Research Article On the Dependence of the Limit Functions on Takeshi Yoshimoto

advertisement
Hindawi Publishing Corporation
Abstract and Applied Analysis
Volume 2013, Article ID 503689, 13 pages
http://dx.doi.org/10.1155/2013/503689
Research Article
On the Dependence of the Limit Functions on
the Random Parameters in Random Ergodic Theorems
Takeshi Yoshimoto
Department of Mathematics, Toyo University, Kawagoe, Saitama 350-8585, Japan
Correspondence should be addressed to Takeshi Yoshimoto; yoshimoto t@toyonet.toyo.ac.jp
Received 16 October 2012; Accepted 28 December 2012
Academic Editor: Haydar Akca
Copyright © 2013 Takeshi Yoshimoto. This is an open access article distributed under the Creative Commons Attribution License,
which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
We study the structure of the ergodic limit functions determined in random ergodic theorems. When the r random parameters
are shifted by the π‘Ÿ0 -shift transformation with π‘Ÿ0 ∈ {1, 2, . . . , π‘Ÿ}, the major finding is that the (random) ergodic limit functions
determined in random ergodic theorems depend essentially only on the π‘Ÿ—π‘Ÿ0 random parameters. Some of the results obtained here
improve the earlier random ergodic theorems of Ryll-Nardzewski (1954), Gladysz (1956), Cairoli (1964), and Yoshimoto (1977) for
positive linear contractions on 𝐿 1 and Woś (1982) for sub-Markovian operators. Moreover, applications of these results to nonlinear
random ergodic theorems for affine operators are also included. Some examples are given for illustrating the relationship between
the ergodic limit functions and the random parameters in random ergodic theorems.
1. A General Argument
The present paper is concerned with the relations between
the limit functions in random ergodic theorems and the
random parameters concomitant to the limit functions. The
first results of the random ergodic theory include Pitt’s random ergodic theorem [1] and Ulam-von Neumann’s random
ergodic theorem [2] concerning a finite number of measurepreserving transformations and Kakutani’s random ergodic
theorem [3] concerning an infinite number of measurepreserving transformations. Furthermore, Kakutani dealt
with the relationship between the random ergodic theorem
and the theory of Markov processes with a stable distribution.
The random ergodic theorem is usually obtained by using
the so-called skew product method as natural extensions of
ergodic theorems and has received a great deal of attention
from the wider point of view including operator-theoretical
treatment. In fact, interesting extensions have been made by
many authors.
It was pointed out by Marczewski (see [4]) that the proof
of Kakutani’s theorem should be found which would not
use the hypothesis that the transformations in question are
one-to-one. Answering this question, Ryll-Nardzewski [4]
improved Kakutani’s theorem to the case of random measurepreserving transformations which are not necessarily one-toone and proved that the limit function is essentially independent of the random parameter. Then, later, Ryll-Nardzewski’s
theorem was generalized by GΕ‚adysz [5] to the case of a
finite number of random parameters. The Ryll-Nardzewski
theorem was extended by Cairoli [6] to the case of positive
linear contractions on 𝐿 1 with an additional condition.
Yoshimoto [7] extended both Gladysz’s theorem and Cairoli’s
theorem to the case of positive linear contractions on 𝐿 1
with a finite number of random parameters. In this paper
we inquire further into the problem of the dependence of
the limit functions upon the random parameters in random
ergodic theorems, and we have an intention of improving the
previous random ergodic theorem of Yoshimoto [7].
In what follows, we suppose that there are a given 𝜎-finite
measure space (𝑆, 𝛽, π‘š) and a probability space (Ω, F, πœ‡). Let
𝐿 𝑝 (π‘š) = 𝐿 𝑝 (𝑆, 𝛽, π‘š), 1 ≤ 𝑝 ≤ ∞ be the usual Banach spaces
of eqivalence classes of 𝛽-measurable functions defined on 𝑆.
From now on, we shall write π‘“πœ” (𝑠) for 𝑓(𝑠, πœ”) if we wish to
regard 𝑓(𝑠, πœ”) as a function of 𝑠 defined on 𝑆 for πœ” arbitrarily
fixed in Ω.
2
Abstract and Applied Analysis
It seems to be worthwhile to include the first random
ergodic theorems which may be stated, respectively, as follows.
exist π‘š-a.e. or in 𝐿 2 (π‘š)-mean with probability 1? Revesz
made the first step toward the study of this problem (see
[8, 9]).
Theorem 1 (see [1, 2]). Let π‘ˆ(𝑠), 𝑉(𝑠) be two given measurepreserving transformations of 𝑆 into itself, which generate all the
combinations of the transformations: π‘ˆ, 𝑉, π‘ˆ(π‘ˆ), π‘ˆ(𝑉), 𝑉(π‘ˆ),
𝑉(𝑉), π‘ˆ(π‘ˆ(𝑉)), π‘ˆ(𝑉(π‘ˆ)), . . .. The ergodic limit exists then for
almost every point 𝑠 of 𝑆 and almost every choice of the infinite
sequence obtained by applying π‘ˆ and 𝑉 in turn at random, for
example, π‘ˆ(𝑠), 𝑉(π‘ˆ(𝑠)), 𝑉(𝑉(π‘ˆ(𝑠))), . . ..
The most general formulation of random ergodic theorems is the following Chacon’s type theorem given by Jacobs
[10].
Exactly speaking, the first step to the theory of random
ergodic theorems was taken by Pitt [1]. The above theorem
was stated by Ulam and von Neumann [2] (independently
of Pitt), but the essence of the contents is the same as the
theorem of Pitt who proved both the pointwise convergence
and the 𝐿 𝑝 (π‘š) (𝑝 ≥ 1, π‘š(𝑆) < ∞) mean convergence
of random averages in question. Ulam and von Neumann
announced the pointwise convergence of random averages in
an abstract form, but the proof has never been published. PittUlam-von Neumann’s random ergodic theorem concerning
a finite number of measure-preserving transformations was
extended by Kakutani [3] to the case of an infinite number of
measure-preserving transformations as follows.
Theorem 2 (Kakutani (1948–1950) [3]). Let Φ = {πœ‘πœ” :
πœ” ∈ Ω} be a 𝛽 ⊗ F-measurable family of measure-preserving
transformations πœ‘πœ” defined on 𝑆, where π‘š(𝑆) = πœ‡(𝑆) = 1. Let
(Ω∗± , F∗± , πœ‡±∗ ) be the two-sided infinite direct product measure
space of (Ω, F, πœ‡). Then, for any 𝑓 ∈ 𝐿 𝑝 (π‘š) (𝑝 ≥ 1), there
exists a (F∗± , πœ‡±∗ )-null set 𝑁∗ such that for any πœ”∗ ∈ Ω∗± − 𝑁∗ ,
Μƒ ∗ (𝑠) ∈ 𝐿 (π‘š) such that
there exists a function 𝑓
𝑝
πœ”
1 𝑛−1
Μƒ ∗ (𝑠)
∑ 𝑓 (πœ‘ℵπ‘˜ (πœ”∗ ) ⋅ ⋅ ⋅ πœ‘ℵ0 (πœ”∗ ) 𝑠) = 𝑓
πœ”
𝑛→∞𝑛
π‘˜=0
lim
π‘š-π‘Ž.𝑒.,
(1)
where ℵ𝑛 (πœ”∗ ) denotes the 𝑛th coordinate of πœ”∗ , and this holds
also in the norm of 𝐿 𝑝 (π‘š).
Kakutani’s paper on the random ergodic theorem was
published in 1950, but the random ergodic theorem had
already been dealt with by Kawada (“Random ergodic theorems”, Suritokeikenkyu (Japanese) 2, 1948). Later, Kawada
reminisced about the circumstances of an affair of his paper. It
is the rights of matter that Kawada’s result is due to Kakutani’s
kind suggestion.
Remark 3. In Kakutani’s random ergodic theorem (as well
as in Pitt-Ulam-von Neumann’ theorem), the sequence
(πœ‘ℵ𝑛 (πœ”∗ ) )𝑛≥0 of measure-preserving transformations on 𝑆 is
chosen at random with the same distribution and independently. In connection with this question, an interesting
problem is the following: if we choose a sequence (πœ‘π‘› )𝑛≥1
at random, not necessarily with the same distribution but
independently from a given set Φ of measure-preserving
transformations on 𝑆, under what condition does the limit
1 𝑛
∑ 𝑓 (πœ‘π‘˜ ⋅ ⋅ ⋅ πœ‘1 𝑠)
𝑛→∞𝑛
π‘˜=1
lim
(2)
Jacobs’ General Random Ergodic Theorem [10]. Let 𝜎 be an
endomorphism of (Ω, F, πœ‡) and let {π‘‡πœ” : πœ” ∈ Ω} be a
strongly F-measurable family of random linear contractions
on 𝐿 1 (π‘š). Let {𝑝𝑛 }𝑛≥0 be a sequence of 𝛽 ⊗ F-measurable
functions defined on 𝑆 × Ω which is admissible for {π‘‡πœ” :
πœ” ∈ Ω}. This means that if β„Ž ∈ 𝐿 1 (π‘š ⊗ πœ‡) and |β„Ž(𝑠, πœ”)| ≤
𝑝𝑛 (𝑠, πœ”) π‘š ⊗ πœ‡-a.e., then |π‘‡πœ” β„ŽπœŽπœ” (𝑠)| ≤ 𝑝𝑛+1 (𝑠, πœ”) π‘š ⊗ πœ‡-a.e.
Then, for any function 𝑓 ∈ 𝐿 1 (π‘š), there exists a πœ‡-null set
𝑁 ∈ F such that for any πœ” ∈ Ω − 𝑁,
∑π‘›π‘˜=0 π‘‡πœ” π‘‡πœŽπœ” ⋅ ⋅ ⋅ π‘‡πœŽπ‘˜ πœ” 𝑓 (𝑠)
Μƒ (𝑠, πœ”))
(= 𝑓
𝑛→∞
∑π‘›π‘˜=0 π‘π‘˜ (𝑠, πœ”)
lim
(3)
exists and is finite π‘š-a.e. on the set {𝑠 : ∑∞
π‘˜=0 π‘π‘˜ (𝑠, πœ”) > 0}
(cf. [11] which includes a further weighted generalization of
Jacobs’theorem).
If all π‘‡πœ” are positive, then Jacobs’ theorem yields the
following Chacon-Ornstein’s type random ergodic theorem
(cf. [10, 12]); for 𝑓 ∈ 𝐿 1 (π‘š) and 𝑔 ∈ 𝐿+1 (π‘š), there exists a
πœ‡-null set 𝑁 such that for any πœ” ∈ Ω − 𝑁,
∑π‘›π‘˜=0 π‘‡πœ” π‘‡πœŽπœ” ⋅ ⋅ ⋅ π‘‡πœŽπ‘˜ πœ” 𝑓 (𝑠)
Μƒ (𝑠, πœ”))
(= 𝑓
𝑔
𝑛 → ∞ ∑𝑛 𝑇 𝑇 ⋅ ⋅ ⋅ 𝑇 π‘˜ 𝑔 (𝑠)
𝜎 πœ”
π‘˜=0 πœ” πœŽπœ”
lim
(4)
exists and is finite π‘š-a.e. on the set {𝑠 : ∑∞
π‘˜=0 π‘‡πœ” π‘‡πœŽπœ” ⋅ ⋅ ⋅
π‘‡πœŽπ‘˜ πœ” 𝑔(𝑠) > 0}. Moreover, if the family {π‘‡πœ” , πœ” ∈ Ω} has a
strictly positive invariant function 𝑔 ∈ 𝐿+1 (π‘š), then for every
𝑓 ∈ 𝐿 1 (π‘š), there exists a πœ‡-null set 𝑁 such that for any
πœ”∈Ω−𝑁
1 𝑛
∑ π‘‡πœ” π‘‡πœŽπœ” ⋅ ⋅ ⋅ π‘‡πœŽπ‘˜ πœ” 𝑓 (𝑠) (= 𝑓∗ (𝑠, πœ”))
𝑛→∞𝑛 + 1
π‘˜=0
lim
(5)
exists and is finite π‘š-a.e. Unfortunately, this CesaΜ€ro-type
result does not hold in general without assuming the existence of a strictly positive invariant function. However, if the
family {π‘‡πœ” : πœ” ∈ Ω} satisfies the norm conditions β€–π‘‡πœ” ‖𝐿 1 (π‘š) ≤
1 and β€–π‘‡πœ” ‖𝐿 ∞ (π‘š) ≤ 1 for all πœ” ∈ Ω, then the above CesaΜ€rotype random ergodic theorem holds even without assuming
the existence of a strictly positive invariant function in 𝐿 1 (π‘š).
Μƒ πœ”) = [𝑓]
Μƒ (𝑠), 𝑓
Μƒ (𝑠, πœ”) =
The above limit functions 𝑓(𝑠,
πœ”
𝑔
∗
∗
Μƒ ] (𝑠), and 𝑓 (𝑠, πœ”) = [𝑓 ] (𝑠) depend generally on
[𝑓
𝑔 πœ”
πœ”
the random parameter πœ”. Our particular interest is in the
relationship between the limit functions and the random
parameters in the case that 𝜎 is the shift transformation of the
random parameter space (Ω∗ , F∗ , πœ‡∗ ) being the one-sided
infinite product of the same probability space (Ω, F, πœ‡).
Abstract and Applied Analysis
3
2. Random Ergodic Theorems
Throughout all that follows, let (Ω∗ , F∗ , πœ‡∗ ) be the one-sided
infinite product measure space of (Ω, F, πœ‡):
∞
∞
∞
(Ω∗ , F∗ , πœ‡∗ ) = (∏Ω𝑛 , ⨂ F𝑛 , ⨂ πœ‡π‘› ) ,
𝑛=1
𝑛=1
𝑛=1
π‘Ÿ
π‘Ÿ
π‘Ÿ
𝑛=1
𝑛=1
𝑛=1
(Ω∗π‘Ÿ , F∗π‘Ÿ , πœ‡π‘Ÿ∗ ) = (∏Ω𝑛 , ⨂ F𝑛 , ⨂ πœ‡π‘› ) ,
(6)
(π‘Ÿ: a positive integer) ,
(Ω𝑛 , F𝑛 , πœ‡π‘› ) = (Ω, F, πœ‡) ,
∗
ℵ𝑛 (πœ‘πœ” ) = ℵ𝑛+1 (πœ” ) ,
𝑛 = 1, 2, . . . .
∗
1 𝑛 π‘˜
Μƒ (𝑠, πœ”∗ )
∑ 𝑇 𝑓 (𝑠, πœ”∗ ) = 𝑓
𝑛→∞𝑛
π‘˜=1
lim
(7)
∗
Then, πœ‘ is clearly a F -measurable and πœ‡ -measurepreserving transformation defined on Ω∗ . Let π‘Ÿ be a fixed
positive integer. For simplicity, we let [πœ”∗ ]π‘Ÿ = (πœ”1 , . . . , πœ”π‘Ÿ )
for any πœ”∗ = (πœ”1 , . . . , πœ”π‘Ÿ , . . .) ∈ Ω∗ . Fix an π‘Ÿ0 ∈ {1, 2, . . . , π‘Ÿ}.
Suppose that to each [πœ”∗ ]π‘Ÿ ∈ Ω∗π‘Ÿ , there corresponds a linear
contraction operator 𝑇[πœ”∗ ] on 𝐿 1 (π‘š). The family {𝑇[πœ”∗ ]π‘Ÿ :
πœ”∗ ∈ Ω∗ } is said to be strongly F∗π‘Ÿ -measurable if for any
β„Ž ∈ 𝐿 1 (π‘š) the function 𝑇[πœ”∗ ]π‘Ÿ β„Ž is strongly F∗π‘Ÿ -measurable
as an 𝐿 1 (π‘š)-valued function defined on Ω∗π‘Ÿ , namely, for the
mapping Ψβ„Ž : [πœ”∗ ]π‘Ÿ → 𝑇[πœ”∗ ]π‘Ÿ β„Ž of F∗π‘Ÿ into 𝐿 1 (π‘š), πœ‡π‘Ÿ∗ ∘ Ψβ„Ž−1
has a separable support (cf. [13]).
Our main result is stated as follows.
Theorem 4. Let π‘Ÿ0 be any fixed integer with 1 ≤ π‘Ÿ0 ≤ π‘Ÿ. Let
{𝑇[πœ”∗ ]π‘Ÿ : πœ”∗ ∈ Ω∗ } be a strongly F∗π‘Ÿ -measurable family of
positive linear contractions on 𝐿 1 (π‘š). Suppose that there exists
a strictly positive 𝐿 1 (π‘š)-function invariant under {𝑇[πœ”∗ ]π‘Ÿ :
πœ”∗ ∈ Ω∗ }. Then, for every 𝑓 ∈ 𝐿 1 (π‘š), there exists a πœ‡∗ -null
set 𝑁∗ ∈ F∗ such that for any πœ”∗ ∈ Ω∗ − 𝑁∗ , there exists a
Μƒ ∗
function 𝑓
[πœ” ]π‘Ÿ−π‘Ÿ ∈ 𝐿 1 (π‘š) such that
Μƒ = 𝑓.
Μƒ
𝑇𝑓
lim
Μƒ ∗
= 𝑓
[πœ” ]π‘Ÿ−π‘Ÿ (𝑠)
0
(8)
π‘š-π‘Ž.𝑒.
Proof. We need the following lemmas.
F∗π‘Ÿ -measurability
Lemma 5 (see [10, 14]). The strong
of the
operator family {𝑇[πœ”∗ ]π‘Ÿ : πœ”∗ ∈ Ω∗ } guarantees that for any
𝑔 ∈ 𝐿 1 (π‘š ⊗ πœ‡∗ ), there exists a uniquely determined 𝛽 ⊗ F∗ measurable version [𝑇[πœ”∗ ]π‘Ÿ π‘”πœ‘π‘Ÿ0 πœ”∗ ](𝑠) of 𝑇[πœ”∗ ]π‘Ÿ π‘”πœ‘π‘Ÿ0 πœ”∗ (𝑠) such
that excepting a πœ‡∗ -null set,
[𝑇[πœ”∗ ]π‘Ÿ π‘”πœ‘π‘Ÿ0 πœ”∗ ] (𝑠) = 𝑇[πœ”∗ ]π‘Ÿ π‘”πœ‘π‘Ÿ0 πœ”∗ (𝑠)
π‘š-a.e.
(9)
Using the measurable version appearing in Lemma 5, we
define
𝑇𝑔 (𝑠, πœ”∗ ) = [𝑇[πœ”∗ ]π‘Ÿ π‘”πœ‘π‘Ÿ0 πœ”∗ ] (𝑠) .
(11)
(10)
(12)
∗
One can easily verify that excepting a suitable πœ‡ -null set
𝑁1∗ ∈ F∗ ,
π‘‡π‘˜ 𝑓 (𝑠, πœ”∗ )
= 𝑇[πœ”∗ ]π‘Ÿ 𝑇[πœ‘π‘Ÿ0 πœ”∗ ]π‘Ÿ ⋅ ⋅ ⋅ 𝑇[πœ‘(π‘˜−1)π‘Ÿ0 πœ”∗ ]π‘Ÿ π‘“πœ‘π‘˜π‘Ÿ0 πœ”∗ (𝑠)
π‘š-a.e.,
(13)
Μƒ πœ”∗ ) does
for all π‘˜ = 1, 2, . . .. Next we wish to show that 𝑓(𝑠,
depend essentially only on (𝑠, [πœ”∗ ]π‘Ÿ−π‘Ÿ0 ). To do this, we define
sub-𝜎-fields ℘𝑛 , 𝑛 = 1, 2, . . ., of 𝛽 ⊗ F∗ by
℘𝑛 = 𝛽 ⊗ F1 ⊗ ⋅ ⋅ ⋅ ⊗ 𝛽𝑛
⊗ F𝑛 ⊗ {0, Ω𝑛+1 } ⊗ ⋅ ⋅ ⋅ ,
𝑛 = 1, 2, . . . ,
(14)
(℘0 = 𝛽 ⊗ {0, Ω1 } ⊗ {0, Ω2 } ⊗ ⋅ ⋅ ⋅) .
It is clear that if π‘š < 𝑛, then ℘π‘š ⊂ ℘𝑛 and that if
𝑓 ∈ 𝐿 1 (π‘š ⊗ πœ‡∗ ), 𝑓𝑛 = ℘𝑛 𝑓, then ℘π‘š 𝑓𝑛 = π‘“π‘š whenever
π‘š ≤ 𝑛. Therefore, the system {𝑓𝑛 , ℘𝑛 : 𝑛 = 1, 2, . . .} forms
a martingale. For each 𝑛, let 𝐸(⋅ | ℘𝑛 ) denote the conditional
expectation operator with respect to the sub-𝜎-field ℘𝑛 . Let
β„Ž(𝑠, πœ”∗ ) be of the form
β„“
β„Ž (𝑠, πœ”∗ ) = πœ‰ (𝑠) ⋅ ∏πœ‚π‘– (πœ”π‘– ) ,
(15)
𝑖=1
0
1 𝑛
∑ 𝑇[πœ”∗ ]π‘Ÿ 𝑇[πœ‘π‘Ÿ0 πœ”∗ ]π‘Ÿ ⋅ ⋅ ⋅ 𝑇[πœ‘(π‘˜−1)π‘Ÿ0 πœ”∗ ]π‘Ÿ π‘“πœ‘π‘˜π‘Ÿ0 πœ”∗ (𝑠)
𝑛→∞𝑛
π‘˜=1
π‘š ⊗ πœ‡-a.e.
Moreover, as is easily checked, we find that
𝑛 = 1, 2, . . . .
Let πœ‘ be the (one-sided) shift transformation defined on
Ω∗ which means that using the coordinate functions ℵ𝑛 (⋅),
∗
From the norm conditions of {𝑇[πœ”∗ ]π‘Ÿ : πœ”∗ ∈ Ω∗ } in
Theorem 4, it turns out that 𝑇 is a linear operator on 𝐿 1 (π‘š ⊗
πœ‡∗ ) with ‖𝑇‖𝐿 1 (π‘š⊗πœ‡∗ ) ≤ 1. Moreover, it is easy to check that
there exists a strictly positive 𝐿 1 (π‘š ⊗ πœ‡∗ )-function invariant
under 𝑇. Thus, for any 𝑓 ∈ 𝐿 1 (π‘š ⊗ πœ‡∗ ), we can apply ChaconOrnstein’s ergodic theorem [12] (cf. Hopf ’s ergodic theorem
Μƒ ∈ 𝐿 (π‘š ⊗ πœ‡∗ )
[15]) to ensure the existence of a function 𝑓
1
such that
where πœ‰ ∈ 𝐿 1 (π‘š) ∩ 𝐿 ∞ (π‘š), πœ‚π‘– ∈ 𝐿 ∞ (Ω𝑖 ), 𝑖 = 1, 2, . . . , β„“.
Then, the linear combinations of functions of the form (15)
are everywhere dense in 𝐿 1 (π‘š ⊗ πœ‡∗ ). Thus, for the question
Μƒ |℘ )=
confronting us, it suffices to prove the relation 𝐸(𝑓
π‘Ÿ−π‘Ÿ0
Μƒ
Μƒ
𝑓 only for the case when 𝑓, is of the form (15).
Lemma 6. It holds that 𝐸(π‘‡π‘š ⋅ | ℘π‘Ÿ−π‘Ÿ0 ) = π‘‡π‘š 𝐸(⋅ | ℘(π‘š−1)π‘Ÿ0 +π‘Ÿ ),
π‘š = 1, 2, . . ..
Proof. It follows that for a sufficiently large β„“ and 𝑛 < β„“, 𝐸(β„Ž |
℘𝑛 ) is such that
𝑛
𝐸 (β„Ž | ℘𝑛 ) (𝑠, πœ”∗ ) = πœ‰ (𝑠) ⋅ ∏πœ‚π‘– (πœ”π‘– )
𝑖=1
β„“
⋅ ∏ ∫ πœ‚π‘— (πœ”) π‘‘πœ‡ (πœ”)
𝑗=𝑛+1 Ω
π‘š ⊗ πœ‡∗ -a.e.
(16)
4
Abstract and Applied Analysis
Thus, excepting a πœ‡∗ -null set 𝑁2∗ , we get
We return to the proof of the theorem. By Lemma 6 and
the martingale convergence theorem (cf. [16, 17]), we have
𝑛
∗
(𝑇∗ 𝐸 (β„Ž | ℘𝑛 ))πœ”∗ (𝑠) = [𝑇[πœ”
∗ ] πœ‰] (𝑠) ∏πœ‚π‘– (πœ”π‘–+π‘Ÿ )
0
π‘Ÿ
𝑖=1
β„“
× ∏ ∫ πœ‚π‘— (πœ”) π‘‘πœ‡ (πœ”)
𝑗=𝑛+1 Ω
π‘š-a.e.
(17)
On the other hand, since excepting a πœ‡∗ -null set
β„“
∗
(𝑇∗ β„Ž)πœ”∗ (𝑠) = [𝑇[πœ”
∗ ] πœ‰] (𝑠) ∏πœ‚π‘– (πœ”π‘–+π‘Ÿ )
0
π‘Ÿ
σ΅„©σ΅„© Μƒ
Μƒ σ΅„©σ΅„©σ΅„©σ΅„©
󡄩󡄩𝐸 (𝑓 | ℘π‘Ÿ−π‘Ÿ0 ) − 𝑓
󡄩𝐿 1 (π‘š⊗πœ‡∗ )
σ΅„©
σ΅„©
Μƒ | ℘ ) − π‘‡π‘š 𝑓
Μƒ σ΅„©σ΅„©σ΅„©σ΅„©
= 󡄩󡄩󡄩󡄩𝐸 (π‘‡π‘š 𝑓
π‘Ÿ−π‘Ÿ0
󡄩𝐿 1 (π‘š⊗πœ‡∗ )
σ΅„©
π‘š Μƒσ΅„©
σ΅„©
Μƒ |℘
≤ σ΅„©σ΅„©σ΅„©σ΅„©π‘‡π‘š 𝐸 (𝑓
󡄩󡄩𝐿 1 (π‘š⊗πœ‡∗ )
(π‘š−1)π‘Ÿ0 +π‘Ÿ ) − 𝑇 𝑓󡄩
σ΅„© Μƒ
Μƒ σ΅„©σ΅„©σ΅„©σ΅„©
| ℘(π‘š−1)π‘Ÿ0 +π‘Ÿ ) − 𝑓
≤ 󡄩󡄩󡄩󡄩𝐸 (𝑓
󡄩𝐿 1 (π‘š⊗πœ‡∗ ) 󳨀→ 0
(25)
as π‘š → ∞, and thus
π‘š-a.e.,
Μƒ
Μƒ | ℘ ) = 𝑓,
𝐸 (𝑓
π‘Ÿ−π‘Ÿ0
(18)
𝑖=1
(26)
Μƒ πœ”∗ ) depends essentially only on
which implies that 𝑓(𝑠,
∗
(𝑠, [πœ” ]π‘Ÿ−π‘Ÿ0 ). Hence, the theorem follows from (11), (26), and
Fubini’s theorem. The proof of Theorem 4 has hereby been
completed.
we have that if 𝑛 ≥ π‘Ÿ − π‘Ÿ0 , then
𝑛
(𝐸 (𝑇∗ β„Ž | ℘𝑛+π‘Ÿ0 ))πœ”∗ (𝑠) = [𝑇[πœ”∗ ]π‘Ÿ πœ‰] (𝑠) ∏πœ‚π‘– (πœ”π‘–+π‘Ÿ0 )
𝑖=1
β„“
× ∏ πœ‚π‘— (πœ”) π‘‘πœ‡ (πœ”)
π‘š-a.e.
𝑗=𝑛+1
(19)
Consequently,
𝑇∗ 𝐸 (β„Ž | ℘𝑛 ) = 𝐸 (𝑇∗ β„Ž | ℘𝑛+π‘Ÿ0 )
π‘š ⊗ πœ‡∗ -a.e.,
(20)
and by approximation,
∗
If we take π‘Ÿ0 = π‘Ÿ in Theorem 4, we have Cairoli’s
Μƒ πœ”∗ ) does not depend essentially on
theorem in which 𝑓(𝑠,
∗
πœ” . The π‘Ÿ random parameter generalization (see [7]) of
Cairoli’s theorem is obtained by taking π‘Ÿ0 = 1 in Theorem 4.
Adapting the (almost) same argument as used in the proof of
Theorem 4, we have the following.
Theorem 7. Let π‘Ÿ0 be any fixed integer with 1 ≤ π‘Ÿ0 ≤ π‘Ÿ. Let
{𝑇[πœ”∗ ]π‘Ÿ : πœ”∗ ∈ Ω∗ } be a strongly F∗π‘Ÿ -measurable family of
linear contractions on 𝐿 1 (π‘š) with ‖𝑇[πœ”∗ ]π‘Ÿ ‖𝐿 (π‘š) ≤ 1 for all
∞
πœ”∗ ∈ Ω∗ . Then, for every 𝑓 ∈ 𝐿 𝑝 (π‘š ⊗ πœ‡∗ ) with 1 ≤ 𝑝 < ∞,
there exists a πœ‡∗ -null set 𝑁∗ ∈ F∗ such that for any πœ”∗ ∈
Μƒ ∗
Ω∗ − 𝑁∗ , there exists a function 𝑓
[πœ” ]π‘Ÿ−π‘Ÿ ∈ 𝐿 𝑝 (π‘š) such that
0
∗
𝑇 𝐸 (⋅ | ℘𝑛 ) = 𝐸 (𝑇 ⋅ | ℘𝑛+π‘Ÿ0 ) ,
(21)
1 𝑛
∑ 𝑇[πœ”∗ ]π‘Ÿ 𝑇[πœ‘π‘Ÿ0 πœ”∗ ]π‘Ÿ ⋅ ⋅ ⋅ 𝑇[πœ‘(π‘˜−1)π‘Ÿ0 πœ”∗ ]π‘Ÿ π‘“πœ‘π‘˜π‘Ÿ0 πœ”∗ (𝑠)
𝑛→∞𝑛
π‘˜=1
lim
Μƒ ∗
= 𝑓
[πœ” ]π‘Ÿ−π‘Ÿ (𝑠)
and so by iteration,
0
𝑇∗π‘š 𝐸 (⋅ | ℘𝑛 ) = 𝐸 (𝑇∗π‘š ⋅ | ℘𝑛+π‘šπ‘Ÿ0 ) ,
π‘š = 0, 1, 2, . . . , 𝑛 ≥ π‘Ÿ − π‘Ÿ0 .
(22)
π‘š = 1, 2, . . . .
(23)
∗
Since 𝑇 is the adjoint operator of 𝑇 , we have thus
𝐸 (π‘‡π‘š ⋅ | ℘π‘Ÿ−π‘Ÿ0 ) = π‘‡π‘š 𝐸 (⋅ | ℘(π‘š−1)π‘Ÿ0 +π‘Ÿ ) ,
π‘š-π‘Ž.𝑒.,
and that if 1 < 𝑝 < ∞, then
In particular,
𝑇∗π‘š 𝐸 (⋅ | ℘π‘Ÿ−π‘Ÿ0 ) = 𝐸 (𝑇∗π‘š ⋅ | ℘(π‘š−1)π‘Ÿ0 +π‘Ÿ ) ,
(27)
π‘š = 1, 2, . . . .
(24)
σ΅„©σ΅„© 𝑛
σ΅„©σ΅„© 1
lim σ΅„©σ΅„©σ΅„© ∑ 𝑇[πœ”∗ ]π‘Ÿ 𝑇[πœ‘π‘Ÿ0 πœ”∗ ]π‘Ÿ ⋅ ⋅ ⋅ 𝑇[πœ‘(π‘˜−1)π‘Ÿ0 πœ”∗ ]π‘Ÿ π‘“πœ‘π‘˜π‘Ÿ0 πœ”∗ (⋅)
𝑛→∞ σ΅„©
σ΅„©σ΅„© 𝑛 π‘˜=1
σ΅„©σ΅„©σ΅„©
σ΅„©σ΅„©
Μƒ ∗
−𝑓
= 0,
(⋅)
σ΅„©σ΅„©
[πœ” ]π‘Ÿ−π‘Ÿ0
σ΅„©σ΅„©
󡄩𝐿 𝑝 (π‘š)
(28)
and that if π‘š(𝑆) < ∞ and 𝑓 ∈ 𝐿(𝑆 × Ω∗ )log+ 𝐿(𝑆 × Ω∗ ), then
σ΅„©σ΅„© 𝑛
σ΅„©σ΅„© 1
lim σ΅„©σ΅„©σ΅„© ∑ 𝑇[πœ”∗ ]π‘Ÿ 𝑇[πœ‘π‘Ÿ0 πœ”∗ ]π‘Ÿ ⋅ ⋅ ⋅ 𝑇[πœ‘(π‘˜−1)π‘Ÿ0 πœ”∗ ]π‘Ÿ π‘“πœ‘π‘˜π‘Ÿ0 πœ”∗ (⋅)
𝑛→∞ σ΅„©
σ΅„©σ΅„© 𝑛 π‘˜=1
σ΅„©σ΅„©σ΅„©
σ΅„©σ΅„©
Μƒ ∗
−𝑓
= 0.
(⋅)
σ΅„©σ΅„©
[πœ” ]π‘Ÿ−π‘Ÿ0
σ΅„©σ΅„©
󡄩𝐿 1 (π‘š)
(29)
Abstract and Applied Analysis
5
Proof. As before, we define 𝑇𝑔(𝑠, πœ”∗ ) = [𝑇[πœ”∗ ]π‘Ÿ π‘”πœ‘π‘Ÿ0 πœ”∗ ](𝑠) for
𝑔 ∈ 𝐿 1 (π‘š ⊗ πœ‡∗ ). From the norm conditions of {𝑇[πœ”∗ ]π‘Ÿ : πœ”∗ ∈
Ω∗ } in Theorem 7, it turns out that 𝑇 is a linear operator
on 𝐿 1 (π‘š ⊗ πœ‡∗ ) with ‖𝑇‖𝐿 1 (π‘š⊗πœ‡∗ ) ≤ 1 and ‖𝑇‖𝐿 ∞ (π‘š⊗πœ‡∗ ) ≤
1. Then, It follows from the Riesz convexity theorem that
‖𝑇‖𝐿 𝑝 (π‘š⊗πœ‡∗ ) ≤ 1 for all 𝑝 with 1 < 𝑝 < ∞. Thus, for any
𝑓 ∈ 𝐿 𝑝 (π‘š ⊗ πœ‡∗ ) (1 ≤ 𝑝 < ∞), we can apply Dunford and
Schwartz’s ergodic theorem [18] to ensure the existence of a
Μƒ ∈ 𝐿 (π‘š ⊗ πœ‡∗ ) such that
function 𝑓
𝑝
1 𝑛
Μƒ (𝑠, πœ”∗ )
lim ∑ π‘‡π‘˜ 𝑓 (𝑠, πœ”∗ ) = 𝑓
𝑛→∞𝑛
π‘˜=1
π‘š ⊗ πœ‡-a.e.,
(30)
Μƒ = 𝑓,
Μƒ
𝑇𝑓
σ΅„©σ΅„© 𝑛
σ΅„©σ΅„©
σ΅„©σ΅„© 1
σ΅„©
Μƒ σ΅„©σ΅„©σ΅„©
lim σ΅„©σ΅„©σ΅„© ∑ π‘‡π‘˜ 𝑓 − 𝑓
= 0,
σ΅„©σ΅„©
𝑛 → ∞σ΅„© 𝑛
σ΅„©σ΅„© π‘˜=1
󡄩󡄩𝐿 (π‘š⊗πœ‡∗ )
𝑝
+
∗
(31)
∗
and that if π‘š(𝑆) < ∞ and 𝑓 ∈ 𝐿(𝑆 × Ω )log 𝐿(𝑆 × Ω ), then
σ΅„©σ΅„© 𝑛
σ΅„©σ΅„©
σ΅„©σ΅„© 1
σ΅„©σ΅„©
π‘˜
Μƒ
σ΅„©
= 0.
lim σ΅„© ∑ 𝑇 𝑓 − 𝑓󡄩󡄩󡄩
𝑛 → ∞σ΅„©
σ΅„©σ΅„© 𝑛 π‘˜=1
σ΅„©σ΅„©
σ΅„©
󡄩𝐿 1 (π‘š⊗πœ‡∗ )
(32)
Now, adapting the (almost) same argument as used in the
proof of Theorem 4, we can find that
∗
π‘š ⊗ πœ‡ -a.e.
(or, ∈ 𝐿 1 (π‘š ⊗ πœ‡∗ )) ,
(34)
where |𝑇| denotes the linear modulus of 𝑇 (see [18, 19]).
Therefore, (27) follows from (30), (33), and Fubini’s theorem.
Equations (28) and (29) follow from (30), (31), (32), (33), (34),
and Fubini’s theorem.
In the setting of measure-preserving transformations,
Theorem 7 is reduced to the random one-parameter result of
Ryll-Nardzewski [4] by taking π‘Ÿ0 = π‘Ÿ = 1 and to the π‘Ÿ random
parameter result of GΕ‚adysz [5] by taking π‘Ÿ0 = 1.
(𝑖)
:
Theorem 8. Let π‘Ÿ1 , . . . , π‘Ÿπœˆ be 𝜈 positive integers. Let {𝑇[πœ”
∗]
π‘Ÿπ‘–
πœ”∗ ∈ Ω∗ }, 𝑖 = 1, 2, . . . , 𝜈, be strongly F(𝑖)
π‘Ÿπ‘– (𝑖 = 1, 2, . . . , 𝜈)measurable families of linear contractions on 𝐿 1 (π‘š) with
(𝑖)
‖𝑇[πœ”
≤ 1. For each 𝑖 = 1, 2, . . . , 𝜈, we set
∗] β€–
π‘Ÿπ‘–
𝑛
1
𝜈
𝐿 ∞ (π‘š)
𝜈
Μƒ ∗ (𝑠) π‘š-a.e. as 𝑛 → ∞, . . . , 𝑛 → ∞
converge to 𝑓
1
𝜈
[πœ” ]π‘ž−1
Μƒ
independently. Here, 𝑓 ∗ (⋅) means that the number of
[πœ” ]π‘ž−1
random parameters is at most π‘ž − 1.
(𝑖)
𝑇𝑖 β„Ž (𝑠, πœ”∗ ) = [𝑇[πœ”
∗ ] β„Žπœ‘π‘Ÿπ‘– πœ”∗ ] (𝑠) ,
π‘Ÿ
𝑖 = 1, 2, . . . , 𝜈.
𝑖
(𝑖)
(𝑖)
(𝑖)
𝑇𝑖 (π‘˜, [πœ”∗ ]π‘Ÿπ‘– ) = 𝑇[πœ”
,
∗ ] 𝑇[πœ‘πœ”∗ ] ⋅ ⋅ ⋅ 𝑇 π‘˜−1 ∗
[πœ‘ πœ” ]
π‘Ÿ
π‘Ÿ
𝑖
π‘Ÿπ‘–
π‘˜ = 1, 2, ⋅ ⋅ ⋅ ,
𝑛
𝑛
1
𝜈
𝜈
1
1
π‘˜
∑ ⋅ ⋅ ⋅ ∑ 𝑇1 1 ⋅ ⋅ ⋅ π‘‡πœˆπ‘˜πœˆ 𝑓 = 𝐸1 ⋅ ⋅ ⋅ 𝐸𝜈 𝑓
𝑛1 ⋅ ⋅ ⋅ π‘›πœˆ π‘˜ =1 π‘˜ =1
(38)
π‘š ⊗ πœ‡∗ -a.e.
converge to 𝐸1 ⋅ ⋅ ⋅ 𝐸𝜈 𝑓 almost everywhere on 𝑆 × Ω∗ as 𝑛1 →
∞, . . . , π‘›πœˆ → ∞ independently, where each 𝐸𝑖 is a projection
of 𝐿 𝑝 (π‘š⊗πœ‡∗ ) onto the manifold 𝑁(𝐼−𝑇𝑖 ) = {𝑔 ∈ 𝐿 𝑝 (π‘š⊗πœ‡∗ ) :
𝑇𝑖 𝑔 = 𝑔} with
σ΅„©σ΅„© 𝑛𝑖
σ΅„©σ΅„©
σ΅„©σ΅„© 1
σ΅„©σ΅„©
π‘˜π‘–
σ΅„©
lim σ΅„©σ΅„©σ΅„© ∑ 𝑇𝑖 𝑓 − 𝐸𝑖 𝑓󡄩󡄩󡄩󡄩
= 0,
𝑛𝑖 → ∞σ΅„© 𝑛
σ΅„©σ΅„©
σ΅„©σ΅„© 𝑖 π‘˜π‘– =1
󡄩𝐿 𝑝 (π‘š⊗πœ‡∗ )
𝑖 = 1, 2, . . . , 𝜈.
(39)
Note here that
𝐸𝜈 𝑓 ∈ 𝑁 (𝐼 − π‘‡πœˆ ) , 𝐸𝜈−1 𝐸𝜈 𝑓 ∈ 𝑁 (𝐼 − π‘‡πœˆ−1 ) , . . . ,
𝐸1 ⋅ ⋅ ⋅ 𝐸𝜈 𝑓 ∈ 𝑁 (𝐼 − 𝑇1 ) .
(40)
Thus, by Lemma 6 and the (𝐿 𝑝 ) martingale convergence
theorem, we find that excepting a πœ‡∗ -null set
𝐸𝜈 𝑓 (𝑠, πœ”∗ ) = (𝐸𝜈 𝑓)[πœ”∗ ]
π‘Ÿπœˆ −1
π‘š-a.e.,
(𝑠)
𝐸𝜈−1 𝐸𝜈 𝑓 (𝑠, πœ”∗ ) = (𝐸𝜈−1 𝐸𝜈 𝑓)[πœ”∗ ]
𝐸1 ⋅ ⋅ ⋅ 𝐸𝜈 𝑓 (𝑠, πœ”∗ ) = (𝐸1 ⋅ ⋅ ⋅ 𝐸𝜈 𝑓)[πœ”∗ ]
(𝑠)
π‘š-a.e.,
max(π‘Ÿ1 ,...,π‘Ÿπœˆ )−1
(𝑇𝑖 (0, [πœ”∗ ]π‘Ÿπ‘– ) = the identity operator) .
(37)
Each 𝑇𝑖 turns out to be a linear contraction on 𝐿 1 (π‘š ⊗ πœ‡∗ )
with ‖𝑇𝑖 ‖𝐿 ∞ (π‘š⊗πœ‡∗ ) ≤ 1. So, it follows from the Riesz convexity
theorem that ‖𝑇𝑖 ‖𝐿 𝑝 (π‘š⊗πœ‡∗ ) ≤ 1 for 1 < 𝑝 < ∞. Hence, from
Dunford-Schwartz’s ergodic theorem [18], we have that for
every 𝑓 ∈ 𝐿 𝑝 (π‘š ⊗ πœ‡∗ ) the multiple averages
max(π‘Ÿπœˆ−1 ,π‘Ÿπœˆ )−1
𝑖
(36)
× (π‘˜πœˆ , [πœ‘π‘˜1 +⋅⋅⋅+π‘˜πœˆ−1 πœ”∗ ]π‘Ÿ ) 𝑓 (𝑠)
(33)
Note here that if 𝑓 ∈ 𝐿 𝑝 (π‘š ⊗ πœ‡∗ ), 1 < 𝑝 < ∞, (or 𝑓 ∈
𝐿(𝑆 × Ω∗ )log+ 𝐿(𝑆 × Ω∗ )), then
1 𝑛
󡄨 󡄨
sup ∑ |𝑇|π‘˜ 󡄨󡄨󡄨𝑓󡄨󡄨󡄨 ∈ 𝐿 𝑝 (π‘š ⊗ πœ‡∗ )
𝑛≥1 𝑛 π‘˜=1
𝑛
𝜈
1
1
∑ ⋅ ⋅ ⋅ ∑ 𝑇1 (π‘˜1 , [πœ”∗ ]π‘Ÿ1 ) ⋅ ⋅ ⋅ π‘‡πœˆ
𝑛1 ⋅ ⋅ ⋅ π‘›πœˆ π‘˜ =1 π‘˜ =1
Proof. As already seen above, we can define the operators
𝑇1 , . . . , π‘‡πœˆ on 𝐿 1 (π‘š ⊗ πœ‡∗ ) as follows: for β„Ž ∈ 𝐿 1 (π‘š ⊗ πœ‡∗ )
and that if 1 < 𝑝 < ∞, then
Μƒ (𝑠, πœ”∗ ) = 𝑓
Μƒ (𝑠, [πœ”∗ ] )
𝑓
π‘Ÿ−π‘Ÿ0
Then, for every 𝑓 ∈ 𝐿 1 (π‘š), there exists a πœ‡∗ -null set 𝑁∗ such
Μƒ ∗ ∈
that for every πœ”∗ ∈ Ω∗ − 𝑁∗ , there exists a function 𝑓
[πœ” ]π‘ž−1
𝐿 1 (π‘š) with π‘ž = max(π‘Ÿ1 , . . . , π‘Ÿπœˆ ) such that the multiple averages
(𝑠)
π‘š-a.e.
(35)
(41)
6
Abstract and Applied Analysis
Finally, observe that excepting a πœ‡∗ -null set, we get
(𝑖)
(𝑖)
(𝑖)
π‘‡π‘–π‘˜ 𝑓 (𝑠, πœ”∗ ) = 𝑇[πœ”
π‘“πœ‘π‘˜ πœ”∗ (𝑠)
∗ ] 𝑇[πœ‘πœ”∗ ] ⋅ ⋅ ⋅ 𝑇 π‘˜−1 ∗
[πœ‘ πœ” ]
π‘Ÿ
π‘Ÿ
𝑖
π‘Ÿπ‘–
𝑖
= 𝑇𝑖 (π‘˜, [πœ”∗ ]π‘Ÿπ‘– ) π‘“πœ”∗ (𝑠)
π‘˜ = 1, 2, . . . ,
π‘š-a.e.,
𝑖 = 1, 2, . . . , 𝜈,
(42)
π‘‡πœˆ (π‘˜πœˆ , [πœ‘π‘˜1 +⋅⋅⋅+π‘˜πœˆ−1 πœ”∗ ]π‘Ÿ )
(43)
𝜈
π‘š-a.e.
π‘˜
π‘˜
(1)
(1)
2
𝑇1 1 𝑇2 2 𝑓 (𝑠, πœ”∗ ) = 𝑇[πœ”
⋅ ⋅ ⋅ 𝑇[πœ‘
∗]
π‘˜1 −1 πœ”∗ ] [𝑇2 𝑓] π‘˜1 ∗ (𝑠)
πœ‘ πœ”
π‘Ÿ
(1)
(1)
(2)
= 𝑇[πœ”
⋅ ⋅ ⋅ 𝑇[πœ‘
⋅⋅⋅
∗]
π‘˜1 −1 πœ”∗ ] 𝑇[πœ‘π‘˜1 πœ”∗ ]
π‘Ÿ
π‘Ÿ1
1
π‘Ÿ2
(2)
𝑇[πœ‘
π‘˜1 +π‘˜2 −1 πœ”∗ ] π‘“πœ‘π‘˜1 +π‘˜2 πœ”∗ (𝑠)
(44)
π‘Ÿ2
= 𝑇1 (π‘˜1 , [πœ”∗ ]π‘Ÿ1 ) 𝑇2 (π‘˜2 , [πœ‘π‘˜1 πœ”∗ ]π‘Ÿ )
π‘š-a.e.
To complete the proof of the above equality, assume that (43)
has already been established for the 𝜈−1 operators 𝑇2 , . . . , π‘‡πœˆ .
Then, it is easily verified by the induction hypothesis that (43)
holds for the 𝜈 operators 𝑇1 , . . . , π‘‡πœˆ . Hence, taking 𝑓(𝑠, πœ”∗ ) =
𝑓(𝑠)𝑒(πœ”∗ ) (𝑒(⋅) = 1), the theorem follows from the above
arguments.
In particular, if the operators in question are commutative
then we have the following.
(𝑖)
:
Theorem 9. Let π‘Ÿ1 , . . . , π‘Ÿπœˆ be 𝜈 positive integers. Let {𝑇[πœ”
∗]
π‘Ÿπ‘–
πœ”∗ ∈ Ω∗ }, 𝑖 = 1, 2, . . . , 𝜈, be a strongly F(𝑖)
π‘Ÿπ‘– (𝑖 = 1, 2, . . . , 𝜈)measurable commuting family of linear contractions on 𝐿 1 (π‘š)
(𝑖)
with ‖𝑇[πœ”
≤ 1. Then, for every 𝑓 ∈ 𝐿 1 (π‘š), there exists
∗] β€–
π‘Ÿπ‘–
𝐿 ∞ (π‘š)
a πœ‡∗ -null set 𝑁∗ such that for every πœ”∗ ∈ Ω∗ − 𝑁∗ , there exists
Μƒ ∗ ∈ 𝐿 (π‘š) with π‘ž = max(π‘Ÿ , . . . , π‘Ÿ ) such that
a function 𝑓
1
1
𝜈
[πœ” ]π‘ž−1
lim
1
𝑛 → ∞ π‘›πœˆ
𝑛−1
𝑛−1
π‘˜1 =0
π‘˜πœˆ =0
∑ ⋅ ⋅ ⋅ ∑ 𝑇1 (π‘˜1 , [πœ”∗ ]π‘Ÿ1 ) ⋅ ⋅ ⋅
π‘‡πœˆ (π‘˜πœˆ , [πœ‘π‘˜1 +⋅⋅⋅+π‘˜πœˆ−1 πœ”∗ ]π‘Ÿ ) 𝑓 (𝑠)
(45)
𝜈
Μƒ ∗ (𝑠)
=𝑓
[πœ” ]π‘ž−1
𝐿 ∞ (π‘š)
𝑛
𝑛
1
𝜈
π‘˜1 +⋅⋅⋅+π‘˜πœˆ−1
π‘‡πœˆ (π‘˜πœˆ , [πœ‘
(46)
∗
πœ” ]π‘Ÿ ) 𝑓 (𝑠)
𝜈
Μƒ ∗ (𝑠) π‘š-a.e. as 𝑒(𝑛) (= (𝑛 , . . . , 𝑛 ) ≥ (1, . . . ,
converge to 𝑓
1
𝜈
[πœ” ]π‘ž−1
1)) → ∞ in a sector of Z𝜈+ .
2
× π‘“ (𝑠, πœ”∗ )
π‘Ÿπ‘–
𝜈
1
1
∑ ⋅ ⋅ ⋅ ∑ 𝑇1 (π‘˜1 , [πœ”∗ ]π‘Ÿ1 ) ⋅ ⋅ ⋅
𝑛1 ⋅ ⋅ ⋅ π‘›πœˆ π‘˜ =1 π‘˜ =1
π‘Ÿ1
1
πœ”∗ ∈ Ω∗ }, 𝑖 = 1, 2, . . . , 𝜈, be strongly F(𝑖)
π‘Ÿπ‘– (𝑖 = 1, 2, . . . , 𝜈)measurable commuting family of linear contractions on 𝐿 1 (π‘š)
(𝑖)
with ‖𝑇[πœ”
≤ 1. Then, for every 𝑓 ∈ 𝐿 1 (π‘š), there exists
∗] β€–
a πœ‡∗ -null set 𝑁∗ such that for every πœ”∗ ∈ Ω∗ − 𝑁∗ , there exists
Μƒ ∗ ∈ 𝐿 (π‘š) with π‘ž = max(π‘Ÿ , . . . , π‘Ÿ ) such that
a function 𝑓
1
1
𝜈
[πœ” ]π‘ž−1
the multiple averages
In fact, we have for two operators 𝑇1 and 𝑇2
π‘˜
(𝑖)
:
Theorem 10. Let π‘Ÿ1 , . . . , π‘Ÿπœˆ be 𝜈 positive integers. Let {𝑇[πœ”
∗]
π‘Ÿπ‘–
π‘˜
𝑇1 1 ⋅ ⋅ ⋅ π‘‡πœˆπ‘˜πœˆ 𝑓 (𝑠, πœ”∗ ) = 𝑇1 (π‘˜1 , [πœ”∗ ]π‘Ÿ1 ) ⋅ ⋅ ⋅
× π‘“πœ”∗ (𝑠)
𝐢0 > 0 such that the ratios 𝑛𝑖 /𝑛𝑗 are bounded by 𝐢0 for
1 ≤ 𝑖, 𝑗 ≤ 𝜈 and all 𝑒(𝑛) = (𝑛1 , . . . , π‘›πœˆ ) (see [20, page 203]).
Appealing to Brunel-Dunford-Schwartz’ theorem (see [20,
Theorem 3.5, page 215]), we find that the multiple averages
(38) converge π‘š-a.e. in a sector Z𝜈+ . In addition, in this case,
Theorem 8 implies that the limit function depends essentially
only on the π‘ž − 1 random parameters. Hence, we have the
following.
π‘š-a.e.
In passing, we make mention of the a.e. convergence for
sectorial restricted random averages. We say that a sequence
𝑒(𝑛) ⊂ Z𝜈+ remains in a sector of Z𝜈+ if there is a constant
A sub-Markovian operator 𝑃∗ on 𝐿 ∞ (π‘š) means that 𝑃∗
is a positive linear contraction on 𝐿 ∞ (π‘š) with 1 subinvariant
under 𝑃∗ (i.e., 𝑃∗ 1 ≤ 1) such that lim𝑛 → ∞ 𝑃∗ β„Žπ‘› =
0 π‘š-a.e. for any (decreasing) sequence (β„Žπ‘› )𝑛≥1 ⊂ 𝐿 ∞ (π‘š)
with lim𝑛 → ∞ β„Žπ‘› = 0 π‘š-a.e. Let 𝑃 denote the positive linear
contraction on 𝐿 1 (π‘š) with 𝑃∗ as the adjoint operator of 𝑃.
When 𝜐 = π‘πœ‡ with a 𝑃-subinvariant function 𝑝 ≥ 0 (𝑝 =ΜΈ 0),
𝜐 is called a 𝑃-subinvariant measure. Here, the subinvariant
function is not necessarily integrable.
∗
: πœ”∗ ∈ Ω∗ } be a strongly F∗π‘Ÿ Now, let Π∗ = {𝑃[πœ”
∗]
π‘Ÿ
measurable family of sub-Markovian operators on 𝐿 ∞ (π‘š).
Let Π = {𝑃[πœ”∗ ]π‘Ÿ : πœ”∗ ∈ Ω∗ } be a strongly F∗π‘Ÿ -measurable
family of positive linear contractions on 𝐿 1 (π‘š), where each
𝑃[πœ”∗ ]π‘Ÿ is the adjoint operator of the corresponding operator
∗
𝑃[πœ”
∗ ] . Then, there exists a positive linear contraction π‘ˆ on
π‘Ÿ
𝐿 1 (π‘š ⊗ πœ‡∗ ) such that for any 𝑓 ∈ 𝐿 1 (π‘š ⊗ πœ‡∗ ),
π‘ˆπ‘“ (𝑠, πœ”∗ ) = [𝑃[πœ”∗ ]π‘Ÿ π‘“πœ‘π‘Ÿ0 πœ”∗ ] (𝑠)
π‘š × πœ‡∗ -a.e
(47)
In addition, if ‖𝑃[πœ”∗ ]π‘Ÿ ‖𝐿 (π‘š) ≤ 1 for all πœ”∗ ∈ Ω∗ , π‘ˆ is
∞
also a contraction on 𝐿 ∞ (π‘š ⊗ πœ‡∗ ) (cf. [10, 14]). The operator
π‘ˆ extends uniquely to a positive linear transformation on the
class of nonnegative 𝛽⊗F∗ -measurable functions defined on
𝑆 × Ω∗ (see [20, page 51]).
Lemma 11. Let the measure π‘š be finite. Assume that π‘š is
𝐸(π‘ˆ⋅ | ℘0 )-subinvariant. Then, π‘ˆ turns out to be a positive
Dunford-Schwartz operator on 𝐿 1 (π‘š ⊗ πœ‡∗ ).
Abstract and Applied Analysis
7
Proof. Since π‘š is assumed to be 𝐸(π‘ˆ⋅ | ℘0 )-subinvariant,
the function 1 (∈ 𝐿 1 (π‘š ⊗ πœ‡∗ )) is 𝐸(π‘ˆ⋅ | ℘0 )-subinvariant.
According to Lemma 1.5 of Woś [21], we get
π‘ˆ1 = 𝐸 (π‘ˆ1 | ℘0 ) ,
(48)
so that 1 is also π‘ˆ-subinvariant. Thus, for 𝑓 ∈ 𝐿 1 (π‘š ⊗ πœ‡∗ ) ∩
𝐿 ∞ (π‘š ⊗ πœ‡∗ ),
σ΅„©σ΅„© σ΅„©σ΅„©
σ΅„© σ΅„©
σ΅„© σ΅„©
σ΅„© σ΅„©
(49)
σ΅„©σ΅„©π‘ˆπ‘“σ΅„©σ΅„©∞ ≤ π‘ˆσ΅„©σ΅„©σ΅„©π‘“σ΅„©σ΅„©σ΅„©∞ = 󡄩󡄩󡄩𝑓󡄩󡄩󡄩∞ π‘ˆ1 ≤ 󡄩󡄩󡄩𝑓󡄩󡄩󡄩∞ .
Since π‘ˆ is a positive contraction on 𝐿 1 (π‘š ⊗ πœ‡∗ ), and π‘ˆ is a
positive Dunford-Schwartz operator on 𝐿 1 (π‘š ⊗ πœ‡∗ ).
Now, the above general theorems can also be applied to
sub-Markovian operators. For example, Theorem 7 yields the
following theorem which extends both Theorems 3.4 and
3.6 of Woś [21] (In proving the key Lemma 1.5 of [21], Woś
showed that if 𝑓 ∈ 𝐿 1 (π‘š ⊗ πœ‡∗ ) does not depend essentially on
parameter πœ”∗ , then π‘ˆπ‘“ = 𝐸(π‘ˆπ‘“ | ℘0 ). Using this equality,
he deduced π‘ˆ1 ≤ 1 from 𝐸(π‘ˆ1 | ℘0 ) ≤ 1 in order to
prove Theorems 2.8 and 3.4 of [21]. But obviously, 1 does not
necessarily belong to 𝐿 1 (π‘š ⊗ πœ‡∗ ) in case where π‘š is 𝜎-finite.
His arguments are of course correct in the case that π‘š is finite
(see, e.g., conditions (i) and (ii) in Theorem 12)).
Theorem 12. Let Π = {𝑃[πœ”∗ ]π‘Ÿ : πœ”∗ ∈ Ω∗ } be a strongly F∗π‘Ÿ measurable family of positive linear contractions on 𝐿 1 (π‘š),
where each 𝑃[πœ”∗ ]π‘Ÿ is the adjoint operator of the corresponding
∗
sub-Markovian operator 𝑃[πœ”
∗ ] . Assume either of the following
π‘Ÿ
conditions:
(i) π‘š is finite and 𝐸(π‘ˆ⋅ | ℘0 )-subinvariant,
(ii) π‘š is 𝜎-finite and ‖𝑃[πœ”∗ ]π‘Ÿ ‖𝐿 (π‘š) ≤ 1 for all πœ”∗ ∈ Ω∗ .
∞
Then, for any 𝑓 ∈ 𝐿 𝑝 (π‘š ⊗ πœ‡∗ ), 1 ≤ 𝑝 < ∞, there exists a
πœ‡∗ -null set 𝑁∗ such that for every πœ”∗ ∈ Ω∗ − 𝑁∗ , there exists
Μƒ ∗
a function 𝑓
[πœ” ]π‘Ÿ−π‘Ÿ ∈ 𝐿 𝑝 (π‘š) such that
0
𝑛
1
lim ∑ 𝑃[πœ”∗ ]π‘Ÿ 𝑃[πœ‘π‘Ÿ0 πœ”∗ ]π‘Ÿ ⋅ ⋅ ⋅ 𝑃[πœ‘(π‘˜−1)π‘Ÿ0 πœ”∗ ]π‘Ÿ π‘“πœ‘π‘˜π‘Ÿ0 πœ”∗ (𝑠)
𝑛→∞𝑛
π‘˜=1
Μƒ ∗
= 𝑓
[πœ” ]π‘Ÿ−π‘Ÿ (𝑠)
0
3. (𝐢, 𝛼)-Type Random Ergodic Theorems
In this section, we establish a (𝐢, 𝛼)-type random ergodic
theorem for measure-preserving transformations on 𝑆. In this
section, we assume that (𝑆, 𝛽, π‘š) is a probability space. For
real 𝛼 > −1 and 𝑛(= 0, 1, 2, . . .), let 𝐴𝛼𝑛 be the (𝐢, 𝛼) coefficient
of order 𝛼, which is defined by the generating function
∞
1
=
𝐴𝛼𝑛 𝑑𝑛 ,
∑
(1 − 𝑑)𝛼+1 𝑛=0
0 < 𝑑 < 1.
(53)
Then, we can easily check that 𝐴𝛼𝑛 is decreasing in 𝑛 for −1 <
𝛼 < 0 and increasing in 𝑛 for 𝛼 > 0. For 𝛼 > −1, we have
𝐴𝛼𝑛 > 0,
𝐴𝛼0 = 1,
𝐴0𝑛 = 1,
𝐴𝛼−1
= 𝐴𝛼𝑛 − 𝐴𝛼𝑛−1 ,
𝑛
𝑛
𝑛+𝛼
𝐴𝛼𝑛 = ∑ 𝐴𝛼−1
𝑛−π‘˜ = ( 𝑛 ) ,
π‘˜=0
𝐴𝛼𝑛 ∼
𝑛𝛼
Γ (𝛼 + 1)
(𝑛 󳨀→ ∞) ,
(54)
and moreover,
𝑛𝛼
(𝑛 + 1)𝛼
≤ 𝐴𝛼𝑛 ≤
Γ (𝛼 + 1)
Γ (𝛼 + 1)
𝐴𝛼−1
𝑛
𝑛𝛼−1
≤
Γ (𝛼)
for 0 < 𝛼 ≤ 1,
(55)
for 𝑛 > 0.
In general, using Hille’s theorem (see [22, Theorem 8]), we
have the following.
(50)
Theorem 13. Let π‘Ÿ0 be any fixed integer with 1 ≤ π‘Ÿ0 ≤ π‘Ÿ. Let
{𝑇[πœ”∗ ]π‘Ÿ : πœ”∗ ∈ Ω∗ } be a strongly F∗π‘Ÿ -measurable family of
linear contractions on 𝐿 𝑝 (π‘š) (1 ≤ 𝑝 < ∞). Let 𝛼 > 0 be fixed.
If for all 𝑓 ∈ 𝐿 𝑝 (π‘š ⊗ πœ‡∗ ) there exists a πœ‡∗ -null set 𝑁∗ such that
∗
for any πœ”∗ ∈ Ω∗ − 𝑁∗ there exists a function 𝑓[πœ”
∈ 𝐿 𝑝 (π‘š)
∗]
π‘Ÿ−π‘Ÿ0
such that
(51)
(i) lim𝑛 → ∞ (1/𝐴𝛼𝑛 ) ∑π‘›π‘˜=0 𝐴𝛼−1
𝑛−π‘˜ 𝑇[πœ”∗ ]π‘Ÿ 𝑇[πœ‘π‘Ÿ0 πœ”∗ ]π‘Ÿ ⋅ ⋅ ⋅
Μƒ ∗ (𝑠) for π‘š-almost all
𝑇[πœ‘π‘˜π‘Ÿ0 πœ”∗ ]π‘Ÿ π‘“πœ‘(π‘˜+1)π‘Ÿ0 πœ”∗ (𝑠) = 𝑓
[πœ” ]π‘Ÿ−π‘Ÿ0
𝑠 ∈ 𝑆,
π‘š-a.e.,
and that if 1 < 𝑝 < ∞, then
σ΅„©σ΅„© 𝑛
σ΅„©σ΅„© 1
lim σ΅„©σ΅„©σ΅„© ∑ 𝑃[πœ”∗ ]π‘Ÿ 𝑃[πœ‘π‘Ÿ0 πœ”∗ ]π‘Ÿ ⋅ ⋅ ⋅ 𝑃[πœ‘(π‘˜−1)π‘Ÿ0 πœ”∗ ]π‘Ÿ π‘“πœ‘π‘˜π‘Ÿ0 πœ”∗ (⋅)
𝑛→∞ 󡄩𝑛
σ΅„©σ΅„© π‘˜=1
σ΅„©σ΅„©
σ΅„©σ΅„©
Μƒ ∗
σ΅„©σ΅„©
−𝑓
=0
(⋅)
[πœ” ]π‘Ÿ−π‘Ÿ0
σ΅„©σ΅„©
󡄩󡄩𝐿 (π‘š)
𝑝
Proof of Theorem 1. In view of Lemma 11, either condition
(i) or condition (ii) guarantees that π‘ˆ is a positive DunfordSchwartz operator on 𝐿 1 (π‘š ⊗ πœ‡∗ ). Hence, we can apply
Theorem 7 to conclude that Theorem 12 follows.
and that if π‘š is finite and 𝑓 ∈ 𝐿(𝑆 × Ω∗ )log+ 𝐿(𝑆 × Ω∗ ), then
σ΅„©σ΅„© 𝑛
σ΅„©σ΅„© 1
lim σ΅„©σ΅„©σ΅„© ∑ 𝑃[πœ”∗ ]π‘Ÿ 𝑃[πœ‘π‘Ÿ0 πœ”∗ ]π‘Ÿ ⋅ ⋅ ⋅ 𝑃[πœ‘(π‘˜−1)π‘Ÿ0 πœ”∗ ]π‘Ÿ π‘“πœ‘π‘˜π‘Ÿ0 πœ”∗ (⋅)
𝑛→∞ σ΅„©
σ΅„©σ΅„© 𝑛 π‘˜=1
(52)
σ΅„©σ΅„©
σ΅„©σ΅„©
Μƒ
−𝑓[πœ”∗ ]π‘Ÿ−π‘Ÿ (⋅) σ΅„©σ΅„©σ΅„©
= 0.
σ΅„©σ΅„©
0
󡄩𝐿 1 (π‘š)
(𝑛+1)
𝑇[πœ”∗ ]π‘Ÿ 𝑇[πœ‘π‘Ÿ0 πœ”∗ ]π‘Ÿ ⋅ ⋅ ⋅
(ii) limπœ† → 1+0 (πœ† − 1) ∑∞
𝑛=0 πœ†
Μƒ ∗ (𝑠),
𝑇[πœ‘π‘˜π‘Ÿ0 πœ”∗ ]π‘Ÿ π‘“πœ‘(π‘˜+1)π‘Ÿ0 πœ”∗ (𝑠) = 𝑓
[πœ” ]π‘Ÿ−π‘Ÿ
0
𝛼
(iii) lim𝑛 → ∞ (1/𝑛 )𝑇[πœ”∗ ]π‘Ÿ 𝑇[πœ‘π‘Ÿ0 πœ”∗ ]π‘Ÿ ⋅ ⋅ ⋅ 𝑇[πœ‘π‘›π‘Ÿ0 πœ”∗ ]π‘Ÿ π‘“πœ‘(𝑛+1)π‘Ÿ0 πœ”∗
(𝑠) = 0,
for π‘š-almost all 𝑠 ∈ 𝑆. Conversely, if (ii) holds but (iii) be
replaced by the condition that there exists a 𝛽⊗F∗ -measurable
8
Abstract and Applied Analysis
∗
= (𝑓 ⋅ 𝑒) (𝑠, πœ”∗ )
function 𝑀(𝑠, πœ”∗ ) finite almost everywhere such that excepting
a πœ‡∗ -null set
󡄨󡄨
󡄨󡄨
󡄨
󡄨󡄨 1 𝑛 𝛽−1
󡄨󡄨 ∑ 𝐴 𝑇[πœ”∗ ] 𝑇[πœ‘π‘Ÿ0 πœ”∗ ] ⋅ ⋅ ⋅ 𝑇[πœ‘π‘˜π‘Ÿ0 πœ”∗ ] π‘“πœ‘(π‘˜+1)π‘Ÿ0 πœ”∗ (𝑠)󡄨󡄨󡄨
󡄨󡄨
󡄨󡄨 𝛽
𝑛−π‘˜
π‘Ÿ
π‘Ÿ
π‘Ÿ
󡄨󡄨
󡄨󡄨 𝐴 𝑛 𝑛=0
≤ π‘€πœ”∗ (𝑠)
π‘š ⊗ πœ‡∗ -a.e.,
󡄨󡄨
󡄨󡄨
󡄨󡄨 1 𝑛
∗ 󡄨󡄨󡄨
sup 󡄨󡄨󡄨 𝛼 ∑ 𝐴𝛼−1
(𝑓
⋅
𝑒)
(𝑠,
πœ”
)
󡄨󡄨
𝑛−π‘˜
󡄨󡄨
𝑛≥0 󡄨󡄨󡄨 𝐴 𝑛 π‘˜=0
󡄨
∈ 𝐿 𝑝 (π‘š ⊗ πœ‡∗ ) (⊂ 𝐿 1 (π‘š ⊗ πœ‡∗ )) ,
(56)
π‘š-a.e.,
1 𝑛 𝛼−1
∑ 𝐴 𝑛−π‘˜ (𝑓 ⋅ 𝑒) (Φ∗π‘˜ (𝑠, πœ”∗ ))
𝑛 → ∞ 𝐴𝛼
𝑛 π‘˜=0
lim
for all 𝑛 ≥ 0, where 𝛽 ≥ 0, then (i) holds for all 𝛼 > 𝛽.
∗
= (𝑓 ⋅ 𝑒) (𝑠, πœ”∗ )
Note here that in Theorem 13, (ii) and (iii) do not
necessarily imply (i) in general.
Theorem 14. Let π‘Ÿ0 be any fixed integer with 1 ≤ π‘Ÿ0 ≤ π‘Ÿ and
{πœ“[πœ”∗ ]π‘Ÿ : πœ”∗ ∈ Ω∗ } be a 𝛽 ⊗ F∗π‘Ÿ -measurable family of measurepreserving transformations on 𝑆. Let 0 < 𝛼 < 1, 𝛼𝑝 > 1, and
𝑓 ∈ 𝐿 𝑝 (π‘š). Then, there exists a πœ‡∗ -null set 𝑁∗ such that for
∗
every πœ”∗ ∈ Ω∗ −𝑁∗ , there exists a function 𝑓[πœ”
(⋅) ∈ 𝐿 𝑝 (π‘š)
∗]
π‘Ÿ−π‘Ÿ0
such that
(59)
Furthermore, applying Lemma 6 to the operator 𝑇∗ induced
by Φ∗ , it follows that
∗
= lim (πœ† − 1) ∑ πœ†−(𝑛+1) 𝑓 (πœ“[πœ‘π‘˜π‘Ÿ0 πœ”∗ ]π‘Ÿ ⋅ ⋅ ⋅ πœ“[πœ”∗ ]π‘Ÿ 𝑠)
=
𝑛=0
∗
𝑓[πœ”
∗]
π‘Ÿ−π‘Ÿ0
(𝑠)
π‘š-π‘Ž.𝑒.,
󡄨󡄨
󡄨󡄨 1 𝑛
lim {∫ 󡄨󡄨󡄨 𝛼 ∑ 𝐴𝛼−1
𝑛−π‘˜ 𝑓 (πœ“[πœ‘π‘˜π‘Ÿ0 πœ”∗ ]π‘Ÿ ⋅ ⋅ ⋅ πœ“[πœ”∗ ]π‘Ÿ 𝑠)
𝑛→∞
𝑆 󡄨󡄨󡄨 𝐴 𝑛 π‘˜=0
∗
−𝑓[πœ”
∗]
π‘Ÿ−π‘Ÿ0
(60)
so that to complete the proof of the theorem, we may take
∗
∗
𝑓[πœ”
(𝑠) = (𝑓 ⋅ 𝑒)[πœ”∗ ]π‘Ÿ−π‘Ÿ (𝑠) .
∗]
π‘Ÿ−π‘Ÿ
0
∞
πœ† → 1+0
∗
(𝑓 ⋅ 𝑒) (𝑠, πœ”∗ ) = (𝑓 ⋅ 𝑒) (𝑠, [πœ”∗ ]π‘Ÿ−π‘Ÿ0 ) ,
𝑛
1
lim
∑ 𝐴𝛼−1
𝑛−π‘˜ 𝑓 (πœ“[πœ‘π‘˜π‘Ÿ0 πœ”∗ ]π‘Ÿ ⋅ ⋅ ⋅ πœ“[πœ”∗ ]π‘Ÿ 𝑠)
𝑛 → ∞ 𝐴𝛼
𝑛 π‘˜=0
π‘š ⊗ πœ‡∗ -a.e.
0
(61)
Remark 15. For example, if π‘Ÿ0 = π‘Ÿ = 1 in Theorem 14, then
∗
∗
𝑓[πœ”
∗ ]π‘Ÿ−π‘Ÿ (𝑠) = 𝑓 (𝑠). It is worthwhile to note that if 0 < 𝛼 < 1
0
and 𝛼𝑝 = 1 (so, 𝑝 = 𝛼−1 > 1), then the pointwise (𝐢, 𝛼)convergence for Φ∗ does not hold in general (see [24]). For
the case of a positive linear contraction on 𝐿 1/𝛼 (π‘š ⊗ πœ‡∗ ), see
Irmisch [25].
In particular, applying Irmisch’s theorem to subMarkovian operators, we have the following.
1/𝑝
󡄨󡄨𝑝
󡄨󡄨
󡄨
(𝑠) 󡄨󡄨 π‘‘π‘š} = 0.
󡄨󡄨
󡄨
(57)
Proof. Define the skew product Φ∗ of πœ™ (the one-sided shift
transformation of Ω∗ ) and {πœ“[πœ”∗ ]π‘Ÿ } as follows:
Theorem 16. Let Π = {𝑃[πœ”∗ ]π‘Ÿ : πœ”∗ ∈ Ω∗ } be a strongly F∗π‘Ÿ measurable family of positive linear contractions on 𝐿 1 (π‘š),
where each 𝑃[πœ”∗ ]π‘Ÿ is the adjoint operator of the corresponding
∗
sub-Markovian operator 𝑃[πœ”
∗ ] . Assume that π‘š is 𝐸(π‘ˆ⋅ | ℘0 )π‘Ÿ
subinvariant. Let 0 < 𝛼 < 1, 𝛼𝑝 > 1 and 𝑓 ∈ 𝐿 𝑝 (π‘š). Then,
there exists a πœ‡∗ -null set 𝑁∗ such that for every πœ”∗ ∈ Ω∗ − 𝑁∗ ,
∗
(⋅) ∈ 𝐿 𝑝 (π‘š) such that
there exists a function 𝑓[πœ”
∗]
π‘Ÿ−π‘Ÿ
0
∗
∗
π‘Ÿ0
∗
Φ (𝑠, πœ” ) = (πœ“[πœ”∗ ]π‘Ÿ 𝑠, πœ‘ πœ” ) ,
∗
∗
(𝑠, πœ” ) ∈ 𝑆 × Ω .
(58)
Let (𝑓 ⋅ 𝑒)(𝑠, πœ”∗ ) = 𝑓(𝑠) ⋅ 𝑒(πœ”∗ ), where 𝑒(πœ”∗ ) = 1 for
all πœ”∗ ∈ Ω∗ . Then, (𝑓 ⋅ 𝑒) ∈ 𝐿 𝑝 (π‘š ⊗ πœ‡∗ ). Since 𝐿 𝑝 (π‘š ⊗
πœ‡∗ ) is reflexive, we see from Yosida-Kakutani’s mean ergodic
theorem [23] and Déniel’s theorem [24] that there exists a
function (𝑓 ⋅ 𝑒)∗ ∈ 𝐿 𝑝 (π‘š ⊗ πœ‡∗ ) such that
1 𝑛 𝛼−1
∑ 𝐴 𝑛−π‘˜ 𝑃[πœ”∗ ]π‘Ÿ 𝑃[πœ‘π‘Ÿ0 πœ”∗ ]π‘Ÿ ⋅ ⋅ ⋅ 𝑃[πœ‘π‘˜π‘Ÿ0 πœ”∗ ]π‘Ÿ
𝑛 → ∞ 𝐴𝛼
𝑛 π‘˜=0
lim
∞
= lim (πœ† − 1) ∑ πœ†−(𝑛+1) 𝑃[πœ”∗ ]π‘Ÿ 𝑃[πœ‘π‘Ÿ0 πœ”∗ ]π‘Ÿ ⋅ ⋅ ⋅ 𝑃[πœ‘π‘˜π‘Ÿ0 πœ”∗ ]π‘Ÿ
πœ† → 1+0
𝑛=0
∗
= 𝑓[πœ”
(𝑠)
∗]
π‘Ÿ−π‘Ÿ
0
π‘š-π‘Ž.𝑒.
(62)
󡄨󡄨
{
󡄨󡄨 1 𝑛
lim { ∬ 󡄨󡄨󡄨 𝛼 ∑ 𝐴𝛼−1
(𝑓 ⋅ 𝑒) (Φ∗π‘˜ (𝑠, πœ”∗ ))
𝑛→∞
󡄨󡄨 𝐴 𝑛 π‘˜=0 𝑛−π‘˜
{𝑆×Ω∗ 󡄨
1/𝑝
󡄨󡄨𝑝
󡄨󡄨
∗
−(𝑓 ⋅ 𝑒) (𝑠, πœ” ) 󡄨󡄨󡄨 π‘‘π‘š ⊗ πœ‡ } = 0,
󡄨󡄨
󡄨
∗
∗
(𝑓 ⋅ 𝑒) (Φ∗ (𝑠, πœ”∗ ))
∗
Proof. In view of Lemma 11, π‘ˆ is a positive linear contraction
on 𝐿 1 (π‘š ⊗ πœ‡∗ ) as well as on 𝐿 ∞ (π‘š ⊗ πœ‡∗ ). Thus, it follows from
the Riesz convexity theorem that β€–π‘ˆβ€–πΏ 𝑝 (π‘š⊗πœ‡∗ ) ≤ 1 for 1 < 𝑝 <
∞. Therefore, we reach the assertion of Theorem 16 through
Theorems 7 and 13 appealed to Irmisch’s theorem [25].
Remark 17. The relations between the random ergodic limit
functions and the random parameters have been investigated
Abstract and Applied Analysis
9
(with satisfactory formulations) only in discrete parameter
cases so far. So, it is very interesting to study the continuous
analogs of the theorems obtained above. But no continuous
results are known from the point of view of the dependence
of the limit functions on the random parameters. Here, it is
worthwhile to notice that Anzai has obtained a continuous
version of Kakutani’s random ergodic theorem for Brownian
motion in continuous parameter cases (see [26]). Let πœ‰ =
(πœ‰π‘‘ )𝑑≥0 (πœ‰π‘‘ = πœ‰π‘‘ (πœ”), πœ‰0 = 0) be a Brownian motion (or Wiener
process) on a probability space (Ω, F, πœ‡). This process has
independent increments; that is, for arbitrary 𝑑1 < 𝑑2 <
⋅ ⋅ ⋅ < 𝑑𝑛 , the random variables πœ‰π‘‘2 − πœ‰π‘‘1 , . . . , πœ‰π‘‘π‘› − πœ‰π‘‘π‘›−1 are
independent. In fact, since the process is Gaussian with
𝐸(πœ‰π‘‘ ) = 0 and 𝐸(πœ‰π‘‘ πœ‰π‘  ) = min(𝑑, 𝑠) by definition, it is sufficient
to verify only that the increments are uncorrelated. Thus, if
𝑠 < 𝑑 < 𝑒 < V, then
𝐸 ((πœ‰π‘‘ − πœ‰π‘  ) (πœ‰V − πœ‰π‘’ ))
=
1
((𝑑 − 𝑠) + (V − 𝑒) − |𝑑 − 𝑠 + V − 𝑒|) = 0.
2
(63)
Let {πœ“π‘‘ : 𝑑 ≥ 0} be an arbitrary ergodic measurable
semiflow on a finite measure space (𝑆, 𝛽, π‘š). Then, Anzai’s
result may be stated as follows: for any 𝑓 ∈ 𝐿 1 (π‘š) and for
almost all πœ” ∈ Ω, there exists a null set 𝑁 = 𝑁𝑓 (πœ”) ∈ F such
that
1 𝛼
∫ 𝑓 (πœ“πœ‰π‘‘ (πœ”) 𝑠) 𝑑𝑑 = ∫ 𝑓 (𝑠) π‘‘π‘š
𝛼→∞𝛼 0
𝑆
lim
(64)
holds for any 𝑠 ∈ 𝑆 − 𝑁. This is an immediate consequence
of the ergodicity of the measure-preserving skew product
semiflow {Ψ𝑑 𝑑 ≥ 0} defined by Ψ𝑑 (πœ”, 𝑠) = (πœ‘π‘‘ πœ”,πœ“πœ‰π‘‘ (πœ”) 𝑠),
where {πœ‘π‘‘ : 𝑑 ≥ 0} is the ergodic semiflow on Ω given by
πœ‰π‘Ž (πœ‘π‘‘ πœ”) = πœ‰π‘Ž+𝑑 (πœ”). It is an interesting and important problem
to generalize Anzai’s result for Brownian motions to the case
of contraction operator quasisemigroups in 𝐿 1 (π‘š) associated
with {πœ‘π‘‘ }. We do not discuss it in the present paper.
4. Applications to Nonlinear Random
Ergodic Theorems
π‘ˆ[πœ”∗ ]π‘Ÿ = 𝑇[πœ”∗ ]π‘Ÿ + [(𝐼 − 𝑇) 𝑔]πœ”∗ ,
πœ”∗ ∈ Ω∗ ,
(66)
where
𝑇𝑓 (𝑠, πœ”∗ ) = [𝑇[πœ”∗ ]π‘Ÿ π‘“πœ‘π‘Ÿ0 πœ”∗ ] (𝑠) ,
(67)
for 𝑓 ∈ 𝐿 1 (π‘š ⊗ πœ‡∗ ) (cf. [11, 27]). For each 𝑓 ∈ 𝐿 1 (π‘š ⊗ πœ‡∗ ),
we define a sequence of random functions {𝑉𝑓 (𝑛, πœ”∗ ) : πœ”∗ ∈
Ω∗ }, 𝑛 = 0, 1, . . ., in 𝐿 1 (π‘š) inductively by
𝑉𝑓 (0, πœ”∗ ) = π‘“πœ”∗ ,
𝑉𝑓 (1, πœ”∗ ) = π‘“πœ”∗ + π‘ˆ[πœ”∗ ]π‘Ÿ 𝑉𝑓 (0, πœ‘π‘Ÿ0 πœ”∗ ) ,
(68)
..
.
𝑉𝑓 (𝑛 + 1, πœ”∗ ) = π‘“πœ”∗ + π‘ˆ[πœ”∗ ]π‘Ÿ 𝑉𝑓 (𝑛, πœ‘π‘Ÿ0 πœ”∗ ) .
Theorem 18. Let {𝑉𝑓 (𝑛, πœ”∗ ) : πœ”∗ ∈ Ω∗ } be the sequence
of random functions associated with 𝑓 ∈ 𝐿 1 (π‘š ⊗ πœ‡∗ ) which
is determined by a random affine system {(𝑇[πœ”∗ ]π‘Ÿ , π‘ˆ[πœ”∗ ]π‘Ÿ , [(𝐼 −
𝑇)𝑔]πœ”∗ , πœ‘π‘Ÿ0 ) : πœ”∗ ∈ Ω∗ } given in 𝐿 1 (π‘š) with ‖𝑇[πœ”∗ ]π‘Ÿ ‖𝐿 (π‘š) ≤
∞
1. Then, there exists a πœ‡∗ -null set 𝑁∗ such that for any πœ”∗ ∈
Μƒ ∗
Ω∗ − 𝑁∗ , there exists a function 𝑓
[πœ” ]π‘Ÿ−π‘Ÿ ∈ 𝐿 1 (π‘š) such that
0
lim
𝑛→∞
The random ergodic theorems obtained above can be applied
to the nonlinear random ergodic theorems for affine systems
(see [27]). An affine operator π‘ˆ on 𝐿 1 (π‘š) is an operator of the
type π‘ˆπ‘“ = 𝑇𝑓 + β„Ž, where 𝑇 is a linear contraction on 𝐿 1 (π‘š),
and β„Ž is a fixed element of 𝐿 1 (π‘š). Then, π‘ˆ is nonlinear and
nonexpansive. The fixed points of π‘ˆ are solutions of Poisson’s
equation for 𝑇, which is (𝐼 − 𝑇)𝑓 = β„Ž. When we assume 𝑇 to
be mean ergodic, the averages
1 𝑛
1 𝑛 π‘˜−1
1 𝑛 π‘˜
∑ π‘ˆ 𝑓 = ∑ π‘‡π‘˜ 𝑓 + ∑ ∑ 𝑇𝑗 β„Ž
𝑛 π‘˜=1
𝑛 π‘˜=1
𝑛 π‘˜=1 𝑗=0
𝑇)πœ‰ will yield almost everywhere convergence of the averages
of π‘ˆπ‘› 𝑓 for any 𝑓 ∈ 𝐿 1 (π‘š). Now, let {(𝑇[πœ”∗ ]π‘Ÿ , π‘ˆ[πœ”∗ ]π‘Ÿ , πœ‰πœ”∗ , πœ‘π‘Ÿ0 ) :
πœ”∗ ∈ Ω∗ } be a random affine system on 𝐿 1 (π‘š) such that
{𝑇[πœ”∗ ]π‘Ÿ : πœ”∗ ∈ Ω∗ } is a strongly measurable family of linear
contractions on 𝐿 1 (π‘š) as well as on 𝐿 ∞ (π‘š) and such that for
some 𝑔 ∈ 𝐿 1 (π‘š ⊗ πœ‡∗ ),
𝑉𝑓 (𝑛, πœ”∗ ) (𝑠)
𝑛+1
Μƒ ∗
=𝑓
[πœ” ]π‘Ÿ−π‘Ÿ (𝑠)
0
π‘š-π‘Ž.𝑒.
(69)
Proof. It follows that there exists a πœ‡∗ -null set 𝑁1∗ such that
for any πœ”∗ ∈ Ω∗ − 𝑁1∗ ,
𝑉𝑓 (𝑛, πœ”∗ ) (𝑠) = π‘“πœ”∗ (𝑠)
𝑛
+ ∑ 𝑇[πœ”∗ ]π‘Ÿ 𝑇[πœ‘π‘Ÿ0 πœ”∗ ]π‘Ÿ ⋅ ⋅ ⋅ 𝑇[πœ‘(π‘˜−1)π‘Ÿ0 πœ”∗ ]π‘Ÿ π‘“πœ‘π‘˜π‘Ÿ0 πœ”∗ (𝑠)
π‘˜=1
(65)
converge if and only if β„Ž ∈ (𝐼 − 𝑇)𝐿 1 (π‘š). If β„Ž ∈ (𝐼 − 𝑇)𝐿 1 (π‘š),
then there exists a unique πœ‰ ∈ (𝐼 − 𝑇)𝐿 1 (π‘š) such that (𝐼 −
𝑇)πœ‰ = β„Ž, and the limit of (1/𝑛) ∑π‘›π‘˜=1 π‘ˆπ‘˜ 𝑓 is 𝐸𝑓+πœ‰, where 𝐸𝑓 is
the limit of (1/𝑛) ∑π‘›π‘˜=1 π‘‡π‘˜ 𝑓. Therefore, iterating π‘ˆπ‘“ = 𝑇𝑓+(𝐼−
− 𝑇[πœ”∗ ]π‘Ÿ 𝑇[πœ‘π‘Ÿ0 πœ”∗ ]π‘Ÿ ⋅ ⋅ ⋅ 𝑇[πœ‘(𝑛−1)π‘Ÿ0 πœ”∗ ]π‘Ÿ π‘”πœ‘π‘›π‘Ÿ0 πœ”∗ (𝑠)
𝑛−1
+ ∑ 𝑇[πœ”∗ ]π‘Ÿ 𝑇[πœ‘π‘Ÿ0 πœ”∗ ]π‘Ÿ ⋅ ⋅ ⋅ 𝑇[πœ‘(π‘˜−1)π‘Ÿ0 πœ”∗ ]π‘Ÿ π‘”πœ‘π‘˜π‘Ÿ0 πœ”∗ (𝑠)
π‘˜=1
+ π‘”πœ”∗ (𝑠) .
(70)
10
Abstract and Applied Analysis
5. Examples
Moreover, we have
∗
lim
𝑛→∞
∗
𝑓 (𝑠, πœ” )
𝑔 (𝑠, πœ” )
= lim
= 0 π‘š-a.e.,
𝑛→∞ 𝑛 + 1
𝑛+1
lim
𝑛→∞𝑛
1
𝑇 ∗ 𝑇 π‘Ÿ0 ∗ ⋅ ⋅ ⋅ 𝑇[πœ‘(𝑛−1)π‘Ÿ0 πœ”∗ ]π‘Ÿ π‘”πœ‘π‘›π‘Ÿ0 πœ”∗ (𝑠)
+ 1 [πœ” ]π‘Ÿ [πœ‘ πœ” ]π‘Ÿ
(71)
= 0 π‘š-a.e.
Therefore, by Theorem 7, there exists a πœ‡∗ -null set 𝑁2∗ such
∗
,
that for every πœ”∗ ∈ Ω∗ − 𝑁2∗ , there exist functions 𝑓[πœ”
∗]
π‘Ÿ−π‘Ÿ0
∗
𝑔[πœ”∗ ]π‘Ÿ−π‘Ÿ ∈ 𝐿 1 (π‘š) such that
0
lim
𝑉𝑓 (𝑛, πœ”∗ ) (𝑠)
𝑛→∞
𝑛+1
∗
= 𝑓[πœ”
(𝑠)
∗]
π‘Ÿ−π‘Ÿ
0
∗
+ 𝑔[πœ”
(𝑠)
∗]
π‘Ÿ−π‘Ÿ
0
(72)
π‘š-π‘Ž.𝑒.
Hence, the theorem follows by putting 𝑁∗ = 𝑁1∗ ∪ 𝑁2∗ and
Μƒ (𝑠, [πœ”∗ ] ) = 𝑓∗ (𝑠, [πœ”∗ ] ) + 𝑔∗ (𝑠, [πœ”∗ ] ) . (73)
𝑓
π‘Ÿ−π‘Ÿ0
π‘Ÿ−π‘Ÿ0
π‘Ÿ−π‘Ÿ0
As far as we are concerned with the ergodic behaviors
of CesaΜ€ro-type processes for nonexpansive operators on 𝐿 𝑝 ,
one can only expect weak convergence in general. In fact,
the pointwise convergence of the (𝐢, 1) averages of nonlinear
and nonexpansive operators on 𝐿 𝑝 may fail to hold. In
addition, these (𝐢, 1) averages do not need converge in
the strong operator topology of 𝐿 𝑝 (see [27, 28]). The socalled nonlinear sums introduced by Wittmann [29] make
it possible to consider the pointwise convergence and the
strong convergence under some additional conditions. To
make the most of advantageous results in the theory of linear
ergodic theorems, it is very rational to consider a class of
affine operators as a model case (cf. [27]). Under the above
setting, observe that
Example 1. Let {πœπœ” : πœ” ∈ Ω} be a 𝛽 ⊗ F-measurable family of
π‘š-measure-preserving transformations on 𝑆, and let πœ™ be πœ‡measure-preserving transformation on Ω. Then, for any 𝑓 ∈
𝐿 1 (π‘š), there exists a πœ‡-null set 𝑁∗ ∈ 𝛽 such that for each
πœ” ∈ Ω − 𝑁∗ , there exists a function π‘“πœ”∗ ∈ 𝐿 1 (π‘š) such that
1 𝑛
∑ 𝑓 (πœπœ™π‘˜−1 πœ” ⋅ ⋅ ⋅ πœπœ” 𝑠) = π‘“πœ”∗ (𝑠)
𝑛→∞𝑛
π‘˜=1
lim
π‘˜
Φ (𝑠, πœ”) = (πœπœ” 𝑠, πœ™πœ”) ,
+ ∑ [𝑇[πœ”∗ ]π‘Ÿ 𝑇[πœ‘π‘Ÿ0 πœ”∗ ]π‘Ÿ ⋅ ⋅ ⋅ 𝑇[πœ‘(𝑗−1)π‘Ÿ0 πœ”∗ ]π‘Ÿ π‘”πœ‘π‘—π‘Ÿ0 πœ”∗ ] (𝑠)
+ 𝑔[πœ”∗ ]π‘Ÿ (𝑠) ,
π‘š {𝑠1 } = ⋅ ⋅ ⋅ = π‘š {𝑠𝑝 } =
πœ‡ {πœ”1 } = ⋅ ⋅ ⋅ = πœ‡ {πœ”2π‘ž } =
πœπœ”2𝑗−1 𝑠𝑖 = {
1 𝑛
∑ π‘ˆ[πœ”∗ ]π‘Ÿ π‘ˆ[πœ‘π‘Ÿ0 πœ”∗ ]π‘Ÿ ⋅ ⋅ ⋅ π‘ˆ[πœ‘(π‘˜−1)π‘Ÿ0 πœ”∗ ]π‘Ÿ π‘“πœ‘π‘˜π‘Ÿ0 πœ”∗ (𝑠)
𝑛→∞𝑛
π‘˜=1
lim
=
∗
𝑓[πœ”
∗]
π‘Ÿ−π‘Ÿ0
(𝑠)
π‘š-π‘Ž.𝑒.
𝑠𝑖+1 ,
𝑠1 ,
(75)
(𝑝: integer, 𝑝 ≥ 2) ,
1
2π‘ž
(π‘ž: integer, 2π‘ž ≥ 𝑝) ,
if 1 ≤ 𝑖 ≤ 𝑝 − 1,
,
if 𝑖 = 𝑝,
πœπœ”2𝑗 = πœπœ”−12𝑗−1 ,
(78)
𝑗 = 1, 2, . . . , π‘ž,
𝑗 = 1, 2, . . . , π‘ž,
if 1 ≤ 𝑗 ≤ 2π‘ž − 1,
if 𝑗 = 2π‘ž.
Thus, for instance, if we consider a function 𝑓(𝑠, πœ”) defined
by
= 0,
Theorem 19. Let {(𝑇[πœ”∗ ]π‘Ÿ , π‘ˆ[πœ”∗ ]π‘Ÿ , [(𝐼 − 𝑇)𝑔]πœ”∗ , πœ‘π‘Ÿ0 ) : πœ”∗ ∈ Ω∗ }
be a random affine system given in 𝐿 1 (π‘š) with ‖𝑇[πœ”∗ ]π‘Ÿ ‖𝐿 (π‘š) ≤
∞
1. If 𝑓 ∈ 𝐿 1 (π‘š ⊗ πœ‡∗ ), then there exists a πœ‡∗ -null set 𝑁∗ such
∗
∈
that for any πœ”∗ ∈ Ω∗ − 𝑁∗ , there exists a function 𝑓[πœ”
∗]
π‘Ÿ−π‘Ÿ0
𝐿 1 (π‘š) such that
1
𝑝
Ω = {πœ”1 , . . . , πœ”2π‘ž } ,
π‘˜ = 1, 2, . . . .
Then, Theorem 7, together with Theorems 2.1 and 2.2 of [27],
yields the following theorem.
(77)
𝑆 = {𝑠1 , . . . , 𝑠𝑝 } ,
π‘“πœ”π‘— (𝑠𝑖 ) = 𝑖 ⋅ 𝛿𝑗𝑖 ,
𝑗=1
(𝑠, πœ”) ∈ 𝑆 × Ω.
If the skew product transformation Φ is ergodic, then the
function π‘“πœ”∗ (𝑠) is constant almost everywhere on 𝑆 × Ω. It
is worthwhile to notice that, in general, the limit function
π‘“πœ”∗ (𝑠) depends on the two variables 𝑠 and πœ”. To illustrate
this, we consider the measure spaces (𝑆, 𝛽, π‘š) and (Ω, F, πœ‡)
and transformations given by
πœ” ,
πœ™πœ”π‘— = { 𝑗+1
πœ”1 ,
(74)
(76)
which follows from Birkhoff ’s ergodic theorem applied to the
so-called skew product Φ of πœ™ and {πœπœ” } defined by
[π‘ˆ[πœ”∗ ]π‘Ÿ π‘ˆ[πœ‘π‘Ÿ0 πœ”∗ ]π‘Ÿ ⋅ ⋅ ⋅ π‘ˆ[πœ‘π‘˜π‘Ÿ0 πœ”∗ ]π‘Ÿ π‘“πœ‘(π‘˜+1)π‘Ÿ0 πœ”∗ ] (𝑠)
= [𝑇[πœ”∗ ]π‘Ÿ 𝑇[πœ‘π‘Ÿ0 πœ”∗ ]π‘Ÿ ⋅ ⋅ ⋅ 𝑇[πœ‘π‘˜π‘Ÿ0 πœ”∗ ]π‘Ÿ π‘“πœ‘(π‘˜+1)π‘Ÿ0 πœ”∗ ] (𝑠)
π‘š-a.e.,
if 𝑗 = 𝑖 (𝑖, 𝑗 = 1, 2, . . . , 𝑝) ,
if 𝑖 = 1, 2, . . . , 𝑝; and 𝑗 = 𝑝 + 1, . . . , 2π‘ž
(79)
where 𝛿𝑗𝑖 denotes the Kronecker delta, then the limit function
π‘“πœ”∗ (𝑠) determined by (76) depends essentially on the two
variables 𝑠 and πœ” (therefore, the limit function π‘“πœ”∗ (𝑠) is not
necessarily independent of the random variable πœ”.) In fact,
one can easily find that under the above setting
1 2𝑛
1
∑ π‘“πœ”1 (πœπœ™2π‘˜−1 πœ”1 ⋅ ⋅ ⋅ πœπœ”1 𝑠𝑝 ) = π‘“πœ”∗1 (𝑠𝑝 ) = ,
𝑛 → ∞ 2𝑛
2
π‘˜=1
lim
𝑝
1 2𝑛
lim
∑ π‘“πœ”π‘ (πœπœ™2π‘˜−1 πœ”1 ⋅ ⋅ ⋅ πœπœ”1 𝑠𝑝 ) = π‘“πœ”∗𝑝 (𝑠𝑝 ) = .
𝑛 → ∞ 2𝑛
2
π‘˜=1
See also Example 3 below.
(80)
Abstract and Applied Analysis
11
The following example given by GΕ‚adysz [5] will be a great
help to understand the subject of this paper.
Example 2 (see Gladysz [5]). In this example, we consider the
measure spaces (𝑆, 𝛽, π‘š) and (Ω, F, πœ‡) taken to be 𝑆 = Ω =
[0, 1), 𝛽 = F = the 𝜎-field of Borel sets, and π‘š = πœ‡ = the
Lebesgue measure. Let π‘Ÿ be a fixed integer with π‘Ÿ ≥ 2, and let
𝛽𝑗 , 𝑗 = 1, 2, . . . , π‘Ÿ, be real constants such that
1
𝛽1 = ⋅ ⋅ ⋅ = π›½π‘Ÿ−1 = ,
π‘Ÿ
1
π›½π‘Ÿ = −1 + .
π‘Ÿ
𝜏[πœ”∗ ]π‘Ÿ 𝑠 = 𝑠 + ∑𝛽𝑖 πœ”π‘–
(82)
Take π‘Ÿ0 = 1. Then, for a function 𝑓 ∈ 𝐿 1 (π‘š) given by 𝑓(𝑠) =
𝑒2πœ‹π‘–π‘  = exp(𝑠), there exists a πœ‡∗ -null set 𝑁∗ such that for each
∗
(⋅) ∈ 𝐿 1 (π‘š) such
πœ”∗ ∈ Ω∗ − 𝑁∗ , there exists a function 𝑓[πœ”
∗]
π‘Ÿ−1
that
π‘š-a.e.,
(83)
where
π‘š-a.e.
(85)
Next, if (for example) π‘Ÿ = 5 and π‘Ÿ0 = 3, we let 𝛽1 , . . . , 𝛽5 be
real constants such that
𝛽1 =ΜΈ 0,
𝛽2 =ΜΈ 0,
𝛽3 = 0,
𝛽4 = −𝛽1 ,
𝛽5 = −𝛽2 .
(86)
Using the skew product transformation Θ defined by
(87)
we have for the function 𝑓(𝑠) = exp(𝑠),
𝑓 (Θπ‘˜ (𝑠, πœ”∗ ))
5
5
5
= exp [𝑠 + ∑ 𝛽𝑗 πœ”π‘— + ∑𝛽𝑗 πœ”3+𝑗 + ⋅ ⋅ ⋅ + ∑ 𝛽𝑗 πœ”3π‘˜+𝑗 ] (88)
𝑗=1
𝑗=1
[ 𝑗=1
]
= exp [𝑠 + 𝛽1 πœ”1 + 𝛽2 πœ”2 ] ,
π‘˜ = 1, 2, . . . .
Hence it follows at once that for almost all πœ”∗ ∈ Ω∗ , there
exists a function 𝑓∗ ∈ 𝐿 1 (π‘š ⊗ πœ‡∗ ) such that
1 𝑛 𝛼−1
∑ 𝐴 𝑛−π‘˜ 𝑓 (𝜏[πœ‘3π‘˜ πœ”∗ ]5 ⋅ ⋅ ⋅ 𝜏[πœ‘3 πœ”∗ ]5 𝜏[πœ”∗ ]π‘Ÿ 𝑠)
𝑛 → ∞ 𝐴𝛼
𝑛 π‘˜=0
lim
𝑓∗ (𝑠, [πœ”∗ ]π‘Ÿ−1 )
= 𝐾 ⋅ exp (𝑠) ⋅ exp [𝛽1 πœ”1 + ⋅ ⋅ ⋅ + (𝛽1 + ⋅ ⋅ ⋅ + π›½π‘Ÿ−1 ) πœ”π‘Ÿ−1 ] ,
𝐻 (πœ”∗ ) = exp [(𝛽2 + ⋅ ⋅ ⋅ + π›½π‘Ÿ ) πœ”1 + ⋅ ⋅ ⋅ + π›½π‘Ÿ πœ”π‘Ÿ−1 ] ,
1 𝑛
∑ 𝐻 (πœ‘π‘˜ πœ”∗ )
𝑛→∞𝑛
π‘˜=1
∗
∗
∗
= ∫ 𝐻 (πœ” ) π‘‘πœ‡ (πœ” )
1 𝑛
∑ 𝑓 (𝜏[πœ‘3π‘˜ πœ”∗ ]5 ⋅ ⋅ ⋅ 𝜏[πœ‘3 πœ”∗ ]5 𝜏[πœ”∗ ]π‘Ÿ 𝑠)
𝑛→∞𝑛 + 1
π‘˜=0
= lim
=
𝐾 = lim
Ω∗
(𝑠)
Θ (𝑠, πœ”∗ ) = (𝜏[πœ”∗ ]5 𝑠, πœ‘3 πœ”∗ ) ,
(mod1) , 𝑠 ∈ 𝑆.
1 𝑛
∗
∑ 𝑓 (𝜏[πœ‘π‘˜−1 πœ”∗ ]π‘Ÿ ⋅ ⋅ ⋅ 𝜏[πœ”∗ ]π‘Ÿ 𝑠) = 𝑓[πœ”
(𝑠)
∗]
π‘Ÿ−1
𝑛→∞𝑛
π‘˜=1
=
𝑛=1
∗
𝑓[πœ”
∗]
π‘Ÿ−1
𝛽1 + 𝛽2 =ΜΈ 0,
𝑖=1
lim
πœ† → 1+0
(81)
Define a 𝛽 ⊗ F∗π‘Ÿ -measurable family {𝜏[πœ”∗ ]π‘Ÿ : πœ”∗ ∈ Ω∗ } of
measure-preserving transformations on 𝑆 by
π‘Ÿ
∞
= lim (πœ† − 1) ∑ πœ†−(𝑛+1) 𝑓 (𝜏[πœ‘π‘›−1 πœ”∗ ]π‘Ÿ ⋅ ⋅ ⋅ 𝜏[πœ‘πœ”∗ ]π‘Ÿ 𝜏[πœ”∗ ]π‘Ÿ 𝑠)
πœ† → 1+0
𝑛=0
∗
= exp [𝑠 + 𝛽1 πœ”1 + 𝛽2 πœ”2 ] = 𝑓[πœ”
(𝑠)
∗]
5−3
(πœ‘ is ergodic)
1
= exp [ (πœ”1 + 2πœ”2 + ⋅ ⋅ ⋅ + (π‘Ÿ − 1) πœ”π‘Ÿ−1 )]
π‘Ÿ
∞
lim (πœ† − 1) ∑ πœ†−(𝑛+1) 𝑓 (𝜏[πœ‘3π‘˜ πœ”∗ ]5 ⋅ ⋅ ⋅ 𝜏[πœ‘3 πœ”∗ ]5 𝜏[πœ”∗ ]π‘Ÿ 𝑠)
∗
= 𝑓[πœ”
∗ ] (𝑠)
2
π‘š-a.e.
( =ΜΈ 0) .
(84)
Supplement. Let 0 < 𝛼 < 1, 𝛼𝑝 > 1, and π‘Ÿ0 = 1. Then, for the
function 𝑓(𝑠) = exp(𝑠) (∈ 𝐿 𝑝 (π‘š)), there exists a πœ‡∗ -null set
𝑁∗ such that for any πœ”∗ ∈ Ω∗ − 𝑁∗ , there exists a function
𝑓∗ ∈ 𝐿 𝑝 (π‘š⊗πœ‡∗ ) such that (if necessary, apply Hille’s theorem
[22])
1 𝑛 𝛼−1
∑ 𝐴 𝑛−π‘˜ 𝑓 (𝜏[πœ‘π‘˜ πœ”∗ ]π‘Ÿ ⋅ ⋅ ⋅ 𝜏[πœ‘πœ”∗ ]π‘Ÿ 𝜏[πœ”∗ ]π‘Ÿ 𝑠)
𝑛 → ∞ 𝐴𝛼
𝑛 π‘˜=0
lim
1 𝑛 𝛼−1
= lim
∑ 𝐴 𝑛−π‘˜ 𝑓 (𝜏[πœ‘π‘˜ πœ”∗ ]π‘Ÿ ⋅ ⋅ ⋅ 𝜏[πœ‘πœ”∗ ]π‘Ÿ 𝜏[πœ”∗ ]π‘Ÿ 𝑠)
𝑛→∞𝑛 + 1
π‘˜=0
(89)
Example 3. In the setting of Example 2, let π‘ž be an 𝛽measurable function with |π‘ž(𝑠)| = 1. Then, for 𝑓 ∈ 𝐿 𝑝 (π‘š ⊗
πœ‡∗ ), 1 ≤ 𝑝 < ∞, there exists a πœ‡∗ -null set 𝑁∗ such that for
any πœ”∗ ∈ Ω∗ − 𝑁∗ , there exists a function π»πœ”∗ ∈ 𝐿 𝑝 (π‘š) such
that
1 𝑛
∑ π‘ž (𝜏[πœ”∗ ]π‘Ÿ 𝑠) π‘ž (𝜏[πœ‘πœ”∗ ]π‘Ÿ 𝜏[πœ”∗ ] 𝑠) ⋅ ⋅ ⋅
𝑛→∞𝑛
π‘˜=1
lim
π‘ž (𝜏[πœ‘π‘˜−1 πœ”∗ ]π‘Ÿ ⋅ ⋅ ⋅ 𝜏[πœ”∗ ]π‘Ÿ 𝑠)
× π‘“ (𝜏[πœ‘π‘˜ πœ”∗ ]π‘Ÿ ⋅ ⋅ ⋅ 𝜏[πœ”∗ ]π‘Ÿ 𝑠) = π»πœ”∗ (𝑠) ,
π»πœ”∗ (𝑠) = π‘ž (𝜏[πœ”∗ ]π‘Ÿ 𝑠) π»πœ‘πœ”∗ (𝜏[πœ”∗ ]π‘Ÿ 𝑠) ,
(90)
12
Abstract and Applied Analysis
for π‘š-almost all 𝑠 ∈ 𝑆. In fact, letting 𝑄(𝑠, πœ”∗ ) = π‘ž(𝑠) ⋅ 𝑒(πœ”∗ )
with 𝑒 = 1 ∈ 𝐿 𝑝 (πœ‡∗ ) and
Φ∗ (𝑠, πœ”∗ ) = (𝜏[πœ”∗ ]π‘Ÿ 𝑠, πœ‘πœ”∗ ) ,
(91)
we have by Gladysz’s theorem ([5], Satz 5) that there exists a
function 𝐹 ∈ 𝐿 𝑝 (π‘š ⊗ πœ‡∗ ) such that
π‘Ÿ
1 𝑛
∑ 𝑄 (Φ∗ (𝑠, πœ”∗ )) 𝑄 (Φ∗2 (𝑠, πœ”∗ )) ⋅ ⋅ ⋅
𝑛→∞𝑛
π‘˜=1
πœ‹[πœ”∗ ]π‘Ÿ 𝑠 = 𝑠 + ∑ 𝛿𝑗 πœ”π‘—
lim
𝑗=1
𝑄 (Φ∗π‘˜ (𝑠, πœ”∗ )) 𝑓 (Φ∗π‘˜+1 (𝑠, πœ”∗ ))
= [𝑄 (𝑠, πœ”∗ )]
−1
⋅ 𝐹 (𝑠, πœ”∗ ) (= π‘ž(𝑠)−1 𝐹 (𝑠, πœ”∗ )) ,
are chosen at random, not necessarily with the same distribution but independently). In general the random system
{(𝑇[πœ”∗ ]π‘Ÿ , πœ‘) : πœ”∗ ∈ Ω∗ } of linear contractions on 𝐿 𝑝 (π‘š) as
given in Theorem 4 plays a role of such an advance stock of
linear contractions on 𝐿 1 (π‘š). To illustrate this, we let 𝑆 =
𝑋 = [0, 1) and consider the 𝛽 ⊗ F∗π‘Ÿ -measurable, π‘š-measurepreserving transformations πœ‹[πœ”∗ ]π‘Ÿ , πœ”∗ ∈ Ω∗ , defined by
(92)
𝐹 (𝑠, πœ”∗ ) = 𝑄 (𝑠, πœ”∗ ) 𝐹 (Φ∗ (𝑠, πœ”∗ ))
In this case, the random system {(πœ‹[πœ”∗ ]π‘Ÿ , πœ‘) : πœ”∗ ∈ Ω∗ } is
taken as a stock of measure-preserving transformations on
𝑆. If πœ”1 , . . . , πœ”π‘Ÿ are linearly independent irrational numbers,
then {πœ‹[πœ”∗ ]π‘Ÿ } is ergodic. Thus, for any 𝑓 ∈ 𝐿 1 (π‘š) with period
1,
lim
almost everywhere on 𝑆 × Ω∗ . Hence, we may take
𝐻 (𝑠, πœ”∗ ) = π‘ž(𝑠)−1 𝐹 (𝑠, πœ”∗ ) ,
(93)
𝐻 (𝑠, πœ”∗ ) = π‘ž (𝜏[πœ”∗ ]π‘Ÿ 𝑠) 𝐻 (Φ∗ (𝑠, πœ”∗ ))
πœ† → 1+0
π‘š ⊗ πœ‡∗ -a.e. (94)
Moreover, since π‘ž is also 𝛽 ⊗ F∗π‘Ÿ -measurable and 𝐻 ∈ 𝐿 1 (π‘š ⊗
πœ‡∗ ) if we consider the function
π‘Ÿ
π‘ž (𝑠, πœ”1 , . . . , πœ”π‘Ÿ ) = π‘ž (𝑠 + ∑ 𝛽𝑖 πœ”π‘– )
(95)
𝑖=1
Then, by Gladysz’s lemma ([5], Hilfssatz 3) there exists a 𝛽 ⊗
F∗π‘Ÿ−1 -measurable function 𝑔 = 𝑔(𝑠, πœ”1 , . . . , πœ”π‘Ÿ−1 ) such that
π‘š ⊗ πœ‡∗ -a.e.
∗
(96)
∗
Consequently, we have for any πœ” ∈ Ω − 𝑁 ,
π»πœ”∗ (𝑠) = 𝑔[πœ”∗ ]π‘Ÿ−1 (𝑠)
π‘š-a.e.
(97)
Remark 20. It is an interesting problem to ask what happens
if we transform a function 𝑓 ∈ 𝐿 𝑝 with a random sequence
𝑇1 , 𝑇2 , . . . , 𝑇𝑛 , . . ., of operators chosen at random from some
stock of linear operators on 𝐿 𝑝 given in advance. What
can we say about the limit lim𝑛 → ∞ (1/𝑛) ∑π‘›π‘˜=1 𝑇1 𝑇2 ⋅ ⋅ ⋅ π‘‡π‘˜ 𝑓?
Unfortunately, we cannot expect any convergence for every
random sequence chosen from the stock. Therefore, it is
desirable to consider how to choose almost every (not every)
random sequence from the stock (cf. Revesz [30] and Yoshimoto [31]). In Pitt [1] and Ulam and von Neumann [2], the
random ergodic theorem for two-measure-preserving transformations π‘ˆ, 𝑉 (cited in Section 1) means the existence of
the a.e. limit of the form lim𝑛 → ∞ (1/𝑛) ∑π‘›π‘˜=1 𝑓(𝑇1 𝑇2 ⋅ ⋅ ⋅ π‘‡π‘˜ 𝑠))
for almost every sequence (𝑇𝑛 )𝑛≥1 of the infinite sequences
obtained by applying π‘ˆ and 𝑉 in turn at random. This is just
the case that the transformations 𝑇𝑛 , 𝑛 ≥ 1, are chosen at
random with the same distribution and independently. See
also Remark 3 (the case that the transformations 𝑇𝑛 , 𝑛 ≥ 1,
∞
π‘Ÿ
𝑛=1
𝑗=1
= lim (πœ† − 1) ∑ πœ†−(𝑛+1) 𝑓 (𝑠 + 𝑛 ∑ 𝛿𝑗 πœ”π‘— )
and then
∗
(98)
π‘Ÿ
1 𝑛
∑ 𝑓 (𝑠 + π‘˜ ∑ 𝛿𝑗 πœ”π‘— )
𝑛→∞𝑛
𝑗=1
π‘˜=1
= π‘ž (𝑠) 𝐹 (Φ∗ (𝑠, πœ”∗ )) ,
𝐻 (𝑠, πœ”∗ ) = 𝑔 (𝑠, πœ”1 , . . . , πœ”π‘Ÿ−1 )
(mod1) , (𝑠 ∈ 𝑆, 𝛿𝑗 =ΜΈ 0) .
1
= ∫ 𝑓 (𝑠) π‘‘π‘š
0
(99)
π‘š-a.e.
This is an immediate consequence of the ergodicity of the
family {πœ‹[πœ”∗ ]π‘Ÿ }. We can state this fact in terms of stochastic
processes. For example, see GΕ‚adysz [5], Satz 3.
Acknowledgment
The author is certainly indebted and very grateful to Mrs.
Caroline Nashat for her kind assistance.
References
[1] H. R. Pitt, “Some generalizations of the ergodic theorem,” Proceedings of the Cambridge Philosophical Society, vol. 38, pp. 325–
343, 1942.
[2] S. M. Ulam and J. von Neumann, “Random ergodic theorem
(abstract),” Bulletin of the American Mathematical Society, vol.
51, p. 660, 1945.
[3] S. Kakutani, “Random ergodic theorems and Markoff processes
with a stable distribution,” in Proceedings of the 2nd Berkeley
Symposium on Mathematical Statistics and Probability, pp. 247–
261, University of California Press, Berkeley, Calif, USA, 1950.
[4] C. Ryll-Nardzewski, “On the ergodic theorems. III. The random
ergodic theorem,” Studia Mathematica, vol. 14, pp. 298–301,
1954.
[5] S. GΕ‚adysz, “Über den stochastischen Ergodensatz,” Studia
Mathematica, vol. 15, pp. 158–173, 1956.
[6] R. Cairoli, “Sur le théoreΜ€me ergodique aléatoire,” Bulletin des
Sciences Mathématiques, vol. 88, pp. 31–37, 1964.
[7] T. Yoshimoto, “On the random ergodic theorem,” Studia Mathematica, vol. 61, no. 3, pp. 231–237, 1977.
[8] P. Revesz, “Some remarks on the random ergodic theorem
I,” Publications of the Mathematical Institute of the Hungarian
Academy of Sciences, vol. 5, pp. 375–381, 1960.
Abstract and Applied Analysis
[9] P. Revesz, “Some remarks on the random ergodic theorem
II,” Publications of the Mathematical Institute of the Hungarian
Academy of Sciences, vol. 6, pp. 205–213, 1961.
[10] K. Jacobs, Lecture Notes on Ergodic Theory, Aarhus University,
Mathematisches Institut, 1962.
[11] T. Yoshimoto, “On multiparameter Chacon’s type ergodic theorems,” Acta Scientiarum Mathematicarum, vol. 78, pp. 489–515,
2012.
[12] R. V. Chacon and D. S. Ornstein, “A general ergodic theorem,”
Illinois Journal of Mathematics, vol. 4, pp. 153–160, 1960.
[13] E. Hille and R. S. Phillips, Functional Analysis and SemiGroups, vol. 31, American Mathematical Society Colloquium,
Providence, RI, USA, 1957.
[14] T. Yoshimoto, Induced contraction semigroups and random
ergodic theorems [Dissertationes Mathematicae], 1976.
[15] E. Hopf, “The general temporally discrete Markoff process,” vol.
3, pp. 13–45, 1954.
[16] S. D. Chatterji, “Martingale convergence and the RadonNikodym theorem in Banach spaces,” Mathematica Scandinavica, vol. 22, pp. 21–41, 1968.
[17] S. D. Chatterji, “Les martingales et leurs applications analytiques,” in École d’Été de Probabilités: Processus Stochastiques,
vol. 307 of Lecture Notes in Mathematics, pp. 27–164, Springer,
Berlin, Germany, 1973.
[18] N. Dunford and J. T. Schwartz, Linear Operators. I, Interscience,
New York, NY, USA, 1957.
[19] R. V. Chacon and U. Krengel, “Linear modulus of linear
operator,” Proceedings of the American Mathematical Society,
vol. 15, pp. 553–559, 1964.
[20] U. Krengel, Ergodic Theorems: With a Supplement by Antoine
Brunel, vol. 6 of de Gruyter Studies in Mathematics, Walter de
Gruyter, Berlin, Germany, 1985.
[21] J. Woś, “Random ergodic theorems for sub-Markovian operators,” Studia Mathematica, vol. 74, no. 2, pp. 191–212, 1982.
[22] E. Hille, “Remarks on ergodic theorems,” Transactions of the
American Mathematical Society, vol. 57, pp. 246–269, 1945.
[23] K. Yosida and S. Kakutani, “Operator-theoretical treatment
of Markoff ’s process and mean ergodic theorem,” Annals of
Mathematics, vol. 42, pp. 188–228, 1941.
[24] Y. Déniel, “On the a.s. CesaΜ€ro-𝛼 convergence for stationary or
orthogonal random variables,” Journal of Theoretical Probability,
vol. 2, no. 4, pp. 475–485, 1989.
[25] R. Irmisch, Punktweise ergodensätze für (C, 𝛼)—Verfahren, 0<
𝛼 <1 [Dissertation, Fachbereich Mathematik], TH Darmstadt,
1980.
[26] H. Anzai, “Mixing up property of Brownian motion,” Osaka
Mathematical Journal, vol. 2, no. 1, pp. 51–58, 1950.
[27] T. Yoshimoto, “Nonlinear random ergodic theorems for affine
operators,” Journal of Functional Analysis, vol. 256, no. 6, pp.
1708–1730, 2009.
[28] U. Krengel and M. Lin, “Order preserving nonexpansive operators in 𝐿 1 ,” Israel Journal of Mathematics, vol. 58, no. 2, pp. 170–
192, 1987.
[29] R. Wittmann, “Hopf ’s ergodic theorem for nonlinear operators,” Mathematische Annalen, vol. 289, no. 2, pp. 239–253, 1991.
[30] P. Revesz, “A random ergodic theorem and its application to the
theory of Markov chains,” in Ergodic Theory, F. B. Wright, Ed.,
pp. 217–250, Academic Press, 1963.
[31] T. Yoshimoto, “Random ergodic theorem with weighted averages,” Zeitschrift für Wahrscheinlichkeitstheorie und Verwante
Gebiete, vol. 30, no. 2, pp. 149–165, 1974.
13
Advances in
Operations Research
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Advances in
Decision Sciences
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Mathematical Problems
in Engineering
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Journal of
Algebra
Hindawi Publishing Corporation
http://www.hindawi.com
Probability and Statistics
Volume 2014
The Scientific
World Journal
Hindawi Publishing Corporation
http://www.hindawi.com
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
International Journal of
Differential Equations
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Volume 2014
Submit your manuscripts at
http://www.hindawi.com
International Journal of
Advances in
Combinatorics
Hindawi Publishing Corporation
http://www.hindawi.com
Mathematical Physics
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Journal of
Complex Analysis
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
International
Journal of
Mathematics and
Mathematical
Sciences
Journal of
Hindawi Publishing Corporation
http://www.hindawi.com
Stochastic Analysis
Abstract and
Applied Analysis
Hindawi Publishing Corporation
http://www.hindawi.com
Hindawi Publishing Corporation
http://www.hindawi.com
International Journal of
Mathematics
Volume 2014
Volume 2014
Discrete Dynamics in
Nature and Society
Volume 2014
Volume 2014
Journal of
Journal of
Discrete Mathematics
Journal of
Volume 2014
Hindawi Publishing Corporation
http://www.hindawi.com
Applied Mathematics
Journal of
Function Spaces
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Optimization
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Download