|
Thus, we find necessary and sufficient conditions for the Mean Ergodic Theorem and the Dunford-Schwartz Pointwise Ergodic Theorem.
Bishop [4] (p.233) put forward the following problem connected with the question of finding a constructive interpretation of ergodic theorems.
Think of X as the union of two equal tanks of fluid, and T as a motion of the fluid, which is supposed to keep the fluid confined to the tank in which it has been placed. Imagine that there may be a small leak, which would in fact allow the fluid in the two tanks to mix, but that we are not able to decide whether a leak actually exists. Since the leak if it exists, is small, there will be little mixing between the tanks after unit time (that is, under the transformation T), but after a long time (that is, under the transformation Tn for some large n) the mixing may be substantial.
He concluded that Birkhoff's Ergodic Theorem is non-constructive.
Bishop proved, using so-called upcrossings, a version of the Chacon-Ornstein Theorem. This theorem is a generalization of Dunford and Schwartz's version of the Pointwise Ergodic Theorem, a result which extends Birkhoff's Ergodic Theorem. In Bishop's ergodic theorem a limit is proved to exists in a constructively very weak sense. Bishop's result is a so-called equal-hypothesis substitute for the ergodic theorem. Bishop considered finding an equal-conclusion substitute to be ‘an important open problem', see [2] (p55). His student Nuber [8][9] found such an equal-conclusion substitute for Birkhoff's Ergodic Theorem. His proof uses measure theoretic techniques and seems to work only for measure-preserving transformations. We use functional analytic techniques to give necessary and sufficient conditions for von Neumann's Mean Ergodic Theorem and the Dunford and Schwartz version of the Pointwise Ergodic Theorem to hold.
In the context of Bishop's constructive mathematics [3] we prove that for the Mean Ergodic Theorem to hold it is sufficient that the projection on the space of invariant functions exists. Conversely, from the convergence of the sequence in the conclusion of that theorem we obtain the projection. We also show that the Mean Ergodic Theorem is sufficient to prove the Dunford and Schwartz version of the Pointwise Ergodic Theorem, and again a converse is also true. The aim of this paper is to make these claims rigorous, see Theorem 16.
The paper is organized as follows. We first prove a Mean Ergodic Theorem. Then the Maximal Ergodic Theorem and Banach's Principle are proved and used to prove the Pointwise Ergodic Theorem. Our presentation loosely follows that of Krengel [7] (p.65,p.159) and Dunford and Schwartz [5]. We use [3] as a general reference for constructive mathematics.
The following definitions will be used throughout this chapter.
Let
T
be an operator on a Banach space
T
k
and the average
A
n
: =
n - 1
k = 0
1 |
n |
When X is a vector space and Y,Z are subspaces such that for every x in X there exist unique y in Y and z in Z such that x = y + z, we write X = Y⊕Z.
Theorem 1. Let T be a contraction
on a Banach space
Proof.
First suppose that
1 |
n |
Let f∈
Now suppose that the sequence
(An)n∈N
converges to an operator P. The equalities T
P = P = P T follow easily
from the definition of the sequence
(An)n∈N.
Consequently, P = 0 on
If z∈
To see that (I - P)x∈
n - 1 |
n |
n - 2 |
n |
1 |
n |
for all n in N. Define
n - 1 |
n |
n - 2 |
n |
1 |
n |
Let
. Classically, the adjoint of an operator is always totally
defined. Constructively this is not the case.
Theorem 2.
Proof. We first prove that
‖T∗x - x‖2 | = | ⟨T∗x - x,T∗x - x⟩ |
= | ‖T∗x‖2 + ‖x‖2 - ⟨x,T∗x⟩ - ⟨T∗x,x⟩ | |
= | ‖T∗x‖2 + ‖x‖2 - ⟨T x,x⟩ - ⟨x,T x⟩ | |
above ( = , [T x = x]) | ‖T∗x‖2 + ‖x‖2 - 2⟨x,x⟩ = ‖T∗x‖2 - ‖x‖2 = 0. |
Consequently, T∗x = x and so
for all y in H. We see that
Suppose that the sequence
(An)n∈N
converges. Theorem 1 shows that
Let (X,μ) be a measure space. A measure-preserving transformation of X is a partial function from a full set to a full set such that for all integrable sets A, τ(A) is integrable and μ(τ(A)) = μ(A). If τ is a measure-preserving transformation, then Tτf: = f∘τ is a contraction on L2. This shows that our result generalizes the possibly more familiar formulation of the Theorem.
Bishop and Bridges [3] (problem 46, p.395) give the following version of the Mean Ergodic Theorem. Let T be a unitary operator on a Hilbert space H; then for all x∈H the sequence (Anx)n∈N converges if and only if the sequence (‖Anx‖)n∈N converges.
Let (
X
,μ) be a measure space. An operator
T
on
L
1
(μ) is an
L1-L∞
contraction
if
T
is a contraction on
L
1
that contracts the
L
∞
-norm on
L
1
∩
L
∞
. When p = 2, this order-theoretic definition
differs from the definition of a positive operator on a
Hilbert space.
Let T be a positive L1-L∞ contraction. Define the operator Mn by
for all n in N. Garcia's proof [7] (p.8) of the following Theorem 3 is constructive. Note however that we do not make any claims about M∞. This operator is defined classically as supk∈NAk, but constructively M∞f may not be a measurable function for all f in L1, i.e. we may not be able to find simple functions approximating M∞f.
In the constructive theory of measure spaces it is not always possible to compute the measure of the set [f<α]≔{x:f(x)<α}. However, we can compute the measure for all but countably many α, such α are called admissible, see [3] for details.
Theorem 3.
Corollary 4.
1 |
α |
The following theorem is sometimes called the little Riesz theorem. The proof we give here is an adaptation of [7] (Lemma 1.7.4).
Proposition 5. If T is a positive L1-L∞ contraction, then T can be uniquely extended to an Lp contraction for all p≥1.
Proof. Because for all p≥1, L1∩L∞ is a dense subset of Lp, it is enough to prove that ‖T f‖p≤‖f‖p for all f∈L1∩L∞. To achieve this goal we will prove that
T
f≤T(fp)
|
(1) |
for all positive simple functions f and p>1.
Assume for a moment that we have done so and let f be a positive simple function and p>1. Then (T f)p≤T(fp), so ‖(T f)p‖1≤‖T(fp)‖1≤‖fp‖1 and thus ‖T f‖p≤‖f‖p. Observe that this inequality trivially holds for p = 1. Consequently, ‖T f‖p>‖f‖p is impossible, i.e. ‖T f‖p≤‖f‖p holds for all p≥1, even if we are unable to decide p = 1 or p>1. It follows that for all p≥1, T is a positive Lp-contraction on the positive simple functions, and consequently also on the simple functions. The simple functions are dense and T is a contraction, so the operator T restricted to the simple functions can be uniquely extended to Lp. This extension agrees with T on L1∩Lp for all p≥1.
We will now prove (
1
) for a positive simple function
f
and
p
>1. We will assume that
Y
: = [
f
>0] is integrable. Since the simple functions
f
with this property are also dense in
L
, we do not loose generality. We note that
f
=
f
χ
Y
. We define
q
: = (1 -
+
p
1 |
p |
ap |
p |
bq |
q |
It follows that for all real numbers c,d>0:
f |
c |
χY |
d |
fp |
cpp |
χ
|
||
dqq |
Consequently,
T(f⋅χY)≤T(fp)
|
(2) |
Let F be a full set on which (2) holds for rational c and d and thus by continuity for all c,d>0. Compute M, m∈R + such that M≥f≥mχY>0. Then f≤m1 - pfp. Let F'⊂F be a full set such that for all x∈F'
1 |
q |
Fix
x
∈
F
'.
If
T
(
f
p
)(
x
) = 0, then
T
(
f
)(
x
)≤
m
1 - p
T
(
f
p
)(
x
) = 0 =
T
(
f
p
)
(
x
).
1
p
If
T
(χ
Y
)(
x
) = 0, then
T
(
f
)(
x
)≤
M
T
(χ
Y
)(
x
) = 0≤
T
(
f
p
)
(
x
).
1
p
If
T
(
f
p
)(
x
)>0 and
T
(χ
Y
)(
x
)>0, then we define
c
: =
T
(
f
p
)
(
x
) and
d
: =
T
(χ
Y
)
1
p
(
x
). The right hand side of (
2
) equals
c
d
(
1
q
1 |
p |
1 |
q |
T(f)(x)≤T(fp)
|
(3) |
1 |
p |
From this point onwards we will assume that a positive L1-L∞ contraction is extended to Lp for all p≥1.
Theorem 6.
‖f‖p.
p
p - 1
Proof. Fix n∈N, p>1,
and f∈L
.
+
p
∫(Mnf)pdμ | = | ∫∫
|
||||
= | p∫∫
|
|||||
above ( = , Fubini) | p∫
|
|||||
above (≤, Wiener) | p∫
|
|||||
= | p∫
|
|||||
above ( = , Fubini) | p∫f(s)∫
|
|||||
= | p∫f(s)∫
|
|||||
= |
|
|||||
≤ |
|
n |
k = 0 |
p |
p - 1 |
A function f:R→R is convex if
whenever λ∈[0,1] and x,y∈R.
Theorem 7. A (total) convex function from R to R has an non-decreasing derivative which is defined in all but countably many points.
Proof. To prove this we will use Bishop's profile theorem, a constructive substitute for the classical lemma stating that every non-decreasing real function is continuous in all but countably many points. Like Bishop we define the set E to be the set of piecewise linear functions
min{z,y} - min{z,x} |
y - x |
whenever x<y. These functions are 0 when z⩽x and 1 when x⩾y. We define
f(y) - f(x) |
y - x |
. This is most easily seen by a geometric argument
considering the graph of a convex function.
Lemma 8.
1 |
μ(X) |
Proof. Let x be in the domain of ϕ'. Then for all real numbers y,
and hence
for all t∈Domf. By integrating we obtain
1 |
μ(X) |
The Lp Ergodic Theorem can be proved classically for finite measure spaces using the pointwise ergodic theorem [14]. Another proof [7] (Thm. 2.1.2) uses a non-constructive compactness argument. Our proof works not only for finite measure spaces, but also for σ-finite measure spaces. We make some preparations.
Let μ be a finite measure. If 1≤p<q and
f∈Lq, then
|f|p is measurable and bounded by
|f|q∨1, hence
|f|p is integrable and by taking
ϕ(x): = x
in Jensen's inequality we obtain
p
q
‖f‖p≤‖f‖qμ(X)
|
(4) |
We see that Lq⊂Lp.
Let μ be σ-finite. If p≥q≥1 and f≤M, then
p |
p |
q |
q |
Consequently,
‖f‖p≤M1
-
|
(5) |
Theorem 9.
Proof. Let f ∈ L p and choose a simple g such that ‖ f - g ‖ p <
ε |
4 |
‖Anf - Amf‖p | ≤ | ‖Anf - Ang‖p + ‖Ang - Amg‖p + ‖Amg - Amf‖p | ||||
≤ |
|
If we show that ‖Ang - Amg‖p→0, when m,n→∞, then (Anf)n∈N is a Cauchy sequence in Lp. That ‖Ang - Amg‖p→0 follows from the fact that the sequence (Ang)n∈N converges in Lq and the inequalities (4) and (5) for the case μ is finite or μ is σ-finite and p≥q. The case μ is σ-finite and 1<p≤q is more difficult. We will now proceed to consider this case. If p≥1, h∈Lp, n,m,l,r∈N and m = l n + r, then
‖Amh‖p - ‖Anh‖p | ≤ |
|
||||
+
|
||||||
≤ | 0 +
|
It follows that the sequence (‖Anh‖p)n∈N is essentially decreasing, that is for each n∈N and each ε>0, there exists N∈N such that ‖Amh‖p≤‖Anh‖p + ε for all m≥N.
Let g∈Lq∩Lp. Let g~ be the limit of Ang in Lq. The function g~ is in Lp, because T contracts the Lp-norm. We will prove that limn→∞Ang = g~ in Lp. By looking at g - g~ we may assume that g~ = 0.
Because the sequence (‖
A
n
g
‖
p
)
n∈N
is essentially decreasing, it is enough to find for each η>0
and
k
∈
N
, an
n
>
k
such that ‖
A
n
g
‖
<η. We will now proceed to find such
n
. Let β =
p
p
η |
4 |
p |
p |
p |
q |
p |
q |
p |
p |
p |
q |
p |
q |
Now
p |
p |
p |
p |
In the former case we are done, so we may assume that
‖An1g‖
>η - β and thus
‖(An1g)χX
- B1‖
p
p
>η - 2β. We choose
B2⊃B1 such that
‖(An1g)χB2
- B1‖
p
p
≥η - 3β. Compute
n2>n1 such that
‖An2g‖
p
p
<βμ(B2)
p
q
- 1. Then
‖An2gχB2
- B1‖
p
q
<β.
p
p
We continue in this way until we find N∈N
such that
‖AnNg‖
<η. That such an N exists we see as
follows. Choose K∈N such that
p
p
p |
p - 1 |
p |
p |
For all N≤K,
‖AnNg‖
>η - β or
‖AnNg‖
p
p
<η. Suppose that for all
N≤K,
‖AnNg‖
p
p
>η - β. Define B0 =
∅ and n0 = 0. Remark that for all
N∈N,
p
p
MN|g| | ≥ | ∑
|
||
≥ | ∑
|
|||
≥ | ∑
|
It follows from the Dominated Ergodic Theorem that for all N≤K,
(
|
≥ | ‖MN|g|‖
|
||||||
≥ | ‖∑
|
|||||||
= | ∑
|
|||||||
> | N(η - 3β) +
‖g‖
|
p |
p |
If μ is σ-finite, it is not true in general that if for some p>1, the sequence An converges to 0 in Lp, then the sequence An converges in L1. To see this let τ(x): = x + 1 on R with Lebesgue measure, T = Tτ and f = χ[0,1]. For all p>1, Anf converges to 0 in Lp, but the sequence does not converge in L1.
The following principle, called Banach's principle, will be used as follows: we prove that a sequence of operators converges almost everywhere (a.e.) on a dense set and then conclude that the sequence converges a.e. on the whole space.
The proof of the following theorem would be easier if we could prove constructively that M∞ is measurable.
Theorem 10.
whenever α‖x‖ is admissible for
M~nx.
Then the set of elements x∈
Proof. Suppose that a sequence zn converges to z in X in norm and that there exists a sequence (fn)n∈N of measurable functions such that for all m∈N, Tmzn→fn a.e. We will prove that the sequence (Tmz)m∈N converges a.e.
For natural numbers a and b,
ω∈Y and x∈
Then
Let ε>0. Choose a natural numbers n such that C (
ε |
2 |
ε |
2 |
For all b≥a≥k,
μ[|Δa,b(⋅,z)|>2ε] | ≤ | μ[|Δa,b(⋅,zn)|>ε] | ||
≤ | μ(D) + μ[2M~b(z - zn)>ε] | |||
≤ | ε + C(
|
Construct ascending sequences (nm)m∈N and (km)m∈N of natural numbers such that
and
for all n≥km. Put
The set E is integrable and
μ(E)≤∑
2ε2 - m = 2ε. For
ω∈ - E and
n,m≥k1:
∞
m = 1
Bishop [4] (p.230) gave a constructive proof of Lebesgue's Differentiation Theorem. Using Banach's principle one can give another constructive proof, see [10] (p.101).
When q≥1 and T is a positive
L1-L∞ contraction, we
define
Theorem 11.
Proof. Suppose that Lq =
αμ[Mnf≥α] | above (≤, [Wiener]) | ∫[Mnf≥α]f | ||||
above (≤, [Jensen]) | (∫[Mnf≥α]fp)
|
|||||
≤ | ‖f‖pμ[Mnf≥α]1
-
|
Consequently,
αμ[Mnf≥α]
≤‖f‖p.
Define M~nf
=
supk≤n|Akf|,
for all f∈Lp.
Substituting α =
β‖f‖p and
observing that M~nf
= Mnf when
f≥0, we obtain
1
p
μ[M~nf≥β‖f‖p]≤β - p. | (6) |
For all f∈Lp, M~nf≤M~n|f|, so inequality (6) holds for all f∈Lp. Banach's principle shows that the sequence (Anf)n∈N converges a.e. for all f in Lp.
The two-step argument above is used because in general we do not have convergence in the L1-norm. The space L1 has an awkward geometrical structure: it is not uniformly convex.
Theorem 12. Let p≥1. If the sequence
(Anf)n∈N
converges a.e. for all f in
Lp, then
Proof. Let f be an element of
L2. Without loss of generality we may assume
that Anf→0 a.e. We
claim that the sequence
(Anf)n∈N
converges weakly to 0. We may assume that f is a simple
function. Since the sequence
(Anf)n∈N
is bounded, the
sequence∫(Anf)g
converges to 0 whenever g is a simple function. The inner
product is continuous on L2, so
⟨Anf,g⟩→0
for all f,g∈L2
Define B n ≔(
n - 1 |
n |
n - 2 |
n |
1 |
n |
Lemma 13.
Lemma 14.
+
1
p
= 1.
1
q
Lemma 15. If N is located in some Lp, then both M and N are located in all Lp, where p>1.
Proof. Locatedness is a property of closed sets, so we may restrict to a set which is dense in all the Lp-spaces, for instance the bounded L1 functions or the simple functions. Moreover, an inhabited set A is located if we can compute the distance ρ(x,A) for each x. Let a be an element of A. Then ρ(x,A) = ρ(x,A∩B(x,ρ(x,a) + 1))), where B(x,r) denotes the ball around x with radius r. Consequently, when considering located sets we may restrict to its bounded subsets.
Theorem 16. Let μ be a
σ-finite measure and T a positive
L1-L∞ contraction.
Denote by An the average
∑
1
n
Ti. The following
statements are equivalent:
n - 1
i = 0
For a finite measure we may replace p>1 by p≥1.
Some of these results have circulated in preprints for some time. They then appeared in my thesis [11]. The results have been used in [1] in the context of reverse mathematics.
I would like to thank Wim Veldman for his advice during the PhD-project. Finally, I would like to thank the referee for comments that helped to improve the presentation of the paper.
[1] Jeremy Avigad and Ksenija Simic. Fundamental notions of analysis in subsystems of second-order arithmetic. Annals of Pure and Applied Logic, to appear.
[2] Errett Bishop. Mathematics as a numerical language. In Intuitionism and Proof Theory (Proceedings of the summer Conference at Buffalo, N.Y., 1968), pages 53–71. North-Holland, Amsterdam, 1970.
[3] Errett Bishop and Douglas Bridges. Constructive analysis, volume 279 of Grundlehren der Mathematischen Wissenschaften. Springer-Verlag, 1985.
[4] Errett A. Bishop. Foundations of constructive analysis. McGraw-Hill Publishing Company, Ltd., 1967.
[5] N. Dunford and J. T. Schwartz. Linear operators. Part I: General theory. Interscience Publishers, 1958.
[6] Hajime Ishihara and Luminiţa Vîţă. Locating subsets of a normed space. Proc. Amer. Math. Soc., 131(10):3231–3239, 2003.
[7] Ulrich Krengel. Ergodic Theorems. Studies in Mathematics. de Gruyter, 1985.
[8] J. A. Nuber. A constructive ergodic theorem. Transactions of the American Mathematical Society, 164:115–137, 1972.
[9] J. A. Nuber. Erratum to ‘A constructive ergodic theorem'. Transactions of the American Mathematical Society, 216:393, 1976.
[10] Karl Petersen. Ergodic theory. Cambridge Studies in Advanced Mathematics. Cambridge University Press, 1983.
[11] Bas Spitters. Constructive and intuitionistic integration theory and functional analysis. PhD thesis, University of Nijmegen, 2002.
[12] Bas Spitters. A constructive view on ergodic theorems. J. Symbolic Logic, 71(2):611–623, 2006.
[13] Bas Spitters. Corrigendum to: “A constructive view on ergodic theorems” J. Symbolic Logic 71 (2006), no. 2, 611–623. J. Symbolic Logic, 71(4):1431–1432, 2006.
[14] Peter Walters. An introduction to ergodic theory, volume 79 of Graduate Texts in Mathematics. Springer-Verlag, 1982.