\stackMath
Chian Yeong ChuahDepartment of Mathematics
The Ohio State University
231 West 18th Avenue
Columbus, OH 43210-1174, USA
ORCid:0000-0003-3776-6555chuah.21@osu.edu,Zhen-Chuan LiuDepartamento de Matemáticas
Universidad Autónoma de Madrid
C/ Francisco Tomás y Valiente, 7 Facultad de Ciencias, módulo 17, 28049 Madrid,Spain.
ORCid: 0000-0002-6092-5473liu.zhenchuan@uam.esand Tao MeiDepartment of Mathematics
Baylor University
1301 S University Parks Dr, Waco, TX 76798, USA.
ORCid: 0000-0001-6191-6184tao_mei@baylor.edu
Abstract.
We give a direct proof of the operator valued Hardy-Littlewood maximal inequality for , which was first proved in [7].
Key words and phrases:
Maximal Inequality, Schatten -class, von Neumann Algebra, -Space
2010 Mathematics Subject Classification:
Primary: 46B28, 46L52. Secondary: 42A45.
1. Introduction
The operator valued Hardy-Littlewood maximal inequality ([7]) has become a basic tool in the study of noncommutative analysis. See e.g. [2, 4, 8] for the recent works which apply this inequality.
The original proof of the inequality contained in [7] reduces the problem to the martingale case where the noncommutative Doob’s maximal inequality due to M. Junge ([5]) is applied. G. Hong reproved this maximal inequality in [3]. He follows Stein’s idea of dominating the Hardy-Littlewood maximal function by maximal averages of heat semigroup operators and applies the noncommutative maximal ergodic theory due to M. Junge and Q. Xu ([6]). Both proofs are indirect which prevent researchers, who are not familiar with the terminology of noncommutative analysis from a good understanding of the theorem. The purpose of this article is to provide a direct and more understandable proof of this inequality for the case .
Recall that, for a locally integrable function , the Hardy-Littlewood maximal function is defined as
where is the interval centered at with length The classical Hardy-Littlewood maximal inequality states that
(1.1) |
for all .
Let be a Banach space, for X-valued functions , their maximal function can be defined by considering the maximal function of the norm of ,
(1.2) |
Apply the classical Hardy-Littlewood maximal inequality to the function , one obtains that
(1.3) |
for all . A shortcoming is that this type of maximal function is scalar-valued and may lose a lot of information of that originally carried.
When with a measurable space, one may define a -valued maximal function as
(1.4) |
and deduce from (1.3) that
(1.5) |
Note that in (1.4), is a -valued function that dominates the average of pointwisely, while in (1.2), is merely a scalar valued function that carries much less information. Can similar results like (1.4) and (1.5) hold for -valued functions when the Banach space is not equipped with a total order but still has a reasonable partial order, e.g. for the Schatten classes?Based on G. Pisier’s work on operator spaces and M. Junge and Junge/Xu’s work on the theory of noncommutative martingales, T. Mei ([7]) proved a maximal inequality like (1.5) for being the noncommutative -spaces. In the case that the Schatten -class and for some , this operator-maximal inequality says that there exists a -valued function such that
(i) as operators for all , i.e. is a self adjoint positive definite operator almost everywhere.
(ii) There exists an absolute constant such that
(1.6) |
A main obstacle for the proof of (1.6) is the lack of a total order. This was overcome by M. Junge using G. Pisier’s duality result for the operator spaces and . This article aims togive a direct proof for (1.6) and a more understandable proof of Pisier’s duality result (Lemma 2.2) for analysts who are not familiar with operator spaces.
2. Preliminary
Let be a separable Hilbert space. We denote the space of bounded linear operators on by . For , we denote by the adjoint operator of , and define by the functional calculus for . We say that is self-adjoint if . We say a self-adjoint operator is positive, denoted by if
(2.1) |
for all . This is equivalent to saying that for some . For two self-adjoint , we write if .The Schatten -classes , are the spaces of so that
Here, is the usual trace for . TheSchatten classes share many properties with the spaces of sequences. In particular, are Banach spaces for and (resp. ) is the dual space of (resp. for ), via the isometric isomorphism,
for (resp. ).
2.1. Noncommutative spaces
A von Neumann algebra, by definition, is a weak- closed subalgebras of .The completeness according to the weak -topology of ensures that it contains the spectral projections of its self-adjoint elements. which is isometrically isomorphic to the subalgebras of the diagonal operators and itself are two basic examples of von Neumann algebras.The usual trace on is a linear functional on the weak- dense subspace satisfying the following properties,
- i)
Tracial: ,
- ii)
Faithful: if and then ,
- iii)
Lower semi-continuous: when is increasing,
- iv)
Semifinite: for any , there exists such that .
This leads to defining semifinite von Neumann algebras as those equipped with a trace , which is an unbounded linear functional satisfying (i)-(iv) for belonging to a weak dense subspace. Note the restriction of the usual trace of on may not be semifinite, and not every von Neumann algebras is semifinite. Given such a pair (which is usually viewed as a noncommutative - space) the noncommutative spaces associated to it are the completion of with finite quasi norm
where is constructed via functional calculus. We set .
The Schatten -classes and the spaces on a semifinite measure space are examples of noncommutative spaces associated with and respectively. Another basic example arises from group von Neumann algebras. Every commutative semifinite von Neumann algebra is isometrically isomorphic to the space of essentially bounded functions on some semifinite measure space .Many basic properties of extend to . In particular, one has the Hölder’s inequality which states that
for . The interpolation properties of extend to as well, and the duality properties of extend to for the range . The elements in may not belong to or in general. They can be understood as unbounded operators affiliated with .For self-adjoint elements in , we say is positive, denoted by if (2.1) holds (the quantity may be though). We write if .
Define as the space of all sequences in such that
(2.2) |
Define as the space of all sequences in such that
(2.3) |
We have the duality that
for , via the isometric isomorphism
On the other hand, isometrically embeds into the dual of as a weak -dense subspace via the same isomorphism for the same range and relation of . We refer to the survey paper [11] for more information on noncommutative spaces.
2.2. Pisier’s norm
Given two positive operators e.g. , the expression does not make sense unless commutes so that the least upper element exists.Nevertheless, the following is a reasonable expression for for sequences of positive elements in ,
(2.4) |
For sequences of positive elements in such that converges in , let
(2.5) |
Here the right hand side is an increasing sequence. G. Pisier ([9] and M. Junge ([5] proved that the expressions (2.4) and (2.5) extend to Banach space norms.111The obvious extension for sequences fails the triangle inequality. Furthermore, they proved a duality result for these two norms (see Lemma 2.2), which is the key for Junge’s proof of the noncommutative Doob’s maximal inequality.
Define as the space of all sequences in which admits a factorization
(2.6) |
where and belongs to the unit ball of .Given , define
(2.7) |
where the infimum is taken over all factorizations of as in (2.6). We denote the subset of consisting of all positive sequences by . Next, we denote the subspace of consisting of all sequences such that for by .
Define as the space of all sequences in which admits a factorization
(2.8) |
such that the series and converges in Given , define
(2.9) |
where the infimum is taken over all possible factorizations (2.8).222There are slightly different definition of the space in the literature. Here, we use the definition given in [10]. We denote the subset of consisting of all positive sequences by .Define to be the space of all sequences with for .G. Pisier ([9]) proved that (2.7) and (2.9) are norms extending (2.4) and (2.5).
Lemma 2.1 (Pisier [9]).
For sequences , we have
(2.10) |
in the sense that both sides are equally finite or both sides are infinite, and
(2.11) |
Proof.
We prove (2.10) first. The left hand side is obviously smaller because we may choose for . To prove that the right hand side is smaller, we assume and assume that there exists a factorization that such that
Then, because. So, converges in because and do. Moreover,
We conclude by taking
For (2.11), assuming , we denote by the projection onto the kernel of . Then, . Let . Then, belongs to the unit ball of and .We see that the first inequality holds. Next, we show the second inequality. We assume that and has a factorization with and .We write with contractions , and consider the new decomposition with . Noting that , we have
with Taking , we conclude (2.11).∎
The following lemma is another key to understanding the proof of the operator Hardy-Littlewood maximal inequality.The result was proved by G. Pisier ([9, 5]). We include an argument for the case of finite sequences below.
Lemma 2.2 ([9, 5]).
The norms (2.4) and (2.5) are in duality. More precisely, for ,
- (i)
For any -tuple in and , we have
(2.12) - (ii)
For any bounded sequence in with , we have
(2.13) - (iii)
embeds isometrically into the dual space of for via the isomorphism
For any in such that for finite positive sequences , there is a (unique) positive sequence in with
such that for any and any ,
Proof.
(i). Let . First, we prove that
(2.14) |
Consider a factorization where and belongs to the unit ball of .Also consider a factorization of where .Then, by Hölder’s inequality and the Cauchy-Schwarz inequality,
Hence, we proved the one side inequality of (i) and (ii) and the first half of (iii).
Now, suppose .Choose with for all . Note
Thus,
Therefore, we proved the other direction of (i).
We now prove the other direction of (ii). By the Hahn-Banach theorem, for any , there exists , such that and .Since is a subspace of , there exists such that. Since the unit ball of is weak -dense in the unit ball of and , we conclude that for any , there exists a in the form
such that and with . On the other hand, we know from (i) that,
We obtain
(2.15) |
We then conclude (ii) by letting .
We now prove the other direction of (iii).Note that is a sub-linear vector space of equipped with a larger norm. So, for any bounded linear functional , its restriction on defines a bounded linear functional . We conclude that there exists such that
(2.16) |
for all . In particular, the expression (2.16) holds for any finite sequences . By (ii), we have
∎
3. Operator-Maximal Inequality
Let be a semifinite von Neumann algebra, e.g. or . Let be the associated noncommutative spaces, e.g. or the Schatten classes .Let be the space of all -valued Bochner-measurable functions on the real line such that
We prove the following operator Hardy-Littlewood maximal inequality for where . The corresponding result for is trivial.
Theorem 3.1.
Given for some , there exists a -valued Bochner-measurable function such that
(i) as operators for all , i.e. almost everywhere.
(ii) There exists an absolute constant such that
(3.1) |
In order to prove main theorem, we prove the dual form of Theorem 3.1. Let be the von Neumann algebra tensor product equipped with the semifinite trace . Then, coincides with the noncommutative spaces associated with the pair for . Let be the averaging operator on defined by
It is easy to verify that is a family of operators from satisfying
- •
;
- •
if ;
- •
for and any .
Lemma 3.2.
For any finite sequence with all , we have
(3.2) |
Proof.
Given a positive sequence with only finitely many non-zero terms and a bijection on , we have that for ,
Note that . So, we have that
By the tracial property of , we have that for any . So
Now, taking the supremum over all bijections on both sides, we get
By dividing the finite number on both sides, we get (3.2).∎
Note that is a family of positive-preserving contractions on . Lemma 3.2 holds trivially if we replace with . We show that this remains true if we replace with for all .We need the following Cauchy-Schwartz inequality. We include a proof for completeness.
Lemma 3.3.
Suppose . Then,we have
(3.3) |
In particular,
(3.4) |
Proof.
Let . Then we have
This implies that
Then, by [1, Prop. 1.3.2], there exists a contraction such that .Thus by Hölder’s inequality,
The lemma is proved.∎
Lemma 3.4.
Under the same assumption of Lemma 3.2, we have that,
(3.5) |
for all finite sequences , , .
Proof.
Assume that . Let