1. Introduction
Analyzing the limit process of Cesàro averages along certain subsequences is a typical way of accessing the ergodic theory. Over the decades, many relevant works have appeared in this direction, gradually forming a fruitful branch. By a dynamical system
, we mean that
is a measured space and
T is a certain linear operator acting on
. Let
be a strictly increasing sequence of non-negative integers, i.e., a subsequence of
. Denote
as all positive integers; then, a series of discussions of the (strong, weak, or almost everywhere) convergence of
leads to various developments, such as the early work of Blum and Hanson [
1], Baxter and Olsen [
2], Bellow and Losert [
3,
4], Bourgain [
5,
6,
7,
8], Jones [
9], Jones and Olsen [
10], Wierdl [
11], as well as others.
Two factors that are often introduced at the start of the discussion are the type of subsequence and the type of action in the dynamical system; either one can be fixed so that an ample argument can be made out of the other. The type of subsequence is of more interest in this paper. One way to classify the subsequences is to consider their densities.
is said to have a
higher (respectively, lower) density δ if
Accordingly, the above sequence
has a
higher (respectively, lower) density δ if the set
has a higher (respectively, lower) density
. We have a quick connection to the weighted ergodic theory if
has nonzero density (Proposition 1.7 [
4]), which is why weighted and subsequential ergodic theorems have sometimes been studied together in the past. In this paper, we follow this line and restrict the system to only consider sequences with a density of one, uniform sequences, and block sequences; we introduce the specific descriptions later.
On the other hand, since the 1970s, noncommutative ergodic theory has been developed step-by-step based on the rapid growth of noncommutative harmonic analysis. A constructive notion of almost-uniform convergence in noncommutative
space was invented by Lance [
12], who substituted exactly the classical almost-everywhere convergence in individual ergodic theorems. Since our discussion concerns the multi-parameter setting, we extend the concept accordingly. Fix
d to be any positive integer and denote
as the
d-parameter index. Let
, where
is a von Neumann algebra with a normal semifinite faithful trace
and
is any Banach space of measurable operators associated with
(the category will be specified in
Section 2). Then by
, we mean that, given any
, there is an
such that for all
,
, and we say
converges to
x. A multi-parameter sequence
is said to
converge bilaterally almost uniformly (respectively, almost uniformly) to
x if for any
, there exists
(the lattice of projections in
) such that
converges to 0 in
. Usually, we denote it as b.a.u. (respectively, a.u.) convergence. Since 2007, Junge and Xu’s [
13] real interpolation method to obtain the strong-type noncommutative Dunford–Schwartz maximal ergodic theorem after Yeadon’s [
14] weak-type
inequality has been seen as a closed answer to the main problem of establishing a noncommutative individual ergodic theorem. Based on the structure, noncommutative ergodic theory has been going forward in some sophisticated directions. Generalizing a dynamical system’s transformation to Lamperty operators [
15] and to group actions [
16] and changing the forms of Cesàro averages to obtain the weighted (even Wiener–Wintner) ergodic theorem [
17,
18,
19] and the subsequential case [
20,
21] are some representative achievements. Inspired by [
22], in this paper, we intend to give the structure of the multi-parameter subsequential ergodic theorem for a noncommutative Dunford–Schwartz system.
Let
be a vector of
d Dunford–Schwartz operators (defined in
Section 2.3). As
(the action can be uniquely extended to any noncommutative Banach space), then
is also called a Dunford–Schwartz function, and a noncommutative multi-parameter Dunford–Schwartz system
is given. Meanwhile, let
be a vector of
d sequences of strictly increasing non-negative integers, i.e., every
is a subsequence of
,
. Naturally, we can give the density
of
as the product of the densities of each
, denoted simply as
. On the other hand, we will need the notation
, which is finite for every positive lower density sequence
, and we denote this as
hereafter.
Similar to the classical notion, we let
be the associated subsequential averaging actions and
for the multi-parameter case.
Corresponding to the classical theory, we can talk about the “goodness” of the multi-parameter subsequence.
Definition 1. Let be a Banach space constructed from , and let be a vector of d linear actions as . We say is a bilaterally good (respectively, good) subsequence in for if for every , converges b.a.u. (respectively, a.u.). Moreover, is a bilaterally good universal (respectively, good universal) subsequence in if it is bilaterally good (respectively, good) for any Dunford–Schwartz operator on .
Consequently, given , it is natural to ask: What kind of subsequences are bilaterally good universal (respectively, good universal)? In this paper, we focus on a type of subsequence that we denote as . We say that if every element of is one of the following:
- (1)
A sequence with a density of one;
- (2)
A recurring uniform sequence;
- (3)
A block sequence with a positive lower density such that .
It will be shown as a corollary of our main result that every is bilaterally good universal in if and good universal in if ; moreover, if is a noncommutative probability space, then is bilaterally good universal in and good universal in .
The single-parameter case of the above question was firstly studied by Litvinov and Mukhamedov in [
20], and their result has been recently extended by O’Brien in [
21], while the original motive comes from commutative works by Brunel and Keane [
23] and Sato [
24,
25]. In Litvinov and O’Brien’s works, they mainly apply the language of “uniform and bilaterally uniform equicontinuity in measure (in short, u.e.m. and b.u.e.m.) at zero” to treat the a.u. and b.a.u. convergence problem in
(
p takes values in
accordingly). However, in this paper, we seek a more “Littlewood–Paley” path, as in [
13], and establish a “maximal” to “individual” procedure that can be expanded in future developments.
Actually, we push the question to a relatively blank region for the subsequential theory, and thus, the above solution can be naturally included. The idea is to consider a “Wiener–Wintner”-type convergence for a certain set of subsequences as follows.
Definition 2. A set of bilaterally good universal (respectively, good universal) subsequences of is said to be of bilaterally subsequential Wiener–Wintner-type (respectively, subsequential Wiener–Wintner-type)—in short, of -bsWW (respectively, -sWW) type—if for any and any , there exists such that and In fact, if every
has a density
, by a characteristic function argument
and we have a transfer
The last term of the above equality is a multi-parameter weighted ergodic average. It shows that whether
is of
-bsWW type is closely relevant to the question of whether the weight function set
is of
-NCbWW (
-noncommutative bilateral Wiener–Wintner) type, which concerns the topic in [
17].
In addition, we give an independent construction and prove that is of -bsWW type for and -sWW type for . Moreover, if is finite, is of -bsWW type and -sWW type.
Nevertheless, for the bigger sets of nonzero density subsequences or even zero density subsequences, the question seems quite sensible and needs more investigation in the future.
3. Noncommutative Maximal Ergodic Inequalities
Usually when proving individual ergodic theorems, e.g., in [
13,
17], maximal inequalities are established as a primary part. In the following, similar to several preparatory works, we give one-parameter estimates of ergodic averages along nonzero density subsequences; then, by a routine iteration argument, multi-parameter maximal ergodic inequalities are obtained. In the process, we also see some close connections between our subsequential case and the weighted case.
Proposition 2. Given , let and suppose for every ; then .
Proof. By the given condition, we know that
thus, we know
which completes the proof. □
We extract the following maximal inequality from the proof of Theorem 3.5 in [
17], and the same argument is valid when
T extends to Dunford–Schwartz operators.
Lemma 4. Let be associated with a noncommutative probability space , and denote ; then for any , there is a constant C such thatholds for any Lemma 5. Let and be a bounded sequence of complex numbers, i.e., and for every k.
Denotethen for every , there exists a constant such thatfor any and any , there is a projection such that Proof.
(i) For every
, it has a linear decomposition
, where
By the triangle inequality and with
for each
j, we know that there is no loss of generality if we add a restriction by considering
. We decompose the mean
as follows:
where
and
. Hence, we have
By Proposition 2, we have
Similarly, we have
From Theorem 4.1 in [
13], we know there exists a constant
such that
Then, by applying the triangle inequality to the norm several times, we obtain
where
(ii) Firstly we take the decomposition
, where
and
. Applying Lemma 3 to each
, we know there is
such that
Taking
, we have
and
Thus, by using the triangle inequality again,
□
Theorem 1. Let , and suppose that the subsequence has a positive lower density.
Then for any , there exists a constant such thatAlso, for any and any , there exists a projection such thatMoreover, if the trace τ is finite, i.e., is a noncommutative probability space, then for any , there exists a constant C such that Proof.
(i) For every
, it has a linear decomposition
, where
Hence, we consider only
Since
then, we have
By the definition of the norm of
, we know
Since
for all
i, by Lemma 5, we have
and then,
(ii) Since we have
we can apply the weak-type result in Lemma 5, and we have the weak-type inequality for the subsequential case.
(iii) For the finite noncommutative Orlicz space case, a similar argument applies (with a minor change in notation in the proof of Lemma 5 and part (i) above); then, by Lemma 4, we finish the proof. □
Theorem 2. Let be a vector of d Dunford–Schwartz operators, and let be a vector of d sequences of strictly increasing non-negative integers, with every having a positive lower density, where
Then, for any , there exists a constant (inherited from Theorem 1) such thatif , we haveMoreover, if the trace τ is finite, i.e., is a noncommutative probability space, given , for any , there are a positive constant C and a projection such that for , we have the following estimates: Proof. For the same reason, we can restrict our consideration to
and apply Theorem 1 with its equivalent formulations, e.g.,
is equivalent to saying there is
satisfying
Thus, the multi-parameter maximal inequalities can be obtained from iterations of the single-parameter case, and the “
” cases are similar to Corollary 4.4 [
13] and Theorem 3.5 [
17]; this can be proved using the same arguments. □
Theorem 3. Let and , and let be a sequence of strictly increasing non-negative integers: has a positive lower density. Then for any and any , there is a projection such that Proof. Since a weighted-version weak-type
inequality is obtained in Theorem 2.1 [
30], by the same argument,
and we immediately obtain the result. □
Remark 1. We point out here that this subsequential version weak-type inequality is a mere induction from Yeadon’s weak-type inequality (check the proof of Theorem 2.1 [30]) plus a “subsequential” argument; thus, it is independent of our “strong-type” result and has a better universal constant. Moreover, it is implied in the proof of Theorem 2.1 [30] that for and any , there is a projection such that Theorem 4. Let , let be a vector of d Dunford–Schwartz operators, and let be a vector of d sequences of strictly increasing non-negative integers: every has a positive lower density, where There exists a positive constant (inherited from Theorem 1) such that for any and any , there is a projection such thatMoreover, for , we have the following estimates: Proof. Let
. By Theorem 1, we have
We use an equivalent formulation: there exists an operator
such that
Applying the previous formulation
times, there exists an operator
such that
Then by the previous theorem and Remark 1, this implies that for
and any
, there is a projection
such that
It is equivalent to say that for any
there is a projection
such that
Given
, we have
, where
and
for each
Hence, we have that for any
, there are
satisfying
Taking
, we have
For
and
(all self-adjoint operators in
), we apply the previous estimate to
. That is, for any
, there is a projection
such that
Then for any
, taking
, by the Kadison–Schwarz inequality
we obtain
Now, given
,
, where
and
. Hence, for any
, there are
such that
By taking
, we obtain the final result. □
4. Noncommutative Wiener–Wintner-Type Subsequential Ergodic Theorems
We give in the following a result that acts as the Banach principle in the theory.
Lemma 6. Let be a vector of d Dunford–Schwartz operators, and let be a family of multi-parameter subsequences satisfying that each is a vector of d sequences of strictly increasing non-negative integers and every has a positive lower density, where
Let . If for a dense subset X of we have and , there exists a projection such thatconverges in for all , then is of -bsWW type. If , is of -sWW type. Moreover, if the trace τ is finite, i.e., is a noncommutative probability space, and X with the above property is dense in , then is of -bsWW type and of -sWW type.
Proof. For
, taking any
, any
, and any
, since
X is dense in
, we can always find one
such that
where
comes from the application of the first maximal inequality in Theorem 4 to the element
: there is a projection
such that
On the other hand, by the assumption, there is a projection
such that
which means that there exists
so that whenever
, we have
Now, take
. Then we have
and
This means that is a Cauchy sequence and thus converges in for all . Therefore, we conclude that is of -bsWW type.
The rest can be shown by similar arguments with the use of the corresponding maximal inequalities in Theorem 4 and Theorem 2. □
As the main result of this paper, we give here the subsequential Wiener–Wintner-type ergodic theorem.
Theorem 5. The subsequence class Δ is of -bsWW type for and -sWW type for . Moreover, if τ is finite, Δ is of -bsWW type and -sWW type.
Proof. We know that has a density of one, a recurring uniform sequence, or a block sequence with positive lower density such that .
Let . induces a canonical splitting on ; that is, According to the decomposition, it is sufficient that we discuss x in each subset separately.
For , here we consider only the typical case ; then or .
When
, the average
; hence, there is a projection
such that
When , the average . Then it turns into the following problem.
For , we consider its dense subset instead.
In the following, we describe the ordinary situation. Let and , , be a subsequence of .
(i) When
has a density of one, by denoting
simply as
, we have
Since
, we have
(ii) When is a uniform sequence, we discuss it as follows.
Let
and
be the apparatus connected with the sequence
. Firstly, by the definition of a uniform subsequence, we have
Since
,
is finite and positive. Next, we just need to estimate
By Lemma 2, for any
, there exist open sets
and
W such that for each point
, there exists a non-negative integer
such that
By the definition of a recurring uniform sequence
, for the neighborhood
W of
as above, there are
and
such that
and
. It is obvious that
Then applying again the result as in Equation (
1), there is
such that for
, we have
Then it follows that for any
, there is a non-negative integer
such that for
So finally, we have that for any
, there is
such that
This also implies that
(iii) When
is a block sequence with a positive lower density such that
,
and since
T is a contraction on
, we obtain
By
, we have
Since Dunford–Schwarz operators are contractions on
, we come to a conclusion summing up the above results: Let
and
; we have
This implies that we have the unit operator
such that
Then, as
is dense in
, from Theorem 6, we know
is of
-bsWW type.
For the and finite Orlicz spaces case, these are the corresponding consequences of applying Theorem 6. □
Corollary 1. Let , and let be the projection onto the fixed-point subspace for , . Let and ; then, for any , there exists such that andand if , there exists such that andConsequently, every is bilaterally good universal in if and good universal if . Moreover, if the trace τ is finite, let ; then for any , there exists such that andand for , there exists such that andthus, every is bilaterally good universal in and good universal in . Proof. We consider only the typical case
. Note that
For
, fix
and decompose
x as
with
Similarly, we decompose
with respect to
, with
Applying
to
x and
to
, we obtain
By the multi-parameter maximal inequality (Theorem 2),
Similarly,
Thus,
Since
is closed in
, it remains to be shown that
Firstly, we consider the one-parameter case. In general, let . From the arguments related to the three classes of sequences in the proof of Theorem 5, we obtain the following results, respectively:
- (i)
- (ii)
- (iii)
In the following, we focus on class (i); the other classes are similar.
Since
for any
, we deduce from Theorem 1 that
belongs to
. Choose a
. Then by Proposition 1, for any
,
As
, the finite sequence
converges to
in
as
. Combining with
and being closed in
, we have
. That is to say,
,
For the two-parameter case, let
. Since
for any
, we deduce from Theorem 2 that
. Then by the interpolation theorem and with
being a contraction, for any
,
By a similar argument as above, we have that
That is to say,
Thus, by Lemma 6.2 [
13], let
; we have that for any
and any
, there exists
such that
and
Combining with Theorem 5, i.e.,
is of
-bsWW type, we have that for any
, there exists a projection
such that
and
For
, we also need only to show
Applying the second part of Theorem 2 and the interpolation theorem for
, we can prove this with a similar argument. Combined with the fact that
is of
-sWW type, there exists a projection
such that
and
In the end, the Orlicz space case can be reasoned analogously. □