• Nie Znaleziono Wyników

Shift spaces and attractors in noninvertible horseshoes by

N/A
N/A
Protected

Academic year: 2021

Share "Shift spaces and attractors in noninvertible horseshoes by"

Copied!
23
0
0

Pełen tekst

(1)

152 (1997)

Shift spaces and attractors in noninvertible horseshoes

by

H. G. B o t h e (Berlin)

Abstract. As is well known, a horseshoe map, i.e. a special injective reimbedding of the unit square I

2

in R

2

(or more generally, of the cube I

m

in R

m

) as considered first by S. Smale [5], defines a shift dynamics on the maximal invariant subset of I

2

(or I

m

). It is shown that this remains true almost surely for noninjective maps provided the contraction rate of the mapping in the stable direction is sufficiently strong, and bounds for this rate are given.

1. Definitions and results. For an integer θ ≥ 2 the set Σ

θ

of all doubly infinite sequences i = (. . . , i

−1

, i

0

, i

1

, . . .), where i

l

∈ {1, . . . , θ}, equipped with the metric

d((. . . , i

−1

, i

0

, i

1

, . . .), (. . . , j

−1

, j

0

, j

1

, . . .)) = X

l=−∞

2

−|l|

|i

l

− j

l

| is a Cantor set. The shift mapping σ : Σ

θ

→ Σ

θ

given by

σ(. . . , i

−1

, i

0

, i

1

, . . .) = (. . . , j

−1

, j

0

, j

1

, . . .) with j

l

= i

l+1

is a homeomorphism which defines a simple but nevertheless nontrivial dy- namics on Σ

θ

; e.g. its periodic points are dense, and there are dense orbits.

Therefore, to ask whether or not a given discrete dynamical system contains a subsystem conjugate to a shift space of this kind is a natural question.

Let R be a topological space with metric d, R

a compact subset of R and f : R

→ R continuous. For k ≥ 1 we define the compact sets

R

k

= {p ∈ R | f

k

(p) is defined}, A

k

= f

k

(R

k

).

Then R

1

= R

⊃ R

2

⊃ R

3

⊃ . . . , A

1

⊃ A

2

⊃ . . . , and we consider the compact sets

R

=

\

k=1

R

k

, A =

\

k=1

A

k

, Z = R

∩ A.

1991 Mathematics Subject Classification: 58F12, 58F15, 58F03.

Key words and phrases: horseshoes, noninvertible maps, shift spaces, attractors.

[267]

(2)

The set A, if not empty, can be regarded as a global attractor of f . Indeed, f (A∩R

) = A, and there is a sequence ε

1

> ε

2

> . . . of real numbers tending to 0 such that for any k > 1 and any p ∈ R

k

we have d(f

k

(p), A) ≤ ε

k

. The set Z is the maximal invariant subset of R, i.e. the maximal set on which f is defined, and f (Z) = Z.

A subset S of R

will be called a shift space in R if for some θ ≥ 2 there is a homeomorphism h : Σ

θ

→ S such that hσ = f h. Obviously, if S is a shift space in R then S ⊂ Z. If Z itself is a shift space in R then we say that f concentrates to a shift space.

Among the best known examples of mappings which concentrate to a shift space are the so called horseshoe mappings (introduced by S. Smale in [5]) which can be defined as follows. Let R

0

= R

m+1

(m ≥ 1) and let R

0

= I

m+1

= I × I

m

be the (m + 1)-dimensional unit cube in R

m+1

which is regarded as the cartesian product of the unit interval I = [0, 1] with the m-dimensional unit cube. To define a horseshoe mapping we fix disjoint subintervals I

1

, . . . , I

θ

in I (θ ≥ 2) and choose f : R

0

→ R

0

so that the following conditions are satisfied, where I

= I

1

∪ . . . ∪ I

θ

:

(i) f (R

0

) ∩ R

0

= f (I

× I

m

).

(ii) For some λ ∈ (0, 1) there are a C

1

mapping ϕ : I

→ I whose restriction to each component I

i

of I

is an expanding C

1

mapping onto I and a C

0

mapping ψ : I

→ [0, 1 − λ]

m

such that

(1) f (t, x) = (ϕ(t), ψ(t) + λx) ((t, x) ∈ I

× I

m

).

(iii) f is injective on I

× I

m

. (See Fig. 1, where m = 2, θ = 3.)

Fig. 1

(3)

It is well known (and not hard to prove) that f concentrates to a shift space Z. Moreover, the global attractor A of f is homeomorphic to the cartesian product I × C

0

of I with a Cantor set C

0

, and each component of A is a C

0

arc running upwards from the bottom {0} × I

m

of R

0

to the top {1} × I

m

. These facts remain true for more general mappings f (see e.g. [3], Ch. III), but they may fail to hold if (iii) is dropped from our assumptions (see Fig. 2, where m = 2, θ = 2).

Fig. 2

This paper is concerned with mappings f satisfying (i) and (ii). If θ and ϕ are fixed we shall show that for “almost all” ψ the mapping f concentrates to a shift space and A has the structure mentioned above even if f is not injective on I

× I

m

, provided λ is sufficiently small.

A natural technical simplification in the definition is obtained by ne- glecting the part of R

0

= R

m+1

outside R

0

= I × I

m

, i.e., we shall start with R = I × I

m

, R

= I

× I

m

and the restriction of the original f to f : R

→ R. Then the whole mapping f is defined by (1). We shall assume that θ, I

, ϕ : I

→ I and λ ∈ (0, 1) are fixed while ψ : I

→ [0, 1 − λ]

m

is variable. Then f is determined by ψ, and sometimes instead of f we shall write f

ψ

. The interval [0, 1 − λ] will be denoted by J.

The maximal subset I

k

of I on which ϕ

k

is defined (k = 0, 1, 2, . . .)

consists of θ

k

disjoint intervals, where I

0

= I ⊃ I

1

= I

⊃ I

2

⊃ I

3

⊃ . . . ,

(4)

and

I

=

\

k=0

I

k

is a Cantor set in I. The Hausdorff dimension dim

H

I

of I

coincides with the box counting dimension dim

B

I

(see [2]) and will be denoted by d

. The end points of the interval I

i

(1 ≤ i ≤ θ) will be denoted by s

i

, t

i

so that ϕ(s

i

) = 0 and ϕ(t

i

) = 1.

To avoid considerable technical difficulties (as e.g. piecewise linear ap- proximations of ϕ and ψ) Theorem 1, Theorem 2 and its corollaries will be restricted to the piecewise linear case; i.e., we shall assume that the restric- tions of ϕ and ψ to the components I

i

of I

are linear mappings onto I or into [0, 1 − λ]

m

, respectively. (See [1], where for nonlinear mapppings in a similar situation the attractor A is considered. Indeed, using the techniques applied there, facts analogous to those stated as Corollary 1 and Corollary 2 can be proved in the nonlinear case provided “full measure in J

2θm

” is replaced by

“open and dense in the space of all C

0

mappings ψ : I

→ J

m

”.)

Now we consider the piecewise linear case. Here d

is determined by

|I

1

|

d

+ . . . + |I

θ

|

d

= 1, where |I

i

| denotes the length of I

i

. Since the map- pings ψ are linear on each interval I

i

they are completely determined by the 2θ points a

i

= ψ(s

i

), b

i

= ψ(t

i

) in J

m

or, equivalently, by the point (a

1

, b

1

, a

2

, b

2

, . . . , a

θ

, b

θ

) in J

2θm

. So all possible mappings ψ are in 1-to-1 correspondence with the points in J

2θm

, and we shall not distinguish be- tween ψ and the corresponding point.

The following sets will play an important role in the piecewise linear case. (A denotes the global attractor of f

ψ

.)

Ψ = {ψ ∈ J

2θm

| f

ψ

does not concentrate to a shift space}, Ψ

A

= {ψ ∈ J

2θm

| f

ψ

|

A∩R

is not injective}.

In Section 2 (Proposition 3) we shall see that Ψ and Ψ

A

are compact, Ψ ⊂ Ψ

A

and that for ψ ∈ J

2θm

A

the global attractor A of f

ψ

is homeomorphic to the cartesian product of an interval with a Cantor set. Moreover, since A∩R

is compact and f

ψ

(A∩R

) = A, for each ψ ∈ J

2θm

A

the restriction f |

A∩R

: A∩R

→ A is a homeomorphism. The main results of this paper are stated in the following two theorems concerning the Hausdorff dimensions of Ψ and Ψ

A

.

Theorem 1. If λ < 1/2 then dim

H

Ψ ≤ 2θm −



m − d

2 log θ log(1/λ)



.

(5)

Theorem 2. If λ < 1/2 then dim

H

Ψ

A

≤ 2θm −



m − 1 − 2 log θ log(1/λ)

 .

Corollary 1. If λ < θ

−2/(m−d)

and λ < 1/2, then the set of all those ψ ∈ J

2θm

for which f

ψ

concentrates to a shift space is open in J

2θm

and has full measure (1 − λ)

2θm

.

Corollary 2. If m > 1, λ < θ

−2/(m−1)

and λ < 1/2, then for all ψ in an open subset of J

2θm

with full measure (1−λ)

2θm

the global attractor A of f

ψ

is the cartesian product of an interval with a Cantor set, and f

ψ

|

A∩R

: A ∩ R

→ A is a homeomorphism.

This corollary can be regarded as a partial answer to a problem of F. Przytycki in [4].

P r o o f o f t h e c o r o l l a r i e s. In these cases dim

H

Ψ < 2θm or dim

H

Ψ

A

< 2θm, respectively, and, by Proposition 3 of Section 2, Ψ and Ψ

A

are compact.

Propositions 1 and 2 in Section 2 will yield some further details.

R e m a r k 1. Our condition λ < 1/2 reflects the fact that two m-dimen- sional cubes in I

m

of edge length at least 1/2 and with edges parallel to those of I

m

must intersect. We do not know whether it is necessary. (Here it is essentially used only in the proof of Lemma 1.)

R e m a r k 2. We do not know whether the bounds for dim

B

Ψ and dim

B

Ψ

A

in the theorems are sharp. As is easily seen all points

ψ =



a

1

, a

1

, a

2

, 1 t a

1

+

 1 − 1

t



a

2

, a

3

, b

3

, . . . , a

θ

, b

θ



belong to Ψ if t ∈ I

\ {0} and to Ψ

A

if t ∈ (0, 1]. Therefore

dim

H

Ψ ≥ 2θm − (2m − d

), dim

H

Ψ

A

≥ 2θm − (2m − 1), but these lower bounds are rather weak, and they do not depend on λ.

The fact stated in Theorem 3 below seems to be more interesting than Remark 2. Though for small m the bounds for θ and λ are by no means exciting (owing to the factor 12 in the theorem) this theorem shows that the exponent −2/(m − 1) in Corollary 2 is optimal at least if m is odd.

Here it is possible, and by the topological methods used in the proof even

natural, to consider the general case, where ϕ and ψ are not necessarily

linear. (Concerning the piecewise linear case see Remark 4 below.) With m,

λ, I

and ϕ : I

→ I fixed (ϕ not necessarily piecewise linear) we consider

the space F of all C

0

mappings ψ : I

→ J

m

with the subspace G consisting

of all those ψ ∈ F for which the restriction of f

ψ

to A∩R

is not injective and

(6)

the attractor A does not have the structure mentioned above. Therefore G corresponds to the set Ψ

A

in the restricted case. By Proposition 3 of Section 2 the set G is closed in F.

Theorem 3. If m ≥ 3 is odd and λ > 12θ

−2/(m−1)

then G has interior points.

For m ≥ 3 and θ ≥ 2 let α(m, θ) be the infimum of all real α

0

> 0 with the following property. For each I

consisting of θ components, each ϕ : I

→ I (not necessarily piecewise linear) and each λ ∈ (α

0

θ

−2/(m−1)

, 1) the set G has interior points in F. Then Theorem 3 is equivalent to

(2) α(m, θ) ≤ 12 (m ≥ 3 odd).

R e m a r k 3. Besides (2) the proof of Theorem 3 will show for odd positive integers m,

(3) lim

m→∞

lim

θ→∞

α(m, θ) ≤ 8,

so that for each ε > 0 the factor 12 in the theorem can be replaced by 8 + ε provided m and θ are sufficiently large.

R e m a r k 4. The proof of Theorem 3 can be modified to show that for m, θ, λ as in the theorem the set G ∩ J

2θm

(J

2θm

⊂ F in the obvious way) has interior points in J

2θm

.

2. Preliminaries. For integers θ ≥ 2 and k

0

≤ k

00

let θ

[k0,k00]

be the set of all sequences (i

k0

, i

k0+1

, . . . , i

k00

) where i

l

∈ {1, . . . , θ}, and let θ

[−∞,k00]

, θ

[k0,∞]

, θ

[−∞,∞]

consist of the sequences which are infinite to the left, to the right or in both directions, respectively. So θ

[−∞,∞]

coincides with the Cantor set Σ

θ

of Section 1, and θ

[−∞,k00]

, θ

[k0,∞]

also have a natural Cantor set structure. The shift map σ : θ

[k0,k00]

→ θ

[k0−1,k00−1]

is defined in the obvious way.

As in Section 1 we assume that I

= I

1

∪ . . . ∪ I

θ

(θ ≥ 2) is the union of θ disjoint closed subintervals of I and that ϕ : I

→ I is a C

1

mapping whose restrictions to the intervals I

i

are expanding mappings onto I. Moreover, for some continuous ψ : I

→ J

m

let f : R

= I

× I

m

→ R = I × I

m

be defined by (1).

The θ

k

components of the domain I

k

of ϕ

k

(k ≥ 1) will be denoted by I

i

(i ∈ θ

[1,k]

), where the indices are chosen so that for k > 1,

I

(i1,...,ik)

⊂ I

(i1,...,ik−1)

, ϕ(I

(i1,...,ik)

) = I

(j1,...,jk−1)

, where j

l

= i

l+1

. For i = (i

1

, i

2

, . . .) ∈ θ

[1,∞]

the intersection T

k=1

I

(i1,...,ik)

contains exactly

one point which will be denoted by t

i

. The sets R

i

= I

i

× I

m

(i ∈ θ

[1,k]

,

(7)

1 ≤ k < ∞) are slices of R = I × I

m

while for i = (i

1

, i

2

, . . .) ∈ θ

[1,∞]

, R

i

=

\

k=1

R

(i1,...,ik)

is the m-dimensional cube {t

i

} × I

m

.

For i ∈ θ

[1,k00]

(1 ≤ k

00

≤ ∞) and 1 ≤ k

0

≤ k

00

, k

0

< ∞, the image f

k0

(R

i

) is well defined and will be denoted by R

σk0(i)

. So R

i

is now defined for all i ∈ θ

[k0,k00]

provided k

0

≤ k

00

, and −∞ < k

0

≤ 1 and 0 ≤ k

00

≤ ∞. By setting

R

i

=

−∞

\

k=0

R

(ik,...,i0,...)

for i = (. . . , i

−1

, i

0

, . . .) ∈ θ

[−∞,k00]

(0 ≤ k

00

≤ ∞) we include the case k

0

= −∞ into our definition.

R

(ik0,...,i0)

R

(ik0,...,ik00)

R

(i1,...,ik00)

Fig. 3

For k

0

and k

00

finite with k

0

≤ 0 the set R

i

is an (m + 1)-dimensional curved prism over an m-dimensional cube with edge length λ

−k0+1

, which for k

00

= 0 has its bottom in {0} × I

m

and its top in {1} × I

m

, while for k

0

≤ 0 and k

00

≥ 1,

R

(ik0,...,ik00)

= R

(ik0,...,i0)

∩ R

(i1,...,ik00)

(see Fig. 3). In the piecewise linear case all these prisms are straight. For i ∈ θ

[−∞,0]

the set R

i

is an arc (or a straight segment in the piecewise linear case) running upwards from a point in {0} × I

m

to a point on {1} × I

m

, and if i ∈ θ

[−∞,∞]

then R

i

contains exactly one point which will be denoted by p

i

. As is easily seen,

(4) f (R

i

) = R

σ(i)

wherever R

i

and R

σ(i)

are defined. Moreover, R

j

⊂ R

i

provided i is a part

of j, i.e., if i can be obtained from j by cancelling digits on one or both

(8)

ends. The domain of f

k

(k ≥ 1) is

R

k

= I

k

× I

m

= [

i∈θ[1,k]

R

i

, and

R

= I

× I

m

=

\

k=1

R

k

is the maximal set on which all iterations f

k

(k ≥ 1) are defined.

The global attractor of f is given by

A = [

i∈θ[−∞,0]

R

i

. The maximal invariant set of f is

Z = [

i∈θ[−∞,∞]

R

i

,

i.e., Z consists of the points p

i

(i ∈ θ

[−∞,∞]

), and by setting h(i) = p

i

we get a surjective mapping h : Σ

θ

= θ

[−∞,∞]

→ Z. As is easily seen, h is continuous, and (4) implies hσ = f h. For t ∈ I and i ∈ θ

[−∞,0]

we define g(t, i) to be the intersection point of {t} × I

m

and R

i

. So we get a surjective continuous mapping g : I × θ

[−∞,0]

→ A.

Proposition 1. The following conditions are equivalent.

(i) f concentrates to a shift space.

(ii) h : Σ

θ

→ Z is a homeomorphism.

(iii) If i, j ∈ θ

[−∞,0]

, i 6= j, then R

i

∩ R

j

∩ R

= ∅.

P r o o f. The equivalence of (ii) and (iii) is an immediate consequence of the following fact. If i

= (. . . , i

−1

, i

0

) ∈ θ

[−∞,0]

then the mapping h

i

: θ

[1,∞]

→ R

i

∩ R

given by h

i

(i

1

, i

2

, . . .) = h(. . . , i

−1

, i

0

, i

1

, . . .) is a homeomorphism.

The implication (ii)⇒(i) follows from (4).

To complete the proof we assume (i) and prove (ii). Since Σ

θ

is compact and h is surjective it is sufficient to show that h is injective.

If for i = (. . . , i

−1

, i

0

, i

1

, . . .), j = (. . . , j

−1

, j

0

, j

1

, . . .) ∈ θ

[−∞,∞]

the positive halves i

+

= (i

1

, i

2

, . . .) and j

+

= (j

1

, j

2

, . . .) are different, then h(i) ∈ R

i+

, h(j) ∈ R

j+

, R

i+

∩ R

j+

= ∅ implies h(i) 6= h(j). If i

+

= j

+

but i 6= j then for some k < 0 the positive halves σ

k

(i)

+

and σ

k

(j)

+

of σ

k

(i) and σ

k

(j) will differ, and we get

h(σ

k

(i)) 6= h(σ

k

(j)).

(9)

By (i), f |

Z

: Z → Z is a homeomorphism, and (f |

Z

)

k

h = hσ

k

for our negative exponent k. So we get (f |

Z

)

k

h(i) 6= (f |

Z

)

k

h(j) and therefore h(i) 6=

h(j).

Proposition 2. The following conditions are equivalent.

(i) f |

A∩R

: A ∩ R

→ A is a homeomorphism.

(ii) g : I × θ

[−∞,0]

→ A is a homeomorphism.

(iii) If i, j ∈ θ

[−∞,0]

, i 6= j, then R

i

∩ R

j

= ∅.

P r o o f. Since g maps each interval I × {i} injectively onto R

i

, the equiv- alence of (ii) and (iii) is obvious.

Now we prove (i)⇒(iii). By (i) for k ≥ 1 the mapping f

k

: A ∩ R

k

→ A is a homeomorphism. To prove (iii) we show that for i = (. . . , i

−1

, i

0

), j = (. . . , j

−1

, j

0

) ∈ θ

[−∞,0]

the existence of a common point p = (t, x) of R

i

and R

j

(t ∈ I, x ∈ I

m

) implies i = j.

For k ≥ 1 there is a unique p

= (t

, x) ∈ A ∩ R

k

such that f

k

(p

) = p.

Here t

∈ I

i

, where i

= (i

1

, . . . , i

k

) ∈ θ

[1,k]

with i

l

= i

l−k

= j

l−k

(1 ≤ l

≤ k). Since k ≥ 1 is arbitrary this shows i

n

= j

n

for all n ≤ 0.

To prove (iii)⇒(i) we assume that all arcs R

i

(i ∈ θ

[−∞,0]

) are disjoint.

Then each component of A∩R

is an arc R

i

∩R

i

(i = (. . . , i

−1

, i

0

) ∈ θ

[−∞,0]

, 1 ≤ i ≤ θ), and f maps this arc injectively onto R

j

, where j = (. . . , j

−1

, j

0

) ∈ θ

[−∞,0]

is given by j

l

= j

l+1

if l < 0 and j

0

= i. So f is injective on each component of A∩R

, and by (iii) different components have disjoint images.

Since A ∩ R

is compact, injectivity of f |

A∩R

together with f (A ∩ R

) = A implies (i).

Proposition 3. Ψ and Ψ

A

are compact, and G is closed in F.

P r o o f. Since the proofs of the three assertions are similar we consider Ψ (in the piecewise linear case) only. For ψ ∈ J

2θm

, f = f

ψ

: R

→ R the cor- responding mapping and 1 ≤ i ≤ θ let Z

i

(ψ) denote the union of all R

i

∩R

, where i = (. . . , i

−1

, i

0

) ∈ θ

[−∞,0]

and i

0

= i. Obviously Z

1

(ψ), . . . , Z

θ

(ψ) are compact and their union is the set Z belonging to f

ψ

.

If f

ψ

concentrates to a shift space then (by Proposition 1(iii)) the sets Z

i

(ψ) are disjoint. We show that the converse also holds. Suppose that Z

1

(ψ), . . . , Z

θ

(ψ) are disjoint, and let i = (. . . , i

−1

, i

0

, i

1

, . . .), j = (. . . , j

−1

, j

0

, j

1

, . . .) ∈ θ

[−∞,∞]

, i 6= j, be given. We have to show that h(i) 6= h(j).

If i

l

6= j

l

for some l ≥ 1, then h(i) and h(j) lie in different components

of R

, and h(i) 6= h(j) is obvious. Now we assume that l

0

≤ 0 is the

maximal index with i

l0

6= j

l0

. Then for i

0

= (. . . , i

0−1

, i

00

, i

01

, . . .) = σ

l0

(i) and

(10)

j

0

= (. . . , j

−10

, j

00

, j

10

, . . .) = σ

l0

(j) we have i

00

6= j

00

but i

0l

= j

l0

if l ≥ 1. The points h(i

0

) and h(j

0

) lie in the same component {t} × I

m

of R

but in different and therefore disjoint sets Z

i0

0

(ψ) and Z

j0

0

(ψ). So h(i

0

) 6= h(j

0

), and since f

−l0

is injective on {t} × I

m

this gives

h(i) = hσ

−l0

(i

0

) = f

−l0

h(i

0

) 6= f

−l0

h(j

0

) = hσ

−l0

(j

0

) = h(j).

To prove that Ψ is compact we show that each point ψ ∈ J

2θm

\ Ψ has a neighbourhood which does not intersect Ψ . If ψ 6∈ Ψ the corresponding sets Z

1

(ψ), . . . , Z

θ

(ψ) are disjoint and since they are compact there is a positive ε such that the distance between any two of them is at least ε.

As is easily seen, the segments R

i

(i ∈ θ

[−∞,0]

) depend continuously on ψ, and this continuity is uniform with respect to i. Therefore, if ψ

0

∈ J

2θm

is sufficiently close to ψ the sets Z

i

0

) belonging to ψ

0

are close to the sets Z

i

(ψ) and hence mutually disjoint. This proves ψ

0

6∈ Ψ .

3. Proof of Theorems 1 and 2. We assume that ϕ : I

→ I, λ ∈ (0, 1/2) and therefore θ, I

k

(1 ≤ k ≤ ∞), I

i

, R

i

(i ∈ θ

[1,k]

, 1 ≤ k ≤ ∞) and t

i

(i ∈ θ

[1,∞]

) are fixed. Let H denote one of the sets I

or I, and let q

= dim

H

H = dim

B

H. We define

Ψ

= {ψ ∈ J

2θm

| R

i

(ψ) ∩ R

j

(ψ) ∩ (H × I

m

) 6= 0

for at least one pair i 6= j ∈ θ

[−∞,0]

}, where R

i

(ψ) denotes the set R

i

which is constructed with the mapping ψ.

Looking at the equivalences between (i) and (iii) of the first two propositions in Section 2 we see that both theorems of Section 1 are combined in

(5) dim

H

Ψ

≤ 2θm −



m − q

2 log θ log(1/λ)

 .

We shall prove (5) at the end of this section after some lemmas are stated and proved.

Besides Ψ

, for 1 ≤ k < ∞, i = (i

1

, . . . , i

k

), j = (j

1

, . . . , j

k

) ∈ θ

[1,k]

, i 6= j, we shall consider the sets

(6)

Ψ

i,j

= {ψ ∈ J

2θm

| R

σk(i)

(ψ) ∩ R

σk(j)

(ψ) ∩ (H × I

m

) 6= ∅}, Ψ

k

= [

i,j∈θ[1,k]

ik6=jk

Ψ

i,j

.

(11)

Since R

(l−k,...,l0)

⊂ R

(l−k+1,...,l0)

, we have Ψ

1

⊃ Ψ

2

⊃ . . . , and together with

Ψ

=

\

k=1

[

i,j∈θ[1,k]

i6=j

Ψ

i,j

the proof of Proposition 3 implies

(7) Ψ

=

\

k=1

Ψ

k

.

For k ≥ 1 and i, j ∈ θ

[1,k]

, i 6= j, we define the mapping π

i,j

: J

2θm

→ I

4m

= (I

m

)

4

by π

i,j

(ψ) = (a, b, c, d), where the points a, b, c, d ∈ I

m

are determined by f

ψk

(s

i

, o) = (0, a), f

ψk

(t

i

, o) = (1, b),

f

ψk

(s

j

, o) = (0, c), f

ψk

(t

j

, o) = (1, d),

with s

i

, t

i

the end points of I

i

such that ϕ

k

(s

i

) = 0 and ϕ

k

(t

i

) = 1, and o = (0, . . . , 0) ∈ I

m

. Therefore (0, a), (1, b) are the end points of the segment f

ψk

(I

i

× {o}) and (0, c), (1, d) those of f

ψk

(I

j

× {o}). Moreover, the segments [(0, a), (1, b)] and [(0, c), (1, d)] are edges of the prisms f

k

(R

i

) = R

σk(i)

and f

k

(R

j

) = R

σk(j)

, respectively, such that for (t, y) ∈ [(0, a), (1, b)] and (t, z) ∈ [(0, c), (1, d)] we have the cubes

(8) R

σk(i)

∩ ({t} × I

m

) = {t} × (y + [0, λ

k

]

m

), R

σk(j)

∩ ({t} × I

m

) = {t} × (z + [0, λ

k

]

m

).

For (a, b, c, d) ∈ (I

m

)

4

= I

4m

we define

π(a, b, c, d) = (c − a, d − b)

and get a mapping π : I

4m

→ [−1, 1]

2m

. Finally, we consider the composition

%

i,j

= ππ

i,j

: J

2θm

→ I

2m

.

Lemma 1. There is a real α

1

> 0 not depending on k, i = (i

1

, . . . , i

k

), j = (j

1

, . . . , j

k

) ∈ θ

[1,k]

such that for any measurable set X in I

4m

,

vol

2θm

−1

i,j

(X)) ≤ α

1

vol

4m

(X)

provided i

k

6= j

k

. (By vol

p

we denote the p-dimensional Lebesgue measure

in R

p

.)

(12)

Lemma 2. There is a real α

2

> 0 such that for any measurable set X in [−1, 1]

2m

,

vol

4m

−1

(X)) ≤ α

2

vol

2m

(X).

Corollary. There is a real α > 0 not depending on k, i = (i

1

, . . . , i

k

), j = (j

1

, . . . , j

k

) ∈ θ

[1,k]

such that for any measurable set X in [−1, 1]

2m

,

vol

2θm

(%

−1

i,j

(X)) ≤ α vol

2m

(X) provided i

k

6= j

k

.

Since the proof of Lemma 2 is trivial it is sufficient to prove Lemma 1.

P r o o f o f L e m m a 1. We start with the remark that π

i,j

can be ex- tended to a linear mapping π

i,j

: R

2θm

→ R

4m

.

The proof will proceed as follows. We define a 4m-dimensional linear subspace L of R

2θm

(depending on i, j) such that π

i,j

|

L

: L → R

4m

is a linear isomorphism and that for any measurable set X in R

4m

we have (9) vol

4m

((π

i,j

|

L

)

−1

(X)) ≤ α

vol

4m

(X),

where

α

=

 1 − λ 1 − 2λ



4m

.

(This is the point where we need λ < 1/2.) Obviously π

i,j

= π

i,j

|

L

π

with a linear projection π

: R

2θm

→ L, and therefore, if X ⊂ I

4m

then

vol

2θm

−1

i,j

(X)) = vol

2θm

−1

i,j

(X) ∩ J

2θm

)

= vol

2θm

∗−1

i,j

|

L

)

−1

(X) ∩ J

2θm

)

≤ (diam J

2θm

)

2θm−4m

vol

4m

((π

i,j

|

L

)

−1

(X))

≤ (diam J

2θm

)

2θm−4m

α

vol

4m

(X), so that the lemma will be proved with

α

2

= (diam J

2θm

)

2θm−4m

 1 − λ 1 − 2λ



4m

, provided (9) is proved.

Thinking of our identification of the mappings ψ : I

→ J

m

with the points in J

2θm

we regard J

2θm

as (J

m

)

and its points as sequences (a

1

, b

1

, . . . , a

θ

, b

θ

), where a

i

, b

i

∈ J

m

. Let J

4m

i,j

denote the 4m-dimensional face of J

2θm

consisting of all (a

1

, b

1

, . . . , a

θ

, b

θ

) with a

i

= b

i

= o for i

k

6=

i 6= j

k

. (Here i

k

, j

k

are the last digits of i, j, respectively, and o denotes the

(13)

point (0, . . . , 0) in R

m

.) Then L is defined to be the 4m-dimensional linear subspace of R

2θm

which contains J

4m

i,j

.

Since π

i,j

is linear there is a real δ such that for any measurable Y in L we have

vol

4m

i,j

(Y )) = δ vol

4m

(Y ),

and, since vol

4m

J

i,j4m

= (1 − λ)

4m

, to prove (9) it is sufficient to show that

vol

4m

i,j

(J

i,j4m

)) ≥

 1 − 2λ 1 − λ



4m

(1 − λ)

4m

= (1 − 2λ)

4m

or that π

i,j

(J

4m

i,j

) contains the cube Q = [λ, 1 − λ]

4m

.

It will be convenient to identify L with R

4m

via the mapping L → R

4m

which is obtained by neglecting in points (x

1

, . . . , x

2θm

) = (a

1

, b

1

, . . . , a

θ

, b

θ

)

∈ L (a

i

, b

i

∈ R

m

) all coordinates not belonging to a

ik

, b

ik

, a

jk

, b

jk

. Then J

4m

i,j

= J

4m

and we have to show

(10) π

i,j

(J

4m

) ⊃ Q.

Starting with the cube Q

= [0, λ]

4m

for each vertex ψ of J

4m

we define the cube Q

ψ

= ψ + Q

. By a simple geometric argument illustrated in Figure 4 it can be proved that any convex set which intersects all 2

4m

cubes Q

ψ

must contain Q. Therefore to prove (10) it is sufficient to show that for any vertex

ψ + Q

ψ + Q

J

4m

Q

Fig. 4

(14)

ψ of J

4m

, π

i,j

(ψ) ∈ Q

ψ

, or, equivalently,

(11) π

i,j

(ψ) − ψ ∈ [0, λ]

4m

.

Assume i

k

< j

k

. For a vertex ψ = (a

ik

, b

ik

, a

jk

, b

jk

) of J

4m

we shall write π

i,j

(ψ) = π

i,j

(ψ) = (a, b, c, d). To prove (11) it is sufficient to prove

(12) a − a

ik

, b − b

ik

, c − a

jk

, d − b

jk

∈ [0, λ]

m

.

We consider a − a

ik

; the remaining cases are analogous. Our identification ψ = (a

1

, b

1

, . . . , a

θ

, b

θ

) made in Section 1 implies, for 1 ≤ i ≤ θ,

f

ψ

(R

i

) ∩ ({0} × I

m

) = f

ψ

({s

i

} × I

m

) = {0} × (a

i

+ [0, λ]

m

).

Therefore we have by the definition of π

i,j

,

(0, a) = f

ψk

(s

i

, o) = f

ψ

f

ψk−1

(s

i

, o)

and, since ϕ(s

(i1,...,il)

) = s

(i2,...,il)

((i

2

, . . . , i

l

) regarded as element of θ

[1,l−1]

), f

ψk−1

(s

i

, o) ∈ {ϕ

k−1

(s

i

)} × I

m

= {s

ik

} × I

m

⊂ R

ik

.

Therefore

(0, a) ∈ f

ψ

(R

ik

) ∩ ({0} × I

m

) = {0} × (a

ik

+ [0, λ]

m

), which proves (12) for a − a

ik

and hence the lemma.

We consider the compact subset

K = {(a, b) ∈ ([−1, 1]

m

)

2

= [−1, 1]

2m

| (1 − t)a + tb = o for some t ∈ H}

of [−1, 1]

2m

.

Lemma 3. Let (a, b, c, d) ∈ I

4m

. Then the segments [(0, a), (1, b)] and [(0, c), (1, d)] intersect in a point (t, x) with t ∈ H and x ∈ I

m

if and only if π(a, b, c, d) ∈ K.

This lemma is an immediate consequence of the definitions of π and K.

Lemma 4. There is a real β > 0 such that for any k ≥ 1 and i, j ∈ θ

[1,k]

, i 6= j, we have

N

λk

i,j

) ⊂ %

−1

i,j

(N

βλk

(K)),

where N

λk

i,j

) denotes the λ

k

-neighbourhood of Ψ

i,j

in J

2θm

while N

βλk

(K) is the βλ

k

-neighbourhood of K in [−1, 1]

2m

.

P r o o f. For an arbitrarily given ψ = (a

1

, b

1

, . . . , a

θ

, b

θ

) ∈ N

λk

i,j

) we choose ψ

0

= (a

01

, b

01

, . . . , a

0θ

, b

0θ

) ∈ Ψ

i,j

so that

|a

0i

− a

i

| ≤ λ

k

, |b

0i

− b

i

| ≤ λ

k

(1 ≤ i ≤ θ).

(15)

A simple geometric argument (by induction with respect to k) shows that for

(a, b, c, d) = π

i,j

(ψ), (a

0

, b

0

, c

0

, d

0

) = π

i,j

0

) each of the distances |a

0

− a|, |b

0

− b|, |c

0

− c|, |d

0

− d| is at most

(13) λ

k

k−1

X

i=0

λ

i

< λ

k

1 − λ < 2λ

k

.

(The last inequality is a consequence of our assumption λ < 1/2. Instead of applying this assumption we could proceed with 1/(1 − λ) instead of 2 and choose β = 4/(1 − λ) + 4

m. Therefore in this proof λ < 1/2 is not essential.) As an immediate consequence of (13) we have

i,j

0

) − π

i,j

(ψ)| < 4λ

k

and from |π(p) − π(q)| < 2|p − q| we get

(14) |%

i,j

0

) − %

i,j

(ψ)| < 8λ

k

.

Since ψ

0

∈ Ψ

i,j

, we can find points t ∈ H and x ∈ I

m

such that (15) (t, x) ∈ R

σk(i)

0

) ∩ R

σk(j)

0

).

Let (t, y) and (t, z) be the points at which {t} × I

m

intersects the segments f

ψk0

(I

i

× {o}) and f

ψk0

(I

j

× {o}), respectively. The end points of these seg- ments are (0, a

0

), (1, b

0

) and (0, c

0

), (1, d

0

) respectively, and (8) together with f

ψk0

(I

i

× {o}) ⊂ R

σk(i)

0

), f

ψk0

(I

j

× {o}) ⊂ R

σk(j)

0

) and (15) implies

(16) |x − y| ≤

m λ

k

, |x − z| ≤ m λ

k

.

Let a

= a

0

+ x − y, b

= b

0

+ x − y, c

= c

0

+ x − z and d

= d

0

+ x − z. Then (a

, b

, c

, d

) ∈ I

4m

, and since (t, x) ∈ [(0, a

), (1, b

)] ∩ [(0, c

), (1, d

)] and t ∈ H, by Lemma 3 we have π(a

, b

, c

, d

) ∈ K.

Applying (16) we get

|(a

0

, b

0

, c

0

, d

0

) − (a

, b

, c

, d

)| ≤ 2 m λ

k

and therefore, by the definition of π,

dist(%

i,j

0

), K) ≤ |π(a

0

, b

0

, c

0

, d

0

) − π(a

, b

, c

, d

)| ≤ 4 m λ

k

. This together with (14) shows %

i,j

(ψ) ∈ N

βλk

(K), where β = 8 + 4

m.

Lemma 5. dim

B

K = m + q

.

(16)

P r o o f. K is the intersection of a cone with [−1, 1]

2m

, i.e., if v ∈ K, γ ∈ R and γv ∈ [−1, 1]

2m

, then γv ∈ K. The full cone is

K = {γv | v ∈ K, γ ∈ R}

= {(a, b) ∈ (R

m

)

2

= R

2m

| (1 − t)a + tb = 0 for some t ∈ H}, and K = K ∩ [−1, 1]

2m

. So it is sufficient to prove

dim

B

K = m + q

.

To describe K we consider the boundary ∂(D

m

× D

m

) = (S

m−1

× D

m

)

∪ (D

m

× S

m−1

) of the topological ball D

m

× D

m

in R

2m

, where D

m

= {a ∈ R

m

| |a| ≤ 1} and S

m−1

= {a ∈ R

m

| |a| = 1}. Then, since

dim

B

K = 1 + dim

B

(∂(D

m

× D

m

) ∩ K), it is sufficient to show

(17) max[dim

B

((S

m−1

× D

m

) ∩ K), dim

B

((D

m

× S

m−1

) ∩ K)] = m − 1 + q

. We consider the first term

(S

m−1

× D

m

) ∩ K =



a, t − 1 t a



a ∈ S

m−1

, t ∈ H ∩ [1/2, 1]

 . Let F = S

m−1

× [1/2, 1], and let χ : F → S

m−1

× D

m

be the mapping given by

χ(a, t) =



a, t − 1 t a

 .

Obviously, χ is an injective C

embedding satisfying χ(S

m−1

×(H∩[1/2, 1]))

= (S

m−1

× D

m

) ∩ K. Then since

dim

B

(S

m−1

× (H ∩ [1/2, 1])) = m − 1 + dim

B

(H ∩ (1/2, 1]), we have

dim

B

((S

m−1

× D

m

) ∩ K) = m − 1 + dim

B

(H ∩ [1/2, 1]) if H ∩ [1/2, 1] 6= ∅.

In the same way we get (D

m

× S

m−1

) ∩ K =

 t t − 1 b, b



b ∈ S

m−1

, t ∈ H ∩ [0, 1/2]

 , dim

B

((D

m

× S

m−1

) ∩ K) = m − 1 + dim

B

(H ∩ [0, 1/2]) if H ∩ [0, 1/2] 6= ∅.

Since q

= max(dim

B

(H ∩ [0, 1/2]), dim

B

(H ∩ [1/2, 1])), this implies (17).

To prove (5) we apply the following result of C. Tricot Jr. [6], in which

dim

B

and dim

B

denote the upper and the lower box counting dimension,

respectively (see e.g. [2]).

Cytaty

Powiązane dokumenty

In the case of arbitrary ζ we use the same arguments based on the obvious relativization of Lemma 1.. General version of the Nadel’s theorem. We assume that the reader is familiar

Math 3CI Even More about solving DiffyQ Symbolicallly Part IV In these problems you are pushed to develop some more symbolic tech- niques for solving ODE’s that extends the

A finite atomistic lattice L is isomorphic to the lattice Sub(P ) for some finite semilattice P iff it satisfies D 2 , has no cycles, is biatomic, has univocally terminating

In 1922, Polish mathematician Stefan Banach proved the following famous fixed point theorem [1]: Let X be a complete metric space with metric d.. This theo- rem called the

Similar result for linear regression and classification for any algorithm, which maintains its weight as a linear combination of instances [Warmuth &amp; Vishwanathan, 2005;

In [Sz], Szymczak introduces a new technique of detection of chaos based on the construction of the Conley index for decompositions of isolated invariant sets.. He defines an index

Proof.. Theorems on continuous mappings. 1 ), intima and suprema of linear pseudo­..

A theorem on generating multimodular pseudotopologies by families of filters is given.. The Orlicz topology is constructed for our