Let f ∈ L1loc (Rn ). Show that mQ (|F ◦ f (x) − mQ (F ◦ f )|) ≤
kF kLip mQ (|f (x) − mQ (f )|).
(2) If f ∈ BMO(Rn ). Then show that F ◦ f ∈ BMO(Rn ) together with the
estimate kF ◦ f k∗ ≤ kF kLip kf k∗ .
228 Morrey Spaces
5.2 Commutators
Commutators will play a key role when we consider elliptic differential
equations. Especially, commutators generated by BMO and singular integral
operators will arise as the solution operators of the elliptic differential equa-
tions. We will consider commutators generated by BMO and singular integral
operators and commutators generated by BMO and fractional integral oper-
ators in Sections 5.2.1 and 5.2.2, respectively.
∞
! 10
p
kf kLp
Z
p0
X
. |a(y) − mQ (y)| dy
j=1
(2j `(Q))n 2j+1 Q\2j Q
∞
X j
. n kak∗ kf kLp
j=1
(2j `(Q)) p
n
. `(Q)− p kak∗ · kf kLp .
Here for the second inequality we used (5.1).
Therefore, this is an appropriate candidate of the definition of [a, T ]f .
To justify the observation above, we will prove the following lemma.
Lp (Rn ), T be a singular integral operator and
S
Lemma 228. Let f ∈
1≤p<∞
a ∈ BMO(Rn ). Given a cube Q, we set
SQ f (x) ≡ (a(x) − mQ (a))T0 f (x) − T [(a − mQ (a))χ2Q · f ](x)
Z
− K(x, y)(a(y) − mQ (a))f (y)dy.
Rn \2Q
Then if Q and R are cubes with Q ⊂ R, then SQ f (x) = SR f (x) for almost
every x ∈ Q(⊂ R).
Proof Note that
SQ f (x) − SR f (x)
Z
= T [(a − mR (a))χ2R\2Q · f ](x) − K(x, y)(a(y) − mR (a))f (y)dy.
2R\2Q
Note that the above definition makes sense because the limit exists for a.e.
x ∈ Rn . In view of Lemma 228, the following proposition is clear.
Proposition 229. Let f ∈ L∞ n
c (R ). If R is a cube and T is a singular integral
operator with the kernel K, then
[a, T ]f (x) = (a(x) − mR (a))T0 f (x) − T [(a − mR (a))χ2R · f ](x)
Z
− K(x, y)(a(y) − mR (a))f (y)dy
Rn \2R
for a.e. x ∈ R.
Proof The proof is left as an exercise for the interested readers. See Exer-
cise 79.
We now prove the boundedness of commutators, where we present an appli-
cation of the sharp maximal inequality.
Lemma 230. Let T be a singular integral operator, and let r > 0, a ∈
BMO(Rn ) and f ∈ L∞ n
c (R ). Then
M ] ([a, T ]f )(x) . kak∗ M (r) ◦ T f (x) + M (r) f (x) (x ∈ Rn ). (5.3)
Proof Now let us estimate M ] ([a, T ]f )(x). To do this, we use the decom-
position of [a, T ]f for a given cube Q above. Firstly, we write
F1 (x) ≡ (a(x) − mQ (a))T0 f (x),
F2 (x) ≡ T [(a − mQ (a))χ2Q · f ](x),
Z
F3 (x) ≡ K(x, y)(a(y) − mR (a))f (y)dy
Rn \2Q
for the sake of simplicity. Let 1 < r < p be an auxiliary parameter fixed
throughout.
We treat F1 by using mQ (|F1 − mQ (F1 )|) ≤ 2mQ (|F1 |). Let us recall that
we have been writing
Z 1t
(t) 1 t 1
mQ (G) ≡ |G(x)| dx , M (t) F (x) ≡ M [|F |t ](x) t
|Q| Q
for t > 0 and F, G ∈ L0 (Rn ). Under this notation we have
(r 0 ) (r)
mQ (|F1 − mQ (F1 )|) . mQ (|a − mQ (a)|) · mQ (|T f |) . kak∗ M (r) [T f ](x).
The most right-hand side matches
√ the first term of (5.3).
As for F2 , we use the L r (Rn )-boundedness of T as well.
mQ (|F2 − mQ (F2 )|) ≤ 2mQ (|F2 |)
(r)
≤ 2mQ (|F2 |)
2
≤ kT [(a − mQ (a))χ2Q · f ]kL√r .
|Q|
BMO spaces and Morrey–Campanato spaces 231
Consider the operator given by WE ≡ sup |Wε f | for each subset E ⊂ (0, ∞), so
ε∈E
that W f (x) = W(0,∞) [|f |](x). By a simple approximation argument, we have
W f (x) = W(0,∞)∩Q [|f |](x). Furthermore, since WEj [|f |](x) ↑ W f (x) for any
∞
sequence of increasing finite subsets {Ej }∞
S
j=1 such that Ej = (0, ∞)∩Q, we
j=1
have only to prove kWE f kLp . kf kLp for any finite subset E ⊂ (0, ∞) with the
BMO spaces and Morrey–Campanato spaces 233
Thus, ME] [{Wε f }ε∈E ](x) . kak∗ M (r) f (x) + kak∗ M (r) ◦ M f (x) for all x ∈ Rn .
If we use the vector-valued sharp maximal inequality, then we have the desired
result.
Although the indicator function of balls is not smooth, we can take advan-
tage of the positivity of the integral kernel.
Corollary 233. Let a ∈ BMO(Rn ), and let 1 < p < ∞. Define
Z
χQ (x)
Ma f (x) ≡ sup |a(x) − a(y)| · |f (y)|dy (x ∈ Rn )
Q∈Q |Q| Q
and
Z
ϕ,ε
T f (x) ≡ Kϕ,ε (x, y)f (y)dy, T ϕ,∗ f (x) = sup |T ϕ,ε f (x)|.
Rn ε>0
where
Fε (z, y) ≡ −(a(y) − mQ (a)) (Kϕ,ε (z, y) − Kϕ,ε (c(Q), y)) χRn \2Q (y)f (y)
+ (a(z) − mQ (a))Kϕ,ε (z, y)f (y)
+ (mQ (a) − a(y))Kϕ,ε (z, y)χ2Q (y)f (y).
For the first term, we will use the John–Nirenberg inequality and the fact that
ϕ is smooth to give
Z
(a(y) − mQ (a)) (Kϕ,ε (z, y) − Kϕ,ε (c(Q), y)) χRn \2Q (y)f (y)dy
Rn
. kak∗ M (r) f (z).
For the second term, we will use the John–Nirenberg inequality again to give
Z Z
1
(a(z) − mQ (a))Kϕ,ε (z, y)f (y)dy dz . kak∗ M (r) [T ϕ,∗ f ](z)
|Q| Q Rn
For the third term, we will use the maximal singular integral operator to
give
Z
(mQ (a) − a(y))Kϕ,ε (z, y)χ2Q (y)f (y)dy . T ϕ,∗ [(a − mQ (a))χ2Q f ](z)
Rn
236 Morrey Spaces
Thus,
M ] [[a, T ]ϕ
∗ f ](x) . kak∗ M
(r)
f (x) + kak∗ M ◦ T ϕ,∗ f (x).
Since a ∈ L∞ (Rn ) and f ∈ L∞ n
c (R ), we are in the position of using the
sharp maximal inequality to have the desired result.
We are oriented to the almost everywhere convergence of truncated com-
mutators. To this end, we will prove that the maximal operators of truncated
commutators are bounded.
Theorem 235. Let a ∈ BMO(Rn ). Also let T be a singular integral operator
with the kernel K. Then the operator
Z
[a, T ]∗ f (x) ≡ sup (a(x) − a(y))K(x, y)f (y)dy (x ∈ Rn )
ε>0 Rn \B(x,ε)
∂αK
(4) max < ∞.
|α|≤2n ∂z α L∞ (Rn ×S n−1 )
Then there exist [a, T ]f ∈ Lp (Rn ) such that lim k[a, T ]ε f − [a, T ]f kLp = 0.
ε↓0
BMO spaces and Morrey–Campanato spaces 237
5.2.3 Exercises
Exercise 79. Prove Proposition 229 using (4.1).
Exercise 80. Let 1 < r < p. Using Lemma 237, the sharp maximal inequality
(r)
and the boundedess of Mα and Iα , prove Theorem 238.
Definition 65 (Lλ,q n
k (R )). Let λ ∈ R and 1 ≤ q ≤ ∞, and fix k ∈ N0 ∪ {−1}.
Then define the Morrey–Campanato space Lλ,q n
k (R ) as the set of all f ∈
q n
Lloc (R ) for which
1
kf kLλ,q ≡ sup inf n ρλ |B(ρ)|− q kf − P kLq (B(x,ρ)) < ∞.
k
(x,ρ)∈Rn+1 P ∈Pk (R )
+
in the above. We will use the following notation for this minimum.
Definition 66. Let λ ∈ R, 1 ≤ q < ∞ and k ∈ N0 ∪ {−1}. Also let f ∈
Lqloc (Rn ), x0 ∈ Rn and ρ > 0.
(1) Define Pk (·; x0 , ρ, f ) ≡ argminP ∈Pk (Rn ) kf − P kLq (B(x0 ;ρ)) , the best
approximation of u up to order k in the ball B(x0 , ρ). Regard this poly-
nomial as a function of the first variable. Other variables play the role
of the parameters.
(2) Define aα (x0 ; ρ, f ) ≡ ∂ α Pk (x0 ; x0 , ρ, f ) for any α ∈ N0 with |α| ≤ k.
Thanks to Remark
n 3, we have the following norm equivalence
o Lλ,q n
k (R ):
1
kf kLλ,q ∼ sup ρλ |B(ρ)|− q kf − PB(x,ρ)
k
f kLq (B(x,ρ)) .
k
(x,ρ)∈Rn+1
+
Proof It follows from the definition of the norm that the natural mapping
[f ] ∈ Lλ,q n λ,q n
max(−1,[−λ]) (R ) 7→ [f ] ∈ Ll (R ) is continuous.
To see that we can construct the inverse mapping, it suffices to prove that
l−1
there exists P = PB(x 0 ,ρ)
f ∈ Pl (Rn ) such that
1
ρλ |B(ρ)|− q kf − PB(x
l−1
0 ,ρ)
kLq (B(x,ρ)) . kf kLλ,q .
l
Proof
(1) This is trivial from the definition of the norms.
(2) This is nothing but Theorem 241 with λ > 0.
(2) Let ρ > 0 and λ < 0. Then whenever α ∈ N0 satisfies |α| < −λ,
f ∈ C [−λ] (Rn ) and k∂ α (f ∗ ψρ ) − ∂ α f kL∞ . ρ−λ−|α| kf kLλ,q .
k
Proof
(1) Since
k∂ α (f ∗ ψρ ) − ∂ α (f ∗ ψρ0 )kL∞
= k∂ α [(f − Pk (·; x0 , ρ, f ))] ∗ ψρ − ∂ α [(f − Pk (·; x0 , ρ, f ))] ∗ ψ2−n kL∞
≤ k(f − Pk (·; x0 , ρ, f )) ∗ ∂ α ψρ kL∞ + k(f − Pk (·; x0 , ρ, f )) ∗ ∂ α ψρ0 kL∞ ,
∞
X
(2) Since f − f ∗ ψρ = (f ∗ ψ2−L ρ − f ∗ ψ2−L+1 ρ ) converges in C [−λ] (Rn ),
L=1
f ∈ C [−λ] (Rn ) and f − f ∗ ψρ satisfy the desired estimate from (1).
242 Morrey Spaces
Thus, f ∈ Lλ,q n
[−λ] (R ).
Proof We may assume 0 < λ < 2 in view of Theorem 245 and the struc-
ture of the class {C˙λ (Rn )}λ>0 .
Let f ∈ Lλ,q n
1 (R ). We know that f is continuous thanks to Lemmas 239
and 240. Let us obtain the norm estimate. Let x, y ∈ Rn be fixed. Suppose
x 6= y. Choose r so that r < |x − y|. Then we have
1 1 1
PB(x+y,r) f (x + y) − 2PB(x,r) f (x) + PB(x−y,r) f (x − y)
1 1 1 1
= PB(x+y,r) f (x + y) − PB(x,|x−y|) f (x + y) − 2(PB(x,r) f (x) − PB(x,|x−y|) f (x))
1 1
+ PB(x−y,r) f (x − y) − PB(x,|x−y|) f (x − y).
Since λ < 0, thanks to Lemmas 239 and 240, we have
1
|PB(x+y,r) 1
f (x + y) − PB(x,|x−y|) f (x + y)| . |x − y|−λ kf kLλ,q ,
k
1
|PB(x,r) 1
f (x) − PB(x,|x−y|) f (x)| . |x − y|−λ kf kLλ,q ,
k
1
|PB(x−y,r) 1
f (x − y) − PB(x,|x−y|) f (x − y)| . |x − y|−λ kf kLλ,q .
k
So, 1
|PB(x+y,r) 1
f (x + y) − 2PB(x,r) 1
f (x) + PB(x−y,r) f (x − y)| . |x − y|−λ kf kLλ,q .
k
Letting r ↓ 0, we obtain |f (x + y) − 2f (x) + f (x − y)| . |x − y|−λ kf kLλ,q
k
5.3.3 Exercises
Exercise 81. Let f ∈ L1loc (Rn ). Show that |mQ (f ) − mk Q (f )| . log(k +
2)kf k∗ for any k ≥ 1 and any cube Q.
244 Morrey Spaces
5.4 Notes
Section 5.1
General remarks and textbooks in Section 5.1
See the textbooks and lecture notes [29, Chapter 5], [104, Chapter 6],
[146, Chapter II, §3], [147], [157, Chapter 3], [223, Chapter 3], [231], [293],
[416, Chapter IV], [382, §3.3] and [432] for the theory of BMO.
Section 5.1.1
Fefferman and Stein defined the space BMO in [125]. Theorem 223 is essen-
tially due to Fefferman and Stein; [125, (1.2)]. See [404] for the definition of
BMO+ (Rn ).
See [363] for the closure of Cc∞ (Rn ) in BMO(Rn ).
Section 5.1.2
See the original paper [221] by John and Nirenberg, where we can find a
variant for functions defined on cubes.
Section 5.2
General remarks and textbooks in Section 5.2
See the textbooks [157, §3.5] and [293].
Section 5.2.1
The boundedness of commutators [b, T ] appearing in this section is due to
Coifman, Meyer and Weiss [84]. Chanillo considered the boundedness property
BMO spaces and Morrey–Campanato spaces 245
of [b, Iα ] in Chanillo [66]. See [84, Theorem 1] for Theorem 231. See [401,
Theorem 2] for Theorem 235 including Lemmas 232 and 234.
Chianrenza, Frasca and Longo pointed out that the commutators in The-
orem 236 arise naturally; see [76, Theorem 2.5] for Theorem 236. We refer to
[464] for the application of Theorem 236. See [161] for its generalization.
We can generalize the above commutators, given by (5.6), as follows: Let
~b = (b1 , b2 , . . . , bm ). Define the multicommutator T~
b
Z m
Y
T~b f (x) ≡ K(x, y)f (y) (bj (x) − bj (y))dy,
Rn j=1
Section 5.2.2
See [36] for example.
Section 5.3
General remarks and textbooks in Section 5.3
The earliest overview of Morrey–Campanato spaces seems to be [353]. The
next one seems [423]. We refer to [258, Chapter 5] for Morrey–Campanato
spaces. We also refer to the survey [359, §4]. See [108] for the characterization
of Morrey spaces and Morrey–Campanato spaces in terms of the heat kernel.
Section 5.3.1
A series of results in this section are basically based on the works by
Campanato [57, 58, 59, 60]. As we saw in Theorem 241, Janson, Taibleson
and Weiss pointed out that the order of the polynomials does not affect so
much the definition of Morrey–Campanato spaces [220, Theorem 1]. Morrey–
Campanato spaces are known to be equivalent to Morrey spaces in many cases.
See [427, 460, 461] for weighted characterizations. Several characterizations of
weighted Morrey–Campanato spaces are obtained by Tang in [427].
The Morrey–Campanato space Lλ,q n θ
k (R ) and the Lipschitz space Lip (R )
n
are equivalent. This equivalence goes back to the works by Campanato and
Meyer; see [58, Theorema, p. 183] and [312, Theorem, p. 718], where both
authors showed that these norms are equivalent to the Lipθ (Rn ) norm. See
[353, p. 72] for an account of these facts. See also [329, Theorem 3.1].
Campanato and Meyers independently showed that Morrey–Campanato
spaces and Morrey spaces coincide in many cases as we saw in Theorem 242
see [57], [59, Theorem 6.II] and [313]. Duong, Xiao and Yan expanded this
idea by the use of the heat kernel in [109].
246 Morrey Spaces
Section 5.3.2
Jonsson, Sjögen and Wallin proved Theorem 244 [222]. Janson, Taible-
son and Weiss pointed out that Morrey–Campanato spaces are isomorphic to
Hölder–Zygmund spaces; see [220, Lemma (3.4)], [220, Theorem 2] and [220,
Theorem 3] for Lemma 243, Theorems 244 and 245, respectively. Grevholm
proved Theorem 246 in [159, Theorem 3.1]. Greenwald also gave a different
proof of Theorem 246 [160]. See the booklet by DeVore and Sharpley [100] and
the lecture note [220] by Janson, Taibleson and Weiss for Theorem 246. We
now aim to consider function spaces related to Morrey–Campanato spaces.
We content ourselves with definitions. We consider CMO and CBMO spaces.
Let d ∈ N0 ∪ {−1}, 1 ≤ p ≤ ∞ and − np ≤ λ < d, and define
Z ! p1
1 1
kf kCMOp,λ,d ≡ sup λ |f (x) − PQd f (x)|p dx
r≥1 r rn Q(r)
Z ! p1
1 1
kf kCBMOp,λ,d ≡ sup λ |f (x) − PQd f (x)|p dx .
r>0 r rn Q(r)
Example 101. Chen and Lau [70] and Garcı́a-Cuerva [141] introduced the
central mean oscillation space CMOp (Rn ) with the norm
Z ! p1
1
kf kCMOp ≡ sup |f (x) − fQ(r) |p dx ,
r≥1 |Q(r)| Q(r)
Z ! p1
1 p
kf kCBMOp ≡ sup |f (x) − fQ(r) | dx .
r>0 |Q(r)| Q(r)
So, CBMOp (Rn ) = CBMOp,0,0 (Rn ). Let C be the space of all constant func-
tions and X /C mean the quotient space of X by C. Then CMOp (Rn ) and
CBMOp (Rn ) are expressed by Bσ (E)(Rn ) and Ḃσ (E)(Rn ), respectively, with
E = Lp (Rn )/C, kf kE(Q(r)) = kf − fQ(r) kLp (Q(r)) and σ = np .
Fu obtained λ-central BMO estimates for commutators of N -dimensional
Hardy operators [131, Theorem 1.1].
Chapter 6
General metric measure spaces
Morrey spaces can describe the local properties of functions over Rn . This
applies to a general metric measure space. We remark that a metric measure
space means a couple (X, d, µ) of a metric space (X, d) and a Borel measure µ.
We do not recall the notion of metric spaces; see [241, §4] for metric measure
spaces. Recall that a Borel measure µ is defined for the σ-field generated by
open sets.
This chapter considers two typical cases. The first is where the Lebesgue
measure is replaced by a general Radon measure on Rn . Recall that a Radon
measure in a metric space (X, d) is a Borel measure such that µ(K) < ∞ for
any compact set K. Here, as examples the fractal sets generated by systems
of iteration functions are envisioned. Although fractal theory is not discussed
due to its vastness, the ternary Cantor set is recalled.
The second is where the underlying space differs from Euclidean space.
Hence, any (X, d) will do as long as (X, d) is a metric space. Graphs are
typical examples. An (unoriented) graph is a pair G = (V, E) comprising a
set V of vertices together with a set E of edges, which are 2-element subsets
of V . The (graph) distance between x and y is the minimal integer k such
that there exist k − 1 elements in between x and y. Since graphs are not the
primary purpose of this book, this chapter does not go into detail. But let us
mention that we use graphs in Examples 109 and 110. One of our motivations
is an extension to the general metric measure setting. Metric measure spaces
have already been discussed in this book: A previous chapter considered the
weighted measure w(x)dx. This chapter is interested in cases where the norm
bound is independent of the weights. Such an estimate can be obtained if
the weighted case is considered. This aspect is examined later in this book.
Section 6.1 establishes the theory of Lebesgue spaces in Euclidean space with
general Radon measures, while Section 6.2 focuses on the theory of Lebesgue
spaces in metric measure spaces.
and
χB(z,r) (x)
Z
M f (x) ≡ sup |f (y)|dµ(y) (x ∈ Rn ).
z∈Rn ,r>0 µ(B(z, r)) B(z,r)
Since we have no information on µ(B(x, r)) and µ(B(z, r)), these two opera-
tors must be treated differently. However, if we have enough information on
the relation between µ(B(x, r)) and µ(B(x, 2r)) for x ∈ Rn and r > 0, then we
can identify M and M 0 as follows: Assume that there exists a constant C > 0
such that the doubling condition µ(B(x, r)) ≤ µ(B(x, 2r)) ≤ Cµ(B(x, r))
holds for all x ∈ Rn and r > 0. Then M 0 f (x) ≤ M f (x) ≤ CM 0 f (x), x ∈ Rn
for any f ∈ L0 (µ). In fact, since
Z
1
M 0 f (x) = sup |f (y)|dµ(y),
z=x,r>0 µ(B(z, r)) B(z,r)
we have M 0 f (x) ≤ M f (x). Meanwhile, since B(z, r) ⊂ B(x, 2r) for all z ∈
B(x, r),
χB(z,r) (x)
Z
M f (x) = sup |f (y)|dµ(y)
z∈Rn ,r>0 µ(B(z, r)) B(z,r)
χB(z,r) (x)
Z
≤ sup |f (y)|dµ(y).
z∈Rn ,r>0 µ(B(z, r)) B(x,2r)
C0
Z
n
µ{x ∈ R : M f (x) > λ} ≤ |f (x)|dµ(x).
λ Rn
Lemma 248. Let k > √ 1, and let {B(xλ , rλ )}λ∈L be a family of balls and
assume that sup rλ ≤ k inf rλ < ∞. Then we can take disjoint subfamilies
λ∈L λ∈L
as in Theorem 247.
Proof As we mentioned, we may suppose that k ≤ 5; otherwise we can
use the classical 5r-covering lemma (Theorem 133).
First of all, a scaling allows us to assume sup rλ = 1. (We are working on
λ∈L
Euclidean space. Hence, we are able to multiply the scalar.)
In this part we divide the family of balls. More precisely we proceed as
follows: Let Q0 be a family of dyadic cubes of side length 1. Here, we are now
Yn
considering cubes of the form Q = [mj , mj +1), where the mj ’s are integers.
j=1
We abbreviate the dyadic cubes in Q0 to “cubes” for short. Let Q0 ≡ [0, 1)n
be a unit cube. We divide the cubes into subfamilies: If →
−
m = (m1 , m2 , . . . , mn )
n
is an element of N0 , we set
Q→ →
− → − 0 →
− n
m ≡ {Q ∈ Q0 : Q − ( p + m) = Q for some p ∈ (4Z) },
−
where 4Z ≡ {4l : l ∈ Z}. Note that the cubes in Q→ m satisfy the following
−
property: Suppose that Q and Q0 both belong to Q→ m and that Q and Q are
− 0
different, then the distance between the two cubes is larger than 3. Hence, if
the center of the ball B is in Q and the center of the ball B 0 is in Q0 , then B
and B 0 are disjoint.
Next we define →
− [
Lm ≡ {λ ∈ L : xλ ∈ Q}.
Q∈Q→
−
m
Taking into account the preceding paragraphs we may assume that all
centers of the balls are in Q0 . In fact once this is proven, by the last paragraph
we can take the balls satisfying the property of this lemma from Q→ m for any
−
→
− n →
− n
m ∈ N0 . For any m ∈ N0 , we obtain families B→
(1) (Nk )
, . . . , B→ . Translation
−
m −
m
shows the number Nk is not dependent on → −m. Hence, our desired family is
(j)
{B→− }→
− n
m m∈N0 ,j≤Nk
.
Hence, in what follows let us assume that all centers of the balls are in Q0
and that sup rλ = 1 by normalization.
λ∈L
First take a ball B(xλ1 , rλ1 ) arbitrarily from the family {B(xλ , rλ )}λ∈L .
1
The assumption k − 2 ≤ inf rλ ensures that the radius of each ball is between
λ∈L
1
k − 2 and 1. Thus, the ball
√ B(xλ1 , krλ1 ) contains all balls B(xλ , rλ ) such that
d(xλ , xλ1 ) is less than k − 1. √
Next take a ball B(xλ2 , rλ2 ) such that d(xλ2 , xλ1 ) ≥ k−1. We may choose
it arbitrarily as long as this condition is satisfied. As in the proceeding para-
graph, the ball√ B(xλ2 , krλ2 ) contains all balls B(xλ , rλ ) such that d(xλ , xλ2 )
is less than k − 1.
General metric measure spaces 251
In this way we repeatedly take a ball B(xλj0 , rλj0 ) such that d(xλj0 , xλj ) ≥
√
k − 1 for all j = 1, 2, . . . , j 0 − 1. This procedure will be stopped at the J0 th
[ J0
[
step where we obtain B(xλ , rλ ) ⊂ B(xλj0 , krλj0 ).
λ∈L j 0 =1
In fact this procedure stops in finite times: Precisely speaking, we claim
that J0 appearing in the last part is bounded by the constant that depends
only on k > 1 and n. Since all radii of the balls are at most 1, all balls are √ con-
tained in [−1, 2]n . By the construction of {xλj }, we have d(xλj , xλj0 ) ≥ k −1
for all j < j 0 ≤ J0 . Thus, we have
( J0 disjoint balls
!) whose radii are more than
√ √
k−1 k−1
. Precisely speaking, B xj , is disjoint. Note
2 2
j=1,2,...,J0
√ !
k−1
that B xj , is contained in [−1, 2]n for all j = 1, 2, . . . , J0 . Hence,
2
√ !n
k−1
we have J0 |B(1)| ≤ 3n . Thus, J0 is bounded by the quantity
2
which depends only on k > 1 and n. We denote this bound by N ; J0 ≤ N. If
J0 is less than N , we formally define Lj ≡ ∅ for j > J0 . Thus, we obtain the
desired families.
Next we prove Theorem 247. That is, we want to eliminate the assumption
√
sup rλ ≤ k inf rλ < ∞.
λ∈L λ∈L
Let B1,p be families obtained from X1 , using the Lemma 248. Suppose we have
obtained the families of the balls Bl,p with l = 1, 2, . . . , j, p = 1, 2, . . . , N and
that Xl with l = 1, 2, . . . , j are defined as the subsets of {B(xλ , rλ )}λ∈L . Then
we define
Xj+1 (6.2)
j+1 j [ [
≡ B = B(xλ , rλ ) : k − 2 < rλ ≤ k − 2 , B ⊂ kB 0 .
1≤p≤N,1≤l≤j B 0 ∈Bl,p
and [ [ [
B⊂ k B.
B∈Xj+1 p=1,2,...,N B∈Bj+1,p
Then we obtain a countable family of balls {Bλ }λ∈Λ0 satisfying two conditions
below:
[ 1 [
(1) Bλ ⊂ Bλ ,
∗
5
λ∈Λ0 λ∈Λ0
1
(2) Bλ is disjoint.
5 λ∈Λ0
Note that (2) will guarantee that any element in A, belongs to a ball in some
Gj . The generation of the ball B is defined to be the number j satisfying
B = Bρ with ρ ∈ Λj .
A trivial but important observation is that different balls Bj and Bk inter-
sect, then either j > k or k < j holds. Let k ∈ N be fixed, keeping the fact
above in mind. Then we set
We will prove that there exists Mn > 0 such that ]Ik ≤ Mn by proving
1
](Ik \ Jk ) . 1 and ]Jk . 1. Let j ∈ Jk . Then rk ≤ rj ≤ 3rk . Since Bj
5 j∈J
is disjoint, it follows that Jk .n 1.
Let j1 , j2 ∈ Ik \ Jk . If the generation of Bj1 and that of Bj2 are different,
then the balls Bj1 and Bj2 meet in a point and c(Bj ) lies sufficiently close
to the boundary of Bj1 and Bj2 . A geometric observation therefore shows the
angle ∠c(Bj1 )c(Bj )c(Bj2 ) & 1.
For each j ∈ Ik \ Jk , consider a point at which Bj and the segment
c(Bj )c(Bk ) meet. Then the observation of the above paragraph says that such
intersection points are torn apart. Even from this observation, we conclude
that the number of generations appearing in Ik \ Jk is bounded by a number
depending on n.
Of course, for each generation, the number of balls that can intersect Bk
is also bounded by a number depending on n. Therefore, ](Ik \ Jk ) .n 1.
In view of this observation we conclude that ]Ik is bounded by a constant
independent of k, say Mn .
We define a mapping σ : N → {1, 2, . . . , Mn + 1} as follows : First we set
σ(j) ≡ j for j ≤ Mn + 1. If j ≥ Mn + 2, then we define inductively
σ(j) (6.3)
−1
≡ min{l = 1, 2, . . . , Mn + 1 : Bj ∩ Bm = ∅ for all m ∈ σ (l) ∩ [1, j − 1]}.
We remark that by the pigeon hole principle the set appearing in (6.3) is not
empty, which guarantees that σ(j) is well defined. From the property of Mn ,
we learn that {Bj }j∈σ−1 (l) is disjoint for each l = 1, 2, . . . , Mn + 1. Thus we
have {{Bj }j∈σ−1 (l) }M n +1
l=1 is the desired family.
General metric measure spaces 255
Thus,
Z π
2
−4r 2 2
sin2 θ 2
π(1 − e )≤C (e−4r − e−4r )dθ.
−π
2
x ∈ B(yx , rx ) (6.4)
and that Z
1
|f (z)|dµ(z) > λ. (6.5)
µ(B(yx , brx )) B(yx ,rx )
Note that sup rx is at most R. Hence, we can apply Theorem 247. Applying
x∈Eb
the theorem with k = b/a > 1, we obtain a countable subset A ⊂ Eb such
that {B(yx , rx )}x∈A satisfies
[ [ b
B(yx , rx ) ⊂ B xj , rj (6.6)
j
a
x∈Eb
General metric measure spaces 257
and X
χB(xj ,rj ) .a,b 1. (6.7)
j
but this fails in general. Here, we will disprove (6.8) with κ = 3. We consider
the case n = 2. We define a measure µ by posing a weight function f given
below :
∞
X 1
f (x) ≡ χRn \B(1) (x) + χB(2−m+1 )\B(2−m ) (x).
m=1
m!
Let µ ≡ f (x)dx.
We disprove (6.8) by reduction to the absurdity. Suppose we have inequal-
ity (6.8) with κ = 3. First of all fix an integer α. We are going to let α tend
to infinity later.
258 Morrey Spaces
where δ0 denotes a dirac measure massed on the origin. In fact, we let (x, y) ∈
B(21−m ). By the rotation invariance of the sets B(2 · 2−m ) and {(x, y) ∈
R2 : M3 δ0 (x, y) > λ}, we may assume that 0 ≤ x < 2−k+1 and y = 0.
Since 0, (x, 0) ∈ B((x/2, 0), (1 + s)x/2) for all s > 0, we have M3 δ0 (x, 0) >
µ(B((x/2, 0), (1+s)x/2))−1 . If s > 0 is sufficiently small, we have M3 δ0 (x, 0) >
µ(B((x/2, 0), (1 + s)x/2))−1 > Rk .
It follows from the claim that we have
Z
|g(x)|dµ(x) ≤ Cµ(B((2−m , 0), 3 · 2−m ))M3 g(0).
B(21−m )
supp(gr )
α
\ [ 2πj 2πj
⊂ B(2−m+1 ) B 2−m+1 cos , 2−m+1 sin ,r .
j=1
α α
M3 µm,α (0)
] 1 ≤ j ≤ α : 2−k+1 cos 2πj 2πj
−k+1
α ,2 sin α ∈ B(y, r)
= sup .
(y,r)∈Rn+1 µ(B(y, 3r))
+
0∈B(y,r)
]J
Then M3 µk,α (0) can be written as M3 µk,p (0) = . max
SJ J⊂{1,2,...,α}
Fixing α, if ]J ≥ 2, we have by geometric observation
that µ(SJ ) ≥
Cα −k −k 1
. Notice also that µ(B((2 , 0), 3 · 2 )) = O . Thus, we have
(m − 1)! m!
µ(B((2−k , 0), 3 · 2−k ))
lim = 0. Furthermore, S{j} ∼ µ(B((2−k , 0), 3 · 2−k ),
k→∞ SJ
where ∼ is independent of α and k. Thus, keeping α fixed, we have
Proof We consider balls; for cubes, a similar technique can be used. Let
us set
n µ(B(x, r))
N ≡ x ∈ R : lim inf =0 . (6.9)
r↓0 rn
We claim that µ does not charge N . In fact, for all x ∈ N and ε > 0, there
exists rx ∈ (0, 1) such that µ(B(x, rx )) ≤ εrx n . Let R > 0 be arbitrary. By
Lemma 268 with δ = 1 (Vitali’s covering lemma), for any compact set K
contained in N ∩ B(R), we can find x1 , x2 , . . . , xL ∈ K such that
L
X L
[
χB(xl ,rxl /3) ≤ 1, K⊂ B(xl , rxl ).
l=1 l=1
where D(µ) stands for the set of all cubes Q in D(Rn ) with µ(Q) > 0. By
using the fact that the dyadic cubes are nested, we can prove the following:
Theorem 254 (Universal weak L1 (µ)-estimate for M D(µ) ). Let µ be a Radon
measure on Rn , λ > 0 and f ∈ L1 (µ). Then
Z
1
µ{x ∈ Rn : M D(µ) f (x) > λ} ≤ |f (y)|dµ(y).
λ {x∈Rn : M D(µ) f (x)>λ}
Proof Mimic the proof of Theorem 134.
A direct consequence of Theorem 254 is the Lp (µ)-boundedness.
Theorem 255 (Universal Lp (µ)-estimate for M D(µ) ). Let µ be a Radon mea-
1
sure on Rn , and let 1 < p < ∞. Then kM D(µ) f kLp (µ) ≤ (p0 ) p kf kLp (µ) for all
f ∈ Lp (µ).
Proof If we integrate the inequality in Theorem 254 against λp−1 dλ over
(0, ∞), we first obtain
Z Z
D(µ) 0
M p
f (x) dµ(x) ≤ p M D(µ) f (x)p−1 |f (x)|dµ(x). (6.11)
Rn Rn
∞
We may assume that f ∈ L (µ) and that µ{f 6= 0} < ∞. In this case,
M D(µ) f ∈ Lp (µ). See Exercise 87. By Hölder’s inequality and (6.11), we obtain
Z Z
M D(µ) f (x)p dµ(x) ≤ p0 |f (x)|p dµ(x). (6.12)
Rn Rn
262 Morrey Spaces
D(Q) D(Q)
If µ is the Lebesgue measure, then we write Mα and Mlog instead of
D(µ;Q) D(µ;Q)
Mα and Mlog ,
respectively.
We have the following universal estimates:
Theorem 257 (Universal estimates over cubes). Let Q be a cube and 0 ≤
α < n.
D(µ;Q)
(1) Let f ∈ L1 (µ), λ > 0, and let Ωλ ≡ {x ∈ Q : Mα f (x) > λ}. Then
n−α
λµ(Ωλ ) n ≤ kf χΩλ kL1 (µ) . (6.14)
n 1 1 α
(2) Let f ∈ Lp (µ) with 1 < p ≤ . Set = − . Then
α q p n
! n−α
n
q
kMαD(µ;Q) f kLq (µ) ≤ n kf kLp (µ) .
q − n−α
In particular, if α = 0, then
kM D(µ;Q) f kLp (µ) ≤ p0 kf kLp (µ) . (6.15)
General metric measure spaces 263
D(µ;Q)
(3) Let f ∈ L1 (µ). Then kMlog f kL1 (µ) ≤ ekf kL1 (µ) .
Proof Throughout the proof, we may assume that f ∈ L∞ (µ) by Fatou’s
lemma.
(1) We partition Ωλ by choosing maximal cubes contained Z in Ωλ : Ωλ =
S 1
Qθ , where each Qθ ∈ D(µ; Q) satisfies α |f (z)|dµ(z) >
θ∈Θ µ(Qθ )1− n Qθ
λ. Thus,
n−α X n−α X
λµ(Ωλ ) n ≤ λ µ(Qθ ) n ≤ kf χQθ kL1 (µ) ≤ kf χΩλ kL1 (µ) .
θ∈Θ θ∈Θ
Thus,
Observe that
0
n (n − α)p q(p − 1)n
q− × = = q.
n−α nu (p − 1)n
By Hölder’s inequality, we obtain
Z
nu n
|f (x)| n−α MαD(µ;Q) f (x)q− n−α dµ(x)
Q
nu n
≤ (kf kLp (µ) ) n−α (kMαD(µ;Q) f kLq )q− n−α .
As a result,
q n n
(kMαD(µ;Q) f kLq (µ) )q ≤ n (kf kLp (µ) ) n−α (kMαD(µ;Q) f kLq )q− n−α .
q− n−α
Thus, by (6.15)
D(Q)
kMlog f kL1 (µ) ≤ kM D(µ;Q),(u) f kL1 (µ)
1
= (kM D(µ;Q) [|f |u ]k u1 ) u
L (µ)
u1
1
≤ kf kL1 (µ) .
1−u
We also define
Z
D(µ) 1
Mlog f (x) ≡ sup χR (x) exp log |f (z)|dµ(z) .
R∈D(µ) µ(R) R
n 1 1 α
(2) Let f ∈ Lp (µ) with 1 < p ≤ . Set = − . Then
α q p n
! n−α
n
q
kMαD(µ) f kLq (µ) ≤ n kf kLp (µ) .
q− n−α
Proof Let Qj ≡ [−2−j , 2j ) for j ∈ N and use Theorem 258. Since the
constants are independent of j, we can let j → ∞ to obtain the desired
results.
Instead of D(Rn ), it is sometimes convenient to use different families. We
do not go into detail; we content ourselves with the case n = 1.
Lemma 259 (Three lattice theorem). Write Zl (R) = {3Q : Q ∈ Dl (R)} for
each l ∈ Z. Then each Zl (R), l ∈ Z, is partitioned into 3 disjoint subfamilies
{Eelk (R)}3k=1 so that the following properties hold, where Elk = {Q ∈ Dl : 3Q ∈
Eelk (R)}.
∞
Eelk (R) which meet at a
S
(1) Let k = 1, 2, 3. For any cubes Q1 , Q2 ∈
l=−∞
point, one is contained in the other.
(2) For all Q ∈ Dl (R), k = 1, 2, 3 and l ∈ Z, there exists Q0 ∈ Elk (R) such
that Q ⊂ 3Q0 ⊂ 5Q.
P
(3) For any k = 1, 2, 3 and l ∈ Z, χQ = 1.
Q∈Eelk
Proof Let
E01 (R) ≡ {. . . , [−3, −2), [0, 1), [3, 4), [6, 7), . . .},
1
E−1 (R) ≡ {. . . , [−2, 0), [4, 6), [10, 12), [16, 18), . . .},
1
E−2 (R) ≡ {. . . , [0, 4), [12, 16), [24, 28), [36, 40), . . .},
1
E−3 (R) ≡ {. . . , [−8, 0), [16, 24), [40, 48), [64, 72), . . .},
E02 (R) ≡ {. . . , [−2, −1), [1, 2), [4, 5), [7, 8), . . .},
2
E−1 (R) ≡ {. . . , [0, 2), [6, 8), [12, 14), [18, 20), . . .},
2
E−2 (R) ≡ {. . . , [4, 8), [16, 20), [28, 32), [40, 44), . . .},
2
E−3 (R) ≡ {. . . , [0, 8), [24, 32), [48, 56), [72, 80), . . .},
E01 (R) ≡ {. . . , [−1, 0), [2, 3), [5, 6), [8, 9), . . .},
1
E−1 (R) ≡ {. . . , [2, 4), [8, 10), [14, 16), [20, 22), . . .},
1
E−2 (R) ≡ {. . . , [8, 12), [20, 24), [32, 36), [44, 48), . . .}
1
E−3 (R) ≡ {. . . , [8, 16), [32, 40), [56, 64), [80, 88), . . .}.
k k
From the definition of E2l (R) and E2l−1 (R) with k = 1, 2, 3 and l ∈ Z, we can
k
guess the defintion of each El (R). We leave the interested readers to check
that the family {Elk (R)}k=1,2,3,l∈Z satisfies (1)–(3).
266 Morrey Spaces
The measure γ is called the Gauss measure. We distort the underlying measure
dx; the ambient space Rn together with the distance remains intact. Here, we
content ourselves with introducing the notion of maximal admissible balls and
the mother of a maximal admissible ball due to Mauceri and Meda.
Definition 67. Let a > 0 and define m(x) ≡ min(1, |x|−1 ).
(1) Define Ba ≡ {B(x, r) : (x, r) ∈ Rn+1
+ , 0 < r ≤ am(x)}.
respectively.
268 Morrey Spaces
a2
(2) For any j ∈ N0 , since aj+1 2 − aj 2 = 2a + from (6.18) we have
aj 2
2a ≤ aj+1 2 − aj 2 ≤ 2a + a2 . (6.19)
j−i
Consequently, for all j ≥ i ≥ 0, by aj 2 − ai 2 = (ak+i 2 − ak+i−1 2 ) and
P
k=1
(6.19), we see that (2) holds.
Lemma 263. Let a > 0 and B be a maximal ball in Ba . Suppose that the
sequence {aj }∞
j=0 is defined as in (6.18). Then, the following hold for each
j ∈ N:
(1) if |c(B)| ≤ 1 + a, then γ(B) ∼ 1;
(2) if aj ≤ |c(B)| < aj+1 for some j ∈ N, then either |c(M (B))| ≤ 1 + a or
aj−1 ≤ |c(M (B))| < aj ;
2 n
(3) if aj ≤ |c(B)| < aj+1 , then γ(B) ∼ e−aj j − 2 ;
(4) if |c(B)| > 1 + a and {M k (B) : 0 ≤ k ≤ k0 } is a chain of maximal
balls in Ba , with the property that M k (B) is the mother of M k−1 (B) for
k ∈ {1, 2, . . . , k0 }, where k0 is the smallest integer such that |cM k0 (B) | ≤
2
n2
γ(B)
1 + a, then e−(a +2a)k 1 − k0k+1 . γ(M k (B)) . e
−2ak
for all k ∈
{1, 2, . . . , k0 }.
Here, the implicit positive constants in (1)–(4) depend only on a and n.
Proof
(1) The proof is left as an exercise; see Exercise 86.
(2) For the purpose of showing |c(M (B))| < aj if |c(M (B))| > 1 + a, we
show that, when |c(M (B))| > 1 + a, we have aj−1 ≤ |c(M (B))| < aj .
This follows from reductio ad absurdum. Indeed, if |c(M (B))| ≥ 1 + a
a
√ |c(M (B))| ≥ aj , then by observing that ϕ(s) = s + s increases on
and
[ a, ∞), we obtain
a
|c(B)| = |c(M (B))| + |c(M (B))|−1 a ≥ aj + = aj+1 ,
aj
which contradicts assumption |c(B)| < aj+1 . Therefore, we have
|c(M (B))| < aj . Likewise, we have aj−1 ≤ |c(M (B))|; the details are
left to the reader. This proves (2).
√ p
(3) Lemma 262 yields 2aj ≤ aj ≤ |c(B)| < aj+1 ≤ (a2 + 2a)(j + 1),
which, combined with |c(B)| ≥ aj > 1, implies r(B) = am(c(B)) =
1
|c(M (B))|−1 a ∼ j − 2 . From aj ≤ |c(B)| < aj+1 and (6.19), it follows
2 2 2 2 2 2 2
that e−a −2a < eaj −aj+1 < eaj −|c(B)| ≤ 1. Hence, eaj −|c(B)| ∼ 1 or,
2 2 1
equivalently, e−|c(B)| ∼ e−aj . This and the aforementioned r(B) ∼ j − 2 ,
2 2 n
yield that γ(B) ∼ e−|c(B)| r(B)n ∼ e−aj j − 2 . Hence, (3) holds.
General metric measure spaces 269
|c(B)| > |c(M (B))| > · · · > |cM k0 −1 (B) | > 1 + a =: a1 . (6.20)
Thus, there exists some j ∈ N such that aj ≤ |c(B)| < aj+1 . Applying
(6.20) and (2) repeatedly, we conclude that, for all k ∈ {1, 2, . . . , k0 },
∞ S
(B) Rn =
S
4Bj,i .
j=0 i∈Ij
Thus, B ∗ does not intersect any ball in Aj , which contradicts (c). Thus,
S ∞ S
S
for all j ∈ N, Aj ⊂ 4Bj,i ⊂ 4Bk,i . This shows the property
i∈Ij k=0 i∈Ik
(B).
(C) By virtue of (a) and (b), together with the definition of Aj , we see
that any ball B from Aj only
S intersects the balls in Aj−1 or Aj+1 .
If B ∈ Aj and B 0 ∈ Aj−1 Aj+1 , by noticing that the width of the
annulus {Aj }∞ 1 1 0
j=1 is decreasing, we see that 2 B ∩ 2 B = ∅. Hence, we
obtain the desired result.
(D) From (b) and a volume argument, it follows easily that ]Aj is bounded
by a positive constant multiple of |aj |2n . That is, ]Ij . j n ∼ (j + 1)n
for j ∈ N by virtue of Lemma 262. Similarly, we see that ]I0 . 1.
The Gauss measure γ is locally doubling in the following sense:
npq
1+ p−q
Lemma 265. Let Cp,q,n ≡ 32 7 log(8 ). If a ball B(x, r) satisfies that
npq
1+ p−q
γ(B(x, 8r)) ≤ 8 γ(B(x, r)), then either |x| < 2r or B(x, r) ∈ BCp,q,n .
Proof It suffices to establish that B(x, r) ∈ BCp,q,n when |x| ≥ 2r. If
x ∈ B(2), then 4m(x) ≥ 2, which together with the fact |x| ≥ 2r gives that
r ≤ 1 ≤ 2m(x). Thus, B(x, r) ∈ B2 ⊂ BCp,q,n .
Now assume that |x| ≥ 2. In this case, m(x) = |x|−1 . Notice that
( 2 )
2 2 2 5
min |y| = (|x| − r) , max |y| = max |x| − r
y∈B(x,r) y∈B (x− 2|x|
3r
x, 41 r ) 4
3r 1
and that B(x − 2|x| x, 4 r) ⊂ B(x, 8r). Consequently, we have
npq n 2 npq
81+ p−q π − 2 e−(|x|−r) |B(x, r)| ≥ 81+ p−q γ(B(x, r)) ≥ γ(B(x, 8r))
3r 1 n 5 2
≥γ B x− x, r ≥ π − 2 e−(|x|− 4 r) |B(x, r)|,
2|x| 4
For any a > 0, denote by Ba the set of all balls B ⊂ X such that r(B) ≤
am(c(B)). Balls in Ba are referred to as admissible balls with scale a. If r(B) =
am(c(B)), then B is called maximal.
Definition 68. The triple (X, d, µ) is called a locally doubling measure metric
space if, for every a > 0, there exist constants Da ∈ (1, ∞) and Ra ∈ (1, ∞)
such that for all B ∈ Ba ,
and
µ(2B) ≥ Ra µ(B). (6.24)
Conditions (6.23) and (6.24) are respectively referred to as the locally doubling
condition and the locally reverse doubling condition.
Example 108. A typical but non-trivial example of the locally doubling
measure metric spaces is the Gauss measure space (Rn , | · |, γ), where
Z
2
−n
γ(E) ≡ π 2 e−|x| dx.
E
For all x ∈ Rn , let m(x) ≡ min{1, |x|−1 }. As above, for any a > 0, we say
that B ∈ Ba if and only if r(B) ≤ am(c(B)). If B ∈ Ba and x ∈ B, then
which implies that m is an admissible function in the sense of (6.22). For all
B ∈ Ba , we have 2B ∈ B2a . Proposition 260 further implies that
Z
n 2
γ(2B) = π − 2 e−|x| dx
2B
n 2
≤ π − 2 e−|c(B)| +4a |2B|
Z
2 n 2
≤ 2n ea +6a π − 2 e−|x| dx
B
n a2 +6a
≤ 2 e γ(B),
r(B) 3
|z| ≤ |z − w| + |w| < + |c(B)| − r(B) = |c(B)| − r(B),
2 2
which implies that
r(B) n 2 r(B)
γ B w, ≥ π − 2 e−(|c(B)|−r(B)) B w, .
2 2
Meanwhile, for all z ∈ B, we have |z| ≥ |c(B)| − |c(B) − z| > |c(B)| − r(B)
n 2
and γ(B) ≤ π − 2 e−(|c(B)|−r(B)) |B|. Thus
6.1.6 Exercises
Exercise 84. Prove (6.1) by integration by parts.
Exercise 85. Let a > 0. Use |x − c(B)| ≤ a min(1, |x|−1 ) for all x ∈ B such
that B is a ball in Ba to show −a2 − 2a ≤ |c(B)|2 − |x|2 ≤ 2a.
Exercise 86. Let a > 0 and define m(x) ≡ min(1, |x|−1 ). Write b ≡ a
a+1 .
Exercise 87.
and
χB(z,r) (x)
Z
M f (x) ≡ sup |f (y)|dµ(y) (x ∈ X).
z∈X,r>0 µ(B(z, r)) B(z,r)
The former is the centered maximal operator, while the latter is the uncentered
maximal operator. Section 6.2.1 considers the boundedness of the maximal
operator. What differs from Section 6.1.2 is the lack of the strong covering
lemma. We have to use the 5r-covering lemma instead. Sections 6.2.1 and
6.2.2 parallel Section 6.1.2. We collect some estimates on the modified maximal
operators. In Section 6.2.3 with an example we explain that the modification
considered in Section 6.2 is actually necessary.
µ(B(x, r)) > 0 for all x ∈ X and r > 0. Define the centered modified maximal
operator
Z
1
Mk,c f (x) = Mk,c,µ f (x) ≡ sup |f (y)|dµ(y)
r>0 µ(B(x, kr)) B(x,r)
χB(y,r) (x)
Z
Mk,uc f (x) = Mk,uc,µ f (x) ≡ sup |f (z)|dµ(z).
y∈X,r>0 µ(B(y, kr)) B(y,r)
We will see how large k must be in order that we have the bounded-
ness of Mk,c . Section 6.2.1 investigates the property of the centered Hardy–
Littlewood maximal operator. We will omit the analysis of the uncentered
Hardy–Littlewood maximal operator. In general metric measure spaces, the
following theorem is fundamental.
Theorem 267. Let (X, d) be a separable metric space. Let µ be a Radon
measure such that µ(B(x, r)) < ∞. Then M2,c is weak-(1, 1) bounded and the
weak-(1, 1) constant is less than or equal to 1:
Z
1
µ{x ∈ X : M2,c f (x) > λ} ≤ |f (x)|dµ(x)
λ X
This extends Vitali’s covering lemma: The lemma is precisely Vitali’s cov-
ering lemma if δ = 1.
Proof We select j1 ∈ {1, 2, . . . , N } so that rj1 = max{r1 , r2 , . . . , rN }. If
[
B(xj , δrj ) ⊂ B(xj1 , (2 + δ)rj1 ),
j∈{1,2,...,N }
we have nothing else to do. Let us assume otherwise in the sequel. We define
Next, we take jq ∈ Λq so that rjq = max rj and This procedure will terminate
j∈Λq
because we are dealing with the finite number of the balls. Suppose we have
stopped in the p-th step after we selected jp and Λp . We will verify that
A ≡ {j1 , j2 , . . . , jp } satisfies all requirements of the lemma.
To verify this we fix j ∈ {1, 2, . . . , N }. We have three possibilities.
(a) j ∈ {j1 , j2 , . . . , jp }.
(b) rj = rjp and j ∈
/ {j1 , j2 , . . . , jp }.
(c) rjk ≥ rj > rjk+1 for some k ∈ {1, 2, . . . , p − 1} and j ∈
/ {j1 , j2 , . . . , jp }.
S
We want to show that B(xj , δrj ) ⊂ B(xj , (2 + δ)rj ); once this is achieved,
j∈A
we have (2). If (a) happens, this inclusion is clear from the definition of
j1 , j2 , . . . , jp . Assume (c) in the sequel. Another possibility (b) can be dealt
with in a similar manner. A geometric observation S together with (c) shows
that j ∈ / Λk and hence B(xj , δrj ) ⊂ B(xk , (2 + δ)rk ) by the
k∈{j1 ,j2 ,...,jk }
definition of Λk and rjk+1 . Thus, our claim is justified.
It remains to show that the balls {B(xj , rj )}j∈A are disjoint. Indeed sup-
pose k < k 0 , so that we have
Combining (6.1) and (6.2), we conclude that {B(xj , rj )}j∈A is disjoint. So,
(1) is verified.
We now prove Theorem 267.
Proof Fix λ > 0. We define Ek ≡ {x ∈ X : Mk,c f (x) > λ} for k > 2. By
Remark 12, it follows that
[
{x ∈ X : Mk,c f (x) > λ} = {x ∈ X : M2,c f (x) > λ}. (6.3)
k>2
and that X
χB(xl ,rxl ) ≤ 1. (6.5)
l∈A
for l ∈ A.
Combining (6.4)–(6.6) gives
!
N
[ [
µ B(xj , δrxj ) ≤ µ B(xl , (2 + δ)rxl )
j=1 l∈A
X
≤ µ(B(xl , (2 + δ)rxl ))
l∈A
X1Z
≤ |f (x)|dµ(x)
λ B(xl ,rx )
l∈A l
1
≤ kf kL1 (µ) . (6.7)
λ
Letting N tend to infinity in (6.7), we derive λµ(Ek ) ≤ kf kL1 (µ) . Tending
k ↓ 2, we conclude the proof of Theorem 267.
278 Morrey Spaces
By Lemma 268 there exists a subfamily of balls {B(xj , arxj )}j∈J with J ⊂
{1, 2, . . . , N } that satisfies the following properties.
X
χB(xj ,rxj ) ≤ 1, (6.11)
j∈J
N
[ [
B(xj , arxj ) ⊂ B(xj , (2 + a)rxj ). (6.12)
j=1 j∈J
k{M7,c fj }∞
j=1 k`q Lp (µ)
.p,q k{fj }∞
j=1 k`q Lp (µ)
. (6.15)
for each k ∈ N.
At this moment, about the definition of the natural number N0 , we keep in
mind that N ∈ M↓ (0, ∞).
Note that
and that
Hence
∞ ∞ ∞
[ X 2πγ 2πγ X (3 · k!)k
µ0 Aj = = . (6.22)
(3 · j!)j (3 · k!)k (3 · j!)j
j=k j=k j=k
!
∞
S
This observation yields the lower bound for µ0 Aj . It remains to obtain
j=k
!
∞
S
the upper bound for µ0 Aj .
j=k
First of all, let us assume that k = 0. Note that µ0 (A0 ) = 2πγ. Hence,
when k = 0, from (6.22), we deduce
∞ ∞
[ [ 1
µ0 Aj = µ0 Aj = 1 ≤ 2πγ 1 + .
j=0
γ
j=k
Proposition 279. There exist a separable space (X, d) and a measure µ such
that Mk,c is bounded if and only if k ≥ 2. Furthermore, the weak-(1, 1) constant
of M2,c is 1 on this space.
Proof We may assume 32 < k < 2, since Mk,c is decreasing as k increases.
Suppose that Mk,c is bounded. We want to derive a contradiction. We begin
1B(0,r)
with constructing an approximation of Dirac delta. Let gr ≡ . Then
µ(B(0, r))
Mk,c gr tends pointwise to Mk,c δ0 as r ↓ 0, where δ0 is a point mass at 0. Let
a be a point defined as follows:
Put Kn ≡ 2 + [log10 3n ]. Define a = an = {3−n χ[1,Kn ] (j)}j∈N ∈ D.
1
Take λ ≡ . We have
µ(B(a, k · 3−n ))
Thus, we have µ(B(0, 3−n )) . λ−1 . That is, µ(B(0, 3−n )) . µ(B(a, k · 3−n )).
By the definition of ν, there are infinitely many integers n such that
√ (2+[log10 3n ])
ν({x + −1y ∈ D : x2 + y 2 ≤ 9−n })
√ . 1.
ν({x + −1y ∈ D : (x − 3−n )2 + y 2 ≤ k 2 · 9−n })
6.2.4 Exercises
Exercise 90. Let (X, d, µ) be a metric measure space.
(1) Show that for any ball B(x, r) and for any y ∈ B(x, r),
Z
1
|f (y)|dµ(y) ≤ M2,c f (x).
µ(B(x, 2r)) B(x,r)
(2) Show that for any ball B(x, r) and for any y ∈ B(x, r),
Z
1
|f (y)|dµ(y) ≤ kf kL∞ (µ) .
µ(B(x, 2r)) B(x,r)
Z ∞
X q
I ≡ sup M22,c fj (x) r hj (x) g(x) dµ(x).
X j=1
(5) Use Proposition 273 and the fact that q > r > 1 to have
∞ Z
X q
I .p,q sup |fj (x)| r M7,c [hj g](x)dµ(x)
j=1 X
Z ∞
X q
= sup |fj (x)| r M7,c [hj g](x) dµ(x). (6.28)
X j=1
and
Z ∞
X q
sup |fj (x)| r M7,c [hj g](x) dµ(x)
X j=1
r1 10
r
Z ∞ ∞
X X 0
q r
≤ sup |fj (x)| (M7,c [hj g](x)) dµ(x)
X j=1 j=1
s00 s10
r
Z ∞
X 0
r
× sup (M7,c [hj g](x)) dµ(x)
X j=1
General metric measure spaces 289
(7) Note that q > p implies s0 > r0 . Use Theorem 270 with parameters
s0 > r0 to obtain
s00 s10
r
Z ∞
X 0
(M7,c [hj g](x))r dµ(x)
X j=1
s00 s10
r
Z ∞
X 0
.p,q (hj (x)g(x))r dµ(x) .
X j=1
Exercise 96. In Theorem 267, prove that we cannot improve the conclusion
in the following sense: Let 0 < ρ < 1. Then there does not exist a constant
Cρ such that
Z
Cρ
µ({x ∈ X : M3−ρ,uc f (x) > λ}) ≤ |f (x)|dµ(x) (6.31)
λ X
6.3 Notes
Section 6.1
General remarks and textbooks in Section 6.1
See standard textbooks [112, Chapter 3], [231], [293] and [459].
See [147, §1.3] for various covering lemmas adapted to quasi-metrics.
Fu, Yang and Yuan considered the commutators in [130, Theorem 1.15].
Futamura, Mizuta and Shimomura consided the Sobolev embedding in the
setting of general metric measure spaces and variable exponents [138, Theo-
rem 4.1]. Garcı́a-Cuerva and Martell obtained the vector-valued estimates for
singular integral operators [143, Section 2] and the existence of the principal
values [144, Theorem 1(f)]. Gorka considered Morrey–Campanato spaces in
the setting of metric measure spaces [154].
Universal estimates for dyadic maximal operators can be found in [223, p.
11, Theorem]; see Theorem 254.
Section 6.1.1
The boundedness of the maximal operator Mκ,uc associated with general
Radon measure goes back to [339, Section 3]. See [338] as well. Later, Sawano
and Terasawa pointed out that the modification rate κ can be decreased [378,
433]. Theorem 247 is [378, Theorem 1.5].
See [68] for the commutators generated by RBMO(µ) and Iα , where Chen
and Sawyer considered a modification constant that differs from the one con-
sidered by Tolsa in [435].
Section 6.1.2
Tolsa pointed out that the modification rate κ in Mκ,uc must satisfy κ > 1
for the boundedness of Mκ,uc ; see [435, p. 127]. This idea was reinvestigated
by Sawano [378].
Section 6.1.3
As we saw in Theorem 253, Tolsa proved that the analogue of the Lebesgue
differentiation theorem is available [435, Remark 2.3]. The covering lemma and
the vector-valued inequality can be found in [378]; see [378, Theorem 1.5] and
[378, Theorem 1.7] for Theorems 247 and 251, respectively.
Section 6.1.4
D(µ)
We considered the universal Lp (µ)-estimate for Mlog defined in (6.13).
See [209, p. 786] for Theorem 256. Theorem 257(3) is recorded as [209, Lemma
General metric measure spaces 291
2.13]. Lerner and Nazarov proposed to distort dyadic cubes as in Lemma 259;
see [92, Lemma 4.8].
Section 6.1.5
Mauceri and Meda considered the admissible balls in [308]. Proposition
260 is due to Mauceri and Meda [308, Proposition 2.1]. Lemma 262 can be
found in [308, p. 298].
Fabes, Gutiérrez and Scotto dealt with singular integral operators associ-
ated with the Gauss measure [119], where we can find a relation between the
Ornstein–Uhlenbeck process and singular integral operators they handle. See
also [174, 175, 183, 443] for more. See [145, Lemma 2.4] and [308, Lemma 5.1]
for more results related to Lemma 264. The corresponding version in the
locally doubling measure metric spaces was proven in [284, Lemma 2.7].
Lemma 265, which guarantess that a ball is doubling under a certain con-
dition, can be found in [281, Lemma 3.9]. Example 108 together with estimate
(6.25) is [278, Lemma 2.1].
Carbonaro, Mauceri and Meda considered the Hardy space H 1 and BMO
associated with the Gauss measure in [62, 63].
We can use the Gauss measure to show the boundedness of the Riesz
transform; see [358].
Section 6.2
General remarks and textbooks in Section 6.2
The textbooks [112, 147, 458] cover the topic exhaustively. A Radon mea-
sure µ satisfying µ(B(x, r)) . rd is called the Frostman measure. This plays
an important role in various aspects in analysis. See [9, 338, 339, 340, 434, 435,
436]. This type of assumption sometimes appears in the analysis of operators
handled in this book. See [112, Chapter 6] for fractional integral operators
on metric measure spaces with general Radon measures, which includes the
analysis of the maximal operator. See [112, Chapter 8, §3] for singular inte-
gral operators. See [233] for fractional integral operators on non-homogeneous
spaces. See [307] for analysis with general Radon measures.
Section 6.2.1
Terasawa proved Theorem 267 when the function (x, r) 7→ µ(B(x, r)),
x ∈ X and r > 0 is continuous in r [433, Theorem 2.4]. Sawano proved
Theorem 267 when X is separable [378, Theorem 2.1] based on Terasawa’s
result. See [378, Corollary 2.1] for Theorem 269.
See [234, 246] for more about the maximal operators on non-homogeneous
spaces.
292 Morrey Spaces
Section 6.2.2
In most cases, the vector-valued extension of the maximal operator associ-
ated with general Radon measures is done. For the modified maximal opera-
tor, see [378, Theorem 1.3] and [378, Theorem 2.1] for Theorems 270 and 274,
respectively. For the dyadic maximal operator associated with general Radon
measures, we refer to the textbook [89, Theorem A.18].
Section 6.2.3
Stempak constructed Examples 109 and 110; see [419, Example 2.1] and
[419, Example 2.2], respectively. See also [378, 418] for more examples showing
that the modification is necessary.
Chapter 7
Weighted Lebesgue spaces
293
294 Morrey Spaces
Although we do not go into the details, the Jacobi hypergroup is the weighted
Lebesgue spaces. Let α ≥ β ≥ −1 with (α, β) 6= (−1, −1). Then we set
µ(t) ≡ (sinh t)1+α (cosh t)1+β dt for t > 0. The space L2 (R+ , ∆) = L2 (R+ , µ)
is the weighted L2 -space called the Jacobi hypergroup.
Section 7.1 focuses on the one-weight (maximal) inequality
Z Z
p
M f (x) w(x)dx . |f (x)|p w(x)dx.
Rn Rn
Generally speaking when the weights on the both sides are the same, the
inequality is called a one-weight inequality or a one-weight norm inequality.
A more general inequality of the form kT f kLq (v) . kf kLp (w) is called a two-
weight inequality or a two-weight norm inequality. Section 7.2 deals with the
two-weight norm inequality of Hardy operators, (fractional) maximal opera-
tors and singular integral operators. We will obtain characterizations different
from those in Section 7.1.
Weighted Lebesgue spaces 295
C
w{x ∈ Rn : M f (x) > λ} ≤ kf kL1 (w) . (7.1)
λ
Here, λ > 0 and f ∈ L0 (Rn ). This is the weighted counterpart of the weak-
(1, 1) inequality of the Hardy–Littlewood maximal operator. Before we inves-
tigate the necessary and sufficient condition, we consider an example. Among
the weights, we are particularly interested in the power weights. Here, by a
power weight we mean a weight w of the form w(x) = |x|α , x ∈ Rn . We are
now interested in the case of w(x) ≡ |x|α with α ∈ R.
This can be checked by distinguishing two cases: c(Q) is far from the origin,
that is, |c(Q)| > 4n`(Q), and otherwise.
We are now oriented to a general case of weights. We answer the question
of characterizing w for which (7.1) holds.
Definition 70. For a weight w, define its A1 -characteristic by
Mw \
[w]A1 ≡ = inf {α > 0 : mQ (w) ≤ α essinfx∈Q w(x)} .
w L∞
Q∈Q
are in the position of using the duality L1 (Rn )-L∞ (Rn ). The above obser-
vations include the proof of kM kL1 (w)→WL1 (w) ∼ [w]A1 . Thus, the proof is
complete.
We now present some examples of A1 -weights. Roughly speaking there are
two ways to create A1 -weights. Theorem 281 below, which uses the maximal
operator of measures, can be used to create some concrete examples, while
Theorem 284 can be used for the theoretical aspects.
Theorem 281. Assume that µ is a Radon measure on Rn such that M µ(x)
is finite for a.e. x ∈ Rn . Then for 0 < δ < 1, w ≡ (M µ)δ is an A1 -weight,
where M µ is given by (4.7). Furthermore, A1 (w) .δ 1.
It matters that the implicit constant above does not depend on µ.
Proof We have to establish mQ (w) . w(x) whenever Q is a cube con-
taining x. We split µ according to 10Q. Write µ|A ≡ µ(A ∩ ·), the restriction
of µ to a measurable subset A. Split µ by µ = µ1 + µ2 with µ1 ≡ µ|10Q and
µ2 ≡ µ|(Rn \ 10Q). We will prove
mQ (M µ1 δ ) . w(x), mQ (M µ2 δ ) . w(x)
for all x ∈ Q. To deal with µ1 we use Example 66, the Kolmogorov inequality,
to obtain
δ
δ δ µ(10Q)
mQ (M µ1 ) . m10Q (M µ1 ) . . w(x).
|10Q|
massed at the origin and δ = na , which is a special case of Theorem 281. Then
w equals M µ modulo multiplicative constants.
We completely characterize the power weights in A1 .
Theorem 283. Let α ∈ R. Then | · |α ∈ A1 if and only if −n < α ≤ 0.
Proof Let α > 0. Then M [| · |α ] ≡ ∞. Thus, this simply paraphrases
Example 114 and Proposition 282 again.
We now invoke the Rubio de Francia iteration algorithm to create more
examples of A1 -weights: As usual let L0 (Rn ) denote the set of all measurable
functions f . Of course, we identify two functions if they are different by a set of
measure zero. Denote by M k the k-fold composition of the Hardy–Littlewood
maximal operator M .
Theorem 284. Let (E(Rn ), k · kE ) be a Banach space of L0 (Rn ) such that
M f ∈ E(Rn ) together with the estimate
kM f kE ≤ Dkf kE (7.2)
n
holds whenever f ∈ E(R ), where D is independent of f . Assume in addition
that any function in E(Rn ) is finite almost everywhere. Let g ∈ E(Rn ) with
∞
Mkg
. Then G belongs to E(Rn ) and satisfies
P
norm 1. Define G ≡ |g| + (2D)k
k=1
the following properties:
kGkE ≤ 2kgkE = 2 (7.3)
and
[G]A1 ≤ 2D. (7.4)
Proof Since the series definining G converges absolutely in the topology
of E(Rn ), G ∈ E(Rn ). This also implies (7.3). Hence, G is finite almost every-
where by our assumption. Note that G is everywhere positive because M g > 0.
Finally (7.4) is easy to check using (7.2) and the sublinearity of f , G ∈ A1 .
We end Section 7.1.1 with the characterization of Morrey spaces in terms
of A1 -weights. The following proposition can be used to express the Morrey
norm in terms of the weighted Lebesgue spaces:
q
Proposition 285. Let 1 < q ≤ p < ∞, and let 0 < θ < p. Then for all
1 1
0 n
f ∈ L (R ), kf k ∼ sup |Q| p−q kf k .
Mp
q ( 1 )
Q∈Q Lq (M 1−θ χ )
Q
1 1
Proof Since M χQ ≥ χQ , kf kMpq ≤ sup |Q| p − q kf k ( 1 ) . To show
Q∈Q Lq (M 1−θ χ )
Q
Here, 1 < p < ∞. We will introduce the class Ap . It should be noted that the
weighted weak-(1, 1) maximal inequality is equivalent to
or equivalently,
1 1
mQ (w− p |g|)w(Q) p ≤ N kgkLp (g ∈ Lp (Rn )).
0 1 p−1 1
By the duality Lp (Rn )-Lp (Rn ), we have mQ (w− p−1 ) p w(Q) p ≤ N. Thus,
w ∈ Ap together with the estimate N p ≥ [w]Ap since Q is arbitrary.
We polish Example 116 having somewhat clarified how the class Ap is
natural.
Example 117. Let w(x) = |x|−n+δ , x ∈ Rn with δ ∈ (0, n). Then [w]Ap ∼
1
( )
δ −1 . We calculate mQ (w)mQp−1 (w−1 ) according to the position of Q. If 0 ∈
/
1 1
( ) ( )
2Q, then mQ (w)mQp−1 (w−1 ) ∼ mQ (w(c(Q)))mQp−1 (w(c(Q))−1 ) = 1. If 0 ∈
1
( p−1 )
2Q, then mQ (w)mQ (w−1 ) ∼ δ −1 `(Q)−δ `(Q)δ = δ −1 .
0
Since Lp (Rn ) and Lp (Rn ) are dual to each other, it is natural to ask
ourselves whether the classes Ap and Ap0 are strongly related. Actually, the
following equality, whose easy proof we omit (see Exercise 97), follows from
the definition.
1
Lemma 287. If w ∈ Ap with 1 < p < ∞, then we have w− p−1 ∈ Ap0 with
1 0
[w− p−1 ]Ap0 = ([w]Ap )p −1 .
1
Motivated by Lemma 287, we occasionally call w− p−1 the dual weight of
w ∈ Ap , 1 < p < ∞.
We apply Lemma 287 to the power weights. Example 118 concludes our
calculation on the power weights.
Weighted Lebesgue spaces 301
Example 118. Let 1 < p < ∞, and let w(x) ≡ |x|−n+δ , x ∈ Rn , with
p0 0
0 < δ < n. Then v(x) ≡ w(x)− p = |x|(p −1)(n−δ) , x ∈ Rn , is an Ap0 -weight
0
with [v]Ap0 ∼ δ −(p −1) . As a result, W (x) ≡ v(x) = |x|(p−1)(n−δ) , x ∈ Rn , is
0 1
an Ap -weight with [W ]Ap ∼ δ −p +1 = δ − p−1 .
We completely characterize the power weights in Ap for 1 < p < ∞. Pay
attention to the endpoint. The condition with p = 1 obtained in Theorem 288
differs from the one in Theorem 283.
Theorem 288. Let α ∈ R, and let 1 < p < ∞. Then | · |α ∈ Ap if and only
if −n < α < n(p − 1).
Proof The necessity follows from Example 116. The sufficiency follows
from Examples 117 and 118 together with the fact that 1 ∈ Ap .
Here, we show that the Ap -condition is sufficient for the weak-(p, p) max-
imal inequality to hold. We need the weak-type estimate later:
Theorem 289. Let w ∈ Ap with p ∈ [1, ∞). Then
Thus, if the estimate λw{x ∈ Rn : M f (x) > λ} ≤ D([w]Ap )A kf kLp (w) holds,
1 1
then δ −1 ≤ Dδ −A(p−1)− p for all δ > 0, indicating that −1 ≥ −A(p − 1) − ,
p
or equivalently, Ap ≥ 1. Suppose we have
then 1 A
λ[w{x ∈ Rn : M f (x) > λ}] p ≤ D0 ([w]A1 ) p kf kLp (w) .
Going through a similar argument using Example 114, we obtain that the
power in the weighted weak-(1, 1) maximal inequality is sharp.
We now consider the weighted strong-(p, p) maximal inequality. As is
announced before, the power of [w]Ap differs from the one in the weighted
weak-(p, p) maximal inequality.
Theorem 290. Let w ∈ Ap with 1 < p < ∞. Then for all f ∈ Lp (w),
1
kM f kLp (w) .p,n ([w]Ap ) p−1 kf kLp (w) .
The proof hinges on the universal estimate. Let W : Rn → (0, ∞) be a
weight. Denote by Mc,W the centered weighted maximal operator given by
Z
Q 1
Mc,W f (x) = Mc,W f (x) ≡ sup |f (y)| W (y)dy. (7.5)
Q∈Q, c(Q)=x W (5Q) Q
Inserting this pointwise estimate into kM f kLp (w) and then using the bound-
edness of Mc,w , we obtain
1 1
kM f kLp (w) .p,n [w]Ap−1
p
kMc,w [(Mc,σ [f · σ −1 ])p−1 w−1 ] p−1 kLp (w)
1
≤ ([w]Ap kMc,w [(Mc,σ [f · σ −1 ])p−1 w−1 ]kLp0 (w) ) p−1
1
.p ([w]Ap k(Mc,σ [f · σ −1 ])p−1 w−1 kLp0 (w) ) p−1
1
= ([w]Ap k(Mc,σ [f · σ −1 ])p−1 kLp0 (σ) ) p−1
1
= [w]Ap−1
p
kMc,σ [f · σ −1 ]kLp (σ) .
Let a ∈ {0, 1/3}n be fixed. Suppose that we have a family Sj = {Qjk }k∈Kj
∞
S
of disjoint cubes in Da for each j ∈ N0 . Recall that S ≡ Sj is an η-sparse
j=0
304 Morrey Spaces
[ [ j X
family with a level structure if Qj+1
k ⊂ Qk and |Qj+1
k ∩R| ≤
k∈Kj+1 k∈Kj k∈Kj+1
η|R| for all R ∈ Sj , j ∈ N. We also remark that a weight w belongs to A2
if and only if w−1 belongs to A2 . Hence, the class A2 is special. Due to this
special property we can prove the following boundedness. We note that the
estimate [w]A2 is optimal in view of the A2 -theorem, which we consider later.
Lemma 292. Let a ∈ {0, 1/3}n . Let {Qjk }j∈N,k∈Kj ⊂ Da be a sparse family
of generalized dyadic cubes with a level structure. As long as the sum makes
sense, define
∞ X
X
Af ≡ mQj (f )χQj
k k
j=1 k∈Kj
for f ∈ L1loc (Rn ). Then kAf kL2 (w) . [w]A2 kf kL2 (w) for all f ∈ L2 (w) and
w ∈ A2 .
Before the proof, a helpful remark may be in order. As is easily seen from
the inequality kf χQ kL1 ≤ kf kL2 (w) kχQ kL2 (w−1 ) , if f ∈ L2 (w) and w ∈ A2 ,
then f ∈ L1loc (Rn ). Hence, at least the definition of mQkj (f ) in Af makes
sense if f ∈ L2 (w).
For the proof we use the following notation. For a weight v, f ∈ L0 (Rn )
and a ∈ {0, 1/3}n , we define
Z
χQ (x)
Da
Mv f (x) ≡ sup |f (y)|v(y)dy (x ∈ Rn ).
Q∈Da v(Q) Q
Proof Write σ ≡ w−1 . We dualize the left-hand side: Matters are reduced
to showing
Z
I≡ Af (x)g(x)dx . [w]A2 kf kL2 (w) kgkL2 (σ)
Rn
∞ X
X
I. mQj (f )mQj (g)|Ekj |. (7.6)
k k
j=1 k∈Kj
Since w ∈ A2 , we have
mQj (f )mQj (g)
k k
Z Z
1 1
≤ [w]A2 · f (x)w(x)σ(x)dx · g(x)σ(x)w(x)dx.
σ(Qjk ) Qjk w(Qjk ) Qjk
Weighted Lebesgue spaces 305
If we insert this estimate into the above expression, we obtain the desired
result.
Example 122. Let 1 < q ≤ p < ∞. Then M is bounded on Mpq (Rn ). In fact,
q
for f ∈ Mpq (Rn ), letting θ = 2p , we obtain
1 1
kM f kMpq ∼ sup |Q| p − q kM f k ( 1 )
Q∈Q Lq (M 1−θ χ )
Q
1
thanks to Proposition 285. Meanwhile, since M ( 1−θ ) χQ ∈ A1 , we have
1 1 1 1
sup |Q| p − q kM f k ( 1 ) . sup |Q| p − q kf k ( 1 ) .
Q∈Q Lq (M 1−θ χQ ) Q∈Q Lq (M 1−θ χ )
Q
If we use Proposition 285 once again, kM f kMpq . kf kMpq . This gives another
proof of Theorem 382 to follow. This simple proof seems to be applicable to
many things. However, as it turns out, in some generalized cases, we need
different discussions.
To conclude this section, we remark that we can consider the local coun-
terpart of the results in this section. The details are omitted.
For the time being we investigate the property of the class A∞ . Let w be a
(1+ε)
weight. If we use Hölder’s inequality, then mQ (w) ≥ mQ (w) for all weights
w. However, since w ∈ L1loc (Rn ) does not imply w ∈ L1+ε n
loc (R ) for any ε > 0,
(1+ε)
the reverse inequality mQ (w) ≤ mQ (w) and its weaker version
(1+ε)
mQ (w) . mQ (w) (7.7)
fails.
However, if the weight belongs to the class A∞ , then the reverse estimate
can be obtained in the following sense:
Theorem 295 (Reverse Hölder inequality). Let w ∈ A∞ . We abbreviate
1 (1+ε)
ε ≡ n+3 > 0. Then mQ (w) ≤ 2mQ (w) for all cubes Q ∈ Q.
2 [w]A∞
Proof By the dilation and the translation, we may assume that Q ∈
D(Rn ). We will use cubes in D(Q). Thus, we may assume
P that w ∈ L∞ (Rn )
by approximating w with functions of the form mR (w)χR for some
R∈Dj (Rn )
D(Q)
j ∈ Z. Let EQ,λ ≡ {x ∈ Q : M [χQ w](x) > λ} for λ > 0. We use the
Layer-Cake formula to obtain
Z Z ∞
I≡ M D(Q) [χQ w](x)ε w(x)dx = ε λε−1 w(EQ,λ )dλ.
Q 0
Let λ > mQ (w). Consider the set {Qj }j∈J(λ) ⊂ D(Q) of all maximal dyadic
cubes in the set EQ,λ whose average of w exceeds λ. Since λ > mQ (w), each
Qj is a proper subset of Q. So thanks to the maximality of each Qj , we have
X X
w(EQ,λ ) = w(Qj ) ≤ 2n λ |Qj | = 2n λ|EQ,λ |.
j∈J(λ) j∈J(λ)
Thus
Z mQ (w) Z ∞
I≤ε λε−1 w(EQ,λ )dλ + 2n ε λε |EQ,λ |dλ
0 mQ (w)
2n ε
Z
≤ mQ (w)1+ε |Q| + M D(Q) [χQ w](x)1+ε dx.
1+ε Q
Since w ∈ A∞ , mR (w) ≤ [w]A∞ exp (mR (log w)) for all R ∈ D(Rn ). This
implies that
D(Q)
M D(Q) w(x) ≤ [w]A∞ Mlog w(x) = [w]A∞ sup χR (x) exp (mR (log w))
R∈D(Q)
Note that
1
n+3 [w]
2n eε[w]A∞ 2n e[w]A∞ ([w]A∞ ) 2
A∞ e 1 1
([w]A∞ )ε = ≤ exp ≤ ,
1+ε 2n+3 [w]A∞ + 1 8 16e 2
since ea ≤ 1 + 2a for 0 < a < 0.1. Thus, it follows that
Z Z
1
w(x)1+ε dx ≤ mQ (w)1+ε |Q| + w(x)1+ε dx.
Q 2 Q
Since w is assumed to be bounded, we obtain the desired result.
The following corollary can be located as the weak version of Corollary
291.
Lemma 296. Let E be a Lebesgue measurable set contained in a cube Q.
n+4
Also let w ∈ A∞ and a > 2n+1+2 [w]A∞ . Assume that |Q| ≤ 2n a−1 |E|.
Then w(Q) .[w]A∞ w(E).
Proof Since w(Q) = w(E) + w(Q \ E), it suffices to show that
1 2n+3 [w]A
w(Q \ E) 1 1+2n+3 [w]A
∞ −
1+2n+4 [w]A
∞
≤2 n+4 [w] =2 ∞ (< 1). (7.8)
w(Q) 21+2 A∞
1
Let ε ≡ as before. First, observe that |Q \ E| ≤ 2n a−1 |Q|. As a
2n+3 [w]A∞
result, by Theorem 295 we obtain
Z 1
1+ε
ε
1+ε
w(Q \ E) ≤ w(x) dx |Q \ E| 1+ε
Q
1 n ε
1+ε
|Q| 1+ε 2
≤ 2 mQ (w) |Q| w(Q)
w(Q) a
n 1+εε
2
= 2 w(Q)
a
n+31
1 1+2 [w]A
∞
≤ 2 1+2 n+4 [w] .
2 A ∞
Recall that two dyadic cubes are disjoint unless one is contained in the other.
Therefore we have
X |Qk+1,j | |Qk,j0 |
w(Ωk+1 ∩ Qk,j0 ) ≤ ≤ .
λk+1 λk+1
j;Qk+1,j ∈D(Qk,j0 )
|Qk,j0 |
If we use (7.10) once again, then w(Ωk+1 ∩ Qk,j0 ) ≤ ≤ αw(Qk,j0 ). We
λk+1
can pass the above inequality to the unweighted one: |Ωk+1 ∩Qk,j0 | ≤ β|Qk,j0 |.
Adding the above inequality over j0 , we obtain
Z ∞
dx X
≤ λk+1 τ |Ωk |
Q w(x)τ
k=−1
∞ kτ
τ
X D
. λ0 β k |Ω0 |
α
k=−1
∞ kτ
τ
X D
≤ λ0 |Q| βk.
α
k=−1
q1
X∞
k{fj }∞
j=0 k`q (Lp (w)) ≡ (kfj kLp (w) )q ,
j=0
q1
X∞
k{fj }∞
j=0 kLp (w,`q ) ≡ |fj |q .
j=0
Lp (w)
Define `q (Lp (w)) and Lp (w, `q ), called weighted vector-valued Lebesgue spaces,
to be the collection of all sequences {fj }∞ j=0 of measurable functions for which
the norms k{fj }∞j=0 k` q (Lp (w)) and k{fj }
∞
j=0 kLp (w,`q ) are finite, respectively. As
applications of Theorem 270, we can prove the following inequality:
Theorem 298 (Weighted vector-valued inequality for M ). Let 1 < p, q < ∞,
1 < r ≤ ∞ and w ∈ Ap . Then k{M fj }∞ ∞
j=0 kLp (w,`q ) .p,q,r k{fj }j=0 kLp (w,`q )
∞ 0 n
for all {fj }j=1 ⊂ L (R ).
Proof Since w ∈ Ap̃ for some p̃ ∈ (1, p) by virtue of Theorem 295, we see
p p̃
that M fj . M22,c,µ [|fj | p̃ ] p , where µ ≡ wdx and M22,c,µ is defined in Section
6.2.1 with k = 22. It remains to combine Theorem 270 with this pointwise
estimate.
−q
1 1
(1) We claim that (u, σ) ≡ (M w) p0 , w ∈ Aα
p,q . In fact, mQ (u) mQ (σ)
q p0 =
− pq0 1 1
mQ ((M w) ) q mQ (w) p0 . If x ∈ Q, then M w(x) ≥ mQ (w). Therefore,
we have
1
1 − p10 1
mQ (u) q mQ (σ) p0 ≤ mQ (w) mQ (w) p0 = 1.
p0
(2) Likewise, we can check that (u, σ) = (w, (M w)− q ) ∈ Aα
p,q .
312 Morrey Spaces
If the weights are radial and monotone, we can calculate the quantity with
ease.
Example 126. Let p, q > 1 and 0 ≤ α < n. Also let v, w be weights. Assume
that there exist ŵ ∈ M↑ [0, ∞) and v̂ ∈ M↓ [0, ∞) such that w(x) = ŵ(|x|),
1 1
v(x) = v̂(|x|) for all x ∈ Rn . Define σ ≡ w− p−1 and σ̂ ≡ w− p−1 . We compare
1 1
I ≡ sup Rα−n kσ̂χ(0,R) kL1 (tn−1 ) q0 kv̂χ(0,R) kL1 (tn−1 ) p and
R>0
1 1
J ≡ sup Rα−n σ(B(x, R)) q0 v(B(x, R)) p .
x∈Rn ,R>0
Since v̂ ∈ M↓ (0, ∞), v(B(x, r)) ≤ v(B(r)) for any x ∈ Rn and r > 0 and,
since ŵ ∈ M↑ (0, ∞), we have σ(B(x, r)) ≤ σ(B(r)). Hence, J ∼ I.
For the class Aα
p,q , we have a weak type estimate.
Theorem 299. Let 1 < p ≤ q < ∞ and 0 ≤ α < n. Then for any pair of
weights (u, σ) ∈ Aα
p,q we have
1
sup λ(u{x ∈ Rn : Mα f (x) > λ}) q . [u, σ]Aαp,q kf kLp (σ1−p ) (7.12)
λ>0
Proof We use Theorem 179. Let Q0 be a fixed dyadic cube. By the mono-
tone convergence theorem, we have only to work in Q0 . Set
X
E ≡ x ∈ Q0 : `(Q)α m3Q (|f |)χEQ (x) > λ
Q∈S
S
for λ > 0 when we have a sparse family S, where EQ ≡ Q \ R
R∈S∩D(Q)\{Q}
for each Q ∈ S. We use Hölder’s inequality to have
1
sup `(Q)α mQ (|f |) u(Q0 ) q . [u, σ]Aαp,q kf kLp (σ1−p ) .
Q∈Q] (Q0 )
1−p 1
(p)
By Hölder’s inequality we have m3Q (|f |) ≤ m3Q (f · σ p )m3Q (σ) p0 . As a
result, we obtain
X h iq
α+ n (p) 1−p 1
λq u(E) . `(Q) p0 m3Q (f · σ p )σ(3Q) p0 u(Q)
Q∈S ∗
X 1−p
q
(p)
. [u, σ]Aαp,q q |Q|m3Q (f · σ p )
Q∈S ∗
pq
X (p) 1−p
. [u, σ]Aαp,q q |Q|m3Q (f · σ p )p .
Q∈S ∗
0 0
for all f ∈ Lp (wp ) ∩ M+ (Rn ) and g ∈ Lq (w−q ) ∩ M+ (Rn ) and all sparse fam-
ilies S. Although we need to take into account the vector ~e, due to similarity,
we assume ~e = 0 and hence S ⊂ D(Rn ).
By Hölder’s inequality, we obtain
Z Z
X β 1 β 1
I= `(Q)α−n σ(Q)v(Q)1− n f (x)dx · v(Q) n g(x)dx
σ(Q) Q v(Q) Q
Q∈S
X β
≤ `(Q)α−n σ(Q)v(Q)1− n inf MσD [f · σ −1 ](x) inf Mv,β
D
[g · v −1 ](x).
x∈Q x∈Q
Q∈S
D
By Theorem 257, the universal estimates of MσD and Mv,β , we obtain
Z p1 Z 10
0 0 q
I. f (x)p σ(x)−p σ(x)dx g(x)q v(x)−q v(x)dx
Rn Rn
= kf kLp (wp ) kgkLq0 (w−q0 ) .
Since Mα . Iα , we can transplant Theorem 300 to Mα .
Theorem 301. Let 1 < q ≤ p < ∞ and 0 < α < n. If w ∈ Aα
p,q , then
kMα f kLq (wq ) . kf kLp (wp ) for all f ∈ L0 (Rn ).
7.1.5 Extrapolation
We have established and will establish that the fundamental operators in
harmonic analysis behave well with respect to these classes of weights; see
Theorems 280, 289, 290 and 298 for the Hardy–Littlewood maximal operator
and Theorem 305 to follow for singular integral operators. When we consider
estimates of Ap -weights, we face the following type of estimate:
kT hkLp (W ) ≤ N ([W ]Ap )khkLp (W ) (h ∈ Lp (W ), W ∈ Ap ), (7.13)
Weighted Lebesgue spaces 315
Since D 1,
Z 10
0 r p0
H(x)p w(x)dx . (khkLp0 (w) ) r0 . (7.17)
Rn
we apply Theorem 302 to F ≡ FJ for each J ∈ N and then apply the monotone
convergence theorem. This is possible, because we have kgkLq (w) ≤ kf kLq (w)
for all (f, g) ∈ FJ and w ∈ Aq as long as D 1.
We present another application. As we did in Lemma 292, define
∞ X
X
Af ≡ mQj (f )χQj
k k
j=1 k∈Kj
7.1.6 A2 -theorem
So far we have seen how the constant behaves in the weighted maximal
estimate in Theorems 280, 289 and 290. We will consider the boundedness
property of singular integral operators. The behavior of the constant had been
unknown and it is known as the A2 -problem/A2 -conjecture. This was settled
down by Petermichl [355, 356] and then was expanded by Hytonen and Lerner
[207, 270]. Nowadays, this theorem is called the A2 -theorem. We establish the
following A2 -theorem here.
318 Morrey Spaces
Earlier it was known that kT f kLp (w) .[w]Ap kf kLp (w) for all f ∈ L∞ n
c (R )
[205].
For the proof of Theorem 305, we need to set up notation. For a sparse
n
family S ≡ {Ukj }j∈N,k∈Kj ⊂ D ≡ D
Q
[aj , bj ) with a level structure and
j=1
∞ X
X
f ∈ L1loc (Rn ) write AD,S f ≡ mU j (f )χU j . To prove Theorem 305, we
k k
j=1 k∈Kj
use the following estimate together with Lemma 218:
Lemma 307. Let m, N ∈ N. Suppose that L ≡ {Qjk }j∈N,k∈Kj ⊂ D(Rn ) is
a sparse family with a level structure and that K0 = {k} is a singleton with
Q0k = [−2N , 2N )n for k ∈ K0 . Let L0 ≡ {Rkj }j∈N,k∈Kj ⊂ D(Rn ) be a family
such that Qjk ⊂ 3Rkj and that Rkj ⊂ 2m Qjk for all j ∈ N and k ∈ Kj . Write
L0 ≡ Rkj if L = Qjk . Fix f ∈ L∞ n + n
c (R ) ∩ M (R ). Define
X
AL,L0 ;m f ≡ mL0 (f )χL .
L∈L
(1) We have kAL,L0 ;m f kL2 . kf kL2 and kA∗L,L0 ;m f kL2 . kf kL2 , where the
implicit constants are independent of f and m.
(2) There exists a constant c(n) that depends only on n such that
(1) We concentrate on the first formula, since the second one can be handled
completely in a similar manner or by the duality argument. To prove
the first formula, we dualize the left-hand side: Matters are reduced to
the inequality
Z
AL,L0 ;m f (x) · g(x)dx . kf kL2 kgkL2
Rn
for all f, g ∈ L∞ n + n
c (R ) ∩ M (R ). Written out in full, the left-hand side
becomes
Z X
AL,L0 ;m f (x) · g(x)dx = mL0 (f )mL (g)|L|.
Rn L∈L
[
We define EL ≡ L \ Qjk0 , if L = Qjk . Then keeping in mind that
k0 ∈Kj+1
{EL }L∈L is disjoint and that 2|EL | ≥ |L|, we have
Z X
AL,L0 ;m f (x) · g(x)dx ≤ 2 m2m L (f )mL (g)|EL |
Rn L∈L
Z
≤2 M f (x)M g(x)dx
Rn
≤ 2kM f kL2 kM gkL2 .
Then it is easy to see that the good part g can be handled easily because
we already know that A∗L,L0 ;m is bounded on L2 (Rn ). To handle the bad
part b, we have only to show
[
λ x ∈ Rn \ 15Q : |A∗L,L0 ;m b(x)| > λ ≤ c(n)mkf kL1 .
Q∈N
We decompose
X X X 1 Z
A∗L,L0 ;m b = A∗L,L0 ;m bQ = bQ (y)dy χL0 .
|L0 | L
Q∈N Q∈N L∈L
320 Morrey Spaces
[
Here and below we fix x ∈ Rn \ 15Q. Let Q ∈ N and L ∈ D(Rn ).
Q∈N
If L ∩ Q = ∅ or L ⊃ Q, then mL (bQ ) = 0. Thus
Z
X X 1
A∗L,L0 ;m b(x) = bQ (y)dy χL0 (x).
|L0 | L
Q∈N L∈L∩D(Q)
= M f + mAD,S f.
By the use of Lemmas 218, we obtain the weighted estimate of the operator
A∗L,L0 ;m .
Proposition 308. Maintain the notation in Lemma 307. Let w ∈ A2 . Then
kAL,L0 ;m f kL2 (w) . m[w]A2 kf kL2 (w) , kA∗L,L0 ;m f kL2 (w) . m[w]A2 kf kL2 (w)
for all f ∈ L∞ n
c (R ).
Proof We may assume that L is a finite set by the use of the monotone
convergence theorem and that f ∈ M+ (Rn ). It suffices to prove the second
formula thanks to the duality L2 (w−1 )-L2 (w) and the fact that w−1 ∈ A2
with [w−1 ]A2 = [w]A2 . Then we calculate
X
m2m L (f )χL . m[w]A2 kf kL2 (w) .
L∈L L2 (w)
n
Proof For each L ∈ L, there exist disjoint 2n -dyadic cubes L1 , L2 , . . . , L2
[2n
m
k m n
such that `(L ) = 2 `(L) for each k = 1, 2, . . . , 2 and that 2 L ⊂ Lk .
k=1
322 Morrey Spaces
X
mL(k) (f )χL . m[w]A2 kf kL2 (w)
L∈L L2 (w)
x1 − y1
Z
1
R1 f (x) = n+1
f (y)dy & .
Rn |x − y| δ|x|n
As a result,
! p1
|x|(p−1)(n−δ)
Z
1
kR1 f kLp (w) & dx & δ − p −1 .
x1 ≥max(2,x2 ,x3 ,...,xn ) |x|pn
Meanwhile ! p1
Z
1
kf kLp (w) = |x|−n+δ
dx ∼ δ− p .
B(1)
By duality
or equivalently,
A
kR1 gkLp0 (σ) .n ([σ]Ap0 ) p0 −1 kgkLp0 (σ) (g ∈ L∞ n
c (R ), σ ∈ Ap0 ).
for almost every x ∈ Rn , where {Qjk }j∈N,k∈Kj is a sparse family with a level
structure. By Lemma 217, we have
∞ X
X ∞ X
χQ0 (x)|T f (x) − Med(T f ; Q0 )| . 2−l m2l Q0 (|f |)χQj (x)
k
l=0 j=1 k∈Kj
for almost every x ∈ Rn . At this step, the singular integral operator T was
controlled by a positive operator. Thus, by Corollary 309,
∞ X
X ∞ X
kχQ0 (T f − Med(T f ; Q0 ))kL2 (w) . 2−l m2l Q0 (|f |)χQj
k
l=0 j=1 k∈Kj
∞
X
. l · 2−l [w]A2 kf kL2 (w)
l=0
∼ [w]A2 kf kL2 (w) .
k{T fj }∞ ∞
j=0 kLp (w,`q ) .q,r k{fj }j=0 kLp (w,`q ) .
7.1.7 Exercises
Exercise 97. Go back to the definition of the Ap /Ap0 -characteristics to show
Lemma 287.
Exercise 98. Let w be a weight on Rn . Consider the uncentered maximal
operator M generated by cubes.
(1) Let f ∈ L0 (Rn ). Then show that f ∈ L1 (w)0 if and only if w−1 f ∈ L∞ .
Furthermore, in this case, kw−1 f kL1 (w)0 = kw−1 f kL∞ .
(2) Show that M maps L1 (w)0 boundedly if and only if w ∈ A1 . Further-
more, in this case, find the operator norm of M .
Exercise 99. Consider the uncentered maximal operator M generated by
cubes.
(1) Let f ∈ L0 (Rn ). Let λ > 0 and define Ω ≡ {x ∈ Rn : M f (x) > λ}.
Show that, for any compact set K contained in Ω, there exists a finite
collection {Qj }Jj=1 of cubes such that mQj (|f |) > λ for all j = 1, 2, . . . , J
SJ
and that K ⊂ Qj .
j=1
(2) Prove Theorem 289 for p > 1 using Hölder’s inequality and the 5r-
covering lemma.
ExerciseZ 100. [135, 450] Let w be a weight. Then show that w ∈ A∞ if and
only if M [wχQ ](x)dx . w(Q) for any cube Q.
Q
or equivalently
kHf kLp (tp dµ) ≤ Ckf kLp (wp ) ,
+
where f ∈ M (0, ∞). If p = q, thanks to Theorem 38 we have rich examples
of monomial weights. In the upper triangle case, 1 < p ≤ q < ∞, we can
completely characterize the boundedness of the integral operators.
Theorem 312. Let 1 < p ≤ q < ∞. Suppose that w ∈ M+ (0, ∞) and that µ
is a continuous Borel measure on (0, ∞). Then
Z ∞ Z v q q1
f (t)dt dµ(v) . kf · wkLp (0,∞)
0 0
We prove Theorem 312 after we prove a lemma. The following lemma can
be located as a sort of integration by parts.
Lemma 313. Let 1 < p < ∞ and µ be a continuous Borel measure. Then
Z
1 p
p0
p dµ(s) ≤ p p µ(t, ∞).
(t,∞)∩supp(µ) µ(s, ∞)
h s i
Proof Let s > t > 0 and N ∈ N. Then define tN (s) ≡ 2N t N . By the
2 t
monotone convergence theorem,
Z Z ∞
1 1
p0
p dµ(s) = lim lim p0
p dµ(s)
(t,∞)∩supp(µ) µ(s, ∞) ε↓0 N →∞ t µ(−ε + tN (s), ∞)
∞
X µ(t + 2−N (k − 1)t, t + 2−N kt)
≤ lim p0
p ,
N →∞
k=1
µ[t + 2−N kt, ∞)
326 Morrey Spaces
for each fixed ε > 0 since if ε > 2−N for N large enough. By the mean-value
theorem,
Z
1
p0
p dµ(s)
(t,∞)∩supp(µ) µ(s, ∞)
X∞ q q
≤ lim p p µ(t + 2−N kt, ∞) − p µ(t + 2−N (k − 1)t, ∞)
N →∞
k=1
p
= p p µ(t, ∞).
We prove Theorem 312.
Proof The “only if” part is easy to check; simply use that
Z r Z ∞ Z r q q1
p
q
f (t)dt · µ(r, ∞) = f (t)dt dµ(v) . kf · wkLp (0,∞)
0 r 0
for all r > 0 and resort to duality. Consequently, we concentrate on the “if”
part.
Z t p10 p
−p0
Let h(t) ≡ w(s) ds for t ∈ (0, ∞). Then for all s > 0,
0
1 1 d 0 0 0 d
= h(s)pp = p0 ph0 (s)h(s)p p−p −1 = p0 h(s)p .
h(s)p0 w(s)p0 p 0
h(s) ds ds
Thus,
t t p10
Kp0
Z Z
1 −p0
ds = p0 h(t)p = p0 w(s) ds ≤ p .
0 h(s) w(s)p0
p 0
0
q
µ(t, ∞)
Meanwhile, by Hölder’s inequality, we have
Z ∞ Z v q
f (t)dt dµ(v)
0 0
Z ∞ Z v q
= f (t)h(t)w(h)h(t)−1 w(t)−1 dt dµ(v)
0 0
Z ∞ Z s pq Z s q0
p p p 1 p
≤ f (t) h(t) w(t) dt dt dµ(s).
0 0 0 h(t)p0 w(t)p0
Recall that we are assuming that p ≤ q. By virtue of Minkowski’s inequality
Z s q0
dt p
which is applied to the weighted measure p 0 p 0 dµ(s), we have
0 h(t) w(t)
Z ∞ Z s q q1
f (t)dt dµ(s)
0 0
# pq p1
Z ∞
"Z
∞ Z s q0
dv p
≤ f (t)p h(t)p w(t)p dµ(s) dt .
0 t 0 h(v)p0 w(v)p0
Weighted Lebesgue spaces 327
As a result,
Z ∞ Z s q q1
f (t)dt dµ(s)
0 0
! pq p1
Z ∞ Z
1 dµ(s)
≤K p0 f (t)p h(t)p w(t)p p0
p dt .
0 (t,∞)∩supp(µ) µ(s, ∞)
Consequently,
(Z q ) q1
∞ Z t
f (s)ds dµ(t) . Kkf · wkLp (0,∞) .
0 0
As is seen by the proof below, Theorems 312 and 314 are equivalent.
Proof Let ν be a measure satisfying ν(E) = µ{t > 0 : t−1 ∈ E} for all
Borel sets E. Then
Z ∞ Z ∞ q Z ∞ Z ∞ q
f (s)ds dµ(t) = f (s)ds dν(t)
0 t 0 t−1
Z ∞ Z t q
−2 −1
= s f (s )ds dν(t).
0 0
Consequently, we have
(Z q ) q1 p1
∞ Z t Z ∞ 2
g(s)ds dν(t) . (g(t)t p0 w(t−1 ))p dt
0 0 0
The proof of Theorem 315 will be given after we prove two lemmas. First,
under our assumption on the function f , we estimate the Hardy operator H
so that we can understand why max in the definition of A comes into play.
Lemma 316. Let 0 < p ≤ 1. Also let f ∈ C 1 (0, ∞) ∩ M↓ (0, ∞) satisfy
Z t p
1
lim f (t) = 0. Then Hf (t) . f (t) + − up f (u)p−1 f 0 (u)du for all
t→∞ t 0
t > 0.
Proof Thanks to the fundamental theorem of calculus and
p1the fact that
Z ∞
1 t
Z
p−1 0
lim f (t) = 0, we have Hf (t) ' − f (u) f (u)du ds. We split
t→∞ t 0 s
the integral into two parts:
Hf (t)
p1 Z Z ∞ p1
1 t
Z Z t
p−1 0 1 t p−1 0
. − f (u) f (u)du ds + − f (u) f (u)du ds.
t 0 s t 0 t
We consider
(Z Z
t t p1 )p
p−1 0
I≡ − χ(0,u) (s)f (u) f (u)du ds .
0 0
For the first term, we use Lemma 13. Recall that 0 < p ≤ q < ∞. Apply-
ing Minkowski’s inequality to the weighted measures f (s)p−1 f 0 (s)sp ds and
t−q dµ(t), we find
(Z Z q ) pq
∞ t p
∼ Ap f (t)p dν(t).
0
For the first term, we use (1.8). Applying the assumption A < ∞, we find
Z t p1
w(t) p−1 0 p
sup − f (u) f (p)(u)u du
t>0 t 0
Z t p p1
p−1 0 p −1
≤ sup − f (u) f (p)(u)u sup s w(s) du
t>0 0 s≥u
Z ∞ Z y p1
p−1 0
≤A − f (y) f (y) v(s)ds dy
0 0
. Akf kLp (v) .
Finally we will list some sufficient conditions for which the Hardy operator
is bounded.
Theorem 318. Let 0 < θ2 ≤ ∞, and let v1 , v2 ∈ M+ (0, ∞).
(1) Let 1 < θ1 ≤ ∞. Then
(Z
∞ Z t θ1 ) θ11
v2 (t) f (s)v1 (s)ds dt . kf kLθ1 (0,∞) (7.3)
0 0
Z ∞ Z t θθ20
+ θ2 θ10 1
for all f ∈ M (0, ∞) if A ≡ v2 (t) v1 (s) ds dt < ∞.
0 0
!
∞
v1 (s)θ2
Z
θ2
(3) Let 0 < θ1 ≤ min(1, θ2 ) and suppose v2 (t) sup 1−θ1 dt <
0 s∈[0,t] sθ2 θ1
(2) Simply mimic the previous case; we leave the proof for interested readers
as Exercise 101.
(3) Thanks to Lemma 14,
Z ∞ Z t θ 2 Z ∞ Z t θθ2
1
θ2 θ1 −1 θ1
v2 (t) g(s)ds dt ≤ v2 (t) s g(s) ds dt.
0 0 0 0
α α
X
LD;S
α f ≡ |Q| n mQ (f )χEQ and cD
∞,α (f ) ≡ sup |Q| n mQ (f ).
Q∈S Q∈D ] (Q0 )
Although we are mainly concerned with the case a = (0, 0, . . . , 0), we state
our main result in terms of Mα instead of MαD .
Weighted Lebesgue spaces 333
Theorem 319 (Cube testing for Mα ). Let v, w be weights, and let 0 ≤ α < n
1
and 1 < p ≤ q < ∞. Define σ ≡ w− p−1 , the dual weight of w. Then
1
kMα [σχQ0 ]kLq (v) . σ(Q0 ) p (7.6)
In particular, condition (7.5) implies that w(x) 6= 0 for almost all x ∈ Rn and
that σ, v ∈ L1loc (Rn ).
Here, we tolerate the case α = 0. It is important that the fundamental
operator Mα is essentially replaced by a somewhat strange operator LSα , whose
role in other fields of mathematics is yet unknown. Also, it matters that LSα
depends on f .
Proof The “only if” part is easy to check. Simply test (7.5) on f = χQ0 σ.
Once we have (7.5), then (7.7) follows immediately from `(Q)α mQ (|f |)χQ ≤
Mα f for all f ∈ L0 (Rn ).
Let us check the “if” part. We will use the fractional maximal operator MαD
defined in Remark 7. Fix a cube Q0 ∈ D(Rn ) to show that kχQ0 MαD f kLq (v) .
kf kLp (w) with the constant independent of Q0 . As we mentioned, although we
need to consider all families Da with a ∈ {0, 1/3}n , we consider D = D(0,0,...,0)
only for simplicity.
P We αmay as well assume that f ∈ M+ (Rn ). We consider
LD;S
α : LD;S
α f = |Q| n m (f )χ
Q EQ and c D
∞,α based on Theorem 179, where
Q∈S
S ⊂ D(Rn ) is a sparse family. We write
X q
V ≡ `(Q)αq (mQ (σ)) χEQ v.
Q∈S
Recall that S and Sλ are subsets of D(Rn ). We define Sλmax to be the set of
all maximal cubes in Sλ keeping this in mind. Then
[
{x ∈ Rn : |mS h(x)| > λ} ⊂ Q.
max
Q∈Sλ
where we used q ≥ p for the second inequality. Since we know that λσ(Q) ≤
khχQ kL1 (σ) for all Q ∈ Sλmax by the definition of Sλ , it follows from the
disjointness of Sλmax that
q
khkL1 (σ) p
n S
V {x ∈ R : |m h(x)| > λ} . .
λ
Next, we will show that cD
∞,α (f )kχQ0 kLq (v) . kf kLp (w) . To this end, we
need to choose a cube Q ∈ D which contains Q0 and we need to show that
α
|Q| n m3Q (|f |)kχQ0 kLq (v) . kf kLp (w) .
α
To this end, since Q0 ⊂ Q, it is enough to show that |3Q| n m3Q (|f |)kχ3Q kLq (v) .
kf kLp (w) . However, this is already done by the previous estimation. In fact,
the family {3Q} is a sparse family.
Thus,
cD S
∞,α (f )kχQ0 kLq (v) + kχQ0 m [f · σ
−1
]kLq (V ) . kf kLp (w) + kf · σ −1 kLp (σ)
∼ kf kLp (w) .
Thus, the proof is complete.
Concerning condition (7.7), we have a clarifying remark.
Weighted Lebesgue spaces 335
Remark 14. Assume that w, v ∈ M+ (Rn ) \ {0}. Then by the Lebesgue dif-
ferentiation theorem there exists x0 ∈ Rn such that
lim mB(x0 ,r) (σ) > 0, lim mB(x0 ,r) (v) > 0.
r↓0 r↓0
α 1 1 1 1
By condition (7.7), lim |B(x0 , r)| n − p + q mB(x0 ,r) (σ) p0 mB(x0 ,r) (v) q ≤ J < ∞.
r↓0
n n
Hence, the condition α ≥ − is necessary for the validity of (7.5).
p q
or
W2 (x) . inf W1 (y) for x ∈ Rn . (7.11)
y∈B(4|x|)\B( 41 |x|)
Then
kT f kLq (W2 q ) ≤ Dkf kLq (W1 q ) (f ∈ L∞ n
c (R ))
and
J2 ≡ sup r−n kW1 −1 kLq0 (B(r)) kM χB(r) W2 kLq (Rn \B(2r)) < ∞. (7.13)
r>0
∞ Z
X
|T [χRn \B(2j+2 ) f ](x)|q W2 (x)q dx . 1 (7.15)
j=−∞ B(2j+1 )\B(2j )
and
∞ Z
X
|T [χB(2j−1 ) f ](x)|q W2 (x)q dx . 1. (7.16)
j=−∞ B(2j+1 )\B(2j )
X∞ Z
≤ sup W2 (x)q · |T [χB(2j+2 )\B(2j−1 ) f ](x)|q dx
j+1 )\B(2j ) B(2j+1 )\B(2j )
j=−∞ x∈B(2
∞
X Z
. sup W2 (x)q · |f (x)|q dx
j+1 )\B(2j ) B(2j+2 )\B(2j−1 )
j=−∞ x∈B(2
∞ Z
X
≤ |f (x)|q sup W2 (z)q dx
j+2 )\B(2j−1 ) 1
j=−∞ B(2 4 |x|≤|z|≤4|x|
X∞ Z
. |f (x)|q W1 (x)q dx = 3.
j=−∞ B(2j+2 )\B(2j−1 )
X∞ Z
. inf W1 (x)q · |T [χB(2j+2 )\B(2j−1 ) f ](x)|q dx
x∈B(2j+1 )\B(2j ) B(2j+1 )\B(2j )
j=−∞
X∞ Z
. inf W1 (x)q · |f (x)|q dx
x∈B(2j+1 )\B(2j ) B(2j+2 )\B(2j−1 )
j=−∞
X∞ Z
= |f (x)|q inf W1 (z)q dx
B(2j+2 )\B(2j−1 ) z∈B(2j+1 )\B(2j )
j=−∞
≤ 3.
Hence we have
Z ∞ Z ∞ q Z ∞
F (s)ds dµ(s) . sq+n−1 F (2−1 s)q v1 (s)ds
0 r 0
7.2.4 Exercises
Exercise 101. Prove Theorem 318(2) using kf kL1 (0,t) ≤ kf kL1 (0,∞) for all
f ∈ M+ (Rn ).
Exercise 102. Using the disjointness of {EQ }Q∈S , prove that mS , given by
(7.8), is bounded on L∞ (Rn ).
Weighted Lebesgue spaces 339
7.3 Notes
Section 7.1
General remarks and textbooks in Section 7.1
The Ap condition initially appeared, in a somewhat different form, in
[369]. On the real line, Muckenhoupt characterized the characterization of
the weights for which the maximal operator M is bounded on Lp (w) if p > 1
and is weak Lp (w)-bounded if p ≥ 1 [324]. For singular integral operators,
Muckenhoupt, Rochberg and Wheenden obtained the same characterization
[205]. The same attempt was done for fractional integral operators in [326].
See standard textbooks [104, Chapter 7], [112], [146, Chapter IV], [147],
[156, Chapter 7], [231], [293, §1.4], [416, Chapter V], [382, §6.1.2] and [432] for
the class Ap . See [293, §3.4] for the class Ap,q .
See the textbooks [104, Chapter 7, §3], [146, Chapter IV, §5], [89] and [382,
§6.1.2.6] for the extrapolation. Especially we can find an exhaustive account
for the recent extrapolation theory in [89].
Section 7.1.1
Muckenhoupt initialized the theory of the A1 class [324]. Example 113 is
from [322, Remark 2.2] and [332, Example 2.3].
Theorem 280, which gives the information on the growth of the operator
norm, is from [206]. Theorem 284 is recorded in [206, Lemma 2.1].
Finally, Example 115 comes from [105, Corollary 4.4].
Section 7.1.2
Muckenhoupt initialized the theory of the Ap class [324]. Theorem 289 can
be found in [39, (2.6)]. See also the proof of [39, Theorem 2.5], where Buckley
applies the weak estimate to obtain the sharp strong estimate, which in turn
proves the sharpness of his weak estimate. See [39, 206, 324] for Theorem 290.
The proof of Theorem 290 in this book hinges on the idea of Lerner [268].
Example 120 is due to Lerner; see [274, Lemma 2.3].
Soria and Weiss showed that more general operators are bounded in [412].
This is a generalization of Theorem 288.
Section 7.1.3
Muckenhoupt initialized the theory of the A∞ class [324]. Theorem 298 is
due to Andersen and John [19, Theorem 3.1].
The reverse Hölder inequality for Ap , 1 < p < ∞ is proved by Coifman
and Fefferman in [82, Theorem IV].
The openness property for Ap , Theorem 297, is shown by Coifman and
Fefferman in [82, Lemma II] (the case of 1 < p < ∞) and [82, Theorem V]
(the case of p = ∞).
340 Morrey Spaces
We refer to [209, Theorem 2.3] for Theorem 295, the reverse Hölder inequal-
ity with precise constant.
Section 7.1.4
Muckenhoupt and Wheeden considered the boundedness of Iα and Mα
from Lp (v p ) to Lq (v q ). See [326] for Definition 74 and Theorem 300. Cruz-
Uribe and Perez considered the weak Lp (v p )-Lp (up ) boundedness of fractional
integral operators in [93, Theorem 1.1]. See Theorem 299.
Section 7.1.5
Curbera, Garcı́a-Cuerva, Martell and Pérez proved the extrapolation the-
orem, Theorem 302 [94].
Section 7.1.6
Coifman and Fefferman showed that kT f kLp (w) .[w]Ap kf kLp (w) for all f ∈
L∞ n
c (R ); see [82, Theorems I and III]. Later on, the estimate was improved.
The solution of this conjecture motivated the study of sharp dependence on
the Ap constants. Petermichl studied the Hilbert transform [355, 356]. Later
on, Hytönen and Lerner considered the decomposition of measurable functions
based on Fujii’s idea in [136]. See [86, 210, 262, 272, 273, 275] for more recent
advances for this field.
We followed [270] for the proof of Lemma 307. Lemma 307(1),(2) and (3)
correspond to [270, Proposition 3.1], [270, Lemma 3.2] and [270, Lemma 3.3].
Based on these steps, Lerner concluded his proof using Lemma 307(4) as we
did in the proof of Theorem 305. See [208] for more about this direction. See
[19, Theorem 5.2] for Theorem 311. See [262] for a more recent approach for
the A2 -theorem.
In Section 7.1.6, for 1 < p < ∞ we showed that w ∈ Ap is sufficient for
singular integral operators to be Lp (Rn )-bounded. See [82, Theorem II] for
the converse.
Section 7.2
General remarks and textbooks in Section 7.2
See the standard textbooks and lecture notes [112, Chapter 2], [146, Chap-
ter VI, §6], [147, Chapters 3, 4 and 9], [223, Chapter 2, §I], [231] and [235,
Chapters 4 and 6].
Section 7.2.1
Muckenhoupt considered the weighted inequality for Hardy operators; see
[325]. Theorems 312 and 314 are well known; see [35, 230, 309, 422, 437] for
Theorems 312 and 314, the case 1 < p ≤ q < ∞. Our formulation of Theorem
Weighted Lebesgue spaces 341
312 is exactly [112, Lemma 2.7.1]. See [47, Lemma 5] for Theorem 318(1), (2),
(3) and [45, Lemma 6] for Theorem 318(4).
Section 7.2.2
Kerman and Sawyer obtained the testing condition in [227, 385].
See [227, Theorem 2.3] and [400, Theorem 1] for the counterpart to
fractional integral operators of Theorem 319. Hunt, Kurtz and Neugebauer
pointed out that this theorem characterizes the Ap -class in [204].
Section 7.2.3
See [47, Theorem 3] for Theorem 320, where they also proved the necessity
of the conditions for the Hilbert transform to be bounded.
We list the definition of some related operators.
Example 129. For suitable f ∈ M+ (Rn ), the weighted Hardy operator Uψ
is defined by Z 1
Uψ f (x) ≡ f (sx)ψ(s)ds (x ∈ Rn ). (7.1)
0
Chapter 8 defines several closed subspaces, two of which coincide. This implies
that Morrey functions are not easy to approximate with nice functions.
Section 8.1 shows that closure in a metric space generally amounts to
measuring the size of the sets which can be approximated in the topology
in question. Section 8.2 gives closed subspaces in terms of closure for actual
approximations in Morrey spaces as well as canonical approximations.
343
344 Morrey Spaces
(2) Let U (Rn ) ⊂ L0 (Rn ) be a linear space with lattice property. For 0 <
Mp n
q (R )
q ≤ p < ∞, define U Mpq (Rn ) ≡ U (Rn ) ∩ Mpq (Rn ) .
The spaces of U we envisage are the following. Due to its importance we
state them as the definitions.
∗
Definition 77 ( Mpq (Rn ), Mpq (Rn ), M
fpq (Rn ), Mpq (Ω) ). Let 0 < q ≤ p < ∞.
(1) The case where U (Rn ) = L∞ (Rn ): The bar subspace Mpq (Rn ) stands for
the closure of L∞ (Rn ) ∩ Mpq (Rn ) with respect to Mpq (Rn ).
(2) The case where U (Rn ) = L0c (Rn ) ≡ {f ∈ L0 (Rn ) : supp(f ) is compact }:
∗
The star subspace Mpq (Rn ) stands for the closure in Mpq (Rn ) of L0c (Rn )∩
Mpq (Rn ).
(3) The case where U (Rn ) = L∞ n ∞ n 0 n
c (R ) = L (R ) ∩ Lc (R ): Denote by
Mq (R ), the tilde subspace, the Mq (R )-closure of Lc (R ) = L∞
f p n p n ∞ n n
c (R )∩
p n
Mq (R ).
(4) Let Ω ⊂ Rn be a measurable set. Define L0 (Ω) to be the subset of
L0 (Rn ) which consists of all measurable functions which vanish almost
everywhere outside Ω. Let U (Rn ) = L0 (Ω). Then we obtain the closed
subspace Mpq (Ω) of all f ∈ Mpq (Rn ) for which f vanishes almost every-
where outside Ω.
Needless to say, L∞ n p n
c (R ) is dense in L (R ) for 0 < p < ∞. Hence, the
above definition is significant only when 0 < q < p < ∞.
Example 131. Let 0 < q ≤ q̃ ≤ p < ∞. We may take U (Rn ) = Mpq̃ (Rn ).
Note that U Mpq (Rn ) differs from Mpq (Rn ) thanks to Theorem 20.
Recall that the Morrey space Mpq (Rn ) with 0 < q ≤ p < ∞ is defined as
the set of all f ∈ Lqloc (Rn ) for which kf kMpq ≡ sup m(f, p, q; r) < ∞, where
r>0
Z ! q1
1 1
m(f, p, q; r) ≡ sup |B(x, r)| p−q q
|f (y)| dy (r > 0).
x∈Rn B(x,r)
(3) The space V (∗) Mpq (Rn ) is defined as the set of all functions f ∈ Mpq (Rn )
Z !
q
such that lim sup |f (y)| χRn \B(N ) (y)dy = 0.
N →∞ x∈Rn B(x,1)
(∗)
(4) The space V0,∞ Mpq (Rn ) is defined as the intersection of these three
spaces.
(∗)
We will show that V0,∞ Mpq (Rn ) coincides with M
fpq (Rn ) in Theorem 326.
There are many closed subspaces. Here, we see that Lp (Rn ) is contained
in any of them.
Proposition 321. Let 0 < q ≤ p < ∞. Then
Proof For example, to check that Lp (Rn ) ⊂ V0 Mpq (Rn ), we argue as fol-
lows: Let f ∈ Lp (Rn ). Then m(f, p, q; r) ≤ m(f, p, p; r) by Hölder’s inequality.
Since L∞ n p n
c (R ) is dense in L (R ), we have
(1) The local bar subspace LMpq (Rn ) denotes the closure of L∞ (Rn ) ∩
LMpq (Rn ) in LMpq (Rn ).
∗
(2) The local star subspace LMpq (Rn ) stands for the closure in LMpq (Rn ) of
L0c (Rn ) ∩ LMpq (Rn ).
(1) Let E be the set defined in Example 11. Then χE ∈ LMpq (Rn ) \
∗
LMpq (Rn ).
∗
(2) χB(1) f ∈ LMpq (Rn ) \ LMpq (Rn ).
∗
(3) f ∈ Mpq (Rn ) \ (LMpq (Rn ) ∪ LMpq (Rn )).
Proposition 324. Let 0 < q ≤ p < ∞. Then Mpq (Rn ) ⊂ V0 Mpq (Rn ).
Proof Simply observe that L∞ (Rn ) ∩ Mpq (Rn ) ⊂ V0 Mpq (Rn ).
∗
We compare Mpq (Rn ) with V (∗) Mpq (Rn ).
∗
Proposition 325. Let 0 < q ≤ p < ∞. Then Mpq (Rn ) ⊂ V (∗) Mpq (Rn ).
Proof Simply observe that L0c (Rn ) ∩ Mpq (Rn ) ⊂ V (∗) Mpq (Rn ).
fpq (Rn ) = V (∗) Mpq (Rn ).
We will show that M 0,∞
(∗) (∗)
Proof Since L∞ n p n fp n
c (R ) ⊂ V0,∞ Mq (R ), we have Mq (R ) ⊂ V0,∞ Mq (R )
p n
p n
by taking the closure with respect to Mq (R ). To verify the converse inclusion,
(∗)
we let f ∈ V0,∞ Mpq (Rn ). Fix R1 , R2 , S > 0 so that R1 < R2 . Then
+ sup m(f, p, q; r)
x∈Rn \B(S),R1 ≤r≤R2
We note that
8.1.4 Exercises
n
Exercise 103. Let 0 < q < p < ∞, and let f (x) ≡ |x|− p , x ∈ Rn .
(1) Disprove that f = lim χB(k) f holds in Mpq (Rn ).
k→∞
takes place in Lp (Rn ) and hence in Mpq (Rn ) for any f ∈ Lp (Rn ). When we
consider the boundedness of operators, the following observation is useful.
Proof Since L∞ n fp n ∞ n
c (R ) is dense in Mq (R ) and Lc (R ) ⊂ L (R ) ⊂
p n
We may replace L∞ n ∞ n
c (R ) with Cc (R ).
Proposition 328. Let 0 < q ≤ p < ∞. Then, Cc∞ (Rn ) is densely contained
fpq (Rn ).
in M
Proof We have only to show that any function in L∞ n
c (R ) lies in the
∞ ∞ ∞
Mpq (Rn )-closureof Cc (R ). Let f ∈ Lc (R ). Then since Cc (Rn ) is dense
n n
We also have the following characterization of Mfp (Rn ), namely any func-
q
p n
tion in M
fq (R ) can be approximated by a linear combination of characteristic
functions over sets of finite measure. Denote by Simple(Rn ) the set of all such
functions.
fpq (Rn ).
Proposition 329. For 0 < q ≤ p < ∞, Simple(Rn ) is dense in M
Proof Let f ∈ M fp (Rn ), and let ε > 0. Then, from the definition of
q
Mq (R ) there exists g ∈ L∞
f p n n
c (R ) such that kf − gkMq < ε. Let E = supp(g).
p
∞
Choose a sequence {gk }k=1 of simple functions such that lim kg −gk kL∞ = 0.
k→∞
k
P k
P
Fix k ∈ N, and write gk = al χEl . Define fk ≡ gk χE = al χE∩El . Since
l=1 l=1
supp(fk ) ⊆ E, we have kg − fk kL∞ = k(g − gk )χE kL∞ ≤ kg − gk kL∞ . By using
the inequality
1
kg − fk kMpq ≤ kg − gk kL∞ kχE kMpq ≤ kg − gk kL∞ |E| p
We characterize the space M fpq (Rn ) for 0 < q ≤ p < ∞. We will use
∞
{Ek }k=1 to denote an arbitrary sequence of measurable subsets of Rn . We will
write Ek → ∅ a.e., if the characteristic functions χEk converge to 0 pointwise
a.e. Namely, Ek → ∅ a.e. if and only if lim χEk (x) = 0 for a.e. x ∈ Rn . Notice
k→∞
that the sets Ek are not required to have finite measure.
Definition 81 (Absolutely continuous norm). Let 0 < q ≤ p < ∞. A function
f in Mpq (Rn ) has absolutely continuous norm in Mpq (Rn ) if kf χEk kMpq → 0
for every sequence {Ek }∞ k=1 satisfying Ek → ∅ a.e., in the sense that
lim χEk (x) = 0 for almost all x ∈ Rn .
k→∞
Here are functions having or not having the absolutely continuous norm.
Example 133. Let 0 < q < p < ∞.
(1) The function f = χ[0,1]n has the absolutely continuous norm. In fact
simply notice that kχEj f kMpq ≤ kχEj f kLp for any decreasing sequence
{Ej }∞
j=1 of measurable sets such that χEj ↓ 0 almost everywhere.
n
(2) The function f ≡ | · |− p χB(1) fails to have the absolutely continuous
norm. This can be checked by using the sequence {B(j −1 )}∞ j=1 .
n
(3) The function f ≡ | · |− p χRn \B(1) , x ∈ Rn , does not have the absolutely
continuous norm. This can be checked by using the sequence {Rn \
B(j)}∞j=1 .
kχFj f kMpq ≤ kfk − f kMpq + kχFj fk kMpq ≤ kfk − f kMpq + kχFj fk kLp .
Approximations in Morrey spaces 351
Letting j → ∞, we obtain
Since k is arbitrary, we have lim kχFj f kMpq = 0. To prove the reverse inclu-
j→∞
sion, we assume that f ∈ Mpq (Rn ) has the absolutely continuous norm. From
the definition of the absolute continuity and the fact that
we have lim kf − χ{|f |≤R}∩B(R) f kMpq = lim kχ{|f |>R}∪(Rn \B(R)) f kMpq = 0.
R→∞ R→∞
fp (Rn ).
We thus conclude f ∈ Mq
Proof Assume that f ∈ Mpq (Rn ) and lim kχRn \B(R) f kMpq = 0. Define
R→∞
fR ≡ χB(R) f . Then fR ∈ L0c (Rn ) and that lim kf − fR kMpq = 0. This shows
R→∞
∗
that f ∈ Mpq (Rn ).
∗
Conversely, assume that f ∈ Mpq (Rn ). Given ε ∈ (0, 1), there exist gε ∈
Lc (Rn ) ∩ Mpq (Rn ) such that kf − gε kMpq < ε. Choose Rε > 0 such that
0
Consequently, for all R > Rε , kχRn \B(R) f kMpq ≤ kf − gε kMpq < ε. This shows
that lim kχRn \B(R) f kMpq = 0.
R→∞
The following proposition relates the truncation by the level set of func-
tions and Mpq (Rn ).
Proposition 332. Let 0 < q ≤ p < ∞. Then
n o
Mpq (Rn ) = f ∈ Mpq (Rn ) : f = lim χ[0,R] (|f |)f in Mpq (Rn ) .
R→∞
352 Morrey Spaces
Proof One inclusion is clear from the definition of Mpq (Rn ). Let us
concentrate on another inclusion. Let f ∈ Mpq (Rn ). Choose {fj }∞j=1 ⊆
Mpq (Rn ) ∩ L∞ (Rn ) such that lim kf − fj kMpq = 0. Then, there exists a
j→∞
subsequence {fjm }∞ ∞
m=1 ⊆ {fj }j=1 such that
gpq (Rn )
Proof Denote by A the right-hand side of (8.3). As before A ⊂ M
gpq (Rn ).
from the definition of M
Conversely, suppose that f ∈ Mgpq (Rn ). Then we can find a sequence
∞ ∞
{fj }j=1 ⊂ Lc (R ) such that fj → f in Mpq (Rn ). Fix j ∈ N. By the quasi-
n
Proof The “if” part is clear because et∆ f is smooth and satisfies
∂ [et∆ f ] ∈ Mpq (Rn ) for any multi-index α thanks to the property of con-
α
volution. Mimic the proof of Theorem 322 for the “only if” part. See Exercise
108 for more details.
The first part of the following corollary, whose proof we omit, is an analogy
to
Corollary 336. Let 1 ≤ q ≤ p < ∞.
(1) Let ψ ∈ Cc∞ (Rn ) and f ∈ Mpq (Rn ), then f ∗ ψ ∈ Mpq (Rn ).
(2) Let ψ ∈ Cc∞ (Rn ) ∩ M+ (Rn ) with kψkL1 = 1. Then f ∈ Mpq (Rn ) belongs
to Mpq (Rn ) if and only if ψj ∗ f → f in Mpq (Rn ) as j → ∞, where
ψj (x) ≡ j n ψ(jx) for x ∈ Rn and j ∈ N.
(3) If f ∈ Mpq (Rn ) and g ∈ L1 (Rn ), then f ∗ g ∈ Mpq (Rn ).
Proof
(1) Simply use ∂ α (f ∗ ψ) = f ∗ ∂ α ψ.
(2) Go through a similar argument to Theorem 335 using (1).
(3) Since kf ∗ gkMpq ≤ kf kMpq kgkL1 , we may assume that g ∈ Cc∞ (Rn ). In
this case, the assertion follows immediately from (1).
and hence
1 1
kf kMpq ≤ lim |B(r)| p − q kf − gkLq (B(r)) ≤ kf − gkMpq .
r→ ∞
354 Morrey Spaces
(3) f ∈
/M / Mpq (Rn ).
fpq (Rn ) since f ∈
n
/ V0 Mpq (Rn ) since f behaves like | · |− p near the origin.
(4) f ∈
(5) f ∈ V∞ Mpq (Rn ) since f is compactly supported.
(7) Since f ∈ C ∞ (Rn ) ∩ L∞ (Rn ) and satisfies ∂ α f ∈ Mpq (Rn ) for all multi-
indexes α, f ∈ Mpq (Rn ).
n
We also consider the truncation of | · |− p using the indicator function
χRn \B(1) .
n
Example 138. Let 0 < q < p < ∞. Define f (x) ≡ |x|− p χRn \B(1) (x) for
x ∈ Rn . Then f ∈ Mpq (Rn ).
n ∗
(1) Since f behaves like | · |− p away from the origin, f ∈
/ Mpq (Rn ).
(1) Using the fact that Gn is made up of infinitely many cubes of volume
fpq (Rn ), showing that
1, we will establish that χGn does not belong to M
∞
Mq (R ) is similar to L (R ) in some sense. In fact, let f ∈ L0c (R) ∩
p n n
|f − χGn | ≥ χ p p .
(k−1+(k!) p−q ,k+(k!) p−q )n
kχF (· + y) − χF kMpq
1 1
. sup kχF (· + y) − χF kLq (Q) + sup |Q| p − q kχF (· + y) − χF kLq (Q) .
Q∈Q Q∈Q
|Q|=1 |Q|≤1
8.2.4 Exercises
Exercise 105. See what happens if we use Example 15. Let 0 < q ! < p <
N
q q
. Also let f = ⊗n
P
∞, and let a ∈ − ,− χ(j−1,j−1+j a ) be a
p−q p j=1
function in Example 15.
∗
/ Mpq (Rn ).
(1) Show that f ∈
∗
/ Mϕ
(3) Use Proposition 331 to conclude that f ∈ n
q (R ).
Exercise 107. Let 0 < q < p < ∞, and let ψ ∈ Cc∞ (Rn ) be such that
−n
χB(1) ≤ ψ ≤ χB(2) . Let f (x) ≡ (1 − ψ(x))|x| p for x ∈ Rn .
∗ n
/ Mpq (Rn ): Hint: f behaves like | · |− p .
(1) Show that f ∈
(2) Show that f ∈ Mpq (Rn ): Hint: Check that f ∈ L∞ (Rn ) ∩ Mpq (Rn ).
(a) Find a small number T (k) > 0 such that for all t ∈ (0, T (k)),
fk − et∆ fk Mp
< 2−k (8.5)
q
holds.
(b) Show that f − et∆ f Mp
≤ 2kf − fk kMpq + fk − et∆ fk Mp
q q
(c) By combining this fact with (8.4) and (8.5), show that
f − et∆ f Mp
< 22−k ,
q
8.3 Notes
Section 8.1
General remarks and textbooks in Section 8.1
As we have seen, there are many non-trivial and important non-dense
subspaces. This aspect seems to have been untouched.
360 Morrey Spaces
Section 8.1.1
The definition of vanishing Morrey spaces at 0 goes back to the work of
Vitanza [444]; see Definition 78. We refer to [361] for commutators acting on
these spaces. See [177, Definition 1.1], [467, Definition 2.23] and [398, Defini-
tion 4.5] for bar subspaces, star subspaces and tilde subspaces, respectively.
Density of simple functions in M fp (Rn ) can be found in [65, Remark 3.2]; see
q
Proposition 329. See [65, Lemmas 3.3 and 6.4] for different characterizations
of Mfpq (Rn ). As an application of the theory of closed subspaces in Morrey
spaces, we can obtain some characterization of closed subpsaces; see [65, Lem-
mas 4.1, 4.2 and 6.2]. We refer to [467, Theorem 2.29] for more. The Morrey
space Mpq (Rn ) does not have Cc∞ (Rn ) as a dense closed subspace; see [438,
Proposition 2.16].
Vanishing Morrey spaces at 0 are studied from various aspects. See [14] and
[15] for fractional integral operators and Marcinkiewicz intergrals associated
with Schrödinger operators, respectively. Cao and Chen handled Toeplitz-type
operators in [61].
Section 8.1.2
Zorko explained that the set of smooth functions in Morrey spaces cannot
approximate Morrey spaces [472, p. 587]. Actually as in Theorem 322, she
pointed out in [472, Proposition 3] that the translation and the convolution
are closely related. We refer to [467, Definition 2.23] for Mpq (Rn ) and to [398,
Theorem 4.3] for M cp (Rn ). See [65, Lemma 2.1] as well as [180, Theorem 6]
q
for Theorem 322. The space L̃p,α (Rn ) in [34] is nothing but the diamond
subspace. See [33, 56] for more.
Section 8.1.3
We refer to [180, Lemma 3] for Proposition 324, where we showed
Mpq (Rn ) ⊂ V0 Mpq (Rn ).
Section 8.2
General remarks and textbooks in Section 8.2
We can find an exhaustive account for absolutely continuous norms in [29,
Chapter 3]. Adams pointed out that any element in diamond subspaces can
be approximated by the smooth functions in his textbook [6].
Section 8.2.1
See [21, Theorem 1.1] for the approximation of Morrey functions by means
of the Fourier-Jacobi expansions. Note that such an expansion goes back to
the original function in the norm topology of Morrey spaces if and only if
Approximations in Morrey spaces 361
the original function belongs to closed subspaces of Morrey spaces. See [398,
Theorem 4.6] for Theorem 330.
Section 8.2.2
Recall that we have charecterized Mpq (Rn ) in Theorem 332. See [65,
Lemma 3.1] for a different characterization of Mpq (Rn ).
Israfilov and Tozman considered the approximation by means of Morrey-
Smilnov classes [213].
Section 8.2.3
Example 140 can be found [181, Example 1.4]. See [277] for analytic Morrey
spaces.
Chapter 9
Predual of Morrey spaces
This chapter is devoted to considering the predual. The dual of Morrey spaces
contains an element which cannot be expressed in terms of locally integrable
functions. However, we show that Morrey spaces are realized as the dual of
Banach function spaces. Furthermore these Banach function spaces are again
realized as the dual of Banach function spaces. One of the reasons why such a
situation occurs is that Morrey spaces are wider than Lebesgue spaces. Hence,
there is difficulty in defining some linear operators.
Given a Banach space X , recall that the dual space X ∗ is made up of
all bounded linear functionals. For x∗ ∈ X ∗ , its dual norm is expressed by
kx∗ kX ∗ ≡ sup{|x∗ (x)| ∈ R : x ∈ X , kxkX = 1} by definition. The space X ∗
k · kX ∗ is a Banach space under the dual norm. Thus, the notation X ∗∗ always
makes sense. Given x ∈ X, we can define Q(x) ∈ X ∗∗ by Q(x)(x∗ ) ≡ x∗ (x) for
x∗ ∈ X ∗ . According to the Hahn–Banach theorem, Theorem 87, the canonical
mapping Q : x ∈ X 7→ Q(x) ∈ X ∗∗ is isometrc. If this isometry Q is surjective,
then we say that X is reflexive. See [310, 1.11.6] for example.
Chapter 9 is interested in predual spaces of Morrey spaces. In functional
analysis, duality plays a key role in justifying the convergence of sequences.
However, the dual space of the Morrey spaces Mpq (Rn ) with 1 < q < p < ∞
cannot be described via coupling. The dual space differs from the Köthe dual.
This is a big difference from Lebesgue spaces.
Section 9.1 introduces the predual of Morrey spaces. The predual of Morrey
spaces has some equivalent characterizations, which are investigated in Section
9.2.
363
364 Morrey Spaces
We start with defining the block spaces in Section 9.1.1 and then we
introduce predual spaces. Section 9.1.2 mainly considers subspaces in predual
spaces. Section 9.1.3 deals with duality. The Fatou property of the function
spaces is one of the important properties. As it turns out, among others it is
difficult to check the Fatou property of predual spaces. We will actually show
that predual spaces enjoy this property in Section 9.1.4. The Fatou prop-
erty of predual spaces has an application. We can specify the Köthe dual of
Morrey spaces in Section 9.1.5. We consider the theory of decomposition in
Section 9.1.6. In particular, as an application of the (pre)duality, we develop
a technique to control the Morrey norm in Section 9.1.6.
Of course we can start with the definition using balls. However, for the sake
of simplicity in later discussions, we use cubes. Note that this can be written
as follows: Z
kf kMpq ≡ sup sup |f (x)A(x)|dx,
(x,r)∈Rn+1
+
A Rn
0
where A moves over all Lq (Rn )-functions supported on a cube Q(x, r) and
1 1
satisfying kAkLq0 ≤ |Q(x, r)| p − q . Our observation justifies the following defi-
nition.
Definition 82 ((p, q)-block). Let 1 ≤ q ≤ p < ∞. A function A ∈ L0 (Rn ) is
a (p, q)-block if there exists a compact cube Q that supports A and
1
− p10
kAkLq0 ≤ |Q| q0 . (9.1)
If we need to specify Q, then we say that b is a (p, q)-block supported on Q.
Some prefer to call (p, q)-blocks (p0 , q 0 )-blocks. Here we will need to con-
sider the predual space, so we have to adopt the name “(p, q)-block”.
− p10
Example 142. Let Q be a cube, and let 1 ≤ q ≤ p < ∞. Then |Q| χQ is
a (p, q)-block.
Other examples of (p, q)-blocks, more precisely how to create (p, q)-blocks
in a standard manner, will be presented below.
By definition if Q supports A, then A(x) = 0 for almost every x ∈ Rn \ Q.
Let 1 ≤ q ≤ p < ∞. As is easily verified by Hölder’s inequality, any
0
(p, q)-block has Lp (Rn )-norm less than 1:
kAkLp0 ≤ 1 (9.2)
for all blocks A. The set of constant multiples of all blocks forms a linear
space. However, this set is unnatural due to the freedom of the choice of cubes
in the definition of (p, q)-blocks. We are thus interested in a natural way to
allow us to consider that the set of all (p, q)-blocks is a Banach space.
0
Definition 83 (Hqp0 (Rn )). Let 1 ≤ q ≤ p < ∞ and define the block space (the
0 0
Zorko space) Hqp0 (Rn ) as the set of all f ∈ Lp (Rn ) for which f is realized as
∞
X
the sum f = λj Aj with some {λj }∞ 1 ∞
j=1 ∈ ` (N) and a sequence {Aj }j=1 of
j=1
0
(p, q)-blocks. Define the norm kf kHp0 for f ∈ Hqp0 (Rn ) as
q0
where λ = {λj }∞
j=1 runs over all admissible expressions:
∞
X
f= λj Aj , {λj }∞ 1
j=1 ∈ ` (N), Aj is a (p, q)-block for all j ∈ N. (9.4)
j=1
366 Morrey Spaces
∞
P
A direct consequence of (9.2) is that the series f = λj Aj in (9.4)
j=1
0
∞
converges in the topology of Lp (Rn ). Thus, the postulate that f =
P
λj Aj
j=1
0
converges in Lp (Rn ) is natural as we will see.
0 0
Proposition 339. Let 1 < p < ∞. Then Hpp0 (Rn ) = Lp (Rn ) with coincidence
of norms.
0
Proof Let f ∈ Lp (Rn ). Fix ε > 0. Then there is a decomposition
∞
X
f = χQ1 f + χQj \Qj−1 f,
j=2
where {Qj }∞j=1 is an increasing sequence of cubes centered at the origin sati-
fying
ε
kχRn \Qk f kLp0 ≤ k
2
for all k ∈ N. Then using this decomposition, we learn kf kHp0 ≤ kf kLp0 +
p0
∞
X ε 0
k
. Thus, f ∈ Hpp0 (Rn ).
2
k=1
0 0
Meanwhile, if f ∈ Hpp0 (Rn ), then f ∈ Lp (Rn ) as we have seen.
0
We give an example of a function in Hqp0 (Rn ).
Example 143. Let 1 < q ≤ p < ∞. Define
1
f (x) ≡ (x ∈ Rn ).
1 + max(|x1 |, |x2 |, . . . , |xn |)M
0 0
Then f ∈ Hqp0 (Rn ) if and only if M p0 > n. Indeed, if f ∈ Hqp0 (Rn ), then
0
f ∈ Lp (Rn ), which forces M p0 > n. Suppose instead that M p0 > n. Define
1 −j n − n +jM −M
b0 ≡ n
χQ(1) · f, bj ≡ 2 p0 p0 χQ(2j )\Q(2j−1 ) · f (j ∈ N).
2
Then, kb0 kL∞ = 2−n and for each j ∈ N
−j pn0 − pn0 +jM −M
2 −j pn0 − pn0 − p10
kbj k L∞ = <2 ≤ |Q(2j )| .
1 + 2(j−1)M
Hence, each bj , j ∈ N0 is a (p, q)-block modulo a multiplicative constant by
Hölder’s inequality or by Example 142. Observe also that
∞
j pn0 + pn0 −jM +M
X
f = 2n b0 + 2 bj .
j=1
Predual of Morrey spaces 367
∞
j pn0 + pn0 −jM +M
Assuming that M p0 > n, we have 2n +
P
2 < ∞. Hence, it
j=1
0
follows that f ∈ Hqp0 (Rn ).
Below we collect some fundamental properties of block spaces.
Lemma 340. Let 1 < q ≤ p ≤ ∞. If A is a (p, q)-block, then kAkHp0 ≤ 1.
q0
1−1
|Q| p q
Proof Set B ≡ kAkLq0 A, assuming that A is not zero almost everywhere.
1 1
Then B is supported on the cube Q and, by virtue of the fact that p + p0 =
1 1
q + q 0 = 1,
1 1 − p10 + q10
kBkLq0 = |Q| p − q = |Q| .
Hence, B is a (p, q)-block. This implies kBkHp0 ≤ 1 according to Lemma 340.
q0
Equating this inequality, we obtain the desired result.
We have the lattice property of block spaces.
Lemma 342 (Lattice property). Let 1 < q ≤ p ≤ ∞. Then, a function f
0 0
belongs to Hqp0 (Rn ) if and only if there exists g ∈ Hqp0 (Rn ) ∩ M+ (Rn ) such that
|f (x)| ≤ g(x) for a.e. x ∈ Rn .
0
Proof Suppose that f ∈ Hqp0 (Rn ). Then there exists a sequence {λk }∞
k=1 ∈
∞ ∞
l1 and a (p, q)-block bk such that f =
P P
λk bk . Letting g ≡ |λk ||bk |,
k=1 k=1
0
we have g ∈ Hqp0 (Rn ) and |f | ≤ g. Conversely, suppose that there exists
0
g ∈ Hqp0 (Rn ) that satisfies |f (x)| ≤ g(x) for a.e. x ∈ Rn . Decompose g as
∞
λ0k b0k where {λ0k }∞ 1 0
P
g= k=1 ∈ ` (N) and bk is a (p, q)-block. Then we see that
k=1
∞
X 1 0
χ{y: g(y)6=0} (x) = λ0k b (x)
g(x) k
k=1
∞ |f (x)|
λ0k fg(x)
(x) 0
bk (x) for almost all x ∈ Rn . Since
P
and, hence, f (x) = ≤1
k=1 g(x)
for a.e. x ∈ Rn , the function (f /g)b0k becomes a (p, q)-block. This proves the
lemma.
368 Morrey Spaces
In the next lemma, we can say that each bk (Q) is supported on 3Q, while, in
Definition 83, each Aj has its support in a cube Qj but there is no information
on the cube Qj .
0
Lemma 343. Let 1 < q ≤ p ≤ ∞ and f ∈ Hqp0 (Rn ) with kf kHp0 < 1. Then
q0
0
λ(Q)b(Q) in the topology of Hqp0 (Rn ),
P
f can be decomposed as f =
Q∈D(Rn )
where {λ(Q)}Q∈D(Rn ) satisfies
X
λ(Q) ≤ 3n (9.5)
Q∈D(Rn )
and each bk is a (p, q)-block. We divide K into the disjoint sets K(Q) ⊂ N,
Q ∈ D(Rn ), as [
K= K(Q)
Q∈D(Rn )
and K(Q) fulfills supp(bk ) ⊂ 3Q and |Qk | ≥ |Q| when k ∈ K(Q). We do this
as follows: Let D = {Q(j) }∞ n
j=1 be an enumeration of D(R ). For each k ∈ K,
we define
jk ≡ min{j : supp(bk ) ⊂ 3Q(j) , |Qk | ≥ |Q(j) |}
for each k. We write K(Q(j) ) ≡ {k ∈ K : jk = j}. We set
1 P
X λk bk (λ(Q) 6= 0),
λ(Q) ≡ 3n
|λk |, b(Q) ≡ λ(Q) k∈K(Q)
k∈K(Q) 0 (λ(Q) = 0).
We now rewrite f as
X X X
f= λk bk = λk bk
k∈K Q∈D(Rn ) k∈K(Q)
−1
X X X X
n n
= 3 |λk | · 3 |λk | λk bk .
Q∈D(Rn ) k∈K(Q) k∈K(Q) k∈K(Q)
Predual of Morrey spaces 369
By (9.6), we have
X X X X
λ(Q) = 3n |λk | = 3n |λk | < 3n ,
Q∈D(Rn ) Q∈D(Rn ) k∈K(Q) k∈K
0
The block space Hqp0 (Rn ) is designed to study the duality of Mpq (Rn ). To
study a predual space of LMpq (Rn ) we can extend our definition naturally.
For example, the notion of local blocks can be defined in a similar fashion.
Definition 84. Let 1 ≤ p ≤ q < ∞. A function A ∈ L0 (Rn ) is a local (p, q)-
block if there exists a cube Q = Q(r) that is centered at the origin, supports
A and satisfies (9.1).
1 1
Example 144. Let 1 ≤ q ≤ p < ∞. For r > 0, |Q(r)| p − q χQ(r) is a local
(p, q)-block.
We now present the definition of local block spaces.
370 Morrey Spaces
0
Definition 85 (LHqp0 (Rn )). Let 1 ≤ p ≤ q < ∞ and define the local block
0 0
space LHqp0 (Rn ) as the set of all f ∈ Lp (Rn ) for which f is realized as the sum
∞ 0
λj Aj converges in Lp (Rn ) with some {λj }∞ 1
P
f= j=1 ∈ ` (N) and a sequence
j=1
0
p
{Aj }∞ n
j=1 of local (p, q)-blocks. Define the norm kf kLHp0 for f ∈ LHq 0 (R ) by
q0
where λ = {λj }∞
j=1 runs over all admissible expressions
∞
X
f= λj Aj , {λj }∞ 1
j=1 ∈ ` (N), Aj is a (p, q)-local block for all j ∈ N.
j=1
0 0
Proof Let f ∈ Lqc (Rn ) ⊂ Hqp0 (Rn ). We may assume that kf kHp0 < 1
q0
and that supp(f ) ⊂ Q0 for some dyadic cube Q0 , since we can decompose f
into 2n sums, each of which is supported on a closure of a quadrant. Then
according to Theorem 344 f can be decomposed as
X
f= λ(Q)b(Q)
Q∈D(Rn )
where X
λ(Q) ≤ 9n .
Q∈D(Rn )
We may assume that each b(Q) is non-negative and that λ(Q) is a non-negative
0
real number, since Hqp0 (Rn ) enjoys the lattice property. Furthermore, we may
assume that supp(b(Q)) ⊂ Q0 . Thus f is non-negative. By the monotone
convergence theorem and the fact that supp(f ) ⊂ Q0 , we have
X
f = lim χ(ε,ε−1 ) (|Q|)λ(Q)b(Q)χQ0
ε↓0
Q∈D(Rn )
0
in Lq (Rn ). Thus, if we set
X
g≡ (1 − χ(ε,ε−1 ) (|Q|))λ(Q)b(Q)χQ0 ,
Q∈D(Rn )
0
then since f ∈ Lq (Rn ), we deduce that g is a (p, q)-block as long as ε is small,
so that X
f =g+ χ(ε,ε−1 ) (|Q|)λ(Q)b(Q)χQ0
Q∈D(Rn )
and
Z
kgkHp0 = max f (x)g(x)dx : f ∈ Mpq (Rn ), kf kMpq = 1 . (9.8)
q0 Rn
for all g ∈ L (Qj ). By the uniqueness of this theorem we can find an Lqloc (Rn )
q
For a fixed cube Q and a fixed function f we set g ≡ χQ sgn(f )|f |q−1 .
1−1
|Q| p q
Notice that the function kgkLq0 g is a (p, q)-block by virtue of Lemma 341.
Thus, we can write
Z q1
1 1 1 1 1
I = |Q| p−q f (x)g(x)dx = |Q| p − q (L(g)) q . (9.10)
Q
1 1
Hence, we have |L(g)| ≤ kLk∗ |Q|− p + q kgkLq0 . As a result we have I ≤ kLk∗ .
This is the desired result. The proof of (9.7) and (9.8) is included in what we
have proven.
0
As the following example shows, `1 (Z) is embedded into Hqp0 (Rn ).
Predual of Morrey spaces 373
∞
P
constant, we have kf kHp0 . |aj |. To check the opposite inequality,
q0 j=−∞
n
we use kf · gkL1 ≤ kgk Mp
q
kf kHp0 , where g(x) ≡ |x|− p for x ∈ Rn . If we
q0
∞
P
use this inequality, then kf kHp0 & |aj |.
q0 j=−∞
In particular,
0 ≤ λk (Q) ≤ 2 · 9n (9.13)
Noticing (9.12), (9.13) and the weak-compactness of the Lebesgue space
0
Lq (Q) (see Corollary 338), we now apply a diagonalization argument or the
Tikhonov compact theorem and, hence, we can select an infinite subsequence
{fkj }∞ ∞
j=1 ⊂ {fk }k=1 that satisfies the following:
X
fkj = λkj (Q)bkj (Q), lim λkj (Q) = λ(Q),
j→∞
Q∈D(Rn )
0
lim bkj (Q) = b(Q) in the weak-topology of Lq (Q),
j→∞
P
where b(Q) is a (p, q)-block supported in Q. We set f0 ≡ λ(Q)b(Q).
Q∈D(Rn )
Then, by Fatou’s lemma and (9.11),
X X
λ(Q) ≤ lim inf λkj (Q) ≤ 2 · 9n , (9.14)
j→∞
Q∈D(Rn ) Q∈D(Rn )
0
which implies f0 ∈ Hqp0 (Rn ).
We will verify that
Z Z
lim fkj (x)dx = f0 (x)dx (9.15)
j→∞ Q0 Q0
for all Q0 ∈ D(Rn ). Once (9.15) is established, we will see that f = f0 and
0
f ∈ Hqp0 (Rn ) by virtue of the Lebesgue differentiation theorem because at
0
least we know that f0 locally in Lq (Rn ). By the definition of fkj and the fact
0
that the sum defining fkj converges in Lp (Rn ), we have
Z ∞
X X Z
fkj (x)dx = λkj (Q) bkj (Q)(x)dx.
Q0 l=−∞ Q∈D(Q0 )∪D ] (Q0 ), |Q|=2ln |Q0 | Q0
Note that
Z
1 1 1 1
bkj (Q)(x)dx ≤ kbkj (Q)kL1 ≤ kbkj (Q)kLq0 |Q| q ≤ |Q| p − q |Q ∩ Q0 | q
Q0
Predual of Morrey spaces 375
Since X
0
fpq (Rn ) is canonically identified with
(2) The dual of LM LHqp0 (Rn ).
for all f ∈ M+ (Rn ) such that ρ(f ) ≤ 1. Other two properties of (P1) are clear
from the corresponding properties of integrals. Consequently, (P1) is verified.
By the monotonicity of integral (P2) is also trivial. Let us check (P3). To
this end, we take a non-negative increasing sequence {gj }∞ j=1 ; 0 ≤ gj ↑ g. If
ρ(g) = 0, then there is nothing to prove. Hence, assume ρ0 (g) > 0. Choose any
non-negative number M ∈ [0, ρ0 (g)).
Then by the definition of ρ0 (g), we can find f ∈ M+ (Rn ) with ρ(f ) ≤ 1
such that Z
M< f (x)g(x)dx ≤ ρ0 (g).
Rn
By virtue of the monotone convergence theorem, if j is large enough
Z
M< f (x)gj (x)dx ≤ ρ0 (gj ) ≤ ρ0 (g).
Rn
Then we first see that |f (x)| < ∞, (a.e. x ∈ Rn ). Splitting f into its real
and imaginary parts and each of these into its positive and negative parts,
we may assume without loss of generality that f ∈ M+ (Rn ). For k = 1, 2, . . .,
0
set fk ≡ min(f, k)χQ(k) . We notice that fk ∈ Hqp0 (Rn ) and kfk kHp0 ≤ 1 by
q0
0
Lemma 342 and (9.19). Since fk ↑ f a.e., it follows that f ∈ Hqp0 (Rn ) and
kf kHp0 ≤ 1 by virtue of Theorem 348. This proves the theorem.
q0
380 Morrey Spaces
0
Theorems 352 and 355 yield Hqp0 0 (Rn ) = (Mpq )00 (Rn ) = Mpq (Rn ). Further-
more, from the fact that (Mpq )00 (Rn ) = Mpq (Rn ), we are able to characterize
the predual of block spaces.
Recall that we defined M fp (Rn ) in Definition 77. The block spaces are
q
fpq (Rn ).
realized as the dual of M
0
Theorem 356. Let 1 < q ≤ p < ∞. Then a predual space of Hqp0 (Rn ) is
fpq (Rn ) in the following sense:
M
0
If g ∈ Hqp0 (Rn ), then Lg : f 7→ Rn f (x)g(x)dx is an element of M fp (Rn )∗ .
R
q
0
Moreover, for any L ∈ M fp (Rn )∗ , there exists g ∈ Hp0 (Rn ) such that L = Lg .
q q
0
Proof The first assertion is clear, since Mpq (Rn ) is the dual of Hqp0 (Rn ).
0
Hence, we will prove that M fpq (Rn )∗ ⊂ Hp0 (Rn ). Thanks to Theorem 355, we
q
need only show M fpq (Rn )∗ .
fpq (Rn )∗ ⊂ (Mpq )0 (Rn ). Suppose that L belongs to M
p 0 n
We will exhibit a function g in (Mq ) (R ) such that
Z
L(f ) = f (x)g(x)dx (f ∈ M fpq (Rn )). (9.20)
Rn
λi (A) ≡ L(χA ), (A ∈ Ai ).
Notice that λi (A) is well defined for all A ∈ Ai because χA belongs to Mfp (Rn ).
q
n
We claim that λm is countably additive on Am for any m ∈ Z . Indeed,
let (Ak )∞
k=1 be a sequence of disjoint sets from Am , and let
l
[ ∞
[ ∞
[
Bl ≡ Ak , (l = 1, 2, . . .), A≡ Ak = Bl .
k=1 k=1 l=1
Rn
convergence theorem and the Fatou property of Morrey norm. This means
that g belongs to (Mpq )0 (Rn ). If g is complex-valued, then the same argument
applied separately to the real and imaginary parts of g shows that each of
these is in (Mpq )0 (Rn ). Hence, that g again belongs to (Mpq )0 (Rn ).
Write, for a function f which can be written as a finite linear combination
of characteristic functions of sets of finite measure,
Z
L(f ) = f (x)g(x)dx
Rn
and observe the continuity of both sides on Mpq (Rn ). Then we conclude from
Theorem 330 that (9.20) holds. This completes the proof of the theorem.
We conclude Section 9.1.5 with an example showing that the dual space
of the Morrey space Mpq (Rn ) with 1 ≤ q < p < ∞ does not coincide with its
Köthe dual. In particular, Mpq (Rn ) with 1 ≤ q < p < ∞ is not reflexive.
Example 148. Let 1 < q < p < ∞ and L : Mpq (Rn ) → C be a bounded
linear functional. Then in view of the embedding Lp (Rn ) ,→ Mpq (Rn ), one has
0
a function g ∈ Lp (Rn ) such that
Z
L(f ) = f (x)g(x)dx, (f ∈ Lp (Rn )).
Rn
382 Morrey Spaces
However, it can happen that L 6= 0 even when g ≡ 0; one can show this by an
example. Recall the set G considered in Example 12. Set
p p
Ik ≡ (k − 1 + k p−q , k + k p−q ).
for k ∈ N. Then,
∞
[ Z
G= Ik , lim χG (x)dx = 1.
k→∞ Ik
k=1
for all f ∈ Mpq (Rn ) and hence all f ∈ Lp (Rn ), then one obtains g ≡ 0 by
virtue of the Lebesgue dominated convergence theorem.
where A is sparse, aQ ∈ PL (Rn )⊥ and |aQ | ≤ χQ for all Q ∈ A and {λQ }Q∈A
satisfies
X
λQ χQ . kf kMpq .
Q∈A
Mp
q
For the proof of Theorem 357, we need to recall some elementary notions.
Let Q be a cube and let L ∈Z N. For all f ∈ Z L1loc (Rn ) there uniquely
exists PQL (f ) ∈ PL (Rn ) such that f (x)g(x)dx = PQL f (x)g(x)dx for all
Q Q
g ∈ PL (Rn ) since {xα }|α|≤k is independent in L2 (Q). The polynomial PQL f is
called the Gram–Schmidt polynomial of order L for Q. We have
Z
L
PQ(1) f (x) = R(x, y)f (y)dy
Q(1)
L −n L · − x0
for some polynomial R. The scaling law PQ(x0 ,r)
f = r PQ(1) f
r
holds (see Proposition 117).
Let A > 4n . We can find a set {Qkj }j∈Jk of dyadic cubes and a collection
{g} ∪ {bkj }j∈Jk of countable collections of L1 (Rn )-functions such that:
(1) (Partition of the set {M D f > Ak }) The family {Qkj }j∈Jk is disjoint and
Ωk ≡ { M D f > Ak } =
P k
Qj . Furthermore,
j∈Jk
(3) (The L∞ (Rn )-condition) The function g k is L∞ (Rn )-bounded: More pre-
cisely, kg k kL∞ ≤ 2n Ak .
(4) (The support condition) The function bkj is supported on the closure of
Qkj ; namely supp(bkj ) ⊂ Qkj .
Corollary 360. Let f ∈ L1loc (Rn ) be such that {M D f > λ} never contains
any quadrant for any λ > 0. Then f = lim (g K − g −K+1 ) in the sense of
K→∞
almost everywhere convergence.
Proof Let K ∈ N. Since f − g K is supported on {M D f > AK }, f =
lim g K . Likewise |g −K | ≤ A−K almost everywhere, f = lim (g K − g −K+1 ).
K→∞ K→∞
We set
akj ≡ χQkj (g k − g k+1 ) (k ∈ Z, j ∈ Jk ).
= 0.
Here for the third line we used the fact that Qkj contains Qk+1
j0 or that
Qkj ∩ Qk+1
j 0 = ∅ for any k ∈ Z, j ∈ J k and j 0
∈ Jk+1 .
We prove Theorem 357 after we translate it into the following form:
Predual of Morrey spaces 385
Theorem 362. Let 1 < q ≤ p < ∞. Then for any f ∈ Mpq (Rn ) there is a
decomposition
∞
X X
f= akj
k=−∞ j∈Jk
∞
X X
|akj | . Ak χQkj (k ∈ Z, j ∈ Jk ), Ak χQkj . kf kMpq
k=−∞ j∈Jk
Mp
q
Proof Note that M D is bounded on Mpq (Rn ), so that we are in the posi-
tion of using the previous observation. What is not contained in the previous
observations is the norm estimate. We do this as follows: First, observe that
Ak χΩk ≤ M D f . Consequently, since the Ωk ’s are nested,
∞
X X ∞
X
Ak χQkj = Ak χΩk
k=−∞ j∈Jk k=−∞ Mp
Mp
q q
. sup Ak χΩk
k∈Z Mp
q
D
≤ kM f kMpq
. kf kMpq .
Here for the last line we used the boundedness of M D on Mpq (Rn ).
There is another side in the theory of decomposition. As an application
of the duality, we obtain the averaging technique.
X As we saw, in some cases,
operators can be decomposed into the sum λQ (f )χFQ , where U is a family
Q∈U
of cubes and FQ is a measurable subset of Q. Hence, we are interested in the
general estimate of the function of the form
X
aQ ,
Q∈U
∞
X
kf kMpq .p,q,s,t λj χQj . (9.22)
j=1
Mp
q
∞
X
kf · gk L1 . λj χQj . (9.23)
j=1
Mp
q
∞
P
from f = λj aj . By the size condition of aj and g, we obtain
j=1
∞ ∞
1
1 1 1 − p10 1 1 1
X X
kf · gkL1 ≤ λj |Q| q − s |Qj | s |Q| q0 = λj |Q| p − s |Qj | s .
j=1 j=1
Note that
∞
X 1
λj χQj ≥ λj0 χQj0 Mp
= |Qj0 | p λj0
q
j=1
Mp
q
∞ ∞ ∞
X 1 1 1 1 X X
kf · gkL1 ≤ |Q| p − s |Qj | s − p λk χQk . λj χQj .
j=1 k=1 Mp
q
j=1
Mp
q
Predual of Morrey spaces 387
q0
If we let κ be the operator norm of the maximal operator M on L t0 (Rn ), then
0 0 1
we obtain κ−1/t χQ M [|g|t ] t0 is a (p, q)-block thanks to Lemma 341. Hence,
we obtain (9.23). This is the desired result.
When q = 1, we may assume t = 1 in Theorem 364.
Theorem 364. Suppose that the parameters p and s satisfy
1 < p < ∞, 1 < s < ∞, p < s.
Assume that {Qj }∞ n ∞ s n ∞
j=1 ⊂ Q(R ), {aj }j=1 ⊂ M1 (R ) and {λj }j=1 ⊂ [0, ∞)
1
fulfill the support condition supp(aj ) ⊂ Qj , the size condition kaj kMs1 ≤ |Qj | s
∞
X ∞
P
and the norm condition λj χQj < ∞. Then f = λj aj converges
j=1 j=1
Mp
1
in L1loc (Rn ) and satisfies
∞
X
kf kMp1 .p,s λj χQj .
j=1
Mp
1
9.1.7 Exercises
0 0
Exercise 109. Let 1 < q ≤ p < ∞. Then show that LHqp0 (Rn ) ⊂ Hqp0 (Rn ).
Exercise 110. [472, Proposition 4] Let 1 < q ≤ p < ∞. Then show that
0
Hqp0 (Rn ) is a Banach space by using (9.2) if necessary.
Exercise 111. [218, Example 2.5] Let 1 < q ≤ p < ∞ and a > 0. Define
1
f (x) ≡ (x ∈ Rn ).
a + |x|M
0
Then similar to Example 143, show that f ∈ Hqp0 (Rn ) if and only if M p0 > n.
Exercise 112. Let 1 ≤ p ≤ q < ∞. Let b1 , b2 be (p, q)-blocks supported on
cubes Q1 and Q2 , respectively. Then show that D−1 (b1 + b2 ) is a (p, q)-block
supported on a bigger cube Q as long as D > 0 is large enough.
Exercise 113. Let Q be a fixed cube, and let 1 < p ≤ q < ∞. Suppose
that we have a sequence {bk }∞k=1 of (p, q)-blocks supported on Q. Assume in
addition that {bk }∞
k=1 converges weakly to b. Then show that b is a (p, q)-block.
Exercise 114. Let 1 < q ≤ p < ∞, and let f ∈ Mpq (Rn ). For each j ∈ N
we set fj ≡ f χB(j) χ[0,j] (|f |). Then show that fj → f as j → ∞ in the
weak-* topology. Hint: Although we need to consider the pairing of f and any
0
g ∈ Hqp0 (Rn ), we can justify that g is a (p, q)-block.
Exercise 115. [95, Lemma 1], [282, Lemma 4.6] Let 0 < α < n and 1 <
p0
q ≤p< α n
. Suppose that we have a bounded sequence {fj }∞ n
j=1 in Hq 0 (R )
n ∞
such that Iα fj (x) converges to F (x) for almost all x ∈ R and that {fj }j=1
0
converges to f in the weak-∗ topology of Hqp0 (Rn ) via Theorem 356.
(1) Let gε (x) ≡ ε−α |x|α−n χB(ε) (x) and hε (x) ≡ ε−α |x|α−n χRn \B(ε) (x) for
x ∈ Rn and ε > 0. Use the weak-∗ convergence of {fj }∞ j=1 to show that
F (x) − Iα f (x) = limj→∞ εα gε ∗ (fj − f )(x) for almost all x ∈ Rn .
(2) Prove that sup kgε ∗ (fj − f )kHp0 < ∞.
j∈N q0
Exercise 117. Let E(µ) be a Banach lattice over a measure space (X, B, µ).
By reexamining the proof of Theorem 352, prove that E(µ) = E 00 (µ) if E(µ)
has the Fatou property.
Exercise 118. [158, Lemma 4.6] Let 0 < p < ∞ and α < max(1, p). Assume
that q ∈ (p, ∞] ∩ [1, ∞]. Suppose that we have a sequence {Qk }∞
k=1 of cubes
and a sequence {Fk }∞ q n + n
k=1 ⊂ L (R ) ∩ M (R ). Then show that
∞
X ∞
X
M (α) χQk · Fk . χQk M (q) Fk .
k=1 Lp k=1 Lp
where the infimum takes over all covers {B(xj , rj )}Nj=1 of E by at most count-
able families of open balls.
As is the case with the Lebesgue integral, it is difficult to calculate H d (E)
for general sets. Here is a special example of E for which we can find H d (E)
exactly.
d
Example 149. Let us show that H d (B(1)) = |B(1)| n for 0 < d ≤ n. It
d
follows from the definition that H d (B(1)) ≤ |B(1)| n . Thus, we need to show
d
H d (B(1)) ≥ |B(1)| n . Let {Bj }∞
j=1 be a ball covering of B(1). Then
nd
∞ ∞
d d
X X
|Bj | n ≥ |Bj | ≥ |B(1)| n .
j=1 j=1
where {Qj }∞
j=1 runs over all countable collections of dyadic cubes satisfying
∞
[
E ⊂ Int Qj . (9.1)
j=1
showing H̃01 ([0, 1) × {a}) ≥ 1. Likewise, we can show H̃01 ([0, 1) × {3, π}) = 1(<
2).
Based on a fundamental geometric observation, we see that H d and H̃0d
are not so different.
Proposition 365. Let 0 < d ≤ n. Then H̃0d ∼ H d .
Proof Let E ⊂ Rn . Let {Qj }∞ n
j=1 ⊂ D(R ) be a countable collection
!
∞
S
satisfying (9.1). We write Bj ≡ B(c(Qj ), 2n`(Qj )). Then E ⊂ Int Qj ⊂
j=1
∞ ∞
Bj . As a result, we see that H d . H̃0d . To end the proof, we may
S S
Qj ⊂
j=1 j=1
thus assume H d (E) < ∞.
When we have a ball B ≡ B(x, r), we can find 6n dyadic cubes
Q1 , Q2 , . . . , Q6n such that
n
6
1 `(Qj ) [
≤ ≤ 2, B(x, r) ⊂ Qj . (9.2)
2 r j=1
n
!
6
S
Thus, we see that B(x, r) = Int(B(x, r)) ⊂ Int Qj . With this in mind,
j=1
we suppose that M ∈ (H d (E), ∞). Then we have a covering {B(xj , rj )}∞ j=1
P∞ d
∞
S
such that M > |B(xj , rj )| n and that E ⊂ B(xj , rj ). Then we can find
j=1 j=1
dyadic cubes {Qj,k }j∈N,k∈N∩[1,6n ] such that
6n
!
[ 1 `(Qj,k )
B(xj , rj ) ⊂ Int Qj,k , ≤ ≤ 2, (j ∈ N).
2 rj
k=1
and that
∞ ∞ ∞
!
[ [ [
E1 ∪E2 ⊂ Int Q1,j ∪Int Q2,k = Int Q1,j ∪ Q2,k . (9.6)
j=1 k=1 j,k=1
Let
Then
∞
[ [ [
Q1,j ∪ Q2,k = Q1,j ∪ Q2,k (9.7)
j,k=1 (j,k)∈A2 ∪A3 ∪A4 (j,k)∈A1 ∪A4
and
∞
[ [ [
Q1,j ∩ Q2,k = Q1,j ∪ Q2,k . (9.8)
j,k=1 (j,k)∈A1 (j,k)∈A2 ∪A3
Thus,
d d
X X
H̃0d (E1 ∪ E2 ) ≤ |Q1,j | n + |Q2,k | n
(j,k)∈A2 ∪A3 ∪A4 (j,k)∈A1 ∪A4
Predual of Morrey spaces 393
and
d d
X X
H̃0d (E1 ∩ E2 ) ≤ |Q1,j | n + |Q2,k | n .
(j,k)∈A1 (j,k)∈A2 ∪A3
∞
[
K ε ≡ {x ∈ Rn : dist(K, x) < ε} ⊂ Int Qj . (9.10)
j=1
!
∞
S
Thus, there exists j0 ∈ N such that Kl ⊂ Int Qj for l ≥ j0 . As a result,
j=1
∞
X
lim H̃0d (Kj ) ≤ `(Qj )d . Since the sequence {Qj }∞
j=1 is arbitrary, we obtain
j→∞
j=1
(9.9).
Like the Lebesgue measure, essentially corresponding to the case where
d = n, the compactness of the sets is absolutely necessary as an example
similar to Example 150 shows.
Lemma 368 (Increasing property for open sets). Let 0 < d ≤ n, and let
{Oj }∞ n d
j=1 be a sequence of open sets in R expanding to O. Then lim H̃0 (Oj ) =
j→∞
!
∞
d
S
H̃0 Oj .
j=1
To this end, we may assume that sup H̃0d (Oj ) < ∞; otherwise (9.11) is trivial.
j∈N
Let ε > 0 be fixed. We choose {Qj,k }∞ n
k=1 ⊂ D(R ) so that
∞
!
[
Oj ⊂ Int Qj,k (9.12)
k=1
and
∞
X
`(Qj,k )d ≤ H̃0d (Oj ) + 2−j ε. (9.13)
j=1
Note that sup `(Qj,k )d ≤ ε + sup H̃0d (Oj ). This means that the size of Qj,k is
j,k∈N j∈N
bounded by a constant independent of j and k. Denote by {Qi }i∈I the disjoint
maximal cubes in {Qj,k }j,k∈N . !
∞ ∞ S
∞
S S S
Then O = Oj ⊂ Int Qj,k = Int Qi from (9.12). Let
j=1 j=1 k=1 i∈I
m ∈ N be fixed. We write
(1)
{Qi }i∈I ∩ {Q1,j }∞
j=1 = {Qi }i∈I1 . (9.14)
Define !
(1)
[
O1,m ≡ Om ∩ Int Qi (m ∈ N).
i∈I1
(1)
We set {Qm,k }k∈K1,m ≡ {Qm,k : k ∈ N, O1,m ∩ Qm,k 6= ∅}. Then
(1)
[
O1,m ⊂ Qm,k . (9.15)
k∈K1,m
(1) (1)
[ [
Qm,k ⊂ Qi . (9.16)
k∈K1,m i∈I1
Let
(1)
{Q1,k }∞ ∗
k=1 \ {Qi }i∈I1 = {Q1,k,m }k∈J1 . (9.17)
From (9.12) with j = 1, (9.14) and (9.17), we learn
! !
[ (1) [
∗
O1 = O1 ∩ Om ∩ Qi ∪ O1 ∩ Q1,k,m .
k∈I1 k∈J1
Predual of Morrey spaces 395
Hence
!! !
(1)
[ [
O1 = O1 ∩ Om ∩ Int Qi ∪ O1 ∩ Q∗1,k,m , (9.18)
k∈I1 k∈J1
since we are assuming that O1 is open. From (9.15) and (9.18), we have
(1)
[ [
O1 ⊂ Qm,k ∪ Q∗1,k,m . (9.19)
k∈K1,m k∈J1
As a result, we have
(1)
X X
H̃0d (O1 ) ≤ `(Qm,k )d + `(Q∗1,k,m )d . (9.20)
k∈K1,m k∈J1
(2)
For j = 2, we mimic the argument above to have a sequence {Qm,k }k∈K2,m
with properties similar to (9.19) and (9.22) above in addition to the property
(1) (2)
{Qm,k }k∈K1,m ∩ {Qm,k }k∈K2,m = ∅. (9.23)
More precisely, we set
(1) (2)
({Qi }i∈I \ {Qm,k }k∈K1,m ) ∩ {Q2,k }∞
k=1 = {Qi,m }i∈I2,m . (9.24)
Define
(2)
[
O2,m ≡ Om ∩ Int Qi,m .
i∈I2,m
(2)
We write {Qm,k }k∈K2,m ≡ {Qm,k : O2,m ∩ Qm,k 6= ∅}. As before, for any
(2) (2)
k ∈ K2,m , Qm,k intersects Qi,m for some i ∈ I2 . In view of (9.24), we have
(9.23). Rearrange the cubes to obtain
(2)
{Q2,k }∞ ∗
k=1 \ {Qi,m }i∈I2,m = {Q2,k }k∈J1 .
(2)
[ [
Going through an argument above, we have O2 ⊂ Qm,k ∪ Q∗2,k
k∈K2,m k∈J2
(2) (2)
`(Qi )d ≤ `(Qm,k )d + 41 ε, corre-
P P
corresponding to (9.19), and
i∈I2,m k∈K2,m
sponding to (9.22).
396 Morrey Spaces
(j)
Continuing this procedure, we can find a collection {Qm,k }k∈Ij,m for j =
(j )
1, 2, . . . , m and a sequence of subsets {Jj }m 1
j=1 of N such that {Qm,k }k∈Ij1 ,m ∩
(j2 ) (j)
[ [
{Qm,k }k∈Ij2 ,m = ∅ for any j1 < j2 ≤ m, that O2 ⊂ Qm,k ∪ Qj,k ,
k∈Kj,m k∈Jj
and that
(j) (j)
X X
`(Qi )d ≤ `(Qm,k )d + 2−j ε.
i∈Ij,m k∈Kj,m
m
(j)
`(Qi )d ≤ 2ε + H̃0d (Om ). Observe that, for any finite
P P
Notice that
j=1 i∈Ij,m
set I0 , there exists an integer M = M (I0 ) such that {Qi }i∈I0 P ⊂ {Qm,k :
m = 1, 2, . . . , M, k = 1, 2, . . .}. Thus, if we let m → ∞, we have `(Qi )d ≤
i∈I
lim H̃0d (Om ) + 2ε, as was to be shown.
m→∞
To this end, we may assume that sup H̃0d (Ej ) < ∞. We fix ε > 0. Then we can
j∈N
choose an open set Oj such that Ej ⊂ Oj and that H̃ d (Oj ) ≤ H̃0d (Ej ) + 2−j ε.
Since
3
H̃0d (O1 ∪ O2 ) + H̃0d (O1 ∩ O2 ) ≤ H̃0d (O1 ) + H̃0d (O1 ) ≤ H̃0d (E1 ) + H̃0d (E2 ) + ε,
4
Here, CN,d is a constant such that CN,d is finite if and only if d < N . Likewise,
d
we can check that kχB kL1 (H d ) ∼ |B| n for any ball B.
One considers the Choquet integral against H̃0d analogously to that against
H keeping in mind that H d and H̃0d are equivalent according to Proposition
d
365.
398 Morrey Spaces
As is the case with the Lebesgue integral, we are not interested in the value
of the integral itself. Let f : Rn 7→ [0, ∞] be any function, and let a > 0, We
mention that we readily obtain
Z Z
d
a · f (x)dH̃0 (x) = a f (x)dH̃0d (x) (9.26)
Rn Rn
by a change of variables.
The following formulas are easy to verify:
Proposition 370. For any increasing sequences {fk }∞
k=1 satisfying 0 ≤ fk ≤
fk+1 , Z Z
lim fk (x)dH̃0d (x) = lim fk (x)dH̃0d (x).
k→∞ Rn Rn k→∞
as was to be shown.
One of the important properties of the Choquet integral is the subadditiv-
ity given in the next theorem:
Predual of Morrey spaces 399
We prove Theorem 371 step by step. The next lemma, showing that addi-
tivity is available for some special cases, is a key to our observation:
Lemma 372. Let E1 ⊃ E2 ⊃ · · · ⊃ EN be a finite decreasing sequence of sets
N
X N
X −1
in Rn . Set f ≡ χEk and h ≡ χEk = f − χEN . Then
k=1 k=1
Z Z ∞
f (x)dH̃0d (x) = H̃0d (h−1 (λ, ∞])dλ + H̃0d (EN ).
Rn 0
as was to be shown.
Lemma 373. Inequality (9.27) remains true when f and g assume their value
in N0 and sup g ≤ 1.
We have
Z Z
f (x)dH̃0d (x) + g(x)dH̃0d (x)
Rn Rn
Z
= h(x)dH̃0d (x) + H̃ d (F1 ∪ EN ) + H̃0d (F1 ∩ EN )
Rn
from the induction assumption to the function h, which satisfies sup h 6= N −1.
Finally, using Lemma 372, we obtain
Z Z
f (x)dH̃0d (x) + g(x)dH̃0d (x)
Rn Rn
Z
≥ (h(x) + χF1 ∪EN (x) + χF1 ∩EN (x))dH̃0d (x)
Rn
Z
≥ (h(x) + χF1 (x) + χEN (x))dH̃0d (x)
n
ZR
≥ (f (x) + g(x))dH̃0d (x),
Rn
where E1 ⊃ E2 ⊃ · · · ⊃ EN and F1 ⊃ F2 ⊃ · · · ⊃ FN .
Predual of Morrey spaces 401
Example 153. Let E be a subset of Rn . Then kχE kL1 (H̃ d ) = H̃0d (E) since
0
as was to be shown.
If we choose d suitably, L1 (H̃0d ) can be considered as the predual of
0
Mp1 (Rn ).
Unlike the duality Hqp0 (Rn )–Mpq (Rn ), the characterization is some-
what indirect because we need to depend on a dense subspace of L1 (H̃0d ).
Predual of Morrey spaces 403
d
. kLk(L1 (H̃ d ))∗ r
0
q
Proof For a ball B, by letting w ≡ |B|−1+ p χB , we obtain
n o
kf kMpq . sup kf kLq (w) : w ∈ M+ (Rn ), kwkL1 (H̃ d ) ≤ 1
0
from the definition of the Morrey norm k·kMpq . To prove the converse relation,
Z
we fix w ∈ M+ (Rn ) with kwkL1 (H̃ d ) ≤ 1 and we set µ(E) ≡ |f (x)|q dx for
0
E
a Borel set E. Then µ is a measure. We observe
Z
|f (x)|q w(x)dx . kµkMp1 ∼ (kf kMpq )q
Rn
The space Vvu (Rn ) denotes the linear space of g ∈ L0 (Rn ) for which there
exists w ∈ L1 (H̃0d ) ∩ B+ (Rn ) which does not vanish almost everywhere on
{g 6= 0}.
Example 154. Let 1 < u < v < ∞. If g ∈ Lvc (Rn ), then g ∈ Vvu (Rn ). In fact,
2
we can take w(x) ≡ κe−|x| , x ∈ Rn with some suitable κ > 0 in Definition
93. If in addition g is supported on a ball B, then kgkVvu ≤ kχkχ B gkLv
Bk 1 d
.
L (H̃0 )
d
− nv
|B| 0
kχB gkLv .
We show that Vvu (Rn ) is a ball Banach function space.
Lemma 378. Let 1 < u ≤ v < ∞. Then Vvu (Rn ) is a ball Banach function
space.
Proof We content ourselves with the completeness of Vvu (Rn ): Let
∞
{gj }∞ u n
P
j=1 ⊂ Vv (R ) satisfy kgj kVvu < ∞. Write λj ≡ kgj kVvu for j ∈ N.
j=1
By the definition of λj there exists wj ∈ B+ (Rn ) such that
Z ! v1
|gj (x)|v wj (x)1−v dx . λj .
{wj 6=0}
∞
P
Let W ≡ λj wj .
j=1
By Hölder’s inequality, we have
q q−1
X∞ X∞ ∞
X
|gj | ≤ λj wj λj |gj |wj 1−q ,
j=1 j=1 j=1
so that q
∞
X ∞
X
|gj | W 1−q ≤ λj |gj |wj 1−q .
j=1 j=1
∞
gj = g in the topology of Vvu (Rn ).
P
Thus,
j=1
0
So far we showed that Vqp0 (Rn ) is a ball Banach function space. We will
0 0
establish that Vqp0 (Rn ) and Hqp0 (Rn ) are the same.
Theorem 379. Let 1 < q ≤ p < ∞. Then a predual space of Mpq (Rn ) is
0
Vqp0 (Rn ) with the equivalence of norms under the following pairing:
Z
0
hf, gi ≡ f (x)g(x)dx (f ∈ Mpq (Rn ), g ∈ Vqp0 (Rn )).
Rn
0
Proof Let g ∈ Vqp0 (Rn ) and f ∈ Mpq (Rn ). Then by Hölder’s inequality
Z q1 Z 10
q
q q0 1−q 0
kf · gkL1 ≤ |f (x)| w(x)dx |g(x)| w(x) dx
Rn Rn
for all Borel functions w that are allowed to vanish only on the set {g = 0}.
Therefore, it follows from Theorem 377 that
Z 10
q
q0 1−q 0
kf · gkL . kf kMq
1 p |g(x)| w(x) dx
Rn
for all such Borel functions w. Thus, from (9.32), we obtain the desired equal-
ity.
0
Conversely, suppose that we have a linear functional L on Vqp0 (Rn ). Then
0
for all balls B and a function g ∈ Lq (Rn ) with support B, we have
q0
− q10 1− p 1 1
|L(g)| ≤ kgkV p0 . |B| 0
kgkLq0 = |B| q − p kgkLq0 ,
q0
thanks to Example 154. Thus, we can mimic the proof of Theorem 347.
As a consequence, we obtain an equivalent expression of the Köthe dual
of Mpq (Rn ).
406 Morrey Spaces
0
Corollary 380. Let 1 < q ≤ p < ∞. Then the Köthe dual of Vqp0 (Rn ) is
0
Mpq (Rn ), or equivalently the Köthe dual of Mpq (Rn ) is Vqp0 (Rn ).
0
Thanks to Theorem 355 and Corollary 380, we see that two spaces Hqp0 (Rn )
0
and Vqp0 (Rn ) are the same.
0
Theorem 381. Let 1 < q ≤ p < ∞. Then the block space Hqp0 (Rn ) coincides
0
with Vqp0 (Rn ) with equivalence of norms.
Proof Simply combine Theorem 355 and Corollary 380.
9.2.4 Exercises
Exercise 119. By a change of variables, prove (9.26).
Exercise 120. For all nonempty sets E ⊂ Rn , show that H 0 (E) ≥ 1, and
that H 0 (E) = 1 if and only if E is bounded. Hint: When H 0 (E) = 1, then we
can find a ball that covers E.
Exercise 121. Let 0 < d ≤ n, and let E ⊂ Rn . Then show that H̃0d (E) =
inf{H̃0d (O) : E ⊂ O, O is open}.
9.3 Notes
Section 9.1
General remarks and textbooks in Section 9.1
Zorko [472], Adams [4] and Kalita [224] studied preduals of Morrey spaces.
Motivated by the work [221], Zorko considered the atomic space [472, p. 589]
unlike what we did in this book. Adams and Xiao investigated the relation
between them [8]. See [298] for the duality of central Herz–Morrey–Musielak–
Orlicz spaces of variable exponents.
See [457] for the duality in terms of the heat kernel.
Section 9.1.1
0
The definition of Hqp0 (Rn ) in Definition 83, originally denoted by H q,ϕ in
[472] in the framework of generalized Morrey spaces (see Section 12.1), is due
0
to Zorko. Examples of functions in Hqp0 (Rn ) can be found in [218]; see [218,
Example 2.4] for Example 143. Example 142, which can be used to calculate
the norm kχQ kHp0 , is due to Komori and Mizuhara [252, Lemma 1]. Although
q0
the position of the cubes in the definition of blocks is not specified, we can
Predual of Morrey spaces 407
use the dyadic grid D(Rn ) for this purpose; see [217, Lemma 2.4] and [397,
Lemma 2.2] for Theorem 344. Note that Theorem 348 goes back to a work
by Izumi, Sato and Yabuta [217, Lemma 2.5] as well as the one by Sawano
and Tanaka [398, Theorem 1.3]. The lattice property, Lemma 342, is [396,
Proposition 3.3].
Theorem 351 is [29, Theorem 2.2] when E(µ) is a Banach function space;
let ρ be a function norm. Then the associate norm ρ0 is itself a function norm.
Theorem 352 is [29, Theorem 2.7] when E(µ) is a Banach function space.
More precisely, every Banach function space E(µ) coincides with its second
associate space E 00 (µ). In other words, a function f belongs to E(µ) if and
only if it belongs to E 00 (µ), and in that case
Theorem 347 was investigated by Zorko [472]; see [224] as well. We refer to
[8] and [151] for more recent characterizations.
Section 9.1.2
0 0
We showed in Theorem 345 that Lqc (Rn ) is dense in Hqp0 (Rn ); see [341,
Lemma 2.6]. If the function is compactly supported, then we have an equiva-
0
lent expression of the norm in Hqp0 (Rn ); see [398, Proposition 5.3] for Theorem
346.
Section 9.1.3
Zorko characterized the predual in [472, Proposition 5]; see Theorem 347.
Section 9.1.4
Kantorovich and Akilov proved that the Fatou property of Banach lattices
is a key for the “bi”-Köthe dual to be back to the original space. [226]; see
Theorem 354.
Section 9.1.5
Köthe and Toeplitz began the study of certain pairs of subspaces of the
space of all real sequences [239, 240, 242]. Their theory has been generalized
by Dieudonné, Cooper and Lorentz and Wertheim [87, 103, 290]. Theorem 352
can be located as a counterpart to ball Banach function spaces of the result
by Lorentz and Luxemberg. We essentially followed the argument used in [29,
0
Theorem 4.1] to specify a predual space of Hqp0 (Rn ). Yang and Yuan showed
that the Morrey space Mpq (Rn ) is not reflexive. We followed [398, Example
5.2] for the proof. See [438, Corollary 2.20] for Theorem 356. The case where
q = 1 can be covered; see [303, Theorem 4.1].
408 Morrey Spaces
Section 9.1.6
We refer to [420, Chapter 8, Lemma 5] for the averaging technique for
Lebesgue spaces. We followed the averaging technique used in [212]; see The-
orem 363. See [388] for the applications to elliptic differential operators and
[180] for the applications to bilinear fractional integral operators. See [315] for
the case of Herz spaces.
See [255] for another type of decomposition of Morrey spaces.
Section 9.2
General remarks and textbooks in Section 9.2
See standard textbooks [6, Chapter 3], [112], [231], [382, §1.1.4] for the
Choquet integral. In particular, we can find Proposition 9.31 in [6, Theorem
5.1]. See [6, Theorem 5.2] for Lemma 375. In [6, Section 5.2] Adams showed
that the various predual spaces defined in this book coincide.
See [156, Appendix J], [415, Chapter I, §3] for the Whitney decomposi-
tion of the domain. Using the capacity, we can consider the notion of quasi-
continuity.
Definition 94 (H̃0d -quasi continuous). A function f is H̃0d -quasi continuous,
if for all ε > 0 there exists an open set such that H̃0d (O) < ε and that f is
continuous outside O.
Section 9.2.1
Yang and Yuan showed that H̃ is subadditive in [462]; see Theorem 366.
Section 9.2.2
The theory of the Choquet integral goes back to Choquet [78]. Orobitg
and Verdera applied the Choquet integral to the boundedness of the Hardy–
Littlewood maximal operators [349]. See [426, Theorem 1] for a generaliza-
tion to fractional maximal operators. Kuznetsov considered a vector-valued
inequality in [261]. Essoh, Fofana and Koua passed this result to Morrey spaces
[114]. Example 151 comes from [349, Lemma. 1]. See the lecture note [4] for
more. It may be interesting that the Hardy-Littlewood maximal operator is
L1 -bounded if the integral is considered in the sense of Choquet; see [4, The-
orem A].
Section 9.2.3
Adams gave a characterization of a predual space of Mp1 (Rn ); see [4, Propo-
sition 1]. The definition of the space Vvu (Rn ) is due to Adams and Xiao [8,
p.1632]; see Definition 93. Adams and Xiao proved Theorems 377, 379 and
381 in [8, Proposition 1 and Theorem 2.2], [8, Theorem 2.2] and [8, Theorem
3.3], respectively.
Chapter 10
Linear and sublinear operators in
Morrey spaces
One of the important problems in the theory of function spaces is the bound-
edness property of operators. This is important because it yields many appli-
cations in various fields of mathematics such as partial differential equations
and potential theory. There is a standard technique which is a local/global
strategy. However, unlike Lebesgue spaces, the operators need to be care-
fully defined in some cases. For example, singular integral operators, dealt
with in Section 10.4, must be handled with care because they are defined by
an approximation of integral kernels. Chapter 10 presents methods to define
operators in Morrey spaces. Likewise, care must be taken when handling com-
mutators in Section 10.5. The Hardy–Littlewood maximal operator (including
the sharp maximal operator) and fractional integral operators are discussed
in Sections 10.1 and 10.3, respectively.
Basically, the boundedness property of operators can be proved by the local
estimate initiated by Burenkov and Guliyev. However, to make the proof self-
contained we consider a direct proof which still recasts the flavor of the local
estimates.
409
410 Morrey Spaces
{fj }∞
j=1 Mp u ≡ k{fj }∞
j=1 k`u Mp
q (` ) q
Theorem 383. Suppose that the parameters p, q and u satisfy 1 < q ≤ p <
∞ and 1 < u ≤ ∞. Then {M fj }∞ ∞
j=1 Mp (`u ) . {fj }j=1 Mp (`u ) for every
q q
sequence of {fj }∞ 0 n
j=1 ⊂ L (R ).
∞ Z
X 1
k{M fj (y)}∞
j=1 k`u . k{fj (z)}∞
j=1 k`u dz (y ∈ Q).
|2k Q| 2k Q
k=1
412 Morrey Spaces
Consequently,
Z ∞ Z
1
−1
X 1 1
|Q| p k{M fj (y)}∞
j=1 k`u dy . |Q| p k{fj (z)}∞
j=1 k`u dz.
Q |2k Q| 2k Q
k=1
Theorem 387. Suppose that the parameters p, q and u satisfy 1 < q ≤ p < ∞
and 1 < u ≤ ∞. Then {M fj }∞ ∞
j=1 LMp (`u ) . {fj }j=1 LMp (`u ) for every
q q
sequence of {fj }∞ 0 n
j=1 ⊂ L (R ).
Proof We omit the proof; see Exercise 123.
We avoided defining LMpq (Rn ) for −∞ < p < 0 < q < ∞. One of the
reasons is that M fails to be bounded as the following example shows.
414 Morrey Spaces
Example 156. Let −∞ < p < 0 < q < ∞. Define the local Morrey norm
k · kLMpq formally by (1.1). Then since | · |θ χRn \B(1) ∈ LMpq (Rn ) whenever
0 < θ < − np , the maximal operator M does not map LMpq (Rn ) boundedly.
10.1.3 Exercises
Exercise 122. For a cube Q ∈ Q, we define Q1 (Q) ≡ {R ∈ Q : R ∩ Q 6=
∅, R∩(Rn \8Q) 6= ∅}. For a function f ∈ Mp1 (Rn ), define Λ ≡ sup mR (|f |).
R∈Q1 (Q)
Let us write Eλ ≡ {x ∈ Q : M f (x) > λ} for λ > Λ. Then for any x ∈ Eλ ,
show that we can find a cube Q0 such that Q0 contains x, that mQ (|f |) > λ
1
and that Q0 ⊂ 8Q. Prove also that |Q| p Λ . kf kMp1 .
Exercise 123.
(1) Reexamine the proof of Theorem 382 to prove Theorem 386.
(2) Prove Theorem 387 by mimicking the proof of Theorem 383.
M2],D ],D
−n−2 f (x) ≡ sup χQ (x)M2−n−2 ,Q f (x) (x ∈ Rn ).
Q∈Q
Theorem 388 (Sharp-maximal inequality for Morrey spaces). Let 0 < q̄ <
q ≤ p < ∞. Then kM2],D 0 n
−n−2 f kMq + kf kMq̄ ∼ kf kMq for all f ∈ L (R ).
p p p
Linear and sublinear operators in Morrey spaces 415
so that we obtain
1 1 1
|Q0 | p − q kf kLq (Q0 ) ≤ |Q0 | p |Med(f ; Q0 )|
∞ X
1 1 X
+ 2|Q | 0 p−q ω2−n−2 (f ; Qjk )χQj .
k
j=1 k∈Kj
Lq (Q0 )
Thus, we obtain
Z q1
1 1 1 1 1
0
|Q | p−q 0
kf kLq (Q0 ) . |Q | |Med(f ; Q )| + |Q | 0 0 p−q M2n,] q
−n−2 (x) dx .
p
Q0
We have to be careful since f does not always satisfy (4.2) as the example
of f = χF shows, where F is the set defined in Example 11.
Proof Although f fails (4.2), we will prove kf · gkL1 ≤ 6kM ] f · M D gkL1
for any (p, q)-blocks g arguing as in Theorem 165. Once this is done, we can
0
use the boundedness of M D on a predual space Hqp0 (Rn ), Theorem 385.
We may assume that g ∈ L∞ n
c (R ) by the truncation. We indicate the
change. We use the same idea assuming that f ∈ L∞ (Rn ) ∩ M+ (Rn ), so that
f ∈ Mpq (Rn ) by assumption. The major change from Theorem 165 is the
control of Ik . Note that
0 n
L (R ).
Proof The proof is omitted; see Exercise 124.
10.2.3 Exercises
Exercise 124. Reexamine the proof of Theorem 388, to prove Theorem 390.
Exercise 125. Let 1 < q ≤ p < ∞.
(1) Show that kf kLMpq .p kM ] f kLMpq for f ∈ L0 (Rn ) with min(M f, 1) ∈
LMpq (Rn ) by reexamining the proof of Theorem 389.
(2) Using (1) and the translation, show that kf kMpq .p kM ] f kMpq for f ∈
L0 (Rn ) with min(M f, 1) ∈ Mpq (Rn ). This will reprove Theorem 389.
Lebesgue spaces because Morrey spaces are equipped with two parameters. We
will consider the boundedness properties of Morrey spaces and local Morrey
spaces in Sections 10.3.1 and 10.3.2. One of the important differences between
the boundedness properties of Morrey spaces and local Morrey spaces is the
difference of the range of the parameter t for which the fractional integral
operator Iα is bounded from Mpq (Rn ) to Mst (Rn ) and Iα is bounded from
LMpq (Rn ) to LMst (Rn ).
n
(2) (Critical case) Assume that 1 ≤ q ≤ p = . Then
α
kI˜α f k∗ . kf kMpq (10.3)
for every f ∈ Mpq (Rn ), where k · k∗ denotes the BMO(R )-norm. n
n
(3) (Supercritical case) Assume that 0 < α− < 1 and that 1 ≤ q ≤ p < ∞.
p
Then
I˜α f α− n
. kf kMpq (10.4)
Lip p
and
mQ (|I˜α f2 − mQ (I˜α f2 )|) . kf k n . (10.6)
Mqα
By using the Lq (Rn )-Lw (Rn ) boundedness of the fractional integral operator
of Iα , which follows from (10.2), we obtain
1 n 1
mQ (|Iα f1 − mQ (Iα f1 )|) . |Q|− w kIα f1 kLw . |Q| α m2Q (|f |q ) q . kf k n .
Mqα
For the proof of the second inequality, we write the left-hand side out in full:
First, we bound the right-hand side with the triangle inequality. Then the
right-hand side is majorized by
ZZZ
1 1 1
2
|f (z)| · n−α
− dxdydz. (10.7)
|Q| Q×Q×(Rn \2Q) |x − z| |y − z|n−α
1 1 |x − y| `(Q)
n−α
− n−α
. n−α+1
. (10.8)
|x − z| |y − z| |z − c(Q)| |z − c(Q)|n−α+1
Linear and sublinear operators in Morrey spaces 419
for all x, y ∈ Q and z ∈ Rn \ 2Q. Thus inserting (10.8) into (10.7) gives us
|f (z)|
Z
˜ ˜
mQ (|Iα f2 − mQ (Iα f2 )|) . `(Q) dz
n |z − c(Q)| n−α+1
R \2Q
|f (z)|
Z
≤ `(Q) n−α+1
dz.
Rn \B(c(Q),`(Q)) |z − c(Q)|
|f (z)|
Z
n−α+1
dz.
Rn \B(c(Q),`(Q)) |z − c(Q)|
Z ∞ Z !
|f (z)|
= (n − α + 1) n−α+2
d` dz
0 B(c(Q),`)\B(c(Q),`(Q)) `
!
∞
|f (z)|
Z Z
≤ (n − α + 1) n−α+2
d` dz.
`(Q) B(c(Q),`) `
kf k αn
Z ∞ Z ! Z ∞
d` d` Mq
|f (z)|dz . kf k n ≤ .
`(Q) B(c(Q),`) `n−α+2 M1α `(Q) `2 `(Q)
Thus, the proof of (10.6) is complete. Hence the proof of (10.3) is concluded.
We will discuss certain inequalities for the fractional maximal operator,
which is defined by
rα
Z
Mα f (x) ≡ sup |f (y)|dy.
r>0 |Q(x, r)| Q(x,r)
n n
kg(λ·)kWMst = λ− s kgkWMst , kf (λ·)kMpq = λ− p kf kMpq ,
for λ > 0, and arithmetic shows Iα [f (λ·)] = λ−α Iα f (λ·) for all f, g ∈ M+ (Rn ).
Consequently, we obtain 1s = p1 − α
n . We may assume that q < p for the purpose
t q 1 1 1
of establishing = . Let R > 1 be the solution of (1 + R)− p = 2 q (1 + R)− q .
s p
n
Then for the set Ej in Example 9, we have kχEj kMpq ∼ (1 + R)−j p .
Let R be any one of the connected components of Ej . Then
Thus, Iα χEj (x) & (1+R)−jα χEj (x). Also, from the definition of Mst , we have
jnq
(1 + R)− pt = kχEj kLt ≤ kχEj kMst from the definition of the Morrey norm.
Consequently, it follows that
jnq jn
. (1 + R)jα kχEj kMpq ∼ (1 + R)j (α− p ) = (1 + R)−
n
(1 + R)− pt s .
qs
Since j ∈ N is arbitrary, it follows that t ≤ .
p
Example 157. We work in R. Suppose that the parameters p, s, t and α
satisfy
1 < p < ∞, 0 < α < 1
and
1 1 s
= − α, t= .
s p p
Let Ek ⊂ [0, 1] be the same set as Example 155. As before, choose r > 0 so
1
that r p = 2r. Then the fractional maximal operator Mα , given by
Z
1
Mα f (x) ≡ sup 1−α |f (x − y)|dy
R>0 R [−R,R]
Consequently,
k
X
t
(kMα χEk k ) &Lt r(k−l)tα (2r)tl (2r)k−l
l=1
k
X
≥ r(k−l)tα rtl/p r(k−l)/p
l=1
kt/p
= kr .
since
1 1 1 lt l 1 1 1 1
α+ = α + = , − ltα − = lt −α− = lt −α− = 0.
tp s p p p p pt p s
As before, this shows that Mα is not bounded from Mp1 (R) to Mst (R). Since
Iα |f | & Mα f , it follows that the Hardy–Littlewood-Sobolev inequality fails;
Iα is not bounded from Mp1 (R) to Mst (R).
One important observation is that we did not use the Hardy–Littlewood–
Sobolev theorem, which asserts Iα maps Lu (Rn ) to Lv (Rn ) boundedly as long
as 1 < u < v < ∞ and v1 = u1 − α p
n , to prove the boundedness from Mq (R ) to
n
s n p n
Mt (R ). Instead, we used the Mq (R )-boundedness of the Hardy–Littlewood
maximal operator.
To clarify the terminology which we use in the sequel, we recall some ter-
minologies. Estimate (10.2), the boundedness of fractional integral operators
on the (classical) Morrey spaces Mpq (Rn ) was studied by Spanne, Adams [2],
Chiarenza and Frasca [75] etc. Basically there are two results on the Morrey
boundedness of fractional integral operators. The first result is due to Spanne,
but was communicated by Peetre in [353, Theorem 5.4].
n
Theorem 394. Let 1 < q ≤ p < . Assume that the parameters s and u
α
satisfy
1 1 α 1 1 α
1 < u ≤ s < ∞, = − , = − . (10.10)
s p n u q n
Then
kIα f kMsu . kf kMpq (10.11)
for every f ∈ Mpq (Rn ) ∩ M+ (Rn ).
Before the proof, let us review Theorem 391 due to Adams. In 1975, as we
saw in Theorem 391, Adams [2] proved that
kIα f kMsu . kf kMpq (f ∈ Mpq (Rn )) (10.12)
as long as q satisfies (10.1).
Proof Arithmetic shows that u defined by (10.10) is less than q defined
by (10.1). Therefore, as was mentioned in [75, Corollary, p. 277], in view of the
embedding in Theorem 19 (together with the sharpness proven in Proposition
393) (10.12) is a better estimate than (10.11).
422 Morrey Spaces
10.3.3 Exercises
n
Exercise 126. Let 0 < α < n, 1 ≤ q ≤ p < α.
r0
(2) Using the Fatou property of Hr00 (Rn ), show that kIα f k p0 . kf k r0
Hp00 Hr00
r0
for all f ∈ Hr00 (Rn ).
424 Morrey Spaces
We use the following lemma to define the operator T on Mpq (Rn ). In many
methods we will present, the following estimate is fundamental:
Lemma 397. Let f ∈ L0 (Rn ). Then for any ball B ≡ B(a, r), we have
Z ∞
X
|K(x, y)f (y)| dy . m2k B (|f |) (10.1)
Rn \2B k=1
for all x ∈ B.
Inequality (10.1) corresponds to Lemma 130 for the maximal operator.
Proof By the size condition and the dyadic decomposition of Rn \ 2B, we
obtain
∞ Z
|f (y)| |f (y)|
Z Z X
|K(x, y)f (y)| dy . n
dy = n
dy.
Rn \2B Rn \2B |x − y| 2k+1 B\2k B |x − y|
k=1
Linear and sublinear operators in Morrey spaces 425
for functions f ∈ L1loc (Rn ) for which the left-hand side of (10.1) converges for
every ball B.
We need to check that (10.2) makes sense despite ambiguity of the choice
of the ball B containing x. We check this as follows:
Proposition 399. Let f ∈ L1loc (Rn ) be such that the left-hand side of (10.1)
converges for every ball B. If B1 and B2 are balls and x ∈ B1 ∩ B2 , then
Z
T [f χ2B1 ](x) + K(x, y)f (y)dy
Rn \2B1
Z
= T [f χ2B2 ](x) + K(x, y)f (y)dy.
Rn \2B2
426 Morrey Spaces
Therefore, for f ∈ L1loc (Rn ), T f (x) is well defined for all x ∈ Rn since the
right-hand side of (10.1) converges for every ball B. In this case we also write
Z
T [f χRn \2B ](x) = K(x, y)f (y)dy (x ∈ B).
Rn \2B
Z
kT f kMpq = T f (x) · g(x)dx.
Rn
for all f ∈ Lp (Rn ) ∩ Lq (Rn ). Then kT f kMpq . kf kMpq for all f ∈ Lp (Rn ). To
prove this, we have only to show that
1 1 1 1
|Q| p − q kT [χ3Q f ]χQ kLq + |Q| p − q kT [f − χ3Q f ]χQ kLq . kf kMpq .
The first estimate follows readily from (10.8). Meanwhile, for almost all x ∈ Q,
we deduce from (10.7)
|T f (x)|
∞ Z
X 1
. |Ω(x − y)| |f (y)|dy
|B(c(Q), 2l `(Q))| B(c(Q),2l `(Q))
l=1
∞
! q̃1 Z ! 10
Z q̃
X 1 q̃ q̃ 0
. |Ω(y)| dy |f (y)| dy
|B(2l+n `(Q))| B(2l `(Q)) B(c(Q),2l `(Q))
l=1
∞
! 10
Z q̃
X 1 q̃ 0
. |f (y)| dy
|B(c(Q), 2l+n `(Q))| B(c(Q),2l `(Q))
l=1
430 Morrey Spaces
|T [f + g](x)| ≤ |T f (x)| + |T g(x)| and |T [kf ](x)| = |k||T f (x)|. We also assume
that
|T f (x) − T g(x)| . |T (f − g)(x)| (10.9)
Assume that there exists a constant c0 > 0 such that, for all f ∈ L∞ n
c (R ),
(10.7) holds, where Ω ∈ L0 (Rn ) is homogeneous of degree zero and Ω ∈
Lp̃ (S n−1 ) for some p̃ ∈ [1, ∞]. Assume also that T is Lq (Rn )-bounded, in the
sense that kT f kLp ≤ Cp kf kLp for all f ∈ Lp (Rn ) ∩ Lq (Rn ). Then kT f kMpq .
kf kMpq for all f ∈ L1c (Rn ) = L1 (Rn ) ∩ L0c (Rn ).
Example 163. Let 1 ≤ q < p < ∞. Then the operator T f ≡ eikf kLq M f
does not satisfy (10.9) and cannot be extended to Morrey space Mpq (Rn ).
One important observation is that we used the Lq (Rn )-boundedness of
singular integral operators to prove the Mpq (Rn )-boundedness. So one of the
advantages of using Morrey spaces is that Morrey spaces can describe the
boundedness property more precisely than Lebesgue spaces.
Proof Simply mimic the proof of Proposition 327; see Exercise 131.
Lemma 405. Let 1 < q ≤ p < ∞, and let T be a singular integral operator.
Then kT f kLMpq . kf kLMpq for all f ∈ Lp (Rn ).
Proof Reexamine the proof of Theorem 401, where Lemma 397 is used;
see Exercise 132.
We can take the same strategy as the definition of singular integral oper-
ators acting on Morrey spaces. What is useful in Theorem 406 is that we can
define singular integral operators on the function space LMpq (Rn ) much larger
than Mpq (Rn ).
Theorem 406. Let 1 < q ≤ p < ∞. Then any singular integral operator T ,
defined initially on Lp (Rn ), extends to a bounded linear operator on LMpq (Rn ).
Proof Mimic Definition 97 as before. See Exercise 133.
Linear and sublinear operators in Morrey spaces 431
10.4.3 Exercises
Exercise 131. Mimic the proof of Proposition 327 to prove Proposition 404.
Exercise 132. Use Lemma 397 to prove Lemma 405.
Exercise 133. Prove Theorem 406 based on Definition 97.
Exercise 134. Let 1 < q ≤ p < ∞. Suppose that we have a linear operator
T : Lq (Rn ) → Lq (Rn ) satisfying kT kLq →Lq = sup{kT f kLq : kf kLq = 1} <
∞. Assume in addition that there exists K ∈ L0 (Rn × Rn ) such that
|K(x, y)| . |x − y|−n
and that Z
T f (x) = K(x, y)f (y)dy (a.e. x ∈
/ supp(f ))
Rn
for all f ∈ Lqc (Rn ) ≡ Lq (R ) ∩ L0c (Rn ). Then show that
n
kT f kMpq .p,q (1 + kT kLq →Lq )kf kMpq (f ∈ Mpq (Rn ) ∩ Lq (Rn )).
(2) K(x, tz) = t−n K(x, z) for every t > 0 and (x, z) ∈ Rn × S n−1 ;
Z
(3) K(x, z)dσ(z) = 0 for all x ∈ Rn ;
S n−1
∂αK
(4) max < ∞.
|α|≤2n ∂z α L∞ (Rn ×S n−1 )
Linear and sublinear operators in Morrey spaces 433
Then for any f ∈ Mpq (Rn ), there exist [a, T ]f ∈ Mpq (Rn ) such that
k[a, T ]f kMpq . kf kMpq .
Proof This is a general fact on singular integral operators together with
the expansion formula used in the proof of Theorem 236. In fact, since q > 1,
we see that the limit
10.5.3 Exercises
Exercise 135. Prove Proposition 410 by mimicking the proof of Theorem
400.
Exercise 136. Prove Theorem 401 by mimicking the proof of Theorem 408.
10.6 Notes
Section 10.1
See the survey [362].
Section 10.1.1
Chiarenza and Frasca showed that the Hardy–Littlewood maximal opera-
tor is weakly and strongly bounded on the Morrey space Mpq (Rn ) [75, Theorem
1(1)] and [75, Theorem 1(2)] when 1 ≤ q ≤ p and when 1 < q ≤ p < ∞, respec-
tively. See [75, Theorem 1] for Theorem 382 as well as the survey paper [391,
Theorem 3.1]. The proof hinges upon the Stein inequality for the Lebesgue
measure Corollary 272 and Proposition 271. Remark 15, concerning the weak-
boundedness of M acting on Morrey spaces, comes from [381].
Theorem 145 is from [394, Theorem 2.2] and [428, Lemma 2.5]. Example
155, asserting that M fails to be Mp1 (Rn )-bounded, comes from [328, Corollary
2.5]. The proof used here comes from [304, Lemma 4.6(i) and Lemma 4.7]. See
[194] for the action of Hardy operators on Morrey spaces.
See [202, Lemma 3.5] for the sharp maximal inequality as well as the
application to singular integral operators.
Di Fazio and Ragusa considered the boundedness of the sharp maximal
operators in [123]. See [247, Theorem 12], [283, Lemma 2.5] and [395, Theo-
rems 1.3 and 1.4] as well. See [383] for weak Morrey spaces.
Section 10.1.2
See [22] for the case of Carleson curves.
Linear and sublinear operators in Morrey spaces 435
Section 10.2
General remarks and textbooks in Section 10.2
It seems that there is no textbook which specilizes in sharp maximal
inequality on Morrey spaces.
Section 10.2.1
Di Fazio and Ragusa initially considered the sharp maximal inequality
for Morrey spaces in connection with elliptic differential equations [123]. See
[247, 333, 395, 440, 441] for more.
Section 10.2.2
See [169, Theorem 4.2] for the sharp maximal inequality for local Morrey
spaces.
Section 10.3
See the textbook [6, Chapter 7], where we can also find an application
of closed subspaces and preduals to the Morrey-boundedness of fractional
integral operators. See also the survey [362].
We refer to [6, Lemma 9.1] for the critical case of Theorem 391. In the
supercritical case, we considered the difference of the kernel and the constant
(average). We can consider the difference of the kernel and the Taylor polyno-
mial; see [282] together with its application to fractional Laplace equations.
Section 10.3.1
The boundedness of fractional integral operators on the (classical) Morrey
spaces Mpq (Rn ) was studied by Spanne, Adams [2], Chiarenza and Frasca [75]
etc. See [2] for the subcritical case of Theorem 391. See [3, Theorem 4.1 and
Remark 4.1] for the critical case of Theorem 391. See [10] for the supercritical
case of Theorem 391. We refer to [402, Theorem 1.1] for the case of bounded
domains, where Serrin gave an explicit bound for the operator norms. We
followed the survey paper [391, Theorem 3.3] for the proof of Theorem 391.
Olsen showed that we can no longer improve the Adams theorem for the
Morrey boundedness of fractional integral operators. See [342, Theorem 10]
for Proposition 393 with n = 3. His proof works for general dimensions. See
also the survey paper [391, Proposition 2.2].
436 Morrey Spaces
in [190, (3)]. Chiarenza and Frasca improved the method; in fact Chiarenza
and Frasca refined (10.1) to:
p p
Iα f . M f s (kf kMpq )1− s (f ∈ M+ (Rn ))
Section 10.3.2
Fu, Li and Lu [133] and Li and Yang [278] proved the boundedness of
fractional integral operators in Herz spaces. See also [306].
Section 10.4
General remarks and textbooks in Section 10.4
We refer to the textbook [6, Chapter 8], where Adams used Proposition 285
as well as Theorem 209. See the survey [264] for applications of Morrey spaces
to Navier–Stokes equations, where singular integral operators come into play.
The strategy taken to define singular integrals via duality is due to Rosenthal
and Schmeisser [370].
Section 10.4.1
Let us start to recall the results on singular integral operators acting on Mor-
rey spaces. Peetre considered the boundedness of singular integral operators in
[353]; see Theorem 401. There are several methods of the definition of singular
integral operators but the definition given here is close to that in [327, (3.1)] and
[198, §3]. There is another definition by means of the predual [396, p. 483]. We
refer to [438, Proposition 2.25] for the boundedness of singular integral opera-
tors. In particular, the boundedness of the Riesz transform is discussed in [438,
Corollary 2.27]; see [371, 438] for more recent information. The boundedness
of singular integrals is also proven there [75, Theorem 3], where Chiarenza and
Frasca employed the A1 -weight technique for the boundedness.
We move on to recall the results of singular integral operators acting on
related function spaces. Komori-Furuya investigated the boundedness of sin-
gular integral operators acting on block spaces in [245, Theorem]; see Theorem
400. See [372] for the concrete case of singular integral operators; they inves-
tigated the Cauchy integral in the predual of Morrey spaces.
Finally, we consider operators related to singular integral operators. Fu
and Lu calculated the operator norm in Morrey spaces of the operators given
by (7.1) and (7.2) [134, Theorem 1.3]. Fu and Lu also characterized a condition
under which Uψb is bounded in Morrey spaces for all b ∈ BMO in terms of ψ
[134, Theorem 1.6]. A similar result for Vψb is obtained in [134, Theorem 1.6].
A passage to the higher order commutator is done in [134, Theorem 3.1].
See [286, Theorem 1.3] for the boundendness of the Bochner–Riesz opera-
tor.
The Schrödinger operator generates singular integral operators; see [75].
Kurata and Sugano investigated such operators acting on Morrey spaces in
[259, 260]; see [285, Theorem 1] for its subsequent result and [285, Theorem
2] for a passage to commutators.
Chen and Jin obtained the boundedness of Marcinkiewicz integrals asso-
ciated with the Schrödinger operator on Morrey spaces [67].
438 Morrey Spaces
Section 10.4.2
Hu, Lu and Yang obtained the boundedness of singular integral operators
on Herz spaces [200]. Later on, Balakishiyev, Burenkov, Guliyev, Gurbuz,
Tararykova and Serbetchi considered singular integral operators in local Mor-
rey spaces [23, 48, 47]. Deringoz, Guliyev and Samko essentially considered the
boundedness of singular integral operators in generalized local Morrey spaces
[99, Theorem 5.5]. See also [132, 301] as well.
Section 10.5
See the survey [362]. Many authors investigated the boundedness property
of commutators on Morrey spaces; see [11, 97, 123, 395].
Section 10.5.1
Many authors investigated the boundedness property of commutators
given by (10.2); see [102, 252, 317, 318, 389, 395] for more details.
Di Fazio and Ragusa considered commutators generated by BMO functions
and singular integral operators. See [123, Theorem 2.1] for Theorems 408 and
409. Shirai considered commutators generated by BMO functions and the
Riesz potential [408]. See [469] for the application of Theorem 409 to parabolic
differential equations. Zhang, Jiang, Sheng and Zhou discussed the uniqueness
and existence of solution.
There are many generalizations. A passage to higher orders is done by Pen
and Liu in [351]. See [201, 286] for commutators generated by the Bochner–
Riesz operator and Lipschitz functions.
The compactness of commutators is also investigated. See [389] for the
sufficient conditions and [71, 72, 73, 74] for the necessary and sufficient con-
ditions. See [20, 341, 343, 430] for more generalization of [71, 72, 73, 74].
Concerning the commutator of the form [b, T ], based on the result by Janso
Linear and sublinear operators in Morrey spaces 439
and Paluszynski [219, 350], Shi and Lu considered the case when b is a Lips-
chitz function or a function in Morrey–Campanato spaces [406, 407]. We can
consider commutators generated by BMO and the maximal operators; see
[97, 456].
Section 10.5.2
Yu and Tao obtained the boundedness of commutators in local Morrey
spaces [468]. See [98, Theorem 3.1] for the boundedness of commutators in
Morrey-type spaces. Guliyev, Rahimova and Omarova considered the bound-
edness of commutators generated by the Littlewood–Paley functions and
BMO [171, Theorem 1–3]. Guliyev, Omarova, Ragusa and Scapellato worked
in the setting of generalized Morrey spaces. [169, Theorem 4.7]. The work
by Hasanov, Muradova and Guliyev [188] and the one by Eroglu, Hasanov
Omarova [113] considered the commutators in the multilinear and generalized
setting. See also [25, 96, 165, 166, 170, 403].
Bibliography
[1] D.R. Adams, Traces of potential arising from translation invariant oper-
ators, Ann. Scuola Norm. Sup. Pisa 25 (1971), 203–217. <80>
[2] D.R. Adams, A note on Riesz potentials, Duke Math. J., 42 (1975),
765–778. <205, xi, 81, 417, 421, 435, 436>
[3] D.R. Adams, Lectures on Lp -potential theory, Volume 2, Departement
of Mathematics, University of Umea, 1981. <240, 435>
[4] D.R. Adams, A note on Choquet integrals with respect to Hausdorff
capacity, in: Function spaces and applications (Lund, 1986), 115–124,
Lecture Notes in Math. 1302, Springer, Berlin, 1988. <217, 406, 408>
[5] D.R. Adams, My love affair with the Sobolev inequality. Sobolev spaces
in mathematics. I, 1–23, Int. Math. Ser. (N. Y.), 8, Springer, New York,
2009. <85>
[6] D.R. Adams, Morrey spaces, Lecture Notes in Applied and Numeri-
cal Harmonic Analysis. Birkhäuser/Springer, Cham, 2015. xv+121 pp.
<143, 288, 290, 324, 364, 365, 360, 408, 434, 435, 437, 438>
[7] D.R. Adams and L.I. Hedberg, Function spaces and potential theory,
Grundlehren der Mathematischen Wissenschaften [Fundamental Prin-
ciples of Mathematical Sciences], 314. Springer-Verlag, Berlin, 1996.
<289, 216, 217>
[8] D.R. Adams and J. Xiao, Nonlinear potential analysis on Morrey spaces
and their capacities, Indiana Univ. Math. J. 53 (2004), 1629–1663. <72,
80, 216, 406, 407, 408>
[9] D.R. Adams and J. Xiao, Morrey potentials and harmonic maps, Comm.
Math. Phys., 308, (2011), 439–456. <291>
[10] D.R. Adams and J. Xiao, Morrey spaces in harmonic analysis, Ark.
Mat., 50 (2012), 201–230. <212, 435>
[11] D.R. Adams and J. Xiao, Regularity of Morrey commutators, Trans.
Amer. Math. Soc., 364 (2012), 4801–4818. <438>
[12] R.A. Adams, Sobolev spaces, Pure and Applied Mathematics, 65, Aca-
demic Press, New York-London, 1975. <85>
441
442 Bibliography
[13] R.A. Adams and J.J.F. Fournier, Sobolev spaces. Volume 140 of Pure
and Applied Mathematics,V Elsevier, Academic Press, New York, 2nd
edition, 2003. <85>
[14] A. Akbulut, V.S. Guliyev, S. Celik and M.N. Omarova, Fractional inte-
gral associated with Schrödinger operator on vanishing generalized Mor-
rey spaces, J. Math. Inequal. 12 (2018), no. 3, 789–805. <360>
[15] A. Akbulut and O. Kuzu, Marcinkiewicz integrals with rough kernel
associated with Schrödinger operator on vanishing generalized Morrey
spaces, Azerb. J. Math. 4 (2014), no. 1, 40–54. <360>
[16] R.M. Alireza, An embedding theorem for Sobolev type functions with
gradients in a Lorentz space, Studia Math. 191 (2009), no. 1, 1–9.
<217>
[17] J. Alvarez, M. Guzmán-Partida and J. Lakey, Spaces of bounded λ-
central mean oscillation, Morrey spaces, and λ-central Carleson mea-
sures, Collect. Math., 51 (2000), 1–47. <82>
[18] J. Alvarez, The distribution function in the Morrey spaces, Proc. Amer.
Math. Soc. 83 (1981), no. 4, 693–699. <31>
[19] K.F. Andersen and R.T. John, Weighted inequalities for vecter-valued
maximal functions and singular integrals, Studia Math., 69 (1980), 19–
31. <339, 340>
[20] R. Arai and E. Nakai, An extension of the characterization of CMO and
its application to compact commutators on Morrey spaces, J. Math. Soc.
Japan 72 (2020), no. 2, 507–539. <438>
[21] A. Arenas and O. Ciaurri, Fourier-Jacobi expansions in Morrey spaces,
J. Math. Anal. Appl. 415 (2014), no. 1, 302–313. <360>
[22] H. Armutcu, A. Eroglu and F. Isayev, Characterizations for the frac-
tional maximal operators on Carleson curves in local generalized Morrey
spaces, Tbilisi Math. J. 13 (2020), no. 1, 23–38. <434>
[23] A.S. Balakishiyev, V.S. Guliyev, F. Gurbuz and A. Serbetci, Sublinear
operators with rough kernel generated by Calderón-Zygmund operators
and their commutators on generalized local Morrey spaces, J. Inequal.
Appl. 2015, 2015:61, 18 pp. <438>
[24] R.J. Bagby, An extended inequality for the maximal function, Proc.
Amer. Math. Soc. 48 (1975), 419–422. <80>
[25] A.S. Balakishiyev, E.A. Gadjieva, F. Gurbuz and A. Serbetci, Bounded-
ness of some sublinear operators and their commutators on generalized
local Morrey spaces, Complex Var. Elliptic Equ. 63 (2018), no. 11, 1620–
1641. <73, 439>
Bibliography 443
[67] D.X. Chen and F.T. Jin, The boundedness of Marcinkiewicz integrals
associated with Schrödinger operator on Morrey spaces, J. Funct. Spaces
2014, Art. ID 901267, 11 pp. <437>
[68] W.G. Chen and E. Sawyer, A note on commutators of fractional integrals
with RBMO(µ) functions, Illinois J. Math. 46 (2002), no. 4, 1287–1298.
<290>
[69] Y.M. Chen, S. Levine and M. Rao, Variable exponent, linear growth
functionals in image restoration, SIAM J. Appl. Math. 66 (2006), 1383–
1406. <84>
[70] Y. Chen and K. Lau, Some new classes of Hardy spaces, J. Funct. Anal.,
84 (1989), 255–278. <246>
[71] Y.P. Chen, Y. Ding and X. Wang, Compactness of commutators of Riesz
potential on Morrey spaces, Potential Anal. 30 (2009), no. 4, 301–313.
<438>
[72] Y.P. Chen, Y. Ding and X.X. Wang, Commutators of Littlewood-Paley
operators on the generalized Morrey space, J. Inequal. Appl. 2010, Art.
ID 961502, 20 pp. <438>
[73] Y.P. Chen, Y. Ding and X. Wang, Compactness for commutators of
Marcinkiewicz integrals in Morrey spaces, Taiwanese J. Math. 15 (2011),
no. 2, 633–658. <438>
[74] Y.P. Chen, Y. Ding and X. Wang, Compactness of commutators for
singular integrals on Morrey spaces, Canad. J. Math. 64 (2012), no. 2,
257–281. <438>
[75] F. Chiarenza and M. Frasca, Morrey spaces and Hardy–Littlewood max-
imal function, Rend. Mat., 7 (1987), 273–279. <141, xi, 421, 434, 435,
436, 437>
[76] F. Chiarenza, M. Frasca and P. Longo, Interior W 2,p estimates for non
divergence elliptic equations with discontinuous coefficients, Ricerche di
Matematica, 15, fasc. 1◦ , (1991), 149–168. <141, 127, 219, 245>
[77] F. Chiarenza, M. Frasca and P. Longo, W 2,p -solvability of the Dirich-
let problem for nondivergence elliptic equations with VMO coefficients,
Trans. Amer. Math. Soc. 336(2) (1993), 841–853. <83>
[78] G. Choquet, Theory of capacities. Ann. Inst. Fourier Grenoble 5, 1953–
1954, 131–295 (1955) <397, 408>
[79] A. Cianchi and L. Pick, Sobolev embeddings into spaces of Campanato,
Morrey, and Hölder type, J. Math. Anal. Appl. 282 (2003), no. 1, 128–
150. <218>
Bibliography 447
[120] Y. Fan and G. Gao, Some estimates of rough singular bilinear fractional
integral, J. Funct. Spaces. Appl. 2012, Article ID 406540, 17 pages.
<16, 83>
[121] X. Fan and D. Zhao, On the spaces Lp(x) (Ω) and W m,p(x) (Ω), J. Math.
Anal. Appl. 263 (2001), no. 2, 424–446. <84>
[122] G. Di Fazio, C.E. Gutierez and E. Lanconelli, Covering theorems,
inequalites on metric spaces and applications to PDE’s, Math. Ann.
341 2 (2008), 255–291. <289>
[123] G. Di Fazio and M.A.G. Ragusa, Commutators and Morrey spaces, Boll.
Un. Mat. Ital. A (7) 5 (1991), no. 3, 323–332. <xi, 434, 435, 438>
[124] C. Fefferman and E.M. Stein, Some maximal inequalities, Amer. J.
Math, 93 (1971), 107–115. <215>
[125] C. Fefferman and E. Stein, H p spaces of several variables, Acta Math.
129 (1972), 137–193. <215, 244>
[126] H. Feichtinger, An elementary approach to Wiener’s third Tauberian
theorem on Euclidean n-space, In: Proceedings of Conference at Cortona
1984, Symposia Mathematica, 29 (1987), 267–301, Academic Press, New
York. <35, 82>
[127] G.B. Folland, Introduction to Partial Differential Equations. Second
Edition. Princeton University Press, Princeton, NJ, 1995. xii+324 pp.
<127>
[128] J. Fournier, Mixed norms and rearrangements: Sobolev’s inequality and
Littlewood’s inequality, Ann. Math. Pura Appl. 148 (1987), 51–76.
<80>
[129] M. Frazier and B. Jawerth, Decomposition of Besov spaces. Indiana
Univ. Math. J. 34 (1985), no. 4, 777–799. <126>
[130] X. Fu, D. Yang and W. Yuan, Generalized fractional integrals and their
commutators over non-homogeneous metric measure spaces, Taiwanese
J. Math. 18 (2014), no. 2, 509–557. <290>
[131] Z.W. Fu, λ-central BMO estimates for commutators of N -dimensional
Hardy operators, J. Inequal. Pure Appl. Math. 9 (2008), no. 4, Article
111, 5 pp. <246>
[132] Z.W. Fu, S.Z. Lu, H.B. Wang and L.G. Wang, Singular integral opera-
tors with rough kernels on central Morrey spaces with variable exponent,
Ann. Acad. Sci. Fenn. Math. 44 (2019), no. 1, 505–522. <438>
[133] Z.W. Fu, Y. Lin and S.Z. Lu, λ-central BMO estimates for commutators
of singular integral operators with rough kernels, Acta Math. Sin. (Engl.
Ser.) 24 (2008), no. 3, 373–386. <437>
Bibliography 451
[134] Z.W. Fu and S.Z. Lu, Weighted Hardy operators and commutators on
Morrey spaces, Front. Math. China 5 (2010), no. 3, 531–539. <437>
[135] N. Fujii, Weighted bounded mean oscillation and singular integrals,
Math. Japon. 22 (1977/1978), 529–534. <324>
[136] N. Fujii, A proof of the Fefferman-Stein-Strömberg inequality for the
sharp maximal functions, Proc. Amer. Soc. 106 (1991), no.2, 371–377.
<216, 340>
[137] Y. Furusho, On inclusion property for certain L(p,λ) spaces of strong
type, Funkcial Ekvac. 23 (1980), no. 2,197–205. <273, 81>
[138] T. Futamura, Y. Mizuta and T. Shimomura, Sobolev embedding for
variable exponent Riesz potentials on metric spaces, Ann. Acad. Sci.
Fenn. Math. 31 (2006), 495–522. <290>
[139] S. Gala On regularity criteria for the three-dimensional micropolar fluid
equations in the critical Morrey-Campanato space, Nonlinear Anal. Real
World Appl. 12 (2011), no. 4, 2142–2150. <81>
[140] G. Gao, Boundedness for commutators of n-dimensional rough Hardy
operators on Morrey-Herz spaces, Comput. Math. Appl. 64 (2012), no.
4, 544–549. <83>
[141] J. Garcı́a-Cuerva, Hardy spaces and Beurling algebras, J. London Math.
Soc. (2), 39, 499–513 (1989) <246>
[142] J. Garcı́a-Cuerva and M.J. L. Herrero, A theory of Hardy spaces assosi-
ated to the Herz spaces, Proc. London Math. Soc. (1994), 69, 605–628.
<82>
[143] J. Garcı́a-Cuerva and J.M. Martell, Weighted inequalities and vector-
valued Carderón–Zygmund operators on nonhomogeneous spaces, Publ.
Mat. 44 (2000), no. 2, 613–640. <290>
[144] J. Garcı́a-Cuerva and J.M. Martell, On the existence of principal values
for the Cauchy integral on weighted Lebesgue spaces for non-doubling
measures, J. Fourier Anal. Appl. 7 (2001), no. 5, 469–487. <290>
[145] J. Garcı́a-Cuerva, G. Mauceri, P. Sjögren and J.L. Torrea, Higher-order
Riesz operators for the Ornstein-Uhlenbeck semigroup, Potential Anal.
10 (1999), 379–407. <291>
[146] J. Garcı́a-Cuerva and J.L. Rubio de Francia, Weighted norm inequalities
and related topics, North-Holland Math. Stud., 116 (1985). <xiii, 143,
214, 215, 218, 244, 339, 340>
[147] I. Genebashvili, A. Gogatishvili, V.M. Kokilashvili and M. Krbec,
Weight theory for integral transforms on spaces of homogeneous type,
Pitman Monographs and Surveys in Pure and Applied Mathematics,
92. Longman, Harlow, 1998. <289, 83, 84, 218, 244, 290, 291, 339, 340,
435>
452 Bibliography
[175] C.E. Gutiérrez, C. Segovia and J.L. Torrea, On higher order Riesz trans-
forms for Gaussian measures, J. Fourier Anal. Appl. 2 (1996), 583–596.
<291>
[176] De Guzman, Real Variable Methods in Fourier Analysis, North-Holland
Mathematical Studies 46, Notas de Matematica (75) (1981). <xiii, 143>
[177] D.I. Hakim, Complex interpolation of certain closed subspaces of Morrey
spaces, Tokyo J. Math. 41 (2018), no. 2, 487–514. <367, 360>
[178] D.I. Hakim and Y. Sawano, Interpolation of generalized Morrey spaces,
Rev. Mat. Complut. 29 (2016), no. 2, 295–340. <70, 72, 365, 366, 367,
81>
[179] D.I. Hakim and Y. Sawano, Calderón’s first and second complex inter-
polation of closed subspaces of Morrey spaces, J. Fourier Analysis and
Applications, 23 (2017), no. 5, 1195–1226. <363, 365, 367, xii>
[180] D.I. Hakim and Y. Sawano, Complex interpolation of vanishing Morrey
spaces, Annals of Functional Analysis, 11 (2020), 643–661. <16, 360,
408>
[181] D.I. Hakim and Y. Sawano, Complex interpolation of various subspaces
of Morrey spaces, Sci. China Math., 63 (2020), 937–964. <365, 361>
[182] G.H. Hardy and J.E. Littlewood, A maximal theorem with function-
theoretic applications, Acta Math. 54 (1930), no. 1, 81–116. <214>
[183] E. Harboure, J.L. Torrea and B. Viviani, Vector-valued extensions of
operators related to the Ornstein–Uhlenbeck semigroup, J. Anal. Math.
91 (2003), 1–29. <291>
[184] G.H. Hardy and J. Littlewood, Some properties of fractional integrals
I, Math. Z. 27 (1927), 565–606. <216>
[185] G.H. Hardy and J. Littlewood, Some properties of fractional integrals
II, Math. Z. 34 (1932), 403–439. <216>
[186] D.D. Haroske and L. Skrzypczak, Embeddings of weighted Morrey
spaces, Math. Nachr. 290 (2017), no. 7, 1066–1086. <210, 81>
[187] P. Harjulehto, P. Hästö, V. Latvala and O. Toivanen, Critical variable
exponent functionals in image restoration, Appl. Math. Lett. 26 (2013),
no. 1, 56–60. <84>
[188] A.A. Hasanov, Sh.A. Muradova and E.A. Gadjieva, Commutators of
multilinear singular integral operators on generalized local Morrey
spaces, Trans. Natl. Acad. Sci. Azerb. Ser. Phys.-Tech. Math. Sci. 35
(2015), no. 4, Mathematics, 61–74. <439>
[189] P. Halmos, Measure Theory, Springer Verlag, New York, (1974). <79>
Bibliography 455
[346] A. Ono, On isomorphism between certain strong L(p,λ) spaces and the
Lipschitz spaces and its applications, Funkcial Ekvac. 21 (1978), no. 3,
261–270. <81>
[347] A. Ono and Y. Furusho, On isomorphism and inclusion property for cer-
tain L(p,λ) spaces of strong type Ann. Mat. Pura Appl. (4) 114 (1977),
289–304. <81>
[348] W. Orlicz, Über konjugierte Exponentenfolgen. Studia Math. 3, 200–212
(1931) <84>
[349] J. Orobitg and J. Verdera, Choquet integrals, Hausdorff content and
the Hardy–Littlewood maximal operator, Bull. London Math. Soc. 30
(1998) 145–150. <408>
[350] M. Paluszynski, Characterization of the Besov spaces via the commuta-
tor operator of Coifman, Rochberg and Weiss, Indiana Univ. Math. J.
44 (1995), 1–17. <439>
[351] M. Peng and L.Z. Liu, Sharp estimates for multilinear integral operators,
Southeast Asian Bull. Math. 34 (2010), no. 1, 163–176. <438>
[352] J. Peetre, On convolution operators leaving Lp,λ spaces invariant, Ann.
Mat. Pura Appl. 72 (1966), 295–304. <218, 438>
[353] J. Peetre, On the theory of Lp,λ spaces, J. Funct. Anal. 4 (1969), 71–87.
<325, 365, 15, 80, 245, 421, 437>
[354] C. Pérez, Sharp Lp -weighted Sobolev inequalities, Ann. Inst. Fourier
(Grenoble) 45 (1995), 809–824. <218>
[355] S. Petermichl, The sharp bound for the Hilbert transform on weighted
Lebesgue spaces in terms of the classical Ap characteristic, Amer. J.
Math. 129 (2007), 1355–1375. <317, 340>
[356] S. Petermichl, The sharp weighted bound for the Riesz transforms, Proc.
Amer. Math. Soc. 136 (2008), 1237–1249. <317, 340>
[357] D. Pettis, On integration in vector spaces, Trans. Amer. Math. Soc. 44
(1938), no. 2, 277–304. <108>
[358] G. Pisier, Riesz transforms: a simpler analytic proof of P.A. Meyer’s
inequality, Lecture Notes in Math. 1321 (1988), 485–501. <291>
[359] H. Rafeiro, N. Samko and S. Samko, Morrey-Campanato spaces:
an overview. Operator theory, pseudo-differential equations, and
mathematical physics, 293–323, Oper. Theory Adv. Appl., 228,
Birkhauser/Springer Basel AG, Basel, 2013. <17, 245>
[360] M.A. Ragusa, Regularity for weak solutions to the Dirichlet problem
in Morrey space, Riv. Mat. Univ. Parma (5) 3 (1995), no. 2, 355–369.
<81>
Bibliography 467
[403] Y. Shen and L.Z. Liu, λ-central BMO estimates for multilinear com-
mutators of Littlewood-Paley operator on central Morrey spaces, Bull.
TICMI 15 (2011), 35–46. <439>
[404] S. Spanne, Some function spaces defined using the mean oscillation over
cubes, Ann. Scuola Norm. Sup. Pisa 19 (3) (1965), 593–608. <72, 244>
(p,Φ)
[405] S. Spanne, Sur l’interpolation entres les espaces Lk , Ann. Scuola
Norm. Sup. Pisa 20 (1966), 625–648. <365, 81, 218>
[406] S. Shi and S. Lu, Some characterizations of Campanato spaces via com-
mutators on Morrey spaces, Pacific J. Math. 264 (2013), no. 1, 221–234.
<439>
[407] S.G. Shi and S.Z. Lu, A characterization of Campanato space via com-
mutator of fractional integral, J. Math. Anal. Appl. 419 (2014), no. 1,
123–137. <439>
[408] S. Shirai, Necessary and sufficient conditions for boundedness of com-
mutators of fractional integral operators on classical Morrey spaces,
Hokkaido Math. J. 35 (2006), no. 3, 683–696. <438>
[409] B. Simon, Real analysis. With a 68 page companion booklet. A Com-
prehensive Course in Analysis, Part 1. American Mathematical Society,
Providence, RI, 2015. xx+789 pp. <93, 107>
[410] B. Simon, Harmonic analysis. A Comprehensive Course in Analysis, Part
3. American Mathematical Society, Providence, RI, 2015. xviii+759 pp.
<85, 127, 214, 215, 218>
[411] S. Sobolev, On a theorem in functional analysis (Russian), Math. Sob.
46 (1938), 471–490. <85, 216>
[412] F. Soria, and G. Weiss, A remark on singular integrals and power
weights, Indiana Univ. Math. J., 43 (1994), 187–204. <339>
[413] G. Stampaccia, L(p,λ) -spaces and interpolation, Comm. Pure Appl.
Math. 17 (1964), 293–306. <365, xii, 80>
[414] E.M. Stein, Intégrales singulières et fonctions différentiable de plusieurs
variables. Lecture Notes, Faculté des Sciences d’Orsay, 1967. <141, 218>
[415] E.M. Stein, Singular Integrals and Differentiability Properties of Func-
tions, Princeton University Press, Princeton, New Jersey, 1970. <141,
289, xiii, 85, 127, 143, 217, 218, 408>
[416] E.M. Stein, Harmonic Analysis, Real-variable Methods, Orthogonality,
and Oscillatory Integrals, Princeton University Press, Princeton, NJ,
1993. <289, xiii, 85, 143, 214, 215, 217, 218, 244, 339, 436>
Bibliography 471
[431] X.X. Tao and H.H. Zhang, On the boundedness of multilinear operators
on weighted Herz–Morrey spaces, Taiwanese J. Math. 15 (2011), no. 4,
1527–1543. <83>
[432] A. Torchinsky, Real variable methods in harmonic analysis. Academic
Press, 1986. <244, 339>
[433] Y. Terasawa, Outer measures and weak type (1,1) estimates of Hardy–
Littlewood maximal operators, Journal of Inequalities and Applications,
Volume 2006 (2006), Article ID 15063, 13 pages. <290, 291>
[434] X. Tolsa, Littlewood-Paley theory and the T (1) theorem with non-
doubling measures, Adv. Math. 164 (2001), no. 1, 57–116. <143, 291>
[435] X. Tolsa, BMO, H 1 , and Calderón–Zygmund operators for non doubling
measures, Math. Ann. 319 (2001), 89–149. <290, 291>
[436] X. Tolsa, Painlevé’s problem and the semiadditivity of analytic capacity,
Acta Math. 190 (2003), no. 1, 105–149. <291>
[437] G. Tomaselli, A class of inequalities, Boll. Un. Mat. Ital. (4) 2 (1969),
622–631. <340>
[438] H. Triebel, Hybrid function spaces, heat and Navier-Stokes equa-
tions, EMS Tracts in Mathematics, 24. European Mathematical Society
(EMS), Zürich, 2014. x+185 pp. <xii, 211, 80, 81, 360, 407, 437>
[439] S. Trudinger, On imbeddings into Orlicz spaces and some applications,
J. Math. Mech. 17 (1967) 473–483. <84, 217>
[440] Y. Tsutsui, Sharp maximal inequalities and bilinear estimates. Har-
monic analysis and nonlinear partial differential equations, 133–146,
Suurikaiseki Kôkyûroku Bessatsu B18, Res. Inst. Math. Sci. (RIMS),
Kyoto, 2010. <81, 435>
[441] Y. Tsutsui, Sharp maximal inequalities and its application to some bilin-
ear estimates. J. Fourier Anal. Appl. 17 (2011), no. 2, 265–289. <213,
435>
[442] N.K. Tumalun, D.I. Hakim, and H. Gunawan, Inclusion between gen-
eralized Stummel Classes and Other Function Spaces, Math. Inequal.
Appl. 23 (2020), no. 2, 547–562. <81>
[443] W. Urbina, On singular integrals with respect to the Gaussian measure,
Ann. Sc. Norm. Sup. Pisa Cl. Sci. 18 (1990), 531–567. <291>
[444] C. Vitanza, Functions with vanishing Morrey norm and elliptic par-
tial differential equations. In: Proceedings of Methods of Real Analysis
and Partial Differential Equations, Capri, pp. 147–150. Springer, 1990.
<360>
Bibliography 473
[445] L.W. Wang and L.S. Shu, Higher order commutators of Marcinkiewicz
integral operator on Herz–Morrey spaces with variable exponent, J.
Math. Res. Appl. 34 (2014), no. 2, 175–186. <83>
[446] G.V. Welland, Weighted norm inequalities for fractional integrals, Proc.
Amer. Math. Soc. 51 (1975), 143–148. <216>
[447] N. Wiener, Generalized harmonic analysis, Acta Math., 55, 117–258
(1930) <82>
[448] N. Wiener, Tauberian theorems, Ann. Math., 33 (1932), 1–100 <82>
[449] N. Wiener, The ergodic theorem. Duke Math. J. 5 (1939), 1–18 <214>
[450] Weighted inequalities for the dyadic square function without dyadic A∞ ,
Duke Math. J. 55 (1987), 19–50. <324>
[451] J.L. Wu, Boundedness of commutators on homogeneous Morrey-Herz
spaces over locally compact Vilenkin groups, Anal. Theory Appl. 25
(2009), no. 3, 283–296. <83>
[452] J.L. Wu, Boundedness for commutators of fractional integrals on Herz–
Morrey spaces with variable exponent, Kyoto J. Math. 54 (2014), no.
3, 483–495. <83>
[453] J.L. Wu, Boundedness of some sublinear operators on Herz–Morrey
spaces with variable exponent, Georgian Math. J. 21 (2014), no. 1, 101–
111. <83>
[454] J.L. Wu, Boundedness for fractional Hardy-type operator on Herz–
Morrey spaces with variable exponent, Bull. Korean Math. Soc. 51
(2014), no. 2, 423–435. <83>
[455] J.L. Wu and Q.G. Liu, Boundedness of some singular integral operators
on Morrey-Herz spaces over locally compact Vilenkin groups, Commun.
Math. Res. 27 (2011), no. 2, 147–155. <83>
[456] C.P. Xie, Some estimates of commutators, Real Anal. Exchange 36
(2010/11), no. 2, 405–415. <439>
[457] M. Xu, The duality between Morrey spaces and its predual spaces char-
acterized by heat kernel, Anal. Theory Appl. 25 (2009), no. 1, 71–78.
<406>
[458] D. Yang, G. Hu and S. Lu, Herz type spaces and their applications,
Mathematics Monograph Series 10. <82, 291>
[459] Da. Yang, Do. Yang and G. Hu, The Hardy space H 1 with non-doubling
measures and their applications, Lecture Notes in Mathematics, 2084,
Springer, Cham, 2013. <290>
474 Bibliography
(µ ⊗ ν)∗ , 4 M D , 149
D(µ;Q)
Aαp,q , 311 Mlog , 262
A∞ , 306 MΦ , 147
Ap , 299 MΩ,α , 216
A1 , 296 Pk (·; x0 , ρ, f ), 239
B(r), xvii Q(j), 229
B(x, r), xvii Q(x, r), xvii
Bp , 150 S n−1 , 122, 196
Bσ (E)(Rn ), 33 U Mpq (Rn ), 343
Ej f , 200 Vvu (Rn ), 403
Gs , 188 (∗)
V0,∞ Mpq (Rn ), 344
IΩ,α , 216
α V (∗) Mpq (Rn ), 344
Kpq (Rn ), 34
0 V0 Mpq (Rn ), 344
L (µ), 2
V∞ Mpq (Rn ), 344
L0 (Rn ), 2
X(ρ), 28
LΦ (µ), 58
[a, T ], 228, 229
Lp (`q , Rn ), 144
[w]A∞ , 306
Lp (µ), 5
[w]Ap , 299
Lp (|t|α ), 294
[w]A1 , 296
Lp (|x|α ), 294
∆2 , 56
Lp (rα ), 294
Φλ,q (0, ∞), 42
Lp (tα ), 294
χ, xviii
Lp (w, `q ), 311
Ḃσ (E)(Rn ), 33
Lploc (Rn ), 7 α
K̇pq , 34
L1 (H̃0d ), 402 ` (Lp (w)), 311
q
Lp,∞ (Rn ), 135 `q (Lp , Rn ), 144
Lp,q (µ), 42 η-function, 133
M , 130 ηj,m , 133
],D
Mλ;Q 0
, 163 λf , 5
k
M , 298 ∇2 , 56
D(Q)
Mα , 262 Mpq (Rn ), 344
D(Q) ∗
Mlog , 262
Mpq (Rn ), 344
D(µ)
Mα , 264
D(µ) Mpq (Rn ), 346
Mlog , 264
D(µ;Q)
ρ0 , 376
Mα , 262 T̃ , 205
475
476 Index
ϕ∗ , 52 LMpq (Rn ), 32
fpq (Rn ), 344
M Med(f ; Q), 164
aα (x0 ; ρ, f ), 239 WLp (µ), 8
f ⊗ g, 22 WL∞ (µ), 8
f ∗ , 39, 164 WLp (Rn ), 135
w ∈ B, 9 WLMpq , 32
w(E), 294 p.v., 197
Mp1 (Rn ), 16 5r-covering lemma, 135
M↓ (I), 38
M↑ (0, a), 50 A∞ -characteristic, 306
M↑ (I), 38 A∞ -constant, 306
Rn+1
+ , 13
A∞ -weight, 306
C γ (Rn ), 78 A1 -characteristic, 296
D(Q), 136, 151 A1 -constant, 296
D(Q0 ), 163 A1 -weight, 296
D(Rn ), 14 Ap -characteristic, 299
D1 (Q), 136 Ap -constant, 299
0
Hqp0 (Rn ), 365 Ap -theorem, 318
Ap -weight, 299
Lλ,q n
k (R ), 238 absolutely continuous norm, 350
p
Mq (Ω), 344 Adams’ theorem, 417
Mpq (Rn ), 13 admissible, 272
P(Rn ), 109 associate norm, 376
PL⊥ (Rn ), 111
Pk (Rn ), 109 ball Banach function norm, 29
Q(Q), xvii ball Banach function space, 29
Q] (Q), xvii Banach function norm, 27
X 0 (ρ), 377 Banach function space, 28
X (ρ0 ), 377 Bessel kernel, 188
X0 ∩ X1 , 89 best approximation, 118, 239
X0 + X1 , 89 Beurling algebra, 82
D+ , 168 [a, T ], 432, 433
D0 , 168
Da , 168 Calderón–Zygmund decomposition,
BCα (Rn ), 78 202
BMO(Rn ), 222 Calderon–Zygmund operator, 195
BMO+ (Rn ), 222 centered modified maximal operator,
Lip(R), 106 275
Lip(Rn ), 74 Choquet integral, 397, 398
LMpq (Rn ), 345 classical Morrey space, 13
∗
LMpq (Rn ), 345 commutator, 228, 432, 433
conjugate function, 52
LMfpq (Rn ), 345
Cube testing for Mα , 333
L logq L(µ), 58
p