• No results found

Computation of the Ruelle-Sullivan map for substitution tilings

N/A
N/A
Protected

Academic year: 2021

Share "Computation of the Ruelle-Sullivan map for substitution tilings"

Copied!
103
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

Computation of the Ruelle-Sullivan Map for

Substitution Tilings

Charles Benjimen Starling B.Sc., University of Victoria, 2003

A

Thesis Submitted in Partial Fulfillment of the Requirements for the Degree of

MASTER OF SCIENCE

in the Department of Mathematics and Statistics

@Charles Benjimen Starling, 2005 University of Victoria

A11 rights reserved. This thesis may not be reproduced in whole or in part, by photocopy or other means, without the permission of the author.

(2)

Abstract

We study the dynamics of tiling spaces through cohomology. An adapta- tion of the ~ e c h - d e ~ h a m theorem allows us to compute the Ruelle-Sullivan map for such spaces and consider its image together with cohomology as a more useful invariant than cohomology alone. Computation of the map is performed for the Penrose tiling and the Octagonal tiling.

(3)

Contents

. .

A b s t r a c t . . .

. . . . . .

. . . . . . . . . . .

.

. . .

. . . . . . 11

Table of Contents

. .

. . . .

. .

. . . .

. . . .

. . . . . . .

. . . . .

iii List of Figures. .

. . . .

. . . . . . . .

.

.

. . . . . . . . . .

. .

. .

v

1 Introduction 1

2 Tilings and Tiling Spaces 5

2.1 Cell Complexes

. . . . .

. . . . . . .

.

. . . . . . . . . .

5 2.2 Tilings

. . . . . . . . . . . . . . . . . . . . . . .

. . .

7 2.3 The Tiling Space .

. . . . . . . .

.

. . . . . . . . .

.

.

. .

8 2.4 Substitution and the Anderson-Putnam Complex

. . . .

.

. .

12 2.5 OT as an Inverse Limit

. . . . . . . .

.

.

. . . . . . .

. . . 15

3 Cohomology 22

3.1 Cohomology in General .

. . . .

.

. .

.

. . . . .

. .

.

. .

22 3.2 Orientation, Incidence Number, and Cellular Cohomology

. .

. 23 3.3 ~ e c h Cohomology

.

. . . . . . . . . .

. . . . .

. . . . . . 27 3.4 Dynamical Cohomology

. . . . .

.

.

. . . . . .

. . . . . . 30

(4)

CONTENTS iv

4.1 Mapping Cellular Cohomology of

ro

to the ~ e c h Cohomology of RT when n = 2

. . .

33

4.2 Mapping cech Cohomology to Dynamical Cohomology to ARn* 40 4.3 Finding a Translation Invariant Measure on RT when n = 2

.

52

. . .

4.4 Computing

(CF,

Otp. tr o a l ( - ) ) 59

5 Computations 72

. . .

5.1 The Octagonal Tiling 72

. . .

5.2 The Penrose Tiling 85

(5)

List of

Figures

. . .

3.1 A typical 2-cell 24

. . .

4.1 O(pi.

V)

34

4.2 Two Possible Arrangements of 3 Vertex Patterns with Non-

. . .

empty Intersection 39

. . .

4.3 Properties of pel. 60

. . .

4.4 pel. EPezrE 61

. . .

4.5 Splitting of a polygon into regions 62

. . .

4.6 O(vi) with the polygon pi 63

4.7 T ( a )

. . .

65

. . .

4.8 Three Vertex Patterns and their Intersections 66

. . .

4.9 hal f ( V l ) 67

. . .

4.10 -dhal f ( V l ) 68

. . .

4.11 l ( f ) and L ( f ) onpi 69

. . .

5.1 Substitution Rule for Octagonal Tiling 73

. . .

5.2 A Patch of an Octagonal Tiling 79

5.3 Cell Complex Generated by the Octagonal Tiling

. . .

80

. . .

(6)

LIST O F FIGURES vi

5.5 Representation of vs on the Cell Complex

. . .

81

5.6 Representation of ud on the Cell Complex

. . .

82

5.7 vl Represented on a Larger Patch

. . .

83

5.8 v4 Represented on a Larger Patch

. . .

84

5.9 Cell Complex and Substitution for the Penrose Tiling

. . .

86

5.10 Lengths Needed for Computations

. . .

90

5.11 Patch of a Penrose Tiling

. . .

91

5.12 First Generator of H 1

(ro)

Represented on a Larger Patch

. . .

92

5.13 First Generator of H 1 ( r o ) represented on Penrose Cell Complex 93 5.14 Fifth Generator of H1(ro) Represented on a Larger Patch

. . .

94

(7)

Chapter

1

Introduction

In the 60's and 70's, patterns were discovered in nature which were aperiodic, yet displayed some long-range order. These were usually crystaline in nature, but went against the laws of crystals as they were known at the time. These were studied with great interest, and eventually this phenomenon came to be called aperiodic order. Tilings are the major example of objects that can display aperiodic order. Much of the study of aperiodic order comes down to the study of certain tilings.

When one thinks of a tiling, what usually comes to mind is a collection of polygons fitting together to cover the plane. The mathematical definition of a tiling extends this to mean a collection of subsets of

Rn

homeomorphic to the closed unit ball in

Rn

whose interiors are pairwise disjoint and whose union is all of

Rn.

From any tiling T we can form an

Rn

action on a topological space

SZT; in this way we study aperiodic order though dynamics. As has been done in [AP] and elsewhere, one way of studying the order in these systems is though the cohomology. This provides some important invariants, but fails

(8)

CHAPTER 1. INTRODUCTION 2

to show the whole picture. In [KP] the authors provide a way of obtaining a map on cohomology which can distinguish between two different Rn-actions with the same cohomology. The goal of this thesis then is t o compute this, the Ruelle-Sullivan map, for tiling systems.

Shortly after Chapter 2 begins we give the definition of a tiling given above with the hypothesis that all the tiles are translates of some member of a finite set {pl,p2,

. . .

,pN); the pi's are called prototiles. We also make some fairly standard constructions, including the notion of translating a tiling - the translate of a tiling is just the tiling obtained by translating each of the component tiles. The first non-intuitive construction is the definition of a .metric on a collection of tilings - this metric basically states that two tilings are close if they agree up t o a small translation on a large ball around the origin. We then form the tiling space QT by taking T, taking all translates of it by vectors in Rn and completing this collection in the metric; the elements of this completion are shown to be tilings themselves. The space QT is shown to be compact if we make a hypothesis on T called finite local complexity - that there are only a finite number of different looking patches in T of any given radius, up t o translation. We then assume further that we have a substitution rule on our prototiles - we have a constant X

>

1 and, for each prototile, a rule for subdividing it into pieces, each of which is another prototile, scaled down by a factor of X-l. This idea extends to patches of tiles and whole tilings, so we can construct the dynamical system with the space !-IT and the substitution map w . Assumptions are made on the substitution so that w is a homeomorphism and thus

(aT,

W ) is a topological dynamical system.

(9)

CHAPTER 1. INTRODUCTION 3

We then construct a cell complex

I'

from $IT as in [AP]. Basically, the n-cells of

I'

are the prototiles with their faces identified if they are adjacent anywhere in any tiling in

nT.

A map

y

based on the substitution is defined on

I?

and is shown to be onto, so we construct the inverse limit limy

J?

and show it to be topologically conjugate to ($IT, w ) under the assumption that the substitution forces its border - this is explained in the section.

After defining cellular, ~ e c h and dynamical cohomology, we begin to con- nect them for the case n = 2 in Chapter 4. To map cellular cocycles to Cech cocycles, an open cover

2

4

is constructed where each open set corresponds to a vertex pattern in T. We then have a map that takes a vertex pattern to the 0-cell at its center, and so this induces a map on the cellular 0-cochains to the ~ e c h 0-cochains. We define a similar map on 1-cochains and extend this to a map on cohomology.

Next comes an adaptation of the ~ e c h - d e ~ h a m theorem [BT] to connect the ~ e c h cohomology to the dynamical cohomology, mapping ~ e c h cocycles to smooth functions on our space. We then define the Ruelle-Sullivan map which takes such functions and integrates them over an invariant probability measure on $IT. In the case of n = 2, these integrate to vectors in Rn*. The philosophy, as suggested in [KP], is that the long range aperiodic order of an aperiodic tiling is given by the its first cohomology group together with the image of the Ruelle-Sullivan map.

To demonstrate this, in Chapter 5 we compute this map for two very

different tilings which have isomorphic first cohomology groups. The first is the octagonal tiling, consisting of two labeled 1, 1,

fi

triangles and a rhomb, along with all their rotations though

y.

We find that the first cohomology

(10)

CHAPTER 1. INTRODUCTION 4 group to be Z5, and we can find a generating set with one of the generators mapping to 0. Three of th remaining generators map to rotates of each other by multiples of

%

while the fifth points in a direction between the first two, that is, at a multiple of

g.

This highlights the symmetry of the tiling through rotations of

$.

The second is the famous kite-and-dart tiling of Penrose. To allow for a substitution rule, the kites and darts have been split into triangles, so that we have 40 prototiles - two differently shaped triangles each given two different labels, and all their rotations though

y.

We compute the first cohomology group to be isomorphic to Z5 and the image of one of its generators under the Ruelle-Sullivan map to be 0. The image of the others are vectors in Rn" which are rotations of each other through different multiples of

E.

Here we see how our map captures some of the rotational symmetry in this aperiodic tiling.

(11)

CHAPTER 2. TILINGS AND TILING SPACES

Chapter

2

Tilings and Tiling Spaces

2.1

Cell Complexes

First let us solidify some terminology. Hereafter let

En =

{x

E Rn

I

1x1

5

1)

un

=

{x

ERn

I

1x1

<

1)

sn-1 - -

{x

E Rn

1

1x1

= 1)

ie, En is the closed unit ball in Rn, Un is its interior and Sn-' is its boundary. A CW-Complex is, roughly speaking, a space built up by the successive adjoining of cells of dimension O,1,2,

. . .

,

etc. To be more precise:

Definition 2.1.1 A CW-Complex o n a Hausdorf space

X

is defined by the prescription of an ascending sequence of closed subspaces

x0cx1cx2c

...

which satisfy the following conditions: ( 1 )

X0

c

X

has the discrete topology.

(12)

CHAPTER 2. TILINGS AND TILING SPACES 6

(2) For n

>

0, Xn is obtained from Xn-l by adjoining a collection {el)xEAn of disjoint sets homeomorphic to Un (called n-cells) such that for each X E A, there exists a continuous map

such that fx maps Un homeomorphically to e: and fx(Sn-')

c

Xn-l.

(3) X is the union of the Xi for i 2 0.

(4) The space X and the subspaces Xi all have the weak topology: A subset A is closed if and only if A

n

?i is closed for all n-cells, en, n = 0 , 1 , 2 .

. .

.

A CW-complex is also called a cell complex. We denote Kn to be the set {el)xEAn of n-cells adjoined t o the complex a t stage n. Also, will denote the closure of an n-cell while 6; will denote

3

- e l and will be called the boundary of el.' We say a CW-Complex is regular if it is a CW-Complex and we can choose each of our f A maps in part (2) of the definition to be homeomorphisms. If K,

#

0

but Ki =

Q)

for all i

>

n, then we say that the CW-Complex is n-dimensional. Also we say that, for two cells e;-l and e l , that e ~ - ' is a face of e l if e;-' C

E x a m p l e

-

If S2 = {(x, y, Z) E

R3

1

x2

+

y2

+

.z2 = I), then define

KO

= {(O,O, -I), (0,0,1))

{(sin r t

,

0, cos ~ t )

I

r E (0, I)), { (- sin r t ,0 , cos r t )

I

r E ( 0 , l ) )

(13)

CHAPTER 2. TILINGS AND TILING SPACES 7

S2

is the boundary of the unit sphere, so we chose vertices to be at points of intersection of

S2

with the z-axis, edges to be lines connecting the two vertices down opposite sides and our 2-cells to be the two determined half-shells. This defines a regular CW-complex on S2, as the edges are both homeomorphic to (0,l) with the homeomorphisms extending to their closures (ie, the two edges do not start and end at the same vertex). Also, the elements of K2 are both homeomorphic to the unit ball in R2, with the homeomorphisms extending to the boundaries which themselves are homeomorphic to S1.

2.2

Tilings

Consider Rn, usual n-dimensional Euclidean space. If A is a subset of Rn, we may translate it by a vector x E Rn,

We shall begin with a finite set {pl, p2, . . . pN) of subsets of

TW"

homeo- morphic to the closed unit ball, which we call prototiles. These prototiles may carry labels to distinguish them, ie, two prototiles may have the same shape but have different labels. We then say that a tile is any subset of Rn which is a translate of one of the pi. Then we define partial tiling and tiling as follows:

Definition 2.2.1 A partial tiling is a collection {tjIj, of subsets of Rn which are translates of prototiles with pairwise disjoint interiors. T h e sup-

port of a partial tiling is defined t o be the union of its tiles; this i s denoted supp(.). A tiling is a partial tiling whose support i s Rn. If T = {tjIjEJ is a tiling, a patch i n T is a subset of T with bounded support.

(14)

CHAPTER 2. TILINGS AND TILING SPACES 8

When n = 1, a tiling can be thought of as a bi-infinite sequence of a finite number of symbols, and when n = 2, it is what one normally thinks of as a tiling; that is, shapes fitting together to cover the plane. If T = { t j ) j , J is a tiling we can, for x E Rn, define the translation of T by x by T

+

x =

{ t j

+

x)j&J-

We also think of a tiling T as a multi-valued function: for u E Rn and U

c

Rn, let

T ( u ) = { t E T

I

u E

t )

T ( U ) =

U

' n u )

uEU

Tilings T and TI are said to agree on U if T ( U ) = T 1 ( U ) .

Definition 2.2.2 A tiling is said to be periodic i f there exists a non-zero x E Rn such that T = T

+

x . A tiling for which no such x exists is called aperiodic.

Periodic tilings are generally not very interesting, so we usually want our tiling to be aperiodic. Unless stated otherwise, tilings from here forward are assumed to be aperiodic.

2.3

The

Tiling Space

RT

If 7 is a collection of tilings, then we can define a metric on I. If T , T' E

7

with T =

{ t j j j E J

and TI = {t:)i,I

,

then define

d ( T , T') = inf{l, E 13 x , X I E Rn 3 1x1

,

lxll

<

E,

(15)

CHAPTER 2. TILINGS AND TILING SPACES 9

This may look complicated, but it is really quite simple: two tilings are close if they agree up to a small translation of a large ball about the origin. To prove that this is a metric, we shall need a lemma.

Lemma 2.3.1 If a and b are positive numbers such that a

+

b

<

1, then

Proof Notice that since a, b, and a

+

b

5

1, we have that 1 - a ( a

+

b)

2

0, and so

0

<

1 - a ( a + b) =+ 0

<

b - a2b - ab2 + a < a - a 2 b - a b 2 + b

+

a

5

(a

+

b)(l - ab).

This implies the result. H

Proposition 2.3.1 d satisfies the conditions of a metric.

Proof That d is symmetric is clear, as the definition is symmetric in T and TI. d(T, T) = 0 for all T because we can always find arbitrarily large balls (and hence arbitrarily small E ) around the origin where T matches up with itself.

1 Conversely, if d(T, TI) = 0, then we must be able to find arbitrary large balls around the origin where T and TI agree up to arbitrarily small translation - this can only be true if T = TI. This d is also always non-negative as it is the inf of a set of positive numbers.

(16)

CHAPTER 2. TILINGS AND TILING SPACES

Now, let R , S and T be tilings. We need to show that

If d(T, R )

+

d ( R , S ) 2 1 then the equality holds because this d(.,.)

<

1 always. So assume d(T, R )

+

d ( R , S )

<

1 and pick E

>

0 small enough so

that d(T, R )

+

d(R, S )

+

E

<

1. Find X T R and xkR with

such that

Likewise, find X R S and xLs with

such that

Since T - X T R agrees with R - xkR on ( B 1 ( 0 ) ) , then we must have

d(T,R)++

that T - X T R - xkR agrees with

R

- xkR - xkR on ( B ~ ( ~ , ; ) + ~ (-xkR)). In a similar way we see that S - X S R - xkR agrees with R - xkR - xkR on

( B ~ ( ~ , A ) + $ ( - x ' ~ ) )

.

This means that S-xSR-xkR agrees with T - xTR-xkR

wherever these two balls overlap. This overlap includes the origin because

1

I-x;,~

,

I-xkR1

<

1 and the radii d(s,R)+Z and ~ ( T , R ) + I are both greater than 1. If r l and r2 denote the largest balls around the origin which are contained in (B d (, I R)+ ~ (-xbR)) and

( B

~ ( s , R ) + $ I (-xkR)) respectively, then

(17)

CHAPTER 2. TILINGS AND TILING SPACES

Now

1

The above lemma implies that rl

2

d(T,R)+ +d(s,R)+g

,

and a symmetric argument shows the same inequality for r2. Thus T - xTR - xkR and S - xSR -

xhR agree on B I (0). Since we have that 1 xTR

+

xkR

I

,

I

S

- X S R - xhR

I

I

d ( T , R ) + d ( S , R ) + e

d(T, R)

+

d(S, R)

+

E by the usual triangle inequality, we have that d(T, S)

5

d(T, R)

+

d(R, S)

+

E which proves the result. H

Thus any collection of tilings can be made into a metric space. One way to produce our collection

7

of tilings is to start with a specific tiling T and let 7 be the set of all translates of T , ie, 7 = T

+

Rn.

Definition 2.3.1 Let

T

be a tiling. Then we define QT to be the metric space obtained by completing T

+

Rn in the above metric.

Strictly speaking, SZT is a space of Cauchy sequences, but we can visualize its elements as tilings. For example, consider the tiling of R2 consisting of unit squares matching up edge-to-edge and vertex-to-vertex, (a checkerboard pattern) with the vertices on the Z2 lattice, except that imagine that the 4 squares centered at the origin are replaced with a single 2

x

2 square. The normal checkerboard tiling (call it C) can be identified with the Cauchy se- quence { T

+

(n, O)):?,

.

This line of thinking leads easily to the identification

(18)

CHAPTER 2. TILINGS AND TILING SPACES 12 When we take a tiling T and form the metric space QT, we might like to ' know whether QT possesses any nice topological properties. It is well known and shown in [AP] that StT is compact if T satisfies the following condition.

Definition 2.3.2 A tiling T is said t o have Finite Local Complexity if, for every

R

>

0 there are only finitely many patches (up to translation) in T

whose radii of their supports are less than

R.

2.4

Substitution and the Anderson-Putnarn

Complex

One of the difficulties encountered in the study of tilings is producing inter- esting examples. One method for doing this is called the substitution method. The substitution method starts with our usual set { p l , p z , . . . p N ) of a finite number of prototiles along with a rule for splitting each prototile into tiles which are smaller copies of the pi's along with an inflation constant X

>

0 which inflates the smaller copies to be the same size as the originals. The simplest example is to take a square with side length 1 and split it into 4 squares of side length one-half. If we then multiply this by X = 2, we end up with 4 copies of our original square. In general, the result of the procedure on pi is denoted w(pi) and it is a partial tiling with support Api. This rule can be extended to the translates of the pis by defining w(pi

+

x ) = w(pi)

+

Ax.

Thus, we can easily define w of a partial tiling - simply divide all the tiles in the patch up according to the rule and inflate everything. This clearly results in a new partial tiling whose support is X times the support of the old. This idea can be easily seen to extend to tilings, ie, w ( T ) is the tiling

(19)

CHAPTER 2. TILINGS AND TILING SPACES 13

obtained by dividing each tile in T according to the rule and then inflating everything - the result of this is also a tiling.

From here onward we shall be dealing with a tiling T, its tiling space OT, and a substitution rule w. We also make some assumptions for our substitution rule w. The first is that w maps flT to itself and that w is one- to-one; this is what's known as recognizability. It is a well known fact (proved in [S]) that if w is one-to-one, then

aT

contains no periodic tilings.

The second assumption is that the substitution is primitive, that is, there is an M E Z+ such that wM(pi) contains a translate of pj for all i , j = l , 2

,...,

N.

The third assumption is that w : QT + is onto. These assumptions lead to the following fact from [AP].

T h e o r e m 2.4.1 Under the hypotheses above,

(aT,

w) is a topological dynam- ical system, that is, w : StT +

aT

is bijective and bicontinuous.

From now on assume that all prototiles are polygons and in the substi- tution they meet vertex to vertex and edge to edge. We are now then ready produce what is known as the Anderson-Putnam complex [AP] for QT. This is started by constructing a Hausdorff space

ro

which is the quotient of the disjoint union of the prototiles obtained by gluing the prototiles together in all ways in which the substitution rule allows them to be adjacent. The inflation map w induces a continuous surjection yo on

ro,

and with respect to which we take the inverse limit to obtain a new space n o . We begin by defining

rk

for k = 0 , l . If t is a tile in a tiling T, we define T ( O ) ( ~ ) =

{ t )

and ~ ( ' ) ( t ) = T(t), that is, ~ ( ~ ) ( t ) is the set of tiles in T that are within k tiles of t. Consider the space

aT

x

Rn

with the product topology. Let--k be

(20)

CHAPTER 2. TILINGS AND TILING SPACES 14

the smallest equivalence relation on OT

x

Rn that takes (TI, ul), (T2, u2) to

(k) (k)

be equivalent whenever Tl (tl) - ul = T2 (t2) - 7.42 for some tiles ti E T ,

such that ui E ti. Now define

rk

= CIT

x

Rn/ - ~ k with the quotient topology. For a simple example, consider again the tiling of the plane by unit squares matching edge-to-edge. There is one prototile, and the identifica- tion described above leads to identifyting the top edge with the bottom edge and the left edge with the right edge. In this example, we see that

ro

is isomorphic to T, the 2-torus.

For the examples investigated in this thesis, the prototiles are going to be 2-cells, so let us see what we have constructed in this case. A point in OT

x

R2 is a tiling T together with a vector u in R2, so if we think of T covering R2, we can think of (T, u) as u E supp(T). If u lies in the interior of a tile

t ,

and t is the translate of a prototile pi, then the equivalence class (T, u)O is the set of all (TI, ul) such that the point ul on the tiling TI is in the interior of a tile tl - which is also a translate of pi - and lies at exactly the same place

u lies on t. Thus, for each prototile pi we can define Pi = {(T, u ) ~ ( T(u) is a translate of pi). The following will be stated without proof.

Claim 2.4.1 The Pi are 2-cells in a cell complex for r o .

To get the rest of this cell complex, imagine drawing all the prototiles, each with the pi label; these are the

Pi.

Next, label the edges in the natural way: start with any edge of any 2-cell, give it a label, then give the same label to any edge on the other 2-cells that may be adjacent to it in any tiling in StT;

do this again for all edges labeled so far until no new labelings can occur. Repeat this for the other edges, and then for the vertices. This defines a CW-complex on To.

(21)

CHAPTER 2. TILINGS AND TILING SPACES 15

To get a similar construction for

rl,

one simply has to start with more 2- cells in the complex. For each prototile, there will be several different 2-cells, each with a different label corresponding to different possible patterns of tiles around it (a tile with such a label will hereafter refered to as a collared tile). Theorem 2.4.2

If

T

has finite local complexity,

rk

is a compact Hausdorf space.

Proof Because T has finite local complexity, we can find an r

>

0 such that every possible pairwise adjacency of the prototiles is represented in the partial tiling T B,(O)

(

)

.

If (TI, ul) is an element of CIT

x Rn

it happens that either ul lies on in the interior of a tile in Tl or on the edge between two tiles. In either case, we can find u in B,(O) such that (T, u) (TI, ul) (in the first case because all prototiles are represented in T

( )

B,(O)

,

in the second case because of our pick of r ) . {T)

x

B,(O) is compact, and so

rk

is the image of a compact set under the quotient map 7rN : OT x Rn -,

rk,

and hence is compact (the quotient map is always continuous with respect to the quotient topology).

A

cell complex is always Hausdorff (see [Ma]), so we are done.

1

2.5

QT

as

an

Inverse

Limit

We aim to show that CIT is isomorphic to a space of sequences in elements of

rk

called an inverse limit; spaces similar to the solenoids discussed in [BS]. To construct this, we first need a surjection on

rk.

Theorem 2.5.1 The inflation map w induces a continuous surjection

(22)

CHAPTER 2. TILINGS AND TILING SPACES 16

Proof Let Tl and T2 be in RT, and assume that ( T I , u l )

-

(T2, u2). Thus

( k )

Tl ( t l ) - ul = ~ Z ( ~ ) ( t 2 ) - u2 for some ul E t 1 E Tl and u2 E t2 E T2, and so we must have that

tl

- ul = t2 - u2. Thus we can choose tiles t', and

t/2

with Xul E t i E w ( { t l ) ) and Xu2 E

t/2

E w ( { t 2 ) ) such that w ( ~ ~ ) ( ~ ) ( t i ) - Xul = ~ ( ~ 2 ) ( ~ ) ( t / 2 ) - XU> Thus, (w(Tl), Xul) N (w(T2), Xu2)

and so yk is well-defined. The map on ilT x Rn that sends ( S , u ) to (w

( S )

,

Xu) is co-ordinatewise continuous and hence continuous, so when we pass to the quotient we see that yk must be continuous. We have that w is invertible on flT, SO (w-'(S), X-lu)k maps to ( S , u ) . Thus, yk is onto.

We now construct the inverse limit space of

rk

with respect to yk. Define

This is a topological space with the relative topology from the product topol- ogy (ie, ilk C

nzl

rk).

Thus, a basis for the topology is the collection of sets of the form B;; = { x E ilk

I

xi E yEy;-"~); i = 1 , 2 , . . . , n ) , where

U

c

rk

is open and n E

N.

We can use the inflation map to define a right shift wk : Rk + ilk by wk(x)i = yk(xi). We can see that wk is invertible with inverse

wk1

( x ) ~ = xi+l.

Before the last theorem of this chapter, we need a standard dynamical definition and a definition of a condition due to Kellendonk.

Definition 2.5.1 Two topological dynamical systems ( X ,

f)

and ( Y , g ) are said to be topologically semi-conjugate if there exists a continuous sur- jection r : X + Y such that T o f = g o T . The systems are said to be

(23)

CHAPTER 2. TILINGS AND TILING SPACES 17

Definition 2.5.2 T h e substitution tiling space (aT, W ) is said to force its

border if there exists a filced positive integer N such that for any tile

t

and tilings TI and T2 in RT containing

t ,

we have that wN(T1)

This says that there is a number of iterations N of the inflation after which the tiles surrounding the image of a tile in two different tilings must be the same. This is always satisfied if the tiles we are dealing with are collared tiles (the.tiles surrounding collared tiles are known after each iteration of the substitution).

Theorem 2.5.2 Let T be a substitution tiling under a substitution rule w

which has recognizability, is primitive, and i s a n onto m a p from OT to itself. T h e n w k : Rk + Rk i s a homeomorphism, and thus (ilk, wk) is a topological

dynamical system. T h e dynamical systems (QT, W ) and (a1, wl) are topologi- cally conjugate. Furthermore, i f T forces its border, then ( a T , W) and

(ao,

wo) are topologically conjugate.

Proof We begin by showing that

(aT,

W) is conjugate to ( a l , wl), and then show that if the substitution forces its border that

(a1,

wl) is conjugate to

(ao,

wo). We must therefore find a homeomorphism between the two spaces that conjugates the actions. What is given below is a sketch of the construction and is also given in [AP].

For any TI E

aT,

define IT : flT -+ R1 by n(T1) = {xi)go where xi = (w-'(TI), 0) We have that yl (xi) = xi-1, so IT is well-defined. Let {xi)go be

any element of 01; we wish to find a tiling TI that maps to it under IT. Since we must have that xo = (TI, O)o, xo specifies the tile

to

in T' which contains

(24)

CHAPTER 2. TILINGS AND TILING SPACES 18 the origin. In the same sense, xl must specify the tile in w-'(T) that contains the origin - tl say. Thus T' contains the partial tiling w(tl). If we continue in this way, we obtain a nested sequence of partial tilings. To have that the limit of these is a tiling (our T'), we recall that in El we are dealing with

"collared tiles", so the tiles around any given patch are determined.

To see that n is one-to-one, suppose we have that n(Tl) = n(T2) for some

TI, T2 E ET. Define

where dist(U, V) is defined as inf {Ilu - v

11

1u E U, v E V) for any sets U, V

c

W. Finite Local Complexity of T implies that this is an inf over a finite set

of positive numbers, and is thus positive. Suppose v E Rn; we show that TI and T2 must agree on a ball around the origin containing v.

Let n E Z+ such that rXn

>

11~11. NOW 7r(Tl) = 7r(T2) as sequences in El, so n(T1), = n(T2),. We can see by finite induction that this reduces to saying that, for some tiles t l and t2 containing the origin, we have W - ~ ( T ~ ) ( ' ) ( ~ ~ ) = w - , ( ~ ~ ) ( l ) ( t ~ ) . Since we are in E l , tl and t2 are collared tiles. Thus, w-"(TI) and w-, (T2) agree at least on B, (0)

,

and hence TI and T2 agree on B,An (0). Since v E B,An (0), we must have that TI and T2 agree everywhere. Thus 7r is one-to-one.

To show that 7r is onto, suppose we have x =

{(Ti,

ui)

l)Z&

El. Define

It needs to be verified that this is a partial tiling, and, using the r defined two paragraphs above, that it is in fact a tiling. Then it is clear that T' E QT

(25)

CHAPTER 2. TILINGS AND TILING SPACES

and x(T') = x.

Bicontinuity is proven using standard methods, and it is easy to check that x o w = wl o 7r to entwine the dynamics. Thus (S1, w) is topologically

conjugate to (01, wl)

.

Now we assume that the substitution rule forces its border, and prove that (ill, wl) is topologically conjugate to

(00,

wo).

If (TI,

ul) N1 (T2, u2) then trivially (TI, ul) -0 (T2, u2), SO the natural map

f

((TI,

41)

= (TI, 4 0

is well-defined for all T' E RT and u E Rn. We clearly have that f is onto, but it is not in general one-to-one. Now

Thus we can define a well-defined map F : R1 + by F(x); = x; for i E

N+.

We claim F is a homeomorphism that conjugates wl and wo.

To show that F is injective, we need that the substitution forces its border. Suppose we have that F ( { x ; ) ~ ~ ) = F ( { x ; ) ~ ~ ) with xi =

(T,,

ui)l and y; =

(T,',

u:)l. We see that if vl E sl E S1 and vz E s 2 E S2 for S1 and S2 E QT,

(0) (0)

and (Sl,vl)o = (S2, v2)0, then

S1

(sl) - vl = S2 (s2) - v2 and the forcing the border condition implies that (wN (S1), XNvl)l = (wN (S2), XNv2)1. By

(26)

CHAPTER 2. TILINGS AND TILING SPACES 20

our hypothesis, F ( { ~ i ) ~ ~ ) = F ({xi)Eo) and in particular (Tj+N, uj+N)0 =

(T;+N, u ; + ~ ) o , by a finite induction we have that ( w N ( ~ j + ~ ) , XN~j+N)l =

( W ~ ( T ~ + ~ ) , X ~ U ; + ~ ) ~ . We know that = W - ~ ( T ) and that U ~ + N =

X-Nuj, SO

(q,

uj)1 = (Tj, u;) and F is one-to-one.

The compactness of OT implies that

F

is onto; the following argument is due to Kellendonk. Say we are given

{(Ti,

ui))zo E Eo. Because

ET

is com- pact, the sequence { W ~ ( T ~ - U , ) ) ~ = ~ has a convergent subsequence { w n k (Tnk -

u n k ) ) g o that converges to some tiling T' E ET. Now, {(w-~(T), O)l)zo is in

El,

and we claim that it maps to {(Ti, ui))go under F. Since (To, ~ 0= ) ~

(wn(Tn), Xnun)0 and wn(Tn - u,) = wn(Tn) - Xnun, we have that (wn(Tn - u,), O)o = (To, 2 ~ ~ ) ~ for a11 n. Because T has Finite Local Complexity, this means that between all the tilings (wn(Tn - u,), only finitely many different

tiles contain the origin. \

Pick an i

>

0. Our subsequence being convergent means that given any

E

>

0, we can find a large enough k so that wnk(Tnk - unk) agrees with T' on

Bt (0). Let

R

be a real number greater than the diameter of each prototile, and chose k so that n k

2

i and wnk(Tnk - unk) agrees with TI on BXiR(0).

nk-i

Then (w-~(T'),O)O = (wnk-'(Tnk -unk),O)o = yo (Tnk,unk)o = (Tijui)~,

which is what we were claiming.

We have easily that F is bicontinuous, and a simple calculation similar to the one above shows that it conjugates the actions of wo and wl. Thus (Eo, wo) is topologically conjugate to (a1, wl) when the substitution forces its border, and since topological conjugacy is an equivalence relation, (Go, wo) is conjugate t o (OT, w). H

(27)

CHAPTER 2. TILINGS AND TILING SPACES 21

Conjugate dynamical systems have the same dynamics and can be thought of as, in a sense, the same system. The elements of

nk

can be seen as tilings in the following way. If {xi)go E i l k , then xo = (TI, 0) for some tiling TI. Furthermore, xo is equivalent to all (S, O)k such that S has the same tile to around the origin as TI - so we can see that the first co-ordinate specifies the tile at the origin. Then by extending this idea to xl = (TI, O)k, we see that xl specifies a patch w(tl) around the origin, with the tile to

c

w(tl). In the case

k

= 1 (or k=O if T forces its border), we see that the limit of this process does indeed specify a tiling in flT.

(28)

CHAPTER 3. COHOMOLOGY

Chapter

3

Cohomology

Cohomology theories are ways to obtain important invariants of a space. The three we will talk about here concern topological spaces.

3.1

Cohomology in General

In general, when we talk about cohomology we mean the following. Let X = {XI, X 2 , . . . } be a sequence of spaces. Define

c i ( x , G) = {f : Xi -+ G).

For some abelian group G. We call Ci(X, G) the group of i-cochains. Suppose we have a sequence of maps di such that

such that hi+, o

bi

= 0 for all i. Then ker di is called the group of i-cocycles while the Imdi-l is called the group of i-coboundaries. The i - t h cohomology

(29)

CHAPTER 3. COHOMOLOGY

group of X with coeficients in G is then defined to be Hi (X, G) = ker

hi/

Im

Elements of Hi(X, G) are still refered to as i-cocycles or merely cocycles. With this, we can define three types of cohomology relevant to tilings.

3.2

Orientation, Incidence Number, and Cel-

lular Cohomology

In a cell complex, orientations are needed on all the cells to define cellular cohomology. If we restrict our attention to 1, 2, and 3 cells for the moment,

it's easy to guess what the orientation of cells would look like: a left or right arrow on a 1-cell, a clockwise or counter-clockwise curl on a 2-cell, or a left or right handed corkscrew in a 3-cell. In cohomology (and homology) theory, we need a way of expressing whether the orientations of cells and their faces

"match up" - this is done with incidence numbers.

In Figure 3.1 the arrows indicate the orientations given to a and the edges e, f and g. If we go around the cell according t o the orientation of a, then we see that the orientations of e and

f

match up t o that of 0, while that of

g does not. In this situation, we would like to define incidence numbers of the pairs (a, e) and (o,

f)

to be +1 and the incidence number of (a, g) to be -1. If e; and eE-' are n and n - 1 cells respectively, then we denote their incidence number by [ei : e g l ] . Note that [e; : e:-'] is defined for arbitrary n and n

-

I cells, but is zero if e:-'

3.

(30)

CHAPTER

3.

COHOMOLOGY

Y

Figure 3.1: A typical 2-cell

Example: Looking at Figure 3.1, we have already established that:

[a : e] = 1

[ a : f ] = 1

[a : g] = -1

In addition t o these, we must relate edges to vertices. If IT is an edge and x is a vertex of the edge, then we define [IT : x] to be

f

1; 1 when the edge points toward the vertex and -1 when it points away. Thus,

[f

:

PI

= 1 [f : 71 = -1 [e : a] = 1 [e :

P]

= -1 [g : a] = 1 [g : y] = -1 [f : a] = 0, etc.

(31)

CHAPTER

3.

COHOMOLOGY

25

It is, in fact, possible to define orientation rigorously to apply to arbitrary dimensions. For the purposes of the spaces presented here, the highest di- mension we will have to deal with is 2, and so our discussion of orientation and incidence will end with the above paragraph.l

We are now ready to define the Cellular Cohomology of a cell complex K. For each i, let

F ( K i , G )

={f

I

f

: K i + G )

where G is an abelian group (in this paper, G will always be either R or Z - for this reason we will often refer to group elements as "numbers"). If K is n-dimensional, let F(K;, G) be the zero group for all i

>

n. Define

A minor problem that arises from this definition is the sum - it may not be finite. There are a couple ways ways to fix this - the first being to define F(Ki, G) to be the finitely supported functions on Ki. This works fine, al- though it is rarely necessary. The other is to impose a mild restriction on our CW-complex stating that, for any n-cell e;, we have that e:-'

c

for only finitely many p. This would make [e\" : e:] = 0 for all but finitely many p.

lIn short, the orientation of a cell e z is derived from a group called the n t h relative homology group of

3

with respect t o d ( e T ) , denoted H , ( g , d(e?)). This group is always

infinite cyclic, and the orientation of eT is defined to be the choice of its generator. A full

(32)

CHAPTER 3. COHOMOLOGY 26

Example: Looking again at our typical 2-cell, Figure 3.1, we see that if

$ E F(Ko, G), we have

We can see that if cp E F ( K l , G), we can think of dl acting on it by taking it to the function that takes a 2-cell and produces a number by adding up the values of cp on the edges multiplied by the respective incidence numbers.

With these sets and maps, we get a chain complex.

"

3

F(Ki, G)

3

F(Ki+l, G )

3

. . .

o

-

F ( K ~ , G ) F ( K ~ , G )

-

...

If we view each F(Ki, G) as a group with the usual addition, then the 8's are all homomorphisms. In addition, the 8 s are related in a very simple way.

Proof Without a definition of orientation and incidence numbers for higher dimension at our disposal, proving this for i

>

2 is impossible. We will prove this for the case i = 0; the others are done in a similar way with when equipped with proper definition of orientation for higher dimensions. Let f E F(Ko, G )

.

We need to show that for all 2-cells e;, we have a

Suppose that e? is surrounded by edges el, ez,

. .

.

,

en and let t(e) and i(e) denote the terminus and initial point of an edge e, respectively. Then

(33)

CHAPTER 3. COHOMOLOGY

But t(ek) = i(ek

+

I), and t (en) = i(el), so the sum collapses,

8 0 Me:) =

-f

(w)

+

[f

(t(e1)) -

f

(w)]

+

[f

(w)

-

f

(w)]

+

. .

*

+

[f

(t(en - 1)) -

f

(i(en))]

+

f

(t(en))

= O

.

( 3 4 Now we can form the cohomology.

Definition 3.2.1 Let K = Uz,Ki be a CW-Complex. Define the ith Cel- lular Cohomology Group of

K ,

denoted Hi(K, G), to be

H ~ ( K ) := ker

&/

Im

Note that this is well defined, as ker

ai

and Im are both groups (the 6"s are homomorphisms) and Imdi-l

c

ker di by Claim 1.2.

3.3

~ e c h

Cohornology

Using the notation from [BT], let X be a topological space, and let U = {Ua)aEJ be an open cover for X , where J is a countable linearly ordered index set. For a

<

b

<

c, denote the pairwise intersections Ua

n

Ub by Uab, triple intersections Ua

n

Ub

n

U, by Uabc etc. Let

u(")

denote the set

(34)

CHAPTER 3. COHOMOLOGY 28 of n-fold intersections of elements of U (0-fold intersections are just the sets themselves, 1-fold intersections are intersections of the form Ua

n

Ub with a

<

b etc). Let

F(u("), G) n E

N

denote the group of functions on the set of n-fold intersections of elements of

U

taking values in the abelian group G. By the 0-fold intersections we mean just the sets themselves. Define boundary maps

&

by

Then, as before, &+18i = 0 and we can form the cohomology of the complex

We denote these groups

kZ

(u,

G) = ker

&/

Im

and call these the ~ e c h Cohomology of the cover

U.

A priori, these groups depend on the cover U. In this regard, we are rescued by a definition and a theorem from [BT]

.

Definition 3.3.1 A good cover for a topological space X is a n open cover of X for which each finite intersection is contractible.

Theorem 3.3.1 IfU and 23 are good covers for a space X , then *(u, G) E

I?(%?,

G) for all i.

(35)

CHAPTER 3. COHOMOLOGY 29 To define the ~ e c h cohomology of a space, we should want a definition which is independent of the cover, good or otherwise. To do this, we follow the lead of [BT] and define the following.

Definition 3.3.2 Let

U

= {Ucu)crEI and

2l

= {Vp)pEJ be open covers of a space X . T h e n we say 'D is a refinement of

U,

written

U

<

'D if there is a map

4

: J --, I such that Vp

c

U+(p).

If

U

<

523, then the map

4

induces a map on Cohomology, via

Q>f

(Va1a2 ... a,) =

f

(Ud(al)d(az)...d(ak)).

It is possible that the indices in the last term are not in the correct order. To deal with this, we adopt the convention to take f (U,(a,),(a,)..,,(ak)) =

sgn(a) f (Uala2,..ak) for any permutation o.

Definition 3.3.3 (BT) A direct system of groups is a collection {Gi)iEI of groups indexed by a directed set I such that for any pair a

<

b there is a group homomorphism f; : G, -, G b satisfying

f," = i d e n t i t y

Whenever we have a direct system of groups, we can form is direct limit. Definition 3.3.4 Let U G i denote the disjoint union of the direct system of groups {Gi)iEI. Introduce an equivalence relation o n

U

Gi by saying that

(36)

CHAPTER 3. COHOMOLOGY 30 g, E G, is equivalent to gb E Gb if for some upper bound c of a and b we have f;(g,) = f,b(gb) in G,. The direct limit of the system, denoted by limiEr Gi,

is the quotient of

U

Gi by this equivalence relation.

Thus, two elements in

U

Gi are equivalent if they are "eventually equal". The direct limit is a group under the operation [g,]

+

[gb] = [f;(ga)

+

f,b(gb)],

where c is an upper bound for a and b and the brackets indicate equivalence classes.

From all this we can see that for each k, {Hk(U, G))u is a direct system of groups.

Definition 3.3.5 The ~ e c h Cohomology of a space X is defined as the direct limit

H ~ ( x , G) = lim H ~ ( U , G)

LL

where the limit is over a directed set of refinements.

3.4

Dynamical Cohomology

Now, take (X, cp) to be a topological Rn-dynamical system, ie, let X be a compact metric Space and cp be a continuous Rn action on X. This just means that for each v E Rn, cp, : X -+ X is a homeomorphism of X and the map sending (x, v) t o v H cp,(x) is jointly continuous; we also have

pv o cpw = %+, for all v, w E Rn. Let C ( X ) denote the algebra of continuous R-valued functions on X . We call f E C ( X ) continuously diflerentible if

O f

- (x) = lim

f

(c~tv(4) -

f

(4

(37)

CHAPTER 3. COHOMOLOGY 31

exists and is back in C ( X ) for all x E X and v E Rd. We say f is smooth if it is infinitely continuously differentiable, and let C m ( X ) denote the set of such functions. We can also take the same approach for a finite dimensional vector space; we let C(X, W) denote the continuous W-valued functions on X . The definition of Cm(X, W) extends naturally.

Let { x l , x2,

. . .

x,) denote the standard basis for Rn, and let Rn* denote the dual space of the real vector space Rn. Then we can always find a basis for {dxl, dx2, . . .

,

dx,) of Rn* such that (xi, d x j ) = Sij where Sij is the Kronecker delta symbol (Sij = 1 if i = j, but = 0 otherwise). This is called the dual basis for Rn* with respect to { x l , 5 2 ,

. . .

2,). Then we make the following

definition (see [BT])

.

Definition 3.4.1 The graded exterior algebra of Rn* is the algebra over R generated by dxl, dx2,

. . . ,

dxnwith the relations

We denote this algebra by ARn*.

Thus ARn*, when viewed as a real vector space, has basis (for 1

5

i

<

j < k I n )

1, dxi, dxidxj, dxidxjdxk,.

.,

dxldx2

. .

.

dxn

We also let AkRn* denote the subspace of ARn* spanned by elements of the form dxi, dxi,

- - -

dxi, for 1

I

il

<

i2

<

-

. . <

ik

5

n.

(38)

CHAPTER 3. COHOMOLOGY

for all v E Rn. This extends to a differential

in the following way: every element of CCm(X, AkRn*) may be written in the form

where I = {il, i2,.

.

.

ik) C {1,2,.

. .

n } and

fl

E Cm(X, R). Furthermore, we have the following relations;

"

O f

-

8

= lim

f

((~tzi (x)) -

f

(x)

df =

C

-d%

axi

t40

i=l axi

t

Proposition 3.4.1 Let d, denote the map di = d : Cm (X, A W * ) -+ CCm (X, Ai+lRn*) T h e n di+l o di = 0

Vi.

Proof See [BT]. Thus, the chain

d

. .

.

4,

c C m ( x ,

nk-lwn*)

5

cyx,

nkw*)

4,

cCm(x,

h k + l w * )

+

. . .

has the property that the image of the d map is contained in the kernel of the previous d map. Thus, we can form the following.

Definition 3.4.2 We define the dynamical cohomology of (X, p ) t o be the groups

(39)

CHAPTER 4. CONNECTING COHOMOLOGY THEORIES 33

Chapter

4

Connecting Cohomology

Theories

4.1

Mapping Cellular Cohomology of

ro

to

the ~ e c h

Cohomology of OT when n

=

2

To construct the Cech Cohomology of our tiling space flT, we must first produce an open cover.

Let pl,p2,.

. .

, p ~ be our N distinct prototiles in tiling T which generates our tiling space StT. Assuming that each of the pi is star-shaped (that is, for each i there exists a point xi E pi such that for any other point y E pi, the line from xi to y is contained in pi) pick a point in the interior of each tile

'

about which it is star-shaped. Next, pick a point in the interior of each edge in the edge set. Because the tiles are star-shaped, these can be connected to the points in the interior of the tile on each tile by a straight line (see Figure

(40)

CHAPTER 4. CONNECTING COHOMOLOGY THEORIES - 34

4.1). This splits the tile into regions, each containing exactly one vertex. For each vertex v around pi we denote its corresponding region by r,. Let

O ( P , , ~ ) =

(

U

B&))

n

pi

X E T ,

for some sufficiently small E . If t is a tile, then t = pi

+

x for some i and

x,

Figure 4.1: O(pi, V)

so define O(t, v) = 0 (pi, v - x)

+

x. Now define, for any tiling T and vertex v in any tile of T ,

O(v) =

U

O(t, v).

v E t

This is an open set in supp(T) =

R2.

Now, for all vertices v in T, look at the sets T(v) - v; we call such a set a vertex pattern. Because of the Finite Local Complexity property, for each v, T(v) - v = T(vi) - vi for some finite

(41)

CHAPTER 4. CONNECTING COHOMOLOGY THEORIES 35

set of vertices v1, v2,

.

.

.

,

V L for which T ( v i ) - vi

#

T ( v j ) - vj if i

#

j , and so

For each i = 1,.

.

.

,

L, define

v,

= {TI : T 1 ( 0 ) = T ( v i ) - v i ) - (O(vi)

-

vi)

= {T'

+

x : T' (0) = T ( v i ) - vi; x E O(vi) - vi)

The set V , consists of all tilings which are translates of tilings which have the pattern T ( v i ) - vi at the origin translated by vectors which keep the origin in r v i . Now we can define

V is our open cover.

As stated earlier, we let

rk

denote the Anderson-Putnam complex of the tiling space OT, and we can also let

rki

denote the i-cells of said complex. We thus get the usual cellular chain complex

0 + F ( r k o ,

z)

5

F(rk1,

z)

5

F(rk2, Z ) + 0

where F

(rk*,

Z ) denotes the set of integer-valued functions on

rk*.

Refering to Figure 1, the first boundary map is defined as

(42)

CHAPTER 4. CONNECTING COHOMOLOGY THEORIES 36

where the sum is over all i - 1 cells.

We now wish to produce a map from the cellular cohomology H1 ( r k ) to the Cech cohomology kl(l;t). This is accomplished by producing 2 maps,

such that the following diagram commutes

where the 8's in the bottom row indicate the usual Cech chain complex. In addition to wanting this diagram t o commute, we also wish to have the kernel of the 8 map from F ( r k l , Z ) to F ( r k z , Z ) to map into the kernel of the

8

map from F(v('), Z) t o F ( v ( ~ ) , Z). This will ensure that the a,'s can be translated to a map cr between H1(l?k) and H ~ R ) .

We will first define a o . It's easy to see that we have a map from V to

rko

- for every vertex pattern look at the vertex at the center of it, and then map

it to that vertex in r k o . We can then take cro to be the dual of this map. Next, we need to decide what al is. Since n = 2, we can assign to the 2-cells arbitrary orientation, so we pick them all to have the same orienta- tion, say clockwise. Suppose we take f E

F(rkl,

Z)

.

Then we want a1 f to be defined on two-fold intersections of our vertex patterns. If a two-fold intersection of vertex patterns is non-empty, then the vertices at the middle

(43)

CHAPTER 4. CONNECTING COHOMOLOGY THEORIES 37 of them must lie on the outside of a common tile t in both vertex patterns. So we define

where the sum is over the edges starting from the vertex at the middle of pattern a to the vertex at the middle of pattern b according to the orientation of the cell. If these two vertices lie on the outside of two different tiles tl and t2, ie, they lie at the beginning and end of an edge which connects tl and t2, then this is not well defined. Since we said earlier that all the tiles must have the same orientation, then if e is the edge in question, [tl : e] must be either plus or minus 1, with [t2 : e] = -[tl : el ([t : el denotes the incidence number of t with respect to edge e). To make our map well-defined, we chose to sum around the tile which has positive orientation number with respect to e.

Claim 4.1.1 The six-term diagram above commutes. In addition, al (ker 8 )

c

ker

8.

(44)

CHAPTER 4. CONNECTING COHOMOLOGY THEORIES 38

where t(e) denotes the terminal point of an edge e and i(e) denotes the initial point of e. The last step follows by collapsing the sum. Notice that if v, and vb were on the same edge that we end up with the same answer, as we pick the 2-cell with which the edge has positive orientation. On the other hand, we have

Now to show that al(ker

8 )

C ker

8.

Take

f

E ker

8.

Then, if a E

rk2,

where the sum is over all edges ei around the tile a. f is in the kernel, so the sum must be zero. Thus f must sum to zero around all tiles a. Thus, if V,

and Vb share an edge, summing around either tile that the edge is a part of will give values negative t o each other, so that multiplying by the incidence number makes them equal. In other words, if V, and Vb share and edge e with the edge adjacent to two tiles t l and t Z , then

where it is understood that one of the sums is a sing ;le term. Now, if V,, 1 and V, are vertex patterns with non-empty intersection,

(45)

CHAPTER 4. CONNECTING COHOMOLOGY THEORIES 39

Figure 4.2: Two Possible Arrangements of 3 Vertex Patterns with Non-empty Intersection

=

Cb

: elf

(4

-

Cb

: elf ( 4

+

C[o

: elf

(4

The figures above show the 2 ways in which 3 vertex patterns could be

arranged around a tile, with a

<

b

<

c. In the first case, we see that we sum around once, as we sum from a to b, then from b to c, then from c to a. In the second, we sum around twice. In either case, our value is an integer multiplied by

xi[o

: ei] f (ei), and this is zero because f E ker 8. Thus cul(ker 8)

c

ker

8.

This proves the claim. W

(46)

CHAPTER 4. CONNECTING COHOMOLOGY THEORIES 40

4.2

Mapping ~ e c h

Cohomology to Dynami-

cal Cohomology to

A R n *

We now describe the mapping ~ e c h cohomology to dynamical cohomology and from dynamical cohomology to AIRn* in the case of a general topological dynamical system - the following section mentions nothing about tilings. Let (X, cp) be a topological Wn-Dynamical System. Then let p be an invariant probability measure on X (these always exist by [GI]). Recall that we can form the dynamical cohomology groups

H~(R", C" (X, W)) = ker di/ Im di-l.

where di : C"

(X,

AiRn*) -, C" (X, Ai+'Wn* ). From this we can define the following map.

Definition 4.2.1 T h e Ruelle-Sullivan current C, associated with p is the linear map

(C,, .) : Cm(X,

nkWn*)

-+ hkWn* defined by

Lemma 4.2.1 Let p be a n invariant probability measure for the action cp. Let f be any cp-smooth function i n CW(X, AkRn*) for some k. Then

(47)

CHAPTER

4.

CONNECTING COHOMOLOGY THEORIES

41

for i = 1,2,

. . .

,

n . Since any invariant measure is the weak-* limit of a convex combination of ergodic measures, we can assume p is ergodic (see

0).

By the Birkhoff ergodic theorem, for almost all x in X we have

= lim

-

where X indicates Lebesgue measure. We have

where dui indicates that dui is omitted. This means that

Ilf

ll,

IS,

(x)dp(x)

I 1.-

5

lim - = 0.

R

I

This proves the result. 1

Since C, is zero on the image of the d maps, it extends to a map on cohomology :

-

:H"(IR",C"(X,E%)) ---,AIRn*

r v > ~

Thus, we have found a map from the dynamical cohomology to AIRn* for any (X, ip). The next step is then to find homomorphism from ~ e c h cohomology to dynamical cohomology, and to compose these into a map from the ~ e c h cohomology to AIRn*, which is called the Ruelle-Sullivan map.

We make the following definition:

Definition 4.2.2 Let

U

= {Ui)i,I be a finite open cover of X . A partition of unity subordiante t o

U

i s a set of positive-valued functions {pi) on X

such that

zi

pi(x) = 1 for all x E X and the support of pi i s contained in

(48)

CHAPTER 4. CONNECTING COHOMOLOGY THEORIES 42 As before, for io

<

il

<

-

.

.

<

ij, we let UiOil...i3 = UiO

n

Uil

n

. . .

n

Uij.

, For any open set U, we note that even if U is not invariant under cp, we have

that, for any continuous function f on U and

v

E Rn, the expression still makes sense as a function on

U

(assuming the limit exists). Thus, we can define Cm(U, AIRY*) as the smooth functions on U with values in ARn*.

Those familiar with the ~ e c h - d e ~ h a m theorem for manifolds and its proof will see how the following is an adaptation of the argument there to our case where X may not be a manifold. It involves constructing a double complex with appropriate maps to show that we can map the cohomologies to each other.

So, we define the double complex

where we always sum only over io

<

il

<

- .

<

i j with nonempty Uioil...ij. If

f

E Kj>'(U), then we denote the ioil

. . .

ij component by fioi, ...,. Also, for notational purposes, we want to set fioil...ij = 0 if Uioil...ij =

0

and fa(io)u(ill...u(ij) =sgn(a) fi,i, ...ij, for any permutation 0.

(49)

CHAPTER 4. CONNECTING COHOMOLOGY THEORIES 43

where

with

&

indicating that il is omitted in the index.

Lemma 4.2.2 The two diflerentials,

8

and d , commute. Proof. Let f E K ~ $ ~ ( u ) . Then for any io

<

il

< - - <

ij+l

(50)

CHAPTER 4. CONNECTING COHOMOLOGY THEORIES 44 While we also have

The above definition of partial differentiation is linear, so we see by exchang- ing the order of summation that the two are equal.

We now want to define a new map on our complex using the partition of unity found above, {pi);. For j

>

0, we define

Lemma 4.2.3 h8

+

8h = 1

(51)

CHAPTER 4. CONNECTING COHOMOLOGY THEORIES 45

Adding the two sees (8hf)i0i1...b cancel with the similar term in the first equation. Thus

that is to say, (ha

+

ah) ( f ) =

f

.

Lemma 4.2.4 The restriction map r defined by

r : C" ( X , A~R"') -+ K O ~ ~ ( U )

(52)

CHAPTER 4. CONNECTING COHOMOLOGY THEORIES Proof. Take f E Cm(X, AkRn*). Then, for any i , we have

Since the restriction of a sum is the sum of the restrictions, and partial

-

differentiation is linear, these are equal.

If f E KOlk(U), f has a unique pre-image under if r when

fi

= f j on Uij

whenever

Ui,

is non-empty. Every f in ker

8

c

KOT~(U) satisfies this, SO we can define r-' on the ker 8's by

where Ui is any member of the cover that contains x. It is easy to see that the map r is bijective onto ker 8 and that r-' its inverse.

Corollary 4.2.1 For any k the following sequence is exact;

(U)

8

a

K~,*(u)

2

~ j + l , k +

. .

.

0 -+

cyx,

hk%tn*)

li

K O I ~ ( U ) +

. . .

and thus has trivial cohomology.

Referenties

GERELATEERDE DOCUMENTEN

In both forementioned cases (L.P. In statistical computation, however, the big source of error is truncation accumulated while forming a so-called matrix of sums of squares and

To a stirred solution of the corresponding compounds 2 (1.0 mmol) in EtOH-water (1:1), LiOH (10.0 mmol) was added and the reaction mixture refluxed overnight.. Then, pH was adjusted

Table 5: Average results of runs of the algorithms in many random generated networks where the new exact method failed to solve the problem (consequently SamIam also failed, as it

Perhaps a more feasible solution lies in adding a type of annotated corpus to corpora already being used by the Translation Machines, and, for this particular case regarding

In het opgravingsareaal dat gelegen is binnen de hoek gevormd door de Reumestraat en de Amandusstraat, troffen we restanten van houtbouw aan (fig.. Een eerste groep

bodemval komen. De aangetroffen soorten moe- ten in de lokale vegetatie geleefd hebben en na hun dood op de bodem van het ven zijn terecht- gekomen. Uiteindelijk raakten ze in

Early versions of the implementation of both sweep-tiling (see 4) and sweep- and-prune (see 3) would sometimes not correctly handle events that occur nearly simultaneously. When a

Hence, it can be concluded that, thermally induced period changes in La/B multilayers are a result of two competing processes, namely, crystallization of lanthanum boride