Analysis in Banach Spaces.
Analysis in Banach Spaces.
Tuomas Hytönen
Jan van Neerven
Mark Veraar
Lutz Weis
Analysis in
Banach Spaces
Volume III: Harmonic Analysis and
Spectral Theory
Ergebnisse der Mathematik und ihrer
Grenzgebiete. 3. Folge / A Series of Modern
Surveys in Mathematics
Volume 76
Series Editors
L. Ambrosio, Pisa, Italy
V. Baladi, Paris, France
G. -M. Greuel, Kaiserslautern, Germany
M. Gromov, Bures-sur-Yvette, France
G. Huisken, Tübingen, Germany
J. Jost, Leipzig, Germany
J. Kollár, Princeton, USA
G. Laumon, Orsay, France
U. Tillmann, Oxford, UK
D. B. Zagier, Bonn, Germany
Ergebnisse der Mathematik und ihrer Grenzgebiete, now in its third sequence, aims
to provide reports, at a high level, on important topics of mathematical research. Each
book is designed as a reliable reference covering a significant area of advanced
mathematics, spelling out related open questions, and incorporating a comprehensive,
up-to-date bibliography.
Tuomas Hytönen • Jan van Neerven
Mark Veraar • Lutz Weis
This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
ix
x Contents
Q Questions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 733
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 783
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 820
Symbols and notations
Sets
Vector spaces
s
Bp,q - Besov space
c0 - space of null sequences
C - space of continuous functions
C0 - space of continuous functions vanishing at infinity
C α - space of Hölder continuous functions
Cb - space of bounded continuous functions
Cc - space of continuous functions with compact support
Cc∞ - space of test functions with compact support
C p - Schatten class
γ(H, X) - space of γ-radonifying operators
xv
xvi Symbols and notations
A - σ-algebra
dfn = fn − fn−1 - nth martingale difference
n - signs in K, i.e., scalars in K of modulus one
εn - Rademacher variables with values in K
E - expectation
F , G , . . . - σ-algebras
Ff - collection of sets in F on which f is integrable
E(·|·) - conditional expectation
γn - Gaussian variables
hI - Haar function
Symbols and notations xvii
µ - measure
kµk - variation of a measure
(Ω, A , P) - probability space
P - probability measure
rn - real Rademacher variables
(S, A , µ) - measure space
σ(f, g, . . . ) - σ-algebra generated by the functions f, g, . . .
σ(C ) - σ-algebra generated by the collection C
τ - stopping time
wα - Walsh function
Operators
Miscellaneous
,→ - continuous embedding
1A - indicator function
a . b - ∃C such that a 6 Cb
a .p,P b - ∃C, depending on p and P , such that a 6 Cb
Symbols and notations xix
C - generic constant
{ - complement
d(x, y) - distance
f ? - maximal function
fe - reflected function
fbb - Fourier transform
f - inverse Fourier transform
f ∗ g - convolution
= - imaginary part
M f - Hardy–Littlewood maximal function
p0 = p/(p − 1) - conjugate exponent
p∗ = max{p, p0 }
℘ - good set-bound
< - real part
s ∧ t = min{s, t}
s ∨ t = max{s, t}
x - generic element of X
x∗ - generic element of X ∗
x ⊗ y - elementary tensor
x+ , x− , |x| - positive part, negative part, and modulus of x
w - weight
wα - power weight t 7→ tα
Standing assumptions
xxi
11
Singular integral operators
[
Tcf (ξ) = K ∗ f (ξ) = K(ξ)
b fb(ξ) =: m(ξ)fb(ξ);
Here, and in many occasions below where we will use the same notation, it is
understood that the supremum is taken over all measurable subsets E of Q
satisfying the stated requirement that |E| 6 λ|Q|. The idea is to quantify how
much f deviates from a constant, if we ignore its (possibly wild) behaviour
on an exceptional set of controlled proportion. The above way of measuring
oscillations is essentially ‘minimal’ in that it can be controlled by average Lq
oscillations for any q > 0:
is called a median of f on the cube (or more general set of finite positive
measure) Q ⊆ Rd . One routinely checks that a median always exists but may
fail to be unique.
|mg | 6 kg1Q\E k∞
|Q ∩ {|g| > |mg |} \ E| > |Q ∩ {g > mg } \ E| > 21 |Q| − |E| > ( 21 − λ)|Q| > 0
and thus kg1Q\E k∞ > |mg |. If mg < 0, the argument is the same, just replac-
ing the second step above by |Q ∩ {g 6 mg } \ E|.
The previous lemma motivates the following:
Indeed, Lemma 11.1.2 says that the usual median is a λ-pseudomedian for
every λ ∈ (0, 21 ). Concerning existence in the general case, we have:
4 11 Singular integral operators
Proof. If oscλ (f ; Q) > 0, this is obvious, since we can always come within any
positive distance from the infimum. So only the case oscλ (f ; Q) = 0 needs
attention. In this case, there we can find a sequence of vectors cn ∈ X and
sets En ⊆ Q with |En | 6 λ|Q| such that k(f − cn )1Q\En k∞ → 0. Since
|En ∪ Em | 6 2λ|Q| < |Q|, any Q \ (En ∪ Em ) has positive measure, and thus
Thus (cn )n>1 is a Cauchy sequence and hence convergent to some c ∈ X. But
then
satisfies |E 0 | 6 λ|Q|.
since |E ε \ E| > |E ε | − |E| > λ|Q| − λ|Q| = 0. Taking the infimum over all
|E| 6 λ|Q|, we contradict the definition of a λ-pseudomedian.
Let us recall and expand the terminology related to dyadic cubes that we
introduced in Chapter 3.
Remark 11.1.7. One might like to replace (i) in Definition 11.1.6 by the “more
intrinsic”
(iii) each Dj is a partition of Rd consisting of left-closed, right-open cubes of
side-length 2−j .
When d = 1, one can check that (i) and (iii) are equivalent. But, for d > 1,
condition (iii) is strictly more general. For instance
where all cubes in the right half-plane are shifted in the y-direction by a fixed
amount α ∈ R relative to the standard dyadic cubes, would qualify for (iii) but
not for (i). The preference over one or the other definition may be a question
of taste; we choose to work with Definition 11.1.6 as stated.
but here and there we will also make use of other systems, which makes it
convenient to deal with a generic system from the beginning. For any given
cube, we may speak of its dyadic subcubes, by which we understand all cubes
obtained by repeatedly bisecting the edges of Q. We will use the notation
D(Q) for the collection of all dyadic subcubes of a cube Q. If Q belongs to a
dyadic system D, then
Remark 11.1.9. The standard dyadic system D 0 has 2d quadrants of the form
S1 × · · · × Sd , where Si ∈ {(−∞, 0), [0, inf ty)} for each i ∈ {1, . . . , d}. It is
also easy to construct dyadic systems, where Rd is the only quadrant.
6 11 Singular integral operators
where the supremum is taken over all dyadic cubes containing x. Here, and
throughout this chapter, unless indicated otherwise, integrals are taken with
respect to Lebesgue measure and are abbreviated in theRabove way to unbur-
den notation. Thus, when g is an integrable function, Q g is shorthand for
g(x) dx. When integrating over all of Rd we will even write g for Rd g.
R R R
Q
Proof of Proposition 11.1.11. If p ∈ [1, ∞), we dualise the left side of (1)
0
against φ ∈ Lp :
Z X Z
X 1 X
a S 1S φ = aS |S|− φ 6 aS |E(S)| inf MD φ
S γ S
S∈S S∈S S∈S
Z X
1 1 X
6 aS 1E(S) MD φ 6 aS 1E(S) kMD φkp0 ,
γ γ p
S∈S S∈S
where kMD φkp0 6 pkφkp0 by Doob’s maximal inequality (Theorem 3.2.2; cf.
the explanations preceding Theorem 3.2.27).
11.1 Local oscillations of functions 7
almost everywhere.
By the maximality of the Q1j , their parent cubes Q
b 1 satisfy the opposite
j
bound |Qb 1 ∩ E| 6 α|Q
b 1 |. Hence in particular
j j
|Q1j ∩ E 0 | 6 |Q
b 1 ∩ E 0 | 6 α|Q
j
b 1 | = 2d α|Q1 |.
j j
Let also n o
Ej1 = Q1j ∩ kf − mf (Q1j )k > 2 oscλ (f ; Q1j )
If 2d α + λ < 1, then Q1j \ (E 0 ∪ Ej1 ) has positive measure, and for any x in
this set, we have both
where each term in the last sum has exactly the same form as the left hand
side and allows to iterate the same consideration.
Assuming that we have proved
N
X −1 X X
1Q0 kf − mf (Q0 )k 6 4 1Qjn oscλ (f ; Qnj ) + 2 1QjN oscλ (f ; QjN )
n=0 j j
X
+ 1QN
j
kf − mf (QN
j )k,
j
Choosing α = 2λ, (11.4) shows that the collection {Qnj }n,j is 21 -sparse, and
with λ = 2−2−d , we also have 2d α + λ = (2d+1 + 1)λ = 2−1 + 2−1−d < 1, as
required. This concludes the proof.
Proof. We assume that the right-hand side is finite, for otherwise there is
nothing to prove. We then assume without loss of generality that S is finite.
In fact, once we have proved the result for finite families,
P in the infinite case
it follows easily that the partial sums of the series S∈S fS (with arbitrary
enumeration) form a Cauchy sequence in Lp (Rd ; X), from which we deduce
the (unconditional) convergence of this series and the asserted norm bound.
0
Concentrating on the finite case, by dualising with g ∈ Lp (Rd ; X ∗ ), it is
equivalent to the estimate
Z X X 1/p
0
hfS , gi dx 6 (1 + γ −1/p p) kfS kpLp (Rd ;X) kgkLp0 (Rd ;X ∗ ) .
S S
We can estimate the second term by Hölder’s inequality and the pairwise
disjointness of the sets ES (S),
Z X X
1ES (S) hfS , gi dx 6 1ES (S) fS p d kgkLp0 (Rd ;X ∗ )
L (R ;X)
S∈S S∈S
X 1/p
= k1ES (S) fS kpLp (Rd ;X) kgkLp0 (Rd ;X ∗ ) .
S∈S
where in the second factor we rearranged the double sum into a single sum,
observing that every S 0 ∈ S is counted at most once as a child of a unique
0
S ∈ S . The second factor is bounded by γ −1/p pkgkp0 thanks to Proposi-
tion 11.1.11(2). Summing up the bounds, we complete the proof of the direct
estimate.
The following lemma describes useful projections and also provides prominent
examples of the functions fS featuring in Proposition 11.1.14.
Lemma 11.1.15. For S ∈ S ⊆ D and f ∈ Lloc
1
(Rd ; X), let
X
PS f := ES 0 f + 1ES (S) f. (11.5)
S 0 ∈chS (S)
|S 0 ∩ Q|
Z
X 1 X
01 X
hf iS 0 = hf iS 0 |S | = f dx,
|Q| |Q| |Q| S0
S 0 ∈chS S 0 ∈chS ,S 0 ⊆Q S 0 ∈chS ,S 0 ⊆Q
and
Z Z Z
1 X 1 1
hPS f iQ = f dx + f dx = f dx,
|Q| S0 |Q| Q∩ES (S) |Q| Q
S 0 ∈chS ,S 0 ⊆Q
The first instance of the interplay of a function and its principal cubes is the
following:
Note that (11.6) is slightly more than the mere (1 − A−1 )-sparseness of S :
it says that the disjoint subsets E(S) ⊆ S in the definition of sparseness may
be chosen as E(S) = ES (S), which is not always the case for an arbitrary
sparse family. For instance, S = {[0, 1), [0, 21 ), [ 12 , 1)} is 12 -sparse, and one can
take for instance E([0, 1)) = [ 41 , 34 ), E([0, 12 )) = [0, 41 ) and E([ 21 , 1)) = [ 43 , 1),
but ES ([0, 1)) = ∅ in this case.
Proof. By maximality, the cubes S 0 ∈ chS (S) are pairwise disjoint. From the
defining condition it follows that
11.1 Local oscillations of functions 13
R R
X X kf kX dx
S0
kf kX dx |S|
|S 0 | 6 6 S =
Ahkf kX iS Ahkf kX iS A
S 0 ∈chS (S) S 0 ∈chS (S)
and hence X 1
|ES (S)| = |S| − |S 0 | > (1 − )|S|.
A
S 0 ∈chS (S)
These two cases confirm the upper bound kPS f kL∞ (Rd ;X) 6 2d Ahkf kX iS .
where the supremum is taken over all cubes containing x ∈ Rd ; a dyadic ver-
#,D
sion M0,λ is obtained by restricting the supremum to dyadic cubes Q ∈ D
only. Via this maximal operator we can obtain a useful oscillatory character-
isation of Lp (Rd ; X), which we will prove in the rest of this section:
−1/p # #
cd kM0,λ f kLp (Rd ) 6 kf − ckLp (Rd ;X) 6 cp kM0,λ f kLp (Rd ) ,
Remark 11.1.19. If we now a priori require that f ∈ Lp0 ,∞ (Rd ; X) for some
p0 ∈ (0, ∞) (unrelated to the exponent p), then the constant c ∈ X guaranteed
by Theorem 11.1.18 is necessarily 0, and thus in fact f ∈ Lp (Rd ; X).
Namely, if f ∈ Lp0 ,∞ (Rd ; X) and f − c ∈ Lp (Rd ; X), it follows that c =
f − (f − c) ∈ Lp0 ,∞ (Rd ; X) + Lp (Rd ; X), thus
|{kck > t}| 6 |{kf k > t/2}| + |{kf − ck > t/2}| < ∞
for all t > 0, which would lead to a contradiction for t ∈ (0, kck).
By Lemma 11.1.1 for any q ∈ (0, ∞), we have
k(f − c)1Q kLq,∞ kf 1Q kLq −1/q
Z 1/q
oscλ (f ; Q) 6 inf 1/q
6 1/q
= λ − kf kq .
c∈X (λ|Q|) (λ|Q|) Q
Taking the supremum over all cubes Q containing a given point, it follows
that 1/q
#
M0,λ f 6 λ−1/q Mq f, Mq f := M (kf kq ) , (11.7)
where M is the Hardy–Littlewood maximal operator. The Lp boundedness of
Mq is an easy combination of some estimates collected from Chapter 3:
Lemma 11.1.20. For all 0 < q < p < ∞, we have
p 1/q
max kMq kLp →Lp , kMq kLp,∞ →Lp,∞ 6 3d/q+d/p .
p−q
Proof. The dyadic (in fact more general martingale) bounds for MqD on Lp
and Lp,∞ for p ∈ (q, ∞), with norm bound (p/(p − q))1/q in each case, have
been treated in Lemma 3.5.17. On the other hand, we recall from (3.36) that
M f 6 3d sup M α f,
α∈{0, 31 , 23 }d
thus 1/p
X
Mq f 6 3d/q sup Mqα f 6 3d/q [Mqα f ]p .
α∈{0, 31 , 23 }d
α∈{0, 13 , 23 }d
Hence
X 1/p p 1/q
kMq f kp 6 3d/q kMqα f kpp 6 3d/q+d/p kf kp ,
p−q
α∈{0, 13 , 23 }d
after which the last step is exactly as in the strong-type case, now using the
weak-type boundedness of the dyadic Mqα instead.
11.1 Local oscillations of functions 15
#
Proposition 11.1.21. The operator M0,λ is bounded from Lp (Rd ; X) to
1/p
Lp (Rd ) and from Lp,∞ (Rd ; X) to Lp,∞ (Rd ), with norm at most cd,λ , where
cd,λ is a constant depending only on d and λ.
The first half of Theorem 11.1.18 is immediate from this proposition (with the
choice λ = 2−2−d so that cd,λ = cd ), combined with the trivial observation
# #
that M0,λ f = M0,λ (f − c) for any constant c ∈ X.
Proof. Let Y ∈ {Lp , Lp,∞ }. By (11.7) and Lemma 11.1.20, we have
p 1/q
#
kM0,λ f kY 6 λ−1/q kMq f kY 6 λ−1/q 3d/q+d/p kf kY .
p−q
With, say, q = 21 p, the right hand side takes the form (λ−2 33d 22 )1/p kf kY .
Towards the deduction of a global Lp estimate from local ones, we record:
Lemma 11.1.22. Let X be a Banach space and p ∈ (0, ∞). Suppose that
f ∈ Lploc (Rd ; X) satisfies
k1Q (f − cQ )kp 6 K
kf − ckp 6 K.
Proof.
S∞ Consider an increasing sequence of cubes Q1 ⊆ Q2 ⊆ . . . such that
d
n=1 n = R . If m 6 n, then
Q
from which the first claimed inequality follows by taking the supremum over
all cubes Q ⊆ Rd .
In the other direction, given a cube Q ⊆ Rd , Lerner’s formula (Theorem
11.1.12) guarantees that
Z
4 X
− kf − mf (Q)k 6 |S| oscλ (f ; S)
Q |Q|
S∈S
4 X # #
6 2|E(S)|kM0,λ f k∞ 6 8kM0,λ f k∞ ,
|Q|
S∈S
and taking the supremum over all cubes Q proves the second bound.
∞
X ∞
X
kαk k∞ |Qk | < ∞ |λk | < ∞ .
k=1 k=1
It is immediate that the two versions of the definition are equivalent via the
correspondence λk = kαk k∞ |Qk | and ak = λ−1k αk .
A disadvantage of this definition is the difficulty of checking the mem-
1
bership of a given function in HD,at (Rd ; X), as doing this via the definition
would require one to construct the atomic decomposition, which might not be
an easy task. The following notion is much more amenable to this:
Definition 11.1.27 (Maximal Hardy space). The maximal Hardy space
1
HD,max (Rd ; X)
consists of all f ∈ L1 (Rd ; X) for which also the (cancellative) dyadic maximal
function
MD f (x) := sup 1Q (x)khf iQ kX
Q∈D
1 d
satisfies MD f ∈ L (R ). The norm in this space is defined as
kf kHD,max
1 := kMD f kL1 (Rd ) .
1
Theorem 11.1.28. Let X be a Banach space. The spaces HD,at (Rd ; X) and
1 d
HD,max (R ; X) are equal with equivalent norms; in fact
d
khkHD,max
1 (Rd ;X) 6 khkHD,at
1 (Rd ;X) 6 6 · 2 · khkHD,max
1 (Rd ;X) .
Proof. Suppose
R first that a ∈ L∞ (Rd ; X) satisfies supp a ⊆ Q for some dyadic
cube and a = 0. Then haiR 6= 0 only if R ( Q, and hence supp MD a ⊆ Q
as well. It follows that
kMD ak1 6 |Q|kMD ak∞ 6 |Q|kak∞ .
P∞
If h = i=1 ai is a series of such function on cubes Qi , then by sublinearity
∞
X ∞
X
khkHD,max
1 (Rd ;E) = kMD hk1 6 kMD ai k1 6 |Qi |kai k∞ ,
i=1 i=1
and taking the infimum over all such representations of h shows that
khkHD,max
1 (Rd ;X) 6 khkHD,at
1 (Rd ;X) .
1
In the other direction, suppose that h ∈ Hmax (Rd ; X). Given λ > 0, let Qλ
be the collection of maximal dyadic cubes Q such that khhiQ kX > λ. Then
X 1 1
|Q| = |{MD h > λ}| 6 kMD hkL1 (Rd ) = khkHD,max
1 (Rd ;X) .
λ λ
Q∈Qλ
Let Qbλ be the collection of maximal dyadic cubes that have a child in Qλ .
Thus these cubes do not belong to Qλ themselves. Hence khhiQ kE 6 λ for
Q ∈ Qbλ , and also
11.1 Local oscillations of functions 19
X X X
|Q| 6 |Q|
b = 2d |Q| = 2d |{M h > λ}|.
Q∈Q
bλ Q∈Qλ Q∈Qλ
Let then
X X
gλ := 1{(S Qbλ ) h + 1Q hhiQ , bλ := 1Q (h − hhiQ ).
Q∈Q
bλ Q∈Q
bλ
1
Corollary 11.1.29. The space HD,at (Rd ; X) = HD,max
1
(Rd ; X) is complete.
20 11 Singular integral operators
1
Proof. It in enough to prove this for HD,max (Rd ; X). Since kf (x)kX 6
d
MD f (x) at a.e. x ∈ R , we have kf kL1 (Rd ;X) 6 kf kHD,max
1 (Rd ;X) . Hence, if
∞ 1
(fn )n=1 is a Cauchy sequence in HD,max (Rd ; X), it is also a Cauchy sequence
in L (R ; X) and thus kfn − f k1 → 0 for some f ∈ L1 (Rd ; X). Since h iQ is
1 d
hence MD (f −h)(x) 6 lim inf n→∞ MD (fn −h)(x), and thus by Fatou’s lemma
1
With h = 0, this shows that f ∈ HD,max (Rd ; X). With h = fm , we find that
lim kMD (f − fm )kL1 (Rd ) 6 lim lim inf kMD (fn − fm )kL1 (Rd ) = 0,
m→∞ m→∞ n→∞
1
and hence fm → f in HD,max (Rd ; X).
1
as an isometric subspace of (HD,at (Rd ; X))∗ .
Proof. Since all norms BMO norms appearing in this proof are dyadic, we
drop the subscript D for the benefit of slightly lighter notation.
Z Z Z
hb, ai i = hb − c, ai i 6 − kb − ckX ∗ |Qi |kai k∞
Qi Qi
Thus
∞ Z
X ∞
X Z ∞ Z
X Z
hb, ai i = lim hbN , ai i = lim hbN , ai i = lim hbN , hi,
N →∞ N →∞ N →∞
i=1 i=1 i=1
where the first two identities use dominated convergence in L1 (Qi ) and in `1 ,
respectively, and the last one follows from the convergence of the series h =
P ∞ 1 d ∞ d ∗ 1 d ∗
i=1 ai in L (R ; E), and the fact that bN ∈ L (R ; E ) ⊆ (L (R ; E)) .
This shows in particular that the pairing of hb, hi is independent of the par-
ticular series representation of h, and hence well defined. Taking the infimum
over all representations in the estimate
∞
X
|hb, hi| 6 kbkBMO |Qi |kai k∞ ,
i=1
we find that
kbk(Hat
1 (Rd ;X))∗ 6 kbkBMO(Rd ;X ∗ ) . (11.8)
For the converse estimate, consider a cube Q and suppose first that s ∈
L1 (Q; X ∗ ) is a simple function, thus measurable with respect to a finite σ-
algebra F of Q. The advantage of this setting is that, for a finite σ-algebra,
0
we have the duality (Lp (F ; X))∗ = Lp (F ; X ∗ ) for an arbitrary Banach space
X and for every p ∈ [1, ∞], including in particular p = ∞. Now inf c∈E ∗ ks −
ckL1 (Q;X ∗ ) is the norm of the equivalence class [s] ∈ L1 (F ; X ∗ )/X ∗ , where
L1 (F ; X ∗ ) = (L∞ (F ; X))∗ .
We claim that the quotient space above is the dual of the subspace
L0∞ (F ; X) ⊆ L∞ (F ; X) of functions with mean zero. In fact, recall from
Proposition B.1.4 that for any subspace Y ⊆ Z, we have the identification
Y ∗ = Z ∗ /Y ⊥ , the quotient of Z ∗ with the annihilator Y ⊥ of Y in Z ∗ . Now
Z = L∞ (F ; X) for a finite σ-algebra F , in which case Z ∗ = L1 (F ; X ∗ ). To
identify Y ⊥ for Y = L0∞ (F R ; X), it is easy to check that the only functions
f ∈ L1 (F ; X ∗ ) for which hf, gi = 0 for all g ∈ L∞0 (F ; X) are the constant
functions. Thus indeed L1 (F ; X ∗ )/X ∗ = (L0∞ (F ; X))∗ , and hence
22 11 Singular integral operators
Z
inf ks − ckL1 (Q;X ∗ ) = k[s]kL1 (F ;X ∗ )/X ∗ = sup hs, gi .
c∈X ∗ g∈L∞0 (F ;X)
kgk∞ 61
Definition 11.2.1. Let X and Y be Banach spaces, p0 ∈ [1, ∞], and consider
(1) We say that T has kernel K, or that K is the kernel of T , if for ev-
ery f ∈ Lpc 0 (Rd ; X) and almost every s at a positive distance from
supp f the following holds: for every functional y ∗ ∈ Y ∗ , the function
t 7→ hK(s, t)f (t), y ∗ i is integrable, and
Z
hT f (s), y i = hK(s, t)f (t), y ∗ i dt.
∗
11.2 Singular integrals and extrapolation of Lp0 bounds 23
∞ L∞ (Rd ;X)
L̄fin (Rd ; X) := L∞ d
fin (R ; X) , where
L∞ d
fin (R ; X)
∞ d
:= {f ∈ L (R ; X) : |{f 6= 0}| < ∞}.
Since this can be done for any ε > 0, we find that f belongs to the L∞ (Rd ; X)-
closure of L∞ d p d ∞ d
fin (R ; X). Since f ∈ L (R ; X) ∩ L (R ; X), this whole intersec-
tion belongs to the said closure, and then so does the closure of this intersec-
tion. This completes the proof.
for some disjoint dyadic cubes Qi . If f is simple, then all biR are also simple.
If f ∈ L1 (Q0 ; X) for some cube Q0 ⊆ Rd and λ > 2−d −Q0 kf k, then the
cubes Qi can be chosen as dyadic subcubes of the initial Q0 , and the function
g to be supported on Q0 .
If f ∈ L1 (S; X) for some quadrant of Rd , then we have Qi ⊆ S.
26 11 Singular integral operators
|Q
bi |
Z Z
1 1
kg(x)kX = khf iQi kX 6 kf kX 6 · kf kX 6 2d · λ
|Qi | Qi |Qi | |Qbi | Q
bi
Proof of Theorem 11.2.5(1). Our plan is to first prove the weak-type result
(1b), and then obtain the strong-type bound (1a) via the Marcinkiewicz In-
terpolation Theorem 2.2.3.
For f ∈ Lp0 (Rd ; X) ∩ L1 (Rd ; X) and λ > 0, we estimate λ|{kT f k > λ}|.
Let f = g + b the Calderón–Zygmund decomposition of f at level αλ
(instead of λ), where α is to be determined. Then
1/p0 1/p0 0 1/p0
kgkp0 6 kgk∞ 0 kgk1 6 (2d αλ)1/p0 kf k1 ,
p0 d p0 d
P in particular g ∈ L (R ; X), and thus b = f − g ∈ L (R ; X). Since b =
so
i bi and the bi are disjointly supported, it follows that each bi also belongs to
Lp0 (Rd ; X) and the identity b = i bi also holds in the sense of convergence
P
in Lp0 (Rd ; X). The assumption that T ∈ L (Lp0 (Rd ; X), Lp0 ,∞ (Rd ; Y )) then
implies that
X X
T f = T (g + b) = T g + T b, Tb = T bi = T bi .
i i
|{kT f k > λ}| 6 |{kT gk > λ/2}| + |{kT bk > λ/2} \ O∗ | + |O∗ |, (11.11)
In order to estimate the ith term here, we denote by zi the common centre
of the cube Qi and the ball Bi . Now the integral representation of T bi (s) is
available at s ∈ {Bi . Explicitly, for each y ∗ ∈ Y ∗ ,
Z Z
hT bi (s), y ∗ i = hK(s, t)bi (t), y ∗ i dt = h[K(s, t) − K(s, zi )]bi (t), y ∗ i dt,
R
where the last step follows from the fact that bi (t) dt = 0. Thus
Z
kT bi (s)kY 6 k[K(s, t) − K(s, zi )]bi (t)kY dt
Qi
and hence
Z Z Z
kT bi (s)kY ds 6 k[K(s, t) − K(s, zi )]bi (t)kY ds dt
{Bi Qi {Bi
28 11 Singular integral operators
Z
6 kKkHör kbi (t)kX dt,
Qi
Ap00 2p0 p0
|{kT gk > λ/2}| 6 kgkp0
p 6 A · (2d αλ)p0 −1 kf k1 ,
(λ/2)p0 0
λp 0 0
so that altogether
(2A · 2d α)p0 cd kf k1
0
|{kT f k > λ}| 6 d
+ 4kKkHör + ,
2 α α λ
where we are still free to choose α > 0. Taking
leads to
kf k1
|{kT f k > λ}| 6 (cd A0 + 4kKkHör ) . (11.13)
λ
If p0 = ∞, we observe that kT gk∞ 6 A0 kgk∞ 6 A0 2d αλ, so that the
same choice of α guarantees that |{kT gk > λ/2}| = 0. Thus, in this case, we
only need to estimate the last two terms in (11.11), and these have exactly
the same bounds in the case p0 < ∞ that was already handled.
We have hence confirmed (11.13) for all f ∈ Lp0 (Rd ; X) ∩ L1 (Rd ; X)
and λ > 0, and this proves the existence of a unique bounded extension
T ∈ L (L1 (Rd ; X), L1,∞ (Rd ; X)) by the density of Lp0 (Rd ; X) ∩ L1 (Rd ; X) in
L1 (Rd ; X). This completes the proof of (1b).
(1b) in case (4): Let then Rd be replaced by a cube Q0 . Note that
and by elementary calculus one verifies that p0p0 −1 6 e for p0 ∈ (1, ∞). Sub-
stituting these estimates, we obtain
n p0 − p o p1
kT kL (Lp (Rd ;X),Lp (Rd ;Y ) 6 2e · cd · (A0 + kKkHör ),
(p0 − p)(p − 1)
which coincides with the claim after redefining cd . Since the Marcinkiewicz
Interpolation Theorem 2.2.3 is valid for general measure spaces, the same
argument applies equally well in the case of a cube or a quadrant as the
underlying domain.
Lemma 11.2.7. Let X and Y be Banach spaces and p0 ∈ [1, ∞]. Let T ∈
L (Lp0 (Rd ; X), Lp0 ,∞ (Rd ; Y )) be an operator with dual Hörmander kernel K.
If B ⊆ Rd is a ball and f ∈ Lp0 (Rd ; X) ∩ L∞ (Rd ; X) is supported in {B, then
for almost all s, s0 ∈ 21 B, we have
Z
hT f (s) − T f (s0 ), y ∗ i = h[K(s, t) − K(s0 , t)]f (t), y ∗ i dt ∀y ∗ ∈ Y ∗ .
{B
Proof.
S∞ Consider an increasing sequence of balls B1 ⊆ B2 ⊆ . . . such that
d
n=1 n = R , and let fn := 1Bn f . Since fn = 1{B fn is compactly supported
B
away from B, for almost every s ∈ 21 B we have
Z
∗
hT fn (s), y i = hK(s, t)fn (t), y ∗ i dt ∀y ∗ ∈ Y ∗ .
{B
It is also clear that fn (t) → f (t) pointwise. On the other hand, the integrand
in (11.16) is pointwise dominated by
which is integrable over t ∈ {B (thus |t−zB | > rB > 2 max{|s−zB |, |s0 −zB |})
by the dual Hörmander condition. Hence
Z
RHS(11.16) → hf (t), [K(s, t)∗ − K(s0 , t)∗ ]y ∗ i dt
{B
The following lemma contains the technical core of the upper extrapolation:
where c = T (1{B f )(z), and z ∈ Q is fixed as one of the (almost all) points of
Q where the conclusion of Lemma 11.2.7 is valid for the function 1{B f . Thus
and hence
A kf k p0
0 ∞
|EΛ | := |{kT (1B f )k > Λ}| 6 cd |Q| 6 λ|Q|
Λ
if we choose Λ := (cd /λ)1/p0 A0 kf k∞ . We conclude that
Taking the supremum over y ∗ in the unit ball of Y ∗ and the essential supre-
mum over s ∈ Q, we arrive at
Hence altogether
#
(in place of M0,λ ◦ T : L∞ (Rd ; X) → L∞ (Rd )) allows us to deduce the relaxed
conclusion
#
M0,λ ◦ T : Lp0 (Rd ; X) ∩ Lp (Rd ; X) → Lp (Rd ), p ∈ (p0 , ∞), (11.17)
f
fet := f − t · 1{kf k>t} ,
kf k
f
fet := f · 1{kf k6t} + t · 1{kf k>t} ,
kf k
and it is immediate to verify that, if f ∈ Lp0 (Rd ; X), these remain in the
space Lp0 (Rd ; X), in addition to the other function space memberships used
in the proof of Theorem 2.2.3.
If θ ∈ (0, 1) is such that 1/p = (1 − θ)/p0 + θ/∞ = (1 − θ)/p0 , the
Marcinkiewicz Interpolation Theorem 2.2.3 shows that the norm of the oper-
ator in (11.17) is at most
since p0 > 1 here. Substituting back (and redefining cd ), we find that the
norm of the operator in (11.17) is at most
p 1/p
0
(cd A0 + 2kKkHör∗ ).
p − p0
Now Theorem 11.1.18, together with Remark 11.1.19 and the a priori
condition that T f ∈ Lp0 ,∞ (Rd ; X), show that
#
kT f kLp (Rd ;Y ) 6 8pkM0,λ (T f )kLp (Rd )
p 1/p
0
6p (cd A0 + 16kKkHör∗ )kf kLp (Rd ;X)
p − p0
34 11 Singular integral operators
for all f ∈ Lp (Rd ; X) ∩ Lp0 (Rd ; X) and p ∈ (p0 , ∞). Since this is a dense
subspace of Lp (Rd ; X), the operator T has a unique extension to this space,
with the same norm estimate above.
The case of a cube or a quadrant in place of Rd follows by the same
argument, since all results quoted are also valid in these settings.
It is also immediate from Lemma 11.2.8 and Proposition 11.1.24 that
#
kT f kBMO(Rd ;X) 6 8kM0,λ (T f )kL∞ (Rd ) 6 (cd A0 + 8kKkHör∗ )kf kL∞ (Rd ;Y )
for all f ∈ Lp0 (Rd ; X) ∩ L∞ (Rd ; X). Since this is not a dense subspace of
L∞ (Rd ; X), extending this estimate, and indeed the very meaning of “T f ”,
to all f ∈ L∞ (Rd ; X) requires an additional effort, to which we turn in Section
11.2.c below.
Proof of Theorem 11.2.5(3). We now assume that K satisfies both Hörmander
and dual-Hörmander conditions, and hence we have access to both cases (1)
and (2) that we already proved. By Theorem 11.2.5(1b), we have
We now use this estimate (rather than the original assumption) as input
to Theorem 11.2.5(2a), i.e., we apply the latter with 1 in place of p0 and
cd (A0 + kKkHör ) in place of A0 . This gives, for all p ∈ (1, ∞), the estimate
1 1/p
kT kL (Lp (Rd ;X),Lp (Rd ;Y ) 6 cd p cd (A0 + kKkHör ) + kKkHör∗
p−1
2 0
6 cd pp (A0 + kKkHör + kKkHör∗ ),
where we estimated
1 1/p p 1/p
6 = (p0 )1/p 6 p0 .
p−1 p−1
The conclusion agrees with the claim, after redefining cd .
The case of a cube or a quadrant in place of Rd is immediate, since both
(1) and (2) of the theorem, which we used above, were already proved in these
cases as well.
The goal of this section is to establish the following theorem, in which indistin-
guishability of BMO(Rd ; X) functions only differing by an additive constant
manifests itself.
Remark 11.2.10.
(1) By the John–Nirenberg inequality, the target space BMOp (Rd ; Y )/Y of Te
is independent of the value of p ∈ [1, ∞); however, the estimate for the
operator norm need not be, and we specifically state it with p = p0 .
(2) The left-hand side of (11.18) could be more pedantically written as “hh, gi,
where h ∈ [Tef ] is arbitrary”: the vanishing integral of g guarantees that
this expression is independent of the choice of h.
(3) The boundedness requirement on T in Theorem 11.2.9 may seem stronger
than in Theorem 11.2.5(2) (where it was only assumed that T maps bound-
edly into the larger space Lp0 ,∞ (Rd ; Y ) and for some p0 in the larger range
[1, ∞]), but this is only superficial, as we can always arrange ourselves to
be in the situation of Theorem 11.2.9 even under the apparently weaker
boundedness hypothesis:
First, if p0 = ∞, there is nothing to prove, as we can simply take
Te = T , which already maps into L∞ (Rd ; Y ) ⊆ BMO(Rd ; Y ). If, on
the other hand, p0 ∈ [1, ∞), Theorem 11.2.5(2a) guarantees that T ∈
L (Lp (Rd ; X), Lp (Rd ; Y )) for all p ∈ (p0 , ∞) ⊆ (1, ∞), and choosing one
such p as a new p0 , we are in the situation assumed in Theorem 11.2.9.
and taking the infimum over c ∈ X on the right, and then the supremum
over all cubes Q ⊆ Rd of the whole chain, we derive (11.20). The dyadic case
follows by taking the supremum over Q ∈ D(S) instead.
In view of Lemma 11.2.12, the construction of an extension
which exactly cancels out the first term in the formula of δ(B 0 ) − δ(B).
Let us then check how ∆T compares with the original T on the intersection
of their domains of definition:
Lemma 11.2.14. If f ∈ Lp0 (Rd ; X) ∩ L∞ (Rd ; X), then
hence
K
X
k vk 6 C. (11.24)
Y
k=1
But the two conditions (11.23) and (11.24) are precisely those of the Bessaga–
Pelczyński Theorem 1.2.40 that guarantee the containment of an isomorphic
copy of c0 in span(vk )∞
k=1 ⊆ Y . This contradicts the assumption on Y .
11.2 Singular integrals and extrapolation of Lp0 bounds 39
Proof. Let B be the ball concentric with Q and with twice the diameter of Q;
hence Q ⊆ 21 B. From the assumption that T ∈ L (Lp0 (Rd ; X), Lp0 (Rd ; Y ))
and f ∈ L∞ (Rd ; X), it is immediate that T (1B f ) ∈ Lp0 (Rd ; Y ) and
so that
k(s, u) 7→ T (1B f )(s) − T (1B f )(u)kLp0 (Q×Q;Y )
6 2|Q|1/p0 kT (1B f )kLp0 (Rd ;Y ) 6 cd A0 |Q|2/p0 kf k∞ ,
The more delicate matter is the integral in (11.21). Certainly this inte-
gral exists, since the dual Hörmander condition guarantees that [K(s, t)∗ −
K(u, t)∗ ]y ∗ is jointly measurable and belongs to L1 ({B, dt; Y ∗ ) uniformly in
(s, u) ∈ Q, while f ∈ L∞ (Rd ; Y ) by assumption. An immediate estimate
with the dual Hörmander condition shows that this integral is bounded by
kKkHör∗ kf k∞ ky ∗ kY ∗ , uniformly in x ∈ Q, and hence defines a Y ∗∗ -valued
function h(s, u) with the pointwise bound
then
40 11 Singular integral operators
∞
X
hy ∗ , h(s, u)i = hhn (s, u), y ∗ i ∀y ∗ ∈ Y ∗ . (11.26)
n=1
n=1 {B
P∞
so the series n=1 hn (s, u) converges absolutely, and hence in norm. Un-
der the mere dual Hörmander condition, but with the assumption that Y
does
P∞ not contain an isomorphic copy of c0 , the needed norm convergence of
n=1 hn (s, u) follows by Proposition 11.2.15 and the bound
∞
X Z
∗
|hhn (s, u), y i| 6 |hf (y), [K(s, t)∗ − K(u, t)∗ ]y ∗ i| dt
n=1 {B
6 2kKkHör∗ ky ∗ kY ∗ kf k∞ < ∞ ∀y ∗ ∈ Y ∗ .
P∞
In both cases, by (11.26), the limit of n=1 hn (s, u) must be h(s, u). Thus, as
a pointwise limit of Y -valued strongly measurable functions, h itself must be
both Y -valued and strongly measurable. Once these qualitative properties are
verified, the quantitative Lp0 (Q × Q; Y ) estimate is immediate by integrating
over Q × Q the already observed pointwise bound (11.25).
Now we are prepared to complete:
p0
Proof of Theorem 11.2.9. The operator ∆T : L∞ (Rd ; X) → Lloc (Rd+d ; Y ) is
well defined by Lemma 11.2.16 and satisfies
Tef := [∆T f (·, u)] (the equivalence class modulo constants), (11.27)
In = hT (1En f ), gi.
For IIn , recalling the construction of Te from (11.27) with u = 0, and then the
definition of ∆T f (s, u) from (11.21) with B = Bn , we have
since Bn ⊆ En . Finally,
Z Z
|IVn | 6 kf kL∞ (Rd ;X) k[K(s, t) − K(0, t)]∗ g(s)kX ∗ dt ds.
Rd {Bn
which concludes the proof of the remaining Claim (b) of Theorem 11.2.9.
Lemma 11.3.3.
∞
i 1 1
X
ωK ( ) 6 (1 + 2d )cK , i
ωK (2−k ) 6 kω i kDini .
2 log 2 K
k=2
11.3 Calderón–Zygmund operators and sparse bounds 43
hence
Z 2−k
du
ω(2−k−1 ) log 2 6 ω(u) ,
2−k−1 u
and thus
∞ ∞ Z 1/2
X X 1 du
ω(2−k ) = ω(2−1−k ) 6 ω(u) .
log 2 0 u
k=2 k=1
i
ci
kωK kDini 6 2d+1 cK 1 + log+ K .
cK
Note that, in concrete situations, the constants cδ or ciK are often much larger
than cK . The point of the bounds in parts (3) and (4) is that these larger
constants contribute to the Dini bounds only logarithmically.
Proof. We will first prove (2); the proof of (1) is analogous.
Z
kK(x, y) − K(x, y 0 )k dx
|x−y|>2|y−y 0 |
Z |y − y 0 | 1
2
6 ωK dx
|x−y|>2|y−y 0 | |x − y| |x − y|d
Z ∞ |y − y 0 | dr Z 12
2 2 dt 2
= σd−1 ωK = σd−1 ωK (t)d = σd−1 kωK kDini
0
2|y−y | r r 0 t
cK cK cK 2d+1 cK
kK(x0 , y) − K(x, y)k 6 + 6 (2 d
+ 1) 6 .
|x0 − y|d |x − y|d |x − y|d |x − y|d
This is the claimed standard estimate, and the Dini estimate follows from part
(3) with δ = 1 and cδ = 2d+1 ciK .
The version with a cube or a quadrant follows with the same argument by
simply restricting all the variables and the integrals to the relevant domain of
definition.
In particular, Dini kernels satisfy both Hörmander and dual Hörmander con-
ditions, and hence all the results of the previous section apply to them:
Corollary 11.3.5 (Calderón–Zygmund). Let X and Y be Banach spaces
and p0 ∈ [1, ∞]. Let T ∈ L (Lp0 (Rd ; X), Lp0 ,∞ (Rd ; Y )) be an operator with
a Calderón–Zygmund kernel K. Then all conclusions of Theorem 11.2.5 hold
2 1
with kKkHör replaced by kωK kDini and kKkHör∗ by kωK kDini in the estimates.
Proof. This follows at once from Theorem 11.2.5, where the same conclusions
are deduced for Hörmander and/or dual Hörmander kernels K, and Lemma
11.3.4, where these assumptions are verified for under the Dini conditions.
Suppose that both T and MT# are bounded from L1 (Rd ; X) to L1,∞ (Rd ). Then
for every boundedly supported f ∈ L1 (Rd ) and ε ∈ (0, 1), there is a (1 − ε)-
sparse family S of dyadic cubes such that, almost everywhere,
8 · 10d · cT X
Z
Tf 6 1S − kf k,
ε 5S S∈S
where
cT := kT k1→1,∞ + kMT# k1→1,∞ . (11.29)
4 · 10d cT
Z X Z X
1Q0 T (15Q0 f ) 6 1Q0 − kf k+ 1Qj − kf k + 1Qj T (15Qj f ),
ε 5Q0 j 5Qj j
and
1Qj T (15Q0 \5Qj f ) 6 1Qj [inf MT# (15Q0 f ) + inf T (15Q0 \5Qj f )]
Qj Qj
(11.32)
6 1Qj [inf MT# (15Q0 f ) + inf {T (15Q0 f ) + T (15Qj f )}]
Qj Qj
where we used sublinearity and the definition of MT# to get the estimates.
Note that no convergence issues arise when viewing the above lines in the
pointwise sense.
The last term in (11.31) already has the correct form, and it remains to
choose the cubes Qj in such a way that we have (11.30) as well as
Z
X cd cT
1Q0 \Sj Qj T (15Q0 f ) + 1Qj T (15Q0 \5Qj f ) 6 1Q0 − kf k.
j
ε 5Q0
11.3 Calderón–Zygmund operators and sparse bounds 47
F (Q) := Q ∩ [{T (15Q f ) > λhkf ki5Q } ∪ {MT# (15Q f ) > λhkf ki5Q }].
|F (Q)| 6 |{T (15Q f ) > λhkf ki5Q }| + |{MT# (15Q f ) > λhkf ki5Q }|
k15Q f k1 5d (11.33)
6 (kT k1→1,∞ + kMT k1→1,∞ ) = cT · |Q|.
λhkf ki5Q λ
|Qj ∩ F (Q0 )|
> 2−d−1 .
|Qj |
The cubes Qj are disjoint, so that
X X |Qj ∩ F (Q0 )| 2 · 10d
|Qj | 6 −d−1
6 2d+1 |F (Q0 )| 6 cT · |Q0 | = ε|Q0 |,
j j
2 λ
2 · 10d
λ := cT .
ε
Substituting back to (11.33), this choice gives in particular that
On the other hand, the maximality of Qj implies that its dyadic parent Q
bj
satisfies the opposite inequality, and hence
|Qj ∩ F (Q0 )| |Q
b j ∩ F (Q0 )| 2−d−1 1
6 6 −d = .
|Qj | −d
2 | Qj |
b 2 2
1Qj T (15Q0 \5Qj f ) 6 1Qj [MT# (15Q0 f )(zj ) + T (15Q0 f )(zj ) + T (15Qj f )(zj )]
6 1Qj [λhkf ki5Q0 + λhkf ki5Q0 + λhkf ki5Qj ],
where we used the bounds for MT# (15Q0 f ) and T (15Q0 f ) on {F (Q0 ) that
follow directly from the definition of these sets, and the analogous bound for
T (15Qj f ) on {F (Qj ). Hence
X X
1Qj T (15Q0 \5Qj f ) 6 1Sj Qj 2λhkf ki5Q0 + 1Qj λhkf ki5Qj ,
j j
and together with (11.31), (11.34) and the choice of λ, this completes the
proof of the lemma.
Iterating the previous lemma, we obtain:
Lemma 11.3.8. Under the assumptions of Theorem 11.3.6, for any cube Q0
1
and f ∈ Lloc (Rd ; X) and ε ∈ (0, 1), there is a (1 − ε)-sparse subcollection
S (Q0 ) ⊆ D(Q0 ) such that, almost everywhere,
8 · 10d cT X
Z
1Q0 T (15Q0 f ) 6 1S − kf k.
ε 5S
S∈S (Q0 )
and cd = 4 · 10d Applying the same estimate to each Q1j in place of Q0 , and
continuing by induction, almost everywhere we obtain
Z N −1 X Z
cd cT X
1Q0 T (15Q0 f ) 6 1Q0 − kf k + 2 1Qnj − kf k
ε 5Q0 n=1 j 5Qnj
Z (11.35)
X X
+ 1Q N
k
− kf k + 1QN k
T (15QN
k
f ),
k 5Qn
k k
In particular,
11.3 Calderón–Zygmund operators and sparse bounds 49
X X
|Qnj | 6 ε |Qn−1
i | 6 . . . 6 εn |Q0 |,
j i
so that the support of the last term in (11.35) becomes negligible in the limit
N → ∞. Thus, almost everywhere, we have
∞ Z
cd cT X X
1Q0 T (15Q0 f ) 6 2 1Qnj − kf k, (11.36)
ε n=0 j 5Qn
j
have measure |Ejn | > (1 − ε)|Qnj |. In other words, the cubes Qnj form a (1 − ε)-
sparse subcollection S (Q0 ) ⊆ D(Q0 ), and (11.36) is precisely the estimate
asserted in the lemma.
In order to pass from the local Lemma 11.3.8 to the global Theorem 11.3.6,
we use:
Proof. Consider all dyadic cubes Q ∈ D with the property that E 6⊆ 2Q.
Clearly all cubes with diam(Q) < 21 diam(E) will satisfy this condition. On the
other hand, every cube Q ∈ D is contained in some Q e ∈ D such that E ⊆ 2Q:e
if we fix some x ∈ Q and then r > 0 large enough so that E ⊆ B(x, r), then
e ⊇ Q with `(Q)
it suffices to take Q e > 2r, since then 2Q e ⊇ B(x, 1 `(Q))
e ⊇ E.
2
Let Q be the collection of maximal dyadic cubes with the property that
E 6⊆ 2Q. Maximality implies disjointness, and from what we just checked, it
follows that every x ∈ Rd is contained in some Q ∈ Q, so these cubes form a
partition of Rd .
Since Q is maximal, its dyadic parent Q b satisfies E ⊆ 2Q.b It remains to
observe that 2Q b ⊆ 5Q to complete the proof.
We now return to:
Proof of Theorem 11.3.6. If f ≡ 0, there is nothing to prove, so fix a non-
zero, compactly supported f ∈ L1c (Rd ; X). Thus E = supp f satisfies 0 <
diam(E) < ∞ as required to apply Lemma 11.3.9. This lemma produces a
partition Q ⊆ D of Rd such that supp f ⊆ 5Q, and thus 15Q f = f , for every
Q ∈ Q. This means that
X X
Tf = 1Q T f = 1Q T (15Q f ).
Q∈Q Q∈Q
50 11 Singular integral operators
Now Lemma 11.3.8 applies to each term on the right, producing (1 − ε)-sparse
subcollections S (Q) ⊆ D(Q) for each Q ∈ Q, and
X cd cT X Z Z
X cd cT X
1Q T (15Q f ) 6 1S − kf k = 1S − kf k,
ε 5S ε 5S
Q∈Q Q∈Q S∈S (Q) S∈S
With Theorem 11.3.6 at our disposal, the following notion should not appear
too alien to the reader:
|E(S)|
Z XZ Z X
AS f · g = − f · − g · |S| 6 inf MD f · inf MD g ·
S S γ
S∈S S S S∈S
Z
1 1 1
6 MD f · MD g 6 kMD f kp · kMD gkp0 6 p0 kf kp · pkgkp0 .
γ γ γ
This shows that kAS kp→p 6 γ −1 pp0 , where γ is the sparseness parameter;
since AS is manifestly positive, it suffices to consider positive functions above,
and the same bound persists for vector-valued functions.
Looking back at the statement of Theorem 11.3.6, it almost says that
T f 6 c · AS kf k under the assumptions of the theorem, but the presence of
the expanded cubes 5S prevents this from being strictly true in the stated
form. While the variant of a sparse operator implicitly appearing in Theorem
11.3.6 would be almost as good as AS for many purposes, the use of the more
symmetric (indeed, self-dual) operators AS as in Definition 11.3.10 is often
preferred.
11.3 Calderón–Zygmund operators and sparse bounds 51
We will need the case N = 5 of the following statement, but we record the gen-
eral formulation for convenience of reference, as the case N = 3 also features
in various applications.
Proposition 11.3.11 (Dilated dyadic cubes). Let N ∈ Z+ be odd. Then
the collection of N -fold concentric dilations {N Q : Q ∈ D(Rd )} can be par-
titioned into N d subcollections D n;N , n ∈ ZNd
, each of which has the same
covering and nestedness properties as D, namely,
[ n;N
D n;N = Dj ,
j∈Z
{N I : I ∈ Dj } = {N 2−j ([0, 1) + m) : m ∈ Z}
= {2−j ([−N 0 , N 0 + 1) + m) : m ∈ Z} N := 2N 0 + 1
= {2−j ([0, N ) + m − N 0 ) : m ∈ Z}
= {2−j ([0, N ) + m) : m ∈ Z}
n m o
= N 2−j [0, 1) + :m∈Z .
N
The sought-after partition of this collection is now achieved as follows: For
each n ∈ ZN = {0, 1, . . . , N − 1} and j ∈ Z, we define
52 11 Singular integral operators
n α(n, j) o
Djn;N := N 2−j [0, 1) + k + :k∈Z (11.37)
N
for appropriate α(n, j) ∈ ZN to be shortly determined. It is clear that each
Djn;N satisfies (1) from the statement of the Proposition, no matter how we
choose α(n, j). To ensure (2), it suffices to check that the left (or equivalently
right) half of any I ∈ Djn;N belongs to Dj+1
n;N
. For a generic I as written above,
the left half will be
α(n, j) 2α(n, j)
N 2−j [0, 21 ) + k + = N 2−j−1 [0, 1) + 2k + .
N N
n;N
For this to be in Dj+1 , it is necessary and sufficient that
where the negative powers are interpreted in the sense of the multiplicative
inverse mod N .
For each j ∈ Z, the map n 7→ 2j n mod N is a bijection on ZN , and thus
N
[ n a o
Djn;N = N 2−j [0, 1) + k + : k ∈ Z, a ∈ ZN
n=0
N
n m o
= N 2−j [0, 1) + : m ∈ Z = {N I : I ∈ Dj },
N
so indeed {N I : I ∈ D} is a disjoint union of the collections D n;N , n ∈ ZN ,
and we already checked that each D n;N has the properties (1) and (2).
Remark 11.3.12 (Shifted dyadic cubes). The cube families D n;N constructed
above are close relatives of the shifted dyadic cubes of Definition 3.2.25, and
they satisfy a variant of the Covering Lemma 3.2.26:
Given an odd N ∈ Z+ , for every cube Q ⊆ Rd , there exist a vector
n ∈ ZdN and a cube D ∈ D n;N such that
N 2N
`(Q) < `(D) 6 `(Q) and Q ⊆ D. (11.40)
N −1 N −1
In fact, let R ∈ D be a cube of side-length `(R) ∈ (`(Q)/(2N 0 ), `(Q)/N 0 ] that
contains the centre zQ of Q, where N = 2N 0 + 1 as before. Then D = N R ∈
D n;N for some n ∈ ZdN , and D contains the cube of side-length 2N 0 `(R) >
`(Q) centred at zQ ; thus D ⊇ Q, and `(D) = N `(R) lies exactly in the range
asserted in (11.40).
11.3 Calderón–Zygmund operators and sparse bounds 53
Also note that both the partition and refinement properties (1) and (2) of
Proposition 11.3.11 of each D n;N , as well as the covering property of every
cube Q ⊆ Rd by a cube in some D n;N , remain invariant if we drop the algebraic
dilation factor N in (11.37), so as to be back to cubes of side-length 2−j .
When N = 3, this reproduces precisely the shifted dyadic cubes of Definition
3.2.25; since 2 ≡ −1 mod 3, (11.39) reduces in this case to the simpler form
α(n, j) = (−1)j n, where reference to modular arithmetic can be avoided.
%
It is now easy to show that the sparse operators with a dilation, AS , may
always be dominated by a finite number of the simple sparse operators AS n .
It is technically convenient to take an odd integer N for the dilation factor.
This causes little loss of generality since, choosing N > %, we can always
dominate Z N dZ
− f6 − f
%Q % NQ
We can now reformulate Theorem 11.3.6 in terms of sparse operators:
Suppose that both T and MT are bounded from L1 (Rd ; X) to L1,∞ (Rd ). Then
for every boundedly supported f ∈ L1 (Rd ), there is a 5−1 -sparse collection
S ⊆ D and, for every n ∈ Zd5 , a 5−d−1 -sparse collection S n ⊆ D n;5 of the
dyadic systems as in Proposition 11.3.11, such that almost everywhere
X
5
T f 6 10d+1 cT AS kf k 6 10d+1 cT AS n kf k,
n∈Zd
5
8 · 10d · cT X
Z
Tf 6 1S − kf k = 10d+1 cT A5S kf k.
4/5 5S
S∈S
This is the first claim, and the second one follows from Lemma 11.3.13.
The goal of this section is to specialise the abstract Theorem 11.3.14 to the
case of Calderón–Zygmund operators in the following form:
where
cd,T 6 cd kT kLp0 (Rd ;X)→Lp0 ,∞ (Rd ;Y ) + cK + kωkDini
with cK and ω as in Definition 11.3.1.
The result remains true if Rd is systematically replaced by a cube or a quad-
rant of Rd , both in the function spaces where the boundedness is considered,
and in the definition of the kernel bounds cK and kωK kDini .
The norm estimate of the latter is the content of the following lemma.
Lemma 11.3.16. Let X and Y be Banach spaces, p0 ∈ [1, ∞], and let T be an
operator with a Dini kernel K : Ṙ2d → L (X, Y ). Then the maximal operator
satisfies
MT# f (x) 6 cd (cK + kωK kDini )M f (x)
and
kMT# kL1 (Rd ;X)→L1,∞ (Rd ) 6 cd (cK + kωK kDini ).
The result remains true if Rd is systematically replaced by a cube Q0 ⊆ Rd
or a quadrant S ⊆ Rd , both in the function spaces where the boundedness is
considered, and in the definition of the kernel bounds cK and kωK kDini .
where
56 11 Singular integral operators
Z
[K(xj , y) − K(x, y)]f (y) dy
{(5Q)
Z
6 √ k[K(xj , y) − K(x, y)]f (y)k dy
{B(x,4 d`(Q))
Z
+ √ k[K(xj , y) − K(x, y)]f (y)k dy =: I + II
B(x,4 d`(Q))\(5Q)
√
where, observing that |xj − x| < d`(Q) 6 14 |x − y| for x, xj ∈ Q and
√
y ∈ {B(x, 4 d`(Q)),
Z |x − x| 1
1 j
I6 √ ω K kf (y)k dy
{B(x,4 d`(Q)) |x − y| |x − y|d
X∞ Z √d`(Q) kf (y)k dy
1
6 √ √ ωK √ √
k=2 2
k d`(Q)6|y−x|<2k+1 d`(Q) 2k d`(Q) (2k d`(Q))d
X∞ Z
6 ωK1
(2−k )cd − √ kf (y)k dy
k=2 B(x,2k+1 d`(Q))
∞
X
6 cd M f (x) 1
ωK (2−k ) 6 cd M f (x)kωK
1
kDini ,
k=2
by Lemma 11.3.3 in the last step. On the other hand, since |xj − y|, |x − y| >
2`(Q) for x, xj ∈ Q and y ∈ / 5Q, we obtain
Z
2
II 6 √ cK d
kf (y)k dy
B(x,2 d`(Q))\(5Q) (2`(Q))
Z
6 cK cd − √ kf (y)k dy 6 cK cd M f (x).
B(x,2 d`(Q))
These bounds give the pointwise estimate for MT# f (x), and the norm estimate
is then immediate from the corresponding bound for the Hardy–Littlewood
maximal operator M .
The case of a cube or a quadrant in place of Rd follows by inspection of
the same argument: if all variables under consideration are restricted like this,
it is evident that only the corresponding restrictions of the kernel conditions
will be needed to make the estimates.
We are now ready to provide the main application of the sparse domination
of Calderón–Zygmund operators: their weighted norm inequalities with an
optimal dependence of the weight. A function w ∈ L1loc (Rd ) is called a weight
if w(x) ∈ (0, ∞) almost everywhere. We recall from Appendix J the following
definition, which we now extend to the local situation as well:
11.3 Calderón–Zygmund operators and sparse bounds 57
Equating the weights inside the operator and on the right hand side, we want
to arrange that σ = σ p w, i.e., that σ = w−1/(p−1) ; this is called the (Lp -)dual
weight of w. With this choice, the previous display reduces to
Thus all three formulations (11.41), (11.42) and (11.43) are equivalent.
We now give the A2 theorem for the sparse operators AS . The simplicity
of this argument is a manifestation of the usefulness of dominating other
operators by the sparse ones.
N 4 2d
kAS kL (L2 (w)) 6 N [w]A2 .
ε
58 11 Singular integral operators
Proof. By the dual weight trick (Remark 11.3.18), with σ := w−1 we need to
prove that
Z
4 2d
ANS (f σ) · gw 6 N [w]A2 kf kL2 (σ) kgkL2 (w) ∀f ∈ L2 (σ), g ∈ L2 (w).
ε
Since AS is a positive operator, both g and h may be taken to be positive,
and there are no subtle convergence issues in the computation that follows.
We first observe that
Z Z
1 σ(Q) 1
hf σiQ = fσ = f σ = hσiQ hf iσQ ,
|Q| Q |Q| σ(Q) Q
where σ(Q) = Q σ and hf iσQ is the average of f with respect to the measure
R
where
hσiN Q hwiQ 6 hσiN Q hwiN Q N d 6 [w]A2 N d .
Hence
|E(Q)|
Z X
AN d
S (f σ) · gw 6 N [w]A2 hf iσN Q hgiw
Q ,
ε
Q∈S
where
hgiw w
Q 6 inf MD g(z)
z∈Q
N 2d max(1, p−1
1
)
kAN
S kL (Lp (w)) 6 cd,p [w]Ap .
ε
Proof. This is an immediate consequence of Theorem 11.3.19 and Rubio de
Francia’s Extrapolation Theorem J.2.1. (In the latter, φpr and cpr should be
replaced by φdpr and cdpr ; the omission of dependence on d is a systematic
typo in Theorem J.2.1 and its proof. This explains a need a constant cd,p
rather than just cp in the statement of the corollary.)
It is also useful to record the following localised version:
By the dual weight trick with σ = w−1 , estimating the left-hand side uniformly
over f ∈ L2 (Q0 , σ) and g ∈ L2 (Q0 , w) of unit norm is equivalent to bounding
N
kAS kL (L2 (Q0 ,w)) .
Term II is dominated by
X Z X |Q0 |Z ∞
X Z
− (1Q0 f ) = − (1Q0 f ) = 2−kd − (1Q0 f ),
|Q| N Q0
Q)Q0 N Q Q)Q0 N Q0 k=1
P∞ P∞
where k=1 2−kd 6 k=1 2−k = 1. Thus
where
w(Q0 )1/2
Z
1Q0 h1Q0 f iN Q0 = f w1/2 σ 1/2
L2 (w) |N Q0 | Q0
w(Q0 )1/2
6 kf kL2 (Q0 ,w) σ(Q0 )1/2
|Q0 |
1/2
6 [w]A2 (Q0 ) kf kL2 (Q0 ,w) 6 [w]A2 (Q0 ) kf kL2 (Q0 ,w) .
σ(N Q ∩ Q0 ) 6 σ(Q),
e w(Q) 6 w(Q),
e e 6 |N Q| = N d |Q|.
|Q|
Hence
σ(N Q ∩ Q0 ) w(Q) σ(Q)
e w(Q)
e
6 N d 6 [w]A2 (Q0 ) N d ,
|N Q| |Q| |Q|
e |Q|e
since Q
e is a cube contained in Q0 . Substituting back, and using sparseness, it
follows that
|E(Q)|
Z X
AN d
S 0 (1Q0 f σ) · gw 6 N [w]A2 (Q0 ) hf iσN Q∩Q0 · hgiw
Q .
Q0 0
ε
Q∈S
C n;N := {N Q ∩ Q0 : Q ∈ D, N Q ∈ D n;N }
are still instances of the Doob maximal operator with respect on abstract
filtered spaces. Repeating the computation (11.44) mutatis mutandis, we then
obtain
62 11 Singular integral operators
X |E(Q)|
hf iσN Q∩Q0 hgiw
Q
ε
Q∈S 0
1 X
6 kMCσ n;N f kL2 (σ) kMD
w
gkL2 (w)
ε d
n∈ZN
1 X 4
6 2kf kL2 (σ) · 2kgkL2 (w) = N d kf kL2 (σ) kgkL2 (w) .
ε d
ε
n∈ZN
Hence
4
I 6 N d [w]A2 (Q0 ) · N d kf kL2 (σ) kgkL2 (w) .
ε
In combination with the bound
II 6 [w]A2 (Q0 ) kf kL2 (σ) kgkL2 (w) .
Recalling that
Z Z Z X
AN
S (1Q 0 f σ) · gw 6 AN
S 0 (1Q0 f σ) · gw + h1Q0 f iN Q · gw
Q0 Q0 Q0 Q)Q
0
and the dual weight trick, we conclude the proof in the case of a cube.
If Q0 is replaced by a quadrant S, we note by density that it suffices to
consider the integrals above compactly supported f and g. But then, if Q0 is a
sufficiently large cube contained in the quadrant and having one corner at the
corner of the quadrant, then such f and g will be supported in Q0 . Thus the
previous considerations apply and give a bound in terms of [w]A2 (Q0 ) , which
is clearly dominated by [w]A2 (S) .
An extension of Proposition 11.3.21 to p 6= 2 follows, in principle, by Rubio de
Francia’s Extrapolation Theorem J.2.1 just like Corollary 11.3.20 from The-
orem 11.3.19. Since Theorem J.2.1 was formulated for global Ap (Rd ) weights
only, we include some remarks about its local version. As a rule, all dyadic
considerations carry over without any change. However, one needs to play a
little attention to the interplay of dyadic and non-dyadic cubes in the local
setting. The following is a local variant of the Covering Lemma 3.2.26:
Lemma 11.3.23. For cubes Q ⊆ Q0 ⊆ Rd , there exist a vector α ∈ {0, 31 , 23 }d
and a dyadic cube
`(P )
log2
D ∈ D α (Q0 ) := {P + α(−1) `(Q0 )
`(P ) : P ∈ D(Q0 )} (11.45)
(the shifted dyadic cubes from Definition 3.2.25) such that
`(D) 6 3`(Q) and Q ⊆ D ⊆ Q0 .
`(P )
log
In (11.45), the point of the factor (−1) 2 `(Q0 ) is simply to alternate between
±1 with each consecutive generation of the dyadic cubes. We refer the reader
to the discussion preceding Lemma 3.2.26 for why such a factor is needed.
11.3 Calderón–Zygmund operators and sparse bounds 63
with
X
MQ0 f 6 3d max α
MQ 0
f 6 3d α
MQ 0
f. (11.46)
1 2
α∈{{0, 3 , 3 }d 1 2
α∈{{0, 3 , 3 }d
Proof. (1) follows by repeating the proof of Theorem J.1.1: the dyadic con-
siderations are unchanged, and in the last step of the proof, one replaces an
application of (3.36) by its localised version (11.46).
The proof of (2) is the same as the proof of Theorem J.2.1, except that the
all references to the maximal operator M are replaced by the local version
MQ0 and, accordingly, all applications of Theorem J.1.1 by case (1) of the
proposition that we already proved. (We note that the φpr and cpr should
be replaced by φdpr and cdpr already in Theorem J.2.1; the omission of the
dependence on d is a systematic typo in Theorem J.2.1 and its proof.)
Proof. Let us first consider the global case. Let f ∈ Lpc (w; X) be supported
on a compact set F . Denoting by σ = w−1/(p−1) the dual weight, we have
Z Z
0 0
kf k = kf kw1/p σ 1/p 6 kf kLp (w) σ(K)1/p < ∞,
K
Recalling the definition of cT , this is the required norm estimate for T re-
stricted to Lpc (w; X); since this subspace is dense in Lp (w; X), it allows to
uniquely extend T to the whole space with the same norm.
The proof in the case of a cube or a quadrant in place of Rd remains the
same, just using the local Corollary 11.3.25 in place of Corollary 11.3.20 to
replace (11.47) by
in Theorem 11.3.26, we have N0 6 ~2,X < ∞, using the notation from the
statement of that theorem. The kernel of the Hilbert transform is K(s, t) =
1
s−t , so that cK = 1 qualifies for the constant in Definition 11.3.1. Moreover,
1 1 s0 − s |s0 − t| 1
− 0 = 0
62 ∀|s − s0 | 6 |s − t|,
s−t s −t (s − t)(s − t) |s − t|2 2
Example 11.3.28 (Power weights). Let α ∈ R, p ∈ (1, ∞), w(x) = |x|α for
x ∈ Rd , and σ(x) = w(x)−1/(p−1) = |x|−α/(p−1) . Then
`(Q)−γ
Z Z
− |x| dx 6 − |x|−γ dx hd
−γ
, γ ∈ [0, d),
Q Q̃ d−γ
Z Z
− |x|γ dx > − |x|γ dx hd,Γ `(Q)γ , γ ∈ [0, Γ ], Γ > 0.
Q Q̃
Qd Qd
Proof. Let Q = i=1 Ii and Q̃ = i=1 I˜d . Then Q is the disjoint union of the
sets QI := i∈I (Ii ∩ I˜i ) × i∈{I (Ij \ I˜j ), where I ranges over all subsets
Q Q
11.3 Calderón–Zygmund operators and sparse bounds 67
√
Z Z p−1 √ −α
d α
sup − w − σ 6 sup (δQ + d)α δQ = 1+
Q:δQ >δ Q Q Q:δQ >δ δ
On the other hand, for any cube Q, it follows from Lemma 11.3.29 that
Z Z p−1 √ 2d d α
− p−1
p−1
sup w − σ 6 sup (δQ + d)α `(Q)α α `(Q)
Q:δQ 6δ Q Q Q:δQ 6δ d − p−1
√ 2d d p−1
= (δ + d)α α
d − p−1
68 11 Singular integral operators
cd,p c0d,p
[w]Ap 6 α p−1 =
(d − p−1 ) [d(p − 1) − α]p−1
For a matching lower bound, it is enough to consider just the unit cube Q, in
which case the estimates of Lemma 11.3.29 apply with Γ = d(p − 1) to give
that
Z Z p−1 p−1 p−1
1 1
[w]Ap > − w − σ hd,p 1 · α h d,p .
Q Q d − p−1 d(p − 1) − α
1
This completes the proof for α ∈ [0, d(p − 1)), noting that 1+α hd,p 1 in this
case.
For α = −γ < 0, we note that
γ
n 1 p0 −1 op−1
[|x|−γ ]Ap = [|x| p−1 ]p−1
Ap0 hd,p 0 γ
d(p − 1) − p−1
p−1 1
= hd,p
d−γ d+α
γ
by applying the previous case to p−1 > 0 and p0 in place of α and p, and
noting that (p − 1)(p0 − 1) = 1.
We are now fully equipped to confirm the sharpness of Corollary 11.3.27.
Proof. Let σ = w−1/(p−1) denote the dual weight. Using the dualised formu-
lation (11.43) of the Lp (w)-boundedness of T = H, and choosing f and g
with positively separated compact supports, so that the kernel representation
is available, we have
ZZ
1 f (y)σ(y)g(x)w(x)
dx dy 6 φ([w]Ap )kf kLp (σ) kgkLp0 (w) (11.48)
π x−y
for all such f and g. If these functions are non-negative with supp f ⊆ R−
and supp g ⊆ R+ , then the integrand is non-negative, and by monotone con-
vergence (11.48) persists even if the supports of f and g meet at the origin.
The crucial point in bounding the Hilbert transform form below is the
following observation: if h(y) = |y|−α 1(−1,0) (y), then for x ∈ (0, 1),
11.3 Calderón–Zygmund operators and sparse bounds 69
1 x
y −α y −α 1 x−α
Z Z
1 1
Hh(x) = dy > dy = , (11.49)
π 0 x+y π 0 2x 2π 1 − α
1
which is essentially h again, but with a factor 1−α that blows up as α → 1−.
We now “test” (11.48) with two choices of (f, g, σ, w), so that (f σ, gw) is
either (|y|−α 1(−1,0) , 1(0,1) ) or (1(−1,0) , |y|−α 1(0,1) ), with α ∈ [0, 1). In either
case (11.49) shows that
1
x−α
Z
1 1 1
LHS(11.48) > dx = ,
2π 0 1−α 2π (1 − α)2
1 1 cp 1
6 (11.48) 6 φ . (11.51)
2π (1 − α)2 (1 − α)p−1 1 − α
Since H 2 = −I, it is clear that kHkL (Lp (w)) > 1, and hence φ(t) > 1 >
c0p t1/(p−1) for t ∈ [1, cp ) as well.
In the second case, we take g = 1(0,1) and w(x) = σ(x)1−p = |x|−α ; thus
0 0
σ(x) = |x|α/(p−1) = |x|α(p −1) and f (x) = 1(−1,0) (x)|x|−α(p −1) . A computa-
tion like (11.50) gives exactly the same final result, only with a slightly differ-
ent intermediate step, and Example 11.3.28 shows that [w]Ap 6 cp /(1 − α).
With this quantity inside φ in (11.51), the substitution t = cp /(1 − α) then
gives
φ(t) > c̃p t ∀t > cp , (11.53)
and the same bound for t ∈ [1, cp ) follows from H 2 = −I as before. The two
lower bounds (11.52) and (11.53) together prove the proposition.
70 11 Singular integral operators
11.4 Notes
Given the emphasis of these volumes in analysis of functions having their
range in a Banach space, we have chosen to keep the consideration related to
the domain of the functions relatively simple, concentrating on the canonical
case of the Euclidean space Rd and, with specific applications in the later
chapters in mind, its rather special subdomains—cubes and quadrants—only.
However, much of this theory could be developed on far more general do-
mains, notably on spaces of homogeneous type (espaces de nature homogène)
introduced by Coifman and Weiss [1971] and extensively studied ever since.
Since our treatment is heavily based on the dyadic cubes on Rd , we recall
that analogous constructions are also available in the mentioned generality.
The construction of a fixed family of sets, sharing the essential properties of
the standard dyadic cubes of Rd , is due to Christ [1990]. We also make use
of “adjacent” and “random” families of dyadic cubes; a reasonably compre-
hensive account of their analogues in spaces of homogeneous type is provided
by Hytönen and Kairema [2012] with several variants and elaborations due to
Auscher and Hytönen [2013], Hytönen and Martikainen [2012], Hytönen and
Tapiola [2014], and Nazarov, Reznikov, and Volberg [2013].
Section 11.1
This section deals with relatively classical topics but with some modern
flavour. In particular, the local oscillation decomposition of Theorem 11.1.12
dates essentially back to Lerner [2010] in the scalar-valued case. The vector-
valued generalisation, introducing the notion of λ-pseudomedian, was first
found by Hänninen and Hytönen [2014]. Our present proof streamlines the
original one.
Proposition 11.1.14 was proved by Katz and Pereyra [1999] in the scalar-
valued case via a multilinear estimate, and by Hänninen and Hytönen [2016]
as stated.
Theorem 11.1.30 on the vector-valued H 1 –BMO duality is essentially from
Bourgain [1986], although the present proof is different. In this circle of ideas,
we have only covered the relatively elementary part of the theory that does
not require any assumptions on the underlying Banach space. Note that The-
orem 11.1.30 says that BMOD (Rd ; X ∗ ) can be identified with an isometric
1
subspace of (HD,at (Rd ; X))∗ . The same proof works in the non-dyadic case,
where arbitrary cubes are allowed both in the definition of BMO and of the
Hardy space atoms. To describe the full dual (Hat 1
(Rd ; X))∗ , Blasco [1988] de-
fines a class of Banach space Y -valued measures BM O(Rd ; Y ). Among other
things, he shows that (Hat1
(Rd ; X))∗ = BM O(Rd ; X ∗ ) for every Banach space
X, whereas BM O(R ; Y ) = BMO(Rd ; Y ), if and only if Y has the Radon–
d
Section 11.2
The material of this section is predominantly classical, and most of the results
would have been available in essentially the present form by the 1980’s, if not
earlier, even in the Banach space valued setting. The scalar-valued origins, of
course, date much further back.
The essence of Theorem 11.2.5 comes from Calderón and Zygmund [1952],
who consider the scalar-valued case (X = Y = L (X, Y ) = C) and Dini kernels
x−y
of the special form K(x, y) = K(x − y) = |x − y|−d Ω |x−y| , where moreover
Ω dσ = 0. In contrast to Theorem 11.2.5, which extrapolates other Lp -
R
S n−1
bounds from an assumed a priori Lp0 -bound, Calderón and Zygmund [1952]
obtained their Lp -boundedness conclusions unconditionally, i.e., they also de-
duce the initial Lp0 -bound for p0 = 2 from their special assumptions on the
kernel. Once this is achieved, the extrapolation to other Lp -bounds is carried
out in much the same way as in the present treatment, particularly in the case
p < p0 . The fact that the extrapolation part of Calderón and Zygmund [1952]
argument remains valid under more general assumptions on the kernel was
observed by Hörmander [1960], who introduced the conditions, now bearing
his name, in Definition 11.2.1 in the case of scalar-valued convolution kernels
K(x, y) = K(x − y). What we have called the (operator-)Hörmander class
Hör was designated as K 1 by Hörmander [1960], who also defines a family of
related conditions K a with a parameter a ∈ [1, ∞]. Just like Hör = K 1 is rel-
evant for the extrapolation of Lp -boundedness, the condition K a permits the
extrapolation of Lp -to-Lq boundedness from one pair (p, q) with p1 − 1q = 1− a1
to other such pairs.
The first Banach space-valued generalisations, which used the operator-
Hörmander conditions, were found by Schwartz [1961] and, apparently inde-
pendently, by Benedek, Calderón, and Panzone [1962]. According to Garcı́a-
Cuerva and Rubio de Francia [1985], the fact that the mere Hörmander con-
dition (involving integrals of kK(s, t)x − K(s0 , t)xkY rather than kK(s, t) −
K(s0 , t)kL (X,Y ) ) is sufficient for results like Theorem 11.2.5 “should have been
observed by anyone trying to adapt the proof of [the Calderón–Zygmund the-
orem] to the vector valued case”, yet they “do not emphasize very much the
interest of this weaker condition since, in most of the applications of vec-
tor valued singular integrals, [the operator Hörmander condition] does hold.”
Rubio de Francia, Ruiz, and Torrea [1986] provided, in their own words, an
“updated review” of Benedek et al. [1962], incorporating several new devel-
opments in singular integrals into the vector-valued theory, and in particular
explicitly dealing with two-variable kernels K(s, t), as we have done here. Our
considerations related to c0 in Theorem 11.2.9 were inspired by Girardi and
Weis [2004].
A version of Theorem 11.2.5 for convolution kernels K(s, t) = K(s − t) is
also presented by Grafakos [2008], where (in contrast to our approach) the
upper extrapolation is achieved by a duality argument, and the interested
reader is referred to this work for details of that approach. Grafakos [2008] is
72 11 Singular integral operators
also explicit about the norm estimate in Theorem 11.2.5(3); this is certainly
well known, but often not spelled out in many references.
Section 11.3
The main body of this section consists of results from the 2010’s. Since the dis-
covery of the original forms of many of these results, there has been significant
activity in generalising and streamlining their proofs, as well as developing en-
tirely new approaches. As a result, our order of presentation deviates from the
historical timeline in favour of a smoother mathematical story. A main result
of this section is certainly the A2 Theorem 11.3.26, but the various Sparse
Domination Theorems 11.3.6, 11.3.14, and 11.3.15, originally developed as
tools for proving the A2 Theorem 11.3.26, have by now established them-
selves as results of intrinsic value and models for desirable type of domination
to search for in other situations.
Prehistory of the A2 theorem
In its scalar-valued and qualitative form (i.e., saying that T is bounded on
Lp (w), but without tracking the estimate for the operator norm), the result
goes back to Hunt, Muckenhoupt, and Wheeden [1973] in the special case
that T is the Hilbert transform (as in Corollary 11.3.27) and to Coifman and
Fefferman [1974] for all standard Calderón–Zygmund operators of convolu-
tion type. The question of sharp dependence of the weighted operator norms
kT kL (Lp (w)) on the weights constant [w]Ap was raised by Buckley [1993],
who settled the case of the Hardy–Littlewood maximal operator (Theorem
J.1.1) and obtained non-matching upper and lower bounds for Calderón–
Zygmund operators. In particular, Proposition 11.3.30 saying that an esti-
1
max(1, p−1 )
mate for kT kL (Lp (w)) can be no better than [w]Ap , is essentially from
Buckley [1993]. In many papers, results of this type a stated in a slightly
weaker form along the lines that “the power of [w]Ap can be no better than
1
max(1, p−1 )”. However, in some related questions, the sharp estimate is known
to exhibit behaviour different from a pure power law.
The question of Buckley [1993] gained new interest through the work of
Astala, Iwaniec, and Saksman [2001], who considered the following problem:
Let O ⊆ C be a domain and k ∈ (0, 1). What is the minimal q such that all
1,q ¯ | 6 k|∂f | (referred to as weakly quasiregular)
functions f ∈ Wloc (O) with |∂f
1,2
must in fact belong to f ∈ Wloc (O) (and then be called simply quasiregular)?
By results of Astala [1994], q > 1 + k suffices; by examples due to Iwaniec
and Martin [1993], q < 1 + k does not, leaving q = 1 + k as the critical case.
Astala, Iwaniec, and Saksman [2001] proved that q = 1+k is still sufficient for
the said self-improvement, under their conjecture that the Beurling–Ahlfors
transform
Z
1 f (y) dA(y)
Bf (z) := − lim , D(z, ε) := {y ∈ C : |y − z| < ε}
π ε→0 C\D(z,ε) (z − y)2
11.4 Notes 73
Shortly after being posed, the conjecture of Astala et al. [2001] was verified by
Petermichl and Volberg [2002], and another proof was found by Dragičević and
Volberg [2003]. Already Petermichl and Volberg [2002] observed that (11.54)
as stated may be derived from its special case p = 2 by keeping track of
the constants in the proof of Rubio de Francia’s extrapolation theorem as
presented, e.g., by Garcı́a-Cuerva and Rubio de Francia [1985]. This idea was
systematised by Dragičević, Grafakos, Pereyra, and Petermichl [2005], whose
results were treated in Appendix J and applied in the section under discussion.
The positive results for the Beurling–Ahlfors transform inspired the ques-
tion of sharp weighted bounds for other operators, and the special role of the
exponent p = 2 as the critical case for extrapolation gave rise to the name
“A2 conjecture”, several further cases of which were settled over the next few
years. In particular, the Hilbert transform (the scalar-valued case of Corollary
11.3.27) and the Riesz transforms were handled by Petermichl [2007, 2008], a
general class of sufficiently smooth odd kernels on R by Vagharshakyan [2010],
and powers of the Beurling–Ahlfors operator by Dragičević [2011]. All these
results relied on
(A) ad hoc representation formulas of special singular integrals in terms of
simple “dyadic shifts” as in the representation of Petermichl [2000] for
the Hilbert transform (see Theorem 5.1.13 and (5.20)), and
(B) Bellman function techniques for sharp weighted bounds of these shifts.
The component (B) behind these results was first challenged by Lacey, Peter-
michl, and Reguera [2010], who replaced it with
(C) “corona decompositions” to verify the “testing conditions” in a
(D) dyadic two-weight T (1) theorem of Nazarov, Treil, and Volberg [2008].
Shortly after, a much simpler alternative to either (B) or (C)–(D) was found
by Cruz-Uribe, Martell, and Pérez [2010], who in turn replaced it by methods
largely similar to the ones that we have used here:
(E) domination of dyadic shifts from (A) (not yet of singular integrals di-
rectly) by the sparse operators AS , and
(F) estimating kAS kL (L2 (w)) as in Theorem 11.3.19, whose proof follows
closely the original one from Cruz-Uribe et al. [2010],
However, component (A) of the original proofs remained unchallenged and,
being somewhat ad hoc for the specific singular integrals considered thus far,
restricted their extension to wider classes of operators.
74 11 Singular integral operators
However, these results were shortly superseded by Hytönen and Lacey [2012]
(arXiv 6/2011) by a new approach combining (G) with elaborations of (E)
and (F) from the approach of Cruz-Uribe et al. [2010]:
(E0 ) domination of the general dyadic shifts from (G) by operators (essentially
like) AN
S , where arbitrarily large N appear, and
0 N
(F ) estimating kAS kL (L2 (w)) with bounds polynomial in log N (which re-
quires much more delicate analysis than Theorem 11.3.19).
11.4 Notes 75
As a curiosity, the term “sparse” in its present usage seems to have been
introduced by Hytönen and Lacey [2012] (line below (∗) on page 2042). This
was pointed out by Andrei Lerner in his survey talk at the “AIM Workshop on
sparse domination of singular integrals” in San José, California, in 10/2017.
Simpler proofs
The difficulties with arbitrarily high shift complexity N , which seemed un-
avoidable in the general A2 theorem until this point, were finally eliminated
by Lerner [2013a,b] (arXiv 2/2012). These papers provide two different proofs
of the same main result, stating that
Although not a necessity for proving the A2 theorem, the possibility of replac-
ing (11.56) by pointwise domination presented itself as a natural question,
which attracted some interest. This was independently achieved by Conde-
Alonso and Rey [2016] (arXiv 9/2014) and Lerner and Nazarov [2019] (also
announced and circulated around the same time in 2014, although in arXiv
only in 8/2015). These results still slightly deviated from Theorem 11.3.15 by
requiring a stronger form of the Dini condition,
Z 1/2
1 dt
ω(t) log2 < ∞.
0 t t
All Dini kernels were first covered by the “elementary” (but not so easy)
proof of Lacey [2017] (arXiv 1/2015), which was further quantified (in terms
of dependence on kωkDini ) by Hytönen, Roncal, and Tapiola [2017] (arXiv
10/2015) and remarkably simplified again by Lerner [2016] (arXiv 12/2015).
In proving Theorem 11.3.15, we have followed the further simplification due
to Lerner and Ombrosi [2020]. One advantage of their approach is a reduc-
tion of the prerequisites from classical Calderón–Zygmund theory necessary
to run their argument. On the technical level, this is achieved by replacing
the maximal operator
of Lerner [2016] by its “sharp” version MT# defined in (11.28). While MT#
can be estimated relatively directly, bounding the larger MT f originally re-
quired non-trivial classical results about the maximal truncations (11.55).
However, it was later observed by Almeida, Betancor, Fariña, and Rodrı́guez-
Mesa [2022] that the bounds for the two operators are actually equivalent
under general assumptions only involving the bounds for T that are used in
the theory anyway. Although not explicitly discussed by Lerner and Ombrosi
[2020], the present vector-valued extensions of their results, leading to Theo-
rems 11.3.15 and 11.3.26, involved little additional effort; this is in contrast to
the first vector-valued A2 theorem by Hänninen and Hytönen [2014]. Further
abstractions are due to Lorist [2021] and Lerner, Lorist, and Ombrosi [2022];
the latter work also explicitly addresses the vector-valued case.
There are some alternative routes to see the sharpness result of Proposition
11.3.30, which goes back to Buckley [1993] well before the matching upper
bounds were known. Luque, Pérez, and Rela [2015] made the curious observa-
tion that this can also be achieved without exhibiting any explicit examples
in the weighted situation, but studying instead the asymptotics of the un-
weighted norms kT kLp →Lp as p → 1 and p → ∞. This depends on a variant
11.4 Notes 77
Further results
For a while, it might have seemed that the new sharp weighted technology
was essentially restricted to the class of Calderón–Zygmund operators. A cer-
tain discouragement against further extensions came from an observation of
Orponen [2013] that if an operator T has a dyadic representation (G) in the
sense of Hytönen [2012], then T must necessarily be a Calderón–Zygmund op-
erator. However, as soon as the role of (G) in the A2 theorem was challenged
by other methods, the door was also open for extensions beyond the standard
Calderón–Zygmund realm. Nevertheless, few could probably have expected
how far this theory could indeed be extended.
As an application of the sharp weighted estimates for Dini kernels discussed
above, Hytönen, Roncal, and Tapiola [2017] (arXiv 10/2015) showed that
rough homogeneous singular integrals
Z
Ω(t/|t|)
TΩ f (s) := p. v. f (s − t) dt, Ω ∈ L∞
0 (S
d−1
).
Rd |t|d
with φ(u) 6 u2 . Although dealing with a class of operators outside the direct
scope of the sparse domination technology of the time, this result may never-
theless be seen as stretching those methods, rather than introducing genuinely
new ones, in that the operator TΩ was decomposed into a series of pieces in
the scope of the previously available tools by following a classical approach to
qualitative versions of similar results by Duoandikoetxea and Rubio de Francia
78 11 Singular integral operators
[1986], and Watson [1990]. A more intrinsic approach has been subsequently
developed by Conde-Alonso, Culiuc, Di Plinio, and Ou [2017], but φ(u) 6 u2
seems to remain the best available bound at the time of writing. In the other
direction, Honzı́k [2023] constructed examples of symbols Ω and weights w
to show that φ(u) > u3/2 ; hence the quantitative behaviour of TΩ is defi-
nitely different from the linear A2 theorem for standard Calderón–Zygmund
operators, but their precise bounds remain open.
Already a few weeks before Hytönen, Roncal, and Tapiola [2017] (late
10/2015 in arXiv), a far-reaching approach to sparse domination of a wide class
of operators had been revealed by Bernicot, Frey, and Petermichl [2016] (early
10/2015 in arXiv). They observed that several operators that act boundedly
in Lp only in some range (p0 , q0 ) ( (1, ∞) (and thus are definitely outside the
Calderón–Zygmund class by Theorem 11.2.5) can be proved to possess sparse
form domination of the type
X Z 1/p0 Z 0
1/q00
|hT f, gi| 6 C |Q| − |f |p0 − |g|q0 .
Q∈S 5Q 5Q
and its dyadic version, where the rectangles are restricted to be dyadic (i.e.,
products of dyadic intervals). Barron, Conde-Alonso, Ou, and Rey [2019] have
shown that it is impossible to dominate the strong dyadic maximal operator
by sparse forms based on rectangles with sides parallel to the axes, which
presents an obstacle to sparse techniques in this setting. While the most obvi-
ous extension of sparse domination is thus excluded, it was shown by Barron
and Pipher [2017] that one canR still obtain a workable substitute
R by replac-
ing the dominating averages −R |f | of f with the averages −R Sf of its dyadic
square function Sf on the right-hand side.
On the other hand, the original dyadic representation (G), while largely su-
perseded by sparse technology in applications to standard Calderón–Zygmund
operators, remains available, after natural modifications, in the product space
theory, as first proved by Martikainen [2012b] in the two-parameter case and
extended to arbitrarily many parameters by Ou [2017]. A vector-valued ap-
proach to this theory has been developed by Hytönen, Martikainen, and Vuori-
nen [2019a].
but the optimal bound for the operator norm kT kL (Lp (w)) in terms of [w]A−
p
remains open. In analogy with the A2 Theorem 11.3.26, it is natural to make:
Partial results for Haar multipliers (see Section 12.1.a) in place of Calderón–
Zygmund operators are obtained by Chen et al. [2020], but beyond that the
conjecture remains open.
Causal operators appear very naturally; e.g., the operator-valued kernel
where | | is (say) the operator norm on L (RN ) (but the choice of the norm
on RN ×N is irrelevant, as they are all equivalent).
With the natural definition
1
kf kLp (W ) := kW p f kLp (Rd ;RN ) ,
is a norm, and hence, by the theorem of John [1948], there is a positive definite
reducing operator [V ]Q,p ∈ RN ×N , such that
Z 1/p √
|[V ]Q,p e| 6 − |V (t)e|p dt 6 N · |[V ]Q,p e|.
Q
The matrix-Ap condition may then be defined by the finiteness of the constant
1
[W ]Ap := sup |[V ]Q,p [V −1 ]Q,p0 |p , V := W p .
Q
1
The reader is invited to check that [V ]Q,p = hV p iQ
p
if N = 1 or p = 2 (but not
in general otherwise), so that the different definitions of Ap are consistent. It
82 11 Singular integral operators
with the important difference that we take the dot product of f (s), g(t) ∈ Rn
first, and only then the absolute value of the result; this allows for critical
directional cancellation compared to (11.57).
The proof of Nazarov, Petermichl, Treil, and Volberg [2017] (arXiv 1/2017),
that standard Calderón–Zygmund operators satisfy (11.58), follows the same
lines as the proof of Theorem 11.3.15 but with important elaborations at a
few selected points, making again use of the ellipsoid theorem of John [1948].
On the other hand, with (11.58) available, Nazarov et al. [2017] can prove the
bound
3/2
kT kL (L2 (W )) 6 cd,T [W ]A2 ,
which remains the best available matrix-weighted estimate for Calderón–
Zygmund operators (or even just for the Hilbert transform) at the time of
writing. A variant of the same results was also obtained by Culiuc, Di Plinio,
and Ou [2018b], seemingly earlier (arXiv 10/2016) but not independently; ac-
cording to their acknowledgment, the concept of domination by convex body
11.4 Notes 83
averages was introduced to these authors by Sergei Treil during his seminar
talk at Brown University in the Spring of 2016.
Since then, further applications and extensions of convex body dom-
ination have been explored by Cruz-Uribe, Isralowitz, and Moen [2018],
Di Plinio, Hytönen, and Li [2020a], Isralowitz, Pott, and Rivera-Rı́os [2021],
Isralowitz, Pott, and Treil [2022], and Muller and Rivera-Rı́os [2022]. Impor-
tantly, Bownik and Cruz-Uribe [2022] extended the Rubio de Francia algo-
rithm (Proposition J.2.2), and its key application to weighted extrapolation
(Theorem J.2.1), to matrix-valued weights, by further development of the
convex body philosophy.
An abstract framework for convex body domination has been proposed by
Hytönen [2023], allowing also Banach space valued functions in the theory.
While genuinely operator-valued weights in infinite dimensions seem to be
out of reach, this framework allows the treatment of RN ×N -valued weights
on spaces of X N -valued functions. In particular, the following simultaneous
extensions of the boundedness of the Hilbert transform on the Banach space
valued L2 (R; X) by Burkholder [1983], and on the matrix-weighted L2 (W ) by
Treil and Volberg [1997], is obtained there.
3/2 2 3/2
and satisfies kHkL (L2 (W ;X N )) 6 cN ~2,X [W ]A2 6 cN β2,X [W ]A2 , where
~2,X = kHkL2 (R;X) and β2,X is the UMD constant.
Singular integrals:
Square functions:
For the range of p as specified and w ∈ Ap , the sharp estimates in Lp (w) are:
1
max( 12 , p−1 )
(4) the strong-type bound is [w]Ap for p ∈ (1, ∞) (Lerner [2011]);
max( 1 , 1 )
(5) the weak-type bound is [w]Ap 2 p for p ∈ [1, ∞) \ {2} (p = 1: Chanillo
and Wheeden [1987], Wilson [2007, 2008]; p ∈ (1, 2): Lacey and Scurry
[2012]; p > 2: Hytönen and Li [2018]);
1 1
(6) the weak-type L2 (w) bound is at most [w]A2 2 (1 + log[w]A1 ) 2 (Domingo-
Salazar, Lacey, and Rey [2016]), but its sharpness seems to remain open
(see Ivanisvili and Volberg [2018] for partial related results).
For matrix-weights, the only known sharp estimates among these exam-
ples, at the time of writing, seem to be the square function bounds (4) for
p ∈ (1, 2]; this was proved by Hytönen, Petermichl, and Volberg [2019b] for
p = 2 and extended by Isralowitz [2020] to p ∈ (1, 2).
12
Dyadic operators and the T (1) theorem
In Chapter 11, we have mainly dealt with a situation, where a bounded linear
operator on some Lp0 (Rd ; X) space is given, and we have then explored its
bounded extensions to other spaces including Lp (Rd ; X) for p 6= p0 . We now
turn to a somewhat different (and often more difficult) question of recognising
such bounded operators to begin with.
Before addressing this question for the Calderón–Zygmund type operators
of the kind studied in Chapter 11, we investigate a number of related objects
in a simpler dyadic model. Besides serving as an introduction to some of the
key techniques, it turns out that these dyadic operators can be, and will be,
also used as building blocks of the proper singular integral operators towards
the end of the chapter.
The dyadic operators will be of two essentially different types. The first
class, which we vaguely refer to as “dyadic singular integrals” in Section 12.1,
consist of a somewhat diverse family of relatives of the prototype dyadic shifts
encountered in Chapter 5, where they we used to represent the prototype sin-
gular integral given by the Hilbert transform. It is thus only natural that a
family of dyadic operators generalising this basic dyadic shift will serve as
building block of the Calderón–Zygmund family of singular integrals gener-
alising the Hilbert transform. Martingale techniques vaguely reminiscent of
those in Section 5.1, but of somewhat higher complexity probably by neces-
sity, will feature in the argument to put the UMD property of the underlying
Banach space into action.
The second class of dyadic operators consists of so-called paraproducts,
which we discuss in Section 12.2. These are new creatures of the non-
convolution realm that we have entered and they will vanish (as we will see) as
soon as we occasionally specialise our considerations to singular integral of the
convolution form. However, for the representation the full class of Calderón–
Zygmund operators they will turn out be quite essential.
The chapter will culminate in a lengthy treatment of the so-called T (1)
theorem, a general criterion for boundedness of singular integral operators.
We will first discuss a version for abstract bilinear form in Section 12.3, and
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 87
T. Hytönen et al., Analysis in Banach Spaces, Ergebnisse der Mathematik und ihrer
Grenzgebiete. 3. Folge / A Series of Modern Surveys in Mathematics 76,
https://doi.org/10.1007/978-3-031-46598-7_2
88 12 Dyadic operators and the T (1) theorem
only then, in the final Section 12.4, turn to the task of checking the assump-
tions of the abstract result for singular integral operators with a Calderón–
Zygmund kernel, of the kind that we met Chapter 11. However, in order to
establish boundedness on Lp (Rd ; X) from scratch, rather than extrapolating
it from another Lp0 (Rd ; X) space where it was already known (as in Chapter
11), somewhat stronger versions of the Calderón–Zygmund conditions will be
needed, and the notion of R-boundedness from Chapter 8 will, once again,
play a prominent role. While the results of this chapter will generically be
established in arbitrary UMD spaces, it turns out that additional information
about type and cotype, as studied in Chapter 7 can be traded against the pre-
cise kernel conditions, so that slightly larger classes of kernels are admissible
under conditions of type and cotype of the underlying space.
It is easy to see that S(C ; X) ⊆ Lp (Rd ; X) for all p ∈ [1, ∞], and that
We begin with what is arguably the simplest class of operators deserving the
name of “dyadic singular integrals”. In essence, we have encountered these
operators already, at least implicitly on the one-dimensional domain space
R1 , where we dealt with operators of the form
X
f 7→ I hf, hI ihI
I∈D
Q = aQ + `(Q)[0, 1)d ,
we have
ch(Q) = {Q0 ∈ Dk+1 : Q0 ⊆ Q}.
For every cube Q, we define the conditional expectation and martingale
difference projections (acting on f ∈ L1loc (Rd ; X))
Z X
EQ f := 1Q − f dx, DQ f := EQ0 f − EQ f. (12.1)
Q Q0 ∈ch(Q)
1
satisfy the following identity for all f ∈ Lloc (Rd ; X):
X X
α
DQ f = hf, hQ ihα
Q =:
α
DQ f.
α∈{0,1}d \{0} α∈{0,1}d \{0}
where
X
hDQ f, hα
Qi =
α
hEQ0 f, hQ i − hEQ f, hQ i
Q0 ∈ch(Q)
X
= hf, EQ0 hα α
Q i = hf, hQ i.
Q0 ∈ch(Q)
The claimed identity follows, since the functionals x∗ ∈ X ∗ separate the points
x ∈ X by the Hahn–Banach theorem.
12.1 Dyadic singular integral operators 91
α
The functions hQ , with α ∈ {0, 1}d \ {0}, are referred to as cancellative Haar
functions, as they all have vanishing mean. In contrast, hQ 0
= |Q|−1/2 1Q is the
non-cancellative Haar function on Q. In the wavelet literature, the cancellative
Haar functions are special cases of mother wavelets, while the non-cancellative
Haar function is the father wavelet.
Lemma 12.1.3. Let X be a Banach space and p ∈ (1, ∞). Then the space of
finite linear combinations of cancellative Haar functions with X-coefficients,
n o
S00 (D; X) := span hQ α
⊗ x : Q ∈ D, α ∈ {0, 1}d \ {0}, x ∈ X , (12.2)
Proof. The filtration generated by the dyadic cubes, (Fk )k∈Z := (σ(D Sk ))k∈Z is
σ-finite with respect to the Lebesgue measure on Rd , and F∞ := σ k∈Z Fk
is the Borel σ-algebra of Rd . Hence Ek f → f in Lp (Rd ; X) as k → ∞ for all
f ∈ Lp (Rd ; X) by the forward convergence of generated martingales (Theorem
3.3.2). On the other hand, F−∞ := k∈Z Fk contains only sets of Lebesgue
T
measure 0 (the empty set) or ∞ (the quadrants, and their unions), which
means (by definition) that the Lebesgue measure is purely infinite on F−∞ .
Thus Ek f → 0 in Lp (Rd ; X) as k → −∞ for all f ∈ Lp (Rd ; X) by the
backward convergence of martingales (Theorem 3.3.5).
Combining these observation P about the limits at ±∞, it follows that func-
M −1
tions of the form EM f − Em f = k=m Dk f are dense in Lp (Rd ; X). Next, we
make the following observations about each Dk appearing in this expansion.
First, for any P ∈ Dm , multiplication with 1P commutes with Dk ; second,
∞
1P Dk f is a finite linear combination of some DQ f , and finally, if (Pi )i=1 is
PN
an enumeration of Dm , then i=1 1Pi f → f in L (R ; X) as N → ∞. Thus
p d
Proposition 12.1.5. Let X be a UMD space, p ∈ (1, ∞), and f ∈ S00 (D; X).
For any any α ∈ {0, 1}d \ {0} and coefficients λQ ∈ K, we have the estimates
X
α
λQ hhQ , f ihα
Q 6 βp,X sup |λQ |kf kLp (Rd ;X) ,
Lp (Rd ;X) Q∈D
Q∈D
X
+
εQ hhα α
Q , f ihQ 6 βp,X kf kLp (Rd ;X) .
Lp (Ω×Rd ;X)
Q∈D
−α γ
Then (Dα Q f, DQ f ) is a martingale difference sequence on Q, as each hQ with
α
γ∈/ {0, α} has average zero on the sets where hQ is constant. Appropriately
−α
enumerated, (Dα Q f, DQ f )Q∈D also forms a martingale difference sequence.
Estimating its martingale transform by a multiplying sequence of 0’s and 1’s,
we obtain
X X
α
λ Q DQ f p d = λQ · DQα
f + 0 · D−α
Q f p d
L (R ;X) L (R ;X)
Q∈D Q∈D
X
α −α
6 βp,X DQ f + DQ f ,
Lp (Rd ;X)
Q∈D
For the other claim, we argue by the contraction principle and the ran-
domised UMD inequality to see that
X X
0 −α
εQ Dα
Qf p d
6 ε α
Q DQ f + εQ DQ f
p d
L (Ω×R ;X) L (Ω×R ;X)
Q∈F Q∈D
X
+ −α
6 βp,X Dα
Q f + DQ f ,
Lp (Rd ;X)
Q∈D
For operator-valued coefficients λQ ∈ L (X, Y ), the following variants of R-
boundedness turn out to be relevant:
following estimates hold for all finitely non-zero families (xQ )Q∈C ⊆ X and
∗
(yQ )Q∈C ⊆ Y ∗ :
X
∗
|Q||hλQ xQ , yQ i|
Q∈C
X X
6 DR p (λ) ε Q x Q 1Q ∗
εQ yQ 1Q ,
Lp (Ω×Rd ;X) Lp0 (Ω×Rd ;Y ∗ )
Q∈C Q∈C
and
X X
ε Q λ Q x Q 1Q 6 E R p (λ) εQ xQ 1Q .
Lp (Ω×Rd ;Y ) Lp (Ω×Rd ;X)
Q∈C Q∈C
−1/2
Proof. We have hα α
Q = sgn(hQ )|Q| 1Q and hence, by the contraction prin-
ciple,
X X
α
ε Q zQ h Q = εQ |Q|−1/2 zQ 1Q
Lp (Ω×Rd ;Z) Lp (Ω×Rd ;Z)
Q∈D Q∈D
for both (zQ , Z) = {(xQ , X), (λQ xQ , Y )}. Using this twice, with both α and
γ, and in between the defining inequality of E R p (λ) for |Q|−1/2 xQ in place
xQ , yields the claim.
These notions are weaker than R-boundedness; we will shortly see that the
converse fails in general.
Lemma 12.1.8. For all Banach spaces X and Y , all operator families λ =
(λQ )Q∈C ⊆ L (X, Y ) and their adjoints λ∗ := (λQ
∗
)Q∈D ⊆ L (Y ∗ , X ∗ ), and
all p ∈ (1, ∞), we have
Proof. The last two estimates are immediate. The first estimate follows by
testing the defining condition of DR p with only one non-zero pair (xQ , yQ ∗
) at
a time. To see that DR p (λ) 6 E R p (λ), for suitable scalars |ηQ | = 1, we have
94 12 Dyadic operators and the T (1) theorem
X XZ
∗ ∗
|Q||hλQ xQ , yQ i| = ηQ hλQ xQ 1Q , yQ 1Q i
Q∈C Q∈C
Z DX X E
∗
=E εQ ηQ λQ xQ 1Q , ε R yR 1R
Q∈C R∈C
X X
∗
6 ε Q η Q λ Q x Q 1Q ε R yR 1R
Lp (Ω×Rd ;X) Lp0 (Ω×Rd ;Y ∗ )
Q∈C R∈C
X X
6 E R p (λ) εQ xQ 1Q ∗
εR yR 1R ,
Lp (Ω×Rd ;X) Lp0 (Ω×Rd ;Y ∗ )
Q∈C R∈C
Proof. Lemma 12.1.9 shows that we have this chain with “6” in place of “=”
throughout. On the other hand, Kahane’s contraction principle guarantees
that Rp (λ) = supQ∈C |λQ |. Thus we have equality throughout.
The following example of DR p -bounded families will play a role in our inves-
tigation of criteria for boundedness of singular integral operators; the uniform
boundedness of the quantities |Q|−1 hT 1Q , 1Q i is classically known as the weak
boundedness property of the operator T .
Example 12.1.10. Suppose that T ∈ L (Lp (Rd ; X), Lp (Rd ; Y )), and define
hT (1Q ), 1Q i ∈ L (X, Y ) by
Z
hT 1Q , 1Q i : x 7→ hT (1Q x), 1Q i = T (1Q x) ∈ Y.
Q
where
X
T εQ ηQ xQ 1Q
Lp (Ω×Rd ;Y )
Q∈C
X
6 kT kL (Lp (Rd ;X),Lp (Rd ;Y )) ε Q x Q 1Q
Lp (Ω×Rd ;X)
Q∈C
While Example 12.1.10 will only play a role later, the weakening of R-
boundedness has the following immediate application:
DR p (λ)
− 6 kHαγ
λ kL (Lp (Rd ;X),Lp (Rd ;Y )) 6 βp,X βp0 ,Y ∗ DR p (λ).
+ +
(12.4)
βp,X βp−0 ,Y ∗
α
For a fixed s ∈ Rd , the sequences (εQ 1Q (s)/|Q|1/2 )Q∈D and (εQ hQ )Q∈D have
equal distribution; thus
X 1Q X
εQ hf, hα
Qi = εQ hf, hα α
Q ihQ
|Q|1/2 Lp (Ω×Rd ;X) Lp (Ω×Rd ;X)
Q∈D Q∈D
+
6 βp,X kf kLp (Rd ;X)
by Proposition 12.1.5 in the last step. Similarly, the last term in (12.5) is
dominated by βp+0 ,Y ∗ kgkLp0 (Rd ;Y ∗ ) . Hence
|hHαγ
λ f, gi| 6 DR p (λ)βp,X βp0 ,Y ∗ kf kLp (Rd ;X) kgkLp (Rd ;Y ∗ ) ,
+ +
0
Then
X
Hαγ
λ f = |Q|1/2 ηQ λQ xQ hγQ ,
Q∈D
X
hHαγ
λ f, gi = ∗
|Q|ηQ hλQ xQ , yQ i,
Q∈D
and hence
X
∗
|Q||hλQ xQ , yQ i| 6 kHαγ
λ kL (Lp (Rd ;X),Lp (Rd ;Y ) kf kLp (Rd ;X) kgkLp (Rd ;Y ∗ ) ,
0
Q∈D
where
X
−
kf kLp (Rd ;X) 6 βp,X εQ |Q|1/2 ηQ xQ hα
Q
Lp (Ω×Rd ;X)
Q∈D
X
−
= βp,X εQ xQ 1Q
Lp (Ω×Rd ;X)
Q∈D
and combining the bounds, we have proved the first estimate in (12.4).
Remark 12.1.12. Under stronger assumptions on the coefficients λ, one can
improve the dependence on the UMD constants:
12.1 Dyadic singular integral operators 97
kHαα
λ kL (Lp (Rd ;X)) 6 βp,X kλk∞ .
where, in the last step, we used the second estimate in Proposition 12.1.5.
Here is a nice class of examples of coefficients satisfying the dyadic R-
boundedness condition:
Then
E R p ((haQ biQ )Q∈D ) 6 βp,Y
+
kak`∞ (L∞ ) kbkL∞ (Rd ;L (X,Y ))
Thus, for α, γ ∈ {0, 1}d \{0}, the Haar multiplier Hαγ
λ extends to a bounded
operator from Lp (Rd ; X) to Lp (Rd ; Y ) of norm at most
Proof. The second claim is immediate from the first one in combination with
Remark 12.1.12(2), so we concentrate on the first one. We may assume by
scaling that kaQ kL∞ (Rd ) 6 1. Then
98 12 Dyadic operators and the T (1) theorem
X X
εQ haQ biQ xQ 1Q = εQ EQ (aQ bxQ 1Q )
Lp (Rd ;Y ) Lp (Rd ;Y )
Q∈D Q∈D
X
+
6 βp,Y b ε Q x Q 1Q
Lp (Rd ;Y )
Q∈D
X
+
6 βp,Y kbkL∞ (Rd ;L (X,Y ) εQ xQ 1Q ,
Lp (Rd ;Y )
Q∈D
is essentially bounded.
Proof. (1)⇒(2): For b ∈ L∞ (0, 1; L (X, Y )), it is clear that the {hbiQ : Q ∈
D([0, 1))} is uniformly bounded. Under the assumption (1), this implies R-
boundedness by Proposition 8.6.1.
(2)⇒(3): This is clear.
(3)⇒(1): From the definition of R-boundedness, it is immediate that
R(T ) = sup({R(F ) : F ⊆ T finite}). So if some collection T is not R-
bounded, it hasSfinite subcollections Fn with R(Fn ) → ∞. Then the count-
∞
able collection n=1 Fn ⊆ T also fails to be R-bounded.
If (1) is not satisfied, then Proposition 8.6.1 says that the unit ball of
B̄L (X,Y ) of L (X, Y ) is not R-bounded. By what we just observed, this means
∞
that we can find a sequence {uk }k=0 ⊆ B̄L (X,Y ) that fails to be R-bounded.
4 1
Let vk := 3 uk − 3 uk+1 and
∞
X
b := vj 1[4−j−1 ,4−j ) .
j=0
Comparison of DR p and E R p
In the rest of this section, we make a further comparison of the two relaxed
notions of R-boundedness from Definition 12.1.6.. When Y is a UMD space—
an assumption that we make a good part of the time—, these notions turn
out to be equivalent. The universal bound DR p (λ) 6 E R p (λ) was already
observed in Lemma 12.1.8. The reverse estimate could be achieved essentially
by concatenating a couple of results that we have treated earlier in these
volumes, but it turns out that a slightly sharper quantitative bound can be
achieved by also revisiting their proofs to establish the following proposition:
Proposition 12.1.15. Let Y be a UMD space and p ∈ (1, ∞). Let E0 :=
{∅, Ω} be the trivial σ-algebra of a probability space (Ω, A , P) supporting a
Rademacher sequence (εn )N N
n=1 , and (Fn )n=1 be a σ-finite filtration of some
measure space (S, F , µ). Then, for all f ∈ Lp (Ω × S; Y ), we have
N
X
εn E(εn f |E0 × Fn ) +
6 βp,Y kf kLp (Ω×S;Y ) .
Lp (Ω×S;Y )
n=1
Thus
E(εn f |E0 × Fn ) = E(εn [E(f |En × Fn ) − E(f |En−1 × Fn )]|E0 × FN )
= E(εn [E(f |G2n ) − E(f |G2n−1 )]|E0 × FN )
= E(εn d2n |E0 × FN ),
where (
E(f |Gk ) − E(f |Gk−1 ), k = 2, . . . , 2N,
dk :=
E(f |G1 ), k = 1,
are martingale differences relative to a filtration (Gk )2N
k=1 on Ω × S defined by
N
X
εn E(εn f |E0 × Fn )
Lp (Ω×S;Y )
n=1
N
X
= ε02n E(εn d2n |E0 × FN )
Lp (Ω 0 ×Ω×S;Y )
n=1
N
X
= E( ε02n εn d2n |E0 × FN )
Lp (Ω 0 ×Ω×S;Y )
n=1
N
X N
X
6 ε02n εn d2n = 0
ε2n d2n
Lp (Ω 0 ×Ω×S;Y ) Lp (Ω 0 ×Ω×S;Y )
n=1 n=1
2N
X 2N
X
6 εk0 dk +
6 βp,Y dk
Lp (Ω 0 ×Ω×S;Y ) Lp (Ω×S;Y )
k=1 k=1
+ +
= βp,Y kE(f |G2N )kLp (Ω×S;Y ) 6 βp,Y kf kLp (Ω×S;Y ) ,
for all fn ∈ Lp (S; Y ), and the K-convexity inequality for UMD spaces (Propo-
sition 4.3.10),
12.1 Dyadic singular integral operators 101
N
X
+
εn E(εn f ) 6 Kp,Y kf kLp (Ω;Y ) , Kp,Y 6 βp,Y , (12.7)
Lp (Ω;Y )
n=1
Proof. We already proved the first inequality in Lemma 12.1.8. For the second
inequality, we first note that, by Fubini’s theorem,
X X
εQ zQ 1Q p d
= εn(Q) zQ 1Q p d
, (12.8)
L (Ω×R ;Z) L (Ω×R ;Z)
Q∈D Q∈D
0
In the Lp (Ω × Rd ; Y ∗ ) norm on the right, we write
X X X
εn(Q) hE(εn(Q) G)iQ 1Q = εn E(E(εn G)|σ(Dn ))1Q
Q∈D n∈Z Q∈Dn
X X
= εn E(E(εn G)|σ(Dn )) = εn E(εn G|{∅, Ω} × σ(Dn )).
n∈Z n∈Z
and we can replace n(Q) by Q on both sides according to (12.8) to obtain the
claimed result.
The fact that the collection D of dyadic cubes enjoys this property underlies
many considerations that we have encountered in these volumes.
12.1 Dyadic singular integral operators 103
Note that the dyadic cubes themselves, contained in this definition as a degen-
erate case with Q2 = Q1 , clearly satisfy this strong nestedness. This notion
is relevant for considerations dealing with Haar functions which, as we recall,
are constant on the dyadic children of their supporting dyadic cubes; thus,
if Q1 ∪ Q2 and R1 ∪ R2 are strongly nested, unequal but intersecting, then
the smaller union is entirely contained in a set of constant value for any Haar
function related to the larger union.
Our first (relatively simple) decomposition into strongly nested subcollec-
tions is the following:
Proof. Step 1 – Let all assumptions of the lemma be in force until further
notice. For each Q ∈ F ∪ φ(F ), we define a label r(Q) ∈ {0, 1, 2} such that
r(Q) 6= r(φ(Q)) for every Q ∈ F unless φ(Q) = Q. This ensures that Q∪φ(Q)
and R ∪ φ(R) are disjoint whenever Q, R ∈ F are two different cubes with
r(Q) = r(R) and `(Q) = `(R).
Indeed, Q 6= R implies φ(Q) 6= φ(R). Since different dyadic cubes of equal
size are disjoint, this implies that Q ∩ R = ∅ = φ(Q) ∩ φ(R). If φ(Q) = Q
or φ(R) = R, this already shows that Q ∪ φ(Q) and R ∪ φ(R) are disjoint. If
φ(Q) 6= Q and φ(R) 6= R, then r(φ(Q)) 6= r(Q) = r(R) implies φ(Q) 6= R and
similarly φ(R) 6= Q. By equal size again, this implies that φ(Q) ∩ R = ∅ =
Q ∩ φ(R), giving the (strong) nestedness when `(Q) = `(R).
104 12 Dyadic operators and the T (1) theorem
To define such r(R), let us denote φ◦0 (Q) = Q, φ◦k (Q) = φ(φ◦(k−1) (Q))
for k > 1. An orbit of φ is a set {φ◦k (Q) : k = 0, . . . , K}, where Q ∈ F and
either φ◦(K+1) (Q) = Q (in this case the orbit is cyclic), or Q ∈ / φ(F ) and
φ◦K (Q) ∈/ F . For all Q ∈ F ∪ φ(F ), we define r(Q) ∈ {0, 1, 2} by alternating
the values 0 and 1 on both non-cyclic orbits and cyclic orbits of even length,
and in addition using the value 2 once on cyclic orbits of odd length. In this
way, we ensure that r(Q) 6= r(φ(Q)) for any Q ∈ F unless Q = φ(Q).
Step 2 – It remains to check the strong nestedness in the case of Q, R ∈ F
with `(Q) < `(R), hence `(Q) < 2−n `(R). If Q ∪ φ(Q) intersect R ∪ φ(R), then
one of P ∈ {Q, φ(Q)} intersects one of S ∈ {R, φ(R)}. Since `(P ) < 2−n `(S)
and the cubes are dyadic, this implies that P (n) ( S. Since φ(Q) ⊆ Q(n) , we
have φ(Q)(n) = Q(n) , and hence Q ∪ φ(Q) ⊆ Q(n) ( S, confirming strong
nestedness in the case of `(Q) < `(R).
In the lack of (12.10), the situation is somewhat more complicated. Suitable
substitute conditions are provided in the following:
Proof. Step 1 – We define the label r(Q) ∈ {0, 1, 2} exactly as in the proof of
Lemma 12.1.20 to ensure that r(Q) 6= r(φ(Q)) unless Q = φ(Q). This gives
the nestedness of the sets Q ∪ φ(Q) for cubes of a fixed sidelength, as before.
Step 2 – We claim that, for each Q ∈ F ∪ φ(F ), there can be at most one
R ∈ F ∪ φ(F ) such that
Q ( R, 3Q(n) 6⊆ RQ , (12.11)
The required second label s(P ) is defined for each P ∈ F as follows. For
all P ∈ F of maximal size, let s(P ) := 0. Recursively, we proceed to the unla-
belled cubes P ∈ F of maximal size. For these cubes, we first check whether
(12.11) occurs with either Q = P or Q = φ(P ), and some R ∈ F ∪ φ(F ).
It could happen that R ∈ F , or R = φ(S) with S ∈ F , or both. We then
require that s(P ) is chosen so that (r(P ), s(P )) ∈
/ {(r(R), s(R)), (r(S), s(S))}.
If S = R, this is clearly one restriction on S(P ). But if S 6= R = φ(S), then
r(R) 6= r(S) by the alternating choice of r along the orbits, and we still get
at most one restriction of the possible value of s(P ). Since different R and S
may arise from the case Q = P and Q = φ(P ) we get altogether at most two
restrictions on s(P ), and we can declare that s(P ) is the smallest remaining
number in {0, 1, 2}.
The next result relaxes the assumptions even further, at the cost of compli-
cating the conclusions:
Lemma 12.1.22. Assume conditions (a) through (d). Then F can be par-
titioned into 144 subcollections Fi , and on each of them we have injections
φi,j : Fi → D, j = 0, 1, 2, 3, where φi,0 (Q) = Q and φi,3 (Q) = φ(Q) such that
each collection
{φi,j (Q) ∪ φi,j+1 (Q) : Q ∈ Fi } (12.12)
is strongly nested.
Proof. The idea is to combine the special cases treated in the two previous
Lemmas 12.1.20 and 12.1.21, which had the additional assumptions (12.10)
and (e), respectively; neither is assumed now.
For every R ∈ D, consider the 2nd cubes Q ∈ D with Q(n) = R. Among
them, there are (2n − 2)d off-boundary cubes Q with 3Q ⊆ R, while the
number of boundary cubes is then
1 nd
2nd − (2n − 2)d = 2nd [1 − (1 − 21−n )d ] 6 2nd · 21−n d 6 2
2
if n > log2 (4d). When this is the case, we can define a permutation ψ : D → D
with `(ψ(Q)) = `(Q), ψ(Q) ⊆ Q(n) (as in (12.10)) such that ψ(Q) is an off-
boundary cube in Q(n) whenever Q is a boundary cube in Q(n) .
Let us first divide F into four subcollection Fu,v , where u, v ∈ {0, 1}, so
that Q ∈ Fu,v is a boundary cube in Q(n) if and only if u = 1, whereas φ(Q)
is a boundary cube in ψ(Q)(n) if and only if v = 1.
Case F0,0 : By Lemma 12.1.21, we can divide F0,0 into nine subcollections Fi
such that {Q ∪ φ(Q) : Q ∈ Fi } is strongly nested. Letting φi,1 = φi,2 = φi,3 =
φ in this case, we trivially have the strong nestedness of {φi,j (Q) ∪ φi,j+1 (Q) :
Q ∈ Fi } for j = 1, 2 (since the collection is simply φ(Fi ) ⊆ D in this case.
106 12 Dyadic operators and the T (1) theorem
Case F0,1 : On the collection F0,1 , we consider the map ψ ◦ φ and observe
that it also satisfies (d); indeed, φ(Q) ⊆ 3Q(n) lies inside one of the dyadic
neighbours of Q(n) , and ψ keeps it inside this same nth generation ancestor.
Since φ(Q) is a boundary cube in φ(Q)(n) for Q ∈ F0,1 by definition of this
collection, ψ(φ(Q)) is off-boundary in φ(Q)(n) = ψ(φ(Q))(n) by definition of
ψ, and hence (F0,1 , ψ ◦ φ) also satisfies (e) in place of (F , φ). Then Lemma
12.1.21 shows that F0,1 can be divided into nine subcollections Fa0 such that
each {Q ∪ ψ(φ(Q)) : Q ∈ Fa0 } is strongly nested. On the other hand, we can
write
Here (F0,1 , ψ) satisfies the assumptions of Lemma 12.1.20, and hence φ(F0,1 )
can be divided into three subcollections Gb such that {R ∪ ψ(R) : R ∈ Gb )} is
strongly nested. This since φ is injective, this induces a decomposition of F0,1
into three subcollections where Fb00 such that {ψ(φ(Q)) ∪ φ(Q) : Q ∈ Fb00 }
is strongly nested. Then, defining Fi = Fa0 ∩ Fb00 for i = (a, b), we find that
both
{Q ∪ ψ(φ(Q)) : Q ∈ Fi }, {ψ(φ(Q)) ∪ φ(Q) : Q ∈ Fi }
are strongly nested, and there is in total 9 · 3 such collections Fi decomposing
F0,1 . So taking φi,1 = ψ ◦ φ and φi,2 = φi,3 , we have the strong nestedness of
the collections in (12.12), the case j = 2 for trivial reasons as in case F0,0 .
Case F1,0 : Similarly, on the collection F1,0 , Lemma 12.1.20 applies to the
mapping ψ to provide three subcollection Fa0 such that {Q ∪ ψ(Q) : Q ∈ Fa0 }
is strongly nested. And Lemma 12.1.21 applies to (ψ(F1,0 ), φ◦ψ −1 ) to provide
nine subcollections Fb00 such that {ψ(Q) ∪ φ(Q) : Q ∈ Fb00 } is strongly nested.
So altogether we have 3 · 9 subcollection Fi = Fa0 ∩ Fb00 such that
are strongly nested. We can hence define φi,1 = ψ, φi,2 = φi,3 = φ to get the
claimed conclusions.
Case F1,1 : Finally, on the collection F1,1 , Lemma 12.1.20 applies to both
(F1,1 : ψ) and to (ψ ◦ φ(F1,1 : ψ −1 ) to provide three subcollections Fa0 and
three other Fb00 such that {Q ∪ ψ(Q) : Q ∈ Fa0 } and {ψ(φ(Q)) ∪ φ(Q) :
Q ∈ Fb00 } are strongly nested. And we check that Lemma 12.1.21 applies to
(ψ(F1,1 ), ψ◦φ◦ψ −1 ) to provide nine subcollections Fc000 such {ψ(Q)∪ψ(φ(Q)) :
Q ∈ Fc000 } is strongly nested. Then with Fi = Fa0 ∩ Fb00 ∩ Fc000 we obtain 32 · 9
subcollections such that {Q∪ψ(Q) : Q ∈ Fi }, {ψ(Q)∪ψ(φ(Q)) : Q ∈ Fi }, and
{ψ(φ(Q)) ∪ φ(Q) : Q ∈ Fi } are strongly nested, and we can define φi,1 = ψ,
φi,2 = ψ ◦ φ, φi,3 = φ in this case.
In total we have divided F into 9 + 2 · 9 · 3 + 9 · 32 = 144 subcollections
Fi with required properties.
Another variant of the conclusion with the same assumptions is as follows:
12.1 Dyadic singular integral operators 107
Lemma 12.1.23. Assume conditions (a) through (d). Then F can be parti-
tioned into 33d+1 subcollections Fi such that each collection
termines φ(Q) and then Q00 . The map Φ is also injective, since Q00 uniquely
determines φ(Q), which (since φ is injective) determines Q and then Q0 . More-
over, we have
Thus F a,m ⊆ D m;3 and Φ satisfy properties (a) and (b) in place of F ⊆ D
and φ, and the scale-separation property (c) is clearly inherited by Φ from φ.
Moreover, the nth D m;3 -ancestor of both Φ(Q0 ) = Q00 and Q0 is clearly 3Q(n)
by construction, and hence Φ satisfies condition (12.10) of Lemma 12.1.20. The
said lemma guarantees that F a,m can be split into 3 subcollections Fja,m , so
that each
{Q0 ∪ Φ(Q0 ) : Q0 ∈ Fja,m } ⊆ D m;3
is strongly nested. Writing i = (a, m, j), and defining
Fi := {Q ∈ F : mQ = m, Q[m] ∈ Fja,m },
these are precisely the collections that we wanted to construct. Since a takes 9d
values, m takes 3d values, and j takes 3 values, the number of these collections
is 9d · 3d · 3 = 33d+1 , as claimed.
108 12 Dyadic operators and the T (1) theorem
Remark 12.1.24. In each of the Lemmas 12.1.20 through 12.1.23, we can drop
assumption (c) at the cost of multiplying the required number of decomposing
subcollections Fi by n + 1.
Proof. For any F ⊆ D, consider the n + 1 subcollection F k := {Q ∈ F :
log2 `(Q) ≡ k mod (n + 1)}. Each of these clearly satisfies (c). Moreover, any
of the other properties (a) through (e) as well as (12.10), if valid for F , is
clearly inherited by each F k . Thus, if F satisfies the assumptions of any of
these lemmas with the possible exception of (c), then each F k satisfies all of
the relevant assumptions, and the lemma in question provides a decomposition
of F k into some Fik with appropriate nestedness conditions. The Sn required
decomposition of the original F is then obtained simply as F = k=0 i Fik ,
S
and clearly the number of collections inS this decomposition is n + 1 times as
many as in the decompositions F k = i Fik given by the lemmas.
(1) If, in addition, Y has type t ∈ [1, p] and X has cotype q ∈ [p, ∞], or one
of them has both, then we also have the estimate
αγ − +
kTφλ k 6 Ad (n + 1)1/t−1/q βp,Y βp,X C(X, Y, p, q, t; λ)
where
n
+
C(X, Y, p, q, t; λ) := min τt,X;p · βp,X · cq,X;p · E R p (λφ−1 ),
+
τt,Y ;p · βp,X · cq,X;p · E R p (λφ−1 ),
+
τt,Y ;p · βp,Y · cq,X;p · E R p (λ),
o
+
τt,Y ;p · βp,Y · cq,Y ;p · E R p (λ)
6 C(X, Y, p, q, t) · Rp (λ),
and
n
+ +
C(X, Y, p, q, t) := min τt,X;p βp,X cq,X;p , τt,Y ;p βp,X cq,X;p ,
o (12.15)
+ +
τt,Y ;p βp,Y cq,X;p , τt,Y ;p βp,Y cq,Y ;p .
(2) If, in addition, λQ 6= 0 only when φ(Q) ⊆ Q(n) , then we have the alterna-
tive norm estimate
αγ 0
kTφλ +
k 6 3 · βp,Y βp,X min{cq,X;p , cq,Y ;p }(n + 1)1/q E R p (λ).
0
(3) For all f ∈ Lp (Rd ; X) and g ∈ Lp (Rd ; Y ∗ ), the extended operator has the
absolutely convergent representation
αγ
XD γ
E
hTφλ f, gi = λQ hf, hα Q i, hg, hφ(Q) i .
Q∈D
When kf kLp (Rd ;X) 6 and kgkLp0 (Rd ;Y ∗ ) 6 1, the corresponding absolute
value series is dominated by the same upper bounds as those given for
αγ
kTφλ kL (Lp (Rd ;X),Lp (Rd ;Y )) above.
Remark 12.1.26. (1) In the prominent special case that X = Y , we have
C(X, X, p, q, t; λ) = C(X, X, p, q, t) · min{E R p (λφ−1 ), E R p (λ)},
+
C(X, X, p, q, t) = τt,X;p · βp,X · cq,X;p .
(2) Case (0) of Theorem 12.1.25 is a special case of (1) using the trivial type
and cotype exponents t = 1, q = ∞ with corresponding constants equal
to one. The role of non-trivial type and cotype is to relax the dependence
on the parameter n. The estimate obtained in case (2) is not strictly
comparable to the other two bounds; its main advantage over the other
two is achieving a quadratic bound in terms of the UMD constants, in
contrast to the cubic bound in the other cases.
110 12 Dyadic operators and the T (1) theorem
where 1 is the constant sequence of all ones. Hence, for the qualitative
conclusion of Theorem 12.1.25, it would suffice to consider just X = Y
and λ = 1, and then combine this special case with Theorem 12.1.11;
however, the reader will quickly realise that this approach would produce
a higher power of the UMD constants in the quantitative conclusion.
Before going into the proof, let us still formulate a corollary in the important
special case when φ : D → D is a bijection:
Corollary 12.1.27. Let φ : D → D be a bijection with `(φ(Q)) = `(Q) and
φ(Q) ⊆ 3Q(n) for some n ∈ N. Let X and Y be a UMD spaces, and let
p ∈ (1, ∞). Suppose that Y has type t ∈ [1, p] and X has cotype q ∈ [p, ∞], or
one of them has both. Let λ = (λQ )Q∈D ⊆ L (X, Y ) be R-bounded, consider
αγ
the mapping Tφλ as in (12.14), and let
αγ αγ
kTφλ k := kTφλ kL (Lp (Rd ;X),Lp (Rd ;Y )) .
where
(2) If, in addition, λQ 6= 0 only when φ(Q) ⊆ Q(n) , then we have the alterna-
tive norm estimate
n 0
o
αγ
kTφλ k 6 3 · βp,X βp,Y min C(n + 1)1/q Rp (λ), C ∗ (n + 1)1/t Rp∗0 (λ) .
where
Proof. The first versions of both bounds (i.e, using the first item of the respec-
tive minimums) above are simply those of Theorem 12.1.25, cases (1) and (2),
±
where we estimated all UMD constants by βp,Z 6 βp,Z . The second versions
of both bounds then follow by duality: When φ : D → D is a bijection, one
directly verifies that
αγ ∗
(Tφλ ) = Tφγα
−1 ,λ∗
φ−1
is an operator of the same form, acting from Q00 (Rd ; Y ∗ ) to Q00 (Rd ; X ∗ ) and
0 0
eventually from Lp (Rd ; Y ∗ ) to Lp (Rd ; X ∗ ). If Z ∈ {X, Y } has type t, then Z ∗
12.1 Dyadic singular integral operators 111
has cotype t0 with ct0 ,Z ∗ ;p0 6 τt,Z;p . (See Proposition 7.1.13; it is formulated
for p = t, but the same short argument is easily modified to give the general
statement.) If a UMD space Z has cotype q, then it has martingale type q
(Proposition 4.3.13), hence Z ∗ has martingale cotype q 0 (Proposition 3.5.29),
and thus cotype q 0 (as observed right before Proposition 4.3.13). Thus we can
apply the case already handled, with (Y ∗ , X ∗ , p0 , t0 , q 0 ) in place of (X, Y, p, q, t),
to get
αγ
kTφλ kL (Lp (Rd ;X),Lp (Rd ;Y )) = kTφγα
−1 ,λ kL (Lp0 (Rd ;Y ∗ ),Lp0 (Rd ;X ∗ ))
φ−1
0 0
6 6 · 34d βp0 ,Y ∗ βp0 ,X ∗ (n + 1)1/q −1/t C(Y ∗ , X ∗ , p0 , t0 , q 0 )Rp0 (λ∗ ).
The claim then follows from βp0 ,Z ∗ = βp,Z and 1/q 0 − 1/t0 = 1/t − 1/q.
The second version of the second bound is obtained from the first version
in the entirely similar way by duality.
Proof of Theorem 12.1.25. Claim (0) is the special case t = 1, q = ∞ of (1),
so we only need to prove the latter of the two. Let F be a finite collection of
dyadic cubes. Then F and φ satisfy the assumptions of Lemma 12.1.23, except
possibly the scale separation (c). By Remark 12.1.24, the lemma still applies
to produce 33d+1 (n + 1) subcollections Fi ⊆ F with the properties given in
γ
Lemma 12.1.23. Let us write xQ = hf, hα Q i. Since the functions (hQ )Q∈F form
a martingale difference sequence, we have
X γ
X
λQ xQ hφ(Q) p d
6 β −
p,Y εQ λQ xQ hγφ(Q) p d
.
L (R ;Y ) L (Ω×R ;Y )
Q∈F Q∈F
From this point on, we have some flexibility as to when we want to “pull out”
the coefficients λQ . For this reason, let us write zQ ∈ Z for a generic choice
of either zQ = λQ xQ ∈ Y or zQ = xQ ∈ X. We then continue with
X X X
εQ zQ hγφ(Q) = εi0 εQ zQ h0φ(Q)
Lp (Ω×Rd ;Z) Lp (Ω 0 ×Ω×Rd ;Z)
Q∈F i Q∈Fi
X X t 1/t
6 τt,Z;p εQ xQ h0φ(Q) ,
Lp (Ω×Rd ;Z)
i Q∈Fi
inequality is due to the fact that the cubes Q[m(i)] and φ(Q)[m(i)] are not
necessarily different.) Hence
2 · 3d
Z
1φ(Q) 6 1φ(Q) |Q| = 1φ(Q) 2 · 3d − 1Q 6 2 · 3d EE(Q) 1Q ,
|E(Q)| E(Q)
where the EE(Q) are conditional expectations associated with a nested family,
and hence with a filtration. This allows us to use Stein’s inequality (Theorem
4.2.23) to the effect that
X
εQ zQ h0φ(Q)
Lp (Ω×Rd ;Z)
Q∈Fi
X
6 2 · 3d εQ zQ EE(Q) h0Q (12.16)
Lp (Ω×Rd ;Z)
Q∈Fi
X
+
6 2 · 3d · βp,Z εQ zQ h0Q
Lp (Ω×Rd ;Z)
Q∈Fi
Then
X X t 1/t
εQ zQ h0Q
Lp (Ω×Rd ;Z)
i Q∈Fi
X X q 1/q
6 (33d+1 (n + 1))1/t−1/q εQ zQ h0Q
Lp (Ω×Rd ;Z)
i Q∈Fi
X X
6 (33d+1 (n + 1))1/t−1/q cq,Z;p ε0i εQ zQ h0Q ,
Lp (Ω 0 ×Ω×Rd ;Z)
i Q∈Fi
If we did not already pull out the coefficients λQ , we do it at this point, after
which we are left with
X
+
ε Q x Q hα
Q 6 βp,X kf kLp (Rd ;X) ,
Lp (Ω×Rd ;X)
Q∈F
where the order of the constants reflects the order of applying the related
estimates: Before pulling out the coefficients λQ , we apply estimates on the
Y side, and after that on the X side. Taking the minimum of the four terms,
we arrive at the assertion of the theorem.
The alternative estimate (2): In order to make efficient use of the additional
assumption φ(Q) ⊆ Q(n) when λQ 6= 0, we will need to modify the preceding
considerations at various points.
Let F be a finite collection of dyadic cubes, and F λ := {Q ∈ F : λQ 6==
0}. Then F λ and φ satisfy the assumptions of Lemma 12.1.20, except possibly
the scale separation (c). By Remark 12.1.24, the lemma still applies to pro-
duce 3(n + 1) subcollections Fiλ ⊆ F λ with the properties given in Lemma
α
12.1.20. Let us write xQ = hf, hQ i. In the first step, we simply use the triangle
inequality:
X γ
X X γ
λQ xQ hφ(Q) p d
6 λQ xQ hφ(Q) p d
.
L (R ;Y ) L (R ;Y )
Q∈F i Q∈Fiλ
The more interesting deviations from the previous case begin now.
Note that hQ α
= |Q|−1/2 (1Q+
α
±
− 1Qα− ) for suitable subsets Qα ⊆ Q with
± 1
|Qα | = 2 |Q|. If Q 6= φ(Q), we see that
1 α 1 −1/2
d+
Q := (h + hγφ(Q) ) = |Q| (1Q+ + − 1 −
Qα ∪φ(Q)− ),
2 Q 2 α ∪φ(Q)γ γ
1 α γ 1 −1/2
d−
Q := (hQ − hφ(Q) ) = |Q| (1Q+ − − 1 −
α ∪φ(Q)γ Qα ∪φ(Q)+ )
2 2 γ
1 α γ
d−
Q := (hQ − hQ ) = |Q|
−1/2
(1Q+ − − 1 −
α ∩Qγ Qα ∩Q+ )
2 γ
where we used the definition of UMD with signs ±εQ multiplying the mar-
tingale differences d± Q , followed by taking an average over the εQ . (It might
appear at first glance that we could have used just the one-sided UMD− prop-
−
erty to arrive at the same conclusion with the smaller constant βp,Y , but this
−
is not the case: an application of the one-sided UMD property would give us
independent random signs, say ε± ±
Q , in front of each dQ , and this is not what
we want.)
For zQ ∈ {xQ , λQ xQ } and Z ∈ {X, Y } we then have
X X
ε Q zQ h α
Q p d
L (Ω×R ;Z)
i Q∈Fiλ
0
X X q 1/q
6 (3(n + 1))1/q ε Q zQ h α
Q
Lp (Ω×Rd ;Z)
i Q∈Fiλ
(12.18)
0 X X
6 (3(n + 1))1/q cq,Z;p ε0i ε Q zQ h α
Q
Lp (Ω×Rd ;Z)
i Q∈Fiλ
0 X
= (3(n + 1))1/q cq,Z;p ε Q zQ h α
Q ,
Lp (Ω×Rd ;Z)
Q∈F λ
12.1 Dyadic singular integral operators 115
using in the last step the fact that the F λ = i Fiλ is a disjoint partition, so
S
the independent random signs εQ with Q ∈ F λ do not “see” the multiplying
signs ε0i . Hence, pulling out the λQ either at the beginning or at the end of
(12.18) (but in any case only after having replaced the translated hγφ(Q) by
hαQ in (12.17), which in contrast to what happened in the previous case of the
proof), we obtain
X γ
X X
λQ xQ hφ(Q) 6 βp,Y ε Q λ Q x Q hα
Q
Lp (Rd ;Y ) Lp (Rd ;Y )
Q∈F i Q∈Fiλ
0 X
6 βp,Y E R p (λ)(3(n + 1))1/q min{cq,X;p , cq,Y ;p } α
ε Q x Q hQ .
Lp (Rd ;X)
Q∈F λ
αγ
is a Haar projection of f ; the action of Tηλ,φ is thus well-defined via the initial
definition on this space. From the previous part of the theorem that we already
proved, we have
X D γ
E
λQ hf, hα
Q i, hg, h φ(Q) i
Q∈F
αγ
6 kTηλ,φ kL (Lp (Rd ;X),Lp (Rd ;Y )) kPF f kLp (Rd ;X) kgkLp (Rd ;Y ∗ ) .
We now apply this estimate with the increasing sequence of finite sets
and this is seen to satisfy kPFN f kLp (Rd ;X) 6 2kf kLp (Rd ;X) and PFN → f in
Lp (Rd ; X) as N → ∞. Thus
X D γ
E
λQ hf, hα
Q i, hg, h φ(Q) i
Q∈D
X D E
γ
= lim λQ hf, hα
Q i, hg, hφ(Q) i
N →∞
Q∈FN
αγ
6 kTηλ,φ kL (Lp (Rd ;X),Lp (Rd ;Y )) lim kPF f kLp (Rd ;X) kgkLp (Rd ;Y ∗ )
N →∞
αγ
= kTηλ,φ kL (Lp (Rd ;X),Lp (Rd ;Y )) kf kLp (Rd ;X) kgkLp (Rd ;Y ∗ ) ,
αγ αγ
where Tηλ,φ has the same norm estimate as Tλ,φ , since
(1) If, in addition, X or Y has cotype q ∈ [p, ∞], then we also have
(
γ 1−1/q − + C(X, Y, p, q) · Rp (λ),
kUφ k 6 Bd (n + 1) βp,Y βp,X +
βp,Y · min{cq,X;p , cq,Y ;p } · E R p (λ),
where
n o
+ + +
C(X, Y, p, q) := min βp,X cq,X;p , βp,Y cq,X;p , βp,Y cq,Y ;p
(12.20)
= C(12.15) (X, Y, p, q, 1).
(2) If, in addition, we have λQ 6= 0 only when φ(Q) ⊆ Q(n) , then we have the
alternative norm estimate
When kf kLp (Rd ;X) 6 1 and kgkLp0 (Rd ;Y ∗ ) 6 1, the corresponding absolute
value series is dominated by the same upper bounds as those given for
γ
kUφλ kL (Lp (Rd ;X),Lp (Rd ;Y )) above.
where 1 is the constant sequence of all ones. Hence, for the qualitative
conclusion of Theorem 12.1.28, it would suffice to consider just X = Y
and λ = 1, and then combine this special case with Theorem 12.1.11;
however, the reader will quickly realise that this approach would produce
a higher power of the UMD constants in the quantitative conclusion.
(4) In contrast to Theorem 12.1.25, our proof of Theorem 12.1.28 does not
allow replacing the assumptions on λ by E R p (λφ−1 ) < ∞. The related
issue of when in the argument, and under what assumptions, we may pull
out the coefficients λQ , is shortly discussed inside the proof.
where each collection {φi,j (Q) ∪ φi,j+1 (Q) : Q ∈ Fi } is strongly nested. But
this implies that each
is (or can be enumerated as) a martingale difference sequence. Note that here
it is important that a smaller union φi,j+1 (Q)∪φi,j (Q) is not just contained in
a larger φi,j+1 (R)∪φi,j (R), but entirely in (a dyadic child of) one of φi,j+1 (R)
or φi,j (R), where the function h0φi,j+1 (R) − hφ0 i,j (R) is constant.
Using this martingale difference property, we can then proceed as in the
proof of Theorem 12.1.25. Let us abbreviate xQ := hf, hγQ i ∈ X and yQ :=
λQ x Q ∈ Y .
12.1 Dyadic singular integral operators 119
X
yQ (h0φ(Q) − hQ
0
)
Lp (Rd ;Y )
Q∈F
2 X
X X
6 yQ (h0φi,j+1 (Q) − h0φi,j (Q) )
Lp (Rd ;Y )
j=0 i Q∈Fi
2 X
X X
−
6 βp,Y εQ yQ (h0φi,j+1 (Q) − hφ0 i,j (Q) )
Lp (Ω×Rd ;Y )
j=0 i Q∈Fi
3
(
−
X X X α0 = α3 = 1,
6 βp,Y αj εQ yQ h0φi,j (Q) ,
j=0 i Q∈Fi
Lp (Ω×Rd ;Y ) α1 = α2 = 2,
where the first and the last steps were simply triangle inequalities.
As in the proof of Theorem 12.1.25, we have some flexibility on when to
pull out the coefficients λQ , and we again proceed with a generic choice of
zQ ∈ Z for either yQ ∈ Y or xQ ∈ X. The norm to be estimated has exactly
the same form as what we estimated (12.16) in the proof of Theorem 12.1.25
(using Lemma 12.1.23 in this step), and we can there read the bound
X
εQ zQ hφ0 i,j (Q)
Lp (Ω×Rd ;Z)
Q∈Fi
X (12.21)
+
6 2 · 3d · βp,Z 0
ε Q zQ h Q .
Lp (Ω×Rd ;Z)
Q∈Fi
It is no later than here that we should to pull out the coefficients λQ , after
which we are left with the final step, based on Proposition 12.1.5, that
120 12 Dyadic operators and the T (1) theorem
X γ +
ε Q x Q hQ 6 βp,X kf kLp (Rd ;X) .
Lp (Ω×Rd ;X)
Q∈F
Rp (λ) · βp,X
+
· cq,X;p ,
+
βp,Y · Rp (λ) · cq,X;p ,
+
βp,Y · cq,Y ;p · Rp (λ).
In the latter two versions, i.e., pulling out the λQ only after making the step
(12.21) with Z = Y , we might as well replace Rp (λ) by E R p (λ), thus leading
to the possible upper bounds
+
βp,Y · E R p (λ) · cq,X;p ,
+
βp,Y · cq,Y ;p · E R p (λ).
(On the other hand, if we wanted to pull out the λQ before step (12.21), and
thus apply (12.21) with Z = X, the coefficient λQ would be multiplying a
Haar function h0φi,j (Q) ; this would lead to a constant of the type E R p (λφ−1 ),
i,j
where φi,j need not be the original φ from the assumptions of the theorem, but
one of the auxiliary mappings produced by Lemma 12.1.22. This would lead
to an unreasonably technical formulation of probably little practical value,
which is why we have not included the resulting alternative upper bound in
the statement of the theorem.)
Altogether, choosing the best of the possible alternative estimates, we
arrive at
X
xQ (h0φ(Q) − h0Q ) p d kf k−1 Lp (Rd ;X)
L (R ;X)
Q∈F
3
−
X 1/q0 +
6 βp,X αj (2 · 3d )(144 · 33d+1 · (n + 1) βp,X ×
j=0
(
C(X, Y, p, q)Rp (λ),
× +
βp,Y min{cq,X;p , cq,Y ;p }E R p (λ),
P3
where C(X, Y, p, q) is as in the statement of the Theorem, and j=0 αj =
1 + 2 + 2 + 1 = 6.
The alternative estimate (2): As in the previous proof of Theorem 12.1.25(2),
we construct some auxiliary martingale differences. The initial considerations
are identical:
12.1 Dyadic singular integral operators 121
1
d2Q := |Q|−1/2 (−1φ(Q) + 2 · 1Q+ ),
3 α
where d2Q has average zero on the sets where d1Q is constant; note that, unlike
in the proof of Theorem 12.1.25(2), the order matters now. Moreover, we can
recover the original functions by
1 −1/2
d1Q + d2Q = = hα
|Q| (1 − 1)1φ(Q) + (1 + 2)1Q+ − 3 · 1Q − Q,
3 α α
1
d1Q − 2d2Q = |Q|−1/2 (1 + 2)1φ(Q) ) + (1 − 4)1Q+ = h0φ(Q) − h0Q .
− 3 · 1Q −
3 α α
hα 1 1 2
Q = dQ = dQ + dQ , h0φ(Q) − h0Q = 0 = 0 · d1Q − 2d2Q .
12.2 Paraproducts
The notion of paraproducts arises from a number of considerations. Here we
choose a point of departure that also motivates their name: they are objects
that arise from a decomposition of the ordinary pointwise product of functions.
While paraproducts certainly look more complicated than the regular product,
it turns out that in certain respects they actually behave better. Another
motivation is the key role that these objects play in the T (1) theorem in
Section 12.3. Some further connections will be discussed in the Notes.
Proposition 12.2.1. Let b ∈ L1loc (Rd ; L (X, Y )), where X and Y are Banach
spaces, and let f ∈ S00 (D; X). Then
X
bf = Hbαγ f + Πb f + Πb∗ f, (12.23)
α,γ∈{0,1}d \{0}
where Hαγ
b are Haar multipliers of the form
X
Hbαγ f := hsgn(hα γ α γ
Q hQ )biQ hf, hQ ihQ ,
Q∈D
X 1Q
Πb∗ f := hb, hα α
Q ihf, hQ i ,
|Q|
Q∈D
α∈{0,1}d \{0}
where the series of Πb∗ f is finitely non-zero, and the non-zero terms in Πb f
are attached to cubes contained in finitely many maximal ones, and the series
converges (at least) conditionally along any decreasing order of the dyadic
cubes contained in these maximal ones.
12.2 Paraproducts 123
This has formally the same structure as Πb f , but with the roles of b and f
reversed, and hence (12.23) could be also written in the form
X
bf = Hαγ ∗
b f + Πf b + Πb f + Πb f,
α,γ∈{0,1}d \{0}
α6=γ
where the series converges (at least) conditionally along any decreasing order
of the dyadic cubes Q ⊆ R, by the Martingale Converge Theorem 3.3.2, since
this is a martingale difference expansion of the function 1R (b − hbiR )x ∈
L1 (Rd ; Y ).
We observe that
α θ
hQ hR = hα θ
Q hhR iQ ∀Q ( R,
whereas
θ θ 1R hα+θ
hR hR = , hα θ
R hR
R
, ∀α 6= θ,
|R| |R|1/2
where we use modulo 2 addition in {0, 1}d . Hence
X X
hb, hα α θ
Q ix ⊗ hQ hR = hb, hα α
Q ihf iQ ⊗ hQ = Πb f,
Q(R Q∈D
α∈{0,1}d \{0} α∈{0,1}d \{0}
124 12 Dyadic operators and the T (1) theorem
θ
observing that hf iQ = hhR iQ x = 0 unless Q ( R. Moreover,
X
hb, hα α θ θ
R ix ⊗ hR hR + hbiR x ⊗ hR
α∈{0,1}d \{0}
1R X hα+θ hb, h0 i
= hb, hθR ix ⊗ + hb, hα
R ix ⊗
R
+ R
x ⊗ hθR
|R| |R|1/2 |R|1/2
α∈{0,1}d \{0,θ}
1R X hα+θ
= hb, hθR ix ⊗ + hb, hα
R ix ⊗
R
.
|R| |R|1/2
α∈{0,1}d \{θ}
Proposition 12.2.3. Let X and Y be UMD spaces and p ∈ (1, ∞). Let
b ∈ L∞ (Rd ; L (X, Y )). Then Λb := Πb + Πb∗ , initially defined on S00 (D; X),
extends to a bounded operator from Lp (Rd ; X) to Lp (Rd ; Y ) of norm
− + +
kΛb kL (Lp (Rd ;X),Lp (Rd ;Y )) 6 1 + (2d − 1)2 βp,Y βp,Y βp,X kbkL∞ (Rd ;L (X,Y )) ,
We will obtain a far better estimate in Theorem 12.2.25, but it seems worth-
while recording this relatively simple bound as an illustration of the techniques
that we have developed thus far.
Proof of Proposition 12.2.3. It is clear that pointwise multiplication by b ∈
L∞ (Rd ; L (X, Y )) defines a bounded operator from Lp (Rd ; X) to Lp (Rd ; Y ),
12.2 Paraproducts 125
for any Banach spaces X, Y and all p ∈ [1, ∞]. Moreover, the Haar multi-
plier Hαγ
b featuring in Proposition 12.2.3 have exactly the form considered in
Proposition 12.1.13, and hence
kHαγ − + +
b f kLp (Rd ;Y ) 6 βp,Y βp,Y βp,X kbkL∞ (Rd ;L (X,Y )) kf kLp (Rd ;X) .
for all f ∈ S00 (D; X), and hence Λb extends to a bounded operator from
Lp (Rd ; X) to Lp (Rd ; Y )) with the asserted norm estimate. Since the claimed
identity holds (by Proposition 12.2.1) for all f ∈ S00 (D; X), and each term is
continuous with respect to the Lp (Rd ; X) norm of f (as we just showed), it is
immediate that this identity extends to all f ∈ Lp (Rd ; X).
As we shall see later, the operator Λb is not only as good as, but actually better
than the pointwise product f 7→ bf , in the sense that it remains a bounded
operator for a broader class of functions b than just the bounded ones. As
the reader will have guessed from the introduced notation, we will also be
interested in the mapping properties of the individual paraproducts Πb and
Πb∗ .
While the paraproduct Πb arouse from our analysis of the pointwise prod-
uct with a multiplier b, in other considerations we will encounter similar series
X
α
Πf = πQ hf iQ ⊗ hα
Q
Q∈D
α∈{0,1}d \{0}
α α
with some coefficient πQ replacing the Haar coefficients hb, hQ i of a function
α α
b above. Formally, we always have πQ = hb, hQ i by choosing
X
α
“ b := Π(1) = πQ ⊗ hα
Q ”,
Q∈D
α∈{0,1}d \{0}
but giving a precise meaning for this series requires non-trivial considerations
in general, and it is hence useful not to insist in the a priori existence of
function b generating the coefficients in this way.
α 1Q
X
1,α 2,α
Λf := πQ hf iQ hα
Q + π Q hf, h Q i . (12.24)
|Q|
Q∈D
α∈{0,1}d \{0}
Of course this covers both Π1 and Π2∗ as special cases, by simply setting some
i,α
of the coefficients πQ equal to zero.
Compared to the operator Λb featuring in Proposition 12.2.3, we now allow
1,α 2,α
possibly different coefficients πQ and πQ in the first and second term above,
as this will be relevant in the T (1) theorem. Via the duality relations
which has exactly the same form as Λ, only with different coefficients, and the
associated bilinear form
X D E D E
1,α 2,α
L(f, g) := πQ hf iQ , hhα α
Q , gi + πQ hf, hQ i, hgiQ . (12.26)
Q∈D
α∈{0,1}d \{0}
In particular, we have
f = x ⊗ hβP , g = y ∗ ⊗ 1R
which is clearly a finite sum. When P = R, the sum above is void, and we get
β 2,β
L(x ⊗ 1R , y ∗ ⊗ hR ) = hπR x, y ∗ i.
On the other hand, assuming the coefficient bound (12.27), the defining series
(12.26) of L(f, g) converges absolutely for all
Proof. We have
1,α
|hπQ x, y ∗ i| = |L(x ⊗ 1Q , y ∗ ⊗ hα
Q )|
and taking the supremum over ky ∗ kY ∗ 6 1 and kxkX 6 1 proves the estimate
for i = 1. The case i = 2 is entirely symmetric. Finally, note that 1/p, 1/q ∈
(0, 1] so that 1/p − 1/q < 1.
To prove the convergence, it is enough to consider f = x⊗1P , g = y ∗ ⊗1R ,
1,α
and moreover, by symmetry, just the first half of L (f, g) with coefficients πQ .
Now
1,α 1,α ∗
|hπQ hf iQ , hg, hα α
Q ii| = |hπQ x, y i|h1P iQ |h1R , hQ i|,
128 12 Dyadic operators and the T (1) theorem
where
1,α |P | |R|
|hπQ x, y ∗ i| 6 C|Q|γ kxkX ky ∗ kY ∗ , h1P iQ 6 , |h1R , hα
Q i| 6 ,
|Q| |Q|1/2
and moreover the last pairing is non-zero only if Q ) R. Hence the absolute
convergence of the series follows from the convergence of
X ∞
X
|Q|γ−3/2 = |R|γ−3/2 2kd(γ−3/2) < ∞,
Q∈D k=1
Q)R
Lemma 12.2.6. Suppose that the series defining Λf converges (even just con-
ditionally) in Lp (Rd ; Y ) for some f = 1R ⊗ x, where R ∈ D and x ∈ X. Then
X 1,α
(1R − ER )Λ(1R ⊗ x) = πQ x ⊗ hα
Q
Q⊆R
α∈{0,1}d \{0}
Proof. We have
(
1,α
1,α 2,α 1Q πQ x ⊗ hα
Q + 0, Q ⊆ R,
1R πQ h1R ⊗ xiQ hα
Q + πQ h1R ⊗ x, hα
Qi = α
|Q| yQ,R ⊗ 1R , Q 6⊆ R,
α
for some yQ,R ∈ Y , which is not difficult to find explicitly, but it is irrelevant
for the present purposes. The assumed convergence in Lp (Rd ; Y ), and the
boundedness of the conditional expectation ER and the pointwise multiplier
1R on Lp (Rd ; Y ) guarantee that we can move (1R − ER ) inside the defining
α α
series. Since ER (yQ,R ⊗ 1R ) = yQ,R ⊗ 1R , we have
X 1,α
(1R − ER )Λ(1R ⊗ x) = πQ x ⊗ hα
Q,
Q⊆R
α∈{0,1}d \{0}
as claimed.
α
Lemma 12.2.7. Let Y be a Banach space, and p ∈ [1, ∞). Let yQ ∈ Y for
all Q ∈ D, α ∈ {0, 1} \ {0}. For each R ∈ D and n ∈ N, consider the sum
d
X
n α
BR := yQ ⊗ hα
Q
Q⊆R
`(Q)>2−n `(R)
α∈{0,1}d \{0}
n
(2) Y has the Radon–Nikodým property, and supn∈N kBR kLp (Rd ;Y ) < ∞.
p
Then there exists a function b ∈ Lloc (Rd ; Y ) such that
1R (b − hbiR ) = BR , α
hb, hR α
i = yR , ∀R ∈ D, α ∈ {0, 1}d \ {0}.
If, moreover, the supremum below is finite, then b ∈ BMOD (Rd ; Y ) and
α
kyQ kY kBR − ckLp (Rd ;Y )
sup 6 kbkBMOpD (Rd ;Y ) = sup inf (12.28)
Q∈D |Q|1/2 R∈D c∈Y |R|1/p
α∈{0,1}d \{0}
n ∞
Proof. It is immediate to verify that (BR )n=0 is a martingale in Lp (Rd ; Y ).
By the Martingale Convergence Theorem 3.3.16, it follows that (2) implies
(1). Hence it suffices to prove the lemma under assumption (1).
We construct the function b via the correspondence established in Lemma
11.2.11. It is enough to construct b|S separately for each quadrant S ⊆ Rd .
So we fix a quadrant S ⊆ Rd , and let
X
∆(s, t) := (hα α α
Q (s) − hQ (t))yQ ,
Q∈D(S)
α∈{0,1}d \{0}
where we need to justify the convergence of this series in some sense. We will
prove that it converges in Lploc (S × S; Y ). To this end, note that any bounded
subset of S × S is contained in R × R for some R ∈ D(S). For s, t ∈ R, only
Q ∈ D(S) with Q ∩ R 6= ∅ can contribute to the series; moreover, if Q ) R,
then hα α α
Q is constant on R, and hence hQ (s) − hQ (t) = 0 for s, t ∈ R. Thus
X
(1R×R ∆)(s, t) = 1R×R (s, t) (hα α α
Q (s) − hQ (t))yQ
Q∈D(R)
α∈{0,1}d \{0}
(12.29)
= 1R×R (s, t)(BR (s) − BR (t)),
and hence b(·) = BR (·) + (b(t) − BR (t)) ∈ Lp (R; Y ) ⊆ L1 (R; Y ). Taking the
average over t ∈ R, it follows that
observing that hhα Q iR = 0 for all Q ⊆ R that appear in the series of BR . Then
it also follows that
hb, hα α α α
R i = h1R (b − hbiR ), hR i = hBR , hR i = yR .
and (12.28) follows from the identity BQ = 1Q (b − hbiQ ), which implies that
= sup inf
0
k1Q (BQ − c0 )kLp (Q;Y )
Q∈D c ∈Y
kb1 kBMOD,so
p
(Rd ;Y ) 6 TΛ .
While rather far from being a sufficient condition for any interesting bounded-
ness results, this weak coefficient bound nevertheless allows us to make sense
of the defining series of the paraproduct on a sufficiently rich class of functions
for our subsequent purposes.
We have the following useful convergence result for truncated paraproducts:
|hm Πf, gi| 6 cd,p CkEm f kLp (Rd ;X) kEm gkLp0 (Rd ;Y ∗ ) −→ 0, (12.32)
m→−∞
2d − 1
where C is the constant in (12.30) and cd,p = .
1 − 2−d/p0
Proof. Let us first consider (2). When 2−m > diam(supp f ), the support
supp f is contained in at most 2d cubes Ri ∈ D. Then in m Πf , we only
need to consider Q ∈ D with Q ) Ri for some (not necessarily unique)
i = 1, . . . , 2d . Then
d
2
X X
α
km Πf kLp (Rd ;Y ) = kπQ hf iQ hα
Q kLp (Rd ;Y ) ,
i=1 Q∈D
Q)Ri
α∈{0,1}d \{0}
where
α
kπQ hf iQ hα α α
Q kLp (Rd ;Y ) 6 kπQ kL (X,Y ) khf iQ kX khQ kLp (Rd )
|Q|1/p
Z Z
α 1 C
= kπQ kL (X,Y ) f 6 f ,
|Q| Ri X |Q|1/2 |Q|1/p0 Ri X
and hence
d
2 Z
X X 1
km Πf kLp (Rd ;Y ) 6 (2d − 1)C f
i=1 Ri X
Q)Ri
|Q|1/p0
d
2 ∞
(2d − 1)C
Z
X X 0
= 1/p0
f 2−kd/p
i=1
|Ri | Ri X
k=1
d
2
2d − 1 X Z
= C |Ri |1/p − f ,
2d/p0−1 i=1 Ri X
where
d
2 Z 2d
X Z p 1/p
d/p0 0
X
1/p
|Ri | − f 62 |Ri | − f = 2d/p kEm f kLp (Rd ;X) .
Ri X Ri X
i=1 i=1
This proves both the convergence of the series and the claimed bound (12.31).
12.2 Paraproducts 133
|hm Πf, gi| = |hm Πf, Em gi| 6 km Πf kLp (Rd ;Y ) kEm gkLp0 (Rd ;Y ∗ ) ,
We will then turn to exploring conditions that ensure the boundedness of the
full paraproduct Π. The obtained necessary conditions serve as a model for
the type of sufficient conditions that we are looking for.
It is convenient to begin with a reduction to finite series. When Y is
reflexive, we have
0
Lp (Rd ; Y ) = Lp (Rd ; Y ∗∗ ) ' (Lp (Rd ; Y ∗ ))∗ .
0
Since S00 (D; Y ∗ ) is dense in Lp (Rd ; Y ∗ ), it is enough to show that the action
0
of Πf is bounded on S00 (D; Y ∗ ) with respect to the norm of Lp (Rd ; Y ∗ ),
uniformly for f in the unit ball of Lp (Rd ; X). Since any fixed g ∈ S00 (D; Y ∗ )
PΠfα, it is enough
only “sees” a finite part of to prove a uniform Lp (Rd ; Y ) esti-
α
mate for the finite sums πQ hf iQ hQ . A key initial estimate in this direction
is the following:
Lemma 12.2.14. Let X be a Banach space, Y be a UMD space, and p ∈
(1, ∞). Let F be a finite collection of dyadic cubes. For all f ∈ Lp (Rd ; X)
and πQα
∈ L (X, Y ), we then have
X X
α − +
πQ hf iQ hα
Q p d
6 β p,Y β p,Y ε Q π α 0
Q h Q f p d .
L (R ;Y ) L (R ;Y )
Q∈F Q∈F
α
Proof. Since (πQ hf iQ hα p d
Q )Q∈F is a martingale difference sequence in L (R ; Y ),
we have
X X
α α − α
πQ hf iQ hQ 6 βp,Y ε Q πQ hf iQ hα
Q .
Lp (Rd ;Y ) Lp (Rd ×Ω;Y )
Q∈F Q∈F
Rewriting the Lp norm on the product Rd × Ω with the help of Fubini’s the-
orem, we observe that at each fixed t ∈ Rd , the sequence of random variables
α
ε Q πQ hf iQ hα
Q (t)
12.2 Paraproducts 135
The previous lemma motivates the following. A background for the nomen-
clature will be discussed in the Notes.
Definition 12.2.15. Let p ∈ (1, ∞). For an indexed family (πQ )Q∈D in a
Banach space Z, we define the Carleson norm
1 X
k(πQ )kCarp (Rd ;Z) := sup sup εQ h0Q πQ .
Q0 ∈D F ⊆D |Q0 |1/p Q⊆Q0
Lp (Q0 ×Ω;Z)
finite
Q∈F
With the help of Theorem 3.2.17 (the John–Nirenberg inequality for adapted
sequences), one can check that any these Carleson norms are actually equiv-
alent for different values of p. We will not need this observation, since the
following proof directly shows that we can use any of these norms in our upper
bound, as we like. Our first sufficient condition for paraproduct boundedness
is stated in terms of this notion as follows:
for f ∈ S00 (Rd ; X) and g ∈ S00 (Rd ; Y ∗ ); the latter guarantees that the sum
α α
is finitely nonzero. We may thus replace πQ by 1F (Q)πQ for some finite set
F ⊆ D, but we do not indicate this explicitly in the notation.
Let P0 be the maximal cubes appearing in this sum. We then construct cube
families Pn inductively as follows. For each P ∈ Pn , let chP (P ) be the
maximal dyadic subcubes P 0 of F such that either
Z Z Z Z
− kf kX > 4− kf kX or − kgkY ∗ > 4− kgkY ∗ .
P0 P P0 P
1 n R kf k R kgk ∗ o
0 X 0 Y
|P 0 | 6 max RP , RP .
4 − kf kX − kgkY ∗
P P
136 12 Dyadic operators and the T (1) theorem
Thus [ 1
EP (P ) := P \ satisfies |EP (P )| > |P |.
2
P 0 ∈chP (P )
Then we let
[ ∞
[
Pn+1 := chP (P ), P := Pn ,
P ∈Pn n=0
For P ∈ P, let
X
PP h := 1P 0 hhiP 0 + 1EP (P ) h.
P 0 ∈chP (P )
0
Since both 1Q and hα
Q are linear combination of Q ∈ chD Q, this implies in
particular that
hf iQ = hPP f iQ , hhα α
Q , gi = hhQ , PP gi, parP Q = P.
With the principal cubes P ∈ P just constructed, we can now rearrange the
sum (12.33) as
X D X E X
hΠ α f, gi = πQα
hPP f iQ hα
Q , PP g =: IP .
P ∈P Q∈D P ∈P
parP Q=P
By Lemma 12.2.14 at the key step introducing the UMD constants, and ap-
plications of Hölder’s inequality and the properties of the principal cubes
elsewhere,
X
α α
IP 6 πQ hPP f iQ hQ kPP gkLq0 (Rd ;Y ∗ )
Lq (Rd ;Y )
Q∈D
parP Q=P
X
− + α 0
6 βq,Y βq,Y ε Q πQ h Q PP f kPP gkLq0 (Rd ;Y ∗ )
Lq (Rd ×Ω;Y )
Q∈D
parP Q=P
X
− + α 0
6 βq,Y βq,Y ε Q πQ hQ kPP f kL∞ (Rd ;X) ×
Lq (Rd ×Ω;L (X,Y ))
Q∈D
parP Q=P
0
× kPP gkL∞ (Rd ;Y ∗ ) |P |1/q
− + α
6 βq,Y βq,Y k(πQ )kCarq (Rd ;L (X,Y )) |P |1/q × 4 · 2d hkf kX iP ×
0
× 4 · 2d hkgkY ∗ iP |P |1/q
− +
= 16 · 4d · βq,Y βq,Y α
k(πQ )kCar2 (Rd ;L (X,Y )) hkf kX iP hkgkY ∗ iP |P |
− +
=: 16 · 4d · βq,Y βq,Y α
k(πQ )kCar2 (Rd ;L (X,Y )) × IIP .
(Note that, in the step that lead to the appearance of the Carleson norm,
α α
we made use of our implicit replacement of πQ by 1F (Q)πQ , for some finite
F ⊆ D, in the beginning of the proof.)
Finally,
138 12 Dyadic operators and the T (1) theorem
X X
IIP 6 2 hkf kX iP hkgkY ∗ iP |EP (P )|
P ∈P P ∈P
X Z
62 MD f · MD g
P ∈P EP (P )
Z
62 M D f · MD g
Rd
6 2kMD f kLp (Rd ) kMD gkLp0 (Rd )
6 2 · p0 kf kLp (Rd ;X) · pkgkLp0 (Rd ;Y ∗ ) ,
This estimate also has a converse, but since it has no immediate use in the
present discussion, we leave the details to an interested reader.
Proof. This is a direct computation
X
0 α
ε Q hQ hb, hQ i p
L (Q0 ×Ω;Z)
Q⊆Q0
Q∈F
X
α α
6 inf ε Q hQ h1Q0 (b − c), hQ i
c∈Z Lp (Q0 ×Ω;Z)
Q⊆Q0
Q∈F
+
6 inf βp,Z k1Q0 (b − c)kLp (Rd ;Z) by Proposition 12.1.5
c∈Z
+
6 βp,Z |Q0 |1/p kbkBMOpD (Rd ;Z) .
Proof. (1) ⇒ (3): The assumed boundedness (1) and duality clearly implies
the testing conditions
kΛ(1Q ⊗ x)kLp (Rd ,X) 6 kΛkL (Lp (Rd ;X)) k1Q ⊗ xkLp (Rd ;X) ,
kΛ (1Q ⊗ x∗ )kLp0 (Rd ,X ∗ ) 6 kΛkL (Lp (Rd ;X)) k1Q ⊗ x∗ kLp0 (Rd ;X ∗ ) .
∗
Condition (3) then follows from Proposition 12.2.8, which also provides the
bounds
max kb1 kBMOD p
(Rd ) , kb2 kBMOp0 (Rd ) 6 kΛkL (Lp (Rd ;X)) .
D
±
where we also used βp,X 6 βp,X . Similarly, we have
using the same bound on the dual side and recalling that βq0 ,X ∗ = βq,X .
(2) ⇒ (1): This is trivial by the triangle inequality.
(3) ⇔ (4): This is the already established equivalence (3) ⇔ (1) specialised
to X = K. The final quantitative bound follows by combining the bounds
already established:
2
X
kΛkL (Lp (Rd ;X)) 6 kΠ̃i kL (Lp (Rd ;X))
i=1
2
X
6 32 · 8d · pp0 βq,X
2
βq,K kbi kBMOqi (Rd ) ,
D
i=1
2
X
6 32 · 8d · pp0 βq,X
2
βq,K kΛkL (Lq (Rd ))
i=1
P2
and i=1 32 = 64.
In this section, we will take a closer look at the special case of the symmetric
i,α α
paraproduct Λb with equal coefficients πQ = hb, hQ i for both i = 1, 2. Our
goal is to obtain a qualitative improvement of the earlier Proposition 12.2.3.
This will require developing modest prerequisites about the projective tensor
product of Banach spaces, and we first turn to this task.
12.2 Paraproducts 141
Definition 12.2.20. For two Banach spaces X and Z, and a bilinear form
λ : X × Z → K, we define
n o
kλkB(X,Z) := sup |λ(x, z)| : kxkX 6 1, kzkZ 6 1 ,
n o
B(X, Z) := λ : X × Z → K bilinear kλkB(X,Z) < ∞ .
Proof. To check thatP hv, λi is well-defined, we need to verify that two different
na
representations v = i=1 xai ⊗ zia , a = 1, 2, result in the same right-hand side.
0 p
To see this, pick a basis (xj )j=1 for span{xia : 1 6 i 6 na , a = 1, 2} and a basis
(zk0 )qk=1 for span{zia : 1 6 i 6 na , a = 1, 2} and expand all xia and zia in the
respective basis. With the help of the Hahn–Banach theorem, pick x∗m ∈ X ∗
and zn∗ ∈ Z ∗ such that hx0j , x∗m i = δj,m and hzk0 , zn∗ , =iδk,n , and consider the
forms λm,n ( ·1 , ·2 ) = h ·1 , x∗m ih ·2 , zn∗ i ∈ B(X, Z) to see that x0j ⊗zk0 are linearly
independent in B(X, Z)∗ . Hence their coefficients must be equal in the two
expansions of v. Make the same expansions on the right-hand side, using the
bilinearity of λ, to find that both expansions lead to linear combinations with
equal coefficients of the values λ(xj0 , zk0 ).
Having verified that the action of λ on X ⊗ Z is well defined, its linearity
is clear. Moreover,
n
X n
X
|λ(xi , zi )| 6 kλkB(X,Z) kxi kX kzi kZ ,
i=1 i=1
for all v ∈ X ⊗ Z. From this estimate, we can uniquely extend the action of
λ to all v ∈ X ⊗Z
b by density, with the estimate
kλk(X⊗Z)∗ 6 kλkB(X,Z) .
On the other hand, we also have
|λ(x, z)| = |hx ⊗ z, λi| 6 kx ⊗ zkX ⊗Z
b kλk(X⊗Z)∗ 6 kxkX kzkZ kλk(X⊗Z)∗ ;
It is clear from the definition that |λ(x, z)| 6 kxkX kzkZ for any λ as in the
last supremum. On the other hand, the Hahn–Banach theorem guarantees the
existence of x∗ ∈ X ∗ and z ∗ ∈ Z ∗ of norm one such that hx, x∗ i = kxkX and
hz, z ∗ i = kzkZ . Then clearly λ( ·1 , ·2 ) = h ·1 , x∗ ih ·2 , z ∗ i has kλkB(X,Z) 6 1
and gives λ(x, z) = kxkX kzkZ .
12.2 Paraproducts 143
Theorem 12.2.25. Let X and Y be UMD spaces and p ∈ (1, ∞). For every
function b ∈ BMOD (Rd ; L (X, Y )), the symmetric paraproduct Λb defines a
bounded operator from Lp (Rd ; X) to Lp (Rd ; Y ) of norm
On the last line, we are using the H 1 –BMO-duality from Theorem 11.1.30;
for f ∈ S00 (D; X) and g ∈ S00 (D; Y ∗ ), the summation is finite, and thus h ∈
L∞c (R ; X ⊗π Y ). Since b ∈ BMOD (R ; L (X, Y )) ⊆ Lloc (R ; L (X, Y )), the
d b ∗ d 1 d
pointwise duality product hb(u), h(u)i is integrable, and one find by dominated
convergence in the defining formula of Theorem 11.1.30 that the duality can be
computed simply as the integral of hb(u), h(u)i over Rd . Thus, an application
of Theorem 11.1.30 followed by Theorem 11.1.28 shows that
khkHmax
1 (Rd ;X⊗Y ∗ ) = kMD hkL1 (Rd ) = sup 1R khhiR kX⊗Y ∗ .
R∈D L1 (Rd )
Thus
Xh i
hf iR = hf iQR ⊗ hgiQR − hf iQ ⊗ hgiQ
Q)R
X
+ (hf iQR − hf iQ ) ⊗ (hgiQR − hgiQ ) =: IR + IIR .
Q)R
The sum IR is telescopic and, since f ∈ S00 (D; X) (we don’t even need the
similar property of g at this point), its terms vanish for all large enough Q.
Thus in fact
IR = hf iR ⊗ hgiR , kIR kX ⊗
b π Y ∗ = khf iR kX khgiR kY ∗
and
sup 1R kIR kX ⊗
bπY ∗ 6 kMD f · MD gkL1 (Rd )
R∈D L1 (Rd )
Thus
X X
kIIR kX ⊗Y
b ∗ 6E εP DP f (u) εQ DQ g(u)
X Y∗
P )R Q)R
X X
6 εP DP f (u) εQ DQ g(u)
Lp (Ω;X) Lp0 (Ω;Y ∗ )
P )R Q)R
X X
6 εP DP f (u) εQ DQ g(u) ,
Lp (Ω;X) Lp0 (Ω;Y ∗ )
P ∈D Q∈D
where the last step was an application of the contraction principle. Thus
12.2 Paraproducts 145
X X
sup kIIR kX ⊗Y
b ∗ 6 εP DP f (u) εQ DQ g(u)
R3u Lp (Ω;X) Lp0 (Ω;Y ∗ )
P ∈D Q∈D
and
sup 1R kIIR kX ⊗Y
b ∗
R∈D L1 (Rd )
X X
6 ε P DP f ε Q DQ g
Lp (Rd ×Ω;X) Lp0 (Rd ×Ω;Y ∗ )
P ∈D Q∈D
+
6 βp,X kf kLp (Rd ;X) · βp+0 ,Y ∗ kgkLp0 (Rd ;Y ∗ ) .
khkHmax
1 b ∗)
(Rd ;X ⊗Y
6 sup 1R kIR kX ⊗Y
b ∗ + sup 1R kIR kX ⊗Y
b ∗
R∈D L1 (Rd ) R∈D L1 (Rd )
0 +
6 (pp + βp,X βp+0 ,Y ∗ )kf kLp (Rd ;X) kgkLp0 (Rd ;Y ∗ ) ,
and altogether we have proved the first estimate claimed in the theorem.
+
The final estimate is seen as follows: First, we have βp,X 6 βp,X and
+
βp0 ,Y ∗ 6 βp0 ,Y ∗ = βp,Y by the observation after Proposition 4.2.3, and Propo-
sition 4.2.17(2). Second, denoting p∗ = max(p, p0 ) > 2, we have βp,Z > βp,R =
p∗ − 1 > 21 p∗ by Theorem 4.5.7, and hence pp0 6 (p∗ )2 6 4βp,X βp,Y .
kΠb kL (L2 (R;`2N )) 6 φ(N )kbkL∞ (R;L (`2N )) for all b ∈ L∞ (R; L (`2N )).
Then
1
φ(N ) > k4N kL (L (`2N )) > (log N − 1),
π
where 4N : L (`2N ) → L (`2N ) is the lower triangle projection defined by
(
ei ⊗ ej , if i > j,
4N (ei ⊗ ej ) :=
0, else
where suggestive notation Π⊗g is defined by the last identity. In the two right-
most expressions, the duality is that between L∞ (R; L (`2N )) and its predual
L1 (R; C 1 (`N
2
)). (We recall from Theorem D.2.6 that (C 1 (H))∗ = L (H) for
any Hilbert space H and from Theorem 1.3.10 that (L1 (R; X))∗ = L∞ (R; X ∗ )
when X ∗ has the Radon–Nikodým property, which the finite-dimensional
(hence reflexive) X = L (`2N ) does by Theorem 1.3.21.)
Thus we deduce that
n o
kΠ⊗g f kL1 (R;C 1 (`N
2 )) = sup |hb, Π⊗g f i| : kbkL∞ (R;L (`2N )) 6 1
n o
= sup |hΠb f, gi| : kbkL∞ (R;L (`2N )) 6 1
2 ) kgkL2 (R;`2 ) .
6 φ(N )kf kL2 (R;`N N
Then
j−1
N X
X
Π⊗g f (t) = ri (t)hu, ei iei ⊗ rj (t)hv, ej iej
j=1 i=1
X
= Dr(t) hu, ei ihv, ej iei ⊗ ej Dr(t)
16i<j6N
N
X
= Dr(t) TN hu, ei ihv, ej iei ⊗ ej Dr(t) = Dr(t) TN (u ⊗ v) Dr(t)
i,j=1
PN
where Dr(t) = i=1 ri (t)ei ⊗ ei and 4̃N is the upper triangle projection
defined by (
ei ⊗ ej , if i < j,
4̃N (ei ⊗ ej ) :=
0, else
and extended by linearity. Since Dr(t) is unitary for every t ∈ [0, 1), it follows
that
2 )) = k4̃N (u ⊗ v)kC 1 (`2 ) .
kΠ⊗g f kL1 (R;C 1 (`2N )) = k4̃N (u ⊗ v)kL1 ([0,1);C 1 (`N N
Dy Lemma D.1.1 and the definition of the Schatten class, every s ∈ C 1 (`N
2
)
has a singular value decomposition
12.3 The T (1) theorem for abstract bilinear forms 147
n
X n
X
s= ak (s)uk ⊗ vk , 2 = kvk k`2 = 1,
kuk k`N N
ak (s) = kskC 1 (`2N )
k=1 k=1
k=1
Xn
6 ak (s)φ(N ) = φ(N )kskC 1 (`2N ) .
k=1
was considered instead. However, the lower bound for the norm of this operator
was achieved by testing with the Hilbert matrix AN = (1{i6=j} (i − j)−1 )Ni,j=1
with vanishing diagonal; hence ∆N (AN ) = TN (AN ), and the same lower
bound follows for ∆N as well.
t : S(D; X) × S(D; Y ∗ ) → K
by letting
Remark 12.3.3 (S(D) vs. S00 (D) in the definition). Since S00 (D; X) is al-
ready dense in Lp (Rd ; X), in order to construct a bounded bilinear form on
0
Lp (Rd ; X) × Lp (Rd ; Y ), it would be sufficient to have an a priori estimate on
S00 (D; X) × S00 (D; Y ∗ ). However, for the type of theorems that we have in
mind, we also like to make assumptions on the action of our bilinear forms
on functions like 1Q ∈ S(D) \ S00 (D), and hence we need to have our initial
bilinear form defined on the larger product S(D; X) × S(D; Y ∗ ). This gives
rise to the following problem, where we take X = Y = K for simplicity, since
the issue is already present in this case:
Suppose that we have a bilinear form t : S(D)2 → K that satisfies the
estimate
12.3 The T (1) theorem for abstract bilinear forms 149
If (f, g) ∈ S00 (D)2 , we have the a priori bound |t(f, g)| = 0, and hence the
unique operator T ∈ L (Lp (Rd )) is given by T = 0. But of course t is not
identically zero on S(D)2 . It is also clear that there cannot possible be any
T ∈ L (Lp (Rd )) with hT f, gi = t(f, g) for all (f, g) ∈ S(D)2 .
To avoid this problem, we make sure to get our a priori estimates on the
full set S(D; X) × S(D; Y ∗ ).
t : S(D; X0 ) × S(D; Y 0 ) → K.
Let C > 0 be a constant and p ∈ (1, ∞). Then the following conditions, each
to hold for every choice of (f, g) ∈ S(D; X0 ) × S(D; Y 0 ), are equivalent:
(1) There is T ∈ L (Lp (Rd ; X), Lp (Rd ; Y )) of norm at most C such that
hT f, gi = t(f, g).
0 0
(2) There is T ∗ ∈ L (Lp (Rd ; Y ∗ ), Lp (Rd ; X ∗ )) of norm at most C such that
α
converges absolutely. We say that t(·, 1) is well-defined if t(hQ , 1) is well-
defined for every Q ∈ D and α ∈ {0, 1}d \ {0}.
We define t(1, hα
Q ) and t(1, ·) analogously.
α
Lemma 12.3.7. If t(hQ , 1) is well-defined, then
(1) for every k ∈ Z with 2−k > `(Q), we have
X
α α
t(hQ , 1) = t(hQ , 1R ),
R∈Dk
where the first equality holds by assumption, and the assumed absolute con-
vergence allows to make the rearrangements and to get the absolute conver-
gence also in the subsequent steps.
(2): Each f ∈ S00 (D) is a linear combination of terms of the form hαQi ,
i
where i ∈ F for some finite index set F . If Q0 ∈ Dj0 is the largest cube
appearing here, then by the previous part of the lemma we know that
X
t(hα
Qi , 1R )
i
R∈Dk
converges absolutely. If the absolute convergence holds for some j and k, the
equality of the corresponding series follows from (12.34).
The case of t(1, ·) is entirely analogous.
As we shall see later, the forms t(1, ·) and t(·, 1) are closely related to
paraproducts. Since the boundedness of paraproducts is tricky, it is use-
ful to be able identify situations, when they can be avoided, i.e., when
t(1, ·) = 0 = t(·, 1).
With this goal in mind, we will now discuss an important case of trans-
lation-invariant bilinear forms. We first check that some natural candidates
for the definition are equivalent:
152 12 Dyadic operators and the T (1) theorem
Lemma 12.3.8. Let Z be a Banach space. The following conditions are equiv-
alent for a bilinear form t : S(D)2 → Z:
(1) t(1Q , 1R ) = t(1Q+̇m , 1R+̇m ) for all Q, R ∈ D with `(Q) = `(R), and all
m ∈ Zd , where Q+̇m := Q + m`(Q).
(2) t(f, g) = t(τh f, τh g) for all f, g ∈ S(D) and all dyadic rational vectors
h, i.e., all h of the form h = m2−k for some m ∈ Zd and k ∈ Z, where
τh f (s) := f (s − h).
If Z = L (X, Y ), these are also equivalent to a variant of (2) for all f ∈
S(D; X) and g ∈ S(D; Y ∗ ) instead.
−1/2
where the coefficients hhα
Q iQ1 +̇γ are equal to ±|Q| , with equally many of
each sign. Now, formally, we have
and hence
X
α
“ (1, hQ )= hhα
Q iQ1 +̇γ t(1, 1Q1 +̇γ )
γ∈{0,1}d
X
α
= hhQ iQ1 +̇γ t(1, 1Q1 ) = 0 · t(1, 1Q1 ) = 0. ”
γ∈{0,1}d
where rearranging the order of the finite sums inside the limit presents no
issues. Here
On the other hand, we have n ∈ [−(2M +1), 2M +1]d \[−(2M −1), (2M −1)]d ,
and the total number of such n ∈ Zd is
154 12 Dyadic operators and the T (1) theorem
(1 + 2(2M + 1))d − (1 + 2(2M − 1))d = (4M + 3)d − (4M − 1)d
6 4d(4M + 3)d−1 ,
and hence
β−γ
kIIM k 6 4d(4M + 3)d−1 × cQ1 (1 + 2M )−d 6 cd cQ1 M −1 .
Remark 12.3.11. It is easy to see from the proof that the decay assumption
(12.35) could be somewhat weakened. We have not strived for maximal gen-
erality at this point, but stated a condition that is both relatively simple to
formulate and easy to verify in our main application to Calderón–Zygmund
singular integrals.
X X X
Dk f = Ek+1 f − Ek f = EQ 0 f − EQ f = DQ f.
Q∈Dk Q0 ∈ch(Q) Q∈Dk
Our starting point for the analysis of a bilinear form is the following useful
identity:
where
t(Ek+1 f, Ek+1 g) = t((Dk + Ek )f, (Dk + Ek )g)
= t(Dk f, Dk g) + t(Dk f, Ek g) + t(Ek f, Dk g) + t(Ek f, Ek g),
and hence
t(Ek+1 f, Ek+1 g) − t(Ek f, Ek g)
= t(Dk f, Dk g) + t(Dk f, Ek g) + t(Ek f, Dk g).
Remark 12.3.13. The upper bound k < M imposed on the summation vari-
ables above is redundant: the condition that f and g are constant on all
Q ∈ DM implies that Dk f = 0 = Dk g for k > M , so that the right side would
remain unchanged if we allow the summations to run to infinity.
The final term in the expansion 12.36 is an error term, and can be controlled
under the following mild conditions, which are obviously necessary for t to
define a bounded operator on Lp :
|t(Em f, Em g)| 6 2d ktkawbp kEm f kLp (Rd ;X) kEm gkLp0 (Rd ;Y ∗ ) −→ 0.
m→−∞
and thus
X
|t(Em f, Em g)| 6 kt(1Q , 1R )kL (X,Y ) khf iQ kX khgiR kY ∗
Q,R∈Dm
X
6 ktkawbp |Q|khf iQ kX khgiR kY ∗
Q,R∈Dm
X X 0
= ktkawbp |Q|1/p khf iQ kX |R|1/p khgiR kY ∗
Q∈Dm R∈Dm
0
X 1/p X 0
1/p0
p
6 ktkawbp 2d/p |Q|khf iQ kX 2d/p |R|khgiR kpY ∗
Q∈Dm R∈Dm
d
= 2 ktkawbp kEm f kLp (Rd ;X) kEm gkLp (Rd ;Y ∗ ) ,
which is the claimed bound.
The other terms in (12.36) can be identified with the various operators that
we have studied in the previous sections:
Definition 12.3.16. Let X, Y be Banach spaces, let t : S(D)2 → L (X, Y )
be a bilinear form, and let t(·, 1) and t(1, ·) be well-defined. We define the
following operators associated with t:
X αγ
Ht := Htα,γ , where Hαγtα,γ
are Haar multipliers (12.3),
0 0
α,γ
X
Tn,t := Tφαγ α,γ ,
n ,tn
where Tφαγ α,γ are Figiel’s operators (12.14)
n ,tn
α,γ∈{0,1}d \{0}
(
φn (Q) := Q+̇n := Q + n`(Q),
with γ
tα,γ α
n (Q) := t(hQ , hQ+̇n ),
X
i
Un,t := Uφα i,α , where Uφα i,α are Figiel’s operators (12.19),
n ,un n ,un
α∈{0,1}d \{0}
(
tn1,α (Q)∗ := t(h0Q+̇n , hα ∗
Q) , i = 1,
with uni,α (Q) :=
t2,α
n (Q) := α
t(hQ 0
, hQ +̇n
), i = 2.
We also define the related paraproducts:
Πt1 := paraproduct with coefficients t(1, hα
Q ),
∗
Πt2 := paraproduct with coefficients t(hα
Q , 1) ,
α,1 α,2
Λt := bi-paraproduct with coefficients πQ = t(1, hα α
Q ) and πQ = t(hQ , 1),
i
Remark 12.3.17. Our indexing of the operators Un,t may appear counterin-
2
tuitive at first sight, as one might like to think of the operators Un,t , which
act on f ∈ L (R ; X) with coefficients t(hQ , hQ+̇n ) ∈ L (X, Y ), as deserv-
p d α 0
1
ing to be the “primary” ones, rather than Un,t , which act on the dual side
p0
Q ) ∈ L (Y , X ). How-
∗ ∗ ∗ ∗
g ∈ L (R ; Y ) with adjoint coefficients t(hQ+̇n , hα
d 0
i
ever, this indexing is chosen, since the operators Un,t naturally arise in par-
allel with the paraproducts Πi of the same index i ∈ {1, 2}—see (12.42) and
(12.43) below—, and it turns out to have some other advantages in the sequel.
With this notation, we can formula Figiel’s decomposition of a bilinear form:
Proposition 12.3.18 (Figiel). Let X, Y be Banach spaces, let t : S(D)2 →
L (X, Y ) be a bilinear form, and let t(·, 1) and t(1, ·) be well-defined. For all
f ∈ S(D; X), g ∈ S(D; Y ∗ ), m ∈ Z,
denoting
u := (I − Em )f ∈ S00 (D; X), v := (I − Em )g ∈ S00 (D; Y ∗ ),
we have the following identity with absolute convergence:
t(f, g) = hHt u, gi + hΠt1 f, vi + hu, Πt2 gi + t(Em f, Em g)+
X n o
1 2
+ hTn,t u, gi + hf, Un,t vi + hUn,t u, gi , (12.38)
d
n∈Z
n6=0
where the operators on the right are as in Definition 12.3.16. If these coeffi-
cients satisfy
α
kt(1, hQ )k, kt(hα
Q , 1)k 6 C|Q|
1/2
, (12.39)
then we have the further identity, with all terms below well defined:
hΠt1 f, vi + hu, Πt2 gi = hΛt f, gi − hm Πt1 f, gi − hf, m Πt2 gi. (12.40)
Remark 12.3.19. Since Hλαγ = Tφαγ 0 ,λ
, we could have incorporated the Haar
multiplier into the second line of (12.38) as hHt u, gi = hT0,t u, gi. But we
prefer to keep it separate, since its treatment will involve some differences
compared to the rest of the Tn,t .
Proof of Proposition 12.3.18. We start with the identity (12.36) of Lemma
12.3.12. Since the sums are finitely nonzero, we are free rearrange as follows,
observing that dyadic cubes Q, R of the same size are necessarily integer (times
side-length) translates of each other:
X X X
t(Dk f, Dk g) = t(DQ f, DR g)
k>m k>m Q,R∈Dk
X X X X X
= t(DQ f, DQ+̇n g) = t(DQ f, DQ+̇n g)
k>m Q∈Dk n∈Zd Q∈D n∈Zd
`(Q)62−m
158 12 Dyadic operators and the T (1) theorem
and we can also switch the order of the last two sums. Observing that u =
(I −Em )f satisfies DQ u = DQ f for `(Q) 6 2−m and DQ u = 0 for `(Q) > 2−m ,
we find that, replacing f by u (and/or g by v) we can drop the restriction
`(Q) 6 2−m in the sum. Moreover, using the convention that summations
over α and γ are always over the set {0, 1}d \ {0},
X XXD γ γ
E
t(DQ u, DQ+̇n g) = t(hα α
Q , hQ+̇n )hhQ , ui, hhQ+̇n , gi
Q∈D α,γ Q∈D
X
= hTφα,γ αγ u, gi = hTn u, gi.
n ,tn
α,γ
Hence
X X X
t(Dk f, Dk g) = hTn u, gi = hHu, gi + hTn u, gi
k>m n∈Zd
(12.41)
n∈Zd
n6=0
For the terms involving Ek , we begin in the same way but then introduce
an additional twist to force some cancellation:
X X X
t(Dk f, Ek g) = t(DQ f, EQ+̇n g)
k>m Q∈D n∈Zd
`(Q)62−m
X X
= t(DQ f, 1Q+̇n (hgiQ+̇n − hgiQ )) + t(DQ f, 1Q+̇n hgiQ )) .
Q∈D n∈Zd
`(Q)62−m
Recalling that only finitely many DQ f with `(Q) 6 2−m are non-zero, we also
get the absolute convergence of
X X X
t(DQ f, 1Q+̇n hgiQ ) = t(DQ f, hgiQ ) =: pm (f, g),
Q∈D n∈Zd Q∈D
`(Q)62−m `(Q)62−m
where adding the summation condition n 6= 0 was for free, since the factor
hgiQ+̇n − hgiQ evidently vanishes when n = 0. Again, replacing f by u allows
us to drop the restrictions to `(Q) 6 2−m both in the sum spelled out above
and in pm (f, g). Moreover,
X
t(DQ u, 1Q+̇n (hgiQ+̇n − hgiQ ))
Q∈D
XXD E
α
= t(hQ α
, h0Q+̇n )hhQ 0
, ui, hhQ 0
+̇n − hQ , gi
α Q∈D
X
= hUφα α,0
2
u, gi = hUn,t u, gi.
n ,tn
α
In the computation above, the fact that u ∈ S00 (D; X) guarantees that all
summations are finite, and the last step is simply the definition of the para-
product via its action of the finitely non-zero Haar expansions in the dual
space. Hence we have verified that
X X
2
t(Dk f, Ek g) = hUn,t u, gi + hu, Πt2 gi, (12.42)
k>m n∈Zd
n6=0
is entirely analogous. Substituting the previous two identities and (12.41) into
(12.36), we obtain the claimed (12.38).
Under the additional assumption (12.39), we know from Corollary 12.2.12
that hΠt1 f, gi is well-defined and bilinear in (f, g) ∈ S(D; X) × S(D; Y ∗ ), and
hence
hΠt1 f, vi = hΠt1 f, gi − hΠt1 f, Em gi = hΠt1 f, gi − hm Πt1 f, gi.
Similarly, hu, Πt2 gi = hf, Πt2 gi − hf, m Πt2 gi, and the previous two identities
combine to give (12.40), noting that hΠt1 f, gi + hf, Πt2 gi = hΛt f, gi.
160 12 Dyadic operators and the T (1) theorem
Lemma 12.3.21. Let X and Y be Banach spaces and p ∈ [1, ∞). Then each
of the following ℘ is a good set-bound on Z = L (X, Y ):
(a) ℘ = U , where U (T ) := sup{kT k : T ∈ T },
(b) ℘ = Rp , the R-bound of order p,
(c) ℘ = Rp∗ , the dual R-bound defined by
Rp∗ (T ) := Rp (T ∗ ), T ∗ := {T ∗ ∈ L (Y ∗ , X ∗ ) : T ∈ T }.
where the first estimate is Proposition 8.1.19(3) and the second is immediate
from Kahane’s contraction principle (cf. the discussion right before Defini-
tion 8.1.1 of R-boundedness). Finally, properties (4) and (5) are contained in
Propositions 8.1.21 and 8.1.22, respectively.
(c): All properties are direct corollaries of the corresponding properties
in (b), since all set operations involved in these properties are well-behaved
under the adjoint operation:
(1) S ⊆ T if and only if S ∗ ⊆ T ∗ ,
12.3 The T (1) theorem for abstract bilinear forms 161
(2) (S ∪ T )∗ = S ∗ ∪ T ∗ and (S + T )∗ = S ∗ + T ∗ ,
(3) if Z ⊆ K, then (Z T )∗ = Z T ∗ ,
(4) (conv T )∗ = conv(T ∗ ) and (abs conv T )∗ = abs conv(T ∗ ),
(5) ( T )∗ = T ∗ .
Remark 12.3.23. Referring to Proposition 12.3.18, one observes that the Figiel
γ
norms impose control on pairings t(hα Q , hQ ), where at least one of the Haar
functions is cancellative, i.e., (α, γ) 6= (0, 0). This is in contrast to the decay
condition (12.35), where α = γ = 0.
∗
Since we also encountered the adjoint function u1,α 1,α
n (Q) := (tn (Q)) , we recall
the following results from the previous volumes:
In particular, if both X and Y are K-convex (resp. UMD spaces), the set-
bounds Rp and Rp∗0 are equivalent on L (X, Y ).
R∈D n∈Zd
`(R)=`(Q)
X
6 ktnα,0 (Q)k|Q|1/2 = kt2,α kFig0 (∞) |Q|1/2 < ∞,
n∈Zd
which shows both that t(·, 1) is well defined and the related bound. The case
of t(1, ·) is analogous.
X 2
X
DR p (tαγ
0 )+ kt(i) kFigσi (Rp ) < ∞,
α,γ i=0
while an inspection of (12.15) shows that C0,1 is larger than C1,1 in general.
(2) Similarly, with (t0 , q0 ) = (t2 , q2 ) = (1, q), we get
Using either choice (1) or (2) in Theorem 12.3.26, its key norm estimate admits
the following form, under the assumption (we recall) that X has cotype q and
Y has type t, or one of them has both,
nX
kT − Λt kL (Lp (Rd ;X),Lp (Rd ;Y )) 6 βp,X βp,Y DR p (t0α,γ )+
α,γ
2
X o
+ Ci Ad kt(0) kFigσi (℘i ) + Bd kt(i) kFigσi (℘i ) ,
i=1
Proof of Theorem 12.3.26. The core of the proof will consist of establishing
claims (a) and (b) under the full set of assumptions (i) through (iv). Assuming
that this is already done, let us see how to conclude the rest of the proof.
The equivalence of (1) and (2) is asserted under the assumptions (i)
through (iii) only. However, the adjacent weak boundedness property (iv)
is clearly necessary for (1) and it is explicitly assumed in (2), so we can as-
sume that this condition is satisfied in any case, and so we are in fact working
under the full set of assumptions (i) through (iv) also in this remaining part
of the proof. Thus the consequences (a) and (b) of this assumption are valid.
In particular, since the bilinear form t − l defines a bounded operator under
this assumption, it is clear that t defines a bounded operator if and only if l
does.
We then turn to the actual proof of (a) and (b) under the assumptions (i)
through (iv). From Lemma 12.3.25, we get that t(·, 1) and t(1, ·), and hence the
two paraproducts, are well defined, and their coefficients satisfy the bounds
(12.44). For f ∈ S(D; X) and g ∈ S(D; Y ∗ ), we then have both identities
(12.38) and (12.40) provided by Proposition 12.3.18. Combined together, they
read as
t(f, g) = hHum , gi + hΛf, gi + Em (f, g)+
X n o
+ hTn um , gi + hf, Un1 vm i + hUn2 um , gi , (12.46)
d
n∈Z
n6=0
12.3 The T (1) theorem for abstract bilinear forms 165
satisfies
2
X
|Em (f, g)| 6 cd,p kt(i) kFig0 (∞) + 2d ktkawbp ×
i=1 (12.47)
× kEm f kLp (Rd ;X) kEm gkLp (Rd ;Y ∗ ) −→ 0
m→−∞
by Lemmas 12.2.11 and 12.3.25 for the paraproduct terms and Lemma 12.3.15
for both the final term and the limit.
Directly from Theorem 12.1.11, we deduce that
X
|hHum , gi| 6 |hHtαγ
α,γ um , gi|
0
α,γ
X (12.48)
+
6 βp,X βp+0 ,Y ∗ DR p (tα,γ
0 )kum kp kgkp ,
α,γ
Note that φn (Q) := Q+̇n satisfies φn (Q) ⊆ 3Q(N ) provided that |n| 6 2N ;
thus in particular for N = dlog+ 2 |n|e; this is relevant in view of applying
Corollary 12.1.27 and Theorem 12.1.28. From Corollary 12.1.27, we deduce
that
X
|hTn um , gi| 6 |hTφαγ α,γ um , gi|
n ,tn
α,γ
using the notation of the statement of the theorem that we are proving. Hence
X
|hTn um , gi| 6 Ad βp,X βp,Y min C0,i kt(0) kFig1/t0 −1/q0 (℘i ) kum kp kgkp0
i=1,2
n∈Zd
n6=0
For the term hf, Un1 vm i, we again apply Theorem 12.1.28 but on the dual
side, with X, Y, p replaced by Y ∗ , X ∗ , p0 . By assumption, Y has type t1 6 p,
and hence Y ∗ has cotype t01 > p0 by Proposition 7.1.13. So we can indeed
apply Theorem 12.1.28 with X, Y, p, q replaced by Y ∗ , X ∗ , p0 , t01 . Recalling
that q1 := ∞, and noting that 1 − 1/t01 = 1/t1 = 1/t1 − 1/q1 , this gives
X
|hf, Un1 vm i| 6 |hf, Uφα ,(t1,α )∗ vm i|
n n
α
Noting that kum kp 6 2kf kp and kvm kp0 6 2kgkp0 , and using the assump-
tion about kt(i) kFig1/ti −1/qi (Rp ) (combined with Proposition 12.3.24 in the case
of Rp0 ((t1,α ∗
n ) )), it follows that the series in (12.46) are term-wise and uni-
formly in m dominated by absolutely convergent series. This allows us to pass
to the limit m → −∞ in (12.46) with dominated convergence to deduce that
D ω := Q+̇ω : Q ∈ D
(2) Conversely, every dyadic system D 0 has this form for some ω ∈ ({0, 1}d )Z .
Proof. Let D 0 be the standard dyadic system, and consider a family of shifts
sj + Dj0 . These clearly satisfy property (i) of Definition 11.1.6. A necessary
and sufficient condition for them to satisfy (ii) of Definition 11.1.6 is that
sj − sj+1 ∈ 2−j−1 Zd .
If D is a dyadic system defined by shifts sj , then D ω is defined by the
shifts sj + ω(j) , where
168 12 Dyadic operators and the T (1) theorem
X
ω(j) := ωk 2−k .
k>j
These satisfy (sj +ω(j) )−(sj+1 −ω(j+1) ) = (sj −sj+1 )+ωj+1 2−j−1 ∈ 2−j−1 Zd ,
and hence D ω is also a dyadic system, as claimed in (1).
Then suppose that D and D 0 are two dyadic systems defined by shifts sj
and s0j , respectively. It is clear that the family Dj = sj + Dj0 only depends on
sj mod 2−j , and hence we may assume without loss of generality that both
sj ∈ [0, 2−j )d and tj := sj0 − sj ∈ [0, 2−j )d . Since both sj − sj+1 ∈ 2−j−1 Zd
and s0j − sj+1
0
∈ 2−j−1 Zd , it follows that also tj − tj+1 ∈ 2−j−1 Zd . Together
with the fact that tj ∈ [0, 2−j )d and tj+1 ∈ [0, 2−j−1 )d , one finds that in
fact tj − tj+1 ∈ 2−j−1 {0, 1}d . Denoting ωj+1 := 2j+1 (tj − tj+1 ) ∈ {0, 1}d , we
obtain X
tj = tj+1 + 2−j−1 ωj+1 = . . . = 2−k ωk = ω(j) ,
k>j
and then
Dj0 = s0j + Dj0 = tj + sj + Dj0 = ω(j) + Dj = Djω ,
as claimed in (2), and this completes the proof.
Moreover, there exists an ω ∈ ({0, 1}d )Z0 such that S00 (D ω ) = S0 (D).
Definition 12.3.34. For θ ∈ {(α, γ), (i, α)}, and n ∈ Zd \ {0}, we define
ω;θ ω;θ
tn,good (R) := 1{R is k(n)-good in D ω } tn (R),
For n ∈ Zd \ {0}, we have k(n) > 2, and hence the notion of “k(n)-good” is
well-defined. For n = 0 we would formally get k(0) = −∞, and “−∞-good”
reduces to the triviality dist(R, {R) > 0; accordingly, for definiteness, we let
ω;θ
t0,good (R) := t0ω;θ (R).
ktω;θ
good kFig (℘) ,
s θ ∈ {(α, γ), (i, α)},
ω;(i)
ktgood kFigs (℘) , i = 1, 2, ktω
good kFigs (℘) .
As we are about to see, these good parts will suffice to control a bounded
extension of the form t, and this also allows us to obtain a better dependence
on the UMD constants. Here is the precise statement:
X 2
X
DR p (tω;α,γ
0 ) + min ktω;(0) kFigσi (Rp ) + ktω;(i) kFigσi (Rp ) 6 C,
i=1,2
α,γ∈{0,1}d \{0} i=1
where all symbols have the same meaning as in (12.46), but with D ω in place
of D. In particular,
uω ω
m = (I − Em )f,
ω
vm ω
= (I − Em )g,
ω ω
where Em = E( |Dm ) satisfy kuω ω
m kp 6 2kf kp and kvm kp0 6 2kgkp0 .
The first and third terms on the right of (12.53) are estimated as in the
proof Theorem 12.3.26. As in (12.47), we have
2
X
ω
|Em (f, g)| 6 cd,p ktω;(i) kFig0 (∞) + 2d ktω kawbp kEm
ω ω
f kp kEm gkp0 → 0
i=1
when m → −∞; note that this convergence is bounded by (iii), (iv), and the
ω ω
easy estimates kEm f kp 6 kf kp and kEm gkp0 6 kgkp0 . Then, as in (12.48),
from Theorem 12.1.11 we get
X
|hHtω uω
m , gi| 6 βp,X βp,Y DR p (t0ω;α,γ )kum
ω
kp kgkp0 .
α,γ
The second term on the right of (12.53) is directly estimated by the uniform
boundedness of the paraproducts Λtω .
We then turn to the more interesting part on the second line of (12.53),
where we begin with some observations. Due to the presence of the truncation
parameter m, all dyadic operators in (12.53) involve cubes of side-length at
most 2−m . On the other hand, due to the constancy of f and g on Q ∈ DM =
DM ω
, their martingale differences are non-zero only on cubes of side-length
strictly larger than 2−M . Hence the right-hand side of (12.53) actually depends
on (ωj )m<j6M only, rather than the infinite sequence (ωj )j6M . Nevertheless,
it will be convenient to also refer to this latter sequence, as we are about to
see.
We compute the expectation of (12.53) with respect to the choice of ω ∈
({0, 1}d )Z6M . As we just observed, this is actually just an arithmetic average
over a finite set of 2d(M −m) elements, so no integrability or measurability
issues arise at this point.
We wish to manipulate this average a little. We note that each of the terms
on the second line of (12.53) take the generic form
∗
X
Φ(Q+̇ω),
Q∈D
where
nXD E
γ γ
Φ(R) ∈ t(hα α
R , hR+̇n )hf, hR i, hg, hR+̇n i ,
α,γ
XD E
α
t(hR , h0R+̇n )hf, hα
R i, hg, h 0
R+̇n − h 0
R i ,
α
12.3 The T (1) theorem for abstract bilinear forms 173
XD γ γ
Eo
0 0 0
t(hR +̇n , hR )hf, hR+̇n − hR i, hg, hR i ,
γ
P∗
and the notation suppresses not only the size condition that 2−M <
−m
`(Q) 6 2 but also an implicit restriction to a fixed finite family of cubes of
each size, depending on the supports of f and g.
Inserting 1 = 2d · E(1{Q+̇ω is k-good} ), it hence follows, using in particular
the independence property established in Lemma 12.3.33(1), that
∗
X ∗
X
E Φ(Q+̇ω) = 2d · E(1{Q+̇ω is k-good} )EΦ(Q+̇ω)
Q∈D Q∈D
∗
X
= 2d
E 1{Q+̇ω is k-good} Φ(Q+̇ω)
Q∈D
∗
X
= 2d · E Φ(Q+̇ω).
Q∈D:
Q+̇ω is k-good
Thus, at the cost of the factor 2d , we can reduce the summation to k-good
cubes only.
Taking the expectation of (12.53) and applying the above observation to
the terms on the second line, with k = k(n) as in Definition 12.3.34, we obtain
t(f, g) = E hHtω um ω
, gi + ltω (f, g) + Emω
(f, g)+
n o
(12.54)
X good ω 1,good ω 2,good ω
+ 2d hTn,t ω um , gi + hf, Un,tω vm i + hUn,t ω um , gi ,
n∈Zd
n6=0
and hence R+̇n ⊆ R(k,ω) . Thus the operators on the right of (12.54) are in the
scope of the sharper special cases of Figiel’s estimates, Corollary 12.1.27(2)
and Theorem 12.1.28(2).
An application of these estimates to (12.54), in the case of Unω,1 on the dual
side and otherwise directly as in Corollary 12.1.27(2) and Theorem 12.1.28(2),
gives
1,good ω
X ω;1,α
|hf, Un,t ω vm i| 6 6βp,X βp,Y c1 (1 + k(n))σ1 ℘1 (tn,good ω
)kf kp kvm kp 0 ,
α
2,good ω
X
|hUn,t ω um , gi| 6 6βp,X βp,Y c2 (1 + k(n))σ2 ℘2 (tω;2,α ω
n,good )kum kp kgkp ,
0
α
174 12 Dyadic operators and the T (1) theorem
good ω
X ω;α,γ
|hTn,t ω um , gi| 6 3βp,X βp,Y min ci (1 + k(n))σi ℘i (tn,good ω
)kum kp kgkp0 .
i=1,2
α,γ
n∈Zd α
n6=0
ω;(0)
X X
(1 + k(n))σi ℘i (tω;α,γ
n,good ) 6 2ktgood kFig i (℘i ) .
σ
n∈Zd α,γ
n6=0
We have thus estimated all terms on the right of (12.54). Let us further
recall that kuω ω p d
m kp 6 2kf kp and um → f in L (R ; X) as m → −∞, with
ω 0 ω
similar results for vm , g and p in place of um , f and p. We can thus pass to
the limit m → −∞ in (12.54) and apply dominated convergence to deduce the
claimed representation formula (12.52). Applying the same estimates above
to (12.52) in place of (12.54), we deduce the claimed norm estimate (a). This
completes the proof of Theorem 12.3.35.
where we used the geometric observation that, for s ∈ Q ⊆ {R, at least half
of any ball of centre s lies in {R. Hence
`(Q)
|Q| vd rd −d−1
ZZ Z
1
ds dt 6 d r · ·r dr
Q×R |s − t|d 0 `(Q) 2
Z ∞
dvd
+d |Q| · |R|r−d−1 dr = |Q| + |R|,
`(Q) 2
Proof. The “only if” part is obvious. For “if”, it suffices to estimate t(1Q , 1R )
for R = Q+̇n and n ∈ {−1, 0, 1}d \ {0}. Then Q ∩ R = ∅, so that we have
access to the kernel representation (12.55), and Lemma 12.4.2 provides us
with the bound
ZZ
cK
kt(1Q , 1R )k 6 d
ds dt 6 18 · |Q| · cK .
Q×R |s − t|
This is one half of the decay condition (12.35). The estimate for t(1Q+̇m , 1Q )
is entirely similar.
Despite the simple observations above, in order to make serious conclusions
about weakly defined singular integrals, we will need the following elaboration
of the earlier Definition 11.3.1:
i 1
ωK (℘; ) 6 (1 + 2d )cK (℘).
2
(3) If K(s, t) = K(s − t) for some K : Rd \ {0} → Z, then
Our goal in this section will be to use these assumptions to control the Haar
γ
coefficients t(hα
Q , hR ), where R = Q + `(Q)n, in the way that was assumed
in the Theorems 12.3.26 and 12.3.35 on bilinear forms. Using the defining
condition (12.55) and bilinearity (noting that hα Q is a linear combination of
1Q0 for Q0 ∈ ch(Q), and likewise hγR ), we have in particular that
ZZ
γ
t(hαQ , h R ) = K(s, t) ds dt, Q ∩ R = ∅.
Q×R
3
√
and, for |n| > 2 d,
178 12 Dyadic operators and the T (1) theorem
3
√
3 d 4 d
n o
γ
℘ t(hα : Q ∈ D 6 ( ) · |n| · ωK (℘;
Q+̇n , hQ )
−d 1
) if γ 6= 0. (12.59)
2 |n|
3
√
3 d 4 d
n o
γ
℘ t(hQ , hQ+̇n ) : Q ∈ D 6 ( ) · |n| · ωK (℘;
α −d 2
) if α 6= 0, (12.60)
2 |n|
Proof. Including momentarily also n = 0 for later use, we have the expansion
γ
X γ
t(hα
Q , hQ+̇n ) = t(1R , 1S )hhα
Q iR hhQ+̇n iS
R∈ch(Q)
S∈ch(Q+̇n)
X γ
= δn,0 t(1R , 1R )hhα
Q iR hhQ+̇n iS
R∈ch(Q)
(12.61)
X γ
+ t(1R , 1S )hhα
Q iR hhQ+̇n iS =: IQ + IIQ .
R∈ch(Q)
S∈ch(Q+̇n)
R6=S
we see that
n t(1 , 1 ) o
U V
IIQ ∈ 2d abs conv : U, V ∈ D, U ∩ V = ∅, `(U ) = `(V ) ,
|U |
where
ZZ ZZ
ds dt
t(1U , 1V ) = K(s, t) ds dt = |s − t|d K(s, t)
U ×V U ×V |s − t|d
n o
∈ 18 · |U | · abs conv |u − v|d K(u, v) : (u, v) ∈ Ṙ2d ,
by Proposition 1.2.12 and Lemma 12.4.2 in the last step. Combining the above
inclusions with the defining properties of good set-bounds (Definition 12.3.20),
we obtain
℘({IIQ : Q ∈ D}) 6 18 · 2d · cK (℘), (12.62)
which coincides with (12.58) when n 6= 0.
For large values of n, we want to obtain a decay, which is not present
in the uniform estimate just established. In this case we apply the kernel
representation combined with the vanishing mean of hαQ (when α 6= 0), to the
result that
12.4 The T (1) theorem for singular integrals 179
ZZ
α γ α γ
t(hQ , hQ +̇n
)= K(s, t)hQ (t)hQ +̇n
(s) ds dt
ZZ
γ
= [K(s, t) − K(s, zQ )]hα
Q (t)hQ+̇n (s) ds dt,
where
√ zQ is the centre of Q. For t ∈ Q and s ∈ Q+̇n, we have |t − zQ | 6
1
2 d`(Q), whereas
1√
|s − zQ | > |zQ+̇n − zQ | − |s − zQ+̇n | > (|n| − d)`(Q),
2
and hence √
|t − zQ | 1
2 d 1 3√
6 √ 6 if |n| > d.
|s − zQ | 1
|n| − 2 d 2 2
In this case we have
ZZ
γ 1
α
t(hQ , hQ+̇n ) ∈ |hα (t)hγQ+̇n (s)| ds dt
|s − zQ |d Q
1
√
n d o
× abs conv |u − v|d [K(u, v) − K(u, v 0 )] : |v − v 0 | 6 2
√ |u − v| ,
|n| − 21 d
1
ZZ
1 1 √d
2 2
6 d
ds dt × ωK √
|Q| Q×(Q+̇n) |s − zQ | |n| − 21 d
1 1 √d 3 d 3 √d
2 2 −d 2
6 √ ωK √ 6 ( ) · |n| ωK 4
(|n| − 21 d)d |n| − 21 d 2 |n|
√
when |n| > 32 d.
γ
The estimate of t(hαQ , hQ+̇n ) with γ 6= 0 is entirely analogous to this, using
regularity in the other variable instead.
Concerning the diagonal n = 0, which was excluded in Lemma 12.4.8, we have
the following estimate:
Lemma 12.4.9. Let X and Y be Banach spaces and p ∈ (1, ∞). Let t : Ṙ2d →
L (X, Y ) be a weakly defined singular integral with the weak DR p -boundedness
property. Then
γ
DR p ({t(hα d
Q , hQ )}Q∈D ) 6 ktkwbp(DRp ) + 18 · 2 · cK (℘), ℘ ∈ {Rp , Rp∗0 }.
180 12 Dyadic operators and the T (1) theorem
where we now need to consider also the term IQ . We estimate the expression
in the definition of DR p ({IQ }Q∈D ):
X X X γ
∗ ∗
|Q||hIQ xQ , yQ i| 6 |Q| |ht(1R , 1R )xQ , yQ i||hhα
Q iR hhQ iR |
Q∈D Q∈D R∈ch(Q)
X X
∗
= |ht(1R , 1R )xQ , yQ i|
Q∈D R∈ch(Q)
X
∗
= |ht(1R , 1R )xR(1) , yR (1) i|
R∈D
X
6 ktkwbp(DRp ) εR xR(1) 1R ×
Lp (Ω×Rd ;X)
R∈D
X
∗
× εR yR (1) 1R .
Lp0 (Ω×Rd ;Y ∗ )
R∈D
Using the usual observation that, by Fubini’s theorem and the fact that only
one R ∈ D of each generation is “seen” at each fixed s ∈ Rd , we can replace
the random εR by εn(R) depending on the generation of R only, or further by
the equidistributed sequence of εn(R(1) ) , we have
X X X
εR zR(1) 1R = εn(Q) zQ 1R
Lp (Ω×Rd ;Z) Lp (Ω×Rd ;Z)
R∈D Q∈D R∈ch(Q)
X X
= εn(Q) zQ 1Q = εQ zQ 1Q
Lp (Ω×Rd ;Z) Lp (Ω×Rd ;Z)
Q∈D Q∈D
∗
for both choices of zQ ∈ {xQ , yQ } and Z ∈ {X, Y }. Hence
and hence, by the obvious triangle inequality for DR p , and its domination by
either ℘ ∈ {Rp , Rp∗0 } according to Lemma 12.1.8, we have
DR p ({t(hQ
α
, hγQ )}Q∈D ) 6 DR p ({IQ }Q∈D ) + DR p ({IIQ }Q∈D )
6 ktkwbp(DRp ) + ℘({IIQ }Q∈D )
6 ktkwbp(DRp ) + 18 · cK (℘)
Proof of Lemma 12.4.10. From Definition 12.3.22 and Lemma 12.4.8, it fol-
lows that
X
kt(0) kFigs (℘) = ktαγ kFigs (℘) ,
α,γ∈{0,1}d \{0}
X X γ
= Q , hQ+̇n ) : Q ∈ D})
(2 + log2 |n|)s ℘({t(hα
α,γ∈{0,1}d \{0} n∈Zd
n6=0
n X √
6 (2d − 1)2 (2 + log2 (3 d)) · 18 · 2d · cK (℘)+ (12.64)
√
|n|<3 d
√
X 3 d −d i 34 d o
s
+ (2 + log2 |n|) ( ) |n| ωK ℘;
√ 2 |n|
|n|>3 d
Since both α 6= 0 6= γ, one can apply either of the estimates (12.59) or (12.60)
of Lemma 12.4.8, and thus take either i ∈ {1, 2} above. Similarly,
X
kt(2) kFigs (℘) = kt2,α kFigs (℘) ,
α∈{0,1}d \{0}
X X
Q , hQ+̇n ) : Q ∈ D})
(2 + log2 |n|)s ℘({t(hα 0
= (12.65)
α∈{0,1}d \{0} n∈Zd
n6=0
where we only have access to estimate (12.60), but not (12.59), of Lemma
12.4.8, now that the second Haar function h0Q+̇n is non-cancellative. The very
last step in (12.65) is of course wasteful, but we make it in order to treat the
right-hand sides of both (12.64) and (12.65) at the same time.
Finally, in complete analogy with (12.65), we also have
182 12 Dyadic operators and the T (1) theorem
X
kt(1) kFigs (℘) = kt1,α kFigs (℘) 6 4d (I + II1 ), (12.66)
α∈{0,1}d \{0}
as we now have access to estimate (12.59), but not (12.60), of Lemma 12.4.8.
It is immediate that
X √
4d I = ad · cK (℘), ad := 4d (2 + log2 (3 d)) · 18 · 2d . (12.67)
√
|n|<3 d
For the other term, we partition the summation over dyadic annuli, in which
the summand is roughly a constant:
∞
X X √ √
4d IIi 6 6d (2 + log2 (3 d) + k)s (3 · 2k d)−d ωK
i
(℘; 2−k−2 ).
√
k=0 3·2k d6|n|
√
<3·2k+1 d
√
The unit-cubes Qn with centres n ∈ Zd are √ disjoint, and for |n| < 3 · 2k+1 d,
they are contained in B(0, (3 · 2k+1 + 21 ) d). Thus
X 1 √ d √ d
1 6 vd (3 · 2k+1 + ) d 6 vd (6.5)d 2kd d , (12.68)
√ 2
|n|<3·2k+1 d
i
Since ωK (℘; u) is non-decreasing, we can finally estimate
Z 2−k−1 i
1 1 s ωK (℘; u) du
(1 + k) s i
ωK (℘; 2−k−2 ) 6 (log2 ) , k = 0, 1, . . . ,
log 2 2−k−2 u log 2 u
and hence
√
d i (13)d vd (2 + log2 (3 d))
4 IIi 6 bd kωK (℘)kDinis , bd := .
log 2
With (12.64), (12.65), (12.66), and (12.67), this concludes the proof. (An
estimate similar to (12.68) could also be used to give a more explicit bound
for the constant ad in (12.67), if desired.)
We have now everything prepared for proving the following:
Theorem 12.4.12 (T (1) theorem for operator-valued kernels). Let
p ∈ (1, ∞) and 1 6 t 6 p 6 q 6 ∞, and suppose that:
12.4 The T (1) theorem for singular integrals 183
kT kL (Lp (Rd ;X),Lp (Rd ;Y )) − sup kΛtω kL (Lp (Rd ;X),Lp (Rd ;Y ))
ω
n
6 βp,X βp,Y 4 sup kt kwbp(DRp ) + c0d c1 cK (Rp∗0) + c2 cK (Rp ) +
d ω
ω
o
+ cd c1 kωK (Rp∗0)kDini1/t + c2 kωK
1 1 2
(Rp )kDini1/q0 ,
where the suprema are over ω ∈ ({0, 1}d )Z0 , the constants cd , c0d depend
only on d, and
Remark 12.3.27, under the (co)type assumption (ii) of Theorem 12.4.12, the
assumption (ii) of Theorem 12.3.26 are satisfied with
where both right-hand sides of are finite by (12.71). With either choice of
(t0 , q0 ) ∈ {(ti , qi )}2i=1 , the resulting finiteness of the left-hand sides coincides
with the assumption on these quantities in (iii) of Theorem 12.3.26.
Summarising, assumptions (i) through (iii) of Theorem 12.4.12, together
with the weak DR p -boundedness property of t, which is either assumed or
implied by the assumptions of each case of Theorem 12.4.12, imply the corre-
sponding assumptions (i) through (iii) of Theorem 12.3.26. Moreover, the con-
dition of adjacent weak boundedness property appearing in Theorem 12.3.26
also follows from these assumptions by Lemma 12.4.3 and the domination of
uniform bounds by either DR p -bounds or ℘i -bounds:
Hence all assumptions, and thus all conclusions of Theorem 12.3.26 are valid
under the assumptions of Theorem 12.4.12. This proves in particular the qual-
itative equivalence (1)⇔(2).
12.4 The T (1) theorem for singular integrals 185
where cd := (Ad + Bd )ad and c0d := (Ad + Bd )bd . This is readily recognised to
coincide with the bound asserted in (a) of the theorem.
(1)⇔(3): This will be an application of Theorem 12.3.35. Assumptions (i)
and (ii) are identical in both theorems.
Concerning assumption (iii), we need to check that the kernel assumptions
(12.69) of the present theorem imply the estimates on Figiel norms of each
bilinear form tω , uniformly in ω ∈ ({0, 1}d )0Z . We already did this for t = t0
above. However, all the lemmas of this section are stated for an arbitrary
dyadic system D, so we may in particular use them with any D ω in place of
D. Moreover, the constants in these estimates are explicit, and clearly inde-
pendent of the particular ω. This proves the qualitative equivalence (1)⇔(3).
(b): For this quantitative estimate, we apply Theorem 12.3.35(a), followed
by (12.72) and (12.73) with tω and D ω in place of t and D, to get
1
kT kL (Lp (Rd ;X),Lp (Rd ;Y )) − sup kΛtω kL (Lp (Rd ;X),Lp (Rd ;Y ))
ω βp,X βp,Y
2
ω;(j)
X X X
6 sup DR p (tω;α,γ
0 ) + 12 · 2d sup ci ktgood kFigσi (℘i )
ω ω
α,γ i=1 j∈{0,i}
2
X
6 4d sup ktω kwbp(DRp ) + sup ci c0d cK (℘i ) + c1d kωK
i
(℘i )kDiniσi ,
ω ω
i=1
where the supremum is over ω ∈ ({0, 1}d )Z0 , the constants cd , c0d depend
only on d, and c1 , c2 are as in (12.70);
(b) the representation formulas (12.45) and (12.52) with Λt = Λtω = 0.
Lemma 12.4.14. Let Z = L (X, Y ) and Φ ∈ Cb ([0, ∞); Z) ∩ C 1 ((0, ∞); Z),
and suppose that
(i) K(u) := 1(0,∞) (u)Φ0 (u) satisfies the Calderón–Zygmund estimate (12.74);
(ii) the range of Φ is R-bounded, Rp (Φ) := Rp ({Φ(u) : u ∈ [0, ∞)}) < ∞;
(iii) a bilinear form t : S(D)2 → Z is defined, for all f, g ∈ S(D), by
ZZ
t(f, g) := lim K(u − v)f (v)g(u) dv du.
ε→0 |u−v|>ε
Then
(1) t is well-defined as a weakly defined singular integral with convolution ker-
nel K(u, v) = K(u − v);
(2) tω satisfies the weak DR p -boundedness property
Proof. (1): Clearly the integral inside the limit is well-defined, since we are
cutting away the singularity. To show the existence of the limit, let first f = 1I
and g = 1J for some intervals I = [aI , bI ) and J = [aJ , bJ ). Then
188 12 Dyadic operators and the T (1) theorem
Z Z bI
K(u − v)f (v) dv = 1(ε,∞) (u − v)Φ0 (u − v) dv
|u−v|>ε aI
Z bI ∧(u−ε)
= 1(aI +ε,∞) (u) Φ0 (u − v) dv
aI
= 1(aI +ε,∞) (u)[Φ((u − bI ) ∨ ε) − Φ(u − aI )].
= Φ((u − bI )+ ) − Φ((u − aI )+ ).
In particular, the limit defining t(f, g) exists for all f, g of the form f = 1I
and g = 1J . By (bi)linearity, it exists for all f, g ∈ S(D).
If f, g ∈ S(D) are disjointly supported, then K(u−v)f (v)g(u) is integrable.
Hence ZZ
t(f, g) = K(u − v)f (v)g(u) dv du
(3): From (12.76) it is evident that t(1I , 1I ) depends only on `(I); since
`(I) = `(I +̇m), it follows that t(1I , 1I ) = t(1I +̇m , 1I +̇m ), as claimed.
12.4 The T (1) theorem for singular integrals 189
tj |dz|
I
j (j) j! j!℘(Φ)
℘(t Φ (t) : t > 0) 6 ℘(Φ) sup j+1
= .
2π t>0 |z−t|=t sin ω (t sin ω) (sin ω)j
℘(Φ)
c̃K (℘) = ℘(|u|K(u) : u 6= 0) = ℘(uΦ0 (u) : u > 0) 6 .
sin ω
Moreover,
where
u
|u − u0 |
Z
u 1 1 su s
dv = u − = 6 = 6 2s
u0 v2 u u0 u0 (1 − w)u 1−s
3℘(Φ) 4
kω̃K kDiniσ 6 1 + log1+σ .
sin ω sin ω
This completes the proof.
12.4 The T (1) theorem for singular integrals 191
observing the cancellation of the two equal terms 1Q (s)1Q (t). We can divide
R \ Q into finitely many cubes P ∈ D of the same size as Q, and then
the integrability of each of the terms on the left against K(s, t) follows from
192 12 Dyadic operators and the T (1) theorem
Hence
I − II
ZZ
t(f, g) = =I= K(s, t)f (t)g(s) dt ds,
2
as required for t to be a weakly defined singular integral with kernel K.
From the defining formula (12.78) it is immediate that t(1Q , 1Q ) = 0, and
hence the quantities featuring in the weak boundedness property of t vanish.
With Q ∈ D ω (which still satisfies 1Q ∈ S(D) for ω ∈ ({0, 1}d )Z0 , by Lemma
12.3.30, the same conclusion extends to tω for all ω ∈ ({0, 1}d )Z0 . The rest of
the corollary is then a direct consequence of Theorem 12.4.12, simply setting
ktkwbp(DRp ) = ktω kwbp(DRp ) = 0. We only need to note that ωK 1
(℘) = ωK 2
(℘)
when K is antisymmetric, which is why a seemingly weaker assumption suffices
in (12.77).
where the supremum is over ω ∈ ({0, 1}d )0Z , the constants cd , cd0 depend
only on d, and c1 , c2 are as in (12.70).
(b) the representation formulas (12.45) and (12.52) with Λt = Λtω = 0.
Proof. This is straightforward by combining (the proofs of) Corollaries 12.4.13
and 12.4.18. In particular, in the proof of Corollary 12.4.18 we observed
that any bilinear form defined as in (iv) of the present corollary will satisfy
t(1Q , 1Q ) = 0 for all Q ∈ D, and hence also t(1Q+̇m , 1Q+̇m ) = 0 = t(1Q , 1Q )
for all m ∈ Zd . This is condition (iv) of Corollary 12.4.13 that was not explic-
itly assumed in the corollary that we are proving.
Remark 12.4.20. As an immediate consequence of Corollary 12.4.13, we obtain
another proof of the essence of Theorem 5.1.13 on the boundedness of the
Hilbert transform H on Lp (R; X) whenever p ∈ (1, ∞) and X is a UMD
space. Indeed, take X = Y , t = 1, and q = ∞, so that the constants in (12.70)
are simply c1 = c2 = 1. Clearly the kernel K(u, v) = π −1 (u − v)−1 of the
Hilbert transform is an antisymmetric convolution kernel, and it is easy to
check the Calderón–Zygmund estimates (12.74) with Dini1 norms. Thus we
obtain the estimate
2
kHkL (Lp (Rd ;X)) 6 c · βp,X ,
with the same quantitative form as (5.24), aside from the unspecified numer-
ical factor above, in contrast to the explicit constant 2 in (5.24). This is quite
natural, considering that (5.24) was obtained by an argument tailored for the
very Hilbert transform, whereas the argument that we just sketched was a
specialisation of a much more general argument to the particular case of H.
The following corollary provides a solution to the Lp extension problem from
Section 2.1 for the important class of Calderón–Zygmund operators:
Theorem 12.4.21 (T (1) theorem for scalar-valued kernels). Let p, s ∈
(1, ∞) and 1 6 t 6 p 6 q 6 ∞, and suppose that:
(i) X is a UMD space with cotype q and type t,
(ii) t : S(D)2 → K is a weakly defined singular integral, whose kernel K :
Ṙ2d → K satisfies the Calderón–Zygmund estimates
2
X
i
cK + kωK kDiniσi < ∞, (12.79)
i=1
Proof. (1) ⇒ (2): For s = p, this is evident by restricting the action of the
operator to a one-dimensional subspace of X. The case of general s ∈ (1, ∞)
follows from the Calderón–Zygmund Theorem 11.2.5 (or even just its classical
scalar-valued version).
(2) ⇒ (3): The weak boundedness property follows from Example 12.1.10:
2
6 (cd kT kL (Ls (Rd )) + σd−1 kωK kDini )
Our assumption (4) also involves ktkwbp < ∞, and Corollary 12.1.9 guarantees
that this coincides with the finiteness of ktkwbp(DRp ) = ktkwbp , when t is scalar-
valued. Thus both assumptions ktkwbp(DRp ) < ∞ and kΛt kL (Lp (Rd ;X)) < ∞
of Theorem 12.4.12(2) are satisfied, hence also the equivalent condition of
Theorem 12.4.12(1), and this coincides with condition (1) of the corollary
that we are proving.
The quantitative estimates: While we have already closed the chain of impli-
cations (1) ⇒ (2) ⇒ (3) ⇒ (4) ⇒ (1), the claimed quantitative bounds require
a direct analysis of the implication (3) ⇒ (1), which relates to the implication
(3) ⇒ (1) of Theorem 12.4.12.
As in the proof of “(4) ⇒ (1)”, under assumption (3), we see that
the paraproducts related to tω are in fact Πtiω = Πbωi ; while the function
bi ∈ BMO(Rd ) ⊆ BMOD ω (Rd ) is independent of ω, the superscript of the
paraproduct signifies the fact that the defining series involves Haar functions
and averages related to Q ∈ D ω . Thus, imitating (12.85) and substituting the
bounds (12.83) and (12.84), we obtain, with s1 := s and s2 := s0 ,
196 12 Dyadic operators and the T (1) theorem
kΛtω kL (Lp (Rd ;X)) = kΠbω1 + (Πbω2 )∗ kL (Lp (Rd ;X))
2
X
6 64 · 8d · pp0 βs,X
2
βs,K kbi kBMOsi (Rd )
i=1 (12.86)
2
X
6 64 · 8d · pp0 βs,X
2
βs,K i
cd kT kL (Ls (Rd )) + σd−1 kωK kDini .
i=1
where we implicitly dominated kbi kBMOsiω (Rd ) 6 kbi kBMOsi (Rd ) in the first es-
D
timate. We now substitute (12.86) and (12.82) into the second norm estimate
in Theorem 12.4.12(b), noting that all R-bounds and DR p -bounds may be
omitted, since they simply reduce to uniform bounds for scalar-valued func-
tions:
Proof. By inspection of the proof of Theorem 12.4.21, the said term only arises
in the estimate of Λtω in (12.86). Under the assumption that b1 = b2 , we have
Λtω = Λbω1 , and we may replace (12.86) by an application of Theorem 12.2.25:
where
1
kb1 kBMO(Rd ) 6 kb1 kBMOs (Rd ) 6 cd kT kL (Ls (Rd )) + σd−1 kωK kDini .
where
Z
(i,k) (i,k) (i,k)
X
(i,k)
S f= AP f, AP f (s) = − aP (s, t)f (t) dt,
P ∈D P
where
Z
(i,k) (i,k)
AP ;R f (s) = − aP ;R (s, u)f (u) du, R ∈ {P } ∪ ch(k) (P ),
R
where the k-goodness of Q guarantees that R := Q+̇n, for |n| 6 2k−2 , shares
with Q the same kth dyadic ancestor R(k) = Q(k) =: P ∈ D. Thus we can
regroup this series under the ancestors P to get
X X D E
γ γ
Tk = tgood (hα
Q , h R )hf, h α
Q i, hg, h R i ,
P ∈D (Q,R)∈Ck (P )
α,γ∈{0,1}d \{0}
where
n 1 1 o
Ck (P ) := (Q, R) : Q, R ∈ ch(k) (P ), `(P ) < |zQ − zR | 6 `(P ) .
8 4
(k)
The subseries under each P ∈ D takes the asserted form hAP f, gi if we define
(0,k)
X γ γ
aP (s, u) := |P | tgood (hα α
Q , hR )hQ (u)hR (s).
(Q,R)∈Ck (P )
α,γ∈{0,1}d \{0}
The cases of Uki are analogous, and lead to representations of the same
form with
(1,k)
X
aP (s, u) := |P | tgood (h0Q , hγR )[h0Q (u) − h0R (u)]hγR (s),
(Q,R)∈Ck (P )
γ∈{0,1}d \{0}
and
(2,k)
X
aP (s, u) := |P | tgood (hα 0 α 0 0
Q , hR )hQ (u)[hR (s) − hQ (s)],
(Q,R)∈Ck (P )
α∈{0,1}d \{0}
12.4 The T (1) theorem for singular integrals 199
where the last summation runs over all relevant Q ∈ chk (P ), for fixed R.
(1,k)
Observe that aP ;R has the factor |R| in front, instead of |P |, due to our
(1,k) R (1,k)
definition of AP ;R f (s) as the average integral −R aP ;R (s, u)f (u) du.
(2,k)
The splitting of aP is entirely analogous; in particular,
(2,k)
X
aP ;Q (s, u) := |Q| α
tgood (hQ α
, h0R )hQ 0
(u)hQ (s), Q ∈ chk (P ).
R:(Q,R)∈Ck (P )
α∈{0,1}d \{0}
It remains to verify that these operators and their kernels satisfy the as-
(0,k) (k) (0,k) (k)
serted properties. The identity AP = DP AP DP is immediate from the
(i,k)
orthogonality of the Haar functions, and the invariance of AP under com-
(k)
position by DP on the side, where the cancellative Haar function appear
(i,k) [0,k)
in aP is justified similarly. Concerning the factors DP , we note that the
are orthogonal projections onto functions supported on P , constant on each
Q ∈ ch(k) (Q), and integrating to zero. Noting the functions hQ 0
− h0R belong
to this class then justifies the remaining parts of the claimed identities.
Concerning the claimed bounds, we note that any given (s, u) ∈ P × P is
contained in exactly one Q × R with Q, R ∈ ch(k) (P ), and moreover,
γ 1Q×R 2kd
|hα
Q ⊗ hR | = = 1Q×R .
|Q|1/2 |R|1/2 |P |
(0,k) (i,k)
The claimed ℘-bounds for aP (s, u), as well as for aP ;P (s, u), then follow
from Lemma 12.4.8, noting that the factor |P | in the definition of these kernels
cancels with the |P1 | above.
(1,k)
For aP ;R with R ∈ ch(k) (P ), all terms in the defining sum are supported
on the same 1R×R , and each individual summand can be estimates by Lemma
12.4.8. We now have the smaller factor |R| in front, but at the same time there
are up to 2kd terms in the sum, all of which accumulate on the same support
now. Since 2kd |R| = |P |, we get the same final bound as before. The case of
(1,k)
aP ;Q with Q ∈ ch(k) (P ) in entirely analogous, and completes the proof.
where
Z
AP ;R f (s) = − aP ;R (s, u)f (u) du, R ∈ {P } ∪ ch(k) (P ),
R
supp aP ;R ⊆ R × R,
n o
kSkShift(℘) := ℘ aP ;R (s, u) : s, u ∈ R ∈ {P } ∪ ch(k) (P ), P ∈ D <∞
The key boundedness properties of these dyadic shifts are contained in the
following:
Theorem 12.4.26. Let X and Y be UMD spaces, and p ∈ (1, ∞). Suppose
that X has cotype q and Y has type t for some 1 6 t 6 p 6 q 6 ∞.
Then for all i ∈ {0, 1, 2} and k ∈ {2, 3, . . .}, all dyadic shifts S of type
(i, k) extends to a bounded operator from Lp (Rd ; X) to Lp (Rd ; Y ). Moreover,
they satisfy the norm estimates
(
kSkShift(Rp ) ct0 ,Y ∗ ;p0 · k 1/t , i = 1,
kSkL (Lp (Rd ;X),Lp (Rd ;Y )) 6 4 · βp,X βp,Y × 0
kSkShift(R∗0 ) cq,X;p · k 1/q , i = 2;
p
and the norm of a shift of type (0, k) is bounded by the minimum of these two
bounds, but with 6 in place of 4.
Proof. We divide the proof into case according to the type of the shift under
consideration.
12.4 The T (1) theorem for singular integrals 201
Shifts of type 1
Let us start with the case i = 1. For f ∈ S00 (Rd ; X) ⊆ Lp (Rd ; X) and
0
g ∈ S00 (Rd ; Y ) ⊆ Lp (Rd ; Y ∗ ), we expand the pairing hSf, gi by separating
the scales according to log2 `(P ) mod k:
k−1 D E
(k) (k)
X X
|hSf, gi| = DP AP f, DP g
j=0 P ∈D
log2 `(P )≡j
mod k
k−1
(k)
X X
6 ε P D P AP f
Lp (Ω×Rd ;Y )
j=0 P ∈D
log2 `(P )≡j
mod k
k−1
(k)
X X
× ε P DP g =: Ij × IIj .
Lp0 (Ω×Rd ;Y ∗ )
P ∈D j=0
log2 `(P )≡j
mod k
(k) P
In Ij , we write out DP = Q∈ch(k) (P ) DQ and note that, in a randomised
sum like here, we are free to replace εQ by εP , since the difference is invisible
to the Lp (Ω; Y ) at a fixed s ∈ Rd . This gives
X X
Ij = ε Q D Q AP f .
Lp (R; Y )
P ∈D Q∈ch(k) (P )
log2 `(P )≡j
mod k
where Z
αP ;Q (s) := − haP ;Q (·, u)iQs − haP ;Q (·, u)iQ du
Q
belongs to the two-fold multiple of the absolute convex hull of the set appear-
ing in the definition of kSkShift(℘) . Thus
[0,k)
X X
IVj = εQ αP ;Q 1Q DP f
Lp (Ω×R; Y )
P ∈D Q∈ch(k) (P )
log2 `(P )≡j
mod k
[0,k)
X X
6 2kSkShift(Rp ) ε Q 1Q D P f
Lp (Ω×R; X)
P ∈D Q∈ch(k) (P )
log2 `(P )≡j
mod k
[0,k)
X
6 2kSkShift(Rp ) ε P DP f ,
Lp (Ω×R; X)
P ∈D
log2 `(P )≡j
mod k
P
using the identity Q∈ch(k) (P ) 1Q = 1P and the interchangeability of εP and
εQ in the random sum in the last step.
[0,k)
Observing that (DP f )log2 `(P )≡j mod k is a martingale difference decom-
position of f for each j ∈ {0, . . . , k − 1} to deduce directly from the definition
of the UMD constants that
+
IVj 6 2kSkShift(Rp ) βp,X kf kLp (Rd ;X) .
where
Z
(k) (k) [0,k)
DP AP ;P f (s) = − DP aP ;P (s, u)DP f (u) du,
P
(k)
and it is understood that DP acts with respect to the s variable.
We will now make use of the tangent martingale construction as in Corol-
lary 4.4.15 and explained just before the statement of the said result: For
every P ∈ D, let TP be a copy of P equipped with the normalised measure
νP := |P |−1 m|PQ, where m is the Lebesgue measure, and consider the prod-
uct space T := P ∈D TP with probability measure ν := ⊗P ∈D νP . Writing a
typical element of T as t = (tP )P ∈D , we then have
12.4 The T (1) theorem for singular integrals 203
Z
(k) (k) [0,k)
DP AP ;P f (s) = DP aP ;P (s, tP )DP f (tP ) dν(t).
T
(k)
Here, DP aP ;P (s, tP ) is the difference of two averages haP ;P (·, tP )iQ , and
hence in twice the absolute convex hull of the set in the definition of kSkShift(℘) .
Thus, the definition of R-boundedness implies that
[0,k)
X
IIIj 6 2kSkShift(Rp ) (s, t) 7→ εP 1P (s)DP f (tP ) p d
.
L (Ω×R ×T ;X)
P ∈D
log2 `(P )≡j
mod k
shows that
[0,k)
X
(s, t) 7→ εP 1P (s)DP f (tP ) 6 βp,X kf kLp (Rd ;X) ,
Lp (Ω×Rd ×T ;X)
P ∈D
log2 `(P )≡j
mod k
and hence
IIIj 6 2kSkShift(Rp ) βp,X kf kLp (Rd ;X) .
Combining this with the estimate for term IVj (and estimating the one-sided
UMD constant by the basic UMD constant), we deduce that
Hence
204 12 Dyadic operators and the T (1) theorem
k−1
X k−1
X
|hSf, gi| 6 Ij × IIj 6 4kSkShift(Rp ) βp,X kf kLp (Rd ;X) IIj ,
j=0 j=0
where
k−1 k−1 10
X X 0 t
IIj 6 k 1/t IIjt
j=0 j=0
k−1 t0 10
(k)
X X t
= k 1/t ε P DP g
Lp0 (Ω×Rd ;Y ∗ )
j=0 P ∈D
log2 `(P )≡j
mod k
(k)
X
6 k 1/t · ct0 ,Y ∗ ;p0 εP DP g
Lp0 (Ω×Rd ;Y ∗ )
P ∈D
Here βp+0 ,Y ∗ 6 βp0 ,Y ∗ = βp,Y by Proposition 4.2.17(2), and ct0 ,Y ∗ ;p0 6 τt,Y ;p by
Proposition 7.1.13 (or its easy extension to deal with the third index in these
constants). This completes the proof for shift of type (1, k).
Shifts of type 2
For a shift of type (2, k), we note that its adjoint S ∗ is a shift of type (1, k),
and hence
Shifts of type 0
Let finally S be a shift of type (0, k). We can then proceed as in the case of
type (1, k) with slight modifications: In view of the eventual application of
the tangent martingale estimate of Corollary 4.4.15, we now separate scales
(k)
by k + 1 levels instead of k, since DP f is only guaranteed to be constant
on Q ∈ ch(k+1) (P ). On the other hand, we now have IVj = 0, and hence
Ij = IIIj .
Following the argument in the case of type (1, k) leads to
(k)
X
IIIj 6 2kSkShift(Rp ) (s, t) 7→ εP 1P (s)DP f (tP ) .
Lp (Ω×Rd ×T ;X)
P ∈D
log2 `(P )≡j
mod k+1
12.4 The T (1) theorem for singular integrals 205
by the contractivity of the conditional expectation in the last step. This last
expression has the same form as what we encountered with shifts of type
(1, k), only with k + 1 in place of k. Thus, by an application of the tangent
martingale inequality of Corollary 4.4.15, we have
[0,k+1)
X
εP 1P (s)DP f (tP ) 6 βp,X kf kLp (Rd ;X) .
Lp (Ω×Rd ×T ;X)
P ∈D
log2 `(P )≡j
mod k+1
Thus,
Ij = IIIj 6 4kSkShift(Rp ) βp,X kf kLp (Rd ;X) ,
which is the same bound as for the corresponding terms in the estimate of
shifts of type (1, k). The rest of the argument is exactly the same, only with
k + 1 in place of k, and leads to the conclusion that
206 12 Dyadic operators and the T (1) theorem
kSkL (Lp (Rd ;X),Lp (Rd ;Y )) 6 4kSkShift(Rp ) βp,X βp,Y ct0 ,Y ∗ ;p0 (k + 1)1/t .
Since the adjoint of a shift of type (0, k) is another shift of the same type, we
also obtain
0
kSkL (Lp (Rd ;X),Lp (Rd ;Y )) 6 4kSkShift(R∗0 ) βp,X βp,Y cq,X;p (k + 1)1/q ,
p
and we can take the minimum of the two bounds. Since k > 2, we can also
make the trivial estimates k + 1 6 23 k and 4 · ( 23 )1/v 6 6 for v ∈ {t, q 0 } so that
in case v > 1.
With the help of the shifts, we can represent any weakly defined singular
integral with appropriate kernel bounds as follows:
kT kL (Lp (Rd ;X),Lp (Rd ;Y )) − sup kΛtω kL (Lp (Rd ;X),Lp (Rd ;Y ))
ω
n
6 βp,X βp,Y 4 sup kt kwbp(DRp ) + c0d cK (Rp ) + cK (Rp∗0 ) +
d ω
ω
o
1
+ cd ct0 ,Y ∗ ;p0 kωK1
(Rp )kDini1/t + cq,X;p kωK 2
(Rp∗0 )kDini1/q0 ,
where the suprema are over ω ∈ ({0, 1}d )0Z , and the constants cd0 , c1d depend
only on d
Proof. We note that the present assumptions coincide with those of Theorem
12.4.12, except that (ii) of the present theorem is slightly stronger than (ii) of
Theorem 12.4.12. Thus the equivalence of (1) and (2) is just repetition from
Theorem 12.4.12.
The first new claim is the dyadic representation formula (a). To see this,
recall that Theorem 12.4.12 gave the representation formula (12.52), repeated
for convenience as
X n good
hT f, gi = E hHtω f, gi + hΛtω f, gi + 2d hTn,tω f, gi+
n∈Zd
n6=0
o
1,good 2,good
+ hf, Un,t ω gi + hUn,t ω f, gi ,
where f, g, and E have the same meaning as in the claimed formula (a). On the
other hand, Lemma 12.4.23 and Remark 12.4.25 inform us that the summation
of the three types of terms over n ∈ Zd \ {0} can be rearranged into a sum
over k > 2 and i ∈ {0, 1, 2} exactly as in the assertion.
From the representation (a), we can then estimate
kT kL (Lp (Rd ;X),Lp (Rd ;Y )) − sup kΛtω kL (Lp (Rd ;X),Lp (Rd ;Y ))
ω
6 sup kHtω kL (Lp (Rd ;X),Lp (Rd ;Y ))
ω
∞
X
+ kSω(i,k) kL (Lp (Rd ;X),Lp (Rd ;Y )) .
k=2
i∈{0,1,2}
Similarly,
∞
X
kSω(2,k) kL (Lp (Rd ;X),Lp (Rd ;Y ))
k=2
6 c0d βp,X βp,Y cK (Rp∗0 ) + cq,X;p kωK
2
(Rp∗0 )kDini1/q0 .
Finally, The sum over shifts of type (0, k) may be estimated by either of the
two bounds above (the different numerical constant in Theorem 12.4.26 is in
any case absorbed into the unspecified dimensional constant).
Remark 12.4.28. The norm estimate obtained in Theorem 12.4.27 via the rep-
resentation in terms of dyadic shifts is essentially the same as that in Theo-
rem 12.4.12 obtained via Figiel’s representation. While the proof of Theorem
12.4.27 partially relied on the proof of Theorem 12.4.12 to avoid repetition, a
larger part of the machinery behind the proof of Theorem 12.4.12, relying in
particular on Figiel’s Theorems 12.1.25 and 12.1.28 concerning his elementary
operators, was replaced in the proof of Theorem 12.4.27 by Theorem 12.4.26 on
the dyadic shifts, which in turn was based on the tangent martingale bounds
of Corollary 4.4.15.
12.5 Notes
Section 12.1
The Haar multipliers Hλ = Hααλ are special cases of martingale transforms dis-
cussed extensively in Volume I; see in particular Sections 3.5 and 4.2.e. In this
framework, the predictable sequences multiplying the martingale differences
X X
Dα
k f := hf, hα α
Q ihQ , D−α
k f := D−α
Q f
Q∈Dk Q∈Dk
12.5 Notes 209
are then
X
vkα = λk 1Qk ∈ L∞ (σ(Dk ); L (X, Y )), vk−α ≡ 0.
Q∈Dk
On the other hand, the Haar multipliers Hαγ λ with α 6= γ already take a
departure from the general theory, and this is even more so with the general
operators of Figiel.
(Note that the conventional indices of dyadic analysis and martingale the-
ory are off by one from each other. In martingale theory, it is customary to
emphasise measurability, and hence the indices of martingale differences agree
with those of the σ-algebra that makes them measurable, while predictable
multipliers are measurable with respect to the “previous” σ-algebra with in-
dex k − 1. In dyadic analysis, the emphasis is on the supporting dyadic cubes,
and hence the “kth” martingale difference Dk f is the sum of the local mar-
tingale differences DQ supported, and averaging to zero, on the dyadic cubes
Q ∈ Dk , but then they are actually measurable only with respect to the “next”
σ(Dk+1 ); at the same time, the “predictable” multipliers are then measurable
with respect to the σ-algebra indicated by their index.)
The relaxed R-boundedness notion DR p of Definition 12.1.6 seems to
be new, but the slightly stronger E R p appears implicitly in Di Plinio, Li,
Martikainen, and Vuorinen [2020b, Remark 6.29], where it is shown that
the family |Q|−1 hT 1Q , 1Q i of Example 12.1.10 has this property when T ∈
L (Lp (Rd ; X), Lp (Rd ; Y )) and X and Y are UMD spaces; this also follows by
combining our Example 12.1.10 (on the DR p property of this family) and
Corollary 12.1.17 (the equivalence of DR p and E R p for UMD spaces). An
advantage of the new DR p is that it allows Example 12.1.10 without any
assumptions on the Banach spaces.
The exact characterisation of the boundedness of the Haar multipliers
Hλ in Theorem 12.1.11 is new; by Lemma 12.1.8 and Propositions 12.1.13
and 12.1.14, the characterising condition is strictly more general than the R-
boundedness condition kx 7→ R({λQ : x ∈ Q ∈ D})k∞ < ∞. This seems
at first to contradict Girardi and Weis [2005], where the necessity of uniform
pointwise R-boundedness for operator-valued martingale transforms is estab-
lished. This apparent contradiction is resolved by observing that, in order
to obtain this necessity of R-boundedness, Girardi and Weis [2005] actually
assume that their transforming sequence (vk )k>1 is allowed to multiply any
∞
subsequence (dfnk )k>1 of the martingale difference sequence (df Pk )k=1 , i.e.,
they assume the boundedness
P of the family of operators f 7→ k>1 vk dfnk
instead of just f 7→ k>1 vk dfk . In the case of Haar multipliers, this would
mean that, for a given sequence λ = (λQ )Q∈D we would consider a family of
operators including in particular all
X
α α
f 7→ λQ(k) hf, hQ ihQ ,
Q∈D
210 12 Dyadic operators and the T (1) theorem
Section 12.2
In analogy with the quote of Stein [1982] on square functions at the beginning
of Chapter 9, also the concept of paraproduct is “not an idea in its pure form,
but rather takes various shapes depending on the uses it is put to”. A friendly
overview to this variety of “shapes and uses” of paraproducts can be found
in Bényi, Maldonado, and Naibo [2010]. Paraproducts were systematically in-
troduced by Bony [1981], but Bényi et al. [2010] convincingly argue that their
12.5 Notes 211
had been settled some years earlier: Independently, Katz [1997] and Nazarov,
Treil, and Volberg [1997b] proved that ψ(N ) . 1 + log N , and the latter
authors also obtained the preliminary lower bound ψ(N ) & (1 + log N )1/2 .
This was improved to ψ(N ) & 1+log N by Nazarov, Pisier, Treil, and Volberg
[2002a]. For a while, there were hopes in the air of obtaining a dimension-free
estimate with BMOD (R; L (`2N )) in place of BMOD so
(R; L (`N
2
)) on the right.
Some indications that made this plausible are discussed in the introduction
of Mei [2006] who, however, destroyed such hopes were by the main result of
that paper, reproduced as Theorem 12.2.26. In combination with the upper
bound by Katz [1997] and Nazarov et al. [1997b] just mentioned, it shows that
1 + log N is the optimal upper bound for kΠb kL (L2 (R;`2N )) /kbkF (R;L (`2N )) for
any of the choices F ∈ {BMOso ∞
D , BMOD , L }.
Further relations between various BMO-type quantities and the norms
of related transformations in infinite-dimensional Hilbert spaces have been
studied by Blasco and Pott [2008, 2010]. Analogous results in the context
of the operator-valued BMOA space of analytic functions are due to Rydhe
[2017].
Section 12.3
We refer the reader to the Notes of the following section for an account of
the T (1) theorem in its more traditional meaning as a boundedness criterion
for Calderón–Zygmund operators (as in the title of David and Journé [1984]).
The section under discussion presents a rather non-canonical approach to this
theory, introduced and described by Figiel [1990] as follows:
Our approach is indirect in the following sense. Rather than trying
to prove that some “classical” operators are bounded, we start from
considering certain rather new operators, which in our opinion have
a basic nature. (All the “singularities” which can occur in our con-
text are neatly packaged inside the basic operators.) Having estab-
lished precise estimates for the norms of those basic operators, we can
take up the “general case”. We just look at the class of those oper-
ators which can be realised as the sum of an absolutely convergent
(in the operator norm) operator series whose summands are simple
compositions of our basic operators. Then it turns out that the choice
was sufficiently efficient for that class to contain so-called generalised
Calderón–Zygmund operators and much more.
A large part of this section, up to and including T (1) Theorem 12.3.26, is an
updated review of Figiel [1990], incorporating a few elaborations:
• the trade-off between the type and cotype properties of the underlying
spaces and the minimal rate of convergence of the coefficients of the bilinear
form, as in Theorem 12.3.26(ii) (which is implicit in the combination of
Figiel [1988, 1990]);
12.5 Notes 213
was already used by Figiel [1990], but it is frequently referred to as the “BCR
algorithm” after Beylkin, Coifman, and Rokhlin [1991]. They explored its ad-
vantages for the numerical evaluation of singular integrals, also making a con-
nection with the T (1) theorem but apparently independently of Figiel [1990].
A decade later in 2002, when two of the present authors started to investigate
a Banach space valued T (1) theorem (eventually published in Hytönen and
Weis [2006b]), they were also initially unaware of the work of Figiel [1990],
which was first brought to their attention by Hans-Olav Tylli. Ever since, the
approach of Figiel [1990] has been highly influential for the development of
the theory of Banach space valued singular integrals.
The second T (1) Theorem 12.3.35, which makes use of a random choice of
the dyadic system D ω , has a history of its own. This method, referred to by
its inventors as “pulling ourselves by hair”, was introduced by Nazarov, Treil,
and Volberg [1997a] to tackle the difficulties in estimating singular integrals
with respect to a non-doubling measure µ, thus going beyond the established
theory in spaces of homogeneous type due to Coifman and Weiss [1971]. Their
original idea consisted of splitting a function into its “good” and “bad” parts,
according to the “good” and “bad” cubes supporting the martingale differ-
ences DQ f : X X
ω ω
fgood := DQ f, fbad := DQ f,
ω
Q∈Dgood ω
Q∈Dbad
and showing that the latter is small, “on average”, with respect to a random
choice of ω:
ω
Ekfbad kL2 (µ) 6 εkf kL2 (µ) .
As a result, it is enough to estimate (an a priori bounded) operator T of
“good” functions only. Namely, if
ω ω ω ω
|hT fgood , ggood i| 6 Ckfgood k2 kggood k2 6 Ckf k2 kgk2 ,
then
214 12 Dyadic operators and the T (1) theorem
ω ω ω ω ω
|hT f, gi| 6 |hT fgood , ggood i| + |hT fgood , gbad i| + |hT fbad , gi|
ω ω
6 Ckf k2 kgk2 + kT kkf k2 kgbad k2 + kT kkfbad k2 kgk2 .
hence
C
kT k 6 C + 2εkT k, kT k 6 .
1 − 2ε
This method was successfully applied and further developed by Nazarov, Treil,
and Volberg [2002b, 2003]. The latter work was extended to Banach space
valued singular integrals with respect to non-doubling measures by Hytönen
[2014]. The first arXiv version of this paper was posted already in 2008, and
hence it was available to provide the backbone for the proof of the A2 theorem
in Hytönen [2012] (arXiv 2010); see the Notes of Chapter 11 for more on the
latter. It was for the purposes of the A2 theorem that a technical elaboration of
the averaging method of Nazarov, Treil, and Volberg [1997a, 2002b, 2003] had
to be invented: “on average”, the bad part is not only small but completely
absent. This allows the replacement of the estimates above by identities of the
type
ω
hT f, gi = EhTgood f, gi.
The observation that one can combine this averaging method with Figiel’s de-
composition of singular integrals in order to simplify the latter, and thereby
obtain sharper quantitative conclusions (notably, a quadratic dependence on
the UMD constant), was then made in Hytönen [2012] (arXiv 2011), where
a version of Theorem 12.3.35 (for scalar kernels and under vanishing para-
product conditions) was first established. The question of obtaining a linear
dependence on the UMD constant is an outstanding open problem already
in the special case of the Hilbert transform (see Problem O.6); but of course
a possible counterexample could be more feasible within the larger class of
operators covered by Theorem 12.3.35. A positive answer has been obtained
for sufficiently smooth even singular integrals on Lp (R; X) by Pott and Sto-
ica [2014]; their result depends on the same averaging trick and the resulting
dyadic representation theorem, but then applies different techniques to com-
plete the estimate.
While our approach to the “random” T (1) Theorem 12.3.35 took a detour
via the “non-random” T (1) Theorem 12.3.26, we should emphasise that this is
by no means necessary; rather, in many recent extensions of the T (1) theorem,
one starts with the randomised set-up from the beginning, and it is often not
even clear whether this could be avoided. We will say more about some of
these extensions later in these Notes. The reasons that we have chosen to
present also the non-random T (1) Theorem 12.3.26 are (at least) two-fold:
On the one hand, we feel that there is some historical documentary value
in providing (probably) the first detailed exposition of the original Banach
12.5 Notes 215
space valued T (1) theorem of Figiel [1990], considering also the number of
other results in the literature relying on this in their proofs (although, in
many cases, one could alternatively apply one or several of the more recent
variants). On the other hand, the non-random T (1) Theorem 12.3.26 is not in
all respects subsumed by the random T (1) Theorem 12.3.35, which makes the
first one applicable in some situations where the latter one is not, and it might
hence be useful for the reader to keep the original T (1) Theorem 12.3.26 in
their toolbox.
While we are not aware of many such applications, here is at least one:
Pseudo-localisation of singular integrals refers to estimates of the form
where
(s) (s)
[ X
Σf,s := {9Q : Q ∈ D, DQ f 6≡ 0}, DQ f := DR f,
R∈chs Q
The first term on the right with decay d + ε is typical, but the second one,
without any ε, is not. However, this term is only supported in a relatively
narrow region of values of the parameter m ∈ Zd , which still allows one to
make favourable estimates of the Figiel norms of t.
A notable aspect of this application is that the construction of the set
Σf,s refers to a fixed dyadic system D, which calls for a Haar expansion of
the operator in terms of this same D, as in the non-random T (1) Theorem
12.3.26, and seems to prevent any effective application of the random systems
D ω , as in the random T (1) Theorem 12.3.35. This suggests that, even after the
successful recent (and very likely future) development of T (1) theorems and
other results based on random dyadic systems, the non-random T (1) Theorem
12.3.26 might not become completely obsolete.
Section 12.4
The classical theory of Calderón and Zygmund [1952] had its focus on con-
volution operators. Their L2 (Rd ) boundedness is amenable to methods of
Fourier analysis, which then serves as a starting point for extrapolation to
216 12 Dyadic operators and the T (1) theorem
While also this T (b) theorem has been extended to UMD spaces by
Hytönen [2006], the need for this is perhaps not as great as in the scalar-
valued case, at least as far as the extension of the boundedness of scalar-
valued Calderón–Zygmund operators to Lp (Rd ; X) is concerned. The reason
for this is that, while it might be difficult to check the T (1) conditions (i) and
(ii) directly, they can nevertheless be verified by the converse direction of the
T (1) theorem, provided that the L2 (Rd ) boundedness of T is already known
by some other method (such as the T (b) theorem). This is, in essence, the
point of the scalar-kernel T (1) Theorem 12.4.21.
Sketch of proof. By the theorem of Coifman, McIntosh, and Meyer [1982], the
operator CA is bounded on L2 (R). It is straightforward to verify that the kernel
of CA is a standard kernel, and hence verifies the assumptions of Theorem
12.4.21 with Dini1 conditions (and associated constants depending only on A),
in which case only trivial type and cotype is needed. Thus Theorem 12.4.21,
with s = p = 2, proves the corollary for p = 2. While we could apply Theorem
12.4.21 with s = 2 and any p ∈ (1, ∞), a better quantitative conclusion for
p 6= 2 is obtained by using case p0 = 2 as input to the Calderón–Zygmund
theorem 11.2.5, which then yields the asserted bound for all p ∈ (1, ∞).
Corollary 12.5.1 seems to have been first stated in Hytönen [2006]; however,
given that it is essentially a concatenation of its scalar case due to Coifman,
McIntosh, and Meyer [1982], and the T (1) theorem of Figiel [1990], it was
probably “known to experts” much earlier. The case when X is a UMD lattice
was established by a different method already by Rubio de Francia [1986].
In a similar way, the extension of the non-homogeneous T (1) theorem of
Nazarov, Treil, and Volberg [2003] to UMD spaces has the following conse-
quence:
Note that CA is (equivalent to) the special case, where µ is the arc-length
measure on the graph {(t, A(t)) : t ∈ R}.
Sketch of proof. The implication (2)⇒(1a) is due to David [1991] and (2)⇒(1b)
due to Melnikov and Verdera [1995] and Mattila, Melnikov, and Verdera
[1996]. The sufficiency of these geometric conditions, (1)⇒(2), was proved
by Tolsa [1999].
The implication ((1a) and (2))⇒(3) follows from an analogue of Theorem
12.4.21 for measures on Rd with the power growth bound µ(B(s, r)) 6 crn
(0 < n 6 d), which is one of the main results of Hytönen [2014]. The implica-
tion (3)⇒(2) is trivial.
This proof sketch highlights the role of T (1) theorems as a device for extending
deep results about the boundedness of specific operators from scalar-valued to
vector-valued spaces, without the need to revisit the details of the original ar-
guments. Indeed, by using the scalar-valued result (2) as an intermediate step,
the equivalence of (1) and (3) is obtained without ever having to deal with
the local curvature condition (1b) in the context of vector-valued functions!
Our operator-kernel T (1) Theorem 12.4.12 is the outcome of a line of
evolution starting with the first such results obtained by Hytönen and Weis
[2006b] and Hytönen [2006], and continued with several variants and exten-
sions addressing
• non-homogeneous measures (Hytönen [2014] (arXiv 2008), Martikainen
[2012a] (arXiv 2010), Hytönen and Vähäkangas [2015]);
• simplifications of the underlying decomposition of the operator (Hytönen
[2012], Hänninen and Hytönen [2016]);
• sharper conclusions under additional symmetry assumptions (Pott and
Stoica [2014], Hytönen [2021]);
• product-space/multiparameter singularities (Di Plinio and Ou [2018], Hytönen,
Martikainen, and Vuorinen [2019a]);
• multilinear operators (Di Plinio, Li, Martikainen, and Vuorinen [2020b],
Airta, Martikainen, and Vuorinen [2022]).
12.5 Notes 219
While these papers extend the theory into several directions that we have not
considered here, many of them also provide valuable pieces of insight into the
basic case of linear Calderón–Zygmund operators on Rd with the Lebesgue
measure, which we have tried to incorporate into the present treatment. De-
spite this extensive background material, some aspects of our present T (1)
Theorem 12.4.12 appear to be new:
(1) For the first time, we are able to state an operator-valued T (1) theorem
that gives a characterisation (as in the scalar-valued T (1) theorem of
David and Journé [1984]), and not just a sufficient condition (as in all
operator-valued papers cited above), for the boundedness of a Calderón–
Zygmund operator with an operator-valued kernel. This depends on two
recent ideas, the combination of which appears here for the first time:
(a) Replacing the (sufficient but not necessary) weak R-boundedness
property of most of the previous contributions by the correct weak
DR p -boundedness property. As discussed in the Notes of Section 12.1,
this idea is from Di Plinio, Li, Martikainen, and Vuorinen [2020b].
(b) Treating the bi-paraproduct Λ = ΠT (1) + ΠT∗ ∗ (1) as a single object,
and making its boundedness into a condition in its own right, rather
than trying (in vain) to force it into a form involving some operator-
valued BMO space. This is implicit in Hytönen [2021].
(2) Recording the quantitative dependence of the estimate in terms of both
the UMD and the (co)type constants, and optimising the argument for
what seems to be the best possible bound currently available. This was
available in important special cases (notably in Hytönen [2012]), and ar-
guably implicit in some other works, but seems to be original as an explicit
statement in the present generality.
As one can see from T (1) Theorem 12.4.12 and its corollaries, the minimal
smoothness of the kernel involves a modulus of continuity kωkDiniσ , where
σ = max(1/t, 1/q 0 ) if X has cotype q and Y has type t, or one of them has
both. In the scalar-valued (or more generally Hilbert space) case, this reduces
to σ = 21 . Incidentally, this appears to be the minimal condition required to
run any known proof of the T (1) theorem, even in the scalar case. As Figiel
[1990] puts it,
it was a nice surprise that such austere methods could in fact lead to
some results which were not less general than their counterparts estab-
lished earlier with no restrictions on the range of admissible methods.
While the original T (1) theorem of David and Journé [1984] and most of its
successors are formulated for Calderón–Zygmund standard kernels, an exten-
sion to Dini-type conditions was obtained shortly after by Yabuta [1985], who
1
proved the theorem under the condition that kω 3 kDini < ∞. It is not obvious
at first sight how this compared to Figiel’s condition kωk 1 < ∞. However,
Dini 2
we may observe that any non-decreasing ω on [0, 1] satisfies
Z 1 Z 1 Z 1 ds α dt
1 α dt 1 α
ω(t) log = ω(t) 1+α ω(t) 1+α
0 t t 0 t s t
Z 1 Z 1 α
1
1 ds dt
6 ω(t) 1+α ω(s) 1+α
0 t s t
Z 1 Z 1
1 dt 1 ds α
6 ω(t) 1+α ω(s) 1+α
0 t 0 s
Z 1 1 ds 1+α
= ω(s) 1+α .
0 s
1 1
With 1+α = 13 , we see that Yabuta’s kω 3 kDini dominates kωkDiniα with α = 2.
(While the Dinis norms were previously defined with log2 in place of log,
and integrating over [0, 21 ] instead of [0, 1], the reader may easily verify that,
extending ω from [0, 21 ] to [0, 1] by ω(t) := ω(min(t, 21 )), these details affect at
most the constants in the final conclusions.)
12.5 Notes 221
The original T (1) theorem of David and Journé [1984] was a characterisation
of boundedness on L2 (Rd ), while we have dealt with extensions of such re-
sults to Lp (Rd ; X). However, the boundedness of a given (singular integral)
operator is basic question arising in several other function spaces as well, and
the T (1) theorem has served as a model for similar results in other spaces.
(See Chapter 14 for information about the functions spaces appearing in this
s
discussion.) Extensions of the T (1) theorems to Besov spaces Ḃp,q were ob-
s s
tained by Lemarié [1985] and to Triebel–Lizorkin spaces Ḃp,q and Ḟp,q by
Frazier, Han, Jawerth, and Weiss [1989]. In these results, p, q ∈ [1, ∞], and
12.5 Notes 223
Tm f = (mfb)∨ = m
b ∗ f = k ∗ f,
where the right-hand side has the formal structure of the operators studied in
Chapter 11, but the question then becomes the correspondence of the condi-
tions on the multiplier m and on the singular convolution kernel k. As we will
see in Section 13.2.a, the function k will be a nice Calderón–Zygmund kernel,
and hence f 7→ k ∗ f will be in the scope of all results of Chapter 11 (notably,
including those dealing with extrapolation of boundedness to the weighted
Lp (w; X) spaces), as soon as m satisfies assumptions like those in the Mihlin
Multiplier Theorem 5.5.10 for sufficiently many derivatives ∂ α m. Moreover,
this result is very general in that it holds for multipliers taking values in arbi-
trary Banach spaces, and then in particular in L (X, Y ) for any Banach spaces
X and Y . However, the required number of derivatives on this level of gener-
ality is higher than that in the Mihlin Multiplier Theorem 5.5.10. Coping only
with the same derivatives as in Mihlin’s theorem turns out to be more deli-
cate and require the use of a Banach space valued Hausdorff–Young inequality
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 225
T. Hytönen et al., Analysis in Banach Spaces, Ergebnisse der Mathematik und ihrer
Grenzgebiete. 3. Folge / A Series of Modern Surveys in Mathematics 76,
https://doi.org/10.1007/978-3-031-46598-7_3
226 13 The Fourier transform and multipliers
0
extends to a bounded operator from `p (Zd ; X) to Lp (Td ; X).
Denoting the norms of the respective extensions (or restrictions) by ϕp,X (Rd ),
ϕp,X (Td ) and ϕp,X (Zd ), we have:
Lemma 13.1.2. Let X be a Banach space and p ∈ (1, ∞). Let f ∈ Lp (Td ; X)
be a trigonometric polynomial, which we identify with its periodic extension
to Rd , and let φ ∈ S (Rd ; X). Then
lim kf (·)φ(ε·)εd/p kLp (Rd ;X) = kf kLp (Td ;X) kφkLp (Rd ) ,
ε↓0
X Z
F [f (·)φ(ε·)εd/p ](ξ) = xk φ(εt)εd/p e2πik·t e−2πiξ·t dt
Rd
k∈Zd
228 13 The Fourier transform and multipliers
b −1 (ξ − k))ε−d/p0
X
= xk φ(ε
k∈Zd
6 ϕp,X (Rd )kf ⊗ φkLp (Rd ;X) = ϕp,X (Rd )kf kLp (R;X) kφkLp (Rd−1 ) .
Choosing f and φ that (almost) achieve equality in the definition of the con-
stants ϕp,X (R) and ϕp,C (Rd−1 ), we obtain the first bound in (13.2).
The first equality in (13.3) is Proposition 2.4.16. The first pair of inequal-
ities and the two equalities in the middle of in (13.3) are all contained in
Proposition 2.4.20 (either as stated or substituting X ∗ in place of X).
Concerning the last pair of inequalities in (13.3), it suffices to prove that
ϕp,X (Rd )
ϕp,X (Td ) 6 , (13.4)
ϕp,C (Rd )
13.1 Bourgain’s theorem on Fourier type 229
since the other bound follows with X ∗ in place of X and using the first equality
in (13.3). To this end, it follows from Lemma 13.1.2 that
b p0 d = lim kF [f (·)φ(ε·)εd/p ]k p0 d
kfbk`p0 (Zd ;X) kφk L (R ) L (R ;X)
ε↓0
(p0 )
holds for every choice of x1 , . . . , xn ∈ X. We abbreviate ϕp,X (n) := ϕp,X (n).
Lemma 13.1.4. Let X be a Banach space and p, q ∈ [1, ∞]. The sequence
(q)
(ϕp,X (n))n>1 is increasing, and
(q) 0
1 6 ϕp,X (n) 6 n1/p , ϕp,X (Z) = lim ϕp,X (n) ∈ [1, ∞].
n→∞
n
Given (xk )k=1 , let x = (xk )k∈Z be its zero
Pnextension. The upper bound
ϕp,X (n) 6 ϕp,X (Z) follows by observing that k=1 ek (t)xk is simply x b(−t).
It only remains to check that ϕp,X (Z) 6 limn→∞ ϕp,X (n). Let x = (xk )k∈Z
be finitely supported, i.e., xk = 0 if |k| > N for some finite N . Now
X 2N
X −1 2N
X −1
x
b= e−k xk = e−N +j xN −j = e−N ej xN −j ,
|k|6N −1 j=1 j=1
hence
2N
X −1 2N
X −1 1/p
kb
xkLp0 (T;X) = ej xN −j 6 ϕp,X (2N − 1) kxN −j kp
Lp0 (T;X)
j=1 j=1
By the density of finitely supported sequences in `p (Z; X), this shows that
ϕp,X (Z) 6 limn→∞ ϕp,X (n), and completes the proof.
The task of proving that a space X has non-trivial Fourier type (assuming
non-trivial type) is hence reduced, in principle, to showing the boundedness
of the sequence (ϕp,X (n))n>1 for some p > 1. Although the proof that we
are about to give is eventually set up slightly differently, this idea serves as a
good motivation for a major part of the subsequent analysis. The proof that
we will present can be roughly divided into the following main steps, treated
in the next four sections:
1. Using type bounds on Sidon sets that partition {1, . . . , n} gives a first mild
improvement ϕ2,X (n) = o(n1/2 ) over the trivial estimate ϕ2,X (n) 6 n1/2 .
2. Comparison with the finite Fourier transform on Zn gives sub-multi-
plicativity and leads to ϕ2,X (n) = O(n1/r−1/2 ) for some r > 1.
3. By a delicate Lemma 13.1.25 of Bourgain, this gives a first uniform bound
(2)
ϕs;X (n) = O(1), but with mismatched exponents s ∈ (1, r) and 2 6= s0 .
4. Standard duality and interpolation, combined with repeating the same
key Lemma 13.1.25 on the dual side, allow us to conclude with p ∈ (1, r).
A thorough reader may recognise some conceptual similarity with the consid-
erations encountered in Section 7.3.b in the context of deducing non-trivial
type (and cotype) from the non-containment of certain subspaces. There we
defined the finite type constant τ2,X (n) as the best constant in the estimate
n
X n
X 1/2
εk x k 6 τ2,X (n) kxk k2 ∀x1 , . . . , xn ∈ X. (13.5)
L2 (Ω;X)
k=1 k=1
These numbers will play a role in the first proof step outlined above.
13.1 Bourgain’s theorem on Fourier type 231
Recall that our goal is deriving non-trivial Fourier P type from non-trivial type.
Thus, from the knowledge that random sums k εk xk can be dominated P by
k(xk )k`p , we would like to conclude that trigonometric sums e x
k k k can
be similarly dominated (though possibly with a different p). An obvious idea
that suggests itself is to try to dominate the trigonometric sum by the random
sum. Indeed, we know from Section 6.5 that this can be done under particular
circumstances if the trigonometric sum is restricted to a special set called a
Sidon set. This leads to the following strategy: Given the initial sum over k ∈
{1, . . . , N }, we want to partition this into sums over Sidon sets on which we can
make estimates, and this partitioning should be done sufficiently economically
so that it allows us to beat the trivial estimate. To carry out this idea, we
need to be able to
1. efficiently recognise Sidon sets, and
2. decompose arbitrary sets into as few as possible Sidon sets.
We now turn to these tasks. Recall from Section 6.5 that a subset Λ ⊆ Z
is called a Sidon set if the following estimate holds uniformly over all finitely
non-zero sequences (cλ )λ∈Λ of complex numbers:
X X
|cλ | 6 C c λ eλ .
∞
λ∈Λ λ∈Λ
Example
P∞ 13.1.6. The sequence {2k : k ∈ N} is quasi-independent. In fact, if
α
k=0 k 2 k
= 0 for a finitely non-zero sequence (αk )∞
k=1 , then
X X
2k = 2k .
k:αk =+1 k:αk =−1
By Example 13.1.6, this gives another proof of the fact that {2k : k ∈ N} is a
Sidon set, but with a slightly weaker constant than Proposition 6.5.3.
Proof. This is based on a variant of the Riesz product method also used in
the proof of Proposition 6.5.3, but the details are somewhat different, and we
will provide a self-contained argument. By considering every finite subset of
the original F , we may assume without loss of generality that F is finite to
begin with. Given parameters % ∈ (0, 1] and ξ = (ξk )k∈F ∈ RF , let then
Y
Rξ (t) := 1 + % cos(2π(kt + ξk ))
k∈F
Y %
= 1 + (ek (t)e1 (ξk ) + e−k (t)e−1 (ξk ))
2
k∈F
X X X
= 2−|α| %|α| exp 2πi αk · kt exp 2πi α k ξk ,
α∈{−1,0,+1}F k∈F k∈F
P
where |α| := k∈F |αk | as usual for multi-indices,. (To relax the notation, we
do not explicate the dependence of Rξ on %.)
From the assumption that F is quasi-independent, it follows that
X
αk · k = 0 only if αk ≡ 0,
k∈F
and hence R bξ (0) = 1. It is also clear from the first line of the definition of
Rξ (t) (recalling that % ∈ (0, 1]) that Rξ (t) > 0, and hence
Z 1
kRξ kL1 (T) = Rξ (t) dt = R
bξ (0) = 1.
0
so that
#F
(m)
X
Rξ (t) = %m R ξ (t), where
m=0
(0) (1) 1X
Rξ (t) = 1, Rξ (t) = ek (t)e1 (ξk ) + e−k (t)e−1 (ξk ) .
2
k∈F
(m) (m)
where R0 is simply Rξ with ξ = 0. It follows that
#F Z 1 #F Z 1 Z 1
(m) (m)
X X
Rξ e j 6 R0 e j = R0 ej 6 kR0 kL1 (T) = 1.
m=0 0 m=0 0 0
and hence
% X X
− %2 |cj | 6 kf kL∞ (T;X) = ck ek .
2 L∞ (T;X)
j∈F k∈F
1
Choosing finally % = 4 completes the proof.
By the previous result, our initial task of decomposing arbitrary sets into
Sidon sets is reduced to decomposing into quasi-independent sets. A first step
in this direction is to know that every set has a quasi-independent subset of
somewhat substantial size.
Lemma 13.1.8. Any finite subset F ⊆ Z\{0} has a quasi-independent subset
F0 ⊆ F of cardinality #F0 > dlog3 #F e.
Proof. Let F0 ⊆ F be a quasi-independent subset of maximal cardinality, and
let nX o
F1 := αk · k : αk ∈ {−1, 0, +1} .
k∈F0
contradicting k0 ∈
/ F1 . Thus αk0 = 0, but then also αk = 0 for all k ∈ F0 , since
F0 is quasi-independent, and this proves that F0 ∪ {k0 } is quasi-independent.
As explained above, this proves that F1 ⊇ F , and hence
#F 6 #F1 6 3#F0 ,
from which the proposition follows, since #F0 > log3 #F is necessarily an
integer.
By recursively removing big quasi-independent subsets, we arrive at the de-
sired decomposition of the initial set:
Lemma 13.1.9. For N ∈ Z+ , let
d(N ) := min{k ∈ Z+ : any subset F ⊆ Z \ {0} of size #F 6 N can be
divided into at most k quasi-independent subsets}.
n
2·3
Then d(3n ) 6 for all n ∈ N. For all n > 1, each of the partitioning
n+1
quasi-independent subsets can be chosen to have size at most n.
13.1 Bourgain’s theorem on Fourier type 235
thus
2 · 3n 2 · 3n
6j <1+ .
n+1 n+1
The remaining set of size at most 3n can then be divided into at most d(3n )
quasi-independent subsets, and by the induction assumption we have
2 · 3n 2 · 3n 4 · 3n
d(3n+1 ) 6 j + d(3n ) < 1 + + =1+ .
n+1 n+1 n+1
For n > 2, we have 1/(n + 1) 6 34 /(n + 2) and 3n /(n + 2) > 49 , and hence
4 16 3n 3n 2 · 3n+1
d(3n+1 ) 6 + <6 = ,
9 3 n+2 n+2 (n + 1) + 1
and this completes the induction step. Note that all quasi-independent subsets
that we constructed in the induction step had either size n + 1, or they came
from the induction assumption, in which case their size is at most n.
In the next remark, we indicate converses to the obtained bounds of Lemmas
13.1.8 and 13.1.9.
Remark 13.1.10. Let F0 ⊆ F := {1, . . . , N } be such that N > 2 and F0 is
quasi-independent. We claim that necessarily #F0 6 2 log2 (N ). Clearly, this
implies d(N ) > 2 logN (N ) . Indeed, write m = #F0 . It suffices to consider m > 2.
2
Let A ⊆ F0 be arbitrary. Then
X X
06 a6 a < N 2.
a∈A a∈F0
We now possess all the ingredients needed for the first estimate of Fourier type
in terms of type, stated in terms of the finite versions of both properties. The
reader may wish to compare the next proposition to Theorem 7.6.12 which
gives a related inequality for the Walsh system.
A 6 2 · 3n /(n + 1)
The following
√ corollary gives the promised improvement over the trivial bound
ϕ2,X (3n ) 6 3n as soon as n is large enough.
13.1 Bourgain’s theorem on Fourier type 237
Corollary 13.1.12. Let X be a Banach space of type p ∈ (1, 2]. Then for all
n > 1, we have
ϕ2,X (3n ) √ 0
√ 6 16 2 · τp,X;2 · n−1/p .
3 n
(q)
As the notation suggests, ϕp,X (Zn ) has an interpretation as the norm of the
Fourier transform (thus, a Fourier type constant) of functions on the finite
group Zn = Z/nZ, but there is no need to insist too much on this point here.
(q) (q)
The difference of the defining inequalities of ϕp,X (n) and ϕp,X (Zn ) is that
p
the L (T; X) integral norm in the former is replaced by a finite Riemann sum
approximation in the latter. We will next develop some tools for comparing
the two kinds of norms. This will involve elements of some fairly classical
Fourier analysis, and we begin with
238 13 The Fourier transform and multipliers
Since
n n
X X ei2π(n+1)t − 1
sin(π(2k + 1)t) = = eiπt ei2πkt = = eiπt
ei2πt − 1
k=0 k=0
sin(π(n + 1)t) sin2 (π(n + 1)t)
= = eiπ(n+1)t = ,
sin(πt) sin(πt)
we obtain the formula for Fn by summing over the formula for Dk . Finally,
2n−1 2n−1 n−1
1 X 1 X X 1
Vn = Dk = Dk − Dk = 2nF2n−1 − nFn−1 .
n n n
k=n k=0 k=0
Lemma 13.1.16. If f is a trigonometric polynomial with deg(f ) < n, then
for all s ∈ R we have
Z 1 n
1X
f (t) dt = f (s + h/n),
0 n
h=1
Proof. It is enough to consider f (t) = ek (t), where |k| < n. We observe that
n
( 2πikn/n
−1
X
2πikh/n e2πik/n ee2πik/n −1 = 0, 0 < |k| < n,
e =
h=1
n, k = 0,
and hence
n n Z 1
1X ek (s) X 2πikh/n
f (s + h/n) = e = ek (s)δk,0 = δk,0 = ek (t) dt.
n n 0
h=1 h=1
On the level of Lp norms, this leads to the following comparison result:
With the usual modification, the result is also true (and entirely trivial) for
p = ∞: of course the supremum over {j/n : j = 1, . . . , n} is dominated by the
supremum over all of T!
Proof. Let
n
X
f (t) := ek (t)xk , m := bn/2c.
k=1
−m = 1 − (m + 1) 6 k 6 n − (m + 1) 6 (2m + 1) − (m + 1) = m,
We now have the desired comparison of the two finite Fourier type constants:
Lemma 13.1.18. For any Banach space X and n ∈ Z+ , we have
(q) (q) (q)
ϕp,X (n) 6 ϕp,X (Zn ) 6 3 · ϕp,X (n).
we obtain
n Z n q n q/p
1X X (q) q X
ek (t)xk dt 6 ϕp,X (Zn ) kxk kp ,
n T
h=1 k=1 k=1
13.1 Bourgain’s theorem on Fourier type 241
(q) (q)
and hence ϕp,X (n) 6 ϕp,X (Zn ).
The other estimate follows at once from the Marcinkiewicz inequality
(Proposition 13.1.17), which is the first step in
n
1 X n
X q 1/q n
X
ek (h/n)xk 63 ek xk
n Lq (T;X)
h=1 k=1 k=1
n
X 1/p
(q)
6 3ϕp,X (n) kxk kp .
k=1
(q)
The following lemma is our reason for considering the quantities ϕp,X (Zn ):
Lemma 13.1.19. For any Banach space X and m, n ∈ Z+ , we have the sub-
multiplicative estimate
(q) (q) (q)
ϕp,X (Zmn ) 6 ϕp,X (Zm )ϕp,X (Zn ), 1 6 p 6 q 6 ∞;
in particular
Proof. The second estimate is an obvious special case with q = p0 > 2 > p.
For the proof of the general estimate, it is convenient to observe that, by
(q)
simple reindexing and modular arithmetic, the condition defining ϕp,X (Zn )
is unchanged if instead of {1, . . . , n} we take all sums over {0, . . . , n − 1}. In
(q)
the defining condition of the constant ϕp,X (Zmn ), we should then sum over
{0, . . . , mn − 1}, and the key trick of the proof is to use a non-symmetric
reindexing of this range for the h and k sums, namely
h = an + b : a = 0, . . . , m − 1, b = 0, . . . , n − 1,
k = um + v : u = 0, . . . , n − 1, v = 0, . . . , m − 1.
Then
hk = (an + b)(um + v) = aumn + avn + bum + bv,
and hence, noting that e2πiau = 1,
Thus
n 1 mn−1
X mn−1
X q o1/q
ek (h/mn)xk
mn
h=0 k=0
n 1 n−1
X 1 m−1
X m−1
X q o1/q
= ev (a/m)yv(b) ,
n m a=0 v=0
b=0
242 13 The Fourier transform and multipliers
n−1
X
yv(b) := ev (b/mn) eu (b/n)xum+v ,
u=0
n 1 n−1 m−1 q/p o1/q
(q)
X X
6 ϕp,X (Zm )q kyv(b) kp
n v=0
b=0
X 1 n−1
n m−1 X n−1 q p/q o1/p
(q)
X
6 ϕp,X (Zm ) eu (b/n)xum+v
v=0
n u=0
b=0
(q)
where we used the defining condition for ϕp,X (Zm ) with the sequences
(b) (q)
(yv )m−1
v=0 for each fixed b = 0, . . . , n − 1, and that for ϕp,X (Zn ) with the
sequences (xmu+v )n−1
u=0 for each fixed v = 0, . . . , m − 1.
Combining the above results with Corollary 13.1.12 of Hinrichs’s inequal-
ity, we achieve the desired power-type improvement over the trivial estimate
ϕ2,X (N ) 6 N 1/2 . One could try to deduce this from Lemma 7.3.19 applied
to ϕ2,X (Zn ). However, this time that does not work since we do not know
whether ϕ2,X (Zn ) is increasing in n. Therefore, we adapt the proof of the
lemma and use the facts that ϕ2,X (n) is increasing and that ϕ2,X (Zn ) is sub-
multiplicative. Our choice of notation r0 below is indicative of the fact that
this is the Hölder conjugate of a (small) exponent r > 1.
where
0 r0
r0 := 3p0 (68 · τp,X;2 )p , C := e 2p0 . (13.10)
Then
For appropriate n and s, we will show that the term within brackets satisfies
[. . .] 6 1. By Corollary 13.1.12, we can estimate
1 1 √ 0 0 0
[. . .] 6 3n( 2 − s ) · 3 · 16 2 · τp,X;2 · 3n/2 · n−1/p =: 3n/s T n−1/p ,
√ 0
where T := 48 2τp,X;2 . Therefore, setting s0 = (1 + eT p )p0 log(3) and taking
0 0
eT p 6 n < eT p + 1 we find that
0
0 0 0 e−1/p
3n/s T n−1/p 6 e1/p T = 1.
T
From the above we conclude that
1 1 0 0
N 1/2−1/s ϕ2,X (N ) 6 3n( s − 2 ) 6 3n/2 = en log(3)/2 6 es /(2p ) .
Proposition 13.1.21. If X has type p and cotype q, then for all n > 1,
Proof. Let (γh )h>1 be a complex Gaussian sequence (i.e., standard indepen-
dent Gaussian random variables). Also let
n
1 X h
ek = √
γ γh ek ( ), k = 1, . . . , n.
n n
h=1
244 13 The Fourier transform and multipliers
n
Then (eγk )k=1 are also independent standard Gaussian random variables (see
Section E.2). Hence, using the natural Gaussian analogue of the finite type
and cotype constants,
n n 2 n n 2
1X X h 1 γ X X h
xk ek ( ) 6 c2,X (n)2 E γh xk ek ( )
n n n n
h=1 k=1 h=1 k=1
n
X 2
= cγ2,X (n)2 E γ
ek xk
k=1
n
X
6 cγ2,X (n)2 τ2,X
γ
(n)2 kxk k2 .
k=1
γ
Since kγk2 = 1, Proposition 7.1.18 informs us that τ2,X 6 τ2,X and cγ2,X 6
γ
c2,X , and the analogous result for the finite constants τ2,X (n) etc. follows by
1 1
the same argument. Finally, Hölder’s inequality implies τ2,X (n) 6 τp,X;2 n p − 2
1 1
and c2,X (n) 6 cq,X;2 n 2 − q .
N
X 1 X
fb(j) = e−1 (nh/N )f (h/N ).
N
j≡n mod N h=1
to find that
N
1 X X X
(e−n f )(h/N ) = fb(j + n) = fb(j),
N
h=1 j≡0 mod N j≡n mod N
1 1 1
In := [n − , n + ), n = 1, . . . , N,
N 2 2
satisfy t0 k ∈ In + Z for some k ∈ F .
Proof. We in fact show that this is true for the “average” choice of t0 ∈ T.
For t ∈ T and n = 1, . . . , N , we denote
The claim is then that N (t0 ) > 81 N for some t0 ∈ T, and we will prove that
Z 1
1
N (t) dt > N, (13.12)
0 8
Z 1 1/2 Z N
1X 1/2
N6 N (t) dt νn (t)2 dt ,
0 0 n=1
Now X X
νn (t) = 1In +Z (kt) = 1I0 +Z (kt − n/N ).
k∈F k∈F
1 1
s(t) := (1 − N |t|)+ , t ∈ [− , ).
2 2
An elementary computation of the Fourier coefficients shows that
(
1 2 1/N, j = 0,
sb(j) = sinc (πj/N ) = (13.14)
N 0, 0 6= j ≡ 0 mod N.
Note that the first equality above is valid for all j ∈ Z, although in the second
we only consider particular cases. Clearly 0 6 sb(j) = O(j −2 ), so that Lemma
13.1.22 applies to f = s. Since s(h/N ) = 1N Z (h) for h ∈ Z, the conclusion of
the lemma takes a particularly clean form, namely
X 1
sb(j) = ∀n = 1, . . . , N. (13.15)
N
j≡n mod N
This confirms (13.13) and hence, as explained in the beginning of the proof,
the assertion of the Lemma.
From the comparison between `pN (X) and `∞
N (X), it is immediate that
(q) (q)
ϕ∞,X (N ) 6 ϕp,X (N ) · N 1/p .
(q)
and ϕp,X (N ) > 1, we have the trivial estimate
(q)
Since F \ F0 ( F is a strictly smaller set and ϕp,X is clearly non-decreasing,
the induction hypothesis applies to show that
(q)
II 6 ϕ∞,X (F \ F0 ) max kxk k
k∈F \F0
(13.16)
(q)
6 Aϕp,X (N )#(F \ F0 )1/p max kxk k.
k∈F
=: I1 (h) + I2 (h),
where (using again the induction hypothesis, now with the smaller set F0 ( F )
(q) ψ(k)
I1 (h) 6 ϕ∞,X (F0 ) max exp(2πikht0 ) − exp(i2πh ) kxk k
k∈F0 N
(q)
ψ(k)
6 Aϕp,X (#F0 )1/p max 2π|h| inf kt0 − − j max kxk k (13.18)
k∈F0 j∈Z N k∈F0
(q) π|h|
6 Aϕp,X (N )N 1/p max kxk k,
N k∈F
13.1 Bourgain’s theorem on Fourier type 249
H := bH 0 c, H 0 := A−q/(q+1) N.
0
Since A > 1, we have H 6 H 0 6 N . Recalling that N > Ap , and noting that
p0 > 1 > q/(q + 1), we also observe that H 0 > 1, and hence H > 1. Thus
this choice of H lies in the admissible range considered above. We also have
H 0 6 H + 1 6 2H, and thus
πH N 1/q πH 0 2N 1/q
A +3 6A +3 = (π + 21/q · 3)A1/(q+1) .
N H N H0
250 13 The Fourier transform and multipliers
(q)
h 1 i (q)
ϕ∞,X (F ) 6 (π + 21/q · 3) · A1/(q+1) + 1 − A ϕp,X (N ) · N 1/p .
8p
To complete the induction step, it remains to check that the quantity in
brackets is at most A, which after easy simplification is the same as
1
(π + 21/q · 3) · A1/(q+1) 6 A.
8p
Clearly this is the case with the choice of A stated in the Lemma.
We are now ready for a first uniform bound on the finite Fourier type con-
stants:
Corollary 13.1.27. Let X be a Banach space, r ∈ (1, 2], and suppose that
ϕ2,X (N ) 6 C · N 1/r−1/2 ∀N ∈ Z+ .
(2) Cr
ϕs,X (N ) 6 3500 ∀N ∈ Z+ .
r−s
Proof. By Lemma 13.1.25 and Remark 13.1.26 with p = q = 2, we have
(2)
ϕ∞,X (F ) 6 1285 · ϕ2,X (N ) · N 1/2 6 1285 · C · N 1/r (13.20)
Thus
N ∞ ∞
(j) (2) (j)
X X X X
ek xk 6 ek xk 6 ϕ∞,X (Fj ) max kxk k
L2 (T;X) L2 (T;X) k∈Fj
k=1 j=1 k∈Fj j=1
13.1 Bourgain’s theorem on Fourier type 251
∞
(2)
X
6 1285 · ϕ2,X (#Fj ) · (#Fj )1/2 · αj−1 by (13.20)
j=1
X∞ ∞
X
6 1285 · C(#Fj )1/r αj−1 6 1285 · Cα−js/r αj−1
j=1 j=1
1−s/r
1285 · C α
= .
α 1 − α1−s/r
The choice α = (s/r)r/(r−s) gives
With the uniform bound of Corollary 13.1.27, we have already covered the
core of the deep implication from non-trivial type to non-trivial Fourier type.
The rest of the argument depends on the more routine techniques of duality
and interpolation, but is still not entirely straightforward. We now turn our
attention to giving these finishing touches to the proof. At the end of this
section, a statement and proof of Bourgain’s theorem will finally be given.
The first duality that we want to use is most elegantly expressed in terms
of the Fourier type constants on the cyclic group ZN :
Proof. Since X is norming for X ∗ , Proposition 1.3.1 shows that `pN (X) is
0
norming for `pN (X ∗ ), so that
N
X N
X p0 1/p0
ek (h/N )x∗k
h=1 k=1
N D
nX N
X N
E X 1/p o
= sup xh , ek (h/N )x∗k : kxh kp 61 ,
h=1 k=1 h=1
Permuting the names of the exponents and using the isometric embedding of
X into X ∗∗ , it also follows that
(q) (q) 0 (p0 )
N 1/q ϕp,X (ZN ) 6 N 1/q ϕp,X ∗∗ (ZN ) 6 N 1/p ϕq0 ,X ∗ (ZN ),
Corollary 13.1.29. Let X be a Banach space, r ∈ (1, 2], and suppose that
ϕ2,X (N ) 6 C · N 1/r−1/2 ∀N ∈ Z+ .
(s0 ) Cr
ϕ∞,X ∗ (F ) 6 1.35 · 107 N 1/s ∀s ∈ (1, r),
r−s
whenever F ⊆ Z is a subset of size #F = N .
Recall from Corollary 13.1.20 that if X has type p ∈ (1, 2], then the assump-
tion is satisfied with C and r as in (13.10).
Proof. By using both estimates of Lemma 13.1.18 with Lemma 13.1.28 in
between, and finally Corollary 13.1.27, we have
0 (s0 ) 0 (s0 ) (2)
N 1/s ϕ2,X ∗ (N ) 6 N 1/s ϕ2,X ∗ (ZN ) = N 1/2 ϕs,X (ZN )
(2) Cr
6 N 1/2 · 3ϕs,X (N ) 6 N 1/2 · 3 · 3500 .
r−s
Then Lemma 13.1.25 and Remark 13.1.26 with p = 2 6 q = s0 show that
(s0 ) (s0 ) Cr 0
ϕ∞,X ∗ (F ) 6 1285 · ϕ2,X ∗ (N ) · N 1/2 < 1.35 · 107 · N 1/2+1/2−1/s .
r−s
whenever F ⊆ Z is a subset of size #F = N .
13.1 Bourgain’s theorem on Fourier type 253
We now come to another form of duality, where we pass from the Fourier
transform on Z to that on the circle T, and it is in this latter setting that our
argument will be completed.
Lemma 13.1.30. Let X be a Banach space, 1 6 s 6 ∞, and suppose that
(s0 )
ϕ∞,X ∗ (F ) 6 K · N 1/s
Proof. Let f ∈ Ls (T; X), let λ > 0, and let F be a finite subset of {k ∈ Z :
kfb(k)k > λ}. (By a periodic analogue of the Riemann–Lebesgue Lemma 2.4.3,
which has essentially the same proof, we could argue that this set is finite to
begin with, but we do not need this here.) Then
1 X b 1 X b
#F 6 kf (k)k = hf (k), x∗−k i
λ λ
k∈F k∈F
for suitable x∗−k ∈ X ∗ of norm one
Z
1 X
= f (t) e−k (t)x∗−k dt
λ T
k∈F
1 X
6 kf kLs (T;X) e k x k s0
λ L (T;X ∗ )
k∈−F
1 (s0 ) 1
6 kf kLs (T;X) ϕ∞,X ∗ (−F ) 6 kf kLs (T;X) K(#F )1/s ,
λ λ
and hence
λ(#F )1−1/s 6 Kkf kLs (T;X) .
Since this is true for any finite F ⊆ {k ∈ Z : kfb(k)k > λ}, it is also true
for F = {k ∈ Z : kfb(k)k > λ} (showing, a posteriori, the finiteness of this
set). Then the supremum over λ > 0 of the left-hand side is precisely the
0
`s ,∞ (Z; X) norm that we wanted to estimate.
From (13.21) and the trivial fact that F is bounded from L1 (T; X) →
`∞ (Z; X), it seems apparent that we should conclude that F is bounded
0
from Lp (T; X) to `p (Z; X) by interpolation. However, the version of the
Marcinkiewicz Interpolation Theorem 2.2.3 covered in the text is not suffi-
cient for this purpose, and we would need the generalisation stated in the
Notes as Theorem 2.7.5. We will give a proof of a quantitative version of the
special case relevant for the present application:
254 13 The Fourier transform and multipliers
Lemma 13.1.31. Let X be a Banach space such that (13.21) holds for some
s ∈ (1, 2]. Then
3K
kF f k`t0 (Z;X) 6 kf kLt (T;X) ∀t ∈ (1, s).
(s − t)1/t0
Then
Z
kf λ kL1 (T;X) = kf kX 6 (Aλγ )1−t kf ktLt (T;X) = (Aλγ )1−t
{kf kX >Aλγ }
and hence
kfbλ k`∞ (Z;X) 6 (Aλγ )1−t 6 θ1 λ,
provided that we choose
−1/(t−1)
γ = −1/(t − 1), A = θ1 .
Then the second term on the right of (13.22) vanishes, and subsequently
Z ∞
t0 0
kF f k`t0 (Z;X) 6 t0 λt −1 #{k : kfbλ (k)k > θ0 λ} dλ
Z0 ∞
0 0 0 0
6 t0 λt −1 (θ0 λ)−s K s kfλ ksLs (T;X) dλ
0
by Lemma 13.1.30
K s0 Z ∞ 0 0 0
s0 /s0
= t0 λt −s −1 kfλ ksLs (T;X) dλ
θ0 0
K s0 Z ∞ 0 0 0
1/s0 s0
6 t0 λt −s −1 kfλ ksX dλ
θ0 0 Ls (T)
so that
1/s0 s0 s0
s0 −(t−1)(t0 −s0 ) ts0 /s
t/s
kf kX = kf kX = kf kLt (T;X) = 1.
Ls (T) Ls (T)
0
Taking θ0 = θ1 = 1
2 and using (t0 )1/t 6 e1/e < 23 , we obtain
0 0 0 0 0
2(t0 )1/t K s /t 3K s /t
kF f k`t0 (Z;X) 6 0 6 .
(t0 − s0 )1/t (t0 − s0 )1/t0
Lemma 13.1.32. Let X be a Banach space, and suppose that there are con-
stants C and r ∈ (1, 2] such that
ϕ2,X (N ) 6 C · N 1/r−1/2
109 · C
ϕt,X 6 .
(r − t)1+1/t0
(s0 ) Cr
ϕ∞,X ∗ (F ) 6 1.35 · 107 N 1/s =: K · N 1/s
r−s
whenever F ⊆ Z is a subset of size #F = N .
By Lemma 13.1.30, it follows that
256 13 The Fourier transform and multipliers
3K 5 · 107 · C · r
ϕt,X (T) := kF kL (Lt (T;X),`t0 (Z;X)) 6 0 6
(s − t) 1/t (r − s)(s − t)1/t0
for all 1 < t < s < r. Optimising the bound with respect to s in this range,
we choose
t2 + (t − 1)r
s= .
2t − 1
With this choice, a computation shows that
t(r − t) 1 (r − t)(t − 1) 1
r−s= > (r − t), s−t= > (r − t)(t − 1).
2t − 1 3 2t − 1 3
Substituting back,
0
7 31+1/t
ϕt,X (T) 6 5 · 10 · C · r ,
(r − t)1+1/t0 (t − 1)1/t0
0
where r 6 2 and 31+1/t 6 33/2 and, for t ∈ (1, 2),
0
(t − 1)1/t = [(t − 1)t−1 ]1/t > [e−1/e ]1/t > e−1/e .
Thus
33/2 · e1/e 109 · C
ϕt,X (T) 6 108 · C 0 6 .
(r − t)1+1/t (r − t)1+1/t0
We are finally ready for the main theorem:
with constants
0
ϕt,X (R) 6 ϕt,X (T) 6 exp 2(68 · τp,X;2 )p .
13.1 Bourgain’s theorem on Fourier type 257
6 κr,2,K {xn }N
n=1 6 κr,2,K {xn }N
n=1 .
Lr (S;`2N ) `2N (Lr (S))
For the reverse estimate, it suffices to pick some non-zero φ ∈ Lr (S) and
observe that the type inequality for xn = an φ ∈ X implies the Kahane–
Khintchine inequality for an ∈ K.
The fact that X has Fourier-type t if t ∈ [1, r0 ] follows from the scalar-
valued Hausdorff–Young inequality and Minkowski’s inequality:
kfbkLt0 (R;Lr (S)) 6 kfbkLr (S;Lt0 (R)) 6 kf kLr (S;Lt (R)) 6 kf kLt (R;Lr (S))
We indicate two alternative proofs of the “only if” part:
258 13 The Fourier transform and multipliers
(1) In Example 2.1.15, it is verified directly that the Fourier transform is not
0 0 0
bounded from Lp (R; `r ) to Lp (R; `r ) for p ∈ (r0 , 2]. By duality, it is also
0
not bounded from Lp (R; `r ) to Lp (R; `r ).
(2) Proposition 7.3.6 says that if X has Fourier type p, then it has cotype p0 .
But Corollary 7.1.6 says that Lr (S) has cotype p0 only for p0 ∈ [r, ∞].
This concludes the verification of the example.
We also record the following simpler variant, which is nevertheless sufficient
for many purposes:
1
Proposition 13.1.35. Let X have type p and cotype q, where p − 1q < 12 . Let
1 1 1 1 h1
:= + − ∈ , 1 .
r 2 p q 2
Then X has every Fourier-type t ∈ (1, r), and
τp,X;2 cq,X;2
ϕt,X (R) 6 ϕt,X (T) 6 109
(r − t)1+1/t0
Proof. By Proposition 13.1.21, we have
1 1 1 1
ϕ2,X (N ) 6 N p − q = N r − 2 , C := τp,X;2 cq,X;2
Thus Lemma 13.1.32 implies the bound for ϕt,X (T), and Proposition 13.1.1
the bound for ϕt,X (R).
The notation MLp (Rd ; X, Y ) stands for the space of all m ∈ L∞ (Rd ; L (X, Y ))
for which Tm extends to a bounded linear operator from Lp (Rd ; X) to
Lp (Rd ; Y ). The connection of Fourier multipliers to integral operators is par-
ticularly simple in the following special case:
13.2 Fourier multipliers as singular integrals 259
c (R ; L (X, Y )).
Let X, Y be Banach spaces and m ∈ L∞ d
Proposition 13.2.1. b
1 1 d
Then for all f ∈ L ∩ L (R ; X), we have
Z
Tm f (x) = k(x − y)f (y) dy,
Rd
b b
where k = m ∈ L1 (Rd ; L (X, Y )).
Such functions exist by Lemma 5.5.21, whose proof also gives the identity
ψ(ξ) b − ϕ(2ξ)
b = ϕ(ξ) b and hence
X
b −j ξ) = ϕ(2
ψ(2 b −N ξ) − ϕ(2
b −L ξ)
L<j6N
b −j ξ)m(ξ),
mj (ξ) := ψ(2 b −N ξ)m(ξ),
mN (ξ) := ϕ(2
(13.23)
X
mN N L
L (ξ) := m (ξ) − m (ξ) = mj (ξ),
L<j6N
i.e., these are supported away from both the origin and infinity. While the
support away from zero is not required by Proposition 13.2.1, it is a conve-
nience for forthcoming considerations due to the special role of the origin in
various multiplier conditions. The next two lemmas describe a precise sense in
which, for many purposes, it is “enough” to study the truncated multipliers
mN .
Lemma 13.2.2. Let X, Y be Banach spaces and m ∈ L∞ (Rd ; L (X, Y )). For
p ∈ (1, ∞), we have m ∈ MLp (Rd ; X, Y ), if and only if mN ∈ MLp (Rd ; X, Y )
uniformly in N , if and only if mN p d
L ∈ ML (R ; X, Y ) uniformly in M and N .
b −N ·) f ) = Tm (ϕ2−N ∗ f ),
TmN f = Tm (Tϕ(2
so that kmN kMLp (Rd ;X,Y ) 6 kϕk1 kmkMLp (Rd ;X,Y ) , and thus
kmN
L kMLp (Rd ;X,Y ) 6 2kϕk1 kmkMLp (Rd ;X,Y ) .
∂/ := ∂/2πi
so that
/j f (ξ) = ξj fb(ξ).
∂d
The deduction of the kernel estimates is easiest when sufficiently many
derivatives are allowed in (13.24); as it turns out, this is somewhat more
than the collection α ∈ {0, 1}d appearing in Mihlin’s Theorem 5.5.10. We
formulate several results for a generic Banach space Z instead of L (X, Y ), as
the operator structure plays no role here; this also makes the formulae slightly
shorter. We say that a collection A of multi-indices is convex, if α ∈ A implies
β ∈ A whenever 0 6 β 6 α.
where the binomial formula was used in the second to last step. The result
follows after dividing both sides by (2π)|α| > 6|α| .
for some A > 0 and all multi-indices α in some convex set. Then
α
kx 7→ ∂/x [(e2πiy·x − 1)f (x)]k∞ 6 (6 + 2|α| )A|y| · A−|α|
for all y ∈ Rd with |y| 6 A−1 , and for the same set of multi-indices.
and hence
α
X
k∂/x [(e2πiy·x − 1)f (x)]k 6 2π|y|A · A−|α| + |y||γ| A−|α|+|γ|
06=γ6α
X
−|α|
6 |y|A · A 2π + (A|y|)|γ|−1 .
06=γ6α
for some A > 0. Then for almost all x, y ∈ Rd with |y| 6 21 |x|, we have
kfbk1 6 cd .
where ωd is the volume of the unit ball in Rd . With α = nei , this shows that
|xi |n |fb(x)| 6 ωd Ad−n for i = 1, . . . , d, which readily gives (13.25).
We observe that fb(x − y) − fb(x) is the Fourier transform of (e2πix·y −
1)f (x), which satisfies the same assumptions as f for |y| 6 A−1 , except for a
multiplicative factor (6 + 2d )A|y|, by Lemma 13.2.4. An application of (13.25)
to this function in place of f hence gives
when A|y| 6 1. On the other hand, if A|y| > 1, then we simply estimate
fb(x−y)− fb(x) by (13.25) and the triangle inequality, recalling the assumptions
that |y| 6 21 |x| and n 6 d + 1:
The last two bounds are both seen to be dominated by the claimed bound in
(13.26).
That fb ∈ L1 (Rd ; Z) is immediate from (13.25) by integrating the estimate
n o
|fb(x)| 6 cd Ad min 1, (A|x|)−d−1 .
c 1 |y|
kk N (x)k 6 , kk N (x − y) − k N (x)k 6 ω ,
|x|d |x|d |x|
X X cd 2(j+1)(d−h)
|k N (x)| 6 |kj (x)| 6 min M
06h6d+1 |x|h
j6N j∈Z
X X cd 2−(j+1)
6 cd 2(j+1)d M + M 6 c0d |x|−d M.
|x|d+1
j:2j+1 61/|x| j:2j+1 >1/|x|
X cd 2(j+1)(d−h)
|k N (x − y) − k N (x)| 6 min min{2j+1 |y|, 1}M
06h6d+1 |x|h
j∈Z
X
(j+1)(d+1)
X cd
6 cd 2 |y|M + |y|M
|x|d+1
j:2j+1 61/|x| j+1
j:1/|x|62 61/|y|
−(j+1)
X cd 2
+ M
|x|d+1
j:2j+1 >1/|y|
Then there is a kernel k ∈ C(Rd \{0}; L (X, Y )) that satisfies the same bounds
as k N in Proposition 13.2.6 and such that
Z
Tm f (x) = k(x − y)f (y) dy
Rd
TmN
L
f = Tm [(ϕ2−N ∗ f ) − (ϕ2−L ∗ f )],
13.2 Fourier multipliers as singular integrals 265
where TmN L
Lp (Rd ; Y ) by Lemma 13.2.2. On
is bounded from Lp (Rd ; X) to b
N
the other hand, kL is a finite sum of kj = mj , where the multipliers mj are
in the scope of Lemma 13.2.5, and hence kL N
∈ L1 (Rd ; L (X, Y )). But then
also f 7→ kN ∗ f is bounded from L (R ; X) to Lp (Rd ; Y ), and the previous
p d
N cd M
kkL (x − y)f (y)k 6 kf (y)k,
|x − y|d
N
by dominated convergence. The pointwise estimates of kL are clearly inherited
by k by the pointwise convergence. This completes the proof for p ∈ (1, ∞).
266 13 The Fourier transform and multipliers
Case p = 1: We can still make use of large parts of the preceding consid-
erations, but some details require a modification. The standard mollifier re-
sult (Proposition 1.2.32) still applies to show that ϕ2−N ∗ f → f , and hence
TmN f → Tm f , in L1 (Rd ; X) as N → ∞, but it no longer guaranteed that
ϕ2−L ∗ f should converge to 0 as L → −∞. Hence, we will separately deal
with TmL f . b
For f ∈ L1 (Rd ; X) ∩ L1 (Rd ; X), we have
Z
kTmL f (x)k = ϕ(2−L ξ)m(ξ)fb(ξ) dξ
Rd
Z
6 kmk∞ kfb(ξ)k dξ 6 ωd (2L+1 )d kmk∞ kf k1 .
|ξ|62L+1
TmN f = Tm N
L
f + Tm L f = k mN
L
∗ f + TmL f.
Since all of the operators acting on f above are bounded from L1 (Rd ; X) to
L1 (Rd ; Y ) + L∞ (Rd ; Y ), the identity continues to hold for all f ∈ L1 (Rd ; X).
Taking the limits N → ∞ and L → −∞, we have TmN f → Tm f in L1 (Rd ; Y )
and TmL f → 0 in L∞ (Rd ; Y ). Along suitable subsequences, we have both
limits almost everywhere, and hence we arrive at the same pointwise limit
Z
N
Tm f (x) = lim kL (x − y)f (y) dy
N →∞ Rd
L→−∞
as in the case p ∈ (1, ∞). The rest of the proof can then be concluded in
the same way as before. Specifically, let us note that the final application
of dominated convergence is justified simple because the product of [y 7→
|x − y|−d ] ∈ L∞ ({B(x, ε)) and f ∈ L1 (Rd ; X) is integrable.
Corollary 13.2.8. Let X, Y be Banach spaces and p0 ∈ [1, ∞). Suppose that
m ∈ MLp0 (Rd ; X, Y ) satisfies
Lemma 13.2.11. Let X be a Banach space with Fourier type p ∈ (1, 2]. Let
f ∈ L∞ d
c ((−A, A) ; X) satisfy
α
k∂/ f k∞ 6 A−|α| ∀α ∈ {0, 1}d
and hence
2 48 1 2 log 2
Φp,X 6 4 1 + 6r0 T 4+ T = ( + 1)( + 1)r0 T 2
log 2 log 2 6r0 T T
0
6 70 · r0 T 2 = 70r0 (68τr,X;2 )r .
Dk = {x ∈ Rd : xi ∈ [2ki , 2ki +1 ) ∀i = 1, . . . , d}
so that obviously X
kfbk1 = k1Dk f k1 .
k∈Zd
Similarly,
For x ∈ Dk , we have |xi | > 2ki , and hence |xβ+γ | > 2k·(β+γ) . We can now
make the following estimate. At a critical point, passing from a norm of the
Fourier transform fb to a norm of f itself, we apply the Fourier type assumption
0 |γ|
to F : Lp (Rγ ; X) → Lp (Rγ ; X), producing the constant ϕp,X (Rγ ) 6 ϕp,X ,
and the trivial boundedness of the Fourier transform F : L1 (Rα+β ; Z) →
L∞ (Rα+β ; Z), with Z = Lq (Rγ ; X) for either q = p or q = p0 , depending on
the (irrelevant) order in which we perform these two steps:
p0 0
0 ϕp,X (ϕp,X )−1/p d
64 d
2 + (1 + log+
2 ϕpp,X ) +
1 − 2−1/p0
0
d
6 4 (3 + p log+
2 ϕp,X + 2p ) 6 (4p0 )d (4 + log+
0 d
2 ϕp,X )
d
0
where we observed that 1 − 2−1/p > 1/(2p0 ), since the function g(u) = u/2 +
2−u satisfies g(u) 6 1 for u = 1/p0 ∈ [0, 21 ], being convex with g(0) = 1 and
g( 12 ) = 1/4 + 2−1/2 < 1.
270 13 The Fourier transform and multipliers
For the second bound, we also use the triangle inequality, but keeping the
restriction to {B. Then
k1{B(0,3|y|) fb(· − y)k1 = k1{B(−y,3|y|) fbk1
6 k1{B(0,2|y|) fbk1 6 k1{(−2r/√d,2r/√d)d ) fbk1 ,
and the same√bound is obvious for fb in place of fb(· − y). Applying (13.28)
with R = 2r/ d produces the required bound.
Proposition 13.2.13. Let X, Y be Banach spaces, and suppose that m ∈
L∞ (Rd ; L (X, Y )) satisfies
Proof of (1). From Lemma 13.2.3 it follows that each Littlewood–Paley trun-
cation mj ∈ L∞c (B(0, 2
j+1
); L (X, Y )) satisfies
α
k∂/ mj k∞ 6 2d M 2−(j+1)|α| ,
which is like the condition of Lemma 13.2.11 with A = 2j+1 and an additional
multiplicative constant 2d M .
Moreover, for x ∈ X, the function mj (·)x ∈ L∞ c (B(0, 2
j+1
); Y ) satisfies
d
the same assumption with constant 2 M kxk, and now the range Y also has
Fourier type p ∈ (1, 2], as required to apply Lemma 13.2.11. In particular,
from (13.30), we conclude that
Z
k(kj (t − s) − kj (t))xkY dt
|t|>3|s|
n 8d2 ϕp,Y o
6 Φdp,Y 2d M kxk min 2, j+1 1/p0 , 8 · 2d 2j+1 r .
(2 r)
272 13 The Fourier transform and multipliers
b
N
Since mN ∈ Lc∞ (Rd ; L (X, Y )) ⊆ L1 (Rd ; L (X, Y )), the kernels k N = m ∈
C0 (Rd ; L (X, Y )) are well defined, and we can estimates
Z
k(k N (t − s) − k N (t))xkY dt
|t|>3|s|
XZ
6 k(kj (t − s) − kj (t))xkY dt
j6N {B
X
d
6 Φp,Y 2d M kxk 8 · 2d 2j+1 r
j:8·2d 2j+1 r62−d−5
X
+ 1
j:2−d−3 62j+1 r6(8d2 ϕ p,Y )p 0
X 8d2 ϕp,Y
+
(2j+1 r)1/p0
2j+1 r>(8d2 ϕp,Y )p0
p0 1
6 Φdp,Y 2d M kxk 2 + (log+ 2
2 (8d ϕp,Y ) + d + 4) + −1/p 0
1−2
d d + 0
6 Φp,Y 2 M kxk 6 + 3d + log2 ϕp,Y p
6 Φdp,Y 2d M kxk · d · Φp,Y 6 (2Φp,Y )d+1 M kxk.
Proof of (2). We note that (13.31) implies a similar bound for the pointwise
adjoint function m∗ = m(·)∗ ∈ L∞ (Rd ; L (Y ∗ , X ∗ )), while the assumption
that X has Fourier type p ∈ (1, 2] implies that X ∗ has the same Fourier
type with ϕp,X ∗ = ϕp,X (Proposition 2.4.16). Thus case (2) follows from the
already proven case (1) applied to (m∗ , Y ∗ , X ∗ ) in place of (m, X, Y ).
Corollary 13.2.14. Let X, Y be Banach spaces with non-trivial Fourier type,
let p0 ∈ [1, ∞), and suppose that m ∈ MLp0 (Rd ; X, Y ) satisfies
kTm f kL2 (Rd ;H2 ) = kmfbkL2 (Rd ;H2 ) 6 M kfbkL2 (Rd ;H1 ) = M kf kL2 (Rd ;H1 ) ,
and thus kmkML2 (Rd ;H1 ,H2 ) 6 M . Since both Hi have Fourier type 2, Corollary
13.2.14 applies to give that m ∈ MLp (Rd ; H1 , H2 ) for all p ∈ (1, ∞).
Note that if m is stably constant in the direction of x, then for every s > 0,
where the last step follows from the assumption (included in the definition of
stably constant) that m is in particular constant in the direction of x.
Example 13.3.3. Suppose that m ∈ C 1 (Rd \{0}) satisfies the first order Mihlin
condition |∇m(x)| 6 M |x|−1 for all x ∈ Rd \ {0}. If m is constant in the
direction of some x, then m is stably constant in this direction. Indeed
Z 1
|m(y + tx) − m(x)| = |m(y + tx) − m(tx)| = y · ∇m(ys + tx) ds
0
1
M |y|
Z
M ds
6 |y| 6 ,
0 |ys + tx| t|x| − |y|
If Tk is the extension of T to Lp0 (Td ; Lp (T(k−1)d ; X)) (Lp (T0 ; X) := X), then
13.3 Necessity of UMD for multiplier theorems 275
r
X r
X
Tk f k 6 kT kL (Lp0 (Td ;X)) fk
Lp (Trd , dt1 ... dtr ;X) Lp (Trd , dt1 ... dtr ;X)
k=1 k=1
for all fk = fk (t1 , . . . , tk ) ∈ Lp0 (Td , dtk ; Lp (T(k−1)d , dt1 . . . dtk−1 ; X)) that
have non-zero Fourier coefficients with respect to tk only in the directions
where m is stably constant.
where
t̄k−1 = (t1 , . . . , tk−1 ) ∈ (Td )k−1 , tk ∈ Td ,
ej (t̄k−1 ) := exp(2πij · t̄k−1 ), e` (tk ) := exp(2πi` · tk ),
and we may choose the same B for all the fk , since there are only finitely many
of them. Then Tk fk has a similar expansion with the (j, `) term multiplied by
m(`).
Let us fix some t̄k := (t̄k−1 , tk ) = (t1 , . . . , tk ) ∈ Tkd for the moment, and
to be chosen below.
We will shortly define an auxiliary function of the new variable t ∈ Td . For
this we need to introduce a couple of product-like operations between vectors
of different lengths. We set
n
k fk : T → X is defined analogously.
The function T]
276 13 The Fourier transform and multipliers
Hence, assuming that N̄k−1 was already chosen, and recalling that j ∈ Z(k−1)d
and ` ∈ Zd with |j|, |`| 6 B take only finitely many different values, we can
choose Nk large enough so that
|m(N̄k−1 j + Nk `) − m(`)| 6 ε
Here the Lp norms are taken with respect to the variable t ∈ Td , and we
recall that the variables t1 , . . . , tr ∈ Td were kept fixed until now. We now
take the Lp norms of (13.33) with respect to t̄r = (t1 , . . . , tr ) ∈ Trd and use
the triangle inequality to get
Z Z r
X p 1/p
(Tk fk )(t̄r + N̄r ⊗ t) dt dt̄r
Trn Tn X
k=1
Z Z r
X p 1/p
6 kT kL (L0p (Tn ;X)) fk (t̄r + N̄r ⊗ t) dt dt̄r
Trn Tn X
k=1
r
X
+ε kfk kA .
k=1
Since there is no more explicit N̄r dependence, we may take ε → 0, and this
gives the assertion.
R
2kmkMLp (Rd ;X) .
βp,X 6 (13.34)
|m(u1 ) − m(u2 )|
Lemma 13.3.6. If m ∈ C(Rd \ {0}) ∩ MLp (Rd ; X), then (m(k))k∈Zd \{0} ∈
MLp0 (Td ; X) and
Proof. This is a slight variant of Proposition 5.7.1, which says that if every
k ∈ Zd is a Lebesgue point of m ∈ L∞ (Rd ), then (m(k))k∈Zd is a Fourier
multiplier on Lp (Td ; X) of at most the norm of the Fourier multiplier m
on Lp (Rd ; X). A slight obstacle is that 0 may fail to be a Lebesgue point
of our m(ξ), no matter how we define m(0). But, if we only consider the
action of these operators on Lp0 (Td ; X), the 0th frequency never shows up,
and one can check that the proof of Proposition 5.7.1 also applies, with trivial
modifications, to the case that each k ∈ Zd \ {0} is a Lebesgue point, giving
exactly what we claimed.
Proof of Theorem 13.3.5. We begin by essentially the same reduction as in
the proofs of both Theorems 5.2.10 and 10.5.1 (the necessity of UMD for the
boundedness of the Hilbert transform and the imaginary powers of the Lapla-
cian, respectively); but we repeat this short step for the reader’s convenience:
By Theorem 4.2.5 it suffices to estimate the dyadic UMD constant. In order to
most conveniently connect this with Fourier analysis, we choose a model of the
Rademacher system (rk )nk=1 , where the probability space is Tdn = Td1 ×· · ·×Tdn
(each Tdk is simply an indexed copy of Td ), and rk = rk (tk ) is a function of the
kth coordinate tk ∈ Tdk only. Moreover, we are free to choose any instance of
such function, as long as it takes both values ±1 on subsets of Td of measure
1
2 . Then it is sufficient to prove that
n
X n
X
k fk 6K fk ,
Lp (Tdn ;X) Lp (Tdn ;X)
k=1 k=1
278 13 The Fourier transform and multipliers
where K is the constant on the right of (13.34), for all signs k = ±1, for
all fk of the form fk = φk (r1 , . . . , rk−1 )rk ; these are precisely the martingale
differences of Paley–Walsh martingales (see Proposition 3.1.10). We use the
convention that Lp (T0 ; X) := X.
Let us then observe that, with suitable choice of the invertible matrices
Aj , j = 1, 2, the multipliers mj (ξ) = m(Aj ξ) (of the same multiplier norm
as the original m) are stably constant in the directions of ±ek , k = 1, 2, and
moreover mj (±ek ) = m(u1 ) if j = k and mj (±ek ) = m(u2 ) if j 6= k. Defining
yet another multiplier m0 = 21 (m1 − m2 ) (of at most the same multiplier
norm as m), we find that m0 is also stably constant in the directions of ±ek ,
k = 1, 2, and moreover m0 (±e1 ) = 21 (m(u1 ) − m(u2 )) =: a and m0 (e2 ) = −a.
If we can prove the claim with m0 , e1 , e2 in place of the original m, u1 , u2 , then
the original claim also follows from
The required condition on fk above is that its Fourier coefficients with respect
to the variable tk should be non-zero only in the directions, where m is stably
constant, i.e., only in the directions ±e1 and ±e2 . Given the product form of
fk , this means more simply that rk should have non-zero Fourier coefficients
only in these directions, which holds in particular if rk is a function of only
the first or only the second coordinate. Note that this gives still (more than)
enough flexibility to make rk equidistributed with a Rademacher variable.
Now, given a sequence (εk )rk=1 , we choose rk to be a function of the first
coordinate if εk = +1, and of the second coordinate if εk = −1. It then follows
that in either case Te(m(j))j∈Zd \{0} rk = aεk rk , and we conclude that
13.3 Necessity of UMD for multiplier theorems 279
n n
X 1 X
k fk = Tk f k
Lp (Tdn ;X) |a| Lp (Tdn ;X)
k=1 k=1
n
2 X
6 kmkMLp (Rd ;X) fk ,
|m(e1 ) − m(e2 )| Lp (Tdn ;X)
k=1
Proof. By the same reductions and notation as in the proof of Theorem 13.3.5,
we now need to check that
n
X n
X
σk f k 6 kmkMLp (Rd ;X) fk ,
Lp (Tdn ;X) Lp (Tdn ;X)
k=1 k=1
has a Fourier series involving only frequencies that are multiples of the vector
nk . By the homogeneity of m again, this means that
and thus
n
X n
X
m(nk )fk = Te(m(j))j∈Zd \{0} fk
Lp (Tdn ;X) Lp (Tdn ;X)
k=1 k=1
n
X
6 kmkMLp (Rd ;X) fk .
Lp (Tdn ;X)
k=1
or in other words
n
X
σk rk φk (r1 , . . . , rk−1 )
Lp (Tdn ;X)
k=1
n
X
6 kmkMLp (Rd ;X) rk φk (r1 , . . . , rk−1 )
Lp (Tdn ;X)
k=1
n
X
+δ krk φk (r1 , . . . , rk−1 )kLp (Tdn ;X) .
k=1
Corollary 13.3.8. Let X be a Banach space, d > 2 and p ∈ (1, ∞). If any of
the following operators is bounded on Lp (Rd ; X), then X is a UMD space:
(1) a second-order Riesz transform Rj Rk , 1 6 j, k 6 d,
(2) their non-zero difference Rj2 − Rk2 , 1 6 j 6= k 6 d,
(3) the Beurling transform B = (R22 − R12 ) + i2R1 R2 (d = 2).
13.3 Necessity of UMD for multiplier theorems 281
Corollary 13.3.9. Let X be a Banach space, let d, k > 1 and p ∈ (1, ∞).
Then there is a constant C such that
Proof. The sufficiency of (1) has been established in Proposition 5.6.10 and
the sufficiency of (2) in Theorem 5.6.11. Moreover, in Theorem 5.6.12, it has
been shown that the UMD property is necessary when k is odd, and that the
boundedness of the second-order Riesz transform R12 is necessary when k is
even and d > 2. By Corollary 13.3.8, the UMD property follows from this,
and hence it is necessary in all cases except (1).
In our final corollary to Theorem 13.3.5, we dispense with the evenness con-
dition.
282 13 The Fourier transform and multipliers
Corollary 13.3.10. Let d > 1 and m ∈ C(Rd \ {0}) be any positively homo-
geneous multiplier (i.e., m(λξ) = m(ξ) for all ξ ∈ Rd \ {0} and λ > 0) that is
not identically constant. If m ∈ MLp (Rd ; X), then X is a UMD space and
R
4kmkMLp (Rd ;X)
βp,X 6 min ,
∈S d−1
u1 ,u2 |m(u1 ) + m(−u1 ) − m(u2 ) − m(−u2 )|
2kmkMLp (Rd ;X) 2
R
βp,X 6 (~p,X )2 6 min ,
u∈S d−1 |m(u) − m(−u)|
The assumption that m is not identically constant, rather than the perhaps
expected “not identically zero”, is necessary: the Fourier multiplier Tm with
m ≡ c coincides with the scalar multiplication f 7→ c · f , whose boundedness
certainly needs no UMD.
Proof. As pointed out right before Proposition 5.3.7, the assumption that
m ∈ MLp (Rd ; X) implies the same property for the reflected function m(ξ)e :=
m(−ξ). Then, by the triangle inequality, the even and odd parts meven :=
1
e and modd := 21 (m − m)
2 (m + m) e are also positively homogeneous multipliers
of at most the same multiplier norm as m. Since m is not identically constant,
and m = meven + modd , at least one of meven or modd is not identically
constant.
If meven is not identically constant, there are two directions u1 , u2 ∈ S d−1
such that meven (u1 ) 6= meven (u2 ) and hence, by evenness,
R
2kmeven kMLp (Rd ;X)
βp,X 6
meven (u1 ) − meven (u2 )
4kmkMLp (Rd ;X)
6 .
m(u1 ) + m(−u1 ) − m(u2 ) − m(−u2 )
(Note that the condition that meven (u1 ) 6= meven (u2 ) is precisely the require-
ment that the denominator is non-zero, and hence can extend the previous
display to all pairs of u1 , u2 ∈ S d−1 ; interpreting 1/0 = ∞, as usual, this only
amounts to adding the triviality βp,X R
6 ∞.)
For the odd part modd , being not identically constant is equivalent to being
not identically zero. If this is the case, there is some direction u ∈ S d−1 such
that m(−u) = −m(u) 6= 0. Writing ξ ∈ Rd as ξ = (ξ · u)u + [ξ − (ξ · u)u], we
consider the invertible linear transformations Aλ ξ = (ξ · u)u + λ[ξ − (ξ · u)u],
13.4 Notes 283
where λ > 0. By Proposition 5.3.8, each modd ◦ Aλ has the same multiplier
norm as modd . As λ → 0, it is clear that Aλ ξ → (ξ · u)u for all ξ ∈ Rd and
thus, by the continuity of m and hence modd ,
(The bound remains valid for all u ∈ S d−1 , reducing to a triviality if m(u) =
m(−u).) By Fubini’s theorem, we find that
13.4 Notes
Section 13.1
The precise quantitative form of the final bound in the comparison of various
Fourier-type constants in Proposition 13.1.1 seems to be new; we were not
aware of this estimate at the time of completing Volume II, where a weaker
0
version was given. The identity ϕp,C (Rd ) = (p1/p (p0 )−1/p )d mentioned below
the said proposition is due to Babenko [1961] in the special case that p0 is an
even integer, and due to Beckner [1975] in full generality.
The main result of this section, Theorem 13.1.33 is from Bourgain [1988a],
with preliminary versions going back to Bourgain [1981, 1982]. The main
theorem of Bourgain [1982] reads as follows: If X is a B-convex Banach space
(which is equivalent to non-trivial type by Proposition 7.6.8), then there are
u, v ∈ (1, ∞) and δ, M ∈ (0, ∞) such that
X 1/u X X 1/v
δ kxγ kuX 6 γxγ 6M kxγ kvX , (13.35)
L2 (G;X)
γ∈Γ γ∈Γ γ∈Γ
when G is either T or the Cantor group {−1, 1}N . For G = T, the leftmost
and rightmost estimates correspond, in our notation, to ϕu1 ,X (T) 6 1/δ1 and
ϕv1 ,X (Z) 6 M1 , respectively. The easy estimate ϕp,X (R) 6 ϕp,X (T) was also
observed by Bourgain [1988b]. In contrast to the case of T, a scaling argument
(substituting f (λ·) in place of f and considering the limit λ → 0 or λ → ∞)
shows that an estimate of the from kfbkLq (R;X) 6 Ckf kLp (R;X) can only hold
for q 0 = p; thus, in order to deduce any Hausdorff–Young inequality on R at
all, the additional steps from the mismatched exponents of Bourgain [1982]
to the dual exponents of Bourgain [1988b] seem to be necessary.
The second half of the argument leading to Bourgain’s Theorem 13.1.33, as
presented in Sections 13.1.c and 13.1.d, is close to the treatment of Bourgain
13.4 Notes 285
[1988b], although we have also benefited from the exposition of these steps by
Pietsch and Wenzel [1998]. On the other hand, the first half of our treatment,
in Sections 13.1.a and 13.1.b, is also based on Pietsch and Wenzel [1998] but
deviates from the original approach of Bourgain [1982]. The beginning of the
argument, leading to Proposition 13.1.11 on “breaking the trivial bound” is
due to Hinrichs [1996], but it also uses a result of Bourgain [1985], Proposition
13.1.7, on the Sidon property of quasi-independent sets.
We have chosen this approach of Hinrichs [1996] and Pietsch and Wenzel
[1998] due to an independent interest, in our opinion, of some of its intermedi-
ate steps, despite the fact that the original argument of Bourgain [1982, 1988b]
seems slightly more efficient in terms of the final quantitative conclusions. In
any case, the main result says that every Banach space of type p ∈ (1, 2] will
0
have Fourier-type r = 1 + (cτp,X;2 )−p , for some absolute constant c. (The ad-
ditional factor 6p0 in our formulation of Theorem 13.1.33 could obviously be
absorbed by choosing a larger constant c.) The difference is in the numerical
value of c, which is 68 in our formulation (up to the lower order factor just
mentioned) and 17 in Bourgain [1982, 1988b].
In our approach, this√ constant comes from the proof of Corollary 13.1.20,
where the estimate 48 2 (≈ 67.88) 6 68 is made. (Since we are clearly off
Bourgain’s constant at this point already,
√ it would seem pointless to insist in
the decimals here.) The constant 48 2, in turn, is produced as
√ √
48 2 = 16 · 2 · 3, where
for all 1 6 i 6= j 6 n}
is obtained. While the left-hand sides are not identical, (13.38) allows Bourgain
[1982] to deduce the Hausdorff–Young inequality with mismatched exponents
as in (13.35) (with X ∗ in place of X) for any v ∈ (1, t), and finally, in Bourgain
[1988b], also the classical Hausdorff–Young inequality (13.37) (again with X ∗
in place of X) with any u1 ∈ (1, v). Since v ∈ (1, t) is arbitrary, one can reach
any u1 ∈ (1, t), and thus in particular the r determined by
0
r0 = (18 · τp,X;2 )p (13.39)
Thus, were the claim in the beginning of the remark true, all these spaces
c
would have the Fourier-type r = c−1 > 1, which is impossible for p ∈ (1, r)
by Example 2.1.15.
13.4 Notes 287
Section 13.2
versions of such results are again well known; for example, a version of Propo-
sition 13.2.7 with d + 2 derivatives (instead of d + 1 in the said proposition)
appears in the book of Stein [1993]. Under this stronger assumption, Stein
[1993] deduces that k ∈ C 1 (Rd \ {0}), while Proposition 13.2.6 gives the
slightly weaker conclusion that k is just barely below Lipschitz, with a modu-
lus of continuity ω(t) = O(t · log(1 + 1/t)). This is still quite enough to derive
like Corollaries 13.2.8, 13.2.9, and 13.2.10 on the boundedness of Fourier mul-
tipliers on weighted Lp (w; X) spaces. Using the result from Stein [1993] in
place of Proposition 13.2.7, a version of Corollary 13.2.10 assuming d + 2
derivatives was formulated by Meyries and Veraar [2015]. In principle, vari-
ants of Propositions 13.2.6 and 13.2.7 sufficient for Corollaries 13.2.8 through
13.2.10 would only require smoothness of order d + ε, but such statements and
proofs are bound to have additional technicalities due to the very formulation
of fractional order smoothness conditions. Various results in this direction, in-
volving kernel bounds for Fourier multipliers with close-to-critical fractional
smoothness, were explored by Hytönen [2004].
To get rid of the ε > 0 altogether, i.e., to deduce useful b (in view of
Calderón–Zygmund extrapolation) kernel estimates for k = m from just d
derivatives of m, one needs to impose assumptions on the Fourier-type of the
underlying spaces. While we have only dealt with the sufficiency of the Fourier-
type assumption in Section 13.2.b, an early result involving both directions,
in dimension d = 1, is the following:
Theorem 13.4.2 (König [1991]). A Banach space X is K-convex if and
only if every f ∈ C 1 (T, X) has Fourier coefficients (fb(n))n∈Z ∈ `1 (Z; X).
Recall that K-convexity is equivalent to non-trivial type by Pisier’s Theo-
rem 7.4.23, and non-trivial type is equivalent to non-trivial Fourier-type by
Bourgain’s Theorem 13.1.33. The proof of “⇒” in Theorem 13.4.2 is then
straightforward from non-trivial Fourier type. For the converse, König [1991]
starts with a concrete counterexample when X = L1 (T), and approximates
this finite versions that can be represented in `1N , with blow-up in the limit
N → ∞. By the Maurey–Pisier Theorem 7.3.8, if X does not have non-trivial
type, then it contains subspaces isomorphic to `1N uniformly, and hence the
said finite examples can also be represented in X. Finally, the closed graph
theorem guarantees that a sequence of examples with blow-up also guarantees
the existence of a single f ∈ C 1 (T, X) with (fb(n))n∈Z ∈
/ `1 (Z; X).
In our formulation of Proposition 13.2.13, the assumed Fourier-type p ∈
(1, 2] only affects the constant in the estimate. However, by more careful
reasoning, one could show that also the number of the required derivative
∂ α m could be reduced as a function of p; roughly speaking, one needs only
derivatives up to order bd/pc + 1, or more generally fractional smoothness of
order d/p + ε, to obtain the same conclusions. Such results can be found in
Hytönen [2004]. In the more general context of various function spaces, this
phenomenon will be explored further in Chapter 14; see Proposition 14.5.3
and take q = ∞ there.
13.4 Notes 289
Our focus in the section under discussion has been exploring conditions
that one bneeds to assume on a multiplier m in order that their associated ker-
nel k = m satisfies the assumptions of one of the extrapolation theorems of
Chapter 11 (so that the a priori boundedness of Tm on one Lp0 (Rd ; X) extends
to other spaces), but similar considerations can also be used to reduce the re-
quired smoothness, as a function of the Fourier-type of the underlying spaces,
in results like Mihlin’s Multiplier Theorem 5.5.10 (where the boundedness of
Tm on Lp (Rd ; X) is deduced “from scratch”). Such results were pioneered
by Girardi and Weis [2003b] and further elaborated by Hytönen [2004]. If m
is scalar-valued, it is also possible to replace Fourier-type by quantitatively
weaker assumptions on type or cotype; see Hytönen [2010].
Section 13.3
The main results of this section, notably Proposition 13.3.4, Theorem 13.3.5,
and Corollary 13.3.8, are essentially from Geiss, Montgomery-Smith, and
Saksman [2010], but we have incorporated some improvements, partially in-
spired by unpublished observations of Alex Amenta that he kindly shared
with us.
These results may be seen as successors, in terms of both statement and
proof, of Theorem 5.2.10 of Bourgain [1983] and Theorem 10.5.1 of Guerre-
Delabrière [1991], which deal with the necessity of UMD for the boundedness
of the Hilbert transform and the imaginary powers (−∆)is of the Laplacian,
respectively. However, none of these three results contains any of the other
two.
Certain elaborations of Corollary 13.3.8 are due to Castro and Hytönen
[2016]. Namely, the identity ∂j ∂k u = −Rj Rk ∆u implies that
d
X
k∂j ∂k ukLp (Rd ;X) 6 C kRi2 ukLp (Rd ;X) , (13.40)
i=1
where C 6 kRj Rk kL (Lp (Rd ;X)) , but C could a priori be much smaller. How-
ever, Castro and Hytönen [2016] show that the seemingly weaker inequality
(13.40) still implies the UMD property with the same control
βp,X 6 2C(13.40) (13.41)
as in Corollary 13.3.8 for kRj Rk kL (Lp (Rd ;X)) . More generally, the same paper
proves the necessity of UMD for any member of a family of inequalities of the
form X
k∂ β ukLp (Rd ;X) 6 C k∂ α ukLp (Rd ;X) ,
α∈A
but the relation between the constants is particularly clean in the example
just mentioned.
It could be of interest to identify more general criteria (subsuming previous
related results) for inequalities of classical/harmonic analysis to
290 13 The Fourier transform and multipliers
k2Rj Rk kL (Lp (Rd ;X)) = kRj2 − Rk2 kL (Lp (Rd ;X)) = βp,X
R
(13.42)
for all 1 6 j 6= k 6 d. The upper bounds for the norms are proved by
representing and estimating the operators by means of stochastic integrals.
Yaroslavstev [2018] obtained further variants of these estimates for related
operators. We plan to detail this in a forthcoming Volume. By (13.41), a
trivial bound, and (13.42), it follows that
Certain substitute results related to the latter are due to Domelevo and
Petermichl [2023c,d]. They construct a new dyadic operator and show that
its boundedness is equivalent to that of the Hilbert transform, with linear de-
pendence between the respective norms in both directions. Analogous results
for the Riesz transforms are obtained in Domelevo and Petermichl [2023a,b].
Further estimates between the Hilbert transform (and variants) and de-
coupling constants related to the UMD constant can be found in Osȩkowski
and Yaroslavtsev [2021].
Corollary 13.3.9 characterises situations in which there is a continuous
embedding H k,p (Rd ; X) ,→ W k,p (Rd ; X). Several related results, including
versions on domains O ⊆ Rd , are due to Arendt, Bernhard, and Kreuter
[2020].
14
Function spaces
kf kBp,q
s (Rd ;X) = (2k ϕk ∗ f )k>0 `q (Lp (Rd ;X))
and
kf kFp,q
s (Rd ;X) := (2ks ϕk ∗ f )k>0 Lp (Rd ;`q (X))
,
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 293
T. Hytönen et al., Analysis in Banach Spaces, Ergebnisse der Mathematik und ihrer
Grenzgebiete. 3. Folge / A Series of Modern Surveys in Mathematics 76,
https://doi.org/10.1007/978-3-031-46598-7_4
294 14 Function spaces
using the notation for Rademacher spaces introduced Section 6.3; the equality
of the latter two norms is obtained by repeating the proof of Theorem 9.4.8
for Rademacher sums. Comparing these norms with the previous two, it is
also of interest to note that equivalent norms are obtained if the εp -norm is
replaced by an εq -norm, by the Kahane–Khintchine inequalities.
In view of their very similar definitions, it comes as no surprise that the
theories of Besov and Triebel–Lizorkin spaces largely parallel each other and
resemble the theory of Bessel potential spaces to some extent. There are some
notable differences however, due to the different orders in which the Lp -norm
and `q -norm are taken; as we have already pointed out, the Triebel–Lizorkin
norm is generally speaking more difficult to handle. The main advantage of
the Besov and Triebel–Lizorkin over the Bessel potential spaces is that they
are often easier to work with, and indeed many basic results for these spaces in
the vector-valued setting do not rely on the geometry of the Banach space X.
This is in stark contrast with the theory of Bessel potential spaces, where the
corresponding results often require geometrical properties such as the UMD
property of X or the Radon–Nikodým property of X ∗ , as we have seen in
Chapter 5.
After establishing notation and proving some preliminary results in Sec-
tion 14.1, the class of Besov spaces is introduced in Section 14.4 via their
Littlewood–Paley decompositions. Several basic aspects of these spaces are
discussed, such as their independence of the inhomogeneous Littlewood–Paley
sequence used in the definition, the density of smooth functions, and Sobolev
type embeddings. We continue with several more advanced results, including
a difference norm characterisation, identification the complex and real inter-
polation spaces, and identification of the dual spaces. In Section 14.5 these
results are used to prove embedding theorems for the spaces γ(L2 (Rd ), X)
introduced in Chapter 9 and to prove R-boundedness of the ranges of smooth
operator-valued functions under type and cotype assumptions. In the same
section we discuss Fourier multiplier results for Besov spaces under (co)type
and Fourier type assumptions.
In Section 14.6 the Triebel–Lizorkin spaces are introduced. Proving the
same basic properties as before is more complicated, especially for the impor-
tant endpoint exponent q = 1, and requires the boundedness of the so-called
Peetre maximal function and the boundedness of Fourier multiplier opera-
14.1 Summary of the main results 295
or subsets thereof. The interpolation results assume that θ ∈ (0, 1) and where
1 1−θ θ 1 1−θ θ
= + , = +
pθ p0 p1 qθ q0 q1
and
respectively.
Identities. Up to equivalent norms we have the following identifications. If
p ∈ [1, ∞), s ∈ (0, 1), then
and, if s ∈ (0, ∞) \ N,
s
Cub (Rd ; X) = B∞,∞
s
(Rd ; X). (Corollary 14.4.26)
S (Rd ; X) ,→ Bp,q
s
(Rd ; X) ,→ S 0 (Rd ; X) (Proposition 14.4.3)
s
Bp,1 (Rd ; X) ,→ Bp,q
s
(Rd ; X) ,→ Bp,∞
s
(Rd ; X) (Proposition 14.4.18)
S (R ; X) ,→
d s
Fp,q (Rd ; X) ,→ S (R ; X)
0 d
(Proposition 14.6.8)
14.1 Summary of the main results 297
s
Fp,1 s
(Rd ; X) ,→ Fp,q s
(Rd ; X) ,→ Fp,∞ (Rd ; X) (Proposition 14.6.13)
k
Fp,1 (Rd ; X) ,→ W k,p (Rd ; X) ,→ k
Fp,∞ (Rd ; X) (Proposition 14.6.13)
s
Fp,1 (Rd ; X) ,→ H s,p (Rd ; X) ,→ s
Fp,∞ (Rd ; X) (Proposition 14.6.13)
s
Bp,p∧q (Rd ; X) ,→ Fp,q
s
(Rd ; X) ,→ Bp,p∨q
s
(Rd ; X). (Proposition 14.6.8)
Sobolev embedding theorem I: If (and only if) either one of the following three
conditions holds: p0 = p1 and s0 > s1 ; p0 = p1 and s0 = s1 and q0 6 q1 ;
p0 < p1 and q0 6 q1 and s0 − pd0 > s1 − pd1 ; then
Sobolev embedding theorem II: Let p0 , p1 ∈ [1, ∞). If (and only if) either one
of the following three conditions holds: p0 = p1 and s0 > s1 ; p0 = p1 and
s0 = s1 and q0 6 q1 ; p0 < p1 and s0 − pd0 > s1 − pd1 (no condition on q0 , q1 );
then
Sobolev embedding theorem III: Let p0 , p1 ∈ (1, ∞). If (and only if) either one
of the following three conditions holds: p0 = p1 and s0 > s1 ; p0 < p1 and
s0 − pd0 > s1 − pd1 ; then
For k ∈ N,
k
B∞,1 (Rd ; X) ,→ Cub
k
(Rd ; X) ,→ B∞,∞
k
(Rd ; X). (Proposition 14.4.18)
d
If p0 ∈ [1, ∞] and s0 , s1 > 0 satisfy s0 − p0 > s1 , then
d d
Jawerth–Franke theorem: If p0 < p1 , and s0 − p0 > s1 − p1 , then
and, if p1 < ∞,
If k > d, then
k k−d
F1,∞ (Rd ; X) ,→ Cub (Rd ; X). (Corollary 14.6.27)
Embeddings under (co)type assumptions: If (and only if) X has type p ∈ [1, 2],
( 1 − 12 )d
p
Bp,p (Rd ; X) ,→ γ(L2 (Rd ), X). (Theorem 14.5.1)
( 1 − 12 )d
γ(L2 (Rd ), X) ,→ Bq,q
q
(Rd ; X). (Theorem 14.5.1)
If X is a UMD Banach space with type p0 ∈ [1, 2] and cotype q0 ∈ [2, ∞], and
if p ∈ (1, ∞), s ∈ R, then
s
Fp,p 0
(Rd ; X) ,→ H s,p (Rd ; X) ,→ Fp,q
s
0
(Rd ; X). (Proposition 14.7.6)
[Bps00 ,q0 (Rd ; X0 ), Bps01 ,q0 (Rd ; X1 )]θ = Bpsθθ ,qθ (Rd ; Xθ ). (Theorem 14.4.30)
s0
[Fp,q (Rd ; X0 ), Fp,q
s1
(Rd ; X1 )]θ = Fp,q
sθ
(Rd ; Xθ ) (Theorem 14.6.23)
14.1 Summary of the main results 299
If s0 6= s1 , then
s0
(Bp,q 0
(Rd ; X), Bp,q
s1
1
(Rd ; X))θ,q = Bp,q
sθ
(Rd ; X) (Theorem 14.4.31)
(H s0 ,p (Rd ; X), H s1 ,p (Rd ; X))θ,q = Bp,q
sθ
(Rd ; X). (Theorem 14.4.31)
s0 s1
(Cub (Rd ; X), Cub (Rd ; X))θ,∞ = B∞,∞
sθ
(Rd ; X). (Corollary 14.4.32)
s0
(Fp,q 0
(Rd ; X), Fp,q
s1
1
(Rd ; X))θ,q = Bp,q
sθ
(Rd ; X). (Proposition 14.6.24)
s
Bp,q (Rd ; X)∗ = Bp−s d ∗
0 ,q 0 (R ; X ) (Theorem 14.4.34)
s
Fp,q (Rd ; X)∗ = Fp−s d ∗
0 ,q 0 (R ; X ). (Theorem 14.6.28)
14.2 Preliminaries
In this section we prepare some, mostly technical, results that will be of use
in our treatments of both Besov and Triebel–Lizorkin spaces.
14.2.a Notation
We start by reviewing some notation that has been introduced in the two
earlier volumes. We use the standard multi-index notation explained in Section
2.5. For the details we refer to the relevant sections (Section 2.4.c for Schwartz
functions, 2.4.d for tempered distributions, 2.5.b and 2.5.d for Sobolev spaces,
and 5.6.a for Bessel potential spaces).
Let X be a Banach space and let d > 1 be an integer. The Schwartz space
S (Rd ; X) is the space of all f ∈ C ∞ (Rd ; X) for which the seminorms
which turns S (Rd ; X) into a complete metric space. Thus S (Rd ; X) has the
structure of a Fréchet space. As a consequence of Lemma 1.2.19 or Lemma
14.2.1, the space Cc∞ (Rd ) ⊗ X is dense in Lp (Rd ; X) for 1 6 p < ∞. We
will prove in Lemma 14.2.1 that Cc∞ (Rd ) ⊗ X is sequentially dense in both
Cc∞ (Rd ; X) and S (Rd ; X).
The space of continuous linear operators
S 0 (Rd ; X) := L (S (Rd ), X)
∂ α f of order |α| 6 k exist and belong to Lp (D; X). Recall that a function
g ∈ L1loc (D) is said to be the weak derivative of order α of f if
Z Z
f (x)∂ α φ(x) dx = (−1)|α| g(x)φ(x) dx for all φ ∈ Cc∞ (D).
D D
the space W k,p (D; X) is a Banach space. For 1 6 p < ∞ and 0 < s < 1,
the Sobolev–Slobodetskii space W s,p (Rd ; X) is the space of all functions f ∈
Lp (Rd ; X) for which the seminorm
Z Z kf (x) − f (y)kp 1/p
[f ]W s,p (D;X) := dx dy
D D |x − y|sp+d
the space W s,p (Rd ; X) is a Banach space. By Theorem 2.5.17, for 1 6 p <∈ ∞
and 0 < s < 1 the real interpolation method gives
with equivalent norms, provided X is a UMD space, 1 < p < ∞, and k > 1 is
an integer. Under the same assumptions, Theorem 5.6.9 gives
kf (x) − f (y)k
[f ]Cbθ (Rd ;X) := sup
x,y∈Rd ,x6=y |x − y|θ
the space Cbθ (Rd ; X) is a Banach space. The Banach space obtained by taking
θ = 1 in these expressions is called the space of Lipschitz continuous functions
and is denoted by Lip(Rd ; X).
For s = k + θ, with k ∈ N and θ ∈ (0, 1), the space Cbs (Rd ; X) is defined as
the space of all f ∈ Cbk (Rd ; X) for which ∂ α f ∈ Cbθ (Rd ; X) for all multi-indices
satisfying |α| 6 k. With the norm
this space is a Banach space. For all s ∈ [0, ∞) we have continuous embeddings
Lemma 14.2.1. The space Cc∞ (Rd ) ⊗ X is sequentially dense in Cc∞ (Rd ; X)
and S (Rd ; X).
compactness, the exists an index j0 ∈ N such that for all j > j0 and all
g ∈ Cc∞ (V ; X) we have φj ∗g ∈ Cc∞ (O; X) and, for all multi-indices α, β ∈ Nd ,
Cc∞ (O)⊗X, and by the above we have gkj → gk in S (Rd ; X). For appropriate
jk > j0 we find that gkjk → g in S (Rd ; X). Since gkjk ∈ Cc∞ (O) ⊗ X, this
proves density in Cc∞ (Rd ; X).
To prove density in S (Rd ; X) let f ∈ S (Rd ; X). By Step 1 there exists
a sequence (fn )n>1 in Cc∞ (Rd ; X) such that fn → f in S (Rd ; X). Using
Step 2, for every n > 1 choose a sequence (fn,k )k>1 in Cc∞ (Rd ) ⊗ X such
that fn,k → fn in Cc∞ (Rd ; X). Then in particular, fn,k → fn in S (Rd ; X).
Since convergence in S (Rd ; X) is governed by countably many seminorms, a
standard diagonal argument allows us to find a subsequence such that fn,kn →
f in S (Rd ; X).
As a corollary to the above lemma we record:
Proposition 14.2.2. For all p ∈ [1, ∞) and s ∈ R the space Cc∞ (Rd ) ⊗ X is
dense in H s,p (Rd ; X).
Lr (Rd ; Y ) and
the supremum being taken over all finite disjoint partitions π of the set A ∈ A ;
the is taken in L (X, Y ). We say that Φ has bounded variation if kΦk(S) < ∞.
For a strongly measurable function f : S → X such that
Z
kf (s)k dkΦk(s) < ∞,
S
14.2 Preliminaries 305
the construction
R of the Bochner integral (see Section 1.2.a) can be repeated
to define S f dΦ as an element of Y satisfying
Z Z
f dΦ 6 kf k dkΦk.
S S
(i) 0 6 ϕ(ξ)
b 6 1, ξ ∈ Rd ,
(ii) ϕ(ξ)
b = 1 if |ξ| 6 1,
(iii) ϕ(ξ)
b = 0 if |ξ| > 23 .
Such functions can be constructed in a similar way as in Lemma 5.5.21.
Remark 14.2.5. If φ ∈ Φ, the function ψ ∈ S (Rd ) given by
ψ(ξ) b − ϕ(2ξ)
b := ϕ(ξ) b
Remark 14.2.6. It is possible to choose the function ϕ is real and even (or
equivalently ϕ
b real and even). In that case it would be possible to use real
Banach spaces in several of the definitions and results of this chapter. For
instance if f ∈ Lp (Rd ; X) or even S 0 (Rd ; X), then ϕ ∗ f can be defined
without using any complex structure.
Definition 14.2.7 (Inhomogeneous Littlewood–Paley sequence). The
inhomogeneous Littlewood–Paley sequence associated with a function ϕ ∈ Φ
is the sequence (ϕk )k>0 in S (Rd ) given by
ϕ
b0 (ξ) := ϕ(ξ),
b k = 0, ξ ∈ Rd ,
(14.3)
ϕ b −k ξ) − ϕ(2
bk (ξ) := ϕ(2 b −k+1 ξ), k > 1, ξ ∈ Rd .
ϕ b1 (2−k+1 ξ),
bk (ξ) = ϕ k > 1, (14.4)
which implies
n
X n−1
X
kϕk k1 = ϕk − ϕk 6 2kϕ0 k1 , k > 1. (14.7)
1
k=0 k=0
14.2 Preliminaries 307
ψ ∗ u = F −1 (ψbfb) (14.13)
b −n ·))g]
lim (·)β ∂ α [(1 − ϕ(2 ∞
= 0.
n→∞
b
L1 (Rd ; X) is a Banach space. It enjoys the scaling invariance property
Note that Cf,d,ε is trivially finite (for all ε > 0) if f ∈ C d+1 (Rd ; X) has
compact support.
P
Proof. In view of (14.5) we have kf kL1 (Rd ;X) 6 j>0 kϕ
bj f kL1 (Rd ;X) , and
b b
The first term is easy to handle. Indeed, since kF kL1 →L∞ 6 1 and 0 6 ϕ
b1 6 1,
Cf,d,ε
kf (2j−1 ξ)k 6 6 2−(j−1)ε Cf,d,ε .
1 + 2(j−1)ε |ξ|ε
Combining this with the previous estimate, this gives the bound T1 6
2−(j−1)ε Cf,d,ε |3B \ B||B|.
For the second term we use the finiteness of Cd := Rd \B |x|−d−1 dx to
R
obtain
T2 6 Cd ξ 7→ |ξ|d+1 F (ϕ
b1 f (2j−1 ·))(ξ) ∞
.
Using that ϕ
b1 is compactly supported we obtain
F (∂ α [ϕ
b1 f (2j−1 ·)]) ∞
b1 f (2j−1 ·)]
6 ∂ α [ϕ 1
b1 f (2j−1 ·)]
.d ∂ α [ϕ ∞
.
b1 ∂ γ [f (2j−1 ·)]k∞ .d
k∂ β ϕ sup k2(j−1)|γ| ∂ γ f (2j−1 ξ)k 6 2−(j−1)ε Cf,d,ε .
16|ξ|63
It follows that T2 .d 2−(j−1)ε Cf,d,ε . This proves (14.15) for j > 1. The case
j = 0 can be shown in a similar way, skipping the dilation step.
For later reference we state the following consequence of Lemma 14.2.11.
Lemma 14.2.12. Let λ > 0 and suppose that f ∈ C d+1+dλe (Rd ; X) has sup-
port in the ball BR = {ξ ∈ Rd : |ξ| 6 R}. Then (1 + | · |)λ fb(·) ∈ L1 (Rd ; X)
and
k(1 + | · |)λ fb(·)kL1 (Rd ;X) 6 CR,d kf kC d+1+dλe (Rd ;X) .
b
For q < ∞, this is just another way of expressing the same spaces with a
different normalisation of the weight, namely Lqw (S; X) = Lq (wq ; X). How-
ever, using the usual modification for q = ∞, the first version reduces to just
L∞ (w; X) = L∞ (S; X) (since dµ and w dµ share the same zero sets), whereas
Lq∞ (S; X) with norm kf kL∞ w (S;X)
= kf wkL∞ (S;X) is a new space with non-
trivial dependence on the weight w.
isometrically, where
1 1−θ θ
= + , w = w01−θ w1θ .
q q0 q1
We first record the simple:
Proof. We assume that kfn − f kLqw0 (S;Y0 )+Lqw1 (S;Y1 ) → 0. Hence, for every
0 1
n, there is a decomposition fn − f = fn0 + fn1 , where kfnj kLqwj (S;Yj ) → 0 for
j
j = 0, 1. By the well known version of the Lemma in just one Lp space, a
subsequence of fn0 converges to 0 almost everywhere in the norm of Y0 . By the
same result, a further subsequence of fn1 also converges to 0 almost everywhere
in the norm if Y1 . Thus, along this last subsequence, fn −f = fn0 +fn1 converges
to 0 almost everywhere in the norm of Y0 + Y1 .
Proof of Theorem 14.3.1. The unweighted version (w0 = w1 = w ≡ 1) of this
result is contained in Theorem 2.2.6. We will reduce the weighted version to
S∞Y := [Y0 , Y1 ]θ . For n ∈ Z+ , we denote
this special case. Let us abbreviate
Sn := {n−1 6 w0 , w1 6 n}. Then n=1 Sn exhausts S, up to a set of measure
zero, by definition of weights.
312 14 Function spaces
q0
Step 1 – Lqw (S; [Y0 , Y1 ]θ ) ⊆ [Lw 0
(S; Y0 ), Lqw11 (S; Y1 )]θ :
q
Let f ∈ Lw (S; Y ), and assume first assume that {f 6= 0} is contained in Sn
for some n ∈ N. Thus
where the equality of space is Theorem 2.2.6, and hence φ = Φ(θ) for some
Φ ∈ H (Lq0 (S; Y0 ), Lq1 (S; Y1 )), where this notation of holomorphic functions
on the unit strip with appropriate boundary behaviour is defined in Section
C.2. The relation φ = Φ(θ) remains valid if we replace Φ(z) by Φ(z)1En , and
hence all the subsequent considerations can be restricted to En . In particular,
multiplication by any power of w0 or w1 is then a bounded operation on any
of the (weighted or not) Lp spaces appearing in this argument. Now
−(1−θ)
f = φw−1 = Φ(θ)w0 w1−θ = F (θ),
−(1−z)
where F (z) := Φ(z)w0 w1−z ∈ H (Lq0 (w0 ; Y0 ), Lq1 (w1 ; Y1 )). Qualita-
tively, the last inclusion is easy from the corresponding relation for Φ and
the restriction of the supports on En , where all multiplications by powers of
wi are bounded. Quantitatively, we have
−(1−j)
kF (j + it)kLqj (wj ;Yj ) = kΦ(j + it)w0 w1−j kLqj (wj ;Yj )
= kΦ(j + it)kLqj (S;Yj ) , j = 0, 1,
kF kH (Lw
q0
(S;Y0 ),Lw1 (S;Y1 )) = kΦkH (Lq0 (S;Y0 ),Lq1 (S;Y1 )) ,
q (14.16)
0 1
and hence
kf k[Lw
q0
(S;Y0 ),Lw1 (S;Y1 )]θ 6 kF kH (Lw0 (S;Y0 ),Lw1 (S;Y1 ))
q q q
0 1 0 1
Taking the infimum over all Φ in this space with φ = Φ(θ), we obtain
kf k[Lqw0 (S;Y0 ),Lqw1 (S;Y1 )]θ 6 kφk[Lq0 (S;Y0 ),Lq1 (S;Y1 )]θ
0 1
Recall that the previous estimate was obtained under the assumption
that f ∈ Lqw (S; Y ) satisfies {f 6= 0} ⊆ Sn . For a general f ∈ Lqw (S; Y ),
this bound holds with either 1Sn f of 1Sn f − 1Sm f in place of f . Since
1Sn f → f in Lqw (S; Y ) by dominated convergence, it follows that 1Sn f
is a Cauchy sequence, and hence convergent, in the interpolation space
q0 q1 q0 q1
[Lw 0
(S; Y0 ), Lw 1
(S; Y1 )]θ and thus in the sum space Lw 0
(S; Y0 ) + Lw 1
(S; Y1 )
by Lemma C.2.5. By Lemma 14.3.2, a subsequence converges almost every-
where to the same limit function. But it is clear that the a.e. limit is f , and
hence
14.3 Interpolation of Lp -spaces with change of weights 313
kf k[Lqw0 (S;Y0 ),Lqw1 (S;Y1 )]θ = lim k1Sn f k[Lqw0 (S;Y0 ),Lqw1 (S;Y1 )]θ
0 1 n→∞ 0 1
We next turn to the case of real interpolation. Recall that for a Banach couple
(X0 , X1 ), the real interpolation space (X0 , X1 )θ,p with p ∈ [1, ∞] and θ ∈
(0, 1), was introduced in Section C.3. Also recall from Theorem C.3.14 that if
p0 , p1 ∈ [1, ∞] satisfy p1 = 1−θ θ
p0 + p1 , then (X0 , X1 )θ,p = (X0 , X1 )θ,p0 ,p1 with
equivalent norms, where the latter denotes the Lions–Peetre interpolation of
X0 and X1 (second mean method). The main result of this section is as follows.
(Lqw00 (S; Y0 ), Lqw11 (S; Y1 ))θ,q0 ,q1 = Lqw (S; (Y0 , Y1 )θ,q0 ,q1 )
isometrically, where
1 1−θ θ
= + , w = w01−θ w1θ .
q q0 q1
In particular,
where tj−θ Φ(t) ∈ Lqj ( dt/t; Lqj (S; Yj )) for j = 0, 1, and (as a consequence)
the improper integral converges in Lq0 (S; Y0 )+Lq1 (S; Y1 ). Multiplying by 1Sn
if necessary, we may assume that each Φ(t) is also supported on Sn .
Choosing the auxiliary weight W := w0−1 w1 , we then have
Z ∞ Z ∞ Z ∞
dt dt dt
f = φw−1 = Φ(t)w−1 = Φ(tW )w−1 =: F (t) .
0 t 0 t 0 t
On Sn , both wj are bounded from above and below. Due to the technical
assumption on the discreteness of their ranges, both these weights, and hence
W , only take finitely many possible value on Sn . Hence
K
X
F (t) = Φ(tW )w−1 = 1Ek Φ(tαk )βk−1 (14.18)
k=1
where in the last step our choice W := w0−1 w1 and the assumption w =
w01−θ w1θ show that W θ−j w−1 wj ≡ 1 for both j = 0, 1 (and indeed having this
identity dictates our choice of the auxiliary W ).
Now, by the Lions–Peetre method, we have
kf k(Lqw0 (S;Y0 ),Lqw1 (S;Y1 ))θ,q 6 sup kt 7→ tj−θ F (t)kLqj ( dt/t;Lqwj (S;Yj ))
0 1 0 ,q1 j
j=0,1
6 lim k1Sn f k q
Lw (S;Y ) = kf k Lqw (S;Y ).
n→∞
unconditionally.
Step 2 – (Lqw00 (S; Y0 ), Lqw11 (S; Y1 ))θ,q0 ,q1 ⊆ Lwq
(S; Y ):
q0 q1
Let f ∈ (Lw0 (S; Y0 ), Lw1 (S; Y1 ))θ,q0 ,q1 . We make the same initial assumptions
on both
R ∞ f and the weights wj as in the previous part. By definition, we have
q
f = 0 F (t) dtt with t
j−θ
F (t) ∈ Lqj ( dt/t; Lwjj (S; Yj )). Working the previous
computations backwards, we find that
Z ∞ Z ∞ Z ∞
dt −1 dt dt
φ := f w = F (t)w = F (tW )w =: Φ(t) ,
0 t 0 t 0 t
where Φ satisfies the relevant measurability conditions (by the structural as-
sumptions on the weights) and the quantitative relation (14.19). We conclude
that
kφk(Lq0 (S;Y0 ),Lq1 (S;Y1 ))θ,q0 ,q1 6 sup ktj−θ Φ(t)kLqj ( dt/t;Lqj (S;Yj ))
j=0,1
k1Sn f kLw
q
(S;Y ) 6 k1Sn f k(Lqw0 (S;Y0 ),Lqw1 (S;Y1 ))θ,q
0 1 0 ,q1
6 kf k q q
(Lw00 (S;Y0 ),Lw11 (S;Y1 ))θ,q0 ,q1 ,
kf kLw
q
(S;Y ) = lim k1Sn f kLqw (S;Y ) 6 kf k(Lqw0 (S;Y0 ),Lqw1 (S;Y1 ))θ,q .
n→∞ 0 1 0 ,q1
kf kLw
q
(S;Y ) = kf kLq 1−θ (S;Y ) 6 ρ(1−θ)+θ kf kLq 1−θ (S;Y )
w0 θ
w1 θ
w0,ρ w1,ρ
6 ρkf k((Lqw0 q
(S;Y0 ),Lw11,ρ (S;Y1 ))θ,q0 ,q1
0,ρ
kyk`pws (Y ) 6 Ckyk1−θ
q
` 0 (Y )
kykθ`qw1 (Y )
,
ws 0 s1
Multiplying with 2ks0 and taking the infimum over all admissible pairs
(y (0) , y (1) ), we find
2ks0 kyk k 6 K(2k(s0 −s1 ) , y)
using the notation of Section C.3. In combination with the identity θ(s1 −s0 ) =
s − s0 and the fact that the K-functional is non-decreasing, this gives
X 1/p
kyk`w
p
s
6 |2k(s−s0 ) K(2k(s0 −s1 ) , y)|p
k>0
XZ 2(k+1)(s0 −s1 )
dt 1/p
6 C0 |t−θ K(2k(s0 −s1 ) , y)|p
k(s0 −s1 ) t
k>0 2
Z ∞
−θ dt 1/p
6 C0 |t K(t, y)|p = C0 kyk(`qw0s (Y ),`qw1s (Y ))θ,p ,
0 t 0 1
1/p
(θp)
where C0 = (1−2−(s 0 −s)p )1/p
if p < ∞. A simple modification of this argument
gives the same result with C0 = 1 if p = ∞.
To prove the reverse inequality it suffices to consider the case q0 = q1 = 1.
Discretising as before, we find
XZ 2k(s0 −s1 )
dt 1/p
kyk(`1ws (Y ),`1ws (Y ))θ,p 6 |t−θ K(t, y)|p
0 1 (k−1)(s0 −s1 ) t
k>0 2
X 1/p
6 C1 |2−θk(s0 −s1 ) K(2k(s0 −s1 ) , y)|p ,
k>0
P∞ P∞
where C2 = k=0 2−k(s0 −s) and C3 = k=1 2−k(s−s1 ) . This gives the inequal-
ity
kyk(`1ws 1
(Y ),`w s
(Y ))θ,p 6 C1 (C2 + C3 )kyk`w
p
s (Y )
.
0 1
The final assertion is immediate from the first assertion and the log-
convexity inequality (L.2).
where ψk (x) := 2k ψ(2k x) and (εk )k∈Z is a Rademacher sequence. With an eye
toward the ensuing discussion we also remark that we have an equivalence of
norms
X
kf kLp (Rd ;X) h ε k ϕk ∗ f , (14.21)
Lp (Ω×Rd ;X)
k∈N
q, these spaces are canonically isometric by Fubini’s theorem. The two choices
lead to the theory of Besov spaces and Triebel–Lizorkin spaces, respectively.
q
The choice `w s
(Z) with the (homogeneous) Littlewood–Paley sequence
(ψk )k∈Z as in (14.20) leads to the so-called homogeneous Besov and Triebel–
Lizorkin spaces. Alternatively, the choice `qws (N) and the use of Littlewood–
Paley sequences (ϕk )k∈N as introduced in Definition 14.2.7 leads to the in-
homogeneous version of these spaces. In what follows we will only present in
the inhomogeneous case. Both classes of spaces are used in applications to
PDE. The advantage of inhomogeneous spaces is that, in the development
of their theory, one can make effective use of Schwartz functions and tem-
pered distributions. The theory of homogeneous spaces is technically more
involved and requires the use of different classes of test functions and equiv-
alence classes of tempered distributions modulo polynomials. Since we have
already encountered Schwartz functions and tempered distributions in many
places, we choose to only develop the theory of inhomogeneous spaces here.
Homogeneous spaces have better scaling properties, and scaling often plays
a crucial role in PDE, but for the purposes of the theory developed here
homogeneous spaces are not essential.
The proofs of (14.20) and (14.21) require the Banach space X to be UMD.
In contrast, in the theory of Besov and Triebel–Lizorkin spaces these norm
equivalences are promoted to definitions, thus eliminating the need of impos-
ing any conditions on X. By taking this approach, most of the fundamental
results in the theory of Besov spaces and Triebel–Lizorkin spaces are true
for arbitrary Banach spaces X. They come with their own versions of the
Mihlin multiplier theorem which does not require the UMD property either,
allowing multipliers without singularities at the origin in case of inhomo-
geneous spaces. The more general multipliers considered in Chapter 5 have
corresponding versions for homogeneous Besov and Triebel–Lizorkin spaces.
Perhaps more surprising is the fact that also for the duality theory of these
spaces no geometrical conditions need to be imposed on X. This contrast
the duality theory for the Bochner spaces, which requires that X ∗ have the
Radon–Nikodým property.
kf kBp,q
s (Rd ;X) := (2ks ϕk ∗ f )k>0 `q (Lp (Rd ;X))
is finite.
14.4 Besov spaces 321
Here, (ϕk )k>0 is the inhomogeneous Littlewood–Paley sequence that has been
fixed throughout the chapter (see Convention 14.2.8). By the discussion of
(14.13), the tempered distribution ϕk ∗ f is a C ∞ -function of polynomial
growth, so that the Lp -norm in the above definition makes sense.
s (Rd ;X) is indeed a norm, suppose that kf kB s (Rd ;X) = 0.
To see that k·kBp,q p,q
Indeed, for q0 6 q1 this follows from (14.22) and the inequality 2ks0 6 2ks1
for k > 0. For q0 > q1 this follows from Hölder’s inequality with q11 = q10 + 1r
and using that k>0 2−k(s0 −s1 )r < ∞.
P
The proof will give explicit constants depending only on s and ϕ0 (in one
direction), respectively s and ψ0 (in the other direction).
Proof. Suppose (ψk )k>0 is another inhomogeneous Littlewood–Paley seq-
uence. Then the analogues of (14.10) and (14.11) hold with ϕj and ψk ; in
particular for all j, k > 0 with |j − k| > 2 we have ϕk ∗ ψj = 0. Using
(14.12) for the sequence (ψk )k>0 , the triangle inequality, Young’s inequality,
and (14.7), we obtain
where we used (14.7). This gives the required estimate in one direction. The
reverse estimate is obtained by reversing the rôles of ϕk and ψk .
k1Bk ∂ β fbkL1 (Rd ;X) 6 k1Bk (1 + | · |2M )−1 kL1 (Rd ) k(1 + | · |2M )∂ β fbkL∞ (Rd ;X)
X
6 |Bk |(1 + 22M (k−1) )−1 [fb]β,δ ,
|δ|62M
using the notation (14.2) for the seminorms defining the Schwartz space. Keep-
ing
P in mind that |Bk | .d 2kd we now choose M = Ms,p,d ∈ N so large that
ks kd 2M (k−1) −1
k>0 2 2 (1 + 2 ) < ∞. With this choice, we obtain the estimate
X
kf kBp,1
s (Rd ;X) .d,p,s [fb]β,δ .
|δ|62M
14.4 Besov spaces 323
X 1 X
X
f (ψ) = fk (ψ` ) = fk (ψk+` ).
k,`>0 `=−1 k>0
Thus, by (14.13),
1 XZ
X
kf (ψ)k 6 kfk (x)k|ψk+` (x)| dx
`=−1 k>0 Rd
1
X
6 (2ks kfk (·)k)k>0 `∞ (Lp (Rd ;X))
(2−ks ψk+` )k>0 `1 (Lp0 (Rd ))
`=−1
6 3 · 2|s| kf kBp,∞
s (Rd ;X) kψkB −s
0 (Rd ) .
p ,1
from this.
Step 3 – To prove density, by Lemma 14.2.1 it suffices to prove the density
of S (Rd ; X) in Bp,q
s
(Rd ; X).
s d
Pn
Fix f ∈ Bp,q (R ; X) and set ζn := k=0 ϕk . By (14.6) we have kζn k1 =
kϕk1 .
s
We will first show that ζn ∗ f → f in Bp,q (Rd ; X). Fix ε > 0 and choose
K ∈ N such that X q
2ksq kϕk ∗ f kL q
p (Rd ;X) < ε .
k>K
k>K
kf − ζn ∗ f kBp,q
s (Rd ;X)
324 14 Function spaces
X 1/q
= 2ksq kϕk ∗ (f − ζn ∗ f )kqLp (Rd ;X)
k>0
K
X 1/q
6 2ksq kϕk ∗ (f − ζn ∗ f )kqLp (Rd ;X) + ε(1 + kϕk1 ).
k=0
The proof requires some preparations. Recall that a sequence (fn )n>1 is said
to converge in S 0 (Rd ; X) if there exists an f ∈ S 0 (Rd ; X) such that fn (φ) →
f (φ) in X for all φ ∈ S (Rd ). Likewise, it is said to be Cauchy in S 0 (Rd ; X)
if (fn (φ))n>1 is a Cauchy sequence in X for all φ ∈ S (Rd ).
Proof. Let (fn )n>1 be a Cauchy sequence in S 0 (Rd ; X). Since X is complete
we may define a linear mapping f : S (Rd ) → X by f (φ) := limn→∞ fn (φ).
We claim that f is continuous. Indeed, for every φ ∈ S (Rd ) the sequence
(fn (φ))n>1 is bounded in X, and therefore the Banach–Steinhaus theorem for
topological vector spaces implies that the sequence (fn )n>1 is equicontinuous.
Hence, given an ε > 0, we can find an open neighbourhood V of 0 in S (Rd )
such that |fn (φ)| 6 ε for all φ ∈ V and n > 1. Taking limits, it follows that
|f (φ)| 6 ε for all φ ∈ V . This means that f is continuous at zero and hence
continuous.
14.4 Besov spaces 325
A normed space E ,→ S 0 (Rd ; X) is said to have the Fatou property if for all
sequences (fn )n>1 in E such that
kϕk ∗ f kLp (Rd ;X) 6 lim inf kϕk ∗ fn kLp (Rd ;X) < ∞.
n→∞
Lemma 14.4.7. Every normed space E ,→ S 0 (Rd ; X) with the Fatou prop-
erty is complete.
where both expressions are infinite whenever one of them is. Here the (ϕk )k>0
are as in Definition 14.4.1, and thus the left-hand side of the above identity
equals kf kBp,q
s (Rd ;X) . The (ψk )k∈Z are as in (14.20). The first expression on
the right-hand side is equal to the homogeneous Besov norm, which we will
not discuss in detail.
To prove the norm equivalence first recall that ψk = ϕk for k > 1. For “.”
it suffices to observe that by Young’s inequality
ϕ0 ∗ f Lp (Rd ;X)
6 kϕ0 k1 kf kLp (Rd ;X) .
s
Conversely, assume that f ∈ Bp,q (Rd ; X). Since ϕ
b0 = 1 on supp(ψbk ) for k 6 0,
we can write
= kf kBp,1
0 (Rd ;X) 6 Cs,q kf kB s (Rd ;X)
p,q
The goal of this section is to prove a version of the Mihlin multiplier theorem
for operator-valued Fourier multipliers acting on vector-valued Besov spaces.
In contrast to the situation in the Lp -setting (cf. Theorems 5.3.18 and 5.5.10),
where we had to assume the UMD property, a variant of the Mihlin theorem
for Besov spaces holds for arbitrary Banach spaces.
We wish to emphasise that the main result, Theorem 14.4.16 below, is
not applicable to multipliers which are non-smooth or even singular near the
origin. This is due to the presence of the term ϕ0 in the definition of in-
homogeneous Littlewood–Paley sequences, whose support contains the origin
in its interior. For instance, the Fourier multiplier associated to the Hilbert
transform does not satisfy the conditions of the theorem.
Unlike in other chapters, we also include the case p = ∞. In order to avoid
density issues, we define ML∞ (Rd ; X, Y ) as the space of Fourier transforms
of operator-valued measures of bounded variation:
14.4 Besov spaces 327
ML∞ (Rd ; X, Y ) := Φ
b : the operator-valued measure
Φ : B(Rd ) → L (X, Y ) is of bounded variation .
Remark 14.4.10. In the scalar case it can be shown that the space ML∞ (Rd ) =
ML∞ (Rd ; C, C) as defined in Definition 14.4.9 coincides with the space of all
m ∈ L∞ (Rd ) for which the quantity
is finite, and that this quantity then equals the norm on ML∞ (Rd ) introduced
above. This provides further motivation for Definition 14.4.9.
In the discussion preceding Lemma 14.2.4 it was observed that for any function
φ ∈ L1 (Rd ; L (X, Y )), an operator-valued measure Φ : B(Rd ) → L (X, Y ) of
bounded variation is obtained by setting
Z
Φ(A) := φ dx,
A
328 14 Function spaces
and that its total variation satisfies kΦk(Rd ) 6 kφkL1 (Rd ;L (X,Y )) . In this way
we obtain contractive embeddings
b
L1 (Rd ; L (X, Y )) ,→ ML∞ (Rd ; X, Y ) ,→ MLp (Rd ; X, Y ).
In particular, if m ∈ C d+1 (Rd ; L (X, Y )) and there exists an ε > 0 such that
Remark 14.4.14. Multipliers with singularities in the origin, such as the mul-
tiplier giving rise to the Hilbert transform, are not covered by Proposition
14.4.11.
where in the last step we applied the Leibniz rule as before and the Fourier
support properties of ϕ1 given by (14.8) and (14.9). Since m(2k ξ) = (2−2k +
|ξ|2 )σ/2 , it is elementary to check that the latter expression is uniformly
bounded in k > 1. A similar argument shows that ϕ0 m ∈ MLp (Rd ; X).
The simple multiplier result of Proposition 14.4.11 is already strong enough to
s
prove the version of Mihlin’s multiplier theorem for Besov spaces Bp,q (Rd ; X)
contained in Theorem 14.4.16 below, valid for arbitrary Banach spaces X and
integrability exponents p, q ∈ [1, ∞]. In the statement of the theorem the end-
points p = ∞ and q = ∞ create some technical difficulties, since we cannot use
the density of the Schwartz functions to define Tm . It is for this reason that in
the theorem we assume that the multiplier m is smooth and has derivatives
of polynomial growth. Many interesting multipliers satisfy this condition, and
to proceed with the development of the theory this version suffices for the
time being. A version which avoids this restriction on m will be presented in
Theorem 14.5.6.
When m ∈ C ∞ (Rd ; L (X, Y )) has derivatives of polynomial growth, one
can define the Fourier multiplier Tm as an operator from S 0 (Rd ; X) into
S 0 (Rd ; Y ) by Tm f := F −1 (mfb). To see that this is well-defined it suffices
to note that mfb ∈ S 0 (Rd ; Y ) for f ∈ S 0 (R; X). In the next theorem, Tm is
s
understood to be the restriction of this operator to Bp,q (Rd ; X). The theorem
s
then asserts that, under Mihlin type conditions on m, it maps Bp,q (Rd ; X)
s d
into Bp,q (R ; Y ).
330 14 Function spaces
kTm f kBp,q
s (Rd ;Y ) = (2ks ϕk ∗ F −1 mfb)k>0 `q (Lp (Rd ;Y ))
1
X
= 2ks F −1 ϕ
bk m ϕ
bk+` fb
k>0 `q (Lp (Rd ;Y ))
`=−1
1
X
6 2ks F −1 (ϕ
bk mfbk+` )k>0 `q (Lp (Rd ;Y ))
`=−1
1
X
6 sup kϕ
bk mkMLp (Rd ;X,Y ) (2ks fk+` )n>0 `q (Lp (Rd ;Y ))
k>0
`=−1
6 2|s| sup kϕ
bk mkMLp (Rd ;X,Y ) kf kBp,q
s (Rd ;X) .
k>0
To complete the proof we must show that supk>0 kϕ bk mkMLp (Rd ;X,Y ) .d Km .
First consider the case k > 1. Since the multiplier norm is invariant under
dilations by Proposition 5.3.8, it suffices to show that
Hence, by Leibniz’s rule the Mihlin condition on m, and the Fourier support
property of ϕ1 given by (14.9), for all |α| 6 d + 1 we have
s
defines an equivalent norm on Bp,q (Rd ; X)
Proof. As a consequence of Proposition 14.4.15 it suffices to prove the equiv-
alence of (14.27) with kJk f kBp,qs−k
(Rd ;X) . This can be deduced from Theorem
14.4.16 by an argument similar to the one in Theorem 5.6.11. In the present
situation it is important to note that the multipliers in the proof of the propo-
sition also satisfy the more restrictive condition (14.25). Below we present a
simplification of the argument of Theorem 5.6.11 adapted to the Besov space
case. Let hξi = (1 + |2πξ|2 )1/2 .
First we prove the estimate
k∂ α f kBp,q
s−k
(Rd ;X) 6 CkJk f kBp,q
s−k
(Rd ;X) .
(2πiξ)α k b
F [∂ α f ](ξ) = (2πiξ)α fb(ξ) = hξi f (ξ) =: mα (ξ)hξik fb(ξ).
hξik
332 14 Function spaces
One checks that mα satisfies the conditions of Theorem 14.4.16, and thus
−1
k∂ α f kBp,q
s−k
(Rd ;X) 6 Cα Cd,p,q kF [h·ik fb]kBp,q
s−k
(Rd ;X)
and therefore
hξi2k b X
hξik fb(ξ) = f (ξ) = cα,k mα (ξ)(2πiξ)α fb(ξ)
hξik
|α|6k
X
= α f (ξ),
cα,k mα (ξ)∂d
|α|6k
(2πiξ)α
where mα (ξ) = hξik
. Applying Theorem 14.4.16 to mα now gives
−1
kJk f kBp,q
s−k
(Rd ;X) = kF [h·ik fb]kBp,q
s−k
(Rd ;X)
X
6 |cα,k |kTmα ∂ α f kBp,qs−k
(Rd ;X)
|α|6k
X
6 Cd,p,k k∂ α f kBp,q
s−k
(Rd ;X) .
|α|6k
kf kBp,∞
0 (Rd ;X) = sup kϕk ∗ f kLp (Rd ;X)
k>0
6 sup kϕk kL1 (Rd ) kf kLp (Rd ;X) 6 2kϕ0 kL1 (Rd ) kf kLp (Rd ;X) ,
k>0
where the last step uses (14.7). This completes the proof of (14.31).
As we already noted, in order to prove the embeddings in (14.30) it suffices
0
to consider the case m = 0. Fix f ∈ B∞,1 (Rd ; X). As before we see that the
P∞
sum k=0 ϕk ∗ f is absolutely convergent in L∞ (Rd ; X). By Lemma 14.2.10
its sum equals f and
∞
X
kf k∞ 6 kϕk ∗ f k∞ = kf kB∞,1
0 (Rd ;X) .
k=0
The most interesting cases are (ii) and (iii), since they can be used to change
the integrability parameter from p0 into p1 .
For the proof of the sufficiency of the three conditions we need two lemmas.
The first provides an Lp -estimate for the derivatives under suitable Fourier
support assumptions. Recall from Lemma 14.2.9 that every f ∈ S 0 (Rd ; X)
with compact Fourier support belongs to C ∞ (Rd ; X) and has at most poly-
nomial growth.
The next lemma provides shows how the Lp -norm of ϕk ∗ ϕk+j scales with k.
Lemma 14.4.21. For all j ∈ Z there exists a constant Cd,j,p > 0 such that
for all k > 0 and k + ` > 0 we have
0
kϕk+` ∗ ϕk kLp (Rd ) = C`,p,d 2kd/p .
(2ks1 ak )k>0 `q 1
6 (2ks0 ak )k>0 `q 0
.
It follows that
Since t > s1 , it follows that the conditions (i) are satisfied, and thus
Next we move to the necessity of the conditions (i), (ii), and (iii). It suffices
to consider the case X = K.
Suppose that we have the continuous embedding stated in the theorem.
By the closed graph theorem there is a constant C = Cd,p0 ,p1 ,q0 ,q1 ,s0 ,s1 such
that for all f ∈ Bps00 ,q0 (Rd ),
where C is a constant independent of n > 1 and the sequence (ak )nk=1 . Once
established, this claim gives q0 6 q1 .
To prove the claim fix a scalar sequence (ak )nk=1 . Applying (14.32) to the
Pn −3k(s0 + pd0 ) Pn −3k(s1 + pd0 )
function f := k=1 2
0 a ϕ
k 3k = k=1 2
1 a ϕ
k 3k gives the
inequality
n q1 1/q1
−3k(s1 + pd0 )
X X
2ms1 q1 2 1 ak ϕm ∗ ϕ3k
Lp1 (Rd )
m>0 k=1
n
(14.35)
X X −3k(s0 + pd0 ) q0 1/q0
ms0 q0
6C 2 2 0 ak ϕm ∗ ϕ3k .
Lp0 (Rd )
m>0 k=1
Let us analyse the expressions on the left-hand and right-hand sides for general
values of p, q, and s. We have ϕm ∗ ϕ3k 6= 0 only for m = 3k + ` with
14.4 Besov spaces 337
` ∈ {−1, 0, 1}. This suggests splitting the sum over m into the sums over
m = 3j + ` for ` ∈ {−1, 0, 1}. Using the lemma, they evaluate as
X n q 1/q
−3k(s+ pd0 )
X
2(3j+`)sq 2 ak ϕ3j+` ∗ ϕ3k
Lp (Rd )
j>0 k=1
n
X q 1/q
−3j(s+ pd0 )
= 2(3j+`)sq 2 aj ϕ3j+` ∗ ϕ3j
Lp (Rd )
j=1
n 1/q
−3j(sq+ dq
X 0
)
= C`,p,d 2(3j+`)sq 2 p0 kaj kq 23jdq/p
j=1
n
X 1/q
= 2`s C`,p,d kaj kq .
j=1
We thus find (using the triangle inequality in `q3 for the upper estimate)
X n q 1/q n
X 1/q
−3k(s+ pd0 )
X
2msq 2 ak ϕm ∗ ϕ3k hd,p,s ka` kq .
Lp (Rd )
m>0 k=1 `=1
In this section we show that Besov spaces with smoothness parameter s > 0
admit a characterisation in terms of difference norms. This characterisation
can be often used to effectively check whether a given concrete function be-
longs to a given Besov space. For example, we check in Corollary 14.4.26
s
that the Besov spaces B∞,∞ (Rd ; X) coincide with certain spaces of s-Hölder
continuous functions.
For functions f : Rd → X and vectors h ∈ Rd , the function ∆h f : Rd → X
is defined by
∆h f (x) := f (x + h) − f (x).
Clearly, the difference operator ∆h thus defined is bounded as an operator on
Lp (Rd ; X) for all 1 6 p 6 ∞, with norm at most 2. We have the following
formula for the powers ∆m m
h = (∆h ) .
Here we used the notation −F := |F1 | F to denote the average over the set F .
R R
(m,τ )
The next theorem implies that if s > 0, then each of the norms ||| · |||B s d
p,q (R ;X)
s
with m > s defines an equivalent norm on Bp,q (Rd ; X).
Theorem 14.4.24 (Difference norms for Besov spaces). Let p, q ∈
[1, ∞], s > 0, τ ∈ [1, ∞], and let m > s be an integer. A function
f ∈ Lp (Rd ; X) belongs to Bp,q
s
(Rd ; X) if and only if [f ]m,τ
s (Rd ;X) < ∞, and
Bp,q
the following equivalence of norms holds:
(m,τ )
kf kBp,q
s (Rd ;X) hd,m,s |||f ||| s
B d .
p,q (R ;X)
Before turning to the details of the proof we give some simple applications.
The first two identify the Sobolev–Slobodetskii spaces and the Hölder spaces
(cf. Section 14.1 for the relevant notation) as Besov spaces.
Corollary 14.4.25 (Sobolev–Slobodetskii spaces). Let p ∈ [1, ∞) and
s ∈ (0, 1). Then
s
Bp,p (Rd ; X) = W s,p (Rd ; X)
with equivalent norms. In fact,
(1,p) 1
[f ]B s d = [f ]W s,p (Rd ;X) . (14.37)
p,p (R ;X) (sp + d)1/p |B1 |
14.4 Besov spaces 339
Proof. By Theorem 14.4.24 it suffices to prove the identity (14.37) for the
seminorms, which follows from Fubini’s theorem and a change of variable:
Z Z Z ∞
(1,p) p
|B1 |p [f ]B s (Rd ;X) = 1{|h|6t} t−sp−d−1 k∆h f (x)kp dt dh dx
p,p d d
R R 0
Z Z
−1
= (sp + d) |h|−sp−d k∆h f (x)kp dh dx
Rd Rd
p
= (sp + d)−1 [f ]W s,p (Rd ;X) .
340 14 Function spaces
The proof of Theorem 14.4.24 makes use the following simple lemma. Recall
the Fourier multiplier notation of Subsection 14.4.b.
Lemma 14.4.28. For non-zero ξ, h ∈ Rd let
e2πih·ξ − 1
mh (ξ) := .
2πih · ξ
Then for all p ∈ [1, ∞] we have mh ∈ MLp (Rd ; X) and kmh kMLp (Rd ;X) 6 1.
Proof. By an elementary computation, the associated Fourier multiplier is
given by
Z 1
Tmh f (x) = f (x − ht) dt = µh ∗ f (x), f ∈ Lp (Rd ; X),
0
R1
where µh (A) = 0 1th∈A dt defines a measure by monotone convergence.
Hence the result follows from (14.24). For p < ∞, one can also use the direct
estimate
Z 1
kTmh f kLp (Rd ;X) 6 kf (· − ht)kLp (Rd ;X) dt = kf kLp (Rd ;X) .
0
Proof of Theorem 14.4.24. Let
Z 1/τ
m,τ
Ip (f, k) := − k∆m τ
2−k h f k dh ,
{|h|61} Lp (Rd )
we obtain
X Z 2−k+1 Z 1/τ q 1/q
(m,τ )
[f ]B s d = t−sq−1 − k∆m τ
h f k dh dt
p,q (R ;X) Lp (Rd )
k∈Z 2−k {|h|6t}
X Z 1/τ q 1/q
6 2d/τ 2ksq − k∆m τ
h f k dh
{|h|62−k+1 } Lp (Rd )
k∈Z
X Z 1/τ q 1/q
= 2d/τ 2(j+1)sq − k∆m τ
2−j h f k dh
{|h|61} Lp (Rd )
j∈Z
X Z 1/τ q 1/q
= 2s+d/τ 2jsq − k∆m τ
2−j h f k dh .
{|h|61} Lp (Rd )
j∈Z
14.4 Besov spaces 341
Similarly,
X Z 1/τ q 1/q
(m,τ )
[f ]B s d > 2−s−1−d/τ 2jsq − k∆2m−j h f kτ dh .
p,q (R ;X) Lp (Rd )
j∈Z {|h|61}
Hence,
(m,τ )
[f ]B s d hd,s,τ (2ks Ipm,τ (f, k))k∈Z `q (Z)
. (14.38)
p,q (R ;X)
In view of (14.36) and (14.38) it thus suffices to prove the two estimates
Throughout the proof of (14.39) and (14.40) we will use the standard algebraic
properties of Lp -multipliers discussed in Section 5.3.a.
Put fj := ϕj ∗ f for j > 0. By Hölder’s inequality,
X
kf kLp (Rd ;X) 6 kfj kLp (Rd ;X) 6 (2−js )j>0 `q0 kf kBp,q
q
(Rd ;X) ,
j>0
0
where the assumption s > 0 implies the finiteness of the `q -norm. To prove
(14.39) and (14.40) it therefore remains to estimate Ipm,∞ (f, k) from above
and Ipm,1 (f, k) from below.
Step 1 – We begin with the proof of (14.39). By Lemma 14.2.10 and the
triangle inequality,
1 X
X
Ipm,∞ (f, k) 6 Ipm,∞ (ϕj ∗ fj+` , k),
`=−1 j>0
where in the last step we used Proposition 14.4.11 with ∂ α ϕ1 ∈ L1 (Rd ). Com-
bining (14.41) with (14.42), estimating the latter using (14.43) and (14.44),
we obtain the estimate
where we applied the discrete version of Young’s inequality and used the
assumption m > s for the finiteness of the `1 norm.
Step 2 – In this step we prove (14.40). For k > 0 let Tk f := 2kd ϕ(2k ·) ∗ f
and Sk f := ϕk ∗ f . By (14.3), for k > 1 we have Sk = Tk − Tk−1 = (I −
Tk−1 ) − (I − Tk ) and therefore
kf kBp,q
s (Rd ;X) = (2ks kSk f kLp (Rd ;X) )k>0 `q
ks
(14.45)
6 kS0 f kLp (Rd ;X) + 2 (2 kTk f − f kLp (Rd ;X) )k>0 `q
.
By Young’s inequality,
14.4 Besov spaces 343
and define the sequence (ϕk )k>0 as in (14.3). For |ξ| 6 1/m and 0 6 j 6 m−1
we have ψ(−(m
b − j)ξ) = 1 and therefore
m−1
X m
m X m
ϕ(ξ)
b = (−1)m+1 (−1)j = (−1)m+1 (−1)j − (−1)m = 1
j=0
j j=0
j
by the binomial theorem, and for |ξ| > 3/2 we have ϕ(ξ) b = 0. Furthermore
the Fourier supports of ϕj and ϕk are disjoint for |j −k| > Nm , where Nm ∈ N
only depends on m (rather than for |j − k| > 2 as in (14.10) in the case of an
inhomogeneous Littlewood–Paley sequence). Thanks to these properties, the
proof of Proposition 14.4.2 may be repeated to see that this system leads to
s
an equivalent norm on Bp,q (Rd ; X).
Let f ∈ Lp (Rd ; X). We claim that
Z
Tk f (x) − f (x) = (−1)m+1 ∆m2−k y f (x)ψ(y) dy (14.47)
Rd
Indeed, taking Fourier transforms in the x-variable and using Lemma 14.4.22
and the fact that ψ(0)
b = 1, we have
Td b b −k ξ) − 1)fb(ξ)
k f (ξ) − f (ξ) = (ϕ(2
m−1
X m
= (−1)m+1 (−1)j ψ(−(m
b − j)2−k ξ) − 1 fb(ξ)
j=0
j
m
X m
= (−1)m+1 (−1)j ψ(−(m
b − j)2−k ξ)
j=0
j
m Z
m+1
X m j −k
= (−1) (−1) e2πi(m−j)2 y·ξ ψ(y) dy
j=0
j R d
Z
−k
= (−1)m+1 (e2πi2 y·ξ − 1)m fb(ξ)ψ(y) dy
d
ZR
= (−1) m+1
F (∆2m−k y f )(ξ)ψ(y) dy
Rd
Fix a real number r > 0, the numerical value of which will be fixed in a
moment. Taking norms in (14.47), using that supx∈Rd (1 + |x|r )|ψ(x)| < ∞,
and writing BR := {ξ ∈ Rd : |ξ| 6 R}, it follows that
kf (x) − Tk f (x)k
Z
6 k∆m
2−k y f (x)ψ(y)k dy
R d
Z X Z
−(j+1)r
.ψ k∆m −k
2 y f (x)k dy + 2 k∆2m−k y f (x)k dy
B1 j>0 B2j+1 \B2j
Z X Z
= k∆2m−k y f (x)k dy + 2−(j+1)(r−d) k∆m
2j+1−k h f (x)k dh
B1 j>0 B1 \B 1
2
X Z
−j(r−d)
6 2 k∆m
2j−k h f (x)k dh.
j>0 B1
14.4.e Interpolation
In order to consider interpolation for Besov spaces, we will now introduce the
so-called retraction and co-retraction operators, which allow us to reduce ques-
tions about the interpolation of Besov spaces to the corresponding questions
about the spaces `qws (Lp (Rd ; X)).
14.4 Besov spaces 345
by
X
R((fk )k>0 ) = ψk ∗ f k , Sf = (ϕk ∗ f )k>0 .
k>0
and
Eθ := (E0 , E1 )θ , Fθ := (F0 , F1 )θ , Xθ := [X0 , X1 ]θ .
By Theorem 2.2.6 and Proposition 14.3.3, Eθ = `qws (Lp (Rd ; Xθ )) isometrically.
Therefore,
[Bps00 ,q0 (Rd ; X0 ), Bps01 ,q0 (Rd ; X1 )]θ = Fθ = RSFθ ⊆ REθ ⊆ Bp,q
s
(Rd ; Xθ ),
kf kFθ = kRSf kFθ 6 CkSf kEθ = CkSf k`rw (Lp (Rd ;Xθ )) = Ckf kBp,q
s (Rd ;X ) ,
θ
s
and
Eθ,q := (E0 , E1 )θ,q , Fθ,q := (F0 , F1 )θ,q .
By Proposition 14.3.5, Eθ,q = `qws (Lp (Rd ; X))with equivalent norms, say with
constants C1 , C2 (depending on θ, p, q, s0 , s1 ), i.e.,
C1−1 kgkEθ,q 6 kgk`w
q p d
s (L (R ;X))
6 C2 kgkEθ,q .
From Theorem C.3.3 it follows that
s0
(Bp,r (Rd ; X), Bp,r
s1 s
(Rd ; X))θ,q = Fθ,q = RSFθ,q ⊆ REθ,q ⊆ Bp,q (Rd ; X),
and for all f ∈ Fθ,q we have
s (Rd ;X) = kSf k`q (Lp (Rd ;X)) 6 C2 kSf kE
kf kBp,q ws θ,q
= C2 kf kFθ,q .
In the converse direction, interpolation R and S by Theorem C.3.3,
s
Bp,q s
(Rd ; X) = RSBp,q (Rd ; X) ⊆ REθ,q ⊆ Fθ,q ,
s
and for all f ∈ Bp,q (Rd ; X) we have
kf kFθ,q = kRSf kFθ,q
6 CkSf kEθ,q . CkSf k`qws (Lp (Rd ;X)) = C3 C1 kf kBp,q
s (Rd ;X) ,
Corollary 14.4.32. Let s0 , s1 ∈ [0, ∞) satisfy s0 6= s1 , let θ ∈ [0, 1], and put
s := (1 − θ)s0 + θs1 . Then
s0 s1
(Cub (Rd ; X), Cub (Rd ; X))θ,∞ = B∞,∞
s
(Rd ; X)
s
with equivalent norms. Moreover, if s ∈
/ N, then B∞,∞ s
(Rd ; X) = Cub (Rd ; X)
with equivalent norms and therefore
s0 s1
(Cub (Rd ; X), Cub (Rd ; X))θ,∞ = (Cub
s
(Rd ; X).
14.4.f Duality
s
The main result of this section identifies the duals of Besov spaces Bp,q (Rd ; X)
for p, q ∈ [1, ∞). It is interesting that no geometric assumptions are needed on
X. This contrasts with the situation for vector-valued Bochner spaces: recall
that, by Theorem 1.3.10, for σ-finite measures spaces one has Lp (S; X) =
0
Lp (S; X ∗ ) if and only if X ∗ has the Radon–Nikodým property.
We start with the preliminary observation that elements in the duals of
Besov spaces can be naturally identified with tempered distributions. Indeed,
s
if g ∈ Bp,q (Rd ; X)∗ , then for all ϕ ∈ S (Rd ) and x ∈ X we have
Proof. The second assertion follows from the first, combined with Corollary
14.4.25.
As a preliminary observation to the proof of the first assertion, we recall
Proposition 2.4.32, which asserts that if g ∈ S 0 (Rd ; X ∗ ) and ζ ∈ S (Rd ),
then ζ ∗ g is in C ∞ (Rd ; X ∗ ) and ∂ α g has polynomial growth for any α ∈ Nd .
Moreover, by Lemma 14.2.10, and the support properties (14.11), (14.12), we
have the identity
XZ 1 XZ
X
g(ζ) = hζ(t), gj (t)i dt = hϕj+` ∗ ζ(t), gj (t)i dt, (14.54)
j>0 Rd `=−1 j>0 Rd
where gj := ϕj ∗ g.
We split the proof of the theorem into three steps.
Step 1 – First let g ∈ Bp−s d ∗
0 ,q 0 (R ; X ). Identifying g with an element of
By Hölder’s inequality,
1 XZ
X
|g(f )| 6 |hfj+` (t), gj (t)i| dt
`=−1 j>0 Rd
1
X
6 2−`s (2(j+`)s fj+` )j>0 `q (Lp (Rd ;X))
(2−js gj )j>0 `q0 (Lp0 (Rd ;X ∗ ))
`=−1
6 3 · 2|s| kf kBp,q
s (Rd ;X) kgk −s
B 0 (Rd ;X ∗ ) .
p ,q 0
zero sequence in S (Rd ) ⊗ X such that k(2js fj )j>0 k`q (Lp (Rd ;X)) 6 1. Put f :=
R(fj )j>0 , where R : `qws (Lp (Rd ; X)) → Bp,q s
(Rd ; X) is the operator considered
in Lemma 14.4.29. Then by (14.54) and the fact that ψbj = ϕ bj−1 +ϕ bj +ϕ
bj+1 = 1
on supp(ϕ bj ) we see that
XZ XZ
g(f ) = hf (t), gj (t)i dt = hfj (t), gj (t)i dt.
j>0 Rd j>0 Rd
Therefore,
XZ
h2js fj (t), 2−js gj (t)i dt = |g(f )| 6 kf kBp,q
s (Rd ;X) kgkB s (Rd ;X)∗
p,q
j>0 Rd
6 kRk kgkBp,q
s (Rd ;X)∗ .
Taking the supremum over all admissible finitely non-zero sequences (fj )j>0 ,
Propositions 1.3.1 and 1.3.3 imply that g belongs to Bp−s d ∗
0 ,q 0 (R ; X ) and
Step 3 – Since the identifications in Steps 2 and 3 are inverse to each other,
they set up a bijective correspondence, and the estimates in the above proof
show that this correspondence is bounded in both directions.
Theorem 14.4.34 permits an extension of Example 14.4.33 to negative smooth-
ness exponents.
f 7→ ζf
14.4 Besov spaces 351
s
defines a bounded operator from Bp,q (Rd ; X) into Bp,q
s
(Rd ; Y ) of norm
kf 7→ ζf kL (Bp,q
s (Rd ;X),B s (Rd ;Y )) .k,s kζkC k (Rd ;L (X,Y )) .
p,q b
(14.55)
To prove this, first assume that q ∈ (1, ∞) and s < 0. From Example 14.4.33
we obtain the boundedness of g 7→ ζ ∗ g from Bp−s d ∗ −s d ∗
0 ,q 0 (R ; Y ) into Bp0 ,q 0 (R ; X ).
Proof. By Proposition 14.4.3 it suffices to show that for every f ∈ Cc∞ (Rd )
there exist fn ∈ Cc∞ (R̈d ) such that fn → f in Bp,q s
(Rd ). Moreover, by the
embedding (14.23) and Theorem 14.4.19 it suffices to prove (2).
In order to prove (2) let fn := ζn f , where ζn (x) = ζ(nx1 , x2 , . . . , xn ) is
multiplication by n in the first coordinate, and where ζ ∈ C ∞ (Rd ) satisfies
ζ = 1 if |x1 | > 2 and ζ = 0 if |x1 | 6 1. Then by Theorem 14.4.31 the following
interpolation inequality holds:
1/p0 1/p
kfn kB 1/p (Rd ) 6 Ckfn kLp (Rd ) kfn kW 1,p (Rd ) .
p,q
Since
352 14 Function spaces
(2) X has cotype q if and only if the identity mapping on Cc∞ (Rd )⊗X extends
to a continuous embedding
( 1 − 12 )d
γ(L2 (Rd ), X) ,→ Bq,q
q
(Rd ; X).
The proof of Theorem 14.5.1 provides quantitative estimates for the norms of
these embeddings. It relies on the following Gaussian version of the Bernstein–
Nikolskii inequality (Lemma 14.4.20).
Here, κ2,p and κq,2 are the Kahane–Khintchine constants introduced in Section
γ
6.2 and τq,X and cγq,X are the Gaussian type and cotype constants of X,
respectively, introduced in Section 7.1.d.
Proof. (1): By a scaling argument it suffices to consider the case t = 21 . By
Example 9.6.5, ∂ α f ∈ γ(Rd ; X) if and only if ξ 7→ ξ α fb ∈ γ(Rd ; X) and in this
case
k∂ α f kγ(Rd ;X) = (2π)|α| kξ 7→ ξ α fb(ξ)kγ(Rd ;X) .
In order to show that ξ 7→ ξ α fb(ξ) ∈ γ(Rd ; X), by Examples 9.1.12 and 9.4.4
it suffices to check fb ∈ γ(Q; X), where Q := [− 12 , 12 ]d ; in that case
The assertion fb ∈ γ(Q; X) is short-hand for the statement that the Pettis
integral operator Ifb : L2 (Q) → X defined by
Z
Ifbg := fb(ξ)g(ξ) dξ, g ∈ L2 (Q),
Q
belongs to γ(L2 (Q), X) (see Section 9.2.a). We will prove the latter by testing
against an orthonormal bases, making use of Theorem 9.1.17.
Let en (ξ) := e2πin·ξ for n ∈ Zd and ξ ∈ Q. These functions define an
orthonormal basis for L2 (Q) and we have
Z
Ifben = fb(ξ)e2πin·ξ dξ = f (n).
Q
It follows from Theorem 9.1.17 that fb ∈ γ(Q, X) and, by the above observa-
tions,
X 1/p
γ
k∂ α f kγ(Rd ;X) 6 π |α| kfbkγ(Q;X) 6 κ2,p τp,X π |α| kf (n)kp .
n∈Zd
To deduce the estimate in the statement of the theorem from it, for h ∈ Q
and s ∈ Rd put fh (s) := f (s + h). Then supp fc
h ⊆ Q and
X 1/p
γ
k∂ α f kγ(Rd ;X) = k∂ α fh kγ(Rd ;X) 6 κ2,p τp,X π |α| kfh (n)kp .
n∈Zd
By Proposition 9.4.13, applied to the decompositions (S2k )k>0 and (S2k+1 )k>0
of Rd \ {0}, for n > m > 0 we obtain
2n n
X 1/p
X 1 1
fk 6 γ
κ2,p τp,X τp,X 22j( p − 2 )pd kf2j kpLp (Rd ;X)
γ(Rd ;X)
k=2m j=m
14.5 Besov spaces, random sums, and multipliers 355
n−1
X 1 1
1/p
γ
+ κ2,p τp,X τp,X 2(2j+1)( p − 2 )pd kf2j+1 kpLp (Rd ;X) .
j=m
( 1 − 12 )d
Since S (Rd ; X) is dense in Bp,p
p
(Rd ; X) by Proposition 14.4.3, the identity
( 1 − 1 )d
mapping on S (Rd ; X) extends to a bounded operator from Bp,p p 2
(Rd ; X)
γ
into γ(Rd ; X) of norm at most 2κ2,p τp,X τp,X . The simple proof that this ex-
tension is injective is left to the reader.
Next we prove the ‘if’ part. Since every Banach space has type 1, the ‘if’
part is trivial for p = 1. In the rest of the proof of (1) we may therefore assume
that p ∈ (1, 2]. We will prove the stronger statement that if for some r ∈ (1, ∞]
the identity operator on S (Rd ; X) extends to a bounded operator, say I,
( 1 − 1 )d
from Bp,rp 2
(Rd ; X) into γ(L2 (Rd ), X), X has type r (and then necessarily
r ∈ (1, 2]).
Let ψ ∈ S (Rd ) be such that kψkL2 (Rd ) = 1 and supp(ψ) b ⊆ {ξ ∈ Rd :
d
b1 (ξ) = 1}. For n > 1, let ψn ∈ S(R ) be defined by
ϕ
b −n+1 ξ).
ψbn (ξ) := 2(−n+1)d/2 ψ(2
N
X 2
kf k2γ(Rd ;X) = E γn xn .
n=1
By putting things together we see that X has type r, with Gaussian type r
γ
constant τr,X 6 kψkp kIk.
(2): This is proved similarly.
356 14 Function spaces
q0
(2) F restricts to a bounded operator from W bsc+1,p
(R ; X) into L (Rd ; X).
d
The case q = ∞ gives sufficient conditions for the Fourier transform to take
values in L1 (Rd ; X). Different conditions guaranteeing this have been dis-
cussed in Lemma 14.2.11, where growth assumptions on the functions and
their derivatives where imposed.
Proof. We start with case (i). Accordingly, let q ∈ [p, ∞] and let X have
Fourier type p
s d d
(1): Let f ∈ Bp,q 0 (R ; X). Put fk := ϕk ∗ f for k > 0. Let I0 = {ξ ∈ R :
|ξ| < 1} and
In := {ξ ∈ Rd : 2n−1 6 |ξ| < 2n }, n > 1.
The sets In thus defined are pairwise disjoint, we have n>0 In = Rd , and
S
1/q0 1 X 1/q0
0 0
X X
kfbkq0 = k1In fbkqq0 6 k1In fbn+` kqq0 ,
n>0 `=−1 n>0
P
where we used that supp(ϕ bk ) ∩ In = ∅ for |n − k| > 2 and that k>0 ϕ
bk = 1.
1 s 1
By Hölder’s inequality with q0 = d + p0 and the Fourier type p assumption,
for ` ∈ {−1, 0, 1} we have
6 ϕp,X (Rd )2(n+1)s kfn+` kp0 6 22s ϕp,X (Rd )2(n+`)s kfn+` kp0 .
0 0
Taking `q -norms on both sides we obtain fb ∈ Lq (Rd ; X) and
(2): This follows from (1) since by Proposition 14.4.18 and Theorem
bsc+1
14.4.19 we have the embeddings W bsc+1,p (Rd ; X) ,→ Bp,∞ (Rd ; X) ,→
s d
Bp,1 (R ; X).
Case (ii): Assume now that q ∈ [2, ∞] and that X has type p and cotype 2.
Using the same notation as in case (i), by Hölder’s inequality with q10 = 1r + 12 ,
Theorem 9.2.10, and Lemma 14.5.2 we have
d/τ
i.e., every m ∈ Bτ,1 (Rd ; L (X, Y )) defines a bounded operator Tm from
Lp (Rd ; X) to Lp (Rd ; Y ).
b
Proof. The result is immediate from the fact that m ∈ L1 (Rd ; L (X, Y )) by
Proposition 14.5.3.
This proves that Tm extends uniquely to Tm ∈ L (L1 (Rd ; X), Lb1 (Rd ; Y )).
Since the second line of (14.58) trivially implies that the kernel m satisfies
the dual Hörmander’s condition, it follows from the Calderón–Zygmund ex-
trapolation theorem (Theorem 11.2.5) that Tm extends uniquely to Tm ∈
L (Lp (Rd ; X), Lp (Rd ; Y )) for all p ∈ [1, ∞). By a duality argument a similar
result can be derived for p = ∞.
It is clear from the above proof that we can replace the Fourier type
conditions by the conditions that Y has type τ2 and cotype 2, and X ∗ has
type τ1 and cotype 2.
s
then there is a bounded operator T : Bp,q (Rd ; X) → Bp,q
s
(Rd ; Y ) with kT k 6
Cd,s,X,Y Km such that T f = F (mfb) for all f ∈ S (R ) ⊗ X.
−1 d
Note that in the case p, q < ∞, one has that T is the unique bounded extension
of Tm : S (Rd ) ⊗ X → S 0 (Rd ; Y ). In the end point case p = ∞ or q = ∞ this
does not make sense since S (Rd ) ⊗ X is not dense in Bp,qs
(Rd ; X). This is the
main reason for the unusual formulation in Theorem 14.5.6.
By a duality argument one can also formulate the (Fourier) (co)type con-
ditions on X ∗ , but the end-point cases require some caution.
s
Proof. For f ∈ Bp,q (Rd ; X) let fk = ϕk ∗ f and mk = ϕ
bk m. Define
14.5 Besov spaces, random sums, and multipliers 359
1 X
X
Tf = Tmk+` fk . (14.59)
`=−1 k>0
We will check that the series converges in S 0 (Rd ; Y ) and defines an element
s
in Bp,q (Rd ; Y ).
The proof follows the lines of Theorem 14.4.16. First we show that mk
bound MLp (Rd ; X, Y ) with a uniform bound in k > 0. First let k > 1. By
invariance under dilations (see Proposition 5.3.8), Corollary 14.5.4, and the
embeddings (14.23) and (14.29), we have
Since mk (2k−1 ·) = ϕb1 (·)m(2k−1 ·), by the support properties of ϕb1 is suffices
α
to bound ∂ [ϕ b1 (ξ)m(2k−1 ξ)] for |α| 6 b τd c + 1, uniformly in k > 1 and 1 6
|ξ| 6 3. This can be done in the same way as in (14.26). The case k = 0 can
be proved in the same way without the dilation argument. We can conclude
that
Next we check the convergence of the series in (14.59). For ζ ∈ S (Rd ) one
P1
has Tmk+` fk (ζ) = j=−1 Tmk+` fk (ζk+j ), where ζk = ϕk ∗ ζ, and thus
1
X
kTmk+` fk (ζ)kY 6 kTmk+` fk kLp (Rd ;Y ) kζk+j kLp0 (Rd )
j=−1
1
X
6 Cd,s,X,Y Km 2sk kfk kLp (Rd ;Y ) 2|s| 2−s(k+j) kζk+j kLp0 (Rd )
j=−1
X X 1
X
kTmk+` fk (ζ)kY 6 Cd,s,X,Y Km 2sk kfk kLp (Rd ;Y ) 2−sk kζk+j kLp0 (Rd )
k>0 k>0 j=−1
|s|
6 3 · 2 Cd,s,X,Y Km kf kBp,q
s (Rd ;X) kζk −s
B 0 (Rd ) ,
p ,q 0
A further consequence of Proposition 14.5.3 is a Fourier multiplier theorem of
a very different nature, in which the multiplier is non-smooth but the domain
and range spaces have different integrability and smoothness exponents.
(quasi-)norm
0
kf kLr0 ,σ (Rd ) := τ 7→ τ 1/r f ∗ (τ ) Lσ (R+ , dτ
τ )
is finite, where
8.1.22,
1 X
X
R(T (t) : t ∈ Rd ) 6 R(ϕk+` ∗ Tk (t) : t ∈ Rd ), (14.63)
`=−1 k>0
provided of course that the operator families occurring in the sums are R-
bounded and their R-bounds are summable. Proving this will occupy us in
the remainder of the proof.
Fix an integer n > 1. Starting from the identity ϕn (t) = 2(n−1)d ϕ1 (2n−1 t)
(see (14.4)), it is elementary to check that the non-increasing rearrangements
satisfy ϕ∗n (τ ) = 2(n−1)d ϕ1∗ (2n−1 τ ). Therefore,
0
kϕn kLr0 ,σ (Rd ) = 2(n−1)d kτ 7→ τ 1/r ϕ1∗ (2n−1 τ )kLσ (R+ , dτ )
τ
1/r 0
=2 (n−1)d/r
kτ 7→ τ ϕ∗1 (τ )kLσ (R+ , dτ ) = 2(n−1)d/r kϕ1 kLr0 ,σ (Rd ) ,
τ
the latter being finite since ϕ1 ∈ S (Rd ). A similar calculation can be done
for n = 0.
For t ∈ Rd define ϕn,t ∈ S (Rd ) by ϕn,t (s) := ϕn (t − s). Then ϕn,t is
identically distributed with ϕn . Letting Tk,ϕn,t ∈ L (X, Y ) be the integral
operator from Proposition 8.5.16, i.e.,
Z
Tk,ϕn,t x := ϕn,t (s)Tk (s)x ds,
Rd
it follows from Proposition 8.5.16 with σ = r0 min{ p10 , 1q } and ψ = ϕn that for
all n > 0 and k > 0 the set {ϕn ∗ Tk (t) : t ∈ Rd } is R-bounded, with R-bound
We have the following variation of this result for the strong operator topology:
Theorem 14.5.9 (Besov functions with R-bounded range – II). Let X
and Y be Banach spaces and assume that Y has type p ∈ [1, 2]. Suppose that
d/p
T : Rd → L (X, Y ) satisfies T x ∈ Bp,1 (Rd ; Y ) for all x ∈ X and
14.5 Besov spaces, random sums, and multipliers 363
X M
X
6 Lp,Y Cϕ 2kd/2 Tk εm xm ,
L2 (Ω;γ(L2 (Rd ),Y ))
k>0 m=1
M
X
X 1 1
6 C0 2kd/2 2k( p − 2 )d Tk εm x m
L2 (Ω;Lp (Rd ;Y ))
k>0 m=1
Z X M
X
6 C0 κ2,1 2kd/p Tk εm x m dP
Ω k>0 Lp (Rd ;Y )
m=1
Z M
X
= C0 κ2,1 T εm xm d/p
dP
Ω Bp,1 (Rd ;Y )
m=1
Z M
X
6 C0 κ2,1 CT εm x m dP
Ω X
m=1
M
X
6 C0 κ2,1 CT εm x m .
L2 (Ω;X)
m=1
Remark 14.5.10.
(1) The method of proof for p = 1 in Theorem 14.5.9 could be extended
to p ∈ (1, 2] if Y has Fourier type p. We have not done this, because
Proposition 7.3.6 shows that having type p is weaker than having Fourier
type p.
(2) In the case p = 1 and d = 1, a variation of the argument in Proposition
8.5.7 actually gives a stronger result than Theorem 14.5.9, namely that if
T x ∈ W d,1 (Rd ; L (X, Y )) for all x ∈ X, then the range of T is R-bounded.
14.6 Triebel–Lizorkin spaces 365
kf kFp,q
s (Rd ;X) = (2ks ϕk ∗ f )k>0 Lp (Rd ;`q (X))
.
The obstruction just noted already makes itself felt if one tries to adapt the
proof that Besov spaces are independent up to an equivalent norm of the in-
homogeneous Littlewood–Paley sequence (ϕk )k>0 to Triebel–Lizorkin spaces.
The encountered difficulty will be resolved by a variant on the Fefferman–
Stein inequality due to Peetre, to which we turn in the present preliminary
subsection.
Throughout this section, unless otherwise stated X is an arbitrary Banach
space. For a strongly measurable function f : Rd → X and r ∈ (0, ∞) we let
the supremum being taken over all Euclidean balls B in Rd that contain x.
Lemma 14.6.1 (Peetre’s maximal inequality). Fix r, t ∈ (0, ∞) and a
multi-index α ∈ Nd , and let f ∈ S 0 (Rd ; X) satisfy
k∂ α f (x − z)k kf (x − z)k
sup t−|α| d/r
6 C1 sup d/r
6 C2 Mr f (x)
z∈Rd (1 + t|z|) z∈Rd (1 + t|z|)
k∂j f (x − z)k
Z
6 c λ (1 + |x − z − y|)−λ (1 + |z|)−d/r kf (y)k dy
(1 + |z|)d/r Rd
Z
6 cλ (1 + |x − z − y|)−d−1 (1 + |x − y|)−d/r kf (y)k dy
Rd
kf (x − y)k
6 C1 sup ,
y∈Rd (1 + |y|)d/r
where C1 = cλ Rd (1 + |y|)−d−1 dy. This gives the first inequality in the state-
R
ε Z 1/r
kg(0)k 6 sup k∇g(y)k + − kg(y)kr dy , (14.66)
2 y∈Qε Qε
1
R R
where we write −Q = |Q| Q
for averages. By scaling it suffices prove (14.66)
for ε = 1.
Fix g ∈ C 1 (Q1 ; X). For all y ∈ Q1 we have kyk 6 21 and
14.6 Triebel–Lizorkin spaces 367
Z 1
g(0) = g(y) + ∇g(ty) · y dt.
0
ε Z 1/r
kf (x − z)k 6 sup k∇f (x − z − y)k + − kf (x − z − y)kr dy .
2 y∈Qε Qε
(14.67)
where we used that (1 + |z|) > 21 (1 + |y + z|) for |y| 6 ε 6 1 and performed a
change of variables. Combining this estimate with the first inequality in the
statement of the lemma, and taking ε ∈ (0, 1] small enough, the result follows.
Step 4 – Next let f ∈ S 0 (Rd ; X) and t > 0. Let fδ = ψ(δ·)f , where
ψ ∈ S (Rd ) satisfies ψ(0) = 1, supp ψb ⊆ {ξ ∈ Rd : |ξ| 6 1} and δ ∈
(0, min{1, t}). Recalling that f ∈ C ∞ (Rd ; X), clearly we have fδ ∈ S (Rd ; X),
fbδ has support in B2t and therefore, by the previous steps, the second in-
equality in the statement of the lemma holds if in the two expressions on the
left-hand side f is replaced by fδ and for the right-hand side we note that
Mr fδ (x) 6 kψk∞ Mr f (x). It remains to let δ → 0 on the left-hand side and
note that fδ (x − z) → f (x − z) and similarly for its derivatives.
Using the pointwise estimate of Lemma 14.6.1, we will now deduce a maximal
inequality in Lp (Rd ; `q ).
compact set with diameter δk > 0. There exists a constant C > 0, depending
only on d, p, q, r, such that
kfk (· − z)k
sup 6 Ckf kLp (Rd ;`q (X)) .
z∈Rd (1 + δk |z|)d/r k>0 Lp (Rd ;`q )
∗ ∗
Proof. We use the short-hand notation f = (fk )k>0 and fd/r = (fk,d/r )k>0 ,
where
∗ kfk (x − z)k
fk,d/r (x) = sup , x ∈ Rd . (14.68)
z∈Rd (1 + δk |z|)d/r
kgk (x − z)k
sup 6 cMr gk (x).
z∈Rd (1 + |z|)d/r
∗ kfk (x − z)k
fk,d/r (x) = sup 6 cMr fk (x).
z∈Rd (1 + δk |z|)d/r
∗ 1/r
kfd/r kLp (Rd ;`q ) 6 ck(Mr fk )k>0 kLp (Rd ;`q ) = c (M (kfk kr ))k>0 Lp/r (Rd ;`q/r )
1/r
.p,q,r c (kfk kr )k>0 Lp/r (Rd ;`q/r )
= ckf kLp (Rd ;`q (X)) .
Theorem 14.6.3. Let X and Y be Banach spaces and let p ∈ [1, ∞), q ∈
[1, ∞], and r ∈ (0, min{p, q}). Let Sk ⊆ Rd , k > 0, be compact
b sets with
diameter δk > 0. Then for all sequences m = (mk )k>0 in L1 (Rd ; L (X, Y ))
and all f = (fk )k>0 ∈ Lp (Rd ; `q (X)) with supp fbk ⊆ Sk for each k > 0 we
have F −1 mF f ∈ Lp (Rd ; `q (Y )) and
= C sup k(1 + | · |)d/r )F −1 [mk (δk ·)]kL1 (Rd ;L (X,Y )) kf kLp (Rd ;`q (X))
k>0
kfk (y)k
Z
kKk ∗ fk (x)k 6 kKk (x − y)k(1 + δk |x − y|)d/r dy
Rd (1 + δk |x − y|)d/r
Z
∗ ∗
6 fn,d/r (x) kKk (x − y)k(1 + δk |x − y|)d/r dy 6 cm fn,d/r (x).
Rd
The required result follows from this by taking Lp (Rd ; `q )-norms and applying
Proposition 14.6.2.
The final identity of the theorem simply follows by a substitution together
with the dilation property δk−1 (F −1 mk )(δk−1 ·) = F −1 [mk (δk ·)] of the Fourier
transform.
and consequently kf k∞ .p0 ,ψ kf kp0 . Since we already knew the result for
p0 > 1, this inequality holds for p0 ∈ (0, ∞). In the remaining case p0 < p1 <
∞, we similarly find that
kf kp1 6 kf k1−p
∞
0 /p1
kf kpp00 /p1 .p0 ,p1 ,ψ kf kp0 .
We now introduce our main characters. Recall that we have fixed a inhomoge-
neous Littlewood–Paley sequence (ϕk )k>0 in Subsection 14.2.c (see Conven-
tion 14.2.8).
kf kFp,q
s (Rd ;X) := (2ks ϕk ∗ f )k>0 Lp (Rd ;`q (X))
is finite.
Fix an arbitrary r ∈ (0, min{p, q}), say r = rp,q = 21 min{p, q}. Applying
Theorem 14.6.3 with δk = 3 · 2k and mk = ψbk to (2ks f`+k )k>0 we obtain
(2ks ψk ∗ f )k>0 Lp (Rd ;`q (X))
6 Cψ,d,p,q,s (2ks f`+k )k>0 Lp (Rd ;`q (X))
0 ks
6 Cψ,d,p,q,s (2 ϕk ∗ f )k>0 Lp (Rd ;`q (X))
.
Since (ψk )k>0 and (ϕk )k>0 were arbitrary, this completes the proof.
The same argument and (14.5) lead to the following useful estimate.
s
Lemma 14.6.7. Let f ∈ Fp,q (Rd ; X), let (ψk )k>0 be a Littlewood–Paley se-
quence, and set
X n
Sn f := ψk ∗ f, n > 0.
k=0
s
Then Sn f ∈ Fp,q (Rd ; X) and there exists a constant C = C(p, q, d, ψ) such
that
kSn f kFp,q
s (Rd ;X) 6 Ckf kF s (Rd ;X) ,
p,q
n > 0.
We have the following analogue of Proposition 14.4.18 for Triebel–Lizorkin
spaces:
Proposition 14.6.8 (Sandwiching with Besov spaces). For all p ∈
[1, ∞), q ∈ [1, ∞], and s ∈ R, we have the natural continuous embeddings
S (Rd ; X) ,→ Fp,q
s
(Rd ; X) ,→ S 0 (Rd ; X),
the first of which is dense if p, q ∈ [1, ∞), and
s
Bp,p∧q (Rd ; X) ,→ Fp,q
s
(Rd ; X) ,→ Bp,p∨q
s
(Rd ; X).
372 14 Function spaces
By combining the first of these inclusions with Lemma 14.2.1 we see that if
p, q ∈ [1, ∞), then Cc∞ (Rd ) ⊗ X is dense in Fp,q
s
(Rd ; X).
s
Proof. First let p > q. For f ∈ Bp,q (Rd ; X) it follows from the triangle in-
equality in Lp/q (Rd ) that
X
kf kqF s (Rd ;X)
= 2ksq kϕk ∗ f kq
p,q Lp/q (Rd )
k>0
X q q
6 2ksq kϕk ∗ f kL p (Rd ;X) = kf kB s d ;X) .
p,q (R
k>0
This gives the first embedding in the second displayed line of the proposi-
s
tion. The second embedding follows from (14.69), which gives Fp,q (Rd ; X) ,→
s d s d
Fp,p (R ; X) = Bp,p (R ; X) continuously. The case p 6 q is handled similarly.
The continuous embeddings in the first line now follow from the corre-
sponding result for Besov spaces contained in Proposition 14.4.3.
Let us finally show that S (Rd ; X) is dense in Fp,q
s
(Rd ; X). The proof is
s
similar toPStep 3 of the proof of Proposition 14.4.3. Let f ∈ Fp,q (Rd ; X) and
n
set ζn := k=0 ϕk . By (14.6) we have kζn k1 6 kϕ0 k1 .
s
We will first show that ζn ∗ f → f in Fp,q (Rd ; X). Let ε > 0 and choose
K > 0 such that
X 1/q
2ksq kϕk ∗ f kq < ε.
Lp (Rd )
k>K
By Young’s inequality,
It follows that
kf − ζn ∗ f kFp,q
s (Rd ;X)
K
X 1/q
6 ε(1 + kϕ0 k1 ) + 2ksq kϕk ∗ f − ζn ∗ ϕk ∗ f kq
Lp (Rd )
k=0
K
X
6 ε(1 + kϕ0 k1 ) + 2ks kϕk ∗ f − ζn ∗ ϕk ∗ f kLp (Rd ;X)
k=0
Proposition 14.6.9. For p ∈ [1, ∞), q ∈ [1, ∞], and s ∈ R, the space
s
Fp,q (Rd ; X) is a Banach space.
Proof. As in the Besov case one proves that for all p ∈ [1, ∞), q ∈ [1, ∞],
s
and s ∈ R, the space Fp,q (Rd ; X) has the Fatou property. Since Triebel–
Lizorkin spaces embed into S 0 (Rd ; X) by Proposition 14.6.8, the completeness
s
of Fp,q (Rd ; X) follows from Lemma 14.4.7.
The main result of this subsection is a version of the Mihlin multiplier theorem
for Triebel–Lizorkin spaces. Before we state it we first prove an important
lifting property as we saw in Proposition 14.4.15 for Besov spaces.
Proposition 14.6.10 (Lifting). Let p ∈ [1, ∞), q ∈ [1, ∞], and s ∈ R. Then
for all σ ∈ R,
s
Jσ : Fp,q s−σ
(Rd ; X) ' Fp,q (Rd ; X) isomorphically. (14.71)
s
Proof. As in Proposition 14.4.15 it suffices to show that Jσ maps Fp,q (Rd ; X)
s−σ d
into Fp,q (R ; X) and is bounded for each σ ∈ R We must show that
(2n(s−σ) ϕn ∗ Jσ f )n>0 belongs to Lp (Rd ; `q (X)). This will be done by ap-
plying the multiplier Theorem 14.6.3 to a multiplier m = (mn )n>0 naturally
associated with Jσ .
Write
X 1
2−nσ ϕn ∗ Jσ f = F −1 mn ϕ bn+` fb,
`=−1
where
mn (ξ) = 2−nσ (1 + 4π 2 |ξ|2 )σ/2 ϕ
bn (ξ).
We have mn ∈ C ∞ (Rd ) and, putting δn = 3 · 2n ,
1 1
bn (δn ·) ⊆ ξ ∈ Rd : 6 |ξ| 6
supp ϕ , (n > 1)
6 2
1
b0 (δ0 ·) ⊆ ξ ∈ Rd : |ξ| 6
supp ϕ .
2
Lemma 14.2.12, applied with λ = d + 1 + dd/re with an arbitrary r = rp,q ∈
(0, min{p, q}), gives the estimate
We continue with the Mihlin multiplier theorem for Triebel–Lizorkin spaces.
Note that the Besov space case was considered in Theorems 14.4.16 and 14.5.6.
s
then there is a bounded operator T : Fp,q (Rd ; X) → Fp,q
s
(Rd ; Y ) with kT k 6
Cd,p,q,s,X,Y Km such that T f = F (mfb) for all f ∈ S (Rd ) ⊗ X.
−1
Note that in the case q < ∞, one has that T is the unique bounded extension
of Tm : S (Rd ) ⊗ X → S 0 (Rd ; Y ).
Proof. We define T in the same was as in (14.59) of the proof of Theorem
14.5.6:
1 X
X
Tf = Tmk+` fk ,
`=−1 k>0
s
where f ∈ Fp,q (Rd ; X), fk = ϕk ∗ f and mk = ϕ s
bk m. Since Fp,q (Rd ; X) ⊆
s
Bp,∞ (Rd ; X) it follows from the proof of Theorem 14.5.6 that the above series
converges in S 0 (Rd ; Y ), and that T g = F −1 (mb
g ) for all g ∈ S (Rd ) ⊗ X.
To prove the required boundedness, note that
1
X
kTm f kFp,q
s (Rd ;Y ) 6 2ks F −1 (mϕ
bk+` ϕ
bk fb)k>0 Lp (Rd ;`q (Y ))
.
`=−1
Proposition 14.6.12. Let p ∈ [1, ∞), q ∈ [1, ∞], and s ∈ R. For all k ∈ N
the expression
X
|||f |||Fp,q
s (Rd ;X) := k∂ α f kFp,q
s−k
(Rd ;X)
|α|6k
s
defines an equivalent norm on Fp,q (Rd ; X).
S (Rd ; X) ,→ Fp,q
s
(Rd ; X) ,→ S 0 (Rd ; X)
and
s
Bp,p∧q (Rd ; X) ,→ Fp,q
s
(Rd ; X) ,→ Bp,p∨q
s
(Rd ; X)
for s ∈ R, p ∈ [1, ∞) and q ∈ [1, ∞]. Moreover, for any q ∈ [1, ∞], it is
immediate from the definitions that
s
Bp,1 (Rd ; X) ,→ Fp,q
s s
(Rd ; X) ,→ Bp,∞ (Rd ; X). (14.72)
The next result compares Triebel–Lizorkin spaces with the Bessel potential
and Sobolev spaces. It can be improved if X is UMD and has type and cotype
properties (see Proposition 14.7.6 below).
Proof. For (14.73) and (14.74), by Propositions 5.6.3, 14.6.10 and 14.6.12
it suffices to consider the special case s = m = 0, for which H 0,p (Rd ; X) =
W 0,p (Rd ; X) = Lp (Rd ; X). It thus remains to show the continuous embeddings
0
Fp,1 (Rd ; X) ,→ Lp (Rd ; X) ,→ Fp,∞
0
(Rd ; X). (14.75)
P The first embedding in (14.75) is true for any p ∈ [1, ∞): writing f =
k>0 ϕk ∗ f it follows that
X
kf kLp (Rd ;X) 6 kϕk ∗ f kLp (Rd ;X) = kf kFp,1
0 (Rd ;X) .
k>0
For the second embedding in (14.75) observe that since ϕ ∈ S (Rd ), it has a
radially decreasing majorant which is integrable. Therefore, by Theorem 2.3.8
there is a constant Cd > 0 such that for all k > 0 and almost all x ∈ Rd ,
kϕk ∗ f (x)k 6 Cd M f (x). Therefore, by the Lp -boundedness of the Hardy–
Littlewood maximal function (Theorem 2.3.2),
kf kFp,∞
0 (Rd ;X) = sup kϕk ∗ f k Lp (Rd )
6 Cd kM f kLp (Rd ) .p Cd kf kLp (Rd ;X) .
k>0
for all sequences of scalars (ak )k>0 for which the expression on the right-hand
side is finite.
To prove the desired estimate, by (14.69) it suffices to consider the case
q0 = q1 = ∞. Taking ak (x) = kϕk ∗ f (x)k with x ∈ Rd in (14.76), raising to
the power p and integrating over Rd , by Hölder’s inequality (with exponents
p0 p1
(1−θ)p and θp ) we obtain
1−θ
kf kFp,q
s (Rd ;X) 6 Cs0 ,s1 ,s kf k s0
F d ;X) kf kθFps1,∞ (Rd ;X)
p0 ,∞ (R 1
as required.
In a similar way one can prove the following variant for the end-point p1 = ∞.
Proof of sufficiency in Theorem 14.6.14. For the sufficiency of (i) first assume
that p0 = p1 , q0 6 q1 , and s0 > s1 . Under these assumptions the result follows
from the fact that
(2ks1 ak )k>0 `q 1
6 (2ks0 ak )k>0 `q 0
.
If p0 = p1 , q0 > q1 , and s0 > s1 , the result follows from (14.23) and (14.72):
kf kFr,r
t (Rd ;X) = kf kB t (Rd ;X) 6 Ckf k s1
r,r Bp ,p (Rd ;X)
1 1
where in the last step we used (14.69). Substituting the latter estimate into
(14.77), we obtain
1/(1−θ)
kf kFps1,1 (Rd ;X) 6 Cs0 ,s1 ,θ C θ/(1−θ) kf kFps0,q (Rd ;X) . (14.78)
1 0 0
Now if q0 < ∞, then the result follows from the density of S (Rd ; X) in
Fps00,q0 (Rd ; X).
Pn
If q0 = ∞ and f ∈ Fps00,∞ (Rd ; X), we let Sn f = k=0 ϕk ∗ f . Then by
Young’s inequality and the fact that ϕj ∗Sn f = 0 for j > n+1, we have Sn f ∈
Bps00 ,1 (Rd ; X). Thus Theorem 14.4.19 implies Sn f ∈ Bps11 ,1 (Rd ; X). More-
over, by Proposition 14.6.8 and (14.69) we also have Sn f ∈ Fps00,1 (Rd ; X) ,→
Fps00,∞ (Rd ; X) and Sn f ∈ Fps11,1 (Rd ; X). Therefore, by (14.78),
kf − fn kF 1/p (Rd ) 6 Ckf − fn kF 1/r (Rd ) = Ckf − fn kB 1/r (Rd ) , r ∈ (1, p),
p,q r,r r,r
(2) If X has cotype q0 , then for all q ∈ (q0 , ∞) and all r ∈ [1, ∞] we have a
continuous embedding
( 1 − 12 )d
γ(L2 (Rd ), X) ,→ Fq,rq (Rd ; X).
Proof. We give the proof of (1), the proof of (2) being similar. Let 1 6 p < p0 .
Let s0 = ( p10 − 12 )d and s = ( p1 − 12 )d. By Theorem 14.6.14 we have a continuous
embedding
s
Fp,r (Rd ; X) ,→ Fps00,p0 (Rd ; X) = Bps00 ,p0 (Rd ; X).
Now the result follows from Theorem 14.5.1.
380 14 Function spaces
∆h f (x) = f (x + h) − f (x)
and ∆m m
h = (∆h ) .
(m,τ )
It will be shown shortly that each of the norms ||| · |||F s d with m > s and
p,q (R ;X)
d d s
s> min{p,q} − defines an equivalent norm on
τ Fp,q (Rd ; X).
The expression for the seminorm simplifies for τ = q ∈ [1, ∞). Indeed, by
Fubini’s theorem we have
1 Z 1/q
(m,q) −(s+d)q m q
[f ]F s (Rd ;X) = |h| k∆ h f (x)k dh .
p,q (sq + d)1/q |B1 | Rd Lp (Rd )
Note that the condition (14.79) holds trivially holds if τ 6 min{p, q}, and in
particular if τ = 1. The condition (14.79) is only used in the proof of “&” of
(14.80).
(m,τ )
For the proof we will use a discretised version of |||f |||F s (Rd ;X) . Put
p,q
14.6 Triebel–Lizorkin spaces 381
Z 1/τ
m,τ
J (f, k)(x) := − k∆2m−k h f (x)kτ dh .
|h|61
As in (14.38) we have
(m,τ )
[f ]F s d hd,s (2ks J m,τ (f, k))k∈Z Lp (Rd ;`q (Z))
.
p,q (R ;X)
Mr := (M (kf kr )(x))1/r
introduced in (14.65).
k∆m
h f (x)k .d,m,r (t|h|)
d/r
Mr (f )(x) if |h| > t−1 ; (14.82)
m m
k∆h f (x)k .d,m,r (t|h|) Mr (f )(x) if |h| 6 t−1 . (14.83)
The estimate (14.82) follows from (14.84) and Lemma 14.4.22, for if |h| > t−1 ,
then
m
m
X m
k∆h f (x)k 6 kf (x + hj)k
j=0
j
To prove (14.83) fix |h| 6 t−1 . Set φ(s) := f (x + sh) for s ∈ R. Then
∆hm f (x) = ∆m 1
1 φ(0). Since for any g ∈ C (R; X) we have k∆1 g(s)k 6
0
supθ∈[s,s+1] kg (θ)k, an induction argument gives
In particular,
X 1/2
k∆hm f (x)k 6 sup kφ(m) (θ)k 6 |h|m sup k∂ α f (x + θh)k2 .
θ∈[0,m] θ∈[0,m]
|α|=m
Substituting this into the previous estimate gives the required estimate.
Proof of Theorem 14.6.20. It remains to prove the inequality & in (14.81).
To begin with, from (i) we have inequality
∞
X
kf kLp (Rd ;X) 6 kfj kX 0 (Rd ;X) .d,p,q,s kf kF s (Rd ;X) ,
= kf kFp,1 p,q
Lp (Rd )
j=0
where fj = ϕj ∗ f as always.
To deal with the seminorm, note that from the assumption (14.79) it fol-
lows that we can find r ∈ (0, ∞) and λ ∈ (0, 1] such that
and therefore
.d,m,s kf kFp,q
s (Rd ;X) .
=: T1 (x) × T2 (x).
14.6.f Interpolation
by
X
R(fk )k>0 = ψk ∗ fk , Sf = (ϕk ∗ f )k>0 .
k>0
Our next aim is an interpolation result which will be used improve the Sobolev
embedding result of Theorems 14.4.19 and 14.6.14.
(Lp0 (Rd ; `qws (X)), Lp1 (Rd ; `qws (X)))θ,p = Lp (Rd ; `qws (X))
which follows from Theorem 2.2.10 and Proposition 14.3.5. The case q = 1
can be deduced from the proof of Theorem 14.4.31 as well. Indeed, since the
operator S of Lemma 14.6.22 is an isometry also for q = 1, we find
kf kFp,1
s (Rd ;X) = kSf kLp (Rd ;`1
w (X))
s
hp,p0 ,p1 ,θ kSf k(Lp0 (Rd ;`1w (X)),Lp1 (Rd ;`1ws (X)))θ,p
s
As an application we can prove some further embedding results.
386 14 Function spaces
Since the embedding Fps00,p0 (Rd ; X) ,→ Bps00 ,p1 (Rd ; X) holds trivially, (14.86)
improves the embedding in Theorem 14.6.14. In a similar way one sees that
(14.87) is an improvement of Theorem 14.6.14. Consequently, it follows from
Theorem 14.6.14 that, under the assumption p0 < p1 , the condition s0 − pd0 >
s1 − pd1 is also necessary for both (14.86) and (14.87).
d d d d
t0 − = s0 − and t1 − = s0 − .
p1 r0 p1 r1
As an interesting consequence of Theorem 14.6.26 we have the following im-
provement of Corollary 14.4.27 (2), extending it to the case p0 = 1. The result
14.6 Triebel–Lizorkin spaces 387
d
is false for integrability exponents p0 > 1. Indeed, if s − p0 > 0 and it would
s− pd
hold that Fps0 ,q (Rd ) ,→ Cub (Rd ) for q = ∞, then it would also hold for
0
all q ∈ [1, ∞). However, by Proposition 14.6.17 this would imply that every
function in Fps0 ,q (Rd ) is zero at x1 = 0, which is of course not true.
s s−d
Corollary 14.6.27. If s > d is an integer, then F1,∞ (Rd ; X) ,→ Cub (Rd ; X)
continuously.
The result also holds in the case where s > d is not integer. However, in this
case Corollary 14.4.27 (2) gives a better result.
Proof. By Theorem 14.6.26 and Proposition 14.4.18,
s s−d s−d
F1,∞ (Rd ; X) ,→ B∞,1 (Rd ; X) ,→ Cub (Rd ; X).
14.6.g Duality
isomorphically.
The proof is similar to that of Theorem 14.4.34. The restriction p, q > 1 comes
in through Lemma 14.6.22.
s s
14.6.h Pointwise multiplication by 1R+ in Bp,q and Fp,q
implies
Therefore,
Z σ Z σ Z σ Z t
ds ds
F (s) = f (s) − I = − f (r) dr − − f (r) dr
t s t s 0 0
Z t (14.88)
= f (σ) − F (σ) − − f (r) dr.
0
e := α − 1 > −1 to
Applying Hardy’s inequality (see Lemma L.3.2(1)) with α
the function s 7→ kF (s)k we obtain
Now the proof is finished as before, this time applying Lemma L.3.2(2) with
e := α − 1 < −1.
α
As an immediate consequence we obtain the following result.
14.6 Triebel–Lizorkin spaces 389
1
with C := 1 + 1 ,
|β− p |
provided the right-hand side is finite.
Here we use the bounded continuous version for f (which exists by Corollary
14.4.27 combined with Propositions 14.6.8 and 14.6.13) respectively. The con-
tinuity of the embeddings in Corollary 14.4.27 gives the closedness of these
subspaces.
We can now prove the following fractional Hardy inequality in terms of
s
the spaces Fp,q and H s,p and their analogues 0 Fp,q
s
and 0 H s,p .
s
Since W s,p (R; X) = Fp,p (R; X) for s ∈ (0, 1), the corollary also covers frac-
tional Sobolev spaces.
Proof. By the embeddings (14.69) and (14.73) it suffices to prove the result
s s
for 0 Fp,∞ (R; X) and Fp,∞ (R; X).
By Proposition Rt 14.6.30, using that for bounded continuousR functions f :
t
R → X we have −0 f (τ ) dτ → f (0) = 0 as t ↓ 0 in case (1) and −0 f (τ ) dτ → 0
as t → ∞ in case (2), we have
Z
k1R+ f kLp (R,|t|−sp dt;X) 6 C x 7→ x−s − kf (x) − f (x − h)k dh p
(0,x) L (R+ )
Z
6 2C x 7→ sup t−s − k∆h f (x)k dh p
t>0 (−t,t) L (R)
(1)
= 2C[f ]F s (R;X) .p,s kf kFp,∞
s (R;X)
p,∞
where in the last step we used Theorem 14.6.20 with m = 1. A similar estimate
holds for f on the negative real axis.
As a consequence we obtain the following result on pointwise multiplication.
Theorem 14.6.32 (Pointwise multiplication by 1R+ ). Let p ∈ [1, ∞),
q ∈ [1, ∞], and s ∈ (0, 1). Each of the two conditions
s
(1) s ∈ (0, 1/p) and f ∈ Fp,q (R; X)
s
(2) s ∈ (1/p, 1) and f ∈ 0 Fp,q (R; X)
s
implies that 1R+ f ∈ Fp,q (R; X) and
Without the condition f (0) = 0, the result is false for s > 1/p. Indeed, this
is clear from the fact that, by combining Corollary 14.4.27 and Proposition
s
14.6.13, we have a continuous embedding Fp,q (R; X) ,→ Cub (R; X). A coun-
terexample to the case s = 1/p will be discussed in Example 14.6.33. It shows
that Propositions 14.6.29, 14.6.30, and Corollary 14.6.31 do not hold for α = 0
and s = 1/p.
Proof. Clearly, k1R+ f kLp (Rd ;X) 6 kf kLp (Rd ;X) . Therefore, using the difference
(1)
norm of Theorem 14.6.20 it remains to estimate [1R+ f ]F s in terms of
p,q (R;X)
(1)
kf kFp,q Fp,q (R;X) . We give the proof for q ∈ [1, ∞); the case q = ∞
s (R;X) and [f ] s
where Kq,s = (sq + q)1/q . Setting z = −x and φp (z) = z 1/p (1 + z)−s−1 , (III)
can be estimated using Young’s inequality for convolutions for the multiplica-
tive group R+ with Haar measure dz z :
Z Z p 1/p
(III) 6 Kq,s (y + z)−s−1 kf (y)k dy dz
R+ R+
Z Z 1 dy p dz 1/p
= Kq,s φp (z/y)y −s+ p kf (y)k
R+ R+ y z
Z 1/p
6 Kq,s kφp kL1 (R+ , dz ) y −sp kf (y)kp dy
z
R+
.p,q,s kf kFp,q
s (R;X) ,
Example 14.6.33. Theorem 14.6.32 is false for s = 1/p even in the scalar-
valued case. Indeed, f ∈ Cc∞ (R) is any function satisfying f ≡ 1 on [−1, 1],
1/p
then for all p ∈ [1, ∞) we have f ∈ Fp,q (R). Let us prove that 1R+ f ∈ /
1/p
Fp,q (R). To this end it suffices to take q = ∞. In case p ∈ (1, ∞) we can use
Theorem 14.6.20 to find
(1)
k1R+ f kF 1/p (Rd ;X) hp k1R+ f k 1/p
p,∞ Fp,∞ (Rd ;X)
Z −x
1
> x 7→ sup t− p −1 |f (x)| dh
t>0 −t Lp (0,1)
1
= x 7→ sup t− p −1 (t − x)
t>x Lp (0,1)
1
−p
&p kx 7→ x kLp (0,1) = ∞.
1 1/r
For p = 1 we note that F1,q (R) ,→ Fr,∞ (R) for all r ∈ (p, ∞) by Theorem
1
14.6.14, and therefore 1R+ f ∈ / F1,q (R).
One could still hope that the boundedness of f 7→ 1R+ f for s = 1/p holds
1/p
on the closure in Fp,q (R) of the smooth functions satisfying f (0) = 0. This
turns out to be false as well. Indeed, in the case q < ∞ the latter space
1/p
coincides with Fp,q (R) by Proposition 14.6.17. If q = ∞, the boundedness is
1/p
also fails, as follows from the previous example and the embedding Fp,∞ (R) ,→
1/r
Fr,r (R) for all r ∈ (p, ∞) contained in Theorem 14.6.14.
6 C 0 kf kFp,q
s (R;X) kgk −s
F 0 0 (R;X ∗ ) .
p ,q
s−ε s+ε s
Proof. First let s > 0. Since (Fp,2 , Fp,2 )1/2,q = Bp,q by Theorem 14.4.31,
the result follows from Theorems 14.6.32 and C.3.3. Here we can allow p = 1
as well.
The result for s < 0 and q ∈ (1, ∞) follows from Theorem 14.4.34 in the
same way as in Corollary 14.6.34. The cases q = 1 and q = ∞ can be obtained
by another real interpolation argument as we did in Example 14.4.35.
The case s = 0 follows by real interpolation between the cases s and −s
for s > 0 small.
In this section we prove Sobolev embeddings and norm estimates for Bessel
potential spaces. Some results will depend on the geometry of X. Real in-
terpolation for H s,p (Rd ; X) has already been considered in Theorem 14.4.31.
Duality for H s,p (Rd ; X) has already been considered in Proposition 5.6.7.
d d
p0 < p 1 and s0 − > s1 − . (14.90)
p0 p1
If s0 , s1 ∈ N, then the same necessary and sufficient conditions give the exis-
tence of a continuous embedding
From Theorem 14.6.14 we see that if either (14.89) or (14.90) holds, then
Fps00,∞ (Rd ; X) ,→ Fps11,1 (Rd ; X). Therefore the required embedding follows from
(14.91) with s = s0 , s1 and p = p0 , p1 .
‘Only if’: If the stated embedding holds, then by (14.91) with s = s0 , s1
and p = p0 , p1 , we also have a continuous embedding Fps00,1 (Rd ; X) ,→
Fps11,∞ (Rd ; X). Therefore, either (14.89) or (14.90) must hold by Theorem
14.6.14.
The corresponding result for Sobolev spaces with integer smoothness can
be proved in the same way, noting that the analogue of (14.91) holds for these
spaces.
Remark 14.7.2. The embedding of Theorem 14.7.1 for Bessel potential spaces
can be restated as the boundedness of J−(s0 −s1 ) = (1 − ∆)−(s0 −s1 ) from
Lp0 (Rd ; X) into Lp1 (Rd ; X). Since J−(s0 −s1 ) is a positive operator by Propo-
sition 5.6.6, we infer from Theorem 2.1.3 that the boundedness in the scalar
case is actually equivalent to boundedness in the vector-valued situation.
If, in Proposition 14.7.3, s0 , s1 > 0 are integers and p ∈ (1, ∞), the same argu-
ment gives that f ∈ W s0 ,p0 (Rd ; X) ∩ W s1 ,p1 (Rd ; X) implies f ∈ W s,p (Rd ; X)
and
The latter estimate extends to p0 ∈ (1, ∞] and p1 ∈ (1, ∞]. Indeed, if only
one of the exponents is infinite, then (14.92) is a consequence of Proposition
14.6.16 and the sandwich results of Propositions 14.4.18 (see (14.29)) and
14.7 Bessel potential spaces 395
d
Now divide both sides by λ|α|− p and pass to the limit λ → ∞.
From the location of the supports of the functions ϕbk one sees three things:
first, that for each ξ ∈ Rd at most three terms in this sum are non-zero
(the sum therefore converges for trivial reasons); second, that k∂ β ϕ
bk k∞ 6
Cβ 2−k|β| ; and third, that
is finite, the outer supremum being taken over all sequences of signs =
(k )k>0 .
By the Mihlin multiplier theorem (Theorem 5.5.10), the Fourier multiplier
operators Tmn associated with mn are bounded on Lp (Rd ; X), with estimates
uniform in n and signs , say sup supn>0 kTmn kL (Lp (Rd ;X)) 6 CX,p,d . Since
n
X
k 2ks ϕk ∗ f = Tmn Js f, (14.93)
k=0
we obtain
n
X
k 2ks ϕk ∗ f 6 CX,p,d kJs f kLp (Rd ;X) = CX,p,d kf kH s,p (Rd ;X) .
Lp (Rd ;X)
k=0
14.7 Bessel potential spaces 397
‘If’: Assume now that f ∈ S 0 (Rd ; X) satisfies |||f |||H s,p (Rd ;X) < ∞. We
claim that k>0 εk 2ks ϕk ∗ f converges in Lp (Ω; Lp (Rd ; X)) and almost surely
P
For k ∈ {0, 1} choose ψk ∈ Cc∞ (R) such that 0 6 ψbk 6 1, supp ψb0 ⊆ {0 6
|ξ| 6 2} and supp ψb1 ⊆ { 14 6 |ξ| 6 4}, and ψbk ≡ 1 on supp ϕ bk . For k > 2 we
−(k−1)
define ψk := ψ1 (2
b b ·). For ω ∈ Ω put
X X
mω := εj (ω)2−js (1 + | · |2 )s/2 ψbj , gω := εk (ω)2ks ϕk ∗ f.
j>0 k>0
As before,
for almost every ω ∈ Ω. Considering finite sums first, one checks that ω 7→
Tmω gω is strongly measurable. Since ω 7→ gω belongs to Lp (Ω; Lp (Rd ; X)), it
follows that so does ω 7→ Tmω gω . By the condition ψbk ≡ 1 on supp ϕ
bk , as in
(14.93) we have Z
Tmω gω dP(ω) = Js f.
Ω
We continue with an embedding result under additional geometric assump-
tions on X. The cases p0 = 1 and q0 = ∞ were proved for general Banach
spaces in Propositions 14.4.18 and 14.6.13.
In combination with Proposition 14.6.13 and Corollary 14.6.18 we obtain:
(2) If X has cotype q0 , then for all q ∈ (q0 , ∞) we have a continuous embedding
1 1
γ(L2 (Rd ), X) ,→ H ( q − 2 )d,q (Rd ; X)
By Theorem 9.2.10, for p0 = 2 assertion (1) also holds for p = 2, and for
q0 = 2 assertion (2) also holds for q = 2.
14.7 Bessel potential spaces 399
Proposition 14.7.8. Let p ∈ (1, ∞), q ∈ [1, ∞], s ∈ R, and m ∈ N. Then the
following assertions hold with A ∈ {B, F }:
(1) If Asp,q (Rd ; X) ,→ H s,p (Rd ; X) continuously, then X has type q.
(2) If H s,p (Rd ; X) ,→ Asp,q (Rd ; X) continuously, then X has cotype q.
(3) If Akp,q (Rd ; X) ,→ W m,p (Rd ; X) continuously, then X has type q.
(4) If m
W m,p (Rd ; X) ,→ Ap,q (Rd ; X) continuously, then X has cotype q.
Proof. (1): By the lifting properties of Propositions 14.4.15, 14.6.10, and
5.6.3, it suffices to consider s = 0. Fix a finitely non-zero sequence (xn )n>1
in X. Let ψ ∈ S (Rd ) be a non-zero function satisfying supp(ψ) b ⊆ [− 1 , − 1 ]d
4 8
and put X n
f (t, ω) := ψ(t) εn (ω)e2πi2 t1 xn ,
n>1
n
where as always (εn )n>1 is a Rademacher sequence. Since (εn e2πi2 t1
)n>1 is
a Rademacher sequence for each t ∈ Rd , we have
Z p
p
X n
Ekf kLp (Rd ;X) = |ψ(t)|p E εn e2πi2 t1 xn dt
Rd n>1
X p (14.94)
= kψkpLp (Rd ) E εn x n .
n>1
kf (·, ω)kAsp,q (Rd ;X) = kψkLp (Rd ) k(xn )n>1 k`q (X) . (14.95)
k(∂ β ψ)fm−j (·, ω)kLp (Rd ;X) 6 k∂ β ψk∞ kfm−j (·, ω)kLp (Rd ;X)
X
6 k∂ β ψk∞ 2−(m−j)n kxn k 6 k∂ β ψk∞ sup kxn k.
n>1
n>1
By the reverse triangle inequality, this shows that there exists a constant
C = C(d, m, p, ψ) such that
X
kf kLp (Ω;W m,p (Rd ;X)) − kψkLp (R) εn xn Lp (Ω;X) 6 C sup kxn k. (14.96)
n>1
n>1
. kf kLp (Ω;Am d
p,q (R ;X))
+ sup kxn k
n>1
(4): This can be proved in the same way as (3). By (14.96) and
P the Kahane
contraction principle, which implies bound supn>1 kxn kp 6 Ek n>1 εn xn kp ,
from the assumption (4) we obtain
X
. kf kLp (Ω;W m,p (Rd ;X)) . εn x n .
Lp (Ω;X)
n>1
Proof. (1)⇒(3) and (2)⇒(3): By Proposition 14.7.8, X has type 2 and cotype
2. Therefore X is isomorphic to a Hilbert space by Theorem 7.3.1.
(3)⇒(2): This is immediate from Proposition 14.7.6 and the fact that
Hilbert spaces are UMD (by Theorem 4.2.14) and have type 2 and cotype 2
(by the result of Example 7.1.2).
(3)⇒(1): This is a special case of the previous implication since Theorem
5.6.11 implies W m,p (Rd ; X) = H m,p (Rd ; X) with equivalent norms.
14.7.c Interpolation
The spaces εp (X) := ε0,p (X) have been introduced in Section 6.3. Clearly
the mapping (xk )k>0 7→ (2ks xk )k>0 defines an isometric isomorphism from
εs,p (X) onto εp (X). For fixed s ∈ R the spaces εs,p (X), 1 < p < ∞, coincide,
with pairwise equivalent norms; this follows from the Kahane–Khintchine in-
equalities as in Proposition 6.3.1. If X does not contain a copy isomorphic
to
P c0 , then Corollary 6.4.12 implies that for any (xk )k>0 in εs,p (X) the sum
ks p
k>0 εk 2 xk converges in L (Ω; X) and almost surely in X, and in this case
X
(xk )k>0 εs,p (X)
= εk 2ks xk .
Lp (Ω;X)
k>0
In particular, the partial sum projections Pn : (xk )k>0 7→ (xk )nk=0 are uni-
formly bounded and strongly convergent to the identity as operators on
εs,p (X).
The next result extends Theorem 7.4.16, which corresponds to the special
case s = 0.
402 14 Function spaces
Lemma 14.7.11. Let X be a UMD space and let p ∈ [1, ∞], q ∈ [1, ∞], and
s ∈ R. For k > 0 set ψk = ϕk−1 + ϕk + ϕk+1 . The operators
defined by X
R(fk )k>0 = ψk ∗ fk , Sf = (ϕk ∗ f )k>0 ,
k>0
As in Examples 14.4.33 and 14.4.35, we can use this corollary to prove bound-
edness of pointwise multiplication by smooth functions:
f 7→ ζf
defines a bounded mapping from H s,p (Rd ; X) into H s,p (Rd ; Y ) of norm .k,s
kζkCbk (Rd ;L (X,Y )) .
Indeed, the pointwise multiplier f 7→ ζf is bounded as a mapping from
W j,p (Rd ; X) into W j,p (Rd ; Y ) for each j ∈ {0, . . . , k}. Therefore, for s ∈ N the
result is immediate from Theorem 5.6.11. If −s ∈ N, then the result follows by
the duality result of Proposition 5.6.7 and Theorem 5.6.11. If s ∈ (0, ∞), then
the result follows by interpolation between the cases j = 0 and j = k by the
complex method [·, ·] ks and applying Theorem C.2.6 and Corollary 14.7.13.
Finally, the case s ∈ (−∞, 0) follows by duality again.
The UMD property of X will only be used through the following proposition.
Proposition 14.7.16. Let p ∈ (1, ∞) and s > 0, and let X be a UMD space.
(1) The operator (−∆)s : S (Rd ; X) → S 0 (Rd ; X) given by
(−∆)s f = |2π · |s fb
kf kH s,p (Rd ;X) hp,X kf kLp (Rd ;X) + k(−∆)s/2 f kLp (Rd ;X) .
14.7 Bessel potential spaces 405
|2πξ|s
Proof. (1): Let m1 (ξ) = (1+|2πξ|2 )s/2
. Using Mihlin’s multiplier Theorem
5.5.10 one can check that m1 ∈ ML (Rd ; X, Y ). Therefore,
p
k(−∆)s f kp = kTm1 Js f kp 6 km1 kMLp (Rd ;X,Y ) kJs f kp 6 Cp,X kf kH s,p (Rd ;X) .
(2): Note that since s > 0, Proposition 5.6.6 gives that H s,p (Rd ; X) ,→
L (Rd ; X) contractively. This combined with (1) gives the estimate “&”.
p
(1+|2πξ|2 )s/2
The estimate . follows similarly. Let m2 (ξ) = 1+|2πξ|s . Then m2 ∈
MLp (Rd ; X, Y ) as before. Therefore,
We need two more preparatory results. The first one is a concrete formula for
(−∆)s/2 f as an integral operator.
Lemma 14.7.17. Let s ∈ (0, 1). For f ∈ S (R; X) we have
f (· + h) − f (·)
Z
(−∆)s/2 f = cs dh, x ∈ R,
R |h|1+s
where the integral on the right-hand side converges absolutely pointwise R, and
as a Bochner integral in Lp (R; X) for any p ∈ [1, ∞). Here cs ∈ R \ {0} is a
constant only depending on s.
Proof. The convergence of the integral for |h| > 1 is immediate. The conver-
R1
gence for |h| < 1 follows by writing f (x + h) − f (x) = 0 f 0 (x + th)h dt.
To prove the stated identity we take Fourier transforms on the right-hand
side and use Fubini’s theorem to obtain
Z 2πihξ
f (· + h) − f (·) −1 b
Z
e
F 1+s
dh dx = 1+s
f (ξ) dh = ks |ξ|s fb(ξ),
R |h| R |h|
where from the fact that the odd part of the integral cancels we see that
ks = 2 R+ cos(2πt)−1
R
t1+s dt is in (−∞, 0). This proves the result with constant
cs = ks−1 (2π)s .
|f (y)|
Z
x 7→ dy p 6 Cp kf kLp (R+ ) .
R+ x + y L (R+ )
406 14 Function spaces
1/p
Proof. Letting ζp (y) = xx+1 , after rewriting the integral, we can use Young’s
inequality for the multiplicative group R+ with Haar measure dx x to obtain
|f (y)|
Z Z Z
dy p dx 1/p
x 7→ dy p = ζp (x/y)y 1/p f (y)
R+ x + y L (R+ ) R+ R+ y x
6 kζp kL1 (R+ , dx ) kf kLp (R+ ) .
x
Proof of Theorem 14.7.15. By Proposition 14.6.17 it suffices to prove the de-
sired estimate for f in the dense class Cc∞ (R \ {0}) ⊗ X. In that case one
actually has g := 1R+ f is in the same class and thus is smooth as well.
We claim that
As soon as we proved the claim, then the result follows. Indeed, applying
Proposition 14.7.16 twice we obtain
To prove (14.98) we only consider the part Lp (R+ ) as the other one is similar.
By elementary considerations
Z ∞ Z Z ∞ Z −x
kf (x + h)k p kf (x + h)k p
1S (x, h) dh dx = dh dx
0 R |h|1+s 0 −∞ |h|1+s
14.8 Notes 407
∞ Z ∞
kf (−y)k
Z p
= dh dy
0 0 (y + h)1+s
∞ ∞
y −s kf (−y)k p
Z Z
6 dh dy
0 0 y+h
(i)
6 Cpp ky 7→ |y|−s f (y)kpLp (R;X)
(ii)
6 Cpp Cp,s
p
kf kpH s,p (R;X) ,
14.8 Notes
Early influential monographs on function spaces are those of Adams [1975] (see
also Adams and Fournier [2003]), Bergh and Löfström [1976], Peetre [1976],
and Triebel [1978]. After these works appeared, a new maximal function argu-
ment was discovered by Peetre [1975] which made it possible to study Besov
and Triebel–Lizorkin spaces in the full range p, q ∈ (0, ∞]. This theory is pre-
sented in detail in the monograph of Triebel [1983] and the more recent works
of Triebel [1992, 2006, 2020, 2013, 2014]; further expositions are due to Ba-
houri, Chemin, and Danchin [2011], Denk and Kaip [2013], Grafakos [2009],
Maz’ya [2011], Runst and Sickel [1996], and Sawano [2018].
Standard references for function spaces in the vector-valued setting in-
clude the works of Amann [1995, 1997, 2019], Triebel [1997], König [1986],
Schmeisser [1987], Schmeisser and Sickel [2001], and Schmeisser and Sickel
[2005]. A unified treatment of Besov and Triebel–Lizorkin spaces and related
classes of function spaces is given by Lindemulder [2021], where the axiomatic
setting of Hedberg and Netrusov [2007] is extended to the vector-valued con-
text. In particular, this covers the weighted and anisotropic settings, and it
allows for Banach function space other than the spaces `q (Lp ) or Lp (`q ) em-
ployed in the construction of the Besov and Triebel–Lizorkin spaces.
The theory of function spaces is a vast topic, and by necessity our
treatment does not cover a number of important topics such as approxi-
mation theory, wavelets, atomic decompositions, weighted spaces, paraprod-
ucts, anisotropic spaces, and typical aspects for bounded domains and man-
ifolds such as traces, extension operators, boundary values, and interpola-
tion with boundary conditions (although some of these topics will be briefly
visited in these notes). Of the omitted themes, we specifically mention the
φ-transform of Frazier and Jawerth [1990], which allows the identification of
Besov and Triebel–Lizorkin spaces with subspaces of appropriate discrete se-
quence spaces. In this identification, the question of boundedness of various
operators on the original function spaces is transformed into the question of
408 14 Function spaces
Section 14.2
Lemma 14.2.1 is taken from Amann [1995]. The other results of this section are
standard in the scalar-valued case, and their extensions to the vector-valued
setting are straightforward.
Section 14.3
The complex and real interpolation results for vector-valued and weighted Lq -
spaces of Theorems 14.3.1 and 14.3.4 extend Theorems 2.2.6 and 2.2.10, where
the unweighted case was treated. The scalar-valued case goes back to Stein
and Weiss [1958], and the extension to the vector-valued weighted setting is
well-known, at least for complex interpolation. The case of real interpolation
is included in the work of Kreı̆n, Petunı̄n, and Semënov [1982], and a dif-
ferent approach based on Stein interpolation for the real method is due to
Lindemulder and Lorist [2022]. The interpolation results for q0 = q1 = ∞ are
false in general. Indeed, already Triebel [1978, 1.18.1] gave an example where
[`∞ ∞ ∞ ns
ws0 (X0 ), `ws1 (X1 )]θ 6= `ws ([X0 , X1 ]θ ) with ws (n) = 2 . Propositions 14.3.3
and 14.3.5 are presented by Triebel [1978], who attributes the real case to Pee-
tre [1967]. More generally, Triebel [1978, Section 1.18] identifies the complex
and real interpolation spaces of `p0 ((Xj )j>1 ) and `p1 ((Yj )j>1 ) for p0 , p1 < ∞
and for sequences of interpolation couples (Xj , Yj )j>1 ; here `p ((Zj )j>1 ) is the
space of all sequences (zj )j>1 with zj ∈ Zj such that (kzj kZj )j>1 belongs to
`p , Z ∈ {X, Y }. Proposition 14.3.3 then follows by taking Xj = 2js X and
Xj = 2js Y . It seems that Proposition 14.3.5 can only be stated for a single
space X unless further assumptions on q0 and q1 are made.
Section 14.4
Our introduction of vector-valued Besov spaces is self-contained up to a mod-
est number of prerequisites from earlier chapters. Part of the section follows
14.8 Notes 409
the presentation by Schmeisser and Sickel [2001]. For the history of Besov
spaces, we refer the reader to Bergh and Löfström [1976] and Triebel [1978,
1983]. Besov spaces appear naturally as real interpolation spaces between Lp
and W k,p (see Theorem 14.4.31). As such, they have important applications
in the theory of evolution equations (see Chapter 18). Moreover, by choosing
the microscopic parameter q suitably, one can often include end-point cases
into the considerations.
In contrast to the theory of the spaces W k,p (Rd ; X) and H s,p (Rd ; X),
where assumptions on the space X such as the Radon–Nikodým property or
the UMD property are often needed, many key results on vector-valued Besov
spaces hold for general Banach spaces X.
Lemma 14.4.5 on the sequential completeness of S 0 (Rd ; X) is a standard
result. It is possible to endow the space Cc∞ (U ; X) with a complete locally
convex topology in such a way that sequential convergence in this topology co-
incides with the ad hoc notion of sequential convergence used here. A detailed
construction is presented by Rudin [1991].
Fourier multipliers
Embedding
Difference norms
classical spaces, as we have done in Corollaries 14.4.25 and 14.4.26 for W s,p
s
and Cub .
In Step 1 of the proof of Theorem 14.4.24 we follow the presentation of
Bergh and Löfström [1976], where the case τ = ∞ was given. Step 2 of the
proof is based on the presentation of Schmeisser and Sickel [2001].
Interpolation
Duality
s
In Theorem 14.4.34, we identified the dual of Bp,q (Rd ; X) with respect to
the duality for S (R ; X) and S (R ; X). Unlike in the Lp -setting treated in
d 0 d
Section 14.5
It is an open problem to characterise the Banach spaces for which these em-
beddings hold (see Problem Q.14).
then F extends boundedly from Lp (Rd , | · |βp ; X) into Lq (Rd , | · |−γq ; X).
412 14 Function spaces
1
In the limiting case γ = max{0, d( min{p,p 0}
+ 1q − 1)}, the above boundedness
of F still holds true under further restrictions on p and q. Surprisingly, if
X has non-trivial Fourier type (equivalently, by Theorem 13.1.33, non-trivial
type), one can allow p = q = 2 by choosing the weights suitably. A similar
result holds in the periodic setting, but the problem is open for more general
orthogonal systems that have been considered by Stein [1956].
R-boundedness
Section 14.6
one can sometimes deduce results about vector-valued Bessel potential spaces
as well. Within the Triebel–Lizorkin scale, one can get closer to H s,p than in
the Besov scale, which often makes Triebel–Lizorkin spaces more useful. For
instance, the sandwich result can be combined with the Sobolev Embedding
Theorem 14.6.14, which allows arbitrary microscopic improvement for Triebel–
Lizorkin spaces. Further flexibility in sandwiching and embedding theorems
can be built in by introducing weights such as |x|γ or |x1 |γ as was done by
Meyries and Veraar [2012, 2014a].
The boundedness of the Peetre maximal function proved in Proposition
14.6.2 appears in the book of Triebel [1997]. This proposition extends results
of Triebel [1983, Theorem 1.6.3] and Triebel [1997, Formula 15.3(iv)] to the
vector-valued setting.
Theorems 14.6.3 and 14.6.11 are presented by Triebel [1997] for scalar-
valued multipliers m. An operator-valued extension is due to Bu and Kim
[2005].
Difference norms
The interpolation and duality results for Triebel–Lizorkin spaces are similar
to their Besov space counterparts. In our presentation, the end-point q = 1 is
excluded, since the Fefferman–Stein inequality for the maximal operator is not
valid in Lp (Rd ; `1 ). This problem can be circumvented by a reduction to in-
terpolation identities for vector-valued Hardy spaces instead of Lp (Rd ; `q (X))
(see Triebel [1983]). The embedding (14.87) of Theorem 14.6.26 is due to
Jawerth [1977], and the one of (14.86) to Franke [1986].
414 14 Function spaces
Section 14.7
The Embedding Theorems 14.7.1, 14.7.3, and 14.7.4 are taken from Schmeisser
and Sickel [2001, 2005]. The end-point cases, where min{p0 , p1 } = 1 <
max{p0 , p1 }, are not completely understood; we refer the reader to Brezis
and Mironescu [2018] for a further discussion.
The Littlewood–Paley theorem 14.7.5 is taken from Meyries and Veraar
[2015], who also consider a weighted setting.
The improved embeddings for Besov, Triebel–Lizorkin, and Bessel poten-
tial spaces under UMD and (co)type assumptions stated in Proposition 14.7.6
are due to Veraar [2013]. The converse result presented in Proposition 14.7.8
seems to be new. In the case p = q, Hytönen and Merikoski [2019] have shown
the following more precise result.
Theorem 14.8.4. For k ∈ N and p ∈ [2, ∞), there is a continuous embedding
k
Bq,q (Rd ; X) ,→ W k,q (Rd ; X)
if and only if X has martingale cotype q.
14.8 Notes 415
Lp –Lq -multipliers
In the scalar-valued case, Lp –Lq Fourier multiplier theorems for p < q first
appeared in the pioneering work of Hörmander [1960]. The scalar-valued case
has the advantage that one can often factor through an L2 -space and use
Plancherel’s identity. In the Banach space-valued case, this is no longer possi-
ble unless additional conditions on the spaces are imposed. The singularities
in Lp –Lq -multiplier theorems for p < q usually behave in a different way from
the case p = q. Often they are absolutely integrable in some appropriate sense,
and then trivially extend to the vector-valued setting by Proposition 2.1.3. A
typical example where this happens is the classical Hardy–Littlewood–Sobolev
inequality on the Lp –Lq -boundedness of f 7→ | · |−s ∗ f .
For operator-valued Lp –Lq -Fourier multipliers, different phenomena arise.
For details and applications to stability of C0 -semigroups we refer the reader
to Rozendaal and Veraar [2018a, 2017, 2018c,b] and the survey by Rozendaal
[2023]. The homogeneous version of Corollary 14.7.7 implies the following
multiplier result of Rozendaal and Veraar [2018a].
Theorem 14.8.5. Let X be a Banach space with type p0 ∈ (1, 2] and let Y
be a Banach space with cotype q0 ∈ [2, ∞). Let p ∈ (1, p0 ) and q ∈ (q0 , ∞),
where we allow p = 2 if p0 = 2 and q = 2 if q0 = 2. Let r ∈ [1, ∞] satisfy
r = p − q . If m : R \ {0} → L (X, Y ) is a strongly measurable function in
1 1 1 d
The proof of this theorem is based on factorisation through γ(L2 (Rd ), X) and
uses the γ-boundedness of the stated operator family. To obtain a homoge-
neous condition on m, one needs the homogeneous version of the γ-Sobolev
embedding. It is not known whether Theorem 14.8.5 holds for p = p0 and
q = q0 . An exception is the case where X and Y are p-convex and q-concave
Banach lattices, respectively; the result then follows from the homogeneous
version of Theorem 14.8.1. Theorem 14.8.5 was used by Rozendaal [2019] to
obtain boundedness of the H ∞ -calculus on fractional domain spaces for strip
type operators. Rozendaal and Veraar [2018a] also prove the following multi-
plier theorem under Fourier type assumptions.
Theorem 14.8.6. Let X be a Banach space with Fourier type p0 ∈ (1, 2] and
let Y be a Banach space with Fourier type q00 ∈ (1, 2]. Let p ∈ (1, p0 ) and
q ∈ (q0 , ∞), and let r ∈ [1, ∞) satisfy 1r = p1 − 1q . If m : Rd \ {0} → L (X, Y )
is a strongly measurable functions and m ∈ Lr,∞ (Rd ; L (X, Y )), then Tm
uniquely extends to a bounded operator from Lp (Rd ; X) to Lq (Rd ; Y )).
The condition m ∈ Lr,∞ (Rd ; L (X, Y )) allows for singularities of the form
|·|−d/r . The proof in the case Cm,r := kmkL (X,Y ) Lr (Rd ) < ∞ with p10 − q10 =
1
r0 is completely straightforward. Indeed, by Hölder’s inequality,
In this chapter we address two strongly interwoven topics: How to verify the
boundedness of the H ∞ -calculus of an operator and how to represent and
estimate its fractional powers. For concrete operators such as the Laplace
operator or elliptic partial differential operators, the fractional domain spaces
can often be identified with certain function spaces considered in Chapter 14
and the imaginary powers of the operator are related to singular integral and
pseudo-differential operators treated in Chapters 11 and 13.
Throughout this chapter, unless otherwise stated, we let A be a sectorial
operator on a Banach space X. We work over the complex scalar field.
the open sector of angle σ in the complex plane; the argument is taken in
the interval (−π, π). A linear operator (A, D(A)) is sectorial if there exists
σ ∈ (0, π) such that the spectrum σ(A) is contained in Σσ and
Here, for z ∈ %(A), the resolvent set of A, R(z, A) := (z − A)−1 denotes the
resolvent of A. In this situation we say that A is σ-sectorial with constant
Mσ,A . The infimum of all σ ∈ (0, π) such that A is σ-sectorial is called the
angle of sectoriality of A and is denoted by ω(A).
By H 1 (Σσ ) we denote the Banach space of all holomorphic functions f :
Σσ → C for which
Definition 15.1.1. For 0 < σ < π we define E(Σσ ) to be the vector space of
holomorphic functions f : Σσ → C of the form
a
f (z) = f0 (z) + + b,
1+z
where f0 ∈ H 1 (Σσ ) ∩ H ∞ (Σσ ) and a, b ∈ C.
We could, more generally, allow functions f0 ∈ H 1 (Σσ ) here, but not much
is gained by doing so because any such function belongs to H ∞ (Σν ) for all
15.1 Extended calculi 421
0 < ν < σ (see Proposition H.1.3). This additional generality would in fact
cause some inconvenience in the statement of the multiplicativity rule (Propo-
sition 15.1.4), where one would be forced to switch to slightly smaller angles.
A further advantage of the present definition is that E(Σσ ) is contained in
H ∞ (Σσ ) as a linear subspace.
and
sup kt 7→ f (eiν t) − c∞ kL1 ((1,∞), dt ) < ∞.
t
|ν|<σ
a
Proof. If f = E(Σσ ) is of the form f (z) = f0 (z) + 1+z + b one may take
c0 = a + b and c∞ = b. In the converse direction, if the bounded holomorphic
function f : Σσ → C has integrable limits c0 and c∞ at 0 and ∞, respectively,
−c∞
then f0 (z) := f (z) − c01+z − c∞ belongs to H 1 (Σσ ) ∩ H ∞ (Σσ ).
The following functions belong to E(Σσ ):
zm
z 7→ for 0 < σ < π and integers n > m > 0;
(1 + z)n
z→
7 exp(−ζz) for 0 < σ < 21 π and ζ ∈ Σ 12 π−σ .
where
a
f (z) = f0 (z) + +b
1+z
with f0 ∈ H 1 (Σσ ) ∩ H ∞ (Σσ ) and a, b ∈ C, and with f0 (A) defined through
the Dunford calculus.
422 15 Extended calculi and powers of operators
Since the constants a and b are uniquely determined by f this is well defined.
For functions in f ∈ H 1 (Σσ ) ∩ H ∞ (Σσ ) the primary calculus of a sectorial
operator A agrees with the Dunford calculus. If A has a bounded H ∞ (Σσ )-
calculus and D(A) ∩ R(A) is dense in X, then for functions f ∈ E(Σσ ) the
definitions of f (A) through the primary calculus agrees with that through the
H ∞ -calculus; this is because in the H ∞ -calculus we have 1+z
1
(A) = (I + A)−1
and 1(A) = I by Theorem 10.2.13.
(f g)(A) = f (A)g(A).
where we used Proposition 10.2.3 to see that ζ(A) = A(I + A)−2 in the
Dunford calculus and hence in the primary calculus. Thus it remains to check
that φ(A)f0 (A) = (φf0 )(A). This follows by applying the resolvent identity
and Cauchy’s theorem to the contour integral representation of the Dunford
calculus:
Z
1
φ(A)f0 (A) = f0 (z)(I + A)−1 R(z, A) dz
2πi Γ
Z
1 f0 (z)
= [R(z, A) − R(−1, A)] dz
2πi Γ 1 + z
Z
1 f0 (z)
= R(z, A) dz
2πi Γ 1 + z
= (φf0 )(A).
zm
(A) = Am (I + A)−n ,
(1 + z)n
m
z
noting that z 7→ (1+z) n belongs to E(Σσ ) for all 0 < σ < π.
By Proposition 15.1.4,
zm z m 1 n−m
(A) = (A) (A)
(1 + z)n 1+z 1+z
= (A(I + A)−1 )m (I + A)m−n = Am (I + A)−n ,
exp(−ζA) := exp(−ζz)(A),
(z n exp(−ζz))(A)x = An exp(−ζA)x.
To see this denote the left-hand side by g(A). By Proposition 15.1.4 and
Example 15.1.5,
1 zn
(I + A)−n g(A) = (A)g(A) = exp(−ζz) (A)
(1 + z)n (1 + z)n
zn
= (A) exp(−ζz)(A) = An (I + A)−n exp(−ζz)(A),
(1 + z)n
from which the claim follows.
The preceding example connects with semigroup theory through Proposition
10.2.7 in Volume II which can be restated in the present language of primary
calculus as follows.
Theorem 15.1.7. Let A be a densely defined sectorial operator on X with
angle ω(A) < 21 π, and let ω(A) < σ < 21 π. Then the bounded holomorphic C0 -
semigroup (S(z))z∈Σ 1 π−σ generated by −A is given by the primary calculus
2
through
S(z) = exp(−zA), z ∈ Σ 21 π−σ ,
where exp(−zA) = exp(−z ·)(A) as in the preceding example.
424 15 Extended calculi and powers of operators
and therefore (ϑf )(A)x = ϑ(A)y by the injectivity of %(A). This shows that
(ϑf )(A)x ∈ R(ϑ(A)), so x ∈ Dϑ (f (A)), and
Interchanging the roles of % and ϑ, one also sees that if x ∈ Dϑ (f (A)), then
x ∈ D% (f (A)). This concludes the proof.
The following observation is an immediate consequence of Proposition 15.1.4.
Lemma 15.1.10. Let A be a sectorial operator on a Banach space X and let
ω(A) < σ < π. If f, g : Σσ → C are holomorphic functions and % ∈ E(Σσ )
and ϑ ∈ E(Σσ ) are regularisers for (f, A) and (g, A), respectively, then %ϑ is
a regulariser for both (f, A) and (g, A).
Proposition 15.1.11. Let A be a sectorial operator on a Banach space X
and let ω(A) < σ < π. Let f : Σσ → C be a holomorphic function such that
the pair (f, A) is regularisable.
(1) the operator f (A) is closed;
(2) if % ∈ E(Σσ ) regularises (f, A), then R(%(A)) ⊆ D(f (A)) and
Proposition 15.1.12. Let A be a sectorial operator on a Banach space X
and let ω(A) < σ < π. Let f, g : Σσ → C be holomorphic functions such that
the pairs (f, A) and (g, A) are regularisable.
426 15 Extended calculi and powers of operators
(1) for all a, b ∈ C the pair (af + bg, A) is regularisable, and for all x ∈
D(f (A)) ∩ D(g(A)) we have x ∈ D((af + bg)(A)) and
(f g)(A)x = f (A)g(A)x.
(f g)(A) = f (A)g(A)
inv(A) = A−1 .
This shows that (%f )(A)g(A)x ∈ R(%(A)) and therefore g(A)x ∈ D(f (A)),
i.e., x ∈ D(f (A)g(A)), and
This shows that (%2 f g)(A)x belongs to R(%2 (A)), so x ∈ D((f g)(A)) by Propo-
sition 15.1.9 and
By part (1) of Proposition 15.1.11 the operator (f g)(A) is closed, and the
above argument shows that it extends f (A)g(A), so f (A)g(A) is closable.
(3): Noting that D((1/f )f )(A) = D(1(A)) = D(I) = X, it follows from
part (2) that if x ∈ D(f (A)), then x ∈ D((1/f )(A)f (A)) and (1/f )(A)f (A)x =
x. Reversing the roles of f and 1/f we also obtain that if x ∈ D((1/f )(A)),
then (1/f )(A)x ∈ D(f (A)) and f (A)(1/f )(A)x = x.
The second assertion follows by considering, e.g., the regulariser %(z) =
z/(1 + z).
As a consequence of what has been shown in the course of the proof of part
(2), and by applying (2) with f and g interchanged, we find that f (A) and
g(A) commute in the following sense: we have
(1) If D(A) is dense in X and rnk f ∈ E(Σσ ), then D(Ak ) is densely contained
in D(f (A)), we have
n o
D(f (A)) = x ∈ X : lim (rnk f )(A)x exists in X ,
n→∞
(2) If D(A) ∩ R(A) is dense in X and ζnk f ∈ E(Σσ ), then D(Ak ) ∩ R(Ak ) is
densely contained in D(f (A)), we have
n o
D(f (A)) = x ∈ X : lim (ζnk f )(A)x exists in X ,
n→∞
where the middle identity follows from the second part of Proposition 15.1.11,
observing that rnk is a regulariser for (f, A). This shows that D(Ak ) is dense
in D(f (A)).
If x ∈ D(f (A)), multiplicativity and the fact that %k rnk f ∈ E(Σσ ) imply
rnk (A)f (A)x = %k (A)−1 rnk (A)(%f )(A)x
= %k (A)−1 (%rnk f )(A)x = (rnk f )(A)x.
Therefore limn→∞ (rnk f )(A)x exists and equals f (A)x.
Conversely, suppose that x ∈ X is such that limn→∞ (rnk f )(A)x =: y exists.
Put zn := rnk (A)(%k f )(A)x. Then zn ∈ D(Ak ), so zn ∈ R(%k (A)). Moreover
zn → (%k f )(A)x, and, by multiplicativity,
%k (A)−1 zn = %k (A)−1 rnk (A)(%k f )(A)x
= %k (A)−1 (%k rnk f )(A)x = (rnk f )(A)x → y.
Since %k (A)−1 is closed it follows that (%k f )(A)x belongs to D(%k (A)−1 ) =
R(%k (A)), and therefore x ∈ D(f (A)).
(2): This is proved in the same way as (1), replacing the use of rn and
Proposition 10.1.7 by ζn and Proposition 10.2.6.
15.1 Extended calculi 429
f (A)g(A) = (f g)(A)
f (Σσ ) ⊆ Στ
for some 0 < τ < π. Suppose furthermore that f (A) is sectorial with
ω(f (A)) < τ . If g : Στ → C is a holomorphic function such that the pairs
(g, f (A)) and (g ◦ f, A) are regularisable, then
f (Σσ ) ⊆ Στ
d∞ d∞
%f = %(f − )+ %
c∞ c∞
%(z) c2∞
−
λ − f (z) c∞ λ − d∞
%(z) c∞ %(z) c∞ (%(z) − c∞ )
6 − +
λ − f (z) c∞ λ − d∞ c ∞ λ − d∞
c∞ (%(z)f (z) − d∞ ) − d∞ (%(z) − c∞ ) c∞ (%(z) − c∞ )
= +
(λ − f (z))(c∞ λ − d∞ ) c ∞ λ − d∞
1 1
6 |c ∞ ||%(z)f (z) − d ∞ | + |d ∞ ||%(z) − c ∞ | + |%(z) − c∞ |
c∞ δ 2 δ
%(·) c2
that %fλ = λ−f (·) has integrable limit c∞ λ−d∞ at ∞.
∞
Replacing c∞ and d∞ by c0 and d0 , in the same way one sees that %fλ has
c20
integrable limit 0 at 0 if c0 = 0, and integrable limit c0 λ−d 0
at 0 if c0 6= 0.
Proof of Theorem 15.1.15. We begin with the proof of the theorem under the
additional assumption made in (i), namely, that g ∈ E(Στ ).
15.1 Extended calculi 431
the operator g(f (A)) being defined by the primary functional calculus of f (A).
Since % regularises (g ◦ f, A), this implies that every x ∈ X belongs to D((g ◦
f )(A)) and
This proves that g(f (A)) = (g ◦ f )(A), and both operators are bounded. This
concludes the proof under the assumption made (i).
For the proof of the theorem under the assumption made in (i), let ϕ ∈
E(Στ ) be a regulariser for the pair (g, f (A)) such that ϕ ◦ f ∈ E(Στ ), and let
ρ be a regulariser for (g ◦ f, A).
We claim that under these circumstances, ρ · (ϕ ◦ f ) regularises (g ◦ f, A).
To this end we must show:
• ρ · (ϕ ◦ f ) · (g ◦ f ) ∈ E(Σσ );
• (ρ · (ϕ ◦ f ))(A) is injective.
The first assertion follows from ρ · (g ◦ f ) ∈ E(Σσ ) (since ρ be a regulariser for
(g ◦ f, A)) and ϕ ◦ f ∈ E(Στ ) (by assumption). For the second assertion we
use the multiplicativity rule of Proposition 15.1.12 (noting that (ϕ ◦ f )(A) =
ϕ(f (A)) by the result of Step 2 and the fact that ϕ ∈ E(Στ )) to see that
The right-hand side is the composition of two injective operators; this is be-
cause ρ is a regulariser for (g ◦ f, A) and ϕ is a regulariser for (g, f (A)). This
proves the claim.
In the following computation, in (i) we use the definition of a regulariser,
in (ii) we apply the result of Step 2 to ϕg ∈ E(Στ ), noting that ϕg satis-
fies the conditions of the theorem since g does, (iii) follows from Proposition
15.1.12, noting that ((ϕg) ◦ f ))(A) = (ϕg)(f (A)) is a bounded operator since
ϕg ∈ E(Στ ), (iv) is a simple rewriting, (v) follows from the definition of a
regulariser, noting that ρ · (ϕ ◦ f ) regularises (g ◦ f, A), (vi) follows by another
application of Proposition 15.1.12, and (vii) uses the result of Step 2 once
again:
(i)
ρ(A)ϕ(f (A))g(f (A)) = ρ(A)(ϕg)(f (A))
(ii)
= ρ(A)((ϕg) ◦ f )(A)
(iii)
= (ρ · ((ϕg) ◦ f ))(A)
(iv)
= (ρ · (ϕ ◦ f ) · (g ◦ f ))(A)
(v)
= (ρ · (ϕ ◦ f ))(A)(g ◦ f )(A)
(vi)
= ρ(A)(ϕ ◦ f )(A)(g ◦ f )(A)
(vii)
= ρ(A)ϕ(f (A))(g ◦ f )(A).
15.1 Extended calculi 433
The identity g(f (A)) = (g ◦ f )(A) follows from this since both ρ(A) and
ϕ(f (A)) are injective.
Our next aim is to relate the extended Dunford calculus with the H ∞ -calculus.
kf (A)k 6 M kf k∞
for all f ∈ H ∞ (Σσ ). In particular, this bound holds for all f ∈ H 1 (Σσ ) ∩
H ∞ (Σσ ). For such functions the extended Dunford calculus agrees with the
Dunford calculus, and therefore the estimate tells us that A has a bounded
H ∞ (Σσ )-calculus.
(2)⇒(1): If x ∈ D(A)∩R(A), then x ∈ R(ζ(A)), say x = ζ(A)y. For the op-
erator f (A) defined through the H ∞ -calculus we have, by the multiplicativity
of the H ∞ -calculus,
where the operator on the right-hand side is again defined by the H ∞ -calculus.
We can also define the operator (ζf )(A) through the primary calculus, and
these two definitions agree (they agree for functions in H 1 (Σσ ) ∩ H ∞ (Σσ )
and for the functions (1 + z)−1 and 1). It follows that
Since D(A) ∩ R(A) is dense in X, this implies that f (A) = ζ(A)−1 (ζf )(A).
The operator on the right-hand side equals the operator f (A) defined through
the extended Dunford calculus, which is therefore bounded.
434 15 Extended calculi and powers of operators
We finish this section with a perturbation result that will be useful in connec-
tion with bounded imaginary powers (see the proof of Lemma 15.3.8).
Theorem 15.1.18. Let A be a densely defined sectorial operator on a Banach
space X and let ω(A) < σ < π. Let f ∈ H ∞ (Σσ ) be given. If the operator
f (A), defined through the extended Dunford calculus of A is bounded, then
also the operator f (A + I), defined through the extended Dunford calculus of
A + I, is bounded and we have
Since kAζ(A + I)k = k(A + I)A(A + I)−2 k 6 Mν,A and Mν,A+I 6 Mν,A , this
proves the estimate
15.1 Extended calculi 435
For functions f ∈ H ∞ (Σσ ) and regulariser %(z) := ζ(z) = z/(1 + z)2 there is
different approach to the extended Dunford calculus via the Cauchy integral
formula, which we outline presently.
Let A be a sectorial operator and let f ∈ H ∞ (Σσ ). For ω(A) < τ < σ 0 < σ,
µ ∈ Σσ0 \ Στ , and x ∈ D(A) ∩ R(A) define
Z
1 1
f (A)x := f (µ)x + f (z) R(z, A) − x dz. (15.4)
2πi ∂Σσ0 z−µ
Let us check that the integrand converges absolutely. Since x ∈ D(A) ∩ R(A)
we may pick y ∈ D(A) with Ay = x. Then
x x kxk
R(z, A)x − = R(z, A)Ay − 6 kR(z, A)Ayk +
z−µ z−µ |z − µ|
which is of the order O(1) as |z| → 0 along ∂Στ , noting that kR(z, A)Ayk =
kR(z, A)[(A − z) + z]yk 6 (1 + kzR(z, A)k)kyk. This establishes the claim.
By an application of Cauchy’s theorem, f (A) is independent of µ ∈ Στ \ Στ 0 .
Since the integrand is an integrable R(A)-valued function, we see that
Note that if f ∈ H 1 (Σσ )∩H ∞ (Σσ ), the above definition of f (A)x agrees with
(10.7).
We will now check that the definition of f (A)x by (15.4) agrees with the
one via Definition 15.1.1 for the regulariser %(z) = ζ(z) = z/(1 + z)2 . Suppose
that x ∈ D(A) ∩ R(A), say x = ζ(A)y. Starting from the latter definition we
have Z
1 zf (z)
f (A)x = (f ζ)(A)y = R(z, A)y dz.
2πi ∂Στ (1 + z)2
Fix ω(A) < τ 0 < τ and µ ∈ Στ \ Στ 0 . To check that (15.4) agrees with
Definition 15.1.1 we must show that
436 15 Extended calculi and powers of operators
f (µ)A(I + A)−2 y
f (z) h z(z − µ)
Z
1 i
= R(z, A)y − (z − µ)R(z, A) − I x dz.
2πi ∂Στ z − µ (1 + z)2
The attentive reader will have noticed that we already used this procedure in
Proposition 10.2.7.
fα (z) := z α := eα log z ,
where we use the branch of the logarithm that is holomorphic in C \ (−∞, 0].
Let 0 < |ν| < σ < π. For z = reiν with r > 0 we have
Aα := fα (A), α ∈ C,
These operators are closed. Moreover, if <λ > 0 and D(A) is dense, then Aα
is densely defined; if <α 6 0 and D(A) ∩ R(A) is dense, then A is injective and
Aα is densely defined. Using the results of Section 10.1.b, These assertions
follow from Proposition 15.1.11, the domain identifications D(An ) = R(%0,n )
and D(An )∩R(An ) = R(%n,n ), and the fact that D(An ) is dense if D(A) dense,
respectively D(An ) ∩ R(An ) is dense in X if D(A) ∩ R(A) is dense in X.
We begin our study of fractional powers with a consistency check.
Proposition 15.2.3. Let A be a sectorial operator on a Banach space X. For
all n = 0, 1, 2, . . . and fn (z) = z n we have
where we used that %n (A) = (I + A)−n in the primary calculus, and that
zn n −n
(%n fn )(A) = (1+z) n (A) = A (I + A) in the primary calculus. This proves
n
that A ⊆ fn (A). In the converse direction, if x ∈ D(fn (A)), then
forcing x ∈ D(An ). This completes the proof of (15.5) for n > 1. For n =
−1, −2, . . . the result follows by applying Proposition 15.1.12(3).
From the definition of the extended Dunford calculus we immediately deduce
the following result.
Proposition 15.2.4. Let A be a sectorial operator on a Banach space X, and
fix an integer k > 1.
(1) For all x ∈ D(Ak ) the function z 7→ Az x is well defined and holomorphic
on {0 < <z < k}.
(2) If A is injective, then for all x ∈ D(Ak ) ∩ R(Ak ) the function z 7→ Az x is
well defined and holomorphic on {−k < <z < k}.
Theorem 15.2.5. Let A be a sectorial operator on a Banach space X, and
let α, α1 , α2 ∈ C.
(1) If A is injective and α ∈ C, then Aα is injective and
Proof. (1): The injectivity of Aα and the identity A−α = (Aα )−1 follow from
Proposition 15.1.12(3). The identity A−α = (A−1 )α follows from Theorem
15.1.15, noting that A−1 is sectorial with the same angle as A.
(2): We consider the regulariser %k (z) = (1 + z)−k , for which we have
R(%k (A)) = D(Ak ).
Let x ∈ D(Aα1 ) and fix an integer k > max{<α2 , <α1 − <α2 }. In order to
prove that x ∈ D(Aα2 ) we must show that ((1 + z)−k z α2 )(A)x ∈ D(Ak ).
Since 2k > <α1 , by the definition of D(Aα1 ) we have ((1+z)−2k z α1 )(A)x ∈
D(A2k ). Using the multiplicativity of the Dunford calculus, this implies that
z k+α2
Ak (I + A)−2k ((1 + z)−k z α2 )(A)x = (A)x
(1 + z)3k
z k−(α1 −α2 ) z α1
= k
(A) (A)x
(1 + z) (1 + z)2k
Proof. Since (λ−cA)−1 = c−1 (c−1 λ−A)−1 , the condition | arg(c)| < π −ω(A)
guarantees that cA is sectorial with ω(cA) 6 ω(A) + |arg c|. Also, for ω(A) <
σ < π − | arg c| and λ ∈ {Σσ+| arg c| we have c−1 λ ∈ {Σσ and
with fα (z) = z α , and ρk (z) := %0,k (z) = (1 + z)−k if <α > 0 and ρk (z) :=
%k,k (z) = z k /(1 + z)2k if <α 6 0. By Cauchy’s theorem we can deform the
path in the above integral to Γ = c · ∂Σω and obtain, by a change of variables,
Z
1
(ρk fα )(cA)x = ρk (z)z α c−1 R(c−1 z, A)x dz
2πi Γ
Z
1
= cα ρk (cz)z α R(z, A)x dz = cα (ρkn (c ·)fα )(A)x.
2πi ∂Σω
(15.6)
If x ∈ D(fα (A)), then (ρk (c ·)fα )(A)x ∈ R(A) (by the definition of D(fα (A)),
since ρk (c ·) is a regulariser for (fα , A)), and (15.6) implies that (ρk fα )(cA)x ∈
R(A) = R(cA). But this implies that x ∈ D(fα (cA)) (by the definition of
D(fα (cA)), since ρk is a regulariser for (fα , cA)). This gives the inclusion
D(fα (A)) ⊆ D(fα (cA)). The same argument in reverse direction gives the
inclusion D(fα (cA)) ⊆ D(fα (A)). Moreover, for any x in this common domain,
the last identity being a consequence of (15.6). Since the right-hand sides are
obviously equal, this gives the result.
Theorem 15.2.7. Let A be a sectorial operator on a Banach space X. If
0 < α < π/ω(A), then Aα is sectorial, we have
ω(Aα ) = αω(A),
If A is R-sectorial and 0 < |α| < π/ωR (A), then Aα is R-sectorial and
µ |µ|1/α µz + |µ|1/α z α
− = = ψτ (|µ|−1/α z).
µ − zα |µ|1/α + z (µ − z α )(|µ|1/α + z)
Hence
1 1 |µ|1/α −1/α
= + ψ τ (|µ| z) .
µ − zα µ |µ|1/α + z
1
Proposition 15.1.12 implies that ( µ−(·) α )(A) is indeed the inverse of (µ −
α α α
(·) )(A) = µ − A . Thus µ ∈ %(A ) and
1 |µ|1/α −1/α
R(µ, Aα ) = + ψ τ (|µ| z) (A)
µ |µ|1/α + z
(15.7)
1
= − |µ|1/α R(−|µ|1/α , A) + ψτ (|µ|−1/α A)
µ
1
using that if λ ∈ {Σσ , then λ−· (A)x = R(λ, A)x, and observing that
−1/α
ψτ (|µ| A) is well defined and bounded by the Dunford calculus of A.
Step 2 – Now let 0 < α < π/ω(A). We will prove that the operator Aα is
sectorial, with ω(Aα ) 6 αω(A).
By Step 1, for τ > αω(A) we have µ ∈ %(A) if | arg µ| > τ . Furthermore,
for σ ∈ (ω(A), τ /α) have
Z
−1/α 1 dz
ψτ (|µ| A) = ψτ (|µ|−1/α z)
2πi ∂Σσ z
Z
1 dz
= ψτ (z) .
2πi ∂Σσ z
0 < <α < <γ < <β or 0 = α < <γ < <β
and for all x ∈ D(Aα ) ∩ D(Aβ ) and ω(A) < σ < π we have
C
kAγ xk 6 kAα xk1−θ kAβ xkθ ,
θ(1 − θ)
Proof. Let m be the smallest integer strictly greater than <β − <α. We will
use
R ∞ the auxiliary function ψ(z) = cz m (1 + z)−2m , where c is chosen so that
ds
0
ψ(s) s = 1. Then the functions
Z 1 Z ∞
ds ds
g(z) := ψ(sz) and h(z) := ψ(sz)
0 s 1 s
h(tA)Aγ x = tα−γ e
h(tA)Aα x.
h(tA)Aα x.
Aγ x = tβ−γ ge(tA)Aβ x + tα−γ e
Therefore,
kAγ xk 6 t<β−<γ ke
g (tA)kkAβ xk + t<α−<γ ke
h(tA)kkAα xk
6 C t<β−<γ kAβ xk + t<α−<γ kkAα xk ,
where the constant C only depends on <β − <α, σ, and A; we used that from
the definition of the primary calculus for it follows that supt>0 kf (tA)k 6 C <
∞ for f ∈ {e h}, using by (10.9) and the sectoriality of A.
g, e
Optimising the choice of t > 0, we arrive at the estimate
" #
θ 1−θ 1 − θ θ
γ
kA xk 6 C + kAα xk1−θ kAβ xkθ .
1−θ θ
Since the term in the square brackets is bounded above by 1/(θ(1 − θ)), this
gives the second estimate.
Remark 15.2.9. It is tempting to believe that
Z 1 Z ∞
ds ds
g(A)x = ψ(sA)x and h(A)x = ψ(sA)x ,
0 s 1 s
444 15 Extended calculi and powers of operators
but these integrals may fail to converges at 0 (the first) and ∞ (the second).
Calderón’s reproducing formula (Proposition 10.2.5) guarantees their conver-
gence (as improper integrals) for elements x ∈ D(A) ∩ R(A) if z 7→ ψ(z)
belongs to H 1 (Σσ ), and for x ∈ D(A) ∩ R(A) if z 7→ ψ(z) log z belongs to
∈ H 1 (Σσ ). The above proof does not depend on these matters; all we needed
there were bounds on the operators g(A) and h(A) that follow directly from
the definitions of these operators through the extended Dunford calculus.
Proof. Let 0 < <α < n. By Theorem 15.2.5 and 15.2.8, applied with θ =
1 − <α/n, for all x ∈ X we have
CMσ,A
kA−α xk = kAn−α (A−n x)k 6 <α
kA−n xk<α/n kxk1−<α/n
n (1− <α
n )
CMσ,A
6 <α <α
kA−1 k<α kxk,
n (1 − n )
Proof. This follows directly from the identity f (A)x = g(Aα )x for x ∈ D(A) ∩
R(A) and f ∈ H ∞ (Σσ ), with g ∈ H ∞ (Σ|α|σ ) given by f (z) = g(z α ).
If A is sectorial, then A + ε is sectorial and boundedly invertible. We conclude
this section a some useful result that applies in this situation.
Proof. The result is clear for α = 1, 2, . . . . Next let α ∈ (0, 1). The functions
(ε + z)α ε + zα
f (z) := − 1, g(z) = −1
ε + zα (ε + z)α
15.2 Fractional powers 445
belong to H 1 (Σσ ) for all 0 < σ < π. For x ∈ D(Ak ) ∩ R(Ak ) with k large
enough, Proposition 15.1.12 gives
Since f (A) and g(A) are bounded, these identities imply D(Aα ) = D((ε+A)α ).
The equivalence of the norms follows from the open mapping theorem.
If β = α + n with n ∈ N and α ∈ (0, 1) then D((ε + A)β ) ⊆ D((ε + A)n )
by Theorem 15.2.5. Thus we obtain
The aim of this section is to prove various integral representations for the
fractional powers of sectorial operators.
If in addition A is densely defined and ω(A) < 21 π, then for all x ∈ D(A) we
have
Z ∞
1
α
A x= s−α S(t)Ax dt,
Γ (1 − α) 0
Note that limz↓0 R(z, A)Ax = 0 for x ∈ D(A) by Proposition 10.1.7, so the first
integral is absolutely convergent. By the same reasoning the second integral is
absolutely convergent. The absolute convergence of the third integral follows
near t = 0 from the fact that x ∈ D(A), and near t = ∞ from the bound
kS(t)Axk 6 Ct−1 kxk (see Theorem G.5.3).
Integrating by parts and using with the identity −αΓ (−α) = Γ (1 − α),
the third identity in Balakrishnan’s theorem may equivalently be presented
as
446 15 Extended calculi and powers of operators
Z ∞
1
α
A x= t−α−1 (S(t)x − x) dt, x ∈ D(A).
Γ (−α) 0
The absolute convergence of this integral follows from the bound kS(t)x−xk =
O(t) as t ↓ 0 for x ∈ D(A).
α
z
Proof. For all ε > 0 the function z 7→ z+ε belongs to H 1 (Σσ ) ∩ H ∞ (Σσ ) and
zα
therefore the operator ( z+ε )(A) can be defined by the Dunford calculus and
is bounded. Fix x ∈ D(A). Then x ∈ D(Aα ), and therefore by multiplicativity
of the extended Dunford calculus (Proposition 15.1.12),
zα
Aα x = (A)(ε + A)x.
z+ε
Similarly,
zα zα
(A)x = (A)(I + A)x.
z+ε (z + ε)(z + 1)
Combining these identities, we compute
zα zα
Aα x = ε (A)(I + A)x + (A)Ax
(z + ε)(z + 1) z+ε
zα
Z
ε
= R(z, A)(I + A)x dz
2πi ∂Σσ (z + ε)(z + 1)
zα
Z
1
+ R(z, A)Ax dz
2πi ∂Σσ z + ε
= (I) + (II).
Noting that z 7→ z α−1 R(z, A)Ax is integrable along ∂Σσ , the term (I) tends
to 0 as ε ↓ 0 by dominated convergence. Also,
Z Z
1 z 1
(II) = z α−1 R(z, A)Ax dz → z α−1 R(z, A)Ax dz
2πi ∂Σσ z + ε 2πi ∂Σσ
The minus sign in the third identity comes from the fact that Γσ is downwards
oriented. The convergence is a consequence of the dominated convergence
theorem.
To prove the third formula we use the identity just proved together with
the Laplace transform representation of the resolvent (Proposition G.4.1) to
get
where Cα = sin(πα) 1
πα 1−α .
(2) If, in addition, A is densely defined and ω(A) < 21 π, and (S(t))t>0 denotes
the bounded analytic C0 -semigroup generated by −A, then for all t > 0 the
operator Aα S(t) is bounded and
We split the integral on the right into two parts and estimate them separately.
First, writing A = (A + t) − t,
Z |λ| Z |λ|
tα−1 (t + A)−1 R(λ, A)Ax dt 6 tα−1 k[I − t(t + A)−1 ]R(λ, A)xk dt
0 0
Z |λ|
−1
6 |λ| tα−1 (1 + M )kλR(λ, A)xk dt
0
M (M + 1) α−1
6 |λ| kxk.
α
Similarly, but now writing A = (A − λ) + λ,
Z ∞ Z ∞
−1
t α−1
(t + A) R(λ, A)Ax dt 6 (1 + M )kxk tα−2 kt(t + A)−1 k dt
|λ| |λ|
M (M + 1) α−1
6 |λ| kxk.
1−α
Turning to the second assertion, by analyticity the operators S(t) map
X into D(A) and supt>0 tkAS(t)k < ∞. The boundedness of the operators
Aα S(t) follows from the boundedness of AS(t) and the inclusion D(A) ⊆
D(Aα ). To prove the estimate, note that for all x ∈ X we have
Z ∞
1
Aα S(t)x = s−α AS(t + s)x ds,
Γ (1 − α) 0
If, in addition, A is densely defined and ω(A) < 21 π, and if (S(t))t>0 denotes
the bounded analytic C0 -semigroup generated by −A, then for all x ∈ R(A)
we have
Z ∞
−α 1
A x= t−α S(t)x dt.
Γ (α) 0
Note that if x = Ay with y ∈ D(A), then R(z, A)x = −y + zR(z, A)y, so the
first integral is absolutely convergent. In the same way it is checked that the
second integral is absolutely convergent. From kS(t)xk = kAS(t)yk = O(1/t)
as t → ∞ (by Theorem G.5.3) we see that the third integral is absolutely
convergent.
Proof. Writing x = Ay with y ∈ D(A) we have A−α x = A1−α y, and Theorem
15.2.13 gives
Z Z
1 1
A−α x = A1−α y = z −α R(z, A)Ay dz = z −α R(z, A)x dz.
2πi ∂Σσ 2πi ∂Σσ
The second identity is proved in the same way. The third follows from the
second by following the lines of the proof of Theorem 15.2.13.
When A boundedly invertible, the identities in the corollary hold for arbitrary
x ∈ X. If in addition A is densely defined, the result extends to arbitrary
<α > 0 as follows:
Theorem 15.2.16. Let A be a densely defined sectorial operator on a Banach
space X with 0 ∈ %(A), and let ω(A) < σ < π. Then for all <α > 0 we have
Z
−α 1
A x= z −α R(z, A)x dz, x ∈ X,
2πi ∂(Σσ \Bε )
Proof. First let x ∈ D(Ak ) with k > <α and set y = ζ(A)−k x = (I +
A)2k A−k x, where ζ(z) = z/(z + 1)2 . The integral
Z
1
Tα x := z −α R(z, A)x dz
2πi ∂(Σσ \Bε )
In the last step, the assumption k > |<λ| was used to justify the change of
contour by Cauchy’s theorem. By the definition of A−α x via the extended
Dunford calculus, the right hand side equals A−α x. This proves the first iden-
tity for x ∈ D(Ak ). Using the second part of Proposition 15.1.13, the general
case follows from it by approximation, noting that Tα is a bounded operator
on X.
Proof. By generalities from semigroup theorem (see Section G.2), the assump-
tions imply that A is densely defined and sectorial with ω(A) 6 21 π. By
Proposition 15.2.7, Aα is densely defined and sectorial of angle 21 πα and con-
sequently −Aα generates a bounded analytic C0 -semigroup by Theorem G.5.2.
15.2 Fractional powers 451
for c > 0. The second and third identity follow from Cauchy’s formula, the
use of which is justified by noting that for z = reiσu with u > 0 we have
α
|esz−tz | = exp(sr cos σ − t<eα(ln r+iσ) )
= exp(sr cos σ − trα cos(ασ)),
α
from which it follows that z 7→ esz−tz is integrable along ∂Σν . In its stated
α
form, Proposition 10.7.2(2) requires gα,t = e−tz to be in H 1 (Σσ ), which is
not the case. The reader may check, however, that the proof still works in
the present situation if we replace integration over ∂Σν by integration over
∂(Σν \ Bε ). For λ > 0 we have
Z ∞ Z c+i∞ Z ∞
−λs 1 α
e fα,t (s) ds = e−λs esz−tz ds dz
0 2πi c−i∞ 0
Z c+i∞ −tzα (15.10)
1 e −tλα
=− dz = e .
2πi c−i∞ z − λ
Using
R∞ the non-negativity of fα,t , upon passing to the limit λ ↓ 0 gives
0
f α,t (s) ds = 1.
Finally, the fact that fα,t is non-negative follows from (15.10), the fact
α
that λ 7→ e−tλ is completely monotone and the Post–Widder real inversion
theorem for the Laplace transform. We refer the reader to the Notes for further
details.
We finish with two examples.
Example 15.2.18 (Fractional derivatives). For 1 < p < ∞, the operator A =
d/dt with domain D(A) = {f ∈ W 1,p (0, T ; X) : f (0) = 0} is sectorial on
Lp (0, T ; X) of angle 21 π and for all <α > 0 and f ∈ Lp (0, T ; X) we have
452 15 Extended calculi and powers of operators
Z x
−α 1
A f (x) = (x − y)α−1 f (y) dy for almost all x ∈ R.
Γ (α) 0
The operators A−α are called the (Liouville) fractional derivatives. In partic-
ular, Z x
A−1 f (x) = f (y) dy
0
−1
The operator V := A is called the Volterra operator. These formulas are
special cases of Theorem 15.2.16 once we note that −A is the generator of the
C0 -semigroup on Lp (0, T ; X) given by
(
f (s − t), s ∈ [0, T ], s > t,
S(t)f (s) =
0, otherwise.
To see that the generator of this semigroup is indeed −A, let us denote the
generator by B for the moment. It is clear that Y := {f ∈ C 1 ([0, T ]; X) :
f (0) = 0} is contained in D(B) and Bf = −f 0 = −Af for all f ∈ Y . Since Y
is also invariant under the semigroup, Y is dense in D(B) by Lemma G.2.4.
But A is a closed operator and Y is also dense in D(A), and therefore B = −A
with equal domains.
H(t)f := kt ∗ f, t > 0,
2
where kt (x) = (4πt)−d/2 e−|x| /(4t) is the heat kernel. It was shown in the same
example that this semigroup extends analytically to {z ∈ C : <z > 0} by the
formula
H(z)f = kz ∗ f, <z > 0,
and that this extension is uniformly bounded and strongly continuous on
every sector Σω with 0 < ω < 21 π. As a consequence, −∆ is a densely defined
sectorial operator of angle ω(∆) = 0.
By Theorem 15.2.17, the operator (−∆)1/2 is densely defined and sectorial
of angle 0 and generates a bounded analytic C0 -semigroup (P (z))z∈Σω for
every 0 < ω < 21 π on Lp (Rd ; X), the so-called Poisson semigroup. By Theorem
15.1.7, in the primary calculus of (−∆)1/2 this semigroup is given by
P (t)f = pt ∗ f, t > 0,
where
Γ ( 21 (d + 1)) t
pt (x) = 1 1
π 2 (d+1) (t2 + |x|2 ) 2 (d+1)
is the Poisson kernel. For d = 1 it takes the simpler form
1 t
pt (x) = .
π t + x2
2
454 15 Extended calculi and powers of operators
where
n
ψn (z) =
, n > 1.
n+z
The remainder of the proof will be devoted to proving the identity
These functions are regularisers for (exp(−t ·), ∆). Once this has been shown
the identity
P (t)f = pt ∗ f, f ∈ Lp (Rd ; X),
follows from Proposition 10.1.7 by passing to the limit n → ∞ in (15.11).
Fixing f ∈ D(∆) ∩ R(∆) and t > 0. Below we will show that
∞ 2
te−t /4s −zs
Z
1/2
e−tz = e ds. (15.12)
0 2π 1/2 s3/2
Assuming this identity for the moment, by Fubini’s theorem and Example
15.1.6 we have
φt (∆)ψn (∆)f = (φt ψn )(∆)f
Z
1 1/2
= e−tz ψn (z)R(z, ∆)f dz
2πi ∂Σσ
Z ∞ 2
te−t /4s −zs
Z
1
= e ψn (z)R(z, ∆)f ds dz
2πi ∂Σσ 0 2π 1/2 s3/2 (15.13)
Z ∞ 2
te−t /4s 1
Z
= e−zs ψn (z)R(z, ∆)f dz ds
0 2π 1/2 s3/2 2πi ∂Σσ
Z ∞ 2
te−t /4s
= exp(−s∆)ψn (∆)f ds.
0 2π 1/2 s3/2
On the other hand,
Z ∞
(∗) 1 t 1 ds
pt (x) = 1 1 s− 2 (d+1) e−1/4s
(4π) 2 (d+1) (t2 + |x|2 ) 2 (d+1) 0 s
Z ∞
t 1 2
+|x|2 )/4s
= 1 s− 2 (d+3) e−(t ds
(4π) 2 (d+1) 0
15.2 Fractional powers 455
∞ 2
te−t /4s
Z
= kt (x) ds,
0 2π 1/2 s3/2
2
where ks (x) = (4πs)−d/2 e−|x| /4s denotes the heat kernel associated with ∆
and (∗) follows from
Z ∞ Z ∞
− 21 (d+1) −1/4s ds 1 1 du 1 1
s e =4 2 (d+1)
u 2 (d+1) e−u = 4 2 (d+1) Γ ( (d + 1)).
0 s 0 u 2
pt ∗ ψn (∆)f
Z ∞Z 2 ∞
te−t /4s
Z
1
= ks (· − y) ψn (z)R(z, ∆)f (y) ds dz dy
−∞ 0 (4π)1/2 s3/2 2πi ∂Σσ
Z ∞ 2 Z ∞
te−t /4s 1
Z
= ψn (z) ks (· − y)R(z, ∆)f (y) dy dz ds
0 2π 1/2 s3/2 2πi ∂Σσ −∞
Z ∞ 2
te−t /4s 1
Z
= ψn (z) exp(−s∆)R(z, ∆)f dz ds
0 2π 1/2 s3/2 2πi ∂Σσ
Z ∞ 2
te−t /4s
= exp(−s∆)ψn (∆)f ds.
0 2π 1/2 s3/2
(15.14)
Renaming the second integration variable and adding the two formulas, the
substitution s = uc − u gives
Z ∞
1 ∞ 1 ∞ −s2
Z Z
c −( c −u)2 c −( c −u)2 1
2
e u du = 1 + 2
e u du = e ds = π 1/2 .
0 u 2 0 u 2 −∞ 2
We will apply this identity with c = 21 tz 1/2 . Completing squares and changing
variables twice, we obtain
∞ 2 ∞
te−t /4s −zs t −(t/2√s−z1/2 √s)2 ds
Z Z
1/2
etz e ds = e √
0 s3/2 s 2 s
Z0 ∞
t −(t/2u−z1/2 u)2
= e du = π 1/2 ,
0 u2
X = N(A) ⊕ R(A).
By (2), the part of A in R(A) is standard sectorial, and the part of A in N(A)
is identically zero.
(1)
γ-bounded bounded
H ∞ -calculus H ∞ -calculus
(3)
(2) (4)
(6) (7)
(8)
almost
γ-sectorial
γ-sectorial
The implications (1), (2), (4), (5), and (8) are trivial. The implication (3)
follows from Theorem 15.3.21, where it is also shown that equivalence holds
when X has Pisier’s contraction property. The implications (1)–(5) are equiv-
alence when X is a Hilbert space. The implication (6) follows from Theorem
15.3.19, and the implication (7) is Theorem 15.3.16.
where we use the branch of the logarithm that is holomorphic on C \ (−∞, 0].
From |ft (z)| = exp(−t arg(z)) it follows that ft ∈ H ∞ (Σσ ) for each 0 < σ < π
and
kft kH ∞ (Σσ ) 6 exp(σ|t|).
Thus if A is a standard sectorial operator with a bounded H ∞ (Σσ )-calculus,
the operators
Ait := ft (A)
are well defined as bounded operators on X. Some examples of operators with
bounded imaginary powers will be discussed in Subsection 15.3.h.
When A is merely standard sectorial, we may use the extended Dunford
calculus to define the operators Ait , t ∈ R, as closed and densely defined
operators in X. This suggests the following definition.
Definition 15.3.4 (BIP). A linear operator A acting in a Banach space X
is said to have bounded imaginary powers (briefly, A has bounded imaginary
powers) if A is standard sectorial and the operators Ait are bounded for all
t ∈ R.
Examples of operators with bounded imaginary powers will be given in Section
15.3.h.
458 15 Extended calculi and powers of operators
Proof. It is evident from the definition through the extended Dunford calculus
that t 7→ Ait x is strongly measurable for all x ∈ X. We have already seen that
Ai0 x = 1(A)x = x for all x ∈ D(A)∩R(A), so Ai0 = I. The identity Ais Ait x =
Ai(s+t) x follows from Proposition 15.1.12. Proposition G.2.7 implies that t 7→
Ait x is continuous for all x ∈ X.
When A has bounded imaginary powers, then by the above result and the
general theory of C0 -(semi)groups, there exist constants M > 1 and ω ∈ R
such that
kAit k 6 M eω|t| .
This allows us to define the abscissa
Proof. (1) and (2): By assumption for all t ∈ R the operators Ait are bounded,
and for <z > 0 the operators A−z are bounded by Corollary 15.2.10. For <z1 >
0 and <z2 > 0 the identity A−z1 A−z2 = A−(z1 +z2 ) follows from Proposition
15.1.12, noting that all operators occurring in this identity are bounded.
We next prove the norm estimate. We begin by noting that
0
CA := sup kA−s k < ∞
s∈[0,1]
0
where CA = CA / max{1, kAk−1 }. This gives the desired bound in (1) for
0 6 <z 6 1.
For z = z 0 + n with n > 1 and 0 6 <z 0 < 1, the estimate in (1) now follows
from
0 0 0
kA−z k = kA−z −n k 6 kA−z kkA−n k 6 CA M −ω|t| kA−1 k<z kA−1 kn
= CA M −ω|t| kA−1 k<z .
(3): Fix an arbitrary integer k > 1 and fix an element x ∈ D(Ak ) ∩ R(Ak ).
We have already seen that z 7→ A−z x is holomorphic on {|<z| < k}; in
particular z 7→ A−z x is continuous on {0 6 <z < k}. The holomorphy on
{|<z| < k} and continuity on {0 6 <z < k} of z 7→ A−z x for general x ∈ X
follows by approximation xn → x with xn ∈ D(Ak ) ∩ R(Ak ), noting that the
above norm estimate implies that the convergence A−z xn → A−z x is locally
uniform on {0 6 <z < k}.
Lemma 15.3.7. If A has bounded imaginary powers, then for all α ∈ C and
t ∈ R we have
Aα Ait = Ait Aα = Aα+it
with equality of domains.
Proof. Since Ait is bounded it is clear that D(Aα ) = D(Ait Aα ). From Propo-
sition 15.1.12(2) we already know the inclusion D(Ait Aα ) ⊆ D(Aα+it ) with
Ait Aα x = Aα+it x for all x ∈ D(Ait Aα ), as well as the equality D(Ait Aα ) =
D(Aα+it )∩D(Aα ). Combining these results, we obtain Ait Aα = Aα with equal
domains.
The second lemma considers bounded imaginary powers for shifted operators:
where S = {z ∈ C : 0 < <z < 1} is the unit strip in the complex plane.
Then f is holomorphic as an X-valued function on S and satisfies f (θ) = x.
Moreover, by Proposition 15.3.6, f is continuous and uniformly bounded on S.
Using the notation introduced in Appendix C, to prove that x ∈ [X, D(Aα )]θ
we must check that f ∈ H (X, D(Aα )). For this it remains to be checked that
t 7→ f (it) belongs to Cb (R; X) and t 7→ f (1+it) belongs to Cb (R; D(Aα )). The
former follows from what has already been said, and for the latter we write
kf (1 + it)kD(Aα ) = kf (1 + it)k + kAα f (1 + it)k. Again by what has already
been said, the function t 7→ f (1 + it) belongs to Cb (R; X). The second term
can be estimated as follows:
2
kAα f (1 + it)k = ke(1+it−θ) Aα A−α(1+it) Aαθ xk
2 2
−t2
= ke(1+it−θ) A−iαt Aαθ xk 6 e(1−θ) M eαω|t| kAαθ xk,
Here we used Lemma 15.3.7, which implies that D(Aα ) = D(Aα+iαt ) and
Aα+iαt y = Aiαt Aα y for y ∈ D(Aα ).
It follows from these estimates that kxn k . kfn kH (X,X) 6 kfn kH (X,D(Aα ))
and kAαθ xn k . kfn kH (X,D(Aα )) , and therefore kxn kD(Aαθ ) . kfn kH (X,D(Aα )) .
Replacing xn by xn − xm in the above argument, we find that the se-
quence (xn )n>1 is Cauchy in D(Aαθ ) and therefore converges to a limit. Since
xn → x in X, this limit must be x. This proves that x ∈ D(Aαθ ) and that
kxkD(Aαθ ) . kf kH (X,D(Aα )) . Taking the infimum with respect to f it follows
that kxkD(Aαθ ) . kxk[X,D(Aα )]θ .
Let us revisit the Laplace operator ∆ on Lp (Rd ; X), where 1 < p < ∞ and
X is a UMD space, with domain D(∆) = H 2,p (R; X). It was already noted
above that −∆ is standard sectorial of angle 0 on Lp (Rd ; X) for all 1 6 p <
1, and by Theorem 10.2.25 it has a bounded H ∞ -calculus of angle 0. As a
consequence, −∆ has bounded imaginary powers. Applying Theorem 15.3.9,
for all 0 < θ < 1 we obtain
D((−∆)θ ) = [Lp (Rd ; X), H 2,p (Rd ; X)]θ with equivalent norms.
462 15 Extended calculi and powers of operators
[Lp (Rd ; X), H 2,p (Rd ; X)]θ = H 2θ,p (Rd ; X) with equivalent norms.
The key lemma is the following representation formula. It expresses the re-
solvent of A in terms of the imaginary powers Ait , and a such it provides the
key insight behind the Clément–Prüss theorem.
the convergence of the principal value integral on the right-hand side being
part of the assertion. Furthermore, for all 0 < s < 1,
Z
1 1
λs As (1 + λA)−1 x = λit Ait x dt. (15.15)
2 R sin(π(s − it))
Proof. We begin with the proof of the first identity. It proceeds in three steps.
Step 1 – First take r = 1 and θ = 0. In this step, for all x ∈ X we will
prove that
15.3 Bounded imaginary powers 463
Z ∞
1 1 π
x+ p.v. A−is x ds
2 2πi −∞ sinh(πs)
Z c+i∞
1 π
= lim A−z x dz,
c↓0 2πi c−i∞ sin(πz)
the convergence of the principal value integral being part of the assertion. Note
that the integrals occurring on right-hand side converge absolutely thanks to
the estimates
| sinh(π(c + it))| = O(eπ|t| ) as t → ±∞
and
kA−c−it xk 6 M eω|t| kA−c xk, t ∈ R,
for all ωBIP (A) < ω < π, with M > 1 a constant depending on ω.
By Cauchy’s theorem we have
Z c+i∞ Z
1 π 1 π
A−z x dz = A−z x dz,
2πi c−i∞ sin(πz) 2πi Γc sin(πz)
where Γc is the (upwards oriented) contour consisting of the union of the two
(1) (3)
half-lines Γc = {is : s 6 −c} and Γc = {is : s > c} and the semi-circle
(2) iϑ 1 1
Γc = {ce : ϑ ∈ [− 2 π, 2 π]}. As c ↓ 0, the contributions along the two
half-lines converge to the principal value integral and the contribution along
the semi-circle converges to 21 x. The latter follows by noting that A−z x → x
as z → 0 in the closed right-half plane, by the continuity of z 7→ A−z x on
that set (see Proposition 15.3.6). Hence
1
2π πceiϕ
Z Z
1 π 1 1
lim dz = lim iϕ
dϕ =
c↓0 2πi (2)
Γc sin(πz) c↓0 2π − 12 π sin(πce ) 2
by the Cauchy theorem, noting that As Ait = As+it by Theorem 15.2.5 in the
first step. This gives the second identity.
Remark 15.3.14. A more direct proof of the second identity can be given as
follows. Starting from the identity
Z ∞
e2πitξ 2e2πsξ
dt = , 0 < s < 1, ξ ∈ R, (15.17)
−∞ sin(π(s − it)) 1 + e2πξ
15.3 Bounded imaginary powers 465
where ω(A) < | arg µ| < ν. Substituting this identity into the right-hand
side of (15.15), a short computation involving Fubini’s theorem, (15.18), and
Cauchy’s theorem gives the result. At the expense of some additional compu-
tations, instead of invoking Proposition 15.1.19 one may also directly use the
definition for Ait x as given in Definition 15.1.8.
Proof of Theorem 15.3.12. (1): First let λ = reiθ with r > 0 and |θ| <
π − ω(A). By subtraction we obtain the identity
Z ∞
−1 −1 1 π
(I + λA) x = (I + A) x + p.v. (λ−is − 1)A−is x ds
2πi −∞ sinh(πs)
for x ∈ D(A2 ) ∩ R(A2 ). The crux is that the term λ−is − 1 is of the order
O(|s|) near s = 0 and can therefore be estimated as |λ−is − 1| . |s| ∧ 1.
Similarly, | sinh(s)| . (|s| ∧ 1)eπ|s| . Therefore the principal value integral is
actually absolutely convergent and bounded in x. As a consequence of this,
the identity extends to arbitrary x ∈ X.
The proof is completed by observing that the integral in the right hand
side of the identity
Z ∞
1 π
(I + λA)−1 x = (I + A)−1 x + (λ−is − 1)A−is x ds
2πi −∞ sinh(πs)
is absolutely convergent for any λ = reiθ with r > 0 and |θ| < π − ωBIP (A).
Indeed, recalling the estimates for λ−is − 1 and sinh(s) mentioned earlier,
choosing ωBIP (A) < ω < π so that |θ| < π − ω we estimate
Z ∞ Z
π −is −is
(λ − 1)A x ds . πe−π|s| Mω eω|s| kxk ds
−∞ sinh(πs) |s|>1
(2): Fix ωBIP (A) < ω < ν < π and choose numbers λn = rn eiθn with
rn > 0 and |θn | < π − ν, as well as vectors xn ∈ X; n = 1, . . . , N . We wish
to show that there exists a constant C, independent of the choices just made,
such that
N
X N
X
εn (I + λn A)−1 xn 6C εn x n ,
L2 (Ω;X) L2 (Ω;X)
n=1 n=1
N
where (εn )n=1 is a Rademacher sequence defined on a probability space (Ω, P).
By a simple approximation argument, there is no loss of generality in assuming
that xn ∈ D(A) ∩ R(A) for all n = 1 . . . , N .
Since ω(A) 6 ωBIP (A) by the Clément–Prüss theorem, Lemma 15.3.13
(with λ = 1), the representation formulas of Lemma 15.3.13 hold for λ = reiθ
with r > 0 and |θ| < π − ν, with x ∈ D(A2 ) ∩ R(A2 ), and
Z ∞
−1 1 1
iθ
(I + re A) x = x + ψθ (s)r−is A−is x ds
2 2πi −∞
Z ∞
1
+ η(s)r−is A−is x ds,
2πi −∞
Z 1
1 ds
+ p.v. r−is A−is x
2πi −1 s
1
=: x + Tr,θ x + Sr x + Rr x,
2
where
π π 1(−1,1) (s)
ψθ (s) = (eθs − 1), η(s) := − .
sinh(πs) sinh(πs) s
Applying this to λ = λn we obtain
N
X
εn (I + λn A)−1 xn
L2 (Ω;X)
n=1
N N
1 X X
6 εn x n + εn Trn ,θn xn
2 n=1 L2 (Ω;X)
n=1
L2 (Ω;X)
N
X N
X
+ ε n Sr n x n + εn R rn x n .
L2 (Ω;X) L2 (Ω;X)
n=1 n=1
Z ∞ N
1 (ω−ν)|s|
X
. Me εn x n ds
2π −∞ n=1
L2 (Ω;X)
N
X
= CA,ν εn x n .
L2 (Ω;X)
n=1
The second term is treated similarly, now using that |η(s)| . e−π|s| :
N Z ∞ N
X 1 X
ε n Sr n x n 6 A−is εn η(s)xn ds
n=1
L2 (Ω;X) 2π −∞ n=1
L2 (Ω;X)
Z ∞ N
1 X
. M eω|s| εn η(s)xn ds
2π −∞ n=1
L2 (Ω;X)
Z ∞ N
1 X
. M e(ω−π)|s| εn x n ds
2π −∞ n=1
L2 (Ω;X)
N
X
0
= CA,ν εn x n .
L2 (Ω;X)
n=1
For estimating the third term we use the UMD property of X through the
boundedness of the Hilbert transform on L2 (R; X).
We begin with a preliminary observation. Let us set Un (s) = (rn A)−is =
−is −is
rn A for brevity. Then by the Kahane contraction principle, for all s ∈ R
we have
N
X N
X
εn Un (s)xn 6 εn A−is xn
L2 (Ω;X) L2 (Ω;X)
n=1 n=1
(15.19)
N
X
ω|s|
6 Me εn x n .
L2 (Ω;X)
n=1
N Z N Z 12 Z
X ds X ds
εn Un (s)xn = εn Un (t) ϕn (t − s) dt
n=1 δ<|s|<1 s n=1 − 12 |s|>δ s
N Z 12 Z 1+t
X ds
− εn Un (s)xn dt
n=1
1
−2 1 s
N Z 12 Z −1+t
X ds
+ εn Un (s)xn dt.
n=1
1
− 2 −1 s
exists in L2 (R; X) by Theorem 5.1.1 and equals πHφn , where H is the Hilbert
transform. As a result we obtain
N N Z 1
X X ds
ε n Rr n x n = εn p.v. Un (s)xn
n=1 n=1 −1 s
N Z
X ds
= εn lim Un (s)xn
n=1
δ↓0 δ<|s|<1 s
N Z 1
X 2
=π εn Un (t)Hϕn (t) dt
n=1 − 12
N Z 1 Z 1+t
X 2 ds
− εn Un (s)xn dt
n=1 − 12 1 s
N Z 1 Z −1+t
X 2 ds
+ εn Un (s)xn dt
n=1 − 12 −1 s
=: (I) + (II) + (III).
It remains to estimate the three terms on the right-hand side. For estimating
+ −
(I) we use that kHkL (L2 (R;X)) 6 2β2,X β2,X (see Theorem 5.1.13). Applying
the Kahane–Khintchine inequality, this gives
N
X Z 1/2
εn Un (t)Hϕn (t) dt
−1/2 L2 (Ω;X)
n=1
15.3 Bounded imaginary powers 469
N
X Z 1/2
h εn Un (t)Hϕn (t) dt
−1/2 L1 (Ω;X)
n=1
Z 1/2 N
X
= εn Un (t)H[1(−1,1) (·)Un (− · )xn ](t) dt
−1/2 n=1 L1 (Ω;X)
Z 1/2 N
X
6 εn Un (t)H[1(−1,1) (·)Un (− · )xn ](t) dt
−1/2 L1 (Ω;X)
n=1
Z 1/2 N
X
6M εn H[1(−1,1) (·)Un (− · )xn ](t) dt
−1/2 L1 (Ω;X)
n=1
Z 1/2 h N
X i
= M1/2 E H 1(−1,1) (·) εn Un (− · )xn (t) dt
−1/2 n=1
h N
X i
6 M1/2 E H 1(−1,1) (·) εn Un (− · )xn
L2 (R;X)
n=1
N
X
+ −
6 2β2,X β2,X M1/2 E εn 1(−1,1) (·)Un (− · )xn
L2 (R;X)
n=1
and, by (15.19),
N
X
E εn 1(−1,1) (·)Un (− · )xn .
L2 (R;X)
n=1
N
X
=E εn (·)Un (− · )xn
L2 (−1,1;X)
n=1
N
X
6 εn (·)Un (− · )xn
L2 (Ω;L2 (−1,1;X))
n=1
N
X
= εn (·)Un (− · )xn
L2 (−1,1;L2 (Ω;X))
n=1
N
X
6 M1/2 εn x n
L2 (−1,1;L2 (Ω;X))
n=1
N
X
= M1/2 εn x n
L2 (Ω;X))
n=1
λAR(λ, A)2 : λ ∈ C \ Σ σ
is uniformly bounded;
(ii) σ-almost γ-sectorial if σ(A) ⊆ Σσ and the set
λAR(λ, A)2 : λ ∈ C \ Σ σ
is γ-bounded.
The operator A is called almost sectorial, respectively almost γ-sectorial if it
is σ-almost sectorial, respectively σ-almost γ-sectorial, for some σ ∈ (0, π).
The identity
λAR(λ, A)2 = [λR(λ, A)]2 − λR(λ, A)
shows that every (γ-)sectorial operator is almost (γ-)sectorial and
15.3 Bounded imaginary powers 471
The above definitions may appear somewhat ad hoc at first sight, but the
motivation to introduce them is as follows. The operators λR(λ, A) used in
the definition of sectoriality can be represented in the primary calculus of A
as
λ
λR(λ, A) = rλ (A) with Rλ (z) = .
λ−z
Indeed, the functions rλ belong to the class E(Σσ ) introduced in Section
15.1.a as long as 0 < σ < |<λ|. They do not belong to H 1 (Σσ ), however, and
this fact is responsible for some of the technical issues encountered in several
proofs. In contrast, the operators λAR(λ, A)2 used in the definition of almost
sectoriality can be represented in the Dunford calculus of A, for we have
λz
λAR(λ, A)2 = reλ (A) with reλ (z) = .
(λ − z)2
Indeed, the functions reλ belong to H 1 (Σσ ) for 0 < σ < |<λ|. Further motiva-
tion will be given in the Notes at the end of the chapter.
The following result gives a version of the (second part of) Clément–Prüss
theorem (Theorem 15.3.12) holds without making any assumptions on the Ba-
nach space X. The price to pay is that only almost γ-sectoriality is obtained:
Proof. Fix ωBIP (A) < θ0 < θ < π and suppose that λ1 , . . . , λn ∈ C are non-
zero and satisfy | arg(λk )| > θ. Note that | arg(µk )| 6 π − θ. Set µk := −1/λk .
Then for all choices x1 , . . . , xn ∈ X we have, by Lemma 15.3.13,
n
1/2
X
E γk λk A1/2 R(λk , A)xk
k=1
n
1/2
X
=E γk µk A1/2 (1 + µk A)−1 xk
k=1
Z n
1 1 X
6 E γk µit it
k A xk dt
2 R | sin(π( 12 − it))| k=1
(∗)
Z (π−(θ−θ 0 ))|t| n
1 e −(π−(θ−θ 0 ))|t| it
it
X
6 e A sup |µ k | E γk xk dt
2 R | sin(π( 12 − it))| 16k6n
k=1
0 n
(∗∗) e(π−(θ−θ ))|t|
Z
1 −θ 0 |t| it
X
6 dt sup e A E γ k xk
2 R | sin(π( 12 − it))| t∈R k=1
n
X
= CE γ k xk ,
k=1
472 15 Extended calculi and powers of operators
where in (∗) we used the contraction principle and in (∗∗) the fact that for
| arg(µ)| 6 π − θ and t ∈ R we have
0 0 0
e−(π−(θ−θ ))|t| Ait |µit | = e−(π−(θ−θ ))|t| e− arg(µ)t Ait 6 e−θ |t| Ait 6 C 0 ,
0
where C 0 := supt∈R ke−θ |t| Ait k is finite since ωBIP (A) < θ0 , and where
0
C0 e(π−(θ−θ ))|t|
Z
C := dt.
2 R | sin(π( 12 − it))|
is γ-bounded as well. Moreover we see that ωeγ (A) 6 θ. This being true for all
ωBIP (A) < θ < π, it follows that ω
eγ (A) 6 ωBIP (A).
{Ait : |t| 6 1}
is γ-bounded.
If A has γ-bounded imaginary powers, the group property Ais Ait = Ai(s+t)
combined with Proposition 8.1.20 (or rather, the elementary bound in the
discussion preceding it) implies that set
{e−ω|t| Ait : t ∈ R}
is γ-bounded for large enough ω > 0. Thus it makes sense to define the abscissa
ensuing proofs will make clear, operators with γ-bounded imaginary powers
can be effectively studied using the continuous square functions introduced in
Section 10.4.b. It is for this reason that our results will be stated for operators
with γ-bounded imaginary powers. The analogous results for operators with
R-bounded imaginary powers automatically follow if the underlying Banach
space has finite cotype.
Proof. Let ωBIP (A) < ν < θ. For each n ∈ Z the singleton {Ain } is γ-bounded,
with γ-bound γ({Ain } = kAin k 6 M eν|n| , where M is a constant independent
of n ∈ Z. By Proposition 8.1.20 (with p = 1 and q = ∞), the set
{e−θ|n| Ain : n ∈ Z}
is γ-bounded. Combined with the fact that {Ais : s ∈ [−1, 1]} is γ-bounded,
by Proposition 8.1.19(3) we obtain that ωγ -BIP (A) < θ.
We have seen in Theorem 15.3.12 that bounded imaginary powers imply sec-
toriality with angle ω(A) 6 ωBIP (A). The next theorem provides the analogue
for γ-bounded imaginary powers.
where C is any finite upper bound for the γ-bounds of the families Γs , s ∈
(0, 21 ). This proves that the set {(I + tA)−1 : t > 0} is γ-bounded.
Applying this reasoning to operators eiθ with 0 < |θ| < π − ωγ-BIP (and
noting that the identity (e±iθ A)it = e∓θ Ait implies that these operators still
have γ-bounded imaginary powers) and using Proposition 8.5.8 to extrapolate
γ-boundedness from the boundary of a sector to the full sector, it follows that
A is γ-sectorial and ωγ (A) 6 ωγ -BIP (A).
Step 2 – We now turn to the proof of the γ-boundedness of the families Γs
uniformly with respect to s ∈ (0, 12 ). We claim that it suffices to prove that
for all f ∈ S (R; X) we have
Z
t 7→ ks (t − u)Ai(t−u) f (u) du 6 Ckf kγ(R;X) , (15.20)
R γ(R;X)
1
where the constant C is independent of 0 < s < 2 and
1
ks (t) := , t ∈ R.
2 sin(π(s − it))
Indeed, suppose that (15.20) has been proved. By Fubini’s theorem and the
second identity of Lemma 15.3.13, for all ξ ∈ R we have
Z Z
ks (t − u)Ai(t−u) f (u) e−2πitξ du dt
R R
Z Z
= ks (t)Ait e−2πitξ dt f (u)e−2πiuξ du = e−2πξs As (1 + e−2πξ A)−1 fb(ξ).
R R
Hence by (15.20) and the fact, observed in Example 9.6.5, that the Fourier
transform extends to an isometry on γ(R; X), we obtain
Since the Fourier transform maps S (R; X) onto itself and this space is dense
in γ(R; X), this estimate extends to all strongly measurable function g : S →
X representing an element of γ(R; X) by density. Then converse to the γ-
multiplier theorem (Proposition 9.5.6) implies that Γs is γ-bounded, with
γ-bound at most C.
Step 3 – To complete the proof of the theorem it remains to prove the
bound (15.20) with a uniform constant C independent of s ∈ (0, 21 ). We start
with the observation that by (15.18) we have
e−sξ
kbs (ξ) = ,
1 + e−ξ
where
X
Ks(n) (u, v) := ks (u − v)1Ij (u)1Ij+n (v).
j∈Z
This sum trivially converges pointwise in (t, v), since each such point is con-
tained in at most one rectangle Ij × Ij+n . We wish to show that the opera-
(n)
tor Ts thus defined extends to a bounded operator on L2 (R), uniformly in
s ∈ (0, 21 ).
For ϕ ∈ Cc (R) we have, by the disjointness of the intervals Ij , monotone
convergence, and a change of variables,
Z X Z 2
kTs(n) ϕk22 = 1Ij (u) ks (u − v)1Ij+n (v)ϕ(v) dv du
R j∈Z R
Z X Z 2
= 1Ij (u) ks (u − v)1Ij+n (v)ϕ(v) dv du
R j∈Z R
XZ Z 2
6 ks (u − v)1Ij+n (v)ϕ(v) dv du
j∈Z R R
X X
2 2
= kkbs 1\
Ij+n ϕk2 6 k1\
Ij+n ϕk2
j∈Z j∈Z
XZ
= |ϕ(u)| du = kϕk22 .
2
j∈Z Ij+n
(n)
This shows that Ts extends to a bounded operator on L2 (R). Moreover,
since
and
1
|ks (u)| 6 . e−π|u| , |u| > 1,
2| sinh(πu)|
by Young’s inequality we have
Z
(n) (n)
kTs k 6 kKs k1 . e−π|u| du . e−2π|n| , |n| > 2.
{|u|>2(|n|−1)}
(n)
By the γ-extension theorem (Theorem 9.6.1), the operators Ts extend to
bounded operators on γ(R; X) and
We may of course relate these constants, but that would only complicate the
estimate below a bit. Fix a Schwartz function f ∈ S (R; X) and an integer
(n)
n ∈ Z. If (u, t) belongs to the support of Ks , then u ∈ Ij and v ∈ Ij+n for
some j ∈ Z, from which it follows that p(u) = p(v) − 2n. Therefore we may
estimate
Z
u 7→ Ks(n) (u, v)Ai(u−v) f (v) dv
R γ(R;X)
Z
= u 7→ Ks(n) (u, v)Ai(u−p(u)+p(v)−v−2n) f (u) du
R γ(R;X)
Z
6 C1 u 7→ Ks(n) (u, v)Ai(p(v)−v−2n) f (v) du
R γ(R;X)
−2π|n| i(p(u)−u−2n)
6 C0 C1 e u 7→ A f (u) γ(R;X)
using the γ-multiplier theorem (Theorem 9.5.1) in the second and fourth step.
Since
X
ks (u − v) = Ks(n) (u, v), u, v ∈ R,
n∈Z
It has already been observed that standard sectorial operators with a bounded
H ∞ -calculus have bounded imaginary powers and ωBIP (A) 6 ωH ∞ (A), the
angle of the H ∞ -calculus of A (see Definition 10.2.10). The following theorem
gives a more precise version of this result.
Moreover,
∞ σ|t|
kAit k 6 Mσ,A e , t ∈ R,
∞
where Mσ,A is the boundedness constant of the H ∞ (Σσ )-calculus of A.
15.3 Bounded imaginary powers 477
with constant C > 0 independent of f . Once this has been shown, we may set
X
f (A) := Aik fk (A),
k∈Z
with convergence in the norm of L (X); here, the operators fk (A) are defined
through the H ∞ (Σσ )-calculus of A. The bound (15.22) implies that
kf (A)k 6 CM kf k∞ . (15.23)
and set Z b
gk (w) := φ(w − t)g(t)e−ikt dt, w ∈ Sσ .
R b
By the Paley–Wiener theorem, φ is an entire function with sufficient decay
to ensure the convergence of the integral for every w ∈ C. Fixing w ∈ Sσ
478 15 Extended calculi and powers of operators
and k ∈ Z, and sing Cauchy’s theorem to shift the path of integration, for
∈ {−1, 1} we may write
Z b
gk (w) := φ(w − t − iν)g(t + iν)e−ik(t+iν) dt, w ∈ Sσ .
R
with Z b
Cσ,ν = sup |φ(x + iy)| dx < ∞.
|y|<σ+ν R
with Cσ,ν as before. It follows that the sum defining f (A) also converges in
H 1 (Σν ). The required consistency now follows by interchanging summation
and integration, along with the fact that (z 7→ z ik fk (z))(A) = Aik fk (A) in the
extended Dunford calculus, hence a posteriori also in the Dunford calculus.
With Theorem 15.3.19 at our disposal we will now investigate the connection
between the γ-boundedness of the imaginary powers Ait and the boundedness
of the H ∞ -calculus of A. In preparation of the next result, it is useful to point
out that in some of these results in Chapter 10 the finite cotype assumption
can be dropped if one defines discrete square functions in terms of Gaussian
sums instead of using Rademacher sums. To be explicit, assuming Definition
10.4.1 to have been modified in this way, the finite cotype assumption can be
dropped in the following results:
• Proposition 10.4.15(2). Indeed, the proof uses the finite cotype assumption
only to pass from Gaussian sums to the Rademacher sums used in the
definition of discrete square functions.
• Theorem 10.4.16(1). Indeed, the finite cotype assumption was only used
to apply Proposition 10.4.15(2).
15.3 Bounded imaginary powers 479
• Proposition 10.4.20. Indeed, the finite cotype assumption was only used
to apply Proposition 10.4.15(2).
The next theorem establishes the connection between γ-bounded imaginary
powers and boundedness of the H ∞ -calculus.
Lemma 15.3.22. Let A be standard sectorial, let ω(A) < σ < π, and suppose
that there is a constant C > 0 such that for all ψ ∈ H 1 (Σσ ) and x∗ ∈
D(A∗ ) ∩ R(A∗ ) we have
Then for all non-zero φ ∈ H 1 (Σσ ) there is a constant cφ > 0 such that for all
x ∈ D(A) ∩ R(A) we have
where we used that ψ(tA)∗ = ψ(tA∗ ). Taking the supremum over all x∗ ∈
D(A∗ ) ∩ R(A∗ ) of normR 6 1, the result now follows from Lemma 10.2.19, with
∞
cφ = inf{kψkH 1 (Σσ ) : 0 φ(t)ψ(t) dtt = 1}.
Proof of Theorem 15.3.21. (1): Fix ωγ-BIP (A) < σ < π. Then the set
{e−σ|t| Ait : t ∈ R} is γ-bounded.
Step 1 – In this step we prove that for all ϑ > σ and x ∈ X the function
t 7→ e−ϑ|t| Ait x belongs to γ(R; X).
By the result of Example 9.4.12 (taking H = C), the function t 7→
e−(ϑ−σ)|t| ⊗ x belongs to γ(R; X) and
1
t 7→ e−(ϑ−σ)|t| ⊗ x γ(R;X)
= kt 7→ e−(ϑ−σ)|t| kL2 (R) kxk h kxk.
(ϑ − σ)1/2
We claim that the functions t 7→ e−ϑ|t| Ait x actually belong to the closed
subspace γ(R; X) of γ∞ (R; X). To prove this, let B be the generator of the C0 -
group (Ait )t∈R . For all x ∈ D(B) and all 0 < a < b < ∞ and −∞ < a < b < 0
the function t 7→ e−ϑ|t| Ait x belongs to C 1 ([a, b]; X), and hence to γ(a, b; X)
by Proposition 9.7.1. Since D(B) is dense in X, the claim now follows from
Corollary 9.5.2.
Step 2 – The formula
cosh(πt) ∞ − 1 +it
Z
1
a− 2 +it = u 2 (u + a)−1 du, a > 0, t ∈ R, (15.25)
π 0
Since ω(A) < π − θ (this is because ω(A) 6 ωBIP (A) = ωγ-BIP (A) by the
Clément–Prüss theorem and Proposition 15.3.18, and ωγ-BIP (A) < ϑ = π − θ
by assumption), we can apply Lemma 10.2.17 (with p = 1) to this identity
and obtain, for all x ∈ X,
e−θ|t| Ait x
t 7→
cosh(πt)
1 eθt
v 7→ eπv+ 2 iθ A1/2 (eiθ e2πv + A)−1 x γ(R;X)
h t 7→ Ait x
cosh(πt) γ(R;X)
−ϑ|t| it
h t 7→ e A x γ(R;X)
1
.A kxk,
ϑ−σ
using (15.24) in the last step. Substituting back ev = u and leaving out terms
of modulus one since they do not affect the γ-norms,
1
u 7→ u1/2 A1/2 (eiθ u + A)−1 x γ(R, du
.A kxk.
u ;X) ϑ−σ
The term in the norm on the left-hand side is of the form φ(u−1 A) with φ(z) =
z 1/2 (eiθ + z)−1 . This function belongs to H 1 (Σϑ0 ) for all 0 < ϑ0 < ϑ = π − θ,
and the estimate can be interpreted as giving the square function estimate
t 7→ φ(tA)x γ(R+ , dt
. kxk, x ∈ X.
t ;X)
Note that up to this point we only have used that A is sectorial and has
bounded imaginary powers (the γ-sectoriality assumption will only be used
482 15 Extended calculi and powers of operators
towards the end of the proof). Because of this, we can apply the same reason-
ing to the part A of A∗ in X := D(A∗ ). Indeed, this operator is sectorial and
has bounded imaginary powers on X and (A )it x∗ = (Ait )∗ x∗ for x∗ ∈ X ;
we leave the easy verification as an exercise to the reader. Together with the
identity φ(tA)∗ x∗ = φ(tA )x∗ , which is equally easy to verify, this gives the
dual square function estimate
The last main result of this chapter is the following characterisation of sectorial
operators on Hilbert spaces with bounded imaginary powers.
Indeed, (15.27) follows by writing x = A−1/2 A1/2 x and using Corollary 15.2.10
to get
kA1/2 xk 6 kxk + kA1/2 xk 6 (kA−1/2 k + 1)kA1/2 xk.
The equivalence (15.28) follows from Proposition K.4.1.
(2)⇒(3): The equality D(A1/2 ) = (H, D(A)) 21 ,2 is an immediate conse-
quence of Peetre’s theorem (Theorem C.4.1), which in the present situation
implies that for each θ ∈ (0, 1) we have
(H, D(A))θ,2 = [H, D(A)]θ with equivalent norms,
and Theorem 15.3.9, which identifies [H, D(A)]1/2 as the fractional domain
space D(A1/2 ) up to an equivalent norm.
(3)⇒(1): On H define
|||x||| := λ 7→ λ1/2 A1/2 (λ + A)−1 x L2 (R+ , dt
.
t ;H)
In view of (15.27) and (15.28) and the assumption in (3), we have the norm
equivalences
|||x||| h kA−1/2 xk(H,D(A)) 1 ,2 h kA−1/2 xkD(A1/2 ) h kxk.
2
15.3.h Examples
It has already been noted that every standard sectorial operator A with a
bounded H ∞ (Σσ )-calculus for some 0 < σ < π has bounded imaginary pow-
ers. Here we wish to highlight two examples:
Example 15.3.24 (Laplacian). Let 1 < p < ∞ and let X be a Banach space. It
was already noted in the discussion preceding Theorem 15.3.11 that if X is a
UMD space, then the negative of the Laplace operator ∆ on Lp (Rd ; X) with
domain D(∆) = H 2,p (Rd ; X) has bounded imaginary powers. In the converse
direction, it was shown in Section 10.5 that if −∆ has bounded imaginary
powers on Lp (Rd ; X), then X is a UMD space.
484 15 Extended calculi and powers of operators
Example 15.3.25 (First derivative). Let 1 < p < ∞ and let X be a UMD
space.
(1) The operator A = d/dx on Lp (R; X) with domain D(A) = W 1,p (R; X)
has bounded imaginary powers with angle 21 π.
(2) The operator A = d/dt on Lp (R+ ; X) with domain D(A) = {f ∈
W 1,p (R+ ; X) : f (0) = 0} has bounded imaginary powers with angle 21 π.
(3) The operator A = d/dt on Lp (0, T ; X) with domain D(A) = {f ∈
W 1,p (0, T ; X) : f (0) = 0} has bounded imaginary powers with angle 21 π
and, more precisely, we have the estimate
1
kAis k .T (1 + s2 )e 2 π|s| , s ∈ R.
For the proofs of (1), (2), and the first part of (3) one may observe that in
each of these three cases A is standard sectorial.
In the case (1), −A generates the translation group on Lp (R; X), and in
the other two cases −A is the generator of the C0 -semigroup on Lp (I; X)
(with I = R+ resp. (0, T )) given by
(
f (s − t), s ∈ I, s > t,
S(t)f (s) =
0, otherwise.
All three semigroups are contractive and, in the scalar-valued case, positive. It
follows that we can apply the Hieber–Prüss theorem (Theorem 10.7.12), which
gives that each of these operators has a bounded H ∞ -calculus of angle 21 π.
It then follows from Theorem 15.3.20 that each of the operators has bounded
imaginary powers.
is finite.
Mπ
6 ,
|y| − π
where we used the elementary inequalities
This proves the theorem under the additional assumption that A is bounded
and has bounded inverse.
Step 2 – To deduce the general case, for ε > 0 we consider the operators
Aε = (A + ε)(I + εA)−1 .
and therefore the operators Aε are uniformly sectorial. In particular the results
of Step 1 apply to Aε , with bounded that are uniform in ε > 0.
Step 3 – Take 0 < ν < π close enough to π so that ∂Σδ is contained in
the resolvent set of each Aε . Noting that R(z, Aε )x → R(z, A)x uniformly on
488 15 Extended calculi and powers of operators
∂Σδ (this follows from uniform sectoriality and similar estimates as above),
we may pass to the limit ε ↓ 0 and obtain
Z
1 1 1 1 1
(A)x = x+ R(z, A) − x dz
λ − log λ − log µ 2πi ∂Σν λ − log z z−µ
Z
1 1 1 1
= lim x+ R(z, Aε ) − x dz
ε↓0 λ − log µ 2πi ∂Σν λ − log z z−µ
Z ∞
1
= lim − (t + Aε )−1 x dt
ε↓0 0 (λ − log t)2 + π 2
Z ∞
1
=− 2 + π2
(t + A)−1 x dt.
0 (λ − log t)
1
Next we show that λ ∈ %(log(A)) and λ−log (A) is a two-sided inverse for
λ−log(A). For x ∈ D(A2 )∩R(A2 ), say x = ζ 2 (A)y, by the general properties of
1
the extended Dunford calculus we have λ−log (A)x ∈ D(A)∩R(A) ⊆ D(log(A))
and
1 ζ
(λ − log(A)) (A)x = ζ(A)(λ − log(A)) (A)y
λ − log λ − log
ζ
= (ζ(λ − log))(A) (A)y
λ − log
ζ
= ζ(λ − log) (A)y
λ − log
= ζ 2 (A)y = x
1
and similarly λ−log (A)(λ − log(A))x = x. By density and closedness, these
identities extend to general x ∈ X and x ∈ D(log(A)), respectively.
Finally, the strip type estimate for A follows from the corresponding esti-
mate for Aε proved above, by letting ε ↓ 0 and using dominated convergence
once more.
One may set up a Dunford calculus and extended Dunford calculus for strip
type operators in much the same way as we did for sectorial operators as
follows. For an operator A of strip type and f ∈ H 1 (Sσ ), where σ > ω S (A),
the Dunford integral
Z
1
f (A)x := f (z)R(z, A)x dx,
2πi ∂Sν
defines a bounded operator f (A) on X. The defining integral converges abso-
lutely and by Cauchy’s theorem it is independent of the choice of ν. Moreover,
Z
1 C 1 C
kf (A)k 6 lim sup |f (z)| |dz| 6 kf kH 1 (Sσ ) .
ν↓ω S (A) 2π ν − ω |=z|=ν π σ − ω S (A)
B it = U (t), t ∈ R.
exists in L2 (R; X) by Theorem 5.1.1 and equals πHφx , where H denotes the
Hilbert transform. As a result we obtain
Z 1 Z
dt dt
p.v. U (t)x = lim U (t)x
−1 t δ↓0 δ<|t|<1 t
Z 12
=π U (s)Hϕx (s) ds
− 12
1 1
Z Z 1+s Z Z −1+s
2 dt 2 dt
− U (t)x ds + U (t)x ds
− 12 1 t − 12 −1 t
=: I + II + III.
With constants C := sup|t|62 kU (t)k and ~2,X := kHkL (L2 (R;X)) , we have
1
Z 2 21 √
kIk 6 πC kHϕx (s)k2 ds 6 πC~2,X kϕx kL2 (R;X) 6 2πC 2 ~2,X kxk.
− 12
we define the bounded operator gb(A) := Φg (A) by the Phillips calculus (see
Section 10.7.a):
15.4 Strip type operators 491
Z ∞
gb(A)x := g(t)U (t)x dt, x ∈ X.
0
Obviously,
g (A)kL (X) 6 M kgkLω1 (R) ,
kb
where kgkLω1 (R) := kt 7→ eω|t| g(t)kL1 (R) and M is in Theorem 15.4.4. The
following lemma enables us to enrich this calculus with certain bounded func-
tions in H ∞ (Sσ ) which have limits for <z → ±∞.
equals Z 1 Z 1
dt dt
h(ξ) =
b e−2πitξ =2 sin(2πtξ)
−1 t 0 t
and its analytic continuation to Sσ satisfies
h(z) = ±π.
lim b
|=z|<σ
<z→∞
Proof. The first assertion follows by elementary computation and the second
from the standard improper integral
Z ∞
dt π
sin t =
0 t 2
f = gb + ab
h+b
1
with a, b ∈ C, g ∈ Lω (R), and h as above, we set
f (A) := gb(A) + ab
h(A) + bI.
We will now exploit the fact that, for 0 6 |ω| < σ < π, the exponential
function z 7→ ez maps the line =z = ω bijectively onto the ray arg(z) = ω.
Thus, it maps the strip Sσ bijectively onto the sector Σσ . From this, we infer
that if λ ∈ {Σσ , then the function
1
rλ (z) := (15.31)
λ − ez
is bounded and holomorphic on Sσ . What is more, this function is of the form
discussed above and therefore rλ (A) is well defined in the primary calculus
(as its derivative satisfies (15.29)).
Remark 15.4.8. In hindsight, one could have introduced the primary calculus
using the functions rλ instead of h. This would restore the symmetry with the
definition of the primary calculus for sectorial operator. Hoever, this would
require an independent construction of the operators rλ (A) = R(λ, B) by
different means.
Rλ − Rµ = (µ − λ)Rλ Rµ , λ, µ ∈ V.
If Rλ0 is injective for some λ0 ∈ V , then there exists a unique closed operator
B on X such that V ⊆ %(B) and Rλ = (λ − B)−1 for all λ ∈ V .
Proof. The resolvent identity implies that any two Rλ and Rµ commute.
If Rλ x = 0, the identity Rλ − Rλ0 = (λ0 − λ)Rλ0 Rλ implies that Rλ0 x = 0
and therefore x = 0. It follows that Rλ is injective. If y ∈ R(Rλ0 ), there is a
unique x ∈ X such that y = Rλ0 x. Then y = Rλ (I − (λ0 − λ)Rλ0 )x ∈ R(Rλ ).
This shows that N := N(Rλ ) = {0} and R := R(Rλ ) are independent of
λ ∈ V . Hence if y ∈ R, then for all λ ∈ V there is a unique xλ ∈ X such that
y = Rλ xλ . Then, by the resolvent identity,
(µ − λ)Rλ Rµ y = Rλ y − Rµ y = Rλ Rµ xµ − Rµ Rλ xλ = Rλ Rµ (xµ − xλ ).
Rλ (λ − B)y = Rλ xλ = y,
1 µ−1
fλ,µ (B) = − (B) (B).
λ − ez µ−1 − e−z
This proves the claim. As a consequence of Proposition 15.3.2, we obtain that
B is standard sectorial.
The identity B it = U (t) follows by writing out the definition of B it in
the extended calculus of B. This results in an expression involving a Dunford
integral containing the resolvent of B. This resolvent can be expressed, via the
definition of the primary calculus, in terms of the Phillips calculus of the C0 -
group U . The details are laborious and are left to the reader. From this, and
the general properties of the extended calculus, it follows that A = log(B).
eA eB x = eC x.
and
15.4 Strip type operators 495
(
ez (1 + ez )−3 , z ∈ Σϑ+ ,
g(z) =
e−2z (1 + e−z )−3 , z ∈ Σϑ− .
For G ∈ {A, B, C} the operators f (G) and g(G) are well defined and bounded
in the bisectorial Dunford calculus. Moreover, by (the bisectorial counterpart
of) Proposition 15.1.12 these operators are injective and ez = f (z)/g(z) im-
plies
eG = g(G)−1 f (G).
Our aim is to prove that
Once we have shown this, from f (A)g(B)−1 ⊆ g(B)−1 f (A) (this follows by
observing that D(g(B)−1 ) = R(g(B)) and using that f (A) and g(B) commute,
we obtain
φ
ch (ξ) = h(−2πξ), ξ ∈ R.
Applying this to G ∈ {−A, −B, −C} and h ∈ {f, g}, and keeping in mind
that −iA, −iB, and −iC generate the groups U (t), V (t), and U (t)V (t), re-
spectively, the identity (15.32) takes the form
Z Z Z
φg (r)φf (t)φf (s)U (t + r)V (s + r) dr ds dt
R R R
Z Z Z
= φf (r)φg (t)φg (s)U (t + r)V (s + r) dr ds dt,
R R R
496 15 Extended calculi and powers of operators
or equivalently,
Z Z Z Z
φg (r)F (t, s)U (t)V (s) ds dt = φg (r)G(t, s)U (t)V (s) ds dt
R R R R
with
Z Z
F (t, s) = φg (r)φf (t−r)φf (s−r) dr, G(t, s) = φf (r)φg (t−r)φg (s−r) dr.
R R
=φcg (x + y)φ
cf (x)φ
cf (y)
= g(−2π(x + y))f (−2πx)f (−2πy)
and similarly
b y) = f (−2π(x + y))g(−2πx)g(−2πy).
G(x,
It is evident from the definitions of f and g that the two right-hand sides
are equal. Therefore F = G be the uniqueness of Fourier transforms. This
completes the proof of (15.32).
Proof. Fix ωBIP (A) + ωBIP (B) < ω < π. Since A and B resolvent commute,
the operators UA (s) = Ais and UB (t) = B it commute for all s, t ∈ R and
U (t) := Ait B −it is a C0 -group satisfying kU (t)k 6 Keω|t| for all t ∈ R and
some K > 1. Hence by Monniaux’s Theorem 15.4.4 we have U (t) = C it ,
t ∈ R, for some standard sectorial operator C having bounded imaginary
15.5 The bisectorial H ∞ -calculus revisited 497
powers. The generators of the C0 -groups equal i log A, −i log B, and i log C.
Since I + C is invertible there is a constant M > 0 such that have
The same argument with the roles of A and B interchanged gives the inequal-
ity
kAxk 6 M kBx + CBxk = M kBx + Axk.
Together, these two inequalities imply the inequality in the statement of the
theorem. This implies the closedness of A + B by a routine argument.
We use the notation introduced in Section 10.6. Specifically, for 0 < ω <
1
2π we define Σω+ := Σω and Σω− := −Σω , and denote by
What distinguishes the theory of bisectorial operators from its sectorial coun-
terpart is the possibility to consider the functions 1Σ + and 1Σ − , both of which
are bounded and holomorphic as functions on the bisector Σ bi = Σ + ∪ Σ − .
If a bisectorial operator A has a bounded H ∞ (Σ bi )-calculus, the operators
1Σ + (A) and 1Σ − (A) are well defined as bounded operators on D(A) ∩ R(A)
and take the role of “spectral projections” associated with the sectors Σ +
and Σ − . (The reason for writing quotations is that one has to be a bit care-
ful here since 0 may belong to the spectrum of A.) From the multiplicativity
of the H ∞ -calculus it follows that the operators 1Σ + and 1Σ − are indeed
projections, and that they are mutually orthogonal in the sense that
1Σ − (A) + 1Σ − (A) = I.
The importance of the operators 1Σ + (A) and 1Σ − (A) stems from their anal-
ogy to the Riesz projections; in particular, their difference
are well defined as bounded projections on D(A) ∩ R(A). As such they are
mutually orthogonal in the sense that
P + P − = P − P + = 0,
P + + P − = I.
It follows that X + is invariant under R(µ, A). Multiplying the first identity on
the left and the second on the right by µ − A+ we see (rµ p+ )(A) is a two-sided
inverse to µ − A+ . It follows that µ ∈ %(A+ ) and
still for x ∈ D(A) ∩ R(A). Among other things it implies that X + is invariant
under (rµ p+ )(A), since we have just proved that (rλ p+ )(A) = P + R(µ, A)
maps into X + . By (15.33) and (15.34) (applied with λ in place of µ), the
right-hand side of (15.35) (hence also the left-hand side) belongs to D(A),
and for x ∈ X + we obtain
It follows that (rµ p+ )(A)x ∈ D(A+ ) and (rµ p+ )(A) is a right inverse of µ−A+
on X + . Also, using (15.34) (again with λ in place of µ) and the fact that
(rλ p+ )(A) and (rµ p+ )(A) in (15.35) commute, for x ∈ D(A+ ) we obtain
that an extended Dunford calculus can be set up for bisectorial operators and
that it enjoys similar properties as in the sectorial case.
Theorem 15.5.2. If A is a standard bisectorial operator with a bounded
H ∞ (Σσbi )-calculus on a Banach space X, then A2 is a standard 2σ-sectorial
operator and
D((A2 )1/2 ) = D(A).
For all x in this common domain we have
k(A2 )1/2 xk h kAxk
with constants independent of x.
Proof. The function a(z) := (z 2 )1/2 is holomorphic on Σσbi and equals z on Σσ+
and −z on Σσ− . Likewise, the function sgn(z) := z/(z 2 )1/2 is holomorphic on
Σσbi and equals 1 on Σσ+ and −1 on Σσ− . Thus, sgn(z)a(z) = z for all z ∈ Σσbi .
By the multiplicativity of the extended functional calculus (cf. Proposition
15.1.12 for the sectorial case), it follows that, if x ∈ D((A2 )1/2 ), then x ∈ D(A)
and Ax = sgn(A)(A2 )1/2 x. Taking norms, we find that
kAxk 6 M k(A2 )1/2 xk,
where M is the boundedness constant of the H ∞ (Σσbi )-calculus of A.
In the same way, the identity a(z) = sgn(z)z implies that, if x ∈
D((A2 )1/2 ), then x ∈ D(A) and (A2 )1/2 x = sgn(A)Ax. Taking norms gives
k(A2 )1/2 xk 6 M kAxk.
It is of some interest to interpret this theorem for the Hodge–Dirac oper-
ator D on L2 (Rd ) ⊕ L2 (Rd ; Cd ) of Example 10.6.5,
0 ∇∗
D= .
∇ 0
where ∇∗ = −div is the adjoint of ∇. Its square is of the form
−∆ 0
D2 = ,
0 ∗
where ∗ equals (at least formally) −∇◦div. Taking g(z) = sgn(z) := z/(z 2 )1/2 ,
then (at least formally)
(−∆)−1/2 0
0 −div 0 ∗
g(D) = D(D2 )−1/2 = · = .
∇ 0 0 ∗ ∇/(−∆)1/2 0
Thus we see the Riesz transform
R = ∇(−∆)1/2
appear as an entry in the functional calculus of D.
502 15 Extended calculi and powers of operators
15.6 Notes
Section 15.1
The idea to use regularising functions to extend the functional calculus to suit-
able classes of unbounded functions goes back to McIntosh [1986]. A compre-
hensive discussion of extended functional calculi is presented by Haase [2006];
see also Haase [2020]. Our treatment in Sections 15.1 and 15.2 is based on
Haase [2006] and Kunstmann and Weis [2004]. The proof of Theorem 15.1.18
is taken from the former reference.
Section 15.2
The first unified account of the theory of fractional powers was undertaken
by Komatsu in a series of papers starting with Komatsu [1966]. This paper
contains the results presented here and much more. Some earlier works on the
subject are due to Balakrishnan [1960], Hille and Phillips [1957], Kato [1960,
1961], Krasnosel0 skiı̆ and Sobolevskiı̆ [1959], Phillips [1952], Watanabe [1961],
and Yosida [1960]. Modern accounts include Albrecht, Duong, and McIntosh
[1996], Denk, Hieber, and Prüss [2003], Dore [1999, 2001], Haase [2006], and
Martı́nez Carracedo and Sanz Alix [2001]. Our treatment barely scratches the
surface of this rich and vast subject, and we have only included the most basic
results needed for the treatment of bounded imaginary powers. Our approach
based on the extended Dunford calculus has the advantage of keeping the
technical details at a minimum, but the price to be paid is that we must make
somewhat restrictive assumptions on the operator A.
Theorem 15.2.8 is from Komatsu [1966], but the proof presented here is
taken from Haase [2006]. Theorem 15.2.13, 15.2.17, and 15.2.16 are due to
Balakrishnan [1960] (see also Yosida [1980]). Some authors take one of the
formulas in the first and third theorem as the definition of the fractional
powers. For further information on the classical approach to fractional powers
via integral representations, we refer the reader to the monographs Butzer
and Berens [1967] and Martı́nez Carracedo and Sanz Alix [2001]. A complete
proof of the non-negativity of the function fα,t in Theorem 15.2.17 can be
found in Yosida [1980, Proposition IX.11.2].
Section 15.3
The next result, a proof of which is given by Kalton, Lorist, and Weis [2023,
Proposition 4.2.4], connects the (almost) γ-sectoriality of A to γ-boundedness
of the associated semigroup.
zAe−zA : z ∈ Σν
For operators A that are diagonal with respect to a Schauder basis, the notion
of γ-almost sectoriality captures a natural property of the basis. In order to
formulate this in the form of a proposition, we first recall that a sequence
P basis of X if every x ∈ X
(xn )n∈Z in a Banach space X is called a Schauder
has a unique representation of the form x = n∈Z cn xn . Associated with a
Schauder basis is its sequence of coordinate projections (Pn )n∈Z on X, defined
by X
Pn cj xj := xn , n ∈ Z,
j∈Z
X n
X
Sn xj := xj , n ∈ N,
j∈Z j=−n
Pn
that is, Sn = k=−n Pk . For any Schauder basis, the set of partial sum
projections is uniformly bounded, and by taking differences the same is seen
to be true for the set of coordinate projections.
On a Banach space X with a Schauder basis (xn )n∈Z , we may consider
the diagonal operator A defined by
Axn := 2n xn , xn ∈ Xn , n ∈ Z,
with its natural maximal domain. It was shown in Proposition 10.2.28 that A
is sectorial of angle ω(A) = 0, and that A has a bounded H ∞ -calculus if and
only if (xn )n∈Z is unconditional. The following result is due to Kalton, Lorist,
and Weis [2023, Proposition 6.1.3].
15.6 Notes 505
Proposition 15.6.2. For the operator A just defined, the following is true:
(1) A is γ-sectorial if and only if the sequence (Sn )n∈N is γ-bounded;
(2) A is almost γ-sectorial if and only if the sequence (Pn )n∈Z is γ-bounded.
We revisit this result in the Notes of Chapter 17 in connection with the prob-
lem of finding examples of sectorial operators that are not R-sectorial.
The Banach space version of this equivalence, due to Kalton, Lorist, and Weis
[2023, Corollary 5.3.9], replaces (15.36) with the condition
the interpolation space on the right-hand side being obtained via the so-called
γ-interpolation method which we briefly outline next.
A discrete version of the γ-interpolation method was already considered
by Kalton, Kunstmann, and Weis [2006], where Rademacher variables were
used instead of Gaussian variables. In that paper, the method was used to
study perturbations of the H ∞ -calculus for various differential operators. The
continuous version of the Gaussian method was introduced by Suárez and
Weis [2006, 2009], where Gaussian interpolation of Bochner spaces Lp (S; X)
and square function spaces γ(S; X), as well as a Gaussian version of ab-
stract Stein interpolation, was studied. An abstract framework covering the
γ-interpolation method, as well as the real and complex interpolation methods,
has been recently developed by Lindemulder and Lorist [2021]. The present
treatment follows the memoir of Kalton, Lorist, and Weis [2023]; theorem
numbers in brackets refer to this memoir. As was pointed out in this refer-
ence, the results in Kalton, Kunstmann, and Weis [2006] and Suárez and Weis
[2006, 2009] were based on a draft version of the memoir.
Let (X0 , X1 ) be an interpolation couple of Banach spaces and let θ ∈ (0, 1).
We call an operator
506 15 Extended calculi and powers of operators
where Tj denotes the operator T from L2 (R, e−2jt dt) into Xj . It is routine to
check that A is complete with respect to this norm.
Denoting eθ : t 7→ eθt , we define (X0 , X1 )θγ as the space of all x ∈ X0 + X1
for which the quantity
Section 15.4
Much of the theory developed in the first three sections of this chapter has an
analogue for strip type operators. The general theory of this class of operators
is developed in the book of Haase [2006], which also treats their connections
with logarithms of sectorial operators. Analogues of the results of Sections
10.3 and 10.4 are presented by Kalton and Weis [2016].
Theorem 15.4.3, on the strip type property and integral representation of
the logarithm log(A) of a standard sectorial A, is due to Nollau [1969]. Our
proof is a variation of the presentation by Haase [2006]. The converse problem
to Theorem 15.4.3, whether the exponent of a striptal operator is sectorial, is
subtle; we refer to Haase [2006] for a counterexample. Theorem 15.4.4, on the
identification of C0 -groups on a UMD space as bounded imaginary powers
of a standard sectorial operator, can be viewed as a partial result in the
positive direction. It was first proved by Monniaux [1999] with a very different
argument based on the notion of analytic generator. The proof presented here
is essentially that of Haase [2009]. Another proof can be found in Haase [2007].
Theorem 15.4.11 about the sum of operators, both of which have bounded
imaginary powers, is due to Dore and Venni [1987] under the slightly stronger
assumption on A and B that they satisfy the resolvent bounds
508 15 Extended calculi and powers of operators
M
k(t + A)−1 k, k(t + B)−1 k 6 , t > 0.
1+t
In its present form, where it is only assumed that
M
k(t + A)−1 k, k(t + B)−1 k 6 , t > 0,
t
the result was obtained by Prüss and Sohr [1990]. The original proof of Dore
and Venni [1987] is ingenious and relatively short, and has been sketched in
the Notes of Chapter 5. It depends on a representation formula for (A + B)−1
in terms of fractional powers of A and B. The refinement of these arguments
by Prüss and Sohr [1990], to obtain the more general case, depends on subtle
approximation arguments for operators A with bounded imaginary powers
which, like the proof presented here, use the functional calculus associated
with the C0 -group (Ait )t∈R and Mellin transform techniques.
The beautiful proof of the Dore–Venni Theorem 15.4.11 presented here
is due to Haase [2007] and fits well in the mainstream of ideas developed in
this volume. This paper also contains our proof of Theorem 15.4.4, which is
originally due to Monniaux [1999] with a different proof based on the notion
of an analytic generator. Our presentation uses some ideas of Haase [2006,
Section 4.2], where a detailed presentation of the theory of strip type operators
if given. With these methods, the operator B = etA can also be defined using
the extended Dunford calculus.
The importance of the Dore–Venni Theorem 15.4.11 is mostly historical,
and the more recent sum-of-operator theorems proved in the next chapter
have turned out to be more versatile in their usage. It is for this reason that
we have contented ourselves with a somewhat sketchy presentation, leaving a
few tedious details to the reader.
Section 15.5
The results of this section follow Duelli [2005] and Duelli and Weis [2005],
where Theorem 15.5.2 (k(A2 )1/2 xk h kAxk) is proved. By the Hieber–Prüss
Theorem 10.7.10, it covers the case where iA generates a bounded C0 -group.
A version of Theorem 15.5.2 (with inhomogeneous estimates) for the case that
iA generates a C0 -group of exponential growth type ω > 0 is due to Haase
[2007].
The spectral projections P ± of Proposition 15.5.1 are studied in more de-
tail by Arendt and Zamboni [2010], Duelli [2005], and Duelli and Weis [2005].
In particular, Arendt and Zamboni [2010] show that, if A is an invertible bi-
sectorial operator and δ > 0 is so small that the closure of the ball B(0, δ)
belongs to the resolvent set of A, then for x ∈ D(A), these projections are
given by
Z
1 dz
P ±x = R(z, A)Ax ,
2πi ∂(Σν± \B(0,δ)) z
15.6 Notes 509
where ω bi (A) < ν < σ is arbitrary. The extended Dunford calculus for bisec-
torial operators, in particular the analogue of Proposition 15.1.12, which was
used in the proof of Theorem 15.5.2, has been studied by Duelli [2005].
For a regularly accretive operator, Kato [1961] defines its real part <A as the
unique maximal accretive operator associated with the sesquilinear form
1
<a : (u, v) 7→ [a(u, v) + a(v, u)]
2
in the above sense. He then proceeds to show that
and that these identities can fail for α > 21 , “but it is not known whether or
not α = 21 can be included”. Kato [1961, Remark 1] writes:
This is perhaps not true in general. But the question is open even
when A is regularly accretive. In this case it appears reasonable to
1 1 1
suppose that both D(A 2 ) and D((A∗ ) 2 ) coincide with D((<A) 2 ) =
D(a), where <A is the real part of A and a is the regular sesquilinear
form which defines A.
As suspected by Kato [1961], a counterexample to the general case of maximal
accretive operators was found shortly after by Lions [1962], but the regularly
accretive case was only disproved a decade later by McIntosh [1972].
What came to be known as Kato’s square root problem was subse-
quently redefined by McIntosh [1982], making the case that, what Kato [1961]
“really had in mind”, was differential operators A = − div B(x)∇, where
B ∈ L∞ (Rd ; Cd ) is such that <(B(x)e|e) > δ > 0 for a.e. x ∈ Rd and all
e ∈ Cd of norm one. For such A, the associated sesquilinear form is
Z
a(u, v) = (B(x)∇u(x)|∇v(x)) dx
Rd
510 15 Extended calculi and powers of operators
with domain D(a) = D(∇) = W 1,2 (Rd ), and the problem thus takes the form
√ ?
k − div B∇ukL2 (Rd ) h k∇ukL2 (Rd ;Cd ) . (15.37)
McIntosh [1982] further suggested that this question was related to Calderón’s
problem about the L2 -boundedness of the Cauchy integral on a Lipschitz
graph (discussed in the Notes of Chapter 12). As pointed out by Alan McIn-
tosh in several oral communications with the authors of this book over the
first decade of this century (the quote within the first sentence of this para-
graph is also from this oral source), before his making this connection, Kato’s
question was generally regarded as being a soft one, several levels easier than
the problem of Calderón, which everyone agreed to be hard. Nevertheless,
the connection suggested by McIntosh [1982] turned out to be a fruitful one,
and the combined efforts of Coifman, McIntosh, and Meyer [1982] led to a
proof of both the boundedness of the Cauchy integral and, what turned out
to be equivalent, McIntosh’s reformulation of Kato’s square root problem in
dimension d = 1.
After this, the status of the redefined square root problem increased sub-
stantially, and important progress was made by Coifman, Deng, and Meyer
[1983], Fabes, Jerison, and Kenig [1985], McIntosh [1985], Alexopoulos [1991],
Journé [1991], Auscher and Tchamitchian [1998], and Auscher, Hofmann,
Lewis, and Tchamitchian [2001], but it took two decades from the one-
dimensional result of Coifman, McIntosh, and Meyer [1982] until a complete
solution was achieved by Hofmann, Lacey, and McIntosh [2002] in dimension
d = 2 and then by Auscher, Hofmann, Lacey, McIntosh, and Tchamitchian
[2002] in all dimensions.
While heavily building on ideas and results about functional calculus of
the second-order operator A = − div B∇, the original solution of the square
root problem was not quite a “pure” functional calculus theorem in the sense
that the gradient featuring in (15.37) does not have the form f (A) of objects
in the functional calculus of A. This “issue” was fixed by a new approach
developed by Axelsson, Keith, and McIntosh [2006] which, in contrast to the
sectorial calculus of second-order operators employed by Auscher et al. [2002],
was based on the bi-sectorial calculus of first-order differential operators, and
promoted the relevance of bi-sectorial operators and bi-sectorial H ∞ -calculus
in subsequent research. Quoting the MathSciNet review of Axelsson et al.
[2006] by Ian Doust, this paper provided “a remarkable consolidation of many
of the ideas that have arisen in the so-called Calderón program”, not only
reproving the square root theorem of Auscher et al. [2002] and several other
results by a unified approach, but also obtaining new geometric applications
to the behaviour of the Hodge–Dirac operator on a Riemannian manifold
under measurable perturbations of the Riemannian metric. In fact, the very
framework of Axelsson et al. [2006] is based on a general notion of perturbed
Hodge–Dirac operators; in the application to the Kato square root problem,
these take the form
15.6 Notes 511
0 − div B
DB := , D(DB ) = D(∇) ⊕ D(div B)
∇ 0
calculus. This work belongs to a circle of ideas that will be treated in Volume
IV.
16
Perturbations and sums of operators
but in concrete cases this definition may be vacuous due to the possibility
that D(A) ∩ D(B) could be unreasonably small or even trivial, i.e., equal to
{0}. Various methods to deal with this problem have been developed, such
as the method of forms. In the context of evolution equations, the two prime
applications one has in mind are cases where either A is the linear operator
governing the equation, e.g., a linear differential operator in the space vari-
ables, and B is the derivative with respect to time, or both A and B are
differential operators in the space variable, typically with B being of lower
order than A. In both of these cases, the resolvent operators R(λ, A) and
R(µ, B) commute and D(A) ∩ D(B) is “large”, in that it contains all elements
of the form R(λ, A)R(µ, B)x with x ∈ X. In fact we have the following result.
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 515
T. Hytönen et al., Analysis in Banach Spaces, Ergebnisse der Mathematik und ihrer
Grenzgebiete. 3. Folge / A Series of Modern Surveys in Mathematics 76,
https://doi.org/10.1007/978-3-031-46598-7_6
516 16 Perturbations and sums of operators
holds for some λ ∈ %(A) and µ ∈ %(B), then it holds for all λ0 ∈ %(A)
and µ0 ∈ %(B) in the connected components of %(A) containing λ and µ,
respectively. This is an easy consequence of the Taylor series identities
∞
X
0
R(λ , A) = (λ − λ0 )n R(λ, A)n+1 ,
n=0
X∞
R(µ0 , B) = (µ − µ0 )n R(µ, B)n+1 ,
n=0
which follow from repeated application of the resolvent identity (see Section
10.1.b). The following definition then suggests itself naturally:
holds for some (or equivalently, all) λ, µ in the connected set {Σσ ∩ {Στ for
some (or equivalently, all) ω(A) < σ < π and ω(B) < τ < π.
we define
Mσ,A := sup{kλR(λ, A)k : λ 6= 0, | arg(λ)| > σ},
fσ,A := sup{kAR(λ, A)k : λ 6= 0, | arg(λ)| > σ}.
M
When A is σ-R-sectorial (the definition being similar), for p ∈ [1, ∞) we set
fRp := Rp ({λR(λ, A) : λ 6= 0, | arg(λ)| > σ}),
M σ,A
fRp := Rp ({AR(λ, A) : λ 6= 0, | arg(λ)| > σ}),
M σ,A
To set the stage for the results to follow, we begin with an elementary per-
turbation result for sectorial operators.
518 16 Perturbations and sums of operators
We first prove the following a priori estimate: For all λ0 > 0 large enough
there exists a constant C > 0 such that
fσ,A kyk + M
kAxk 6 M fσ,A kyk + M
fσ,A kBxk 6 M fσ,A δkAxk + M
fσ,A Kkxk.
C1
kxk 6 kyk =: Dkyk,
|λ| − C2 K
provided we take λ0 > C2 K sufficiently large (in order that |λ| > C2 K). Such
choices of λ0 imply that |λ − λ0 | 6 Cσ |λ| and, together with (16.5),
n
X n
X n
X
εj Axj 6 C0 ε j yj + C0 K εj x j ,
p p p
j=1 j=1 j=1
where C1 = 1 + (1 + δ)C0 .
Next we claim that there exist D > 0 and λ0 > 0 such that
and
n
X n
X n
X
2C1 DK εj x j 6D εj |λj |xj =D εj λj x j .
p p p
j=1 j=1 j=1
n
X
6 C1 D ε j yj ,
p
j=1
Using the inequality a1−θ bθ 6 (1 − θ)a + θb, for all ε > 0 we obtain
1 1
kxk1−θ kxkθD(A) 6 (1 − θ)ε− 1−θ kxk + ε θ kxkD(A) .
The result now follows by combining the estimates and choosing ε > 0 small
enough.
16.2 Perturbation theorems 523
The same proof works if one assumes that B : (X, D(A))θ,p → X is a bounded
operator for some θ ∈ (0, 1) and p ∈ [1, ∞], or that B : [X, D(A)]θ → X is a
bounded operator for some θ ∈ (0, 1). A similar remark applies to Theorem
16.2.8 below.
D(Aα ) ⊆ D(B)
and
kBxk 6 akAα xk + bkxk, x ∈ D(A),
for suitable real numbers a, b > 0 and α ∈ (0, 1). If A has a bounded H ∞ (Σσ )-
calculus in X for some ω(A) < σ < π, then A+B +ν has a bounded H ∞ (Σσ )-
calculus in X for all sufficiently large ν > 0.
where the contour Γη = ∂Ση with ω(A) < η < σ is chosen as usual. Near the
origin, the integrand is bounded since we assumed that 0 ∈ %(A + B + ν). For
large values of |λ| the integrand may be estimated pointwise by
since
M1 := sup{kλR(λ, A + B + ν)k : | arg λ| = η}
is finite by sectoriality of A + B + ν and
In the last line of the statement of the theorem, recall the notation M fθ,A =
sup{kAR(λ, A)k : λ 6= 0, k arg(λ)| > θ}.
If X has the triangular contraction property, in particular if X is a UMD
space, then by Theorem 10.3.4 we have ωR (A) 6 ωH ∞ (A) and therefore the
theorem applies.
At the end of the section, an example will be presented which shows that
the additional assumptions (ii) and (iii) cannot be omitted.
We will reduce the theorem to the following technical lemma.
Lemma 16.2.9. Let A be a densely defined sectorial operator with a bounded
H ∞ (Σσ )-calculus and let B a densely defined R-sectorial operator on X. Let
ω(A) < σ < π and ωR (B) < τ < π, and set µ := max{σ, τ }. Suppose there
exists a holomorphic function M : {| arg(λ)| > µ} → L (X) with R-bounded
range and a β ∈ (0, 1) such that
R(λ, B) = R(λ, A) + Aβ R(λ, A)M (λ)A1−β R(λ, A), | arg(λ)| > µ. (16.10)
Then B has a bounded H ∞ (Σµ )-calculus.
16.2 Perturbation theorems 527
Proof. Our aim is to prove that there exists a function φ ∈ H 1 (Σµ ) and a
constant C > 0 such that for all integers N > 1, all scalars −N , . . . , N of
modulus one, and all t > 0 we have
X
n φ(t2n B) 6 C.
|n|6N
Once we have this, it follows from Proposition 10.4.11 (and tracking angles in
its proof) that B has a bounded H ∞ (Σµ )-calculus.
Let µ < ν < π and consider the function ψν ∈ H 1 (Σµ ) given by
z 1/2
ψν (z) = , z ∈ Σµ ,
(eiν − z)1/2 (2eiν − z)1/2
By (16.10),
By Corollary 15.2.14, the right-hand side has decay of order |λ|−1 as |λ| → ∞
in the complement of Σµ . Hence, by Cauchy’s theorem and taking µ < τ < ν,
Z
1
φν (t2n B) = φν (λ)R(λ, t2n B) dλ
2πi ∂Στ
= φν (t2n A) + t2n Aβ R(eiν , t2n A)M (t−1 2−n eiν )A1−β R(eiν , t2n A)
− t2n+1 Aβ R(2eiν , t2n A)M (t−1 2−n 2eiν )A1−β R(2eiν , t2n A)
= φν (t2n A) + t2n Aβ R(eiν , t2n A)M (t−1 2−n eiν )A1−β R(eiν , t2n A)
− t2n−1 Aβ R(eiν , t2n−1 A)M (t−1 2−(n−1) eiν )A1−β R(eiν , t2n−1 A)
= φν (t2n A) + φβ,ν (t2n A)M (t−1 2−n eiν )φ1−β,ν (t2n A)
− φβ,ν (t2n−1 A)M (t−1 2−(n−1) eiν )φ1−β,ν (t2n−1 A)
= (I) + (II) + (III),
(16.11)
zα
φα,ν (z) := .
eiν − z
In the penultimate identity of (16.11) we used the identity
where the implicit constant in the last step is the R-boundedness constant of
M . Similarly, shifting the index by one and using the contraction principle,
we estimate (III) as follows:
X
n φβ,ν (t2n A)M (t−1 2−n eiν )φ1−β,ν (t2n A)x
|n|6N
X X
.M E εn φ1−β,ν (t2n−1 A)x sup εn φβ,ν (t2n−1 A∗ )x∗
kx∗ k61
|n|6N |n|6N
X X
6E n
εn φ1−β,ν (t2 A)x sup E εn φβ,ν (t2n A∗ )x∗ .
kx∗ k61
|n|6N +1 |n|6N +1
Taking the supremum over N > 1 and t > 0, this proves the square function
bound
16.2 Perturbation theorems 529
X
sup sup n φν (t2n B)x 6 Ckxk + 2C 0 sup kxkφ1−β ,A sup kx∗ kφβ ,A∗
|n|>N t>0 t>0 kx∗ k61
|n|6N
6 C 00 kxk,
where the estimate in the last step follows from the boundedness of the
H ∞ (Σσ )-calculus of A through Theorem 10.4.4.
Proof of Theorem 16.2.8. By the second part of Theorem 16.2.4, assump-
tion (i) implies that A + B is σ-R-sectorial operator provided the small-
ness condition on C0 in (i) holds. Moreover, if we impose C0 < 1, then
for all x ∈ D(A + B) = D(A) we have kAxk 6 k(A + B)xk + kBxk 6
k(A + B)xk + C0 kAxk and therefore kAxk 6 (1 − C0 )−1 k(A + B)xk, while at
the same time k(A + B)xk 6 kAxk + kBxk 6 (1 + C0 )kAx. We conclude that
again provided C0 is small enough, for then kBR(λ, A)k 6 C0 kAR(λ, A)k 6
C0 kλR(λ, A) − Ik 6 C0 (1 + Mσ∨τ,A ) < 1 and the series converges absolutely.
First we assume that (i) and (iii) hold. For the time being, we do not
assume that 0 ∈ %(A) (in which case A−1 is bounded by Corollary 15.2.10),
but only assume that A is invertible and B maps D(A1−α ) into D(A−α ). Then
U := A−α BAα−1 is bounded on X of norm kU k = C1 and we have
converges in operator norm and defines a holomorphic function for | arg λ| >
σ ∨ τ . We then can rewrite (16.12) in the form
X
R(λ, A + B) = R(λ, A) + Aα R(λ, A) [U AR(λ, A)]k U A1−α R(λ, A)
k>0
By the R-sectoriality of A and Proposition 8.1.24, the set {M (λ) : | arg(λ)| >
σ} is R-bounded. Thus we derived the representation required in Lemma
16.2.9 and we can conclude that A+B also has a bounded H ∞ (Σσ∨τ -calculus.
530 16 Perturbations and sums of operators
with equivalent norms, with equivalence constants which may be chosen in-
dependent of θ ∈ (0, 1). Thus we obtain that B acts as a bounded operator
from X1−θα to X−θα with norm . C01−θ C1θ , 0 < θ < 1.
We can choose θ so small that B satisfies (iii) for α0 = θα with C10 <
−1
M
f
σ∨τ,A no matter how big C1 was. This completes the proof of the case (iii).
Finally assume that (ii) holds for some α ∈ (0, 1). By Proposition 15.1.12
we have Aα−1 ⊆ Aα A−1 and A1+α ⊆ Aα A (in fact we have equality in the
second case by Theorem 15.2.5), and therefore
kAα−1 Bxk = kAα A−1 Bxk 6 C1 kA1+α A−1 Bxk = C1 kAα Bxk
implies that (iii) holds for the exponent 1 − α ∈ (0, 1) and x ∈ D(A1+α ). Since
D(A1+α ) is dense in D(Aα ) by Proposition 15.1.13, (iii) holds for the exponent
1 − α ∈ (0, 1) and x ∈ D(Aα ).
We conclude this section with an example, due to McIntosh and Yagi, shows
that boundedness of the H ∞ -calculus is not preserved by small relatively
bounded perturbations even when X is a Hilbert space. This shows that the
additional assumptions (ii) or (iii) in Theorem 16.2.7 cannot be omitted, no
matter how small the constant on (i) is chosen.
bounded bisectorial H ∞ -calculus, and this is not the case. We conclude that
A2 + Cε does not have a bounded H ∞ -calculus.
Let us proceed to the construction of the operators A and Bε . Fix ε > 0.
Omitting subscripts ε in what follows, for n = 1, 2, . . . we will construct
bounded operators An and Bn on a finite-dimensional Hn with the following
properties for any 0 < σ < 21 π:
• An and An + Bn are σ-bisectorial with kBn k 6 εkAn k;
• A2n and (An + Bn )2 = A2n + Cn with kCn k 6 2εkA2n k;
• the spectra of An and An + Bn is contained in (−∞, 1] ∪ [1, ∞);
• the resolvents of An and An + Bn satisfy
for all λ ∈ C \ R;
• An and An + Bn have contractive, respectively bounded, H ∞ (Σσ± )-calculi;
• the spectral projections 1Σσ± (An + Bn ) have norm > n.
The counterexample with the stated properties is obtained by taking
M M M M
H= Hn , A := Tn , B := Bn , C := Cn .
n>1 n>1 n>1 n>1
(n)
Sn = (sjk )N
j,k=0 given by
n
(n) (n) ε
tjk = 2j δjk , sjk := δj6=k .
π(k − j)
Then Tn is self-adjoint and Sn Tn is skew-adjoint. The self-adjoint matrix iSn
is the Nn × Nn Toeplitz matrix with generating function εθ/π, θ ∈ (−π, π),
that is, we have
sjk = fbj−k , j, k = 0, . . . , Nn .
Since the norm of a Toeplitz matrix with bounded real-valued generating
function f is bounded by kf kL∞ (T) , we see that kSn k 6 ε.
(n)
The matrix Zn = (zjk )N
j,k=0 given by
n
(n) 2k ε
zjk = δj6=k
π(k − j)(2j + 2k )
532 16 Perturbations and sums of operators
has norm
Nn −1
2N n ε X 1
kZn k > kZeNn k =
π j=0 (Nn − j)(2j + 2Nn )
ε 2Nn 1 1 (16.13)
> N −1 N
+ + · · · + 1
π (2 n + 2 n ) N n Nn − 1
2ε
> log(Nn + 1) > n,
3π
where (en )N
n=0 denote the standard unit vectors in C
n Nn +1
, and it satisfies
T n Z n + Z n T n = Sn T n . (16.14)
with
kCn k 6 kTn Sn Tn − Sn Tn2 k 6 2kSn kkTn k2 6 2εkA2n k,
where we used that Tn∗ = −Tn , so Tn is normal and therefore kTn k2 = kT 2 k.
Furthermore, one checks that σ(An ) = σ(An + Bn ) and
R(λ, Tn ) R(λ, Tn )Sn Tn R(λ, −Tn )
R(λ, An + Bn ) = (16.15)
0 R(λ, −Tn )
σ(An + Bn ) = σ(An )
= σ(Tn ) ∪ σ(−Tn ) = {1, 2, 4, . . . , 2Nn } ∪ {−1, −2, −4, . . . , −2Nn }
Indeed, using (16.14) and (16.15), for 0 < ν < σ we formally compute
1Σσ± (An + Bn )
Z
1
= R(z, An + Bn ) dz
2πi ∂Σν±
Z
1 R(z, Tn ) R(z, Tn )Sn Tn R(z, −Tn )
= dz
2πi ∂Σν± 0 R(z, −Tn )
Z
1 R(z, Tn ) R(z, Tn )(Tn Zn + Zn Tn )R(z, −Tn )
= dz
2πi ∂Σν± 0 R(z, −Tn )
Z
1 R(z, Tn ) R(z, Tn )Zn + Zn R(z, −Tn )
= dz
2πi ∂Σν± 0 R(z, −Tn )x
(∗) I Zn
= = Pn+ ,
0 0
and the reverse triangle inequality (16.16) holds. If in addition X has the
triangular contraction property, then A + B is R-sectorial and
(Theorem 16.3.6).
16.3 Sum-of-operator theorems 535
then the reverse triangle inequality (16.16) holds (Theorem 16.3.14). The
same conclusion holds if X is a Hilbert space, A has bounded imaginary
powers and B is densely defined, and ωBIP (A) + ω(B) < π (Theorem
16.3.15).
To conclude this section we provide an example of the type of applications
that will be studied in depth in the next two chapters and which indeed have
motivated the development of the abstract approach to sums of operators
presented here.
Suppose that −A generates a C0 -semigroup on a Banach space X and
consider the inhomogeneous abstract Cauchy problem
(
u0 (t) + Au(t) = f (t), t ∈ [0, T ],
(ACP)
u(0) = 0.
Du := u0
It will be checked in the next chapter (see Section 17.3.c) that this operator
is sectorial of angle 21 π. Using this operator, we can rewrite (ACP) as the
abstract operator equation
(D + A)u
e =f
(Af
e )(t) := A(f (t)), t ∈ (0, T ).
In the next chapter (see Propositions 17.3.14 and 17.3.15) we prove that the
following assertions are equivalent:
536 16 Perturbations and sums of operators
(1) the inverse triangle inequality (16.16) holds, i.e., there is a constant C > 0
such that
kAuk
e p + kDukp 6 Ck(A
e + D)ukp , e ∩ D(D);
u ∈ D(A)
(2) A
e + D is closed;
(3) e + D boundedly invertible;
A
(4) A has maximal Lp -regularity on (0, T ).
For the problem (ACP), maximal Lp -regularity means that the unique mild
solution of the problem, which is given in terms of the semigroup S generated
by −A as Z t
u(t) = S(t − s)f (s) ds
0
then the operator A + B with its natural domain D(A + B) = D(A) + D(B)
has a closed extension to a sectorial operator C which satisfies
Furthermore,
(1) If A or B is injective, then C is injective;
(2) If A and B are densely defined, then C is densely defined;
(3) If A and B are densely defined and A or B is standard sectorial, then C
is standard sectorial.
16.3 Sum-of-operator theorems 537
If (2) holds (and hence if (3) holds), then C equals the closure of A + B.
The proof of this theorem will be given shortly. We first pause a brief moment
to explain why the condition
enters naturally in this theorem. Variants of this condition appear in all sum-
of-operator theorems we are about to encounter. Arguing naively, one would
like to realise the operator sum A + B through a ‘bivariate’ extended Dunford
calculus as (z +w)(A+B), where z +w is short-hand for the function (z, w) 7→
z + w. With this notation, to prove sectoriality of A + B one must estimate
the norms of
λ
λR(λ, A + B) = (A, B)
λ − (z + w)
for all λ ∈ C in the complement of a sector Σω containing all sums z + w with
z ∈ Σσ and w ∈ Στ , where ω(A) < σ < π and ω(B) < τ < π as usual. But
the algebraic sum Σσ + Στ is a sector only if σ + τ 6 π! Under this condition,
Σσ + Στ = Σmax{σ,τ } . In contrast, when σ + τ > π the reader may check that
Σσ + Στ = C. Clearly, the condition σ + τ 6 π forces ω(A) + ω(B) < π, and
in that case we may replace σ and τ by slightly smaller values to arrange that
σ + τ < π. Incidentally, this heuristic argument also shows that the inequality
ω(A + B) 6 max{ω(A), ω(B)} is natural to expect.
Let us now turn to the proof Theorem 16.3.2. Let A and B be resolvent
commuting sectorial operators in X satisfying ω(A) + ω(B) < π, and let
ω(A) < σ < π and ω(B) < τ < π be such that σ + τ < π. The construction of
the sectorial operator C extending A+B is based on the following observation,
which makes use of the primary calculus involving the spaces E(Σ) introduced
in Section 15.1.a. For a holomorphic function h ∈ E(Σσ ) ⊗ E(Στ ) of the form
N
X
h(z, w) = fn (z)gn (w)
n=1
It is not difficult that the operator h(A, B) is well defined, in the sense that it
does not depend on the particular representation of h. We now observe that
z+w 1 1 1 1
h(z, w) := = 1− + 1− .
(1 + z)(1 + w) 1+z 1+w 1+z 1+w
1
ρ(z, w) :=
(1 + z)(1 + w)
as a regulariser for the function (z, w) 7→ z + w, we define
C := (I + A)(I + B)h(A, B)
with domain
Choose ω(A) < σ < π and ω(B) < τ < π in such a way that σ + τ < π.
As was already observed above, the condition σ + τ < π implies that
Σσ + Στ := {z + w : z ∈ Σσ , w ∈ Στ } = Σmax{σ,τ } .
λ λ2 λzw
= + .
λ − (z + w) (λ − z)(λ − w) (λ − (z + w))(λ − z)(λ − w)
These functions are holomorphic on Σσ × Στ and one may check that
and the sectoriality of C with angle ω(C) 6 ω now easily follows from the
fact that value of the integral on the right-hand side of (16.19) is uniformly
bounded with respect to µ on the arc {|µ| = 1, | arg(µ)| > max{σ, τ }.
Since the choices ω(A) < σ < π and ω(B) < τ < π and max{σ, τ } < ω < π
were arbitrary, it follows that ω(C) 6 max{ω(A), ω(B)}.
It remains to prove the assertions (1)–(3).
(1): Suppose that is injective and let x ∈ D(C) be such that Cx = 0. By
the definition of C, this means that h(A, B)x = (I + B)−1 (I + A)−1 y for some
y ∈ X and Cx = (I + A)(I + B)h(A, B)x = y = 0. Consider the function
z
g(z, w) = , z, w ∈ Σσ .
(1 + z)2 (z + w)
By the primary calculus, for fixed z ∈ Σσ we have
Borrowing some terminology from the next subsection, this function belongs
to H 1 (Σσ ; A ), where A is the set of operators in L (X) commuting with the
resolvent of A, and we may define a bounded operator g(A, B) through the
Dunford integral
Z
1
g(A, B) := z(1 + z)−2 (z + B)−1 (z − A)−1 dz.
2πi ∂Σν
In view of
(z + B)−1 (z − A)−1 C
= (z + B)−1 (z − A)−1 [(I + A)(I + B)h(A, B)]
= [(1 + z)(z − A)−1 − I][(1 − z)(z + B)−1 + I]h(A, B)
= [(1 + z)(z − A)−1 − I][(1 − z)(z + B)−1 + I](A + B)(I + A)−1 (I + B)−1
= [(1 + z)(z − A)−1 − I][I − (I + A)−1 ][(1 − z)(z + B)−1 + I](I + B)−1
+ [(1 + z)(z − A)−1 − I][(1 − z)(z + B)−1 + I][I − (I + B)−1 ](I + A)−1
= (z − A)−1 − (z + B)−1 ,
where the last line follows by the resolvent identity. By Cauchy’s theorem,
Z Z
z
−1 −1
z
2
(z − A) − (z + B) dz = 2
(z − A)−1 dz
∂Σν (1 + z) ∂Σν (1 + z)
540 16 Perturbations and sums of operators
= A(I + A)−2 x
It follows that
In its stated form, the proposition will be useful in the proof of Theorem
16.3.10. It is clear from the proof that the proposition could be stated with
%(A), %(B), and R(λ, C) replaced by more general operators φ(A), ψ(B), and
f (A) under suitable conditions on the functions φ, ψ, and f . We leave the
details to the interested reader.
Proof. It has already been observed that every λ 6∈ Σmax{νA ,νB } belongs to
the resolvent set of C, and by (16.18) (using the notation introduced there)
we have R(λ, C) = λR(λ, A)R(λ, B) + gλ (A, B), where
zw 1 λ
gλ (z, w) := = − .
(λ − (z + w))(λ − z)(λ − w) (λ − (z + w)) (λ − z)(λ − w)
(16.20)
%(A)R(λ, C)%(B)
Z Z
1
= %(z)%(w)R(z, A)R(λ, C)R(w, B) dw dz,
(2πi)2 ∂ΣνA ∂ΣνB
we see that this results in the sum of three integrals, where (I) corresponds
to the contribution λR(λ, A)R(λ, B), and (II) and (III) correspond to the
splitting of gλ by (16.20). By a simple computation involving Fubini’s theorem
and Cauchy’s theorem, the integrals (I) and (III) cancel, and the integral (II)
equals the one in the statement of the lemma.
with C = kA(A + B)−1 k. From this one also obtains the estimate
and together these estimates give (16.16), with implied constant 1 + 2C.
In what follows, A always denotes a sectorial operator on a Banach space
X, and we fix ω(A) < σ < π. Let A be a closed sub-algebra of L (X) resolvent
commuting with A, i.e.,
where ω(A) < ν < σ. The resulting operator is independent of the particular
choice of ν, and it satisfies
Mσ,A
kf (A)k 6 kF kH 1 (Σσ ;A ) ,
π
where Mσ,A = supλ∈{Σσ kλR(λ, A)k.
As in Proposition 10.2.2, this calculus is multiplicative and satisfies the
following convergence property: if Fn , F ∈ H 1 (Σσ ; A ) are uniformly bounded
and satisfy Fn (z)x → F (z)x for all z ∈ Σσ and x ∈ X, then for all g ∈ H 1 (Σσ )
we have
is R-bounded.
16.3 Sum-of-operator theorems 543
Parts (2) and (3) are analogues of the corresponding results in Theorem
10.2.13 and the proofs are similar. The proof of (1), which is the non-trivial
part of the theorem, is based on an extension of Lemma 10.3.13, which states
that if A is a sectorial operator on a Banach space X and F ∈ H 1 (Σσ ; A ) is
given, with ω(A) < σ < π, then for all ω(A) < ν < σ we have
Z
1 dz
F (A) = z 1/2 F (z)φz (A) , (16.21)
2πi ∂Σν z
where φz (λ) := λ1/2 /(z − λ). The proof is identical to that of Lemma 10.3.13;
all one needs to do is to replace H 1 (Σσ ) by H 1 (Σσ ; A ) throughout, and
so is the justification of the well-definedness of the operators φz (A) and the
convergence of integral on the right-hand side of (16.21).
We also need the following strengthening of Lemma 10.3.8:
Proof of Theorem 16.3.4. Let ω(A) < ν < σ < π and let F be as in (1). As in
the proof of Theorem 10.3.4(3), for x ∈ D(A) ∩ R(A) and F ∈ H 1 (Σσ ; A ) ∩
RH ∞ (Σσ ; A ) we find
X X 1 Z 2
−iν/2 dt
F (A)x = e F (e−iν 2j t)φe−iν (t−1 2−j A)x ,
=±1
2πi 1 t
j∈Z
and in the operator-valued Dunford calculus the operators G(A) and (F G)(A)
are given by
We have already observed in Proposition 16.3.1 that (16.22) implies the closed-
ness of A + B. The standard sectoriality of A + B now follows from Theorem
16.3.2.
546 16 Perturbations and sums of operators
Step 2 – We now assume that A and B are densely defined, but not nec-
essarily standard sectorial. Then the operators Aε := A + ε and Bε : +B + ε
are standard sectorial, and we may apply the above reasoning with Fε (z) :=
Bε (z +Bε )−1 and Gε (z) := ζn (z)2 (z +Bε )ζn (Bε )2 . This results in the estimate
λR(λ, A + B)
Z Z
1
= λ2 R(λ, A)R(λ, B) + fλ (z, w)R(z, A)R(w, B) dz dw.
(2πi)2 ∂Στ ∂Σσ
Outside the closure of Σσ+τ the operators λ2 R(λ, A)R(λ, B) are R-bounded,
by the R-sectoriality of A (which follows from the second part of Theorem
10.3.4) and B (by assumption). The operators corresponding to the Dunford
integral with fλ are R-bounded by Theorem 8.5.2; the integrability properties
required to apply theorem have already been observed in the proof of Theorem
16.3.2 (see (16.19)).
Since standard sectorial operators with a bounded H ∞ -calculus on a Banach
space with the triangular contraction property are R-sectorial with ωR (A) =
ωH ∞ (A) (see Corollary 10.4.10), we have the following corollary.
Corollary 16.3.7. Let A and B be resolvent commuting densely defined (re-
spectively, standard) sectorial operators with bounded H ∞ -calculi satisfying
ωH ∞ (A) + ωH ∞ (B) < π on a Banach space with the triangular contraction
property. Then A + B is a densely defined (respectively, standard) sectorial
operator and (16.22) holds.
is R-bounded.
Ψ f (λ1 , ·, . . . , ·) = f (λ1 , A2 , . . . , An ) ∈ T
for all λ1 ∈ Σσ1 .
λ1 7→ f (λ1 , A2 , . . . , An )
1 n−1 Z Z Yn
= ... f (λ1 , . . . , λn )R(λj , Aj ) dλ2 . . . dλn
2πi ∂Σν2 ∂Σνn j=2
for every fixed λ1 ∈ Σσ1 and all x ∈ X. Now apply the convergence property
of Φ1 to Fm (λ) = fm (λ, A2 , . . . , An ) and F (λ) = f (λ, A2 , . . . , An ).
The final R-boundedness assertion follows directly from the final assertion
of Theorem 16.3.4.
As an application we have the following variant of Corollary 16.3.7. This is
result is actually true for Banach space X with the triangular contraction
property; we refer to the Notes for a discussion of this fact.
For the proof we need a technical proposition. For the sake of its formulation,
the joint Dunford calculus of two resolvent commuting sectorial operators A
and B will be denoted by ΦA,B : f 7→ f (A, B), for functions f ∈ H 1 (Σσ ) ×
H 1 (Στ ). Likewise, the operator-valued Dunford calculus of A will be denoted
by ΦA : F 7→ F (A), for operator-valued functions F ∈ H 1 (Σσ ; A ) where A
is set of operators resolvent commuting with A.
which equals
but also
Z 1 Z
1
= %(z) %(w)f (z + w)R(w, B) dw R(z, A) dz
2πi ∂ΣνA 2πi ∂ΣνB
= ΦA (z 7→ %(z)f (z + B)%(B)).
Proof of Theorem 16.3.10. Since X has Pisier’s contraction property, A and
B admit R-bounded operator-valued H ∞ -calculi by Theorem 10.3.4. Choose
ω(A) < σA < π and ω(B) < σB < π such that σA + σB < π, and let
max{σA , σB } < σ < π. For f ∈ H ∞ (Σσ ), the function
F (z, w) := f (z + w)
belongs to H ∞ (ΣσA ) × H ∞ (ΣσB ). Since B has an R-bounded H ∞ -calculus,
the set {F (z, B) : z ∈ ΣσA } is an R-bounded subset of the set A of bounded
operators resolvent commuting with A. Applying the operator-valued calculus
of A we obtain a bounded operator F (A, B) on D(A) ∩ R(A). By Proposition
16.3.11,
%(A)f (A + B)%(B) = %(A)F (A, B)%(B),
where %(z) = z/(1 + z)2 . Since %(A) is injective and %(B) has dense range (we
assumed that A and B are standard sectorial), we conclude that f (A + B) =
F (A, B) is a bounded operator on D(A) ∩ R(A). The bound kf (A, B)k .
kf kH ∞ (Σσ ) follows by tracing the steps of the proof.
16.3 Sum-of-operator theorems 551
The main result of this section (Theorem 16.3.13) provides a version of The-
orem 16.3.6 in which no assumption on the Banach space is needed, the as-
sumption on a A is weakened to sectoriality, and the assumption on B is
strengthened to having an absolute calculus.
We denote
ωabs (A) := inf σ ∈ (0, π) : A admits an absolute calculus on Σσ .
Mν,A
kh(tA)g(tA)xk 6 kF k∞ kkg(tA)yk.
π
By the absolute calculus, this implies
Mν,A Mν,A
kF (A)yk = kxk 6 M kF k∞ kyk.
π π
By the same argument as in the proof of Theorem 10.2.13, this proves the
first assertion.
The second assertion follows by taking F (λ) = f (λ)IX .
Comparing this result with the Dore–Venni theorem, where both A and B are
assumed to have bounded imaginary powers, we observe that here, bounded-
ness of the imaginary powers is imposed only on A.
16.3 Sum-of-operator theorems 553
In this subsection and the next, we show connect the absolute calculus with
the theory of real interpolation. The crucial observation is contained in the
following theorem.
Theorem 16.3.16 (Lp -bounds imply absolute calculus). Let A be a sec-
torial operator in a Banach space X and let ω(A) < σ < π. Let 1 6 p 6 ∞,
and suppose that there exist φ ∈ H 1 (Σσ ) ∩ H ∞ (Σσ ) such that
kxkφ,p := t 7→ φ(tA)x Lp (R+ , dt
, x ∈ D(A) ∩ R(A),
t ;X)
with Z ∞
dt
G(λ) := g(tλ)R(λ, A)ψ(tA) .
0 t
Applying Young’s inequality for L1 (R+ , dt
t ) twice (after parametrising ∂Σν
−1
and substituting s 7→ s ), we obtain that φ(·A)x ∈ Lp (R+ , dt
t ; X) and
Mν,A
φ(·A)x Lp (R+ , dt
6 kφkH 1 (Σσ ) kgkH 1 (Σσ ) ψ(·A)x Lp (R+ , dt
.
t ;X) π t ;X)
Now that we know that t 7→ φ(tA)x belongs to Lp (R+ , dt t ; X), the opposite
norm estimate is obtained by reversing the roles of φ and ψ. This proves the
theorem for x ∈ D(A) ∩ R(A).
For x ∈ D(A) ∩ R(A) the result follows by approximation, noting that
ψ(tA)xkLp (R+ , dt ;X) 6 Ckx by a closed graph argument.
t
6 g(·A)y Lp (R+ , dt
h f (·A)y Lp (R+ , dt
h kyk
t ;X) t ;X)
The theorem should be compared with the first part of Proposition K.4.1,
which asserts that If A is a sectorial operator in X, then
n o
(X, D(A))θ,p = x ∈ X : λ 7→ λθ kGR(λ, G)xk ∈ Lp (R+ , dλ
λ )
Lp ((0, 1), dt
t ; X) for every t > 0. Hence, as in the proof of Proposition 10.2.5
we have
Z 1
ds
h(tA)x = (f φ)(stA) . (16.26)
0 s
Next, noting the identities
Z ∞ Z ∞
ds ds
zg(z) = zf (sz)φ(sz) = s−1 f1 (sz)φ(sz) ,
1 s 1 s
We can now apply the first part of Hardy’s inequality (Lemma L.3.2) to obtain
that t 7→ t−θ σx (t) ∈ Lp (R+ , dt
t ; X) and
1
kt 7→ t−θ σx (t)kLp (R+ , dt ;X) .σ,A kt 7→ t−θ (f φ)(tA)xkLp (R+ , dt ;X)
t θ t
558 16 Perturbations and sums of operators
1
.σ,A kf kH 1 (Σσ ) kt 7→ t−θ φ(tA)xkLp (R+ , dt ;X) .
θ t
kt−θ K(t, x; X, D(A))kLp (R+ , dt ) .σ,A kxk + kt−θ φ(tA)xkL1 (R+ , dt ;X) .
t t
and therefore the second term can be estimated in the same way at the third
term appearing in the second line of (16.27).
Lemma 16.3.21. Under the assumptions of the theorem, for every α < 1 − θ
the norm
Z ∞ p dt 1/p zα
kxkα = t−θ ζα (tA)xk with ζα (z) = , (16.28)
0 t (1 + z)2α
Since the same estimate holds with s and t interchanged, this gives (16.29).
Second, for all s > 0, t ∈ [s, 2s], and x ∈ X we have
16.4 Notes
Section 16.1
The problem of defining the sum of two unbounded operators A and B can
be approached from various angles. Besides the direct approach of defining
A + B as (the closure of) the operator given on D(A) ∩ D(B) by (A + B)x =
Ax + Bx, which works well if A and B have commuting resolvents, various
other approaches can be taken. When A and B generate uniformly bounded
C0 -semigroups S and T respectively, conditions can be formulated in order
that the limit in the Trotter product formula
c(x, y) = (Cx|y).
Like always, subtle domain questions have to be taken care of. A detailed treat-
ment is given in Kato [1995]; for a gentle introduction see, e.g., Van Neerven
[2022].
Section 16.2
for all x in this common domain, then if one of the operators has a bounded
H ∞ -calculus, then so does the other. Notice that there are no smallness as-
sumptions here.
The basic idea of the proofs of Theorem 16.2.8, the comparison theorem
just quoted, and their variants is to use the relative bounded or the equivalence
of norms of (16.32) to show the equivalence of the discrete square function
norms X
x 7→ εj φ(2j A)x
j∈Z
562 16 Perturbations and sums of operators
with the corresponding ones for B. Here, φ is usually of the form φ(z) =
z α (1 + z)n with α < n. The two conditions (ii) and (iii) of Theorem 16.2.8
and (16.32) correspond to the two sides of the square function estimate.
In Kalton, Kunstmann, and Weis [2006] it is also explained how to use
these perturbation theorems to establish the boundedness of the H ∞ -calculus
for rather general classes of elliptic operators on H s,p (Rd ) or H s,p (D) for
smooth domains D ⊆ Rd with Lopatinskii–Shapiro boundary conditions. The
idea is to compare them to constant coefficients. A related approach is used in
Denk, Hieber, and Prüss [2003]. For more recent results the reader is referred
to the Notes of Chapter 17.
Let us mention two further topics related to the H ∞ -calculus and its per-
turbations.
for all x ∈ Lp (S). Given a second standard sectorial operator A on Lp (S), the
following R-boundedness condition expresses that A is “close” to B in terms
of the “ Littlewood–Paley pieces” φ(2j A) and φ(2j B):
R φ(s2j+k A)ψ(t2k B) : j ∈ Z . 2−βk (16.33)
for some β > 0 and all k ∈ Z and s, t ∈ [1, 2]. Kunstmann and Weis [2017]
contains the following theorem:
way one can, for example, derive the boundedness of the H ∞ -calculus of the
Stokes operator on the Helmholtz space Lp0 (D) from the boundedness of the
H ∞ -calculus of the Laplace operator on Lp (D), for bounded Lipschitz do-
mains D ⊆ Rd with d > 3 and | 21 − p1 | < 2d 1
(see Kunstmann and Weis
[2017]).
with (1 − θ)α + θβ = γ for α 6= β and θ ∈ (0, 1). This is not true anymore in
Banach spaces, where complex interpolation is related to boundedness of the
imaginary powers, rather than the boundedness of the H ∞ -calculus. However,
such an identification is possible with the help of the γ-interpolation method
introduced in the Notes to Section 15.3. It is shown in Kalton, Kunstmann,
and Weis [2006, Section 5.3] that a standard γ-sectorial operator A on a
Banach space X with non-trivial type has a bounded H ∞ -calculus if and only
if
(Ẋα , Ẋβ )γθ = Ẋδ
with (1 − θ)α + θβ = δ for α 6= β and θ ∈ (0, 1). Even when A does not
have a bounded H ∞ -calculus, the spaces (Ẋα , Ẋβ )γθ can be identified with
γ
certain square function spaces Hs,A which are defined as the completion of
m m
D(A ) ∩ R(A ), m > |s| + 1, with respect to the norm
for some φ ∈ H 1 (σσ ) such that z 7→ z −s φ(z) still belongs to H 1 (σσ ). Complete
proofs can be found in Kalton, Lorist, and Weis [2023].
Section 16.3
(see also Lancien, Lancien, and Le Merdy [1998, Remark 6.5] and Albrecht,
Franks, and McIntosh [1998]) that any sectorial operator on a Hilbert space
with a bounded H ∞ -calculus has a bounded operator-valued H ∞ -calculus.
In these papers the “right” method of proof for Theorem 16.3.4 was already
found, but the crucial ingredient of R-boundedness was still missing.
Theorem 16.3.6 is due to Kalton and Weis [2001]. In the next chapter, the
connections of Theorem 16.3.6 with maximal Lp -regularity will be discussed
in detail. As we will see in Volume IV, the operator-valued functional calculus
of Theorem 16.3.4 can be used to give a short proof for stochastic maximal Lp -
regularity; see Van Neerven, Veraar, and Weis [2015b]. In Clément and Prüss
[2001] it is shown that if A is an injective operator generating a bounded C0 -
group on a UMD space X, and B is an invertible closed linear operator in X
resolvent commuting with A such that ±iB is R-sectorial, then the operator
A + B with domain D(A) ∩ D(B) is closed and invertible. If B is also sectorial
with angle ω(B) < 12 π, then A + B is sectorial as well, and ω(A + B) < 21 π.
Theorem 16.3.9 is due to Lancien, Lancien, and Le Merdy [1998], Lancien
and Le Merdy [1998], who extend an earlier result of Albrecht [1994] on Lp -
spaces with 1 < p < ∞.
That Theorem 16.3.10 holds more generally for Banach spaces with the
triangular contraction property was shown by Le Merdy [2003].
The absolute functional calculus was introduced in Kalton and Kucherenko
[2010], where Theorems 16.3.13, 16.3.14, and 16.3.20 were proved. The defi-
nition of the absolute calculus may be a little off-putting if one is accustomed
to thinking in terms of spectral theory, but the benefits of this notion is con-
siderable:
• It implies an operator-valued H ∞ -calculus and sum-of-operators theorem
without the complexities of R-boundedness, just as in Hilbert spaces (see
Theorems 16.3.13 and 16.3.14).
• It leads to a simple sufficient condition for the abstract functional calculus
of a sectorial operator A in terms of the equivalence
Z ∞ dt 1/p
kxk h kφ(tA)xkp (16.34)
0 t
1 ∞
for functions φ, ψ ∈ H (Σσ ) ∩ H (Σσ ) satisfying the conditions of Theorem
16.3.19. Interestingly, this equivalence of norms remains true under somewhat
weaker conditions on φ and ψ; see Haase [2006, Theorem 6.4.2]. The proof
follows the lines of the equivalence of continuous square functions in Chapter
10, with simplifications due to the fact that various subtleties in the handling
of γ-norms can now be avoided. This more general version of the equivalence
of norms covers the function φ(z) = z/(z + 1) which is implicit in the first
part of Proposition K.4.1.
Corollary 16.3.22 is a classical result due to Dore [1999]. This result was
subsequently generalised to standard sectorial operators in Dore [2001], where
it was shown that such operators have a bounded H ∞ -calculus on (X, D(A) ∩
R(A))θ,p ; see also Kalton and Kucherenko [2010], who establish their absolute
functional calculus.
The present chapter is one of the central ones of this book project. Maximal
regularity provides a link between the general theory of operator-valued sin-
gular integrals and the theory of H ∞ -functional calculus with the regularity
theory for evolution equations. In some cases progress on these topics was
even motivated by applications to maximal regularity.
Maximal Lp -regularity for the Cauchy problem
(
u0 (t) + Au(t) = f (t), t ∈ I,
(17.1)
u(0) = 0,
ku0 kLp (I;X) + kAukLp (I;X) 6 Cku0 + AukLp (I;X) = Ckf kLp (I;X) . (17.2)
This a priori estimate is often crucial to fixed point methods for more general
classes of non-linear partial differential equations with A as their linear main
part. This will be explained in more detail in Chapter 18.
Unfortunately, maximal Lp -regularity for A may fail even if −A is the
generator of an analytic semigroup on Lq (0, 1) (see Section 17.4.c). Proving
maximal Lp -regularity usually requires quite sophisticated methods. In earlier
chapters, we have prepared for three approaches to maximal regularity:
• Operator-valued singular integral operators:
Formally using the variation of constants formula for the solution of (17.2),
we have
Z t
Au(t) = Ae−(t−s)A f (s) ds.
0
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 569
T. Hytönen et al., Analysis in Banach Spaces, Ergebnisse der Mathematik und ihrer
Grenzgebiete. 3. Folge / A Series of Modern Surveys in Mathematics 76,
https://doi.org/10.1007/978-3-031-46598-7_7
570 17 Maximal regularity
Since kAe−tA k . 1t (and often even h), it is clear that we have to deal
with singular integrals to obtain maximal Lp -regularity. The details of this
approach will be presented in Section 17.3.a.
L1loc (I; X)
It is clear from condition (ii) in the above definition that a strong solution u
is equal almost everywhere to a continuous function on I. In what follows, we
will always work with this version, which satisfies the identity of (ii) for all
t ∈ I and satisfies
u(0) = x.
If (0, T0 ) is a bounded interval contained in I and u is an strong solution of
(ACPx ) on I, then by taking the supremum over [0, T0 ] in (ii) we obtain
v 0 (s) = (λ + A)−1 S(t − s)(u0 (s) + Au(s)) = (λ + A)−1 S(t − s)f (s)
for almost all s ∈ (0, t). This implies v ∈ W 1,p (0, t; X) and, by Proposition
2.5.9,
Z t
−1
v(t) − (λ + A) S(t)x = v(t) − v(0) = v 0 (s) ds
0
Z t
−1
= (λ + A) S(t − s)f (s) ds.
0
where Rwe used the definition of a mild solution and Hille’s theorem; the iden-
t
tity A r S(s − r)x ds = −[S(t − r)x − x] which as used in this computation
is justified by Proposition K.1.7.
(2)⇒(3): Since u is almost everywhere differentiable with u0 ∈ L1loc (I; X),
and (λ + A)−1 S(t)x = S(t)(λ + A)−1 x is differentiable with derivative A(λ +
A)−1 S(t)x, (17.4) and Proposition 2.5.9 imply
574 17 Maximal regularity
A(λ + A)−1 u(t) = −(λ + A)−1 [u0 (t) − f (t)] almost all t ∈ I.
Combining this with A(λ + A)−1 u(t) = u(t) − λ(λ + A)−1 u(t) we find that
u ∈ D(A) almost everywhere and Au = −u0 + f ∈ L1loc (I; X).
(3)⇒(1): Since u ∈ D(A) almost everywhere and Au ∈ L1loc (I; X), (17.4)
implies
Z t Z t
−1 −1
(λ + A) Au(s) ds − A(λ + A) S(s)x ds
0 0
h Z t i
= −(λ + A)−1 u(t) − S(t)x − f (s) ds .
0
−1
Rt
It follows from this that A(λ+A) 0
S(s)x ds ∈ D(A), which in turn implies
Rt
that 0 S(s)x ds ∈ D(A). Applying λ + A on both sides, we obtain
Z t Z t Z t
Au(s) ds + A S(s)x ds = −u(t) + S(t)x + f (s) ds.
0 0 0
Rt
Since A 0
S(s)x ds = S(t)x − x, this shows that u is a strong solution.
The following proposition provides a large class of functions f : [0, T ] → X
for which the mild solution u = S ∗ f to (ACP0 ) is a strong solution.
Proposition 17.1.4. Let A be sectorial of angle < 21 π and let S be the ana-
lytic semigroup generated by −A. Then for all f ∈ C α ([0, T ]; X) with α > 0,
the mild solution u = S ∗ f to (ACP0 ) satisfies
u ∈ C([0, T ]; X) ∩ L∞ (0, T ; D(A)).
In particular, u is a strong solution to (ACP0 ).
Proof. In the discussion following Definition 17.1.2 we have already seen that
u ∈ C([0, T ]; X). In the remainder of the proof there is no loss of generality
in assuming that 0 < α < 1. For t ∈ [0, T ] we can write
Z t Z t
u(t) = S(t − r)(f (r) − f (t)) dr + S(t − r)f (t) ds =: u1 (t) + u2 (t).
0 0
Since by (K.4) we have sups∈(0,T ] ksAS(s)k < ∞, we find that u1 (t) ∈ D(A)
and
Z t
kAu1 (t)k 6 kAS(t − r)(f (r) − f (t))k dr
0
Z T
6 [f ]C α ([0,T ];X) rα−1 kAS(r)k dr.
0
From Proposition K.1.11 we see that u2 (t) ∈ D(A) and Au2 (t) = (I−S(t))f (t),
and thus
kAu2 (t)k 6 sup kI − S(t)k sup kf (t)k.
t∈[0,T ] t∈[0,T ]
17.2 Maximal Lp -regularity 575
kukLp (0,T ;X) 6 T (kAukLp (0,T ;X) + kf kLp (0,T ;X) ). (17.6)
Similarly,
kvkW 1,p (0,T ;D(A)) 6 kvkLp (0,T ;X) + kv 0 kLp (0,T ;X)
6 T kv 0 kLp (I;X) + kv 0 kLp (I;X) 6 (T + 1)kvk0 Ẇ 1,p (I;X) .
A
Thus the maximum of the two left-hand sides has this same upper bound.
17.2 Maximal Lp -regularity 577
for every bounded interval [0, T ] ⊆ I. ¯ Hence un (t) → u(t) for t ∈ [0, T ]
¯
and, [0, T ] ⊆ I being arbitrary, for t ∈ I. (Here we are thinking of con-
tinuous versions of the Sobolev functions in question.) On the other hand,
Aun ∈ Lp (I; X) is a Cauchy sequence, hence convergent to some limit
w ∈ Lp (I; X), and thus a subsequence converges almost everywhere. If t ∈ I
is such a point, then un (t) → u(t) while Aun (t) → w(t). Since A is closed, it
follows that u(t) ∈ D(A) and Au = w ∈ Lp (I; X). This shows that the limit
u ∈ 0 Ẇ 1,p (I; X) in fact belongs to 0 ẆA1,p (I; X). Since un → u in 0 Ẇ 1,p (I; X)
and Aun → w = Au in Lp (I; X), it follows that un → u ∈ 0 ẆA1,p (I; X).
We now come to the principal notion to be studied in this chapter. Here,
and in the remainder of this subsection, we assume that 1 6 p 6 ∞ unless
explicitly stated otherwise.
The least admissible constant in this definition will be called the maximal
reg
Lp -regularity constant of A on I and will be denoted by Mp,A (I).
p
If A has maximal L -regularity on I, then for bounded subintervals
(0, T ) ⊆ I, the estimate (17.5) implies that if uf is an Lp -solution of (ACP0 )
with f ∈ Lp (I; X), then its restriction to (0, T ) satisfies
reg
kuf kLp (0,T ;X) 6 T 1/p kuf kC([0,T ];X) 6 T (Mp,A (I) + 1)kf kLp (I;X) . (17.10)
is an isomorphism with
1 reg
kf kLp (I;X) 6 kM f k0 Ẇ 1,p (I;X) 6 (Mp,A (I) + 1)kf kLp (I;X)
2 A
The existence and uniqueness of the Lp -solution uf implies that the mapping
f 7→ uf is linear from Lp (I; X) into 0 ẆA1,p (I; X).
Conversely, given u ∈ 0 ẆA1,p (I; X), we have f := u0 + Au ∈ Lp (I; X) with
kf kLp (I;X) 6 ku0 kLp (I;X) + kAukLp (I;X) 6 2kuk0 Ẇ 1,p (I;X) ,
A
kvk0 Ẇ 1,p (I;X) 6 kvkLp (0,T ;D(A))∩W 1,p (0,T ;X) 6 (T + 1)kvk0 Ẇ 1,p (I;X) .
A A
Proof. This first bound is immediate by combining (17.8) and (17.12). For
the comparison of the spaces, the first estimate is clear, and the second one
is (17.8).
The next result gives another sufficient condition, also allowing infinite inter-
vals I, that Lp -solutions u do in fact belong to Lp (I; X).
kuk0 Ẇ 1,p (I;X) 6 kukLp (I;D(A))∩0 W 1,p (I;X) 6 (kA−1 k + 1)kuk0 Ẇ 1,p (I;X) .
A A
p p
In this situation, for all f ∈ L (I; X), the unique L -solution uf to (ACP0 )
belongs to this space and satisfies
reg
kuf kLp (I;D(A))∩W 1,p (I;X) 6 (kA−1 k + 1)(Mp,A (I) + 1)kf kLp (I;X) .
Proof. The boundary value v(0) = 0 is assumed in both spaces, and the first
norm estimate is evident. The second estimate follows by taking the maximum
of
kukW 1,p (R+ ;X) 6 kukLp (R+ ;X) + ku0 kLp (R+ ;X)
6 kA−1 k kAukLp (R+ ;X) + ku0 kLp (R+ ;X)
6 (kA−1 k + 1)kuk0 Ẇ 1,p (R+ ;X)
A
and
From these bounds and (17.12), the final claim concerning uf follows.
• I = (0, T ) is bounded;
• 0 ∈ %(A).
For f ∈ Lp (I; X) let uf denote the corresponding Lp -solution in Lp (I; D(A))∩
W 1,p (I; X). Then
1,p
0 ẆA (I; X) ' Lp (I; D(A)) ∩ 0 W 1,p (I; X);
kvk0 Ẇ 1,p (I;X) 6 kvkLp (I;D(A))∩W 1,p (I;X) 6 (min{kA−1 k, T } + 1)kvk0 Ẇ 1,p (I;X) ;
A A
Proof. By Proposition 17.2.2, we have uf ∈ 0 ẆA1,p (I; X). The existence and
uniqueness of uf for every f ∈ Lp (I; X) guarantees that M : f 7→ uf is a
linear mapping from Lp (I; X) to 0 ẆA1,p (I; X). Since A is closed, 0 ẆA1,p (I; X)
is a Banach space by Proposition 17.2.3.
To prove that M is bounded, by the closed graph theorem, it suffices to
check that it is closed. To this end, let fn → f in Lp (I; X) and un := M fn → v
in 0 ẆA1,p (I; X). Since u0n +Aun = fn , taking limits gives v 0 +Av = f . It follows
that v is an Lp -solution corresponding to f . The uniqueness of Lp -solutions
therefore gives v = M f .
Thus M is closed, hence bounded, and therefore
AM f = M Af
(I + D)Cf = f.
Indeed, for f ∈ Lp (I; X) the product rule and an integration by parts give
h Z t i
DCf (t) = ∂t e−t es f (s) ds
0
Z t
−t
= −e es f (s) ds + f (t) = −Cf (t) + f (t).
0
where the first identity (valid more generally for all g ∈ Lp (I; X)) is a re-
statement of the equation that a solution M g with datum g must satisfy,
and the second one follows by noting that g ∈ 0 ẆA1,p (I; X) has the correct
boundary value g(0) = 0 by definition, and evidently solves the equation with
datum (A + D)g, so it must be the Lp -solution by uniqueness. Therefore, if
we can prove that M and D commute in a suitable sense, the desired result
will follow.
First observe that, by Hille’s theorem (Theorem 1.2.4),
from which it follows that v = M (I + D)−1 f . This gives the required result.
With these preparations out of the way, we can turn to the proof of the
lemma. It remains to show that for f ∈ Lp (R+ ; D(A)) one has M Af = AM f .
By the previous observations,
(1 + D)−1 M Af = M (1 + D)−1 Af
= M A(1 + D)−1 f
= M (A + D)(1 + D)−1 f − M D(1 + D)−1 f
= (1 + D)−1 f − D(1 + D)−1 M f
= (1 + D)−1 (f − DM f )
= (1 + D)−1 (AM f ).
1
where f ∈ Lloc (I; X) and x ∈ X; as before, I is either a bounded interval
(0, T ) or R+ = (0, ∞). An obvious necessary condition for the existence of
an Lp -solution u is that f should be in Lp (I; X); this is immediate from the
definition of an Lp -solution, which imply that u0 and Au belong to Lp (I; X),
and we then have the trivial bound
k · k(X,D(A))1− 1 ,p h k · kTr
(X,D(A))
p 1− 1 ,p
p
of Theorem L.2.3.
kxkTr
(X,D(A)) 1 6 (1 + 1/T ) max kukLp (0,T ;D(A)) , kukW 1,p (0,T ;X) ;
1− ,p
p
and
1
kt 7→ t1− p v(t)kLp (R+ , dt ;D(A)) = kvkLp (R+ ;D(A))
t
1,p ¯
are finite. By assumption, we know that x = u(0), where u = uf ∈ Wloc (I; X);
however, it might be that I is only a finite interval, and even if I = R+ , the
norm kukLp (R+ ;X) 6 kukLp (R+ ;D(A)) need not be finite.
To fix both problems at once, let (0, T ) ⊆ I be finite, and let φ(t) :=
(1 − t/T )+ . Then v(t) := φ(t)u(t) is defined on all R+ , and it still satisfies
v(0) = u(0) = x. Since v 0 = φ0 u + φu0 , where |φ0 | 6 T1 1(0,T ) and |φ| 6 1(0,T ) ,
we obtain
kvkLp (R+ ;D(A)) 6 kukLp (0,T ;D(A))
6 kukLp (0,T ;X) + kAukLp (0,T ;X)
and
17.2 Maximal Lp -regularity 585
1 1
kv 0 kLp (R+ ;X) 6 kukLp (0,T ;X) + ku0 kLp (0,T ;X) 6 + 1 kukW 1,p (0,T ;X)
T T
The Lp -norms of Au and u0 above are clearly finite for an Lp -solution to
(ACPx ), and so is the Lp -norm of u on a finite (0, T ) ⊆ I by
Z t
kukLp (0,T ;X) = t 7→ x + u0 (s) ds
0 Lp (0,T ;X)
The claimed bound dealing with a bounded (0, T ) ⊆ I follows from this at
once. The case of I = R+ follows by taking the limit T → ∞, noting that the
term T1 kukLp (0,T ;X) → 0 in this case.
We first prove this for 1 < p 6 ∞; the case p = 1 will be established below
Lemma 17.2.22; there, also the precise estimate for kAukL1 (I;X) is stated (as it
turns out, the constant 3 gets replaced by the supremum over I of the norms
of the semigroup generated by −A).
Proof for 1 < p 6 ∞. Uniqueness of Lp -solutions follows from maximal Lp -
regularity. For the proof of existence we first consider the case 1 < p < ∞.
By Theorem L.2.3, for every ε > 0 we can find a function g ∈ Lp (R+ ; D(A)) ∩
W 1,p (R+ ; X) such that g(0) = x and
v 0 + Av = f + g 0 + Ag, v(0) = 0.
which maps the data (x, f ) to the solution u; this remains correct for I = R+
if 0 ∈ %(A). This extends Corollary 17.2.9 to the case of non-zero initial data.
u(t) := eλt x − x,
f (t) := u0 (t) + Au(t) = (λ + A)eλt x − Ax.
keλ(·) − 1kLp (I) kAxk 6 C keλ(·) kLp (I) k(λ + A)xk + T 1/p kAxk .
(17.15)
By scaling and using keλ(·) − 1kLp (I) > keλ(·) kLp (I) − k1kLp (I) , we find that
(keλT (·) kLp (0,1) − 1)kAxk 6 C keλT (·) kLp (0,1) k(λ + A)xk + kAxk . (17.16)
Let r0 > 0 be such that ker0 (·) kLp (0,1) − 1 > 2C. For <λ > r0 /T , we obtain
Put
Z
Rλ x := 2<λ e−λt MI (fλ ⊗ x)(t) dt. (17.18)
I
If we could let T → ∞ then this would give ARλ x = −λRλ x + x. This is not
possible, however, since I = (0, T ) is fixed. Instead, we will take <λ so large
that the remainder term has norm 6 1/2. To this end let
Suppose that <λ > 1/T . By (17.10), (17.17), and the fact that for any 0 <
α 6 1 the function t 7→ tα e−t/2 is bounded above on R+ by (2α/e)α 6 1,
0
kQλ xk 6 2<λ e−<λT T 1/p (C + 1)kfλ ⊗ xkp + e−2<λT kxk
0
6 2(<λT )1/p e−<λ T (C + 1)kxk + e−2 kxk (17.23)
− 21 <λ T
6 2e (C + 1)kxk + e−2 kxk.
In combination with Lemma G.1.4, this proves that large enough translates
of −A are sectorial of angle < 21 π. The results of Appendix K imply that
these translates, and hence −A itself, generate analytic semigroups on X in
the sense of Definition K.1.2.
Proof of Theorem 17.2.15(2) with non-sharp constant. The above proof for
I = (0, T ) can be repeated for I = R+ , with MI replaced by MR+ , up to
equation (17.22) which now reads
We can still apply (17.10) to this equation, and in combination with (17.17)
we arrive at the bound given in (17.23). This time we can pass to the limit
T → ∞ and arrive at the identity ARλ x = −λRλ x + x. Therefore, Rλ is a
right-inverse of λ + A. To see that it is a left-inverse it suffices to check that
ARλ = Rλ A. The latter follows from (17.19) and Lemma 17.2.12.
Sectoriality follows from (17.20) with a slightly worse bound than stated
in (17.14). The bound will be further improved below.
Having analytic semigroups available, we can now deduce maximal regularity
on sub-intervals by using the equivalence result for strong and mild solutions
described in Proposition 17.1.3. Recall from Definition 17.1.2 that if S =
(S(t))t>0 is a locally bounded strongly measurable semigroup S on X, then
for x ∈ D(A) the continuous function u : [0, ∞) → X defined by
Z t
u(t) := S(t)x + S(t − s)f (s) ds, t ∈ I,
0
Proof. (2) ⇒ (1) is clear from Lemma 17.2.16, and the estimate “>” follows
from (17.25).
(1) ⇒ (2): Let f ∈ Lp (R+ ; X). For every 0 < T < ∞ there exists a unique
p
L -solution uT to (ACP0 ) on (0, T ). By uniqueness, uT |[0,S] = uS for every
S 6 T . Therefore, we can construct a function u : [0, ∞) → X such that
u|[0,T ] = uT for every 0 < T < ∞. Then u is a strong solution to (ACP0 ) on
R+ . Moreover, uniqueness follows by taking restrictions to [0, T ]. Since
kAukLp (0,T ;X) 6 Ckf kLp (0,T ;X) 6 Ckf kLp (R+ ;X) ,
reg
where C = supT >0 Mp,A (0, T ), the desired result follows from the monotone
convergence theorem by passing to the limit T → ∞.
Now that we know that maximal Lp -regularity of A implies that −A generates
an analytic semigroup it is of interest to relate Lp -solutions to mild solutions.
p
is well defined, maps F into L (I; X), and there is a constant C > 0 such
that
kV f kLp (I;X) 6 Ckf kLp (I;X) , f ∈ F.
In this situation, V uniquely extends to a bounded operator on Lp (I; X) such
that
V f = Auf ,
where uf is the mild and Lp -solution to (ACP0 ) associated with f , and the
reg
least admissible constant C in the above inequality coincides with Mp,A (I).
592 17 Maximal regularity
Remark 17.2.20 (On the choice of the space F ). In the above formulation it
is implicit that one should have
Z t
S(t − s)f (s) ds ∈ D(A)
0
for almost all t ∈ I. By Proposition 17.1.4 this holds for instance for F =
C α (I; X) with α > 0 (which is dense if p ∈ [1, ∞)). In the important case
when D(A) = X, it can be useful to take F = Lp (I; D(A)) (which is dense
if p ∈ [1, ∞)); in this case the operator A can even be pulled through the
integral. This choice doesn’t work for p = ∞ even when D(A) = X, because
the space L∞ (I; D(A)) in general fails to be dense in L∞ (I; X). As we will
see in Section 17.2.f, this problem can be circumvented by considering the
equivalent notion of maximal C-regularity.
Remark 17.2.21 (Connection with singular integrals). For analytic semigroups
S with generator −A, the bound
kAS(t)k = O(1/t) as t ↓ 0
(apply (K.4) to λ + A for λ large enough) shows that in those cases where
it is possible to pull A through the integral in the definition of V , a singular
convolution integral is obtained. This aspect of the theory will be further
studied at later stage in this chapter (see, for example, Theorems 17.2.31,
17.2.39, and Section 17.3).
Proof of Theorem 17.2.19. ‘Only if’: Every f ∈ Lp (I; X) gives rise to a unique
Lp -solution uf which, by Proposition 17.1.3, equals the mild solution. There-
fore, the mapping f 7→ Auf = V f extends to a bounded operator on Lp (I; X)
by maximal Lp -regularity.
‘If’: Let uf be the mild solution associated with f ∈ F . The assumptions
on F imply that uf takes values in D(A) and that Au ∈ Lp (I; X), so uf is an
Lp -solution by Proposition 17.1.3. Since V f = Auf for all f ∈ F , maximal Lp -
regularity follows from the boundedness of V , closedness of A, and Proposition
17.2.10.
We continue with a useful special property of maximal L1 -regularity which will
be needed several times below: it leads to a characterisation of maximal L1 -
regularity in Theorem 17.3.11, and it allows us to give a proof of Proposition
17.2.14 for p = 1, and it will be used in Proposition 17.2.32.
Lemma 17.2.22. Let A have maximal L1 -regularity on I, and let S be the
analytic semigroup generated by −A. Then
Z
reg
kAS(t)xk dt 6 M1,A (I)CI,S kxk, x ∈ X, (17.26)
I
Remark 17.2.23. A little more can be said. For t > s > 0 with t ∈ I we have
Z t Z t Z t
kS(s)x − S(t)xk = A S(r)x dr = AS(r)x dr 6 kAS(r)xk dr
s s s
By this estimate and the Bochner integrability of AS(·), the limit S(0+)x :=
limt↓0 S(t)x exists and it satisfies
Z t
kS(0+)x − S(t)xk 6 kAS(r)xk dr.
0
594 17 Maximal regularity
In order to prepare for some results in the next sections, we conclude the
present section by introducing a constant related to the maximal regularity
constant with an additional parameter λ; in applications, this additional flex-
ibility can sometimes be exploited.
From Theorem 17.2.19 we recall that if A is a linear operator in X such
that −A generates an analytic semigroup S on a Banach space X, then A has
maximal Lp -regularity on some bounded interval (0, T ) if and only if
Z t
V f (t) := A S(t − s)f (s) ds, t ∈ [0, T ],
0
is well defined for all functions f in some dense subspace of Lp (I; X) and
extends to a bounded operator V on Lp (I; X). When I = (0, T ) is a bounded
interval, in these circumstances for all λ ∈ C the operator Vλ defined by
Z t
Vλ f (t) := A Sλ (t − s)f (s) ds, t ∈ [0, T ],
0
has the same properties, where Sλ (t) := e−λt S(t) is the rescaled semigroup.
Indeed, in this case we have
kVλ f kLp (I;X) 6 max{1, e−<λT }kV kL (Lp (I;X)) ks 7→ e−λs f (s)kLp (I;X)
If the mild solution uλ exists and takes values in D(A) for all for all f ∈ F ,
where F is some dense subspace of Lp (I; X), then the mapping f 7→ Vλ f
defined by (17.27) uniquely extends to a bounded linear mapping on Lp (I; X)
and we may define
reg
Mp,A,λ (I) := kVλ kL (Lp (I;X)) .
reg reg
Thus if A has maximal Lp -regularity on I, then Mp,A,0 (I) = Mp,A (I). For
other values of λ, we will derive various bounds in Theorem 17.2.24 and Propo-
reg reg
sition 17.2.27. We should emphasise that, in general, Mp,A,λ (I) 6= Mp,A+λ (I);
17.2 Maximal Lp -regularity 595
Rt
the latter would be the norm the operator Ṽλ f (t) := (A+λ) 0 Sλ (t−s)f (s) ds,
while in (17.27), only the semigroup is rescaled, but not the operator A.
If α, β ∈ R and Vα extends to a bounded operator on Lp (I; X), then so
does Vα+iβ and
kVα+iβ f kLp (I;X) = kVα eiβ· f kLp (I;X) , f ∈ Lp (I; X). (17.29)
reg reg
Clearly, this implies Mp,A,α+iβ (I) = Mp,A,α (I) for any α, β ∈ R.
reg reg 2M
Mp,A,λ (R+ ) 6 2 max{e<λT /2 , e−<λT }Mp,A (I) + . (17.31)
T (<λ − µ)
reg reg 2M
Mp,A (R+ ) 6 2Mp,A (I) + . (17.32)
Tω
596 17 Maximal regularity
Two remarks are in order. First of all, in (17.31) no claim is made with
regard to maximal Lp -regularity on R+ ; this is done only in (17.32). Secondly,
in situations where A has maximal Lp -regularity on every bounded interval
(0, T ) one can optimise the choice of T in (17.31) by letting it depend on the
parameter λ (see, for instance, the proof of Proposition 17.2.27).
Proof. Let Sλ (t) = e−λt S(t), where λ = α + iβ ∈ C with α > µ. For f ∈
Lp (R+ ; X) the expression for the mild solution
uf,λ := Sλ ∗ f
defines a continuous function uf,λ : [0, ∞) → X. We will show that uf,λ takes
values in D(A) almost everywhere and that
where in the last step we applied Young’s inequality. This proves (17.33).
We conclude this section by showing the necessity of the invertibility assump-
tion in Proposition 17.2.8.
If I = (0, T ) is bounded and <λ > 0, we can apply either (1) or (2), but the
first estimate, which is independent of λ, will be sharper. Both cases can be
combined into
reg
kAukLp (I;X) 6 Mp,A (I)e(<λ)− T kf kLp (I;X) ,
2
If <λ > 0, the choice T = <λ gives the bound
reg reg
Mp,A,λ (R+ ) 6 2eMp,A (R+ ) + M =: K, <λ > 0. (17.35)
and hence
and hence kAxk 6 Cµ k(µ + 1 + A)xk for x ∈ D(A). Therefore, using the
λ-analogue of (17.25) and (17.35), for all λ ∈ C with <λ > µ + 1 we obtain
reg reg
Mp,A,λ (I) 6 Mp,A,λ (R+ )
reg
6 Cµ Mp,A+µ+1,λ−µ−1 (R+ )
reg
6 Cµ (2eMp,A+µ+1 (R+ ) + M ) =: Kµ .
kuj kLp (I;X) 6 kSλj kL1 (0,T ;L (X)) kf kLp (I;X) , j ∈ {1, 2}.
Next let I = R+ and <λ > 0. The previous considerations apply with
(<λ)− = 0; the boundedness of Lλ now follows from Young’s inequality, which
gives the estimate
M |λ|
kλLλ k 6 kt 7→ λM e−tλ kL1 (R+ ) =
<λ
where kS(t)k 6 M for all t > 0. This gives the estimate
satisfies
Extrapolation of integrability
Note that we here specifically apply the Definition 11.2.1(4) with I = (0, T )
or I = R+ ; these are a one-dimensional cube and a quadrant, respectively.
By Theorem 17.2.15, −A generates an analytic semigroup (S(t))t>0 , which
is bounded if I = R+ . By (K.4), C := supt∈I ktAS(t)k is a finite quantity.
By Proposition K.3.1, C > 1/e unless A is bounded, hence 4C > 1, and log+
may be replaced by log in the Dini and Hörmander bounds of the lemma.
Proof. (1): By Theorem 17.2.19 the operator V extends to a bounded operator
on Lp0 (I; X). To show that V has kernel K, let f ∈ Lp0 (I; X) be compactly
supported and let F = supp(f ) be its support. Fix t ∈ I such that δ :=
dist(t, F ) > 0. Then
Z Z
kK(t, s)f (s)k ds 6 sup kAS(t − s)k kf (s)k ds < ∞.
I s∈[0,t−δ] F
(2): It is clear that for s, t ∈ I, kK(s, t)k 6 C/|s − t|, thus we can take
cK = C in Definition 11.3.1. Moreover, for all s, t ∈ I and r ∈ {s, t}, we have
Duality
The following duality result for maximal Lp -regularity implies assertion (5) of
Theorem 17.2.26. As before, we assume that I = (0, T ) is a bounded interval
or I = R+ .
For the proof of this theorem we need the following lemma, which is a vari-
ant of Corollary 1.3.2 except for the fact that we interchange the roles of X
and X ∗ for the sake of the application we have in mind; the corresponding
version without this interchange is true as well, with the same proof (or as an
application, by passing through the bidual).
In (17.40), we deliberately avoid denoting the space by Cb (I; Y ) for the fol-
lowing reason: if Y happens to have a norm k kY of its own, the notation
Cb (I; Y ) refers to functions f : I → Y that are continuous with respect to
k kY , while in (17.40), we think of continuity relative to the norm k kX of the
ambient space X.
Proof. We will show that kgkL1 (I;X ∗ ) = supf |hf, gi|, where the supremum is
taken over all f in (17.40) of norm 6 1. The estimate “>” is clear. To prove
the converse estimate, the density of simple functions (Lemma 1.2.19) and the
regularity of the LebesgueP measure imply that it suffices to consider simple
n
functions of the form g = j=1 1Ij x∗j , where I1 , . . . , In are disjoint bounded
intervals contained in I, and x∗1 , . . . , x∗n are elements of X ∗ . Let ε ∈ (0, 1)
R arbitrary and choose functions ϕj ∈ Cc (I) such that 0 6 ϕj 6 1Ij and
be
ϕ (t) dt > (1 − ε)[IJ | for each j. Choose yj ∈ Y of norm kyj k 6 1 such
Ij j
Pn
that kx∗j k 6 (1 + ε)hyj , x∗j i. Then f := j=1 ϕj yn has norm 6 1 and
n
X Z
hf, gi = hyj , x∗j i ϕj (t) dt
j=1 I
n
1−εX 1−ε
> |Ij |kx∗j k = kgkL1 (I;X ∗ ) .
1 + ε j=1 1+ε
Proof of Proposition 17.2.32. (1)⇒(2) for bounded intervals I = (0, T ): Let
A have maximal Lp -regularity on I = (0, T ). By Theorem 17.2.15 A is sec-
torial of angle < π/2 and −A generates an analytic semigroup (S(t))t>0 . As
a consequence (see Remark K.1.12), −A∗ generates an analytic semigroup as
well, and it is given by (S ∗ (t))t>0 , where S ∗ (t) := (S(t))∗ . To prove that A∗
0
has maximal Lp -regularity on I we will use Theorem 17.2.19.
606 17 Maximal regularity
We begin with the case p ∈ (1, ∞). Fix f ∈ Lp (0, T ; D(A)) (where, to be
sure, we think of D(A) as a Banach space with respect to the graph norm
kxkD(A) := kxkX + kAxkX ) and g ∈ C 1 ([0, T ]; X ∗ ), and write fe(t) := f (T − t)
and ge(s) = g(T − s) for s, t ∈ I. Then the function S ∗ fe takes values in D(A),
and by Proposition 17.1.4 we have S ∗ ∗ g ∈ C([0, T ]; X) ∩ L∞ (0, T ; D(A)). By
Fubini’s theorem and suitable substitutions,
Z TD Z t E
∗ ∗
hf, A S ∗ gi = Af (t), S ∗ (t − s)g(s) ds dt
0 0
Z T D Z s E
= A S(s − t)fe(t) dt, ge(s) ds = hAS ∗ fe, gei,
0 0
where the brackets in the first and last term refer to the duality between
0
Lp (I; X) and Lp (I; X ∗ ), the latter viewed as a closed subspace of the dual of
p
L (I; X) (see Proposition 1.3.1). Therefore,
Since the embedding Lp (I; D(A)) ,→ Lp (I; X) is dense (recall that D(A) was
assumed to be densely defined in Proposition 17.2.32 that we are proving),
this estimate extends to all f ∈ Lp (I; X) by density. Thus by Proposition
1.3.1
reg
kA∗ S ∗ ∗ gkLp0 (I;X ∗ ) 6 Mp,A (I)kgkLp0 (I;X ∗ ) .
0
Therefore, A∗ has maximal Lp -regularity, and Mpreg reg
0 ,A∗ (I) 6 Mp,A (I) by The-
1 ∗ p0 ∗
orem 17.2.19 and the density of C ([0, T ]; X ) in L ([0, T ]; X ).
In the case p = ∞, the above proof can be modified by using function from
the subspace {f ∈ C([0, T ]; X) : f takes values in D(A)}, which is norming
0
for Lp (0, T ; X ∗ ) = L1 (0, T ; X ∗ ) by Lemma 17.2.33.
Finally, let p = 1. Let g ∈ L∞ (I; X ∗ ) and f ∈ L1 (I; D(A)). By Theorem
17.2.31, A has maximal Lq -regularity on I for all q ∈ (1, ∞). Therefore, by the
previous proof, A∗ has maximal Lq -regularity for all q ∈ (1, ∞). In particular,
Theorem 17.2.19 implies that S ∗ ∗ g takes values in D(A∗ ) almost everywhere
on I. As before one can check that hf, A∗ S ∗ ∗ gi = hAS ∗ fe, gei, and thus
reg
|hf, A∗ S ∗ ∗ gi| 6 M1,A (I)kf kL1 (I;X) kgkL∞ (I;X ∗ ) .
complete the proof, first let p ∈ [1, ∞). Fix functions f ∈ Lp (I; D(A)) and
0 0
g ∈ Lp (I; X ∗ ), and define fe and ge as before. By the maximal Lp -regularity
of A∗ on I, the convolution S ∗ ∗ ge takes values in D(A∗ ) almost everywhere
on (0, T ). Using Fubini’s theorem and the same substitutions as before, we
obtain
Z TZ t
hAS ∗ f, gi = hS(t − s)Af (s), g(t)i ds dt
0 0
Z T Z s
= g (t)i dt ds = hf, A∗ S ∗ ∗ gei.
hAfe(s), S ∗ (s − t)e
0 0
reg
Therefore, the maximal Lp -regularity of A and the bound Mp,A (I) 6 Mpreg
0 ,A∗ (I)
follow as before.
If p = ∞, one can take f ∈ L∞ (I; X) and g ∈ L1 (I; D(A∗ )). The density
of D(A∗ ) in X ∗ follows from Lemma 17.2.22 and Remark 17.2.23. Therefore,
we can argue similarly as in the p = 1 case of the implication (1)⇒(2) by
reversing the roles of (A, p, f ) and (A∗ , p0 , g).
(2)⇔(1) for I = R+ : The case I = R+ follows from the previous cases,
and the following identities from Proposition 17.2.18:
Mpreg reg
0 ,A (R+ ) = sup Mp0 ,A (0, T ),
reg
Mp,A reg
∗ (R+ ) = sup Mp,A∗ (0, T ).
T >0 T >0
Earlier in this section, we have seen that maximal Lp0 -regularity for one ex-
ponent p0 ∈ [1, ∞] extrapolates to maximal Lp -regularity for all exponents
p ∈ (1, ∞). We shall now consider several extrapolation results to weighted
spaces. Recall that a weight is a measurable function w : (0, ∞) → (0, ∞) such
that w ∈ L1loc [0, ∞). Here we use the notation L1loc (I; X) introduced at the
beginning of this chapter, taking I = (0, ∞). To be explicit, the local integra-
bility condition asks for w to be integrable on every bounded subinterval of
(0, ∞).
Let I = (0, T ) be a bounded interval or I = R+ = (0, ∞).
Given a weight w on (0, ∞) and an exponent p ∈ [1, ∞), the space
Lp (I, w; X) can be defined as the Banach space of strongly measurable func-
tions f : I → X such that t 7→ w1/p (t)f (t) belongs to Lp (I; X). With respect
to the natural norm on this space, for functions belonging to this space we
have the identity
kf kLp (I,w;X) = kw1/p f kLp (I;X) .
In order to also cover the exponent p = ∞ it is useful to proceed slightly
differently.
608 17 Maximal regularity
p
Definition 17.2.34. For weights w ∈ Lloc ([0, ∞)) we define Lpw (I; X) as the
Banach space of strongly measurable functions f : I → X such that t 7→
w(t)f (t) belongs to Lp (I; X). With respect to the natural norm on this space,
we then have the identity
kf kLw
p
(I;X) = kwf kLp (I;X) .
In what follows we will always make the assumption that the weight w is
chosen in such a way that, as sets, we have the inclusion
p 1
Lw (I; X) ⊆ Lloc (I; X), (17.41)
using again the notation introduced at the beginning of the chapter. For p ∈
0
[1, ∞], the inclusion (17.41) holds if and only w−1 ∈ Lploc (I).
With the spaces Lpw (I; X) at hand, we may define the notion of maximal
Lw -regularity in the obvious manner, replacing all occurrences of Lp in Defini-
p
tion 17.2.4 by Lpw . We leave it to the reader to check that several basic results
on maximal Lp -regularity extend mutatis mutandis to maximal Lpw -regularity;
this includes Propositions 17.2.5, 17.2.10, and Theorem 17.2.19.
The present section will be concerned with identifying situations in which
maximal Lpw -regularity can be inferred from maximal Lq -regularity (where
the exponents p and q are possibly different). We start with two elementary
results in this direction for power weights tα and exponential weights e−λt .
From the point of view of applications to evolution equations, these suffice in
most cases; some of them are also valid in some of the end-point cases p = 1
and p = ∞. After that, we consider the more complicated case of Ap -weights,
restricting ourselves to p ∈ (1, ∞) in that case.
In order to treat extrapolation to power weights, we need the following
lemma. At the expense of additional technicalities, the implicit assumption of
strong measurability with respect to the uniform operator norm can be relaxed
somewhat, but in the applications we have in mind, it is always fulfilled.
Lemma 17.2.35. Let X and Y be a Banach spaces, and let p ∈ (1, ∞] and
α ∈ R be fixed. Let k : R+ × R+ → [0, ∞) be given by
|1 − (t/s)α |
k(t, s) = 1t6=s K(t, s),
|t − s|
|1 − tα |
c(t) = 1J (t)t1/p ,
|t − 1|
where J = (1, ∞) in case (1) and J = (0, ∞) \ {1} in case (2). Then
Z
kTk f (t)k 6 1D (t, s)kk(t, s)kkf (s)k ds
R+
|1 − (t/s)α | 1/p
Z
−1/p ds
6 kKk∞ t 1J (t/s)(t/s)1/p s kf (s)k .
R+ |(t/s) − 1| s
−1/p
= kKk∞ t c ? g(t),
where g(s) = s1/p kf (s)k and we used the convolution product ? of the group
(R+ , ·) with Haar measure ds/s. Therefore, Young’s inequality implies
On the interval (1/2, 3/2), c is uniformly bounded by the mean value theorem
R 3/2
and therefore 1/2 c(t) dtt < ∞. On the interval (0, 1/2), c is only non-zero in
case (2), and in this situation α > −1/p implies
Z 1/2 Z 1/2
dt 1
c(t) 62 t p −1 max{1, tα } dt < ∞.
0 t 0
Using Lemma 17.2.35 we can prove the equivalence of weighted and un-
weighted maximal Lp -regularity in case of power weights. The reader may
check that the proof below can be extended to various other classes of integral
operators. Actually, a similar argument has already been used in Proposition
10.2.31 for proving weighted Lp -boundedness of the Hilbert transform.
610 17 Maximal regularity
We do not know about a possible converse of (2), but it does not seem to be
particularly relevant for applications.
Proof. (1): Using the scaling properties of the weights wα , one can check that
Theorem 17.2.15 extends to the weighted setting. Therefore, −A generates an
analytic semigroup, which is bounded in case I = R+ . In view of Theorem
17.2.19 and its weighted extension, it suffices to prove that the operator V
introduced in this theorem is bounded on Lp (I; X) if and only if it is bounded
on Lpwα (I; X).
The boundedness of V on Lpwα (I; X) is equivalent to the boundedness of
the operator Vα on Lp (I; X), where
Z
Vα f (t) := A 1t>s tα S(t − s)s−α f (s) ds.
I
For λ ∈ %(A),
1
This implies that V is bounded on Lw α
(I; X) with norm at most C.
The proof actually shows that, in case −A generates a bounded analytic semi-
group, we could also allow any α 6 −1/p in Proposition 17.2.36. Since this
is unimportant for applications to evolution equations and since such weights
are not locally integrable near zero (which would lead to problems in Theorem
17.2.15), we do not elaborate on this any further.
We will apply the above result to the inhomogeneous problem
(
u0 (t) + Au(t) = f, t ∈ I,
(17.42)
u(0) = x,
where I = (0, T ) or I = R+ . The aim is to extend Proposition 17.2.14 to
the weighted setting. In the absence of weights, this proposition was proved
using the trace method. The proof presented here is different and uses the
semigroup generated by −A.
We fix p, p0 ∈ [1, ∞] satisfying p1 + p10 = 1, and consider the weight
wα (t) = tα for α ∈ (−1/p, 1/p0 ) ∪ {0}.
We define
X1− p1 −α,p := (X, D(A))1− p1 −α,p for α ∈ (−1/p, 1/p0 )
As in Proposition 17.2.13 one checks that the conditions on the data (f, x)
are also necessary if α + p1 ∈ (0, 1).
By choosing α appropriately, we can allow any u(0) = x ∈ (X, D(A))θ,p for
θ ∈ (0, 1). On the other hand, if p ∈ (1, ∞), we can apply Theorem L.4.1 to see
that u ∈ C(I; X1− p1 ,p ) (i.e, the solution regularises instantaneously from its
initial value u(0) into X1− p1 ,p . This agrees with the behaviour of Lp -solutions
without weights. Note that an estimate for kukLpwα (I;X) is not included in
(17.43), but can be obtained in the case of a bounded interval I = (0, T ) by
the same method as in (17.5).
612 17 Maximal regularity
kAux kLw
p
α (I;X)
6 CkxkX1− 1 −α,p .
p
Note that the extremal case α = 0 and p = ∞ is trivial. For the extremal case
α = 0 and p = 1 we use Lemma 17.2.22.
Clearly, u = uf + ux is a strong solution to (17.42), and combining the
estimates gives (17.43).
For later use we observe that, by restriction and uniqueness, the optimal
constant C in the estimate is monotone with respect to this interval I.
Next we consider the case of exponential weights. Here, only the case
I = R+ is of interest, as t 7→ eλt is uniformly bounded from below and above
on bounded subsets of R+ .
It follows that the function u defined by u(t) = eλt v(t) is the unique strong
solution to (17.42) with initial condition x = 0 and by the previous estimate
we obtain
reg
kAukLp−λ(·) (R+ ;X) = kAvkLp (R+ ;X) 6 Mp,A (R+ )kf kLp−λ(·) (R+ ;X) .
e e
17.2 Maximal Lp -regularity 613
As in Theorem 17.2.19 we see that for all f ∈ Lp (I, w; X) there exists a unique
Lp (w)-solution u of (ACP0 ), which satisfies Au = V f , and hence the claimed
estimate for solutions follows by combining the previous operator norm bound
with
The least admissible constant in this definition will be called the maximal
reg
C-regularity constant of A on I and will be denoted by Mcont,A (I).
Proof. (1): This follows by repeating the proof of the corresponding result for
maximal Lp -regularity (Proposition 17.2.5) verbatim.
(2): Fix f ∈ Cb (I; X). It follows from the definition of a strong solution
that uf (t) ∈ D(A) for almost all t ∈ I and uf ∈ Cb (I; X). By the definition
of maximal C-regularity, the almost everywhere defined function Auf is equal
almost everywhere, say outside the null set Nf , to a function F ∈ Cb (I; X).
Now let t ∈ I be arbitrary and choose a sequence tn → t, with every tn in
I \ Nf . Then uf (tn ) → uf (t) in X by continuity, and Auf (tn ) → F (t) in X.
Hence by closedness, uf (t) ∈ D(A) and similarly Auf (t) = F (t). This proves
the first assertion. The second is now evident from the fact that both uf and
Auf are well defined pointwise and continuous as X-valued functions.
As in the case of maximal Lp -regularity, in order to check maximal C-
regularity for a given closed operator A it suffices to prove existence, unique-
ness, and the estimate (17.44) for strong solutions corresponding to inhomo-
geneities f in a dense subspace F on Cb (I; X). The reader may check that
all results in Section 17.2.a have an analogue for maximal C-regularity, with
similar proofs. In particular we record the following analogue of Proposition
17.2.7:
17.2 Maximal Lp -regularity 615
The following is the analogue of Dore’s Theorem 17.2.15 and can be proved
in exactly the same way.
Proof. If I = (0, T ), the result is clear from the density of Cb (I; D(A))
in Cb (I; X). Next let I = R+ . We already observed that u := S ∗ f
is continuous with values in D(A). Let f ∈ Cb (I; X) and fix T > 0.
Choose fn ∈ Cb ([0, T ]; D(A)) such that fn → f |[0,T ] in Cb ([0, T ]; X). Let
f (t) = f (T ∧ t) and f n (t) = fn (T ∧ t). Then f n → f in Cb ([0, ∞); X).
It follows that un := S ∗ f n is a Cauchy sequence and hence convergent in
Cb ([0, T ]; D(A)). Since un → u := S ∗ f in X pointwise on [0, T ], it follows
that u := S ∗ f ∈ Cb ([0, T ]; D(A)) and
This implies that A has maximal C-regularity. This part of the proof does not
use the density of D(A).
Next suppose that A is densely defined and has maximal C-regularity.
As in the proof of Proposition 17.2.32 one shows that A∗ has maximal L1 -
reg
regularity with constant at most Mcont,A (I). Now Proposition 17.2.32 implies
reg reg reg
that A has maximal L∞ -regularity with M∞,A (I) = M1,A ∗ (I) 6 Mcont,A (I).
Conditions for maximal L∞ and C-regularity will be given in Section 17.3.c.
For later applications we also state a version of Proposition 17.2.36 for
maximal C-regularity with power weights. For this purpose, for α ∈ R we let
wα (t) := tα and consider the Banach space
n o
Cwα (I; X) := f ∈ C(I \ {0}; X) : sup ktα f (t)k < ∞
t∈I
17.2 Maximal Lp -regularity 617
with norm
Proof. Fixing α ∈ [0, 1), we can repeat the argument of Proposition 17.2.36,
using the additional fact that in Lemma 17.2.35(1) the operator Tk maps
L∞ (I; X) into Cb (I; Y ) if K is continuous (using notation introduced in the
lemma). To check the latter, let
|1 − (t/s)α |
c(t, s) := 1t>s
|t − s|
and recall from the proof of Lemma 17.2.35 that c(t, ·) ∈ L1 (I). We have
Z
kTk f (t + h) − Tk f (t)kY 6 kf k∞ c(t, s)kK(t + h, s) − K(t, s)k ds
I
Z
+ kf k∞ kKk∞ |c(t + h, s) − c(t, s)| ds.
I
To estimate the second term here, the mean value theorem applied to xα − 1
with x > 1, gives
Z t+h Z t+h t+h Z t+h
s −1 α
c(t + h, s) ds 6 α ds = ds → 0
t t t+h−s t s
as h ↓ 0. For the first term, the triangle inequality and the mean value theorem
lead to
(t + h)α − tα 1 1
|c(t + h, s) − c(t, s)| 6 + ((t/s)α − 1) −
sα (t + h − s) t+h−s t−s
tα−1 h h
6α α + ((t/s)α − 1)
s (t + h − s) (t + h − s)(t − s)
618 17 Maximal regularity
=: I + II.
α−1 α−1
Since sα t(t+h−s)
h
6 t sα is integrable, part I can be handled via dominated
convergence. For part II, we estimate
Z t/2
4h t/2
Z
II ds 6 2 ((t/s)α − 1) ds → 0,
0 t 0
and, the mean value theorem and dominated convergence theorem,
Z t Z t Z t
t h αh
II ds 6 α −1 ds = ds → 0.
t/2 t/2 s (t + h − s)(t − s) t/2 s(t + h − s)
The case h < 0 can be treated in a similar way.
Next we present a version of Corollary 17.2.37 for maximal C-regularity. In
Chapter 18 we also need the variants of the on subspaces of Cwα , which we
first introduce. For α ∈ [0, 1), I = (0, T ] or I = (0, ∞) let
Cwα ,0 (I; X) := {u ∈ Cwα (I; X) : lim tα ku(t)k = 0},
t↓0
1
Cw α
(I; X) := {u ∈ Cwα (I; X) : u0 ∈ Cwα (I; X)}.
We first prove the following a priori estimate: for all λ > ω large enough there
exists a constant D > 0 such that
−λt
where Sλ (t) = e S(t) is the semigroup generated by −A − λ. Hence, for
λ > ω,
Z ∞
(λ − ω)kukp 6 M (λ − ω)e−(λ−ω)t dt · (kf kp + kBukp )
0
6 M (kf kp + δkAukp + Kkukp ) 6 C1 kf kp + C1 Kkukp ,
where in the last step we used (17.48) and took C1 = M (δC0 + 1). For λ >
ω + C1 K this gives
kAukp 6 C2 kf kp , (17.50)
where E is the space introduced earlier in this proof; we used the easy obser-
vation that the domain of λ + A and A coincide as sets, and the associated
graph norms are equivalent. Thus L0 : E → F = Lp (I; X) is surjective.
From Lemma 16.2.2 we deduce that L1 is surjective, and hence invertible
by (17.47). Therefore, for each f ∈ Lp (I; X) there exists a unique u ∈ E with
u0 + (λ + A + B)u = f and (17.47) we obtain
Proof. Under the stated assumptions, the relative smallness condition (17.46)
on B was checked in the proof of Corollary 16.2.5. Thus the claims are imme-
diate from Proposition 17.2.49.
As an application of the previous perturbation results, we prove well-posedness
for a non-autonomous Cauchy problem. This result can be seen as an extension
of Corollary 17.2.37 to the time-dependent setting. Let X0 and X1 be Banach
spaces such that X1 ,→ X0 with a continuous embedding, and let A : [0, T ] →
L (X1 , X0 ) be strongly measurable in the strong operator topology. Consider
the time-dependent inhomogeneous problem
(
u0 (t) + A(t)u(t) = f (t), t ∈ (0, T )
(17.52)
u(0) = x.
(ii) for all t ∈ [0, T ] the operator A(t), viewed as an operator acting in X0
with domain D(A(t)) = X1 , has maximal Lp -regularity on (0, T ) with
reg
M := sup Mp,A(t) (0, T ) < ∞.
t∈[0,T ]
Set wα (t) := tα . Under these assumptions, for all x ∈ (X0 , X1 )1− p1 −α,p and
f ∈ Lpwα (0, T ; X0 ) there exists a unique strong solution
kukLw
p
α (0,T ;X1 )
+ kukWw1,p (0,T ;X0 )
α
Assumption (i) says that the graph norms of A(t), viewed as an operator
acting in X0 with domain D(A(t)) = X1 , are equivalent to the norm of X1 ,
uniformly with respect to t ∈ [0, T ]; Assumption (ii) says that these operators
have maximal Lp -regularity on (0, T ), uniformly with respect to t ∈ [0, T ].
The fact that u(t), for t > 0, takes values in X1− p1 ,p , which is a smaller
space than the initial value space X1− p1 −α,p if α > 0, is referred to by saying
that the solution instantaneously regularises if α > 0.
The proof of the theorem uses an iteration argument, in which the interval
[0, T ] is subdivided into smaller intervals [tn , tn+1 ]. On each of these, the
problem (17.52) is solved by a fixed point argument, taking u(tn ) as the initial
value for solving (17.52) on [tn , tn+1 ]. For this to work, it is crucial that
u(tn ) belong to the correct trace space; the bookkeeping needed to achieve
this is made possible by maximal Lp -regularity and is effected through a
judicious choice of the fixed point space. In the process, we ascertain that
upon completion of the (n + 1)th step, a suitable a priori inequality that
encodes all this information is being extended from [0, tn ] to [0, tn+1 ].
Implicit in this reasoning are two observations, the easy proofs of which
we leave to the reader:
• Gluing strong solutions on subintervals gives a strong solution on the full
interval. This follows by iterating the definition of a strong solution.
• Gluing Sobolev functions on subintervals gives a Sobolev function on the
full interval. This can be seen, e.g., through the characterisation of W 1,p -
functions as indefinite integrals of Lp -functions (see Section 2.5.c).
Proof. For each τ ∈ (0, T ] let Mτ : Lp (0, T ; X0 ) → Lp (0, T ; X1 ), f 7→ uf be
the solution map defined in Corollary 17.2.9, associated with A(τ ) in place of
A. (Note that the subscript τ of Mτ refers only to the operator A(τ ), whereas
the time interval under consideration is the same (0, T ) for all τ .)
Since A(τ ) has maximal Lp -regularity on (0, T ) with constant M , it follows
from (17.53) and (17.10) that
kuf kLp (0,T ;X1 ) 6 L kA(τ )uf kLp (0,T ;X0 ) + kuf kLp (0,T ;X0 )
6 L M kf kLp (0,T ;X0 ) + T (M + 1)kf kLp (0,T ;X0 )
6 L(1 + T )(M + 1)kf kLp (0,T ;X0 ) ,
624 17 Maximal regularity
and hence
kMτ k 6 (1 + T )(M + 1)L =: M
f1 .
By the extrapolation of maximal regularity with power weights (Proposition
17.2.36), A(0) has maximal Lpwα -regularity on (0, T ) with a certain constant
M0 and therefore in a similar way as before we see that the associated mapping
M0 : L wp
α
(0, T ; X0 ) → Lw p
α
(0, T ; X1 ) is bounded by with constant M
f0 .
Let M f = max{M f1 }. By uniform continuity, we can choose δ > 0 such
f0 , M
that, for all s, t ∈ [0, T ],
1
|t − s| < δ ⇒ kA(t) − A(s)kL (X1 ,X0 ) < .
2M
f
Finally, since u0 = −Au + f it is enough to estimate the first term on the left
with the right-hand side.
17.2 Maximal Lp -regularity 625
The main theorem of this section, Theorem 17.3.1, provides a necessary and
sufficient condition for maximal Lp -regularity on R+ in the setting of UMD
Banach spaces.
(2) If X is a UMD space, p ∈ (1, ∞) and A is R-sectorial with angle ωR (A) <
1 p
2 π, then A has maximal L -regularity on R+ with
reg 2
Mp,A (R+ ) 6 400~p,X βp,X (M + 1)2 ,
Remark 17.3.2. If A 6= 0, then lim inf t→0 kA(it + A)−1 k > 1, and in particular
the constant in Theorem 17.3.1(2) satisfies M > 1. To see this, let x ∈ D(A)
be so that both Ax 6= 0. Thus yt := (it + A)x is non-zero when |t| is small
enough. Now
|t|kxk
kA(it + A)−1 k > 1 − .
kyt k
is well defined, takes values in Lp (R+ ; X), and extends to a bounded op-
erator on Lp (R+ ; X);
(2) the mapping f → V f , defined for functions f ∈ F (R) by
Z ∞
V f (t) := A S(t − s)f (s) ds,
−∞
where we set S(−t) = 0 for t > 0, is well defined, takes values in Lp (R; X),
and extends to a bounded operator on Lp (R; X).
reg
In this case we have kV k = kV k = Mp,A (R+ ). Furthermore, the equivalence
of (1) and (2) and norm identity hold for p = ∞ if we additionally assume
that S be uniformly exponentially stable and set F (I) = L∞ (I; X).
In the above result, it is part of the assumptions that the convolutions take
values in D(A) almost everywhere. If f is sufficiently regular, one can pull
the operator A inside the integral by Hille’s theorem (Theorem 1.2.4). By
Theorem 17.2.19, (1) holds if and only if A has maximal Lp -regularity on R+ .
Proof. (2)⇒(1): This implication hold trivially, and so does the bound kV k 6
kV k.
(1)⇒(2) for p ∈ [1, ∞): Without loss of generality we can assume that
F (I) are the functions in Lp (I; X) with support in a bounded set.
628 17 Maximal regularity
Fix f ∈ F (R) and pick r > 0 so that the support of f is in (−r, ∞). Let
fr : R+ → X be defined by fr (s) := f (s − r). Then for all t > −r we have
Z ∞ Z t+r
V f (t) = AS(t+r−s)f (s−r) ds = AS(t+r−s)fr (s) ds = V fr (t+r).
−∞ 0
whereas V f (t) = 0 for t < −r since the support of f is in (−r, ∞) and S(t) = 0
for t < 0. Hence
Z ∞ Z ∞
p p
kV f kLp (R;X) = kV fr (t + r)k dt = kV fr (t)kp dt
−r 0
= kV fr kpLp (R+ ;X) 6 kV kp kfr kpLp (R+ ;X) = kV kp kf kpLp (R;X) .
Lemma 17.3.4. Let A be sectorial of angle < 21 π, and let S be the bounded
analytic semigroup generated by −A. In S 0 (R; L (X)) the following identity
holds:
F (A(1 + A)−1 1R+ S) = A(1 + A)−1 (2πi · +A)−1 .
Somewhat informally the lemma states that F (1R+ S) = (2πi · +A)−1 . The
terms A(1 + A)−1 cannot be left out, however, since AS and (2πi · +A)−1
need not be locally integrable.
Proof. For δ > 0 and t ∈ R let kδ (t) := 1(0,∞) (t)e−δt S(t). The function
kδ belongs to L∞ (R; L (X)), which we view as continuously embedded in
S 0 (R; L (X)). Then, for all Schwartz functions φ ∈ S (R; X),
Z
−1 b
A(1 + A) k0 (φ) = A(1 + A)−1 k0 (t)φ(t)
b dt
R
Z
(i)
= lim A(1 + A)−1 kδ (t)φ(t) b dt
δ↓0 R
= lim A(1 + A)−1 kbδ (φ)
δ↓0
Z
(ii)
= lim A(1 + A)−1 (2πiξ + δ + A)−1 φ(ξ) dξ
δ↓0 R
Z
(iii)
= A(1 + A)−1 (2πiξ + A)−1 φ(ξ) dξ,
R
where (i) and (iii) follow by dominated convergence and (ii) by the definition
of a generator (Definition G.2.1).
b
We recall from Section 2.4.a that L1 (R; X) denotes the space of functions that
are the inverse Fourier transform of a function in L1 (R; X).
By Proposition 10.3.3 (of which the proof extends to the case of non-dense
D(A)), the R-boundedness of {λ(λ + A)−1 : λ ∈ C+ \ {0}} implies the R-
sectoriality of A with angle ωR (A) < 21 π.
If p = ∞, then by Theorem 17.2.31, A has maximal Lq -regularity on R+
for any q ∈ (1, ∞), and R-sectoriality follows from the previous case.
(2): If A is R-sectorial of angle ωR (A) < 12 π, then the function m(ξ) =
A(i2πξ + A)−1 has R-bounded range with R-bound M as in the assumptions.
Moreover, m ∈ C 1 (R \ {0}; L (X)), and for all ξ ∈ R \ {0} we have
This proves the convolution estimate kAS ∗ f kL1 (I;X) 6 Ckf kL1 (I;X) , which
completes the proof of (1) and provides the Bochner integrability of s 7→
AS(t − s)f (s) on (0, t).
Turning to case p = ∞, we point out that we now impose a stronger as-
sumption of the analytic semigroup, involving the prefix ‘C0 -’, compared to
Theorem 17.3.11.
Theorem 17.3.12 (Kalton–Portal, maximal L∞ -regularity). Let I =
(0, T ) or I = R+ . Let −A be the generator of an analytic C0 -semigroup
(S(t))t>0 on a Banach space X, and suppose that supt∈I kS(t)k < ∞. Then
the following assertions are equivalent:
(1) A has maximal L∞ -regularity on I;
(2) A has maximal C-regularity on I;
(3) there exists a constant C > 0 such that
kxk 6 lim sup kS(t)xk + C sup ktAS(t)xk, x ∈ X,
t→T t∈I
where we set T = ∞ if I = R+ .
If these conditions are satisfied, and if we denote
MI,k := sup k(tA)k S(t)k, k = 0, 1, 2,
t∈I
reg
then (3) holds with C = (1 + 4M∞,A (I)MI,0 ), and
reg reg
M∞,A (I) = Mcont,A (I) 6 MI,1 log(2) + CMI,2 .
17.3 Characterisations of maximal Lp -regularity 635
Remark 17.3.13.
(1) If T < ∞, we have lim supt→T kS(t)xk = kS(T )xk, while if T = ∞, then
for all x ∈ R(A) we have lim supt→∞ kS(t)xk = 0.
(2) The quantities MI,k are finite under the assumptions of Theorem 17.3.12,
although only case k = 0 is explicitly postulated:
For k = 1, let first I = R+ . Then we are assuming that (S(t))t>0 is
a bounded analytic C0 -semigroup, and the finiteness of MI,1 is an ap-
plication of Theorem G.5.3. If I = (0, T ) with T < ∞, then (S(t))t>0
is only assumed to be an analytic C0 -semigroup, but not necessarily
bounded. (The finiteness of MI,0 is automatic for any semigroup.) How-
ever, Sλ (t) = e−λt S(t) will be a bounded analytic C0 -semigroup, if λ > 0
is sufficiently large. Then Theorem G.5.3 applies to show that
and then
MI,1 6 sup eλt kt(λ + A)Sλ (t)k + ktλS(t)k 6 eλT Mλ + T λMI,0 < ∞.
t∈I
Proof of Theorem 17.3.12. The equivalence of (1) and (2) was already ob-
tained in Theorem 17.2.46. Note that MI,k < ∞ for each k > 0. This is clear
if T < ∞ and follows from the uniform boundedness of the semigroup if k > 1.
(3)⇒(2): Let f ∈ Cb (I; D(A)) and fix t ∈ I. Since u := S ∗ f takes values
in D(A), we can use the assumption to obtain
For the first part of (17.57), the bound kAS(σ)k 6 MI,1 /σ gives
Z t
kS(r)Au(t)k 6 kAS(t − s + r)f (s)k ds
0
Z t
1
6 MI,1 dskf kCb (I;X)
0 t−s+r
t
= MI,1 log 1 + kf kCb (I;X) .
r
As r ↑ T , the logarithm tends to log(1 + Tt ), which vanishes if T = ∞. For
the second part of (17.57), the bound kA2 S(σ)k 6 MI,2 /σ 2 gives
Z t
2
krA S(r)u(t)k 6 krA2 S(t − s + r)f (s)k ds
0
636 17 Maximal regularity
Z t
r
6 MI,2 dskf kCb (I;X)
0 (t − s + r)2
t
= MI,2 kf kCb (I;X)
t+r
Therefore, (17.57) becomes
t
kAu(t)k 6 MI,1 log 1 + + CMI,2 kf kCb (I;X) .
T
Now Lemma 17.2.45 implies that A has maximal C-regularity.
(1)⇒(3): To obtain the desired estimate, it suffices to consider I = (0, T )
with T < ∞. The case I = R+ can then be deduced from the relation of max-
imal regularity on finite intervals and R+ established in Proposition 17.2.18.
Moreover, the generator −A of a C0 -semigroup has a dense domain by Propo-
sition G.2.3(4), and hence it suffices to consider x ∈ D(A).
Proposition 17.2.32 implies that A∗ has maximal L1 -regularity with con-
reg
stant at most M∞,A (I). Therefore, Theorem 17.3.11 gives
Z
reg
kA∗ S(t)∗ x∗ k dt 6 M∞,A (I)MI,1 kx∗ k, x∗ ∈ X. (17.58)
I
By (17.58),
Z T
reg
htAS(t/2)x, A∗ S(t/2)∗ x∗ i dt 6 4M∞,A (I)MI,1 sup ktAS(t)xk kx∗ k,
0 t∈(0,T )
and hence
reg
|hx, x∗ i| 6 kS(T )xk kx∗ k + (1 + 4M∞,A (I)MI,1 ) sup ktAS(t)xk kx∗ k.
t∈(0,T )
admits a unique Lp -solution u of (17.59) such that Au, u0 ∈ Lp (I; X). Here,
I = (0, T ) or I = R+ , and u0 is the weak derivative of u on I. Let
1,p
(I; X) := u ∈ W 1,p (I; X) : u(0) = 0 ,
0W
recalling that functions in W 1,p (I; X) admit a unique version that is contin-
uous on I.
Let D denote the weak derivative, viewed as a closed operator on X e =
Lp (I; X) with domain
D(D) := 0 W 1,p (IX).
We will also consider D as an operator on Xe = Cb (I; X) with domain
kAuk
e p + kDukp 6 M k(A
e + D)ukp , u ∈ D(A)
e ∩ D(D);
(4) A
e + D is closed.
kuf kD(A)∩D(D)
e
e + D)uf kp = kuf kp + kf kp 6 (M + 1)kf kp ,
= kuf kp + k(A
e p + kDukp 6 M k(A
kAuk e + D)ukp , u ∈ D(A)
e ∩ D(D);
Moreover, letting M0 (I) denote the least admissible constant in (2), one has
reg
M0 (I) 6 2Mp,A (I) + 1,
reg reg
Mp,A (0, T ) 6 M0 (0, T ), and Mp,A (R+ ) 6 M0 (R+ ) + 2MS (I).
17.3 Characterisations of maximal Lp -regularity 639
On the other hand, given u ∈ D(A) e ∩ D(D) we may take f := (A e + D)u. Then
uf = u, and substituting this in the above inequality gives the desired result.
(2)⇒(1): First consider the Lp -setting with 1 6 p < ∞. For f ∈ D(A))
e set
Rt
uf (t) = 0 S(t − s)f (s) ds. Note that Auf + Duf = f . In the case I = (0, T ),
one has uf ∈ D(A)e ∩ D(D) and (2) implies
e f kp 6 M k(A
kAuf kp = kAu e + D)uf kp = kf kp ,
and conclude the required result by density of D(A))e in Lp (0, T ; X). In the
p
case I = R+ it need not be true that uf ∈ L (R+ ; X), and thus we cannot
apply the estimate of (2). In order to get around this problem, fix T > 0 and
define vT : [0, ∞) → X by vT := φT uf , where φT : [0, ∞) → [0, 1] is the
piecewise linear function satisfying connecting the points (0, 1), (T, 1), and
(2T, 0), and which is zero on the interval (2T, ∞).
We have vT ∈ D(A) e ∩ D(D) and
e + D)uf + φT0 uf ,
e + D)vT = φT (A
(A
and therefore
2p −1
where cpp = p 6 2. Combining the estimates and letting T → ∞, we obtain
640 17 Maximal regularity
kAuf kLp (R+ ;X) 6 M kf kLp (R+ ;X) + MS cp kf kLp (R+ ;X) .
In (3) and (4), the angles are in fact equal to π/2. This will not be needed
later on.
Proof. We begin by observing that the operator −D generates the C0 -
semigroup S of right-shifts on Lp (0, T ; X),
f (s − t), if s − t ∈ I;
S(t)f (s) =
0, if s − t ∈
/ I.
(X, e θ,p = (Lp (I; X), Lp (I; D(A)))θ,p = Lp (I; (X, D(A))θ,p ).
e D(A))
Therefore, Theorem 16.3.20 implies that for λ ∈ R large enough λ + A eθ,p has
p
an absolute functional calculus on L (I; (X, D(A))θ,p ) of angle < π/2. Since
Proposition 17.3.16 gives that D is sectorial of angle π2 , we can apply Theorem
16.3.14 and Proposition 17.3.14 to conclude the required result.
(2): For 1 6 p < ∞ the result follows from (1) and Theorem 17.2.24. Now
let p = ∞. We use the equivalent norm of Theorem L.2.4 for (X, D(A))θ,∞ .
For all t > 0,
Z t
kAS ∗ f (t)kX 6 kAS(t − s)f (s)kX ds
0
Z t
.A,θ k(t − s)θ−1 dskf kL∞ (0,T ;(X,D(A))θ,∞ )
0
644 17 Maximal regularity
The least admissible constant C will be called the maximal Lp -regularity con-
reg
stant and will be denoted by Mp,A (R).
reg
Unlike in the definition of Mp,A (I) for bounded intervals I = (0, T ) or R+ , an
a priori bound on kukLp (R;X) is also included in the estimate (17.61) through
the use of the graph norm. Also, unlike in the case I = (0, T ) or I = R+ , in the
above definition the uniqueness is assumed a priori. In fact, when uniqueness
holds, then A is injective. Indeed, suppose that Ax = 0. Then, setting u = x,
one has u0 + Au = 0, and therefore u = 0 by uniqueness. Conversely, if A is
injective, then uniqueness is immediate from the estimate (17.61).
646 17 Maximal regularity
u0 + (λ + A)u = f (17.62)
satisfies
C
λ0 e−α|·| kukX + e−α|·| kAukX 6 ke−α|·| f kLp (R;X) .
Lp (R) 1 − |α|λ−1
0 C
v 0 + Av + α sgn(·)v = fα ,
v 0 + (λ + A)v = fα − α sgn(·)g.
|α|C
kT (g1 ) − T (g2 )kY 6 |α|Ckg1 − g2 kLp (R;X) 6 kg1 − g2 kY , (17.65)
λ0
where the constant on the right is < 1 by our assumptions. Therefore, T has
a unique fixed point v ∈ Y , and this is our required Lp -solution. Moreover,
by (17.65) and the assumed estimate of Theorem 17.3.29,
|α|C
kvkY 6 kT (v) − T (0)kY + kT (0)kY 6 kvkY + Ckfα kLp (R;X) .
λ0
C
It follows that kvkY 6 Kkfα kLp (R;X) with K = 1−|α|Cλ−1
, and therefore
0
as claimed.
Proof of Theorem 17.3.29. Fix α ∈ (0, λ0 /C), and let Y = Lp (R; D(A)) be
normed as in (17.64).
Let x ∈ X, s ∈ R, and ξ ∈ R be fixed, and for t ∈ R set fs (t) := eiξ(t+s) x.
Then e−α|·| fs ∈ Lp (R; X), and by Proposition 17.3.30 there exists a unique
Lpe−α|·| -solution us to
u0s + (λ + A)us = fs ,
and we have
C
ke−α|·| us kY 6 ke−α|·| fs kLp (R;X) . (17.66)
1 − αCλ−1
0
Since fs (t) = f0 (s+t) = eiξs f0 (t), it follows that us (t) = u0 (s+t) = eiξs u0 (t).
Therefore u0 (s) = eiξs y with y := u0 (0). Since e−α|·| u0 ∈ Y , we must have
y ∈ D(A), and consequently
MR f := uf , (17.67)
u0 + (λ + A)u = f (17.69)
satisfies
reg
In particular, Mp,A (R+ ) 6 C.
u0 + (λ + A)u = f on R (17.74)
satisfies
reg
λkuλ kLp (R;X) + kAuλ kLp (R;X) 6 (M + Mp,A (R+ ))kf kLp (R;X) , (17.75)
where M is such that kS(t)k 6 M for all t > 0, with (S(t))t>0 the bounded
analytic semigroup generated by A.
The required bound for AMRλ can be deduced from this by the same argument
as in Lemma 17.3.3.
From Theorem 17.3.32 and Proposition 17.3.33, it follows that, in order to
establish maximal Lp -regularity of A on R+ , it suffices to prove maximal
Lp -regularity of λ + A on R for all λ > 0, along with the estimate (17.70)
for some n ∈ N (or equivalently, for n = 0). The advantage of considering
problems on R is that one can apply the Fourier transform in the time variable,
and initial value conditions do not play any role. In particular, this leads to
simplifications in studying evolution equation with inhomogeneous boundary
values and initial values. This will be further explained in the Notes at the
end of the chapter.
associated with the functions m0 (ξ) := m(ξ) and m1 (ξ) := Am(ξ) extend
to a bounded operators from Lp (R; X) to Lp (R; X).
In this situation, the Lp -solution u to (17.60) satisfies u = Tm f , and we have
reg
max{N0 , N1 } 6 Mp,A (R) = kTm kL (Lp (R;X),Lp (R;D(A)) 6 N0 + N1 ,
Proof. We start by making some observations that will be needed in the proofs
of both implications. Let f ∈ S (R; X) be fixed. Then the function ξ 7→
(2πiξ+A)−1 fb(ξ) belongs to S (R; D(A)). Thus we may define u ∈ S (R; D(A))
by
u and F (Au) = Ab
F (u0 ) = (2πi·)b u
Now the bounded extension is obtained by the density of S (R; X) in Lp (R; X).
(2)⇒(1): By the preliminary observations, for any given f ∈ S (R; X) the
function u defined by (17.76) satisfies u0 + Au = f in S (R; X), and u = Tm f
is in S (R; X). Therefore, u is an Lp -solution to (17.60) and it satisfies
17.3 Characterisations of maximal Lp -regularity 655
kukLp (R;D(A)) = kTm f kLp (R;D(A)) 6 kTm kL (Lp (R;X),Lp (R;D(A))) kf kLp (R;X) .
If v is another Lp -solution to (17.60), these observations also imply that v =
Tm f = u. Therefore, any Lp -solution satisfies the required a priori estimate.
reg
The equality Mp,A (R) = kTm kL (Lp (R;X),Lp (R;D(A))) follows from a com-
bination of the bounds established in the course of proving (1)⇒(2) and
(2)⇒(1).
(2)⇔(3): Since kxkD(A) := kxkX + kAxkX , it follows that
kTm0 f kLp (R;X) = kTm f kLp (R;X) 6 kTm f kLp (R;D(A)) ,
kTm1 f kLp (R;X) = kATm f kLp (R;X) 6 kTm f kLp (R;D(A)) ,
kTm f kLp (R;D(A)) 6 kTm f kLp (R;X) + kATm f kLp (R;X) ,
which readily implies both implications and the related estimates.
The next theorem presents an extrapolation result for maximal Lp -regularity
on R.
Theorem 17.3.35 (Extrapolation). Let A be a densely defined bisectorial
operator that has maximal Lp0 -regularity on R for some p0 ∈ [1, ∞). Then for
all p ∈ (1, ∞) and Muckenhoupt weights w ∈ Ap , the operator A has maximal
Lp (w)-regularity on R. Moreover, the strong solution of (17.60) satisfies
1
max(1, p−1 )
kukLp (R,w;D(A)) 6 Cp0 ,p (M 2 + Mpreg
0 ,A
(R))[w]Ap kf kLp (R,w;X) ,
Next we present a duality result for maximal regularity on R. Note that unlike
in Proposition 17.2.32 we do not consider the end-points p = 1 and p = ∞.
Proposition 17.3.37 (Duality). Let X be a Banach space, and let 1 < p <
∞. Suppose that A is a closed and densely defined linear operator in X. Then
A has maximal Lp -regularity on R if and only if its adjoint A∗ has maximal
0
Lp -regularity on R, and in this case
1 reg
M (R) 6 Mpreg reg
0 ,A∗ (R) 6 2Mp,A (R).
2 p,A
Proof. By Proposition 17.3.34, maximal Lp -regularity of A on R is equivalent
to the boundedness of the two multipliers mk (ξ) = Ak (i2πξ + A)−1 (k = 0, 1),
0
on Lp (R; X). Similarly, maximal Lp -regularity of A∗ on R is equivalent to
the boundedness of the two multipliers (A∗ )k (i2πξ + A∗ )−1 (k = 0, 1) on
0
Lp (R; X ∗ ). It is evident that these are the pointwise adjoints mk (ξ)∗ of the
mk (ξ). By Proposition 5.3.7, if m ∈ MLp (R; X, X), then its pointwise adjoint
0
satisfies m∗ ∈ MLp (R; X ∗ , X ∗ ) with
and one can extract from the short proof that the converse implication and
estimate are also true. A combination of these results proves the proposition
at hand, and it is also easy to get the quantitative statement
1
X 1
X
Mpreg
0 ,A∗ (R) 6 kTm∗k kL (Lp0 (R;X ∗ )) = kTmk kL (Lp (R;X))
k=0 k=0
reg
6 2 max kTmk kL (Lp (R;X)) 6 2Mp,A (R).
k=0,1
Theorem 17.4.1. Let X be a Banach space and let 1 < p, q < ∞ and α > 0,
let λ > 0, and consider the operator A = λ + (−∆)α/2 as a closed and densely
defined operator on Lq (Rd ; X). Let I = R+ or I = (0, T ) with T ∈ (0, ∞).
The following assertions are equivalent:
(1) A has maximal Lp -regularity on I;
(2) X is a UMD space.
R reg
In this situation we have βp,X 6 2Mp,A (R+ ).
By inspection of the proof, the reader can also track a quantitative estimate
in the other direction. However, the present method is sub-optimal for this
purpose, and we therefore refrain from being more explicit about the result.
Proof. (2)⇒(1): By Theorem 10.2.25 (which, in turn, was proved with the help
of the Mihlin Multiplier Theorem 5.5.10), −∆ has a bounded H ∞ -calculus of
angle 0. Moreover, the proof of Theorem 10.2.25 shows that this functional
calculus is explicitly given by Fourier multiplier through the formula
where
λ + |2πξ|α
M (η, ξ) = .
2πiη + λ + |2πξ|α
Invoking the same scaling argument as in Proposition 5.5.2, we find that
the multiplier M2 defined by
|ξ|α
M2 (η, ξ) =
iη + |ξ|α
reg
satisfies kM2 kMLp (R×Rd ;X) 6 Mp,A (R+ ). Clearly M2 satisfies M2 (η, 0) = 0
and M (0, ξ) = 1 for every η 6= 0 and ξ 6= 0. Thus Theorem 13.3.5 applies to
show that
R reg
βp,Y 6 2kM2 kMLp (R×Rd ;X) 6 2Mp,A (R+ ),
Example 17.4.2. The aim of this example is to prove that the operator −∆,
viewed as a closed and densely defined operator on L1 (Rd ), fails to have
maximal Lp -regularity on any bounded interval (0, T ) for any p ∈ [1, ∞].
Suppose, for a contradiction, that −∆ has maximal Lp -regularity on (0, T ).
By Theorem 17.2.24, for large enough λ > 0, the operator λ − ∆ has maximal
Lp -regularity on R+ . Then, by Theorem 17.3.1(1), λ−∆ is R-sectorial of angle
< π/2. By Proposition 10.3.3, this implies that the family {e−(λ−∆)t : t > 0}
is R-bounded, and by the contraction principle this implies that the family
{et∆ : t ∈ [0, 1]} is R-bounded. We will now show that the latter is not the
case.
σ 2 |y|2
Set fσ (y) := σ d e− 2 . Completing squares, one obtains
σd
Z
|x−y|2 σ 2 |y|2
et∆ fσ (x) = d/2
e− 2t e− 2 dy
(2πt) Rd
d |σx|2
Z 2
σ − 2(σ2 t+1) − σ t+1 |y− σ2x |2
= d/2
e e 2t t+1 dy
(2πt) Rd
σd −
|σx|2
= 2 d/2
e 2(σ2 t+1) .
(σ t + 1)
660 17 Maximal regularity
N
|x|2 1/2
Z X 1 − 22n +1
= e dx
Rd n=0
(22n + 1)d
For m > 0 consider the disjoint annuli Am = {x : 2m < |x| < 2m+1 }. Splitting
the integral and estimating, we obtain
N Z |x|2
X 1 − 2(22m +1)
kF kL1 (Rd ) > e dx
m=0 Am (22m + 1)d/2
N
2md |2m x|2
Z
−
X
= e 2(22m +1) dx > (N + 1)cd ,
m=0 A0 (22m + 1) d/2
where cd > 0 depends only on the dimension d. On the other hand, by Fubini’s
theorem and Hölder’s inequality in Ω, we find
N
X √ √
εn f σ 6 N kfσ kL1 (Rd ) = N (2π)d/2 .
L1 (Ω;L1 (Rd ))
n=0
obtain several results which show that positive results for p = 1 and p = ∞
are rare, and geometric restrictions on the underlying spaces are required.
From Theorem 17.2.46 we recall that maximal L∞ -regularity coincides
with maximal C-regularity in the case that A is densely defined, and therefore
all results in this section pertaining to maximal L∞ -regularity also hold for
maximal C-regularity.
We begin with a theorem due to Baillon, which implies that in Banach
spaces without an isomorphic copy of c0 , only bounded operators A can have
maximal L∞ -regularity. Examples of Banach spaces without an isomorphic
copy of c0 are reflexive Banach spaces and L1 -spaces. To see the latter, one
may for instance note that L1 -spaces have cotype 2 (Proposition 7.1.4), but
c0 does not (Corollary 7.1.10).
Proof. It suffices to consider the case that I = (0, T ) with T ∈ (0, ∞). Reason-
ing by contradiction, we assume that A is an unbounded linear operator with
maximal L∞ -regularity and construct an isomorphic copy of c0 inside X. This
will be done by means of the Bessaga–Pelczyński theorem (Theorem 1.2.40),
which asserts that if (xn )n>1 is a sequence in a Banach space X satisfying
inf n>1 kxn k > 0 and
Xk
j xj 6 C
j=1
for all k > 1 and all signs 1 , . . . , k ∈ {−1, 1}, with a uniform constant C,
then the closed linear span of (xn )n>1 contains a subspace isomorphic to c0 .
By Theorem 17.2.15, −A generates an exponentially bounded analytic
semigroup (S(t))t>0 . By Proposition K.3.1,
1
lim sup tkAS(t)k > ,
t↓0 e
and hence there exists a sequence T > t1 > t2 > . . . ↓ 0 and a sequence
(yj )j>1 in X of norm one vectors such that
1
tj kAS(tj )yj k > , j > 1.
2e
By restriction to [0, t1 ] and passing to a subsequence, we may assume that
tj
t1 = T and tj+1 6 for all j > 1.
2j
Setting M := supt∈[0,T ] kS(t)k and fj (s) := S(s)yj , we have
662 17 Maximal regularity
Z tj
tj kAS(tj )yj k = A S(tj − s)fj (s) ds 6 C sup kfj (t)k 6 CM,
0 t∈[0,T ]
reg
where C = M∞,A (I).
Set xj := tj AS(tj )yj for j > 0. By the preceding estimates,
1
6 inf kxj k 6 sup kxj k 6 CM.
2e j>0 j>0
k
X Z k
X
= j S(s − T + tj )yj dt = (tj − tj+1 )j S(s − T + tj )yj .
j=1 Ij j=1
Since fn vanishes for n > N it is clear that A∗ uf (t) ∈ D(A∗ ), and by Young’s
inequality
Z ∞X Z t
kA∗ uf kL1 (R+ ;`1 ) = ne−n(t−s) fn (s) ds dt
0 n>1 0
XZ ∞ Z t
= ne−n(t−s) fn (s) ds dt
n>1 0 0
664 17 Maximal regularity
X
6 kne−n· kL1 (R+ ) kfn kL1 (R+ )
n>1
X
= kfn kL1 (R+ ) 6 kf kL1 (R+ ;`1 ) ,
n>1
interchanging sum and integral in the last step. From the extrapolation result
of Theorem 17.2.31 we find that A∗ has maximal Lp -regularity for all p ∈
[1, ∞). Finally, we note that by Theorem 17.4.4, A∗ cannot have maximal
L∞ -regularity, for `1 does not contain a closed subspace isomorphic to c0 .
Example 17.4.7. On `q with q ∈ (1, ∞), we consider the closed densely defined
operator A : (yn )n>1 7→ (nyn )n>1 with maximal domain
Let I = R+ or I = (0, T ) with T ∈ (0, ∞) and let p ∈ [1, ∞]. We will show
that A has maximal Lp -regularity if and only if p ∈ (1, ∞). Indeed, the fact
that maximal Lp -regularity fails for p = ∞ and p = 1 follows from Theorem
17.4.4 and Corollary 17.4.5. For p = q we argue as in Example 17.4.6. For all
f ∈ Lq (R+ ; `q ), by Young’s convolution inequality we have
Z ∞ XZ t q XZ ∞ Z t q
ne−n(t−s) fn (s) ds dt = ne−n(t−s) fn (s) ds dt
0 n>1 0 n>1 0 0
17.4 Examples and counterexamples 665
X q q
6 kne−n(·) kL 1 (R ) kfn kLq (R )
+ +
n>1
(2) Two basic sequences (xn )n>1 and (yn )n>1 in X are said to be equivalent
if for any scalar sequence (cn )n>1 it is true that
X X
cn xn converges if and only if cn yn converges.
n>1 n>1
gives a bijection between [(xn )n>1 ] and [(yn )n>1 ]. Since the coordinate projec-
tions of a Schauder basis are bounded, this bijection is closed, and therefore
bounded by the closed graph theorem. Applying the same argument to its
inverse, it follows that this bijection is an isomorphism.
The following lemma isolates the crucial property of Lq (0, 1) that will be
needed in the proof of Theorem 17.4.8.
Lemma 17.4.10. The Haar basis of Lq (0, 1) is unconditional for every 1 <
q < ∞, and non-homogeneous for q 6= 2.
Using the notation of Section 9.1.h, the (L2 -normalised) Haar basis is given
as (hn )n>1 , where h1 ≡ 1 and hn := φj,k for n > 2, where n = 2j + k with
j = 0, 1, 2, . . . and k = 1, . . . , 2j , and
Proof of Lemma 17.4.10. Unconditionality for 1 < q < ∞ of the Haar basis
is proved in Corollaries 4.5.8 and 4.5.16.
Consider the subsequence (φn,2 )n>2 . The supports of the functions gn
are disjoint, contained in (0, 21 ), and therefore they span a closed sub-
space isomorphic to `q . On the other hand, the functions in the sequence
(φj,k )j=1,2,3,...; k=2j−1 +1,...2j have support in ( 21 , 1) and their linear span con-
P 2j
tains the ‘( 21 , 1)-Rademacher’ functions rj = k=2j−1 +1 φj,k , j > 1, whose
closed span is isomorphic to `2 by the Khintchine inequalities. Since it can be
shown (a proof is included in the final paragraph of this section) that `q does
not contain a closed subspace isomorphic to `2 unless q = 2, it follows that
for 1 < q < ∞ with q 6= 2, the Haar basis of Lq (0, 1) is unconditional and
non-homogeneous.
With this lemma at hand, Theorem 17.4.8 is seen to be a special case of the
following more general result:
17.4 Examples and counterexamples 667
By invoking additional results from the theory of Schauder bases, Kalton and
Lancien proved an even stronger result, namely, that if X is a Banach space
with an unconditional basis (xn )n>1 with the property every sectorial operator
of angle zero on X is R-sectorial, then X is isomorphic to `2 . We will comment
on this result Notes at the end of the chapter.
Proof of Theorem 17.4.11. Let (en )n>1 be a non-homogeneous unconditional
basis for X. By renorming, this basis can be made into an 1-unconditional
basis, i.e., an unconditional basis satisfying
X X
λn cn en 6 kλk`∞ cn en
n>1 n>1
will do. This renorming preserves non-symmetry of the basis, and will be used
in the proof below to eliminate inessential unconditionality constants from
several estimates. As a side-remark, a Banach space with a 1-unconditional
basis is a Banach lattice with respect to the coordinate-wise ordering.
P (emj )j>1 and (enj )j>1P
Choose disjoint subsequences , as well as a scalar
sequence (aj )j>1 , such that j>1 aj emj converges and j>1 aj enj does not
converge.
The construction below will produce a sectorial operator A on the closed
linear span Y of the two sequences (emj )j>1 and (enj )j>1 that has all the
stated properties. The desired example on the whole space X is then obtained
by taking the direct sum with the identity operator on the closed linear span
Z of the remaining basis vectors (note that X = Y ⊕ Z with contractive
projections thanks to the 1-unconditionality).
Since the subsequences (emj )j>1 and (enj )j>1 do not overlap, the first
be relabelled as (e2j )j>1 and the second as (e2j−1 )j>1 . Define the sequence
(fj )j>1 in X by
(
ej (j odd)
fj =
ej + ej−1 (j even)
so that e2j = f2j − f2j−1 . It is elementary to check that (fj )j>1 is a Schauder
basis for its closed linear span. Consider the diagonal operator
X X
A: cj fj 7→ 2j cj fj
j>1 j>1
668 17 Maximal regularity
with its natural domain. By Lemma 10.2.27 (restricting to the positive in-
tegers as index set) this operator is sectorial of zero angle, and moreover
0 ∈ %(A). The analytic C0 -semigroup (S(t))t>0 generated by −A is bounded
and strongly continuous on every sector of angle strictly less than 21 π.
Assume, for a contradiction, that A is R-sectorial. Let λ = (λj )j>1 be a
sequence of positive real numbers to be chosen shortly. Since e2j = f2j −f2j−1 ,
we obtain
`
X
εj aj λj (λj + A)−1 (f2j − f2j−1 )
j=k
` λ a
X j j λj a j
= εj f 2j − f 2j−1
λj + 22j λj + 22j−1
j=k
` λ a
X j j λj a j
= εj (e 2j + e 2j−1 ) − e 2j−1
λj + 22j λj + 22j−1
j=k
` λ a λ
X j j j λj
= εj e 2j + a j − e 2j−1 .
λj + 22j λj + 22j λj + 22j−1
j=k
Now take λj := 22j . The above identity then takes the form
` ` 1
X X 1
εj aj 22j (22j + A)−1 (f2j − f2j−1 ) = εj aj e2j + aj e2j−1 .
2 6
j=k j=k
The 1-conditionality of (ej )j>1 and the assumed R-boundedness imply that
` `
X 1 1 X 1 1
aj e2j − aj e2j−1 = εj aj e2j + aj e2j−1
2 6 2 6 L2 (Ω;X)
j=k j=k
`
X
6 R({λ(λ + A)−1 : λ > 0}) εj aj e2j
L2 (Ω;X)
j=k
`
X
= R({λ(λ + A)−1 : λ > 0}) aj e2j ,
j=k
where R({λ(λ + A)−1 : λ > 0}) denotes the R-bound of the set {λ(λ +
A)−1 : λ > 0}. Since the right-hand side tends to P
0 as k, ` → ∞, so does
the left-hand side. By the convergence of the sum j>1 aj e2j , this in turn
implies
X`
lim lim aj e2j−1 = 0.
k,`→∞ k,`→∞
j=k
P`
This contradicts the fact that j=k aj e2j−1 fails to converge.
17.4 Examples and counterexamples 669
On closed subspaces of `q
In this paragraph we complete the proof that the Haar basis of Lq (0, 1) is
non-homogeneous for q ∈ [1, ∞) unless q = 2 (Lemma 17.4.10). The missing
piece of information was the following result.
The proof of this proposition is based on two lemmas, for which we need to
introduce some terminology. Recall that a sequence in a Banach space X is
said to be a basic sequence if it is a Schauder basis for its closed linear span,
and that two basic sequences (xn )n>1 and (yn )n>1 are said to be equivalent
if for any scalar sequence (cn )n>1 it is true that
X X
cn xn converges if and only if cn yn converges.
n>1 n>1
The basis constant of a basic sequence (xn )n>1 is defined as the (finite) number
supN >1 kPN k, where PN is the projection in the closed span [(xn )n>1 ] of
(xn )n>1 defined by
X XN
PN cn xn := c n xn .
n>1 n=1
m −1
NX
ym = c n xn ,
m=Nm−1
Lemma 17.4.14. Let (xn )n>1 be a basic sequence in a Banach space X and
let (yn )n>1 be a sequence in X which satisfies
X 1
kxn − yn k 6 ,
3K
n>1
670 17 Maximal regularity
where K is the basis constant of (xn )n>1 . Then (yn )n>1 is a basic sequence
equivalent to (xn )n>1 , and the mapping
X X
cn xn 7→ c n yn
n>1 n>1
This implies that kT xk > 13 kxk for all x ∈ [(xn )n>1 ], and therefore T is an
isomorphism from [(xn )n>1 ] onto its range in [(yn )n>1 ], which is dense since
it contains all finite linear combinations of the yn . It follows that T is an
isomorphism from [(xn )n>1 ] onto [(yn )n>1 ].
Finally, let y ∈ [(yn )n>1 ] and let x := T −1 y. Since (xn )P
n>1 is a Schauder
basisPfor [(xn )n>1 ] we have a unique representation x = n>1 cn xn . Then
y = n>1 cn yn and this representation is unique. This proves that (yn )n>1 is
a Schauder basis for [(yn )n>1 ], i.e., (yn )n>1 is a basic sequence.
where (∗) follows from the fact that 43 6 kun k 6 54 for all n > 1. This
computation shows that [(un )n>1 ] ' [(en )n>1 ] = `q isomorphically, and since
also [(yn )n>1 ] ' [(un )n>1 ] isomorphically we conclude that Z := [(yn )n>1 ] '
`q isomorphically.
17.5 Notes
Parabolic maximal regularity estimates of the form ku0f kp + kAuf kp 6 Ckf kp
can be traced to back to the Ladyženskaja, Solonnikov, and Ural0 ceva [1968].
Early contributions to the abstract operator-theoretic framework include
Sobolevskiı̆ [1964], De [1964], Grisvard [1969], Da Prato and Grisvard [1975].
More recent expositions can be found in Amann [1995], Kunstmann and Weis
[2004], Denk, Hieber, and Prüss [2003], Prüss and Simonett [2016].
There is an extensive literature of applications of maximal Lp -regularity
to non-linear parabolic problems; we refer to Prüss and Simonett [2016] and
references therein. An abstract approach to quasi-linear and semi-linear evo-
lution equations based on maximal Lp -regularity will be presented in Chapter
18. A small sample of the applications in the literature, which counts several
hundreds of papers, is given in the notes of that chapter.
Section 17.1
Most of the semigroups in this chapter are analytic semigroups which are not
necessarily strongly continuous. These are widely used in the literature, and
an extensive treatment can be found in Lunardi [1995].
672 17 Maximal regularity
Section 17.2
Calderón, and Panzone [1962, Theorem 2 ]. This result is not immediately ap-
plicable however, since it is phrased for kernels which are locally integrable on
Rd . In the present context one needs a corresponding result for kernels which
are locally integrable on Rd \{0}. The approach taken in the main text circum-
vents this problem by using the more general extrapolation result of Chapter
11. Under the assumption that −A generate an analytic C0 -semigroup, the
extrapolation result of part (4) of Theorem 17.2.31 is due to Cannarsa and
Vespri [1986] and Coulhon and Lamberton [1986]; see also Sobolevskiı̆ [1964];
the quantitative bound appears to be new. An extrapolation result for maxi-
mal Lp -regularity to the vector-valued Hardy space H 1 (I; X) is presented in
Hytönen [2005].
The duality result of Theorem 17.2.265 is due to Kalton and Portal [2008],
who proved it for maximal Lp -regularity on R+ via discrete maximal regular-
ity. The precise bounds for arbitrary intervals of the duality result obtained
in Proposition 17.2.32 seem to be new.
Weighted extrapolation
Maximal C-regularity
Section 17.3
According to Coulhon and Lamberton [1986], the following question was asked
by Brezis:
“à quelles conditions sur l’espace de Banach E a-t-on la régularité Lp
pour tout A générateur d’un semi-groupe analytique borné sur E ?”
[Under which conditions on the Banach space E does every bounded analytic
semigroup generator have maximal Lp -regularity?]
Corollary 17.3.8, which is due to De [1964], asserts maximal Lp -regularity
for negative generators of bounded analytic C0 -semigroups on Hilbert spaces.
In Coulhon and Lamberton [1986] it was shown that the Poisson semigroup
has maximal Lp -regularity on Lq (Rd ; X) if and only if X is a UMD space (see
Theorem 17.4.1 for the precise formulation). Lamberton [1987] subsequently
showed that if −A generates a bounded analytic C0 -semigroup S on a space
L2 (Ω) such that the operators S(t) extend to contractions on Lq (Ω) for every
q ∈ [1, ∞], then A has maximal Lp -regularity for all p ∈ (1, ∞).
A breakthrough was made by Dore and Venni [1987], where it was shown
that every operator A in a UMD Banach space X with bounded imaginary
powers of angle ωBIP (A) < π/2 has maximal Lp -regularity for p ∈ (1, ∞). The
approach was based on their preliminary version of (what is nowadays called)
the Dore–Venni theorem (Theorem 15.4.11), in which the additional condition
that the operators be invertible was made. This invertibility condition was
17.5 Notes 675
removed in Prüss and Sohr [1990], and Corollary 17.3.7 follows from this
paper.
It took approximately ten years Brezis’s question was settled definitively:
• Kalton and Lancien [2000] showed that Brezis’s question, in the form
stated, has a negative answer. For every q ∈ (1, ∞) \ {2}, they gave an
example of an operator −A generating a bounded analytic C0 -semigroup
on Lq (0, 1) such that A does not have maximal Lp -regularity on finite time
intervals (see Theorem 17.4.8).
• It was shown in Weis [2001a,b] that a sectorial operator A on a UMD
Banach space X has maximal Lp -regularity if and only if A is R-sectorial
(see Theorem 17.3.1).
The negative general answer of Kalton and Lancien [2000] will be further dis-
cussed below. The characterisation of maximal Lp -regularity in UMD spaces of
Weis [2001a,b] is based on the operator-valued Fourier multiplier theorem for
UMD-valued multipliers (Theorem 5.3.18), which was proved in the same pa-
per. This work led to many follow-up studies in which maximal Lp -regularity
was proved by checking R-sectoriality, some of which will be discussed be-
low. In particular, Weis’s result implies the sufficient conditions for maximal
Lp -regularity stated in Corollaries 17.3.7 and 17.3.9. Corollary 17.3.9 resem-
bles the previously mentioned result of Lamberton [1987], which instead of
positivity, assumes contractivity on Lq for all q ∈ [1, ∞]. One can check that
under the latter condition the argument in the proof of Corollary 17.3.9 can
be repeated: the R-sectoriality with ωR (A) < π/2 proved in Theorem 10.7.13
was based on Akcoglu’s maximal ergodic Theorem 10.7.14. The latter was
stated and proved for semigroups which are positive and contractive for some
q ∈ (1, ∞), but is also valid if contractivity is assumed for all q ∈ [1, ∞] (see
Dunford and Schwartz [1958, Theorem VIII.7.7]).
The perturbation result of Corollary 17.3.10 improves Proposition 17.2.49
for UMD spaces, since the condition of the relative bound is weaker. A similar
result under a different condition can be found in Kunstmann and Weis [2001].
The characterisations of maximal Lp -regularity for the endpoints p = 1
and p = ∞ of Theorems 17.3.11 and 17.3.12 are to due to Kalton and Portal
[2008], who proved these results via discrete versions of maximal Lp -regularity.
It seems that their proof of Theorem 17.3.12 also assumes that the operator
A be densely defined.
The alternative proof of Theorem 17.3.1 presented in Theorem 17.3.18 is
due to Kalton and Weis [2001], and is based on their sum-of-operator method
explained in Theorem 16.3.6. Proposition 17.3.14 seems to be folklore. The
fact that only the weaker result of Proposition 17.3.15 holds in the case I = R+
seems to be less known. The properties of the derivative operator Du = u0 on
Lp (I; X) with a Dirichlet condition at the left end-point collected in Proposi-
tion 17.3.16 are standard. Extensions to weighted spaces Lp (R+ , tγ d, t; X) can
be found in Prüss and Simonett [2004] and Meyries and Schnaubelt [2012b]. A
direct proof based on the Mihlin multiplier theorem and the duality between
676 17 Maximal regularity
Then A has maximal Lp -regularity on Lq (Ω, w) for any p, q ∈ (1, ∞) and any
w ∈ Aq .
The above result can be seen as another way to answer to Brezis question on
Lr -spaces: if the semigroup is bounded analytic on all Lr (Ω, w) with w ∈ Ar ,
then one has maximal Lp -regularity. Theorem 17.5.1 was used by Haller, Heck,
and Hieber [2003] to prove maximal Lp -regularity for a large class of elliptic
systems on Rd .
Fröhlich [2007] used Theorem 17.5.1 to prove maximal Lp -regularity for the
Stokes operator on C 1,1 -domains, and extends the papers Farwig and Sohr
[1997], Fröhlich [2003] where Rd and the half-space were considered, respec-
tively. Currently, such results under the weakest conditions on the domain are
due to Kunstmann and Weis [2017], and state that for any θ ∈ (0, π/2) the
Stokes operator has a bounded H ∞ -calculus of angle < π/2 (and thus maxi-
mal Lp -regularity) on bounded Lipschitz domains for all q ∈ (1, ∞) satisfying
| 1q − 12 | < 2d
1
+ ε Here ε > 0 only depends on the Lipschitz domain, and the
angle θ. The condition on q comes from the sectoriality result proved by Shen
[2012].
There is a large number of maximal Lp -regularity (or even H ∞ -calculus)
results for operators of Stokes type arising in fluid dynamics. Some of the
recent ones include Choudhury, Hussein, and Tolksdorf [2018], Giga, Gries,
Hieber, Hussein, and Kashiwabara [2017], Hieber and Prüss [2020], Prüss
[2018], Shibata [2020], Shibata and Shimizu [2008], Simonett and Wilke
[2022a], Tolksdorf [2018, 2020], Tolksdorf and Watanabe [2020], Watanabe
[2022, 2023]. Applications of maximal Lp -regularity to equations of Navier–
Stokes type are too numerous to list here, but some are mentioned in Chapter
18.
where the result is stated for analytic semigroup which are not necessarily
continuous. A direct proof for 1 6 p < ∞ can be found in Prüss and Si-
monett [2016]. The proof for p = ∞ is taken from Lunardi [1995]. Examples
of operators with maximal L1 -regularity on real interpolation spaces of the
form (X, D(A))θ,1 can be obtained from Corollary 17.3.20 or by more direct
arguments. This often leads to maximal regularity results for operators acting
s s
the Besov spaces Bq,1 or their homogeneous counterparts Ḃq,1 . These spaces
appear naturally in fluid dynamics; see, e.g., Danchin, Hieber, Mucha, and
Tolksdorf [2020], Ogawa and Shimizu [2016, 2021], Xu [2022], and references
therein.
Our proof of Corollary 17.3.20 uses the fact that Lp (I; (X, D(A))θ,p ) equals
the real interpolation space (Lp (I; X), Lp (I; D(A)))θ,p up to an equivalent
norm. Alternative results can be obtained if one replaces Lp (I; X) by a real
s
interpolation space such as the Besov space Bp,q (I; X), or even by other spaces
s
such as the Triebel–Lizorkin spaces Fp,q (I; X). To prove maximal regularity
results in such spaces at least two methods are available which will be dis-
cussed briefly. We will only provide the details in the case I = R in order to
avoid difficulties with extension operators and initial value conditions.
The first method is due to Kalton and Kucherenko [2010], and is simi-
lar as in Corollary 17.3.20. However, this time the absolute calculus is used
for Du = u0 instead of A. It follows from Proposition 14.4.17 that D is
s s+1
a closed operator on Bp,q (R; X) with domain Bp,q (R; X) for any s ∈ R,
p, q ∈ [1, ∞]. An application of the Mihlin multiplier theorem (Theorem
14.4.16) for Besov spaces shows that for all µ > 0 the operator µ + D is an
invertible sectorial operator of angle 6 π/2. Moreover, by Theorem 14.4.31,
s s0 s1
Bp,q (R; X) = (Bp,q (R; X), Bp,q (R; X))θ,q if (1 − θ)s0 + θs1 = s. Therefore,
Theorem 16.3.14 implies that D has an absolute function calculus of an-
gle 6 π/2. From Theorem 16.3.20 we see that, at least for p, q < ∞, for
any densely defined sectorial operator A such that A − µ is also sectorial,
the operator A + D = (A − µ) + (D + µ) is invertible as an operator from
s s+1
Bp,q (R; D(A)) ∩ Bp,q (R; X). This implies a maximal regularity result for A
s
in Bp,q (R; X).
The second method is due to Amann [1997], Girardi and Weis [2003a], and
Bu and Kim [2005], and uses a Fourier multiplier Mihlin’s multiplier theorem
for Besov and Triebel Lizorkin spaces (Theorems 14.4.16 and 14.6.11 respec-
tively). Indeed, one can repeat the argument of Theorem 17.3.1(2), except
that one needs 0 ∈ %(A) to avoid problems with the Mihlin condition at zero
in Theorems 14.4.16 and 14.6.11. An extension of these results to Besov–Orlicz
spaces was recently given in Ondreját and Veraar [2020], where it was used
to study temporal regularity of stochastic convolutions.
Section 17.4
Theorem 17.5.2 (Kalton, Lorist, and Weis [2023]). For an order con-
tinuous Banach function space X the following assertions are equivalent:
(1) {λ(λ − ∆)−1 : λ > 0} is R-bounded on Lp (Rd ; X);
(2) X is a UMD space.
Combining their proof with the dilation argument in Theorem 17.4.1, one can
show that the same equivalence holds when the negative Laplacian −∆ is
replaced by A = λ + (−∆)α/2 . Notice that condition (1) is equivalent to R-
sectoriality. It is unknown whether Theorem 17.5.2 holds for general Banach
spaces X.
Theorem 17.4.4 is the main result of Baillon [1980]. Our presentation combines
ideas of Baillon [1980] and Eberhardt and Greiner [1992] (in the latter paper
we do not understand how the closed graph theorem is applied, for the space
of piecewise continuous functions is not complete). Example 17.4.6(1), which
shows that the conclusion of Baillon’s theorem fails for c0 , is taken from Dore
[2000] and seems to be an example due to Kato (See the MathSciNet review of
the paper Baillon [1980]). Corollary 17.4.5 and Examples 17.4.6(2) and 17.4.7
are due to Guerre-Delabrière [1995].
Before the statement of Corollary 17.4.5, we mentioned that the duals of
L∞ (S) and certain C(K)-spaces have cotype 2. This follows from the fact
that these duals are abstract L1 -spaces and thus isometric to an L1 -space, see
17.5 Notes 681
Lindenstrauss and Tzafriri [1979, Theorem 1.b.2 and the proof of Theorem
1.b.6].
A simple proof of the result, due to Sobczyk [1941], that isomorphic copies
of c0 in separable Banach spaces are always complemented can be found in
Lindenstrauss and Tzafriri [1977, Theorem 2.f.5]. That X contains a comple-
mented copy of `1 if X ∗ contains a copy of c0 is due to Bessaga and Pelczyński
[1958]; see also Lindenstrauss and Tzafriri [1977, Proposition 2.e.8]. Variations
of Baillon’s theorem and connections to control theory can be found in Jacob,
Schwenninger, and Wintermayr [2022], where it is shown that if −A generates
an analytic C0 -semigroup S, then the boundedness and well-definedness of the
operator ΦT : L∞ (0, T ; X) → X given by ΦT (f ) = AS ∗ f (T ) characterises
boundedness of A. This shows that one cannot replace the essential supremum
by a supremum in the definition of maximal L∞ -regularity. From the proof of
Corollary 17.3.20 for p = ∞ one sees that in certain situations ΦT is bounded
when S is not strongly continuous on (X, D(A))θ,∞ .
After partial results by several authors, Brezis’s question was finally solved
to the negative in Kalton and Lancien [2000], who constructed counterexam-
ples in separable Banach lattices not isomorphic to a Hilbert space, and in
particular in Lq -spaces with 1 < q < ∞, q 6= 2. This paper implies part (1)
of Theorem 17.4.8. More general counterexamples were constructed in Kalton
and Lancien [2002]. The construction presented here is due to Fackler [2014,
2016] and has the additional merits of being explicit and solving to the nega-
tive the extrapolation problem for maximal Lp -regularity; in particular, this
work implies part (2) of Theorem 17.4.8. Our proof of Theorem 17.4.11 is a
simplified version of the proof in Fackler [2016]. The result in this paper is
stated with ‘non-homogeneous’ replaced by ‘non-symmetric’, and then invokes
a result from Singer [1970] to reduce matters to the non-homogeneous setting
up to a permutation of one of the two subsequences. To deduce from this
the isomorphic characterisation of `2 mentioned after the statement of The-
orem 17.4.11, one uses a result of Lindenstrauss and Zippin [1969] (see also
McArthur [1972, Theorem 7.6], Fackler [2016, Proposition 5.5]) which states
that if X is a Banach space with an unconditional basis, and if X is not isomor-
phic to c0 , `1 , or `2 , then X has a normalised unconditional, non-symmetric
basis. The spaces c0 and `1 can be excluded, because on these spaces it is
possible to give simple direct constructions of sectorial operators that are
not R-sectorial (Kalton and Lancien [2000], Fackler [2016, Propositions 4.2
and 4.3]). For more on the theory of unconditional bases in Banach spaces
the reader may consult Lindenstrauss and Tzafriri [1977], Singer [1970]. Our
presentation of Proposition 17.4.12 follows Lindenstrauss and Tzafriri [1977].
In connection with Theorem 17.4.11 it is of interest to observe that the
diagonal operator A featuring in the proof is R-sectorial (respectively, almost
R-sectorial) if and only if the partial sum projections (respectively, the coor-
682 17 Maximal regularity
dinate projections) of the basis on which it acts are R-bounded; see Kalton,
Lorist, and Weis [2023, Proposition 6.1.3].
Several counterexamples to maximal Lp -regularity were presented by
Le Merdy [1999] in case X = L1 (T) and X = C(T). These are connected to
Examples 17.4.2 and 17.4.3 on L1 (Rd ) and C0 (Rd ) which are motivated by this
paper. The mentioned paper also contains a counterexample in X = K (`2 ),
the space of compact operators on `2 , of a densely defined invertible secto-
rial operator A of angle ω(A) < 21 π admitting bounded imaginary powers
with kAit k = 1 for all t ∈ R but without maximal Lp -regularity on bounded
intervals for any 1 < p < ∞. Some further counterexamples and standard
constructions for sectorial and Ritt operators can be found in Arnold and
Le Merdy [2019].
Abstract results showing that a large class of weakly compact C0 -semi-
groups on C(K) and L1 (Ω) fail to be R-bounded (and thus do not have max-
imal Lp -regularity) can be found in Hoffmann, Kalton, and Kucherenko [2004].
Further extension are given in Kalton and Kucherenko [2008], Kucherenko and
Weis [2005].
Constant domains
where B is the unit ball in Rd . It follows from their results that there exist
constants m, M > 0 and a measurable a : [0, T ] × Ω → R with m 6 a 6 M
on [0, T ] × B, such that for q > 1 close to 1 there is no maximal Lp regularity
for (17.78) on X0 = Lq (B). Another counterexample for a similar equations
but on Ω = R, can be obtained from Krylov [2016], where also the range of q
is analysed for which maximal Lq -regularity in Lq (R) does hold.
For differential operators with coefficients that are measurable in time and
VMO in space (with uniform estimates in time), Krylov [2008, Theorem 4.3.7
and Chapter 7] shows that the usual conditions such as uniform ellipticity
are sufficient for maximal Lp -regularity in Lq (Rd ) with p > q. The proof is
based on several sophisticated maximal function techniques. By a duality ar-
gument, the condition p > q can be removed in case of constant coefficients in
space. Generalisations to higher order elliptic systems (and certain boundary
conditions) have been obtained in Dong and Kim [2011] for p = q.
A more abstract operator-theoretic approach was taken by Gallarati and
Veraar [2017a,b], where arbitrary p, q ∈ (1, ∞) were considered and certain
R-boundedness conditions on the evolution family and commutator condi-
tions were assumed. In the special case of elliptic operators with coefficients
that are measurable in time, they use their abstract results to prove maximal
Lp -regularity with Ap -weights in time (and possibly also in space) in Lq (Rd ).
Afterwards they use the method of “freezing the coefficients” to extend to co-
efficients which are measurable in time and continuous in space (uniformly in
time). Using Rubio de Francia’s extrapolation theory (see Theorem J.2.1) they
then obtain maximal Lp -regularity on Lq (Rd ) for all p, q ∈ (1, ∞). Dong and
Kim [2018] unified the above results and proved estimates involving parabolic
weights and coefficients which are measurable in time and VMO in space. In
Dong and Krylov [2019] and Krylov [2020], Rubio de Francia’s extrapolation
theory was used to obtain regularity estimates in mixed Lp -norms for fully
684 17 Maximal regularity
Maximal γ-regularity
Lp -regularity and maximal γ-regularity are equivalent. This points the way
to using maximal regularity techniques beyond the UMD setting.
In what follows, the reader is assumed to be familiar with the basic theory
of γ-radonifying operators as presented in Section 9.1, whose notation and
terminology we follow here. As in the definition of maximal Lp -regularity, our
starting point is the inhomogeneous abstract Cauchy problem
(
u0 (t) + Au(t) = f (t), t ∈ I,
(ACP0 )
u(0) = 0,
Note that
Z t
f (r) dr 6 (t − s)1/2 kf kγ(a,b;X) , a 6 s 6 t 6 b.
s
Rt
It follows that t 7→ a
f (s) ds ∈ C([a, b]; X) and
Z t
t 7→ f (s) ds 6 (b − a)1/2 kf kγ(a,b;X) .
a C([a,b];X)
Rt
It follows from Theorem 9.6.1 that t 7→ a f (s) ds also belongs to γ(a, b; X).
Indeed, it is trivial to check that for functions f ∈ L2 (a, b) ⊗ X this mapping
686 17 Maximal regularity
observing that the indefinite integral, as an operator on L2 (a, b), has norm
(b − a)1/2 .
In analogy with Theorem 17.2.15, whose proof can be repeated almost
verbatim, we have:
Theorem 17.5.5 (Maximal γ-regularity for the problem (ACP0 )). Let
A be densely defined γ-sectorial on a Banach space X with ωγ (A) < 21 π, and
let (S(t))t>0 be the semigroup on X generated by −A. Then A has maximal
γ-regularity on I if and only if for every f ∈ γ(L2 (I), X) the mild solution
uf takes values in D(A) almost everywhere and Au ∈ γ(L2 (I); X). In this
situation, mild solutions and γ-solutions agree.
The following theorem shows that the UMD assumption of Theorem 17.3.1
can be lifted if maximal Lp -regularity is replaced by maximal γ-regularity.
γ-solution u(t) takes values in D(A1/2 ) for all t > 0, the resulting function
uf : R+ → D(A1/2 ) is uniformly continuous and satisfies
Miscellaneous topics
A : Y → L (X1 , X0 ) and F : Z → X0
A ∈ L (X1 , X0 ) and F : Z → X0 ,
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 689
T. Hytönen et al., Analysis in Banach Spaces, Ergebnisse der Mathematik und ihrer
Grenzgebiete. 3. Folge / A Series of Modern Surveys in Mathematics 76,
https://doi.org/10.1007/978-3-031-46598-7_8
690 18 Nonlinear parabolic evolution equations in critical spaces
The space on the left-hand side is the usual space in which solutions lie when
maximal Lp -regularity techniques are applicable. For the space Z one can take
either take Y , or more generally (X0 , X1 )β,1 with β ∈ [1 − p1 , 1), the latter
requiring polynomial growth restrictions on F . In practice, we often split F
into two parts F = FTr + Fc , where
FTr : Y → X0 , Fc : Z → X0 (18.2)
with Y and Z as before. Here the subscript Tr stands for trace space and c
stands for critical. The word critical is also used in the title of the chapter.
In Section 18.2 we will give a definition of a criticality using only evolution
equation terminology. Surprisingly, this often coincides with criticality from a
PDE perspective.
The following simple example explains why the additional flexibility in
choosing Z may be expected to be useful.
Example 18.0.1. On Rd consider the equation
(
∂t u − a(u)∆u = −u3 ,
(18.3)
u(0) = u0 ,
in order to define F (u). Conditions for this are given by Corollary L.4.7 which
2
in the current situation α = 0, h = 3 and thus θ = 1 − 3p lead to the
2θ,q 3q d 2
requirement H ,→ L , which holds if and only if q + p 6 3, which is even
weaker than what we saw before. We will come back to this point in Examples
18.1.3 and 18.3.1.
In Section 18.1 we start with the study of local existence and uniqueness
for semi-linear equations, where the function F is defined on the trace space
Y = (X0 , X1 )1− p1 with p ∈ (1, ∞). Here we can admit initial values u0 which
belong to the space Y . We present this setting separately, as it allows us to
introduce some important techniques in the simplest possible setting.
In Sections 18.2 we turn to the study of local well-posedness in the tech-
nically more demanding quasi-linear setting. At the same time, we improve
on the assumptions needed to make things work: it is possible to allow expo-
nents p ∈ [1, ∞] and functions F of the form FTr + Fc as in (18.2), with FTr
defined on Y as before and Fc on a larger space Z. Furthermore, we work in
a weighted setting in time. This has at least three important advantages:
(i) it allows initial data u0 belonging to the space (X0 , X1 )1−α− p1 , where
α > 0 is a parameter associated with the weight;
(ii) global existence of solutions can be proved under milder blow-up criteria;
(iii) it allows the inclusion of the endpoint p = ∞ (inclusion of the endpoint
p = 1 is possible for different reasons).
Blow-up criteria will be discussed in Section 18.2.d. After presenting an illus-
trating example in Section 18.3, the final Section 18.4 presents long term and
692 18 Nonlinear parabolic evolution equations in critical spaces
even global well-posedness results for small initial data in the case F = Fc
(i.e., FTr = 0).
and refer to this space as the trace space associated with the problem (18.4).
The following definition extends the notion of Lp -solutions to the present
setting.
one has
Then for all R > 0 there exists a T > 0 such that for all u0 ∈ X1− p1 ,p satisfying
ku0 kX1− 1 ,p 6 R the problem (18.4) has a unique Lp -solution u. Moreover,
p
u ∈ C([0, T ]; X1− p1 ,p ).
kukLp (0,T ;X1 )∩W 1,p (0,T ;X0 ) 6 Cp,A,T (kf kLp (0,T ;X0 ) + ku0 kX1− 1 ,p ). (18.6)
p
Note that zu0 ∈ MRp (0, T ) for every T < ∞. By (18.1) (and its proof) and
(18.6) (with f = 0) we have
Let Φ : B1T (u0 ) → MRp (0, T ) be defined by Φ(v) := u, where u is the unique
Lp -solution to the problem
(
u0 + Au = F (v),
u(0) = u0 .
This unique solution exists by the discussion preceding the proof; note that
F (v) ∈ C([0, T ]; X0 ) by the continuity of F and (18.1).
For later purpose we observe that for all v1 , v2 ∈ B1T (u0 ), Corollary L.4.6
(using v1 (0) − v2 (0) = 0 to get T -independent constants) implies that
First we check that Φ maps B1T (u0 ) into itself. For all v ∈ B1T (u0 ) one has
ku − zu0 kMRp (0,T ) 6 Cp,A,T kF (v)kLp (0,T ;X0 ) 6 Cp,A,1 T 1/p CR,F .
Therefore, for 0 < T 6 (Cp,A,1 CR,F )−p ∧ 1 we find that u ∈ B1T (u0 ).
To check that Φ is a uniform contraction, let vi ∈ B1T (u0 ) for i ∈ {1, 2}.
Using the maximal Lp -regularity estimate (18.6) for the equation which u1 −u2
satisfies, and (18.5), we find that
kΦ(v1 ) − Φ(v2 )kMRp (0,T ) 6 Cp,A,T kF (v1 ) − F (v2 )kLp (0,T ;X0 )
6 Cp,A,1 T 1/p ψ(NR )kv1 − v2 kC([0,T ];X1− 1 ,p )
p
1/p
6 Cp,A,1 T ψ(NR )Cp kv1 − v2 kMRp (0,T ) ,
where in the last step we used (18.7). Therefore, combining both conditions
on T it follows that for T = 21 ((Cp,A,1 ψ(NR )Cp )−p ∧ (Cp,A,1 CR,F )−p ∧ 1) the
mapping Φ is a uniform contraction on B1T (u0 ) with
1
kΦ(v1 ) − Φ(v2 )kMRp (0,T ) 6 kv1 − v2 kMRp (0,T ) .
2
By the Banach fixed point theorem, the restriction of Φ to B1T (u0 ) has a
unique fixed point u ∈ B1T (u0 ). From the definition of Φ, it is immediate that
u is an Lp -solution to (18.4).
It remains to prove the uniqueness. Uniqueness is clear on B1T (u0 ), but
we still need to prove uniqueness in the larger set MRp (0, T ). Let u1 , u2 ∈
MRp (0, T ) be Lp -solutions to (18.4). Then for every t ∈ [0, T ], by Corollary
L.4.6 (with t-independent constant), (18.6), and the remarks below it, and
(18.5),
where N is such that kui kC([0,T ];X1− 1 ,p ) 6 N for i ∈ {1, 2}. Therefore, apply-
p
ing Gronwall’s inequality to ku1 (t) − u2 (t)kpX , we find that u1 ≡ u2 on
1− 1 ,p
p
[0, T ].
Here is a simple example to which Theorem 18.1.2 can be applied. Further
examples will be given in Section 18.3.
Example 18.1.3. Let A ∈ L (X1 , X0 ), where
X0 = H s,q (Rd ) and X1 = H s+2,q (Rd )
with s ∈ (−2, 0] and q ∈ (1, ∞). In the present situation, Theorem 14.4.31
s+2− 2
shows that X1− p1 ,p = Bq,p p (Rd ). In order to have a concrete equation in
mind note that one for instance could take A to be a second order differential
operator such as −∆, and we could consider the PDE
(
∂t u − ∆u = f (u),
u(0) = u0 ,
where in the last step we used (14.22) and Proposition 14.4.18. Taking v = 0
and using the assumption f (0) = 0, one also obtains
kF (u)kX0 6 LCkukX1− 1 ,p .
p
The above estimates on F are not optimal and the condition on the exponents,
namely, s+2− p2 − dq > 0 turns out to be far from sharp. We also notice that in
s+2− 2
the example we can only treat rather smooth initial values u0 ∈ Bq,p p (Rd )
(in particular they need to be Hölder continuous). This turns out to be far
from sharp. Both these sharpness issues will be addressed in the next section.
We will make several changes to the simple setting considered in Section 18.1.
Besides the fact that the operator A now depends on the solution u, the
changes are as follows:
• The non-linearity is of the form
F = FTr + Fc ,
where FTr plays a similar role as in Section 18.1, and Fc is the so-called
critical part of F . We assume that both FTr and Fc are defined on a suitable
subset of Xσ,p (see (18.10) below) with σ ∈ [0, 1 − p1 ], and that Fc satisfies
a suitable polynomial growth condition.
• Weights in time are added (see Corollaries 17.2.37 and 17.2.48). This will
enable us to reduce the smoothness conditions on the initial data. At the
same time, this makes it possible to formulate flexible conditions for global
existence.
• The full range p ∈ [1, ∞] will be considered.
In Example 18.3.1 we will see that the new setting takes care of the issues
raised in the discussion after Example 18.1.3.
18.2.a Setting
X1 ,→ X0 .
Without loss of generality we will always assume that the constant in the
embedding is > 1.
698 18 Nonlinear parabolic evolution equations in critical spaces
We further fix
p ∈ [1, ∞]
and
α ∈ [0, p10 ) ∪ {0},
Note that Xθ,r = (X0 , X1 )θ,r if θ ∈ (0, 1) and r ∈ [1, ∞), because in these
ranges X0 ∩ X1 = X1 is dense in (X0 , X1 )θ,r by Corollary C.3.15. For θ = 0,
Xθ,r = (X0 , X1 )θ,r holds for all r ∈ [1, ∞] by definition.
Remark 18.2.1. There is some flexibility with regard to the choice of the spaces
Xθ in (18.10). These spaces will appear only in the assumptions on the non-
linearity Fc through (18.11) below. The only requirement needed is that Xθ,1
continuously embeds into this space. In the above definition one could for
instance take Xθ to be Xθ,r , [X0 , X1 ]θ , or D((ω + A(u0 ))θ ) for ω ∈ R large
enough.
In addition to the above-stated assumptions on the spaces X0 , X1 and the
parameters p, α, σ, we make the following structural assumptions on the
operator A and the non-linearity F .
(i) FTr : X1 ∩ Oσ,p → X0 and there exists an LTr > 0 such that
(ii) Fc : X1 ∩ Oσ,p → X0 and there exist m > 1, βj ∈ (σ, 1), ρj > 0 for
j ∈ {1, . . . , m}, and Lc > 0 such that
m
ρ ρ
X
kFc (u) − Fc (v)kX0 6 Lc (1 + kukXjβ + kvkXjβ )ku − vkXβj
j j
j=1
(18.11)
1 + ρj σ
βj 6 , j ∈ {1, . . . , m}. (18.12)
1 + ρj
We will now proceed to the main theorems on local well-posedness for the
quasi-linear problem
(
u0 + A(u)u = F (u), on (0, T ),
(18.13)
u(0) = v0 ,
where we recall that the Banach space Cwα ,0 ((0, T ]; X1 ) was defined before
∞
Corollary 17.2.48, and is a closed subspace of Lw α
(0, T ; X1 ). Similar assertions
1
hold for its C -variant.
p
Definition 18.2.5. Let Assumption 18.2.2 hold. A function u ∈ MRα (0, T )
p
is called a Lwα -solution to (18.13) on (0, T ) if u takes values in Oσ,p ,
A(u)u, F (u) ∈ L1 (0, T ; X0 ), and for all t ∈ [0, T ] we have
Z t Z t
u(t) − v0 + A(u(s))u(s) ds = F (u(s)) ds.
0 0
The main result of this section is the following local well-posedness for quasi-
linear equations.
Theorem 18.2.6 (Local well-posedness for quasi-linear problems).
Let Assumption 18.2.2 hold. If, for some u0 ∈ Oσ,p , the operator A(u0 ) has
maximal Lp -regularity (maximal C-regularity if p = ∞) on finite time inter-
vals, then there exist T > 0 and ε > 0 such that for all
This shows that for α > 0, the solution u instantaneously (that is, for
t ∈ (0, T ]) regularises from X1−α− p1 ,p to X1− p1 ,p . By similar arguments, an
analogous continuous dependence as in (18.14) holds in C([0, T ]; X1−α− p1 ,p )
and in the weighted space Cwα ((0, T ]; X1− p1 ,p ).
The parameters T , ε, and C in Theorem 18.2.6 depend on the choice
of u0 in general. The parameters T and ε need to be small enough for the
conclusions of the theorem to hold. This has several reasons. First of all, ε
must be small because we need BXσ,p (u0 , ε) to be contained in Oσ,p . More
importantly, the proof uses the maximal regularity of A(u0 ) to obtain local
well-posedness of (18.13) with initial value v0 , via a perturbation argument
involving the smallness of ku0 − v0 kXσ,p .
702 18 Nonlinear parabolic evolution equations in critical spaces
The time T must be small for two reasons. First of all, we need to assure
that u maps [0, T ] to Oσ,p . Secondly, in the proof of the theorem we also
need T to be small in order to be able to use fixed point arguments. This
is hardly surprising: already in the familiar setting of ordinary differential
equations, blow-up can occur in the presence of locally Lipschitz continuous
non-linearities F . Theorem 18.2.15 will provide conditions under which one
can extend the time interval of existence and uniqueness to the full interval
[0, ∞). For a special class of semi-linear equations, Theorem 18.2.17 will give
large-time well-posedness for small initial data.
The proof Theorem 18.2.6 uses a fixed point argument similar to the one of
Theorem 18.1.2. However, the proof is technically more demanding due to the
quasi-linear structure of the problem, the presence of the additional term Fc ,
the use of weights, and the admission of the full range p ∈ [1, ∞]; some new
ideas are needed to deal with these difficulties.
We will use the following abbreviations to keep the formulas at a reasonable
length. For k ∈ {0, 1} and j ∈ {1, . . . , m} we let, with notation introduced
earlier,
p
Lwα (0, T ; Xk ) if p < ∞
Ek :=
Cwα ,0 ((0, T ]; Xk ) if p = ∞
(
(ρj +1)p
Lwα/(ρ j +1)
(0, T ; Xβj∗ ) if p < ∞
Yj :=
Cwα/(ρj +1) ,0 ((0, T ]; Xβj∗ ) if p = ∞,
1+ρj σ
where βj∗ := 1+ρj . Assumption 18.2.2 implies that βj 6 βj∗ .
Lemma 18.2.7. Let Assumption 18.2.2 hold. Then for all T > 0 we have
continuous embeddings
and there exists a constant M1,T > 0 such that for all u ∈ MRpα (0, T ) and
j ∈ {1, . . . , m} we have
These constants may be chosen so that supT >1 M1,T < ∞. For functions
u ∈ MRpα (0, T ) satisfying u(0) = 0, the constants M1,T can be replaced by a
constant M1 independent of T > 0.
Proof. For p ∈ [1, ∞), the embeddings and estimates follow from Corollar-
ies L.4.6 and L.4.7, where for p = 1 we additionally use Remark L.4.2 and
Proposition L.4.5.
18.2 Local well-posedness for quasi-linear evolution equations 703
For p = ∞, the same can be done if Yj is replaced by Cwα/(ρj +1) ((0, T ]; Xβj∗ ).
∞
To get the embedding into its closed subspace Yj , recall u ∈ MRα (0, T ) =
1
Cwα,0 ((0, T ]; X1 ) ∩ Cwα,0 ((0, T ]; X0 ). Then, for all j ∈ {1, . . . , m}, by (L.19),
1−βj∗ ρj
where λ = α = 1+ρj . Hence
and
kFTr (u) − FTr (v)kC([0,T ];X0 ) 6 LTr ku − vkC([0,T ];Xσ,p ) ,
m
ρ ρ ρ
X
Cβjj ,X Lc T δj + kukYjj + kvkYjj ku − vkYj ,
kFc (u) − Fc (v)kE0 6
j=1
(18.17)
αρj ρj
where δj = 1+ρj + (1+ρj )p .
This gives the required estimate for A. Taking v ≡ u0 ∈ Xσ,p fixed, one also
sees that the function t 7→ A(u(t))z(t) belongs to E0 .
By Assumption 18.2.2(3),
This implies the estimate for FTr in (18.17); the assumptions on FTr and the
continuity of u : [0, T ] → Xσ,p (see Lemma 18.2.7) imply that t 7→ FTr (u(t))
belongs to C([0, T ]; X0 ).
Next, we have u, v ∈ Yj by Lemma 18.2.7. Moreover, for all j ∈ {1, . . . , m},
704 18 Nonlinear parabolic evolution equations in critical spaces
ρ ρ
(1+kukXjβ∗ + kvkXjβ∗ )ku − vkXβ∗
j j j Lp
wα (0,T )
(i)
ρ ρ
6 (1 + kukXjβ∗ + kvkXjβ∗ )k j(ρ +1)p/ρj ku − vk j (ρ +1)p
j j Lwαρ (0,T ) Lwα/(ρ (0,T ;Xβ ∗ )
j /(ρj +1) j +1) j
(ii)
ρ ρ
6 T δj +
kukYjj + kvkYjj ku − vkYj
1 j ρ
where in (i) we applied Hölder’s inequality with (1+ρ j)
+ (1+ρ j)
= 1 and
in (ii) the definition of Yj and the triangle inequality. The estimate for Fc
in (18.17) now follows from Assumption 18.2.2 and the inequality kxkXβj 6
Cβj ,X kxkXβ∗ (see Proposition L.1.1(2)), where we used that the embedding
j
constant in X1 ,→ X0 is > 1, and thus Cβj ,X > 1.
The estimate for FTr immediately extends to p = ∞. The estimates for
A and Fc also extend to p = ∞ if we replace E0 = Cwα ,0 ((0, T ]; X0 ) by
L∞wα (0, T ; X0 ). In order to obtain the estimates in the E0 -norm, it remains
to prove that t 7→ A(u(t))z(t) and t 7→ Fc (u(t)) are continuous on (0, T ] and
tα kA(u(t))z(t)kX0 and tα kFc (u(t))kX0 are bounded and tend to zero as t ↓ 0.
To prove continuity for A, we observe that for s, t ∈ (0, T ]
kA(u(t))z(t) − A(u(s))z(s)kX0
6 k(A(u(t)) − A(u(s))z(t)kX0 + kA(u(s))(z(t) − z(s))kX0
6 Lku(t) − u(s)kXσ,∞ kz(t)kX1 + kA(u(s))kL (X1 ,X0 ) k(z(t) − z(s))kX1
L
6 α ku(t) − u(s)kXσ,∞ kzkE1 + kA(u(s))kL (X1 ,X0 ) k(z(t) − z(s))kX1 .
t
The latter tends to zero if t → s, and the desired continuity follows. To prove
the bound and convergence of tα kA(u(t))z(t)kX0 , we observe that by (18.18),
applied with v ≡ x ∈ X1 ∩ Oσ,p ,
kA(u(t))z(t)kX0 6 kA(x)z(t)kX0 + kA(u(t))z(t) − A(x)z(t)kX0
6 kA(x)kL (X1 ,X0 ) kz(t)kX1 + Lku − xkC([0,T ];Xσ,p ) kz(t)kX1 .
Since z ∈ Cwα ,0 ((0, T ]; X1 ), this implies the desired boundedness and conver-
gence.
To prove continuity for Fc , note that by Assumption 18.2.2, for s, t ∈ (0, T ]
we have
m
ρ ρ
X
kFc (u(t)) − Fc (u(s))kX0 6 Lc (1 + ku(t)kXjβ + ku(s)kXjβ )ku(t) − u(s)kXβj.
j j
j=1
BrT (v0 ) = {v ∈ MRpα (0, T ) : v(0) = v0 , kv − zu0 kMRpα (0,T ) 6 r}. (18.19)
Note that BrT (v0 ) is a closed subset of MRpα (0, T ) by the continuity of the
trace at zero (see Lemma 18.2.7).
To prove local well-posedness for (18.13), we will apply the Banach fixed
point theorem to the mapping Φv0 : BrT (v0 ) → BrT (v0 ) defined by Φv0 (v) = u,
p
where u is the Lw α
-solution to
(
u0 + A(u0 )u = (A(u0 ) − A(v))v + F (v), on (0, T ),
(18.20)
u(0) = v0 .
Below we will first ensure that BrT (v0 ) ⊆ Oσ,p for ε, r > 0 small enough, so
that A(v) and F (v) are well-defined. Then from its definition, it is clear that
Φv0 maps BrT (v0 ) to MRpα (0, T ). Below we will check that for ε, r, T > 0 small
enough, Φv0 is well defined as a mapping from BrT (v0 ) to itself by using the
maximal regularity assumption on A(u0 ) and the mapping properties of A
p
and F . Note that a function u is an Lw α
-solution to (18.13) if and only if u
p
is an Lwα -solution (Cwα -solution if p = ∞) to (18.20) with u = v. Before we
turn to the fixed point argument we need several preparatory lemmas.
Choose ε0 > 0 such that BXσ,p (u0 , ε0 ) ⊆ Oσ,p . Fix T1 > 0 such that
By Corollaries 17.2.37 and 17.2.48, there is a constant CT1 such that for every
v0 ∈ Xσ,p we have
The constant CT1 will depend on T1 in general, but this will not create prob-
lems since T1 is fixed.
706 18 Nonlinear parabolic evolution equations in critical spaces
In order to show that A(v) and F (v) in (18.20) are well defined, we need
to check that v(t) ∈ Oσ,p for all t ∈ (0, T ) when ε ∈ (0, ε0 ), r ∈ (0, 1], and
T ∈ (0, T1 ] are small enough. This is taken care of in the next lemma.
Lemma 18.2.9. Let Assumption 18.2.2 hold, and let ε0 > 0 be chosen as
before (18.21). For small enough r ∈ (0, 1] and ε ∈ (0, ε0 ) the following holds:
For all v0 ∈ BXσ,p (u0 , ε), all T ∈ (0, T1 ], and all v ∈ BrT (v0 ), one has kv −
u0 kC([0,T ];Xσ,p ) < ε0 , and thus v(t) ∈ Oσ,p for all t ∈ [0, T ].
Proof. For notational convenience we write k · k∞,T = k · kC([0,T ];Xσ,p ) . For all
v ∈ BrT (v0 ),
This implies the required result for all r, ε > 0 small enough.
In the next lemma we collect some estimates for A, FTr , and Fc , which will
be used to ensure that Φv0 maps BrT (v0 ) to itself.
where Cε,r,T (u0 ) and Cε,T (u0 ) are independent of v0 and v, Cε,r,T (u0 ) and
Cε,T (u0 ) are non-decreasing in each of the variables ε, r, and T , and satisfy
Cε,r,T (u0 ), Cε,T (u0 ) → 0 as ε, r, T ↓ 0.
kFTr (v)kX0 6 kFTr (v) − FTr (u0 )kX0 + kFTr (u0 )kX0
6 LTr kv − u0 kXσ,p + kFTr (u0 )kX0
6 LTr ε0 + kFTr (u0 )kX0 ,
where in the last step we used Lemma 18.2.9. Taking Lpwα -norms, we obtain
1
kFTr (v)kE0 6 T α+ p (LTr ε0 + kFTr (u0 )kX0 ).
kvkYj 6 kv − zu0 kYj + kzu0 kYj 6 M1 r + (M1 + M1,T1 )CT1 ε + kzu0 kYj .
kFc (v)kE0
6 kFc (v) − Fc (zu0 )kE0 + kFc (zu0 )kE0
m
ρ
X
Cβjj ,X Lc T δj + 2(M1 r + C
eT ε + kj,T (u0 ))ρj (M1 r + C
6 eT ε) + kc,T (u0 ),
1 1
j=1
where have set C eT = (M1 + M1,T )CT , kj,T (u0 ) = kzu kY , and kc,T (u0 ) =
1 1 1 0 j
kFc (zu0 )kE0 . Note that kj,T (u0 ) → 0 and kc,T (u0 ) → 0 as T ↓ 0 since zu0 ∈
MRpα (0, T ) ⊆ Yj and since Fc (zu0 ) ∈ E0 by Lemma 18.2.8.
708 18 Nonlinear parabolic evolution equations in critical spaces
The estimate kFc (v)kE0 6 Cε,r,T (u0 )r + Cε,T (u0 ) in the statement of the
lemma now follows, with constants
m
ρ
X
Cβjj ,X Lc T δj + 2(M1 r + C
eT ε + kj,T (u0 ))ρj M1
Cε,r,T (u0 ) = 1
j=1
eT εM −1 Cε,1,T (u0 ) + kc,T (u0 ),
Cε,T (u0 ) = C 1 1
Remark 18.2.11. In the last part of the proof one does not have kj,T (u0 ) → 0
∞
and kc,T (u0 ) → 0 as T ↓ 0 if one were to use maximal Lw α
-regularity or data
u0 in (X0 , X1 )σ,∞ rather than in the closed subspace Xσ,∞ . This is one of
the reasons for working with maximal Cwα -regularity and data in Xσ,∞ . It is
also clear from the above proof that α = 0 leads to difficulties if p = ∞. For
1
example, the estimate for FTr (v) in Lemma 18.2.10 contains a factor T α+ p
which does not vanish in the limit T ↓ 0 if α = 0 and p = ∞.
The final lemma contains Lipschitz variations of the above estimates, which
will be used to show that Φv0 is a uniform contraction.
kv − u0 k∞,T
6 kv − zv0 k∞,T + kzv0 − zu0 k∞,T + kT (u0 )
6 M1 kv − zv0 kMRαp (0,T ) + M1,T1 kzv0 − zu0 kMRαp (0,T1 ) + kT (u0 )
6 M1 kv − zu0 kMRαp (0,T ) + (M1 + M1,T1 )kzv0 − zu0 kMRαp (0,T1 ) + kT (u0 )
6 M1 r + (M1 + M1,T1 )CT1 ε + kT (u0 ).
(18.25)
and
kv1 − v2 kYj 6 (M1,T1 + M1 )CT1 kv1,0 − v2,0 kXσ,p + M1 kv1 − v2 kMRpα (0,T ) .
After these preparations we are ready to turn to the proof of Theorem 18.2.6.
It will be useful to recall the maximal regularity estimate which follows from
Corollaries 17.2.37 and 17.2.48: for all f ∈ E0 and v0 ∈ Xσ,p there exists a
unique Lpwα -solution (Cwα -solution if p = ∞) to the problem
(
u0 + A(u0 )u = f on (0, T ),
u(0) = v0 ,
This constant CT also depends on A(u0 ) and p, but we can choose it in such
a way that CT 6 CT1 whenever T < T1 ; this follows from a weighted version
of (17.25).
Proof of Theorem 18.2.6. Fix ε ∈ (0, ε0 ) and r ∈ (0, 1] be as in Lemma
18.2.9, and let T ∈ (0, T1 ]. Let BrT (v0 ) be as in (18.19). Let Φv0 : BrT (v0 ) →
p
MRα (0, T ) be defined by Φv0 (v) := u, where u is the Lpwα -solution (Cwα -
solution if p = ∞) to the problem
(
u0 + A(u0 )u = (A(u0 ) − A(v))v + F (v),
(18.29)
u(0) = v0 .
Then v takes values in Oσ,p by Lemma 18.2.9, and we have (A(v) − A(u0 ))v ∈
E0 and F (v) ∈ E0 by Lemma 18.2.10. Below Theorem 18.2.6 we have already
observed that local existence and uniqueness follow if we can show that Φv0
has a unique fixed point.
Since u − zu0 satisfies (18.29) with v0 replaced by v0 − u0 , by the maximal
regularity estimate (18.28) applied on (0, T1 ) (see (18.21) for the definition of
T1 ) we have
ku − zu0 kMRpα (0,T ) 6 CA,T1 ku0 − v0 kXσ,p + k(A(u0 ) − A(v))v + F (v)kE0
eε,r,T r + C
6 CA,T1 ε + C eε,T ,
applying Lemma 18.2.10 in the last step, and where C eε,r,T and Ceε,T are con-
stants such that Cε,r,T → 0 as ε, r, T ↓ 0. Therefore, for r, ε, T > 0 small
e
enough we obtain ku − zu0 kMRpα (0,T ) 6 r, and thus u ∈ BrT (v0 ).
Next, fix vj,0 ∈ BXσ,p (u0 , ε) and vj ∈ BrT (vj,0 ) for j ∈ {1, 2}. Then u =
Φv1,0 (v1 ) − Φv2,0 (v2 ) solves the problem
(
u0 + A(u0 )u = (A(u0 ) − A(v1 ))v1 − (A(u0 ) − A(v2 ))v2 + F (v1 ) − F (v2 ),
u(0) = v1,0 − v2,0 .
where
and
Choosing ε > 0, r > 0, and T > 0 so small that CA,T1 Lε,r,T (u0 ) 6 1/2, we
obtain
1
kΦv1,0 (v1 ) − Φv2,0 (v2 )kMRαp (0,T ) 6 kv1 − v2 kMRpα (0,T )
2 (18.30)
+ (CA,T1 + 1)kv1,0 − v2,0 kXσ,p .
1
kuv1,0 − uv2,0 kMRpα (0,T ) 6 kuv1,0 − uv2,0 kMRpα (0,T )
2
+ (CA,T1 + 1)kv1,0 − v2,0 kXσ,p
which implies
Te.
r
Let εe := min ε, 8(CA,T +1) and set
1
n ro
Te := inf t ∈ [0, T ] : kuu0 − zu0 kMRpα (0,t) > ,
2
712 18 Nonlinear parabolic evolution equations in critical spaces
We claim that for every v0 ∈ BXσ,p (u0 , εe), the element uv0 ∈ MRpα (0, Te)
is the unique Lpwα -solution to (18.13). To show this, we will prove the slightly
stronger result (which will play a key role in the construction of the maximal
solution in Section 18.2.d) that, for an τ > 0, if v ∈ MRpα (0, τ ) is an Lpwα -
solution to (18.13), then v ≡ uv0 on [0, Te ∧ τ ]. This will give the theorem for
Te instead of T .
Let
τv := inf{t ∈ [0, Te ∧ τ ] : kv − zu0 kMRpα (0,t) > r},
kv − zu0 kMRpα (0,τv ) = kuv0 − zu0 kMRpα (0,τv ) 6 kuv0 − zu0 kMRpα (0,Te) < r,
Proof. Let us say that an Lpwα -solution v to (18.13) on (0, T ) has the unique-
p
ness property if for any τ > 0 and any Lw α
-solution u to (18.13) on (0, τ ), we
have v ≡ u on [0, T ∧ τ ]. Let Tmax (v0 ) be the supremum of all T > 0 such
p
that there exists an Lw α
-solution to (18.13) on (0, T ) with the uniqueness
property. Then Tmax (v0 ) > 0 by Theorem 18.2.6. Note that the uniqueness
property was established as part of the uniqueness proof. It follows that there
exists a maximal Lpwα -solution u : [0, Tmax (v0 )) → X0 to (18.13).
The final assertion in the theorem is called a blow-up criterion. Blow-up cri-
teria can be used to prove global well-posedness. In typical applications, as-
suming Tmax (v0 ) < ∞, energy estimates can be used to show that v∗ :=
limt↑Tmax (v0 ) v(t) exists in Oσ,p . This contradicts Theorem 18.2.15 and thus
leads to Tmax (v0 ) = ∞, i.e., global existence. Further blow-up criteria are
discussed in the Notes.
Proof. Assuming that T0 := Tmax (v0 ) < ∞ and that v∗ := limt↑T0 v(t) exists
in Xσ,p with v∗ ∈ Oσ,p , a contradiction will be derived.
The idea is to restart the problem at time T0 with initial value v∗ and apply
Theorem 18.2.6 to extend v to a larger time interval [0, T0 + δ]. However, it is
not self-evident that v ∈ MRpα (0, T0 + δ). This problem will be overcome by
using a compactness argument.
From the continuity of v and the assumption that the limit v∗ at t = T0
exist, it follows that the set
is compact in Xσ,p . By Theorem 18.2.6, for all x ∈ K there exists an open ball
p
B(x, εx ) ⊆ Oσ,p such that for initial values from B(x, εx ) we can find an Lw α
-
p
solution in MRα (0, tx ) for some tx > 0. Since K is compact, the open cover
{B(x, εx ) : x ∈ K} has a finite sub-cover {B(xn , εxn ) : n = 1, . . . N }. Let
δ := minn=1,...,N txn . Then for all x ∈ K there exists a unique Lpwα -solution
ux ∈ MRpα (0, δ) to the problem
714 18 Nonlinear parabolic evolution equations in critical spaces
(
u0 + A(u)u = F (u),
(18.31)
u(0) = x.
is well defined, belongs to MRpα (0, T + 21 δ), and is an Lpwα -solution to (18.13)
on (0, T0 + 21 δ). This contradicts the maximality of T0 .
Under the conditions of Theorem 18.2.15, one can leave out the uniqueness
from the second bullet in Definition 18.2.13. This excludes the existence of an
(non-unique) Lpwα -solution u ∈ MRpα (0, T ) which extends v.
Remark 18.2.16. Let Assumption 18.2.2 hold, and suppose that for all u0 ∈
Oσ,p the operator A(u0 ) has maximal Lp -regularity (C-regularity if p = ∞)
on finite time intervals. Let v0 ∈ Oσ,p and let v : [0, Tmax (v0 )) → X0 be
the maximal solution provided by Theorem 18.2.14. Now suppose that u ∈
MRpα (0, T ) is an Lpwα -solution to (18.13) for some T > 0. We claim that
T 6 Tmax (v0 ) and u ≡ v on (0, T ). To see this, first note that by the uniqueness
property of the proof of Theorem 18.2.14 one has u = v on [0, T ∧ Tmax (v0 )).
Thus it remains to show T 6 Tmax (v0 ). Suppose that T > Tmax (v0 ). Since
u ∈ MRpα (0, T ), it follows from Lemma 18.2.7 that
and v∗ ∈ Oσ,p . This contradicts Theorem 18.2.15 and thus the claim follows.
As a consequence of Theorem 18.2.15 we obtain the following criteria for global
well-posedness for (18.13) in the semi-linear case.
(3) sup kv(t)kXσ,p < ∞ and Assumption 18.2.2 holds in the sub-
t∈[0,Tmax (v0 ))
critical case,
then Tmax (v0 ) = ∞, and thus the Lpwα -solution v exists globally.
Proof. The existence of the maximal solution has already been observed in
Theorem 18.2.14.
We start with a preliminary observation. Fix ρ > 0 and T ∈ (0, ∞),
and set β ∗ := 1 − (α + p1 )(1 − ρ+1 1
). We claim that for all β ∈ (σ, β ∗ ] and
u ∈ L∞ (0, T ; Xσ,p ) ∩ Lpwα (0, T ; X1 ) we have
kukLhp
w (0,T ;(X0 ,X1 )β,1 ) 6 CT kukλL∞ (0,T ;Xσ,p ) kuk1−λ
Lp
w (0,T ;X1 )
, (18.32)
α/h α
1−β
where h = ρ + 1, λ ∈ (0, 1) is given by λ = 1
α+ p
, and where CT also depends
on α, h, p and is non-decreasing in T . From the assumption on β it follows
1
that λ ∈ [1 − 1+ρ , 1). Moreover, if β < β ∗ , one even has λ > 1 − ρ+1
1
. To prove
(18.32), note that by (C.6), Theorem L.3.1, and (L.2),
Therefore, maximal Lp -regularity of A implies that v ∈ MRpα (0, Tmax (v0 )). In
particular, limt↑Tmax (v0 ) v(t) exists in Xσ,p (see Lemma 18.2.7). This contra-
dicts Theorem 18.2.14. It follows that Tmax (v0 ) = ∞.
(2): This can be proved similarly, this time using maximal C-regularity.
(3): Suppose, for a contradiction, that Tmax (v0 ) < ∞. Let Oσ,p be as in
the proof of (1), and let T ∈ (0, Tmax (v0 )). As before, it suffices to prove
v ∈ MRp (0, Tmax (v0 )).
716 18 Nonlinear parabolic evolution equations in critical spaces
kvkMRp (0,T ) 6 C(kv0 kXσ,p + kFc (v)kLpwα (0,T ;X0 ) + kFTr (v)kLw
p
α (0,T ;X0 )
),
(18.33)
kFc (v)kLpwα (0,T ;X0 ) 6 kFc (v) − Fc (x)kLpwα (0,T ;X0 ) + kFc (x)kLpwα (0,T ;X0 )
m
ρ ρ ρ
X
6 Lc T δj + kvkYej + kxkYej kv − xkYej
j j j
j=1
m
ρ +1
X
6 Lc Cj,x + kvkYej ,
j
j=1
where in the last step we used Young’s inequality in the form aρ b 6 aρ+1 +
bρ+1 , and the constant Cj,x depends on Tmax (v0 ) but not on T . Let
M := sup kv(t)kXσ,p .
t∈[0,Tmax (v0 ))
1−βj
By (18.32) with h = ρj + 1, β = βj , and λj = 1 ,
α+ p
we find that
Setting ε = (2CLc m)−1 and letting T tend to Tmax (v0 ), it follows that v ∈
MRp (0, Tmax (v0 )).
18.3 Examples and comparison 717
and
s+2σ
Xσ,p = Bq,p (Rd ),
where p ∈ (1, ∞) (extensions to the end-points are possible, but not considered
here for simplicity) and σ ∈ (0, 1/p0 ] are arbitrary but fixed for the moment.
Suppose now that f ∈ C 1 (R) satisfies
d
s>− . (18.35)
q0
718 18 Nonlinear parabolic evolution equations in critical spaces
d 1 d dq(ρ + 1)
s + 2β − = s− and q 6 .
q ρ+1 q d − qs
Thus we arrive at the conditions
ρ d dρ
β= − s and s > − . (18.36)
2(ρ + 1) q q
Sobolev embeddings can also be applied in sub-optimal cases, but here we wish
to demonstrate certain optimality and scaling behaviour which is present only
if all Sobolev embeddings are sharp.
Combining (18.36) with the (sub)criticality condition (18.12), we obtain
d
ρ − s 6 2 + 2ρσ,
q
and criticality holds if
1d 1
σ= −s − .
2 q ρ
Since σ ∈ (0, 1/p0 ] we arrive the following condition on (q, s) to obtain a
critical setting:
1d 1 1
0< − s − 6 0. (18.37)
2 q ρ p
If (18.37) holds for some p, then it also holds for all larger values of p,
and one can take the limit p → ∞. Thus (18.35), (18.36), (18.37), and the
assumption s ∈ (−2, 0] imply
n d 2 do d 2 dρ
max −2+ − , −2, − 0 < s < − , and − 6 s 6 0. (18.38)
q ρ q q ρ q
In the converse direction, if (18.38) holds, then (18.37) holds for large enough
p, so the existence of a triple (p, q, s) satisfying the aforementioned conditions
is equivalent to (18.38).
Elementary computations show that we can find pairs (s, q) satisfying these
conditions holds if and only if
2 2 d
ρ> and < . (18.39)
d ρ(ρ + 1) q
In this case, the corresponding critical space for the initial data is given by
18.3 Examples and comparison 719
d
s+2σ −2
Xσ,p = Bq,p (Rd ) = Bq,p
q ρ
(Rd ). (18.40)
In (18.40) the limiting case where q = 21 dρ(ρ + 1) shows that we can ‘almost’
−2/(ρ+1)
treat initial data from the space Bq,p (Rd ). The less important so-called
microscopical tuning parameter p in (18.40) needs to be so large that (18.37)
holds.
Unlike in Example 18.1.3, it now becomes possible to take the special
structure of f into account. The space of initial data which we could consider
s+2− 2
in the example was Bq,p p (Rd ) with s ∈ (−2, 0] and s + 2 − p2 − dq > 0. Under
these restrictions, the smoothness parameter satisfies s+2− p2 > dq , which leads
to a much smaller class of initial data than considered in (18.40). Introducing
weights in the set-up of Example 18.1.3, does not change anything.
Indeed, the step where Hölder’s inequality is used can then be replaced by
k(1 + |u|ρ + |v|ρ )(u − v)kr 6 C`(1 + kukρ(ρ+1)r + kvkρ(ρ+1)r )ku − vk(ρ+1)r .
We finish this section with an example illustrating how Theorems 18.2.6 and
18.2.15 can be applied to obtain local and global well-posedness for certain
concrete PDEs.
This equation fits into the setting discussed in Example 18.3.1 with X0 =
H s,q (Rd ) and X1 = H s+2,q (Rd ) for suitable (q, s). Indeed, taking ρ = 2, one
checks that (18.39) holds if 1 < q < 3d. Let s ∈ (−2, 0] be such that (18.38)
holds with ρ = 2, and set σ := 21 ( dq − s) − 12 . Choose p ∈ (1, ∞) so large that
(18.37) holds. Then, by Example 18.3.1, F (u) = −u3 satisfies the Assumption
18.2.2. We choose to include the linear part of −u3 + u into the operator A.
Another possibility would be to put it into F as well, and consider ρ1 = 2 and
ρ2 > 0 arbitrary small.
From Example G.5.6 it follows that for s = 0 the operator Au = −∆u − u
on X0 , with domain X1 , is sectorial of angle zero. Moreover, by Theorems
17.4.1 and 17.2.26, A has maximal Lp -regularity on finite time intervals for
all p ∈ (1, ∞). Since the Bessel potentials (1 − ∆)t/2 commute with ∆, the
maximal Lp -regularity extends to the full range s ∈ R.
From now on we view (18.41) as an abstract problem of the form (18.13).
In particular, we say that (18.41) admits a (maximal) (p, q, s, σ)-solution if
(18.13) has a (maximal) Lpwα -solution. Applying Theorems 18.2.6 and 18.2.14,
d
−1
q
it follows that for every u0 ∈ Oσ,p = Xσ,p = Bq,p (Rd ) (see (18.40)), the prob-
lem (18.41) admits a maximal (p, q, s, σ)-solution (u, Tmax (u0 )). Moreover,
u ∈ Ww1,p
α
(0, T ; H s,q (Rd )) ∩ Lpwα (0, T ; H s+2,q (Rd ))
d
−1 2
s+2− p
(18.42)
q
∩ C([0, T ]; Bq,p (Rd )) ∩ C([τ, T ]; Bq,p (Rd ))
for all 0 < τ < T < Tmax (u0 ), where we used the instantaneous regularisation
stated in (18.15).
Global well-posedness can often be obtained via Theorem 18.2.17, but to
apply it to the rough initial data considered in the above example requires
first performing a (weighted) bootstrap argument to obtain enough regular-
ity in space and time. After that, suitable energy estimate can be applied.
Bootstrapping regularity will not be discussed here (a concise discussion of
this technique is included in the Notes). Instead, we will only prove global
well-posedness for sufficiently smooth initial data. This is done in the next
example. In particular, all initial data u0 ∈ Lq (Rd ) for q ∈ (d, 2d) are covered
if d ∈ {2, 3, 4, 5, 6}.
Example 18.3.5 (Global well-posedness for the Allen-Cahn equation). Consider
again the problem (18.41) in dimension d > 2. In order to obtain that u
takes values in H 1,q (Rd ), the smallest value of s which we can allow (without
bootstrapping) is s = −1. Let q ∈ ( d2 , 2d) and p ∈ (2, ∞) are such that
d 2 d 1
q + p 6 2 (see (18.37)), and set ρ := 2, σ := 2q , and α := 1 − p − σ. These
choices form a special case of Example 18.3.4, and in particular they lead to
a critical setting.
18.3 Examples and comparison 721
d
−1
Let u0 ∈ Bq,p q
(Rd ); note that this space contains Lq (Rd ) if q > d. By
the result of Example 18.3.4, the problem (18.41) admits a (unique) maximal
(p, q, s, σ)-solution, and for all 0 < τ < T < Tmax (u0 ) we have
1− 2
u ∈ Lp (τ, T ; H 1,q (Rd )) ∩ C([τ, T ]; Bq,p p (Rd )).
We will show global existence, i.e., Tmax (u0 ) = ∞, under the more restrictive
conditions
2q
max{d, 2d − 6} < q < 2d and 2 < p 6 . (18.43)
2d − q
For d = 2 we can take q ∈ (2, 4) and p ∈ (2, 2q/(4 − q)]. For d = 3 we can
take q ∈ (3, 6) and p ∈ (2, 2q/(6 − q)]. We do not claim this is optimal, and we
expect that by further bootstrapping some of these conditions can be omitted.
Step 1 – Assuming that Tmax (u0 ) < ∞, we will derive a contradiction
with Theorem 18.2.17(1). For the latter it suffices to use Step 2 below. How-
ever, we prefer to show the techniques to check Theorem 18.2.17(1) since
this can be useful for other situations. This boils down to showing that
u ∈ Lpwα (0, T ; H 1,q (Rd )) and
sup ku(t)kq
d −1 < ∞.
t∈[0,Tmax (v0 )) Bq,p (Rd )
By (18.42), both assertions are clear on [0, τ ] for any τ < Tmax (v0 ). Thus it
suffices to show that, for some τ > 0,
Step 2 – We show the second part of (18.44). Since dq − 1 < 0, by the easy
embeddings of (14.23) and Proposition 14.4.18, it is enough to show that
The idea will be to apply the chain rule of Lemma 18.3.6 below. For this we
need that u3 ∈ L1 (τ, T ; Lq ) for 0 < τ < T < Tmax (u0 ). To see this, note that
by Sobolev embedding with θ − dq = − 3q d
and interpolation,
3(1−θ)
ku3 kLq = kuk3L3q 6 Ckuk3H θ,q 6 C 0 kukLq kuk3θ
H 1,q .
we see that
Z tZ
q q
ku(t)kL q (Rd ) = ku(τ )kLq (Rd )
− q(q − 1) |u|q−2 |∇u|2 dx dr
τ R d
Z tZ
+q |u|q−2 (−u4 + u2 ) dx dr (18.45)
τ Rd
Z t
q q
6 ku(τ )kLq (Rd ) + q ku(r)kL q (Rd ) dr.
τ
q
Therefore, by Gronwall’s lemma applied to t 7→ ku(t)kL q (Rd ) ,
q q q(t−τ )
ku(t)kL q (Rd ) 6 ku(τ )kLq (Rd ) e .
Since we assumed Tmax (u0 ) < ∞, this implies the desired bound
q q qTmax (u0 )
N := sup ku(t)kL q (Rd ) 6 ku(τ )kLq (Rd ) e < ∞. (18.46)
t∈[τ,Tmax (u0 ))
3qd
where q0 = q+d . To prove that the latter is finite, note that by Sobolev
d dq
embedding with θ − 2 = − 2q 0
(then θ ∈ (0, 1] by (18.43) and 2q0 /q > 2 since
q < 2d),
q/2
kukLq0 (Rd ) = k|u|q/2 kL2q0 /q
6 C0 k|u|q/2 kH θ,2
18.3 Examples and comparison 723
1−θ
6 C1 k|u|q/2 kL 2 (Rd ) k|u|
q/2 θ
kW 1,2 (Rd )
6 C2 k|u| kL2 (Rd ) + k|u|q/2 k1−θ
q/2
k∇|u|q/2 kL
θ
2 (Rd )
L2 (Rd )
θZ θ/2 i
q(1−θ)/2 q
h
q/2
= C2 kukLq (Rd ) + kukLq (Rd ) θ |u|q−2 |∇u|2 dx .
2 Rd
qθ
h Z θ/2 i
6 C2 N 1/2 + N (1−θ)/2 θ |u|q−2 |∇u|2 dx ,
2 Rd
where we used (18.46). Therefore, u ∈ L3p (τ, Tmax (u0 ); Lq0 (Rd )) follows if we
can check that
Z Tmax (u0 ) Z 3pθ/q
|u|q−2 |∇u|2 dx dt < ∞.
τ Rd
q
The latter follows from (18.47) since our choice of θ satisfies θ 6 3p , which
follows from (18.43).
Lemma 18.3.6 (Chain rule in the weak setting). Let q ∈ [2, ∞) and
p ∈ (1, ∞). Suppose that u ∈ C([τ, T ]; Lq (Rd )) ∩ Lp (τ, T ; H 1,q (Rd )), G ∈
0
Lp (τ, T ; Lq (Rd ; Rd )), and g ∈ L1 (τ, T ; Lq (Rd )) are such that for all t ∈ [τ, T ]
Z t Z t
u(t) = u(τ ) + ∇ · G(s) ds + g(s) ds, (18.48)
τ τ
1+ρ σ
for all u, v ∈ X1 ∩ Oσ,p , where βj ∈ (σ, 1), ρj > 0 are such that βj 6 1+ρjj
for j ∈ {1, . . . , m}. Then for every T ∈ (0, ∞) there exist ε > 0 such that for
each kv0 kXσ,p 6 ε, the problem
(
u0 + Au = F (u), on (0, T ),
(18.53)
u(0) = v0 ,
has a unique Lpwα -solution uv0 ∈ MRpα (0, T ). Moreover, there is a C > 0 such
that for all kv0 kXσ,p , kv1 kXσ,p 6 ε,
kuv0 − uv1 kMRpα (0,T ) 6 Ckv0 − v1 kXσ,p . (18.54)
If additionally, A has maximal Lp -regularity (C-regularity if p = ∞) on R+
and 0 ∈ %(A), then the above holds with (0, T ) replaced by R+ .
726 18 Nonlinear parabolic evolution equations in critical spaces
p
Proof. In the proof we use the notation Ej = Lw α
(0, T ; Xj ). Let u0 = 0 and
set T1 = T . Without loss of generality we may assume T > 1 and r 6 1. Let
Φv0 : BrT (v0 ) → MRpα (0, T ) be defined by Φv0 (v) := u, where u is the unique
Lpwα -solution to
(
u0 + Au = F (v),
u(0) = v0 .
Note that for r ∈ (0, 1] and ε > 0 small enough, v takes values in Oσ,p by
Lemma 18.2.9, and by Lemma 18.2.10 we have F (v) ∈ E0 . Below Theorem
18.2.6, it has already been observed that local existence and uniqueness follow
if we can show that Φv0 has a unique fixed point.
p
By the maximal regularity estimate (18.28) we have u ∈ MRα (0, T ), u(0) =
v0 , and
where the estimate for F (v) follows from Lemmas 18.2.7 and 18.2.8, the con-
stant C can be taken T -independent since T > 1, and we used (18.19) with
u0 = 0 and zu0 = 0. Note that the terms T δj can be avoided due to the more
restrictive condition (18.52). The above estimate shows that for r, ε > 0 small
enough, kukMRpα (0,T ) 6 r, and thus u ∈ BrT (v0 ).
Next, fix vj,0 ∈ BXσ,p (u0 , ε) and vj ∈ BrT (vj,0 ) for j ∈ {1, 2}. Then u =
Φv1,0 (v1 ) − Φv2,0 (v2 ) solves the problem
(
u0 + Au = F (v1 ) − F (v2 ),
u(0) = v1,0 − v2,0 .
kukMRαp (0,T ) 6 CA,T kF (v1 ) − F (v2 )kE0 + CA,T kv1,0 − v2,0 kXσ,p ,
By (18.55), Φv0 : BrT (v0 ) → BrT (v0 ) is a strict contraction, and thus it has
a unique fixed point uv0 ∈ BrT (v0 ). This is the required solution to (18.53).
Moreover, (18.55) implies that for v1,0 , v2,0 ∈ BXσ,p (u0 , ε)
1
kuv1,0 − uv2,0 kMRpα (0,T ) 6 kuv1,0 − uv2,0 kMRpα (0,T )
2
+ CA,T kv1,0 − v2,0 kXσ,p ,
and thus
kuv1,0 − uv2,0 kMRpα (0,T ) 6 2CA,T kv1,0 − v2,0 kXσ,p .
which gives (18.54).
In case A has maximal regularity on R+ and 0 ∈ %(A), then (18.28) holds
with (0, T ) replaced by R+ . Moreover, one can check that Lemma 18.2.9 holds
with (0, T1 ) replaced by R+ . Therefore, one can repeat the above argument
on the half line.
18.5 Notes
Critical spaces
presented which works for all p ∈ [1, ∞] and all admissible weights, with p = ∞
corresponding to maximal C-regularity. Moreover, we do not need geometric
conditions on X0 such as the UMD property, or further conditions on A(u0 )
besides maximal Lp - or C-regularity. In part of the existing literature, the
spaces Xβj appearing in (18.11) are taken as the complex interpolation spaces
[X0 , X1 ]βj . Taking the real interpolation spaces (X0 , X1 )βj ,1 leads to a less
restrictive condition on Fc and is easier to work with in the proofs.
The case p = 1 of Theorem 18.2.6 seems to be new. It is important to
observe that for p = 1 one is forced to take σ = α = 0, which in turn forces
the X0 -valued trace part FTr to be defined on an open subset Oσ,p of the
same space X0 . For non-linearities of the form F = FTr , this requirement rules
out many interesting examples of non-linearities. However, by allowing non-
linearities with a critical part, i.e., non-linearities of the form F = FTr + Fc ,
many interesting examples can be covered even when p = 1, the point being
that it suffices to have Fc locally Lipschitz with respect to the norms of the
smaller spaces X1/(1+ρj ) (with the ρj ’s as in Assumption 18.2.2). On the other
hand, according to Theorem 17.4.5, operators with maximal L1 -regularity are
rare. An exception is the case where X0 itself is a real interpolation space in
which case the Da Prato–Grisvard theorem applies (see Corollary 17.3.20).
It should be observed that a more flexible condition on Fc could be used
in (18.11), namely
m
ρ ρ
X
kFc (u) − Fc (v)kX0 6 Lc (1 + kukXjϕ + kvkXjϕ )ku − vkXβj , (18.56)
j j
j=1
The formulation (18.56) allows for different space regularity for u, v, and u−v
on the right-hand side (see Agresti and Veraar [2022a] and Prüss, Simonett,
and Wilke [2018]). However, in all known examples, it suffices to take ϕj = βj
(as we do in the main text) in order to obtain the sharpest results. Note that
by taking ϕj = βj , (18.57) reduces to the sub-criticality condition (18.12).
of α). In the semi-linear case, the blow-up criteria can be further weakened as
was done in Theorem 18.2.17. In case of semi-linear functions F of quadratic
type, blow-up criteria appear in Prüss, Simonett, and Wilke [2018, Section
2.1]. Some of these were extended, for a more general class of semi-linearities
F , in to a stochastic setting in Agresti and Veraar [2022b, Theorem 4.11].
Simplifying this to the deterministic setting, one arrives at the following result:
Applications
The examples considered in Section 18.3 are very basic, and local/global
well-posedness is well known for a broad class of initial values. The examples
are merely chosen to demonstrate the abstract theorems of Section 18.2 in a
simple setting. The method to check the blow criteria in Example 18.3.5 is
taken from Agresti and Veraar [2023a], where these techniques are used in
several examples.
An extension of the results of Section 18.2 to stochastic quasi-linear evo-
lution equations in critical spaces was recently obtained in Agresti and Ver-
aar [2022a,b], where completely new proofs where required. Applications to
stochastic PDE can be found in these works, as well as in Agresti and Ver-
aar [2021, 2022c, 2023b,a], Agresti [2022], Agresti, Hieber, Hussein, and Saal
[2022a,b].
Q
Questions
Calderón–Zygmund operators
The extrapolation theory for the boundedness of Calderón–Zygmund opera-
tors developed in Chapter 11 is in many ways analogous and parallel to the
extrapolation of Lp inequalities for martingale transforms that we discussed
in Section 3.5. Specifically, the quantitative statement of Calderón–Zygmund
Theorem 11.2.5(3) is analogous to the estimate (3.48) of Martingale Extrapo-
lation Theorem 3.5.4; in both cases, the Lp norm of an operator is controlled
by the Lq norm multiplied by the factor pp0 = p+p0 , which exhibits the correct
blow up of these norms as p → 1 or p → ∞. However, in the case of martingale
transforms with respect to a Paley–Walsh filtration, (3.49) gives a more pre-
0
cise estimate with the factor pq + pq0 . While this has the same rate of blow up
as p → 1 or p → ∞, it gives a better estimate if the “starting point” q is either
large or close to 1. (For instance, think of the case that q is large and p = 2q.)
Given the well-behaved nature of the Lebesgue measure with respect to which
the Calderón–Zygmund singular integrals are integrated, it seems reasonable
to expect that the behaviour of these operators should be as good as that of
martingale transforms with respect to a Paley–Walsh filtration; however, this
is not reflected in the quantitative estimates of Calderón–Zygmund Theorem
11.2.5. We therefore pose the question:
Problem Q.1. Under the assumptions of Calderón–Zygmund Theorem 11.2.5
(or even just for more regular operators with a standard kernel), is there an
estimate
? p p0
kT kL (Lp (Rd ;X),Lp (Rd ;Y )) 6 cd + 0 kT kL (Lq (Rd ;X),Lq (Rd ;Y )) + CK ,
q q
for all p, q ∈ (1, ∞), where cd depends only on the dimension and CK only on
the kernel K of the operator T , either via the quantities kKkHör and kKkHör∗
from Definition 11.2.1 of a Hörmander kernel, or cK and ωK from Defini-
tion 11.3.1 of a Calderón–Zygmund kernel? In particular, does the Hilbert
transform satisfy an estimate
© Springer Nature Switzerland AG 2023 733
T. Hytönen et al., Analysis in Banach Spaces, Ergebnisse der Mathematik und ihrer
Grenzgebiete. 3. Folge / A Series of Modern Surveys in Mathematics 76,
https://doi.org/10.1007/978-3-031-46598-7
734 Q Questions
? p p0
~p,X 6 C + ~q,X , ~p,X := kHkL (Lp (R;X)) (Q.1)
q q0
with some universal constant C?
The prospective estimate (Q.1) would be an analogue of Theorem 4.2.7, which
gives a similar bound for the UMD constants βp,X and βq,X in place of ~p,X
and ~q,X .
Recall that Problem O.6 asks about a possible linear dependence between
βp,X and ~p,X for p = 2, which remains wide open. Note that if this linear
relation was true for all p ∈ (1, ∞), and moreover with constants indepen-
dent of p, then (Q.1) would immediately follow via this linear relation from
Theorem 4.2.7 for the UMD constants. Thus, (Q.1) could be thought of as a
simpler model problem related to the presumably difficult Problem O.6.
As we have seen in Chapter 11, and many more examples can be found in
the literature, sparse domination of operators is very efficient in capturing
essential information about their boundedness properties on various function
spaces, particularly in view of sharp weighted norm inequalities. As discussed
in the Notes of that Chapter, convex body domination provides a useful elab-
oration in view of applications like matrix-weighted inequalities and commu-
tator estimates. The proofs of existing convex body domination results follow
the same broad outline as their sparse domination counterparts, but require
elaborations at critical points of the argument. Rather than reworking each
sparse domination proof for the convex body improvement, it would be useful
to have a general statement guaranteeing that one implies the other—or, to
know that such a statement is impossible, justifying the need of case-by-case
study. This raises the question:
Problem Q.2. Does sparse domination imply convex body domination?
More precisely, let 1 6 p < r < q 6 ∞ and
T ∈ L (Lr (Rd ; X), Lr (Rd ; Y )).
For each N ∈ Z+ , consider the following property:
For some constants ε ∈ (0, 1) and α, C ∈ [1, ∞), for all f = (fn )N n=1 ∈
L∞c (R d
; X) N
and g = (g ) N
n n=1 ∈ L ∞
c (R d
; Y ∗ N
) (or some other suitable
test function spaces), there is an ε-sparse collection S ⊆ D such that
|hT f, gi|
N
(Q.2)
X C ZZ X
6 sup hfn (s), φ(s)ihψ(t), gn (t)i ds dt,
|Q| φ,ψ αQ×αQ n=1
Q∈S
0
where the supremum is over all φ ∈ Lp (αQ; X ∗ ) and ψ ∈ Lq (αQ; Y )
such that
Q Questions 735
Z Z
p0
− kφ(s)kX ∗ ds 6 1, − kψ(t)kqY dt 6 1.
αQ αQ
If this property holds for N = 1, does it follow (a) for N = 2, or even (b) for
all N ∈ Z+ ? If not, what is a counterexample? Does the implication hold
• under additional assumptions on the operator T ?
• with a relaxed conclusion with p + ε and q − ε in place of p and q?
Since sparse domination implies (weighted) boundedness, this leads to the fol-
lowing natural question whether their results remains true for general UMD
spaces. In particular this would provide a subclass of operators for which we
can answer Problem Q.3.
Problem Q.4. Let X be a UMD space and let p ∈ (1, ∞). Let ε > 0, and
T ∈ L (Lp (Rd )) be an operator such that for any f, g ∈ Lc∞ (Rd ) there exists
a ε-sparse collection of cubes S such that
Z XZ Z
|T f | · |g| dx 6 − |f | dx− |g| dx |S|.
Rd S∈S S S
such that
Z XZ Z
kT ⊗ IX f kX · |g| dx 6 CX − kf k dx− |g| dx |S|?
Rd S∈S S S
Paraproducts
Problem Q.5. What is the largest class of Banach spaces X such that
By what we just discussed, this class should at least include all UMD spaces
and all spaces of martingale type 2.
Q Questions 737
T (1) theorems
In the various T (1) theorems proved in Sections 12.3 and 12.4 and discussed
in the related Notes, we have seen the following:
(1) Under assumptions on the Haar coefficient of a bilinear form with respect
to a fixed dyadic system, the induced operator satisfies Lp (Rd ; X) bounds
that are cubic in the UMD constant βp,X (see Theorem 12.3.26).
(2) Under assumptions on the Haar coefficient of a bilinear form with re-
spect to an ensemble of dyadic system, the induced operator satisfies
Lp (Rd ; X) bounds that are quadratic in the UMD constant βp,X (see The-
orem 12.3.35). Such assumptions are verified by weakly defined Calderón–
Zygmund operators with standard kernels (see Theorem 12.4.12), and in
particular by extensions of scalar-valued Calderón–Zygmund operators on
Lp (Rd ) (see Theorem 12.4.21).
(3) Under additional symmetry and smoothness assumptions, the induced
operator satisfies Lp (Rd ; X) bounds that are linear in the UMD constant
βp,X . These results were not covered in the present volume, but can be
found in the works of Geiss et al. [2010] for a class of Fourier multipliers,
and of Pott and Stoica [2014] for a class of Calderón–Zygmund operators
in dimension d = 1.
A major problem is the following:
Problem Q.6. Is there an upper bound that is linear in the UMD constant for
all operators in the scope of Theorems 12.3.26, 12.3.35, 12.4.12, and 12.4.21?
If not, what is a counterexample?
Problem Q.7. Can any of the bounds in (1) through (3) be improved? In par-
ticular, does T (1) Theorem 12.3.26 with a single dyadic system allow bounds
that are quadratic in βp,X ? Do Figiel’s elementary operators from Theorems
12.1.25 and 12.1.28 admit such bounds? If not, what is a counterexample?
Given the relatively wide scope of estimates that one can prove with bounds
quadratic in the UMD constant (in contrast to the somewhat restricted class of
linear estimates currently available), Problem Q.7 would appear to be more
approachable than the presumably very hard Problem Q.6. In view of the
currently different quantitative bounds in (1) and (2) above, Problem Q.7 has
a “philosophical” dimension concerning the role of random dyadic systems
738 Q Questions
Problem Q.8. Is the interdependence between the required decay of the Haar
coefficients (resp. modulus of smoothness of the kernel) and type and cotype
of the spaces in T (1) Theorems 12.3.26 and 12.3.35 (resp. 12.4.12 and 12.4.21)
sharp? If not, what are the sharp forms of these theorems? What are exam-
ples of bilinear forms or operators satisfying weaker forms of the assumptions
and failing the conclusions of these theorems concerning boundedness of the
induced operator?
This problem is relevant and open even in the scalar-valued case with type
and cotype 2, in which case the existing T (1) theorems require Figiel or Dini
norms of order 21 . Nevertheless, the fact that the Banach space valued theory
“interprets” the seemingly arbitrary number 21 as max( 1t , q10 ) with the type
and cotype exponents t and q could suggest the construction of possible coun-
terexamples through extremal situations for the type and cotype inequalities
even in the scalar-valued case.
On the side of the open-ended Problem Q.8 we pose the more provocative:
Problem Q.9. Is any of the T (1) theorems valid for all classical Dini kernels?
If not, what is an example of an operator with a Dini kernel, satisfying all
assumptions of a T (1) theorem yet failing to be bounded on L2 (Rd ) or on
L2 (Rd ; X) for some UMD space X?
Fourier type
Problem Q.10. If a Banach space X has type p and cotype q with p1 − 1q < 12 ,
does it follow that X has Fourier type r determined by 1r = p1 − 1q + 12 ? If not,
what is a counterexample?
In Theorem 14.7.15 we have seen that for UMD spaces 1R+ d acts as pointwise
s,p 0
multiplier on H (R; X) for p ∈ (1, ∞) and −1/p < s < 1/p. The same result
holds for general dimension d, and there exist at least three different proofs
of this fact (see Meyries and Veraar [2015], Lindemulder [2017], Lindemulder,
Meyries, and Veraar [2018]). However, it is not known whether the UMD
property or any other condition on X is necessary.
Problem Q.12. Let p ∈ (1, ∞) and s ∈ (0, 1/p). Characterise those Banach
spaces X for which 1Rd+ acts as pointwise multiplier on H s,p (Rd ; X).
k
Theorem 14.8.4, due to Hytönen and Merikoski [2019], states that Bp,p (Rd ; X)
k,p d
embeds continuously into W (R ; X) if and only if X has martingale cotype
p (see Section 3.5.d).
Problem Q.13. Let p ∈ (1, 2), k ∈ N and s ∈ R. Characterise those Banach
spaces X for which one has continuous embeddings
k
W k,p (Rd ; X) ,→ Bp,p (Rd ; X) (Q.3)
s,p d s
H (R ; X) ,→ Bp,p (Rd ; X). (Q.4)
s
A similar question can be asked with Bp,p (Rd ; X) replaced by Fp,qs
(Rd ; X) or
s d
Bp,q (R ; X). Moreover, by Proposition 14.7.8, each of the embeddings (Q.3)
and (Q.4) implies that X has type p. In case of UMD Banach spaces X, type
p is also sufficient for these embeddings. Therefore, it is natural to conjecture
that (Q.3) and (Q.4) are both equivalent to martingale type p.
Theorem 14.5.1, due to Kalton, Van Neerven, Veraar, and Weis [2008],
characterises the Banach spaces X for which the Sobolev embedding
( 1 − 12 )d
p
Bp,p (Rd ; X) ,→ γ(L2 (Rd ), X)
holds as the Banach space that have type p. A corresponding result for the
converse embedding, with ‘p’ replaced by ‘q’, characterises Banach spaces with
cotype q. It is natural to ask for similar embeddings for Bessel potential spaces
and Triebel-Lizorkin spaces.
Problem Q.14. Let p ∈ (1, 2) and r ∈ (p, ∞). Characterise those Banach
spaces X for which one has continuous embeddings
1 1
H ( p − 2 )d,p (Rd ; X) ,→ γ(L2 (Rd ), X) (Q.5)
1
(p − 12 )d
Fp,r (Rd ; X) ,→ γ(L2 (Rd ), X). (Q.6)
In Corollaries 14.6.18 and 14.7.7 it was shown that having type r > p is
sufficient. Moreover, it is known that each of the embeddings implies type p,
that and the embedding (Q.5) holds for p-convex Banach lattices; see Veraar
Q Questions 741
Functional calculus
Several open problems related to the H ∞ -functional calculus have been stated
in Volume II. Theorem 15.3.9 is a classical result of Seeley [1971] and states
that for a sectorial operator A with bounded imaginary powers, the domain
of fractional powers D(Aθ ) coincides with the complex interpolation space
[X, D(A)]θ . A converse holds for Hilbert spaces X, where boundedness of imag-
inary powers characterises the boundedness of the H ∞ -calculus. For Banach
spaces, boundedness of imaginary powers does not imply the boundedness of
the H ∞ -calculus (see Example 10.2.32).
Maximal regularity
via the results in Geiss, Montgomery-Smith, and Saksman [2010]. Our lower
estimate (Q.7) was obtained via an anisotropic extension of part of their
result. It seems an interesting problem to try to extend the techniques in the
latter paper to find a full analogue of the estimates (Q.8) for the maximal
Lp -regularity constant of the Laplace operator:
reg
Problem Q.17. Let Mp,−∆ (R+ ) be the maximal Lp -regularity constant of
−∆ on Lp (Rd ; X) on R+ , where X is a UMD space. We ask if exist universal
constants 0 < c 6 C < ∞ such that
reg
c max{βp,X , ~p,X } 6 Mp,−∆ (R+ ) 6 C(βp,X + ~p,X ).
In Theorem 17.4.8 we have seen that there are operators A such that −A
generates an analytic semigroup on Lq , but A does not have maximal Lp -
regularity on finite time intervals. This provided a negative answer to Brezis’s
question as explained in the notes of Chapter 17. The question still remains
open for differential operators.
Problem Q.18. Let O ⊆ Rd be open, and let A be a differential operator
on Lp (O) with p ∈ (1, ∞) such that −A generates an analytic semigroup on
Lp (O). Does A have maximal Lp -regularity?
Some evidence in favor of having maximal Lp -regularity can be found in
Blunck and Kunstmann [2002], Kunstmann [2008] for operators in divergence
and non-divergence form respectively.
In Theorem 17.2.15 we have seen that maximal Lp -regularity of A implies
that −A generates an analytic semigroup. Such a result seems unavailable for
time-dependent operators A without any further conditions. Because of the
time-dependence, generation of a semigroup has to be replaced by generation
of an evolution family. A family of bounded operators (S(t, s))06s6t6T on a
Banach space X is called an evolution family if
(1) S(t, t) = I for all t ∈ [0, T ];
(2) S(t, s)S(s, r) = S(t, r) for all 0 6 r 6 s 6 t 6 T ;
It is called strongly continuous if S : {(t, s) ∈ [0, T ]2 : s 6 t} → L (X) is
strongly continuous. References for the theory of evolution families include
Amann [1995], Engel and Nagel [2000], Lunardi [1995], Pazy [1983], Tanabe
[1979] and Yagi [2010].
Problem Q.19. Let X0 and X1 be Banach spaces, with X1 continuously
and densely embedded in X0 , and let p ∈ [1, ∞]. Suppose that A : [0, T ] →
L (X1 , X0 ) is strongly measurable in the uniform operator topology, and that
there exist a constant 0 < C < ∞ such that
Suppose further that A has maximal Lp -regularity in the sense that for all
f ∈ Lp (0, T ; X0 ) there exists a unique strong solution u ∈ W 1,p (0, T ; X0 ) ∩
Q Questions 743
Lp (0, T ; X1 ) with u(0) = 0 to the problem u0 (t) + A(t)u = f (t) for t ∈ [0, T ].
Does there exists a strongly continuous evolution family (S(t, s))06s6t6T on
X0 such that for all f ∈ Lp (0, T ; X0 ) the strong solution u is equals
Z t
u(t) = S(t, s)f (s) ds, t ∈ [0, T ]?
0
The problem is also open if additionally one assumes that for each for fixed
t0 ∈ [0, T ] the operator −A(t0 ) generates an analytic semigroup on X0 . By a
result of Prüss and Schnaubelt [2001], for A ∈ C([0, T ]; L (X1 , X0 )) with the
property that −A(t0 ) generates an analytic semigroup group in X0 for each
fixed t0 ∈ [0, T ], an associated evolution family (S(t, s))06s6t6T .
The following converse to Problem Q.19 is also open.
Problem Q.20. Let X0 and X1 be UMD spaces, with X1 continuously and
densely embedded in X0 . Suppose that A : [0, T ] → L (X1 , X0 ) is strongly
measurable in the uniform operator topology, and that there exist a constant
0 < C < ∞ such that
Suppose that for all 0 6 s < t 6 T we have S(t, s) ∈ L (X0 , X1 ) and the
families
{S(t, s) : 0 6 s 6 t 6 T }
{(t − s)A(r)S(t, s) : 0 6 s 6 t 6 T, r ∈ [0, T ]}
for every x ∈ X as a Bochner integral, and one may use that bounded
operators can be pulled through such integrals;
K.1 Measurable semigroups 747
• S ∗ (t)Y ⊆ Y for all t > 0: for then one has, for all x∗ ∈ Y ,
hS(t)x − S(s)x, x∗ i
= hR(λ, G)S(t)y − R(λ, G)S(s)y, x∗ i
Z ∞ Z ∞
−λr ∗
= e hS(t + r)y, x i dr − e−λr hS(s + r)y, x∗ i dr
0 0
Z ∞ Z ∞
= eλt e−λr hS(r)y, x∗ i dr − eλs e−λr hS(r)y, x∗ i dr
t s
Z ∞ Z t
−λr ∗
λt
= (e − e ) λs
e hS(r)y, x i dr − eλs e−λr hS(r)y, x∗ i dr.
t s
Fox fixed t > 0, the right-hand side is a continuous function of s > 0. Hence
this is also true for the left-hand side, from which we infer that s 7→ hS(s)x, x∗ i
is continuous. Dividing by t − s and letting s → t, by the continuity just
observed we obtain
D S(t)x − S(s)x E Z ∞
∗
lim , x = λe λt
e−λr hS(r)y, x∗ i dr − hS(t)y, x∗ i
s→t t−s t
748 K Measurable semigroups
(3): Repeating the argument of (2) with S(s)x replaced by x, and using
that Y is norming, the estimate (K.1) implies, for λ > ω and t > 0,
where the first identity follows from the fact that G generates S and the second
from the fact that G0 generates S0 := S|D(G) , both in the sense of Definition
K.1.2; by strong continuity, there is no need to evaluate against functionals
in Y . Since the left-hand side belongs to D(G), so does the right-hand side x.
Applying λ − G to both sides, we obtain the identity (λ − G0 )x = (λ − G)x.
Since the former belongs to D(G), so does the latter. This proves that x ∈
D(G1 ) and G0 x = G1 x.
In the converse direction, if x ∈ D(G1 ), then writing x = R(λ, G)y with
y ∈ X gives, for all x∗ ∈ Y ,
Z ∞
∗
hx, x i = e−λt hS(t)y, x∗ i dt
0
K.1 Measurable semigroups 749
Z ∞
= e−λt hS(t) (λ − G)x, x∗ i dt
0 | {z }
∈D(G)
Z ∞
= e−λt hS0 (t)(λ − G)x, x∗ i dt = hR(λ, G0 )(λ − G)x, x∗ i,
0
where the last step follows from the fact that G0 generates S0 = S|D(G) in the
sense of Definition K.1.2. It follows that x = R(λ, G0 )(λ − G)x ∈ D(G0 ).
Proof. We have kSµ (t)k 6 M e(ω−<µ)t , {λ ∈ C : <λ > ω − <µ} ⊆ %(Gµ ), and
if <λ > ω − <µ, then <(λ + µ) > ω, and therefore for all x ∈ X and x∗ ∈ Y
we have
Z ∞ Z ∞
−λt ∗
e hSµ (t)x, x i dt = e−(λ+µ)t hS(t)x, x∗ i dt
0 0
= hR(λ + µ, G)x, x∗ i = hR(λ, Gµ )x, x∗ i.
The following proposition provides an analogue of Proposition G.2.3(3), which
states that if S is aR C0 -semigroup on X with generator A, then for all x ∈ X
t
and t > 0 one has 0 S(s)x ds ∈ D(A) and
Z t
A S(s)x ds = S(s)x − x,
0
The difficulty in the present set-up is that the integrals of the semigroup orbits
make no a priori sense. Establishing that the integrals do indeed exist in X
as “weak Y -integrals” is part of our task in proving the proposition. In the
strongly measurable case, all this poses no problems and in the proposition
below one can simply take
Z t
0
xt = S(s)x ds
0
and
Gxt0 = S(t)x − x.
If in addition x ∈ D(G), we furthermore have
Z t
hGx0t , x∗ i = hS(s)Gx, x∗ i ds, x∗ ∈ Y.
0
then integrating the first identity over [0, t] with t > 0, applying x∗ , and
subtracting the second identity, we are left with the identity thx, x∗ i = 0,
valid for all x∗ ∈ Y , and therefore x = 0. But for x = 0, uniqueness is clear.
Step 2 – For the proof of existence, fix an arbitrary λ > ω and consider
the rescaled semigroup Sλ generated by Gλ as in Proposition K.1.6. This
semigroup is uniformly exponentially stable, by which we mean that it satisfies
(K.1) with a negative exponent. In particular, Gλ is boundedly invertible. The
0
element xt,λ := Gλ−1 (Sλ (t)x − x) belongs to D(Gλ ) = D(G) and satisfies
0
Gλ xt,λ = Sλ (t)x − x.
belongs to D(G). Indeed, using the notation used in the proof of Proposition
K.1.5 it belongs to D(G0 ), and we have D(G0 ) ⊆ D(G). Since also x0t,λ ∈ D(G)
(by Step 2), we conclude from the representation (K.3) that x0t belongs to
D(G), and we have
Z t
Gx0t = eλt Gx0t,λ − λG eλs x0s,λ ds
0
Z t
−1 λt
= GGλ (S(t)x − e x) − λG eλs (Sλ (s)G−1 −1
λ x − Gλ x) ds
0
Z t
−1 λt
= (I + λGλ )(S(t)x − e x) − λG S(s)G−1
λ x ds
0
+ (eλt − 1)GG−1
λ x
(∗)
= (I + λG−1 λt −1 −1
λ )(S(t)x − e x) − λ(S(s)Gλ x − Gλ x)
+ (eλt − 1)(I + λG−1
λ )x
= S(t)x − x,
Finally, applying this with G−1 x in place of x and noting that (G−1 x)0t =
G−1 x0t , we obtain
Z t
hx0t , x∗ i = hS(s)x, x∗ i ds.
0
for the unique element x0 ∈ D(G) satisfying the conclusions of the proposition.
With this notation, we have the following result.
Proposition K.1.8. Under the assumptions of Proposition K.1.7, for all
s, t > 0 and x ∈ X we have
Z t Z t
S(s) S(r)x dr = S(r + s)x dr,
0 0
R t+s Rs
where the right-hand side is shorthand for 0
S(r)x dr − 0
S(r)x dr.
Proof. As in the proof of Proposition K.1.7, this is easy for the rescaled semi-
group Sλ . Indeed, for all ∈ X and x∗ ∈ Y we have
0 0
Gλ Sλ (s)xt,λ = Sλ (s)Gλ xt,λ
0 0
= Sλ (t)[Sλ (s)x] − [Sλ (s)x] = Gλ xt+s,λ − Gλ xs,λ .
These identities imply that S(s)x0t,λ satisfies the two properties of Proposition
K.1.7 with x replaced by S(s)x. By uniqueness, in the notation introduced
above this gives the desired identity
Z t Z t
Sλ (s) Sλ (r)x dr = Sλ (r + s)x dr.
0 0
The general case can again be deduced from the rescaled case, by similar
arguments as before. We leave the details to the reader.
K.1 Measurable semigroups 753
We proceed with some important examples. The first demonstrates the con-
sistency of Definition K.1.2 with the corresponding Definition G.2.1 for C0 -
semigroups.
Remark K.1.12. Suppose that A is a densely defined sectorial operator of an-
gle ω(A) < 21 π. By Theorem G.5.2, −A generates a bounded analytic C0 -
semigroup S in the sense of Definition G.2.1. Since the adjoint operator A∗ is
sectorial and has the same angle, combination of the above two results shows
that −A∗ generates the bounded analytic (but not necessarily C0 ) semigroup
S ∗ in the sense of Definition K.1.2.
The ‘only if’ part is obvious. To prove the ‘if’ part, we show that kS(t)k 6
M e−ωt with
Indeed, let t > 0 be fixed and write t = (n + θ)t0 with θ ∈ [0, 1) and n ∈ N.
Then, with the above choices of ω and M ,
Proof. The implication (1)⇒(2) part is clear. To prove the converse implica-
tion (2)⇒(1), by the closed graph theorem there exists a constant C > 0 such
that kS(·)xkLp (R+ ;X) 6 Ckxk for all x ∈ X. Let M > 1 and ω > 0 be such
that kS(t)k 6 M eωt for all t > 0. Then
t
1 − e−ωpt
Z
kS(t)xkp = e−wp(t−s) kS(t − s)S(s)xkp ds 6 M p C p kxkp .
ωp 0
Therefore kS(t)k 6 CKt−1/p and the result follows from the preliminary
observation.
Below we will see that S ∗ f can often be interpreted as the so-called mild
solution to an abstract Cauchy problem.
Proof. (1)⇒(2): If kS(t)k 6 M e−ωt for all t > 0, with ω > 0, then kS∗f (t)k 6
φ ∗ f (t) for all t > 0, where φ(s) = M e−ωs 1R+ (s). Taking Lp -norms, Young’s
inequality gives S ∗ f ∈ Lp (R+ ; X) and
kS ∗ f kLp (R+ ;X) 6 kφkL1 (R+ ) kf kLp (R+ ;X) 6 M ω −1 kf kLp (R+ ;X) .
(2)⇒(1): By the closed graph theorem there exists a constant C > 0 such
that
kS ∗ f kLp (R+ ;X) 6 Ckf kLp (R+ ;X) , f ∈ Lp (R+ ; X).
First consider p ∈ [1, ∞). Let M > 1 and µ > 0 be such that kS(t)k 6 M eµt
for all t > 0. Fix ε > 0 and x ∈ X, and set f (t) := e−(µ+ε)t S(t)x for t > 0.
Then
M
kf kLp (R+ ;X) 6 kxk
(εp)1/p
and
t
1 − e−(µ+ε)t
Z
S ∗ f (t) = S(t − s)f (s) ds = S(t)x.
0 µ+ε
Fixing any τ > 0 we obtain
using that kf (t)k 6 M e−εt kxk in the last inequality. Proposition K.2.1 now
gives the required uniformly exponential stability.
For p = ∞ the above argument can be repeated to give the bound
kS(t)xk = t−1 kS ∗ fx (t)k 6 t−1 Ckfx kL∞ (R+ ;X) 6 t−1 Kkxk.
As a consequence, for any integer n > 1, S(t) maps X into D(An ) and
The following result shows that the quantity (K.4) cannot be arbitrarily
small unless the operator A is bounded. This fact plays a role in a construction
of a counterexample in Chapter 17 (see Theorem 17.4.4).
Proposition K.3.1. Let A be sectorial of angle < 12 π and let S be the locally
bounded strongly measurable semigroup generated by −A. If
1
lim sup tkAS(t)k < ,
t↓0 e
then A is bounded.
Proof. For all t > 0 and x ∈ X we have −AS(t)x = S 0 (t)x. To see this, we first
use Hille’s theorem to move A into the integral, then we write A = (A − z) + z
and use that Z
1
e−zt x dz = 0
2πi Γ
by Cauchy’s theorem, and note that
Z
1
ze−zt R(z, A)x dz = S 0 (t)x
2πi Γ
by differentiation under the integral sign. As a consequence, for all fixed t > 0
the assumption of the lemma implies
t 0 t 1
lim sup S < .
n→∞ n n e
By induction, S (n) (t) = (−A)n S(t) = (−AS( nt ))n x = (S 0 ( nt ))n . The in-
n
equality nn /n! 6 en implies that lim supt→∞ nn! k(S 0 ( nt ))n k < 1, and therefore
there exists a δ > 0 such that for every t > 0 the series
758 K Measurable semigroups
∞ ∞
X 1 X (z − t)n nn t 0 t n
S(z)
e := (z − t)n S (n) (t) = S
n=0
n! n=0
tn n! n n
implies that
Z t+h Z h 1 Z t −1
1 1
(S(h) − I) = S(s) ds − S(s) ds S(s) ds .
h h t 0 t 0
(1)
Proof. Denote the two sets on the right-hand side in (1) and (2) by Xθ,p
(2)
and Xθ,p , respectively. We will write K(t, x) := K(t, x; X, D(A)) for the K-
functional of the real interpolation method (see Appendix C).
For (1) we will prove continuous inclusions
(1)
(X, D(A))θ,p ⊆ Xθ,p ⊆ (X, D(A))θ,p ,
It follows that
(1)
We next prove the inclusion Xθ,p ⊆ (X, D(A))θ,p . Suppose x ∈ X is such
that λ 7→ λθ A(λ + A)−1 x belongs to Lp (R+ , dt
t ; X). Then, using the decom-
position x = A(λ + A)−1 x + λ(λ + A)−1 x ∈ X + D(A), we obtain
kλθ A(λ + A)−1 xkLp (R+ , dt ;X) > kλ 7→ λθ A(λ + A)−1 xkLp ((0,r), dt ;X)
t t
1
> kxkkλ 7→ λθ kLp ((0,r), dt ) =: Cθ,p,A kxk.
2 t
K.4 An interpolation result 761
(2) (2)
To prove the inclusion Xθ,p ⊆ Xθ,p , suppose that x ∈ X is such that
t 7→ t−θ (S(t)x − x) ∈ Lp (R+ , dtt ; X). Using the identity A(λ + A)
−1
=
−1
λ(λ + A) x − x and the Laplace transform representation of the resolvent
(Proposition G.4.1), for λ > 0 we have
Z ∞
−1
A(λ + A) = λe−λt (S(t)x − x) dt.
0
R∞
where Cθ,p = 0 µθp+1 e−µ dµ
µ = Γ (θp+1). This gives the desired inclusion for
R ∞ θ+1 −λt θ
1 6 p < ∞. For p = ∞ we note that for all λ > 0 we have 0 λ e t dt =
R ∞ θ −s
0
s e ds = Γ (1 + θ), and therefore
with Cθ = Γ (1 + θ).
Assume finally that kS(t)k 6 M e−ωt for all t > 0, with M > 1 and
ω > 0. Choose R = RM,ω > 0 so large that kS(t)k 6 21 for t > R. Then
kS(t)x − xk > 21 kxk for t > R. If 1 6 p < ∞, it follows that
Z ∞
−θ p dt
t 7→ t (S(t)x − x) p dt > t−θp (kxk/2)p =: CR,θ,p kxkp .
L (R+ , t ;X) R t
This gives the equivalence with the homogeneous norm for 1 6 p < ∞. For
p = ∞ we simply note that
K.5 Notes
L.1 Preliminaries
tK(t−1 , x; X0 , X1 ) = K(t, x; X1 , X0 ).
K(t, x) = K(t, x; X0 , X1 ).
For 0 < θ < 1 and 1 6 p 6 ∞, the real interpolation space (X0 , X1 )θ,p is the
Banach space defined by
(X0 , X1 )θ,p := x ∈ X0 + X1 : kxkθ,p < ∞ ,
with norm
One has (X, X)θ,p = X and (X0 , X1 )θ,p = (X1 , X0 )1−θ,p with identical norms.
By Lemma C.3.12, X0 ∩ X1 is dense in (X0 , X1 )θ,p whenever 0 < θ < 1 and
1 6 p < ∞. In what follows we let
(X0 ,X1 )θ,∞
Xp,∞ := X0 ∩ X1 (L.1)
kxkθ,p 6 kxk1−θ θ
X0 kxkX1 , x ∈ X0 ∩ X1 . (L.2)
In (C.6) we have seen that for all 0 < θ < 1 and 1 6 p0 6 p1 6 ∞ we have
the continuous inclusion
The next result shows that more can be said in the special case when X1 ⊆ X0
with continuous inclusion mapping (we write X1 ,→ X0 in this situation).
X1 ,→ (X0 , X1 )θ,p ,→ X0
with
kxkX0 6 C θ kxkθ,p 6 CkxkX1 , x ∈ X1 .
(2) For all 0 < θ0 < θ1 < 1, and all 1 6 p0 , p1 , p 6 ∞ we have a continuous
embedding
with
min{1, t}kxkX0 6 kx0 kX0 + tkx1 kX0 6 kx0 kX0 + Ctkx1 kX1 .
using the contractivity of the inclusion (L.3). Also, by (L.2) and the inequality
just proved,
1−θ 1−θ
kxkθ,p 6 kxkX 0
kxkθX1 6 C θ(1−θ) kxkθ,p kxkθX1 .
Definition L.2.1 (Trace method). For p ∈ [1, ∞] and θ ∈ (0, 1), the space
(X0 , X1 )Tr
θ,p
where the infimum over u extends over all strongly measurable functions
u : (0, ∞) → X0 + X1 with the above three properties. Note that if
u : (0, ∞) → X0 + X1 is strongly measurable and satisfies (i) and (ii), then
u ∈ W 1,1 ((0, T ); X0 + X1 ) for all 0 < T < ∞; in particular, u is equal al-
most everywhere to a (uniquely defined) continuous function from [0, ∞) to
X0 + X1 . In condition (iii), we refer to this version when imposing u(0) = x.
We continue with a technical result which shows that, in the definition of
the trace method, we may restrict ourselves to functions u ∈ C 1 ((0, ∞); X0 ∩
X1 ) without changing the norm defined by (L.4). Indeed, this is due to the
fact that the constant ε > 0 can be taken arbitrarily small in both inequalities
in (L.6).
Proposition L.2.2. For ε > 0 let gε : [0, ∞) → R be the ‘tent shaped’ piece-
wise linear function which is identically zero on [0, 1] and [1 + 2ε, ∞) and
whose graph connects the points (1, 0), (1 + ε, ε−1 ) and (1 + 2ε, 0) linearly. Let
ϕε (t) := tgε (t). Let u : (0, ∞) → X0 + X1 be strongly measurable and satisfy
conditions (i) and (ii) of Definition L.2.1, and define uε : (0, ∞) → X0 + X1
by
Z ∞ Z ∞
dτ dτ
uε (t) := ϕε (t/τ )u(τ ) = ϕε (τ )u(t/τ ) . (L.5)
0 τ 0 τ
Then uε ∈ C([0, ∞); X0 + X1 ) ∩ C 1 ((0, ∞); X0 ∩ X1 ), we have uε (0) = u(0),
and for all 0 < θ < 1,
and
2(1 + 2ε)2
kt 7→ t2−θ u0ε (t)kLp (R+ , dt ;X1 ) 6 kt 7→ t1−θ u(t)kLp (R+ , dt ;X1 ) .
t ε t
−θ
where we used (L.7) and the fact that t 6 1 on the support of ϕε . Similarly,
Finally, in view of
Z ∞
dτ
t2−θ u0ε (t) = (t/τ )2−θ ϕ0ε (t/τ ) · τ 1−θ u(τ ) ,
0 τ
where the last equality follows by exact computation of the L1 -norm in the
preceding line.
(X0 , X1 )Tr
θ,p = (X0 , X1 )θ,p
Proof. Let x ∈ (X0 , X1 )Trθ,p , choose u such that Definition L.2.1(i)-(iii) hold,
and let uε with 0 < ε < 1 be as in Proposition L.2.2.
Setting v(t) = tu0ε (t), for j ∈ {0, 1} we find
The existence of the limits follows from the convergence of the integral. From
Proposition L.2.2 we see that limt→∞ uε (t) = 0 in X1 . From the definition of
uε we obtain limt→0 uε (t) = uε (0) = u(0) = x.
From Theorem C.3.14 it follows that x ∈ (X0 , X1 )θ,p and, using the nota-
tion of the theorem,
2
(1 + 2ε)
6 8Cθ max kt 7→ t1−θ u(j) (t)kLp (R+ , dt ;Xj ) .
ε j∈{0,1} t
With ε = 1/2, taking the infimum over all admissible functions u gives the
bound
kxkθ,p 6 64Cθ kxk(X0 ,X1 )Tr
θ,p
.
In the converse direction, suppose that x ∈ (X0 , X1 )θ,p and fix ε > 0. By
R ∞ C.3.14 there exists a strongly measurable v : (0, ∞) → X0 ∩ X1 such
Theorem
that 0 v(t) dt
t = x and for j ∈ {0, 1},
kt 7→ t1−θ u0 (t)kLp (R+ , dt ;X0 ) 6 kt 7→ t−θ v(t)kLp (R+ , dt ;X0 ) 6 12(1 + ε)kxkθ,p .
t t
If, in addition to the assumptions already made, we also assume that 0 ∈ %(A),
then S is uniformly exponentially stable (because A − ε is sectorial in that
case and hence generates a bounded analytic semigroup), and we can use
u(t) = S(t)x in the above proof instead. This gives
and consequently
= max{M1 , M2 }kxk(X,D(A))θ,p .
Since also kxk 6 Ckxk(X,D(A))θ,p , this concludes the proof for the case I = R+ .
In case I = (0, T ), we use a simple scaling argument. For this let M > 0
and ω > 0 be such that e−ωt kS(t)k 6 M e−t and e−ωt ktAS(t)k 6 M e−t for
all t > 0, and set Sω (t) = e−ωt S(t). In both of the implications below we will
use that (X, D(ω − A))1− p1 ,p = (X, D(A))1− p1 ,p with equivalent norms.
First suppose that |||x|||(0,T ) := kt1−θ AS(·)xkLp ((0,T ), dt ;X) < ∞. In order
t
to show that x ∈ (X, D(A))θ,p with the desired norm estimate, note
kt 7→ t1−θ ASω (t)xkLp (R+ , dt ;X) 6 |||x|||(0,T ) + M kt 7→ t−θ e−t kLp ((T,∞), dt ) kxk
t t
1
6 |||x|||(0,T ) + M T −θ− p kxk.
Similarly, we obtain
L.3 Reiteration 771
kt 7→ t1−θ Sω (t)xkLp (R+ , dt ;X) 6 M kt 7→ t1−θ e−t kLp (R+ , dt ) kxk =: M Lθ,p kxk.
t t
Therefore, from the case I = R+ and the above observation we see that
x ∈ (X, D(A))θ,p and
L.3 Reiteration
Next we will prove the reiteration theorem.
Theorem L.3.1 (Reiteration). Let p ∈ [1, ∞], and fix 0 6 θ0 < θ1 6 1 and
λ ∈ (0, 1). Suppose that Y0 , Y1 are Banach spaces with continuous embeddings
Proof. First let x ∈ (Y0 , Y1 )λ,p . Suppose that x0 ∈ Y0 and x1 ∈ Y1 are such
that x = x0 + x1 . Then
In the converse direction, suppose that x ∈ (X0 , X1 )θ,p and fix ε > 0. Let
u and uε be as in Definition L.2.1 and Proposition L.2.2, and set vε (t) :=
uε (t1/(θ1 −θ0 ) ). By the substitution s = t1/(θ1 −θ0 ) , the estimate
1
= t 7→ t−λ+1/(θ1 −θ0 ) u0ε (t1/(θ1 −θ0 ) ) Lp (R+ , dt
θ1 − θ0 t ;Y0 )
6 B s 7→ (s 1−θ
ku0ε (s)k)1−θ
X0 (s
0 2−θ
ku0ε (s)kX1 )θ0
Lp (R+ , ds
s )
1−θ0 θ0
6 B s 7→ s1−θ u0ε (s) Lp (R+ , ds
s 7→ s2−θ u0ε (s) Lp (R+ , ds
s ;X0 ) s ;X1 )
1−θ0 θ0
6 BCε s 7→ s1−θ u0 (s) Lp (R+ , ds ;X0 ) s 7→ s1−θ u(s) Lp (R+ , ds ;X1 )
s s
n o
1−θ0 θ0
6 BCε max s 7→ s1−θ u0 (s) Lp (R+ , ds ;X0 ) , s 7→ s1−θ u(s) Lp (R+ , ds ;X1 )
s s
2
2(1+2ε)
applying Proposition L.2.2 in the penultimate
R ∞ 0 step, with Cε := ε . By
same substitution and writing uε (s) = s uε (r) dr (see (L.8)), it follows from
Hardy’s inequality (see Lemma L.3.2(2) below) and (L.10) that
θ1 − θ0
6 ks 7→ s1−θ+θ1 u0ε (s)kLp (R+ , ds ;Y1 )
θ1 − θ s
B
6 ks 7→ (s1−θ ku0ε (s)kX0 )1−θ1 (s2−θ ku0ε (s)kX1 )θ1 kLp (R+ , ds )
1−λ s
B
6 ks 7→ s1−θ u0ε (s)kLp (R1 , ds ;X ) ks 7→ s2−θ u0ε (s)kL1p (R , ds ;X )
1−θ θ
1−λ + s 0 + s 1
BCε
6 ks 7→ s1−θ u0 (s)k1−θ 1
Lp (R+ , ds
ks 7→ s1−θ u(s)kθL1p (R , ds ;X )
1−λ s ;X0 ) + s 1
BCε n
1−θ0 θ0
o
6 max s 7→ s1−θ u0 (s) Lp (R+ , ds ;X0 ) , s 7→ s1−θ u(s) Lp (R+ , ds ;X1 )
1−λ s s
2
with Cε := 2(1+2ε)
ε as before.
Combining these estimates and taking the infimum over all admissible
functions u, we obtain the bound
BCε
kxk(Y0 ,Y1 )Tr 6 kxk(X0 ,X1 )Tr .
λ,p 1−λ θ,p
1
Setting ε = 2 gives Cε = 16, and using the estimate from Theorem L.2.3, we
obtain
1 o
n1
kxk(Y0 ,Y1 )λ,p 6 64 max kxk(Y0 ,Y1 )Tr
,
λ 1−λ λ,p
1024B n1 1 o
6 max , kxk(X0 ,X1 )Tr .
1−λ λ 1−λ θ,p
In the above proof we used Hardy’s inequality.
Lemma L.3.2 (Hardy’s inequality). Let p ∈ [1, ∞].
(1) If α > −1 and f : R+ → [0, ∞] is measurable, then
Z s
−α 1 1
s 7→ s f (t) dt p 6 kf kLp (R+ ,t−αp dt
s 0 ds
L (R+ , s ) |α + 1| t )
1 ∞
Z
1
s 7→ s−α f (t) dt 6 kf kLp (R+ ,t−αp dt
s s Lp (R+ , ds ) |α + 1| t )
s
∞
(2): This is proved similarly, this time setting v(s) := 1s s f (t) dt and
R
noting that Z ∞
s−α v(s) = s−α f (θs) dθ.
1
Ww1,p
α
(I; X0 ) ∩ Lpwα (I; X1 ).
L.4 Mixed derivatives and Sobolev embedding 775
Ww1,p
α
(R+ ; X0 ) ∩ Lpwα (R+ ; X1 ) ,→ Cb ([0, ∞); (X0 , X1 )θ,p ) (L.12)
and
1/p 1−1/p
sup tα ku(t)k(X0 ,X1 )1− 1 ,p 6 K1− p1 ku0 kLpw kukLpw . (L.13)
t>0 p α (R+ ;X0 ) α (R+ ;X1 )
The above result is often applied in the setting where we have a continuous
embedding X1 ,→ X0 . In that case, (L.13) shows instantaneous regularisation
in case α > 0.
Remark L.4.2. The estimate (L.11) extends to the case p = 1 and α = 0
(in whichR case we have θ = 0, Xθ,1 = X0 , and w0 ≡ 1). Indeed, writing
∞
u(t) = − t u0 (s) ds and using density of compactly supported functions, we
see that W 1,1 (R; X0 ) ,→ C([0, ∞); X0 ) continuously, and
1
ku(1−j) kLpwα (R+ ;Xj ) = kt 7→ t1−θ− p u(1−j) (t)kLp (R+ ;Xj )
= kt 7→ t1−θ u(1−j) (t)kLp (R+ , dt ;Xj ) .
t
By the argument after Definition L.2.1, this implies that u ∈ W 1.1 (0, T ); X0 +
X1 ) for all T > 0, and therefore u has a version belonging to C([0, ∞; X0 +X1 ).
We will denote this version by u again. By Definition L.2.1, applied to x :=
u(0), we obtain that u(0) ∈ (X0 , X1 )θ,p and, reversing the above identities,
6 Kθ ku0 k1−θ
p kukθLw
p ,
Lw α (R+ ;X0 ) α (R+ ;X1 )
which is (L.11).
Step 3 – In the remainder of the proof we assume that p ∈ (1, ∞). In the
present step we prove the continuous embedding (L.12). Applying (L.15) to
u(· + t) − u(· + s) with 0 6 s 6 t, we obtain
Multiplying on both sides with tα (where now α can be allowed to take any
value in [0, 1/p0 )) and using (L.17), we obtain
1/p 1−1/p
tα ku(t)k(X0 ,X1 )1− 1 ,p 6 K1− p1 tα ku0 (· + t)kLp (R+ ;X0 ) ku(· + t)kLp (R+ ;X1 )
p
1/p 1−1/p
6 K1− p1 ku0 kLpw kukLpw ,
α (R+ ;X0 ) α (R+ ;X1 )
We also need a variant of Theorem L.4.1 in the special case of weighted spaces
of continuous functions. For α ∈ [0, 1), and I = (0, T ] or I = (0, ∞), define
Corollary L.4.3 (Mixed derivatives and traces). Let α ∈ (0, 1), set θ :=
1 − α. Then we have a continuous embedding
1
Cw α
((0, ∞); X0 ) ∩ Cwα ,0 ((0, ∞); X1 ) ,→ Cb ([0, ∞))
1
Proof. Fix an arbitrary Cw α
((0, ∞); X0 ) ∩ Cwα ,0 ((0, ∞); X1 ). It is clear that
u : (0, ∞) → X0 ∩ X1 is continuous, and we have already seen in Theorem
L.4.1 that u : [0, ∞) → (X0 , X1 )θ,p is continuous. It follows that on (0, ∞), u
takes values in Xp,∞ . Also, by (L.11) (applied with p = ∞),
To complete the proof, it thus suffices to show that ku(·+τ )−ukCwα (R+ ;X1 ) →
0 as τ ↓ 0. To this end let ε > 0, and choose δ0 ∈ (0, 1) such that for all
t ∈ (0, 2δ0 ] we have tα ku(t)kX1 < ε. By the continuity of u on (0, ∞) as an
X0 ∩ X1 -valued function and the support condition, u is uniformly continuous
on [δ0 , ∞) as an X0 ∩X1 -valued function, hence also as an X1 -valued function.
778 L The trace method for real interpolation
Therefore, we can find δ1 > 0 such that for all τ ∈ (0, δ1 ) and t > δ0 we have
ku(t + τ ) − u(t)kX1 < ε. Setting δ = min{δ0 , δ1 }, we find that for all τ ∈ (0, δ),
Ww1,p
α
(R+ ; X0 ) ∩ Lpwα (R+ ; X1 ) ,→ Lrwα/h (R+ ; (X0 , X1 )θ,1 ),
and there exists a constant Cp,α,h , depending only on (p, α, h), such that
We conclude that
λ(1−σ)
kukLrw (R+ ;(X0 ,X1 )θ,1 ) 6 C 0 Kσλ ku0 kLpw kuk1−λ+λσ
Lp
α/h α (R+ ;X0 ) wα (R+ ;X1 )
1
1− h
= C 0 Kσ ku0 k1−θ
Lp
kukθLpw ,
w (R+ ;X0 )
α α (R+ ;X1 )
0
where Cα,p = 2 + 3(1 − αp0 )1/p if p > 1 and Cα,p = 5 if p = 1.
p
The assertions (2), (3), and (4) also hold with Lw α
(0, T ; X) and Ww1,p
α
(0, T ; X)
1
replaced by Cwα ((0, T ]; X) and Cwα ((0, T ]; X), respectively.
Often we will need (4) in order to ensure that the bounds are T -independent
(note that the bound in (3) involves th term T −1 ). This is not possible without
the condition u(0) = 0 if p < ∞. Indeed, let u ∈ Ww1,p α
(0, T ; X) be such that
ku(0)k = 1. Then by Sobolev embedding
T ↓ 0.
Proof of Proposition L.4.5. (1), (2), and (3): First let T = 1. Let φ : [1, ∞) →
[0, 1] be given by φ(t) = max{3 − 2t, 0}. Set
u(t)
on (0, 1),
E1 u(t) := φ(t)u(2 − t) on (1, 23 ),
on [ 23 , ∞).
0
k(E1 u)0 kLpwα (R+ ;X) 6 2ku0 kLpwα (0,1;X) + 3kukLpwα (0,1;X) if u ∈ Ww1,p
α
(0, 1; X).
p
This implies the result for T = 1. In the general case let u ∈ Lw α
(0, T ; X) or
1,p
u ∈ Wwα (0, T ; X), and set uT (t) := u(tT ). Define the extension operator by
To prove the final assertion, let ε > 0. We can repeat the proof of (3)
and (4) with a slight deformation of φ, worsening the constant by at most an
additive term ε. It then remains to let ε ↓ 0.
and, if p < ∞,
Ww1,p
α
(0, T ; X0 ) ∩ Lpwα (0, T ; X1 ) ,→ C([0, T ]; (X0 , X1 )θ,p ),
Ww1,p
α
(0, T ; X0 ) ∩ Lpwα (0, T ; X1 ) ⊆ C((0, T ); (X0 , X1 )1− p1 ,p ).
−1
6 Kθ (2 + 3T )kuk1−θ
Ww1,p kukθLpw (0,T ;X1 ) .
α (0,T ;X0 ) α
From this estimate it is clear that if addition u(0) = 0 holds, then the constant
2 + 3T −1 can be replaced by the constant Cα,p of Proposition L.4.5.
The remaining cases are proved in the same way, using in addition Corol-
lary L.4.3 and Theorem L.4.4.
In the same way we obtain the following result from Theorem L.4.4.
Ww1,p
α
(0, T ; X0 ) ∩ Lpwα (0, T ; X1 ) ,→ Lw
r
α/h
(0, T ; (X0 , X1 )θ,1 ),
L.5 Notes
The trace method of Section L.2 is due to J. L. Lions in a classical series
of papers. We follow the presentation in Lunardi [2009] and Triebel [1978],
where a detailed historical account is given. A version of the trace method
with fractional smoothness was recently obtained in Agresti, Lindemulder,
and Veraar [2023]. The reiteration Theorem L.3.1 for real interpolation is due
to Lions and Peetre [1964]. A unified presentation of the reiteration method,
which covers many interpolation methods (including the real and complex
method), can be found in Lindemulder and Lorist [2021].
The mixed derivative result of Theorem L.4.1 is a standard consequence
of the trace method. The end-point case of Corollary L.4.3 is less standard,
but important in evolution equations, and can be found in Lunardi [1995].
The mixed derivative result of Theorem L.4.4 is also a simple consequence
of the trace method, although the simple proof presented here may be new.
Fractional versions have been proved in Agresti and Veraar [2022a].
The presentation of Proposition L.4.5, which is a standard result on ex-
tension operators for bounded intervals (0, T ), follows this reference. Other
constructions can be found in Meyries and Schnaubelt [2012b]. A discussion
on extension operators for more general domains in Rd can be found in the
notes of Chapter 14. In Corollaries L.4.6 and L.4.7 the extension operators are
used to obtain versions of the mixed derivative results on bounded intervals.
In Chapter 18 it is important that the embedding constants can be taken
independent of the size of the interval if one works with functions vanishing
at zero.
References
ume 98 of Monogr. Textbooks Pure Appl. Math., pages 1–19. Dekker, New
York, 1986. 70, 211, 219
Bourgain, J. Vector-valued Hausdorff–Young inequalities and applications. In
Geometric aspects of functional analysis (1986/87), volume 1317 of Lecture
Notes in Math., pages 239–249. Springer-Verlag, Berlin, 1988a. 283, 287,
738, 739
Bourgain, J. Vector-valued Hausdorff–Young inequalities and applications. In
Geometric aspects of functional analysis (1986/87), volume 1317 of Lecture
Notes in Math., pages 239–249. Springer, Berlin, 1988b. 284, 285, 286
Bownik, M. and D. Cruz-Uribe. Extrapolation and factorization of matrix
weights. arXiv:2210.09443, 2022. 83
Brezis, H. and P. Mironescu. Gagliardo–Nirenberg, composition and products
in fractional Sobolev spaces. J. Evol. Equ., 1(4):387–404, 2001. 413
Brezis, H. and P. Mironescu. Gagliardo–Nirenberg inequalities and non-
inequalities: the full story. Ann. Inst. H. Poincaré C Anal. Non Linéaire,
35(5):1355–1376, 2018. 413, 414
Bu, F., T. P. Hytönen, D. Yang, and W. Yuan. Matrix-weighted Besov-type
and Triebel–Lizorkin-type spaces. arXiv:2304.00292, 2023. 223, 417
Bu, S. and J.-M. Kim. Operator-valued Fourier multiplier theorems on Triebel
spaces. Acta Math. Sci. Ser. B (Engl. Ed.), 25(4):599–609, 2005. 413, 679
Buckley, S. M. Estimates for operator norms on weighted spaces and reverse
Jensen inequalities. Trans. Amer. Math. Soc., 340(1):253–272, 1993. 72, 76
Bui, H.-Q. Weighted Besov and Triebel spaces: interpolation by the real
method. Hiroshima Math. J., 12(3):581–605, 1982. 416
Bui, H.-Q., M. Paluszyński, and M. H. Taibleson. A maximal function charac-
terization of weighted Besov–Lipschitz and Triebel–Lizorkin spaces. Studia
Math., 119(3):219–246, 1996. 416
Bui, H.-Q., M. Paluszyński, and M. Taibleson. Characterization of the Besov–
Lipschitz and Triebel–Lizorkin spaces. The case q < 1. In Proceedings of the
conference dedicated to Professor Miguel de Guzmán (El Escorial, 1996),
volume 3, pages 837–846, 1997. 416
Burkholder, D. L. A geometric condition that implies the existence of cer-
tain singular integrals of Banach-space-valued functions. In Conference on
harmonic analysis in honor of Antoni Zygmund, Vol. I, II (Chicago, Ill.,
1981), Wadsworth Math. Ser., pages 270–286. Wadsworth, Belmont, CA,
1983. 83
Butzer, P. L. and H. Berens. Semi-groups of operators and approximation,
volume 145 of Grundlehren der Mathematischen Wissenschaften. Springer-
Verlag New York Inc., New York, 1967. 502
Calderón, A.-P. Commutators of singular integral operators. Proc. Nat. Acad.
Sci. U.S.A., 53:1092–1099, 1965. 211, 216
Calderón, A.-P. Cauchy integrals on Lipschitz curves and related operators.
Proc. Nat. Acad. Sci. U.S.A., 74(4):1324–1327, 1977. 216
Calderón, A.-P. and A. Zygmund. On the existence of certain singular inte-
grals. Acta Math., 88:85–139, 1952. 71, 215, 287
790 REFERENCES
Frey, D., A. McIntosh, and P. Portal. Conical square function estimates and
functional calculi for perturbed Hodge–Dirac operators in LP . J. Anal.
Math., 134(2):399–453, 2018. 512
Friedman, A. Partial differential equations. Holt, Rinehart and Winston, Inc.,
New York-Montreal, Que.-London, 1969. 727
Fröhlich, A. Stokes- und Navier–Stokes-Gleichungen in gewichteten Funk-
tionenräumen. PhD thesis, TU Darmstadt, 2001. 677
Fröhlich, A. The Stokes operator in weighted Lq -spaces. I. Weighted estimates
for the Stokes resolvent problem in a half space. J. Math. Fluid Mech., 5
(2):166–199, 2003. 677
Fröhlich, A. The Stokes operator in weighted Lq -spaces. II. Weighted resolvent
estimates and maximal Lp -regularity. Math. Ann., 339(2):287–316, 2007.
677
Gallarati, C. and M. C. Veraar. Maximal regularity for non-autonomous equa-
tions with measurable dependence on time. Potential Anal., 46(3):527–567,
2017a. 683, 743
Gallarati, C. and M. C. Veraar. Evolution families and maximal regularity
for systems of parabolic equations. Adv. Differential Equations, 22(3-4):
169–190, 2017b. 683, 743
Gallarati, C., E. Lorist, and M. C. Veraar. On the `s -boundedness of a family
of integral operators. Rev. Mat. Iberoam., 32(4):1277–1294, 2016. 743
Ganguly, P. and S. Thangavelu. On the lacunary spherical maximal function
on the Heisenberg group. J. Funct. Anal., 280(3):Paper No. 108832, 32,
2021. 79
Garcı́a-Cuerva, J. and J. L. Rubio de Francia. Weighted norm inequalities and
related topics, volume 116 of North-Holland Mathematics Studies. North-
Holland Publishing Co., Amsterdam, 1985. 71, 73
Garcı́a-Cuerva, J., J. L. Torrea, and K. S. Kazarian. On the Fourier type of
Banach lattices. In Interaction between functional analysis, harmonic anal-
ysis, and probability (Columbia, MO, 1994), volume 175 of Lecture Notes in
Pure and Appl. Math., pages 169–179. Dekker, New York, 1996. 287, 738,
739
Garcı́a-Cuerva, J., K. S. Kazaryan, V. I. Kolyada, and J. L. Torrea. The
Hausdorff–Young inequality with vector-valued coefficients and applica-
tions. Uspekhi Mat. Nauk, 53(3(321)):3–84, 1998. translation in Russian
Math. Surveys 53 (1998), no. 3, 435–513. 411, 739
Garling, D. J. H. Random martingale transform inequalities. In Probability
in Banach spaces 6 (Sandbjerg, 1986), volume 20 of Progr. Probab., pages
101–119. Birkhäuser Boston, Boston, MA, 1990. 284
Garoni, C. and S. Serra-Capizzano. Generalized locally Toeplitz sequences:
theory and applications. Vol. I. Springer, Cham, 2017. 561
Geiss, S. A counterexample concerning the relation between decoupling con-
stants and UMD-constants. Trans. Amer. Math. Soc., 351(4):1355–1375,
1999. 284
REFERENCES 797
1986), volume 14 of Proc. Centre Math. Anal. Austral. Nat. Univ., pages
210–231. Austral. Nat. Univ., Canberra, 1986. 502, 503
McIntosh, A. and A. Yagi. Operators of type ω without a bounded H∞
functional calculus. In Miniconference on operators in analysis (Sydney,
1989), volume 24 of Proc. Centre Math. Anal. Austral. Nat. Univ., pages
159–172. Austral. Nat. Univ., Canberra, 1990. 503, 561
Mei, T. Notes on matrix valued paraproducts. Indiana Univ. Math. J., 55(2):
747–760, 2006. 211, 212
Mei, T. An extrapolation of operator-valued dyadic paraproducts. J. Lond.
Math. Soc. (2), 81(3):650–662, 2010. 211
Mei, T. and J. Parcet. Pseudo-localization of singular integrals and noncom-
mutative Littlewood–Paley inequalities. Int. Math. Res. Not. IMRN, (8):
1433–1487, 2009. 215
Melnikov, M. S. and J. Verdera. A geometric proof of the L2 boundedness of
the Cauchy integral on Lipschitz graphs. Int. Math. Res. Not. IMRN, (7):
325–331, 1995. 218
Meyer, Y. La minimalité de le espace de Besov Ḃ00,1 et la continuité de
opérateurs définis par des intégrales singulières, volume 4 of Monografı́as
de Matemáticas. Univ. Autónoma de Madrid, 1986. 221, 738
Meyries, M. and R. Schnaubelt. Maximal regularity with temporal weights
for parabolic problems with inhomogeneous boundary conditions. Math.
Nachr., 285(8-9):1032–1051, 2012a. 678
Meyries, M. and R. Schnaubelt. Interpolation, embeddings and traces of
anisotropic fractional Sobolev spaces with temporal weights. J. Funct.
Anal., 262(3):1200–1229, 2012b. 675, 782
Meyries, M. and M. C. Veraar. Sharp embedding results for spaces of smooth
functions with power weights. Studia Math., 208(3):257–293, 2012. 413,
414, 416, 417
Meyries, M. and M. C. Veraar. Traces and embeddings of anisotropic function
spaces. Math. Ann., 360(3-4):571–606, 2014a. 413
Meyries, M. and M. C. Veraar. Characterization of a class of embeddings for
function spaces with Muckenhoupt weights. Arch. Math. (Basel), 103(5):
435–449, 2014b. 416, 417
Meyries, M. and M. C. Veraar. Pointwise multiplication on vector-valued
function spaces with power weights. J. Fourier Anal. Appl., 21(1):95–136,
2015. 288, 414, 415, 416, 740
Mielke, A. Über maximale Lp -Regularität für Differentialgleichungen in
Banach- und Hilbert-Räumen. Math. Ann., 277(1):121–133, 1987. 678
Mihlin, S. G. On the multipliers of Fourier integrals. Dokl. Akad. Nauk SSSR
(N.S.), 109:701–703, 1956. 287
Mihlin, S. G. Fourier integrals and multiple singular integrals. Vestn. Leningr.
Univ. (Ser. Mat. Mekh. Astron.), 12(7):143–155, 1957. 287
Monniaux, S. A perturbation result for bounded imaginary powers. Arch.
Math. (Basel), 68(5):407–417, 1997. 502
810 REFERENCES
Neerven, J. M. A. M. van and P. Portal. The Weyl calculus for group gener-
ators satisfying the canonical commutation relations. J. Operator Theory,
83(2):253–298, 2020. 567
Neerven, J. M. A. M. van, M. C. Veraar, and L. W. Weis. Conditions for
stochastic integrability in UMD Banach spaces. In Banach spaces and their
applications in analysis, pages 125–146. Walter de Gruyter, Berlin, 2007.
741
Neerven, J. M. A. M. van, M. C. Veraar, and L. W. Weis. Maximal γ-
regularity. J. Evol. Equ., 15(2):361–402, 2015a. 684, 687
Neerven, J. M. A. M. van, M. C. Veraar, and L. W. Weis. Stochastic integra-
tion in Banach spaces – a survey. In Stochastic analysis: A series of lectures,
volume 68 of Progress in Probability. Birkhäuser Verlag, 2015b. 564
Neerven, J. M. A. M. van, P. Portal, and H. Sharma. Spectral multiplier
theorems for abstract harmonic oscillators on UMD lattices. C. R. Math.
Acad. Sci. Paris, 361:835–846, 2023. 567
Nollau, V. Über den Logarithmus abgeschlossener Operatoren in Banachschen
Räumen. Acta Sci. Math. (Szeged), 30:161–174, 1969. 507
Oberlin, R. Sparse bounds for a prototypical singular Radon transform.
Canad. Math. Bull., 62(2):405–415, 2019. 79
Ogawa, T. and S. Shimizu. End-point maximal L1 -regularity for the Cauchy
problem to a parabolic equation with variable coefficients. Math. Ann., 365
(1-2):661–705, 2016. 679
Ogawa, T. and S. Shimizu. Maximal L1 -regularity of the heat equation and
application to a free boundary problem of the Navier–Stokes equations near
the half-space. J. Elliptic Parabol. Equ., 7(2):509–535, 2021. 679
Ondreját, M. and M. C. Veraar. On temporal regularity of stochastic con-
volutions in 2-smooth Banach spaces. Ann. Inst. Henri Poincaré Probab.
Stat., 56(3):1792–1808, 2020. 679
Orponen, T. On the Haar shift representations of Calderón–Zygmund opera-
tors. Proc. Amer. Math. Soc., 141(8):2693–2698, 2013. 77
Osȩkowski, A. and I. Yaroslavtsev. The Hilbert transform and orthogonal
martingales in Banach spaces. Int. Math. Res. Not. IMRN, (15):11670–
11730, 2021. 291
Ou, Y. Multi-parameter singular integral operators and representation theo-
rem. Rev. Mat. Iberoam., 33(1):325–350, 2017. 79
Parcet, J. Pseudo-localization of singular integrals and noncommutative
Calderón–Zygmund theory. J. Funct. Anal., 256(2):509–593, 2009. 215
Pazy, A. Semigroups of linear operators and applications to partial differential
equations, volume 44 of Applied Mathematical Sciences. Springer-Verlag,
New York, 1983. 727, 742, 762
Peetre, J. Sur les espaces de Besov. C. R. Acad. Sci. Paris Sér. A-B, 264:
A281–A283, 1967. 408
Peetre, J. On spaces of Triebel–Lizorkin type. Ark. Mat., 13:123–130, 1975.
407
812 REFERENCES
Sobczyk, A. Projection of the space (m) on its subspace (c0 ). Bull. Amer.
Math. Soc., 47:938–947, 1941. 681
Sobolevskiı̆, P. E. Coerciveness inequalities for abstract parabolic equations.
Dokl. Akad. Nauk SSSR, 157:52–55, 1964. Translation in Soviet Math. Dokl.
5 (1964), 894–897. 671, 672, 673
Stein, E. M. Interpolation of linear operators. Trans. Amer. Math. Soc., 83:
482–492, 1956. 412
Stein, E. M. Note on singular integrals. Proc. Amer. Math. Soc., 8:250–254,
1957. 673
Stein, E. M. The development of square functions in the work of A. Zygmund.
Bull. Amer. Math. Soc. (N.S.), 7(2):359–376, 1982. 210
Stein, E. M. Harmonic analysis: real-variable methods, orthogonality, and
oscillatory integrals, volume 43 of Princeton Mathematical Series. Princeton
University Press, Princeton, NJ, 1993. 288
Stein, E. M. and G. Weiss. Interpolation of operators with change of measures.
Trans. Amer. Math. Soc., 87:159–172, 1958. 408
Strichartz, R. S. Multipliers on fractional Sobolev spaces. J. Math. Mech.,
16:1031–1060, 1967. 415
Štrkalj, Ž. R-Beschränktheit, Summensätze abgeschlossener Operatoren und
operatorwertige Pseudodifferentialoperatoren. PhD thesis, Universität Karl-
sruhe, 2000. 566, 684
Suárez, J. and L. W. Weis. Interpolation of Banach spaces by the γ-method. In
Methods in Banach space theory, volume 337 of London Math. Soc. Lecture
Note Ser., pages 293–306. Cambridge Univ. Press, Cambridge, 2006. 505
Suárez, J. and L. W. Weis. Addendum to “Interpolation of Banach spaces by
the γ-method”. Extracta Math., 24(3):265–269, 2009. 505
Tanabe, H. Equations of evolution, volume 6 of Monographs and Studies
in Mathematics. Pitman (Advanced Publishing Program), Boston, Mass.,
1979. 682, 684, 727, 742
Tanabe, H. Functional analytic methods for partial differential equations,
volume 204 of Monographs and Textbooks in Pure and Applied Mathematics.
Marcel Dekker Inc., New York, 1997. 684
Taylor, M. E. Pseudo differential operators, volume Vol. 416 of Lecture Notes
in Mathematics. Springer-Verlag, Berlin-New York, 1974. 408
Taylor, M. E. Partial differential equations I. Basic theory, volume 115 of
Applied Mathematical Sciences. Springer, New York, second edition, 2011a.
408
Taylor, M. E. Partial differential equations II. Qualitative studies of linear
equations, volume 116 of Applied Mathematical Sciences. Springer, New
York, second edition, 2011b. 408
Taylor, M. E. Partial differential equations III. Nonlinear equations, volume
117 of Applied Mathematical Sciences. Springer, New York, second edition,
2011c. 408
Titchmarsh, E. C. Introduction to the theory of Fourier integrals. Chelsea
Publishing Co., New York, third edition, 1986. 503
REFERENCES 817
A + B, 515 MD , 6
s
Bp,q (Rd ; X), 320 Mσ,A , 420
∞
Cc (U ; X), 303 M
fσ,A , 517
Cbθ (Rd ; X), 302 fRp , 517
M σ,A
Cbk (Rd ; X), 302 fRp , 517
Cubs
(Rd ; X), 302 M σ,A
Cub (Rd ; X), 302
k ℘ (good set-bound), 160
E(Σσ ), 420 R̈d , 351
s
Fp,q (Rd ; X), 370 %(A), 420
H (Σσ ; A ), 542
∞ RH ∞ (Σσ ; A ), 542
H p (Sϑ ), 485 S 0 (Rd ; X), 300
s,p d S (Rd ; X), 300
RH (R ; X), 301 Sϑ , 485
−, 6
Lpw (I, w; X), 607 σ(A), 420
b pw (I; X), 608
L Σω , 420
Σωbi , 497
L1 (Rd ; X), 308
W k,p (D; X), 300
Lip(Rd ; X), 302
W s,p (Rd ; X), 301
L∞ d
fin (R ; X), 24 1,p
reg 0 ẆA , 576
Mcont,A (I), 614
reg [w]Ap , 57
Mp,A,λ (I), 594 (X0 , X1 )Tr
reg θ,p , 765
Mp,A (I), 577 Xθ,∞ , 764
MRp , 693 ω(A), 420
#
M0,λ , 13 ω bi (A), 497
UMD
and R-boundedness of λ(λ − ∆)−1 ,
680
and maximal Lp -regularity, 657
necessity for multipliers, 277
unconditional, 1-, 667