16.01.2013 Views

DOMINANCE AND REGULARITY IN COXETER GROUPS A ...

DOMINANCE AND REGULARITY IN COXETER GROUPS A ...

DOMINANCE AND REGULARITY IN COXETER GROUPS A ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>DOM<strong>IN</strong>ANCE</strong> <strong>AND</strong> <strong>REGULARITY</strong> <strong>IN</strong> <strong>COXETER</strong> <strong>GROUPS</strong><br />

A Dissertation<br />

Submitted to the Graduate School<br />

of the University of Notre Dame<br />

in Partial Fulfillment of the Requirements<br />

for the Degree of<br />

Doctor of Philosophy<br />

by<br />

Tom Edgar<br />

Graduate Program in Mathematics<br />

Notre Dame, Indiana<br />

July 2009<br />

Matthew Dyer, Director


<strong>DOM<strong>IN</strong>ANCE</strong> <strong>AND</strong> <strong>REGULARITY</strong> <strong>IN</strong> <strong>COXETER</strong> <strong>GROUPS</strong><br />

Abstract<br />

by<br />

Tom Edgar<br />

The main purpose of this thesis is to generalize many of the results of Dyer,<br />

[18]. Dyer introduces a large family of finite state automata that are demonstrated<br />

to recognize many natural subsets of Coxeter groups. In the current work, we<br />

consider generalized length functions in Coxeter systems and show that these<br />

length functions lead to a larger family of finite state automata, which in turn<br />

recognize a larger class of natural subsets of Coxeter groups. By a well-known<br />

result, these subsets (called regular subsets) will then have a rational multivariate<br />

Poincaré series.<br />

Dominance order on the set of reflections (or positive roots) of a Coxeter system<br />

plays a major role in the study of regular subsets. We draw the connection between<br />

dominance in root systems and a notion of closure in root systems. We describe<br />

some conjectures, due to Dyer, [17], and introduce a conjectural normal form for<br />

Coxeter group elements related to closure and dominance.<br />

In [14], Dyer introduced the notion of the imaginary cone to characterize domi-<br />

nance. We investigate the imaginary cone in the special case of hyperbolic Coxeter<br />

systems. We show that any infinite, irreducible, non-affine Coxeter system has<br />

universal reflection subgroups of arbitrarily large rank. This allows us to deduce<br />

some consequences about the growth type of Coxeter systems.


For Laura, Sue, Kristin, and Mark<br />

ii


CONTENTS<br />

FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vi<br />

ACKNOWLEDGMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . vii<br />

SYMBOLS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix<br />

CHAPTER 1: <strong>IN</strong>TRODUCTION . . . . . . . . . . . . . . . . . . . . . . . 1<br />

CHAPTER 2: BACKGROUND MATERIAL . . . . . . . . . . . . . . . . 4<br />

2.1 Coxeter Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4<br />

2.2 Root Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5<br />

2.3 Reflection Subgroups . . . . . . . . . . . . . . . . . . . . . . . . . 6<br />

2.4 The Weak Order on T and the Set of Elementary Reflections . . . 7<br />

2.5 Dominance Order on T . . . . . . . . . . . . . . . . . . . . . . . . 9<br />

CHAPTER 3: GENERALIZED LENGTH FUNCTIONS . . . . . . . . . . 13<br />

3.1 Commutative Monoids . . . . . . . . . . . . . . . . . . . . . . . . 14<br />

3.2 A General Length for Coxeter Systems . . . . . . . . . . . . . . . 15<br />

3.3 Infinite Heights and Generalized Partitions for Reflections . . . . 18<br />

3.4 Infinite Height and the Weak Order . . . . . . . . . . . . . . . . . 19<br />

3.5 The Root Poset . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20<br />

3.6 Monotonicity Properties in Dominance Order on T . . . . . . . . 22<br />

3.7 A Partition of the Reflections . . . . . . . . . . . . . . . . . . . . 25<br />

3.8 Recurrence Formulae . . . . . . . . . . . . . . . . . . . . . . . . . 25<br />

3.9 Monoids with Poset Structure . . . . . . . . . . . . . . . . . . . . 26<br />

3.10 General Length and the Reflection Cocycle . . . . . . . . . . . . . 29<br />

CHAPTER 4: F<strong>IN</strong>ITE STATE AUTOMATA . . . . . . . . . . . . . . . . 32<br />

4.1 Regular Languages . . . . . . . . . . . . . . . . . . . . . . . . . . 32<br />

4.2 Automata . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34<br />

4.3 Canonical Automata for Coxeter Systems . . . . . . . . . . . . . . 35<br />

iii


4.4 Some Subsets of S ∗ . . . . . . . . . . . . . . . . . . . . . . . . . . 37<br />

4.5 The Boolean Algebra of Regular Sets of S ∗ . . . . . . . . . . . . . 39<br />

4.6 The Reflection Cocycle and Automata . . . . . . . . . . . . . . . 42<br />

CHAPTER 5: REGULAR SUBSETS OF <strong>COXETER</strong> <strong>GROUPS</strong> . . . . . . 44<br />

5.1 Boolean Algebra of Regular Sets in W . . . . . . . . . . . . . . . 44<br />

5.2 Coxeter Groups are Automatic . . . . . . . . . . . . . . . . . . . . 45<br />

5.3 Different Types of Regularity . . . . . . . . . . . . . . . . . . . . 45<br />

5.4 p-Regularity for Different p . . . . . . . . . . . . . . . . . . . . . . 47<br />

5.5 Some Properties of p-Completely Regular Sets . . . . . . . . . . . 48<br />

5.6 Poincaré Series and Examples of Regular Subsets of W . . . . . . 50<br />

5.7 More Regular Subsets of W . . . . . . . . . . . . . . . . . . . . . 52<br />

5.8 Regularity in Direct Products . . . . . . . . . . . . . . . . . . . . 56<br />

5.9 Product Regularity . . . . . . . . . . . . . . . . . . . . . . . . . . 57<br />

5.10 Multivariate Product Regularity . . . . . . . . . . . . . . . . . . . 58<br />

CHAPTER 6: <strong>REGULARITY</strong> <strong>IN</strong> T <strong>AND</strong> MIXED PRODUCTS . . . . . 63<br />

6.1 Regularity of sets of Reflections . . . . . . . . . . . . . . . . . . . 63<br />

6.2 Types of Regularity in T . . . . . . . . . . . . . . . . . . . . . . . 65<br />

6.3 p-Complete Regularity in T . . . . . . . . . . . . . . . . . . . . . 66<br />

6.4 Maps of Boolean Algebras . . . . . . . . . . . . . . . . . . . . . . 72<br />

6.5 Regularity in Mixed Products . . . . . . . . . . . . . . . . . . . . 79<br />

6.6 General Regularity in Mixed Products . . . . . . . . . . . . . . . 80<br />

6.7 Multivariate Poincaré Series . . . . . . . . . . . . . . . . . . . . . 82<br />

CHAPTER 7: HECKE ALGEBRA MODULES . . . . . . . . . . . . . . . 84<br />

7.1 The Generic Iwahori-Hecke Algebra . . . . . . . . . . . . . . . . . 84<br />

7.2 A Module for H . . . . . . . . . . . . . . . . . . . . . . . . . . . 85<br />

7.3 The Hecke Algebra Modules from hW,p,M∞ . . . . . . . . . . . . . 86<br />

7.4 The 0-Hecke Algebra . . . . . . . . . . . . . . . . . . . . . . . . . 89<br />

CHAPTER 8: <strong>IN</strong>ITIAL SECTIONS OF REFLECTION ORDERS . . . . 90<br />

8.1 Classification of Twisted Bruhat Orders . . . . . . . . . . . . . . . 90<br />

8.2 Proof of the Classification . . . . . . . . . . . . . . . . . . . . . . 92<br />

CHAPTER 9: CLOSURE <strong>IN</strong> THE ROOT SYSTEM . . . . . . . . . . . . 101<br />

9.1 2-closure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101<br />

9.2 More types of subsets of Φ + . . . . . . . . . . . . . . . . . . . . . 102<br />

9.3 A Conjectural Normal Form for Elements of W . . . . . . . . . . 103<br />

iv


CHAPTER 10: HYPERBOLIC <strong>COXETER</strong> <strong>GROUPS</strong> . . . . . . . . . . . 108<br />

10.1 Coxeter Graphs . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108<br />

10.2 Symmetric Matrices and Sylvester’s Law . . . . . . . . . . . . . . 109<br />

10.3 Hyperbolic Coxeter Systems . . . . . . . . . . . . . . . . . . . . . 109<br />

10.4 Non-affine Coxeter Systems . . . . . . . . . . . . . . . . . . . . . 110<br />

10.5 A Class of Coxeter Systems . . . . . . . . . . . . . . . . . . . . . 112<br />

10.6 A Smaller Class of Coxeter Systems . . . . . . . . . . . . . . . . . 112<br />

10.7 Parabolic Subsystems of Non-affine Coxeter Systems . . . . . . . 114<br />

10.8 Characterizations of Hyperbolic Coxeter Systems . . . . . . . . . 114<br />

CHAPTER 11: LARGE REFLECTION SUB<strong>GROUPS</strong> . . . . . . . . . . . 116<br />

11.1 Convex Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116<br />

11.2 A Metric on Convex Sets . . . . . . . . . . . . . . . . . . . . . . . 117<br />

11.3 Approximation of a Sphere by Polytopes . . . . . . . . . . . . . . 118<br />

11.4 Cones in Coxeter Systems . . . . . . . . . . . . . . . . . . . . . . 120<br />

11.5 Topology on Rays in Cones . . . . . . . . . . . . . . . . . . . . . 121<br />

11.6 Cones in Hyperbolic Coxeter Systems . . . . . . . . . . . . . . . . 122<br />

11.7 Universal Coxeter Systems . . . . . . . . . . . . . . . . . . . . . . 123<br />

11.8 Large Reflection Subgroups . . . . . . . . . . . . . . . . . . . . . 125<br />

CHAPTER 12: EXPONENTIAL GROWTH <strong>IN</strong> <strong>COXETER</strong> <strong>GROUPS</strong> . . 129<br />

12.1 Growth Types . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130<br />

12.2 Properties of W(3) . . . . . . . . . . . . . . . . . . . . . . . . . . . 130<br />

12.3 Parabolic Quotients . . . . . . . . . . . . . . . . . . . . . . . . . . 131<br />

12.4 Proof of Proposition 12.3.3 . . . . . . . . . . . . . . . . . . . . . . 132<br />

REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134<br />

v


FIGURES<br />

3.1 This diagram shows the idea of generalized length in � B2 = 〈r, s, t |<br />

r 2 = s 2 = t 2 = (rt) 2 = 1, (rs) 4 = (st) 4 = 1〉. Each group element<br />

corresponds to one of the alcoves, with the yellow triangle<br />

representing the identity. Each reflection corresponds to one of the<br />

affine hyperplanes, and the color of the hyperplane distinguishes the<br />

conjugacy class. There are three conjugacy classes. The standard<br />

length counts the number of affine hyperplanes separating w from<br />

the yellow alcove. The generalized length counts this same number<br />

but retains additional information about the conjugacy class<br />

of each affine hyperplane separating w from the identity. For the<br />

w specified above we have l(w) = 12 while lp(w) = (3, 6, 3) where<br />

p represents the function taking distinct values on each conjugacy<br />

class. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17<br />

5.1 This diagram shows the tiling of the plane corresponding to � B2 as<br />

in Figure 3.1. Each of the gray regions is an intersection of two half<br />

spaces, and thus each is p-completely regular for any p. Then, the<br />

total gray region is a union of two p-completely regular sets and so<br />

is p-completely regular itself. Therefore, this set also has a rational<br />

multivariate Poincaré series. . . . . . . . . . . . . . . . . . . . . . 53<br />

8.1 This is a schematic diagram of the root system for (W, S) infinite.<br />

Φ + consists of the roots which are non-negative linear combinations<br />

of αr and αs. The dotted ray through the origin represents a limit<br />

line of roots. If C1 and C2 are not both satisfied by Φ + then Ψ is<br />

pictured above with the roots in Ψ labeled by •, and those not in<br />

Ψ labeled by ◦. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97<br />

vi


ACKNOWLEDGMENTS<br />

First and foremost, I would like to thank my advisor, Matthew Dyer. I am<br />

always amazed by how much he knows and is willing to teach me, whether it be<br />

for an hour in his office or over the computer thousands of miles away. Without<br />

his endless patience, encouragement, and ideas, I would not have finished this<br />

work. I would also like to thank his family, Annette and Katy, for their wonderful<br />

hospitality and amusing conversations.<br />

I would like to thank the members of my thesis committee, Mark Colarusso,<br />

Sam Evens, and Michael Gekhtman, who took much of their valuable time to read<br />

the thesis and provide thoughtful comments. I would further like to thank Sam<br />

Evens for his conversations about research throughout my fifth year.<br />

Additionally, I would like to thank many of the mathematicians I have had the<br />

pleasure of working with and learning from during my years as a student. I would<br />

like to thank Lorelei Koss, Dave Richeson, and Barry Tesman for their excellent<br />

teaching and for inspiring me to choose a career in mathematics. I thank Anton<br />

Betten for his guidance in my master’s program and for teaching me a great deal<br />

of interesting combinatorics. I would like to acknowledge and thank Brian Hall<br />

for the many enlightening conversations about root systems and mathematics in<br />

general. I would like to thank fellow graduate students Benjamin Jones, Don<br />

Brower, and John Wallbaum for many interesting and helpful conversations; I<br />

would moreover like to thank all the Notre Dame graduate students for the good<br />

vii


times spent here.<br />

Without my family, I would not be where I am today. I would like to express<br />

deep gratitude to Margot and Dick for treating us like their own children during<br />

our time in Indiana. I thank my sister, Kristin, for her loving support and will-<br />

ingness to just call and talk. I thank my mom, Sue, for everything she has done<br />

for me in my life. She is an amazing woman and someone I will always look up<br />

to. I would also like to thank my dad, Mark. He taught me how to work hard,<br />

and he always believed in me. I have never met a better man, and I will forever<br />

miss him.<br />

Finally, and most of all, I thank my beautiful wife, Laura. Her love and support<br />

have been an integral part of my success as a graduate student. She makes life<br />

fun, and I couldn’t imagine what it would be like if she were not around. I owe<br />

her immensely for her willingness to help me achieve this goal.<br />

viii


SYMBOLS<br />

W Coxeter group<br />

S Coxeter group generators<br />

l Standard length function on W<br />

lp<br />

General length function on W<br />

T Reflections<br />

N Reflection cocycle<br />

Φ Root system<br />

Φ + Positive roots<br />

Π Simple roots<br />

( , ) Bilinear pairing on root system<br />

≤A<br />

Twisted Bruhat order on group<br />

W ′ Reflection subgroup<br />

M Monomials in Z[X]<br />

� Dominance order on reflections<br />

≤ Weak order on T ; Order on M<br />

ℓp<br />

General length function on S ∗<br />

ix


CHAPTER 1<br />

<strong>IN</strong>TRODUCTION<br />

In this dissertation, we investigate some of the applications of dominance order<br />

on the root system (or reflections) of a Coxeter group. Dominance was first defined<br />

by Brink and Howlett in [5], where it is used to show that Coxeter groups are<br />

automatic. Brink further investigates the dominance minimal roots in [4].<br />

Despite this work, the dominance order is still poorly understood in general.<br />

More recently, in [15] Dyer has employed the idea of using the imaginary cone<br />

as a combinatorial tool for studying dominance in the root system. Our general<br />

goal will be to further the understanding of the role dominance plays in different<br />

aspects of study in Coxeter systems. The dissertation is organized as follows.<br />

Chapter 2 introduces the basic background information necessary to develop<br />

the results in the later chapters. These ideas will be used throughout the thesis.<br />

The thesis is broken up into three independent parts, which are loosely related<br />

through their connections to dominance order.<br />

In Chapter 3, we examine general length functions on Coxeter systems. These<br />

give rise to a class of height functions, which we demonstrate can be understood<br />

by studying the weak order and give a nice characterization of dominance order.<br />

These height functions allow us to define certain finite sets of reflections that parti-<br />

tion the entire set of reflections. We begin Chapter 4 by defining regular languages<br />

and finite state automata. We are then able to use the partition of the reflections<br />

1


described in Chapter 3 to define a class of canonical finite automata for the Cox-<br />

eter system. These automata generalize the automata defined by Dyer in [18]. In<br />

Chapter 5, we use the automata defined in Chapter 4 to describe large classes of<br />

regular subsets of Coxeter groups, i.e. sets that can be recognized by a finite state<br />

automaton. We introduce different types of regularity; in particular, our notion<br />

of p-complete regularity allows us to easily describe certain regular subsets of a<br />

Coxeter group. Regularity has some interesting and important implications for<br />

(multivariate) Poincaré series of subsets of a Coxeter system, and we will discuss<br />

these briefly. Due to the fact that many natural subsets of Coxeter systems cannot<br />

be described as completely regular in the sense of Chapter 5, we use Chapter 6<br />

to define an alternate type of regularity, which will allow us to demonstrate the<br />

regularity of many subsets of the set of reflections in a Coxeter system. We end<br />

with a characterization of regular sets inside of products W n × T m where W is a<br />

Coxeter system with T the set of reflections. Finally, in Chapter 7, we describe<br />

how the finite sets that give regularity also lead to a large family of modules for<br />

the generic Iwahori-Hecke algebra.<br />

Next, in Chapters 8 and 9, we discuss a notion of closure in the root system<br />

(or reflections). Chapter 8 classifies possible partial orders on a Coxeter system,<br />

generalizing Bruhat order, using certain closed subsets of the roots. These results<br />

appear in [20] as well. Chapter 9 focuses on discussing closure in general and other<br />

types of subsets of roots. We outline some conjectures, posed by Dyer (see [17]),<br />

involving closure in the root system. We also describe an application of these<br />

conjectures related to closure of the sets defined in Chapter 4, which would lead<br />

to a normal form for elements of the Coxeter group.<br />

Finally, in Chapters 10-12, we investigate some properties of the imaginary<br />

2


cone, especially in hyperbolic Coxeter systems. In Chapter 10, we provide some<br />

characterizations of hyperbolic Coxeter systems. As a byproduct of our charac-<br />

terizations, we obtain a simple, case-free proof of the well known fact that any<br />

Coxeter system contains a hyperbolic subsystem as a standard parabolic subsys-<br />

tem. Chapter 11 focuses on studying the imaginary cone in the case of hyperbolic<br />

Coxeter systems. We use the imaginary cone to show that any Coxeter system con-<br />

tains universal reflection subgroups of arbitrarily large rank. Furthermore, in the<br />

hyperbolic case, the positive spans of the simple roots of the universal reflection<br />

subgroups are shown to approximate the imaginary cone (using an appropriate<br />

topology on the set of roots). This answers in the affirmative a question due to<br />

Dyer [15] in the special case of hyperbolic Coxeter systems. Lastly, Chapter 12<br />

uses the results from Chapter 11 describing large universal reflection subgroups<br />

in any Coxeter system to extend the results of [29] regarding exponential growth<br />

in parabolic quotients in Coxeter groups.<br />

3


CHAPTER 2<br />

BACKGROUND MATERIAL<br />

We begin with a chapter introducing the basic objects and terminology used<br />

throughout the thesis.<br />

2.1 Coxeter Groups<br />

Definition 2.1.1. We define a Coxeter system to be a pair (W, S) consisting of<br />

a group W and a set of generators S ⊂ W subject only to relations of the form<br />

(ss ′ ) m(s,s′ ) = 1 where m(s, s) = 1 and m(s, s ′ ) = m(s ′ , s) ≥ 2 for s �= s ′ . If s and<br />

s ′ have no such relation, we say that m(s, s ′ ) = ∞.<br />

We may sometimes abuse terminology and call (W, S), or W , a Coxeter group.<br />

For general results about Coxeter groups, we refer the reader to [23] or [2]. In<br />

general, S does not need to be finite. For many of the results we will state we<br />

need S to be finite. We shall specify when this is the case.<br />

Every element w ∈ W can be written as a product of the generators w =<br />

s1s2 · · · sm, each si ∈ S. If m is minimal among all such expressions for w we will<br />

say that s1s2 · · · sn is a reduced expression for w and we say that w has length n,<br />

i.e. l(w) = n. So l : W → N denotes the standard length function on (W, S).<br />

We let T denote the set of W -conjugates of S, T := �<br />

w∈W wSw−1 , and call<br />

T the set of reflections. Then, following [19] we define N : W → P(T ) to be<br />

4


the “reflection cocycle” given by N(w) = {t ∈ T | l(tw) < l(w)}. We call N(w) a<br />

cocycle because it satisfies the cocycle condition:<br />

N(xy) = N(x) + xN(y)x −1<br />

where + represents symmetric difference on the power set of T , A + B = (A ∪<br />

B) \ (A ∩ B). The existence of this cocycle implies that W acts on P(T ) by<br />

w · A = N(w) + wAw −1 for A ⊆ T .<br />

2.2 Root Systems<br />

Without loss of generality, we will assume that (W, S) is realized in its standard<br />

reflection representation on a real vector space V with Π denoting the set of simple<br />

roots and (−, −) : V × V → R representing the associated bilinear form given<br />

by (α, β) = − cos π<br />

m(sα,sβ) if m(sα, sβ) �= ∞ and (α, β) ≤ −1 otherwise. Then<br />

we denote the roots and positive roots by Φ and Φ + respectively. Recall that<br />

Φ = Φ + ˙∪ − Φ + , and we may sometimes write Φ − := −Φ + .<br />

For α ∈ Φ we let sα ∈ T represent the corresponding reflection. Recall that<br />

the set of reflections are in bijective correspondence with the set of positive roots<br />

under the bijection α ↦→ sα : Φ + → T . For t ∈ T , we will let αt ∈ Φ + denote the<br />

unique positive root α with sα = t. We recall the following two facts<br />

l(wt) < l(w) ⇔ w(αt) ∈ Φ − for t ∈ T and w ∈ W (2.2.1)<br />

l(tw) < l(w) ⇔ w −1 (αt) ∈ Φ − for t ∈ T and w ∈ W (2.2.2)<br />

5


We can then describe the W -action on Φ in terms of the length function l by<br />

the following formula<br />

for any w ∈ W and t ∈ T .<br />

2.3 Reflection Subgroups<br />

⎧<br />

⎪⎨ 1, if l(wt) < l(w)<br />

w(±αt) = ±ɛαwtw−1, ɛ =<br />

⎪⎩ −1, if l(wt) > l(w)<br />

(2.2.3)<br />

We call a subgroup W ′ of W a reflection subgroup of W if it is generated by<br />

the reflections it contains, W ′ = 〈W ′ ∩ T 〉. It was shown by Dyer ([19]) and<br />

Deodhar ([10]) independently that any reflection subgroup is also a Coxeter sys-<br />

tem; moreover, any reflection subgroup has a canonical set of Coxeter generators<br />

χ(W ′ ) = {t ∈ T | N(t) ∩ W ′ = {t}}. We say a reflection subgroup is dihedral<br />

if it is generated by two distinct reflections or equivalently |χ(W ′ )| = 2. We let<br />

l(W ′ ,χ(W ′ )) : W ′ → N be the length function for (W ′ , χ(W ′ )) and T ′ = W ′ ∩ T<br />

be the set of reflections of W ′ . Due to [19], N(W ′ ,χ(W ′ ))(x) = N(x) ∩ W ′ for all<br />

reflection subgroups W ′ where N(W ′ ,χ(W ′ )) is the reflection cocycle for (W ′ , χ(W ′ )).<br />

Any dihedral reflection subgroup, W ′ = 〈sα, sβ〉, is contained in a unique<br />

maximal dihedral reflection subgroup, namely 〈sγ| γ ∈ Φ ∩ (Rα + Rβ)〉. We let<br />

M = {W ′ < W | W ′ is a maximal dihedral reflection subgroup}<br />

M∞ = {W ′ ∈ M | |W ′ | = ∞}.<br />

6


For any reflection t ∈ T , we get the following set:<br />

The previous remarks imply that<br />

Mt = {W ′ ∈ M | t ∈ W ′ }.<br />

T \ {t} = ˙�<br />

where ˙∪ represents a disjoint union.<br />

W ′ ∈Mt<br />

((W ′ ∩ T ) \ {t}) (2.3.1)<br />

Since any reflection subgroup is also a Coxeter system (W ′ , χ(W ′ )), we let ΦW ′,<br />

Φ +<br />

W ′, ΠW ′ ⊆ Φ be the set of roots, positive roots, and simple roots for (W ′ , χ(W ′ ))<br />

sitting inside the root system for (W, S). We have that ΦW ′ = {α ∈ Φ| sα ∈ W ′ }.<br />

For any w ∈ W , there exists a unique x ∈ W ′ w with N(x)∩W ′ = ∅ (likewise a<br />

unique x ∈ wW ′ with N(x −1 )∩W ′ = ∅). According to [19], we have χ(w −1 W ′ w) =<br />

x −1 χ(W ′ )x. Let w = yx with y ∈ W ′ . Then α ↦→ w −1 (α) = x −1 y −1 (α) induces a<br />

bijection ΦW ′<br />

α → y −1 (α) : ΦW ′<br />

∼=<br />

→ Φw−1W ′ w. This bijection is obtained by composing the bijection<br />

∼=<br />

→ ΦW ′ given by the action of y−1 ∈ W ′ on the root system of<br />

∼=<br />

→ Φx−1W ′ x = Φw−1W ′ w give by the action<br />

W ′ and the bijection α → x −1 (α) : ΦW ′<br />

of x. Since N(x) ∩ W ′ = ∅, the latter bijection preserves positive roots as well.<br />

2.4 The Weak Order on T and the Set of Elementary Reflections<br />

Much of the terminology of this section is described in terms of the root system<br />

in [2]. We will instead use terminology associated to the set of reflections when<br />

possible. For any reflection t ∈ T , we define the standard height to be h(t) :=<br />

h(W,S)(t) = l(t)−1<br />

2 .<br />

At this point, we define a partial order on the set of reflections. For any<br />

7


t, t ′ ∈ T , we set t → t ′ if there is some s ∈ S (it is unique if it exists) such that<br />

t ′ = sts and l(t ′ ) = l(t) + 2 (or equivalently h(t ′ ) = h(t) + 1). Let ≤ be the<br />

partial order on T given as the reflexive transitive closure of →, and call (T, ≤)<br />

the reflection poset or the weak order on T . It is clear that h is the rank function<br />

for (T, ≤).<br />

From the formula given in (2.3.1), we see that for t ∈ T we have<br />

h(t) = �<br />

W ′ ∈Mt<br />

h(W ′ ,χ(W ′ ))(t)<br />

where h(W ′ ,χ(W ′ )) is the height function for (W ′ , χ(W ′ )). Now, we use this charac-<br />

terization of height to introduce the “∞-height” of a reflection as follows.<br />

h ∞ (t) := �<br />

W ′ ∈Mt<br />

|W ′ |=∞<br />

h(W ′ ,χ(W ′ ))(t). (2.4.1)<br />

We now introduce a special subset of the reflections. Let T0 := {t ∈ T | h ∞ (t) =<br />

0}. We call this set the set of elementary reflections. Using the bijection between<br />

T and Φ + , this set is in bijective correspondence with the elementary roots defined<br />

in [5]. In that paper, Brink and Howlett also demonstrate that this set is finite.<br />

For another description of this fact, see [2].<br />

More recently, Dyer has shown similar sets to be finite. More precisely, we<br />

define, for any n ∈ N, Tn := {t ∈ T | h ∞ (t) = n}. For these sets, we have the<br />

following theorem, due to Dyer, found in [14].<br />

Theorem 2.4.1. Let (W, S) be a Coxeter system with reflections T . If S is a<br />

finite set, then Tn is a finite set for all n ∈ N.<br />

These sets of reflections, as well as others like it, will be defined in more<br />

8


generality in Chapter 3, and we will make extensive use of these types of sets in<br />

Chapter 5 in order to demonstrate nice regularity properties of Coxeter systems.<br />

We will also use these sets to define certain modules for the generic Iwahori-Hecke<br />

algebra associated to (W, S) in Chapter 7.<br />

2.5 Dominance Order on T<br />

The set of elementary reflections also has a nice characterization in terms of<br />

another partial order on the set of reflections; it is the set of minimal elements of<br />

this partial order. Dominance order was initially defined in [5]. The elementary<br />

reflections and the dominance order are studied in more detail in [4]. The basic<br />

properties of dominance can be found in [2],[4], or [5], but we will include a few<br />

results here.<br />

We will first define dominance on the root system associated to (W, S) and use<br />

the canonical bijection of positive roots with reflections to transport the definition<br />

to the reflections.<br />

We say that α ∈ Φ dominates β ∈ Φ written β � α if, for all w ∈ W ,<br />

w(α) ∈ Φ − implies that w(β) ∈ Φ − .<br />

Lemma 2.5.1. 1. If α, β ∈ Φ, then α dominates β in Φ if and only if for<br />

some (respectively every) reflection subgroup W ′ of W with sα, sβ ∈ W ′ , α<br />

dominates β in ΦW ′ (with respect to W ′ ).<br />

2. � is a partial order on Φ.<br />

3. Multiplication by −1 is an order-reversing bijection of (Φ, �) with itself.<br />

4. The W -action on Φ is as a group of poset automorphisms of (Φ, �).<br />

9


Proof. First, suppose that β, α ∈ Φ and that sα and sβ are contained in a reflection<br />

subgroup W ′ . According to section 2.3, we have that for any w ∈ W , there is<br />

x ∈ wW ′ with N(x −1 ) ∩ W ′ = ∅, and α ↦→ x(α) is a bijection, which preserves<br />

positive roots, between ΦW ′ and Φ wW ′ w −1. Let w = xy with y ∈ W ′ . Suppose<br />

that β � α in ΦW ′ with respect to W ′ . Then if w(α) ∈ Φ − , we have xy(α) ∈ Φ −<br />

and so y(α) ∈ Φ −<br />

W ′. Thus, w(β) = xy(β) ∈ x(Φ −<br />

W ′) = Φ −<br />

W ′ ⊂ Φ − and so β � α<br />

in Φ. If β � α in Φ, then by definition, w(α) ∈ Φ − implies w(β) ∈ Φ − for all<br />

w ∈ W and so this is true for all w ′ ∈ W as required for 1. We now describe<br />

the dominance order on dihedral Coxeter systems. Let (W, S) be dihedral, with<br />

S = {r, s}. Then if W is finite, all the (finite) elements of Φ are incomparable<br />

in �. If W is infinite, then the dominance order is given by the coarsest partial<br />

order satisfying the relations<br />

and<br />

· · · − αsrsrs ≺ −αsrs ≺ −αs ≺ αr ≺ αrsr ≺ αrsrsr ≺ · · ·<br />

· · · − αrsrsr ≺ −αrsr ≺ −αr ≺ αs ≺ αsrs ≺ αsrsrs ≺ · · · .<br />

In either of these cases, it is clear that (Φ, �) is a partial order. For general<br />

(not necessarily dihedral) (W, S), it is clear that � is reflexive and transitive by<br />

definition. Then, antisymmetry in general follows by using antisymmetry in a<br />

reduction to a dihedral reflection subgroup. Namely, suppose β � α � β with<br />

α �= β in Φ, and then consider W ′ = 〈sα, sβ〉. Due to 1, we see that β � α � β<br />

in ΦW ′, and by antisymmetry in the dihedral case, we get that α = β, which<br />

is a contradiction. Finally, 3 and 4 are simple consequences of the definition of<br />

dominance. In particular if β � α then w(β) � w(α) since for any u ∈ W we have<br />

10


uw(α) ∈ Φ − implies that uw(β) ∈ Φ − .<br />

Then, as stated before, we use the canonical bijection between Φ + and T to<br />

define the notion of dominance on the set of reflections. We have that sβ � sα if<br />

and only if β � α in Φ + . This is equivalent to t ′ � t if for all w ∈ W we have<br />

l(wt) < l(w) implies that l(wt ′ ) < l(w); in this case, we say that t dominates t ′ .<br />

We will write (T, �) when referring to the set of reflections partially ordered by<br />

dominance.<br />

We recall that in a poset, we say an element a covers an element b if b < a and<br />

there is no z ∈ P with b < z < a.<br />

Lemma 2.5.2. Let (W, S) be a Coxeter system.<br />

1. If s, t ∈ T0 generate an infinite subgroup, then {s, t} is the set of canonical<br />

Coxeter generators for the maximal dihedral reflection subgroup of (W, S)<br />

containing {s, t}.<br />

2. If α, β ∈ Φ + are distinct elementary roots (i.e. sα and sβ are in T0) and<br />

〈sα, sβ〉 is an infinite dihedral reflection subgroup then β covers −α in dom-<br />

inance order.<br />

Proof. Due to the definition of h ∞ in equation (2.4.1), we know that t ∈ T0 if and<br />

only if<br />

0 = h ∞ (t) = �<br />

W ′ ∈Mt<br />

|W ′ |=∞<br />

h(W ′ ,χ(W ′ ))(t)<br />

which implies that h(W ′ ,χ(W ′ ))(t) = 0 for all infinite dihedral reflection subgroups<br />

W ′ . Therefore, by the definition of the height, we see that t ∈ χ(W ′ ) for all<br />

infinite, maximal dihedral reflection subgroups of W with t ∈ W ′ ∩ T . We see<br />

that 1 follows from this fact. To prove 2, since 〈sα, sβ〉 is infinite then −α ≺ β<br />

11


y the description of dominance in infinite dihedral reflection subgroups given in<br />

2.5.1. Suppose that γ ∈ Φ and −α ≺ γ ≺ β. Since β is elementary, we cannot have<br />

γ ∈ Φ + ; similarly, since α is elementary and −γ ≺ α we cannot have −γ ∈ Φ + so<br />

that γ �∈ Φ − . This contradicts the fact that Φ = Φ + ˙∪Φ − .<br />

Remark 2.5.3. In [14], Dyer uses the notion of the imaginary cone in a Coxeter<br />

system (see Chapter 11) to characterize dominance. He demonstrates that every<br />

covering α ′ ≺ β ′ in dominance order is in the W -orbit of a covering α ≺ β as in<br />

part 2 of 2.5.2.<br />

12


CHAPTER 3<br />

GENERALIZED LENGTH FUNCTIONS<br />

In this chapter, we begin by introducing a family of generalized length functions<br />

on a Coxeter system. These length functions replace the standard length function<br />

on (W, S) by taking values in N k for some k ≥ 1 instead of simply taking values in<br />

N. In a similar manner to section 2.4, these length functions will allow us to define<br />

corresponding height functions and, more importantly, a class of “infinite-height”<br />

functions on the set of reflections. These height functions lead to a partition of the<br />

set of reflections with interesting applications. Due to Theorem 2.4.1, these sets<br />

of reflections turn out to be finite. We will use the finiteness property to develop<br />

many canonical automata in Chapter 4.<br />

We proceed by interpreting these height functions in terms of the poset (T, ≤)<br />

(weak order) as defined in section 2.4. In turn, we further obtain monotonicity<br />

properties of these length and height functions with respect to the dominance<br />

order.<br />

Finally, we demonstrate a basic recurrence formula for the reflection cocycle<br />

on (W, S) in terms of this new partition of T . We end with a few results that will<br />

be important in the following chapters. Many of the results generalize some of the<br />

results in [18].<br />

13


3.1 Commutative Monoids<br />

At this point we want to define a commutative monoid M. We have two<br />

equivalent definitions for M which we introduce below. Both definitions will be<br />

useful for different applications, but it will be clear from context how we want to<br />

view M since one definition will have multiplicative notation and the other will<br />

have additive notation.<br />

Let X be a finite set of indeterminates and Z[X] be the corresponding integral<br />

polynomial ring. We say m ∈ Z]X] is a monomial if m = �<br />

x∈X x ax where ax ∈ N<br />

for all x ∈ X. Let M := MX be the set of all monomials of Z[X], that is<br />

M := {m| m ∈ Z[X]; m is a monomial}. For any m ∈ M, m = �<br />

x∈X x ax ,<br />

define the degree of m to be deg(m) := �<br />

x∈X ax.<br />

If we fix an order for X = {x1, ..., xk}, we can also consider the commutative<br />

monoid N k defined by the map ζ : M → N k given by m = x a1<br />

1 x a2<br />

2 · · · x ak<br />

k ↦→<br />

(a1, ..., ak). Since ζ is an isomorphism, we will abuse notation by setting N k = M.<br />

The manner in which we view M will be clear depending on whether we use<br />

additive or multiplicative notation. We note that for m ∈ M, deg(m) = �<br />

x∈X ax<br />

is independent of how we view M. For any monomial m = (a1, ..., ak) we let<br />

m(i) := ai.<br />

For X as above, let X −1 = {x −1 | x ∈ X}. We can define M ′ := {m| m ∈<br />

Z[X, X −1 ]; m is a monomial} ∼ = Z k . This set has a natural group structure. We<br />

will also view M as a subset of M ′ ; in the case when we view M additively,<br />

this will explain using the notation m − n (and will explain using division when<br />

viewing M multiplicatively).<br />

We place a partial order ≤M on M as follows. For any m ∈ M and n ∈ M,<br />

we say m ≤M n if and only if n<br />

m ∈ M ⊂ M′ (additively this means n − m ∈<br />

14


M ⊂ M ′ ) . For x ∈ X, we say that m ✁x n if n<br />

m<br />

= x and call this a covering<br />

relation of type x. It is clear that if m ✁x n and m ≤M m ′ ≤M n then m ′ = m or<br />

m ′ = n. From this point on, we will write ≤:=≤M; this will not cause confusion<br />

with ≤ on T as the context will make clear which ≤ is intended.<br />

Now, by an ordered generalized partition of a monomial m we will mean a<br />

tuple M = (m1, m2, ...), mn = 0 for n ≫ 0, with deg(mi) ≥deg(mj) for all i < j<br />

along with �<br />

i mi = m. Next, we define an equivalence relation, ∼, on the set of<br />

ordered generalized partitions in the following way. Let (m1, m2, ...) ∼ (n1, n2, ...)<br />

if there exists a permutation of the natural numbers, σ, such that mi = nσ(i)<br />

for all i. For (m1, m2, ...) an ordered partition, we let M = [(m1, m2, ...)] be<br />

the equivalence class of the ordered partition. We call these equivalence classes<br />

generated by this relation unordered generalized partitions. For any unordered<br />

partition M = [(m1, m2, ...)], we define |M| := �<br />

i mi. All the terminology of the<br />

partial order on M can be phrased additively if necessary.<br />

At this point, we can define a partial order on the set of unordered generalized<br />

partitions. Let M and N be any two generalized partitions. Then we say that<br />

M ≤ N if there exists (m1, m2, ...) ∈ M and (n1, n2, ...) ∈ N such that mi ≤ ni in<br />

M for all i. For any x ∈ X, we define the covering relation of type x by M ✁x N<br />

if there exists (m1, m2, ...) ∈ M and (n1, n2, ...) ∈ N and j ∈ N such that mi = ni<br />

for all i �= j and mj ✁x nj in M.<br />

3.2 A General Length for Coxeter Systems<br />

Suppose we have a fixed Coxeter system (W, S). Let M be the additive,<br />

commutative monoid described in section 3.1 where M is generated by a finite<br />

set X = {x1, ..., xk} of indeterminates. Suppose that we are given a surjective<br />

15


function p : T → X satisfying p(t) = p(t ′ ) if t is conjugate to t ′ , i.e. if there exists<br />

v ∈ W with t = vt ′ v −1 . For any reflection t ∈ T , we say that t is of type xi if<br />

p(t) = xi where xi ∈ X (for simplicity we will usually say that t is of type p(t)).<br />

Until further notice, we will view M with additive notation and we will, for ease,<br />

let p(t) represent ζ(p(t)).<br />

We define a family of “length” functions in the following way. Let W ′ be any<br />

reflection subgroup of W and w ∈ W ′ where w = r1...rn is a reduced expression<br />

with ri ∈ χ(W ′ ). Then we define lp,W ′ : W ′ → M by lp,W ′(w) = lp,W ′(r1...rn) =<br />

p(r1) + p(r2) + · · · + p(rn). See figure 3.1 for an example.<br />

Similarly, for any reflection r ∈ W ′ ∩ T with r = r1...rmrm+1...r2m+1, we define<br />

a corresponding “height function” hp,W ′ : T → M by hp,W ′(r) = p(r1) + p(r2) +<br />

· · · + p(rm). Finally, for any r ∈ W ′ ∩ T where r = r1...rmrm+1...r2m+1, we define<br />

c(r) := p(r) = p(rm+1), which is called the center of r. Note that for t ∈ T ,<br />

lp,W ′(t) = c(t) + 2hp,W ′(t).<br />

From these height functions, we get the following functions on reflections as<br />

well. For any union O ⊂ M of W -orbits of maximal dihedral reflection subgroups<br />

on M , we define<br />

and<br />

hW,p,O : t ↦→ �<br />

W ′ ∈Mt<br />

W ′ ∈O<br />

hp,W ′(t) (3.2.1)<br />

lW,p,O : t ↦→ c(t) + 2hW,p,O(t). (3.2.2)<br />

In this thesis, we will only consider the case where O = M∞.<br />

Remark 3.2.1. We note that, considering terminology from section 2.4, we have<br />

h ∞ (t) = deg(hW,p,M∞(t)). Also, if M = N and q(t) = 1 for all t ∈ T , then<br />

16


Figure 3.1. This diagram shows the idea of generalized length in � B2 =<br />

〈r, s, t | r 2 = s 2 = t 2 = (rt) 2 = 1, (rs) 4 = (st) 4 = 1〉. Each group element<br />

corresponds to one of the alcoves, with the yellow triangle representing the<br />

identity. Each reflection corresponds to one of the affine hyperplanes, and<br />

the color of the hyperplane distinguishes the conjugacy class. There are<br />

three conjugacy classes. The standard length counts the number of affine<br />

hyperplanes separating w from the yellow alcove. The generalized length<br />

counts this same number but retains additional information about the<br />

conjugacy class of each affine hyperplane separating w from the identity.<br />

For the w specified above we have l(w) = 12 while lp(w) = (3, 6, 3) where<br />

p represents the function taking distinct values on each conjugacy class.<br />

17<br />

w


h ∞ (t) = hW,q,M∞(t).<br />

Lemma 3.2.2. Let (W, S) be a Coxeter system, and let ≤∅ be ordinary Bruhat<br />

order on (W, S) (see Chapter 8). Let x, y ∈ W with x ≤∅ y. Then for all reflection<br />

subgroups W ′ ⊆ W with x, y ∈ W ′ , the following hold.<br />

1. lp,W ′(x) ≤ lp,W ′(y), and<br />

2. hp,W ′(x) �> hp,W ′(y) if x, y ∈ T ∩ W ′ .<br />

3. hp,W ′(x) ≤ hp,W ′(y) if x, y ∈ T ∩ W ′ and c(x) = c(y).<br />

Proof. We know that x ≤∅ y in W if and only if x ≤∅ y in W ′ for all x, y ∈ W ′ ⊆<br />

W . The subword characterization of Bruhat order and the definition of lp,W in<br />

terms of reduced expressions imply 1. Finally, 2 and 3 follow from 1.<br />

3.3 Infinite Heights and Generalized Partitions for Reflections<br />

According to equation (3.2.1) we can attach an unordered generalized partition<br />

to each reflection t, MW,p,M∞(t), where each part, mi, is given by some distinct<br />

hp,W ′(t) with W ′ ∈ M∞.<br />

Lemma 3.3.1. Let t ∈ T and s ∈ S with lp(sts) = lp(t) + 2p(s). Then<br />

1. hW,p,M∞(sts) = hW,p,M∞(t) if 〈s, t〉 is finite and hW,p,M∞(sts) = hW,p,M∞(t)+<br />

p(s) if 〈s, t〉 is infinite.<br />

2. MW,p,M∞(sts) = MW,p,M∞(t) if 〈s, t〉 is finite. If 〈s, t〉 is infinite then<br />

MW,p,M∞(t) ✁p(s) MW,p,M∞(sts).<br />

Proof. Let W ′ be a maximal dihedral reflection subgroup. According to 2.3 if<br />

s �∈ W ′ , then N(s) ∩ W ′ = ∅ so (W ′ , χ(W ′ )) ∼ = (sW ′ s, sχ(W ′ )s) and thus<br />

18


hp,sW ′ s(sts) = hp,W ′(t). If s ∈ W ′ , then s ∈ χ(W ′ ) and sW ′ s = W , and it<br />

follows that hp,W ′(sts) = hp,W ′(t) + p(s). Now, since Msts = {sW ′ s| W ′ ∈ Mt}, if<br />

〈s, t〉 ∈ Mt ∩ M∞ then we have<br />

Otherwise we have<br />

hW,p,M∞(sts) = �<br />

W ′ ∈Msts<br />

W ′ ∈M∞<br />

= p(s) + �<br />

hW,p,M∞(sts) = �<br />

W ′ ∈Msts<br />

W ′ ∈M∞<br />

= �<br />

W ′ ∈Mt<br />

W ′ ∈M∞<br />

�<br />

hp,W ′(sts) =<br />

W ′ ∈Mt<br />

W ′ ∈M∞<br />

W ′ ∈Mt<br />

W ′ ∈M∞<br />

hp,sW ′ s(sts)<br />

hp,W ′(t) = p(s) + hW,p,M∞(t).<br />

�<br />

hp,W ′(sts) =<br />

W ′ ∈Mt<br />

W ′ ∈M∞<br />

hp,W ′(t) = hW,p,M∞(t).<br />

Based on these results, the lemma follows clearly.<br />

3.4 Infinite Height and the Weak Order<br />

hp,sW ′ s(sts)<br />

Consider the Hasse diagram of the weak order (T, ≤) as a directed graph with<br />

vertex set T and directed edges (t, t ′ ) from t to t ′ for all t, t ′ ∈ T with t → t ′ .<br />

Furthermore, we can regard this graph as a labeled directed graph by labeling<br />

the edges as follows. For any edge (t, t ′ ), we know there is a unique s ∈ S with<br />

t ′ = sts. From this, we can label the edge (t, t ′ ) with label p(s), and we call this<br />

edge an edge of type p(s). We say that an edge (t, t ′ ) is p(s)-short if it is labeled<br />

with p(s) and if hW,p,M∞(t ′ ) = hW,p,M∞(t), i.e. if 〈t, s〉 is finite. We say that the<br />

edge is p(s)-long if it is labeled with p(s) and hW,p,M∞(t ′ ) = hW,p,M∞(t) + p(s),<br />

i.e. if 〈t, s〉 is infinite. Based on this, we see that for any t ≤ t ′ in (T, ≤), the<br />

19


number of p(s)-long edges (xi−1, xi) in a chain t = x0 → x1 → ... → xn = t ′ is<br />

the m(j) = hW,p,M∞(t ′ )(j) − hW,p,M∞(t)(j) where p(s) = xj. Furthermore, for any<br />

fixed t, then maximal number of p(s)-long edges in chains x0 → x1 → ... → xn = t<br />

is hW,p,M∞(t)(j) and this maximum is realized if x0 ∈ S.<br />

Corollary 3.4.1. The map t ↦→ MW,p,M∞(t) is an order preserving map from the<br />

reflection poset of (W, S) to the poset of generalized partitions.<br />

Proof. This follows directly from part 2 of Lemma 3.3.1.<br />

Remark 3.4.2. Notice that for t ≤ t ′ , we have that c(t) = c(t ′ ) by definition of the<br />

weak order.<br />

3.5 The Root Poset<br />

The reflections endowed with weak order as a poset is isomorphic to a subposet<br />

of another poset which we briefly introduce. We call this poset the root poset<br />

denoted (Φ, ≤). Define → on Φ by α → β if and only if there is some γ ∈ Π such<br />

that β = sγ(α) and (γ|α) < 0, i.e. β − α ∈ R>0γ. Then, we let ≤ be the reflexive<br />

transitive closure of →.<br />

We can consider the Hasse diagram of (Φ, ≤) as a directed graph with vertex set<br />

Φ and edges (α, β) whenever α → β. We label all edges by their type p(sγ) where<br />

sγ is uniquely determined by β = sγ(α) where γ ∈ Π. We say an edge is p(sγ)-long<br />

if it is labeled with p(sγ) and if (α|γ) ≤ −1, otherwise we call it p(sγ)-short. It<br />

is clear that the relations (Φ + , →) and (Φ + , ≤) are obtained by transporting the<br />

relations from (T, →) and (T, ≤) by means of the bijection sα ↦→ α from T to Φ + .<br />

We note that this correspondence also respects edge lengths and types, i.e. the<br />

edge (α, β) is p(s)-long if and only if the corresponding edge (sα, sβ) is p(s)-long.<br />

20


Using this correspondence, we translate the basic properties of the reflection<br />

poset into corresponding properties of the root poset. In particular, we introduce<br />

the following length function on the roots.<br />

⎧<br />

⎪⎨ lW,p,M∞(sα) if α ∈ Φ<br />

LW,p,M∞(α) =<br />

⎪⎩<br />

+<br />

−lW,p,M∞(sα) if α ∈ Φ− Lemma 3.5.1. 1. For α ≤ β in Φ, the interval [α, β] is finite.<br />

2. The map α ↦→ −α is an order-reversing bijection of (Φ, ≤) with itself.<br />

(3.5.1)<br />

3. Let α = α0 → α1 → ... → αn = β be a maximal chain in (Φ, ≤). Then for<br />

any xi ∈ X, the number of xi-long edges (αi−1, αi) in the chain (regarded<br />

as a path from α to β in the Hasse diagram) is equal to m(i) where m =<br />

LW,p,M∞<br />

(β)−LW,p,M∞ (α)<br />

.<br />

2<br />

Proof. First, 1 follows from the corresponding fact that intervals are finite in<br />

(T, ≤), and we remark that 2 is clearly true since −α → α for all α ∈ Π.<br />

Now, suppose that α, β ∈ Φ + , then sα ≤ sβ in (T, ≤). Let d := sα ≤ sα1 ≤<br />

· · · ≤ sαn−1 ≤ sβ be the maximal chain from sα to sβ. Then according to 3.4,<br />

hW,p,M∞(sβ)(i) − hW,p,M∞(sα)(i) is the number of xi-long edges in the chain. By<br />

definition of lW,p,M∞ we have that<br />

hW,p,M∞(sβ) − hW,p,M∞(sα) = lW,p,M∞(sβ) − c(sβ)<br />

2<br />

� �<br />

lW,p,M∞(sα) − c(sα)<br />

−<br />

2<br />

and by Remark 3.4.2 we know that since sα ≤ sβ then c(sα) = c(sβ) and the result<br />

follows in this case. Similarly, if α, β ∈ Φ − then −α, −β ∈ Φ + and −β ≤ −α and<br />

the proof follows from above.<br />

This leaves us with the case that α ∈ Φ − and β ∈ Φ + (notice we cannot have<br />

21


α ∈ Φ + and β ∈ Φ − ). Let αi ∈ Φ − and αi+1 ∈ Φ + , then since αi → αi+1 we<br />

must have −αi = αi+1 (since the only positive root made negative by sγ is γ for<br />

γ ∈ Π). In this case, we have that the edge (αi, αi+1) is p(sαi+1 )-long. Note, that<br />

since α ≤ αi+1 ≤ β we have c(sα) = c(sβ) = c(sαi+1 ) = p(sαi+1 ) by Remark 3.4.2<br />

and the definition of c. Then, using the previous argument, we know that the<br />

number of xi-long edges in α → α1 → · · · → αi is equal to the number of xi-long<br />

edges in −αi → −αi−1 → · · · → −α which is hW,pM∞(sα)(i). Also, the number of<br />

xi-long edges in αi+1 → αi+2 → · · · → β is hW,p,M∞(sβ)(i). Now we have the total<br />

number of xi long edges is given by m(i) where<br />

m : = hW,pM∞(sβ) + hW,pM∞(sα) + c(sαi+1 )<br />

= lW,p,M∞(sβ) − c(sβ)<br />

2<br />

= lW,p,M∞(sα) + lW,p,M∞(sβ)<br />

.<br />

2<br />

+ lW,p,M∞(sα) − c(sα)<br />

2<br />

3.6 Monotonicity Properties in Dominance Order on T<br />

+ c(sαi+1 )<br />

Recall that using the canonical bijection between Φ + and T , given by α ↦→ sα,<br />

we transport the dominance order on Φ + to dominance order on T . We have that<br />

sβ � sα if and only if β � α. This is equivalent to t ′ � t if for all w ∈ W we have<br />

l(wt) < l(w) implies that l(wt ′ ) < l(w); in this case, we say that t dominates t ′ .<br />

Lemma 3.6.1. Let r, t ∈ T .<br />

(a) t dominates r in T if and only if for some (respectively every) reflection<br />

group W ′ of W with t, r ∈ W ′ , t dominates r in dominance order on the set<br />

W ′ ∩ T of reflections of (W ′ , χ(W ′ )).<br />

22


(b) t dominates r if and only if either (1) t = r, or (2) 〈t, r〉 is infinite and<br />

hp,W (rtr) < hp,W (t).<br />

(c) t dominates r if and only if either (1) t = r, or (2) 〈t, r〉 is infinite and<br />

lp,W (rt) < lp,W (t) and hp,W (r) �> hp,W (t).<br />

(d) If r � t then t = r if and only if hp,W (r) = hp,W (t).<br />

(e) Let t ∈ T , xi ∈ X, and m = hW,p,M∞(t). Then, m(i) = |{t ′ ∈ T |t ′ ≺<br />

t; p(t ′ ) = xi}|.<br />

(f) If r � t, then hW,p,M∞(r) ≤ hW,pM∞(t).<br />

Proof. First of all, (a) follows from part 1 of Lemma 2.5.1 via the bijection between<br />

positive roots and T . We simply transfer the structure of dominance by this<br />

bijection and the result holds.<br />

Next, (e) is clear for finite dihedral groups. Now, suppose we have an infinite<br />

dihedral group (W, S) with S = {r, s}. Then, as in 2.5.1, we know that the<br />

dominance order on T in this case is the coarsest partial order satisfying<br />

and<br />

s ≺ srs ≺ srsrs ≺ srsrsrs ≺ · · ·<br />

r ≺ rsr ≺ rsrsr ≺ rsrsrsr ≺ · · · .<br />

Using this description, (e) is readily checked. Then, for general (W, S), for any<br />

23


xi ∈ X we have<br />

hW,p,M ∞(t)(i) = �<br />

W ′ ∈Mt<br />

W ′ ∈M∞<br />

= �<br />

W ′ ∈Mt<br />

W ′ ∈M∞<br />

= �<br />

W ′ ∈Mt<br />

W ′ ∈M∞<br />

hp,W ′(t)(i)<br />

|{t ′ ∈ W ′ ∩ T | t ′ ≺W ′ ,χ(W ′ ) t; p(t ′ ) = xi}|<br />

| t ′ ∈ W ′ ∩ T | t ′ ≺ t; p(t ′ ) = xi}|<br />

= |{t ′ ∈ T | t ′ ≺ t ′ ; p(t ′ ) = xi}|<br />

where the first equality is the definition, the second equality follows from the<br />

dihedral case, the third equality follows from (a), and the final equality follows<br />

from equation (2.3.1).<br />

Now, we can also prove (b) and (c) using this reduction to rank 2 as well. We<br />

note that by the description for dihedral reflection groups above, (b) and (c) hold<br />

by inspection. For the general case of (b) with arbitrary (W, S), suppose that<br />

t, r ∈ T with r � t and t �= r. Let W ′ = 〈r, t〉. Since r � t in W ′ , according to [4]<br />

we have that W ′ is infinite. The dihedral case shows that we have hp,W ′(rtr) <<br />

hp,W ′(t). This implies that rtr


≤∅ t or t ≤∅ r (by inspection of infinite dihedral case). Due to the fact that<br />

hp,W (r) �> hp,W (t), we must have r ≤∅ t and so hp,W ′(r) �> hp,W ′(t) by Lemma<br />

3.2.2; consequently, r � t in W ′ and thus in W by (a). Finally, (d) follows from<br />

(b) and (f) follows from (e) directly.<br />

3.7 A Partition of the Reflections<br />

At this point, we define certain subsets of the set of reflections, T , of a given<br />

Coxeter group (W, S). These sets are analogs of Tn from section 2.4. For any<br />

monomial m ∈ M, define Tm := {t ∈ T | hW,p,M∞(t) = m}.<br />

Proposition 3.7.1. Suppose S is finite. Then Tm is finite for any m ∈ M.<br />

Proof. This follows from Theorem 2.4.1 since Tm ⊂ T deg(m) and deg(m) ∈ N.<br />

Recall (M, ≤) from section 3.1. For any m ∈ M, let T≤m := �<br />

n≤m Tn.<br />

3.8 Recurrence Formulae<br />

We now further extend the results of [18]. Recall the reflection cocycle N :<br />

W → P(T ). For any m ∈ M and w ∈ W , we define Nm(w) := N(w) ∩ T≤m.<br />

Proposition 3.8.1. Let w ∈ W , s ∈ S, and m ∈ M.<br />

(a) If s �∈ Nm(w) then Nm(sw) = ({s} ∪ sNm(w)s) ∩ T≤m.<br />

(b) If s ∈ Nm(w) and m ≥ p(s), then Nm−p(s)(sw) = (sNm(w)s \ {s}) ∩<br />

T≤m−p(s)<br />

Proof. Recall that N(sw) = {s} + sN(w)s so the containment of the right hand<br />

side into the left hand side is clear for both parts. Now assume that s �∈ Nm(w).<br />

25


Then s �∈ N(w). Let t ∈ Nm(sw) = N(sw) ∩ T≤m = ({s} ∪ sN(w)s) ∩ T≤m, and<br />

then let t ′ = sts. If t = s then we clearly have t = s ∈ ({s} ∪ sNm(s)s) ∩ T≤m. If<br />

t �= s, then t ∈ sN(w)s and so t ′ ∈ N(w). If t = t ′ then we already have assumed<br />

t ′ = t ∈ T≤m and so this proves that t ′ ∈ Nm(w) so that t ∈ sNm(w)s and thus t<br />

is in the right hand side. If 〈s, t〉 is finite or if lp(t) = lp(t ′ )+2p(s) then by Lemma<br />

3.3.1 we know that hW,p,M∞(t ′ ) ≤ hW,p,M∞(t) ≤ m by assumption that t ∈ T≤m;<br />

thus, t ′ ∈ Nm(w) and t ∈ sNm(w)s as well as the right hand side. This only<br />

leaves the case where 〈s, t〉 is infinite and lp(t ′ ) = lp(t) + 2p(s). Let W ′ = 〈t, s〉.<br />

Since s ∈ S, then we know that χ(W ′ ) = {r, s} for some r ∈ T . By assumption,<br />

s �∈ N(t) and t �= 1 ∈ W ′ so we must have r ∈ N(t). Therefore, we know by the<br />

proof of Lemma 2.5 that t dominates r. We have assumed that t ∈ N(sw) and the<br />

dominance then implies that r ∈ N(sw) as well. Also, by our first assumption, we<br />

know that s ∈ N(sw) as well, but this forms a contradiction since we now have<br />

〈t, s〉 = 〈r, s〉 ⊂ N(sw) (see [17]). However N(sw) is finite, and we know that<br />

〈t, s〉 is infinite. Thus, we have a contradiction proving (a).<br />

To prove part (b), assume that s ∈ Nm(w). Thus, we have that s ∈ N(w)<br />

and so by symmetric difference we must have N(sw) = sN(w)s \ {s}. Let t ∈<br />

Nm−p(s)(sw) = (sN(w)s \ {s}) ∩ T≤m−p(s). Set t ′ := sts. Clearly we have that<br />

hW,p,M∞(t ′ ) ≤ m − p(s) + p(s) = m so t ′ ∈ T≤m. Also, t ′ ∈ N(w) by assumption<br />

that t ∈ sN(w)s \ {s}. Thus, t ′ ∈ Nm(w) and so t = st ′ s ∈ sNm(w)s \ {s}, and<br />

we already have assumed that t ∈ T≤m−p(s).<br />

3.9 Monoids with Poset Structure<br />

Let (P, ≤) be any poset and let M be an additively written commutative<br />

monoid. Suppose we have a partial order ≤M on M such that 0 ≤M m for all<br />

26


m ∈ M and m ≤ n if and only if m + k ≤ n + k for all m, n, k ∈ M. In this<br />

case, we say ≤M is compatible with the monoid structure on M. Furthermore,<br />

let q : P → M \ {0} be a map. Define an operator | − | : P fin (P ) → M in the<br />

following way, where P fin (P ) represents all of the finite subsets of P . For finite<br />

A ⊂ P , we have<br />

|A| := �<br />

q(a) (3.9.1)<br />

a∈A<br />

Remark 3.9.1. Notice that if M = N with the usual total order and q(x) = 1 for<br />

all x ∈ P then this is the usual cardinality of a finite set.<br />

An ideal I of P is a subset of P satisfying the following property: if x ∈ I<br />

and y ≤ x in P then y ∈ I as well. Any subset X ⊂ P is contained in a<br />

smallest (under inclusion) ideal of P called the ideal generated by X. We say<br />

an ideal, I, is principal if I is generated by a single element {x} with x ∈ P .<br />

We say that P is lower finite if every principal ideal is finite. If we have a lower<br />

finite poset, we define h ′ : P → M by h ′ (x) = |{z ∈ P | z < x}|. For any<br />

Γ ⊂ M define PΓ := {z ∈ P | h ′ (z) ∈ Γ}. We will consider certain cases like<br />

Γ =≤M m := {n ∈ M| n ≤M m} and Γ =


The fact that the right hand side implies the left is clear. So assume that |I| �≤M<br />

m. Now, we define B := {I ′ ⊆ I| I ′ is an ideal; |I ′ | �≤M m}. Since I is finite, B<br />

is clearly finite and nonempty since I ⊂ B. Let I0 ∈ B be a minimal element of B<br />

(with respect to inclusion). We note that |∅| = 0 and due to the fact that m ≥ 0<br />

for all m ∈ M, we must have |I0| �= ∅. Then, for any maximal element t ∈ I0, we<br />

know that J := I0 \ {t} is an ideal satisfying J � I0 ⊂ I. So by the minimality of<br />

I0, we must have |J| ≤M m. Similarly, let K = {x ∈ I0| {y ∈ I0| y > x} = ∅} be<br />

the set of maximal elements in I0, then J ′ = I0 \ K must also have |J ′ | ≤M m.<br />

So, for any t ∈ I0 we must have that h ′ (t) = |{s ∈ P | s < t}| ≤M |I0 \ K| =<br />

|J ′ | ≤M m. Therefore, t ∈ P≤Mm. So we have shown that I0 ⊂ I ∩ P≤Mm and<br />

thus |I ∩ P≤Mm| ≥M |I0|. If |I ∩ P≤Mm| ≤M m, then this would contradict the<br />

fact that |I0| �≤M m; thus, we have shown (a).<br />

To show (b), we will use (a). First, suppose that |I| = m. This implies<br />

that |I| �≤M n for all n ✁M m (where ✁M represents the covering relation for<br />

≤M). Using part (a), we have that |I ∩ P≤Mn| �≤M n for all n ✁M m. Putting<br />

these all together we get that |I ∩ ∪k✁MmP≤Mk| �≤M n for all n ✁M m. Thus,<br />

|I ∩ P


Suppose, for the sake of contradiction, that |I| >M m. This implies that |I| �≤M m<br />

and so by (a) we conclude that |I ∩ P≤Mm| �≤M m. However, this contradicts the<br />

assumption that I ∩ P m.<br />

5. lp,W (xy) = lp,W (x) + lp,W (y) − 2m iff |Nm(x −1 ) ∩ Nm(y)| = m and<br />

∪n✁m (Nn(x −1 ) ∩ Nn(y)) = Nm(x −1 ) ∩ Nm(y).<br />

Proof. First, we note that if w = r1...rn is a reduced expression, then N(w) =<br />

{ti| i ∈ {1, ..., n}} where ti := r1...ri−1riri−1...r1 (see [19]). From this we see clearly<br />

that p(ri) = p(ti) so that we have lp,W = |N(w)|. Also, we remark that N(x) is an<br />

ideal of T under dominance order by definition, and so for any finite family (xi)i∈I<br />

of elements of W , the intersection ∩iN(xi) is a finite ideal. From these remarks,<br />

(1) and (2) follow directly from the previous lemma.<br />

29


The points (3) and (5) also follow using the cocycle condition N(xy) = N(x)+<br />

xN(y)x −1 where + denotes symmetric difference. This condition along with the<br />

interpretation of N(x) in terms of lp,W (x) imply that<br />

lp,W (xy) = lp,W (x) + lp,W (y) − 2|N(x −1 ) ∩ N(y)|.<br />

Then let I = N(x −1 ) ∩ N(y) and the Lemma implies (3) and (5).<br />

Finally, if |Nm(x −1 ) ∩ Nm(y)| > m then by (3) we know lp,W (xy) �≥ lp,W (x) +<br />

lp,W (y) − 2m. However, we also know that |N(x −1 ) ∩ N(y)| = k ≥ |Nm(x −1 ) ∩<br />

Nm(y)| > m due to the containment of sets. So since<br />

we have that<br />

as required for (4).<br />

lp,W (x) + lp,W (y) + 2m < lp,W (x) + lp,W (y) + 2k<br />

lp,W (xy) = lp,W (x) + lp,W (y) − 2k < lp,W (x) + lp,W (y) − 2m<br />

We finish this section with a lemma that has the same ideas as the previous<br />

lemma.<br />

Lemma 3.10.2. Suppose that r, s ∈ S and x, y ∈ W with lp,W (xr) > lp,W (x) and<br />

lp,W (ys) > lp,W (y). If ysy −1 � xrx −1 , then lp,W (y −1 x) ≤ lp,W (x) + lp,W (y) − 2n<br />

where n = hW,p,M∞(ysy −1 ).<br />

Proof. We have s = y −1 ysy −1 y � y −1 xrx −1 y since lp,W (ys) > lp,W (y). Therefore,<br />

we know that lp,W (y −1 xr) = lp,W (y −1 x) + lp(r). Thus, by the exchange condition<br />

we know that lp,W (sy −1 xr) ≥ lp,W (y −1 x) + lp(r) − lp(s).<br />

30


Now, ysy −1 � xrx −1 and xrx −1 ∈ N(xr) so we must have ysy −1 ∈ N(xr).<br />

Also, ysy −1 ∈ N(ys) by definition. So<br />

ysy −1 ∈ N(xr) ∩ N(ys) = N(xr) ∩ N((sy −1 ) −1 ).<br />

Therefore, we see that if n = hW,p,M∞(ysy −1 ), we have |N((sy −1 ) −1 ) ∩ N(xr)| ≥<br />

n + p(s) using Lemma 3.6.1. As in the proof of (4) from the previous lemma, this<br />

implies that lp,W (sy −1 xr) ≤ lp,W (sy −1 ) + lp,W (xr) − 2(n + p(s)). Putting together<br />

the information we have, we get<br />

lp,W (y −1 x) ≤ lp,W (sy −1 xr) − lp(r) + lp(s)<br />

≤ lp,W (xr) + lp,W (sy −1 ) − 2(n + p(s)) − lp(r) + lp(s)<br />

≤ lp,W (x) + lp(r) + lp,W (y) + lp(s) − 2n − 2p(s) − lp(r) + lp(s)<br />

= lp,W (x) + lp,W (y) − 2n.<br />

31


CHAPTER 4<br />

F<strong>IN</strong>ITE STATE AUTOMATA<br />

In this chapter, we begin by introducing some standard terminology involving<br />

finite state automata. For the basic theory of using automata in groups, we use<br />

[21]. In [5], Brink and Howlett described a method of creating a finite state<br />

automata that would recognize any finitely generated Coxeter system. For any<br />

finite Coxeter system, the corresponding automata can be described by using the<br />

Cayley graph of the group.<br />

More recently, Dyer ([18]) has demonstrated a much larger class of finite state<br />

automata that can recognize many different subsets of Coxeter systems as well.<br />

We generalize the results of Dyer using generalized length functions from Chapter<br />

3. These length functions give rise to large class of finite state automata, which<br />

are more powerful than the family of automata introduced by Dyer. We then<br />

describe the corresponding languages (subsets of the free monoid S ∗ ) recognized<br />

by our automata. These automata will be used in Chapter 5 to show that many<br />

subsets of Coxeter systems are regular.<br />

4.1 Regular Languages<br />

We use this section to introduce the idea of a regular language of a monoid.<br />

Let M be a monoid. By a (left) M-set, we mean a set A along with a function<br />

32


M × A → A given by (m, a) ↦→ ma such that 1Ma = a for all a ∈ A if 1M is the<br />

identity of M and m1(m2a) = (m1m2)a for all a ∈ A and mi ∈ M. In this case,<br />

we say that M acts on A.<br />

A subset M ′ of M is called (monoid) regular if there is a left M-set A with<br />

A finite, an element a0 ∈ A and a subset Af ⊂ A with the property that M ′ =<br />

{m ∈ M| ma0 ∈ Af}. The following lemma is well known (see [21]). We include<br />

a proof here for completeness.<br />

Lemma 4.1.1. The family of regular subsets of M forms a Boolean subalgebra<br />

R(M) of P(M). If M is finitely (or countably) generated, then R(M) is count-<br />

able.<br />

Proof. Suppose that M1 and M2 are both monoid regular subsets of M. We will<br />

demonstrate that M c 1 = M \ M1, M1 ∪ M2 and M1 ∩ M2 are all monoid regular as<br />

well.<br />

Since Mi is monoid regular, there exists M-sets A1 and A2 with ai ∈ Ai and<br />

(Ai)f ⊂ Ai such that Mi = {m ∈ M| ma0 ∈ (Ai)f}. Now, A1 is finite and so<br />

A ′ f = A1 \ (A1)f is also finite. Then M c 1 = {m ∈ M| ma1 ∈ A ′ f } so M c 1 is regular.<br />

Next, let B = A1×A2 and let m(a, a ′ ) = (ma, ma ′ ) be the action of M on B. Note<br />

that B is a finite set. Then M1 ∩ M2 = {m ∈ M| ma1 ∈ (A1)f; ma2 ∈ (A2)f} =<br />

{m ∈ M| m(a1, a2) ∈ (A1)f × (A2)f}. Let Bf = (A1)f × (A2)f} and then M1 ∩ M2<br />

is regular. Finally, M1 ∪ M2 = {m ∈ M| ma1 ∈ (A1)f; or ma2 ∈ (A2)f}. Let<br />

B ′ f = (A1)f ×A2 ∪A1 ×(A2)f ⊂ B and then M1 ∪M2 = {m ∈ M| m(a1, a2) ∈ B ′ f }<br />

so M1 ∪ M2 is regular.<br />

Now, let S be a set, called an alphabet. Elements of the free monoid S ∗ on<br />

S are called words. We call subsets of S ∗ languages. Let N be an additively<br />

written commutative monoid and let map L ′ : S → N be a map. Then, for<br />

33


any word sn...s1 ∈ S ∗ , we can define the length ℓ : S ∗ → N by ℓ(sn...s1) :=<br />

L ′ (sn) + · · · + L ′ (s1). For example, if we let N = N and L ′ (s) = 1 for all s ∈ S,<br />

then ℓ(sn...s1) = n. Also, for any word w = sn...s1 ∈ S ∗ we define the transpose<br />

† : S ∗ → S ∗ by †(w) := w † := s1...sn.<br />

Now, if A is a finite S ∗ -set, then we can describe the action by restriction of<br />

the multiplication map S ∗ × A → A to a function µ : S × A → A. Typically, µ<br />

is called the transition function of the S ∗ -set A. Since the transition function µ<br />

can be an arbitrary function, we get a bijection between functions µ : S × A → A<br />

where A is a set and S ∗ -sets A. If S is a finite set, then we define a regular<br />

language on S to be a monoid regular subset of S ∗ . Therefore, a regular language<br />

is determined by a tuple (µ : S × A → A, a0, Af) where A is a finite set, µ is<br />

a function, a0 ∈ A and Af ⊆ A. The language determined by previous tuple is<br />

L = {m ∈ S ∗ | ma0 ∈ Af} where the S ∗ action is determined by µ.<br />

4.2 Automata<br />

We can equivalently describe the monoid regular subsets of S ∗ by the termi-<br />

nology of finite state automata. A finite state automaton for the alphabet S is a<br />

tuple (A, a0, µ : S × A → A, Af) such that A is a finite set with a distinguished<br />

element a0 called the initial state, a finite subset Af ⊆ A called the accepting<br />

states (or final states) and a transition function µ. The automaton reads a word<br />

w = sn...s1 from right to left, one letter at a time. The machine begins in the<br />

initial state a0; upon reading a letter si, the state moves from the current state,<br />

a, to the state µ(si, a). We say the automaton accepts a word w = sn...s1 if after<br />

reading w, the current state of the automaton is a member of Af. The set of words<br />

accepted by the automaton is called the language accepted by the automaton. It<br />

34


is clear that the language accepted by the automaton (A, a0, µ : S × A → A, Af)<br />

corresponds to the regular subset of S ∗ associated to A viewing A as an S ∗ -set.<br />

Example 4.2.1. Consider the affine Weyl group of type � A1 = 〈r, s | r 2 = s 2 = 1〉.<br />

We know that the set of reduced expressions in � A1 is given by<br />

{e, r, s, rs, sr, rsr, srs, rsrs, srsr, ...}.<br />

Below, we have a finite state automaton that recognizes the set of reduced expres-<br />

sions of words in � A1. A is the vertex set of the given graph, a0 is the blue vertex,<br />

and Af is the set of non-red vertices. µ : S × A → A is given by the labeled edges.<br />

��<br />

•<br />

��<br />

��<br />

��<br />

��<br />

��<br />

��<br />

��<br />

r ��<br />

��<br />

��<br />

��r<br />

��<br />

��<br />

��<br />

��<br />

��<br />

��<br />

��<br />

����<br />

• �<br />

r<br />

s<br />

�<br />

��<br />

•<br />

��<br />

��<br />

��<br />

�<br />

�<br />

��<br />

��<br />

��<br />

��<br />

�<br />

s �<br />

��<br />

��<br />

��s<br />

��<br />

��<br />

��<br />

��<br />

���<br />

��<br />

��<br />

�<br />

•<br />

4.3 Canonical Automata for Coxeter Systems<br />

At this point, we want to describe some finite state automata for Coxeter<br />

systems.<br />

Almost all objects defined in the rest of this chapter depend on the choice of p<br />

as in 3.2. We indicate when defining anything if it is dependent on p by including<br />

an appropriate subscript or superscript. However, when using the terminology,<br />

if there is no confusion about which function p we are referring to, we will leave<br />

off the corresponding subscript or superscript for ease of notation. Also, since<br />

(W, S) will be fixed, we let lp := lp,W . Recall, that lp takes values in M (which<br />

35<br />

r,s


depends on p) and that M has a partial order, which is described in section 3.1.<br />

In this chapter, we view M additively; therefore, M ∼ = N k and M ⊂ M ′ ∼ = Z k<br />

(see section 3.1).<br />

Let (W, S) be a Coxeter system and S ∗ be the free monoid on S. We denote<br />

the natural monoid homomorphism from S ∗ to W extending the identity map on<br />

S by ν : S ∗ → W . Let ℓp : S ∗ → M be the length of w = s1 · · · sn ∈ S ∗ given by<br />

ℓp(s1 · · · sn) = p(s1) + · · · + p(sn). Note ℓp(w) ≥ lp,W (ν(w)).<br />

We associate to (W, S) an S ∗ -set A = A(W,S). As a set we have<br />

A(W,S,p) := A = {∞} ∪ {(m, A)| m ∈ M, A ⊆ T≤m}.<br />

The S ∗ -action can be described by its restriction to a function S × A → A<br />

given by (s, ∞) ↦→ ∞ and<br />

⎧<br />

⎪⎨<br />

(s, (m, A)) ↦→<br />

⎪⎩<br />

∞, if p(s) �≤ m and s ∈ A<br />

� �<br />

m − p(s), (sAs \ {s}) ∩ T≤(m−p(s)) , if p(s) ≤ m and s ∈ A<br />

(m, (sAs ∪ {s}) ∩ T≤m) , if s �∈ A<br />

We may define sub-S ∗ -sets An = A(W,S,p,n) of A for n ∈ M defined by<br />

An := {∞} ∪ {(m, A)| m ≤ n, A ⊆ T≤m}<br />

We know that if S is finite then An is a finite S ∗ -set.<br />

36


Lemma 4.3.1. Let x ∈ W , w ∈ S ∗ , and m ∈ M. Then w · ∞ = ∞ and<br />

⎧<br />

⎪⎨ ∞, if k �≤ m<br />

w · (m, Nm(x)) ↦→<br />

⎪⎩ (m − k, Nm−k(ν(w)x)) , if k ≤ m<br />

in A , where k := lp(x)+ℓp(w)−lp(ν(w)x)<br />

2<br />

Proof. This follows by induction on ℓp(w) since it is clearly true for each w =<br />

s ∈ S. Then suppose that w = sw ′ and the result is known for sw ′ . Then, by<br />

definition of the S-action on pairs, the result is clear.<br />

Finally, from the previous lemma, we can define<br />

M = M(W,S,p) := {∞} ∪ {(m, Nm(x))| m ∈ M, x ∈ W }<br />

and M(W,S,p,n) = Mn = M ∩ An for n ∈ M, which are all sub-S ∗ -sets of A . We<br />

call these the canonical and n-canonical S ∗ -sets for (W, S) respectively.<br />

4.4 Some Subsets of S ∗<br />

Let Aa,b := {w ∈ S ∗ | w · a = b} for any a, b ∈ M. By Lemma 4.3.1, we see<br />

that A∞,∞ = S ∗ and A∞,x = ∅ for x �= ∞. We now intend to describe the sets<br />

Aa,b. To do this, we introduce some subsets of S ∗ in the following way.<br />

To begin with, for any word w ∈ S ∗ , we define<br />

Lp(w) := ℓp(w) − lp(ν(w))<br />

2<br />

∈ M.<br />

For any m ∈ M, we can then define S ∗ m := S ∗ p,m = {w ∈ S ∗ | Lp(w) = m}. The<br />

elements of S ∗ m will be called the m-nonreduced words of (W, S). If S is finite,<br />

37


then for any w ∈ W , there are only finitely many m-nonreduced expressions of w,<br />

namely the elements of the set ν −1 (w) ∩ S ∗ m.<br />

Next, for x, y ∈ W and n ∈ M ′ , we define more general sets<br />

F p x,y,n := {w ∈ S ∗ | lp(yν(w)x) = lp(y) + ℓp(w) + lp(x) − 2n} (4.4.1)<br />

and also define, for m ∈ M ′<br />

With this notation, we see that S ∗ m = F1,1,m.<br />

F p x,y,n,m := S ∗ m ∩ F p<br />

x,y,n+m. (4.4.2)<br />

Remark 4.4.1. It is clear from the definition that F p x,y,n = ∅ if n �∈ M, i.e. if<br />

n < 0.<br />

Example 4.4.2. For s ∈ S, we have that the set F1,s,lp(s) is given by<br />

{w ∈ S ∗ |lp(sw) = lp(s) + ℓp(w) − 2lp(s) = ℓp(w) − lp(s)},<br />

which can also be described by<br />

[ν −1 ({w ∈ W | lp(sw) < lp(w)}) ∩ S ∗ 0] ˙∪[ν −1 ({w ∈ W | lp(sw) > lp(w)}) ∩ S ∗ lp(s)].<br />

Next for any x ∈ W , k ∈ M ′ and finite subsets A, B of T , we let<br />

H p<br />

x,k,A,B := {w ∈ S∗ | lp(ν(w)x) = ℓp(w)+lp(x)−2k; N(ν(w)x)∩A = B} (4.4.3)<br />

so it is clear the H p<br />

x,k,A,B<br />

such that N(y) ∩ A = B.<br />

= ∅ unless k ∈ M, B ⊆ A, and there exists a y ∈ W<br />

38


According to Lemma 4.3.1, we see that for m ∈ M and x ∈ W we get<br />

and then for m, n ∈ M,<br />

A(m,Nm(x)),∞ = �<br />

k�≤m<br />

Fx,1,k<br />

⎧<br />

⎪⎨ ∅ if n �≤ m<br />

A(m,Nm(x)),(n,B) =<br />

⎪⎩ Hx,m−n,T≤n,B if n ≤ m<br />

where both sides are empty unless B is of the form Nn(y) for some y ∈ W .<br />

4.5 The Boolean Algebra of Regular Sets of S ∗<br />

(4.4.4)<br />

At this point, we assume that S is finite. Then, for a, b ∈ M, we can view<br />

Aa,b in the following way. We choose m large enough so that a, b ∈ Mm. Then<br />

Aa,b is the regular language accepted by an automaton where Mm is the state set,<br />

the transition function described in 4.3.1, initial state a, and single accept state<br />

b. At this stage, we characterize the Boolean subalgebra of S ∗ generated by all<br />

Aa,b where a �= ∞, b �= ∞.<br />

Lemma 4.5.1. The following Boolean subalgebras of S ∗ are all equal:<br />

1. The Boolean subalgebra Dp of P(S ∗ ) generated by all sets Aa,b where a, b ∈<br />

M \ {∞}.<br />

2. The Boolean subalgebra D ′ p of P(S ∗ ) generated by all sets Hx,k,A,B.<br />

3. The Boolean subalgebra D ′′<br />

p of P(S ∗ ) generated by all sets Fx,y,k,m.<br />

4. The Boolean subalgebra D ′′′<br />

p of P(S ∗ ) generated by all sets Fx,y,k.<br />

39


Moreover, Dp is stable under the automorphism † : S ∗ → S ∗ given by †(sn....s1) =<br />

s1...sn.<br />

Proof. According to equation (4.4.4), we know that Dp is generated by all the sets<br />

Hx,k,T≤n,B where x, y ∈ W , n, k ∈ M and B ⊂ T is finite. Suppose that k ∈ M,<br />

x ∈ W , and A, B ⊆ T are finite sets. We choose an element n ∈ M such that<br />

A ⊆ T≤n. Then<br />

Hx,k,A,B = �<br />

B ′ ⊆T≤n<br />

B ′ ∩A=B<br />

Hx,k,T≤n,B ′<br />

and so Hx,k,A,B is generated by a finite union of Aa,b for some a, b ∈ M \ {∞}.<br />

Thus Dp = D ′ p.<br />

Next, we note that for w ∈ Fx,y,k,m we get (since w ∈ S ∗ m)<br />

lp(yν(w)x) = lp(y) + ℓp(w) + lp(x) − 2(k + m)<br />

= lp(y) + lp(ν(w)) + 2m + lp(x) − 2(k + m)<br />

and so lp(yν(w)x) = lp(y) + lp(ν(w)) + lp(x) − 2k. This implies that Fx,y,k,m = ∅<br />

unless 0 ≤ k ≤ lp(y) + lp(x). This shows that S ∗ m = ∪0≤k≤lp(x)+lp(y)Fx,y,k,m and<br />

that we also have<br />

Fx,y,k =<br />

�<br />

k−lp(x)−lp(y)≤m≤k<br />

Fx,y,k−m,m<br />

thus showing that D ′′<br />

p = D ′′′<br />

p since it is clear that Fx,y,k,m = S ∗ m ∩ Fx,y,k+m =<br />

F1,1,m ∩ Fx,y,k+m.<br />

Now, we show that D ′′′<br />

p ⊆ D ′ p. For fixed x ∈ W and A ⊆ T finite, we have<br />

S ∗ = �<br />

�<br />

k∈M B⊆A<br />

Hx,k,A,B<br />

Recall from section 3.10 the terminology |A| ∈ M for A ⊆ T finite. Let y ∈ W .<br />

40


Then for w ∈ Hx,k,A,B we have<br />

lp(yν(w)x) = lp(y) + lp(ν(w)x) − 2|N(ν(w)x) ∩ N(y −1 )|<br />

From this we can see that<br />

= lp(x) + ℓp(w) + lp(y) − 2k − 2|N(ν(w)x) ∩ N(y −1 )|.<br />

Fx,y,m =<br />

�<br />

�<br />

k,n∈M:k+n=m B⊆N(y−1 ):|B|=n<br />

H x,k,N(y −1 ),B.<br />

This union is finite since there are only finitely many pairs k, n in M less than m<br />

and N(y −1 ) is a finite set, and so D ′′′<br />

p ⊆ D ′ p.<br />

have<br />

So we now must show that D ′ p ⊆ D ′′′ p. Let x, y ∈ W and w ∈ S ∗ , then we<br />

Fx,1,k ∩ Fx,y,m = {w ∈ W | lp(ν(w)x) = ℓp(w) + lp(x) − 2k and<br />

lp(yν(w)x) = lp(y) + ℓp(w) + lp(x) − 2m},<br />

(4.5.1)<br />

and by replacing ℓp(w) + lp(x) = lp(ν(w)x) + 2k in the second equation (using the<br />

first equation) we get<br />

Fx,1,k ∩ Fx,y,m = {w ∈ W | lp(ν(w)x) = ℓp(w) + lp(x) − 2k and<br />

lp(yν(w)x) = lp(ν(w)x) + lp(y) − 2m + 2k}.<br />

(4.5.2)<br />

Now, we know that lp(ν(w)x) ≤ lp(y) + lp(yν(w)x) ≤ lp(y) + lp(y) + lp(ν(w)x)<br />

consequently 0 ≤ lp(y)+lp(yν(w)x)−lp(ν(w)x) ≤ 2lp(y). Substituting appropriate<br />

equations from (4.5.1) and (4.5.2), we note that Fx,1,k ∩ Fx,y,m is empty unless<br />

0 ≤ 2lp(y) − 2m + 2k ≤ 2lp(y) and this implies that k ≤ m ≤ k + lp(y). We also<br />

have ∪m∈MFx,y,m = S ∗ . Now, let y from above be a given reflection y = t. Then<br />

41


we have that the following sets<br />

and<br />

Lx,1,k,t := {w ∈ Fx,1,k| t ∈ N(ν(w)x)} = Fx,1,k ∩<br />

L ′ x,1,k,t := {w ∈ Fx,1,k| t �∈ N(ν(w)x)} = Fx,1,k ∩<br />

�<br />

m: k≤m≤k+lp(t)<br />

0≤lp(t)≤2m−2k<br />

�<br />

m: k≤m≤k+lp(t)<br />

lp(t)≥2m−2k∈M<br />

Fx,t,m<br />

Fx,t,m<br />

are clearly in D ′′′<br />

p since these intersections are finite. If B ⊆ A, we thus get<br />

Hx,k,A,B =<br />

� �<br />

t∈B<br />

Lx,1,k,t<br />

� ⎛<br />

∩ ⎝ �<br />

t∈A\B<br />

L ′ x,1,k,t<br />

and as we have noted before, Hx,k,A,B = ∅ if B �⊆ A. Therefore, Hx,k,A,B is in D ′′′<br />

p ,<br />

which is what we were trying to show.<br />

Finally, suppose w ∈ Fx,y,m for some m. Now, we have that ν(†(w)) = ν(w) −1 .<br />

Since lp(yν(w)x) = lp(y) + ℓp(w) + lp(x) − 2m then we can reverse this to see that<br />

we must also have lp(x −1 ν(†(w))y −1 ) = lp(x −1 ) + ℓp(†(w)) + lp(y −1 ) − 2m and so<br />

†(w) ∈ F y −1 ,x −1 ,m, which shows that Dp is stable under †.<br />

4.6 The Reflection Cocycle and Automata<br />

Recall that for A ⊆ T a finite subset, we write |A| = �<br />

t∈A p(t). Thus, we<br />

note that lp(w) = |N(w)|. Now, we state a lemma that describes the essence<br />

of the previous proof. The problem of determining, for all w ∈ W , the value of<br />

lp(xw)−lp(w) for all x in some given finite subset of W is equivalent to finding, for<br />

42<br />

⎞<br />

⎠ ,


all w ∈ W , the intersection N(w) ∩ A for some given finite subset A of reflections.<br />

Recall that M ′ = Z k if M = N k .<br />

Lemma 4.6.1. (a) Let x1, ..., xn be elements of W , and l1, ..., ln ∈ M ′ . Let<br />

A := ∪ n i=1N(x −1<br />

i ). Then for w ∈ W , lp(xiw)−lp(w) = li for all i ∈ {1, ..., n}<br />

if and only if |(N(w) ∩ A) ∩ N(x −1 lp(xi)−li<br />

i )| = 2<br />

(b) Let A be a finite subset of T . Then for w ∈ W , we have<br />

for all i ∈ {1, ..., n}.<br />

N(w) ∩ A = {t ∈ A| lp(tw) = lp(w) − k for some k ∈ M≤lp(t)}<br />

Proof. Recall that N(xy) = N(x) + xN(y)x −1 = x(N(x −1 ) + N(y))x −1 . This<br />

implies that<br />

lp(xy) = lp(x) + lp(y) − 2|N(x −1 ) ∩ N(y)|<br />

where |−| is as above. Therefore, for w ∈ W , if xi ∈ W , we have lp(xiw)−lp(w) =<br />

lp(x) − 2|N(x −1<br />

i ) ∩ N(w)|. Since N(w) ∩ N(x−1 ) = (N(w) ∩ A) ∩ N(x −1<br />

i ), then<br />

lp(xiw) − lp(w) = li if and only if lp(xi) − 2|(N(w) ∩ A) ∩ N(x −1<br />

i )| = li, and (a)<br />

follows.<br />

Then, (b) follows since N(w) = {t ∈ T | lp(tw) < lp(w)}. We have lp(tw) =<br />

lp(t) + lp(w) − 2|N(t) ∩ N(w)|, we see that lp(t) − 2|N(t) ∩ N(w)| = −k for some<br />

k ∈ M.<br />

43


CHAPTER 5<br />

REGULAR SUBSETS OF <strong>COXETER</strong> <strong>GROUPS</strong><br />

This chapter is devoted to the definition of various notions of regularity in<br />

Coxeter systems. We introduce three different types of regularity. The strongest<br />

type of regularity, p-complete regularity, will be examined in detail, and we will<br />

describe many subsets of Coxeter systems that are p-completely regular. The set<br />

of m-nonreduced expressions of elements of any p-completely regular subset of W<br />

is a regular language in S ∗ for all m ∈ M. This implies that the multivariate<br />

Poincaré series of p-completely regular subsets are rational, providing a natural<br />

explanation (and strengthening) of known rationality results on Poincaré series of<br />

many natural subsets of W . We also use the results in this section to describe<br />

regularity of subsets in products of Coxeter systems.<br />

5.1 Boolean Algebra of Regular Sets in W<br />

For any x, y ∈ W and m ∈ M ′ , we can now define a subset<br />

Wx,y,m := W p x,y,m = {w ∈ W | lp(ywx) = lp(y) + lp(w) + lp(x) − 2m} ⊂ W.<br />

We have that W =<br />

�<br />

0≤k≤lp(x)+lp(y)<br />

Wx,y,k. Notice that Fx,y,k,m = ν −1 (Wx,y,k) ∩ S ∗ m<br />

so that Dp is the Boolean subalgebra of P(S ∗ ) generated by all intersections<br />

ν −1 (Wx,y,k) ∩ S ∗ m.<br />

44


Remark 5.1.1. Again, it is clear that Wx,y,m = ∅ if m �∈ M. This allows us to<br />

only consider sets Wx,y,m with m ≥ 0.<br />

Let Bp(W ) denote the Boolean subalgebra of P(W ) generated by all the sets<br />

Wx,y,k with x, y ∈ W and k ∈ M. Then Dp consists of all the sets ∪m∈Mν −1 (Bm)∩<br />

S ∗ m where Bm ∈ Bp(W ) for all m and Bm is either ∅ for all but finitely many<br />

m or Bm is W for all but finitely many m. Thus, we see that Dp is uniquely<br />

determined by the sets Bp(W ) together with the set of m-nonreduced expressions<br />

of w, for all m ∈ M and w ∈ W .<br />

Example 5.1.2. Let J ⊆ S be a spherical subset, i.e. WJ is a finite standard<br />

parabolic subgroup. Then, let wJ be the longest element of WJ and suppose<br />

lp(wJ) = k. Then D(J) := {w ∈ W | DL(w) = J} ∈ Bp(W ) since it is the set<br />

{w ∈ W | lp(wJw) = lp(wJ) + lp(w) − 2k = lp(w) − lp(wJ)} = W1,wJ ,k. This set is<br />

known as the descent class associated to J.<br />

5.2 Coxeter Groups are Automatic<br />

The following result is due to Brink and Howlett, [5], in their proof that Coxeter<br />

groups are automatic.<br />

Theorem 5.2.1. Fix a total order of S and give S ∗ the corresponding reverse<br />

lexicographic order �L. Then for w ∈ W define ˙w as the minimal (under �L)<br />

element of the non-empty and finite set ν −1 (w) ∩ S ∗ 0. Then ˙ W = { ˙w| w ∈ W } is<br />

a regular subset of S ∗ .<br />

5.3 Different Types of Regularity<br />

Fix any map µ : W → S ∗ with the following properties:<br />

45


(i) For all w ∈ W , µ(w) ∈ S ∗ 0<br />

(ii) ν ◦ µ = IdW .<br />

(iii) µ(W ) is a regular subset of S ∗ .<br />

Properties (i) and (ii) require that for w ∈ W , we have µ(w) is a reduced<br />

expression of W , and we know such a map exists due to Theorem 5.2.1.<br />

We say that a subset A of W is µ-regular if µ(A) is a regular subset of S ∗ .<br />

Definition 5.3.1. 1. A subset A of W is weakly regular if there is some map<br />

w ↦→ ˙w : A → S ∗ 0 such that ν( ˙w) = w for all w ∈ A and { ˙w| w ∈ A} is a<br />

regular subset of S ∗ .<br />

2. A subset of W is regular (resp. p-strongly regular) if ν −1 (A)∩S ∗ m is a regular<br />

subset of S ∗ for m = 0 (resp. for all m ∈ M).<br />

3. A subset of W is called p-completely regular if A ∈ Bp(W ).<br />

Example 5.3.2. Due to example 5.1.2, D(J) is p-completely regular.<br />

Lemma 5.3.3. (a) Any p-strongly regular subset of W is regular.<br />

(b) If A is a regular subset of W , then A is µ-regular for any map µ satisfying<br />

conditions (i)-(iii) above.<br />

(c) If A is µ-regular for some µ, then A is weakly regular.<br />

Proof. Parts (a) and (c) follow directly from the definitions. To show part (b),<br />

we assume that A is a regular subset of W . This means that ν −1 (A) ∩ S ∗ 0 is a<br />

regular subset of S ∗ . Additionally, since µ satisfies (i)-(iii), we know that µ(W ) is<br />

a regular subset of S ∗ as well. Now, (b) follows since µ(A) = µ(W )∩(ν −1 (A) ∩ S ∗ 0)<br />

which is a finite intersection of regular sets and is thus regular.<br />

46


5.4 p-Regularity for Different p<br />

Suppose that p : T → Xp and p ′ : T → Xp ′ are two different functions<br />

satisfying the properties of section 3.2 with corresponding sets of indeterminates<br />

Xp and Xp ′ respectively. Additionally, let Mp and Mp ′ be the corresponding<br />

sets of monomials for Xp and Xp ′. Since Mp and Mp ′ are the free commutative<br />

monoids on Xp and Xp ′ respectively, then for any map ϕ : Xp → Xp ′ we get<br />

a corresponding map, denoted ˜ϕ, with ˜ϕ : Mp → Mp ′ given by the universal<br />

property. We note that deg(m) = deg( ˜ϕ(m)) for any m ∈ Mp.<br />

Proposition 5.4.1. Let ϕ : Xp → Xp ′ be a surjective map such that ϕ(p(t)) =<br />

p ′ (t) for all t ∈ T . Then the following are true:<br />

1. Dp ′ ⊂ Dp, and<br />

2. Bp ′(W ) ⊂ Bp(W ).<br />

Proof. For q ∈ {p, p ′ }, Dq is generated by the sets F q x,y,m with x, y ∈ W and<br />

m ∈ Mq according to Lemma 4.5.1. Now, since ϕ is surjective then ˜ϕ is surjective<br />

as well. Let m ∈ Mp ′. Then define Q := ˜ϕ−1 (m). By surjectivity, Q �= ∅ and<br />

since ˜ϕ preserves degree, |Q| < ∞. Then, it follows that<br />

F p′<br />

x,y,m = �<br />

k∈Q<br />

F p<br />

x,y,k ∈ Dp<br />

and therefore Dp ′ ⊂ Dp, proving 1. Then, 2 is analogous replacing F q x,y,m by<br />

W q x,y,m.<br />

47


5.5 Some Properties of p-Completely Regular Sets<br />

sets.<br />

In this section we describe some of the basic properties of p-completely regular<br />

Proposition 5.5.1. (a) Any p-completely regular set is p-strongly regular (and<br />

thus also regular and µ-regular for any µ satisfying (i)-(iii) in section 5.3.<br />

(b) For any finite subsets B, C ⊆ A of T the subset<br />

GA,B,C := {w ∈ W | N(w) ∩ A = B, N(w −1 ) ∩ A = C}<br />

of W is p-completely regular.<br />

(c) If A ∈ Bp(W ) then A −1 := {w −1 | w ∈ A} ∈ Bp(W ).<br />

(d) Bp(W ) is closed under action by automorphisms of (W, S) that commute<br />

with p, i.e. automorphisms satisfying θ(p(t)) = p(θ(t)) for all t ∈ T .<br />

(e) Bp(W ) is closed under left and right multiplication by elements of W .<br />

(f) Bp(W ) contains all finite subsets of W .<br />

(g) Bp(W ) is countable.<br />

Proof. Suppose that A is p-completely regular. Then as we have seen ν −1 (A) ∩<br />

S ∗ m ∈ Dp for all m ∈ M. But Dp contains only regular languages by Lemma<br />

4.5.1 part (1), which proves (a). To prove (b), we let Wt := {w ∈ W | lp(tw) <<br />

lp(w)} = � lp(t)−c(t)<br />

lp(w)} =<br />

�<br />

W1,t,k ∈ Bp(W ) since both are finite unions. We also have sets<br />

0≤k< lp(t)+c(t)<br />

2<br />

Xt := {w ∈ W | lp(wt) < lp(w)} and X ′ t := {w ∈ W | lp(wt) > lp(w)}, and both<br />

48


are in Bp(W ) using finite unions of appropriate sets Wt,1,k. Now, since A is finite,<br />

we get<br />

YB := �<br />

Wt ∩ �<br />

t∈B<br />

t∈A\B<br />

YC := �<br />

Xt ∩ �<br />

t∈C<br />

t∈A\C<br />

and then GA,B,C := YB ∩ YC. In particular, this process describes an application,<br />

in detail, of Lemma 4.6.1.<br />

Next, (c) holds since (Wx,y,k) −1 = {w −1 | lp(ywx) = lp(y)+lp(w)+lp(x)−2k} =<br />

{w ∈ W | lp(yw −1 x) = lp(y) + lp(w) + lp(x) − 2k} = {w ∈ W | lp(x −1 wy −1 ) =<br />

lp(x −1 ) + lp(w) + lp(y −1 ) − 2k} = W y −1 ,x −1 ,k.<br />

Now, for any automorphism, θ, of (W, S) commuting with p, we have lp(w) =<br />

lp(θ(w)) for all w ∈ W . Then, we see that if w ∈ Wx,y,k then lp(θ(ywx)) =<br />

lp(ywx) = lp(y) + lp(x) + lp(w) − 2k = lp(θ(y)) + lp(θ(x)) + lp(θ(w)) − 2k so that<br />

θ(Wx,y,k) ⊂ Wθ(x),θ(y),k. Using the same argument with the automorphism θ −1<br />

W ′<br />

t<br />

demonstrates that in face θ(Wx,y,k) = Wθ(x),θ(y),k, proving (d).<br />

Next, (e) holds using the following observations. Let a ∈ W and b ∈ W<br />

and x, y ∈ W . Then lp(ya −1 ) = lp(y) + lp(a −1 ) − 2P for some P ≥ 0 and<br />

lp(b −1 x) = lp(x) + lp(b −1 ) − 2Q for some Q ≥ 0. If w ∈ Wx,y,k we have lp(ywx) =<br />

lp(y) + lp(w) + lp(x) − 2k. Then awb ∈ W satisfies lp(ya −1 awbb −1 x) = lp(y) +<br />

lp(w)+lp(x)−2k = lp(ya −1 )−lp(a −1 )+2P+lp(w)−lp(b −1 )+lp(b −1 x)+2Q−2k =<br />

lp(ya −1 ) + lp(awb) − 2n + lp(b −1 x) + 2P + 2Q − 2k where 0 ≤ n ≤ lp(a) + lp(b)<br />

and awb ∈ W b −1 ,a −1 ,n. Therefore we have<br />

aWx,y,kb =<br />

�<br />

0≤n≤lp(a)+lp(b)<br />

X ′ t<br />

(W b −1 x,ya −1 ,k−P−Q+n ∩ W b −1 ,a −1 ,n)<br />

49


where we recall that Wu,v,k = ∅ if k < 0.<br />

Finally, (f) holds since we have GS,∅,∅ = {1} ∈ Bp(W ) by (b). Then for any<br />

w ∈ W , w{1} = {w} ∈ Bp(W ) by (e), and so any finite union of singleton<br />

elements of W is in Bp(W ). Then, (g) follows from the nature of the fact that<br />

there are countably many sets Wx,y,k and we can only take finite unions and<br />

intersections.<br />

Example 5.5.2. For any J ⊆ S spherical, the subgroup WJ is p-completely<br />

regular by (f). Since W = W1,1,0, then W is p-completely regular. If (W, S) is<br />

finite, we can easily demonstrate that (W, S) is regular by letting the Cayley graph<br />

of (W, S) be the corresponding S ∗ -set.<br />

Remark 5.5.3. Given a completely regular set A ⊆ W , described explicitly as<br />

a finite union of finite intersections of the sets W p<br />

x,y,k , we could construct an<br />

automaton on S accepting ν −1 (A) ∩ S ∗ n for any n ∈ M using Lemma 4.5.1. In<br />

addition, following from Lemma 5.3.3, we could construct an automaton on S<br />

accepting µ(A) for any µ satisfying the (i)-(iii) from 5.3.3.<br />

5.6 Poincaré Series and Examples of Regular Subsets of W<br />

The previous result leads to some results about the rationality of Poincaré<br />

series of subsets of Coxeter groups. First we recall the definition of a Poincaré<br />

series.<br />

Definition 5.6.1. Let u be an indeterminate and consider the power series ring<br />

Z[[u]]. For a subset A ⊆ W we define the Poincaré series of A to be<br />

P (A; W ) = �<br />

u deg(lp(w)) .<br />

w∈A<br />

50


If we consider the multiplicative version of M generated by X, then we can con-<br />

sider the completion Z[[X]] of Z[X]. We define the multivariate Poincaré series of<br />

A to be<br />

Pp(A; W ) = �<br />

lp(w).<br />

w∈A<br />

The following examples are previously known in the case where lp = l is the<br />

standard length; we use our results to show the examples are p-completely regular<br />

for general p as well.<br />

Example 5.6.2. 1. For any finitely generated reflection subgroup W ′ and any<br />

standard parabolic subgroup WJ of (W, S), any double coset W ′ wWJ has<br />

a unique element of minimal length determined by the conditions N(w) ∩<br />

χ(W ′ ) = ∅ and N(w −1 ) ∩ J = ∅ where χ(W ′ ) is the set of canonical Coxeter<br />

generators of (W, S). In particular the set of these shortest coset represen-<br />

tatives, denoted W ′ \ W/WJ is thus given by the following:<br />

W ′ \ W/WJ = �<br />

C⊂χ(W )<br />

Gχ(W ′ ),∅,C ∩ �<br />

B⊂J<br />

GJ,B,∅<br />

and so is p-completely regular. This applies to shortest coset representatives<br />

W ′ \ W of W and to shortest double coset representatives WK \ W/WJ of<br />

standard parabolic subgroups.<br />

2. The weak right order ≤ on W is defined by x ≤ y if N(x) ⊆ N(y) or<br />

equivalently if lp(x −1 y) = lp(y)−lp(x). Thus, if we fix x ∈ W with lp(x) = m,<br />

the set<br />

W 1,x −1 ,m = {y ∈ W | lp(x −1 y) = lp(y) + lp(x) − 2lp(x)} = {y ∈ W | y ≥ x}<br />

51


is the upper rays in weak right order, and so this set is p-completely regular.<br />

A similar argument shoes that the upper rays in weak left order are also<br />

p-completely regular.<br />

3. As we have seen in the proof of Proposition 5.5.1 the sets {w ∈ W | lp(tw) <<br />

lp(w)} and {w ∈ W | lp(wt) < lp(w)} are p-completely regular for fixed<br />

t ∈ T . These spaces (and their complements) are known as left and right<br />

half spaces respectively. It follows that any finite intersection and union of<br />

left (or right) half spaces is p-completely regular. See Figure 5.1 for example.<br />

All of the sets in the previous examples thus have rational Poincaré series and<br />

multivariate Poincaré series due to a well-known result, called the transfer matrix<br />

method (see [2] or [28]), which states that regular languages have rational Poincaré<br />

series.<br />

5.7 More Regular Subsets of W<br />

The sets described in this section are shown to be p-completely regular in [18]<br />

for the case where p : T → N with p(t) = 1 for all t. We now prove that these<br />

sets are also p-completely regular for general p.<br />

Recall that W acts as a group of permutations on the set T × {±1} such that<br />

for s ∈ S, ɛ ∈ {±1}, and t ∈ T \{s} we have s(s, ɛ) = (s, −ɛ) and s(t, ɛ) = (sts, ɛ).<br />

We note that as a W × {±1}-set, T × {±1} is isomorphic to the standard root<br />

system Φ, with Φ + corresponding to T × {1} and simple roots Π corresponding to<br />

S × {1}. With this terminology, we have the following p-completely regular sets.<br />

Proposition 5.7.1. 1. For u, v ∈ T , W (u, v) = {w ∈ W | wuw −1 = v} is<br />

p-completely regular.<br />

52


Figure 5.1. This diagram shows the tiling of the plane corresponding to<br />

�B2 as in Figure 3.1. Each of the gray regions is an intersection of two<br />

half spaces, and thus each is p-completely regular for any p. Then, the<br />

total gray region is a union of two p-completely regular sets and so is<br />

p-completely regular itself. Therefore, this set also has a rational multivariate<br />

Poincaré series.<br />

53


2. For α, β ∈ T × {±1}, W (α, β) = {w ∈ W | w(α) = β} is p-completely<br />

regular.<br />

3. For finite sets Γ, ∆ ⊂ T ×{±1}, the set {w ∈ W | w(Γ) = ∆} is p-completely<br />

regular.<br />

4. For J, K ⊆ S, {w ∈ W | wJw −1 = K} is p-completely regular.<br />

5. For J, K ⊆ S, {w ∈ W | w((WJ ∩T )×{1}) = (WK ∩T )×{1}} is p-completely<br />

regular.<br />

Proof. We first prove 1 for u = r and v = s with r, s ∈ S. In this case, we have<br />

W (r, s) = {w ∈ W | wrw −1 = s} = {w ∈ W | wr = sw}. Now, W (r, s) will be<br />

empty (and thus p-completely regular) unless r is conjugate to s. So, if r and<br />

s are conjugate, then lp(r) = lp(s) and according to the exchange condition, we<br />

have that W (r, s) = Z1 ∪ Z2 where<br />

and<br />

Z1 = {w ∈ W | lp(wr) = lp(w) − lp(r) = lp(sw); lp(swr) = lp(w)}<br />

Z2 = {w ∈ W | lp(wr) = lp(w) + lp(r) = lp(sw); lp(swr) = lp(w)}.<br />

Then, each Zi is p-completely regular due to the following descriptions.<br />

and<br />

Z1 = Wr,1,lp(r) ∩ W1,s,lp(s) ∩ Wr,s,lp(r)<br />

Z2 = Wr,1,0 ∩ W1,s,0 ∩ Wr,s,lp(r).<br />

54


Now, in general, let u = xrx −1 and let v = ysy −1 , then we know that W (r, s)<br />

is p-completely regular by above and<br />

yW (r, s)x −1 = {ywx −1 ∈ W | wrw −1 = s}<br />

= {w ′ ∈ W | y −1 w ′ xrx −1 w ′−1 y = s}<br />

= {w ′ ∈ W | w ′ uw ′−1 = v} = W (u, v)<br />

is p-completely regular since p-completely regular sets are closed under left and<br />

right multiplication by elements of W by Proposition 5.5.1, proving 1.<br />

Now, according to equation (2.2.3), we know that the W -action on T × {±1}<br />

is given by w(t, ɛ) = (wtw −1 , τw,tɛ) where τw,t ∈ {±1} is given by τw,t = 1 if<br />

lp(wt) > lp(w) and τw,t = −1 if lp(wt) < lp(w). Then, let α = (u, ɛ) and β = (v, ɛ ′ )<br />

then<br />

W (α, β) = W (u, v) ∩ {w ∈ W | τw,uɛ = ɛ ′ }.<br />

Now, W (u, v) is p-completely regular by 1. If ɛ = ɛ ′ then τw,u must be 1 and<br />

so {w ∈ W | τw,u = 1} = {w ∈ W | lp(wu) > lp(w)} is p-completely regular by<br />

Example 3 in 5.6. Likewise if ɛ �= ɛ ′ then τw,u must be -1 and so {w ∈ W | τw,u =<br />

−2} = {w ∈ W | lp(wu) < lp(w)} is p-completely regular by the same example.<br />

Thus, {w ∈ W | τw,uɛ = ɛ ′ } is p-completely regular so that W (α, β) also is as<br />

required.<br />

Finally, to demonstrate 3, we note that the set described is empty unless<br />

|Γ| = |∆|. Let {α1, ..., αn} = Γ and {β1, ..., βn} = ∆. Then we have<br />

YΓ,∆ := {w ∈ W | w(Γ) = ∆} = ∪σ∈Sn ∩ n i=1 W (αi, βσ(i)),<br />

where Sn denotes the symmetric group. Then, YΓ,∆ is p-completely regular by 2.<br />

55


Similarly, we have 4 following from 1 in the exact same way replacing W (αi, βσ(i))<br />

with W (si, rσ(i)). Finally, according to section 2.3, we know that w((WJ ∩ T ) ×<br />

{1}) = (WK ∩ T ) × {1} if and only if w(J × {1}) = K × {1}, so 5 follows from 3<br />

directly using Γ = J × {1} and ∆ = K × {1} since both sets are finite.<br />

Example 5.7.2. The centralizer of a reflection, t, is CW (t) = {w ∈ W | w −1 tw =<br />

t}. Due to the previous Proposition, this set is p-completely regular. So the cen-<br />

tralizer of any finitely generated reflection subgroup is also p-completely regular.<br />

5.8 Regularity in Direct Products<br />

We now turn to looking at direct products of Coxeter systems. We have the<br />

following result describing the p-completely regular subsets of a Coxeter system<br />

in terms of the p-completely regular subsets of its irreducible components.<br />

Lemma 5.8.1. Suppose that (W, S) is the direct product of Coxeter systems<br />

(Wi, Si) for i = 1, ..., n. We identify W = W1 × · · · × Wn and S = ˙∪Si. Then<br />

Bp(W ) is equal to the Boolean subalgebra Bp(W1) ∗ · · · ∗ Bp(Wn) of P(W ) gen-<br />

erated by all subsets of W of the form A1 × · · · × An with Ai ∈ Bp(Wi).<br />

Proof. This is clearly true for n = 1. Now, suppose it is true for n = k, we will<br />

show that it is true for n = k + 1. We consider W ′ = W1 × · · · × Wk and Wk+1<br />

as subgroups of W . For any x, y ∈ W , we write x = x ′ xk+1 and y = y ′ yk+1 where<br />

56


x ′ , y ′ ∈ W ′ and xk+1, yk+1 ∈ Wk+1. Let m ∈ M. Then<br />

W p x,y,m = {w ∈ W | lp(ywx) = lp(y) + lp(w) + lp(x) − 2m}<br />

= {w ′ wk+1 ∈ W ′ × Wk+1 |<br />

lp(y ′ w ′ x ′ yk+1wk+1xk+1) =<br />

lp(y ′ ) + lp(yk+1) + lp(w ′ ) + lp(wk+1) + lp(x ′ ) + lp(xk+1) − 2m}<br />

= �<br />

W<br />

k1,k2∈M<br />

k1+k2=m<br />

′p<br />

x ′ ,y ′ p<br />

× (Wk+1)<br />

,k1 xk+1,yk+1,k2<br />

(5.8.1)<br />

so that Bp(W ) ⊆ Bp(W ′ ) ∗ Bp(Wk+1). Also, equation (5.8.1) implies that<br />

W p<br />

x ′ ,y ′ ′p<br />

= W ,k1 x ′ ,y ′ ,k1 × Wk+1 and W p<br />

xk+1,yk+1,k2 = W ′ × (Wk+1) p<br />

. Therefore,<br />

xk+1,yk+1,k2<br />

we have<br />

W ′p<br />

x ′ ,y ′ p<br />

p<br />

× (Wk+1) = W ,k1 xk+1,yk+1,k2 x ′ ,y ′ p<br />

∩ W ,k1 xk+1,yk+1,k2<br />

so that Bp(W ′ ) ∗ Bp(Wk+1) ⊆ Bp(W ). However, by induction, we know that<br />

Bp(W ′ ) = Bp(W1) ∗ · · · ∗ Bp(Wk) and the result follows.<br />

5.9 Product Regularity<br />

Complete regularity of a subset in a product W = W1 × · · · × Wn implies<br />

regularity of many subsets of S ∗ and so regularity of various subsets of S ∗ 1 ×· · ·×S ∗ n.<br />

Then we regard each S ∗ i as a submonoid of S ∗ and each Wi as a subgroup of W ;<br />

then we can define natural homomorphisms νi : S ∗ i → Wi. We say that a subset of<br />

S ∗ 1 × · · · × S ∗ n is product regular if it is in the Boolean algebra of subsets generated<br />

by the products A1 × · · · × An with each Ai regular in S ∗ i .<br />

Lemma 5.9.1. For any p-completely regular subset A of W1 × · · · × Wn, and any<br />

57


n-tuple (k1, ..., kn) ∈ M n , the subset<br />

{(x1, ..., xn) ∈ (S1) ∗ k1 × · · · × (Sn) ∗ kn | (ν1(x1), ..., νn(xn)) ∈ A}<br />

is product regular in (S1) ∗ × · · · × (Sn) ∗ .<br />

Proof. For any subset A of W , we define<br />

A ′ := {x1, ..., xn) ∈ (S1) ∗ k1 × · · · × (Sn) ∗ kn | (ν1(x1), ..., νn(xn)) ∈ A}.<br />

If A is a product then A = A1 × · · · × An, then we have<br />

A ′ = � (S1) ∗ k1<br />

∩ ν−1 1 (A1) � × · · · × � (Sn) ∗ kn ∩ ν−1 n (An) � .<br />

In this, if each Ai is p-completely regular (or just p-strongly regular), then each<br />

(Si) ∗ ki<br />

∩ ν−1<br />

i (Ai) is regular in (Si) ∗ by Lemma 5.3.3 and thus A ′ is product regular.<br />

Now if A = ∪mBm, where the union is finite, then A ′ = ∪mB ′ m. Now, this implies<br />

that A ′ is product regular for any p-completely regular set since any such set A is<br />

a finite union of products of p-completely regular sets.<br />

5.10 Multivariate Product Regularity<br />

For any Coxeter system (W, S), define a subset of W n to be p-completely<br />

regular if it is completely regular as a subset of the n-fold product Coxeter system<br />

(W, S) n . We have the following result.<br />

Lemma 5.10.1. Let n, m be natural numbers, m ∈ M, f1, ..., fm ∈ {1, ..., n},<br />

ɛ1, ..., ɛm ∈ {±1}, x0, x1, ..., xm ∈ W and A, B, C be finite subsets of W satisfying<br />

B, C ⊆ A. Then let L be the subset of W n consisting of n-tuples satisfying the<br />

58


following conditions:<br />

(i) If<br />

v := xm(wfm) ɛm xm−1(wfm−1) ɛm−1 xm−2 · · · x1(wf1) ɛ1 x0<br />

and k := � m<br />

i=0 lp(xi) + � m<br />

i=1 lp(wfi ), then lp(v) = k − 2m.<br />

(ii) N(v) ∩ A = B and N(v −1 ) ∩ A = C.<br />

Then L is a p-completely regular subset of W n .<br />

Proof. We will prove this by induction on m. To make things more tractable, we<br />

denote L by<br />

L = L(n, m, m, (fj), (ɛj), (xj), A, B, C)<br />

to show that L depends on its parameter set.<br />

First, suppose that m is fixed. If<br />

is p-completely regular, then<br />

L(n, m, m, (fj), (ɛj), (xj), ∅, ∅, ∅)<br />

L(n, m, m, (fj), (ɛj), (xj), A, B, C)<br />

is also p-completely regular using Lemma 4.6.1 (similar to the proof of Proposition<br />

5.5.1 part (b)).<br />

Now, suppose that m = 1. The set<br />

L(1, 1, m, (1), (1), (x0, x1), ∅, ∅, ∅) = W p x0,x1,m<br />

and thus is p-completely regular by definition. Also, as in the proof of Proposition<br />

59


5.5.1 part (c), we have<br />

L(1, 1, m, (1), (−1), (x0, x1), ∅, ∅, ∅) = L(1, 1, m, (1), (1), (x −1<br />

1 , x −1<br />

0 ), ∅, ∅, ∅).<br />

It follows that the sets<br />

for ɛ ∈ {±1} are p-completely regular.<br />

Let n ∈ N. Then any set<br />

with k ∈ {1, ..., n} is of the form<br />

L(1, 1, m, (1), (ɛ), (x0, x1), ∅, ∅, ∅)<br />

L ′ = L(n, 1, m, (k), (ɛ), (x0, x1), ∅, ∅, ∅)<br />

L ′ = W k−1 × L(1, 1, m, (1), (ɛ), (x0, x1), ∅, ∅, ∅) × W n−k<br />

and thus is p-completely regular. Therefore, we have shown that any set<br />

is p-completely regular.<br />

L(n, 1, m, (k), (ɛ), (x0, x1), A, B, C)<br />

At this point, we have proved that the set L is p-completely regular if m = 1,<br />

and if m = 0 it is clear that L is p-completely regular. So, we now move to the<br />

induction step. We will show that for fixed m, p-complete regularity of all sets of<br />

the form<br />

L(n, m, m, (fj), (ɛj), (xj), A, B, C)<br />

60


implies p-complete regularity of any sets of the form<br />

L ′′ := L(n, m + 1, m, (fj), (ɛj), (xj), ∅, ∅, ∅).<br />

Now, suppose that w = (w1, ..., wn) ∈ W n is an element of L ′′ above. Then we<br />

define the following two words:<br />

and<br />

vm := xm(wfm) ɛm xm−1(wfm−1) ɛm−1 xm−2 · · · x1(wf1) ɛ1 x0<br />

vm+1 := xm+1(wfm+1) ɛm+1 xm(wfm) ɛm xm−1(wfm−1) ɛm−1 xm−2 · · · x1(wf1) ɛ1 x0<br />

and two p-lengths km := � m<br />

i=0 lp(xi) + � m<br />

i=1 lp(wfi ) and km+1 := km + lp(xm+1) +<br />

lp(wfm+1). Let v := xm+1(wfm+1) ɛm+1 so that vm+1 = vvm, and let k := lp(xm+1) +<br />

lp(wfm+1) so that km+1 = km + k. Using these equalities, we have<br />

lp(vm+1) = lp(v) + lp(vm) − 2|N(v −1 ) ∩ N(vm)|<br />

= km+1 − k − km + lp(v) + lp(vm) − 2|N(v −1 ) ∩ N(vm)|<br />

�<br />

k − lp(v)<br />

= km+1 − 2<br />

+<br />

2<br />

km − lp(vm)<br />

+ |N(v<br />

2<br />

−1 �<br />

) ∩ N(vm)| .<br />

By definition, lp(v) = lp(xm+1) + lp(wfm+1) − 2|N(x −1<br />

m+1) ∩ N((wfm+1) ɛm+1 )| ≤ k,<br />

and similarly we have lp(vm) ≤ km. We know that w ∈ L ′′ so we must have<br />

lp(vm+1) = km+1 − 2m.<br />

According to Corollary 3.10.1 part (4) we know that lp(vm+1) < km+1 − 2m if<br />

lp(v) < k − 2m, lp(vm) < km − 2m or |Nm(v −1 ) ∩ Nm(vm)| > m. Otherwise, we<br />

61


have<br />

�<br />

k − lp(v)<br />

lp(vm+1) = km+1 − 2<br />

+<br />

2<br />

km − lp(vm)<br />

+ |Nm(v<br />

2<br />

−1 �<br />

) ∩ Nm(vm)| .<br />

(5.10.1)<br />

Suppose lp(v) = k − 2m ′ and lp(vm) = km − 2m ′′ . Then the equation (5.10.1) is<br />

lp(vm+1) = km+1 − 2 � m ′ + m ′′ + |Nm(v −1 ) ∩ Nm(vm)| � .<br />

Since lp(vm+1) = km+1 − 2m, then we have the following:<br />

where<br />

L ′′ =<br />

�<br />

m ′ ,m ′′ ,B,C<br />

B,C⊂T≤m<br />

m ′ +m ′′ +|B∩C|=m<br />

Lm ′ ,m ′′ ,B,C<br />

Lm ′ ,m ′′ ,B,C = L(n, m, m ′ , (fj) m j=1, (ɛj) m j=1, (xj) m j=0, T≤m, B, ∅)<br />

∩ L(n, 1, m ′′ , (fm+1), (ɛm+1), (1, xm+1), T≤m, ∅, C)<br />

both of which are p-completely regular by induction. Thus L ′′ is a finite union of<br />

p-completely regular sets and so L ′′ is p-completely regular as needed.<br />

62


CHAPTER 6<br />

<strong>REGULARITY</strong> <strong>IN</strong> T <strong>AND</strong> MIXED PRODUCTS<br />

The p-complete regularity in W as we have described has been useful for show-<br />

ing that many natural subsets of W are regular, and therefore have rational<br />

Poincaré series. However, there are many natural subsets of W , in particular<br />

subsets of T , that can be shown to not be p-complete regular in W (or even µ-<br />

regular in W ). Nonetheless, we would like to be able to show that certain subsets<br />

of T are regular in some sense. We will use the fact that every reflection has a<br />

palindromic reduced expression to introduce a notion of p-complete regularity in<br />

T . We can then extend our notion of regularity to regularity in products of T<br />

or even mixed products of W and T . We include some examples of p-completely<br />

regular subsets of T , and the regularity implies that these subsets have rational<br />

(multivariate) Poincaré series.<br />

6.1 Regularity of sets of Reflections<br />

Let X ⊂ S ∗ be the set of all non-empty words in S ∗ . We see that X is<br />

clearly a regular subset of S ∗ being the complement of the set consisting of the<br />

empty word. We say that a subset of X is regular if it is regular as a subset of<br />

S ∗ . Define η : X → T by η(rn+1rn · · · r1) = r1 · · · rnrn+1rn · · · r1. Let ℓ ′ p(x) =<br />

2ℓp(x) − lp(rn+1) = 2ℓp(x) − p(η(x)). Let � ′ L<br />

lexicographic order on S ∗ to X.<br />

63<br />

be the restriction of some reverse


Proposition 6.1.1. 1. For any m ∈ M let Xm := {x ∈ X| lp(η(x)) = ℓ ′ p(x)−<br />

2m}. Then Xm is a regular subset of X.<br />

2. For t ∈ T , define ˙xt as the minimal (under the order � ′ L ) element of the<br />

non-empty finite set η −1 (t) ∩ X0. Then ˙<br />

T = { ˙xt| t ∈ T } is a regular subset<br />

of X.<br />

Proof. We prove 1 first. Let m ∈ M, and let x ∈ Xm. Then, we can write<br />

x = rx ′ with r ∈ S and x ′ ∈ S ∗ k<br />

for some k ∈ M. By definition, we know that<br />

lp(ν(x ′ )) = ℓp(x ′ ) − 2k. Since x ∈ Xm, and substituting the previous formula we<br />

get<br />

lp(ν(x ′ ) −1 rν(x ′ )) = lp(η(x)) = ℓ ′ p(x) − 2m = 2ℓp(x ′ ) + lp(r) − 2m<br />

= 2(2lp(ν(x ′ )) + 2k) + lp(r) − 2m<br />

= 2lp(ν(x ′ )) + lp(r) − 2(m − 2k)<br />

. (6.1.1)<br />

Equation (6.1.1) implies that m − 2k ≥ 0 so 2k ≤ m; furthermore, (6.1.1) also<br />

implies that ν(x ′ ) ∈ Ar,m−2k where<br />

Ar,m−2k := L(1, 2, m − 2k, (1, 1), (1, −1), (e, r, e), ∅, ∅, ∅)<br />

where e ∈ W is the identity, and L is as in 5.10. According to Lemma 5.10.1<br />

Ar,m−2k is p-completely regular. Therefore, we have x ′ ∈ S ∗ k ∩ ν−1 (Ar,m−2k) which<br />

is regular by Lemma 5.5.1. So, for any x ∈ Xm, there is an r ∈ S and k ∈ M<br />

with 2k ≤ m such that x ∈ r(S ∗ k ∩ ν−1 (Ar,m−2k)). Thus<br />

Xm ⊆ �<br />

�<br />

�<br />

r<br />

r∈S<br />

0≤2k≤m<br />

64<br />

S ∗ k ∩ ν −1 (Ar,m−2k)<br />


Now, suppose that x ∈ r(S ∗ k ∩ ν−1 (Ar,m−2k)) for some r ∈ S and k ∈ M with<br />

2k ≤ m. We have from above that<br />

Ar,m−2k = {w ∈ W | lp(w −1 rw) = 2lp(w) + lp(r) − 2(m − 2k)}.<br />

Thus x = rx ′ with x ′ ∈ ν −1 (Ar,m−2k) and lp(ν(x ′ )) = 2ℓp(x ′ ) − 2k. This gives that<br />

lp(η(x)) = lp(ν(x ′ ) −1 rν(x ′ )) = 2lp(ν(x ′ )) + lp(r) − 2(m − 2k)<br />

so that x ∈ Xm. Thus we have shown<br />

Xm = �<br />

�<br />

�<br />

r<br />

r∈S<br />

= 2(ℓp(x ′ ) − 2k) + lp(r) − 2(m − 2k)<br />

= 2ℓp(x ′ ) + lp(r) − 2m = 2ℓ ′ p(x) − 2m,<br />

0≤2k≤m<br />

S ∗ k ∩ ν −1 (Ar,m−2k)<br />

and the right hand side is a finite union of regular sets as already shown hence<br />

Xm is regular as well.<br />

Part (2) follows from and unpublished result due to Bob Howlett, which is the<br />

analog of 5.2.1 for T instead of W . The result is mentioned in [18].<br />

6.2 Types of Regularity in T<br />

Fix any map τ : T → X, denoted t → xt for t ∈ T , with the following<br />

properties:<br />

(i) For all t ∈ T , τ(t) ∈ X0.<br />

(ii) η ◦ τ = IdT .<br />

(iii) ˙<br />

T := τ(T ) is a regular subset of X.<br />

65<br />

�<br />

,


We know that maps, τ, satisfying (i)-(iii) above exist due to the previous propo-<br />

sition. We say a subset A of T is τ-regular if τ(A) is regular in X.<br />

Definition 6.2.1. A subset A of T is regular (respectively p-strongly regular) in<br />

T if η −1 (A) ∩ Xm is a regular subset of X for m = 0 (respectively all m ∈ M).<br />

Lemma 6.2.2. (a) Any p-strongly regular subset of T is regular.<br />

(b) If A is a regular subset of T then A is τ-regular for any map satisfying<br />

(i)-(iii) above.<br />

Proof. Part (a) is a restatement of the definition since 0 ∈ M. For part (b), we<br />

have that for regular subset A ⊆ T and τ satisfying (i)-(iii) above, we get that<br />

τ(A) = (η −1 (A) ∩ X0) ∩ τ(T ) which is an intersection of regular sets and thus<br />

regular.<br />

6.3 p-Complete Regularity in T<br />

We now want to describe the analog of Bp(W ) when dealing with regular<br />

subsets of T . To do this, we first introduce a few subsets of P(T ).<br />

Definition 6.3.1. 1. For x ∈ W and m ∈ M, we let<br />

R p x,m := {t ∈ T | lp(xtx −1 ) = 2lp(x) + lp(t) − 2m}.<br />

Then, let B ′′<br />

p(T ) be the Boolean subalgebra of P(T ) generated by all sets<br />

R p x,m for x ∈ W and m ∈ M.<br />

2. For t ∈ T and A ⊆ T , we define At := {w ∈ W | w −1 tw ∈ A}. Then,<br />

let B ′ p(T ) be the family of all subsets A of T such that for all r ∈ S,<br />

Ar ∈ Bp(W ).<br />

66


3. For any t ∈ T and A ⊂ T and m ∈ M, we define<br />

A p<br />

t,m := {w ∈ W | w −1 tw ∈ A; lp(w −1 tw) = 2lp(w) + lp(t) − 2m}.<br />

We let Bp(T ) be the family of all subsets A of T such that for all r ∈ S<br />

and m ∈ M we have Ar,m ∈ Bp(W ). We call the elements of Bp(T ) the<br />

p-completely regular subsets of T .<br />

Properties of these families of subsets are described in the following proposi-<br />

tion. Again, for ease of notation, we omit the superscript p since p is understood<br />

to be fixed.<br />

Proposition 6.3.2. (a) If A ∈ B ′ p(T ) then At ∈ Bp(W ) for all t ∈ T . Like-<br />

wise, if A ∈ Bp(T ) then At,m ∈ Bp(W ) for all t ∈ T and m ∈ M.<br />

(b) B ′′<br />

p(T ), B ′ p(T ) ⊂ Bp(T ) are Boolean subalgebras of P(T ) closed under con-<br />

jugation by elements of W and closed under permutations of T induced by<br />

automorphisms, θ, of (W, S) that commute with p.<br />

(c) B ′ p(T ) contains all finite subsets of T and all conjugacy classes of simple<br />

reflections.<br />

(d) Any set A which is p-completely regular in T is p-strongly regular in T .<br />

Proof. Suppose A ∈ B ′ p(T ). Then Ar ∈ Bp(W ) for all r ∈ S. Now, let t ∈ T be<br />

given. Then, t = u −1 su for some u ∈ W and s ∈ S. Now we have<br />

At = {w ∈ W | w −1 tw ∈ A} = {w ∈ W | w −1 u −1 suw ∈ A}<br />

= {u −1 v ∈ W | v −1 sv ∈ A} = u −1 · {v ∈ W | v −1 sv ∈ A} = u −1 As,<br />

67


and u −1 As ∈ Bp(W ) since As ∈ Bp(W ) and by Proposition 5.5.1 Bp(W ) is<br />

closed under left multiplication by elements of W , thus the first part of (a) is true.<br />

For the second part of (a), we fix u ∈ W and s ∈ S such that t = u −1 su and<br />

lp(t) = 2lp(u) + lp(s). Now, we consider the sets<br />

Bk := u −1 As,m−2k ∩ W1,u,k ∈ Bp(W ).<br />

Notice, if w ∈ Bk then w = u −1 v with v ∈ As,m−2k, and so we have w −1 tw =<br />

v −1 uu −1 suu −1 v = v −1 sv ∈ A. Also, we have<br />

lp(w −1 tw) = lp(v −1 sv) = 2lp(v) + lp(s) − 2(m − 2k) = 2lp(uw) + lp(s) − 2(m − 2k),<br />

but w ∈ W1,u,k so lp(uw) = lp(u) + lp(w) − 2k so that we get<br />

lp(w −1 tw) = 2(lp(u) + lp(w) − 2k) + lp(s) − 2(m − 2k)<br />

= 2lp(w) + 2lp(u) + lp(s) − 2m<br />

= 2lp(w) + lp(t) − 2m.<br />

Therefore Bk ⊆ At,m. Similarly, At,m ∩ W1,u,k ⊆ Bk. We note that Bk = ∅ unless<br />

2k ≤ m, and thus<br />

At,m =<br />

�<br />

k:0≤2k≤m<br />

So At,m ∈ Bp(W ) since it is a finite union of sets in Bp(W ), finishing the proof<br />

of (a).<br />

For the first part of (b), we notice that for A, B ∈ B ′ p(T ) or A, B ∈ Bp(T ) and<br />

r ∈ S, it is clear that (A ∪ B)r = Ar ∪ Br, (A ∩ B)r = Ar ∩ Br and (A c )r = (Ar) c<br />

so that B ′ p(T ) and Bp(T ) are clearly Boolean subalgebras of P(T ) and B ′′<br />

p(T )<br />

68<br />

Bk.


is a Boolean subalgebra by definition. Now, suppose that A ∈ B ′ p(T ). Then<br />

Ar ∈ Bp(W ) by definition. Also, for any r ∈ S and m ∈ M, we have the set<br />

Br,m = {w ∈ W | lp(w −1 rw) = 2lp(w) + lp(r) − 2m}<br />

= L(1, 2, m, (1, 1), (1, −1), (e, r, e), ∅, ∅, ∅)<br />

following notation from section 5.10. Thus, by Lemma 5.10.1 we have Br,m ∈<br />

Bp(W ) for all r ∈ S and m ∈ M. Finally, we see that Ar,m = Ar ∩ Br,m, and<br />

thus Ar,m ∈ Bp(W ) for all r ∈ S and m ∈ M so that A ∈ Bp(T ).<br />

Now, we show that R p x,m ∈ Bp(T ) for all x ∈ W and m ∈ M. For any r ∈ S<br />

and n ∈ M we have<br />

(R p x,m)r,n = {w ∈ W | w −1 rw ∈ Tx,m; lp(w −1 rw) = 2lp(w) + lp(r) − 2n}<br />

= {w ∈ W | lp(xw −1 rwx −1 ) = 2lp(x) + lp(w −1 rw) − 2m;<br />

lp(w −1 rw) = 2lp(w) + lp(r) − 2n}<br />

= {w ∈ W | lp(xw −1 rwx −1 ) = 2lp(x) + 2lp(w) + lp(r) − 2(m + n);<br />

lp(w −1 rw) = 2lp(w) + lp(r) − 2n}<br />

= L(1, 2, m + n, (1, 1), (−1, 1), (x, r, x −1 ), ∅, ∅, ∅) ∩ Br,n<br />

using notation from the proof of Lemma 5.10.1, and by the result of that lemma,<br />

we get that (R p x,m)r,n ∈ Bp(W ). Thus, R p x,m ∈ Bp(T ) so that B ′′<br />

p(T ) ⊂ Bp(T ) as<br />

required.<br />

Next, suppose that A ∈ B ′ p(T ). Then Ar ∈ Bp(W ) for all r ∈ S. Let w ∈ W<br />

69


e fixed. Then wAw −1 ∈ B ′ p(T ) since for r ∈ S we have<br />

(wAw −1 )r = {u ∈ W | u −1 ru ∈ wAw −1 } = {u ∈ W | w −1 u −1 ruw ∈ A}<br />

= {vw −1 ∈ W | v −1 rv ∈ A} = {v ∈ W | v −1 rv ∈ A}w −1 = Arw −1<br />

is in Bp(W ) because Bp(W ) is closed under multiplication by elements of W .<br />

For A ∈ Bp(T ) and w ∈ W , we see that for r ∈ S and m ∈ M we have<br />

(wAw −1 )r,m = (wAw −1 )r ∩ �<br />

0≤k≤lp(w)<br />

(B(w, r, 2k) ∩ W w −1 ,1,k)w −1<br />

(6.3.1)<br />

where B(w, r, 2k) = {v ∈ W | lp(wv −1 rvw) = 2lp(w) + 2lp(v) + lp(r) − 2(m + 2k)}.<br />

Indeed, if u ∈ (wAw −1 )r,m, then u ∈ (wAw −1 )r and lp(u −1 ru) = 2lp(u) + lp(r) −<br />

2m. By the previous argument, u = vw −1 , and then we get lp(wv −1 rvw −1 ) =<br />

2lp(vw −1 ) + lp(r) − 2m = 2(lp(v) + lp(w) − 2k) + lp(r) − 2m where k ≤ lp(w) (and<br />

we note that this means v ∈ W w −1 ,1,k). Hence, we have inclusion of the left hand<br />

side in the right hand side of (6.3.1). However, working the equation backwards,<br />

since k ≤ lp(w) we get inclusion the other way proving that Bp(T ) is closed under<br />

conjugation.<br />

Suppose that θ is an automorphism of (W, S). Then we see if A ∈ B ′ p(T ) we<br />

have for any r ∈ S<br />

θ(A)r = {u ∈ W | u −1 ru ∈ θ(A)} = {u ∈ W | θ −1 (u −1 ru) ∈ A}<br />

= {u ∈ W | θ −1 (u −1 )θ −1 (r)θ −1 (u) ∈ A} = {θ(v) ∈ W | v −1 θ −1 (r)v ∈ A}<br />

= θ(A θ −1 (r))<br />

and so θ(A)r ∈ Bp(W ) since Bp(W ) is closed under automorphisms. Notice<br />

this holds for any automorphism. For A ∈ Bp(T ), and θ an automorphism of<br />

70


(W, S) commuting with p, recall that from the proof of Proposition 5.5.1 that<br />

θ(lp(w)) = lp(w). Then for any r ∈ S and m ∈ M we have<br />

θ(A)r,m = {u ∈ W | u −1 ru ∈ θ(A); lp(u −1 ru) = 2lp(u) + lp(r) − 2m<br />

= θ(A)r ∩ {u ∈ W | lp(u −1 ru) = 2lp(u) + lp(r) − 2m}.<br />

We know that θ(A)r ∈ Bp(W ) from above, and {u ∈ W | lp(u −1 ru) = 2lp(u) +<br />

lp(r) − 2m} ∈ Bp(W ) by Lemma 5.10.1. Thus, we have proved (b).<br />

To show (c), we first note that {w ∈ W | w −1 rw = t} ∈ Bp(W ) for all r ∈ S by<br />

Proposition 5.7.1. Therefore {t} ∈ B ′ p(T ) for all t ∈ T , and hence any finite subset<br />

of T is p-completely regular as well. Also, if A = {wrw −1 | w ∈ W } is the conjugacy<br />

class of r ∈ S, then for any s ∈ A we have As = {w ∈ W | w −1 sw ∈ A} = W if<br />

s is conjugate to r and As = ∅ if s is not conjugate to r. Since W ∈ Bp(W ) and<br />

∅ ∈ Bp(W ), then As ∈ Bp(W ) for all s ∈ S and so A ∈ B ′ p(T ).<br />

Finally, we show that (d) is true. We have to show that Bm := η −1 (A) ∩ Xm<br />

is regular for all m ∈ M if A is p-completely regular. If x ∈ η −1 (A) then x = rx ′<br />

for some r ∈ S and η(x) = ν(x ′ ) −1 rν(x ′ ). Thus, η −1 (A) = ∪r∈S{r}B ′ r where<br />

B ′ r = {x ∈ S ∗ | ν(x) −1 rν(x) ∈ A}. Since x ∈ Xm, we also know that lp(η(x)) =<br />

lp(ν(x ′ ) −1 rν(x ′ )) = 2ℓp(x ′ ) + lp(r) − 2m. Therefore, we see that Bm = ∪r∈SrBr,m<br />

where<br />

Br,m := {x ′ ∈ S ∗ | ν(x ′ ) −1 rν(x ′ ) ∈ A; lp(ν(x ′ ) −1 rν(x ′ )) = 2ℓp(x ′ ) + lp(r) − 2m}.<br />

Hence, to show that Bm is regular, we need to show that each Br,m is regular (as<br />

Bm is then a finite union of regular sets).<br />

Now, if x ′ ∈ S ∗ k we know that ℓp(x ′ ) = lp(ν(x ′ )) + 2k. Therefore, we know that<br />

71


Br,m ∩ S∗ k is given by the set<br />

{x ′ ∈ S ∗ |ν(x ′ ) −1 rν(x ′ ) ∈ A; lp(ν(x ′ ) −1 rν(x ′ )) = 2lp(x ′ ) + lp(r) − 2(m − 2k)},<br />

which is equal to ν−1 (Ar,m−2k) ∩ S∗ k . Now, we know that since A is p-completely<br />

regular in T then Ar,n ∈ Bp(W ) for all n ∈ M so that Ar,m−2k is p-completely<br />

regular in W for all k. Thus, by Proposition 5.5.1 we know that ν −1 (Am−2k) ∩ S ∗ k<br />

is regular. Also, we see that Ar,m−2k = ∅ unless 0 ≤ 2k ≤ m. Therefore we have<br />

Br,m = ∪k∈M(Br,m ∩ S ∗ k) = ∪0≤k≤2m(Br,m ∩ S ∗ k)<br />

which is a finite union and thus regular, completing the proof of (d).<br />

6.4 Maps of Boolean Algebras<br />

In this section, we prove some technical facts about Boolean subalgebras that<br />

will be useful for describing regularity in mixed products of T and W .<br />

Suppose that X1, X2, ..., Xn are sets with each Xi �= ∅. Let X := X1×· · ·×Xn.<br />

For each i let<br />

be the natural projection and let<br />

πi : X → Xi<br />

π c i : X → X1 × · · · × ˆ Xi × · · · × Xn<br />

be the natural projection where ˆ Xi means we omit Xi. For each i let Bi be a<br />

Boolean subalgebra of P(Xi). Then let B := B1 ∗ · · · ∗ Bn represent the Boolean<br />

subalgebra of P(X) generated by all A1 × · · · × An with Ai ∈ Bi.<br />

72


For any Y ∈ P(X1) ∗ · · · ∗ P(Xn) = P(X), we define an equivalence relation<br />

on Xi, ≡Y,i, given by a ≡Y,i b if π c i<br />

� π −1<br />

i (a) ∩ Y � = π c i<br />

� π −1<br />

i (b) ∩ Y � .<br />

Lemma 6.4.1. Suppose X = X1 × · · · × Xn, B1 ∗ · · · ∗ Bn are as above, and<br />

Y ∈ P(X). Then the following hold.<br />

(a) For each i, there exist finitely many equivalence classes of ≡Y,i, denoted<br />

Y1,i, Y2,i, ..., Yni,i.<br />

(b) Define a rectangle to be a set of the form Yj1,1 ×Yj2,2 ×· · ·×Yjn,n (where each<br />

set in the product is one of the finitely many equivalence classes in (a)). If<br />

a point of Y is in a rectangle, then that rectangle is a subset of Y , and thus,<br />

Y is a union of (finitely many) rectangles.<br />

(c) Y ∈ B1 ∗ · · · ∗ Bn if and only if Yj,i ∈ Bi for all i ∈ {1, .., n} and all<br />

j ∈ {1, ..., ni}.<br />

(d) Let Fi be a subset of Xi for each i such that Fi∩Yj,i �= ∅ for all j ∈ {1, ..., ni}.<br />

Consider F ′ := Y ∩F of F := F1×· · ·×Fn, and let ≡F ′ ,i be the corresponding<br />

equivalence relation on each Fi (in the same manner as ≡Y,i on Xi). Then<br />

the equivalence classes for ≡F ′ ,i are precisely the intersections Yj,i ∩ Fi. In<br />

particular, the rectangles for Y ∩ F that are contained in Y ∩ F are the<br />

intersections with F of the rectangles for Y which are contained in Y .<br />

Proof. Since Y ∈ P(X), we write<br />

Y = �<br />

Ak,1 × · · · × Ak,n<br />

k∈K<br />

where K is a finite index set. Then define Ci to be the Boolean subalgebra of<br />

P(Xi) generated by the finite set of elements {Ak,i| k ∈ K}. For each k ∈ K,<br />

73


let A r k,i := Ak,i and let A c k,i = Xi \ Ak,i (the complement). Then for a sequence<br />

ɛ := (ɛ1, ..., ɛ|K|) with each ɛi ∈ {r, c} we define Ai,ɛ := ∩k∈KA ɛk<br />

k,i . There are<br />

only finitely many such sequences, and so we call the finite set {Ai,ɛ| Ai,ɛ �= ∅}<br />

the set of atoms of Ci. We necessarily have Ai,ɛ ∩ Ai,ɛ ′ = ∅ for ɛ �= ɛ′ , and<br />

each set J ∈ Ci is a finite union of atoms. Suppose now that a, b ∈ Ai,ɛ for<br />

some ɛ. Let J ⊆ K such that ɛj = r for j ∈ J and ɛj = c for j �∈ J. Then<br />

{Aj,i| a ∈ Aj,i} = {Aj,i| j ∈ J} = {Aj,i| b ∈ Aj,i}. Therefore,<br />

π c i<br />

� π −1<br />

i (a) ∩ Y � = �<br />

j∈J<br />

Aj,1 × · · · × ˆ Aj,i × · · · × Aj,n = π c � −1<br />

i πi (b) ∩ Y � ;<br />

and thus a ≡Y,i b so that each atom Ai,ɛ is contained in an equivalence class of<br />

≡Y,i. Therefore, the atoms give a finer partition of Xi than ≡Y,i. Since there are<br />

only finitely many atoms, there are only finitely many equivalence classes of ≡Y,i<br />

proving (a).<br />

Now, suppose that Yj1,1 ×Yj2,2 ×· · ·×Yjn,n is a rectangle and that (a1, ..., an) ∈<br />

Y ∩ (Yj1,1 × Yj2,2 × · · · × Yjn,n). Let (b1, ..., bn) ∈ Yj1,1 × Yj2,2 × · · · × Yjn,n, as well.<br />

Then, since a1 ≡Y,1 b1, we know that (b1, a2, ..., an) ∈ Y . Working inductively,<br />

since ai ≡Y,i bi for all i, we have (b1, ..., bi, ai+1, ..., an) ∈ Y for all i. In particular,<br />

(b1, ..., bn) ∈ Y . Therefore, Yj1,1 × Yj2,2 × · · · × Yjn,n ⊆ Y , and so Y is a finite union<br />

of rectangles proving (b).<br />

Next, if Yji,i ∈ Bi for all i ∈ {1, ..., n} and ji ∈ {1, ..., ni} then Yj1,1 × Yj2,2 ×<br />

· · · × Yjn,n ∈ B1 ∗ · · · ∗ Bn for all rectangles and since Y is a finite union of<br />

rectangles by (b), we have Y ∈ B1 ∗ · · · ∗ Bn. Conversely, if Y ∈ B1 ∗ · · · ∗ Bn,<br />

then from the proof of (a), each Ak,i ∈ Bi, and so each Ai,ɛ ∈ Bi and so Ci ⊆ Bi.<br />

However, each Yji,i ∈ Ci ⊆ Bi as required for (c).<br />

Finally, let Fi, F , and F ′ be as in the statement of (d). Let a ≡Y,i b and let<br />

74


a, b ∈ Fi. Then we know that<br />

Since both a, b ∈ Fi, this implies that<br />

π c i<br />

π c i<br />

� −1<br />

πi (a) ∩ Y )� = π c � −1<br />

i πi (b) ∩ Y )� .<br />

� −1<br />

πi (a) ∩ (Y ∩ F ))� = π c � −1<br />

i πi (b) ∩ (Y ∩ F ))� ,<br />

and thus a ≡F ′ ,i b. Therefore Yj,i ∩ Fi ⊆ F ′ k,i for some k.<br />

Now, suppose that a ≡F ′ ,i b. Then a, b ∈ Fi and we have<br />

π c i<br />

� −1<br />

πi (a) ∩ (Y ∩ F ))� = π c � −1<br />

i πi (b) ∩ (Y ∩ F ))� .<br />

Let (x1, ..., a, ..., xn) ∈ π −1<br />

i (a)∩Y . Then, since Yj,m ∩Fm �= ∅, we have xm ≡Y,m x ′ m<br />

with x ′ m ∈ Fm for all m �= i. Therefore, we see that (x ′ 1, ..., ai, ..., x ′ n) ∈ π −1<br />

i (a) ∩<br />

(Y ∩ F ). Since a ≡F ′ ,i b, then (x ′ 1, ..., b, ..., x ′ n) ∈ π −1<br />

i (b) ∩ (Y ∩ F ), and using the<br />

equivalence xm ≡Y,m x ′ m in reverse, this implies that (x1, ..., b, ..., xn) ∈ π −1<br />

i (b)∩Y .<br />

The exact same argument starting with b instead of a shows<br />

π c i<br />

� π −1<br />

i (a) ∩ Y )� = π c i<br />

� −1<br />

πi (b) ∩ Y )�<br />

so that a ≡Y,i b, and thus Fk,i ⊆ Yj,i ∩ F for some j. Therefore, we have equality<br />

of equivalence classes. Equality of rectangles follows directly.<br />

Next, we assume we additionally have another family of sets X ′ 1, ..., X ′ n. Let<br />

X ′ := X ′ 1 × · · · × X ′ n, and suppose that we are given a subset Hi ⊆ Hom(Xi, X ′ i) ×<br />

Bi. For each i, we define a Boolean algebra B ′ i (which depends on Hi) of P(X ′ i)<br />

consisting of sets A ⊆ X ′ i such that for all (hi, Bi) ∈ Hi we have h −1<br />

i (A)∩Bi ∈ Bi.<br />

75


Then, let B ′ := B ′ 1 ∗ · · · ∗ B ′ n be the corresponding Boolean subalgebra of P(X ′ ).<br />

Define H := {((h1 × · · · × hn), (B1 × · · · × Bn))| (hi, Bi) ∈ Hi; ∀i}. Additionally,<br />

we set<br />

B ′′ := {Z ⊆ X ′ | h −1 (Z) ∩ B ∈ B, for all (h, B) ∈ H}.<br />

Lemma 6.4.2. Suppose that for all i ∈ {1, ..., n} there exists a finite subset H ′ i ⊂<br />

Hi such that ∪(hi,Bi)∈H ′ i hi(Bi) = Xi. Then B ′′ ⊆ P(X ′ 1) ∗ · · · ∗ P(X ′ n).<br />

Proof. Define H ′ := {((h1 × · · · × hn), (B1 × · · · × Bn))| (hi, Bi) ∈ H ′ i; ∀i} ⊆ H.<br />

Let Z ∈ B ′′ . Clearly we have<br />

�<br />

(h,B)∈H ′<br />

h � h −1 (Z) ∩ B � ⊆ Z, (6.4.1)<br />

and since each h −1 (Z) ∩ B ∈ P(X) and h : P(X) → P(X ′ ), we have that the<br />

left hand side of (6.4.1) is in P(X ′ ). Now, suppose that (z1, ..., zn) ∈ Z. By<br />

assumption, we can write zi = hi(bi) for some (hi, Bi) ∈ H ′ i with bi ∈ Bi. Then,<br />

let h := h1 × · · · × hn and B := B1 × · · · × Bn so that (h, B) ∈ H ′ by construction.<br />

Then, b := (b1, ..., bn) ∈ h −1 (Z) ∩ B and so (z1, ..., zn) ∈ h (h −1 (Z) ∩ B) for some<br />

(h, B) ∈ H ′ . Therefore, we have<br />

Z ⊆ �<br />

(h,B)∈H ′<br />

h � h −1 (Z) ∩ B � ,<br />

and thus Z ∈ P(X ′ ) since we have equality in (6.4.1).<br />

Lemma 6.4.3. With notation as above, the following hold.<br />

1. B ′ ⊆ B ′′ .<br />

2. Suppose that for all i ∈ {1, ..., n} and for all h ∈ Hi we have that h is<br />

76


surjective; furthermore, assume that for each hi, we have<br />

�<br />

Bi: (hi,Bi)∈Hi<br />

Then B ′′ ⊆ B ′ (and thus B ′ = B ′′ ).<br />

Bi = Xi. (6.4.2)<br />

Proof. We begin by showing 1. Let C ′ i ∈ B ′ i. Then if (hi, Bi) ∈ Hi we have<br />

h −1<br />

i (C′ i) ∩ Bi ∈ Bi. So we have that<br />

(h1 × · · · × hn) −1 (C ′ i × · · · × C ′ n) ∩ (B1 × · · · × Bn)<br />

= (h −1<br />

1 (C ′ 1) ∩ B1) × · · · × (h −1<br />

n (C ′ n) ∩ Bn) ∈ B1 ∗ · · · ∗ Bn = B.<br />

Therefore, C ′ 1 × · · · × C ′ n ∈ B ′′ . Since B ′′ is closed under finite unions and<br />

intersections, 1 follows.<br />

Now, suppose that for all i and for all h ∈ Hi, we have that h is surjective.<br />

Let Z ∈ B ′′ . According to Lemma 6.4.2, Z ∈ P(X ′ ), and thus, by Lemma 6.4.1,<br />

we have<br />

Z = �<br />

Zk,1 × · · · × Zk,n<br />

k∈K<br />

where K is a finite set, and each Zk,1 ×· · ·×Zk,n is a rectangle (in the terminology<br />

of 6.4.1). Now, let (h, B) ∈ H with h = (h1 × · · · × hn). We have,<br />

V := h −1 (Z) = �<br />

k∈K<br />

A quick calculation shows that if a ∈ Xi then<br />

π c i<br />

h −1<br />

1 (Zk,1) × · · · × h −1<br />

n (Zk,n).<br />

� −1<br />

πi (a) ∩ h−1 (Z) � = h −1 � π c � −1<br />

i πi (hi(a)) ∩ Z �� . (6.4.3)<br />

77


Using equation (6.4.3) and the assumption that all hi are surjective, then if a ≡V,i<br />

b, we get hi(a) ≡Z,i hi(b) and if a ≡Z,i b, we get a ′ ≡V,i b ′ for all a ′ ∈ h −1<br />

i (a) and<br />

b ′ ∈ h −1<br />

i (b). Therefore, each h−1 1 (Zk,1) × · · · × h−1 n (Zk,n) is a rectangle for h−1 (Z)<br />

(recalling the definition of a rectangle from 6.4.1).<br />

Now, fix h = (h1 × · · · × hn). Consider the sets Qi := {D (j)<br />

i | (hi, D (j)<br />

i ) ∈ Hi}.<br />

By assumption<br />

if D (ji)<br />

i<br />

h −1 (Z) ∩ (D (j1)<br />

1<br />

× · · · × D (jn)<br />

n ) ∈ B<br />

∈ Qi for all i. Then, we let ˜ Bi = ∪j∈JD (j)<br />

i<br />

Since these unions are finite, we have<br />

Z ′ := h −1 (Z) ∩ ( ˜ B1 × · · · × ˜ Bn) ∈ B.<br />

with J finite and all D(j) i ∈ Qi.<br />

Choose each ˜ Bi large enough so that ˜ Bi ∩ Vj,i �= ∅ for all j ∈ {1, ..., ni} (all<br />

equivalence classes of ≡V,i); we note that this is possible by our assumption (6.4.2).<br />

Then, we have<br />

Z ′ = �<br />

k∈K<br />

(h −1<br />

1 (Zk,1) ∩ ˜ B1) × · · · × (h −1<br />

n (Zk,n) × ˜ Bn) ∈ B.<br />

By Lemma 6.4.1 (d), we know that (h −1<br />

1 (Zk,1) ∩ ˜ B1) × · · · × (h −1<br />

n (Zk,n) × ˜ Bn) is<br />

a rectangle. Hence, by Lemma 6.4.1 (c), we have that each h −1<br />

i (Zk,i) ∩ ˜ Bi ∈ Bi.<br />

Finally, since Bi ∈ Bi for all i, we know that<br />

h −1<br />

i (Zk,i) ∩ Bi = h −1<br />

i (Zk,i) ∩ ˜ Bi ∩ Bi ∈ Bi.<br />

Therefore, for all i ∈ {1, ..., n} and k ∈ K, we have Zi,k ∈ B ′ i and thus Z ∈ B ′ as<br />

required.<br />

78


6.5 Regularity in Mixed Products<br />

We now extend our notions of p-complete regularity on T and W to the notion<br />

of a p-complete regularity of products W n × T m with n, m ∈ N.<br />

As done previously when dealing with Cartesian products, we say that a subset<br />

of a product B1 × · · · × Bn with each Bi equal to T or W is p-completely regular<br />

if it is in the Boolean subalgebra of P(B1 × · · · × Bn) generated by products<br />

A1 × · · · × An where each Ai is p-completely regular in W if Bi = W and each Aj<br />

is p-completely regular in T for Bj = T . To avoid trivial situations, if n = 0, we<br />

let B1 × · · · × Bn be the singleton set and declare all subsets to be p-completely<br />

regular. For ease of notation in the next following lemma, we assume that B =<br />

B1 × · · · × Bn = T k × W n−k with Bi = T for i ∈ {1, ..., k} and Bi = W otherwise.<br />

Lemma 6.5.1. Let B = B1 ×· · ·×Bn be as above, with each Bi = T for 1 ≤ i ≤ k<br />

and Bi = W for k < i ≤ n. For any family ¯t = (t1, ..., tk) in T k , we define<br />

f¯t : W n → B1 × · · · × Bn by f¯t(w1, ..., wn) = (u1, ..., un) where<br />

⎧<br />

⎪⎨ w<br />

ui :=<br />

⎪⎩<br />

−1<br />

i tiwi; if 1 ≤ i ≤ k<br />

wi; otherwise.<br />

Then a subset X ⊆ B1 × · · · × Bn is p-completely regular if and only if for all ¯s =<br />

(s1, ..., sk) ∈ Sk and (m1, ..., mk) ∈ Mk , f −1<br />

¯s (X)∩ � Ts1,m1 × · · · × Tsk,mk × W n−k�<br />

is p-completely regular in W n .<br />

Proof. This lemma follows from the lemmas in Section 6.4 because the map and<br />

sets satisfy the following. Using terminology from that section, we have H ′ i :=<br />

∪ ¯s∈S k(f¯s, T¯s,0), which is finite. Since each conjugacy class is p-completely regular,<br />

and T is a finite union of conjugacy classes, we reduce to the case where we replace<br />

79


Bi = T by a conjugacy class in T so that fsi is surjective. Also, for any ¯s ∈ Sk , we<br />

have ∪m∈MTsi,m equal to the conjugacy class of si, so we satisfy (6.4.2). Finally,<br />

by definition A ⊆ T is p-completely regular in T if and only if f −1<br />

s (A)∩Ts,m = As,m<br />

is p-completely regular in W for all s ∈ S and m ∈ M.<br />

Remark 6.5.2. We see that the above proof would work for mixed products B =<br />

B1 × · · · × Bn. We simply used the appropriate ordering for ease of notation in<br />

the statement.<br />

Remark 6.5.3. Let B ∼ = W n−k ×T k and B ′ ∼ n<br />

= W ′ −k ′<br />

×T k′ be products as above, so<br />

that B×B ′ ∼ n+n<br />

= W ′ −k−k ′<br />

×T k+k′ is another such product. Let R be a p-completely<br />

regular subset of B × B ′ . Then for b ′ ∈ B ′ the set Bb ′ ,R := {b ∈ B| (b, b ′ ) ∈ R}<br />

is p-completely regular in B. Also {b ∈ B| (b, b ′ ) ∈ R for some b ′ ∈ B ′ } and<br />

{b ∈ B| (b, b ′ ) ∈ R for all b ′ ∈ B ′ } are both p-completely regular.<br />

6.6 General Regularity in Mixed Products<br />

Now, we generalize Lemma 5.10.1 to the case of mixed products of copies of<br />

W and T .<br />

Consider B = B1 ×· · ·×Bn ∼ = W n−k ×T k as above. Then, we let F be the free<br />

group on n generators X1, ..., Xn. For any element b = (b1, ..., bn) ∈ B, we define<br />

a group homomorphism αb : F → W determined by αb(Xi) = bi. Now, fix an<br />

element g ∈ F , and suppose g = X n1<br />

i1<br />

for all i. Then αb(g) = b n1<br />

i1<br />

· · · Xnm<br />

im is a reduced word for g with ni ∈ Z<br />

· · · bnm.<br />

Then, we can define a length function on F ,<br />

im<br />

dependent on b, by lb p : F → M by lb p(g) = �m |nj|lp(bij j=1 ) where g is as above and<br />

|nj| represents the absolute value. Now, we can see that l b p(g) − lp(αb(g)) ∈ 2M.<br />

So for any g ∈ F and b ∈ B, we call the quantity l b p(g) − lp(αb(g)) the g-deficiency<br />

of b.<br />

80


Theorem 6.6.1. For any m ∈ M and g ∈ F , the set<br />

is a p-completely regular subset of B.<br />

Bg,m := {b ∈ B| lp(αb(g)) = l b p(g) − 2m}<br />

Proof. Fix m ∈ M and g ∈ F with g = X n1<br />

i1<br />

· · · Xnm<br />

im . Let D := {i1, ..., im}.<br />

Then let C := {j1, ..., jq} ⊂ D be such that Bj = T for all j ∈ C and Bj = W<br />

for all j ∈ D \ C. Then for any ¯r ∈ S k , we consider W ¯r g,m := f −1<br />

¯r (Bg,m). Let<br />

Zi := W if Bi = W and Zi = Tri,ni otherwise, and let (n1, ..., nk) ∈ M k . For<br />

w := (w1, ..., wn) ∈ W ¯r g,m ∩ (Z1 × · · · × Zn), we have<br />

where xik = wik if ik �∈ D and xik<br />

Kvw := �<br />

ik:ik∈D<br />

vw := x n1<br />

i1<br />

· · · xnm<br />

im<br />

= w−1<br />

ik rik wik if ik ∈ D. Thus, we have<br />

�<br />

(2|nik |lp(wik ) + 1) +<br />

Then, since ¯r is fixed, Lemma 5.10.1 proves that<br />

ik:ik�∈D<br />

(|nik |lp(wik ))<br />

W ¯r g,m ∩(Z1 ×· · ·×Zn) = {(w1, ..., wn) ∈ W n | lp(vm) = Kvm −2m}∩(Z1 ×· · ·×Zn)<br />

is p-completely regular. Therefore, by Lemma 6.5.1, since W ¯r g,m ∩ (Z1 × · · · × Zn)<br />

is p-completely regular for all ¯r ∈ S k and (n1, ..., nk) ∈ M k , then we have Bg,m is<br />

p-completely regular.<br />

81


6.7 Multivariate Poincaré Series<br />

Recall the notation from 5.6. In [8], it is shown that the Poincaré series for<br />

the set of reflections, P (T ; W ), is rational. Due to the results of this chapter,<br />

we obtain the stronger result that the multivariate Poincaré series of the set of<br />

reflections as well as any p-completely regular subset of T will be rational.<br />

Furthermore, if we have a mixed product B := B1 × · · · Bm+n ∼ = W m × T n ,<br />

let M1, ..., Mm+n be the corresponding multiplicative monomials for each Bi gen-<br />

erated by X1, ..., Xm+n with corresponding length functions lpi : Bi → Mi. Let<br />

Z[[X1, ..., Xm+n]] be the completion of the polynomial ring Z[X1, ..., Xm+n]. Then<br />

according to the transfer matrix method ([2] or [28]), for any p-completely regular<br />

subset of A ⊆ B, we have that the multivariate Poincaré series<br />

P (A; B) =<br />

�<br />

(x1,...,xm+n)∈A<br />

lp1(x1) · · · lpm+n(xm+n)<br />

is rational. In fact, P (A; B) is a finite sum of products, each of the form f1 · · · fm+n<br />

with each fi a rational expression in the variables Xi, i.e. fi ∈ Q(Xi).<br />

Example 6.7.1. Let lp : W → M, lp1 : W → M1, lp2 : W → M2, and<br />

lp3 : T → M3 be generalized length functions with M considered additively and<br />

M1, M2, and M3 considered multiplicatively.<br />

1. According to Theorem 6.6.1, for n ∈ M,<br />

Add(n) := {(x, y) ∈ W × W | lp(xy) = lp(x) + lp(y) − 2n}<br />

82


is p-completely regular in W × W . Therefore,<br />

P (Add(n); W × W ) =<br />

�<br />

(x,y)∈Add(n)<br />

lp1(x)lp2(y) =<br />

�<br />

x∈M1;x ′ ∈M2<br />

ax,x ′xx′<br />

has a rational expression, where ax,x ′ ∈ N is the number of pairs (x, y) ∈<br />

W × W with lp(xy) = lp(x) + lp(y) − 2n and lp1(x) = x and lp2(y) = x ′ .<br />

2. Also by Theorem 6.6.1, for n ∈ M,<br />

Un := {(x, t) ∈ W × T | lp(xtx −1 ) = 2lp(x) + lp(t) − 2n}<br />

is p-completely regular in W × T . It follows that<br />

P (Un; W × T ) = �<br />

(x,t)∈Un<br />

lp1(x)lp3(t)<br />

has a rational expression. Moreover, according to Remark 6.5.3, for any<br />

x ∈ W , the set<br />

Un(x) := {t ∈ T | lp(xtx −1 ) = 2lp(x) + lp(t) − 2n}<br />

is p-completely regular in T so<br />

has a rational expression.<br />

P (Un(x); T ) = �<br />

83<br />

t∈Un(x)<br />

lp3(t)


CHAPTER 7<br />

HECKE ALGEBRA MODULES<br />

The recurrence formulae given in section 3.7 look very similar to recurrence<br />

formulae for the generic Iwahori-Hecke Algebra. In this chapter, we will further<br />

discuss this algebra, and we introduce a large family of modules for the Iwahori-<br />

Hecke algebra. These modules arise from the sets T≤m. In, [18], Dyer uses modules<br />

like these to prove a weak form of Lusztig’s conjecture about the boundedness<br />

of the a-function. There is some hope that deeper regularity properties of the<br />

modules described in this section may lead to more results about the a-function.<br />

Throughout this chapter, we fix an arbitrary Coxeter system, (W, S), with<br />

reflections T , reflection cocycle N : W → P(T ), and with generalized length<br />

function lp,W where p is a surjective function from T to the set of indeterminates<br />

X as in section 3.2. In addition to X, let V be another set of indeterminates<br />

equipped with a surjective function q : T → V satisfying q(t) = q(t ′ ) if t is<br />

conjugate to t ′ . For ease of notation, we denote p(r) = pr and q(r) = qr for any<br />

r ∈ S.<br />

7.1 The Generic Iwahori-Hecke Algebra<br />

Let R denote the integral polynomial ring R = Z[X, V ] in indeterminates<br />

X ∪ V . We consider the generic Iwahori-Hecke algebra H of (W, S) over R with<br />

84


generators (as a unital R-algebra) tr for r ∈ S subject to the braid relations of<br />

(W, S)<br />

trts · · ·<br />

� �� �<br />

= tstr · · ·<br />

� �� �<br />

m(r,s) factors m(r,s) factors<br />

and the quadratic relations t 2 r = pr +qrtr. As an R-module, H is free with R-basis<br />

{tw}w∈W . The multiplication is determined by the following formulae: t1 = IdH<br />

and<br />

⎧<br />

⎪⎨ trw<br />

if r �∈ N(w)<br />

trtw =<br />

.<br />

⎪⎩ prtrw + qrtw if r ∈ N(w)<br />

By these formulae, it is clear that the set {tw| w ∈ W } is a spanning set for<br />

H as a Z[X, V ]-module. In fact, this set is also a Z[X, V ]-basis for H (see [26]).<br />

7.2 A Module for H<br />

We regard R as a free Z[V ] module with basis corresponding to M := MX,<br />

i.e. monomials in Z[X]. We will denote monomials in MX by boldface letters, e.g.<br />

m. For any m ∈ M, we denote the dual basis element of R † := HomZ[V ](R, Z[V ])<br />

by m ∗ so that m ∗ (n) = 0 if n �= m and m ∗ (m) = 1.<br />

R † is an R-module with the standard R-structure given by (rf)(r ′ ) = f(rr ′ )<br />

for f ∈ R † and r, r ′ ∈ R. Then, we note that<br />

prm ∗ ⎧ � �<br />

⎪⎨<br />

∗<br />

m if pr ≤ m<br />

pr<br />

=<br />

⎪⎩ 0 otherwise<br />

Let R ′ be the R-submodule of R † spanned by the dual basis elements of MX.<br />

Then R ′ is free as a Z[V ]-module with basis MX.<br />

Next, we can form the H -module H ′ := H ⊗R R ′ . For ease, we denote the<br />

85


element tw ⊗ m ∗ of H ′ by tw,m for m ∈ M and w ∈ W . These elements form a<br />

Z[V ]-basis of H ′ . We thus have<br />

for all r ∈ S and<br />

for all r ∈ S.<br />

⎧<br />

⎪⎨<br />

trtw,m =<br />

⎪⎩<br />

⎧<br />

⎪⎨ tw,<br />

prtw,m =<br />

⎪⎩<br />

m if pr ≤ m<br />

pr<br />

0 otherwise<br />

trw,m<br />

if r �∈ N(w)<br />

trw, m<br />

pr + qrtw,m if r ∈ N(w) and pr ≤ m<br />

qrtw,m<br />

if r ∈ N(w) and pr �≤ m<br />

For each n ∈ M, the Z[V ]-span of the elements tw,m with w ∈ W and m ≤ n<br />

is a H -submodule, called H ′<br />

n of H ′ . Provided S is finite, then H ′<br />

n has finite<br />

rank as a Z[V ]-module. For n ≥ m > 0, multiplication by m induces a short<br />

exact sequence of H -modules:<br />

0 −→ �<br />

k


tN(w)∩T≤m,m is an H -module epimorphism.<br />

2. For m ∈ M, H ′′<br />

m = ρ(H ′ m) is a submodule of H ′′ . So H ′′<br />

m is a submodule<br />

of H ′′<br />

n if m ≤ n.<br />

Proof. We define Z[V ]-linear endomorphisms p ′ r and θr on H ′′ for r ∈ S by<br />

and<br />

where<br />

⎪⎨<br />

θr(tA,m) =<br />

p ′ r(tA,m) =<br />

⎧<br />

⎪⎩<br />

tA ′ ,m<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

tA∩T ≤ m pr , m<br />

pr<br />

if pr ≤ m<br />

0 otherwise<br />

if r �∈ N(w)<br />

tA ′′<br />

r , m<br />

pr + qrtA,m if r ∈ N(w) and pr ≤ m<br />

qrtA,m<br />

if r ∈ N(w) and pr �≤ m<br />

A ′ := (rAr ∪ {r}) ∩ T≤m and A ′′<br />

r := (rAr \ {r}) ∩ T≤ m<br />

pr<br />

Now suppose that tw,m ∈ H ′ . Then on one hand we have<br />

⎧<br />

⎪⎨ ρ(tw,<br />

ρ(pr(tw,m)) =<br />

⎪⎩<br />

m<br />

pr ) if pr<br />

⎫<br />

⎪⎬ ≤ m<br />

ρ(0) otherwise<br />

⎪⎭ =<br />

⎧<br />

⎪⎨ tN(w)∩T≤ m ,<br />

pr<br />

⎪⎩<br />

m<br />

⎫<br />

⎪⎬<br />

if pr ≤ m<br />

pr<br />

⎪⎭<br />

0 otherwise<br />

and on the other hand we have that<br />

p ′ r(ρ(tw,m)) = p ′ ⎧<br />

⎪⎨ tN(w)∩T≤ m ,<br />

pr<br />

r(tN(w)∩T≤m,m) =<br />

⎪⎩<br />

m if pr ≤ m<br />

pr<br />

0 otherwise<br />

Thus, we see that p ′ r(ρ(tw,n)) = ρ(pr(tw,n)) and extending by linearity we get<br />

p ′ r(ρ(h)) = ρ(pr(h)) for all h ∈ H ′ .<br />

Next, let tw,m ∈ H ′ and tr ∈ H . Again we compute the following:<br />

87<br />

⎫<br />

⎪⎬<br />

⎪⎭ .


⎧<br />

⎪⎨<br />

ρ(trw,m) if r �∈ N(w)<br />

ρ(tr(tw,m)) =<br />

⎪⎩<br />

⎧<br />

⎪⎨<br />

=<br />

⎪⎩<br />

ρ(trw, m<br />

pr + qrtw,m) if r ∈ N(w) and pr ≤ m<br />

ρ(qrtw,m) if r ∈ N(w) and pr �≤ m<br />

tN(rw)∩T≤m,m<br />

tN(rw)∩T ≤ m pr , m<br />

pr<br />

qrtN(w)∩T≤m,m<br />

by linearity of ρ on Z[V ]. Also,<br />

⎪⎨<br />

θr(ρ(tw,m)) = θr(tN(w)∩T≤m,m) =<br />

if r �∈ N(w)<br />

+ qrtN(w)∩T≤m,m if r ∈ N(w) and pr ≤ m<br />

⎧<br />

⎪⎩<br />

tB ′ ,m<br />

if r ∈ N(w) and pr �≤ m<br />

if r �∈ N(w)<br />

tB ′′ m<br />

r , pr + qrtB,m if r ∈ N(w) and pr ≤ m<br />

qrtB,m<br />

if r ∈ N(w) and pr �≤ m<br />

where B = N(w) ∩ T≤m. Then B ′ = (rBr ∪ {r}) ∩ T≤m = N(rw) ∩ T≤m since<br />

r �∈ N(w) then N(rw) = {r} ∪ rN(w)r. Finally we have B ′′<br />

r = rBr \ {r} ∩ T≤ m<br />

pr =<br />

N(rw) ∩ T≤ m<br />

pr<br />

since r ∈ N(w) then N(rw) = rN(w)r \ {r}. These all follow from<br />

Proposition 3.8.1. Therefore we have shown that θr(ρ(tw,n)) = ρ(tr(tw,n)) for all<br />

tw,n and so extending by linearity we get θr(ρ(h)) = ρ(tr(h)) for all h ∈ H ′ .<br />

Now, ρ is clearly an epimorphism,, and so we thus have an H -module structure<br />

on H ′′ in which pr acts by p ′ r for all r ∈ S, qr acts via the natural Z[V ]-module<br />

structure on H ′′ for all r ∈ S, and tr acts by θr for all r ∈ S. Hence, 1 is true.<br />

Then 2 follows from the definition of ρ.<br />

88


7.4 The 0-Hecke Algebra<br />

Now, we regard Z[V ] as the R-module R/XR. Then for any m ∈ M we<br />

define � H ′′<br />

m := H ′′<br />

m/( � ′′<br />

n


CHAPTER 8<br />

<strong>IN</strong>ITIAL SECTIONS OF REFLECTION ORDERS<br />

In this chapter, we recall the definition due to Dyer (c.f. [11], [12], or [13])<br />

of reflection orders, which are certain total orders of the set of reflections T ,<br />

and of their initial sections, which are subsets of T that are known to lead to<br />

partial orders (twisted Bruhat orders) on W that are similar to Bruhat order.<br />

In particular, using the initial section ∅ ⊆ T we get the Bruhat order on W , and<br />

using T ⊆ T , we get the reverse Bruhat order on W . In this chapter, we determine<br />

all subsets of T that give rise to partial orders on W in the same manner. The<br />

subsets of T that have this property are those which are “locally” (at each dihedral<br />

reflection subgroup) initial sections; it has been conjectured by Dyer ([12]) that<br />

these sets coincide with the initial sections. Reflection orders are closely related to<br />

a notion of closure in the positive root system. In Chapter 9, we will examine the<br />

relationship between closure in root systems and dominance order on the positive<br />

roots. For general references about the properties of reflection orders and initial<br />

sections, see [2], [11],[12], and [13].<br />

8.1 Classification of Twisted Bruhat Orders<br />

Let (W, S) be a Coxeter system. Recall that W acts on P(T ) by w · A =<br />

N(w) + wAw −1 for any A ⊆ T and w ∈ W , where + represents symmetric<br />

90


difference. Now, for any A ⊆ T , we follow [11] by defining a directed graph Ω(W,A)<br />

with the vertex set of Ω(W,A) equal to W and edge set E(W,A) = {(tw, w)| t ∈ w·A}.<br />

We also have the following equivalent definition of E(W,A).<br />

E(W,A) = {(tw, w)| t ∈ N(w) + wAw −1 } = {(tw, w)| w −1 tw ∈ N(w −1 ) + A}<br />

= {(wt ′ , w)| t ′ ∈ N(w −1 ) + A},<br />

and we will find it convenient to use this description, E(W,A) = {(wt, w)| t ∈<br />

N(w −1 ) + A}. For x ∈ W , we have that t ∈ w · A if and only if t ∈ (wx −1 ) · x · A,<br />

and thus the map w ↦→ wx −1 defines an isomorphism Ω(W,A) ∼ = Ω(W,x·A). In<br />

addition, for A ⊆ T we can define a length function lA : W → Z in the following<br />

way:<br />

lA(v, w) = l(wv −1 �<br />

�<br />

) − 2�[N(vw<br />

−1 �<br />

�<br />

) ∩ v · A] � ∈ Z<br />

and then set lA(w) = lA(1, w). We can define a pre-order ≤A for any A ⊆ T given<br />

by the following: v ≤A w if and only if there exist t1, ..., tn ∈ T with w = vt1...tn<br />

such that ti �∈ [N((vt1...ti−1) −1 ) + A] for all i = 1, ..., n.<br />

Remark 8.1.1. The multivariate analog of such single variable length functions,<br />

lA, similar to that of Chapter 3, has not been studied but should be of interest.<br />

Definition 8.1.2. Following [12], a total order on the set T ′ in a dihedral Coxeter<br />

system, (W ′ , S ′ ) with S = {r, s}, is called a dihedral reflection order if it is either<br />

≺R or ≺ ′ R where ≺R and ≺ ′ R are defined by r ≺R rsr ≺R · · · ≺R srs ≺R s and<br />

s ≺ ′ R srs ≺′ R · · · ≺′ R rsr ≺′ R<br />

r. Then, for any Coxeter system (W, S) we call<br />

a total order ≺R on T a reflection order if the restriction ≺R � � W ′ ∩T is a dihedral<br />

reflection order for all dihedral reflection subgroups W ′ of (W, S) with respect to<br />

the canonical generators χ(W ′ ).<br />

91


Remark 8.1.3. We can also define a reflection order in terms of the root system Φ<br />

associated to the Coxeter system. This definition can be found in [2], and we will<br />

discuss this definition in more detail in Chapter 9.<br />

Recall from [12] that an initial section of a reflection order is a subset A ⊆ T<br />

such that there is a reflection order ≺R with the property that a ≺R b for all<br />

a ∈ A and b ∈ T \ A. It is shown in [11] that ≤A is a partial order of W if A is<br />

an initial section of a reflection order. Our main result in this chapter, Theorem<br />

8.1.4 below, describes all subsets A of T for which ≤A is a partial order.<br />

Let A(W,S) be the set of initial sections of reflection orders of T . Now, we<br />

define Â(W,S) = {A ⊆ T | A ∩ W ′ ∈ A(W ′ ,χ(W ′ )) ∀W ′ ⊆ W dihedral}. It has been<br />

conjectured by Dyer that A = Â. We now come to the main result:<br />

Theorem 8.1.4. Let (W, S) be any Coxeter system. The following are equivalent:<br />

1. Ω(W,A) is acyclic.<br />

2. ≤A is a partial order.<br />

3. Ω(W,A) has no cycle of length four.<br />

4. A ∈ Â.<br />

5. lA(xt) < lA(x) for all x ∈ W and t ∈ N(x −1 ) + A.<br />

8.2 Proof of the Classification<br />

In the following proofs, for any positive root α ∈ Φ + , let tα ∈ T be the<br />

corresponding reflection, and for any reflection t ∈ T , let αt ∈ Φ + be the cor-<br />

responding positive root. To begin with we investigate the dihedral case. Sup-<br />

pose (W, S) is dihedral, i.e. S = {r, s}. There is a bijection between subsets<br />

92


A ⊆ T and subsets Ψ ⊂ Φ such that Ψ ∪ −Ψ = Φ and Ψ ∩ −Ψ = ∅ given by<br />

A = AΨ = {tα| α ∈ Ψ ∩ Φ + }. We note that A−Ψ = AΨ + T .<br />

Lemma 8.2.1. For any AΨ ⊆ T , w · AΨ = Aw(Ψ).<br />

Proof. It is enough to show that this is true for r ∈ S. Now we know that<br />

r · AΨ = {r} + rAΨr = {r} + {rtαr| α ∈ Ψ ∩ Φ + } = {r} + {tr(α)| α ∈ Ψ ∩ Φ + } =<br />

{r} + {tα| α ∈ r(Ψ) ∩ r(Φ + )}. If αr ∈ Ψ then −αr ∈ r(Ψ) ∩ r(Φ + ) and so r ∈<br />

{tα| α ∈ r(Ψ) ∩ r(Φ + )}. This implies that r · AΨ = {tα| α ∈ r(Ψ) ∩ Φ + }. If αr �∈ Ψ<br />

then r �∈ {tα| α ∈ r(Ψ)∩r(Φ + )}. But αr ∈ r(Ψ) so {r}+{tα| α ∈ r(Ψ)∩r(Φ + )} =<br />

{tα| α ∈ r(Ψ) ∩ Φ + }. Thus, in both cases r · AΨ = {tα| α ∈ r(Ψ) ∩ Φ + }.<br />

Since we are considering (W, S) dihedral, we choose an orientation of the plane<br />

spanned by Φ. Then for any root α ∈ Φ we can define the neighbor of α, denoted<br />

nbr(α), to be the root lying directly next to α if we traverse the root system<br />

clockwise. Recall that S = {r, s}. Interchanging r and s if necessary, we assume<br />

without loss of generality that αr = nbr(−αs). We note that this implies that<br />

nbr(αs) �∈ Φ + .<br />

Lemma 8.2.2. Let (W, S) be dihedral. Suppose α ∈ Φ + and γ := nbr(α) ∈ Φ + .<br />

Then tαtγ = sr.<br />

Proof. Let (rsr...)n denote an element with n simple reflections listed (we note<br />

that l((rsr...)n) is not necessarily n). With this terminology, t ∈ T can be written<br />

t = (rsr...)2i+1 or t = (srs...)2i+1 for some i ≥ 0. Under the chosen orientation<br />

for Φ + , if tα = (rsr...)2i+1 then we can write tγ = (rsr...)2i+3 so that tαtγ =<br />

(rsr...)2i+1(rsr...)2i+3 = sr. Otherwise, if tα = (srs...)2i+1 (i ≥ 1) then tγ =<br />

(srs...)2i−1 and so tαtγ = sr.<br />

93


Now, given A = AΨ, we introduce two conditions that a positive system, Γ + ,<br />

of Φ can have:<br />

C1: There are roots α, nbr(α) ∈ Γ + such that α �∈ Ψ and nbr(α) ∈ Ψ.<br />

C2: There are roots β, nbr(β) ∈ Γ + such that β ∈ Ψ with nbr(β) �∈ Ψ.<br />

Lemma 8.2.3. Let (W, S) be dihedral and let A = AΨ ⊆ T .<br />

1. If the positive system r(Φ + ) has condition C1, then there exists a path 1 →<br />

x → sr in Ω(W,A).<br />

2. If the positive system r(Φ + ) has condition C2, then there exists a path sr →<br />

x → 1 in Ω(W,A).<br />

Proof. Replacing A by A + T reverses the orientation of edges in Ω(W,A) and so 2<br />

follows from 1.<br />

We now prove 1. There are two cases to consider. If α �= αs, then both α<br />

and γ := nbr(α) ∈ Φ + . So tα �∈ A and tγ ∈ A. Also, since tα �∈ {r, s}, it<br />

follows that tγ ∈ N(tα). So we have that tγ �∈ N(tα) + A. Thus, we have a path<br />

1 → tα → tαtγ = sr, where the last equality follows from Lemma 8.2.2. Now, if<br />

α = αs then nbr(α) = −αr. Since −αr ∈ Ψ we know that αr �∈ Ψ which implies<br />

r �∈ A. Also, s �∈ A and clearly r �∈ N(s). Together, we see that there is a path<br />

1 → s → sr in Ω(W,A).<br />

Lemma 8.2.4. For (W, S) dihedral, let A = AΨ ⊆ T . If there are no 4-cycles in<br />

Ω(W,A) then there is no positive system, Γ + ⊂ Φ satisfying C1 and C2.<br />

Proof. Suppose there is a positive system Γ + satisfying C1 and C2. It is clear that<br />

−Γ + also satisfies C1 and C2. Since Γ + satisfies both C1 and C2, then w(Γ + )<br />

94


will also satisfy both conditions (w with even length respects both conditions and<br />

w with odd length interchanges the conditions). Thus, we can find a w ∈ W such<br />

that r(Φ + ) = w(Γ + ) or r(Φ + ) = w(−Γ + ). So we have that r(Φ + ) satisfies C1 and<br />

C2 with respect to Aw(Ψ) = w · AΨ. Since Ω(W,A) is isomorphic to Ω(W,w·A), we can<br />

assume without loss of generality that r(Φ + ) satisfies C1 and C2 with respect to<br />

AΨ. Thus, Lemma 8.2.3 implies that we have a path sr → u → 1 → v → sr in<br />

Ω(W,A).<br />

Lemma 8.2.5. Let (W, S) be dihedral and A = AΨ ⊆ T . If there is no positive<br />

system, Γ + , of Φ satisfying C1 and C2, then A ∈ A(W,S).<br />

Proof. Let A �∈ A(W,S). Recall that the only two reflection orders on (W, S) are<br />

≺R and ≺ ′ R described in Definition 8.1.2. Since A is not an initial section of either<br />

of these orders, without loss of generality (replacing A with T \ A if necessary)<br />

we can find t0 ∈ A and t1, t2 ∈ T \ A such that t1 ≺R t0 and t2 ≺ ′ R t0, i.e.<br />

t2 ≺R t0 ≺R t1.<br />

Now, if (W, S) is finite, we can list all of the reflections and thus can find<br />

t ′ , t ′′ ∈ T such that t2 �R t ′ ≺R t0 �R t ′′ ≺R t1 with αt ′ �∈ Ψ and nbr(αt ′) ∈ Ψ,<br />

and αt ′′ ∈ Ψ and nbr(αt ′′) �∈ Ψ. Thus Φ+ satisfies C1 and C2.<br />

If (W, S) is infinite, then T = Tr ∪ Ts (disjoint union) where Tr = {t ∈ T | r ∈<br />

N(t)} and Ts = {t ∈ T | s ∈ N(t)}. Suppose Φ + does not satisfy C1 and C2.<br />

We cannot have t1, t2, t0 ∈ Tu for some u ∈ {r, s} since this would imply, by<br />

the reasoning for (W, S) finite, that Φ + satisfies C1 and C2. So, we can assume<br />

that t2, t0 ∈ Tr and t1 ∈ Ts (the case t2 ∈ Tr and t0, t1 ∈ Ts is exactly similar).<br />

Additionally, since C1 and C2 aren’t both satisfied by Φ + , we may assume the<br />

following conditions hold (see Figure 8.1):<br />

1. αt0 = nbr(αt2),<br />

95


2. t ∈ T \ A if t �R t2,<br />

3. t ∈ T \ A if t ∈ Ts and t �R t1,<br />

4. t ∈ A if t ∈ Tr and t0 �R t,<br />

5. t ∈ A if t1 ≺R t.<br />

Now, consider the positive system Γ + with simple roots αt0 and −nbr(αt0). Then<br />

Γ + satisfies C1 using the roots αt2 and αt0 = nbr(αt2). Also, by above, αt1 �∈ Ψ<br />

and nbr(αt1) ∈ Ψ (note that even if t1 = s this is true since nbr(αs) = −αr ∈ Ψ<br />

because by above αr �∈ Ψ). Using this, we see that Γ + satisfies C2 using the roots<br />

−αt1 and nbr(−αt1) = −nbr(αt1) (again see Figure 8.1).<br />

With the dihedral case taken care of, we proceed to the general case. For the<br />

remainder of the chapter, we assume that (W, S) is a general Coxeter system.<br />

Proposition 8.2.6. For all w ∈ W and A ∈ Â(W,S), w · A ∈ Â(W,S).<br />

Proof. It suffices to check the condition for w = r ∈ S. Let W ′ be a dihedral<br />

reflection subgroup. If r ∈ W ′ , then (r · A) ∩ W ′ = N(W ′ ,χ(W ′ ))(r) + r(A ∩ W ′ )r ∈<br />

A(W ′ ,χ(W ′ )) by [12]. Now, if r �∈ W ′ , then (r · A) ∩ W ′ = r(A ∩ rW ′ r)r. However,<br />

conjugation by r defines an isomorphism (W ′ , χ(W ′ )) ∼ = (rW ′ r, rχ(W ′ )r) in this<br />

case, and by assumption A∩rW ′ r ∈ A(rW ′ r,rχ(W ′ )r), thus (r·A)∩W ′ ∈ A(W ′ ,χ(W ′ )).<br />

Before, we prove Theorem 8.1.4, we provide one more important proposition,<br />

which is proven in [11], but we include a proof here as well.<br />

Proposition 8.2.7. Let A ∈ Â(W,S), x ∈ W , t ∈ T . Then lA(x, xt) > 0 iff<br />

t �∈ N(x −1 ) + A.<br />

96


αr<br />

. . .<br />

. . .<br />

αt0<br />

. . .<br />

αt2 αt1<br />

. . . . . .<br />

Figure 8.1. This is a schematic diagram of the root system for (W, S) infinite.<br />

Φ + consists of the roots which are non-negative linear combinations<br />

of αr and αs. The dotted ray through the origin represents a limit line<br />

of roots. If C1 and C2 are not both satisfied by Φ + then Ψ is pictured<br />

above with the roots in Ψ labeled by •, and those not in Ψ labeled by ◦.<br />

97<br />

. . .<br />

. . .<br />

. . .<br />

αs


Proof. We begin by proving the statement for all A ∈ Â(W,S) and t ∈ T with x = 1,<br />

i.e. lA(1, t) = lA(t) > 0 if and only if t �∈ A. If (W, S) is dihedral with S = {r, s},<br />

then we see that, without loss of generality, t = (rsr...)2n+1 with l(t) = 2n + 1.<br />

Then, N(t) = {r, rsr, ..., (rsr...)2n+1, ..., (rsr...)4n+1}. Since A ∈ Â(W,S), then if<br />

(rsr...)2i+1 ∈ A either (rsr...)2j+1 ∈ A for all j ≤ i or (rsr...)2j+1 ∈ A for all<br />

i ≤ j ≤ 2n. In particular |N(t) ∩ A| ≥ n + 1 if and only if t = (rsr...)2n+1 ∈ A.<br />

Therefore, in this case lA(t) > 0 if and only if t �∈ A.<br />

Suppose that (W, S) is not dihedral. Then if W ′ is any dihedral reflection<br />

subgroup containing t, since A ∈ Â(W,S) then A ′ := A ∩ W ′ ∈ Â(W ′ ,χ(W ′ )), and so<br />

we have lA ′(t) makes sense in the Coxeter system (W ′ , χ(W ′ )). Now, by equation<br />

2.3.1, we get that<br />

and<br />

which implies that<br />

and<br />

N(t) \ {t} = ˙�<br />

N(t) \ {t} ∩ A = ˙�<br />

l(t) − 1 = |N(t) \ {t}| = �<br />

W ′ ∈Mt<br />

W ′ ∈Mt<br />

W ′ ∈Mt<br />

lA(t) − 1 = �<br />

N(W ′ ,χ(W ′ ))(t) \ {t}<br />

� N(W ′ ,χ(W ′ ))(t) \ {t} � ∩ A,<br />

|N(W ′ ,χ(W ′ ))(t) \ {t}| = �<br />

W ′ ∈Mt<br />

W ′ ∈Mt<br />

l(W ′ ,χ(W ′ ))(t) − 1<br />

(lA∩W ′(t) − 1). (8.2.1)<br />

Now, t �∈ A, if and only if t �∈ A ∩ W ′ for all W ′ ∈ Mt. This occurs if and only<br />

if lA∩W ′(t) > 0 for all W ′ ∈ Mt and then by equation (8.2.1) is true if and only if<br />

we get lA(t) > 0 as required.<br />

98


Now, if x ∈ W ,<br />

lA(x, xt) = l(xtx −1 ) − 2|N(xtx −1 ) ∩ x · A|<br />

= l(xtx −1 ) − 2|N(xtx −1 ) ∩ 1 · (x · A)| = lx·A(1, xtx −1 )<br />

so lA(x, xt) = lx·A(xtx −1 ). Therefore, lA(x, xt) > 0 if and only if lx·A(xtx −1 ) > 0.<br />

Also, t �∈ N(x −1 ) + A if and only if xtx −1 �∈ xN(x −1 )x −1 + xAx −1 = N(x) +<br />

xAx −1 = x · A. So t �∈ N(x −1 ) + A if and only if xtx −1 �∈ x · A.<br />

Putting together the previous statements, we have that the statement of the<br />

proposition is equivalent to the statement lx·A(xtx −1 ) > 0 if and only if t �∈ x · A.<br />

Following Proposition 8.2.6, we let A ′ = x · A ∈ Â(W,S) and t ′ = xtx −1 , then this<br />

case is equivalent to lA ′(t′ ) > 0 if and only if t ′ �∈ A ′ which is true due to the x = 1<br />

case above.<br />

At this point, we can use the previous lemmas and propositions to prove The-<br />

orem 8.1.4:<br />

Proof. (1)⇔(2)⇒(3) is clear.<br />

(3)⇒(4): According to section 2.3, we know that for any dihedral subgroup W ′<br />

of W , N(W ′ ,χ(W ′ ))(w) = N(W,S)(w) ∩ W ′ . This implies Ω(W ′ ,A∩W ′ ) is a subgraph of<br />

Ω(W,A). So, if A �∈ Â(W,S), Lemma 8.2.4 and Lemma 8.2.5 imply that there exists<br />

some dihedral subgroup W ′ of W with Ω(W ′ ,A∩W ′ ) containing a cycle with four<br />

edges.<br />

(4)⇒(5): Suppose that A ∈ Â(W,S), x ∈ W and t ∈ N(x −1 )+A. Then Proposition<br />

8.2.7 implies that lA(x, xt) < 0. But by [11], lA(1, x) + lA(x, xt) = lA(1, xt) and<br />

this implies lA(xt) − lA(x) = lA(x, xt) < 0.<br />

(5)⇒(1): Suppose Ω(W,A) has a cycle. This means that xt1...tn = x for some n > 0<br />

(all ti ∈ T and with ti �∈ N((xt1...ti−1) −1 ) + A for all i). By assumption, lA(x) <<br />

99


lA(xt1) < lA(xt1t2) < ... < lA(xt1t2...tn) = lA(x) which is a contradiction.<br />

Remark 8.2.8. In case (W, S) is finite, say of order m, we can give a much simpler<br />

proof of the equivalence of (1), (2), (4), and (5). By the proof above, we have<br />

(4)⇒(5)⇒(1)⇔(2). If (4) fails, then w · A �= T for all w ∈ W (or else A = w −1 · T<br />

which is an initial section). So we can choose t �∈ w · A. It follows that we can<br />

recursively choose t1, ..., tm such that t1 �∈ A and ti �∈ ti−1 · · · t1 · A for i = 2, ..., m.<br />

This gives us the following path in Ω(W,A): 1 → t1 → . . . → tm · · · t1. However, this<br />

path has m + 1 elements and W has m elements, so there must be two elements<br />

in the path that are the same thus creating a cycle.<br />

100


CHAPTER 9<br />

CLOSURE <strong>IN</strong> THE ROOT SYSTEM<br />

This chapter recalls a notion of closure on the positive roots that has been<br />

studied by Dyer [16]. A conjectural relation, due to Dyer, between these closures<br />

and the sets T≤m (m ∈ N), as defined in section 2.4 and generalized in Chapter<br />

3, would lead to a new and potentially interesting normal form for elements of<br />

Coxeter groups, via known special cases of other conjectures of Dyer ([17]) involv-<br />

ing initial sections and these closures. The results of the previous chapter afford<br />

the possibility of bringing the study of twisted Bruhat orders to bear on these<br />

conjectures.<br />

9.1 2-closure<br />

Let (W, S) be a fixed Coxeter system. Recall that we denote the roots and<br />

positive roots by Φ and Φ + respectively. We begin by describing a notion of<br />

closure on the subsets of positive roots, Φ + . This closure is similar to the notion<br />

of closure in the crystallographic root system of a finite Weyl group as described<br />

in [3].<br />

Definition 9.1.1. Let Γ ⊆ Φ + . We say that Γ is 2-closed if for all α, β ∈ Γ, we<br />

have (R≥0α + R≥0β) ∩ Φ + ⊆ Γ.<br />

101


Remark 9.1.2. Let TΓ ⊆ T be the subset of reflections corresponding to Γ via the<br />

canonical bijection. Then, 2-closure is the same as the following. TΓ is 2-closed<br />

if for all tα, tβ ∈ TΓ with W ′ = 〈tα, tβ〉, {t ∈ W ′ ∩ T | tα ≺R t ≺R tβ} ∪ {t ∈<br />

W ′ ∩ T | tβ ≺R t ≺R tα} ⊂ TΓ for all reflection orders ≺R on T (see Chapter 8).<br />

Definition 9.1.3. We say a subset Γ ⊆ Φ + is biclosed if Γ and Φ + \ Γ are both<br />

2-closed.<br />

In particular, initial sections of reflection orders are biclosed subsets. Moreover,<br />

Â(W,S) as defined in Chapter 8 is the set of all subsets of T corresponding to<br />

biclosed subsets of Φ + under the natural bijection Φ + ↔ T .<br />

9.2 More types of subsets of Φ +<br />

We introduce three more types of subsets of Φ + . These notions are due to<br />

Dyer, and further information about them can be found in [17].<br />

Definition 9.2.1. Let Γ ⊆ Φ + .<br />

1. We say Γ is balanced if for all α ∈ Γ and W ′ ∈ Msα, then β ∈ Φ +<br />

W ′ such<br />

that l(W ′ ,χ(W ′ ))(sβ) < l(W ′ ,χ(W ′ ))(sα) implies that β ∈ Γ.<br />

2. We say Γ is bipedal if for all α ∈ Γ and W ′ ∈ Msα with α �∈ ΠW ′, ΠW ′ ⊂ Γ.<br />

3. We say Γ is unipedal if for all α ∈ Γ and W ′ ∈ Msα, then ΠW ′ = {β, γ}<br />

implies that either β ∈ Γ or γ ∈ Γ.<br />

The following proposition is then clear from the definitions.<br />

Proposition 9.2.2. Suppose Γ ⊆ Φ + .<br />

1. If Γ is balanced then Γ is bipedal.<br />

102


2. If Γ is bipedal, then Γ is unipedal.<br />

3. If Φ + \ Γ is closed, then Γ is unipedal.<br />

4. If Γ ⊆ Φ + is bipedal and ∆ ⊆ Φ + is unipedal, then Γ ∩ ∆ is unipedal.<br />

For any subset Γ ⊆ Φ + , we let ¯ Γ denote the 2-closure of Γ, that is the smallest<br />

2-closed set containing Γ. The following is conjectured by Dyer in [17].<br />

Conjecture 9.2.3. If Γ ⊆ Φ + is unipedal, then ¯ Γ is biclosed.<br />

This would strengthen the following conjecture, also in [17], parts of which<br />

were raised in [13, Remark 2.1.4].<br />

Conjecture 9.2.4. Let A := {Γ ⊆ Φ + | Γ is biclosed}. Then A is a complete<br />

lattice with �<br />

i∈I Γi = �<br />

i∈I Γi for an arbitrary family {Γi}i∈I ⊂ A.<br />

The last conjecture extends the known fact that W under weak order is a meet<br />

semi-lattice (see [2] for example). In fact, [17] proves the following:<br />

Theorem 9.2.5. Let Φw := Φ + ∩ w(−Φ + ). If Γ ⊆ Φw for some w ∈ W and Γ is<br />

unipedal, then ¯ Γ = Φx ⊆ Φw for some x ∈ W .<br />

The previous theorem provides a new formula for the join (when defined) of<br />

elements of W in weak order. For x, y, z ∈ W , x∨y = z if and only if Φx ∪ Φy = Φz.<br />

There is a similar (proven) formula for meet, which we do not give.<br />

9.3 A Conjectural Normal Form for Elements of W<br />

Our conjectural normal form for elements of W depends on the following con-<br />

jecture of Dyer in [17].<br />

103


Conjecture 9.3.1. Let m ∈ N and let Φ +<br />

≤m = {α ∈ Φ+ | h ∞ (sα) ≤ m} be the<br />

subset of positive roots corresponding to T≤m. Then Φ +<br />

≤m is balanced.<br />

In fact, we will only need the weaker conjecture that Φ +<br />

≤m is bipedal (see part<br />

1 of Proposition 9.2.2). Then we have the following theorem.<br />

Theorem 9.3.2. Let m ∈ N such that Φ +<br />

≤m<br />

is bipedal.<br />

1. For any x ∈ W , there is a unique x ′ ∈ W such that Nm(x ′ ) = Nm(x),<br />

and any y ∈ W with Nm(y) ⊇ Nm(x) can be written y = x ′ y ′′ with l(y) =<br />

l(x ′ ) + l(y ′′ ).<br />

2. Any 1 �= w ∈ W can be uniquely written as w = w1 · · · wn for certain<br />

wi ∈ W \ {1} with i = 1, ..., n and n ≥ 1 such that l(w) = l(w1) + · · · + l(wn)<br />

and wi = (wi · · · wn) ′ for each i where x ↦→ x ′ is as in 1.<br />

Proof. Let x ∈ W . Then Φx is biclosed and thus unipedal by part 3 of Proposition<br />

9.2.2. By assumption, Φ +<br />

≤m is bipedal and so Φx ∩ Φ +<br />

≤m<br />

is unipedal by part 4 of<br />

9.2.2. Therefore, according to Theorem 9.2.5, we see that Φx ∩ Φ +<br />

≤m<br />

= Φx ′ for<br />

some x ′ ∈ W . If y ∈ W with Nm(y) ⊇ Nm(x), then Φx ∩ Φ +<br />

≤m ⊆ Φy ∩ Φ +<br />

≤m ⊆ Φy.<br />

Therefore we get that<br />

Φx ′ = Φx ∩ Φ +<br />

≤m ⊆ Φy = Φy,<br />

and so y = x ′ y ′′ with l(y) = l(x ′ ) + l(y ′′ ) due to well known properties of the weak<br />

order on (W, S), which can be found in [2]. This proves 1, and 2 follows from 1<br />

inductively.<br />

The next proposition shows that the previous theorem holds in many basic<br />

Coxeter systems.<br />

Proposition 9.3.3. Let (W, S) be a Coxeter system, and let m ∈ N.<br />

104


1. If (W, S) is finite or affine, then Φ +<br />

≤m<br />

is bipedal.<br />

2. If (W, S) is right-angled, that is m(s, s ′ ) ∈ {1, 2, ∞} for all (s, s ′ ) ∈ S × S,<br />

then Φ +<br />

≤m<br />

is bipedal.<br />

Proof. If (W, S) is finite, then Φ +<br />

≤m = Φ+ for all m ∈ N, and so the result is<br />

trivial. Suppose that (W, S) is affine. Then there is a well-known description of<br />

the root system of (W, S) in terms of the root system, Ψ, of the associated finite<br />

Weyl group, (W ◦ , S ◦ ). We remind the reader of this now (see [24]). We let U<br />

denote the span of Ψ, and let ∆ be the simple roots for Ψ. We define V := U +Rδ<br />

to be a vector space with U as a subspace of codimension one. Then, we have<br />

Φ = {α + nδ| α ∈ Ψ; n ∈ Z},<br />

Φ + := {α + nδ| α ∈ Ψ + ; n ≥ 0} ∪ {−α + nδ| α ∈ Ψ + ; n ≥ 1},<br />

and Π = ∆∪{δ − �α} is the set of simple roots for Φ where �α is the highest root for<br />

Ψ. It is well known that every reflection, sα+nδ (α ∈ Ψ), is contained in a unique<br />

infinite maximal dihedral reflection subgroup, namely the group generated by<br />

{sα, sδ−α}. It follows from inspection in the dihedral case then that for α ∈ Ψ + we<br />

get h ∞ (sα+nδ) = h ∞ (s−α+(n+1)δ) = n. Now, let m ∈ N, and suppose α+eδ ∈ Φ≤m.<br />

Let sα+eδ ∈ W ′ be a dihedral reflection subgroup with ΠW ′ = {β + nδ, γ + kδ}.<br />

Then, we have α + eδ = a(β + nδ) + b(γ + kδ) with a, b ∈ Z>0. Therefore,<br />

e = an + bk, and so an + bk ≤ m with a, b ∈ Z>0. This implies that n ≤ m<br />

and k ≤ m, and thus h ∞ (β + nδ) ≤ n ≤ m and h ∞ (γ + kδ) ≤ k ≤ m. Hence<br />

ΠW ′ ⊂ Φ≤m as required.<br />

To show 2, suppose that (W, S) is right-angled. We know that every dihedral<br />

reflection subgroup either has order 4 or ∞. Then for any t ∈ T and W ′ ∈ Mt<br />

105


with |W ′ | = 4, we must have t ∈ χ(W ′ ) and so h(W ′ ,χ(W ′ ))(t) = 0. Therefore, we<br />

must have<br />

h ∞ (t) = �<br />

W ′ ∈Mt<br />

|W ′ |=∞<br />

h(W ′ ,χ(W ′ ))(t) = �<br />

W ′ ∈Mt<br />

h(W ′ ,χ(W ′ )) = h(t).<br />

Now suppose α ∈ Φ +<br />

≤m and W ′ ∈ Msα. Let {β, γ} = ΠW ′. Then, by previous<br />

statement and the fact that sβ ≤∅ sα in Bruhat order, we get that<br />

h ∞ (sβ) = h(sβ) ≤ h(sα) = h ∞ (sα) ≤ m,<br />

and so β ∈ Φ +<br />

≤m , and similarly for γ as required.<br />

This normal form described by Theorem 9.3.2 could be called the m-automatic<br />

normal form for reasons we now describe. Suppose m ∈ N is fixed. Let p : T → N<br />

be given by p(t) = 1 for all t ∈ T . Then lp = l, and we can construct the m-<br />

canonical automata from Chapter 5, Mm. If we input a reduced expression for w ∈<br />

W in to the automata starting at (m, Nm(1)), we end on the node corresponding<br />

to (m, Nm(w)). According to the conjecture, there is a unique minimal element<br />

x ∈ W with (m, Nm(x)) = (m, Nm(w)). This element is w1. Then, we repeat this<br />

process with w −1<br />

1 w and so on until we are finished. Since this normal form comes<br />

from the m-canonical automata, the name is justified.<br />

Remark 9.3.4. Let p be as in Chapter 3 with associated set M. For any m ∈ M,<br />

we can define Φ +<br />

≤m := {α ∈ Φ+ | hW,p,M∞(sα) ≤ m}. Following Conjecture 9.3.1, it<br />

is reasonable to conjecture that Φ +<br />

≤m is balanced (or weaker, bipedal). We include<br />

the following example to show that this is in fact not the case in general.<br />

Example 9.3.5. Let (W, S) be the affine Weyl group of type ˜ B2. That is S =<br />

106


{r, s, t} with m(r, s) = m(s, t) = 4 and m(r, t) = 2. Due to well known facts,<br />

there are 8 elementary reflections. We list this set below.<br />

T0 = {r, rsr, srs, s, sts, tst, t, rtstr}.<br />

Now, there are three conjugacy classes of reflections. Define p : T → N 3 by<br />

p(u) = (1, 0, 0) if u is conjugate to r, p(u) = (0, 1, 0) if u is conjugate to s, and<br />

p(u) = (0, 0, 1) if u is conjugate to t; let lp : W → N 3 and hW,p,M∞ : T → N 3 be<br />

the corresponding length and infinite height functions.<br />

Next, let Φ +<br />

≤(0,1,0) be as defined in the previous remark. Define<br />

W ′ := 〈rstsr, rsrtstrsr, srstsrs, s〉<br />

so that χ(W ′ ) = {rstsr, s}. Then the reflections of W ′ have the following infinite<br />

heights: hW,p,M∞(s) = (0, 0, 0), hW,p,M∞(rstsr) = (1, 0, 0), hW,p,M∞(rsrtstrsr) =<br />

(0, 1, 0), and hW,p,M∞(srstsrs) = (1, 0, 0). Then clearly, αrsrtstrsr ∈ Φ +<br />

≤(0,1,0) and<br />

rsrtstrsr ∈ W ′ , but αrstsr ∈ ΠW ′ and αrstsr �∈ Φ +<br />

≤(0,1,0) . Therefore Φ+<br />

(0,1,0)<br />

is not<br />

bipedal. We note that (according to Proposition 9.3.3) this example does not<br />

provide a counterexample to Conjecture 9.3.1.<br />

107


CHAPTER 10<br />

HYPERBOLIC <strong>COXETER</strong> <strong>GROUPS</strong><br />

The standard classification of finite reflection groups (or finite Coxeter sys-<br />

tems) and affine Coxeter systems can be found in [23] or [24]. Affine Coxeter<br />

systems are in some natural sense the smallest infinite Coxeter groups (e.g. they<br />

have only polynomial growth, see Chapter 12). In this chapter, we will consider<br />

the “next” class of infinite Coxeter systems known as hyperbolic Coxeter systems.<br />

There is a very well known, case-by-case, classification of hyperbolic Coxeter sys-<br />

tems, which can be found in [23, chapter 6]. In this chapter, we discuss several<br />

characterizations of hyperbolic Coxeter systems by properties of their Coxeter<br />

graphs. According to [6], it is well known that any irreducible, infinite, non-affine<br />

Coxeter system contains a standard parabolic subsystem that is a hyperbolic Cox-<br />

eter system; however, we could not find a proof in the literature, and the proof of<br />

the characterizations we include allows us to provide a case-free proof of this fact.<br />

For the remainder of this chapter, we will assume that S is a finite set.<br />

10.1 Coxeter Graphs<br />

Fix a Coxeter system, (W, S). We let Γ := Γ(W,S) be the associated Coxeter<br />

graph; that is, Γ is the undirected, labeled graph with vertex set S and edges<br />

(si, sj) whenever m(si, sj) ≥ 3 and each edge labeled by the corresponding order<br />

108


m(si, sj). To this graph, we associate the matrix of the bilinear form defined in<br />

section 2.2. That is, A is a symmetric matrix with Ai,j := Asi,sj = (αi, αj). We<br />

say that (W, S) is irreducible if Γ is connected. For J ⊆ S, let ΓJ be the full<br />

subgraph of Γ with vertices corresponding to J. We may abuse notation and say<br />

that s ∈ S − J is connected to J if s is connected to ΓJ.<br />

10.2 Symmetric Matrices and Sylvester’s Law<br />

For any real symmetric matrix over a finite dimensional vector space, we define<br />

the inertia of A to be the triple (a, b, c) where a represents the number of positive<br />

eigenvalues of A, b represents the number of negative eigenvalues of A, and c is the<br />

number of zero eigenvalues of A. We will abuse notation and say that a Coxeter<br />

system has inertia (a, b, c) if the corresponding matrix A has inertia (a, b, c). We<br />

say that an n × n symmetric matrix A is positive definite if it has inertia (n, 0, 0).<br />

We say that a n × n symmetric matrix is positive semidefinite if it has inertia<br />

(a, 0, c), i.e. it has only non-negative eigenvalues. If A is positive definite or<br />

positive semi-definite then we say A is of positive type. We state the following<br />

fact of real symmetric matrices (see for example [22]):<br />

Proposition 10.2.1 (Sylvester’s law of inertia). If A is an n × n real symmetric<br />

matrix with inertia (a, b, c), then A is congruent to the diagonal matrix D =<br />

Ia ⊕ −Ib ⊕ 0c.<br />

10.3 Hyperbolic Coxeter Systems<br />

Let (W, S) be an irreducible Coxeter system with associated graph Γ (so Γ<br />

is connected) and matrix A (representing the bilinear form as in 2.2). Suppose<br />

S is finite and |S| = n. Following [23], we know that W is finite if and only<br />

109


if A is positive definite, that is, if A has inertia (n, 0, 0). Also, we know that<br />

W is affine if and only if the inertia is (n − 1, 0, 1). We define an irreducible<br />

Coxeter system to be hyperbolic if it has inertia (n − 1, 1, 0) and (v, v) < 0<br />

for all v ∈ C := {λ ∈ V | (λ, αs) > 0 ∀s ∈ S}. Also, we have the following<br />

characterization of hyperbolic Coxeter systems:<br />

Proposition 10.3.1. [23, Prop. 6.8] Let (W, S) be an irreducible Coxeter system,<br />

with graph Γ and associated bilinear form (−, −) with matrix A. Then (W, S) is<br />

hyperbolic if and only if the following two conditions are satisfied:<br />

1. (−, −) (or equivalently A) is non-degenerate, but not positive definite.<br />

2. For each s ∈ S, the Coxeter graph Γ ′ , obtained by removing s from Γ, is of<br />

positive type.<br />

10.4 Non-affine Coxeter Systems<br />

At this point, we prove a quick technical lemma that will be needed in what is to<br />

follow. From this point on, we may use (W, S), W and Γ = Γ(W,S) interchangeably<br />

to represent the corresponding Coxeter system. There is no confusion in doing this<br />

as every finite rank (W, S) has a corresponding Coxeter graph and every simple<br />

finite, undirected, edge-labeled graph (labels in N≥3 ∪ {∞}) gives rise to a finite<br />

rank Coxeter system. We say a graph Γ is of finite (resp. affine) type if the<br />

corresponding Coxeter system is finite (resp. affine).<br />

Lemma 10.4.1. Suppose that (W, S) is an infinite, irreducible, non-affine Coxeter<br />

system. Let |S| = n. Suppose that there exists s ∈ S such that the standard<br />

parabolic subsystem (WS\{s}, S \ {s}) is a connected affine Coxeter system. Then<br />

the inertia of (W, S) is (n − 1, 1, 0).<br />

110


Proof. Let (W, S) be as in the statement. Thus, there exists s ∈ S such that<br />

S \ {s} is of connected affine type. By removing another simple reflection s ′ we<br />

must get a finite Coxeter system of rank n − 2, and so (W, S) must have at least<br />

n − 2 positive eigenvalues.<br />

Now, we order S so that S = {s1, ..., sn−1, sn} where sn = s. Then, we let A ′<br />

be the (n − 1) × (n − 1) leading principal minor of the matrix A associated to<br />

(W, S) and Γ. Since A ′ is the matrix for (WS\{s}, S \ {s}) which is of affine type<br />

of rank n − 1, it has inertia (n − 2, 0, 1) and there exists an isotropic root called<br />

δ. Additionally, we know that δ = � n−1<br />

i=1 kiαsi with ki > 0 for all i (c.f. [24]). By<br />

Proposition 10.2.1, we know that there exists an invertible matrix C such that<br />

C tr A ′ C = In−2 ⊕ 01. We can lift this to a matrix D = C ⊕ I1 so that<br />

⎛<br />

B := D tr ⎜<br />

AD = ⎜<br />

⎝<br />

In−2 0<br />

.<br />

0<br />

a1<br />

.<br />

a n−2<br />

0 · · · 0 0 c<br />

a1 · · · an−2 c 1<br />

where c = (αs, δ) and ai ∈ R for all i. Since (W, S) is irreducible, Γ is con-<br />

nected and so there exists some i �= n such that (αsn, αsi ) < 0. Thus by the<br />

characterization of δ above we see that (αsn, δ) �= 0, i.e. c �= 0.<br />

Next, we want to find the determinant of B. To do this, we expand the matrix<br />

along the (n−1)’st row noting that we have to multiply c by −1 due to its position<br />

in the matrix. We thus get that det(B) = −c · det(B ′ ) where<br />

⎛<br />

B ′ ⎜<br />

:= ⎜<br />

⎝<br />

In−2 0<br />

.<br />

0<br />

a1 · · · an−2 c<br />

111<br />

⎞<br />

⎟<br />

⎠<br />

⎞<br />

⎟<br />


Now, we can find the determinant of B ′ easily by expanding along the (n − 1)’st<br />

column of B ′ (where we will multiply c by +1 due to its position in the matrix.<br />

This gives us that det(B ′ ) = c det(In−2). Putting the results together we see that<br />

det(B) = −c det(B ′ ) = −c 2 det(In−2) = −c 2 .<br />

Since B has at least n−2 positive eigenvalues and it has negative determinant (due<br />

to the fact that c �= 0), we know that exactly one of the remaining eigenvalues must<br />

be negative. Indeed, if both were negative or both were positive, the determinant<br />

of B would be positive. Thus B has inertia (n − 1, 1, 0). Using Proposition 10.2.1,<br />

we can conclude that A must have inertia (n − 1, 1, 0) as well.<br />

10.5 A Class of Coxeter Systems<br />

We define a class of Coxeter systems as follows.<br />

Definition 10.5.1. Let H be the class of all irreducible, finite rank, infinite,<br />

non-affine Coxeter systems, (W, S), with Coxeter graph Γ such that every proper<br />

subgraph Γ ′ ⊂ Γ is the disjoint union of graphs that are Coxeter graphs of only<br />

finite or affine type.<br />

Lemma 10.5.2. Every hyperbolic Coxeter system is in H.<br />

Proof. By Proposition 10.3.1 part (2) we see that any hyperbolic Coxeter system<br />

must be in H. In particular, the hyperbolic Coxeter systems are precisely those<br />

in H with inertia (n − 1, 1, 0).<br />

10.6 A Smaller Class of Coxeter Systems<br />

We now define a certain subclass of H that will be interesting to study.<br />

112


Definition 10.6.1. Let h ⊆ H be the class of all Coxeter systems, (W, S), in H<br />

that have Γ satisfying exactly one of the following (it will be clear that a graph<br />

cannot satisfy both):<br />

(h1) There exists s ∈ S such that Γ − {s} is a disjoint union of Coxeter graphs<br />

of finite type.<br />

(h2) Γ − {s} is a connected Coxeter graph of affine type for all s ∈ S.<br />

Example 10.6.2. Notice that any infinite rank 3 Coxeter system can be described<br />

by one of the two following graphs (note: we color a vertex ◦ if we intend to refer<br />

to it below):<br />

(a)<br />

•�<br />

���<br />

l �<br />

� k<br />

�<br />

�<br />

◦ •<br />

m<br />

(b)<br />

m k<br />

• ◦ •<br />

with l ≥ m ≥ k ≥ 3. For arbitrary m and k, if we remove vertex ◦ from graph<br />

(b) we will be left with the Coxeter graph for a dihedral Coxeter system of order<br />

4. Thus, graph (b) is clearly of type (h1) (provided it is infinite and non-affine).<br />

For graph (a), if k �= ∞, then removing vertex ◦ leaves us with the Coxeter<br />

graph for a finite dihedral Coxeter system, and thus (a) is of type (h1) (provided<br />

it is not affine). If k = ∞ then we necessarily have m = ∞ and l = ∞. Therefore,<br />

∞<br />

removing any vertex from (a) gives us the graph • • , which is clearly of affine<br />

type. Thus any infinite, non-affine irreducible rank 3 Coxeter system is in h.<br />

Lemma 10.6.3. Any Coxeter system in h is hyperbolic.<br />

Proof. Let (W, S) be in h ⊆ H and let |S| = n, i.e. the rank of (W, S) is n. Let Γ<br />

be the Coxeter graph of (W, S) and let A be the matrix associated to (−, −) (or<br />

similarly to Γ).<br />

113


Case 1: (W, S) is of type (h1). Then, since there exists s ∈ S such that (WS−{s}, S −<br />

{s}) is of finite type and so the associated bilinear form and matrix A ′ must<br />

have inertia (n−1, 0, 0). Thus A must have at least n−1 positive eigenvalues.<br />

If A has inertia (n, 0, 0) (resp. (n − 1, 0, 1)) then (W, S) would be of finite<br />

type (resp. affine type) and thus (W, S) �∈ H, which is a contradiction (since<br />

h ⊆ H by definition). Therefore A must have inertia (n − 1, 1, 0) forcing<br />

(W, S) to be hyperbolic by Proposition 10.3.1.<br />

Case 2: (W, S) is of type (h2). Then since S −{s} is of connected affine type Lemma<br />

10.4.1 implies that the inertia of (W, S) is (n − 1, 1, 0). In particular (−, −)<br />

is non-degenerate. Furthermore, since for each r ∈ S we have Γ − {r} is of<br />

positive type, Proposition 10.3.1 implies that (W, S) is hyperbolic.<br />

10.7 Parabolic Subsystems of Non-affine Coxeter Systems<br />

Lemma 10.7.1. Any infinite, irreducible, finite rank, non-affine Coxeter system<br />

(W, S) contains a standard parabolic subsystem in the class H.<br />

Proof. Let (W ′ , S ′ ) be a minimal rank, irreducible, infinite, non-affine parabolic<br />

subsystem of (W, S), which must exist. Then since (W ′ , S ′ ) is minimal rank, if we<br />

remove any vertex s then (WS ′ \{s}, S ′ \ {s}) cannot contain an infinite, non-affine<br />

parabolic subsystem. Thus, (WS ′ \{s}, S ′ \ {s}) must be of finite or affine type.<br />

10.8 Characterizations of Hyperbolic Coxeter Systems<br />

Lemma 10.8.1. Let (W, S) be in the class H. Then (W, S) is in the class h.<br />

114


Proof. Suppose that (W, S) is not of type (h1) and let Γ := ΓS. Then because<br />

(W, S) is in class H, we know that (WS\{s}, S \ {s}) is of affine type for every<br />

s ∈ S. Suppose that s ∈ S and ΓS\{s} is not connected. Let Γ0, Γ1, ..., Γn be<br />

the connected components (so n ≥ 1). Now, since ΓS\{s} is of affine type we can<br />

assume without loss of generality that Γ0 is of affine type. Then Γ ′ = Γ0 ∪ {s} is<br />

irreducible (since (W, S) is irreducible) and is properly contained in Γ0 ∪{s}∪{s1}<br />

where s1 ∈ Γ1. Thus Γ ′ is properly contained in Γ. However, Γ ′ cannot be of finite<br />

or affine type or else its subgraph Γ0 would be finite, contradicting that Γ0 is<br />

affine. Therefore if ΓS\{s} is not connected then Γ properly contains Γ ′ which is<br />

infinite and non-affine, contradicting the assumption that (W, S) is in H. Thus,<br />

we have shown that if (W, S) is not of type (h1), then for every s ∈ S, Γ \ {s} is<br />

of connected affine type, i.e. (W, S) is of type (h2).<br />

Theorem 10.8.2. The following hold:<br />

1. The classes h and H both coincide with the class of hyperbolic Coxeter sys-<br />

tems.<br />

2. Any finite rank, irreducible, infinite, non-affine Coxeter system contains a<br />

standard parabolic subsystem of hyperbolic type.<br />

Proof. By Lemma 10.5.2, we know that all hyperbolic Coxeter systems are con-<br />

tained in H. Then, by Lemma 10.8.1 we know that H is contained in h (so clearly<br />

h and H are equal). Finally, by Lemma 10.6.3 we know that every element of h<br />

is a hyperbolic Coxeter system. Thus we have equality throughout and 1 follows.<br />

For 2, we use part 1 of the theorem just proved along with Lemma 10.7.1.<br />

115


CHAPTER 11<br />

LARGE REFLECTION SUB<strong>GROUPS</strong><br />

By a universal Coxeter system, we mean one with no braid relations (the un-<br />

derlying group is a free product of cyclic groups of order 2). The main result<br />

of this chapter is that all irreducible, non-affine Coxeter systems have universal<br />

reflection subgroups of arbitrarily large rank. To prove this, we use the imagi-<br />

nary cone, which was introduced by Dyer ([14]) in order to prove a conjectured<br />

characterization of coverings in dominance order (see Remark 2.5.3). Dyer raised<br />

the question of whether the imaginary cone of an infinite, irreducible, non-affine<br />

Coxeter system can be approximated by imaginary cones of universal reflection<br />

subgroups. We show this is the case for hyperbolic Coxeter groups, and then our<br />

main result follows from Chapter 10. Many of the auxiliary results in this chapter<br />

can be found in [15]. Since we will define a few different cones, we need some<br />

terminology regarding convexity and polyhedral cones. For a reference on such<br />

terms, consult [1].<br />

11.1 Convex Sets<br />

In this section, we describe our terminology involved with convex sets. Let V<br />

be a real vector space.<br />

Definition 11.1.1. We say that x is a convex combination of x1, ...xn ∈ V if there<br />

exists λ1, ..., λn ∈ R such that<br />

116


1. x = λ1x1 + · · · + λnxn<br />

2. λ1 + · · · + λn = 1<br />

3. λi ≥ 0 for all i.<br />

For any set M we call the set of all convex combinations of elements of M the<br />

convex hull of M and we denote it by conv(M).<br />

For any two elements x, y ∈ V , we denote<br />

[x, y] := {λx + (1 − λ)y| 0 ≤ λ ≤ 1}.<br />

We say a set C ⊂ V is convex if, for all x, y ∈ C (x �= y), [x, y] ⊂ C.<br />

11.2 A Metric on Convex Sets<br />

Assume that V is finite dimensional, so that V ∼ = R n . Then V has the usual<br />

distance function as a metric d : V × V → R. Let Y ⊆ V and x ∈ V . Define<br />

Then for any X ⊆ V and Y ⊆ V , we define<br />

d(x, Y ) = inf d(x, y) (11.2.1)<br />

y∈Y<br />

d(X, Y ) = sup d(x, Y ) (11.2.2)<br />

x∈X<br />

Definition 11.2.1. Let K := {X ⊆ V | X is compact}.<br />

The following proposition can be found in [25, §21 VII].<br />

117


Proposition 11.2.2. Consider the distance function defined in (11.2.2). We<br />

create a new distance function on K given by dK(X, Y ) = max{d(X, Y ), d(Y, X)}.<br />

Then dK defines a metric on K.<br />

The metric defined above is known as the Hausdorff metric (or Hausdorff<br />

distance).<br />

11.3 Approximation of a Sphere by Polytopes<br />

Now, suppose that S d is the d-sphere sitting inside of R d+1 . We want to show<br />

that S d can be approximated by the boundary of a convex hull of a finite number<br />

of points. For any set A ⊆ R n , let ∂A denote the boundary of A. To make the<br />

previous statement more concrete, we have the following lemma.<br />

Lemma 11.3.1. There exists a sequence of compact, convex sets Pm with Pm =<br />

conv({x1, ..., xm}) where {x1, ..., xm} ⊂ S d and m < ∞ such that dK(∂Pm, S d ) → 0<br />

as m → ∞, i.e. ∂Pm → S d in the metric space K.<br />

Proof. Let ɛ > 0 be given. Now, for any point z ∈ R d+1 , define Ba(z) to be the<br />

open ball of radius a centered at z. Then we define<br />

U := {B ɛ<br />

2 (z)| z ∈ Sd }.<br />

Clearly S d ⊂ �<br />

U∈U U. Since Sd is compact, we can find a finite open subcover of<br />

U call it U ′ := {U1, ..., Um} (where m = mɛ is dependent on ɛ).<br />

Now, each Ui is a ball of radius ɛ/2 centered at xi ∈ S d . Thus, we have a finite<br />

set of points {x1, ..., xm} ⊂ S d . Let Pm := conv({x1, ..., xm}). Let ∂Pm denote the<br />

boundary of Pm.<br />

118


Suppose y ∈ ∂Pm. Now, there exists a minimal set I := {xi1, ..., xik } ⊂<br />

{x1, ..., xm} such that y ∈ conv({xi1, ..., xik } = conv(I) ⊂ ∂Pm. Then, we can<br />

choose a supporting affine hyperplane H of Pm such that I ⊂ Pm ∩ H. Let H +<br />

and H − represent the positive and negative half-spaces associated to H. Suppose,<br />

without loss of generality, that Pm ⊂ H − . Let H ′ := int(H + ) and let u be an<br />

outer unit normal vector of H (u points into H + ). By assumption, Pm ∩ H ′ = ∅<br />

and so ∂Pm ∩ H ′ = ∅ and {x1, ..., xm} ∩ H ′ = ∅.<br />

Next, for the sake of contradiction, we suppose that d(y, S d ) ≥ ɛ. Then let<br />

L = H + ɛ<br />

ɛ<br />

u be the affine hyperplane parallel to H with d(H, L) = 2 2 and L ⊂ H′ .<br />

It follows that S d ∩ L + �= ∅, where L + is the positive half-space associated to L.<br />

Indeed, if S d ∩L + = ∅, then S d ⊆ L − and so d(Pm ∩H, S d ) ≤ d(H, L). This would<br />

imply that<br />

d(y, S d ) ≤ sup d(x, S<br />

x∈Pm∩H<br />

d ) = d(Pm ∩ H, S d ) ≤ d(H, L) = ɛ<br />

2<br />

which is not true by assumption. Thus, we let z ∈ Sd ∩ L + . Then B ɛ (z) ⊂ H′<br />

2<br />

by construction. However, z ∈ Sd so z ∈ Ui for some Ui ∈ U ′ so xi ∈ B ɛ (z) ⊂ H′<br />

2<br />

which is a contradiction. Therefore, d(y, S d ) < ɛ.<br />

< ɛ,<br />

Since y is arbitrary, we can conclude that d(∂Pm, S d ) < ɛ. Now, suppose<br />

y ∈ S d , then y ∈ Ui for some i. Thus,<br />

d(y, ∂Pm) = inf d(y, z) ≤ d(y, xi) <<br />

z∈∂Pm<br />

ɛ<br />

2 .<br />

Again, since y is arbitrary, we can conclude that d(Sd , ∂Pm) < ɛ . Thus we have<br />

2<br />

determined that dK(∂Pm, Sd ) = max{d(Sd , ∂Pm), d(∂Pm, Sd )} < max{ ɛ , ɛ} = ɛ<br />

2<br />

as desired.<br />

119


11.4 Cones in Coxeter Systems<br />

Recall, for any reflection subgroup W ′ , we let χ(W ′ ) be the set of canonical<br />

generators of W ′ as a Coxeter system. Then, according to section 2.3, we have a<br />

corresponding set of roots, positive roots, and simple roots for the Coxeter system<br />

(W ′ , χ(W ′ )) given by<br />

ΦW ′ = {α ∈ Φ| sα ∈ W ′ }, Φ +<br />

W ′ := ΦW ′ ∩ Φ+ , ΠW ′ = {α ∈ Φ+ | sα ∈ χ(W ′ )}<br />

respectively.<br />

We now introduce several cones that appear inside of the vector space V in-<br />

troduced in section 2.2. For any W ′ a reflection subgroup of W , we define the<br />

fundamental chamber for W ′ on the vector space V by<br />

CW ′ = {v ∈ V | (v, α) ≥ 0 for all α ∈ ΠW ′}.<br />

Then we denote the Tits cone of W ′ by XW ′ where this is defined by<br />

�<br />

XW ′ =<br />

w∈W ′<br />

w(CW ′).<br />

In addition to these we define the following:<br />

�<br />

KW ′ := R≥0ΠW ′ ∩ −CW ′, ZW ′ :=<br />

w∈W ′<br />

w(KW ′)<br />

where ZW ′ is known as the imaginary cone of W ′ on V (we only define it if S is<br />

finite). The definition of the imaginary cone here is taken from [15], which models<br />

it loosely on one characterization of imaginary cones of Kac-Moody Lie algebras<br />

120


([24]). The results we use are analogs for general W of results proved by Kac for<br />

imaginary cones of Kac-Moody Lie algebras (for which, however, the proofs use<br />

imaginary roots, which have no counterpart here). We will refer to CW , XW , KW<br />

and ZW simply by C , X , K and Z .<br />

11.5 Topology on Rays in Cones<br />

By a ray through v in V we mean a set {λv| λ ∈ R≥0}. We let R denote the<br />

set of rays of V . We can give R a topology and a metric in the following way. Let<br />

B be compact convex body containing 0 in V , and let ∂B denote the boundary of<br />

B. Define the map ϕ : R → ∂B by ϕ(x) = x∩∂B. It is clear that ϕ is a bijection.<br />

We thus declare ϕ to be a homeomorphism between R and ∂B, and this gives R a<br />

topology and a metric corresponding to ∂B in V , with the topology independent<br />

of B. We also introduce the following subsets of R. Let R+ := {R≥0α| α ∈ Φ + }<br />

and let R0 := R+ \ R+. We say a ray, R≥0α, is positive if (α, α) > 0 and isotropic<br />

if (α, α) = 0. We recall some facts about these sets (both proven in [15]).<br />

Proposition 11.5.1. 1. R+ consists of positive rays and is discrete in the<br />

subspace topology.<br />

2. R0 consists of isotropic rays and is closed in R.<br />

3. R0 is the set of limit rays of R+.<br />

In addition, we have the following theorem.<br />

Theorem 11.5.2. The cone Z is the convex closure of �<br />

This theorem has the following immediate corollary.<br />

121<br />

r∈R0<br />

r ∪ {0}.


Corollary 11.5.3. If W ′ is a finitely generated reflection subgroup of W , then<br />

ZW ′ ⊆ ZW .<br />

Remark 11.5.4. A much stronger fact, proven in [15], which we will not use, is<br />

that in the situation of the corollary above we actually have ZW ′ ⊆ ZW .<br />

11.6 Cones in Hyperbolic Coxeter Systems<br />

Recall that we say (W, S) is a hyperbolic Coxeter system if the bilinear form<br />

(−, −) on V has inertia (n − 1, 1, 0) and every proper standard parabolic Coxeter<br />

system is of finite or affine type (see Chapter 10).<br />

Suppose we fix a hyperbolic Coxeter system (W, S). We know that V has<br />

a basis x0, ..., xn such that (xi, xj) = δi,j, (xi, xi) = 1 for all i ∈ {1, ..., n}, and<br />

(x0, x0) = −1. Then, according to [15], we have (after replacing x0 by −x0 if<br />

necessary) that<br />

Z = L<br />

where L = { � n<br />

i=0 λixi| λ0 ≥ �� n<br />

i=1 λ2 i }.<br />

Now, consider the setup as in section 11.5. Let RΠ ⊂ R be the set of rays<br />

spanned by non-zero elements of R≥0Π. Since (W, S) is hyperbolic, the bilinear<br />

form is non-degenerate, and thus the interior of C is non-empty (see [15]). Take<br />

ρ ∈ int(C ) so that (ρ, α) > 0 for all α ∈ Π. Therefore, we may define a map<br />

τ : R≥0Π \ {0} → H := {v ∈ V | (v, ρ) = 1}<br />

given by τ(v) = v/(ρ, v). Then the image of τ is the set P := {v ∈ V | (ρ, v) = 1},<br />

which is a convex polytope in the affine hyperplane H with points (ρ, α) −1 α for<br />

α ∈ Π as vertices. From τ, we introduce a map on R as follows. For r ∈ R, we<br />

122


let ˆτ(r) = τ(x) for all x ∈ r \ {0} (note ˆτ is independent of choice of x ∈ r \ {0}).<br />

With terminology from section 11.5 we may take B so that P ⊂ ∂B.<br />

Now, we can see that Z ∩ H is a disk. Also, by Theorem 11.5.2 we see that<br />

Z ∩ H = conv ��<br />

r∈R0 r� ∩ H.<br />

Lemma 11.6.1. R0 is the set of rays in the boundary of Z .<br />

Proof. First, a ray r ∈ R0 is isotropic by Proposition 11.5.1, and thus it is in<br />

±∂(L )) = ±∂(Z ) due to the description of L above. However, r must also be<br />

in Z , so r ∈ ∂(Z ).<br />

For the reverse implication, the statements above imply it is enough to show<br />

that the result is true inside of H, that is the convex hull of limit points of<br />

(images of) roots in H includes the entire boundary sphere of � Z = L � ∩ H.<br />

We let the boundary of Z ∩ H be denoted by S := ∂ˆτ(Z ), and we let A :=<br />

ˆτ � conv ��<br />

r∈R0 r�� . Then A is the convex hull of the points �<br />

ˆτ(r). But then<br />

the statement is clear since S is a sphere and is contained in A since Z ∩ H =<br />

conv ��<br />

r∈R0 r� ∩H. So every point in S must also be in A or else we would not have<br />

r∈R0<br />

that the boundary of A is a sphere (which is true since (W, S) is hyperbolic).<br />

11.7 Universal Coxeter Systems<br />

Suppose we have a hyperbolic Coxeter system (W, S) as in section 11.6. Then<br />

with the same setup there, we know that Z = L and intersecting with H we get<br />

that S = ∂ˆτ(Z ) is a sphere in H.<br />

Lemma 11.7.1. Suppose that x, y ∈ S. Then there exist αx, αy ∈ Φ + such that<br />

τ(αx), τ(αy) ∈ ˆτ (R+) are sufficiently close to x and y respectively; furthermore,<br />

αx and αy satisfy (αx, αy) < −1. In particular, if S ′ := {sαx, sαy} then (W ′ , S ′ ) is<br />

an indefinite infinite dihedral reflection subgroup of (W, S) and χ(W ′ ) = S ′ .<br />

123


Proof. Since x, y ∈ S then x and y are limit points of ˆτ(R+) and therefore we can<br />

find (images of) positive roots arbitrarily close to x and y. Thus, we pick positive<br />

roots αx ∈ Φ + and αy ∈ Φ + such that τ(αx) and τ(αy) are close enough to x<br />

and y respectively to ensure that {aτ(αx) + bτ(αy)| a, b ∈ R>0} ∩ (Z ∩ H) �= ∅.<br />

Again, this is possible since S is a sphere and x and y are two points (x �= y)<br />

on the sphere and so the relative interior of the line connecting x and y must be<br />

included in the relative interior of the disk bounded by the sphere which is Z ∩H.<br />

Now, we pick a point in the interior of the imaginary cone, v = aτ(αx) + bτ(αy) ∈<br />

int(Z ∩ H) ⊂ int(Z ) with a, b ∈ R>0. Then (v, v) < 0 necessarily and so (−, −)<br />

restricted to Span{αx, αy} cannot be of finite type or affine type. Thus, (W ′ , S ′ )<br />

cannot be a finite dihedral subgroup. Therefore, we see that |(αx, αy)| > 1. Now<br />

if (αx, αy) > 1 then by [4, Proposition 2.2] (and without loss of generality) we<br />

must have αx dominates αy. However, this is impossible since, by investigation<br />

of dominance in infinite dihedral groups ([14]), we know that αx can dominate αy<br />

only if [αx, αy] ∩ ZW ′ = ∅, which is not the case here. Thus, (αx, αy) < −1 and<br />

so S ′ = {sαx, sαy} is the canonical set of generators for the infinite dihedral group<br />

generated by S ′ , (W ′ , S ′ ).<br />

Now, by Lemma 11.3.1 recall that for arbitrary m we can find m points<br />

{x1, ..., xm} ⊂ S such that Pm := conv{x1, ..., xm} along with the property that<br />

dK(Pm, S) → 0 as m → ∞. Then, by Lemma 11.7.1 above, since m is finite, we<br />

can find αxi ∈ Φ+ with τ(αxi ) sufficiently close to xi such that S ′ i,j := {sαx i , sαx j }<br />

is the set of canonical generators for the (infinite) dihedral group it generates,<br />

(W ′<br />

i,j, S ′ i,j). Now, let S ′ := {sαx 1 , ..., sαxm } and W ′ be the reflection subgroup<br />

generated by S ′ .<br />

Proposition 11.7.2. Then (W ′ , S ′ ) is a universal Coxeter system with χ(W ′ ) =<br />

124


S ′ .<br />

Proof. Consider the Coxeter system (W ′ , S ′ ). Then by [19, Proposition 3.5], S ′ =<br />

χ(W ′ ) if and only if {s, s ′ } = χ(〈s, s ′ 〉) for all s, s ′ ∈ S ′ and this follows from<br />

Lemma 11.7.1. Finally, since m(s, s ′ ) = ∞ for all s �= s ′ ∈ S ′ then (W ′ , S ′ ) is a<br />

universal Coxeter system.<br />

11.8 Large Reflection Subgroups<br />

With the terminology as in section 11.6, we have H containing P . Thus, we<br />

can view ˆτ(Z ) as a subset of P . Additionally, we can view any subset of RΠ as a<br />

subset of P . Thus, we will abuse notation by defining dK on any compact subset<br />

of RΠ in the following way: for X, Y ⊂ RΠ, dK(X, Y ) = dK(ˆτ(X), ˆτ(Y )). We put<br />

the previous results together to obtain the following result.<br />

Proposition 11.8.1. Suppose (W, S) is hyperbolic. There exists a sequence of fi-<br />

nite rank, universal reflection subgroups, (W ′ m, S ′ m), of W with dK(R≥0ΠW ′ , Z ) →<br />

m<br />

0 as m → ∞ and dK(Z W ′ , Z ) → 0 as m → ∞.<br />

m<br />

Proof. Let ɛ > 0 be given. Then we can pick m large enough so that dK(Pm, S) <<br />

ɛ/2 by Lemma 11.3.1 where Pm = conv{x1, ..., xm} for some {x1, ..., xm} ⊂ S.<br />

Thus, dK(Pm, Z ∩ H) < ɛ/2. According to Lemma 11.7.2 and the discussion<br />

before it, we can choose {αx1, ..., αxm} ⊂ Φ + so that τ(αxi ) is sufficiently close to<br />

xi (in particular so that dK(τ(αxi ), xi) < ɛ/2) with the property that (W ′ m, S ′ m) is<br />

a universal Coxeter system where S ′ m := {sαx 1 , ..., sαxm } and W ′ m = 〈S ′ m〉. Now, let<br />

Qm = conv{τ(αx1), ..., τ(αxm)}. Then Qm = R≥0ΠW ′ m ∩ H and dK(Qm, Pm) < ɛ/2<br />

125


since dK(τ(αxi ), xi) < ɛ/2 (by assumption above). Therefore, we see that<br />

dK(R≥0ΠW ′, Z ) = dK(Qm, Z ∩ H)<br />

≤ dK(Qm, Pm) + dK(Pm, Z ∩ H)<br />

= dK(Qm, Pm) + dK(Pm, S) < ɛ/2 + ɛ/2 = ɛ<br />

as desired. It remains to show that we can take dK(ZW ′ , Z ) → 0 also.<br />

m<br />

We temporarily fix i �= j. Then since (αxi , αxj ) < −1, we have<br />

Z 〈sαx i ,sαx j 〉 = {z ∈ R≥0αxi<br />

= R≥0βxi,xj + R≥0βxj,xi<br />

+ R≥0αxj | (z, z) ≤ 0}<br />

(which is a two-dimensional cone) for some isotropic linearly independent vectors<br />

βxi,xj<br />

R≥0βxj,xi<br />

and βxj,xi . We choose our previous notation so that βxi,xj ∈ R≥0αxi +<br />

and βxj,xi ∈ R≥0βxi,xj + R≥0αxj .<br />

Then, from [14], since 〈sαx i , sαx j 〉 is dihedral, we have<br />

Z 〈sαx i ,sαx j 〉 = Z ∩ � R≥0αxi<br />

+ R≥0αxj<br />

Thus, τ(βxi,xj ) and τ(βxj,xi ) are the points given by the intersection<br />

This implies that we can choose αxi<br />

[τ(αxi ), τ(αxj )] ∩ S = {τ(βxi,xj ), τ(βxj,xi )}.<br />

� .<br />

and αxj so that all of the following hold:<br />

dK(τ(αxi ), xi) < ɛ/4 dK(τ(αxj ), xj) < ɛ/4<br />

dK(τ(βxi,xj ), xi) < ɛ/4 dK(τ(βxj,xi ), xj) < ɛ/4.<br />

126<br />

(11.8.1)


Next, by Corollary 11.5.3, we know that<br />

Z 〈sαx i ,sαx j 〉 ⊆ Z W ′ m<br />

⊆ Z<br />

for all pairs i �= j. Each of these are convex cones, and so we also have<br />

�<br />

i�=j<br />

� �<br />

R≥0 R≥0βxi,xj + R≥0βxj,xi ⊆ Z W ′ ⊆ Z . (11.8.2)<br />

m<br />

Now, let ɛ > 0. We choose m large enough and αx1, ..., αxm sufficiently close to<br />

x1, ..., xm respectively so that (11.8.1) holds and so that dK(R≥0ΠW ′ , Z ) < ɛ/2<br />

m<br />

where W ′ m = 〈sαx , ..., sαxm 〉 (which is possible by the first part of the proposition).<br />

1<br />

Then (11.8.1) implies that dK(αxi , βxi,xj ) < ɛ/2 for all j �= i so that<br />

and thus<br />

dK<br />

dK([τ(βxi,xj ), τ(βxj,xi )], [τ(αxi ), τ(αxj )]) < ɛ/2<br />

�<br />

� � �<br />

R≥0βxi,xj + R≥0βxj,xi ,<br />

i�=j<br />

m�<br />

k=1<br />

Together, since �m k=1 R≥0αxk = R≥0ΠW ′ , we get that<br />

m<br />

dK<br />

dK<br />

� �<br />

� � �<br />

R≥0βxi,xj + R≥0βxj,xi , Z ≤<br />

i�=j<br />

�<br />

� � �<br />

R≥0βxi,xj + R≥0βxj,xi ,<br />

i�=j<br />

m�<br />

k=1<br />

R≥0αxk<br />

�<br />

R≥0αxk<br />

�<br />

< ɛ<br />

2 .<br />

+ dK(R≥0ΠW ′ ɛ ɛ<br />

, Z ) < + m 2 2<br />

Then, equation (11.8.2) implies that dK(Z W ′ , Z ) < ɛ as well, finishing the proof.<br />

m<br />

The next theorem follows from the previous propositions and Chapter 10.<br />

127<br />

= ɛ.


Theorem 11.8.2. Let (W, S) be a finite rank, irreducible, infinite, non-affine<br />

Coxeter system. Then, for any m ∈ N, W contains a reflection subgroup W ′ with<br />

|χ(W ′ )| = m such that (W ′ , χ(W ′ )) is a universal Coxeter system.<br />

Proof. Let m ∈ N. By Theorem 10.8.2, (W, S) contains a hyperbolic parabolic<br />

subsystem (WJ, J). By Proposition 11.7.2, we have that (WJ, J) contains a uni-<br />

versal Coxeter system (W ′ , S ′ ) of rank m and thus (W, S) also contains (W ′ , S ′ )<br />

proving the theorem.<br />

128


CHAPTER 12<br />

EXPONENTIAL GROWTH <strong>IN</strong> <strong>COXETER</strong> <strong>GROUPS</strong><br />

For any finitely generated group, we can define the growth function of the group<br />

by determining the number of elements of the group of all lengths (with respect<br />

to the generating set). In [6], de la Harpe demonstrates that irreducible, infinite,<br />

non-affine Coxeter systems have what is known as exponential growth by showing<br />

that any such Coxeter system must contain a free non-abelian subgroup. Also, in<br />

[27], Margulis and Vinberg prove the stronger result that such a W must contain<br />

a finite index subgroup which surjects onto a non-abelian free group. Finite and<br />

affine Coxeter systems are known to have polynomial growth.<br />

Even more recently, in [29], Viswanath uses an alternate method to show that<br />

any irreducible, infinite, non-affine simply-laced Coxeter system has exponential<br />

growth. He also obtains the stronger result that for such a system (W, S) and<br />

any J � S the quotient W/WJ also has exponential growth. His method involves<br />

finding a particular universal rank 3 reflection subgroup inside of (W, S). In [29,<br />

Remark 3], the author notes that it would be interesting to know if this result<br />

holds in the non-simply laced case as well. We intend to answer this question<br />

using our results from the previous two chapters.<br />

129


12.1 Growth Types<br />

To begin with, we introduce the basic notions of growth types in finitely gen-<br />

erated groups. Let (W, S) be a finitely generated Coxeter system. We follow [7]<br />

or [29] for general terminology and results.<br />

Definition 12.1.1. Let a := (ak)k≥0 be a sequence of non-decreasing natural<br />

numbers. We define the exponential growth rate of the sequence to be ω(a) :=<br />

lim sup k→∞ a 1/k<br />

k .<br />

For each k ∈ N, we define a subset W≤k := {w ∈ W | l(w) ≤ k}. Then, for any<br />

subset F ⊆ W with 1 ∈ F , we define F≤k := F ∩ W≤k. Finally, for any k ∈ N and<br />

F ⊆ W , we define the number γF,k := |F≤k| and the sequence γF := (γF,k)k≥0.<br />

Then, let ω(F ) := ω(γF ) = lim sup k→∞ γ 1/k<br />

F,k .<br />

Then, we say that a subset F has exponential growth if ω(F ) > 1 and subex-<br />

ponential growth otherwise. If a subset F has subexponential growth, and there is<br />

C ∈ R>0 and d ∈ Z≥0 with γF,k ≤ Ck d for all k ≥ 0, then we say F has polynomial<br />

growth. If F is of subexponential growth and not of polynomial growth, F has<br />

intermediate growth.<br />

We note that if P (F, W ) = �<br />

w∈F ul(w) = �<br />

k≥0 aku k is the Poincaré series of<br />

F , then the coefficients ak = γF,k − γF,k−1 for k ≥ 1 and a0 = γF,0.<br />

12.2 Properties of W(3)<br />

The proof that quotients have exponential growth requires some properties of<br />

the universal Coxeter system W(3) = 〈a, b, c| a 2 = b 2 = c 2 = 1〉 with Coxeter<br />

diagram<br />

•�<br />

∞ ���∞<br />

• •<br />

�<br />

�<br />

�<br />

�<br />

∞<br />

130


These properties are all known and listed in [29], but we remind the reader here<br />

for completeness.<br />

W(3) has the following properties:<br />

1. The Poincaré series of W(3) is P (W(3)) = 1+q<br />

1−2q .<br />

2. Each w ∈ W(3) has a unique reduced expression.<br />

3. The conjugacy classes of a, b, and c are all distinct.<br />

4. For x �= y ∈ W(3)/(W(3)){a} =: W {a}<br />

(3) , we know that xax−1 �= yay −1 . Also,<br />

P (W {a}<br />

(3) , W(3)) = P (W (3))<br />

1+q<br />

1 = 1−2q so there are exactly 2k elements x ∈ W {a}<br />

(3)<br />

with l(x) = k. For each such x, we have l(xax −1 ) ≤ 2l(x) + 1 = 2k + 1.<br />

5. Property 4 implies that for k ≥ 0 there are more than 2 k elements t ∈<br />

{xax−1 | x ∈ W {a}<br />

(3) } with l(t) ≤ 2k + 1.<br />

6. For any reflection t ∈ {xax−1 | x ∈ W {a}<br />

(3) }, properties 2 and 3 imply that a<br />

appears in any reduced expression for t.<br />

12.3 Parabolic Quotients<br />

We now prove a generalization of [29, Theorem 1].<br />

Theorem 12.3.1. Let (W, S) be an irreducible, infinite, non-affine Coxeter sys-<br />

tem. Then for all J � S, W J = W/WJ has exponential growth.<br />

Our proof is similar to the one found in [29], however, we need to make some<br />

changes so that proof works for all Coxeter systems (as opposed to only simply-<br />

laced Coxeter systems).<br />

We first recall an important result due to Deodhar, [9].<br />

131


Theorem 12.3.2. Let (W, S) be a Coxeter system, T the set of reflections, and<br />

J ⊆ S. If t1, t2 ∈ T \ WJ with t1 �= t2, then t1WJ �= t2WJ, that is, distinct<br />

reflections in T \ WJ lie in distinct left cosets of WJ.<br />

According to [29], to prove Theorem 12.3.1, we need to prove the following<br />

proposition.<br />

Proposition 12.3.3. Let (W, S) be an irreducible, infinite, non-affine Coxeter<br />

system and let J � S. Then there exists M ∈ N such that for all k ≥ 0 there are<br />

at least 2 k elements t ∈ T \ WJ such that l(t) ≤ M(2k + 1).<br />

[29].<br />

If this proposition holds, Theorem 12.3.1 holds by the same argument given in<br />

12.4 Proof of Proposition 12.3.3<br />

Proof. For all x ∈ W , any reduced expression for x uses the same simple reflec-<br />

tions. We write supp(x) = {r ∈ S| r appears in all reduced expression for x}. It<br />

is also well known that x ∈ WJ if and only if supp(x) ⊆ J.<br />

Now, let (W, S) be an irreducible, infinite, non-affine Coxeter system and let<br />

s ′ �∈ J. According to Theorem 11.8.2, (W, S) contains a reflection subgroup W ′<br />

such that W ′ ∼ = W(3). Suppose χ(W ′ ) = {t1, t2, t3}. Without loss of generality,<br />

we may assume that s ′ ∈ supp(t1). Indeed, if s ′ �∈ supp(ti) for all i, then let<br />

Z ⊂ S be the smallest subset such that W ′ ⊂ WZ. Then, take a minimal path<br />

s ′ = sn − sn−1 − · · · − s1 − z in the Coxeter diagram with si �∈ Z for all i and<br />

z ∈ Z; let the word corresponding to the path be x = s1 · · · sn ∈ W , and by<br />

reordering we assume z ∈ supp(t1). Now, we define a sequence of reflections in<br />

the following way. Let r0 = t1 and ri = si · · · s1t1s1 · · · si for all i ≥ 1. Let αri<br />

132<br />

∈ Φ+


e the root corresponding to ri. Now, since ({s1, ..., si−1} ∪ Z) ∩ ({si}) = ∅ for<br />

all i ≥ 1 and si is connected to si−1, we have (αri−1 , αsi ) < 0 for all i and so<br />

l(ri) = l(ri−1) + 2 for all i ≥ 1. This implies that l(x −1 t1x) = 2l(x) + l(t1);<br />

hence, s ′ = sn ∈ supp(x −1 tx). Then, xW ′ x −1 is a Coxeter system, isomorphic to<br />

W(3) with χ(xW ′ x −1 ) = {xtx −1 , xt2x −1 , xt3x −1 } = {t ′ 1, t ′ 2, t ′ 3} with s ′ ∈ supp(t ′ 1)<br />

as required.<br />

Next, by the properties listed in Section 12.2, we know that there are at least<br />

2 k elements t ∈ W ′ ∩ T such that l(W ′ ,χ(W ′ ))(t) ≤ 2k + 1 such that t1 ∈ supp(t).<br />

Define M := max{l(ti), l(t2), l(t3)}. Then, we have l(w) ≤ M(l(W ′ ,χ(W ′ ))(w)) for<br />

all w ∈ W ′ . This implies that for each t ∈ W ′ ∩T above, we have l(t) ≤ M(2k+1).<br />

For each t ∈ W ′ ∩ T described above, t1 appears in a reduced expression for t;<br />

in terms of roots this means αt = �3 i=1 ciαti with c1 > 0. Since s ′ ∈ supp(t1), we<br />

have αt1 = �<br />

s∈S csαs with cs ′ > 0. Therefore, it follows that αt = �<br />

s∈S dsαs with<br />

ds ′ > 0 so that s′ ∈ supp(t). Thus, we have t �∈ WJ. This implies that t ∈ T \ WJ.<br />

Therefore, by Theorem 12.3.2, we get that there are at least 2 k elements t ∈ T \WJ<br />

such that l(t) ≤ M(2k + 1). The proposition follows.<br />

133


REFERENCES<br />

1. Alexander Barvinok. A course in convexity, volume 54 of Graduate Studies<br />

in Mathematics. American Mathematical Society, Providence, RI, 2002.<br />

2. Anders Björner and Francesco Brenti. Combinatorics of Coxeter groups, volume<br />

231 of Graduate Texts in Mathematics. Springer, New York, 2005.<br />

3. Nicolas Bourbaki. Lie groups and Lie algebras. Chapters 4–6. Elements of<br />

Mathematics (Berlin). Springer-Verlag, Berlin, 2002. Translated from the 1968<br />

French original by Andrew Pressley.<br />

4. Brigitte Brink. The set of dominance-minimal roots. J. Algebra, 206(2):371–<br />

412, 1998.<br />

5. Brigitte Brink and Robert B. Howlett. A finiteness property and an automatic<br />

structure for Coxeter groups. Math. Ann., 296(1):179–190, 1993.<br />

6. Pierre de la Harpe. Groupes de Coxeter infinis non affines. Exposition. Math.,<br />

5(1):91–96, 1987.<br />

7. Pierre de la Harpe. Topics in geometric group theory. Chicago Lectures in<br />

Mathematics. University of Chicago Press, Chicago, IL, 2000.<br />

8. Ronald de Man. The generating function for the number of roots of a Coxeter<br />

group. J. Symbolic Comput., 27(6):535–541, 1999.<br />

9. Vinay V. Deodhar. On the root system of a Coxeter group. Comm. Algebra,<br />

10(6):611–630, 1982.<br />

10. Vinay V. Deodhar. A note on subgroups generated by reflections in Coxeter<br />

groups. Arch. Math. (Basel), 53(6):543–546, 1989.<br />

11. M. J. Dyer. Hecke algebras and shellings of Bruhat intervals. II. Twisted<br />

Bruhat orders. In Kazhdan-Lusztig theory and related topics (Chicago, IL,<br />

1989), volume 139 of Contemp. Math., pages 141–165. Amer. Math. Soc.,<br />

Providence, RI, 1992.<br />

12. M. J. Dyer. Hecke algebras and shellings of Bruhat intervals. Compositio<br />

Math., 89(1):91–115, 1993.<br />

134


13. M. J. Dyer. Quotients of twisted Bruhat orders. J. Algebra, 163(3):861–879,<br />

1994.<br />

14. Matthew Dyer. Imaginary cone and dominance order in Coxeter groups. in<br />

preparation.<br />

15. Matthew Dyer. Imaginary cone and reflection subgroups of Coxeter groups.<br />

in preparation.<br />

16. Matthew Dyer. On rigidity of abstract root systems of Coxeter groups.<br />

preprint.<br />

17. Matthew Dyer. Questions on initial sections of reflection orders. in preparation.<br />

18. Matthew Dyer. Regular subsets of Coxeter groups. in preparation.<br />

19. Matthew Dyer. Reflection subgroups of Coxeter systems. J. Algebra,<br />

135(1):57–73, 1990.<br />

20. Tom Edgar. Sets of reflections defining twisted Bruhat orders. J. Algebraic<br />

Combin., 26(3):357–362, 2007.<br />

21. David B. A. Epstein, James W. Cannon, Derek F. Holt, Silvio V. F. Levy,<br />

Michael S. Paterson, and William P. Thurston. Word processing in groups.<br />

Jones and Bartlett Publishers, Boston, MA, 1992.<br />

22. Leslie Hogben, editor. Handbook of linear algebra. Discrete Mathematics and<br />

its Applications (Boca Raton). Chapman & Hall/CRC, Boca Raton, FL, 2007.<br />

Associate editors: Richard Brualdi, Anne Greenbaum and Roy Mathias.<br />

23. James E. Humphreys. Reflection groups and Coxeter groups, volume 29 of<br />

Cambridge Studies in Advanced Mathematics. Cambridge University Press,<br />

Cambridge, 1990.<br />

24. Victor G. Kac. Infinite-dimensional Lie algebras. Cambridge University Press,<br />

Cambridge, third edition, 1990.<br />

25. K. Kuratowski. Topology. Vol. I. New edition, revised and augmented. Translated<br />

from the French by J. Jaworowski. Academic Press, New York, 1966.<br />

26. G. Lusztig. Hecke algebras with unequal parameters, volume 18 of CRM Monograph<br />

Series. American Mathematical Society, Providence, RI, 2003.<br />

27. G. A. Margulis and È. B. Vinberg. Some linear groups virtually having a free<br />

quotient. J. Lie Theory, 10(1):171–180, 2000.<br />

135


28. Richard P. Stanley. Enumerative combinatorics. Vol. 1, volume 49 of Cambridge<br />

Studies in Advanced Mathematics. Cambridge University Press, Cambridge,<br />

1997. With a foreword by Gian-Carlo Rota, Corrected reprint of the<br />

1986 original.<br />

29. Sankaran Viswanath. On growth types of quotients of Coxeter groups by<br />

parabolic subgroups. Comm. Algebra, 36(2):796–805, 2008.<br />

136

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!