Operator
Operator
Richard M. Aron
Mohammad Sal Moslehian
Ilya M. Spitkovsky
Hugo J. Woerdeman
Editors
Operator
and Norm
Inequalities and
Related Topics
Trends in Mathematics
Trends in Mathematics is a series devoted to the publication of volumes arising
from conferences and lecture series focusing on a particular topic from any area of
mathematics. Its aim is to make current developments available to the community as
rapidly as possible without compromise to quality and to archive these for reference.
Proposals for volumes can be submitted using the Online Book Project Submission
Form at our website www.birkhauser-science.com.
All contributions should undergo a reviewing process similar to that carried out by
journals and be checked for correct use of language which, as a rule, is English.
Articles without proofs, or which do not contain any significantly new results,
should be rejected. High quality survey papers, however, are welcome.
Mathematics Subject Classification: 46L08, 15A09, 47A30, 47A55, 47B35, 47B38, 47B32, 30H10,
42B35, 44A15, 46A16, 16W80, 46L10, 46L54, 47L55; Primary: 15A42, 15A45, 47A63, 47A64, 46L30,
47A60, 47B49, 47A30, 46B20, 46C15, 52A21, 46C50, 47B38, 47A10, 47A11, 47B48, 46E05, 46E10,
46E15, 46E40, 47E38, 47H30, 46B04, 26D10, 34A40, 35A23, 47B37, 46L05, 42B35, 35B40, 45A07,
45G10; Seconday: 47A64, 47A30, 81P17, 15A45, 47A63, 47B65, 47A56, 47A60, 47B10, 47B15,
47B20, 47B44, 47B47, 05C20, 46B10, 46B28, 46C05, 47L25, 47L30, 47B6, 47A53, 47A55, 06F20,
47H07, 47J05, 46B07, 46B20, 34L10, 46L60, 42B30, 60G42, 42B25, 46E30, 47H06, 47H20
© The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature Switzerland
AG 2022
This work is subject to copyright. All rights are solely and exclusively licensed by the Publisher, whether
the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse
of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and
transmission or information storage and retrieval, electronic adaptation, computer software, or by similar
or dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or
the editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.
This book is published under the imprint Birkhäuser, www.birkhauser-science.com by the registered
company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Preface
v
vi Preface
There are several ways to extend the notion of orthogonality from inner product
spaces to the framework of normed spaces. The most developed one is the Birkhoff–
James orthogonality. It was introduced by Birkhoff [Duke Math. J. 1 (1935), 169–
172] and extensively studied by R.C. James [Duke Math. J. 12 (1945), 291–302]; cf.
C. Alsina, J. Sikorska, M.S. Tomás [Norm derivatives and characterizations of inner
product spaces. World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, 2010].
Chapter “Birkhoff–James Orthogonality: Characterizations, Preservers, and
Orthogonality Graphs” reviews the Birkhoff–James orthogonality starting from
historical perspectives throughout the current development and presents several
characterizations of Birkhoff–James orthogonality in classical Banach spaces, C ∗ -
algebras, and Hilbert C ∗ -modules. In addition, some characterizations of preservers
of Birkhoff–James orthogonality are given.
Chapter “Approximate Birkhoff-James Orthogonality in Normed Linear Spaces
and Related Topics” is an introduction to approximate Birkhoff–James orthogonal-
ity in real normed spaces and its characterizations.
Preface vii
This part mainly studies inequalities concerning closed range, normal, and Toeplitz
operators (cf. A. Böttcher and B. Silbermann [Analysis of Toeplitz operators.
Second edition. Prepared jointly with Alexei Karlovich. Springer Monographs in
Mathematics. Springer-Verlag, Berlin, 2006] and I. Gohberg, S. Goldberg, and M.
A. Kaashoek [Basic classes of linear operators. Birkhäuser Verlag, Basel, 2003]).
In Chapter “Normal Operators and Their Generalizations”, some aspects of local
spectral theory and Fredholm theory of certain classes of operators that generalize
normal operators on Hilbert spaces are studied.
Chapter “On Wold Type Decomposition for Closed Range Operators” surveys
Wold-type decomposition for closed range operators satisfying certain operator
inequalities. Several results on left invertible operators close to isometries are listed
and extended to the case of regular operators.
Chapter “(Asymmetric) Dual Truncated Toeplitz Operators” considers properties
of asymmetric dual truncated Toeplitz operators acting between the orthogonal
complements of two model spaces.
Chapter “Boundedness of Toeplitz Operators in Bergman-Type Spaces” is
devoted to the open problem of characterization of the bounded Toeplitz operators
Ta in Bergman spaces. Based on the structure of the Bergman spaces, a characteri-
zation of the boundedness and compactness is presented in the case of operators in
spaces with weighted sup-norms.
This part deals with miscellaneous inequalities concerning topological and geomet-
rical properties of various Banach spaces and operator algebras (cf. W.B. Johnson
and J. Lindenstrauss (ed.) [Handbook of the geometry of Banach spaces. Vol. I.
North-Holland Publishing Co., Amsterdam, 2001]).
In Chapter “Disjointness Preservers and Banach-Stone Theorems”, the so-called
weak and strong Banach-Stone theorems are given. In addition, it is proved
that in many cases lattice isomorphisms (Kaplansky’s Theorem), ring isomor-
phisms (Gelfand-Kolmogorov Theorem), multiplicative isomorphisms (Milgram’s
Theorem), isometries (Banach-Stone Theorem), and nonvanishing preservers are
⊥-isomorphisms.
viii Preface
xi
xii Contents
Abstract Several inequalities have been established in the context of Hilbert spaces
operators or operator algebras. Our discussion will be limited to matrices. Important
inequalities in mathematics and other sciences, such as Golden-Thompson inequal-
ity or von Neumann trace inequality, and extensions, are revisited. Our main goal is
to emphasize the link between majorization theory and other relevant inequalities.
Notation
N Set of natural numbers
N0 Set of nonnegative integer numbers
R Set of real numbers
R+0 Set of nonnegative real numbers
C Set of complex numbers
Rn Vector space of real n-tuples
Cn Vector space of complex n-tuples
· Euclidean norm; spectral norm or operator norm
N. Bebiano ()
CMUC, Departamento de Matemática, Faculdade de Ciências e Tecnologia da, Universidade de
Coimbra, Coimbra, Portugal
e-mail: [email protected]
R. Lemos
CIDMA, Department of Mathematics, University of Aveiro, Aveiro, Portugal
e-mail: [email protected]
G. Soares
CMAT-UTAD, Universidade de Trás-os-Montes e Alto Douro, Vila Real, Portugal
e-mail: [email protected]
1 Introduction
k
k
xi ≤ yi , k = 1, . . . , n, (1)
i=1 i=1
and equality occurs in (1) for k = n. Further, if (1) holds, then x is said to be weakly
majorized or submajorized by y and the notation x ≺w y is used. We remark that
x ≺ y is equivalent to
n
n
xi ≥ yn , k = 1, . . . , n, (2)
i=k i=k
k
k
xi ≤ yi , k = 1, . . . , n, (3)
i=1 i=1
and equality occurs in (3) for k = n. If x, y > 0, i.e., all the components of x, y are
positive, this is clearly equivalent to
n
n
xi ≥ yn , k = 1, . . . , n, (4)
i=k i=k
this justifying the log-majorization terminology. When equality between the prod-
ucts of all the components of x and y is not required, the following parallel notations
are used:
Proposition 1.4 Let x, y ∈ Rn have all the components positive and f be a non-
decreasing continuous function on an interval containing all the components of x, y,
such that f (et ) is convex. Then
In particular, f (t) = t in the previous proposition shows that the weak log-
majorization ≺wlog is stronger than the weak majorization ≺w .
2 Matrix Majorization
is the spectral radius of A. Further, considering the Euclidean norm x of a vector
x ∈ Cn , let
λ1 (A) ≥ · · · ≥ λn (A).
s1 (A) ≥ · · · ≥ sn (A).
A norm ||| · ||| in Mn (C) is said to be unitarily invariant if |||U AV ||| = |||A||| for any
A, U, V ∈ Mn (C) with U, V unitary. Examples of unitarily invariant norms are the
Schatten p-norms given by
n 1
p
p
1
Ap = si (A) = tr |A|p p , p ≥ 1,
i=1
8 N. Bebiano et al.
k
A(k) = si (A), k = 1, . . . , n,
i=1
also called Frobenius norm, Hilbert-Schmidt norm or Schur norm, is the norm
induced by to the Frobenius or Hilbert-Schmidt inner product in Mn (C):
The notion of majorization gives a mean for comparing two probability distri-
butions or two density matrices, that is positive semidefinite matrices of trace one,
using the eigenvalues, in an elegant way. It arises in fields like computer science,
economics or quantum mechanics.
Important sources on majorization for eigenvalues and singular values of matri-
ces are [21, 46, 47, 55, 72] and two survey articles of T. Ando [2, 3].
For simplicity, if A, B ∈ Mn (C) have real eigenvalues, then λ(A) ≺ λ(B) and
λ(A) ≺w λ(B) are abreviated to A ≺ B and A ≺w B, respectively.
The main diagonal entries and the eigenvalues of a Hermitian matrix are related
through majorization. This classical result due to I. Schur [84] can be briefly stated
as follows.
Theorem 2.1 (Schur Majorization Theorem, 1923) If A ∈ Mn (C) is Hermitian,
then In ◦ A ≺ A.
In 1954, A. Horn [51] proved the converse, giving rise to the next fundamental
result, which received considerable attention and led to generalizations in several
directions.
Theorem 2.2 (Schur-Horn Theorem) Let x, y ∈ Rn . There exists a Hermitian
matrix with prescribed diagonal entries and prescribed eigenvalues arranged,
respectively, in x and y if and only if x ≺ y.
After this, Horn’s subsequent work on the eigenvalues of sums of Hermitian
matrices culminated in the inequalities conjectured in [53]. The solution to Horn’s
conjecture appeared in two papers, one by A. Klyachko [60] and the other one by
A. Knutson and T. Tao [61].
Another relevant result in matrix majorization is due to Ky Fan [28].
Theorem 2.3 (Ky Fan Dominance Theorem, 1951) Let A, B ∈ Mn (C). Then the
following are equivalent statements:
i. |A| ≺w |B|;
Log-majorization Type Inequalities 9
ii. |||A||| ≤ |||B||| for any unitarily invariant norm ||| · ||| in Mn (C).
If A, B ∈ Mn (C) have nonnegative eigenvalues, A ≺log B stands for
Abbreviated notations for the weaker versions, involving either ≺wlog or ≺wlog, are
analogously used. Clearly, if A, B have positive eigenvalues, then
On the other hand, some classical determinantal inequalities can find their majoriza-
tion counterparts.
As usual, A > 0 means that A is a positive definite matrix and A ≥ B means that
A − B is a positive semidefinite matrix. Real-valued continuous functions f defined
on a real interval , such that
for all Hermitian A, B ∈ Mn (C) with spectra in and all n ∈ N, are said to
be operator monotone on . A useful and fundamental tool for treating operator
inequalities is Löwner-Heinz inequality. Löwner’s original proof [69] used an
integral representation for operator monotone functions and an alternative proof was
given by Heinz [44]. It states that
A≥B≥0 ⇒ Aα ≥ B α , (6)
that is, f (t) = t α is operator monotone on R+ 0 , for α ∈ [0, 1]. In general, (6) is not
true for α > 1.
For k = 1, . . . , n and nk = nk , the kth compound or kth antisymmetric tensor
power of A ∈ Mn (C) is the matrix A∧k ∈ Mnk (C) with entries given by the
minors det A(i, j), where the index sets i, j ⊂ {1, . . . , n} have cardinality k and
are lexicographically ordered. As usual, A(i, j) denotes the submatrix of A that lies
in rows and columns indexed, respectively, by i, j. Some essential properties of these
matrices [21] are listed below:
∧k ∧k B ∧k (Binet-Cauchy formula);
P1. ∧k ∗ = A ∗ ∧k
(AB)
P2. A = (A ) ;
∧k r
P3. A = (Ar )∧k , r > 0;
−1
P4. A∧k = (A−1 )∧k if A is invertible;
10 N. Bebiano et al.
k
P5. λi A∧k = λij (A), where i = (i1 , . . . , ik ) and 1 ≤ i1 < · · · < ik ≤ n;
j =1
k
P6. A∧k = s1 A∧k = sj (A), k = 1, . . . , n.
j =1
n
1 1
λi (A) = | det(A)| = det(A) det(A) 2 = det(A∗ A) 2 = det |A|
i=1
The von Neumann’s trace inequality was first published in 1937 by von Neumann
[96] with a complicated proof. Other proofs were given in 1959 and subsequently
in 1975, based on doubly stochastic matrices, by Mirsky [76, 77]. However, these
proofs only work in the finite dimensional case. A simple proof, which also extends
to the infinite dimensional setting, was finally obtained in 1991 by R. D. Grigorieff
[41].
Theorem 3.1 (von Neumann’s Inequality, 1937) Let A, B ∈ Mn (C). Then
n
|tr(AB)| ≤ si (A)si (B)
i=1
Apply the antisymmetric tensor power technique to the previous inequality, that is,
replace A, B by their kth compounds, k = 1, . . . , n, and use P6. Equality for k = n
is immediate by properties of the determinants.
Remark 3.3 For A, B ∈ Mn (C), Horn and Weyl’s log-majorizations stated before,
the second applied to the product AB, imply the corresponding weak majorizations,
so we easily find for k = 1, . . . , n that
k
k
k
k
λi (AB) ≤ |λi (AB)| ≤ si (AB) ≤ si (A)si (B). (8)
i=1 i=1 i=1 i=1
Mirsky [76] and Theobald [88]. Ruhe [82] reobtained it under the more restrictive
assumption of positive semidefiniteness of both matrices.
Theorem 3.4 If A, B ∈ Mn (C) are Hermitian, then
n
n
λi (A)λn−i+1 (B) ≤ tr(AB) ≤ λi (A)λi (B). (9)
i=1 i=1
n
n
n
xi yn−i+1 ≤ xi yσ (i) ≤ xi yi
i=1 i=1 i=1
n
n
min αj + βσ (j ) ≤ det(A + B) ≤ max αj + βσ (j ) .
σ ∈Sn σ ∈Sn
j =1 j =1
det(A + B) = det(A0 + U ∗ B0 U ),
where A0 , B0 are the diagonal forms of A, B. Since the unitary group is compact and
the determinant is continuous, det(A0 +V ∗ B0 V ) attains its maximum and minimum
values for some unitary matrix V ∈ Mn (C). Take
U = ei S
= In + i S + O( 2 ),
det(A0 + U ∗ V ∗ B0 V U )
to the first order in , it can be easily shown that V ∗ B0 V commutes with the inverse
of A0 + V ∗ B0 V . Thus V ∗ B0 V commutes with A0 and the theorem follows. If
A0 + V ∗ B0 V is singular, then the result follows by a limiting argument.
Remark 3.8 A natural generalization of Fiedler’s Theorem would be the following.
If A, B ∈ Mn (C) are normal matrices with eigenvalues α1 , . . . , αn , β1 , . . . , βn ,
respectively, then det(A + B) lies in the convex hull of the products
n
αj + βσ (j ) , σ ∈ Sn .
j =1
k
k
sij (A) sn−ij +1 (B) ≤ sj (AB), k = 1, . . . , n,
j =1 j =1
equivalently,
k
k
sij (AB) ≤ sj (A) sij (B), k = 1, . . . , n,
j =1 j =1
The next result [9, 21] has a simple proof, using majorization theory.
Theorem 3.10 If A, B ≥ 0 have eigenvalues a1 ≥ · · · ≥ an and b1 ≥ · · · ≥ bn ,
respectively, then
n
n
aj2 + bj2 ≤ | det(A + iB)|2 ≤ aj2 + bn−j
2
+1 .
j =1 j =1
n
n
= aj2 1 + λ2j (A−1 B)
j =1 j =1
and
1 1
λj (A−1 B) = sj2 A− 2 B 2 , j = 1, . . . , n.
k
1 1
k
1 1
k
1 1
sn−j +1 (A− 2 )sj (B 2 ) ≤ sj (A− 2 B 2 ) ≤ sj (A− 2 )sj (B 2 )
j =1 j =1 j =1
−2 2 1 1 bj 1 1 bj
2
sn−j +1 (A )sj (B 2 ) = , sj2 (A− 2 )sj2 (B 2 ) = , j = 1, . . . , n.
aj an−j +1
b1 bn b1 bn
,..., ≺log λ(A−1 B) ≺log ,..., .
a1 an an a1
Thus,
n bj2
n
n 2
bn−j +1
1+ ≤ 1 + λ2j (A−1 B) ≤ 1+
j =1
aj2 j =1 j =1
aj2
A + A∗ A − A∗
Re A = and Im A =
2 2i
are Hermitian matrices, the next corollary is easy to derive.
Corollary 3.12 If A ∈ Mn (C) is such that Re A > 0 then
eA+B = eA eB .
tr eH +K+L ≤ tr eH eK eL
obviously holds, but this is not true in general as the next counter-example, due to
C. J. Thompson [90], shows.
Example Consider the Pauli matrices
01 0 −i 1 0
σ1 = , σ2 = , σ3 = ,
10 i 0 0 −1
sinh a
eA = cosh a I2 + I2 .
a
H = σ1 , K = σ2 , L = (σ3 − σ2 − σ1 ).
Log-majorization Type Inequalities 17
In this case,
tr eH +K+L = 2 cosh
tr eH eK eL < tr eH +K+L .
for some unitary matrices U, V . This result has application in quantum computing.
We observe that Thompson’s result was obtained before the long-standing Horn’s
conjecture on eigenvalues of sums of Hermitian matrices has been solved (see [20]
for more details).
Several trace inequalities may be strengthened in the set up of majorization. This
is the case of the Golden-Thompson inequality. In fact, it was proved by Lenard [66]
and Thompson [90] that
eH +K ≺w e 2 eK e 2
H H
for A, B ≥ 0, firstly used to derive inequalities for the moments of the eigenvalues
of the Schrödinger Hamiltonian [68].
Theorem 4.2 (Araki’s Log-majorization, 1990) Let A, B ≥ 0. Then
1 1 r r
(A 2 BA 2 )r ≺log A 2 B r A 2 , r ≥ 1, (11)
18 N. Bebiano et al.
or equivalently
1
q q 1 p p p
(A 2 B q A 2 ) q ≺log A 2 B p A 2 , 0 < q ≤ p. (12)
1 1 r r
Proof It is clear that (A 2 BA 2 )r and A 2 B r A 2 have the same determinant. Assum-
ing A invertible, let us prove that
1 1 r r
λ1 ((A 2 BA 2 )r ) ≤ λ1 (A 2 B r A 2 ). (13)
r r 1 1
To do so we may prove that A 2 B r A 2 ≤ In implies that A 2 BA 2 ≤ In , because
both sides of (13) have the same order of homogeneity for A and B, so that we can
multiply A, B by a positive constant. Since B r ≤ A−r , for r ≥ 1, then Löwner-
Heinz inequality implies B ≤ A−1 . If A is not invertible, by a continuity argument,
we obtain (13). By properties P1 and P3, then
∧k 1 ∧k 1 1 ∧k
(A ) 2 (B )(A∧k 2 )r = (A 2 BA 2 )r .
1
eH +K ≺log eH eK , (15)
H H
since e 2 eK e 2 and eH eK have the same eigenvalues. From the previous results, we
can see that Golden-Thompson inequality is strengthened to
pH pK pH 1
eH +K ≤ e 2 e e 2 p , p > 0,
for any unitarily invariant norm, and the right hand side decreases to the left hand
side as p converges to 0.
5 Ando-Hiai Inequality
−1
A∇α B = (1 − α)A + αB and A!α B = (1 − α)A−1 + αB −1
are the weighted arithmetic and harmonic means, respectively; Awl B = A and
Awr B = B are the left and right trivial operator means, respectively.
Kubo and Ando proved that there is a one-to-one correspondence between
operator connections and operator monotone functions on R+ 0.
Theorem 5.1 ([62]) For each operator connection σ , there exists a unique ope-
rator monotone function f : R+ +
0 → R0 , satisfying
holds, with the right hand side defined via functional calculus, and extended to
A, B ≥ 0 as follows
A σ B = lim (A + In ) σ (B + In ).
→0+
Let α ∈ [0, 1]. In particular, associated with the operator monotone function
f (t) = t α , the α-weighted geometric mean is
1 1 1 α 1
A α B = A 2 A − 2 B A − 2 A 2 .
It is easy to see that Aα B = B1−α A and when A commutes with B, then
Aα B = A1−α B α . The geometric mean, simply denoted by , is the mean
1
corresponding to f (t) = t 2 . It is the unique positive semidefinite solution of the
Riccati equation XA−1 X = B, also characterized [80] as
AX
A B = max X ∈ Hn : ≥0 . (16)
XB
1 1
Further, there is a unitary matrix U such that AB = A 2 U B 2 .
Ando and Hiai [5] proved the following interesting result, concerning the
weighted geometric mean.
Theorem 5.2 (Ando-Hiai Inequality, 1994) For A, B ≥ 0 and α ∈ [0, 1],
or, equivalently,
p 1 1
A α B p p ≺log Aq α B q q , 0 < q ≤ p. (18)
Aα B ≤ In . (19)
1 1
Let C = A− 2 BA− 2 . By continuity, we may assume that A, B are invertible.
It follows from (19) that A ≤ C −α . By Löwner-Heinz inequality, we have
Log-majorization Type Inequalities 21
A ≤ C −α . In this case,
1 1
Ar α B r = A 2 A α (CAC) 1− C A 2
1 1
≤ A 2 C −α α (C 2−α 1− C A 2
1 1
= A 2 C αA 2 = Aα B,
by the joint monotonicity of the weighted geometric means α and 1− . Therefore,
Ar α B r ≤ In . We have just proved that λ1 (Aα B) ≤ 1 implies λ1 (Ar α B r ) ≤ 1.
Thus,
and applying the antisymmetric tensor power trick, having also in mind that
det Ar α B r = det(Aα B)r ,
This proves that (17) also holds for r > 2. Now, for 0 < q ≤ p, the result easily
follows.
The following corollary complements the previous log-majorizations of Golden-
Thompson type [5].
Corollary 5.3 If H, K are Hermitian matrices and α ∈ [0, 1], then
1
epH α epK p ≺log e(1−α)H +αK , p > 0.
that is, the Lie-Trotter like formula for the weighted geometric mean obtained in
[45, Lemma 3.3].
The corresponding inequality for unitarily invariant norms holds, with the left
hand side norm decreasing to the right hand side as p converges to 0.
A celebrated development of Löwner-Heinz inequality established by T. Furuta
[35] is the next order preserving operator inequality, which was motivated by a
previous conjecture by Chan and Kwong [25].
22 N. Bebiano et al.
Furuta and many other researchers refined and generalized (20) and applied these
results to produce new inequalities [36].
The essential part of Furuta inequality is the case q = p+r1+r , which can be
formulated for invertible A, using the weighted geometric mean, as follows:
Fujii and Kamei [33] showed that Ando-Hiai inequality is equivalent to Furuta
inequality. Next, we show this direct implication. Indeed, let A ≥ B > 0 and p ≥ 1.
Firstly, if 0 ≤ r ≤ 1, then Ar ≥ B r by Löwner-Heinz inequality. Consequently,
A−r p+r
r B ≤ B
p −r r
p+r B p = In .
A−r p+r
r B ≤ In .
p
that is, the essential part of Furuta inequality (21) holds. The remaining part follows
readily from Löwner-Heinz inequality.
Extensions of Furuta inequality and Ando-Hiai log-majorization were given by
Furuta [37] and afterwards by other authors. Nowadays, the multivariate geometric
mean as settled in [22, 78], following a Riemannian geometric approach, is often
called the Karcher mean [63]. It is also called Cartan mean and Riemannian mean.
Log-majorization Type Inequalities 23
Using the previous techniques of Ando and Hiai, Bebiano, Lemos and Providência
[12, Theorem 2.1] obtained the next log-majorization of Araki’s type.
Theorem 6.1 (BLP Inequality, 2005) For A, B ≥ 0,
1+q 1+q 1 r r q 1
A 2 Bq A 2 ≺log A 2 A 2 B r A 2 r A 2 , 0 < q ≤ r. (22)
that is,
r r q
A−1 ≥ A 2 B r A 2 r ≥ 0.
− r/q+r r r r q r r q
A−(1+q) = A r/q ≥ A− 2 A 2 B r A 2 r q A− 2 r = B q ,
that is,
1+q 1+q
A 2 Bq A 2 ≤ In
and then (23) holds. Using Lemma 2.4, the result follows from (23) replacing A, B
by A∧k , B ∧k , respectively, for k = 1, . . . , n, by properties P1, P3, P5.
For convenience of notation, for α ∈ R, let
B = A 2 (A− 2 BA− 2 )α A 2
1 1 1 1
A α
A1−α B α ≺log A α B.
A αB ≺log A1−α B α .
and the left hand side converges to the right hand side as p converges to 0.
Proof The log-majorization (25) implies the trace inequality
p p q
tr (As Aq B q ) ≤ tr As (A 2 B pA 2 ) p , 0 ≤ q ≤ p, s ≥ 0,
occuring trace equality when q = 0. Taking the derivatives of the right and left hand
sides of the previous inequality at q = 0, observing that
d q q
A B q=0
= log A + log B, (27)
dq
d p p p pq p p
1
p
(A 2 B A 2 ) q=0
= log A 2 B p A 2 , p > 0, (28)
dq
yields a trace inequality. Multiplying both hand sides of the obtained trace inequality
by q provides (26). By the parametric Lie-Trotter formula, we may see that (28)
converges to (27) as p converges to 0.
The case q = s = 1 in Theorem 6.5 is due to Hiai and Petz [45]. It was later
complemented in [5]. Using relative entropy terminology, Theorem 6.5 for q = s,
replacing B by B −1 , may be written in the condensed form
s
S(As , B s ) ≤ − tr Ŝ(Ap |B p )As−p , s ≥ 0, p > 0,
p
this providing an upper bound for the relative entropy S(A, B) when s = 1.
Fujii, Nakamoto and Tominaga [34] improved BLP inequality as follows.
Theorem 6.6 If A, B ≥ 0, p ≥ 1, q ≥ 0, then
1+q 1+q 1+q p+q 1 q q1 1
A 2 B A 2 p(1+q) ≤ A 2 A 2 B q+p A 2 p A 2 .
Proof If α = 0 or α = 1, the result is trivial. Let 0 < α < 1. Clearly, Aα B and
A1−α B α have the same determinant. Let us prove that
A−(1−α) ≥ B α .
−(1−α) 1+αr r r α
A ≥ A−(1−α) 2 B A−(1−α) 2 .
Taking r = 1
1−α yields
1 1 α
A−1 ≥ A− 2 BA− 2 ,
−1
A σ ⊥ B = B −1 σ A−1
1 −1 1 1 1
A 2 fσ (M) fρ (M) fσ (M) A 2 ≤ A 2 fτ (M) A 2 = A τ B, (32)
1 1
where M = A− 2 BA− 2 . Therefore
λ1 (A σ B)(A ρ B)−1 (A σ B)(A τ B)−1 ≤ 1. (33)
Moreover,
1
s1 (A τ ⊥ B) 2 D (A σ B)∼ D (A ρ ⊥ B) 2
1
28 N. Bebiano et al.
Now, it is clear that (34) is not greater than (33) and the implication (31) holds. If
A, B ≥ 0, by a continuity argument, the result is obtained.
If the nonzero operator connections σ, τ, ρ satisfy fσ2 ≥ fτ fρ , then (30) holds
with each connection replaced by its dual [65].
Applying Theorem 7.1 for A, B ≥ 0, C ∈ Hn∼ and σ ≤ τ = ρ yields
λ1 (A τ ⊥ B) C ∗ (A σ B)∼ C ≤ λ1 (A C ∗ B ∼ C). (35)
≺log (A σ ⊥ B) C ∗ (A σ B)∼ C.
The last inequality in (36) is the case τ, ρ as the trivial operator means wl , wr and
σ = in Theorem 7.1. The first inequality in (36) follows after taking square roots
of the obtained eigenvalues from the case τ = σ = with ∼ deleted in (35),
then replacing C by C ∗ (A B)∼ C. Applying Weyl’s trick to (36) and observing the
equality of the determinants of the matrices involved, a log-majorization is obtained.
Next, replace A by A σ ⊥ B, B by A σ B in that log-majorization and use the identity
(A σ ⊥ B) (A σ B) = A B.
Corollary 7.3 If A, B ≥ 0, C ∈ Hn∼ and 0 ≤ α ≤ β ≤ 1, then
1 1
(A1−α B) 2 C ∗ (A β B)∼ C (A 1+α−β B) 2 ≺log A C ∗ B ∼ C. (37)
2
1 1
s1 (A 1−α B) 2 C ∗ (A β B)∼ C (A 1+α−β B) 2 ≤ λ1 (A C ∗ B ∼ C).
2
Replace A, B, C by their kth compounds and apply properties P1–P6. The determi-
nants of the matrices in both hand sides of (37) are equal.
Log-majorization Type Inequalities 29
Remark 7.4 Let A, B ≥ 0, C ∈ Hn∼ and α ∈ [0, 1]. If σ = α in Corollary 7.2 and
β = 1 in Corollary 7.3, then
(A1−α B) 2 C ∗ (A B)∼ C (A α B) 2
1 1
(A σ ⊥ B) C ∗ (A σ B)∼ C ≺log A C ∗ B ∼ C
The first inequality in (39) follows from (36) when the symbol ∼ is absent and
C = (AB)r . Concerning the second, if A > 0 and λ1 (AB)r+1 ≤ 1, then B ≤ A−1
implies
Thus, the last inequality in (39) is true. By continuity, it remains valid for A ≥ 0.
After applying Weyl’s trick to (39), the obtained log-majorization implies the log-
majorization between the corresponding square roots.
30 N. Bebiano et al.
k
1 1
k r+1
si A 2 (A B)r B 2 ≤ si 2 (AB), k = 1, . . . , n,
i=1 i=1
In this section, Ando and Visick’s inequalities [4, 94] for the Hadamard product
of positive definite matrices, which settled affirmatively Bapat and Johnson’s
conjecture [56], are revisited and weighted interpolations are presented.
We recall that a map : Mm (C) → Mn (C) is called positive if A ≥ 0 implies
(A) ≥ 0 and it is called unital if (Im ) = In .
Lemma 8.1 ([1]) If : Mm (C) → Mn (C) is a unital positive linear map and f is
operator monotone on R+
0 , then
f ( (A)) ≥ (f (A)), A ≥ 0.
The proof presented below of Ando and Visick’s inequalities follows Ando’s
arguments [4]. We state these results in the following condensed form, where ∼ is
either omitted or acts as the transpose.
Theorem 8.2 If A, B > 0, then A ◦ B ≺wlog AB ∼ , that is,
n
n
λi (A ◦ B) ≥ λi (AB ∼ ), k = 1, . . . , n. (40)
i=k i=k
Proof There exits a unital positive linear map such that (X ⊗ Y ) = X ◦ Y for
all X, Y ∈ Mn (C). For A, B > 0,
n
n
n
λi (log(A ◦ B)) ≥ λi (In ◦ (H + K)) ≥ λi (H + K)
i=k i=k i=k
n
n
n
n
λi (A ◦ B) ≥ eλi (In ◦(H +K)) ≥ eλi (H +K) = λi eH +K
i=k i=k i=k i=k
n
n
≥ λi (eH eK ) = λi (AB), k = 1, . . . , n.
i=k i=k
(see also [95, Theorem 1]). According to the previous proof, we can write
1 1
A ◦ B ≺wlog lim Ap ◦ B p p ≺wlog lim Ap (B ∼ )p p ≺log AB ∼ .
p→0 p→0
32 N. Bebiano et al.
n
n
−r
λi (A ◦ B)−r ≤ λi AB ∼ , k = 1, . . . , n,
i=k i=k
and deduced Theorem 8.2 from it. In fact, this is equivalent to Theorem 8.2 as shown
by Bebiano and Perdigão [10]. One of the implications is a trivial consequence of
the following limit:
λ−r − 1
lim = −log λ, λ > 0.
r→0 r
Considering the function f (t) = log(1 + ert ), which is convex and increasing for
t > 0, with , r > 0, by Proposition 1.3 ii, we obtain
n
n
−r
1 + λi (A ◦ B) ≤ 1 + λi (AB ∼ )−r , k = 1, . . . , n.
i=k i=k
Theorem 8.4 Let A, B > 0 and D ∈ Mn (C) be a diagonal matrix, assumed real
when ∼ is absent. If α ∈ [0, 1], then
n
n
λi (A o B) |D|2 ≥ λi (AB) D (AB)∼ D (41)
i=k i=k
n
≥ λi (A1−α B) D (Aα B)∼ D
i=k
n
≥ λi A D B ∼ D , k = 1, . . . , n,
i=k
Further, if C = D ∈ Hn∼ in Corollary 7.2 with σ = α , α ∈ [0, 1], and in (38), the
result is obtained.
If ∼ is deleted and D = In , then (41) was previously given by Ando [4,
Theorem 2] and, in this case, the remaining inequalities were obtained by Hiai and
Lin [48]. The complete version in Theorem 8.4 is derived in [65]. The inequalities
in (41) hold for A, B > 0, k = 1, . . . , n and D diagonal, but
n
n
λi (A o B) |D|2 ≥ λi (AB) D ∗ (AB)∼ D (42)
i=k i=k
does not remain true, in general, when D is replaced by any Hermitian matrix.
Example Consider
10 21 1 1+i
A= , B= , D= .
01 11 1 − i −3
1
In this case, AB = B 2 and (42) with ∼ absent does not hold, because
1 2
λ2 ((A ◦ B) D 2 ) ≈ 3.783 ≤ λ2 B 2 D ≈ 4.095.
9 Indefinite Inequalities
n
per A = aj σ (j ) , σ ∈ Sn .
σ ∈Sn j =1
Although permanents and determinants have similar definitions and share some
common properties, they exhibit substancial differences, such as the non-
multiplicativity of the permanent.
In 1926, van der Waerden raised a question [93] and motivated a conjecture:
the minimum of the permanent of a n-square doubly stochastic matrix is nn!n and
34 N. Bebiano et al.
equality occurs when every entry of the matrix equals n1 . This conjecture attracted
the attention of mathematicians all over the world, although it remained open for
more than fifty years. The proof of this famous conjecture by G. P. Egoritjev [26] in
1981, also proved by Falikman [27], is based on an inequality for permanents, which
is a special case of a result of A. D. Alexandroff on positive definite quadratic forms.
In what follows we write per A = per(a1 , . . . , an ) with ai the ith column of A.
Theorem 9.1 (Alexandroff Permanental Inequality) For a1 , . . . , an ∈ Rn ,
the space Rn is no longer Euclidean but Lorentzian, accordingly as the length of the
vector (x1 , . . . , xn ) is
That is, we are dealing now with a so called indefinite inner product space.
In this section, we present miscellaneous indefinite inequalities obtained in this
set up. First, we introduce some definitions and notations.
Let J be a Hermitian involutive matrix, that is, J ∗ = J and J 2 = In . Consider
C endowed with the indefinite inner product induced by J , given by [x, y] = y ∗ J x
n
In ≥J Aα ≥J B α .
The case α = 12 in Theorem 9.2 is due to Ando [6, Theorem 6], being the cases
α = 0 and α = 1 trivially satisfied. Motivated by these results, the Furuta inequality
of indefinite type in (43) and (44) was established by Sano [83, Theorem 3.4] and
Bebiano et al. [14, Theorem 2.1], respectively.
Theorem 9.3 (Furuta Inequality of Indefinite Type) Let A, B ∈ Mn (C) be J -
Hermitian with nonnegative spectra, μIn ≥J A ≥J B (or A ≥J B ≥J μIn ) for
some μ > 0. Then for each r ≥ 0,
p+r 1
r r q
A q ≥J A 2 B p A 2 (43)
and
1 p+r
r r q
B 2 Ap B 2 ≥J B q (44)
For an arbitrary J -Hermitian matrix A ∈ Mn (C), we denote by σJ± (A) the set
of eigenvalues of A with eigenvectors x, such that x ∗ J x = ±1. We say that A is
J -unitarily diagonalizable if every eigenvalue of A belongs to either σJ+ (A) or to
σJ− (A). In this case, σJ+ (A) and σJ− (A) have r and n − r eigenvalues, respectively.
Consider a J -Hermitian matrix A, whose eigenvalues α1 ≥ · · · ≥ αr belong to
σJ+ (A) and αr+1 ≥ · · · ≥ αn belong to σJ− (A). In this case, the eigenvalues of A
are said to not interlace if either αr > αr+1 or αn > α1 , otherwise, they are said to
interlace.
Theorem 9.4 Let J = Ir ⊕ −In−r , 0 < r < n, and A, C ∈ Mn (C) be
non-scalar J -Hermitian and J -unitarily diagonalizable matrices with eigenvalues
α1 ≥ · · · ≥ αr (c1 ≥ · · · ≥ cr ) in σJ+ (A) (σJ+ (C)) and αr+1 ≥ · · · ≥ αn
(cr+1 ≥ · · · ≥ cn ) in σJ− (A) (σJ− (C)). If the eigenvalues of A and C do not interlace,
then statements i.and ii. hold.
36 N. Bebiano et al.
n
tr(CA) ≤ ci αi .
i=1
r
n
ci αr−i+1 + ci αn+r−i+1 ≤ tr(CA).
i=1 i=r+1
Several other inequalities of indefinite type have been studied. For instance,
just to mention a few, we refer some spectral inequalities for the trace of the
exponential or the logarithmic of J -Hermitian matrices [15], operator inequalities
associated with Furuta inequality of indefinite type [16], a reversed Heinz-Kato-
Furuta inequality [17] and indefinite versions of some determinantal inequalities
[19], including a Fiedler-type theorem for the determinant of J -positive matrices
[18].
Recently, Matharu, Malhotra and Moslehian [74] defined a J -mean associated
with a positive matrix monotone function f on (0, ∞), such that f (1) = 1, for J -
Hermitian matrices with spectra in (0, ∞). Fundamental properties of this J -mean,
such as the power monotonicity and an indefinite version of Ando-Hiai inequality
[74, Theorem 3.11] were obtained.
Acknowledgments The authors would like to thank the referees for their valuable comments
which helped to improve the presentation of this chapter. The work of the first author was partially
supported by the Centre for Mathematics of the University of Coimbra - UIDB/00324/2020,
funded by the Portuguese Government through FCT/MCTES. The work of the second author was
supported by Portuguese funds through the Center for Research and Development in Mathematics
and Applications (CIDMA) and the Portuguese Foundation for Science and Technology (FCT-
Fundação para a Ciência e a Tecnologia), project UIDB/04106/2020. The work of the third author
was financed by CMAT-UTAD and Portuguese funds through the Portuguese Foundation for Sci-
ence and Technology (FCT-Fundação para a Ciência e a Tecnologia), reference UIDB/00013/2020.
References
1. T. Ando, Concativity of certain maps of positive definite matrices and applications to Hadamard
products. Linear Algebra Appl. 26, 203–241 (1979)
2. T. Ando, Majorization, doubly stochastic matrices and comparison of eigenvalues. Linear
Algebra Appl. 118, 163–248 (1989)
3. T. Ando, Majorizations and inequalities in matrix theory. Linear Algebra Appl. 199, 17–67
(1994)
4. T. Ando, Majorization relations for Hadamard products. Linear Algebra Appl. 223/224, 57–64
(1995)
5. T. Ando, F. Hiai, Log-majorization and complementary Golden-Thompson type inequality.
Linear Algebra Appl. 197, 113–131 (1994)
Log-majorization Type Inequalities 37
6. T. Ando, Löwner inequality of indefinite type. Linear Algebra Appl. 385, 73–80 (2004)
7. H. Araki, On an inequality of Lieb and Thirring. Lett. Math. Phys. 19, 167–170 (1990)
8. T. Azizov, I. Iokhvidov, Linear Operators in Spaces with an Indefinite Metric (Nauka, Moscow,
1986). English Translation: Wiley, New York, 1989
9. N. Bebiano, Contradomínios Numéricos Generalizados: variações sobre este tema, PhD Thesis,
Universidade de Coimbra, 1984
10. N. Bebiano, C. Perdigão, Extremal matrices in certain determinantal inequalities. Linear
Multilinear Algebra 44, 261–276 (1998)
11. N. Bebiano, J. da Providência Jr., R. Lemos, Matrix inequalities in statistical mechanics. Linear
Algebra Appl. 376, 265–273 (2004)
12. N. Bebiano, R. Lemos, J. da Providência, Inequalities for quantum relative entropy. Linear
Algebra Appl. 401, 159–172 (2005)
13. N. Bebiano, H. Nakazato, J. da Providência, R. Lemos, G. Soares, Inequalities for J -Hermitian
matrices. Linear Algebra Appl. 407, 125–139 (2005)
14. N. Bebiano, R. Lemos, J. da Providência, G. Soares, Further developments of Furuta inequality
of indefinite type. Math. Inequal. Appl. 13, 523–535 (2010)
15. N. Bebiano, R. Lemos, J. da Providência, G. Soares, Trace inequalities for logarithms and
powers of J -Hermitian matrices. Linear Algebra Appl. 432, 3172–3182 (2010)
16. N. Bebiano, R. Lemos, J. da Providência and G. Soares, Operator inequalities for J -
contractions. Math. Inequal. Appl. 12, 883–897 (2012)
17. N. Bebiano, R. Lemos, J. da Providência, On a reverse Heinz-Kato-Furuta inequality. Linear
Algebra Appl. 437, 1892–1905 (2012)
18. N. Bebiano, J. da Providência, A Fiedler-type theorem for the determinant of J -positive
matrices. Math. Inequal. Appl. 19, 663–669 (2016)
19. N. Bebiano, J. da Providência, Determinantal inequalities for J -accreative dissipative matrices.
Studia Universitatis Babeş-Bolyai Mathematica 62, 119–125 (2017)
20. R. Bhatia, Linear Algebra to Quantum Cohomology: the story of Alfred Horn’s inequalities,
Am. Math. Monthly 108, 289–318 (2001)
21. R. Bhatia, Matrix Analysis. Graduate Texts in Mathematics, vol. 169 (Springer, New York,
1997)
22. R. Bhatia, J. Holbrook, Riemannian geometry and matrix geometric means. Linear Algebra
Appl. 413, 594–618 (2006)
23. R. Bhatia, Y. Lim, T. Yamazaki, Some norm inequalities for matrix means. Linear Algebra
Appl. 501, 112–122 (2016)
24. G. Birkhoff, Tres observaciones sobre el algebra lineal. Rev. Universidad Nacional de
Tucumán, Ser. A 5, 147–151 (1946)
25. N.N. Chan, K. Kwong, Hermitian matrix inequalities and a conjecture. Am. Math. Monthly
92, 533–541 (1985)
26. G.P. Egorychev, The solution of van der Waerden’s problem for permanents. Adv. Math. 42,
299–305 (1981)
27. D.I. Falikman, The proof of the van der Waerden’s conjecture regarding the permanent of
doubly stochastic matrices. Mat. Zametki 29(6), 931–938 (1981); Math. Notes 29, 475–479
(1981)
28. K. Fan, Maximum properties and inequalities for the eigenvalues of completely continuous
operators. Proc. Nat. Acad. Sci. U.S.A. 37, 760–766 (1951)
29. M. Fiedler, Bounds for the determinant of the sum of Hermitian matrices. Proc. Am. Math.
Soc. 30(1), 27–31 (1971)
30. P.J. Forrester, C.J. Thompson, The Golden-Thompson inequality: historical aspects and random
matrix applications. J. Math. Phys. 55, 023503 (2014)
31. J.I. Fujii, T. Furuta, Löwner-Heinz, Cordes and Heinz-Kato inequalities. Math. Jpn. 38, 73–78
(1993)
32. J.I. Fujii, E. Kamei, Relative operator entropy in non-commutative information theory. Math.
Jpn. 34, 341–348 (1989)
38 N. Bebiano et al.
33. M. Fujii, E. Kamei, Ando-Hiai inequality and Furuta inequality. Linear Algebra Appl. 416,
541–545 (2006)
34. M. Fujii, R. Nakamoto, M. Tominaga, Generalized Bebiano-Lemos-Providência inequalities
and their reverses. Linear Algebra Appl. 426, 33–39 (2007)
35. T. Furuta, A ≥ B ≥ 0 assures (B r Ap B r )1/q ≥ B {p+2r)/q for r ≥ 0, p ≥ 0, q ≥ 1 with
(1 + 2r)q ≥ p + 2r. Proc. Am. Math. Soc. 101, 85–88 (1987)
36. T. Furuta, Invitation to Linear Operators: From Matrices to Bounded Linear Operators on a
Hilbert Space (CRC Press, Boca Raton, 2001)
37. T. Furuta, An extension of the Furuta inequality and Ando-Hiai log-majorization. Linear
Algebra Appl. 219, 139–155 (1995)
38. T. Furuta, Operator inequality implying generalized Bebiano-Lemos-Providência one, Linear
Algebra Appl. 426, 342–348 (2007)
39. T. Furuta, Extensions of inequalities for unitarily invariant norms via log-majorization. Linear
Algebra Appl. 436, 3463–3468 (2012)
40. S. Golden, Lower bounds for Helmholtz functions. Phys. Rev. B 137, 1127–1128 (1965)
41. R.D. Grigorieff, Note on von Neumann’s trace inequality. Math. Nachr. 151, 327–328 (1991)
42. F. Hansen, Multivariate extensions of the Golden-Thompson inequality. Ann. Funct. Anal. 6(4),
301–310 (2015)
43. G.H. Hardy, J.E. Littlewood, G. Pólya, Inequalities (Cambridge University Press, Cambridge,
1952)
44. E. Heinz, Beiträge zur Störungstheorie der Spektralzerlegung. Math. Ann. 123, 415–438
(1951)
45. F. Hiai, D. Petz, The Golden-Thompson trace inequality is complemented. Linear Algebra
Appl. 181, 153–185 (1993)
46. F. Hiai, Matrix analysis: matrix monotone functions, matrix means and majorization. Interdis-
cipl. Inf. Sci. 16(2), 139–248 (2010)
47. F. Hiai, D. Petz, Introduction to Matrix Analysis and Applications. Universitext (Springer, New
York, 2014)
48. F. Hiai, M. Lin, On an eigenvalue inequality involving the Hadamard product. Linear Algebra
Appl. 515, 313–320 (2017)
49. F. Hiai, Y. Seo, S. Wada, Ando-Hiai type inequalities for multivariate operator means. Linear
Multilinear Algebra 67, 2253–2281 (2019)
50. F. Hiai, Y. Seo, S. Wada, Ando-Hiai type inequalities for operator means and operator
perspectives. Int. J. Math. 31(1), 2050007 (2020)
51. A. Horn, Doubly stochastic matrices and the diagonal of a rotation matrix. Am. J. Math. 76,
620–630 (1954)
52. A. Horn, On the eigenvalues of a matrix with prescribed singular values. Proc. Am. Math. Soc.
5, 4–7 (1954)
53. A. Horn, Eigenvalues of sums of Hermitian matrices. Pac. J. Math. 12, 225–241 (1962)
54. R.A. Horn, C.R. Johnson, Matrix Analysis (Cambridge University Press, Cambridge, 1985)
55. R.A. Horn, C.R. Johnson, Topics in Matrix Analysis (Cambridge University Press, Cambridge,
1991)
56. C.R. Johnson, R.B. Bapat, A weak multiplicative majorization conjecture for Hadamard
products. Linear Algebra Appl. 104, 246–247 (1988)
57. M. Kian, M.S. Moslehian, Y. Seo, Variants of Ando-Hiai inequality for operator power means.
Linear Multilinear Algebra 69, 1694–1704 (2021)
58. M. Kian, M.S. Moslehian, Y. Seo, Variants of Ando-Hiai type inequalities for deformed means
and applications. Glasgow Math. J. 63, 622–639 (2021)
59. M. Kian, M.S. Moslehian, Matrix inequalities related to power means of probability measures.
Linear Multilinear Algebra (2021), https://doi.org/10.1080/03081087.2021.1930991
60. A. Klyachko, Stable bundles, representation theory and Hermitian operators. Selecta Math.,
New Series 4, 419–445 (1998)
61. A. Knutson, T. Tao, The Honeycomb model of the Berenstein-Zelevinsky cone I: proof of the
saturation conjecture. J. Am. Math. Soc. 12, 1055–1090 (1999)
Log-majorization Type Inequalities 39
62. F. Kubo, T. Ando, Means of positive linear operators. Math. Ann. 246, 205–224 (1980)
63. J. Lawson, Y. Lim, Karcher means and Karcher equations of positive definite operators. Trans.
Am. Math. Soc. Ser. B. 1, 1–22 (2014)
64. R. Lemos, G. Soares, Some log-majorizations and an extension of a determinantal inequality.
Linear Algebra Appl. 547, 19–31 (2018)
65. R. Lemos, G. Soares, Spectral inequalities for Kubo-Ando operator means. Linear Algebra
Appl. 607, 29–44 (2020)
66. A. Lenard, Generalization of the Golden-Thompson inequality tr(eA eB ) ≥ tr(eA+B ). Indiana
Univ. Math. J. 21, 457–467 (1971)
67. E. Lieb, Convex trace functions and the Wigner-Yanase-Dyson conjecture. Adv. Math. 11,
267–288 (1973)
68. E.H. Lieb, W.E. Thirring, Studies in Mathematical Physics, ed. by E. Lieb, B. Simon, A.S.
Wightmer (Princeton University Press, Princeton, 1976), pp. 269–303
69. K. Löwner, Über monotone Matrixfunktionen. Math. Z. 38, 177–216 (1934)
70. M. Marcus, An eigenvalue for the product of normal matrices. Am. Math. Monthly 63, 173–
174 (1956)
71. M. Marcus, Derivations, Plücker relations and the numerical range. Indiana Univ. Math. J. 22,
1137–1149 (1973)
72. A.W. Marshall, I. Olkin, B.C. Arnold, Inequalities: Theory of Majorization and Its Applica-
tions, 2nd edn. (Springer, New York, 2011)
73. J.S. Matharu, J.S. Aujla, Some inequalities for unitarily invariant norms. Linear Algebra Appl.
436, 1623–1631 (2012)
74. J.S. Matharu, C. Malhotra, M.S. Moslehian, Indefinite matrix inequalities via matrix means.
Bull. Sci. Math. 171, 103036 (2021)
75. A. Matsumoto, R. Nakamoto, M. Fujii, Reverse of Bebiano-Lemos-Providência inequality
and Complementary Furuta inequality (Inequalities on Linear Operators and its Applications),
Departmental Bulletin Paper, Kyoto University, 2008, 91–98
76. L. Mirsky, On the trace of matrix products. Math. Nachr. 20, 171–174 (1959)
77. L. Mirsky, A trace inequality of John von Neumann. Monatshefte für Mathematik 79, 303–306
(1975)
78. M. Moakher, A differential geometric approach to the geometric mean of symmetric positive-
definite matrices. SIAM J. Matrix Anal. Appl. 26, 735–747 (2005)
79. G.N. de Oliveira, Normal matrices (research problem). Linear Multilinear Algebra 12, 153–154
(1982)
80. W. Pusz, S.L. Woronowicz, Functional calculus for sesquilinear forms and the purification
map. Rep. Math. Phys. 8, 159–170 (1975)
81. H. Richter, Zur abschatzung von matrizen-normen. Math. Nachr. 18, 178–187 (1958)
82. A. Ruhe, Perturbation bounds for means of eigenvalues and invariant subspaces. BIT Numer.
Math. 10, 343–354 (1970)
83. T. Sano, Furuta inequality of indefinite type. Math. Inequal. Appl. 10, 381–387 (2007)
84. I. Schur, Über eine Klasse von Mittelbildungen mit Anwendungen auf die Determinantenthe-
orie. Sitzungsber. Berl. Math. Ges. 22, 9–20 (1923)
85. D. Sutter, M. Berta, M. Tomamichel, Multivariate trace inequalities. Commun. Math. Phys.
352, 37–58 (2017)
86. K. Symanzik, Proofs and refinements of an inequality of Feynman. J. Math. Phys. 6, 1155–
1156 (1965)
87. T. Tao, The Golden-Thompson inequality - What’s new.terrytao.wordpress.com/2010/07/15/
the-golden-thompson-inequality (2010)
88. C.M. Theobald, An inequality for the trace of the product of two symmetric matrices. Math.
Proc. Cambridge Philos. Soc. 77, 265–267 (1975)
89. C.J. Thompson, Inequality with applications in statistical mechanics. J. Math. Phys. 6, 1812–
1813 (1965)
90. C.J. Thompson, Inequalities and partial orders on matrix spaces. Indiana Univ. Math. J. 21,
469–480 (1971)
40 N. Bebiano et al.
91. R.C. Thompson, Proof of a conjectured exponential formula. Linear Multilinear Algebra 19,
187–197 (1986)
92. H. Umegaki, Condition expectation in an operator algebra IV. Kodai Math. Semin. Rep. 14,
59–85 (1962)
93. B.L. van der Waerden, Aufgabe 45. Jber. Deutsch. Math. Verein. 35, 117 (1926)
94. G. Visick, A weak majorization involving the matrices A ◦ B and AB. Linear Algebra Appl.
223/224, 731–744 (1995)
95. G. Visick, Majorizations of Hadamard products of matrix powers. Linear Algebra Appl. 269,
233–240 (1998)
96. J. von Neumann, Some matrix-inequalities and metrization of matrix-space. Tomsk Univ. Rev.
1, 286–300 (1937)
97. J. Wang, Y. Li, H. Sun, Lie-Trotter Formula for the Hadamard Product. Acta Math. Sci. 40,
659–669 (2020)
98. H. Weyl, Inequalities between the two kinds of eigenvalues of a linear transformation Proc.
Natl. Acad. Sci. U. S. A. 35, 408–411 (1949)
99. T. Yamazaki, The Riemannian mean and matrix inequalities related to the Ando-Hiai inequality
and chaotic order. Oper. Matrices 6, 577–588 (2012)
100. T. Yamazaki, The Ando-Hiai inequality for the solution of the generalized Karcher equation
and related results. J. Math. Anal. Appl. 479, 531–545 (2019)
101. X. Zhan, Matrix Inequalities. Lecture Notes in Mathematics (Springer, Berlin, 2002)
102. L. Zou, An arithmetic geometric mean inequality for singular values and its applications.
Linear Algebra Appl. 528, 25–32 (2017)
Ando-Hiai Inequality: Extensions and
Applications
Abstract The Ando-Hiai inequality says that if A#α B ≤ I for a fixed α ∈ [0, 1]
and positive invertible operators A, B on a Hilbert space, then Ar #α B r ≤ I for
1 1 1 1
r ≥ 1, where #α is the α-geometric mean defined by A#α B = A 2 (A− 2 BA− 2 )α A 2 .
This chapter is devoted by extensions and applications of Ando-Hiai inequality. It
is closely related to Furuta inequality, Bebiano-Lemos-Providência inequality and
grand Furuta inequality. Consequently they are given useful extensions.
1 Introduction
A#α B ≤ I ⇒ Ar #α B r ≤ I (r ≥ 1).
M. Fujii ()
Department of Mathematics, Osaka Kyoiku University, Kashiwara, Osaka, Japan
e-mail: [email protected]
R. Nakamoto
3-4-13 Daihara-cho, Hitachi, Japan
e-mail: [email protected]
It holds for positive operators A, B on a Hilbert space, and is called the Ando-Hiai
inequality, denoted by (AH) simply .
Now the original proof of (AH) is a typical application of Löwner-Heniz
inequality (LH), i.e.,
A ≥ B ≥ 0 ⇒ Ar ≥ B r (0 ≤ r ≤ 1).
and
r r 1 r r 1
(ii) (A 2 Ap A 2 ) q ≥ (A 2 B p A 2 ) q
Related to Furuta inequality, see [4, 5, 15, 16, 31] and [9].
After publishing (AH), Furuta himself [17] presented so-called “grand Furuta
inequality” which interpolates (AH) to his inequality (FI), see also [8] and [9].
Grand Furuta Inequality (GFI) If A ≥ B > 0 and t ∈ [0, 1], then
1−t+r
[A 2 (A− 2 B p A− 2 )s A 2 ] (p−t)s+r ≤ A1−t +r
r t t r
Ando-Hiai Inequality 43
A #α B ≤ I ⇒ Ar # αr+(1−α)
αr B≤I (r ≥ 1);
A #α B ≤ I ⇒ A # α+(1−α)s
α Bs ≤ I (s ≥ 1).
It is known that both one-sided versions are equivalent, and that they are
alternative expressions of (FI).
2 Extensions
This section is based on our paper [11]. First of all, a binary operation α is defined
by the same formula as the α-geometric mean for α ∈ [0, 1], that is,
A α B = A 2 (A− 2 BA− 2 )α A 2
1 1 1 1
for A, B > 0.
Recently (AH) is extended by Seo [29] and [23] as follows: For α ∈ [−1, 0],
A α B ≤ I for A, B > 0 implies Ar α B r ≤ I for r ∈ [0, 1].
So, following our previous work, we present two variable version of it. For this,
we mention the following useful identity on the binary operation : For β ∈ R and
positive invertible operators X and Y ,
X β Y = X(X−1 −β Y
−1
)X. (1)
A βC = A (A− #−β C −1 )A
≤ A (C −α #−β C −1 )A
= A C α(1−2 )A
≤ A A1−2 A = A.
A−r = A− 2 (A − 12
≤ A− 2 AA− 2 = I.
1 1 1
βB β C)A
α1 = 2−αα
. Hence it follows from the preceding paragraph that for r ∈ [ 12 , 1],
1
I ≥ (A− 2 )r = A− 2
r
β1 B β1 B,
α1 r αr/2
where β1 = α1 r+(1−α1 ) = αr/2+(1−α) . This means that the desired inequality holds
for r ∈ [ 14 , 12 ]. Finally we have the conclusion by the induction.
Lemma 2.2 If A α B ≤ I for α ∈ [−1, 0] and A, B > 0, then A β B s ≤ I for
s ∈ [ −2α
1−α , 1], where β = α+(1−α)s .
α
so that
A β B −s = B − 2 (D β B )B − 2 ≤ B − 2 BB − 2 = I.
1 1 1 1
Theorem 2.3 If A α B ≤ I for α ∈ [−1, 0] and A, B > 0, then Ar βB
s ≤ I for
r ∈ [0, 1] and s ∈ [ −2αr
1−α , 1], where β = αr+(1−α)s .
αr
Then we have
1 ≥ Ar γ B s = Ar αr Bs
γ +(1−γ )s αr+(1−α)s
A−r # 1+r B p ≤ A
p+r
holds for p ≥ 1.
More precisely, the conclusion in above is improved by
A−r # 1+r B p ≤ B (≤ A)
p+r
A−r 1+r Bp ≤ A
p+r
As well as (FI), (GFI) has also mean theoretic expression as follows:
If A ≥ B > 0 and t ∈ [0, 1], then
A−r+1 r
r+(p−1)s
(A#s B p ) ≤ A
−2r
holds for p ≤ −1, r ∈ [0, 1] and s ∈ [ p−1 , 1].
Proof Theorem 2.3 says that if A−1 α B ≤ I , then A−r β B s ≤ I for r ∈ [0, 1]
and s ∈ [ −2αr
1−α , 1], where β = αr+(1−α)s . So the assumption is that B1 = C ≤ A,
αr α
1 1
where C = A 2 BA 2 . On the other hand, putting α = 1
p, it follows that
1 1
I ≥ A−r αr
αr+(1−α)s
B s = A−r r
r+(p−1)s
(A− 2 B1 p A− 2 )s
or equivalently
A ≥ A−r+1 r
r+(p−1)s
(A#s B1 p ),
p 1 1
which is the conclusion by B1 = C = A 2 BA 2 .
Ando-Hiai Inequality 47
−2r
holds for p ≤ −1, r ∈ [0, t] and s ∈ [ p−t , 1].
We can prove it under a restriction as follows:
Theorem 2.8 If A ≥ B > 0 and t ∈ [0, 1], then
−t −2r−(1−t )
holds for p ≤ −1, r ∈ [0, t] and s ∈ [max{ p−t , p−t }, 1].
−t
Proof First of all, we note that r + (p − t)s ≤ t + (p − t)s ≤ 0 by s ≥ p−t . So we
1−t +r 1−t +r
have r+(p−t )s ≤ 0. On the other hand, −1 ≤ r+(p−t )s is obtained by −(r + (p −
−2r−(1−t )
t)s) ≥ 1 − t + r since the assumption s ≥ p−t . Namely −(1−t +r)
r+(p−t )s ∈ [0, 1].
Hence we have
The second inequality in above is shown as follows: The exponent −(p − t)s − t
of B is nonnegative as mentioned first. Thus, if −(p − t)s − t ≤ 1, the second
inequality holds. On the other hand, if −(p − t)s − t ≥ 1, then the Furuta inequality
assures that
t−r 1−t+r
B −(p−t )s−t A
t−r
(A 2 2 ) −(p−t)s−r ≤ A1−t +r ,
or equivalently
The following theorems show that Theorem 2.8 holds for the critical points s =
−t −2r−(1−t )
p−t , p−t .
48 M. Fujii and R. Nakamoto
−2r−(1−t )
holds for p ≤ −1, r ∈ [0, t] and s = p−t .
1−t +r
Proof First of all, we note that r+(p−t )s = −1 and X −1 Y = XY −1 X. Therefore
the conclusion is arranged as
A−r+t −1 (At #s B p ) ≤ A,
and so
t t 1 P +t
(A 2 B P A 2 ) q ≤ A q
(A 2 B −p A 2 )s ≤ A1+2r−t ,
t t
−t
holds for p ≤ −1, r ∈ [0, t] and s = p−t .
−t −2r−(1−t )
By Theorem 2.8, we have to consider the case p−t < p−t , that is, 0 ≤
t −r < 1
2 can be assumed. Hence we have
1−t +r 1
=1− < −1.
r + (p − t)s t −r
Ando-Hiai Inequality 49
1−t +r
As a special case, we take t = 23 , r = 13 and p = −2. Then s = 1
4 and r+(p−t )s =
−2, so that the statement in this case is arranged as follows:
If A ≥ B > 0, then
1 2
−2
A3 −2 (A 3 # 1 B )≤A
4
holds? It is proved by using Furuta inequality twice: First of all, since A ≥ B > 0,
(FI) ensures that
1 1 5 5 1 1 1 1
(A 3 B 2 A 3 ) 8 ≤ A 3 ; (A 3 B 2 A 3 ) 8 ≤ A 3 .
So we have
−2
) = A 6 (A− 6 (A 3 # 1 B −2 )A− 6 )−2 A 6
1 2 1 1 2 1 1
A3 −2 (A 3 # 1 B
4 4
1 1 2 1 1
= A 6 (A 6 (A− 3 # 1 B 2 )A 6 )2 A 6
4
= A 6 (A− 3 # 1 A 6 B 2 A 6 )2 A 6
1 1 1 1 1
4
1 1 1 1 1 1 1
= A 6 (A− 6 (A 3 B 2 A 3 ) 4 A− 6 )2 A 6
= (A 3 B 2 A 3 ) 4 A− 3 (A 3 B 2 A 3 ) 4
1 1 1 1 1 1 1
1 1 1 1 1
≤ (A 3 B 2 A 3 ) 4 − 8 + 4
1 1 3
≤ (A 3 B 2 A 3 ) 8
≤ A,
as desired.
To prove Theorem 2.10, we cite a lemma obtained by the Furuta inequality.
Lemma 2.11 If A ≥ B > 0, t ≥ 0 and p ≤ −1, then
1+t
(A 2 B −p A 2 ) −p+t ≤ A1+t ;
t t
t t
in particular, (A 2 B −p A 2 )s ≤ At holds for s = t
−p+t .
To show Theorem 2.10, we reformulate it as follows:
Theorem 2.12 If A ≥ B > 0, t ≥ c−1
c+1 for some c ≥ 2, 1 ≥ t > r ≥ 0 with
t − r = c+1
1
and p ≤ −1, then
1
A c+1 −c (At #s B p ) ≤ A
50 M. Fujii and R. Nakamoto
holds for s = t
−p+t .
1 α α s 1
A c+1 −c (At #s B p ) = A 2 Y c A 2 , and X t = X −p+t ≤ A, in particular, Xs ≤ At
st
and X t ≤ At for 0 ≤ t ≤ 1 + t by Lemma 2.11.
(1) First we suppose that 2n ≤ c < 2n + 1 for some n, i.e., c = 2n + for
some ∈ [0, 1). Since α(c − 2) ≤ t − α = r by α(c − 1) ≤ t, we have
α ≤ α(2(n − 1) + ) = α(c − 2) ≤ r and so
α −r α(2(n − k) + ) − r
−1 ≤ ≤ ≤0
t t
= Y n Y Y n = Y n (A− 2 Xs A− 2 ) Y n
r r
Yc
≤ Y n (A− 2 At A− 2 ) Y n = Y n Aα Y n
r r
by Xs ≤ At and (LH)
= Y n−1 A− 2 Xs Aα −r
Xs A− 2 Y n−1
r r
α −r
≤ Y n−1 A− 2 X2s+(α −r) st
A− 2 Y n−1
r r
by Xs ≤ At , ∈ [−1, 0]
t
−2r
≤ Y n−1 A2t+α Y n−1 by putting t = 2t + α − r ≤ 1 + t
≤ Y Aα(2(n−1)+ ) Y
≤ Aα(2n+ )
= Aαc .
Hence we have
1 α α
A c+1 −c (At #s B p ) = A 2 Y c A 2 ≤ Aαc+α = A,
as desired.
Ando-Hiai Inequality 51
Y 1+ ≤ Aα(1+ ).
It is proved as follows:
Y 1+ = (A− 2 Xs A− 2 )1+
r r
= A− 2 X 2 (X 2 A−r X 2 ) X 2 A− 2
r s s s s r
≤ A− 2 X 2 (X 2 X− t X 2 ) X 2 A− 2
r s s sr s s r
= A− 2 Xs+(s− t A− 2
r sr r
)
≤ A− 2 At +α A− 2 = Aα(1+ ).
r r
Y c = Y n Y 1+ Y n ≤ Y n Aα(1+ ) Y n
= Y n−1 A− 2 Xs Aα(1+
r
Xs A− 2 Y n−1
r
)−r
as desired.
52 M. Fujii and R. Nakamoto
Recently Ito and Kamei [20] give an improvement to the above results. For this,
they rewrite it as follows:
and
−r
=A # 1−r B
1+r
−r
≤ A # 1−r A = A1−2r .
1+r
−r −p
= A (B # 1+p Ar )A−r ≥ A−r AA−r = A1−2r .
r+p
by (SF).
Ando-Hiai Inequality 53
1−t +2r
holds for p ≥ 1 and p+t ≤ s ≤ 1.
1−t +r
Proof Noting that 0 ≤ (p+t )s−r ≤ 1 and 0 ≤ t − r ≤ 1 hold, we have
Next we show Ar−t # 1−t+r (A−t #s B p ) ≤ A1−2(t −r) by dividing into three
(p+t)s−r
cases:
(i) If (p + t)s − t ≥ 1 holds, then
≤ Ar−t # 1−t+r (B −t #s B p )
(p+t)s−r
= Ar−t # 1−t+r B
1+t−r
≤A r−t
# 1−t+r A = A1−2(t −r).
1+t−r
≤ Ar−t # 1−t+r (B −t #s B p )
(p+t)s−r
so that we have
Therefore we obtain the desired result since B 1−2(t −r) ≤ A1−2(t −r) holds if
0 ≤ t − r ≤ 12 and A1−2(t −r) ≤ B 1−2(t −r) holds if 12 ≤ t − r ≤ 1.
3 Applications
X β Y = X(X−1 −β Y −1 )X.
This means that if β ∈ [−1, 0], then β looks like an operator mean in some sense.
We also rewrite Lemma 2.1 and Theorem 2.6 for convenience.
Lemma 3.1 If A α B ≤ 1 for α ∈ [−1, 0] and A, B > 0, then Ar β B ≤ 1 for
r ∈ [0, 1], where β = αr+(1−α)
αr
.
It is reformulated as Furuta type as follows:
Theorem 3.2 If A ≥ B > 0, then
A−r 1+r Bp ≤ A
p+r
Ando-Hiai Inequality 55
A−r # 1+r B p ≤ A
p+r
by (LH).
As an application, we show a generalization of BLP inequality in the below. For
this, we need the following lemma:
Lemma 3.4 Suppose that A, B > 0.
(1) If Ar 1 B p+r ≤ A1+r for some p ≤ −1 and r ∈ [−1, 0], then B 1+r ≤ A1+r .
p
(2) If Ar 1 B p+r ≤ A1+r for some p ∈ [0, 1] and r ≤ −1, then B 1+r ≤ A1+r .
p
1
Proof (1) Since the assumption is rephrased as B1 = (A− 2 B p+r A− 2 ) p ≤ A, it
r r
p+r p+r
Theorem 3.8 Suppose that A ≥ 0, 0 < mI ≤ B ≤ MI for some M > m > 0 and
h= Mm . Then
1 r r 1 1
A 2 (A 2 B p+r A 2 ) p A 2
p p − p1
≤ A− 2 (A 2 B p+r A 2 )−1 A− 2
r r
Ando-Hiai Inequality 57
p + r − p1 1+r
Ar 1 B p+r ≤ K(h1+s , − ) A
p 1+s
− p1
Ar 1 B p+r ≤ K(h, −(p + r)) A1+r
p
in general.
By the way, Theorem 3.7 is expressed that
(1) hq − 1 (2) hq − 1
j1 = jq,h = , j 2 = j =
q(hq − hq−1 ) q,h
q(h − 1)
and
⎧
⎪
⎪ (1 − λ)M if 0 < λ < j1
⎪
⎨ 1
β(m, M, q; λ) = q−1 M q −mq q−1 λ(Mmq −mM q )
⎪ + M q −mq if λ ∈ [j1 , j2 ]
⎪
⎪
q λq(M−m)
⎩(1 − λ)m if λ > j2 .
1 r r 1 1 p(1+s)
A 2 (A 2 B p+r A 2 ) p A 2 p+r
Proof We first refer [14, Theorem 6]: If A1 > 0, 0 < m1 I ≤ B1 ≤ M1 I for some
M1 > m1 > 0 and q > 1, then for each λ > 0
q q q 1
A1 B1 A1 q ≤ λA1 B1 A1 + β(m1 , M1 , q; λ)A1 2 .
1+s
We apply it for A1 = A 2 , B1 = B 1+s and q = − p+r
1+s . Then it follows that
1 r r 1 1 p(1+s)
A 2 (A 2 B p+r A 2 ) p A 2 p+r
1 r r 1 1 −p
= A 2 (A 2 B p+r A 2 ) p A 2 q
p p 1
≤ A− 2 (A 2 B p+r A 2 )−1 A− 2 q
r r
by (AC)
p+r p+r 1
= A− 2 B −(p+r) A− 2 q
1
q q q
= A1 B1 A1 q
t−r t−r 1
B2 = [A 2 (Ar # 1 B (p−t )s+r )A 2 ] p ≤ A.
s
At −r #
p
1−t+r (At s B2 ) ≤ A.
(p−t)s+r
Moreover we have
p t−r t−r
At s B2 = At s [A 2 (Ar # 1 B (p−t )s+r )A 2 ]
s
s
60 M. Fujii and R. Nakamoto
t−r t−r
=A 2 B (p−t )s+r A 2 ,
so that
A ≥ At −r #
p
1−t+r (At s B2 )
(p−t)s+r
= At −r #
t−r t−r
1−t+r A 2 B (p−t )s+r A 2
(p−t)s+r
t−r
B 1−t +r A
t−r
=A 2 2 ,
as desired.
As a corollary, we have a norm inequality of BLP type corresponding to (GFI):
Corollary 3.13 Suppose that A, B > 0 and t ∈ [0, 1]. Then
1−t+r 1−t+r
A 2 B 1−t +r A 2 ≤ I . Since the assumption is equivalent to
−t −2r−(1−t )
holds for p ≤ −1, r ∈ [0, t] and s ∈ [max{ p−t , p−t }, 1].
Consequently, as similar to (GFI) itself, we have the following operator inequal-
ity and norm inequality.
Theorem 3.14 Suppose that A, B > 0 and t ∈ [0, 1] satisfy
−t −2r−(1−t )
for some p ≤ −1, r ∈ [0, t] and s ∈ [max{ p−t , p−t }, 1]. Then B 1−t +r ≤
A 1−t +r .
Corollary 3.15 Suppose that A, B > 0 and t ∈ [0, 1]. Then
(p−t)s+r 1
B 1−t +r A ps(1−t+r) ≤ A 2 {A− 2 (A 2 B (p−t )s+r A 2 ) s A− 2 } p A 2
1−t+r 1−t+r 1 t r r 1 t 1
A 2 2
−t −2r−(1−t )
holds for p ≤ −1, r ∈ [0, t] and s ∈ [max{ p−t , p−t }, 1].
Proof of Theorem 3.14 The proof is similar to that of Theorem 3.12. As in the
proof of it, we have
t−r t−r 1
B2 = [A 2 (Ar 1 B (p−t )s+r )A 2 ] p ≤ A.
s
At −r #
p
1−t+r (At s B2 ) ≤ A.
(p−t)s+r
p t−r t−r
Moreover, since we have At s B2 = A 2 B (p−t )s+r A 2 , it follows that
A ≥ At −r # B2 ) = At −r #
p t−r t−r
1−t+r (At s 1−t+r A 2 B (p−t )s+r A 2 .
(p−t)s+r (p−t)s+r
By multiplying A−
t−r
2 on both sides, we obtain the conclusion.
In succession, we discuss some inequalities on the logarithm. The chaotic order
A " B for A, B > 0 is defined by log A ≥ log B. It is weaker than the
usual Löwner order A ≥ B. The Furuta inequality for chaotic order is known in
[6]:
Chaotic Furuta Inequality (CFI)
If A " B for A, B > 0, then for each r ≥ 0,
r r 1 r r 1
(i) (B 2 Ap B 2 ) q ≥ (B 2 B p B 2 ) q
and
r r 1 r r 1
(ii) (A 2 B p A 2 ) q ≥ (A 2 Ap A 2 ) q
A−r # p+r
r B ≤ I
p
and B p # p A−r ≤ I
p+r
hold for p ≥ 0.
As an application of (CFI), we obtain that if A " B for A, B > 0, then
A−r # 1+r B p ≤ B
p+r
1
= A 2 log(A− 2 B p+r A− 2 ) p A 2 , the assumption is
r r r r
p S(A |B
1 r p+r )
Proof Since
rephrased as
1
log A ≥ log(A− 2 B p+r A− 2 ) p .
r r
1
Putting B1 = (A− 2 B p+r A− 2 ) p , we have A " B1 . Hence it follows from (CFI)
r r
that
so that Ar ≥ I # p+r
r B p+r = B r .
Next we show a generalization of the above, which is type of (GFI).
Theorem 3.17 Let A, B > 0 and t, r ≥ 0 be given. Then, if
We here recall that the following operator inequality of type of (GFI), see [9,
Theorem 3.16]: If A " X for A, X > 0, then
(p+t)s+r r t t r 1
A q ≥ [A 2 (A 2 Xp A 2 )s A 2 ] q
t+r
At +r ≥ [A 2 (A 2 Xp A 2 )s A 2 ] (p+t)s+r
r t t r
t+r
= [A 2 (A 2 [A− B (p+t )s+r )A−
r t t+r t+r t r
2 (Ar 1 2 ]A 2 )s A 2 ] (p+t)s+r
s
t+r
= [A 2 (A− 2 (Ar
r r
B (p+t )s+r )A− 2 )s A 2 ] (p+t)s+r
r r
1
s
t+r
= [A 2 (A− 2 B (p+t )s+r A− 2 A 2 ] (p+t)s+r
r r r r
= B t +r ,
k
k
λi (A) ≥ λi (B) for k = 1, · · · , n − 1
i=1 i=1
and
n
n
λi (A) = λi (B),
i=1 i=1
(5) λ
i=1 i (A) = λ1 k (A)) k = 1, · · · , n if A > 0.
(C
Cosequently, for A, B > 0, A (log) B if and only if det A = det B and
λ1 (Ck (A)) ≥ λ1 (Ck (B)) for k = 1, · · · , n. Incidentally we note that Ck (A α B) =
Ck (A) α Ck (B) for A, B > 0 by (2)-(4), so that matrix inequalities of Ando-Hiai
type implies log-majorization inequalities corresponding to them.
The following log-majorization inequality corresponds to Theorem 2.3:
If A α B ≤ 1 for α ∈ [−1, 0] and positive invertible operators A and B, then
Ar β B s ≤ 1 for r ∈ [0, 1] and s ∈ [ −2αr
1−α , 1], where β = αr+(1−α)s .
αr
−t −2r−(1−t )
holds for p ≤ −1, r ∈ [0, t] and s ∈ [max{ p−t , p−t }, 1].
1 1
Putting A1 = A−1 , B1 = A− 2 B p A− 2 and p = α1 , it implies that
+r
A1 α B1 ≤ I #⇒ A1−t
1 β (A1−t
1 s B1 ) ≤ I.
4 Concluding Remarks
A, B > 0, A#α B ≥ I ⇒ Ar #α B r ≥ I.
A, B > 0, A σf B ≥ I ⇒ Ar σf B r ≥ I for r ≥ 1,
Another improvement of (AH) is posed by Seo [28]: For a fixed α ∈ (0, 1),
There are deep discussion on Ando-Hiai inequality for n-variable operator means
in [34] and [22].
References
22. M. Kian, M.S. Moslehian, Y. Seo, Variants of Ando-Hiai inequality for operator power means.
Linear Multilinear Algebra 69, 1694–1704 (2021)
23. M. Kian, Y. Seo, Norm inequalities related to the matrix geometric mean of negative power.
Sci. Math. Jpn. (in Editione Electronica), e-2018. article 2018-7
24. F. Kubo, T. Ando, Means of positive linear operators. Math. Ann. 246, 205–224 (1980)
25. K. Löwner, Über monotone Matrix function. Math. Z. 38, 177–216 (1934)
26. R. Nakamoto, Y. Seo, A complement of the Ando-Hiai inequality and norm inequalities for the
geometric mean. Nihonkai Math. J. 18, 43–50 (2007)
27. G.K. Pedersen, Some operator monotone functions. Proc. Am. Math. Soc. 36, 309–310 (1972)
28. Y. Seo, On a reverse of Ando-Hiai inequality. Banach J. Math. Anal. 4, 87–91 (2010)
29. Y. Seo, Matrix trace inequalities related to the Tsallis relative entropy of negative order. J.
Math. Anal. Appl. 472, 1499–1508 (2019)
30. Y. Seo, M. Tominaga, A complement of the Ando-Hiai inequality. Linear Algebra Appl. 429,
1546–1554 (2008)
31. K. Tanahashi, Best possibility of the Furuta inequality. Proc. Am. Math. Soc. 124, 141–146
(1996)
32. S. Wada, Some ways of constructing Furuta-type inequalities. Linear Algebra Appl. 457, 276–
286 (2014)
33. S. Wada, When does Ando-Hiai inequality hold? Linear Algebra Appl. 540, 234–243 (2018)
34. T. Yamazaki, The Riemannian mean and matrix inequalities related to the Ando-Hiai inequality
and chaotic order. Oper. Matrices 6, 577–588 (2012)
Relative Operator Entropy
Abstract The relative operator entropy S(A|B) is an operator version of (the minis
of) the Kullback-Leibler divergence in the information theory. It is introduced an
extension, called solidarities A s B, of the Kubo-Ando operator means A m B and
moreover it is a tangent vector of the path of geometric means A #t B. So we discuss
mean-like properties and geometric ones in the manifold of the positive invertible
operators. In fact, this path A #t B is a geodesic and S(A|B) is the initial tangent
vector in this manifold with the principal fiber bundle, say the CPR geometry. The
former defines the multivariate power mean and the latter the Karcher mean. Related
to the quantum information theory, we discuss the Tsallis operator entropy and its
trace as the secant vector between A and A #t B.
1 Introduction
The relative operator entropy is derived from the Kubo-Ando operator means and
the Uhlmann relative operator entropy [18, 19]. So we begin with the Kubo-Ando
means [35] and the solidarities as their generalization in Sect. 2. We note that a
solidarity can be defined among the positive invertible operators, while it is not
always defined for non-invertible ones.
J. I. Fujii
Department of Educational Collaboration (Sciences; Mathematics and Information), Osaka
Kyoiku University, Kashiwara, Osaka, Japan
e-mail: [email protected]
Y. Seo ()
Department of Mathematics Education, Osaka Kyoiku University, Kashiwara, Osaka, Japan
e-mail: [email protected]
The theory of operator means is started at Ando’s lecture note [4] and established
as the Kubo-Ando theory [35]. For positive operators on a Hilbert space, the theory
of operator means is defined axiomatically: An (operator) connection m is a binary
operation on positive operators satisfying the following axioms:
• monotonicity: A1 ≤ A2 and B1 ≤ B2 imply A1 m B1 ≤ A2 m B2 .
• semi-continuity: An ↓ A and Bn ↓ B imply An m Bn ↓ A m B.
• transformer inequality: T ∗ (A m B)T ≤ (T ∗ AT ) m (T ∗ BT ).
An operator mean is a connection m satisfying
• normalization: A m A = A.
It is easy to show that the transformer inequality becomes equality if T is invertible.
For an operator mean m, the representing function fm (x) = 1 m x is operator
monotone:
1
A m◦ B = B m A, f ◦ (x) = xf
x
1
A m∗ B = (A−1 m B −1 )−1 , f ∗ (x) =
f (1/x)
x
A m⊥ B = (B −1 m A−1 )−1 , f ⊥ (x) = .
f (x)
A mt B − A
A sm B = lim .
t %0 t
f (1/ε) f (x)
lim F (ε) = lim = lim = lim f (x)
ε→0 ε→0 1/ε x→∞ x x→∞
by the l’Hospital theorem where the limit always exists since f is monotone
nonincreasing. In general, f is not defined at 0, but this fact shows F can be always
defined.
Lemma 2.1 If f is operator monotone on (0, ∞), then the corresponding solidarity
s is defined by f (x) = 1 s x and the transpose F (x) = x s 1 is operator concave
on [0, ∞). Conversely, if F is operator concave on [0, ∞) and F (0) 0, then
F (x) = x s 1 defines a solidarity., i.e., f (x) = 1 s x is operator monotone on
(0, ∞).
Proof Suppose f is operator monotone. We may assume A and B are positive
1 1 1 1
invertible operators. Then (A + B)− 2 A 2 and (A + B)− 2 B 2 are contractions. Since
f is also operator concave, the Jensen inequality of Davis-Hansen-Pedersen type
[10, 28, 29] shows
1
Multiplying (A + B) 2 from both sides in the above inequality, we have
Af (A−1 ) + Bf (B −1 ) A+B
≤ f (2(A + B)−1 ),
2 2
that is, F (A)+F
2
(B)
≤ F A+B 2 , which shows the operator concavity of F .
Conversely suppose F is operator concave and F (0) 0. Let A B for positive
invertible operators A and B. Then B − 2 A 2 is contractive and then we also have
1 1
1 1 1 1 1 1
B − 2 f (A)B − 2 = B − 2 A 2 F (A−1 )A 2 B − 2
≤ B − 2 A 2 F (A−1 )A 2 + (B − A) 2 F (0)(B − A) 2 B − 2
1 1 1 1 1 1
≤ F (B −1 + 0) = F (B −1 ),
Relative Operator Entropy 73
so that
f (A) ≤ BF (B −1 ) = f (B),
F (α)A + f (1/α)B ≥ C
A s B ≤ F (α)A + f (1/α)B,
F (α)A + f (1/α)B ≥ C
−1 −1
F (α)A + f (1/α)Bε ≥ C, that is, F (α)Xε + f (1/α) ≥ Bε 2 CBε 2 .
The left hand of the latter inequality in the above attains η(Xε ) = −Xε log Xε as the
infimum, that is, A s Bε ≥ C. Therefore A s Bε is bounded below and A s B exists
with A s B ≥ C.
74 J. I. Fujii and Y. Seo
Remark 2.3 In the proof of the above in [17], the only if part is a little ambiguous.
The above proof is a complete version of [17, Theorem 1]. In the case of relative
operator entropy, we showed it in [22].
If F is nonnegative, then F is also operator monotone, so that F defines the
transpose of some operator mean. When s is operator mean, the above property is
clear, so that we may assume that F (x0 ) = 0 for some x0 0. If the set of zero
points {x0 } consists of 0 only, then F is nonpositive and so is f . In the case x0 > 0,
F (x0 )
f (1/x0) = = 0.
x0
A mt B − A ∂
A sm B = lim = A mt B
t %0 t ∂t t =0
if the limit exists. Since fm (1) = 1, we have f (1) = F (1) = 0 for the
corresponding solidarity s.
if the limit (in the strong operator topology) exists as a bounded operator. But, in
general, S(A|B) does not always exist. On the other hand, based on the fact that
xt − 1
% log t as t ↓ 0, it follows that A#t B−A
t is monotone-decreasing as t ↓ 0,
t
so that another equivalent definition of Uhlmann’s type is the derivative one for the
path of geometric means A#t B:
A#t B − A
S(A|B) = s-lim
t ↓0 t
if the limit exists. If A and B are commuting and S(A|B) is defined, then
T ∗ S(A|B)T = S(T ∗ AT |T ∗ BT ).
76 J. I. Fujii and Y. Seo
But it is too strong for the existence of S(A|B). In fact, A is not majorized by A2 if
σ (A) = [0, 1], while we easily see S(A|A2 ) = A log A.
Another candidate is the kernel inclusion
ker A ⊃ ker B,
which is weaker than the range inclusion. In fact, the kernel condition does not
guarantee the existence: For B with σ (B) = [0, 1] where 0 is not an eigenvalue, it
follows that S(I |B) = log B diverges while both kernels are trivial.
The third condition between the above ones is B-absolute continuity in the sense
of Ando’s Lebesgue decomposition [3]:
A = [B]A ≡ s-lim A : nB
n→∞
where A : B defined by
is the parallel addition [2], which is the half of the harmonic mean A h B [4]. Kosaki
[34] showed that
1 1
[B]A = A 2 PM A 2
1
M = {y | A 2 y ∈ ranB}.
Relative Operator Entropy 77
This result implies A = [B]A = limt ↓ 0 A#t B and hence B-absolute continuity
guarantees the continuity of A#t B at t = 0 and it is a necessary condition for the
existence of S(A|B) as the above derivative. In fact, this continuity is in the norm
topology:
Lemma 3.4 If S(A|B) exists, then A#t B converges uniformly to A for t ↓ 0.
Proof Since there are scalars c1 and c2 with
1
B + (log α)A > ∃c ∈ R (∀α > 0)
α
1 1
(3) B-absolute continuity: A = [B]A = A 2 PM A 2 = lim A#t B .
t↓0
(4) kernel inclusion: ker A ⊃ ker B.
Remark 3.7 If both ranges of A and B are closed, in particular, for the case of
matrices, the above conditions in Theorem 3.6 are mutually equivalent since the
1
relation ranA 2 = ranA = (ker A)⊥ holds for all positive operators A.
4 CPR Geometry
The path of geometric means mt and the relative operator entropy are important
concepts in the geometric structure of the manifold of the positive invertible
operators discussed by Corach et al. [8, 9] which we call it CPR geometry. In this
78 J. I. Fujii and Y. Seo
section, we mention the general structure of the geometry of fiber bundles to read
easily for readers not familiar with it. Here the CPR geometry represents the one
on the Finsler manifold A + , the positive invertible elements in a unital C*-algebra
A , which we review in the below. Corach himself reformulated it in [8]: The base
manifold is A + with the tangent vector bundle A h (the selfadjoint operators in A ).
As we confirm later, the invertible elements G = G(A ) is the principal fiber bundle
(of A + ). Thus the total space P = {G, A + , U, π} is defined by
projection π: G → A + , g $→ gg ∗
structure group the unitary group U = U(A )
1
principal fiber π −1 (A) = A 2 U
This definition of this fiber homeomorphic to U is consistent. In fact, for each unitary
U , we obtain
π(A 2 U ) = A 2 U (A 2 U )∗ = A 2 U U ∗ A 2 = A.
1 1 1 1 1
G = {A 2 U |A ∈ A + , U ∈ U}.
1
G
g
π −1 (A)
π∗
π
+
∗
A = π(g) = gg
Vg
G
Γ
g Hg
π −1 (A)
γ
+
A = π(g)
we regard G as an upper structure of the base space A + , the kernel ker π∗ is the
vectors along π∗ ‘vertically’. So the subspace ker π∗ ∩Tg of the tangent space Tg (G)
at g ∈ G is called the vertical space Vg . In this case, since the Lie algebra of a unitary
group is the skew-hermitians, we have
Vg = {gX | X∗ = −X}.
In Tg (G), if a ‘horizontal space’ Hg is given (and it is compatible for the right action
U; HgU = Hg U ), then it is called the principal fiber G has a connection (in the
sense of manifold). In CPR geometry, it is naturally determined by Hg = gA h
considering the vertical one Vg . In fact, the compatibility of the right action is
shown: Take gH ∈ Hg with H = H ∗ . Then, for all U ∈ U, gH U = gU U ∗ H U
with U ∗ H U ∈ A h , which shows Hg U = HgU .
Now, for a (differentiable) curve γ (t) on A + , consider a horizontal lift (t) ∈
−1
π (γ (t)).
The term ‘lift’ means
˙
and the term ‘horizontal’ means the tangent vector (t) ∈ H(t ). In other words,
˙
(t)−1 (t) is hermitian, i.e.,
This is equivalent to ˙ ∗ =
˙ ∗ . Then we have
˙ ∗ + ˙ ∗ = 2
γ̇ = ˙ ∗
and hence
γ (t) = A #t B.
Proof The transport equation follows from:
˙ −1
= A 2 (log C)C 2 C − 2 A− 2 = A 2 (log C)A− 2
1 t t 1 1 1
2(t)(t)
1 1 1 1 1 1
γ̇ (t)γ (t)−1 = A 2 (log C)C t A 2 A− 2 C −t A− 2 = A 2 (log C)A− 2 .
In order to obtain further results in this geometry, we identify the tangent vector
bundle, which is A h , as the associated vector bundle for G. Note that G acts on A ∈
A + by A $→ gAg ∗ and hence X → gXg ∗ ≡ ρ(g)X. Regard A h as the tangent
vector space, consider the associated bundle: it is the quotient bundle G ×ρ A h of
G × A h with the equivalence relation of the right action by f ∈ G
Roughly speaking, at the point π(g), we see ρ(g −1 )X = g −1 X(g ∗ )−1 as a tangent
vector. Considering the connection of G, we reflects it on the tangent bundle via a
horizontal lift: If γ is a path on A + and is a horizontal lift of it, then we observe
−1 (t)γ̇ (t)((t)−1 )∗ instead of the tangent vector γ̇ (t), and thereby the translation
(t)(0)−1 : (0) $→ (t) yield the parallel displacement of a tangent vector X
along γ from 0 to t is
Note that
˙ ) = 1
( −1 ˙ −1 = γ̇ γ −1
2
Relative Operator Entropy 81
˙ ) = − −1
( −1 ˙ −1 .
So the covariant derivative Dt of a tangent field X(t) along the curve γ (t) in A +
is given by
O = Dt γ̇ = γ̈ − γ̇ γ −1 γ̇
d
log f (t) = f (t)f −1 (t).
dt
Proof For arbitrary ε > 0, there is δε > 0 with
f (t + δε ) − f (t)
− f (t)
δε < ε,
82 J. I. Fujii and Y. Seo
or equivalently
log f (t + δε ) − log f (t) log(f (t) + h) − log f (t) log(f (t) + h) − log f (t) h
≤ =
δε δε h δε
log(f (t) + h) − log f (t)
= (f (t) + ε) −→ f (t)−1 f (t)
h
we have f (0) = I and f also satisfies the geodesic equation (Indeed, it is a curve
from I to A− 2 BA− 2 ). By f f −1 = (f f −1 )2 , we have
1 1
(f f −1 ) = f f −1 + f (f −1 ) = f f −1 − (f f −1 )2 = 0
The parallel transport of the tangent vector X at γ (t1 ) to that at γ (t2 ) along the
geodesic γ is described as Ptt12 X. Now by Lemma 4.1, we can obtain the parallel
transform along the geodesic:
Relative Operator Entropy 83
P01 X = A 2 C 2 A− 2 XA− 2 C 2 A 2
1 1 1 1 1 1
1 1
for C = A− 2 BA− 2 .
1 1 1
In particular, the case B = I , we have A− 2 XA− 2 = ρ −1 (A 2 )X. This
manifold is a symmetric space and then a homogeneous space; A + = G/U. Thus
properties around the identity I reflect on those around other points. Moreover,
every symmetric space is geodesic complete, that is, the domain [0, 1] is extended
to R:
for t ∈ R.
From this viewpoint, the definition
1 1
L(X; A) = XA = A− 2 XA− 2
at each point A ∈ A makes the above manifold A + a Finsler space with a Finsler
metric as in the CPR main result. Since XA is equivalent to the operator norm
X, it is a Finsler metric if
Finsler condition: L(Pt X; γ (t)) = Pt Xγ (t ) = Xγ (0) = L(X; γ (0))
holds for all curves γ and parallel transports Pt along γ . The CPR geometry does not
always determine a unique Finsler metric. In fact, we show each unitarily invariant
norm ||| ||| also gives a Finsler metric for the CPR geometry:
Theorem 4.5 For a unitarily invariant norm ||| ||| on A , a function
1 1
L||| ||| (X; A) = |||X|||A = |||A− 2 XA− 2 |||
condition is satisfied by
1 1
= |||γ (0)− 2 Xγ (0)− 2 ||| = |||X|||γ (0).
84 J. I. Fujii and Y. Seo
A Finsler metric L||| ||| (X; A), which is called a unitarily invariant Finsler one, is
homogeneous like L(X; A):
Theorem 4.6 For any invertible operator Y ,
= ||| (Y ∗ AY )− 2 Y ∗ XY (Y ∗ AY )−1 Y ∗ XY (Y ∗ AY )− 2 |||
1 1
1 1
= ||| (Y ∗ AY )− 2 Y ∗ XA−1 XY (Y ∗ AY )− 2 |||
= ||| A− 2 XY (Y ∗ AY )−1 Y ∗ XA− 2 |||
1 1
1 1
= ||| A− 2 XA−1 XA− 2 ||| = |||A− 2 XA− 2 |||
1 1
Thus this geometric consideration says that the Karcher equation should be
w(n)S(An |X) = O
n
as a barycenter of the terminal tangent vectors, which is the geometric meaning for
the Karcher equation, cf. [37, 38]. It is also consistent considering the power mean
equation
X = w(n) (An #t X) ,
or
An #t X − X
w(n) = O.
n
t
Relative Operator Entropy 85
S(A) = Tr [η(A)],
where the entropy function η(t) = −t log t. As for the Shannon entropy, it is
extremely useful to define a quantum version of the relative entropy. Suppose that A
and B are density matrices in Sn . As a quantum generalization of the relative entropy
due to Kullback and Leibler [36], Umegaki firstly introduced in the setting of von
Neumann algebra [45] in 1962 the quantum relative entropy of A with respect to B,
which is defined by
Tr [A(log A − log B)] if supp A ⊂ supp B,
SU (A|B) = (1)
+∞ otherwise.
We call (1) the Umegaki relative entropy. In [1], for any A and B in Sn , the Tsallis
relative entropy of A to B is defined by
1 − Tr [A1−α B α ]
Dα (A|B) = = Tr [A1−α (lnα A − lnα B)] (2)
α
for any 0 < α ≤ 1, where lnα t = t α−1 is the α-logarithmic function. The Tsallis
α
relative entropy (2) is a 1-parameter extension of (1), and Ruskai and Stillinger
in [41] showed the following relation between the Tsallis relative entropy and the
Umegaki relative entropy:
The quantum quantity Tr [A(log A1/2 B −1 A1/2)] is firstly proposed by Belavkin and
Staszewski [6] in the framework of C∗ -algebra. Since we treat SF K (A|B) as the
minus of the trace of the relative operator entropy S(A|B), we call (3) the FK
relative entropy, or the BS relative entropy in [39, pp125]. If A and B commute,
then we have SU (A|B) = SF K (A|B). Generally, two quantum formulations of
the relative entropy are different. In fact, Hiai and Petz [30] showed the following
relation:
1
SU (A|B) ≤ − Tr [A1−q S(Aq |B q )] for all q > 0. (5)
q
In fact, if q = 1 in (5), then we have the Hiai-Petz inequality (4) and as q → 0 the
right-hand side of (5) converges to the Umegaki relative entropy SU (A|B).
Moreover, Yanagi et al. [47] have been advancing research on the Tsallis relative
operator entropy as an operator generalization of the Tsallis relative entropy, which
is regarded as a 1-parameter extension of the relative operator entropy: For positive
definite matrices A and B in Pn , the Tsallis relative operator entropy is defined by
A α B − A
Tα (A|B) = for 0 < α ≤ 1.
α
We recall the notation α for the binary operation
that have formula in common with α . Then the Tsallis relative entropy of negative
order is defined by
A α B−A
Tα (A|B) = for α ∈ [−1, 0).
α
For convenience, we denote another quantum Tsallis relative entropy of order α ∈
[−1, 1]\{0} by
NTα (A|B) = −Tr Tα (A|B)
for positive definite matrices A and B in Pn . For two relative entropies, we know
the following two relations:
Aα B
Aα B − A
Tα (A|B) =
α −NTα (A|B)
We have the following properties of the quantum relative entropy NTα of order
α ∈ [−1, 1]\{0}, also see [21]:
Theorem 5.1 Let A and B be positive definite matrices in Pn and α ∈ [−1, 1]\{0}.
Then the following properties of the quantum Tsallis relative entropy NTα hold:
(1) (Non-negativity) NTα (A|B) ≥ 0 if A ≥ B.
(2) (Pseudoadditivity)
NTα (A1 ⊗ A2 |B1 ⊗ B2 ) = NTα (A1 |B1 ) + NTα (A2 |B2 ) + αNTα (A1 |B1 )NTα (A2 |B2 ).
(3) (Joint convexity) NTα ( j λj Aj | j λj Bj ) ≤ j λj NTα (Aj |Bj ).
(4) (Monotonicity) For any trace-preserving positive linear map
Proof For (1), if α ∈ [−1, 0), then it follows that (1 − α)A + αB ≤ A α B and so
A α B −A (1 − α)A + αB − A
NTα (A|B) = −Tr ≥ −Tr
α α
= −Tr [B − A] ≥ 0.
Tα (A1 ⊗ A2 |B1 ⊗ B2 )
1
= [(A1 ⊗ A2 ) α (B1 ⊗ B2 ) − A1 ⊗ A2 ]
α
1 1 1
= Tα (A1 |B1 ) ⊗ (A2 α B2 ) + (A1 α B1 ) ⊗ Tα (A2 |B2 ) + A1 ⊗ Tα (A2 |B2 )
2 2 2
1
+ Tα (A1 |B1 ) ⊗ A2
2
1 1
= Tα (A1 |B1 ) ⊗ (A2 α B2 ) − Tα (A1 |B1 ) ⊗ A2 + Tα (A1 |B1 ) ⊗ A2
2 2
1 1
+ (A1 α B1 ) ⊗ Tα (A2 |B2 ) − A1 ⊗ Tα (A2 |B2 ) + A1 ⊗ Tα (A2 |B2 )
2 2
= αTα (A1 |B1 ) ⊗ Tα (A2 |B2 ) + Tα (A1 |B1 ) ⊗ A2 + A1 ⊗ Tα (A2 |B2 ).
( λj Aj ) α λj Bj ) − λj Aj
(
NTα ( λj Aj | λj Bj ) = −Tr
α
j j
λj (Aj α Bj ) − λj Aj
≤ −Tr
α
⎡ ⎤
= −Tr ⎣ λj Tα (Aj |Bj )⎦ = λj NTα (Aj |Bj )
j j
A α B −A
= −Tr
α
= NTα (A|B)
for all p ≥ q > 0 and 0 < q ≤ 1, and any unitarily invariant norm |||·|||.
Proof By the antisymmetric tensor technique, in order to prove (9), it suffices to
show that
q
p p 1−q 1−q
λ1 (A 2 A− 2 B p A− 2
1 p 1
A 2 ) ≤ λ1 (A 2 Bq A 2 ) (10)
1 1
or equivalently, replacing A and B by A q−1 and B q respectively
p p p−1
B ≤A #⇒ A q−1 q B q ≤ A q−1
p
for all 0 < q ≤ p and 0 < q ≤ 1, because both sides of (10) have the same order of
homogeneity for A, B, so that we can multiply A, B by a positive constant.
90 J. I. Fujii and Y. Seo
p +r−p r
Put r = 1−qp
> 0 and p = p
q ≥ 1. Then p−1
q−1 = p . It follows from the
Furuta inequality (8) that
p p
A q−1 q B q = A−r 1 B p
p p
= A−r p +r A−r 1+r B p by the multiplicity of α
p (1+r) p +r
p +r−p r p−1
=A p = A q−1
1
Dα (A|B) ≤ − Tr [A1−q T αq (Aq |B q )]
q
≤ −Tr
α
" q q q α q #
A1−q A 2 (A− 2 B q A− 2 ) q A 2 − Aq
= −Tr α
q q
1−q
A
= −Tr T αq (Aq |B q )
q
A1−α B α − A
Dα (A|B) = −Tr [ ]
α
A1/2(A−q/2 B q A−q/2 )α/q A1/2 − A
≤ −Tr [ ]
α
1−q
A
= −Tr T αq (A |B )
q q
q
6 Concluding Remarks
There are many related topics on the relative operator entropy and the Tsallis relative
operator entropy, see [32, 40] and [11]. Among others, we present generalizations
of the relative operator entropy and the Tsallis relative operator entropy for positive
invertible operators on a Hilbert space due to Isa et al. [31]. We treat Tα (A|B) in
A x B −A
which the range of α is extended from [−1, 1] to R, that is, Tx (A|B) ≡
x
(x ∈ R\{0}) and T0 (A|B) ≡ lim Tx (A|B) = S(A|B).
x→0
We regard the Tsallis relative operator entropy Tx (A|B) as the average rate of
change of the path A t B over the interval [0, x] and relative operator entropy
S(A|B) as the rate of change of the path at t = 0. Based on this viewpoint, we define
the n-th Tsallis relative operator entropy and the n-th relative operator entropy.
We begin by defining the first relative operator entropy S[1] (A|B) and the first
Tsallis relative operator entropy Tx[1] (A|B) as S(A|B) and Tx (A|B), respectively,
that is,
1 1 1 1
S[1] (A|B) ≡ S(A|B) = A 2 (log A− 2 BA− 2 )A 2 and
1 1
(A− 2 BA− 2 )x − I 1
1
Tx[1] (A|B) ≡ Tx (A|B) = A 2 A 2 (x ∈ R\{0}).
x
92 J. I. Fujii and Y. Seo
ax − 1
The corresponding functions to S[1] (A|B) and Tx[1] (A|B) are log a and
x
ax − 1
for a > 0, respectively. Since lim = log a, it follows that T0[1] (A|B) ≡
x→0 x
lim Tx[1] (A|B) = S[1] (A|B). Next, we define the second Tsallis relative operator
x→0
entropy Tx[2] (A|B) as the average rate of change of Tx[1] (A|B) over the interval [0, x],
that is,
a x − 1 − x log a a x − 1 − x log a
Since its corresponding function is and lim =
x2 x→0 x2
1
(log a)2 , we define the second relative operator entropy as
2
1 1 1 1 1 1
S[2] (A|B) ≡ A 2 (log A− 2 BA− 2 )2 A 2 = S(A|B)A−1 S(A|B),
2 2
1 1 1
S[n] (A|B) ≡ A 2 (log A− 2 BA− 2 )n A 2 = A(A−1 S(A|B))n
1 1 1
n! n!
and the n-th Tsallis relative operator entropy Tx[n] (A|B) are inductively defined by
and for n ≥ 2,
dn x
Since a = a x (log a)n holds for a > 0, we have
dx n
dn 1 1 1 1 1 1
A x B = A 2 (A− 2 BA− 2 )x (log A− 2 BA− 2 )n A 2 = (A x B)(A−1 S(A|B))n
dx n
1 dn
and so S[n] (A|B) = A x B .
n! dx n x=0
Then we have
n−1
A x B =A+ x S (A|B) + x n Tx[n] (A|B),
k [k]
k=1
which is the Taylor’s expansion of A x B around 0. We remark that the k-th relative
operator entropy S[k] (A|B) appears as the coefficient of the x k -term and the n-th
Tsallis relative operator entropy Tx[n] (A|B) appears in the residual term. So we can
call Tx[n] (A|B) the n-th residual relative operator entropy.
There are deep discussion on the n-th relative operator entropy and the n-th
Tsallis relative operator entropy in [31] and [44].
Acknowledgments The authors are partially supported by JSPS KAKENHI Grant Number
JP19K03542.
References
11. S.S. Dragomir, A survey of recent inequalities for relative operator entropy, in Operations
Research, Engineering, and Cyber Security, Springer Optim. Appl., vol. 113 (Springer, Cham,
2017), pp. 199–229
12. J.I. Fujii, Izumino’s view of operator means. Math. Japon. 33, 671–675 (1988)
13. J.I. Fujii, Operator means and the relative operator entropy, in Operator Theory and Complex
Analysis (Sapporo, 1991), Oper. Theory Adv. Appl., vol. 59 (Birkhäuser, Basel, 1992), pp.
161–172
14. J.I. Fujii, Structure of Hiai-Petz parametrized geometry for positive definite matrices. Linear
Algebra Appl. 432 , 318–326 (2010)
15. J.I. Fujii, Path of quasi-means as a geodesic. Linear Algebra Appl. 434, 542–558 (2011)
16. J.I. Fujii, Interpolationality for symmetric operator means. Sci. Math. Japon. 75, 267–274
(2012)
17. J.I. Fujii, M. Fujii, Y. Seo, An extension of the Kubo-Ando theory: Solidarities. Math. Japon.
35, 509–512 (1990)
18. J.I. Fujii, E. Kamei, Relative operator entropy in noncommutative information theory. Math.
Japon. 34, 341–348 (1989)
19. J.I. Fujii, E. Kamei, Uhlmann’s interpolational method for operator means. Math. Japon. 34,
541–547 (1989)
20. J.I. Fujii, E. Kamei, Interpolational paths and their derivatives. Math. Japon. 36, 557–560
(1994)
21. J.I. Fujii, Y. Seo, Tsallis relative operator entropy with negative parameters. Adv. Oper. Theory
1(2), 219–236 (2016)
22. J.I. Fujii, Y. Seo, The relative operator entropy and the Karcher mean. Linear Algebra Appl.
542, 4–34 (2018)
23. M. Fujii, Y. Seo, Matrix trace inequalities related to the Tsallis relative entropies of real order.
J. Math. Anal. Appl. 498, Article 124877 (2021)
24. M. Fujii, J. Mićić Hot, J. Pečarić, Y. Seo, Recent Developments of Mond-Pečarić Method in
Operator Inequalities (Element, Zagreb, 2012)
25. S. Furuichi, Matrix trace inequalities on the Tsallis entropies. J. Inequal. Pure Appl. Math. 9,
Article 1, 7pp. (2008)
26. S. Furuichi, K. Yanagi, K. Kuriyama, Fundamental properties of Tsallis relative entropy. J.
Math. Phys. 45, 4868–4877 (2004)
27. S. Furuichi, K. Yanagi, K. Kuriyama, A note on operator inequalities of Tsallis relative opeartor
entropy. Linear Algebra Appl. 407, 19–31 (2005)
28. F. Hansen, An operator inequality. Math. Ann. 246, 249–250 (1980)
29. F. Hansen, G.K. Pedersen, Jensen’s inequality for operators and Löwner’s theorem. Math. Ann.
258, 229–241 (1982)
30. F. Hiai, D. Petz, The Golden-Thompson trace inequality is complemented. Linear Algebra
Appl. 181, 153–185 (1993)
31. H. Isa, E. Kamei, H. Tohyama, M. Watanabe, The n-th relative operator entropies and the n-th
operator divergences. Ann. Funct. Anal. 11(2), 298–313 (2020)
32. S. Kim, H. Lee, Relative operator entropy related with the spectral geometric mean. Anal.
Math. Phys. 5(3), 233–240 (2015)
33. S. Kobayashi, K. Nomizu, Foundations of Differential Geometry (John Wiley & Sons, New
York, vol. 1 (1963), vol. 2 (1969))
34. H. Kosaki, Remarks on Lebesgue-type decomposition of positive operators. J. Oper. Theory
11, 137–143 (1984)
35. F. Kubo, T. Ando, Means of positive linear operators. Math. Ann. 246, 205–224 (1980)
36. S. Kullback, R. Leibler, On information and sufficiency. Ann. Math. Stat. 22, 79–86 (1951)
37. J. Lawson, Y. Lim, Karcher means and Karcher equations of positive definite operators. Trans.
Am. Math. Soc., Ser. B 1, 1–22 (2014)
38. Y. Lim, M. Pálfia, Matrix power means and the Karcher mean. J. Funct. Anal. 262, 1498–1514
(2012)
Relative Operator Entropy 95
39. M. Ohya, D. Petz, Quantum Entropy and Its Use (Springer, Berlin Heidelberg, 2004), Corrected
Second Printing
40. M. Raïssouli, M.S. Moslehian, S. Furuichi, Relative entropy and Tsallis entropy of two
accretive operators. C. R. Math. Acad. Sci. Paris 355(6), 687–693 (2017)
41. M.B. Ruskai, F.H. Stillinger, Convexity inequalities for estimating free energy and relative
entropy. J. Phys. A 23, 2421–2437 (1990)
42. Y. Seo, Matrix trace inequalities on Tsallis relative entropy of negative order. J. Math. Anal.
Appl. 472, 1499–1508 (2019)
43. C.E. Shannon, A mathematical theory of communication. Bell Syst. Tech. J. 27, 623–656
(1948)
[n]
44. H. Tohyama, E. Kamei, M. Watanabe, The n-th residual relative operator entropy Rx,y (A|B).
Adv. Oper. Theory 6(1), No. 18, 11 pp. (2021)
45. H. Umegaki, Conditional expectation in an operator algebra, IV (Entropy and information),
Kodai Math. Sem. Rep. 14, 59–85 (1962)
46. J. von Neumann, Mathematical Foundations of Quantum Mechanics (Princeton University
Press, Princeton, NJ, 1955). (Originally appeared in German in 1932)
47. K. Yanagi, K. Kuriyama, S. Furuichi, Generalized Shannon inequalities based on Tsallis
relative operator entropy. Linear Algebra Appl. 394, 109–118 (2005)
Matrix Inequalities and
Characterizations of Operator
Monotone Functions
1 Introduction
Throughout this chapter, M n stands for the algebra of n × n matrices over C and Pn
denotes the cone of positive definite elements in M n . Denote by In the identity
matrix of M n . For a Hermitian matrix A with eigenvalues in the domain of a
function f , the matrix f (A) is defined by means of the functional calculus.
Definition 1.1 A continuous function f on I (⊂ R) is called n-monotone, if
T. H. Dinh ()
Department of Mathematics, Troy University, Troy, AL, USA
e-mail: [email protected]
H. Osaka
Department of Mathematical Sciences, Ritsumeikan University, Kusatsu, Shiga, Japan
e-mail: [email protected]
O. E. Tikhonov
Institute of Computer Mathematics and Information Technologies, Kazan Federal University,
Kazan, Russia
e-mail: [email protected]
for any pair of self-adjoint matrices A, B ∈ M n with σ (A), σ (B) ⊂ I , where σ (A)
stands for the spectrum of A.
Definition 1.2 A continuous function f on I (⊂ R) is called n-convex if the
inequality
holds for all self-adjoint matrices A, B ∈ M n with σ (A), σ (B) ⊂ I and for all
λ ∈ [0, 1]. Also, f is called a n-concave on I if (−f ) is n-convex on I .
Let f : I (⊂ R) → R. We call a function f operator convex if f is n-convex for
any n ∈ N, and operator monotone if f is n-monotone for any n ∈ N.
Operator monotone functions were firstly introduced and studied by Löwner in
1934 [35]. He has completely characterized all operator monotone functions as the
class of Pick functions. These functions play an essential role in the theory of
analytic functions. He has also proved that if f is operator monotone on [0, ∞),
then there exists a positive measure μ on [0, ∞) such that
$ ∞ st
f (t) = a + bt + dμ(s),
0 t +s
where a is a real number and b ≥ 0. According to this representation, for r ∈ [0, 1],
the function f (t) = t r is operator monotone on [0, ∞). This is the well-known
Löwner-Heinz’s inequality [42, Theorem 1.1] for positive semidefinite matrices
which states that for any 0 ≤ A ≤ B and r ∈ [0, 1], Ar ≤ B r .
In [31], Hansen and Perdesen considered some basic equivalent assertions for
operator monotone functions and operator convex functions. They showed that for
a strictly positive, continuous function f on (0, ∞), the following statements are
equivalent:
(a) f is operator concave.
t
(b) is operator monotone.
f (t)
In [37], Osaka and Tomiyama discussed some similar assertions at each level n
in order to see clearly the inside of the double piling structure of matrix monotone
functions and of matrix convex functions. More precisely, the main results in [37]
are in the following.
Characterizations of Operator Monotone Functions 99
f (C AC) ≤ C f (A)C.
f (t )
(iii) The function g(t) = t is n-monotone on (0, α).
Then
where notation (A)m ⇒ (B)n means that “if (A) holds for the matrix algebra M m ,
then (B) holds for the matrix algebra M n ”.
In 1980, Kubo and Ando [34] introduced the theory of operator means. Let
B(H)+ be the set of positive invertible operators in a Hilbert space H. A binary
operation σ : B(H)+ × B(H)+ → B(H)+ , (A, B) $→ Aσ B, is called a connection
if the following requirements are fulfilled:
(I) If A ≤ C and B ≤ D, then Aσ B ≤ Cσ D;
(II) C ∗ (Aσ B)C ≤ (C ∗ AC)σ (C ∗ BC);
(III) If An % A and Bn % B, then An σ Bn % Aσ B.
Further, a mean is a connection satisfying the normalized condition:
(IV) 1σ 1 = 1.
Kubo and Ando showed that there exists an affine order-isomorphism from the
class of connections to the class of positive operator monotone functions, which is
given by σ $→ fσ (t) = 1σ t. Let f be an operator monotone function, then for
positive definite matrices A and B,
In 1996, Petz [38] introduced the theory of monotone metric in quantum infor-
mation theory which was based on operator monotone functions. Therefore, such
functions are important in matrix analysis, quantum information and other areas as
well. The authors refer readers to the books of William Donoghue [29], Barry Simon
[41] and Bhatia [5] for more details about operator monotone functions.
It is well-known that if σ is a symmetric matrix Kubo-Ando mean, i.e., Aσ B =
Bσ A, then the representing function fσ satisfies the following inequalities
2x 1+x
≤ fσ (x) ≤ .
1+x 2
100 T. H. Dinh et al.
where A!B = 2(A−1 + B −1 )−1 is the harmonic mean of A and B, and A∇B =
(A + B)/2 is the arithmetic mean of A and B. Obviously, if f : [0, ∞) → [0, ∞)
is operator monotone, we have
For any normal state φ on B(H), a positive kernel operator Sφ with Tr (Sφ ) = 1
such that φ(A) = Tr (Sφ A) (A ∈ B(H)) is uniquely defined. If f is an operator
Characterizations of Operator Monotone Functions 101
monotone function, and φ is a positive linear functional on B(H), then for any 0 ≤
A ≤ B,
x s y 1−s + x 1−s y s
Gs (x, y) =
2
the Heinz means and by
x+y
Hs (x, y) = s + (1 − s)x 1/2 y 1/2
2
the Heron means.
The family of Heron means and Heinz means are clearly interpolations between
the arithmetic and the geometric means. In [6], Bhatia obtained a relation between
the Heinz mean and the Heron mean which states that for t ∈ [0, 1],
where α(s) = 2s − 1.
Proof The implication follows from (7) and monotonicity, so we only need to show
the converse. Given two positive numbers a ≤ b, it suffices to show that there exist
positive numbers x and y such that
a x s y 1−s + x 1−s y s
= √
b α(s) (x + y) + 2(1 − α(s)2 ) xy
2
(y/x)α(s)/2 + (y/x)−α(s)/2
=
α(s)2 ((y/x)1/2 + (y/x)−1/2) + 2(1 − α(s)2 )
cosh(α(s)c)
= ,
α(s)2 cosh(c) + (1 − α(s)2 )
cosh(αc)
fα (c) =
α 2 cosh(c) + (1 − α2 )
Continuity and the Intermediate Value Theorem imply that the function fα :
[0, ∞) → (0, 1] is surjective. Moreover, we can show that the function fα :
[0, ∞) → (0, 1] is also injective. To do this, it is enough to show that the function
is monotonic on [0, ∞). So, note that
d
fα (c) ≤ 0
dc
Characterizations of Operator Monotone Functions 103
d
gα (c) = 2α(−1 + α 2 ) cosh(cα) sinh(c/2)2
dc
which is clearly non-positive when c ≥ 0. Hence, the function fα : [0, ∞) → (0, 1]
is bijective. To obtain a solution for (9), fix s ∈ (0, 1/2) ∪ (1/2, 1) and set c =
−1
fα(s) (a/b). With this, we can obtain the desired x and y satisfying (9).
Let 0 ≤ p ≤ 1 ≤ q. It is well-known that for non-negative numbers a and b,
√ a p + bp 1/p
a+b a q + bq 1/q
ab ≤ ≤ ≤ ,
2 2 2
or,
√
ab ≤ μ(p, a, b) ≤ μ(1, a, b) ≤ μ(q, a, b), (10)
1/p
where μ(p, a, b) = a +b
p p
2 is the power mean (or, binomial mean). Using
similar arguments as in the proof of Theorem 2.1 one can obtain the following
theorem.
Theorem 2.2 Let f be a continuous function on [0, ∞). For 0 ≤ p ≤ 1 ≤ q,
suppose that one of the following inequalities holds for all any non-negative
numbers a ≤ b:
√
(1) f (a) ≤ f ( ab);
(2) f (μ(1, a, b)) ≤ f (b),
a s b1−s + a 1−s bs a+b √
(3) f ≤ f |2s − 1| + (1 − |2s − 1|) ab .
√ 2 2
(4) f ( ab) ≤ f (μ(p, a, b));
(5) f (μ(p, a, b)) ≤ f (μ(1, a, b)) ;
(6) f (μ(1, a, b)) ≤ f (μ(q, a, b)) .
Then the function f is increasingly monotone on [0, ∞).
104 T. H. Dinh et al.
Notice that from (7) we have the following inequalities for matrix means:
As B + A1−s B
f (AB) ≤ f ; (11)
2
A+B A+B
f α(s)2 + (1 − α(s)2 )AB ≤f .
2 2
Proof It is obvious that (i) implies (ii), (iii), and (iv). Let us show that (iii) implies
(i) first and then we show (ii) implies (i). That would complete the proof since
(iv) implies (i) follows from [1, Proposition 4.1] since the matrix Heron mean is
symmetric.
Suppose (12) holds for any positive definite matrices A and B. We need to show
that for any 0 < X ≤ Y ,
f (X) ≤ f (Y ).
Firstly, let us consider the case when Y = In . We now show that there exist positive
definite matrices A0 , B0 such that
1−s
1/2 C0 + C0
s
A0 s B0 + A0 1−s B0 1/2
= A0 A0 = X (13)
2 2
Characterizations of Operator Monotone Functions 105
and
−1/2 −1/2
where C0 = A0 B0 A0 . From (14), we get
−1/2
1/2 In + C0 1/2
A0 = α(s)2 + (1 − α(s)2 )C0 .
2
√ −1
x s + x 1−s 1+x
f (x) = α(s)2 + (1 − α(s)2 ) x
2 2
is bijective and takes values in (0, 1]. Therefore, for any 0 < X ≤ In there exists
a unique matrix C0 satisfying (15). Hence, the matrix A0 is obtained from (14) and
1/2 1/2
the matrix B0 equals A0 C0 A0 .
In general, for 0 < X ≤ Y we have 0 < Y −1/2 XY −1/2 ≤ In . By the above
arguments, we can find A0 , B0 ∈ M+ n such that
A0 s B0 + A0 1−s B0
= Y −1/2 XY −1/2
2
and
A0 + B0
α(s)2 + (1 − α(s)2 )A0 B0 = In .
2
Remark 2.4 There are numerous papers on the matrix Heron mean and the matrix
Heinz mean and related questions. We refer the readers to [8, 16, 20, 22, 24–26] and
the references therein.
or
Definition 2.5 ([4]) Let f : R+ → R+ , where R+ = (0, ∞). We say that f ∈ Fop
if it satisfies the following conditions:
1. f is operator monotone,
2. tf (t −1 ) = f (t) for all t ∈ R+ , and
3. f (1) = 1.
Notice, that functions in f ∈ Fop are in one-to-one correspondence with
symmetric means.
Definition 2.6 For f, g ∈ Fop , define
t + 1 f (t)
ψ(t) = , t > 0.
2 g(t)
in these particular cases, both of which are operator monotone. It is shown in [4] that
Fop forms a lattice under 0. It is worth noting that this order is stronger than the
regular point-wise order ≤. That is, if f 0 g then f ≤ g because ψ(t) ≤ 1+t 2 (t ∈
R+ ).
Characterizations of Operator Monotone Functions 107
It is shown in [4, Proposition 2.1] that f ∈ Fop implies that f has an integral
representation of the form
1 + t H (t )
f (t) = e (16)
2
where
$ 1 (λ2 − 1)(1 − t)2
H (t) = h(λ) dλ
0 (t + λ)(1 + tλ)(λ + 1)2
ϕ(t) = t −1 f (t 2 ).
Then,
√
1. If · ≺ f then, as a real function, ϕ is monotonically decreasing on (0, 1) and
monotonically
√ increasing on (1, ∞).
2. If · f then ϕ is monotonically increasing on (0, 1) and monotonically
decreasing on (1, ∞).
Proof Consider the derivative
ϕ (t) = −t −2 f (t 2 ) + 2f (t 2 ).
depending on the interval and the order relationship considered. Based on (16) we
consider
if and only if
1−t
H (t) ≶ , respectively.
2t (1 + t)
108 T. H. Dinh et al.
An easy calculation shows that when h(λ) is substituted by the constant function
1/2, the integral becomes
$
1 1 (1 − λ2 )(1 − t 2 ) 1−t
dλ = .
2 0 (t + λ) (1 + tλ)
2 2 2t (1 + t)
So now√we apply [4, Theorem 2.4] to determine the monotonicity of ϕ in each case.
So, let · ≺ f and t ∈ (0, 1). In this case h(λ) ≤ 1/2 and the integrand,
(1 − λ2 )(1 − t 2 )
h(λ) ≥ 0
(t + λ)2 (1 + tλ)2
1−t
H (t) ≤ ,
2t (1 + t)
X = A#B and Y = Aσ B.
Proof Note that if we show that for In ≤ X−1/2 Y X−1/2 := Y0 ≤ γ In we can find
positive operators A0 and B0 such that:
In = A0 #B0 and Y0 = A0 σ B0 ,
Y0 = A0 σ A−1
0 .
for any positive operators√A and B, then the function g is operator monotone on
R+ . If, on the other hand, · f and
Since γ0i X ≤ γ0i+1 X ≤ γ γ0i X, Lemma 2.9 implies that there exist positive operators
Ai and Bi such that:
and so
where h : [−1,
√ 0] → [0, 1] is a measurable function. We notice that if h(λ) = 1/2,
then f (t) = t.
Definition 2.12 For f, g ∈ E, we say f 1sa g if and only if fg −1 is operator
monotone.
In the following, we show that this so defined relation satisfies the same
properties as the order defined in [4] on Fop that we introduced earlier in this section.
Characterizations of Operator Monotone Functions 111
and
So, clearly f 1sa g if and only if hf ≥ hg a.e. Therefore, 1sa defines an order
relation on E. Moreover, it is easy to see that for f ∈ E implies that 1 0sa f (t) 0sa
t. Indeed, 1 0sa f (t) follows from the monotonicity of f and f (t) 0sa t follows
from the monotonicity of f (tt ) .
We can also define the meet and join of any two elements in a similar fashion as
in [4]. Let f, g ∈ E then define:
$ 0 1 t
f ∧ g = exp + min{hf (λ), hg (λ)} dλ,
−1 λ−t 1 − λt
$ 0 1 t
f ∨ g = exp + max{hf (λ), hg (λ)} dλ.
−1 λ−t 1 − λt
As an immediate result we obtain that E with 0sa is a lattice. That is, for any two
f, g ∈ E,
f ∧ g 0sa f 0sa f ∨ g.
t
f (t) → f † (t) =
f (t)
1 f (t)
f † (t −1 ) = = = (f † (t))−1 ,
tf (t −1 ) t
and
f 0sa g #⇒ g † 0sa f † .
112 T. H. Dinh et al.
ϕ(t) = t −1 f (t 2 ).
Then,
√
1. If √· 0sa f then, as a real function, ϕ is monotonically increasing on R+ .
2. If · 1sa f then ϕ is monotonically decreasing on R+ .
Proof As before, to show the monotonicity of ϕ as a real function, it suffices to
show
depending on the interval and the order relationship considered. With the integral
expression of f , (19) becomes:
$ 0 1 1
2t + h(λ) dλ ≶ 1.
−1 (λ − t) 2 (1 − λt)2
Using the same arguments as in Lemma 2.9 and Theorem 2.10 we can show the
following result.
Theorem 2.14 Let σ √be some self-adjoint operator mean on R+ with representing
function f such that · ≺sa f and let A and B be positive operators such that
A < B. Then, if
There is yet another class of means to consider. Let τ and τ ⊥ be the means
represented by operator monotone functions g and g † , respectively. Kubo and Ando
showed in [34, Theorem 5.4] that if an operator mean σ with representing function
f satisfies
In this section, we prove Theorem 2.10 for general symmetric Kubo-Ando means.
We used monotonicity of the function ϕ on certain intervals to obtain bijectivity,
thus obtaining a well-defined ϕ −1 when restricted to the appropriate intervals. With
this function, we were able to solve the problem in Lemma 2.9, which then allowed
us to obtain the desired characterization. With a little care, it is possible to obtain
the same result when ϕ is only surjective on the prescribed intervals.
114 T. H. Dinh et al.
√
We recall some of our notation from Sect. 2.2. Suppose that · ≶ f , respectively,
as before define ϕ(t) = t −1 f (t 2 ). Then, we have
With this we can show a lemma which is stronger than Lemma 2.9.
Lemma 2.16 Let σ √ be some symmetric
√ operator mean on R+ with representing
function f such that · < f (resp. · > f ) and let γ = limt →∞ f (t 2 )/t. Then, if
X and Y are positive definite matrices such that X ≤ Y < γ X (resp. γ X < Y ≤ X),
then there exist positive matrices A and B such that
X = A#B and Y = Aσ B.
√
Proof As in the proof of Lemma 2.9, we show the lemma when · < f . In this
case, it suffices to show that given In ≤ Y0 = U diag({λi (Y0 )}) U ∗ ≤ γ In , we can
find A0 ≥ 0 such that
Y0 = A0 σ A−1
0 = ϕ(A0 ).
While ϕ(t) is not necessarily bijective in this case, it is continuous on [1, ∞) and
ϕ(1) = f (1) = 1. Therefore, the restriction of ϕ to some subset of [1, ∞) is
surjective onto [1, γ ).
Since σ (Y0 ) ⊂ [1, γ ), surjectivity of the restriction of ϕ implies that the set
satisfies
Now, let p be a real number, and a, b be positive. According to the relation (1) the
power mean μ(p, a, b) is corresponding to the monotone function f as
1/p
1 + tp
fμ,p (t) = .
2
Then for positive definite matrices A and B, the Kubo-Ando matrix power mean is
defined as
1/p
1 + (A−1/2BA−1/2 )p
Pμ (p, A, B) = A1/2 A1/2 .
2
Therefore, from the chain of inequalities (10) for an operator monotone function f
on (0, ∞) we have
1/p 1/q
Ap + B p A+B Aq + B q
≤ ≤ (26)
2 2 2
whenever A and B are positive semidefinite. It is worth noting that the inequalities
in (26) were discussed by Audenaert and Hiai [2], where they obtained conditions
on p and q such that (26) holds true. In [10, 18] the authors also studied new types
of operator convex functions and related inequalities.
116 T. H. Dinh et al.
for any positive definite matrices A and B. Then f is operator monotone on (0, ∞).
Proof Suppose that the inequality (27) holds, it suffices to show that for any 0 ≤
X ≤ Y , there exist two positive semidefinite matrices A and B such that
Firstly, let us consider the case when X = In . We now show that there exist positive
definite matrices A0 and B0 such that A0 B0 = In and
−1/2
B0 A−1/2 A0 = A0 fμ,p A−2
1/2 1/2
Y = A0 fμ,p A0 0 . (28)
x p + x −p
1/p
Since the function h(x) = xfμ,p (x −2 ) = is surjective from
2
(0, ∞) to [1, ∞), we obtain that for any In ≤ Y there exists a matrix A0 > 0
satisfying (28). The matrix B0 is equal to A−1
0 .
In general, for 0 < X ≤ Y we have In ≤ X−1/2 Y X−1/2 . By the above
arguments, we can find positive semidefinite matrices A0 and B0 such that A0 B0 =
In and
−1/2
X−1/2 Y X−1/2 = A0 fμ,p A0 B0 A−1/2 A0
1/2 1/2
= Pμ (p, A0 , B0 ) .
A+B
f ≤ f A1/2 fμ,q A−1/2BA−1/2 A1/2 , (29)
2
A+B
f A1/2fμ,p A−1/2BA−1/2 A1/2 ≤ f , (30)
2
for any positive definite matrices A and B. Then f is operator monotone on (0, ∞).
Characterizations of Operator Monotone Functions 117
To prove Theorem 2.19, we will need the following lemma which is the inverse
problem for the arithmetic mean and the power mean. Recently, in [27] we also
study the inverse problem for the generalized contraharmonic means.
Lemma 2.20 Suppose X and Y are positive definite matrices satisfying X ≤ Y <
γ X (resp. γ X < Y ≤ X) where γ = 21−1/q (resp. γ = 21−1/p ), then there exist
positive matrices A and B such that
A+B
X= and Y = Pμ (q, A, B) (resp. Y = Pμ (p, A, B)),
2
where 0 < p ≤ 1 ≤ q.
Proof We show the lemma when X ≤ Y < γ X, the remaining case can be obtained
similarly. Firstly, let us consider the case when X = In . Then it suffices to show that
given In ≤ Y = U diag ({λi (Y )}) U ∗ ≤ γ In , we can find A0 , B0 ≥ 0 such that
A0 + B0
In = and Y = Pμ (q, A0 , B0 ).
2
Or, equivalently, there exists 0 < A0 ≤ 2In such that Y = Pμ (q, A0 , 2In − A0 ) =
ϕ(A0 ), where
1/q
x q + (2 − x)q
ϕ(x) = x 1/2fμ,q x −1/2(2 − x)x −1/2 x 1/2 = .
2
Note that ϕ is continuous on [0, 2] and surjective from [0, 2] onto [1, γ ]. Since
λi (Y ) ∈ [1, γ ], for each i, we can choose δi (Y ) ∈ [0, 2] such that ϕ (δi (Y )) =
λi (Y ). The matrix
satisfies
We have
mt mt +1
γ0 < λr < · · · < λt−1 +1 ≤ γ0 ≤ γ0 t−1
+1
< · · · < λ1 ≤ γ01 +1 .
< λt−1 < · · · < λt−2 +1 ≤ γ0 t−1
It follows that
mt mt
I < γ0 I < γ02 I < · · · < γ0 I = γ0(E1 + E2 + · · · + Er )
m
≤ λr Er + · · · + λt−1 +1 Et−1 +1 + γ0 t Et−1 + Et−1 −1 + · · · + E1
m +1
≤ λr Er + · · · + λt−1 +1 Et−1 +1 + γ0 t Et−1 + Et−1 −1 + · · · + E1
...
mt−1
≤ λr Er + · · · + λt−1 +1 Et−1 +1 + γ0 Et−1 + Et−1 −1 + · · · + E1
m
≤ λr Er + · · · + λt−2 +1 Et−2 +1 + γ0 t−1 Et−2 + Et−2 −1 + · · · + E1
m +1
≤ λr Er + · · · + λt−2 +1 Et−2 +1 + γ0 t−1 Et−2 + Et−2 −1 + · · · + E1
...
m2
≤ λr Er + · · · + λ2 +1 E2 +1 + γ0 E1 + E1 −1 + · · · + E1
m +1
≤ λr Er + · · · + λ2 +1 E2 +1 + γ0 2 E1 + E1 −1 + · · · + E1
...
m1
≤ λr Er + · · · + λ2 +1 E2 +1 + γ0 E1 + E1 −1 + · · · + E1
m1 +1
≤ λr Er + . . . λ1 E1 = Y0 ≤ γ0 I.
Characterizations of Operator Monotone Functions 119
After multiplying each term of the chain of inequalities on both sides by X1/2 , let
Zk be the k-th expression of the chain, we obtain the following chain inequalities
m1 +1
0 ≤ X = Z1 ≤ Z2 ≤ · · · ≤ Zm−1 ≤ Y ≤ Zm = γ0 X.
Consequently,
In [2] Audenaert and Hiai determined values of p and q such that the following
inequality holds true
1/p 1/q
Ap + B p Aq + B q
≤
2 2
1/p 1/q
Ap + B p A+B Aq + B q
≤ ≤ (31)
2 2 2
If this system has a positive solution, then we may use the result to characterize
operator monotone function. Unfortunately, inequalities in (31) do not characterize
operator monotone functions, in general.
120 T. H. Dinh et al.
Proposition 2.21 For any q > 1, there exists a non-monotone operator function
satisfying
1/q
A+B Aq + B q
f ≤f
2 2
However, when p = 1/2 we were able to solve the inverse problem and establish a
new characterization of operator monotone functions.
Theorem 2.22 Let f be a continuous function on [0, ∞) that satisfies the following
inequality
2
A1/2 + B 1/2 A+B
f ≤f ,
2 2
Proof Firstly, we show that f (X) ≤ f (Y ) for any positive semidefinite matrices
X, Y with 0 ≤ X ≤ Y ≤ 2X. Indeed, we need to solve the following system
⎧
⎪ 2
⎨ A +B
1/2 1/2
⎪
=X
2 (33)
⎪
⎩ A + B = Y.
⎪
2
Subtracting the first equation from the second, we obtain
2
A1/2 − B 1/2
Y −X = .
2
2 2
A = X1/2 + (Y − X)1/2 , B = X1/2 − (Y − X)1/2 .
0 ≤ X = Z1 ≤ Z2 ≤ Z3 ≤ · · · ≤ Zm−1 ≤ Y ≤ Zm
with Zk ≤ Zk+1 < 2Zk for all k = 1, 2, . . . , m. Therefore, combining with the
previous arguments, we get
122 T. H. Dinh et al.
Powers-Størmer’s inequality [3, 39] asserts that for s ∈ [0, 1] the following
inequality
holds for any pair of positive matrices A, B. This is a key inequality to prove the
upper bound of Chernoff bound, in quantum hypothesis testing theory [3]. This
inequality was first proved in [3], using an integral representation of the function
t s . After that, N. Ozawa [32, Proposition 1.1] gave a much simpler proof for the
same inequality, using fact that, for s ∈ [0, 1], the function f (t) = t s (t ∈
[0, +∞)) is an operator monotone. Recently, Ogata in [36] extended this inequality
to standard von Neumann algebras. The motivation of this section is that if the
function f (t) = t s is replaced by another operator monotone function (this class is
intensively studied) then Tr (A + B − |A − B|) may get smaller upper bound that is
used in quantum hypothesis testing. Based on Ozawa’s proof we formulate Powers-
Størmer’s inequality for an arbitrary operator monotone function on (0, +∞).
Lemma 3.1 ([31, Theorem 2.5]) Let f be a strictly positive, continuous function
on [0, ∞). If f is 2n-monotone, then for any positive semidefinite A and a
contraction C in M n
C ∗ f (A)C ≤ f (C ∗ AC).
Lemma 3.2 ([14]) Let f be a strictly positive, continuous function on (0, ∞). Then
f is n-monotone if and only if − f 1(t ) is n-monotone.
Proof For any t1 , t2 , · · · , tn ∈ (0, ∞) we have
f (t )−f (t )
Since f is n-monotone, [ iti −tj j ] is positive semidefinite by Löwner [35],
hence, we have
⎛ ⎞ ⎛ ⎞
(− f (t1 i ) ) − (− f (t1j ) ) 1
− 1
⎝ ⎠ = − ⎝ f (ti ) f (tj ) ⎠
ti − tj ti − tj
1 f (ti ) − f (tj )
=− −
f (ti )f (tj ) ti − tj
Characterizations of Operator Monotone Functions 123
1 f (ti ) − f (tj )
= [ ]◦[
f (ti )f (tj ) ti − tj
≥ 0,
Hence, the function − f (tt ) is n-monotone. Therefore, from Lemma 3.2 we conclude
that
1 t
− =
− f (tt ) f (t)
is n-monotone.
Theorem 3.4 ([12]) Let f be a strictly positive, 2n-monotone function on (0, ∞).
Then for any pair of positive definite matrices A, B ∈ M n such that A ≤ B
1 1
Tr (A) + Tr (B) − Tr (|A − B|) ≤ 2Tr (f (A) 2 g(B)f (A) 2 ), (35)
where g(t) = t
f (t ) .
Proof Let A, B be positive matrices in M n such that A ≤ B. We may assume that
A and B are invertible. For operator A − B let us denote by P = (A − B)+ and
Q = (A − B)− its positive and negative part, respectively. Then we have
A − B = P − Q and |A − B| = P + Q, (36)
A + Q = B + P. (37)
124 T. H. Dinh et al.
1 1
ϕ(A + B) − ϕ(|A − B|) ≤ 2ϕ(f (A) 2 g(B)f (A) 2 ) (39)
for any positive invertible A, B ∈ B(H). Then both functions f and g on (0, ∞)
are operator monotone.
Proof Suppose that g is not operator monotone. Then there exist n ∈ N and
invertible positive matrices A, B in M n with A ≤ B such that g(A) ≤ g(B).
Hence,
1 1
A ≤ f (A) 2 g(B)f (A) 2 .
1 1
Put A = [aij ] and f (A) 2 g(B)f (A) 2 = [bij ] = B . Note that for any unitary U in
Mn
1 1
= f (U AU ∗ ) 2 g(U BU ∗ )f (U AU ∗ ) 2 .
Hence without loss of generality, we can assume that a11 > b11 by Lemma 3.6.
Let Sϕ be a density operator on H such that ϕ(X) = Tr(Sϕ X) for all X ∈ B(H).
By Lemma 3.7, there exists system {ξi }∞
an orthogonal ∞
i=1 ⊂ H and {λi }i=1 ⊂ [0, 1)
∞ n n
such that i=1 λi = 1 and i=1 aii λi > i=1 bii λi .
Let consider the following canonical inclusion:
n
n
ρ : M n −→ |ξi ξi | B (H) |ξi ξi |
i=1 i=1
n
ρ([xij ]) = xij |ξi ξj |.
i,j =1
Put
∞
∞
C = ρ(A) + |ξi ξi | and D = ρ(B) + |ξi ξi |.
i=n+1 i=n+1
126 T. H. Dinh et al.
1 1 1 1
n
ρ(f (A) 2 )ρ(g(B))ρ(f (A) 2 ) = ρ(f (A) 2 g(B)f (A) 2 ) = bij |ξi ξj |.
i=1
n ∞
= bij |ξi ξj | + |ξi ξi |.
i,j =1 i=n+1
Consequently,
∞
ϕ(C) = Tr(Sϕ (ρ(A) + |ξi ξi |))
i=n+1
n ∞
= (Sϕ ρ(A)ξi |ξi ) + λi
i=1 i=n+1
n ∞
= aii λi + λi
i=1 i=n+1
n ∞
> bii λi + λi
i=1 i=n+1
n ∞
1 1
= (Sϕ ρ(f (A) 2 g(B)f (A) 2 )ξi |ξi ) + λi
i=1 i=n+1
Characterizations of Operator Monotone Functions 127
∞
1 1
= Tr(Sϕ (ρ(f (A) 2 g(B)f (A) 2 ) + |ξi ξi |))
i=n+1
1 1
= ϕ(f (C) 2 g(D)f (C) 2 ).
A+B 1
− Aσf B ≤ |A − B|
2 2
for any positive semidefinite matrices A and B satisfying the condition AB + BA ≥
0. Using the same method as in Sect. 2, the first author [7] also obtained a new
characterizations of operator monotone functions using the matrix Powers-Størmer
inequality. It was showed that for a nonnegative function f on [0, ∞), if
A+B 1
f ≤f AB + A1/2|In − A−1/2BA−1/2 |A1/2
2 2
for any positive definite matrices A and B, then f is operator monotone on [0, ∞).
Finally, we show that if the monotonicity inequality holds for at least one normal
state on the algebra B(H), where H is some infinite-dimensional Hilbert space, then
we also have a new characterization of operator monotone functions.
Theorem 3.10 ([11]) Let ϕ be a normal state on B(H). The following condition is
sufficient (and, evidently, necessary) for a continuous function f : → R (where
is a subset of R) to be operator monotone function:
(∗): for any A, B ∈ B(H)sa such that σ (A), σ (B) ⊂ ,
Proof Let f be a continuous function that satisfies the condition (∗), and suppose
that f is not an operator monotone function on . Therefore, there exists a
natural number n, Hermitian matrices A = [aij ]ni,j =1 and B = [bij ]ni,j =1
such that σ (A ), σ (B ) ⊂ , A ≤ B , and α11 > β11 , where [αij ]ni,j =1 =
f (A ), [βij ]ni,j =1 = f (B ).
Let ξk be the eigenvectors of the density operator Sϕ corresponding to eigen-
values (possibly, zero ones) λk . In the space H we choose
an orthonormal system
{ξk }nk=1 of eigenvectors of the operator Sϕ such that nk=1 αkk λk > nk=1 βkk λk .
128 T. H. Dinh et al.
We can always do this, because the sum of all eigenvalues of the operator Sϕ , taking
into account their multiplicities, equals one, therefore we can choose λ1 sufficiently
large, and λk (k = 2, 3, · · · , n) arbitrarily small. Let us complement the system
{ξk }nk=1 to the orthonormal basis {ξk }nk=1 ∪ {ξk }k∈K (where K is a set of indexes)
consisting of eigenvectors of Sϕ . Consider H as the direct sum H1 H2 of the n-
dimensional Hilbert space H1 with the basis {ξk }nk=1 and the Hilbert spaceH2 with
the basis {ξk }k∈K . Choose some η0 ∈ and put A = A η0 E , B = B ξ0 E ,
where E is the unit operator in the space H2 . Then A, B ∈ B(H) ; σ (A), σ (B) ⊂
sa
and A ≤ B; but
n
ϕ(f (A)) = Tr (Sϕ f (A)) = αkk λk + f (ξ0 )λk
k=1 k∈K
n
> βkk λk + f (ξ0 )λk = ϕ(f (B))
k=1 k∈K
References
1. T. Ando, F. Hiai, Operator log-convex functions and operator means. Math. Ann. 350(3), 611–
630 (2011)
2. K.M.R. Audenaert, F. Hiai, On matrix inequalities between the power means: Counterexam-
ples. Linear Algebra Appl. 439(5), 1590–1604 (2013)
3. K.M.R. Audenaert, J. Calsamiglia, L.I. Masanes, R. Munoz-Tapia, A. Acin, E. Bagan, F.
Verstraete, Discriminating States: The quantum Chernoff bound. Phys. Rev. Lett. 98, 160501
(2007)
4. K.M.R. Audenaert, L. Cai, F. Hansen, Inequalities for quantum skew information. Lett. Math.
Phys. 85, 135–146 (2008)
5. R. Bhatia, Matrix Analysis, Graduate Texts in Mathematics (Springer, New York, 1997)
6. R. Bhatia, Interpolating the arithmetic-geometric mean inequality and its operator version.
Linear Algebra Appl. 413, 355–363 (2006)
7. T.H. Dinh, On characterization of operator monotone functions. Linear Algebra Appl. 487,
260–267 (2015)
8. T.H. Dinh, Some inequalities for the matrix Heron mean. Linear Algebra Appl. 528, 321–330
(2017)
9. T.H. Dinh, H. Osaka, Interpolation functions and inequalities. Banach J. Math. Anal. 9(1),
67–74 (2015)
10. T.H. Dinh, B.K. Vo, Some inequalities for operator (p, h)-convex functions. Linear Multilinear
Algebra 66(3), 580–592 (2018)
Characterizations of Operator Monotone Functions 129
11. T.H. Dinh, O.E. Tikhonov, To the theory of operator monotone and operator convex functions.
Izv. Vyssh. Uchebn. Zaved. Mat. 54(3), 9–14 (2010) [Translation: Russian Mathematics. 54(3),
7–11 (2010)]
12. T.H. Dinh, H. Osaka, H.M. Toan, On generalized Powers-Størmer’s inequality. Linear Algebra
Appl. 438(1), 242–249 (2013)
13. T.H. Dinh, M.T. Ho, H. Osaka, Interpolation classes and matrix means. Banach J. Math. Anal.
9(3), 140–152 (2015)
14. T.H. Dinh, H. Osaka, J. Tomiyama, Characterization of operator monotone functions by
Powers-Størmer type inequalities. Linear Multilinear Algebra 63(8), 1577–1589 (2015)
15. T.H. Dinh, H. Osaka, B.K. Vo, A generalized reverse Cauchy inequality for matrices. Linear
Multilinear Algebra 64(7), 1415–1423 (2016)
16. T.H. Dinh, M.S. Moslehian, C. Conde, P. Zhang, An extension of the Polya-Szego operator
inequality. Expo. Math. 35(2), 212–220 (2017)
17. T.H. Dinh, M.T. Ho, H.B. Du, On some matrix mean inequalities with Kantorovich constant.
Sci. Math. Jap. 80(2), 139–151 (2017)
18. T.H. Dinh, T.D. Dinh, B.K. Vo, A new type of operator convexity. Acta Math. Viet. 4(43),
595–605 (2018)
19. T.H. Dinh, R. Dumitru, J. Franco, New characterizations of operator monotone functions.
Linear Algebra Appl. 546, 169–186 (2018)
20. T.H. Dinh, B.K. Vo, T.Y. Tam, In-sphere property and reverse inequalities for matrix means.
Electr. J. Linear Algebra 35(35), Article 3 (2019)
21. T.H. Dinh, O.E. Tikhonov, L. Veselova, Inequalities for the extended part of a von Neumann
algebra, related to operator monotone and operator convex functions. Ann. Funct. Anal. 10(3),
425–432 (2019)
22. T.H. Dinh, M.T. Ho, C.T. Le, B.K. Vo. Two trace inequalities for operator functions. Math.
Ineq. Appl. 22(3), 1021–1026 (2019)
23. T.H. Dinh, H. Osaka, S. Wada, Functions preserving operator means. Ann. Funct. Anal. 11,
1203–1219 (2020)
24. T.H. Dinh, R. Dumitru, J. Franco, On the matrix Heron means and Renyi divergences. Linear
Multilinear Algebra (2020). https://doi.org/10.1080/03081087.2020.1763239
25. T.H. Dinh, H.R. Moradi, M. Sababheh, On the Pólya-Szegö operator inequality. Hacettepe J.
Math. Stat. 49(5), 1744–1752 (2020)
26. T.H. Dinh, A.V. Le, C.T. Le, N.Y. Phan. The matrix Heinz mean and related divergence.
Hacettepe J. Math. Stat. 49(5), 1744–1752 (2020)
27. T.H. Dinh, C.T. Le, B.K Vo. The inverse problem for generalized contraharmonic means,
accepted to Russian Math (2022)
28. T.H. Dinh, C.T. Le, V.T. Nguyen, B.K. Vo. Matrix power means and new characterizations of
operator monotone functions, accepted to Linear Multilinear Algebra (2022)
29. W. Donoghue, Monotone Matrix Functions and Analytic Continuation (Springer, New York,
1974)
30. F. Hansen, Selfadjoint means and operator monotone functions. Math. Ann. 256, 29–35 (1981)
31. F. Hansen, G.K. Pedersen, Jensen’s inequality for operator and Löwner’s theorem. Math. Ann.
258, 229–241 (1982)
32. V. Jakšic, Y. Ogata, C.A. Pillet, R. Seiringer, Quantum hypothesis testing and non-equilibrium
statistical mechanics. Rev. Math. Phys. 24(6), 1230002, 67 pp. (2012)
33. Y. Kapil, C. Conde, M.S. Moslehian, M. Singh, M. Sababheh. Norm inequalities related to the
Heron and Heinz means. Mediterr. J. Math. 14(5), paper no. 213, 18 pp. (2017)
34. F. Kubo, T. Ando, Means of positive linear operators. Math. Ann. 246(3), 205–224 (1980)
35. C. Löwner, Über monotone matrix funktionen. Math. Z. 38, 177–216 (1934)
36. Y. Ogata, A generalization of Powers-Størmer inequality. Lett. Math. Phys. 97(3), 339–346
(2011)
37. H. Osaka, J. Tomiyama, Double piling structure of matrix monotone functions and of matrix
convex functions. Linear Algebra Appl. 431, 1825–1832 (2009)
38. D. Petz, Monotone metrics on matrix spaces. Linear Algebra Appl. 244, 81–96 (1996)
130 T. H. Dinh et al.
39. R.T. Powers, E. Størmer, Free states of the canonical anti-commutation relations. Commun.
Math. Phys. 16, 1–33 (1970)
40. W. Pusz, S.L. Woronowicz. Functional calculus for sesquilinear forms and the purification
map. Rep. Math. Phys. 5, 159–170 (1975)
41. B. Simon, Löwner’s Theorem on Monotone Matrix Functions, Grundlehren der mathematis-
chen Wissenschaften, vol. 354 (Springer Nature Switzerland, Cham, 2019)
42. X. Zhan, Matrix Inequalities (Springer, 2002)
Perspectives, Means and their
Inequalities
1 Introduction
H. Osaka ()
Department of Mathematics Sciences, Ritsumeikan University, Kusatsu, Japan
e-mail: [email protected]
S. Wada
Department of Information and Computer Engineering, National Institute of Technology,
Kisarazu College, Kisarazu, Japan
e-mail: [email protected]
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 131
R. M. Aron et al. (eds.), Operator and Norm Inequalities and Related Topics,
Trends in Mathematics, https://doi.org/10.1007/978-3-031-02104-6_5
132 H. Osaka and S. Wada
Through this chapter, H is a Hilbert space. B(H )sa is the set of bounded self-adjoint
operators, B(H )+ is the set of bounded positive operators on H , and B(H )++ is the
set of invertible elements in B(H )+ . We also write A ≥ 0 when A ∈ B(H )+ , and
A > 0 when A ∈ B(H )++ .
To form the perspective function of a convex map is a useful trick in convex
analysis [34, 35]. For a real valued continuous function f on (0, ∞), a two-variable
operator function Pf defined by
is called the operator perspective of f [13–15, 32, 33]. In this section, some
properties of operator perspectives are given.
Example 2.1 Consider the operator convex function f (t)= − log t defined on
(0, ∞). Then the perspective function Pf is given by
for A, B > 0.
Note that when A and B are commuting positive definite matrices, then
where S(A, B) is relative entropy defined by S(A, B) = Tr(A log A)−Tr(A log B).
Perspectives, Means and their Inequalities 133
2.1 Homogeneity
Corollary 2.1
Remark 2.1 Let f be a continuous function on (0, ∞) which has the analytic
continuation. If we regard f (AB −1 ) as the holomorphic function calculus for
AB −1 , a function (A, B) $→ f (AB −1 )B has the homogeneity property. Some
authors define the operator perspective of non-self-adjoint operators in this way (cf.
[8]).
2.2 Convexity
Definition 2.1 Let be a convex subset of B(H )sa × B(H )sa and let P be a
function from to B(H )sa . The function P is said to be jointly convex if
holds for all A, B ∈ B(H )sa with σ (A), σ (B) ⊂ J and α ∈ [0, 1].
The function f is said to be operator concave if the reverse inequality holds. It
is known that a positive valued function f on (0, ∞) is operator concave if and
only if f is operator monotone (i.e., f (A) ≤ f (B) holds if 0 < A ≤ B) [6,
Theorem V.2.5].
The following result is well-known [24].
Proposition 2.2 Let f be a real valued continuous function on [0, ∞) with f (0) ≤
0. Then f is operator convex if and only if the inequality
Perspectives, Means and their Inequalities 135
Let f be an operator convex function on (0, ∞). For an arbitrary > 0, f (t)(:=
f (t + ) − f ( )) satisfies the assumption of the above result. So the following is
obtained.
Corollary 2.2 Let f be an operator convex function on (0, ∞). Then the perspec-
tive function Pf is jointly convex on B(H )++ × B(H )++ .
Proof Let (A1 , B1 ), (A2 , B2 ) be in B(H )++ × B(H )++ and let α ∈ [0, 1]. Put
A := αA1 + (1 − α)A2 and B := αB1 + (1 − α)B2 , then we have
Thus
holds. So the equation Pf˜ (A, B) = Pf (B, A) always holds for an arbitrary f .
136 H. Osaka and S. Wada
We next consider the case if f is an operator convex function on (0, ∞). Using
the last theorem, the map
In the following, OC, OC0 , and OM denote, respectively, all the (real valued)
operator convex functions on (0, ∞), all the functions in OC that take the value
0 at 1, and all the (real valued) operator monotone functions on (0, ∞).
Pf (A + tI, B + tI ) = Pf (A + t I + (t − t )I, B + t I + (t − t )I )
≤ Pf (A + t I, B + t I ) + Pf ((t − t )I, (t − t )I )
= Pf (A + t I, B + t I ) + 0 · I (0 ≤ t < t).
Pf (A + X, B + X)x | x
= supPf (A + X + I , B + X + I )x | x
≤ sup Pf (A + X , B + X ) + Pf (X − X + I , X − X + I ) x | x
= Pf (A + X , B + X )x | x
Example 2.1 Take either f (t) = − log t or f (t) = t log t. From the proposition
above, the sequence Pf (A + (1/n)I, B + (1/n)I ) is increasing for all A, B ≥ 0.
Moreover, if A, B are invertible, Pf (A + (1/n)I, B + (1/n)I ) 3 Pf (A, B).
Let f ∈ OC with |f (0+)| < ∞. Then f0 (:= f − f (0+)) is operator convex and
f0 (0+) = 0. Thus there exists h ∈ OM such that f0 (t) = th(t) and f˜0 = h(1/t)
[1], which implies
because that
≤ B 2 h(B 2 A−1
1 1 1 1
1 B )B
2 2
= Pf0 (B, A1 ).
Pk (B1 , A1 ) ≤ Pk (B2 , A1 )
≤ Pk (B2 , A2 ) − k(0+)(A2 − A1 )
≤ Pk (B2 , A2 ) (A1 ≤ A2 , B1 ≤ B2 ).
Pf (A + I, B + I )
for A, B ≥ 0.
From the above discussion, if an operator convex function f on (0, ∞) satisfies
|f (0+)| < ∞, then the function −f˜0 is operator monotone. For an arbitrary α ∈
(0, 1), we put g and gα by g(t) := −f˜0 (t) and gα (t) := g(t + α) − g(α). Then
gα is a non-negative operator monotone function. The perspective function Pgα is
138 H. Osaka and S. Wada
calculated as follows:
converges strongly.
Corollary 2.6 ([32, Corollary 6.4]) Let f be an operator convex function on
(0, ∞). Then the followings are equivalent:
(1) Pf (A + I, B + I ) converges strongly as ↓ 0 for every A, B ≥ 0 such that
αB ≤ A for some α > 0;
(2) |f˜(0+)| < ∞.
As we said above, the function g defined as g(t)(:= −f˜0 (t)) is operator
monotone. If |g(0+)| = |f˜(0+)| < ∞, then g − g(0+) is a non-negative operator
monotone function. Thus
converges strongly as ↓ 0 .
In the next statement, we say that a binary operation σ on B(H )++ is an operator
connection if it can be described as Aσ B = Pg (B, A) for some non-negative
function g ∈ OM.
Perspectives, Means and their Inequalities 139
Pf (A, B) = −Aσ B + λA + μB
Remark 2.6 It is obvious that the perspective function is continuous w.r.t. operator
norm topology (i.e., if An − A → 0, Bn − B → 0, then Pf (An , Bn ) −
Pf (A, B) → 0). Moreover, if f is operator convex, then, from [33, Theorem 6.1],
the following property (upper continuity) holds:
properties of this functional calculus in the case when the function f and the pair
(A, B) are restricted.
Let A, B be positive operators with AB = BA. Then N(:= A + iB) is a
normal operator. So there is an isometric *-isomorphism ϕN from the function space
C(σ (N)) of continuous functions on σ (N) onto the C*-algebra C ∗ (N) generated
by N (see [11, Corollary I.3.3]).
Note that σ (N) ⊂ σ (A) + iσ (B) and C(σ (A) × σ (B)) 4 C(σ (A) + iσ (B)).
So, for ∈ C(σ (A) × σ (B)), we can define a functional calculus (A, B) by
(A, B) := ϕN ˆ |σ (N) ,
ϕN : Bb (σ (N)) → B(H )
f ) on [0, ∞)
Remark 3.1 We shall mainly treat the bounded Borel function (= 2
which implies
A = (A + B)1/2 CC ∗ (A + B)1/2 .
Proof It is enough to show the uniqueness. If (R, S) and (R , S ) are such pairs,
then
and
for all polynomial p. Thus, for every continuous function f on [0, ∞),
and
Proposition 3.2 If A > 0, then for every continuous function f on [0, ∞),
Corollary 3.1 If B > 0, then for every continuous function f on [0, ∞),
RS
(A + B)1/2 (A + B)1/2 = B − B(A + B)−1 B = (A : B).
R+S
(A + B)1/2 S(log R − log S)(A + B)1/2 = B 1/2 log(B −1/2 AB −1/2 )B 1/2 .
for every continuous function f on [0, ∞). We shall discuss later the corresponding
mapping F of (A, B), namely
AB
z | z + (A + B)u | u = Ax | x + By | y
A+B
AB
inf (Ax | x + By | y) = z | z. (3)
x+y=z A+B
As stated in the last subsection, for every (A, B) = (0, 0), there uniquely exists the
commuting pair (R, S) of positive contractive operators such that (2) holds. This
pair (R, S) satisfy the following:
Corollary 3.2 Let H0 = ran(A + B)1/2 . For z ∈ H0 ,
Proof From the preceding proposition, for every ∈ (0, 1), there exists x , y ∈ H0
such that x + y = z and
Let δ := (R : S)z | z+ −(Rx | x + Sy | y ) and let K > be a positive
number such that
Rx | x
= R(x − x ) | x + Rx | x − x + Rx | x
≤ ( /K)R(x + x ) + Rx | x
≤ ( /K)(2x + /K) + Rx | x
< δ/2 + Rx | x .
RS
(A : B) = (A + B)1/2 (A + B)1/2 .
R+S
(A : B)z | z
= inf (Ax | x + By | y)
x+y=z
RS
= w | w
R+S
RS
= (A + B)1/2 (A + B)1/2 z | z.
R+S
Perspectives, Means and their Inequalities 145
RS
A − (A : B) = (A + B)1/2 R − (A + B)1/2 (4)
R+S
= (A + B)1/2 R 2 (A + B)1/2 . (5)
Similarly, put
ek (ek−1 − ek )
e0 := R + S(= I ), e1 := R, ek+1 := ek − (k = 1, 2, . . .).
ek + (ek−1 − ek )
Put f (t) := (t, 1) and f˜ (t) := (1, t). It is obvious that f˜ (t) = (1, t) =
t (1/t, 1) = tf (1/t) for t > 0. From the continuity of , f and f˜ (t) are
continuous function on (0, ∞) and
Let A, B are invertible positive operators. Then R, S are also invertible, and so,
from Corollary 3.1,
(A + B)1/2 C = U (C ∗ (A + B)C)1/2 ,
U ∗ U = Pran|(A+B)1/2 C|
and
U U ∗ = Pran((A+B)1/2 C) = Pran(A+B)1/2 .
The last equality comes from the assumption. Note that by Proposition 3.1 there
exists a commuting pair (R, S) of positive contractive operators on ker(A + B)⊥
such that
So, we have
and similarly
C ∗ BC = (C ∗ AC + C ∗ BC)1/2 U ∗ SU (C ∗ AC + C ∗ BC)1/2 .
Thus
The next theorem plays an important role in the study of PW-functional calculus.
Theorem 3.1 ([33, Theorem 6.1]) Let be a real valued homogeneous contin-
uous function on [0, ∞)2 and let An , Bn , A, B are in B(H )+ . If An % A and
Bn % B, then (An , Bn ) strongly converges to (A, B).
We need some lemmas. Here, (Rn , Sn ) is the pair of positive contractive operators
on ran(A + B)1/2 corresponding to (An , Bn ), namely,
An = Tn Rn Tn , A = T RT , Tn % T
Rn T x | T y
= Rn (T − Tn )x | T y + Rn Tn x | (T − Tn )y + Rn Tn x | Tn y
148 H. Osaka and S. Wada
Using the fact that = An − (An : Bn ) strongly converges to A − (A :
Tn Rn2 Tn
B) = T R 2 T (see (5)), we have the following.
Lemma 3.2 Rn2 ξ | η → R 2 ξ | η for ξ, η ∈ ran(A + B)1/2 .
Proof It is enough to show the case if ξ, η ∈ ran(A + B)1/2 . By the similar
calculation in the above proof, we have
|Rn2 T x | T y − R 2 T x | T y|
≤ (T − Tn )xT y + Tn x(T − Tn )y
+ (Tn Rn2 Tn − T R 2 T )xy
= (T − Tn )xT y + Tn x(T − Tn )y
+ (An − (An : Bn ) − (A − (A : B)))xy.
tends to 0 as n → ∞.
Assume the statement holds for some k. Then for ξ ∈ ran(A + B)1/2 ,
Proof of Theorem 3.1 We use the notations Tn and T defined in the proof of
Lemma 3.1. Put f (t) := (t, 1 − t). Then, f is a continuous function on [0, 1]
and
and
By using Proposition 3.5 and Theorem 2.1, we have xα ( ) | x ≥ 0 for all α ∈
[0, 1]. Thanks to Theorem 3.1, this sequence converges to xα | x≥ 0 as % 0.
150 H. Osaka and S. Wada
From Theorem 2.1, it is clear that the former properties of an operator perspective
are inherited by PW-functional calculus.
Corollary 3.4 Let f be a positive valued continuous function on (0, ∞) with (6).
Then the map (A, B) $→ f (A, B) is jointly concave if and only if f is operator
monotone on (0, ∞).
Corollary 3.5 Let f be a continuous function on (0, ∞) with (6). Then
f (A, B) = f˜ (B, A) holds for all (A, B) ∈ B(H )+ × B(H )+ .
Corollary 3.7 Let α, β > 0 and let f ∈ OC with (6). If f (α/β) = 0, then the map
X ≥ 0 $→ f (A + αX, B + βX) is decreasing.
We denote by OM+ the set of all positive (non-negative) operator monotone
functions on (0, ∞). If f ∈ OM+ , then f˜ is also in OM+ , so f satisfies (6).
Corollary 3.8 Let f ∈ OM+ . If 0 ≤ A ≤ C, 0 ≤ B ≤ D, then 0 ≤ f (A, B) ≤
f (C, D).
Proof Let > 0. The facts 0 < A ≤ C , 0 < B ≤ D and Proposition 2.5 imply
the following :
0 ≤ Pf (A , B ) = f (A , B ),
and
0≤ f (C ,D ) − f (A , B ),
holds.
Proof For 1, 2, 3 > 0,
= Pf (C 3 A 1 C 3 + 2 I, C 3 B 1C 3 + 2I ) (Proposition 3.5)
We next treat the domain (B(H )+ × B(H )+ )≥ . Let f be a real valued continu-
ous function on (0, ∞) with |f˜(0+)| < ∞ and let α, c > 0. Assume that positive
operators A, B satisfy A ≥ cB. Then, from the above lemma, we have R ≥ cS and
R > 0, which implies 0 ≤ RS ≤ (1/c)I .
The functions f˜ and f˜α extend to continuous functions on [0, 1/c], where
f˜(0) := f˜(0+) and f˜α (0) := f˜α (0+) = f˜(0+) . Here, we define a two variable
function ϕ(α, t) on [0, 1] × [0, 1/c] by
⎧
⎨f˜ (t) = tf 1
+α , if α ∈ [0, 1], t ∈ (0, 1/c],
α t
ϕ(α, t) :=
⎩f˜α (0+), if α ∈ [0, 1], t = 0.
152 H. Osaka and S. Wada
This function is continuous on the compact set [0, 1] × [0, 1/c]. So, that is
uniformly continuous, i.e., for every > 0, there exists δ > 0 such that if
d((α, t), (α , t )) = |(α − α )2 + (t − t )2 |1/2 < δ, then |ϕ(α, t) − ϕ(α , t )| < .
If α(= d((α, t), (0, t))) < δ, then
X − fα (A, B)
S( )
0≤ ≤ (1/c)I
R( )
where f˜ − f˜α ∞ := max{|f˜(t) − f˜α (t)| : t ∈ [0, 1/c]}. Thus the desired result
is obtained by Theorem 3.1 and Proposition 3.9.
Corollary 3.11 Let f be a continuous function on (0, ∞). Then the followings are
equivalent:
(1) Pf (A , B ) converges strongly as % 0 for all (A, B) ∈ (B(H )+ × B(H )+ )≥ ;
(2) |f˜(0+)| < ∞.
Proof (2)⇒(1). Immediate from the above proposition. (1)⇒(2). Take (A, B) =
(I, 0).
Corollary 3.12 Let f be a continuous function on (0, ∞). Then the followings are
equivalent:
(1) Pf (A , B ) converges strongly as % 0 for all (A, B) ∈ (B(H )+ × B(H )+ )≤ ;
(2) |f (0+)| < ∞.
Example 3.3 Take f (t) = t α (α ∈ R). Then the followings are equivalent:
(1) Pf (A , B ) converges strongly as % 0 for all (A, B) ∈ (B(H )+ × B(H )+ )≤
(resp. for all (A, B) ∈ (B(H )+ × B(H )+ )≥ ) ;
(2) α ≥ 0 (resp. α ≤ 1).
Remark 3.4 When the Hilbert space H is finite dimensional, (A, B) ∈ (B(H )+ ×
B(H )+ )≥ if and only if P(ker A)⊥ ≥ P(ker B)⊥ . Since the “only if” part is obvious,
we show the “if” part. The positive operators A, B have the following spectral
decomposition:
n
A= αi Pi (α1 ≥ α2 ≥ · · · ≥ αn > 0),
i=1
m
B= βj Qj (β1 ≥ β2 ≥ · · · ≥ βm > 0).
j =1
n
m
αn m
αn
m
αn
A ≥ αn Pi ≥ αn Qj = β1 Qj ≥ βj Qj = B.
β1 β1 β1
i=1 j =1 j =1 j =1
154 H. Osaka and S. Wada
Aσ B = (C −1 CACC −1 )σ (C −1 CBCC −1 )
≥ C −1 ((CAC)σ (CBC)) C −1
≥ C −1 C(Aσ B)CC −1 = Aσ B.
Lemma 4.2 Let σ ∈ and A, B ∈ B(H )+ . If an orthogonal projection P
commutes with A, B, then ((AP )σ (BP ))P = (Aσ B)P = P (Aσ B) holds.
Proof Since the condition (i) and (ii) hold, we have
P (Aσ B)P ≤ (P AP )σ (P BP ) ≤ Aσ B.
This implies
Proof Let P be an orthogonal projection. It follows from the fact that P commutes
with I and tI and the preceding lemma that P commutes with I σ (tI ). Thus I σ (tI )
reduces all closed subspaces in H .
Put f (t) := I σ (tI ). From the statement (i), f (t) is right continuous on [0, ∞).
On the other hand, from Lemma 4.1, f (t)/t (= ((1/t)I )σ I ) is left continuous on
(0, ∞). Combining them, f is continuous on [0, ∞).
Lemma 4.4 f (t)(:= I σ (tI )) is an operator monotone function on [0, ∞).
show the case that there exist orthogonal projections {Pi } such that
ProofWe first
A = i αi Pi , i Pi = I and Pi Pj = 0 (i = j ). From Lemma 4.2,
I σ A = (I σ A)( Pi ) = (I σ A)Pi
i i
= (Pi σ (αi Pi )) = (I σ (αi I ))Pi
i i
= f (αi )Pi = f (A).
i
For general A, there exist a sequence An of the above form such that An % A. So,
I σ An = f (An ) converges to I σ A(= f (A)) strongly. From the statement (i), it
follows that f is operator monotone.
Proof of Theorem 4.1 Let σ ∈ . From the above lemmas, f (t)(:= I σ (tI )) is
operator monotone and f (A) = I σ A for all A ≥ 0. Thus we have
hold for all A, B ∈ B(H )+ and > 0. Taking the strong limit of each side, Aσ B =
f˜ (A, B) holds for all A, B ∈ B(H )+ .
Let us prove the second half of the theorem. The equivalences
σ = 0 ⇐⇒ fσ = 0
and
σ1 ≤ σ2 ⇐⇒ fσ1 ≤ fσ2
are obvious. So, it is enough to show that the map σ $→ f is surjective. For
every f ∈ OM+ , put Aσ B := f˜ (A, B). This binary operation σ satisfies the
156 H. Osaka and S. Wada
statements (i) and (ii) by Corollary 3.8 and Corollary 3.9, respectively. The proof of
the statement (iii) comes from Theorem 3.1 and Corollary 3.7 This implies σ ∈ .
From the above argument, a normalized positive valued operator monotone function
on (0, ∞) is identified with an operator mean. The binary operation σ satisfying
Aσ B = A (resp. Aσ B = B) is a trivial example for an operator mean and is
denoted by l (resp. r).
It is known that a necessary and sufficient condition for a positive continuous func-
tion f on (0, ∞) to be operator monotone is that f has an integral representation as
follows :
$
x:t
f (t) = dm(x) (t > 0), (7)
[0,∞] x : 1
where m is a positive finite Borel measure on [0, ∞] (See [6, V.53, p144]).
Example 4.1 We give some examples of the correspondence m ↔ f .
sin απ x α−1
· dx ←→ f#α (t) := t α (0 < α < 1),
π 1+x
Example 4.2 We above introduced f∇α , f#α and f!α . The corresponding operator
means can be written as follows:
A∇α B = (1 − α)A + αB
= (A + B)1/2 ((1 − α)R + αS)(A + B)1/2 ,
Perspectives, Means and their Inequalities 157
RS
= (A + B)1/2 (A + B)1/2 .
αR + (1 − α)S
Remark 4.1 In the following, we write f!0 (t) := 1, f#0 (t) := 1, f!1 (t) := t and
f#1 (t) := t.
The operator means ∇α , #α and !α are called the arithmetic mean , the geometric
mean and the harmonic mean. By a simple calculation, we have f!α ≤ f#α ≤ f∇α ,
which implies !α ≤ #α ≤ ∇α for all α ∈ [0, 1].
In the following, we denote ∇ := ∇1/2 , # := #1/2 and ! :=!1/2.
Example 4.3 Since the power functions t α (0 ≤ α ≤ 1) are in OM+ 1 , the integral
)1 α
0 t dα = (t − 1)/ log t is also in OM+ . The corresponding operator mean
1
is denoted by λ and is called the logarithmic mean. The relation between the
logarithmic mean and the operator means stated above is # ≤ λ ≤ ∇.
Remark 4.2 The upper and lower bounds for the logarithmic mean have been
studied [18, 57]. The following is curious in this respect [38, 42]:
Let’s talk about statements that characterize the geometric mean. Recall the
fundamental formula for the operator geometric mean:
We first show that it is the unique positive solution of the Riccati equation.
Proposition 4.1 Let A, B, X be positive operators. If A is invertible, then X =
A#B if and only if XA−1 X = B.
158 H. Osaka and S. Wada
Proof
Proposition 4.2 Let A, B, X be positive operators. Consider the following state-
ments
AX
(1) ≥ 0;
XB
(2) XA−1 X ≤ B (∀ > 0);
(3) X ≤ A#B.
Then (1) ⇐⇒ (2) ⇒ (3) hold.
−1/2 −1/2
Proof (1) ⇐⇒ (2). Let δ > 0. Put S := A XBδ . Since
" −1/2 # " −1/2 #
I S A A X A
= −1/2 −1/2 ≥ 0,
S∗ I Bδ X Bδ Bδ
we have
* + * +
I −S x x I S x x
= ≥0
−S ∗ I y y S ∗ I −y −y
(2)⇒ (3).
XA−1 X ≤ B
⇐⇒ (A−1/2 XA−1/2 )2 = A−1/2 XA−1 XA−1/2 ≤ A−1/2 BA−1/2
#⇒ A−1/2 XA−1/2 ≤ (A−1/2 BA−1/2 )1/2
⇐⇒ X ≤ A #B.
Perspectives, Means and their Inequalities 159
⎡ ⎤
2−1 I
⎢ ⎥
f˜ (P , Q) = (P + Q)1/2 ⎣ f˜ (1, 0)I ⎦ (P + Q)
1/2
f˜ (0, 1)I
.
where a := fσ (0+) and b := fσ (0+).
1
4.2.4 Transforms on OM+
f˜σ (t) = tfσ (1/t), fσ∗ (t) = fσ (1/t)−1 , fσ⊥ (t) = t/fσ (t),
which implies the binary operations (σ̃ , σ ∗ , σ ⊥ ) are in 1 . We show some properties
of these transforms (σ̃ , σ ∗ , σ ⊥ ) on 1 . It follows from the fact σ̃˜ = σ , (σ ∗ )∗ =
σ and (σ ⊥ )⊥ = σ that these transforms are bijective map on 1 . By a simple
calculation, we have the following.
Proposition 4.4 Let σ1 , σ2 ∈ 1 . If σ1 ≤ σ2 , then
Corollary 4.1
! = ∇ ∗ ≤ λ∗ = λ⊥ ≤ #∗ = # ≤ λ ≤ ∇.
t +f
Remark 4.3 The injective map f $→ fˆ := 1+f is called the Barbour transform on
OM+ . This map plays an important role in the analysis of OM+ ([40, 46]). The
Barbour transform has the following properties:
+ = OM+ \{1}, + = {f ∈ OM+ | f! ≤ f ≤ f∇ },
1 1 1
OM OM
f/⊥ = (fˆ)⊥ , (0 0
f˜) = (fˆ)∗ , (f ∗ ) = 1̂ 1
f (f ∈ OM+ ).
Let f be a positive function on (0, ∞). It is known [6, Theorem V. 25] that f is
operator monotone if and only if f is operator concave. Using this, the following is
obtained.
Perspectives, Means and their Inequalities 161
for all x > 0, which implies 1 − f (1) ≥ f (0+) ≥ 0 and f (1) ≥ 0. So σ ≤ ∇f (1) .
Applying this argument to f ∗ , we have σ ∗ = σf ∗ ≤ ∇(f ∗ ) (1) = ∇f (1) = (!f (1) )∗ ,
which implies !f (1) ≤ σ .
dfσ
Remark 4.4 We call the positive number α := dx x=1 the weight of σ .
Proposition 4.6
{σ ∈ 1 | σ̃ = σ } {σ ∈ 1 | ! ≤ σ ≤ ∇}.
+1
Proof Put f (t) := 3tt +3 . Then the operator mean σf which corresponds f is in
{σ ∈ | ! ≤ σ ≤ ∇}\{σ ∈ 1 | σ̃ = σ } .
1
t +f
Remark 4.5 Recall the Barbour transform f $→ fˆ = 1+f . For f ∈ OM+
1 with
{f ∈ OM+
1
\{1} | f = f ∗ } = {fˆ | f = f˜, f ∈ OM+ }.
is a typical example.
162 H. Osaka and S. Wada
Let f ∈ OM+
1 . The two variable positive function
f˜ (s, t) := sf (t/s) (s, t > 0)
is homogeneous and is monotone for each variable. Furthermore, it satisfies
and
The function pr,α (s, t) is increasing w.r.t. r and (s, t) $→ pr,α (s, t) is an operator
mean for all α ∈ [0, 1] if and only if r ∈ [−1, 1]. The map r $→ pr,α is a path
connecting familiar operator means. For example,
An operator mean pr,α (r ∈ [−1, 1]) is called the power mean [45].
Proposition 4.7 Let A, B, X be positive invertible operators and let (r, α) ∈
[−1, 1] × [0, 1]. Then X = A σpr,α B if and only if
where σpr,α is an operator mean which corresponds to the function t $→ pr,α (1, t),
fr (t) := (t r − 1)/r (r = 0) and f0 (t) := log t.
The proof is left to the reader.
Corollary 4.2 Let A, B, X be positive operators and let r > 0. Then
(1 − α) fr (A , X( )) + α fr (B , X( ))(= 0)
is discussed in [41, Theorem 5.6]. The existence and uniqueness of the positive
solution X is not trivial in general, whereas the proof in the case when n = 2 is very
easy. When r = 0, the solution X is a multivariate extension of the geometric mean
and it is called the Karcher mean.
and
1/(s−t )
1 ss
S1 (s, t) := lim Sα (s, t) = .
α→1 e tt
In [51], V.E. Szabó define the following function and discuss its operator mono-
tonicity:
(t α1 − 1)(t α2 − 1) · · · (t αn − 1)
t $→ t γ ,
(t β1 − 1)(t β2 − 1) · · · (t βn − 1)
This function is called the power difference mean and is an operator mean if and
only if p ∈ [−1, 2] [18, 30, 44, 52].
The Lehmer mean defined by
sp + t p s 2p − t 2p s p−1 − t p−1
lp (s, t) := = ·
s p−1 + t p−1 s 2p−2 −t 2p−2 sp − t p
is introduced in [25, 26]. We call this the Petz-Hasegawa mean. This function is an
operator mean if and only if p ∈ [−1, 2] ([5, 25]). An elementary calculus shows
that this function equals the harmonic mean if p = −1, 2, the logarithmic mean if
p = 0, 1 and the power mean p1/2,1/2 if p = 1/2.
Remark 4.8 Let (a, b) be a pair of real numbers with |a|, |b| ≤ 2. Put
b ta − 1
ma,b (t) := .
a tb − 1
1 is that (a, b) is
A necessary and sufficient condition for this function to be in OM+
in the following set [44]:
5 Operator Inequalities
In this section, we study some inequalities involving an operator mean. A linear map
: B(H ) → B(K) is said to be positive if (A) ∈ B(K)+ for every A ∈ B(H )+ .
For a positive map , by estimating (IH ), some properties of becomes clear.
The continuity of a positive map is guaranteed by the fact = (IH ) (cf.
[47]). A positive map is said to be strictly positive (resp. unital ) if (IH ) ∈
B(K)++ (resp. (IH ) = IK ). Since a positive map preserves the order relation,
we have
for all A ∈ B(H )++ , where α := min sp(A). So, a positive unital map is strictly
positive.
In what follows, we denote P(H, K) (resp. P(H, K)1 ) the set of strictly positive
(resp. unital positive) map from B(H ) to B(K).
If A is a positive definite matrix having the spectral decomposition: A =
i αi Pi , then, for ∈ P(H, K)1 ,
IK (A−1 )−1 (A) − (A−1 )−1 0 IK 0
0 IK 0 (A−1 ) (A−1 )−1 IK
(A) IK
=
IK (A−1 )
αi (Pi )
(Pi ) αi 1
= = ⊗ (Pi ) ≥ 0.
(Pi ) αi−1 (Pi ) 1 αi−1
i i
By the similar argument, the following two lemmas (Choi’s inequality and Kadi-
son’s inequality) are obtained.
Lemma 5.1 ([9, Theorem 2.1]) Let ∈ P(H, K)1 . Then
(A−1 ) ≥ (A)−1
(A2 ) ≥ (A)2
Proposition 5.1 Let ∈ P(H, K)1 and let f be an operator convex function on
the open interval (α, β). Then,
f ( (A)) ≤ (f (A))
A2 −1 −1 1 1
= (A) + 2 (IH ) + 2
IH − xA x x x IH − xA
−1 −1 1 1
≥ (A) + 2 (IH ) + 2
x x x IK − x (A)
(A)2
= .
IK − x (A)
A2 (A)2
= (A2 ) ≥ (A)2 = .
IH − xA IK − x (A)
β−α α+β
In the general case, if we put g(t) := f 2 t + 2 , then g is operator convex
on (−1, 1) and we have
2 α+β
f ( (A)) = g (A) − IK
β −α β −α
2 α+β
=g A− IH
β−α β −α
2 α+β
≤ g A− IH = (f (A)).
β−α β −α
Theorem 5.1 Let ∈ P(H, K) and let f be a real valued operator convex
function on (0, ∞). Then
Corollary 5.1 Let σ be an operator mean and let ∈ P(H, K). Then
(Aσ B) ≤ (A σ B )
= − (P−fσ (B , A ))
168 H. Osaka and S. Wada
≤ −P−fσ ( (B ), (A ))
= (A )σ (B )
= ( (A) + (IH ))σ ( (B) + (IH )).
Let t be a positive real number with t ≥ 1 (resp. t ≤ 1). The function [1, ∞) 8 r $→
t r is monotone increasing (resp. decreasing). A certain numerical mean (a, b) $→
a σf b(:= af (b/a)) also satisfies this property. If the positive function f satisfies
f (t)r ≤ f (t r ) (r ≥ 1), then
a r σf br ≥ (a σf b)r (r ≥ 1)
which is equivalent to
a σf b ≥ 1 ⇒ a r σf br ≥ 1 (r ≥ 1).
In [4], Ando and Hiai prove that an operator version of this inequality holds for
the weighted geometric mean. In this section, we study the similar statement
A, B > 0, Aσ B ≥ I ⇒ Ar σ B r ≥ I (9)
and
A, B > 0, Aσ B ≤ I ⇒ Ar σ B r ≤ I. (10)
For a positive operator X, we denote the minimum value of sp(X) by λmin (X).
The following inequality holds for an arbitrary operator mean.
Proposition 5.2 Let σ be an operator mean and A, B > 0. Then
fσ (C r )
Ar σ B r ≥ λmin λmin (Aσ B)r−1 (Aσ B) for 1 ≤ r ≤ 2,
fσ (C)r
−1 −1
where C := A 2 BA 2 and fσ (t) := 1σ t (t > 0).
Perspectives, Means and their Inequalities 169
Proof We first show the case when λmin (Aσ B) = 1. Since Aσ B ≥ I , we have
fσ (C) ≤ A. Set := 2 − r. Then
1
Ar σ B r = A 2 fσ (A− 2 + 2 CA 2 B − A 2 CA− 2 + 2 )A 2
r r 1 1 1 r 1 r
1 1 1 1 1 1
= A 2 fσ (A− 2 + 2 CA 2 (A− 2 C −1 A− 2 ) A 2 CA− 2 + 2 )A 2
r r r r
−1+ −1+
= A 2 A 2 fσ (A 2 C[A# C −1 ]CA 2 )A 2 A 2
1 1− 1− 1
1
2 3 1
= A 2 A1− σ {C[A# C −1 ]C} A 2
1 1 1 −1 1
≥A 2 σ C # C C A2
fσ (C)1− fσ (C)
1 fσ (C 2− ) 1 1 fσ (C r ) 1
= A2 A2 = A2 A2
fσ (C)1− fσ (C)r−1
fσ (C r ) 1 1 fσ (C r )
≥ λmin r
A 2 fσ (C)A 2 = λmin (Aσ B).
fσ (C) fσ (C)r
fσ (D r )
(A/α)r σ (B/α)r ≥ λmin (A/α)σ (B/α),
fσ (D)r
fσ (C r )
Ar σ B r ≥ λmin α r−1 (Aσ B).
fσ (C)r
Using this result, the condition fσ (t r ) ≥ fσ (t)r (t > 0, 2 ≥ r ≥ 1) clearly
implies
A, B ≥ 0, Aσf B ≤ I ⇒ Ar σf B r ≤ I (r ≥ 1).
In what follows, we denote the set of all functions in OM+ 1 satisfying (i) in
Corollary 5.2 (resp. (i) in Corollary 5.4) by PMI (resp. PMD). Note that
f ∈ P MI ⇐⇒ f ∗ ∈ P MD.
So
Since the function t $→ tα (α ∈ [0, 1]) is in P MI ∩ P MD, we have the following.
Perspectives, Means and their Inequalities 171
Corollary 5.6 (The Ando–Hiai Inequality [4]) Let α ∈ [0, 1]. Then the follow-
ings hold.
Remark 5.10 Statements (11) and (12) are equivalent. Assume (11). Then, for
A, B > 0 with A#α B ≤ I , we have A−1 #α B −1 ≥ I . So A−r #α B −r ≥ I holds
by (11). This implies Ar #α B r ≤ I .
By using the Ando–Hiai inequality, the essential part of the Furuta inequality [19]
is obtained.
Proposition 5.3 If A ≥ B > 0, then
A−r # 1+r B p ≤ B
p+r
for p ≥ 1 and r ≥ 0.
Proof It is enough to assume r > 0. We first show
A−r # p+r
r B ≤ I
p
(p ≥ 1, r > 0). (13)
Note that there exists s ∈ (0, 1] such that r/s ≥ 1. Set q := p/r. It follows from
A ≥ B > 0 that
A−s # 1 B sq ≤ B −s # 1 B sq = I
1+q 1+q
which implies
A−r # p+r
r B = A
p −s(r/s)
# 1 B sq(r/s) ≤ I
1+q
= B # p−1 B p #
p p A−r
p p+r
172 H. Osaka and S. Wada
≤ B # p−1 I = B.
p
p
Proposition 5.4 (The Furuta Inequality) If A ≥ B ≥ 0, then
1/q p+r
Ar/2B p Ar/2 ≤A q
1/q p+r
Ar/2B p Ar/2 ≤A q
p+r
Thus, by taking q(1+r) power of each side,
p+r 1+r
1/q q(1+r) p+r p+r
Ar/2B p Ar/2 = Ar/2B p Ar/2 ≤A q .
Proof of Proposition 5.4 Suppose (p, q, r) satisfies the condition. If p = 0 or r =
0, then the desired inequality clearly holds. So we assume p > 0 and r > 0. Put
s := min{p, 1}(≤ 1). Then p := (p/s) ≥ 1 and r := (r/s) > 0. From the above
lemma,
1+r
s+r
r r p +r
r r p+r s p
A B A
2 p 2 = (A ) (B ) (A )
s 2 s 2
(p +r ) p1+r
+r
≤ (As )
= As+r .
4 5
Since q ≥ max p+r 1+r , 1 p+r
, we have 1 ≥ q(s+r) . Thus, by taking p+r
q(s+r) power of
each side, the desired inequality is obtained.
Perspectives, Means and their Inequalities 173
(Ap σf˜ B p )1/p = Pf (Ap , B p )1/p ≤ exp(f (1) log A + (1 − f (1)) log B).
Proposition 5.6 ([32]) Let α ∈ [0, 1] and P MDα be the set of all f ∈ OM+
1 such
that f (1) = α. If A, B > 0. Then the followings are equivalent:
(i)(1 − α) log A + α log B ≤ 0;
(ii)Ap σf B p ≤ I for all p > 0 and for all f ∈ P MDα ;
(iii)Ap #α B p ≤ I for all p > 0;
(iv) p $→ Ap σf B p is a decreasing map from [0, ∞) into B(H )++ for all f ∈
P MDα ;
(v) p $→ Ap #α B p is a decreasing map from [0, ∞) into B(H )++ .
Proof The equivalence of (i) − (iii) is immediate from the above argument.
(iii) ⇒ (iv). From Corollary 5.4, we have Ar σf B r ≤ Aσf B (r ∈ [1, 2]). Thus
for all r ≥ 1, Ar σf B r ≤ Aσf B holds. Using this,
Combining this and the fact that Ap σf B p ≤ I = A0 σf B 0 (p > 0), the desired
result is obtained.
(iv) ⇒ (v). Immediate.
(v) ⇒ (i). From Proposition 5.5,
174 H. Osaka and S. Wada
Aμ # βμ B λ ≥ I for λ ≥ 1 and μ ≥ 1.
αλ+βμ
αλ βμ
log(W −1/2 Aμ W −1/2 ) + log(W −1/2 B λ W −1/2 ) ≥ 0,
αλ + βμ αλ + βμ
The perspective function for a real valued continuous function is often appeared in
the theory of convex analysis. The operator perspective of invertible positive oper-
ators is a non-commutative analogy of this. In Sect. 2, we state some fundamental
properties of an operator perspective such as the equivalence between the operator
convexity of the representation function and the joint convexity of the perspective
[13, 15].
In Sect. 3, we study a two-variable functional calculus (A, B) of positive
operators A, B developed by W. Pusz and S. L. Woronowicz (PW-functional
calculus for short)[49]. We introduce some important properties in [27, 33]. We
defined the PW-functional calculus for a real valued continuous function with the
condition (6). The set of such functions does not contain some important functions
such as log t and t log t, but (A, B) can often be calculated in some restricted
domain (Sect. 3.2.5).
The operator connection in the sense of Kubo and Ando [39] can be considered
as the PW-functional calculus for a positive operator monotone function on [0, ∞).
Their axiomatization not only teaches us the essence of an operator connection, but
also serves as a tool for checking whether a given binary operation is an operator
connection. In Sect. 4, we introduce some fundamental properties of an operator
mean and give some examples.
Thanks to the basic results of the operator perspective described in the previous
section, some of the properties of the operator mean become apparent by taking
limits.
Transformations ( ˜, ∗ , ⊥ ) on the set of operator means are well-known [39]. In
Sect. 4.2, the Barbour transform is introduced. Though it was first defined in [40],
but before that, its sprouting can be seen in the papers of some authors [7, 43].
The fixed point with respect to the transformation σ $→ σ ∗ is called a self-
adjoint operator mean. F. Hansen’s work concerning a characterization of that
was groundbreaking [22]. We briefly mention the relations between a self-adjoint
operator mean and a symmetric one proved by H. Osaka and S. Wada [46].
Inequalities involving operator perspectives and positive linear maps have been
studied(e.g.,[1, 2, 9]). In Sect. 5.1, we discuss about how the operator convexity and
the operator concavity of a representing function are reflected in the inequalities.
Almost all statements are classical and fundamental.
For an operator mean σ , some statements comparing Aσ B and Ar σ B r are
treated in Sect. 5.2. A central result (Proposition 5.2) is stated in [4, Theorem 2.1].
Functions satisfying the condition (i) in Corollary 5.2(resp. Corollary 5.4) is said
to be power monotone increasing (resp. decreasing) [54] and has been studied in
[31, 32, 55].
The Furuta inequality was developed as a generalization of the Löwner-Heinz
inequality: A ≥ B ≥ 0 ⇒ As ≥ B s ≥ 0 (s ∈ [0, 1]) [19]. Though the original
proof is elegant and not very difficult, but we gave a slightly longer proof using the
Ando–Hiai inequality in Sect. 5.3. The proof is given by M. Fujii and E. Kamei
176 H. Osaka and S. Wada
[17]. They showed that the essential part of the Furuta inequality(Proposition 5.3)
and the Ando–Hiai inequality (Corollary 5.6) imply each other.
For A, B ∈ B(H )++ , a (weaker) order defined by log A ≥ log B is called the
chaotic order. Many equivalent conditions have been studied [16, 20, 21, 32, 53].
Among them, we introduce some statements that come from the Ando–Hiai
inequality [32].
The last statement given here is developed by T. Furuta, M. Yanagida and T.
Yamazaki [21]. We show this result using the method developed by T. Yamazaki in
the theory of the multivariate Ando–Hiai type inequality [56].
Acknowledgments The first author’s research is supported by KAKENHI grant No. JP20K03644.
References
1. T. Ando, Topics on Operator Inequalities, Lecture Note (Hokkaido University, Sapporo, 1978)
2. T. Ando, Concavity of certain maps on positive definite matrices and applications to Hadamard
products. Linear Algebra Appl. 26, 203–241 (1979)
3. T. Ando, On some operator inequalities. Math. Ann. 279(1), 157–159 (1987)
4. T. Ando, F. Hiai, Log majorization and complementary Golden–Thompson type inequalities.
Linear Algebra Appl. 197, 113–131 (1994)
5. Á. Besenyei, The Hasegawa–Petz mean: properties and inequalities. J. Math. Anal. Appl.
391(2), 441–450 (2012)
6. R. Bhatia, Matrix Analysis. Graduate Texts in Mathematics, vol. 169 (Springer, New York,
1997)
7. J.L. Brenner, W.A. Newcomb, O.G. Ruler, An inequality: problem 85–20. SIAM Rev. 28, 573
(1987)
8. H.J. Carlin, G.A. Noble, Circuit properties of coupled dispersive lines with applications to
wave guide modelling, in Proceedings on Network and Signal Theory, ed. by J.K. Skwirzynki,
J.O. Scanlan. (Peter Pergrinus, Stevenage, 1973), pp. 258–269
9. M.-D. Choi, A Schwarz inequality for positive linear maps on C ∗ -algebras. Illinois J. Math.
18, 565–574 (1974)
10. J.B. Conway, A Course in Operator Theory. Graduate Studies in Mathematics, vol. 21
(American Mathematical Society, Providence, RI, 2000)
11. K.R. Davidson, C*-Algebras by Example. Fields Institute Monographs, vol. 6 (American
Mathematical Society, Providence, RI, 1996)
12. B. Dacorogna, P. Maréchal, The role of perspective functions in convexity, polyconvexity, rank-
one convexity and separate convexity. J. Convex Anal. 15(2), 271–284 (2008)
13. E.G. Effros, A matrix convexity approach to some celebrated quantum inequalities. Proc. Natl.
Acad. Sci. U. S. A. 106(4), 1006–1008 (2009)
14. E. Effros, F. Hansen, Non-commutative perspectives. Ann. Funct. Anal. 5(2), 74–79 (2014)
15. A. Ebadian, I. Nikoufar, M.E. Gordji, Perspectives of matrix convex functions. Proc. Natl.
Acad. Sci. U. S. A. 108(18), 7313–7314 (2011)
16. M. Fujii, T. Furuta, E. Kamei, Furuta’s inequality and its application to Ando’s theorem. Linear
Algebra Appl. 179, 161–169 (1993)
17. M. Fujii, E. Kamei, Ando–Hiai inequality and Furuta inequality. Linear Algebra Appl. 416(2–
3), 541–545 (2006)
18. J.I. Fujii, M. Fujii, Y. Seo, An extension of the Kubo–Ando theory: solidarities. Math. Jpn. 35,
509–512 (1990)
Perspectives, Means and their Inequalities 177
47. V.I. Paulsen, Completely Bounded Maps and Operator Algebras (Cambridge University Press,
Cambridge, 2002)
48. G.K. Pedersen, Analysis Now. Graduate Texts in Mathematics, vol. 118 (Springer, New York,
1989)
49. W. Pusz, S.L. Woronowicz, Functional calculus for sesquilinear forms and the purification
map. Rep. Math. Phys. 5, 159–170 (1975)
50. K.B. Stolarsky, Generalizations of the logarithmic mean. Math. Mag. 48, 87–92 (1975)
51. V.E.S. Szabó, A class of matrix monotone functions. Linear Algebra Appl. 420, 79–85 (2007)
52. S.-E. Takahasi, M. Tsukada, K. Tanahashi, T. Ogiwara, An inverse type of Jensen’s inequality.
Math. Jpn. 50, 85–91 (1999)
53. M. Uchiyama, Some exponential operator inequalities. Math. Inequal. Appl. 2, 469–471 (1999)
54. S. Wada, Some ways of constructing Furuta-type inequalities. Linear Algebra Appl. 457, 276–
286 (2014)
55. S. Wada, When does Ando–Hiai inequality hold?. Linear Algebra Appl. 540, 234–243 (2018)
56. T. Yamazaki, The Riemannian mean and matrix inequalities related to the Ando–Hiai
inequality and chaotic order. Oper. Matrices 6, 577–588 (2012)
57. Z.-H. Yang, New sharp bounds for logarithmic mean and identric mean. J. Inequal. Appl.
2013(1), 116 (2013)
Cauchy–Schwarz Operator and Norm
Inequalities for Inner Product Type
Transformers in Norm Ideals of Compact
Operators, with Applications
for strongly square integrable operator families {At }t ∈, {Bt∗ }t ∈ and symmetri-
cally norming functions , such that the associated unitarily invariant norm is
nuclear, Q∗ or arbitrary, under some additional commutativity conditions. The
applications of this and complementary inequalities for Q and Schatten–von
Neumann norms to Aczél–Bellman, Grüss–Landau, arithmetic–geometric, Young,
Minkowski, Heinz, Zhan, Heron, and generalized derivation norm inequalities are
also presented.
1 Introduction
In this paper we will denote respectively by B(H) and K(H) the spaces of all
bounded and all compact linear operators on a separable, complex Hilbert space H.
Each symmetrically norming (s.n.) function ϒ, defined on sequences of complex
numbers, gives rise to the associated symmetric or a unitary invariant (u.i.) norm
||·||ϒ on operators, defined on the naturally associated norm ideal Cϒ (H) ⊂ K(H).
Most important example of s.n. functions is the trace s.n. function (denoted also by
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 179
R. M. Aron et al. (eds.), Operator and Norm Inequalities and Related Topics,
Trends in Mathematics, https://doi.org/10.1007/978-3-031-02104-6_6
180 D. R. Jocić and M. Lazarević
def ∞
1 or 1 ), defined by (λn )∞
n=1 = n=1 |λn |. For any s.n. function ϒ, there is their
adjoint s.n. function, which we denote by ϒ ∗. For any p ≥ 1 a s.n. function ϒ could
(p)
be p modified and this p-modification ϒ also represents a new s.n. functions, for
which the corresponding ideal of compact operators will be denoted by Cϒ (p) (H).
(p) ∗
Similarly, the ideal of compact operators, associated to dual s.n. function ϒ
we will denote by Cϒ (p)∗ (H) and by Cϒ (p) (H) we will denote the closure of
◦
finite rank operators in the ||·||ϒ (p) norm. The Schatten–von Neumann trace classes
Cp (H) = C(p) (H) are the most important and the best known examples of norm
def
ideals associated to degree p modified (i.e. its s.n. function ) norms. Amongst them,
C1(H) is also known as the class of nuclear operators, while C2(H) is known as the
Hilbert–Schmidt class. For the norm in Cp (H) we will use the simplified notation
(p) p (2)
||·||p . For p ≥ 2, all norms ||·||ϒ (p) are also known as Q-norms, as ϒ = (ϒ ( 2 ) )
p
and ϒ ( 2 ) is also a s.n. function, while their associated dual norms ||·||ϒ (p) ∗ are also
known as Q∗ -norms.
If (, M, μ) is a space with a measure μ on σ -algebra M, consisting of
(measurable) subsets of , then we will refer to a function A : → B(H) : t $→ At
as to a ([M]) weakly∗ -measurable whenever t $→ At g, h is a ([M]) measurable
for all g, h ∈ H. If, in addition, those functions are [μ] integrable on , then A is
called ([μ]) weakly∗ -integrable on , in which ) case there is the unique (known as
Gel’fand or weak∗ -integral and denoted by At dμ(t)) operator in B(H), having
the property
*$ + $
At dμ(t)h, k = At h, k dμ(t) for all h, k ∈ H.
)
Note that At dμ(t) also satisfies the definition of Pettis integral. For a more
complete account about weak∗ -integrability of operator valued (o.v.) functions
the interested reader is referred to [5, p.53], [15, p.320] and [20, lemma 1.2].
Let also L2 (, μ, H)) denote the space of all (weakly) measurable functions
f : → H such that f (t)2 dμ(t) < +∞, and similarly, let L2G (, μ,B(H))
denote
) the space of all weak∗ -measurable functions A : → B(H) such that
||At f || dμ(t) < +∞ for all f ∈ H. In this case we say that A is ([μ])
2
square integrable (s.i.), and by analogy, a family {An }∞ n=1 in B(H) will be called
a (strongly) square summable (s.s.) if ∞ n=1 ||A n h|| 2 < +∞ for all h ∈ H. Also,
A∗A is Gel’fand integrable iff A ∈ L2G (, μ,B(H)), as shown (amongst others) in
[15, ex. 2]. If a family {At }t ∈ consists of mutually commuting normal operators,
we will refer to it as to a m.c.n.o. family. If for A ∈ L2G (, μ,B(H)) the associated
family {At }t ∈ is a m.c.n.o. family, then we will refer to {At }t ∈ (and A) as to s.i.
and m.c.n.o. family (o.v. function).
Cauchy–Schwarz Inequalities for i.p.t. Transformers 181
∞ ∞
For sequences {An }n=1 , {Bn }n=1 in B(H) and a norm ideal Cϒ (H) the associated
transformer ∞ n=1 An ⊗Bn is called σ -elementary transformer on Cϒ (H) iff for
every X ∈ Cϒ (H) there is Y ∈ Cϒ (H), such that Y = [w] ∞ n=1 An XBn =
def
z−z∗
2i > 0}, C– = {z ∈ C : 9z < 0} and 2Z (H) for the Hilbert space of all sequences
def
The inequality (1) is a special case Xt ··= X, Ct ··= At and Dt ··= Bt for all t ∈
in [15, th. 3.1(a)], inequalities (2), (3) and (a3) are proved in [26, lemma 2.1], while
the inequality (4) is proved in [26, cor. 2.3].
Cauchy–Schwarz Inequalities for i.p.t. Transformers 183
For elementary operators the inequality (2) is further refined in [19, lemma 2.2],
saying that
Lemma 2.2 If X ∈ B(H) and {A1 , . . . , AN }, {B1 , . . . , BN } are families in B(H)
for N ∈ N, then for all c ≥ || Nn=1 A∗n An ||1/2 > 0 we have the identity
N 2
N
N 1 −1
N 2
A∗n XBn + A∗n An A∗m XBm − cXBn
2
An cI + c2 I −
n=1 n=1 n=1 m=1
N
= c2 |XBn |2 .
n=1
N 2
N
N
and thus necessary and sufficient conditions for equality to take place in (5) are
Nthere ∗exists D∈N B(H)
that that An D = XBn for all n = 1, . . . , N and
such
∗ A D = 0.
A
n=1 n nA − A
n=1 n n
The fundamental role in investigation of i.p.t. transformers is played by the
following list of Cauchy–Schwarz norm inequalities.
Theorem
) 2.3 Let p ≥ 2, ,ϒ be s.n. functions and X ∈ B(H), then
A
t XBt dμ(t) ∈ C (H) :
) 1 ) 1
(a) if A, B ∗ ∈ L2G (, μ,B(H)) and ∗
At At dμ(t) 2X Bt Bt∗ dμ(t) 2 ∈ C (H),
in which case
$ $ 1
2
$ 1
2
At XBt dμ(t) ≤ A∗t At dμ(t) X Bt Bt∗ dμ(t) , (6)
$ $ 1
2
$ 1
2
At XBt dμ(t) ≤ A∗t At dμ(t) X Bt∗ Bt dμ(t) , (7)
184 D. R. Jocić and M. Lazarević
u.e.2s.i.m.c.n.o. family; )
(c) if under conditions of (b) operator Bt∗ Bt dμ(t) is invertible, then
$ $ − 12 $ 1
2
At XBt dμ(t) Bt∗ Bt dμ(t) ≤ A∗t At dμ(t) X ,
(8)
) 1
(d) if A∗, B ∗ ∈ L2G (, μ,B(H)) and X Bt Bt∗ dμ(t) 2 ∈ C (H),
$ $ 1
2
$ 1
2
At XBt dμ(t) ≤ At A∗t dμ(t) X Bt Bt∗ dμ(t) ,
(9)
u.e.2s.i.m.c.n.o. family; )
(e) if under conditions of (d) operator At A∗t dμ(t) is invertible, then
$ − 12 $ $ 1
2
At A∗t dμ(t) At XBt dμ(t) ≤ X Bt Bt∗ dμ(t) ,
(10)
The inequality:
• (6) in the case (a1) of Theorem 2.3 is exactly the case p ··= q ··= r ··= 1 of the
inequality (24) in [15, th. 3.3];
• (6) in the case (a2) of Theorem 2.3 is the inequality (32) in [26, th. 3.1(d)]. If
X ∈ Cϒ (p)∗(H), then (6) is the inequality (3.3) in [18, th. 3.1(c)];
Cauchy–Schwarz Inequalities for i.p.t. Transformers 185
• (6) in the case (a3) of Theorem 2.3 is the inequality (23) in [15, th. 3.2]. In the
case of the counting measure μ of ··= N, the inequality (6) was proved earlier
in [13, th. 2.2]. If X ∈ C|||·|||(H), then (6) is the inequality (3.4) in [18, th. 3.1(d)] ;
• (7) in the case (b1) (resp. (b2)) of Theorem 2.3 is exactly the inequality (28) in
[26, th. 3.1(b)] (resp. the special case Ct ··= A∗t and Dt ··= Bt for all t ∈ of the
inequality (33) in [21, lemma 3.4]). In the case (b3) the inequality (7) is exactly
the inequality (3.1) in [18, th. 3.1(a)];
• (9) in the case (d1) (resp. (d2)) of Theorem 2.3 is exactly the inequality (29) in
[26, th. 3.1(b)] (resp. the special case Ct ··= A∗t and Dt ··= Bt for all t ∈ of the
inequality (34) in [21, lemma 3.4]). In the case (d3) the inequality (7) is exactly
the inequality (3.2) in [18, th. 3.1(b)];
• (8) in the case (c1) (resp. (c2)) in Theorem 2.3 derives similarly by applying
the inequality (28) in [26, th. 3.1(b)] to L2G (, μ,B(H)) families {At }t ∈ and
6 ) − 17
Bt Bt∗Bt dμ(t) 2 t ∈ (instead of {Bt }t ∈) (resp. the inequality (33) in [21,
lemma 3.4] to s.i. families {Ct∗ }t ∈ , {Dt }t ∈, given by Ct ··= A∗t and Dt ··=
) −1
Bt Bt∗Bt dμ(t) 2 for all t ∈ );
• (10) in the case (e1) (resp. (e2)) in Theorem 2.3 proves by analogy to the proof
of the case (c1) (resp. (c2)).
We conclude our list of Cauchy–Schwarz inequalities for i.p.t. transformers
within the context of Schatten–von Neumann ideals.
Theorem 2.4 If μ is a σ -finite measure on and A, A∗, B, B ∗ ∈ L2G (, μ,B(H)),
then for all 1 ≤ p, q, r < +∞ satisfying p2 = q1 + 1r and X ∈ Cp(H)
$
At XBt dμ(t) ≤ (11)
p
8 8
$ $ q−1 $ $ r−1
A∗t At A∗t dμ(t) Bt∗ Bt dμ(t) Bt∗ dμ(t)
2q 2r
At dμ(t) X Bt ,
p
) ∗
) ∗
If, in addition, At At dμ(t) and Bt Bt dμ(t) are invertible, then
$ 2q − 2
1 1 $ $ 2r − 2
1 1
$ p−2
1 1 $ $ − 12
At A∗t dμ(t) At XBt dμ(t) Bt∗ Bt dμ(t) ≤
p
$ 1
p
A∗t At dμ(t) X , (15)
p
$ − 12 $ $ p−2
1 1
The inequality (11) was proved in [15, th. 3.3], while the inequality (12) (resp. (13))
represents the special case q ··= +∞ (resp. r ··= +∞), which actually follows from
the proof of [15, th. 3.3]. In the case of the counting measure μ of ··= N, the
special case p ··= q ··= r of the inequality (11) was proved earlier in [14, th. 2.1].
The inequality (14) follows directly by applying the inequality (11) to the
6) ∗
1 −1 7
t t ∈ instead of {At }t ∈ and to the s.i. family
s.i. family 2q 2 A
At At dμ(t)
6 ) ∗
1 17
−2
Bt Bt Bt dμ(t) 2r
t ∈
instead of {Bt }t ∈ .
Similarly, the inequality (15) (resp. (16)) follows immediately from the
6) 1 − 1 7
inequality (12) (resp. (13)) applied to s.i. families At A∗t dμ(t) p 2At t ∈ and
6 ) − 1 7 6) − 1 7
Bt Bt∗ Bt dμ(t) 2 t ∈ (resp. to s.i. families ∗
At At dμ(t)
2A
t t ∈ and
6 ) 1 17
−
Bt Bt∗ Bt dμ(t) p 2 t ∈ .
Cauchy–Schwarz Inequalities for i.p.t. Transformers 187
Aczél–Bellman u.i. norm inequality for i.p.t. transformers was presented in [15,
th. 4.1], saying that
Theorem 3.1 If μ is a σ -finite measure) on , {At }t ∈ and {B) t }t ∈ are m.c.n.o.
families in L2G (, μ,B(H)), such that A∗t At dμ(t) and Bt Bt∗ dμ(t) are
)
contractions and X − At XBt dμ(t) ∈ C|||·|||(H) for some X ∈ B(H), then
$ 1
2
$ 1
2
$
I− A∗t At dμ(t) X I − Bt Bt∗ dμ(t) ≤ X− At XBt dμ(t) . (17)
This complements the Cauchy–Schwarz u.i. norm inequality [15, th. 3.2], also
generalizing the previous inequality in [13, th. 2.3] in two directions.
In the sequel, we will need the following terminology.
Definition 3.2 For an analytic function f on some neighborhood of zero, let
def √ −1
Rf = lim supn→∞ n |cn | denotes the radius of convergence for its Taylor
∞
(power) series (f (z) =) n=0 cn zn , with 0 < Rf ≤ +∞. For A ∈ L2G (, μ,B(H))
and for an analytic function f with non-negative Taylor coefficients, satisfying
) def ) ∗ n 1
r A∗t ⊗At dμ(t) = inf At ⊗At dμ(t) (I ) n ≤ Rf (when transformer
) ∗
n∈N
At ⊗At dμ(t) acts on B(H)) and f (0) > 0, we define its associated generalized
spectral defect operator f,A by
) −1/2
f,A = lim f ρ 2 A∗t ⊗At dμ(t) (I )
def
[s ]
.
ρ31
For a function f : D → C : z $→ ∞ ·
n=0 z (in which case Rf = 1) A ·= f,A
n
was introduced by [21, def. 2.1], while in the case that μ is the counting measure
on N operator f,A was given in [23, def. 2.3]. In both cases above it follows from
) −1/2
[21, lemma 2.2] and [23, rem. 2, rem. 7] that f,A = f A∗t At dμ(t) if
A ··= {At }t ∈ is s.i. and m.c.n.o. family.
The above definition helps us in formulation of the following generalization of
Theorem 3.1, which reformulates [21, th. 3.5]:
Theorem 3.3 Let ϒ be a s.n. function and p ≥ 2. If A∗, B ∈ L2G (, μ,B(H)) are
) )
such that r At ⊗A∗t dμ(t) ≤ 1, r Bt∗ ⊗Bt dμ(t) ≤ 1 and at least one of
families {At }t ∈ or {Bt }t ∈ is a m.c.n.o. family, then for X ∈ Cϒ (p)(H)
$
XB
A∗ ϒ
(p) X− At XBt dμ(t) (p)
.
ϒ
188 D. R. Jocić and M. Lazarević
)
Similarly, if p ≥ 2 and ∞ ∗ ∗ 2
n ||Bt1 · · · Btn h|| dμ (t1 , . . . , tn ) < +∞ for all
n
) n=1
h ∈ H, then r Bt ⊗Bt∗ dμ(t) ≤ 1 and for all X ∈ Cp(H)
$
1− p2 −2
A∗ XB p
≤ X− At XBt dμ(t) B ∗p .
p
)
If ∞ 2 ∗ ∗
n=1 n ||At1· · ·Atn h|| + ||Bt1 · · ·Btn h||
2 dμn (t , . . . , t ) < +∞ for all
1 n
) ∗
h ∈ H and p, q, r ≥ 1 are such that p = q + r , then r At ⊗At dμ(t) ≤ 1,
2 1 1
)
r Bt ⊗Bt∗ dμ(t) ≤ 1 and for all X ∈ Cp(H)
$
1− 1 1− 1r −1 −1
A∗ q XB p
≤ A q X − At XBt dμ(t) B ∗r .
p
In this section we will show that Aczél–Bellman type norm inequalities in the
previous section are just a part of a wider family of Cauchy–Schwarz norm
inequalities.
Theorem 4.1 Let ,ϒ be s.n. functions, let p 2, let f be an analytic function
with non-negative Taylor coefficients and X ∈ C (H). Then
9 9
∞ : ∞ : ∞
: :
f An⊗Bn X ; f An⊗An (I ) X;f
∗ Bn⊗Bn∗ (I ) ,
n=1 n=1 n=1
∞
both {An }∞
if ∗ ∞
n=1 and {Bn }n=1 are s.s. families such that
∗
n=1 An An < Rf ,
∞ ∗
n=1 Bn Bn < Rf and one of additional pair of conditions are satisfied:
(p) ∗
(a) ··= ϒ and at least one of {An }∞ ∗ ∞
n=1 or {Bn }n=1 is a m.c.n.o. family,
(b) ||·|| ··= |||·||| and both {An }∞
n=1 and {B ∗ }∞ are m.c.n.o. families.
n n=1
(c) If ··= ϒ , {A∗n }∞ ∞
(p)
{B }
∞ ∗ n n=1 are s.s. families, such that
n=1 and
∞ ∗ <R ,
A
n=1 n n A f n=1 Bn Bn < Rf , then:
9
∞ : ∞ ∞
:
1
2
f An ⊗Bn X ≤ ;f A∗n An X f Bn∗ ⊗Bn (I )
(p) (p)
n=1 ϒ n=1 ϒ n=1
∞
(d1) if f Bn∗ ⊗Bn (I ) is invertible in addition to conditions (c) and (c1);
n=1
9
∞ − 12 ∞ : ∞
:
f An ⊗A∗n (I ) f An ⊗Bn X ≤ X;f Bn∗ Bn
(p) (p)
n=1 n=1 ϒ n=1 ϒ
∞
(d2) if, in addition to conditions (c) and (c2), f An ⊗A∗n (I ) is invertible.
n=1
190 D. R. Jocić and M. Lazarević
Theorem 4.1(a) (resp. (b)) is a reformulation of [23, th. 2.4(a)] (resp. [23,
th. 2.2]), while cases (c1) and (c2) are based on [23, th. 2.4(b)]. The case (d1)
(resp. (d2)) follows by applying the inequality (8) in Theorem 2.3(c2) (resp.
the inequality
<∞ (10) in Theorem 2.3(e2))6 to the counting measure
7 μ on ··=
√ √ √
{0} m=1 N and to families { c0 I } ∪ cm An(m) · · · An(m) 1≤k≤m and { c0 I } ∪
m
m 1 (m)
1≤nk
6√ 7
cm Bn(m) · · · Bn(m) 1≤k≤m , combined with the arguments already used in the proof
1 m (m)
1≤nk
of [23, th. 2.2].
For two-side multipliers A⊗B Theorem 4.1 remains valid under some additional
subnormality conditions for A and B ∗, as stated in [18, th. 3.2]:
def
Theorem 4.2 Let , ϒ be s.n. functions, p ≥ 2, f (z) = ∞ n
n=0 cn z be analytic
function with non-negative Taylor coefficients cn ≥ 0 and A, B ∈ B(H) have their
1/2
spectra σ (A) ∪ σ (B) ⊂ D 0, Rf . Then for all X ∈ C (H)
∞
1
cn AnXB n = f (A⊗B)X
≤ f (A∗⊗A)(I )X
f (B ∗⊗B)(I ) 2
n=0
∞
1
2
∞
1
2
∗n n ∗n n
= cn A A X cn B B
n=0 n=0
∞
∞
1
2
cn AnXB n ≤ f (A∗A)X
cn B ∗nB n
n=0 n=0
(b) if ··= ϒ
(p)
and A is quasinormal;
∞
1
cn AnXB n ≤ f (A⊗A∗ )(I ) 2 X f (B ⊗ B ∗ )(I )
n=0
∞
1
2
∞
1
2
n ∗n n ∗n
= cn A A X cn B B
n=0 n=0
∞
∞
1
2
cn AnXB n ≤ cn AnA∗n X f (BB ∗)
n=0 n=0
Cauchy–Schwarz Inequalities for i.p.t. Transformers 191
∞
cn AnXB n ≤ f (A∗ ⊗A)(I )X f (B ⊗B ∗ )(I )
n=0
∞
1
2
∞
1
2
∗n n n ∗n
= cn A A X cn B B
n=0 n=0
(p) ∗
(e1) if ··= ϒ and A or B ∗ is subnormal,
(e2) if ||·|| ··= |||·||| and both A and B ∗ are subnormal,
(e3) if ||·|| ··= ||·||1 ;
∞
∞
2
1
(p) ∗
(f1) if ··= ϒ and B ∗ is quasinormal,
(f2) if ||·|| ··= |||·|||, B ∗ is quasinormal and A is subnormal;
∞
∞
1
2
n ∗n
cn A XBn n
≤ f (A∗A)X cn B B
n=0 n=0
(p) ∗
(g1) if ··= ϒ and A is quasinormal,
(g2) if ||·|| ··= |||·|||, A is quasinormal and B ∗ is subnormal;
∞
cn AnXB n ≤ f (A∗A)X f (BB ∗ )
n=0
∞
∞ 2 2
cm A (m) · · · A (m) h + B ∗(m) · · · B ∗(m) h < +∞, then
m=1 (m) (m) n1 nm n1 nm
n1 ,...,nm =1
∞
1− 1 1− 1 −1 −1
f,Aq∗ f An ⊗Bn (X)f,Br ≤ f,A
q r
Xf,B ∗
p
. (18)
p
n=1
192 D. R. Jocić and M. Lazarević
4
∞
∞
∞
∞ 5
(b) If max A∗n An , An A∗n , Bn∗ Bn , Bn Bn∗ < Rf , then
n=1 n=1 n=1 n=1
∞
∞
∞
q−1 1
2q
f An⊗Bn X ≤ f A∗n⊗An f An⊗A∗n (I )
n=1 p n=1 n=1
∞
∞
r−1 1
2r
×X f Bn⊗Bn∗ f Bn∗⊗Bn (I ) , (19)
n=1 n=1 p
∞ 1 1 ∞ ∞ 1 1
2q − 2 2r − 2
f An⊗A∗n (I ) f An ⊗Bn (X) f Bn∗⊗Bn (I )
n=1 n=1 n=1 p
∞
1 ∞
1
2q 2r
≤ f A∗n⊗An (I ) X f Bn⊗Bn∗ (I ) . (20)
n=1 n=1 p
The part (a), including the inequality (18), is a simple reformulation of [23,
th. 2.9], while the inequality (19) in part (b) is also a reformulation of [23, th. 2.8].
∞ ∗ −1/2
According to [23, rem. 7] we can conclude that f,A = f n=1 An⊗An (I )
based on the conditions given in part (b). With analogous formulas for f,A∗ , f,B
and f,B ∗ and by the fact that conditions in (b) provides the fulfilment of all
requirements in (a), then the inequality (19) becomes the proclaimed inequality (20).
The next is a list of selected inequalities from [23, cor. 2.5, cor. 2.6]
Corollary 4.4 Let ϒ be a s.n. function, let p ≥ 2, α, β ∈ [−1, 1], γ 0 and let
A, B ∈ B(H), X ∈ Cϒ (p)(H) and Y ∈ Cϒ (p)∗(H).
(a) If A and B are strict contractions, i.e., max{||A||, ||B ||} < 1, with B being
additionally normal, then
(A⊗B)X+ (I − A ⊗ B) log I − A⊗B X (p)
≤
ϒ
∞
AnA∗n 1/2
X B ∗B + (I − B ∗B) log I − B ∗B ,
n(n − 1) ϒ
(p)
n=2
γ γ
I − A⊗B Y + α I + βA⊗B Y ϒ (p) ∗ ≤
∞
1/2 γ γ
(−1)n+αβ n γn A∗nAn Y I − B ∗B + α I + βB ∗B .
(p) ∗
n=0 ϒ
Cauchy–Schwarz Inequalities for i.p.t. Transformers 193
∞
1/2
1+αβ n
exp(A∗A) + α exp(βA∗A)X (p) n! B ∗nB n ,
ϒ
n=0
∞
1/2
1+αβ n
exp(A∗A) + α exp(βA∗A) Y n! B nB ∗n ,
(p) ∗
n=0 ϒ
n
Ak ⊗B k
X − (I − A⊗B) exp X
k ϒ
(p)
k=1
9
:
:
n
(A∗A)k
≤ ;I − (I − A∗A) exp X
k ϒ
(p)
k=1
n
B ∗k ⊗B k 1/2
∗
× I − (I − B ⊗B) exp (I ) ,
k
k=1
n
Ak ⊗B k
Y − (I − A⊗B) exp Y
k (p) ∗
k=1 ϒ
9
:
:
n
(A∗A)k
≤ ;I − (I − A∗A) exp Y
k
k=1
9
:
:
n
B k ⊗ B ∗k
× ;I − (I − B⊗B ∗ ) exp (I ) .
k ϒ
(p) ∗
k=1
Also, the list of selected inequalities from [18, th. 3.4] is displayed in
Theorem 4.5 Let ϒ be a s.n. function, let p ≥ 2, let α, β ∈ [−1, 1] and let
A, B, C, D, T ∈ B(H), with A, T ∗ being subnormal and C, D ∗ being quasinormal,
such that their spectra satisfy σ (A) ∪ σ (B) ∪ σ (C) ∪ σ (D) ∪ σ (T ) ⊂ D. Then for
all X ∈ Cϒ (p)(H), Y ∈ Cϒ (p)∗(H) and Z ∈ Cϒ (H)
α log I + βC⊗B X − log I − C⊗B X (p)
≤
ϒ
∞
1/2
1−α(−β)n
α log(I + βC ∗ C) − log I − C ∗ C X n B ∗nB n ,
(p)
ϒ n=1
194 D. R. Jocić and M. Lazarević
arcsin(C⊗B)X + α log βC⊗B + I + β 2 C 2 ⊗B 2 X (p)
≤
ϒ
=
arcsin(C ∗ C) + α log βC ∗ C + I + β 2 (C ∗ C)2 X
(p)
ϒ
∞
1/2
(1+(−1)nαβ 2n+1 )(2n)!
× 22n (n!)2 (2n+1)
(B ∗ )2n+1 B 2n+1 ,
n=0
π βπ
tan C⊗B X + α tanh C⊗B X ≤
2 2 ϒ
(p)
=
π ∗ βπ ∗
tan C C + α tanh C C X
2 2 ϒ
(p)
∞
1/2
(22n −1)π 2n−1 (|B 2n |+αβ
2n−1 B )
× 2(2n)!
2n
(B ∗ )2n−1 B 2n−1 ,
n=1
α log I + βA⊗B Y − log I − A⊗B Y ϒ
(p) ∗ ≤
∞
1/2 ∞
1/2
1−α(−β)n 1−α(−β)n
n A∗nAn Y n B n B ∗n ,
(p) ∗
n=1 n=1 ϒ
α log I + βA⊗T Z − log I − A⊗T Z ≤
∞
1/2 ∞
1/2
1−α(−β)n 1−α(−β)n
n A∗nAn Z n T n T ∗n ,
n=1 n=1
arcsin(C⊗D)Z + α log βC⊗D + I + β 2 C 2 ⊗D 2 Z ≤
=
arcsin(C ∗ C) + α log βC ∗ C + I + β 2 (C ∗ C)2 Z
=
× arcsin(DD ∗ ) + α log βDD ∗ + I + β 2 (DD ∗ )2 ,
π βπ
tan C⊗D Z + α tanh C⊗D Z ≤
2 2
= =
π ∗ βπ ∗ π βπ
tan C C + α tanh C C Z tan DD ∗ + α tanh DD ∗ .
2 2 2 2
Cauchy–Schwarz Inequalities for i.p.t. Transformers 195
$ $ $
At XBt dμ(t) − At dμ(t)X Bt dμ(t) ≤
p
$ $ $ q−1 $ 2 1
2 2 2q
A∗t − A∗t dμ(t) dμ(t) At − At dμ(t) dμ(t) X
$ $ $ r−1 $ 2 1
2 2 2r
× Bt − Bt dμ(t) dμ(t) Bt∗ − Bt∗ dμ(t) dμ(t) .
p
For some other types of u.i. norms the counterpart of Theorem 5.1 says:
Theorem 5.2 Let μ be a probability measure on , p ≥ 2, ,ϒ be s.n. functions
and X ∈ B(H).
(a) If A, B ∗ ∈ L2G (, μ,B(H)), then
$ $ $
At XBt dμ(t) − At dμ(t)X Bt dμ(t) ≤ (21)
8$ $ 8$ $
2 2
A∗t At dμ(t) − At dμ(t) X Bt Bt∗ dμ(t) − Bt dμ(t) ,
∗ ∗
(c) If A , B ∈ L2G (, μ,B(H)), then
$ $ $
At XBt dμ(t) − At dμ(t)X Bt dμ(t) ≤ (23)
$ $ $ 8 $
2 1/2 2
At A∗t dμ(t) − A∗t dμ(t) X Bt Bt∗ dμ(t) − Bt∗ dμ(t) ,
As we can confine to the case in which the righthand side of the inequality (21) is
finite, so (21) in the case (a1) is a special case p ··= q ··= r ··= 1 of Theorem 5.1.
The proof of [20, th. 2.1] provides the proof for the inequality (21) in the case (a3),
together with the proof for the case (a2), with the only difference that we now apply
(to the same families) the Cauchy–Schwarz inequality [26, th. 3.1(d)] instead of
[15, th. 3.2]. For σ -elementary transformers the case (a2) was proved earlier in [32,
th. 2.10].
The case (b1) (resp. (c1)) was proved in [32, th. 2.6(21)] (resp. [32, th. 2.6(20)],
while the case (b2) (resp. (c2)) was proved in [32, th. 2.7(28)] (resp. [32, th. 2.7(27)].
Definition 5.3 For a bounded field of operators A ··= {At }t ∈ its radius of
its essential range is given by r∞ (A) = infB∈B(H) sup esst ∈ ||At − B || =
def
infB∈B(H) ||A − B ||∞ = minB∈B(H) ||A − B ||∞ , while its essential diameter is
given by diam∞ (A) = sup esss,t ∈||As − At ||.
def
The inequality (24) ( resp. (25)) was proved in [20, th. 2.8] (resp. [20, cor. 2.9]).
The operator Grüss–Landau inequality is given by:
Theorem 5.5 Let μ be a probability measure on , A∗, B ∈ L2G (, μ,B(H)), X ∈
B(H) and η ∈ [0, 1]. Then
$ $ $ 2η
At XBt dμ(t) − At dμ(t)X Bt dμ(t) ≤ (26)
$ $ 2 η $ $ 2 η
At A∗t dμ(t) − A∗t dμ(t) Bt∗ X∗XBt dμ(t) − X Bt dμ(t) .
Specially, if A ··= {At }t ∈ and B ··= {Bt }t ∈ are (bounded) self-adjoint fields
satisfying C ≤ At ≤ D and E ≤ Bt ≤ F for all t ∈ and some bounded self-
adjoint operators C, D, E, F, such that C, D (resp. E, F ) commutes with At (resp.
Bt ) for all t ∈ , satisfying CD = DC and EF = F E, then
$ $ $ 2η
At Bt dμ(t) − At dμ(t) Bt dμ(t) ≤
$ $ η $ η $ η
D− At dμ(t) At dμ(t) − C F− Bt dμ(t) Bt dμ(t) − E
1
≤ 2η ||D − C ||2η (F − E)2η . (27)
4
The inequality (26) (resp. (27)) was proved in [32, th. 2.1] (resp. [32, th. 2.2]).
The refined Grüss–Landau operator and norm inequalities are presented in:
N
Theorem 5.6 If {α1 , . . . , αN } are in (0, 1], satisfying n=1 αn = 1 for some
N ∈ N, and if X, {A1 , . . . , AN } and {B1 , . . . , BN } are in B(H), then for all
N −1 ∗
N 2 12
c≥ n=1 αn An An − n=1 An >0
N
N
N 2
αn−1 A∗n XBn − A∗n X Bn +
n=1 n=1 n=1
N
N
2 1 −1
−1
Am −αn−1 An ) cI + c2 I − αn−1 A∗n An +
2
αm αn (αm An
1≤m<n≤N n=1 n=1
198 D. R. Jocić and M. Lazarević
N
N
N 2
× αn−1 A∗n XBn − A∗n X Bn −1
− cX(αm Bm − αn−1 Bn )
n=1 n=1 n=1
N
N
2
= c2 αn−1 Bn∗ X∗XBn − X Bn . (28)
n=1 n=1
N
N
N p
αn−1A∗n XBn − A∗n X Bn ≤ (29)
(q )
n=1 n=1 n=1 ϒ
N
N
N p
αn−1A∗n XBn − A∗n X Bn +
n=1 n=1 n=1
p p
N
N
2 1 −1
−1
Am − αn−1An ) cI + c2 I − αn−1A∗n An +
2
αm2 αn2 (αm An
1≤m<n≤N n=1 n=1
N
N
N p
× αn−1A∗n XBn − A∗n X −1
Bn −cX(αm Bm − αn−1Bn ) q
(p)
n=1 n=1 n=1 ϒ
p
N
N
N 2
≤c p
αn−1 Bn∗ X∗XBn − Bn∗ XX ∗
Bn . (30)
q
n=1 n=1 n=1 ( )
ϒ 2
If {Bn }Nn=1 is additionally a m.c.n.o. family, then the expression appearing in (30)
N 2 1/2 p
further estimates by cp X Nn=1 αn−1 Bn∗ Bn − n=1 Bn ϒ
(q ) .
The identity (28) is a subject of [22, lemma 2.1], while the chain of inequalities (29)–
(30) was proven in [22, th. 2.2] and was further generalized in [22, th. 2.3].
√
I − A∗A f (I )Y − f (A⊗B)(Y ) ≤
I − f (A)∗ f (A)(Y − AYB)(I − B ∗B)−1/2 I − f (B)∗ f (B)
1/2
,
(36)
The inequality
• (31) in the case (a1) (resp. (a2) and (a3)) was proved in [24, th. 2.1(a3)] (resp.
[24, th. 2.1(a2)] and [24, th. 2.1(a1)]);
• (32) in the case (a1) (resp. (a2) and (a3)) was proved in [24, th. 2.1(b3)] (resp.
[24, th. 2.1(b2)] and [24, th. 2.1(b1)]);
• (33) in cases (b1) and (b2) is just the inequality (6) in [25, th. 2.1(a4)];
• (34) in cases (b1) and (b2) is a reformulation of the inequality (9) in [25,
th. 2.1(b4)];
• (35) in cases (c1) and (c2) is just the inequality (7) in [25, th. 2.1(a4)];
• (36) in cases (c1) and (c2) is a reformulation of the inequality (10) in [25,
th. 2.1(b4)].
Parts of Theorem 6.1 remains valid for the subclass of functions in disc algebra
A(D) possessing non-negative Taylor coefficients, if at least one of operators A or
B ∗ is quasinormal, as presented in [18, th. 3.6]:
def ∞
Theorem 6.2 Let ϒ be a s.n. function, p ≥ 2, f (z) = n
n=0 cn z be an analytic
∞
function with non-negative Taylor coefficients cn ≥ 0, 0 < n=0 cn < +∞ and let
A, B, X, Y ∈ B(H) be such that A, B ∗ are contractions and at least one of them is
quasinormal. If AY − Y B ∈ Cϒ (p)∗(H), then
√ √
I − A∗A f (A)Y − Yf (B) I − BB ∗ ϒ (p) ∗ ≤
f (1)I − |f (A)|2 /f (1)(AY − Y B) f (1)I − |f (B ∗ )|2 /f (1) ϒ
(p) ∗ .
− 12 1
A+A∗
B +B ∗
2
ϕ ϕ(A)X − Xϕ(B) ≤ (AX − XB)ϕ ,
2 2
(38)
− 12 1
B +B ∗ A+A∗ 2
ϕ(A)X − Xϕ(B) ϕ ≤ ϕ (AX − XB) ,
2 2
(39)
minσ ( A∗ +A ) ϕ > 0. Thus, for every h ∈ H there is a positive measure μh , such that
> A∗+A2 ? ) ∗
ϕ ( 2 )h, h = σ ( A∗ +A ) ϕ (λ)dμh (λ) ≥ mA ||h||2 , implying that ϕ A 2+A ≥
∗ 2
mA I, so ϕ A 2+A is invertible, as proclaimed. Similarly we also conclude that
∗
ϕ B+B
2 is positive and invertible.
• The inequality (37) in the case (a1) (resp. (a2) and (a3)) is stated at the end of
[24, th. 3.1(c)] (resp. is given by the inequality (37) in [24, th. 3.1(c)] and by the
inequality (39) in [24, th. 3.1(d)]). The rest of the statement in the part (a) is based
on estimates
A∗+A A∗ +A−1 A∗ +A B+B ∗ B+B ∗−1 B+B ∗
ϕ 2 ≤ 2 ϕ 2 , ϕ 2 ≤ 2 ϕ 2 , (41)
presented in the proof of [24, th. 3.1(c),(d)], additionally combined with the (double)
monotonicity property [26, (1)].
• The inequality (38) in the case (b1) (resp. (b2)) is stated at the end of [24,
th. 3.1(a)] (resp. is given by the inequality (35) in [24, th. 3.1(a)]).
• The inequality (39) in the case (c1) (resp. (c2)) is stated at the end of [24,
th. 3.1(b)] (resp. is given by the inequality (36) in [24, th. 3.1(b)]).
• The first inequality in (40) in the case (d) is again based on the inequalities in (41),
combined with application of the monotonicity property [26, (1)].
The proof of the second inequality in (40) in the case (d1) for ϕ ∈ P(−1, 1)
relays on the estimate
− 21 − 12
A∗+A B+B ∗
ϕ 2 ϕ(A)X − Xϕ(B) ϕ 2 =
$ 1 − 12 − 12
A∗+A B+B ∗
ϕ (0) ϕ 2 (I − tA)−1 (AX − XB)(I − tB)−1 ϕ 2 dμ(t)
−1
$ 1
1
∗ − 21 ∗ − 12 2
≤ ϕ A 2+A
ϕ (0) (I − tA) −1
(I − tA ) ∗ −1
dμ(t)ϕ A 2+A (42)
−1
$ 1
1
− 21 − 21 2
B+B ∗ B+B ∗
×||AX − XB || ϕ 2 ϕ (0) (I − tB ∗ )−1(I − tB )−1 dμ(t)ϕ 2 ,
−1
∗ −1/2
ϕ A 2+A X) instead of X. The second inequality in (40) in the case (d3) is given
by the inequality (42) in [24, th. 3.1(d)].
The previous Theorem 7.2 provides the following generalization of some well
known norm inequality to normal operators, which are either accretive or they have
strictly contractive real parts.
Corollary 7.3 Let A, B ∈ B(H) be normal, such that AX − XB ∈ C|||·|||(H) for
some X ∈ B(H).
(a) If A and B are accretive operators, then for all θ ∈ [0, 1]
A∗ +A1−θ B+B ∗1−θ
2
2 (Aθ X − XB θ ) 2
2 ≤ θ |||AX − XB |||, (43)
√
A∗ + A log(A)X − X log(B) B + B ∗ ≤ 2|||AX − XB |||. (44)
in [30, (4.1)] and another important generalization of the first inequality in [8,
Hilfssatz 3.] and Heinz-type means inequality [9, (5.3.1)] from positive to self-
adjoint operators is given by the special case p ··= 1 of [12, lemma 3.2]:
Lemma 7.4 For self-adjoint A, B ∈ B(H) and X ∈ B(H), p ≥ 1 and u.i. norms
|||·|||, the function f : [0, p] → [0, +∞], defined by
is convex and symmetric on [0, p], non-increasing on [0, p/2] and non-decreasing
on [p/2, p], implying that for all s ∈ [0, p]
The inequality (46) follows from the first part of the above lemma, as f (s) ≤
max[0,p] f = f (0) = f (p) = |A|p−1AX + XB|B|p−1 for all s ∈ [0, p]. The
1+2β
inequality (45) follows from inequality (46) by taking p ··= 1, A ··= H 2s
and
1+2β
B ··= −K 2s
for s ∈ (0, 1). Inequality (46) plays the important role in the proofs
of the norm inequality for self-adjoint derivations
in [12, th. 3.1] and the perturbation norm inequality in [12, th. 3.3]:
It is also noted in [24, rem. 3] that the special case of (44) for strictly positive
operators A and B can also be derived from geometric-logarithmic mean inequality
[30, (5.2)].
R+
206 D. R. Jocić and M. Lazarević
will denote the operator valued Laplace transform of f (in A), and, similarly, if a
function t $→ e−t A Xe−t B f (t) is Gel’fand integrable on R+ , then
$
LB(B(H)) f (A⊗I + I ⊗B)(X) = e−t A Xe−t B f (t) dt
def
R+
will denote the Laplace transformer of f (in a generalized derivation A⊗I + I ⊗B,
evaluated in X).
In the sequel, we will use the simplified notation Lf instead of both LB(H) f and
LB(B(H)) f, as their (exact) meaning will be clear from the context. As it is usual
for linear transformations, we will also often write Lf (A⊗I + I ⊗B)X instead of
Lf (A⊗I + I ⊗B)(X), except in the case X ··= I in which the brackets will not be
omitted.
For Laplace transformers we have the following reformulation of [28, th. 2.2]:
Theorem 8.2 Let ,ϒ be s.n. functions, p ≥ 2, A, B ∈ B(H), f, g : R+ → C be
some Lebesgue measurable functions and X ∈ C (H).
) ∗
(a) If R+ (||e−t Ah||2 |f (t)|2 + ||e−t B h||2 |g(t)|2 )dt < +∞ for all h ∈ H, then
1 1
L(fg)(A,B )X
≤ (L|f |2(A∗,A )(I )) 2 X(L|g|2(B,B ∗)(I )) 2
, (47)
1 1
L(fg)(A,B )X
≤ (L|f |2 (A∗,A )(I )) 2 X
L|g|2 (B ∗,B )(I ) 2 ,
(48)
∗ ∗
(c) If R+ (||e−t A h||2 |f (t)|2 + ||e−t B h||2 |g(t)|2 )dt < +∞ for all h ∈ H, then
1 1
L(fg)(A,B )X
≤ L|f |2 (A,A∗)(I ) 2 X(L|g|2 (B,B ∗)(I )) 2
,
(49)
X − L(fg)(A,B )X
≥
)
(b) If R+ (||e−t Ah||2 |f |2−2α ∗ |g|2−2β (t) + ||e−t B h||2 |f |2α ∗ |g|2β (t)) dt < +∞
for all h ∈ H, then
∗ ∗
(c) If R+ (||e−t A h||2 |f |2−2α ∗ |g|2−2β (t) + ||e−t B h||2 |f |2α ∗ |g|2β (t))dt < +∞
for all h ∈ H, then
As L |f |2−2α ∗ |g|2−2β (A∗,A )(I ) = L |f |2−2α ∗ |g|2−2β (A∗ + A) if A is normal
and L |f |2α ∗ |g|2β (B,B ∗)(I ) = L |f |2α ∗ |g|2β (B + B ∗) if B is normal,
inequalities (51), (52) and (53) can be simplified in those cases.
Jocić et al. [27, lemma 2.5] guarantees the correctness of the following
Definition 8.5 For accretive operators A, B ∈ B(H) and X ∈ B(H) we define
√ √
V −(A, B)X = A∗ + Ae−T A Xe−T B B + B ∗,
def
[s ] lim
T →+∞
−
(A, B)X = lim e−T A Xe−T B .
def
[s ]
T →+∞
The previous Definition 8.5 is useful for the formulation of the next:
Theorem 8.6 Let α ∈ (0, 1), p ≥ 2, ϒ be a s.n. function and A, B, X ∈ B(H),
such that A and B are accretive.
(a) If AX + XB ∈ C1(H), then
√ √
A∗ + AX B + B ∗ − V −(A, B)X 1 ≤ (54)
1/2 1/2
I − −(A∗, A)I (AX + XB) I − −(B, B ∗)I 1 ≤ ||AX + XB ||1 .
√ √
Consequently, the left side of (54) can be replaced by A∗ + AX B + B ∗ 1
under additional conditions required in (b1) or (b2) of [27, lemma 2.7].
Cauchy–Schwarz Inequalities for i.p.t. Transformers 209
X(B + B ∗)1/2 ϒ
(p) ≤ (A∗ + A) −1/2 (AX + XB)PR(B+B ∗) (p)
ϒ
(A∗ + A)1/2 X ϒ
(p) ≤ PR(A∗+A) (AX + XB)(B + B ∗)−1/2 (p)
ϒ
(c) Specially, in the case of the Hilbert-Schmidt class C2 (H) (i.e. if ··= and
p ··= 2), the requirement of normality for A or B can be omitted, with (55)
and (56) still remaining valid, if PR(B+B ∗) (resp. PR(A∗+A) ) is replaced by
1/2 1/2
I − −(B, B ∗)I (resp. I − −(A∗, A)I ).
(d) If at least one of operators A or B is additionally normal and AX + XB ∈
Cϒ (p)∗(H), then
√ √
A∗ + AX B + B ∗ ϒ
(p) ∗ ≤
1 1
I − −(A∗, A)I (AX + XB) I − −(B, B ∗)I 2
2
ϒ
(p) ∗ ≤ ||AX + XB ||ϒ (p) ∗ .
√
p A B
(A∗ + A)1/pX(B + B ∗)1/p ≤ p
p p (2 − 2α)(2α) X+X .
p p
210 D. R. Jocić and M. Lazarević
For the proof see [27, th. 2.9] and its proof, as well as [27, rem. 2.2]. In addition, for
the inequality (57) see [27, th. 2.8(d)] and its proof.
Thus, Theorem 8.6 provides extensions of the Young’s norm inequality [15,
cor. 4.1] in various direction, where the norm inequality (58) in part (e) extends
it from positive definite to normal accretive operators.
The next Theorem 8.7 represents the main part of [27, th. 2.10].
Theorem 8.7 Let X ∈ C|||·|||(H), (αn )∞
n=1 be a sequence
∞
∞ in (0, 1), (tn )n=1 be a
summable sequence in (0, +∞), f : C → C : z $→ n=1 (1 + tn z) and let each
of the absolutely summable (in B(H)) families {An }∞ ∞
n=1 and {Bn }n=1 consists of
normal, accretive and mutually commuting operators. Then
∞
1/2 ∞
1/2
|||X||| ≤ (I + A∗n + An ) X (I + Bn + Bn∗)
n=1 n=1
∞
≤ (I ⊗I + An ⊗I + I ⊗Bn )X .
n=1
Corollary 8.8 was earlier presented in [27, cor. 2.13]. For its connections with
[31, prop. 21(1)] see [27, rem. 2.5].
Cauchy–Schwarz Inequalities for i.p.t. Transformers 211
A2X + 2AXB + XB 2
,
X − 2CXD + C 2XD 2
,
M
M 1/2
N
N 1/2
∗m M −m
m A A X n B N −n B ∗n
m=0 n=0
√ +N )/2
(M
M+N M+N
−k
≤ M+N
(M−1)!(N−1)! 2 A 2 XB k ,
k
2 −1 !
k=0
We start this section with the shortened version of [17, th. 2.5], which provides
Minkowski inequality for p-modified u.i. norms.
Theorem 9.1 Let p ≥ 2, {Bn }n∈N be a sequence in B(H), such that n∈N Bn
in the strong operator topology and let {βn }n∈N be a sequence in (0, +∞)
converges
satisfying n∈N βn < +∞. Then
1
p p p p 1− p1 1
Bm Bn p
1−p p
Bn + 12 βm2 βn2 − ≤ βn βn |Bn |p .
βm βn
n∈N m,n∈N n∈N n∈N
If 0 = Bn ∈ C|||·||| (H) for all n ∈ N and n∈N |||Bn |||(p) < +∞, then
(p)
1
p p p
Bm Bn p p
Bn ≤ Bn + 12 |||Bm |||(p)
2
|||Bn |||(p)
2
−
(p)
|||Bm |||(p) |||Bn |||(p)
n∈N n∈N m,n∈N
1− p1
1
p
≤ |||Bn |||(p) |||Bn |||(p)
1−p
|Bn |p ≤ |||Bn |||(p) . (61)
n∈N n∈N n∈N
Here we used the notation |||·|||(p) = ||·||ϒ (p) , where s.n. function ϒ is uniquely
def
determined by ||·||ϒ = |||·|||, as well as the monotonicity property for u.i. norms to
get the first (of three) inequality in (61).
We also have the following refined norm inequality for bi-infinite operator
matrices, given as a part of [17, th. 2.8]:
Theorem 9.2 If 2 ≤ p < +∞, then for all [Am,n ]m,n∈Z ∈ C|||·||| (2Z (H))
(p)
2 p
p p
[Am,n ]m,n∈Z ≤ [Am,n ]m,n∈Z + 12 [Ak,n ]n∈Z 2 [Al,n ]n∈Z 2
(p) (p) (p)
k,l∈Z
2 2 p 2
[Ak,n ]n∈Z [Al,n ]n∈Z p 2
× 2
− 2
≤ [Am,n ]n∈Z ≤
[Ak,n ]n∈Z [Al,n ]n∈Z m∈Z
(p)
(p) (p)
2
p p |A∗k,m |2 |A∗l,m |2 p p
[A∗n,m ]n∈Z + 12 |||A∗k,m |||(p) |||A∗l,m |||(p)
p 2 2
−
|||A∗k,m |||(p)
2 |||A∗l,m |||(p)
2
m∈Z k,l∈Z
≤ |||Am,n |||(p)
2
.
m,n∈Z
Cauchy–Schwarz Inequalities for i.p.t. Transformers 213
p
If q ≥ p = if C|||·||| (2Z (H)) is the associated ideal of compact operators
def
p−1 , (q)
∗ ( 2 (H)), then for all [A
and |||·|||(q)∗ is a norm on its dual space C|||·||| Z m,n ]m,n∈Z ∈
(q)
∗ (2 (H))
C|||·||| Z
(q)
p
p
p
Am,n m,n∈Z ≥ Am,n ≥ Am,n .
(q)∗ n∈Z (q)∗ (q)∗
m∈Z m,n∈Z
N
z ∈ C and n=0 cn AnXB N −n ∈ C|||·|||(H), then
N
N N
2 |cN |
N
|sin(θn /2)|rnαn AN − n=1 αn XB n=1 αn
n=1
N
N
N
def
N
2
(d) If a sequence {cn }Nn=−N in C satisfies cn eint = eit − rn eiθn for all
n=−N n=1
t ∈ R, cN = 0 and Nn=−N cn AN +nXB N −n ∈ C|||·|||(H), then
N
N
N N
(e) If n=0 n einη+i(N −n)θ AnXB N −n ∈ C|||·|||(H), then
N N
2N cos η−θ AN −
N
n=1 αn XB n=1 αn
2
N
N
N
≤ (2 − 2αn )(2αn ) n einη+i(N −n)θAnXB N −n and
n=1 n=0
N +1 N N
sin 2 η◦
η AN − n=1 αn XB n=1 αn
sin 2◦
N
N
N inη◦ AnXB N −n
if n=0 e ∈ C|||·|||(H) for some η◦ ∈ (0, 2π).
2N
(f) If θ◦ ∈ (−π, π) and 2N i(N −n)θ◦ AnXB 2N −n − T AN XB N ∈ C (H) for
n e |||·|||
n=0
2N
0 ≤ T < 22N cos2N θ2◦ , then 2N
n ei(2N −n)θ◦AnXB 2N −n ∈ C|||·|||(H) and
n=0
2N
2N
n ei(N −n)θ◦ − T ANXB N ≤
n=0
2N
2N
1− θ
T
n ei(N −n)θ◦AnXB 2N −n ≤
22N cos2N 20
n=0
2N
2N
n ei(N −n)θ◦AnXB 2N −n − T ANXB N .
n=0
Some relations between Young and Heron means norm inequalities are investi-
gated in [27, cor. 2.12], which we present here in the next reduced form:
Corollary 9.4 A, B, X ∈ B(H), A ≥ 0, B ≥ 0, 0 ≤ β < 1/2 ≤ α ≤ α , such
√ If √
that (1 − α ) AX B + α (AX + XB)/2 ∈ C|||·|||(H), then
√ √ $
β 1 1−β−t t−β t−β 1−t−β β 1
AX B ≤ 2(1−1−2β) A 2 + 4 A 2 XB 2 +A 2 XB 2 dt B 2 +4 ≤
[β,1−−β]
$
3 1 1 3
A 4 − 2 XB 4 + 2 + A 4 + 2 XB 4 − 2 dt ≤
t t t t
1
2(1−−2β)
[β,1−−β]
3 β β β β
4−2
XB 4 + 2 + A 4 + 2 XB 4 − 2 ≤
1 1 3
1
2 A
1−β 1 1 β β 1 1 1−β
1
4 A 2 A 2 X + XB 2 B 2 + A 2 A 2 X + XB 2 B 2 ≤
Cauchy–Schwarz Inequalities for i.p.t. Transformers 215
√ √
1
4 AX + 2 AX B + XB ≤
√ √ √ √ α
(1 − α) AX B + α2 (AX+XB) ≤ (1 − α ) AX B + 2 (AX+XB) .
For some additional insight in this topics see also [4, 11], [27, rem. 2.3,rem. 2.4],
[29, 34] and references therein.
w)
∗
where A∗t Bt dμ(t) denotes w∗ or Gel’fand integral of B(H) valued w∗ (or
Gel’fand) integrable function A∗B : → B(H) : t $→ A∗t Bt . A wider class of
semi-inner product modules is treated in [15, th. 2.1] and [20, lemma 1.3]. Thus,
) ∗
w∗
) ∗
w∗
At ⊗Bt dμ(t) : B(H) → B(H) : X $→ At XBt dμ(t) = A, X·B, where
X·B : → B(H) : t $→ XBt , explaining the origin of the name for inner product
type transformers
considered in this and previous papers.
If M,·,· is a semi-inner product module over a C∗ -algebra A, then Cauchy–
Schwarz inequality asserts that
2
A, B = B, AA, B ≤ ||A, A||B, B for all A, B ∈ M. (63)
2
A, X·B = X·B, AA, X·B ≤ ||A, A||X·B, X·Bfor all A, B ∈ M, X ∈ A,
which for M ··= L2G (, μ,B(H)) becomes the Cauchy–Schwarz inequality (2).
In [1] semi-inner product A-modules are considered with respect to alternated
semi-inner products. For an arbitrary C ∈ M another semi-inner product on
M defines correctly by ·,·C : M×M → A : (A, B) $→ ||C, C||A, B −
A, CC, B. By applying (63) to X·B instead of B and to ·,·C instead of ·,·
216 D. R. Jocić and M. Lazarević
gives
2
||C, C||A, X·B − A, CC, X·B
2 2
≤ ||C, C||A, A − C, A ||C, C||X·B, X·B − C, X·B (64)
for all A, B ∈ M and X ∈ A. A special case M ··= L2G (, μ,B(H)) for a
probability measure μ on , A∗, B ∈ L2G (, μ,B(H)) and Ct ··= I for all t ∈
provides the Grüss-Landau operator inequality (26) for η ··= )1, which then implies
the validity of Theorem 5.5 for all η ∈ [0, 1]. If in addition At dμ(t) = 0 then
we have the strengthened Cauchy–Schwarz inequality
) 2 ) ∗
) ∗ ∗
) 2
At XBt dμ(t) ≤ At At dμ(t) Bt X XBt dμ(t) − X Bt dμ(t) ,
)
which improves the inequality (2) in Theorem 2.1 if At dμ(t) = 0. Another
special case X ··= I coincides with the inequality [1, (2.7)]. More generally, in
the special case when C, A = 0 the inequality (64) provides i.t.p. transformer’s
Ostrowski inequality on a semi-inner product over C∗-algebra
2 ||A,A|| 2
A, X·B ≤ ||C,C|| ||C, C||X·B, X·B − C, X·B , (65)
which also generalizes the inequality (63). Again, the special case X ··= I coincides
with the inequality [1, (2.8)]. Alternatively, if C, B = 0 and X ··= I, then (64)
implies another Ostrowski inequality on C∗ -modules
2 |C,A|2
A, B ≤ A, A − ||C,C|| B, B, (66)
Here, C#D = C 1/2 (C −1/2 DC −1/2 )1/2 C 1/2 denotes the operator geometric mean
def
which applied to X·B instead of B for X ∈ A leads to a refined version for the
inequality (64).
Like all C∗ -algebras, B(H) itself represents a Hilbert C∗ -module over B(H) with
inner products A, B = A∗B, for all A, B ∈ B(H), so the special case T ··= I
def
in (67) says |A B| ≤ (V ∗A∗AV )#B ∗B, which generalizes the matrix inequality
∗
Acknowledgments Authors were partially supported by MPNTR grant No. 174017, Serbia
References
27. D.R. Jocić, D. Krtinić, M. Lazarević, Extensions of the arithmetic-geometric means and
Young’s norm inequalities to accretive operators, with applications. Linear Multilinear Algebra
(2021). https://doi.org/10.1080/03081087.2021.1900049
28. D.R. Jocić, D. Krtinić, M. Lazarević, Laplace transformers in norm ideals of compact
operators. Banach J. Math. Anal. 15, 67 (2021). https://doi.org/10.1007/s43037-021-00149-
3
29. R. Kaur, M.S. Moslehian, M. Singh, C. Conde, Further refinements of the Heinz inequality.
Linear Algebra Appl. 447, 26–37 (2014)
30. H. Kosaki, Positive Definiteness of Functions with Applications to Operator Norm Inequalities.
Memoirs of the American Mathematical Society, vol. 212, no. 997 (American Mathematical
Society, Providence, 2011)
31. G. Larotonda, Norm inequalities in operator ideals. J. Funct. Anal. 255, 3208–3228 (2008)
32. M. Lazarević, Grüss-Landau inequalities for elementary operators and inner product type
transformers in Q and Q* norm ideals of compact operators. Filomat 33, 2447–2455 (2019)
33. R. Nakayama, Y. Seo, R. Tojo, Matrix Ostrowski inequality via the matrix geometric mean. J.
Math. Inequal. 14, 1375–1382 (2020)
34. J. Zhao, J. Wu, Some operator inequalities for unitarily invariant norms. Ann. Funct. Anal. 8,
240–247 (2017)
Norm Estimations for the Moore-Penrose
Inverse of the Weak Perturbation
of Hilbert C ∗ -Module Operators
LZ,T = ET T † + IK − T T † , RF,T = F T † T + IH − T † T .
1 Introduction
Let Mm×n (C) be the set of all m × n complex matrices, In and 0n be the identity
matrix and zero matrix in Mn (C), respectively. For every A ∈ Mm×n (C), let A† and
A denote the Moore-Penrose inverse and the spectral norm of A, respectively.
C. Fu
Health School Attached to Shanghai University of Medicine & Health Sciences, Shanghai, PR
China
D. Du · L. Huang · Q. Xu ()
Department of Mathematics, Shanghai Normal University, Shanghai, PR China
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 221
R. M. Aron et al. (eds.), Operator and Norm Inequalities and Related Topics,
Trends in Mathematics, https://doi.org/10.1007/978-3-031-02104-6_7
222 C. Fu et al.
Given any M, T ∈ Mm×n (C), if ran(M) = ran(T ), then clearly there exist two
matrices E ∈ Mm (C) and F ∈ Mn (C) such that
where LE,T and RF,T are given by (13) below, in which H = Cn and K = Cm .
Based on the expression (2), the terms of the strong perturbation and the weak
perturbation were introduced in [21], through which representations for M † as
well as norm estimations for M † − T † were carried out therein. Recently, another
expression of M was introduced in [17] as
M = ET T † · T · (F T † T )∗ .
As a result, the representation theory for M † obtained in [21] has been improved in
[17], which leads to some new norm estimation for M † − T † in [4].
The Moore-Penrose inverse associated to the rank-preserving perturbation and
the stable perturbation is considered originally for the additive perturbation [7,
11, 13, 18, 19], which can also be dealt with in the multiplicative perturbation
case. Given a multiplicative perturbation (12) in the matrix case, as shown in
Norm Estimations for the M-P Inverse of the Weak Perturbation 223
[4, Lemma 2.1 and Corollary 2.1], the relationship between various types of
multiplicative perturbations for matrices can be figured out as follows:
M is a stable perturbation of T
⇑
M is a weak perturbation of T
⇑
M − T · T † < 1 #⇒ M is a strong perturbation of T
>
M is a weak perturbation of T and ran(M) = ran(T ).
The diagram above, together with [4, Examples 2.1 and 2.2], indicates that the weak
perturbation is actually parallel to the rank-preserving perturbation. It is notable that
much progress has been made on norm estimations for M † − T † in the case of
rank-preserving perturbation, yet little has been done up to now in the case of the
weak perturbation. The purpose of this paper is, in the general setting of Hilbert
C ∗ -module operators, to make some generalizations of the main results originally
obtained in [4, 17] for matrices.
Let A be a C ∗ -algebra, H and K be two Hilbert A -modules, and M be a
weak perturbation of T ∈ L(H, K) given by (12). Formulas for M † , MM † and
M † M are derived in Theorem 4.5, hence a generalization of [17, Corollary 3.6] is
obtained from the matrix case to the case of Hilbert C ∗ -module operators. Based on
these formulas, some norm computations of the associated operators can be carried
out in Lemmas 5.1 and 5.2 by using certain C ∗ -algebraic techniques employed in
[4]. Consequently in Theorem 5.3, an elegant estimation is derived in the weak
perturbation case as
6 7
max MM † − T T † , M † M − T † T ≤ T † · M − T . (3)
Thus, a generalization of [4, Theorem 3.1] is obtained from the matrix case to
the case of Hilbert C ∗ -module operators. Note that [4, Theorem 3.2] indicates an
interesting phenomenon that the similar estimation
6 7
max MM † − T T † , M † M − T † T ≤ M † · M − T (4)
may be false for a general weak perturbation. It is gratifying that (4) is always
true for every strong perturbation (see Theorem 5.4 for the details). So in the
strong perturbation case, another elegant estimation for M † − T † is derived
in Theorem 5.6 by using both (3) and (4). This shows theoretically that some
well-known results, associated to the stable additive perturbation satisfying a norm
inequality (28) [11, 13, 18, 19], have been generalized in this paper to the strong
224 C. Fu et al.
perturbation case, since as is shown in Sect. 3.2, every stable additive perturbation
is essentially a strong perturbation whenever this widely used norm inequality
is satisfied. Furthermore, some new results concerning the strong perturbation of
operators are also obtained; see Theorem 5.7 for the details.
The paper is organized as follows. In Sect. 2, we recall some basic knowledge
about the Moore-Penrose inverse of Hilbert C ∗ -module operators. In Sect. 3, we
study the relationship between various types of multiplicative perturbations. In
Sects. 4 and 5, we focus on the study of representations and norm estimations for the
Moore-Penrose inverse associated to the multiplicative perturbation, respectively.
The notations of “⊕” and “ ” are used in this paper with different meanings for
the sake of reader’s convenience. Given Hilbert A -modules H1 and H2 , let
4 5
H1 ⊕ H2 = (h1 , h2 )T : hi ∈ Hi , i = 1, 2 ,
On the other hand, if both H1 and H2 are closed submodules of a Hilbert A -module
H such that H1 ∩ H2 = {0}, then we write
H1 H2 = {h1 + h2 : hi ∈ Hi , i = 1, 2} .
Definition 2.1 ([12, 15]) Let A ∈ L(H, K). The Moore-Penrose inverse of A,
written A† , is the unique element X ∈ L(K, H ) which satisfies
Lemma 2.2 ([16, Theorem 1.3]) For every A ∈ L(H, K), A† exists if and only if
R(A) is closed.
Lemma 2.3 (cf. [6, Theorem 3.2] and [15, Remark 1.1]) Let A ∈ L(H, K).
Then the closedness of any one of the following sets implies the closedness of the
remaining three sets:
If R(A) is closed, then R(A) = R(AA∗ ), R(A∗ ) = R(A∗ A) and the following
orthogonal decompositions hold:
Remark 2.4 ([14, Section 1]) Let A ∈ L(H, K). If A† exists, then A is said to be
Moore-Penrose invertible (briefly, M-P invertible). In such case,
(A† )∗ = (A∗ )† , (A∗ A)† = A† (A∗ )† , R(A† ) = R(A∗ ), N (A† ) = N (A∗ ). (6)
Lemma 2.5 (cf. [18, Lemma 4.1]) For every A ∈ L(K, H ), let
0 A
ρ(A) = ∈ L(H ⊕ K)sa . (7)
A∗ 0
0 (A† )∗
ρ(A)† = . (8)
A† 0
AA∗ 0
ρ(A)ρ(A)∗ = , (9)
0 A∗ A
so R ρ(A)ρ(A)∗ is closed if and only if both R(AA∗ ) and R(A∗ A) are closed.
Thus, from Lemmas 2.2 and 2.3 we conclude that ρ(A)† exists if and only if A†
exists, and in which case the verification of (8) follows directly from four equations
in (5).
Lemma 2.6 (cf. [18, Lemma 4.1]) For every A ∈ L(K, H ), let ρ(A) be defined by
(7). Then
Norm Estimations for the M-P Inverse of the Weak Perturbation 227
If furthermore LE,T and RF,T are both invertible, then M is said to be a strong
perturbation of T .
Definition 3.2 ([4, 21]) The operator M given by (12) is said to be a weak
perturbation of T if both LE,T and RF,T defined by (13) are M-P invertible, and
the following three conditions are satisfied:
(i) T T † L†E,T (IK − T T † ) = 0;
†
(ii) T † T RF,T (IH − T † T ) = 0;
(iii) L†E,T LE,T T = T RF,T
†
RF,T .
Definition 3.3 ([4, 18, 19]) The operator M given by (12) is said to be a stable
perturbation of T if R(M) ∩ R(T )⊥ = {0}.
Lemma 3.4 (cf. [4, Lemma 2.1]) Let M be the multiplicative perturbation of T ∈
L(H, K) given by (12). Then the following statements are valid:
(i) If M is a semi-strong perturbation of T , then M is a weak perturbation of T ;
(ii) If M is a weak perturbation of T , then M is a stable perturbation of T ;
(iii) If T † · M − T < 1, then both L∗E,T and RF,T
∗ are injective, where LE,T
and RF,T are defined by (13).
Proof
(i) Suppose that M is a semi-strong perturbation of T . Then
hence by (15) we conclude that u = 0. This completes the proof that R(M) ∩
R(T )⊥ = {0}.
(iii) Suppose that T † · M − T < 1. Let
∗
α = T † (LE,T T RF,T − T ).
α = T † (M − T ) ≤ T † · M − T < 1.
∗ ), since (I − T † T )R ∗
Given every x ∈ N (RF,T F,T = IH − T T , we have
†
H
x = T T x, hence
†
∗
x = T † (LE,T T RF,T − T ) x ≤ αx,
∗ ) = {0}.
which happens only if x = 0. This completes the proof that N (RF,T
† † ∗
Similarly, due to IK − T T = (IK − T T )LE,T and
E 0
ρ(M) = Hρ(T )H ∗ with H = (16)
0 F
is a strong (weak) perturbation of ρ(T ), where ρ(T ) and ρ(M) are defined by (7).
Proof Direct computation yields (16). Also, from (8) we can obtain
LE,T 0
LH,ρ(T ) = Hρ(T )ρ(T )† + IK⊕H − ρ(T )ρ(T )† = , (17)
0 RF,T
230 C. Fu et al.
B† 0
A† = . (18)
−CB † IK
It follows that
∗
UQ XUP = (X11 + X21 )P + (X12 + X22 )(IH − P ). (21)
An interpretation of the weak (strong) perturbation for matrices was given in [4,
Remark 2.1], which can also be carried out for operators.
Remark 3.8 Let M be the multiplicative perturbation of T ∈ L(H, K) given by (12)
such that both LE,T and RF,T defined by (13) are M-P invertible. Let PT ∈ L(H )
and QT ∈ L(K) be two projections defined by
PT = T † T and QT = T T † , (22)
−1
where T11 = T |H1 ,K1 is invertible such that T11 = T † |K1 ,H1 . Then
∗ IK1 0 IH1 0
UQ T T T † UQ T
= and UPT T † T UP∗T = , (24)
0 0 0 0
which lead to
∗ B0 D0
UQT ET T † UQ T
= and UPT F T † T UP∗T = (25)
C0 G0
232 C. Fu et al.
for some B ∈ L(K1 ), C ∈ L(K1 , K1⊥ ), D ∈ L(H1 ) and G ∈ L(H1 , H1⊥ ). It follows
that
∗ B 0 ∗ D 0
UQT LE,T UQT = and UPT RF,T UPT = . (26)
C IK ⊥ G IH ⊥
1 1
Note that
∗
† ∗
†
UQT LE,T UQ T
= UQT L†E,T UQ T
and UPT RF,T UP∗T = UPT RF,T
†
UP∗T ,
∗ B † BT 11 0
= UQ T
UP T .
0 0
Similarly, we have
† ∗ T11 D † D 0
T RF,T RF,T = UQ T
UP T .
0 0
Put
IH 0 M11 0
E= ∗ † and T = .
M12 M11 IK 0 S
With the conditions as above, norm estimations for M † were considered in [19] and
[18] for Hilbert space operators and Hilbert C ∗ -module operators, respectively. In
this subsection, we will show that M can be in fact expressed as a strong perturbation
of T .
Let PT and QT be defined by (22) such that UQT T UP∗T and UPT T † UQ ∗ are
T
∗
given by (23), where H1 = R(T ), K1 = R(T ) and T11 ∈ L(H1 , K1 ) is invertible.
Put
M11 M12
UQT MUP∗T = and 11 = M11 − T11 .
M21 M22
Then
−1
T11 · 11 = T † · 11 ≤ T † · UQT · · UP∗T < 1,
−1
−1 −1 −1
therefore M11 is invertible such that M11 = IH1 + T11 11 T11 , hence
M11 0
UQT MUP∗T = W1 W2 , (29)
0 S
234 C. Fu et al.
−1
where S = M22 − M21 M11 M12 ∈ L(H1⊥ , K1⊥ ) and
−1
IK1 0 IH1 M11 M12
W1 = −1 and W2 = . (30)
M21 M11 IK ⊥ 0 IH ⊥
1 1
M11 0 0 0
W1 = for every x ∈ H1⊥ and y ∈ K1⊥ \ {0},
0 S x y
M11 0
UQT MUP∗T = W1 W2
0 0
−1
IK1 0 T11 + 11 0 IH1 M11 M12
= −1
M21 M11 0 0 0 0 0
∗
B0 T11 0 D0
= ,
C0 0 0 G0
−1
where B = IK1 and D = (IH1 + T11 11 )∗ . It follows from (21) that M = ET F ∗ ,
where
−1
E = (IK1 + M21 M11 )QT ,
−1 −1 −1 ∗
F = IH1 + T11 11 + (IH1 + T11 11 )M11 M12 PT .
Lemma 4.1 ([10, Proposition 1.3.5]) Let x and y be two positive elements in a
C ∗ -algebra. If x ≤ y, then x ≤ y.
Definition 4.2 Given every B ∈ L(H ) and C ∈ L(H, K), let ∈ L(H ) be defined
by
= (B, C) = B ∗ B + C ∗ C. (31)
Lemma 4.3 (cf. [17, Lemma 2.4]) Given every B ∈ L(H ) and C ∈ L(H, K), let
= (B, C) be defined by (31) and let be defined by
B0
= ∈ L(H ⊕ K).
C0
† B ∗ † C ∗
† = . (32)
0 0
0
Proof Note that ∗ = , so by Lemmas 2.2 and 2.3 we know that
00
( † )∗ = † and † = † .
B ∗ x = u + v, (33)
v, v = v, B ∗ x − u = Bv, x − v, u = 0,
† B ∗ = B ∗ and † C ∗ = C ∗ . (34)
236 C. Fu et al.
B † = B and C † = C. (35)
In view of (34) and (35), the four Penrose equations for and † stated in (5) are
satisfied.
Lemma 4.4 Let B ∈ L(H ) and C ∈ L(H, K) be such that B is M-P invertible and
C = CB † B. Then the operator = (B, C) defined by (31) is also M-P invertible
such that
† = † = B † B.
PB = BB † and QB = B † B,
and put H1 = R(PB ) = R(B) and H2 = R(QB ) = R(B ∗ ). Then by (20) we have
−1
∗ B11 0 B11 0
UPB BUQ B
= and UQB B † UP∗B = , (36)
0 0 0 0
−1
where B11 = B|H2 ,H1 is invertible such that B11 = B † |H1 ,H2 . It follows that
∗ IH2 0
UQB B † BUQ B
= UQB B † UP∗B · UPB BUQ
∗
B
= . (37)
0 0
W 0
C ∗ C = UQ
∗
B
UQB , where W ∈ L(H2 ) is positive. (38)
0 0
∗ B
Furthermore, from (36) we know that B11 11 is invertible in L(H2 ), which
∗
obviously leads to the invertibility of B11 B11 + W in L(H2 ). In view of (31), (36)
and (38), we have
∗ B +W 0
B11
∗ 11
= UQ B
UQ B , (39)
0 0
∗ IH2 0
† = † = UQ B
UQB = B † B.
0 0
Theorem 4.5 (cf. [17, Corollary 3.6]) Suppose that T ∈ L(H, K) is M-P invert-
ible. Let M be a weak perturbation of T given by (12). Then
ET T † , F T † T , T T † ET T † and T † T F T † T
∗ B0 D0
1 = UQ T
UQT and 2 = UP∗T UP T ,
0 0 0 0
∗ B† 0 D† 0
1† = UQ T
UQT and 2† = UP∗T UP T .
0 0 0 0
C = CB † B and G = GD † D. (43)
We may then combine (25), (43), Lemmas 4.3 and 4.4 to conclude that
† ∗ † ∗
∗ E B E C
(ET T † )† = UQ T
UQ T , (44)
0 0
F† D ∗ F† G∗
(F T † T )† = UP∗T UP T , (45)
0 0
E = B ∗ B + C ∗ C and F = D ∗ D + G∗ G (46)
238 C. Fu et al.
such that
† †
E E = E E = B † B and F† F = F F† = D † D. (47)
M = ET T † · T · (F T † T )∗ (48)
∗ B0 T11 0 D ∗ G∗
= UQ T
UP T
C0 0 0 0 0
∗ BT11 D ∗ BT11 G∗
= UQ UP T , (49)
T CT11 D ∗ CT11 G∗
−1 † ∗ † ∗
DF† 0 T11 0 E B E C
M † = UP∗T † UQ T
GF 0 0 0 0 0
−1 † ∗ −1 † ∗
DF† T11 E B DF† T11 E C
= UP∗T −1 † ∗ −1 † ∗ UQT . (50)
GF† T11 E B GF† T11 E C
Then we may combine (49), (50), (46), (47) and (27) to get an expression of MM †
as
∗ 11 12
MM † = UQ T
UQ T ,
21 22
where
−1 † ∗ −1 † ∗
11 = BT11 (D ∗ D + G∗ G)F† T11 E B = BT11 F F† T11 E B
−1 † ∗ −1 † ∗ † ∗
= B · T11 D † D · T11 E B = B · B † BT11 · T11 E B = BE B ,
† ∗ † ∗ † ∗
12 = BE C , 21 = CB † BE B = CE B and 22 = CF† C ∗ .
Therefore,
† ∗ † ∗
∗ BE B BE C
MM = †
UQ T † ∗ † ∗ UQT = ET T † · (ET T † )† (51)
CE B CE C
by (25) and (32). As a result, (MM † )∗ = MM † . Moreover, (51) together with (48)
yields
In this section, we study norm estimations for the Moore-Penrose inverse. First, we
present a technical result, which was originally obtained in [4] for matrices.
Lemma 5.1 ([4, Lemma 3.3]) Let B ∈ L(H ) be M-P invertible and C ∈ L(H, K)
be such that C = CB † B. Then
CB † 2
C † C ∗ = and IH − B † B ∗ = θ (B, C), (52)
1 + CB † 2
Proof Following the notations as in the proof of Lemma 4.4, we know that is
M-P invertible such that † is given by (40), which can be expressed alternately as
−1 ∗ )−1 0
∗ B11 (IH1 + Z)−1 (B11
† = UQ B
UQ B , (54)
0 0
∗ −1 −1
) W 2 ∈ L(H2 , H1 ), Z = SS ∗ = (B11
∗ −1
1
S = (B11 ) W B11 ∈ L(H1 ). (55)
240 C. Fu et al.
Z0
(CB † )∗ (CB † ) = (B † )∗ (C ∗ C)B † = UP∗B UP B ,
0 0
Since Z is positive, we have Z = max {t : t ∈ σ (Z)}, where σ (Z) is the spectrum
of Z in L(H1 ). Note that the function from t to 1+tt
is monotonically increasing
on [0, +∞), so by the spectral theory for normal elements in a C ∗ -algebra [10,
Section 1], we have
4 t 5 Z
Z(IH1 + Z)−1 = max : t ∈ σ (Z) = . (57)
1+t 1 + Z
Now, we prove the first equation in (52). By (40), (38) and (54)–(56), we have
∗ ∗
† 1 1 1 1
C C =
† ∗
C
2 · C † 2 = C † 2
·C
2†
† 1 ∗ † 1
= 2 = ∗ −1 ∗ −1
(B11 B11 + W ) 2 W (B11 B11 + W ) 2
2C C
∗ − 1 1 ∗ − 1 1 ∗
=
B B
11 11 + W 2W 2 · B B
11 11 + W 2W 2
∗ − 1 1 ∗ ∗ − 1 1
=
B B
11 11 + W 2W 2 · B B
11 11 + W 2W 2
1 1
∗ 1 −1 1
= W 2 (B11 B11 + W )−1 W 2 = W 2 B11 (IH1 + Z)−1 (B11
∗ −1
) W 2
S ∗ S
= S ∗ (IH1 + SS ∗ )−1 · S = (IH2 + S ∗ S)−1 S ∗ · S =
1 + S ∗ S
SS ∗ Z CB † 2
= = = .
1 + SS ∗ 1 + Z 1 + CB † 2
Norm Estimations for the M-P Inverse of the Weak Perturbation 241
Finally, we prove the second equation in (52). First, we consider the case that B
is not surjective. In this case, H1 = H and thus IH ⊥ = 0. Therefore, by (36) and
1
(54) we have
∗ IH1 − (IH1 + Z)−1 0
IH − B B †
=UP∗B UP B
0 IH ⊥
1
Z(IH1 + Z)−1 0
=UP∗B UP B .
0 IH ⊥
1
Z CB † 2
IH − B † B ∗ = Z(IH + Z)−1 = = .
1 + Z 1 + CB † 2
Our next technical result is as follows.
Lemma 5.2 Let M be a weak perturbation of T ∈ L(H, K) given by (12) and let
LE,T and RF,T be defined by (13). Put
Then
MM † − T T † = T T † (IK − MM † ) = θL , (62)
πL
MM † (IK − T T † ) = , (63)
1 + πL2
242 C. Fu et al.
M † M − T † T = T † T (IH − M † M) = θR , (64)
πR
M † M(IH − T † T ) = . (65)
1 + πR2
MM † − T T † = M † M(IH − T † T )
= M † (M − T )(IH − T † T )
≤ M † · M − T · IH − T † T
≤ M † · M − T . (66)
Next, we consider the general case that M is given by (12). By Lemma 3.6
we know that ρ(M) = Hρ(T )H ∗ is also a strong perturbation of ρ(T ), where
ρ(T ), ρ(M) and H are given by (7) and (16), respectively. Using (9), (11), (66) and
(10), we obtain
6 7
max MM † − T T † , M † M − T † T = ρ(M)ρ(M)† − ρ(T )ρ(T )†
≤ ρ(M)† · ρ(M) − ρ(T )
= M † · M − T .
Remark 5.5 It is remarkable that upper bound (4) may be false for a general weak
perturbation. For a counterexample, see [4, Theorem 3.2].
Norm Estimations for the M-P Inverse of the Weak Perturbation 243
where
λ = M † · T † · M − T .
Therefore,
√
5+1
M − T = π(M − T ) ≤
† † † †
λ.
2
This completes the proof of (67) in the case that H = K.
Next we consider the general case that H = K. Let M B=EB· T
B· F
B∗ ∈ L(H ⊕K),
where
B= 0 0 0 0 F 0
E , TB = B=
and F . (72)
0E T 0 0 0
Then
0 T† 0 0 0 M†
TB† = B=
,M B† =
,M , (73)
0 0 M0 0 0
BTBTB† + IH ⊕K − T
BTB† = IH 0
B TB = E
LE, , (74)
0 LE,T
(77)
−1 −1
∗
RR −1 ,M = F T † T + IH − F T † T (IH − M † M) RF,T (IH − T † T ),
F,T
(78)
−1
(IK − MM † ) L −1 MM † = πL , (79)
LE,T ,M
9
: πL 6 72
L −1 ≤ : ; + max (ET T † )† , IK − T T † , (80)
LE,T ,M
1 + πL2
L − IK ≤ LE,T − IK · (ET T † )† , (81)
L−1
E,T ,M
−1 6 7
LL−1 ,M ≤ (1 + πL + πL2 ) · max ET T † 2 , 1 + πL2 , (82)
E,T
−1
LL−1 ,M − IK ≤ 1 + πL2 · LE,T − IK , (83)
E,T
where LE,T , RF,T and πL are defined by (13) and (58) respectively, and
LL−1 = L−1
E,T MM + IK − MM ,
† †
(84)
E,T ,M
−1
RR −1 ,M = RF,T M † M + IH − M † M. (85)
F,T
Proof
(1) Note that (76) can be derived directly from (2). Put
PM = MM † , KM = R(PM ) = R(M)
IKM A12
UPM AUP∗M = , (89)
0 A22
in which
(IK − T T † )(IK − PM )ξ = S ∗ ξ = 0,
0 0 IK1 − B −1 B ∗ −B −1 C ∗
= ·
0 IK ⊥ −C −1 B ∗ IK ⊥ − C −1 C ∗
1 1
(B ∗ )−1 −(B ∗ )−1 C ∗ 0 0
· ·
0 IK ⊥ 0 IK ⊥
1 1
0 0 ∗
= = UQT (IK − T T † )UQ ,
0 IK ⊥ T
1
where = B ∗ B + C ∗ C. Therefore,
∗
(IK − T T † )(IK − PM ) L−1 † †
E,T (IK − T T ) = IK − T T . (92)
Norm Estimations for the M-P Inverse of the Weak Perturbation 247
It follows that
∗
A22 (IK − PM ) L−1 †
E,T (IK − T T ) = S, (93)
⊥.
which means that A22 is surjective, since by (86) and (91) we have R(S) = KM
⊥
Therefore, A22 is invertible in L(KM ), and then (93) gives
−1 ∗
A−1
22 S = (IK − PM ) LE,T (IK − T T ).
†
(94)
It follows from (89) that the operator A defined by (88) is invertible such that
∗
+ PM · (IK − T T † ) · (IK − PM )(L−1 †
E,T (IK − T T )
∗
= ET T † · (IK − PM )(L−1 E,T (IK − T T ) + PM (IK − T T ).
† †
(97)
Therefore, it can be deduced by (96), (95), (21), (86), (91), (94) and (97) that
−1 2 3
LL−1 = PM − A12A−1
22 (IK − PM ) + A −1
22 (IK − PM ) LE,T
E,T ,M
where A22 is formulated by (90) and S is defined by (91). Note that both IK −
PM and IK − T T † are idempotent, so by (90) we have
Y Y ∗ = A−1 †
22 (IK − PM )(IK − T T ) − (IK − PM ) + IK
= −IK ⊥ + A−1
22 . (99)
M
Since the operator A22 defined by (90) is positive definite and A22 ≤ 1, we
know that A−1
22 is also positive definite. Then (99) indicates that each spectral
−1
point λ of A22 satisfies λ ≥ 1, which means that
Y 2 = Y Y ∗ = A−1
22 − 1. (100)
A−1 ∗ † † ∗ † † † ∗ ∗ †
22 = (SS ) = (S ) S = S (S ) = (S S) . (101)
1
(IK ⊥ − C † C ∗ )−1 = .
1 1 − C † C ∗
∗ ∗ IKM 0
SPM |KM A22 0 0
00
= .
Y 0
So we may combine the equation above with (100), (101), (102) and (103) to
conclude that
X = Y = A−1 22 − 1 = (S ∗ S)† − 1 = (1 + π 2 ) − 1 = π .
L L
which gives
∗IK1 − B −1 B ∗ + −1 −B −1 C ∗
ZZ =
−C −1 B ∗ IK ⊥ − C −1 C ∗
1
IK1 − B −1 B ∗ −B −1 C ∗ −1 0
= +
−C −1 B ∗ −C −1 C ∗ 0 IK ⊥
1
∗ −1 0
=UQT (T T † − PM )UQ + .
T 0 IK ⊥
1
Note that
−1 0 2 3∗
∗ ∗
= UQT (ET T † )† UQ · UQT (ET T † )† UQ ,
0 0 T T
250 C. Fu et al.
Similarly,
hence
(CB −1 )∗
W ∗W =
CB IK ⊥ + CB (IK1 + )(CB −1 )∗
−1 −1
1
0
=W1 W1∗ . (106)
0 IK ⊥ + CB −1 (CB −1 )∗
1
Note that CB −1 = πL , so by (10) we can get ρ (CB −1 )∗ = πL , where
ρ (CB −1 )∗ is defined by (7). It follows from (104) that
0 0
W1 W1∗ = −1 ∗
IK1 ⊕K ⊥ + ρ (CB ) +
1 0 CB −1 (CB −1 )∗
≤ 1 + πL + πL2 .
Norm Estimations for the M-P Inverse of the Weak Perturbation 251
B − IK1 0
W − IK1 ⊕K ⊥ = · W2 , (107)
1 C 0
IK1 (CB −1 )∗
where W2 = . As is shown above, W2 = W2 W2∗ =
0 0
1 + πL . Therefore (107) yields
2
−1
LL−1 ,M − IK = W − IK1 ⊕K ⊥ ≤ LE,T − IK · 1 + πL2 .
E,T 1
Acknowledgments This work was supported by the National Natural Science Foundation of
China (11971136) and a grant from Shanghai Municipal Science and Technology Commission
(18590745200).
References
1. A. Böttcher, I.M. Spitkovsky, A gentle guide to the basics of two projections theory. Linear
Algebra Appl. 432, 1412–1459 (2010)
2. N. Castro-González, J. Ceballos, F.M. Dopico, J.M. Molera, Accurate solution of structured
least squares problems via rank-revealing decompositions. SIAM J. Matrix Anal. Appl. 34,
1112–1128 (2013)
3. N. Castro-González, M.F. Martı́nez-Serrano, J. Robles, Expressions for the Moore-Penrose
inverse of block matrices involving the Schur complement. Linear Algebra Appl. 471, 353–
368 (2015)
4. C. Fu, C. Song, G. Wang, Q. Xu, Norm estimations for the Moore-Penrose inverse of the weak
perturbation of matrices. Linear Multilinear Algebra. 70, 125–237 (2022)
5. N. Hu, W. Luo, C. Song, Q. Xu, Norm estimations for the Moore-Penrose inverse of
multiplicative perturbations of matrices. J. Math. Anal. Appl. 437, 498–512 (2016)
6. E.C. Lance, Hilbert C ∗ -modules–A Toolkit for Operator Algebraists (Cambridge University
Press, Cambridge, 1995)
7. Z. Li, Q. Xu, Y. Wei, A note on stable perturbations of Moore-Penrose inverses. Numer. Linear
Algebra Appl. 20, 18–26 (2013)
252 C. Fu et al.
L. Arambašić
Department of Mathematics, Faculty of Science, University of Zagreb, Zagreb, Croatia
e-mail: [email protected]
A. Guterman · S. Zhilina
Department of Mathematics and Mechanics, Lomonosov Moscow State University, Moscow,
Russia
Moscow Center for Fundamental and Applied Mathematics, Moscow, Russia
Moscow Center for Continuous Mathematical Education, Moscow, Russia
e-mail: [email protected]
B. Kuzma ()
University of Primorska, Koper, Slovenia
IMFM, Ljubljana, Slovenia
Moscow Center for Fundamental and Applied Mathematics, Moscow, Russia
e-mail: [email protected]
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 255
R. M. Aron et al. (eds.), Operator and Norm Inequalities and Related Topics,
Trends in Mathematics, https://doi.org/10.1007/978-3-031-02104-6_8
256 L. Arambašić et al.
1 Introduction
In a real or complex inner product space (H, ·, ·), two vectors x, y are orthogonal
(denoted by x ⊥ y) if x, y = 0. The familiar, and mostly trivial, properties of
orthogonality relation are
Homogeneity x ⊥ y implies (αx) ⊥ (βy) for scalars α, β.
Left additivity x1 ⊥ y and x2 ⊥ y imply (x1 + x2 ) ⊥ y.
Right additivity x ⊥ y1 and x ⊥ y2 imply x ⊥ (y1 + y2 ).
Symmetricity x ⊥ y implies y ⊥ x.
Would it be possible to extend the orthogonality relation from inner product
spaces to a more general normed space (X , · )? The lack of the inner product
in such spaces suggests that, if there is any chance to succeed, one should better
express orthogonality with the help of the norm alone. This is possible in many ways
and it gives rise to several nonequivalent extensions of orthogonality. We present a
few of them, each with its own virtues and vices. Unless explicitly stated otherwise,
our claims will hold equally well for real and complex normed spaces. Thus, we
will occasionally omit specifying the underlying field and will simply refer that
something holds for all scalars or that there exists a scalar with a particular property,
etc. Let F denote the real or the complex field.
In 1934 Roberts [76] observed that in inner product spaces x ⊥ y if and only if
for every scalar λ. Clearly, this condition can be verified in every normed space and
gives rise to Roberts orthogonality: x ⊥R y if (1) holds.
It is an easy exercise to see that Roberts orthogonality is homogeneous and
symmetric. It does not have an additivity property in general—say, in three-
dimensional real space equipped with supremum norm, (R3 , · ∞ ), the vector
x = (1, 1, 1) is Roberts orthogonal to y1 = (1, −1, 0) and to y2 = (0, 2, −2) but
not to y1 + y2 .
Further vices of Roberts orthogonality are that the condition (1) is very restric-
tive, so much that in some normed spaces x ⊥R y only when either x or y is
a zero vector. An example of such a (real, two-dimensional) space was given by
James [45]; it equals the set of all real polynomials of degree at most two restricted
to unit interval [0, 1] and vanishing at 0, and equipped with the supremum norm.
In the same paper, James also proved the next proposition.
Birkhoff–James Orthogonality: Characterizations, Preservers, and Orthogonality Graphs 257
Proposition 1.1 (Cf. [45, Corollary 4.7]) The following conditions are equivalent
for a real and at least two-dimensional normed space X :
(i) X is an inner product space.
(ii) In every two-dimensional plane ⊆ X and for every x ∈ there exists a
nonzero y ∈ which is Roberts orthogonal to x.
The proof will be given within Sect. 1.3.
Similar characterization holds also for complex normed spaces provided that
we take for two-dimensional real-linear planes. Namely, each complex normed
space (X , · ) can be considered as a real normed space (denoted temporarily as
(X , · R )) by restricting the scalars, and the following general result applies:
Proposition 1.2 (Cf. [31, Theorem 7.2]) A complex normed space (X , · ) is an
inner product space if and only if the associated real normed space (X , · R ) is
an inner product space. When this happens, the inner products ·, · and ·, ·R are
related by the equations a, bR = Rea, b and a, b = a, bR − iia, bR .
Sketch of the Proof If ·R is induced by a real inner product ·, ·R , then 2x2 =
(1 + i)x2 = x2 + ix2 + 2ix, xR , so ix, xR = 0. Linearization gives
further x, iyR + ix, yR = 0. From here it is straightforward that x, y :=
x, yR − iix, yR is an inner product over C which induces the given norm.
James in [45] observed that one does not need the full set of scalars in order that (1)
be equivalent to the usual orthogonality in real inner product spaces; it suffices that
x + y = x − y. (2)
n→+∞
nx + n+α y − nx
1
+ n1 y ≤ n( n+α
1
− n1 )y = ( n+α
n
− 1)y −−−−→ 0.
Therefore,
n→+∞
(n + α)x + y − nx + y = (n + α)x + n+α y − nx
1
+ n1 y −−−−→ αx.
258 L. Arambašić et al.
x + λy ≥ x
for every scalar λ. This orthogonality was more thoroughly investigated by James
in [46, 47], and because of this it is often known as Birkhoff–James (B-J for short)
orthogonality. We remark, however, that B-J orthogonality can be traced back at
least as far as Carathéodory (see [5]).
One can see from its very definition that B-J orthogonality is homogeneous,
but we prefer to verify this from its equivalent geometrical reformulation (see
Proposition 1.4 below), due to James [47, Corollary 2.2], which will reveal much
more than just homogeneity. To do that we need to introduce some terminology.
Recall that F denotes either the field of real or the field of complex numbers; also,
given a vector x ∈ (X , · ) in a normed space X over F, the linear functional
f : X → F for which
Fig. 1 Birkhoff–James
orthogonality: x ⊥ y
260 L. Arambašić et al.
Proof Let (H, ·, ·) be an inner product space and H its completion. By the
Riesz representation theorem and the Bunyakovsky–Cauchy–Schwarz inequality—
the part which claims that equality can hold only for linearly dependent vectors [42,
p. 4], the only supporting functional at a normalized vector x ∈ H ⊆ H is given by
·, x.
From the equivalence stated in Proposition 1.4, the homogeneity of B-J orthog-
onality is evident; it is also clear that in every two-dimensional plane and for
every x ∈ there exists y ∈ with x ⊥ y. That is, similarly to isosceles,
B-J orthogonality is always nontrivial. It is easily seen that B-J orthogonality is
in general nonsymmetric, that is, x ⊥ y does not always imply y ⊥ x (e.g., in the
space (F2 , · ∞ ), where · ∞ is the maximum norm, we have (1, 1) ⊥ (1, 0)
but (1, 0) ⊥ (1, 1)). We will discuss the properties of B-J orthogonality more
thoroughly in Sect. 2.
Proposition 1.4 yields a simple procedure to find all vectors y in a normed space
which are B-J orthogonal to a given normalized x, namely:
C
Nx := {y ∈ X ; x ⊥ y} = ker f.
f ∈X ∗
f =f (x)=1
But one can do slightly better: Singer [82] (see also his monograph [84, Lemma 1.3,
p. 169]) observed that if M ⊆ X is an n-dimensional subspace of a real normed
space X , and f is a bounded linear functional on X such that f |M = 1, then
f coincides on M with some convex combination of at most n extremal points
of the dual norm. This works also for complex normed spaces, except that in this
case one requires at most 2n − 1 extremal points of the dual norm. By taking
M = span{x, y}, one obtains the following result, which was the starting point of
Li and Schneider’s [60] investigation into B-J orthogonality on rectangular matrices
equipped with Schatten p-norm (see also the next chapter).
Proposition 1.6 (Cf. [60, Proposition 2.1]) Let X be a normed space over F and
let x, y ∈ X be linearly independent. Then the following are equivalent:
(i) x ⊥ y.
(ii) There exist h (h ≤ 2 if F = R, respectively, h ≤ 3 if F = C) extremal points
f1 , . . . , fh in unit sphere of the dual norm and h numbers λ1 , . . . , λh > 0 with
h
i=1 λi = 1 such that
h
λi fi (y) = 0 and f1 (x) = · · · = fh (x) = x.
i=1
them; more detailed study can be found in survey papers [3] and [5]. One possibility
is Pythagorean orthogonality introduced in [47] by x ⊥P y if
Day [31] obtained another nice characterization of inner product spaces based on
the (non)equivalence between Pythagorean and isosceles orthogonalities.
Proposition 1.8 (Cf. [31, Theorems 5.1 and 5.2]) The following are equivalent for
a normed space X .
(i) X is an inner product space.
(ii) Isosceles orthogonality implies Pythagorean orthogonality.
(iii) Pythagorean orthogonality implies isosceles orthogonality.
Sketch of the Proof (i) #⇒ (ii) and (iii). Given any x, y ∈ X , both x ⊥I y and
x ⊥P y are equivalent to Rex, y = 0.
(ii) #⇒ (i). Choose any normalized x, y ∈ X . Then (x + y) ⊥I (x − y), so by
the assumptions also (x + y) ⊥P (x − y), that is,
We first assume that X is a real normed space. It can be shown [31, Theorem 2.2]
that a two-dimensional real normed space Y is an inner product space (equivalently,
the norm’s unit sphere is an ellipse) if and only if x + y2 + x − y2 = 4 for all
normalized vectors x, y ∈ Y.
Consider now two arbitrary vectors u, v ∈ X and a two-dimensional subspace
Y ⊆ X which contains them. Then Y is an inner product space, so u + v2 + u −
v2 = 2(u2 + v2 ), and thus parallelogram identity holds in X . Therefore, by
Jordan–von Neumann characterization [49], X is an inner product space.
Assume now that X is a complex normed space. We have already proved that X
is a real inner product space, so it remains to apply Proposition 1.2.
(iii) #⇒ (i). Assume that x ⊥P y, that is, x +y2 = x2 +y2 . Then we also
have x ⊥I y, so x +y = x −y, and thus x −y2 = x2 +y2 , which means
that x ⊥P (−y). By [31, Lemma 5.3], X is a real inner product space whenever the
conditions x ⊥P y and x ⊥P (−y) are equivalent. Since Pythagorean orthogonality
does not depend on the field F, for a complex normed space X it remains to apply
Proposition 1.2.
In 1957 Singer [83] introduced a homogeneous version of isosceles orthogonality
y
by x ⊥S y if either x · y = 0 or x x
⊥I y . It is additive in dimension two.
The question of its additivity in general was finally put to the rest by Lin [61] who
found that in higher dimensions Singer orthogonality is additive if and only if the
norm is induced by the inner product.
In 1984 Alonso [2] introduced area orthogonality in real normed spaces by
x ⊥A y if either x · y = 0 or x, y are linearly independent and the lines Rx,
Ry divide the norm’s unit ball of the plane span{x, y} 4 R2 into four parts of equal
Birkhoff–James Orthogonality: Characterizations, Preservers, and Orthogonality Graphs 263
area; together with Benítez they obtained yet another interesting characterization of
real inner product spaces, whose proof we omit, as follows:
Proposition 1.9 (Cf. [4, Proposition 3]) The following are equivalent for a real
normed space X :
(i) X is an inner product space.
(ii) B-J orthogonality implies area orthogonality.
(iii) Area orthogonality implies B-J orthogonality.
An interesting equivalent description of orthogonality in inner product spaces
is presented in [40]. It does not rely neither on the inner product nor on the norm
but rather on isometries of the norm. To state it properly, we require the following
terminology. Given a vector x, denote by x := {λx; λ ≥ 0} the closed ray in
direction of x. Observe that 0 = {0} and that, for x nonzero, x \ {0} = {λx; λ > 0}
is an open ray.
Proposition 1.10 (Cf. [40]) Two vectors x, y in a real inner product space H are
orthogonal (in a classical sense) if and only if there exists a rigid motion, i.e., an
isometry T of H such that
−
→ − → → −−→
T x ∪ Ty ∪ −
x \ (−
→
y \ {0}) = −
→
x ∪ (−x). (3)
Two vectors x, y in a complex inner product space H are orthogonal if and only if
there exists an isometry T for which, simultaneously, (3) and
−−−→ − → −−→ −−→ −−−→
T (ix) ∪ T y ∪ (ix) \ (−
→
y \ {0}) = (ix) ∪ (−ix) (3 )
are satisfied.
The condition in (3) is clearly homogeneous and a moment’s thought reveals that
an isometry T satisfies (3) if and only if the isometry T = −T satisfies (3) with
x and y swapped. Hence, the relation based on (3) is always symmetric. However,
in general normed spaces it might be trivial, i.e., if vectors x, y satisfy (3), then at
least one of them must be zero. This can happen when the isometry group consists
of scalar operators only. Such normed spaces do exist: Davis [29] was the first to
construct separable Banach space over reals such that its only isometries (surjective
or not) are ±I . We also refer to Gordon and Loewy [38, Theorem 3.1] who, by
answering a question of Lindenstrauss, showed much more: Any finite subgroup of
(real) orthogonal n-by-n matrices which contains −I is the isometry group of some
norm on Rn . Moreover, any real or complex Banach space can be renormed, so that
its isometry group consists of scalar operators only, see Jarosz [48].
The following proposition presents several other nice criteria for a normed space
to be an inner product space.
264 L. Arambašić et al.
Proposition 1.11 Let X be a real or complex normed space. Then the following
conditions are equivalent:
(i) X is an inner product space.
(ii) x = y = 1 implies αx+βy = βx+αy for all α, β ∈ R (Ficken [36]).
(iii) There exists some γ ∈ R, γ = 0, 1, such that x = y = 1 implies x +
γ y = γ x + y (Lorch [62]).
(iv) x = y = 1 and x ⊥ y imply (x + y) ⊥ (x − y) (Oman [70,
Theorem 5.21]).
The interested reader is referred to a monograph by Amir [7] for hundreds of
additional conditions on the norm which force it to be induced by an inner product.
at x in direction of y:
Observe that D− (x; y) = −D+ (x; −y). We show next their geometric significance.
It will be beneficial to denote by J (x) the set of all supporting functionals at a
vector x. Recall that φ ∈ J (x) if and only if φ(x) = x and φ = 1.
Proposition 2.2 (Cf. [33, Lemma 1 and Theorem 15]) Let x be a nonzero vector
in X . Then
Sketch of the Proof If φ ∈ J (x), then φ = Re φ = 1 and φ(x) = x. So, for
t ∈ R sufficiently close to zero, x +ty ≥ | Re φ(x +ty)| = x+t Re φ(y). After
dividing by t and keeping an eye on its sign we get limt 30 x+ty−x
t ≤ Re φ(y) ≤
x+ty−x
limt %0 t .
To show that both inequalities are achieved, define two real-linear functionals on
spanR {x, y} by ψ± (αx+βy) := αx+βD± (x; y), (α, β ∈ R). Since t $→ x+ty
is convex, the function β $→ x+βy−x
β increases to D− (x; y) as β 3 0 and
decreases to D+ (x; y) as β % 0, respectively. Thus, with β ≥ 0 we have
with a similar derivation and the same conclusion also for β < 0. Moreover,
ψ+ (−x − βy) = −x − βD+ (x; y) ≤ −x + (x − βy − x) = x − βy −
2x ≤ − x − βy. Thus, ψ+ (αx + βy) ≤ αx + βy (α, β ∈ R). Combined with
ψ+ (x) = x, we get ψ+ = 1. By the Hahn–Banach theorem, we can enlarge
the domain of ψ+ to X without affecting its norm, and (if X is a complex space)
make it into a complex-linear functional φ+ ∈ J (x) with Re φ+ = ψ+ .
One argues similarly with ψ− to construct φ− ∈ J (x) with Re φ− (y) = ψ− (y) =
D− (x; y).
Proposition 2.3 (Cf. [47, Theorem 3.1]) Let X be a real normed space, x, y ∈ X .
Then x ⊥ (y − αx) if and only if D− (x; y) ≤ αx ≤ D+ (x; y).
Sketch of the Proof Apply Propositions 1.4 and 2.2.
Definition 2.4 Let (X , · ) be a real or complex normed space. The norm is
Gateaux differentiable at a point x if for all y ∈ X there exists D(x; y) =
limR8t →0 x+ty−x
t .
266 L. Arambašić et al.
so φ(y) = D(x; y). It follows that the Gateaux derivative y $→ D(x; y) is the
unique supporting functional at x, and thus x is a smooth point.
Assume now that x is not a smooth point, that is, there exist two different
supporting functionals φ1 , φ2 at x. Then there exists some y ∈ X such that
φ1 (y) < φ2 (y). Hence it follows from Proposition 2.2 that
where u, v := uv t is the standard scalar product of row vectors in Rn and ∂x :=
∂· ∂·
∂ξ1 , . . . , ∂ξn (x) is the norm’s differential evaluated at vector x.
A complex normed space (Cn , · ) can be regarded as a real normed space
(R , · ) by restricting the scalars. We write x = (ξ1 , . . . , ξn ) ∈ Cn and
2n
where u, v := uv ∗ is the standard scalar product of row vectors in Cn . This can
be simplified if one introduces complex partial derivatives
∂ · ∂ · ∂ · ∂ · ∂ · ∂ ·
:= −i and := +i ;
∂ξk ∂Re ξk ∂ Im ξk ∂ξk ∂Re ξk ∂ Im ξk
then we can identify the C-linear supporting functional fx with the norm’s complex
conjugate differential
∂ · ∂ ·
∂C x := ,..., (x)
∂ξ1 ∂ξn
(5)
∂ · ∂ · ∂ · ∂ ·
= +i ,..., +i (x).
∂ Re ξ1 ∂ Im ξ1 ∂ Re ξn ∂ Im ξn
We will need two additional properties of orthogonality which hold trivially for
orthogonality in inner product spaces.
Left uniqueness For any x, y ∈ X , x = 0, there exists at most one α ∈ F such
that (αx + y) ⊥ x.
Right uniqueness For any x, y ∈ X , x = 0, there exists at most one α ∈ F such
that x ⊥ (αx + y).
Theorem 2.6 (Cf. [46, Theorems 1 and 2], [47, Theorems 4.1–4.3 and 5.1]) Let
X be a real or complex normed space.
(i) B-J orthogonality in X is left unique if and only if the norm on X is strictly
convex.
(ii) B-J orthogonality in X is right unique if and only if it is right additive if and
only if the norm on X is smooth.
268 L. Arambašić et al.
for any λ ∈ (−ε, ε), so f (y) = 1. Hence f ((1 − λ)x + λy) = 1 for all
λ ∈ (−ε, 1+ε), and thus f is also supporting at (1−λ)x+λy. Since f (x−y) =
0, we obtain from Proposition 1.4 that ((1 − λ)x + λy) ⊥ (x − y) for all
λ ∈ (−ε, 1 + ε). Therefore, B-J orthogonality in X is not left unique.
Assume now that B-J orthogonality in X is not left unique. Then there exists
a two-dimensional subspace Y ⊆ X , two linearly independent normalized
vectors x, y ∈ Y and a normalized vector z ∈ Y such that x ⊥ z and y ⊥ z.
Consider two supporting functionals fx , fy : Y → F at x and y, respectively,
such that fx (z) = fy (z) = 0. Since Y is two-dimensional, fy is a scalar
multiple of fx , so we may assume that fx = fy = f . Hence
for all λ ∈ [0, 1]. Therefore, the norm’s unit sphere contains a line segment
[x, y], and the norm on X is not strictly convex.
(ii) We first show that if B-J orthogonality is right unique, then for any nonzero
x ∈ X there exists a unique supporting functional fx at x. Assume from
the contrary that f1 and f2 are two distinct supporting functionals for some
normalized vector x ∈ X . Then there exists y ∈ X such that f1 (y) = f2 (y).
We have f1 (y−f1 (y)x) = 0, so x ⊥ (y−f1 (y)x). Similarly, f2 (y−f2 (y)x) =
0, so x ⊥ (y − f2 (y)x). Hence B-J orthogonality in X is not right unique.
It follows from Proposition 1.4 that uniqueness of the supporting functional
fx for any nonzero x ∈ X implies right additivity of B-J orthogonality in X .
We next show that right additivity implies right uniqueness.
Assume that B-J orthogonality is right additive and that x ⊥ (αx + y),
x ⊥ (βx + y) for some α, β ∈ F and x, y ∈ X . Then x ⊥ −(βx + y), so
x ⊥ ((αx + y) − (βx + y)) = (α − β)x. This is only possible for α − β = 0,
so B-J orthogonality in X is right unique.
Birkhoff–James Orthogonality: Characterizations, Preservers, and Orthogonality Graphs 269
D
n
N := ker fk . (7)
k=1
z $→ ∂z
and it is continuous there by [77, Theorem 25.5]. Also, − z = z implies that
∂ − z = −∂z (8)
for all z ∈ . Note that the Euclidean unit n-sphere Sn ⊂ Rn+1 is homeomorphic
(ζ1 ,...,ζn+1 ,0,...,0)
to (via the radial projection r : (ζ1 , . . . , ζn+1 ) $→ (ζ1 ,...,ζn+1 ,0,...,0)
). In particular,
by Borsuk–Ulam’s theorem applied to the function
> ? > ?
z $→ x1 , ∂r(z) , . . . , xn , ∂r(z) : Sn → Rn , (9)
which is odd (by (8)) and continuous, there exists a point y ∈ ⊆ Rn+1 ⊕ 0n
on the unit> sphere? where the function (9) is equal to 0n . By Proposition 2.5,
g : x $→x, ∂y is a supportingfunctional at y, and for all γ1 , . . . , γn ∈ R we
have g( nk=1 γk xk ) = 0, so y ⊥ nk=1 γk xk . Since y ∈ N , we finally obtain that y
satisfies condition (6).
Case 2: X is a complex normed space.
Recall that dimC X = 2n + 1 and that each z ∈ N satisfies xk ⊥ z in a complex
normed space X . It remains to find an element in N which satisfies the second
condition in (6). To do this, we consider X and N as real normed spaces. We may
apply the construction from the real case to 2n vectors x1 , ix1 , . . . , xn , ixn , since
dimR X = 2(2n + 1) ≥ 2(2n) + 1 and dimR N ≥ 2(n + 1) ≥ (2n) + 1. This
gives us a normalized
n vector y ∈ N such that, for all α1 , . . . , αn , β1 , . . . , βn ∈ R,
it holds
y ⊥ k=1 (α k xk + βk ixk ) in a real vector space X . This is equivalent to
y ⊥ nk=1 γk xk for all γ1 , . . . , γn ∈ C in a complex vector space X , so y satisfies
both conditions in (6).
Example 2.4 in [13] demonstrates that the lower bound on dim X in Theorem 2.9
is exact. We state two immediate corollaries.
Corollary 2.10 Let X be a normed space over F with dim X ≥ 3. Then for every
normalized vector x there is a normalized vector y with x ⊥
⊥ y.
Corollary 2.11 Let X be a normed space over F with dim X ≥ 5. Then for every
two normalized vectors x, y ∈ X there is a normalized vector z ∈ X with x ⊥
⊥
z⊥⊥ y.
Another corollary to Theorem 2.9 is that in infinite-dimensional spaces we can
find infinitely many pairwise B-J orthogonal nonzero vectors.
Corollary 2.12 Let X be an infinite-dimensional normed space over F. Then there
exists an infinite number of nonzero vectors which are pairwise B-J orthogonal.
Proof We construct a sequence (xn )n of pairwise B-J orthogonal normalized
vectors in X recursively. Let x1 ∈ X be an arbitrary normalized vector. Assume
now that we already have x1 , . . . , xn . Since dim X > 2n + 1, Theorem 2.9 implies
Birkhoff–James Orthogonality: Characterizations, Preservers, and Orthogonality Graphs 273
How does B-J orthogonality look like in classical Banach spaces? For Hilbert spaces
it coincides with the usual inner product orthogonality. The next important example
we provide the answer for is B(H), the space of bounded linear operators on a
complex Hilbert space H (equipped with the usual operator norm). Let us begin
with two examples.
Example
(a) Let A, B ∈ B(H) be operators with orthogonal ranges, that is, such that A∗ B =
0. Then, by Pythagorean identity, for all λ ∈ C it holds
so A ⊥ B.
(b) Let A ∈ B(H) be an operator which is not bounded from below. Let (xn )n
be a sequence of normalized vectors such that limn→∞ Axn = 0. Then
limn→∞ (I + λA)xn = 1 = I for all λ ∈ C, so I + λA ≥ I and
I ⊥ A.
274 L. Arambašić et al.
Stampfli first proved that W0 (A) is a closed convex subset of C, and this was used
in the proof of the equivalence d(A, CI ) = A ⇔ 0 ∈ W0 (A).
Later, in his study on the distance to finite-dimensional subspaces in operator
algebras, Magajna [63] introduced, for A, B ∈ B(H), the notion of the maximal
numerical range of B ∗ A relative to A in the following way
4
WA (B ∗ A) = μ ∈ C; ∃(xn )n ∈ H, xn = 1, lim B ∗ Axn , xn = μ,
n→∞
5
lim Axn = A .
n→∞
Proof We may assume that B = 1. By using the same arguments as in [86,
Lemma 2], it can be shown that the set WA (B ∗ A) is a nonempty, closed and convex
subset of C.
Suppose that 0 ∈ WA (B ∗ A). Let (xn )n be a sequence of normalized vectors in
H such that limn→∞ B ∗ Axn , xn = 0 and limn→∞ Axn = A. Then for each n
we have
We will show that A − μB < A which will imply that A ⊥ B.
Let x ∈ H be a normalized vector. If x ∈ S then
limn→∞ (|A|xn − Axn ) = 0, so (xn )n is convergent as well. Then (ii) holds with
x := lim xn . The converse is obvious.
In the case when dim H < ∞, this corollary can also be stated in terms of
matrices (see [20, Theorem 1.1]). Moreover, Li and Schneider used Proposition 1.6
to generalize it to rectangular (real or complex) matrices Mm×n (F) with Schatten
p-norm
=
n √ p 1/p
Ap := p
(σi (A))p = tr A∗ A ,
i=1
where σ1 (A) ≥ σ2 (A) ≥ · · · ≥ σn (A) ≥ 0 are the singular values of A (the square
roots of eigenvalues of A∗ A). Note that A∞ = σ1 (A) is the spectral norm. By
[79, Theorem 9], Schatten norms · p and · q on Mm×n (F) with 1/p + 1/q = 1
are dual to each other; the duality is given by (A, B) $→ tr(B ∗ A).
Except for p = 1 and p = ∞, Schatten p-norm is differentiable, hence smooth.
√ p 1/p
This can be seen directly for p > 2, because X $→ tr X∗ X is a
∗
composition of real differentiable functions: X $→ H (X) = X X is clearly real
differentiable, while
=
n n
∗
H =U λi (H )Eii U $→ tr(|H |p/2) := |λi (H )|p/2
p p
i=1 i=1
is also a differentiable map on the real space of Hermitian matrices by, e.g.,
Lewis’s
[59, Theorem 1.1]. One should mention that the differentiability of H $→
p
tr(|H |p/2) follows also from Rellich’s result [75] (a simplified proof can be
found in Kato’s monograph [53, Theorem II.6.8]). Namely, for each Hermitian
H, H ∈ Mn (C), there exist n continuously differentiable functions in a real variable
t which represent n eigenvalues (counted with multiplicities) of H + tH . It is then
immediate that H $→ |H |p/2 has a continuous directional derivative in direction
H , so it must be differentiable. We remark that [59, Theorem 1.1] gives a much
more general differentiability result for matrix functions. The differentiability for
1 < p < 2 was proved in [15], but the reader is again referred to Lewis [58,
Corollary 2.6] for a more general result.
Let us introduce a bit more notation: Given a matrix A ∈ Mn (F), we denote by
WC (A) := {x ∗ Ax; x ∈ Cn , x ∗ x = 1}
WR (A) := {x ∗ Ax; x ∈ Rn , x ∗ x = 1}
be the restricted numerical range. Recall that WC (A) is convex, and so is WR (A)
because it is an interval (it is an image of a continuous function x $→ x ∗ Ax whose
domain is the Euclidean (n − 1)-sphere).
Birkhoff–James Orthogonality: Characterizations, Preservers, and Orthogonality Graphs 277
0 ∈ WF (U ∗ BA∗ U ).
Sketch of the Proof (i) #⇒ (iv). By Proposition 2.1, there exist extreme points
Fk = yk xk∗ ∈ Mm×n (F) (xk = yk = 1) for 1 ≤ k ≤ h and positive scalars λk
summing to one, such that
h
A∞ = Fk , A := tr(Fk∗ A) = yk∗ Axk and λk Fk , B = 0. (10)
k=1
1 1
U vk = yk and xk = A∗ yk = V vk ; V = A∗ U ∈ Mn×r (F).
A∞ A∞
Thus, by the second condition in (10), 0 = hk=1 λk vk∗ U ∗ BV vk , which is a point in
the convex hull of WF (U ∗ BV ). Due to its convexity, 0 ∈ WF (U ∗ BV ), giving (iv).
(iv) #⇒ (iii). Choose a normalized vector v ∈ Fr with 0 = v ∗ U ∗ BA∗ U v, then
y := U v and x := A1 ∞ A∗ U v are normalized vectors satisfying (ii).
The remaining implications (iii) #⇒ (ii) #⇒ (i) are straightforward.
Proposition 3.5 (Cf. [60, Theorem 3.2]) Let m ≤ n and let p ∈ (1, ∞). The
following are equivalent for A, B ∈ (Mm×n (F), · p ), equipped with Schatten
p-norm:
(i) A ⊥ B.
(ii) tr(P p−1 U B ∗ ) = 0 where A = P U is the polar decomposition (P ∈ Mm (F) is
positive semidefinite and U ∈ Mm×n (F) satisfies U U ∗ = Im ).
278 L. Arambašić et al.
m ∗
Sketch of the Proof Assume that A = 0 and write P = i=1 σi (A)xi xi for some
orthonormal basis (xi )i . One can calculate that
1
m
1
T := p P p−1
U = p σi (A)p−1 xi xi∗ U
Ap Ap i=1
k
A(k) := σi (A) = max | tr(U ∗ AV )|; A ∈ Mn (C).
U,V ∈Mn×k (C)
i=1 U ∗ U =V ∗ V =Ik
The special importance of Ky Fan norms is their dominance property: Given two
matrices A, B ∈ Mn (C), then A ≤ B for every unitarily invariant norm if
and only if this inequality holds for all n Ky Fan norms (see [43, Corollary 3.5.9]).
Grover in [39] gave the following characterization of B-J orthogonality in Ky Fan
norms, which we state without a proof:
Proposition 3.6 (Cf. [39, Theorem 1.1 and Theorem 3.2]) Let 1 ≤ k ≤ n, let
A = U |A| be a polar decomposition of A ∈ Mn (F), and let B ∈ Mn (F). If there
exist k orthonormal vectors u1 , . . . , uk ∈ Fn such that
k
|A|ui = σi (A)ui for i = 1, . . . , k and ui , U ∗ Bui = 0,
i=1
then A is B-J orthogonal to B in · (k) . If in addition σk (A) > 0, then the converse
is also true.
The condition (ii) from Corollary 3.3 can be restated as:
(ii ) There exists a normalized vector x ∈ Fn such that Ax = A and Ax ⊥ Bx.
Now (ii ), unlike (ii), makes sense not only in inner product spaces but also in
general normed spaces. It is easy to see that, in any norm · on Fn , (ii ) implies
(i) of Corollary 3.3 (i.e., A ⊥ B) provided that Mn (F) is equipped with the induced
operator norm T := supx=1 T x. Bhatia and Šemrl conjectured in [20] that,
conversely, (i) implies (ii ) in any norm on Fn .
Birkhoff–James Orthogonality: Characterizations, Preservers, and Orthogonality Graphs 279
The case of commutative C ∗ -algebras was discussed by Kečkić. Recall that for
each unital commutative C ∗ -algebra A there is a compact Hausdorff space K such
that A is ∗-isomorphic to the C ∗ -algebra C(K) of all continuous complex valued
functions on K with the maximum norm f = max{|f (t)|; t ∈ K}, see [23,
II.2.2.4 and II.1.1.3.(2)]. If A is a nonunital commutative C ∗ -algebra, then there
is a noncompact locally compact Hausdorff space such that A is isomorphic
to C0 (), the C ∗ -algebra of all continuous complex functions on vanishing at
“infinity”. If K = ∪ {s0 } is the one-point compactification of , then we can
identify C0 () with the C ∗ -subalgebra {f ∈ C(K); f (s0 ) = 0} of C(K). In this
way, since B-J orthogonality of two elements “happens” in the subspace spanned by
these two elements, it is enough to obtain the characterization of B-J orthogonality
in C(K).
Let us begin with some examples.
280 L. Arambašić et al.
Example
(a) Let f, g ∈ C(K) be such that there is t0 ∈ K satisfying |f (t0 )| = f and
g(t0 ) = 0. Then
so f ⊥ g.
(b) Let f, g ∈ C(K). It follows from the first example that f ⊥ (f 2 − |f |2 )g.
Observe the similarity with Corollary 3.2.
(c) Let f, g ∈ C([0, 1]) be defined as f (t) = 1 and g(t) = e2πit . It is easy to see
that f ⊥ g. This example shows that the sufficient conditions stated in the first
example are not necessary.
Theorem 3.7 (Cf. [55, Corollary 2.1]) Let C(K) be the Banach space of all
continuous complex valued functions on a compact Hausdorff space K, with the
norm f = max{|f (t)|; t ∈ K}. For f ∈ C(K) define
Ef = {t ∈ K; |f (t)| = f }.
that is, if and only if it holds that, under every rotation around the origin, the set
{e−i arg f (t )g(t); t ∈ Ef } contains at least one value with nonnegative real part.
Since f (t)g(t) = f e−i arg f (t ) g(t) for each t ∈ Ef , this is equivalent to (ii).
) Since the convex hull of the set F is the set of points of the form
K f (t)g(t) dλ(t) where λ is a probability measure supported on a finite subset
Birkhoff–James Orthogonality: Characterizations, Preservers, and Orthogonality Graphs 281
For c0 , that is, the space of all complex-valued sequences which converge to zero,
and equipped with supremum norm, the characterization of B-J orthogonality is, as
expected, easier.
Proposition 3.9 (Cf. [54, Example 1.7]) The following are equivalent for x =
(xn )n , y = (yn )n ∈ c0 :
(i) x ⊥ y.
(ii) There does not exist an acute open angle D = {z; α < arg z < β} with
β − α < π, such that xn yn ∈ D for all those n for which |xn | = x holds.
Some other classical spaces are p and Lp (μ) with p ∈ [1, ∞). Here Lp (μ)
denotes the space of complex valued functions on a measurable space with a
positive measure μ whose p-th degree is summable. In these spaces B-J orthog-
onality was characterized by James [47] and Kečkić [54] by means of supporting
functionals and Gateaux derivatives of the norm. We remark that James considered
the real case and = [0, 1] with Lebesgue measure μ only, but the proof can be
transferred easily to the complex case and an arbitrary measurable space .
282 L. Arambašić et al.
Proposition 3.10 (Cf. [47, Example 8.1]) Let x = (xj )j ∈N and y = (yj )j ∈N
belong to p with p ≥ 1. Then x ⊥ y if and only if one of the following conditions
holds:
(i) p = 1 and
xj
yj ≤ |yj |;
|xj |
xj =0 xj =0
Then for each n, φn : T $→ T xn , xn is a state. Recall that the set of states is closed
in w∗ -topology, hence compact by the Banach–Alaoglu theorem.
Proposition 3.12 (Cf. [10, Theorem 2.7], [21, Proposition 4.1]) The following is
equivalent in a C ∗ -algebra A :
(i) a ⊥ b.
(ii) There exists a state φ on A such that
which gives a ⊥ b.
Conversely, suppose that a ⊥ b. By Gelfand–Naimark theorem, we embed A
into B(H) and then use a sequence of states φn : T $→ T xn , xn provided by
Theorem 3.1. Due to w∗ -compactness, this sequence has a subsequence which w∗ -
converges to a desired state φ.
By using the linking algebra of a Hilbert C ∗ -module, this result can be extended
from C ∗ -algebras to Hilbert C ∗ -modules. The concept of a Hilbert C ∗ -module has
been introduced by Kaplansky [52] and Paschke [71] in an investigation of right
modules over a C ∗ -algebra which possess a C ∗ -valued inner product respecting the
module action. More precisely, a Hilbert C ∗ -module X over a C ∗ -algebra A is a
right A -module equipped with an A -valued inner product · , · : X × X → A
which satisfies
(1) x, αy + βz = αx, y + βx, z for x, y, z ∈ X , α, β ∈ C,
(2) x, ya = x, ya for x, y ∈ X , a ∈ A,
(3) x, y∗ = y, x for x, y ∈ X ,
(4) x, x ≥ 0 for x ∈ X ; if x, x = 0 then x = 0,
284 L. Arambašić et al.
√
and which is a Banach space with respect to the norm defined as x = x, x.
We say that a Hilbert C ∗ -module X over a C ∗ -algebra A is full if the inner products
of elements from X span a dense subset in A, in short, if X , X = A . A left
Hilbert C ∗ -module can be defined in a similar way. Every right Hilbert A -module
is also a left Hilbert C ∗ -module over a C ∗ -algebra K(X ) of ‘compact’ operators
on X , that is, the C ∗ -algebra spanned by the operators
θx,y , x, y ∈ X , defined as
θx,y (z) = xy, z. It is easy to show that x = θx,x .
Besides Hilbert spaces, one of the most important examples of Hilbert C ∗ -
modules are C ∗ -algebras. If A is a C ∗ -algebra, then we can regard it as a Hilbert
C ∗ -module over itself with the algebra multiplication as a (right) module action and
an inner product defined as a, b = a ∗ b.
The linking algebra L(X ) of a Hilbert A -module X is defined as the C ∗ -algebra
of all ‘compact’ operators acting on the Hilbert A -module A ⊕ X . It can be written
in the matrix form
Ta ly
L(X ) = ; a ∈ A , x, y ∈ X , T ∈ K(X ) ,
rx T
Ta 0
and (X∗ Y ) = 0. If φ : A → C is defined by the formula φ(a) = ,
0 0
we easily see that φ is a state on A satisfying φ(x, x) = x2 and φ(x, y) = 0.
The converse is similar to the C ∗ -algebra case.
If H and K are Hilbert spaces, then the Banach space B(H, K) of all bounded
linear operators from H into K is a Hilbert C ∗ -module over B(H). Therefore, as a
corollary of the previous theorem we get a generalization of Theorem 3.1.
Corollary 3.14 Let A, B ∈ B(H, K). Then A ⊥ B if and only if there is a
sequence of normalized vectors (xn )n in H such that limn→∞ Axn = A and
limn→∞ B ∗ Axn , xn = 0.
Birkhoff–James Orthogonality: Characterizations, Preservers, and Orthogonality Graphs 285
We remark that Singla [85, Theorem 1.3] very recently extended Corollary 3.14 and
classified when an operator A ∈ B(H, K), which is not far away from compact
operators, is B-J orthogonal to a subspace B ⊆ B(H, K). This was achieved as
a consequence of his study of norm’s directional derivatives D+ at elements in a
C ∗ -algebra (which are not far from some, possibly nonproper, closed ideal).
x, y = 0 #⇒ x ⊥s y #⇒ x ⊥ y.
if and only if x − y0 , y = 0 for each y ∈ Y. This holds also for general normed
spaces:
Proposition 4.1 Let X be a normed space. Among all the vectors from a subspace
Y ⊆ X a vector y0 ∈ Y is the best approximate to x in a sense of (11) precisely
when x − y0 is B-J orthogonal to Y.
Proof From the definition of B-J orthogonality we have that (x − y0 ) ⊥ Y if and
only if (x − y0 ) + y ≥ x − y0 for every y ∈ Y.
In infinite-dimensional Banach spaces the sum of two closed subspaces may fail
to be closed.
For a classical concrete example, consider the Hilbert space H := 2
with en n≥0 as a standard orthonormal basis, let M be a closed subspace spanned
by pairwise orthogonal vectors e2n + n+1 1
e2n+1 n and let N be a closed subspace
spanned by orthonormal vectors e2n n . Then M ∩ N = 0 and M + N is dense
in 2 , since
it contains en for any n ≥ 0. However, M + N does not contain the
vector n≥0 n+1 1
e2n+1 ∈ 2 , because otherwise the only way to write it would be
n≥0 (e2n + n+1 e2n+1 ) − n≥0 e2n ∈
1
n≥0 e2n , but / 2 .
However, it is easily seen that, in Hilbert spaces, the sum of two orthogonal,
closed subspaces is again closed. Again, this can be generalized easily to general
Banach spaces (see Anderson [8, Remark 1.3]). To avoid misunderstanding, we say
288 L. Arambašić et al.
A − U AU ∗ = (A − λI ) − U (A − λI )U ∗ ≤ 2A − λI .
A0 + λI = A + (λ + λ0 )I ≥ A + λ0 I = A0 ,
NX − XN + S ≥ S.
This result has been greatly extended in various directions, see [25, 34, 56].
It was shown in [10] that B-J orthogonality provides a convenient criterion for
two elements of a normed space X to satisfy the equality in the triangle inequality.
Proposition 4.4 (Cf. [10, Proposition 4.1]) Let X be a real or complex normed
space, x, y ∈ X . Then the following conditions are equivalent:
(i) x + y = x + y;
(ii) x ⊥ (yx − xy);
(iii) y ⊥ (yx − xy).
Proof It follows from the Hahn–Banach theorem that x + y = x + y if and
only if there exists a norm-one linear functional f : X → F such that f (x) = x
and f (y) = y, see [69, Theorem 2]. If x = 0, then the latter condition can be
restated as f (x) = x and f (yx − xy) = 0. Such a linear functional f exists
290 L. Arambašić et al.
if and only if x ⊥ (yx − xy), so the equivalence (i) ⇔ (ii) is proved. Similarly
for (i) ⇔ (iii).
This proposition, together with the obtained characterizations of B-J orthogo-
nality in different normed spaces, gives characterizations of the case of equality in
the triangle inequality, which were directly obtained earlier in [9] and [69]. Let us
formulate the version for C ∗ -algebras.
Corollary 4.5 (Cf. [69, Theorem 1], [9, Remark 2.2]) Let A be a C ∗ -algebra and
a, b ∈ A be nonzero. Then the equality a + b = a + b holds if and only if
there is a state φ on A such that φ(a ∗ b) = ab.
Proof Suppose that a + b = a + b holds. Then, by the previous proposition,
a ⊥ (ba − ab) and, by Proposition 3.12, there is a state φ on A such that
φ(a ∗ a) = a2 and φ(a ∗ (ba − ab)) = 0. These two relations give φ(a ∗ b) =
ab.
Conversely, suppose there is a state φ on A such that φ(a ∗ b) = ab. From
the Bunyakovsky-Cauchy-Schwarz inequality applied to (a, b) $→ φ(a ∗ b) we get
wherefrom φ(a ∗ a) = a2 and φ(b∗ b) = b2 . Thus, φ(a ∗ (ba − ab)) =
bφ(a ∗ a) − aφ(a ∗ b) = 0, so, by Proposition 3.12, a ⊥ (ba − ab).
Among other applications of B-J orthogonality we mention also Cheng–
Mashreghi–Ross’s [28] bounds for the zeros of an analytic function on a disk,
where B-J orthogonality on p with p ∈ (1, ∞) is used, see Proposition 3.10 above.
It is an easy exercise to prove that a linear map between inner product spaces which
preserves orthogonality must be a scalar multiple of an isometry. This result can
also be generalized to B-J orthogonality. Koldobsky [57] was the first to show that
linear preservers of B-J orthogonality on real normed spaces are scalar multiples of
isometries. Later, Blanco and Turnšek [24] extended his result to complex normed
spaces. We present a simplified proof due to Wójcik [94, Theorem 2.1]. To counter
nonsmooth norms, Wójcik relied on the following standard identities:
Proposition 5.1 (Cf. [33, Theorem 18]) In the notations of Sect. 2.1, for any
nonzero x, y ∈ X it holds that
x + ty − x
≤ Re φt (y) = D+ (x + ty; y). (13)
t
x+ty−x
Moreover, for each fixed t > 0 and each x we have D+ (x; y) ≤ t . By
replacing x with x + ty we get, in combination with (13),
Both the left and the right sides converge to D+ (x; y) as t % 0, which gives the
second estimate of (12); the first one is completely similar.
Theorem 5.2 (Cf. [24, 57, 94]) Let T : X → Y be a (conjugate) linear map
between two normed spaces X , Y over the field F. Then the following are equiv-
alent:
(i) T preserves B-J orthogonality, i.e., x ⊥ y #⇒ T x ⊥ T y for every x, y ∈ X .
(ii) T is a scalar multiple of an isometry, i.e., there is γ ≥ 0 such that T x =
γ x for every x ∈ X .
Sketch of the Proof Injectivity is straightforward: Indeed, assume T c = 0 for some
nonzero c. Notice that 0 = c +(−1)c ≥ c, so there exists a small enough ε > 0
such that if x < ε, then (c + x) + (−1)c ≥ c + x, i.e., (c + x) ⊥ c. By
Proposition 1.4, we can find a scalar λ = λx with (c + x) ⊥ (λ(c + x) + c), giving
T x = T (c + x) ⊥ T (λ(c + x) + c) = λσ T x, where σ : F → F is either identity or
a complex conjugation. This is possible only if λ = 0 (contradicting (c + x) ⊥ c)
or if T x = 0. Hence, if T is not injective, then T = 0.
Choose linearly independent x, y ∈ X and fix φ ∈ J (x). By Proposition 1.4 with
α := φ(y)
x , we have x ⊥ (αx − y), so also T x ⊥ (α T x − T y). By Proposition 1.4
σ
x 2 3
= D− (T x; T y) , D+ (T x; T y) .
T x
T x
0 = D+ (T x; T y) − D+ (x; y)
x
T (x + ty) − T x T x x + ty − x
= lim −
t %0 t x t
T (x + ty) · x − T x · x + ty
= lim
t %0 t · x
b(x + ty) − b(x) b(x + ty) − b(x)
= lim · x + ty = x · lim .
t %0 t t %0 t
b(x+ty)−b(x)
Likewise one shows that limt 30 t = 0. Hence, the function b is
constant.
Blanco and Turnšek considered in [24] (possibly nonlinear) bi-preservers of B-J
orthogonality on projective Banach space PX := {[x] = Fx; x ∈ X \ {0}}. Their
result is the following:
Theorem 5.3 (Cf. [24, Corollary 3.4]) Let X be an infinite-dimensional, reflexive,
smooth Banach space. If : PX → PX is a bijective map such that
([x]) = [U x].
The smoothness assumption is indispensable here, for example, in Pc0 there exists
bijective bi-preservers which are not induced by (conjugate) linear map, let alone
isometry (see [24, Example 3.5]). It turns out, however, that the assumption about
infinite dimensionality is not required. We will show this in our last chapter.
We also remark that there do exist complex reflexive Banach spaces which
are conjugate-linear isometric but are not even isomorphic. The examples were
constructed by Bourgain [27] and Kalton [51].
Birkhoff–James Orthogonality: Characterizations, Preservers, and Orthogonality Graphs 293
Below we collect, mostly without proofs, some partial answers; we refer to [14]
for proofs. The first one is merely a restatement of Corollary 2.12.
Proposition 6.1 A normed space X over F is infinite-dimensional if and only if
ˆ ) contains an infinite clique.
(X
Di-orthograph alone can compute the dimension of the underlying space.
Proposition 6.2 (Cf. [14, Lemma 2.3]) Let X be a normed space over F and
n ≥ 2. Then the following statements are equivalent.
ˆ )
(i) For any (n − 1)-tuple of lines (x1 , . . . , xn−1 ), one can always find xn ∈ (X
with xi ⊥ xn , 1 ≤ i ≤ n − 1.
(ii) dim X ≥ n.
Corollary 6.3 A normed space X has dimension n < ∞ if and only if its di-
orthograph (Xˆ ) satisfies item (i) with k = 1, . . . , n − 1 and does not satisfy item
(i) for larger k.
Di-orthograph can detect the presence of nonsmooth points. Compare with
Theorem 2.6(ii) where smoothness was related to right uniqueness and right
additivity of B-J orthogonality.
Proposition 6.4 (Cf. [14, Lemma 2.5]) Let X be a normed space over F with 2 ≤
dim X = n < ∞. Then the norm in X is nonsmooth if and only if there exist
ˆ ) and two additional distinct lines yn , yn such that
n − 1 lines x1 , . . . , xn−1 ∈ (X
xi ⊥ xj for 1 ≤ i < j ≤ n − 1, and xi ⊥ yn , xi ⊥ yn for 1 ≤ i ≤ n − 1.
294 L. Arambašić et al.
Di-orthograph alone can also detect if the norm is strictly convex (compare again
with Theorem 2.6 (i) or (iv)). Given a vertex z in a di-orthograph , ˆ let Nz :=
ˆ ˆ ˆ
{v ∈ ; (z, v) ∈ E()} denote its neighborhood; here E() is the set of all
ˆ Note that Nz = ∅ if the underlying normed space is at least
directed edges of .
two-dimensional.
Proposition 6.5 (Cf. [14, Lemma 2.6]) A normed space X with dim X ≥ 2 over
ˆ X)
ˆ ) → 2(
F is strictly convex if and only if the function (X which maps a vertex
z to its neighborhood Nz is injective.
How much information on the norm is encoded in the di-orthograph (X ˆ )? Clearly,
if A : X → Y is a linear bijective isometry between two Banach spaces, then
A induces an isomorphism of di-orthographs (X ˆ ) and (Y).
ˆ We show next the
converse of this fact. Note that this problem is closely related to characterization of
preservers of B-J orthogonality which was discussed in Sect. 5. Namely, bijective bi-
preservers of B-J orthogonality between projective Banach spaces P(X ) and P(Y)
are exactly isomorphisms between (X ˆ ) and (Y).
ˆ
In the lemma below, a curve is a subset in C, which is the image of a path, i.e.,
the image of a continuous map r : [a, b] → C where [a, b] ⊆ R is an interval with
at least two different points. We say that a function f : C → C is bounded from
above on a subset ⊆ C if supz∈ |f (z)| < ∞.
Lemma 6.6 (Cf. [14, Lemma 3.1]) Suppose that a nonzero ring homomorphism
σ : C → C is bounded from above on a curve ⊆ C with more than one point.
Then σ is continuous, hence either identity or a complex conjugation.
Sketch of the Proof The full proof is relatively long and can be found in [14].
We first manipulate by a finite number of rotations/translations/taking unions to
obtain a closed curve ˆ which separates the complex plane. Observe that the ring
homomorphism σ remains bounded from above on . ˆ Then we use the fact, inspired
ˆ ˆ
by a paper of Simon and Taylor [81], that − := {γ1 − γ2 ; γ1 , γ2 ∈ } ˆ has a
ˆ ˆ
nonempty interior. Since σ is clearly bounded from above also on − , it is hence
bounded on an open set, and hence it must be continuous (see, e.g., [1, Corollary 5,
p. 15]).
We remark that, with the help of dimension theory for separable metric spaces
(see, e.g., a monograph by Hurewitz and Wallman [44]), an even more general result
was obtained by Shchepin E[80]: If 1 , . . . , n are n ≥ 1 compact connected subsets
n
in Rn such that (i) 0 ∈ i=1 i and (ii) there exist n linearly independent points
ai ∈ i , then the sum ni=1 i := {γ1 + · · · + γn ; γi ∈ i } ⊆ Rn is of dimension
at least n and therefore it contains an open ball (see [44, Theorem IV.3]).
Birkhoff–James Orthogonality: Characterizations, Preservers, and Orthogonality Graphs 295
It was shown by Rätz [74] (see also Sundaresan [87, Lemma 1]) that B-J
orthogonality in real normed spaces is Thalesian, that is, if x, y are B-J orthogonal
vectors in a real normed space, then for every λ0 > 0 there exists a scalar α
such that (x + αy) ⊥ (λ0 x − αy). This fact was used by Wójcik [94, Proof of
Theorem 3.1] to give an alternative proof that linear maps which preserve B-J
orthogonality between real normed spaces are scalar multiples of isometries. The
main idea of the proof of the next lemma comes from Wójcik’s paper (see [94,
Proof of Theorem 3.1]) and uses a partial extension of Thalesian property for B-J
orthogonality in complex normed spaces: we show that if normalized vectors x, y
are mutually B-J orthogonal, then there exists a curve ⊆ C with more than one
point, such that for every λ ∈ we can find α ∈ C with (x + αy) ⊥ (λx − αy).
Recall that an additive map : X → Y between F-vector spaces X and Y is σ -
quasilinear if (λx) = σ (λ) (x) holds where σ : F → F is a field homomorphism
(if σ is surjective, such maps are semilinear).
Proposition 6.7 (Cf. [14, Lemma 3.4]) Let X , Y be smooth normed complex
spaces of dimension at least two, and let σ : C → C be a field homomorphism.
If a nonzero σ -quasilinear map : X → Y preserves B-J orthogonality, then σ is
identity or a complex conjugation.
Sketch of the Proof By Proposition 2.13(i), there exist two mutually B-J orthogo-
nal normalized vectors x, y ∈ X such that (x) = 0. Identify the two-dimensional
subspace spanC {x, y} with (C2 , · ), so that x = (1, 0) and y = (0, 1). By (5), a
C-linear supporting functional at a point (1, α) ∈ C2 equals
∗
∂(1,α) ∂(1,α)
fα = ∂z1 , ∂z2 , (zk = xk + iyk ).
∗
∂(1,α) ∂(1,α)
functional equals the limit of ∂z1 , ∂z2 as R 8 α → ∞. Hence,
∂(1,α)
∂z2
wα = ∂(1,α)
αx − αy
∂z1
∂(1,α)
∂z2
λ(α) := ∂(1,α)
α,
∂z1
Proof By Proposition 6.1, dim X < ∞ if and only if dim Y < ∞. It then follows
from Corollary 6.3 that dim X = dim Y.
Birkhoff–James Orthogonality: Characterizations, Preservers, and Orthogonality Graphs 297
but their images T [x], T [y], T [z] span a three-dimensional space. Then we can
choose a norm-attaining normalized functional g : Y → F which annihilates T [x]
and T [y] but not T [z]. Clearly, the line spanned by such g belongs to the range of S,
so [g] = S[f ], and f annihilates [x] and [y] but not the line [z] contradicting (15).
Indeed, T is a bijective morphism of projective spaces. By the (nonsurjective
version of) Fundamental Theorem of Projective Geometry (see Faure [35]), there
exist a field isomorphism σ : F → F and a σ -quasilinear map V : X → Y such that
T [x] = [V x] for x ∈ X \ {0}. Clearly, V must be orthogonality preserving, so, by
Proposition 6.7, σ is either identity or a complex conjugation. By Theorem 5.2, the
(conjugate) linear orthogonality preserving map V : (X , · X ) → (Y, · Y ) is a
scalar multiple of an isometry (i.e., there exists a scalar μ ∈ C such that U := μV
satisfies U xY = xX ).
We remark that, if X is finite-dimensional and smooth, then in the theorem
above the smoothness of Y need not be assumed in advance; it follows from
Proposition 6.4 (after Corollary 6.3 establishes that dim X = dim Y). Note also that
Theorem 6.8 is not valid for dim X = dim Y = 2 because of the Radon planes; see
also [14, Example 4.1] or [88, Theorem 4.1]. Here is a restatement of Theorem 6.8
in terms of preservers of B-J orthogonality.
Corollary 6.9 Let X , Y be normed spaces over F, with 3 ≤ dim X < ∞. Assume
also that X is smooth. If there exists a bijection : PX → PY such that
We have similar results valid in Banach spaces rather than their projectivizations.
Let us show first that, in smooth spaces, B-J orthogonality can check for linear
independence among two vectors:
Lemma 6.10 Let x, y be nonzero vectors in a smooth normed space X over F.
Then the following conditions are equivalent:
(i) Fx = Fy.
(ii) There exists a vector z ∈ X such that z ⊥ x but z ⊥ y.
Proof (i) #⇒ (ii). Choose a normalized linear functional f which attains its norm
at some vector z ∈ X and which satisfies f (x) = 0 and f (y) = 0. Then f is the
unique supporting functional at z, and (ii) follows from Proposition 1.4.
(ii) #⇒ (i). Follows from homogeneity of B-J orthogonality for any normed
space X .
Example The above lemma does not hold in nonsmooth spaces. Say, in (R3 , · ∞ )
we have that u = (1, 1/2, 0) and v = (1, 1/3, 0) are independent, and yet they are
B-J orthogonal to the same vectors: Nu = Nv = {0} × R2 , and the same vectors are
B-J orthogonal to them: v N = u N = {(a, b, c); (|c| ≥ max{|a|, |b|}) ∨ (a + b =
0 ∧ |c| ≤ |b|)}, and also they are mutually B-J orthogonal to the same set of vectors,
i.e., to {(0, b, c); |b| ≤ |c|}.
Corollary 6.11 Let X , Y be smooth normed spaces over F, with 3 ≤ dim X < ∞.
If there exists a bijection : X → Y such that
x ⊥ y ⇐⇒ (x) ⊥ (y),
Acknowledgments The authors would like to express their deep gratitude to Professor Rajna
Rajić for many insightful conversations and suggestions while preparing this manuscript. They are
also grateful to the referee for suggesting additional references.
The work of the second and the fourth authors is supported by the RSF grant 21-11-00283.
The third author acknowledges the financial support from the Slovene Research Agency, ARRS
(research programs No. P1-0222, No. P1-0285, and research project No. N1-0210).
References
23. B. Blackadar, Operator Algebras. Theory of C ∗ -Algebras and von Neumann Algebras.
Encyclopaedia of Mathematical Sciences, vol. 122 (Springer, Berlin and Heidelberg, 2006)
24. A. Blanco, A. Turnšek, On maps that preserve orthogonality in normed spaces. Proc. Roy. Soc.
Edinburgh Sect. A 136(4), 709–716 (2006)
25. A. Blanco, A. Turnšek, On the converse of Anderson’s theorem. Linear Algebra Appl. 424(2–
3), 384–389 (2007)
26. F. Bohnenblust, A characterization of complex Hilbert spaces. Portugal Math. 3, 103–109
(1942)
27. J. Bourgain, Real isomorphic complex Banach spaces need not be complex isomorphic. Proc.
Am. Math. Soc. 96(2), 221–226 (1986)
28. R. Cheng, J. Mashreghi, W.T. Ross, Birkhoff–James orthogonality and the zeros of an analytic
function. Comput. Methods Funct. Theory 17, 499–523 (2017)
29. W.J. Davis, Separable Banach spaces with only trivial isometries. Rev. Roumaine Math. Pures
Appl. 16, 1051–1054 (1971)
30. M.M. Day, Polygons circumscribed about convex closed curves. Trans. Am. Math. Soc. 62,
315–319 (1947)
31. M.M. Day, Some characterizations of inner-product spaces. Trans. Am. Math. Soc. 62, 320–
337 (1947)
32. R. Deville, G. Godefroy, V. Zizler, Smoothness and Renormings in Banach Spaces. Pitman
Monographs and Surveys in Pure and Applied Mathematics, vol. 64 (Longman Scientific &
Technical, Harlow, 1993)
33. S.S. Dragomir, S. Boriotti, D. Dennis, Semi-inner Products and Applications (Nova Science
Publishers, Hauppauge, 2004)
34. H.-K. Du, Another generalization of Anderson’s theorem. Proc. Am. Math. Soc. 123(9), 2709–
2714 (1995)
35. C.-A. Faure, An elementary proof of the fundamental theorem of projective geometry. Geom.
Dedicata. 90, 145–151 (2002)
36. F.A. Ficken, Note on the existence of scalar products in normed linear spaces. Ann. of Math.
45(2), 362–366 (1944)
37. G. P. Gehér, An elementary proof for the non-bijective version of Wigner’s theorem. Phys. Lett.
A 378(30–31), 2054–2057 (2014)
38. Y. Gordon, R. Loewy, Uniqueness of () spaces. Math. Ann. 241(2), 159–180 (1979)
39. P. Grover, Orthogonality of matrices in the Ky Fan k-norms. Linear Multilinear Algebra 65(3),
496–509 (2017)
40. P. Grover, S. Singla, Birkhoff–James orthogonality and applications: a survey, in Operator
Theory, Functional Analysis and Applications. Operator Theory: Advances and Applications,
vol. 282 (Springer International Publishing, New York, 2021), pp. 293–315
41. P. Hájek, J. Talponen, Smooth approximations of norms in separable Banach spaces. Q. J.
Math. 65, 957–969 (2014)
42. P.R. Halmos, A Hilbert Space Problem Book (Springer, Berlin, 1982)
43. R.A. Horn, C.R. Johnson, Topics in Matrix Analysis (Cambridge University Press, Cambridge,
1991)
44. W. Hurewics, H. Wallman, Dimension Theory (Princeton University Press, Princeton, 1941)
45. R.C. James, Orthogonality in normed linear spaces. Duke Math. J. 12, 291–302 (1945)
46. R.C. James, Inner product in normed linear spaces. Bull. Am. Math. Soc. 53, 559–566 (1947)
47. R.C. James, Orthogonality and linear functionals in normed linear spaces. Trans. Am. Math.
Soc. 61, 265–292 (1947)
48. K. Jarosz, Any Banach space has an equivalent norm with trivial isometries. Isr. J. Math. 64,
49–56 (1988)
49. P. Jordan, J. von Neumann, On inner products in linear metric spaces. Ann. Math. 36(2), 719–
723 (1935)
50. S. Kakutani, Some characterizations of Euclidean space. Jap. J. Math. 16, 93–97 (1939)
51. N.J. Kalton, An elementary example of a Banach space not isomorphic to its complex
conjugate. Canad. Math. Bull. 38(2), 218–222 (1995)
Birkhoff–James Orthogonality: Characterizations, Preservers, and Orthogonality Graphs 301
52. I. Kaplansky, Modules over operator algebras. Am. J. Math. 75, 839–858 (1953)
53. T. Kato, Perturbation Theory for Linear Operators (Springer, Berlin, 1980)
54. D.J. Kečkić, Orthogonality in S1 and S∞ spaces and normal derivations. J. Oper. Theory
51(1), 89–104 (2004)
55. D.J. Kečkić, Orthogonality and smooth points in C (K) and Cb (). Eur. Math. J. 3(4), 44–52
(2012)
56. F. Kittaneh, Normal derivations in norm ideals. Proc. Am. Math. Soc. 123(6), 1779–1785
(1995)
57. A. Koldobsky, Operators preserving orthogonality are isometries. Proc. Roy. Soc. Edinburgh
Sect. A 123, 835–837 (1993)
58. A.S. Lewis, The convex analysis of unitarily invariant matrix functions. J. Convex Anal. 2(1–
2), 173–183 (1995)
59. A.S. Lewis, Derivatives of spectral functions. Math. Oper. Res. 21(3), 576–588 (1996)
60. C.K. Li, H. Schneider, Orthogonality of matrices. Linear Algebra Appl. 347, 115–122 (2002)
61. P.K. Lin, A remark on the Singer-orthogonality in normed linear spaces. Math. Nachr. 160,
325–328 (1993)
62. E.R. Lorch, On certain implications which characterize Hilbert spaces. Ann. Math. 49(3), 523–
532 (1948)
63. B. Magajna, On the distance to finite-dimensional subspaces in operator algebras. J. London
Math. Soc. 47(2), 516–532 (1993)
64. V.M. Manuilov, E.V. Troïtsky, Hilbert C ∗ -Modules. Translations of Mathematical Mono-
graphs, vol. 22 (American Mathematical Society, Providence, 2005)
65. H. Mizuguchi, The differences between Birkhoff and isosceles orthogonalities in Radon planes.
Extracta Math. 32(2), 173–208 (2017)
66. L. Molnár, Orthogonality preserving transformations on indefinite inner product spaces:
generalization of Uhlhorn’s version of Wigner’s theorem. J. Funct. Anal. 194(2), 248–262
(2002)
67. M.S. Moslehian, A. Zamani, Characterizations of operator Birkhoff-James orthogonality.
Canad. Math. Bull. 60(4), 816–829 (2017)
68. G.J. Murphy, C ∗ -Algebras and Operator Theory (Academic Press, San Diego, 1990)
69. R. Nakamoto, S.-E. Takahasi, Norm equality condition in triangular inequality. Sci. Math. Jpn.
55(3), 463–466 (2002)
70. J.A. Oman, Characterizations of inner product spaces, Ph.D. Thesis. Michigan State University,
Michigan, MI, 1969
71. W.L. Paschke, Inner product modules over B ∗ -algebras. Trans. Am. Math. Soc. 182, 443–468
(1973)
72. K. Paul, D. Sain, Birkhoff-James orthogonality and its application in the study of geometry of
Banach space, in Advanced Topics in Mathematical Analysis (CRC Press, Boca Raton, 2019),
pp. 245–284
73. R.R. Phelps, Convex Functions, Monotone Operators and Differentiability (Springer, Berlin,
1989)
74. J. Rätz, On orthogonally additive mapping. Aequationes Math. 28, 35–49 (1985)
75. F. Rellich, Perturbation Theory of Eigenvalue Problems. Lecture Notes (New York University,
New York, 1953)
76. B.D. Roberts, On the geometry of abstract vector spaces. Tôhoku Math. J. 39, 42–59 (1934)
77. R.T. Rockafellar, Convex Analysis (Princeton University Press, Princeton, 1970)
78. S. Roy, S. Bagchi, D. Sain, Birkhoff–James orthogonality in complex Banach spaces and
Bhatia–Šemrl Theorem revisited (2021). arXiv:2109.12775
79. R. Schatten, Norm Ideals of Completely Continuous Operators (Springer, Berlin, 1960)
80. E.V. Shchepin, The dimension of a sum of curves. Uspekhi Mat. Nauk 30(4)(184), 267–268
(1975) (in Russian)
81. K. Simon, K. Taylor, Interior of sums of planar sets and curves. Math. Proc. Cambridge Philos.
Soc. 168(1), 119–148 (2020)
302 L. Arambašić et al.
82. I. Singer, Sur le L-problème de la théorie des moments dans les espaces de Banach. Acad.
Repub. Popul. Romîne. Bul. Şti. Secţ. Şti. Mat. Fiz. 9, 19–28 (1957)
83. I. Singer, Angles abstraits et fonctions trigonométriques dans les espaces de Banach. Acad.
Repub. Popul. Romîne. Bul. Şti. Secţ. Şti. Mat. Fiz. 9, 29–42 (1957)
84. I. Singer, Best Approximation in Normed Linear Spaces by Elements of Linear Subspaces
(Springer, New York, 1970)
85. S. Singla, Gateaux derivative of C ∗ norm. Linear Algebra Appl. 629, 208–218 (2021)
86. J.G. Stampfli, The norm of a derivation. Pac. J. Math. 33, 737–747 (1970)
87. K. Sundaresan, Orthogonality and nonlinear functionals on Banach spaces. Proc. Am. Math.
Soc. 34, 187–190 (1972)
88. R. Tanaka, On Birkhoff–James orthogonality preservers between real non-isometric Banach
spaces (2021). arXiv:2108.00655
89. R. Tanaka, Nonlinear equivalence of Banach spaces based on Birkhoff–James orthogonality. J.
Math. Anal. Appl. 505(1), 125444 (2022)
90. A.E. Taylor, A geometric theorem and its application to biorthogonal systems. Bull. Am. Math.
Soc. 53, 614–616 (1947)
91. A. Turnšek, On operators preserving James’ orthogonality. Linear Algebra Appl. 407, 189–195
(2005)
92. E.P. Wigner, Gruppentheorie und ihre Anwendung auf die Quantenmechanik der Atomspek-
trum (Fredrik Vieweg und Sohn, Braunschweig, 1931)
93. P. Wójcik, The Birkhoff orthogonality in pre-Hilbert C ∗ -modules. Oper. Matrices 10(3), 713–
729 (2016)
94. P. Wójcik, Mappings preserving B-orthogonality. Indag. Math. 30(1), 197–200 (2019)
95. K. Zi˛etak, On the characterization of the extremal points of the unit sphere of matrices. Linear
Algebra Appl. 106, 57–75 (1988)
Approximate Birkhoff-James
Orthogonality in Normed Linear Spaces
and Related Topics
Jacek Chmieliński
J. Chmieliński ()
Department of Mathematics, Pedagogical University of Krakow, Kraków, Poland
e-mail: [email protected]
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 303
R. M. Aron et al. (eds.), Operator and Norm Inequalities and Related Topics,
Trends in Mathematics, https://doi.org/10.1007/978-3-031-02104-6_9
304 J. Chmieliński
In an inner product space (H, ·|·) with the standard orthogonality relation: x⊥y ⇔
x|y = 0, we have an equally natural notion of the approximate orthogonality or,
more specifically, the ε-orthogonality (with a given ε ∈ [0, 1)):
i.e., x⊥ε y. Since the relations ⊥ and ⊥ε are here symmetric, the roles of x, y and z
in the above statement can be interchanged.
Although inner product spaces are the most natural venue for orthogonality,
analogous relations may be considered also in normed linear spaces (cf. e.g. the
survey [1]). Among the most natural, with clear geometrical background, are the
isosceles orthogonality:
x⊥i y ⇐⇒ x + y = x − y
Many other, not considered here, orthogonality relations are defined, including
axiomatic definitions in linear spaces or other structures (cf. [24, 26, 44]).
Birkhoff-James Orthogonality One of the most important orthogonality relation
in a normed space, is the Birkhoff-James orthogonality (sometimes called the
Birkhoff orthogonality). This concept is well known, extensively studied (cf.
[1, 2, 6, 29–31]) and crucial for the present paper.
Let X be a real normed linear space and x, y ∈ X . The Birkhoff-James
ortohogonality (BJ-orthogonality) of x and y (in the given order) is defined as
follows:
For a fixed x ∈ X \ {0} we consider the (always nonempty) set of its supporting
functionals:
where X ∗ denotes the dual space. With the Hahn-Banach theorem behind, the
following characterization can be given [30, Corollary 2.2]:
x + ty2 − x2
ρ± (x, y) = lim
t →0± 2t
x + ty − x
= x lim , x, y ∈ X .
t →0± t
Since not every norm is generated by an inner product, the following definition
is sometimes very much helpful. Due to [25, 35] (see also the monograph [23]) each
norm in a linear space X (over K ∈ {R, C}) admits a semi-inner product generating
this norm, that is a functional [·|·] : X × X → K satisfying the conditions:
306 J. Chmieliński
x⊥s y ⇐⇒ [y|x] = 0
and ρ can
It is known (cf. [23]) that for a real normed space X the values of ρ+ −
be obtained as the supremum or infimum, respectively, over the values taken by all
semi-inner products in X on a given pair of vectors. Namely,
ρ+ (x, y) = sup{[y|x] : [·|·] is a semi-inner product on X } (4)
and
ρ− (x, y) = inf{[y|x] : [·|·] is a semi-inner product on X }. (5)
The objective of this part of the paper is to define and develop the concept
of an approximate Birkhoff-James orthogonality. Various attempts and several
characterizations will be discussed. While many of the facts presented here have
been already known, there is some novelty in this section, both in results and in the
presentation and proofs. For the convenience of the reader, we tried to make this
part as detailed and self-contained as possible; only in few places we refer to results
from the outside. From now on, let X be a real normed space with dim X ≥ 2.
Approximate Birkhoff-James Orthogonality 307
There are at least two notions of ε-BJ-orthogonality (with given ε ∈ [0, 1)). The
first one was given by Dragomir [22]
x⊥ B y ⇐⇒ ∀ λ ∈ R : x + λy ≥ (1 − ε)x
ε
One can check that if the norm comes from an inner product then (6) coincides
exactly with (1), whereas ⊥ B gives ⊥ with η = 1 − (1 − ε)2 . The latter causes
η
ε
that it is sometimes more convenient
√ to use a modification of the Dragomir’s
definition, replacing 1 − ε by 1 − ε2 :
x⊥εD y ⇐⇒ ∀ λ ∈ R : x + λy ≥ 1 − ε2 x.
Some relationships between ⊥εB and ⊥ B were established in [10, 19, 38]. In
ε 2
particular, if X is a real normed space and ε ∈ 0, 12 , then (cf. [19])
x⊥εB y #⇒ x⊥ B y.
2ε
x⊥ B y #⇒ x⊥ηB y
ε
−1 ε
with η = δX ∗ 2 .
In the sequel we will consider mainly the approximate BJ-orthogonality ⊥εB ,
defined by (6).
It is known [18, Theorem 3.2] that if X is a real normed space, then for any semi-
inner product on X and ε ∈ [0, 1), there is ⊥εs ⊂ ⊥εB (in particular ⊥s ⊂ ⊥B ) and in
a smooth space both relations coincide, that is ⊥εs = ⊥εB (cf. also [50]).
In our considerations we will also use the following characterization of the ε-
BJ-orthogonality (actually, it is a special case of a more general result given in [18,
Theorem 3.1]):
Lemma 2.1 Let X be a real normed linear space and let ε ∈ [0, 1). Then, for
arbitrary x, y ∈ X we have
x⊥εB y ⇐⇒ ρ− (x, y) − εx y ≤ 0 ≤ ρ+ (x, y) + εx y. (7)
308 J. Chmieliński
x + λy2 − x2
−εx y ≤ for λ > 0
2λ
and
x + λy2 − x2
≤ εx y for λ < 0.
2λ
x + λy2 − x2
lim ≤ 2εxy
λ→0− λ
x + λy2 − x2
lim < 2(ε + γ )xy.
λ→0− λ
x + λy2 − x2
∀ λ ∈ [δ1 , 0) : < 2(ε + γ )xy,
λ
whence
Analogously, from the second inequality in (7) we get (for the same γ as above and
for some δ2 > 0)
Since γ was arbitrarily chosen from the interval (0, 1), letting γ → 0+ in (10) we
obtain
Obviously, the above inequality holds true also for λ = 0, thus finally we get x⊥εB y.
Now we formulate a useful auxiliary result (originally stated in [15]).
Lemma 2.2 Let X be a real normed space and let x, y ∈ X be such that x⊥εB y
(with some ε ∈ [0, 1)). Then, for each n ∈ N there exists a semi-inner product [·|·]n
in X such that
1
| [y|x]n | ≤ ε + x y. (11)
n
Proof Applying (7) and (4)–(5), for each integer n we may choose semi-inner
products [·|·]n and [·|·]n such that
1 1
[y|x]n < ε + x y and − ε+ x y < [y|x]n .
n n
1 1
− ε+ x y ≤ λn [y|x]n + (1 − λn ) [y|x]n ≤ ε + x y.
n n
1 1
− ε+ x y ≤ [y|x]n ≤ ε + x y.
n n
Characterizations of the Approximate Orthogonality The notion of approxi-
mate BJ-orthogonality ⊥εB , as defined by (6), can be characterized in various ways.
Now, we collect the already known as well as new characterizations in one theorem.
Theorem 2.3 (Characterization of ⊥εB ) Let X be a real normed linear space,
x, y ∈ X and let ε ∈ [0, 1). The following conditions are equivalent (each of them
can be treated as a definition of x⊥εB y):
(i) ∃ c > 0 ∀ λ ∈ [−c, c] : x + λy2 ≥ x2 − 2εx λy,
(ii) ∀ λ ∈ R : x + λy2 ≥ x2 − 2εx λy,
(iii) ∃ z ∈ Lin {x, y} : x⊥B z, z − y ≤ εy,
(iv) ∃ ϕ ∈ J (x) : |ϕ(y)| ≤ εy,
310 J. Chmieliński
λ0 1 1
x2 ≤ f =f λ0 + 1 − ·0
n n n
1 1
≤ f (λ0 ) + 1 − f (0)
n n
1 1
< x2 + 1 − x2 = x2 ,
n n
[y|x]n
zn := − x + y ∈ Lin {x, y},
x2
it is easy to see that x⊥s,n zn and since ⊥s,n ⊂ ⊥B , it follows that x⊥B zn .
Applying (11) we estimate zn − y and we get
1
x⊥B zn and zn − y ≤ ε + y. (12)
n
Notice that zn ≤ 2y whence the elements of the sequence (zn )n=1,2,... belong to
a closed ball in a two-dimensional space Lin {x, y}. Thus there exists z ∈ Lin {x, y}
and a subsequence (znk ) convergent to z. Finally, (12) and continuity of the norm
yield x⊥B z and z − y ≤ εy.
(iii)⇒(iv) Assuming (iii) and applying (3), there exists ϕ ∈ J (x) such that
ϕ(z) = 0. Therefore, |ϕ(y)| = |ϕ(z) − ϕ(y)| ≤ z − y ≤ εy, as claimed.
Approximate Birkhoff-James Orthogonality 311
so in (13) (or (14)) one can restrict to considering |λ| < (2 − ε) x
y .
Proposition 2.7 For an arbitrary real normed space X , x, y ∈ X and ε ∈ [0, 1)
we have:
ε
x ⊥B y ⇐⇒ x⊥ B y and x⊥εB y.
ε
ε
Proof Assume that x ⊥B y. The assertion is trivial for x = 0 or y = 0 so we assume
that x, y ∈ X \ {0}. If |λ| ≤ x
y , then −ελy ≥ −εx and from (14) we have
This part of the paper is a survey on known results, with the aim to give
at least an impression of various places where the notions of BJ-orthogonality
and its approximate counterpart can be considered or applied. Actually, we will
concentrate ourselves here on approximate orthogonality. The role of the (exact)
BJ-orthogonality in studies on the geometry of Banach spaces is well described,
e.g., in [42].
This section is organized as follows. At first, we discuss the natural extension
from general normed spaces to the space of linear bounded operators. Then we
consider an issue belonging to linear preserver problems, namely preservation or
approximate preservation (by linear operators) of the BJ-orthogonality relation. In
the third subsection we are pointing out some recent studies on the approximate
symmetry of the BJ-orthogonality, which is a topic very much connected to
approximate orthogonality as well as to the geometry of the considered space.
Approximate Birkhoff-James Orthogonality 313
MT = {x ∈ SH : T x = T },
where SH stands for the unit sphere in H. We begin with presenting a canonical
result given by Bhatia and Šemrl [8] (and independently by Paul [41]).
Theorem 3.1 ([8, Theorem 1.1, Remark 3.1]) Let H be a Hilbert space and let
T , A ∈ B(H). Then, the following conditions are equivalent:
(1) T ⊥B A;
(2) ∃ (xn )∞
n=1 ⊂ SH : limn→∞ T xn = T , limn→∞ T xn |Axn = 0.
Moreover, if dim H < ∞, then each of the above conditions is equivalent to:
(3) ∃ x0 ∈ MT : T x0 ⊥Ax0.
The above result was developed by various authors. Benítez, Fernández and
Soriano [7] showed that the equivalence (1)⇔(3) is valid if and only if H is a Hilbert
space (cannot be replaced by a Banach space). Generalizations of Theorem 3.1 have
been obtained, e.g., by Arambašić and Rajić [4], Sain, Paul and Hait [45, 46], Grover
[27] and Wójcik [51].
In [15] authors provided the first extension of the result of Bhatia and Šemrl to
approximate orthogonality in B(H).
Theorem 3.2 ([15, Theorem 3.2]) Let H be a real Hilbert space, let T , A ∈ B(H)
and let ε ∈ [0, 1). Then, the following conditions are equivalent:
(1) T ⊥εB A;
(2) ∃ (xn )∞
n=1 ⊂ SH :
Moreover, if dim H < ∞, then each of the above conditions is equivalent to:
(3) ∃ x0 ∈ MT : | T x0 |Ax0 | ≤ εT A.
If dim H < ∞ and, additionally, MT ⊂ MA , then each of the above three conditions
is equivalent also to:
(4) ∃ x0 ∈ MT : T x0 ⊥ε Ax0 .
314 J. Chmieliński
∃ x0 ∈ MT : T x0 ⊥ε Ax0 .
(b) There exist two sequences (xn ), (yn ) of unit vectors and two sequences of
positive real numbers (εn ), (δn ) such that
(i) εn → 0, δn → 0, T xn → T , T yn → T as n → ∞;
(ii) T xn +λAxn 2 ≥ (1−εn2 )T xn 2 −2ε 1 − εn2 T xn λA for all λ ≥ 0;
(iii) T yn +λAyn 2 ≥ (1−δn2 )T yn 2 −2ε 1 − δn2 T yn λA for all λ ≤ 0.
Under some additional conditions on the norm attainment set, further character-
izations of the approximate BJ-orthogonality of operators between normed linear
spaces were obtained in particular by Mal et al. [36]. For bilinear operators the topic
was studied recently by Khurana and Sain [33]. For a positive operator A ∈ B(H)
acting on a Hilbert space H, Sen et al. [48] defined A-orthogonality in H and A-
BJ-orthogonality in B(H). Next, they introduced the notions of (ε, A)-orthogonality
and (ε, A)-BJ-orthogonality in H, and finally (ε, A)-BJ-orthogonality in B(H). In
particular, some characterizations of those approximate orthogonalities were given.
Investigations concerning BJ-orthogonality in semi-Hilbertian spaces were carried
on also by Zamani [52].
Now, let us consider the space C(K) of all real continuous mappings defined
on a locally compact topological space K endowed with the supremum norm. A
subspace C0 (K) of C(K) defined by
has the property that for f ∈ C0 (K) the set Mf := {t ∈ K : |f (t)| = f } is always
nonempty and compact. In [15] a characterization of approximate BJ-orthogonality
on C0 (K) was given.
Theorem 3.8 ([15, Theorem 3.6]) Let f, g ∈ C0 (K), f = 0 = g. Assume that Mf
is connected. Then, the following conditions are equivalent:
(a) f ⊥εB g;
(b) ∃ t1 ∈ Mf : |g(t1 )| ≤ εg.
A linear mapping T between two inner product spaces H and K which preserves
orthogonality, i.e., such that
x⊥y #⇒ T x⊥T y, x, y ∈ H,
x⊥y #⇒ T x⊥εB T y, x, y ∈ X .
Then
x⊥B y #⇒ y⊥εB x.
In [19] there were given several conditions sufficient for approximate symmetry
of ⊥B . On the other hand, there were given examples of (classes of) normed
Approximate Birkhoff-James Orthogonality 317
3.4 Varia
Apart from mentioned above, there are other areas of research where the notion of
approximate BJ-orthogonality is involved. Without giving too much details we only
signal their presence.
Approximate Birkhoff-James Orthogonality in Hilbert Modules In an inner
product module over a C ∗ -algebra one can define an orthogonality relation by
using both the inner product as well as the corresponding norm. This gives rise
to various types of orthogonality (see Arambašić and Rajić [4, 5]). Approximate
318 J. Chmieliński
BJ-orthogonality in such realm has been also considered by Moslehian and Zamani
[39].
Orthogonality Sets Two notions of approximate Birkhoff-James orthogonality sets
have been introduced by Sain et al. [47]. Given x ∈ X and ε ∈ [0, 1) one can
consider
A geometrical description of these two sets in an arbitrary normed space was given:
each of them is a union of two-dimensional normal cones.
A similar notion of Birkhoff–James ε-orthogonality sets for matrices and for
matrix polynomials, based on the Dragomir’s definition of the approximate BJ-
orthogonality, were defined and studied in [20] (see also [21, 32, 40]).
The list of topics and results connected with the approximate BJ-orthogonality
which we dealt with in this section is by no means complete. It was our purpose just
to show that the considered concepts may be applied in a variety of further studies.
References
1. J. Alonso, C. Benitez, Orthogonality in normed linear spaces: a survey. Part I: main properties.
Extracta Math. 3(1), 1–15 (1988). Part II: Relations between main orthogonalities. Extracta
Math. 4(3), 121–131 (1989)
2. J. Alonso, H. Martini, S. Wu, On Birkhoff orthogonality and isosceles orthogonality in normed
linear spaces. Aequationes Math. 83(1–2), 153–189 (2012)
3. C. Alsina, J. Sikorska, M. Santos Tomás, Norm Derivatives and Characterizations of Inner
Product Spaces (World Scientific, Hackensack, 2010)
4. Lj. Arambašić, R. Rajić, The Birkhoff-James orthogonality in Hilbert C ∗ -modules. Linear
Algebra Appl. 437(7), 1913–1929 (2012)
5. Lj. Arambašić, R. Rajić, On three concepts of orthogonality in Hilbert C ∗ -modules. Linear
Multilinear Algebra 63(7), 1485–1500 (2015)
6. G. Birkhoff, Orthogonality in linear metric spaces. Duke Math. J. 1(2), 169–172 (1935)
7. C. Benítez, M. Fernández, M.L. Soriano, Orthogonality of matrices. Linear Algebra Appl.
422(1), 155–163 (2007)
8. R. Bhatia, P. Šemrl, Orthogonality of matrices and some distance problems. Linear Algebra
Appl. 287(1–3), 77–85 (1999)
9. A. Blanco, A. Turnšek, On maps that preserve orthogonality in normed spaces. Proc. R. Soc.
Edinburgh Sect. A 136(4), 709–716 (2006)
10. J. Chmieliński, On an ε-Birkhoff orthogonality. J. Inequal. Pure Appl. Math. 6(3) (2005). Art.
79
11. J. Chmieliński, Linear mappings approximately preserving orthogonality. J. Math. Anal. Appl.
304(1), 158–169 (2005)
12. J. Chmieliński, Stability of the orthogonality preserving property in finite-dimensional inner
product spaces. J. Math. Anal. Appl. 318(2), 433–443 (2006)
13. J. Chmieliński, Orthogonality Preserving Property and Its Ulam Stability. Functional Equa-
tions in Mathematical Analysis, Springer Optimization and Its Applications, vol. 52 (Springer,
New York, 2012), pp. 33–58
Approximate Birkhoff-James Orthogonality 319
42. K. Paul, D. Sain Birkhoff-James Orthogonality and Its Application in the Study of Geometry
of Banach Space. Advanced Topics in Mathematical Analysis (CRC Press, Boca Raton, FL,
2019), pp. 245–284
43. K. Paul, D. Sain, A. Mal, Approximate Birkhoff-James orthogonality in the space of bounded
linear operators. Linear Algebra Appl. 537, 348–357 (2018)
44. J. Rätz, On orthogonally additive mappings. Aequationes Math. 28(1–2), 35–49 (1985)
45. D. Sain, K. Paul, Operator norm attainment and inner product spaces. Linear Algebra Appl.
439(8), 2448–2452 (2013)
46. D. Sain, K. Paul, S. Hait, Operator norm attainment and Birkhoff-James orthogonality. Linear
Algebra Appl. 476, 85–97 (2015)
47. D. Sain, K. Paul, A. Mal, On approximate Birkhoff-James orthogonality and normal cones in
a normed space. J. Convex Anal. 26(1), 341–351 (2019)
48. J. Sen, D. Sain, K. Paul, On approximate orthogonality and symmetry of operators in semi-
Hilbertian structure. Bull. Sci. Math. 170 (2021). Paper No. 102997
49. A. Turnšek, On mappings approximately preserving orthogonality. J. Math. Anal. Appl. 336(1),
625–631 (2007)
50. P. Wójcik, Characterization of smooth spaces by approximate orthogonalities. Aequationes
Math. 89(4), 1189–1194 (2015)
51. P. Wójcik, Orthogonality of compact operators. Expo. Math. 35(1), 86–94 (2017)
52. A. Zamani, Birkhoff-James orthogonality of operators in semi-Hilbertian spaces and its
applications. Ann. Funct. Anal. 10(3), 433–445 (2019)
Orthogonally Additive Operators
on Vector Lattices
1 Introduction
The work of famous mathematicians Drewnowski et al. [15, 16, 32, 33] has led to the
appearance and study of orthogonally additive operators (OAOs) on vector lattices.
OAOs generalize linear ones (see for a definition below) and naturally appear in
different areas of modern mathematics, e.g. partial differential equations, convex
geometry, dynamical systems and stochastic processes [31, 50, 57]. It is worth
mentioning that some classical operators of nonlinear analysis including Uryson,
Hammerstein and Nemyskii operators are orthogonally additive in appropriate
M. Pliev ()
Southern Mathematical Institute of the Russian Academy of Sciences, Vladikavkaz, Russia
North Caucasus Center for Mathematical Research of the Vladikavkaz Scientific Center of the
Russian Academy of Sciences, Vladikavkaz, Russia
e-mail: [email protected]
M. Popov
Institute of Exact and Technical Sciences, Pomeranian University in Słupsk, Słupsk, Poland
Vasyl Stefanyk Precarpathian National University, Ivano-Frankivsk, Ukraine
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 321
R. M. Aron et al. (eds.), Operator and Norm Inequalities and Related Topics,
Trends in Mathematics, https://doi.org/10.1007/978-3-031-02104-6_10
322 M. Pliev and M. Popov
function spaces [35]. The theory of OAOs is being developed by different authors
in many directions [1, 4, 5, 19–21, 47, 49].
The aim of this paper is to discuss some known results and state open problems
which can be useful for further development of the theory. Let us describe the
content of this article. In the next section, we briefly present necessary background
on vector lattices and orthogonally additive operators. In Sect. 3, we discuss the
lateral partial order on vector lattices, which is useful for the study of OAOs.
Section 4 presents several extension theorems. In Sect. 5 we provide results on the
order structure of different classes of OAOs, including elegant formulas of Riesz-
Kantorovich type for the lattice operations over OAOs. In Sect. 6, we discuss on
C-compact and AM-compact OAOs. In particular, we show that, under some mild
conditions, the set of all C-compact OAOs is a projection band in OAr (E, F ) and
present a solution to the domination problem for AM-compact abstract Uryson
operators. Section 7 is devoted to the relationships between different partial order
continuities, which are clear for linear operators, however become involved for
OAOs. In Sect. 8 we present some theorems on narrow OAOs, including a deep
result on the representation of regular operators as the sum of a pseudo-embedding
and a diffuse (= narrow) operator, both for linear and orthogonally additive settings.
In Sect. 9, we endow the vector lattice of order bounded OAOs with a norm, such
that the set of all OAOs having finite norm becomes a Banach lattice in which the
subspace of all linear bounded operators is contractive complemented by means of
plenty projections called linear sections of OAOs. Final section contains some open
problems. Several results are provided with proofs, which are not new.
The standard reference books on the theory of vector and Banach lattices are
[8, 37]. All vector
<n lattices we consider below are supposed to be Archimedean. We
n
write x = i=1 x i to express that x = i=1 xi and xi ⊥xj for all i = j . In
particular, for n = 2 we use the notation x = x1 x2 . We say that y is a fragment
(a component) of x ∈ E and use the notation y ? x, if y ⊥ (x − y). The set of
all fragments of x ∈ E is denoted by Fx . We say that x1 , x2 ∈ Fx are mutually
complemented if x = x1 x2 . It is a standard exercise to show that ? is a partial
order on E, called the lateral order (see [38] for a detailed study of this order).
Definition 2.1 Let E be a vector lattice and X a real vector space. A function
T : E → X is called an orthogonally additive operator (OAO in short) provided
T (x + y) = T (x) + T (y) for any disjoint elements x, y ∈ E.
It is not hard to check that T (0) = 0. The set of all OAOs from E to X is a real
vector space with respect to the natural linear operations.
Orthogonally Additive Operators on Vector Lattices 323
Since for every x ∈ R the set Fx contains only two elements {0, x}, one has T ∈
P(E, F ). On the other hand, for the order bounded subset (−1, 1) ⊂ R, the subset
T (−1, 1) is order unbounded and hence T ∈ / U(E, F ).
Proposition 2.3 Let E, F be vector lattices. Then OA+ (E, F ) ⊆ P(E, F ).
Proof Given any T ∈ OA+ (E, F ), x ∈ E and y ∈ Fx , one has
T (x) = T y (x − y) = T (y) + T (x − y) ≥ T y ≥ 0,
The given partial order ≤ on a vector lattice E induces another partial order ? on
E, which was formally introduced and studied in [38].
Definition 3.1 Let E be a vector lattice. The partial order ? on E we call the lateral
order on E. A subset G ⊆ E is said to be laterally bounded in E if G ⊆ Fx for some
x ∈ E. We do not mention here “from above” because every subset is automatically
laterally bounded from below by zero.
The lateral supremum and infimum
withrespect to the lateral order ? on E are
denoted using the bold symbols , ∪ and , ∩ respectively.
Proposition 3.2 ([38, 52]) Let E be a vector lattice and e ∈ E. Then the following
assertions hold.
1. The set Fe of all fragments of e is a Boolean algebra with zero 0, unit e with
respect to the operations ∪ and ∩. Moreover, x ∪ y = (x+ ∨ y+ ) − (x− ∨ y− )
and x ∩ y = (x+ ∧ y+ ) − (x− ∧ y− ) for all x, y ∈ Fe .
2. Assume e ≥ 0. Then the following holds.
(a) The lateral order ? on Fe coincides with the lattice order ≤.
(b) Let a nonempty subset A of Fe have a lateral supremum a = A
(respectively, a lateral infimum a = A).
(i) If y = sup A (respectively, y = inf A) exists in E then y = a.
(ii) If, moreover, E has the principal projection property then sup A (respec-
tively, inf A) exists in E and by (i) equals a.
Remark that there
exist a vector lattice E, an element e ∈ E+ and subsets A and
B of Fe such that A and B exist, while sup A and inf B do not exist in E [52,
Example 1.2].
A vector lattice E is said to be
• C-completeif every nonempty laterally bounded subset G of E has a lateral
supremum G ∈ E;
• laterally complete if every disjoint family from E+ has a supremum.
If a vector lattice E is either Dedekind complete or laterally complete then E is
C-complete [38, Corollary 5.8]. The Banach lattice C[0, 1] is a C-complete vector
lattice which is neither Dedekind complete, nor laterally complete.
The following statement is easy to prove (cf. [8, Theorem 1.49], [18]).
Proposition 3.3 Let E be a C-complete vector lattice. Then for every x ∈ E the
Boolean algebra Fx is Dedekind complete.
The lateral order is of great importance for the study of OAOs.
326 M. Pliev and M. Popov
By (1) of Proposition 3.2, every finite laterally bounded subset of E has a lateral
supremum and lateral infimum. However, there is a vector lattice E and a two-point
subset {x, y} of E which (being laterally bounded from below by 0) has no lateral
infimum [38, Example 3.11]. A vector lattice E is said to have the intersection
property if every two-point subset {x, y} of E has a lateral
infimum x ∩ y. Remark
that if x ∩ y exists for some x, y ∈ E then x ∩ y = (Fx ∩ Fy ) is the ?-maximal
common fragment of x and y [52, Proposition 1.12]. The intersection property is a
lateral analogue of the principal projection property, see Proposition 4.9.
The following result describes the relationships between the intersection property
and some other known properties of vector lattices.
Theorem 3.14 Let E be a vector lattice.
1. If E has the principal projection property then E possesses the intersection
property. Moreover,(∀x, y ∈ E) Fx∩y = Fx ∩ Fy .
2. The C-completeness of E implies the intersection property of E.
3. The vector lattice C[0, 1] is C-complete and does not have the principal
projection property. As a consequence, the intersection property does not imply
the principal projection property.
4. There exists a vector lattice with the principal projection property which is not
C-complete. As a consequence, the intersection property does not imply the C-
completeness.
Item (1) follows from [38, Theorem 3.13]. (2) is a part of [52, Proposition 1.12].
(3) The C-completeness of C[0,1] is proved in [43, Proposition 4.2], and the fact
that C[0, 1] fails the principal projection property is well known and easily seen. (4)
A corresponding example is provided in [39, Proposition 2.5].
Definition 4.1 Let E be a vector lattice and I a lateral ideal of E. We say that a
subset D of I is absolutely order bounded in I, provided
is a positive OAO from E to F , that is, T1I ∈ OA+ (E, F ). Moreover, T1I x = T x for
all x ∈ I.
Now we present a refinement of [24, Theorem 3]. Given a vector lattice E, an
OAO T : E → E is said to be laterally non-expanding, if T (x) ? x for all x ∈ E.
Obviously, every laterally non-expanding OAO preserves disjointness. A laterally
non-expanding projection (that is, T 2 = T ) is called a lateral retraction. A subset A
of E is called a lateral retract if A is the image of some lateral retraction T : E →
E, that is, T (E) = A. A lateral band A of E, which is a lateral retract, is called a
projection lateral band, and the lateral retraction of E onto A is called the lateral
band projection of E onto A.
Theorem 4.4 ([27, Theorem 2.6]) Let E be a vector lattice.
1. For each lateral retract A in E there is a unique lateral retraction of E onto A.
2. Every lateral retraction is horizontally continuous.
3. Every lateral retract in E is a lateral band.
The following theorem generalizes Theorem 3 of [24] and asserts, in particular,
that every lateral band in a C-complete vector lattice is a lateral retract, and hence,
a projection lateral band.
Theorem 4.5 Every lateral band B of a C-complete vector lattice E is a lateral
retract, and the function pB : E → E defined by setting for every x ∈ E
pB (x) = (Fx ∩ B) = x B (2)
Hence, pB is an OAO.
The following theorem provides a partial case of formula (2) for principal
lateral bands, however proved under a less restrictive assumption on E to have the
intersection property.
Theorem 4.6 ([52, Theorem 1.6]) Let E be a vector lattice with the intersection
property. Then for every e ∈ E the function Qe : E → E defined by setting
Qe x = x ∩ e for all x ∈ E
Two fundamental results in this direction were obtained by Mazón and Segura de
León in 1990.
Theorem 5.1 ([35, Theorem 3.2]) Let E and F be vector lattices with F Dedekind
complete. Then U(E, F ) is a Dedekind complete vector lattice. Moreover, for each
S, T ∈ U(E, F ) and x ∈ E the following conditions hold:
1. (T ∨ S)(x) = sup{T (y) + S(z) : x = y z}.
2. (T ∧ S)(x) = inf{T (y) + S(z) : x = y z}.
3. T+ (x) = sup{T y : y ? x}.
4. T− (x) = − inf{T y : y ? x}.
5. |T (x)| ≤ |T |(x).
The second one represents the lattice operations on U(E, F ) in terms of directed
systems.
Theorem 5.2 ([36, Lemma 3.2]) Let E and F be vector lattices with F Dedekind
complete. Then for all T , S ∈ U(E, F ) and x ∈ E one has
4 <n 5
n
1. i=1 T (y i ) ∧ S(y i ) : x = y i ; n ∈ N ↓ (S ∧ T )(x).
i=1
4 <n 5
n
2. i=1 T (yi ) ∨ S(yi ) : x = yi ; n ∈ N ↑ (S ∨ T )(x).
i=1
4 <
n 5
n
3. i=1 |T (y i )| : x = y i ; n ∈ N ↑ |T |(x).
i=1
Similar results, obtained by the first named author and Ramdane in 2018, concern
a more wide class P(E, F ) of C-bounded OAOs.
Theorem 5.3 ([47, Theorem 3.6]) Let E and F be vector lattices with F
Dedekind complete. Then P(E, F ) is a Dedekind complete vector lattice. Moreover,
P(E, F ) = OAr (E, F ) and for all S, T ∈ P(E, F ) and x ∈ E conditions 1–5
from Theorem 5.1 hold.
The following proposition strengthens the inclusion U(E, F ) ⊂ P(E, F ).
Proposition 5.4 ([47, Proposition 3.7]) Let E, F be vector lattices with F
Dedekind complete. Then U(E, F ) is an order ideal of P(E, F ).
The next example shows that U(E, F ) need not be a band in P(E, F ).
332 M. Pliev and M. Popov
FT = {S ∈ U+ (E, F ) : S ∧ (T − S) = 0}.
Orthogonally Additive Operators on Vector Lattices 333
Proposition 5.7 ([11, Lemma 3.6]) Let E, F be vector lattices with F Dedekind
complete, ρ ∈ B(F ), T ∈ U+ (E, F ) and D be a lateral ideal. Then π D T is a
positive abstract Uryson operator and ρπ D T ∈ FT .
If D = Fx then the operator D x
n π Tx is denoted by π T . Let F be a vector lattice.
Any fragment of the form i=1 ρi π i T , n ∈ N, where ρ1 , . . . , ρn is a finite family
of mutually disjoint order projections of F , is called an elementary fragment of T .
The set of all elementary fragments of T we denote by AT . The following theorem
describes the structure of FT for a positive abstract Uryson operator T .
Theorem 5.8 ([42, Theorem 3.12]) Let E, F be vector lattices, F Dedekind
↑↓↑
complete and T ∈ U+ (E, F ). Then FT = AT .
Remark that, for linear positive operators a similar theorem and its modifications
were proved by de Pagter, Aliprantis and Burkinshaw, Kusraev and Strizhevski, see
[7, 13, 29].
In this subsection, following [43] we show that the set of all C-compact regular
OAOs from a vector lattice E to a Banach lattice F with an order continuous norm
is a band in the vector lattice of all OAOs from E to F .
334 M. Pliev and M. Popov
which are valid in OAr (E, F ) we obtain that COAr (E, F ) is a sublattice of
OAr (E, F ).
(c) Now we show that, if 0 ≤ Tλ ↑ T in OAr (E, F ) and any Tλ ∈ COAr (E, F )
then T ∈ COAr (E, F ). Indeed, take x ∈ E and ε > 0. Since the Banach lattice
F is order continuous, it follows from Tλ x ↑ T x that T x − Tλ0 x < 4ε for
some λ0 . We claim that, moreover, T y − Tλ0 y < ε4 for any y ∈ Fx . Indeed,
consider x = y z for some z ∈ E. Then
0 ≤ T y − Tλ 0 y ≤ T y − Tλ 0 y + T z − Tλ 0 z = T x − Tλ 0 x
implies T y − Tλ0 y ≤ T x − Tλ0 x. Since Tλ0 ∈ COAr (E, F ), there exists
a finite subset D of Fx with the property that for any y ∈ Fx there exists u ∈ D
satisfying
ε
Tλ0 u − Tλ0 y < .
2
So we obtain T u − T y ≤ T u − Tλ0 u + T y − Tλ0 y + Tλ0 u − Tλ0 y < ε,
which establishes the relative compactness of T (Fx ) in F .
(d) Finally we prove that COAr (E, F ) is an order ideal in OAr (E, F ). Let 0 ≤
R ≤ T , where R ∈ OAr (E, F ) and T ∈ COAr (E, F ). Then R ∈ IT and
by the Freudenthal’s spectral theorem [8, Theorem 2.8], there exists a sequence
(Sn )n∈N in OAr (E, F ) of T -step-functions with 0 ≤ Sn ↑ R. Taking into
account that1 Sn ∈ COAr (E, F ) for all n ∈ N and, what has been established
in c), we deduce that R ∈ COAr (E, F ). So, COAr (E, F ) is a band in
OAr (E, F ).
(e) Due to the Dedekind completeness of OAr (E, F ), it is a projection band.
1 This follows from the fact that together with T each fragment Ti of T belongs to COAr (E, F ).
336 M. Pliev and M. Popov
Throughout this section, let E, F be vector lattices with F Dedekind complete. Here
we discuss the relationships between the order continuity of an abstract Uryson
operator T : E → F and its modulus |T |, as well as some partial order continuities,
like horizontal-to-order and uniformly-to-order ones. Similar questions for linear
operators are simpler, see the next proposition.
Orthogonally Additive Operators on Vector Lattices 337
Proposition 7.1 ([35, Proposition 3.9]) Let E be a vector lattice with the principal
projection property, F a Dedekind complete vector lattice and S : E → F a regular
linear operator. Then the following assertions hold:
1. if S is horizontally-to-order continuous then S is order continuous;
2. if S is horizontally-to-order σ -continuous then S is order σ -continuous.
This is not longer true for OAOs due to the following example.
Example Let 0 ≤ p ≤ ∞. There exists a horizontally-to-order continuous
orthogonally additive functional f : Lp → R which is not order continuous.
Moreover, f is not uniformly-to-order continuous.
Define a function ϕ : R → [−1, 1] by setting ϕ(t) = t as |t| ≤ 1 and ϕ(t) = 0
for |t| > 1. Then define f : Lp → R by setting
$
f (x) = ϕ x(t) dμ(t).
[0,1]
Detailed proof that f is as desired the reader can find in [22, Example 2.1].
We say that a function f : E → F is uniformly-to-order continuous if for every
o
net (xα ) in E and every x ∈ E the condition xα ⇒ x implies f (xα ) −→ f (x).
Assume T ∈ U(E, F ). Then the function T B : E → F defined by setting
is a positive abstract Uryson operator (that is, TB ∈ U(E, F )+ ) [35, Proposition 3.4].
Following [52], we say that T B is the envelope of T . Remark that the envelope has
the following properties (see propositions 3.4 and 3.5 of [52] for details).
Proposition 7.2 Let E, F be vector lattices with F Dedekind complete and S, T ∈
U(E, F ). Then
1. T (x) ≤ TB(x) for all x ∈ E;
2. if x ≤ y for x, y ∈ E then TB(x) ≤ TB(y);
3. if 0 ≤ S ≤ T then BS(x) ≤ T B(x) for all x ∈ E;
4. S + T (x) ≤ BS(x) + TB(x) for all x ∈ E;
5. if, moreover, E has the principal projection property then B
T = TB.
Theorem 7.3 ([22, Theorem 2.2]) Let E be a vector lattice with the principal
projection property, F a Dedekind complete vector lattice and T ∈ U(E, F ).
Consider the following statements.
(i) T is order continuous.
(ii) |T | is order continuous.
(iii) The envelope T B of T is horizontally-to-order continuous.
Then
⎧
⎨ 0 for − ∞ < x ≤ 0,
|f |(x) = x for 0 < x ≤ 1,
⎩
1 for 1 < x < +∞.
Narrow linear operators were introduced and studied in 1990 by Plichko and the
second named author in [41] as a generalization of compact operators defined on
symmetric function spaces. But actually these operators were investigated by dif-
ferent mathematicians earlier. In 2009 narrow operators were naturally generalized
to linear operators defined on vector lattices in [34] (see also [53] and references
therein). After the monograph [53] was published, the notion was (not less naturally)
generalized to OAOs in paper by the authors [44], and then developed in some other
papers (see e.g., [21]). A recent paper by the authors [46] contains a representation
theorem for regular operators (two versions for both linear and orthogonally additive
settings) which generalizes a number of known results in this direction.
Definition 8.1 An OAO T : E → F between vector lattices E, F is said to be
order narrow if for every e ∈ E there is a net of decompositions e = eα eα such
Orthogonally Additive Operators on Vector Lattices 339
that the net T (eα ) − T (eα ) α order converges to zero. An OAO T : E → G from a
vector lattice E to a Banach space G is called narrow if for every e ∈ E and every
ε > 0 there is a decomposition e = e e such that T (e ) − T (e ) < ε. An OAO
T : E → V from a vector lattice E to a linear space V is called strictly narrow if
for every e ∈ E there is a decomposition e = e e such that T (e ) = T (e ).
Obviously, every strictly narrow operator is both narrow and order narrow
for suitable range lattices. Let us briefly demonstrate of why is every “small”
operator strictly narrow. Let E be a Banach function space on [0, 1]. Then every
horizontally-to-norm continuous OAO f : E → R is strictly narrow. Indeed,
given any e ∈ E, the function ϕ : [0, 1] → R defined by ϕ(t) = f (e · 1[0,t ] ) is
continuous, ϕ(0) = 0 and ϕ(1) = f (e). Choosing s ∈ [0, 1] so that ϕ(s) = f (e)/2,
we obtain e = e · 1[0,s] e · e · 1(s,1] and
f (e)
f e · 1[0,s] = = f e · 1(s,1] .
2
Theorem 8.2 ([53, Proposition 2.1]) If a Banach function space E on a finite
measure space (, , μ) has an absolutely continuous norm on the unit (i.e.,
limμ(A)→0 1A = 0) and X is a Banach space then every compact and every
AM-compact linear operator T : E → X is narrow.
A strictly narrow operator need not have “small” range and can be “very non-
compact”: under the same assumptions on T = (, , μ), for every rearrangement
invariant Banach space E on T there exists a strictly narrow linear projection of
E onto a subspace E0 which is isometrically isomorphic to E [53, Theorem 4.17].
The assumption on E to have an absolutely continuous norm on the unit is essential:
there are nonnarrow continuous linear functionals on L∞ [53, Example 10.12].
The following theorem extends Theorem 8.2 to OAOs.
Theorem 8.3 ([44, Theorem 3.2]) Let E be an atomless Dedekind complete vector
lattice and X a Banach space. Then every orthogonally additive horizontally-to-
norm continuous C-compact operator T : E → X is narrow.
operators on vector lattices [34] and the authors’ representation of order bounded
OAOs [44]. The main contribution of the result under presentation is ridding of
the order continuity assumption on an operator and extending to OAOs. Earliest
of the above mentioned results are formulated using the language of measures
and integral operators on suitable function spaces, and the later ones used lattice
terminology, which is shorter and allows obtaining more general results. The general
idea of all representation theorems is to split the vector lattice X(E, F ) of all
operators T : E → F from a certain class into a direct sum of orthogonal bands
X(E, F ) = Y (E, F ) ⊕ Z(E, F ), where operators from Y (E, F ) are in some sense
atomic and operators from Z(E, F ) are kind of continuous, which then yields a
unique desirable representation of every operator T = Ta + Tc .
Any representation theorem for linear operators can be stated as follows: let
H(E, F ) be the band generated by all lattice homomorphisms (in other words, by all
disjointness preserving operators) from E to F , and D(E, F ) = H(E, F )d be the
disjoint complement of H(E, F ) in Lb (E, F ). Then, by the Dedekind completeness
of F , we obtain the following decomposition of Lb (E, F ) into orthogonal bands
The next two combined theorems by Enflo, Kalton, Rosenthal and Starbird
characterize summands of (6) (cf. theorems 7.38 and 7.39 in [53]).
Theorem 8.4 ([54, Theorem 3.2]) For every operator T ∈ L(L1 ) the following
assertions are equivalent:
1. T ∈ H(L1 );
2. T equals a pointwise absolutely convergent2 series T = ∞n=1 Tn of disjointness
preserving operators Tn ∈ L(L1 ).
Moreover, every nonzero operator T ∈ H(L1 ) possesses the following property: for
every ε > 0 there exists a measurable subset A of [0, 1] such that the restriction
T L (A) is an into isomorphism with
1
(i) T L (A) ≥ T − ε;
1
−1
(ii) T · T
L1 (A)
< 1 + ε.
L1 (A)
Before stating of the second theorem, we provide with a definition of the Enflo-
Starbird function for OAOs, which is the same as for linear operators.
Definition 8.5 Let E, F be vector lattices with F Dedekind complete and T ∈
OAr (E, F ). We define a function λT : E+ → F+ , called the Enflo-Starbird
function of T , by setting for all x ∈ E+
G
F
m
λT (x) = inf sup |T |xi : x = x i , x i ∈ E+ , m ∈ N .
1≤i≤m i=1
Theorem 8.6 ([26, 54]) For every T ∈ L(L1 ) the following assertions are
equivalent:
1. T ∈ D(L1 );
2. T is narrow;
3. the Enflo-Starbird function λT of T equals zero;
4. for every measurable subset A of [0, 1] the restriction T L1 (A)
is not an into
isomorphism.
Implication (4) ⇒ (3) and the technique of λ-function is due to Enflo and Starbird
[17]; equivalences (1) ⇔ (3) ⇔ (4) can be deduced from Kalton’s results [26], and
the final equivalence (1) ⇔ (2), as well as noticing of the entire theorem the reader
can find in Rosenthal’s paper [54].
In view of Theorem 8.4, elements of H(E, F ) we call pseudo-embeddings, and
following terminology of Weis [58] and Huijsmans-de Pagter [25], elements of
D(E, F ) we call diffuse operators. Using this terminology, we provide below one
of the main results of [34] generalizing Theorem 8.6 to vector lattices (see also [53,
Theorem 10.40]).
Theorem 8.7 ([34]) Let E, F be Dedekind complete vector lattices such that E
is atomless and F is an ideal of some order continuous Banach lattice. Then for
every regular order continuous operator T : E → F the following assertions are
equivalent:
1. T ∈ D(E, F );
2. T is order narrow;
3. the Enflo-Starbird λ-function of T is zero: λT = 0.
Hence, each regular order continuous operator T : E → F is uniquely
represented in the form T = Ta + Tc where Ta is a sum of an order absolutely
342 M. Pliev and M. Popov
3 Weis’ representation theorem [58] was proved under the assumption of order continuity.
Orthogonally Additive Operators on Vector Lattices 343
T ∈ L+ + −
b (E, F ) and all x ∈ E, and by ϕ(T ) = ϕ(T ) − ϕ(T ) for an arbitrary
+ −
T ∈ Lb (E, F ), that is, ϕ(T ) x = T |x| − T |x| for all x ∈ E. If E has the
principal projection property then ϕ is a lattice monomorphism, that is, an injective
lattice homomorphism.
which we call the absolute norm. The following statement, in particular, asserts that
it is a norm.
Theorem 9.2 ([52, Theorem 3.2]) Let E be a normed lattice and F a Dedekind
complete Banach lattice. Then AB(E, F ) is a normed lattice with respect to the
absolute norm, and is a sublattice of U(E, F ).
The following example (with non-obvious proof) shows that the normed lattice
AB(E, F ) need not be norm complete.
Example ([52, Example 3.3]) Let E = F = L1 [0, 1]. Then the normed space
AB(E, F ) is not norm complete.
344 M. Pliev and M. Popov
which we call the uniform norm. Obviously, UB(E, F ) ⊆ AB(E, F ) and T abs ≤
T u = TBabs for every T ∈ UB(E, F ).
There inclusion UB(E, F ) ⊂ AB(E, F ) is strict for the class of AL-spaces [52,
Example 3.8].
Theorem 9.4 ([52, Theorem 3.9]) Let E be a normed lattice and F a Dedekind
complete Banach lattice. Then UB(E, F ) is a Dedekind complete Banach lattice
with respect to the uniform norm, and is a sublattice of U(E, F ).
L ⊕ L = {x + y : x ∈ L , y ∈ L }
is a level as well.
A level L in a vector lattice E is said to be positive (respectively, negative)
provided L ⊂ E+ (respectively, x ≤ 0 for each x ∈ L). The relation L ≥ 0
(respectively, L ≤ 0) means that the level L is positive (respectively, negative).
Proposition 9.7 ([52, Proposition 2.5]) Every level L in a vector lattice E admits
a unique decomposition into a direct sum of levels L = L+ ⊕ L− , where L+ ≥ 0
and −L− ≥ 0. In particular, for any principal level L = Fe one has (Fe )+ = F(e+ )
and (Fe )− = F(−e− ) .
The first theorem concerns the very general case and asserts the existence of an
extended linear section without any additional properties.
Theorem 9.9 ([52, Theorem 4.2]) Let E be a vector lattice, F a vector space and
T : E → F an OAO. Then for every level L of E there exists an extended linear
section S : E → F of T by L.
Definition 9.10 Let E, F be vector lattices and D ⊆ E. A function f : D → F is
said to be vertically order σ -continuous on D if D is an ideal of E and for every
w ∈ D+ , every x ∈ Ew and every increasing sequence (xn )∞ n=1 in Ew such that
o
0 ≤ x − xn ≤ n1 w one has f (xn ) −→ f (x).
One can easily show that every regular linear operator T : E → F is vertically
order σ -continuous on E, once F is Archimedean [52, Proposition 4.6].
Next is the main technical tool for the construction of linear sections.
Theorem 9.11 ([52, Theorem 4.7]) Let E, F be vector lattices. Assume E has the
principal projection property, F is Dedekind complete and T ∈ U(E, F ). Then for
every level L of E there is a unique regular linear section S = L (T ) : EL ⊕ Ld →
F of T by L. Moreover, if L ≥ 0 then S + = (L (T ))+ = L (T + ). In particular, if
T is positive as an OAO and L ≥ 0 then S is positive as a linear operator.
Definition 9.12 Let E be a vector lattice with the principal projection property, F
a Dedekind complete vector lattice, T ∈ U(E, F ) and L a level of E. The regular
linear section S = L (T ) : EL ⊕Ld → F of T by L, the existence and uniqueness
of which Theorem 9.11 asserts, is called the canonical linear section of T by L.
Theorem 9.11 yields the following properties of the canonical linear section
S = L (T ) of T by L:
1◦ S is a regular linear operator.
If, in addition, L ≥ 0 then
2◦ S + = (L (T ))+ = L (T + );
3◦ if T ≥ 0 as an OAO then S ≥ 0 as a linear operator.
The next property is expressed in the following theorem.
Theorem 9.13 ([52, Theorem 4.12]) Let E be a vector lattice with the principal
projection property, F a Dedekind complete vector lattice, T ∈ U(E, F ) and L a
positive level of E. If T is horizontally-to-order continuous (horizontally-to-order
σ -continuous) on L then the canonical linear section S = L (T ) of T by L is order
continuous (order σ -continuous) on its domain.
The following theorem asserts that the canonical linear section is a linear operator
from UB(E, F ) to Lr (EL ⊕ Ld , F ) with some useful properties.
Theorem 9.14 ([52, Theorem 4.15]) Let E, F be vector lattices. Assume E has the
principal projection property, F is Dedekind complete and L is a level in E. Then the
corresponding canonical linear section as a mapping L : UB(E, F ) → Lr (EL ⊕
Ld , F ) is a disjointness preserving linear operator. If, moreover, L ≥ 0 then L
Orthogonally Additive Operators on Vector Lattices 347
Observe that, under the above assumptions on E and F , the Banach space
L(E, F ) is a subspace of UB(E, F ).
Recall that a Banach lattice F is called a KB-space if every increasing norm
bounded sequence in F+ is norm convergent (equivalently, if the canonical image
of F in its second dual F is a band [8, Theorem 4.60]). By [8, Theorem 4.75], every
KB-space F satisfies (8) for any AL-space E. Nevertheless, there is a strictly wider
class of Banach lattices F than the KB-spaces (including e.g. infinite dimensional
L∞ (μ)-spaces which are not KB-spaces), possessing (8) for any AL-space E.
348 M. Pliev and M. Popov
10 Open Problems
Order projections in different spaces of OAOs were studied in [2, 3, 12, 48],
Problem 10.6 Obtain a formula for the order projection of OAr (E, F ) onto the
band generated by an arbitrary positive OAO T : E → F .
Problem 10.7 Obtain a formula for the order projection in OAr (E, F ) onto the
band by all disjointness preserving OAOs from in E to F .
References
4. N. Abasov, M. Pliev, On extensions of some nonlinear maps in vector lattices. J. Math. Anal.
Appl. 455(1), 516–527 (2017)
5. N. Abasov, M. Pliev, Dominated orthogonally additive operators in lattice-normed spaces. Adv.
Oper. Theory 4(3), 251–264 (2019)
6. Yu. Abramovich, Z. Chen, A. Wickstead, Regular-norm balls can be closed in the strong
operator topology. Positivity 1(1), 75–96 (1997)
7. C.D. Aliprantis, O. Burkinshaw, The components of a positive operator. Math. Z. 184(2), 245–
257 (1983)
8. C.D. Aliprantis, O. Burkinshaw, Positive Operators (Springer, Dordrecht, 2006)
9. J. Appell, P.P. Zabrejko, Nonlinear Superposition Operators (Cambridge University Press,
Cambridge, 1990)
10. J. Appell, J. Banas, N. Merentes, Bounded Variation and Around (De Gruyter, Berlin, 2014)
11. M.A. Ben Amor, M. Pliev, Laterally continuous part of an abstract Uryson operator. Int. J.
Math. Anal. 7(58), 2853–2860 (2013)
12. R. Chill, M. Pliev, Atomic operators in vector lattices. Mediter. J. Math. 17, article 138 (2020)
13. B. de Pagter, The components of a positive operator. Indag. Math. 48(2), 229–241 (1983)
14. P.G. Dodds, D.H. Fremlin, Compact Operators in Banach Lattices. Izrael J. Math. 34, 287–320
(1974)
15. L. Drewnowski, W. Orlicz, On orthogonally-additive functionals. Bull. Acad. Polon. Sci. Ser.
Sci. Math. Astron. Phys. 16, 883–888 (1968)
16. L. Drewnowski, W. Orlicz, Continuity and representation of orthogonally-additive functionals.
Bull. Acad. Polon. Sci. Ser. Sci. Math. Astron. Phys. 17, 647–653 (1969)
17. P. Enflo, T. Starbird, Subspaces of L1 containing L1 . Stud. Math. 65(2), 203–225 (1979)
18. N. Erkursun-Özcan, M. Pliev, On orthogonally additive operators in C-complete vector lattices.
Banach. J. Math. Anal. 16(1), article number 6 (2022)
19. W.A. Feldman, A characterization of non-linear maps satisfying orthogonality properties.
Positivity 21(1), 85–97 (2017)
20. W.A. Feldman, A factorization for orthogonally additive operators on Banach lattices. J. Math.
Anal. Appl. 472(1), 238–245 (2019)
21. O. Fotiy, A. Gumenchuk, I. Krasikova, M. Popov, On sums of narrow and compact operators.
Positivity 24(1), 69–80 (2020)
22. O. Fotiy, I. Krasikova, M. Pliev, M. Popov, Order continuity of orthogonally additive operators.
Results Math. 77, article 5 (2022)
23. A.I. Gumenchuk, Lateral continuity and orthogonally additive operators. Carpathian Math.
Publ. 7(1), 49–56 (2015)
24. A.I. Gumenchuk, M.A. Pliev, M.M. Popov, Extensions of orthogonally additive operators. Mat.
Stud. 42(2), 214–219 (2014)
25. C.B. Huijsmans, B. de Pagter, Disjointness preserving and diffuse operators. Compos. Math.
79(3), 351–374 (1991)
26. N.J. Kalton, The endomorphisms of Lp (0 ≤ p ≤ 1). Indiana Univ. Math. J. 27(3), 353–381
(1978)
27. A. Kamińska, I. Krasikova, M. Popov, Projection lateral bands and lateral retracts. Carpathian
Math. Publ. 12(2), 333–339 (2020)
28. M.A. Krasnosel’skii, P.P. Zabrejko, E.I. Pustil’nikov, P.E. Sobolevski, Integral Operators in
Spaces of Summable Functions (Noordhoff, Leiden, 1976)
29. A.G. Kusraev, M.A. Strizhevski, Lattice-normed spaces and dominated operators, Studies on
Geometry and Functional Analysis. Trudy Inst. Mat. (Novosibirsk), Novosibirsk 7, 132–158
(1987) (in Russian)
30. S. Kwapien, On the form of a linear operator in the space of all measurable functions. Bull.
Acad. Polon. Sci. Sér. Sci. Math. Astronom. Phys. 21, 951–954 (1973)
31. H. Le Dret, Nonlinear Elliptic Partial Differential Equations (Springer, Berlin, 2018)
32. M. Marcus, V. Mizel, Representation theorem for nonlinear disjointly additive functionals and
operators on Sobolev spaces. Trans. Am. Math. Soc. 226, 1–45 (1977)
Orthogonally Additive Operators on Vector Lattices 351
33. M. Marcus, V. Mizel, Extension theorem of Hahn-Banach type for nonlinear disjontly additive
functionals and operators in Lebesgue spaces. J. Funct. Anal. 24, 303–335 (1977)
34. O.V. Maslyuchenko, V.V. Mykhaylyuk, M.M. Popov, A lattice approach to narrow operators.
Positivity 13(3), 459–495 (2009)
35. J.M. Mazón, S. Segura de León, Order bounded ortogonally additive operators. Rev. Roumane
Math. Pures Appl. 35(4), 329–353 (1990)
36. J.M. Mazon, S. Segura de Leon, Uryson operators. Rev. Roumane Math. Pures Appl. 35(5),
431–449 (1990)
37. P. Meyer-Nieberg, Banach Lattices (Springer, Berlin, 1991)
38. V. Mykhaylyuk, M. Pliev, M. Popov, The lateral order on Riesz spaces and orthogonally
additive operators. Positivity 25(2), 291–327 (2021)
39. V. Mykhaylyuk, M. Pliev, M. Popov, The lateral order on Riesz spaces and orthogonally
additive operators. II. Preprint
40. V. Orlov, M. Pliev, D. Rode, Domination problem for AM-compact abstract Uryson operators.
Arch. Math. 107(5), 543–552 (2016)
41. A.M. Plichko, M.M. Popov, Symmetric function spaces on atomless probability spaces.
Dissertationes Math. (Rozprawy Mat.) 306, 1–85 (1990)
42. M. Pliev, Domination problem for narrow orthogonally additive operators. Positivity 21(1),
23–33 (2017)
43. M. Pliev, On C-compact orthogonally additive operators. J. Math. Anal. Appl. 494(1), 124594
(2021)
44. M. Pliev, M. Popov, Narrow orthogonally additive operators. Positivity 18(4), 641–667 (2014)
45. M. Pliev, M. Popov, Extension of abstract Urysohn operators. Sib. Math. J. 57(3), 552–557
(2016)
46. M. Pliev, M. Popov, Representation theorems for regular linear and orthogonally additive
operators. Preprint, 26 p. (2020)
47. M. Pliev, K. Ramdane, Order unbounded orthogonally additive operators in vector lattices.
Mediter. J. Math. 15(2), article 55 (2018)
48. M. Pliev, M. Weber, Disjointness and order projections in the vector lattices of abstract Uryson
operators. Positivity 20(3), 695–707 (2016)
49. M.A. Pliev, F. Polat, M.R. Weber, Narrow and C-compact orthogonally additive operators in
lattice-normed spaces. Results Math. 74, article 157 (2019)
50. A. Ponosov, E. Stepanov, Atomic operators, random dynamical systems and invariant mea-
sures. St. Petersburg Math. J. 26(4), 607–642 (2015)
51. M. Popov, Horizontal Egorov property of Riesz spaces. Proc. Am. Math. Soc. 149(1), 323–332
(2021)
52. M. Popov, Banach lattices of orthogonally additive operators. J. Math. Anal. Appl. 514(1),
126279 (2022)
53. M. Popov, B. Randrianantoanina, Narrow Operators on Function Spaces and Vector Lattices
(De Gruyter, Berlin, 2013)
54. H.P. Rosenthal, Embeddings of L1 in L1 . Contemp. Math. 26, 335–349 (1984)
55. S. Segura de León, Bukhvalov type characterizations of Urysohn operators. Stud. Math. 99(3),
199–220 (1991)
56. S. Segura de León, Characterizations of Hammerstein operators. Rend. Math. 7(12), 689–707
(1992)
57. P. Tradacete, I. Villanueva, Valuations on Banach lattices. Int. Math. Res. Not. 2020(1), 287–
319 (2020)
58. L. Weis, On the representation of order continuous operators by random measures. Trans. Am.
Math. Soc. 285(2), 535–563 (1984)
Part III
Inequalities Related to Types of Operators
Normal Operators and their
Generalizations
Pietro Aiena
1 Introduction
P. Aiena ()
Dipartimento d’Ingegneria, Università di Palermo (Italia), Palermo, Italy
e-mail: [email protected]
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 355
R. M. Aron et al. (eds.), Operator and Norm Inequalities and Related Topics,
Trends in Mathematics, https://doi.org/10.1007/978-3-031-02104-6_11
356 P. Aiena
Since this chapter concerns the spectral theory of bounded linear operators, we
always assume that the Banach spaces, or the Hilbert spaces, are complex infinite-
dimensional. If X, Y are Banach spaces, by L(X, Y ) we denote the Banach space of
all bounded linear operators from X into Y . Recall that if T ∈ L(X, Y ), the norm of
T is defined by
T x
T := sup .
x =0 x
f (x) := T x, y.
Hence
ker T := {x ∈ X : T x = 0},
while the range of T is denoted by T (X). In the sequel, for every bounded operator
T ∈ L(X, Y ), we shall denote by α(T ) the nullity of T , defined as α(T ) :=
dim ker T , while the deficiency β(T ) of T is the dimension of the cokernel of T (X),
i.e., β(T ) := dim Y/T (X) = codim T (X). The spectrum of T ∈ L(X) defined as
σ (T ) := {λ ∈ C : λI − T is not bijective}.
These two spectra are nonempty and dual one to each other, i.e., σa (T ∗ ) = σs (T )
and σs (T ∗ ) = σa (T ). Both spectra σap (T ) and σsu (T ) contain the boundary ∂σ (T )
of the spectrum, see [4, Theorem 1.12]. Furthermore, it is easily seen that for Hilbert
space operators for the adjoint T ∗ and the dual T we have:
X = T p (X) ⊕ ker T p .
with ak , bk ∈ X, and λ ∈ D(λ0 , δ) \ {λ0 }. The coefficients are given by the formulas
$ $
1 f (λ) 1
ak = dλ, and bk = f (λ)(λ − λ0 )k−1 dλ,
2πi (λ − λ0 )k+1 2πi
for all 0 < |λ − λ0 | < δ. The coefficients are calculated according the formulas
$
1 Rλ
Qk = dλ (3)
2πi (λ − λ0 )k+1
$
1
Pk = Rλ (λ − λ0 )k−1 dλ, (4)
2πi
is defined in a similar way as we have done for the elements of the Banach algebra
L(X).
Normal Operators and their Generalizations 361
In the particular case that of functions which are equal to 1 on certain parts of
σ (T ) and equal to 0 on others we get idempotent operators. To see this, suppose
that σ is a spectral set (i.e. σ and σ (T ) \ σ are both closed) and := 1 ∪ 2 is an
open covering of σ (T ) such that 1 ∩ 2 = ∅ and σ ⊆ 1 , define h(λ) := 1 for λ
on 1 and h(λ) := 0 for λ on 2 . Consider the operator Pσ := h(T ). It is easy to
check that Pσ2 = Pσ , so Pσ is a projection called the spectral projection associated
with σ , and obviously
$
1
Pσ = (λI − T )−1 dλ, (5)
2πi
where is a curve enclosing σ and which separates σ from the remaining part of
the spectrum.
Let us consider again the case of an isolated point λ0 of σ (T ). Then {λ0 } is a
spectral set, so we can consider the spectral projection P0 associated with {λ0 }. It is
easy to check that if Pk are defined according (4) then
P1 = P0 , Pk = (T − λ0 I )n−1 P0 (k = 1, 2, . . . ) (6)
Equation (6) show that either Pk = 0, or that there exists a natural p such that
Pk = 0 for k = 1, . . . , p but Pk = 0 for k > p. In the second case the isolated point
λ0 is pole of order p of T .
The spectral sets produce the following decomposition see [39, §49].
Theorem 2.8 If σ is a spectral set (possibly empty) of T ∈ L(X) then the projection
in (5) generates the decomposition X = Pσ (X) ⊕ ker Pσ . The subspaces Pσ (X)
and ker Pσ are invariant under every f (T ) with f ∈ H(σ (T )); the spectrum of the
restriction T |Pσ (X) is σ and the spectrum of T | ker Pσ is σ (T ) \ σ .
The proof of the following basic result may be found in [39, Proposition 50.2].
Theorem 2.9 Let T ∈ L(X). Then λ0 ∈ σ (T ) is a pole of R(λ, T ) if and only if
0 < p(λ0 I − T ) = q(λ0 I − T ) < ∞. Moreover, if p := p(λ0 I − T ) = q(λ0 I − T )
then p is the order of the pole, every pole λ0 ∈ σ (T ) is an eigenvalue of T , and if
P0 is the spectral projection associated with {λ0 } then
In the following result, due to Schmoeger [54], we show that for an isolated point
λ0 of σ (T ) the quasi-nilpotent part H0 (λ0 I −T ) and the analytical core K(λ0 I −T )
may be precisely described as a range or a kernel of a projection.
362 P. Aiena
Theorem 2.10 Let T ∈ L(X), where X is a Banach space, and suppose that λ0
is an isolated point of σ (T ). If P0 is the spectral projection associated with {λ0 },
then:
(i) P0 (X) = H0 (λ0 I − T );
(ii) ker P0 = K(λ0 I − T ). Consequently,
X = H0 (λ0 I − T ) ⊕ K(λ0 I − T ).
H0 (λ0 I − T ) = ker(λ0 I − T )p ,
and
Note that if T is of Kato type then also T is of Kato type. More precisely, the
pair (N ⊥ , M ⊥ ) is a GKD for T with T |N ⊥ semi-regular and T |M ⊥ nilpotent, see
Theorem 1.43 of [1].
For many reasons the most satisfactory generalization to the general Banach space
setting of the normal operators on a Hilbert space is the concept of decomposable
operator. In fact the class of these operators possesses a spectral theorem and a rich
lattice structure for which it is possible to develop what it is called a local spectral
theory, i.e. a local analysis of their spectra. Decomposability may be defined in
several ways, for instance by means of the concept of glocal spectral subspace.
Definition 3.1 For an arbitrary bounded linear operator on a Banach space T ∈
L(X) and a closed subset F of C, the glocal spectral subspace XT (F ) is defined as
the set of all x ∈ X such that there is an analytic X-valued function f : C \ F → X
for which
(λI − T )f (λ) = x
for all λ ∈ C \ F .
The quasi-nilpotent part may be described as a glocal subspace, indeed we have
If the function f is defined on the set ρT (x) then f is called a local resolvent
function of T at x. The set ρT (x) is called the local resolvent of T at x. The local
spectrum σT (x) of T at the point x ∈ X is defined to be the set
σT (x) := C \ ρT (x).
364 P. Aiena
Definition 3.3 For every subset F of C the local spectral subspace of an operator
T ∈ L(X) associated with F ⊆ C is the set
XT (F ) := {x ∈ X : σT (x) ⊆ F }.
Note that T has SVEP if and only if XT (F ) = XT (F ) for every closed subset
F ⊆ C, see [4, Theorem 2.23]. Obviously, if F1 ⊆ F2 ⊆ C then XT (F1 ) ⊆ XT (F2 )
and
XT (F ) = XT (F ∩ σ (T ).
K(T ) = XT (C \ {0}) = {x ∈ X : 0 ∈
/ σT (x)},
and dually,
From the equality (1) it is easily follows that for Hilbert space operators we have
The proof of the following theorem may be found in [1, Theorem 3.16, Theo-
rem 3.17].
Theorem 3.5 Let T ∈ L(X), X a Banach space and suppose that λI −T is of Kato
type . Then we have:
(i) T has SVEP at λ ⇔ p(λI − T ) < ∞.
(ii) T ∗ has SVEP at λ ⇔ q(λI − T ) < ∞.
In particular the equivalences (i) and (ii) hold for semi-Fredholm operators.
Recall that a bounded operator K ∈ L(X) is said to be algebraic if there exists a
non-constant polynomial h such that h(K) = 0. Trivially, every nilpotent operator
is algebraic and it is well-known that if K n (X) has finite dimension for some n ∈ N
then K is algebraic. An operator T ∈ L(X) is said to be Riesz if λI − T is Fredholm
for every λ = 0.
The SVEP is also stable under algebraic commuting or Riesz commuting
perturbations, see [5, 6]:
Theorem 3.6 Let T ∈ L(X) and K be algebraic which commutes with T . If T
has SVEP then T + K has SVEP. An anologous result holfd for Riesz commuting
perturbations.
Definition 3.7 A bounded operator T is said to be decomposable if, for any open
covering {U1 , U2 } of the complex plane C there are two closed T -invariant subspaces
Y1 and Y2 of X such that Y1 + Y2 = X and σ (T |Yk ) ⊆ Uk for k = 1, 2.
A bounded operator T ∈ L(X) on a Banach space X is said to have the Dunford
property (C) if every glocal spectral subspace is closed. We have
see [43], where property (δ) means that for every open covering (U, V ) of C we
have X = XT (U ) + XT (V ). Standard examples of decomposable operators are
normal operators on Hilbert spaces and operators which have totally disconnected
spectra, as for instance compact operators.
366 P. Aiena
Two important properties in local spectral theory related to property (C) are the
so-called property (β). This property has been introduced by Bishop [16] and is
defined as follows. Let U be an open subset of C and denote by H(U, X) the Fréchet
space of all analytic functions f : U → X with respect the pointwise vector space
operations and the topology of locally uniform convergence. T ∈ L(X) has Bishop’s
property (β) if for every open U ⊆ C and every sequence (fn ) ⊆ H(U, X) for
which (λI − T )fn (λ) converges to 0 uniformly on every compact subset of U , then
also fn → 0 in H(U, X).
Examples of operators which have property (β), are provided by the weighted
right shift on 2 (N) for which the weight sequence is increasing, see [43]. Note that
and
see [43].
Let T denote the dual of T . Property (β) and property (δ) are dual each
other, i.e., T ∈ L(X) satisfies (β) (respectively (δ)) if and only if T satisfies (δ)
(respectively, (β)), see [43]. Consequently,
Examples of operators satisfying property (β) but not decomposable may be found
among multipliers of semi-simple commutative Banach algebras, see [43].
In the sequel, for every set F ⊆ C we set F := {λ : λ ∈ F } and F cl the closure
of F . In the case of Hilbert space operators the dual T may be replaced by the
adjoint T ∗ . Let x ∈ HT ∗ (F ), for some closed F ⊆ C. Then there exists an analytic
function f : C → H such that (λI − T ∗ )f (λ) = x for all λ ∈ C \ F . From (1) we
know that U (λI − T ∗ ) = (λ̄I − T )U , so
H = U H = U HT ∗ (V cl ) + U HT ∗ (W cl ) = HT (V¯cl ) + HT (W¯cl ),
Normal Operators and their Generalizations 367
and hence, T has property (δ). An analogous argument shows that if T has property
(δ) then T ∗ has property (δ), so we have.
By duality, T has property (β) ⇔ T = (T ∗ )∗ has property (δ), and hence, by (8),
if and only if (T ∗ ) has property (δ), from which we conclude that
Consequently,
T is decomposable ⇔ T ∗ is decomposable.
Lemma 3.8 Let T ∈ L(X), X a Banach space, and λ ∈ ρ(T ). Then λ(λI −
T )−1 x → x for every x ∈ X as |λ| → +∞.
Proof Fix x ∈ X and define f (λ) := (λI − T )−1 x : ρ(T ) → X. It is known that
f (λ) → 0 when |λ| → +∞. We have, for every λ ∈ ρ(T ),
Both the functions f (λ) and g(λ) are defined in D0 \ {0} and for μ ∈ D0 \ {0} we
have
Define
f (μ̄), y if μ = D0 ,
h(μ) :=
x, g(μ) if μ ∈ D0 ,
The function h(μ) is well-defined and is analytic on C. Since f ( μ̄) = (μ̄I −T ∗ )−1 x
for all μ̄ ∈ ρ(T ∗ ), see [4, Remark 2.11] and f (μ̄) → 0 for |μ̄| → +∞, then
h(μ) → 0 as |μ| → +∞, so, by the classical Liouville theorem, h ≡ 0 on C. From
368 P. Aiena
Lemma 3.8 we have also have μ̄(μ̄I − T ∗ )−1 x = −x, as |μ| → +∞, μ̄ ∈ ρ(T ∗ ),
hence
so x ∈ K(T )⊥ , as desired.
Next we want to show that if T is decomposable then H0 (T ∗ ) = K(T )⊥ . To
do this we need some preliminary results. Suppose that M is a closed T -invariant
subspace of a Banach space X and denote by T /M : X/M → X/M the canonical
quotien mapping defined on the quotient X/M by (T /M)(x + M) := T x + M.
For an open disc D of C centered at 0, let D denote its closure.
Lemma 3.10 Suppose that T ∈ L(X), X a Banach space, is decomposable. If
M := XT (C \ D) then σ (T /M) is contained in D.
Proof This follows as a particular case of Theorem 1.2.23, part (b), of [43], by
taking F = C \ D.
If Y is a closed T -invariant subspace by T |Y we denote the restriction of T to Y .
Lemma 3.11 Suppose that T ∈ L(H ) is decomposable. If D is an open disc
centered at 0 then HT (C \ D)⊥ ⊆ HT ∗ (D).
Proof Let M := XT (C \ D). Recall that M is a closed invariant subspace of T ,
since a decomposable operator has property (C), while M ⊥ is a closed subspace
invariant under T ∗ . We show first that
σ (T ∗ |M) ⊆ D (9)
σ (T ∗ |M ⊥ ) = σ (T /M)∗ = σ (T /M).
σ (T ∗ |M ⊥ ) = σ (T ∗ |XT (C \ D)⊥ ) ⊆ D,
since D is the closure of D = D. Thus, the inclusion (9) is proved. From part (e) of
[43, Proposition 1.2.16] we then obtain HT (C \ D)⊥ = M ⊥ ⊆ HT ∗ (D).
Theorem 3.12 Let T ∈ L(H ) be decomposable then H0 (T ∗ ) = K(T )⊥ and
H0 (T ) = K(T ∗ )⊥ .
Normal Operators and their Generalizations 369
Proof To show the equality H0 (T ∗ ) = K(T )⊥ , let {Dα }α denote the set of all closed
discs of C centered at 0. Since T has SVEP we have
D
H0 (T ∗ ) = HT ∗ ({0}) = HT ∗ ({0}) = HT ∗ (Dα ),
α
see [4, Theorem 2.13, part (iv)]. To show the equality H0 (T ∗ ) = K(T )⊥ we need to
prove, by Theorem 3.9, the inclusion K(T )⊥ ⊆ H0 (T ∗ ), and for this it suffices to
prove that K(T )⊥ ⊆ HT ∗ (D), where D is any closed disc centered at 0. Evidently,
HT (C \ D) ⊆ HT (C \ {0}) = K(T ),
H0 (T ) = H0 ((T ∗ )∗ ) = K(T ∗ )⊥ .
Remark 3.13 It should be noted that the identity K(T ) = H0 (T ∗ )⊥ in general does
not hold even if T is decomposable. For instance, if T ∈ L(H ) is Riesz operator
which has infinite spectrum then T is decomposable, but K(T ) is not closed, since in
this case σ (T ) would be finite, see [50]. Hence K(T ) = H0 (T ∗ )⊥ , since H0 (T ∗ )⊥
is closed.
Corollary 3.14 If T ∈ L(H ) is self-adjoint then H0 (T ) = K(T )⊥ .
Proof T is decomposable and T = T ∗ .
Perhaps the most important class of operators in Hilbert spaces is given by the
normal operators defined on a Hilbert space. Recall that if H is a complex infinite-
dimensional Hilbert space a bounded linear operator T ∈ L(H ) is said to be normal
if
T T ∗ = T ∗T (10)
Normal operators have several important spectral properties that will be next
recalled.
(A) Every isolated spectral point of a normal operator T is a simple pole of the
resolvent. If every isolated spectral point of an operator T on a Banach space
is a pole then T is said to be polaroid.
370 P. Aiena
σ (T ) \ σw (T ) = π00 (T ), (11)
where
i.e, the spectral points for which λI − T ∈ W (X) are exactly the eigenvalues
which have finite multiplicity. An operator T ∈ L(X), X a Banach space, for
which the equality (11) holds is said to satisfy Weyl theorem.
(D) T is decomposable; in particular both T and T ∗ have SVEP.
Normal operators may be generalized in several ways:
Hyponormal Operators This class of operators on Hilbert spaces is defined
whenever the condition of normality (10) is relaxed to the inequality
T ∗T ≥ T T ∗. (12)
T ∗ x ≤ T x for all x ∈ H.
T ∗ T x, x ≥ T T ∗ x, x for allx ∈ H,
or equivalently
T ∗ x2 = T ∗ x, T ∗ x = T T ∗ x, x ≤ T ∗ T x, x = T x, T x = T x2 .
Thus, T ∗ x ≤ T x.
A routine computation shows that a weighted right shift on the Hilbert space
2 (N) is hyponormal if and only if the corresponding weight sequence is increasing.
Normal Operators and their Generalizations 371
T (T ∗ T ) = (T ∗ T )T .
A very easy example of quasi-normal operator is given by the unilateral right shift
R on the Hilbert space 2 (N). Recall that a such operator is defined by
and obviously R(R ∗ R) = (RR ∗ )R. Note that R is not normal, since
RR ∗ = RL = R ∗ R = LR = I.
for details see Furuta [33, p. 105]. In the sequel we show some relevant properties
of hyponormal operators.
Theorem 4.1 Let T ∈ L(H ) be hyponormal. Then we have:
(i) λI − T is hyponormal for every λ ∈ C.
(ii) If M is a closed invariant subspace of H then T |M is hyponormal.
Proof
(i) We have
thus, λI − T is hyponormal.
(ii) IfPM is the projection of T onto M, then (T |M)∗ = (PM T ∗ )|M. For every
x ∈ M then we have
thus T |M is hyponormal.
372 P. Aiena
Lemma 4.2 Let T ∈ L(H ) be a self-adjoint operator such that λI ≤ T for some
λ ≥ 0. Then T is invertible. In particular, if I ≤ T then 0 ≤ T −1 ≤ I .
Proof To show the first assertion, observe that by the Schwarz inequality we have
T xx ≥ (T x, x) ≥ cx2 ,
T −1 (T − I ) = I − T −1 ≥ 0,
thus T −1 ≤ I .
It is easily seen that if T is self-adjoint then ST S ∗ is also self-adjoint for every
S ∈ L(H ). Moreover, if T is positive then ST S ∗ ≥ 0 for all S ∈ L(H ).
Theorem 4.3 If T ∈ L(H ) is an invertible hyponormal operator then its inverse
T −1 is also hyponormal.
Proof Suppose that T is hyponormal. Then T ∗ T − T T ∗ ≥ 0 and hence, as noted
above, the product
T −1 (T ∗ T − T T ∗ )(T −1 )∗
T −1 (T ∗ T )(T −1 )∗ − I ≥ 0,
and hence
T −1 (T ∗ T )(T −1 )∗ ≥ I,
0 ≤ [T −1 (T ∗ T )(T −1 )∗ ]−1 ≤ I.
Normal Operators and their Generalizations 373
S := I − T ∗ (T −1 (T ∗ )−1 )T
is positive, so
T −1 ST −1 ≥ 0,
(T −1 )∗ T −1 − T −1 (T −1 )∗ ≥ 0.
Hence T −1 is hyponormal.
Paranormal Operators An operator T ∈ L(X) on a Banach space X is said to be
paranormal if
T k+1 x T k+2 x
≤
T k x T k+1 x
T n x T x T 2 x T n x
= ···
x x T x T n−1 x
T n+1 x T n+2 x T 2n x T 2n x
≤ · · · = .
T n x T n+1 x T 2n−1 x T n x
We claim that h(T ) = h(0)I . To see that let us consider the two possibilities: h(0) =
0 or h(0) = 0.
If h(0) = 0 then h(T ) is quasi-nilpotent, so from the implication (14), we deduce
that h(T ) = 0, hence the equality h(T ) = h(0)I trivially holds.
Suppose the other case h(0) = 0, and set h1 (T ) := h(0)1
h(T ). Clearly, h1 (T ) has
spectrum {1} and h1 (T ) = 1. Moreover, h1 (T ) is invertible and also its inverse
h1 (T )−1 has norm 1. The operator h1 (T ) is then doubly power-bounded and by
a classical theorem due to Gelfand, see [43, Theorem 1.5.14] for a proof, it then
follows that h1 (T ) = I , and hence h(T ) = h(0)I , as claimed.
Now, from the equality h(0)I − h(T ) = 0, we see that there exist some natural
n ∈ N and μ ∈ C for which
n
0 = h(0)I − h(T ) = μ T m (λi I − T ) with λi = 0,
i=1
T x2 = (T x, T x) = (T ∗ T x, x) ≤ T ∗ (T x)x
≤ T (T x)x = T 2 xx,
Therefore, H0 (λI − T ) ⊆ ker(λI − T ), and since the reverse inclusion holds for
every operator, then we have H0 (λI − T ) = ker(λI − T ).
Equation (15) has a remarkable conseguence. To see this, observe that if λ ∈
iso σ (T ), where iso K denotes the isolated points of a set K ⊆ C, then by
Theorem 2.10, we have
and hence
Therefore,
while
Let λ ∈ D0 and μ ∈ D0 be two distinct complex numbers such that f (λ) and f (μ)
are non-zero. Since ker (μI − T ) ⊥a ker (λI − T ), then
and hence
Therefore,
by
from which it follows that Ax ∈ H0 (λI − S) = ker (λI − S). Hence A(λI − T )x =
(λI − S)Ax = 0 and, since A is injective, this implies that (λI − T )x = 0, i.e.
x ∈ ker (λI − T ). Therefore H0 (λI − T ) = ker (λI − T ) for all λ ∈ C.
The more general case of H (p)-operators is proved by using a similar argument.
0 ≤ PM ≤ I,
Normal Operators and their Generalizations 381
R := |T |1/2W |T |1/2
log (T ∗ T ) ≥ log (T T ∗ ).
382 P. Aiena
T1 := |R|1/2 V |R|1/2 .
(T T )p ≥ (T T )p .
T T ∗ ≤ MT ∗ T .
|T̂ | ≥ |T | ≥ |T̂ ∗ |.
T ∗ |T ∗ |2p T ≤ T ∗ |T |2p T .
|T 2 | ≥ |T |2 .
Every log-hyponormal operator is a class A operator [34], but the converse is no true,
see [33, p. 176]. Every class A operator is paranormal (an example of a paranormal
operator which is not a class A operator can be found in [33, p. 177]). Therefore
every class A operator, as well as every algebraically class A operator, is polaroid.
Quasi-Class A Operators An operator T ∈ L(H ) is said to be a quasi-class A
operator if T ∗ |T 2 |T ≥ T ∗ |T |2 T . The quasi-class A operators contains the class of
al p-quasinormal operators and the class of all class A operators. In [31] it is given
an example of a quasi-class A operator which is not paranormal. Every quasi-class
A operator has SVEP, since p(λI − T ) ≤ 1 for all λ ∈ C, while every non-zero λ0
isolated point of the spectrum is a simple pole of T and H0 (λ0 I −T ) = ker(λ0 I −T ),
see [31]. It has been observed in [31, Example 0.2] that a quasi-class A operator need
not to be normaloid.
∗-Paranormal Operators A bounded operator T ∈ L(H ) is said to be ∗-
paranormal if T ∗ x2 ≤ T 2 x for every unit vector x ∈ H . Paranormality
is independent of ∗-paranormality and, evidently, hyponormal operators are both
paranormal and ∗-paranormal. It is known [13] that
T ∗ k |T ∗ |2p T k ≤ (T ∗ k |T |2p T k .
Evidently,
1. a (1, 1)-quasihyponormal operator is quasihyponormal;
2. a (p, 1)-quasihyponormal operator is p-quasihyponormal;
3. a (p, 0)-quasihyponormal operator is p-hyponormal if 0 < p < 1 and
hyponormal if p = 1.
In [57, Theorem 6] it has been proved that every (p, k)-quasihyponormal
operator T ∈ L(H ) is hereditarily polaroid. Moreover, every isolated point λ = 0 is
a simple pole of the resolvent.
It should be noted that the class of totally ∗-paranormal operators, as well as
the class of M-hyponormal operators, are independent of the classes (p, k)-quasi-
hyponormal.
σ (f (T )) = f (σ (T )) = {f (0)}.
Normal Operators and their Generalizations 385
We claim that f (T ) = f (0)I . To see this, let us consider the two possibilities:
f (0) = 0 or f (0) = 0.
If f (0) = 0 then f (T ) is quasi-nilpotent and f (T ) is normaloid, and hence
f (T ) = 0. The equality f (T ) = f (0)I then trivially holds.
Suppose the other case f (0) = 0, and set f1 (T ) := f (0) 1
f (T ). Clearly,
σ (f1 (T )) = {1} and f1 (T ) = 1. Further, f1 (T ) is invertible and is T HN .
This easily implies that its inverse f1 (T )−1 has norm 1. The operator f1 (T )
is then doubly power-bounded and, by a classical theorem due to Gelfand (see
[43, Teorem 1.5.14] for an elegant proof), it then follows that f1 (T ) = I , and
consequently f (T ) = f (0)I , as claimed.
Now, define g(λ) := f (0) − f (λ). Clearly, g(0) = 0, and g may have only a
finite number of zeros in σ (T ). Let {0, λ1 , . . . , λn } be the set of all zeros of g, where
λi = λj , for all i = j , and λi has multiplicity ni ∈ N. Write
n
g(λ) = μλm (λi I − T )ni h(λ),
i=1
n
0 = g(T ) = μ T m (λi I − T )ni h(T ) with λi = 0,
i=1
where all the operators λi I − T and h(T ) are invertible. This, obviously, implies
that T m = 0, i. e. T is nilpotent.
If T ∈ L(X) the numerical range of T is defined as
In the case of Hilbert space operator the numerical range may be described as the
set
r(T ) ≤ w(T ) ≤ T .
386 P. Aiena
The next nontrivial result has been proved in Sinclair [55], we omit the not simple
proof.
Theorem 5.3 Let T ∈ L(X) and suppose that 0 is in the boundary of the numerical
range of T . Then the kernel of T is orthogonal (in the sense of Birkoff-James) to the
range of T .
In the case of paranormal operators we have:
Theorem 5.4 Suppose that T ∈ L(X) is totally hereditarily normaloid , λ, μ ∈ C,
with λ = 0 and λ = μ. Then ker (λI − T ) ⊥ ker (μI − T ), i.e. ker (λI − T ) is
orthogonal to ker (μI − T ) in the Birkhoff and James sense.
Proof Suppose first that |λ| ≥ |μ| and let x ∈ ker (λI − T ), y ∈ ker (μI − T ).
Then T x = λx and T y = μy. Denote by M the subspace generated by x and y and
set TM := T |M. Clearly, σ (T |M) = {λ, μ} and being T |M normaloid then
so that μ(T |M) = |λ|. Consequently, λ belongs to the boundary of the numerical
range of T |M and hence, by Theorem 5.3, ker(λI − T |M) ⊥ (λI − T |M)(M).
Evidently, λ and μ are poles of the resolvent of T |M having order 1. Denoting by Pλ
and Pμ the spectral projections for T |M associated with {λ} and {μ}, respectively,
we then have
1 1
σ (T |M)−1 = { , },
λ μ
xi − xj ≥ 1.
Normal Operators and their Generalizations 387
is T HN .
We recall now some elementary algebraic facts. Suppose that T ∈ L(X) and
X = M ⊕ N, with M and N closed subspaces of X, M invariant under T . With
respect to this decomposition of X it is known that T may be represented by a
AB
upper triangular operator matrix , where A ∈ L(M), C ∈ L(N) and
0 C
x
B ∈ L(N, M). It is easily seen that for every x = ∈ M we have T x = Ax,
0
so A = T |M. Let us consider now the case of operators T acting on a Hilbert space
H , and suppose that T k (H ) is not dense in H . In this case we can consider the
nontrivial orthogonal decomposition
⊥
H = T k (H ) ⊕ T k (H ) , (16)
⊥
where T k (H ) = ker(T ∗ )k , T ∗ the adjoint of T . Note that the subspace T k (H ) is
T -invariant, since
T (T k (H )) ⊆ T (T k (H )) = T k+1 (H ) ⊆ T k (H ).
Thus we can represent, with respect the decomposition (16), T as an upper triangular
operator matrix
T1 T2
, (17)
0 T3
388 P. Aiena
⊥
where T1 = T |T k (H ). Moreover, T3 is nilpotent. Indeed, if x ∈ T k (X) , an easy
computation yields
0
T kx = T = T3 k x.
x
⊥
Hence T3 k x = 0, since T k x ∈ T k (H ) ∪ T k (H ) = {0}. Therefore we have:
Theorem 5.9 Suppose that T ∈ L(H ) and T k (H ) non dense in H . Then,
T1 T2
according the decomposition (16), T = is quasi-T HN if and only if
0 T3
T1 is T HN . Furthermore,
The class of (1, k)- quasiparanormal operators has been studied in [48]. If T k (H )
T1 T2
is not dense then, in the triangulation T = , T1 = T |T k (H ) is n-
0 T3
quasiparanormal, and hence T HN , see [62].
2. An extension of class A operators is given by the class of all k-quasiclass A
operators, where T ∈ L(H ), H a separable infinite dimensional Hilbert space,
is said to be a k-quasiclass A operator if
T ∗ k (|T |2 − |T |2 )T k ≥ 0.
This class of operators contains the class of all quasi- ∗-paranormal operators
(which corresponds to the value k = 1). Every k-quasi *-paranormal operator is
quasi-T HN . Indeed, if T k has dense range then T is ∗-paranormal and hence
T HN . If T k does not have dense range then T may be decomposed, according
T1 T2
the decomposition H = T k (H ) ⊕ ker T ∗ k , as T = , where T1 =
0 T3
T |T k (H ) is ∗-paranormal, hence T HN , see [49, Lemma 2.1].
Every (p, k)-quasihyponormal operator T with respect to the decomposition
T1 T2
H = T k (H ) ⊕ ker T ∗ k , may be represented as a matrix T = , where
0 0
T1 := T |T k (H ) is k-hyponormal (hence paranormal) and consequently T HN ,
see [40]. The next result generalizes the result of Lemma 5.2.
Theorem 5.10 Suppose that T ∈ L(H ), H a Hilbert space, is analytically quasi-
T HN and quasi-nilpotent. Then T is nilpotent.
Proof Suppose first that T is quasi-nilpotent and k-quasi T HN . If T k (H ) is dense
then T is T HN , so T is nilpotent by Theorem 5.2. Suppose that T k (H ) is not dense
T1 T2
and write T = , where T1 is T HN , T3 k = 0, and σ (T ) = σ (T1 ) ∪ {0}.
0 T3
Since σ (T ) = {0} and σ (T1 ) is not empty, we then have σ (T1 ) = {0}, thus T1 is
0 T2
a quasi-nilpotent T HN operator and hence T1 = 0. Therefore T = . An
0 T3
easy computation yields that
k+1 k+1
0 T2 0 T2 T3k
T k+1 = = = 0,
0 T3 0 T3k+1
so that T is nilpotent.
Finally, suppose that T is quasi-nilpotent and analytically k-quasi T HN . Let h
be analytic on an open neighborhood of σ (T ), and non-constant on the components
of its domain, be such that h(T ) is quasi-T HN . We claim that h(T ) is nilpotent.
If h(T )k has dense range then h(T ) is T HN and hence, by Lemma 5.2, h(T )
is nilpotent. Suppose that h(T )k has not dense range. Then with respect the
⊥
decomposition X = h(T )k (H ) ⊕ h(T )k (H ) , the operator h(T ) has a triangulation
390 P. Aiena
AB
h(T ) = , such that A = h(T )|h(T )k (H ) is T HN and
0 C
n
h(λ) = μλ (λi I − T )ni g(λ),
n
i=1
where g(λ) has no zeros in σ (T ) and λi = 0 are the other zeros of g with
multiplicity ni . Hence
n
h(T ) = μ T n (λi I − T )ni g(T ),
i=1
where all λi I − T and g(T ) are invertible. Since h(T ) is nilpotent then also T is
nilpotent.
Theorem 5.11 If T ∈ L(H ) is an analytically quasi T HN operator, then T is
polaroid.
Proof We show that for every isolated point λ of σ (T ) we have p(λI − T ) =
q(λI − T ) < ∞. Let λ be an isolated point of σ (T ). Then H = H0 (λI − T ) ⊕
K(λI − T ), by Theorem 2.10. Furthermore, since σ (T |H0 (λI − T )) = {λ}, while
σ (T |K(λI − T )) = σ (T ) \ {λ}, so the restriction λI − T |H0 (λI − T ) is quasi-
nilpotent and λI −T |K(λI −T ) is invertible. Since λI −T |H0 (λI −T ) is analytically
quasi T HN , then Lemma 5.10 implies that λI −T |H0 (λI −T ) is nilpotent. In other
worlds, λI − T is an operator of Kato Type.
Now, both T and the dual T ∗ have SVEP at λ, since λ is isolated in σ (T ) =
σ (T ∗ ), and this implies, by Theorem 3.5, that both p(λI − T ) and q(λI − T ) are
finite. Therefore, λ is a pole of the resolvent.
Lemma 5.12 Suppose that T ∈ L(X) admits, with respect to the decomposition
T1 T2
X = M ⊕ N, the representation T = , where T3 is nilpotent. Then T has
0 T3
SVEP if and only if T1 has SVEP.
Normal Operators and their Generalizations 391
λI − T1 −T2 f1 (λ)
0 = (λI − T )f (λ) =
0 −λI − T3 f2 (λ)
Then (λI − T3 )f2 (λ) = 0 and (λI − T1 )f1 (λ) − T2 f2 (λ) = 0. Since a nilpotent
operator has SVEP then f2 (λ) = 0, and consequently (λI − T1 )f1 (λ) = 0. But T1
has SVEP at λ0 , so f1 (λ) = 0 and hence f (λ) = 0 on U . Thus, T has SVEP at λ0 .
Since λ0 is arbitrary then T has SVEP.
Conversely, suppose that T has SVEP. Since T1 is the restriction of T to M and
the SVEP from T is inherited by the restriction to closed invariant subspaces, then
T1 has SVEP.
Theorem 5.13 If T ∈ L(H ) is analytically quasi T HN , then T is hereditarily
polaroid and hence has SVEP.
Proof Let f ∈ Hnc (σ (T )) such that f (T ) is quasi T HN . If M is a closed T -
invariant subspace of X, we know that f (T )|M is quasi T HN , by Lemma 5.8, and
f (T )|M = f (T |M), so f (T |M) is polaroid, by Theorem 5.11, and consequently,
T |M is polaroid, see [4, Theorem 4.19].
σw (T ) := {λ ∈ C : λI − T ∈
/ W (X)}.
σuw (T ) := {λ ∈ C : λI − T ∈
/ W+ (X)},
σlw (T ) := {λ ∈ C : λI − T ∈
/ W− (X)}.
σb (T ) := {λ ∈ C : λI − T ∈
/ B(X)}.
Following Coburn [21], we say that Weyl’s theorem holds for T ∈ L(X) (in symbol,
(W )) if
σ (T ) \ σw (T ) = π00 (T ), (18)
where
σbw (T ) := {λ ∈ C : λI − T ∈
/ BW (X)}.
Normal Operators and their Generalizations 393
Another version of Weyl’s theorem has been introduced by Berkani and Koliha ([14]
as follows: T ∈ L(X) is said to verify generalized Weyl’s theorem, (in symbol
(gW )) if
where
An operator T ∈ L(X) is said to be Drazin invertible if p(T ) and q(T ) are finite,
and this happens if and only if T is invertible or 0 is a pole of the resolvent. The
Drazin spectrum id denoted by σd (T ). Note that (gW )) holds for T if and only if T
satisfies generalized Browder’s theorem, (i.e., σbw = σd , or equivalently, Browder’s
theorem, see [4, Theorem 5.15]) and E(T ) = (T ), where (T ) is the set of all
poles of the resolvent of T . Note that generalized Weyl’s theorem entails Weyl’s
theorem, see [4, Theorem 6.60].
The following result shows that in presence of SVEP the polaroid condition
entails Weyl type theorems.
Theorem 6.1 Let T ∈ L(X) be polaroid and suppose that either T or T has SVEP.
Then both T and T satisfy generalized Weyl’s theorem.
Proof If T is polaroid also T is polaroid, and Weyl’s theorem and generalized
Weyl’s theorem for T , or T , are equivalent, see [10, Theorem 3.7]. The assertion
then follows from [10, Theorem 3.3].
Remark 6.2 In the case of a Hilbert space operator T ∈ L(H ) it is more
appropriated to consider the Hilbert adjoint T ∗ instead of the dual T . Note that
T satisfies (gW ) if and only if T ∗ does. This easily follows from the well known
equalities, σw (T ∗ ) = σw (T ), where E is the conjugate of E ⊆ C, σb (T ∗ ) =
σb (T ), E(T ∗ ) = E(T ), and (T ∗ ) = (T ). Furthermore, T satisfies SVEP
if and only if T ∗ satisfies SVEP, so, in the statement of Theorem 6.1, T may be
replaced by the Hilbert adjoint T ∗ .
In [3] it is shown that if T is hereditarily polaroid and has SVEP, and K is an
algebraic operator which commutes with T then T + K is polaroid and T + K is
a-polaroid.
The following perturbation result has been proved in [3, Theorem 3.12].
Theorem 6.3 Suppose that T ∈ L(X) and K ∈ L(X) an algebraic operator
commuting with T ∈ L(X). If T ∈ L(X), or T , has SVEP and T , or T , is
hereditarily polaroid, then f (T + K) and f (T + K ) ) satisfies (gW ) for every
f ∈ Hnc (σ (T + K)).
394 P. Aiena
T ∗ + K ∗ is a-polaroid ⇔ T + K is a-polaroid,
References
1. P. Aiena, Fredholm and Local Spectral Theory, with Application to Multipliers (Kluwer,
Dordecht, 2004)
2. P. Aiena, Algebraically paranormal operators on Banach spaces. Banach J. Math. Anal. 7(2),
136–145 (2013)
3. P. Aiena, E. Aponte, Polaroid type operators under perturbations. Stud. Math. 214(2), 121–136
(2013)
4. P. Aiena, Fredholm and Local Spectral Theory II, with Application to Weyl-Type Theorems.
Springer Lecture Notes of Mathematics, vol. 2235 (Springer, Berlin, 2018)
5. P. Aiena, V. Muller, The localized single-valued extension property and Riesz operators. Proc.
Am. Math. Soc. 143(5), 2051–2055 (2015)
6. P. Aiena, M.M. Neumann, On the stability of the localized single-valued extension property
under commuting perturbations. Proc. Am. Math. Soc. 141(6), 2039–2050 (2013)
7. P. Aiena, P. Peña, A variation on Weyl’s theorem. J. Math. Anal. Appl. 324, 566–579 (2006)
8. P. Aiena, S. Triolo, Weyl type theorems on Banach spaces under compact perturbations.
Mediterr. J. Math. 15(3), 1–18 (2018). https://doi.org/10.1007/s00009-018-1176-y
9. P. Aiena, S. Triolo, Some remarks on the spectral properties of Toeplitz operators. Mediterr. J.
Math. 16, 135 (2019). http://dx.doi.org/10.1007/s00009-019-1397-8
10. P. Aiena, E. Aponte, E. Bazan, Weyl type theorems for left and right polaroid operators. Integr.
Equ. Oper. Theory 66, 1–20 (2010)
11. A. Aluthge, On p-hyponormal operators for 1 < p < 1. Integr. Equ. Oper. Theory 13, 307–315
(1990)
12. T. Ando, Operators with a norm condition. Acta Sci. Math. 33, 169–178 (1972)
13. S.C. Arora, J.K. Thukral, On a class of operators. Glasnik Math. 21, 381–386 (1986)
Normal Operators and their Generalizations 395
14. M. Berkani, J.J. Koliha, Weyl type theorems for bounded linear operators. Acta Sci. Math.
69(1–2), 359–376 (2003)
15. M. Berkani, M. Sarih, On semi B-Fredholm operators. Glasgow Math. J. 43, 457–465 (2001)
16. E. Bishop. A duality theorem for an arbitrary operator. Pac. J. Math. 9, 379–397 (1959)
17. M. Chō, Spectral properties of p-hyponormal operators. Glasgow Math. J. 436, 117–122
(1994)
18. M. Chō, J.I. Lee, p-hyponormality is not translation-invariant. Proc. Am. Math. Soc. 131(10),
3109–3111 (2003)
19. M. Chō, I.H. Jeon, J.I. Lee, Spectral and structural properties of log-hyponormal operators.
Glasgow Math. J. 42, 345–350 (2000)
20. N.N. Chourasia, P.B. Ramanujan, Paranormal operators on Banach spaces. Bull. Austral. Math.
Soc. 21, 161–168 (1980)
21. L.A. Coburn, Weyl’s theorem for nonnormal operators. Michigan Math. J. 13(3), 285–288
(1966)
22. J.B. Conway, The Theory of Subnormal Operators. Mathematical Survey and Monographs,
vol. 36 (American Mathematical Society, Providence, 1992)
23. I. Colojoară, C. Foiaş, Theory of Generalized Spectral Operators (Gordon and Breach, New
York, 1968)
24. D.S. Djordjević, Operators obeying a-Weyl’s theorem. Publ. Math. Debrecen 55(3–4), 283–
298 (1999)
25. N. Dunford, J.T. Schwartz, Linear Operators. Part I (1967), Part II (1967), Part III (Wiley, New
York, 1967)
26. R.G. Douglas Banach Algebra Techniques in Operator Theory. Graduate Texts in Mathematics,
vol. 179, 2nd edn. (Springer, New York, 1998)
27. B.P. Duggal, Isolated eigenvalues, poles, compact perturbation of Banach space operators.
Oper. Matrices Debrecen 13(4), 966–973 (2019)
28. B.P. Duggal, S.V. Djordjević, Weyl’s theorems and continuity of the spectra in the class of
p-hyponormaloperators. Stud. Math. 143(1), 23–32 (2000)
29. B.P. Duggal, S.V. Djordjevíc, Generalized Weyl’s theorem for a class of operators satisfying a
norm condition. Math. Proc. Royal Irish Acad. 104A, 75–81 (2004)
30. B.P. Duggal, H. Jeon, Remarks on spectral properties of p-hyponormal and log-hyponormal
operators. Bull. Kor. Math. Soc. 42, 541–552 (2005)
31. B.P. Duggal, I.H. Jeon, H. Kim, On Weyl’s theorem for quasi-class A operators. J. Korean
Math. Soc. 43, 899–909 (2006)
32. D.R. Farenick, W.Y. Lee, Hyponormality and spectra of Toeplitz operators. Trans. Am. Math.
Soc. 348(10), 4153–4174 (1996)
33. T. Furuta, Invitation to Linear Operators (Taylor and Francis, London, 2001)
34. T. Furuta, M. Ito, T. Yamazaki, A subsclass of paranormal operators including class of log-
hyponormal and several related classes. Sci. Math. 1, 389–403 (1998)
35. S.R. Garcia, M. Putinar, Complex symmetric operators and applications I. Trans. Am. Math.
Soc. 358, 1285–1315 (2006)
36. S.R. Garcia, M. Putinar, Complex symmetric operators and applications II. Trans. Am. Math.
Soc. 359, 3913–3931 (2007)
37. Y.M. Han, A.-H. Kim, A note on ∗-paranormal operators. Integr. Equ. Oper. Theory 49, 435–
444 (2004)
38. Y.M. Han, J.I. Lee, D. Wang, Riesz idempotent and Weyl’s theorem for w-hyponormal
operator. Integr. Equ. Oper. Theory 53, 51–60 (2005)
39. H. Heuser, Functional Analysis (Marcel Dekker, New York, 1982)
40. I.H. Kim, On (p, k)-quasihyponormal operators. Math. Inequal. Appl. 4, 629–638 (2004)
41. E. Ko, On p-hyponormal operators. Proc. Am. Math. Soc. 128(3), 775–780 (2000)
42. D. Lay, A. Taylor, Introduction to Functional Analysis (Wiley, New York, 1980)
43. K.B. Laursen, M.M. Neumann, An Introduction to Local Spectral Theory. London Mathemat-
ical Society Monographs, vol. 20 (Clarendon Press, Oxford, 2000)
44. W.Y. Lee, Weyl’spectra of operator matrices. Proc. Am. Math. Soc. 129, 131–138 (2001)
396 P. Aiena
45. M.Y. Lee, S.H. Lee, Some generalized theorems on p-quasihyponormal operators for 0 < p <
1. Nihonkai Math. J. 8, 109–115 (1997)
46. C. Lin, Y. Ruan, Z. Yan, w-hyponormal operators are subscalar. Integr. Equ. Oper. Theory 50,
165–168 (2004)
47. M. Mbekhta, Local spectrum and generalized spectrum. Proc. Am. Math. Soc. 112, 457–463
(1991)
48. S. Mecheri, Bishop’s property (β) and Riesz idempotent for k-quasi- paranormal operators.
Banach J. Math. Anal. 6, 147–154 (2012)
49. S. Mecheri, On a new class of operators and Weyl type theorems. Filomat 27(4), 629–636
(2013)
50. T.L. Miller, V.G. Miller, M.M. Neumann, Operators with closed analytic core. Rend. Circolo
Mat. Palermo. 51(3), 495–502 (2003)
51. V. Müller, Spectral Theory of Linear Operators and Spectral Systems on Banach Algebras.
Operator Theory: Advances and Applications, 2nd edn. (Birkhäuser, Berlin, 2007)
52. M. Putinar, Hyponormal operators are subscalar. J. Operator Theory 12, 385–395 (1984)
53. W. Rudin, Functional Analysis, 2nd edn. (McGraw-Hill, New York, 1991)
54. C. Schmoeger, On isolated points of the spectrum of a bounded operator. Proc. Am. Math. Soc.
117, 715–719 (1993)
55. A.M. Sinclair, Eigenvalues in the boundary of the numeraical range. Pac. J. Math. 81, 231–234
(1970)
56. K. Tanahashi, On log-hyponormal operators. Integral Equ. Oper. Theory 34, 364–372 (1999)
57. K. Tanahashi, A. Uchiyama, M. Chō, Isolated points of the spectrum of (p, k)-quasi-
hyponormal operators. Linear Algebra Appl. 382, 221–229 (2004)
58. F.H. Vasilescu, Analytic Functional Calculus and Spectral Decompositions (Editura
Academiei/D. Reidel Publishing Company, Bucharest/Dorrecht, 1982)
59. P. Vrbová, On local spectral properties of operators in Banach spaces. Czechoslov. Math. J.
23(98), 483–492 (1973)
60. H. Widom, On the spectrum of Toeplitz operators. Pac. J. Math. 14, 365–375 (1964)
61. D. Xia, Spectral Theory of Hyponormal Operators. (Birkhauser, Boston, 1993)
62. J.T. Yuan, G.X. Ji, On (n, k)-quasi paranormal operators. Stud. Math. 209, 289–301 (2012)
On Wold Type Decomposition for Closed
Range Operators
1 Introduction
Let us denote first by H a Hilbert space and by L(H) the algebra of all bounded
linear operators on H. For an operator T ∈ L(H), we denote by R(T ) and N(T )
the
E∞range and the kernel sub-spaces
H of T respectively. We also write R ∞ (T ) =
∞
k=0 R(T ) and N (T ) =
k n
n≥0 N(T ) for the generalized range and the
generalized kernel of T respectively. We will say that T is a pure operator if
R ∞ (T ) = {0}.
A subspace E ⊂ H is said to be T −invariant if T (E) ⊆ E, the lattice of all
closed T −invariant sub-spaces in H will be denoted by Lat (T , H). For any given
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 397
R. M. Aron et al. (eds.), Operator and Norm Inequalities and Related Topics,
Trends in Mathematics, https://doi.org/10.1007/978-3-031-02104-6_12
398 H. Ezzahraoui et al.
I
subspace E of H, we set [E]T = ∞ k
k=0 T (E) for the smallest closed T -invariant
subspace of H containing E. A subspace E is called a reducing subspace for T if it
is invariant for both T and T ∗ in which T ∗ is the adjoint operator of T .
Recall that the operator T ∈ L(H) is bounded below if there is c > 0 satisfying
T x ≥ cx for every x ∈ H. It is not difficult to see that an operator T is bounded
below if and only if it is one to one and has a closed range. This is also equivalent
to T ∗ T invertible and hence to T is left invertible. A standard left inverse of T is
given by L = (T ∗ T )−1 T ∗ . The reduced minimum modulus γ (T ) of T is defined by
the formula:
M = [M A T M]T .
Later on, S.Shimorin introduced a more general property named the Wold-type
decomposition which generalizes the classical Wold decomposition and proved in
[24, Theorem 3.6]. S. Shimorin utilized a concept from the paper of Richter from
On Wold Type Decomposition for Closed Range Operators 399
1988, who showed that an analytic 2-concave operator has the wandering subspace
property (see [20]). Shimorin obtain a weak analog of the Wold decomposition
theorem, representing operator close to isometry in some sense as a direct sum of a
unitary operator and a shift operator acting in some reproducing kernel Hilbert space
of vector-valued holomorphic functions defined on a disc. The construction of the
Shimorin’s model for a left-invertible analytic operator becomes as a powerful tool
in the model theory of left-invertible operators.
be a Hilbert space endowed with some orthonormal basis (en )n≥0 . The weighted
shift Sω on Hω is defined by Sω en = ω(n)en+1 . We devote this section to some
classical results concerning Beurling theorems for weighted shift operators.
The hardy space H 2 (D) of analytic functions on the unit disc D is given as the
Hilbert space
⎧ ⎫
⎨ ⎬
H 2 (D) := f (z) = an zn : f 2H 2 (D) := |an |2 < ∞ .
⎩ ⎭
n≥0 n≥0
E = [E A Mz E]Mz .
The standard Bergman space L2a (D) of square area integrable analytic functions on
the unit disc is given as the Hilbert space
$
1 1
L2a (D) = {f (z) = an zn : f 2 = |f (z)|2 dA(z) = |an |2 < ∞.}
2π D n+1
n≥0 n≥0
The Dirichlet space D(D) consists of analytic functions on the unit disc D is
$
D(D) = {f (z) = an zn : D(f ) := |f (z)|2 dA(z) < ∞},
n≥0 D
On Wold Type Decomposition for Closed Range Operators 401
here dA(z) = 1
π rdrdt denotes normalized area measure on D. A norm on D(D) is
defined by
∞
f 2D = f 2H 2 (D) + D(f ) = (n + 1)|an |2 .
n=0
1 n
{en (z) = z : n ≥ 0}
n+1
is a canonical orthonormal basis. The main theorem of Beurling type in the case of
Dirichlet space is given by Richter in [20]. In this space, every z-invariant subspace
E of D(D) is generated by an extremal function. More precisely, E = φD(mφ ),
where φ is a normalized extremal function, mφ is a certain absolutely continuous
measure on the unit circle T, and D(mφ ) is a Dirichlet-type space associated with
mφ . Moreover, E := E A zE = Cφ is a one dimensional wandering subspace such
that
E = [E A Mz E]Mz .
In the Hardy space on the bi-disc H 2 (D2 ), Beurling theorem fails in general. Indeed,
W. Rudin provided two examples showing that none of the equalities in Beurling
theorem hold. Recall that a closed subspace E ⊂ H 2 (D2 ) is invariant under the bi-
shift M(z1 ,z2 ) = (Mz1 , Mz2 ) if and only if (z1 E +z2 E) ⊂ E. Again E A(z1 E +z2 E)
is a wandering subspace.
Example ([21]) The invariant subspace [z1 − z2 ]M(z1 ,z2 ) is not of the form θ H 2 (D2 )
for any two variable inner function θ ∈ H 2 (D2 ).
Example ([21]) Let E be the set of all functions f ∈ H 2 (D2 ) which have a zero
of order greater than or equal to n at (1 − n−3 , 0) for n = 1, 2, · · · . Then E is a
not finitely generated invariant subspace of the bi-shift, i.e., there exists no finite set
f1 , f2 , · · · , fn ∈ H 2 (D2 ) such that E = [f1 , f2 , · · · , fn ]M(z1 ,z2 ) .
We also have the next result.
Theorem 1.3 ([13, Theorem 3.6]) There exists a nontrivial function f ∈ H 2 (D2 )
such that [f ]M(z1 ,z2 ) A (z1 [f ]M(z1 ,z2 ) + z2 [f ]M(z1 ,z2 ) ) does not generate [f ]. The
402 H. Ezzahraoui et al.
E = [E A Mz E]Mz .
Motivated by the previous discussion, the next definition has been introduced in
several papers.
Definition 1.4 We shall say that an operator T ∈ L(H) admits Wold-type decom-
position, if R ∞ (T ) is closed and,
(i) R ∞ (T ) is reducing for T for which the restriction operator T|R ∞ (T ) is unitary.
(ii) H = [H A T H]T ⊕ R ∞ (T ).
Definition 1.5 An operator T ∈ L(H) is said to have the Wandering subspace
property if H = [H A T H]T and we say the Beurling-type theorem holds for T if
M = [M A T M]T for every M ∈ Lat (T , H).
It is clear that if Beurling-type theorem holds for T , it will follow that T admits
the Wandering subspace property. Also for a pure operator, T has Wandering space
property if and only if it admits Wold decomposition. From the preceding remarks
the Hardy shift, the Bergman shift and the Dirichlet shift satisfy the Beurling-type
theorem. We discuss below the contributions of several authors that have been inter-
ested in the class of operators satisfying the Beurling-type theorem. The problem of
describing all weighted shifts that satisfy Beurling-type theorem remains open.
The case of left invertible operators has been widely studied in the two last
decades. It is always assumed that T satisfies some operator inequalities close to
isometries. A pioneer result goes to S. Richter (see [20]), that provides a sufficient
condition on an operator S ∈ L(H) to admit the Wandering subspace property. More
precisely,
Theorem 1.6 ([20, Theorem 1]) Let S ∈ L(H) be pure such that
If M ∈ Lat (S, H), then there exists a wandering subspace B for S such that
N
M= S n B.
n≥0
In particular, Richter’s result states that the Dirichlet shift satisfies the Wandering
subspace property.
On Wold Type Decomposition for Closed Range Operators 403
we will get 1 ≤ c.
404 H. Ezzahraoui et al.
In 2009, S. Sun and D. Zheng (see [25]) gave another proof of the Beurling-type
theorem by proving some new identities in the Bergman space and later, K. J. Izuchi,
K. H. Izuchi and Y. I. Izuchi used these ideas in [14] to prove the next theorem.
Theorem 1.11 ([14, Theorem 1.1]) Let T ∈ L(H) . If T satisfies the following
conditions:
(i) T x2 + T ∗2 T x2 ≤ 2T ∗ T x2 for all x ∈ H;
(ii) T is bounded below;
(iii) T ≤ 1;
(iv) T ∗k x −→ 0 as k −→ ∞ for every x ∈ H.
Then H = [H A T H]T
The main purpose of this survey is to present the abstract approach to the problem.
We extend the previous results to the more general class of regular operators
introduced by M. Mbekhta in [17] and developed in [8].
We devote Sect. 2 to some well known properties of regular operators and the
basic tools of this class of operators. Section 3 is focused on the generalization of
the previous results to the class of regular operators. More precisely, we give under
the same conditions on orbits, an extension of Wold-type decomposition for regular
operators. See Theorem 3.9).
Section 4 is devoted to the duality between a bi-regular operator T and its Cauchy
dual ω(T ). This duality is reflected in terms of extended Wold-type decomposition.
Some applications and examples are widely given. In particular, we apply our results
to regular bilateral weighted shifts.
2 Regular Operators
operator T † is called the Moore-Penrose inverse of T , and has been widely studied
in the literature. It is usually defined as the unique solution of the following four
operator equations:
T T †T = T , T †T T † = T †, (T T † )∗ = T T † , (T † T )∗ = T † T . (3)
1
T † = .
γ (T )
It is clear that T is regular if and only if T ∗ is regular, that all injective operators
with closed range and that all surjective operators are regular. We give next some
classical known facts on regular operators. We refer to the corresponding books and
papers for proofs and further information.
Following Saphar [22], the algebraic core C(T ) of T , is the greatest subspace M
of H for which T (M) = M. In terms of sequences, we have
Proposition 2.3 ([7, Proposition 1]) Let T ∈ L(H) be a regular operator. We have
R ∞ (T ) = T (R ∞ (T )) = C(T ).
(i) R ∞ (T ) is closed,
(ii) If R ∞ (T ) = {0}, then T is left invertible,
For further information, We refer to [1], [3] and [15].
The next proposition is given in [3] and will be useful in the sequel,
Proposition 2.5 ([7, Proposition 3]) Let T ∈ L(H) be regular. If S is such that
T ST = T , then
T n S n T n = T n for every n ≥ 1.
x = x − SnT nx
= n−1 k k
k=0 S PN(T ) T x.
Since PN(T ) T k x ∈ N(T ) and by our assumption S k N(T ) ⊆ R(T ) for every k ≥ 0,
we obtain x ∈ R(T ) and hence N(T n ) ⊆ R(T ).
is bijective. In particular,
is one to one.
(iii) R(T )⊥ 4 R(T n ) ∩ R(T n+1 )⊥ for every n ≥ 0.
(bij ect ion)
Corollary 2.12 For T ∈ L(H) regular and for every integer n ≥ 1, we have
We also have
Corollary 2.13 Let T ∈ L(H) be regular such that dim R(T )⊥ = 1. Then there is
a sequence of orthogonal wandering vectors {ek }k≥1 such that
J
H= C.ek ⊕ R ∞ (T ).
k≥1
Proof By corollary 2.12, dim R(T n ) ∩ R(T n+1 )⊥ = 1 for every n ≥ 0. Let en be
a nonzero vector in R(T n−1 ) ∩ R(T n )⊥ for n ≥ 1. Since
In the sequel, consider T ∈ L(H) such that γ (T ) ≥ 1. Fix the next notations
DT = (T ∗ T − T † T )1/2 .
(ii)
Proof We notice first that since γ (T ) > 0, the operator T † exists. Suppose that the
inequality (4) holds and let x ∈ N(T )⊥ . Since N(T )⊥ = R(T † ), we get T † T x =
PR(T † ) x = x and so we have the result. Conversely, let x ∈ H T † T x ∈ N(T )⊥ . By
substituting T † T x ∈ N(T )⊥ for x in (5) and by using the identity T T † T = T we
obtain (4).
In the case of expansive operators, these inequalities are equivalent to Inequality (2)
introduced in [19].
We extend next some known results of left invertible operators to our setting.
Lemma 3.2 ([7, Lemma 1]) Let T ∈ L(H) be such that γ (T ) ≥ 1 and let n ≥ 1
be an integer. For every x ∈ H, we have
n−1
n
x =
2
PE (T ) x + (T ) x +
† i 2 † n 2
DT (T † )i x2 , (6)
i=0 i=1
where PE = I − T T † .
Lemma 3.3 Let T ∈ L(H) be with closed range and n ≥ 1. We have,
(i) x − T n (T † )n x = n−1
i=0 T PE (T ) x for every x ∈ H;
i † i
(ii) x − (T † )n T n x = n−1
i=0 (T ) PE † T x for every x ∈ H.
† i i
Proof We have:
n−1
I − T n (T † )n = T i (T † )i − T i+1 (T † )i+1
i=0
n−1
= T i (I − T T † )(T † )i
i=0
n−1
= T i PE (T † )i .
i=0
[E]T ⊂ R ∞ (T )⊥ .
H = [E]T + R ∞ (T )
with E = H A T H.
412 H. Ezzahraoui et al.
We will show next that under the assumptions of Theorem 3.8, we have H =
[E]T ⊕ R ∞ (T ). We collect next some additional results provided by the same
assumptions.
Theorem 3.9 ([7, Theorem 8]) Let T ∈ L(H) be a regular operator with γ (T ) ≥
1. Under the assumptions of Theorem 3.8, the following assertions hold.
(i) The subspace R ∞ (T ) is reducing for T ,
†
(T|R ∞ (T ) )† = T|R ∗ ∗
(ii) ∞ (T ) = T|R ∞ (T ) = (T|R ∞ (T ) ) ,
(iii) T † is regular,
(iv) the restriction
is a unitary operator,
(v) H has an orthogonal decomposition
H = [E]T ⊕ R ∞ (T ).
H = [E † ]T † ⊕ R ∞ (T † ).
We derive that:
Corollary 3.11 ([7, Corollary 8]) Let T ∈ L(H). Under the assumptions of
Theorem 3.8, we have
T †n = T n† on R ∞ (T ) ∩ R ∞ (T ∗ ).
H = [E]T ⊕ R ∞ (T ),
On Wold Type Decomposition for Closed Range Operators 413
T|R ∞ (T ) : R ∞ (T ) → R ∞ (T )
We apply our results to non necessarily left invertible weighted shift operators. To
this purpose, we assume that H is a Hilbert space and (en )n∈Z is an orthonormal
basis of H.
Let (ωn )n∈Z be a bounded sequence and Sω : H −→ H be the bilateral weighted
bounded shift defined by Sω en = ωn en+1 . It well known that
Sω is one to one if and
only if ωn = 0 for every n. Indeed, let x ∈ H such that x = n∈Z xn en . We have
T x2 = |xn |2 |ωn |2 ≥ inf |ωn |2 x2 .
n∈Z
n∈Z
ωk2 .ωk+1
2 2
· · · ωk+n−1 − 1 ≤ cn (ωk2 − 1) ∀k ∈ Z∗ ; n ≥ 2. (8)
Now, Let x ∈ H such that x = k∈Z ak ek . Clearly, we have
Sωn x2 = |ak |2 ωk2 · · · ωk+n−1
2
, n≥1
k =0
and
Sω† Sω x2 = |ak |2 .
k =0
(a) Since Sω† x2 = k =1 ω2 |ak |
1 2 and ωk ≥ 1 for all integers k = 0, we
k−1
conclude that Sω† is a contraction (and so γ (Sω ) ≥ 1). Now, from Inequality (8),
we get
where c = 1.
(b) From Theorem 3.9, we conclude that Sω† is a regular operator.
(c) By Theorem 3.9, the subspace R ∞ (Sω ) is reducing for Sω and the restriction
of Sω on R ∞ (Sω ) ∩ N(S ⊥
Iω ) 6is a7 unitary operator. On the other
I hand, clearly
we have R ∞ (Sω ) = e and since N(S ) ⊥ =
k =0 {ek }, we get
I j ≤06 7
j ω
∞ ⊥ I
R (Sω ) ∩ N(Sω ) = j <0 ej and so Sω | j<0 {ej } is unitary.
Remark 3.14 If we assume that ωk = 1 for all k = 0 and ω0 = 0, then we have
Sω† = Sω∗ , in this case; Sω is called a partial isometry.
Example ([7]) Let H be a Hilbert space and (en )n∈Z an orthonormal basis of H.
Let Sω : H −→ H be a bilateral shift defined by Sω (en ) = ωn en+1 where (ωk )k∈Z
is defined by:
⎧
⎪
⎨1 fork < 0,
ωk = 0 for k = 0,
⎪
⎩ k+1 for k ≥ 1.
k
On Wold Type Decomposition for Closed Range Operators 415
√
The sequence (ωk )k∈Z is bounded. More precisely, we have 1 ≤ ωk ≤ 2 for all
k = 0. Take cn = n. Clearly, we have
1 1
= = ∞.
cn n
n≥2 n≥2
ωk2 .ωk+1
2
· · · ωk+n−1
2
− 1 ≤ cn (ωk2 − 1).
For k ≥ 1, we have
ωk2 .ωk+1
2
· · · ωk+n−1
2
− 1 ≤ cn (ωk2 − 1).
Thus, the inequality (8) is satisfied. Consequently, the operator Sω satisfies all
properties of Proposition 3.13.
operator Tα = α(S ∗
I ⊗ S)P , with S is the bilateral shift and P is the orthogonal
projection onto {ei ⊗ ej : i, j ∈ Z, i = j }, defined on H by
I implies that if i = j we have ai,j = 0. Thus, x =
This i∈Z ai,i ei,i ∈
{e
i∈Z i,i }.
Since Tα ei+1,i−1 = α(1 − δi+1,i−1 )ei,i = αei,i and Tα ei+2,i−2 = αei+1,i−1 ,
an induction argument shows that
1 n
ei,i = T ei+n,i−n for every n ≥ 1.
αn α
I
Hence N(Tα ) ⊆ R ∞ (Tα ). On the other hand, since R(Tα ) = j =i+2 {ei,j },
then R(Tα ) is closed. Therefore, Tα is regular.
Now, simple computations shows that
and thus
N6 7
N(Tα∗ ) = ei,i+2 .
i∈Z
1 ∗ 1
Tα† ei,j , em,n = Tα† ei,j , T em−1,n+1 = Tα Tα† ei,j , em−1,n+1 .
α α α
On Wold Type Decomposition for Closed Range Operators 417
Since j = i + 2, we have ei,j ∈ R(Tα ) and thus Tα Tα† ei,j = PR(Tα ) ei,j =
ei,j . This implies Tα† ei,j , em,n = α1 ei,j , em−1,n+1 = α1 ei+1,j −1 , em,n , and
finally
1 1
Tα† ei,j = (1 − δi+2,j )ei+1,j −1 = 2 T ∗ ei,j . (10)
α α
The Cauchy dual of an operator T ∈ L(H) with closed range is introduced in [10]
by the next formula
ω(T ) := T †∗ .
R ∞ (ω(T )) ⊆ R ∞ (T ).
(ii)
1
Q(T x + T † T y)2 ≤ (1 + )(x2 + cT y2 ), ∀x, y ∈ H, (12)
c
where Q is the orthogonal projection of H onto R(T ). In particular, if T satisfies (1),
then T is bounded below and ω(T ) is concave.
On Wold Type Decomposition for Closed Range Operators 419
Definition 4.8 We shall say that an operator T ∈ L(H) admits the extended Wold-
type decomposition if
(i) R ∞ (T ) is closed and reduces T ,
(ii) T|R ∞ (T )∩N(T )⊥ : R ∞ (T ) ∩ N(T )⊥ −→ R ∞ (T ) is unitary,
(iii) H = [E]T ⊕ R ∞ (T ).
Notice in passing that if T satisfies the extended Wold-type decomposition and
if T is one to one or T|R ∞ (T ) is unitary, then T admits the classical Wold type
decomposition property investigated in [24].
For example, for a regular operator T and under the assumptions of Theorem 3.8,
by Theorem 3.9, T admits the extended Wold-type decomposition.
Remark 4.9
(i) Corollary 3.5 implies that a regular operator T is pure if and only if ω(T ) has
the wandering subspace property.
(ii) If T is an operator with closed range, then R(ω(T )) is closed. So, since
ω((ω(T )) = T and from (i) in Corollary 4.6 we have
R ∞ (T )⊥ ⊆ [E]ω(T ) .
It is not clear whether the inclusion in Corollary 3.5, (i) can be replaced by the
equality if T is an arbitrary regular operator. However, in the following example, we
show that the equality holds if T is a regular weighted shift on l 2 (Z).
Example Let H be a Hilbert space endowed with an orthonormal basis (en )n∈Z and
let α = (α)n∈Z be a bounded sequence. The weighted shift Sα on H associated with
α is the bounded linear operator Sα : H −→ H defined by Sα en = αn en+1 . It is
proved in [8] that if Sα is regular, then there exits at most n0 ∈ Z such that αn0 = 0.
Let Sα be a regular bilateral weighted shift such that α0 = 0. It is easy to see that
−1
n α
(Πi=1 / {1, · · · , n} ;
k−i ) ek−n , if k ∈
(Sᆠ)n ek =
0 , if k ∈ {1, · · · , n} .
420 H. Ezzahraoui et al.
I 6 7
We derive that N((Sα† )n ) = 1≤j ≤n ej for every n ≥ 1, and in particular, we
have
Since i + 1 ∈ {1, · · · , n}, we get (Sα† )n Sαi e1 = 0, and thus Sαi e1 ∈ N((Sα† )n ) for
I 6 i 7
every 0 ≤ i ≤ n − 1. Hence n−1 i=0 Sα e1I ⊂ N((S
† )n )). On the other hand, from
6 α7
[8, Proposition 8], we have N((Sα† )n ) ⊂ n−1 i
i=0 Sα e1 , and thus
N4
n−1 5
N((Sα† )n ) = Sαi e1 .
i=0
Therefore
N
N((Sα† )n ) = [e1 ]Sα .
n≥0
I 6 7 I 6 7
It is easy to check that j ≤0 ej = R ∞ (Sα ) = R ∞ (ω(Sα )) and that j ≥1 ej =
[e1 ]Sα = R ∞ (Sα )⊥ .
Proposition 4.10 If T ∈ L(H) is a bi-regular operator, then
In particular,
H = [E]T ⊕ R ∞ (T ).
(ii) ∞
∗
PR (T ) T T = PR (T ) T T ;
∞
†
(iii) T ∗ PR ∞ (T ) = T † PR ∞ (T ) ;
(iv) T ∗ T PR ∞ (T ) = T † T PR ∞ (T ) ;
(v) T|R ∞ (T )∩N(T )⊥ : R ∞ (T ) ∩ N(T )⊥ −→ R ∞ (T ) is unitary.
(i) T is bi-regular;
(ii) R ∞ (T † ) = R ∞ (T ∗ ).
Proof
(i) Follows from Remark 4.14.
(ii) From T is bi-regular, we obtain R ∞ (T † ) and R ∞ (T ∗ ) are closed. By Propo-
sition 4.10 and since ω(T ∗ ) = T † , we have R ∞ (T † )⊥ = [E † ]T ∗ and
R ∞ (T ∗ )⊥ = [E † ]T † . On the other hand, because E † = N(T ) ⊆ R ∞ (T ),
we get T ∗n x = T †n x for every x ∈ E † and for every n ≥ 0. Then we have
[E † ]T † = [E † ]T ∗ , which proves that R ∞ (T † ) = R ∞ (T ∗ ).
422 H. Ezzahraoui et al.
Since T is regular if and only if its adjoint T ∗ is regular, then it is clear from
Proposition 4.15 that if T is regular and one of the conditions (i)–(v) in Remark 4.14
is satisfied for T and for T ∗ , then R ∞ (ω(T )) = R ∞ (T ). So, by Proposition 4.10,
we get the following result.
Corollary 4.16 Let T ∈ L(H) be regular. If one of the conditions (i)–(v) in
Remark 4.14 is fulfilled for T and for T ∗ , then T and T ∗ admit the extended Wold-
type decomposition.
The duality between T and ω(T ) is reflected in terms of extended Wold-type
decomposition as follows.
Proposition 4.17 Let T ∈ L(H) be bi-regular. Then T admits the extended Wold-
type decomposition if and only if ω(T ) admits it. In this case, we have
T ∗ x = T † x.
ω(T )∗ x = T † x = T ∗ x = x, ∀x ∈ R ∞ (T ).
I
(i)#⇒(vii): Let A = ∞ ∞
i=1 T E and B = R (ω(T )). It follows from (i) and
i
(ii) that A, B = {0} and by Proposition 4.10 we have A ⊥ B. Since R(T ) is closed
and E ⊥ T i E ∀i ≥ 1, by Proposition 4.10 again we obtain
∞
N
R(T ) = E ⊥ = ([E]T ⊕ R ∞ (ω(T ))) A E = T i E ⊕ R ∞ (ω(T )) = A ⊕ B.
i=1
I I∞ i I
Since T is continuous, T A = T ∞ i=1 T E ⊆
i
I
i=2 T E ⊆ ∞ i=1 T E = A. On
i
∞
the other hand, E ⊥ T E ∀i ≥ 1, thus A ⊕ E = i=1 T E ⊕ E = [E]T . So, by
i i
T x = a + b = a − a2 + T b1 , and then
T x = a1 + T b1 , (15)
and
1
n
lim inf ω−k−1 · · · ω−k−n > 0.
n→+∞ k≥0
426 H. Ezzahraoui et al.
(Sω† )n = (Sωn )† .
On the other hand, we provide provides an example disapproving the equality. For
1 −1 1 0
T = we have T † = 12 and (T † )2 T 2 = 12 T † T which is not a
0 0 −1 0
projection. Thus (T † )2 = (T 2 )† .
The equality (T n )† = (T † )n may fail even for left invertible operators as shown
by the examples below.
Recall that for T left invertible, we have T † = (T ∗ T )−1 T ∗ . Since T n is also left
invertible, we get T n† = (T n∗ T n )−1 T ∗n and T †n = ((T ∗ T )−1 T ∗ )n .
Problem 5.2 When is the restriction of a bi-regular operator is bi-regular?
As for regular operators the restriction to the kernel is not bi-regular. It is
interesting to see for which condition on the invariant subspace M of T , T|M is
bi-regular. For example, if T is regular and E is an invariant subspace such that
N ∞ (T ) ⊂ E, then T|E is regular. Is this fact true for bi-regular operators?
Problem 5.3 Is every regular operator bi-regular?
In contrast with T is regular if and only if T ∗ is regular. We do not know if T is
regular operator if and only if T † is regular. A positive answer is given when T is
regular such that (T † )n = (T n )† every n ≥ 1. In particular, regular weighted shifts
are bi-regular.
References
1. P. Aiena, Fredholm and Local Spectral Theory with Applications to Multipliers (Kluwer,
Dordecht, 2004)
2. A. Aleman, S. Richter, C. Sundberg, Beurling’s theorem for the Bergman space. Acta Math.
177(2), 275–310 (1996)
3. C. Badea, M. Mbekhta, Operators similar to partial isometries. Acta Sci. 71, 663–680 (2005)
4. A. Beurling, On two problems concerning linear transformations in Hilbert space. Acta Math.
81, 239–255 (1949)
5. S. Chavan, On operators Cauchy dual to 2-hyperexpansive operators. Proc. Edinburgh Math.
Soc. 50, 637–552 (2007)
6. N.C. Dincic, D.S. Djordjevic, Basic reverse order law and its equivalencies. Aequationes Math.
85(3), 505–517 (2013)
On Wold Type Decomposition for Closed Range Operators 427
Abstract Multiplication operators on the space L2 (T) on the unit circle T with
Lebesgue measure are classical operators. So are Toeplitz operators on the Hardy
space H 2 ⊂ L2 (T). Sarason’s paper (Oper Metrices 1:491–526, 2007) has started
investigations of truncated Toeplitz operators (TTO), i.e., compressions of these
multiplication operators to model spaces. If operators act between two different
model spaces they are called asymmetric truncated Toeplitz operators (ATTO).
Naturally the compressions of multiplication operators between orthogonal comple-
ments of model spaces can be investigated. They are called dual truncated Toeplitz
operators (DTTO), or asymmetric dual truncated Toeplitz operators (ADTTO) if
orthogonal complements to different model spaces are considered. In this chapter
the properties of ADTTO are presented.
TTO and ATTO are natural generalizations of Toeplitz matrices which appear in
many contexts, such as in the study of finite–interval convolution equations, signal
processing, control theory, probability and diffraction problems [10, 11, 19]. Model
spaces, which provide the natural setting for TTO and ATTO, have generated
enormous interest and they are relevant in connection with a variety of topics
such as the Schrödinger operator, classical extremal problems in control theory,
M. C. Câmara
Center for Mathematical Analysis, Geometry and Dynamical Systems Mathematics Department,
Instituto Superior Técnico, Universidade de Lisboa Av. Rovisco Pais, Lisboa, Portugal
e-mail: [email protected]
K. Kliś-Garlicka · M. Ptak ()
Department of Applied Mathematics, University of Agriculture, Kraków, Poland
e-mail: [email protected]; [email protected]
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 429
R. M. Aron et al. (eds.), Operator and Norm Inequalities and Related Topics,
Trends in Mathematics, https://doi.org/10.1007/978-3-031-02104-6_13
430 M. C. Câmara et al.
Hankel operators and Toeplitz matrices (see for instance [12] and [10]). Natural
conjugations, which model spaces and the whole L2 (T) possess (see [3]), make
model spaces even more natural in the context of physics [11]. Their orthogonal
complements in L2 (T) also appear in numerous applications. In the equivalent
setting of the real line [9, 14], using time and frequency as the natural variables,
and taking the inner function θ = θλ with θλ = exp(iλξ ) for ξ ∈ R, they appear
via the Fourier transform, for instance, as high frequency signals, which are of
decisive importance in electronics, or as outputs of high–pass filters. DTTO and
ADTTO, acting on these spaces have realizations, for example, in long distance
communication links with several regenerators along the path that cancel low–
frequency noise using high–pass filters, or in the description of wave propagation in
the presence of finite–length obstacles. The chapter is mostly based on the papers
[4–7, 18].
Let D = {z ∈ C : |z| < 1} be the unit disk and T the unit circle. Denote by
L2 (T) := L2 (∂D) the space of measurable and square integrable functions on T
with respect to the normalized Lebesgue measure, by H 2 denote the classical Hardy
space, and let H−2 = L2 (T) A H 2 . Let P + be the orthogonal projection from L2 (T)
onto H 2 and P − = IL2 (T) − P + .
Let ϕ ∈ L∞ (T). Recall that the multiplication operator on L2 (T) is defined as
Mϕ f = ϕ f for f ∈ L2 (T). The Toeplitz operator is defined by Tϕ f = P + (ϕf ),
and the operator Hϕ f = P − (ϕf ), for f ∈ H 2 , is called the Hankel operator (with
symbol ϕ). The space of all Toeplitz operators is denoted by T (H 2 ) and space of
all Hankel operators is denoted H(H 2 , H−2 ).
Recall that the operator J , defined by J : L2 (T) → L2 (T), Jf (z) = z̄f (z),
z ∈ T, is an antilinear involution. Moreover, J −1 = J = J (by we denote the
antilinear adjoint) and J (H 2 ) = H−2 , J (H−2 ) = H 2 . More properties of antilinear
operators can be found for example in [17]. Let θ be an inner function. Note that
the multiplication operator Mθ maps H 2 bijectively onto θ H 2 and Mθ−1 = Mθ̄ .
Moreover, each of the operators J , Mθ and Mθ̄ preserves L∞ (T).
The following properties can be easily verified.
Proposition 1.1 Let θ be an inner function. Then
(a) f1 , f2 = θf1 , θf2 = Jf2 , Jf1 for f1 , f2 ∈ L2 (T);
(b) PθH 2 = Mθ P + Mθ̄ ;
(c) P − = J P +J ;
(d) Mθ (f1 ⊗ f2 )Mθ̄ = θf1 ⊗ θf2 for f1 , f2 ∈ L2 (T);
(e) J (f1 ⊗ f2 )J = Jf1 ⊗ Jf2 for f1 , f2 ∈ L2 (T);
(f) Mθ J Mθ = J ;
(g) J Mϕ = Mϕ̄ J for ϕ ∈ L∞ (T).
In particular J 1 = z̄ and J (1 ⊗ 1)J = z̄ ⊗ z̄. Here, for f1 , f2 ∈ L2 (T), f1 ⊗ f2
denotes the operator defined on L2 (T) by (f1 ⊗ f2 )(f ) = f, f2 f1 ,
For a nonconstant inner function θ denote by Kθ the model space defined
as the orthogonal complement of θ H 2 in H 2 , i.e., Kθ = H 2 A θ H 2, and let
Pθ : L2 (T) → Kθ be the orthogonal projection and let Pθ⊥ = IL2 (T) − Pθ be the
(Asymmetric) Dual Truncated Toeplitz Operators 431
orthogonal projection from L2 (T) onto (Kθ )⊥ . Hence we have following natural
decompositions
Cθ Mϕ Cθ = Mϕ̄ . (2)
Define also
Bϕθ,α = Pα⊥ Mϕ|Kθ ∩L∞ (T) , Dϕθ,α = Pα⊥ Mϕ|K ⊥ ∩L∞ (T) .
θ
Hence, according to the decomposition (3), the action of the operator Mϕ is given
by the matrix
" # " #
α,θ ∗
Pα Mϕ|Kθ ∩L∞ (T) Pα Mϕ|K ⊥ ∩L∞ (T) g Aθ,α
ϕ (Bϕ̄ ) g
θ = . (4)
Pα⊥ Mϕ|Kθ ∩L∞ (T) Pα⊥ Mϕ|K ⊥ ∩L∞ (T) h Bϕθ,α Dϕθ,α h
θ
If Aθ,α
ϕ extends to the whole Kθ as a bounded operator, it is called an asymmetric
truncated Toeplitz operator (ATTO). Similarly, if Bϕθ,α extends to a bounded
operator from Kθ to Kα⊥ , it is called an asymmetric big truncated Hankel operator
(ATHO) (see [13, 15]), and if Dϕθ,α extends to the whole Kθ⊥ as a bounded operator,
it is called an asymmetric dual truncated Toeplitz operator (ADTTO).
Let us fix the notation
T (Kθ , Kα ) = {Aθ,α
ϕ : ϕ ∈ L (T) and Aϕ is bounded},
2 θ,α
In case θ = α we will use the shorter notation Aθϕ , Bϕθ , Dϕθ and T (Kθ ) and
T (Kθ , Kθ⊥ ), T (Kθ⊥ ), respectively.
The following basic properties of restrictions of multiplication operators hold.
⊥ ⊥ ⊥
ϕ ∈ T (Kθ , Kα ), Bϕ ∈ T (Kθ , Kα ), Dϕ ∈ T (Kθ , Kα ).
Lemma 2.2 Let Aθ,α θ,α θ,α
Proposition 2.3 Let α, θ be inner functions and ϕ ∈ L2 (T). Assume that Aθ,α
ϕ ∈
⊥ ⊥ ⊥
T (Kθ , Kα ), Bϕ ∈ T (Kθ , Kα ), Dϕ ∈ T (Kθ , Kα ). Then
θ,α θ,α
θ,α
ϕ = Aα ϕ̄θ̄ Cθ ;
(a) Cα Aθ,α
(b) Cα Dϕθ,α = Dαθ,α C ;
ϕ̄ θ̄ θ
(c) Cα Bϕθ,α = Bαθ,α C .
ϕ̄ θ̄ θ
Now, since Kθ and Kθ⊥ are invariant for Cθ (the same holds for α), for g ∈ Kθ ∩
L∞ (T), h ∈ Kθ⊥ ∩ L∞ (T) using matrix representation (4) we get
" #" #
α,θ ∗
Cα |Kα 0 Aθ,α
ϕ (Bϕ̄ ) g
0 Cα |Kα⊥ Bϕθ,α Dϕθ,α h
" θ,α α,θ ∗ # " #
Aα ϕ̄ θ̄ (Bᾱϕθ ) Cθ |Kθ 0 g
= θ,α θ,α .
Bα ϕ̄ θ̄ Dα ϕ̄ θ̄ 0 C θ |K ⊥ h
θ
Hence
θ,α
ϕ g = Aα ϕ̄ θ̄ Cθ g for g ∈ Kθ ∩ L∞ (T);
(a) Cα Aθ,α
(b) Cα Dϕθ,α h = Dαθ,α C h for h ∈ Kθ⊥ ∩ L∞ (T);
ϕ̄ θ̄ θ
(c) Cα (Bϕ̄α,θ )∗ h = (Bᾱϕθ
α,θ ∗
) Cθ h for h ∈ Kθ⊥ ∩ L∞ (T);
(d) Cα Bϕθ,α g = Bαθ,α C g for g ∈ Kθ ∩ L∞ (T).
ϕ̄ θ̄ θ
Corollary 2.4 Let θ be an inner function and ϕ ∈ L2 (T).
Assume that Aθϕ ∈
θ ⊥
T (Kθ ), Dϕ ∈ T (Kθ ). Then Aϕ and Dϕ are Cθ –symmetric, i.e., Cθ Aθϕ =
θ θ Aθϕ̄ Cθ
and Cθ Dϕθ = Dϕ̄θ Cθ .
The following properties of the operators Bzθ and Bz̄θ can be verified.
Lemma 2.5 Let θ be an inner function. Then
(a) Bzθ = θ ⊗ 1 k0θ ;
(b) (Bzθ )∗ = 1k0θ ⊗ θ ;
(c) Bz̄ = z̄ ⊗ k0θ ;
θ
Since
we have
Dzθ Dz̄θ f, g = f, g − f, θ zk̃0θ , g = f, g − f, θ Pθ⊥ (zk̃0θ ), g.
Calculating (c) we will use Lemma 2.5 and the formula for multiplication of rank-
one operators (see [17])
Bzθ (Bz̄θ )∗ = (θ ⊗ 1
k0θ )(k0θ ⊗ z̄) = k0θ , k̃0θ θ ⊗ z̄ = θ (0) θ ⊗ z̄.
It was shown that an asymmetric truncated Toeplitz operator can be bounded even if
it has no bounded symbol. The same is true for an asymmetric big truncated Hankel
operator. In the case of bounded dual asymmetric Toeplitz operators the symbol is
always bounded and unique.
Proposition 3.1 Let ϕ ∈ L2 (T). Then Dϕθ,α is bounded if and only if ϕ ∈ L∞ (T).
Moreover, in that case, Dϕθ,α = ϕ∞ .
Proof Let f ∈ H ∞ . Then θf ∈ θ H ∞ ⊂ θ H 2 and
Dϕθ,α (θf )2 =(P − + αP + ᾱ)(θf )2 = P − (θf )2 + αP + ᾱθf 2
If the operator Dϕθ,α is bounded, then there is a constant C > 0 such that
Dϕθ,α (θf ) Cf . Hence Tᾱϕθ f Cf for every f ∈ H ∞ , which implies
that Tᾱϕθ is bounded and in consequence ᾱϕθ ∈ L∞ (T). Thus ϕ ∈ L∞ (T) and
ϕ∞ = Tᾱϕθ Dϕθ,α .
If now ϕ ∈ L∞ (T), then for any f ∈ Kθ⊥ we have
An important consequence of Proposition 3.2 is that the only symbol for the zero
asymmetric dual truncated Toeplitz operator is ϕ = 0. What follows is that each
ADTTO has a unique symbol.
Proposition 3.3 Let α, θ be inner functions and let ϕ ∈ L∞ (T). If Dϕθ,α is
invertible, then ϕ is invertible in L∞ (T).
Proof By Douglas [8, Corollary 4.24] we know that the operator Mϕ is invertible in
L2 (T) if and only if ϕ is invertible in L∞ (T). Assume that Dϕθ,α is invertible, then
there is a constant c > 0 such that
Mϕ zk g = ϕzk g = ϕθg Pα⊥ ϕθg = Dϕθ,α θg cθg = czk g.
Since the set {zk g : k ∈ Z, g ∈ H 2 } is dense in L2 (T), we have that for x ∈ L2 (T)
Mϕ x cx.
Since (Dϕθ,α )∗ = Dϕ̄α,θ is also invertible, thus Mϕ̄ x cf , we conclude that
Mϕ is invertible by Douglas [8, Corollary 4.9].
Since the unilateral shift S is unitarily equivalent to the Toeplitz operator Tz we are
able to describe the commutant of the unilateral shift as
Now considering the compressions Aθz and Dzθ , it is natural to try to describe
the commutants {Aθz } and {Dzθ } . In the more general asymmetric setting we are
searching for all operators intertwining Aθz and Aαz in the case of two model spaces
or searching for all operators intertwining Dzθ and Dzα in the case of orthogonal
complements of two model spaces, i.e., we try to describe the following sets of
operators
Thus we have
Clearly,
Bϕθ,α k0θ = Pα⊥ (ϕk0θ ) and Bϕ̄α,θ k̃0α = Pθ⊥ (ϕ̄ k̃0α ).
Thus we have
Clearly,
Bϕθ,α k0θ = Pα⊥ (ϕk0θ ) and Bϕ̄α,θ k̃0α = Pθ⊥ (ϕ̄ k̃0α ).
Theorem 4.2 Let α, θ be nonconstant inner functions and ϕ ∈ L∞ (T), ϕ = 0.
Then Dϕθ,α Dzθ = Dzα Dϕθ,α if and only if one of the following holds
(a) α(0) = 0 = θ (0) and ϕ ∈ α
gcd(α,θ) Kz·gcd(α,θ) , or
(b) α = θ and ϕ ∈ (k0θ )−1 Kzθ .
Proof Applying Theorem 4.1
Since
Since the functions k0α , k0θ are bounded from below and analytic (in consequence we
have (k0α )−1 , (k0θ )−1 ∈ H ∞ ), we get by (10) that ϕ ∈ H 2 . Let γ = gcd(α, θ ). Then,
by (11),
ϕk0α − cα = θ̄ αh = θ̄ α
γ̄ γ h ∈ H 2.
Hence h is divisible by θ
γ and since h ∈ Kθ = K θ ⊕ γθ Kγ , we have
γ
h= θ
γ h1 with h1 ∈ Kγ .
Therefore, by (11),
ᾱ α
γ̄ ϕk0 = cγ + h1 ∈ H 2 .
ϕ= α
γ ϕ1 with ϕ1 ∈ H 2 \ {0}.
ϕ1 k0θ = cγ + g1 , g1 ∈ Kγ , (12)
ϕ1 k0α = cγ + h1 , h1 ∈ Kγ . (13)
Moreover,
ϕ= α
γ ϕ1 ∈ α
γ (Kγ ⊕ Cγ ) = γ Kz·γ ;
α
440 M. C. Câmara et al.
2. if α = θ , then (12) and (13) become the same condition, equivalent to ϕk0θ ∈
Kθ ⊕ Cθ = Kzθ , which leads to ϕ ∈ (k0θ )−1 Kzθ .
Corollary 4.3 Let θ be a nonconstant inner function and let ϕ ∈ L∞ (T), ϕ = 0.
Then Dϕθ ∈ {Dzθ } if and only if ϕ ∈ (k0θ )−1 Kzθ .
Example Let a ∈ D; we denote by Ba the Blaschke factor with zero at a, i.e.,
Ba (z) = 1− āz . Let θ = zBb and α = zBa with a = b, a = 0, b = 0, a, b ∈ D.
a−z
Then gcd(θ, α) = z. By Theorem 4.2 for ϕ ∈ L∞ (T) we have Dϕθ,α ∈ I(Kθ⊥ , Kα⊥ )
if and only if
4 5
aa0 +(aa1 −a0 )z−a1 z2
ϕ ∈ Ba Kz2 = 1−āz : a0 , a1 ∈ D .
(a) Dϕθ,α − Dzα Dϕθ,α Dz̄θ = Pα⊥ (ϕ z̄k̃0θ ) ⊗ θ + α ⊗ Pθ⊥ (ϕ̄ z̄k̃0α );
(b) Dϕθ,α − Dz̄α Dϕθ,α Dzθ = Pα⊥ (ϕ z̄k0θ ) ⊗ z̄ + z̄ ⊗ Pθ⊥ (ϕ̄ z̄k0α ).
Proof By Theorem 4.1 we have
Dϕθ,α Dzθ Dz̄θ − Dzα Dϕθ,α Dz̄θ = α ⊗ Dzθ Pθ⊥ (ϕ̄ k̃0α ) − Pα⊥ (ϕk0θ ) ⊗ Dzθ z̄.
Dzθ Pθ⊥ (ϕ̄ k̃0α ) = Pθ⊥ (Dzθ (ϕ̄ z̄(α − α(0))) = Pθ⊥ (ϕ̄(α − α(0))) = Pθ⊥ (ϕ̄ z̄k̃0α ).
To prove (b) write (a) for α ϕ̄ θ̄ ∈ L∞ (T) and apply Cα and Cθ , then
Corollary 5.2 Let Dϕθ ∈T (Kθ⊥ ). Then
(a) Dϕθ − Dzθ Dϕθ Dz̄θ = Pθ⊥ (ϕ z̄k̃0θ ) ⊗ θ + θ ⊗ Pθ⊥ (ϕ̄ z̄k̃0θ );
(b) Dϕθ − Dz̄θ Dϕθ Dzθ = Pθ⊥ (ϕ z̄k0θ ) ⊗ z̄ + z̄ ⊗ Pθ⊥ (ϕ̄ z̄k0θ ).
1−|a|2 |a|2 −1
Example Let θ = Ba with a ∈ D, then k0θ = 1−āz and k̃0θ = 1−āz . Consider
ϕ ≡ 1, then
Pθ⊥ (ϕ z̄k̃0θ ) = Pθ⊥ (z̄k̃0θ ) = Pθ⊥ (|a|2 − 1)(z̄ + ā 1−1āz ) = (|a|2 − 1)z̄.
Thus
D1θ − Dzθ D1θ Dz̄θ = Pθ⊥ (z̄k̃0θ ) ⊗ θ + θ ⊗ Pθ⊥ (z̄k0θ ) = (|a|2 − 1)(θ ⊗ z̄ + z̄ ⊗ θ ).
442 M. C. Câmara et al.
and
Pθ⊥ (ϕ̄ z̄k̃0θ ) = (|a|2 − 1)Pθ⊥ z̄2 (1 − a z̄) 1−1āz = (|a|2 − 1) p(z̄),
6 A Characterization of ADTTO
for some ψ, χ ∈ Kθ . In other words, the left hand side of (15) can be expressed as
an operator of rank at most two. In this section our aim is to give similar expressions
for operators from T (Kθ⊥ , Kα⊥ ) using operators of rank at most two.
The characterization (15) proved in [19] for truncated Toeplitz operators imme-
diately gives a symbol of the truncated Toeplitz operator A = Aθψ+χ̄ . Moreover, the
relation between ψ and χ is simple, see [19, Corollary after Theorem 3.1]. However,
for any asymmetric dual truncated Toeplitz operator, the functions μ = Pα⊥ (ϕ z̄k̃0θ ),
ν = Pθ⊥ (ϕ̄ z̄k̃0α ) in the formula (a) in Proposition 5.1, strongly and in a very
complicated way depend on each other. Moreover, in case of dual truncated Toeplitz
operators, having the rank-two operator on the right hand side of (5.1), μ⊗θ +α ⊗ν
with μ, ν ∈ Kθ⊥ , we are far from obtaining the symbol of D. For this reason, to
answer a natural question when an operator D ∈ B(Kθ⊥ , Kα⊥ ) is a ADTTO and to
(Asymmetric) Dual Truncated Toeplitz Operators 443
find its symbol, we will consider the matrix decomposition of D. This will be done
in Theorem 6.7.
First the compressions of ADTTO’s to certain subspaces of Kθ⊥ and Kα⊥ will
be considered. Let θ, α be two inner functions. Using the decompositions Kθ⊥ =
θ H 2 ⊕ H−2 and Kα⊥ = αH 2 ⊕ H−2 one can write each operator D ∈ B(Kθ⊥ , Kα⊥ ) as
a matrix
" #
PαH 2 D|θH 2 PαH 2 D|H 2
D= −
.
P − D|θH 2 P − D|H 2
−
where
and
Let us denote
(b) Γˆϕ|θH
θ −
∞ = P Mϕ|θH ∞ = Hϕθ Mθ̄|θH ∞ ;
and
From Proposition 6.1 (a), (b) we obtain, in particular, that:
Corollary 6.3 For ϕ1 , ϕ2 ∈ L∞ (T),
(a) T̂ϕθ1 T̂ϕθ2 = Mθ Tϕ1 Tϕ2 Mθ̄|θH 2 ;
(b) Ťϕ1 Ťϕ2 = J Tϕ̄1 Tϕ̄2 J ;
(c) T̂z̄θ T̂zθ = IθH 2 ;
(d) T̂zθ T̂z̄θ = Mθ (I − 1 ⊗ 1)Mθ̄|θH 2 = IθH 2 − θ ⊗ θ |θH 2 ;
(e) Ťz̄ Ťz = J (I − 1 ⊗ 1)J|H 2 = IH 2 − z̄ ⊗ z̄|H 2 ;
− − −
(f) Ťz Ťz̄ = IH 2 .
−
T̂ϕθ1 T̂ϕθ2 = Mθ Tϕ1 Tϕ2 Mθ̄|θH 2 and Ťϕ1 Ťϕ2 = J Tϕ̄1 Tϕ̄2 J.
446 M. C. Câmara et al.
It thus follows from the properties of classical Toeplitz operators that if one of the
functions ϕ1 , ϕ2 belongs to H ∞ , then
It is well known that classical Toeplitz and Hankel operators can be characterized
in terms of compressions of Mz to H 2 and H−2 .
Recall that
(A) if T ∈ B(H 2 ), then T ∈ T (H 2) if and only if T = Tz∗ T Tz and in that case
T = Tϕ with ϕ = T (1) + T ∗ (1) − T ∗ 1, 1, (it is Brown–Halmos result, see
[10, Theorem 4.16]);
(B) if H ∈ B(H 2 , H−2 ), then H ∈ H(H 2, H−2 ) if and only if P − zH = H Tz and in
that case P − ϕ = H (1), see [16, Theorem 1.8, Chapter 1].
As a consequence of Proposition 6.1 we get the following.
Theorem 6.5 Let θ and α be two nonconstant inner functions.
(a) Let T̂ ∈ B(θ H 2 , αH 2 ). Then T̂ ∈ T (θ H 2 , αH 2 ) if and only if T̂ = T̂z̄α T̂ T̂zθ .
In that case T̂ = T̂ϕθ,α with ϕ = θ̄ T̂ (θ ) + α T̂ ∗ (α) − α θ̄ T̂ θ, α ∈ L∞ (T).
(b) Let Ť ∈ B(H−2 ). Then Ť ∈ T (H−2 ) if and only if Ť = Ťz Ť Ťz̄ and in that case
Ť = Ťϕ with ϕ = zŤ z̄ + z̄Ť ∗ z̄ − Ť z̄, z̄ ∈ L∞ (T).
(c) Let Γˆ ∈ B(θ H 2 , H−2 ). Then Γˆ ∈ T (θ H 2 , H−2 ) if and only if Ťz Γˆ = Γˆ T̂zθ
and in that case Γˆ = Γˆϕθ with P − (θ ϕ) = Γˆ θ . Moreover, there exists such a ϕ
belonging to L∞ (T).
(d) Let Γˇ ∈ B(H−2 , αH 2 ). Then Γˇ ∈ T (H−2 , αH 2 ) if and only if Γˇ Ťz̄ = T̂z̄α Γˇ and
in that case Γˇ = Γˇϕα with P − (α ϕ̄) = Γˇ ∗ α. Moreover, there exists such a ϕ
belonging to L∞ (T).
Proof By Proposition 6.4 (a) and (A), if T̂ ∈ T (θ H 2 , αH 2 ), then T (H 2 ) 8
Mᾱ T̂ Mθ|H 2 = Tz∗ Mᾱ T̂ Mθ Tz . Equivalently
= α ᾱ T̂ θ + θ̄ T̂ ∗ α − θ̄ T ∗ α, 1 θ̄ = θ̄ T̂ θ + α T̂ ∗ α − α θ̄ T ∗ α, θ .
To show (b) note that by Proposition 6.4 (b) and (A), if Ť ∈ T (H−2 ), then T (H 2 ) 8
J Ť J|H 2 = Tz∗ J T̂ J Tz . Equivalently Ť = (J Tz̄ J )Ť J Tz J|H 2 = Ťz Ť Ťz̄ . In that case
−
(Asymmetric) Dual Truncated Toeplitz Operators 447
To prove (c) we apply Proposition 6.4 (c) and (B). We have that Γˆ ∈
T (θ H 2 , H−2 ) if and only if P − zΓˆ Mθ|H 2 = Γˆ Mθ Tz . Equivalently,
In that case Γˆ = Γˆϕθ where θ ϕ is a symbol for the Hankel operator Γˆ Mθ|H 2 (by
Proposition 6.1 (c)), thus P − (θ ϕ) = Γˆ Mθ (1) = Γˆ θ .
To obtain the last condition we apply Proposition 6.4 (d) and (B). Note
that Γˇ ∈ T (H−2 , αH 2 ) if and only if Γˇ ∗ Mα|H 2 ∈ H(H 2 , H−2 ). Equivalently,
P − zΓˇ ∗ Mα|H 2 = Γˇ ∗ Mα Tz . Hence Ťz Γˇ ∗ = Γˇ ∗ Mα Tz Mᾱ|θH 2 . Finally, Ťz Γˇ ∗ =
Γˇ ∗ T̂zα , which is the same as Γˇ Ťz̄ = T̂z̄ Γˇ . In that case Γˇ = Γˇϕα where α ϕ̄ is a symbol
of the Hankel operator (Mᾱ Γˇ )∗ = Γˇ Mα|H 2 , so P − (α ϕ̄) = Γˇ ∗ Mθ 1 = Γˇ ∗ α.
We will now consider operators of the form
" #
T̂ϕθ,α Γˇϕα
D= 1 2 .
Γˆϕθ3 Ťϕ4
with ϕi ∈ L2 (T) for i = 1, 2, 3, 4. Note that if D given above is bounded, then T̂ϕθ,α
1
and Ťϕ4 are also bounded and so, as mentioned above, necessarily ϕ1 , ϕ4 ∈ L∞ (T).
On the other hand, even though for bounded D the compressions Γˇϕα2 and Γˆϕθ3 are
also bounded, the functions ϕ2 and ϕ3 may not belong to L∞ (T) (but there exist
ψ2 , ψ3 ∈ L∞ (T) such that Γˇϕα2 = Γˇψα2 and Γˆϕθ3 = Γˆψθ3 ).
We will now study relations of the operators (17) and (18) with respect to the
conjugation Cθ (see (1)). Recall that Cθ can be expressed as Cθ = Mθ J = J Mθ̄ ,
hence we obtain the following.
Proposition 6.6 For ϕ ∈ L∞ (T),
(a) T̂ϕθ,α = Cα Ťα ϕ̄ θ̄ Cθ|θH 2 = Cα Ťα Ťϕ̄ Ťθ̄ Cθ|θH 2 ;
(b) Ťϕ = (P − Cα Mθ̄ )|θH 2 T̂ϕ̄α,θ (Mθ Cα )|H 2 ;
−
(c) Γˆϕθ = Cθ Γˇ θ Cθ|θH 2 .
ϕ̄
Proof The proof of (c) can be found in [2]. A slight modification of the proof of [2,
Proposition 22] (for α = θ ) gives
where the last equality follows from (20). Hence (a) holds.
448 M. C. Câmara et al.
− − −
= P Mθ Cα Mα Tᾱϕ̄θ Mθ̄ Cθ P Mᾱ|H 2 = P Mθ Cα T̂ϕ̄θ,α Cθ Mᾱ|H 2 .
− −
Mθ Cα = Mα Cθ = Cθ Mᾱ = Cα Mθ̄ .
Note that for arbitrary ϕ ∈ L2 (T)
equalities (a)–(c) in Proposition 6.6 hold on
the set of bounded functions.
Finally, we are ready to give the following characterization of ADTTO’s.
Theorem 6.7 Let θ and α be inner functions and let D ∈ B(Kθ⊥ , Kα⊥ ). Then the
operator D is an asymmetric dual truncated Toeplitz operator, D ∈ T (Kθ⊥ , Kα⊥ ), if
and only if the following conditions hold:
(a) PαH 2 D|θH 2 = T̂z̄α PαH 2 D|θH 2 T̂zθ ;
∗
(b) P − D|H 2 = (P − Cα Mθ̄ )|θH 2 PαH 2 D|θH 2 (Mθ Cα )|H 2 ;
− −
(c) P − D|θH 2 T̂zθ = Ťz P − D|θH 2 and (PαH 2 D|H 2 )∗ T̂zα = Ťz (PαH 2 D|H 2 )∗ ;
− −
(d) P − (D(θ )) = P − (θ αD ∗ (α)) and P − (D ∗ (α)) = P − (θ αD(θ )).
In that case, D = Dϕθ,α with ϕ ∈ L∞ (T) given by
Note that (a) follows from Theorem 6.5 (a). Moreover, (b) is satisfied by Proposi-
tion 6.6 (b) and (c) is satisfied by Theorem 6.5.
Moreover,
and
−
= (P Cα Mθ̄ )|θH 2 T̂ϕ̄α,θ (Mθ Cα )|H 2 = Ťϕ .
−
By (c) and Theorem 6.5 (c)–(d) there are functions ψ, χ ∈ L∞ (T) such that
P − D|θH 2 = Γˆψθ with P − (θ ψ) = P − D|θH 2 (θ ) and (PθH 2 D|H 2 )∗ = Γˆχθ with
−
P − (αχ) = (PαH 2 D|H 2 )∗ (α). We will now use (d) to show that
−
hence (α ϕ̄ − αχ) ⊥ H−2 . Thus, by (19), we proved that P − D|θH 2 = Γˆϕθ and
(PαH 2 D|H 2 )∗ = Γˆϕ̄θ , that is, PαH 2 D|H 2 = (Γˆϕ̄θ )∗ = Γˇϕθ . Therefore,
− −
" #
T̂ϕθ,α Γˇϕα
Dϕθ = .
Γˆϕθ Ťϕ
Remark 6.8 By (21), the symbol ϕ ∈ L∞ (T) of an asymmetric dual truncated
Toeplitz operator D can be obtained by calculating D(θ ) and D ∗ (α). The symbol
ϕ can also be calculated using D(z̄) and D ∗ (z̄). To see this let ϕ = ϕ − + ϕ + ,
ϕ − ∈ H−2 , ϕ + ∈ H 2 and let ϕ̂(0) denote the 0–th Fourier coefficient of ϕ. Then, by
the fact that Cα Dϕθ,α Cθ|K ⊥ = Dαθ,α
ϕ̄ θ̄
, we have
θ
Note that the decomposition (24) is orthogonal while (21) in general is not.
Remark 6.9 Let α and θ be nonconstant inner functions. If D ∈ B(Kθ⊥ , Kα⊥ ), then
D is an asymmetric dual truncated Toeplitz operator with an analytic symbol if and
only if D satisfies conditions of Theorem 6.7 and moreover P − (zD(z̄)) = 0. The
last condition means that D(z̄) ⊥ z̄H−2 .
It is a classical result of Brown and Halmos [1] that the product of two Toeplitz
operators is zero if and only if at least one of them has a zero symbol. The product
of two Toeplitz operators Tϕ Tψ (ϕ, ψ ∈ L∞ (T)) is a Toeplitz operator if and only
if ϕ̄ or ψ is analytic. They also gave necessary and sufficient conditions for the
(Asymmetric) Dual Truncated Toeplitz Operators 451
Note that
J Hϕ J = Hϕ∗ (29)
and
J Ťϕ J = Tϕ . (30)
(a) ϕ, ψ ∈ H 2 ;
(b) ϕ̄, ψ̄ ∈ H 2 ;
(c) either ϕ or ψ is constant.
1
(1−|λ|2 ) 2
Let λ ∈ D be fixed and let Kλ = 1−zλ̄
be the normalized reproducing kernel.
Denote, for each function f ∈ L2 (T), by f+ = P + f and f− = P − f .
Lemma 7.2 Let θ be a nonconstant inner function, and let ϕ, ψ ∈ L∞ (T). Then
if and only if ϕ̄ ∈ H 2 or ψ ∈ H 2 .
452 M. C. Câmara et al.
Proof Assume Tϕ Tψ = Tθ̄ϕ Tθψ . Since Tθ̄ Tϕ = Tθ̄ϕ and Tψ Tθ = Tθψ , then Tϕ Tψ =
Tθ̄ n ϕ Tθ n ψ , for each positive integer n. Hence
Tϕ Tψ Kλ , Kλ = Tθ̄ n ϕ Tθ n ψ Kλ , Kλ .
Now we will use the properties of Berezin transform. Similarly as in [20, proof of
Theorem 4.3], we have
ϕ+ (λ)ψ− (λ) − [θ̄ n ϕ]+ (λ)[θ n ψ]− (λ) = ([θ̄ n ϕ]− [θ n ψ]+ − ϕ− ψ+ )Kλ , Kλ
+ [θ̄ n ϕ]+ (λ)[θ n ψ]+ (λ) + [θ̄ n ϕ]− (λ)[θ n ψ]− (λ) − ϕ+ (λ)ψ+ (λ) − ϕ− (λ)ψ− (λ).
The right hand side of the above equation is harmonic on D. Thus the left hand side
is harmonic on D too. For any harmonic function h(λ) we have hKλ , Kλ = h(λ).
Therefore
(ϕ+ ψ− − [θ̄ n ϕ]+ [θ n ψ]− )Kλ , Kλ = ϕ+ (λ)ψ− (λ)[θ̄ n ϕ]+ (λ)[θ n ψ]− (λ).
E
∞
An isometry T on H 2 is pure if T n H 2 = {0}. A Toeplitz operator with an
n=0
analytic symbol is a pure isometry if and only if its symbol is a nonconstant inner
function. Thus
∞
J
H2 = θ n Kθ
0
∞
∞
ϕ= θ j xj + z̄ θ̄ l ȳ l
j =0 l=0
∞
∞
with ϕ2 = xj 2 + yl 2 . Hence
j =0 l=0
⎛ ⎞ ⎛ ⎞
∞
n−1
P + θ̄ n ϕ = P + θ̄ n θ j xj = P + ⎝θ̄ n θ j xj ⎠ + Tθ̄ n ⎝ θ j xj ⎠
j =0 j n j =0
(Asymmetric) Dual Truncated Toeplitz Operators 453
Therefore
⎛ ⎞
P + θ̄ n ϕ = [θ̄ n ϕ]+ = P + ⎝θ̄ n θ j xj ⎠
j n
and
1
P + θ̄ n ϕ θ j xj = ( xj 2 ) 2 .
j n j n
and
Moreover,
$
|[θ̄ ϕ]+ [θ ψ]− Kλ , Kλ | =
n n
[θ̄ n ϕ]+ [θ n ψ]− |Kλ |2 dm
∂D
1 + |λ| + n
P θ̄ ϕ ψ,
1 − |λ|
so
thus
Since
Hence
which gives
Thus
Proof Note that (a) is equivalent to (b), since (Dϕθ )∗ = Dϕ̄θ . Using (16) in that case
we obtain
θ
T̂ϕ̄θ 0 T̂ψ̄ 0
Dϕ̄ =
θ
, Dψ̄ =
θ
,
Γˆϕ̄θ Ťϕ̄ Γˆψ̄θ Ťψ̄
and
θ
T̂ϕψ 0
θ
Dϕψ = .
Γˆϕψ
θ
Ťϕψ
Γˆϕ̄θ T̂ψ̄θ (θf ) + Ťϕ̄ Γˆψ̄θ (θf ) = P − (ϕ̄θ P + (θ̄ ψ̄θf )) + Ťϕ̄ P − (ψ̄θf )
and
Γˆϕψ
θ
(θf ) = P − (ϕψθf ) = Hθϕψ f.
Note that Hθ ϕ̄ = Hϕ̄ Tθ and by (26) Tθ Tψ̄ = Tθ ψ̄ − Hθ̄∗ Hψ̄ . Therefore by (26)
= Hϕ̄ (Tθ ψ̄ − Hθ̄∗ Hψ̄ ) + Ťϕ̄ Hθ ψ̄ = Hϕ̄ Tθ ψ̄ + Ťϕ̄ Hθ ψ̄ − Hϕ̄ Hθ̄∗ Hψ̄
= Hθ ϕ̄ ψ̄ − Hϕ̄ Hθ̄∗ Hψ̄ . (35)
λ−z
ωλ (z) =
1 − λ̄z
Tω∗λ Hf∗1 Hf2 Hf∗3 Ťωλ − Hf∗1 Hf2 Hf∗3 = − (Hf∗1 Hf2 Hf∗3 J Kλ ) ⊗ (J Kλ )
− (J Hf1 Kλ ) ⊗ (J Tωλ Tω∗λ Hf∗3 Hf2 Kλ )
Tz∗ Hϕ̄∗ Hθ̄ Hψ̄∗ Ťz − Hϕ̄∗ Hθ̄ Hψ̄∗ = − (Hϕ̄∗ Hθ̄ Hψ̄∗ J 1) ⊗ (J 1)
(Tz∗ Hϕ̄∗ Hθ̄ 1) ⊗ (Ťz∗ Hψ̄ 1) = (J Hϕ̄ 1) ⊗ (J Tz Tz∗ Hψ̄∗ Hθ̄ 1). (36)
(Asymmetric) Dual Truncated Toeplitz Operators 457
Note that Ťz∗ Hψ̄ 1 = z̄(ψ(z) − ψ(0)) and J Hϕ̄ 1 = z̄(ϕ(z) − ϕ(0)) are not zero. Let
f ∈ Kθ⊥ be such that f, Ťz∗ Hψ̄ 1 = 0. Then
(Tz∗ Hϕ̄∗ Hθ̄ 1) ⊗ (Ťz∗ Hψ̄ 1)f = f, Ťz∗ Hψ̄ 1Tz∗ Hϕ̄∗ Hθ̄ 1
and
(J Hϕ̄ 1) ⊗ (J Tz Tz∗ Hψ̄∗ Hθ̄ 1)f = f, J Tz Tz∗ Hψ̄∗ Hθ̄ 1J Hϕ̄ 1,
so by (36)
(λ1 J Hϕ̄ 1) ⊗ (Ťz∗ Hψ̄ 1) =(J Hϕ̄ 1) ⊗ (J Tz Tz∗ Hψ̄∗ Hθ̄ 1);
(J Hϕ̄ 1) ⊗ (λ̄1 Ťz∗ Hψ̄ 1) =(J Hϕ̄ 1) ⊗ (J Tz Tz∗ Hψ̄∗ Hθ̄ 1).
Hence
Since J Tz J = Ťz∗ and the kernel of Tz is zero, the kernel of Ťz∗ is also zero.
Thus (38) is equivalent to
so
and
λ1 J Hψ̄ 1 = λ1 J P − ψ̄ = λ1 P + (z̄ψ).
458 M. C. Câmara et al.
Since the set {kw : w ∈ D} is linearly dense in H 2 , we see that Hϕ̄ Hθ̄∗ Hψ̄ = 0.
As a conclusion of Theorem 7.1, Lemma 7.3 and Lemma 7.5 we have the
following.
Theorem 7.6 Let θ be a nonconstant inner function and ϕ, ψ ∈ L∞ (T). Then
Dϕθ Dψθ = Dϕψ
θ if and only if one of the following conditions hold
Theorem 7.7 Let θ be a nonconstant inner function and let ϕ, ψ ∈ L∞ (T). Then
Comparing Theorems 7.6 and 7.7 note that if Dϕθ Dψθ = Dϕψ
θ
, then the operators Dϕθ
θ
and Dψ commute. However, the converse is not true.
Example Let α, β be nonconstant inner functions and let θ = αβ. Note that then by
Theorem 7.7 the operators Dθθ and Dαθ commute. On the other hand, ᾱ θ̄ θ = ᾱ is not
analytic, hence, by Theorem 7.6, Dθθ Dαθ = Dθα
θ .
Acknowledgments The work of the first author was partially supported by FCT/Portugal through
UID/MAT/04459/2020. The research of the second and the third authors was financed by the
Ministry of Science and Higher Education of the Republic of Poland.
References
1. A. Brown, P. Halmos, Algebraic properties of Toeplitz operators J. Reine Angew. Math. 213,
89–102 (1964)
2. M.C. Câmara, K. Kliś-Garlicka, B. Łanucha, M. Ptak, Compressions of multiplications
operators and their characterizations. Results Math. 75, 157 (2020). https://doi.org/10.1007/
s00025-020-01283-4
3. C. Câmara, K. Kliś-Garlicka, B. Łanucha, M. Ptak, Conjugations in L2 (T) and their invariants.
Anal. Math. Phys. 10, 22 (2020). https://doi.org/10.1007/s13324-020-00364-5
4. C. Câmara, K. Kliś-Garlicka, B. Łanucha, M. Ptak, Intertwining property for compressions of
multiplication operators (2020). arXiv:2012.05330
5. M.C. Câmara, K. Kliś-Garlicka, B. Łanucha, M. Ptak, Invertibility, Fredholmness and kernels
of dual truncated Toeplitz operators. Banach J. Math. Anal. 14, 1558–1580 (2020)
6. M.C. Câmara, K. Kliś-Garlicka, B. Łanucha, M. Ptak, Shift invariance and reflexivity of
compressions of multiplication operators. Forum Math. (2022). https://doi.org/10.1515/forum-
2021-0129
7. X. Ding, Y. Sang, Dual truncated Toeplitz operators. J. Math. Anal. Appl. 461, 929–946 (2018)
8. R.G. Douglas, Banach Algebra Techniques in Operator Theory, vol. 179 (Springer, Berlin,
2012)
9. P.L. Duren, Theory of H p Spaces. Pure and Applied Mathematics, vol. 38 (Academic Press,
New York, 1970)
10. S.R. Garcia, J.E. Mashreghi, W. Ross, Introduction to Model Spaces and Their Operators.
Cambridge Studies in Advanced Mathematics, vol. 148 (Cambridge University Press, Cam-
bridge, 2016)
11. S.R. Garcia, M. Putinar, Complex symmetric operators and applications. Trans. Am. Math.
Soc. 358, 1285–1315 (2006)
460 M. C. Câmara et al.
12. S. Garcia, W.T. Ross, Model Spaces: A Survey. Invariant Subspaces of the Shift Operator.
Contemp. Math. vol. 638 (Am. Math. Soc., Providence, RI, 2015), pp. 197–245
13. S. Garcia, W.T. Ross, W.R. Wogen, C*-Algebras Generated by Truncated Toeplitz Operators,
vol. 8 (Math and Computer Science Faculty Publications, 2014). http://scholarship.richmond.
edu/mathcs-faculty-publications/8
14. P. Koosis, Introduction to H p Spaces, 2nd edn. (Cambridge University Press, Cambridge,
1998)
15. P. Ma, F. Yan, D. Zhang, Zero, finite rank, and compact big truncated Hankel operators on
model spaces. Proc. Am. Math. Soc. 146, 5235–5242 (2018)
16. V.V. Peller, Hankel Operators and Their Applications (Springer, New York, 2003)
17. M. Ptak, K. Simik, A. Wicher, C–normal operators. Electron. J. Linear Algebra 36, 67–79
(2020)
18. Y. Sang, Y. Qin, X. Ding, A theorem of Brown–Halmos type for dual truncated Toeplitz
operators. Ann. Funct. Anal. 11, 271–284 (2020)
19. D. Sarason, Algebraic properties of truncated Toeplitz operators, Oper. Matrices 1, 491–526
(2007)
20. K. Stroethoff, The Berezin transform and operators on spaces of anlytivc functions. Banach
Center Publ. Linear Oper. 85, 361–380 (1997)
21. D. Xia, D. Zheng, Products of Hankel operators. Integr. Equ. Oper. Theory 29, 339–363 (1997)
Boundedness of Toeplitz Operators
in Bergman-Type Spaces
The focus of this article is on recent results on the boundedness of Toeplitz operators
on weighted Bergman spaces of holomorphic functions, mainly on the open unit
disk D of the complex plane C, although some of the results are also formulated on
the unit ball BN of CN , N = 2, 3, . . . . The related question on the compactness is
only considered when it can be dealt with parallel to boundedness, and certain more
special recent results for compactness will remain out of this review.
We will concentrate on two circles of ideas. First, we deal with Toeplitz
operators with oscillating symbols and weak Carleson-type sufficient conditions for
J. Taskinen ()
Department of Mathematics and Statistics, University of Helsinki, Helsinki, Finland
e-mail: [email protected]; [email protected]
J. A. Virtanen
Department of Mathematics and Statistics, University of Reading, Reading, England
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 461
R. M. Aron et al. (eds.), Operator and Norm Inequalities and Related Topics,
Trends in Mathematics, https://doi.org/10.1007/978-3-031-02104-6_14
462 J. Taskinen and J. A. Virtanen
boundedness. The starting point of this direction of research is the article [30]. The
second approach applies to operators with radial symbols, and it is based on the
results on the structure of weighted Bergman spaces which were pioneered in the
works of W. Lusky [17–19] and adapted to the study of Toeplitz operators recently in
the papers [4, 5]. This led to a characterization of the boundedness and compactness
of Toeplitz operators in weighted H ∞ -spaces.
Let us present the basic notation and definitions. The notation concerning the
spaces on the unit ball BN will only be needed and thus given at the end of Sect. 2.
The normalized area measure on D is denoted by dA = π −1 rdrdθ , where r and θ
are the polar coordinates of z = reiθ ∈ C. Given 1 ≤ p < ∞ and the real parameter
α > −1 we define the weighted area measure by dAα (z) = (1 + α)(1 − r 2 )α dA(z)
and set
4 $ 5
p
Lpα (D) = g : D → C measurable : gp,α := |g|p dAα < ∞ and
D
in the case α = 0 these spaces are denoted by Lp (D) and Ap (D), respectively. Here,
v(z) = (1 − |z|2 )α are called standard weights.
We will also consider more general weighted Bergman spaces and their analogue,
weighted Hardy space Hv∞ corresponding to p = ∞. In general, by a weight
v we mean a continuous function D →]0, ∞[ which is radial, vanishing on the
boundary and decreasing with the radius, i.e. there holds v(z) = v(|z|) for all z ∈ D,
lim|z|→1 v(z) = 0 and v(r) ≥ v(s) if 1 > s > r > 0. We denote vdA = dAv and,
for 1 ≤ p < ∞,
4 $ 5
p
Lpv (D) = g : D → C measurable : gp,v := |g|p dAv < ∞ and
D
and
h∞
v (D) = {g : D → C : g harmonic, gv := sup |g(z)|v(|z|) < ∞}
z∈D
and
Hv∞ (D) = {g ∈ h∞
v : g holomorphic };
Boundedness of Toeplitz Operators in Bergman-Type Spaces 463
Ta f = Pα Ma f, (1)
but the assumptions made so far do not always suffice to guarantee that (1) makes
sense, since Ma might map f outside L2α (D). In the case a is a bounded function,
there is no problem with the definition, since Pα can be written with the help of the
Bergman kernel as the integral operator
$
f (w)
Pα f (z) = dAα (w) ,
(1 − zw)2+α
D
hence
$
a(w)f (w)
Pα Ma f = dAα (w), (2)
(1 − zw)2+α
D
and for every z ∈ D, these integrals converge for all f ∈ L1α (D). Moreover, it is
p
known that Pα is a bounded operator in the space Lα (D), when 1 < p < ∞, which
p p
yields the boundedness of Ta : Aα (D) → Aα (D) for bounded symbols.
It is not difficult to construct unbounded symbols a which still induce bounded
Toeplitz operators, but the characterization of symbols a ∈ L1 (D) such that
p p
Ta : Aα (D) → Aα (D) is well-defined and bounded is a well-known open problem.
Let us mention some partial results on it. The characterization of boundedness and
compactness of Toeplitz operators with nonnegative symbols in terms of Carleson
type measures first appeared in [24].
D. Luecking [15] proved that a Toeplitz operator Ta with a nonnegative symbol
a ∈ L1 (D) is bounded in A2 (D), if and only if the average
$
−1
|B(z, r)| a(w) dA(w)
B(z,r)
is a bounded function of z. Here B(z, r) denotes a disk in the Bergman metric, with
center z and some fixed radius r > 0. Toeplitz operators with radial symbols in
464 J. Taskinen and J. A. Virtanen
the space A2α (D) and analogues on higher dimensional domains were thoroughly
considered in [10]: in this case the operator is unitarily equivalent with a sequence
space multiplier, see also (44) below, and thus the boundedness properties can be
determined. A partial generalization to the case p = 2 was established in [21]. The
Berezin transform
$
f (w)
B(f )(z) = (1 − |z|2 )2 dA(w), z ∈ D, (3)
|1 − zw̄|4
D
is a useful tool for the theory of Toeplitz operators, although it will not be used in
this article. N. Zorboska proved in [39] for symbols a of bounded mean oscillation
that the Toeplitz operator Ta : A2 (D) → A2 (D) is bounded if and only if B(a) is
bounded. The results of [15] and [39] generalize to other Ap (D)-spaces, 1 < p <
∞, as well, see e.g. [30]. Here is a non-exhaustive list of other works dealing with
the boundedness and compactness of Toeplitz operators in Bergman-type spaces:
[8], [10], [9], [12], [11], [15], [16], [21], [22], [25], [27], [28], [29], [30], [34], [35],
[36], [37], [39]. The monograph [38] is a standard reference for the topic, and we
also mention the survey article [31].
In this article we will review in Sect. 2 the results of [30], [33], [12]. These consist
of sufficient, weak Carleson-type conditions for the boundedness and compactness
of Toeplitz operators in reflexive Bergman spaces with standard weights, both on the
unit disk and the unit ball. Sections 3–6 are mainly based on the recent works [4, 5],
which deal with operators on Hv∞ (D)-spaces with quite general classes of weights.
Theorem 5 of Sect. 4 states that there is a bounded harmonic symbol f for which Tf
is unbounded in Hv∞ (D) for any radial weight v satisfying our general assumptions.
The main result of Sect. 5, Theorem 7 contains a necessary and sufficient condition
for the boundedness of Tf in Hv∞ (D), as well as the corresponding result for the
compactness. These conditions are slightly abstract, and thus in Sect. 6 we derive
some more concrete, easily formulated sufficient conditions based on the results of
Sect. 5.
We conclude this section by a remark on the definition of Toeplitz operators as
an improper integral. Here, we fix α > −1 and assume the symbol a is radial.
Formula (4) will be considered in detail in Sect. 2 even for more general, non-radial
symbols. The proof of Proposition 1 is taken here from [14], although some versions
of it have probably been known for specialists for a long time.
Proposition 1 Let a be a radial symbol, i.e. a(z) = a(|z|) for almost all z ∈ D,
belonging to L1α (D), α > −1, and let g(z) = ∞ n
n=0 gn z be a holomorphic function
on D. Then, the defining integral (2) of Ta g exists in the improper sense as the limit
$
a(w)g(w)
Ta g(z) = lim dAα (w), (4)
ρ→1 (1 − zw)2+α
|w|<ρ
Boundedness of Toeplitz Operators in Bergman-Type Spaces 465
and in particular the power series on the right converges for all z ∈ D.
Here and in the next we denote by B and Γ Euler’s beta- and gamma-functions,
n!Γ (c)
B(n + 1, c) = , c > 0,
Γ (n + 1 + c)
and for 0 < ρ ≤ 1
√
$ρ
√
βa,α (ρ, n) = (α + 1) t n (1 − t)α a( t)dt, (6)
0
where the integral converges by the assumptions that a is radial and belongs to
L1α (D).
Proof of Proposition 1 We start by the remark that for all m ∈ N0 , the integral
$
g(w)w m a(w)dAα (w)
D
exists in the improper sense for every holomorphic g on the disk D. Namely, the
rotational symmetry of a and the usual orthogonality relations of trigonometric
polynomials yield for all m ∈ N0
$ $ρ
g(w)w a(w)dAα (w) = 2(α + 1)gm
m
r 2m+1 a(r)(1 − r 2 )α dr. (7)
|w|<ρ 0
Clearly, the limit exists, when ρ → 1. For every 0 < ρ < 1, z ∈ D, we obtain by (7)
$
a(w)g(w)
dAα (w)
(1 − zw)2+α
|w|<ρ
$ ∞
(zw)n
= g(w) a(w)dAα (w)
(α + 1)B(n + 1, α + 1)
|w|<ρ n=0
∞
βa,α (ρ, n)gn
= zn . (8)
(α + 1)B(n + 1, α + 1)
n=0
466 J. Taskinen and J. A. Virtanen
n!Γ (α + 1)
B(n + 1, α + 1) ≥ ≥ CL n−L (9)
(n + L)!
$1
βa,α (ρ, n) ≤ βa,α (1, n) = 2(α + 1) t 2n (1 − t 2 )α a(t)dt ≤ Cα (10)
0
for another constant Cα > 0, for all ρ and n, since a ∈ L1α (D). Moreover, since g is
1
a holomorphic function on D, we have lim supn→∞ |gn | n ≤ 1, hence,
1
βa,α (1, n)gn n 1 1
lim sup ≤ lim sup CL Cα nL n · lim sup |gn | n ≤ 1.
n→∞ (α + 1)B(n + 1, α + 1) n→∞ n→∞
1
D = {ρeiφ | r ≤ ρ ≤ 1 − (1 − r) , θ ≤ φ ≤ θ + π(1 − r)} (11)
2
)
for all 0 < r < 1, θ ∈ [0, 2π]. Let |D| := D dA and, for w = ρeiφ ∈ D(r, θ ), let
$ρ $φ
1
âD (w) := a(-eiϕ )-dϕd-. (12)
|D|
r θ
We will study symbols a for which there exists a constant C > 0 such that
Ta : Ap → Ap , defined by
∞
Ta f (z) = Fn f (z) (17)
n=1
is bounded for all 1 < p < ∞, and there is a constant Cα , independent of a, such
that
The main step of the proof consists of writing the integral (16) in polar
coordinates and performing a double integration by parts (once with respect to
both coordinates) such that there appear integrals of a and derivatives of f (w)(1 −
|w|2 )α (1 − zw̄)2+α . The former can be estimated by using the assumption (13)
and the latter by using bounds for the maximal Bergman projection and well
known arguments and estimates related with hyperbolic geometry. One obtains a
representation for the integral (2) as a pointwise convergent sum of the integrals (16)
as in (17). We refer to [30] for the details. Improved versions of the proof appear in
[33] and [12], and they yield our next theorem, although we do not repeat the proof
here. We remark that every Toeplitz operator
$
a(w)f (w)
Taρ f (z) = dAα (w) (19)
(1 − zw̄)2+α
|w|<ρ
p p
is bounded Aα (D) → Aα (D), since the support of the symbol is contained in a
compact subset of D.
Theorem 2 Let 1 < p < ∞ and 1/p + 1/q = 1, and let the symbol a be as in
p p
Theorem 1. Then, the generalized Toeplitz operator Ta : Aα (D) → Aα (D), defined
in (17), can be written as
p
for all f ∈ Aα (D). The limit converges with respect to the strong operator topology.
Moreover, the transposed operator Ta∗ : Aα (D) → Aα (D) (with respect to the
q q
q
for f ∈ Aα (D) and for almost all z ∈ D, and the limit also converges in the strong
operator topology.
Boundedness of Toeplitz Operators in Bergman-Type Spaces 469
The limits in (20), (21) cannot in general converge in the operator norm, since
the operators Taρ are compact. We mention that, when α = 0, the above results are
formulated in [33] also for little Hankel operators
$
a(w)f (w)
ha f (z) = dA(w) , z ∈ D. (22)
(1 − z̄w)2
D
Here, one also defines using the same decomposition of the unit disk as above
$
a(w)f (w)
Hn f (z) = dA(w) , z ∈ D, (23)
(1 − z̄w)2
Dn
and defines the generalized little Hankel operator ha f (z) as ∞ n=1 Hn f (z). Then,
if (13) holds for the symbol a, one obtains that ha : Ap (D) → Lp (D) is bounded
for all 1 < p < ∞, the operator norm of ha has the same bound as in (18), and
finally, the operator ha and its transpose have representations as improper integrals
similar to those in (20), (21).
The definition (17) of a generalized Toeplitz operator depends on the geometry
of the special decomposition (14) of the unit disk, but Theorem 2 largely removes
this unsatisfactory feature, since the improper integral in (20) is quite a natural one.
We remark that in the literature there are versions of the result, which use different
subdomains of the unit disk. In [36] the condition (13) is replaced by a similar one
on Carleson squares
6 7
Shα (eiθ ) = ρeiφ : 1 − h < ρ < 1, |φ − θ | < παh
Theorems 1 and 2, first proved in [30] and [33], have been generalized to the case
of Toeplitz operators on the Bergman space of the unit ball of CN in the recent work
[12], but even presenting the results leads to non-trivial technical challenges. We
do not directly need the Euclidean space R3 here, but since that dimension is still
within the capabilities of the human imagination, we ask the reader to think about a
radially symmetric decomposition of the unit ball of R3 : that is indeed a challenge,
since decomposing the ball surface into finitely many identical squares in spherical
coordinates (corresponding to intervals [θn , θn ] in (14)–(15)) is impossible. For
example, starting to fill the ball surface from the equator with spherical squares with
one side parallel to the meridians, one runs into difficulties at latest when trying to
fill the north and south caps.
The results of [12] are formulated for measures with standard weights and thus
the proofs contain new information even in the case N = 1, since the earlier papers
only contained the unweighted case. The basic idea of the proof is the same as in
[30] and [33], but new non-trivial technical considerations are nevertheless needed.
Let us review the approach of [12] superficially without going into all technical
details. For α > −1, we define the weighted Lebesgue measure dVα on the unit
ball BN , N ∈ N, by dVα (z) = cα (1 − |z|2 )α dV (z), where dV is the unweighted
)N-dimensional (real) Lebesgue measure and cαpis a normalizing constant such that
BN dV α = 1. For 1 ≤ p < ∞, we denote by L α (BN ) the Lp -space with respect to
p
the measure dVα and by Aα (BN ) the weighted Bergman space of all holomorphic
p
functions in Lα (BN ). We also denote by Pα the orthogonal projection from L2α (BN )
p p
onto A2α (BN ). It is known to be a bounded operator Lα (BN ) onto Aα (BN ) for all
1 < p < ∞.
In the following it is useful to work with real variables by identifying CN with
Rn , n = 2N, so that BN equals Bn in real coordinates. Accordingly, any point
x ∈ Bn with modulus |x| = r can be written as
n−1
ξ = (r, θ2 , · · · , θn ) ∈ [0, 1[× [0, π[×[0, 2π[ =: Qn ,
j =2
x y ⇐⇒ x1 ≤ y1 , | π2 − x2 | ≥ | π2 − y2 |, . . . , | π2 − xn−1 | ≥ | π2 − yn−1 |,
xn ≤ yn . (24)
(j ) (j )
On each Qj we pick up the smallest and largest points x (j ) = x1 , . . . , xn
(j ) (j )
and y (j ) = y1 , . . . , yn with respect to the given ordering, hence, there holds
(j ) (j )
Qj = Q x , y , where we denote, for a, b ∈ Qn with a b,
6 7
Q(a, b) = x ∈ Rn : a x b , B(a, b) = σ Q(a, b) . (25)
Note that for x, y ∈ [0, 1) × [0, π2 ]n−2 × [0, 2π] the order relation “” coincides
with the usual partial order of points in Rn , which is then mirrored to all of Qn to
account for the construction of the sets Qj and Bj . In particular, the x (j ) and y (j )
are two opposite corners of Qj and we have Bj = B(x (j ) , y (j ) ).
Let a : BN → C be a locally integrable function and 1 < p < ∞. The
generalized Toeplitz operator is defined by
∞
∞
Ta f (z) := Ta (χj f )(z) = Pα (aχj f )(z), (26)
j =1 j =1
p
if the series converges for almost every z ∈ BN and all f ∈ Aα (BN ). Here χj
denotes the characteristic function of the set Bj . The boundedness of the Bergman
p p
projection Pα in Lα (BN ) implies that Ta f = Pα (af ) whenever af ∈ Lα (BN ).
In particular, if a is bounded, then Ta is just the standard Toeplitz operator. As in
the one-dimensional case, a “weak” Carleson-type condition (28) implies that Ta
becomes a well-defined bounded linear operator and the definition coincides with
the integral definition, when it is interpreted as an improper integral. Accordingly,
given a locally integrable a : BN → C, we define for all j ∈ N
$
aj := sup
B a dVα (27)
y∈Bj
B(x (j) ,y)
)
and denote |B|α = B dVα for all measurable subsets B ⊂ BN .
472 J. Taskinen and J. A. Virtanen
B
aj ≤ Ca |Bj |α (28)
p
for all j ∈ N, then the series (26) converges almost everywhere and in Lα (BN ) and
p p
defines a bounded linear operator Aα (BN ) → Aα (BN ) with Ta ≤ Cα Ca , for
some constant Cα > 0 independent of a.
Given the symbol a as above and 0 < ρ < 1, we define aρ (z) = a(z) for |z| ≤ ρ
p
and aρ (z) = 0 for ρ < |z| < 1; then every operator Taρ is bounded on Aα (BN ),
since the supports of the symbols are compact subsets of the unit ball, or also by
the previous theorem. As in the one-dimensional case, the assumption (28) allows
the following representation of the Toeplitz operator, which does not depend on the
decomposition (Bj )j ∈N .
Theorem 4 Let 1 < p < ∞ and 1/p + 1/q = 1, and suppose that a ∈ L1loc
satisfies (28). Then
Ta f = lim Taρ f
ρ→1
for all f ∈ Aα (BN ) and the transpose operator Ta∗ : Aα (BN ) → Aα (BN ) can be
p q q
expressed as
p
for f ∈ Aα (BN ).
p
The transpose is defined respect to the standard duality of Aα (BN )-spaces.
It would probably be possible and technically easier to formulate and prove a
result analogous to Theorem 3 by using a rectangular Whitney decomposition of
BN instead of the one described here, but there would then be the disadvantage
that the spherical symmetry would be lost and the condition for the boundedness
would depend on the particular choice of the decomposition. In particular, it might
be difficult or impossible to prove Theorem 4 with that approach.
From now on we will deal with Toeplitz operators in spaces on D with quite
general weights v satisfying the basic assumptions of Sect. 1. A typical, important
example of weights considered in this section is the exponentially decreasing
v(r) = exp(−1/(1 − r)). Because of such examples we need again to pay attention
to the definition of Toeplitz operators in the spaces Av (D) and Hv∞ (D), namely,
p
Boundedness of Toeplitz Operators in Bergman-Type Spaces 473
there is the problem that the Bergman projection may not be bounded. Actually we
will show that this is always the case for p = ∞ for any weight, see Theorem 5, but
even in the reflexive case there may be problems in this respect: in [7] it was shown
that for the above mentioned exponential weight v(z), the orthogonal projection
p
L2v (D) → A2v (D) is bounded in Lv if and only if p = 2. Moreover, in [19] W. Lusky
proved that the mere existence of a bounded projection from L∞ ∞
v (D) onto Hv (D)
is equivalent to v satisfying condition (B) of Definition 2, below. For example, the
exponential weight v satisfies (B), but there also exist natural weights which do not,
like v(z) = (1 − log(1 − |z|))−1 (see the statement after Theorem 1.2. of [19] and
Example 2.4. of the same paper for other examples).
Yet, even in the spaces Hv∞ (D) and Av (D) with general weights, the definition
p
of the Toeplitz operator involves the orthogonal projection Pv : L2v (D) → A2v (D). It
will be useful to consider the integral kernel of Pv , the so called Bergman kernel. In
the next we follow well-known arguments, see e.g. [7]. ) We denote the inner product
in the Hilbert spaces L2v (D) and A2v (D) by f, g = D f g dAv . Then, the functions
−1/2
ek (z) = Γ2k zk , where k ∈ N0 and
$1
Γk = 2π r k+1 v(r)dr, (29)
0
form an orthonormal basis of A2v (D). We remark that the numbers Γk satisfy for all
0 < - < 1 and some constant Cv,- > 0 the following lower bound
Γk ≥ Cv,- -k (30)
for every k ∈ N0 . This follows from (29) by considering the integral e.g. over the
interval [-, 1 − (1 − -)/2] only.
p
Convergence in the space Av (D), 1 < p < ∞, with respect to the norm · p,v
p p
implies pointwise convergence (hence Av (D) is a closed subspace of Lv (D) ), and
thus the point evaluation functionals at any point of D are bounded functionals on
p
Av (D). Consequently, we find the Bergman kernel by using the Riesz representation
theorem, which allows us to choose the family of functions Kz ∈ A2v (D), z ∈ D,
such that
$
g(z) = g, Kz = g(w)Kz (w) dAv (w) (31)
D
for all g ∈ A2v (D). The integral operator defined by the right hand side can be
extended to L2v (D), and it actually defines the orthogonal projection from L2v (D)
474 J. Taskinen and J. A. Virtanen
onto A2v (D), i.e. the Bergman projection Pv . Using the orthonormal basis (ek )∞
k=0
we can write for all z ∈ D
∞
$
∞ k k
z w
Pv g(z) = g, ek ek (z) = g(w)dAv (w). (32)
Γ2k
k=0 D k=0
Here, the order of the summation and the integral can be changed, because (30)
leads for any fixed z ∈ D to the estimate
zk w k |z| k
≤ cv,- 2 , (33)
Γ2k -
and we can choose here -2 > |z| so that the sum on the right-hand side of (32)
converges well enough. Moreover, the estimate (33) implies that for every z ∈ D the
Bergman kernel Kz is a bounded function:
∞
∞
$
zn
(Tf g)(z) = fg, en en (z) = f (w)g(w)w n v(w)dA , (36)
Γ2n
n=0 n=0 D
Boundedness of Toeplitz Operators in Bergman-Type Spaces 475
where the series converges in L2v (D). However, the formula also holds for all g ∈
Hv∞ (D) (since we are assuming f ∈ L1 (D)) and the product fgv thus belongs to
L1 (D), and one can commute the summation and integration in (36), due to (33).
In the latter case, the sum (36) converges uniformly for z in compact subsets of the
disk.
For example, the standard weights v(r) = (1 − r 2)α , α > 0 are normal, whereas the
weights of exponential type, v(r) = exp(−α/(1−r)β ), α, β > 0, are not. The Riesz
projection P maps harmonic functions into holomorphic ones and it is defined by
∞
P ak r |k| eikθ = ak r k eikθ , r ∈ [0, 1), θ ∈ [0, 2π]. (39)
k∈Z k=0
For every m > 0 we denote by rm be a point where the function r $→ r m v(r) attains
its absolute maximum on [0, 1]. Due to the general assumptions on the weights it is
easily seen that rn ≥ rm if n ≥ m and limm→∞ rm = 1; see for example [17] for
details.
We now turn to questions on the boundedness of Toeplitz operators Tf with
harmonic symbols f . In the case f is even holomorphic, the operator Tf is just the
multiplier Mf , and it is quite plain that Tf is bounded, if and only if f ∈ H ∞ (D),
i.e., f is a bounded function. Due to the generality of the weights, the details
of this claim are exposed in [4, Section 2]. Allowing the symbol to be just a
harmonic function changes the situation dramatically. The basic reason for this is
the unboundedness of the Riesz and Bergman projections with respect to the sup-
476 J. Taskinen and J. A. Virtanen
norm, but one can develop this idea as far as the following result. We repeat that in
all of our results the weights v must satisfy the general assumptions made in Sect. 1.
Theorem 5 There is a bounded harmonic function f : D → C such that Tf is not
a bounded operator Hv∞ (D) → Hv∞ (D) for any weight v on D.
This result implies the following conclusion.
Corollary 1 For any weight v, the Bergman projection Pv is not a bounded
mapping L∞ ∞
v (D) → Lv (D).
$π/2
1 eikπ/2 − e−ikπ/2 ei|k|π/2 − e−i|k|π/2
ak = e−ikt dt = =
2π 2kπi 2|k|πi
−π/2
(−1)j
(2j +1)π , if |k| = 2j + 1 for some j ∈ N0 ,
=
0 for other k ∈ Z \ {0}.
Next we define the test functions, which will be used in showing the unbounded-
ness of the Toeplitz operator: we set
r m eimθ
gm (z) = m v(r )
, z = reiθ ∈ D
rm m
for all m ∈ N0 , where the definition of the maximum point rm was given in the
beginning of the section so that we obviously have gm v = 1. We next show that
for all m ∈ N0 there holds
m ∞
Γ2m zk zk
Tf gm (z) = bk−m m v(r )
+ b k−m m v(r )
(41)
Γ2k rm m rm m
k=0 k=m+1
∞ |k| e ikθ
where f (z) = k=−∞ bk r and Γk is as in (29). Indeed, this follows from
and (36).
Let us now turn to the final proof showing that Tf is unbounded on Hv∞ (D). We
define
∞
(−1)j
f1 (z) = z2j +1 + z̄2j +1
2j + 1
j =0
and note that it suffices to show that Tf1 is unbounded since Tf = T1/2 + π −1 Tf1
and T1/2 (multiplication by constant 1/2) is bounded. Fix a positive integer m, say
m = 4m0 for m0 ∈ N. Then
odd if k is odd odd if j is odd
k − m is and j − 2m0 is
even if k is even even if j is even.
m ∞
Γ2m zk zk
Tf1 gm (z) = bk−m m
+ bk−m m . (42)
Γ2k rm v(rm ) rm v(rm )
k=0, k=m+1,
k odd k odd
478 J. Taskinen and J. A. Virtanen
Recall that b4j +1−m = 1/|4(j − m0 ) + 1|. So if we take θ = 0 then all summands
in the preceding sum are non-negative. Hence
∞ ∞ 4j +1
rm 1 rm (rm
4 )j rm
log = ≤
5 1 − rm
4 5 j 4j + 1
j =1 j =0
4j +1
rm v(rm )
= b4j +1−m m v(r )
≤ S(Tf1 (gm ))(rm )v(rm )
rm m
m+1≤4j +1<∞
since trivially by the definition of the operator S we have sup|z|=r |(Sf )(z)| ≤
sup|z|=r |f (z)|. Since limm→∞ rm = 1, the left-hand side of the preceding estimate
grows to the infinity, when m → ∞. Hence Tf1 and also Tf cannot be bounded.
∞
$1 $2π
zn
Tf g(z) = f (r)g(reiθ )r n+1 e−inθ v(r) dθ dr
Γ2n
n=0 0 0
∞
$1 ∞
zn
= f (r)r 2n+1 v(r)gn dr = γn gn zn (44)
Γ2n
n=0 0 n=0
Boundedness of Toeplitz Operators in Bergman-Type Spaces 479
where g = n gn z
n and
$1
1
γn = r 2n+1 v(r)f (r)dr. (45)
Γ2n
0
We expose here the approach based mainly on the works [17], [19] and [20]
dealing with the condition (B), below, which according to Theorem 1.1 of [19]
characterizes those radial weights such that the space Hv∞ (D) is isomorphic to the
Banach space ∞ . Examples of weights satisfying (B) are all normal weights (38),
in particular the standard weights, and the weights of exponential type v(r) =
exp(−γ /(1 − r)β ); see [19].
The very definition of condition (B) is somewhat technical and we cannot quite
avoid other technical considerations in this survey either, however, one can follow
our presentation without going into the depth of the arguments just by keeping in
mind that condition (B) associates
∞to the weight an increasing sequence of indices
(mn )∞
n=1 ⊂ (0, ∞) and radii rmn n=1 ⊂ (0, 1) such that mn → ∞ and rn → 1 as
n → ∞, and moreover, gives the very useful equivalent representation in Theorem 6
for the weighted sup-norm. We recall that the numbers rm ∈]0, 1[ were defined in
the beginning of Sect. 4.
Definition 2 The weight v satisfies the condition (B), if
Note that here m and n need not be integers. We now fix a number b > 2: it
is shown in Lemma 5.1. of [19] that it is then possible to choose, by induction, an
increasing, unbounded sequence (mn )∞n=1 ⊂ (0, ∞) such that
mn mn+1
rmn v(rmn ) rmn+1 v(rmn+1 )
b = min , .
rmn+1 v(rmn+1 ) rmn v(rmn )
where k ∈ Z and m0 = 0. Here [r] is the largest integer not greater than r. With the
help of these numbers we define the coefficient multipliers of de la Valleé Poisson
480 J. Taskinen and J. A. Virtanen
∞ |k| e ikθ ,
type, acting on harmonic functions f (z) = k=−∞ fk r by
∞
∞
Wn : fk r |k| eikθ $→ wnk fk r |k| eikθ
k=−∞ k=−∞
for all 0 ≤ r ≤ 1 and g ∈ h∞ v (D). See (37) for the notation. The inequality (47)
follows e.g. by combining an inequality in Theorem 1 of [20] with Lemma 3.3.
of [19].
The operators Wn are important, since they decompose the space Hv∞ (D) into
finite dimensional blocks with a useful representation for the norm. The result is
from Theorem 1 of [20], see also Propositions 4.1. and 5.2. of [19].
Theorem 6 Let v satisfy (B). Then there are constants c1 , c2 > 0 such that, for all
g ∈ h∞
v (D),
c1 sup M∞ (Wn g, rmn )v(rmn ) ≤ gv ≤ c2 sup M∞ (Wn g, rmn )v(rmn ) (48)
n∈N n∈N
and
for all n ∈ N.
Moreover, it follows from Theorem 6 that if the numbers fk ∈ C, k ∈ Z satisfy
|k| ikθ
sup sup wnk fk rm n
e v(rmn ) < ∞, (50)
n∈N θ∈[0,2π] m
n−1 <|k|≤mn+1
∞ |k| ikθ
then the series defining the harmonic function f (reiθ ) = k=−∞ fk r e
∞
converges uniformly on compact subsets of D and f belongs to hv (D) and gv is
bounded by a constant depending on the weight v. For this statement, see Remark
1, (iii) of [20].
Examples If v is normal then one can take mn = 2kn for suitable fixed k > 0
(see [19, Example 2.4], and [17]). For v(r) = exp(−α/(1 − r)β ) one can take
mn = β 2 (β/α)1/β n2+2/β − β 2 n2 , see [2].
We now formulate one of the main results of this section, the characterization
of boundedness and compactness for coefficient multipliers. The case of Toeplitz
operators with radial symbols follows easily from this. The result was already
Boundedness of Toeplitz Operators in Bergman-Type Spaces 481
proven for a more restricted class of weights in Theorem 4.1 of [18]. We will assume
that a sequence (γk )∞ k=0 of complex numbers is given, and consider the formal series
f (θ ) = ∞ k=0 γ k e ikθ , which may or may not converge. The formal series W f is
n
then naturally defined as
∞
Wn f (θ ) = wnk γk eikθ
k=0
where the numbers wnk are as in (46). We denote by Mf the coefficient multiplier
∞
Mf g(z) = γk gk r k eikθ , z = reiθ (51)
k=0
for harmonic functions g(z) = ∞ |k| ikθ . By definition, M g is holomor-
k=−∞ gk r e f
phic, if the series (51) converges.
Theorem 7 Let the weight v satisfy condition (B). Then Mf maps h∞
v (D) into
Hv∞ (D) and is bounded, if and only if
$2π
sup |(Wn f )(θ )|dθ < ∞. (52)
n∈N
0
$2π
|(Wn f )(θ )|dθ → 0 as n → ∞. (53)
0
$2π
1
MWn f g(z) = Wn f (θ − ψ)g(reiψ )dψ, z = reiθ ∈ D.
2π
0
$2π
1
MWn f g(z) v(r) ≤ |(Wn f )(θ )|dθ gv (54)
2π
0
482 J. Taskinen and J. A. Virtanen
for all g ∈ h∞
v (D), Hence,
for all n and r, where the constant C > 0 is the supremum on the left-hand
side of (52). According to the remark concerning (50) the series on the right-hand
side of (51) converges uniformly on compact subsets of D, defines an element of
Hv∞ (D) and is bounded by gv . This means that Mf maps h∞ v (D) continuously
into Hv∞ (D).
As for the compactness of the operator Mf under the assumption (53), one takes
a sequence (gj )∞ ∞
j =1 contained in the closed unit ball of hv (D) and converging to
0 uniformly on compact subsets of D. One needs to show that Mf maps such a
sequence into a one converging to 0 with respect to the norm; see for example [26,
Section 2.4]. Roughly speaking, one can improve the boundedness proof to get this,
by using the assumption (53) together with the assumption on the convergence in
the compact-open topology. One needs a more sophisticated use of Theorem 6.
As usual, the proof for the necessity of the condition (52) for the boundedness
requires a careful enough choice of appropriate test functions. To this end we
fix an arbitrary 0 < ε < 1 as well as n ∈ N and ϕ ∈ [0, 2π]. Using
the
Fejer |k| approximation theorem we find a trigonometric polynomial g(z) =
k∈Z gk r eikθ , depending on n, ϕ and ε, such that
Wn f (ϕ − θ ) ε
g(rmn eiθ ) − < (55)
|Wn (ϕ − θ )|v(rmn ) v(rmn )
As a consequence,
$2π $2π
1 1
|(Wn f )(θ )|dθ = |(Wn f )(ϕ − θ )|dθ
2π 2π
0 0
$2π
1
≤ (Wn f )(ϕ − θ )g(rmn eiθ )dθ v(rmn ) + ε
2π
0
$2π
1
= f (ϕ − θ )(Wn g)(rmn eiθ )dθ v(rmn ) + ε
2π
0
Using Theorem 6 and (47), (56) we find a constant C > 0 such that
Wn gv ≤ c2 M∞ (Wn g, rmn )v(rmn ) ≤ c2 dM∞ (g, rmn )v(rmn ) ≤ 2Cc2 .
)
2π
Hence supn |(Wn f )(θ )|dθ < ∞.
0
The proof for the necessity of the condition (53) for the compactness of Mf
needs a number of additional technical details.
Since Riesz projection P , (39), is bounded by the assumptions of Theorem 7, it
follows that the boundedness and compactness of Mf : Hv∞ (D) → Hv∞ (D) are
also equivalent to (52) and (53), respectively.
Let us turn back to Toeplitz operators. Let Ta be a Toeplitz operator on Hv∞ (D)
with a given radial symbol a ∈ L1 (D), i.e. a(z) = a(|z|) for almost every z. Then,
defining
$1 ∞
1
γk = r 2k+1 v(r)a(r)dr, k ∈ N0 and fa (θ ) = γk eikθ , (58)
Γ2k
0 k=0
it was shown in (44)–(45) that Ta coincides with the Taylor multiplier with
coefficients (γk )∞
k=0 . The previous theorem thus yields the main result on the
boundedness and compactness.
Theorem 8 Let the weight satisfy (B). If a ∈ L1 is radial then Ta is bounded as an
operator Hv∞ (D) → Hv∞ (D) if and only if
$2π
sup |(Wn fa )(θ )|dθ < ∞, (59)
n
0
$2π
lim |(Wn fa )(θ )|dθ = 0. (60)
n→∞
0
We finally recall that Theorems 1.1 and 3.3 of the article [21] contain necessary
p
and sufficient conditions for the boundedness and compactness of Ta : Av (D) →
p
Av (D) for 1 < p < ∞, with minimal assumptions on the radial weights
v. However, the characterization is in terms of the boundedness of coefficient
multipliers in Hardy spaces, which is another open problem.
484 J. Taskinen and J. A. Virtanen
It is not difficult to see that (61) holds for example for the important classes of stan-
dard, normal and exponential weights. For normal weights, condition (61) with =
1/2 follows from Lemma 4.5. of [3]. In the case v(r) = exp(−1/(1−r)) it is known
)1
that 0 r m v(r)dr, m > 1, is proportional to the quantity m−3/4 exp(−Bm1/2) for
some constant B > 0 independent of m (see e.g. Lemma 2.2. in [7] or Lemma 4.28
in [1]). Hence, assuming < 1/2 we obtain
$1
r n−n v(r)dr ≤ C(n − n )3/4 exp(−B(n − n )1/2 )
0
$1
3/4
≤Cn exp(−Bn1/2
+C )≤C r n v(r)dr
0
−1 1/2 1 −1 −2
(n − n )1/2 = n1/2 (1 − n ) = n1/2 1 − n + O(n2 )
2
1 −1/2 −3/2
= n1/2 − n + O(n2 ) ≥ n1/2 − C
2
for all n. Thus, (61) holds. The same argument works for the more general weights
v(r) = exp(−α/(1 − r)β ), α, β > 0.
Boundedness of Toeplitz Operators in Bergman-Type Spaces 485
It was proven in [19] that normal and exponential weights satisfy (B).
Theorem 9 Let v satisfy (B) and (61) and assume that the symbol a ∈ L1 is real
valued and radial. The operator Ta is a bounded operator Hv∞ (D) → Hv∞ (D) in
any of the following cases:
(i) The restriction a|[δ,1[ is differentiable (with respect to r) for some δ ∈]0, 1[
and there holds
lim sup a (r) < ∞ or lim inf a (r) > −∞, (62)
r→1 r→1
and, in addition,
(ii) The restriction a|[δ,1[ is differentiable for some δ ∈]0, 1[, a satisfies (62) and,
for some constant C > 0, there holds the bound
C
|a (r)| ≤ 2 for r ∈]δ, 1[. (64)
(1 − r) log(1 − r)
Here, the case of complex valued symbols can be treated in the same way as in
the previous theorem.
The item (i) in both Theorems 9 and 10 follows from Theorem 8. We do not
present the proof but only refer to [5]. Recall that the coefficients of the series fa
)1
in (59), (60) are given in (58), which involves integrals 0 r n a(r)v(r)dr: the proofs
of (i) of Theorems 9 and 10 are based on quite technical estimates and calculations
with these expressions.
However, it is not so difficult to see that the sufficient condition (ii) essentially
implies (i) in Theorem 9. Assume a is real-valued and that (64) holds. For all r ∈
]δ, 1[ we get by the change of the integration variable log(1 − s) =: x and dx/ds =
−1/(1 − s) that
$1 $1 $
log(1−r)
1 1 C
|a (s)|ds ≤ C 2 ds = C dx = .(67)
(1 − s) log(1 − s) x2 | log(1 − r)|
r r −∞
$1
a(1) = a (s)ds + a(δ) = lim a(r) .
r→1
δ
$1
C
|a(r) − a(1)| = a (s)ds ≤ , (68)
| log(1 − r)|
r
which means that the function a −a(1) satisfies (63). Note that the Toeplitz operator
with the constant symbol a(1) is bounded as it is just a constant multiplier.
It is plain that (iii) implies (ii) in Theorem 9.
Also, as regards to Theorem 10, the assumptions in (ii) imply those of (i).
Namely, if (66) holds, then we can repeat the calculation (67)–(68) so that the
constant C is replaced by a positive function C(r) with C(r) → 0 as r → 1. Then,
we see from the analogue of (68) that the function a − a(1) even satisfies (65). If in
addition a(1) = 0 then also a satisfies (65). Note that if limr→1 a(r) = a(1) = 0,
then Ta is a compact perturbation of a non-zero multiple of the identity which is not
compact, and thus it cannot be a compact operator.
In [5] it is shown that if v is a normal weight, the assumptions on a in the previous
theorems can be relaxed, namely the boundedness of Ta : Hv∞ (D) → Hv∞ (D)
follows just from (63) and the compactness from (65) without any smoothness
Boundedness of Toeplitz Operators in Bergman-Type Spaces 487
assumptions on the symbol. Also, in the case of exponential weights v(r) = exp −
α/(1−r)β ), α, β > 0, the smoothness requirements on a can be dropped, namely, if
where the numbers γn are as in (45). The idea of trying to characterize the
p p
boundedness and compactness of Ta : Av (D) → Av (D) for 2 < p < ∞ (or
1 < p < 2) by interpolating does not seem to work, but one can derive a sufficient
p
condition similar to (52) for the boundedness of Ta in Av (D).
To formulate and sketch the proof of the result we need some modifications of
the notions that were used in the case of weighted sup-norms. We again assume
that the weight v satisfies condition (B). First, instead of the de la Valleé
Poisson
just to use the Dirichlet projections Qn g(z) = nk=0 gk zk for
operators it is enough
holomorphic g(z) = ∞ k
k=0 gk z . It is known that there are constants cp > 0 with
Mp (Qn g, r) ≤ cp Mp (g, r) for all 0 < r < 1, 1 < p < ∞, where cp does not
) 2π
depend on g, n or r and we write Mp (g, r)p = (2π)−1 0 |g(reiθ )|p dθ .
Analogously with the case of weighted sup-norms one picks up suitable increas-
ing numerical sequences (n )∞ ∞
n=1 with 1 = 0 and limn→∞ n = ∞ and (sn )n=1 ⊂
(0, 1) with limn→∞ sn = 1 and then defines the operators
Zn = Q[n+1 ] − Q[n ] , n ∈ N.
These are used to derive an equivalent form of the weighted Lp -norm: for some
p
constants c2 ≥ c1 > 0, for every f ∈ Av (D), there holds
∞
p p 1/p
c1 f p,v ≤ ωn Mp (Zn f, sn ) ≤ c2 f p,v , (72)
n=1
where the numbers ωn are determined by the weight. The details of the definitions
of the various parameters and proof of (72) can be found in [13] for p = 1 and in
[20] for 1 < p < ∞. Examples and calculations in concrete cases can be found in
488 J. Taskinen and J. A. Virtanen
α 1/β 1
n = β 1+1/β α −1/β n2+2/β − βn2 , sn = 1 − . (73)
β n2/β
Proposition
∞ 2 Let the weight satisfy (B), let a ∈ L1 be a radial function and let
fa (θ ) = ikθ be as in (45). Then the Toeplitz operator T : Ap (D) →
k=0 γk e a v
p
Av (D) is a well-defined, bounded operator, if
$2π
sup |(Zn fa )(θ )|dθ < ∞, (74)
n∈N
0
p p
and Ta : Av (D) → Av (D) is compact, if
$2π
|(Zn fa )(θ )|dθ → 0 as n → ∞. (75)
0
$2π
(Zn Mf g)(z) = (MZn f g)(z) = Zn f (θ − ψ)Zn g(reidψ )dψ,
0
$2π
Mp (Zn Mf g, r) ≤ |(Zn f )(θ )|dθ Mp (Zn g, r) (76)
0
The inequality Mf gp,v ≤ Cgp,v thus follows by applying (74) and (72) to
both Mf gp,v and gp,v , and this implies the boundedness of Ta .
Assume next (75) holds, and let (gj )∞
j =1 be a sequence which is contained in the
p
unit ball of Av (D) and which converges to 0 uniformly on compact subsets of D,
) 2π
and assume ε > 0 is given. We choose N ∈ N such that 0 |(Zn f )(θ )|dθ < ε.
The convergence of the sequence in the compact-open topology can be used to find
a large enough J ∈ N such that
ε ε
sup |Zn Mf gj (z)|v(z) < ⇒ Mp (Zn Mf gj , rmn ) <
|z|≤rmn 2πNωn Nωn
Boundedness of Toeplitz Operators in Bergman-Type Spaces 489
N ∞
p p p
Mf gj p,v ≤ ωn Mp (Zn Mf gj , rmn )p + ωn Mp (Zn Mf gj , rmn )p
n=1 n=N+1
∞
p p
≤ ε+ε ωn Mp (Zn gj , rmn )p ≤ 2εgj p,v ≤ 2ε.
n=N+1
Acknowledgments JT was supported in part by the Väisälä Foundation of the Finnish Academy
Sciences and Letters. JV was supported in part by the Engineering and Physical Sciences Research
Council grant EP/T008636/1
References
1. H. Arroussi, Function and operator theory on large Bergman spaces, Ph.D. thesis, Universitat
de Barcelona (2016)
2. J. Bonet, W. Lusky, J. Taskinen, Solid hulls and cores of weighted H ∞ -spaces. Rev. Mat.
Compl. 31, 781–804 (2018)
3. J. Bonet, W. Lusky, J. Taskinen, Solid hulls and solid cores of weighted Bergman spaces.
Banach J. Math. Anal. 13(2), 468–485 (2019)
4. J. Bonet, W. Lusky, J. Taskinen, On boundedness and compactness of Toeplitz operators in
weighted H ∞ -spaces. J. Funct. Anal. 278(10), 108456 (2020)
5. J. Bonet, W. Lusky, J. Taskinen, On the boundedness of Toeplitz operators with radial symbols
over weighted sup-norm spaces of holomorphic functions. J. Math. Anal. Appl. 493(1), 124515
(2021)
6. A. Dieudonne, Bounded and compact operators on the Bergman space L1a in the unit ball of
Cn . J. Math. Anal. Appl. 388(1), 344–360 (2012)
7. M. Dostanic, Unboundedness of the Bergman projections on Lp spaces with exponential
weights. Proc. Edinb. Math. Soc. 47, 111–117 (2004)
8. M. Engliš, Toeplitz operators and weighted Bergman kernels. J. Funct. Anal. 255(6), 1419–
1457 (2008)
9. S. Grudsky, N. Vasilevski, Bergman-Toeplitz operators: radial component influence. Integr. Eq.
Oper. Theory 40, 16–33 (2001)
10. S. Grudsky, A. Karapetyants, N. Vasilevski, Toeplitz operators on the unit ball in CN with
radial symbols. J. Oper. Theory 49, 325–346 (2003)
11. R. Hagger, J.A. Virtanen, Compact Hankel operators with bounded symbols. J. Oper. Theory
86, 317–329 (2021)
12. R. Hagger, C. Liu, J. Taskinen, J.A. Virtanen, Toeplitz operators on the unit ball with locally
integrable symbols. Integr. Equ. Oper. Theory 94, 17 (2022)
13. A. Harutyunyan, W. Lusky, On L1 −subspaces of holomorphic functions. Stud. Math. 198,
157–175 (2010)
14. A. Karapetyants, J. Taskinen, Toeplitz operators with radial symbols on general analytic
function spaces. Submitted
490 J. Taskinen and J. A. Virtanen
15. D.H. Luecking, Trace ideal criteria for Toeplitz operators. J. Funct. Anal. 73(2), 345–368
(1987)
16. D.H. Luecking, Finite rank Toeplitz operators on the Bergman space. Proc. Am. Math. Soc.
136(5), 1717–1723 (2008)
17. W. Lusky, On weighted spaces of harmonic and holomorphic functions. J. Lond. Math. Soc.
51, 309–320 (1995)
18. W. Lusky, Growth conditions for harmonic and holomorphic functions, Functional Analysis,
in Proceedings of the First International Workshop, ed. by S. Dierolf, S. Dineen, P. Domanski
(1996), pp. 281–291
19. W. Lusky, On the isomorphism classes of weighted spaces of harmonic and holomorphic
functions. Stud. Math. 175, 19–45 (2006)
20. W. Lusky, J. Taskinen, Bounded holomorphic projections for exponentially decreasing weights.
J. Funct. Spaces Appl. 6, 59–70 (2008)
21. W. Lusky, J. Taskinen, Toeplitz operators on Bergman spaces and Hardy multipliers. Stud.
Math. 204, 137–154 (2011)
22. P. Mannersalo, Toeplitz operators with locally integrable symbols on Bergman spaces of
bounded simply connected domains. Complex Var. Elliptic Eq. 61(6), 854–874 (2016)
23. P. Mannersalo, Toeplitz operators on Bergman spaces of polygonal domains. Proc. Edinb. Mat.
Soc. 62, 1115–1136 (2019)
24. G. McDonald, C. Sundberg, Toeplitz operators on the disc. Indiana Univ. Math. J. 28(4), 595–
611 (1979)
25. A. Perälä, J. Taskinen, J.A. Virtanen, Toeplitz operators with distributional symbols on
Bergman spaces. Proc. Edinb. Math. Soc. 54(2), 505–514 (2011)
26. J.H. Shapiro, Composition Operators and Classical Function Theory (Springer, New York,
1993)
27. K. Stroethoff, Compact Toeplitz operators on Bergman spaces. Math. Proc. Camb. Philos. Soc.
124, 151–160 (1998)
28. K. Stroethoff, D. Zheng, Toeplitz and Hankel operators on Bergman spaces. Trans. AMS
329(2), 773–794 (1992)
29. D. Suárez, The essential norm of operators in the Toeplitz algebra on Ap (Bn ). Indiana Univ.
Math. J. 56(5), 2185–2232 (2007)
30. J. Taskinen, J.A. Virtanen, Toeplitz operators on Bergman spaces with locally integrable
symbols. Rev. Math. Iberoamericana 26(2), 693–706 (2010)
31. J. Taskinen, J.A. Virtanen, New results and open problems on Toeplitz operators in Bergman
spaces. New York J. Math. 17, 147–164 (2011)
32. J. Taskinen, J.A. Virtanen, Weighted BMO and Toeplitz operators on the Bergman space A1 .
J. Oper. Theory 68(1), 131–140 (2012)
33. J. Taskinen, J.A. Virtanen, On generalized Toeplitz and little Hankel operators on Bergman
spaces. Arch. Math. 110(2), 155–166 (2018)
34. N.L. Vasilevski, Bergman type spaces on the unit disk and Toeplitz operators with radial
symbols, Reporte Interno 245, Departamento de Matemáticas, CINVESTAV del I.P.N., Mexico
City, 1999
35. N.L. Vasilevski, Commutative algebras of Toeplitz operators on the Bergman space, in
Operator Theory: Advances and Applications, vol. 185 (Birkhäuser, 2008)
36. F. Yan, D. Zheng, Bounded Toeplitz operators on Bergman space. Banach J. Math. Anal. 13(2),
386–406 (2019)
37. K. Zhu, Positive Toeplitz operators on weighted Bergman space. J. Oper. Theory 20, 329–357
(1988)
38. K. Zhu, Operator Theory in Function Spaces, Mathematical Surveys and Monographs, 2nd
edn., vol. 138 (American Mathematical Society, Providence, 2007)
39. N. Zorboska, Toeplitz operators with BMO symbols and the Berezin transform. Int. J. Math.
Math. Sci. 46, 2929–2945 (2003)
Part IV
Inequalities in Various Banach Spaces
Disjointness Preservers
and Banach-Stone Theorems
Research of the second author is supported by the Ministry of Education - Singapore, under its
Academic Research Fund Tier 1 (AcRF project no. RG24/19(S)).
D. H. Leung ()
Department of Mathematics, National University of Singapore, Singapore, Singapore
e-mail: [email protected]
W. K. Tang
School of Physical and Mathematical Sciences, Nanyang Technological University, Singapore,
Singapore
e-mail: [email protected]
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 493
R. M. Aron et al. (eds.), Operator and Norm Inequalities and Related Topics,
Trends in Mathematics, https://doi.org/10.1007/978-3-031-02104-6_15
494 D. H. Leung and W. K. Tang
1 Introduction
Let be a topological space. Denote by C() the vector space of all (real-valued)
continuous functions on . The space C() carries with it many structures. Indeed,
it is an algebra under pointwise addition and multiplication. It is also a vector lattice
under pointwise supremum and infimum. Finally, if is compact Hausdorff, the
space C() is a Banach space with the norm f = sup{|f (ω)| : ω ∈ }.
In the first half of the twentieth century, three remarkable theorems appeared that
characterize the space in terms of each of these structures of C().
Theorem 2.1 (Banach-Stone) Let , be compact Hausdorff spaces and let T :
C() → C() be a linear isometry. There are a homeomorphism ϕ : → and
a function h ∈ C() so that |h(σ )| = 1 for all σ ∈ and that
Theorem 2.1 was proved by Banach [9] for the case of compact metric spaces.
The theorem was extended to compact Hausdorff spaces by Stone [38].
Theorem 2.2 (Gelfand-Kolmogorov [24]) Let , be compact Hausdorff spaces
and let T : C() → C() be an algebra isomorphism. There is a homeomorphism
ϕ : → such that
Theorem 2.3 (Kaplansky [28]) Let , be compact Hausdorff spaces and let
T : C() → C() be a vector lattice isomorphism. There are a homeomorphism
ϕ : → and a function h ∈ C() so that h(σ ) > 0 for all σ ∈ and that
Theorem 2.3 is a special case of Kaplansky’s result. For a discussion of the result
in its full generality, see Sect. 4.1. In the intervening three quarters of a century, a
large number of extensions and generalizations of these results have been obtained.
A particularly fruitful concept that unifies the three classical theorems is that of
disjointness preserving operators. Let , be topological spaces. Two functions
f, g ∈ C(), respectively, C(), are said to be disjoint if the pointwise product
fg = 0. In terms of the lattice structure, f and g are disjoint if and only if
|f | ∧ |g| = 0. Suppose that A() and A() are vector subspaces of C() and
C() respectively. A linear operator T : A() → A() is disjointness preserving
496 D. H. Leung and W. K. Tang
This proves that Tf and T g are disjoint. Hence T is disjointness preserving. The
same applies to T −1 by symmetry.
The classical theorems of Banach-Stone, Gelfand-Kolmogorov and Kaplansky
can now be unified and extended by the next result.
Theorem 2.6 ([27]) Let and be compact Hausdorff spaces and let T :
C() → C() be a linear biseparating map. There are a homeomorphism
ϕ : → and a function h ∈ C() so that h(σ ) = 0 for all σ ∈ and
that
above. We will give a detailed proof of Theorem 2.6 that seems to us to be most
amenable to generalization. For the remainder of the section, let , and T be as
in Theorem 2.6. For any function f ∈ C(), the support of f , supp f , is the closure
of the set {ω : f (ω) = 0}. Similarly for functions in C().
Proposition 2.7 ([5, Lemma 4]) If f, g ∈ C() and supp f ⊆ supp g, then
supp Tf ⊆ supp T g.
Proof Otherwise, there are f, g ∈ C() with supp f ⊆ supp g, yet supp Tf ⊆
supp T g. Hence there exists σ0 ∈/ supp T g so that Tf (σ0 ) = 0. Choose h ∈ C() so
that h(σ0 ) = 0 and that h is disjoint from T g. Since T −1 is disjointness preserving,
T −1 h and g are disjoint. Thus T −1 h and f are disjoint. Therefore, h and Tf are
disjoint, which contradicts the fact that h and Tf are both nonzero at σ0 .
For any σ ∈ , set
1 1
Un = |f |−1 ( , ), n ∈ N.
(3n + 5) (3n + 1)2
2
H H H
Then ω ∈ n Un = n U2n−1 ∪ n U2n . Without loss of generality, assume that
H
ω ∈ n U2n−1 . Set
1 1
Vn = |f |−1 ( , ), n ∈ N.
(6n + 3) (6n − 3)2
2
Then (Vn ) is a sequence of disjoint open sets so that U2n−1 ⊆ Vn for all n. For
each n, choose a function hn ∈ C() so that 0 ≤ hn ≤ 1, hn = 1 on U2n−1
and hn = 0 outside Vn . The sequence of functions (nhn f )n is pairwise disjoint
and nhn f ≤ (6n−3)n
2 → 0. Hence the sum g := nhn f converges in C().
For each n, g − nf = 0 on the set U2n−1 . By definition of ϕ, this H implies that
T (g − nf )(σ ) = 0 for all σ ∈ ϕ −1 (U2n−1 ). Choose a net (ωα ) in n U2n−1 that
converges to ω. Let nα ∈ N be such that ωα ∈ U2nα −1 . Set σα = ϕ −1 (ωα ). Since
Banach-Stone Theorems 499
f (ω) = 0, ω ∈
/ Un for any n. Thus limα nα = ∞. Note that (σα ) converges to σ .
Therefore,
1
Tf (σ ) = lim Tf (σα ) = lim T g(σα ) = 0,
α α nα
Results in Sect. 2 may serve to convince the reader that biseparating maps are worthy
of study in their own right. Indeed, plenty of results concerning biseparating maps
have been obtained in the past 30 years or so. Most of these are in the context of
linear or at least additive maps. Since surjective additive maps between vector spaces
are linear maps over the field of rational numbers, the results remain mainly “linear”
in character. Very recently, several papers [15, 17, 18] appeared that took the study of
isomorphisms of disjointness structure, or ⊥-isomorphisms, to very general settings.
It is shown that even for function spaces with minimal structure, analysis of ⊥-
isomorphisms can still bear fruitful results. The aim of this section is to describe this
general approach to ⊥-isomorphisms. Earlier results on biseparating maps, in spaces
of (vector-valued) continuous functions, uniformly continuous functions, Lipschitz
functions and differentiable functions, will be seen as consequences. Applications
to theorems of Banach-Stone type will be considered in the next section.
3.1 ⊥-Isomorphisms
Let , X be Hausdorff topological spaces and let A(, X) be a subset of C(, X),
the set of continuous functions f : → X. For f, g ∈ A(, X), let
Weak regularity is a basic assumption to ensure that there are sufficient functions
in A(, X) and A(, Y ) to yield a nontrivial theory. The following simple yet
important result is noted and used in [15]. Its ancestry can be traced back to at
least [5, Lemma 4].
Proposition 3.1 Let T : A(, X) → A(, Y ) be a bijection, where A(, X) and
A(, Y ) are h- and T h-weakly regular respectively. Then T is a ⊥h -isomorphism
if and only if it is a ⊆h -isomorphism.
A set U in is a regular open set if U = int U . All sets of the form σh (f ) are
regular open sets. Denote the collection of all regular open sets in by RO().
RO() is a Boolean algebra with 0 = ∅, 1 = , lattice operations U ∧ V = U ∩ V ,
U ∨ V = int U ∪ V and negation ¬U = int(\U ). See [38]. If is a regular
topological space, then RO() is a basis for the topology on .
Let T : A(, X) → A(, Y ) be a ⊆h -isomorphism, where A(, X) and
A(, Y ) are h- and T h-weakly regular respectively. Define a map θh : RO() →
RO() by
C
θh (U ) = int {σT h (Tf ) : f ∈ A(, X), σh (f ) ⊆ U }. (1)
By (1) and the fact that T is a ⊆-isomorphism, θh2 (σh2 (f )) = σT h2 (Tf ). Then
∅ = σh2 (f ) = θh−1
2
(σT h2 (Tf )) ⊆ U . Hence σh2 (f ) is a nonempty set disjoint from
¬U . By condition (L), there exist g ∈ A(, X) and a nonempty set V ∈ RO(),
V ⊆ σh2 (f ), so that
h1 on ¬U ,
g=
f on V .
Now
1. θf (V ) ⊆ θf (σh2 (f )) = θf (σf (h2 )) = σTf (T h2 ) = σT h2 (Tf ).
2. T g = T h1 on θh1 (¬U ) = ¬θh1 (U ) = int[θh1 (U )c ] ⊇ σT h2 (Tf ).
3. T g = Tf on θf (V ).
4. T h1 = T h2 on θh2 (U ) ⊇ σT h2 (Tf ).
502 D. H. Leung and W. K. Tang
Here, we have applied Theorem 3.2 for items 2–4. It follows that Tf = T h2 on
θf (V ). However, θf (V ) is a nonempty open subset of σT h2 (Tf ) = int [Tf = T h2 ].
So we have reached a contradiction. Therefore, θh2 (U ) ⊆ θh1 (U ). Since θh1 (U ) is
a regular open set, θh2 (U ) ⊆ θh1 (U ). The lemma follows by symmetry.
Proof of Theorem 3.3 Taking into account Theorem 3.2, it suffices to show that
θh1 = θh2 for any h1 , h2 ∈ A(, X). Suppose that there exists U ∈ RO() so that
θh2 (U ) ⊆ θh1 (U ), so that in fact θh2 (U ) ⊆ θh1 (U ). By condition (L) for A(, Y ),
there exists g ∈ A(, Y ) and a nonempty set V ∈ RO(), V ⊆ θh2 (U )\θh1 (U ), so
that
T h1 on θh1 (U ),
g=
T h2 on V .
U ∩ θh−1
2
(V ) = θg (θh1 (V )) ∩ θg (V ) = ∅.
However, θh−1
2
(V ) is a nonempty subset of θh−1 2
(θh2 (U )) = U . The contradiction
shows that θh2 (U ) ⊆ θh1 (U ). The reverse inclusion follows by symmetry.
In Theorem 3.3, say that θ is associated with T . We list a few examples of sets of
functions satisfying condition (L). Another example is given in Lemma 4.2 below.
Example (a) Let be a completely regular Hausdorff space and let X be a convex
set in a Hausdorff topological vector space. The space C(, X) consists of all
continuous functions from into X.
(b) Let be a metric space and let X be a convex set in a normed space. Denote by
U (, X), U∗ (, X), Lip(, X), Lip∗ (, X), respectively, the set of uniformly
continuous functions, the set of bounded uniformly continuous functions, the
set of Lipschitz functions and the set of bounded Lipschitz functions from to
X.
(c) Let be an open set in a Banach space Z and let X be a Banach space. For p ∈
N ∪ {∞}, denote by C p (, X ) the space of all p-times continuously (Fréchet)
differentiable X -valued functions on . To ensure that there are “sufficiently
many” functions in C p (, X ), we assume that there exists a bump function in
C p (Z), i.e., a function ξ ∈ C p (Z) that has nonempty bounded support in Z.
To see that all of the spaces A(, X) above satisfy condition (L), first observe
that if ω0 ∈ , U ∈ RO() and ω0 ∈ / U , then there exist ξ : → [0, 1], V ∈
RO(X) containing ω0 so that ξ = 0 on V and ξ = 1 on U . Moreover, for the
situation in (b), we can choose ξ to be Lipschitz, and for case (c), we can choose
Banach-Stone Theorems 503
ξ ∈ C p (). Given h1 , h2 ∈ A(, X), it is easy to verify that f (ω) = ξ(ω)h1 (ω) +
(1−ξ(ω))h2 (ω) defines a function in A(, X) that is equal to h1 on U and h2 on V .
Remark 3.5 Condition (L) is a condition on A(, X), respectively, A(, Y ), that
guarantees that the Boolean isomorphisms θh are independent of h. Alternatively,
we may impose conditions on T to warrant the same outcome. For example,
if X, Y are Hausdorff topological groups, then C(, X) is a topological group
under pointwise group operations. Suppose that A(, X), A(, Y ) are subgroups
of C(, X), C(, Y ) respectively and T : A(, X) → A(, Y ) is a group
isomorphism as well as a ⊥h isomorphism for some h ∈ A(, X). Then routine
verification shows that T is a ⊥-isomorphism and for any k ∈ A(, X), θh (U ) =
θk (U ) for all U ∈ RO(X). In particular, the situation occurs if X and Y are
Hausdorff topological vector spaces, A(, X), A(, Y ) are respective subspaces of
C(, X), C(, Y ), and T : A(, X) → A(, Y ) is an additive ⊥0 -isomorphism.
3.3 Representation
In many cases, it is possible to improve Theorems 3.7 and 3.8 by giving a functional
representation of the ⊥-isomorphism T .
Proposition 3.9 Let , , X, Y be Hausdorff spaces. Suppose that T :
A(, X) → A(, Y ) is a bijection, where A(, X), A(, Y ) are subsets of
C(, X) and C(, Y ) respectively. Assume that there is a homeomorphism
ψ : → that is associated with T . If f, g ∈ A(, X), ω0 ∈
and there exists h ∈ A(, X) so that ω0 ∈ int [h = f ] ∩ int [h = g], then
Tf (ψ(ω0 )) = T g(ψ(ω0 )).
Proof Let U = int[h = f ] and V = int[h = g]. By assumption, T h = Tf
on ψ(U ) and T h = Tf on ψ(V ). Since ψ is a homeomorphism and ω0 ∈
U ∩ V , ψ(ω0 ) ∈ ψ(U ) ∩ ψ(V ). By continuity of Tf, T g and T h, Tf (ψ(ω0 )) =
T h(ψ(ω0 )) = T g(ψ(ω0 )).
From the example following Theorem 3.3, the spaces listed there all satisfy
condition (L).
Theorem 3.10 Let , be a first countable compact Hausdorff topological space
and let X, Y be convex sets in Hausdorff topological vector spaces, with X, Y
containing more than one point. If T : C(, X) → C(, Y ) is a ⊥-isomorphism,
then there are a homeomorphism ψ : → and a function : × X → Y so
that
Sketch of Proof As mentioned above, C(, X) and C(, Y ) both satisfy condition
(L). Hence there is a homeomorphism ψ : → associated with T by
Theorem 3.7. For any x ∈ X, let gx ∈ C(, X) be the constant function with
value x. Define : × X → Y by (ω, x) = (T gx )(ψ(ω)). Let ω0 ∈ and
506 D. H. Leung and W. K. Tang
f ∈ C(, X). Set x = f (ω0 ). Using the first countability of , one can easily
construct h ∈ C(, X) so that ω0 ∈ int [h = f ] ∩ int [h = gx ]. By Proposition 3.9,
4 Applications
If A(, X) and A(, Y ) are lattices (in the pointwise order), then an order
isomorphism is a lattice isomorphism. Following [28], we say that A(, X) is
X-normal if for any disjoint closed sets F1 , F2 in and any x1 , x2 ∈ X, there
exists f ∈ A(, X) so that f = xi on Fi , i = 1, 2. The following statement is
Kaplansky’s Theorem in its full generality.
Theorem 4.1 (Kaplansky [28]) Let , be compact Hausdorff spaces and let
X, Y be totally ordered sets with the order topology. If C(, X) and C(, Y ) are X-
and Y -normal respectively and there exists an order isomorphism T : C(, X) →
C(, Y ), then and are homeomorphic.
We will see that Kaplansky’s Theorem as well as similar results on other function
spaces can be derived from considerations in Sect. 3. A function f ∈ C(, X) is
bounded if there are x1 , x2 ∈ X so that x1 ≤ f (ω) ≤ x2 for all ω ∈ . Clearly,
if is compact Hausdorff, or if X has both largest and smallest elements, then all
functions in C(, X) are bounded.
Lemma 4.2 Let be a regular topological space and let X be a totally ordered set
with the order topology. Suppose that A(, X) is a X-normal sublattice of C(, X)
that consists of bounded functions. Then A(X, E) satisfies condition (L).
Proof Let h1 , h2 ∈ A(, X), U ∈ RO() and ω ∈ / U . Since is regular, there
exists V ∈ RO() containing ω so that V ∩ U = ∅. There are x1 , x2 ∈ X so that
x1 ≤ h1 (ω), h2 (ω) ≤ x2 for all ω ∈ . By X-normality, there are k1 , k2 ∈ A(, X)
so that
x2 on V x1 on V
k1 = and k2 = .
x1 on U x2 on U
σT g (T l) ∩ σT g (k) ⊆ σT g (T l) ∩ U = σT g (T l) ∩ σT g (T h) = ∅.
Linear and nonlinear lattice and order isomorphisms have been well studied in a
variety of function spaces. Garrido and Jaramillo [20] showed that the unital vector
lattices U () and U∗ () determine up to uniform homeomorphism. In [21], the
same authors showed that as a unital vector lattice, Lip() determines up to
Lipschitz homeomorphism. For Lipschitz spaces defined on Banach spaces, F. and J.
Cabello Sánchez showed that Lip(R) and Lip∗ (R) are isomorphic as vector lattices.
However, if X is a Banach space of dimension > 1 and Y is a Banach space, then
Lip∗ (X ) is not isomorphic as a vector lattice to Lip(Y) [14].
As a lattice alone (i.e., disregarding linearity), Shirota [37] proved that if U∗ ()
and U∗ () are lattice isomorphic, with , complete metric spaces, then is
uniformly homeomorphic to . In the same paper, the claim was also made for
lattice isomorphisms T : U () → U (); but the proof contains a gap. The gap
was repaired by F. Cabello Sánchez [11] and F. and J. Cabello Sánchez [13]. The
same authors also showed that if T : C p () → C p () is an order isomorphism,
where p ∈ N∪{∞} and , are manifolds modeled on Banach spaces that support
C p -bump functions, then and are homeomorphic [12]. A unified treatment of
order isomorphisms between functions spaces can be found in [32].
One can also consider the situation for Theorem 4.4 away from the confines
of compact Hausdorff spaces. A completely regular Hausdorff space has a
“largest” compactification, the Stone-Čech compactification β, characterized by
the fact that every continuous function f from into a compact Hausdorff
space X has a unique continuous extension fB : β → X. A good source of
information concerning the Stone-Čech compactification is [39]. For the purpose
of extending the Gelfand-Kolmogorov Theorem, Hewitt [26] introduced the class of
realcompact spaces. Let R∞ be the one point compactification of R. The (Hewitt)
realcompactification υ consists of all ω0 ∈ β such that for any continuous
real-valued function f on , its continuous extension fB : β → R∞ satisfies
fB(ω0 ) ∈ R. is realcompact if = υ. It is known that a space is realcompact if
and only if it is homeomorphic to a closed subspace of R for some index set ; see,
e.g., [22]. Hewitt showed that for realcompact spaces, C() as a ring determines
uniquely up to homeomorphism. The result was generalized by Araujo et al. [5],
510 D. H. Leung and W. K. Tang
and subsequently by Araujo to vector valued functions [1, 2]. If X and Y are vector
spaces, denote the set of all linear bijections from X onto Y by I (X , Y).
Theorem 4.6 ([2]) Let and be realcompact spaces and let X and Y be normed
spaces. If T : C(, X ) → C(, Y) is a linear biseparating map (i.e., linear ⊥-
isomorphism), then there are a homeomorphism ϕ : → and a function J :
→ I (X , Y) so that
Let be a Hausdorff topological space. The space C() of all real valued
continuous functions on is a (unital) ring under pointwise operations. The subring
C∗ () consists of the bounded functions. Clearly, the (pointwise) order on these
rings is determined by the ring structure since f ≥ 0 if and only if f is a square in
the ring. It follows immediately that results from Sect. 4.1 give rise to corresponding
results concerning ring isomorphisms. In particular, we cite Hewitt’s generalization
of the theorem of Gelfand-Kolmogorov as a consequence of Theorem 4.8.
Theorem 4.9 ([26]) Let and be realcompact spaces. Suppose that C() and
C() are isomorphic as rings, then and are homeomorphic.
If is a complete metric space, then Lip∗ () is a ring under pointwise
operations. Ring isomorphisms between such rings were described in [21]. Let
, be metric spaces. A function ψ : → is Lipschitz in the small if there
exist r, K > 0 so that d(ψ(ω1 ), ψ(ω2 )) ≤ Kd(ω1 , ω2 ) whenever ω1 , ω2 ∈ and
d(ω1 , ω2 ) < r. ψ is a LS-homeomorphism if it is a homeomorphism so that both ψ
and ψ −1 are Lipschitz in the small.
Banach-Stone Theorems 511
Theorem 4.10 (Garrido and Jaramillo) Let , be complete metric spaces. The
following are equivalent.
1. Lip∗ () and Lip∗ () are isomorphic as unital rings.
2. Lip∗ () and Lip∗ () are isomorphic as unital vector lattices.
3. and are LS-homeomorphic.
Garrido et al. showed that the ring of smooth functions C ∞ (M) determines
the manifold M up to smooth diffeomorphism. For notions and notation regarding
global analysis on infinite dimensional manifolds, refer to [29].
Theorem 4.11 (Garrido et al. [23]) Let M and N be paracompact Banach
manifolds modeled on C ∞ -smooth Banach spaces. The rings C ∞ (M) and C ∞ (N)
are isomorphic if and only if M and N are C ∞ -diffeomorphic.
Instead of ring isomorphisms, one can disregard linearity and consider maps that
preserve multiplication alone.
Proposition 4.12 Let , be Hausdorff spaces and let A(), A() be unital
subrings of C() and C() respectively. Assume that either
1. and are compact and A(), A() satisfy condition (L); or
2. and are complete metric spaces and A(), A() satisfy conditions (L) and
(Ls ).
If T : A() → A() is a multiplicative isomorphism, i.e., T is a bijection so that
T (fg) = Tf · T g for all f, g ∈ A(), then there is a homeomorphism ψ : →
that is associated with T in the sense defined before Theorem 3.7. In particular, T
is a ⊥-isomorphism.
Proof By assumption, the constant functions belong to A() and A(). Let Let
0, 2 denote the constant functions with values 0, 2 respectively. Then
2 · T 0 = T (T −1 2 · 0) = T 0.
f ⊥0 g ⇐⇒ fg = 0 ⇐⇒ Tf · T g = T 0 = 0 ⇐⇒ Tf ⊥T 0 T g.
4.4 Isometry
The study of isometries is probably the most well developed part among theorems
of Banach-Stone type. Here we restrict ourselves to a much abridged survey. Further
information can be found in the two-volume monograph [19].
Behrends [10] introduced the use of centralizers into Banach-Stone considera-
tions. Let X be a Banach space and denote the set of extreme points of the ball
in X ∗ by ext X ∗ . A bounded linear operator S : X → X is a multiplier if every
x ∗ ∈ ext X ∗ is an eigenvector of S ∗ , i.e., S ∗ x ∗ = aS (x ∗ )x ∗ for some scalar aS (x ∗ ).
If R, S are multipliers, say that R is an adjoint of S if aR (x∗) = aS (x ∗ ) for all
x ∗ ∈ X ∗ . The centralizer Z(X ) of X consists of all multipliers S for which an
adjoint exists. Note that for real Banach spaces, the centralizer is the same as the
set of all multipliers. Multiples of the identity operator are always present in the
centralizer. Say that X has trivial centralizer if there are no other operators in Z(X ).
Many classes of Banach spaces have trivial centralizers; refer to [10, 19].
Theorem 4.14 (Behrends) Let X and Y be Banach spaces which have trivial
centralizers. Suppose further that and are locally compact Hausdorff spaces
and that there exists a surjective linear isometry T : C0 (, X ) → C0 (, Y). Then
Banach-Stone Theorems 513
where
f (ω1 ) − f (ω2 )
f ∞ = sup f (ω) and L(f ) = sup .
ω∈ ω1 =ω2 d(ω1 , ω2 )
Araujo and Dubarbie [6] gave a complete description of isometries between vector-
valued spaces of Lipschitz functions. We state a special case of their result here.
Define an equivalence relation on by x ∼ y if there are x = x1 , . . . , xn = y
in so that d(xi , xi+1 ) < 2, 1 ≤ i < n. The equivalence classes are called the
2-components of .
Theorem 4.16 Let , be complete metric spaces and let X , Y be Banach spaces.
Assume that T : Lip(, X ) → Lip(, Y) is a surjective linear isometry so that for
all σ ∈ , there is a constant function f ∈ Lip(, X ) so that Tf (σ ) = 0. Then
514 D. H. Leung and W. K. Tang
In this part, assume that , are regular topological spaces and X, Y are Hausdorff
spaces. Let A(, X) and A(, Y ) be subsets of C(, X) and C(, Y ) respectively.
Given n ∈ N and h ∈ A(, X), a bijection T : A(, X) → A(, Y ) is a ∩nh -
isomorphism if for any f1 , . . . , fn ∈ A(, X),
D
n D
n
[fi = h] = ∅ ⇐⇒ [Tfi = T h] = ∅.
i=1 i=1
by Hernández and Ródenas [25]. Further results were given in [16, 31].
Proposition 4.17 Let , be regular topological spaces. Suppose that A(, X)
and A(, Y ) satisfy condition (L) and that there exists k ∈ A(, X) so that
[k = h] = ∅. If T : A(, X) → A(, Y ) is a ∩2h -isomorphism, then it is a
⊥h -isomorphism.
Proof First of all, since T is a ∩2h -isomorphism, it is a ∩1h -isomorphism. Thus [k =
h] = ∅ implies [T k = T h] = ∅. Suppose that there are f, g ∈ A(X, E) so that
f ⊥h g but Tf ⊥T h T g. There exists σ0 ∈ where Tf (σ0 ), T g(σ0 ) = T h(σ0 ).
Since is a regular topological space, there exists V ∈ RO() containing σ0 so
that V ⊆ [Tf = T h] ∩ [T g = T h]. Then σ0 ∈ / int V c and int V c ∈ RO(). As
A(, Y ) satisfies condition (L), there exist l ∈ A(, Y ) and W ∈ RO(Y ) so that
σ0 ∈ W , l = T k on int V c and l = T h on W . Now
[Tf = T h] ∪ [l = T h] ⊇ V ∪ int V c = .
Banach-Stone Theorems 515
3
∈ [f = 1 − h] ∩ [g = 1 − h] = [Tf = T h] ∩ [T g = T h].
4
g ≤ f ≤ h #⇒ T g ≤ Tf ≤ T h or
g ≤ f ≤ h #⇒ T h ≤ Tf ≤ T g
In the first case, we say that T is order preserving with respect to f and in the
second case, T is anti-order preserving with respect to f .
Banach-Stone Theorems 517
References
16. L. Dubarbie, Maps preserving common zeros between subspaces of vector-valued continuous
functions. Positivity 14, 695–703 (2010)
17. X. Feng, Nonlinear biseparating operators on vector-valued function spaces, PhD thesis,
National University of Singapore, 2018
18. X. Feng, D.H. Leung, Nonlinear biseparating maps. Preprint (2020). arXiv:2009.11570
19. R.J. Fleming, J.E. Jamison, Isometries on Banach Spaces, vols. 1 and 2 (Chapman & Hall,
2007)
20. M.I. Garrido, J.A. Jaramillo, A Banach-Stone theorem for uniformly continuous functions.
Monatsh. Math. 131, 189–192 (2000)
21. M.I. Garrido, J.A. Jaramillo, Homomorphisms on function lattices. Monatsh. Math. 141, 127–
146 (2004)
22. M.I. Garrido, J.A. Jaramillo, Variations on the Banach-Stone theorem. Extracta Math. 17, 351–
383 (2002)
23. M.I. Garrido, J.A. Jaramillo, J.A. Prieto, Banach-Stone theorems for Banach manifolds. Rev.
Real Acad. Cienc. Exact. Fis. Natur. Madrid 94, 525–528 (2000)
24. I. Gelfand, A.N. Kolmogoroff, On rings of continuous functions on a topological space. C. R.
(Doklady) URSS 22, 11–15 (1939)
25. S. Hernández, A.M. Ródenas, Automatic continuity and representation of group homomor-
phisms defined between groups of continuous functions. Topology Appl. 154, 2089–2098
(2007)
26. E. Hewitt, Rings of real-valued continuous functions, I. Trans. Am. Math. Soc. 64, 54–99
(1948)
27. K. Jarosz, Automatic continuity of separating linear isomorphisms. Canad. Math. Bull. 33,
139–144 (1990)
28. I. Kaplansky, Lattices of continuous functions. Bull. Am. Math. Soc. 53, 617–623 (1947)
29. A. Kriegl, P.W. Michor, The Convenient Setting of Global Analysis. AMS Mathematical
Surveys and Monographs, vol. 53 (AMS, Providence, 1997)
30. H. König, V. Milman, Operator functional equations in analysis, in Asymptotic Geometric
Analysis, ed. by M. Ludwig et al., Fields Institute Communications, vol. 68, 189–209 (2013)
31. D.H. Leung, W.-K. Tang, Banach–Stone Theorems for maps preserving common zeros.
Positivity 14, 17–42 (2010)
32. D.H. Leung, W.-K. Tang, Nonlinear order isomorphisms on function spaces. Dissertationes
Math. 517, 1–75 (2016)
33. L. Li, N.-C. Wong, Kaplansky theorem for completely regular spaces. Proc. Am. Math. Soc.
142, 1381–1389 (2014)
34. L. Li, N.-C. Wong, Banach-Stone theorems for vector valued functions on completely regular
spaces. J. Math. Anal. Appl. 395, 265–274 (2012)
35. A.N. Milgram, Multiplicative semigroups of continuous functions. Duke Math. J. 16, 377–383
(1940)
36. J. Mrčun, P. Šemrl, Multiplicative bijections between algebras of differentiable functions. Ann.
Acad. Sci. Fenn. Math. 32, 471–480 (2007)
37. T. Shirota, A generalization of a theorem of I. Kaplansky. Osaka Math. J. 4, 121–132 (1952)
38. M.H. Stone, Applications of the theory of Boolean rings to general topology. Trans. Am. Math.
Soc. 41(3), 375–481 (1937)
39. R.C. Walker, The Stone-Čech Compactification (Springer, 1974)
40. N. Weaver, Isometries of noncompact Lipschitz spaces. Canad. Math. Bull. 38, 242–249 (1995)
The Bishop–Phelps–Bollobás Theorem:
An Overview
Abstract In this survey, we provide an overview from 2008 to 2021 about the
Bishop–Phelps–Bollobás theorem.
S. Dantas
Departament de Matemàtiques and Institut Universitari de Matemàtiques i Aplicacions de
Castelló (IMAC), Universitat Jaume I, Castelló, Spain
e-mail: [email protected]
D. García · M. Maestre () · Ó. Roldán
Departamento de Análisis Matemático, Facultad de Ciencias Matemáticas, Universidad de
Valencia, Burjasot, Valencia, Spain
e-mail: [email protected]; [email protected]; [email protected]
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 519
R. M. Aron et al. (eds.), Operator and Norm Inequalities and Related Topics,
Trends in Mathematics, https://doi.org/10.1007/978-3-031-02104-6_16
520 S. Dantas et al.
space. The space C0 (L) is the space of all real or complex continuous functions
defined on L with limit zero at infinity. We denote by H ∞ (D) the algebra of all
bounded analytic functions on the open unit disc D in the complex place C. For
a complex Banach space X , we denote by Au (BX ; Y) (respectively, Ab (BX ; Y))
the set of all Y-valued uniformly (respectively, bounded) continuous functions on
BX that are holomorphic on the interior of BX . The symbol Aw∗ u (BX ) stands for
the unital algebra of all w∗ -uniformly continuous functions from BX into C which
are holomorphic on the interior of BX endowed with the supremum norm · ∞ .
Throughout Sect. 2.6, M will denote a pointed metric space, that is, a metric space
with a distinguished point 0 ∈ M and such that M\{0} = ∅, and d will denote the
metric of M. If Y is a real Banach space, then Lip0 (M, Y) stands for the space of all
Lipschitz mappings f : M → Y with f (0) = 0 equipped with the Lipschitz norm.
If Y = R, we may omit Y in the notation and simply write Lip0 (M). Finally, F (M)
denotes the Lipschitz-free space associated to M (we refer the reader to the survey
[111] and references therein for a solid background in Lipschitz-free spaces). We
denote by P(N X ; Y) the Banach space of all N-homogeneous polynomials from X
into Y .
This survey is mainly motivated by two results due to three mathematicians:
the Americans Errett Albert Bishop (1928–1983) and Robert Ralph Phelps (1926–
2013), and the Hungarian Béla Bollobás (1943-). In fact, the whole story initiates
with Robert Clarke James (1918–2003) who provided one of the most famous
characterizations for reflexive spaces in Banach space theory known nowadays
as the James theorem (see [117, 118]). In fact, in any first course in Functional
Analysis, one is able to construct easily a bounded linear functional which never
attains its norm and this opens the gate for a very natural question: when does a
linear functional attain its norm? James, with outstanding techniques, proved that a
Banach space X is reflexive if and only if every bounded linear functional attains its
norm. It is not difficult to prove by using the Hahn-Banach theorem that when X is
reflexive, every functional attains its norm. The real deal with the James theorem is
of course the converse.
Since James characterized reflexive Banach spaces as those where every func-
tional attained its norm, the concept of subreflexivity arose to denote the normed
spaces for which the set of norm-attaining functionals is dense in the dual. Although
there exist incomplete normed spaces which are not subreflexive (see [117]), in
1961, Bishop and Phelps showed that every Banach space is subreflexive; in other
words, they proved that in any Banach space X , given a functional x ∗ ∈ X and
an arbitrary positive number ε > 0, it is always possible to find a new functional
y ∗ ∈ X which attains its norm and satisfies y ∗ − x ∗ < ε (see [41]). This means
that, for every Banach space X , the set NA(X ) is norm dense in the dual space X .
Now, in order to be fair with the title of this survey, it remains to fit Bollobás’
name somehow. This is so because Bollobás proved a strengthening of the Bishop–
Phelps’ result known nowadays as the Bishop–Phelps–Bollobás theorem [42]. The
original Bollobás’ result states the following.
The Bishop–Phelps–Bollobás Theorem: An Overview 521
Theorem 1.1 ([42, Theorem 1]) Let X be a Banach space. Suppose that x ∈ SX
and x ∗ ∈ SX satisfy
ε2
|x ∗ (x) − 1| ≤ ,
2
First of all, let us notice that it is clear that the Bishop–Phelps theorem is a particular
case of the Bollobás’ theorem. Indeed, Theorem 1.1 above contains much more
information than the denseness of the functionals which attains their norms: it gives
a simultaneous control of the involved points and functionals in a quantitative way
and that is the great difference between these two theorems. At this point, it is
worth mentioning the recent paper [63] where the authors sought the best possible
constants in the Bollobás theorem (see, in particular, [63, Theorem 2.1]). We will
be referring to the next result (extracted from [63, Corollary 2.4]) as the Bishop–
Phelps–Bollobás Theorem, which is the sharpest version of Theorem 1.1 (see Sect. 3
for more information).
Theorem 1.2 (Bishop–Phelps–Bollobás Theorem) Let X be a Banach space. Let
ε ∈ (0, 2) and suppose that x ∈ BX and x ∗ ∈ BX satisfy
ε2
Re x ∗ (x) > 1 − .
2
norms. For instance, this happens when X is reflexive (see [137, Theorem 1]) or
when Y has the property β of Lindenstrauss (see [137, Proposition 3]). In other
words, putting some extra conditions in the involved spaces, we can get versions of
the Bishop–Phelps theorem for operators.
As a careful and curious reader may imagine, after Lindenstrauss’ paper, a vast
research on the topic has been done during the past sixty years in several directions.
We would like to name just a few of them which provided a great impact on
extensions of the Bishop–Phelps theorem: J. Bourgain, R. E. Huff, J. Johnson,
W. Schachermayer, J. J. Uhl, J. Wolfe, and V. Zizler continued the study on the
set of operators which attain their norms ([43, 116, 120, 144, 148, 151]); M. D.
Acosta, R. M. Aron, F. J. Aguirre, Y. S. Choi, and R. Payá ([8, 35, 70]) considered
some problems in the same line involving bilinear mappings; the second and the
third authors of the present survey considered it for homogeneous polynomials
(see [21, 36]); and more recently several problems on norm-attainment of Lipschitz
mappings were also tackled (see for instance [54, 64, 111, 121]). We suggest the
interested reader to check the excellent survey [4] from María Dolores Acosta to
know more about the norm-attaining theory (up to 2006).
In this survey, we will be interested in (generalizations of) the Bishop–Phelps–
Bollobás theorem. Let us notice that so far we have mentioned only possible
versions of the Bishop–Phelps theorem for operators and nothing related to the
Bollobás theorem for this class of functions was considered yet. The first time this
question was addressed and systematically studied was in 2008, when the authors
from [9] considered Theorem 1.2 for operators and provided several conditions
under which this theorem holds in this more general manner.
The next definition is the main one of the present work and it is exactly what we
will be discussing throughout the next sections.
Definition 1.3 (Bishop–Phelps–Bollobás Property) Let X and Y be Banach
spaces. We say that the pair (X , Y) has the Bishop–Phelps–Bollobás property for
operators (BPBp, for short) if given ε > 0, there exists η(ε) > 0 such that whenever
T ∈ B(X , Y) with T = 1 and x ∈ SX satisfy
We notice that the Banach spaces c0 () and ∞ () satisfy property β of
Lindenstrauss with ρ = 0 by using their biorthogonal systems. W. Schachermayer
introduced a dual version of the previous property that is satisfied by spaces like 1 .
Definition 2.3 ([145]) A Banach space Y satisfies property α of Schachermayer if
there exist A = {yγ : γ ∈ } ⊆ SY , A∗ = {yγ∗ : γ ∈ } ⊆ SY , and 0 ≤ ρ < 1
such that the following hold:
(a) yγ∗ (yγ ) = 1 for every γ ∈ ,
(b) |yγ∗ (yβ )| ≤ ρ < 1 for every γ = β,
(c) BY is the closed absolutely convex hull of A.
J. Lindenstrauss also introduced in [137] two properties to denote spaces for
which the density of norm-attaining operators was granted. Namely, a Banach space
X has property A of Lindenstrauss if NA(X , Y) is always dense in B(X , Y) for
all Banach spaces Y, and a Banach space Y has property B of Lindenstrauss
if NA(X , Y) is always dense in B(X , Y) for all Banach spaces X . It is worth
noting that property β of Lindenstrauss implies property B and property α of
Schachermayer implies property A (see [137, 145]). The concepts of universal
domain and range arose to represent those Banach spaces satisfying properties A
and B, respectively. This concept was extended to the BPBp as follows.
Definition 2.4 ([34, Definition 1.2]) Let X , Y be Banach spaces. We say that
(a) X is a universal BPB domain space if, for every Banach space Z, the pair
(X , Z) has the BPBp.
(b) Y is a universal BPB range space if, for every Banach space Z, the pair (Z, Y)
has the BPBp.
Now, we have the two following results.
Theorem 2.5 ([9, Theorem 2.2]) Suppose that Y has property β of Lindenstrauss.
Then, the pair (X , Y) has the BPBp for operators for every Banach space X . In
other words, Y is a universal BPB range space.
The next result is due to Sun Kwang Kim and Han Ju Lee (see also Theorem 2.39
below).
Theorem 2.6 ([125, Theorem 3.1]) Suppose that X is uniformly convex. Then,
(X , Y) has the BPBp for operators for every Banach space Y. In other words, X is
a universal BPB domain space.
Remark 2.7 It is worth mentioning that Theorems 2.1, 2.5, and 2.6 were all
generalized by considering bounded closed convex sets instead of the unit ball
BX (see [67, Theorems 3.1, 3.2 and Corollary 3.4], respectively). We strongly
recommend the reader to take a look at that interesting paper due to Dong Hoon
Cho and Yun Sung Choi, which contains a more general approach.
Let us notice that the authors in [9] were not interested in calculating the optimal
constants in Theorem 2.5 (see the proof of [9, Theorem 2.2]). Nevertheless, by
The Bishop–Phelps–Bollobás Theorem: An Overview 525
using some ideas from [54] (and also from [61, 62]), Vladimir Kadets and Mariia
Soloviova studied in [123] estimates for these constants when the range space
satisfies property β of Lindenstrauss (see Theorem 3.4). As far as we know no
further research was done in this direction when X is taken to be uniformly convex.
Regarding sharpness of the constants in Bishop–Phelps–Bollobás like theorems, we
send the interested reader to Sect. 3 below.
We have already seen that reflexivity plays an important role when it comes to
operators which attain their norms. Indeed, by using the James theorem, it is not
difficult to construct a linear operator which never attains its norm when the domain
space is non-reflexive. On the other hand, Lindenstrauss proved that reflexive spaces
satisfy property A, that is, if X is reflexive, then NA(X , Y) is dense in B(X , Y)
for every Banach space Y. Therefore, it is natural to wonder whether the same
happens with the BPBp. However, this is not the case, as shown in the example after
Theorem 2.1. Actually, there exists a reflexive space X such that the pair (X , X )
does not have the BPBp. For this, take any Y which is reflexive and strictly convex
but not uniformly convex and consider the reflexive space X = 21 ⊕1 Y. This space
does the job (we send the interested reader to carefully check [34, Theorem 2.1,
Corollary 3.3, and Example 3.4]).
Now, going back to the non-reflexive setting, the authors in [9] gave a complete
characterization for all Banach spaces Y such that the pair (1 , Y) has the Bishop–
Phelps–Bollobás property. They defined the approximate hyperplane series property
(AHSP, for short) and proved that (1 , Y) has the BPBp for operators if and only if Y
has the AHSP (see [9, Theorem 4.1]). This property is very technical and we will not
treat it here. We send the interested reader in the recent progress on the AHSP (and
its variants) to the already mentioned recent survey [6] (and the references therein),
where M. D. Acosta exposes interesting facts about it in Section 3 (the interested
reader may also check for instance the papers [10, 22, 60, 78, 108, 114] and their
references for more information). For the sake of completeness, in the next theorem
we exhibit some classical Banach spaces Y for which the pairs (1 , Y) satisfy the
BPBp.
Theorem 2.8 The pair (1 , Y) has the BPBp for operators when
(a) Y is finite-dimensional.
(b) Y is uniformly convex.
(c) Y = C0 (L).
(d) Y = L1 (μ), where μ is any measure.
(e) Y = A(D).
(f)Y = H ∞ (D).
(g) Y has the property β of Lindenstrauss.
(h) Y = L1 (μ, X ), where μ is a σ -finite measure and X is as (a)-(g).
(i)Y = C0 (L, X ), where L is a locally compact Hausdorff space and X is as
(a)-(g).
(j) Y = Au (BX ), the algebra of all uniformly continuous and holomorphic
mappings on the open unit ball of a complex Banach space X .
526 S. Dantas et al.
Notice that Theorem 2.8.(g) is a particular case of Theorem 2.5 (see also the more
general result [80, Proposition 2.10]). For the proofs of Theorem 2.8 we send the
reader to [80, Section 2]. Item (i) from the previous theorem can be found in [114,
Corollary 2.10], and item (j) can be found in [75, Corollary 2.17].
Note that the previous result shows once more a difference between the classical
norm-attaining theory and the BPBp, since we know that if X has the Radon-
Nikodým property, then for all Banach spaces Y, NA(X , Y) is dense in B(X , Y)
(see [43]), but there are Banach spaces Y such that (1 , Y) fails the BPBp, and 1
satisfies the Radon-Nikodým property (see also [34, Section 3]). In fact, the authors
of this survey are not aware of works studying when (X , Y) has the BPBp if X
has the Radon-Nikodým property besides some particular cases. We do not know
for example for what spaces Y we have that (J, Y) satisfies the BPBp, where J
is the James’ space (see Question 6, and see [103, 138] for background on James’
space).
Still on the AHSP, we have the following result which gives several examples on
when the pair (L1 (μ), Y) satisfies the BPBp for operators: (L1 (μ), Y) has the BPBp
if Y has the Radon-Nikodým property and the AHSP. We again send the reader to
[6, pgs. 16-19] for a more complete discussion on the recent progress on the AHSP.
Theorem 2.9 ([77, Theorem 2.2]) Suppose that Y has the Radon-Nikodým prop-
erty. Let μ be a σ -finite measure. Then, the pair (L1 (μ), Y) has the BPBp for
operators if and only if Y has the AHSP.
On the other hand, Richard Aron, Yun Sung Choi, and the second and third
authors of this survey proved that (L1 (μ), L∞ [0, 1]) has the BPBp for operators
(see [33, Theorems 2.3 and 2.4]). This result was extended by Yun Sung Choi, Sun
Kwang Kim, Han Ju Lee, and Miguel Martín (see [79, Theorem 4.1]).
Theorem 2.10 ([79, Theorem 4.1] (see also [33, Theorems 2.3 and 2.4])) Let
μ be an arbitrary measure and let ν be a localizable measure. Then, the pair
(L1 (μ), L∞ (ν)) has the BPBp for operators. In particular, (L1 (μ), L∞ [0, 1]) has
the BPBp for operators.
There is one more result that Theorem 2.9 above does not cover. Indeed, when
the range space is an L1 -space, we have the following general result.
Theorem 2.11 ([79, Theorem 3.1]) Let μ, ν be arbitrary measures. Then, the pair
(L1 (μ), L1 (ν)) has the BPBp for operators.
Regarding Lp -spaces, we borrow [79, Corollary 1.3] to summarize all the results
on the pairs (Lp , Lq ) that satisfy the BPBp for operators including Theorems 2.10
and 2.11. Let us notice that (b) below is an immediate consequence of Theorem 2.6.
Item (c) follows from [11, Theorem 2.5] (see Theorem 2.13).
Theorem 2.12 ([79, Corollary 1.3]) Let μ, ν be any measures. Then, the pair
(Lp (μ), Lq (ν)) has the BPBp for operators when
(a) p = 1 and 1 ≤ q < ∞.
(b) 1 < p < ∞ and 1 ≤ q ≤ ∞.
The Bishop–Phelps–Bollobás Theorem: An Overview 527
Theorem 2.18 ([127, Theorem 2.2]) If Y is uniformly convex, then the pair
(C(K), Y) has the BPBp for operators in the real case.
In the complex case we should highlight the following nice result due to M. D.
Acosta from 2016 (see [5]), which generalizes Theorem 2.17.(b).
Theorem 2.19 ([5, Theorem 2.4]) The pair (C0 (L), Y) has the BPBp for opera-
tors in the complex case whenever L is a locally compact Hausdorff and Y is any
C-uniformly convex complex space. In particular, the pairs
(a) (C0 (L), Lp (ν)),
(b) (L∞ (μ), Lp (ν))
satisfy the BPBp for operators in the complex case for any positive measure μ and
1 ≤ p < ∞ and for every measure ν.
Let us note that Theorem 2.19.(b) was not covered by Theorem 2.12 above. It
is worth mentioning that it seems to be an open problem whether or not the pair
(c0 , 1 ) satisfies the Bishop–Phelps–Bollobás property for operators in the real case
(see Question 5).
Whenever the pair (c0 , 1 ) has the BPBp for operators, then (n∞ , 1 ) satisfies it
uniformly for every n ∈ N (see [34, Theorem 2.1]), and the converse also holds. This
fact motivated M. D. Acosta and José L. Dávila to characterize the Banach spaces Y
such that the pairs of the form (n∞ , Y) satisfy the BPBp for operators (see [17, 18];
see also [14]) for a fixed n ∈ N. In order to do so, they considered a geometric
property: the approximate hyperplane sum property for n∞ . A complete treatment
of this property is done in Acosta’s survey [6] in pages 22 and 23, and we strongly
suggest the reader to check this and the references therein. Here, we highlight the
consequences of such a property and exhibit the specific Banach spaces Y such that
the pair (n∞ , Y) satisfies the BPBp for operators. We send the reader to check [6,
Proposition 4.9] and the paragraph just after that. Let us notice that some of the
items in Theorem 2.20 follow immediately from already mentioned results in this
survey, nevertheless we include them for the sake of completeness.
Theorem 2.20 ([17, Theorem 3.3]) Let n ∈ N be fixed with n ≥ 2. The pair
(n∞ , Y) has the BPBp for operators when
(a) Y is finite-dimensional.
(b) Y is uniformly convex.
(c) Y has property β of Lindenstrauss.
(d) Y ⊆ C(K) is a uniform algebra.
(e) Y = L1 (μ) for every positive measure μ.
(f) Y = C0 (L, Z), where Z is one of the spaces in (a)-(e).
Note that from the previous result, the pairs (n∞ , 1 ) satisfy the BPBp for
operators for all n, as desired. However, it is not known if they satisfy it uniformly,
since the mappings η obtained in the proof depend on n, so the question of whether
or not (c0 , 1 ) has the BPBp for operators remains open despite that.
The Bishop–Phelps–Bollobás Theorem: An Overview 529
We have been discussing some results and questions where the domain space
was c0 or n∞ . Let us remark that, actually, some results are known when the domain
space is Asplund. Richard M. Aron, Bernardo Cascales and Olena Kozhushkina
showed in [32] the following result.
Theorem 2.21 ([32, Corollaries 2.6 and 2.7]) (X , Y) has the BPBp for operators
in the following cases:
(a) X is Asplund and Y = C0 (L) for any locally compact Hausdorff L.
(b) X is any Banach space and Y = C0 (L), where L is a scattered locally compact
Hausdorff space.
Soon after, Bernardo Cascales, Antonio J. Guirao and Vladimir Kadets extended
Theorem 2.21.(a) to the case where the range space is a uniform algebra (see [55]).
Theorem 2.22 ([55, Theorem 3.6]) (X , Y) has the BPBp for operators if X is
Asplund and Y ⊂ C(K) is a uniform algebra.
For positive results on the BPBp for operators when the range is an operator
space, such as K(X , C(K)) or W(X , C(K)), we send the reader to [13, Theorem
3.1 and Corollary 3.2].
In this section we consider the BPBp when restricted to some particular classes of
operators. We may define in a natural way when a pair of Banach spaces (X , Y)
satisfies the BPBp for some class. For instance, we say that (X , Y) has the BPBp
for compact operators when one starts with a compact operator T in Definition 1.3
satisfying (1) and end up with another compact operator S satisfying conditions (2).
Theorem 2.23 ([79, Corollary 5.3]) Let K be a compact Hausdorff space K and
μ be a finite measure. Consider the real Banach space L1 (μ) and C(K).
(a) The pair (L1 (μ), C(K)) has the BPBp for Bochner representable operators.
(b) The pair (L1 (μ), C(K)) has the BPBp for weakly compact operators.
In fact, we have a more complete scenario when it comes to finite-rank, compact,
weakly compact, and Radon-Nikodým operators when the domain is an L1 -space
due to María Dolores Acosta, Julio Becerra Guerrero, Domingo García, Sun Kwang
Kim, and Manuel Maestre.
Theorem 2.24 ([13, Proposition 2.2, Theorem 2.3, and Corollary 2.4]) Let μ be
a finite measure such that L1 (μ) is infinite-dimensional. The pair (L1 (μ), Y) has
the BPBp for
(1) finite-rank operators,
(2) compact operators,
530 S. Dantas et al.
Before going on, let us give some attention to an important tool for complex spaces
used in the proofs of previous results: a Urysohn type lemma for uniform algebras
proved by Cascales, Guirao, and Kadets (see [55]).
Lemma 2.26 ([55, Lemma 2.7]) Let A ⊆ C(K) be a unital uniform algebra and
0 its Choquet boundary. Then, for any open subset U of K with U ∩ 0 = ∅ and
for 0 < ε < 1, there exist f ∈ A and t0 ∈ U ∩ 0 satisfying
(a) f (t0 ) = f ∞ = 1.
(b) |f (t)| < ε for every t ∈ K \ U .
(c) |f (t)| + (1 − ε)|1 − f (t)| ≤ 1 for every t ∈ K.
We send the interested reader to [128] and [38] for related (and inspired by [55,
Lemma 2.7]) Urysohn type lemmas for holomorphic functions. This lemma is also
used in [75] to obtain results concerning the numerical index, the Daugavet equation,
lushness and the AHSP on uniform algebras and also in [86] to give a version
of Bishop–Phelps–Bollobás theorem for the unital uniform algebra Aw∗ u (BX ) for
some complex Banach space X .
Focusing on compact operators, the first three authors of this survey, together with
Miguel Martín, studied the BPBp for this class of operators in a systematic way [88].
Although this survey is focused on the Bollobás theorem, it is worth mentioning that
Martín gave a negative answer to an old open question on whether every compact
operator can be approximated by norm-attaining operators [139] opening the gate
for further research on this topic (see also [140] and the references therein).
Bearing in mind our Sect. 2.1, we already have a long list of pairs of Banach
spaces (X , Y) that satisfy the BPBp for compact operators. Indeed, when analyzing
the proofs of such results, when one starts with a compact operator, the new
operator that we construct there which satisfy the Bollobás’ conditions (2) is trivially
compact. We borrow [88, Examples 1.5] and the results we have mentioned already,
and list them in the following theorem.
Theorem 2.27 ([88, Examples 1.5]) The pair (X , Y) has the BPBp for compact
operators when
(a) X is arbitrary and Y has property β of Lindenstrauss.
(b) X is uniformly convex and Y is arbitrary.
(c) X is arbitrary and Y ⊆ C(K) is a uniform algebra.
532 S. Dantas et al.
Corollaries 2.11 and 2.12]; see also [30, Proposition 2.13 and Corollary 2.14]
for a partial converse).
• If X and Y are finite dimensional Banach lattices ([29, Corollary 2.12]).
Theorem 2.35 ([29]) X has the BPBp for positive functionals when:
• X is uniformly monotone for orthogonal elements (this actually characterizes
having a stronger property than the BPBp for positive functionals) ([29, Theorem
2.9]; see also [29, Remark 2.8]).
• X is a finite-dimensional Banach lattice ([29, Corollary 2.13]).
• X is strongly monotone and has the hereditary norm-attaining property ([29,
Theorem 2.16]). In particular, X = C(K) (K compact Hausdorff), X = M(K)
(K compact Hausdorff) and X = Lp (μ) (1 ≤ p < ∞, μ positive measure) have
this property ([29, Corollary 3.2]).
They also provide some examples of Banach lattices that do not satisfy the BPBp
for positive functionals (see [29, Section 3]).
In this section we will see some important results on the line of the Bishop–
Phelps–Bollobás property for multilinear mappings which were obtained in the past
few years focused on what has come after [4]. Throughout this section, we write
X1 , . . . , XN , Y for arbitrary Banach spaces.
The Bishop–Phelps–Bollobás Theorem: An Overview 535
To start with, we need to adapt Definition 1.3 for this new context.
Definition 2.37 We say that (X1 , . . . , XN ; Y) has the Bishop–Phelps–Bollobás
property for multilinear mappings (BPBp for multilinear mappings, for short) if
given ε > 0, there exists η(ε) > 0 such that whenever A ∈ B(X1 , . . . , XN ; Y)1
with A = 1 and (x1 , . . . , xN ) ∈ SX1 × . . . × SXN satisfy
First of all, we need to justify why the study of such a property for multilinear
mappings is relevant. In 2009, Yun Sung Choi and Hyun Gwi Song proved that
the Bishop–Phelps–Bollobás theorem does not hold for bilinear forms on 1 × 1 .
Actually, they showed that given the bilinear form T (ei , ej ) := 1 − δij , there is no
norm-attaining bilinear form S on 1 ×1 satisfying S −T < 1 (see [81, Theorem
2]). We highlight it below.
Theorem 2.38 ([81, Theorem 2]) The triple (1 , 1 ; K) fails the BPBp for bilinear
forms.
Nevertheless, Theorem 2.39 below gives a positive result. In particular, if X is
uniformly convex, then (X , Y) has the BPBp for operators for every Banach space
Y as in Theorem 2.6.
Theorem 2.39 ([15, Theorem 2.2]) Let X1 , . . . , XN be uniformly convex Banach
spaces. Then, (X1 , . . . , XN ; Y) has the BPBp for multilinear mappings for every
Banach space Y.
The authors of [15] also gave a complete characterization for the triple (1 ×
Y; K) to have the Bishop–Phelps–Bollobás property for bilinear forms by using the
AHSP for a pair (X , X ) which we will not treat here. We send the interested reader
to [6, Definition 5.7] and the references therein. We give the most relevant (and
specific) consequences of it in the next theorem.
Theorem 2.40 ([15, Theorem 3.6]) The triple (1 , Y; K) has the BPBp for bilin-
ear forms when
(a) Y is uniformly smooth.
(b) Y is finite-dimensional.
(c) Y = C0 (L) and, in particular, when Y = C(K) or Y = c0 .
(d) Y = K(H), where H is a Hilbert space.
1 We send the reader back to the first paragraph of the Introduction to check our notation.
536 S. Dantas et al.
The reader can find the proofs of Theorem 2.40.(a)–(d) in [15, Propositions 4.1,
4.2, 4.4 and 4.7], respectively.
Following the line of Theorem 2.38 above, we have the following negative result.
Theorem 2.41 ([15, Proposition 4.8]) The triple (1 , L1 (μ); K) fails the BPBp for
bilinear forms whenever L1 (μ) is infinite-dimensional.
It is worth mentioning that Yun Sung Choi showed in 1997 that the set
of all norm-attaining bilinear forms on L1 [0, 1] × L1 [0, 1] is not dense in
B(L1 [0, 1], L1 [0, 1]; K) (see [70, Theorem 3]). This means, in particular, that
this triple cannot satisfy the BPBp for bilinear forms.
Theorem 2.42 The triple (L1 [0, 1], L1 [0, 1]; K) fails the BPBp for bilinear forms.
On the other hand, the same authors of [15] together with Choi, Kim, and
Lee provided a characterization for the triple (L1 (μ), Y; K) to have the BPBp
for bilinear forms in [12]. Let us notice that we have an additional assumption in
[12, Theorem 2.6]: the Banach space Y is assumed to be Asplund. The tools and
techniques used to prove the following result are also based on the AHSP.
Theorem 2.43 ([12, Theorem 2.6 and Corolary 2.7]) Let μ be a σ -finite measure
such that L1 (μ) is infinite-dimensional. Then, (L1 (μ), Y; K) has the BPBp for
bilinear forms when
(a) Y is uniformly smooth.
(b) Y is finite-dimensional.
(c) Y = c0 .
Remark 2.44 In the vein of Theorems 2.40 and 2.43, it is worth mentioning that
there are positive results on Banach function spaces over a measure space. Indeed,
Lucía Agud, José M. Calabuig, Sebastián Lajara, and Enrique A. Sánchez Pérez
gave some applications to the BPBp (for operators and bilinear forms) from their
results on Gâteaux and Fréchet smoothness, and the uniform smoothness of Lp (m),
where m : −→ X is a vector measure (we invite the reader to visit Section 4 of
[31] and check the necessary background in Sections 1, 2 and 3 of that paper).
There is another positive result when it comes to the BPBp for bilinear forms on
C0 (L1 ) × C0 (L2 ) in the complex case due to Kim, Lee, and Martín, where L1 , L2
are locally Hausdorff topological spaces (see [132]). Indeed, we have the following
result.
Theorem 2.45 ([132, Theorem 2 and Corollary 3]) (C0 (L1 ), C2 (L2 ); K) has the
BPBp for bilinear forms in the complex case. In particular, so does (c0 , c0 ; K) (in
the complex case).
We send the reader to [69] for some generalizations on the previous result.
We also have the following positive result when one of the factors is c0 and the
other one is an p -space.
The Bishop–Phelps–Bollobás Theorem: An Overview 537
Theorem 2.46 ([124, Corollary 2.9]) The triple (c0 , p ; K) has the BPBp for
bilinear forms whenever 1 < p < ∞.
At this point, one might wonder what happens with Theorem 2.46 when c0
is replaced by 1 . The answer to this is that we have an analogous result, and it
follows immediately from Theorem 2.8.(b) and [87, Proposition 2.6]. This was also
observed explicitly in [82, Corollary 1.1]. We highlight this in the following result.
Theorem 2.47 ([82, Corollary 1.1]) The triple (1 , p ; K) has the BPBp for
bilinear forms whenever 1 < p < ∞.
We have seen in Theorem 2.27 that when Y is an isometric L1 -predual, we
have that the pair (X , Y) has the BPBp for compact operators. In turns out that, by
using the metric approximation property on the L1 -predual, we have the following
technical result for (compact) multilinear mappings which provides some positive
results for the BPBp for this class of functions.
Theorem 2.48 ([87, Theorem 2.9]) Suppose that Y is an L1 -predual and that the
N-tuple (X1 , . . . , XN ; K) has the BPBp for multilinear forms. Then, the (N + 1)-
tuple (X1 , . . . , XN ; Y) has the BPBp for compact multilinear mappings.
In particular, we have the following corollary.
Corollary 2.49 Suppose that Y is an L1 -predual. The triple (X , Z; Y) has the
BPBp for compact bilinear mappings when
(a) X = C0 (L1 ) and Z = C0 (L2 ) in the complex case.
(b) both X , Z are uniformly convex Banach spaces.
(c) X = 1 and
(c.1) Z is uniformly smooth.
(c.2) Z is finite-dimensional.
(c.3) Z = C0 (L) and, in particular, Z = C(K) or Z = c0 .
(c.4) Z = K(H), where H is a Hilbert space.
(d) X = L1 (μ), where μ is a σ -finite measure such that L1 (μ) is infinite-
dimensional and
(d.1) Z is uniformly smooth.
(d.2) Z is finite-dimensional.
(d.3) Z = c0 .
(e) X = c0 and Z = p for 1 < p < ∞.
Let us observe that items (c) and (d) of Corollary 2.49 are immediate consequences
of Theorem 2.40 and Theorem 2.43, respectively. Item (b) follows from Theo-
rem 2.39 and item (a) from Theorem 2.45. Item (e) follows from Theorem 2.46.
Remark 2.50 In [131], Kim, Lee, and Martín defined a more general geometric
property than the ASHP in order to characterize the pairs (X , Y) such that
(1 (X ), Y) satisfies the BPBp for operators. In the same directions, in [87]
538 S. Dantas et al.
we defined the analogous property for bilinear forms with the idea to give a
characterization for the triples of the form (1 (X ), Y; K) to have the BPBp for
bilinear forms.
For symmetric bilinear forms (Hermitian forms), the only positive known result
is the following one:
Theorem 2.51 ([104]) Let H be a Hilbert space.
• (H, H) has the BPBp for continuous symmetric bilinear forms ([104, Theorems
3.2 and 3.4]).
• If H is complex, then (H, H) has the BPBp for continuous Hermitian forms
([104, Corollary 2.2]).
For negative results on the norm-attaining multilinear mappings (and, in particu-
lar, for the Bollobás version for this class of functions) we send the interested reader
to [8, 47, 119].
As in the case of operators (see Theorem 2.6) and multilinear mappings (see
Theorem 2.39), we have the following universal result for polynomials.
Theorem 2.53 ([12, Theorem 3.1]) Let X be a uniformly convex Banach space.
Then, (X ; Y) has the BPBp for N-homogeneous polynomials for every Banach
space Y.
The technique used to prove Theorem 2.53 is based on the original Lindenstrauss
argument. It is not known whether the analogous result for symmetric N-linear
forms holds (see Question 8).
The Bishop–Phelps–Bollobás Theorem: An Overview 539
In this section, we consider the Bollobás theorem for holomorphic functions. Let
us start by saying that it seems that not much has been done in this direction.
Nevertheless, from our point of view, holomorphic functions deserve special
attention and the reason is quite simple: it requires a complete different approach
and interesting techniques.
To start with, we highlight a non-linear version of the Bishop–Phelps–Bollobás
theorem. This was done in [86]. It was proven that a Bishop–Phelps–Bollobás type
theorem holds on Aw∗ u (BX ) whenever X is either a uniformly convex or a locally
c-uniformly convex, order-continuous sequence space. For necessary notation, we
send the reader to the first paragraph of Sect. 1.
Theorem 2.55 ([86, Theorem 1]) Let X be a complex Banach space. Suppose that
s is norm dense in SX . Then, given ε > 0, there exists η(ε) > 0 such that
whenever f ∈ Aw∗ u (BX ) with f ∞ = 1 and x0∗ ∈ SX satisfy
there are g ∈ Aw∗ u (BX ) with g∞ = 1 and x1∗ ∈ SX such that
In Theorem 2.55, s represents the set of all strong peak points for Aw∗ u (BX ).
Besides that, notice that in Theorem 2.55 we are requiring that the set s is dense
in SX ; for results when this happens, we refer to [86, Proposition 5 and 6]. For
necessary background we send the reader to [86, pages 8 and 9].
540 S. Dantas et al.
In the same direction, Kim and Lee proved similar results to Theorem 2.55 for
the spaces Au (BX ) and Ab (BX ) under some additional conditions on the complex
Banach space X .
Theorem 2.56 ([128, Corollary 8]) Suppose that X is either a locally uniformly
convex space or a locally c-uniformly convex, order-continuous sequence space.
Let Y be a Banach space. If A is one of Au (BX ) or Ab (BX ), then for ε ∈ (0, 1),
whenever a norm-1 element f in A(BX ; Y) and an element x0 ∈ BX satisfy
ε
f (x0 ) > 1 − ,
6
A quick glance at the definition of · L suggests that the most natural way of
defining a norm-attaining Lipschitz mapping is the following. We say that f ∈
Lip0 (M, Y) strongly attains its norm if there exist x, y ∈ M, x = y, such that
f (x) − f (y)
f L = .
d(x, y)
The Bishop–Phelps–Bollobás Theorem: An Overview 541
The subset of all Lipschitz mappings in Lip0 (M, Y) which attain their norms
strongly is denoted by SNA(M, Y) (although the notations SA(M, Y) and
LipSNA(M, Y) have also been used in the literature). Since there is a natural
concept of norm attainment, it makes sense to study density problems or even
Bishop–Phelps–Bollobás type properties in this scenario. Nevertheless, in [121,
Theorem 2.3] it was shown that if X is a geodesic pointed metric space (in particular,
if X is any Banach space), then SNA(X , R) is never dense in Lip0 (X , R) and later
this was extended to metric length spaces [54].
Theorem 2.58 ([54, Theorem 2.2]) Let M be a complete length pointed metric
space. Then, the set SNA(M, R) is not dense in Lip0 (M, R).
In fact, this result combined with Proposition 2.69 show that if M is a metric
length space and Y is any other Banach space, a BPBp-like property is not possible
(actually, in this case we do not even get density, see [65, Proposition 4.2]).
However, this inconvenience has not stopped researchers from developing a rich
theory on denseness of norm-attaining Lipschitz mappings by considering domains
that are not Banach spaces or using weaker norm attainment concepts (see, for
instance, [68, Section 1] for a clean exposition on several and the relations between
them). As a matter of fact, as we are going to show throughout this section, many
authors came up with several definitions in order to get positive results in this setting.
As far as the authors of this survey know, the study of the density of norm-
attaining Lipschitz mappings was initiated independently in [112] and [121], and
since then many authors have contributed to the topic. We refer the interested reader
to [54, 64–66, 68, 83, 106, 107, 112, 121] and also the nice survey [111, Section
5] for a solid background on the topic. Four of those works also include Bishop–
Phelps–Bollobás type properties for Lipschitz mappings. The rest of the section
will focus on those results.
We start with a work by Vladimir Kadets, Miguel Martín, and Mariia Soloviova,
who focused their study on the case where M is a Banach space, Y = R, and
dealt with some weaker forms of norm attainment (see [121]). First, for the set of
continuous seminorms on X , Sem(X ), they got a variation of a BPBp-like result.
Proposition 2.59 ([121, Proposition 3.4]) Let X be a Banach space. Then for
every ε > 0, there is δ > 0 such that for every p0 ∈ Sem(X ) with p0 = 1
and every x0 ∈ SX with p0 (x0 ) > 1 − δ, there exist p ∈ Sem(X ) with p = 1 and
x ∈ SX such that
Note that, in particular, this implies the uniform density of the set of norm
attaining seminorms.
We now need the definition of a natural (but weaker) form of norm attainment in
Lipschitz functionals in the setting of Banach spaces.
542 S. Dantas et al.
Definition 2.60 ([121, Definition 1.3]) Let X be a real Banach space. A Lipschitz
functional g ∈ Lip0 (X ) attains its norm at the direction u ∈ SX if there is a
sequence of pairs {(xn , yn )} in X × X , with xn = yn , such that
xn − yn g(xn ) − g(yn )
lim =u and lim = g.
n→∞ xn − yn n→∞ xn − yn
In this case, we say that g attains its norm directionally. The set of all those f ∈
Lip0 (X ) that attain their norm directionally is denoted by DA(X ).
This kind of norm attainment is a natural approach since it coincides with the
usual norm attainment if g is linear. Also, if X is finite-dimensional, then DA(X ) =
Lip0 (X ). In [121], the authors defined a BPBp-like property for Lipschitz mappings
involving this kind of norm attainment.
Definition 2.61 ([121, Definition 1.4]) A Banach space X has the directional
Bishop–Phelps–Bollobás property for Lipschitz functionals (X ∈ LipBPB, for
short), if for every ε > 0, there is δ > 0 such that for every f ∈ Lip0 (X ) with
f = 1 and for every x, y ∈ X with x = y satisfying
f (x) − f (y)
> 1 − δ,
x − y
there is g ∈ Lip0 (X ) with g = 1 and there is u ∈ SX such that g attains its norm
at the direction u,
x−y
g − f < ε, and − u
x − y < ε.
We also need to define another kind of norm attainment and its associated BPBp-
like property.
Definition 2.62 ([121, Definition 1.3]) Let X be a real Banach space. A Lipschitz
functional g ∈ Lip0 (X ) attains its norm at a point v ∈ X at the direction u ∈ SX if
there is a sequence of pairs {(xn , yn )} in X × X , with xn = yn and limn→∞ xn =
limn→∞ yn = v, such that
xn − yn g(xn ) − g(yn )
lim =u and lim = g.
n→∞ xn − yn n→∞ xn − yn
In this case, we say that g attains its norm locally-directionally. The set of all those
f ∈ Lip0 (X ) that attain their norm locally-directionally is denoted by LDA(X ).
Definition 2.63 ([121, Definition 1.4]) A Banach space X has the local directional
Bishop–Phelps–Bollobás property for Lipschitz functionals (X ∈ LLipBPB for
short), if for every ε > 0, there is δ > 0 such that for every f ∈ Lip0 (X ) with
The Bishop–Phelps–Bollobás Theorem: An Overview 543
f (x) − f (y)
> 1 − δ,
x − y
there is g ∈ Lip0 (X ) with g = 1 and there are v ∈ X and u ∈ SX such that g
attains its norm at the point v at the direction u,
x−y
g − f < ε, − u
x − y < ε, and dist(v, [x, y]) < ε.
With the help of some lemmas (that might be of interest in themselves) involving
the LipBPB and the LLipBPB (see [121, Lemmas 4.1 and 4.4]), the authors showed
the following.
Theorem 2.64 ([121, Theorem 5.3]) Every uniformly convex Banach space X has
the local directional Bishop–Phelps–Bollobás property for Lipschitz functionals.
Some time after, Rafael Chiclana and Miguel Martín did a systematic study of a
vector-valued BPBp property for Lipschitz mappings for the strong norm attainment
with the domain being a metric space. We begin with the following definition.
Definition 2.65 ([65, Definition 1.1]) Let M be a pointed metric space and let Y
be a Banach space. We say that the pair (M, Y) has the Lipschitz Bishop–Phelps–
Bollobás property (Lip-BPB property for short) if given ε > 0, there is η(ε) > 0
such that for every norm-one F ∈ Lip0 (M, Y) and every p, q ∈ M, p = q such
that
If this holds for a class of linear operators from F (M) to Y, we will say that the pair
(M, Y) has the Lip-BPB property for that class.
The authors also give a reformulation of that definition in [65, Remark 1.2.(a)]
in terms of linear operators associated to the Lipschitz mappings. The first result for
this property is the following.
Theorem 2.66 ([65, Theorem 2.1]) Let M be a finite pointed metric space and let
Y be a Banach space. If (F (M), Y) has the BPBp, then (M, Y) has the Lip-BPB
property.
544 S. Dantas et al.
f (xn ) − f (yn ) xn − yn
→ z with z = f , → u and xn , yn → x.
xn − yn xn − yn
We denote by LDirA(X , Y) the set of every f ∈ Lip0 (X , Y) which attains its norm
locally directionally at some point x ∈ X in some direction u ∈ SX toward some
point z ∈ Y.
Definition 2.72 ([121, Definition 1.4]) A pair of Banach spaces (X , Y) is said to
have the local directional Bishop–Phelps–Bollobás property for Lipschitz mappings
(in short, LDirA-BPBp) if for every ε > 0, there is η > 0 such that whenever
f ∈ SLip0 (X ,Y ) and x, y ∈ X × X with x = y satisfy
f (x) − f (y)
> 1 − η,
x − y
Their main result extends [121, Theorem 5.3] in the case of compact operators.
Theorem 2.73 ([68, Theorem 4.1]) Let X and Y be Banach spaces such that X
is uniformly convex and (F (X ), Y) has the BPBp for compact operators. Then, the
pair (X , Y) has the LDirA-BPBp for Lipschitz compact mappings. In fact, we have
something more: for every ε > 0, there exists η > 0 such that for any positive
function ρ : X̃ → R and whenever f ∈ SLip0K (X ,Y ) and (x, y) ∈ X̃ satisfy
f (x) − f (y)
> 1 − η,
x − y
There is also a slightly different version of Theorem 2.73 where the domain space
is a Hilbert space.
Theorem 2.74 ([68, Theorem 4.5]) Let H be a Hilbert space and let Y be a
Banach space. Suppose that (F (M), Y) has the BPBp for compact operators. Then
for every ε > 0, there exists η > 0 such that whenever f ∈ SLip0K (H,Y ) and
(x, y) ∈ H̃ satisfy
f (x) − f (y)
> 1 − η,
x − y
there exist g ∈ SLip0K (H,Y ), z ∈ SY and x ∈ H such that g attains its norm locally
x−y
directionally at the point x in the direction x−y toward z, g − f < ε, and
dist(x, [x, y]) < ε max{x, y}.
In particular, the same holds if Y has property β of Lindenstrauss or if Y is an
L1 (μ) space (see [68, Corollary 4.6]), and the same is also true if Y is a finite-
dimensional polyhedral Banach space (see [68, Corollary 4.12]).
Finally, they got some stronger versions of Theorems 2.73 and 2.74 for the case
when X = R (see [68, Propositions 4.7 and 4.8, Corollaries 4.9 and 4.10]).
To conclude this section, let us note that a very recent work by Rafael Chiclana
and Miguel Martín ([66]) studied stability properties for the Bishop–Phelps–
Bollobás property for Lipschitz mappings (see Definition 2.65). Let us start by
defining the sum of a family of pointed metric spaces.
Definition 2.75 ([66, Definition 2.1], from [149, Definition 1.13]) Given a family
of pointed metric spaces {(Mi , di )}i∈I , the (metric) sum of the family is the disjoint
union of all Mi ’s, identifying the base points, endowed with the following metric d:
d(x, y) = di (x, y) if both x, y ∈ <Mi , and d(x, y) = di (x, 0) + dj (0, y) if x ∈ Mi ,
y ∈ Mj and i = j . We will write i∈I Mi to denote the sum of the family of metric
spaces.
<
Proposition 2.76 ([66, Proposition 2.2]) Let M = M1 M2 be the sum of two
pointed metric spaces and let Y be a Banach space. If the pair (M, Y) has the
Lip-BPB property, then so do (M1 , Y) and (M2 , Y).
The version of this proposition for compact Lipschitz mappings remains true (see
[66, Proposition 2.6]). However, none of the converses hold (see [66, Example 2.4]).
They also provided an extension of [65, Proposition 4.4]. We will include the
result for the sake of completeness, but we refer the reader to [56] for the necessary
definitions and background (see also [66, Section 3]) .
Theorem 2.77 ([66, Theorem 3.5]) Let M be a pointed metric space such that
(M, R) has the Lip-BPB property. Let Y be a Banach space in ACKρ with associated
1-norming set ⊆ BY , and let ε > 0. Then, there exists η(ε, ρ) > 0 such that if
we take T̂ ∈ L(F (M), Y) a -flat operator with T L = 1 and m ∈ Mol(M)
satisfying T̂ (m) > 1 − η(ε, ρ), then there exist an operator Ŝ ∈ L(F (M), Y)
The Bishop–Phelps–Bollobás Theorem: An Overview 547
As a consequence, if M is a pointed metric space such that (M, R) has the Lip-
BPB property, then we get several pairs of spaces satisfying the Lip-BPB property
for -flat operators (see [66, Corollary 3.6] for the details). The inconvenience of
having to work with -flat operators disappears if we restrict the results to the
compact operator case.
Proposition 2.78 ([66, Proposition 3.12]) Let M be a pointed metric space such
that (M, R) has the Lip-BPB property and let Y be an ACKρ Banach space. Then,
the pair (M, Y) has the Lip-BPB for Lipschitz compact mappings.
And again, a series of consequences can be derived from this (see [66, Corollary
3.13] for the details).
Using some results from [66] and [65, Proposition 4.16], we get the following
result.
Corollary 2.79 ([66, Corollary 3.17]) Let M be a pointed metric space and let Y
be a Banach space such that the pair (M, Y) has the Lip-BPB property for Lipschitz
compact mappings.
(1) For every compact Hausdorff topological space K, the pair (M, C(K, Y)) has
the Lip-BPB property for Lipschitz compact mappings.
(2) For 1 ≤ p < ∞, if the pair (M, p (Y)) has the Lip-BPB property for Lipschitz
compact mappings, then so does (M, Lp (μ, Y)) for every positive measure μ.
(3) For every σ -finite positive measure μ, the pair (M, L∞ (μ, Y)) has the Lip-BPB
property for Lipschitz compact mappings.
Finally, they study stability results for absolute sums of codomains.
Proposition 2.80 ([66, Proposition 4.3]) Let M be a pointed metric space, Y be
a Banach space and Y1 be an absolute summand of Y. If the pair (M, Y) has the
Lip-BPB property with a function ε $→ η(ε), then so does (M, Y1 ) with the same
function.
As a consequence, we get an universal mapping η for universal Lip-BPB domain
spaces.
Corollary 2.81 ([66, Corollary 4.4]) Let M be a pointed metric space such that
(M, Y) has the Lip-BPB property for all Banach spaces Y. Then, there exists a
function ηM (ε), which only depends on M, such that the pair (M, Y) has the Lip-
BPB property witnessed by the function ηM (ε) for evey Banach space Y.
A result in the same direction is the following.
Proposition 2.82 ([66, Proposition 4.6]) Let M be a pointed metric space, Y a
Banach space, and K a compact Hausdorff topological space. If (M, C(K, Y)) has
548 S. Dantas et al.
the Lip-BPB property witnessed by a function η(ε), then (M, Y) has the Lip-BPB
property witnessed by the same function.
It is worth noting that the last 3 results are still valid in the case of compact
operators (see [66, Proposition 4.8]
and the discussion before it for the details).
Let Y = [ i∈I Yi ]c0 or Y = [ i∈I Yi ]∞ for some family of Banach spaces
{Yi }i∈I and let M be a pointed metric space. By Proposition 2.80, if (M, Y) has the
Lip-BPB property, then all pairs (M, Yi ) have it with the same function η. We have
the following result regarding the reverse implication:
Proposition 2.83 ([66, Proposition 4.9]) Let M be a pointed metric space, let
{Y
i } i∈I be a family of Banach spaces and let Y = [ i∈I Yi ]c0 or Y =
[ i∈I Yi ]∞ . Assume that (M, Yi ) has the Lip-BPB property with the function ηi (ε)
for each i ∈ I . If inf{ηi (ε) : i ∈ I } > 0 for every ε > 0, then (M, Y) has the Lip-
BPB property.
Finally, let us notice that the compact operators version of the previous result is
also true (see [66, Proposition 4.11]).
Brailey Sims, in his 1972 Ph.D. dissertation (see [146]), raised a question that is,
in nature, related to the one that Lindenstrauss tackled in 1963 [137]: the norm-
denseness of the set of numerical radius attaining operators on a Banach space
X (we will define this concept shortly). Ever since, many authors have made
contributions regarding this question, such as Ira David Berg, Brailey Sims (see
[40]), Carmen Silvia Cardassi (see [50–52]), María Dolores Acosta, Francisco
José Aguirre, Rafael Payá, and Manuel Ruiz Galán (see for instance [2, 3, 7, 23–
25, 141]). It is also worth noting for the interested reader that M. D. Acosta initiated
a systematic study of this question in her nice Ph.D. dissertation [1].
Similar to what happened with the study of norm-attaining operators, it is a
natural question whether or not we can have Bishop–Phelps–Bollobás type theorems
for the numerical radius on some Banach spaces. The study of this question was first
addressed in 2013 by Antonio José Guirao and Olena Kozhushkina in [115], where,
paralleling the work [9], they introduced and studied the Bishop–Phelps–Bollobás
property for numerical radius. We need some background before proceeding.
Given a Banach space X , the set of states of X is (X ) := {(x, x ∗ ) ∈ SX ×
SX : x ∗ (x) = 1}. Given an operator T ∈ B(X ), its numerical radius is defined
as ν(T ) = sup{|x ∗ (T (x))| : (x, x ∗ ) ∈ (X )}. It is immediate to check that ν is a
seminorm, and that for all T ∈ B(X ), we have 0 ≤ ν(T ) ≤ T . The numerical
index of a Banach space X is defined as the following number n(X ) = inf{ν(T ) :
The Bishop–Phelps–Bollobás Theorem: An Overview 549
If, moreover, S can be chosen so that ν(S) = 1, we say that X has the Bishop–
Phelps–Bollobás property for the numerical radius (abbreviated BPBp-nu, although
some authors use the notation BPBp-ν as well).
If the conditions from the previous definition hold within a subclass of B(X ), we
say that X has the (weak)-BPBp-nu for that class of operators (see [19]).
Acosta exhibits a nice overview of the main known results about the BPBp-nu in
her survey [6, Section 6]. For the sake of completeness, we will briefly list some of
the results about it without further detail. The following results are true both in the
real and complex settings unless specified otherwise.
Theorem 2.85 Let X , Y be a Banach spaces, K be a compact Hausdorff topologi-
cal space and μ be any measure.
1. If is any index set, then the spaces c0 () and 1 () have the BPBp-nu ([115]).
2. L1 (R) has the BPBp-nu ([101, Theorem 9]).
3. If X is finite-dimensional, then it has the BPBp-nu ([130, Proposition 2]).
4. L1 (μ) has the BPBp-nu ([130, Theorem 9]).
5. If X is both uniformly convex and uniformly smooth, then it has the weak-BPBp-
nu ([130, Proposition 4]).
6. If n(X ) > 0, or if n (X ) > 0, then X has the BPBp-nu if and only if it has the
weak-BPBp-nu ([130, Proposition 6] and [134, Theorem 3.2]). In particular,
all the Lp (μ) spaces have the BPBp-nu if 1 < p < ∞ ([130, Example 8] and
[134, Theorem 2.3]).
7. If K admits local compensation (see [37, Definition 2.1]), then the real space
C(K) has the BPBp-nu ([37, Theorem 2.2]). In particular, if K is metric, the
real space C(K) has the BPBp-nu (see [37, Section 3]).
8. If X is strongly lush (see [133, Definition 1.2]; strongly lush spaces include
C(K) spaces, L1 (μ) spaces and finite-codimensional subspaces of C[0, 1]) and
550 S. Dantas et al.
presented a tool (based in [120, Lemma 3.1]) that in particular allows to carry the
BPBp for compact operators from some projections of a space to the space itself
(check Sect. 2.2). In order to get a somewhat similar result for the numerical radius,
one needs to control things both in the space and in its dual. The most general result
obtained in this direction is the following lemma.
Lemma 2.86 ([105, Lemma 2.1]) Let X be a Banach space satisfying that
nK (X ) > 0. Suppose that there is a mapping η : (0, 1) −→ (0, 1) such that
given δ > 0, x1∗ , . . . , xn∗ ∈ BX and x1 , . . . , x ∈ BX , we can find norm one
operators P1: X −→ P1(X ), i : P1(X ) −→ X such that for P := i ◦ P1: X −→ X ,
the following conditions are satisfied:
(i)P ∗ (xj∗ ) − xj∗ < δ, for j = 1, . . . , n.
(ii)P (xj ) − xj < δ, for j = 1, . . . , .
(iii)P1 ◦ i = IdP1(X ) .
(iv) P1(X ) satisfies the Bishop–Phelps–Bollobás property for numerical radius for
compact operators with the mapping η.
(v) Either P is an absolute projection and i is the natural inclusion, or
nK (P1(X )) = nK (X ) = 1.
Then, X satisfies the BPBp-nu for compact operators.
Throughout [105, Section 2], Lemma 2.86 is used to show that if a Banach space
X with positive compact index can be suitably projected into some net of spaces
that have the BPBp-nu for compact operators with a common mapping η, then
sometimes it is possible to show that X also has that property (see [105, Proposition
2.2]). This is used for instance to show that if nK (X ) > 0, then if the spaces n∞ (X ),
n ∈ N, all have the BPBp-nu for compact operators with the same η, then c0 (X )
also has the property [105, Corollary 2.3], and this is actually an equivalence, since
the converse implication was already known (see [88, Proposition 4.3]). Another not
so direct consequence of Lemma 2.86 is that if a Banach space X satisfies that X is
isometrically isomorphic to 1 , then X has the BPBp-nu for compact operators (see
[105, Corollary 2.6]). Let us highlight these two results.
Corollary 2.87 ([105, Corollary 2.3]) Let X be a Banach space with nK (X ) > 0.
Then, the following statements are equivalent:
(i) The space c0 (X ) has the BPBp-nu for compact operators.
(ii) There is a function η : (0, 1) −→ (0, 1) such that all the spaces n∞ (X ), with
n ∈ N, have the BPBp-nu for compact operators with the function η.
Moreover, if X is finite-dimensional, these properties hold whenever c0 (X ) or
∞ (X ) have the BPBp-nu.
Corollary 2.88 ([105, Corollary 2.6]) Let X be a Banach space such that X is
isometrically isomorphic to 1 . Then X has the BPBp-nu for compact operators.
Finally, in [105, Section 3], a series of tools involving some topological
procedures and finding some suitable projections, was developed to study the BPBp-
552 S. Dantas et al.
The vast majority of results we have seen so far were focused on finding pairs of
Banach spaces for which a Bishop–Phelps–Bollobás theorem is valid. However,
The Bishop–Phelps–Bollobás Theorem: An Overview 553
(almost) none of the proofs investigate the sharpness of the constants associated
to those theorems. This interesting question initiated by Bollobás himself (see [42,
Remark after Theorem 1]) has also been studied in the recent years. In this section
we will briefly discuss some of the results obtained in this direction.
In order to do so, let us define the moduli of Bishop–Phelps–Bollobás in its
general form.
Definition 3.1 ([123, Definition 4]) The Bishop–Phelps–Bollobás modulus of a
pair of Banach spaces (X , Y) is the function (X , Y, ·) : (0, 1) → R+ whose
value in η ∈ (0, 1) is defined as the infimum of those ε > 0 such that for every
(x, T ) ∈ BX × BB(X ,Y ) with
T (x) > 1 − η,
(see [63, Theorem 2.1 and Corollary 2.4]), that is, given
√ an η, the best ε one can get
that works for all Banach spaces at once is at most 2η. However, the authors did
554 S. Dantas et al.
not stop there: not only they gave this upper bound, but they also showed that it is
actually sharp, that is, there are Banach spaces where it can not be improved. We
show just some of these examples.
Note that they also found spaces for which that upper bound could be improved
such that when X is uniformly non-square (see [123, Theorem 2]), or when X = 21
(see [123, Theorem 3]). Note that 21 attains the maximum possible modulus in
the case of functionals, but this is not the case for operators. On the other hand,
the authors do not provide any example of a pair of Banach spaces for which
the estimation from the previous theorem is sharp. They did, however, show the
existence of pairs of Banach spaces for which a lower bound of the modulus was
reasonably close to the upper bound.
Theorem 3.5 ([123, Theorem 4]) For every Banach space Y,
S
(21 , Y, η) ≥ min{ 2η, 1},
In this section, we treat stronger properties than the BPBp, namely the Bishop–
Phelps–Bollobás operator property and the Bishop–Phelps–Bollobás point prop-
erty. Let us take a brief moment here to explain where the motivation to study such
properties comes from.
In [125], Sun Kwang Kim and Han Ju Lee proved the following characterization
for uniform convexity.
Theorem 4.1 ([125, Theorem 2.1]) A Banach space X is uniformly convex if and
only for every ε > 0, there exists η(ε) > 0 such that for every x ∗ ∈ SX and x ∈ BX
such that
Let us notice that Theorem 4.1 is a Bollobás type theorem where the functional x ∗
does not change; in other words, the same functional that almost attains the norm at
a point, actually attains its norm at a nearby point. At a first glance, the analogous
property for bounded linear operators looks really restrictive in the sense that not
many spaces would satisfy it. To make sure we are speaking the same language as
the reader, let us give a name to such a (possible) property.
Definition 4.2 ([84, Definition 2.8]) We say that the pair (X , Y) has the Bishop–
Phelps–Bollobás operator property (BPBop, for short) for operators if given ε > 0,
there is η(ε) > 0 such that whenever T ∈ B(X , Y) with T = 1 and x ∈ SX
satisfy
Notice now that Theorem 4.1 says simply that X is uniformly convex if and only
(X , K) has the BPBop for linear functionals. It turns out that, for spaces X , Y with
dimension bigger or equal to 2, the BPBop for operators is never possible. Indeed,
after several negative results presented in [84], the first author together with Kadets,
Kim, Lee, and Martín proved the following result.
Theorem 4.3 ([94, Theorem 2.1]) Let X , Y be real Banach spaces of dimension
greater or equal to 2. There exist (Tn )∞
n=1 ⊆ NA(X , Y) ∩ SB(X ,Y ) and x0 ∈ SX
such that Tn (x0 ) −→ 1 as n → ∞ and
4 5
inf dist (x0 , {x ∈ SX : Tn (x) = 1}) > 0.
n∈N
Let us notice that the BPBpp implies immediately the BPBp. Moreover, as expected,
it turns out that, for linear functionals, the Banach space X is uniformly smooth if
and only if the pair (X , K) has the BPBpp for linear functionals (see [95, Theorem
2.1]). This implies (see [95, Proposition 2.3]) that whenever a pair (X , Y) satisfies
the BPBpp for operators, the domain space X must be uniformly smooth. For
that reason, we can see the great connection between the BPBpp and uniform
smoothness as well as why the pairs of the form (1 , Y) and (c0 , Y) always fail such
a property. In particular, the BPBp and the BPBpp are not equivalent properties.
Next we will give the first positive results about the point property for operators.
To start with, we present the following result, which says that Hilbert spaces are
universal domain spaces. This is a consequence of the fact that Hilbert spaces have
transitive norms, that is, if x, y ∈ SH satisfy x − y < ε, then there exists a linear
isometry R : H −→ H such that R(x) = y and R − IdH < ε, where IdH is the
identity operator on H.
Theorem 4.5 ([95, Theorem 2.5]) Let H be a Hilbert space. The pair (H, Y) has
the BPBpp for operators for every Banach space Y.
It is worth mentioning that there exists a more general result than Theorem 4.5
due to Cabello-Sánchez et al. [45]. They use the BPBpp as a tool to deal with Banach
space X whose group of isometries acts micro-transitively on SX . In fact, they
introduce a weakening of the micro-transitive, the uniform micro-semitransitive
norms (see [45, Definition 2.2]). As a consequence of some results related to the
BPBpp, they were able to prove that every Banach space whose norm is uniformly
micro-semitransitive (in particular, if it is micro-transitive) is both uniformly convex
and uniformly smooth (see [45, Corollary 2.13]). In fact, the only known spaces
so far about micro-transitivity are Hilbert spaces. It is also worth mentioning that
we do not know if micro-transitivity is a different property than uniform micro-
semitransitivity.
Coming back to the point property and bearing in mind Theorem 4.5, it is natural
to wonder whether the analogous result holds for Lp (μ)-spaces. In other words, is
it true that the pair (Lp (μ), Y) satisfies the BPBpp for every Banach space Y? This
question was addressed in [93] (where the authors called the BPBpp as pointwise
BPB property). See also [95, Remark 2.6]. The answer to this question is not positive
in general, as the following theorem shows.
Theorem 4.6 ([93, Corollary 3.6]) Let 2 < p < ∞ and let μ be a positive
measure. If dim(Lp (μ)) ≥ 2, then there exists a Banach space Y such that
(Lp (μ), Y) fails to have the BPBpp for operators.
Concerning Theorem 4.6, as far as we know, it is still open what happens when
1 < p < 2 (see Question 10). Nevertheless, we have the following list of pairs
satisfying the BPBpp for operators.
558 S. Dantas et al.
Theorem 4.7 ([94, Proposition 4.2 and Corollary 4.3]) Let X be an arbitrary
uniformly smooth Banach space. The pair (X , Y) has the BPBpp when
(a) Y is a uniform algebra (in particular, when Y = C(K) and Y = C0 (L)).
(b) Y has property β of Lindenstrauss.
Now another natural question arises. Do the pairs of the form (X , Lp (μ)) or
(X , p ) for 1 < p < ∞ satisfy the BPBpp for operators for every uniformly smooth
Banach space? The answer is again negative and can be checked in the following
result. Let us put some emphasis on item (b) in Theorem 4.8 below: the pair (Xp , 2p )
must have the BPBp for operators by Theorem 2.6 although this is no longer the case
for the BPBpp.
Theorem 4.8 ([94, Theorem 4.4, Corollary 4.8, and Theorem 4.9])
(a) Let 1 < p < ∞. Then, there exists a uniformly smooth Banach space X such
that (X , Lp (μ)) fails to have the BPBpp for operators. In particular, there are
uniformly smooth Banach spaces X , Z such that (X , p ) and (Z, np ) for n ≥ 2
fail the BPBpp.
(b) For each 2 ≤ p < ∞, there is a uniformly convex and uniformly smooth Banach
space Xp such that (Xp , 2p ) fails the BPBpp for operators.
It is worth mentioning that, similarly to how the BPBp has been studied for
compact operators, the authors of [94] considered the BPBpp for compact operators.
We send the reader to Section 5 of that paper. At this point, however, it is worth
noting that it seems to be unknown whether the BPBpp and the BPBpp for compact
operators are equivalent properties (see Question 11).
Up to this point, the properties considered were uniform in nature. In this section
we are going to tackle properties in which this uniform character is somehow
lost. We will treat weakenings of the Bishop–Phelps–Bollobás (point and operator)
properties. Besides their own interest, these properties were recently used succes-
sively as a tool in two different (in principle not connected) occasions. Indeed,
on the one hand, they were used to defined exactly when the projective norm on
X⊗ Bπ Y, the symmetric projective norm on ⊗ Bπ,s,N X , and the supremum norms
on P(N X ; Y) and B(X1 , . . . , XN ; Y) are (uniformly) strongly subdifferentiable
(see [90, 96–99]). On the other hand, one of these properties was used as a
tool to study norm attainment on X ⊗ Bπ Y (which is naturally connected to an
important problem on norm-attaining theory (see Question 2)) and ⊗ Bπ,s,N X (see
[89, 92]).
The Bishop–Phelps–Bollobás Theorem: An Overview 559
Surprisingly, as the reader will see in a moment, there are several positive results
on property Lo,o . Beforehand, let us justify why the study of property Lo,o is not
merely an attempt of forcing a property without much sense.
As we have mentioned already, Miguel Martín [139] proved that there are
compact operators which cannot be approximated by norm-attaining operators.
Perhaps the most important question at this very moment on norm-attaining theory is
to known whether every finite-rank operator can be approximated by norm-attaining
ones (see Question 2). Since the space of nuclear operators N(X , Y) from X into Y
satisfies that F(X , Y) ⊆ N(X , Y) ⊆ K(X , Y), it seems to be completely reasonable
to consider the class N(X , Y) and address the analogous problem in this setting.
This was done by the first and last author of this manuscript together with Luis
Carlos García Lirola, Mingu Jung, and Abraham Rueda Zoca in the recent papers
[89, 92]. Indeed, the authors used the Lo,o as a tool to provide positive results on
the denseness of nuclear operators as well as tensors in the (symmetric) projective
tensor product between Banach spaces.
Let us recall that, in the scalar-valued case, the BPBop and the BPBpp are dual
properties in the sense that (X , K) has the BPBpp if and only if (X , K) has the
BPBop (see [96, Proposition 2.2]). It seems to be natural also to consider the “local”
version of the point property and this was done for the first time by the first author
in a joint work with Sun Kwang Kim, Han Ju Lee, and Martin Mazzitelli.
2The symbol Lo,o comes from the fact that it is a local property and the double “o” means the
operator property together with the requirement that η depends on an operator. In the paper [84],
property Lo,o was called property 1 and in [143] strong BPB or sBPBp.
560 S. Dantas et al.
Definition 5.2 ([96, Definition 2.1]) 3 We say that (X , Y) has the Lp,p if given
ε > 0 and x ∈ SX , there exists η(ε, x) > 0 such that whenever T ∈ B(X , Y) with
T = 1 satisfies
Therefore, properties Lo,o and Lp,p must walk parallel with each other. It turns
out that they are closely related to the strong subdifferentiability (usually denoted by
SSD) of the norm of the Banach space as was observed by Gilles Godefroy, Vicente
Montesinos, and Václav Zizler (see [113]).
Theorem 5.3 ([96, Theorem 2.3]) Let X be a Banach space.
(a) (X , K) has the Lp,p if and only if X is SSD.
(b) (X , K) has the Lo,o if and only if X is reflexive and X is SSD.
Theorem 5.3 (together with [96, Proposition 2.6] (see also [143])) yields big
differences between properties Lp,p and the BPBpp as well as property Lo,o
and the BPBop (recall that the BPBpp and the BPBop for linear functionals
give characterizations for uniformly smooth and uniformly convex Banach spaces,
respectively). We suggest the interested reader to go to [97, page 47] or to the
discussion in [97, pages 305 and 306] to find all necessary results and references
about SSD.
Example The pair (X , K) has the
(a) Lp,p for linear functionals (but not the BPBpp) when
(a1) X = c0 .
(a2) X is the predual of the Lorentz space d∗ (w, 1).
(a3) X is the space of functions of vanishing mean oscillation (VMO), the
predual of the Hardy space H 1 .
(a4) X is n1 or n∞ when n ≥ 2.
(b) Lo,o for linear functionals (but not the BPBop) when
(b1) X is n1 or n∞ .
(b2) X is the space (⊕∞ k
k=1 ∞ )2 .
Before moving forward to the results about operators, let us highlight one result
on the Lo,o for linear functionals. The first author of this survey together with
3 Analogous to the Lo,o , we can justify the notation for property Lp,p .
The Bishop–Phelps–Bollobás Theorem: An Overview 561
Abraham Rueda Zoca showed that there is a strong connection between this property
and the compact operators (see also [143, Theorem 2.12] and [147, Theorem 2.3]).
Theorem 5.4 ([99, Theorem A]) Let X be a reflexive space. The following are
equivalent.
(a) (X , K) has the Lo,o for linear functionals.
(b) (X , Y) has the Lo,o for compact operators for every Y.
(c) X is SSD.
Contrary to the BPBp (where one has to assume finite-dimensionality on both
domain and range spaces, see Theorem 2.1), property Lo,o does not require such
strong assumptions as we can see in the next result.
Theorem 5.5 ([84, Theorem 2.4]) Let X be finite-dimensional. Then, the pair
(X , Y) has the Lo,o for every Banach space Y.
We will come back to the Lo,o in a moment. Now, we sum up all the known
results for bounded linear operators when it comes to the Lp,p . We suggest the
interested reader to check [96, Propositions 2.8, 2.9, 2.10, Theorem 2.12] and [97,
Theorem 3.6].
Theorem 5.6 ([96, Section 2]) The pair (X , Y) has the Lp,p when
(a) X and Y are finite-dimensional.
(b) (X , K) satisfies property Lp,p and Y has property β of Lindenstrauss.
(c) X = n1 and Y is
(c1) Y is finite-dimensional.
(c2) Y is uniformly convex.
(c3) Y = C0 (L).
(c4) Y = L1 (μ), where μ is a positive measure.
(c5) Y = A(D).
(c6) Y = H ∞ (D).
(c7) Y has the property β of Lindenstrauss.
(c8) Y = L1 (μ, X ), where μ is a σ -finite measure and X is as (c1)-(c7).
(d) X = c0 and Y = Lp (μ) whenever μ is a positive measure and 1 ≤ p < ∞.
In the same direction as properties Lo,o and Lp,p , the authors in [96] considered
also a local Bishop–Phelps–Bollobás property, where the η that appears in Defini-
tion 1.3 depends on a norm-one point or operator. Following the same notation, we
set Lo to mean that we have a local BPBp when η depends on an operator and we
set Lp when we have a local BPBp when η depends on a point. In that paper, the
authors were interested in differentiating all of these properties and, as far as we
know, there is not much done in the direction of properties Lo and Lp (see [96,
Proposition 3.4 and Proposition 4.5]).
In Fig. 1, we sum up all the implications that hold and next justify why all
these properties are different from each other. We do not know whether property
Lp implies the denseness of the norm-attaning operators (see Question 13).
562 S. Dantas et al.
(1) BPBp BPBop: this follows immediately from Theorem 4.3 and any positive
result for operators from Sect. 2.1. In the functional case the same happens:
recall that the BPBop for linear functionals characterizes the uniformly convex
Banach spaces and, on the other hand, the Bishop–Phelps–Bollobás theorem
holds for every Banach space.
(2) BPBp BPBpp: recall that the BPBpp for functional characterizes uniformly
smooth Banach spaces (see [95, Theorem 2.1]) and, in the operator case, if
(X , Y) has the BPBpp, then X must be uniformly smooth (see [95, Proposition
2.3]). Therefore, we can take any positive result on the BPBp from Sect. 2.1
where the domain is not uniformly smooth and we are done.
(3) Lp,p BPBpp: in the functional case, the BPBpp characterizes uniformly
smooth Banach spaces (see [95, Theorem 2.1]) and the Lp,p characterizes the
strong subdifferentiability of the norm (see [96, Theorem 2.3]). Therefore, the
pair (c0 , K) has the Lp,p but it cannot satisfy the BPBpp.
(4) Lo,o BPBop: By Theorem 5.5, the pair (X , Y) always satisfies property
Lo,o whenever X is finite-dimensional and Y is arbitrary. On the other hand, in
the functional case, the BPBop characterizes uniformly convex spaces ([125,
Theorem 2.1]) and in the operator case it does make any sense (Theorem 4.3).
(5) Lp Lp,p : Since the BPBp implies property Lp , we have (1 , Y) satisfies
Lp in many cases (see, for instance, Theorem 2.8). Nevertheless, these pairs
cannot have Lp,p since if (X , Y) has Lp,p , then X must be SSD (see [97,
Corollary 2.4]).
(6) Lp BPBp: It is known that the set NA(( ∞ k=2 k )2 , Y) is dense in
2
∞ 2
B(( k=2 k )2 , Y) [137] for every Banach space Y. By [96, Proposition 3.4],
the pair (( ∞ 2 Y) satisfies property Lp for every Banach space Y.
k=2 k )2 ,
On the other hand, if (( ∞ k=2 k )2 , Y) had the BPBp for every Banach space
2
Y, then for every ε ∈ (0, 1), there would exista ε-dense uniformly strongly
exposed family of the unit sphere of the space ( ∞ 2
k=2 k )2 (see [34, Corollary
∞ 2
3.6]) by using the fact that ( k=2 k )2 is superreflexive. Nevertheless, by
[96, Lemma 5.1 and Proposition 5.2], there exists ε0 ∈ (0, 1) such that
The Bishop–Phelps–Bollobás Theorem: An Overview 563
We can adapt Definitions 5.1 and 5.2 in a natural way to the context of bilinear
mappings. This was done for the first time in [96] (and more recently in [99]) with
the aim of classifying when the projective norm in the projective tensor product
between Banach spaces is strongly subdifferentiable. We invite the reader to go to
[96, Definition 2.1] for the formal definitions.
Concerning the Lo,o for bilinear mappings, we have the following general
characterization which yields several particular interesting cases. It deals with the
reflexivity of the projective tensor product X ⊗Bπ Y and allows us to relate property
Lo,o in different classes of functions (for linear functionals, operators, and bilinear
forms).
Theorem 5.7 ([99, Theorem B]) Let X be a strictly convex Banach space of a
Banach space satisfying the Kadec-Klee property. Let Y be an arbitrary Banach
space and assume that either X or Y satisfies the approximation property. The
following statements are equivalent.
Bπ Y, K) has the Lo,o for linear functionals.
(a) (X ⊗
B
(b) X ⊗π Y is reflexive and both (X , K) and (Y, K) have the Lo,o for linear
functionals.
(c) (X , Y; K) has the Lo,o for bilinear forms.
As a consequence, we have the following list of examples. The symbol q stands
for the conjugate index of q.
Example ([97, Proposition 2.2.(a), Lemma 2.6, and Theorem 2.7.(b)]) The triple
(X , Y; K) has the Lo,o for bilinear mappings whenever
564 S. Dantas et al.
6 Open Questions
In this section, we provide open questions on the different topics that we have treated
in this survey.
1. Bounded linear operators
It is known that when both X , Y are finite-dimensional Banach spaces, the pair
(X , Y) satisfies the BPBp for operators (see Theorem 2.1) and the proof of such
a result is done by contradiction using the compactness of the unit balls of both
spaces. The following is due to Richard Aron.
Question 1 (Richard Aron) To provide a direct (constructive) proof for the fact
that (X , Y) has the BPBp for operators whenever X and Y are finite-dimensional
spaces.
Still in the finite-dimensional vein, we have the following question. It is worth
noting that the following question is not known even when the range space is R2
endowed with the Euclidean norm.
Question 2 Is it true that all finite-rank operators can be approximated by norm-
attaining ones?
A characterization is known for the Banach spaces Y such that the pairs of
the form (1 , Y) satisfy the Bishop–Phelps–Bollobás property (see Theorem 2.8)
through the AHSP. In the same line, we have the following question.
The Bishop–Phelps–Bollobás Theorem: An Overview 565
S (X , R, η)
√
Question 15 ([123, Problem 1]) Is it true that ≤ min{ 2η, 1} for all
real Banach spaces X ?
In this section, we present some further research that has been done in the past
few years. Moreover, we present possible new lines that the interested reader could
follow. We divide the present section into subsections depending on the specific
direction that we are considering.
Stability Results There are plenty of results on stability of the Bishop–Phelps–
Bollobás property when it comes to (absolute) direct sums. This provides more
examples of pairs of Banach spaces satisfying the BPBp (for operators, in particular)
as well as counterexamples. It is worth mentioning that stability results appear quite
commonly when one starts working on the BPBp; for this, we suggest, for instance,
references [27, 69, 72, 131] and also [34, Section 2].
The Bollobás Theorem for Operators on X ⊗ Bπ Y After the papers [89] and [92],
it seems to be natural to ask when a Bollobás theorem for operators defined on tensor
products holds. In particular, the question of when it is possible to get the BPBp for
pairs of the forms (X ⊗Bπ Y, Z ⊗Bπ W ) seems to have its own interest.
The BPBp-nu for Multilinear Mappings As far as we known, Theorem 2.91,
which says that L1 (μ) fails the Bishop–Phelps–Bollobás property for numerical
radius for multilinear mappings, is the only result in this line. Perhaps, more research
in this direction would provide interesting results. Other classes of mappings for
which denseness of numerical radius attaining mappings have been studied include
N-homogeneous polynomials (see for instance [16, 73]) and holomorphic mappings
(see for instance [136, 150]), but as far as we know, no BPBp-nu property has been
studied for those classes of mappings.
Minimum Norm-Attaining Operators A relatively new line of research studies
the operators that attain their minimum norms. For T ∈ B(X , Y), its minimum
norm is defined as the number m(T ) := infx∈SX T (x). Bollobás type theorems in
this line could have their own relevance. We suggest the reader references [39, 53,
57, 58, 135].
Group Invariant Version of Bollobás Theorem Two very recent papers (see
[85, 102]) consider versions of the Bishop–Phelps theorem for group invariant
functionals and operators. In [102], Javier Falcó proved that a Bollobás theorem
in this context does not hold in general (see [102, Example, page 1611]) but he
also proved a possible extension for it (see [102, Theorem 5]). These properties for
operators have their own interest.
568 S. Dantas et al.
The Set of Operators that Satisfy a BPB Theorem Very recently, the first and
fourth authors of this survey, in a joint work with Mingu Jung, studied Bollobás
type theorems from a different perspective (see [91]): instead of looking for spaces
satisfying the Bollobás theorem, they studied the set of operators for which some
Bollobás type theorem are valid ([91, Definition 1.1]). As far as we know there is
no further research in this direction.
Strongly Subdifferentiability of ⊗ Bπ,s,N X and P(N X ) In a upcoming paper,
the first author together with Mingu Jung, Martin Mazzitelli, and Jorge Tomás
Rodríguez, study the strongly subdifferentiability of the symmetric projective tensor
product [90]. To do so, they study the Bishop–Phelps–Bollobás point property for
N-homogeneous polynomials as a tool to provide when the symmetric projective
Bπ,s,N X and P(N X ) have strongly differentiable norms.
tensor product ⊗
Related Local Properties The reader can easily notice that one can consider
the BPBop when η depends on a norm-one point x as well as the BPBpp when
η depends on a norm-one operator T (see Definitions 5.1 and 5.2). This yields
properties Lo,p and Lp,o , respectively. There are not many results in this line and we
invite the reader to check the recent paper [98], where the main aim of the authors
is to distinguish all of them from each other. On the other hand, not much is known
about the differences between the BPBp and its local versions, Lp and Lo (see [96,
Section 3]).
The following tables gather and summarize known results about the Bishop–Phelps–
Bollobás property for pairs of classical Banach spaces. In the pdf version of this
document, each cell is hyperlinked to the corresponding result within this survey.
The first column will represent domain spaces and the first row will represent range
spaces.
Remark 8.1 As an important note, the table will be mostly focused on positive
results, that is, when a pair of spaces has the BPBp. Negative results will be
reserved exclusively for universal domains and ranges, and may not be exhaustive.
For instance, in any pair where NA(X , Y) is not dense in B(X , Y), the BPBp
automatically fails, but even in pairs where every operator attains its norm, we can
find counterexamples (see Example after Theorem 2.1 for a 2-dimensional space
failing to be a universal BPB domain). Actually, many classical Banach spaces fail
to be universal BPB domains, such as c0 , 1 , ∞ L1 (μ), C(K) or any n1 or n∞ with
n > 1 (except for maybe particular cases), and it is known also that 1 , p , L1 (μ),
Lp (μ), and C(K) spaces are not universal BPB range spaces in general (except for
some particular cases). We encourage the interested reader to check the nice paper
[34], where an exposition of this problem is shown, and see also the excellent survey
[4] about norm-attaining operators.
The Bishop–Phelps–Bollobás Theorem: An Overview 569
Tables 1 and 2 use the following notation. Unless otherwise mentioned, 1 <
p, q < ∞, m, n > 1, μ, ν are any measures, K1 , K2 are any compact Hausdorff
spaces, L1 , L2 are any locally compact Hausdorff spaces, and F.D. will denote finite-
dimensional Banach spaces. Besides, we have the following additional notation.
• Symbol means that the pair has the BPBp in general, possibly under some
extra conditions specified in the subindexes.
570 S. Dantas et al.
Acknowledgments The authors of this survey would like to thank the anonymous referee for
the many useful suggestions and comments. The authors would also like to thank Ramón Aliaga,
Mingu Jung, Vladimir Kadets, Miguel Martín, Martin Mazzitelli, Jorge Tomás Rodríguez, and
Abraham Rueda Zoca for fruitful conversations on the topic of this survey.
Sheldon Dantas was supported by Spanish AEI Project PID2019 - 106529GB -
I00/MCIN/AEI/10.13039/501100011033. Domingo García and Manuel Maestre were
supported by project MTM 2017-83262-C2-1-P/MCIN/AEI/10.13039/501100011033 (FEDER)
and by PROMETEU/2021/070. Óscar Roldán was supported by the Spanish Minis-
terio de Universidades, grant FPU17/02023, and by project MTM2017-83262-C2-1-
P/MCIN/AEI/10.13039/501100011033 (FEDER).
References
1. M.D. Acosta, Operadores que alcanzan su radio numérico, Ph.D. thesis, Univ. of Granada,
1990
2. M.D. Acosta, Denseness of numerical radius attaining operators: renorming and embedding
results. Indiana Univ. Math. J. 40(3), 903–914 (1991)
3. M.D. Acosta, Every real Banach space can be renormed to satisfy the denseness of numerical
radius attaining operators. Isr. J. Math. 81(3), 273–280 (1993)
4. M.D. Acosta, Denseness of norm attaining operators. Rev. R. Acad. Cienc. Exactas Fís. Nat.
Ser. A Mat. RACSAM 100(1–2), 9–30 (2006)
5. M.D. Acosta, The Bishop–Phelps–Bollobás property for operators on C(K). Banach J. Math.
Anal. 10(2), 307–319 (2016)
6. M.D. Acosta, On the Bishop–Phelps–Bollobás property, in Function Spaces XII, 13–32.
Banach Center Publ., vol. 119 (Polish Acad. Sci. Inst. Math., Warsaw, 2019)
7. M.D. Acosta, F.J. Aguirre, R. Payá, A space by W. Gowers and new results on norm and
numerical radius attaining operators. Acta Univ. Carolin. Math. Phys. 33(2), 5–14 (1992)
8. M.D. Acosta, F.J. Aguirre, R. Payá, There is no bilinear Bishop–Phelps theorem. Isr. J. Math.
93, 221–227 (1996)
9. M.D. Acosta, R.M. Aron, D. García, M. Maestre, The Bishop–Phelps–Bollobás theorem for
operators. J. Funct. Anal. 254(11), 2780–2799 (2008)
10. M.D. Acosta, R.M. Aron, F.J. García-Pacheco, The approximate hyperplane series property
and related properties. Banach J. Math. Anal. 11(2), 295–310 (2017)
11. M.D. Acosta, J. Becerra-Guerrero, Y.S. Choi, M. Ciesielski, S.K. Kim, H.J. Lee,
M.L. Lourenço, M. Martín, The Bishop–Phelps–Bollobás property for operators between
spaces of continuous functions. Nonlinear Anal. 95, 323–332 (2014)
The Bishop–Phelps–Bollobás Theorem: An Overview 571
12. M.D. Acosta, J. Becerra-Guerrero, Y.S. Choi, D. García, S.K. Kim, H.J. Lee, M. Maestre, The
Bishop–Phelps–Bollobás property for bilinear forms and polynomials. J. Math. Soc. Japan
66(3), 957–979 (2014)
13. M.D. Acosta, J. Becerra-Guerrero, D. García, S.K. Kim, M. Maestre, Bishop–Phelps–
Bollobás property for certain spaces of operators. J. Math. Anal. Appl. 414(2), 532–545
(2014)
14. M.D. Acosta, J. Becerra-Guerrero, D. García, S.K. Kim, M. Maestre, The Bishop–Phelps–
Bollobás property: a finite-dimensional approach. Publ. Res. Inst. Math. Sci. 51(1), 173–190
(2015)
15. M.D. Acosta, J. Becerra-Guerrero, D. García, M. Maestre, The Bishop–Phelps–Bollobás
theorem for bilinear forms. Trans. Am. Math. Soc. 365(11), 5911–5932 (2013)
16. M.D. Acosta, J. Becerra-Guerrero, M. Ruiz-Galán, Numerical-radius-attaining polynomials.
Q. J. Math. 54(1), 1–10 (2003)
17. M.D. Acosta, J.L. Dávila, A basis of Rn with good isometric properties and some applications
to denseness of norm attaining operators. J. Funct. Anal. 279(6), 108602, 26 pp. (2020)
18. M.D. Acosta, J.L. Dávila, M. Soleimani-Mourchehkhorti, Characterization of the Banach
spaces Y satisfying that the pair (4∞ , Y ) has the Bishop–Phelps–Bollobás property for
operators. J. Math. Anal. Appl. 470(2), 690–715 (2019)
19. M.D. Acosta, M. Fakhar, M. Soleimani-Mourchehkhorti, The Bishop–Phelps–Bollobás
property for numerical radius of operators on L1 (μ). J. Math. Anal. Appl. 458(2), 925–936
(2018)
20. M.D. Acosta, D. García, S.K. Kim, M. Maestre, The Bishop–Phelps–Bollobás property for
operators from c0 into some Banach spaces. J. Math. Anal. Appl. 445(2), 1188–1199 (2017)
21. M.D. Acosta, D. García, M. Maestre, A multilinear Lindenstrauss theorem. J. Funct. Anal.
235(1), 122–136 (2006)
22. M.D. Acosta, M. Mastyło, M. Soleimani-Mourchehkhorti, The Bishop–Phelps–Bollobás and
approximate hyperplane series properties. J. Funct. Anal. 278(9), 2673–2699 (2018)
23. M.D. Acosta, R. Payá, Denseness of operators whose second adjoints attain their numerical
radii. Proc. Am. Math. Soc. 105(1), 97–101 (1989)
24. M.D. Acosta, R. Payá, Numerical radius attaining operators and the Radon-Nikodým
property. Bull. Lond. Math. Soc. 25(1), 67–73 (1993)
25. M.D. Acosta, M. Ruiz-Galán, Reflexive spaces and numerical radius attaining operators.
Extracta Math. 15, 247–255 (2000)
26. M.D. Acosta, M. Soleimani-Mourchehkhorti, Bishop–Phelps–Bollobás property for pos-
itive operators between classical Banach spaces, in The Mathematical Legacy of Victor
Lomonosov. Adv. Anal. Geom., vol. 2 (De Gruyter, Berlin, 2020), pp. 1–13
27. M.D. Acosta, M. Soleimani-Mourchehkhorti, Stability results of properties related to the
Bishop–Phelps–Bollobás property for operators. Sci. China Math. 64(5), 1011–1028 (2021)
28. M.D. Acosta, M. Soleimani-Mourchehkhorti, Bishop–Phelps–Bollobás property for positive
operators when the domain is L∞ . Bull. Math. Sci. 11, (2), 16 pp. (2021). Paper no. 2050023
29. M.D. Acosta, M. Soleimani-Mourchehkhorti, Bishop–Phelps–Bollobás property for positive
functionals (2021). arXiv:2106.05935
30. M.D. Acosta, M. Soleimani-Mourchehkhorti, Bishop–Phelps–Bollobás property for positive
operators when the domain is C0 (L) (2021). arXiv:2108.01638
31. L. Agud, J.M. Calabuig, S. Lajara, E.A. Sánchez-Pérez, Differentiability of Lp of a vector
measure and applications to the Bishop–Phelps–Bollobás property. Rev. R. Acad. Cienc.
Exactas Fís. Nat. Ser. A Mat. RACSAM 111(3), 735–751 (2017)
32. R.M. Aron, B. Cascales, O. Kozhushkina, The Bishop–Phelps–Bollobás theorem and Asplund
operators. Proc. Am. Math. Soc. 139(10), 3553–3560 (2011)
33. R.M. Aron, Y.S. Choi, D. García, M. Maestre, The Bishop–Phelps–Bollobás theorem for
L(L1 (μ), L∞ [0, 1]). Adv. Math. 228(1), 617–628 (2011)
34. R.M. Aron, Y.S. Choi, S.K. Kim, H.J. Lee, M. Martín, The Bishop–Phelps–Bollobás version
of Lindenstrauss properties A and B. Trans. Am. Math. Soc. 367(9), 6085–6101 (2015)
572 S. Dantas et al.
35. R.M. Aron, C. Finet, E. Werner, Some remarks on norm-attaining n-linear forms, in Function
Spaces (Edwardsville, IL, 1994). Lecture Notes in Pure and Appl. Math., vol. 172 (Dekker,
New York, 1995), pp. 19–28
36. R.M. Aron, D. García, M. Maestre, On norm attaining polynomials. Publ. Res. Inst. Math.
Sci. 39(1), 165–172 (2003)
37. A. Avilés, A.J. Guirao, J. Rodríguez, On the Bishop–Phelps–Bollobás property for numerical
radius in C(K)-spaces. J. Math. Anal. Appl. 419(1), 395–421 (2014)
38. N. Bala, K. Dhara, J. Sarkar, A. Sensarma, A Bishop–Phelps–Bollobás theorem for bounded
analytic functions (2021). arXiv:2109.10125
39. N. Bala, G. Ramesh, A Bishop–Phelps–Bollobás type property for minimum attaining
operators. Oper. Matrices 15(2), 497–513 (2021)
40. I.D. Berg, B. Sims, Denseness of operators which attain their numerical radius. J. Aust. Math.
Soc. Ser. A 36(1), 130–133 (1984)
41. E. Bishop, R.R. Phelps, A proof that every Banach space is subreflexive. Bull. Am. Math.
Soc. 67, 97–98 (1961)
42. B. Bollobás, An extension to the theorem of Bishop, Phelps. Bull. Lond. Math. Soc. 2, 181–
182 (1970)
43. J. Bourgain, On dentability and the Bishop–Phelps property. Isr. J. Math. 28(4), 265–271
(1977)
44. S.A. Buss, Versiones locales y uniformes del Teorema de Bishop–Phelps–Bollobás, Bache-
lor’s thesis, National University of Comahue, 2019
45. F. Cabello-Sánchez, S. Dantas, V. Kadets, S.K. Kim, H.J. Lee, M. Martín, On Banach spaces
whose group of isometries acts micro-transitively on the unit sphere. J. Math. Anal. Appl.
488(1), 124046, 14 pp. (2020)
46. Á. Capel, M. Martín, J. Merí, Numerical radius attaining compact linear operators. J. Math.
Anal. Appl. 445(2), 1258–1266 (2017)
47. D. Carando, S. Lassalle, M. Mazzitelli, On the polynomial Lindenstrauss theorem. J. Funct.
Anal. 263(7), 1809–1824 (2012)
48. D. Carando, M. Mazzitelli, Bounded holomorphic functions attaining their norms in the
bidual. Publ. Res. Inst. Math. Sci. 51(3), 489–512 (2015)
49. D. Carando, J.T. Rodríguez, Symmetric multilinear forms on Hilbert spaces: Where do they
attain their norm?. Linear Algebra Appl. 563, 178–192 (2019)
50. C.S. Cardassi, Numerical radius attaining operators, in Banach Spaces (Columbia, Mo.,
1984). Lecture Notes in Math., vol. 1166 (Springer, Berlin, 1985), pp. 11–14
51. C.S. Cardassi, Density of numerical radius attaining operators on some reflexive spaces. Bull.
Aust. Math. Soc. 31(1), 1–3 (1985)
52. C.S. Cardassi, Numerical radius-attaining operators on C(K). Proc. Am. Math. Soc. 95(4),
537–543 (1985)
53. X. Carvajal, W. Neves, Operators that attain their minima. Bull. Braz. Math. Soc. (N.S.) 45(2),
293–312 (2014)
54. B. Cascales, R. Chiclana, L.C. García-Lirola, M. Martín, A. Rueda-Zoca, On strongly norm
attaining Lipschitz maps. J. Funct. Anal. 277(6), 1677–1717 (2019)
55. B. Cascales, A.J. Guirao, V. Kadets, A Bishop–Phelps–Bollobás type theorem for uniform
algebras. Adv. Math. 240, 370–382 (2013)
56. B. Cascales, A.J. Guirao, V. Kadets, M. Soloviova, -flatness and Bishop–Phelps–Bollobás
type theorems for operators. J. Funct. Anal. 274(3), 863–888 (2018)
57. U.S. Chakraborty, Some remarks on minimum norm attaining operators. J. Math. Anal. Appl.
492(2), 124492, 14 pp. (2020)
58. U.S. Chakraborty, Some Bishop–Phelps–Bollobás type properties in Banach spaces with
respect to minimum norm of bounded linear operators. Ann. Funct. Anal. 12(3), 15 pp. (2021).
Paper no. 46
59. L.X. Cheng, Q.J. Cheng, K.K. Xu, W. Zhang, Z.M. Zheng, A Bishop–Phelps–Bollobás
theorem for Asplund operators. Acta Math. Sin. (Engl. Ser.) 36(7), 765–782 (2020)
The Bishop–Phelps–Bollobás Theorem: An Overview 573
60. L. Cheng, D. Dai, Y. Dong, A sharp operator version of the Bishop–Phelps theorem for
operators from 1 to CL-spaces. Proc. Am. Math. Soc. 141(3), 867–872 (2013)
61. M. Chica, V. Kadets, M. Martín, J. Merí, Further properties of the Bishop–Phelps–Bollobás
moduli. Mediterr. J. Math. 13(5), 3173–3183 (2016)
62. M. Chica, V. Kadets, M. Martín, J. Merí, M. Soloviova, Two refinements of the Bishop–
Phelps–Bollobás modulus. Banach J. Math. Anal. 9(4), 296–315 (2015)
63. M. Chica, V. Kadets, M. Martín, S. Moreno-Pulido, F. Rambla-Barreno, Bishop–Phelps–
Bollobás moduli of a Banach space. J. Math. Anal. Appl. 412(2), 697–719 (2014)
64. R. Chiclana, L.C. García-Lirola, M. Martín, A. Rueda-Zoca, Examples and applications of
the density of strongly norm attaining Lipschitz maps. Rev. Mat. Iberoam. 37(5), 1917–1951
(2021)
65. R. Chiclana, M. Martín, The Bishop–Phelps–Bollobás property for Lipschitz maps. Nonlinear
Anal. 188, 158–178 (2019)
66. R. Chiclana, M. Martín, Some stability properties for the Bishop–Phelps–Bollobás property
for Lipschitz maps. Stud. Math. 264(2), 121–147 (2022)
67. D.H. Cho, Y.S. Choi, The Bishop–Phelps–Bollobás theorem on bounded closed convex sets.
J. Lond. Math. Soc. (2) 93(2), 502–518 (2016)
68. G. Choi, Y.S. Choi, M. Martín, Emerging notions of norm attainment for Lipschitz maps
between Banach spaces. J. Math. Anal. Appl. 483(1), 123600, 24 pp. (2020)
69. G. Choi, S.K. Kim, The Bishop–Phelps–Bollobás property on the space of c0 -sum, Mediterr.
J. Math. 19(2), 16 pp. (2022). Paper No. 72
70. Y.S. Choi, Norm attaining bilinear forms on L1 [0, 1]. J. Math. Anal. Appl. 211(1), 295–300
(1997)
71. Y.S. Choi, S. Dantas, M. Jung, The Bishop–Phelps–Bollobás properties in complex Hilbert
spaces. Math. Nachr. 294(11), 2105–2120 (2021)
72. Y.S. Choi, S. Dantas, M. Jung, M. Martín, The Bishop–Phelps–Bollobás property and
absolute sums. Mediterr. J. Math. 16(3), 24 pp. (2019). Paper no. 73
73. Y.S. Choi, D. García, S.G. Kim, M. Maestre, Norm or numerical radius attaining polynomials
on C(K). J. Math. Anal. Appl. 295(1), 80–96 (2004)
74. Y.S. Choi, D. García, S.G. Kim, M. Maestre, The polynomial numerical index of a Banach
space. Proc. Edinb. Math. Soc. (2) 49(1), 39–52 (2006)
75. Y.S. Choi, D. García, S.K. Kim, M. Maestre, Some geometric properties of disk algebras. J.
Math. Anal. Appl. 409(1), 147–157 (2014)
76. Y.S. Choi, S.G. Kim, Norm or numerical radius attaining multilinear mappings and polyno-
mials. J. Lond. Math. Soc. (2) 54(1), 135–147 (1996)
77. Y.S. Choi, S.K. Kim, The Bishop–Phelps–Bollobás theorem for operators from L1 (μ) to
Banach spaces with the Radon–Nikodým property. J. Funct. Anal. 261(6), 1446–1456 (2011)
78. Y.S. Choi, S.K. Kim, The Bishop–Phelps–Bollobás property and lush spaces. J. Math. Anal.
Appl. 390(2), 549–555 (2012)
79. Y.S. Choi, S.K. Kim, H.J. Lee, M. Martín, The Bishop–Phelps–Bollobás theorem for
operators on L1 (μ). J. Funct. Anal. 267(1), 214–242 (2014)
80. Y.S. Choi, S.K. Kim, H.J. Lee, M. Martín, On Banach spaces with the approximate hyperplane
series property. Banach J. Math. Anal. 9(4), 243–258 (2015)
81. Y.S. Choi, H.G. Song, The Bishop–Phelps–Bollobás theorem fails for bilinear forms on 1 ×
1 . J. Math. Anal. Appl. 360(2), 752–753 (2009)
82. D. Dai, The Bishop–Phelps–Bollobás theorem for bilinear mappings. Adv. Math. (China)
44(1), 105–110 (2015)
83. A. Dalet, G. Lancien, Some properties of coarse Lipschitz maps between Banach spaces.
North-West. Eur. J. Math. 3, 41–62 (2017)
84. S. Dantas, Some kind of Bishop–Phelps–Bollobás property. Math. Nachr. 290(5–6), 774–784
(2017)
85. S. Dantas, J. Falcó, M. Jung, Group invariant operators and some applications on norm-
attaining theory (2021). arXiv:2110.02066
574 S. Dantas et al.
86. S. Dantas, D. García, S.K. Kim, U.Y. Kim, H.J. Lee, M. Maestre. A nonlinear Bishop–Phelps–
Bollobás type theorem. Q. J. Math. 70(1), 7–16 (2019)
87. S. Dantas, D. García, S.K. Kim, H.J. Lee, M. Maestre, On the Bishop–Phelps–Bollobás
theorem for multilinear mappings. Linear Algebra Appl. 532, 406–431 (2017)
88. S. Dantas, D. García, M. Maestre, M. Martín, The Bishop–Phelps–Bollobás property for
compact operators. Can. J. Math. 70(1), 53–73 (2018)
89. S. Dantas, L.C. García-Lirola , M. Jung, A. Rueda-Zoca, On norm-attainment in (symmetric)
tensor products (2021). arXiv:2104.06841
90. S. Dantas, M. Jung, M. Mazzitelli, J.T. Rodríguez, On the strong subdifferentiability of the
homogeneous polynomials and (symmetric) tensor products (in preparation)
91. S. Dantas, M. Jung, Ó. Roldán, Norm-attaining operators which satisfy a Bollobás type
theorem. Banach J. Math. Anal. 15(2), 26 pp. (2021). Paper no. 40
92. S. Dantas, M. Jung, Ó. Roldán, A. Rueda-Zoca, Norm-attaining tensors and nuclear operators.
Mediterranean J. Math. (to be formally accepted). arXiv:2006.09871
93. S. Dantas, V. Kadets, S.K. Kim, H.J. Lee, M. Martín, On the pointwise Bishop–Phelps–
Bollobás property for operators. Can. J. Math. 71(6), 1421–1443 (2019)
94. S. Dantas, V. Kadets, S.K. Kim, H.J. Lee, M. Martín, There is no operatorwise version of the
Bishop–Phelps–Bollobás property. Linear Multilinear Algebra 68(9), 1767–1778 (2020)
95. S. Dantas, S.K. Kim, H.J. Lee, The Bishop–Phelps–Bollobás point property. J. Math. Anal.
Appl. 444(2), 1739–1751 (2016)
96. S. Dantas, S.K. Kim, H.J. Lee, M. Mazzitelli, Local Bishop–Phelps–Bollobás properties. J.
Math. Anal. Appl. 468(1), 304–323 (2018)
97. S. Dantas, S.K. Kim, H.J. Lee, M. Mazzitelli, Strong subdifferentiability and local Bishop–
Phelps–Bollobás properties. Rev. R. Acad. Cienc. Exactas Fís. Nat. Ser. A Mat. RACSAM
114(2), 16 pp. (2020). Paper no. 47
98. S. Dantas, S.K. Kim, H.J. Lee, M. Mazzitelli, On some local Bishop–Phelps–Bollobás
properties, in The Mathematical Legacy of Victor Lomonosov. Adv. Anal. Geom., vol. 2, (De
Gruyter, Berlin, 2020), pp. 109–121
99. S. Dantas, A. Rueda-Zoca, A characterization of a local vector valued Bollobás theorem.
Results Math. 76(4), 14 pp. (2021). Paper no. 167
100. A. Defant, K. Floret, Tensor Norms and Operator Ideals. North-Holland Mathematics
Studies, vol. 176 (Elsevier, Amsterdam, 1993)
101. J. Falcó, The Bishop–Phelps–Bollobás property for numerical radius on L1 . J. Math. Anal.
Appl. 414(1), 125–133 (2014)
102. J. Falcó, A group invariant Bishop–Phelps theorem. Proc. Am. Math. Soc. 149(4), 1609–1612
(2021)
103. H. Fetter, B. Gamboa de Buen, The James Forest. London Mathematical Society Lecture
Notes Series, vol. 236 (Cambridge University Press, Cambridge, 1997)
104. D. García, H.J. Lee, M. Maestre, The Bishop–Phelps–Bollobás property for Hermitian forms
on Hilbert spaces. Q. J. Math. 65(1), 201–209 (2014)
105. D. García, M. Maestre, M. Martín, Ó. Roldán, On the compact operators case of the Bishop–
Phelps–Bollobás property for numerical radius. Results Math. 76(3), 23 pp. (2021). Paper no.
122
106. L.C. García-Lirola, C. Petitjean, A. Procházka, A. Rueda Zoca, Extremal structure and
Duality of Lipschitz free space. Mediterr. J. Math. 15(2), 23 pp. (2018). Paper no. 69
107. L.C. García-Lirola, A. Procházka, A. Rueda Zoca, On the structure of spaces of vector-valued
Lipschitz functions. Stud. Math. 239(3), 249–271 (2017)
108. F.J. García-Pacheco, The AHSP is inherited by E-summands. Adv. Oper. Theory 2(1), 17–20
(2017)
109. F.J. García-Pacheco, S. Moreno-Pulido, The Bishop–Phelps–Bollobás modulus for function-
als on classical Banach spaces. Adv. Oper. Theory 4(1), 1–23 (2019)
110. F.J. García-Pacheco, S. Moreno-Pulido, The Bishop–Phelps–Bollobás modulus for operators.
Acta Sci. Math. (Szeged) 85(1–2), 189–201 (2019)
The Bishop–Phelps–Bollobás Theorem: An Overview 575
111. G. Godefroy, A survey on Lipschitz-free Banach spaces. Comment. Math. 55(2), 89–118
(2015)
112. G. Godefroy, On norm attaining Lipschitz maps between Banach spaces. Pure Appl. Funct.
Anal. 1(1), 39–46 (2016)
113. G. Godefroy, V. Montesinos, V. Zizler, Strong subdifferentiability of norms and geometry of
Banach spaces. Comment. Math. Univ. Carolin. 36(3), 493–502 (1995)
114. T. Grando, M.L. Lourenço, On a function module with approximate hyperplane series
property. J. Aust. Math. Soc. 108(3), 341–348 (2020)
115. A.J. Guirao, O. Kozhushkina, The Bishop–Phelps–Bollobás property for numerical radius in
1 (C). Stud. Math. 218(1), 41–54 (2013)
116. R.E. Huff, Dentability and the Radon-Nikodým property. Duke Math. J. 41, 111–114 (1974)
117. R.C. James, Reflexivity and the supremum of linear functionals. Ann. Math. 66, 159–169
(1957)
118. R.C. James, Characterizations of reflexivity. Stud. Math. 23, 205–216 (1964)
119. M. Jiménez Sevilla, R. Payá, Norm attaining multilinear forms and polynomials on preduals
of Lorentz sequence spaces. Stud. Math. 127(2), 99–112 (1998)
120. J. Johnson, J. Wolfe, Norm attaining operators. Stud. Math. 65(1), 7–19 (1979)
121. V. Kadets, M. Martín, M. Soloviova, Norm-attaining Lipschitz functionals. Banach J. Math.
Anal. 10(3), 621–637 (2016)
122. V. Kadets, M. Soloviova, A modified Bishop–Phelps–Bollobás theorem and its sharpness.
Mat. Stud. 44(1), 84–88 (2015)
123. V. Kadets, M. Soloviova, Quantitative version of the Bishop–Phelps–Bollobás theorem for
operators with values in a space with the property β. Mat. Stud. 47(1), 71–90 (2017)
124. S.K. Kim, The Bishop–Phelps–Bollobás theorem for operators from c0 to uniformly convex
spaces. Isr. J. Math. 197(1), 425–435 (2013)
125. S.K. Kim, H.J. Lee, Uniform Convexity and Bishop–Phelps–Bollobás property. Can. J. Math.
66(2), 373–386 (2014)
126. S.K. Kim, H.J. Lee, Simultaneously continuous retraction and Bishop–Phelps–Bollobás type
theorem. J. Math. Anal. Appl. 420(1), 758–771 (2014)
127. S.K. Kim, H.J. Lee, The Bishop–Phelps–Bollobás property for operators from C(K) to
uniformly convex spaces. J. Math. Anal. Appl. 421(1), 51–58 (2015)
128. S.K. Kim, H.J. Lee, A Urysohn-type theorem and the Bishop–Phelps–Bollobás theorem for
holomorphic functions. J. Math. Anal. Appl. 480(2), 123393, 8 pp. (2019)
129. S.K. Kim, H.J. Lee, P.K. Lin, The Bishop–Phelps–Bollobás property for operators from
L∞ (μ) to uniformly convex Banach spaces. J. Nonlinear Convex Anal. 17(2), 243–249
(2016)
130. S.K. Kim, H.J. Lee, M. Martín, On the Bishop–Phelps–Bollobás property for numerical
radius. Abstr. Appl. Anal. 2014, 479208, 15 pp. (2014)
131. S.K. Kim, H.J. Lee, M. Martín, The Bishop–Phelps–Bollobás theorem for operators from 1
sums of Banach spaces. J. Math. Anal. Appl. 428(2), 920–929 (2015)
132. S.K. Kim, H.J. Lee, M. Martín, Bishop–Phelps–Bollobás property for bilinear forms on
spaces of continuous functions. Math. Z. 283(1–2), 157–167 (2016)
133. S.K. Kim, H.J. Lee, M. Martín, On the Bishop–Phelps–Bollobás theorem for operators and
numerical radius. Stud. Math. 233(2), 141–151 (2016)
134. S.K. Kim, H.J. Lee, M. Martín, J. Merí, On a second numerical index for Banach spaces.
Proc. R. Soc. Edinburgh Sect. A 150(2), 1003–1051 (2020)
135. S.H. Kulkarni, G. Ramesh, On the denseness of minimum attaining operators. Oper. Matrices
12(3), 699–709 (2018)
136. H.J. Lee, Denseness of numerical radius attaining holomorphic functions. J. Inequal. Appl.
981453, 5 pp. (2009)
137. J Lindenstrauss, On operators which attain their norm. Isr. J. Math. 1, 139–148 (1963)
138. J. Lindenstrauss, L. Tzafriri, Classical Banach spaces. I (Springer, Berlin, New York, 1977)
139. M. Martín, Norm-attaining compact operators. J. Funct. Anal. 267(5), 1585–1592 (2014)
576 S. Dantas et al.
140. M. Martín, The version for compact operators of Lindenstrauss properties A and B. Rev. R.
Acad. Cienc. Exactas Fís. Nat. Ser. A Mat. RACSAM 110(1), 269–284 (2016)
141. R. Payá, A counterexample on numerical radius attaining operators. Isr. J. Math. 79(1), 83–
101 (1992)
142. R.A. Ryan, Introduction to Tensor Products of Banach Spaces. Springer Monographs in
Mathematics (Springer, London, 2002)
143. D. Sain, Smooth points in operator spaces and some Bishop–Phelps–Bollobás type theorems
in Banach spaces. Oper. Matrices 13(2), 433–445 (2019)
144. W. Schachermayer, Norm attaining operators on some classical Banach spaces. Pac. J. Math.
105(2), 427–438 (1983)
145. W. Schachermayer, Norm attaining operators and renormings of Banach spaces. Isr. J. Math.
44(3), 201–212 (1983)
146. B. Sims, On numerical range and its applications to Banach algebras, Ph.D. thesis, Univ. of
Newcastle, 1972
147. J. Talponen, Note on a kind of Bishop–Phelps–Bollobás property for operators (2017).
arXiv:1707.03251
148. J.J. Uhl, Norm attaining operators on L1 [0, 1] and the Radon-Nikodým property. Pacific J.
Math. 63(1), 293–300 (1976)
149. M. Weaver, Lipschitz Algebras (World Scientific Publishing Co., Inc., River Edge, NJ, 1999)
150. R. Zarghami, Coincidence the sets of norm and numerical radius attaining holomorphic
functions on finite-dimensional spaces. Acta Univ. Apulensis Math. Inform. 25, 229–233
(2011)
151. V. Zizler, On some extremal problems in Banach spaces. Math. Scand. 32, 214–224 (1973)
A New Proof of the Power Weighted
Birman–Hardy–Rellich Inequalities
Abstract In this chapter, we introduce a new method for proving the power-
weighted Birman–Hardy–Rellich integral inequalities,
$ ∞ $ ∞
2 2
dx x α f (m) (x) A(, α) dx x α−2 f (m−) (x) ,
0 0
−
A(, α) = 4 (2j − 1 − α)2 .
j =1
The new method of proof simultaneously establishes both the existence of such
inequalities and their optimality (i.e., sharpness of the constant A(, α) on the space
C0∞ ((0, ∞)) of infinitely differentiable functions of compact support in (0, ∞)).
We also note that these inequalities are strict, that is, equality holds if and only if
f ≡ 0.
F. Gesztesy ()
Department of Mathematics, Baylor University, Waco, TX, USA
e-mail: [email protected]
https://www.baylor.edu/math/index.php?id=935340
I. Michael
Department of Mathematics, Louisiana State University, Baton Rouge, LA, USA
e-mail: [email protected]
https://www.math.lsu.edu/~imichael
M. M. H. Pang
Department of Mathematics, University of Missouri, Columbia, MO, USA
e-mail: [email protected]
https://www.math.missouri.edu/people/pang
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 577
R. M. Aron et al. (eds.), Operator and Norm Inequalities and Related Topics,
Trends in Mathematics, https://doi.org/10.1007/978-3-031-02104-6_17
578 F. Gesztesy et al.
.
Extensions to homogeneous Sobolev spaces, that is, f ∈ H0m (0, ρ); x α dx , as
.
well as the vector-valued case, where f ∈ H0m (0, ρ); x α dx; H , ρ ∈ (0, ∞)∪{∞},
with H a complex, separable Hilbert space, are also discussed.
1 Introduction
The primary aim in this paper is to provide a new proof of the optimal version of
the power-weighted sequence of Birman–Hardy–Rellich inequalities of the form,
$ ∞ $ ∞
2 2
dx x α f (m) (x) A(, α) dx x α−2 f (m−) (x) ,
0 0 (1)
m ∈ N, ∈ {1, . . . , m}, α ∈ R, f ∈ C0∞ ((0, ∞)),
where
A(, α) = 4− (2j − 1 − α)2 , ∈ N, α ∈ R. (2)
j =1
The novelty of our proof lies in the fact that both the existence and the optimality of
the constants A(, α) in (1) are established simultaneously.
Moreover, we also prove these inequalities in the context of homogeneous
.
Sobolev spaces, that is, for f ∈ H0m (0, ρ); x α dx , as well as in the vector-valued
.
case for f ∈ H0m (0, ρ); x α dx; H , ρ ∈ (0, ∞) ∪ {∞}, and the Sobolev space
H0m ((ρ, ∞); x α dx; H), ρ ∈ (0, ∞), where H is a complex, separable Hilbert space.
We recall that the special case α = 0 appeared in work of Birman in 1961
(English translation in 1966) [3] (see also [9, pp. 83–84]). The case m = 1 in (1)
represents Hardy’s celebrated inequality [11], [12, Sect. 9.8] (see also [16, Chs. 1,
3, App.]), the case m = 2 is due to Rellich [24, Sect. II.7] (actually, in the multi-
dimensional context). The inequalities (1) are known to be strict, that is, equality
holds in (1) if and only if f = 0 on (0, ∞). Moreover, they are known to be optimal,
that is, the constants A(, α) in (1) are sharp, although, this must be qualified as
different authors frequently prove sharpness for different function spaces. In the
present one-dimensional context at hand, sharpness of (1) is often proved in an
integral form (rather than the currently presented differential form) where f (m) on
the left-hand side is replaced by F and f on the right-hand side by m repeated
integrals over F . For pertinent one-dimensional sources, we refer, for instance, to
[2, p. 3–5], [4], [6, p. 104–105], [8, 10, 11], [12, p. 240–243], [16, Ch. 3], [17,
p. 5–11], [19, 20, 23]. We also note that higher-order Hardy inequalities, including
various weight functions, are discussed in [5], [15, Sect. 5], [16, Chs. 2–5], [17,
A New Proof of the Power Weighted Birman–Hardy–Rellich Inequalities 579
Chs. 1–4], [18], and [22, Sect. 10] (however, with the exception of [5], Birman’s
sequence of inequalities, i.e., (1) for α = 0, is not mentioned in these sources).
There exists a wealth of multi-dimensional investigations of Hardy, Rellich, etc.,
inequalities on domains ⊆ Rn , n ∈ N, n 2, which, when specialized to a ball
in Rn and spherically symmetric functions f , yields one dimensional inequalities of
the Birman–Hardy–Rellich-type with various weight functions. Since we included
a very detailed bibliography in [7], including such multi-dimensional sources, we
refrained from repeating it here and just focused on the one dimensional literature.
Briefly turning to the contents of each section, we introduce our new proof,
a variant of a combination of transformations studied by Hartman [13], [14,
p. 324–325] and Müller-Pfeiffer [21, p. 200–207], in Sect. 2. Generalizations to
homogeneous Sobolev spaces and to the vector-valued case (replacing complex-
valued f ( · ) by f ( · ) ∈ H, with H a complex, separable Hilbert space) appear in
Sect. 3. For background on the vector-valued case we refer to Appendix A.
Throughout this paper, H represents a complex, separable Hilbert space with
corresponding scalar product ( · , · )H (linear in the second factor) and associated
norm · H .
m
A(m, α) = 4−m (2j − 1 − α)2 , m ∈ N, α ∈ R. (3)
j =1
Then,
$ ∞ $ ∞
2 2
α
dx x f (m)
(x) A(m, α) dx x α−2m f (x) ,
0 0 (4)
m ∈ N, α ∈ R, f ∈ C0∞ ((0, ∞)).
Moreover, the constant A(m, α) in (4) is optimal and the inequality is strict, that is,
equality holds in (4) if and only if f ≡ 0.
580 F. Gesztesy et al.
Proof Let C ∈ (0, ∞) and define Q as the operator in L2 ((0, ∞); dx) given by
dm dm
Q = (−1)m m x α m − Cx α−2m
. (5)
dx dx C0∞ ((0,∞))
Utilizing
$ $
b
2
b (m)
dx x α f (m) (x) = (−1)m dx x α f (m) (x) f (x),
a a (6)
m ∈ N, α ∈ R, f ∈ C0∞ ((a, b)), 0 a < b ∞,
Thus, to establish (4) and, simultaneously, optimality of the constant A(m, α), we
will show that
yields
(m)
2m
x α f (m) (x) = e−[(2m+1−α)/2]t c (m, α)w() (t), (12)
=0
One notes that the solutions (14) are linearly independent due to (9).
Thus, recalling (10)–(12), it follows that the solutions of
2m
c (m, α)w() (t) = 0, t ∈ R, (15)
=0
and
Continuing iteratively, one concludes that the linearly independent solutions of (15)
are of the form
are linearly independent solutions of (15). The zeros of the characteristic polynomial
of (15) are thus the constant factors in the exponents of (21). Hence, the character-
582 F. Gesztesy et al.
2m
Pm,α (λ) = c (m, α)λ
=0
Thus, the coefficients c (m, α), = 0, 1, . . . , 2m, satisfy the following properties:
$ ∞
m
= dt w(t) (−1) |c2j (m, α)|w (t) − Cw(t) .
j (2j )
(24)
−∞ j =0
Hence, Q 0 in L2 ((0, ∞); dx) if and only if the constant coefficient operator S
in L2 (R; dt), defined by
m
d 2j
S= (−1) |c2j (m, α)| 2j − C
j
, (25)
dt C0∞ (R)
j =0
m
T in L2 (R; dξ ) by the polynomial j =1 |c2j (m, α)|ξ + |c0 (m, α)| − C , that is,
2j
m
(T v)(ξ ) = |c2j (m, α)|ξ 2j v(ξ ) + |c0 (m, α)| − C v(ξ ),
j =1
(26)
$ ∞
v ∈ dom(T ) = u ∈ L2 (R; dξ ) dξ ξ |u(ξ )| < ∞ .
4m 2
−∞
Indeed,
m−
A(, α)A(m − , α − 2) = 4− (2j − 1 − α)2 4−(m−) [2k − 1 − (α − 2)]2
j =1 k=1
m−
= 4−m (2j − 1 − α)2 [2(k + ) − 1 − α]2
j =1 k=1
m
= 4−m (2j − 1 − α)2 (2k − 1 − α)2
j =1 k=+1
m
= 4−m (2j − 1 − α)2
j =1
3 Some Generalizations
Proof Using the notations in the proof of Lemma A.3, Theorem 2.2 implies,
$ $ ∞
∞
α (m)
2 ∞
(x)H =
2
dx x f dx x α f (m) (x), ϕk H
0 0 k=1
∞ $ ∞
(m) 2
= dx x α fk (x)
k=1 0
∞
$ ∞ (m−) 2
A(, α) dx x α−2 fk (x)
k=1 0
∞ $
∞ 2
= A(, α) dx x α−2 f (m−) k (x)
k=1 0
$ ∞ ∞
2
= A(, α) dx x α−2
f (m−) (x), ϕk H
0 k=1
$ ∞ 2
= A(, α) dx x α−2 f (m−) (x)H . (33)
0
Definition 3.2 Let m ∈ N, α ∈ R.
(i) Define
.
H m (0, ∞); x α dx; H = f : (0, ∞) → H f (j ) ∈ ACloc ((0, ∞); H);
$
∞ 2
j = 0, 1, . . . , m − 1; dx x α f (m) (x)H < ∞ . (34)
0
We also introduce,
$ .
∞ 2
|||f |||2m,α = dx x α f (m) (x)H , f ∈ H m (0, ∞); x α dx; H . (35)
0
.
= f ∈ H m (0, ∞); x α dx; H there exists a Cauchy sequence {fn }∞
n=1
in C0∞ ((0, ∞); H), ||| · |||m,α such that, for all 0 < a < b < ∞,
$ b
2
lim
dx fn (x) − f (x) H = 0 . (37)
n→∞ a
Remark 3.3 To see that (36) and (37) describe the same space, we first note that
∞ .m
if {fn }∞
n=1 ⊆ C0 ((0, ∞); H), ||| · |||m,α and f ∈ H (0, ∞); x α dx; H
satisfy (36), then, by putting = m in (36), (37) is satisfied. Next, suppose
.
f ∈ H m (0, ∞); x α dx; H satisfies (37). By Lemma A.4 there exists a unique
.
g ∈ H m (0, ∞); x α dx; H such that for = 0, 1, . . . , m,
$ ∞ 2
lim dx x α−2 fn(m−) (x) − g (m−) (x)H = 0. (38)
n→∞ 0
for all 0 < a < b < ∞, hence f = g since (37) is true by assumption. Thus, (38)
implies (36). D
.
Clearly, C0∞ ((0, ∞)); H) ⊆ H m (0, ∞); x α dx; H . Moreover, by Lemma A.4,
.
H0m (0, ∞); x α dx; H can be identified with the completion of C0∞ ((0, ∞); H)
with respect to the norm ||| · |||m,α , that is,
. ||| · |||m,α
H0m (0, ∞); x α dx; H = C0∞ ((0, ∞); H) . (40)
.m ∞
∞ Let f ∈ H0 (0, ∞); x dx; H and let {fn }n=1 be a Cauchy sequence in
Proof α
C0 ((0, ∞); H), ||| · |||m,α satisfying (36). Then, for all ∈ {1, . . . , m},
$ $
∞ 2 ∞ 2
dx x α f (m) (x)H = lim dx x α fn(m) (x)H
0 n→∞ 0
$ ∞ 2
A(, α) lim dx x α−2fn(m−) (x)H
n→∞ 0
$ ∞
2
= A(, α) dx x α−2f (m−) (x)H , (42)
0
m $
∞ 2
0 j m − 1; f 2m,α = dx x α f (j ) (x)H < ∞ . (43)
j =0 0
(ii) Define
H0m ((0, ∞); x α dx; H) = f ∈ H m ((0, ∞); x α dx; H) there exists a Cauchy
∞
sequence {fn }∞
n=1 in C0 ((0, ∞); H), · m,α such that
lim fn − f m,α = 0 (44)
n→∞
= f ∈ H m ((0, ∞); x α dx; H) there exists a Cauchy sequence {fn }∞
n=1 in
C0∞ ((0, ∞); H), · m,α such that for any 0 < a < b < ∞ we have
$ b
lim dx fn (x) − f (x)2H = 0 . (45)
n→∞ a
Remark 3.6 To see that (44) and (45) describe the same space,
we first note
that if {fn }∞
n=1 ⊆ C ∞ ((0, ∞); H), ·
0 m,α and f ∈ H m (0, ∞); x α dx; H
H m (0, ∞); x α dx; H satisfies (45). By Lemma A.6 there is a unique g ∈
H m (0, ∞); x α dx; H such that
$ ∞ 2
lim dx x α fn (x) − g(x)H lim fn − gm,α = 0, (46)
n→∞ 0 n→∞
By the assumption that (45) is true, this gives f = g. Thus (46) implies (44). D
∞
Clearly
H) ⊆ H (0, ∞); x dx; H and · m,α is a norm on
C0 ((0, ∞)); m α
H m (0, ∞); x α dx; H . By Lemma A.6, H0m ((0, ∞); x α dx; H) can be identified
with the completion of C0∞ ((0, ∞); H) with respect to · m,α , that is,
· m,α
H0m ((0, ∞); x α dx; H) = C0∞ ((0, ∞); H) . (48)
lim fn(m−)
k
(x) = f (m−) (x) for a.e. x ∈ (0, ∞). (51)
k→∞
$ ∞ 2
= lim dx x α fn(m)
k
(x)
H
k→∞ 0
$ ∞ 2
= dx x α f (m) (x)H (by (44)). (52)
0
Next, we consider bounded intervals (0, ρ), 0 < ρ < ∞, and recall a simplified
version of [7, Theorem 3.1 (iii)].
Lemma 3.8 Let γ ∈ (0, ∞), ρ ∈ (0, γ ), and set
m
m
B(m, α) = 4−m (2j − 1 − α)2 > 0, m ∈ N, α ∈ R. (53)
k=1 j =1,j =k
Then,
$ ρ $ ρ
dx x α−2 f (m−) (x) [ln(γ /x)]−2,
2 2
dx x f α (m)
(x) B(, α)
0 0 (54)
m ∈ N, ∈ {1, . . . , m}, α ∈ R, f ∈ C0∞ ((0, ρ)).
∞ $
ρ
dx x α−2 f (m−) k (x) [ln(γ /x)]−2
2
= B(, α)
k=1 0
590 F. Gesztesy et al.
$ ∞
ρ (m−) 2
= B(, α) dx x α−2 f (x), ϕk H [ln(γ /x)]−2
0 k=1
$ ρ 2
= B(, α) dx x α−2 f (m−) (x)H [ln(γ /x)]−2 . (56)
0
Definition 3.10 Let m ∈ N, ρ ∈ (0, ∞), α ∈ R.
(i) Define
H0m ((0, ρ); x α dx; H) = f : (0, ρ) → H f (j ) ∈ ACloc ((0, ρ)), 0 j m − 1;
∞
there exists a Cauchy sequence {fn }∞
n=1 in C0 ((0, ρ); H), · m,α
$ ρ
(k) 2
such that lim α (k)
dx x fn (x) − f (x) H = 0, 0 k m
n→∞ 0
(57)
= f : (0, ρ) → H f (j ) ∈ ACloc ((0, ρ)), 0 j m − 1; there exists a
∞
Cauchy sequence {fn }∞
n=1 in C0 ((0, ρ); H), · m,α such that for any
$ b
0 < a < b < ρ, we have lim dx fn (x) − f (x)2H = 0 . (58)
n→∞ a
(ii) Define
.
H0m (0, ρ); x α dx; H = f : (0, ρ) → H f (j ) ∈ ACloc ((0, ρ)), 0 j m − 1;
∞
there exists a Cauchy sequence {fn }∞ n=1 in C0 ((0, ρ); H), ||| · |||m,α
$ ρ
2
such that lim dx x α fn(m) (x) − f (m) (x)H = 0 (59)
n→∞ 0
= f : (0, ρ) → H f (j ) ∈ ACloc ((0, ρ)), 0 j m − 1; there exists a
∞
Cauchy sequence {fn }∞
n=1 in C0 ((0, ρ); H), ||| · |||m,α such that for
$ b
any 0 < a < b < ρ, we have lim dx fn (x) − f (x)2H = 0 . (60)
n→∞ a
Remark 3.11
(i) An argument similar to that from Remark 3.6 shows that the conditions (57)
and (58) describe the same space.
A New Proof of the Power Weighted Birman–Hardy–Rellich Inequalities 591
(ii) Lemma A.7 together with (i) show that the conditions (59) and (60) describe
the same space. D
By Lemma A.7 one has
.
H0m ((0, ρ); x α dx; H) = H0m (0, ρ); x α dx; H , m ∈ N, α ∈ R, ρ ∈ (0, ∞).
(61)
lim fn(m−)
k
(x) = f (m−) (x) for a.e. x ∈ (0, ρ). (64)
k→∞
We now establish analogous results on the exterior domain (ρ, ∞), ρ ∈ (0, ∞).
592 F. Gesztesy et al.
.
For f, g ∈ H m ((ρ, ∞); x α dx; H) let
$ ∞
f, gm,α = dx x α (f (m) (x), g (m) (x))H . (67)
ρ
.
Then · , · m,α is an inner product on H m ((ρ, ∞); x α dx; H). In fact,
.m
H ((ρ, ∞); x α dx; H), · , · m,α is a Hilbert space.
Proof The proof is analogous to the argument from [7, Proposition B.1].
.m
Definition 3.14 For m ∈ N, ρ ∈ (0, ∞), α ∈ R let H0 ((ρ, ∞); x α dx; H) denote
.
the closure of C0∞ ((ρ, ∞); H) in the space H m ((ρ, ∞); x α dx; H), · , · m,α .
.
Lemma 3.15 Let f ∈ H0m ((ρ, ∞); x α dx; H). Then there is a sequence {fn }∞
n=1 ⊂
C0∞ ((ρ, ∞); H) such that
$ ∞ 2
lim dx x α fn(m) (x) − f (m) (x)H = 0 (68)
n→∞ ρ
lim f (k) (x) = f (k) (x) for a.e. x ∈ (ρ, ∞). (69)
n→∞ n
Proof The proof is analogous to the argument from [7, Corollary B.2].
Theorem 3.16 Let ρ ∈ (0, ∞). Then,
$ $
∞ 2 ∞ 2
dx f (m) (x)H A(, α) dx x α−2 f (m−) (x)H ,
ρ ρ (70)
.m
m ∈ N, ∈ {1, . . . , m}, α ∈ R, f ∈ H0 ((ρ, ∞); x α dx; H).
.
for all = 1, . . . , m, α ∈ R, m ∈ N, and f ∈ H0m ((ρ, ∞); x α dx; H).
A New Proof of the Power Weighted Birman–Hardy–Rellich Inequalities 593
Optimality of A(, α), and strictness of the inequalities in this section follow
from Theorem 2.2.
For the remainder of this appendix, H denotes a separable complex Hilbert space.
Definition A.1
(i) Let a, b ∈ R, a < b. A function f : [a, b] → H is said to be absolutely
continuous, denoted by f ∈ AC([a, b]; H), if for every ε > 0 there exists
δ > 0 such that for every finite collection {(aj , bj )}N
j =1 of disjoint subintervals
N
in [a, b] with j =1 (bj − aj ) < δ, one has
N
f (bj ) − f (aj )H < ε. (A.1)
j =1
such that
$ x
f (x) = f (x0 ) + dt g(t), x ∈ (c, d). (A.2)
x0
Lemma A.3 Let m ∈ N, α ∈ R, and assume that A(m, α) > 0. Then ||| · |||m,α ,
defined in (35), is a norm on C0∞ ((0, ∞); H).
Proof We only need to show that if f ∈ C0∞ ((0, ∞); H) and |||f |||m,α = 0, then
f = 0. Let {ϕk }∞ ∞
k=1 be an orthonormal basis of H. For f ∈ C0 ((0, ∞); H) and
k ∈ N we write
Suppose f ∈ C0∞ ((0, ∞); H) and |||f |||m,α = 0. Then, for all k ∈ N, fk ∈
C0∞ ((0, ∞)) and, applying Birman’s inequalities to fk , we have
$ $
∞
α (m)
2 ∞
(x)H
2
0= |||f |||2m,α = dx x f dx x α f (m) k (x)
0 0
$ ∞ $ ∞
(A.6)
(m) 2 2
= dx x α fk (x) A(m, α) dx x α−2m
fk (x) .
0 0
Hence, f = 0.
Lemma A.4 Let m ∈ N and α ∈ R. Suppose that A(m, α) > 0 and let {fn }∞ n=1
be a Cauchy sequence in (C0∞ ((0, ∞); H), ||| · |||m,α ). Then there exists a unique
.
f ∈ H m (0, ∞); x α dx; H , defined in Definition 3.2, such that, for all 0 m,
$ ∞ 2
lim dx x α−2fn(m−) (x) − f (m−) (x)H = 0. (A.8)
n→∞ 0
}∞
(m−)
hence {fn n=1 is a Cauchy sequence in L ((0, ∞); x
2 α−2dx; H). Thus there
(j )
g0 (x) = gj (x) for a.e. x > 0, 1 j m. (A.12)
(j )
lim fnk (x) = gj (x), x ∈ K. (A.13)
k→∞
6 7 1/2
max a −[α−2(m−1)], x −[α−2(m−1)] |x − a|1/2
$ ∞ 2 1/2
× dt t α−2(m−1) fnk (t) − g1 (t)H
0
−→ 0 . (A.15)
k→∞
Similarly, for 0 j m − 1,
(j )
gj (x) = lim fnk (x) (x ∈ K)
k→∞
$ x $ x
(j ) (j +1)
= lim fnk (a) + dt {fnk (t) − gj +1 (t)} + dt gj +1 (t) ,
k→∞ a a
(A.18)
and, by (A.11),
$ $
x 6 (j +1) 7 x (j +1)
dt fnk (t) − gj +1 (t) dt fnk (t) − gj +1 (t)H
a H a
$ x (j +1) 1/2
dt fnk (t) − gj +1 (t)2H |x − a|1/2
a
$ x (j +1) 2 1/2
= dt t −[α−2(m−j −1)] t α−2(m−j −1) fnk (t) − gj +1 (t)H |x − a|1/2
a
6 7 1/2
max a −[α−2(m−j −1)], x −[α−2(m−j −1)] |x − a|1/2
$ ∞ (j +1) 2 1/2
× dt t α−2(m−j −1) fnk (t) − gj +1 (t)H
0
−→ 0 . (A.19)
k→∞
A New Proof of the Power Weighted Birman–Hardy–Rellich Inequalities 597
.
Putting f = g0 , we have f ∈ H m (0, ∞); x α dx; H , by (A.20) and (A.21), (A.8)
follows from (A.11).
The uniqueness of f follows from (A.8) with = m.
Remark A.5 The condition (36) or (37) in Definition 3.2 (ii) in the context
.
of H0m (0, ∞); x α dx; H is necessary
to ensure that representation
of an
element in the completion of C0∞ ((0, ∞); H), ||| · |||m,α by a function in
.
H m (0, ∞); x α dx; H is unique.
To illustrate this point, consider the following example with m = 1, α = 0 and
H = C: Let g ∈ C0∞ ((0, ∞)) and put
.
but Fj ∈ H 1 (0, ∞); dx; C for all j ∈ N and Fj = Fk for j = k.
This kind of “non-uniqueness” phenomenon is due to ||| · |||m,α not being a norm
.
on H m (0, ∞); x α dx; H . D
Next, we turn to the case A(m, α) = 0.
Lemma A.6 Let m ∈ N, α ∈ R. With the notation established in Definition 3.5 (i),
let {fn }∞
n=1 be a Cauchy
sequence in C0∞ ((0, ∞); H), · m,α . Then there exists
a unique f ∈ H (0, ∞); x α dx; H such that
m
∞
Proof Since {fn }∞
(j ) ∞
n=1 is a Cauchy sequence in C0 ((0, ∞); H), · m,α , {fn }n=1
is a Cauchy sequence in L2 ((0, ∞); x α dx; H) for 0 j m, therefore there exists
598 F. Gesztesy et al.
(j )
g0 (x) = gj (x) for a.e. x ∈ (0, ∞), 1 j m. (A.26)
(j )
lim fnk (x) = gj (x), x ∈ K. (A.27)
k→∞
(j )
gj (x) = lim fnk (x) (x ∈ K)
k→∞
$ x $ x
(j ) (j +1)
= lim fnk (a) + dt {fnk (t) − gj +1 (t)} + dt gj +1 (t) ,
k→∞ a a
(A.28)
and, by (A.25),
$ $
x 6 (j +1) 7 x (j +1)
dt fnk (t) − gj +1 (t) dt fnk (t) − gj +1 (t)H
a H a
$ x (j +1) 2 1/2
dt fnk (t) − gj +1 (t)H |x − a|1/2
a
$ x (j +1) 2 1/2
= dt t −α t α fnk (t) − gj +1 (t)H |x − a|1/2
a
$
6 7 1/2 ∞ (j +1) 2 1/2
max a −α , x −α dt t α fnk (t) − gj +1 (t)H |x − a|1/2
a
−→ 0 . (A.29)
k→∞
Putting f = g0 , we have f ∈ H m ((0, ∞); x α dx; H), by (A.30) and (A.31), (A.24)
follows from (A.25).
Finally, uniqueness of f is a consequence of · m,α being a norm on the space
H m ((0, ∞); x α dx; H).
Lemma A.7 Let m ∈ N, α ∈ R, ρ ∈ (0, ∞). Then
$ ρ 2 1/2
|||f |||m,α = dx x α f (m) (x)H (A.32)
0
and
m $
ρ 2 1/2
f m,α = dx x α f (k) (x)H (A.33)
k=0 0
600 F. Gesztesy et al.
References
1 Introduction
N. Teofanov
Department of Mathematics and Informatics, University of Novi Sad, Novi Sad, Serbia
e-mail: [email protected]
J. Toft ()
Department of Mathematics, Linnæus University, Växjö, Sweden
e-mail: [email protected]
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 601
R. M. Aron et al. (eds.), Operator and Norm Inequalities and Related Topics,
Trends in Mathematics, https://doi.org/10.1007/978-3-031-02104-6_18
602 N. Teofanov and J. Toft
and were thereafter reconsidered by many authors in different contexts. Let us give
a brief, and unavoidably incomplete account on the related results.
In Sect. 3 we formulate in Theorems 3.5 and 3.7 bilinear versions of more
general multiplication and convolution results in [54, Section 3]. The contents of
Theorems 3.5 and 3.7 in the unweighted case for modulation spaces M p,q can be
summarized as follows.
Proposition 1.1 Let pj , qj ∈ (0, ∞], j = 0, 1, 2,
1 1 1 1 1
θ1 = max 1, , , and θ2 = max 1, , .
p0 q1 q2 p1 p2
Then
1 1 1 1 1 1
M p1 ,q1 · M p2 ,q2 ⊆ M p0 ,q0 , + = , + = θ1 + ,
p1 p2 p0 q1 q2 q0
1 1 1 1 1 1
M p1 ,q1 ∗ M p2 ,q2 ⊆ M p0 ,q0 , + = θ2 + , + = .
p1 p2 p0 q1 q2 q0
The general multiplication and convolution properties in Sect. 3 also overlap with
results by Bastianoni, Cordero and Nicola in [2], by Bastianoni and Teofanov in [1],
and by Guo et al. in [32].
The multiplication relation in Proposition 1.1 for pj , qj ≥ 1 was obtained
already in [17] by Feichtinger. It is also obvious that the convolution relation was
well-known since then (though a first formal proof of this relation seems to be given
first in [48]). In general, these convolution and multiplication properties follow the
rules
and
which goes back to [17] in the Banach space case and to [25] in the quasi-Banach
case. See also [19] and [42] for extensions of these relations to more general Banach
function spaces and quasi-Banach function spaces, respectively.
In Sect. 3 we basically review some results from [54]. To make this survey self-
contained we give the proof of Theorem 3.7 in unweighted case. In contrast to [32],
we do not deduce any sharpness for our results.
To show Proposition 1.1 in the quasi-Banach setting, apart from the usual use of
Hölder’s and Young’s inequalities, additional arguments are needed. In our situation
we discretize the situations in similar ways as in [2] by using Gabor analysis for
modulation spaces, and then apply some further arguments, valid in non-convex
An Excursion to Multiplications and Convolutions on Modulation Spaces 603
analysis. This approach is slightly different compared to what is used in [32] which
follows the discretization technique introduced in [55], and which has some traces
of Gabor analysis.
We refer to [54] for a detailed discussion on the uniqueness of multiplications
and convolutions in Proposition 1.1.
In Sect. 4 we apply the results from previous parts in the framework of the so
called Gabor product. It is introduced in [14] in order to derive a phase space
analogue to the usual convolution identity for the Fourier transform. The main
motivation is to use such kind of products in a phase-space formulation of certain
nonlinear equations. As noticed in [14], among other interesting characteristics
of phase-space representations, the initial value problem in phase-space may be
well-posed for more general initial distributions. This means that the phase-space
formulation could contain solutions other than the standard ones. We refer to [11–
13], where the phase-space extensions are explored in different contexts. Here we
illustrate this approach by considering the nonlinear cubic Schrödinger equation,
which appear for example in Bose-Einstein condensate theory [35]. We also refer to
[4, Chapter 7] for an overview of results related to well-posedness of the nonlinear
Schrödinger equations in the framework of modulation spaces, see also [3, 38, 39].
2 Preliminaries
In what follows we let F be the Fourier transform which takes the form
$
(F f )(ξ ) = fB(ξ ) ≡ (2π)− 2
d
f (x)e−ix,ξ dx
Rd
when f ∈ L1 (Rd ). Here · , · denotes the usual scalar product on Rd . The same
notation is used for the usual dual form between test functions and corresponding
(ultra-)distributions. We recall that map F extends uniquely to a homeomorphism
on the space of tempered distributions S (Rd ), to a unitary operator on L2 (Rd ) and
restricts to a homeomorphism on the Schwartz space of smooth rapidly decreasing
functions S (Rd ), cf. (29). We also observe with our choice of the Fourier transform,
604 N. Teofanov and J. Toft
the usual convolution identity for the Fourier transform takes the forms
F (f · g) = (2π)− 2 fB∗ B
g and F (f ∗ g) = (2π) 2 fB· B
d d
g (1)
when f, g ∈ S (Rd ).
In several situations it is convenient to use a localized version of the Fourier
transform, called the short-time Fourier transform, STFT for short. The short-time
Fourier transform of f ∈ S (Rd ) with respect to the fixed window function φ ∈
S (Rd ) is defined by
Here ( · , · )L2 denotes the unique continuous extension of the inner product on
L2 (Rd ) restricted to S (Rd ) into a continuous map from S (Rd ) × S (Rd ) to C.
We observe that using certain properties for tensor products of distributions,
(cf. [33, 52]). If in addition f ∈ Lp (Rd ) for some p ∈ [1, ∞], then
$
(Vφ f )(x, ξ ) = (2π)− 2
d
f (y)φ(y − x)e−iy,ξ dy. (2)
Rd
x $→ Vφ f (x, ξ )
ξ $→ Vφ f (x, ξ )
resembles on fB(ξ ).
An Excursion to Multiplications and Convolutions on Modulation Spaces 605
As for the ordinary Fourier transform, there are several mapping properties which
hold true for the short-time Fourier transform. As an elegant way to approach such
properties in the framework of distributions, we may follow ideas given in [24] by
Folland.
In fact, let T be the semi-conjugated tensor map
T (f, φ) = f ⊗ φ, (3)
$
(F2 F )(x, ξ ) = (2π)− 2 F (x, y)e−iy,ξ dy.
d
(5)
Rd
Then
when f, φ ∈ S (Rd ).
We observe that the mappings
For short-time Fourier transform, the Parseval identity is replaced by the so-
called Moyal identity, also known as the orthogonality relation given by
(Vφ f, Vψ g)L2 (R2d ) = (ψ, φ)L2 (Rd ) (f, g)L2 (Rd ) , (14)
1 1 1 1
+ = + = 1.
p p q q
An Excursion to Multiplications and Convolutions on Modulation Spaces 607
By Moyal’s identity (14) it follows that if φ ∈ S (Rd ) \ {0}, then the identity
operator on S (Rd ) is given by
Id = φ−2
L2
· Vφ∗ ◦ Vφ , (16)
provided suitable mapping properties of the (L2 -)adjoint Vφ∗ of Vφ can be estab-
lished. Obviously, Vφ∗ fullfils
when F ∈ S (R2d ). We may now use mapping properties like (11)–(12) to extend
the definition of Vφ∗ F when F and φ belong to various classes of function and
distribution spaces. For example, by (11), (10) and (12), it follows that the map
are continuous.
If φ ∈ S (Rd ) satisfies φL2 = 1, then (16) shows that Vφ∗ ◦ Vφ is the identity
operator on S (Rd ). If we swap the order of this composition we get certain types
608 N. Teofanov and J. Toft
of projections. In fact, for any φ ∈ S (Rd ) \ {0}, let Pφ be the operator given by
Pφ ≡ φ−2
L2
· Vφ ◦ Vφ∗ . (20)
Pφ2 = φ−2
L2
· Vφ ◦ φ−2 ∗ ∗ −2 ∗
2 · Vφ ◦ Vφ ◦ Vφ = φL2 · Vφ ◦ Vφ = Pφ .
O L PQ R
The identity operator
Hence,
Pφ F = Vφ f,
Pφ F = Vφ ◦ φ−2
L2
· Vφ∗ ◦ Vφ f = Vφ f,
which shows that any element in Vφ (S (Rd )) equals an element in Pφ (S (R2d )),
i.e. Pφ (S (R2d )) = Vφ (S (Rd )). This gives the last identity in (22). In the same
way, the first two identities are obtained.
Remark 2.2 Let F ∈ S (R2d ). Then it follows from the last identity in (22) that
F = Vφ f for some f ∈ S (Rd ), if and only if
F = Pφ F. (23)
f = (φ−2
L2
) · Vφ∗ F. (24)
An Excursion to Multiplications and Convolutions on Modulation Spaces 609
Pφ F = φ−2
L2
· Vφ φ ∗V F, F ∈ S (R2d ), (25)
R2d
(26)
(F ∗V G) ∗V H = F ∗V (G ∗V H ), (27)
Before defining the Gelfand-Shilov spaces, we recall that the Schwartz space
S (Rd ) consists of all (complex-valued) smooth functions f ∈ C ∞ (Rd ) such that
sup |x β ∂ α f (x)| ≤ Cα,β , (29)
x∈Rd
for some constants Cα,β > 0, which only depend on the multi-indices α, β ∈ Nd .
The Schwartz space possess several convenient properties, and is heavily used in
mathematics, science and technology. For example, the Schwartz space is invariant
610 N. Teofanov and J. Toft
under Fourier transformation. By duality the same holds true for its (L2 -)dual
S (Rd ), the set of tempered distributions on Rd .
On the other hand, we observe that there are no conditions on the growths of
the constants Cα,β with respect to α, β ∈ Nd . This implies that in the context of
the spaces S (Rd ) and S (Rd ), it is almost impossible to investigate important
properties like analyticity or related regularity properties which are stronger than
pure smoothness. For investigating such stronger regularity properties, we need to
modify S (Rd ) and the estimate (29) by imposing suitable growth conditions on the
constants Cα,β . This leads to the definition of Gelfand-Shilov spaces, [26, 40].
We only discuss Fourier invariant Gelfand-Shilov spaces and their properties.
Let 0 < s ∈ R be fixed. We have two different types of Gelfand-Shilov spaces. The
Gelfand-Shilov space Ss (Rd ) of Roumieu type with parameter s > 0 consists of all
f ∈ C ∞ (Rd ) such that
sup |x β ∂ α f (x)| ≤ Ch|α+β| (α!β!)s , (30)
x∈Rd
for some constants C, h > 0. In the same way, the Gelfand-Shilov space s (Rd ) of
Beurling type with parameter s > 0 consists of all f ∈ C ∞ (Rd ) such that for every
h > 0, there is a constant C = Ch > 0 such that (30) holds. Hence, in comparison
with the definition of Schwartz functions, we have limited ourself to constants Cα,β
in (29) which are not allowed to grow faster than those of the form
Ch|α+β| (α!β!)s
One has that S1 (Rd ) consists of real analytic functions, and that 1 (Rd ) consists
of smooth functions on Rd which are extendable to entire functions on Cd . The
topologies of Ss (Rd ) and s (Rd ) are defined by the semi-norms
|x β ∂ α f (x)|
f Ss,h ≡ sup . (31)
h|α+β| (α!β!)s
Let Ss,h (Rd ) be the Banach space which consists of all f ∈ C ∞ (Rd ) such that
(Rd ) be the (L2 -)dual of S (Rd ). If s ≥ 1 ,
f Ss,h in (31) is finite, and let Ss,h s,h 2
then the Gelfand-Shilov distribution space Ss (Rd ) of Roumieu type is the projective
(Rd ) with respect to h > 0. If instead s > 1 , then the Gelfand-Shilov
limit of Ss,h 2
(Rd ) with
distribution space s (Rd ) of Beurling type is the inductive limit of Ss,h
respect to h > 0. Consequently, for admissible s we have
D C
Ss (Rd ) =
Ss,h (Rd ) and s (Rd ) =
Ss,h (Rd ).
h>0 h>0
It can be proved that Ss (Rd ) and s (Rd ) are the (strong) duals to Ss (Rd ) and
s (Rd ), respectively.
We have the following embeddings and density properties for Gelfand-Shilov
and Schwartz spaces
for some r > 0 (for every r > 0). Here g1 g2 means that g1 (θ ) ≤ c · g2 (θ ) holds
uniformly for all θ in the intersection of the domains of g1 and g2 and for some
constant c > 0, and we write g1 E g2 when g1 g2 g1 .
The analysis in [8, 15] can also be applied on the Schwartz space, from which it
follows that an element f ∈ S (Rd ) belongs to S (Rd ), if and only if
Remark 2.3 Several properties in Sects. 2.1–2.3 in the background of S (Rd ) and
S (Rd ) also hold for the Gelfand-Shilov spaces and their distribution spaces. Let
s ≥ 12 . By similar arguments which lead to Proposition 2.1 and (13), it follows that
and
It follows that (14) makes sense after each S in (15) are replaced by Ss . Let
φ ∈ Ss (Rd ) \ {0} be fixed. Then by similar arguments which lead to (19) give that
the mappings
are continuous. For Pφ in (20) we have that (21) still holds true and that (22) can be
completed with
Pφ (Ss (R2d )) = Vφ (Ss (Rd )) and Pφ (Ss (R2d )) = Vφ (Ss (Rd )). (38)
We also have that the twisted convolution in (26) is continuous from Ss (R2d ) ×
Ss (R2d ) to Ss (R2d ) and uniquely extendable to a continuous map Ss (R2d ) ×
Ss (R2d ) or Ss (R2d ) × Ss (R2d ) to Ss (R2d ), and that the formulae (25)–(28) still
hold true after each S is replaced by Ss in the attached assumptions.
If instead s > 12 , then similar facts hold true with s in place of Ss above, at
each occurrence.
Remark 2.4 In similar ways as characterizing Gelfand-Shilov spaces in terms of
Fourier estimates (see (33)), we may also use the short-time Fourier transform to
perform similar characterizations. Moreover, the short-time Fourier transform can
in addition be used to characterize spaces of Gelfand-Shilov distributions.
In fact, let φ ∈ Ss (Rd ) \ {0} (φ ∈ s (Rd ) \ {0}) be fixed and let f be a Gelfand-
Shilov distribution on Rd . Then the following is true:
1. f ∈ Ss (Rd ) (f ∈ s (Rd )), if and only if
1 1
e−r(|x|
s +|ξ | s )
|Vφ f (x, ξ )| (39)
If ω and v are weights on Rd such that (41) holds, then ω is also called v-moderate.
We note that (41) implies that ω fulfills the estimates
for some constant C > 0, then for v0 = C 1/2 v, one has that v0 ∈ PE (Rd ) is
submultiplicative and v E v0 (see e.g. [17, 19, 28]).
We also recall from [29] that if v is positive and locally bounded and satis-
fies (44), then v(x) ≤ C0 er0 |x| for some positive constants C0 and r0 . In fact, if
x ∈ Rd ,
for some r > 0. In particular, (42) shows that for any ω0 ∈ PE (Rd ), there is a
constant r > 0 such that
If (41) holds, then there is a smallest positive even function v0 such that (41)
holds with C = 1. We remark that this v0 is given by
ω(x + y) ω(−x + y)
v0 (x) = sup , ,
y∈Rd ω(y) ω(y)
1
lim v(nx) n = 1 (47)
n→∞
An Excursion to Multiplications and Convolutions on Modulation Spaces 615
log(v(nx))
lim = 0, (47)
n→∞ n
and is usually called the GRS condition, or Gelfand-Raikov-Shilov condition.
A more restrictive condition on v compared to (47) is given by the Beurling-
Domar condition
∞
log(v(nx))
< ∞. (48)
n2
n=1
In fact, it is clear that the ordering in (49) holds. On the other hand, if r > 0 and
θ ∈ (0, 1), then due to
θ
er|x| ∈ PBD (Rd ) \ P(Rd ),
F Lp,q (R2d ) ≡ GF,ω,p Lq (Rd ) , where GF,ω,p (ξ ) = F ( · , ξ )ω( · , ξ )Lp (Rd )
(ω)
and
F Lp,q (R2d ) ≡ HF,ω,q Lp (Rd ) , where HF,ω,q (x) = F (x, · )ω(x, · )Lq (Rd ) ,
∗,(ω)
p,q
when F is (complex-valued) measurable function on R2d . Then L(ω) (R2d )
p,q
(L∗,(ω) (R2d )) consists of all measurable functions F such that F Lp,q < ∞
(ω)
(F Lp,q < ∞).
∗,(ω)
In similar ways, let 1 , 2 be discrete sets, ω be a positive function on
1 × 2 and 0 (1 × 2 ) be the set of all formal (complex-valued) sequences
c = {c(j, k)}j ∈1 ,k∈2 . Then the discrete Lebesgue spaces, i.e. the Lebesgue
sequence spaces
p,q p,q
(ω) (1 × 2 ) and ∗,(ω) (1 × 2 )
cp,q (1 ×2 ) ≡ Gc,ω,p q (2 ) , where Gc,ω,p (k) = F ( · , k)ω( · , k)p (1 )
(ω)
and
cp,q ≡ Hc,ω,q p (1 ) , where Hc,ω,q (j ) = c(j, · )ω(j, · )q (2 ) ,
∗,(ω) (1 ×2 )
Next we discuss extended Hölder and Young relations for multiplications and
convolutions on discrete Lebesgue spaces. The Hölder and Young conditions on
Lebesgue exponent are then
1 1 1
≤ + , (51)
q0 q1 q2
An Excursion to Multiplications and Convolutions on Modulation Spaces 617
respectively
1 1 1 1 1
≤ + − max 1, , . (52)
p0 p1 p2 p1 p2
Notice that, when p1 , p2 ∈ (0, 1), then (52) becomes p0 ≥ max{p1 , p2 }, while
for p1 , p2 ≥ 1 it reduces to the common Young condition
1 1 1
1+ ≤ + .
p0 p1 p2
ω0 (j ) ≤ ω1 (j )ω2 (j ), j ∈ $, (53)
respectively
$ = { n1 e1 + · · · + nd ed ; (n1 , . . . , nd ) ∈ Zd },
q
a1 · a2 q0 ≤ a1 q1 a2 q2 , aj ∈ (ωj j ) ($), j = 1, 2; (55)
(ω0 ) (ω1 ) (ω2 )
2. if (54) holds, then the map (a1 , a2 ) $→ a1 ∗ a2 from 0 ($) × 0 ($) to 0 ($)
p p p
extends uniquely to a continuous map from (ω11 ) ($) × (ω22 ) ($) to (ω00 ) ($), and
p
a1 ∗ a2 p0 ≤ a1 p1 a2 p2 , aj ∈ (ωj j ) ($), j = 1, 2. (56)
(ω0 ) (ω1 ) (ω2 )
The assertion 1. in Proposition 2.5 is the standard Hölder’s inequality for discrete
Lebesgue spaces. The assertion 2. in that proposition is the usual Young’s inequality
for Lebesgue spaces on lattices in the case when p0 , p1 , p2 ∈ [1, ∞]. A proof of
Proposition 2.5 is given in Appendix A in [54].
618 N. Teofanov and J. Toft
In this section we introduce modulation spaces, and recall their basic properties,
in particular in the context of Gelfand-Shilov spaces. Notice that we permit
the Lebesgue exponents to belong to the full interval (0, ∞] instead of the
most common choice [1, ∞], and general moderate weights which may have a
(sub)exponential growth. Here we also recall some facts on Gabor expansions for
modulation spaces.
Then we deduce multiplication and convolution estimates on modulation spaces.
There are several approaches to multiplication and convolution in the case when the
involved Lebesgue exponents belong to [1, ∞] (see [9, 17, 19, 32, 43, 48]). Here we
consider the case when these exponents belong to (0, ∞) (see also [1, 2, 25, 41, 42,
50]). In addition, and in order to keep the survey style of our exposition, we focus
on the bilinear case, and refer to [54] for extension of these results to multi-linear
products.
p,q
is finite. The topology of M(ω) (Rd ) is defined by the (quasi-)norm · M p,q ;
(ω)
p,q
2. The modulation space (of Wiener amalgam type) W(ω) (Rd ) consists of all f ∈
1 (Rd ) such that
p,q
is finite. The topology of W(ω) (Rd ) is defined by the (quasi-)norm · W p,q .
(ω)
p,q p,q
For convenience we set M p,q = and W p,q =
M(ω) when the weight ω is
W(ω)
trivial, i.e. when ω(x, ξ ) = 1 for every x, ξ ∈ Rd . We also set
p p,p p,p
M(ω) ≡ M(ω) ( = W(ω) ) and M p ≡ M p,p ( = W p,p ).
An Excursion to Multiplications and Convolutions on Modulation Spaces 619
and
D p,q
D p,q
S (Rd ) = M(vr,∞ ) (Rd ) = W(vr,∞ ) (Rd ), (63)
r>0 r>0
C C
S (Rd ) =
p,q p,q
M(1/vr,∞ ) (Rd ) = W(1/vr,∞ ) (Rd ), (64)
r>0 r>0
D D
Ss (Rd ) =
p,q p,q
M(1/vr,s ) (Rd ) = W(1/vr,s ) (Rd ) (65)
r>0 r>0
and
C C
s (Rd ) =
p,q p,q
M(1/vr,s ) (Rd ) = W(1/vr,s ) (Rd ). (66)
r>0 r>0
The topologies of the spaces on the left-hand sides of (61)–(66) are obtained by
replacing each intersection by projective limit with respect to r > 0 and each union
with inductive limit with respect to r > 0.
The relations (61)–(66) are essentially special cases of [49, Theorem 3.9], see
also [31, 45, 46]. In order to be self-contained we here give a proof of (62).
Proof of (62) Since
∞ p,q p,q ∞
M(v 2r,s )
(Rd ) ⊆ M(vr,s ) (Rd ), W(vr,s ) (Rd ) ⊆ M(v r,s )
(Rd ),
it suffices to prove the result for p = q = ∞. Let φ ∈ 1 (Rd ) \ {0} be fixed. First
suppose that
∞ ∞
f ∈ M(v r,s )
(Rd ) = W(v r,s )
(Rd ).
Then it follows from the definition of modulation space norm that (39) holds for
some r > 0. By Remark 2.4 it follows that f ∈ Ss (Rd ), and we have proved
C
∞
M(v r,s )
(Rd ) ⊆ Ss (Rd ). (67)
r>0
Suppose instead that f ∈ Ss (Rd ). Then (39) holds for some r > 0, giving that
∞ (Rd ). Hence (67) holds with reversed inclusion, and the result follows.
f ∈ M(v r,s )
1,1 d
Example 3.3 Let p = q = 1 and ω = 1. Then M(ω) (R ) = M 1 (Rd ) is the
Feichtinger algebra, probably the most prominent example of a modulation space.
We refer to a recent survey [34] for a detailed account on M 1 (Rd ), and to [14,
Lemma 11] for a list of its basic properties.
An Excursion to Multiplications and Convolutions on Modulation Spaces 621
2,2 d
Familiar examples arise when p = q = 2 and ω = 1. Then M(ω) (R ) =
M 2 (Rd ) = L2 (Rd ), and
2,2
M(ωs)
(Rd ) = H s (Rd ), s ∈ R,
where ωs (ξ ) = ξ s , and H s (Rd ) is the Sobolev space (also known as the Bessel
potential space) of distributions f ∈ S (Rd ) such that
$
f 2H s := ξ 2s |fB(ξ )|2 dξ < ∞,
Rd
2,2
cf. [28, Proposition 11.3.1]. Furthermore, if vs (x, ξ ) = (x, ξ )s , then M(v s)
(Rd ) =
Qs (R ), s ∈ R, [7, Lemma 2.3]. Here Qs denotes the Shubin-Sobolev space, [44].
d
1 1
s +|ξ | s )
ω(x + y, ξ + η) ω(x, ξ )er(|x| , (68)
for some s > 1 and r > 0. We observe that this implies that
in vew of (42), (61) and (66). For more general approaches we refer to [19, 27, 28,
42, 50].
Since s > 1, it follows from Sections 1.3 and 1.4 in [33] that there are φ, ψ ∈
s (Rd ) with values in [0, 1] such that
2 3 3 3d 2 1 1 3d
supp φ ⊆ − , , φ(x) = 1 when x ∈ − , (70)
4 4 4 4
2 3 3 3d
supp ψ ⊆ [−1, 1]d , ψ(x) = 1 when x ∈ − , (71)
4 4
and
φ( · − j ) = 1. (72)
j ∈Zd
622 N. Teofanov and J. Toft
Let f ∈ s (Rd ). Then x $→ f (x)φ(x − j ) belongs to s (Rd ) and is supported in
j + [− 34 , 34 ]d . Hence, by periodization it follows from Fourier analysis that
f (x)φ(x − j ) = c(j, ι)eix,ι , x ∈ j + [−1, 1]d , (73)
ι∈πZd
where
d
π
c(j, ι) = 2−d (f, φ( · − j )ei · ,ι ) =
2
Vφ f (j, ι), j ∈ Zd , ι ∈ πZd .
2
Since ψ = 1 on the support of φ, (73) gives
π d
(73)
2
f (x)φ(x − j ) = Vφ f (j, ι)ψ(x − j )eix,ι , x ∈ Rd ,
2
ι∈πZd
π d
2
f (x) = Vφ f (j, ι)ψ(x − j )eix,ι , x ∈ Rd , (74)
2
(j,ι)∈$
where
$ = Zd × (πZd ), (75)
which is the Gabor expansion of f with respect to the Gabor pair (φ, ψ) and lattice
$, i.e. with respect to the Gabor atom φ and the dual Gabor atom ψ. Here the series
converges in s (Rd ). By duality and the fact that compactly supported elements in
s (Rd ) are dense in s (Rd ) we also have
π d
2
f (x) = Vψ f (j, ι)φ(x − j )eix,ι , x ∈ Rd , (76)
2
(j,ι)∈$
π d
Vφ f (j, ι)T (ψ( · − j )ei · ,ι )(x)
2
(Tf )(x) =
2
(j,ι)∈$
An Excursion to Multiplications and Convolutions on Modulation Spaces 623
and
π d
T (ψ( · − j )ei · ,ι )(x) = (Vφ (T (ψ( · − j )ei · ,ι )))(k, κ)ψ(x − k)eix,κ .
2
2
(k,κ)∈$
π d
2
(Tf )(x) = (A · Vφ f )(j, ι)ψ(x − j )eix,ι , (77)
2
(j,ι)∈$
d
π
(T (ψ( · − j )ei · ,ι ), φ( · − k)ei · ,κ)L2 (Rd )
2
a(j , k) =
2
when j = (j, ι) and k = (k, κ). (78)
By the Gabor analysis for modulation spaces we get the following restatement of
[54, Proposition 1.8]. We refer to [17, 19–21, 25, 27, 28, 50] for details.
Proposition 3.4 Let s > 1, p, q ∈ (0, ∞], ω ∈ PE (R2d ) be such that (68) holds
for some r > 0, φ, ψ ∈ s (Rd ) with values in [0, 1] be such that (70), (71) and (72)
hold true, and let f ∈ s (Rd ). Then the following is true:
p,q
1. f ∈ M(ω) (Rd ), if and only if Vφ f p,q (Zd ×πZd ) < ∞;
(ω)
p,q
2. f ∈ M(ω) (Rd ), if and only if Vψ f p,q (Zd ×πZd ) < ∞;
(ω)
3. the quasi-norms
φ0 = (2π)− 2 φ1 φ2
d
(79)
624 N. Teofanov and J. Toft
where
Fj = Vφj fj , j = 0, 1, 2. (81)
$
f1 (y)φ1 (y − x)f2 (y)φ2 (y − x)e−iy,ξ dy = (2π) 2 (Vφ1 φ2 (f1 f2 ))(x, ξ ).
d
=
where Fj are given by (81), and that we may extract f0 = f1 ∗ f2 from (82).
Next we discuss convolutions and multiplications for modulation spaces, and
start with the following convolution result for modulation spaces. For multiplica-
tions of elements in modulation spaces we need to swap the conditions for the
involved Lebesgue exponents compared to (51) and (52). That is, these conditions
become
1 1 1 1 1 1 1 1 1
≤ + , ≤ + − max 1, , , , (85)
p0 p1 p2 q0 q1 q2 p0 q1 q2
or
1 1 1 1 1 1 1 1
≤ + , ≤ + − max 1, , . (86)
p0 p1 p2 q0 q1 q2 q1 q2
An Excursion to Multiplications and Convolutions on Modulation Spaces 625
respectively
p ,q
f1 f2 M p0 ,q0 f1 M p1 ,q1 f2 M p2 ,q2 , fj ∈ M(ωjj ) j (Rd ), j = 1, 2. (89)
(ω0 ) (ω1 ) (ω2 )
p ,q
f1 f2 W p0 ,q0 f1 W p1 ,q1 f2 W p2 ,q2 , fj ∈ W(ωjj ) j (Rd ), j = 1, 2. (90)
(ω0 ) (ω1 ) (ω2 )
The corresponding results for convolutions are the following. Here the conditions
on the involved Lebesgue exponents are swapped as
1 1 1 1 1 1 1 1 1
≤ + − max 1, , , , ≤ + (91)
p0 p1 p2 q0 p1 p2 q0 q1 q2
or
1 1 1 1 1 1 1 1
≤ + − max 1, , , ≤ + (92)
p0 p1 p2 p1 p2 q0 q1 q2
p ,q
f1 ∗ f2 M p0 ,q0 f1 M p1 ,q1 f2 M p2 ,q2 , fj ∈ M(ωjj ) j (Rd ), j = 1, 2. (93)
(ω0 ) (ω1 ) (ω2 )
626 N. Teofanov and J. Toft
p ,q
f1 ∗ f2 W p0 ,q0 f1 W p1 ,q1 f2 W p2 ,q2 , fj ∈ W(ωjj ) j (Rd ), j = 1, 2. (94)
(ω0 ) (ω1 ) (ω2 )
in view of (84).
We have
and
Then
and
in view of (A.5) and Proposition A.1 in Appendix A (see also [25, Theorem 3.3])).
By (84) we have
G(x, λ) ≤ a1 (k1 , λ)a2 (k2 , λ)(χk1 +Q ∗ χk2 +Q )(x)
k1 ,k2 ∈Zd
(97)
≤ a1 (k1 , λ)a2 (k2 , λ)χk1 +k2 +Qd,2 (x).
k1 ,k2 ∈Zd
We observe that
where
and
If we apply the p0 quasi-norm on (98) with respect to the l variable, then
Proposition 2.5 (2) and the fact that I is finite set give
b( · , λ)p0 ≤ (a1 ( · , λ) ∗ a2 ( · , λ))( · − 2e0 + m) p0
m∈I
≤ (a1 ( · , λ) ∗ a2 ( · , λ))( · − 2e0 + m)p0
m∈I
E a1 ( · , λ) ∗ a2 ( · , λ)p0
By applying the q0 quasi-norm and using Proposition 2.5 (1) we now get
A combination of this estimate with (95) and (96) gives that f1 ∗ f2 is well-defined
and that (93) holds.
The uniqueness now follows from that (93) holds for f1 , f2 ∈ S (Rd ), and that
S (Rd ) is dense in M p,q (Rd ) when p, q < ∞.
In this section we give an illustration how the multiplication properties for modula-
tion spaces can be used when treating certain nonlinear problems. We consider the
Gabor product which is connected to such multiplication properties. It is introduced
in [14] in order to derive a phase space analogue to the usual convolution identity
for the Fourier transform. We will prove a formula related to (80), and then use
results from previous section to extend the Gabor product initially defined on
M 1 (R2d )×M 1 (R2d ) to some other spaces. Finally, we show how the Gabor product
gives rise to a phase-space formulation of the qubic Schrödinger equation.
Definition 4.1 Let φ ∈ M 1 (Rd ) \ {0}, and let F1 , F2 ∈ M 1 (R2d ). Then the Gabor
product φ is given by
F1 φ F2 (x, ξ )
$$$
= (2π) −d −ix,ξ
e B
φ (ζ − ξ )eix,ζ F1 (y, η)F2 (y, ζ − η) dydηdζ. (99)
R3d
An Excursion to Multiplications and Convolutions on Modulation Spaces 629
In the proof of [14, Lemma 13] it is justified that the Gabor product in (99) is
well-defined, and that
is a continuous map.
The Gabor product is particularly well-suited in the context of the STFT.
Theorem 4.2 Let φ, φ1 , φ2 ∈ M 1 (Rd )\ {0}. Then
where
$
G(y, ζ ) = (Vφ1 f1 )(y, η)(Vφ2 f2 )(y, ζ − η) dη.
Rd
where
$
H (z) = φ2 (z − y)φ1 (z − y) dy = (φ2 , φ1 )L2 .
Rd
Hence, by evaluating the integral with respect to ζ , and using Fourier’s inversion
formula, we get
This can be considered as a phase space form of the Young convolution inequality.
Next we discuss continuity of the Gabor product on certain spaces involving
superpositions of short-time Fourier transforms. In the end we deduce properties
p,q
similar to [14, Theorem 29]. Instead of modulation spaces of the form M(ω) (Rd ),
p, q ∈ [1, ∞), ω ∈ PE (R2d ), here we consider modulation spaces of Wiener
p,q
amalgam types W(ω) (Rd ), and allow the “quasi-Banach” choice for Lebesgue
parameters, i.e. p and q are allowed to be smaller than one.
An Excursion to Multiplications and Convolutions on Modulation Spaces 631
(N),p,q
be an orthonormal basis of L2 (Rd ). Then let VG ,ω (R2d ) be the closure of
N G
(N)
VG (R ) = 2d
Vφn fn ; fn ∈ 1 (R ) d
(105)
n=1
p,q
with respect to the L∗,(ω) (R2d ) norm. In particular, if N = 1, φ = φ1 and p, q ≥ 1,
then this reduces to the closure
p,q p,q
Pφ (L∗,(ω) (R2d )) = Vφ (W(ω) (Rd ))
of
N
F = Vφn fn . (106)
n=1
F1 φ F2 Lp0 ,q0 F1 Lp1 ,q1 F2 Lp2 ,q2 , (107)
∗,(ω0 ) ∗,(ω1 ) ∗,(ω2 )
(N),p ,q
for all Fj ∈ VG ,ω j j (R2d ), j = 1, 2.
j
In particular, if Fj = Vφ fj , j = 1, 2, and φL2 = 1, then (107) reduces to
Vφ f1 φ Vφ f2 Lp0 ,q0 = f1 f2 W p0 ,q0 f1 W p1 ,q1 f2 W p2 ,q2 . (108)
∗,(ω0 ) (ω0 ) (ω1 ) (ω2 )
632 N. Teofanov and J. Toft
∂ψ
i + ψ + λ|ψ|2 ψ = 0, (109)
∂t
subject to the initial condition:
ψ(x, 0) = ϕ(x).
d
∂2
= .
j =1
∂xj2
∂F
d
2
i − 1 φF
ξj + Dxj F + λF φF = 0. (110)
∂t
j =1
1 is given by
and F
1(x, ξ ) = F (x, −ξ ).
F (111)
space formulation “contains” the solutions of the standard NLSE, but it is richer,
as it admits other solutions. We refer to [11–13], where phase-space extensions are
explored in several different contexts.
Let us conclude by noticing that (110) contains the triple product. Thus, its
qualitative analysis calls for a multilinear extension of Theorems 3.6 and 4.3. Then
the conditions (86) and (87) become more involved, see [54]. Such analysis demands
a more technical tools and arguments and goes beyond the scope of this survey
article.
when
a0 (j ) ≡ f · ω0 Lr (j +Qd ) , j ∈ Zd ,
when
consists of all measurable f ∈ Lrloc (Rd ) such that F Wr (ω0 ,p ) is finite, and the
Wiener amalgam spaces
p,q p,q
Wr (ω, p,q ) = Wr (ω, p,q (Z2d )) and Wr (ω, ∗ ) = Wr (ω, ∗ (Z2d ))
consist of all measurable F ∈ Lrloc (R2d ) such that F Wr (ω,p,q ) respectively
F Wr (ω,p,q
∗ )
are finite. We observe that Wr (ω0 , p ) is often denoted by
r p
W (L , (ω) ) in the literature (see e. e. [17, 19, 25, 41]).
634 N. Teofanov and J. Toft
Obviously, Wr (ω0 , p ) and Wr (ω, p,q ) increase with p, q, decrease with r, and
W(ω, p,q ) 3→ L(ω) (R2d ) ∩ 1 (R2d ) 3→ L(ω) (R2d ) 3→ Wr (ω, p,q )
p,q p,q
(A.3)
and
with
and
We refer to [54, Proposition 3.6] for the proof of Proposition A.1 and to [19, 21,
28, 41, 42, 54] for some facts about the operators Pφ and Vφ∗ ,
For p, q ≥ 1, i.e. the case when all spaces are Banach spaces, proofs of
Proposition A.1 can be found in e.g. [28] as well as in abstract forms in [19]. In the
general case when p, q > 0, we refer to [25, 42], since proofs of Proposition A.1
are essentially given there.
References
17. H.G. Feichtinger, Modulation spaces on locally compact abelian groups. Technical report,
University of Vienna, Vienna, 1983; also in ed. by M. Krishna, R. Radha, S. Thangavelu.
Wavelets and Their Applications (Allied Publishers Private Limited, NewDehli, 2003),
pp. 99–140
18. H.G. Feichtinger, Modulation spaces: looking back and ahead. Sampl. Theory Signal Image
Process. 5, 109–140 (2006)
19. H.G. Feichtinger, K.H. Gröchenig, Banach spaces related to integrable group representations
and their atomic decompositions, I. J. Funct. Anal. 86, 307–340 (1989)
20. H.G. Feichtinger, K.H. Gröchenig, Banach spaces related to integrable group representations
and their atomic decompositions, II. Monatsh. Math. 108, 129–148 (1989)
21. H.G. Feichtinger, K.H. Gröchenig, Gabor frames and time-frequency analysis of distributions.
J. Funct. Anal. 146, 464–495 (1997)
22. H.G. Feichtinger, G. Narimani, Fourier multipliers of classical modulation spaces. Appl.
Comput. Harmon. Anal. 21, 349–359 (2006)
23. C. Fernandez, A. Galbis, J. Toft, Characterizations of GRS-weights, and consequences in
time-frequency analysis. J. Pseudo-Differ. Oper. Appl. 6, 383–390 (2015)
24. G.B. Folland, Harmonic Analysis in Phase Space (Princeton University Press, Princeton,
1989)
p,q
25. Y.V. Galperin, S. Samarah, Time-frequency analysis on modulation spaces Mm , 0 < p, q ≤
∞. Appl. Comput. Harmon. Anal. 16, 1–18 (2004)
26. I.M. Gelfand, G.E. Shilov, Generalized Functions, II–III (Academic Press, NewYork, 1968).
Reprinted by AMS (2016)
27. K.H. Gröchenig, Describing functions: atomic decompositions versus frames. Monatsh.
Math. 112, 1–42 (1991)
28. K. Gröchenig, Foundations of Time-Frequency Analysis (Birkhäuser, Boston, 2001)
29. K. Gröchenig, Composition and spectral invariance of pseudodifferential operators on
modulation spaces. J. Anal. Math. 98, 65–82 (2006)
30. K. Gröchenig, Weight functions in time-frequency analysis, in ed. by L. Rodino, M.W. Wong.
Pseudodifferential Operators: Partial Differential Equations and Time-Frequency Analysis.
Fields Institute Communications, vol. 52 (American Mathematical Society, Providence,
2007), pp. 343–366
31. K. Gröchenig, G. Zimmermann, Spaces of test functions via the STFT. J. Funct. Spaces Appl.
2, 25–53 (2004)
32. W. Guo, J. Chen, D. Fan, G. Zhao, Characterizations of some properties on weighted
modulation and wiener amalgam spaces. Michigan Math. J. 68, 451–482 (2019)
33. L. Hörmander, The Analysis of Linear Partial Differential Operators, vol I–III. (Springer,
Berlin, 1983, 1985)
34. M.S. Jakobsen, On a (no longer) new segal algebra: a review of the feichtinger algebra. J.
Fourier Anal. Appl. 24, 1579–1660 (2018)
35. P.G. Kevrekidis, D.J. Frantzeskakis, R. Carretero-Gonzalez (eds.), Emergent Nonlinear
Phenomena in Bose-Einstein Condensation (Springer, Berlin, 2008)
36. E.H. Lieb, Integral bounds for radar ambiguity functions and Wigner distributions. J. Math.
Phys. 31, 594–599 (1990)
37. E.H. Lieb, J.P. Solovej, Quantum coherent operators: a generalization of coherent states. Lett.
Math. Phys. 22, 145–154 (1991)
38. T. Oh, Y. Wang, Global well-posedness of the one-dimensional cubic nonlinear Schrödinger
equation in almost critical spaces. J. Diff. Eq. 269, 612–640 (2020)
39. T. Oh, Y. Wang, On global well-posedness of the modified KdV equation in modulation
spaces. Discrete Continuous Dyn. Syst. 41, 2971–2992 (2021)
40. S. Pilipović, Tempered ultradistributions. Boll. U.M.I. 7, 235–251 (1988)
41. H. Rauhut, Wiener amalgam spaces with respect to quasi-Banach spaces. Colloq. Math. 109,
345–362 (2007)
42. H. Rauhut, Coorbit space theory for quasi-Banach spaces. Stud. Math. 180, 237–253 (2007)
An Excursion to Multiplications and Convolutions on Modulation Spaces 637
1 Introduction
G.H. Hardy and J.E. Littlewood have their names associated to dozens of inequal-
ities and when we mention the Hardy–Littlewood inequalities it is natural that
researchers of different fields conceive different results. In this work the Hardy–
Littlewood inequalities are the main theorems of the paper [32] and their m-linear
generalizations. In some sense the starting point of this cycle of ideas rests on the
works of Orlicz, Littlewood, Bohnenblust and Hille in the beginning of the 1930’s
(see [17, 33]). These results show how the sums of the coefficients are dominated
by the norms of m-linear forms in c0 spaces and, in 1934, Hardy and Littlewood
extended these inequalities to bilinear forms in p spaces. We recall that an operator
D. Núñez-Alarcón
Departamento de Matemáticas, Universidad Nacional de Colombia, Bogotá, Colombia
e-mail: [email protected]
D. M. Pellegrino ()
Departamento de Matemática, Universidade Federal da Paraíba, João Pessoa, PB, Brazil
e-mail: [email protected]
A. B. Raposo Jr.
Departamento de Matemática, Universidade Federal do Maranhão, São Luís, MA, Brazil
e-mail: [email protected]
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 639
R. M. Aron et al. (eds.), Operator and Norm Inequalities and Related Topics,
Trends in Mathematics, https://doi.org/10.1007/978-3-031-02104-6_19
640 D. Núñez-Alarcón et al.
⎛ ⎛ ⎞ λ ⎞ λ1
n n 2
⎜ 2 ⎟
⎝ ⎝ T ej1 , ej2 ⎠ ⎠ ≤ Cp1 ,p2 T , (1)
j1 =1 j2 =1
⎛ ⎛ ⎞ λ ⎞ λ1
2
⎜
n
n
2 ⎟
⎝ ⎝ T ej1 , ej2 ⎠ ⎠ ≤ Cp1 ,p2 T , (2)
j2 =1 j1 =1
The Hardy-Littlewood Inequalities in Sequence Spaces 641
and
⎛ ⎞1
μ
n
n
⎝ T ej1 , ej2
μ⎠
≤ Cp1 ,p2 T , (3)
j1 =1 j2 =1
for all bilinear forms T : np1 × np2 → K and all positive integers n. Moreover, the
exponents λ and μ are optimal.
Theorem 1.2 ([32, Theorems 2 and 4]) Let p1 , p2 ∈ [2, ∞], with 1/2 < 1/p1 +
1/p2 < 1. There is a constant Cp1 ,p2 such that the inequalities (1) and (2) are still
true, and
⎛ ⎞1
λ
n
n
λ
⎝ T ej1 , ej2 ⎠ ≤ Cp1 ,p2 T , (4)
j1 =1 j2 =1
for all bilinear forms T : np1 × np2 → K and all positive integers n. Moreover, λ is
optimal.
Theorem 1.3 ([32, Theorem 5]) Let p1 , p2 ∈ (1, ∞] be such that 1/p1 + 1/p2 <
1. Then
⎛ ⎛ ⎞ λ ⎞ λ1
p∗
⎜
n
n
p ∗ 2
⎟
⎝ ⎝ T ej1 , ej2 2 ⎠ ⎠ ≤ T
j1 =1 j2 =1
for all non-negative (i.e., T ej1 , ej2 ≥ 0 for all (j1 , j2 ) ∈ N × N) bilinear forms
T : np1 × np2 → K and all positive integers n.
Theorem 1.4 ([32, Theorems 3 and 4]) Let p1 , p2 ∈ (1, ∞] be such that 1/p1 +
1/p2 < 1. If pk1 ∈ (1, 2] and pk2 ∈ (2, ∞], for {k1 , k2 } = {1, 2}, then there is a
constant Cp1 ,p2 such that
⎛ ⎞1
n
n λ
λ
⎝ T ej1 , ej2 ⎠ ≤ Cp1 ,p2 T , (5)
j1 =1 j2 =1
for all bilinear forms T : np1 × np2 → K and all positive integers n. Moreover, λ is
optimal.
642 D. Núñez-Alarcón et al.
The constant and, in some sense, the exponents in Theorem 1.4 were improved
by Ozikiewicz and Tonge in [39]:
Theorem 1.5 ([39, Theorem 5]) Let p1 , p2 ∈ (1, ∞] be such that 1/p1 + 1/p2 <
1. If pk1 ∈ (1, 2] and pk2 ∈ (2, ∞], for {k1 , k2 } = {1, 2}, then
⎛ ⎛ ⎞ λ ⎞ λ1
pk∗
⎜
n
n
pk∗ 2
⎟
⎝ ⎝ T ej1 , ej2 2 ⎠ ⎠ ≤ T ,
jk1 =1 jk2 =1
for all positive integers n and all bilinear forms T : np1 × np2 → K.
Several natural questions related to the previous results arise:
• How do these inequalities behave for multilinear forms?
• What are the best constants Cp1 ,p2 ?
• What happens when 1/p1 + 1/p2 ≥ 1?
In the next sections we present modern proofs of the Hardy–Littlewood inequal-
ities and discuss the three issues mentioned above. More precisely, in Sect. 2 we
present some auxiliary results that will be used throughout the text. Sections 3–6
are devoted to present proofs of m-linear versions of Theorems 1.1–1.5. In Sects. 7
and 8 we try to present the state-of-the-art of the investigation related to the optimal
constants of the Hardy–Littlewood inequalities and the case 1/p1 +· · · +1/pm ≥ 1.
Finally, in the final section we present connections between the Hardy–Littlewood
inequalities and the Gale–Berlekamp switching game.
2 Preliminary Results
From now on, E ∗ denotes the topological dual of a Banach space E and BE denotes
n
the closed unit ball of E. Moreover, denotes the multiple summation
B
ik =1
n
n
n
n
··· ···
i1 =1 ik−1 =1 ik+1 =1 im =1
and, for all pm = (p1 , . . . , pm ) ∈ [1, ∞]m , and each k ∈ {1, . . . , m}, let us define
1 1 1 1
1/pm ≥k
:= + ··· + and 1/pm := 1/pm ≥1
= + ··· + .
pk pm p1 pm
⎜
2
n n
⎟ √ m−1
⎜ ⎝ A(ei1 , . . . , eim ) ⎠
2 ⎟ ≤ 2 A.
⎝ ⎠
ik =1 iBk =1
√ m−1
Proof For the sake of convenience, we shall denote M = 2 and ai =
A(ei1 , . . . , eim ) . Let
6 7
l := card j : pj = ∞ .
When l < m note that the conditions (i) and (ii) reduce to precisely 1/pm ≤
1/2. We begin by proving by induction on l ∈ {0, . . . , m − 1} that, for any
pm = (p1 , . . . , pm ) ∈ [2, ∞]m with 1/pm ≤ 1/2, we have
⎛ ⎛ ⎞ 1 ×λp ⎞ λp1m
m
n n 2
⎜ ⎝ 2⎠ ⎟
⎝ ai ⎠ ≤ MA (6)
ik =1 iBk =1
for all m-linear forms A : np1 × · · · × npm → K, and all k ∈ {1, . . . , m}.
For l = 0, let n ∈ N and let A : n∞ × · · · × n∞ → K be an m-linear form. Then,
λpm = 1 and, by the Khinchin inequality for multiple sums (see [24, p. 455]) we
have
⎛ ⎛ ⎞ 1 ×λp ⎞ λp1m ⎛ ⎞1
2 m 2
⎜
n
n
⎟
n
n
⎝ ⎝ 2⎠
ai ⎠ = ⎝ ai ⎠ ≤ M A
2
ik =1 iBk =1 ik =1 iBk =1
6 7
6 for card 7j : pj = ∞ = l − 1 and let us
Let us assume that the result is valid
prove that it is also valid when card j : pj = ∞ = l. If k is an index such that
pk = ∞, we fix x ∈ npk and consider
defined by
(k) n
where xz(k) = xj zj . By applying the induction hypothesis to Ak , we know
j =1
that
⎛ ⎛ ⎞ 1 ×λk ⎞ λ1k
n n 2
⎜ ⎝ ⎟
ai2 ⎠
λk
⎝ xik ⎠
ik =1 iBk =1
⎛ ⎞ 1
⎛ ⎞ λk λk
2
⎜
n
n
⎟
=⎜ ⎝ A(ei1 , . . . , eik−1 , xeik , eik+1 , . . . , eim ) ⎠ ⎟
2
⎝ ⎠
ik =1 iBk =1
⎛ ⎞ 1
⎛ ⎞ λk λk
⎜
2
n n
⎟
=⎜ ⎝ Ak (ei1 , . . . , eim ) ⎠ ⎟
2
⎝ ⎠
ik =1 iBk =1
where we have set 1/λk := 1/λpm + 1/pk . Since (pk /λk )∗ = λpm /λk , we get
⎛ ⎛ ⎞ λ1
⎞ 1 ×λp ⎞ λp1m
1
⎛ ⎛ ⎞ 1 ×λk × ∗ ×
pk k pk ∗
2 m 2 λk
⎜
n
n
⎟ ⎜
n
n
⎟ λk
⎝ ⎝ ai2 ⎠ ⎠ = ⎜ ⎝ a 2⎠ ⎟
⎝ i ⎠
ik =1 B
ik =1 ik =1 iBk =1
⎛⎛ ⎞
⎞ 1 ×λk ⎞n
1
⎛ n λk
⎜⎜ 2
2
⎟
⎟
=⎜
⎝⎝⎝ a ⎠ ⎠ ⎟
⎠
i
B
ik =1
ik =1 (pk /λk )∗
The Hardy-Littlewood Inequalities in Sequence Spaces 645
⎛ ⎛ ⎞ 1 ×λk ⎞ λ1k
2
⎜
n
n
⎟
=⎝ sup yik ⎝ ai2 ⎠ ⎠
y∈Bn B
(pk /λk ) ik =1 ik =1
⎛ ⎛ ⎞ 1 ×λk ⎞ λ1k
n n 2
⎜ λk ⎝ 2⎠ ⎟
= ⎝ sup xik ai ⎠
x∈Bn i =1
pk k iBk =1
≤ MA,
where the last inequality holds by (7). This shows that (6) holds when k is an index
such that pk = ∞. Let us also show that it also holds when k is an index such that
pk = ∞. Without loss of generality, let us suppose that p1 = ∞. The induction
hypothesis applied to A1 and
1 1 1
= +
λ1 λpm p1
now give
⎛ ⎞ 1 ×λ1
2
n
n
∀x ∈ Bnp , ⎝ ai2 xi1
2⎠
≤ M λ1 Aλ1 . (8)
1
ik =1 iBk =1
⎛ ⎞1
2
n
Sik = ⎝ ai2 ⎠ .
iBk =1
646 D. Núñez-Alarcón et al.
n λ pm
We shall show that ik =1 Sik ≤ M λpm Aλpm . We write
n
λp
n
λp −2
n
Sik m = Sik m ai2
ik =1 ik =1 iBk =1
n
n
ai2
= 2−λpm
ik =1 iBk =1 Sik
n
n 2/s
ai 2/s ∗
= 2−λpm
ai
i1 =1 iB1 =1 Sik
⎛ ⎛ ⎞ t ⎞ 1t ⎛ ⎛ ⎞ t ∗∗ ⎞ t ∗
1
s
⎜
n
n
ai2 ⎜ ⎝ 2⎠ ⎟
n n s
≤ ⎝ ⎝ ⎠ ⎟⎠ ⎝ ai ⎠ .
(2−λpm )s
i1 =1 iB1 =1 Sik i1 =1 iB1 =1
Choosing
2 − λ1 t∗ λp
s= and ∗
= m
2 − λpm s 2
⎛ ⎞1 ⎛ ⎞ 1∗
⎛ ⎞ λpm t ⎛ ⎞ λpm t
λ1 2
n
⎜
n
n
ai2 ⎟ ⎜
n
n
⎟
λ pm
Sik ≤ ⎜ ⎝ ⎠ ⎟ ⎜ ⎝ ai2 ⎠ ⎟
⎝ (2−λ1 ) ⎠ ⎝ ⎠
ik =1 i1 =1 B S
i1 =1 k
i i1 =1 iB1 =1
⎛ ⎞1
⎛ ⎞ λpm t
⎜
λ1
λpm
n n
ai2 ⎟
≤ (MA) t∗ ⎜ ⎝ ⎠ ⎟ .
⎝ (2−λ1 ) ⎠
i1 =1 S
iB=1 ik
1
The Hardy-Littlewood Inequalities in Sequence Spaces 647
It remains to control the last sum appearing on the right hand side of the previous
inequality. By (the converse of) the Hölder inequality, it is sufficient to show that
⎛ ⎞
n
n
ai2
⎝ ⎠ yi1 ≤ (MA)λ1
(2−λ1 )
i1 =1 B S
i =1 ik
1
∗
for any y ∈ Bn ∗
. Since λpm /λ1 = p1 /λ1 , this means that we shall prove
(λpm /λ1 )
that
n
n
ai2 λ1
2−λ1
xi1 ≤ (MA)λ1
i1 =1 iB=1 Sik
1
n
n
ai2 λ1
n
n
ai2 λ1
xi1 = xi1
i1 =1 B S 2−λ1
i =1 ik
2−λ1
ik =1 iB=1 Sik
1 k
⎛ ⎞ 2−λ1 ⎛ ⎞ λ1
2
n n n 2
ai2
≤ ⎝ ⎠ ⎝ ai2 xi1
2⎠
.
B
Si2k
ik =1 ik =1 iBk =1
The second result is a technical lemma that appears in [11, Lemma 3.1]:
Lemma 2.2 ([11, Lemma 3.1]) Let m ≥ 1, qm = (q1 , . . . , qm ) ∈ (0, ∞)m , and
pm = (p1 , . . . , pm ) ∈ (1, ∞]m , with 1/pm < 1. If there is a constant Cpm ,qm such
that
⎛ ⎞1
⎛ ⎛ ⎞ qm−1 ⎞ qq1 q1
2
⎜ qm
⎟
⎜ ⎝
n n
⎜ ⎟ ⎟
⎜ ⎝· · · T (ej1 , . . . , ejm )qm ⎠ ···⎠ ⎟ ≤ Cpm ,qm T
⎝ ⎠
j1 =1 jm =1
for all n ∈ N and all non-negative m-linear forms T : np1 × · · · × npm → K, then
1/qk ≤ 1 − 1/pm ≥k
,
⎛ ⎞1
q
n
⎝ T (ej1 ) q⎠
≤ Cp,q T
j1 =1
for all non-negative linear forms T : np → K. For each n, consider the non-negative
n ∗
linear form Tn (x) = xj . By the Hölder inequality, we have Tn ≤ n1/p . On
j =1
the other hand
⎛ ⎞1
q
n
1
⎝ Tn (ej ) ⎠ = n q
q
j =1
and, since n is arbitrary, we conclude that 1/q ≤ 1/p∗ and the case m = 1 is done.
Now, let us proceed by induction. Suppose that the result is valid for m − 1 and let
1/pm < 1. Thus, 1/pm ≥2 < 1 and the induction hypothesis combined with a
simple argument of summability tells us that, if there is a constant Cpm ,qm such that
⎛ ⎞1
⎛ ⎛ ⎞ qm−1 ⎞ qq1 q1
2
⎜ qm
⎟
⎜ ⎝
n n
⎜ ⎟ ⎟
⎜ ⎝· · · T (ej1 , . . . , ejm )qm ⎠ ···⎠ ⎟ ≤ Cpm ,qm T
⎝ ⎠
j1 =1 jm =1
The Hardy-Littlewood Inequalities in Sequence Spaces 649
for all n ∈ N and all non-negative m-linear forms T : np1 × · · · × npm → K, then
1/qk ≤ 1 − 1/pm ≥k
For each n consider the non-negative m-linear form Bn : np1 × · · · × npm → K given
by
n
(1) (2) (m)
Bn (x (1) , . . . , x (m) ) = xj xj · · · xj .
j =1
n
(1) (2) (m)
Bn = sup xj xj · · · xj
x (1) ,...,x (m) ≤1 j =1
⎛ ⎛ ⎞ 1 ⎞
m λpm
⎜ n
x (k)
n
⎝ λpm ⎠ ⎟
≤ sup ⎝ j |1| ⎠
x (1) ,...,x (m) ≤1 j =1 np
k=1 k j =1
≤ n1/λpm .
and, since n is arbitrary, we conclude that 1/q1 ≤ 1/λpm and the proof is completed.
In this section we prove the following extended version of Theorem 1.1 for m-linear
forms.
650 D. Núñez-Alarcón et al.
Theorem 3.1 ([5, Theorem 1.3]) Let m ≥ 2, let pm = (p1, . . . , pm ) ∈ [2, ∞]m be
m
such that 0 ≤ 1/pm ≤ 1/2, and qm = (q1 , . . . , qm ) ∈ λpm , 2 . The following
statements are equivalent:
(i) There is a constant Cpm ,qm such that, for each n ∈ N and each m-linear form
T : np1 × · · · × npm → K we have
⎛ ⎞1
⎛ ⎛ ⎞ qm−1 ⎞ qq1 q1
2
⎜ ⎟
⎜ ⎝
n n qm
⎜ qm ⎠ ⎟ ⎟
⎜ ⎝· · · T ei1 , . . . , eim ···⎠ ⎟ ≤ Cpm ,qm T .
⎝ ⎠
i1 =1 im =1
(9)
(ii)
m+1
1/qm ≤ − 1/pm . (10)
2
Proof (i) ⇒ (ii). The Kahane–Salem–Zygmund inequality (see, for instance, [3])
assures the existence of an m-linear form T0 : np1 ×· · ·×npm → K with coefficients
±1 satisfying
T0 ≤ Km n 2 −
m+1
|1/pm |
for a certain constant Km . Plugging this m-linear form into (9) we arrive at (ii).
(ii) ⇒ (i). Let n ∈ N and let T : np1 × · · · × npm → K be an m-linear form. By
Proposition 2.1, we have
⎛ ⎞ 1
⎛ ⎞ λpm λpm
⎜
2
n n
⎟ √ m−1
⎜ ⎝ T (ei1 , . . . , eim ) ⎠
2 ⎟ ≤ 2 T .
⎝ ⎠
ik =1 iBk =1
for all n and all k ∈ {1, . . . , m}. Using a well-known result of Minkowski (see [31,
Corollary 5.4.2]), since λpm ≤ 2, we can interchange the position of the exponent 2
√ m−1
and obtain m inequalities with the same constant 2 and exponents
⎧
⎪
⎪ q1,1 , q2,1, q3,1, . . . , qm,1 = λpm , 2, 2, . . . , 2 ,
⎪
⎨ q1,2 , q2,2, q3,2, . . . , qm,2 = 2, λp , 2, . . . , 2
m
⎪ ..
⎪
⎪ .
⎩
q1,m , q2,m , . . . , qm−1,m , qm,m = 2, 2, . . . , 2, λpm .
The Hardy-Littlewood Inequalities in Sequence Spaces 651
2λpm − λpm qj
θj = ,
2qj − λpm qj
1 θk m
=
qj qj,k
k=1
and, by the Hölder inequality for mixed sums (see [1, Theorem 2.49]), we conclude
that
⎛ ⎞1
⎛ ⎛ ⎞ qm−1 ⎞ qq1 q1
2
⎜ ⎟
⎜ ⎝
qm
⎜
n n
qm ⎠ ⎟ ⎟ √ m−1
⎜ ⎝· · · T ei1 , . . . , eim ···⎠ ⎟ ≤ 2 T .
⎝ ⎠
i1 =1 im =1
Observe that, if we take
2m
q1 = · · · = qm = ∈ λpm , 2 ,
m + 1 − 2 1/pm
we obtain
m+1
1/qm = − 1/pm ,
2
providing the following extension of (3) to the multilinear setting:
Theorem 3.2 ([45, Theorem B]) Let m ≥ 2, and let pm = (p1 , . . . , pm ) ∈ [2, ∞]m
be such that 0 ≤ 1/pm ≤ 1/2. There is a constant Cpm ≥ 1 such that, for every
positive integer n and each m-linear form T : np1 × · · · × npm → K, we have
⎛ ⎞ m+1−2|1/pm |
2m
n
2m
⎝ T ei1 , . . . , eim m+1−2|1/pm | ⎠ ≤ Cpm T .
i1 ,...,im =1
Moreover, the exponent 2m/ m + 1 − 2 1/pm is optimal.
652 D. Núñez-Alarcón et al.
An
m-linear
operator between Banach spaces T : E1 × · · · × Em → F is multiple
r; pm -summing (see e.g. [18, 34]), where pm = (p1 , . . . , pm ), if there is C > 0
such that
⎛ ⎞1 ⎛ ⎞ 1
n r r
m
n pk
(j ) n
for every xij ⊆ Ej , with j = 1, . . . , m.
ij =1
We denote by m
, . . . , Em ; F )
the space composed by all multiple
(r;pm ) (E1
r; pm -summing m-linear operators T : E1 × · · · × Em → F . In the case of linear
operators, this notion reduces to the well-known concept of absolutely summing
operators (see [28]).
The following extension of Theorem 1.2 is due to Dimant and Sevilla-Peris ([29,
Proposition 4.1]); the proof presented here follows the lines of [9, Theorem 2.2].
Theorem 4.1 ([29, Proposition 4.1]) Let m ≥ 2, and let pm = (p1 , . . . , pm ) ∈
[2, ∞]m be such that 1/2 ≤ 1/pm < 1. There is a constant Cpm such that
⎛ ⎞ 1
λpm
n
⎝ T ei1 , . . . , eim
λpm ⎠
≤ Cpm T ,
i1 ,...,im =1
for all m-linear forms T : np1 ×· · ·×npm → K and all positive integers n. Moreover,
the exponent λpm is optimal.
Proof The optimality of the exponent λpm is a straightforward consequence of
Lemma 2.2. Let
G
1 1
r
s = min r : there exist r indexes k1 , . . . , kr such that ≤ <1 .
2 pki
i=1
For the sake of simplicity let us suppose that pk1 = p1 , . . . , pks = ps . Observe that
1 1 1
+ ···+ < .
p1 ps−1 2
The Hardy-Littlewood Inequalities in Sequence Spaces 653
Let n ∈ N and let A : np1 × · · · × nps → K be s-linear. Let us fix x ∈ nps and
consider
√ s−1
≤ 2 Axnps
1 λs−1 λs−1
∗ =1− = ,
(ps /λs−1 ) ps λs
⎛ ⎛ ⎞ λs ⎞ λs
1
n n 2
⎜ ⎝ 2⎠ ⎟
⎝ A(ei1 , . . . , eis ) ⎠
is =1 i1 ,...,is−1 =1
⎛⎛ ⎞ 1
⎛ ⎞λs−1 ⎞n λs−1
⎜⎜ n ⎟
⎟
=⎜ ⎝⎝ A(ei1 , . . . , eis ) ⎠ ⎟
2
⎝
⎠
⎠
i1 ,...,is−1 =1 ∗
is =1 (ps /λs−1 )
⎛ ⎞ 1
⎛ ⎞λs−1 λs−1
⎜
n
n
⎟
=⎜ yis ⎝ A(ei1 , . . . , eis ) ⎠ ⎟
2
⎝ sup ⎠
y∈Bn is =1 i1 ,...,is−1 =1
ps
λs−1
654 D. Núñez-Alarcón et al.
⎛ ⎛ ⎞λs−1 ⎞ λs−1
1
⎜
n
n
⎟
= ⎝ sup xis
λs−1 ⎝ A(ei1 , . . . , eis ) ⎠
2
⎠
x∈Bn i =1 i1 ,...,is−1 =1
ps s
√ s−1
≤ 2 A.
By the Khinchin inequality (see [28, Theorem 1.10]), with constant 1, because λs ≥
2, we have, for every n and all (s +1)-linear forms Ts+1 : np1 ×· · ·×nps ×n∞ → K,
⎛ ⎛ ⎞ λs ⎞ λs
1
2
⎜
n
n
2 ⎟
⎝ ⎝ Ts+1 ei1 , . . . , eis+1 ⎠ ⎠
i1 ,...,is =1 is+1 =1
⎛ ⎛ ⎞ λs ⎞ λs
1
$ λs λs
⎜
n
⎜
1
n
⎟ ⎟
≤⎜
⎝ ⎝ Ts+1 ei1 , . . . , eis+1 ris+1 (t) dt ⎠ ⎟ ⎠
i1 ,...,is =1 0 is+1 =1
⎛ ⎛ ⎞ ⎞1
$ 1
n
n
λs λs
⎜ ⎟
=⎝ Ts+1 ⎝ei1 , . . . , eis , eis+1 ris+1 (t)⎠ dt ⎠
0 i ,...,i =1 is+1 =1
1 s
⎛ ⎛ ⎞ ⎞1
λs λs
⎜
n
n
⎟
≤ ⎝ sup Ts+1 ⎝ei1 , . . . , eis , eis+1 ris+1 (t)⎠ dt ⎠
t ∈[0,1] i ,...,i =1 is+1 =1
1 s
⎛ ⎛ ⎞ ⎞1
λs λs
⎜
n n
⎟
= sup ⎝ Ts+1 ⎝ei1 , . . . , eis , eis+1 ris+1 (t)⎠ dt ⎠
t ∈[0,1] i1 ,...,is =1 is+1 =1
The Hardy-Littlewood Inequalities in Sequence Spaces 655
⎛ ⎞
√ s−1 n
≤ 2 sup T
s+1
⎝·, . . . , ·, e r
is+1 is+1 (t) ⎠
t ∈[0,1] is+1 =1
√ s−1
= 2 Ts+1 .
Here and henceforth, rj (t) denotes the j -th Rademacher function. Thus, from the
previous inequality together with canonical inclusion of p spaces,
⎛ ⎞1
n λs
⎝ Ts+1 ei1 , . . . , eis+1
λs ⎠
i1 ,...,is+1 =1
⎛ ⎛ ⎞ λs ⎞ λs
1
λs
⎜
n n
λs ⎟
=⎝ ⎝ Ts+1 ei1 , . . . , eis+1 ⎠ ⎠
i1 ,...,is =1 is+1 =1
⎛ ⎛ ⎞ λs ⎞ λs
1
n n 2
⎜ 2 ⎟
≤⎝ ⎝ Ts+1 ei1 , . . . , eis+1 ⎠ ⎠
i1 ,...,is =1 is+1 =1
√ s−1
≤ 2 Ts+1 ,
for every n and all (s + 1)-linear forms Ts+1 : np1 × · · · × nps × n∞ → K. Using the
canonical isometric isomorphisms for the spaces of weakly summable sequences
(see [28, Proposition 2.2]) we know that this is equivalent to assert that (see [29,
p. 308])
s+1
(E , . . . , Es+1 ; K) = L(E1 , . . . , Es+1 ; K)
(λs ;p1∗ ,...,ps∗ ,1) 1
by the inclusion theorem for multiple summing forms proved in [2, 13], we have
s+1 (E1 , . . . , Es+1 ; K)
λs+1 ;p1∗ ,...,ps+1
∗ = L(E1 , . . . , Es+1 ; K)
656 D. Núñez-Alarcón et al.
for all (s + 1)-linear forms T : np1 × · · · × nps × nps+1 → K and all positive integers
n. The proof is completed by a standard induction argument.
From now on, for all pm = (p1 , . . . , pm ) ∈ [1, ∞]m , with 1/pm < 1, and each
k ∈ {1, . . . , m}, let us define
1/λpm ,k = 1 − 1/pm ≥k
.
for all non-negative m-linear forms T : np1 × · · · × npm → K and all positive
integers n.
(ii) There is a constant Cpm ,qm such that
⎛ ⎛ ⎞ qq1 ⎞ q1
1
⎛ ⎞ qm−1 2
⎜ ⎜
qm
⎟ ⎟
n n n
⎜ ⎝ T (ej1 , . . . , ejm )qm ⎠ ···⎠ ⎟
⎝ ⎝ · · · ⎠ ≤ Cpm ,qm T
jσ1 =1 jσ2 =1 jσm =1
The Hardy-Littlewood Inequalities in Sequence Spaces 657
for all non-negative m-linear forms T : np1 × · · · × npm → K and all positive
integers n.
(iii) The exponents q1 , . . . , qm satisfy
qk ≥ λtm ,k ,
for all k ∈ {1, . . . , m}, with tm = pσ1 , . . . , pσm .
Proof To simplify the notation we will consider σj = j for all j ; the other cases
are similar. We observe that, in this case, λtm ,k = λpm ,k for all k ∈ {1, . . . , m}.
(i) ⇒ (ii) is obvious and (ii) ⇒ (iii) follows from Lemma 2.2. So, it
remains to prove (iii) ⇒ (i). In the case m = 1 the result is immediate, it holds
with constant 1 and does not need the non-negative assumption. Let us show the
general case m, supposing that the result holds for m − 1; so we suppose that if
pm−1 = (p1 , . . . , pm−1 ) ∈ (1, ∞]m−1 is such that 1/pm−1 < 1, then
⎛ ⎛ ⎞ μμ1 ⎞ μ1
1
⎛ ⎞ μm−2 2
μm−1
⎜n
⎜ ⎝
n
⎟ ⎟
⎜ T (ej1 , . . . , ejm−1 )μm−1 ⎠ ···⎠ ⎟
⎝ ⎝· · · ⎠ ≤ T ,
j1 =1 jm−1 =1
where
(μ1 , . . . , μm−1 ) = λpm−1 ,1 , . . . , λpm−1 ,m−1 ,
for all positive integers n and all non negative (m − 1)-linear forms T : np1 × · · · ×
npm−1 → K. We recall that for an m-linear form D : np1 × · · · × npm → K, we have
n
(1) (m)
= sup D ej1 , . . . , ejm xj1 · · · xjm
x (i) np ≤1; 1≤i≤m j1 ,...,jm =1
i
⎛ ∗ ⎞ 1
∗
pm pm
⎜
n
n
⎟
= sup ⎝ D ej1 , . . . , ejm xj(1)
1
· · · xj(m−1)
m−1
⎠ .
x (i) np ≤1; 1≤i≤m−1 jm =1 j/
m =1
i
(12)
658 D. Núñez-Alarcón et al.
Suppose that pm = (p1 , . . . , pm ) ∈ (1, ∞]m satisfies 1/pm < 1. In this case
1 1 1 1 1
+ ···+ <1− = = ∗
p1 pm−1 pm λpm ,m pm
n
∗
T (ej1 , . . . , ejm−1 ) = D(ej1 , . . . , ejm )pm , (13)
jm =1
n
(1) (m−1)
sup T (ej1 , . . . , ejm−1 )xj1 · · · xjm−1
x (i) nr ≤1; 1≤i≤m−1 j/
m =1
i
n
(1) (m−1)
= sup T (ej1 , . . . , ejm−1 )xj1 · · · xjm−1
x (i) + nr ≤1; 1≤i≤m−1 j/
m =1
i
n ∗
pm
= sup T (ej1 , . . . , ejm−1 ) xj(1)
1
· · · xj(m−1)
m−1
x (i) + np ≤1; 1≤i≤m−1 j/=1
m
i
n
n
∗ (1)
∗
(m−1) pm
= sup D(ej1 , . . . , ejm )pm xj1 · · · xjm−1
x (i) + np ≤1; 1≤i≤m−1 jm =1 j/=1
m
i
⎛ ⎞pm∗
n
n
= sup ⎝ D(ej1 , . . . , ejm )xj1 · · · xjm−1 ⎠
(1) (m−1)
for all x (k) ∈ Bnr , with k = 1, . . . , m − 1, where the last equality holds by (12).
k
The Hardy-Littlewood Inequalities in Sequence Spaces 659
Note that
m−1
1 1 m−1
1
−1 m−1
1
∗
= pm = 1− < 1.
rk pk pm pk
k=i k=i k=i
and
(β1 , . . . , βm−1 ) = λrm−1 ,1 , . . . , λrm−1 ,m−1
we have
⎛ ⎛ ⎞ αα1 ⎞ α1 ×pm
1 ∗
⎛ ⎞ αm−1 2
⎜ ⎜ ⎝
αm
⎟ ⎟
n n
∗
⎜ ⎝· · · D(ej1 , . . . , ejm )pm ⎠ ···⎠ ⎟
⎝ ⎠
j1 =1 jm =1
⎛ ⎛ ⎞ αα1 ⎞ α1 ×pm
1 ∗
⎛ ⎞ αm−2 2
αm−1
⎜n
⎜ ⎝
n αm−1
⎟ ⎟
=⎜
⎝ ⎝· · · T (e j1 , . . . , e jm−1 ) αm ⎠ ···⎠ ⎟ ⎠
j1 =1 jm−1 =1
⎛ ⎛ ⎞ β1 ⎞ β1
1
⎛ ⎞ βm−2 β2
⎜n
⎜
n βm−1
⎟ ⎟
⎜ ⎜ ⎟
=⎜ ⎝· · ·
⎝ T (ej1 , . . . , ejm−1 ) βm−1 ⎠
···⎟
⎠ ⎟ .
⎝ ⎠
j1 =1 jm−1 =1
⎛ ⎞λ 1 ∗
×pm
⎛ ⎞ λλpm ,1 pm ,1
⎛ ⎞ λpm ,m−1 pm ,2
⎜ n λpm ,m ⎟
⎜ ⎜ n
∗ ⎟ ⎟
⎜ ⎜· · · ⎝ D(ej1 , . . . , ejm )pm ⎠ ···⎟ ⎟
⎜ ⎝ ⎠ ⎟
⎝j =1 jm =1 ⎠
1
660 D. Núñez-Alarcón et al.
n
(1) (m−1)
≤ sup T (ej1 , . . . , ejm−1 )xj1 · · · xjm−1
x (i) nr ≤1; 1≤i≤m−1 j/
m =1
i
∗
≤ Dpm ,
for all m-linear forms T : np1 × · · · × npm → K and all positive integers n.
(ii) There is a constant Cpm ,qm such that
⎛ ⎛ ⎞ qq1 ⎞ q1
1
⎛ ⎞ qm−1 2
⎜ ⎜
qm
⎟ ⎟
n n n
⎜ ⎝ qm ⎠
···⎠ ⎟
⎝ ⎝ · · · T (ej1 , . . . , ejm ) ⎠ ≤ Cpm ,qm T
jσ1 =1 jσ2 =1 jσm =1
for all m-linear forms T : np1 × · · · × npm → K and all positive integers n.
(iii) The exponents q1 , . . . , qm satisfy
qk ≥ λtm ,k ,
for all k ∈ {1, . . . , m}, with tm = pσ1 , . . . , pσm .
Proof To simplify the notation we will consider σj = j for all j ; the other cases
are similar. (i) ⇒ (ii) is immediate and (ii) ⇒ (iii) follows from Lemma 2.2. Let
us prove (iii) ⇒ (i). Let T : np1 × · · · × npm → K be an m-linear form. We define
The Hardy-Littlewood Inequalities in Sequence Spaces 661
2
A(ej1 , . . . , ejm ) = T (ej1 , . . . , ejm ) ,
⎛ ∗
pm
⎞ 2
∗
pm
⎜
2
n n
⎟
= sup ⎜ 2 (1) (m−1)
T (ej1 , . . . , ejm ) xj1 · · · xjm−1 ⎟
⎝ ⎠
x (i) ∈Bn jm =1 j/
si m =1
1≤i≤m−1
⎛ ∗
pm
⎞ 2
∗
pm
2
⎜n
n
(m−1) 2 ⎟
= sup ⎜ T (ej1 , . . . , ejm )
2 (1)
xj1 · · · xjm−1 ⎟
⎝ ⎠
x (i) ∈Bn jm =1 j/
pi m =1
1≤i≤m−1
⎛ ⎞ 2
⎛ ⎞ pm∗ ∗
pm
2
⎜n
n
(m−1) 2 ⎟
≤ sup ⎜ ⎝ T (ej1 , . . . , ejm )xj1 · · · xjm−1 ⎠
(1) ⎟ .
⎝ ⎠
x (i) ∈Bn jm =1 j/
pi m =1
1≤i≤m−1
We use the Khinchin inequality for multiple sums (see [24, p. 455]), with constant
∗ ≥ 2, to obtain
1 because pm
⎛ ⎞ 2
⎛ ⎞ pm∗ ∗
pm
2
⎜n
n
(m−1) 2 ⎟
sup ⎜ ⎝ T (ej1 , . . . , ejm )xj1 · · · xjm−1 ⎠
(1) ⎟
⎝ ⎠
x (i) ∈Bn jm =1 j/
pi m =1
1≤i≤m−1
⎛ ⎞ 2
⎛ ∗ ⎞ pm∗∗ ∗
pm
⎜n $
n
m−1
pm pm
⎟
⎜ ⎜ (k) ⎟ ⎟
≤ sup ⎜ ⎝ T (ej1 , . . . , ejm ) rjk (tk ) xjk dt ⎠ ⎟
x (i) ∈Bn ⎝ I m−1 ⎠
pi jm =1 j/
m =1 k=1
1≤i≤m−1
662 D. Núñez-Alarcón et al.
⎛ ⎞ 2
⎛ ∗ ⎞ pm∗∗ ∗
pm
⎜$
pm pm
⎟
⎜
n n m−1
⎜ (k) ⎟ ⎟
≤ sup ⎜ ⎝ T (e j1 , . . . , e jm ) rjk (tk ) xjk ⎠ dt ⎟
x (i) ∈Bn ⎝ I m−1 ⎠
pi jm =1 j/
m =1 k=1
1≤i≤m−1
$ G
2
∗
∗ pm
≤ sup T pm
dt : x (i) ≤ 1; 1 ≤ i ≤ m − 1
I m−1 np
i
= T 2 ,
A ≤ T 2 .
and
⎛ ⎞
1
m
1 2 2 2 − pm 1
2 = 2 ⎝1 − ⎠=1− + ···+ + =
λpm ,i pj pi pm−1 pm λsm ,i
j =i
for each i ∈ {1, . . . , m}. Combining these facts with Theorem 5.1, we obtain
⎛ ⎞λ 2
⎛ ⎞ λλpm ,1 pm ,1
⎛ ⎞ λp ,m−1
m pm ,2
⎜ pm∗ ⎟
⎜ n ⎜ n
∗ ⎟ ⎟
⎜ ⎜· · · ⎝ T (ej1 , . . . , ejm )
pm ⎠ ···⎟ ⎟
⎜ ⎝ ⎠ ⎟
⎝j1 =1 jm =1 ⎠
⎛ ⎞λ 2
⎛ ⎞ 2×λpm ,1 pm ,1
⎛ ⎞ 2×λp ,m−1
m 2×λ pm ,2
⎜ 2×pm∗ ⎟
⎜ n ⎜ n ∗
pm ⎟ ⎟
=⎜ ⎜· · · ⎝ ⎠ ···⎟ ⎟
2× 2
⎜ ⎝ T (ej1 , . . . , ejm ) ⎠ ⎟
⎝j1 =1 jm =1 ⎠
The Hardy-Littlewood Inequalities in Sequence Spaces 663
⎛ ⎞λ 1
⎛ ⎞ λλsm ,1 sm ,1
⎛ ⎞ λsm ,m−1
⎜
λsm ,m
sm ,2
⎟
⎜ n ⎜ n
⎟ ⎟
=⎜
⎜
⎜· · · ⎝
⎝ A(ej1 , . . . , ejm )λsm ,m ⎠ ···⎟
⎠
⎟
⎟
⎝j1 =1 jm =1 ⎠
≤ A ≤ T 2 .
Since
and
λtm ,1 = λpm ,
then, the extension of Theorem 1.4, with optimal exponent and optimal constant,
follows immediately from the previous theorem:
Theorem 6.2 ([11, Theorem 3.2] and [29, Proposition 4.1]) Let m ≥ 2,
pm = (p1 , . . . , pm ) ∈ (1, ∞]m be such that pi ∈ (1, 2] for some i and 1/pm < 1.
Then
⎛ ⎞ 1
n λpm
⎝ T (ej1 , . . . , ejm )
λpm ⎠
≤ T (15)
j1 ,...,jm =1
for all m-linear forms T : np1 ×· · ·×npm → K and all positive integers n. Moreover,
the exponent λpm is optimal.
The optimal exponent λpm in the above result was obtained in [29] and the
optimal constant 1 was obtained in [11].
The paper of Hardy and Littlewood does not investigate summability of the
coefficients of bilinear forms T : np1 × np2 → K when 1/p1 + 1/p2 ≥ 1. In fact,
in this case it is simple to verify that there is no finite exponent q so that there is a
constant C satisfying
⎛ ⎞1
n
n q
⎝ T ej1 , ej2
q⎠
≤ C T
j1 =1 j2 =1
664 D. Núñez-Alarcón et al.
for all bilinear forms T : np1 ×np2 → K and all positive integers n. This observation
seems to trivialize the problem since it is obvious that
sup T ej1 , ej2 ≤ T .
j1 ,j2 ∈N
Until very recently the investigation of the m-linear case has followed this vein and
ignored the case
However, if we consider the problem from a broader perspective, we observe that the
case (16) hides subtleties. In fact, under an anisotropic viewpoint (allowing different
exponents for different indexes) the problem is no longer trivial since a Hardy–
Littlewood type inequality
⎛ ⎞1
⎛ ⎛ ⎞ qm−1 ⎞ qq1 q1
2
⎜ ⎟
⎜ ⎝
n n qm
⎜ qm ⎠ ⎟ ⎟
⎜ ⎝· · · T ei1 , . . . , eim ···⎠ ⎟ ≤ C T
⎝ ⎠
i1 =1 im =1
⎛ ⎛ ⎞ s1 s2 ⎞ s2
1
⎛ ⎞ 1
sm sm−1 3
⎜n
⎜ ⎝
n
⎟ ⎟
sup ⎜ ⎟ ≤ Cm T
sm ⎠
⎝ ⎝ · · · T ej1 , . . . , ejm ···⎠ ⎠
j1 j2 =1 jm =1
(17)
for all m-linear forms T : nm × · · · × nm → K, and all positive integers n, with
2m(m − 1)
sk =
mk − 2k + 2
for all k = 2, . . . , m. Moreover, s1 = ∞ and s2 = m are sharp and, for m > 2 the
optimal exponents sk satisfying (17) fulfill
m
sk ≥ .
k−1
The Hardy-Littlewood Inequalities in Sequence Spaces 665
There are still several open problems related to the Hardy–Littlewood inequalities
for m-linear forms. The first class of open problems are related to the optimal
exponents. As we can easily observe, there are some cases not covered by the
previous theorems and the optimal exponents of these cases are not known, in
general. At least for bilinear forms and p1 , p2 ∈ [2, ∞], the optimal exponents
are known for the non-critical cases:
Theorem 8.1 ([44, Theorem 5.1]) Let p1 , p2 ∈ [2, ∞] with 1/p1 +1/p2 < 1, and
q1 , q2 > 0. For {k1 , k2 } = {1, 2} the following assertions are equivalent:
(i) There is a constant C such that
⎛ ⎛ ⎞ q1 ⎞ q11
q2
⎜
n
n
q2 ⎠ ⎟
⎝ ⎝ T (ej1 , ej2 ) ⎠ ≤ C T ,
jk1 =1 jk2 =1
for all bilinear forms T : np1 × np2 → K and all positive integers n.
(ii) The exponents q1 , q2 satisfy (q1 , q2 ) ∈ [λ, ∞) × [pk∗2 , ∞) and
1 1 3 1 1
+ ≤ − + .
q1 q2 2 p1 p2
T (x, y) = x1 y1 + x1 y2 + x2 y1 − x2 y2
√
has norm 2, we conclude that the constant √ 2 is sharp (see [30]). A similar
argument shows that (for real scalars) 2 is also the optimal constant for (9)
with (m, p1 , p2 ) = (2, ∞, ∞) , regardless of the q1 , q2 ∈ [1, 2]. The case
m > 2, with p1 = · · · = pm = ∞ and real scalars, with exponents
(q1 , . . . , qm ) ∈ {(1, 2, . . . , 2) , . . . , (2, 2, . . . , 2, 1)} was solved in [41, 43] and
666 D. Núñez-Alarcón et al.
the optimal constants for a more general family of exponents were obtained in
[22]. The best known estimates for case for m > 2, with pm = (∞, . . . , ∞)
and qm = m+1 2m 2m
, . . . , m+1 can be found in [15] (upper bounds) and [30] (lower
bounds):
⎧
⎪ j
⎪
⎪ 1− 1 446381 m 32 − j1 2−2j
⎪
⎪ ≤ ≤ √ for m ≥ 14 and real scalars,
⎪
⎪ 2 m C pm ,qm 2 55440
⎪
⎪ j =14
π
⎨
1 m 1
The precise asymptotic growth of the upper bounds provided above is also
calculated in [15]:
1−γ
Cpm ,qm ≤ κm 2 ≤ κm0.212 for complex scalars,
2−log 2−γ
Cpm ,qm ≤ κm 2 ≤ κm0.365 for real scalars,
by
⎛ ⎞ m+1
n 2m
2m
⎝ T ein11 , . . . , einkk
m+1
⎠ ,
i1 ,...,ik =1
k log k
lim = 0.
m→∞ m
The Hardy-Littlewood Inequalities in Sequence Spaces 667
Kn
lim = 1.
n→∞ (log n) /n
The constants of the polynomial Hardy–Littlewood inequalities for real scalars were
investigated in [20]. In particular, in [20] it was shown that for the case of real
scalars the constants are no longer sub-exponential, and behave differently than in
the complex case; applications can be found in [27].
668 D. Núñez-Alarcón et al.
Designed independently by Elwyn Berlekamp and David Gale in the 1960’s, the
Gale–Berlekamp switching game consists of an n × n square matrix of light bulbs
set-up at an initial light configuration. The goal is to turn off as many lights as
possible using n row and n column switches, which invert the state of each bulb in
the corresponding row or column.
For an initial pattern of lights , let i() denote the smallest final number of
on-lights achievable by row and column switches starting from . We define
which represents the smallest possible number of remaining on-lights, starting from
“the worst” initial pattern. Sometimes the problem is posed as to find the maximum
of the difference between the number of lights that are on and the number that are
off, often denoted by Gn . Since
1 2
Rn = n − Gn ,
2
both formulations are equivalent.
Originally, the problem introduced by Berlekamp asks for the exact value of R10 .
In 2004, Carlson and Stolarski proved that R10 = 35 (see [21]). Up to now, the exact
value of Rn is known only for small values of n due to involving combinatorial
arguments. The exact values of Rn for n up to 12 were obtained by Carlson and
Stolarski [21] (see also [19]).
In this section we show how the Hardy–Littlewood cycle of ideas is associated
to the Gale–Berlekamp switching game and how it is useful to provide asymptotic
bounds for Gn . We initially notice that byassociating +1 to the on-lights and −1 to
n
the off-lights from the array of lights aij i,j =1 , we have
⎧ ⎫
⎨
n ⎬
Gn = min max aij xi yj : aij = −1 or + 1 .
⎩xi ,yj ∈{−1,1} ⎭
i,j =1
The Hardy-Littlewood Inequalities in Sequence Spaces 669
Now, we shall observe that Gn is precisely the norm of the bilinear form A : n∞ ×
n∞ → R defined by
n
A (x, y) = aij xi yj . (19)
i,j =1
It n
isnwell-known that the extreme points of the closed unit ball of ∞ are precisely
xj j =1 such that xj = 1 for all j = 1, . . . , n. Finally, we recall that for all
finite-dimensional Banach spaces E we have
for all bilinear forms T : E × E → R, where ext (BE ) represents the extreme points
of the closed unit ball BE of E. In fact, since BE is compact, there exist x, y ∈ BE
such that |T (x, y)| = T . By the Minkowski Theorem, we know that BE is the
(1) (2) 1 (1)
convex hull of ext (BE ). Hence, there exist λi , λj ∈ [0, 1] with ki=1 λi =
k2 (2) (1) (2)
j =1 λj = 1 and zi , zj ∈ ext (BE ) such that
k1
(1) (1)
k2
(2) (2)
x= λi zi and y= λj zj .
i=1 j =1
Therefore
⎛ ⎞
k1
k2
T = |T (x, y)| = T ⎝ (1) (1)
λi zi , λj zj ⎠
(2) (2)
i=1 j =1
k1
k2
(1) (2) (1) (2)
≤ λi λj T zi , zj
i=1 j =1
(1) (2)
and we conclude that there exist zi0 , zj0 ∈ ext BE such that
(1) (2)
T zi0 , zj0 = T .
670 D. Núñez-Alarcón et al.
The above results yield that Gn = A for A as in (19). Now, as a straightforward
consequence of the Hardy–Littlewood inequalities for m = 2 and p1 = p2 = ∞,
we have
Gn ≥ 2−1/2n3/2 .
On the other
√ hand,
the Kahane–Salem–Zygmund inequality (see [16]) tells us that
Gn ≤ 2 5 log 9 n3/2 and we conclude that
1 Gn
√ ≤ 3/2 ≤ 2 5 log 9.
2 n
Gn
0.797 + o (1) ≤ ≤ 1 + o (1) .
n3/2
References
1. R.A. Adams, J.J.F. Fournier, Sobolev Spaces. Pure and Applied Mathematics Series, 2nd edn.
(Elsevier, Amsterdam, 2003)
2. N. Albuquerque, L. Rezende, Anisotropic regularity principle in sequence spaces and applica-
tions. Commun. Contemp. Math. 20(7), 1750087, 14 (2018)
3. N. Albuquerque, L. Rezende, Asymptotic estimates for unimodular multilinear forms with
small norms on sequence spaces. Bull. Braz. Math. Soc. 52(1), 23–39 (2021)
4. N. Albuquerque, F. Bayart, D. Pellegrino, J.B. Seoane-Sepúlveda, Sharp generalizations of the
multilinear Bohnenblust-Hille inequality. J. Funct. Anal. 266(6), 3726–3740 (2014)
5. N. Albuquerque, F. Bayart, D. Pellegrino, J.B. Seoane-Sepúlveda, Optimal Hardy-Littlewood
type inequalities for polynomials and multilinear operators. Isr. J. Math. 211(1), 197–220
(2016)
6. N. Albuquerque, G. Arajo, W. Cavalcante, T. Nogueira, D. Núñez, D. Pellegrino, P. Rueda, On
summability of multilinear operators and applications. Ann. Funct. Anal. 9(4), 574–590 (2018)
7. N. Alon, J.H. Spencer, The Probabilistic Method. Wiley Series in Discrete Mathematics and
Optimization, 4th edn. (Wiley, Hoboken, 2016). xiv+375 pp.
8. F.C. Alves, D. Serrano-Rodríguez, Complex Bohnenblust-Hille inequality whose monomials
have indices in an arbitrary set. São Paulo J. Math. Sci. 14(1), 242–248 (2020)
9. G. Araújo, K. Câmara, Universal bounds for the Hardy-Littlewood inequalities on multilinear
forms. Results Math. 73(3), paper 124, 10 (2018)
10. G. Araújo, D. Pellegrino, On the constants of the Bohnenblust-Hille and Hardy-Littlewood
inequalities. Bull. Braz. Math. Soc. 48(1), 141–169 (2017)
11. R.M. Aron, D. Núñez-Alarcón, D. Pellegrino, D. Serrano-Rodríguez, Optimal exponents for
Hardy-Littlewood inequalities for m-linear operators. Linear Algebra Appl. 531, 399–422
(2017)
The Hardy-Littlewood Inequalities in Sequence Spaces 671
Abstract They are many faces of C ∗ -algebras whose symmetries encode important
aspects of their structures. We show that in surprisingly different situations these
symmetries are implemented by Jordan *-isomorphisms and lead to full Jordan
invariants. In this respect we study the following structures: 1. One dimensional
projections in a Hilbert space with transition probability and orthogonality relation
(Wigner type theorems). 2. Projection lattices of von Neumann algebras and AW∗ -
algebras (Dye type theorems) 3. Abelian C∗ -subalgebras with set theoretic inclusion
(Bohrification program in quantum theory) 4. Measures on state spaces endowed
with the Choquet order.
1 Introduction
ψ(ab) = ψ(a)ψ(b)
ψ(a ∗ ) = ψ(a)∗ .
J. Hamhalter ()
Department of Mathematics, Faculty of Electrical Engineering, Czech Technical University in
Prague, Prague, Czech Republic
e-mail: [email protected]
E. Turilova
Institute of Mathematics and Mechanics, Kazan Federal University, Kazan, Russia
e-mail: [email protected]
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 673
R. M. Aron et al. (eds.), Operator and Norm Inequalities and Related Topics,
Trends in Mathematics, https://doi.org/10.1007/978-3-031-02104-6_20
674 J. Hamhalter and E. Turilova
It turns out that many features of operator algebras relevant to their inner structure
and applications in physics are preserved by Jordan *-isomorphism rather than *-
isomorphisms. A Jordan *-isomorphism is a bijective linear map ϕ : A → B
between C ∗ -algebras satisfying the following conditions for all a ∈ A.
ϕ(a 2) = ϕ(a)2
ϕ(a ∗ ) = ϕ(a)∗ .
As we can see Jordan morphisms are well behaved only with respect to individual
elements and involve *-morphisms as a special case. They are ubiquitous in
preserver theory of operator algebras. Even, the very Banach space structure
of a C ∗ -algebra is in fact complete Jordan invariant. Indeed, famous Kadison
isometry theorem and its ramifications say that if two C ∗ -algebras are isometric
as Banach spaces then there is a Jordan *-isomorphism between them. Moreover,
unital surjective linear isometries between C ∗ -algebras are precisely Jordan *-
isomorphisms. We would like to review some aspects of this surprising universality
of Jordan morphisms by reviewing some recent results the authors witnessed and
have pleasure to deal with during last decade. We start at elementary level in the
first section and then move to more advanced aspects of symmetries in operator
algebras. The second section is devoted to famous Wigner theorem [18]. Based
on our results in [1, 16], we provide elementary proof on symmetries on Hilbert
spaces that preserve transition probability between one dimensional projections
(rays). These symmetries are given by Jordan *-isomorphisms that are in this case
implemented by unitary or antiunitary map. The core of our approach consists
in elementary matrix algebra arguments proving Kadison’s result to the effect
that Jordan *-isomorphism on a matrix algebra is either *-isomorphism or *-
antiisomorphism. We also show that quantum logic version of Wigner theorem due
to Uhlhorn can be quickly derived from Gleason theorem. In the second section we
analyze Dye theorem that is a generalization of Wigner theorem to a much general
context of von Neumann algebras. In Jordan formulation Dye theorem says that
any bijection between (almost all) projection lattices of von Neumann algebras
preserving orthogonality in both directions extends to a Jordan *-isomorphism
between algebras themselves. In other words, the projection lattice is a full Jordan
invariant. We show that this deep result can be quickly derived from Generalized
Gleason theorem due to Bunce and Wright. In the second part of Sect. 3 we outline
the proof of Dye’s theorem for AW∗ -algebras [7]. These algebras are algebraic
generalizations of von Neumann algebras due to Kaplansky. They are more natural
from the point of view of quantum theory. Also they seem to be useful from purely
mathematical point of view. For example, there is one-one correspondence between
complete Boolean algebras and projection lattices of abelian AW∗ -algebras. The
technique of proving Dye theorem for AW∗ -algebras must be different in combining
insights of von Neumann, Dye, and Heunen and Reyes [12] on one side and Gleason
theorem for Type I finite AW∗ -algebras due to J. Hamhalter [7] on the other side.
In Sect. 4 we consider yet another order structure associated with C ∗ -algebras. It is
Symmetries of C ∗ -algebras and Jordan Morphisms 675
the poset of abelian subalgebras order by set theoretic inclusion. This structure has
driven attention of many mathematical physicists in connection with Bohrification
program in quantum theory [14]. In this approach quantum system can only be seen
through its classical subsystems. In operator algebraic model of quantum theory
quantum system is given by a C ∗ -algebra A and its classical subsystems are given by
abelian C ∗ -subalgebras of A. Therefore the poset C(A) of abelian C ∗ -subalgebras
of A embodies the Bohr’s doctrine. We show that C(A) is a complete Jordan
invariant for many algebras and synthesize various results along this line. Finally,
in concluding Sect. 5 we review recent results discovering new complete Jordan
invariant—the Choquet order structure on orthogonal measures on state spaces of
C ∗ -algebras. We show intimate connection of this structure with C(A) and describe
preservers of the Choquet order in terms of Jordan *-isomorphisms.
We shall now recall basic notions and introduce the notation. For all details on
operator algebras and their applications to physics the reader is advised to consult
monographs [2, 13, 14, 17]. Let A be a C ∗ -algebra. If it has the unit it is called unital.
We shall denote by As , and A+ the set of all self-adjoint and positive elements,
respectively. By the symbol Z(A) we shall represent the center of A. It is the set of
all elements commuting with every element of A. By the symbol P (A) we shall
denote the poset of projections in A, that is the set of self-adjoint idempotents
ordered by relation: p ≤ q if p = pq = qp. If the algebra A is unital with unit
1, then orthocomplement of projection p is the projection 1 − p. Two projections
p, q are called orthogonal, in symbols p ⊥ q, if pq = 0. The bijection between
projections posets is called orthoisomorphism if it preserves the orthogonality in
both directions. The central cover c(p) of projection p is the smallest projection
c(p) that lies in the center and is bigger than p. Projection is called faithful if its
central cover is 1. Finally, projection p is called abelian if pAp is an abelian C ∗ -
algebra. Two projections p and q are said to be equivalent if p = v ∗ v and q = vv ∗
for some element v ∈ A. In case of a unital C ∗ -algebra, the unitary element is
an element u whose inverse is u∗ . By the symbol B(H ) we shall denote the C ∗ -
algebra of all bounded operators acting on a Hilbert space H . If S ⊂ B(H ), then
S will represent the commutant of S, that is the set of all operators commuting
with all elements in S. A vector ξ ∈ H is called separating for S ⊂ B(H ) if
aξ = 0 implies a = 0 for all a ∈ S. Moreover, ξ is called biseparating for S
if it is separating both for S and S . Von Neumann algebra is a C ∗ -algebra that
has a predual. AW∗ -algebra A is a C ∗ -algebra that is Baer∗ ring, which means
that for any nonempty set S ⊂ A there is unique projection p ∈ A such that
S 0 = {x ∈ A : sx = 0 for all s ∈ S} = pA.
Let us have two C ∗ -algebras A and B. A linear map ϕ : A → B that preserves
the product and *-operation is called *-homomorphism. If it is a bijection we
call it *-isomorphism. A linear map ϕ : A → B that reverses the product (that
is ϕ(ab) = ϕ(b)ϕ(a) for all a, b ∈ A) and preserves *-operation is called *-
antihomomorphism. If it is a bijection we call it *-antiisomorphism. A linear map
ϕ : A → B that preserves the squares (ϕ(a 2) = ϕ(a)2 for all a ∈ A) and
*-operation is called a Jordan *-homomorphism. If it is a bijection we call it a
676 J. Hamhalter and E. Turilova
Jordan *-isomorphism. We shall now collect a few folklore results about Jordan
*-morphisms that will be used frequently in this work.
Proposition 1.1 Let us have C ∗ -algebras A and B. Let ϕ : A → B be a Jordan
*-homomorphism. Then the following holds:
(i) ϕ preserves the triple products:
The previous theorem has nontrivial extension to nearly all von Neumann
algebras. Let us recall that von Neumann algebra has Type I2 direct summand if
it has a direct summand *-isomorphic to the algebra C(X, M2 (C)) of all continuous
functions on a hyperstonean space X with values in the algebra of two by two
matrices.
Theorem 1.4 (Gleason-Bunce-Wright) Let M be a von Neumann algebra not
having Type I2 direct summand and X be a Banach space. Then any bounded
finitely additive measure μ : P (M) → X extends to a bounded linear functional
T : M → X.
We shall need the following n generalization of Jordan *-homomorphisms.
Definition 1.5 Let A and B be unital C ∗ -algebras. A map ψ : A → B is called
a quasi Jordan *-homomorphisms if the following conditions are satisfied for all
a, b ∈ A.
(i) ψ is homogeneous
(ii) ψ(a + b) = ψ(a) + ψ(b) whenever a and b commute.
(iii) ψ(a ∗ ) = ψ(a)∗ .
(iv) ψ(a 2 ) = ψ(a)2 .
When ψ is a bijection such that both ψ and ψ −1 are quasi Jordan *-
homomorphisms, we call ψ quasi Jordan *-isomorphism.
It is a consequence of Theorem 1.4 that any quasi Jordan *-homomorphism
defined on a von Neumann algebra without Type I2 direct summand is linear. This
does not hold for Type I2 von Neumann algebras (see [5]) for a more detailed
discussion).
2 Wigner Theorem
n
A= λi Pi λi ∈ R, Pi ∈ P1 (H ) . (1)
i=1
Put
n
ϕ̂(A) = λi ϕ(Pi ) .
i=1
The key argument of the proof is to show that the this definition is correct, that is,
not depending on the expression (1). To this end let us take another decomposition
m
A= μj Qj μi ∈ R, Qi ∈ P1 (H ) .
j =1
n
n
tr λi ϕ(Pi )ϕ(R) = λi tr(ϕ(Pi )ϕ(R)) = (2)
i=1 i=1
n
= λi tr(Pi R) = tr(AR) . (3)
i=1
n
m
T = λi ϕ(Pi ) S= μj ϕ(Qj ) .
i=1 j =1
Symmetries of C ∗ -algebras and Jordan Morphisms 679
tr (S − T )ϕ(R) = 0 .
n
A= λi Pi ,
i=1
n
n
ϕ̂(A2 ) = ϕ̂ λ2i Pi = λ2i ϕ(Pi ) = ϕ̂(A)2 .
i=1 i=1
10 00
Pe1 = Pe2 = .
00 01
680 J. Hamhalter and E. Turilova
As we know ϕ(Pe1 ) and ϕ(Pe2 ) are orthogonal rank one projections projecting on
linear span of unit orthogonal vectors f1 and f2 , respectively. Let us take a unitary
operator V satisfying conditions Vf1 = e1 and Vf2 = e2 . Then, for i = 1, 2,
ab ax
ϕ =
cd yd
00
A= ,
10
0y
ϕ(A) = ,
x0
xy 0
0=
0 xy
00 00
ϕ =
10 α0
Symmetries of C ∗ -algebras and Jordan Morphisms 681
ab a αb
ϕ = . (4)
cd αc d
01 0α
Since ϕ preserves spectra and sends matrix to , we conclude that
10 α0
|α| = 1. Let us now take a unitary map
U : (z1 , z2 ) ∈ C2 → (αz1 , z2 ) ,
that is
α0
U= .
01
ab a αb ab
U U∗ = =ϕ .
cd αc d cd
00 0α
ϕ =
10 00
for some α ∈ C. In the same way as before we can establish that α is a complex unit
and
ab a αc
ϕ = . (5)
cd αb d
U : (z1 , z2 ) ∈ C2 → (αz1 , z2 ) ∈ C2 .
ab
A= ,
cd
the action of U A∗ U ∗ on (z1 , z2 ) gives (az1 + αcz2 , αbz1 + dz2 ) which is precisely
the action of ϕ(A) on (z1 , z2 ). This complements the proof.
682 J. Hamhalter and E. Turilova
The algebra M2 (C) describes a two-level quantum system. Its state space has a
natural geometric interpretation. Let us identify each state with a positive matrix of
trace one (so called density matrix). Every density matrix T can be represented by
a vector r = (r1 , r2 , r3 ) in the three dimensional unit ball in the following way:
1
T = (I + r1 σ1 + r2 σ2 + r3 σ3 ) ,
2
where σ1 , σ2 , σ3 are Pauli spin matrices;
01 0 −i 1 0
σ1 = , σ2 = , σ3 = .
10 i 0 0 −1
Ad V (T ) = V T V ∗ , T ∈ B(H ) .
One can directly check that Pξ A∗ APξ = Aξ 2 Pξ . By this, we can continue the
previous computation and obtain
The foregoing identity allows us to define in a correct way the unitary map V :
Mn (C) → Mn (C) by
V Aξ = ϕ(A)ξ , A ∈ B(H) .
Symmetries of C ∗ -algebras and Jordan Morphisms 683
We shall prove that V implements ϕ. Any vector z ∈ Cn can be written in the form:
z = ϕ(B)ξ where B ∈ Mn (C). Then we have
V AV ∗ z = V AV ∗ ϕ(B)ξ =
As
⎛ ⎞ ⎛ ⎞
000 000
A2 = ⎝ 0 0 0 ⎠ , we have that α,β,γ (A2 ) = ⎝ 0 0 0 ⎠ .
100 β00
giving immediately β = αγ .
684 J. Hamhalter and E. Turilova
α,β,γ = AdU .
ij
Mn (C) := {Aij : A ∈ Mn (C)}
We have that A2 = 0. By Proposition 2.2 (and its proof) there are complex numbers
α, β of modulus one such that
⎛ ⎞
00β
ϕ(A) = ⎝ α 0 0 ⎠ .
000
Symmetries of C ∗ -algebras and Jordan Morphisms 685
Then
⎛ ⎞
00 0
0 = ϕ(A2 ) = ⎝ 0 0 αβ ⎠ ,
00 0
which is a contradiction. Suppose that M312 (C) has negative orientation and M313 (C)
has positive orientation. Then, similarly, there are complex numbers α and β of
modulus one such that
⎛ ⎞
0α0
ϕ(A) = ⎝ 0 0 0 ⎠ .
β 00
Then
⎛ ⎞
0 0 0
0 = ϕ(A2 ) = ⎝ 0 0 0 ⎠ ,
0 βα 0
which is a contradiction. Now we shall prove that M312 (C) and M323 (C) have the
same orientation. For this we consider action of ϕ on the matrix
⎛ ⎞ ⎛ ⎞
000 000
A = ⎝ 1 0 0 ⎠ and its square A2 = ⎝ 0 0 0 ⎠ .
010 100
Suppose that M312 (C) has positive and M323 (C) has negative orientation. Then there
are complex numbers α, γ of modulus one such that
⎛ ⎞
000
ϕ(A) = ⎝ α 0 γ ⎠ .
000
This way the result is established for n ≤ 3. Let us now tackle general case of
n ≥ 3. We shall show that one of the following statements are true
To this end let us fix a rank-one projection P . There exists a rank-one projection
Q that does not commute with P . The projection E = P ∨ Q is a rank-two
projection and the same holds for ϕ(E). As ϕ acts on EMn (C)E as a Jordan ∗-
isomorphism (that must preserve commutativity in both directions), we have that
ϕ(P )ϕ(Q) = ϕ(Q)ϕ(P ). Besides, we know from the case n = 2, that there are two
(mutually exclusive) possibilities: ϕ(P Q) = ϕ(P )ϕ(Q) and ϕ(P Q) = ϕ(Q)ϕ(P ).
Suppose the first one holds. Take any rank-one projection R that is not orthogonal
to P . Then the projection F = P ∨ Q ∨ R has rank at most three, and the same
holds for its image under ϕ. Based on our result for n ≤ 3, ϕ must act as ∗-
isomorphisms on F Mn (C)F . Therefore ϕ(P R) = ϕ(P )ϕ(R). As the same holds
for rank-one projections R orthogonal to P , we conclude that ϕ(P R) = ϕ(P )ϕ(R)
for all rank-one projections R. If ϕ(P Q) = ϕ(Q)ϕ(P ), then, similarly, we can show
that ϕ(P R) = ϕ(R)ϕ(P ) for all rank-one projections R.
In summary, we have proved that for any rank-one projection P one of the
following two statements is true.
We call P to have a positive orientation if (8) holds and negative orientation if (9)
holds. We shall show that (6) holds or (7) holds by demonstrating that all rank-one
projections have the same orientation. Suppose, for a contradiction, that there is
a rank-one projection P with positive orientation and rank-one projection Q with
negative orientation. As P = Q, we can find a rank-one projection R not commuting
either with P or Q. Then ϕ(RQ) = ϕ(QR)∗ = (ϕ(R)ϕ(Q))∗ = ϕ(Q)ϕ(R) =
ϕ(R)ϕ(Q). Therefore R has negative orientation. Considering now the pair P , R,
we have for the same reason that P has negative orientation—a contradiction.
Finally, as rank-one projections span the whole algebra, we can see that ϕ is either
∗-isomorphism (case (6)) or ∗-antiisomorphism (case (7)). Now Proposition 2.3
concludes the proof.
The foregoing results on the structure of Jordan *-isomorphisms allows us to
establish Wigner theorem both for finite and infinite dimensional Hilbert space.
Theorem 2.6 (Wigner Theorem for Finite Quantum Systems) Let ϕ : P1 (H ) →
P1 (H ), where H is a finite dimensional Hilbert space, be a map preserving
Symmetries of C ∗ -algebras and Jordan Morphisms 687
The infinite-dimensional variant of the previous theorem can be found in [1].
Proof The crucial idea is to find a bounded linear extension of ϕ. For this consider
two identical linear combinations of projections in B(H ), i.e.
n
m
λi Pi = μj Qj ,
i=1 j =1
to B(H ) by Theorem 1.3. By applying this extension to the equality above we obtain
n
m
f λi ϕ(Pi ) = f μj ϕ(Qj ) .
i=1 j =1
n
m
λi ϕ(Pi ) = μj ϕ(Qj ) .
i=1 j =1
n
n
T λi Pi = λi ϕ(Pi ) .
i=1 i=1
It can be verified easily that T is bounded and so can be extended to a bounded linear
operator (denoted by the same letter) T : B(H ) → B(H ). As this map preserves
projection, it has to be Jordan *-homomorphism by Proposition 1.2. Arguing in the
same way for the inverse map ϕ −1 gives us that T is a Jordan *-isomorphism. By
the previous discussion (see e.g. Theorem 2.5 for a finite dimensional case) this map
is implemented by unitary or antiunitary operator.
Theorem 2.8 (Uhlhorn) Let H be a Hilbert space with dimH ≥ 3. Let ϕ :
P (H ) → P (H ) be an orthoisomorphism. Then there is either a unitary or an
antiunitary operator U acting on H implementing ϕ.
Proof It is enough to show that ϕ extends to an orthoisomorphism of P (H ). Let us
take any projection P acting on B(H ). Suppose that
P = sup Qα = sup Rβ
because both these suprema have the same orthocomplement. This allows to extend
ϕ to an orthoisomorphism on P (H ) by putting
3 Dye Theorem
and
ϕ(1 − p) = 1 − ϕ(p) .
(1 + aa ∗)−1 (1 + aa ∗)−1 a
.
a ∗ (1 + aa ∗)−1 a ∗ (1 + aa ∗)−1 a
Further by eii we shall denote the matrix having at position (i, i) the unit and zeros
elsewhere. It turns out that each pij (a) is a projection and even that projections
of this form generate the projection lattice of Mn (A) as a complete orthomodular
lattice (see e.g. [12, Lemma 4.1]).
There is an elegant bridge between linear structure of A and lattice operations in
Mn (A), which is a great discovery going back to J. von Neumann (see [4, Lemma 4,
Lemma 3(i)]). Let us recall that by a lattice polynomial in variables p1 , . . . , pk we
mean a formal expressions that is a result of finitely many lattice operations (∨, ∧)
performed on elements in {p1 , . . . , pk }.
Lemma 3.3 There exist lattice polynomials P , Q and R such that for any elements
a, b, c of a C ∗ -algebra A, where c is invertible, any integer n ≥ 3, and any distinct
indices 1 ≤ i, j, k ≤ n, the following holds.
(i) pij (a + b) = P (pij (a), pij (b), pik (c), eii , ejj , ekk ).
(ii) pij (−ab) = Q(pik (a), pkj (b), eii , ejj )
(iii) pij (−a ∗ ) = R(pj i (a), eii , ejj ).
As a consequence of Lemma 3.3 any lattice morphism of Mn (A) that preserves
projections of type pij (a) and eii , induces a *-ring morphism on the underlying
algebra A. This is an important ingredient in proving Dye theorem as well in
establishing the following deep result of Heunen and Reyes [12, Theorem 4.6].
Theorem 3.4 (Heunen and Reyes) Let A and B be AW ∗ -algebras. Let
f : P (A) → P (B)
direct sum
C ⊕ M2 (D) ,
μ : P (A) → X ,
T : A → X.
μ : P (B) → X ,
T : B → X.
Proof of Theorem 3.2 We have now all ingredients to prove the (nonbijective) Dye
theorem for AW∗ -algebras. Let us have an AW∗ -algebra A, not having Type I2
direct summand, and another AW∗ -algebra B. Consider a map ϕ : P (A) → P (B)
that preserves suprema of arbitrary projections and orthocomplements. Fix now two
projections e, f ∈ A. Theorem 3.7 assures us that there is a Jordan map J mapping
the algebra C = AW ∗ (1, e, f ) to B that coincides with ϕ on P (C). By the algebraic
properties of the Jordan morphism we can conclude that (10) is satisfied. Now we
can use Theorem 3.4 to prove Theorem 3.2.
692 J. Hamhalter and E. Turilova
Let A be a unital C ∗ -algebra. By the symbol C(A) we shall denote the structure of
all abelian C ∗ -subalgebras of A containing the unit of A. When endowed with set
theoretic inclusion, C(A) becomes the poset with the least element span{1}. Infima
in this poset are given by set theoretic intersections of subalgebras. On the other
hand, supremum of two elements E and F exists in C(A) if and only if E and F
mutually commute. Similarly we shall denote by C0 (A) the poset of all unital abelian
C ∗ -subalgebras (not necessarily containing the unit of A). Let P be a subposet of
C0 (A) and Q be a subposet of C0 (B). The map ϕ : P → Q is said to be implemented
by a map : A → B if
ϕ : C(A) → C(B)
ψ :A→B
such that
This result has a topological background. In fact, any unital abelian C ∗ -algebra
is *-isomorphic to C(X), where X is a compact Hausdorff space. Moreover, by
Banach-Stone theorem any *-isomorphism ψ : C(X) → C(Y ) is given by the
homeomorphism τ : Y → X in the following way
There are interrelations between algebraic structure of abelian C ∗ -algebra C(X) and
topology of X. For example, for any closed ideal I of C(X) there is a unique closed
subset F of X such that
I = {f ∈ C(X) : f is zero on F}
The key role in studying the poset C(C(X)) is played by so called ideal subalgebras,
that is by C ∗ -subalgebras of C(X) generated by a closed proper ideal I of C(X) and
the unit 1. More precisely, for each proper ideal algebra C of C(X) there is a closed
subset F of X with at least two points such that
C = {f ∈ C(X) : f is constant on F } .
It can be shown that any order isomorphism of subalgebras structures preserves ideal
subalgebras. Therefore it induces an order isomorphism between closed subsets of
X. An important part of proving Theorem 4.1 is therefore establishing the form
of isomorphisms of the poset of closed subsets. This is achieved in the following
theorem (see [6, Theorem 2.3]). Let us denote by the symbol F (X) the poset of all
closed subsets of X with at least two points ordered by set theoretic inclusion.
Theorem 4.2 Let X and Y be compact Hausdorff spaces. Suppose that X is not a
singleton. Let
ψ : F (X) → F (Y )
τ :X→Y
such that
is an order isomorphisms.
There is a natural question whether the converse holds as well. The answer is
in the positive. It was proved by the first author in [6, Theorem 3.4]. For further
treatment and alternative proofs we recommend [14, 15].
Theorem 4.4 (Hamhalter) Let A and B be unital C ∗ -algebras. Then for any order
isomorphism
ϕ : C(A ) → C(B)
ψ :A→B
such that
Except for the poset C(A), it is also natural to consider the poset C0 (A) of
all abelian C ∗ -subalgebras (unital or not) of A. The poset C0 (A) has different
properties. For example, let us consider A = C2 . The poset C(A) consists of two
elements span {(1, 1)} and A, related by span {(1, 1)} ≤ A. In case of C0 (A) we have
the largest element A, the smallest element {0} and three incomparable elements
span {(1, 1)}, span {(1, 0)}, and span {(0, 1)}. The group of order isomorphisms is
just the group of permutations of these three point set. As Jordan *-isomorphisms
are unital, they implement only those order automorphisms of C0 (A) that leave the
element span{(1, 1)} fixed. Therefore one cannot hope that order isomorphisms are
implemented by (quasi) Jordan automorphisms in this case. However careful anal-
ysis shows that implementation is possible if we assume that order isomorphisms
preserve the subalgebra generated by the unit. Given a unital C ∗ -algebra A we shall
denote by O the one dimensional subalgebra span{1}.
The authors have obtained the following result [8]:
Theorem 4.5 (Hamhalter and Turilova) Let A and B be unital C ∗ -algebras
isomorphic neither to C ⊕ C nor to M2 (C). Let
ϕ : C0 (A) → C0 (B)
ϕ(O) = O .
ψ :A→B
such that
Vξ = span{Pξ , 1 − Pξ }} ,
ϕ : C(M) → C(N )
ψ :M→N
such that
Based on Theorem 4.5 we can now obtain the nonunital version of Theorem 4.6
proved in [8].
Theorem 4.7 (Hamhalter and Turilova) Let M be a von Neumann algebra
without Type I2 direct summand. Let N be another von Neumann algebra. Let
ϕ : C0 (M) → C0 (N )
ϕ(O) = O .
ψ :M→N
such that
ϕ : V(M) → V(N)
ψ :M→N
implementing ϕ:
Original proof of this result is based on Dye Theorem. It can be also proved
quickly by using Generalised Gleason theorem as in the previous subsection. We
have generalized this result to AW∗ -algebras. Since we do not have Gleason type
theorem to our disposal in this case we have to rely fully on Dye theorem for AW∗ -
algebras discussed in Sect. 2. The following result may be found in [7, Theorem 4.6]
for C(M) or in [15, Theorem 9.2.8] for V(N ). Having an AW∗ -algebra M we shall
denote by V(M) the poset of abelian AW∗ -subalgebras of M containing the unit of
M.
Theorem 4.9 (Hamhalter, Lindenhovous) Let M be an AW∗ -algebra without
Type I2 direct summand and N be another AW∗ -algebra.
(i) Let
ϕ : V(M) → V(N )
ψ :M→N
698 J. Hamhalter and E. Turilova
implementing ϕ:
(ii) Let
ϕ : C(M) → C(N )
ψ :M→N
implementing ϕ:
Besides the order structure of all abelian C ∗ -subalgebras and abelian von Neu-
mann subalgebras we can also consider the simplest structure of finite dimensional
abelian subalgebras. Each algebra of this type is isomorphic to the power Cn and
corresponds to decomposition of the unit into sum of orthogonal projections. Let
Cf in (A) be the set of all finite dimensional abelian C ∗ -subalgebras of a unital
C ∗ -algebra A containing the unit and ordered by set theoretic inclusion. Using
Dye theorem one can show once more that order isomorphism of this structure is
implemented by Jordan *-isomorphism (see [9, Proposition 3.5]).
Theorem 4.10 (Hamhalter and Turilova) Let M be a von Neumann algebra
without Type I2 direct summand. Let N be another von Neumann algebra. Let
ϕ : Cf in (M) → Cf in (A)
In the previous section we could see that for algebras not containing Type I2 part, the
structure of abelian subalgebras implies that given algebras are isomorphic as Jordan
algebras. For algebras of Type I2 we know that not all order isomorphisms between
the structure of abelian subalgebras are implemented by Jordan maps. However,
it is surprising that such algebras are even *-isomorphic if they have isomorphic
posets of abelian subalgebras. In fact, we have shown that C(A) is a complete Jordan
invariant for all von Neumann algebras (see Theorem 2.3 in [11].)
Symmetries of C ∗ -algebras and Jordan Morphisms 699
Theorem 4.11 (Hamhalter and Turilova) Let M and N be von Neumann alge-
bras. The following assertions are equivalent
(i) V(M) and V(N ) are isomorphic.
(ii) Cf in (M) and Cf in (N ) are isomorphic.
(iii) P(M) and P(N ) are orthoisomorphic.
(iv) M and N are isomorphic as Jordan algebras.
Proof First observe that (i) implies (ii). This is due to the fact that finite dimensional
abelian subalgebras can be characterized as those elements in V(M) that have
only finitely many elements beneath. This is of course preserved by any order
isomorphism. Therefore any order isomorphism between V(M) and V(N ) restricts
to an order isomorphism between Cf in (M) and Cf in (N ).
Now we focus on proving that (ii) implies (iii). It has been proved in [3,
Lemma 3.1] that Cf in (M) is isomorphic to the poset of finite Boolean subalgebras
B f in (P(M)) of P(M) by the map
N = wN ⊕ (1 − w)N .
P(Z(zM)) and P(Z(wN )). Therefore Z(zM) and Z(wN ) are isomorphic as C ∗ -
algebras (it follows from the fact than any orthoisomorphism between projection
lattices of abelian von Neumann algebras extends to a *-isomorphism). According
to the structure theory of finite homogeneous von Neumann algebras of Type In
such algebras are ∗-isomorphic if they have ∗-isomorphic centers. This concludes
the proof of given implication.
Finally, the implication (iv) ⇒ (i) is easy.
In this section we shall present our main results on complete Jordan invariants based
on Choquet order structure of decompositions of states. Let us first recall basic
definitions. For details on Choquet theory on state spaces we refer the reader to
monograph [17]. Let us have a compact Hausdorff space X. By a Radon measure
μ on X we mean an element of the dual space C(X)∗ . By celebrated Riesz
representation theorem there is a bijection between Radon measures and regular
Borel measures on X. In this correspondence μ ∈ C(X)∗ can be canonically
identified with a regular Borel measure μ on X in the sense of the formula
$
μ(f ) = f (ω) dμ(ω) f ∈ C(X) .
X
It is known that the relation ≺C is a partial order on the set of positive Radon
measures (see e.g. [2, Proposition 4.1.3, p.325], [17, Definition 6.5,p. 233]). The
order ≺C is called the Choquet order.
Now we turn to the situation when K is a state space of a C ∗ -algebra and establish
some new results for Choquet order in this situation. Let ϕ be a state on a C ∗ -algebra
A. The triple (πϕ , ξϕ , Hϕ ) will represent the GNS data of ϕ. By Mϕ we shall denote
the von Neumann algebra generated by πϕ (A). Then Mϕ = πϕ (A) . Let Cϕ be the
(real) space of all functionals in A∗ spanned by positive functionals dominated by
ϕ. In other words,
Cϕ = span {ψ : 0 ≤ ψ ≤ ϕ} .
It is well known that there is a bijective positive map between Cϕ and πϕ (A) ,
sending each element ψ ∈ Cϕ to an operator aψ ∈ Mϕ such that, for each a ∈ A,
(see e.g. [17, Proposition IV 3.10, p. 201]). Let μ ∈ Mϕ+ (S(A)). Take f ∈
L∞ (S(A), μ). Then, according to the previous discussion, there is a unique element,
θμ (f ) ∈ Mϕ such that, for each a ∈ A,
$
θμ (f )πϕ (a)ξϕ , ξϕ = f (ω)a(ω) dμ(ω) .
S(A)
The map θμ is a unital weak∗ to weak∗ continuous map from L∞ (S(A), μ) into
von Neumann algebra Mϕ (see e.g. [17, Proposition 6.18, p. 238]). The measure
μ ∈ Mϕ+ (S(A)) is called orthogonal if, for each Borel set E ⊂ S(A), the positive
functionals ϕE and ϕE c on A given by
$ $
ϕE (a) = a(ω) dμ(ω) ϕS(A)\E (a) = a(ω) dμ(ω)
E S(A)\E
Let us denote by Oϕ (A) the set of all orthogonal measures having barycenter ϕ.
f in
Of course, δx ∈ Oϕ (A). Let us denote by Oϕ (A) the set of all finitely supported
orthogonal measures in Oϕ (A).
As an example, let us look at Oϕ (M2 (C)). The state space of M2 (C) can be
identified with all matrices of the form
1 1 + β1 β2 + iβ3
,
2 β2 − iβ3 1 − β1
where (β1 , β2 , β3 ) is a point in the three dimensional unit ball. (See also example
after Proposition 2.2.) In this way the state space is affine isomorphic to the unit
three dimensional ball. For simplicity, let ϕ be the normalized trace. It corresponds
to the origin. Any orthogonal measure on the state space that has barycenter
ϕ is a convex combination of two Dirac measures concentrated at vector states
(pure states) corresponding to orthogonal unit vectors in C2 . When using the ball
representation of the state space, we can see that these orthogonal states correspond
to antipodal points on the unit sphere. The set Oϕ (M2 (C))) can be then viewed as a
set of measures on the unit ball that are concentrated at two antipodal points on the
unit sphere and that assign mass 1/2 to each of them.
Going back to general situation, we shall denote by ϕ the map
ϕ : Oϕ (A) → V(Mϕ ) : μ → Cμ .
One of the basic theorems we shall use in this note is the following Tomita
theorem (see e.g. [17, Prop. 6.23, p. 241, Theorem 6.25 p.244]). It establishes a
one-to-one correspondence between orthogonal measures and abelian subalgebras
that preserves the Choquet order.
Theorem 5.1 (Tomita Theorem) The map ϕ : μ → Cμ is a bijection of
Oϕ (A) onto V(Mϕ ). Moreover, the following conditions are equivalent for μ, ν ∈
Oϕ (A):
(i) μ ≺ ν
(ii) Cμ ⊂ Cν .
In particular, the posets (Oϕ (A), ≺) and (V(Mϕ ), ⊂) are order isomorphic.
In [9] we showed that discrete version of Tomita’s theorem holds for finitely
supported measures as well.
Theorem 5.2 (Hamhalter and Turilova) The map ϕ : μ → Cμ is an order
isomorphism of Oϕ (A) onto Cf in (Mϕ ).
f in
F (μ) = −1
ψ J [ϕ (μ)] .
F (μ) = −1
ψ J [ϕ (μ)]
f in
for each μ ∈ Oϕ (A).
Proof Let us prove (i), the statement (ii) can be proved analogously. By Theo-
rem 5.1 we know that order isomorphism F induces order isomorphism between
V(Mϕ ) and V(Mψ ). Suppose that Mϕ has no Type I2 direct summand. By
employing Theorem 4.8, we see that this isomorphism is implemented by a Jordan
*-isomorphism J : Mϕ → Mψ . This shows the form of F . By Theorem 9.1.3 in
[13] if a von Neumann algebra is of type I , then the same holds for its commutant.
Therefore, if Mϕ has no nonzero Type I direct summand, then Mϕ has no Type I2
direct summand and we apply the previous reasoning.
The assumption on Type I2 direct summand in the previous theorem is essential
as the following example demonstrates.
Example Let A = M2 (C) and let ϕ be a faithful state on A. Then there is a Choquet
order isomorphism on Oϕ (A) that is not induced by any Jordan ∗-isomorphism on
Mϕ .
Proof We can write ϕ = λ1 ωe1 + λ2 ωe2 , where e1 , e2 is a standard orthonormal
basis of C2 and λ1 and λ2 are positive non-zero numbers with sum one. It can be
verified easily that GNS data reads as follows:
H ϕ = C2 ⊕ C2 ,
1/2 1/2
ξϕ = λ1 e1 ⊕ λ2 e2 ,
704 J. Hamhalter and E. Turilova
a0
πϕ (a) = .
0a
αI βI
,
γ I δI
j (x) = J x ∗ J x∈M
Acknowledgment This work was supported by the project OPVVV CAAS CZ.02.1.01/0.0/0.0/
16_019/0000778
References
1. J. Barvínek, J. Hamhalter, Linear algebraic proof of Wigner Theorem and its consequences.
Math. Slovaca (2), 67 (2017)
2. O. Brateli, D.W. Robinson, Operator Algebras and Quantum Statistical Mechanics, vol. 1
(Springer, Berlin, 1997)
3. A. Döring, J. Harding, Abelian subalgebras and the Jordan structure of von Neumann algebras.
Houston J. Math. (2) 42, 559–568 (2016)
4. H.A. Dye, On the geometry of projections in certain operator algebras. Ann. Math. (1) 61,
73–89 (1955)
5. J. Hamhalter, Quantum Measure Theory (Kluwer Academic, Boston, 2003)
6. J. Hamhalter, Isomorphisms of ordered structures of abelian C ∗ -subalgebras of C ∗ -algebras. J.
Math. Anal. Appl. 383, 391–399 (2011)
7. J. Hamhalter, Dye’s Theorem and Gleason’s Theorem for AW*-algebras. J. Math. Anal. Appl.
422, 1103–1115 (2015)
8. J. Hamhalter, E. Turilova, Automorphisms of ordered structures of abelian parts of operator
algebras and their role in quantum theory. Int. J. Theor. Phys. (10) 53, 3333–3345 (2014)
9. J. Hamhalter, E. Turilova, Orthogonal measures on state spaces and context structures of
quantum theory. Int. J. Theor. Phys. 55, 3353–3365 (2016)
10. J. Hamhalter, E. Turilova, Choquet order and Jordan maps. Lobachevskii J. Math. (3) 39,
340–347 (2018)
11. J. Hamhalter, E. Turilova, Jordan invariants of von Neumann algebras given by abelian
subalgebras and Choquet order on state spaces. Int. J. Theor. Phys. (2) 60, 1–11 (2021)
12. C. Heunen, M. Reyes, Active lattices determine AW*-algebras. J. Math. Anal. Appl. 416, 289–
313 (2014)
13. R.V. Kadison, J.R. Ringrose, Theory of Operator Alegebras I, II (Academic, New York, 1986)
14. K. Landsman, Foundations of Quantum Theory, From Classical Concepts to Operator
Algebras. Springer Open, Fundamental Theories of Physics, vol. 188 (Springer, Cham, 2017)
15. B. Lindenhovius, C (A). Ph.D. thesis, Radbound University, Nijmegen, 2016
16. R. Simon, N. Mukunda, S. Chaturvedi, V. Srinivasan, J. Hamhalter, Comment on: Two
elementary proofs of the Wigner Theorem on symmetry in quantum mechanics [Phys. Letter
A 327 (2008) 6847]. Phys. Lett. A 378, 2332–2335 (2014)
17. M. Takesaki, Theory of Operator Algebras I, II, III (Springer, Berlin, 2001)
18. E.P. Wigner, Group Theory and Its Application to the Quantum Mechanics of Atomic Spectra
(Academic, New York, 1959)
Part V
Inequalities in Commutative and
Noncommutative Probability Spaces
Mixed Norm Martingale Hardy Spaces
and Applications in Fourier Analysis
Ferenc Weisz
Abstract We consider martingale Hardy spaces defined with the help of mixed
Lp -norm. Five mixed normed martingale Hardy spaces will be investigated: Hps ,
HpS , HpM
, Pp and Qp . We give two different generalizations of Doob’s maximal
inequality for mixed-norm Lp spaces. We prove also two versions of atomic
decompositions. Several martingale inequalities and the generalization of the well-
known Burkholder-Davis-Gundy inequality are also presented. The dual spaces of
the mixed-norm martingale Hardy spaces are given as the mixed-norm BMOr ( α)
spaces. This implies the John-Nirenberg inequality BMO1 ( α ) ∼ BMOr (α ) for
1 < r < ∞. As an application in Fourier-analysis, we verify the boundedness of the
Fejér maximal operator from Hp to Lp , whenever 1/2 < p < ∞. As a consequence
of the boundedness, we get some almost everywhere and norm convergence results.
1 Introduction
Since 1970, the theory of Hardy spaces has been developed very quickly (see
e.g. Fefferman and Stein [19], Stein [81], Grafakos [34]). Fefferman [18] proved
that the dual space of the Hardy space is equivalent to the space of functions of
bounded mean oscillation (BMO). John and Nirenberg [55] obtained their famous
inequality, i.e., that the BMOp spaces are equivalent. One year later, Fefferman
and Stein [19] characterized the dual space of Hp (0 < p < 1) as a Lipschitz
space. The most powerful technique in the theory of Hardy spaces, the so-called
F. Weisz ()
Department of Numerical Analysis, Eötvös L. University, Budapest, Hungary
e-mail: [email protected]
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 709
R. M. Aron et al. (eds.), Operator and Norm Inequalities and Related Topics,
Trends in Mathematics, https://doi.org/10.1007/978-3-031-02104-6_21
710 F. Weisz
atomic decomposition was given in Coifman and Weiss [14, 15]. Recently several
papers were published about the generalization of Hardy spaces. For example,
Hardy spaces with variable exponents were considered in Nakai and Sawano [68],
Yan, et al. [99], Jiao et al. [54], Liu et al. [63] and [64]. Moreover Musielak-Orlicz-
Hardy spaces were studied in Yang et al. [100]. The mixed norm classical Hardy
spaces have been developed in Cleanthous et al. [11] and intensively studied by
Huang et al. in [41–44, 46].
Parallel, a similar theory was evolved for different types of martingale Hardy
spaces Hps , HpS , HpM , Pp and Qp (see e.g. Garsia [23], Long [65] and Weisz [87]).
In the celebrated work of Burkholder and Gundy [6], it was proved that the Lp
norms of the maximal function and the quadratic variation, that is the spaces HpM
and HpS , are equivalent for 1 < p < ∞. In the same year, Davis [16] extended this
result for p = 1. For martingale Hardy spaces, Weisz [87] worked out the theory
of atomic decomposition. Some boundedness results, duality theorems, martingale
inequalities and interpolation results can be proved with the help of the atomic
decomposition. A martingale analogue of H1 -BMO duality can be found in the
books Garsia [23], Long [65] and Weisz [87]. For dyadic martingales, Herz [37]
obtained the dual space of Hp (0 < p < 1). In 1990, Weisz [86] characterized the
dual space of Hp (0 < p < 1) for general martingales via atomic decomposition.
For a regular stochastic basis, the BMOp spaces are equivalent in the martingale
case, too. Recently, these results were extended to more general cases. Jiao et al.
investigated martingale Hardy-Lorentz spaces in [51, 52] and variable martingale
Hardy spaces in [49, 50, 53]. Martingale Musielak–Orlicz Hardy spaces were
investigated in Xie et al. [96–98]. The theory of martingale Hardy spaces can be well
applied in Fourier analysis (see Gát [25, 26], Goginava [30, 31] or Weisz [87, 90]).
The mixed Lebesgue spaces were introduced in 1961 by Benedek and Panzone
[3] (see also Hörmander [40]). They considered the Descartes product (, F , P)
of the probability spaces (i , F i , Pi ), where = di=1 i , F is generated by
d d
i=1 F and P is generated by i=1 P . The mixed Lp -norm of the measurable
i i
function f is defined as a number obtained after taking successively the Lp1 -norm
of f in the variable x1 , the Lp2 -norm in the variable x2 , . . ., the Lpd -norm in the
variable xd . Some basic properties of the spaces Lp were proved in [3], such as the
well known Hölder’s inequality and the duality theorem. Mixed-norm Lebesgue and
Hardy spaces were investigated in a great number of papers (e.g. in [1, 9–13, 27, 28,
35, 38, 39, 41–44, 46, 48, 56–60, 80]).
In this paper we will introduce five mixed normed martingale Hardy spaces: Hps ,
HpS , HpM
, Pp and Qp . In Sect. 3, Doob’s inequality will be proved, that is, we will
show that
sup |En f | ≤ C f p
n∈N p
for all f ∈ Lp , where 1 < p < ∞. We present also another version of Doob’s
inequality. In Sect. 5, we give the atomic decomposition for the five mixed normed
Mixed Martingale Hardy Spaces and Applications in Fourier Analysis 711
martingale Hardy spaces. Using the atomic decomposition and Doob’s inequality,
several martingale inequalities will be proved in Sect. 6. We will show that, if
the stochastic basis (Fn ) is regular, then the five martingale Hardy spaces are
equivalent. As a consequence of Doob’s inequality, the generalization of the well-
known Burkholder-Davis-Gundy inequality can be shown. In the next section, we
prove that the dual of HpM is BMO2 ( α ) and, if the stochastic basis is regular, then
M
the dual of Hp is BMOr ( α ), where 0 < p ≤ 1, α = 1/p − 1 and 1 < r < ∞.
Consequently, we obtain the generalization of the John and Nirenberg theorem
for mixed normed martingale spaces: if 0 ≤ α < ∞ and 1 < r < ∞, then
BMO1 ( α ) = BMOr (α ) with equivalent norms.
In the one-dimensional case, Paley [71] (see also Schipp, Wade, and Simon [75]
and Weisz [90]) proved the Lp -norm convergence of the partial sums of the Walsh-
Fourier series of f in case of 1 < p < ∞. There is no convergence result for
p ≤ 1 (see [33, 75]). Using summability methods, such as Fejér means, the L1 -
norm convergence can be reached for functions in L1 , too. It was proved by Fine
[21] that the Fejér means of the one-dimensional Walsh-Fourier series converges
almost everywhere to the function if f ∈ L1 . Schipp [72] obtained the same result
by proving the weak type inequality of the maximal operator σ∗ of the Fejér means.
By interpolation, this implies that σ∗ is bounded on Lp (1 < p < ∞). Next Fujii
[22] extended this and showed that σ∗ is bounded from the dyadic Hardy space H1
to L1 (see also Schipp and Simon [74]). Later the author (see [88]) generalized this
further and proved that σ∗ is bounded from Hp to Lp for 1/2 < p < ∞. The
boundedness does not hold for 0 < p ≤ 1/2 (see [78]).
In the two-dimensional case, Weisz considered the Fejér maximal operator over
a cone and he proved in [89] that σ∗ is bounded from Hp to Lp for 1/2 < p <
∞. Gát [24] and Weisz [89] proved that the Fejér means of the two-dimensional
Walsh-Fourier series converge to the function almost everywhere if we consider
the convergence over the diagonal, or more generally, over a cone. This result was
proved for trigonometric Fourier series by Marcinkievicz and Zygmund [67] and
Weisz [92]. Similar results were obtained in numerous other papers (see, e.g., Gát
[25, 26] and Goginava [29–31]).
In this paper, we generalize the previous results for mixed normed martingale
Hardy spaces. We will prove that the Fejér maximal operator defined over a cone
is bounded from Hp to Lp (1/2 < p < ∞). As a consequence, we get some
convergence results, such as almost everywhere and norm convergence of the
multi-dimensional Fejér means defined over a cone. This result generalizes the well-
known theorem of Gát [24] and Weisz [89]. Some summability results for classical
mixed norm Hardy spaces Hp (Rd ) and for Fourier transforms can be found in
[45, 93].
We denote by C a positive constant, which can vary from line to line, and denote
by Cp a constant depending only on p. The symbol A ∼ B means that there exist
constants α, β > 0 such that αA ≤ B ≤ βA and A B means that there exist
C > 0 such that A ≤ CB.
712 F. Weisz
is finite, with the usual modification if pj = ∞ for some j ∈ {1, . . . , d}. If for
some 0 < p ≤ ∞, p = (p, . . . , p), then we get back the classical Lebesgue space
Lp . Under r < p ≤ q, we mean that for all i = 1, . . . , d, r < pi ≤ q, where
0 ≤ r < q ≤ ∞. For a vector p, we will use the notations
p− := min {p1 , . . . , pd } .
The
conjugate exponent vector of p will be denoted by (p) , that is, (p)
=
p1 , . . . , pd , where 1/pi + 1/pi = 1 (i = 1, . . . , d). For α > 0, p/α :=
(p1 /α, . . . , pd /α). Benedek and Panzone [3] proved the next two basic results for
the mixed Lebesgue space.
Theorem 2.1 If 1 ≤ p ≤ ∞, then for all f ∈ Lp and g ∈ L(p)
,
$
|fg| dP ≤ f p g(p)
.
Moreover,
$
f p = sup fg dP .
g(p)
≤1
Similarly to the Lebesgue spaces, the following result holds for the dual of Lp .
Mixed Martingale Hardy Spaces and Applications in Fourier Analysis 713
3 Doob’s Inequality
fn ≤ Rfn−1 .
then f is an Lp -bounded martingale, briefly f ∈ Lp . We define the Doob’s maximal
function by
Of course,
M(f ) ≤ Md ◦ Md−1 ◦ . . . ◦ M1 (f ),
Mi (f ) := sup Ein f .
n∈N
M(f )p ≤ Cp f p .
For 1 < p < ∞ and for the classical Hardy-Littlewood inequality, Theorem 3.4
was shown in Bagby [2]. This theorem implies easily the next generalization of
Doob’s inequality, that is to say, the maximal operator M is bounded on Lp in case
1 < p < ∞ (see [84]).
Theorem 3.5 Under the same conditions as in Theorem 3.4, for all f ∈ Lp ,
L(p1 ,∞) (cf. Lemma 3.5 in [41] and Lemma 4.8 in [69]). A weighted version of
Doob’s inequality can be found in Chen et al. [9].
We can easily modify the definition of the maximal operator. For a constant q
and f ∈ Lq , let
1/q
Mq (f ) := sup En (|f |q ) .
n∈N
Mq (f )p ≤ Cp f p .
where 0 < q < ∞. Now, we can show that under some conditions, this operator is
bounded on Lp , too.
Theorem 3.7 ([94]) Let 0 < q < ∞ and 0 < p < ∞ or
This theorem was proved in [45] for the classical maximal function (for a part of
this theorem see also [70]).
716 F. Weisz
For n ∈ N and a martingale f = (fn )n∈N , the martingale differences are defined by
dn f := fn − fn−1 , f0 := f−1 := 0.
The map ν : → N ∪ {∞} is called a stopping time relative to (Fn ) if for all
n ∈ N, {ν = n} ∈ Fn . For a martingale f = (fn ) and a stopping time ν, the stopped
martingale is defined by
n
fnν = dm f χ{ν≥m} .
m=0
Let us define the quadratic variation and the conditional quadratic variation of a
martingale f relative to (, F , P, (Fn )n∈N ) by
m 1/2 ∞ 1/2
Sm (f ) := |dn f | 2
, S (f ) := |dn f | 2
n=0 n=0
m 1/2 ∞
1/2
sm (f ) := En−1 |dn f | 2
, s (f ) := En−1 |dn f |
2
.
n=0 n=0
The set of the sequences (λn )n∈N of non-decreasing, non-negative and adapted
functions with λ∞ := limn→∞ λn is denoted by $. With the help of the previous
operators, we introduce five mixed normed martingale Hardy as follows:
4 5
HpM
:= f = (fn )n∈N : f H M := M(f )p < ∞ ;
p
4 5
HpS := f = (fn )n∈N : f H S := S(f )p < ∞ ;
p
4 5
Hps := f = (fn )n∈N : f H s := s(f )p < ∞ ;
p
6
Qp := f = (fn )n∈N : ∃ (λn )n∈N ∈ $,
7
such that Sn (f ) ≤ λn−1 , λ∞ ∈ Lp ,
6
Pp := f = (fn )n∈N : ∃ (λn )n∈N ∈ $,
7
such that |fn | ≤ λn−1 , λ∞ ∈ Lp .
Define
For a constant p, we get back the well known martingale Hardy spaces HpM , HpS ,
Hps , Qp and Pp investigated exhaustively in [87].
The following corollary comes from Theorem 3.5. It is well-known for martin-
gale Hardy spaces with p = (p, . . . , p) (see e.g. [87]).
Corollary 4.1 If 1 < p < ∞ or p satisfies (1), then HpM
is equivalent to Lp .
5 Atomic Decomposition
If s(a) in (ii) is replaced by S(a) (resp. M(a)), then the function a is called
∞)-atom (resp. (M, p,
(S, p, ∞)-atom).
If p is a constant, then (ii) reads as follows:
s(a)χ{τ <∞} ≤ P(τ < ∞)−1/p .
∞
Every function from the Hardy space Hps (0 < p ≤ 1) can be decomposed into the
sum of atoms.
Theorem 5.2 ([87]) Let p be a constant with 0 < p ≤ 1. A martingale f =
(fn )n∈N ∈ Hps if and only if there exist a sequence (a k )k∈Z of (s, p, ∞)-atoms and
a sequence (μk )k∈Z of real numbers such that
fn = μk En a k a. e. (n ∈ N)
k∈Z
and
1/p
p
f Hps ∼ inf μk , (2)
k∈N
In the present form the theorem does not hold for 1 < p < ∞ and it cannot be
extended to mixed norm Hardy spaces. It is easy to see that for 0 < p ≤ 1, (2) can
be written as
1/p
p p 1/p
μ χ
k {τk <∞}
f Hps ∼ inf μk = inf
.
k∈N k∈N χ{τk <∞} p
p
Writing the p-norm instead of the p-norm, we can generalize this form of the atomic
decomposition to mixed norm Hardy spaces (see [84]).
Theorem 5.3 Let 0 < p < ∞. A martingale f = (fn )n∈N ∈ Hps if and only if
∞)-atoms and a sequence (μk )k∈Z of real
there exist a sequence (a k )k∈Z of (s, p,
numbers such that
fn = μk En a k a. e. (n ∈ N) (3)
k∈Z
and
t 1/t
μk χ{τk <∞}
f H s
∼ inf ,
p
k∈Z χ{τk <∞} p
p
where 0 < t ≤ min {p− , 1} and the infimum is taken over all decompositions of the
form (3).
If we replace the space Hps by Pp (resp. by Qp ) and the (s, p, ∞)-atoms by
(M, p, ∞)-atoms (resp. by (S, p, ∞)-atoms), then the theorem holds, too.
If the stochastic basis (Fn ) is regular and 0 < t < min {p− , 1}, then the same
S
holds for the space HpM as for Pp and the same for Hp as for Qp .
Proof We will sketch the proof for Hps , only. Assume that f ∈ Hps and let us define
the following stopping times:
4 5
τk := inf n ∈ N : sn+1 (f ) > 2k .
where
τ τ
fn k+1 − fn k
μk := 3 · 2k χ{τk <∞} p and ank := .
μk
Mixed Martingale Hardy Spaces and Applications in Fourier Analysis 719
s (f τk+1 ) + s (f τk ) −1
s ak ≤ ≤ χ{τk <∞} p ,
μk
∞)-atom.
thus a k is an (s, p,
Since
lim s f − f τk = lim s f τk = 0
k→∞ k→−∞
almost everywhere, by the dominated convergence theorem (see e.g. [3]) we get that
m
k
f − μk a ≤ f − f τm+1 H s + f τ−l H s → 0
p p
k=−l Hps
Denote by
4 5
Ok := {τk < ∞} = s(f ) > 2k .
Then for all k ∈ Z, Ok+1 ⊂ Ok . Moreover, for all x ∈ and for all 0 < t ≤ 1,
t
t
3 · 2k χOk (x) ≤C 3 · 2k χOk \Ok+1 (x) .
k∈Z k∈Z
≤ C s(f )χOk \Ok+1
k∈Z p
= C s(f )p .
where the infimum is taken over all atomic decompositions of the form (3). There
are also corresponding equivalences for the other Hardy spaces. From Theorem 5.3,
we get immediately the next corollary.
Corollary 5.4 If the stochastic basis (Fn ) is regular, then
If s(a) in (ii) is replaced by S(a) (resp. M(a)), then the function a is called simple
r)-atom (resp. simple (M, p,
(S, p, r)-atom).
r)-atoms are more complicated than
The atomic decomposition via simple (s, p,
the atomic decomposition via (s, p, ∞)-atoms. To this, we need the condition that
every σ -algebra is generated by countably many atoms.
Theorem 5.6 ([94]) Suppose that every σ -algebra (Fn )n is generated by countably
many atoms. Let 1 < r ≤ ∞ and
⎧
⎪
⎪ p1 < r1 , r2 , . . . , rd ,
⎨
p2 < r2 , . . . , rd ,
⎪
⎪ · · ·
⎩
pd < rd .
A martingale f = (fn )n∈N ∈ Hps if and only if there exist a sequence (a k,j,i )k,j,i of
r)-atoms associated with (Ik,j,i )k,j,i ⊂ A(Fj ), which are disjoint for
simple (s, p,
fixed k, and a sequence (μk,j,i )k∈Z,j ∈N,i of positive real numbers such that
∞
fn = μk,j,i En a k,j,i a.e. (n ∈ N) (5)
k∈Z j =0 i
and
⎛ ⎛ ⎞t ⎞1/t
∞
⎝ ⎝
μ χ
k,j,i Ik,j,i
⎠⎠
f H s ∼ inf ,
p χIk,j,i p
k∈Z j =0 i
p
where 0 < t < min {p− , 1} and the infimum is taken over all decompositions of the
form (5).
If we replace the space Hps by Pp (resp. by Qp ) and the simple (s, p, r)-atoms
by simple (M, p, r)-atoms (resp. by simple (S, p,
r)-atoms), then the theorem holds,
too.
If the stochastic basis (Fn ) is regular, then the same holds for the space HpM as
for Pp and the same for HpS as for Qp .
Proof We sketch the proof, only. Besides Theorem 3.7, the basic idea of the proof
is to decompose the sets {τk = j } into the union of atoms (Ik,j,i )i ⊂ Fj such that
C
Ik,j,i = {τk = j } ∈ Fj ,
i
722 F. Weisz
where the stopping times τk were defined in the proof of Theorem 5.3. Note that for
fixed k, j , the atoms (Ik,j,i )i ⊂ Fj are disjoint. We can show that
τ
n−1
τ
fn = (fn − fn ) =
k+1 τk
χIk,j,i (fn k+1 − fnτk )
k∈Z k∈Z j =0 i
n−1
k,j,i
= μk,j,i an ,
k∈Z j =0 i
where
τ
fn k+1 − fn k
τ
= 3 · 2 χIk,j,i p
k,j,i
μk,j,i k
and an = χIk,j,i .
μk,j,i
Similarly to (4), we obtain that
⎛ ⎞1/t
∞
t
f H s ∼ inf ⎝ 3 · 2k χIk,j,i ⎠ , (6)
p
k∈Z j =0 i
p
where the infimum is taken over all atomic decompositions of the form (5). The
corresponding equivalences for the other Hardy spaces hold, too.
6 Martingale Inequalities
Fτ = {F ∈ F : F ∩ {τ ≤ n} ∈ Fn , n ≥ 1}.
Mixed Martingale Hardy Spaces and Applications in Fourier Analysis 723
Tf p ≤ C f H s .
p
If we replace the spaces Hrs and Hps by HrM and Pp (resp. by HrS and Qp ) and the
∞)-atoms by (M, p,
(s, p, ∞)-atoms (resp. by (S, p, ∞)-atoms), then the theorem
holds, too.
∞)-atoms a, (S, p,
It is easy to see that for all (s, p, ∞)-atoms a or (M, p,
∞)-
atoms a and A ∈ Fτ , s(aχA ) = s(a)χA , S(aχA ) = S(a)χA and M(aχA ) =
M(a)χA . This means that the operators s, S and M satisfy condition (7). Applying
the preceding theorem to these operators, we [84] obtain
Theorem 6.2 We have the following martingale inequalities:
(i)
f H M ≤ C f H s , f H S ≤ C f H s (0 < p < 2) .
p p p p
(ii)
(iii)
(iv)
(v)
Proof Let f ∈ Hps . The σ -sublinear operator M is bounded from H2s to L2 (see
e.g. Weisz [87]), that is Mf 2 ≤ C f H s . So we can apply Theorem 6.1 with the
2
choice r = 2 and p := (p1 , . . . , pd ), where pi < 2 and we get that
The operator S is also bounded from H2s to L2 (see [87]), hence using
Theorem 6.1 we obtain
f H S ≤ C f H s (0 < p < 2) .
p p
By Burkholder-Gundy and Doob’s inequality, for all 1 < r < ∞, S(f )r ≈
M(f )r ≈ f r (see Weisz [87]). Using this, the previous inequality and
Theorem 6.1, we have
For f = (fn )n∈N ∈ Qp there exists a sequence (λn )n∈N for which Sn (f ) ≤ λn−1
and λ∞ ∈ Lp . Using the inequality |fn | ≤ Mn−1 (f ) + λn−1 and the preceding
inequality, we get that
Similarly, if f = (fn )n∈N ∈ Pp , then |fn | ≤ λn−1 with a suitable sequence
(λn )n∈N for which λ∞ ∈ Lp . Since
n 1/2
Sn (f ) = |dk f | 2
≤ Sn−1 (f ) + |dn f | ≤ Sn−1 (f ) + 2λn−1 ,
k=0
we have that
Mixed Martingale Hardy Spaces and Applications in Fourier Analysis 725
We know (see e.g. Weisz [87]) that Sn (f ) ≤ R 1/2 sn (f ) if the stochastic basis is
regular. Using the definition of Qp and the fact that sn (f ) ∈ Fn−1 we get
By the last inequality of Theorem 6.2, we obtain that Qp = Hps . The next corollary
follows from Theorem 6.2 and Corollary 5.4 (see [84]).
Corollary 6.3 If the stochastic basis (Fn ) is regular, then the five Hardy spaces are
equivalent, that is
for some k ∈ {1, . . . , d}. Then for all non-negative, measurable function sequence
(fn )n∈N ,
En (fn ) ≤ C fn .
n∈N p n∈N p
f H s ≤ C f H S .
p p
The so called Davis decomposition (Lemma 6.6) holds also for mixed norm spaces.
The main idea of the proof of this decomposition is Theorem 6.4 (see [84]).
726 F. Weisz
Lemma 6.6 Suppose that 1 < p < ∞ or p satisfies (8). If f ∈ HpS , then there
exists h ∈ Gp and g ∈ Qp such that f = h + g and
HpS = HpM
In this section, we study the dual spaces of mixed normed martingale Hardy
spaces. To this end, we have to suppose that every σ -algebra (Fn )n is generated
by countably many atoms.
Definition 7.1 Suppose that every σ -algebra (Fn )n is generated by countably many
atoms. Let 0 ≤ α < ∞ and 1 ≤ q < ∞. Define BMOq ( α ) as the space of functions
f ∈ Lq for which
f BMOq (α) = sup sup χI −11 χI (q) (f − fn )χI q < ∞.
n≥0 I ∈A(Fn ) α +1
lϕ (f ) = E(f ϕ) (f ∈ L2 ).
2)-
and the convergence holds also in the L2 -norm, where a k,j,i is a simple (s, p,
atom and μk,j,i = 3 · 2k χIk,j,i p . Hence
∞
lϕ (f ) = E(f ϕ) = μk,j,i E(a k,j,i ϕ).
k∈Z j =0 i
Observe that
∞
≤ μk,j,i a k,j,i 2 (ϕ − ϕj )χIk,j,i 2
k∈Z j =0 i
∞
P(Ik,j,i )1/2
≤ μk,j,i (ϕ − ϕj )χIk,j,i 2
χI
k∈Z j =0 i k,j,i p
∞
μk,j,i ϕBMO2 (α) . (9)
k∈Z j =0 i
Indeed, since 0 < p ≤ 1, the inequality holds for all Lpi spaces (i = 1, . . . , d),
hence it holds also for Lp spaces. Taking into account this inequality, (6) and
Theorem 5.6, we conclude that
∞
∞
μk,j,i 2k χIk,j,i p
k∈Z j =0 i k∈Z j =0 i
∞
k
≤
2 χ
Ik,j,i f Hps .
k∈Z j =0 i
p
l(f ) = lϕ (f ) = E(f ϕ) (f ∈ L2 ).
(ϕ − ϕj )χI
a= .
(ϕ − ϕj )χI 2 χI 1 χI −1
2
α +1
l(f ) = E(f ϕ) (f ∈ L∞ ).
Mixed Martingale Hardy Spaces and Applications in Fourier Analysis 729
We can verify the following result similarly to Theorem 7.2 (see [94]).
Theorem 7.3 Suppose that every σ -algebra (Fn )n is generated by countably many
atoms. If 0 < p ≤ 1 and α = 1/p − 1, then
∗
Pp 1 = BMO1 (
α)
α ) = BMOr (
BMO1 ( α)
where
∞
n= nk 2k , (0 ≤ nk < 2).
k=0
fB(n) := E(f wn ) (n ∈ N)
Since wn is Fk measurable for n < 2k , it can immediately be seen that this limit
does exist. We remember that if f ∈ L1 , then Ek f → f in the L1 -norm as k → ∞,
hence
Thus the Walsh-Fourier coefficients of f ∈ L1 are the same as the ones of the
martingale (Ek f )k≥0 obtained from f .
Mixed Martingale Hardy Spaces and Applications in Fourier Analysis 731
n−1
sn f := fB(k)wk (n ∈ N).
k=0
It is a basic question, whether the function f can be reconstructed from the partial
sums of its Fourier series. It is easy to see that, for any martingale f = (fn ),
s2n f = fn (n ∈ N).
where 1 ≤ p < ∞ and f ∈ Lp . This result was generalized by Paley [71] and
Schipp et al. [75, Theorem 4.1].
Theorem 8.1 If f ∈ Lp for some 1 < p < ∞, then
and
One of the deepest results in harmonic analysis is Carleson’s result (see Carleson
[8], Hunt [47], Billard [4], Sjölin [79]). Using tree martingales, Schipp [73] gave a
nice proof for the theorem (see also [76, 87]).
Theorem 8.2 If f ∈ Lp for some 1 < p < ∞, then
sup |sn f | ≤ Cp f Lp
n∈N Lp
and
lim sn f = f a.e.
n→∞
Though Theorems 8.1 and 8.2 are not true for p = 1, with the help of some
summability methods they can be generalized for these endpoint cases. Obviously,
summability means have better convergence properties than the original Fourier
series. Summability is intensively studied in the literature. We refer at this time
732 F. Weisz
only to the books Stein and Weiss [82], Butzer and Nessel [7], Trigub and Belinsky
[85], Grafakos [34] and Weisz [90–92, 95] and the references therein.
The best known summability method is the Fejér method. In 1904 Fejér [20]
investigated the arithmetic means of the partial sums of the trigonometric Fourier
series, the so called Fejér means and proved that if the left and right limits f (x − 0)
and f (x + 0) exist at a point x, then the Fejér means converge to (f (x − 0) + f (x +
0))/2. One year later Lebesgue [62] extended this theorem and obtained that every
integrable function is Fejér summable almost everywhere.
We define the Fejér summability means by the arithmetic means of the partial
sums:
1
n−1 n
j B
σn f := sk f = 1− f (j )wj .
n n
k=0 j =0
and
To obtain almost everywhere convergence for the Fejér means, we introduce the
maximal operator of the Fejér means:
σ∗ f := sup |σn f |.
n∈N
Fujii [22] proved that σ∗ is bounded from H1 to L1 (see also Schipp and
Simon [74]). Later, using the atomic decomposition, the author [88] (see also [90])
generalized this result to all 1/2 < p < ∞:
Theorem 8.4 If 1/2 < p ≤ ∞ and f ∈ Hp , then
For p ≤ 1/2, the theorem does not hold (see Simon and Weisz [78], Simon
[77]). We get the next weak type (1, 1) inequality from Theorem 8.4 by interpolation
(Weisz [88, 90]). It was originally proved by Schipp [72].
Corollary 8.5 If f ∈ L1 , then
This weak type (1, 1) inequality and the density argument of Marcinkiewicz and
Zygmund [67] imply the next corollary, which was proved by Fine [21] and later
Schipp [72].
Corollary 8.6 If f ∈ L1 , then
lim σn f = f a.e.
n→∞
d
x ) :=
wn ( wnk (xk )
k=1
fB(
n) := E(f wn ) n = (n1 , . . . , nd ) ∈ Nd .
1 −1
n d −1
n
sn f := ··· fB(k)
w
k n ∈ Nd )
(
k1 =0 kd =0
and
1
n1
nd
σn f := d ··· sk f n ∈ Nd ),
(
k=1 nk k1 =1 kd =1
734 F. Weisz
respectively. We will investigate the convergence of the Fejér means over the
diagonal, or more generally, over cones. For α ≥ 0 let us define the cone
ni
α := n = (n1 , . . . , nd ) ∈ N : 2−α ≤ ≤ 2α (i, j = 1, . . . , d) .
nj
For the almost everywhere convergence, we have to investigate the Fejér maximal
operator,
σ∗ f := sup |σn f | .
n ∈α
Using Theorem 5.3, we can verify that σ∗ is bounded from Hp to Lp for 1/2 < p <
∞.
Theorem 9.1 ([83]) If α ≥ 0 and 1/2 < p < ∞, then for all f ∈ Hp ,
For a constant p, this theorem is due to the author [89, 90]. There are counterex-
amples for the boundedness of σ∗ if p ≤ 1/2 (Goginava and Nagy [32]). Since the
Walsh polynomials are dense in Hp , the following consequences of Theorem 9.1
can be proved by a density argument in the usual way (see [83]).
Corollary 9.2 If α ≥ 0 and 1/2 < p < ∞, then for all f ∈ Hp , σn f (
x ) converges
for almost every x ∈ [0, 1)d and in the Lp -norm as n → ∞ and n ∈ α .
If I ∈ Fk is a dyadic cube with length 2−dk , then the restriction of the martingale
f to I is defined by
f χI := (fn χI : n ≥ k) .
Corollary 9.3 Let α ≥ 0, 1/2 < p < ∞ and f ∈ Hp . If there exists a dyadic cube
I , such that the restricted martingale f χI ∈ L1 (I ), then
lim x ) = f (
σn f ( x) for a.e. x ∈ I and in the Lp (I )-norm.
n ∈α ,
n→∞
lim x ) = f (
σn f ( x) for a.e. x ∈ [0, 1)2 and in the Lp -norm.
n ∈α ,
n→∞
Recall that H1 ⊂ L1 and Hp ∼ Lp for 1 < p < ∞. Next we generalize the
preceding corollary. Theorem 9.1 and interpolation imply (Weisz [89, 90])
Mixed Martingale Hardy Spaces and Applications in Fourier Analysis 735
This corollary was proved independently by Gát [24] and Weisz [89].
Acknowledgment This research was supported by the Hungarian National Research, Develop-
ment and Innovation Office—NKFIH, KH130426.
References
16. B.J. Davis, On the integrability of the martingale square function. Isr. J. Math. 8, 187–190
(1970)
17. J.L. Doob, Semimartingales and subharmonic functions. Trans. Am. Math. Soc. 77, 86–121
(1954)
18. C. Fefferman, Characterizations of bounded mean oscillation. Bull. Am. Math. Soc. 77,
587–588 (1971)
19. C. Fefferman, E.M. Stein, H p spaces of several variables. Acta Math. 129, 137–194 (1972)
20. L. Fejér, Untersuchungen über Fouriersche Reihen. Math. Ann. 58, 51–69 (1904)
21. N.J. Fine, Cesàro summability of Walsh-Fourier series. Proc. Nat. Acad. Sci. USA 41, 558–
591 (1955)
22. N. Fujii, A maximal inequality for H 1 -functions on a generalized Walsh-Paley group. Proc.
Am. Math. Soc. 77, 111–116 (1979)
23. A.M. Garsia, Martingale Inequalities. Seminar Notes on Recent Progress. Math. Lecture
Note. (Benjamin, New York, 1973)
24. G. Gát, Pointwise convergence of the Cesàro means of double Walsh series. Ann. Univ. Sci.
Budapest Sect. Comput. 16, 173–184 (1996)
25. G. Gát, Almost everywhere convergence of Fejér means of two-dimensional triangular Walsh-
Fourier series. J. Fourier Anal. Appl. 24(5), 1249–1275 (2018)
26. G. Gát, Cesàro means of subsequences of partial sums of trigonometric Fourier series. Constr.
Approx. 49(1), 59–101 (2019)
27. A.G. Georgiadis, J. Johnsen, M. Nielsen, Wavelet transforms for homogeneous mixed-norm
Triebel-Lizorkin spaces. Monatsh. Math. 183(4), 587–624 (2017)
28. A.G. Georgiadis, M. Nielsen, Pseudodifferential operators on mixed-norm Besov and Triebel-
Lizorkin spaces. Math. Nachr. 289(16), 2019–2036 (2016)
29. U. Goginava, On some (Hp,q , Lp,q )-type maximal inequalities with respect to the Walsh-
Paley system. Georgian Math. J. 7, 475–488 (2000)
30. U. Goginava, Almost everywhere summability of multiple Walsh-Fourier series. J. Math.
Anal. Appl. 287, 90–100 (2003)
31. U. Goginava, Maximal operators of (C, α)-means of cubic partial sums of d-dimensional
Walsh-Fourier series. Anal. Math. 33, 263–286 (2007)
32. U. Goginava, Maximal operators of Fejér means of double Walsh-Fourier series. Acta Math.
Hungar. 115, 333–340 (2007)
33. B. Golubov, A. Efimov, V. Skvortsov, Walsh Series and Transforms (Kluwer Academic,
Dordrecht, 1991)
34. L. Grafakos, Classical and Modern Fourier Analysis (Pearson Education, Upper Saddle
River, NJ, 2004)
35. J. Hart, R.H. Torres, X. Wu, Smoothing properties of bilinear operators and Leibniz-type
rules in Lebesgue and mixed Lebesgue spaces. Trans. Am. Math. Soc. 370(12), 8581–8612
(2018)
36. C. Herz, Bounded mean oscillation and regulated martingales. Trans. Am. Math. Soc. 193,
199–215 (1974)
37. C. Herz, Hp -spaces of martingales, 0 < p ≤ 1. Z. Wahrscheinlichkeitstheorie Verw. Geb.
28, 189–205 (1974)
38. K.-P. Ho, Strong maximal operator on mixed-norm spaces. Ann. Univ. Ferrara Sez. VII, Sci.
Mat. 62(2), 275–291 (2016)
39. K.-P. Ho, Mixed norm Lebesgue spaces with variable exponents and applications. Riv. Mat.
Univ. Parma (N.S.) 9(1), 21–44 (2018)
40. L. Hörmander, Estimates for translation invariant operators in Lp spaces. Acta Math. 104,
93–140 (1960)
41. L. Huang, J. Liu, D. Yang, W. Yuan, Atomic and Littlewood-Paley characterizations of
anisotropic mixed-norm Hardy spaces and their applications. J. Geom. Anal. 29, 1991–2067
(2019)
42. L. Huang, J. Liu, D. Yang, W. Yuan, Dual spaces of anisotropic mixed-norm Hardy spaces.
Proc. Am. Math. Soc. 147(3), 1201–1215 (2019)
Mixed Martingale Hardy Spaces and Applications in Fourier Analysis 737
43. L. Huang, J. Liu, D. Yang, W. Yuan, Identification of anisotropic mixed-norm Hardy spaces
and certain homogeneous Triebel-Lizorkin spaces. J. Approx. Theory 258, 105459 (2020)
44. L. Huang, J. Liu, D. Yang, W. Yuan, Real-variable characterizations of new anisotropic
mixed-norm hardy spaces. Commun. Pure Appl. Anal. 19(6), 3033–3082 (2020)
45. L. Huang, F. Weisz, D. Yang, W. Yuan, Summability of Fourier transforms on mixed-norm
Lebesgue spaces via associated Herz spaces. Anal. Appl. (2021). https://doi.org/10.1142/
S0219530521500135
46. L. Huang, D. Yang, On function spaces with mixed norms - a survey. J. Math. Study 54, 1–75
(2021)
47. R.A. Hunt, On the convergence of Fourier series, in Orthogonal Expansions and their
Continuous Analogues, Proc. Conf. Edwardsville, IL, 1967 (Illinois Univ. Press, Carbondale,
1968), pp. 235–255
48. Y. Jiang, W. Sun, Adaptive sampling of time-space signals in a reproducing kernel subspace
of mixed Lebesgue space. Banach J. Math. Anal. 14(3), 821–841 (2020)
49. Y. Jiao, F. Weisz, L. Wu, D. Zhou, Variable martingale Hardy spaces and their applications in
Fourier analysis. Diss. Math. 550, 1–67 (2020)
50. Y. Jiao, F. Weisz, L. Wu, D. Zhou, Dual spaces for variable martingale Lorentz-Hardy spaces.
Banach J. Math. Anal. 15, 53 (2021)
51. Y. Jiao, L. Wu, A. Yang, R. Yi, The predual and John-Nirenberg inequalities on generalized
BMO martingale space. Trans. Am. Math. Soc. 369, 537–553 (2017)
52. Y. Jiao, G. Xie, D. Zhou, Dual spaces and John-Nirenberg inequalities of martingale Hardy-
Lorentz-Karamata spaces. Q. J. Math. 66, 605–623 (2015)
53. Y. Jiao, D. Zhou, Z. Hao, W. Chen, Martingale Hardy spaces with variable exponents. Banach
J. Math 10, 750–770 (2016)
54. Y. Jiao, Y. Zuo, D. Zhou, L. Wu, Variable Hardy-Lorentz spaces H p(·),q (Rn ). Math. Nachr.
292, 309–349 (2019)
55. F. John, L. Nirenberg, On functions of bounded mean oscillation. Commun. Pure Appl. Math.
14, 415–426 (1961)
56. J. Johnsen, S. Munch Hansen, W. Sickel, Characterisation by local means of anisotropic
Lizorkin-Triebel spaces with mixed norms. Z. Anal. Anwend. 32(3), 257–277 (2013)
57. J. Johnsen, S. Munch Hansen, W. Sickel, Anisotropic, mixed-norm Lizorkin-Triebel spaces
and diffeomorphic maps. J. Funct. Spaces 2014, 15 (2014). Id/No 964794
58. J. Johnsen, S. Munch Hansen, W. Sickel, Anisotropic Lizorkin-Triebel spaces with mixed
norms – traces on smooth boundaries. Math. Nachr. 288(11–12), 1327–1359 (2015)
59. J. Johnsen, W. Sickel, A direct proof of Sobolev embeddings for quasi-homogeneous
Lizorkin-Triebel spaces with mixed norms. J. Funct. Spaces Appl. 5(2), 183–198 (2007)
60. J. Johnsen, W. Sickel, On the trace problem for Lizorkin–Triebel spaces with mixed norms.
Math. Nachr. 281(5), 669–696 (2008)
61. R.H. Latter, A characterization of H p (Rn ) in terms of atoms. Stud. Math. 62, 92–101 (1978)
62. H. Lebesgue, Recherches sur la convergence des séries de Fourier. Math. Ann. 61, 251–280
(1905)
63. J. Liu, F. Weisz, D. Yang, W. Yuan, Variable anisotropic Hardy spaces and their applications.
Taiwan. J. Math. 22, 1173–1216 (2018)
64. J. Liu, F. Weisz, D. Yang, W. Yuan, Littlewood-Paley and finite atomic characterizations of
anisotropic variable Hardy-Lorentz spaces and their applications. J. Fourier Anal. Appl. 25,
874–922 (2019)
65. R. Long, Martingale Spaces and Inequalities (Peking University Press and Vieweg
Publishing, Beijing, Braunschweig, 1993)
66. S. Lu, Four Lectures on Real H p Spaces (World Scientific, Singapore, 1995)
67. J. Marcinkiewicz, A. Zygmund, On the summability of double Fourier series. Fund. Math.
32, 122–132 (1939)
68. E. Nakai, Y. Sawano, Hardy spaces with variable exponents and generalized Campanato
spaces. J. Funct. Anal. 262(9), 3665–3748 (2012)
69. T. Nogayama, Mixed Morrey spaces. Positivity 23(4), 961–1000 (2019)
738 F. Weisz
70. T. Nogayama, T. Ono, D. Salim, Y. Sawano, Atomic decomposition for mixed Morrey spaces.
J. Geom. Anal. 31(9), 9338–9365 (2021)
71. R.E.A.C. Paley, A remarkable system of orthogonal functions. Proc. Lond. Math. Soc. 34,
241–279 (1932)
72. F. Schipp, Über gewissen Maximaloperatoren. Ann. Univ. Sci. Budapest Sect. Math. 18,
189–195 (1975)
73. F. Schipp, Universal contractive projections and a.e. convergence, in Probability Theory
and Applications, Essays to the Memory of József Mogyoródi, ed. by J. Galambos, I. Kátai
(Kluwer Academic, Dordrecht, 1992), pp. 47–75
74. F. Schipp, P. Simon, On some (H, L1 )-type maximal inequalities with respect to the Walsh-
Paley system, in Functions, Series, Operators, Proc. Conf. in Budapest, 1980. Coll. Math.
Soc. J. Bolyai, vol. 35 (North Holland, Amsterdam, 1981), pp. 1039–1045
75. F. Schipp, W.R. Wade, P. Simon, J. Pál, Walsh Series: An Introduction to Dyadic Harmonic
Analysis (Adam Hilger, Bristol, New York, 1990)
76. F. Schipp, F. Weisz, Tree martingales and a.e. convergence of Vilenkin-Fourier series. Math.
Pannon. 8, 17–36 (1997)
77. P. Simon, Cesàro summability with respect to two-parameter Walsh systems. Monatsh. Math.
131, 321–334 (2000)
78. P. Simon, F. Weisz, Weak inequalities for Cesàro and Riesz summability of Walsh-Fourier
series. J. Approx. Theory 151, 1–19 (2008)
79. P. Sjölin, An inequality of Paley and convergence a.e. of Walsh-Fourier series. Ark. Mat.
7(6), 551–570 (1969)
80. A. Stefanov, R.H. Torres, Calderón-Zygmund operators on mixed Lebesgue spaces and
applications to null forms. J. Lond. Math. Soc. II. Ser. 70(2), 447–462 (2004)
81. E.M. Stein, Harmonic Analysis: Real-variable Methods, Orthogonality and Oscillatory
Integrals (Princeton Univ. Press, Princeton, NJ, 1993)
82. E.M. Stein, G. Weiss, Introduction to Fourier Analysis on Euclidean Spaces (Princeton Univ.
Press, Princeton, N.J., 1971)
83. K. Szarvas, F. Weisz, Applications of mixed martingale Hardy spaces in Fourier analysis J.
Math. Anal. Appl. 492, 124403 (2020)
84. K. Szarvas, F. Weisz, Mixed martingale Hardy spaces. J. Geom. Anal. 31, 3863–3888 (2021)
85. R.M. Trigub, E.S. Belinsky, Fourier Analysis and Approximation of Functions (Kluwer
Academic, Dordrecht, 2004)
86. F. Weisz, Martingale Hardy spaces for 0 < p ≤ 1. Probab. Theory Relat. Fields 84, 361–376
(1990)
87. F. Weisz, Martingale Hardy Spaces and their Applications in Fourier Analysis. Lecture Notes
in Math., vol. 1568 (Springer, Berlin, 1994)
88. F. Weisz, Cesàro summability of one- and two-dimensional Walsh-Fourier series. Anal. Math.
22, 229–242 (1996)
89. F. Weisz, Cesàro summability of two-dimensional Walsh-Fourier series. Trans. Am. Math.
Soc. 348, 2169–2181 (1996)
90. F. Weisz, Summability of Multi-dimensional Fourier Series and Hardy Spaces. Mathematics
and Its Applications (Kluwer Academic, Dordrecht, 2002)
91. F. Weisz, Summability of multi-dimensional trigonometric Fourier series. Surv. Approx.
Theory 7, 1–179 (2012)
92. F. Weisz, Convergence and Summability of Fourier Transforms and Hardy Spaces. Applied
and Numerical Harmonic Analysis (Springer, Birkhäuser, Basel, 2017)
93. F. Weisz, Summability in mixed-norm Hardy spaces. Ann. Univ. Sci. Budapest Sect. Comput.
48, 233–246 (2018)
94. F. Weisz, Dual spaces of mixed-norm martingale Hardy spaces. Commun. Pure Appl. Anal.
20, 681–695 (2021)
95. F. Weisz, Lebesgue Points and Summability of Higher Dimensional Fourier Series. Applied
and Numerical Harmonic Analysis (Springer, Birkhäuser, Basel, 2021)
Mixed Martingale Hardy Spaces and Applications in Fourier Analysis 739
96. G. Xie, Y. Jiao, D. Yang, Martingale Musielak-Orlicz Hardy spaces. Sci. China Math. 62(8),
1567–1584 (2019)
97. G. Xie, F. Weisz, D. Yang, Y. Jiao, New martingale inequalities and applications to Fourier
analysis. Nonlinear Anal. 182, 143–192 (2019)
98. G. Xie, D. Yang, Atomic characterizations of weak martingale Musielak-Orlicz Hardy spaces
and their applications. Banach J. Math. Anal. 13(4), 884–917 (2019)
99. X. Yan, D. Yang, W. Yuan, C. Zhuo, Variable weak Hardy spaces and their applications. J.
Funct. Anal. 271, 2822–2887 (2016)
100. D. Yang, Y. Liang, L.D. Ky, Real-Variable Theory of Musielak-Orlicz Hardy Spaces. Number
2182 in Lecture Notes in Mathematics (Springer, Berlin, 2017)
The First Eigenvalue for Nonlocal
Operators
Julio D. Rossi
Abstract In this chapter we present some results concerning the first eigenvalue
for a nonlocal operator in convolution form with a smooth kernel. Given a bounded
domain ⊂ RN and a smooth kernel J , we deal with the eigenvalue problem
$
J (x − y)(u(y) − u(x)) dy = −λ1 u(x), x ∈ ,
A
First, let us briefly introduce the prototype of nonlocal problem that will be
considered along this chapter. Let J : RN → R be a nonnegative, radial, continuous
function with
$
J (z) dz = 1.
RN
J. D. Rossi ()
Departamento de Matemática, FCEyN, Universidad de Buenos Aires, Buenos Aires, Argentina
e-mail: [email protected]
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 741
R. M. Aron et al. (eds.), Operator and Norm Inequalities and Related Topics,
Trends in Mathematics, https://doi.org/10.1007/978-3-031-02104-6_22
742 J. D. Rossi
and variations of it, have been recently widely used to model diffusion processes.
More precisely, as stated in [2, 20], if u(x, t) is thought of as a density at the point x
at time t and J (x − y) is thought
) of as the probability distribution of jumping from
location y to location x, then RN J (y − x)u(y, t) dy = (J ∗ u)(x, t) is the rate at
) individuals are arriving at position x from all other places and −u(x, t) =
which
− RN J (y − x)u(x, t) dy is the rate at which they are leaving location x to travel
to all other sites. This consideration, in the absence of external or internal sources,
leads immediately to the fact that the density u satisfies Eq. (1).
Equation (1) is called nonlocal diffusion equation since the to evaluate the right
hand side at a point x and time t one needs to know u in a neighborhood of x to
compute the convolution term J ∗ u. This equation shares many properties with the
classical heat equation, ut = u, such as: bounded stationary solutions are constant,
a maximum principle holds for both of them and, even if J is compactly supported,
perturbations propagate with infinite speed [20]. However, there is no regularizing
effect in general.
Let us fix a bounded domain in RN . For local problems the two most
common boundary conditions for local PDEs are Dirichlet and Neumann. When
looking at boundary conditions for nonlocal problems, one has to modify the usual
formulations for local problems.
Concerning the homogeneous Dirichlet boundary conditions for nonlocal prob-
lems we consider
⎧ $
⎪
⎪
⎪ ut (x, t) = N J (x − y)u(y, t) dy − u(x, t), x ∈ , t > 0,
⎪
⎨ R
In this model we have that diffusion takes place in the whole RN but we impose that
u vanishes outside . In the biological interpretation, we have a hostile environment
outside , any individual that jumps outside dies instantaneously. This is the
analogous of what is called homogeneous Dirichlet boundary conditions for the
heat equation. However, the boundary datum is not understood in the usual sense,
since we are not imposing that u is continuous up to ∂.
The First Eigenvalue for Nonlocal Operators 743
In this model we have that the integral term )takes into account the diffusion inside
. In fact, as we have explained, the integral J (x − y)(u(y, t) − u(x, t)) dy takes
into account the individuals arriving or leaving position x from other places. Since
we are integrating in , we are imposing that diffusion takes place only in . The
individuals may not enter nor leave the domain. This is the analogous of what is
called homogeneous Neumann boundary conditions in the literature.
These nonlocal problems has been used to model very different applied situ-
ations, for example in biology [12, 23, 29], image processing [22, 27], particle
systems [9], coagulation models [21], nonlocal anisotropic models for phase tran-
sition [1], mathematical finances [26], etc. Besides the interest for the applications
there is also a great amount of work dealing with purely mathematical issues. For
example, see [3, 4, 6, 7, 10–13, 15–20, 24, 25], and references therein.
Associated with these evolution problems there are two eigenvalue problems that
play a central role in determining the asymptotic behaviour of the solutions as t →
+∞. They are given by
⎧ $
⎨ φ(x) − J (x − y)φ(y) dy = λ1 (x), x ∈ ;
⎩
φ(x) = 0, x ∈ ,
and
$
− J (x − y)(ϕ(y) − ϕ(x)) dy = β1 ϕ(x), x ∈ .
The analysis of these eigenvalue problems is our main concern in this chapter. The
proofs of the main results are taken from the references [2, 14, 28], but we include
here some details to make this chapter self-contained.
We have to mention the close relation between this kind of evolution problems
and probability theory. In fact, when one looks at a Levy process [8] the nonlocal
operator that appears naturally is a fractional power of the Laplacian. This is out of
the scope of this chapter and we refer to [5] for a reference concerning the interplay
between nonlocal partial differential equations and probability. Although we are not
dealing with probability issues, let us explain briefly why the concrete problem (1)
has a clear probabilistic interpretation. Let (E, E) be a measurable space and P :
E × E → [0, 1] be a probability transition on E. Then, we define a transition
744 J. D. Rossi
+∞ n
t
Pt (x, A) = e−t P (n) (x, A) t ∈ R+ ,
n!
n=0
Now, consider the Markov process (Zt )t ≥0 associated to the transition function
(Pt )t ≥0, and denote by μt its distribution. Then the family (μt )t ≥0 satisfies a linear
equation of the form
$
∂
μt = P (y, ·)μt (dy) − μt .
∂t
Now, let us introduce with some detail the two eigenvalue problems that we discuss
in this chapter (Dirichlet or Neumann boundary conditions). We will deal mainly
with the L2 formulation that gives linear eigenvalue problems, and at the end of
the chapter we will comment briefly on the results and difficulties for nonlinear
eigenvalues.
The First Eigenvalue for Nonlocal Operators 745
Here and in what follows we denote by u the extension by zero of u to the whole
RN , that is,
u(x) x ∈ ,
u(x) =
0 x ∈ RN \ .
conversely, it is easy to check that if φ1 > 0 is a solution to (6) (with λ1 the smallest
eigenvalue) then it is a minimizer of (5). Hence, we look for the first eigenvalue
of (6), which is equivalent to
$
(1 − λ1 )φ1 (x) = J (x − y)φ 1 (y) dy, x ∈ . (7)
RN
Hence we are looking for the largest eigenvalue, μ = 1 − λ1 , of S. Since the kernel
is smooth, S is compact and then this eigenvalue is attained at some function φ1 (x)
that turns out to be an eigenfunction for our original problem (6). By taking |φ1 |
instead of φ1 in (5) we may assume that φ1 ≥ 0 in . Indeed, one simply has to use
the fact that (a − b)2 ≥ (|a| − |b|)2 .
Let us present some properties of the eigenvalue problem (6).
Proposition 2.1 ([2, 14]) Let λ1 the first eigenvalue of (6) and denote by φ1 (x) a
corresponding non-negative eigenfunction. Then φ1 (x) is strictly positive in and
λ1 is a positive simple eigenvalue with λ1 < 1.
Proof Since J (0) > 0 and J is continuous we have that B(0, d) ⊂ supp(J ) for
some d > 0. Let us assume, for simplicity, that supp(J ) = B(0, 1). First, observe
746 J. D. Rossi
The regularity of the eigenfunctions and the previous analysis shows that C ∗ is
nontrivial and bounded. Moreover from its definition, there must exists x ∗ ∈
such that φ2 (x ∗ ) = C ∗ φ1 (x ∗ ). Define φ(x) = C ∗ φ1 (x) − φ2 (x). From the linearity
of (6), φ is a non-negative eigenfunction associated to λ1 with φ(x ∗ ) = 0. From the
The First Eigenvalue for Nonlocal Operators 747
If u0 is continuous, positive and bounded, then there exist positive constants C and
C ∗ such that
and
Therefore
$ $
−2λ1 t
u (x, t) dx ≤ e
2
u20 (x) dx
On the other hand, we have that Cφ1 (x) is a solution of (11) for every C ∈ R and
moreover, it follows from the above eigenfunction analysis that the set of stationary
solutions of (11) is given by S∗ = {Cφ1 , C ∈ R}.
Define now for every t > 0, the function
is non-decreasing. These properties imply that the following two limits exist,
and also provide the compactness of the orbits which is necessary to pass to the limit
(after extracting subsequences if needed) in order to obtain that v(·, t +tn ) → w(·, t)
as tn → ∞ uniformly on compact subsets in × R+ and w(x, t) is a continuous
function which satisfies (11). Let us recall the concept of ω-limit set of the trajectory
u(t) that begins with u(0) = u0 ,
4 5
ω(u0 ) := g ∈ L2 () : ∃tn → ∞ with u(tn ) → g in L2 () .
Moreover, C ∗ (t) plays a role of a Lyapunov function and this fact allows to
conclude that ω(u0 ) ⊂ S∗ , the set of stationary solutions of (11), and the uniqueness
of the profile. In more detail, assume that g ∈ ω(u0 ) does not belong to S∗ and
The First Eigenvalue for Nonlocal Operators 749
consider w(x, t) the solution of (11) with initial data g(x) and define
or
$ $
J (x − y)u(x)u(y) dy dx
λ1 () = 1 − sup $ (13)
u∈L2 ()
u=0
u2 (x) dx
or
$ $
J (x − y)(u(x) − u(y))2 dy dx
1 RN RN
λ1 () = inf $ . (14)
2 u∈L2 ()
u=0 u2 (x) dx
Proof Recall that we introduced the compact operator S : L2 () → L2 () given
by
$ $
S(u)(x) := J (x − y)u(y) dy = J (x − y)u(y) dy, x∈
RN
λ1 = 1 − S.
750 J. D. Rossi
Then, the variational characterizations (12) and (13) are can be obtained at once
since
$ $ 2
and
$ $
J (x − y)u(x)u(y) dy dx
| Su, u |
S = sup = sup $ ,
u∈L2 () |u|2L2 () u∈L2 ()
u2 (x) dx
u=0 u=0
$
where A(x) = J (x − y) dy.
Proof Taking u ≡ 1 as test function in (12), we obtain
$ 1/2
1
1 − λ1 () ≥ A2 (x) dx . (16)
||
Finally, (15) follows from (16) and (17). This concludes the proof of the corollary.
The estimates (15) obtained in Corollary 2.4 are not sharp: if the domain
contains a ball of radius 2, say, then the right-hand side in (15) equals one, so the
estimate is useless.
Now, we turn our attention to the dependence of the first eigenvalue on the
domain . A first consequence of the variational characterization, (5), is the strict
monotonicity of λ1 ().
Theorem 2.5 ([28]) The principal eigenvalue, λ1 (), is decreasing with respect to
the domain, that is, if 1 ⊂ 2 , then λ1 (1 ) > λ1 (2 ).
Proof of Theorem 2.5 We notice that L2 (1 ) ⊂ L2 (2 ), provided we extend all
functions of the first space by zero outside 1 . Hence we have, thanks to the
characterization (12), that λ1 (1 ) ≥ λ1 (2 ). To show that the inequality is strict,
we notice that if λ1 (1 ) = λ1 (2 ), then we obtain an associated eigenfunction
which is positive in 1 , but zero in 2 \ 1 , which contradicts the strong maximum
principle.
Next, we analyze perturbations δ of a fixed domain , where δ is a small
parameter, and consider the issues of continuity and differentiability of λ1 (δ ) with
respect to δ. We assume that the perturbed domain verifies δ = (δ, ), where
: (−ε, ε) × → RN takes the form
Proof of Theorem 2.6 We first notice that for small δ we can always assume 1 ⊂
δ ⊂ 2 for some smooth domains 1 and 2 not depending on δ. Thanks to
Theorem 2.5 this implies
for z ∈ , where vδ (w) = uδ (w + (δ, w)) and (δ, w) = det(I +D (δ, w)). We
select vδ with the normalization |vδ |L2 () = 1. Then, for every sequence δn → 0,
we have a subsequence—still denoted by δn —such that vδn ' v weakly in L2 ().
Since
for almost every w ∈ . Using the dominated convergence theorem, we also have
the convergence in L2 ().
On the other hand, since λ1 (δn ) is bounded, we may pass to a further
subsequence to have λ1 (δn ) → μ, where 0 < μ < 1, thanks to (19). Then,
setting δ = δn in (20) and passing to the limit we have that the convergence of vδn
to v0 is strong in L2 (). Therefore, |v0 |L2 () = 1. By (20) we finally have
$
J (x − y)v0 (y) dy = (1 − μ)v0 (x), x ∈ ,
Theorem 2.7 Let λ(δ) = λ1 (δ ) be the principal eigenvalue in δ , and assume
δ = (δ, ), where is of the form (18) with ∈ C 1 ((−ε, ε) × ) for some
ε > 0 and (0, ·) = 0. Then λ(δ) is differentiable with respect to δ at δ = 0, and
$ * +
∂
λ (0) = −(1 − λ1 ()) u20 (x) (0, x), ν(x) dS(x), (21)
∂ ∂δ
where
and (z) = det(I + D (δ, z)) and D stands for differentiation with respect to the
second variable. By our regularity assumptions we have that
where
> (0, z) − (0, w)
?
K(z, w) = ∇J (z − w), u0 (z)u0 (w)
> (0, z)
?
+J (z − w)u0 (w) ∇u0 (z),
> (0, w)
?
+J (z − w)u0 (z) ∇u0 (w),
+J (z − w)u0 (z)u0 (w) div( (0, z))
and stands for differentiation with respect to δ. Integrating (24) with respect to z
and w in , we get,
$ $
A(z, w) (z)(w) dz dw
$ $ $ $
= J (z − w)u0 (z)u0 (w) dz dw + δ K(z, w) dz dw + o(δ)
$ $
= μ(0) + δ K(z, w) dz dw + o(δ).
(25)
and thus
$
μ(δ) − μ(0) >
?
lim inf ≥ μ(0) u20 (z) (0, z), ν(z) dS(z).
δ→0+ δ ∂
The remaining limits, lim supδ→0+ , lim infδ→0− and lim supδ→0− of the incremen-
tal quotients μ(δ)−μ(0)
δ can be proved with similar calculations (we only remark
that for the upper estimate the continuity of uδ is needed), and therefore we finally
conclude that
$
μ(δ) − μ(0) > ?
lim = μ(0) u20 (z) (0, z), ν(z) dS(z).
δ→0 δ ∂
C
k
δ = {x ∈ RN : dist(x, i ) < δ}. (28)
i=1
756 J. D. Rossi
We have δ = (δ, ), where (δ, x) = x + δ 1(x). Moreover, the derivative with
respect to δ, 1 = ∂∂δ (0, ·). verifies 1 = ν on the components i while 1 = 0 on
the remaining components of the boundary. Hence, we obtain that λ1 (δ ) decreases
linearly as δ goes to zero.
Corollary 2.8 Let be a bounded C 1 domain of RN , and assume δ is the
perturbation of given by (28). Then λ(δ) = λ1 (δ ) is differentiable with respect
to δ at δ = 0, and
k $
λ (0) = −(1 − λ1 ()) u20 (x) dS(x) < 0,
i=1 i
where u0 is the positive eigenfunction of λ1 () normalized with |u0 |L2 () = 1.
Having established the smoothness and monotonicity of λ1 (), we proceed with
the analysis of its asymptotic behavior both for small and large domains . In this
context n → RN means that the sequence of sets n contains balls BRn (centered
at a fixed point) with radii Rn → +∞.
Theorem 2.9 ([28]) For the principal eigenvalue λ1 () we have λ1 () → 1 when
|| → 0 and λ1 (n ) → 0 when n → RN .
Proof of Theorem 2.9 We make use of Corollary 2.4. First, notice that if || → 0,
the integral in the second inequality in (15) goes to zero, and thus λ1 () → 1.
To prove that λ1 (n ) → 0 when n → RN , we first show that λ1 (BR ) → 0
when R → ∞. According to (15) we have
$ $ 1/2
2
1
λ1 (BR ) ≤ 1 − J (x − y) dy dx ,
|BR | BR BR
J (x − y) ≤ J (0) + e
if x, y ∈ γ . Then
$ $ $
(1 − λ1 (γ )) uγ (x) dx = J (x − y)uγ (y) dy dx
γ $ $ γ γ $
≤ (J (0) + e) uγ (y) dy dx = (J (0) + e)||γ N uγ (y) dy.
γ γ γ
It follows that
1 − λ1 (γ )
lim sup ≤ J (0)||.
γ →0+ γN
The reverse inequality for the liminf can be proved in an analogous way. This
completes the proof of (30).
Let us prove now (31), which is much more involved. The first step is to show
that λ1 (γ ) ≤ Cγ −2 for a certain positive constant. Indeed, we will show the more
precise estimate,
Let
) 2φ be the positive eigenfunction of the Laplacian in , normalized by
φ (x) dx = 1 and extended by zero outside . Taking as a test function
φγ (x) = φ(x/γ ) in the variational characterization (14), we obtain
$ $ 2
1 x y
J (x − y) φ −φ dy dx
2 RN RN γ γ
λ1 (γ ) ≤ $ 2
.
x
φ dx
γ γ
Taking into account that the function φ belongs to W 1,∞ (RN ), we have
z 1
$ 1* z
+
φ w+ − φ(w) = ∇φ w + s , z ds
γ γ 0 γ
1
$ $ $ 1* z
+ 2
γ 2 λ1 (γ ) ≤ J (z) ∇φ w + s , z ds dw dz. (34)
2 B1 RN 0 γ
)
We notice that the integrals B1 J (z)zi zj dz vanish by symmetry when i = j , while
they are all equal to 2A(J ) when i = j . Thus (35) implies (33).
Now let ϕγ be a positive eigenfunction
) associated to λ1 (γ ), and set ψγ (x) =
ϕγ (γ x), x ∈ . We normalize ψγ by ψγ2 (x) dx = 1. According to the variational
characterization (14), we have
$ $
2λ1 (γ ) = Jγ (x − y)(ψγ (x) − ψγ (y))2 dxdy,
1
1
Indeed, thanks to (33), we may assume that γn2 λ1 (γn ) → λ0 ≥ 0. We notice that
ψn satisfies
Note that all the integrals in what follows may be considered in RN , since v and
ψn vanish outside . Thanks to Fubini’s theorem, the integrals in the left-hand side
of (38) can be rewritten to obtain
$ $
γnN J (γn (x − y))(v(y) − v(x))ψn (x) dx dy
RN RN $ (39)
= −λ1 (γn ) ψn (x)v(x) dx,
RN
since J has unit integral. Letting z = −γn (x − y) in the first integral of (39), we get
$ $
z
J (z) v x + − v(x) ψn (x) dx dz
RN RN $ γ n (40)
= −λ1 (γn ) ψn (x)v(x) dx.
RN
z
v x+ − v(x)
γn
N $
1 ∂v 1 1
N
∂ 2v sz
= (x)zi + 2 (1 − s) x+ zi zj ds,
γn ∂xi γn 0 ∂xi ∂xj γn
i=1 i,j =1
The First Eigenvalue for Nonlocal Operators 761
$ $
N
∂v
J (z) γn (x)zi
R N R N ∂x i
i=1 ⎞
N $ 1
∂ 2v sz
+ (1 − s) x+ zi zj ds ⎠ ψn (x) dxdz
0 ∂xi ∂xj γn
$ i,j =1
= −γn2 λ1 (γn ) ψn (x)v(x)dx.
Next we analyze the integrals involving the first derivatives of v. Notice that
$ $ $ $
∂v ∂v
J (z) (x)zi ψn (x)dxdz = (x)ψn (x) J (z)zi dz dx = 0
RN RN ∂xi R N ∂x i RN
∂ 2v sz ∂ 2v
x+ → (x)
∂xi ∂xj γn ∂xi ∂xj
$
N $
1 ∂ 2v
(x)ψ(x) J (z)zi zj dz dx = A(J )v(x)ψ(x).
2 RN i,j =1 ∂xi ∂xj RN
Thus
$ $
A(J ) v(x)ψ(x) dx = −λ0 ψ(x)v(x) dx. (42)
RN RN
According to (36), we may integrate by parts in the integral of the left-hand side
in (42) to obtain
$ $
A(J ) ∇v(x)∇ψ(x)dx = λ0 ψ(x)v(x) dx.
RN RN
762 J. D. Rossi
Since v ∈ C0∞ () is arbitrary, and ψ ∈ H01 () with ψ ≡ 0, we have that ψ is a
positive eigenfunction associated to − in . Thus λ0 = A(J )σ1 (), and since the
sequence γn was arbitrary, the theorem is proved.
For the Neumann problem (3) the first eigenvalue is zero (with an eigenfunction
given by a constant). Hence we look for the first nontrivial eigenvalue, given by
$ $
1
J (x − y)(u(y) − u(x))2 dy dx
2
β1 (J, ) = inf) $ (43)
u∈L2 (), u=0
(u(x))2 dx
Remark that
$ $
m= inf J (x − y)(u(x) − u(y)) dy u(x) dx = β1 .
u∈H, uL2 () =1
The First Eigenvalue for Nonlocal Operators 763
m > 0.
Proof Let
$
A(x) = J (x − y) dy.
For every ε small let us choose two disjoint balls of radius ε contained in ,
B(x1,ε , ε) and B(x2,ε , ε) in such a way that xi,ε → x0 as ε → 0. By using
and
$ $
J (x − y)uε (y) uε (x) dy dx
lim
= 0.
ε→0 2|B(0, ε)|
Proof Since
$ t$
u(x, t) − u0 (x) = J (x − y) (u(y, s) − u(x, s)) dy ds,
0
The corresponding stationary problem to (3) is described by the equation
$
0= J (x − y)(ϕ(y) − ϕ(x)) dy. (46)
Proposition 2.14 Every stationary solution of (3) is constant in , and, since the
total mass is preserved, the unique stationary solution with the same mass as u0 is
$
1
ϕ= u0 (x) dx.
||
K = max ϕ(x)
x∈
6 7
and consider the set A = x ∈ : ϕ(x) = K . The set A is clearly closed and
non empty. We claim that it is also open in . Let x0 ∈ A, then
$
0= J (x0 − y)(ϕ(y) − ϕ(x0 )) dy,
therefore, since ϕ(y) ≤ ϕ(x0 ), this implies ϕ(y) = ϕ(x0 ) for all y ∈ ∩ B(x0 , d),
for any B(0, d) ⊂ supp(J ). Hence A is open as claimed. Consequently, as is
connected, A = and ϕ is constant.
Theorem 2.15 ([2]) For every u0 ∈ L2 () the solution u(x, t) of (3) satisfies
$ $
u(·, t) − 1 ≤e −β1 t 1
u0
u0
|| L2 () u0 − || 2 , (47)
L ()
Proof Let
$ $ 2
1 1
H (t) = u(x, t) − u0 dx.
2 ||
Differentiating, using (43) and the conservation of the total mass, we obtain
$ $
1
H (t) = − J (x − y)(u(y, t) − u(x, t))2 dy dx
2 $ $ 2
1
≤ −β1 u(x, t) − u0 dx.
||
766 J. D. Rossi
Hence
Therefore, integrating,
We seek for an exponential estimate in L∞ of the decay of w(x, t). The linearity of
the equation implies that w(x, t) is a solution of (3) and satisfies
$ t $
−A(x)t −A(x)t
w(x, t) = e w0 (x) + e e A(x)s
J (x − y)w(y, s) dy ds,
0
)
where A(x) = J (x − y)dx. By using (47) and Hölder’s inequality it follows that
$ t
|w(x, t)| ≤ e−A(x)t |w0 (x)| + Ce−A(x)t eA(x)s−β1s ds.
0
Therefore, w(x, t) decays to zero exponentially fast and, moreover, (48) holds
thanks to Lemma 2.12.
An analysis of the dependence of the first eigenvalue β1 with respect to the
domain is left open.
Here we briefly comment on the Lq case. We refer to [2] for details. The main
difficulty is to show that there exists an eigenfunction associated with the usual
minimization of the corresponding Raleigh quotients. Below, we just show that the
associated optimal constants are positive.
The First Eigenvalue for Nonlocal Operators 767
B0 = {x ∈ J \ : d(x, ) ≤ r/2},
B1 = {x ∈ : d(x, B0 ) ≤ r/2},
j −1
Bj = {x ∈ \ ∪k=1 Bk : d(x, Bj −1 ) ≤ r/2}, j = 2, 3, . . .
Observe that we can cover by a finite number of non null sets {Bj }ljr=1 . Now
$ $ $ $
J (x − y)|u(y) − u(x)|q dy dx ≥ J (x − y)|u(y) − u(x)|q dy dx,
RN RN Bj Bj−1
for j = 1, . . . , lr , and
$ $
J (x − y)|u(y) − u(x)|q dy dx
Bj $ B$j−1 $ $
1
≥ q J (x − y)|u(x)|q dy dx − J (x − y)|u(y)|q dy dx
2 Bj Bj−1 B B $
$ $ $
j j−1
1
= q J (x − y) dy |u(x)| dx −
q
J (x − y) dx |u(y)|q dy
2 Bj Bj−1
$ $ Bj−1 Bj
≥ αj |u(x)| dx − β
q
|u(y)| dy,
q
Bj Bj−1
where
$
1
αj = min J (x − y) dy > 0
2q x∈Bj Bj−1
Hence
$ $ $ $
J (x − y)|u(y) − u(x)|q dydx ≥ αj |u(x)|q dx − β |u(y)|q dy.
RN RN Bj Bj−1
C
m
= (B(x̂i , r) ∩ ). (52)
i=1
The First Eigenvalue for Nonlocal Operators 769
Let us argue by contradiction. Suppose that (51) is false. Then, there exists un ∈
Lq (), with un Lq () = 1, and satisfying
$ $ $
1/q
1≥n J (x − y)|un (y) − un (x)| dydx q
+ un ∀n ∈ N.
Consequently,
$ $
lim J (x − y)|un (y) − un (x)|q dy dx = 0 (53)
n
and
$
lim un = 0. (54)
n
Let
and
$
fn (x) = J (x − y)|un (y) − un (x)|q dy.
fn → 0 in L1 ().
Fn → 0 en Lq ( × ).
Let x̂1 ∈ B(x1 , δ) \ (B1 ∪ B2 ), then there exists a subsequence, denoted equal,
such that
So, successively, for x̂m ∈ B(xm , δ) \ (B1 ∪ B2 ), there exists a subsequence, again
denoted equal, such that
Now, by (52),
= (B(x̂1 , r) ∩ ) ∪ (∪m
i=2 (B(x̂i , r) ∩ )).
Hence, since is a bounded domain, there exists i2 ∈ {2, .., m} such that
(B(x̂1 , r) ∩ ) ∩ (B(x̂i2 , r) ∩ ) = ∅.
Consequently
λ1 = λ2 = . . . = λm = λ.
The First Eigenvalue for Nonlocal Operators 771
Fn (x̂1 , .) → 0 in Lq ().
un → λ in Lq (B(x̂i , r) ∩ ).
un → λ in Lq ().
By (54), λ = 0, so
un → 0 in Lq (),
References
1 Introduction
I. Cho
Dept. of Math. & Stat., St. Ambrose Univ., Davenport, IA, USA
e-mail: [email protected]
P. Jorgensen ()
Dept. of Math., Univ. of Iowa, Iowa City, IA, USA
e-mail: [email protected]
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 773
R. M. Aron et al. (eds.), Operator and Norm Inequalities and Related Topics,
Trends in Mathematics, https://doi.org/10.1007/978-3-031-02104-6_23
774 I. Cho and P. Jorgensen
multiplication-identity) 1B ∈ B, satisfying
1B · T = T = T · 1B , ∀T ∈ B,
and
ψ (1B ) = 1 = ψ 1nB , ∀n ∈ N.
Note that, even though a given topological ∗-probability space (B, ψ) is not unital,
our main results of the text hold under non-unitality. Throughout this paper, all
given (noncommutative) free probability spaces are assumed to be unital just for
convenience.
An element T ∈ B is said to be a free random variable if we understand it as an
element of (B, ψ). For example, a self-adjoint free random variable S ∈ (B, ψ) is a
self-adjoint element S ∈ B satisfying S = S ∗ in B, where S ∗ is the adjoint of S in
B.
For any arbitrary free random variables T1 , . . . , Ts ∈ (B, ψ), for s ∈ N, the free
distribution of T1 , . . . , Ts is characterized by the joint free moments,
n
r
ψ Til l = ψ Tir11 Tir22 ...Tirnn ,
l=1
of {T , T ∗ }, for all (i1 , . . . , in ) ∈ {1, . . . , s}n and (r1 , . . . , rn ) ∈ {1, ∗}n , for all
ψ
n ∈ N, where k• (.) is the free cumulant on B in terms of the linear functional ψ,
by the Möbius inversion (e.g., [17, 22–25]). In (1), the set NC(n) is the lattice of all
“noncrossing” partitions over a discrete set {1, . . . , n}, with its maximal partition,
1n = {(1, . . . , n)} ,
the single-block partition with its block (1, . . . , n), for all n ∈ N; and μ is the
Möbius functional of [23] satisfying
μ (0n , 1n ) = (−1)n+1 cn , ∀n ∈ N,
and
μ (θ, 1n ) = 0,
θ∈NC(n)
Comparing Banach Spaces for Systems of Free Random Variables Followed by. . . 775
where
(2n)!
cn = , ∀n ∈ N,
n!(n + 1)!
and 0n = {(1), (2), . . . , (n)}is the minimal partition of NC(n), having its n-many
blocks (1), (2), . . . , (n), and [θ1 , θ2 ] is the interval in NC(n) under the partial
ordering,
θ1 ≤ θ2 , ⇐⇒ ∀U ∈ θ1 , ∃V ∈ θ2 , s.t., U ⊆ V ,
1 2k 1 (2k)! (2k)!
ck = = = , (4)
k+1 k k+1 k!(2k − k)! k!(k + 1)!
are the k-th Catalan numbers for all k ∈ N0 = N ∪ {0}; (ii) to study free-
distributional data in (B, ψ) induced by the free random variables of (i); (iii)
to investigate structure theorems of a unital C ∗ -probability space generated by
countable-infinitely many free random variables of (i), (iv) to consider certain
actions of Banach-space operator on the free random variables of (iii); and (v) to
characterize how the actions of (iv) deform the free distributions of (ii).
1.1 Background
In both classical and free probability theory, the semicircular law is an important
topic (e.g., [1, 2, 5–8, 10–12, 20, 21, 28, 30]). The (classical, or free) distributions of
semicircular elements, called the semicircular law, are well-known in statistical lan-
guage. In particular, operators satisfying the semicircular law (under a fixed linear
functional) have been studied in free probability, and they are well-characterized in
analytic, or combinatorial free-probabilistic senses (e.g., [1, 17, 18, 21, 28–30]).
Semicircular elements play a key role in operator-algebraic free theory by the
(free) central limit theorem(s) (e.g., see [2, 17, 19, 28–30]). Roughly speaking, the
semicircular law is the noncommutative counterpart of the classical Gaussian (or,
the normal) distribution in commutative theory. From combinatorial approaches
(e.g., [17, 22, 23, 25]), the free distributions of semicircular elements are universally
characterized by the Catalan numbers {ck }∞ k=1 of (4). More precisely, the semicir-
cular law is characterized by the free-moment sequence,
∞
ωn c n2 = (0, c1 , 0, c2 , 0, c3 , ...) .
n=1
1.2 Motivation
Recently, from the analysis on p-adic number fields Qp , semicircular elements are
constructed (e.g., [5, 10]), for primes p. It shows connections among number theory,
Comparing Banach Spaces for Systems of Free Random Variables Followed by. . . 777
operator algebra and quantum statistical physics (e.g., [26, 27]) via free probability.
Motivated by the constructions of [5, 10], semicircular elements are generated, as
Banach-space operators (e.g., [13, 14]), by |Z|-many orthogonal projections in a
C ∗ -algebra (e.g., [6–8, 11, 12]), different from classical approaches. Independently,
the joint free distributions of mutually free, multi semicircular elements are re-
characterized both combinatorially and analytically in [9]. Such re-characterizations
and the combinatorial techniques of [9] are applied in this paper (See Sect. 3 below).
1.3 Overview
2 Preliminaries
For fundamental free probability theory, see e.g., [3, 15–17, 28–30]. Free proba-
bility is a noncommutative operator-algebraic version of classical measure theory
and statistical analysis. Roughly, a topological ∗-probability space (B, ψ) is a
noncommutative free-probabilistic analogue of a classical measure space (X, ρ)
of a measurable space X and its (bounded) measure ρ. In particular, if (B, ψ) is
unital, satisfying ϕ (1B ) = 1, then it is a noncommutative counterpart of a classical
probability space (X, ρ), equipped with the probability measure ρ, satisfying
ρ (X) = 1.
Free probability is an important branch of operator algebra theory (e.g., [9,
17, 20–22, 28, 29]), and it provides interesting applications in both mathematical
and scientific fields (e.g., [4–8, 10–12, 24, 25]). Here, we use combinatorial free
probability of e.g., [17, 22, 23, 25].
Let (A, ϕ) be a unital topological ∗-probability space. As we discussed at
the beginning, the free distribution of a self-adjoint free random variable a is
characterized by
∞
the free-moment sequence ϕ(a n ) n=1 ,
778 I. Cho and P. Jorgensen
or,
by (1) and (2) (e.g., [17, 22, 23, 25]), where k• (.) is the free cumulant on A in terms
of ϕ under the Möbius inversion of [22].
Definition 1 A self-adjoint free random variable x ∈ (A, ϕ) is said to be
semicircular, if
ϕ(x n ) = ωn c n2 , ∀n ∈ N, (6)
where ωn are in the sense of (3), and ck are the k-th Catalan numbers (4) for all
k ∈ N0 .
By the Möbius inversion, a free random variable x is semicircular in (A, ϕ), if
and only if
for all n ∈ N, where δ is the Kronecker delta. So, one can use the definition (6)
and the characterization (7) alternatively. i.e., the semicircular law, which is the
free distributions of semicircular elements, is characterized by the free-moment
sequence,
In this section, we study joint free distributions of mutually free, multi semicircular
elements in a unital C ∗ -probability space (A, ϕ). Suppose there are N-many
semicircular elements x1 , . . . , xN in (A, ϕ), for N ∈ N, and assume that they are
mutually free in (A, ϕ). By the self-adjointness of x1 , . . . , xN in A, the (joint) free
Comparing Banach Spaces for Systems of Free Random Variables Followed by. . . 779
distribution, say
denot e
ρ = ρx1 ,...,xN , (10)
∞ 6 7
∪ ∪ ϕ xi1 xi2 ...xin (11)
n=1 (i1 ,...,in )∈{1,...,N}n
denot e
Is = (i1 , . . . , is ) ∈ {1, . . . , N}s , (12)
with its cardinality s, without considering repetition. i.e., even though ij1 = ij2 as
entries of the sequence Is of (12) for j1 = j2 ∈ {1, . . . , s}, regard them as distinct
elements of the set [Is ] of (13).
Then, from the set [Is ] of (13), define a “noncrossing” partition π (i1 ) in the
noncrossing-partition lattice NC ([Is ]) (e.g., [17, 22, 23, 25]) as follows; (i) starting
from the very first entry i1 of Is , construct the maximal block U1 satisfying
[Is ] \ U1 ,
construct the second maximal block U2 of π (i1 ) containing the entry, as in (14), and
do these processes until end to have the noncrossing partition π (i1 ), and (iii) such a
resulted partition π (i1 ) must be “maximal” in NC ([Is ]), under the partial ordering
on NC ([Is ]) (e.g., see [22, 23, 25]), satisfying both (i) and (ii). For example, if
I10 = (1, 1, 2, 2, 1, 1, 1, 2, 1, 2)
780 I. Cho and P. Jorgensen
and
with
i1 = i2 = i5 = i6 = i7 = i9 = 1 and i3 = i4 = i8 = i10 = 2,
in NC([I8 ]), satisfying the conditions (i), (ii) and (iii). Remark here that, even
though i3 = i4 = i8 = i10 = 2, one cannot construct the block (i3 , i4 , i8 , i10 ) in
π (i1 ), because this block has two crossings with the first block (i1 , i2 , i5 , i6 , i7 , i9 ),
so, to avoid the crossings, we need separated blocks (i3 , i4 ), (i8 ) and (i10 ).
Now, similar to the noncrossing partition π (i1 ) for the first entry i1 of Is ,
construct noncrossing partitions,
similarly satisfying the above conditions (i), (ii) and (iii) by replacing i1 to il , for all
l = 2, . . . , s. i.e., π (il ) are the maximal partitions containing the block containing
all identical entries of Is with il , for all l = 1,. . ., s. It is
not
hard to check that if
il1 = il2 in {1, . . . , N}, as entries of Is , then π il1 = π il2 in NC ([Is ]). Thus, if
ik1 , . . . , ikn are mutually distinct in {1, . . . , N} as entries of Is , for n ≤ s, then the
corresponding partitions,
π ik1 , . . . , π ikn
“can” be distinct. Remark that, sometimes, some of them can be identically same in
NC ([Is ]); for instance, if
let
J = (1, 1, 1, 1, 2, 2) = (i1 , . . . , i6 ) ,
then
in NC ({i1 , . . . , i6 }).
Now, suppose π (il ) ∈ NC ([Is ]), for l = 1, . . . , s, is the noncrossing partition
induced by the s-tuple Is of (12),
π (il ) = {U1 , . . . , Ut },
Comparing Banach Spaces for Systems of Free Random Variables Followed by. . . 781
where t ≤ s and Uk ∈ π(Is ) are the blocks of (i) and (ii), satisfying (iii), for k = 1,
. . . , t. Then the partition π (il ) is regarded as the join partition (e.g., [17, 22, 23]),
where 1|Uk | are the maximal elements of NC (Uk ), for all k = 1, . . . , t, by regarding
each block Uk as an independent discrete finite sets.
By collecting all such partitions, define the subset ([Is ]) of NC ([Is ]) by
([Is ]) = {π (il ) : l = 1, . . . , s} .
i.e., all partitions of e ([Is ]) have their blocks with even cardinalities, where 2N =
{2n : n ∈ N}.
Let Is be in the sense of (12), and let xi1 , . . . , xis be the corresponding
semicircular elements of (A, ϕ) induced by Is , without considering repetition in
the free semicircular family {x1 , . . . , xN }. Define a free random variable X[Is ] by
def s
X[Is ] = xil ∈ (A, ϕ). (16)
l=1
s
Theorem 2 Let Is be an s-tuple (12), and let X[Is ] = xil be the corresponding
l=1
free random variable (16) of (A, ϕ). Then
ϕ (X [Is ]) = ϕθ xi1 , . . . , xis ,
θ∈ e ([Is ])
with
ϕθ xi1 , . . . , xis = c |V | , (17)
2
V ∈θ
ϕ (X [Is ]) = 0 ⇐⇒ e ([Is ]) = ∅,
{sn }∞
n=1 = {s1 , s2 , s3 , ...},
to
{xn }∞
n=0 = {x0 , x1 , x2 , ...},
X = {xn }∞
n=0 = {xn }n∈N0 ,
free-homo
(A1 , ϕ1 ) −→ (A2 , ϕ2 ). (18)
free-homo
Definition 3 Suppose (A1 , ϕ1 ) −→ (A2 , ϕ2 ) in the sense of (18), by a free-
homomorphism : A1 → A2 . If is a ∗-isomorphism, then it is called a free-
isomorphism, and (A1 , ϕ1 ) is said to be free-isomorphic to (A2 , ϕ2 ). We denote this
relation by
free-iso
(A1 , ϕ1 ) = (A2 , ϕ2 ). (19)
Comparing Banach Spaces for Systems of Free Random Variables Followed by. . . 783
By the definitions (18) and (19), if two C ∗ -probability spaces are free-
isomorphic, then they are understood as a same unital C ∗ -probability space.
Y ∪ Y ∗ of A, with Y ∗ = {y ∗ ∈ A : y ∈ Y }.
denot e
Xϕ = (X, ϕ = ϕ |X ) (20)
in (A, ϕ).
Now, let (B, ψ) be a unital C ∗ -probability space, containing a family S =
{yn }n∈Z of mutually free, |Z|-many semicircular elements, and let
denot e
Sψ = (S, ψ = ψ |S ) (21)
∗-iso ∞ ∗-iso ∞
X = C ∗ ({xn }) = C ∗ {xn } , (22)
n=0 n=0
in (A, ϕ), where () in the first ∗-isomorphic relation of (22) is the free-probabilistic
free product of [17, 22, 29, 30], and the () in the second ∗-isomorphic relation of
(22) is the pure-algebraic free product inducing the noncommutative free words in
∞
∪ {xn } = X.
n=0
def ∗-iso
X = C ∗ (X) = C ∗ {xn }n∈N0 = C ∗ ({xn }) . (23)
n∈N0
784 I. Cho and P. Jorgensen
Therefore, the first ∗-isomorphic relation of (22) holds by (23). i.e., all elements
of X are the limits of linear combinations of free reduced words (under operator
multiplication on A) in X by [17, 22, 23, 29, 30].
So, if we consider all noncommutative free words in the family X = {xn }n∈N0 ,
then they have their unique operator forms in X, which are the free reduced words
by (23). It shows that the second ∗-isomorphic relation of (22) holds, too.
By (22), one can understand the C ∗ -probability space Xϕ of (20) as an indepen-
dent free-probabilistic structure,
Sψ = C ∗ {sj } , ψ |C ∗ ({sj }) , (25)
j ∈Z j ∈Z
and
where
and
−N = {−n : n ∈ N}.
G : X → S,
by
:X→S
satisfying
n1 = n2 , n2 = n3 , . . . , nN−1 = nN in N0 ,
N
if one has a free reduced word T = xnkll ∈ Xϕ , where xn1 , . . . , xnN ∈ X, for
l=1
k1 , . . . , kN ∈ N, for N ∈ N, then
N N
(T ) = xnkll = xnkll
l=1 l=1
by the multiplicativity of
N k
= (xnl ) l
l=1
by the multiplicativity of
N k
= s l , (30)
l=1 g(nl )
in Sψ , by (29).
Since (n1 , . . . , nN ) ∈ NN
0 is an alternating N-tuple in N0 , the N-tuple,
(g(n1 ), . . . , g(nN )) ∈ ZN ,
786 I. Cho and P. Jorgensen
is an alternating N-tuple in Z, too, by (27) and (28). i.e., the formula (30) shows that
the images (T ) of all free reduced words T ∈ Xϕ with their lengths-N become
free reduced words of Sψ with the same lengths-N.
Lemma 6 The multiplicative linear transformation : X → S of (29) is a ∗-
isomorphism. i.e.,
∗-iso
X = S, (31)
∗-iso
where = means “being ∗-isomorphic to.”
Proof Remark that all elements of X (or, of S) are the limits of linear combinations
of free reduced words in X (resp., in S) by (22) (resp., by (25)). So, the multiplicative
linear transformation of (29) is bijective and bounded. Observe that, for any xn ∈
X ⊂ X, and t ∈ C,
(txn )∗ = t xn
∗
since sg(n) = sg(n) , by the semicircularity
∗
= tsg(n) = ((txn ))∗ ,
free-iso
Xϕ = Sψ . (33)
Proof By (31), there exists a ∗-isomorphism of (29) from X onto S. By (22) and
(25), it suffices to show that the ∗-isomorphism preserves the free distributions of
generators of Xϕ to those of generators of Sψ .
Let xn ∈ X ⊂ Xϕ . Then
5 Free-Distributional Data on Xϕ
Let Xϕ = (X, ϕ) be the C ∗ -probability space (25), “identified with (24) by (33),”
generated by the free semicircular family X = {xj }j ∈Z .
Theorem 8 Let Is = (i1 , . . . , is ) be an arbitrary s-tuple in Zs , for s ∈ N, like in
(12), and let π(Is ) ∈ NC ({i1 , . . . , is }) be the noncrossing partition (15) induced by
Is . If X[Is ] be a free random variable (16) of Xϕ , then the free-distributional data
ϕ (X[Is ]) is characterized by the formula (17).
7 s ]) on Xϕ are obtained
Proof Under hypothesis, the free-distributional data ϕ6(X[I
by (17), since all elements of the generator set X = xj j ∈Z of Xϕ are mutually
free, semicircular elements.
6 Certain Free-Isomorphisms on Xϕ
By (24), (25) and (33), our C ∗ -probability space Xϕ is a representative of all unital
C ∗ -probability spaces generated by mutually free, |N|-many semicircular elements.
As we assumed in Sects. 4 and 5, we let Xϕ be the C ∗ -probability space (25)
generated by a free semicircular family X = {xj }j ∈Z of mutually free, |Z|-many
semicircular elements.
6.1 Shifts on Z
h(j ) = j + 1, (34)
for all j ∈ Z.
Remark that, by (34), one can define a function h on N0 by
h = g −1 ◦ h ◦ g on N0 , (34’)
788 I. Cho and P. Jorgensen
for all n ∈ Z, where (◦) is the usual functional composition, and h−1 is the inverse
of h,
h−1 (j ) = j − 1, ∀j ∈ Z.
h(n) (j ) = j + n,
for all j, n ∈ Z. And h(n) are invertible with their inverse h(−n) , for all n ∈ Z.
Definition 9 We call the bijections h(n) of (35), the n-th shifts on Z, for n ∈ Z.
Let h(n) be the n-th shifts (35) on Z, for all n ∈ Z. Let k ∈ Z, and define a
“multiplicative” linear transformation λk acting on Xϕ by the morphism satisfying
λk xj = xh(k) (j ) = xj +k , ∀xj ∈ X ⊂ Xϕ . (36)
N n
By the multiplicativity of the morphism λk of (36), if T = x l is a free reduced
6 7 l=1 jl
words of Xϕ with its length-N in X = xj j ∈Z , then
N n N n
λk (T ) = λk xjl l = x l , (37)
l=1 l=1 jl +k
(j1 + k, . . . ,jN + k) ∈ ZN
is alternating in Z, too. So, the computation (37) says that the morphism λk of (36)
assign free reduced words to free reduced words preserving lengths in Xϕ .
Also, by (36) and (37), one has
⎧
⎪
⎪ 1Xϕ , the identity map on Xϕ if k = 0
⎪
⎪
⎪
⎪
⎨
O · λ · λPQ· ...... · λR
λ if k > 0
λ =
k
⎪
⎪ k -times
⎪
⎪ −1
· −1
· ...... · λ−1R if k < 0,
⎪
⎪
λ
O λ PQ
⎩
|k|-times
for all k ∈ Z, by (35) and (36), where (·) is the multiplication (or, composition) of
linear transformations.
Observe that, for t ∈ C, and xj ∈ X ⊂ X,
∗ ∗
λk txj = t xj +k = txj∗+k = λk (txj ) ,
implying that
∗
λk T ∗ = λk (T ) , for all T ∈ Xϕ , (38)
ϕ (X[Is ]) = ϕ λk (X[Is ]) in Xϕ ,
790 I. Cho and P. Jorgensen
by Theorem 8 (or, (17)) and (39), where X[Is ] are in the sense of (16). Thus,
ϕ (T ) = ϕ λk (T ) , for all T ∈ Xϕ ,
Group
λ = (Z, +) , the infinite abelian cyclic group, (41)
Group
where “ = ” means “being group-isomorphic.”
Proof Let λk1 , λk2 ∈ λ. Then, by (36) and (37),
in λ, for all k1 , k2 , k3 ∈ Z.
Observe that the set λ contains λ0 = 1Xϕ , the identity map on Xϕ , by (34) and
(36), satisfying
showing that every element λk ∈ λ has its unique (·)-inverse λ−k . Thus, the set λ of
(40) forms a subgroup of Aut Xϕ . Clearly,
(j ) = λj , for all j ∈ Z,
for all k ∈ Z.
Definition 12 The group λ of (40), acting on Xϕ via the group-action θ of (42), is
called the integer-shift group on Xϕ .
We here study how the integer-shift group λ of (40) affects the free probability on
Xϕ , under the group-action θ of (42).
Theorem 13 Let λ be the integer-shift group, and let θ be the group-action (42) of
λ acting on Xϕ . Then the free probability on Xϕ is preserved by θ , in the sense that:
θ λk (T ) = λk (T ) , in Xϕ ,
and hence,
ϕ θ λk (T ) = ϕ λk (T ) = ϕ (T ) ,
for all k ∈ Z, by Theorem 10. So, the action θ preserves the free probability on Xϕ .
k
Notation From below, we denote the images θ λ (T ) ∈ Xϕ of T ∈ Xϕ simply
by λk (T ), for all λk ∈ λ.
Let λ ⊂ Aut (Xϕ ) be the integer-shift group (40) acting on the C ∗ -probability
space Xϕ (via the canonical action θ of (42)) generated by the free family {xj }j ∈Z
of semicircular elements xj ’s. In this section, we construct some free random
variables in a certain C ∗ -probability space, containing Xϕ , whose free distributions
are followed by the semicircular law.
Definition 14 Let (B, ψ) be an arbitrary topological ∗-probability space. A free
random variable y ∈ (B, ψ) is followed by the semicircular law, if
n
ψ y rl
= ωn c n2 , (44)
l=1
for all (r1 , . . . , rn ) ∈ {1, ∗}n , for all n ∈ N, where ωn are in the sense of (3) for all
n ∈ N, and ck are the k-th Catalan numbers (4) for all k ∈ N0 .
By the definition (44), if a self-adjoint free random variable y is followed by the
semicircular law in (B, ψ), then it is nothing but a semicircular element. i.e., all
semicircular elements are followed by the semicircular law in the sense of (44), but
not all such free random variables are semicircular. In the text, we focus on studying
“non-self-adjoint” free random variables followed by the semicircular law.
Let Γ be an arbitrary discrete group, and let H be the group Hilbert space,
H = l 2 (Γ ) , the l 2 -space,
Comparing Banach Spaces for Systems of Free Random Variables Followed by. . . 793
and
> ?
ξg = ξg , ξg = 1, (46)
2 2
where is the infinite sum under l 2 -topology6induced
7 by (46).
Note that, for any Hilbert-space vectors ξg g∈Γ = B ∪ {e}, the following
multiplication-rule holds;
in H by (47).
The relation (48) shows that the group Γ is acting on the operator algebra B (H )
via a group-action m,
m (g) = mg ∈ B (H ) , ∀g ∈ Γ , (49)
def 6 7
M = mg ∈ B (H ) : g ∈ Γ ,
794 I. Cho and P. Jorgensen
M = C ∗ (M) of B (H ) ,
def
(50)
Λ = C ∗ (λ) in B (Hλ ) ,
def
where
Hλ = l 2 (λ) (51)
defined by
tk mλk (S) = tk λk (S) , (52)
k∈Z k∈Z
for all S ∈ Xϕ , where B Xϕ is the operator space (in the sense of [13]), consisting
of all bounded linear transformations on Xϕ , by regarding Xϕ as a Banach space
with its C ∗ -norm. Indeed, is a well-defined algebra-action of Λ acting on (the
Banach space) Xϕ , since
and
S ∗ = ( (S))∗ , (53)
by (42), (48) and (52), for all S1 , S2 , S ∈ Xϕ . i.e., by (52) and (53), all operators
of Λ are understood to be Banach-space operators acting on the Banach space, our
C ∗ -probability space Xϕ .
Proposition 17 If T = tk mλk ∈ Λ is a shift operator, and xj ∈ X is a
k∈Z
generating semicircular element of Xϕ , then
ϕ (T ) xjn = ωn c n2 tk , (54)
k∈Z
for all n ∈ N.
Proof For any n ∈ N, and xj ∈ X ⊂ Xϕ , one has
ϕ tk mλk n
xj =ϕ k n
tk λ xj
k∈Z k∈Z
by (52)
= tk ϕ xjn+k = tk ϕ xjn
k∈Z k∈Z
by (43)
= ωn c n2 tk .
k∈Z
def
X = $⊗X (55)
796 I. Cho and P. Jorgensen
def
τ (S ⊗ T ) = ϕ ( (S) (T )) = ϕ (S (T )) , (56)
denote
Xτ = (X , τ ) ,
τ (I ) = ϕ λ0 (1X ) = ϕ (1X ) = 1.
denote
uk,j = λk ⊗ xj ∈ Xτ , for k, j ∈ Z, (57)
n n
τ uk,j =τ λk ⊗ xj = τ λkn ⊗ xjn
on Xϕ , by (43), (54) and (56), since all generating shift operators λk ∈ Λ are free-
isomorphisms on Xϕ .
Lemma 19 Let uk,j = λk ⊗ xj be a generating operator (57) of the shift-
semicircular C ∗ -probability space Xτ , for k, j ∈ Z. Then
n
τ uk,j = ωn c n2 = ϕ xjn , (59)
for all n ∈ N.
Proof The free-distributional data (59) is obtained by (58).
Comparing Banach Spaces for Systems of Free Random Variables Followed by. . . 797
u0,j = λ0 ⊗ xj in Xτ ,
are followed by the semicircular law in the sense of (44), because they are
semicircular in Xτ . So, we are now interested in the cases where
k = 0 in Z.
∗ ∗
uk,j = λk ⊗ xj∗ = λ−k ⊗ xj = u−k,j , in Xτ . (60)
i.e., if k = 0, then the generating operators uk,j are not self-adjoint in Xτ , and
hence, they cannot be semicircular in Xτ , for all j ∈ Z.
Lemma 21 Let uk,j ∈ Xτ be a generating free random variable for k ∈ Z×
= Z \ {0}, and j ∈ Z. Then
n
τ u∗k,j = ωn c n2 = ϕ xjn , (61)
for all n ∈ N.
798 I. Cho and P. Jorgensen
∗
Proof If k = 0 in Z× and j ∈ Z, and hence, if ukj = u−k
j in Xτ , then
∗ n n
τ ukj =τ u−k
j = ϕ xjn−kn = ωn c n2 = ϕ xjn ,
in Xτ . First, observe that, if (r1 , . . . , rn ) ∈ {1, ∗}n is a “mixed” n-tuple of {1, ∗},
for n ∈ N>1 = N \ {1}, in the sense that: there exists at least one i0 ∈ {r1 , . . . , rn },
such that i0 = rm in {1, ∗}, for some m ∈ {1, . . . , n}, then
n
rl
uk,j = λ#(1)k−#(∗)k ⊗ xjn ,
l=1
where
and
n
r
uk,j l = λ(#(1)−#(∗))k ⊗ xjn ,
l=1
by (63)
= ϕ xjn+(#(1)−#(∗))k = ϕ xjn = ωn c n2 ,
for all n ∈ N, by (59) and (61). Therefore, the free-distributional data (64) holds.
By (59), (61) and (64), we obtain the following result.
Theorem 23 Every generating free random variable uk,j of (57) are followed by
the semicircular law in the sense of (44) in the shift-semicircular C ∗ -probability
space Xτ , for all k, j ∈ Z. i.e.,
n
rl
τ uk,j = ωn c n2 = ϕ xjn , (65)
l=1
(1, 1, . . . , 1) , or (∗, ∗, . . . , ∗) .
Then, by (59) and (61), the free-distributional data (65) holds. Meanwhile, if
(r1 , . . . , rn ) ∈ {1, ∗}n is mixed for n ∈ N>1 , then the formula (65) holds too, by
(64).
6 Therefore,
∗the free-distributional
7 data (65) holds as the joint free moments of
uk,j , uk,j = u−k,j , on Xτ . Equivalently, the generating free random variable
uk,j is followed by the semicircular law in Xτ , for all k, j ∈ Z.
The above theorem shows that there do exist free random variables in a C ∗ -
probability space followed by the semicircular law in the sense of (44).
Theorem 24 Let (B, ψ) be a unital C ∗ -probability space containing mutually free,
semicircular elements {y1 , . . . , yN }, for N ∈ N∞ = N ∪ {∞}. Then there exists a
800 I. Cho and P. Jorgensen
C ∗ -probability space (B, τ ) and a free random variable y ∈ (B, τ ), such that y is
followed by the semicircular law.
Proof Suppose a unital C ∗ -probability space (B, ψ) contains mutually free |N|-
many semicircular elements, Y = {y1 , y2, y3 , ...}. Then the C ∗ -subalgebra
C ∗ (Y)
∗ ∗
of B induces a C -probability space C (Y) , ψ = ψ |C ∗ (Y ) , which is free-
isomorphic to our C ∗ -probability
6 7 space Xϕ = (X, ϕ) generated by the free
semicircular family X = xj j ∈Z , by (33). And hence, there exists the correspond-
ing shift-semicircular C ∗ -probability space Xτ = (X , τ ), containing infinitely
many generating free random variables uk,j of (57) followed by the semicircular
law by (65). i.e., if
N = |N| = ∞, in N∞ ,
and
τ = ψ ∞ , on B,
where () is the free product of C ∗ -algebras (e.g., [17, 30]). Remark that all free
factors {B[i]}∞
i=1 , identified with BN , are free from each other (e.g., [30]) in (B, τ ),
and hence, the C ∗ -probability space (B, τ ) contains its free semicircular family,
∞
Y = {yi1 , . . . , yiN } ,
i=1
in a free factor B[i] = BN , for all i ∈ N. Under the possible rearrangement, one can
let
6 7
Y = yj j ∈N .
0
Therefore, by (33) and (65), this theorem also holds even if N < ∞ in N∞ .
In conclusion, if a unital C ∗ -probability space (B, ψ) contains mutually free N-
many semicircular elements for N ∈ N∞ , then one can have free random variables
in a certain C ∗ -probability space (B, τ ), followed by the semicircular law.
The above theorem shows not only that there do exist free random variables
followed by the semicircular law, but also how to construct them. It also shows
there are sufficiently many such free random variables. Motivated by Theorem 24,
one can verify that, whenever a semicircular element x, and a free-isomorphism
β exist for a unital C ∗ -probability space containing x, one can construct the free
random variables,
4 5
βk ⊗ x
k∈Z
Xτ = (Λ ⊗ X, τ ) ,
generated by the free random variables followed by the semicircular law. Through-
out this section, we let
denote
ul = ukl ,jl = λkl ⊗ xjl ∈ Xτ (66)
in Xτ .
Theorem 25 Let xj1 , . . . , xjN ∈ X be generating semicircular elements of the C ∗ -
probability space Xϕ (where j1 , . . . , jN are not necessarily distinct in Z), and let
λk1 , . . . , λkN ∈ λ be integer-shifts generating the shift algebra Λ (where k1 , . . . , kN
802 I. Cho and P. Jorgensen
are not necessarily distinct in Z), inducing the generating free random variables
N
u1 , . . . , uN of (66) in the shift-semicircular C ∗ -probability space Xτ . If w = xjl
l=1
N
is a free random variable of Xϕ , and if W(r1 ,...,rN ) = url l is a free random variable
l=1
of Xτ , where url l ∈ Xτ are in the sense of (66), or (67), for l = 1, . . . , N , then
τ W(r1 ,...,rN ) = ϕ (w) , (68)
ϕ (w) = 7, in C,
N
r
W(r1 ,...,rN ) = ul l ∈ Xτ ,
l=1
in Z, for all l = 1, . . . , N .
Then, there exists k(r1 ,...,rN ) ∈ Z, such that
N
λil = λk(r1 ,...,rN ) ∈ λ,
l=1
denote
ukl ,j = β kl ⊗ yil ∈ (ΛB ⊗ B, τ )
τ (S ⊗ T ) = ψ (S(T )) ,
for all (r1 , . . . , rn ) ∈ {1, ∗}n , for all n ∈ N, where the right-hand side of (70) is
determined by (17).
Proof The proof of the free-distributional data (70) is similar to that of (68) by
Theorem 24. Indeed, there exists k0 ∈ Z, such that
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
n
n
n
n
ukl ,il = ⎝ β εl kl ⎠ ⊗ ⎝ yil ⎠ = β k0 ⊗ ⎝ yil ⎠ ,
rl
l=1 j =1 j =1 j =1
in (ΛB ⊗ B, τ ), where
1 if rl = 1
εl =
−1 if rl = ∗,
804 I. Cho and P. Jorgensen
Xτ = (Λ ⊗ X, τ ) ,
for all (r1 , . . . , rn ) ∈ {1, ∗}n and (i1 , . . . , in ) ∈ {1, . . . , s}n , for all n ∈ N, where
ϕ
k•τ (.) (respectively, k• (.)) is the free cumulant on Xτ (respectively, on Xϕ ) in terms
of the linear functional τ on Xτ (respectively, ϕ on Xϕ ), by the semicircularity (7)
of the generating operators xj ∈ X of Xϕ .
Proposition 27 Let uk,j ∈X be a generating free random variable of Xτ , followed
by the semicircular law, for k, j ∈ Z. Then
knτ urk,j
1
, . . . , urk,j
n
= δn,2 = knϕ xj , . . . , xj , (72)
by (71) and (68), for all (r1 , . . . , rn ) ∈ {1, ∗}n . In particular, the second equality
holds by the freeness of xj1 and xj2 in Xϕ . Indeed, by assumption, the semicircular
6elements
7 xj1 and xj2 are distinct in the generating free semicircular family X =
xj j ∈Z in Xϕ , implying the freeness of them (e.g., [22, 23]). Therefore,
for all mixed n-tuples (l1 , . . . , ln ) ∈ {1, 2}n , for all (r1 , . . . , rn ) ∈ {1, ∗}n , for all
n ∈ N>1 . Equivalently, two subsets,
6 7 6 7
u1 , u∗1 and u2 , u∗2
are free in Xτ , i.e., if j1 = j2 in Z, then two free random variables u1 and u2 are
free in Xτ (e.g., [30]).
(⇐) Assume now that j1 = j = j2 in Z, and hence, ul = ukl ,j = λkl ⊗ xj ∈ X
in Xτ . Then, for any (mixed, or non-mixed) (l1 , . . . , ln ) ∈ {1, 2}n , for n ∈ N, we
806 I. Cho and P. Jorgensen
have
⎛ ⎞
⎜ ⎟
knτ url11 , . . . , urlnn = knϕ ⎝xj , xj , . . . , xj ⎠ = δn,2 ,
O PQ R
n-times
even though l1 = l2 in {1, 2}. It shows that such two free random variables u1 and
u2 are not free in Xτ . i.e., if j1 = j2 in Z, then u1 and u2 are not free in Xτ .
The above theorem shows that the generator set X of Xτ is decomposed to be
X = Xj in Xτ ,
j ∈Z
with
4 5
Xj = uk,j = λk ⊗ xj ∈ X : k ∈ Z , ∀j ∈ Z, (73)
∗-iso
Xτ = C ∗ Xj ,
j ∈Z
where
4 5
Xj = uk,j = λk ⊗ xj ∈ X : k ∈ Z , ∀j ∈ Z, (74)
By the structure theorem (74), one obtains the following result, too.
Theorem 31 The shift-semicircular C ∗ -probability space Xτ satisfies that
∗-iso ∗
Xτ = Λ ⊗ CX {xj } ,
j ∈Z
and hence,
∗-iso ∗
Xτ = Λ Λ ⊗ CX {xj } , (75)
j ∈Z
∗
where CX (Z) are the C ∗ -subalgebras of Xϕ generated by the subsets Z ∪ Z ∗ of
Xϕ , and where “Λ ” is the amalgamated free product with its amalgamation over
the group C ∗ -algebra Λ (in the sense of [22]).
∗-iso
Proof By (74), we have Xτ = C ∗ Xj , where Xj are in the sense of (73)
j ∈Z
for all j ∈ Z. So, by definition,
∗-iso ∗ 6 7 6 7
C ∗ Xj = CB(H ∗
) (λ) ⊗ CX xj
∗
= Λ ⊗ CX xj ,
6 7
where λ = λk k∈Z is the integer-shift group, generating the group algebra Λ in the
operator algebra B (H ), where H is the group Hilbert space (45) of λ. Therefore,
the first ∗-isomorphic relation of (75) holds.
By the definition of amalgamated free products with amalgamations of [22], the
second ∗-isomorphic relation (75) holds, too, since
∗-iso ∗
∗-iso ∗
Xτ = Λ ⊗ CX {xj } = Λ⊗ CX {xj } ,
j ∈Z j ∈Z
6
∗ {x }
7
by understanding Λ as the common C ∗ -subalgebras Λ ⊗ CX j j ∈Z
of Xτ .
X = Λ ⊗ X,
6 7
X = xj j ∈Z , equipped with the linear functional τ , satisfying
τ (T ⊗ S) = ϕ (T (S)) ,
denote
uk,j = λk ⊗ xj ∈ Xτ , for all k, j ∈ Z, (76)
t
8.1 Banach-Space Operators Ts,l ∈ B (Xτ )
Let B (Xτ ) be the operator space consisting of all Banach-space operators acting
on our shift-semicircular C ∗ -probability space Xτ , by regarding X = Λ ⊗ X as a
Banach space. Define an element Ts,lt
∈ B (Xτ ) by
def
t
Ts,l = Mt λs ⊗ λl , on Xτ ,
satisfying
t
Ts,l uk,j = Mt λs ⊗ λl λk ⊗ xj
(77)
= tλs λk ⊗ λl xj = tλs+k ⊗ xj +l ,
i.e.,
t
Ts,l uk,j = t λk+s ⊗ xj +l = tuk+s,j +l , (78)
is the generator set of all generating free random variables (76) of Xτ . In (77), the
tensor factor Mt λs ∈ B (Λ) of Ts,lt is a multiplication operator acting on the shift-
Mt λs (S) = tλs S, in Λ, ∀S ∈ Λ,
Comparing Banach Spaces for Systems of Free Random Variables Followed by. . . 809
and
by (77) and (78). Note that an operator uk+s,j +l ∈ X , in the far-right-hand side
is a generating free random variable of Xτ , followed by the semicircular law, by
(65). So, the element u ∈ Xτ is a scalar multiple of uk+n,j +l , and hence, the free
distribution of u may be affected by the semicircularity. Indeed, observe that, for
any
one has
n
n
n
τ uri =τ t ri urk+s,j
i
+l
i=1 i=1 i=1
where
t if ri = 1
t ri =
t if ri = ∗,
810 I. Cho and P. Jorgensen
= t #(1) t #(∗) ωn c n2 ,
where
and
for all (r1 , . . . , rn ) ∈ {1, ∗}n , for all n ∈ N, by (79). Moreover, since t ∈ R× ,
n
τ uri
= ωn t #(1)+#(∗) c n2 = ωn t n c n2 ,
i=1
implying that
n n
2
τ uri = ωn t 2 c n2 , (82)
i=1
where
N
N
so = si , and lo = li , in Z,
i=1 i=1
in B (Xτ ). i.e.,
N
t
Tsii,li = Tstoo,lo in B (Xτ ) ,
i=1
with
N
N
N
to = ti , so = si , and lo = li . (83)
i=1 i=1 i=1
N
to = ti ∈ C× ,
i=1
such that
n
ri
τ TN uk,j = ωn to#(1) to #(∗) c n2 , (84)
i=1
for all (r1 , . . . , rn ) ∈ {1, ∗}n , for all n ∈ N, for any fixed generating free random
variables uk,j ∈ X of Xτ , for all k, j ∈ Z, where #(1) and #(∗) are in the sense of
(79).
Proof Under hypothesis, we have
TN = Tstoo,lo ∈ B (Xτ ) ,
deform the original free-distributional data on Xτ . Recall again that all generating
free random variables
4 5
X = uk,j = λk ⊗ xj : k, j ∈ Z
denote
u0 = uk0 ,j0 ∈ X (85)
Mu0 (T ) = u0 T in Xτ , ∀T ∈ Xτ . (86)
Mw (y) = wy, ∀y ∈ Xτ ,
in Xτ , by (85).
Let’s consider the tensor-factor xj0 xj ∈ Xϕ of the image (87) of Mu0 uk,j ∈
Xτ . Suppose first that j = j0 in Z. Then
xj0 xj = xj20 ∈ Xϕ ,
So, if uj0 = Mu0 uk,j0 = λk0 +k ⊗ xj20 ∈ Xτ , then
n
τ urji0 = τ λ#(1)−#(∗) ⊗ xj2n
0
i=1 (89)
= ϕ λ#(1)−#(∗) xj2n
0
= ϕ xj2n
0
= cn ,
by (88), for all (r1 , . . . , rn ) ∈ {1, ∗}n , for all n ∈ N, since λl ∈ λ ⊂ Λ are free-
isomorphisms on Xϕ , for all l ∈ Z.
Lemma 39 Let Mu0 ∈ B (Xτ ) be the multiplication operator (86) with its symbol
u0 ∈ Xτ of (85). Then, for any generating free random variables uk,j0 ∈ X ⊂ Xτ ,
for all k ∈ Z, and fixed j0 ∈ Z, we have
n
r
τ Mu0 uk,j0 i = cn , (90)
i=1
by (87)
n
ri
ri (k0 +k)
=τ λ ⊗ xj0 xj
i=1
where
k0 + k if ri = 1
ri (k0 + k) =
− (k0 + k) if ri = ∗,
and
⎧
ri ⎨ xj0 xj if ri = 1
xj0 xj =
⎩
xj xj0 if ri = ∗,
816 I. Cho and P. Jorgensen
where
since λN(r1 ,...,rn ) is a free-isomorphism on Xϕ . Note here that the C-quantity (91) is
characterized by Theorem 8, or (17).
Lemma 40 Let u0 = Mu0 uk,j ∈ Xτ , where Mu0 ∈ B (Xτ ) is the multiplication
operator (86) with its symbol u0 = uk0 ,j0 ∈ X ⊂ Xτ of (77), and suppose j = j0
in Z. Then
n
n ri
τ ur0i =ϕ xj0 xj , (92)
i=1 i=1
where j = j0 in Z. Then the images uj0 and u0 are no longer followed by the
semicircular law. In particular,
n
τ urji0 = cn , the n-th Catalan number, (93)
i=1
Comparing Banach Spaces for Systems of Free Random Variables Followed by. . . 817
and
n
n ri
τ ur0i =ϕ xj0 xj ,
i=1 i=1
characterized by Theorem 8 (or, (17)), for all (r1 , . . . , rn ) ∈ {1, ∗}n , for all n ∈ N.
Proof The free-distributional data (93) are obtained by (90) and (92). Therefore,
the free random variables uj0 and u0 are not followed by the semicircular law in
Xτ .
The above theorem illustrates how our multiplication operators deform the free
probability on the shift-semicircular C ∗ -probability space Xτ . In particular, the
generating free random variables of Xτ , followed by the semicircular law, are no
longer followed by the semicircular law, under the action of our multiplications
operators by (93). For example, if
u0 = Mu0 uk,j = λk0 +k ⊗ xj0 xj ∈ Xτ
since ϕ(xj ) = ω1 c 1 = 0, for all n ∈ N, by (93). Note that, the formula (94) is
2
obtained by Theorem 8. Indeed, the free reduced word,
n
W = xj0 xj = xj0 xj xj0 xj ...xj0 xj ∈ Xϕ
in the lattice NC ({i1 , . . . , i2n }) of all noncrossing partitions over {i1 , . . . , i2n }.
Similar to (94), under the same hypothesis, we have
n
τ u∗0 = 0, for all n ∈ N,
818 I. Cho and P. Jorgensen
since
∗ n n
u0 = xj xj0 = xj xj0 xj xj0 ...xj xj0 ∈ Xϕ
= τ λ0 ⊗ xj xj20 xj = ϕ xj xj20 xj
where xj xj20 xj = xj xj0 xj0 xj ∈ Xϕ is a free reduced word with its length-3
by (93), because the free random variable w = xj xj20 xj ∈ Xϕ induces its corre-
sponding noncrossing partition,
πw = {(i1 , i4 ), (i2 , i3 )} ,
since
Mu1 Mu2 uk,j = Mu1 λk2 λk ⊗ xjl xj
= λk1 λk2 λk ⊗ xj1 xj2 xj = λk1 λk2 λk ⊗ xj1 xj2 xj
= λk1 λk2 ⊗ xj1 xj2 λk ⊗ xj = (u1 u2 ) uk,j
= Mu1 u2 uk,j ,
Let Mui = Muki ,ji ∈ B (Xτ ) be the multiplication operators with their symbols
ui = uki ,ji ∈ X , for i = 1, . . . , N , for N ∈ N. Then, by (95),
N
Mui = M
N = M⎛
N
⎞
. (96)
i=1
ui k
⎜ i=1 i ⎟
N
i=1 ⎝λ ⎠⊗ xji
i=1
by (96)
⎛
N
⎞ N
k+ ki
= ⎝λ i=1 ⎠⊗ xji xj , (97)
i=1
in Xτ .
Theorem 42 Let Mui = Muki ,ji ∈ B (Xτ ) be the multiplication operators with
their symbols uki ,ji ∈ X , for i = 1, . . . , N , for N ∈ N, and let uk,j ∈ X be an
N
arbitrary generating free random variable of Xτ . If M = Mui ∈ B (Xτ ), and if
i=1
u = M uk,j ∈ Xτ , then
N
τ (u) = ϕ xj1 xj = ϕ xj1 xj2 ...xjN xj , (98)
i=1
where
N
ko = k + ki ∈ Z,
i=1
820 I. Cho and P. Jorgensen
and
denote
M = M⎛
N
⎞ = M ko ⊗xjN , on Xτ ,
⎜ i=1
ki
⎟ λ
⎝λ ⎠⊗xjN
with
N
λko ∈ Λ, with ko = ki ∈ Z, (100)
i=1
and xjN ∈ Xϕ is the free reduced word with its length-1 for a fixed generating
semicircular element xj of Xϕ .
So, one has
u = M uk,j = λko λk ⊗ xjN xj = λko +k ⊗ xjN+1 ,
n
for K = εi (ko + k) ∈ Z, with
i=1
1 if ri = 1
εi =
−1 if ri = ∗,
References
1. M. Ahsanullah, Some inferences on semicircular distribution. J. Stat. Theory Appl. 15(3), 207–
213 (2016)
2. H. Bercovici, D. Voiculescu, Superconvergence to the central limit and failure of the Cramer
theorem for free random variables. Probab. Theory Related Fields 103(2), 215–222 (1995)
3. M. Bozejko, W. Ejsmont, T. Hasebe, Noncommutative probability of type D. Int. J. Math.
28(2), 1750010, 30 (2017)
4. M. Bozheuiko, E.V. Litvinov, I.V. Rodionova, An extended anyon Fock space and non-
commutative Meixner-type orthogonal polynomials in the infinite-dimensional case. Usp.
Math. Nauk. 70(5), 75–120 (2015)
5. I. Cho, Semicircular families in free product banach ∗-algebras induced by p-adic number
fields over primes p. Complex Anal. Oper. Theory 11(3), 507–565 (2017)
6. I. Cho, Acting semicircular elements induced by orthogonal projections on von Neumann
algebras. Mathematics 5, 74 (2017). https://doi.org/10.3390/math5040074
7. I. Cho, Semicircular-like laws and the semicircular law induced by orthogonal projections.
Complex Anal. Oper. Theory 12, 1657–1695 (2018)
8. I. Cho, Free stochastic integrals for weighted-semicircular motion induced by orthogonal
projections, in Advanced Topics in Mathematical Analysis. Monograph Ser., Appl. Math. Anal:
Theo., Methods & Appl. (Taylor & Fransis, Abington, 2019)
9. I. Cho, J. Dong, Catalan numbers and free distributions of mutually free multi semicircular
elements. Complex Anal. Oper. Theory. Preprint (2021). Submitted
10. I. Cho, P.E.T. Jorgensen, Semicircular elements induced by p-adic number fields. Opusc. Math.
35(5), 665–703 (2017)
11. I. Cho, P.E.T. Jorgensen, Banach ∗-algebras generated by semicircular elements induced by
certain orthogonal projections. Opusc. Math. 38(4), 501–535 (2018)
12. I. Cho, P.E.T. Jorgensen, Semicircular elements induced by projections on separable Hilbert
spaces, in Operator Theory: Advances & Applications (OT, vol. 275). Linear Systems, Signal
Processing & Hypercomplex Analysis (2019), pp. 167–209
13. A. Connes, Noncommutative Geometry (Academic Press, San Diego, CA, 1994). ISBN: 0-12-
185860-X
14. P.R. Halmos, Hilbert Space Problem Books. Grad. Texts in Math., vol. 19 (Springer, Berlin,
1982). ISBN: 978-0387906850
822 I. Cho and P. Jorgensen
15. I. Kaygorodov, I. Shestakov, Free generic Poisson fields and algebras. Commun. Algebra 46(4)
(2018). https://doi.org/10.1080/00927872.2017.1358269
16. L. Makar-Limanov, I. Shestakov, Polynomials and poisson dependence in free Poisson algebras
and free poisson fields. J. Algebra 349(1), 372–379 (2012)
17. A. Nica, R. Speicher, Lectures on the Combinatorics of Free Probability, London Math. Soc.
Lecture Note Ser., vol. 335, 1st edn. (Cambridge Univ. Press., Cambridge, 2006). ISBN-
13:978-0521858526
18. I. Nourdin, G. Peccati, R. Speicher, Multi-Dimensional Semicircular Limits on the Free Wigner
Chaos. Progr. Probab., vol. 67 (2013), pp. 211–221
19. V. Pata, The central limit theorem for free additive convolution. J. Funct. Anal. 140(2), 359–380
(1996)
20. F. Radulescu, Random matrices, amalgamated free products and subfactors of the C ∗ -algebra
of a free group of nonsingular index. Invent. Math. 115, 347–389 (1994)
21. F. Radulescu, Free Group factors and Hecke operators, notes taken by N. Ozawa, in Proceed.
24-th Conference in Oper. Theo., Theta Advanced Series in Math. (Theta Foundation, 2014)
22. R. Speicher, Combinatorial theory of the free product with amalgamation and operator-valued
free probability theory. Am. Math. Soc. Mem. 132(627) (1998)
23. R. Speicher, A conceptual proof of a basic result in the combinatorial approach to freeness.
Infin. Dimens. Anal. Quantum Probab. Relat. Top. 3, 213–222 (2000)
24. R. Speicher, U. Haagerup, Brown’s spectrial distribution measure for R-diagonal elements in
finite Von Neumann algebras. J. Funct. Anal. 176(2), 331–367 (2000)
25. R. Speicher, T. Kemp, Strong Haagerup inequalities for free R-diagonal elements. J. Funct.
Anal. 251(1), 141–173 (2007)
26. V.S. Vladimirov, p-Adic quantum mechanics. Commun. Math. Phys. 123(4), 659–676 (1989)
27. V.S. Vladimirov, I.V. Volovich, E.I. Zelenov, p-Adic Analysis and Mathematical Physics. Ser.
Soviet & East European Math., vol. 1 (World Scientific, Singapore, 1994). ISBN: 978-981-02-
0880-6
28. D. Voiculescu, Free probability and the Von Neumann algebras of free groups. Rep. Math.
Phys. 55(1), 127–133 (2005)
29. D. Voiculescu, Aspects of free analysis. Jpn. J. Math. 3(2), 163–183 (2008)
30. D. Voiculescu, K. Dykema, A. Nica, Free Random Variables. CRM Monograph Series, vol. 1.
(Am. Math. Soc., Providence, 1992). ISBN-13: 978-0821811405