A constructive view on ergodic theorems

Bas Spitters

Radboud University

Nijmegen, the Netherlands

. The author was partially supported by the Netherlands Organization for Scientific Research (NWO). The current paper is an updated version of [12] including the corrections in [13].

Abstract. Let T be a positive L1-L contraction. We prove that the following statements are equivalent in constructive mathematics.

  1. The projection in L2 on the space N: = cl{x - T x:xL2} exists;
  2. The sequence (Tn)nN Cesàro-converges in the L2 norm;
  3. The sequence (Tn)nN Cesàro-converges almost everywhere.

Thus, we find necessary and sufficient conditions for the Mean Ergodic Theorem and the Dunford-Schwartz Pointwise Ergodic Theorem.

1.Introduction

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.

2.The mean Ergodic Theorem

The following definitions will be used throughout this chapter.

Let T be an operator on a Banach space X . Define the sum S n : = ∑

n - 1
k = 0
T k and the average A n : =
1
n
S n . Define the subspaces M : = { fX : T f = f } and N : = cl{ x - T x : xX }. An operator T is called a contraction if ‖ T x ‖≤‖ x ‖ whenever xX .

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 = YZ.

Theorem 1. Let T be a contraction on a Banach space X. The sequence (An)nN converges if and only if X = MN, in which case limn→∞An = PM, where PM denotes the projection on M parallel to N.

Proof. First suppose that X = MN . Let h = h M + h N , where h MM and h NN . We claim that A n h converges to h M . First consider f = g - T g for some gX ; then the sequence A n f =

1
n
( g - T n g ) converges to 0. When fN , then there exists gX such that ‖ f - ( g - T g )‖<ε, so for all nN , ‖ A n ( f - ( g - T g ))‖<ε. Hence for large n , ‖ A n f ‖<2ε. Consequently, the sequence A n f converges to 0 whenever fN and, using the notation above, A n h converges to h M .

Let fX; then there exist fM in M and fN in N such that f = fM + fN. Consequently, Anf = fM + AnfN which converges to fM when n tends to ∞.

Now suppose that the sequence (An)nN converges to an operator P. The equalities T P = P = P T follow easily from the definition of the sequence (An)nN. Consequently, P = 0 on N and

P2 = limn→∞AnPabove ( = , [T P = P])limn→∞P = P.

If zMN and ε>0, then there exists uX such that ‖z - (u - T u)‖<ε. Hence for all nN, ‖An(z - (u - T u))‖<ε. Because An(u - T u) converges to 0 and for all nN, Anz = z we see that ‖z‖≤ε. Consequently, MN = {0}.

To see that (I - P)xN observe that

( I - T )(
n - 1
n
I +
n - 2
n
T + ⋯ +
1
n
T n - 2 ) = I - A n

for all n in N. Define

y n : = (
n - 1
n
I +
n - 2
n
T + ⋯ +
1
n
T n - 2 ) x ;

then (I - T)yn→(I - P)x. Consequently, X = MN.

Let H be a Hilbert space and let T be an operator on H . Let xH . There exists a vector x such that langle T y , x rangle = langle x , x rangle if and only if the functional y ↦⟨ T y , x ⟩ is normable. This follows from the Riesz representation theorem [ 3 ] (p.419). If such a vector exists we will denote it by T x even if the adjoint is not totally defined

. Classically, the adjoint of an operator is always totally defined. Constructively this is not the case.

1 .

Theorem 2. [Mean Ergodic Theorem]Let T be a contraction on a Hilbert space H. Then the sequence (An)nN converges if and only if N is located; in this case the sequence (An)nN converges to the orthogonal projection PM on M.

Proof. We first prove that M and N are orthogonal. Suppose that xM, i.e. T x = x. We claim that the map y↦⟨T y,x⟩ is normable. Since |langleT y,xrangle|≤‖x‖‖y‖ whenever yH, ‖x‖ is an upper bound on the norm. On the other hand this upper bound is attained at x. It follows that x∈DomT and ‖Tx‖ = ‖x‖. Now,

Tx - x2 = Tx - x,Tx - x
= Tx2 + ‖x2 - ⟨x,Tx⟩ - ⟨Tx,x
= Tx2 + ‖x2 - ⟨T x,x⟩ - ⟨x,T x
above ( = , [T x = x]) Tx2 + ‖x2 - 2⟨x,x⟩ = ‖Tx2 - ‖x2 = 0.

Consequently, Tx = x and so

x,(I - T)y⟩ = ⟨(I - T)x,y⟩ = 0

for all y in H. We see that M and N are orthogonal.

Suppose that the sequence (An)nN converges. Theorem 1 shows that H = MN. Because M and N are also orthogonal, M and N are located.

Conversely, suppose that N is located. We know that MN. We will prove that NM. Let xN. Then ⟨(I - T)y,x⟩ = 0 whenever yH. It follows that ⟨y,x⟩ = ⟨T y,x⟩, i.e. x = Tx. By a similar argument as above we see that T x = x. We conclude that M = N, so by Theorem 1 the sequence (An)nN converges.

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 xH the sequence (Anx)nN converges if and only if the sequence (‖Anx‖)nN converges.

3.Maximal Ergodic Theorems

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 1L that is, ‖ f1 ≤‖ T f1 and for all real numbers m , | T f |≤ m whenever fL 1 and | f |≤ m . An operator T on an ordered vector space is positive

. When p = 2, this order-theoretic definition differs from the definition of a positive operator on a Hilbert space.

2 if T f ≥0 whenever f ≥0. When τ is a measure-preserving transformation, then T τ is a positive L 1 - L contraction.

Let T be a positive L1-L contraction. Define the operator Mn by

Mnf: = supknAkf

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 supkNAk, but constructively Mf may not be a measurable function for all f in L1, i.e. we may not be able to find simple functions approximating Mf.

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. [Hopf's Maximal Ergodic Theorem] Let T be a positive contraction on L1(μ). Let n be a natural number. If α≥0 is admissible for Mnf, then

[Mnf≥α]f≥0.

Corollary 4. [Wiener [7] (p.51)]Let T be a positive L1-L contraction. Let fL1, nN and α>0 be admissible for Mnf. Then for all n,

μ[ M n f ≥α]≤
1
α
[Mnf≥α] f .

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, L1L is a dense subset of Lp, it is enough to prove that ‖T fp≤‖fp for all fL1L. To achieve this goal we will prove that

T fT(fp)
1
p
(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)pT(fp), so ‖(T f)p1≤‖T(fp)‖1≤‖fp1 and thus ‖T fp≤‖fp. Observe that this inequality trivially holds for p = 1. Consequently, ‖T fp>‖fp is impossible, i.e. ‖T fp≤‖fp 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 L1Lp 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

+
p
, we do not loose generality. We note that f = f χ Y . We define q : = (1 -
1
p
) - 1 as usual. For all real numbers a , b ≥0:

a b
ap
p
+
bq
q
.

It follows that for all real numbers c,d>0:

f
c
χY
d
fp
cpp
+
χ
q
Y
dqq
.

Consequently,

T(f⋅χY)≤T(fp)
c d
cpp
+ T
q
Y
)
d c
dqq
a.e.
(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, mR + such that MfmχY>0. Then fm1 - pfp. Let F'⊂F be a full set such that for all xF'

f(x)≤M,(T f)(x)≤m1 - p(T fp)(x), (T f)(x)≤M(TχY)(x) and (TχY)
1
q
(x)≤1.

Fix xF '.

If T ( f p )( x ) = 0, then T ( f )( x )≤ m 1 - p T ( f p )( x ) = 0 = T ( f p )

1
p
( x ).

If TY )( x ) = 0, then T ( f )( x )≤ M TY )( x ) = 0≤ T ( f p )
1
p
( x ).

If T ( f p )( x )>0 and TY )( x )>0, then we define c : = T ( f p )
1
p
( x ) and d : = TY )
1
q
( x ). The right hand side of ( 2 ) equals c d (
1
p
+
1
q
) = c d . Because f = f χ Y we obtain:

T(f)(x)≤T(fp)
1
p
TY)
1
q
(x)≤T(fp)
1
p
(x).
(3)

We conclude that in any case T(f)(x)>T(fp)
1
p
(x) is impossible. It follows that (3) holds for all xF'. Consequently, (1) holds and we have thus completed the proof.

From this point onwards we will assume that a positive L1-L contraction is extended to Lp for all p≥1.

Theorem 6. [Dominated Ergodic Theorem] Let T be a positive L1-L contraction. Then for all nN, p>1 and fLp: ‖Mnfp

p
p - 1
fp.

Proof. Fix nN, p>1, and fL

+
p
.

∫(Mnf)pdμ = ∫∫
Mnf(s)
0
pαp - 1dαdμ(s)
= p∫∫
0
αp - 1χ[Mnf≥α](s)dαdμ(s)
above ( = , Fubini) p
0
αp - 1μ[Mnf≥α]dα
above (≤, Wiener) p
0
αp - 2[Mnf≥α]f dμdα
= p
0
∫αp - 2χ[Mnf≥α](s)f(s)dμ(s)dα
above ( = , Fubini) pf(s)∫
0
αp - 2χ[Mnf≥α](s)dαdμ(s)
= pf(s)∫
Mnf(s)
0
αp - 2dαdμ(s)
=
p
p - 1
f(Mnf)p - 1dμ
p
p - 1
fpMnf
p - 1
p
.

This last inequality follows from Hölder's inequality and the fact that M n f ≤∑
n
k = 0
A k f , so that M n fL p . For all fL p , M n fM n | f | and thus ‖ M n fp ≤‖ M n | f |‖ p
p
p - 1
fp .

A function f:RR is convex if

fx + (1 - λ)y)≤λf(x) + (1 - λ)f(y),

whenever λ∈[0,1] and x,yR.

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

h x y ( z )≔
min{z,y} - min{z,x}
y - x

whenever x<y. These functions are 0 when zx and 1 when xy. We define

Λ( h x y )≔
f(y) - f(x)
y - x
.

Then Λ is increasing

. This is most easily seen by a geometric argument considering the graph of a convex function.

3 on the set E with the order inherited from the the functions from R to R that is, ( E ,Λ) is a profile. The profile theorem ensures that all but countably many points are smooth. In the present case this means that the function is differentiable at these points.

Lemma 8. [Jensen's inequality] Let (X,μ) be a finite measure space and ϕ:RR be a convex function. If f,ϕ∘f are in L1, then

ϕ(
1
μ(X)
f )μ( X )≤∫ϕ∘ f .

Proof. Let x be in the domain of ϕ'. Then for all real numbers y,

ϕ(y)≥ϕ'(x)(y - x) + ϕ(x),

and hence

ϕ(f(t))≥ϕ'(x)(f(t) - x) + ϕ(x)

for all t∈Domf. By integrating we obtain

∫ϕ∘f≥ϕ'(x)((∫f) - xμ(X)) + ϕ(x)μ(X).

If we take a sequence ( x n ) nN in Domϕ' tending to
1
μ(X)
f , we obtain the inequality above.

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 fLq, then |f|p is measurable and bounded by |f|q∨1, hence |f|p is integrable and by taking ϕ(x): = x

p
q
in Jensen's inequality we obtain

fp≤‖fqμ(X)
1
p
-
1
q
.
(4)

We see that LqLp.

Let μ be σ-finite. If pq≥1 and fM, then

f
p
p
= ∫|f|p = ∫|f|q|f|p - qMp - qf
q
q
.

Consequently,

fpM1 -
q
p
f
q
p
q
.
(5)

Theorem 9. [Lp Ergodic Theorem] Let p,q≥1 and (X,μ) be a finite measure space or let p>1,q≥1 and (X,μ) a σ-finite measure space. Let T be a positive L1-L contraction. If the sequence (An)nN converges in Lq, then it converges in Lp.

Proof. Let fL p and choose a simple g such that ‖ f - gp <

ε
4
. Let M be a bound for g . For all n , mN :

Anf - Amfp Anf - Angp + ‖Ang - Amgp + ‖Amg - Amfp
ε
4
+ ‖Ang - Amgp +
ε
4
.

If we show that ‖Ang - Amgp→0, when m,n→∞, then (Anf)nN is a Cauchy sequence in Lp. That ‖Ang - Amgp→0 follows from the fact that the sequence (Ang)nN converges in Lq and the inequalities (4) and (5) for the case μ is finite or μ is σ-finite and pq. The case μ is σ-finite and 1<pq is more difficult. We will now proceed to consider this case. If p≥1, hLp, n,m,l,rN and m = l n + r, then

Amhp - ‖Anhp
1
l
‖(I + Tn + ⋯ + T(l - 1)n)Anhp - ‖Anhp
+
1
m
Tl n(I + ⋯ + Tr - 1)hp
0 +
r
m
hp
n
m
hp.

It follows that the sequence (‖Anhp)nN is essentially decreasing, that is for each nN and each ε>0, there exists NN such that ‖Amhp≤‖Anhp + ε for all mN.

Let gLqLp. 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 gp ) nN is essentially decreasing, it is enough to find for each η>0 and kN , an n > k such that ‖ A n g

p
p
<η. We will now proceed to find such n . Let β =
η
4
. Take an integrable set B 1 such that ‖ g χ X - B1
p
p
<β. We recall that A n g →0 in L q . We can thus compute n 1 such that‖ A n1 g
p
q
<βμ( B 1 )
p
q
- 1
. By ( 4 )

‖(An1gB1
p
p
≤‖(An1gB1
p
q
μ(B1)1 -
p
q
<β.

Now

An1g
p
p
orAn1g
p
p
>η - β.

In the former case we are done, so we may assume that ‖An1g

p
p
>η - β and thus ‖(An1gX - B1
p
p
>η - 2β. We choose B2B1 such that ‖(An1gB2 - B1
p
p
≥η - 3β. Compute n2>n1 such that ‖An2g
p
q
<βμ(B2)
p
q
- 1
. Then ‖An2gχB2 - B1
p
p
<β.

We continue in this way until we find NN such that ‖AnNg

p
p
<η. That such an N exists we see as follows. Choose KN such that

((
p
p - 1
) p - 1)‖ g
p
p
+ β≤ K (η - 3β).

For all NK, ‖AnNg

p
p
>η - β or ‖AnNg
p
p
<η. Suppose that for all NK, ‖AnNg
p
p
>η - β. Define B0 = ∅ and n0 = 0. Remark that for all NN,

MN|g|
N
i = 0
MN|gBi + 1 - Bi
N
i = 0
Ani|gBi + 1 - Bi
N
i = 0
AnigχBi + 1 - Bi.

It follows from the Dominated Ergodic Theorem that for all NK,

(
p
p - 1
)pg
p
p
MN|g|‖
p
p
‖∑
N
i = 0
(AnigBi + 1 - Bi
p
p
=
N
i = 0
‖(AnigBi + 1 - Bi
p
p
> N(η - 3β) + ‖g
p
p
- β.

It follows that there exists NK such that ‖AnNg
p
p
<η. Consequently, Angg~ in Lp. Finally, we observe that even if we do not know whether pq or p>q we can show that the sequence An converges in Lp. To see this we observe that either p<q + ε or p>q, the latter case has been treated before. In the former case, the sequence also converges in Lq&upl;ε as we proved before, and we can thus proceed as above.

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.

4.The Pointwise Ergodic Theorem

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. [Banach's Principle] Let (Y,μ) be a measure space. Let (Tn)nN be a sequence of linear operators from a Banach space X to the space μ-measurable real functions. Define for each nN, an operator M~n by M~nx: = supkn|Tkx| for all xX. Suppose that there exists a positive decreasing function C from R to R such that limα→∞C(α) = 0 and for all x and n:

μ[M~nx≥α‖x‖]<C(α),

whenever α‖x‖ is admissible for M~nx. Then the set of elements xX for which the sequence (Tnx)nN converges a.e. is closed.

Proof. Suppose that a sequence zn converges to z in X in norm and that there exists a sequence (fn)nN of measurable functions such that for all mN, Tmznfn a.e. We will prove that the sequence (Tmz)mN converges a.e.

For natural numbers a and b, ω∈Y and xX put

Δa,b(ω,x): = supan,mb|Tnx(ω) - Tmx(ω)|.

Then

a,b(⋅,z) - Δa,b(⋅,zn)|≤|Δa,b(⋅,z - zn)|≤2M~b(z - zn).

Let ε>0. Choose a natural numbers n such that C (

ε
2
z - z n ‖)<ε and choose a natural number k and an integrable set DY such that μ( D )<ε and for all ak :

{ω:| T a z n (ω) - f n (ω)|>
ε
2
}⊂ D .

For all bak,

μ[|Δa,b(⋅,z)|>2ε] μ[|Δa,b(⋅,zn)|>ε]
+ μ[|Δa,b(⋅,z) - Δa,b(⋅,zn)|>ε]
μ(D) + μ[2M~b(z - zn)>ε]
ε + C(
ε
2
z - zn‖)≤2ε.

Construct ascending sequences (nm)mN and (km)mN of natural numbers such that

C(ε2 - m - 1z - znm‖)<ε2 - m

and

μ[|Tnznm - fnm|>ε2 - m]<ε2 - m

for all nkm. Put

E: = cupm{ω:|Δkm,km + 1(ω,z)|>ε2 - m}.

The set E is integrable and μ(E)≤∑

m = 1
2ε2 - m = 2ε. For ω∈ - E and n,mk1:

|Tnz(ω) - Tmz(ω)|≤ε.

Consequently, the sequence (Tnz)nN converges a.e.

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 Mq = {fLq:T f = f} and Nq = cl{T f - f:fLq}.

Theorem 11. [Pointwise Ergodic Theorem] Let μ be a σ-finite measure. Let T be a positive L1-L contraction. If Lq = MqNq, for some q≥1, then for all p≥1, the sequence (Anf)nN converges a.e. for all fLp.

Proof. Suppose that Lq = MqNq, for some q≥1. First let p>1. Theorem 9 and Theorem 1 show that Lp = MpNp. Suppose that f = g + T h - h, where gLp, T g = g and hLLp. The set of these f is dense in Lp. Because T contracts the L-norm, limn→∞Anf = g a.e. For all admissible α>0,

αμ[Mnf≥α] above (≤, [Wiener]) [Mnf≥α]f
above (≤, [Jensen]) (∫[Mnf≥α]fp)
1
p
μ[Mnf≥α]1 -
1
p
fpμ[Mnf≥α]1 -
1
p
.

Consequently, αμ[Mnf≥α]

1
p
≤‖fp. Define M~nf = supkn|Akf|, for all fLp. Substituting α = β‖fp and observing that M~nf = Mnf when f≥0, we obtain

μ[M~nf≥β‖fp]≤β - p. (6)

For all fLp, M~nfM~n|f|, so inequality (6) holds for all fLp. Banach's principle shows that the sequence (Anf)nN converges a.e. for all f in Lp.

Finally we remove the assumption that p is strictly greater than 1. Let p≥1. The set Lp + 1Lp is dense in Lp , so one can apply Banach's principle since inequality (6) above holds for all p≥1.

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)nN converges a.e. for all f in Lp, then N = cl{x - T x:xL2} is located in L2.

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)nN converges weakly to 0. We may assume that f is a simple function. Since the sequence (Anf)nN 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,gL2 that is, (Anf)nN converges weakly to 0.

Define B n ≔(

n - 1
n
I +
n - 2
n
T + ⋯ +
1
n
T n - 2 ). By an argument similar to the proof of Theorem 1 we see that for each x , ( I - T ) B n x converges weakly to an element in the space wcl{ y - T y : yL 2 } and thus

L2 = M⊕wcl{(I - T)x:xL2},

where wcl denotes the weak closure. Since (I - T)Bnx = (I - An)x is a bounded sequence and the weak closure of a bounded convex inhabited subset coincides with its strong closure, by Lemma 5.2.4 in [11], we see that N = wcl{y - T y:yL2}. It follows that the display sum above is orthogonal and thus that N is located.

Lemma 13. [6]Let E be a uniformly convex and uniformly smooth Banach space and C a bounded convex subset of E. Then C is located if and only if sup{f(z) : zC } exists for each normable linear functional f on E .

Lemma 14. [3][6]For p>1, the space Lp is uniformly convex and uniformly smooth and each normable functional may be represented by a functional f↦∫f g, for some g in Lq, where

1
p
+
1
q
= 1.

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,AB(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.

We will now apply Lemma 14. Since the simple functions are dense in all the Lp-spaces, if sup{(z,f):zC} exists for all f in Lp, then this supremum exists for all f in set of the simple functions, and thus for all f in any Lp', where p'>1. It follows that A is located in Lp'. In particular, if N are located in Lp, then it is located in L2. Since these sets are orthogonal, they are thus both located in L2. Consequently, they are both located in any Lp'.

Theorem 16. Let μ be a σ-finite measure and T a positive L1-L contraction. Denote by An the average

1
n
n - 1
i = 0
Ti. The following statements are equivalent:

  1. The set N = cl{x - T x:xL2} is located in L2;
  2. The sequence (An)nN converges in L2;
  3. For all p>1, the sequence (An)nN converges in Lp;
  4. There is p>1 such that the sequence (An)nN converges in Lp;
  5. For all p≥1 and fLp, the sequence (Anf)nN converges a.e.;
  6. There is p≥1 such that for all fLp, the sequence (Anf)nN converges a.e.

For a finite measure we may replace p>1 by p≥1.

Proof. (1)⇔(2) is Theorem 2. (4)⇒(3) is Theorem 9. (3)⇒(2) is trivial. (2)⇒(4) is trivial. (4)⇒(5) follows from Theorem 1 and Theorem 11. (5)⇒(6) is trivial. (6)⇒(1) follows from Theorem 12.

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.

Bibliography

[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.