Combinatorics: The Art of Counting 147046280X, 9781470462802

This book is a gentle introduction to the enumerative part of combinatorics suitable for study at the advanced undergrad

192 79 2MB

English Pages 304 [328] Year 2020

Report DMCA / Copyright

DOWNLOAD PDF FILE

Table of contents :
Cover
Title page
Copyright
Contents
Preface
List of Notation
Chapter 1. Basic Counting
1.1. The Sum and Product Rules for sets
1.2. Permutations and words
1.3. Combinations and subsets
1.4. Set partitions
1.5. Permutations by cycle structure
1.6. Integer partitions
1.7. Compositions
1.8. The twelvefold way
1.9. Graphs and digraphs
1.10. Trees
1.11. Lattice paths
1.12. Pattern avoidance
Exercises
Chapter 2. Counting with Signs
2.1. The Principle of Inclusion and Exclusion
2.2. Sign-reversing involutions
2.3. The Garsiaโ€“Milne Involution Principle
2.4. The Reflection Principle
2.5. The Lindstrรถmโ€“Gesselโ€“Viennot Lemma
2.6. The Matrix-Tree Theorem
Exercises
Chapter 3. Counting with Ordinary Generating Functions
3.1. Generating polynomials
3.2. Statistics and ๐‘ž-analogues
3.3. The algebra of formal power series
3.4. The Sum and Product Rules for ogfs
3.5. Revisiting integer partitions
3.6. Recurrence relations and generating functions
3.7. Rational generating functions and linear recursions
3.8. Chromatic polynomials
3.9. Combinatorial reciprocity
Exercises
Chapter 4. Counting with Exponential Generating Functions
4.1. First examples
4.2. Generating functions for Eulerian polynomials
4.3. Labeled structures
4.4. The Sum and Product Rules for egfs
4.5. The Exponential Formula
Exercises
Chapter 5. Counting with Partially Ordered Sets
5.1. Basic properties of partially ordered sets
5.2. Chains, antichains, and operations on posets
5.3. Lattices
5.4. The Mรถbius function of a poset
5.5. The Mรถbius Inversion Theorem
5.6. Characteristic polynomials
5.7. Quotients of posets
5.8. Computing the Mรถbius function
5.9. Binomial posets
Exercises
Chapter 6. Counting with Group Actions
6.1. Groups acting on sets
6.2. Burnsideโ€™s Lemma
6.3. The cycle index
6.4. Redfieldโ€“Pรณlya theory
6.5. An application to proving congruences
6.6. The cyclic sieving phenomenon
Exercises
Chapter 7. Counting with Symmetric Functions
7.1. The algebra of symmetric functions, Sym
7.2. The Schur basis of Sym
7.3. Hooklengths
7.4. ๐‘ƒ-partitions
7.5. The Robinsonโ€“Schenstedโ€“Knuth correspondence
7.6. Longest increasing and decreasing subsequences
7.7. Differential posets
7.8. The chromatic symmetric function
7.9. Cyclic sieving redux
Exercises
Chapter 8. Counting with Quasisymmetric Functions
8.1. The algebra of quasisymmetric functions, QSym
8.2. Reverse ๐‘ƒ-partitions
8.3. Chain enumeration in posets
8.4. Pattern avoidance and quasisymmetric functions
8.5. The chromatic quasisymmetric function
Exercises
Appendix. Introduction to Representation Theory
A.1. Basic notions
Exercises
Bibliography
Index
Back Cover
Recommend Papers

Combinatorics: The Art of Counting
 147046280X, 9781470462802

  • Commentary
  • decrypted from 1590975CF33E19DA768D956D8DE9D996 source file
  • 0 0 0
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up
File loading please wait...
Citation preview

GRADUATE STUDIES I N M AT H E M AT I C S

210

Combinatorics: The Art of Counting Bruce E. Sagan

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

10.1090/gsm/210

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Combinatorics: The Art of Counting

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

GRADUATE STUDIES I N M AT H E M AT I C S

210

Combinatorics: The Art of Counting Bruce E. Sagan

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Marco Gualtieri Bjorn Poonen Gigliola Sta๏ฌƒlani (Chair) Je๏ฌ€ A. Viaclovsky 2020 Mathematics Subject Classi๏ฌcation. Primary 05-01; Secondary 06-01.

For additional information and updates on this book, visit www.ams.org/bookpages/gsm-210

Library of Congress Cataloging-in-Publication Data Names: Sagan, Bruce Eli, 1954- author. Title: Combinatorics : the art of counting / Bruce E. Sagan. Description: Providence, Rhode Island : American Mathematical Society, [2020] | Series: Graduate studies in mathematics, 1065-7339 ; 210 | Includes bibliographical references and index. Identi๏ฌers: LCCN 2020025345 | ISBN 9781470460327 (paperback) | ISBN 9781470462802 (ebook) Subjects: LCSH: Combinatorial analysisโ€“Textbooks. | AMS: Combinatorics โ€“ Instructional exposition (textbooks, tutorial papers, etc.). | Order, lattices, ordered algebraic structures โ€“ Instructional exposition (textbooks, tutorial papers, etc.). Classi๏ฌcation: LCC QA164 .S24 2020 | DDC 511/.6โ€“dc23 LC record available at https://lccn.loc.gov/2020025345

Copying and reprinting. Individual readers of this publication, and nonpro๏ฌt libraries acting for them, are permitted to make fair use of the material, such as to copy select pages for use in teaching or research. Permission is granted to quote brief passages from this publication in reviews, provided the customary acknowledgment of the source is given. Republication, systematic copying, or multiple reproduction of any material in this publication is permitted only under license from the American Mathematical Society. Requests for permission to reuse portions of AMS publication content are handled by the Copyright Clearance Center. For more information, please visit www.ams.org/publications/pubpermissions. Send requests for translation rights and licensed reprints to [email protected]. c 2020 by the American Mathematical Society. All rights reserved.  The American Mathematical Society retains all rights except those granted to the United States Government. Printed in the United States of America. โˆž The paper used in this book is acid-free and falls within the guidelines 

established to ensure permanence and durability. Visit the AMS home page at https://www.ams.org/ 10 9 8 7 6 5 4 3 2 1

25 24 23 22 21 20

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

To Sally, for her love and support

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Contents

Preface

xi

List of Notation Chapter 1.

Basic Counting

xiii 1

ยง1.1. The Sum and Product Rules for sets

1

ยง1.2. Permutations and words

4

ยง1.3. Combinations and subsets

5

ยง1.4. Set partitions

10

ยง1.5. Permutations by cycle structure

11

ยง1.6. Integer partitions

13

ยง1.7. Compositions

16

ยง1.8. The twelvefold way

17

ยง1.9. Graphs and digraphs

18

ยง1.10. Trees

22

ยง1.11. Lattice paths

25

ยง1.12. Pattern avoidance

28

Exercises

33

Chapter 2.

Counting with Signs

41

ยง2.1. The Principle of Inclusion and Exclusion

41

ยง2.2. Sign-reversing involutions

44

ยง2.3. The Garsiaโ€“Milne Involution Principle

49

ยง2.4. The Reflection Principle

52

vii Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

viii

Contents

ยง2.5. The Lindstrรถmโ€“Gesselโ€“Viennot Lemma

55

ยง2.6. The Matrix-Tree Theorem

59

Exercises

64

Chapter 3.

Counting with Ordinary Generating Functions

71

ยง3.1. Generating polynomials

71

ยง3.2. Statistics and ๐‘ž-analogues

74

ยง3.3. The algebra of formal power series

81

ยง3.4. The Sum and Product Rules for ogfs

86

ยง3.5. Revisiting integer partitions

89

ยง3.6. Recurrence relations and generating functions

92

ยง3.7. Rational generating functions and linear recursions

96

ยง3.8. Chromatic polynomials

99

ยง3.9. Combinatorial reciprocity

106

Exercises

109

Chapter 4.

Counting with Exponential Generating Functions

117

ยง4.1. First examples

117

ยง4.2. Generating functions for Eulerian polynomials

121

ยง4.3. Labeled structures

124

ยง4.4. The Sum and Product Rules for egfs

128

ยง4.5. The Exponential Formula

131

Exercises

134

Chapter 5.

Counting with Partially Ordered Sets

139

ยง5.1. Basic properties of partially ordered sets

139

ยง5.2. Chains, antichains, and operations on posets

145

ยง5.3. Lattices

148

ยง5.4. The Mรถbius function of a poset

154

ยง5.5. The Mรถbius Inversion Theorem

157

ยง5.6. Characteristic polynomials

164

ยง5.7. Quotients of posets

168

ยง5.8. Computing the Mรถbius function

174

ยง5.9. Binomial posets

178

Exercises

183

Chapter 6.

Counting with Group Actions

189

ยง6.1. Groups acting on sets

189

ยง6.2. Burnsideโ€™s Lemma

192

ยง6.3. The cycle index

197

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Contents

ix

ยง6.4. Redfieldโ€“Pรณlya theory

200

ยง6.5. An application to proving congruences

205

ยง6.6. The cyclic sieving phenomenon

209

Exercises

213

Chapter 7.

Counting with Symmetric Functions

219

ยง7.1. The algebra of symmetric functions, Sym

219

ยง7.2. The Schur basis of Sym

224

ยง7.3. Hooklengths

230

ยง7.4.

235

๐‘ƒ-partitions

ยง7.5. The Robinsonโ€“Schenstedโ€“Knuth correspondence

240

ยง7.6. Longest increasing and decreasing subsequences

244

ยง7.7. Differential posets

248

ยง7.8. The chromatic symmetric function

253

ยง7.9. Cyclic sieving redux

256

Exercises

259

Chapter 8.

Counting with Quasisymmetric Functions

267

ยง8.1. The algebra of quasisymmetric functions, QSym

267

ยง8.2. Reverse ๐‘ƒ-partitions

270

ยง8.3. Chain enumeration in posets

274

ยง8.4. Pattern avoidance and quasisymmetric functions

276

ยง8.5. The chromatic quasisymmetric function

279

Exercises

283

Appendix. Introduction to Representation Theory

287

ยงA.1. Basic notions

287

Exercises

292

Bibliography

293

Index

297

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Preface

Enumerative combinatorics has seen an explosive growth over the last 50 years. The purpose of this text is to give a gentle introduction to this exciting area of research. So, rather than trying to cover many different topics, I have chosen to give a more leisurely treatment of some of the highlights of the field. My goal has been to write the exposition so it could be read by a student at the advanced undergraduate or beginning graduate level, either as part of a course or for independent study. The reader will find it similar in tone to my book on the symmetric group. I have tried to keep the prerequisites to a minimum, assuming only basic courses in linear and abstract algebra as background. Certain recurring themes are emphasized, for example, the existence of sum and product rules first for sets, then for ordinary generating functions, and finally in the case of exponential generating functions. I have also included some recent material from the research literature which, to my knowledge, has not appeared in book form previously, such as the theory of quotient posets and the connection between pattern avoidance and quasisymmetric functions. Most of the exercises should be doable with a reasonable amount of effort. A few unsolved conjectures have been included among the problems in the hope that an interested student might wish to tackle one of them. They are, of course, marked as such. A few words about the title are in order. It is in part meant to be a tip of the hat to Donald Knuthโ€™s influential series of books The art of computer programing, Volumes 1โ€“3 [51โ€“53], which, among many other things, helped give birth to the study of pattern avoidance through its connection with stack sorting; see Exercise 36 in Chapter 1. I hope that the title also conveys some of the beauty found in this area of mathematics, for example, the elegance of the Hook Formula (equation (7.10)) for the number of standard Young tableaux. In addition I should mention that, due to my own preferences, this book concentrates on the enumerative side of combinatorics and mostly ignores the important extremal and existential parts of the field. The reader interested in these areas can consult the books of Flajolet and Sedgewick [25] and of van Lint [95].

xi Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

xii

Preface

This book grew out of the lecture notes which I have compiled over years of teaching the graduate combinatorics course at Michigan State University. I would like to thank the students in these classes for all the feedback they have given me about the various topics and their presentation. I am also indebted to the following colleagues, some of whom taught from a preliminary version of this book, who provided me with suggestions as well as catching numerous typographical errors: Matthias Beck, Moussa Benoumhani, Andreas Blass, Seth Chaiken, Sylvie Corteel, Georges Grekos, Richard Hensh, Nadia Lafreniรจre, Duncan Levear, and Tom Zaslavsky. Darij Grinberg deserves special mention for providing copious comments and corrections as well as providing a number of interesting exercises. I also received valuable feedback from four anonymous referees. Finally, I wish to express my appreciation of Ina Mette, my editor at the American Mathematical Society. Without her gentle support and persistence, this text would never have seen the light of day. Because I typeset this document myself, all errors can be blamed on my computer. East Lansing, Michigan, 2020

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

List of Notation

Symbol

Definition

๐ด(๐ท) ๐ด(๐บ) ๐’œ(๐บ) ๐‘Ž(๐บ) ๐ด([๐‘›], ๐‘˜) ๐ด(๐‘›, ๐‘˜) ๐ด๐‘› (๐‘ž) ๐’œ(๐‘ƒ) Asc ๐‘ asc ๐‘ Asc ๐œ‹ asc ๐œ‹ Av๐‘› (๐œ‹) ๐›ผ๐‘Ÿ ๐›ผฬ„ ๐›ผ(๐ถ) ๐ต(๐บ) ๐ต(๐‘‡) ๐ต๐‘› ๐ตโˆž ๐ต(๐‘›)

arc set of digraph ๐ท adjacency matrix of graph ๐บ set of acyclic orientations of ๐บ number of acyclic orientations of ๐บ set of permutations ๐œ‹ in ๐”–๐‘› having ๐‘˜ descents Eulerian number, cardinality of ๐ด([๐‘›], ๐‘˜) Eulerian polynomial atom set of poset ๐‘ƒ ascent set of a proper coloring ๐‘ ascent number of a proper coloring ๐‘ ascent set of permutation ๐œ‹ ascent number of permutation ๐œ‹ the set of permutations in ๐”–๐‘› avoiding ๐œ‹ reversal of composition ๐›ผ expansion of composition ๐›ผ rank composition of chain ๐ถ incidence matrix of graph ๐บ set of partitions of the set ๐‘‡ Boolean algebra on [๐‘›] poset of subsets of โ„™ ๐‘›th Bell number

Page 21 60 103 103 121 121 122 169 279 279 76 76 29 32 274 275 61 10 140 178 10

xiii Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

xiv

List of Notation

Symbol

Definition

โ„‚ ๐‘ ๐‘– (๐‘”) ๐ถ๐ฟ๐‘› co ๐‘‡ ๐ถ๐‘› ๐ถ๐‘› ๐‘๐‘ฅ (๐‘ƒ) ๐ถโˆž ๐ถ(๐‘›) ๐‘([๐‘›], ๐‘˜) ๐‘(๐‘›, ๐‘˜) ๐‘๐‘œ (๐ฟ, ๐‘˜) โ„‚๐‘‹ โ„‚[๐‘ฅ] โ„‚[[๐‘ฅ]] ๐’ž(๐œ‹) ๐’ž๐‘š (๐œ‹) Des ๐‘ƒ Des ๐œ‹ des ๐œ‹ ๐ท๐‘› ๐ทโˆž ๐ท(๐‘›) ๐’Ÿ(๐‘›) ๐’Ÿ(๐‘‰) ๐’Ÿ(๐‘‰, ๐‘˜) deg ๐‘š deg ๐‘ฃ ฮ”๐‘“(๐‘›) ๐›ฟ๐‘ฅ,๐‘ฆ ๐›ฟ(๐‘ฅ, ๐‘ง) ๐ธ(๐บ) ๐ธ(๐ฟ) ๐ธ(๐ฟ) ๐ธ๐‘› ๐‘’๐‘› ๐ธ(๐‘ก) Exc ๐œ‹ exc ๐œ‹

complex numbers number of cycles of length ๐‘– in group element ๐‘” claw poset with ๐‘› atoms content of tableau ๐‘‡ cycle with ๐‘› vertices chain poset of length ๐‘› column insertion of element ๐‘ฅ into tableau ๐‘ƒ chain poset on โ„• Catalan number set of permutations in ๐”–๐‘› with ๐‘˜ cycles signless Stirling number of the first kind ordered ๐‘˜ cycle decompositions of permutations of ๐ฟ vector space generated by set ๐‘‹ over โ„‚ polynomial algebra in ๐‘ฅ over โ„‚ formal power series algebra in ๐‘ฅ over โ„‚ set of functions compatible with ๐œ‹ set of functions compatible with ๐œ‹ bounded by ๐‘š descent set of tableau ๐‘ƒ descent set of permutation ๐œ‹ descent number of permutation ๐œ‹ lattice of divisors of ๐‘› divisibility poset on โ„™ derangement number set of Dyck paths of semilength ๐‘› set of all digraphs on vertex set ๐‘‰ set of all digraphs on vertex set ๐‘‰ with ๐‘˜ edges degree of a monomial degree of vertex ๐‘ฃ in a graph forward difference operator of ๐‘“(๐‘›) Kronecker delta delta function of poset incidence algebra edge set of graph ๐บ set structure on label set ๐ฟ nonempty set structure on label set ๐ฟ Euler number ๐‘›th elementary symmetric function generating function for elementary symmetric functions set of excedances of permutation ๐œ‹ number of excedances of permutation ๐œ‹

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Page 1 197 169 225 19 139 245 178 26 12 12 127 248 71 81 236 236 271 75 76 140 181 43 26 21 21 219 20 162 7 159 18 125 125 120 221 221 122 122

List of Notation

Symbol

Definition

Fix ๐‘“ ๐‘“๐‘› ๐น๐‘› ๐”ฝ๐‘ž ๐‘“(๐‘ฅ) ๐‘“๐‘† (๐‘ฅ) ๐น(๐‘›) ๐น(๐‘ฅ) ๐น๐’ฎ (๐‘ฅ) ๐น๐‘† ๐น๐›ผ ๐‘“๐œ† ฮฆ ๐œ™ ๐บโงต๐‘’ ๐บ/๐‘’ GL(๐‘‰) ๐’ข(๐‘‰) ๐’ข(๐‘‰, ๐‘˜) ๐บ๐‘ฅ ๐ป๐‘ = ๐ป๐‘–,๐‘— โ„Ž๐‘ = โ„Ž๐‘–,๐‘— โ„‹๐‘› โ„Ž๐‘› ๐ป(๐‘ก) ideg ๐‘ฃ Inv ๐œ‹ inv ๐œ‹ โ„(๐‘ƒ) ๐ผ(๐‘†) ISF(๐บ; ๐‘ก) ISF๐‘š (๐บ) isf๐‘š (๐บ) ๐‘–๐œ† (๐บ) ๐’ฅ(๐‘ƒ) ๐พ๐‘› ๐พ๐‘› ๐พ๐œ†,๐œ‡

fix point set of a function ๐‘“ Fibonacci number Fibonacci number Galois field with ๐‘ž elements ordinary generating function weight-generating function for weighted set ๐‘† binomial poset ๐‘›-interval factorial function exponential generating function exponential generating function for structure ๐’ฎ fundamental quasisymmetric for set ๐‘† fundamental quasisymmetric for composition ๐›ผ number of standard Young tableaux of shape ๐œ† fundamental map on permutations bijection between subsets and compositions graph ๐บ with edge ๐‘’ deleted graph ๐บ with edge ๐‘’ contracted general linear group over vector space ๐‘‰ set of all graphs on vertex set ๐‘‰ set of all graphs on vertex set ๐‘‰ with ๐‘˜ edges stabilizer of element ๐‘ฅ under the action of group ๐บ hook of cell ๐‘ = (๐‘–, ๐‘—) hooklength of cell ๐‘ = (๐‘–, ๐‘—) set of hook diagrams with ๐‘› cells ๐‘›th complete homogeneous symmetric function complete homogeneous generating function in-degree of vertex ๐‘ฃ in a digraph inversion set of permutation ๐œ‹ inversion number of permutation ๐œ‹ incidence algebra of poset ๐‘ƒ lower-order ideal generated by ๐‘† in a poset increasing spanning forest generating function of ๐บ set of ๐‘š-edge increasing spanning forests of ๐บ number of ๐‘š-edge increasing spanning forests of ๐บ number of independent type ๐œ† partitions in graph ๐บ distributive lattice of lower-order ideals of poset ๐‘ƒ complete graph with ๐‘› vertices lattice of compositions of ๐‘› number of tableaux of shape ๐œ† and content ๐œ‡

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

xv

Page 44 3 2 79 81 86 178 117 125 269 269 225 122 16 100 101 287 20 20 191 230 230 278 221 221 21 74 74 158 143 105 105 105 254 151 19 140 225

xvi

List of Notation

Symbol

Definition

๐ฟ(๐บ) โ„’(๐บ) โ„’(๐‘ƒ) โ„“(๐ถ) โ„“(๐œ†) โ„“(๐œ‹) lim ๐‘“๐‘˜ (๐‘ฅ)

Laplacian of graph ๐บ bond lattice of graph ๐บ set of linear extensions of ๐‘ƒ length of chain ๐ถ in a poset length of an integer partition ๐œ† length of a permutation ๐œ‹ limit of a sequence of formal power series

62 167 238 147 15 4 84

lds ๐œ‹ lis ๐œ‹ ๐ฟ๐‘› (๐‘ž) ๐ฟโˆž (๐‘ž) ๐ฟ(๐‘‰) ๐œ†(๐น) ๐œ†! maj ๐œ‹ ๐‘€(๐‘›) ๐‘€(๐‘ƒ) ๐‘€๐›ผ ๐‘š๐œ† ๐œ‡(๐‘ƒ) ๐œ‡(๐‘ฅ) ๐œ‡(๐‘ฅ, ๐‘ง) โ„• NBC๐‘˜ (๐บ) nbc๐‘˜ (๐บ) ๐’ฉโ„ฐ(๐‘š, ๐‘›) odeg ๐‘ฃ ๐’ช๐‘ฅ ๐‘‚(๐‘”) ๐‘œ(๐‘”) โ„™ ๐‘ƒโˆ— ๐’ซ๐ถ(๐บ) ๐‘ƒ(๐บ; ๐‘ก) Par ๐‘ƒ Par๐‘š ๐‘ƒ ๐‘ƒ๐‘› ๐‘ƒ(๐‘›) ๐‘(๐‘›) ๐‘๐‘› ๐‘ƒ(๐‘ก)

length of a longest decreasing subsequence of ๐œ‹ length of a longest increasing subsequence of ๐œ‹ lattice of subspaces of ๐”ฝ๐‘›๐‘ž poset of subspaces of vector space ๐‘‰โˆž over ๐”ฝ๐‘ž lattice of subspaces of ๐‘‰ type of partition induced by edge set ๐น multiplicity factorial of partition ๐œ† major index of permutation ๐œ‹ Mertens function monomial quasisymmeric function for poset ๐‘ƒ monomial quasisymmetric function monomial symmetric function Mรถbius function value on a poset ๐‘ƒ one-variable Mรถbius function evaluated at ๐‘ฅ two-variable Mรถbius function on the interval [๐‘ฅ, ๐‘ง] nonnegative integers set of no broken circuit sets of ๐‘˜ edges of ๐บ number of no broken circuit sets of ๐‘˜ edges of ๐บ set of ๐‘-๐ธ lattice paths from (0, 0) to (๐‘š, ๐‘›) out-degree of vertex ๐‘ฃ in a digraph orbit of an element ๐‘ฅ under action of a group big oh notation applied to function ๐‘” order of a group element ๐‘” positive integers dual of poset ๐‘ƒ set of proper colorings of ๐บ with the positive integers chromatic polynomial of graph ๐บ set of ๐‘ƒ-partitions set of ๐‘ƒ-partitions bounded by ๐‘š path with ๐‘› vertices set of partitions of the integer ๐‘› number of partitions of the integer ๐‘› ๐‘›th power sum symmetric function power sum symmetric generating function

245 244 140 178 140 255 254 76 183 275 268 220 154 154 157 1 102 102 26 21 190 182 210 1 142 279 100 238 238 19 13 13 221 221

๐‘˜โ†’โˆž

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Page

List of Notation

xvii

Symbol

Definition

๐‘ƒ(๐‘›, ๐‘˜) ๐‘(๐‘›, ๐‘˜) ๐‘ƒ(๐‘†) ๐‘ƒ(๐‘†, ๐‘˜) ๐‘ƒ((๐‘†, ๐‘˜)) ๐‘ƒ(๐œ‹) ๐’ซ(๐‘ข; ๐‘ฃ) ฮ ๐‘› ฮ (๐’ฎ) ฮ ๐‘’ (๐’ฎ) ฮ ๐‘œ (๐’ฎ) โ„š ๐‘„(๐‘›) ๐‘ž(๐‘›) ๐‘„(๐‘›, ๐‘˜) ๐‘ž(๐‘›, ๐‘˜) QSym QSym๐‘› ๐‘„(๐œ‹) ๐‘„๐‘› (ฮ ) โ„ โ„›๐ถ(๐œ‹) rk ๐‘ƒ Rk๐‘˜ ๐‘ƒ rk ๐‘ฅ โ„›(๐‘˜, ๐‘™) RPar ๐‘ƒ โ„›(๐‘ƒ) rpp๐‘› (๐œ†) rpar(๐‘ƒ; ๐ฑ) ๐‘Ÿ๐‘ฅ (๐‘ƒ) ๐œŒ(๐น) ๐œŒ โˆถ ๐บ โ†’ GL(๐‘‰) ๐’ฎ(๐ฟ) ๐”– ๐”–๐‘› ๐‘†๐‘“(๐‘›) sgn sh ๐‘‡ ๐‘ (๐‘›, ๐‘˜)

set of partitions of ๐‘› into at most ๐‘˜ parts number of partitions of ๐‘› into at most ๐‘˜ parts permutations of a set ๐‘† permutations of length ๐‘˜ of a set ๐‘† words of length ๐‘˜ over a set ๐‘† insertion tableau of ๐œ‹ set of directed paths from ๐‘ข to ๐‘ฃ in a digraph partition lattice on [๐‘›] partition structure on structure ๐’ฎ even partition structure on structure ๐’ฎ odd partition structure on structure ๐’ฎ rational numbers set of compositions of the integer ๐‘› number of compositions of the integer ๐‘› set of compositions of ๐‘› into ๐‘˜ parts number of partitions of ๐‘› into ๐‘˜ parts algebra of quasisymmetric functions quasisymmetric functions of degree ๐‘› recording tableau of ๐œ‹ quasisymmetric function for patterns ฮ  real numbers set of functions reverse compatible with ๐œ‹ rank of a ranked poset ๐‘ƒ ๐‘˜th rank set of a ranked poset ๐‘ƒ rank of an element ๐‘ฅ in a ranked poset set of partitions contained in a ๐‘˜ ร— ๐‘™ rectangle set of reverse ๐‘ƒ-partitions reduced incidence algebra of a binomial poset number of shape ๐œ† reverse plane partitions of ๐‘› generating function for reverse ๐‘ƒ-partitions row insertion of element ๐‘ฅ into tableau ๐‘ƒ vertex partition induced by edge set ๐น representation of group ๐บ labeled structure on label set ๐ฟ pattern poset symmetric group on [๐‘›] summation operator applied to function ๐‘“(๐‘›) sign function on a signed set shape of tableau ๐‘‡ signed Stirling number of the first kind

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Page 15 15 4 4 5 242 56 140 131 133 133 1 16 16 16 16 268 268 242 277 1 270 147 147 147 79 271 179 233 271 241 255 287 124 140 11 162 44 225 13

xviii

List of Notation

Symbol

Definition

๐‘†(๐‘‡, ๐‘˜) ๐‘†(๐‘›, ๐‘˜) ๐‘†๐‘œ (๐ฟ, ๐‘˜) ๐’ฎ๐‘‡(๐บ) st std ๐œŽ Supp ๐‘ฅ supp ๐‘ฅ Sym Sym๐‘› SYT(๐œ†) SSYT(๐œ†) ๐‘ ๐œ† ๐‘‡๐‘–,๐‘— ๐’ฏ๐‘› ๐‘ˆ(๐‘†) ๐‘‰(๐ท) ๐‘‰(๐บ) ๐‘‰โˆž ๐‘ค ๐‘˜ (๐‘ƒ) ๐‘Š ๐‘˜ (๐‘ƒ) ๐‘Š๐‘› wt ๐ฑ ๐ฑ๐‘ ๐ฑ๐‘“ ๐ฑ๐‘‡ ๐‘‹๐‘” ๐‘‹(๐บ; ๐ฑ) ๐‘‹(๐บ; ๐ฑ, ๐‘ž) ๐‘Œ โ„ค ๐œ(๐‘ฅ, ๐‘ง) ๐œ(๐‘ ) ๐‘ง(๐‘”) ๐‘(๐บ) #๐‘† |๐‘“| |๐‘†| |๐‘‡| ๐‘†โŠŽ๐‘‡

set of partitions of the set ๐‘‡ into ๐‘˜ blocks Stirling number of the second kind set of ordered partitions of the set ๐ฟ into ๐‘˜ blocks set of spanning trees of graph ๐บ statistic on a set standardization of the permutation ๐œŽ support set of ๐‘ฅ in a product of claws size of support set of ๐‘ฅ in a product of claws algebra of symmetric functions symmetric functions of degree ๐‘› set of standard Young tableaux of shape ๐œ† set of semistandard Young tableaux of shape ๐œ† Schur function element in cell (๐‘–, ๐‘—) of tableau ๐‘‡ set of monomino-domino tilings of a row of ๐‘› squares upper-order ideal generated by ๐‘† in a poset vertex set of digraph ๐ท vertex set of graph ๐บ vector space with a countably infinite basis over ๐”ฝ๐‘ž Whitney number of the first kind for a poset ๐‘ƒ Whitney number of the second kind for a poset ๐‘ƒ walk with ๐‘› vertices weight function on a set a countably infinite set of variables monomial for a coloring ๐‘ of a graph monomial for a function ๐‘“ monomial for a tableau ๐‘‡ fixed points of group element ๐‘” acting on set ๐‘‹ chromatic symmetric function of graph ๐บ chromatic quasisymmetric function of graph ๐บ Youngโ€™s lattice set of integers zeta function in the incidence algebra of a poset Riemann zeta function cycle index of group element ๐‘” cycle index of group ๐บ cardinality of the set ๐‘† size (sum of values) of a function cardinality of the set ๐‘† sum of entries of tableau ๐‘‡ disjoint union of sets ๐‘† and ๐‘‡

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Page 10 10 127 59 74 28 173 173 220 220 224 225 225 225 3 143 21 18 178 156 156 19 86 219 253 270 225 192 253 280 140 1 159 182 197 197 1 236 1 233 1

List of Notation

xix

Symbol

Definition

|๐œ†| ๐œ†โŠข๐‘› ๐‘†ร—๐‘‡ ๐‘ƒโŠŽ๐‘„ ๐‘ƒโŠ•๐‘„ ๐‘ƒร—๐‘„ [๐‘”] [๐‘”]๐ต [๐‘›] [๐‘›]๐‘ž [๐‘›]๐‘ž ! [๐‘ฅ๐‘› ]๐‘“(๐‘ฅ) ๐‘›โ†“๐‘˜ 2๐‘† (๐‘†๐‘˜)

sum of the parts of partition ๐œ† ๐œ† is a partition of ๐‘› (Cartesian) product of sets ๐‘† and ๐‘‡ disjoint union of posets ๐‘ƒ and ๐‘„ ordinal sum of posets ๐‘ƒ and ๐‘„ (Cartesian) product of posets ๐‘ƒ and ๐‘„ linear transformation for group element ๐‘” matrix in basis ๐ต for group element ๐‘” set of integers {1, 2, . . . , ๐‘›} ๐‘ž-analogue of nonnegative integer ๐‘› ๐‘ž-analogue of ๐‘›! coefficient of ๐‘ฅ๐‘› in ๐‘“(๐‘ฅ) ๐‘› falling factorial with ๐‘˜ factors set of subsets of ๐‘† set of ๐‘˜-element subsets of ๐‘†

(๐‘›๐‘˜)

binomial coefficient

[๐‘›๐‘˜] ๐‘ž

๐‘ž-binomial coefficient

[๐‘‰๐‘˜ ] {{๐‘Ž, ๐‘Ž, . . . }} {{๐‘Ž2 , . . . }} ((๐‘†๐‘˜)) ๐œ’(๐บ) ๐œ’(๐‘”) ๐‘ฅโ‹–๐‘ฆ ๐‘ฆโ‹—๐‘ฅ 0ฬ‚ 1ฬ‚ [๐‘ฅ, ๐‘ฆ] ๐‘ฅโˆง๐‘ฆ โ‹€๐‘‹ ๐‘ฅโˆจ๐‘ฆ ๐‘ˆ +๐‘‰ ๐‘“โˆ—๐‘” ๐œ’(๐‘ƒ; ๐‘ก) ๐‘ƒ/ โˆผ ๐œ”๐‘›

๐‘˜-dimensional subspaces of vector space ๐‘‰ multiset individual element notation multiset multiplicity notation set of ๐‘˜-element multisubsets of ๐‘† chromatic number of ๐บ character of group element ๐‘” ๐‘ฅ is covered by ๐‘ฆ in a poset ๐‘ฆ covers ๐‘ฅ in a poset the minimum element of a poset the maximum element of a poset closed interval from ๐‘ฅ to ๐‘ฆ in a poset meet of ๐‘ฅ and ๐‘ฆ in a poset meet of the subset ๐‘‹ in a poset join of ๐‘ฅ and ๐‘ฆ in a poset sum of subspaces ๐‘ˆ and ๐‘‰ convolution of ๐‘“ and ๐‘” in the incidence algebra characteristic polynomial of a ranked poset ๐‘ƒ quotient of poset ๐‘ƒ by equivalence relation โˆผ primitive ๐‘›th root of unity

79 8 8 9 99 291 140 140 142 142 143 148 149 149 149 158 164 169 210

Robinsonโ€“Schensted map

242

Robinsonโ€“Schenstedโ€“Knuth map

244

RS

๐œ‹ โ†ฆ (๐‘ƒ, ๐‘„) RSK

๐‘€ โ†ฆ (๐‘‡, ๐‘ˆ)

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Page 13 13 1 145 146 146 287 287 4 75 75 83 4 5 6 7 77

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

10.1090/gsm/210/01

Chapter 1

Basic Counting

In this chapter we will develop the most elementary techniques for enumerating sets. Even though these methods are relatively basic, they will presage more complicated things to come. We denote the integers by โ„ค and parameters such as ๐‘› and ๐‘˜ are always assumed to be integral unless otherwise indicated. We also use the notation โ„• and โ„™ for the nonnegative and positive integers, respectively. As usual, โ„š, โ„, and โ„‚ stand for the rational numbers, real numbers, and complex numbers, respectively. Finally, whenever taking the cardinality of a set we will assume it is finite.

1.1. The Sum and Product Rules for sets The Sum and Product Rules for sets are the basis for much of enumeration. And we will see various extensions of them later to ordinary and exponential generating functions. Although the rules are very easy to prove, we will include the demonstrations because the results are so useful. Given a finite set ๐‘†, we will use either of the notations #๐‘† or |๐‘†| for its cardinality. We will also write ๐‘† โŠŽ ๐‘‡ for the disjoint union of ๐‘† and ๐‘‡, and usage of this symbol implies disjointness even if it has not been previously explicitly stated. Finally, our notation for the (Cartesian) product of sets is ๐‘† ร— ๐‘‡ = {(๐‘ , ๐‘ก) โˆฃ ๐‘  โˆˆ ๐‘†, ๐‘ก โˆˆ ๐‘‡}. Lemma 1.1.1. Let ๐‘†, ๐‘‡ be finite sets. (a) (Sum Rule) If ๐‘† โˆฉ ๐‘‡ = โˆ…, then |๐‘† โŠŽ ๐‘‡| = |๐‘†| + |๐‘‡|. (b) (Product Rule) For any finite sets |๐‘† ร— ๐‘‡| = |๐‘†| โ‹… |๐‘‡|. Proof. Let ๐‘† = {๐‘ 1 , . . . , ๐‘ ๐‘š } and ๐‘‡ = {๐‘ก1 , . . . , ๐‘ก๐‘› }. For part (a), if ๐‘† and ๐‘‡ are disjoint, then we have ๐‘† โŠŽ ๐‘‡ = {๐‘ 1 , . . . , ๐‘ ๐‘š , ๐‘ก1 , . . . , ๐‘ก๐‘› } so that |๐‘† โŠŽ ๐‘‡| = ๐‘š + ๐‘› = |๐‘†| + |๐‘‡|. 1 Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

2

1. Basic Counting

For part (b), we induct on ๐‘› = |๐‘‡|. If ๐‘‡ = โˆ…, then ๐‘† ร— ๐‘‡ = โˆ… so that |๐‘† ร— ๐‘‡| = 0 as desired. If |๐‘‡| โ‰ฅ 1, then let ๐‘‡ โ€ฒ = ๐‘‡ โˆ’ {๐‘ก๐‘› }. We can write ๐‘† ร— ๐‘‡ = (๐‘† ร— ๐‘‡ โ€ฒ ) โŠŽ (๐‘† ร— {๐‘ก๐‘› }). Also ๐‘† ร— {๐‘ก๐‘› } = {(๐‘ 1 , ๐‘ก๐‘› ), . . . , (๐‘ ๐‘š , ๐‘ก๐‘› )}, which has |๐‘†| = ๐‘š elements since the second component is constant. Now by part (a) and induction |๐‘† ร— ๐‘‡| = |๐‘† ร— ๐‘‡ โ€ฒ | + |๐‘† ร— {๐‘ก๐‘› }| = ๐‘š(๐‘› โˆ’ 1) + ๐‘š = ๐‘š๐‘›, which finishes the proof.

โ–ก

In combinatorial choice problems, one is often given either the option to do one operation or another, or to do both. Suppose there are ๐‘š ways of doing the first operation and ๐‘› ways of doing the second. If there is no common operation, then the Sum Rule tells us that the number of ways to do one or the other is ๐‘š + ๐‘›. And if doing the first operation has no effect on doing the second, then the Product Rule gives a count of ๐‘š๐‘› for doing the first and then the second. More generally if there are ๐‘š ways of doing the first operation and, no matter which of the ๐‘š is chosen, the number of ways to continue with the second operation is ๐‘›, then again there are ๐‘š๐‘› ways to do both. (The actual ๐‘› second operations available may depend on the choice of the first, but not their number.) So in practice one translates from English to mathematics by replacing โ€œorโ€ with addition and โ€œandโ€ with multiplication. Another important concept related to cardinalities is that of a bijection. A bijection between sets ๐‘†, ๐‘‡ is a function ๐‘“ โˆถ ๐‘† โ†’ ๐‘‡ which is both injective (one-to-one) and surjective (onto). If ๐‘†, ๐‘‡ are finite, then the existence of a bijection between them implies that |๐‘†| = |๐‘‡|. (One can extend this notion to infinite sets, but we will have no cause to do so here.) In combinatorics, one often uses bijections to prove that two sets have the same cardinality. See, for just one of many examples, the proof of Theorem 1.1.2 below. We will illustrate these ideas with one of the most famous sequences in all of combinatorics: the Fibonacci numbers. As is sometimes the case, there is an amusing (if somewhat improbable) story attached to the sequence. One starts at the beginning of time with a pair of immature rabbits, one male and one female. It takes one month for rabbits to mature. In every subsequent month a pair gives birth to another pair of immature rabbits, one male and one female. If rabbits only breed with their birth partner and live forever (as I said, the story is somewhat improbable), how many pairs of rabbits are there at the beginning of month ๐‘›? Let us call this number ๐น๐‘› . It will be convenient to let ๐น0 = 0. Since we begin with only one pair, ๐น1 = 1. And at the beginning of the second month, the pair has matured but produced no offspring, so ๐น2 = 1. In subsequent months, one has all the rabbits from the previous month, counted by ๐น๐‘›โˆ’1 , together with the newborn pairs. The number of newborn pairs equals the number of mature pairs from the previous month, which equals the total number of pairs from the month before which is ๐น๐‘›โˆ’2 . Thus, applying the Sum Rule, (1.1)

๐น๐‘› = ๐น๐‘›โˆ’1 + ๐น๐‘›โˆ’2 for ๐‘› โ‰ฅ 2 with ๐น0 = 0 and ๐น1 = 1

where we can start the recursion at ๐‘› = 2 rather than ๐‘› = 3 due to letting ๐น0 = 0. The ๐น๐‘› are called the Fibonacci numbers. It is also important to note that some authors Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

1.1. The Sum and Product Rules for sets

3

Figure 1.1. ๐’ฏ3

define this sequence by letting (1.2)

๐‘“0 = ๐‘“1 = 1 and ๐‘“๐‘› = ๐‘“๐‘›โˆ’1 + ๐‘“๐‘›โˆ’2 for ๐‘› โ‰ฅ 2.

So it is important to make sure which flavor of Fibonacci is being discussed in a given context. One might wonder if there is an explicit formula for ๐น๐‘› in addition to the recursive one above. We will see that such an expression exists, although it is far from obvious how to derive it from what we have done so far. Indeed, we will need the theory of ordinary generating functions discussed in Chapter 3 to derive it. Another thing which might be desired is a combinatorial interpretation for ๐น๐‘› . A combinatorial interpretation for a sequence of nonnegative integers ๐‘Ž0 , ๐‘Ž1 , ๐‘Ž2 , . . . is a sequence of sets ๐‘† 0 , ๐‘† 1 , ๐‘† 2 , . . . such that #๐‘†๐‘› = ๐‘Ž๐‘› for all ๐‘›. Such interpretations often give rise to very pretty and intuitive proofs about the original sequence and so are highly desirable. One could argue that the story of the rabbits already gives such an interpretation. But we would like something more amenable to mathematical manipulation. Suppose we are given a row of squares. We are also given two types of tiles: dominos which can cover two squares and monominos which can cover one. A tiling of the row is a set of tiles which covers each square exactly once. Let ๐’ฏ๐‘› be the set of tilings of a row of ๐‘› squares. See Figure 1.1 for a list of the elements of ๐’ฏ3 . There is a simple relationship between tilings and Fibonacci numbers. Theorem 1.1.2. For ๐‘› โ‰ฅ 1 we have ๐น๐‘› = #๐’ฏ๐‘›โˆ’1 . Proof. It suffices to prove that both sides of this equation satisfy the same initial conditions and recurrence relation. When the row contains no squares, it only has the empty tiling so ๐’ฏ0 = 1 = ๐น1 . And when there is one square, it can only be tiled by a monomino so ๐’ฏ1 = 1 = ๐น2 . For the recursion, the tilings in ๐’ฏ๐‘› can be divided into two types: those which end with a monomino and those which end with a domino. Removing the last tile shows that these tilings are in bijection with those in ๐’ฏ๐‘›โˆ’1 and those in ๐’ฏ๐‘›โˆ’2 , respectively. Thus #๐’ฏ๐‘› = #๐’ฏ๐‘›โˆ’1 + #๐’ฏ๐‘›โˆ’2 as desired. โ–ก To see the power of a good combinatorial interpretation, we will now give a simple proof of an identity for the ๐น๐‘› . Such identities are legion. See, for example, the book of Benjamin and Quinn [10]. Corollary 1.1.3. For ๐‘š โ‰ฅ 1 and ๐‘› โ‰ฅ 0 we have ๐น๐‘š+๐‘› = ๐น๐‘šโˆ’1 ๐น๐‘› + ๐น๐‘š ๐น๐‘›+1 . Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

4

1. Basic Counting

Proof. By the previous theorem, the left-hand side counts the number of tilings of a row of ๐‘š + ๐‘› โˆ’ 1 squares. So it suffices to show that the same is true of the right. Label the squares 1, . . . , ๐‘š + ๐‘› โˆ’ 1 from left to right. We can write ๐’ฏ๐‘š+๐‘›โˆ’1 = ๐’ฎ โŠŽ ๐’ฏ where ๐’ฎ contains those tilings with a domino covering squares ๐‘š โˆ’ 1 and ๐‘š, and ๐’ฏ has the tilings with ๐‘šโˆ’1 and ๐‘š in different tiles. The tilings in ๐’ฏ are essentially pairs of tilings, the first covering the first ๐‘š โˆ’ 1 square and second covering the last ๐‘› squares. So the Product Rule gives |๐’ฏ| = |๐’ฏ๐‘šโˆ’1 | โ‹… |๐’ฏ๐‘› | = ๐น๐‘š ๐น๐‘›+1 . Removing the given domino from the tilings in ๐’ฎ again splits each tiling into a pair with the first covering ๐‘š โˆ’ 2 squares and the second ๐‘› โˆ’ 1. Taking cardinalities results in |๐’ฎ| = ๐น๐‘šโˆ’1 ๐น๐‘› . Finally, applying the Sum Rule finishes the proof. โ–ก The demonstration just given is called a combinatorial proof since it involves counting discrete objects. We will meet other useful proof techniques as we go along. But combinatorial proofs are often considered to be the most pleasant, in part because they can be more illuminating than demonstrations just involving formal manipulations.

1.2. Permutations and words It is always important when considering an enumeration problem to determine whether the objects being considered are ordered or not. In this section we will consider the most basic ordered structures, namely permutations and words. If ๐‘† is a set with #๐‘† = ๐‘›, then a permutation of ๐‘† is a sequence ๐œ‹ = ๐œ‹1 . . . ๐œ‹๐‘› obtained by listing the elements of ๐‘† in some order. If ๐œ‹ is a permutation, we will always use ๐œ‹๐‘– to denote the ๐‘–th element of ๐œ‹ and similarly for other ordered structures. We let ๐‘ƒ(๐‘†) denote the set of all permutations of ๐‘†. For example, ๐‘ƒ({๐‘Ž, ๐‘, ๐‘}) = {๐‘Ž๐‘๐‘, ๐‘Ž๐‘๐‘, ๐‘๐‘Ž๐‘, ๐‘๐‘๐‘Ž, ๐‘๐‘Ž๐‘, ๐‘๐‘๐‘Ž}. Clearly #๐‘ƒ(๐‘†) only depends on #๐‘†. So often we choose the canonical ๐‘›-element set [๐‘›] = {1, 2, . . . , ๐‘›}. We can also consider ๐‘˜-permutations of ๐‘† which are sequences ๐œ‹ = ๐œ‹1 . . . ๐œ‹๐‘˜ obtained by linearly ordering ๐‘˜ distinct elements of ๐‘†. Here, ๐‘˜ is called the length of the permutation and we write โ„“(๐œ‹) = ๐‘˜. Again, we use the same terminology and notation for other ordered structures. The set of all ๐‘˜-permutations of ๐‘† is denoted ๐‘ƒ(๐‘†, ๐‘˜). By way of illustration, ๐‘ƒ({๐‘Ž, ๐‘, ๐‘, ๐‘‘}, 2) = {๐‘Ž๐‘, ๐‘๐‘Ž, ๐‘Ž๐‘, ๐‘๐‘Ž, ๐‘Ž๐‘‘, ๐‘‘๐‘Ž, ๐‘๐‘, ๐‘๐‘, ๐‘๐‘‘, ๐‘‘๐‘, ๐‘๐‘‘, ๐‘‘๐‘}. In particular, if #๐‘† = ๐‘›, then ๐‘ƒ(๐‘†, ๐‘›) = ๐‘ƒ(๐‘†). Also ๐‘ƒ(๐‘†, ๐‘˜) = โˆ… for ๐‘˜ > ๐‘› since in this case it is impossible to pick ๐‘˜ distinct elements from a set with only ๐‘›. And ๐‘ƒ(๐‘†, 0) = {๐œ–} where ๐œ– is the empty sequence. To count permutations it will be convenient to introduce the following notation. Given nonnegative integers ๐‘›, ๐‘˜, we can form the falling factorial ๐‘›โ†“๐‘˜ = ๐‘›(๐‘› โˆ’ 1) . . . (๐‘› โˆ’ ๐‘˜ + 1). Note that ๐‘˜ equals the number of factors in the product. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

1.3. Combinations and subsets

5

Theorem 1.2.1. For ๐‘›, ๐‘˜ โ‰ฅ 0 we have #๐‘ƒ([๐‘›], ๐‘˜) = ๐‘›โ†“๐‘˜ . In particular #๐‘ƒ([๐‘›]) = ๐‘›! . Proof. Since ๐‘ƒ([๐‘›]) = ๐‘ƒ([๐‘›], ๐‘›), it suffices to prove the first formula. Given ๐œ‹ = ๐œ‹1 . . . ๐œ‹๐‘˜ โˆˆ ๐‘ƒ([๐‘›], ๐‘˜), there are ๐‘› ways to pick ๐œ‹1 . Since ๐œ‹2 โ‰  ๐œ‹1 , there remains ๐‘› โˆ’ 1 choices for ๐œ‹2 . Since the number of choices for ๐œ‹2 does not depend on the actual element chosen for ๐œ‹, one can continue in this way and apply a modified version of the Product Rule to obtain the result. โ–ก Note that when 0 โ‰ค ๐‘˜ โ‰ค ๐‘› we can write (1.3)

๐‘›โ†“๐‘˜ =

๐‘›! . (๐‘› โˆ’ ๐‘˜)!

But for ๐‘˜ > ๐‘› the product ๐‘› โ†“๐‘˜ still makes sense, even though the product cannot be expressed as a quotient of factorials. Indeed, if ๐‘˜ > ๐‘›, then zero is a factor and so ๐‘› โ†“๐‘˜ = 0, which agrees with the fact that ๐‘ƒ([๐‘›], ๐‘˜) = โˆ…. In the special case ๐‘˜ = 0 we have ๐‘› โ†“๐‘˜ = 1 because it is an empty product. Again, this reflects the combinatorics in that ๐‘ƒ([๐‘›], 0) = {๐œ–}. One of the other things to keep track of in a combinatorial problem is whether elements are allowed to be repeated or not. In permutations we have no repetitions. But the case when they are allowed is interesting as well. A ๐‘˜-word over a set ๐‘† is a sequence ๐‘ค = ๐‘ค 1 . . . ๐‘ค ๐‘˜ where ๐‘ค ๐‘– โˆˆ ๐‘† for all ๐‘–. Note that there is no assumption that the ๐‘ค ๐‘– are distinct. We denote the set of ๐‘˜-words over ๐‘† by ๐‘ƒ((๐‘†, ๐‘˜)). Note the use of the double parentheses to denote the fact that repetitions are allowed. Note also that ๐‘ƒ(๐‘†, ๐‘˜) โŠ† ๐‘ƒ((๐‘†, ๐‘˜)), but usually the inclusion is strict. To illustrate ๐‘ƒ(({๐‘Ž, ๐‘, ๐‘, ๐‘‘}, 2)) = ๐‘ƒ({๐‘Ž, ๐‘, ๐‘, ๐‘‘}, 2) โŠŽ {๐‘Ž๐‘Ž, ๐‘๐‘, ๐‘๐‘, ๐‘‘๐‘‘}. The proof of the next result is almost identical to that of Theorem 1.2.1 and so is left to the reader. When a result is given without proof, this is indicated by a box at the end of its statement. Theorem 1.2.2. For ๐‘›, ๐‘˜ โ‰ฅ 0 we have #๐‘ƒ(([๐‘›], ๐‘˜)) = ๐‘›๐‘˜ .

โ–ก

1.3. Combinations and subsets We will now consider unordered versions of the combinatorial objects studied in the last section. These are sometimes called combinations, although the reader may know them by their more familiar name: subsets. Given a set ๐‘†, we let 2๐‘† denote the set of all subsets of ๐‘†. Notice that 2๐‘† is a set, not a number. For example, 2{๐‘Ž,๐‘,๐‘} = {โˆ…, {๐‘Ž}, {๐‘}, {๐‘}, {๐‘Ž, ๐‘}, {๐‘Ž, ๐‘}, {๐‘, ๐‘}, {๐‘Ž, ๐‘, ๐‘}}. The reason for this notation should be made clear by the following result. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

6

1. Basic Counting

Theorem 1.3.1. For ๐‘› โ‰ฅ 0 we have #2[๐‘›] = 2๐‘› . Proof. By Theorem 1.2.2 we have 2๐‘› = #๐‘ƒ(({0, 1}, ๐‘›)). So it suffices to find a bijection ๐‘“ โˆถ 2[๐‘›] โ†’ ๐‘ƒ(({0, 1}, ๐‘›)), and there is a canonical one. In particular, if ๐‘† โŠ† [๐‘›], then we let ๐‘“(๐‘†) = ๐‘ค 1 . . . ๐‘ค ๐‘› where, for all ๐‘–, 1 if ๐‘– โˆˆ ๐‘†, ๐‘ค๐‘– = { 0 if ๐‘– โˆ‰ ๐‘†. To show that ๐‘“ is bijective, it suffices to find its inverse. If ๐‘ค = ๐‘ค 1 . . . ๐‘ค ๐‘› โˆˆ ๐‘ƒ(({0, 1}, ๐‘›)), then we let ๐‘“โˆ’1 (๐‘ค) = ๐‘† where ๐‘– โˆˆ ๐‘† if ๐‘ค ๐‘– = 1 and ๐‘– โˆ‰ ๐‘† if ๐‘ค ๐‘– = 0 where 1 โ‰ค ๐‘– โ‰ค ๐‘›. It is easy to check that the compositions ๐‘“ โˆ˜ ๐‘“โˆ’1 and ๐‘“โˆ’1 โˆ˜ ๐‘“ are the identity maps on their respective domains. This completes the proof. โ–ก The proof just given is called a bijective proof and it is a particularly nice kind of combinatorial proof. This is because bijective proofs can relate different types of combinatorial objects, sometimes revealing unexpected connections. Also note that we proved ๐‘“ bijective by finding its inverse rather than showing directly that it was oneto-one and onto. This is the preferred method as having a concrete description of ๐‘“โˆ’1 can be useful later. Finally, when dealing with functions we will always compose them right-to-left so that (๐‘“ โˆ˜ ๐‘”)(๐‘ฅ) = ๐‘“(๐‘”(๐‘ฅ)). We now want to count subsets by their cardinality. For a set ๐‘† we will use the notation ๐‘† ( ) = {๐‘‡ โŠ† ๐‘† โˆฃ #๐‘‡ = ๐‘˜}. ๐‘˜ As an example, {๐‘Ž, ๐‘, ๐‘} ( ) = {{๐‘Ž, ๐‘}, {๐‘Ž, ๐‘}, {๐‘, ๐‘}}. 2 As expected, we now find the cardinality of this set. Theorem 1.3.2. For ๐‘›, ๐‘˜ โ‰ฅ 0 we have #(

๐‘›โ†“๐‘˜ [๐‘›] . )= ๐‘˜ ๐‘˜!

Proof. Cross-multiplying and using Theorem 1.2.1 we see that it suffices to prove #๐‘ƒ([๐‘›], ๐‘˜) = ๐‘˜! โ‹…#(

[๐‘›] ). ๐‘˜

To see this, note that we can get each ๐œ‹1 . . . ๐œ‹๐‘˜ โˆˆ ๐‘ƒ([๐‘›], ๐‘˜) exactly once by running through the subsets ๐‘† = {๐‘ 1 , . . . , ๐‘ ๐‘˜ } โŠ† [๐‘›] and then ordering each ๐‘† in all possible ways. The number of choices for ๐‘† is #([๐‘›] ) and, by Theorem 1.2.1 again, the number ๐‘˜ of ways of permuting the elements of ๐‘† is ๐‘˜!. So we are done by the Product Rule. โ–ก Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

1.3. Combinations and subsets

7

1 1 1 1 1

1 2

3 4

1 3

6

1 4

1

Figure 1.2. Rows 0 through 4 of Pascalโ€™s triangle

Given ๐‘›, ๐‘˜ โ‰ฅ 0, we define the binomial coefficient ๐‘›โ†“๐‘˜ ๐‘› [๐‘›] . ( ) = #( ) = ๐‘˜ ๐‘˜ ๐‘˜!

(1.4)

The reason for this name is that these numbers appear in the binomial expansion which will be studied in Chapter 3. Often you will see the binomial coefficients displayed in a triangular array called Pascalโ€™s triangle which has (๐‘›๐‘˜) as the entry in the ๐‘›th row and ๐‘˜th diagonal. When ๐‘˜ > ๐‘› it is traditional to omit the zeros. See Figure 1.2 for rows 0 through 4. (We apologize to the reader for not writing out the whole triangle, but this page is not big enough.) For 0 โ‰ค ๐‘˜ โ‰ค ๐‘› we can use (1.3) to write ๐‘›! ๐‘› , ( )= ๐‘˜ ๐‘˜! (๐‘› โˆ’ ๐‘˜)!

(1.5)

which is pleasing because of its symmetry. We can also extend the binomial coefficients to ๐‘˜ < 0 by letting (๐‘›๐‘˜) = 0. This is in keeping with the fact that ([๐‘›] ) = โˆ… in this case. ๐‘˜ In the next theorem, we collect various basic results about binomial coefficients which will be useful in the sequel. In it, we will use the Kronecker delta function defined by 1 if ๐‘ฅ = ๐‘ฆ, ๐›ฟ๐‘ฅ,๐‘ฆ = { 0 if ๐‘ฅ โ‰  ๐‘ฆ. Also note that we do not specify the range of the summation variable ๐‘˜ in (c) and (d) because it can be taken as either 0 โ‰ค ๐‘˜ โ‰ค ๐‘› or ๐‘˜ โˆˆ โ„ค since the extra terms in the larger sum are all zero. Both viewpoints will be useful on occasion. Theorem 1.3.3. Suppose ๐‘› โ‰ฅ 0. (a) The binomial coefficients satisfy the initial condition 0 ( ) = ๐›ฟ ๐‘˜,0 ๐‘˜ and recurrence relation ๐‘› ๐‘›โˆ’1 ๐‘›โˆ’1 ( )=( )+( ) ๐‘˜ ๐‘˜โˆ’1 ๐‘˜ for ๐‘› โ‰ฅ 1. (b) The binomial coefficients are symmetric, meaning that ๐‘› ๐‘› ( )=( ). ๐‘˜ ๐‘›โˆ’๐‘˜ Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

8

1. Basic Counting

(c) We have ๐‘› โˆ‘ ( ) = 2๐‘› . ๐‘˜ ๐‘˜ (d) We have ๐‘› โˆ‘(โˆ’1)๐‘˜ ( ) = ๐›ฟ๐‘›,0 . ๐‘˜ ๐‘˜ Proof. (a) The initial condition is clear. For the recursion let ๐’ฎ1 be the set of ๐‘† โˆˆ ([๐‘›] ) ๐‘˜ [๐‘›] [๐‘›] with ๐‘› โˆˆ ๐‘†, and let ๐’ฎ2 be the set of ๐‘† โˆˆ ( ๐‘˜ ) with ๐‘› โˆ‰ ๐‘†. Then ( ๐‘˜ ) = ๐’ฎ1 โŠŽ ๐’ฎ2 . But if ๐‘› โˆˆ ๐‘†, then ๐‘† โˆ’ {๐‘›} โˆˆ ([๐‘›โˆ’1] ). This gives a bijection between ๐’ฎ1 and ([๐‘›โˆ’1] ) so that ๐‘˜โˆ’1 ๐‘˜โˆ’1 ๐‘›โˆ’1 ๐‘›โˆ’1 and this implies #๐’ฎ1 = (๐‘˜โˆ’1). On the other hand, if ๐‘› โˆ‰ ๐‘†, then ๐‘† โˆˆ ([๐‘›โˆ’1] #๐’ฎ ) 2 = ( ๐‘˜ ). ๐‘˜ Applying the Sum Rule completes the proof. [๐‘›] (b) It suffices to find a bijection ๐‘“ โˆถ ([๐‘›] ) โ†’ (๐‘›โˆ’๐‘˜ ). Consider the map ๐‘“ โˆถ 2[๐‘›] โ†’ 2[๐‘›] ๐‘˜ by ๐‘“(๐‘†) = [๐‘›] โˆ’ ๐‘† where the minus sign indicates difference of sets. Note that the composition ๐‘“2 is the identity map so that ๐‘“ is a bijection. Furthermore ๐‘† โˆˆ ([๐‘›] ) if and ๐‘˜ [๐‘›] only if ๐‘“(๐‘†) โˆˆ (๐‘›โˆ’๐‘˜ ). So ๐‘“ restricts to a bijection between these two sets.

(c) This follows by applying the Sum Rule to the equation 2[๐‘›] = โจ„๐‘˜ ([๐‘›] ). ๐‘˜ (d) The case ๐‘› = 0 is easy, so we assume ๐‘› > 0. We will learn general techniques for dealing with equations involving signs in the next chapter. But for now, we try to prove the equivalent equality ๐‘› ๐‘› โˆ‘ ( ) = โˆ‘ ( ). ๐‘˜ ๐‘˜ ๐‘˜ odd ๐‘˜ even Let ๐’ฏ1 be the set of ๐‘‡ โˆˆ 2[๐‘›] with #๐‘‡ odd and let ๐’ฏ2 be the set of ๐‘‡ โˆˆ 2[๐‘›] with #๐‘‡ even. We wish to find a bijection ๐‘” โˆถ ๐’ฏ1 โ†’ ๐’ฏ2 . Consider the operation of symmetric difference ๐‘† ฮ” ๐‘‡ = (๐‘† โˆ’ ๐‘‡) โŠŽ (๐‘‡ โˆ’ ๐‘†). It is not hard to see that (๐‘† ฮ” ๐‘‡) ฮ” ๐‘‡ = ๐‘†. Now define ๐‘” โˆถ 2[๐‘›] โ†’ 2[๐‘›] by ๐‘”(๐‘‡) = ๐‘‡ ฮ” {๐‘›} so that, by the previous sentence, ๐‘”2 is the identity. Furthermore, ๐‘” reverses parity and so restricts to the desired bijection. โ–ก As with the case of permutations and words, we want to enumerate โ€œsetsโ€ where repetitions are allowed. A multiset ๐‘€ is an unordered collection of elements which may be repeated. For example ๐‘€ = {{๐‘Ž, ๐‘Ž, ๐‘Ž, ๐‘, ๐‘, ๐‘}} = {{๐‘, ๐‘Ž, ๐‘, ๐‘Ž, ๐‘, ๐‘Ž}}. Note the use of double curly brackets to denote a multiset. We will also use multiplicity notation where ๐‘Ž๐‘š denotes ๐‘š copies of the element ๐‘Ž. Continuing our example ๐‘€ = {{๐‘Ž3 , ๐‘, ๐‘2 }}. As with powers, an exponent of one is optional and an exponent of zero indicates that there are no copies of that element in the multiset. The cardinality of a multiset is its number of elements counted with multiplicity. So in our example #๐‘€ = 2 + 1 + 3 = 6. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

1.3. Combinations and subsets

9

If ๐‘† is a set, then ๐‘€ is a multiset on ๐‘† if every element of ๐‘€ is an element of ๐‘†. We let ((๐‘†๐‘˜)) be the set of all multisets on ๐‘† of cardinality ๐‘˜ and ๐‘› [๐‘›] (( )) = #(( )). ๐‘˜ ๐‘˜ To illustrate {๐‘Ž, ๐‘, ๐‘} (( )) = { {{๐‘Ž, ๐‘Ž}}, {{๐‘Ž, ๐‘}}, {{๐‘Ž, ๐‘}}, {{๐‘, ๐‘}}, {{๐‘, ๐‘}}, {{๐‘, ๐‘}} } 2 and so ((32)) = 6. Theorem 1.3.4. For ๐‘›, ๐‘˜ โ‰ฅ 0 we have ๐‘› ๐‘›+๐‘˜โˆ’1 (( )) = ( ). ๐‘˜ ๐‘˜ Proof. We wish to find a bijection ๐‘“ โˆถ ((

[๐‘› + ๐‘˜ โˆ’ 1] [๐‘›] )) โ†’ ( ). ๐‘˜ ๐‘˜

Given a multiset ๐‘€ = {{๐‘š1 โ‰ค ๐‘š2 โ‰ค ๐‘š3 โ‰ค โ‹ฏ โ‰ค ๐‘š๐‘˜ }} on [๐‘›], let ๐‘“(๐‘€) = {๐‘š1 < ๐‘š2 + 1 < ๐‘š3 + 2 < โ‹ฏ < ๐‘š๐‘˜ + ๐‘˜ โˆ’ 1}. Now the ๐‘š๐‘– + ๐‘– โˆ’ 1 are distinct, and the fact that ๐‘š๐‘˜ โ‰ค ๐‘› implies ๐‘š๐‘˜ + ๐‘˜ โˆ’ 1 โ‰ค ๐‘› + ๐‘˜ โˆ’ 1. It follows that ๐‘“(๐‘€) โˆˆ ([๐‘›+๐‘˜โˆ’1] ) and so the map is well-defined. It should now be easy ๐‘˜ for the reader to construct an inverse, proving that ๐‘“ is bijective. โ–ก As with the binomial coefficients, we extend ((๐‘›๐‘˜)) to negative ๐‘˜ by letting it equal zero. In the future we will do the same for other constants whose natural domain of definition is ๐‘›, ๐‘˜ โ‰ฅ 0 without comment. We do wish to comment on an interesting relationship between counting sets and multisets. Note that definition (1.4) is well-defined for any complex number ๐‘› since the falling factorial is just a product, and in particular it makes sense for negative integers. In fact, if ๐‘› โˆˆ โ„•, then (1.6)

(

โˆ’๐‘› (โˆ’๐‘›)(โˆ’๐‘› โˆ’ 1) โ‹ฏ (โˆ’๐‘› โˆ’ ๐‘˜ + 1) )= ๐‘˜ ๐‘˜! ๐‘›(๐‘› + 1) โ‹ฏ (๐‘› + ๐‘˜ โˆ’ 1) ๐‘˜! ๐‘› = (โˆ’1)๐‘˜ (( )) ๐‘˜ = (โˆ’1)๐‘˜

by Theorem 1.3.4. This kind of situation where evaluation of an enumerative formula at negative arguments yields, up to sign, another enumerative function is called combinatorial reciprocity and will be studied in Section 3.9. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

10

1. Basic Counting

1.4. Set partitions We have already seen that disjoint unions are nice combinatorially. So it should come as no surprise that set partitions also play an important role. A partition of a set ๐‘‡ is a set ๐œŒ of nonempty subsets ๐ต1 , . . . , ๐ต๐‘˜ such that ๐‘‡ = โจ„๐‘– ๐ต๐‘– , written ๐œŒ โŠข ๐‘‡. The ๐ต๐‘– are called blocks and we use the notation ๐œŒ = ๐ต1 / . . . /๐ต๐‘˜ leaving out all curly brackets and commas, even though the elements of the blocks, as well as the blocks themselves, are unordered. For example, one set partition of ๐‘‡ = {๐‘Ž, ๐‘, ๐‘, ๐‘‘, ๐‘’, ๐‘“, ๐‘”} is ๐œŒ = ๐‘Ž๐‘๐‘“/๐‘๐‘’/๐‘‘/๐‘” = ๐‘‘/๐‘’๐‘/๐‘”/๐‘๐‘“๐‘Ž. We let ๐ต(๐‘‡) be the set of all ๐œŒ โŠข ๐‘‡. To illustrate, ๐ต({๐‘Ž, ๐‘, ๐‘}) = {๐‘Ž/๐‘/๐‘, ๐‘Ž๐‘/๐‘, ๐‘Ž๐‘/๐‘, ๐‘Ž/๐‘๐‘, ๐‘Ž๐‘๐‘}. The ๐‘›th Bell number is ๐ต(๐‘›) = #๐ต([๐‘›]). Although there is no known expression for ๐ต(๐‘›) as a simple product, there is a recursion. Theorem 1.4.1. The Bell numbers satisfy the initial condition ๐ต(0) = 1 and the recurrence relation ๐‘›โˆ’1 ๐ต(๐‘›) = โˆ‘ ( )๐ต(๐‘› โˆ’ ๐‘˜) ๐‘˜โˆ’1 ๐‘˜ for ๐‘› โ‰ฅ 1. Proof. The initial condition counts the empty partition of โˆ…. For the recursion, given ๐œŒ โˆˆ ๐ต([๐‘›]), let ๐‘˜ be the number of elements in the block ๐ต containing ๐‘›. Then there are (๐‘›โˆ’1 ) ways to pick the remaining ๐‘˜ โˆ’ 1 elements of [๐‘› โˆ’ 1] to be in ๐ต. And the ๐‘˜โˆ’1 number of ways to partition [๐‘›] โˆ’ ๐ต is ๐ต(๐‘› โˆ’ ๐‘˜). Summing over all possible ๐‘˜ finishes the proof. โ–ก We may sometimes want to keep track of the number of blocks in our partitions. So define ๐‘†(๐‘‡, ๐‘˜) to be the set of all ๐œŒ โŠข ๐‘‡ with ๐‘˜ blocks. The Stirling numbers of the second kind are ๐‘†(๐‘›, ๐‘˜) = #๐‘†([๐‘›], ๐‘˜). We will introduce Stirling numbers of the first kind in the next section. For example ๐‘†({๐‘Ž, ๐‘, ๐‘}, 2) = {๐‘Ž๐‘/๐‘, ๐‘Ž๐‘/๐‘, ๐‘Ž/๐‘๐‘} so ๐‘†(3, 2) = 3. Just as with the binomial coefficients, the ๐‘†(๐‘›, ๐‘˜) for 1 โ‰ค ๐‘˜ โ‰ค ๐‘› can be displayed in a triangle as in Figure 1.3. And like the binomial coefficients, these Stirling numbers satisfy a simple recurrence relation. 1 1 1 1 1

1 3

7 15

1 6

25

1 10

1

Figure 1.3. Rows 1 through 5 of Stirlingโ€™s second triangle

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

1.5. Permutations by cycle structure

11

Theorem 1.4.2. The Stirling numbers of the second kind satisfy the initial condition ๐‘†(0, ๐‘˜) = ๐›ฟ ๐‘˜,0 and recurrence relation ๐‘†(๐‘›, ๐‘˜) = ๐‘†(๐‘› โˆ’ 1, ๐‘˜ โˆ’ 1) + ๐‘˜๐‘†(๐‘› โˆ’ 1, ๐‘˜) for ๐‘› โ‰ฅ 1. Proof. By now, the reader should be able to explain the initial condition without difficulty. For the recursion, the elements ๐œŒ โˆˆ ๐‘†([๐‘›], ๐‘˜) are of two flavors: those where ๐‘› is in a block by itself and those where ๐‘› is in a block with other elements. Removing ๐‘› in the first case leaves a partition in ๐‘†([๐‘› โˆ’ 1], ๐‘˜ โˆ’ 1) and this is a bijection. This accounts for the summand ๐‘†(๐‘›โˆ’1, ๐‘˜โˆ’1). Removing ๐‘› in the second case leaves ๐œŽ โˆˆ ๐‘†([๐‘›โˆ’1], ๐‘˜), but this map is not a bijection. In particular, given ๐œŽ, one can insert ๐‘› into any one of its ๐‘˜ blocks to recover an element of ๐‘†([๐‘›], ๐‘˜). So the total count is ๐‘˜๐‘†(๐‘› โˆ’ 1, ๐‘˜) for this case. โ–ก

1.5. Permutations by cycle structure The ordered analogue of a decomposition of a set into a partition is the decomposition of a permutation of [๐‘›] into cycles. These are counted by the Stirling numbers of the first kind. The symmetric group is ๐”–๐‘› = ๐‘ƒ([๐‘›]). As the name implies, ๐”–๐‘› has a group structure defined as follows. If ๐œ‹ = ๐œ‹1 . . . ๐œ‹๐‘› โˆˆ ๐”–๐‘› , then we can view this permutation as a bijection ๐œ‹ โˆถ [๐‘›] โ†’ [๐‘›] where ๐œ‹(๐‘–) = ๐œ‹๐‘– . From this it follows that ๐”–๐‘› is a group where the operation is composition of functions. Given ๐œ‹ โˆˆ ๐”–๐‘› and ๐‘– โˆˆ [๐‘›], there is a smallest exponent โ„“ โ‰ฅ 1 such that ๐œ‹โ„“ (๐‘–) = ๐‘–. This and various other claims below will be proved using digraphs in Section 1.9. In this case, the elements ๐‘–, ๐œ‹(๐‘–), ๐œ‹2 (๐‘–), . . . , ๐œ‹โ„“โˆ’1 (๐‘–) are all distinct and we write ๐‘ = (๐‘–, ๐œ‹(๐‘–), ๐œ‹2 (๐‘–), . . . , ๐œ‹โ„“โˆ’1 (๐‘–)) and call this a cycle of length โ„“ or simply an โ„“-cycle of ๐œ‹. Cycles of length one are called fixed points. As an example, if ๐œ‹ = 6514237 and ๐‘– = 1, then we have ๐œ‹(1) = 6, ๐œ‹2 (1) = 3, ๐œ‹3 (1) = 1 so that ๐‘ = (1, 6, 3) is a cycle of ๐œ‹. We now iterate this process: if there is some ๐‘— โˆˆ [๐‘›] which is not in any of the cycles computed so far, we find the cycle containing ๐‘— and continue until every element is in a cycle. The cycle decomposition of ๐œ‹ is ๐œ‹ = ๐‘ 1 . . . ๐‘ ๐‘˜ where the ๐‘๐‘— are the cycles found in this process. Continuing our example, we could get ๐œ‹ = (1, 6, 3)(2, 5)(4)(7). To distinguish the cycle decomposition of ๐œ‹ from its description as ๐œ‹ = ๐œ‹1 . . . ๐œ‹๐‘› we will call the latter the one-line notation for ๐œ‹. This is also distinct from two-line notation, which is where one writes (1.7)

๐œ‹=

1 ๐œ‹1

2 ๐œ‹2

... ๐‘› . . . . ๐œ‹๐‘›

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

12

1. Basic Counting

1 1 2 6 24

1 3

11 50

1 6

35

1 10

1

Figure 1.4. Rows 1 through 5 of Stirlingโ€™s first triangle

Note that an โ„“-cycle can be written in โ„“ different ways depending on which of its elements one starts with; for example (1, 6, 3) = (6, 3, 1) = (3, 1, 6). Furthermore, the distinct cycles of ๐œ‹ are disjoint. So if we think of the cycle ๐‘ as the permutation of [๐‘›] which agrees with ๐œ‹ on the elements of ๐‘ and has all other elements as fixed points, then the cycles of ๐œ‹ = ๐‘ 1 . . . ๐‘ ๐‘˜ commute where we consider the product as a composition of permutations. Returning to our running example, we could write ๐œ‹ = (1, 6, 3)(2, 5)(4)(7) = (4)(1, 6, 3)(7)(2, 5) = (5, 2)(3, 1, 6)(7)(4). As mentioned above, we defer the proof of the following result until Section 1.9. Theorem 1.5.1. Every ๐œ‹ โˆˆ ๐”–๐‘› has a cycle decomposition ๐œ‹ = ๐‘ 1 . . . ๐‘ ๐‘˜ which is unique up to the order of the factors and cyclic reordering of the elements within each ๐‘ ๐‘– . We are now in a position to proceed parallel to the development of set partitions with a given number of blocks in the previous section. For ๐‘› โ‰ฅ 0 we denote by ๐‘([๐‘›], ๐‘˜) the set of all permutations in ๐”–๐‘› which have ๐‘˜ cycles in their decomposition. Note the difference between โ€œ๐‘˜ cyclesโ€ referring to the number of cycles and โ€œ๐‘˜-cyclesโ€ referring to the length of the cycles. The signless Stirling numbers of the first kind are ๐‘(๐‘›, ๐‘˜) = #๐‘([๐‘›], ๐‘˜). So, analogous to what we have seen before, ๐‘(๐‘›, ๐‘˜) = 0 for ๐‘˜ < 0 or ๐‘˜ > ๐‘›. To illustrate the notation, ๐‘([4], 1) = {(1, 2, 3, 4), (1, 2, 4, 3), (1, 3, 2, 4), (1, 3, 4, 2), (1, 4, 2, 3), (1, 4, 3, 2)} so ๐‘(4, 1) = 6. In general, as you will be asked to prove in an exercise, ๐‘([๐‘›], 1) = (๐‘›โˆ’1)!. Part of Stirlingโ€™s first triangle is displayed in Figure 1.4. We also have a recursion. Theorem 1.5.2. The signless Stirling numbers of the first kind satisfy the initial condition ๐‘(0, ๐‘˜) = ๐›ฟ ๐‘˜,0 and recurrence relation ๐‘(๐‘›, ๐‘˜) = ๐‘(๐‘› โˆ’ 1, ๐‘˜ โˆ’ 1) + (๐‘› โˆ’ 1)๐‘(๐‘› โˆ’ 1, ๐‘˜) for ๐‘› โ‰ฅ 1. Proof. As usual, we concentrate on the recurrence. Given ๐œ‹ โˆˆ ๐‘([๐‘›], ๐‘˜), we can remove ๐‘› from its cycle. If ๐‘› was a fixed point, then the resulting permutations are counted by ๐‘(๐‘› โˆ’ 1, ๐‘˜ โˆ’ 1). If ๐‘› was in a cycle of length at least two, then the permutations obtained upon removal are in ๐‘([๐‘› โˆ’ 1], ๐‘˜). So one must find the number of Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

1.6. Integer partitions

13

ways to insert ๐‘› into a cycle of some ๐œŽ โˆˆ ๐‘([๐‘› โˆ’ 1], ๐‘˜). There are โ„“ places to insert ๐‘› in a cycle of length โ„“. So the total number of insertion spots is the sum of the cycle lengths of ๐œŽ, which is ๐‘› โˆ’ 1. โ–ก The reader may have guessed that there are also (signed) Stirling numbers of the first kind defined by ๐‘ (๐‘›, ๐‘˜) = (โˆ’1)๐‘›โˆ’๐‘˜ ๐‘(๐‘›, ๐‘˜). It is not immediately apparent why one would want to attach signs to these constants. We will see one reason in Chapter 5 where it will be shown that the ๐‘ (๐‘›, ๐‘˜) are the Whitney numbers of the first kind for the lattice of set partitions ordered by refinement. Here we will content ourselves with proving an analogue of part (d) of Theorem 1.3.3. Corollary 1.5.3. For ๐‘› โ‰ฅ 0 we have โˆ‘ ๐‘ (๐‘›, ๐‘˜) = { ๐‘˜

1 if ๐‘› = 0 or 1, 0 if ๐‘› โ‰ฅ 2.

Proof. The cases when ๐‘› = 0 or 1 are easy to verify, so assume ๐‘› โ‰ฅ 2. Since ๐‘ (๐‘›, ๐‘˜) = (โˆ’1)๐‘›โˆ’๐‘˜ ๐‘(๐‘›, ๐‘˜) and (โˆ’1)๐‘› is constant throughout the summation, it suffices to show that โˆ‘๐‘˜ (โˆ’1)๐‘˜ ๐‘(๐‘›, ๐‘˜) = 0. Using Theorem 1.5.2 and induction on ๐‘› we obtain โˆ‘(โˆ’1)๐‘˜ ๐‘(๐‘›, ๐‘˜) = โˆ‘(โˆ’1)๐‘˜ ๐‘(๐‘› โˆ’ 1, ๐‘˜ โˆ’ 1) + โˆ‘(โˆ’1)๐‘˜ (๐‘› โˆ’ 1)๐‘(๐‘› โˆ’ 1, ๐‘˜) ๐‘˜

๐‘˜

๐‘˜

= โˆ’ โˆ‘(โˆ’1)

๐‘˜โˆ’1

๐‘(๐‘› โˆ’ 1, ๐‘˜ โˆ’ 1) + (๐‘› โˆ’ 1) โˆ‘(โˆ’1)๐‘˜ ๐‘(๐‘› โˆ’ 1, ๐‘˜)

๐‘˜

๐‘˜

= โˆ’0 + (๐‘› โˆ’ 1)0 =0 as desired.

โ–ก

Note the usefulness of considering the sums in the preceding proof as over ๐‘˜ โˆˆ โ„ค rather than 0 โ‰ค ๐‘˜ โ‰ค ๐‘›. This does away with having to consider any special cases at the values ๐‘˜ = 0 or ๐‘˜ = ๐‘›.

1.6. Integer partitions Just as one can partition a set into blocks, one can partition a nonnegative integer as a sum. Integer partitions play an important role not just in combinatorics but also in number theory and the representation theory of the symmetric group. See the appendix at the end of the book for more information on the latter. An integer partition of ๐‘› โ‰ฅ 0 is a multiset ๐œ† of positive integers such that the sum of the elements of ๐œ†, denoted |๐œ†|, is ๐‘›. We also write ๐œ† โŠข ๐‘›. These elements are called the parts. Since the parts of ๐œ† are unordered, we will always list them in a canonical order ๐œ† = (๐œ†1 , . . . , ๐œ†๐‘˜ ) which is weakly decreasing. We let ๐‘ƒ(๐‘›) denote the set of all Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

14

1. Basic Counting

partitions of ๐‘› and ๐‘(๐‘›) = #๐‘ƒ(๐‘›). For example, ๐‘ƒ(4) = {(1, 1, 1, 1), (2, 1, 1), (2, 2), (3, 1), (4)} so that ๐‘(4) = 5. Note the distinction between ๐‘ƒ([๐‘›]), which is a set of set partitions, and ๐‘ƒ(๐‘›), which is a set of integer partitions. Sometimes we will just say โ€œpartitionโ€ if the context makes it clear whether we are partitioning sets or integers. We will use multiplicity notation for integer partitions just as we would for any multiset, writing ๐œ† = (1๐‘š1 , 2๐‘š2 , . . . , ๐‘›๐‘š๐‘› ) where ๐‘š๐‘– is the multiplicity of ๐‘– in ๐œ†. There is no known product formula for ๐‘(๐‘›). In fact, there is not even a simple recurrence relation. One can use generating functions to derive results about these numbers, but that must wait until Chapter 3. Here we will just introduce a useful geometric device for studying ๐‘(๐‘›). The Ferrers or Young diagram of ๐œ† = (๐œ†1 , . . . , ๐œ†๐‘˜ ) โŠข ๐‘› is an array of ๐‘› boxes into left-justified rows such that row ๐‘– contains ๐œ†๐‘– boxes. Dots are also sometimes used in place of boxes and in this case some authors use โ€œFerrers diagramโ€ for the dot variant and โ€œYoung diagramโ€ for the corresponding array of boxes. We often make no distinction between a partition and its Young diagram. The Young diagram of ๐œ† = (5, 5, 2, 1) is shown in Figure 1.5. We should warn the reader that we are writing our Young diagrams in English notation where the rows are numbered from 1 to ๐‘˜ from the top down as in a matrix. Some authors prefer French notation where the rows are numbered from bottom to top as in a Cartesian coordinate system. The conjugate or transpose of ๐œ† is the partition ๐œ†๐‘ก whose Young diagram is obtained by reflecting the diagram of ๐œ† about its main diagonal. This is done in Figure 1.5, showing that (5, 5, 2, 1)๐‘ก = (4, 3, 2, 2, 2). There is also another way to express the parts of the conjugate.

๐œ† = (5, 5, 2, 1) =

= โ€ข โ€ข

โ€ข

โ€ข

โ€ข

โ€ข

โ€ข

โ€ข

โ€ข

โ€ข

โ€ข

โ€ข

โ€ข ๐œ†๐‘ก =

Figure 1.5. A partition, its Young diagram, and its conjugate

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

1.6. Integer partitions

15

Proposition 1.6.1. If ๐œ† = (๐œ†1 , . . . , ๐œ†๐‘˜ ) is a partition and ๐œ†๐‘ก = (๐œ†๐‘ก1 , . . . , ๐œ†๐‘ก๐‘™ ), then, for 1 โ‰ค ๐‘— โ‰ค ๐‘™, ๐œ†๐‘—๐‘ก = #{๐‘– โˆฃ ๐œ†๐‘– โ‰ฅ ๐‘—}. Proof. By definition, ๐œ†๐‘—๐‘ก is the length of the ๐‘—th column of ๐œ†. But that column contains a box in row ๐‘– if and only if ๐œ†๐‘– โ‰ฅ ๐‘—. โ–ก The number of parts of a partition ๐œ† is called its length and is denoted โ„“(๐œ†). At this point the reader is probably expecting a discussion of those partitions of ๐‘› with โ„“(๐œ†) = ๐‘˜. As it turns out, it is a bit simpler to consider ๐‘ƒ(๐‘›, ๐‘˜), the set of all partitions ๐œ† of ๐‘› with โ„“(๐œ†) โ‰ค ๐‘˜, and ๐‘(๐‘›, ๐‘˜) = #๐‘ƒ(๐‘›, ๐‘˜). Note that the number of ๐œ† โŠข ๐‘› with โ„“(๐œ†) = ๐‘˜ is just ๐‘(๐‘›, ๐‘˜) โˆ’ ๐‘(๐‘›, ๐‘˜ โˆ’ 1). So in some sense the two viewpoints are equivalent. But it will be easier to state our results in terms of ๐‘(๐‘›, ๐‘˜). Note also that ๐‘(๐‘›, 0) โ‰ค ๐‘(๐‘›, 1) โ‰ค โ‹ฏ โ‰ค ๐‘(๐‘›, ๐‘›) = ๐‘(๐‘›, ๐‘› + 1) = โ‹ฏ = ๐‘(๐‘›). Because of this behavior, it is best to display the ๐‘(๐‘›, ๐‘˜) in a matrix, rather than a triangle, keeping in mind that the entries in the ๐‘›th row eventually stabilize to an infinite repetition of the constant ๐‘(๐‘›). Part of this array will be found in Figure 1.6. We also assume that ๐‘(๐‘›, ๐‘˜) = 0 if ๐‘› < 0 or ๐‘˜ < 0. Unlike ๐‘(๐‘›), one can write down a simple recurrence relation for ๐‘(๐‘›, ๐‘˜). Theorem 1.6.2. The ๐‘(๐‘›, ๐‘˜) satisfy ๐‘(0, ๐‘˜) = {

0 if ๐‘˜ < 0, 1 if ๐‘˜ โ‰ฅ 0

and ๐‘(๐‘›, ๐‘˜) = ๐‘(๐‘› โˆ’ ๐‘˜, ๐‘˜) + ๐‘(๐‘›, ๐‘˜ โˆ’ 1) for ๐‘› โ‰ฅ 1 Proof. We skip directly to the recursion. Note that since conjugation is a bijection, ๐‘(๐‘›, ๐‘˜) also counts the partitions ๐œ† = (๐œ†1 , . . . , ๐œ†๐‘™ ) โŠข ๐‘› such that ๐œ†1 โ‰ค ๐‘˜. It will be convenient to use this interpretation of ๐‘(๐‘›, ๐‘˜) for the proof. We have two possible cases. If ๐œ†1 = ๐‘˜, then ๐œ‡ = (๐œ†2 , . . . , ๐œ†๐‘™ ) โŠข ๐‘› โˆ’ ๐‘˜ and ๐œ†2 โ‰ค ๐œ†1 = ๐‘˜. So these partitions are counted by ๐‘(๐‘› โˆ’ ๐‘˜, ๐‘˜). The other possibility is that ๐œ†1 โ‰ค ๐‘˜ โˆ’ 1. And these ๐œ† are taken care of by the ๐‘(๐‘›, ๐‘˜ โˆ’ 1) term. โ–ก

0 1 2 3 4

0 1 0 0 0 0

1 1 1 1 1 1

2 1 1 2 2 3

3 1 1 2 3 4

4 1 1 2 3 5

5 1 1 2 3 5

Figure 1.6. The values ๐‘(๐‘›, ๐‘˜) for 0 โ‰ค ๐‘› โ‰ค 4 and 0 โ‰ค ๐‘˜ โ‰ค 5

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

16

1. Basic Counting

1.7. Compositions Recall that integer partitions are really unordered even though we usually list them in weakly decreasing fashion. This raises the question about what happens if we considered ways to write ๐‘› as a sum when the summands are ordered. This is the notion of a composition. A composition of ๐‘› is a sequence ๐›ผ = [๐›ผ1 , . . . , ๐›ผ๐‘˜ ] of positive integers called parts such that โˆ‘๐‘– ๐›ผ๐‘– = ๐‘›. We write ๐›ผ โŠง ๐‘› and use square brackets to distinguish compositions from integer partitions. This causes a notational conflict between [๐‘›] as a composition of ๐‘› and as the integers from 1 to ๐‘›, but the context should make it clear which interpretation is meant. Let ๐‘„(๐‘›) be the set of compositions of ๐‘› and ๐‘ž(๐‘›) = #๐‘„(๐‘›). So the compositions of 4 are ๐‘„(4) = {[1, 1, 1, 1], [2, 1, 1], [1, 2, 1], [1, 1, 2], [2, 2], [3, 1], [1, 3], [4]}. So ๐‘ž(4) = 8, which is a power of 2. This, as your author is fond of saying, is not a coincidence. Theorem 1.7.1. For ๐‘› โ‰ฅ 1 we have ๐‘ž(๐‘›) = 2๐‘›โˆ’1 . Proof. There is a famous bijection ๐œ™ โˆถ 2[๐‘›โˆ’1] โ†’ ๐‘„(๐‘›), which we will use to prove this result. This map will be useful when working with quasisymmetric functions in Chapter 8. Given ๐‘† = {๐‘ 1 , . . . , ๐‘ ๐‘˜ } โŠ† [๐‘› โˆ’ 1] written in increasing order, we define (1.8)

๐œ™(๐‘†) = [๐‘ 1 โˆ’ ๐‘ 0 , ๐‘ 2 โˆ’ ๐‘ 1 , . . . , ๐‘ ๐‘˜ โˆ’ ๐‘ ๐‘˜โˆ’1 , ๐‘ ๐‘˜+1 โˆ’ ๐‘ ๐‘˜ ]

where, by definition, ๐‘ 0 = 0 and ๐‘ ๐‘˜+1 = ๐‘›. To show that ๐œ™ is well-defined, suppose ๐œ™(๐‘†) = [๐›ผ1 , . . . , ๐›ผ๐‘˜+1 ]. Since ๐‘† is increasing, ๐›ผ๐‘– = ๐‘ ๐‘– โˆ’ ๐‘ ๐‘–โˆ’1 is a positive integer. Furthermore ๐‘˜+1

๐‘˜+1

โˆ‘ ๐›ผ๐‘– = โˆ‘ (๐‘ ๐‘– โˆ’ ๐‘ ๐‘–โˆ’1 ) = ๐‘ ๐‘˜+1 โˆ’ ๐‘ 0 = ๐‘›. ๐‘–=1

๐‘–=1

Thus ๐œ™(๐‘†) โˆˆ ๐‘„(๐‘›) as desired. To show that ๐œ™ is bijective, we construct its inverse ๐œ™โˆ’1 โˆถ ๐‘„(๐‘›) โ†’ 2[๐‘›โˆ’1] . Given ๐›ผ = [๐›ผ1 , . . . , ๐›ผ๐‘˜+1 ] โˆˆ ๐‘„(๐‘›), we let ๐œ™โˆ’1 (๐›ผ) = {๐›ผ1 , ๐›ผ1 + ๐›ผ2 , ๐›ผ1 + ๐›ผ2 + ๐›ผ3 , . . . , ๐›ผ1 + ๐›ผ2 + โ‹ฏ + ๐›ผ๐‘˜ }. It should not be hard for the reader to prove that ๐œ™โˆ’1 is well-defined and the inverse of ๐œ™. โ–ก As usual, we wish to make a more refined count by restricting the number of constituents of the object under consideration. Let ๐‘„(๐‘›, ๐‘˜) be the set of all compositions of ๐‘› with exactly ๐‘˜ parts and let ๐‘ž(๐‘›, ๐‘˜) = #๐‘„(๐‘›, ๐‘˜). Since the ๐‘ž(๐‘›, ๐‘˜) will turn out to be previously studied constants, we will forgo the usual triangle. The result below follows easily by restricting the function ๐œ™ from the previous proof, so the demonstration is omitted. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

1.8. The twelvefold way

17

Theorem 1.7.2. The composition numbers satisfy ๐‘ž(0, ๐‘˜) = ๐›ฟ ๐‘˜,0 and ๐‘›โˆ’1 ๐‘ž(๐‘›, ๐‘˜) = ( ) ๐‘˜โˆ’1 โ–ก

for ๐‘› โ‰ฅ 1.

1.8. The twelvefold way We now have all the tools in place to count certain functions. There are 12 types of such functions and so this scheme is called the twelvefold way, an idea which was introduced in a series of lectures by Gian-Carlo Rota. The name was suggested by Joel Spencer and should not be confused with the twelvefold path of Buddhism! We will consider three types of functions ๐‘“ โˆถ ๐ท โ†’ ๐‘…, namely, arbitrary functions, injections, and surjections. We will also permit the domain ๐ท and range ๐‘… to be of two types each: either distinguishable, which means it is a set, or indistinguishable, which means it is a multiset consisting of a single element repeated some number of times. Thus the total number of types of functions under consideration is the product of the number of choices for ๐‘“, ๐ท, and ๐‘… or 3 โ‹… 2 โ‹… 2 = 12. Of course, a function where the domain or range is a multiset is not really well-defined, even though the intuitive notion should be clear. To be precise, when ๐ท is a multiset and ๐‘… is a set, suppose ๐ทโ€ฒ is a set with |๐ทโ€ฒ | = |๐ท|. Then a function ๐‘“ โˆถ ๐ท โ†’ ๐‘… is an equivalence class of functions ๐‘“ โˆถ ๐ทโ€ฒ โ†’ ๐‘… where ๐‘“ and ๐‘” are equivalent if #๐‘“โˆ’1 (๐‘Ÿ) = #๐‘”โˆ’1 (๐‘Ÿ) for all ๐‘Ÿ โˆˆ ๐‘…. The reader can come up with the corresponding notions for the other cases if desired. We will assume throughout that |๐ท| = ๐‘› and |๐‘…| = ๐‘˜ are both nonnegative integers. We will collect the results in the chart in Table 1.1. We first deal with the case where both ๐ท and ๐‘… are distinguishable. Without loss of generality, we can assume that ๐ท = [๐‘›]. So a function ๐‘“ โˆถ ๐ท โ†’ ๐‘… can be considered as a word ๐‘ค = ๐‘“(1)๐‘“(2) . . . ๐‘“(๐‘›). Since there are ๐‘˜ choices for each ๐‘“(๐‘–), we have, by Theorem 1.2.2, that the number of such ๐‘“ is #๐‘ƒ(([๐‘˜], ๐‘›)) = ๐‘˜๐‘› . If ๐‘“ is injective, then ๐‘ค becomes a permutation, giving the count #๐‘ƒ([๐‘˜], ๐‘›) = ๐‘˜โ†“๐‘› from Theorem 1.2.1. For surjective functions, we need a new concept. If ๐ท is a set, then the kernel of a function ๐‘“ โˆถ ๐ท โ†’ ๐‘… is the partition ker ๐‘“ of ๐ท whose blocks are the nonempty subsets of the form ๐‘“โˆ’1 (๐‘Ÿ) for ๐‘Ÿ โˆˆ ๐‘…. For example, if ๐‘“ โˆถ {๐‘Ž, ๐‘, ๐‘, ๐‘‘} โ†’ {1, 2, 3} is given by ๐‘“(๐‘Ž) = ๐‘“(๐‘) = 2, Table 1.1. The twelvefold way

๐ท

๐‘…

arbitrary ๐‘“

injective ๐‘“

surjective ๐‘“

dist.

dist.

๐‘˜๐‘›

๐‘˜โ†“๐‘›

๐‘˜! ๐‘†(๐‘›, ๐‘˜)

indist.

dist.

(๐‘›+๐‘˜โˆ’1 ) ๐‘›

(๐‘›๐‘˜)

(๐‘›โˆ’1 ) ๐‘˜โˆ’1

dist.

indist.

โˆ‘๐‘—=0 ๐‘†(๐‘›, ๐‘—)

๐›ฟ(๐‘› โ‰ค ๐‘˜)

๐‘†(๐‘›, ๐‘˜)

indist.

indist.

๐‘(๐‘›, ๐‘˜)

๐›ฟ(๐‘› โ‰ค ๐‘˜)

๐‘(๐‘›, ๐‘˜) โˆ’ ๐‘(๐‘›, ๐‘˜ โˆ’ 1)

๐‘˜

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

18

1. Basic Counting

๐‘“(๐‘) = 3, and ๐‘“(๐‘‘) = 1, then ker ๐‘“ = ๐‘Ž๐‘/๐‘/๐‘‘. If ๐‘“ is to be surjective, then the function can be specified by picking a partition of ๐ท for ker ๐‘“ and then picking a bijection ๐‘” from the blocks of ker ๐‘“ into ๐‘…. Continuing our example, ๐‘“ is completely determined by its kernel and the bijection ๐‘”(๐‘Ž๐‘) = 2, ๐‘”(๐‘) = 3, and ๐‘”(๐‘‘) = 1. The number of ways to choose ker ๐‘“ = ๐ต1 / . . . /๐ต๐‘˜ is ๐‘†(๐‘›, ๐‘˜) by definition. And, using the injective case with ๐‘› = ๐‘˜, the number of bijections ๐‘” โˆถ {๐ต1 , . . . , ๐ต๐‘˜ } โ†’ ๐‘… is ๐‘˜ โ†“๐‘˜ = ๐‘˜!. So the total count is ๐‘˜! ๐‘†(๐‘›, ๐‘˜). Now suppose ๐ท is indistinguishable and ๐‘… is distinguishable where we assume ๐‘… = [๐‘˜]. Then one can think of ๐‘“ โˆถ ๐ท โ†’ ๐‘… as a multiset ๐‘€ = {{1๐‘š1 , . . . , ๐‘˜๐‘š๐‘˜ }} on ๐‘… where ๐‘š๐‘– = #๐‘“โˆ’1 (๐‘–). It follows that โˆ‘๐‘– ๐‘š๐‘– = #๐ท = ๐‘›. So, by Theorem 1.3.4, the number of all such ๐‘“ is ๐‘˜ ๐‘›+๐‘˜โˆ’1 (( )) = ( ). ๐‘› ๐‘› If ๐‘“ is to be injective, then we are picking an ๐‘›-element subset of ๐‘… = [๐‘˜] giving a count of (๐‘›๐‘˜). If ๐‘“ is to be surjective, then ๐‘š๐‘– โ‰ฅ 1 for all ๐‘– so that [๐‘š1 , . . . , ๐‘š๐‘˜ ] is a composition of ๐‘›. It follows from Theorem 1.7.2 that the number of functions is ๐‘ž(๐‘›, ๐‘˜) = (๐‘›โˆ’1 ). ๐‘˜โˆ’1 To deal with the case when ๐ท = [๐‘›] is distinguishable and ๐‘… is indistinguishable, we introduce a useful extension of the Kronecker delta. If ๐‘† is any statement, we let (1.9)

๐›ฟ(๐‘†) = {

1 if ๐‘† is true, 0 if ๐‘† is false.

Returning to our counting, ๐‘“ is completely determined by its kernel, which is a partition of [๐‘›]. If we are considering all ๐‘“, then the kernel can have any number of blocks up to and including ๐‘˜. Summing the corresponding Stirling numbers gives the corresponding entry in Table 1.1. If ๐‘“ is injective, then for such a function to exist we must have ๐‘› โ‰ค ๐‘˜. And in that case there is only one possible kernel, namely the partition into singleton blocks. This count can be summarized as ๐›ฟ(๐‘› โ‰ค ๐‘˜). For surjective ๐‘“ we are partitioning [๐‘›] into exactly ๐‘˜ blocks, giving ๐‘†(๐‘›, ๐‘˜) possibilities. If ๐ท and ๐‘… are both indistinguishable, then the nonzero numbers of the form ๐‘š๐‘– = #๐‘“โˆ’1 (๐‘Ÿ) for ๐‘Ÿ โˆˆ ๐‘… completely determine ๐‘“. And these numbers form a partition of ๐‘› = #๐ท into at most ๐‘˜ = #๐‘… parts. Recalling the notation of Section 1.6, the total number of such ๐‘“ is ๐‘(๐‘›, ๐‘˜). The line of reasoning for injective functions follows that of the previous paragraph with the same resulting answer. Finally, for surjectivity we need exactly ๐‘˜ parts, which is counted by ๐‘(๐‘›, ๐‘˜) โˆ’ ๐‘(๐‘›, ๐‘˜ โˆ’ 1).

1.9. Graphs and digraphs Graph theory is a substantial part of combinatorics. We will use directed graphs to give the postponed proof of the existence and uniqueness of the cycle decomposition of permutations in ๐”–๐‘› . A labeled graph ๐บ = (๐‘‰, ๐ธ) consists of a set ๐‘‰ of elements called vertices and a set ๐ธ of elements called edges where an edge consists of an unordered pair of vertices. We will write ๐‘‰(๐บ) and ๐ธ(๐บ) for the vertex and edge set of ๐บ, respectively, if we wish to emphasize the graph involved. Geometrically, we think of the vertices as nodes and the edges as line segments or curves joining them. Conventionally, in graph theory an Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

1.9. Graphs and digraphs

19

๐‘ฃ

๐‘ค

๐‘ฆ

๐‘ฅ

Figure 1.7. A graph ๐บ

edge connecting vertices ๐‘ฃ and ๐‘ค is written ๐‘’ = ๐‘ฃ๐‘ค rather than ๐‘’ = {๐‘ฃ, ๐‘ค}. In this case we say that ๐‘’ contains ๐‘ฃ and ๐‘ค, or that ๐‘’ has endpoints ๐‘ฃ and ๐‘ค. We also say that ๐‘ฃ and ๐‘ค are neighbors. For example, a drawing of the graph ๐บ with vertices ๐‘‰ = {๐‘ฃ, ๐‘ค, ๐‘ฅ, ๐‘ฆ} and edges ๐ธ = {๐‘ฃ๐‘ค, ๐‘ฃ๐‘ฅ, ๐‘ฃ๐‘ฆ, ๐‘ค๐‘ฅ, ๐‘ฅ๐‘ฆ} is displayed in Figure 1.7. If #๐‘‰ = 1, then there is only one graph with vertex set ๐‘‰ and such a graph is called trivial. Call graph ๐ป a subgraph of ๐บ, written ๐ป โŠ† ๐บ, if ๐‘‰(๐ป) โŠ† ๐‘‰(๐บ) and ๐ธ(๐ป) โŠ† ๐ธ(๐บ). In this case we also say that ๐บ contains ๐ป. There are several types of subgraphs which will play an important role in what follows. A walk of length โ„“ in ๐บ is a sequence of vertices ๐‘Š โˆถ ๐‘ฃ 0 , ๐‘ฃ 1 , . . . , ๐‘ฃ โ„“ such that ๐‘ฃ ๐‘–โˆ’1 ๐‘ฃ ๐‘– โˆˆ ๐ธ for 1 โ‰ค ๐‘– โ‰ค โ„“. We say that the walk is from ๐‘ฃ 0 to ๐‘ฃ โ„“ , or is a ๐‘ฃ 0 โ€“๐‘ฃ โ„“ walk, or that ๐‘ฃ 0 , ๐‘ฃ โ„“ are the endpoints of ๐‘Š. We call ๐‘Š a path if all the vertices are distinct and we usually use letters like ๐‘ƒ for paths. In particular, we will use ๐‘Š๐‘› or ๐‘ƒ๐‘› to denote a walk or a path having ๐‘› vertices, respectively. In our example graph, ๐‘ƒ โˆถ ๐‘ฆ, ๐‘ฃ, ๐‘ฅ, ๐‘ค is a path of length 3 from ๐‘ฆ to ๐‘ค. Notice that length refers to the number of edges in the path, which is one less than the number of vertices. A cycle of length โ„“ in ๐บ is a sequence of distinct vertices ๐ถ โˆถ ๐‘ฃ 1 , ๐‘ฃ 2 , . . . , ๐‘ฃ โ„“ such that we have distinct edges ๐‘ฃ ๐‘–โˆ’1 ๐‘ฃ ๐‘– for 1 โ‰ค ๐‘– โ‰ค โ„“, and subscripts are taken modulo โ„“ so that ๐‘ฃ 0 = ๐‘ฃ โ„“ . Returning to our running example, ๐ถ โˆถ ๐‘ฃ, ๐‘ฅ, ๐‘ฆ is a cycle in ๐บ of length 3. In a cycle the length is both the number of vertices and the number of edges. The notation ๐ถ๐‘› will be used for a cycle with ๐‘› vertices and we will call this an ๐‘›-cycle. We also denote by ๐พ๐‘› the complete graph which consists of ๐‘› vertices and all possible (๐‘›2) edges between them. A copy of a complete graph in a graph ๐บ is often called a clique. There is a close relationship between some of the parts of a graph which we have just defined. Lemma 1.9.1. Let ๐บ be a graph and let ๐‘ข, ๐‘ฃ โˆˆ ๐‘‰. (a) Any walk from ๐‘ข to ๐‘ฃ contains a path from ๐‘ข to ๐‘ฃ. (b) The union of any two different paths from ๐‘ข to ๐‘ฃ contains a cycle. Proof. We will prove (a) and leave (b) as an exercise. Let ๐‘Š โˆถ ๐‘ฃ 0 , . . . , ๐‘ฃ โ„“ be the walk. We will induct on โ„“, the length of ๐‘Š. If โ„“ = 0, then ๐‘Š is a path. So assume โ„“ โ‰ฅ 1. If ๐‘Š is a path, then we are done. If not, then some vertex of ๐‘Š is repeated, say ๐‘ฃ ๐‘– = ๐‘ฃ๐‘— for ๐‘– < ๐‘—. Then we have a ๐‘ขโ€“๐‘ฃ walk ๐‘Š โ€ฒ โˆถ ๐‘ฃ 0 , ๐‘ฃ 1 , . . . , ๐‘ฃ ๐‘– , ๐‘ฃ๐‘—+1 , ๐‘ฃ๐‘—+2 , . . . , ๐‘ฃ โ„“ which is shorter than ๐‘Š. By induction, ๐‘Š โ€ฒ contains a path ๐‘ƒ and so ๐‘Š contains ๐‘ƒ as well. โ–ก Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

20

1. Basic Counting

To state our first graphical enumeration result, let ๐’ข(๐‘‰) be the set of all graphs on the vertex set ๐‘‰. We will also use ๐’ข(๐‘‰, ๐‘˜) to denote the set of all graphs in ๐’ข(๐‘‰) with ๐‘˜ edges. Theorem 1.9.2. For ๐‘› โ‰ฅ 1 and ๐‘˜ โ‰ฅ 0 we have #๐’ข([๐‘›]) = 2( 2 ) ๐‘›

and #๐’ข([๐‘›], ๐‘˜) = (

(๐‘›2) ). ๐‘˜

Proof. If ๐‘‰ = [๐‘›] is given, then a graph ๐บ with vertex set ๐‘‰ is completely determined by its edge set. Since there are ๐‘› vertices, there are (๐‘›2) possible edges to choose from. So the number of ๐บ in ๐’ข([๐‘›]) is the number of subsets of these edges, which, by Theorem 1.3.1, is the given power of 2. The proof for ๐’ข([๐‘›], ๐‘˜) is similar, just using the definition (1.4). โ–ก A graph is unlabeled if the vertices in ๐‘‰ are indistinguishable. If the type of graph is clear from the context or does not matter for the particular application at hand, we will omit the adjectives โ€œlabeledโ€ and โ€œunlabeledโ€. The enumeration of unlabeled graphs is much more complicated than for labeled ones. So this discussion is postponed until Section 6.4 where we will develop the necessary tools. If ๐บ is a graph and ๐‘ฃ โˆˆ ๐‘‰, then the degree of ๐‘ฃ is deg ๐‘ฃ = the number of ๐‘’ โˆˆ ๐ธ containing ๐‘ฃ. In our running example deg ๐‘ฃ = deg ๐‘ฅ = 3 and deg ๐‘ค = deg ๐‘ฆ = 2. There is a nice relationship between vertex degrees and the cardinality of the edge set. The demonstration of the next result illustrates an important method of proof in combinatorics, counting in pairs. Theorem 1.9.3. For any graph ๐บ we have โˆ‘ deg ๐‘ฃ = 2|๐ธ|. ๐‘ฃโˆˆ๐‘‰

Proof. Consider ๐‘ƒ = {(๐‘ฃ, ๐‘’) | ๐‘ฃ is contained in ๐‘’}. Then #๐‘ƒ = โˆ‘ (number of ๐‘’ containing ๐‘ฃ) = โˆ‘ deg ๐‘ฃ. ๐‘ฃโˆˆ๐‘‰

๐‘ฃโˆˆ๐‘‰

On the other hand #๐‘ƒ = โˆ‘ (number of ๐‘ฃ contained in ๐‘’) = โˆ‘ 2 = 2|๐ธ|. ๐‘’โˆˆ๐ธ

๐‘’โˆˆ๐ธ

Equating the two counts finishes the proof. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

โ–ก

1.9. Graphs and digraphs

21

๐‘ฃ

๐‘ค

๐‘ฆ

๐‘ฅ

Figure 1.8. A digraph ๐ท

Theorem 1.9.3 is often called the Handshaking Lemma because of the following interpretation. Suppose ๐‘‰ is the set of people at a party and we draw an edge between person ๐‘ฃ and person ๐‘ค if they shake hands during the festivities. Then adding up the number of handshakes given by each person gives twice the total number of handshakes. It is often useful to have specified directions along the edges. A labeled directed graph, also called a digraph, is ๐ท = (๐‘‰, ๐ด) where ๐‘‰ is a set of vertices and ๐ด is a set of arcs which are ordered pairs of vertices. We use the notation ๐‘Ž = ๐‘ฃ๐‘ค โƒ— for arcs and say that ๐‘Ž goes from ๐‘ฃ to ๐‘ค. To illustrate, the digraph with ๐‘‰ = {๐‘ฃ, ๐‘ค, ๐‘ฅ, ๐‘ฆ} and ๐ด = {๐‘ฃ๐‘ค, โƒ— ๐‘ค๐‘ฃ, โƒ— ๐‘ค๐‘ฅ, โƒ— ๐‘ฆโƒ—๐‘ฃ, ๐‘ฆโƒ—๐‘ฅ} is drawn in Figure 1.8. We use ๐‘‰(๐ท) and ๐ด(๐ท) to denote the vertex set and arc set, respectively, of a digraph ๐ท when we wish to be more precise. Directed walks, paths, and cycles are defined for digraphs similarly to their undirected cousins in graphs, just insisting the ๐‘ฃโƒ— ๐‘–โˆ’1 ๐‘ฃ ๐‘– โˆˆ ๐ด for ๐‘– in the appropriate range. So, in our example digraph, ๐‘ƒ โˆถ ๐‘ฆ, ๐‘ฃ, ๐‘ค, ๐‘ฅ is a directed path and ๐ถ โˆถ ๐‘ฃ, ๐‘ค is a directed cycle. Note that ๐‘ค, ๐‘ฅ, ๐‘ฆ, ๐‘ฃ is not a directed path because the arc between ๐‘ฅ and ๐‘ฆ goes the wrong way. Let ๐’Ÿ(๐‘‰) and ๐’Ÿ(๐‘‰, ๐‘˜) be the set of digraphs and the set of digraphs with ๐‘˜ arcs, respectively, having vertex set ๐‘‰. The next result is proved in much the same manner as Theorem 1.9.2 so the demonstration is omitted. Theorem 1.9.4. For ๐‘› โ‰ฅ 1 and ๐‘˜ โ‰ฅ 0 we have #๐’Ÿ([๐‘›]) = 2๐‘›(๐‘›โˆ’1) and ๐‘›(๐‘› โˆ’ 1) #๐’Ÿ([๐‘›], ๐‘˜) = ( ). ๐‘˜

โ–ก

In a digraph ๐ท there are two types of degrees. Vertex ๐‘ฃ โˆˆ ๐‘‰ has out-degree and in-degree odeg ๐‘ฃ = the number of ๐‘Ž โˆˆ ๐ด of the form ๐‘Ž = ๐‘ฃ๐‘ค, โƒ— ideg ๐‘ฃ = the number of ๐‘Ž โˆˆ ๐ด of the form ๐‘Ž = ๐‘ค๐‘ฃ, โƒ— respectively. In Figure 1.8, for example, odeg ๐‘ฃ = 1 and ideg ๐‘ฃ = 2. The next result will permit us to finish our leftover business from Section 1.5. The union of digraphs ๐ท โˆช ๐ธ is the digraph with vertices ๐‘‰(๐ท โˆช ๐ธ) = ๐‘‰(๐ท) โˆช ๐‘‰(๐ธ) and arcs ๐ด(๐ท โˆช ๐ธ) = ๐ด(๐ท) โˆช ๐ด(๐ธ).

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

22

1. Basic Counting

Lemma 1.9.5. Let ๐ท = (๐‘‰, ๐ด) be a digraph. We have odeg ๐‘ฃ = ideg ๐‘ฃ = 1 for all ๐‘ฃ โˆˆ ๐‘‰ if and only if ๐ท is a disjoint union of directed cycles. Proof. The reverse implication is easy to see since the out-degree and in-degree of any vertex ๐‘ฃ of ๐ท would be the same as those degrees in the directed cycle containing ๐‘ฃ. But in such a cycle odeg ๐‘ฃ = ideg ๐‘ฃ = 1. For the forward direction, pick any ๐‘ฃ = ๐‘ฃ 1 โˆˆ ๐‘‰. Since odeg ๐‘ฃ 1 = 1 there must exist a vertex ๐‘ฃ 2 with ๐‘ฃโƒ— โƒ— 1 ๐‘ฃ 2 โˆˆ ๐ด. By the same token, there must be a ๐‘ฃ 3 with ๐‘ฃ 2 ๐‘ฃ 3 โˆˆ ๐ด. Continue to generate a sequence ๐‘ฃ 1 , ๐‘ฃ 2 , . . . in this manner. Since ๐‘‰ is finite, there must be two indices ๐‘– < ๐‘— such that ๐‘ฃ ๐‘– = ๐‘ฃ๐‘— . Let ๐‘– be the smallest such index and let ๐‘— be the first index after ๐‘– where repetition occurs. Thus ๐‘– = 1, for if not, then we have ๐‘ฃโƒ— ๐‘–โˆ’1 ๐‘ฃ ๐‘– , ๐‘ฃโƒ— ๐‘—โˆ’1 ๐‘ฃ ๐‘– โˆˆ ๐ด, contradicting the fact that ideg ๐‘ฃ ๐‘– = 1. By definition of ๐‘—, we have a directed cycle ๐ถ โˆถ ๐‘ฃ 1 , ๐‘ฃ 2 , . . . , ๐‘ฃ๐‘—โˆ’1 . Furthermore, no vertex of ๐ถ can be involved in another arc since that would make its out-degree or in-degree too large. Continuing in this manner, we can decompose ๐ท into disjoint directed cycles. โ–ก Sometimes it is useful to allow loops in a graph which are edges of the form ๐‘’ = ๐‘ฃ๐‘ฃ. Similarly, we can permit loops as arcs ๐‘Ž = ๐‘ฃโƒ—๐‘ฃ in a digraph. Another possibility is that we would want multiple edges, meaning that one could have more than one edge between a given pair of vertices, making ๐ธ into a multiset. Multiple arcs are defined similarly. If we make no specification for our (di)graph, then we are assuming that it has neither loops nor multiple edges. We will now prove Theorem 1.5.1. Proof (of Theorem 1.5.1). To any ๐œ‹ โˆˆ ๐”–๐‘› we associate its functional digraph ๐ท๐œ‹ which has ๐‘‰ = [๐‘›] and an arc ๐šค๐šฅโƒ— โˆˆ ๐ด if and only if ๐œ‹(๐‘–) = ๐‘—. Now ๐ท๐œ‹ is a digraph with loops. Because ๐œ‹ is a function we have odeg ๐‘– = 1 for all ๐‘– โˆˆ [๐‘›]. And because ๐œ‹ is a bijection we also have ideg ๐‘– = 1 for all ๐‘–. The proof of the previous lemma works equally well if one allows loops. So ๐ท๐œ‹ is a disjoint union of cycles. But cycles of the digraph ๐ท๐œ‹ correspond to cycles of the permutation ๐œ‹. Thus the cycle decomposition of ๐œ‹ exists. It is also easy to check that the cycles of ๐ท๐œ‹ produced by the algorithm in the demonstration of necessity in Lemma 1.9.5 are unique. This implies the uniqueness statement about the cycles of ๐œ‹ and so we are done. โ–ก

1.10. Trees Trees are a type of graph which often occurs in practice, even in domains outside of mathematics. For example, trees are used as data structures in computer science, or to model evolution in genetics. A graph ๐บ is connected if, for every pair of vertices ๐‘ฃ, ๐‘ค โˆˆ ๐‘‰, there is a walk in ๐บ from ๐‘ฃ to ๐‘ค. By Lemma 1.9.1(a), this is equivalent to there being a path from ๐‘ฃ to ๐‘ค in ๐บ. The connected components of ๐บ are the maximal connected subgraphs. If ๐บ is connected, there is only one component. Call ๐บ acyclic if it contains no cycles. A forest is another name for an acyclic graph. The connected components of a forest are called trees. So a graph ๐‘‡ is a tree if it is both connected and acyclic. Figure 1.9 contains five trees ๐‘‡1 , . . . , ๐‘‡5 . Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

1.10. Trees

23

1

1

1

4

6

5

3 2 ๐‘‡1 ๐‘ค1 = 4 ๐‘™1 = 5

6

4

3 2 ๐‘‡2 ๐‘ค2 = 6 ๐‘™2 = 3

1 6

4

2 ๐‘‡3 ๐‘ค3 = 1 ๐‘™3 = 6

4

1 4

2 ๐‘‡4 ๐‘ค4 = 4 ๐‘™4 = 2

๐‘‡5 ๐‘ค1 = 4 ๐‘™2 = 3

Figure 1.9. The Prรผfer algorithm

A leaf in a graph ๐บ is a vertex ๐‘ฃ having deg ๐‘ฃ = 1. The next result will show that nontrivial trees have leaves (regardless of the time of year). Further, it should be clear from this lemma why leaves are a useful tool for induction in trees. In order to state it, we need the following notation. If ๐บ is a graph and ๐‘Š โŠ† ๐‘‰, then ๐บ โˆ’ ๐‘Š is the graph on the vertex set ๐‘‰ โˆ’ ๐‘Š whose edge set consists of all edges in ๐ธ with both endpoints in ๐‘‰ โˆ’ ๐‘Š. If ๐‘Š = {๐‘ฃ} for some ๐‘ฃ, then we write ๐บ โˆ’ ๐‘ฃ for ๐บ โˆ’ {๐‘ฃ}. In Figure 1.9, ๐‘‡2 = ๐‘‡1 โˆ’ 5. Similarly, if ๐น โŠ† ๐ธ, then ๐บ โˆ’ ๐น is the graph with ๐‘‰(๐บ โˆ’ ๐น) = ๐‘‰(๐บ) and ๐ธ(๐บ โˆ’ ๐น) = ๐ธ(๐บ) โˆ’ ๐ธ(๐น). If ๐น consists of a single edge, then we use a similar abbreviation as for subtracting vertices. Lemma 1.10.1. Let ๐‘‡ be a tree with #๐‘‰ โ‰ฅ 2. (a) ๐‘‡ has (at least) 2 leaves. (b) If ๐‘ฃ is a leaf of ๐‘‡, then ๐‘‡ โ€ฒ = ๐‘‡ โˆ’ ๐‘ฃ is also a tree. Proof. (a) Let ๐‘ƒ โˆถ ๐‘ฃ 0 , . . . , ๐‘ฃ โ„“ be a path of maximum length in ๐‘‡. Since ๐‘‡ is nontrivial, ๐‘ฃ 0 โ‰  ๐‘ฃ โ„“ . We claim that ๐‘ฃ 0 , ๐‘ฃ โ„“ are leaves and we will prove this for ๐‘ฃ 0 as the same proof works for ๐‘ฃ โ„“ . Suppose, towards a contradiction, that deg ๐‘ฃ 0 โ‰ฅ 2. Then there must be a vertex ๐‘ค โ‰  ๐‘ฃ 1 such that ๐‘ฃ 0 ๐‘ค โˆˆ ๐ธ. We now have two possibilities. If ๐‘ค is not a vertex of ๐‘ƒ, then the path ๐‘ƒ โ€ฒ โˆถ ๐‘ค, ๐‘ฃ 0 , . . . , ๐‘ฃ โ„“ is longer than ๐‘ƒ, a contradiction to the definition of ๐‘ƒ. If ๐‘ค = ๐‘ฃ ๐‘– for some 2 โ‰ค ๐‘– โ‰ค โ„“, then the portion of ๐‘ƒ from ๐‘ฃ 0 to ๐‘ฃ ๐‘– together with the edge ๐‘ฃ 0 ๐‘ฃ ๐‘– forms a cycle in ๐‘‡, again a contradiction. (b) It is clear that ๐‘‡ โ€ฒ is still acyclic since removing vertices cannot create a cycle. To show it is connected, take ๐‘ฅ, ๐‘ฆ โˆˆ ๐‘‰(๐‘‡ โ€ฒ ). So ๐‘ฅ, ๐‘ฆ are also vertices of ๐‘‡. Since ๐‘‡ is connected, Lemma 1.9.1(a) implies that there is a path ๐‘ƒ from ๐‘ฅ to ๐‘ฆ in ๐‘‡. If this path is also in ๐‘‡ โ€ฒ , then we will be done. But if ๐‘ƒ goes through ๐‘ฃ, then, since there is a unique vertex ๐‘ฃโ€ฒ adjacent to ๐‘ฃ, ๐‘ƒ would have to pass through ๐‘ฃโ€ฒ just before and just after ๐‘ฃ. This contradicts the fact that the vertices of ๐‘ƒ are distinct. โ–ก There are a number of characterizations of trees. We collect some of them here as they will be useful in the sequel. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

24

1. Basic Counting

Theorem 1.10.2. Let ๐‘‡ be a graph with #๐‘‰ = ๐‘› and #๐ธ = ๐‘š. The following are equivalent conditions for ๐‘‡ to be a tree: (a) ๐‘‡ is connected and acyclic. (b) ๐‘‡ is acyclic and ๐‘› = ๐‘š + 1. (c) ๐‘‡ is connected and ๐‘› = ๐‘š + 1. (d) For every pair of vertices ๐‘ข, ๐‘ฃ there is a unique path from ๐‘ข to ๐‘ฃ. Proof. We will prove the equivalence of (a), (b), and (c). The equivalence of (a) and (d) is left as an exercise. To prove that (a) implies (b), it suffices to show by induction on ๐‘› that ๐‘› = ๐‘š + 1. This is trivial if ๐‘› = 1. If ๐‘› โ‰ฅ 2, then, by Lemma 1.10.1, ๐‘‡ has a leaf ๐‘ฃ. Induction applies to ๐‘‡ โ€ฒ = ๐‘‡ โˆ’ ๐‘ฃ so that its vertex and edge cardinalities are related by ๐‘›โ€ฒ = ๐‘šโ€ฒ + 1. But ๐‘› = ๐‘›โ€ฒ + 1 and ๐‘š = ๐‘šโ€ฒ + 1 so that ๐‘› = ๐‘š + 1. To see why (b) implies (c), consider the connected components ๐‘‡1 , . . . , ๐‘‡๐‘˜ of ๐‘‡. Since ๐‘‡ is acyclic, each of these components is a tree. Also, from the implication (a) โŸน (b), we have that ๐‘›๐‘– = ๐‘š๐‘– + 1 for 1 โ‰ค ๐‘– โ‰ค ๐‘˜ where ๐‘›๐‘– = #๐‘‰(๐‘‡๐‘– ) and ๐‘š๐‘– = #๐ธ(๐‘‡๐‘– ). Adding these equations together and using the fact that โˆ‘๐‘– ๐‘›๐‘– = ๐‘› and โˆ‘๐‘– ๐‘š๐‘– = ๐‘š we obtain ๐‘› = ๐‘š + ๐‘˜. But we are given that ๐‘› = ๐‘š + 1. So we must have ๐‘˜ = 1 . This means that ๐‘‡ only has one component and so is connected. We prove that (c) implies (a) by contradiction. So suppose that ๐‘‡ contains a cycle ๐ถ and let ๐‘’ = ๐‘ข๐‘ฃ โˆˆ ๐ธ(๐ถ). We claim that ๐‘‡ โˆ’ ๐‘’ is still connected. For if ๐‘ฅ, ๐‘ฆ are any two vertices of ๐‘‡ โˆ’ ๐‘’, then there is a walk ๐‘Š from ๐‘ฅ to ๐‘ฆ in ๐‘‡. If ๐‘Š does not contain ๐‘’, then ๐‘Š is still in ๐‘‡ โˆ’ ๐‘’. If ๐‘Š does contain ๐‘’, then replace ๐‘’ in ๐‘Š with the path ๐ถ โˆ’ ๐‘’ to form a new walk ๐‘Š โ€ฒ from ๐‘ฅ to ๐‘ฆ in ๐‘‡ โˆ’ ๐‘’. We can keep removing edges in this way until the resulting graph ๐‘‡ โ€ฒ is acyclic. Since ๐‘‡ โ€ฒ is still connected, it is a tree. And by the first implication we have ๐‘›โ€ฒ = ๐‘šโ€ฒ + 1. But ๐‘›โ€ฒ = ๐‘› and ๐‘šโ€ฒ < ๐‘š so that ๐‘› < ๐‘š + 1, the desired contradiction. โ–ก Let ๐’ฏ(๐‘‰) be the set of all trees on the vertex set ๐‘‰. There are quite a number of different proofs of the beautiful formula below for #๐’ฏ(๐‘‰), many of which are in Moonโ€™s book on the subject [64]. Theorem 1.10.3. For ๐‘› โ‰ฅ 1 we have #๐’ฏ([๐‘›]) = ๐‘›๐‘›โˆ’2 . Proof. The result is trivial if ๐‘› = 1, so assume ๐‘› โ‰ฅ 2. By Theorem 1.2.2 it suffices to find a bijection ๐‘“ โˆถ ๐’ฏ([๐‘›]) โ†’ ๐‘ƒ(([๐‘›], ๐‘› โˆ’ 2)). There is a famous algorithm for constructing ๐‘“ which is called the Prรผfer algorithm. An example will be found in Figure 1.9. Given ๐‘‡ โˆˆ ๐’ฏ([๐‘›]), to determine ๐‘“(๐‘‡) = ๐‘ค 1 . . . ๐‘ค ๐‘›โˆ’2 we will build a sequence of trees ๐‘‡ = ๐‘‡1 , ๐‘‡2 , . . . , ๐‘‡๐‘›โˆ’1 by removing vertices from ๐‘‡ as follows. Since the vertices of ๐‘‡ are labeled 1, . . . , ๐‘› it makes sense to talk about, e.g., a maximum vertex because of the ordering on the integers. Given ๐‘‡๐‘– , we find the leaf ๐‘™๐‘– โˆˆ ๐‘‰(๐‘‡๐‘– ) such that ๐‘™๐‘– is maximum and let ๐‘‡๐‘–+1 = ๐‘‡๐‘– โˆ’ ๐‘™๐‘– . By the previous lemma, ๐‘‡๐‘–+1 will also be a tree. Since ๐‘™๐‘– is a leaf, it is adjacent to a unique vertex ๐‘ค ๐‘– in ๐‘‡๐‘– and we let ๐‘ค ๐‘– be the ๐‘–th element of ๐‘“(๐‘‡). Now each ๐‘ค ๐‘– โˆˆ [๐‘›] and ๐‘“(๐‘‡) has length ๐‘› โˆ’ 2 by definition. So ๐‘“(๐‘‡) โˆˆ ๐‘ƒ(([๐‘›], ๐‘› โˆ’ 2)). Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

1.11. Lattice paths

25

To show that ๐‘“ is a bijection, we find its inverse. Given ๐‘ค โˆˆ ๐‘ƒ(([๐‘›], ๐‘› โˆ’ 2)), we will first construct a permutation ๐‘™ = ๐‘™1 . . . ๐‘™๐‘›โˆ’2 โˆˆ ๐‘ƒ([๐‘›], ๐‘› โˆ’ 2) where ๐‘™๐‘– will turn out to be the leaf removed from ๐‘‡๐‘– to form ๐‘‡๐‘–+1 . We construct the ๐‘™๐‘– inductively by letting (1.10)

๐‘™๐‘– = max([๐‘›] โˆ’ {๐‘™1 , . . . , ๐‘™๐‘–โˆ’1 , ๐‘ค ๐‘– , . . . , ๐‘ค ๐‘›โˆ’2 }).

Finally we construct ๐‘“โˆ’1 (๐‘ค) = ๐‘‡ by letting ๐‘‡ have edges ๐‘’ ๐‘– = ๐‘™๐‘– ๐‘ค ๐‘– for 1 โ‰ค ๐‘– โ‰ค ๐‘› โˆ’ 2 as well as the edge ๐‘’ ๐‘›โˆ’1 = ๐‘™๐‘›โˆ’1 ๐‘™๐‘› where [๐‘›] โˆ’ {๐‘™1 , . . . , ๐‘™๐‘›โˆ’2 } = {๐‘™๐‘›โˆ’1 , ๐‘™๐‘› }. To show that ๐‘“โˆ’1 (๐‘ค) = ๐‘‡ is a tree, note first that ๐‘™1 is a leaf of ๐‘‡ because ๐‘™1 is attached to ๐‘ค 1 but to none of the other vertices of ๐‘‡ by (1.10) and the definition of ๐‘’ ๐‘›โˆ’1 . Consider ๐‘คโ€ฒ = ๐‘ค 2 . . . ๐‘ค ๐‘›โˆ’2 and apply the algorithm for ๐‘“โˆ’1 to ๐‘คโ€ฒ using the ground set [๐‘›] โˆ’ {๐‘™1 } instead of [๐‘›]. By induction, the result is a tree ๐‘‡ โ€ฒ . And ๐‘‡ is formed by adding ๐‘™1 as a leaf to ๐‘‡ โ€ฒ , which makes ๐‘‡ a tree as well. To show that ๐‘“ and ๐‘“โˆ’1 are inverses we will show that ๐‘“โˆ’1 โˆ˜ ๐‘“ is the identity map, leaving the proof for ๐‘“ โˆ˜ ๐‘“โˆ’1 to the reader. Suppose ๐‘“(๐‘‡) = ๐‘ค 1 . . . ๐‘ค ๐‘›โˆ’2 . Also let โ€ฒ the sequence of leaves removed during the construction of ๐‘“(๐‘‡) be ๐‘™1โ€ฒ . . . ๐‘™๐‘›โˆ’2 . Then โ€ฒ by definition of the algorithm, the edges of ๐‘‡ are exactly ๐‘™๐‘– ๐‘ค ๐‘– for 1 โ‰ค ๐‘– โ‰ค ๐‘› โˆ’ 2 โ€ฒ โ€ฒ โ€ฒ } = {๐‘™๐‘›โˆ’1 , ๐‘™๐‘›โ€ฒ }. Comparing this with the definition ๐‘™๐‘›โ€ฒ where [๐‘›] โˆ’ {๐‘™1โ€ฒ , . . . , ๐‘™๐‘›โˆ’2 and ๐‘™๐‘›โˆ’1 โˆ’1 of ๐‘“ we see that it suffices to show that ๐‘™๐‘– = ๐‘™๐‘–โ€ฒ for all ๐‘– and that this will follow if one can prove the equality holds for 1 โ‰ค ๐‘– โ‰ค ๐‘› โˆ’ 2. Since ๐‘™๐‘–โ€ฒ is a leaf in ๐‘‡๐‘– , it cannot โ€ฒ . Of the remaining vertices, those be any of the previously removed leaves ๐‘™1โ€ฒ , . . . , ๐‘™๐‘–โˆ’1 which are among ๐‘ค ๐‘– , . . . , ๐‘ค ๐‘›โˆ’2 are not currently leaves since they are attached to future leaves which are to be removed. And conversely those not among the ๐‘ค ๐‘– , . . . , ๐‘ค ๐‘›โˆ’2 must be leaves; otherwise, they would be listed as some ๐‘ค๐‘— for ๐‘— โ‰ฅ ๐‘– once all their adjacent leaves were removed. Hence the leaves of ๐‘‡๐‘– are precisely the elements of โ€ฒ [๐‘›] โˆ’ {๐‘™1โ€ฒ , . . . , ๐‘™๐‘–โˆ’1 , ๐‘ค ๐‘– , . . . , ๐‘ค ๐‘›โˆ’2 }. Since we always remove the leaf of maximum value, we see that the rule for choosing ๐‘™๐‘–โ€ฒ is exactly the same as the one in (1.10). So ๐‘™๐‘– = ๐‘™๐‘–โ€ฒ as desired. โ–ก

1.11. Lattice paths Lattice paths lead to many interesting counting problems in combinatorics. They are also important in probability and statistics; see the book of Mohanty [63] for examples. Consider the integer lattice in the plane โ„ค2 = {(๐‘ฅ, ๐‘ฆ) | ๐‘ฅ, ๐‘ฆ โˆˆ โ„ค}. A lattice path is a sequence of elements of โ„ค2 written ๐‘ƒ โˆถ (๐‘ฅ0 , ๐‘ฆ0 ), (๐‘ฅ1 , ๐‘ฆ1 ), . . . , (๐‘ฅโ„“ , ๐‘ฆโ„“ ). Just as in graph theory, we say the path has length โ„“ and goes from (๐‘ฅ0 , ๐‘ฆ0 ) to (๐‘ฅโ„“ , ๐‘ฆโ„“ ), which are called its endpoints. Unlike graph-theoretic paths, we do not assume the (๐‘ฅ๐‘– , ๐‘ฆ ๐‘– ) are distinct. To illustrate the notation, if we assume that the left-hand path in Figure 1.10 starts at the origin, then it would be written ๐‘ƒ โˆถ (0, 0), (0, 1), (0, 2), (1, 2), (1, 3), (2, 3), (3, 3), (3, 4), (4, 4). Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

26

1. Basic Counting

Figure 1.10. Dyck paths

The step between (๐‘ฅ๐‘–โˆ’1 , ๐‘ฆ ๐‘–โˆ’1 ) and (๐‘ฅ๐‘– , ๐‘ฆ ๐‘– ) on ๐‘ƒ is the vector ๐‘ ๐‘– = [๐‘ฅ๐‘– โˆ’ ๐‘ฅ๐‘–โˆ’1 , ๐‘ฆ ๐‘– โˆ’ ๐‘ฆ ๐‘–โˆ’1 ]. Note the use of square versus round brackets to distinguish steps from vertices of the path. Note that ๐‘ƒ is determined up to translation by its steps and that it is determined completely by its steps and initial vertex. If no initial vertex is specified, it is assumed to be the origin. We let ๐ธ = [1, 0] and ๐‘ = [0, 1], calling these east and north steps, respectively. The path on the left in Figure 1.10 could also be represented ๐‘ƒ โˆถ ๐‘๐‘๐ธ๐‘๐ธ๐ธ๐‘๐ธ. For our first enumerative result, we use the notation ๐’ฉโ„ฐ(๐‘š, ๐‘›) for the set of lattice paths from (0, 0) to (๐‘š, ๐‘›) only using steps north and east. We call lattice paths using only ๐‘ and ๐ธ steps northeast paths. Theorem 1.11.1. For ๐‘š, ๐‘› โ‰ฅ 0 we have #๐’ฉโ„ฐ(๐‘š, ๐‘›) = (

๐‘š+๐‘› ). ๐‘š

Proof. Let ๐‘ƒ be a northeast lattice path from (0, 0) to (๐‘š, ๐‘›). Then ๐‘ƒ has ๐‘š + ๐‘› total steps. And once ๐‘š of them are chosen to be ๐ธ, the rest must be ๐‘. The result follows. โ–ก

We will be particularly concerned with a special type of northeast path. A Dyck path of semilength ๐‘› is a northeast lattice path which begins at (0, 0), ends at (๐‘›, ๐‘›), and never goes below the line ๐‘ฆ = ๐‘ฅ. The first path in Figure 1.10 is of this type. Note that ๐‘› is called the semilength because the Dyck path itself has 2๐‘› steps. We let ๐’Ÿ(๐‘›) denote the set of Dyck paths of semilength ๐‘›. This should cause no confusion with the notation ๐’Ÿ(๐‘‰) for the set of digraphs on the vertex set ๐‘‰ because in the former notation ๐‘› is a nonnegative integer while in the latter it is a set. We now define that Catalan numbers to be ๐ถ(๐‘›) = #๐’Ÿ(๐‘›). The Catalan numbers are ubiquitous in combinatorics. In fact, Stanley has written a book [92] containing 214 different combinatorial interpretations of ๐ถ(๐‘›). A few of these are listed in the exercises. The Catalan numbers satisfy a nice recursion. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

1.11. Lattice paths

27

Theorem 1.11.2. We have the initial condition ๐ถ(0) = 1 and recurrence relation ๐ถ(๐‘›) = ๐ถ(0)๐ถ(๐‘› โˆ’ 1) + ๐ถ(1)๐ถ(๐‘› โˆ’ 2) + ๐ถ(2)๐ถ(๐‘› โˆ’ 3) + โ‹ฏ + ๐ถ(๐‘› โˆ’ 1)๐ถ(0) for ๐‘› โ‰ฅ 1. Proof. The initial condition counts the trivial path of a single vertex. For the recursion, take ๐‘ƒ โˆถ ๐‘ฃ 0 , . . . , ๐‘ฃ 2๐‘› โˆˆ ๐’Ÿ(๐‘›) where ๐‘ฃ ๐‘– = (๐‘ฅ๐‘– , ๐‘ฆ ๐‘– ) for all ๐‘–. Let ๐‘— > 0 be the smallest index such that ๐‘ฃ 2๐‘— is on the line ๐‘ฆ = ๐‘ฅ. Such an index exists since ๐‘ฃ 2๐‘› = (๐‘›, ๐‘›) satisfies this condition. Also note that no vertex of odd subscript is on ๐‘ฆ = ๐‘ฅ since the number of north steps and the number of east steps preceding that vertex cannot be equal. It follows that ๐‘ƒ1 , the portion of ๐‘ƒ from ๐‘ฃ 1 to ๐‘ฃ 2๐‘—โˆ’1 , stays above ๐‘ฆ = ๐‘ฅ + 1. So the number of choices for ๐‘ƒ1 is ๐ถ(๐‘— โˆ’ 1). Furthermore, if ๐‘ƒ2 is the portion of ๐‘ƒ from ๐‘ฃ 2๐‘— to ๐‘ฃ 2๐‘› , then ๐‘ƒ2 is (a translation of) a Dyck path of semilength ๐‘› โˆ’ ๐‘—. So the number of choices for ๐‘ƒ2 is ๐ถ(๐‘› โˆ’ ๐‘—). Thus the total number of such ๐‘ƒ is ๐ถ(๐‘— โˆ’ 1)๐ถ(๐‘› โˆ’ ๐‘—). Summing over 1 โ‰ค ๐‘— โ‰ค ๐‘› finishes the proof. โ–ก There is an explicit expression for the Catalan numbers. But to derive this formula it will be convenient to use a second set of paths counted by ๐ถ(๐‘›). Call the steps ๐‘ˆ = [1, 1] and ๐ท = [1, โˆ’1] up and down, respectively. An updown path is one using only ฬƒ such steps. It should be clear that if we let ๐’Ÿ(๐‘›) be the set of updown lattice paths from ฬƒ (0, 0) to (2๐‘›, 0) never going below the ๐‘ฅ-axis, then #๐’Ÿ(๐‘›) = #๐’Ÿ(๐‘›) = ๐ถ(๐‘›). In fact one can get from the paths in one set to those in the other by rotation and dilation of the plane. The two paths in Figure 1.10 correspond under this map and the second one would be represented as ๐‘ƒ โˆถ ๐‘ˆ๐‘ˆ๐ท๐‘ˆ๐ท๐ท๐‘ˆ๐ท. Theorem 1.11.3. For ๐‘› โ‰ฅ 0 we have ๐ถ(๐‘›) =

1 2๐‘› ( ). ๐‘›+1 ๐‘›

Proof. We rewrite the right-hand side as 1 2๐‘› (2๐‘›)! 1 2๐‘› + 1 = ( )= ( ). ๐‘›+1 ๐‘› 2๐‘› + 1 ๐‘› ๐‘›! (๐‘› + 1)! Let ๐’ซ be the set of all updown paths starting at (0, 0) and ending at (2๐‘› + 1, โˆ’1). Such paths have 2๐‘› + 1 steps of which ๐‘› are up (forcing the other ๐‘› + 1 to be down) so that #๐’ซ = (2๐‘›+1 ). Our strategy will be to find a partition ๐œŒ of ๐’ซ such that ๐‘› (1) #๐ต = 2๐‘› + 1 for every block ๐ต of ๐œŒ and ฬƒ (2) there is a bijection between the blocks of ๐œŒ and the paths in ๐’Ÿ(๐‘›). ฬƒ It will then follow that #๐’Ÿ(๐‘›) is equal to the number of blocks of ๐œŒ, which is #๐’ซ/(2๐‘› + 1), giving the desired equality. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

28

1. Basic Counting

To determine ๐œŒ, we will take any ๐‘ƒ โˆˆ ๐’ซ and describe the block ๐ต containing ๐‘ƒ. We will refer to the ๐‘ฆ-coordinate of a vertex ๐‘ฃ of ๐‘ƒ as its height, written ht ๐‘ฃ. Suppose ๐‘ƒ has step representation ๐‘ƒ โˆถ ๐‘ 1 ๐‘ 2 . . . ๐‘ 2๐‘›+1 . Define the ๐‘Ÿth rotation of ๐‘ƒ to be the path ๐‘ƒ๐‘Ÿ โˆถ ๐‘ ๐‘Ÿ+1 ๐‘ ๐‘Ÿ+2 . . . ๐‘ 2๐‘›+1 ๐‘ 1 ๐‘ 2 . . . ๐‘ ๐‘Ÿ where all paths start at the origin. Let ๐ต = {๐‘ƒ0 , . . . , ๐‘ƒ2๐‘› }. So to show that ๐ต has the correct cardinality, we must prove that the ๐‘ƒ๐‘– are all distinct. Suppose to the contrary that two are equal. By renumbering if necessary, we can assume that ๐‘ƒ0 = ๐‘ƒ๐‘— for some 1 โ‰ค ๐‘— โ‰ค 2๐‘›. Take ๐‘— be be minimum. Iterating this equality, we get ๐‘ƒ0 = ๐‘ƒ๐‘— = ๐‘ƒ2๐‘— = . . . . These equalities and the fact that ๐‘— is as small as possible imply that ๐‘ƒ = ๐‘ƒ0 is the concatenation of ๐‘ƒโ€ฒ โˆถ ๐‘ 1 . . . ๐‘ ๐‘— with itself, say ๐‘˜ times for some ๐‘˜ โ‰ฅ 2. Suppose ๐‘ƒ โ€ฒ ends at height โ„Ž. Then ๐‘ƒ must end at height ๐‘˜โ„Ž and so ๐‘˜โ„Ž = โˆ’1. This forces ๐‘˜ = 1, which is a contradiction. To finish the proof, we must show that the blocks of ๐œŒ are in bijection with the ฬƒ paths in ๐’Ÿ(๐‘›). Let ๐’Ÿฬƒ โ€ฒ (๐‘›) denote the set of paths obtained by appending a down step to ฬƒ each path in ๐’Ÿ(๐‘›). So ๐œŒ partitions ๐’ซ โŠ‡ ๐’Ÿฬƒ โ€ฒ (๐‘›). Thus it suffices to show that there is a ฬƒ unique path from ๐’Ÿ โ€ฒ (๐‘›) in each block ๐ต of ๐œŒ. Let ๐ต be generated by rotating a path ๐‘ƒ as in the previous paragraph and let ๐‘ƒ โˆถ ๐‘ฃ 0 . . . ๐‘ฃ 2๐‘›+1 be the lattice point representation of ๐‘ƒ. Let โ„Ž be the minimum height of a vertex of ๐‘ƒ, and among all vertices of ๐‘ƒ of height โ„Ž let ๐‘ฃ ๐‘Ÿ be the left most. We claim that ๐‘ƒ๐‘Ÿ โˆˆ ๐’Ÿฬƒ โ€ฒ (๐‘›) and no other ๐‘ƒ๐‘  is in this set for ๐‘  โˆˆ {0, 1, . . . , ๐‘›} โˆ’ {๐‘Ÿ}. We will prove the first of these two claims and leave the second, whose demonstration is similar, as an exercise. Since ๐‘ฃ ๐‘Ÿ is translated to the origin and has smallest height in ๐‘ƒ, the translations of all ๐‘ฃ ๐‘– for ๐‘– โ‰ฅ ๐‘Ÿ will lie weakly above the ๐‘ฅaxis. As for the ๐‘ฃ ๐‘– with ๐‘– < ๐‘Ÿ, they must be translated so the ๐‘ฃ ๐‘Ÿ becomes the last vertex of ๐‘ƒ๐‘Ÿ which is of height โˆ’1. But since ๐‘ฃ ๐‘Ÿ was the first vertex of minimum height in ๐‘ƒ, the vertices before it must be translated to have height greater than โˆ’1 and so must also lie weakly above the ๐‘ฅ-axis. It follows that only the last vertex of ๐‘ƒ๐‘Ÿ is below the ๐‘ฅ-axis, which is what we wished to prove. โ–ก

1.12. Pattern avoidance Pattern avoidance is a relatively recent area of study in combinatorics. It has seen strong growth in part because of its connections to algebraic geometry and computer science. For more information about this topic, see the books of Bรณna [18] or Kitaev [48]. Let ๐‘† be a set of integers with #๐‘† = ๐‘˜ and consider a permutation ๐œŽ โˆˆ ๐‘ƒ(๐‘†). The standardization of ๐œŽ is the permutation std ๐œŽ โˆˆ ๐‘ƒ([๐‘˜]) obtained by replacing the smallest element of ๐œŽ by 1, the next smallest by 2, and so on. For example, if ๐œŽ = 263, then std ๐œŽ = 132. Given ๐œŽ โˆˆ ๐”–๐‘› and ๐œ‹ โˆˆ ๐”–๐‘˜ in one-line notation, we say that ๐œŽ contains a copy of ๐œ‹ if there is a subsequence ๐œŽโ€ฒ of ๐œŽ such that std ๐œŽโ€ฒ = ๐œ‹. Note that a subsequence need not consist of consecutive elements of ๐œ‹. In this case, ๐œ‹ is called the pattern. To illustrate, ๐œŽ = 425613 contains the pattern ๐œ‹ = 132 since ๐œŽโ€ฒ = 263 standardizes to ๐œ‹. On the other hand, we say that ๐œŽ avoids ๐œ‹ if it has no subsequence ๐œŽโ€ฒ with std ๐œŽ = ๐œ‹. Continuing our example, one can check that ๐œŽ avoids 4321 since ๐œŽ does not contain a decreasing subsequence of length four. There is an equivalent Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

1.12. Pattern avoidance

29

๐œŽ=

๐œ‹=

Figure 1.11. The diagrams for ๐œ‹ = 132 and ๐œŽ = 425613

definition of pattern containment which the reader will see in the literature. If ๐‘†, ๐‘‡ are sets with #๐‘† = #๐‘‡ = ๐‘˜, then call ๐œŽ = ๐œŽ1 . . . ๐œŽ ๐‘˜ โˆˆ ๐‘ƒ(๐‘†) and ๐œ = ๐œ1 . . . ๐œ๐‘˜ โˆˆ ๐‘ƒ(๐‘‡) order isomorphic if ๐œŽ ๐‘– < ๐œŽ๐‘— is equivalent to ๐œ๐‘– < ๐œ๐‘— for all ๐‘–, ๐‘—. It is easy to see that ๐œŽ contains a copy of ๐œ‹ if and only if ๐œŽ contains a subsequence order isomorphic to ๐œ‹. To study patterns, it will be useful to have a geometric model of a permutation analogous to its permutation matrix. Again, the integer lattice will come into play. Given ๐œŽ = ๐œŽ1 . . . ๐œŽ๐‘› โˆˆ ๐”–๐‘› , its diagram is the set of points (๐‘–, ๐œŽ ๐‘– ) โˆˆ โ„ค2 for 1 โ‰ค ๐‘– โ‰ค ๐‘›. In displaying the diagram, the lower-left corner is always assumed to have coordinates (1, 1). Using our running example, the diagrams for ๐œ‹ = 132 and ๐œŽ = 425613 are shown in Figure 1.11. The points corresponding to the copy 263 of ๐œ‹ in ๐œŽ have been enlarged to emphasize how easily one can see pattern containment using diagrams. From an enumerative point of view, avoidance often turns out to be easier to work with than containment. So given ๐œ‹ โˆˆ ๐”–๐‘˜ , we consider Av๐‘› (๐œ‹) = {๐œŽ โˆˆ ๐”–๐‘› | ๐œŽ avoids ๐œ‹}. Note that many authors use ๐”–๐‘› (๐œ‹) instead of Av๐‘› (๐œ‹) for this set. Call ๐œ‹ and ๐œ‹โ€ฒ Wilf equivalent, written ๐œ‹ โ‰ก ๐œ‹โ€ฒ , if # Av๐‘› (๐œ‹) = # Av๐‘› (๐œ‹โ€ฒ ) for all ๐‘› โ‰ฅ 0. It is easy to see that this is an equivalence relation on ๐”–๐‘› . We will prove that any two permutations in ๐”–3 are Wilf equivalent, although this is not as startling as it might first sound. Certain Wilf equivalences follow easily from manipulation of diagrams. Consider the dihedral group of the square (1.11)

๐ท = {๐œŒ0 , ๐œŒ90 , ๐œŒ180 , ๐œŒ270 , ๐‘Ÿ0 , ๐‘Ÿ1 , ๐‘Ÿโˆ’1 , ๐‘Ÿโˆž }

where ๐œŒ๐œƒ is rotation by ๐œƒ degrees counterclockwise and ๐‘Ÿ๐‘š is reflection across a line of slope ๐‘š. If ๐œŽ contains a copy ๐œŽโ€ฒ of ๐œ‹ and ๐‘“ โˆˆ ๐ท, then ๐‘“(๐œŽ) contains a copy ๐‘“(๐œŽโ€ฒ ) of ๐‘“(๐œ‹). Using ๐‘“โˆ’1 , we see that the converse of the previous assertion is also true. It follows that ๐œŽ avoids ๐œ‹ if and only if ๐‘“(๐œŽ) avoids ๐‘“(๐œ‹). We have proven the following result. Lemma 1.12.1. For any ๐œ‹ โˆˆ ๐”–๐‘˜ and any ๐‘“ โˆˆ ๐ท we have ๐œ‹ โ‰ก ๐‘“(๐œ‹). Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

โ–ก

30

1. Basic Counting

๐œŽโ€ฒ ๐œŽ=

๐œŽโ€ณ

๐‘— Figure 1.12. Decomposing ๐œŽ โˆˆ Av๐‘› (132)

The equivalences in this lemma are called trivial Wilf equivalences. In particular, in ๐”–3 one sees by repeatedly applying ๐œŒ90 that 132 โ‰ก 231 โ‰ก 213 โ‰ก 312 and 123 โ‰ก 321. In fact, all six permutations are Wilf equivalent and their avoidance sets are counted by the Catalan numbers. We start with 132 from the first set of equivalent permutations. Theorem 1.12.2. For ๐‘› โ‰ฅ 0 we have # Av๐‘› (132) = ๐ถ(๐‘›). Proof. We will induct on ๐‘›, using the initial condition and recurrence relation for ๐ถ(๐‘›) given in Theorem 1.11.2. As usual, we concentrate on the latter. Pick ๐œŽ = ๐œŽ1 . . . ๐œŽ๐‘› โˆˆ Av๐‘› (132) and suppose ๐œŽ๐‘— = ๐‘›. So we can write ๐œŽ = ๐œŽโ€ฒ ๐‘›๐œŽโ€ณ where ๐œŽโ€ฒ = ๐œŽ1 . . . ๐œŽ๐‘—โˆ’1 and ๐œŽโ€ณ = ๐œŽ๐‘—+1 . . . ๐œŽ๐‘› . Clearly ๐œŽโ€ฒ and ๐œŽโ€ณ must avoid 132 since they are subsequences of ๐œŽ. We also claim that min ๐œŽโ€ฒ > max ๐œŽโ€ณ so that we can think of the diagram of ๐œŽ decomposing as in Figure 1.12. Indeed, if there is ๐‘  โˆˆ ๐œŽโ€ฒ and ๐‘ก โˆˆ ๐œŽโ€ณ with ๐‘  < ๐‘ก, then ๐œŽ contains ๐‘ ๐‘›๐‘ก, which is a copy of 132, a contradiction. Thus ๐œŽโ€ฒ and ๐œŽโ€ณ are permutations of {๐‘›โˆ’1, ๐‘›โˆ’2, . . . , ๐‘›โˆ’๐‘—+1} and [๐‘›โˆ’๐‘—], respectively, both of which avoid 132. Conversely, if the diagram of ๐œŽ has the form given in Figure 1.12 with ๐œŽโ€ฒ , ๐œŽโ€ณ avoiding 132, then ๐œŽ must avoid 132. This is a case-by-case proof by contradiction, considering where the elements of a copy of 132 could lie in the diagram if one existed. We leave the details to the reader. To finish the count, from what we have shown and induction there are ๐ถ(๐‘— โˆ’ 1) choices for ๐œŽโ€ฒ and ๐ถ(๐‘› โˆ’ ๐‘—) for ๐œŽโ€ณ . Taking their product and summing over ๐‘— โˆˆ [๐‘›] shows that there are ๐ถ(๐‘›) choices for ๐œŽ. โ–ก Next we will tackle 123, but to do so we will need some new concepts. The leftright minima of ๐œŽ = ๐œŽ1 . . . ๐œŽ๐‘› โˆˆ ๐”–๐‘› are the ๐œŽ ๐‘– satisfying ๐œŽ ๐‘– < min{๐œŽ1 , ๐œŽ2 , . . . , ๐œŽ ๐‘–โˆ’1 }. For example ๐œŽ = 698371542 has left-right minima ๐œŽ1 = 6, ๐œŽ4 = 3, and ๐œŽ6 = 1. The indices ๐‘– such that ๐œŽ ๐‘– is a left-right minimum are called the left-right minimum positions. If necessary to distinguish from the positions, the ๐œŽ๐‘— themselves are called the left-right Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

1.12. Pattern avoidance

31

minimum values. Reading the left-right minima in order from left to right, the positions and values always satisfy (1.12)

1 = ๐‘– 1 < ๐‘– 2 < โ‹ฏ < ๐‘–๐‘™

and

๐‘š1 > ๐‘š 2 > โ‹ฏ > ๐‘š ๐‘™ = 1

for some ๐‘™ โ‰ฅ 1. We will need to determine, given a set of values and positions, whether a permutation exists with left-right minima having these values and positions. To do this, we introduce the dominance order on compositions, which is also useful in other areas of combinatorics and representation theory. A weak composition of ๐‘› is a sequence ๐›ผ = [๐›ผ1 , . . . , ๐›ผ๐‘™ ] of nonnegative integers with โˆ‘๐‘– ๐›ผ๐‘– = ๐‘›. So in weak compositions zero is permitted as a part and we will use 0 subscripts on notation for compositions when used for weak compositions. If ๐›ผ, ๐›ฝ โŠง0 ๐‘›, then ๐›ผ is dominated by ๐›ฝ, written ๐›ผ โŠด ๐›ฝ, if we have ๐›ผ1 + ๐›ผ2 + โ‹ฏ + ๐›ผ๐‘— โ‰ค ๐›ฝ1 + ๐›ฝ2 + โ‹ฏ + ๐›ฝ๐‘— for all ๐‘— โ‰ฅ 1, where ๐›ผ๐‘— = 0 if ๐‘— > โ„“(๐›ผ) and similarly for ๐›ฝ. To illustrate, [2, 2, 1, 1] โŠด [3, 1, 2] because 2 โ‰ค 3, 2+2 โ‰ค 3+1, 2+2+1 โ‰ค 3+1+2, and 2+2+1+1 = 3+1+2+0. Since ๐›ผ, ๐›ฝ โŠง0 ๐‘› the last inequality always becomes an equality. In the next result, the reader will notice a similarity between the construction of ๐œ„ and ๐œ‡ and the map ๐œ™ defined by (1.8). Lemma 1.12.3. Let ๐œŽ โˆˆ ๐”–๐‘› . (a) We have ๐œŽ โˆˆ Av๐‘› (123) if and only if its subsequence of non-left-right minima is decreasing. (b) There exists ๐œŽ โˆˆ Av๐‘› (123) with left-right minima positions and values given by (1.12) if and only if ๐œ„ โŠด ๐œ‡ where ๐œ„ = (๐‘–2 โˆ’ ๐‘–1 โˆ’ 1, ๐‘–3 โˆ’ ๐‘–2 โˆ’ 1, . . . , ๐‘–๐‘™+1 โˆ’ ๐‘–๐‘™ โˆ’ 1), ๐œ‡ = (๐‘š0 โˆ’ ๐‘š1 โˆ’ 1, ๐‘š1 โˆ’ ๐‘š2 โˆ’ 1, . . . , ๐‘š๐‘™โˆ’1 โˆ’ ๐‘š๐‘™ โˆ’ 1), and ๐‘–๐‘™+1 = ๐‘š0 = ๐‘› + 1. In this case, ๐œŽ is unique. Proof. (a) We will prove this statement in its contrapositive form. Suppose first that ๐œŽ contains a copy ๐œŽ ๐‘– ๐œŽ๐‘— ๐œŽ ๐‘˜ of 123. Then ๐œŽ๐‘— , ๐œŽ ๐‘˜ cannot be left-right minima since ๐œŽ ๐‘– is smaller than both and to their left in ๐œŽ. Since ๐œŽ๐‘— < ๐œŽ ๐‘˜ , the non-left-right minima subsequence contains an increase. Conversely, suppose ๐œŽ๐‘— < ๐œŽ ๐‘˜ with ๐‘— < ๐‘˜ and both non-left-right minima. Let ๐œŽ ๐‘– be the left-right minimum closest to ๐œŽ๐‘— on its left. We have that ๐œŽ ๐‘– exists since ๐œŽ begins with a left-right minimum. Then ๐œŽ ๐‘– < ๐œŽ๐‘— < ๐œŽ ๐‘˜ , giving a copy of 123. (b) Clearly if ๐œŽ exists, then it must be unique since the positions and values of its left-right minima are given by (1.12) and the rest of the elements can only be arranged in one way by (a). We can attempt to build ๐œŽ satisfying the given conditions as follows. An example will be found following the proof. Start with a row of ๐‘› blank positions. Now fill in the values ๐‘š1 > โ‹ฏ > ๐‘š๐‘™ at the positions ๐‘–1 < โ‹ฏ < ๐‘–๐‘™ . Filling in the rest of the positions with the elements of ๐‘† = [๐‘›] โˆ’ {๐‘š1 , . . . , ๐‘š๐‘™ } (the set of non-left-right minima) in decreasing order gives a ๐œŽ avoiding 123 since ๐œŽ is a union of two decreasing subsequences. So the only question is whether doing this will result in a permutation Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

32

1. Basic Counting

having the ๐‘š๐‘— as its left-right minima. We have that ๐‘š1 is always a left-right minimum regardless of the other entries. Now ๐‘š๐‘—+1 will be the next left-right minimum after ๐‘š๐‘— if and only if the blanks before position ๐‘–๐‘—+1 are filled with elements larger than ๐‘š๐‘— . Note that ๐œ„๐‘— = ๐‘–๐‘—+1 โˆ’ ๐‘–๐‘— โˆ’ 1 is the number of spaces between positions ๐‘–๐‘— and ๐‘–๐‘—+1 . Also ๐œ‡๐‘— = ๐‘š๐‘—โˆ’1 โˆ’๐‘š๐‘— โˆ’1 is the number of ๐‘  โˆˆ ๐‘† with ๐‘š๐‘— < ๐‘  < ๐‘š๐‘—โˆ’1 . It follows that ๐œ„1 +โ‹ฏ+๐œ„๐‘— is the number of blanks before position ๐‘–๐‘—+1 and ๐œ‡1 + โ‹ฏ + ๐œ‡๐‘— is the number of elements of ๐‘† greater than ๐‘š๐‘— . So filling in the spaces preserves the left-right minima if and only if the inequalities for ๐œ„ โŠด ๐œ‡ are satisfied. This completes the proof. โ–ก Suppose we want to see if there is ๐œŽ โˆˆ Av9 (123) with left-right minima 6 > 3 > 1 in positions 1 < 4 < 6. We start off with the diagram (1.13)

๐œŽ=6

3 1

.

We wish to check whether filling the blanks with the remaining elements of [9] in decreasing order will result in a permutation which has the initial elements as leftright minima. One way to do this is just to fill the blanks and verify that the desired elements become left-right minima: ๐œŽ = 6 9 8 3 7 1 5 4 2. Another way is to use the ๐œ„ and ๐œ‡ compositions. Note that ๐œ„1 = 4 โˆ’ 1 โˆ’ 1 = 2 is the number of blanks between ๐‘š1 = 6 and ๐‘š2 = 3 in the original diagram. Similarly ๐œ‡1 = 10โˆ’6โˆ’1 = 3 is the number of elements of ๐‘† = [9]โˆ’{6, 3, 1} greater than ๐‘š1 = 6. In order to fill the blanks between 6 and 3 so that 6 is a left-right minimum, the numbers used must all be greater than 6. This is possible exactly when ๐œ„1 โ‰ค ๐œ‡1 . Similarly ๐œ„1 + ๐œ„2 โ‰ค ๐œ‡1 + ๐œ‡2 ensures that one can fill the blanks to the left of ๐‘š3 = 1 with numbers greater than ๐‘š2 = 3, and so forth. So checking whether ๐œ„ โŠด ๐œ‡ also determines whether ๐œŽ has the correct left-right minima. We will need an analogue of Lemma 1.12.3 for elements of Av๐‘› (132). To state it, we define the reversal of a weak composition ๐›ผ = [๐›ผ1 , ๐›ผ2 , . . . , ๐›ผ๐‘™ ] to be ๐›ผ๐‘Ÿ = [๐›ผ๐‘™ , ๐›ผ๐‘™โˆ’1 , . . . , ๐›ผ1 ]. Lemma 1.12.4. Let ๐œŽ โˆˆ ๐”–๐‘› . (a) We have ๐œŽ โˆˆ Av๐‘› (132) if and only if, for every left-right minimum ๐‘š, the elements of ๐œŽ to the right of and greater than ๐‘š form an increasing subsequence. (b) There exists ๐œŽ โˆˆ Av๐‘› (132) with left-right minima positions and values given by (1.12) if and only if ๐œ‡๐‘Ÿ โŠด ๐œ„๐‘Ÿ where ๐œ„, ๐œ‡ are as given in Lemma 1.12.3. In this case, ๐œŽ is unique Proof. Much of the proof of this result is similar to the demonstration of Lemma 1.12.3 and so will be left as an exercise. Here we will only present the construction of ๐œŽ โˆˆ Av๐‘› (132) from its diagram of left-right minima and blanks. Again, an example follows the explanation. We keep the notation of the proof of the previous lemma. We start by filling the blanks to the right of ๐‘š๐‘™ = 1 with the elements ๐‘  โˆˆ ๐‘† such that ๐‘š๐‘™ < ๐‘  < ๐‘š๐‘™โˆ’1 in increasing order and as far left as possible (so they will be consecutive). Next we fill in the remaining blanks to the right of ๐‘š๐‘™โˆ’1 with those ๐‘  โˆˆ ๐‘† such that ๐‘š๐‘™โˆ’1 < ๐‘  < ๐‘š๐‘™โˆ’2 so that they form an increasing subsequence which is as far left as possible given the spaces already filled. Continue in this manner until all blanks are occupied. โ–ก Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Exercises

33

Suppose we wish to fill in the diagram (1.13) so that ๐œŽ avoids 132. If ๐‘š3 < ๐‘  < ๐‘š2 , 3 1 2 . Similarly, then ๐‘  = 2, so we put 2 just to the right of ๐‘š3 = 1 to get ๐œŽ = 6 ๐‘š1 < ๐‘  < ๐‘š2 is satisfied by ๐‘  = 4, 5 so we put these elements following ๐‘š2 = 3 in increasing order using the left-most blanks available to obtain ๐œŽ = 6 34125 . Finally, we do the same for the elements greater than ๐‘š1 = 6 to get the end result ๐œŽ = 6 7 8 3 4 1 2 5 9. We need one last observation before we achieve our goal of showing all elements of ๐”–3 are Wilf equivalent. Suppose ๐›ผ = [๐›ผ1 , . . . , ๐›ผ๐‘™ ] and ๐›ฝ = [๐›ฝ1 , . . . , ๐›ฝ ๐‘™ ] are weak compositions of ๐‘›. We claim ๐›ผ โŠด ๐›ฝ if and only if ๐›ฝ๐‘Ÿ โŠด ๐›ผ๐‘Ÿ . To see this, note that the inequality ๐›ผ1 +โ‹ฏ+๐›ผ๐‘— โ‰ค ๐›ฝ1 +โ‹ฏ+๐›ฝ๐‘— is equivalent to ๐‘›โˆ’(๐›ฝ1 +โ‹ฏ+๐›ฝ๐‘— ) โ‰ค ๐‘›โˆ’(๐›ผ1 +โ‹ฏ+๐›ผ๐‘— ). But ๐‘›โˆ’(๐›ผ1 +โ‹ฏ+๐›ผ๐‘— ) = ๐›ผ๐‘Ÿ +๐›ผ๐‘Ÿโˆ’1 +โ‹ฏ+๐›ผ๐‘—+1 and similarly for ๐›ฝ. Making this substitution we get the necessary inequalities for ๐›ฝ๐‘Ÿ โŠด ๐›ผ๐‘Ÿ and all steps are reversible. Finally, we say that a bijection ๐‘“ โˆถ ๐‘† โ†’ ๐‘‡ preserves property ๐‘ƒ if ๐‘  โˆˆ ๐‘† having property ๐‘ƒ is equivalent to ๐‘“(๐‘ ) having property ๐‘ƒ for all ๐‘  โˆˆ ๐‘†. Theorem 1.12.5. For ๐‘› โ‰ฅ 0 and any ๐œ‹ โˆˆ ๐”–3 we have # Av๐‘› (๐œ‹) = ๐ถ(๐‘›). Proof. By Theorem 1.12.2 and the discussion just before it, it suffices to show that we have # Av๐‘› (123) = ๐ถ(๐‘›). This will be true if we can find a bijection ๐‘“ โˆถ Av๐‘› (123) โ†’ Av๐‘› (132). In fact, ๐‘“ will preserve the values and positions of left-right minima. Suppose ๐œŽ โˆˆ Av๐‘› (123) has its positions and values given by (1.12). By Lemma 1.12.3 there is a unique such ๐œŽ and we must also have ๐œ„ โŠด ๐œ‡. But, as noted just before this theorem, this is equivalent to ๐œ‡๐‘Ÿ โŠด๐œ„๐‘Ÿ . So, using Lemma 1.12.4, there is a unique ๐œŽโ€ฒ โˆˆ Av๐‘› (132) having the given positions and values of its left-right minima and we let ๐‘“(๐œŽ) = ๐œŽโ€ฒ . Because of the existence and uniqueness of ๐œŽ and ๐œŽโ€ฒ , this is a bijection. โ–ก Note that the description of ๐‘“ in the previous proof can be made constructive. Given ๐œŽ โˆˆ Av๐‘› (123), we remove its non-left-right minima and rearrange them using the algorithm in the proof of Lemma 1.12.4. So, using our running example, ๐‘“(698371542) = 678341259.

Exercises (1) Prove each of the following identities for ๐‘› โ‰ฅ 1 in two ways: one inductive and one combinatorial. ๐‘›

(a) โˆ‘ ๐น๐‘– = ๐น๐‘›+2 โˆ’ 1. ๐‘–=1 ๐‘›

(b) โˆ‘ ๐น2๐‘– = ๐น2๐‘›+1 โˆ’ 1. ๐‘–=1 ๐‘›

(c) โˆ‘ ๐น2๐‘–โˆ’1 = ๐น2๐‘› . ๐‘–=1

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

34

1. Basic Counting

(2) Prove that if ๐‘˜, ๐‘› โˆˆ โ„™ with ๐‘˜ โˆฃ ๐‘› (meaning ๐‘˜ divides evenly into ๐‘›), then ๐น ๐‘˜ โˆฃ ๐น๐‘› . (3) Given ๐‘š โˆˆ โ„™, show that the sequence of Fibonacci numbers is periodic modulo ๐‘š; that is, there exists ๐‘ โˆˆ โ„™ such that ๐น๐‘›+๐‘ โ‰ก ๐น๐‘› (mod ๐‘š) for all ๐‘› โ‰ฅ 0. The period modulo ๐‘š is the smallest ๐‘ such that this congruence holds. Note that it is an open problem to find the period of the Fibonacci sequence for an arbitrary ๐‘š. (4) The Lucas numbers are defined by ๐ฟ0 = 2, ๐ฟ1 = 1, and ๐ฟ๐‘› = ๐ฟ๐‘›โˆ’1 + ๐ฟ๐‘›โˆ’2 for ๐‘› โ‰ฅ 2. Prove the following identities for ๐‘š, ๐‘› โ‰ฅ 1. (a) ๐ฟ๐‘› = ๐น๐‘›โˆ’1 + ๐น๐‘›+1 . (b) Let ๐’ž๐‘› be the set of tilings of ๐‘› boxes arranged in a circle with dominos and monominos. Show that #๐’ž๐‘› = ๐ฟ๐‘› . (c) ๐ฟ๐‘š+๐‘› = ๐น๐‘šโˆ’1 ๐ฟ๐‘› + ๐น๐‘š ๐ฟ๐‘›+1 . (d) ๐น2๐‘› = ๐น๐‘› ๐ฟ๐‘› . (5) Prove Theorem 1.2.2. (6) Check that the two maps defined in the proof of Theorem 1.3.1 are inverses. (7) (a) Prove Theorem 1.3.3(b) using equation (1.5). (b) Give an inductive proof of Theorem 1.3.3(c). (c) Give an inductive proof of Theorem 1.3.3(d). (8) Let ๐‘†, ๐‘‡ be sets. (a) Show that ๐‘† ฮ” ๐‘‡ = (๐‘† โˆช ๐‘‡) โˆ’ (๐‘† โˆฉ ๐‘‡). (b) Show that (๐‘† ฮ” ๐‘‡) ฮ” ๐‘‡ = ๐‘†. (9) Given nonnegative integers satisfying ๐‘›1 + ๐‘›2 + โ‹ฏ + ๐‘›๐‘š = ๐‘›. the corresponding multinomial coefficient is (1.14)

(

๐‘› ๐‘›! . )= ๐‘›1 ! ๐‘› 2 ! . . . ๐‘› ๐‘š ! ๐‘›1 , ๐‘›2 , . . . , ๐‘›๐‘š

We extend this definition to negative ๐‘›๐‘– by letting the multinomial coefficient be zero if any ๐‘›๐‘– < 0. Note that when ๐‘š = 2 we recover the binomial coefficients as (

๐‘› ๐‘› ) = ( ). ๐‘˜, ๐‘› โˆ’ ๐‘˜ ๐‘˜

(a) Find and prove analogues of Theorem 1.3.3(a), (b), and (c) for multinomial coefficients. (b) A permutation of a multiset ๐‘€ = {{1๐‘›1 , 2๐‘›2 , . . . , ๐‘š๐‘›๐‘š }} is a linear arrangement of the elements of ๐‘€. Let ๐‘ƒ(๐‘€) denote the set of permutations of ๐‘€. For example ๐‘ƒ({{12 , 22 }}) = {1122, 1212, 1221, 2112, 2121, 2211}. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Exercises

35

1 1 1 1 1 1

0

0

1

1

1

1 1

1 1

0 0

1 1

1 0

0 0

1

1 0

1 1

1 1

1 0

1

1 1

1

Figure 1.13. Pascalโ€™s triangle modulo 2

Prove that ๐‘› #๐‘ƒ({{1๐‘›1 , 2๐‘›2 , . . . , ๐‘š๐‘›๐‘š }}) = ( ) ๐‘›1 , ๐‘› 2 , . . . , ๐‘› ๐‘š in three ways: (i) combinatorially, (ii) by induction on ๐‘›, (iii) by proving that ๐‘› ๐‘› ๐‘› โˆ’ ๐‘›1 ( ) = ( )( ) ๐‘›1 , ๐‘› 2 , . . . , ๐‘› ๐‘š ๐‘›1 ๐‘›2 , . . . , ๐‘›๐‘š and then inducting on ๐‘š. (10) (a) Prove the Pascal triangle is fractal modulo 2. Specifically, if one replaces each binomial coefficient by its remainder on division by 2, then, for any ๐‘˜ โ‰ฅ 0, the triangle consisting of rows 0 through 2๐‘˜ โˆ’ 1 is repeated on the left and on the right in rows 2๐‘˜ through 2๐‘˜+1 โˆ’1 with an inverted triangle of zeros in between. See Figure 1.13 for the first eight rows. Hint: Induct on ๐‘˜. (b) Formulate and prove an analogous result modulo ๐‘ for any prime ๐‘. (11) Find the inverse for the map in the proof of Theorem 1.3.4, proving that it is welldefined and the inverse to the given function. (12) For ๐‘› โ‰ฅ 0 define the ๐‘›th Fibotorial to be the product ๐น๐‘›! = ๐น1 ๐น2 . . . ๐น๐‘› . Also, for 0 โ‰ค ๐‘˜ โ‰ค ๐‘› define a Fibonomial coefficient by ๐น! ๐‘› ( ) = ! ๐‘›! . ๐‘˜ ๐น๐‘˜ ๐น๐‘›โˆ’๐‘˜ ๐น Note that from this definition it is not clear that this is an integer. (a) Show that the Fibonomial coefficients satisfy the initial conditions (๐‘›0) = ๐น (๐‘›๐‘›) = 1 and recurrence ๐น

๐‘› ๐‘›โˆ’1 ๐‘›โˆ’1 ( ) = ๐น๐‘›โˆ’๐‘˜+1 ( ) + ๐น ๐‘˜โˆ’1 ( ) ๐‘˜ ๐‘˜โˆ’1 ๐‘˜ ๐น

๐น

๐น

for 0 < ๐‘˜ < ๐‘›. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

36

1. Basic Counting

(b) Show that (๐‘›๐‘˜) is an integer for all 0 โ‰ค ๐‘˜ โ‰ค ๐‘›. ๐น (c) Find a combinatorial interpretation of (๐‘›๐‘˜) . ๐น

(13) For ๐‘› โ‰ฅ 1 show that the Stirling numbers of the second kind have the following values. (a) ๐‘†(๐‘›, 1) = 1. (b) ๐‘†(๐‘›, 2) = 2๐‘›โˆ’1 โˆ’ 1. (c) ๐‘†(๐‘›, ๐‘›) = 1. ๐‘› (d) ๐‘†(๐‘›, ๐‘› โˆ’ 1) = ( ). 2 ๐‘› ๐‘› (e) ๐‘†(๐‘›, ๐‘› โˆ’ 2) = ( ) + 3( ). 3 4 (14) For ๐‘› โ‰ฅ 1 show that the signless Stirling numbers of the first kind have the following values. (a) ๐‘(๐‘›, 1) = (๐‘› โˆ’ 1)!. ๐‘› 1 (b) ๐‘(๐‘› + 1, 2) = ๐‘›! โˆ‘ . ๐‘– ๐‘–=1 (c) ๐‘(๐‘›, ๐‘›) = 1. ๐‘› (d) ๐‘(๐‘›, ๐‘› โˆ’ 1) = ( ). 2 ๐‘› ๐‘› (e) ๐‘(๐‘›, ๐‘› โˆ’ 2) = 2( ) + 3( ). 3 4 (15) Call an integer partition ๐œ† self-conjugate if ๐œ†๐‘ก = ๐œ†. Show that the number of selfconjugate ๐œ† โŠข ๐‘› equals the number of ๐œ‡ โŠข ๐‘› having parts which are distinct (no part can be repeated) and odd. Hint: Use Young diagrams and try to guess a bijection inductively by first seeing what it has to be for small ๐‘›. Then try to construct a bijection for ๐‘› + 1 which will be consistent in some way with the one for previous values. Finally try to describe your bijection in a noninductive manner. (16) The main diagonal of a Young diagram is the set of squares starting with the one at the top left and moving diagonally right and down. So in Figure 1.5, the main diagonal of ๐œ† consists of two squares. Prove the following. (a) If ๐œ† is self-conjugate as defined in the previous exercise, then |๐œ†| โ‰ก ๐‘‘ (mod 2) where ๐‘‘ is the length (number of squares) of the main diagonal. (b) Let ๐‘ ๐‘‘ (๐‘›) be the number of partitions of ๐‘› whose diagonal has length ๐‘‘. Then ๐‘ ๐‘‘ (๐‘›) = โˆ‘ ๐‘(๐‘š, ๐‘‘)๐‘(๐‘› โˆ’ ๐‘š โˆ’ ๐‘‘ 2 , ๐‘‘). ๐‘šโ‰ฅ0

(17) Define ๐‘๐‘’ (๐‘›, ๐‘˜) to be the number of ๐œ† โŠข ๐‘› having exactly ๐‘˜ parts. Prove the following under the assumption that ๐‘› โ‰ฅ 4, where โŒŠโ‹…โŒ‹ is the round-down function. (a) ๐‘๐‘’ (๐‘›, ๐‘˜) = ๐‘(๐‘› โˆ’ ๐‘˜, ๐‘˜). (b) ๐‘๐‘’ (๐‘›, 1) = 1. (c) ๐‘๐‘’ (๐‘›, 2) = โŒŠ๐‘›/2โŒ‹. (d) ๐‘๐‘’ (๐‘›, ๐‘› โˆ’ 2) = 2. (e) ๐‘๐‘’ (๐‘›, ๐‘› โˆ’ 1) = 1. (f) ๐‘๐‘’ (๐‘›, ๐‘›) = 1. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Exercises

37

(18) Finish the proof of Theorem 1.7.1. (19) Prove Theorem 1.7.2. (20) Consider a line of ๐‘› copies of the integer 1. One can now put slashes in the spaces between the 1โ€™s and count the number of 1โ€™s between each pair of adjacent slashes to form a composition of ๐‘›. For example, if ๐‘› = 6, then we start with 1 1 1 1 1 1. One way of inserting slashes is 1 1/1/1 1 1, which corresponds to the composition 2 + 1 + 3 = 6. Give alternate proofs of Theorems 1.7.1 and 1.7.2 using this idea. (21) A weak composition of ๐‘› into ๐‘˜ parts is a sequence of ๐‘˜ nonnegative integers summing to ๐‘›. Find a formula for the number of weak compositions of ๐‘› into ๐‘˜ parts and then prove it in three different ways: (a) by using a variant of the map ๐œ™ defined in (1.8), (b) by finding a relation between weak compositions and compositions and then using the statement of Theorem 1.7.2 (as opposed to their proofs as in part (a)), (c) by modifying the construction in the previous exercise. (22) Show that the last two columns in Table 1.1 agree when ๐‘“ is bijective, that is, when ๐‘› = ๐‘˜. (23) Prove Lemma 1.9.1(b). (24) A graph ๐บ = (๐‘‰, ๐ธ) is regular if all of its vertices have the same degree. If deg ๐‘ฃ = ๐‘Ÿ for all vertices ๐‘ฃ, then ๐บ is regular of degree ๐‘Ÿ. (a) Show that if ๐บ is regular of degree ๐‘Ÿ, then ๐‘Ÿ|๐‘‰| |๐ธ| = , 2 (b) Call ๐บ bipartite if there is a set partition of ๐‘‰ = ๐‘‰1 โŠŽ ๐‘‰2 such for all ๐‘ข๐‘ฃ โˆˆ ๐ธ we have ๐‘ข โˆˆ ๐‘‰1 and ๐‘ฃ โˆˆ ๐‘‰2 or vice versa. Show that a bipartite graph regular of degree ๐‘Ÿ โ‰ฅ 1 has |๐‘‰1 | = |๐‘‰2 |. (25) A graph ๐บ is planar if it can be drawn in the plane โ„2 without edge crossings. In this case the regions of ๐บ are the topologically connected components of the settheoretic difference โ„2 โˆ’ ๐บ. Let ๐‘… be the set of regions of ๐บ. If ๐‘Ÿ โˆˆ ๐‘…, then let deg ๐‘Ÿ be the number of edges on the boundary of ๐‘Ÿ. Show that โˆ‘ deg ๐‘Ÿ โ‰ค 2|๐ธ|. ๐‘Ÿโˆˆ๐‘…

Find, with proof, a condition on the cycles of ๐บ which is equivalent to having equality. (26) Two graphs ๐บ, ๐ป are isomorphic, written ๐บ โ‰… ๐ป, if they are equal as unlabeled graphs. The complement of a graph ๐บ = (๐‘‰, ๐ธ) is the graph ๐บฬ„ with vertices ๐‘‰ and with ๐‘ข๐‘ฃ an edge of ๐บฬ„ if and only if ๐‘ข๐‘ฃ โˆ‰ ๐ธ. Call ๐บ self-complementary if ๐บ โ‰… ๐บ.ฬ„ (a) Show that there exists a self-complementary graph with ๐‘› vertices if and only if ๐‘› โ‰ก 0 (mod 4) or ๐‘› โ‰ก 1 (mod 4). (b) Show that in a self-complementary graph with ๐‘› vertices where ๐‘› โ‰ก 1 (mod 4) there must be at least one vertex of degree (๐‘› โˆ’ 1)/2. Hint: Show that the number of vertices of degree (๐‘› โˆ’ 1)/2 must be odd. (27) Prove Theorem 1.9.4. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

38

1. Basic Counting

(28) Prove that in any digraph ๐ท = (๐‘‰, ๐ด) we have โˆ‘ ideg ๐‘ฃ = โˆ‘ odeg ๐‘ฃ = |๐ด|. ๐‘ฃโˆˆ๐‘‰

๐‘ฃโˆˆ๐‘‰

(29) Prove the equivalence of (a) and (d) in Theorem 1.10.2. Hint: Use Lemma 1.9.1(b). (30) Consider a sequence of nonnegative integers ๐‘‘ โˆถ ๐‘‘1 , . . . , ๐‘‘๐‘› . Call ๐‘‘ a degree sequence if there is a graph with vertices ๐‘ฃ 1 , . . . , ๐‘ฃ ๐‘› such that deg ๐‘ฃ ๐‘– = ๐‘‘๐‘– for all ๐‘–. (a) Let ๐‘‡ be a tree with ๐‘› vertices and arrange its degree sequence in weakly decreasing order. Prove that for 1 โ‰ค ๐‘– โ‰ค ๐‘› we have ๐‘›โˆ’1 ๐‘‘๐‘– โ‰ค โŒˆ โŒ‰. ๐‘– (b) Let ๐‘‡ be a tree with ๐‘› vertices and let ๐‘˜ โ‰ฅ 2 be an integer. Suppose the degree sequence of ๐‘‡ satisfies ๐‘‘๐‘– = 1 or ๐‘˜ for all ๐‘–. Prove that ๐‘‘๐‘– = ๐‘˜ for exactly (๐‘› โˆ’ 2)/(๐‘˜ โˆ’ 1) indices ๐‘–. (31) Finish the proof of Theorem 1.10.3. (32) Consider ๐‘› cars ๐ถ1 , . . . , ๐ถ๐‘› passing in this order by a line of ๐‘› parking spaces numbered 1, . . . , ๐‘›. Each car ๐ถ๐‘– has a preferred space number ๐‘ ๐‘– in which to park. If ๐ถ๐‘– gets to space ๐‘ ๐‘– and it is free, then it parks. Otherwise it proceeds to the next empty space (which will have a number greater than ๐‘ ๐‘– ) and parks there if such a space exists. If no such space exists, it does not park. Call ๐‘ = (๐‘ 1 , . . . , ๐‘๐‘› ) a parking function of length ๐‘› if all the cars end up in a parking space. (a) Show that ๐‘ is a parking function if and only if its unique weakly increasing rearrangement ๐‘‘ = (๐‘‘1 , . . . , ๐‘‘๐‘› ) satsifies ๐‘‘๐‘– โ‰ค ๐‘– for all ๐‘– โˆˆ [๐‘›]. (b) Use a counting argument to show that the number of parking functions of length ๐‘› is (๐‘› + 1)๐‘›โˆ’1 . Hint: Consider parking where there are ๐‘› + 1 spaces arranged in a circular manner and ๐‘› + 1 is an allowed preference for cars. (c) Reprove (b) by finding a bijection between parking functions of length ๐‘› and trees on ๐‘› + 1 vertices. Hint: Let ๐‘‡ be a tree on ๐‘› + 1 vertices labeled 0, . . . , ๐‘› and call vertex 0 the root of the tree. Draw ๐‘‡ in the plane so that the vertices connected to the root, called the rootโ€™s children, are in increasing order read left to right. Continue to do the same thing for the children of each child of the root, and so forth. Create a permutation ๐œ‹ by reading the children of the root left to right, then the grandchildren of the root left to right, etc. Finally, orient each edge of ๐‘‡ so that it points from a vertex to its parent and call this set of arcs ๐ด. Map ๐‘‡ to ๐‘ = (๐‘ 1 , . . . , ๐‘๐‘› ) where ๐‘๐‘– = {

if ๐šค0โƒ— โˆˆ ๐ด, if ๐šค๐œ‹ โƒ—๐‘— โˆˆ ๐ด.

1 1+๐‘—

(33) Consider ๐ธ๐‘Š-lattice paths along the ๐‘ฅ-axis which are paths starting at the origin and using steps ๐ธ = [1, 0] and ๐‘Š = [โˆ’1, 0]. (a) Show that if an ๐ธ๐‘Š-lattice path has length ๐‘› and ends at (๐‘˜, 0), then ๐‘› and ๐‘˜ have the same parity and |๐‘˜| โ‰ค ๐‘›. (b) Show that the number of ๐ธ๐‘Š-lattice paths of length ๐‘› ending at (๐‘˜, 0) is ๐‘› ( ๐‘›+๐‘˜ ). 2

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Exercises

39

(c) Show that the number of ๐ธ๐‘Š-lattice paths of length 2๐‘› ending at the origin and always staying on the nonnegative side of the axis is ๐ถ(๐‘›). (34) Show that the Catalan numbers ๐ถ(๐‘›) also count the following objects: (a) ballot sequences which are words ๐‘ค = ๐‘ค 1 . . . ๐‘ค 2๐‘› containing ๐‘› ones and ๐‘› twos such that in any prefix ๐‘ค 1 . . . ๐‘ค ๐‘– the number of ones is always at least as great as the number of twos, (b) sequences of positive integers 1 โ‰ค ๐‘Ž 1 โ‰ค ๐‘Ž 2 โ‰ค . . . โ‰ค ๐‘Ž๐‘› with ๐‘Ž๐‘– โ‰ค ๐‘– for 1 โ‰ค ๐‘– โ‰ค ๐‘›, (c) triangulations of a convex (๐‘› + 2)-gon using nonintersecting diagonals, (d) noncrossing partitions ๐œŒ = ๐ต1 / . . . /๐ต๐‘˜ โŠข [๐‘›] where a crossing is ๐‘Ž < ๐‘ < ๐‘ < ๐‘‘ such that ๐‘Ž, ๐‘ โˆˆ ๐ต๐‘– and ๐‘, ๐‘‘ โˆˆ ๐ต๐‘— for ๐‘– โ‰  ๐‘—. (35) Fill in the details of the proof of Theorem 1.11.3. (36) A stack is a first-in first-out (FIFO) data structure with two operations. One can put something on the top of a stack, called pushing, or take something from the top of the stack, called popping. A permutation ๐œŽ = ๐œŽ1 . . . ๐œŽ๐‘› โˆˆ ๐”–๐‘› is considered sorted if its elements have been rearranged to form the permutation ๐œ = 12 . . . ๐‘›. Consider the following algorithm for sorting ๐œŽ. Start with an empty stack and an empty output permutation ๐œ. At each stage there are two options. If the stack is empty or the current first element ๐‘  of ๐œŽ is smaller than the top element of the stack, then one pushes ๐‘  onto the stack. If ๐œŽ has become empty or the top element ๐‘ก of the stack is smaller than the first element of ๐œŽ, then one pops ๐‘ก from the stack and appends it to the end of ๐œ. An example showing the sorting of ๐œŽ = 3124 will ๐œ

stack

๐œŽ

๐œ–

๐œ–

3124

๐œ–

3

124

1 ๐œ–

3

24

1

3

24

2 1

3

4

12

3

4

123

๐œ–

4

123

4

๐œ–

1234

๐œ–

๐œ–

Figure 1.14. A stack-sorting algorithm

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

40

1. Basic Counting

be found in Figure 1.14. Note that the input permutation ๐œŽ is on the right and the output permutation ๐œ is on the left so that the head of ๐œŽ and the tail of ๐œ are nearest the stack. (a) Show that this algorithm sorts ๐œŽ if and only if ๐œŽ โˆˆ Av๐‘› (231). (b) Show that if there is a sequence of pushes and pops which sorts ๐œŽ, then it must be the sequence given by the algorithm. (37) Suppose ๐œ‹ = ๐œ‹1 . . . ๐œ‹๐‘˜ โˆˆ ๐”–๐‘˜ . Prove the following descriptions of actions of elements of ๐ท in terms of one-line notation. (a) ๐‘Ÿโˆž (๐œ‹) = ๐œ‹๐‘˜ . . . ๐œ‹1 โ‰” ๐œ‹๐‘Ÿ , the reversal of ๐œ‹. (b) ๐‘Ÿ0 (๐œ‹) = (๐‘˜ + 1 โˆ’ ๐œ‹1 ) . . . (๐‘˜ + 1 โˆ’ ๐œ‹๐‘˜ ) โ‰” ๐œ‹๐‘ , the complement of ๐œ‹. (c) ๐‘Ÿ1 (๐œ‹) = ๐œ‹โˆ’1 , the group-theoretic inverse of ๐œ‹. (38) Finish the proof of Theorem 1.12.2. (39) Given any set of permutations ฮ  we let Av๐‘› (ฮ ) = {๐œŽ โˆˆ ๐”–๐‘› โˆฃ ๐œŽ avoids all ๐œ‹ โˆˆ ฮ }. If ๐œ‹ = ๐œ‹1 ๐œ‹2 . . . ๐œ‹๐‘š โˆˆ ๐”–๐‘š is a permutation and ๐‘› โˆˆ โ„•, then we can construct a new permutation ๐œ‹ + ๐‘› = ๐œ‹1 + ๐‘›, ๐œ‹2 + ๐‘›, . . . , ๐œ‹๐‘š + ๐‘›. Given permutations ๐œ‹, ๐œŽ of disjoint sets, we denote by ๐œ‹๐œŽ the permutation obtained by concatenating them. Define two other concatenations on ๐œ‹ โˆˆ ๐”–๐‘š and ๐œŽ โˆˆ ๐”–๐‘› , the direct sum ๐œ‹ โŠ• ๐œŽ = ๐œ‹(๐œŽ + ๐‘š) and skew sum ๐œ‹ โŠ– ๐œŽ = (๐œ‹ + ๐‘›)๐œŽ. Finally for ๐‘› โ‰ฅ 0 we use the notation ๐œ„๐‘› = 12 . . . ๐‘› for the increasing permutation of length ๐‘›, and ๐›ฟ๐‘› = ๐‘› . . . 21 for the decreasing one. Prove the following. (a) Av๐‘› (213, 321) = {๐œ„๐‘˜1 โŠ• (๐œ„๐‘˜2 โŠ– ๐œ„๐‘˜3 ) โˆฃ ๐‘˜1 + ๐‘˜2 + ๐‘˜3 = ๐‘›}. (b) Av๐‘› (132, 213) = {๐œ„๐‘˜1 โŠ– ๐œ„๐‘˜2 โŠ– โ‹ฏ โˆฃ โˆ‘๐‘– ๐‘˜๐‘– = ๐‘›}. (c) Av๐‘› (132, 213, 321) = {๐œ„๐‘˜1 โŠ– ๐œ„๐‘˜2 โˆฃ ๐‘˜1 + ๐‘˜2 = ๐‘›}. (d) Av๐‘› (132, 231, 312) = {๐›ฟ ๐‘˜1 โŠ• ๐œ„๐‘˜2 โˆฃ ๐‘˜1 + ๐‘˜2 = ๐‘›}. (e) Av๐‘› (132, 231, 321) = {(1 โŠ– ๐œ„๐‘˜1 ) โŠ• ๐œ„๐‘˜2 โˆฃ ๐‘˜1 + ๐‘˜2 = ๐‘› โˆ’ 1}. (f) Av๐‘› (123, 132, 213) = {๐œ„๐‘˜1 โŠ– ๐œ„๐‘˜2 โŠ– โ‹ฏ โˆฃ โˆ‘๐‘– ๐‘˜๐‘– = ๐‘› and ๐‘˜๐‘– โ‰ค 2 for all ๐‘–}. (40) Finish the proof of Lemma 1.12.4.

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

10.1090/gsm/210/02

Chapter 2

Counting with Signs

In the previous chapter, we concentrated on counting formulae where all of the terms were positive. But there are interesting things to say when one permits negative terms as well. This chapter is devoted to some of the principal techniques which one can use in such a situation.

2.1. The Principle of Inclusion and Exclusion The Principle of Inclusion and Exclusion, or PIE, is one of the classical methods for counting using signs. After presenting the Principle itself, we will give an application to derangements which are permutations having no fixed points. In the Sum Rule, Lemma 1.1.1(a), we assumed that the sets ๐‘†, ๐‘‡ are disjoint. Of course, it is easy to see that for any finite sets ๐‘†, ๐‘‡ we have (2.1)

|๐‘† โˆช ๐‘‡| = |๐‘†| + |๐‘‡| โˆ’ |๐‘† โˆฉ ๐‘‡|.

Indeed, |๐‘†| + |๐‘‡| counts ๐‘† โˆฉ ๐‘‡ twice and so to count it only once we must subtract the cardinality of the intersection. But one could ask if there is a similar formula for the union of any number of sets. It turns out that it is often more useful to consider these sets as subsets of some universal set ๐‘† and count the number of elements in ๐‘† which are not in any of the subsets, similar to the viewpoint used in pattern avoidance. To set up notation, let ๐‘† be a set and let ๐‘† 1 , . . . , ๐‘†๐‘› โŠ† ๐‘†. We wish to find a formula for |๐‘† โˆ’ โ‹ƒ๐‘– ๐‘† ๐‘– |. When ๐‘› = 1 we clearly have |๐‘† โˆ’ ๐‘† 1 | = |๐‘†| โˆ’ |๐‘† 1 |. And for ๐‘› = 2 equation (2.1) yields |๐‘† โˆ’ (๐‘† 1 โˆช ๐‘† 2 )| = |๐‘†| โˆ’ |๐‘† 1 | โˆ’ |๐‘† 2 | + |๐‘† 1 โˆฉ ๐‘† 2 |. 41 Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

42

2. Counting with Signs

๐‘†

๐‘† ๐‘†1

๐‘†1

๐‘†2

Figure 2.1. The PIE for ๐‘› = 1, 2

Venn diagrams showing the shaded region counted for these two cases are given in Figure 2.1. The reader may have already guessed the generalization for arbitrary ๐‘›. This type of enumeration where one alternately adds and subtracts cardinalities is sometimes called a sieve. Theorem 2.1.1 (Principle of Inclusion and Exclusion, PIE). If ๐‘† is a finite set with subsets ๐‘† 1 , . . . , ๐‘†๐‘› , then (2.2)

๐‘› | | |๐‘† โˆ’ ๐‘† ๐‘– || = |๐‘†| โˆ’ โˆ‘ |๐‘† ๐‘– | + | โ‹ƒ | | 1โ‰ค๐‘–โ‰ค๐‘› ๐‘–=1

๐‘›

โˆ‘

|๐‘† ๐‘– โˆฉ ๐‘†๐‘— | โˆ’ โ‹ฏ + (โˆ’1)๐‘› |

1โ‰ค๐‘– โ‹ฏ > ๐‘๐‘—โˆ’1 > ๐‘๐‘— = min ๐ต๐‘— . Now in ๐œŒโ€ฒ we have ๐ต๐‘–โ€ฒ = {๐‘๐‘– } for ๐‘– โ‰ค ๐‘— with ๐‘1 > โ‹ฏ > ๐‘๐‘— . As a consequence, no ๐ต๐‘–โ€ฒ โ€ฒ can be split or merged for ๐‘– < ๐‘— and so ๐œ„(๐œŒโ€ฒ ) will merge ๐ต๐‘—โ€ฒ into ๐ต๐‘—+1 . Thus ๐œ„(๐œŒโ€ฒ ) = ๐œŒ as desired. It is clear that ๐œ„ is sign reversing since ๐œ„(๐œŒ) has one more or one fewer block than ๐œŒ. So we just need to find the fixed points. But if ๐œŒ โˆˆ Fix ๐œ„, then all ๐œŒโ€™s blocks contain a single element; otherwise one could be split. It follows that ๐œŒ = ({๐‘1 }, . . . , {๐‘๐‘› }). Furthermore, none of the blocks can be merged and so ๐‘1 > โ‹ฏ > ๐‘๐‘› . But this forces our set composition to be ๐œŒ = ({๐‘›}, {๐‘› โˆ’ 1}, . . . , {1}) and sgn ๐œŒ = (โˆ’1)๐‘› , completing the proof. โ–ก To illustrate, suppose ๐‘› = 8. As we have done previously, we will dispense with brackets and commas in sets. Consider ๐œŒ = (๐ต1 , . . . , ๐ต5 ) = (5, 3, 147, 2, 68). Then ๐ต3 is splittable and splitting it results in ๐œŽ(๐œŒ) = (5, 3, 1, 47, 2, 68). Also, ๐ต4 can be merged into ๐ต5 in ๐œŒ since ๐ต4 = {2} and 2 < min ๐ต5 = 6. Merging these two blocks gives ๐œ‡(๐œŒ) = (5, 3, 147, 268). To decide which operation to use we start with ๐ต1 . It cannot be split, having only one element. And it cannot be merged with ๐ต2 since 5 > min ๐ต2 = 3. Similarly ๐ต2 cannot be split or merged with ๐ต3 . But we have already seen that ๐ต3 can be split so that ๐œ„(๐œŒ) = (5, 3, 1, 47, 2, 68) = ๐œŒโ€ฒ . To check that ๐œ„(๐œŒโ€ฒ ) = ๐œŒ is similar. Involutions involving merging and splitting often come up when finding formulae for antipodes in Hopf algebras. One can consult the papers of Benedetti-Bergeron [7], Benedetti-Hallam-Machacek [8], Benedetti-Sagan [9], or Bergeron-Ceballos [12] for examples. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

2.3. The Garsiaโ€“Milne Involution Principle

49

2.3. The Garsiaโ€“Milne Involution Principle So far we have used sign-reversing involutions to explain cancellation in alternating sums. But can they also furnish a bijection for proving that two given sets have the same cardinality? The answer in certain cases is โ€œyesโ€ and the standard technique for doing this is called the Garsiaโ€“Milne Involution Principle. Garsia and Milne [30] introduced this method to give the first bijective proof of the Rogersโ€“Remanujan identities, famous formulas which involve certain sets of integer partitions. Since then the Involution Principle has found a number of other applications. See, for example, the articles of Remmel [73] or Wilf [100]. In order to prove the Garsiaโ€“Milne result, we will need a version of Lemma 1.9.5 which applies to a slightly wider class of digraphs. Since the demonstration of the next result is similar to that of the earlier one, we leave the pleasure of proving it to the reader. Lemma 2.3.1. Let ๐ท = (๐‘‰, ๐ด) be a digraph. We have odeg ๐‘ฃ, ideg ๐‘ฃ โ‰ค 1 for all ๐‘ฃ โˆˆ ๐‘‰ if and only if ๐ท is a disjoint union of directed paths and directed cycles. โ–ก The basic idea of the Involution Principle is that, under suitable conditions, if one has two signed sets each with their own sign-reversing involution, then we can use a bijection between these sets to create a bijection between their fixed-point sets. So let ๐‘† and ๐‘‡ be disjoint signed sets with sign-reversing involutions ๐œ„ โˆถ ๐‘† โ†’ ๐‘† and ๐œ… โˆถ ๐‘‡ โ†’ ๐‘‡ such that Fix ๐œ„ โŠ† ๐‘† + and Fix ๐œ… โŠ† ๐‘‡ + . Furthermore, suppose we have a bijection ๐‘“ โˆถ ๐‘† โ†’ ๐‘‡ which preserves signs in that sgn ๐‘“(๐‘ ) = sgn ๐‘  for all ๐‘  โˆˆ ๐‘†. A picture of this setup can be found in Figure 2.3. Note that although all arrows are really doubleheaded, we have only shown them in one direction because of what is to come. And the circular arrows on the fixed points have been ignored. We now construct a map ๐น โˆถ Fix ๐œ„ โ†’ Fix ๐œ… as follows. To define ๐น(๐‘ ) for ๐‘  โˆˆ Fix ๐œ„ we first compute ๐‘“(๐‘ ) โˆˆ ๐‘‡ + . If ๐‘“(๐‘ ) โˆˆ Fix ๐œ…, then we let ๐น(๐‘ ) = ๐‘“(๐‘ ). If not, we apply the functional composition ๐œ™ = ๐‘“ โˆ˜ ๐œ„ โˆ˜ ๐‘“โˆ’1 โˆ˜ ๐œ… to ๐‘“(๐‘ ). Remembering that we compose from right to left, this takes ๐‘“(๐‘ ) to ๐‘‡ โˆ’ , ๐‘† โˆ’ , ๐‘† + , and ๐‘‡ + in that order. If this brings us to an element of Fix ๐œ…, then we let ๐น(๐‘ ) = ๐œ™(๐‘“(๐‘ )). Otherwise we apply ๐œ™ as many times as necessary, say ๐‘š, to arrive at an element of Fix ๐œ… and define ๐น(๐‘ ) = ๐œ™๐‘š (๐‘“(๐‘ )).

(2.8)

Continuing the example in Figure 2.3 we see that ๐‘“(๐‘ ) = ๐‘ข โˆ‰ Fix ๐œ…. So we apply ๐œ™, which takes ๐‘ข to ๐‘ฃ, ๐‘Ÿ, ๐‘ž, and ๐‘ก in turn. Since ๐‘ก โˆˆ Fix ๐œ… we let ๐น(๐‘ ) = ๐‘ก. Of course, we have to worry whether this is all well-defined; e.g., does ๐‘š always exist? And we also need to prove that ๐น is a bijection. This is taken care of by the next theorem. Theorem 2.3.2 (Garsiaโ€“Milne Involution Principle). With the notation of the previous paragraph, the map ๐น โˆถ Fix ๐œ„ โ†’ Fix ๐œ… is a well-defined bijection. Proof. Recall the notion of a functional digraph as used in the proof from Section 1.9 of Theorem 1.5.1. Define the following functions by restriction of their domains: ๐‘“ = ๐‘“|๐‘†+ ,

๐‘” = ๐‘“โˆ’1 |๐‘‡ โˆ’ ,

๐œ„ = ๐œ„|๐‘†โˆ’ ,

๐œ… = ๐œ…|๐‘‡ + โˆ’Fix ๐œ… .

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

50

2. Counting with Signs

๐‘†+

๐‘‡+

๐‘ž

๐‘“

๐œ„

๐‘  Fix ๐œ„

๐‘Ÿ

๐‘ข

๐‘“

๐œ…

๐‘ก Fix ๐œ…

๐‘ฃ

๐‘“โˆ’1 ๐‘†โˆ’

๐‘‡โˆ’ Figure 2.3. An example of the Garsiaโ€“Milne construction

Consider ๐ท which is the union of the functional digraphs for ๐‘“, ๐‘”, ๐œ„, and ๐œ…. It is easy to verify from the definitions that ๐‘ฅ โˆˆ ๐‘‰(๐ท) has in-degrees and out-degrees given by the following table depending on the subset of ๐‘† โˆช ๐‘‡ containing ๐‘ฅ: subset Fix ๐œ„ Fix ๐œ… (๐‘† โˆ’ Fix ๐œ„) โˆช (๐‘‡ โˆ’ Fix ๐œ…)

odeg ๐‘ฅ 1 0 1

ideg ๐‘ฅ 0 1 1

For example, if ๐‘ฅ โˆˆ Fix ๐œ„, then the only arc containing ๐‘ฅ comes from ๐‘“ and so odeg ๐‘ฅ = 1 and ideg ๐‘ฅ = 0. On the other hand, if ๐‘ฅ โˆˆ ๐‘† + โˆ’ Fix ๐œ„, then ๐‘ฅ has an arc going out from ๐‘“ and one coming in from ๐œ„ giving odeg ๐‘ฅ = ideg ๐‘ฅ = 1. Now ๐ท satisfies the hypothesis of the forward direction of Lemma 2.3.1. It follows that ๐ท is a disjoint union of directed paths and directed cycles. Each directed path must start at a vertex with out-degree 1 and in-degree 0 and end at a vertex with these degrees switched. Furthermore, all other vertices have out-degree and in-degree both 1. From these observations and the chart, it follows that these paths define a 1-to-1 correspondence between the vertices of Fix ๐œ„ and those of Fix ๐œ…. Furthermore, from the definition of ๐ท we see that each path corresponds exactly to a functional composition ๐œ™๐‘š ๐‘“(๐‘ ) for ๐‘  โˆˆ Fix ๐œ„ and some ๐‘š โ‰ฅ 0. So ๐น is the bijection defined by these paths. โ–ก

Before we give an application of the previous theorem, we should mention an approach which can be useful in setting up the necessary sets and bijections. Here is one way to try to find a bijection ๐น โˆถ ๐‘‹ โ†’ ๐‘Œ between two finite sets ๐‘‹, ๐‘Œ . (i) As with the PIE, construct a set ๐ด with subsets ๐ด1 , . . . , ๐ด๐‘› such that ๐‘‹ = ๐ด โˆ’ โ‹ƒ ๐ด๐‘– . Similarly construct ๐ต and ๐ต1 , . . . , ๐ต๐‘› for ๐‘Œ . (ii) Use the method of our second proof of the PIE to set up a sign-reversing involution ๐œ„ on the set ๐‘† as given by (2.6). Similarly construct ๐œ… on a set ๐‘‡. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

2.3. The Garsiaโ€“Milne Involution Principle

51

(iii) Find a bijection ๐‘“ โˆถ ๐‘† โ†’ ๐‘‡ of the form ๐‘“(๐‘Ž, ๐ผ) = (๐‘, ๐ผ) which is well-defined in that ๐‘Ž โˆˆ ๐ด๐ผ if and only if ๐‘ โˆˆ ๐ต๐ผ . Recall that Fix ๐œ„ = (๐‘Ž, โˆ…) where ๐‘Ž โˆˆ ๐ด โˆ’ โ‹ƒ ๐ด๐‘– . Thus Fix ๐œ„ โŠ† ๐‘† + as needed to apply the Involution Principle, and there is a natural bijection between Fix ๐œ„ and ๐‘‹. Note also that ๐‘“ is automatically sign preserving since sgn(๐‘Ž, ๐ผ) = (โˆ’1)#๐ผ = sgn(๐‘, ๐ผ). So once these three steps have been accomplished, Theorem 2.3.2 guarantees that we have a bijection ๐‘‹ โ†’ ๐‘Œ . As already remarked, the Involution Principle is useful in proving integer partition identities. Say that partition ๐œ† = (๐œ†1 , ๐œ†2 , . . . , ๐œ†๐‘˜ ) has distinct parts if ๐œ†1 > ๐œ†2 > โ‹ฏ > ๐œ†๐‘˜ (as opposed to the usual weakly decreasing condition). On the other hand, say that ๐œ† has odd parts if all the ๐œ†๐‘– are odd. The next result is a famous theorem of Euler. As has become traditional, an example follows the proof. Theorem 2.3.3 (Euler). Let ๐‘ƒ๐‘‘ (๐‘›) be the set of partitions of ๐‘› with distinct parts and let ๐‘ƒ๐‘œ (๐‘›) be the set of partitions of ๐‘› with odd parts. For ๐‘› โ‰ฅ 0 we have #๐‘ƒ๐‘‘ (๐‘›) = #๐‘ƒ๐‘œ (๐‘›).

Proof. It suffices to show that there is a bijection ๐‘ƒ๐‘‘ (๐‘›) โ†’ ๐‘ƒ๐‘œ (๐‘›). To apply the PIE to ๐‘ƒ๐‘‘ (๐‘›) we can take ๐ด = ๐‘ƒ(๐‘›), the set of all partitions of ๐‘›, with subsets ๐ด1 , . . . , ๐ด๐‘› where ๐ด๐‘– = {๐œ† โŠข ๐‘› โˆฃ ๐œ† has (at least) two copies of the part ๐‘–}. Note that ๐ด๐‘– = โˆ… if ๐‘– > ๐‘›/2, but this does no harm and keeps the notation simple. It should be clear from the definitions that ๐‘ƒ๐‘‘ (๐‘›) = ๐ด โˆ’ โ‹ƒ ๐ด๐‘– . Similarly, for ๐‘ƒ๐‘œ (๐‘›) we let ๐ต = ๐‘ƒ(๐‘›) with subsets ๐ต๐‘– = {๐œ‡ โŠข ๐‘› โˆฃ ๐œ‡ has a part of the form 2๐‘–} for 1 โ‰ค ๐‘– โ‰ค ๐‘›. Again, it is easy to see that ๐‘ƒ๐‘œ (๐‘›) = ๐ต โˆ’ โ‹ƒ ๐ต๐‘– . The construction of ๐‘†, ๐œ„, ๐‘‡, and ๐œ… are now exactly the same as in the second proof of the PIE. So it suffices to construct an appropriate bijection ๐‘“ โˆถ ๐‘† โ†’ ๐‘‡. Given (๐œ†, ๐ผ) โˆˆ ๐‘†, we replace, for each ๐‘– โˆˆ ๐ผ, a pair of ๐‘–โ€™s in ๐œ† by a part 2๐‘– to form ๐œ‡. So if ๐œ† โˆˆ ๐ด๐‘– , then ๐œ‡ โˆˆ ๐ต๐‘– for all ๐‘– โˆˆ ๐ผ and the map ๐‘“(๐œ†, ๐ผ) = (๐œ‡, ๐ผ) is well-defined. It is also easy to construct ๐‘“โˆ’1 , taking an even part 2๐‘– in ๐œ‡ and replacing it with two copies of ๐‘– to form ๐œ† as ๐‘– runs over ๐ผ. Appealing to Theorem 2.3.2 finishes the proof. โ–ก Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

52

2. Counting with Signs

To illustrate this demonstration, suppose we start with (6, 2, 1) โˆˆ ๐‘ƒ๐‘‘ (9). For the pairs in ๐‘† and ๐‘‡, we will dispense with delimiters and commas as usual. So ๐‘“

๐‘“โˆ’1

๐œ…

๐œ„

(621, โˆ…) โ†ฆ (621, โˆ…) โ†ฆ (621, 3) โ†ฆ (3321, 3) โ†ฆ (3321, โˆ…) ๐‘“

๐œ…

๐‘“

๐œ…

๐‘“

๐œ…

๐‘“โˆ’1

๐œ„

โ†ฆ (3321, โˆ…) โ†ฆ (3321, 1) โ†ฆ (33111, 1) โ†ฆ (33111, 13) ๐‘“โˆ’1

๐œ„

โ†ฆ (621, 13) โ†ฆ (621, 1) โ†ฆ (6111, 1) โ†ฆ (6111, โˆ…) ๐‘“โˆ’1

๐œ„

โ†ฆ (6111, โˆ…) โ†ฆ (6111, 3) โ†ฆ (33111, 3) โ†ฆ (33111, โˆ…) ๐‘“

โ†ฆ (33111, โˆ…). ๐น

It follows that we should map (6, 2, 1) โ†ฆ (3, 3, 1, 1, 1). Clearly one might like to find a more efficient bijection if one exists. This issue will be further explored in the exercises.

2.4. The Reflection Principle The Reflection Principle is a geometric method for working with certain combinatorial problems involving lattice paths. In particular, it will permit us to give a very simple proof of the binomial coefficient formula for the Catalan numbers. It is also useful in proving unimodality, an interesting property of real number sequences. Consider the integer lattice โ„ค2 and northeast paths in this lattice. Suppose we are given a line in the plane of the form ๐ฟ โˆถ ๐‘ฆ = ๐‘ฅ + ๐‘ for some ๐‘ โˆˆ โ„ค. Note that the reflection in ๐ฟ of any northeast path is again a northeast path. If ๐‘ƒ is a path from ๐‘ข to ๐‘ƒ

๐‘ฃ, then we write ๐‘ƒ โˆถ ๐‘ข โ†’ ๐‘ฃ or ๐‘ข โ†’ ๐‘ฃ. Suppose ๐‘ƒ โˆถ ๐‘ข โ†’ ๐‘ฃ intersects ๐ฟ and let ๐‘ฅ be its

๐‘ฃ

๐ฟ

๐ฟ

๐‘ƒ2 ๐‘ฃโ€ฒ ๐‘ฅ

๐‘ฅ

ฮฅ๐ฟ

โ†’

๐‘ข

๐‘ƒ1

๐‘ƒ2โ€ฒ

๐‘ข

Figure 2.4. The map ฮฅ๐ฟ

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

๐‘ƒ1

2.4. The Reflection Principle

53

last (northeast-most) point of intersection. Then ๐‘ƒ can be written as the concatenation ๐‘ƒ1

๐‘ƒ2

๐‘ƒ โˆถ ๐‘ข โ†’ ๐‘ฅ โ†’ ๐‘ฃ. For example, on the left in Figure 2.4 we have the path ๐‘ƒ = ๐ธ๐ธ๐ธ๐‘๐‘๐‘๐‘๐ธ๐‘ with ๐‘ƒ1 = ๐ธ๐ธ๐ธ๐‘๐‘ and ๐‘ƒ2 = ๐‘๐‘๐ธ๐‘. We now define a new path ๐‘ƒ1

๐‘ƒ2โ€ฒ

ฮฅ๐ฟ (๐‘ƒ) โˆถ ๐‘ข โ†’ ๐‘ฅ โ†’ ๐‘ฃโ€ฒ where ๐‘ƒ2โ€ฒ and ๐‘ฃโ€ฒ are the reflections of ๐‘ƒ2 and ๐‘ฃ in ๐ฟ, respectively. Returning to our example, ๐‘ƒ2โ€ฒ = ๐ธ๐ธ๐‘๐ธ, which is obtained from ๐‘ƒ2 by merely interchanging north and east steps. So ฮฅ๐ฟ (๐‘ƒ) = ๐ธ๐ธ๐ธ๐‘๐‘๐ธ๐ธ๐‘๐ธ as on the right in Figure 2.4. This is the fundamental map for using the Reflection Principle. To state it precisely, let ๐’ฉโ„ฐ(๐‘ข; ๐‘ฃ) denote the set of northeast paths from ๐‘ข to ๐‘ฃ and let ๐’ฉโ„ฐ๐ฟ (๐‘ข; ๐‘ฃ) be the subset of paths which intersect ๐ฟ. If ๐‘ข is omitted, then it is assumed that ๐‘ข = (0, 0). Also, be sure to distinguish the notation ๐’ฉโ„ฐ(๐‘ข; ๐‘ฃ) for the northeast paths from ๐‘ข to ๐‘ฃ and ๐’ฉโ„ฐ(๐‘š, ๐‘›) for the northeast paths from (0, 0) to (๐‘š, ๐‘›). The former contains a semicolon where the latter has a comma. Theorem 2.4.1 (Reflection Principle). Given ๐ฟ โˆถ ๐‘ฆ = ๐‘ฅ + ๐‘ for ๐‘ โˆˆ โ„ค and ๐‘ฃ โˆˆ โ„ค2 , we let ๐‘ฃโ€ฒ be the reflection of ๐‘ฃ in ๐ฟ. Then the map ฮฅ๐ฟ โˆถ ๐’ฉโ„ฐ๐ฟ (๐‘ข; ๐‘ฃ) โ†’ ๐’ฉโ„ฐ๐ฟ (๐‘ข; ๐‘ฃโ€ฒ ) is a bijection. Proof. In fact, we can show that ฮฅ๐ฟ is an involution on ๐’ฉโ„ฐ๐ฟ (๐‘ข; ๐‘ฃ) โˆช ๐’ฉโ„ฐ๐ฟ (๐‘ข; ๐‘ฃโ€ฒ ). This follows from the fact that reflection in ๐ฟ is an involution and that the set of intersection points does not change when passing from ๐‘ƒ โˆฉ ๐ฟ to ฮฅ๐ฟ (๐‘ƒ) โˆฉ ๐ฟ. โ–ก As a first application of Theorem 2.4.1, we will give a simpler, although not as purely combinatorial, proof of Theorem 1.11.3. We restate the formula here for reference: 1 2๐‘› ๐ถ(๐‘›) = ( ). ๐‘›+1 ๐‘› Proof. Recall that ๐ถ(๐‘›) counts the set ๐’Ÿ(๐‘›) of northeast Dyck paths from (0, 0) to (๐‘›, ๐‘›). From Theorem 1.11.1 we know that the total number of all northeast paths ๐‘ƒ from the origin to (๐‘›, ๐‘›) is #๐’ฉโ„ฐ(๐‘›, ๐‘›) = (

2๐‘› ). ๐‘›

Note that ๐‘ƒ does not stay weakly above ๐‘ฆ = ๐‘ฅ if and only if ๐‘ƒ intersects the line ๐ฟ โˆถ ๐‘ฆ = ๐‘ฅ โˆ’ 1. And by the Reflection Principle, such paths are in bijection with ๐’ฉโ„ฐ๐ฟ ((0, 0); (๐‘› + 1, ๐‘› โˆ’ 1)) since (๐‘› + 1, ๐‘› โˆ’ 1) is the reflection of (๐‘›, ๐‘›) in ๐ฟ. But all paths from (0, 0) to (๐‘› + 1, ๐‘› โˆ’ 1) cross ๐ฟ since these two points are on opposite sides of the line. Thus, using Theorem 1.11.1 again, #๐’ฉโ„ฐ๐ฟ ((0, 0); (๐‘› + 1, ๐‘› โˆ’ 1)) = #๐’ฉโ„ฐ(๐‘› + 1, ๐‘› โˆ’ 1) = (

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

2๐‘› ). ๐‘›+1

54

2. Counting with Signs

So subtracting the number of non-Dyck paths from the total number of paths in ๐’ฉโ„ฐ(๐‘›, ๐‘›) gives ๐ถ(๐‘›) = ( =

2๐‘› 2๐‘› )โˆ’( ) ๐‘› ๐‘›+1

(2๐‘›)! (2๐‘›)! โˆ’ ๐‘›! ๐‘›! (๐‘› + 1)! (๐‘› โˆ’ 1)!

= (1 โˆ’

=

2๐‘› ๐‘› )( ) ๐‘›+1 ๐‘›

1 2๐‘› ( ) ๐‘›+1 ๐‘› โ–ก

as desired.

The Reflection Principle can also be used to prove that certain sequences have a property called unimodality. A sequence of real numbers ๐‘Ž0 , ๐‘Ž1 , . . . , ๐‘Ž๐‘› is said to be unimodal if there is an index ๐‘š such that ๐‘Ž0 โ‰ค ๐‘Ž1 โ‰ค โ‹ฏ โ‰ค ๐‘Ž๐‘š โ‰ฅ ๐‘Ž๐‘š+1 โ‰ฅ โ‹ฏ โ‰ฅ ๐‘Ž๐‘› . So this is the next most complicated behavior after being weakly increasing or weakly decreasing. In fact the latter are the special cases of unimodality where ๐‘š = ๐‘› or ๐‘š = 0. Many sequences arising in combinatorics, algebra, and geometry are unimodal. See the survey articles of Stanley [89], Brenti [20], or Brรคndรฉn [19] for more details. The term โ€œunimodalโ€ comes from probability and statistics where one thinks of the ๐‘Ž๐‘– as giving you the distribution of a random variable taking values in {0, 1, . . . , ๐‘›}. Then a unimodal distribution has only one hump. We have already met a number of unimodal sequences, although we have not remarked on the fact. Here is the simplest. Theorem 2.4.2. For ๐‘› โ‰ฅ 0 the sequence ๐‘› ๐‘› ๐‘› ( ), ( ), . . . , ( ) 0 1 ๐‘› is unimodal. Proof. Because the binomial coefficients are symmetric, Theorem 1.3.3(b), it suffices to prove that this sequence is increasing up to its halfway point. So we want to show ๐‘› ๐‘› ( )โ‰ค( ) ๐‘˜ ๐‘˜+1 for ๐‘˜ < โŒŠ๐‘›/2โŒ‹. From Theorem 1.11.1, we know that ๐‘› ( ) = #๐’ฉโ„ฐ(๐‘˜, ๐‘› โˆ’ ๐‘˜) ๐‘˜

and

(

๐‘› ) = #๐’ฉโ„ฐ(๐‘˜ + 1, ๐‘› โˆ’ ๐‘˜ โˆ’ 1). ๐‘˜+1

So it suffices to find an injection ๐‘– โˆถ ๐’ฉโ„ฐ(๐‘˜, ๐‘› โˆ’ ๐‘˜) โ†’ ๐’ฉโ„ฐ(๐‘˜ + 1, ๐‘› โˆ’ ๐‘˜ โˆ’ 1). Let ๐ฟ be the perpendicular bisector of the line segment from (๐‘˜, ๐‘›โˆ’๐‘˜) to (๐‘˜+1, ๐‘›โˆ’๐‘˜โˆ’1). It is easy to Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

2.5. The Lindstrรถmโ€“Gesselโ€“Viennot Lemma

55

check that ๐ฟ has the form ๐‘ฆ = ๐‘ฅ + ๐‘ for ๐‘ โˆˆ โ„ค. From the Reflection Principle, we have a bijection ฮฅ๐ฟ โˆถ ๐’ฉโ„ฐ๐ฟ (๐‘˜, ๐‘› โˆ’ ๐‘˜) โ†’ ๐’ฉโ„ฐ๐ฟ (๐‘˜ + 1, ๐‘› โˆ’ ๐‘˜ โˆ’ 1). But since ๐‘˜ < โŒŠ๐‘›/2โŒ‹ the lattice points (0, 0) and (๐‘˜, ๐‘›โˆ’๐‘˜) are on opposite sides of ๐ฟ so that ๐’ฉโ„ฐ๐ฟ (๐‘˜, ๐‘›โˆ’๐‘˜) = ๐’ฉโ„ฐ(๐‘˜, ๐‘›โˆ’๐‘˜). Furthermore ๐’ฉโ„ฐ๐ฟ (๐‘˜ + 1, ๐‘› โˆ’ ๐‘˜ โˆ’ 1) โŠ† ๐’ฉโ„ฐ(๐‘˜ + 1, ๐‘› โˆ’ ๐‘˜ โˆ’ 1). So extending the range of ฮฅ๐ฟ provides the desired injection. โ–ก It turns out that the Stirling number sequences ๐‘(๐‘›, 0), ๐‘(๐‘›, 1), . . . , ๐‘(๐‘›, ๐‘›)

and

๐‘†(๐‘›, 0), ๐‘†(๐‘›, 1), . . . , ๐‘†(๐‘›, ๐‘›)

are also unimodal. But this is not so easy to prove directly. One reason for this is that these sequences are not symmetric like the one for the binomial coefficients. And there is no known simple expression for the index ๐‘š where they achieve their maxima. Instead it is better to use another property of real sequences, called log-concavity, which can imply unimodality. This is one of the motivations for the next section.

2.5. The Lindstrรถmโ€“Gesselโ€“Viennot Lemma The lemma in question is a powerful technique for dealing with certain determinantal identities. It was first discovered by Lindstrรถm [57] and then used to great effect by Gessel and Viennot [31] as well as many other authors. Like the Reflection Principle, this method uses directed paths. On the other hand, it uses multiple paths and is not restricted to the integer lattice. In particular, when there are two paths, then log-concavity results can be obtained. A sequence of real numbers ๐‘Ž0 , ๐‘Ž1 , . . . , ๐‘Ž๐‘› is called log-concave if, for all 0 < ๐‘˜ < ๐‘›, we have (2.9)

๐‘Ž2๐‘˜ โ‰ฅ ๐‘Ž๐‘˜โˆ’1 ๐‘Ž๐‘˜+1 .

As usual, we can extend this to all ๐‘˜ โˆˆ โ„ค by letting ๐‘Ž๐‘˜ = 0 for ๐‘˜ < 0 or ๐‘˜ > ๐‘›. Logconcave sequences, like unimodal ones, are ubiquitous in combinatorics, algebra, and geometry. See the previously cited survey articles of Stanley, Brenti, and Brรคndรฉn for details. For example, a row of Pascalโ€™s triangle or either of the Stirling triangles is logconcave. The name โ€œlog-concaveโ€ comes from the following scenario. Suppose that we have a function ๐‘“ โˆถ โ„ โ†’ โ„ which is concave down. So if one takes any two points on the graph of ๐‘“, then the line segment connecting them lies weakly below ๐‘“. Taking the points to be (๐‘˜ โˆ’ 1, ๐‘“(๐‘˜ โˆ’ 1)) and (๐‘˜ + 1, ๐‘“(๐‘˜ + 1)) and comparing the ๐‘ฆ-coordinate of the midpoint of the corresponding line segment with that coordinate on ๐‘“ gives (๐‘“(๐‘˜ โˆ’ 1) + ๐‘“(๐‘˜ + 1))/2 โ‰ค ๐‘“(๐‘˜). Now if ๐‘“(๐‘ฅ) > 0 for all ๐‘ฅ and the function log ๐‘“(๐‘ฅ) is concave down, then substituting into the previous inequality and exponentiating gives ๐‘“(๐‘˜ โˆ’ 1)๐‘“(๐‘˜ + 1) โ‰ค ๐‘“(๐‘˜)2 just like the definition of log-concavity for sequences. It turns out that log-concavity and unimodality are related. Proposition 2.5.1. Suppose that ๐‘Ž0 , ๐‘Ž1 , . . . , ๐‘Ž๐‘› is a sequence of positive reals. If the sequence is log-concave, then it is unimodal. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

56

2. Counting with Signs

Proof. To show a sequence is unimodal it suffices to show that after its first strict decrease, then it continues to weakly decrease. But ๐‘Ž๐‘˜โˆ’1 > ๐‘Ž๐‘˜ is equivalent to ๐‘Ž๐‘˜โˆ’1 /๐‘Ž๐‘˜ > 1 for positive ๐‘Ž๐‘˜ . Rewriting (2.9) as ๐‘Ž๐‘˜ /๐‘Ž๐‘˜+1 โ‰ฅ ๐‘Ž๐‘˜โˆ’1 /๐‘Ž๐‘˜ we see that if ๐‘Ž๐‘˜โˆ’1 /๐‘Ž๐‘˜ > 1, then ๐‘Ž๐‘™โˆ’1 /๐‘Ž๐‘™ > 1 for all ๐‘™ โ‰ฅ ๐‘˜. So the sequence is unimodal. โ–ก Even though log-concavity implies unimodality for positive sequences, it is paradoxically often easier to prove log-concavity rather than proving unimodality directly. This comes in part from the fact that the log-concave condition is a uniform one for all ๐‘˜, as opposed to unimodality where one must know where the maximum of the sequence occurs. We can rewrite (2.9) as ๐‘Ž2๐‘˜ โˆ’ ๐‘Ž๐‘˜โˆ’1 ๐‘Ž๐‘˜+1 โ‰ฅ 0, or in terms of determinants as | ๐‘Ž๐‘˜ | | ๐‘Ž๐‘˜โˆ’1

(2.10)

๐‘Ž๐‘˜+1 | | โ‰ฅ 0. ๐‘Ž๐‘˜ |

To prove that the determinant is nonnegative, we could show that it counts something and that is exactly what the Lindstrรถmโ€“Gesselโ€“Viennot Lemma is set up to do. We will first consider the case of 2 ร— 2 determinants and at the end of the section indicate how to do the general case. As a running example, we will show how to prove log-concavity of the sequence of binomial coefficients considered in Theorem 2.4.2. Let ๐ท be a digraph which is acyclic in that it contains no directed cycles. Given two vertices of ๐‘ข, ๐‘ฃ โˆˆ ๐‘‰(๐ท), we let ๐’ซ(๐‘ข; ๐‘ฃ) denote the set of directed paths from ๐‘ข to ๐‘ฃ. We will assume that ๐‘ข, ๐‘ฃ are always chosen so that ๐‘(๐‘ข; ๐‘ฃ) = #๐’ซ(๐‘ข; ๐‘ฃ) is finite even if ๐ท itself is not. To illustrate, let ๐ท be the digraph with vertices โ„ค2 and arcs from (๐‘š, ๐‘›) to (๐‘š + 1, ๐‘›) and to (๐‘š, ๐‘› + 1) for all ๐‘š, ๐‘› โˆˆ โ„ค. Then ๐’ซ(๐‘ข; ๐‘ฃ) is just the set of northeast lattice paths from ๐‘ข to ๐‘ฃ, denoted ๐’ฉโ„ฐ(๐‘ข; ๐‘ฃ) in the previous section. We will continue to use the notation for general paths from that section for any acyclic digraph. We also extend that notation as follows. Given a directed path ๐‘ƒ โˆถ ๐‘ข โ†’ ๐‘ฃ and vertices ๐‘ฅ coming ๐‘ƒ

before ๐‘ฆ on ๐‘ƒ, we let ๐‘ฅ โ†’ ๐‘ฆ be the portion of ๐‘ƒ between ๐‘ฅ and ๐‘ฆ. Continuing the general exposition, suppose we are given ๐‘ข1 , ๐‘ข2 โˆˆ ๐‘‰ called the initial vertices and ๐‘ฃ 1 , ๐‘ฃ 2 โˆˆ ๐‘‰ which are the final vertices. We wish to consider determinants of the form (2.11)

| ๐‘(๐‘ข1 ; ๐‘ฃ 1 ) ๐‘(๐‘ข1 ; ๐‘ฃ 2 ) | | | = ๐‘(๐‘ข1 ; ๐‘ฃ 1 )๐‘(๐‘ข2 ; ๐‘ฃ 2 ) โˆ’ ๐‘(๐‘ข1 ; ๐‘ฃ 2 )๐‘(๐‘ข2 ; ๐‘ฃ 1 ). | ๐‘(๐‘ข2 ; ๐‘ฃ 1 ) ๐‘(๐‘ข2 ; ๐‘ฃ 2 ) |

Note that ๐‘(๐‘ข1 ; ๐‘ฃ 1 )๐‘(๐‘ข2 ; ๐‘ฃ 2 ) counts pairs of paths (๐‘ƒ1 , ๐‘ƒ2 ) โˆˆ ๐’ซ(๐‘ข1 ; ๐‘ฃ 1 ) ร— ๐’ซ(๐‘ข2 ; ๐‘ฃ 2 ) โ‰” ๐’ซ12 and similarly for ๐‘(๐‘ข1 ; ๐‘ฃ 2 )๐‘(๐‘ข2 ; ๐‘ฃ 1 ) and ๐’ซ(๐‘ข1 ; ๐‘ฃ 2 ) ร— ๐’ซ(๐‘ข2 ; ๐‘ฃ 1 ) โ‰” ๐’ซ21 . Returning to our example, if we wish to show 2

๐‘› ๐‘› ๐‘› ( ) โˆ’( )( ) โ‰ฅ 0, ๐‘˜ ๐‘˜โˆ’1 ๐‘˜+1 Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

2.5. The Lindstrรถmโ€“Gesselโ€“Viennot Lemma

57

๐‘ฃ2

๐‘ฃ2 ๐‘ฃ1

๐‘ฃ1

โ„ฆ

โ†’

๐‘ฅ

๐‘ข2

๐‘ฅ

๐‘ข2 ๐‘ข1

๐‘ข1 Figure 2.5. The Lindstrรถmโ€“Gesselโ€“Viennot Involution

then we could take ๐‘ข1 = (1, 0), ๐‘ข2 = (0, 1), ๐‘ฃ 1 = (๐‘˜ + 1, ๐‘› โˆ’ ๐‘˜), ๐‘ฃ 2 = (๐‘˜, ๐‘› โˆ’ ๐‘˜ + 1). ๐‘› It follows from Theorem 1.11.1 that ๐‘(๐‘ข1 ; ๐‘ฃ 1 ) = ๐‘(๐‘ข2 ; ๐‘ฃ 2 ) = (๐‘›๐‘˜), while ๐‘(๐‘ข1 ; ๐‘ฃ 2 ) = (๐‘˜โˆ’1 ) ๐‘› and ๐‘(๐‘ข2 ; ๐‘ฃ 1 ) = (๐‘˜+1). More specifically, if ๐‘› = 7 and ๐‘˜ = 3, then in Figure 2.5 we have a pair of paths in ๐’ซ21 counted by (27)(47) on the left and another pair in ๐’ซ12 counted by 2

(37) on the right. For readablity, the grid for the integer lattice has been suppressed, leaving only the vertices of โ„ค2 . To prove that the determinant (2.11) is nonnegative, we will construct a signreversing involution ฮฉ on the set ๐’ซ โ‰” ๐’ซ12 โˆช ๐’ซ21 where sgn(๐‘ƒ1 , ๐‘ƒ2 ) = {

+1 if (๐‘ƒ1 , ๐‘ƒ2 ) โˆˆ ๐’ซ12 , โˆ’1 if (๐‘ƒ1 , ๐‘ƒ2 ) โˆˆ ๐’ซ21 .

We will construct ฮฉ so that every pair in ๐’ซ21 is in a 2-cycle with a pair in ๐’ซ12 . Furthermore, the remaining fixed points in ๐’ซ12 will be exactly the path pairs in ๐’ซ which do not intersect. It follows that (2.11) is just the number of nonintersecting path pairs in ๐’ซ and therefore must be nonnegative. To define ฮฉ, consider a path pair (๐‘ƒ1 , ๐‘ƒ2 ) โˆˆ ๐’ซ. If ๐‘ƒ1 โˆฉ ๐‘ƒ2 is empty, then this pair is in ๐’ซ12 , since every pair in ๐’ซ21 intersects. So in this case we let ฮฉ(๐‘ƒ1 , ๐‘ƒ2 ) = (๐‘ƒ1 , ๐‘ƒ2 ), a fixed point. If ๐‘ƒ1 โˆฉ ๐‘ƒ2 โ‰  โˆ…, then consider the list of intersections ๐‘ฅ1 , . . . , ๐‘ฅ๐‘ก in the order in which they are encountered on ๐‘ƒ1 . We claim they must also be encountered in this order on ๐‘ƒ2 . For if there were intersections ๐‘ฅ, ๐‘ฆ such that ๐‘ฅ comes before ๐‘ฆ on ๐‘ƒ1 and ๐‘ฆ ๐‘ƒ1

๐‘ƒ2

comes before ๐‘ฅ on ๐‘ƒ2 , then one can show that the directed walk ๐‘ฅ โ†’ ๐‘ฆ โ†’ ๐‘ฅ contains a directed cycle, as the reader will be asked to do in the exercises. This contradicts the assumption that ๐ท is acyclic. So there is a well-defined notion of a first intersection Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

58

2. Counting with Signs

๐‘ฅ = ๐‘ฅ1 . We now let ฮฉ(๐‘ƒ1 , ๐‘ƒ2 ) = (๐‘ƒ1โ€ฒ , ๐‘ƒ2โ€ฒ ) where ๐‘ƒ1

๐‘ƒ2

๐‘ƒ2

๐‘ƒ1

๐‘ƒ1โ€ฒ = ๐‘ข1 โ†’ ๐‘ฅ โ†’ ๐‘ฃ 2 , ๐‘ƒ2โ€ฒ = ๐‘ข2 โ†’ ๐‘ฅ โ†’ ๐‘ฃ 1 , if (๐‘ƒ1 , ๐‘ƒ2 ) โˆˆ ๐’ซ12 , and similarly if (๐‘ƒ1 , ๐‘ƒ2 ) โˆˆ ๐’ซ21 with ๐‘ฃ 1 and ๐‘ฃ 2 reversed. An illustration of this map is shown in Figure 2.5. Because the set of intersections in (๐‘ƒ1 , ๐‘ƒ2 ) is the same as in (๐‘ƒ1โ€ฒ , ๐‘ƒ2โ€ฒ ), the first intersection remains the same and this makes ฮฉ an involution. It is also clear from its definition that it changes sign. We have proved the following lemma and corollary. Lemma 2.5.2. Let ๐ท be an acyclic digraph. Let ๐‘ข1 , ๐‘ข2 , ๐‘ฃ 1 , ๐‘ฃ 2 โˆˆ ๐‘‰(๐ท) be such that each pair of paths (๐‘ƒ1 , ๐‘ƒ2 ) โˆˆ ๐’ซ21 intersects. Then | ๐‘(๐‘ข1 ; ๐‘ฃ 1 ) ๐‘(๐‘ข1 ; ๐‘ฃ 2 ) | | | = number of nonintersecting pairs (๐‘ƒ1 , ๐‘ƒ2 ) โˆˆ ๐’ซ12 . | ๐‘(๐‘ข2 ; ๐‘ฃ 1 ) ๐‘(๐‘ข2 ; ๐‘ฃ 2 ) | โ–ก

In particular, the determinant is nonnegative. Corollary 2.5.3. For ๐‘› โ‰ฅ 0 the sequence ๐‘› ๐‘› ๐‘› ( ), ( ), . . . , ( ) 0 1 ๐‘› is log-concave.

Lemma 2.5.2 can be extended to ๐‘› ร— ๐‘› determinants as follows. Let ๐‘ข1 , . . . , ๐‘ข๐‘› and ๐‘ฃ 1 , . . . , ๐‘ฃ ๐‘› be ๐‘›-tuples of distinct vertices in an acyclic digraph. For ๐œ‹ โˆˆ ๐”–๐‘› , we let ๐’ซ๐œ‹ = {(๐‘ƒ1 , . . . , ๐‘ƒ๐‘› ) โˆฃ ๐‘ƒ๐‘– โˆถ ๐‘ข๐‘– โ†’ ๐‘ฃ ๐œ‹(๐‘–) for all ๐‘– โˆˆ [๐‘›]} and ๐’ซ=

โ‹ƒ

๐’ซ๐œ‹ .

๐œ‹โˆˆ๐”–๐‘›

To make ๐’ซ into a signed set, recall from abstract algebra that the sign of ๐œ‹ โˆˆ ๐”–๐‘› is sgn ๐œ‹ = (โˆ’1)๐‘›โˆ’๐‘˜ if ๐œ‹ has ๐‘˜ cycles in its disjoint cycle decomposition. There are other ways to define sgn ๐œ‹, but they are all equivalent. One crucial property of this sign function is that if ๐ด = [๐‘Ž๐‘–,๐‘— ] is a matrix, then det ๐ด = โˆ‘ (sgn ๐œ‹)๐‘Ž1,๐œ‹(1) ๐‘Ž2,๐œ‹(2) . . . ๐‘Ž๐‘›,๐œ‹(๐‘›) . ๐œ‹โˆˆ๐”–๐‘›

Now if (๐‘ƒ1 , . . . , ๐‘ƒ๐‘› ) โˆˆ ๐’ซ๐œ‹ , then we let sgn(๐‘ƒ1 , . . . , ๐‘ƒ๐‘› ) = sgn ๐œ‹. To extend the involution ฮฉ, call ๐‘ƒ = (๐‘ƒ1 , . . . , ๐‘ƒ๐‘› ) intersecting if there is some pair ๐‘ƒ๐‘– , ๐‘ƒ๐‘— which intersects. Given an intersecting ๐‘ƒ, we find the smallest ๐‘– such that ๐‘ƒ๐‘– intersects another path of ๐‘ƒ and let ๐‘ฅ be the first intersection of ๐‘ƒ๐‘– with another path. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

2.6. The Matrix-Tree Theorem

59

Now take the smallest ๐‘— > ๐‘– such that ๐‘ƒ๐‘— goes through ๐‘ฅ. We now let ฮฉ(๐‘ƒ) = ๐‘ƒ โ€ฒ where ๐‘ƒ โ€ฒ is ๐‘ƒ with ๐‘ƒ๐‘– , ๐‘ƒ๐‘— replaced by (2.12)

๐‘ƒ๐‘–

๐‘ƒ๐‘—

๐‘ƒ๐‘—

๐‘ƒ๐‘–

๐‘ƒ๐‘–โ€ฒ

= ๐‘ข๐‘– โ†’ ๐‘ฅ โ†’ ๐‘ฃ ๐œ‹(๐‘—) ,

๐‘ƒ๐‘—โ€ฒ

= ๐‘ข๐‘— โ†’ ๐‘ฅ โ†’ ๐‘ฃ ๐œ‹(๐‘–) ,

respectively. One now needs to check that ฮฉ is a sign-reversing involution. As before, nonintersecting path families ๐‘ƒ are fixed points of ฮฉ. Modulo the details about ฮฉ, we have now proved the following. Lemma 2.5.4 (Lindstrรถmโ€“Gessleโ€“Viennot). Let ๐ท be an acyclic digraph. Consider two sequences of vertices ๐‘ข1 , . . . , ๐‘ข๐‘› , ๐‘ฃ 1 , . . . , ๐‘ฃ ๐‘› โˆˆ ๐‘‰(๐ท) such that every ๐‘ƒ โˆˆ ๐’ซ๐œ‹ is intersecting for ๐œ‹ โ‰  id, the identity permutation. We have det[๐‘(๐‘ข๐‘– ; ๐‘ฃ๐‘— )]1โ‰ค๐‘–,๐‘—โ‰ค๐‘› = number of nonintersecting ๐‘ƒ โˆˆ ๐’ซid . โ–ก

In particular, the determinant is nonnegative.

This theorem also has something to say about real sequences. Any sequence ๐‘Ž0 , . . . , ๐‘Ž๐‘› has a corresponding Toeplitz matrix which is the infinite matrix ๐ด = [๐‘Ž๐‘—โˆ’๐‘– ]๐‘–,๐‘—โ‰ฅ0 . So ๐‘Ž โŽก 0 0 โŽข ๐ด=โŽข โŽข 0 โŽฃ โ‹ฎ

๐‘Ž1 ๐‘Ž0 0 โ‹ฎ

๐‘Ž2 ๐‘Ž1 ๐‘Ž0 โ‹ฎ

โ‹ฏ ๐‘Ž2 ๐‘Ž1 โ‹ฎ

๐‘Ž๐‘› โ‹ฏ ๐‘Ž2 โ‹ฎ

0 ๐‘Ž๐‘› โ‹ฏ โ‹ฎ

0 0 ๐‘Ž๐‘› โ‹ฎ

0 ... โŽค 0 ... โŽฅ . 0 ... โŽฅ โŽฅ โ‹ฎ โ‹ฎ โŽฆ

The sequence is called Pรณlya frequency, or PF for short, if every square submatrix of ๐ด has a nonnegative determinant. Notice that, in particular, we get the determinants in (2.10) so that PF implies log-concave. Lemma 2.5.4 can be used to prove that a sequence is PF in much the same way that Lemma 2.5.2 can be used to prove that it is log-concave. The reader should now have no difficulty in proving the following result. Theorem 2.5.5. For ๐‘› โ‰ฅ 0 the sequence ๐‘› ๐‘› ๐‘› ( ), ( ), . . . , ( ) 0 1 ๐‘› is PF.

โ–ก

2.6. The Matrix-Tree Theorem We end this chapter with another application of determinants. There are many places where these animals abide in enumerative combinatorics and a good survey will be found in the articles of Krattenthaler [54,55]. Here we will be concerned with counting spanning trees using a famous result of Kirchhoff called the Matrix-Tree Theorem. A subgraph ๐ป โŠ† ๐บ is called spanning if ๐‘‰(๐ป) = ๐‘‰(๐บ). So a spanning subgraph is completely determined by its edge set. A spanning tree ๐‘‡ of ๐บ is a spanning subgraph which is a tree. Clearly for a spanning tree to exist, ๐บ must be connected. Let ๐’ฎ๐‘‡(๐บ) be the set of spanning trees of ๐บ. If one considers the graph ๐บ on the left in Figure 2.6, Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

60

2. Counting with Signs

๐‘ฃ

๐‘’ ๐‘”

๐‘“

๐‘ฆ

๐‘ค

๐‘–

โ„Ž

๐‘ฅ ๐‘’

๐‘ฃ

๐‘ค ๐‘”

๐‘“

๐‘ฆ

โ„Ž

๐‘ฅ

๐‘–

Figure 2.6. A graph ๐บ, its spanning trees, and an orientation

then the list of its eight spanning trees is in the middle of the figure (shrunk to half size so they will fit and without the vertex and edge labels). To develop the tools needed to prove our main theorem, we first need to make some remarks about combinatorial matrices. We will often have occasion to create matrices whose rows and columns are indexed by sets rather than numbers. If ๐‘†, ๐‘‡ are sets, then an ๐‘† ร— ๐‘‡ matrix ๐‘€ is constructed by giving a linear order to the elements of ๐‘† and to those of ๐‘‡ and using them to index the rows and columns of ๐‘€, respectively. So if (๐‘ , ๐‘ก) โˆˆ ๐‘† ร— ๐‘‡, then ๐‘š๐‘ ,๐‘ก is the entry in ๐‘€ in the row indexed by ๐‘  and the column indexed by ๐‘ก. The reader may have noted that such a matrix depends not just on ๐‘†, ๐‘‡, but also on their linear orderings. However, changing these orderings merely permutes rows and columns in ๐‘€ which will usually have no effect on the information we wish to extract from it. If ๐บ = (๐‘‰, ๐ธ) is a graph, then there are several important matrices associated with it. The adjacency matrix of ๐บ is the ๐‘‰ ร— ๐‘‰ matrix ๐ด = ๐ด(๐บ) with ๐‘Ž๐‘ฃ,๐‘ค = {

1 if ๐‘ฃ๐‘ค โˆˆ ๐ธ, 0 otherwise.

Using the ordering ๐‘ฃ, ๐‘ค, ๐‘ฅ, ๐‘ฆ, the graph on the left in Figure 2.6 has adjacency matrix

๐ด=

๐‘ฃ ๐‘ค ๐‘ฅ ๐‘ฆ

โŽก โŽข โŽข โŽข โŽฃ

๐‘ฃ 0 1 1 1

๐‘ค 1 0 1 0

๐‘ฅ 1 1 0 1

๐‘ฆ 1 0 1 0

โŽค โŽฅ . โŽฅ โŽฅ โŽฆ

The adjacency matrix is always symmetric since ๐‘ฃ๐‘ค and ๐‘ค๐‘ฃ denote the same edge. It also has zeros on the diagonal since our graphs are (usually) loopless. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

2.6. The Matrix-Tree Theorem

61

A second matrix associated with ๐บ is its incidence matrix, ๐ต = ๐ต(๐บ), which is the ๐‘‰ ร— ๐ธ matrix with entries ๐‘๐‘ฃ,๐‘’ = {

1 if ๐‘ฃ is an endpoint of ๐‘’, 0 otherwise.

Returning to our example, the graph has

๐ต=

๐‘ฃ ๐‘ค ๐‘ฅ ๐‘ฆ

โŽก โŽข โŽข โŽข โŽฃ

๐‘’ 1 1 0 0

๐‘“ 1 0 0 1

๐‘” 1 0 1 0

โ„Ž 0 1 1 0

๐‘– 0 0 1 1

โŽค โŽฅ . โŽฅ โŽฅ โŽฆ

By construction, row ๐‘ฃ of ๐ต contains deg ๐‘ฃ ones, and every column contains 2 ones. We will also need the diagonal ๐‘‰ ร— ๐‘‰ matrix ๐ถ(๐บ) which has diagonal entries ๐‘๐‘ฃ,๐‘ฃ = deg ๐‘ฃ. These three matrices are nicely related. Proposition 2.6.1. For any graph ๐บ we have ๐ต๐ต ๐‘ก = ๐ด + ๐ถ. Proof. The (๐‘ฃ, ๐‘ค) entry of ๐ต๐ต๐‘ก is the inner product of rows ๐‘ฃ and ๐‘ค of ๐ต. If ๐‘ฃ = ๐‘ค, then this is, using the notation (1.9), 2 โˆ‘ ๐‘๐‘ฃ,๐‘’ = โˆ‘ ๐›ฟ(๐‘ฃ is an endpoint of ๐‘’)2 = deg ๐‘ฃ = ๐‘๐‘ฃ,๐‘ฃ ๐‘’

๐‘’

since 02 = 0 and 12 = 1. Similarly, if ๐‘ฃ โ‰  ๐‘ค, then the entry is โˆ‘ ๐‘๐‘ฃ,๐‘’ ๐‘๐‘ค,๐‘’ = โˆ‘ ๐›ฟ(๐‘ฃ is an endpoint of ๐‘’) โ‹… ๐›ฟ(๐‘ค is an endpoint of ๐‘’) ๐‘’

๐‘’

= ๐›ฟ(๐‘ฃ๐‘ค โˆˆ ๐ธ) = ๐‘Ž๐‘ฃ,๐‘ค , โ–ก

which completes the proof.

Interestingly, to compute the number of spanning trees of ๐บ we will have to turn ๐บ into a digraph. An orientation of ๐บ is a digragh ๐ท with ๐‘‰(๐ท) = ๐‘‰(๐บ) and, for each edge ๐‘ฃ๐‘ค โˆˆ ๐ธ(๐บ), either the arc ๐‘ฃ๐‘ค โƒ— or the arc ๐‘ค๐‘ฃ โƒ— in ๐ด(๐ท). In this case ๐บ is called the underlying graph of ๐ท. The digraph on the right in Figure 2.6 is an orientation of our running example graph ๐บ. The adjacency matrix of a digraph is defined just as for graphs and will not concern us here. But we will need the directed incidence matrix, ๐ต = ๐ต(๐ท), defined by

๐‘๐‘ฃ,๐‘Ž =

โƒ— for some ๐‘ค, โŽงโˆ’1 if ๐‘Ž = ๐‘ฃ๐‘ค 1 if ๐‘Ž = ๐‘ค๐‘ฃ โƒ— for some ๐‘ค, โŽจ 0 otherwise. โŽฉ

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

62

2. Counting with Signs

For the digraph in Figure 2.6 we have

๐ต=

๐‘ฃ ๐‘ค ๐‘ฅ ๐‘ฆ

โŽก โŽข โŽข โŽข โŽฃ

๐‘’ ๐‘“ ๐‘” โ„Ž ๐‘– โˆ’1 โˆ’1 1 0 0 โŽค 1 0 0 1 0 โŽฅ 0 0 โˆ’1 โˆ’1 1 โŽฅ โŽฅ 0 1 0 0 โˆ’1 โŽฆ

.

Here are the two properties of ๐ต(๐ท) which will be important for us. Proposition 2.6.2. Let ๐ท be a digraph and let ๐ต = ๐ต(๐ท). (a) If the rows of ๐ต are ๐›1 , . . . , ๐›๐‘› , then ๐›1 + โ‹ฏ + ๐› ๐‘› = ๐ŸŽ where ๐ŸŽ is the zero vector. (b) If ๐ท is an orientation of a graph ๐บ, then (2.13)

๐ต๐ต ๐‘ก = ๐ถ(๐บ) โˆ’ ๐ด(๐บ).

Proof. For (a), just note that every column of ๐ต contains a single 1 and a single โˆ’1, which will cancel in the sum. The proof of (b) is similar to that for Proposition 2.6.1 and so is left to the reader. โ–ก It is interesting to note that although the matrix ๐ต on the left-hand side of (2.13) depends on ๐ท, the right-hand side only depends on the underlying graph ๐บ. The matrix ๐ฟ(๐บ) = ๐ถ(๐บ) โˆ’ ๐ด(๐บ) is called the Laplacian of ๐บ and controls many combinatorial aspects of the graph. Returning to our example, we have 3 โˆ’1 โˆ’1 โˆ’1 โŽก โŽค โˆ’1 2 โˆ’1 0 โŽฅ โŽข ๐ฟ(๐บ) = โŽข . 3 โˆ’1 โŽฅ โŽข โˆ’1 โˆ’1 โŽฅ 0 โˆ’1 2 โŽฆ โŽฃ โˆ’1 Note that the sum of the rows of ๐ฟ = ๐ฟ(๐บ) is zero since, for all ๐‘ฃ โˆˆ ๐‘‰, column ๐‘ฃ contains deg ๐‘ฃ on the diagonal and then deg ๐‘ฃ other nonzero entries which are all โˆ’1. So det ๐ฟ = 0. But removing the last row and column of the previous displayed matrix and taking the determinant gives 3 โˆ’1 โˆ’1 2 โˆ’1 ] = 8. det[ โˆ’1 โˆ’1 โˆ’1 3 The reader may recall that 8 was also the number of spanning trees of ๐บ. This is not a coincidence! But before we can prove the implied theorem, we need one more result. Let ๐‘€ be an ๐‘† ร— ๐‘‡ matrix and let ๐ผ โŠ† ๐‘† and ๐ฝ โŠ† ๐‘‡. Let ๐‘€๐ผ,๐ฝ denote the submatrix of ๐‘€ whose rows are indexed by ๐ผ and columns by ๐ฝ. In ๐ต(๐บ) for our example graph ๐บ with ๐ผ = {๐‘ฃ, ๐‘ฅ} and ๐ฝ = {๐‘“, ๐‘”, ๐‘–} we would have ๐ต๐ผ,๐ฝ = [

1 0

1 1

0 ]. 1

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

2.6. The Matrix-Tree Theorem

63

If ๐ผ = ๐‘†โˆ’{๐‘ } for some ๐‘  โˆˆ ๐‘† and ๐ฝ = ๐‘‡ โˆ’{๐‘ก} for some ๐‘ก โˆˆ ๐‘‡, then we use the abbreviation ๐‘€๐‘ ,ฬ‚ ๐‘ก ฬ‚ for ๐‘€๐ผ,๐ฝ . In this case when ๐‘† = ๐‘‡ = [๐‘›], the (๐‘–, ๐‘—) cofactor of ๐‘€ is ๐‘š๐‘–,ฬ‚ ๐‘— ฬ‚ = (โˆ’1)๐‘–+๐‘— det ๐‘€๐‘–,ฬ‚ ๐‘— .ฬ‚ We will need the following famous result about determinants called the Cauchyโ€“ Binet Theorem. Since this is really a statement about linear algebra rather than combinatorics, we will just outline a proof in the exercises. Theorem 2.6.3 (Cauchyโ€“Binet Theorem). Let ๐‘„ be an [๐‘š] ร— [๐‘›] matrix and let ๐‘… be [๐‘›] ร— [๐‘š]. Then det ๐‘„๐‘… =

โˆ‘ det ๐‘„ [๐‘š],๐พ โ‹… det ๐‘…๐พ,[๐‘š] . ๐พโˆˆ([๐‘›] ) ๐‘š

โ–ก

Note that in the special case ๐‘š = ๐‘› this reduces to the well-known statement that det ๐‘„๐‘… = det ๐‘„ โ‹… det ๐‘…. Theorem 2.6.4 (Matrix-Tree Theorem). Let ๐บ be a graph with ๐‘‰ = [๐‘›], ๐ธ = [๐‘š], and let ๐ฟ = ๐ฟ(๐บ). We have for any ๐‘–, ๐‘— โˆˆ [๐‘›] #๐’ฎ๐‘‡(๐บ) = โ„“๐‘–,ฬ‚ ๐‘— .ฬ‚ Proof. We will do the case when ๐‘– = ๐‘— = ๐‘› as the other cases are similar. So โ„“๐‘›,ฬ‚ ๐‘›ฬ‚ = (โˆ’1)๐‘›+๐‘› det ๐ฟ๐‘›,ฬ‚ ๐‘›ฬ‚ = det ๐ฟ๐‘›,ฬ‚ ๐‘›ฬ‚ . Let ๐ท be any orientation of ๐บ and ๐ต = ๐ต(๐ท). By Proposition 2.6.2(b), we have that ๐ฟ = ๐ถ(๐บ) โˆ’ ๐ด(๐บ) = ๐ต๐ต ๐‘ก . It follows that ๐ฟ๐‘›,ฬ‚ ๐‘›ฬ‚ = ๐ต๐‘Š ,๐ธ (๐ต๐‘Š ,๐ธ )๐‘ก where ๐‘Š = [๐‘› โˆ’ 1]. Applying the Cauchyโ€“Binet Theorem we get โ„“๐‘›,ฬ‚ ๐‘›ฬ‚ =

โˆ‘ det ๐ต๐‘Š ,๐น โ‹… det(๐ต๐‘Š ,๐น )๐‘ก = โˆ‘ (det ๐ต๐‘Š ,๐น )2 . ๐ธ ๐ธ ๐นโˆˆ(๐‘›โˆ’1 ๐นโˆˆ(๐‘›โˆ’1 ) )

So the theorem will be proved if we can show that (2.14)

det ๐ต๐‘Š ,๐น = {

ยฑ1 0

if the edges of ๐น are a spanning tree of ๐บ, otherwise.

Note that ๐ต๐‘Š ,๐น is the incidence matrix of the digraph ๐ท๐น having ๐‘‰(๐ท๐น ) = ๐‘‰ and ๐ด(๐ท๐น ) = ๐น but with the row of vertex ๐‘› removed. We say that ๐ท๐น is a tree if its underlying graph is one. We first consider the case when ๐ท๐น is not a tree. We know #๐น = ๐‘› โˆ’ 1 so, by Theorem 1.10.2, ๐ท๐น must be disconnected. Thus there is a component of ๐ท๐น not containing the vertex ๐‘›. And the sum of the row vectors of ๐ต๐‘Š ,๐น corresponding to that component is ๐ŸŽ by Proposition 2.6.2(a). Thus det ๐ต๐‘Š ,๐น = 0 in this case. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

64

2. Counting with Signs

Now suppose that ๐ท๐น is a tree. To prove this case of (2.14), it suffices to permute the rows and columns of ๐ต๐‘Š ,๐น so that the matrix becomes lower triangular with ยฑ1 on the diagonal. Such a permutation corresponds to a relabeling of the vertices and edges of ๐ท๐น . If ๐‘› = 1, then ๐ต๐‘Š ,๐น is the empty matrix which has determinant 1. If ๐‘› > 1, then, by Lemma 1.10.1, ๐ท๐น has at least two leaves. So in particular there is a leaf in ๐‘Š = [๐‘› โˆ’ 1]. By relabeling ๐ท๐น we can assume that ๐‘ฃ = 1 is the leaf and ๐‘Ž = 1 is the sole arc containing ๐‘ฃ. It follows that the first row of ๐ต๐‘Š ,๐น has ยฑ1 in the (1, 1) position and zeros elsewhere. Now we consider ๐ท๐น โˆ’ ๐‘ฃ and recurse to finish constructing the matrix. โ–ก We can use this theorem to rederive Cayleyโ€™s result, Theorem 1.10.3, enumerating all trees on a given vertex set. To do so, consider the complete graph ๐พ๐‘› with vertex set ๐‘‰ = [๐‘›]. Clearly the number of trees on ๐‘› vertices is the same as the number of spanning trees of ๐พ๐‘› . The Laplacian ๐ฟ(๐พ๐‘› ) consists of ๐‘› โˆ’ 1 down the diagonal with โˆ’1 everywhere else. So ๐ฟ๐‘›,ฬ‚ ๐‘›ฬ‚ is the same matrix but with dimensions (๐‘› โˆ’ 1) ร— (๐‘› โˆ’ 1). Add all the rows of this matrix to the first row. The result is a first row which is all ones since every column consists of an ๐‘› โˆ’ 1 as well as ๐‘› โˆ’ 2 minus ones in some order. Next add the first row to each of the other rows. This will cancel all the minus ones in those rows as well as changing each diagonal entry from ๐‘› โˆ’ 1 to ๐‘›. Now the matrix is upper triangular and, since elementary row operations do not change the determinant, we have that โ„“๐‘›,ฬ‚ ๐‘›ฬ‚ is the product of the diagonal entries which consist of a one and ๐‘› โˆ’ 2 copies of ๐‘›. Cayleyโ€™s Theorem follows.

Exercises (1) Let ๐‘› be a positive integer and let ๐‘1 , . . . , ๐‘ ๐‘˜ be distinct primes. Prove that the number of integers between 1 and ๐‘› not divisible by any of the ๐‘ ๐‘– is ๐‘›โˆ’ โˆ‘ โŒŠ 1โ‰ค๐‘–โ‰ค๐‘˜

๐‘› ๐‘› ๐‘› โŒ‹ โˆ’ โ‹ฏ + (โˆ’1)๐‘˜ โŒŠ โŒ‹+ โˆ‘ โŒŠ โŒ‹. ๐‘๐‘– ๐‘ ๐‘ ๐‘ ๐‘ . . . ๐‘๐‘˜ ๐‘– ๐‘— 1 2 1โ‰ค๐‘– ๐‘›. For example [

[4]! 4 ]= 2 [2]! [2]! =

[4][3] [2][1]

=

(1 + ๐‘ž + ๐‘ž2 + ๐‘ž3 )(1 + ๐‘ž + ๐‘ž2 ) (1 + ๐‘ž)

= 1 + ๐‘ž + 2๐‘ž2 + ๐‘ž3 + ๐‘ž4 .

(3.5)

It is not at all clear from the definition just given that this is actually a polynomial in ๐‘ž rather than just a rational function. But this follows easily using induction and our next result. Note that this theorem gives two ๐‘ž-analogues for the ordinary binomial recursion. This illustrates a general principle that ๐‘ž-analogues are not necessarily unique as we have also seen in the inv and maj interpretations of [๐‘›]๐‘ž !. Theorem 3.2.3. We have [

0 ] = ๐›ฟ0,๐‘˜ ๐‘˜ ๐‘ž

and, for ๐‘› โ‰ฅ 1, [

๐‘› ๐‘›โˆ’1 ๐‘›โˆ’1 ] = ๐‘ž๐‘˜ [ ] +[ ] ๐‘˜ ๐‘ž ๐‘˜ ๐‘˜โˆ’1 ๐‘ž ๐‘ž =[

๐‘›โˆ’1 ๐‘›โˆ’1 ] + ๐‘ž๐‘›โˆ’๐‘˜ [ ] . ๐‘˜ ๐‘˜โˆ’1 ๐‘ž ๐‘ž

Proof. The initial condition is trivial. We will prove the first recursion for the ๐‘žbinomial, leaving the other as an exercise. Using the definition in terms of ๐‘ž-factorials and finding a common denominator gives ๐‘ž๐‘˜ [

[๐‘› โˆ’ 1]! ๐‘›โˆ’1 ๐‘›โˆ’1 ]+[ ]= (๐‘ž๐‘˜ [๐‘› โˆ’ ๐‘˜] + [๐‘˜]) ๐‘˜ ๐‘˜โˆ’1 [๐‘˜]! [๐‘› โˆ’ ๐‘˜]! =

[๐‘› โˆ’ 1]! โ‹… [๐‘›] [๐‘˜]! [๐‘› โˆ’ ๐‘˜]!

=[

๐‘› ] ๐‘˜

as desired. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

โ–ก

78

3. Counting with Ordinary Generating Functions

Figure 3.1. The Young diagrams for (5, 5, 2, 1) โŠ‡ (3, 2, 2)

We will now give a ๐‘ž-analogue of the Binomial Theorem (Theorem 3.1.1). Let ๐‘ž, ๐‘ก be two variables. Theorem 3.2.4. For ๐‘› โ‰ฅ 0 we have ๐‘›

๐‘˜ ๐‘› (1 + ๐‘ก)(1 + ๐‘ž๐‘ก)(1 + ๐‘ž2 ๐‘ก) โ‹ฏ (1 + ๐‘ž๐‘›โˆ’1 ๐‘ก) = โˆ‘ ๐‘ž( 2 ) [ ] ๐‘ก๐‘˜ . ๐‘˜ ๐‘˜=0 ๐‘ž

(3.6)

Proof. We will induct on ๐‘› where the case ๐‘› = 0 is easy to check. For ๐‘› > 0 we can use the second recursion in the previous result and the induction hypothesis to write โˆ‘ ๐‘ž( 2 ) [ ๐‘˜

๐‘˜

๐‘˜ ๐‘˜ ๐‘› ๐‘›โˆ’1 ๐‘›โˆ’1 ] ๐‘ก ๐‘˜ = โˆ‘ ๐‘ž( 2 ) [ ] ๐‘ก๐‘˜ + โˆ‘ ๐‘ž( 2 )+๐‘›โˆ’๐‘˜ [ ] ๐‘ก๐‘˜ ๐‘˜ ๐‘ž ๐‘˜ ๐‘˜โˆ’1 ๐‘ž ๐‘˜ ๐‘˜ ๐‘ž

๐‘˜โˆ’1 ๐‘›โˆ’1 = (1 + ๐‘ก)(1 + ๐‘ž๐‘ก) โ‹ฏ (1 + ๐‘ž๐‘›โˆ’2 ๐‘ก) + ๐‘ž๐‘›โˆ’1 ๐‘ก โˆ‘ ๐‘ž( 2 ) [ ] ๐‘ก๐‘˜โˆ’1 ๐‘˜ โˆ’ 1 ๐‘˜ ๐‘ž

= (1 + ๐‘ก)(1 + ๐‘ž๐‘ก) โ‹ฏ (1 + ๐‘ž๐‘›โˆ’2 ๐‘ก) + ๐‘ž๐‘›โˆ’1 ๐‘ก(1 + ๐‘ก)(1 + ๐‘ž๐‘ก) โ‹ฏ (1 + ๐‘ž๐‘›โˆ’2 ๐‘ก) = (1 + ๐‘ก)(1 + ๐‘ž๐‘ก) โ‹ฏ (1 + ๐‘ž๐‘›โˆ’1 ๐‘ก), which is what we wished to prove.

โ–ก

There are many combinatorial interpretations of the ๐‘ž-binomial coefficients. We will content ourselves with presenting two of them here. If ๐œ† = (๐œ†1 , . . . , ๐œ†๐‘˜ ) and ๐œ‡ = (๐œ‡1 , . . . , ๐œ‡๐‘™ ) are integer partitions, then we say that ๐œ† contains ๐œ‡, written ๐œ† โŠ‡ ๐œ‡, if ๐‘˜ โ‰ฅ ๐‘™ and ๐œ†๐‘– โ‰ฅ ๐œ‡๐‘– for ๐‘– โ‰ค ๐‘™. Equivalently, the Young diagram of ๐œ† contains the Young diagram of ๐œ‡ if they are placed so that their northwest corners align. As an example, (5, 5, 2, 1) โŠ‡ (3, 2, 2) and Figure 3.1 shows the diagram of ๐œ† with the squares of ๐œ‡ shaded inside. The notation ๐œ‡ โŠ† ๐œ† should be self-explanatory. Given ๐œ‡ โŠ† ๐œ†, one also has the corresponding skew partition (3.7)

๐œ†/๐œ‡ = {(๐‘–, ๐‘—) โˆˆ ๐œ† โˆฃ (๐‘–, ๐‘—) โˆ‰ ๐œ‡}.

The cells of the skew partition in Figure 3.1 are white. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

3.2. Statistics and ๐‘ž-analogues

79

The ๐‘˜ ร— ๐‘™ rectangle is the integer partition whose multiplicity notation is (๐‘˜๐‘™ ). Consider the set of partitions contained in this rectangle โ„›(๐‘˜, ๐‘™) = {๐œ† โˆฃ ๐œ† โŠ† (๐‘˜๐‘™ )}. Recalling that |๐œ†| is the sum of the parts of ๐œ†, we consider the generating function โˆ‘๐œ†โˆˆโ„›(๐‘˜,๐‘™) ๐‘ž|๐œ†| . For example, if ๐‘˜ = ๐‘™ = 2, then we have ๐œ† โŠ† (22 ) ๐‘ž|๐œ†|

โˆ… 1

(1) ๐‘ž

(2) ๐‘ž2

(12 ) ๐‘ž2

(2, 1) ๐‘ž3

(22 ) , ๐‘ž4

which gives ๐‘ž|๐œ†| = 1 + ๐‘ž + 2๐‘ž2 + ๐‘ž3 + ๐‘ž4 .

โˆ‘ ๐œ†โˆˆโ„›(2,2)

The reader will have noticed the similarity to (3.5), which is not an accident. Theorem 3.2.5. For ๐‘˜, ๐‘™ โ‰ฅ 0 we have ๐‘ž|๐œ†| = [

โˆ‘ ๐œ†โˆˆโ„›(๐‘˜,๐‘™)

๐‘˜+๐‘™ ] . ๐‘˜ ๐‘ž

Proof. We induct on ๐‘˜ where the case ๐‘˜ = 0 is left to the reader. If ๐‘˜ > 0 and ๐œ† โŠ† (๐‘˜๐‘™ ), then there are two possibilities. Either ๐œ†1 < ๐‘˜ in which case ๐œ† โŠ† ((๐‘˜ โˆ’ 1)๐‘™ ) or ๐œ†1 = ๐‘˜ so that ๐œ† can be written as ๐œ† = (๐‘˜, ๐œ†โ€ฒ ) where ๐œ†โ€ฒ is the partition containing the parts of ๐œ† other than ๐œ†1 . So ๐œ†โ€ฒ โŠ† (๐‘˜๐‘™โˆ’1 ). Notice that in this case |๐œ†| = |๐œ†โ€ฒ | + ๐‘˜. We now use induction and Theorem 3.2.3 to obtain โˆ‘

๐‘ž|๐œ†| =

๐œ†โˆˆโ„›(๐‘˜,๐‘™)

โˆ‘

๐‘ž|๐œ†| +

๐œ†โˆˆโ„›(๐‘˜โˆ’1,๐‘™)

โˆ‘

โ€ฒ

๐‘ž|๐œ† |+๐‘˜

๐œ†โ€ฒ โˆˆโ„›(๐‘˜,๐‘™โˆ’1)

=[

๐‘˜+๐‘™โˆ’1 ๐‘˜+๐‘™โˆ’1 ] + ๐‘ž๐‘˜ [ ] ๐‘˜โˆ’1 ๐‘˜

=[

๐‘˜+๐‘™ ], ๐‘˜ โ–ก

which finishes the proof.

For our second combinatorial interpretation of the Gaussian polynomials we will need some linear algebra. Let ๐‘ž be a prime power and let ๐”ฝ๐‘ž be the Galois field with ๐‘ž elements. Let ๐‘‰ be a vector space of dimension dim ๐‘‰ = ๐‘› over ๐”ฝ๐‘ž . We will use ๐‘Š โ‰ค ๐‘‰ to indicate that ๐‘Š is a subspace of ๐‘‰. Let [

๐‘‰ ] = {๐‘Š โ‰ค ๐‘‰ โˆฃ dim ๐‘Š = ๐‘˜}. ๐‘˜

The subspaces of dimension ๐‘˜ are in bijective correspondence with ๐‘˜ ร— ๐‘› row-reduced echelon matrices of full rank, that is, with no zero rows. For example, if ๐‘› = 4 and Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

80

3. Counting with Ordinary Generating Functions

๐‘˜ = 2, then the possible matrices are [

0 0 0 0

1 0

0 0 1 ], [ 1 0 0

โˆ— 0

0 0 1 ], [ 1 0 0

0 1

โˆ— ], โˆ—

[

1 0

โˆ— 0 1 โˆ— ], [ 0 1 0 0

0 1

โˆ— 1 ], [ โˆ— 0

โˆ— โˆ—

โˆ— ], โˆ—

โˆ— 0

0 1

where the stars represent arbitrary elements of ๐”ฝ๐‘ž . So the number of subspaces corresponding to one of these star diagrams is ๐‘ž๐‘  where ๐‘  is the number of stars. Thus #[

๐”ฝ4๐‘ž ] = 1 + ๐‘ž + 2๐‘ž2 + ๐‘ž3 + ๐‘ž4 , 2

which should look very familiar at this point! Note however that, in contrast to previous cases, this actually represents an integer rather than a polynomial since ๐‘ž is a prime power. Of course, this example generalizes. Because of this result people sometimes talk half-jokingly about sets being vector spaces over the (nonexistent) Galois field with one element. Theorem 3.2.6. If ๐‘‰ is a vector space over ๐”ฝ๐‘ž of dimension ๐‘›, then #[

๐‘‰ ๐‘› ]=[ ] . ๐‘˜ ๐‘˜ ๐‘ž

Proof. Given ๐‘Š โ‰ค ๐‘‰ with dim ๐‘Š = ๐‘˜, we first count the number of possible ordered bases (๐ฏ1 , ๐ฏ2 , . . . , ๐ฏ๐‘˜ ) for ๐‘Š. Note that since dim ๐‘‰ = ๐‘› we have #๐‘‰ = #๐”ฝ๐‘›๐‘ž = ๐‘ž๐‘› . We can pick any nonzero vector for ๐ฏ1 so the number of choices is ๐‘ž๐‘› โˆ’ 1. For ๐ฏ2 we can choose any vector in ๐‘‰ which is not in the span of ๐ฏ1 , which gives ๐‘ž๐‘› โˆ’ ๐‘ž possibilities. Continuing in this way, the total count will be (๐‘ž๐‘› โˆ’ 1)(๐‘ž๐‘› โˆ’ ๐‘ž)(๐‘ž๐‘› โˆ’ ๐‘ž2 ) . . . (๐‘ž๐‘› โˆ’ ๐‘ž๐‘˜โˆ’1 ). By a similar argument, the number of different ordered bases which span a given ๐‘Š of dimension ๐‘˜ is (๐‘ž๐‘˜ โˆ’ 1)(๐‘ž๐‘˜ โˆ’ ๐‘ž)(๐‘ž๐‘˜ โˆ’ ๐‘ž2 ) . . . (๐‘ž๐‘˜ โˆ’ ๐‘ž๐‘˜โˆ’1 ). So the number of possible ๐‘Šโ€™s is ๐‘˜ (๐‘ž๐‘› โˆ’ 1)(๐‘ž๐‘› โˆ’ ๐‘ž) . . . (๐‘ž๐‘› โˆ’ ๐‘ž๐‘˜โˆ’1 ) ๐‘ž( 2 ) (๐‘ž๐‘› โˆ’ 1)(๐‘ž๐‘›โˆ’1 โˆ’ 1) . . . (๐‘ž๐‘›โˆ’๐‘˜+1 โˆ’ 1) = ๐‘˜ (๐‘ž๐‘˜ โˆ’ 1)(๐‘ž๐‘˜ โˆ’ ๐‘ž) . . . (๐‘ž๐‘˜ โˆ’ ๐‘ž๐‘˜โˆ’1 ) ๐‘ž( 2 ) (๐‘ž๐‘˜ โˆ’ 1)(๐‘ž๐‘˜โˆ’1 โˆ’ 1) . . . (๐‘ž โˆ’ 1)

=

(๐‘ž โˆ’ 1)๐‘˜ [๐‘›][๐‘› โˆ’ 1] . . . [๐‘› โˆ’ ๐‘˜ + 1] (๐‘ž โˆ’ 1)๐‘˜ [๐‘˜][๐‘˜ โˆ’ 1] . . . [1]

=[

๐‘› ], ๐‘˜

as advertised. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

โ–ก

3.3. The algebra of formal power series

81

There is a beautiful proof of this result due to Knuth [50] using row-reduced echelon matrices as in the previous example. The reader will be asked to supply the details in the exercises.

3.3. The algebra of formal power series We now wish to generalize the concept of generating function from finite to countably infinite sequences. To do so, we will have to use power series. But we wish to avoid the questions of convergence which come up when using analytic power series. Instead, we will work in the algebra of formal power series. This will mean that we have to be careful since, in an algebra, one is only permitted to apply an operation like addition or multiplication a finite number of times. But there is another concept of convergence which will take care of this issue. We should note that there is a whole branch of combinatorics which uses analytic techniques to extract useful information about a sequence, such as its rate of growth, from the corresponding power series. For information about this approach, see the book of Flajolet and Sedgewick [25]. A formal power series is an expression of the form โˆž

๐‘“(๐‘ฅ) = ๐‘Ž0 + ๐‘Ž1 ๐‘ฅ + ๐‘Ž2 ๐‘ฅ2 + โ‹ฏ = โˆ‘ ๐‘Ž๐‘› ๐‘ฅ๐‘› , ๐‘›=0

where the ๐‘Ž๐‘› are complex numbers. We also say that ๐‘“(๐‘ฅ) is the ordinary generating function or ogf for the sequence ๐‘Ž๐‘› , ๐‘› โ‰ฅ 0. Often we will leave out the adjective โ€œordinaryโ€ in this chapter since we will not have met any other type of generating function yet. Note that these series are considered formal in the sense that the powers of ๐‘ฅ are just place holders and we are not permitted to substitute a value for ๐‘ฅ. Because of this rule, analytic convergence is not an issue and we can happily talk about formal power series such as โˆ‘๐‘›โ‰ฅ0 ๐‘›! ๐‘ฅ๐‘› which converge nowhere except at ๐‘ฅ = 0. We will use the notation โ„‚[[๐‘ฅ]] = { โˆ‘ ๐‘Ž๐‘› ๐‘ฅ๐‘› โˆฃ ๐‘Ž๐‘› โˆˆ โ„‚ for all ๐‘› โ‰ฅ 0} . ๐‘›โ‰ฅ0

The set is an algebra, the algebra of formal power series, under the three operations of addition, scalar multiplication, and multiplication defined by โˆ‘ ๐‘Ž๐‘› ๐‘ฅ๐‘› + โˆ‘ ๐‘๐‘› ๐‘ฅ๐‘› = โˆ‘ (๐‘Ž๐‘› + ๐‘๐‘› )๐‘ฅ๐‘› , ๐‘›โ‰ฅ0

๐‘›โ‰ฅ0

๐‘›โ‰ฅ0

๐‘ โˆ‘ ๐‘Ž๐‘› ๐‘ฅ๐‘› = โˆ‘ (๐‘๐‘Ž๐‘› )๐‘ฅ๐‘› , ๐‘›โ‰ฅ0

๐‘›โ‰ฅ0

โˆ‘ ๐‘Ž๐‘› ๐‘ฅ๐‘› โ‹… โˆ‘ ๐‘๐‘› ๐‘ฅ๐‘› = โˆ‘ ๐‘๐‘› ๐‘ฅ๐‘› , ๐‘›โ‰ฅ0

where ๐‘ โˆˆ โ„‚ and

๐‘›โ‰ฅ0

๐‘›โ‰ฅ0 ๐‘›

๐‘๐‘› = โˆ‘ ๐‘Ž๐‘˜ ๐‘๐‘›โˆ’๐‘˜ . ๐‘˜=0

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

82

3. Counting with Ordinary Generating Functions

The reader may object that, as mentioned earlier, in an algebra one is only permitted a finite number of additions yet the very elements of โ„‚[[๐‘ฅ]] seem to involve infinitely many. But this is an illusion. Remember that ๐‘ฅ is a formal parameter so that the expression โˆ‘๐‘› ๐‘Ž๐‘› ๐‘ฅ๐‘› is only meant to be a mnemonic device which gives intuition to the definitions of the three algebra operations, especially that of multiplication. We could just as easily have defined โ„‚[[๐‘ฅ]] to be the set of all complex vectors (๐‘Ž0 , ๐‘Ž1 , ๐‘Ž2 , . . . ) subject to the operation of vector addition (๐‘Ž0 , ๐‘Ž1 , ๐‘Ž2 , . . . ) + (๐‘0 , ๐‘1 , ๐‘2 , . . . ) = (๐‘Ž0 + ๐‘0 , ๐‘Ž1 + ๐‘1 , ๐‘Ž2 + ๐‘2 , . . . ) and similarly for the other two. What is true is that one is only permitted to add or multiply a finite number of elements of โ„‚[[๐‘ฅ]]. So one can only perform operations which will alter the coefficient of a given power of ๐‘ฅ a finite number of times. Now given a sequence of complex numbers ๐‘Ž0 , ๐‘Ž1 , ๐‘Ž2 , . . . , we associate with it the ordinary generating function ๐‘“(๐‘ฅ) = ๐‘Ž0 + ๐‘Ž1 ๐‘ฅ + ๐‘Ž2 ๐‘ฅ2 + โ‹ฏ โˆˆ โ„‚[[๐‘ฅ]]. We will sometimes say that this series counts the objects enumerated by the ๐‘Ž๐‘› if appropriate. As with generating polynomials, the reason for doing so is to exploit properties of โ„‚[[๐‘ฅ]] to obtain information about the original sequence. We will often write this generating function as โˆ‘๐‘› ๐‘Ž๐‘› ๐‘ฅ๐‘› , assuming the that range of indices is ๐‘› โ‰ฅ 0. Let us start with a simple example. Consider the sequence 1, 1, 1, . . . with generating function โˆ‘๐‘› ๐‘ฅ๐‘› . We would like to simplify this as a geometric series to (3.8)

1 + ๐‘ฅ + ๐‘ฅ2 + โ‹ฏ =

1 . 1โˆ’๐‘ฅ

But what does the right-hand side even mean since 1/(1 โˆ’ ๐‘ฅ) appears to be a rational function and so not an element of โ„‚[[๐‘ฅ]]? The way out of this conundrum is to remember that given an element ๐‘Ž in an algebra ๐ด, it is possible for ๐‘Ž to have an inverse, namely an element ๐‘Žโˆ’1 such that ๐‘Ž โ‹… ๐‘Žโˆ’1 = 1 where 1 is the identity element of ๐ด. So to prove (3.8) in this setting we must show that โˆ‘๐‘› ๐‘ฅ๐‘› and 1 โˆ’ ๐‘ฅ are inverses. This is easily done by using the distributive law: (1 โˆ’ ๐‘ฅ)(1 + ๐‘ฅ + ๐‘ฅ2 + โ‹ฏ) = (1 + ๐‘ฅ + ๐‘ฅ2 + โ‹ฏ) โˆ’ ๐‘ฅ(1 + ๐‘ฅ + ๐‘ฅ2 + โ‹ฏ) = (1 + ๐‘ฅ + ๐‘ฅ2 + โ‹ฏ) โˆ’ (๐‘ฅ + ๐‘ฅ2 + ๐‘ฅ3 + โ‹ฏ) = 1. This example illustrates a general principle that often well-known results about analytic power series carry over to their formal counterparts, although some work may be required to check that this is true. For the most part, we will assume the truth of a standard formula in this setting without further comment. But it would be wise to also give a couple of examples to show that caution may be needed. One illustration is that the expression 1/๐‘ฅ has no meaning in โ„‚[[๐‘ฅ]] because ๐‘ฅ does not have an inverse. For suppose we have ๐‘ฅ๐‘“(๐‘ฅ) = 1 for some formal power series ๐‘“(๐‘ฅ). Then on the left-hand side the constant coefficient is 0 while on the right it is 1, a contradiction. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

3.3. The algebra of formal power series

83

As another example, consider the sequence 1/๐‘›! for ๐‘› โ‰ฅ 0. We would like to write ๐‘ฅ๐‘› ๐‘›! ๐‘›โ‰ฅ0

๐‘’๐‘ฅ = โˆ‘

for the corresponding generating function but, again, run into the problem that ๐‘’๐‘ฅ is not a priori an element of โ„‚[[๐‘ฅ]]. The solution this time is to define ๐‘’๐‘ฅ to be a formal symbol which stands for this power series. Then, of course, to be complete we would need to verify formally that all the usual rules of exponents hold such as ๐‘’2๐‘ฅ = (๐‘’๐‘ฅ )2 . We will not take the time to do this. But we will point out a case where the rules do not hold. In particular, in โ„‚[[๐‘ฅ]] one cannot write ๐‘’1+๐‘ฅ = ๐‘’๐‘’๐‘ฅ . This is because the left-hand side is not well-defined. Indeed, when expanding โˆ‘๐‘› (1 + ๐‘ฅ)๐‘› /๐‘›! there are infinitely many additions needed to compute the coefficient of any given power of ๐‘ฅ which, as we have already noted, is not permitted. Although we will not verify every specific analytic identity needed for formal power series in this text, it would be good to have some general results about which operations are permitted in โ„‚[[๐‘ฅ]]. First we deal with the issue of when a formal power series is invertible. Theorem 3.3.1. If ๐‘“(๐‘ฅ) = โˆ‘๐‘› ๐‘Ž๐‘› ๐‘ฅ๐‘› , then ๐‘“(๐‘ฅ)โˆ’1 exists in โ„‚[[๐‘ฅ]] if and only if ๐‘Ž0 โ‰  0. Proof. For the forward direction, suppose ๐‘“(๐‘ฅ)๐‘”(๐‘ฅ) = 1 where ๐‘”(๐‘ฅ) = โˆ‘๐‘› ๐‘๐‘› ๐‘ฅ๐‘› . Taking the constant coefficient on both sides gives ๐‘Ž0 ๐‘0 = 1. So ๐‘Ž0 โ‰  0. Now assume ๐‘Ž0 โ‰  0. We will construct an inverse ๐‘”(๐‘ฅ) = โˆ‘๐‘› ๐‘๐‘› ๐‘ฅ๐‘› . We want ๐‘“(๐‘ฅ)๐‘”(๐‘ฅ) = 1. Comparing coefficients of ๐‘ฅ๐‘› on both sides we see that we wish to have ๐‘Ž0 ๐‘0 = 1 and, for ๐‘› โ‰ฅ 1, ๐‘Ž0 ๐‘๐‘› + ๐‘Ž1 ๐‘๐‘›โˆ’1 + โ‹ฏ + ๐‘Ž๐‘› ๐‘0 = 0. Since ๐‘Ž0 โ‰  0 we can take ๐‘0 = 1/๐‘Ž0 . By the same token, when ๐‘› โ‰ฅ 1 we can solve for ๐‘๐‘› in the previous displayed equation giving a recursive formula for its value. Thus we can construct such a ๐‘”(๐‘ฅ) and are done. โ–ก Our example with ๐‘’๐‘ฅ shows that we also need to be careful about substitution. We wish to define the substitution of ๐‘”(๐‘ฅ) into ๐‘“(๐‘ฅ) = โˆ‘๐‘› ๐‘Ž๐‘› ๐‘ฅ๐‘› to be ๐‘“(๐‘”(๐‘ฅ)) = โˆ‘ ๐‘Ž๐‘› ๐‘”(๐‘ฅ)๐‘› . ๐‘›โ‰ฅ0

But now the right-hand side is an infinite sum of formal power series, not just formal variables. To be able to talk about such sums, we need to introduce a notion of convergence in โ„‚[[๐‘ฅ]]. It will be convenient to have the notation that for a formal power series ๐‘“(๐‘ฅ) [๐‘ฅ๐‘› ]๐‘“(๐‘ฅ) = the coefficient of ๐‘ฅ๐‘› in ๐‘“(๐‘ฅ), which we have usually been calling ๐‘Ž๐‘› . Suppose that we have a sequence ๐‘“0 (๐‘ฅ), ๐‘“1 (๐‘ฅ), ๐‘“2 (๐‘ฅ), . . . of formal power series. We say that this sequence converges to ๐‘“(๐‘ฅ) โˆˆ โ„‚[[๐‘ฅ]] Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

84

3. Counting with Ordinary Generating Functions

and write lim ๐‘“๐‘˜ (๐‘ฅ) = ๐‘“(๐‘ฅ),

๐‘˜โ†’โˆž

if, for any ๐‘›, the coefficient of ๐‘ฅ๐‘› in the sequence is eventually constant and equals the coefficient of ๐‘ฅ๐‘› in ๐‘“(๐‘ฅ). Formally, given ๐‘›, there exists a corresponding ๐พ such that [๐‘ฅ๐‘› ]๐‘“๐‘˜ (๐‘ฅ) = [๐‘ฅ๐‘› ]๐‘“(๐‘ฅ) for all ๐‘˜ โ‰ฅ ๐พ. Otherwise we say that the sequence diverges or that the limit does not exist. As an illustration, consider the sequence ๐‘“0 (๐‘ฅ) = 1, ๐‘“1 (๐‘ฅ) = 1 + ๐‘ฅ, ๐‘“2 (๐‘ฅ) = 1 + ๐‘ฅ + ๐‘ฅ2 , . . . so that ๐‘“๐‘˜ (๐‘ฅ) = 1 + ๐‘ฅ + โ‹ฏ + ๐‘ฅ๐‘˜ . Then this sequence has a limit; namely lim ๐‘“๐‘˜ (๐‘ฅ) = โˆ‘ ๐‘ฅ๐‘› =

๐‘˜โ†’โˆž

๐‘›โ‰ฅ0

1 . 1โˆ’๐‘ฅ

To prove this, note that given ๐‘› we can let ๐พ = ๐‘›. So for ๐‘˜ โ‰ฅ ๐‘› we have [๐‘ฅ๐‘› ]๐‘“๐‘˜ = [๐‘ฅ๐‘› ]๐‘“๐‘› = 1. On the other hand, consider the sequence ๐‘“0 (๐‘ฅ) = 1 + ๐‘ฅ, ๐‘“1 (๐‘ฅ) = 1/2 + ๐‘ฅ/2, ๐‘“2 (๐‘ฅ) = 1/4 + ๐‘ฅ/4, . . . and in general ๐‘“๐‘˜ (๐‘ฅ) = 1/2๐‘˜ + ๐‘ฅ/2๐‘˜ . This sequence does not converge in โ„‚[[๐‘ฅ]] since for any ๐‘› we have that [๐‘ฅ]๐‘“๐‘˜ (๐‘ฅ) is always different for different ๐‘˜. This is in contrast to the analytic situation where this sequence converges to zero. As in analysis, we now use convergence of sequences to define convergence of series. Given ๐‘“0 (๐‘ฅ), ๐‘“1 (๐‘ฅ), ๐‘“2 (๐‘ฅ), . . . , we say that their sum exists and converges to ๐‘“(๐‘ฅ), written โˆ‘๐‘˜โ‰ฅ0 ๐‘“๐‘˜ (๐‘ฅ) = ๐‘“(๐‘ฅ), if lim ๐‘ ๐‘˜ (๐‘ฅ) = ๐‘“(๐‘ฅ)

๐‘˜โ†’โˆž

where (3.9)

๐‘ ๐‘˜ (๐‘ฅ) = ๐‘“0 (๐‘ฅ) + ๐‘“1 (๐‘ฅ) + โ‹ฏ + ๐‘“๐‘˜ (๐‘ฅ)

is the ๐‘˜th partial sum. Divergence is defined as expected. Note that this definition is consistent with our notation for formal power series since given a sequence ๐‘Ž0 , ๐‘Ž1 , ๐‘Ž2 , . . . , we can let ๐‘“๐‘˜ (๐‘ฅ) = ๐‘Ž๐‘˜ ๐‘ฅ๐‘˜ and then prove that โˆ‘๐‘˜โ‰ฅ0 ๐‘“๐‘˜ (๐‘ฅ) = ๐‘“(๐‘ฅ) where ๐‘“(๐‘ฅ) = โˆ‘๐‘˜โ‰ฅ0 ๐‘Ž๐‘˜ ๐‘ฅ๐‘˜ . To state a criterion for convergence of series, it will be useful to define the minimum degree of ๐‘“(๐‘ฅ) = โˆ‘๐‘› ๐‘Ž๐‘› ๐‘ฅ๐‘› to be mdeg ๐‘“(๐‘ฅ) = smallest ๐‘› such that ๐‘Ž๐‘› โ‰  0 if ๐‘“(๐‘ฅ) โ‰  0, and let mdeg ๐‘“(๐‘ฅ) = โˆž if ๐‘“(๐‘ฅ) = 0. It turns out that to show a sum of power series converges, it suffices to take a limit of integers. Theorem 3.3.2. Given ๐‘“0 (๐‘ฅ), ๐‘“1 (๐‘ฅ), ๐‘“2 (๐‘ฅ), . . . โˆˆ โ„‚[[๐‘ฅ]], then โˆ‘๐‘˜โ‰ฅ0 ๐‘“๐‘˜ (๐‘ฅ) exists if and only if lim (mdeg ๐‘“๐‘˜ (๐‘ฅ)) = โˆž.

๐‘˜โ†’โˆž

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

3.3. The algebra of formal power series

85

Proof. We will prove the forward direction, leaving the other implication as an exercise. We are given that the sequence ๐‘ ๐‘˜ (๐‘ฅ) as defined by (3.9) converges. So given ๐‘›, there is a ๐พ such that [๐‘ฅ๐‘› ]๐‘ ๐พ (๐‘ฅ) = [๐‘ฅ๐‘› ]๐‘ ๐พ+1 (๐‘ฅ) = [๐‘ฅ๐‘› ]๐‘ ๐พ+2 (๐‘ฅ) = โ‹ฏ . But for ๐‘— โ‰ฅ 0 we have ๐‘ ๐พ+๐‘— (๐‘ฅ) = ๐‘ ๐พ (๐‘ฅ) + ๐‘“๐พ+1 (๐‘ฅ) + ๐‘“๐พ+2 (๐‘ฅ) + โ‹ฏ + ๐‘“๐พ+๐‘— (๐‘ฅ). It follows that [๐‘ฅ๐‘› ]๐‘“๐‘˜ (๐‘ฅ) = 0 for ๐‘˜ > ๐พ. Now given ๐‘›, take ๐‘ to be the maximum of all the ๐พ-values associated to integers less than or equal to ๐‘›. From what we have shown, this forces mdeg ๐‘“๐‘˜ (๐‘ฅ) > ๐‘› for ๐‘› > ๐‘. But by definition of a limit of real numbers, this means lim๐‘˜โ†’โˆž (mdeg ๐‘“๐‘˜ (๐‘ฅ)) = โˆž. โ–ก We are now on a firm footing with our definition of substitution as we know what it means for a sum of power series to converge. We can use the previous result to give a simple criterion for convergence when substituting one generating function into another. Theorem 3.3.3. Given ๐‘“(๐‘ฅ), ๐‘”(๐‘ฅ) โˆˆ โ„‚[[๐‘ฅ]], then the composition ๐‘“(๐‘”(๐‘ฅ)) exists if and only if (1) ๐‘“(๐‘ฅ) is a polynomial or (2) ๐‘”(๐‘ฅ) has zero constant term. Proof. If ๐‘“(๐‘ฅ) is a polynomial, then ๐‘“(๐‘”(๐‘ฅ)) is a finite sum and so obviously converges. So assume ๐‘“(๐‘ฅ) = โˆ‘๐‘› ๐‘Ž๐‘› ๐‘ฅ๐‘› is not polynomial. If ๐‘”(๐‘ฅ) has no constant term, we have mdeg ๐‘Ž๐‘› ๐‘”(๐‘ฅ)๐‘› โ‰ฅ ๐‘›. So the limit in the previous theorem is infinity and ๐‘“(๐‘”(๐‘ฅ)) is well-defined. To finish the proof, consider the remaining case where [๐‘ฅ0 ]๐‘”(๐‘ฅ) โ‰  0. Since ๐‘“(๐‘ฅ) is not a polynomial, there are an infinite number of ๐‘› such that ๐‘Ž๐‘› โ‰  0. But for these ๐‘› we have mdeg ๐‘Ž๐‘› ๐‘”(๐‘ฅ)๐‘› = 0. So the desired limit cannot be infinity and ๐‘“(๐‘”(๐‘ฅ)) does not exist in โ„‚[[๐‘ฅ]]. โ–ก We will also find it useful to consider certain infinite products. We approach their convergence just as we did for infinite sums. Given a sequence ๐‘“0 (๐‘ฅ), ๐‘“1 (๐‘ฅ), ๐‘“2 (๐‘ฅ), . . . , we say that their product exists and converges to ๐‘“(๐‘ฅ), written โˆ๐‘˜โ‰ฅ0 ๐‘“๐‘˜ (๐‘ฅ) = ๐‘“(๐‘ฅ), if lim ๐‘ ๐‘˜ (๐‘ฅ) = ๐‘“(๐‘ฅ)

๐‘˜โ†’โˆž

where ๐‘ ๐‘˜ (๐‘ฅ) = ๐‘“0 (๐‘ฅ)๐‘“1 (๐‘ฅ) . . . ๐‘“๐‘˜ (๐‘ฅ). We have the following result whose proof is similar enough to that of Theorem 3.3.2 that we will leave it to the reader. Theorem 3.3.4. Let ๐‘“0 (๐‘ฅ), ๐‘“1 (๐‘ฅ), ๐‘“2 (๐‘ฅ), . . . be power series with zero constant terms. Then โˆ๐‘˜โ‰ฅ0 (1 + ๐‘“๐‘˜ (๐‘ฅ)) exists if and only if lim (mdeg ๐‘“๐‘˜ (๐‘ฅ)) = โˆž.

๐‘˜โ†’โˆž

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

โ–ก

86

3. Counting with Ordinary Generating Functions

Let us end this section by showing how the previous result can give simple verifications that a product does or does not exist. Consider โˆ๐‘˜โ‰ฅ1 (1 + ๐‘ฅ๐‘˜ ). In this case ๐‘“๐‘˜ (๐‘ฅ) = ๐‘ฅ๐‘˜ and mdeg ๐‘ฅ๐‘˜ = ๐‘˜. So the desired limit is infinity and this product exists. As we will see in Section 3.5, it counts integer partitions with distinct parts. By contrast, the product โˆ๐‘˜โ‰ฅ0 (1 + ๐‘ฅ/2๐‘˜ ) does not converge since mdeg ๐‘ฅ/2๐‘˜ = 1.

3.4. The Sum and Product Rules for ogfs Just as for sets there is a Sum Rule and a Product Rule for ordinary generating functions. In order to state these results, we need the idea of a weight-generating function. This approach makes it possible to construct generating functions for various sequences in a very combinatorial manner. As a first application, we make a more deep exploration of the Binomial Theorem. Let ๐‘† be a set. Then a weighting of ๐‘† is a function wt โˆถ ๐‘† โ†’ โ„‚[[๐‘ฅ]]. Most often if ๐‘  โˆˆ ๐‘†, then wt ๐‘  will just be a monomial reflecting some property of ๐‘ . For example, if st is any statistic on ๐‘†, then we could take wt ๐‘  = ๐‘ฅst ๐‘  . For a more concrete illustration which we will continue to use throughout this section, let ๐‘† = 2[๐‘›] and define for ๐‘‡ โˆˆ ๐‘† wt ๐‘‡ = ๐‘ฅ|๐‘‡| .

(3.10)

Given a weighted set ๐‘†, we can form the corresponding weight-generating function ๐‘“(๐‘ฅ) = ๐‘“๐‘† (๐‘ฅ) = โˆ‘ wt ๐‘ . ๐‘ โˆˆ๐‘†

We must be careful that this sum exists in โ„‚[[๐‘ฅ]], and if it does, then we say ๐‘† is a summable set. Of course, when ๐‘† is finite then it is automatically summable. To illustrate for ๐‘† = 2[3] we have ๐‘‡

โˆ…

{1}

{2}

{3}

{1, 2}

{1, 3}

{2, 3}

{1, 2, 3}

wt ๐‘‡

1

๐‘ฅ

๐‘ฅ

๐‘ฅ

๐‘ฅ2

๐‘ฅ2

๐‘ฅ2

๐‘ฅ3

so that ๐‘“๐‘† (๐‘ฅ) = 1 + 3๐‘ฅ + 3๐‘ฅ2 + ๐‘ฅ3 . More generally, for ๐‘† = 2[๐‘›] we have ๐‘“๐‘† (๐‘ฅ) = โˆ‘ ๐‘ฅ|๐‘‡| ๐‘‡โˆˆ2[๐‘›] ๐‘›

= โˆ‘

โˆ‘ ๐‘ฅ๐‘˜

๐‘˜=0 ๐‘‡โˆˆ([๐‘›]) ๐‘˜ ๐‘›

๐‘› = โˆ‘ ( )๐‘ฅ๐‘˜ ๐‘˜ ๐‘˜=0 and we have recovered the generating function for a row of Pascalโ€™s triangle. The following theorem will permit us to manipulate weight-generating functions with ease. For the Sum Rule, if ๐‘†, ๐‘‡ are disjoint weighted sets, then we weight ๐‘ข โˆˆ ๐‘† โŠŽ๐‘‡ Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

3.4. The Sum and Product Rules for ogfs

87

using ๐‘ขโ€™s weight in ๐‘† or in ๐‘‡ depending on whether ๐‘ข โˆˆ ๐‘† or ๐‘ข โˆˆ ๐‘‡, respectively. For arbitrary ๐‘†, ๐‘‡ we weight ๐‘† ร— ๐‘‡ by letting wt(๐‘ , ๐‘ก) = wt ๐‘  โ‹… wt ๐‘ก. Lemma 3.4.1. Let ๐‘†, ๐‘‡ be summable sets. (a) (Sum Rule) The set ๐‘† โˆช ๐‘‡ is summable. If ๐‘† โˆฉ ๐‘‡ = โˆ…, then ๐‘“๐‘†โŠŽ๐‘‡ (๐‘ฅ) = ๐‘“๐‘† (๐‘ฅ) + ๐‘“๐‘‡ (๐‘ฅ). (b) (Product Rule) The set ๐‘† ร— ๐‘‡ is summable and ๐‘“๐‘†ร—๐‘‡ (๐‘ฅ) = ๐‘“๐‘† (๐‘ฅ) โ‹… ๐‘“๐‘‡ (๐‘ฅ). Proof. (a) Since ๐‘† is summable, given any ๐‘› โˆˆ โ„•, there are only a finite number of ๐‘  โˆˆ ๐‘† such that wt ๐‘  has a nonzero coefficient of ๐‘ฅ๐‘› . And the same is true of ๐‘‡. It follows that only finitely many elements of ๐‘† โˆช ๐‘‡ have such a coefficient, which mean this set is summable. To prove the desired equality, we compute as follows: ๐‘“๐‘†โŠŽ๐‘‡ (๐‘ฅ) = โˆ‘ wt ๐‘ข = โˆ‘ wt ๐‘ข + โˆ‘ wt ๐‘ข = ๐‘“๐‘† (๐‘ฅ) + ๐‘“๐‘‡ (๐‘ฅ). แต†โˆˆ๐‘†โŠŽ๐‘‡

แต†โˆˆ๐‘†

แต†โˆˆ๐‘‡

(b) The statement about summability of ๐‘† ร— ๐‘‡ is safely left as an exercise. Computing the weight-generating function gives ๐‘“๐‘†ร—๐‘‡ (๐‘ฅ) =

โˆ‘

wt(๐‘ , ๐‘ก) = โˆ‘ wt ๐‘  โ‹… โˆ‘ wt ๐‘ก = ๐‘“๐‘† (๐‘ฅ) โ‹… ๐‘“๐‘‡ (๐‘ฅ), ๐‘ โˆˆ๐‘†

(๐‘ ,๐‘ก)โˆˆ๐‘†ร—๐‘‡

๐‘กโˆˆ๐‘‡

โ–ก

so we are done.

We can now use these rules to derive various generating functions in a straightforward manner. We begin by reproving the Binomial Theorem as stated in Theorem 3.1.1. We have already seen that the summation side is the weight-generating function for ๐‘† = 2[๐‘›] . For the product side, it will be useful to reformulate ๐‘† in terms of multiplicity notation. Specifically, consider ๐‘† โ€ฒ = {๐‘‡ โ€ฒ = (1๐‘š1 , 2๐‘š2 , . . . , ๐‘›๐‘š๐‘› ) โˆฃ ๐‘š๐‘– = 0 or 1 for all ๐‘–} weighted by wt ๐‘‡ โ€ฒ = ๐‘ฅโˆ‘๐‘– ๐‘š๐‘– . Clearly we have a bijection ๐‘“ โˆถ ๐‘† โ†’ ๐‘†โ€ฒ given by ๐‘“(๐‘‡) = (1๐‘š1 , 2๐‘š2 , . . . , ๐‘›๐‘š๐‘› ) where ๐‘š๐‘– = {

0 if ๐‘– โˆ‰ ๐‘‡, 1 if ๐‘– โˆˆ ๐‘‡.

(In fact, this is the map used in the proof of Theorem 1.3.1.) Furthermore, this bijection is weight preserving in that wt ๐‘“(๐‘‡) = wt ๐‘‡. As a concrete example, if ๐‘› = 5 and ๐‘‡ = {2, 4, 5}, then ๐‘“(๐‘‡) = (10 , 21 , 30 , 41 , 51 ) and wt ๐‘“(๐‘‡) = ๐‘ฅ3 = wt ๐‘‡. The advantage of using ๐‘† โ€ฒ is that it is clearly a weighted product of the sets {๐‘–0 , ๐‘–1 } for ๐‘– โˆˆ [๐‘›] where wt ๐‘–0 = 1 and wt ๐‘–1 = ๐‘ฅ. So we can write, where for distinct elements ๐‘Ž and ๐‘ we use the shorthand ๐‘Ž โŠŽ ๐‘ for {๐‘Ž} โŠŽ {๐‘}, ๐‘† โ€ฒ = {10 , 11 } ร— {20 , 21 } ร— โ‹ฏ ร— {๐‘›0 , ๐‘›1 } = (10 โŠŽ 11 ) ร— (20 โŠŽ 21 ) ร— โ‹ฏ ร— (๐‘›0 โŠŽ ๐‘›1 ). Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

88

3. Counting with Ordinary Generating Functions

Translating this expression using both parts of Lemma 3.4.1, we obtain ๐‘›

๐‘› โˆ‘ ( )๐‘ฅ๐‘˜ = ๐‘“๐‘† (๐‘ฅ) ๐‘˜ ๐‘˜=0 = ๐‘“๐‘†โ€ฒ (๐‘ฅ) = (wt 10 + wt 11 )(wt 20 + wt 21 ) โ‹ฏ (wt ๐‘›0 + wt ๐‘›1 ) = (1 + ๐‘ฅ)๐‘› . Since this was the readerโ€™s first example of the use of weight-generating functions, we were careful to write out all the details. However, in practice one is usually more concise, for example, making no distinction between ๐‘† and ๐‘† โ€ฒ with the understanding that they yield the same weight-generating function and so can be considered the same set in this situation. We also usually omit checking summability, assuming that such details could be filled in if necessary. We are now ready for a more substantial example, namely the Binomial Theorem for negative exponents. Theorem 3.4.2. If ๐‘› โˆˆ โ„•, then 1 ๐‘› = โˆ‘ (( ))๐‘ฅ๐‘˜ . (1 โˆ’ ๐‘ฅ)๐‘› ๐‘˜โ‰ฅ0 ๐‘˜ Proof. The summation side suggests that we should consider ๐‘† = {๐‘‡ โˆฃ ๐‘‡ is a multiset on [๐‘›]} with weight function given by (3.10). We are rewarded for our choice since ๐‘“๐‘† (๐‘ฅ) = โˆ‘ wt ๐‘‡ = โˆ‘ ๐‘‡โˆˆ๐‘†

๐‘˜โ‰ฅ0

๐‘› โˆ‘ ๐‘ฅ๐‘˜ = โˆ‘ (( ))๐‘ฅ๐‘˜ . ๐‘˜ ๐‘˜โ‰ฅ0 ๐‘‡โˆˆ(([๐‘›] )) ๐‘˜

We now write ๐‘† = {(1๐‘š1 , 2๐‘š2 , . . . , ๐‘›๐‘š๐‘› ) โˆฃ ๐‘š๐‘– โ‰ฅ 0 for all ๐‘–} = (10 โŠŽ 11 โŠŽ 12 โŠŽ . . . ) ร— (20 โŠŽ 21 โŠŽ 22 โŠŽ . . . ) ร— โ‹ฏ ร— (๐‘›0 โŠŽ ๐‘›1 โŠŽ ๐‘›2 โŠŽ . . . ) with weight function wt ๐‘–๐‘˜ = ๐‘ฅ๐‘˜ . Using Lemma 3.4.1 yields ๐‘“๐‘† (๐‘ฅ) = (wt 10 + wt 11 + wt 12 + โ‹ฏ) โ‹ฏ (wt ๐‘›0 + wt ๐‘›1 + wt ๐‘›2 + โ‹ฏ) = (1 + ๐‘ฅ + ๐‘ฅ2 + โ‹ฏ)๐‘› 1 = (1 โˆ’ ๐‘ฅ)๐‘› and the theorem is proved.

โ–ก

There are several remarks which should be made about this result. First of all, contrast it with our first version of the Binomial Theorem. In Theorem 3.1.1 we are counting subsets of [๐‘›] where repeated elements are not allowed and the resulting generating function is (1 + ๐‘ฅ)๐‘› . In Theorem 3.4.2 we are counting multisets on [๐‘›] so that repetitions are allowed and these are counted by 1/(1 โˆ’ ๐‘ฅ)๐‘› . We will see another example of this in the next section. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

3.5. Revisiting integer partitions

89

We can also make Theorem 3.4.2 look almost exactly like Theorem 3.1.1. Indeed, if ๐‘› โ‰ค 0, then by Theorem 3.4.2 and equation (1.6) (with โˆ’๐‘› substituted for ๐‘›) we have (1 + ๐‘ฅ)๐‘› =

1 โˆ’๐‘› ๐‘› = โˆ‘ (( ))(โˆ’๐‘ฅ)๐‘˜ = โˆ‘ ( )๐‘ฅ๐‘˜ . ๐‘˜ (1 โˆ’ (โˆ’๐‘ฅ))โˆ’๐‘› ๐‘˜โ‰ฅ0 ๐‘˜ ๐‘˜โ‰ฅ0

This is exactly like Theorem 3.1.1 except that we have an infinite series whereas for positive ๐‘› we have a polynomial. Analytically, the Binomial Theorem makes sense for any ๐‘› โˆˆ โ„‚ as long as |๐‘ฅ| < 1 so that the series converges. In โ„‚[[๐‘ฅ]] one can make sense of (1 + ๐‘ฅ)๐‘› for any rational ๐‘› โˆˆ โ„š; see Exercise 12 of this chapter, and prove the following. Theorem 3.4.3. For any ๐‘› โˆˆ โ„š we have ๐‘› (1 + ๐‘ฅ)๐‘› = โˆ‘ ( )๐‘ฅ๐‘˜ . ๐‘˜ ๐‘˜โ‰ฅ0

โ–ก

3.5. Revisiting integer partitions The theory of integer partitions is one place where ordinary generating functions have played a central role. In this context and others it will be necessary to consider infinite products. But then, as we have seen in the previous section, we must take care that these products converge. There is a corresponding restriction on the sets which we can use to construct weight-generating functions. We begin by discussing this matter. Let ๐‘† be a weighted set. We say that ๐‘† is rooted if there is an element ๐‘Ÿ โˆˆ ๐‘† called the root satisfying (1) wt ๐‘Ÿ = 1 and (2) if ๐‘  โˆˆ ๐‘† โˆ’ {๐‘Ÿ}, then wt ๐‘  has zero constant term. For example, the sets (๐‘›0 , ๐‘›1 , ๐‘›2 , . . . ) used in the proof of Theorem 3.4.2 were rooted with ๐‘Ÿ = ๐‘›0 since wt ๐‘›0 = 1 and wt ๐‘›๐‘˜ = ๐‘ฅ๐‘˜ for ๐‘˜ โ‰ฅ 1. Given a sequence ๐‘† 1 , ๐‘† 2 , ๐‘† 3 . . . of rooted sets with ๐‘† ๐‘– having root ๐‘Ÿ ๐‘– , their direct sum is defined to be ๐‘†1 โŠ• ๐‘†2 โŠ• ๐‘†3 โŠ• โ‹ฏ = {(๐‘ 1 , ๐‘ 2 , ๐‘ 3 , . . . ) โˆฃ ๐‘ ๐‘– โˆˆ ๐‘† ๐‘– for all ๐‘– and ๐‘ ๐‘– โ‰  ๐‘Ÿ ๐‘– for only finitely many ๐‘–}. Note that when the number of ๐‘† ๐‘– is finite, then their direct sum is the same as their product. But when their number is infinite the root condition kicks in. Note that, because of this condition, we have a well-defined weighting on โŠ•๐‘–โ‰ฅ1 ๐‘† ๐‘– given by wt(๐‘ 1 , ๐‘ 2 , ๐‘ 3 , . . . ) = โˆ wt ๐‘ ๐‘– ๐‘–โ‰ฅ1

since the product has only finitely many factors not equal to 1. In addition, the Product Rule in Lemma 3.4.1 has to be modified appropriately to get convergence. But the proof is similar to the former result and so is left as an exercise. Theorem 3.5.1. Let ๐‘† 1 , ๐‘† 2 , ๐‘† 3 , . . . be a sequence of summable, rooted sets such that lim [mdeg(๐‘“๐‘†๐‘– (๐‘ฅ) โˆ’ 1)] = โˆž.

๐‘–โ†’โˆž

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

90

3. Counting with Ordinary Generating Functions

Then the direct sum ๐‘† 1 โŠ• ๐‘† 2 โŠ• ๐‘† 3 โŠ• โ‹ฏ is summable and ๐‘“๐‘†1 โŠ•๐‘†2 โŠ•๐‘†3 โŠ•โ‹ฏ (๐‘ฅ) = โˆ ๐‘“๐‘†๐‘– (๐‘ฅ).

โ–ก

๐‘–โ‰ฅ1

We will now prove a theorem of Euler giving the generating function for ๐‘(๐‘›), the number of integer partitions of ๐‘›. The reader should contrast the proof with that given for counting multisets in Theorem 3.4.2, which has evident parallels. Theorem 3.5.2. We have โˆ‘ ๐‘(๐‘›)๐‘ฅ๐‘› = โˆ ๐‘›โ‰ฅ0

๐‘–โ‰ฅ1

1 . 1 โˆ’ ๐‘ฅ๐‘–

Proof. Motivated by the sum side we consider the set ๐‘† of all integer partitions ๐œ† of all numbers ๐‘› โ‰ฅ 0 with weight wt ๐œ† = ๐‘ฅ|๐œ†| ,

(3.11)

recalling that |๐œ†| is the sum of the parts of ๐œ†. It follows that ๐‘“๐‘† (๐‘ฅ) = โˆ‘ wt ๐œ† = โˆ‘ โˆ‘ ๐‘ฅ๐‘› = โˆ‘ ๐‘(๐‘›)๐‘ฅ๐‘› . ๐œ†โˆˆ๐‘†

๐‘›โ‰ฅ0 |๐œ†|=๐‘›

๐‘›โ‰ฅ0

We now express ๐‘† as a direct sum, using multiplicity notation, as ๐‘† = {(1๐‘š1 , 2๐‘š2 , 3๐‘š3 , . . . ) โˆฃ ๐‘š๐‘– โ‰ฅ 0 for all ๐‘– and only finitely many ๐‘š๐‘– โ‰  0} = (10 โŠŽ 11 โŠŽ 12 โŠŽ โ‹ฏ) โŠ• (20 โŠŽ 21 โŠŽ 22 โŠŽ โ‹ฏ) โŠ• (30 โŠŽ 31 โŠŽ 32 โŠŽ โ‹ฏ) โŠ• โ‹ฏ . Note that since we want the exponent on wt ๐œ† to be the sum of its parts, and ๐‘–๐‘˜ represents a part ๐‘– repeated ๐‘˜ times, we must take wt ๐‘–๐‘˜ = ๐‘ฅ๐‘–๐‘˜ in contrast to the weight used in the proof of Theorem 3.4.2. Translating into generating functions by using the previous theorem gives ๐‘“๐‘† (๐‘ฅ) = โˆ(wt ๐‘–0 + wt ๐‘–1 + wt ๐‘–2 + wt ๐‘–3 + โ‹ฏ) ๐‘–โ‰ฅ1

= โˆ(1 + ๐‘ฅ๐‘– + ๐‘ฅ2๐‘– + ๐‘ฅ3๐‘– + โ‹ฏ) ๐‘–โ‰ฅ1

=โˆ ๐‘–โ‰ฅ1

1 , 1 โˆ’ ๐‘ฅ๐‘–

which is the desired result.

โ–ก

The reader will notice that the previous proof actually shows much more. In particular, the factor 1/(1 โˆ’ ๐‘ฅ๐‘– ) is responsible for keeping track of the parts equal to ๐‘– in ๐œ†. We can make this precise as follows. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

3.5. Revisiting integer partitions

91

Proposition 3.5.3. Given ๐‘› โˆˆ โ„• and ๐‘ƒ โŠ† โ„™, let ๐‘๐‘ƒ (๐‘›) be the number of partitions of ๐‘› all of whose parts are in ๐‘ƒ. (a) We have โˆ‘ ๐‘๐‘ƒ (๐‘›)๐‘ฅ๐‘› = โˆ ๐‘›โ‰ฅ0

๐‘–โˆˆ๐‘ƒ

1 . 1 โˆ’ ๐‘ฅ๐‘–

(b) In particular, for ๐‘˜ โˆˆ โ„™, โˆ‘ ๐‘ [๐‘˜] (๐‘›)๐‘ฅ๐‘› = ๐‘›โ‰ฅ0

1 . (1 โˆ’ ๐‘ฅ)(1 โˆ’ ๐‘ฅ2 ) โ‹ฏ (1 โˆ’ ๐‘ฅ๐‘˜ )

Proof. For (a), one uses the ideas in the proof of Theorem 3.5.2 except that the elements of ๐‘† only contain components of the form ๐‘Ÿ๐‘š๐‘Ÿ for ๐‘Ÿ โˆˆ ๐‘ƒ. And (b) follows immediately from (a). โ–ก Instead of restricting the set of parts of a partition, we can restrict the number of parts. Recall that ๐‘(๐‘›, ๐‘˜) is the number of ๐œ† โŠข ๐‘› with length โ„“(๐œ†) โ‰ค ๐‘˜. Corollary 3.5.4. For ๐‘˜ โ‰ฅ 0 we have โˆ‘ ๐‘(๐‘›, ๐‘˜)๐‘ฅ๐‘› = ๐‘›โ‰ฅ0

1 . (1 โˆ’ ๐‘ฅ)(1 โˆ’ ๐‘ฅ2 ) โ‹ฏ (1 โˆ’ ๐‘ฅ๐‘˜ )

Proof. From the previous result, it suffices to show that there is a size-preserving bijection between the partitions counted by ๐‘ [๐‘˜] (๐‘›) and those counted by ๐‘(๐‘›, ๐‘˜). The map ๐œ† โ†’ ๐œ†๐‘ก is such a map. Indeed, ๐œ† only uses parts in [๐‘˜] if and only if ๐œ†1 โ‰ค ๐‘˜. In terms of Young diagrams, this means that the first row of ๐œ† has length at most ๐‘˜. It follows that the first column of ๐œ†๐‘ก has length at most ๐‘˜, which is equivalent to ๐œ†๐‘ก having at most ๐‘˜ parts. โ–ก In the previous section we pointed out a relationship between the generating functions for sets and for multisets. The same holds for integer partitions. Let ๐‘ ๐‘‘ (๐‘›) be the number of partitions of ๐‘› into distinct parts as defined in Section 2.3. Theorem 3.5.5. We have โˆ‘ ๐‘ ๐‘‘ (๐‘›)๐‘ฅ๐‘› = โˆ(1 + ๐‘ฅ๐‘– ). ๐‘›โ‰ฅ0

๐‘–โ‰ฅ1

Proof. Up to now we have been writing out most of the gory details of our proofs by weight-generating function as the reader gets familiar with the method. But by now it should be sufficient just to write out the highlights. We begin by letting ๐‘† be all partitions of all ๐‘› โˆˆ โ„• into distinct parts and using the weighting in (3.11). It is routine to show that ๐‘“๐‘† (๐‘ฅ) results in the sum side of the theorem. To get the product side we write ๐‘† = {(1๐‘š1 , 2๐‘š2 , 3๐‘š3 , . . . ) โˆฃ ๐‘š๐‘– = 0 or 1 for all ๐‘–, only finitely many ๐‘š๐‘– โ‰  0} =

โจ

(๐‘–0 โŠŽ ๐‘–1 ).

๐‘–โ‰ฅ1

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

92

3. Counting with Ordinary Generating Functions

The generating function translation is ๐‘“๐‘† (๐‘ฅ) = โˆ(1 + ๐‘ฅ๐‘– ) ๐‘–โ‰ฅ1

โ–ก

and we are done.

As mentioned in the introduction to this chapter, one of the reasons for using generating functions is that they can give quick and easy proofs for various results. Here is an example where we reprove Eulerโ€™s distinct parts-odd parts result, Theorem 2.3.3, which we restate here for convenience. Let ๐‘๐‘œ (๐‘›) be the number of partitions of ๐‘› into parts all of which are odd. Theorem 3.5.6 (Euler). For all ๐‘› โ‰ฅ 0, ๐‘๐‘œ (๐‘›) = ๐‘ ๐‘‘ (๐‘›). Proof. It suffices to show that these two sequence have the same generating function. Using Theorems 3.5.3(a) and 3.5.5 as well as multiplying by a strange name for one, we get โˆ‘ ๐‘ ๐‘‘ (๐‘›)๐‘ฅ๐‘› = (1 + ๐‘ฅ)(1 + ๐‘ฅ2 )(1 + ๐‘ฅ3 ) โ‹ฏ ๐‘›โ‰ฅ0

= (1 + ๐‘ฅ)(1 + ๐‘ฅ2 )(1 + ๐‘ฅ3 ) โ‹ฏ

(1 โˆ’ ๐‘ฅ)(1 โˆ’ ๐‘ฅ2 )(1 โˆ’ ๐‘ฅ3 ) โ‹ฏ (1 โˆ’ ๐‘ฅ)(1 โˆ’ ๐‘ฅ2 )(1 โˆ’ ๐‘ฅ3 ) โ‹ฏ

=

(1 โˆ’ ๐‘ฅ2 )(1 โˆ’ ๐‘ฅ4 )(1 โˆ’ ๐‘ฅ6 ) โ‹ฏ (1 โˆ’ ๐‘ฅ)(1 โˆ’ ๐‘ฅ2 )(1 โˆ’ ๐‘ฅ3 ) โ‹ฏ

=

1 (1 โˆ’ ๐‘ฅ)(1 โˆ’ ๐‘ฅ3 )(1 โˆ’ ๐‘ฅ5 ) โ‹ฏ

= โˆ‘ ๐‘๐‘œ (๐‘›)๐‘ฅ๐‘› , ๐‘›โ‰ฅ0

which completes this short and slick proof.

โ–ก

3.6. Recurrence relations and generating functions The reader may have noticed that many of the combinatorial sequences described in Chapter 1 satisfy recurrence relations. If one has a sequence defined by a recursion, then generating functions can often be used to find an explicit expression for the terms of the sequence. It is also possible to glean information from the generating function derived from a recurrence which is hard to extract from the recurrence itself. This section is devoted to exploring these ideas. We start with a simple algorithm using generating functions to solve a recurrence relation. Given a sequence ๐‘Ž0 , ๐‘Ž1 , ๐‘Ž2 , . . . defined by a recursion and boundary conditions, we wish to find a self-contained formula for the ๐‘›th term. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

3.6. Recurrence relations and generating functions

93

(1) Multiply the recurrence by ๐‘ฅ๐‘› ; usually a good choice for ๐‘› is the largest index of all the terms in the recurrence. Sum over all ๐‘› โ‰ฅ ๐‘‘ where ๐‘‘ is the smallest index for which the recurrence is valid. (2) Let ๐‘“(๐‘ฅ) = โˆ‘ ๐‘Ž๐‘› ๐‘ฅ๐‘› ๐‘›โ‰ฅ0

and express the equation in step (1) in terms of ๐‘“(๐‘ฅ) using the boundary conditions. (3) Solve for ๐‘“(๐‘ฅ). (4) Find ๐‘Ž๐‘› as the coefficient of ๐‘ฅ๐‘› in ๐‘“(๐‘ฅ). We note that partial fraction expansion can be a useful way to accomplish step (4). For a simple example, suppose that our sequence is defined by ๐‘Ž0 = 2 and ๐‘Ž๐‘› = 3๐‘Ž๐‘›โˆ’1 for ๐‘› โ‰ฅ 1. Calculating the first few values we get ๐‘Ž1 = 2 โ‹… 3, ๐‘Ž2 = 2 โ‹… 32 , ๐‘Ž3 = 2 โ‹… 33 . So it is easy to guess and then prove by induction that ๐‘Ž๐‘› = 2 โ‹… 3๐‘› . We would now like to obtain this result using generating functions. Step (1) is easy as we just write ๐‘Ž๐‘› ๐‘ฅ๐‘› = 3๐‘Ž๐‘›โˆ’1 ๐‘ฅ๐‘› and then sum to get โˆ‘ ๐‘Ž๐‘› ๐‘ฅ๐‘› = โˆ‘ 3๐‘Ž๐‘›โˆ’1 ๐‘ฅ๐‘› . ๐‘›โ‰ฅ1

๐‘›โ‰ฅ1

Letting ๐‘“(๐‘ฅ) be as in step (2) we see that โˆ‘ ๐‘Ž๐‘› ๐‘ฅ๐‘› = ๐‘“(๐‘ฅ) โˆ’ ๐‘Ž0 = ๐‘“(๐‘ฅ) โˆ’ 2 ๐‘›โ‰ฅ1

and โˆ‘ 3๐‘Ž๐‘›โˆ’1 ๐‘ฅ๐‘› = 3๐‘ฅ โˆ‘ ๐‘Ž๐‘›โˆ’1 ๐‘ฅ๐‘›โˆ’1 = 3๐‘ฅ๐‘“(๐‘ฅ) ๐‘›โ‰ฅ1

๐‘›โ‰ฅ1

where the last equality is obtained by substituting ๐‘› for ๐‘› โˆ’ 1 in the sum. For step (3) we have 2 . (3.12) ๐‘“(๐‘ฅ) โˆ’ 2 = 3๐‘ฅ๐‘“(๐‘ฅ) โŸน ๐‘“(๐‘ฅ) โˆ’ 3๐‘ฅ๐‘“(๐‘ฅ) = 2 โŸน ๐‘“(๐‘ฅ) = 1 โˆ’ 3๐‘ฅ As far as step (4), we can now expand 1/(1 โˆ’ 3๐‘ฅ) as a geometric series (that is, use (3.8) and substitute 3๐‘ฅ for ๐‘ฅ) to obtain ๐‘“(๐‘ฅ) = 2 โˆ‘ 3๐‘› ๐‘ฅ๐‘› = โˆ‘ 2 โ‹… 3๐‘› ๐‘ฅ๐‘› . ๐‘›โ‰ฅ0

๐‘›โ‰ฅ0

๐‘›

Extracting the coefficient of ๐‘ฅ we see that ๐‘Ž๐‘› = 2 โ‹… 3๐‘› as expected. In the previous example, it was easier to guess the formula for ๐‘Ž๐‘› and then prove it by induction rather than use generating functions. However, there are times when it is impossible to guess the solution this way, but generating functions still give a straightforward method for obtaining the answer. An example of this is given by the Fibonacci sequence. Our result will be slightly nicer if we use the definition of this sequence given by (1.1). Following the algorithm, we write โˆ‘ ๐น๐‘› ๐‘ฅ๐‘› = โˆ‘ (๐น๐‘›โˆ’1 + ๐น๐‘›โˆ’2 )๐‘ฅ๐‘› . ๐‘›โ‰ฅ2

๐‘›โ‰ฅ2

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

94

3. Counting with Ordinary Generating Functions

Writing ๐‘“(๐‘ฅ) = โˆ‘๐‘›โ‰ฅ0 ๐น๐‘› ๐‘ฅ๐‘› we obtain โˆ‘ ๐น๐‘› ๐‘ฅ๐‘› = ๐‘“(๐‘ฅ) โˆ’ ๐น0 โˆ’ ๐น1 ๐‘ฅ = ๐‘“(๐‘ฅ) โˆ’ ๐‘ฅ ๐‘›โ‰ฅ2

and โˆ‘ (๐น๐‘›โˆ’1 + ๐น๐‘›โˆ’2 )๐‘ฅ๐‘› = ๐‘ฅ(๐‘“(๐‘ฅ) โˆ’ ๐น0 ) + ๐‘ฅ2 ๐‘“(๐‘ฅ) = (๐‘ฅ + ๐‘ฅ2 )๐‘“(๐‘ฅ). ๐‘›โ‰ฅ2

Setting the expressions for the left and right sides equal and solving for ๐‘“(๐‘ฅ) yields ๐‘ฅ ๐‘“(๐‘ฅ) = . 1 โˆ’ ๐‘ฅ โˆ’ ๐‘ฅ2 For the last step we wish to use partial fractions and so must factor 1 โˆ’ ๐‘ฅ โˆ’ ๐‘ฅ2 . Using the quadratic formula, we see that the denominator has roots ๐‘Ÿ1 =

โˆ’1 + โˆš5 2

and

๐‘Ÿ2 =

โˆ’1 โˆ’ โˆš5 . 2

It follows that 1 โˆ’ ๐‘ฅ โˆ’ ๐‘ฅ2 = (1 โˆ’

๐‘ฅ ๐‘ฅ )(1 โˆ’ ) ๐‘Ÿ1 ๐‘Ÿ2

since both sides vanish at ๐‘ฅ = ๐‘Ÿ1 , ๐‘Ÿ2 and both sides have constant term 1. So we have the partial fraction decomposition ๐‘ฅ ๐ด ๐ต (3.13) ๐‘“(๐‘ฅ) = = + (1 โˆ’

๐‘ฅ )(1 ๐‘Ÿ1

โˆ’

๐‘ฅ ) ๐‘Ÿ2

(1 โˆ’

๐‘ฅ ) ๐‘Ÿ1

(1 โˆ’

๐‘ฅ ) ๐‘Ÿ2

for constants ๐ด, ๐ต. Clearing denominators gives ๐‘ฅ = ๐ด(1 โˆ’

๐‘ฅ ๐‘ฅ ) + ๐ต(1 โˆ’ ). ๐‘Ÿ2 ๐‘Ÿ1

Setting ๐‘ฅ = ๐‘Ÿ1 reduces this equation to ๐‘Ÿ1 = ๐ด(1 โˆ’ ๐‘Ÿ1 /๐‘Ÿ2 ) and solving for ๐ด shows that ๐ด = 1/โˆš5. Similarly letting ๐‘ฅ = ๐‘Ÿ2 yields ๐ต = โˆ’1/โˆš5. Plugging these values back into (3.13) and expanding the series, ๐‘“(๐‘ฅ) =

1 ๐‘ฅ๐‘› ๐‘ฅ๐‘› โ‹… โˆ‘ ๐‘›. ๐‘› โˆ’ โˆš5 ๐‘›โ‰ฅ0 ๐‘Ÿ1 โˆš5 ๐‘›โ‰ฅ0 ๐‘Ÿ2 1

โ‹…โˆ‘

By rationalizing denominators one can check that 1/๐‘Ÿ1 = (1 + โˆš5)/2 and 1/๐‘Ÿ2 = (1 โˆ’ โˆš5)/2. So taking the coefficient of ๐‘ฅ๐‘› on both sides of the previous displayed equation gives (3.14)

๐น๐‘› =

๐‘›

๐‘›

1 + โˆš5 1 1 โˆ’ โˆš5 ) โˆ’ ) . ( 2 2 โˆš5 โˆš5 1

(

This example shows the true power of the generating function method. It would be impossible to guess the formula in (3.14) from just computing values of ๐น๐‘› . In fact, it is not even obvious that the right-hand side is an integer! Our algorithm can be used to derive generating functions in the case where one has a triangle of numbers rather than just a sequence. Here we illustrate this using the Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

3.6. Recurrence relations and generating functions

95

Stirling numbers. Recall that the signless Stirling numbers of the first kind satisfy the recurrence relation and boundary conditions in Theorem 1.5.2. Translating these to the signed version gives ๐‘ (0, ๐‘˜) = ๐›ฟ0,๐‘˜ and ๐‘ (๐‘›, ๐‘˜) = ๐‘ (๐‘› โˆ’ 1, ๐‘˜ โˆ’ 1) โˆ’ (๐‘› โˆ’ 1)๐‘ (๐‘› โˆ’ 1, ๐‘˜) for ๐‘› โ‰ฅ 1. We wish to find the generating function ๐‘“๐‘› (๐‘ฅ) = โˆ‘๐‘˜ ๐‘ (๐‘›, ๐‘˜)๐‘ฅ๐‘˜ where we are using the fact that ๐‘ (๐‘›, ๐‘˜) = 0 for ๐‘˜ < 0 or ๐‘˜ > ๐‘› to sum over all integers ๐‘›. Applying our algorithm we have ๐‘“๐‘› (๐‘ฅ) = โˆ‘ ๐‘ (๐‘›, ๐‘˜)๐‘ฅ๐‘˜ ๐‘˜

= โˆ‘[๐‘ (๐‘› โˆ’ 1, ๐‘˜ โˆ’ 1) โˆ’ (๐‘› โˆ’ 1)๐‘ (๐‘› โˆ’ 1, ๐‘˜)]๐‘ฅ๐‘˜ ๐‘˜

= ๐‘ฅ๐‘“๐‘›โˆ’1 (๐‘ฅ) โˆ’ (๐‘› โˆ’ 1)๐‘“๐‘›โˆ’1 (๐‘ฅ) = (๐‘ฅ โˆ’ ๐‘› + 1)๐‘“๐‘›โˆ’1 (๐‘ฅ), giving us a recursion for the sequence of generating functions ๐‘“๐‘› (๐‘ฅ). From the boundary condition for ๐‘ (0, ๐‘˜) we have ๐‘“0 (๐‘ฅ) = 1. It is now easy to guess a formula for ๐‘“๐‘› (๐‘ฅ) by writing out the first few values and proving that pattern holds by induction to obtain the theorem below, which also follows easily from Theorem 3.1.2. Theorem 3.6.1. For ๐‘› โ‰ฅ 0 we have โˆ‘ ๐‘ (๐‘›, ๐‘˜)๐‘ฅ๐‘˜ = ๐‘ฅ(๐‘ฅ โˆ’ 1) โ‹ฏ (๐‘ฅ โˆ’ ๐‘› + 1).

โ–ก

๐‘˜

In an entirely analogous manner, one can obtain a generating function for the Stirling numbers of the second kind. Because of the similarity, the proof is left to the reader. Theorem 3.6.2. For ๐‘˜ โ‰ฅ 0 we have โˆ‘ ๐‘†(๐‘›, ๐‘˜)๐‘ฅ๐‘› = ๐‘›

๐‘ฅ๐‘˜ . (1 โˆ’ ๐‘ฅ)(1 โˆ’ 2๐‘ฅ) โ‹ฏ (1 โˆ’ ๐‘˜๐‘ฅ)

โ–ก

Comparing the previous two results, the reader will note a similar relationship as between generating functions for objects without repetitions (sets, distinct partitions) and those where repetitions are allowed (multisets, ordinary partitions). As already mentioned, this will be explained in Section 3.9. So far, all the generating functions we have derived from recurrences have been rational functions. This is because the recursions are linear and we will prove a general result to this effect in the next section. We will end this section by illustrating that more complicated generating functions, for example algebraic ones, do arise in practice. Let us consider the Catalan numbers ๐ถ(๐‘›) and the generating function ๐‘(๐‘ฅ) = โˆ‘๐‘›โ‰ฅ0 ๐ถ(๐‘›)๐‘ฅ๐‘› . Using the recursion and boundary condition in Theorem 1.11.2 and computing in the way we have become accustomed to, we obtain ๐‘(๐‘ฅ) = 1 + โˆ‘ ๐ถ(๐‘›)๐‘ฅ๐‘› = 1 + โˆ‘ ( ๐‘›โ‰ฅ1

โˆ‘

๐ถ(๐‘–)๐ถ(๐‘—))๐‘ฅ๐‘› = 1 + ๐‘ฅ๐‘(๐‘ฅ)2 .

๐‘›โ‰ฅ1 ๐‘–+๐‘—=๐‘›โˆ’1

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

96

3. Counting with Ordinary Generating Functions

Writing ๐‘ฅ๐‘(๐‘ฅ)2 โˆ’ ๐‘(๐‘ฅ) + 1 = 0 and solving for ๐‘(๐‘ฅ) using the quadratic formula yields ๐‘(๐‘ฅ) =

1 ยฑ โˆš1 โˆ’ 4๐‘ฅ . 2๐‘ฅ

Two things seem to be wrong with this formula for ๐‘(๐‘ฅ). First of all, we donโ€™t know whether the plus or minus solution is the correct one. And second, we seem to have left the ring of formal power series because we are dividing by ๐‘ฅ which has no inverse. Both of these can be solved simultaneously by choosing the sign so that the numerator has no constant term. Then one can divide by ๐‘ฅ simply by reducing the power of each term in the top by one. By Theorem 3.4.3 we see that the generating function for โˆš1 โˆ’ 4๐‘ฅ = (1 โˆ’ 4๐‘ฅ)1/2 has constant term (1/2 ) = 1. So the correct sign is 0 negative and we have proved the following. Theorem 3.6.3. We have โˆ‘ ๐ถ(๐‘›)๐‘ฅ๐‘› = ๐‘›โ‰ฅ0

1 โˆ’ โˆš1 โˆ’ 4๐‘ฅ . 2๐‘ฅ

โ–ก

One can use this generating function to rederive the explicit expression for ๐ถ(๐‘›) in Theorem 1.11.3, and the reader will be asked to carry out the details in the exercises.

3.7. Rational generating functions and linear recursions The reader may have noticed in the previous section that, both in the initial example and for the Fibonacci sequence, the solution of the recursion for ๐‘Ž๐‘› was a linear combination of functions of the form ๐‘Ÿ๐‘› where ๐‘Ÿ varied over the reciprocals of the roots of the denominator of the corresponding generating function. This happens for a wide variety of recursions which we will study in this section. Before giving a theorem which characterizes this situation, we will study one more example to illustrate what can happen. Consider the sequence defined by ๐‘Ž0 = 1, ๐‘Ž1 = โˆ’4, and (3.15)

๐‘Ž๐‘› = 4๐‘Ž๐‘›โˆ’1 โˆ’ 4๐‘Ž๐‘›โˆ’2

for ๐‘› โ‰ฅ 2.

Following the usual four-step program, we have, for ๐‘“(๐‘ฅ) = โˆ‘๐‘›โ‰ฅ0 ๐‘Ž๐‘› ๐‘ฅ๐‘› , ๐‘“(๐‘ฅ) โˆ’ 1 + 4๐‘ฅ = โˆ‘ ๐‘Ž๐‘› ๐‘ฅ๐‘› ๐‘›โ‰ฅ2

= 4๐‘ฅ โˆ‘ ๐‘Ž๐‘›โˆ’1 ๐‘ฅ๐‘›โˆ’1 โˆ’ 4๐‘ฅ2 โˆ‘ ๐‘Ž๐‘›โˆ’2 ๐‘ฅ๐‘›โˆ’2 ๐‘›โ‰ฅ2

๐‘›โ‰ฅ2 2

= 4๐‘ฅ(๐‘“(๐‘ฅ) โˆ’ 1) โˆ’ 4๐‘ฅ ๐‘“(๐‘ฅ). Solving for ๐‘“(๐‘ฅ) and evaluating the constants in the partial fraction expansion yields ๐‘“(๐‘ฅ) =

1 โˆ’ 8๐‘ฅ 1 โˆ’ 8๐‘ฅ 3 4 = โˆ’ = . 1 โˆ’ 2๐‘ฅ (1 โˆ’ 2๐‘ฅ)2 1 โˆ’ 4๐‘ฅ + 4๐‘ฅ2 (1 โˆ’ 2๐‘ฅ)2

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

3.7. Rational generating functions and linear recursions

97

Taking the coefficient of ๐‘ฅ๐‘› on both sides using the Theorem 3.4.2 (interchanging the roles of ๐‘› and ๐‘˜) together with the fact that ๐‘˜ ๐‘›+๐‘˜โˆ’1 ๐‘›+๐‘˜โˆ’1 (( )) = ( )=( ) ๐‘› ๐‘› ๐‘˜โˆ’1

(3.16) gives a final answer of

๐‘›+1 ๐‘› ๐‘Ž๐‘› = 4 โ‹… 2๐‘› โˆ’ 3( )2 = (1 โˆ’ 3๐‘›)2๐‘› . 1

(3.17)

So now, instead of a constant times ๐‘Ÿ๐‘› we have a polynomial in ๐‘› as the coefficient. And that polynomial has degree less than the multiplicity of 1/๐‘Ÿ as a root of the denominator. These observations generalize. Consider a sequence of complex numbers ๐‘Ž๐‘› for ๐‘› โ‰ฅ 0. We say that the sequence satisfies a (homogeneous) linear recursion of degree ๐‘‘ with constant coefficients if there is a ๐‘‘ โˆˆ โ„™ and constants ๐‘ 1 , . . . , ๐‘ ๐‘‘ โˆˆ โ„‚ with ๐‘ ๐‘‘ โ‰  0 such that (3.18)

๐‘Ž๐‘›+๐‘‘ + ๐‘ 1 ๐‘Ž๐‘›+๐‘‘โˆ’1 + ๐‘ 2 ๐‘Ž๐‘›+๐‘‘โˆ’2 + โ‹ฏ + ๐‘ ๐‘‘ ๐‘Ž๐‘› = 0.

To simplify things later, we have put all the terms of the recursion on the left-hand side of the equation and made ๐‘Ž๐‘›+๐‘‘ the term of highest index rather than ๐‘Ž๐‘› . One can also consider the nonhomogeneous case where one has a summand ๐‘ ๐‘‘+1 which does not multiply any term of the sequence, but we will have no cause to do so here. It turns out that the sequences satisfying a recursion (3.18) are exactly the ones having rational generating functions. Theorem 3.7.1. Given a sequence ๐‘Ž๐‘› for ๐‘› โ‰ฅ 0 and ๐‘‘ โˆˆ โ„™, the following are equivalent. (a) The sequence satisfies (3.18). (b) The generating function ๐‘“(๐‘ฅ) = โˆ‘๐‘›โ‰ฅ0 ๐‘Ž๐‘› ๐‘ฅ๐‘› has the form (3.19)

๐‘“(๐‘ฅ) =

๐‘(๐‘ฅ) ๐‘ž(๐‘ฅ)

where ๐‘ž(๐‘ฅ) = 1 + ๐‘ 1 ๐‘ฅ + ๐‘ 2 ๐‘ฅ2 + โ‹ฏ + ๐‘ ๐‘‘ ๐‘ฅ๐‘‘

(3.20)

and deg ๐‘(๐‘ฅ) < ๐‘‘. (c) We can write ๐‘˜

๐‘Ž๐‘› = โˆ‘ ๐‘ ๐‘– (๐‘›)๐‘Ÿ๐‘–๐‘› ๐‘–=1

where the ๐‘Ÿ ๐‘– are distinct, nonzero complex numbers satisfying ๐‘˜

(3.21)

1 + ๐‘ 1 ๐‘ฅ + ๐‘ 2 ๐‘ฅ2 + โ‹ฏ + ๐‘ ๐‘‘ ๐‘ฅ๐‘‘ = โˆ(1 โˆ’ ๐‘Ÿ ๐‘– ๐‘ฅ)๐‘‘๐‘– ๐‘–=1

and ๐‘ ๐‘– (๐‘›) is a polynomial with deg ๐‘ ๐‘– (๐‘›) < ๐‘‘๐‘– for all ๐‘–. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

98

3. Counting with Ordinary Generating Functions

Proof. We first prove the equivalence of (a) and (b). Showing that (a) implies (b) is essentially an application of our algorithm. Multiplying (3.18) by ๐‘ฅ๐‘›+๐‘‘ and summing over ๐‘› โ‰ฅ 0 gives 0 = โˆ‘ ๐‘Ž๐‘›+๐‘‘ ๐‘ฅ๐‘›+๐‘‘ + ๐‘ 1 ๐‘ฅ โˆ‘ ๐‘Ž๐‘›+๐‘‘โˆ’1 ๐‘ฅ๐‘›+๐‘‘โˆ’1 + โ‹ฏ + ๐‘ ๐‘‘ ๐‘ฅ๐‘‘ โˆ‘ ๐‘Ž๐‘› ๐‘ฅ๐‘› ๐‘›โ‰ฅ0

๐‘›โ‰ฅ0

๐‘›โ‰ฅ0 ๐‘‘โˆ’2

๐‘‘โˆ’1

= [๐‘“(๐‘ฅ) โˆ’ โˆ‘ ๐‘Ž๐‘› ๐‘ฅ๐‘› ] + ๐‘ 1 ๐‘ฅ [๐‘“(๐‘ฅ) โˆ’ โˆ‘ ๐‘Ž๐‘› ๐‘ฅ๐‘› ] + โ‹ฏ + ๐‘ ๐‘‘ ๐‘ฅ๐‘‘ ๐‘“(๐‘ฅ) ๐‘›=0

๐‘›=0

= ๐‘ž(๐‘ฅ)๐‘“(๐‘ฅ) โˆ’ ๐‘(๐‘ฅ) where ๐‘ž(๐‘ฅ) is given by (3.20) and ๐‘(๐‘ฅ) is the sum of the remaining terms, which implies deg ๐‘(๐‘ฅ) < ๐‘‘. Solving for ๐‘“(๐‘ฅ) completes this direction. To prove (b) implies (a), cross multiply (3.19) and use (3.20) to write ๐‘(๐‘ฅ) = ๐‘ž(๐‘ฅ)๐‘“(๐‘ฅ) = (1 + ๐‘ 1 ๐‘ฅ + ๐‘ 2 ๐‘ฅ2 + โ‹ฏ + ๐‘ ๐‘‘ ๐‘ฅ๐‘‘ )๐‘“(๐‘ฅ). Since deg ๐‘(๐‘ฅ) < ๐‘‘ we have that [๐‘ฅ๐‘›+๐‘‘ ]๐‘(๐‘ฅ) = 0 for all ๐‘› โ‰ฅ 0. So taking the coefficient of ๐‘ฅ๐‘›+๐‘‘ on both sides of the previous displayed equation gives the recursion (3.18). We now show that (b) and (c) are equivalent. The fact that (b) implies (c) again follows from the algorithm. Specifically, using equations (3.19), (3.20), and (3.21), as well as partial fraction expansion, we have (3.22)

๐‘“(๐‘ฅ) =

๐‘(๐‘ฅ) ๐‘˜ โˆ๐‘–=1 (1

โˆ’ ๐‘Ÿ ๐‘– ๐‘ฅ)๐‘‘๐‘–

๐‘˜

๐‘‘๐‘–

=โˆ‘โˆ‘ ๐‘–=1 ๐‘—=1

๐ด๐‘–,๐‘— (1 โˆ’ ๐‘Ÿ ๐‘– ๐‘ฅ)๐‘—

for certain constants ๐ด๐‘–,๐‘— . But by Theorem 3.4.2 and equation (3.16) we have that [๐‘ฅ๐‘› ]

1 ๐‘— ๐‘›+๐‘—โˆ’1 ๐‘› = (( ))๐‘Ÿ๐‘–๐‘› = ( )๐‘Ÿ๐‘– ๐‘› ๐‘—โˆ’1 (1 โˆ’ ๐‘Ÿ ๐‘– ๐‘ฅ)๐‘—

where (๐‘› + ๐‘— โˆ’ 1)(๐‘› + ๐‘— โˆ’ 2) โ‹ฏ (๐‘› + 1) ๐‘›+๐‘—โˆ’1 )= ๐‘—โˆ’1 (๐‘— โˆ’ 1)! is a polynomial in ๐‘› of degree ๐‘— โˆ’ 1 for any given ๐‘—. Now taking the coefficient of ๐‘ฅ๐‘› on both sides of (3.22) gives (

๐‘˜

๐‘‘

๐‘– ๐‘›+๐‘—โˆ’1 ๐‘Ž๐‘› = โˆ‘ [ โˆ‘ ๐ด๐‘–,๐‘— ( )] ๐‘Ÿ๐‘–๐‘› . ๐‘—โˆ’1 ๐‘–=1 ๐‘—=1

Calling the polynomial inside the brackets ๐‘ ๐‘– (๐‘›), we have derived the desired expansion. The proof that (c) implies (b) essentially reverses the steps of the forward direction. So it is left as an exercise. โ–ก We note that the preceding theorem is not just of theoretical significance but is also very useful computationally. In particular, because of the equivalence of (a) and (c), one can solve a linear, constant coefficient recursion in a more direct manner without having to deal with generating functions. To illustrate, suppose we have a sequence Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

3.8. Chromatic polynomials

99

satisfying (3.18). Then we know that the solution in (c) is in terms of the ๐‘Ÿ ๐‘– which are the reciprocals of the roots of ๐‘ž(๐‘ฅ) as given by (3.20). To simplify things, we consider the polynomial ๐‘Ÿ(๐‘ฅ) = ๐‘ฅ๐‘‘ ๐‘ž(1/๐‘ฅ) = ๐‘ฅ๐‘‘ + ๐‘ 1 ๐‘ฅ๐‘‘โˆ’1 + ๐‘ 2 ๐‘ฅ๐‘‘โˆ’2 + โ‹ฏ + ๐‘ ๐‘‘ . Comparison with (3.21) shows that the ๐‘Ÿ ๐‘– are the roots of ๐‘Ÿ(๐‘ฅ). We now find the ๐‘ ๐‘– (๐‘›) by solving for the coefficients of these polynomials using the initial conditions. To be quite concrete, consider again the example (3.15) with which we began this section. Since ๐‘Ž๐‘› โˆ’4๐‘Ž๐‘›โˆ’1 +4๐‘Ž๐‘›โˆ’2 = 0 for ๐‘› โ‰ฅ 2 we factor ๐‘Ÿ(๐‘ฅ) = ๐‘ฅ2 โˆ’4๐‘ฅ+4 = (๐‘ฅโˆ’2)2 . So ๐‘Ž๐‘› = ๐‘(๐‘›)2๐‘› where deg ๐‘(๐‘›) < 2. It follows that ๐‘(๐‘›) = ๐ด + ๐ต๐‘› for constants ๐ด, ๐ต. Plugging in ๐‘› = 0 we get 1 = ๐‘Ž0 = ๐ด20 = ๐ด. Now letting ๐‘› = 1 gives โˆ’4 = ๐‘Ž1 = (1 + ๐ต)21 or ๐ต = โˆ’3. Thus ๐‘Ž๐‘› is again given as in (3.17). But this solution is clearly simpler than the first one given. This is called the method of undetermined coefficients. Of course, the advantage of using generating functions is that they can be used to solve recursions even when they are not linear and constant coefficient. There is a striking resemblance between the theory we have developed in this section and the method of undetermined coefficients for solving linear differential equations with constant coefficients. This is not an accident and the material in this section may be considered as part of the theory of finite differences which is a discrete analogue of the theory of differential equations. We will have more to say about finite differences when we study Mรถbius inversion in Section 5.5.

3.8. Chromatic polynomials Sometimes generating functions or polynomials appear in unexpected ways. We now illustrate this phenomenon using the chromatic polynomial of a graph. Let ๐บ = (๐‘‰, ๐ธ) be a graph. A (vertex) coloring of ๐บ from a set ๐‘† is a function ๐‘ โˆถ ๐‘‰ โ†’ ๐‘†. We refer to ๐‘† as the color set. Figure 3.2 contains a graph which we will be using as our running example together with two colorings using the set ๐‘† = {white, gray, black}. We say that ๐‘ is proper if, for all edges ๐‘ข๐‘ฃ โˆˆ ๐ธ we have ๐‘(๐‘ข) โ‰  ๐‘(๐‘ฃ). The first coloring in Figure 3.2 is proper while the second is not since the edge ๐‘ฃ๐‘ฅ has both endpoints colored gray. The chromatic number of ๐บ, denoted ๐œ’(๐บ), is the minimum cardinality of a set ๐‘† such that there is a proper coloring ๐‘ โˆถ ๐‘‰ โ†’ ๐‘†. In our example, ๐œ’(๐บ) = 3 because we have displayed a proper coloring with three colors in Figure 3.2 (black, white, and gray), and one cannot use fewer colors because of the triangle ๐‘ข๐‘ฃ๐‘ฅ.

๐‘ข

๐‘ฃ

๐‘ฅ

๐‘ค

๐บ=

Figure 3.2. A graph and two colorings

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

100

3. Counting with Ordinary Generating Functions

The chromatic number is an important invariant in graph theory. But by its definition, it belongs more to extremal combinatorics (which studies structures which minimize or maximize a constraint) than the enumerative side of the subject. Although we will not have much more to say about ๐œ’(๐บ) here, we would be remiss if we did not state one of the most famous mathematical theorems in which it plays a part. Call a graph planar if it can be drawn in the plane without any pair of edges crossing. Theorem 3.8.1 (The Four Color Theorem). If ๐บ is planar, then ๐œ’(๐บ) โ‰ค 4.

โ–ก

Note that this result is in stark contrast to ordinary graphs which can have arbitrarily large chromatic number. Complete graphs, for example, have ๐œ’(๐พ๐‘› ) = ๐‘›. The Four Color Theorem caused quite a stir when it was proved in 1977 by Appel and Haken (with the help of Koch) [1, 2]. For one thing, it had been the Four Color Conjecture for over 100 years. Also their proof was the first to make heavy use of computers to do the calculations for all the various cases and the demonstration could not be completely checked by a human. We now turn to the enumerating graph colorings. Let ๐‘ก โˆˆ โ„•. The chromatic polynomial of ๐บ is defined to be ๐‘ƒ(๐บ; ๐‘ก) = the number of proper colorings ๐‘ โˆถ ๐‘‰ โ†’ [๐‘ก]. This concept was introduced by George Birkhoff [13]. It is not clear at this point why ๐‘ƒ(๐บ; ๐‘ก) should be called a polynomial, but let us compute it for the graph in Figure 3.2. Consider coloring the vertices of ๐บ in the order ๐‘ข, ๐‘ฃ, ๐‘ค, ๐‘ฅ. There are ๐‘ก choices for the color of ๐‘ข. This leaves ๐‘ก โˆ’ 1 possibilities for ๐‘ฃ since it cannot be the same color as ๐‘ข. By the same token, the number of choices for ๐‘ค is ๐‘ก โˆ’ 1. Finally, ๐‘ฅ can be colored in ๐‘ก โˆ’ 2 ways since it cannot have the colors of ๐‘ข or ๐‘ฃ and these are different. So the final count is (3.23)

๐‘ƒ(๐บ) = ๐‘ƒ(๐บ; ๐‘ก) = ๐‘ก(๐‘ก โˆ’ 1)(๐‘ก โˆ’ 1)(๐‘ก โˆ’ 2) = ๐‘ก4 โˆ’ 4๐‘ก3 + 5๐‘ก2 โˆ’ 2๐‘ก.

This is a polynomial in ๐‘ก, the number of colors! Before proving that this is always the case, we have a couple of remarks. First of all, there is a close relationship between ๐‘ƒ(๐บ; ๐‘ก) and ๐œ’(๐บ); namely ๐‘ƒ(๐บ; ๐‘ก) = 0 if 0 โ‰ค ๐‘ก < ๐œ’(๐บ) but ๐‘ƒ(๐บ; ๐œ’(๐บ)) > 0. This follows from the definitions of ๐‘ƒ and ๐œ’ since the latter is the smallest nonnegative integer for which proper colorings of ๐บ exist and the former counts such colorings. Secondly, it is not always possible to compute ๐‘ƒ(๐บ; ๐‘ก) in the manner above and express it as a product of factors ๐‘ก โˆ’ ๐‘˜ for integers ๐‘˜. For example, consider the cycle ๐ถ4 with vertices labeled clockwise as ๐‘ข, ๐‘ฃ, ๐‘ค, ๐‘ฅ. If we try to use this method to compute ๐‘ƒ(๐ถ4 ; ๐‘ก), then everything is fine until we get to coloring ๐‘ฅ, for ๐‘ฅ is adjacent to both ๐‘ข and ๐‘ค. But we cannot be sure whether ๐‘ข and ๐‘ค have the same color or not since they themselves are not adjacent. It turns out that the same ideas can be used both for proving that ๐‘ƒ(๐บ; ๐‘ก) is always a polynomial in ๐‘ก and to rectify the difficulty in computing ๐‘ƒ(๐ถ4 ; ๐‘ก). Consider a graph ๐บ = (๐‘‰, ๐ธ) and an edge ๐‘’ โˆˆ ๐ธ. The graph obtained by deleting ๐‘’ from ๐บ is denoted ๐บ โงต ๐‘’ and has vertices ๐‘‰ and edges ๐ธ โˆ’ {๐‘’}. The middle graph in Figure 3.3 is obtained from our running example by deleting ๐‘’ = ๐‘ฃ๐‘ฅ. The graph obtained by contracting ๐‘’ in ๐บ is Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

3.8. Chromatic polynomials

101

๐‘ข ๐บ=

๐‘ฃ

๐‘ฃ

๐‘ฅ

๐‘ค

๐บโงต๐‘’=

๐‘’ ๐‘ฅ

๐‘ข

๐‘ค ๐‘ข ๐บ/๐‘’ =

๐‘ฃ๐‘’ ๐‘ค

Figure 3.3. Deletion and contraction

denoted ๐บ/๐‘’ and is obtained by shrinking ๐‘’ to a new vertex ๐‘ฃ ๐‘’ , making ๐‘ฃ ๐‘’ adjacent to the vertices which were adjacent to either endpoint of ๐‘’ and leaving all other vertices and edges of ๐บ the same. Contracting ๐‘ฃ๐‘ฅ in our example graph results in the graph on the right in Figure 3.3. The next lemma is crucial in the study of ๐‘ƒ(๐บ; ๐‘ก). It is ideally set up for induction on #๐ธ since both ๐บ โงต ๐‘’ and ๐บ/๐‘’ have fewer edges than ๐บ. Lemma 3.8.2 (Deletion-Contraction Lemma). If ๐บ is a graph, then for any ๐‘’ โˆˆ ๐ธ we have ๐‘ƒ(๐บ; ๐‘ก) = ๐‘ƒ(๐บ โงต ๐‘’; ๐‘ก) โˆ’ ๐‘ƒ(๐บ/๐‘’; ๐‘ก). Proof. We will prove this in the form ๐‘ƒ(๐บ โงต ๐‘’) = ๐‘ƒ(๐บ) + ๐‘ƒ(๐บ/๐‘’). Suppose ๐‘’ = ๐‘ข๐‘ฃ. Since ๐‘’ is no longer present in ๐บ โงต ๐‘’, its proper colorings are of two types: those where ๐‘(๐‘ข) โ‰  ๐‘(๐‘ฃ) and those where ๐‘(๐‘ข) = ๐‘(๐‘ฃ). If ๐‘(๐‘ข) โ‰  ๐‘(๐‘ฃ), then properly coloring ๐บ โงต ๐‘’ is the same as properly coloring ๐บ. So there are ๐‘ƒ(๐บ) colorings of the first type. There is also a bijection between the proper colorings of ๐บ โงต ๐‘’ where ๐‘(๐‘ข) = ๐‘(๐‘ฃ) and those of ๐บ/๐‘’; namely color ๐‘ฃ ๐‘’ with the common color of ๐‘ข and ๐‘ฃ and leave all the other colors the same. It follows that there are ๐‘ƒ(๐บ/๐‘’) colorings of the second type and the lemma is proved. โ–ก We can now easily show that ๐‘ƒ(๐บ; ๐‘ก) lives up to its name. Theorem 3.8.3. We have that ๐‘ƒ(๐บ; ๐‘ก) is a polynomial in ๐‘ก for any graph ๐บ. Proof. We induct on #๐ธ. If ๐บ has no edges, then clearly ๐‘ƒ(๐บ; ๐‘ก) = ๐‘ก#๐‘‰ which is a polynomial in ๐‘ก. If #๐ธ โ‰ฅ 1, then pick ๐‘’ โˆˆ ๐ธ. By deletion-contraction ๐‘ƒ(๐บ; ๐‘ก) = ๐‘ƒ(๐บ โงต ๐‘’; ๐‘ก) โˆ’ ๐‘ƒ(๐บ/๐‘’; ๐‘ก). And by induction we have that both ๐‘ƒ(๐บ โงต ๐‘’; ๐‘ก) and ๐‘ƒ(๐บ/๐‘’; ๐‘ก) are polynomials in ๐‘ก. Thus the same is true of their difference. โ–ก We can also use Lemma 3.8.2 to compute the chromatic polynomial of ๐ถ4 . Recall our notation of ๐‘ƒ๐‘› and ๐พ๐‘› for paths and complete graphs on ๐‘› vertices, respectively. Now picking any edge ๐‘’ โˆˆ ๐ธ(๐ถ4 ) we can use deletion-contraction and then determine the polynomials of the resulting graphs via coloring vertex by vertex to obtain ๐‘ƒ(๐ถ4 ) = ๐‘ƒ(๐‘ƒ4 ) โˆ’ ๐‘ƒ(๐พ3 ) = ๐‘ก(๐‘ก โˆ’ 1)3 โˆ’ ๐‘ก(๐‘ก โˆ’ 1)(๐‘ก โˆ’ 2) = ๐‘ก(๐‘ก โˆ’ 1)(๐‘ก2 โˆ’ 3๐‘ก + 3). Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

102

3. Counting with Ordinary Generating Functions

Note that the quadratic factor has complex roots, thus substantiating our claim that ๐‘ƒ(๐บ) does not always have roots which are integers. One can use induction and Lemma 3.8.2 to prove a host of results about ๐‘ƒ(๐บ; ๐‘ก). Since these demonstrations are all similar, we will leave them to the reader. We will use a nonstandard way of writing down the coefficients of this polynomial, which will turn out to be convenient later. Theorem 3.8.4. Let ๐บ = (๐‘‰, ๐ธ) and write ๐‘ƒ(๐บ; ๐‘ก) = ๐‘Ž0 ๐‘ก๐‘› โˆ’ ๐‘Ž1 ๐‘ก๐‘›โˆ’1 + ๐‘Ž2 ๐‘ก๐‘›โˆ’2 โˆ’ โ‹ฏ + (โˆ’1)๐‘› ๐‘Ž๐‘› . (a) ๐‘› = #๐‘‰. (b) mdeg ๐‘ƒ(๐บ; ๐‘ก) = the number of components of ๐บ. (c) ๐‘Ž๐‘– โ‰ฅ 0 for all ๐‘– and ๐‘Ž๐‘– > 0 for 0 โ‰ค ๐‘– โ‰ค ๐‘› โˆ’ mdeg ๐‘ƒ(๐บ; ๐‘ก). โ–ก

(d) ๐‘Ž0 = 1 and ๐‘Ž1 = #๐ธ.

Now that we know ๐‘ƒ(๐บ; ๐‘ก) is a polynomial we can ask if there is any combinatorial interpretation for its coefficients, the reverse of our approach up to now, which has been to start with a sequence and then find its generating function. Put a total order on the edge set ๐ธ, writing ๐‘’ < ๐‘“ if ๐‘’ is less than ๐‘“ in this order and similarly for other notation. If ๐ถ is a cycle in ๐บ, then the corresponding broken circuit ๐ต is the set of edges obtained from ๐ธ(๐ถ) by removing the smallest edge in the total order. Returning to the graph in Figure 3.2, let ๐‘ = ๐‘ข๐‘ฃ, ๐‘ = ๐‘ข๐‘ฅ, ๐‘‘ = ๐‘ฃ๐‘ค, ๐‘’ = ๐‘ฃ๐‘ฅ and impose the order ๐‘ < ๐‘ < ๐‘‘ < ๐‘’. The only cycle has edges ๐‘, ๐‘, ๐‘’ and the corresponding broken circuit is ๐ต = {๐‘, ๐‘’} which are the edges of a path. Say that a set of edges ๐ด โŠ† ๐ธ contains no broken circuit or is an NBC set if ๐ด 6โŠ‡ ๐ต for any broken circuit ๐ต. Let NBC๐‘˜ = NBC๐‘˜ (๐บ) = {๐ด โŠ† ๐ธ โˆฃ #๐ด = ๐‘˜ and ๐ด is an NBC set} and nbc๐‘˜ = nbc๐‘˜ (๐บ) = # NBC๐‘˜ (๐บ). In our example graph ๐‘˜

NBC๐‘˜ (๐บ)

nbc๐‘˜ (๐บ)

0 1 2 3 4

{โˆ…} { {๐‘}, {๐‘}, {๐‘‘}, {๐‘’} } { {๐‘, ๐‘}, {๐‘, ๐‘‘}, {๐‘, ๐‘’}, {๐‘, ๐‘‘}, {๐‘‘, ๐‘’} } { {๐‘, ๐‘, ๐‘‘}, {๐‘, ๐‘‘, ๐‘’} } โˆ…

1 4 5 2 0

Comparison of the last column of this table with the coefficients of ๐‘ƒ(๐บ; ๐‘ก) presages our next result, which is due to Whitney [99]. It is surprising that the conclusion does not depend on the total order given to the edges. The proof we give is based on the demonstration of Blass and Sagan [16]. Theorem 3.8.5. If #๐‘‰ = ๐‘›, then, given any ordering of ๐ธ, ๐‘›

๐‘ƒ(๐บ; ๐‘ก) = โˆ‘ (โˆ’1)๐‘˜ nbc๐‘˜ (๐บ) ๐‘ก๐‘›โˆ’๐‘˜ . ๐‘˜=0

Proof. Identify each ๐ด โˆˆ NBC๐‘˜ (๐บ) with the associated spanning subgraph. Then ๐ด is acyclic since any cycle contains a broken circuit. It follows from Theorem 1.10.2 that Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

3.8. Chromatic polynomials

103

๐‘ข

๐‘ฃ

๐‘ฅ

๐‘ค

๐บ=

Figure 3.4. Two orientations of a graph

๐ด is a forest with ๐‘› โˆ’ ๐‘˜ component trees. Hence nbc๐‘˜ (๐บ) ๐‘ก๐‘›โˆ’๐‘˜ is the number of pairs (๐ด, ๐‘) where ๐ด โˆˆ NBC๐‘˜ (๐บ) and ๐‘ โˆถ ๐‘‰ โ†’ [๐‘ก] is a coloring constant on each component of ๐ด. We call such a coloring ๐ด-improper. Make the set of such pairs into a signed set by letting sgn(๐ด, ๐‘) = (โˆ’1)#๐ด . So the theorem will be proved if we exhibit a signreversion involution ๐œ„ on these pairs whose fixed points have positive sign and are in bijection with the proper coloring of ๐บ. Define the fixed points of ๐œ„ to be the (๐ด, ๐‘) such that ๐ด = โˆ… and ๐‘ is proper. These pairs clearly have the desired characteristics. For any other pair, ๐‘ is not a proper coloring so there must be an edge ๐‘’ = ๐‘ข๐‘ฃ with ๐‘(๐‘ข) = ๐‘(๐‘ฃ). Let ๐‘’ be the smallest such edge in the total order. Now define ๐œ„(๐ด, ๐‘) = (๐ด ฮ” {๐‘’}, ๐‘) โ‰” (๐ดโ€ฒ , ๐‘). It is clear that ๐œ„ reverses signs. And it is an involution because ๐‘ does not change, and so the smallest monochromatic edge is the same in a pair and its image. We just need to check that ๐œ„ is well-defined. If ๐ดโ€ฒ = ๐ด โˆ’ {๐‘’}, then obviously ๐ด is still an NBC set and ๐‘ is ๐ดโ€ฒ -improper. If ๐ดโ€ฒ = ๐ด โˆช {๐‘’}, then, since ๐‘’ joined two vertices of the same color, ๐‘ is still ๐ดโ€ฒ -improper. But assume, towards a contradiction, that ๐ดโ€ฒ is no longer NBC. Then ๐ดโ€ฒ โŠ‡ ๐ต where ๐ต is a broken circuit, and ๐‘’ โˆˆ ๐ต since ๐ด is NBC. Since ๐‘ is ๐ดโ€ฒ -improper, all edges in ๐ต have vertices colored ๐‘(๐‘ข). But ๐‘’ is the smallest edge having that property, and so the smaller edge removed from a cycle to get ๐ต cannot exist. Thus ๐ดโ€ฒ is NBC, ๐œ„ is well-defined, and we are done with the proof. โ–ก One of the amazing things about the chromatic polynomial is that it often appears in places where a priori it has no business being because no graph coloring is involved. We now give two illustrations of this. Recall from Section 2.6 that an orientation ๐‘‚ of a graph ๐บ is a digraph with the same vertex set obtained by replacing each edge ๐‘ข๐‘ฃ of ๐บ by one of the possible arcs ๐‘ขโƒ—๐‘ฃ or ๐‘ฃโƒ—๐‘ข. See Figure 3.4 for two orientations of our usual graph. Call ๐‘‚ acyclic if it does not contain any directed cycles and let ๐’œ(๐บ) and ๐‘Ž(๐บ) denote the set and number of acyclic orientations of ๐บ, respectively. The first of the orientations just given is acyclic while the second is not. The total number of orientations of the cycle ๐‘ข, ๐‘ฃ, ๐‘ฅ, ๐‘ข is 23 and the number of those which produce a cycle is 2 (clockwise and counterclockwise). Since neither orientation of ๐‘ฃ๐‘ค can produce a cycle, we see that ๐‘Ž(๐บ) = 2(23 โˆ’ 2) = 12. We now do something very strange and plug ๐‘ก = โˆ’1 into the chromatic polynomial (3.23) and get ๐‘ƒ(๐บ; โˆ’1) = (โˆ’1)(โˆ’2)(โˆ’2)(โˆ’3) = 12. Although it is not at all clear what it means to color a graph with โˆ’1 colors, we have just seen an example of the following celebrated theorem of Stanley [85]. Theorem 3.8.6. For any graph ๐บ with #๐‘‰ = ๐‘› we have ๐‘ƒ(๐บ; โˆ’1) = (โˆ’1)๐‘› ๐‘Ž(๐บ). Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

104

3. Counting with Ordinary Generating Functions

Proof. We induct on #๐ธ where the base case is an easy check. It suffices to show that (โˆ’1)๐‘› ๐‘ƒ(๐บ; โˆ’1) and ๐‘Ž(๐บ) satisfy the same recursion. Using the Deletion-Contraction Lemma for the former, we see that we need to show ๐‘Ž(๐บ) = ๐‘Ž(๐บ โงต ๐‘’) + ๐‘Ž(๐บ/๐‘’) for a fixed ๐‘’ = ๐‘ข๐‘ฃ โˆˆ ๐ธ. Consider the map ๐œ™ โˆถ ๐ด(๐บ) โ†’ ๐ด(๐บ โงต ๐‘’) โ€ฒ

โ€ฒ

which sends ๐‘‚ โ†ฆ ๐‘‚ where ๐‘‚ is obtained from ๐‘‚ by removing the arc corresponding to ๐‘’. Clearly ๐‘‚โ€ฒ is still acyclic so the function is well-defined. We claim ๐œ™ is onto. Suppose to the contrary that there is some ๐‘‚โ€ฒ โˆˆ ๐ด(๐บ โงต ๐‘’) such that adding back ๐‘ขโƒ—๐‘ฃ creates a directed cycle ๐ถ, and similarly with ๐‘ฃโƒ—๐‘ข creating a cycle ๐ถ โ€ฒ . Then (๐ถ โˆ’ ๐‘ขโƒ—๐‘ฃ) โˆช (๐ถ โ€ฒ โˆ’ ๐‘ฃโƒ—๐‘ข) is a closed, directed walk which must contain a directed cycle by Exercise 14(b) of Chapter 1. But this third directed cycle is contained in ๐‘‚โ€ฒ , which is the desired contradiction. If ๐‘‚โ€ฒ โˆˆ ๐ด(๐บ โงต ๐‘’), then by definition of the map #๐œ™โˆ’1 (๐‘‚โ€ฒ ) โ‰ค 2. And from the previous paragraph #๐œ™โˆ’1 (๐‘‚โ€ฒ ) โ‰ฅ 1. So ๐‘Ž(๐บ) = ๐‘ฅ + 2๐‘ฆ where ๐‘ฅ = #{๐‘‚โ€ฒ โˆฃ ๐œ™โˆ’1 (๐‘‚โ€ฒ ) = 1} and ๐‘ฆ = #{๐‘‚โ€ฒ โˆฃ ๐œ™โˆ’1 (๐‘‚โ€ฒ ) = 2}. Since ๐‘Ž(๐บ โงต ๐‘’) = ๐‘ฅ + ๐‘ฆ it suffices to show that ๐‘Ž(๐บ/๐‘’) = ๐‘ฆ. We will do this by constructing a bijection ๐œ“ โˆถ {๐‘‚โ€ฒ โˆˆ ๐ด(๐บ โงต ๐‘’) โˆฃ ๐œ™โˆ’1 (๐‘‚โ€ฒ ) = 2} โ†’ ๐ด(๐บ/๐‘’). Let ๐‘Œ be the domain of ๐œ“. If there are a pair of edges ๐‘ค๐‘ข, ๐‘ค๐‘ฃ โˆˆ ๐ธ(๐บ), then any ๐‘‚โ€ฒ โˆˆ ๐‘Œ contains either both ๐‘ค๐‘ข โƒ— and ๐‘ค๐‘ฃ โƒ— or both ๐‘ข๐‘ค โƒ— and ๐‘ฃ๐‘ค. โƒ— This is because in all other cases adding back one of the orientations of ๐‘’ would create an orientation of ๐บ with a cycle, contradicting the fact that ๐œ™โˆ’1 (๐‘‚โ€ฒ ) = 2. So define ๐‘‚โ€ณ = ๐œ“(๐‘‚โ€ฒ ) to be the orientation of ๐บ/๐‘’ which agrees with ๐‘‚โ€ฒ on all arcs not containing the new vertex ๐‘ฃ ๐‘’ , and on any edge of the form ๐‘ค๐‘ฃ ๐‘’ uses the same orientation as either ๐‘ค๐‘ข or ๐‘ค๐‘ฃ. (As just shown, these two orientations are either both towards or both away from ๐‘’). Proving that ๐œ“ is a well-defined bijection is left as an exercise. โ–ก We should mention that Stanley actually proved a stronger result giving a combinatorial interpretation to ๐‘ƒ(๐บ; โˆ’๐‘ก) for all negative integers โˆ’๐‘ก. See Exercise 28(b) for details. So, as we saw with the binomial coefficients in (1.6), we have another instance of combinatorial reciprocity. We will study this phenomenon more generally in the next section. For our second example of the protean nature of the chromatic polynomial, we will need to assume that our graphs ๐บ have vertex set [๐‘›] so that one has a total order on the vertices. Let ๐น be a spanning forest of ๐บ and root each component tree ๐‘‡ of ๐น at its smallest vertex ๐‘Ÿ. Say that ๐น is increasing if the integers on any path starting at ๐‘Ÿ form an increasing sequence for all roots ๐‘Ÿ. In Figure 3.5 the reader will find the usual 1

2

1 ๐น1 =

2

1 ๐น2 =

2

3

4

3

4

3

๐บ= 4

Figure 3.5. A graph and two spanning forests

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

3.8. Chromatic polynomials

105

graph, now labeled with [4], and two spanning forests. We have that ๐น1 is increasing as any singleton node is increasing, and in the nontrivial tree the only maximal path from the root is 1, 2, 4, which is an increasing sequence. On the other hand ๐น2 is not increasing because of the path 1, 4, 2. For a graph ๐บ = (๐‘‰, ๐ธ) we define ISF๐‘š (๐บ) = {๐น โˆฃ ๐น is an increasing spanning forest of ๐บ with ๐‘š edges} and isf๐‘š (๐บ) = # ISF๐‘š (๐บ). If #๐‘‰ = ๐‘›, then consider the corresponding generating polynomial ๐‘›

isf(๐บ) = isf(๐บ; ๐‘ก) = โˆ‘ (โˆ’1)๐‘š isf๐‘š (๐บ)๐‘ก๐‘›โˆ’๐‘š . ๐‘š=0

Let us compute this for our example graph. Any tree with zero or one edge is increasing so that isf0 (๐บ) = 1 and isf1 (๐บ) = #๐ธ = 4. Any of the pairs of edges of ๐บ form an increasing forest except for the pair giving ๐น2 in Figure 3.5. So isf2 (๐บ) = (42) โˆ’ 1 = 5. Similarly one checks that isf3 (๐บ) = 2. And isf4 (๐บ) = 0 since ๐บ itself is not a forest. So isf(๐บ; ๐‘ก) = ๐‘ก4 โˆ’ 4๐‘ก3 + 5๐‘ก2 โˆ’ 2๐‘ก = ๐‘ก(๐‘ก โˆ’ 1)2 (๐‘ก โˆ’ 2) = ๐‘ƒ(๐บ; ๐‘ก). We cannot always have isf(๐บ) = ๐‘ƒ(๐บ) because the former depends on the labeling of the vertices (even though our notation conceals that fact) while the latter does not. So we will put aside deciding when they are equal for now and concentrate on the factorization over โ„ค which we have just seen and which, as we will see, is not a coincidence. In fact, the roots will be the cardinalities of the edge sets defined by (3.24)

๐ธ ๐‘˜ = {๐‘–๐‘˜ โˆˆ ๐ธ โˆฃ ๐‘– < ๐‘˜}

for 1 โ‰ค ๐‘˜ โ‰ค ๐‘›. In our example ๐ธ1 = โˆ… (since there is no vertex smaller than 1), ๐ธ2 = {12}, ๐ธ3 = {23}, and ๐ธ4 = {14, 24}. Lemma 3.8.7. If ๐บ has ๐‘‰ = [๐‘›], then a spanning subgraph ๐น is an increasing forest if and only if it is obtained by picking at most one edge from each ๐ธ ๐‘˜ for ๐‘˜ โˆˆ [๐‘›]. Proof. For the forward direction assume, towards a contradiction, that ๐น contains both ๐‘–๐‘˜ and ๐‘—๐‘˜ with ๐‘–, ๐‘— < ๐‘˜. So if ๐‘Ÿ is the root of the tree containing ๐‘–, ๐‘—, ๐‘˜, then, by the increasing condition, ๐‘– must be the vertex preceding ๐‘˜ on the unique ๐‘Ÿโ€“๐‘˜ path. But the same must be true of ๐‘—, which is a contradiction. For the reverse implication, we must first verify that ๐น is acyclic. But if ๐น contains a cycle ๐ถ, then let ๐‘˜ be its maximum vertex. It follows that there are ๐‘–๐‘˜, ๐‘—๐‘˜ โˆˆ ๐ธ(๐ถ) and, by the maximum requirement, ๐‘–, ๐‘— < ๐‘˜. This contradicts the assumption in this direction. Similarly one can show that if ๐น is not increasing, then one can produce two edges from the same ๐ธ ๐‘˜ and so we are done. โ–ก It is now a simple matter to prove the following result of Hallam and Sagan [41]. The proof given here was obtained by Hallam, Martin, and Sagan [40]. Theorem 3.8.8. If ๐บ has ๐‘‰ = [๐‘›], then ๐‘›

isf(๐บ; ๐‘ก) = โˆ(๐‘ก โˆ’ |๐ธ ๐‘˜ |). ๐‘˜=1

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

106

3. Counting with Ordinary Generating Functions

Proof. The coefficient of ๐‘ก๐‘›โˆ’๐‘š in the product is, up to sign, the sum of all terms of the form |๐ธ๐‘–1 ||๐ธ๐‘–2 | โ‹ฏ |๐ธ๐‘–๐‘š | where the ๐‘–๐‘— are distinct indices. But this product is the number of ways to pick one edge out of each of the sets ๐ธ๐‘–1 , ๐ธ๐‘–2 , . . . , ๐ธ๐‘–๐‘š . So, by the previous lemma, the sum is the number of increasing forests with ๐‘š edges, finishing the proof. โ–ก Returning to the question of when the chromatic and increasing spanning forest polynomials are equal, we need the following definition. Graph ๐บ has a perfect elimination order (peo) if there is a total ordering of ๐‘‰ as ๐‘ฃ 1 , ๐‘ฃ 2 , . . . , ๐‘ฃ ๐‘› such that, for all ๐‘˜, the set of vertices coming before ๐‘ฃ ๐‘˜ in this order and adjacent to ๐‘ฃ ๐‘˜ form the vertices of a clique (complete subgraph of ๐บ). This definition may seem strange at first glance, but it has been useful in various graph-theoretic contexts. Returning to our running example graph we see that the order 1, 2, 3, 4 is a peo since 1 is adjacent to no earlier vertex, 2 and 3 are both adjacent to a single previous vertex which is a ๐พ1 , and 4 is adjacent to 1 and 2 which form an edge also known as a ๐พ2 . We can now prove another result from [40]. Lemma 3.8.9. Let ๐บ have ๐‘‰ = [๐‘›]. Write the edges of ๐บ as ๐‘–๐‘— with ๐‘– < ๐‘— and order them lexicographically. For all ๐‘š โ‰ฅ 0 we have ISF๐‘š (๐บ) โŠ† NBC๐‘š (๐บ). Furthermore, we have equality for all ๐‘š if and only if the natural order on [๐‘›] is a peo. Proof. To prove the inclusion we suppose that ๐น is an increasing spanning forest which contains a broken circuit ๐ต and derive a contradiction. By the lexicographic ordering of the edges, ๐ต must be a path of the form ๐‘ฃ 1 , ๐‘ฃ 2 , . . . , ๐‘ฃ ๐‘™ where ๐‘ฃ 1 = min{๐‘ฃ 1 , . . . , ๐‘ฃ ๐‘™ } and ๐‘ฃ 2 > ๐‘ฃ ๐‘™ . So there must be a smallest index ๐‘– โ‰ฅ 2 such that ๐‘ฃ ๐‘– > ๐‘ฃ ๐‘–+1 . It follows that ๐‘ฃ ๐‘–โˆ’1 , ๐‘ฃ ๐‘–+1 < ๐‘ฃ ๐‘– so that the two corresponding edges of ๐ต contradict Lemma 3.8.7. For the forward direction of the second statement we must show that if ๐‘–, ๐‘— < ๐‘˜ and ๐‘–๐‘˜, ๐‘—๐‘˜ โˆˆ ๐ธ(๐บ), then ๐‘–๐‘— โˆˆ ๐ธ(๐บ). By Lemma 3.8.7 again, {๐‘–๐‘˜, ๐‘—๐‘˜} is not the edge set of an increasing spanning forest. So, by the assumed equality, this set must contain a broken circuit. Since there are only two edges, this set must actually be a broken circuit, and ๐‘–๐‘— โˆˆ ๐ธ(๐บ) must be the edge used to complete the cycle. The reverse implication is left as an exercise. โ–ก From this result we immediately conclude the following. Theorem 3.8.10. Let ๐บ have ๐‘‰ = [๐‘›]. Then isf(๐บ; ๐‘ก) = ๐‘ƒ(๐บ; ๐‘ก) if and only if the natural order on [๐‘›] is a peo. โ–ก

3.9. Combinatorial reciprocity When plugging a negative parameter into a counting function results in a sign times another enumerative function, then this is called combinatorial reciprocity. This concept was introduced and studied by Stanley [86]. We have already seen two examples Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

3.9. Combinatorial reciprocity

107

of this in equation (1.6) and Theorem 3.8.6 (and, more generally, Exercise 28 of this chapter). Here we will make a connection with recurrences and rational generating functions. See the text of Beck and Sanyal [5] for a whole book devoted to this subject. Before stating a general theorem, we return to the example with which we began Section 3.6. This was the sequence defined by ๐‘Ž0 = 2 and ๐‘Ž๐‘› = 3๐‘Ž๐‘›โˆ’1 for ๐‘› โ‰ฅ 1. One can extend the domain of this recursion to all integral ๐‘›, in which case one gets 2 = ๐‘Ž0 = 3๐‘Žโˆ’1 so that ๐‘Žโˆ’1 = 2/3. Then 2/3 = ๐‘Žโˆ’1 = 3๐‘Žโˆ’2 yielding ๐‘Žโˆ’2 = 2/32 , and so forth. An easy induction shows that for ๐‘› โ‰ค 0 we have ๐‘Ž๐‘› = 2 โ‹… 3๐‘› just as for ๐‘› โ‰ฅ 0. We can also compute the generating function for the negatively indexed part of the sequence, where it is convenient to start with ๐‘Žโˆ’1 , which is the geometric series 2๐‘ฅ 2๐‘ฅ/3 2๐‘ฅ๐‘› = = . ๐‘› 3 1 โˆ’ ๐‘ฅ/3 3โˆ’๐‘ฅ ๐‘›โ‰ฅ1

โˆ‘ ๐‘Žโˆ’๐‘› ๐‘ฅ๐‘› = โˆ‘ ๐‘›โ‰ฅ1

Comparing this to ๐‘“(๐‘ฅ) = โˆ‘๐‘›โ‰ฅ0 ๐‘Ž๐‘› ๐‘ฅ๐‘› as found in (3.12) we see that โˆ’๐‘“(1/๐‘ฅ) =

โˆ’2 2๐‘ฅ = = โˆ‘ ๐‘Ž ๐‘ฅ๐‘› . 1 โˆ’ 3/๐‘ฅ 3 โˆ’ ๐‘ฅ ๐‘›โ‰ฅ1 โˆ’๐‘›

The reader should have some qualms about writing ๐‘“(1/๐‘ฅ) since we have gone to great pains to point out that ๐‘ฅ has no inverse in โ„‚[[๐‘ฅ]]. Indeed, if we use the definition that ๐‘“(๐‘ฅ) = โˆ‘๐‘›โ‰ฅ0 ๐‘Ž๐‘› ๐‘ฅ๐‘› , then ๐‘“(1/๐‘ฅ) = โˆ‘๐‘›โ‰ฅ0 ๐‘Ž๐‘› /๐‘ฅ๐‘› , which is not a formal power series! But if ๐‘“(๐‘ฅ) can be expressed as a rational function ๐‘“(๐‘ฅ) = ๐‘(๐‘ฅ)/๐‘ž(๐‘ฅ) where deg ๐‘(๐‘ฅ) โ‰ค deg ๐‘ž(๐‘ฅ) โ‰” ๐‘‘, then we can make sense of this substitution as follows. Since ๐‘ž(๐‘ฅ) has degree ๐‘‘ we have that ๐‘ฅ๐‘‘ ๐‘ž(1/๐‘ฅ) is also a polynomial and is invertible since its constant coefficient is nonzero (Theorem 3.3.1). Furthermore, ๐‘ฅ๐‘‘ ๐‘(1/๐‘ฅ) is also a polynomial since ๐‘‘ โ‰ฅ deg ๐‘(๐‘ฅ). So we can define (3.25)

๐‘“(1/๐‘ฅ) =

๐‘ฅ๐‘‘ ๐‘(1/๐‘ฅ) ๐‘ฅ๐‘‘ ๐‘ž(1/๐‘ฅ)

and stay inside the formal power series ring. With this convention, the following result makes sense. Theorem 3.9.1. Suppose that ๐‘Ž๐‘› is a sequence defined for all ๐‘› โˆˆ โ„ค and satisfying the linear recurrence relation with constant coefficients (3.18) for all such ๐‘›. Letting ๐‘“(๐‘ฅ) = โˆ‘๐‘›โ‰ฅ0 ๐‘Ž๐‘› ๐‘ฅ๐‘› , we have โˆ‘ ๐‘Žโˆ’๐‘› ๐‘ฅ๐‘› = โˆ’๐‘“(1/๐‘ฅ). ๐‘›โ‰ฅ1

Proof. By (3.25), to prove the theorem it suffices to show that ๐‘ฅ๐‘‘ ๐‘ž(1/๐‘ฅ) โˆ‘ ๐‘Žโˆ’๐‘› ๐‘ฅ๐‘› = โˆ’๐‘ฅ๐‘‘ ๐‘(1/๐‘ฅ). ๐‘›โ‰ฅ1

Note that by (3.20) we have ๐‘ฅ๐‘‘ ๐‘ž(1/๐‘ฅ) = ๐‘ฅ๐‘‘ + ๐‘ 1 ๐‘ฅ๐‘‘โˆ’1 + ๐‘ 2 ๐‘ฅ๐‘‘โˆ’2 + โ‹ฏ + ๐‘ ๐‘‘ . Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

108

3. Counting with Ordinary Generating Functions

So if ๐‘š โ‰ฅ 1, then, using (3.18) and the fact that ๐‘ฅ๐‘‘ ๐‘(1/๐‘ฅ) has degree at most ๐‘‘, [๐‘ฅ๐‘š+๐‘‘ ]๐‘ฅ๐‘‘ ๐‘ž(1/๐‘ฅ) โˆ‘ ๐‘Žโˆ’๐‘› ๐‘ฅ๐‘› = ๐‘Žโˆ’๐‘š + ๐‘ 1 ๐‘Žโˆ’๐‘šโˆ’1 + ๐‘ 2 ๐‘Žโˆ’๐‘šโˆ’2 + โ‹ฏ + ๐‘ ๐‘‘ ๐‘Žโˆ’๐‘šโˆ’๐‘‘ ๐‘›โ‰ฅ1

=0 = [๐‘ฅ๐‘š+๐‘‘ ](โˆ’๐‘ฅ๐‘‘ ๐‘(1/๐‘ฅ)). Similarly we can prove that this equality of coefficients continues to hold for โˆ’๐‘‘ โ‰ค ๐‘š โ‰ค 0. This completes the proof. โ–ก To illustrate this theorem, we consider the negative binomial expansion. So if one fixes ๐‘› โ‰ฅ 1, then using Theorem 3.4.2 and equation (3.16) ๐‘“(๐‘ฅ) โ‰”

๐‘› ๐‘›+๐‘˜โˆ’1 ๐‘˜ 1 = โˆ‘ (( ))๐‘ฅ๐‘˜ = โˆ‘ ( )๐‘ฅ . ๐‘› ๐‘˜ ๐‘›โˆ’1 (1 โˆ’ ๐‘ฅ) ๐‘˜โ‰ฅ0 ๐‘˜โ‰ฅ0

Note that since ๐‘› is fixed we are thinking of (๐‘›+๐‘˜โˆ’1 ) as a function of ๐‘˜. Substituting ๐‘›โˆ’1 โˆ’๐‘˜ for ๐‘˜ in the binomial coefficient, we wish to consider the corresponding generating function ๐‘›โˆ’๐‘˜โˆ’1 ๐‘˜ ๐‘”(๐‘ฅ) = โˆ‘ ( )๐‘ฅ . ๐‘›โˆ’1 ๐‘˜โ‰ฅ1 We note that (๐‘›โˆ’๐‘˜โˆ’1 ) = 0 for 1 โ‰ค ๐‘˜ < ๐‘› since then 0 โ‰ค ๐‘› โˆ’ ๐‘˜ โˆ’ 1 < ๐‘› โˆ’ 1. So ๐‘ฅ๐‘› can ๐‘›โˆ’1 be factored out from ๐‘”(๐‘ฅ) and, using (1.6) and the above expression for the negative binomial expansion, ๐‘”(๐‘ฅ) = ๐‘ฅ๐‘› โˆ‘ ( ๐‘˜โ‰ฅ๐‘›

= ๐‘ฅ๐‘› โˆ‘ ( ๐‘—โ‰ฅ0

๐‘› โˆ’ ๐‘˜ โˆ’ 1 ๐‘˜โˆ’๐‘› )๐‘ฅ ๐‘›โˆ’1

โˆ’๐‘— โˆ’ 1 ๐‘— )๐‘ฅ ๐‘›โˆ’1

๐‘—+1 = (โˆ’1)๐‘›โˆ’1 ๐‘ฅ๐‘› โˆ‘ (( ))๐‘ฅ๐‘— ๐‘› โˆ’ 1 ๐‘—โ‰ฅ0 = (โˆ’1)๐‘›โˆ’1 ๐‘ฅ๐‘› โˆ‘ ( ๐‘—โ‰ฅ0

=

๐‘›+๐‘—โˆ’1 ๐‘— )๐‘ฅ ๐‘›โˆ’1

(โˆ’1)๐‘›โˆ’1 ๐‘ฅ๐‘› . (1 โˆ’ ๐‘ฅ)๐‘›

On the other hand, we could apply Theorem 3.9.1 and write ๐‘”(๐‘ฅ) =

โˆ’1 โˆ’๐‘ฅ๐‘› (โˆ’1)๐‘›โˆ’1 ๐‘ฅ๐‘› = = ๐‘› ๐‘› (1 โˆ’ 1/๐‘ฅ) (๐‘ฅ โˆ’ 1) (1 โˆ’ ๐‘ฅ)๐‘›

giving the same result but with less computation. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Exercises

109

Exercises (1) Prove that for ๐‘› โˆˆ โ„• ๐‘›

๐‘› โˆ‘ 2๐‘˜ ( ) = 3๐‘› ๐‘˜ ๐‘˜=0 in two ways. (a) Use the Binomial Theorem. (b) Use a combinatorial argument. (c) Generalize this identity by replacing 2๐‘˜ by ๐‘๐‘˜ for any ๐‘ โˆˆ โ„•, giving both a proof using the Binomial Theorem and one which is combinatorial. (2) For ๐‘š, ๐‘›, ๐‘˜ โˆˆ โ„• show that (

๐‘š+๐‘› ๐‘š ๐‘› ) = โˆ‘ ( )( ) ๐‘˜ โˆ’๐‘™ ๐‘˜ ๐‘™ ๐‘™โ‰ฅ0

in three ways: by induction, using the Binomial Theorem, and using a combinatorial argument. (3) Let ๐‘ฅ1 , . . . , ๐‘ฅ๐‘š be variables. Prove the multinomial coefficient identity โˆ‘ ๐‘›1 +โ‹ฏ+๐‘›๐‘š

๐‘› ๐‘› ๐‘› ( )๐‘ฅ1 1 โ‹ฏ ๐‘ฅ๐‘š๐‘š = (๐‘ฅ1 + โ‹ฏ + ๐‘ฅ๐‘š )๐‘› , ๐‘› , . . . , ๐‘› 1 ๐‘š =๐‘›

in three ways: (a) by inducting on ๐‘›, (b) by using the Binomial Theorem and inducting on ๐‘š, (c) by a combinatorial argument. (4) (a) Prove Theorem 3.1.2. (b) Use this generating function to rederive Corollary 1.5.3. (5) (a) Recall that an inversion of ๐œ‹ โˆˆ ๐‘ƒ([๐‘›]) is a pair (๐‘–, ๐‘—) with ๐‘– < ๐‘— and ๐œ‹๐‘– > ๐œ‹๐‘— and we call ๐œ‹๐‘– the maximum of the inversion. The inversion table of ๐œ‹ is ๐ผ(๐œ‹) = (๐‘Ž1 , ๐‘Ž2 , . . . , ๐‘Ž๐‘› ) where ๐‘Ž๐‘˜ is the number of inversions with maximum ๐‘˜. Show that 0 โ‰ค ๐‘Ž๐‘˜ < ๐‘˜ for all ๐‘˜ and that ๐‘›

inv ๐œ‹ = โˆ‘ ๐‘Ž๐‘˜ . ๐‘˜=1

(b) Let โ„๐‘› = {(๐‘Ž1 , ๐‘Ž2 , . . . , ๐‘Ž๐‘› ) โˆฃ 0 โ‰ค ๐‘Ž๐‘˜ < ๐‘˜ for all ๐‘˜}. Show that the map ๐œ‹ โ†ฆ ๐ผ(๐œ‹) is a bijection ๐‘ƒ([๐‘›]) โ†’ โ„๐‘› . (c) Use part (b) and weight-generating functions to rederive Theorem 3.2.1. (6) Let st โˆถ ๐”–๐‘› โ†’ โ„• be a statistic on permutations. Recall that for a permutation ๐œ‹ we let Av๐‘› (๐œ‹) denote the set of permutations in ๐”–๐‘› avoiding ๐œ‹. Say that permutations st

๐œ‹, ๐œŽ are st-Wilf equivalent, written ๐œ‹ โ‰ก ๐œŽ, if for all ๐‘› โ‰ฅ 0 we have equality of the Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

110

3. Counting with Ordinary Generating Functions

generating functions ๐‘žst ๐œ =

โˆ‘ ๐œโˆˆAv๐‘› (๐œ‹)

โˆ‘

๐‘žst ๐œ .

๐œโˆˆAv๐‘› (๐œ)

(a) Show that if ๐œ‹ and ๐œŽ are st-Wilf equivalent, then they are Wilf equivalent. (b) Show that inv

132 โ‰ก 213 and inv

231 โ‰ก 312 and that there are no other inv-Wilf equivalences between two permutations in ๐”–3 . (c) Show that maj

132 โ‰ก 231 and maj

213 โ‰ก 312 and that there are no other maj-Wilf equivalences between two permutations in ๐”–3 . (7) (a) Prove that [

๐‘› ๐‘› ]=[ ] ๐‘˜ ๐‘›โˆ’๐‘˜

in three ways: using the ๐‘ž-factorial definition, using the interpretation in terms of integer partitions, and using the interpretation in terms of subspaces. (b) Prove the second recursion in Theorem 3.2.3 in two ways: by mimicking the proof of the first recursion and by using the first recursion in combination with part (a). (c) If ๐‘† โŠ† [๐‘›], then let ฮฃ๐‘† be the sum of the elements of ๐‘†. Give two proofs of the following ๐‘ž-analogue of the fact that #([๐‘›] ) = (๐‘›๐‘˜): ๐‘˜ โˆ‘ ๐‘žฮฃ๐‘† = ๐‘ž( ๐‘†โˆˆ([๐‘›] ) ๐‘˜

๐‘˜+1 2

)[ ๐‘› ] , ๐‘˜ ๐‘ž

one proof by induction and the other using Theorem 3.2.5. (d) Give two other proofs of Theorem 3.2.6: one by inducting on ๐‘› and one by using Theorem 3.2.5. (8) (a) Reprove Theorem 3.2.4 in two way: using integer partitions and using subspaces. (b) The Negative ๐‘ž-Binomial Theorem states that 1 ๐‘›+๐‘˜โˆ’1 ๐‘˜ = โˆ‘[ ]๐‘ก . ๐‘˜ (1 โˆ’ ๐‘ก)(1 โˆ’ ๐‘ž๐‘ก)(1 โˆ’ ๐‘ž2 ๐‘ก) . . . (1 โˆ’ ๐‘ž๐‘›โˆ’1 ๐‘ก) ๐‘˜โ‰ฅ0 Give three proofs of this result: inductive, using integer partitions, and using subspaces. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Exercises

111

(9) (a) Given ๐‘›1 +๐‘›2 +โ‹ฏ+๐‘›๐‘š = ๐‘›, define the corresponding ๐‘ž-multinomial coefficient to be [๐‘›]๐‘ž ! ๐‘› [ ] = ๐‘›1 , ๐‘›2 , . . . , ๐‘›๐‘š ๐‘ž [๐‘›1 ]๐‘ž ! [๐‘›2 ]๐‘ž ! . . . [๐‘›๐‘š ]๐‘ž ! if all ๐‘›๐‘– โ‰ฅ 0 or zero otherwise. Prove that [

๐‘› ] ๐‘›1 , ๐‘› 2 , . . . , ๐‘› ๐‘š ๐‘ž ๐‘š

= โˆ‘ ๐‘ž๐‘›1 +๐‘›2 +โ‹ฏ+๐‘›๐‘–โˆ’1 [ ๐‘–=1

๐‘›โˆ’1 ] . ๐‘›1 , . . . , ๐‘›๐‘–โˆ’1 , ๐‘›๐‘– โˆ’ 1, ๐‘›๐‘–+1 , . . . , ๐‘›๐‘š ๐‘ž

(b) Define inversions, descents, and the major index for permutations (linear orderings) of the multiset ๐‘€ = {{1๐‘›1 , 2๐‘›2 , . . . , ๐‘š๐‘›๐‘š }} exactly the way as was done for permutations without repetition. Let ๐‘ƒ(๐‘€) be the set of permutations of ๐‘€. Prove that โˆ‘

๐‘žinv ๐œ‹ =

๐œ‹โˆˆ๐‘ƒ(๐‘€)

โˆ‘

๐‘žmaj ๐œ‹ = [

๐œ‹โˆˆ๐‘ƒ(๐‘€)

๐‘› ] . ๐‘›1 , ๐‘› 2 , . . . , ๐‘› ๐‘š ๐‘ž

(c) Let ๐‘‰ be a vector space over ๐”ฝ๐‘ž of dimension ๐‘›. Assume ๐‘† = {๐‘ 1 < โ‹ฏ < ๐‘ ๐‘š } โŠ† {0, 1, . . . , ๐‘›}. Then a flag of type ๐‘† is a chain of subspaces ๐น โˆถ ๐‘Š1 < ๐‘Š2 < โ‹ฏ < ๐‘Š๐‘š โ‰ค ๐‘‰ such that dim ๐‘Š ๐‘– = ๐‘ ๐‘– for all ๐‘–. The reason for this terminology is that when ๐‘› = 2 and ๐‘† = {0, 1, 2}, then ๐น consists of a point (the origin) contained in a line contained in a plane which could be viewed as a drawing of a physical flag with the point being the hole in the ground, the line being the flag pole, and the plane being the cloth flag itself. Give two proofs that #{๐น of type ๐‘†} = [

๐‘› ] , ๐‘ 1 , ๐‘ 2 โˆ’ ๐‘ 1 , ๐‘ 3 โˆ’ ๐‘ 2 , . . . , ๐‘ ๐‘š โˆ’ ๐‘ ๐‘šโˆ’1 , ๐‘› โˆ’ ๐‘ ๐‘š ๐‘ž

one by mimicking the proof of Theorem 3.2.6 and one by induction on ๐‘š. (10) (a) Prove that in โ„‚[[๐‘ฅ]] we have ๐‘’๐‘˜๐‘ฅ = (๐‘’๐‘ฅ )๐‘˜ for any ๐‘˜ โˆˆ โ„•. (b) Define formal power series for the trigonometric functions using their usual Taylor expansions. Prove that in โ„‚[[๐‘ฅ]] we have sin2 ๐‘ฅ + cos2 ๐‘ฅ = 1 and sin 2๐‘ฅ = 2 sin ๐‘ฅ cos ๐‘ฅ. (c) If one is given a sequence ๐‘Ž0 , ๐‘Ž1 , ๐‘Ž2 , . . . and defines ๐‘“๐‘˜ (๐‘ฅ) = ๐‘Ž๐‘˜ ๐‘ฅ๐‘˜ , then show that โˆ‘๐‘˜โ‰ฅ0 ๐‘“๐‘˜ (๐‘ฅ) = ๐‘“(๐‘ฅ) where ๐‘“(๐‘ฅ) = โˆ‘๐‘˜โ‰ฅ0 ๐‘Ž๐‘˜ ๐‘ฅ๐‘˜ . (d) Prove the backwards direction of Theorem 3.3.2. (e) Use Theorems 3.3.1 and 3.3.3 to reprove that 1/๐‘ฅ and ๐‘’1+๐‘ฅ are not well-defined in โ„‚[[๐‘ฅ]]. (f) Prove Theorem 3.3.4 (11) Prove that if ๐‘†, ๐‘‡ are summable sets, then so is ๐‘† ร— ๐‘‡. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

112

3. Counting with Ordinary Generating Functions

(12) Say that ๐‘“(๐‘ฅ) โˆˆ โ„‚[[๐‘ฅ]] has a square root if there is ๐‘”(๐‘ฅ) โˆˆ โ„‚[[๐‘ฅ]] such that ๐‘“(๐‘ฅ) = ๐‘”(๐‘ฅ)2 . (a) Prove that ๐‘“(๐‘ฅ) has a square root if and only if mdeg ๐‘“(๐‘ฅ) is even. (b) Show that as formal power series (1 + ๐‘ฅ)1/2 = โˆ‘ ( ๐‘˜โ‰ฅ0

1/2 ๐‘˜ )๐‘ฅ . ๐‘˜

(c) Show that as formal power series (๐‘’๐‘ฅ )

1/2

๐‘ฅ๐‘˜ . 2๐‘˜ ๐‘˜! ๐‘˜โ‰ฅ0

=โˆ‘

(d) Generalize the previous parts of this exercise to ๐‘šth roots for ๐‘š โˆˆ โ„™. (13) Give a second proof of Theorem 3.4.2 by using induction. (14) Prove Theorem 3.4.3. (15) Prove Theorem 3.5.1. (16) (a) Finish the proof of Theorem 3.5.3(a). (b) Give a second proof of Theorem 3.5.3(b) by using Theorem 3.2.5. (c) Show that the generating function for the number of partitions of ๐‘› with largest part ๐‘˜ equals the generating function for the number of partitions of ๐‘› with exactly ๐‘˜ parts and that both are equal to the product ๐‘ฅ๐‘˜ . (1 โˆ’ ๐‘ฅ)(1 โˆ’ ๐‘ฅ2 ) โ‹ฏ (1 โˆ’ ๐‘ฅ๐‘˜ ) (d) The Durfee square of a Young diagram ๐œ† is the largest square partition (๐‘‘ ๐‘‘ ) such that (๐‘‘ ๐‘‘ ) โŠ† ๐œ†. Use this concept to prove that 2

๐‘ฅ๐‘‘ . (1 โˆ’ ๐‘ฅ)2 (1 โˆ’ ๐‘ฅ2 )2 โ‹ฏ (1 โˆ’ ๐‘ฅ๐‘‘ )2 ๐‘‘โ‰ฅ0

โˆ‘ ๐‘(๐‘›)๐‘ฅ๐‘› = โˆ‘ ๐‘›โ‰ฅ0

(17) Let ๐‘Ž๐‘› be the number of integer partitions of ๐‘› such that any part ๐‘– is repeated fewer than ๐‘– times and let ๐‘๐‘› be the number of integer partitions of ๐‘› such that no part is a square. Use generating functions to show that ๐‘Ž๐‘› = ๐‘๐‘› for all ๐‘›. (18) Given ๐‘š โ‰ฅ 2, use generating functions to show that the number of partitions of ๐‘› where each part is repeated fewer than ๐‘š times equals the number of partitions of ๐‘› into parts not divisible by ๐‘š. Note that bijective proofs of this result were given in Exercise 15 of Chapter 2. (19) (a) Show that ๐น๐‘› is the closest integer to ๐‘›

1 + โˆš5 ( ) 2 โˆš5 1

for ๐‘› โ‰ฅ 1. (b) Prove that ๐น๐‘›2 = {

๐น๐‘›โˆ’1 ๐น๐‘›+1 + 1 ๐น๐‘›โˆ’1 ๐น๐‘›+1 โˆ’ 1

if ๐‘› is odd, if ๐‘› is even

in two ways: using equation (3.14) and by a combinatorial argument. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Exercises

(20) (a) (b) (c) (d) (e)

113

Use the algorithm in Section 3.6 to rederive Theorem 3.1.1. Complete the proof of Theorem 3.6.1. Give a second proof of Theorem 3.6.1 using Theorem 3.1.2 Prove Theorem 3.6.2. Let ๐‘  be the infinite matrix with rows and columns indexed by โ„• and with ๐‘ (๐‘›, ๐‘˜) being the entry in row ๐‘› and column ๐‘˜. Similarly define ๐‘† with entries ๐‘†(๐‘›, ๐‘˜). Show that ๐‘†๐‘  = ๐‘ ๐‘† = ๐ผ where ๐ผ is the โ„• ร— โ„• identity matrix. Hint: Use Theorems 3.6.1 and 3.6.2.

(21) Given ๐‘› โˆˆ โ„• and ๐‘˜ โˆˆ โ„ค, define the corresponding ๐‘ž-Stirling number of the second kind by ๐‘†[0, ๐‘˜] = ๐›ฟ0,๐‘˜ and, for ๐‘› โ‰ฅ 1, ๐‘†[๐‘›, ๐‘˜] = ๐‘†[๐‘› โˆ’ 1, ๐‘˜ โˆ’ 1] + [๐‘˜]๐‘ž ๐‘†[๐‘› โˆ’ 1, ๐‘˜]. (a) Show that ๐‘ฅ๐‘˜ . (1 โˆ’ [1]๐‘ž ๐‘ฅ)(1 โˆ’ [2]๐‘ž ๐‘ฅ) โ‹ฏ (1 โˆ’ [๐‘˜]๐‘ž ๐‘ฅ)

โˆ‘ ๐‘†[๐‘›, ๐‘˜]๐‘ฅ๐‘› = ๐‘›โ‰ฅ0

(b) All set partitions ๐œŒ = ๐ต1 /๐ต2 / . . . /๐ต๐‘˜ in the rest of this problem will be written in standard form, which means that 1 = min ๐ต1 < min ๐ต2 < โ‹ฏ < min ๐ต๐‘˜ . An inversion of ๐œŒ is a pair (๐‘, ๐ต๐‘— ) where ๐‘ โˆˆ ๐ต๐‘– for some ๐‘– < ๐‘— and ๐‘ > min ๐ต๐‘— . We let inv ๐œŒ be the number of inversions of ๐œŒ. For example, ๐œŒ = ๐ต1 /๐ต2 /๐ต3 = 136/25/4 has inversions (3, ๐ต2 ), (6, ๐ต2 ), (6, ๐ต3 ), and (5, ๐ต3 ) so that inv ๐œŒ = 4. Show that ๐‘†[๐‘›, ๐‘˜] = โˆ‘ ๐‘žinv ๐œŒ . ๐œŒโˆˆ๐‘†([๐‘›],๐‘˜)

(c) A descent of a set partition ๐œŒ is a pair (๐‘, ๐ต๐‘–+1 ) where ๐‘ โˆˆ ๐ต๐‘– and ๐‘ > min ๐ต๐‘— . We let des ๐œŒ be the number of descents of ๐œŒ. In the previous example, ๐œŒ has descents (3, ๐ต2 ), (6, ๐ต2 ), and (5, ๐ต3 ) so that des ๐œŒ = 3. The descent multiset of ๐œŒ is denoted Des ๐œŒ and is the multiset {{1๐‘‘1 , 2๐‘‘2 , . . . , ๐‘˜๐‘‘๐‘˜ โˆฃ for all ๐‘–, ๐‘‘๐‘– = number of descents (๐‘, ๐ต๐‘–+1 )}}. The major index of ๐œŒ is maj ๐œŒ =

โˆ‘ ๐‘– = ๐‘‘1 + 2๐‘‘2 + โ‹ฏ + ๐‘˜๐‘‘๐‘˜ . ๐‘–โˆˆDes ๐œŒ

In our running example Des ๐œŒ = {{12 , 21 }} so that maj ๐œŒ = 1 + 1 + 2 = 4. Show that ๐‘†[๐‘›, ๐‘˜] = โˆ‘ ๐‘žmaj ๐œŒ . ๐œŒโˆˆ๐‘†([๐‘›],๐‘˜)

(22) Given ๐‘› โˆˆ โ„• and ๐‘˜ โˆˆ โ„ค, define the corresponding signless ๐‘ž-Stirling number of the first kind by ๐‘[0, ๐‘˜] = ๐›ฟ0,๐‘˜ and, for ๐‘› โ‰ฅ 1, ๐‘[๐‘›, ๐‘˜] = ๐‘[๐‘› โˆ’ 1, ๐‘˜ โˆ’ 1] + [๐‘› โˆ’ 1]๐‘ž ๐‘[๐‘› โˆ’ 1, ๐‘˜] (a) Show that โˆ‘ ๐‘[๐‘›, ๐‘˜]๐‘ฅ๐‘˜ = ๐‘ฅ(๐‘ฅ + [1]๐‘ž )(๐‘ฅ + [2]๐‘ž ) โ‹ฏ (๐‘ฅ + [๐‘› โˆ’ 1]๐‘ž ). ๐‘˜โ‰ฅ0

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

114

3. Counting with Ordinary Generating Functions

(b) The standard form of ๐œ‹ โˆˆ ๐”–๐‘› is ๐œ‹ = ๐œ…1 ๐œ…2 โ‹ฏ ๐œ…๐‘˜ where the ๐œ…๐‘– are the cycles of ๐œ‹, min ๐œ…1 < min ๐œ…2 < โ‹ฏ < min ๐œ…๐‘˜ , and each ๐œ…๐‘– is written beginning with min ๐œ…๐‘– . Define the cycle major index of ๐œ‹ to be maj๐‘ ๐œ‹ = maj ๐œ‹โ€ฒ where ๐œ‹โ€ฒ is the permutation in one-line form obtained by removing the cycle parentheses in the standard form of ๐œ‹. If, for example, ๐œ‹ = (1, 7, 2)(3, 6, 8)(4, 5), then ๐œ‹โ€ฒ = 17236845 so that maj๐‘ ๐œ‹ = 2 + 6 = 8. Show that ๐‘[๐‘›, ๐‘˜] = โˆ‘ ๐‘žmaj๐‘ ๐œ‹ . ๐œ‹โˆˆ๐‘([๐‘›],๐‘˜)

(23) Reprove the formula ๐ถ(๐‘›) =

2๐‘› 1 ( ) ๐‘›+1 ๐‘›

by using Theorem 3.6.3. (24) (a) Show that if ๐‘˜, ๐‘™ โˆˆ โ„• are constants, then (๐‘›+๐‘™ ) is a polynomial in ๐‘› of degree ๐‘˜. ๐‘˜ (b) Show that the polynomials ๐‘› ๐‘›+1 ๐‘›+2 ( ), ( ), ( ), . . . 0 1 2 form a basis for the algebra of polynomials in ๐‘›. (c) Use part (a) to complete the proof of Theorem 3.7.1. (25) Redo the solution for the first recursion in Section 3.6 as well as the one for ๐น๐‘› using the method of undetermined coefficients. (26) Prove for the ๐‘›-cycle that ๐‘ƒ(๐ถ๐‘› ; ๐‘ก) = (๐‘ก โˆ’ 1)๐‘› + (โˆ’1)๐‘› (๐‘ก โˆ’ 1) in two ways: using deletion-contraction and using NBC sets. (27) Prove Theorem 3.8.4 using induction and give a second proof of parts (b)โ€“(d) using NBC sets. (28) (a) Complete the proof of Theorem 3.8.6. (b) Let ๐บ = (๐‘‰, ๐ธ) be a graph and ๐‘ก โˆˆ โ„™. Call an acyclic orientation ๐‘‚ and a (not necessarily proper) coloring ๐‘ โˆถ ๐‘‰ โ†’ [๐‘ก] compatible if, for all arcs ๐‘ขโƒ—๐‘ฃ of ๐‘‚, we have ๐‘(๐‘ข) โ‰ค ๐‘(๐‘ฃ). Show that if #๐‘‰ = ๐‘›, then ๐‘ƒ(๐บ; โˆ’๐‘ก) = (โˆ’1)๐‘› (number of compatible pairs (๐‘‚, ๐‘)). (c) Show that Theorem 3.8.6 is a special case of part (b). (29) Finish the proofs of Lemma 3.8.7 and Lemma 3.8.9. (30) (a) Call a permutation ๐œŽ which avoids ฮ  = {231, 312, 321} tight. Show that ๐œŽ is tight if and only if ๐œŽ is an involution having only 2-cycles of the form (๐‘–, ๐‘– + 1) for some ๐‘–. (b) Let ๐บ be a graph with ๐‘‰ = [๐‘›]. Call a spanning forest ๐น of ๐บ tight if the sequence of labels on any path starting at a root of ๐น avoids ฮ  as in part (a). Let TSF๐‘š (๐บ) = {๐น โˆฃ ๐น is a tight spanning forest of ๐บ with ๐‘š edges}. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Exercises

115

Show that if ๐บ has no 3-cycles, then for all ๐‘š โ‰ฅ 0 TSF๐‘š (๐บ) โŠ† NBC๐‘š (๐บ). (c) A candidate path in a graph ๐บ with no 3-cycles is a path of the form ๐‘Ž, ๐‘, ๐‘, ๐‘ฃ 1 , ๐‘ฃ 2 , . . . , ๐‘ฃ ๐‘š = ๐‘‘ such that ๐‘Ž < ๐‘ < ๐‘, ๐‘š โ‰ฅ 1, and ๐‘ฃ ๐‘š is the only ๐‘ฃ ๐‘– smaller than ๐‘. A total order on ๐‘‰(๐บ) is called a quasiperfect ordering (qpo) if every candidate path satisfies the following condition: either ๐‘Ž๐‘‘ โˆˆ ๐ธ(๐บ), or ๐‘‘ < ๐‘ and ๐‘๐‘‘ โˆˆ ๐ธ(๐บ). Consider the generating function tsf(๐บ; ๐‘ก) = โˆ‘ (โˆ’1)๐‘š tsf๐‘š (๐บ)๐‘ก๐‘›โˆ’๐‘š . ๐‘šโ‰ฅ0

Show that tsf(๐บ; ๐‘ก) = ๐‘ƒ(๐บ; ๐‘ก) if and only if the natural order on [๐‘›] is a qpo. (31) Fill in the details of the case โˆ’๐‘‘ โ‰ค ๐‘š โ‰ค 0 in the proof of Theorem 3.9.1. (32) (a) Extend the Fibonacci numbers ๐น๐‘› to all ๐‘› โˆˆ โ„ค by insisting that their recursion continue to hold for ๐‘› < 0. Show that if ๐‘› โ‰ฅ 0, then ๐นโˆ’๐‘› = (โˆ’1)๐‘›โˆ’1 ๐น๐‘› . (b) Find โˆ‘๐‘›โ‰ฅ1 ๐นโˆ’๐‘› ๐‘ฅ๐‘› in two ways: by using part (a) and by using Theorem 3.9.1.

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

10.1090/gsm/210/04

Chapter 4

Counting with Exponential Generating Functions

Given a sequence ๐‘Ž0 , ๐‘Ž1 , ๐‘Ž๐‘› , . . . of complex numbers, one can associate with it an exponential generating function where ๐‘Ž๐‘› is the coefficient of ๐‘ฅ๐‘› /๐‘›!. In certain cases it turns out that the exponential generating function is easier to deal with than the ordinary one. This is particularly true if the ๐‘Ž๐‘› count combinatorial objects obtained from some set of labels. We give a method for dealing with such structures which again give rise to Sum and Product Rules as well as an Exponential Formula unique to this setting.

4.1. First examples Given ๐‘Ž0 , ๐‘Ž1 , ๐‘Ž๐‘› , . . . where ๐‘Ž๐‘› โˆˆ โ„‚ for all ๐‘›, the corresponding exponential generating function (egf) is ๐‘ฅ ๐‘ฅ2 ๐‘ฅ๐‘› ๐น(๐‘ฅ) = ๐‘Ž0 + ๐‘Ž1 + ๐‘Ž2 + โ‹ฏ = โˆ‘ ๐‘Ž๐‘› . 1! 2! ๐‘›! ๐‘›โ‰ฅ0 In order to distinguish these from the ordinary generating functions (ogfs) in the previous chapter, we will often use capital letters for egfs and lowercase ones for ogfs. The use of the adjective โ€œexponentialโ€ is because in the simple case when ๐‘Ž๐‘› = 1 for all ๐‘›, the corresponding egf is ๐น(๐‘ฅ) = โˆ‘๐‘›โ‰ฅ0 ๐‘ฅ๐‘› /๐‘›! = ๐‘’๐‘ฅ . To illustrate why egfs may be useful in studying a sequence, consider ๐‘Ž๐‘› = ๐‘›! for ๐‘› โ‰ฅ 0. The ogf is ๐‘“(๐‘ฅ) = โˆ‘๐‘›โ‰ฅ0 ๐‘›! ๐‘ฅ๐‘› and this power series cannot be simplified. On the other hand, the egf is ๐‘ฅ๐‘› 1 ๐น(๐‘ฅ) = โˆ‘ ๐‘›! = ๐‘›! 1 โˆ’ ๐‘ฅ ๐‘›โ‰ฅ0 which can now be manipulated if necessary. To get some practice in using egfs, we will now compute some examples. One technique which occurs often in determining generating functions is the interchange of 117 Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

118

4. Counting with Exponential Generating Functions

summations. As an example, consider the derangement numbers ๐ท(๐‘›) and the formula which was given for them in Theorem 2.1.2. So ๐‘›

โˆ‘ ๐ท(๐‘›) ๐‘›โ‰ฅ0

๐‘ฅ๐‘› 1 ๐‘ฅ๐‘› = โˆ‘ ๐‘›! ( โˆ‘ (โˆ’1)๐‘˜ ) ๐‘›! ๐‘˜! ๐‘›! ๐‘›โ‰ฅ0 ๐‘˜=0 (โˆ’1)๐‘˜ โˆ‘ ๐‘ฅ๐‘› ๐‘˜! ๐‘˜โ‰ฅ0 ๐‘›โ‰ฅ๐‘˜

=โˆ‘

(โˆ’1)๐‘˜ ๐‘ฅ๐‘˜ ๐‘˜! 1 โˆ’ ๐‘ฅ ๐‘˜โ‰ฅ0

=โˆ‘

=

1 (โˆ’๐‘ฅ)๐‘˜ โˆ‘ 1 โˆ’ ๐‘ฅ ๐‘˜โ‰ฅ0 ๐‘˜!

=

๐‘’โˆ’๐‘ฅ . 1โˆ’๐‘ฅ

We will be able to give a much more combinatorial derivation of this formula once we have introduced the theory of labeled structures in Section 4.3. For now, we just record the result for future reference. Theorem 4.1.1. We have โˆ‘ ๐ท(๐‘›) ๐‘›โ‰ฅ0

๐‘ฅ๐‘› ๐‘’โˆ’๐‘ฅ = . ๐‘›! 1โˆ’๐‘ฅ

โ–ก

If the given sequence is defined by a recurrence relation, then one can use a slight modification of the algorithm in Section 3.6 to compute its egf. One just multiplies by ๐‘ฅ๐‘› /๐‘›!, rather than ๐‘ฅ๐‘› , and sums. Note that the largest index in the recursion may not be the best choice to use as the power on ๐‘ฅ because of the following considerations. Given ๐‘“(๐‘ฅ) = โˆ‘๐‘›โ‰ฅ0 ๐‘Ž๐‘› ๐‘ฅ๐‘› , then its formal derivative is the formal power series ๐‘“โ€ฒ (๐‘ฅ) = โˆ‘ ๐‘›๐‘Ž๐‘› ๐‘ฅ๐‘›โˆ’1 . ๐‘›โ‰ฅ0

This derivative enjoys most of the usual properties of the ordinary analytic derivative such as linearity and the product rule. One similarly defines formal integrals. Note that if one starts with an egf ๐น(๐‘ฅ) = โˆ‘๐‘›โ‰ฅ0 ๐‘Ž๐‘› ๐‘ฅ๐‘› /๐‘›!, then (4.1)

๐น โ€ฒ (๐‘ฅ) = โˆ‘ ๐‘›๐‘Ž๐‘› ๐‘›โ‰ฅ0

๐‘ฅ๐‘›โˆ’1 ๐‘ฅ๐‘›โˆ’1 ๐‘ฅ๐‘› = โˆ‘ ๐‘Ž๐‘›+1 = โˆ‘ ๐‘Ž๐‘› ๐‘›! ๐‘›! (๐‘› โˆ’ 1)! ๐‘›โ‰ฅ0 ๐‘›โ‰ฅ1

which is just the egf for the same sequence shifted up by one. So it can simplify things if the subscript on an element of the sequence is greater than the exponent of the corresponding power of ๐‘ฅ. Let us consider the Bell numbers and their recurrence relation given in Theorem 1.4.1. Let ๐ต(๐‘ฅ) = โˆ‘๐‘›โ‰ฅ0 ๐ต๐‘› ๐‘ฅ๐‘› /๐‘›!. It will be convenient to replace ๐‘› by ๐‘› + 1 and ๐‘˜ by ๐‘˜ + 1 in the recursion before multiplying by ๐‘ฅ๐‘› /๐‘›! and summing. So using (4.1) and Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

4.1. First examples

119

the summation interchange trick we obtain ๐ต โ€ฒ (๐‘ฅ) = โˆ‘ ๐ต(๐‘› + 1) ๐‘›โ‰ฅ0

๐‘ฅ๐‘› ๐‘›!

๐‘›

๐‘› ๐‘ฅ๐‘› = โˆ‘ ( โˆ‘ ( )๐ต(๐‘› โˆ’ ๐‘˜)) ๐‘›! ๐‘˜ ๐‘›โ‰ฅ0 ๐‘˜=0 ๐‘›

1 ๐ต(๐‘› โˆ’ ๐‘˜)๐‘ฅ๐‘› ๐‘˜! (๐‘› โˆ’ ๐‘˜)! ๐‘˜=0

=โˆ‘ โˆ‘ ๐‘›โ‰ฅ0

=โˆ‘ ๐‘˜โ‰ฅ0

๐‘ฅ๐‘›โˆ’๐‘˜ ๐‘ฅ๐‘˜ โˆ‘ ๐ต(๐‘› โˆ’ ๐‘˜) ๐‘˜! ๐‘›โ‰ฅ๐‘˜ (๐‘› โˆ’ ๐‘˜)!

๐‘ฅ๐‘˜ ๐ต(๐‘ฅ) ๐‘˜! ๐‘˜โ‰ฅ0

=โˆ‘

= ๐‘’๐‘ฅ ๐ต(๐‘ฅ). We now have a differential equation to solve, but we must take some care to make sure this can be done formally. To this end we define the formal natural logarithm by (4.2)

ln(1 + ๐‘ฅ) = ๐‘ฅ โˆ’

๐‘ฅ2 ๐‘ฅ3 (โˆ’1)๐‘›โˆ’1 ๐‘ฅ๐‘› + โˆ’โ‹ฏ= โˆ‘ 2 3 ๐‘› ๐‘›โ‰ฅ1

which can be thought of as formally integrating the geometric series for 1/(1 + ๐‘ฅ). Note that since ln(1+๐‘ฅ) has an infinite number of terms, to have ln(1+๐‘“(๐‘ฅ)) well-defined it must be that ๐‘“(๐‘ฅ) has constant term 0 by Theorem 3.3.3. In other words, for an infinite series ๐‘”(๐‘ฅ), we have ln ๐‘”(๐‘ฅ) is only defined if the constant term of ๐‘”(๐‘ฅ) is 1. Luckily this is true of ๐ต(๐‘ฅ) so we can separate variables above to get ๐ต โ€ฒ (๐‘ฅ)/๐ต(๐‘ฅ) = ๐‘’๐‘ฅ and then formally integrate to get ln ๐ต(๐‘ฅ) = ๐‘’๐‘ฅ + ๐‘ for some constant ๐‘. By definition (4.2) a natural log has no constant term so we must take ๐‘ = โˆ’1. Solving for ๐ต(๐‘ฅ) we obtain the following result. Again, a more combinatorial proof will be given later. Theorem 4.1.2. We have โˆ‘ ๐ต(๐‘›) ๐‘›โ‰ฅ0

๐‘ฅ๐‘› ๐‘ฅ = ๐‘’๐‘’ โˆ’1 . ๐‘›!

โ–ก

We end this section by discussing certain permutations whose descent sets have a nice structure. To compute their egf we will need to define the formal sine power series by ๐‘ฅ3 ๐‘ฅ5 ๐‘ฅ2๐‘›+1 sin ๐‘ฅ = ๐‘ฅ โˆ’ + โˆ’ โ‹ฏ = โˆ‘ (โˆ’1)๐‘› 3! 5! (2๐‘› + 1)! ๐‘›โ‰ฅ0 and

1 sin ๐‘ฅ , tan ๐‘ฅ = . cos ๐‘ฅ cos ๐‘ฅ Note that sec ๐‘ฅ and tan ๐‘ฅ are well-defined by Theorem 3.3.1. cos ๐‘ฅ = (sin ๐‘ฅ)โ€ฒ ,

sec ๐‘ฅ =

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

120

4. Counting with Exponential Generating Functions

Call a permutation ๐œ‹ โˆˆ ๐‘ƒ([๐‘›]) alternating if (4.3)

๐œ‹1 > ๐œ‹2 < ๐œ‹3 > ๐œ‹4 < โ‹ฏ

or equivalently if Des ๐œ‹ consists of the odd number in [๐‘›]. The ๐‘›th Euler number is ๐ธ๐‘› = the number of alternating ๐œ‹ โˆˆ ๐‘ƒ([๐‘›]). For example, when ๐‘› = 4 then the alternating permutations are 2143, 3142, 3241, 4132, 4231 so ๐ธ4 = 5. A permutation is complement alternating if ๐œ‹1 < ๐œ‹2 > ๐œ‹3 < ๐œ‹4 > โ‹ฏ or equivalently ๐œ‹๐‘ is alternating where ๐œ‹๐‘ is the complement of ๐œ‹ as defined in Exercise (37)(b) of Chapter 1. Clearly ๐ธ๐‘› also counts the number of complement alternating ๐œ‹ โˆˆ ๐‘ƒ([๐‘›]). More generally, define any sequence of integers to be alternating using (4.3) and similarly for complement alternating. We have the following result for the Euler numbers. Theorem 4.1.3. We have ๐ธ0 = ๐ธ1 = 1 and, for ๐‘› โ‰ฅ 1, ๐‘›

๐‘› 2๐ธ๐‘›+1 = โˆ‘ ( )๐ธ ๐‘˜ ๐ธ๐‘›โˆ’๐‘˜ . ๐‘˜ ๐‘˜=0 Also โˆ‘ ๐ธ๐‘› ๐‘›โ‰ฅ0

๐‘ฅ๐‘› = sec ๐‘ฅ + tan ๐‘ฅ. ๐‘›!

Proof. To prove the recurrence it will be convenient to consider the set ๐‘† which is the union of all permutations which are either alternating or complement alternating in ๐‘ƒ([๐‘› + 1]). So #๐‘† = 2๐ธ๐‘›+1 . Pick ๐œ‹ โˆˆ ๐‘† and suppose ๐œ‹๐‘˜ = ๐‘› + 1. Then ๐œ‹ factors as a word ๐œ‹ = ๐œ‹โ€ฒ (๐‘› + 1)๐œ‹โ€ณ . Suppose first that ๐œ‹ is alternating. Then ๐‘˜ is odd, ๐œ‹โ€ฒ is alternating, and ๐œ‹โ€ณ is complement alternating. The number of ways of choosing the elements of ๐œ‹โ€ฒ is (๐‘›๐‘˜) and the remaining ones are used for ๐œ‹โ€ณ . The number of ways of arranging the elements for ๐œ‹โ€ฒ in alternating order is ๐ธ ๐‘˜ , and for ๐œ‹โ€ณ it is ๐ธ๐‘›โˆ’๐‘˜ . So the total number of such ๐œ‹ is (๐‘›๐‘˜)๐ธ ๐‘˜ ๐ธ๐‘›โˆ’๐‘˜ where ๐‘˜ is odd. Similar considerations show that the same formula holds for even ๐‘˜ when ๐œ‹ is complement alternating. The summation side of the recursion follows. Now let ๐ธ(๐‘ฅ) be the egf for the ๐ธ๐‘› . Multiplying the recurrence by ๐‘ฅ๐‘› /๐‘›! and summing over ๐‘› โ‰ฅ 1 one obtains the differential equation and boundary condition 2๐ธ โ€ฒ (๐‘ฅ) = ๐ธ(๐‘ฅ)2 + 1

and

๐ธ(0) = 1

where, to make things well-defined for formal power series, ๐ธ(0) is an abbreviation for the constant term of ๐ธ(๐‘ฅ). One now obtains the unique solution ๐ธ(๐‘ฅ) = sec ๐‘ฅ + tan ๐‘ฅ, either by separation of variables or by verifying that this function satisfies the differential equation and initial condition. โ–ก

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

4.2. Generating functions for Eulerian polynomials

121

4.2. Generating functions for Eulerian polynomials In Section 3.2 we saw that the inv and maj statistics have the same distribution. It turns out that there are statistics that have the same distribution as des and these are called Eulerian. The polynomials having this distribution have nice corresponding generating functions of both the ordinary and exponential type. We will discuss them in this section. A whole book devoted to this topic has been written by Petersen [69]. Given ๐‘›, ๐‘˜ โˆˆ โ„• with 0 โ‰ค ๐‘˜ < ๐‘›, the corresponding Eulerian number is ๐ด(๐‘›, ๐‘˜) = #{๐œ‹ โˆˆ ๐”–๐‘› โˆฃ des ๐œ‹ = ๐‘˜}. As usual, we let ๐ด(๐‘›, ๐‘˜) = 0 if ๐‘˜ < 0 or ๐‘˜ โ‰ฅ ๐‘› with the exception ๐ด(0, 0) = 1. For example, if ๐‘› = 3, then we have ๐‘˜ des ๐œ‹ = ๐‘˜

0

1

2

123

132, 213, 231, 312

321

so that ๐ด(3, 0) = 1,

๐ด(3, 1) = 4,

๐ด(3, 2) = 1.

Be sure not to confuse these Eulerian numbers with the Euler numbers introduced in the previous section. Also, some authors use ๐ด(๐‘›, ๐‘˜) to denote the number of permutations in ๐”–๐‘› having ๐‘˜ โˆ’ 1 descents. Some elementary properties of the ๐ด(๐‘›, ๐‘˜) are given in the next result. It will be convenient to let ๐ด([๐‘›], ๐‘˜) = {๐œ‹ โˆˆ ๐”–๐‘› โˆฃ des ๐œ‹ = ๐‘˜}. Theorem 4.2.1. Suppose ๐‘› โ‰ฅ 0. (a) The Eulerian numbers satisfy the initial condition ๐ด(0, ๐‘˜) = ๐›ฟ ๐‘˜,0 and recurrence relation ๐ด(๐‘›, ๐‘˜) = (๐‘˜ + 1)๐ด(๐‘› โˆ’ 1, ๐‘˜) + (๐‘› โˆ’ ๐‘˜)๐ด(๐‘› โˆ’ 1, ๐‘˜ โˆ’ 1) for ๐‘› โ‰ฅ 1. (b) The Eulerian numbers are symmetric in that ๐ด(๐‘›, ๐‘˜) = ๐ด(๐‘›, ๐‘› โˆ’ ๐‘˜ โˆ’ 1). (c) We have โˆ‘ ๐ด(๐‘›, ๐‘˜) = ๐‘›! . ๐‘˜

Proof. We leave all except the recursion as an exercise. Suppose ๐œ‹ โˆˆ ๐ด([๐‘›], ๐‘˜). Then removing ๐‘› from ๐œ‹ results in ๐œ‹โ€ฒ โˆˆ ๐ด([๐‘› โˆ’ 1], ๐‘˜) or ๐œ‹โ€ณ โˆˆ ๐ด([๐‘› โˆ’ 1], ๐‘˜ โˆ’ 1) depending on the relative size of the elements to either side of ๐‘› in ๐œ‹. A permutation ๐œ‹โ€ฒ will result if either the ๐‘› in ๐œ‹ is in the space corresponding to a descent of ๐œ‹โ€ฒ , or at the end of ๐œ‹โ€ฒ . So a given ๐œ‹โ€ฒ will result ๐‘˜+1 times by this method, which accounts for the first term in the sum. Similarly, one obtains a ๐œ‹โ€ณ from ๐œ‹ if ๐‘› is either in the space of an ascent or at the beginning. So the total number of repetitions in this case is ๐‘› โˆ’ ๐‘˜ and the recurrence is proved. โ–ก Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

122

4. Counting with Exponential Generating Functions

The ๐‘›th Eulerian polynomial is ๐ด๐‘› (๐‘ž) = โˆ‘ ๐ด(๐‘›, ๐‘˜)๐‘ž๐‘˜ = โˆ‘ ๐‘ždes ๐œ‹ . ๐‘˜โ‰ฅ0

๐œ‹โˆˆ๐”–๐‘›

Any statistic having distribution ๐ด๐‘› (๐‘ž) is said to be an Eulerian statistic. One of the other famous Eulerian statistics counts excedances. An excedance of a permutation ๐œ‹ โˆˆ ๐”–๐‘› is an integer ๐‘– such that ๐œ‹(๐‘–) > ๐‘–. This gives rise to the excedance set Exc ๐œ‹ = {๐‘– โˆฃ ๐œ‹(๐‘–) > ๐‘–} and excedance statistic exc ๐œ‹ = # Exc ๐œ‹. To illustrate, if ๐œ‹ = 3167542, then ๐œ‹(1) = 3, ๐œ‹(3) = 6, and ๐œ‹(4) = 7 while ๐œ‹(๐‘–) โ‰ค ๐‘– for other ๐‘–. So Exc ๐œ‹ = {1, 3, 4} and exc ๐œ‹ = 3. Making a chart for ๐‘› = 3 as we did for des gives ๐‘˜ 0 1 2 exc ๐œ‹ = ๐‘˜

123

132, 213, 312, 321

231

so that the number of permutations in each column is given by the ๐ด(3, ๐‘˜), even though the sets of permutations in the two tables are not necessarily equal. In order to prove that ๐ด(๐‘›, ๐‘˜) also counts permutations by excedances, we will need a map that is so important in enumerative combinatorics that it is sometimes called the fundamental bijection. Before we can define this function, we will need some more concepts. Similar to what was done in Section 1.12, call an element ๐œ‹๐‘– of ๐œ‹ โˆˆ ๐”–๐‘› a left-right maximum if ๐œ‹๐‘– > max{๐œ‹1 , ๐œ‹2 , . . . , ๐œ‹๐‘–โˆ’1 }. Note that ๐œ‹1 and ๐‘› are always left-right maxima and that the left-right maxima increase left-to-right. To illustrate, the left-right maxima of ๐œ‹ = 51327846 are 5, 7, and 8. The left-right maxima of ๐œ‹ determine the left-right factorization of ๐œ‹ into factors ๐œ‹๐‘– ๐œ‹๐‘–+1 . . . ๐œ‹๐‘—โˆ’1 where ๐œ‹๐‘– is a left-right maximum and ๐œ‹๐‘— is the next such maximum. In our example ๐œ‹, the factorization is 5132, 7, and 846. Recall that since disjoint cycles commute, there are many ways of writing the disjoint cycle decomposition ๐œ‹ = ๐‘ 1 ๐‘ 2 โ‹ฏ ๐‘ ๐‘˜ . We wish to distinguish one which is analogous to the left-right factorization. The canonical cycle decomposition of ๐œ‹ is obtained by writing each ๐‘ ๐‘– so that it starts with max ๐‘ ๐‘– and then ordering the cycles so that max ๐‘ 1 < max ๐‘ 2 < โ‹ฏ < max ๐‘ ๐‘˜ . To illustrate, the permutation ๐œ‹ = (7, 1, 8)(2, 4, 5, 3)(6) written canonically becomes ๐œ‹ = (5, 3, 2, 4)(6)(8, 7, 1). The fundamental map is ฮฆ โˆถ ๐”–๐‘› โ†’ ๐”–๐‘› where ฮฆ(๐œ‹) is obtained by replacing each left-right factor ๐œ‹๐‘– ๐œ‹๐‘–+1 . . . ๐œ‹๐‘—โˆ’1 by the cycle (๐œ‹๐‘– , ๐œ‹๐‘–+1 , . . . , ๐œ‹๐‘—โˆ’1 ). For example ฮฆ(51327846) = (5, 1, 3, 2)(7)(8, 4, 6) = 35261874. Note that from the definitions it follows that the cycle decomposition of ฮฆ(๐œ‹) obtained will be the canonical one. It is also easy to construct an inverse for ฮฆ: given ๐œŽ โˆˆ ๐”–๐‘› , we construct its canonical cycle decomposition and then just remove the parentheses Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

4.2. Generating functions for Eulerian polynomials

123

and commas to get ๐œ‹. These maps are inverses since the inequalities defining the leftright factorization and canonical cycle decomposition are the same. We have proved the following. Theorem 4.2.2. The fundamental map ฮฆ โˆถ ๐”–๐‘› โ†’ ๐”–๐‘› is a bijection.

โ–ก

Corollary 4.2.3. For ๐‘›, ๐‘˜ โ‰ฅ 0 we have ๐ด(๐‘›, ๐‘˜) = number of ๐œ‹ โˆˆ ๐”–๐‘› with ๐‘˜ excedances. Proof. A coexcedance of ๐œ‹ โˆˆ ๐”–๐‘› is ๐‘– โˆˆ [๐‘›] such that ๐œ‹(๐‘–) < ๐‘–. Note that the distribution of coexcedances over ๐”–๐‘› is the same as for excedances. Indeed, we have a bijection on ๐”–๐‘› defined by ๐œ‹ โ†ฆ ๐œ‹โˆ’1 . And this bijection has the property that the number of excedances of ๐œ‹ is the number of coexcedances of ๐œ‹โˆ’1 because one obtains the twoline notation for ๐œ‹โˆ’1 (as defined in Section 1.5) by taking the two-line notation for ๐œ‹, interchanging the top and bottom lines, and then permuting the columns until the first row is 12 . . . ๐‘›. We now claim that if ฮฆ(๐œ‹) = ๐œŽ where ฮฆ is the fundamental bijection, then des ๐œ‹ is the number of coexcedances of ๐œŽ which will finish the proof. But if we we have a descent ๐œ‹๐‘– > ๐œ‹๐‘–+1 in ๐œ‹, then ๐œ‹๐‘– , ๐œ‹๐‘–+1 must be in the same factor of the left-right factorization. So in ฮฆ(๐œ‹) we have a cycle mapping ๐œ‹๐‘– to ๐œ‹๐‘–+1 . This makes ๐œ‹๐‘– a coexcedance of ๐œŽ. Similar ideas show that no ascent of ๐œŽ gives rise to a coexcedance of ๐œŽ and so we are done. โ–ก We will now derive two generating functions involving the Eulerian polynomials, one ordinary and one exponential. Theorem 4.2.4. For ๐‘› โ‰ฅ 0 we have ๐ด๐‘› (๐‘ž) = โˆ‘ (๐‘š + 1)๐‘› ๐‘ž๐‘š . (4.4) (1 โˆ’ ๐‘ž)๐‘›+1 ๐‘šโ‰ฅ0 Proof. We will count descent partitioned permutations ๐œ‹ which consist of a permutation ๐œ‹ โˆˆ ๐”–๐‘› which has bars inserted in some of its spaces either between elements or before ๐œ‹1 or after ๐œ‹๐‘› , subject to the restriction that the space between each descent ๐œ‹๐‘– > ๐œ‹๐‘–+1 must have a bar. For example, if ๐œ‹ = 2451376, then we could have ๐œ‹ = 24|5||137|6|. Let ๐‘(๐œ‹) be the number of bars in ๐œ‹. We will show that both sides of (4.4) are the generating function ๐‘“(๐‘ž) = โˆ‘๐œ‹ ๐‘ž๐‘(๐œ‹) . First of all, given ๐œ‹, what is its contribution to ๐‘“(๐‘ž)? First we must put bars in the descents of ๐œ‹ which results in a factor of ๐‘ždes ๐œ‹ . Now we can choose the rest of the bars by putting them in any of the ๐‘› + 1 spaces of ๐œ‹ with repetition allowed which, by Theorem 3.4.2, gives a factor of 1/(1 โˆ’ ๐‘ž)๐‘›+1 . So ๐‘“(๐‘ž) = โˆ‘ ๐œ‹โˆˆ๐”–๐‘›

๐‘ždes ๐œ‹ ๐ด๐‘› (๐‘ž) = (1 โˆ’ ๐‘ž)๐‘›+1 (1 โˆ’ ๐‘ž)๐‘›+1

which is the first half of what we wished to prove. On the other hand, the coefficient of ๐‘ž๐‘š in ๐‘“(๐‘ž) is the number of ๐œ‹ which have exactly ๐‘š bars. One can construct these permutations as follows. Start with ๐‘š bars Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

124

4. Counting with Exponential Generating Functions

which create ๐‘š + 1 spaces between them. Now place the numbers 1, . . . , ๐‘› between the bars, making sure that the numbers between two consecutive bars form an increasing sequence. So we are essentially placing ๐‘› distinguishable balls into ๐‘š + 1 distinguishable boxes since the ordering in each box is fixed. By the twelvefold way, there are (๐‘š + 1)๐‘› ways of doing this, which completes the proof. โ–ก We can now use the ogf just derived to find the egf for the polynomials ๐ด๐‘› (๐‘ฅ). Theorem 4.2.5. We have โˆ‘ ๐ด๐‘› (๐‘ž)

(4.5)

๐‘›โ‰ฅ0

๐‘žโˆ’1 ๐‘ฅ๐‘› = . ๐‘›! ๐‘ž โˆ’ ๐‘’(๐‘žโˆ’1)๐‘ฅ

Proof. Multiply both sides of the equality in the previous theorem by the quantity (1 โˆ’ ๐‘ž)๐‘› ๐‘ฅ๐‘› /๐‘›! and sum over ๐‘›. The left-hand side becomes (1 โˆ’ ๐‘ž)๐‘› ๐ด๐‘› (๐‘ž)๐‘ฅ๐‘› 1 ๐‘ฅ๐‘› โˆ‘ ๐ด๐‘› (๐‘ž) . = ๐‘›+1 1 โˆ’ ๐‘ž ๐‘›โ‰ฅ0 ๐‘›! (1 โˆ’ ๐‘ž) ๐‘›! ๐‘›โ‰ฅ0 โˆ‘

And the right side is now โˆ‘ โˆ‘ ๐‘ž๐‘š ๐‘›โ‰ฅ0 ๐‘šโ‰ฅ0

[(1 โˆ’ ๐‘ž)๐‘ฅ(๐‘š + 1)]๐‘› (1 โˆ’ ๐‘ž)๐‘› (๐‘š + 1)๐‘› ๐‘ฅ๐‘› = โˆ‘ ๐‘ž๐‘š โˆ‘ ๐‘›! ๐‘›! ๐‘šโ‰ฅ0 ๐‘›โ‰ฅ0 = โˆ‘ ๐‘ž๐‘š ๐‘’(1โˆ’๐‘ž)๐‘ฅ(๐‘š+1) ๐‘šโ‰ฅ0

=

๐‘’(1โˆ’๐‘ž)๐‘ฅ 1 โˆ’ ๐‘ž๐‘’(1โˆ’๐‘ž)๐‘ฅ

1 . ๐‘’(๐‘žโˆ’1)๐‘ฅ โˆ’ ๐‘ž Setting the two sides equal and solving for the desired generating function completes the proof. โ–ก =

4.3. Labeled structures There is a method for working combinatorially with exponential generating functions which we will present in the following sections. It is based on Joyalโ€™s theory of species [47]. His original method used the machinery of categories and functors. But for the type of enumeration we will be doing, it is not necessary to use this level of generality. An exposition of the full theory can be found in the textbook of Bergeron, Labelle, and Leroux [11]. A labeled structure is a function ๐’ฎ which assigns to each finite set ๐ฟ a finite set ๐’ฎ(๐ฟ) such that (4.6)

#๐ฟ = #๐‘€ โŸน #๐’ฎ(๐ฟ) = #๐’ฎ(๐‘€).

We call ๐ฟ the label set and ๐’ฎ(๐ฟ) the set of structures on ๐ฟ. We let ๐‘ ๐‘› = #๐’ฎ(๐ฟ) Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

4.3. Labeled structures

125

for any ๐ฟ of cardinality ๐‘›, and this is well-defined because of (4.6). We sometimes use ๐’ฎ(โ‹…) as an alternative notation for the structure ๐’ฎ. We also have the corresponding egf ๐น๐’ฎ = ๐น๐’ฎ(โ‹…) (๐‘ฅ) = โˆ‘ ๐‘ ๐‘› ๐‘›โ‰ฅ0

๐‘ฅ๐‘› . ๐‘›!

Although these definitions may seem very abstract, we have already seen many examples of labeled structures. It is just that we have not identified them as such. The rest of this section will be devoted to putting these examples in context. A summary can be found in Table 4.1. To start, consider the labeled structure defined by ๐’ฎ(๐ฟ) = 2๐ฟ . So ๐’ฎ assigns to each label set ๐ฟ the set of subsets of ๐ฟ. To illustrate ๐’ฎ({๐‘Ž, ๐‘}) = {โˆ…, {๐‘Ž}, {๐‘}, {๐‘Ž, ๐‘}}. Clearly ๐’ฎ satisfies (4.6) with ๐‘ ๐‘› = #2[๐‘›] = 2๐‘› . So the associated generating function is (4.7)

๐น2โ‹… (๐‘ฅ) = โˆ‘ 2๐‘› ๐‘›โ‰ฅ0

๐‘ฅ๐‘› = ๐‘’2๐‘ฅ . ๐‘›!

We will also want to specify the size of the subsets under consideration by using the structure ๐’ฎ(๐ฟ) = (๐ฟ๐‘˜) for some fixed ๐‘˜ โ‰ฅ 0. Now we have ๐‘ ๐‘› = (๐‘›๐‘˜) and, using the fact that this binomial coefficient is zero for ๐‘› < ๐‘˜, (4.8)

๐‘› ๐‘ฅ๐‘› ๐‘›! ๐‘ฅ๐‘› ๐‘ฅ๐‘›โˆ’๐‘˜ ๐‘ฅ๐‘˜ ๐‘ฅ ๐‘ฅ๐‘˜ โˆ‘ โ‹… = ๐น( โ‹… ) (๐‘ฅ) = โˆ‘ ( ) = โˆ‘ = ๐‘’ . ๐‘˜ ๐‘›! ๐‘˜! ๐‘›โ‰ฅ๐‘˜ (๐‘› โˆ’ ๐‘˜)! ๐‘˜! ๐‘˜! (๐‘› โˆ’ ๐‘˜)! ๐‘›! ๐‘˜ ๐‘›โ‰ฅ0 ๐‘›โ‰ฅ๐‘˜

It will be convenient to have a map which adds no extra structure to the label set. So define the labeled structure ๐ธ(๐ฟ) = {๐ฟ}. Note that ๐ธ returns the set consisting of ๐ฟ itself, not the set consisting of the elements of ๐ฟ. Consequently ๐‘ ๐‘› = 1 for all ๐‘› and ๐น๐ธ = ๐‘’๐‘ฅ . The use of ๐ธ for this labeled structure reflects both the fact that its egf is the exponential function and also that the French word for โ€œsetโ€ is โ€œensembleโ€. (Joyal is a francophone.) We will also need to specify that a set be nonempty by defining ๐ธ(๐ฟ) = {

{๐ฟ} โˆ…

if ๐ฟ โ‰  โˆ…, if ๐ฟ = โˆ….

Note that ๐ธ(โˆ…) = {โˆ…} while ๐ธ(โˆ…) = โˆ…. For ๐ธ we clearly have ๐‘ ๐‘› = 1 for ๐‘› โ‰ฅ 1 and ๐‘ 0 = 0. It is also obvious that (4.9)

๐น๐ธ = ๐‘’๐‘ฅ โˆ’ 1.

For partitions of sets we will use the structure ๐ฟ โ†ฆ ๐ต(๐ฟ) where ๐ต(๐ฟ) is defined as in Section 1.4. So ๐‘ ๐‘› = ๐ต(๐‘›) and, by Theorem 4.1.2, the egf is ๐น๐ต = โˆ‘ ๐ต(๐‘›) ๐‘›โ‰ฅ0

๐‘ฅ๐‘› ๐‘ฅ = ๐‘’๐‘’ โˆ’1 . ๐‘›!

We will be able to give a combinatorial derivation of this fact once we derive the Exponential Formula in Section 4.5, rather than using the recursion and formal manipulations as we did before. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

126

4. Counting with Exponential Generating Functions

Table 4.1. Labeled structures

๐’ฎ(๐ฟ)

Counts

๐‘ ๐‘›

egf

2๐ฟ

subsets

2๐‘›

โˆ‘ 2๐‘› ๐‘›โ‰ฅ0

๐‘ฅ๐‘› = ๐‘’2๐‘ฅ ๐‘›!

๐ฟ ( ) ๐‘˜

๐‘˜-subsets

๐‘› ( ) ๐‘˜

๐‘› ๐‘ฅ๐‘› ๐‘ฅ๐‘˜ ๐‘’๐‘ฅ โˆ‘( ) = ๐‘˜ ๐‘›! ๐‘˜! ๐‘›โ‰ฅ0

๐ธ(๐ฟ)

sets

1

๐‘ฅ๐‘› = ๐‘’๐‘ฅ ๐‘›! ๐‘›โ‰ฅ0

๐ธ(๐ฟ)

nonempty sets

1 โˆ’ ๐›ฟ๐‘›,0

๐‘ฅ๐‘› = ๐‘’๐‘ฅ โˆ’ 1 ๐‘›! ๐‘›โ‰ฅ1

๐ต(๐ฟ)

set partitions

๐ต๐‘›

โˆ‘ ๐ต๐‘›

โˆ‘

โˆ‘

๐‘›โ‰ฅ0

๐‘†(๐ฟ, ๐‘˜)

๐‘†๐‘œ (๐ฟ, ๐‘˜)

๐”–(๐ฟ)

set partitions with ๐‘˜ blocks

๐‘†(๐‘›, ๐‘˜)

โˆ‘ ๐‘†(๐‘›, ๐‘˜) ๐‘›โ‰ฅ0

ordered version of ๐‘†(๐ฟ, ๐‘˜)

๐‘˜! ๐‘†(๐‘›, ๐‘˜)

permutations

๐‘›!

๐‘ฅ๐‘› 1 = (๐‘’๐‘ฅ โˆ’ 1)๐‘˜ ๐‘›! ๐‘˜!

๐‘˜! โˆ‘ ๐‘†(๐‘›, ๐‘˜) ๐‘›โ‰ฅ0

โˆ‘ ๐‘›! ๐‘›โ‰ฅ0

๐‘ฅ๐‘› = (๐‘’๐‘ฅ โˆ’ 1)๐‘˜ ๐‘›!

๐‘ฅ๐‘› 1 = ๐‘›! 1โˆ’๐‘ฅ

๐‘(๐‘›, ๐‘˜)

๐‘ฅ๐‘› 1 1 โˆ‘ ๐‘(๐‘›, ๐‘˜) = (ln ) ๐‘›! ๐‘˜! 1 โˆ’ ๐‘ฅ ๐‘›โ‰ฅ0

ordered version of ๐‘(๐ฟ, ๐‘˜)

๐‘˜! ๐‘(๐‘›, ๐‘˜)

๐‘ฅ๐‘› 1 ๐‘˜! โˆ‘ ๐‘(๐‘›, ๐‘˜) = (ln ) ๐‘›! 1 โˆ’ ๐‘ฅ ๐‘›โ‰ฅ0

permutations with 1 cycle

(๐‘› โˆ’ 1)!

โˆ‘ (๐‘› โˆ’ 1)!

๐‘(๐ฟ, ๐‘˜)

permutations with ๐‘˜ cycles

๐‘๐‘œ (๐ฟ, ๐‘˜)

๐‘(๐ฟ)

๐‘ฅ๐‘› ๐‘ฅ = ๐‘’๐‘’ โˆ’1 ๐‘›!

๐‘›โ‰ฅ1

๐‘˜

๐‘˜

๐‘ฅ๐‘› 1 = ln ๐‘›! 1โˆ’๐‘ฅ

Just as with subsets, we will restrict our attention to partitions with a given number of blocks by using ๐ฟ โ†ฆ ๐‘†(๐ฟ, ๐‘˜). Now we have ๐‘ ๐‘› = ๐‘†(๐‘›, ๐‘˜), a Stirling number of the second kind. But we have yet to find a closed form for the egf โˆ‘๐‘›โ‰ฅ0 ๐‘†(๐‘›, ๐‘˜)๐‘ฅ๐‘› /๐‘›! to verify that entry in Table 4.1. We will be able to do this easily once we have the sum and product rules for egfs presented in the next section. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

4.3. Labeled structures

127

We will sometimes work with set partitions where there is a specified ordering on the blocks and use the notation ๐‘†๐‘œ (๐ฟ, ๐‘˜) = {(๐ต1 , ๐ต2 , . . . , ๐ต๐‘˜ ) โˆฃ ๐ต1 /๐ต2 / . . . /๐ต๐‘˜ โŠข ๐ฟ}. We call these ordered set partitions or set compositions. Note that the ordering is of the blocks themselves, not of the elements in each block, so that ({1, 3}, {2}) โ‰  ({2}, {1, 3}) but ({1, 3}, {2}) = ({3, 1}, {2}). Clearly the labeled structure ๐ฟ โ†ฆ ๐‘†๐‘œ (๐ฟ, ๐‘˜) has ๐‘ ๐‘› = ๐‘˜! ๐‘†(๐‘›, ๐‘˜) and a similar statement can be made for the egf. We will also need weak set compositions where we will allow empty blocks. One can look at labeled structures on permutations analogously to what we have just seen for set partitions. In this context, consider a permutation of ๐ฟ to be a bijection ๐œ‹ โˆถ ๐ฟ โ†’ ๐ฟ decomposed into cycles as we did when ๐ฟ = [๐‘›] in Section 1.5. Let ๐’ฎ(๐ฟ) = ๐”–(๐ฟ) be the labeled structure of all permutations of ๐ฟ so that ๐‘ ๐‘› = ๐‘›! and ๐น๐”– = โˆ‘๐‘› ๐‘›! ๐‘ฅ๐‘› /๐‘›! = 1/(1 โˆ’ ๐‘ฅ). We have the associated structures ๐‘(๐ฟ, ๐‘˜) = {๐œ‹ = ๐‘ 1 ๐‘ 2 โ‹ฏ ๐‘ ๐‘˜ โˆฃ ๐œ‹ is a permutation of ๐ฟ with ๐‘˜ cycles ๐‘ ๐‘– } with ordered variant ๐‘๐‘œ (๐ฟ, ๐‘˜) = {(๐‘ 1 , ๐‘ 2 , . . . , ๐‘ ๐‘˜ ) โˆฃ the ๐‘ ๐‘– are the cycles of a permutation of ๐ฟ}. Using the signless Stirling numbers of the first kind we see that the sequences enumerating these two structures are ๐‘(๐‘›, ๐‘˜) and ๐‘˜! ๐‘(๐‘›, ๐‘˜), respectively. Again, we will wait to evaluate the corresponding egfs. Finally, we will find the special case ๐‘(๐ฟ) โ‰” ๐‘(๐ฟ, 1) of having only one cycle particularly useful. In this case, the enumerator is easy to compute. Proposition 4.3.1. We have #๐‘([๐‘›]) = {

(๐‘› โˆ’ 1)! 0

if ๐‘› โ‰ฅ 1, if ๐‘› = 0.

Proof. The empty permutation has no cycles so that ๐‘(โˆ…) = 0. Suppose ๐‘› โ‰ฅ 1 and consider (๐‘Ž1 , ๐‘Ž2 , . . . , ๐‘Ž๐‘› ) โˆˆ ๐‘([๐‘›]). Then the number of such cycles is the number of ways to order the ๐‘Ž๐‘– divided by the number of orderings which give the same cycle, namely ๐‘›! /๐‘› = (๐‘› โˆ’ 1)!. โ–ก

It follows from the previous proposition that ๐น๐‘ = โˆ‘ (๐‘› โˆ’ 1)! ๐‘›โ‰ฅ1

๐‘ฅ๐‘› ๐‘ฅ๐‘› 1 =โˆ‘ = ln ๐‘›! ๐‘› 1 โˆ’ ๐‘ฅ ๐‘›โ‰ฅ1

by definition (4.2). Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

128

4. Counting with Exponential Generating Functions

4.4. The Sum and Product Rules for egfs Just as with sets and ogfs, there is a Sum Rule and a Product Rule for egfs. To derive these results, we will first need corresponding rules for labeled structures. Suppose ๐’ฎ and ๐’ฏ are labeled structures. If ๐’ฎ(๐ฟ) โˆฉ ๐’ฏ(๐ฟ) = โˆ… for any finite set ๐ฟ, then we say ๐’ฎ and ๐’ฏ are disjoint. In this case we define their disjoint union structure, ๐’ฎ โŠŽ ๐’ฏ, by (๐’ฎ โŠŽ ๐’ฏ)(๐ฟ) = ๐’ฎ(๐ฟ) โŠŽ ๐’ฏ(๐ฟ). It is easy to see, and we will prove in Proposition 4.4.1 below, that ๐’ฎ โŠŽ ๐’ฏ satisfies the definition of a labeled structure. As examples of this concept, suppose #๐ฟ = ๐‘›. Then 2๐ฟ can be partitioned into the subsets of ๐ฟ having size ๐‘˜ for 0 โ‰ค ๐‘˜ โ‰ค ๐‘›. In other words ๐ฟ ๐ฟ ๐ฟ 2๐ฟ = ( ) โŠŽ ( ) โŠŽ โ‹ฏ โŠŽ ( ). 0 1 ๐‘›

(4.10)

Note that to make a statement about all ๐ฟ regardless of cardinality we can write 2๐ฟ = โจ„๐‘˜โ‰ฅ0 (๐ฟ๐‘˜) since (๐ฟ๐‘˜) = โˆ… for ๐‘˜ > #๐ฟ. Similarly, we have (4.11)

๐ต(๐ฟ) = ๐‘†(๐ฟ, 0) โŠŽ ๐‘†(๐ฟ, 1) โŠŽ โ‹ฏ โŠŽ ๐‘†(๐ฟ, ๐‘›)

and (4.12)

๐”–(๐ฟ) = ๐‘(๐ฟ, 0) โŠŽ ๐‘(๐ฟ, 1) โŠŽ โ‹ฏ โŠŽ ๐‘(๐ฟ, ๐‘›).

To define products, let ๐’ฎ and ๐’ฏ be arbitrary labeled structures. Their product, ๐’ฎร—๐’ฏ, is defined by (๐’ฎ ร— ๐’ฏ)(๐ฟ) = {(๐‘†, ๐‘‡) โˆฃ ๐‘† โˆˆ ๐’ฎ(๐ฟ1 ), ๐‘‡ โˆˆ ๐’ฏ(๐ฟ2 ) with (๐ฟ1 , ๐ฟ2 ) a weak composition of ๐ฟ}. Intuitively, we carve ๐ฟ up into two subsets in all possible ways and put an ๐’ฎ-structure on the first subset and a ๐’ฏ-structure on the second. Again, we will show that this is indeed a labeled structure in Proposition 4.4.1. Strictly speaking, ๐’ฎ ร— ๐’ฏ should be a multiset since it is possible that the same pair (๐‘†, ๐‘‡) could arise from two different ordered partitions. However, in the examples we will use, this will never be the case. And the theorems we will prove about the product will still be true in the more general context if we count with multiplicity. In order to give some examples using products, we will need a notion of equivalence of structures. Say that labeled structures ๐’ฎ and ๐’ฏ are equivalent, and write ๐’ฎ โ‰ก ๐’ฏ if #๐’ฎ(๐ฟ) = #๐’ฏ(๐ฟ) for all finite ๐ฟ. Sometimes we will write ๐’ฎ(๐ฟ) โ‰ก ๐’ฏ(๐ฟ) for this concept if the context makes inclusion of a generic label set ๐ฟ convenient. Clearly if ๐’ฎ โ‰ก ๐’ฏ, then ๐น๐’ฎ (๐‘ฅ) = ๐น๐’ฏ (๐‘ฅ). As a first illustration of these concepts, we claim that (4.13)

2โ‹… โ‰ก (๐ธ ร— ๐ธ)(โ‹…).

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

4.4. The Sum and Product Rules for egfs

129

To see this, note that subsets ๐‘† โˆˆ 2๐ฟ are in bijection with weak compositions via the map ๐‘† โ†” (๐‘†, ๐ฟ โˆ’ ๐‘†). So #2๐ฟ = #(๐ธ ร— ๐ธ)(๐ฟ) as we wished to show. These same ideas demonstrate that we have ๐‘†๐‘œ (โ‹…, 2) = (๐ธ ร— ๐ธ)(โ‹…) and more generally, for any ๐‘˜ โ‰ฅ 0, ๐‘˜

๐‘†๐‘œ (โ‹…, ๐‘˜) = ๐ธ (โ‹…).

(4.14) In a similar manner, we obtain

๐‘๐‘œ (โ‹…, ๐‘˜) = ๐‘๐‘˜ (โ‹…).

(4.15)

It is time to prove the Sum and Product Rules for labeled structures. In so doing, we will also be showing that they satisfy the definition for a labeled structure (4.6). Proposition 4.4.1. Let ๐’ฎ, ๐’ฏ be labeled structures and let ๐‘ ๐‘› = #๐’ฎ(๐ฟ),

๐‘ก๐‘› = #๐’ฏ(๐ฟ)

where #๐ฟ = ๐‘›. (a) (Sum Rule) If ๐’ฎ and ๐’ฏ are disjoint, then #(๐’ฎ โŠŽ ๐’ฏ)(๐ฟ) = ๐‘ ๐‘› + ๐‘ก๐‘› . (b) (Product Rule) For any ๐’ฎ, ๐’ฏ ๐‘›

๐‘› #(๐’ฎ ร— ๐’ฏ)(๐ฟ) = โˆ‘ ( )๐‘ ๐‘˜ ๐‘ก๐‘›โˆ’๐‘˜ . ๐‘˜ ๐‘˜=0 Proof. For part (a) we have #(๐’ฎ โŠŽ ๐’ฏ)(๐ฟ) = #(๐’ฎ(๐ฟ) โŠŽ ๐’ฏ(๐ฟ)) = #๐’ฎ(๐ฟ) + #๐’ฏ(๐ฟ) = ๐‘ ๐‘› + ๐‘ก๐‘› . Now consider part (b). In order to construct (๐‘†, ๐‘‡) โˆˆ (๐’ฎ ร— ๐’ฏ)(๐ฟ) we must first pick a weak composition ๐ฟ = ๐ฟ1 โŠŽ ๐ฟ2 . This is equivalent to just picking ๐ฟ1 as then ๐ฟ2 = ๐ฟ โˆ’ ๐ฟ1 . So if #๐ฟ1 = ๐‘˜, then there are (๐‘›๐‘˜) ways to perform this step. Next we must put an ๐’ฎ-structure on ๐ฟ1 and a ๐’ฏ-structure on ๐ฟ2 which can be done in ๐‘ ๐‘˜ ๐‘ก๐‘›โˆ’๐‘˜ ways. Multiplying together the two counts and summing over all possible ๐‘˜ yields the desired formula. โ–ก As application of this result, note that applying the Sum Rule to (4.10) just gives 2๐‘› = โˆ‘๐‘˜ (๐‘›๐‘˜) which is Theorem 1.3.3(c). And if we apply the Product Rule to (4.13), we get ๐‘›

๐‘›

๐‘› ๐‘› 2๐‘› = โˆ‘ ( ) โ‹… 1 โ‹… 1 = โˆ‘ ( ) ๐‘˜ ๐‘˜ ๐‘˜=0 ๐‘˜=0 again. Somewhat more interesting formulas are derived in Exercise 8(b) of this chapter. We can now translate Proposition 4.4.1 into the corresponding rules for exponential generating functions. This will permit us to fill in the entries in Table 4.1 which were postponed in the previous section. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

130

4. Counting with Exponential Generating Functions

Theorem 4.4.2. Let ๐’ฎ, ๐’ฏ be labeled structures. (a) (Sum Rule) If ๐’ฎ and ๐’ฏ are disjoint, then ๐น๐’ฎโŠŽ๐’ฏ (๐‘ฅ) = ๐น๐’ฎ (๐‘ฅ) + ๐น๐’ฏ (๐‘ฅ). (b) (Product Rule) For any ๐’ฎ, ๐’ฏ ๐น๐’ฎร—๐’ฏ (๐‘ฅ) = ๐น๐’ฎ (๐‘ฅ) โ‹… ๐น๐’ฏ (๐‘ฅ). Proof. Let ๐‘ ๐‘› = ๐’ฎ([๐‘›]) and ๐‘ก๐‘› = ๐’ฏ([๐‘›]). Using the Sum Rule in Proposition 4.4.1 gives ๐น๐’ฎ (๐‘ฅ) + ๐น๐’ฏ (๐‘ฅ) = โˆ‘ ๐‘ ๐‘› ๐‘›โ‰ฅ0

๐‘ฅ๐‘› ๐‘ฅ๐‘› + โˆ‘ ๐‘ก๐‘› ๐‘›! ๐‘›โ‰ฅ0 ๐‘›!

= โˆ‘ (๐‘ ๐‘› + ๐‘ก๐‘› ) ๐‘›โ‰ฅ0

๐‘ฅ๐‘› ๐‘›!

= โˆ‘ #(๐’ฎ โŠŽ ๐’ฏ)([๐‘›]) ๐‘›โ‰ฅ0

๐‘ฅ๐‘› ๐‘›!

= ๐น๐’ฎโŠŽ๐’ฏ (๐‘ฅ). Now using the Product Rule of the same proposition yields ๐น๐’ฎ (๐‘ฅ)๐น๐’ฏ (๐‘ฅ) = ( โˆ‘ ๐‘ ๐‘› ๐‘›โ‰ฅ0

๐‘ฅ๐‘› ๐‘ฅ๐‘› )( โˆ‘ ๐‘ก๐‘› ) ๐‘›! ๐‘›โ‰ฅ0 ๐‘›!

๐‘›

๐‘ ๐‘˜ ๐‘ก โ‹… ๐‘›โˆ’๐‘˜ )๐‘ฅ๐‘› ๐‘˜! (๐‘› โˆ’ ๐‘˜)! ๐‘˜=0

= โˆ‘(โˆ‘ ๐‘›โ‰ฅ0

๐‘›

๐‘› ๐‘ฅ๐‘› = โˆ‘ ( โˆ‘ ( )๐‘ ๐‘˜ ๐‘ก๐‘›โˆ’๐‘˜ ) ๐‘›! ๐‘˜ ๐‘›โ‰ฅ0 ๐‘˜=0 = โˆ‘ #(๐’ฎ ร— ๐’ฏ)([๐‘›]) ๐‘›โ‰ฅ0

๐‘ฅ๐‘› ๐‘›!

= ๐น๐’ฎร—๐’ฏ (๐‘ฅ), โ–ก

which completes the proof.

As an illustration of how this result can be used, we can apply the Sum Rule to (4.10), keeping in mind the comment following the equation, to write ๐น2โ‹… (๐‘ฅ) = โˆ‘ ๐น(โ‹…) (๐‘ฅ). ๐‘˜ ๐‘˜โ‰ฅ0

We can check this using (4.7) and (4.8): ๐‘ฅ๐‘˜ ๐‘ฅ ๐‘ฅ๐‘˜ ๐‘’ = ๐‘’๐‘ฅ โˆ‘ = ๐‘’๐‘ฅ โ‹… ๐‘’๐‘ฅ = ๐‘’2๐‘ฅ = ๐น2โ‹…(๐‘ฅ). ๐‘˜! ๐‘˜! ๐‘˜โ‰ฅ0 ๐‘˜โ‰ฅ0

โˆ‘ ๐น(โ‹…) (๐‘ฅ) = โˆ‘

๐‘˜โ‰ฅ0

๐‘˜

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

4.5. The Exponential Formula

131

One can also apply the Product Rule to (4.13) and obtain ๐น2โ‹…(๐‘ฅ) = ๐น๐ธ (๐‘ฅ)๐น๐ธ (๐‘ฅ). Again, this yields a simple identity; namely ๐‘’2๐‘ฅ = ๐‘’๐‘ฅ โ‹… ๐‘’๐‘ฅ . The true power of Theorem 4.4.2 is that it can be used to derive egfs which are more complicated to prove by other means. For example, applying the Product Rule to equation (4.14) along with (4.9) yields ๐น๐‘†๐‘œ (โ‹…,๐‘˜) (๐‘ฅ) = ๐น๐ธ (๐‘ฅ)๐‘˜ = (๐‘’๐‘ฅ โˆ’ 1)๐‘˜ , a new entry for Table 4.1. Furthermore, since ๐น๐‘†๐‘œ (โ‹…,๐‘˜) (๐‘ฅ) = ๐‘˜! ๐น๐‘†(โ‹…,๐‘˜) (๐‘ฅ) we obtain (๐‘’๐‘ฅ โˆ’ 1)๐‘˜ . ๐‘˜! This permits us to give another derivation of the egf for ๐ต(๐‘›). Using the Sum Rule and (4.11) gives ๐น๐‘†(โ‹…,๐‘˜) (๐‘ฅ) =

(๐‘’๐‘ฅ โˆ’ 1)๐‘˜ ๐‘ฅ = ๐‘’๐‘’ โˆ’1 . ๐‘˜! ๐‘˜โ‰ฅ0

๐น๐ต (๐‘ฅ) = โˆ‘ ๐น๐‘†(โ‹…,๐‘˜) (๐‘ฅ) = โˆ‘ ๐‘˜โ‰ฅ0

These same ideas can be used to derive the egfs for permutations with a given number of cycles, as the reader is asked to do in the exercises.

4.5. The Exponential Formula Often in combinatorics and other areas of mathematics there are objects which can be broken down into components. For example, the components of set partitions are blocks and the components of permutations are cycles. The exponential formula determines the egf of a labeled structure in terms of the egf for its components. It can also be considered as an analogue of the Product Rule for egfs where one carves ๐ฟ into an arbitrary number of subsets (rather than just 2) and the subsets are unordered (rather than ordered). To make these ideas precise, let ๐’ฎ be a labeled structure satisfying (4.16)

๐’ฎ(๐ฟ) โˆฉ ๐’ฎ(๐‘€) = โˆ… if ๐ฟ โ‰  ๐‘€.

The corresponding partition structure, ฮ (๐’ฎ), is defined by (ฮ (๐’ฎ))(๐ฟ) = {{๐‘† 1 , ๐‘† 2 , . . . } โˆฃ for all ๐ฟ1 /๐ฟ2 / . . . โŠข ๐ฟ with ๐‘† ๐‘– โˆˆ ๐’ฎ(๐ฟ๐‘– ) for all ๐‘–}. Intuitively, to form (ฮ (๐’ฎ))(๐ฟ) we partition the label set ๐ฟ in all possible ways and then put a structure from ๐’ฎ on each block of the partition, again in all possible ways. Condition (4.16) is imposed so that each element of (ฮ (๐’ฎ))(๐ฟ) can only arise in one way from this process. To illustrate, (4.17)

๐ต(๐ฟ) = {๐ฟ1 /๐ฟ2 / . . . โŠข ๐ฟ} โ‰ก (ฮ (๐ธ))(๐ฟ)

since ๐ฟ๐‘– is the only element of ๐ธ(๐ฟ๐‘– ) for any ๐‘–. In much the same way, we see that (4.18)

๐”–(๐ฟ) = {๐‘ 1 ๐‘ 2 โ‹ฏ โˆฃ ๐‘ ๐‘– a cycle on ๐ฟ๐‘– for all ๐‘– for all ๐ฟ1 /๐ฟ2 / . . . โŠข ๐ฟ} โ‰ก (ฮ (๐‘))(๐ฟ).

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

132

4. Counting with Exponential Generating Functions

There is a simple relationship between the egf for ฮ (๐’ฎ) and the egf for ๐’ฎ which is the labeled structure defined by ๐’ฎ(๐ฟ) = {

๐’ฎ(๐ฟ) โˆ…

if ๐ฟ โ‰  โˆ…, if ๐ฟ = โˆ….

So if ๐‘ ๐‘› = #๐’ฎ([๐‘›]), then ๐น๐’ฎ (๐‘ฅ) = โˆ‘ ๐‘ ๐‘› ๐‘›โ‰ฅ1

๐‘ฅ๐‘› . ๐‘›!

We need ๐น๐’ฎ (๐‘ฅ) to have a zero constant term so that the composition in the next result will be well-defined, see Theorem 3.3.3. Theorem 4.5.1 (Exponential Formula). If ๐’ฎ is a labeled structure satisfying (4.16), then ๐นฮ (๐’ฎ) (๐‘ฅ) = ๐‘’๐น๐’ฎ (๐‘ฅ) . Proof. We have

๐น๐’ฎ (๐‘ฅ)๐‘˜ . ๐‘˜! ๐‘˜โ‰ฅ0

๐‘’๐น๐’ฎ (๐‘ฅ) = โˆ‘

From the Product Rule for egfs in Theorem 4.4.2 we see that ๐น๐’ฎ (๐‘ฅ)๐‘˜ is the egf for putting ๐’ฎ-structures on partitions of the label set into ๐‘˜ ordered, nonempty blocks. So, by (4.16), ๐น๐’ฎ (๐‘ฅ)๐‘˜ /๐‘˜! is the egf for putting ๐’ฎ-structures on partitions of the label set into ๐‘˜ unordered, nonempty blocks. Now using the Sum Rule for egfs, again from Theorem 4.4.2, it follows that โˆ‘๐‘˜โ‰ฅ0 ๐น๐’ฎ (๐‘ฅ)๐‘˜ /๐‘˜! is the egf for putting ๐’ฎ-structures on partitions of the label set into any number of unordered, nonempty blocks. But this is exactly the structure ฮ (๐’ฎ) and so we are done. โ–ก As a first application of the Exponential Formula, consider (4.17). In this case ๐’ฎ = ๐ธ and ๐น๐ธ (๐‘ฅ) = ๐‘’๐‘ฅ โˆ’ 1. So, applying the previous theorem, ๐น๐ต (๐‘ฅ) = ๐นฮ (๐ธ) (๐‘ฅ) = ๐‘’๐น๐ธ (๐‘ฅ) = ๐‘’๐‘’

๐‘ฅ

โˆ’1

.

Even though we already knew this generating function, this proof is definitely the simplest both computationally and conceptually. We can use (4.18) in a similar manner. Now ๐’ฎ = ๐‘ and ๐น๐‘ (๐‘ฅ) = ln(1/(1 โˆ’ ๐‘ฅ)) = ๐น๐‘ (๐‘ฅ) since the original egf already has no constant term. Applying the Exponential Formula gives 1 ๐น๐”– (๐‘ฅ) = ๐นฮ (๐‘) (๐‘ฅ) = ๐‘’๐น๐‘ (๐‘ฅ) = ๐‘’ln(1/(1โˆ’๐‘ฅ)) = 1โˆ’๐‘ฅ which at least agrees with what we computed previously for this egf, even though this in now a more roundabout way of getting it. But with Theorem 4.5.1 in hand it is easy to get more refined information about permutations or other labeled structures. For example, suppose we wish to give a simpler and more combinatorial derivation for the egf of the derangement numbers ๐ท(๐‘›) found in Theorem 4.1.1. The corresponding structure is defined by ๐’Ÿ(๐ฟ) = derangements on ๐ฟ. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

4.5. The Exponential Formula

133

In order to express ๐’Ÿ(๐ฟ) as a partition structure we need to permit only cycles of length two or greater. So let ๐‘(๐ฟ) if #๐ฟ โ‰ฅ 2, ๐’ฎ(๐ฟ) = { โˆ… otherwise. It follows that ๐’Ÿ โ‰ก ฮ (๐’ฎ). Furthermore, ๐‘ ๐‘› = {

(๐‘› โˆ’ 1)! 0

if ๐‘› โ‰ฅ 2, otherwise,

so that ๐น๐’ฎ (๐‘ฅ) = โˆ‘ (๐‘› โˆ’ 1)! ๐‘›โ‰ฅ2

๐‘ฅ๐‘› 1 ๐‘ฅ๐‘› =โˆ‘ = ln( ) โˆ’ ๐‘ฅ. ๐‘›! ๐‘› 1 โˆ’ ๐‘ฅ ๐‘›โ‰ฅ2

Applying the Exponential Formula gives โˆ‘ ๐ท(๐‘›)

(4.19)

๐‘›โ‰ฅ0

๐‘ฅ๐‘› 1 ๐‘’โˆ’๐‘ฅ = ๐น๐’Ÿ (๐‘ฅ) = ๐นฮ (๐’ฎ) (๐‘ฅ) = exp(ln( . ) โˆ’ ๐‘ฅ) = ๐‘›! 1โˆ’๐‘ฅ 1โˆ’๐‘ฅ

One can mine even more information from Theorem 4.5.1 since the proof shows that each of the summands ๐น๐’ฎ (๐‘ฅ)๐‘˜ /๐‘˜! has a combinatorial meaning. Define the hyperbolic sine and cosine functions to be the formal power series sinh ๐‘ฅ = ๐‘ฅ +

๐‘ฅ3 ๐‘ฅ5 ๐‘ฅ2๐‘›+1 + +โ‹ฏ= โˆ‘ 3! 5! (2๐‘› + 1)! ๐‘›โ‰ฅ0

and cosh ๐‘ฅ = 1 +

๐‘ฅ2 ๐‘ฅ4 ๐‘ฅ2๐‘› . + +โ‹ฏ= โˆ‘ 2! 4! (2๐‘›)! ๐‘›โ‰ฅ0

It is easy to see that for any formal power series ๐‘“(๐‘ฅ) = โˆ‘๐‘›โ‰ฅ0 ๐‘Ž๐‘› ๐‘ฅ๐‘› we can extract the series of odd or even terms by (4.20)

โˆ‘ ๐‘Ž2๐‘›+1 ๐‘ฅ2๐‘›+1 = ๐‘›โ‰ฅ0

๐‘“(๐‘ฅ) โˆ’ ๐‘“(โˆ’๐‘ฅ) 2

โˆ‘ ๐‘Ž2๐‘› ๐‘ฅ2๐‘› =

and

๐‘›โ‰ฅ0

๐‘“(๐‘ฅ) + ๐‘“(โˆ’๐‘ฅ) . 2

It follows that ๐‘’๐‘ฅ โˆ’ ๐‘’โˆ’๐‘ฅ and 2 Define the odd partition structure ฮ ๐‘œ (๐’ฎ) by (4.21)

sinh ๐‘ฅ =

cosh ๐‘ฅ =

๐‘’๐‘ฅ + ๐‘’โˆ’๐‘ฅ . 2

(ฮ ๐‘œ (๐’ฎ))(๐ฟ) = {{๐‘† 1 , ๐‘† 2 , . . . } โˆˆ (ฮ (๐’ฎ))(๐ฟ) โˆฃ ๐ฟ partitioned into an odd number of blocks} and similarly define the even partition structure ฮ ๐‘’ (๐’ฎ). A proof like that of the Exponential Formula can be used to demonstrate the following. Theorem 4.5.2. If ๐’ฎ is a labeled structure satisfying (4.16), then ๐นฮ ๐‘œ (๐’ฎ) (๐‘ฅ) = sinh ๐น๐’ฎ (๐‘ฅ) and ๐นฮ ๐‘’ (๐’ฎ) (๐‘ฅ) = cosh ๐น๐’ฎ (๐‘ฅ). Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

โ–ก

134

4. Counting with Exponential Generating Functions

Now suppose we wish to find the egf for ๐‘Ž๐‘› which is the number of permutations of [๐‘›] that have an odd number of cycles. As before ๐’ฎ = ๐‘ with ๐น๐‘ (๐‘ฅ) = ๐น๐‘ (๐‘ฅ) = ln(1/(1 โˆ’ ๐‘ฅ)). Using Theorem 4.5.2 and then (4.21) we see that

โˆ‘ ๐‘Ž๐‘› ๐‘›โ‰ฅ0

๐‘ฅ๐‘› = sinh ๐น๐‘ (๐‘ฅ) ๐‘›! =

๐‘’ln(1/(1โˆ’๐‘ฅ)) โˆ’ ๐‘’โˆ’ ln(1/(1โˆ’๐‘ฅ)) 2

=

1 1 โˆ’ (1 โˆ’ ๐‘ฅ)) ( 2 1โˆ’๐‘ฅ

=๐‘ฅ+

1 โˆ‘ ๐‘ฅ๐‘› . 2 ๐‘›โ‰ฅ2

Extracting the coefficient of ๐‘ฅ๐‘› /๐‘›! from the first and last sums above yields

(4.22)

๐‘Ž๐‘› = {

๐‘›! /2 1

if ๐‘› โ‰ฅ 2, if ๐‘› = 1.

Of course, once one has obtained such a simple answer, one would like a purely combinatorial explanation and the reader is encouraged to find one in Exercise 12(c) of this chapter.

Exercises (1) (a) Use the recursion for the derangement numbers in Exercise 4 of Chapter 2 to reprove Theorem 4.1.1. (b) Use the recursion for the derangement numbers in Exercise 5 of Chapter 2 to reprove Theorem 4.1.1. (2) (a) Finish the proof of Theorem 4.2.1. (b) Finish the proof of Corollary 4.2.3. (c) Give two proofs of the identity

(๐‘š + 1)๐‘› = โˆ‘ ๐ด(๐‘›, ๐‘˜)( ๐‘˜โ‰ฅ0

๐‘š+๐‘›โˆ’๐‘˜ ), ๐‘›

one using equation (4.4) and one using descent partitioned permutations. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Exercises

135

(d) Give a combinatorial proof of the following formula for ๐ด(๐‘›, ๐‘˜): ๐‘˜+1

๐ด(๐‘›, ๐‘˜) = โˆ‘ (โˆ’1)๐‘– ( ๐‘–=0

๐‘›+1 )(๐‘˜ โˆ’ ๐‘– + 1)๐‘› . ๐‘–

Hint: Use the Principle of Inclusion and Exclusion and descent partitioned permutations. (e) Give a combinatorial proof of the following recursion for ๐‘› โ‰ฅ 1: ๐‘›โˆ’2

๐‘›โˆ’1 ๐ด๐‘› (๐‘ž) = ๐ด๐‘›โˆ’1 (๐‘ž) + ๐‘ž โˆ‘ ( )๐ด๐‘– (๐‘ž)๐ด๐‘›โˆ’๐‘–โˆ’1 (๐‘ž). ๐‘– ๐‘–=0 Hint: Factor each ๐œ‹ โˆˆ ๐”–๐‘› as ๐œ‹ = ๐œŽ๐‘›๐œ. (f) Use (e) to give a second proof of (4.5). (3) Let ๐ผ โŠ‚ โ„™ be finite and let ๐‘š = max ๐ผ if ๐ผ is nonempty or ๐‘š = 0 if ๐ผ = โˆ…. For ๐‘› > ๐‘š define the corresponding descent polynomial ๐‘‘(๐ผ; ๐‘›) to be the number of ๐œ‹ โˆˆ ๐”–๐‘› such that Des ๐œ‹ = ๐ผ. (a) Prove that ๐‘‘([๐‘˜]; ๐‘›) = (๐‘›โˆ’1 ). ๐‘˜ (b) If ๐ผ โ‰  โˆ…, then let ๐ผ โˆ’ = ๐ผ โˆ’ {๐‘š}. Prove that ๐‘› ๐‘‘(๐ผ; ๐‘›) = ( )๐‘‘(๐ผ โˆ’ ; ๐‘š) โˆ’ ๐‘‘(๐ผ โˆ’ ; ๐‘›). ๐‘š

(c) (d) (e) (f)

(g)

Hint: Consider the set of ๐œ‹ โˆˆ ๐”–๐‘› such that Des(๐œ‹1 ๐œ‹2 . . . ๐œ‹๐‘š ) = ๐ผ โˆ’ and ๐œ‹๐‘š+1 < ๐œ‹๐‘š+2 < โ‹ฏ < ๐œ‹๐‘› . Use part (b) to show that ๐‘‘(๐ผ; ๐‘›) is a polynomial in ๐‘› having degree deg(๐ผ; ๐‘›) = ๐‘š. Reprove the fact that ๐‘‘(๐ผ; ๐‘›) is a polynomial in ๐‘› using the Principle of Inclusion and Exclusion. Since ๐‘‘(๐ผ; ๐‘›) is a polynomial in ๐‘›, its domain of definition can be extended to all ๐‘› โˆˆ โ„‚. Show that if ๐‘– โˆˆ ๐ผ, then ๐‘‘(๐ผ; ๐‘–) = 0. Show that the complex roots of ๐‘‘(๐ผ; ๐‘›) all lie in the circle |๐‘ง| โ‰ค ๐‘š in the complex plane and also all have real part greater than or equal to โˆ’1. Note: This seems to be a difficult problem. (Conjecture) Show that the complex roots of ๐‘‘(๐ผ; ๐‘›) all lie in the circle ||๐‘ง โˆ’ ๐‘š + 1 || โ‰ค ๐‘š โˆ’ 1 . | 2 | 2 Note that this conjecture implies part (f).

(4) (a) Derive the generating function โˆ‘ ๐‘†(๐‘›, ๐‘˜)๐‘ก๐‘˜ ๐‘›,๐‘˜โ‰ฅ0

๐‘ฅ๐‘› ๐‘ฅ = ๐‘’๐‘ก(๐‘’ โˆ’1) ๐‘›!

in two ways: using the recursion for the ๐‘†(๐‘›, ๐‘˜) and using the generating functions in Table 4.1. (b) Rederive the egf for the Bell numbers ๐ต(๐‘›) using part (a). (5) (a) Find a formula for โˆ‘๐‘›,๐‘˜โ‰ฅ0 ๐‘(๐‘›, ๐‘˜)๐‘ก๐‘˜ ๐‘ฅ๐‘› /๐‘›! and prove it in two ways: using the recursion for the ๐‘(๐‘›, ๐‘˜) and using the generating functions in Table 4.1. (b) Rederive the egf for the permutation structure ๐”–(โ‹…) using part (a). Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

136

4. Counting with Exponential Generating Functions

(6) (a) Let ๐‘–๐‘› be the number of involutions in ๐”–๐‘› . Show that ๐‘–0 = ๐‘–1 = 1 and for ๐‘› โ‰ฅ 2 ๐‘–๐‘› = ๐‘–๐‘›โˆ’1 + (๐‘› โˆ’ 1)๐‘–๐‘›โˆ’2 . (b) Show that โˆ‘ ๐‘–๐‘› ๐‘›โ‰ฅ0

๐‘ฅ๐‘› 2 = ๐‘’๐‘ฅ+๐‘ฅ /2 ๐‘›!

in two ways: using the recursion in part (a) and using the Exponential Formula. (c) Given ๐ด โŠ† โ„™, let ๐‘†(๐‘›, ๐ด) be the number of partitions of [๐‘›] all of whose block sizes are elements of ๐ด. Use the Exponential Formula to find and prove a formula for โˆ‘โ‰ฅ0 ๐‘†(๐‘›, ๐ด)๐‘ฅ๐‘› /๐‘›!. (d) Repeat part (c) for ๐‘(๐‘›, ๐ด), the number of permutations of [๐‘›] all of whose cycles have lengths which are elements of ๐ด. (7) Fill in the details for finding the egf and solving the differential equation in the proof of Theorem 4.1.3. (8) (a) Use (4.14) and the Product Rule for labeled structures to show that ๐‘†(๐‘›, 2) = 2๐‘›โˆ’1 โˆ’ 1. (b) Use (4.15) and the Product Rule for labeled structures to show that ๐‘›

1 . ๐‘˜ ๐‘˜=1

๐‘(๐‘› + 1, 2) = ๐‘›! โˆ‘

(9) (a) Use the Theorem 4.4.2 to derive the egfs in Table 4.1 for the structures ๐‘๐‘œ (โ‹…, ๐‘˜) and ๐‘(โ‹…, ๐‘˜). (b) Use part (a) to rederive the egf for the structure ๐”–(โ‹…). (10) (a) Suppose ๐’ฎ is a labeled structure satisfying (4.16) and ๐’ฏ is any labeled structure. Their composition, ๐’ฏ โˆ˜ ๐’ฎ, is the structure such that (๐’ฏ โˆ˜ ๐’ฎ)(๐ฟ) = {({๐‘† 1 , ๐‘† 2 , . . . }, ๐‘‡) โˆฃ for all ๐ฟ1 /๐ฟ2 / . . . โŠข ๐ฟ with ๐‘† ๐‘– โˆˆ ๐’ฎ(๐ฟ๐‘– ) for all ๐‘– and ๐‘‡ โˆˆ ๐’ฏ({๐‘† 1 , ๐‘† 2 , . . . })} . Prove that ๐น๐’ฏโˆ˜๐’ฎ (๐‘ฅ) = ๐น๐’ฏ (๐น๐’ฎ (๐‘ฅ)). (b) Use (a) to reprove the Exponential Formula. (11) Let โ„ฑ(๐ฟ) be the labeled structures consisting of all forests with ๐ฟ as vertex set. Show that ๐‘ฅ๐‘› โˆ‘ #โ„ฑ([๐‘›]) = exp(โˆ‘ ๐‘›๐‘›โˆ’2 ๐‘ฅ๐‘› /๐‘›! ). ๐‘›! ๐‘›โ‰ฅ0 ๐‘›โ‰ฅ1 (12) (a) Prove the identities (4.20). (b) Prove Theorem 4.5.2. (c) Reprove (4.22) by finding a bijection between the permutations of [๐‘›], ๐‘› โ‰ฅ 2, which have an odd number of cycles and those which have an even number of cycles. (d) Find a formula for the number of permutations of [๐‘›] having an even number of cycles in two ways: by using Theorem 4.5.2 and by using (4.22). Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Exercises

137

(13) Let ๐‘Ž๐‘› be the number of permutations in ๐”–๐‘› that have an even number of cycles, all of them of odd length. (a) Use a parity argument to show that if ๐‘› is odd, then ๐‘Ž๐‘› = 0. (b) Use egfs to show that if ๐‘› is even, then ๐‘Ž๐‘› = (

๐‘› ๐‘›! ) . ๐‘›/2 2๐‘›

(c) Use part (b) to show that if ๐‘› is even, then the probability that in tossing a fair coin ๐‘› times exactly ๐‘›/2 heads occur is the same as the probability that a permutation chosen uniformly at random from ๐”–๐‘› has an even number of cycles, all of them of odd length. (d) Reprove part (c) by giving, when ๐‘› is even, a bijection between pairs (๐‘†, ๐œ‹) [๐‘›] where ๐‘† โˆˆ (๐‘›/2 ) and ๐œ‹ โˆˆ ๐”–๐‘› and pairs (๐‘‡, ๐œŽ) where ๐‘‡ โˆˆ 2[๐‘›] and ๐œŽ โˆˆ ๐”–๐‘› has an even number of cycles, all of them of odd length. (14) Let ๐‘—๐‘› be the number of involutions in ๐”–๐‘› which have no fixed points. (a) Give a combinatorial proof that ๐‘—2๐‘›+1 = 0 and that ๐‘—2๐‘› = 1 โ‹… 3 โ‹… 5 โ‹ฏ (2๐‘› โˆ’ 1). (b) Use the Exponential Formula to find a simple expression for the exponential ๐‘ฅ๐‘› generating function โˆ‘๐‘›โ‰ฅ0 ๐‘—๐‘› ๐‘›! . (c) Use the exponential generating function from (b) to give a second derivation of the formula in (a).

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

10.1090/gsm/210/05

Chapter 5

Counting with Partially Ordered Sets

Partially ordered sets, known as โ€œposetsโ€ for short, give a fruitful way of ordering combinatorial objects. In this way they provide new perspectives on objects we have already studied, such as interpreting various combinatorial invariants as rank functions. They also give us new and powerful tools to do enumeration such as the Mรถbius Inversion Theorem which generalizes the Principle of Inclusion and Exclusion.

5.1. Basic properties of partially ordered sets A partially ordered set or poset is a pair (๐‘ƒ, โ‰ค) where โ€œ๐‘ƒโ€ is a set and โ€œโ‰คโ€ is a binary relation on ๐‘ƒ satisfying the following axioms for all ๐‘ฅ, ๐‘ฆ, ๐‘ง โˆˆ ๐‘ƒ: (a) (reflexivity) ๐‘ฅ โ‰ค ๐‘ฅ, (b) (antisymmetry) if ๐‘ฅ โ‰ค ๐‘ฆ and ๐‘ฆ โ‰ค ๐‘ฅ, then ๐‘ฅ = ๐‘ฆ, and (c) (transitivity) if ๐‘ฅ โ‰ค ๐‘ฆ and ๐‘ฆ โ‰ค ๐‘ง, then ๐‘ฅ โ‰ค ๐‘ง. Often we will refer to the poset as just ๐‘ƒ since the partial order will be obvious from context. We will also use notation for the standard order on โ„ in the setting of posets in the obvious way. For example ๐‘ฅ < ๐‘ฆ means ๐‘ฅ โ‰ค ๐‘ฆ but ๐‘ฅ โ‰  ๐‘ฆ, or using ๐‘ฆ โ‰ฅ ๐‘ฅ as equivalent to ๐‘ฅ โ‰ค ๐‘ฆ. We say that ๐‘ฅ, ๐‘ฆ โˆˆ ๐‘ƒ are comparable if ๐‘ฅ โ‰ค ๐‘ฆ or ๐‘ฆ โ‰ค ๐‘ฅ. Otherwise they are incomparable. A poset where every pair of elements is comparable is called a total order. There are standard partial orders on many of the combinatorial objects we have already studied and we list some of them here. โ€ข The chain of length ๐‘› is (๐ถ๐‘› , โ‰ค) where ๐ถ๐‘› = {0, 1, . . . , ๐‘›} and ๐‘– โ‰ค ๐‘— is the usual ordering of the integers. So ๐ถ๐‘› is a total order. Note that ๐ถ๐‘› has ๐‘›+1 elements and in some texts this would be referred to as an (๐‘› + 1)-chain. Although 139 Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

140

5. Counting with Partially Ordered Sets

๐ถ๐‘› was also used as the notation for the graphical cycle with ๐‘› vertices, the context should make it clear which object is meant. โ€ข The Boolean algebra is (๐ต๐‘› , โŠ†) where ๐ต๐‘› = 2[๐‘›] and ๐‘† โŠ† ๐‘‡ is set containment. Be sure not to confuse ๐ต๐‘› with the ๐‘›th Bell number ๐ต(๐‘›). โ€ข The divisor lattice is (๐ท๐‘› , โˆฃ) where ๐ท๐‘› consists of all the positive integers which divide evenly into ๐‘› and ๐‘Ž โˆฃ ๐‘ means that ๐‘Ž divides evenly into ๐‘ (in that ๐‘/๐‘Ž is an integer). So in ๐ท12 we have 2 โ‰ค 6 but 2 โ‰ฐ 3. Note the distinction between ๐ท๐‘› and the derangement number ๐ท(๐‘›). โ€ข The lattice of partitions is (ฮ ๐‘› , โ‰ค) where ฮ ๐‘› is the set of all partitions of [๐‘›] and ๐œŒ โ‰ค ๐œ means that every block of ๐œŒ is contained in some block of ๐œ, called the refinement ordering. For example, in ฮ 6 we have 14/2/36/5 โ‰ค 1245/36 because {1, 4}, {2}, and {5} are all contained in {1, 2, 4, 5} and {3, 6} is contained in itself. โ€ข Youngโ€™s lattice is (๐‘Œ , โ‰ค) where ๐‘Œ is the set of all integer partitions and ๐œ† โ‰ค ๐œ‡ is containment of Young diagrams as defined in Section 3.2. โ€ข The lattice of compositions is (๐พ๐‘› , โ‰ค) where ๐พ๐‘› is the set of all compositions of ๐‘› and ๐›ผ โ‰ค ๐›ฝ is refinement of compositions: ๐›ผ can be obtained from ๐›ฝ by replacing each ๐›ฝ ๐‘– by a composition [๐›ผ๐‘— , ๐›ผ๐‘—+1 , . . . , ๐›ผ๐‘˜ ] โŠง ๐›ฝ ๐‘– . For example, in ๐พ11 we have [2, 3, 2, 1, 3] โ‰ค [2, 5, 4] because the first 2 in [2, 5, 4] was replaced by itself, the 5 was replaced by [3, 2], and the 4 by [1, 3]. On the other hand [2, 3, 1, 2, 3] โ‰ฐ [2, 5, 4]. Again, there is a notational overlap with ๐พ๐‘› as the complete graph on ๐‘› vertices, but the two will never appear together. โ€ข The pattern poset is (๐”–, โ‰ค) where ๐”– is the set of all permutations and ๐œ‹ โ‰ค ๐œŽ means that ๐œŽ contains ๐œ‹ as a pattern. โ€ข The subspace lattice is (๐ฟ(๐‘‰), โ‰ค) where ๐ฟ(๐‘‰) is the set of subspaces of a finite vector space ๐‘‰ over ๐”ฝ๐‘ž and ๐‘ˆ โ‰ค ๐‘Š means ๐‘ˆ is a subspace of ๐‘Š. If ๐‘‰ = ๐”ฝ๐‘›๐‘ž , then we often denote this poset by ๐ฟ๐‘› (๐‘ž). There are also important partial orders on permutations which reflect their group structure (strong and weak Bruhat order) and on certain subgraphs of a graph (the bond lattice). We will define the latter later when it is needed. Often a poset (๐‘ƒ, โ‰ค) is represented by a certain (di)graph which can be easier to work with than just using the axioms. If ๐‘ฅ, ๐‘ฆ โˆˆ ๐‘ƒ, then we say that ๐‘ฅ is covered by ๐‘ฆ or ๐‘ฆ covers ๐‘ฅ, written either ๐‘ฅ โ‹– ๐‘ฆ or ๐‘ฆ โ‹— ๐‘ฅ, if ๐‘ฅ < ๐‘ฆ and there is no ๐‘ง โˆˆ ๐‘ƒ with ๐‘ฅ < ๐‘ง < ๐‘ฆ. The Hasse diagram of ๐‘ƒ is the graph with vertices ๐‘ƒ and an edge from ๐‘ฅ up to ๐‘ฆ if ๐‘ฅ โ‹– ๐‘ฆ. Note that this is actually a digraph where all the arcs are directed up and so are just written as edges with this understanding. Hasse diagrams for examples from the above list are given in Figure 5.1. For those which are infinite, only the bottom of the poset is displayed. Note that in the case of the subspace lattice ๐‘‰ = ๐”ฝ23 the subspaces are listed using their row-echelon forms. We will make no distinction between a poset and its Hasse diagram if no confusion will result by blurring the distinction. Finally, certain posets such as the real numbers under their normal ordering have no covers. So in such cases it does not make sense to try to draw a Hasse diagram. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

5.1. Basic properties of partially ordered sets

141

{1, 2, 3}

3 2

{1, 2}

{1, 3}

1

{1}

{2}

{2, 3} {3}

0 โˆ…

๐ถ3

๐ต3 123

18 9

6

12/3

13/2

1/23

3

2 1

1/2/3

๐ท18

ฮ 3

โ‹ฎ

โ‹ฎ [4]

โ‹ฎ

123 132 213 12

[3, 1]

[2, 2]

[2, 1, 1]

[1, 2, 1]

[1, 3] [1, 1, 2]

โˆ…

[1, 1, 1, 1]

๐‘Œ

๐พ4

21

1 0

[

231 312 321 [ 0 1 ]

0 ] 1

[ 1 0 ]

[ 1 1 ]

1 โˆ…

{(0, 0)}

๐”–

๐ฟ2 (3) Figure 5.1. A zoo of Hasse diagrams

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

[ 1 2 ]

142

5. Counting with Partially Ordered Sets

๐‘

๐‘‘

๐‘Ž

๐‘

Figure 5.2. Minimal and maximal elements

There are certain parts of a poset ๐‘ƒ to which we will often refer. A minimal element of ๐‘ƒ is ๐‘ฅ โˆˆ ๐‘ƒ such that there is no ๐‘ฆ โˆˆ ๐‘ƒ with ๐‘ฆ < ๐‘ฅ. Note that ๐‘ƒ can have multiple minimal elements. The poset in Figure 5.2 has minimal elements ๐‘Ž, ๐‘. Dually, define a maximal element of ๐‘ƒ to be ๐‘ฅ โˆˆ ๐‘ƒ with no ๐‘ฆ > ๐‘ฅ in ๐‘ƒ. The example poset just cited has maximal elements ๐‘, ๐‘‘. By way of contrast, ๐‘ƒ has a minimum element if there is ๐‘ฅ โˆˆ ๐‘ƒ such that ๐‘ฅ โ‰ค ๐‘ฆ for every ๐‘ฆ โˆˆ ๐‘ƒ. A minimum element is unique if it exists because if ๐‘ฅ and ๐‘ฅโ€ฒ are both minimum elements, then ๐‘ฅ โ‰ค ๐‘ฅโ€ฒ and ๐‘ฅโ€ฒ โ‰ค ๐‘ฅ which forces ๐‘ฅ = ๐‘ฅโ€ฒ by antisymmetry. In this case the minimum element is often denoted 0.ฬ‚ All of the posets in Figure 5.1 have a 0.ฬ‚ In fact, in ๐ท๐‘› we have 0ฬ‚ = 1, a rare instance where one can write that zero equals one and be mathematically correct! Again, there is the dual notion of a maximum element ๐‘ฅ โˆˆ ๐‘ƒ which satisfies ๐‘ฅ โ‰ฅ ๐‘ฆ for all ๐‘ฆ โˆˆ ๐‘ƒ. A maximum is unique if it exists and is denoted 1.ฬ‚ The following result sums up the existence of minimum and maximum elements for the posets in Figure 5.1. Its proof is sufficiently easy that it is left as an exercise. Proposition 5.1.1. We have the following minimum and maximum elements. โ€ข In ๐ถ๐‘› we have 0ฬ‚ = 0, 1ฬ‚ = ๐‘›. โ€ข In ๐ต๐‘› we have 0ฬ‚ = โˆ…, 1ฬ‚ = [๐‘›]. โ€ข In ๐ท๐‘› we have 0ฬ‚ = 1, 1ฬ‚ = ๐‘›. โ€ข In ฮ ๐‘› we have 0ฬ‚ = 1/2/ . . . /๐‘›, 1ฬ‚ = [๐‘›]. โ€ข In ๐‘Œ we have 0ฬ‚ = โˆ… and no 1.ฬ‚ โ€ข In ๐พ๐‘› we have 0ฬ‚ = [1๐‘› ], 1ฬ‚ = [๐‘›]. โ€ข In ๐”– we have 0ฬ‚ = โˆ… and no 1.ฬ‚ โ€ข In ๐ฟ(๐‘‰) we have 0ฬ‚ is the zero subspace, 1ฬ‚ = ๐‘‰.

โ–ก

As one can tell from the previous set of definitions, it is sometimes useful to reverse inequalities in a poset. So if ๐‘ƒ is a poset, then we define its dual ๐‘ƒโˆ— to have the same underlying set with ๐‘ฅ โ‰ค ๐‘ฆ in ๐‘ƒโˆ— if and only if ๐‘ฆ โ‰ค ๐‘ฅ in ๐‘ƒ. The Hasse diagram of ๐‘ƒ โˆ— is thus obtained by reflecting the one for ๐‘ƒ in a horizontal axis. As is often the case in mathematics, we analyze a structure by looking at its substructures. In posets ๐‘ƒ, these come in several varieties. A subposet of ๐‘ƒ is a subset ๐‘„ โŠ† ๐‘ƒ with the inherited partial order; namely ๐‘ฅ โ‰ค ๐‘ฆ for ๐‘ฅ, ๐‘ฆ โˆˆ ๐‘„ if and only if ๐‘ฅ โ‰ค ๐‘ฆ Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

5.1. Basic properties of partially ordered sets

143

{1, 2, 3, 5, 6}

{1, 2, 3, 5}

{1, 2, 3, 6}

{1, 3, 5, 6}

[{1, 3}, {1, 2, 3, 5, 6}] = {1, 2, 3}

{1, 3, 5}

{1, 3, 6}

{1, 3} Figure 5.3. An interval in ๐ต7

in ๐‘ƒ. In such a case we will sometimes use a subscript to make precise which poset is being considered, as in ๐‘ฅ โ‰ค๐‘ƒ ๐‘ฆ. Note that some authors call this an induced subposet and use the term โ€œsubposetโ€ when ๐‘„ satisfies the weaker condition that ๐‘ฅ โ‰ค๐‘„ ๐‘ฆ implies ๐‘ฅ โ‰ค๐‘ƒ ๐‘ฆ (but not necessarily conversely). There are several especially important subposets. Assume ๐‘ฅ, ๐‘ฆ โˆˆ ๐‘ƒ. Then the corresponding closed interval is [๐‘ฅ, ๐‘ฆ] = {๐‘ง โˆˆ ๐‘ƒ โˆฃ ๐‘ฅ โ‰ค ๐‘ง โ‰ค ๐‘ฆ}. Note that [๐‘ฅ, ๐‘ฆ] = โˆ… unless ๐‘ฅ โ‰ค ๐‘ฆ. For example, the Hasse diagram of the interval [{1, 3}, {1, 2, 3, 5, 6}] in ๐ต7 is displayed in Figure 5.3. Note that if one removes the labels, then this diagram is exactly like the one for ๐ต3 in Figure 5.1. We will explain this formally when we introduce the concept of isomorphism below. Open and half-open intervals in a poset are defined as expected. A subset ๐ผ โŠ† ๐‘ƒ is a lower-order ideal if ๐‘ฅ โˆˆ ๐ผ and ๐‘ฆ โ‰ค ๐‘ฅ imply that ๐‘ฆ โˆˆ ๐ผ. For example, if ๐‘ƒ has a 0,ฬ‚ then any interval of the form [0,ฬ‚ ๐‘ฅ] is a lower-order ideal. If ๐‘† โŠ† ๐‘ƒ is any subset, then the lower-order ideal generated by ๐‘† is ๐ผ(๐‘†) = {๐‘ฆ โˆˆ ๐‘ƒ โˆฃ ๐‘ฆ โ‰ค ๐‘ฅ for some ๐‘ฅ โˆˆ ๐‘†}. We often leave out the set braces in ๐‘†, for example, writing ๐ผ(๐‘ฅ, ๐‘ฆ) for the ideal generated by ๐‘† = {๐‘ฅ, ๐‘ฆ} โŠ† ๐‘ƒ. If #๐‘† = 1, then the order ideal is called principal. If ๐‘ƒ has a 0,ฬ‚ then ๐ผ(๐‘ฅ) = [0,ฬ‚ ๐‘ฅ]. In Youngโ€™s lattice, ๐ผ(๐œ†) is just all the partitions contained in ๐œ†. So when we have a rectangle ๐œ† = (๐‘˜๐‘™ ), then ๐ผ(๐œ†) = โ„›(๐‘˜, ๐‘™), the set which came into play when discussing the ๐‘ž-binomial coefficients in Section 3.2. upper-order ideals ๐‘ˆ as well as those generated by a set, ๐‘ˆ(๐‘†), are defined by reversing all the inequalities. We will sometimes abbreviate โ€œlower-order idealโ€ to โ€œorder idealโ€ or even just โ€œideal,โ€ whereas for upper-order ideals both adjectives will always be used. Some simple properties of ideals are given in the next proposition. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

144

5. Counting with Partially Ordered Sets

Proposition 5.1.2. Let ๐‘ƒ be a poset. (a) We have that ๐ผ โŠ† ๐‘ƒ is a lower-order ideal if and only if ๐‘ƒ โˆ’ ๐ผ is an upper-order ideal. (b) If ๐‘ƒ is finite and ๐ผ is a lower-order ideal, then ๐ผ = ๐ผ(๐‘†) where ๐‘† is the set of maximal elements of ๐ผ. (c) If ๐‘ƒ is finite and ๐‘ˆ is an upper-order ideal, then ๐‘ˆ = ๐‘ˆ(๐‘†) where ๐‘† is the set of minimal elements of ๐‘ˆ. Proof. We will prove (b) and leave the other two parts to the reader. We will prove the equality by proving the two corresponding set containments. Suppose ๐‘ฅ โˆˆ ๐ผ and consider the set ๐‘‹ = {๐‘ฆ โˆˆ ๐ผ โˆฃ ๐‘ฆ โ‰ฅ ๐‘ฅ}. This subset of ๐‘ƒ is nonempty since ๐‘ฅ โˆˆ ๐‘‹. So ๐‘‹ has at least one maximal element ๐‘ฆ since ๐‘ƒ is finite. In fact, ๐‘ฆ must be maximal in ๐ผ since, if not, there is ๐‘ง > ๐‘ฆ with ๐‘ง โˆˆ ๐ผ. But then by transitivity ๐‘ง > ๐‘ฅ so that ๐‘ง โˆˆ ๐‘‹. This contradicts the maximality of ๐‘ฆ in ๐‘‹. So ๐‘ฆ โˆˆ ๐‘† and ๐‘ฅ โˆˆ ๐ผ(๐‘†) showing that ๐ผ โŠ† ๐ผ(๐‘†). To show ๐ผ(๐‘†) โŠ† ๐ผ, take ๐‘ฆ โˆˆ ๐ผ(๐‘†). By definition ๐‘ฆ โ‰ค ๐‘ฅ for some ๐‘ฅ โˆˆ ๐‘† and ๐‘† โŠ† ๐ผ. So ๐‘ฆ โ‰ค ๐‘ฅ where ๐‘ฅ โˆˆ ๐ผ which forces ๐‘ฆ โˆˆ ๐ผ by definition of lower-order ideal. โ–ก

To define isomorphism, we need to consider maps on posets. Assume posets ๐‘ƒ, ๐‘„. Then a function ๐‘“ โˆถ ๐‘ƒ โ†’ ๐‘„ is order preserving if ๐‘ฅ โ‰ค๐‘ƒ ๐‘ฆ โŸน ๐‘“(๐‘ฅ) โ‰ค๐‘„ ๐‘“(๐‘ฆ). For example, the map ๐‘“ โˆถ ๐ถ๐‘› โ†’ ๐ต๐‘› by ๐‘“(๐‘–) = [๐‘–] is order preserving because ๐‘– โ‰ค ๐‘— implies that ๐‘“(๐‘–) = [๐‘–] โŠ† [๐‘—] = ๐‘“(๐‘—). We say that ๐‘“ is an isomorphism or that ๐‘ƒ and ๐‘„ are isomorphic, written ๐‘ƒ โ‰… ๐‘„, if ๐‘“ is bijective and both ๐‘“ and ๐‘“โˆ’1 are order preserving. It is important to show that ๐‘“โˆ’1 , and not just ๐‘“, is order preserving. For consider the two posets in Figure 5.4. Define a bijection ๐‘“ โˆถ ๐‘ƒ โ†’ ๐‘„ by ๐‘“(๐‘Ž) = ๐‘Žโ€ฒ and ๐‘“(๐‘) = ๐‘โ€ฒ . Then ๐‘“ is order preserving vacuously since there are no order relations in ๐‘ƒ. But clearly we do not want ๐‘“ to be an isomorphism since ๐‘ƒ and ๐‘„ have different (unlabeled) Hasse diagrams. This is witnessed by the fact that ๐‘“โˆ’1 is not order preserving: we have ๐‘Žโ€ฒ โ‰ค ๐‘โ€ฒ but ๐‘“โˆ’1 (๐‘Žโ€ฒ ) = ๐‘Ž โ‰ฐ ๐‘ = ๐‘“โˆ’1 (๐‘โ€ฒ ). There are a number of isomorphisms involving the posets in Figure 5.1. Some of them are collected in the next result.

๐‘โ€ฒ ๐‘„=

๐‘ƒ= ๐‘Ž

๐‘ ๐‘Žโ€ฒ

Figure 5.4. Two nonisomorphic posets

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

5.2. Chains, antichains, and operations on posets

145

Proposition 5.1.3. We have the following isomorphisms. In all cases we assume the intervals used are nonempty. (a) ๐ถ๐‘›โˆ— โ‰… ๐ถ๐‘› . (b) If ๐‘–, ๐‘— โˆˆ ๐ถ๐‘› , then [๐‘–, ๐‘—] โ‰… ๐ถ๐‘—โˆ’๐‘– . (c) ๐ต๐‘›โˆ— โ‰… ๐ต๐‘› . (d) If ๐‘†, ๐‘‡ โˆˆ ๐ต๐‘› , then [๐‘†, ๐‘‡] โ‰… ๐ต|๐‘‡โˆ’๐‘†| . (e) ๐ท๐‘›โˆ— โ‰… ๐ท๐‘› . (f) If ๐‘™, ๐‘š โˆˆ ๐ท๐‘› , then [๐‘™, ๐‘š] โ‰… ๐ท๐‘š/๐‘™ . (g) If ๐‘› is a product of ๐‘˜ distinct primes, then ๐ท๐‘› โ‰… ๐ต๐‘˜ . (h) If ๐œŒ โˆˆ ฮ ๐‘› has ๐‘˜ blocks, then [๐œŒ, 1]ฬ‚ โ‰… ฮ ๐‘˜ . (i) For all ๐‘› we have ๐พ๐‘› โ‰… ๐ต๐‘›โˆ’1 . (j) If ๐‘‰, ๐‘Š are vector spaces over ๐”ฝ๐‘ž of the same dimension, then ๐ฟ(๐‘‰) โ‰… ๐ฟ(๐‘Š). (k) If ๐‘‹, ๐‘Œ โˆˆ ๐ฟ(๐‘‰), then [๐‘‹, ๐‘Œ ] โ‰… ๐ฟ(๐‘‹/๐‘Œ ) where ๐‘‹/๐‘Œ is the quotient vector space. Proof. We will prove the statement about [๐‘†, ๐‘‡] in ๐ต๐‘› and leave the rest of the isomorphisms as exercises. Let ๐‘‡ โˆ’ ๐‘† = {๐‘ก1 , ๐‘ก2 , . . . , ๐‘ก๐‘› } where ๐‘› = |๐‘‡ โˆ’ ๐‘†|. Define a map ๐‘“ โˆถ [๐‘†, ๐‘‡] โ†’ ๐ต๐‘› as follows. If ๐‘‹ โˆˆ [๐‘†, ๐‘‡], then ๐‘‹ = ๐‘† โŠŽ ๐‘‹ โ€ฒ where ๐‘‹ โ€ฒ โŠ† ๐‘‡ โˆ’ ๐‘†. If ๐‘‹ โ€ฒ = {๐‘ก ๐‘– , . . . , ๐‘ก๐‘— }, then let ๐‘“(๐‘‹) = {๐‘–, . . . , ๐‘—}. This is well-defined since, by definition, ๐‘–, . . . , ๐‘— โˆˆ [๐‘›]. To show that ๐‘“ is a bijection, we construct its inverse. Assume ๐ผ = {๐‘–, . . . , ๐‘—} โˆˆ ๐ต๐‘› . Then let ๐‘“โˆ’1 (๐ผ) = ๐‘† โŠŽ {๐‘ก ๐‘– , . . . , ๐‘ก๐‘— }. The proof that this is well-defined is similar to that of ๐‘“, and the fact that these are inverses is clear from their definitions. Finally, we need to show that ๐‘“ and ๐‘“โˆ’1 are order preserving. If ๐‘‹ โ‰ค ๐‘Œ in [๐‘†, ๐‘‡], then ๐‘‹ โ€ฒ โŠ† ๐‘Œ โ€ฒ in ๐‘‡ โˆ’ ๐‘†. It follows that ๐‘“(๐‘‹) โ‰ค ๐‘“(๐‘Œ ) in ๐ต๐‘› . Thus ๐‘“ is order preserving. As far as ๐‘“โˆ’1 , take ๐ผ โ‰ค ๐ฝ in ๐ต๐‘› . Then the corresponding sets ๐‘‡๐ผ and ๐‘‡๐ฝ in ๐‘‡ โˆ’ ๐‘† gotten by using ๐ผ and ๐ฝ for subscripts satisfy ๐‘‡๐ผ โŠ† ๐‘‡๐ฝ . It follows that ๐‘“โˆ’1 (๐ผ) = ๐‘† โŠŽ ๐‘‡๐ผ โŠ† ๐‘† โŠŽ ๐‘‡๐ฝ = ๐‘“โˆ’1 (๐ฝ) which shows that ๐‘“โˆ’1 is order preserving.

โ–ก

5.2. Chains, antichains, and operations on posets We will now consider three operations for building posets. Chains and the related notion of antichains will play important roles. Given posets (๐‘ƒ, โ‰ค๐‘ƒ ) and (๐‘„, โ‰ค๐‘„ ) with ๐‘ƒ โˆฉ ๐‘„ = โˆ…, their disjoint union is the poset whose elements are ๐‘ƒ โŠŽ ๐‘„ with the partial order ๐‘ฅ โ‰ค๐‘ƒโŠŽ๐‘„ ๐‘ฆ if (a) ๐‘ฅ, ๐‘ฆ โˆˆ ๐‘ƒ and ๐‘ฅ โ‰ค๐‘ƒ ๐‘ฆ or (b) ๐‘ฅ, ๐‘ฆ โˆˆ ๐‘„ and ๐‘ฅ โ‰ค๐‘„ ๐‘ฆ. So one just takes the relations in ๐‘ƒ and ๐‘„ and does not add any new ones. To illustrate, the poset on the left in Figure 5.4 is the disjoint union ๐‘ƒ = {๐‘Ž} โŠŽ {๐‘}. If one takes both posets in this figure, then ๐‘ƒ โŠŽ ๐‘„ is the poset on {๐‘Ž, ๐‘, ๐‘Žโ€ฒ , ๐‘โ€ฒ } with ๐‘Žโ€ฒ < ๐‘โ€ฒ being the only Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

146

5. Counting with Partially Ordered Sets

strict order relation. An important example of disjoint union is the ๐‘›-element antichain ๐ด๐‘› which consists of a set of ๐‘› elements with no strict order relations. So ๐‘ƒ = {๐‘Ž} โŠŽ {๐‘} is a copy of ๐ด2 . Another way to combine disjoint posets is their ordinal sum ๐‘ƒ + ๐‘„ which has elements ๐‘ƒ โŠŽ ๐‘„ and ๐‘ฅ โ‰ค๐‘ƒ+๐‘„ ๐‘ฆ if one of (a), (b), (c) hold where (a) and (b) are as in the previous paragraph and the third possibility is (c) ๐‘ฅ โˆˆ ๐‘ƒ, ๐‘ฆ โˆˆ ๐‘„. Intuitively, one takes the relations in ๐‘ƒ and ๐‘„ and also makes everything in ๐‘ƒ smaller than everything in ๐‘„. As an example using chains, we have ๐ถ๐‘š + ๐ถ๐‘› โ‰… ๐ถ๐‘š+๐‘›+1 . Note that in general ๐‘ƒ + ๐‘„ โ‰‡ ๐‘„ + ๐‘ƒ as the use of the adjective โ€œordinalโ€ suggests. Our third method to produce new posets from old ones is via products. Given two (not necessarily disjoint) posets (๐‘ƒ, โ‰ค๐‘ƒ ) and (๐‘„, โ‰ค๐‘„ ), their (direct or Cartesian) product has underlying set ๐‘ƒ ร— ๐‘„ = {(๐‘ฅ, ๐‘ฆ) โˆฃ ๐‘ฅ โˆˆ ๐‘ƒ, ๐‘ฆ โˆˆ ๐‘„} together with the partial order (๐‘ฅ, ๐‘ฆ) โ‰ค๐‘ƒร—๐‘„ (๐‘ฅโ€ฒ , ๐‘ฆโ€ฒ ) if ๐‘ฅ โ‰ค๐‘ƒ ๐‘ฅโ€ฒ and ๐‘ฆ โ‰ค๐‘„ ๐‘ฆโ€ฒ . We let ๐‘ƒ๐‘› denote the ๐‘›-fold product of ๐‘ƒ. One can obtain the Hasse diagram for ๐‘ƒ ร— ๐‘„ by replacing each vertex of ๐‘„ with a copy of ๐‘ƒ and then, for each edge between two vertices of ๐‘„, connecting each pair of vertices having the same first coordinate in the corresponding two copies of ๐‘ƒ. Illustrations of this can be found in Figure 5.1. For example, ๐ท18 looks like a rectangle because it is isomorphic to ๐ถ1 ร— ๐ถ2 . Also, ๐ต3 โ‰… ๐ถ13 , which is why the Hasse diagram looks like the projection of a 3-dimensional cube into the plane. Both of these isomorphisms generalize, and there is one for ฮ ๐‘› as well. Proposition 5.2.1. We have the following product decompositions. (a) We have ๐ต๐‘› โ‰… ๐ถ1๐‘› . ๐‘›

๐‘›

๐‘›

(b) If the prime factorization of ๐‘› is ๐‘› = ๐‘1 1 ๐‘2 2 โ‹ฏ ๐‘๐‘˜ ๐‘˜ , then ๐ท๐‘› โ‰… ๐ถ๐‘›1 ร— ๐ถ๐‘›2 ร— โ‹ฏ ร— ๐ถ๐‘›๐‘˜ . (c) If ๐œŒ โ‰ค ๐œ in ฮ ๐‘› , then [๐œŒ, ๐œ] โ‰… ฮ ๐‘›1 ร— ฮ ๐‘›2 ร— โ‹ฏ ร— ฮ ๐‘›๐‘˜ where ๐œ = ๐‘‡1 /๐‘‡2 / . . . /๐‘‡๐‘˜ and ๐‘›๐‘– is the number of blocks of ๐œŒ contained in ๐‘‡๐‘– for all ๐‘–. Proof. (a) Consider the map ๐‘“ used in the proof of Theorem 1.3.1. Stated in the language of posets, we see that ๐‘“ โˆถ ๐ต๐‘› โ†’ ๐ถ1๐‘› . And we have already shown that ๐‘“ is bijective. So we only need to prove that ๐‘“ and ๐‘“โˆ’1 are order preserving. Suppose ๐‘†, ๐‘‡ โŠ† [๐‘›] with ๐‘“(๐‘†) = (๐‘ฃ 1 , . . . , ๐‘ฃ ๐‘› ) and ๐‘“(๐‘‡) = (๐‘ค 1 , . . . , ๐‘ค ๐‘› ). Now ๐‘† โŠ† ๐‘‡ if and only if ๐‘– โˆˆ ๐‘† implies ๐‘– โˆˆ ๐‘‡. By the definition of ๐‘“, this is equivalent to ๐‘ฃ ๐‘– = 1 implying ๐‘ค ๐‘– = 1. But this means that ๐‘ฃ ๐‘– โ‰ค ๐‘ค ๐‘– for all ๐‘– since if ๐‘ฃ ๐‘– = 0, then ๐‘ฃ ๐‘– โ‰ค ๐‘ค ๐‘– is automatic. Thus we have shown that ๐‘† โ‰ค ๐‘‡ in ๐ต๐‘› if and only if (๐‘ฃ 1 , . . . , ๐‘ฃ ๐‘› ) โ‰ค (๐‘ค 1 , . . . , ๐‘ค ๐‘› ) in ๐ถ1๐‘› which is what we wished to prove. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

5.2. Chains, antichains, and operations on posets

147

(b) The map for ๐ท๐‘› is similar. Define ๐‘” โˆถ ๐ท๐‘› โ†’ โจ‰๐‘– ๐ถ๐‘›๐‘– by ๐‘”(๐‘‘) = (๐‘‘1 , ๐‘‘2 , . . . , ๐‘‘๐‘˜ ) ๐‘‘

where ๐‘‘ = โˆ๐‘– ๐‘๐‘– ๐‘– . The reader can now verify that this is a well-defined isomorphism of posets. (c) The construction of the isomorphism is messy but not conceptually difficult. Consider blocks of ๐œŒ as being single elements and then aggregate all those lying in a given block of ๐œ together. Again, the reader can fill in the details. โ–ก So chains can help us understand various posets by taking products. There is another important way in which chains can be used to decompose certain posets. Let ๐‘ƒ be a poset and ๐ถ โˆถ ๐‘ฅ0 < ๐‘ฅ1 < โ‹ฏ < ๐‘ฅ๐‘› be a chain in ๐‘ƒ. We say that ๐ถ is a chain of length ๐‘› from ๐‘ฅ0 to ๐‘ฅ๐‘› and also use the term ๐‘ฅ0 โ€“๐‘ฅ๐‘› chain. We let โ„“(๐ถ) denote the length of ๐ถ. Call ๐ถ maximal if it is not strictly contained in a larger chain of ๐‘ƒ. If [๐‘ฅ0 , ๐‘ฅ๐‘› ] is finite, this is equivalent to ๐‘ฅ๐‘– < ๐‘ฅ๐‘–+1 being a cover for all 0 โ‰ค ๐‘– < ๐‘›. The chain ๐ถ is saturated if it is not strictly contained in a larger chain from ๐‘ฅ0 to ๐‘ฅ๐‘› . Equivalently, ๐ถ is saturated if it is maximal in [๐‘ฅ0 , ๐‘ฅ๐‘› ]. For example, in ๐ท18 the chain 3 < 6 < 18 is a chain of length 2 from 3 to 18 which is saturated since each inequality is a cover. But this chain is not maximal since it is contained in the larger chain 1 < 3 < 6 < 18. Some posets can be written as a disjoint union of certain subposets called ranks as follows. Suppose ๐‘ƒ is a poset which is locally finite in that the cardinality of any interval [๐‘ฅ, ๐‘ฆ] of ๐‘ƒ is finite. All the posets in Figure 5.1 are locally finite even though ๐‘Œ and ๐”– are not finite. The poset of real numbers with its usual total order is not locally finite. Let ๐‘ƒ be locally finite and have a 0.ฬ‚ Then ๐‘ƒ is ranked if for any ๐‘ฅ โˆˆ ๐‘ƒ all saturated chains from 0ฬ‚ to ๐‘ฅ have the same length. In this case we call this common length the rank of ๐‘ฅ and it is denoted rk ๐‘ฅ or rk๐‘ƒ ๐‘ฅ if we wish to be specific about the poset. For ๐‘˜ โˆˆ โ„•, the ๐‘˜th rank set of ๐‘ƒ is (5.1)

Rk๐‘˜ ๐‘ƒ = {๐‘ฅ โˆˆ ๐‘ƒ โˆฃ rk ๐‘ฅ = ๐‘˜}.

If ๐‘ƒ is finite, then we define its rank to be rk ๐‘ƒ = max{๐‘˜ โˆฃ Rk๐‘˜ ๐‘ƒ โ‰  โˆ…}. All the posets in Figure 5.1 are ranked and we will describe their rank sets shortly. An example of a poset which is not ranked is in Figure 5.5. This is because there are two ฬ‚ 1ฬ‚ chains, namely 0ฬ‚ < ๐‘ < 1ฬ‚ which is of length 2 and 0ฬ‚ < ๐‘Ž < ๐‘ < 1ฬ‚ which saturated 0โ€“ is of length 3. We will now list the rank information for the posets in Figure 5.1. Note how many of the combinatorial concepts which were introduced in earlier chapters occur naturally in this context. These results are easily proved, so the demonstrations can be filled in by the reader. Proposition 5.2.2. All the posets in Figure 5.1 are ranked with the following rank functions. (a) If ๐‘˜ โˆˆ ๐ถ๐‘› , then rk(๐‘˜) = ๐‘˜, so Rk๐‘˜ (๐ถ๐‘› ) = {๐‘˜}. Also, rk(๐ถ๐‘› ) = ๐‘›. (b) If ๐‘† โˆˆ ๐ต๐‘› , then rk(๐‘†) = #๐‘†, so Rk๐‘˜ (๐ต๐‘› ) = ([๐‘›] ). Also, rk(๐ต๐‘› ) = ๐‘›. ๐‘˜ Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

148

5. Counting with Partially Ordered Sets

1ฬ‚

๐‘ ๐‘ ๐‘Ž

0ฬ‚ Figure 5.5. An unranked poset ๐‘‘

๐‘‘

(c) If ๐‘‘ โˆˆ ๐ท๐‘› has prime factorization ๐‘‘ = ๐‘1 1 โ‹ฏ ๐‘๐‘Ÿ ๐‘Ÿ , then rk(๐‘‘) = ๐‘‘1 + โ‹ฏ + ๐‘‘๐‘Ÿ . So Rk๐‘˜ (๐ท๐‘› ) is the set of ๐‘‘ โˆฃ ๐‘› with ๐‘˜ primes in their prime factorization, counted with multiplicity. Also, rk(๐ท๐‘› ) is the total number of primes dividing ๐‘›, counted with multiplicity. (d) If ๐œŒ โˆˆ ฮ ๐‘› has ๐‘ blocks, then rk(๐œŒ) = ๐‘› โˆ’ ๐‘, so Rk๐‘˜ (ฮ ๐‘› ) = ๐‘†([๐‘›], ๐‘› โˆ’ ๐‘˜). Also, rk(ฮ ๐‘› ) = ๐‘› โˆ’ 1. (e) If ๐œ† โˆˆ ๐‘Œ , then rk(๐œ†) = |๐œ†|, so Rk๐‘˜ (๐‘Œ ) = ๐‘ƒ(๐‘˜). (f) If ๐›ผ โˆˆ ๐พ๐‘› has ๐‘ parts, then rk(๐›ผ) = ๐‘› โˆ’ ๐‘, so Rk๐‘˜ (๐พ๐‘› ) = ๐‘„(๐‘›, ๐‘› โˆ’ ๐‘˜). Also, rk(๐พ๐‘› ) = ๐‘› โˆ’ 1. (g) If ๐œ‹ โˆˆ ๐”–, then rk(๐œ‹) = |๐œ‹|, so Rk๐‘˜ (๐”–) = ๐”–๐‘˜ . (h) If ๐‘Š โˆˆ ๐ฟ(๐‘‰), then rk(๐‘Š) = dim ๐‘Š, so Rk๐‘˜ (๐‘‰) = [๐‘‰๐‘˜ ]. Also, rk(๐ฟ(๐‘‰)) = dim ๐‘‰. โ–ก

5.3. Lattices The reader will have noticed that several of our example posets are called โ€œlatticesโ€. This is an important class of partially ordered sets with the property that pairs of elements have greatest lower bounds and least upper bounds. It is also common to study lattices whose elements satisfy certain identities using these two operations. In this section we will prove a theorem characterizing the lattices satisfying a distributive law. If ๐‘ƒ is a poset and ๐‘ฅ, ๐‘ฆ โˆˆ ๐‘ƒ, then a lower bound for ๐‘ฅ, ๐‘ฆ is a ๐‘ง โˆˆ ๐‘ƒ such that ๐‘ง โ‰ค ๐‘ฅ and ๐‘ง โ‰ค ๐‘ฆ. For example, if ๐‘†, ๐‘‡ โˆˆ ๐ต๐‘› , then any set contained in both ๐‘† and ๐‘‡ is a lower bound. We say that ๐‘ฅ, ๐‘ฆ have a greatest lower bound or meet if there is an element in ๐‘ƒ, denoted ๐‘ฅ โˆง ๐‘ฆ, which is a lower bound for ๐‘ฅ, ๐‘ฆ, and ๐‘ฅ โˆง ๐‘ฆ โ‰ฅ ๐‘ง for all lower bounds ๐‘ง of ๐‘ฅ and ๐‘ฆ. Returning to ๐ต๐‘› , we have ๐‘† โˆง ๐‘‡ = ๐‘† โˆฉ ๐‘‡. In fact, one can remember the notation for meet as just a squared-off intersection symbol. Note that if the meet of ๐‘ฅ, ๐‘ฆ exists, then it is unique. Indeed, if ๐‘ง, ๐‘งโ€ฒ are both greatest lower bounds of ๐‘ฅ, ๐‘ฆ, then we have ๐‘ง โ‰ฅ ๐‘งโ€ฒ since ๐‘งโ€ฒ is a lower bound and ๐‘ง is the greatest lower bound. But interchanging the roles of ๐‘ง and ๐‘งโ€ฒ also gives ๐‘งโ€ฒ โ‰ฅ ๐‘ง. So ๐‘ง = ๐‘งโ€ฒ by antisymmetry. Note also that it is possible for the meet not to exist. For example, in the poset of Figure 5.2, ๐‘Ž โˆง ๐‘ does Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

5.3. Lattices

149

not exist because this pair has no lower bound. Also, ๐‘ โˆง ๐‘‘ does not exist but this is because the pair has both ๐‘Ž and ๐‘ as lower bounds, but there is no lower bound larger than both ๐‘Ž and ๐‘. One can extend these definitions in the obvious way from pairs of elements to any nonempty set of elements ๐‘‹ = {๐‘ฅ1 , . . . , ๐‘ฅ๐‘› } โŠ† ๐‘ƒ. In this case the meet is denoted ๐‘ฅ. ๐‘‹= โ‹€ โ‹€ ๐‘ฅโˆˆ๐‘‹

One can also reasonably define the meet of the empty set as long as ๐‘ƒ has a 1.ฬ‚ Indeed, for any ๐‘ฅ โˆˆ ๐‘ƒ we would want ๐‘ฅโˆง(

โ‹€

โˆ…) =

({๐‘ฅ} โˆช โˆ…) = {๐‘ฅ} = ๐‘ฅ. โ‹€ โ‹€

But the only element ๐‘ฆ of ๐‘ƒ such that ๐‘ฅ โˆง ๐‘ฆ = ๐‘ฅ for all ๐‘ฅ is ๐‘ฆ = 1.ฬ‚ So we let โˆงโˆ… = 1.ฬ‚ The concepts of upper bound and least upper bound are obtained by reversing the inequalities in the definitions of the previous paragraph. If the least upper bound of ๐‘ฅ, ๐‘ฆ exists, then it is denoted ๐‘ฅ โˆจ ๐‘ฆ and is also called their join. A lattice is a poset such that every pair of elements has a meet and a join. Note that this is different from the use of the term โ€œlatticeโ€ as in โ€œlattice pathโ€ which in this case refers to the discrete subgroup โ„ค2 of โ„2 . The context should make it clear which meaning is intended. All the poset families in Figure 5.1 are lattices except for the pattern poset ๐”– which has subposets isomorphic to Figure 5.2 between its second and third ranks. To describe the meets and joins we need the following terminology. If ๐‘, ๐‘‘ โˆˆ โ„™, then gcd(๐‘, ๐‘‘) and lcm(๐‘, ๐‘‘) denote their greatest common divisor and least common multiple, respectively. Given two Young diagrams ๐œ†, ๐œ‡, we then take their intersection ๐œ† โˆฉ ๐œ‡ or union ๐œ† โˆช ๐œ‡ by aligning them as in Figure 3.1 and then taking the intersection or union of their sets of squares, respectively. If ๐‘ˆ, ๐‘Š are vector subspaces of ๐‘‰, then their sum is ๐‘ˆ + ๐‘Š = {๐‘ข + ๐‘ค โˆฃ ๐‘ข โˆˆ ๐‘ˆ, ๐‘ค โˆˆ ๐‘Š}. We leave the verification of the next result to the reader. Proposition 5.3.1. The posets ๐ถ๐‘› , ๐ต๐‘› , ๐ท๐‘› , ฮ ๐‘› , ๐‘Œ , ๐พ๐‘› , and ๐ฟ(๐‘‰) are all lattices for all ๐‘› and ๐‘‰ of finite dimension over some ๐”ฝ๐‘ž . In addition, we have the following descriptions of their meets and joins. (a) If ๐‘–, ๐‘— โˆˆ ๐ถ๐‘› , then ๐‘– โˆง ๐‘— = min{๐‘–, ๐‘—} and ๐‘– โˆจ ๐‘— = max{๐‘–, ๐‘—}. (b) If ๐‘†, ๐‘‡ โˆˆ ๐ต๐‘› , then ๐‘† โˆง ๐‘‡ = ๐‘† โˆฉ ๐‘‡ and ๐‘† โˆจ ๐‘‡ = ๐‘† โˆช ๐‘‡. (c) If ๐‘, ๐‘‘ โˆˆ ๐ท๐‘› , then ๐‘ โˆง ๐‘‘ = gcd(๐‘, ๐‘‘) and ๐‘ โˆจ ๐‘‘ = lcm(๐‘, ๐‘‘). (d) Suppose ๐œŒ, ๐œ โˆˆ ฮ ๐‘› . Then ๐œŒ โˆง ๐œ is the partition whose blocks are the nonempty intersections of the form ๐ต โˆฉ๐ถ for blocks ๐ต โˆˆ ๐œŒ, ๐ถ โˆˆ ๐œ. Also, ๐œŒโˆจ๐œ is the partition such that ๐‘, ๐‘ are in the same block of the join if and only if there is a sequence of blocks ๐ท1 , . . . , ๐ท๐‘š where each ๐ท๐‘– is a block of either ๐œŒ or ๐œ, ๐‘ โˆˆ ๐ท1 , ๐‘ โˆˆ ๐ท๐‘š , and ๐ท๐‘– โˆฉ ๐ท๐‘–+1 โ‰  โˆ… for all ๐‘–. (e) If ๐œ†, ๐œ‡ โˆˆ ๐‘Œ , then ๐œ† โˆง ๐œ‡ = ๐œ† โˆฉ ๐œ‡ and ๐œ† โˆจ ๐œ‡ = ๐œ† โˆช ๐œ‡. (f) If ๐‘ˆ, ๐‘Š โˆˆ ๐ฟ(๐‘‰), then ๐‘ˆ โˆง ๐‘Š = ๐‘ˆ โˆฉ ๐‘Š and ๐‘ˆ โˆจ ๐‘Š = ๐‘ˆ + ๐‘Š. Next we will give a list of some elementary properties of lattices. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

โ–ก

150

5. Counting with Partially Ordered Sets

Proposition 5.3.2. Let ๐ฟ be a lattice. Then the following are true for all ๐‘ฅ, ๐‘ฆ, ๐‘ง โˆˆ ๐ฟ. (a) (Idempotent law) ๐‘ฅ โˆง ๐‘ฅ = ๐‘ฅ โˆจ ๐‘ฅ = ๐‘ฅ. (b) (Commutative law) ๐‘ฅ โˆง ๐‘ฆ = ๐‘ฆ โˆง ๐‘ฅ and ๐‘ฅ โˆจ ๐‘ฆ = ๐‘ฆ โˆจ ๐‘ฅ. (c) (Associative law) (๐‘ฅ โˆง ๐‘ฆ) โˆง ๐‘ง = ๐‘ฅ โˆง (๐‘ฆ โˆง ๐‘ง) and (๐‘ฅ โˆจ ๐‘ฆ) โˆจ ๐‘ง = ๐‘ฅ โˆจ (๐‘ฆ โˆจ ๐‘ง). (d) (Absorption law) ๐‘ฅ โˆง (๐‘ฅ โˆจ ๐‘ฆ) = ๐‘ฅ = ๐‘ฅ โˆจ (๐‘ฅ โˆง ๐‘ฆ). (e) ๐‘ฅ โ‰ค ๐‘ฆ โŸบ ๐‘ฅ โˆง ๐‘ฆ = ๐‘ฅ โŸบ ๐‘ฅ โˆจ ๐‘ฆ = ๐‘ฆ. (f) If ๐‘ฅ โ‰ค ๐‘ฆ, then ๐‘ฅ โˆง ๐‘ง โ‰ค ๐‘ฆ โˆง ๐‘ง and ๐‘ฅ โˆจ ๐‘ง โ‰ค ๐‘ฆ โˆจ ๐‘ง. (g) ๐‘ฅ โˆง (๐‘ฆ โˆจ ๐‘ง) โ‰ฅ (๐‘ฅ โˆง ๐‘ฆ) โˆจ (๐‘ฅ โˆง ๐‘ง) and ๐‘ฅ โˆจ (๐‘ฆ โˆง ๐‘ง) โ‰ค (๐‘ฅ โˆจ ๐‘ฆ) โˆง (๐‘ฅ โˆจ ๐‘ง). (h) If ๐‘‹ is a finite, nonempty subset of ๐ฟ, then โ‹€ ๐‘‹ and โ‹ ๐‘‹ exist. (i) If ๐ฟ is finite, then ๐ฟ has a 0ฬ‚ and a 1.ฬ‚ (j) The dual ๐ฟโˆ— is a lattice. (k) If ๐‘€ is a poset and there is an isomorphism ๐‘“ โˆถ ๐ฟ โ†’ ๐‘€, then ๐‘€ is also a lattice. Furthermore, for ๐‘ฅ, ๐‘ฆ โˆˆ ๐ฟ, ๐‘“(๐‘ฅ โˆง ๐‘ฆ) = ๐‘“(๐‘ฅ) โˆง ๐‘“(๐‘ฆ)

and

๐‘“(๐‘ฅ โˆจ ๐‘ฆ) = ๐‘“(๐‘ฅ) โˆจ ๐‘“(๐‘ฆ).

Proof. The proofs of these results are straightforward. So we will give a demonstration for the first inequality in (g) and leave the rest to the reader. By definition of lower bound, ๐‘ฅ โˆง ๐‘ฆ โ‰ค ๐‘ฅ. And similarly ๐‘ฅ โˆง ๐‘ฆ โ‰ค ๐‘ฆ โ‰ค ๐‘ฆ โˆจ ๐‘ง. So by definition of greatest lower bound, ๐‘ฅโˆง๐‘ฆ โ‰ค ๐‘ฅโˆง(๐‘ฆโˆจ๐‘ง). Similar reason gives ๐‘ฅโˆง๐‘ง โ‰ค ๐‘ฅโˆง(๐‘ฆโˆจ๐‘ง). Using the previous two inequalities and the definition of least upper bound yields (๐‘ฅ โˆง ๐‘ฆ) โˆจ (๐‘ฅ โˆง ๐‘ง) โ‰ค ๐‘ฅ โˆง (๐‘ฆ โˆจ ๐‘ง) which is what we wished to prove. โ–ก

It is certainly possible for the inequalities in (g) of the previous proposition to be strict. For example, using the lattice in Figure 5.5 we have ๐‘ โˆง (๐‘Ž โˆจ ๐‘) = ๐‘ โˆง 1ฬ‚ = ๐‘ while (๐‘ โˆง ๐‘Ž) โˆจ (๐‘ โˆง ๐‘) = ๐‘Ž โˆจ 0ฬ‚ = ๐‘Ž. However, when we have equality in one of these two inequalities, it is forced in the other. Proposition 5.3.3. Let ๐ฟ be a lattice. Then ๐‘ฅ โˆง (๐‘ฆ โˆจ ๐‘ง) = (๐‘ฅ โˆง ๐‘ฆ) โˆจ (๐‘ฅ โˆง ๐‘ง) for all ๐‘ฅ, ๐‘ฆ, ๐‘ง โˆˆ ๐ฟ if and only if ๐‘ฅ โˆจ (๐‘ฆ โˆง ๐‘ง) = (๐‘ฅ โˆจ ๐‘ฆ) โˆง (๐‘ฅ โˆจ ๐‘ง) for all ๐‘ฅ, ๐‘ฆ, ๐‘ง โˆˆ ๐ฟ. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

5.3. Lattices

151

Proof. For the forward direction we have (๐‘ฅ โˆจ ๐‘ฆ) โˆง (๐‘ฅ โˆจ ๐‘ง) = [(๐‘ฅ โˆจ ๐‘ฆ) โˆง ๐‘ฅ] โˆจ [(๐‘ฅ โˆจ ๐‘ฆ) โˆง ๐‘ง] = ๐‘ฅ โˆจ [(๐‘ฅ โˆจ ๐‘ฆ) โˆง ๐‘ง] = ๐‘ฅ โˆจ [(๐‘ฅ โˆง ๐‘ง) โˆจ (๐‘ฆ โˆง ๐‘ง)] = [๐‘ฅ โˆจ (๐‘ฅ โˆง ๐‘ง)] โˆจ (๐‘ฆ โˆง ๐‘ง) = ๐‘ฅ โˆจ (๐‘ฆ โˆง ๐‘ง), โ–ก

which is the desired equality. The proof of the other direction is similar.

A lattice which satisfies either of the two equalities in Proposition 5.3.3 is called a distributive lattice, and these equations are called the distributive laws. Of the lattices in Proposition 5.3.1, ฮ ๐‘› is not distributive for ๐‘› โ‰ฅ 3. In fact, taking ๐‘ง, ๐‘ฆ, ๐‘ง to be the three elements of rank 1 in ฮ 3 gives a strict inequality in the defining relation. Similarly, ๐ฟ(๐‘‰) is not distributive for dim ๐‘‰ โ‰ฅ 2. All the other lattices are distributive as the reader will be asked to show in the exercises. Proposition 5.3.4. The posets ๐ถ๐‘› , ๐ต๐‘› , ๐ท๐‘› , ๐‘Œ , ๐พ๐‘› are distributive lattices for all ๐‘›.

โ–ก

We will now prove a beautiful theorem characterizing finite distributive lattices due to Birkhoff [14]. It says that every such lattice is essentially a set of lower-order ideals partially ordered by inclusion. Given a poset ๐‘ƒ, consider ๐’ฅ(๐‘ƒ) = {๐ผ โˆฃ ๐ผ is a lower-order ideal of ๐‘ƒ}. {๐‘Ÿ, ๐‘ , ๐‘ก, ๐‘ข} ๐‘ข {๐‘Ÿ, ๐‘ , ๐‘ก} ๐‘ 

{๐‘Ÿ, ๐‘ก, ๐‘ข}

๐‘ก {๐‘Ÿ, ๐‘ } ๐‘Ÿ

{๐‘Ÿ, ๐‘ก} {๐‘Ÿ}

๐‘ƒ โˆ… ๐’ฅ(๐‘ƒ) Figure 5.6. A poset and its associated distributive lattice

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

152

5. Counting with Partially Ordered Sets

Turn ๐’ฅ(๐‘ƒ) into a poset by letting ๐ผ โ‰ค ๐ฝ if ๐ผ โŠ† ๐ฝ. For example, a poset ๐‘ƒ and the corresponding poset ๐’ฅ(๐‘ƒ) are shown in Figure 5.6. Our first order of business is to show that ๐’ฅ(๐‘ƒ) is always a distributive lattice. Proposition 5.3.5. If ๐‘ƒ is any poset, then ๐’ฅ(๐‘ƒ) is a distributive lattice. Proof. It suffices to show that if ๐ผ, ๐ฝ โˆˆ ๐’ฅ(๐‘ƒ), then so are ๐ผ โˆฉ ๐ฝ and ๐ผ โˆช ๐ฝ. Indeed, this suffices because these are the greatest lower bound and least upper bound if one considers all subsets of ๐‘ƒ, and set intersection and union satisfy the distributive laws. We will demonstrate this for ๐ผ โˆฉ ๐ฝ since the argument for union is similar. We need to show that ๐ผ โˆฉ ๐ฝ is a lower-order ideal of ๐‘ƒ. So take ๐‘ฅ โˆˆ ๐ผ โˆฉ ๐ฝ and ๐‘ฆ โ‰ค ๐‘ฅ. Now ๐‘ฅ โˆˆ ๐ผ and ๐‘ฅ โˆˆ ๐ฝ. Since both sets are ideals, this implies ๐‘ฆ โˆˆ ๐ผ and ๐‘ฆ โˆˆ ๐ฝ. So ๐‘ฆ โˆˆ ๐ผ โˆฉ ๐ฝ which is what we needed to show. โ–ก The amazing thing is that every finite distributive lattice is of the form ๐’ฅ(๐‘ƒ) for some poset ๐‘ƒ. In order to prove this, we need a way that, given a distributive lattice ๐ฟ, we can identify the ๐‘ƒ from which it was built. This is done using certain elements of ๐‘ƒ which we now define. Let ๐ฟ be a finite lattice so that ๐ฟ has a 0ฬ‚ by Proposition 5.3.2(i). Call ๐‘ฅ โˆˆ ๐ฟ โˆ’ {0}ฬ‚ join irreducible if ๐‘ฅ cannot be written as ๐‘ฅ = ๐‘ฆ โˆจ ๐‘ง where ๐‘ฆ, ๐‘ง < ๐‘ฅ. Equivalently, if ๐‘ฅ = ๐‘ฆ โˆจ ๐‘ง, then ๐‘ฆ = ๐‘ฅ or ๐‘ง = ๐‘ฅ. Let Irr(๐ฟ) = {๐‘ฅ โˆˆ ๐ฟ โˆฃ ๐‘ฅ is join irreducible}. It turns out the join irreducibles are easy to spot in the Hasse diagram of ๐ฟ. Also, the join irreducibles under a given element join to give that element. Proposition 5.3.6. Let ๐ฟ be a finite lattice. (a) Element ๐‘ฅ โˆˆ ๐ฟ is join irreducible if and only if ๐‘ฅ covers exactly one element. (b) For any ๐‘ฅ โˆˆ ๐ฟ, if we let (5.2)

๐ผ๐‘ฅ = {๐‘Ÿ โ‰ค ๐‘ฅ โˆฃ ๐‘Ÿ โˆˆ Irr(๐ฟ)}, then ๐‘ฅ = โ‹ ๐ผ๐‘ฅ .

Proof. (a) First note that, by definition, 0ฬ‚ is not join irreducible. And 0ฬ‚ covers no elements, so the proposition is true in that case. Now assume ๐‘ฅ โ‰  0.ฬ‚ For the forward direction suppose, towards a contradiction, that ๐‘ฅ covers (at least) two elements ๐‘ฆ, ๐‘ง. But then ๐‘ฅ = ๐‘ฆ โˆจ ๐‘ง since ๐‘ฅ is clearly an upper bound for ๐‘ฆ, ๐‘ง and there can be no smaller bound because of the covering relations. This contradicts the fact that ๐‘ฅ is join irreducible. For the converse, let ๐‘ฅ cover a unique element ๐‘ฅโ€ฒ . If ๐‘ฅ = ๐‘ฆ โˆจ ๐‘ง with ๐‘ฆ, ๐‘ง < ๐‘ฅ, then we must have ๐‘ฆ, ๐‘ง โ‰ค ๐‘ฅโ€ฒ because ๐‘ฅ covers no other element. This forces ๐‘ฆ โˆจ ๐‘ง โ‰ค ๐‘ฅโ€ฒ < ๐‘ฅ which is the desired contradiction. (b) We induct on the number of elements in [0,ฬ‚ ๐‘ฅ]. If ๐‘ฅ = 0,ฬ‚ then the statement is true because ๐ผ๐‘ฅ = โˆ… and the empty join equals 0ฬ‚ just as the empty meet equals 1.ฬ‚ If ๐‘ฅ > 0,ฬ‚ then there are two cases. If ๐‘ฅ โˆˆ Irr(๐ฟ), then ๐‘ฅ is the maximum element of ๐ผ๐‘ฅ so โ‹ ๐ผ๐‘ฅ = ๐‘ฅ by Proposition 5.3.2(e). If ๐‘ฅ โˆ‰ Irr(๐ฟ), then, by part (a), ๐‘ฅ covers more than Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

5.3. Lattices

153

one element. So we can write ๐‘ฅ = ๐‘ฆ โˆจ ๐‘ง for any pair of elements ๐‘ฆ, ๐‘ง covered by ๐‘ฅ. By induction, ๐‘ฆ = โ‹ ๐ผ๐‘ฆ and ๐‘ง = โ‹ ๐ผ๐‘ง where ๐ผ๐‘ฆ , ๐ผ๐‘ง โŠ† ๐ผ๐‘ฅ . It follows that ๐‘ฅ = ๐‘ฆ โˆจ ๐‘ง = ( ๐ผ๐‘ฆ ) โˆจ ( ๐ผ๐‘ง ) = (๐ผ โˆช ๐ผ ) โ‰ค ๐ผ โ‰ค๐‘ฅ โ‹ โ‹ โ‹ ๐‘ฆ ๐‘ง โ‹ ๐‘ฅ where the last inequality follows since ๐‘Ÿ โ‰ค ๐‘ฅ for all ๐‘Ÿ โˆˆ ๐ผ๐‘ฅ . The previous displayed inequalities force ๐‘ฅ = โ‹ ๐ผ๐‘ฅ as desired. โ–ก Using this proposition, the reader can see immediately in Figure 5.6 that ๐’ฅ(๐‘ƒ) has four join irreducibles which form a subposet isomorphic to ๐‘ƒ. Birkhoffโ€™s theorem says this always happens. Theorem 5.3.7 (Fundamental Theorem of Finite Distributive Lattices). If ๐ฟ is a finite distributive lattice, then ๐ฟ โ‰… ๐’ฅ(๐‘ƒ) where ๐‘ƒ = Irr(๐ฟ). Proof. We need to define two order-preserving maps ๐‘“ โˆถ ๐ฟ โ†’ ๐’ฅ(๐‘ƒ) and ๐‘” โˆถ ๐’ฅ(๐‘ƒ) โ†’ ๐ฟ which are inverses of each other. If ๐‘ฅ โˆˆ ๐ฟ, then let ๐‘“(๐‘ฅ) = ๐ผ๐‘ฅ as defined by (5.2). Note that this is well-defined since ๐ผ๐‘ฅ is an ideal: if ๐‘Ÿ โˆˆ ๐ผ๐‘ฅ and ๐‘  โ‰ค ๐‘Ÿ is irreducible, then ๐‘  โ‰ค ๐‘Ÿ โ‰ค ๐‘ฅ so that ๐‘  โˆˆ ๐ผ๐‘ฅ . Also, ๐‘“ is order preserving since ๐‘ฅ โ‰ค ๐‘ฆ implies ๐ผ๐‘ฅ โŠ† ๐ผ๐‘ฆ by an argument similar to the one just given. For the inverse map, we let ๐‘”(๐ผ) = โ‹ ๐ผ for ๐ผ โˆˆ ๐’ฅ(๐‘ƒ). Clearly this is an element of ๐ฟ since ๐‘ƒ โŠ‚ ๐ฟ and so ๐‘” is well-defined. It is also order preserving since if ๐ผ โŠ† ๐ฝ, then ๐‘”(๐ฝ) =

โ‹

๐ฝ=(

โ‹

๐ผ) โˆจ (

โ‹

(๐ฝ โˆ’ ๐ผ)) โ‰ฅ

โ‹

๐ผ = ๐‘”(๐ผ).

It remains to prove that ๐‘“ and ๐‘” are inverses. If ๐‘ฅ โˆˆ ๐ฟ, then, using Proposition 5.3.6(b), ๐‘”(๐‘“(๐‘ฅ)) = ๐‘”(๐ผ๐‘ฅ ) =

โ‹

๐ผ๐‘ฅ = ๐‘ฅ.

Now consider ๐ผ โˆˆ ๐’ฅ(๐‘ƒ) and let ๐‘ฅ = ๐‘”(๐ผ) = โ‹ ๐ผ. Then ๐‘“(๐‘”(๐ผ)) = ๐ผ๐‘ฅ and we must show ๐ผ = ๐ผ๐‘ฅ . For the containment ๐ผ โŠ† ๐ผ๐‘ฅ , take ๐‘Ÿ โˆˆ ๐ผ. So ๐‘Ÿ โ‰ค โ‹ ๐ผ = ๐‘ฅ. But this means ๐‘Ÿ โˆˆ ๐ผ๐‘ฅ by definition (5.2), which gives the desired subset relation. To show ๐ผ๐‘ฅ โŠ† ๐ผ, take ๐‘Ÿ โˆˆ ๐ผ๐‘ฅ . We have that โ‹ ๐ผ = ๐‘ฅ = โ‹ ๐ผ๐‘ฅ . So ๐‘Ÿ โˆง (โ‹ ๐ผ) = ๐‘Ÿ โˆง (โ‹ ๐ผ๐‘ฅ ) and, applying the distributive law, (5.3)

โ‹

{๐‘Ÿ โˆง ๐‘  โˆฃ ๐‘  โˆˆ ๐ผ} =

{๐‘Ÿ โˆง ๐‘  โˆฃ ๐‘  โˆˆ ๐ผ๐‘ฅ }. โ‹

Since ๐‘Ÿ โˆˆ ๐ผ๐‘ฅ , the set on the right in (5.3) contains ๐‘Ÿ โˆง ๐‘Ÿ = ๐‘Ÿ. Furthermore, every element of that set is of the form ๐‘Ÿ โˆง ๐‘  โ‰ค ๐‘Ÿ. It follows that the right-hand side of the equality in (5.3) is just ๐‘Ÿ. But ๐‘Ÿ is join irreducible, so there must be some element ๐‘  โˆˆ ๐ผ on the left in (5.3) such that ๐‘Ÿ โˆง ๐‘  = ๐‘Ÿ. By Proposition 5.3.2(e) this forces ๐‘Ÿ โ‰ค ๐‘  โˆˆ ๐ผ. Since ๐ผ is an ideal, we have ๐‘Ÿ โˆˆ ๐ผ which is the final step of the proof. โ–ก

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

154

5. Counting with Partially Ordered Sets

5.4. The Mรถbius function of a poset The Mรถbius function is a fundamental invariant of any locally finite poset. A special case of this function was first studied in number theory. But the generalization to partially ordered sets is both more powerful and also in some ways more intuitive. It is of great use to enumerators because, as we will see in the next section, it permits one to invert certain summation formulas to get information about the summands. Let ๐‘ƒ be a locally finite poset with a 0.ฬ‚ The (one-variable) Mรถbius function of ๐‘ƒ is a map ๐œ‡ โˆถ ๐‘ƒ โ†’ โ„ค defined inductively by if ๐‘ฅ = 0,ฬ‚

1 (5.4)

๐œ‡(๐‘ฅ) = { โˆ’ โˆ‘ ๐œ‡(๐‘ฆ)

otherwise.

๐‘ฆ 1 so that ๐‘ฅ > ๐‘“(๐‘ฅ) > 0ฬ‚๐‘„ . Now, by induction, ๐œ‡๐‘ƒ (๐‘ฅ) = โˆ’ โˆ‘ ๐œ‡๐‘ƒ (๐‘ฆ) = โˆ’

โˆ‘

๐‘ฆ ๐‘˜๐‘. Show that ๐‘(๐‘›, ๐‘˜) โ‰ก 0 (mod ๐‘). Hint: Use part (a). (c) Let ๐‘ be prime. Given ๐‘› โ‰ฅ ๐‘˜ โ‰ฅ 0, write ๐‘› = ๐‘›โ€ณ ๐‘ + ๐‘›โ€ฒ where 0 โ‰ค ๐‘›โ€ฒ < ๐‘ and ๐‘˜ = ๐‘˜โ€ณ (๐‘ โˆ’ 1) + ๐‘˜โ€ฒ where 0 โ‰ค ๐‘˜โ€ฒ < ๐‘ โˆ’ 1. Then โ€ณ โ€ณ ๐‘› ๐‘(๐‘›, ๐‘› โˆ’ ๐‘˜) โ‰ก (โˆ’1)๐‘˜ ( โ€ณ )๐‘(๐‘›โ€ฒ , ๐‘›โ€ฒ โˆ’ ๐‘˜โ€ฒ ) (mod ๐‘). ๐‘˜

Hint: Use part (a). (d) Consider two polynomials ๐‘“(๐‘ฅ), ๐‘”(๐‘ฅ) โˆˆ โ„ค[๐‘ฅ] and let ๐‘š โˆˆ โ„™. We say that ๐‘“(๐‘ฅ) is congruent to ๐‘”(๐‘ฅ) modulo ๐‘š if every coefficient of ๐‘“(๐‘ฅ) โˆ’ ๐‘”(๐‘ฅ) is divisible by ๐‘š. If ๐‘ is prime, show that ๐‘ฅโ†“๐‘ โ‰ก ๐‘ฅ๐‘ โˆ’ ๐‘ฅ (mod ๐‘) where ๐‘ฅโ†“๐‘ = ๐‘ฅ(๐‘ฅ โˆ’ 1) โ‹ฏ (๐‘ฅ โˆ’ ๐‘ + 1). Hint: Use Theorem 3.1.2 and part (a). (23) Finish the proof of Theorem 6.5.6 (24) (a) Show that if ๐บ acts on ๐‘‹, then it also acts on ๐‘†(๐‘‹, ๐‘˜), the set of partitions of ๐‘‹ into ๐‘˜ blocks. (b) Let ๐‘ be prime and let ๐‘› โ‰ฅ ๐‘. Show for the Stirling numbers of the second kind that ๐‘†(๐‘›, ๐‘˜) โ‰ก ๐‘†(๐‘› โˆ’ ๐‘, ๐‘˜ โˆ’ ๐‘) + ๐‘†(๐‘› โˆ’ ๐‘ + 1, ๐‘˜) (mod ๐‘). (c) If ๐‘ is prime and ๐‘˜ โˆˆ โ„ค, then show ๐‘†(๐‘, ๐‘˜) โ‰ก {

1 if ๐‘˜ = 1 or ๐‘ 0 otherwise

(mod ๐‘).

(d) Suppose ๐‘ is prime and 0 โ‰ค ๐‘˜ โ‰ค ๐‘›. If ๐‘— satisfies ๐‘๐‘— โ‰ค ๐‘˜ < ๐‘๐‘—+1 , then ๐‘†(๐‘› + ๐‘๐‘— (๐‘ โˆ’ 1), ๐‘˜) โ‰ก ๐‘†(๐‘›, ๐‘˜) (mod ๐‘). Hint: Use part (b). (25) (a) Recall from Exercise 10 in Chapter 1 that Pascalโ€™s triangle is fractal modulo 2. Give a second proof of this by using Lucasโ€™s Congruence (Theorem 6.5.6) to show that if 0 โ‰ค ๐‘› < 2๐‘š and 0 โ‰ค ๐‘˜ โ‰ค ๐‘› + 2๐‘š , then ๐‘› โŽง( ) if 0 โ‰ค ๐‘˜ โ‰ค ๐‘› โŽช ๐‘˜ โŽช ๐‘› + 2๐‘š if ๐‘› < ๐‘˜ < 2๐‘š (mod 2). ( )โ‰ก 0 ๐‘˜ โŽจ โŽช โŽช( ๐‘› ) if 2๐‘š โ‰ค ๐‘˜ โ‰ค ๐‘› + 2๐‘š โŽฉ ๐‘˜ โˆ’ 2๐‘š (b) Find and prove the analogue of part (a) for an arbitrary prime ๐‘ โ‰ฅ 2. (26) Use Proposition 6.5.8 to prove an analogue of Proposition 6.5.9 for any ๐‘› โˆˆ โ„™. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Exercises

217

(27) (a) Prove that for any group ๐บ the poset ๐ฟ(๐บ) is a lattice. In particular, if ๐ป, ๐พ โˆˆ ๐ฟ(๐บ), then ๐ป โˆง ๐พ = ๐ป โˆฉ ๐พ and ๐ป โˆจ ๐พ is the subgroup of ๐บ generated by ๐ป, ๐พ. (b) Let โ„ค๐‘ be the integers modulo ๐‘ and consider the direct sum โ„ค๐‘๐‘› of ๐‘› copies of โ„ค๐‘ . Show that if ๐‘ is prime, then ๐ฟ(โ„ค๐‘๐‘› ) โ‰… ๐ฟ๐‘› (๐‘), the subspace lattice of dimension ๐‘› over the Galois field with ๐‘ elements. (c) Use part (b) to prove a congruence modulo ๐‘2 for the binomial coefficients. (28) Prove Lemma 6.6.1(b). (29) Prove Lemma 6.6.3. (30) Let ๐‘ โˆˆ โ„™ and ๐‘” = (1, 2, . . . , ๐‘). Consider the group ๐บ = โŸจ๐‘”โŸฉ acting on the set of multisets ๐‘‹ = (([๐‘›] )). ๐‘˜ (a) Suppose ๐‘€ = {{1๐‘š1 , 2๐‘š2 , . . . , ๐‘›๐‘š๐‘› }}. Show that ๐‘”๐‘€ = ๐‘€ if and only if ๐‘š1 = ๐‘š2 = โ‹ฏ = ๐‘š๐‘ . (b) Show that if ๐‘ is prime, then ๐‘› ๐‘›โˆ’๐‘ (( )) โ‰ก โˆ‘ (( )) (mod ๐‘). ๐‘˜ ๐‘˜ โˆ’ ๐‘š๐‘ ๐‘šโ‰ฅ0 (31) (a) Prove that ๐‘’2๐‘˜๐œ‹๐‘–/๐‘› is a primitive ๐‘›th root of unity if and only if gcd(๐‘˜, ๐‘›) = 1. (b) Prove that the CSP is exhibited by the triple [๐‘›] ๐‘› ] ). (( ), โŸจ(1, 2, . . . , ๐‘›)โŸฉ, [ ๐‘˜ ๐‘ž ๐‘˜

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

10.1090/gsm/210/07

Chapter 7

Counting with Symmetric Functions

A formal power series is symmetric if it is invariant under permutation of its variables. We have already seen generating functions of this type arise naturally in Theorem 6.4.2. Symmetric functions have many other connections to combinatorics some of which we will discuss in this chapter. In particular, we will see that they arise when studying log-concavity, Young tableaux, various posets, chromatic polynomials, and the cyclic sieving phenomenon. They are also intimately connected with group representations. The appendix to this book contains a summary of the facts we will need in this regard, and more information can be found in Saganโ€™s text [79]. For a wealth of information about symmetric functions in general, see Macdonaldโ€™s book [60].

7.1. The algebra of symmetric functions, Sym In this section we formally define the algebra of symmetric functions and introduce some of its standard bases. Along the way, we prove the Fundamental Theorem of Symmetric Functions and show how the coefficients of a polynomial can be expressed as a symmetric function of its roots. Let ๐ฑ = {๐‘ฅ1 , ๐‘ฅ2 , ๐‘ฅ3 , . . . } be a countably infinite set of commuting variables. Con๐œ† ๐œ† ๐œ† sider the algebra of formal power series โ„‚[[๐ฑ]]. A monomial ๐‘š = ๐‘ฅ๐‘–11 ๐‘ฅ๐‘–22 โ‹ฏ ๐‘ฅ๐‘–๐‘™๐‘™ has 6 5 degree deg ๐‘š = โˆ‘๐‘– ๐œ†๐‘– . For example, deg(๐‘ฅ2 ๐‘ฅ4 ๐‘ฅ8 ) = 5 + 1 + 6 = 12. We say that ๐‘“(๐ฑ) โˆˆ โ„‚[[๐ฑ]] is homogeneous of degree ๐‘› if deg ๐‘š = ๐‘› for all monomials ๐‘š appearing in ๐‘“(๐ฑ). A weaker condition is that ๐‘“(๐ฑ) have bounded degree which means that there is an ๐‘› with deg ๐‘š โ‰ค ๐‘› for all ๐‘š appearing in ๐‘“(๐ฑ). To illustrate, ๐‘“(๐ฑ) = โˆ‘๐‘– ๐‘š. For example, (7.3)

(1, 2)(๐‘ฅ12 + 2๐‘ฅ1 ๐‘ฅ23 + 5๐‘ฅ14 ๐‘ฅ3 โˆ’ ๐‘ฅ3 ๐‘ฅ4 ) = ๐‘ฅ22 + 2๐‘ฅ13 ๐‘ฅ2 + 5๐‘ฅ24 ๐‘ฅ3 โˆ’ ๐‘ฅ3 ๐‘ฅ4 .

Call ๐‘“(๐ฑ) symmetric if ๐œ‹๐‘“ = ๐‘“ for all ๐œ‹ in every symmetric group ๐”–๐‘š . Equivalently, ๐œ† ๐œ† ๐œ† ๐œ† ๐œ† ๐œ† any two monomials ๐‘ฅ๐‘–11 ๐‘ฅ๐‘–22 โ‹ฏ ๐‘ฅ๐‘–๐‘™๐‘™ and ๐‘ฅ๐‘—11 ๐‘ฅ๐‘—22 โ‹ฏ ๐‘ฅ๐‘—๐‘™๐‘™ , where ๐‘–1 , . . . , ๐‘–๐‘™ are distinct and similarly for ๐‘—1 , . . . , ๐‘— ๐‘™ , have the same coefficient in ๐‘“(๐ฑ) since one can always find a permutation ๐œ‹ such that ๐œ‹(๐‘–๐‘˜ ) = ๐‘— ๐‘˜ for all ๐‘˜. Another equivalent description is that any ๐œ† ๐œ† ๐œ† ๐œ† ๐œ† ๐œ† monomial ๐‘ฅ๐‘–11 ๐‘ฅ๐‘–22 โ‹ฏ ๐‘ฅ๐‘–๐‘™๐‘™ has the same coefficient as ๐‘ฅ1 1 ๐‘ฅ2 2 โ‹ฏ ๐‘ฅ๐‘™ ๐‘™ in ๐‘“(๐ฑ). To illustrate, (7.4)

๐‘“(๐ฑ) = 4๐‘ฅ15 + 4๐‘ฅ25 + 4๐‘ฅ35 + โ‹ฏ โˆ’ 6๐‘ฅ12 ๐‘ฅ22 โˆ’ 6๐‘ฅ12 ๐‘ฅ32 โˆ’ 6๐‘ฅ22 ๐‘ฅ32 โˆ’ โ‹ฏ

is symmetric. Let Sym๐‘› = Sym๐‘› (๐ฑ) = {๐‘“ โˆˆ โ„‚[[๐ฑ]] โˆฃ ๐‘“ is symmetric and homogeneous of degree ๐‘›}. The algebra of symmetric functions is Sym = Sym(๐ฑ) =

โจ

Sym๐‘› (๐ฑ).

๐‘›โ‰ฅ0

Alternatively, Sym(๐ฑ) is the set of all symmetric power series in โ„‚[[๐ฑ]] of bounded degree. This is because elements of the direct sum can only have a finite number of components which are nonzero. So the series in (7.4) is in Sym, but the ones in (7.1) and (7.3) are not. There are a number of interesting bases for Sym. We start with those functions obtained by symmetrizing a monomial. If ๐œ† = (๐œ†1 , ๐œ†2 , . . . , ๐œ†๐‘™ ) is a partition, then the associated monomial symmetric function is ๐œ†

๐œ†

๐œ†

๐‘š๐œ† = ๐‘š๐œ† (๐ฑ) = โˆ‘ ๐‘ฅ๐‘–11 ๐‘ฅ๐‘–22 โ‹ฏ ๐‘ฅ๐‘–๐‘™๐‘™ where the sum is over all distinct monomials having exponents ๐œ†1 , ๐œ†2 , โ‹ฏ , ๐œ†๐‘™ . We will often drop the parentheses and commas in the subscript ๐œ† as well as use multiplicity notation. As examples, in (7.4) we have ๐‘“ = 4๐‘š(5) โˆ’ 6๐‘š(2,2) = 4๐‘š5 โˆ’ 6๐‘š22 , and (7.5)

๐‘š21 = ๐‘ฅ12 ๐‘ฅ2 + ๐‘ฅ1 ๐‘ฅ22 + ๐‘ฅ12 ๐‘ฅ3 + ๐‘ฅ1 ๐‘ฅ32 + ๐‘ฅ22 ๐‘ฅ3 + ๐‘ฅ2 ๐‘ฅ32 + โ‹ฏ .

We must verify that we have defined a basis for Sym. Theorem 7.1.1. The ๐‘š๐œ† as ๐œ† varies over all partitions form a basis for Sym. Consequently dim Sym๐‘› = ๐‘(๐‘›), the number of partitions of ๐‘›. Proof. The second sentence follows immediately from the first. And it is clear that the ๐‘š๐œ† are independent since no two contain a monomial in common. So it suffices to show that every ๐‘“ โˆˆ Sym can be written as a linear combination of the ๐‘š๐œ† . Suppose ๐œ† ๐œ† ๐œ† ๐‘ฅ๐‘–11 ๐‘ฅ๐‘–22 โ‹ฏ ๐‘ฅ๐‘–๐‘™๐‘™ is a monomial appearing in ๐‘“ and having coefficient ๐‘ โˆˆ โ„‚. Without loss Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

7.1. The algebra of symmetric functions, Sym

221

of generality we can assume the indices have been arranged so that ๐œ†1 โ‰ฅ ๐œ†2 โ‰ฅ โ‹ฏ โ‰ฅ ๐œ†๐‘™ and let ๐œ† = (๐œ†1 , ๐œ†2 , . . . , ๐œ†๐‘™ ). Since ๐‘“ is symmetric, every monomial in ๐‘“ with exponents ๐œ†1 , ๐œ†2 , . . . , ๐œ†๐‘™ appears with coefficient ๐‘. So ๐‘“ โˆ’ ๐‘๐‘š๐œ† is still symmetric and contains no monomials with these exponents. Since ๐‘“ is of bounded degree, we can repeat this process a finite number of times until we reach the zero power series. It follows that ๐‘“ will be a linear combination of the monomial symmetric functions which appear during this algorithm. โ–ก There are three bases which are formed multiplicatively in that one first defines ๐‘“๐‘› for ๐‘› โˆˆ โ„™ and then sets (7.6)

๐‘“ ๐œ† = ๐‘“ ๐œ†1 ๐‘“ ๐œ†2 โ‹ฏ ๐‘“ ๐œ†๐‘™

where ๐œ† = (๐œ†1 , ๐œ†2 , . . . , ๐œ†๐‘™ ). Specifically, for ๐‘› โ‰ฅ 1 we define the ๐‘›th power sum symmetric function ๐‘๐‘› = ๐‘š(๐‘›) = โˆ‘ ๐‘ฅ๐‘–๐‘› , ๐‘–โ‰ฅ1

the ๐‘›th elementary symmetric function โˆ‘

๐‘’ ๐‘› = ๐‘š(1๐‘› ) =

๐‘ฅ๐‘–1 โ‹ฏ ๐‘ฅ๐‘–๐‘› ,

๐‘–1 ๐‘ 2 > โ‹ฏ > ๐‘ ๐‘˜ } โŠ‚ โ„•. Since ๐‘“ is weakly decreasing on ๐œ‹ by condition C1, those ๐‘Ÿ with ๐‘“(๐‘Ÿ) = ๐‘ 1 must come first in ๐œ‹. Furthermore, such ๐‘Ÿ must be arranged in increasing order since, if not, then there would be a descent which would force two of these ๐‘Ÿ to have distinct images by C2. Similar considerations show that the next elements in ๐œ‹ must be those such that ๐‘“(๐‘Ÿ) = ๐‘ 2 in increasing order, and so forth. Since all of the choices made in constructing ๐œ‹ are forced on us by the definition, the permutation is unique and we have proved the first statement of the proposition. As for (7.18), the uniqueness statement just proved shows that the union is disjoint. And existence of a compatible ๐œ‹ for each ๐‘“ shows containment of the left-hand side in the right. The other containment is trivial since each ๐’ž(๐œ‹) consists of functions ๐‘“ โˆถ [๐‘›] โ†’ โ„•. โ–ก Define the size of ๐‘“ โˆถ [๐‘›] โ†’ โ„• to be ๐‘›

|๐‘“| = โˆ‘ ๐‘“(๐‘–).

(7.19)

๐‘–=1

Continuing our example |๐‘“| = 21 + 21 + 21 + 20 + 20 + 10 + 0 + 0 = 113. It will also be instructive to consider the following subsets of ๐’ž(๐œ‹) where the maximum of a function is the maximum of the values in its image ๐’ž๐‘š (๐œ‹) = {๐‘“ โˆˆ ๐’ž(๐œ‹) โˆฃ max ๐‘“ โ‰ค ๐‘š}. There are nice generating functions associated with ๐’ž(๐œ‹) and ๐’ž๐‘š (๐œ‹). Lemma 7.4.2. For any ๐œ‹ โˆˆ ๐”–๐‘› we have (7.20)

โˆ‘ ๐‘ฅ|๐‘“| = ๐‘“โˆˆ๐’ž(๐œ‹)

๐‘ฅmaj ๐œ‹ (1 โˆ’ ๐‘ฅ)(1 โˆ’ ๐‘ฅ2 ) โ‹ฏ (1 โˆ’ ๐‘ฅ๐‘› )

and (7.21)

โˆ‘ #๐’ž๐‘š (๐œ‹)๐‘ฅ๐‘š = ๐‘šโ‰ฅ0

๐‘ฅdes ๐œ‹ . (1 โˆ’ ๐‘ฅ)๐‘›+1

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

7.4. ๐‘ƒ-partitions

237

Proof. We will prove the first equality as the demonstration of the second is similar and so is left as an exercise. The basic idea behind the proof is the same as used in the demonstration of Theorem 1.3.4 where one adds or subtracts sufficient amounts to turn weak inequalities into strict ones or vice versa. An example will follow the proof. Let ๐œ‹ = ๐œ‹1 ๐œ‹2 . . . ๐œ‹๐‘› with Des ๐œ‹ = {๐‘‘1 < ๐‘‘2 < โ‹ฏ < ๐‘‘๐‘˜ }. We will construct a bijection ๐œ™ โˆถ ๐’ž(๐œ‹) โ†’ ฮ›๐‘› where ฮ›๐‘› is the set of all partitions ๐œ† satisfying the length restriction โ„“(๐œ†) โ‰ค ๐‘›. Given ๐‘“ โˆˆ ๐’ž(๐œ‹), we will construct a sequence of functions ๐‘“ = ๐‘“0 , . . . , ๐‘“๐‘˜ where ๐‘“๐‘– will remove the strict inequality restriction at index ๐‘‘๐‘– from ๐‘“๐‘–โˆ’1 . So let ๐‘“1 be obtained from ๐‘“ by subtracting one from each of ๐‘“(๐œ‹1 ), . . . , ๐‘“(๐œ‹๐‘‘1 ) and leaving the other ๐‘“ values the same. Similarly, ๐‘“2 is constructed from ๐‘“1 by subtracting one from ๐‘“1 (๐œ‹1 ), . . . , ๐‘“1 (๐œ‹๐‘‘2 ) with other values constant, and so forth. By the end, the only restrictions on ๐‘“๐‘˜ are that we have ๐‘“๐‘˜ (๐œ‹1 ) โ‰ฅ โ‹ฏ โ‰ฅ ๐‘“๐‘˜ (๐œ‹๐‘› ) โ‰ฅ 0. So the nonzero images of ๐‘“๐‘˜ form a partition ๐œ† โˆˆ ฮ›๐‘› and we let ๐œ™(๐‘“) = ๐œ†. This is a bijection as its inverse is easy to construct. Furthermore, from the definition of the algorithm, it follows that ๐‘˜

|๐‘“| = |๐‘“1 | + ๐‘‘1 = โ‹ฏ = |๐œ†| + โˆ‘ ๐‘‘๐‘– = |๐œ†| + maj ๐œ‹. ๐‘–=1

Now appealing to Corollary 3.5.4 we have โˆ‘ ๐‘ฅ|๐‘“| = โˆ‘ ๐‘ฅ|๐œ†|+maj ๐œ‹ = ๐œ†โˆˆฮ›๐‘›

๐‘“โˆˆ๐’ž(๐œ‹)

๐‘ฅmaj ๐œ‹ (1 โˆ’ ๐‘ฅ)(1 โˆ’ ๐‘ฅ2 ) โ‹ฏ (1 โˆ’ ๐‘ฅ๐‘› ) โ–ก

as desired.

Using our running example, the vector of values of the initial function on ๐œ‹ is given by ๐‘“ = (21, 21, 21, 20, 20, 10, 0, 0). Since Des ๐œ‹ = {3, 6} our first step is to subtract one from the first 3 values to obtain ๐‘“1 = (20, 20, 20, 20, 20, 10, 0, 0). Next we subtract one from the first 6 values so that ๐‘“2 = (19, 19, 19, 19, 19, 9, 0, 0). Taking the nonzero components gives ๐œ† = (19, 19, 19, 19, 19, 9). We now have all the tools needed to find generating functions for partitions whose parts are distributed over a poset. Let ๐‘ƒ be a partial order on the set [๐‘›]. In order to distinguish the usual total order on integers from the partial order in ๐‘ƒ, we will use ๐‘– โ‰ค ๐‘— for the former and ๐‘– โŠด ๐‘— for the latter. So in the poset on the left in Figure 7.8 we have 3 โŠฒ 2, but 2 < 3 as integers. If ๐‘ƒ is a poset on [๐‘›], then a ๐‘ƒ-partition is a map

1

4

0

4

2

5

2

3

5

1

3

7

Figure 7.8. A poset ๐‘ƒ on [4] on the left and a ๐‘ƒ-partition on the right

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

238

7. Counting with Symmetric Functions

๐‘“ โˆถ ๐‘ƒ โ†’ โ„• such that PP1 ๐‘– โŠด ๐‘— implies ๐‘“(๐‘–) โ‰ฅ ๐‘“(๐‘—) and PP2 ๐‘– โŠด ๐‘— and ๐‘– > ๐‘— implies ๐‘“(๐‘–) > ๐‘“(๐‘—). So PP1 says that ๐‘“ is weakly decreasing on ๐‘ƒ, while PP2 means that ๐‘“ is strictly decreasing on โ€œdescentsโ€ of ๐‘ƒ. A ๐‘ƒ-partition for the poset in Figure 7.8 is shown on the right with the values of ๐‘“ circled. Note that by transitivity, it suffices to assume that PP1 and PP2 hold when ๐‘– is covered by ๐‘—. Let Par ๐‘ƒ = {๐‘“ โˆถ ๐‘ƒ โ†’ โ„• โˆฃ ๐‘“ is a ๐‘ƒ-partition}. For the poset in Figure 7.8 we have Par ๐‘ƒ = {๐‘“ โˆถ [4] โ†’ โ„• โˆฃ ๐‘“(1) โ‰ฅ ๐‘“(2), ๐‘“(3) > ๐‘“(2), ๐‘“(2) โ‰ฅ ๐‘“(4)}. We will also need the Jordanโ€“Hรถlder set of an (arbitrary) poset ๐‘ƒ which is โ„’(๐‘ƒ) = {๐œ‹ โˆฃ ๐œ‹ is a linear extension of ๐‘ƒ}. Note that if ๐‘ƒ is a poset on [๐‘›], then โ„’(๐‘ƒ) โŠ† ๐”–๐‘› . Continuing the Figure 7.8 example, โ„’(๐‘ƒ) = {1324, 3124}. The next result, while not hard to prove, is crucial, as its name suggests. Lemma 7.4.3 (Fundamental Lemma of ๐‘ƒ-Partitions). Let ๐‘ƒ be a poset on [๐‘›]. Then we have ๐‘“ โˆˆ Par ๐‘ƒ if and only if ๐‘“ โˆˆ ๐’ž(๐œ‹) for some ๐œ‹ โˆˆ โ„’(๐‘ƒ). Thus Par ๐‘ƒ =

โจ„

๐’ž(๐œ‹).

๐œ‹โˆˆโ„’(๐‘ƒ)

Proof. We will just prove the forward implication as the reverse is similar. And the proof of the equation for Par ๐‘ƒ is also omitted as it follows the same lines as for (7.18). So suppose ๐‘“ โˆˆ Par ๐‘ƒ. We know from the first part of Lemma 7.4.1 that ๐‘“ is compatible with a unique ๐œ‹ โˆˆ ๐”–๐‘› . So we just need to show that ๐œ‹ โˆˆ โ„’(๐‘ƒ); that is, if ๐‘– โŠฒ ๐‘—, then ๐‘– should appear before ๐‘— in ๐œ‹. Assume, to the contrary, that we have (7.22)

๐œ‹ = ... ๐‘— ... ๐‘– ...

being the order of the two elements in ๐œ‹. Since ๐‘“ โˆˆ Par ๐‘ƒ and ๐‘– โŠฒ ๐‘— we must have ๐‘“(๐‘–) โ‰ฅ ๐‘“(๐‘—) by condition PP1. But C1 and the form of ๐œ‹ in (7.22) force ๐‘“(๐‘—) โ‰ฅ ๐‘“(๐‘–). Thus ๐‘“(๐‘–) = ๐‘“(๐‘—). This equality together with (7.22) and C2 imply ๐‘— < ๐‘–. But now we have ๐‘– โŠฒ ๐‘— and ๐‘– > ๐‘— as well as ๐‘“(๐‘–) = ๐‘“(๐‘—) which contradicts P2. This is the desired contradiction. โ–ก We now translate the previous result in terms of generating functions. Just as with compatible functions, use the notation Par๐‘š ๐‘ƒ = {๐‘“ โˆˆ Par ๐‘ƒ โˆฃ max ๐‘“ โ‰ค ๐‘š}. Theorem 7.4.4. For any poset ๐‘ƒ on [๐‘›] we have (7.23)

โˆ‘ ๐‘ฅ|๐‘“| = ๐‘“โˆˆPar ๐‘ƒ

โˆ‘๐œ‹โˆˆโ„’(๐‘ƒ) ๐‘ฅmaj ๐œ‹ (1 โˆ’ ๐‘ฅ)(1 โˆ’ ๐‘ฅ2 ) โ‹ฏ (1 โˆ’ ๐‘ฅ๐‘› )

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

7.4. ๐‘ƒ-partitions

239

and โˆ‘ | Par๐‘š ๐‘ƒ| ๐‘ฅ๐‘š =

(7.24)

โˆ‘๐œ‹โˆˆโ„’(๐‘ƒ) ๐‘ฅdes ๐œ‹

๐‘šโ‰ฅ0

(1 โˆ’ ๐‘ฅ)๐‘›+1

.

Proof. We prove (7.23), leaving (7.24) as an exercise. Using the previous lemma and then (7.20) yields โˆ‘ ๐‘ฅ ๐‘“โˆˆPar ๐‘ƒ

|๐‘“|

=

โˆ‘

โˆ‘ ๐‘ฅ

|๐‘“|

๐œ‹โˆˆโ„’(๐‘ƒ) ๐‘“โˆˆ๐’ž(๐œ‹)

=

โˆ‘๐œ‹โˆˆโ„’(๐‘ƒ) ๐‘ฅmaj ๐œ‹ (1 โˆ’ ๐‘ฅ)(1 โˆ’ ๐‘ฅ2 ) โ‹ฏ (1 โˆ’ ๐‘ฅ๐‘› ) โ–ก

which is what we wished to demonstrate.

As a check, consider the poset ๐‘ƒ which is the chain 1 โŠฒ 2 โŠฒ . . . โŠฒ ๐‘›. Then the only inequalities satisfied by ๐‘“ โˆˆ Par ๐‘ƒ are ๐‘“(1) โ‰ฅ ๐‘“(2) โ‰ฅ โ‹ฏ โ‰ฅ ๐‘“(๐‘›) โ‰ฅ 0. So ๐‘“ corresponds to a partition ๐œ† with at most ๐‘› parts. On the other hand โ„’(๐‘ƒ) consists of the single permutation ๐œ‹ = 12 . . . ๐‘› with maj ๐œ‹ = 0. So (7.23) becomes (7.25)

โˆ‘ ๐‘ฅ|๐œ†| = โ„“(๐œ†)โ‰ค๐‘›

1 (1 โˆ’ ๐‘ฅ)(1 โˆ’ ๐‘ฅ2 ) โ‹ฏ (1 โˆ’ ๐‘ฅ๐‘› )

which agrees with Corollary 3.5.4. Of course, this cannot be considered a new proof of the corollary since it was used in the demonstration of (7.20). But at least it suggests we havenโ€™t made any mistakes! Another case is the chain ๐‘› โŠฒ ๐‘› โˆ’ 1 โŠฒ . . . โŠฒ 1. Now the ๐‘“ โˆˆ Par ๐‘ƒ satisfy the inequalities ๐‘“(๐‘›) > ๐‘“(๐‘› โˆ’ 1) > โ‹ฏ > ๐‘“(1) โ‰ฅ 0. The set โ„’(๐‘ƒ) still contains a unique element, but it is ๐œ‹ = ๐‘› . . . 21 which has ๐‘› maj ๐œ‹ = 1 + 2 + โ‹ฏ + (๐‘› โˆ’ 1) = ( ). 2 Plugging into (7.23) we see that ๐‘ฅ( 2 ) . (1 โˆ’ ๐‘ฅ)(1 โˆ’ ๐‘ฅ2 ) โ‹ฏ (1 โˆ’ ๐‘ฅ๐‘› ) ๐‘›

(7.26)

โˆ‘ ๐‘ฅ|๐‘“| = ๐‘“โˆˆPar ๐‘ƒ

The reader may find it instructive to write down the bijection which permits one to derive this equation from (7.25). This map is a special case of the one used in the proof of (7.20) but it is easier to see what is going on in this simple case. The time has come to fulfill our promise from the beginning of this section to derive the Hook Formula, (7.10), from the generating function for reverse plane partitions, (7.14). Let ๐œ† be the Young diagram of a partition of ๐‘›. We turn ๐œ† into a poset ๐‘ƒ๐œ† by partially ordering the cells of ๐œ† componentwise: (๐‘–, ๐‘—) โŠด (๐‘–โ€ฒ , ๐‘—โ€ฒ ) whenever ๐‘– โ‰ค ๐‘–โ€ฒ and ๐‘— โ‰ค ๐‘—โ€ฒ . See Figure 7.9 for an example where ๐œ† = (4, 3, 1). So ๐‘ƒ๐œ† is formed from the Young diagram of ๐œ† by rotating 135โˆ˜ counterclockwise and imposing a grid of covers. Note that #โ„’(๐‘ƒ๐œ† ) = ๐‘“๐œ† because there is a simple bijection between SYT ๐‘‡ of shape ๐œ† and linear extensions of ๐‘ƒ๐œ† : each tableau ๐‘‡ corresponds to a linear extension of the cells ๐‘ 1 , . . . , ๐‘๐‘› where they are ordered so that ๐‘ ๐‘– is the cell of ๐‘‡ containing ๐‘– for 1 โ‰ค ๐‘– โ‰ค ๐‘›. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

240

7. Counting with Symmetric Functions

๐‘ƒ๐œ†

๐œ†

๐‘ƒ๐œ†โˆ—

Figure 7.9. A Young diagram and associated posets

Now consider the poset dual ๐‘ƒ๐œ†โˆ— and label its elements with the numbers in [๐‘›] in any way which corresponds to a linear extension of ๐‘ƒ๐œ†โˆ— as described in the previous paragraph for ๐‘ƒ๐œ† . It is easy to see that the ๐‘ƒ๐œ†โˆ— -partitions are precisely the reverse plane partitions of shape ๐œ†. (The use of the dual corresponds to these plane partitions being โ€œreverseโ€.) And clearly we still have #โ„’(๐‘ƒ๐œ†โˆ— ) = ๐‘“๐œ† . Combining (7.14) and (7.23) yields. ๐‘(๐‘ฅ) 1 = โˆ‘ rpp๐‘› (๐œ†)๐‘ฅ๐‘› = โ„Ž๐‘–,๐‘— (1 โˆ’ ๐‘ฅ)(1 โˆ’ ๐‘ฅ2 ) โ‹ฏ (1 โˆ’ ๐‘ฅ๐‘› ) 1 โˆ’ ๐‘ฅ ๐‘›โ‰ฅ0 (๐‘–,๐‘—)โˆˆ๐œ† โˆ

where ๐‘(๐‘ฅ) = โˆ‘๐œ‹โˆˆโ„’(๐‘ƒ โˆ— ) ๐‘ฅmaj ๐œ‹ . Thus ๐œ†

๐‘›! (1 โˆ’ ๐‘ฅ)(1 โˆ’ ๐‘ฅ2 ) โ‹ฏ (1 โˆ’ ๐‘ฅ๐‘› ) = โ„Ž๐‘–,๐‘— ๐‘ฅโ†’1 โ„Ž โˆ โˆ(๐‘–,๐‘—)โˆˆ๐œ† 1 โˆ’ ๐‘ฅ (๐‘–,๐‘—)โˆˆ๐œ† ๐‘–,๐‘—

๐‘“๐œ† = ๐‘(1) = lim which is the Hook Formula.

7.5. The Robinsonโ€“Schenstedโ€“Knuth correspondence This section will be devoted to proving the following important identity. Theorem 7.5.1. For any given ๐‘› โ‰ฅ 0 we have (7.27)

โˆ‘ (๐‘“๐œ† )2 = ๐‘›! . ๐œ†โŠข๐‘›

From the point of view of representation theory this is just the special case of equation (A.9) in the appendix where ๐บ = ๐”–๐‘› and the dimensions are given by (A.7). However, we wish to give a bijective proof of (7.27). This map was discovered in two very different forms by Robinson [75] and Schensted [81]. It is the latter description which will be presented here. We will also see that this algorithm and the corresponding identity can be generalized from standard to semistandard Young tableaux. To prove (7.27), it suffices to construct a bijection (7.28)

RS

๐œ‹ โ†ฆ (๐‘ƒ, ๐‘„)

between permutations ๐œ‹ โˆˆ ๐”–๐‘› and pairs of SYT (๐‘ƒ, ๐‘„) of the same shape ๐œ† โŠข ๐‘›. The heart of this construction will be a method of inserting a positive integer into a tableau. A partial Young tableau (PYT) is a filling of a shape with distinct positive integers such that the rows and columns increase. A PYT is standard precisely when its entries are Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

7.5. The Robinsonโ€“Schenstedโ€“Knuth correspondence

241

๐‘ƒ= 1 3 6 9 2 8 10 4 7 1 3 6 9 โ†๐Ÿ“ 2 8 10 4 7

1 3 ๐Ÿ“ 9 2 8 10 โ† ๐Ÿ” 4 7

1 3 5 9 2 ๐Ÿ” 10 4 โ†๐Ÿ– 7

1 3 5 9 2 6 10 4 ๐Ÿ– 7

= ๐‘Ÿ5 (๐‘ƒ).

Figure 7.10. Inserting ๐‘ฅ = 5 into a partial tableau ๐‘ƒ

[๐‘›] for some ๐‘›. A partial tableau ๐‘ƒ is shown at the top in Figure 7.10. Now given a PYT ๐‘ƒ and a positive integer ๐‘ฅ โˆ‰ ๐‘ƒ, we insert ๐‘ฅ into ๐‘ƒ using the following algorithm: RS1 Set ๐‘… โ‰” the first row of ๐‘ƒ. RS2 While ๐‘ฅ is less than some element of row ๐‘…, let ๐‘ฆ be the leftmost such element and replace ๐‘ฆ by ๐‘ฅ in ๐‘…. Repeat this step with ๐‘… โ‰” the row below ๐‘… and ๐‘ฅ โ‰” ๐‘ฆ. RS3 Now ๐‘ฅ is greater than every element of ๐‘…, so place ๐‘ฅ at the end of this row and terminate. In step RS2 we say that ๐‘ฅ bumps ๐‘ฆ. An example of inserting 5 into the PYT in Figure 7.10 is given in the second row of the figure. Elements being bumped are written in boldface and the notation ๐‘… โ† ๐‘ฅ means that ๐‘ฅ is being inserted in row ๐‘…. If ๐‘ƒ โ€ฒ is the result of inserting ๐‘ฅ into ๐‘ƒ by rows, then we write ๐‘Ÿ๐‘ฅ (๐‘ƒ) = ๐‘ƒ โ€ฒ . The reader should check that this operation is well-defined in that ๐‘ƒ โ€ฒ is still a PYT. There is a second operation needed to describe the map (7.28) which will be used for the second component. An outer corner of a shape ๐œ† (or of a tableau of that shape) is a cell (๐‘–, ๐‘—) โˆ‰ ๐œ† such that ๐œ† โˆช {(๐‘–, ๐‘—)} is the Young diagram of a partition. The outer corners of ๐‘„ in Figure 7.11 are (1, 5), (2, 3), (3, 2), and (5, 1). Suppose we have a partial tableau ๐‘„, ๐‘ฆ > max ๐‘„, and (๐‘–, ๐‘—) which is an outer corner of ๐‘„. The tableau ๐‘„โ€ฒ obtained by placing ๐‘ฆ in ๐‘„ at (๐‘–, ๐‘—) has all the entries of ๐‘„ together with ๐‘„โ€ฒ๐‘–,๐‘— = ๐‘ฆ. The choice of an outer corner and the condition on ๐‘ฆ ensure that ๐‘„โ€ฒ is still a PYT. See Figure 7.11 for an example of a placement. We are now ready to describe (7.28). Consider ๐œ‹ as being given in two-line notation (1.7) 1 2 ... ๐‘› . ๐œ‹= ๐œ‹1 ๐œ‹2 . . . ๐œ‹๐‘› We will construct a sequence of pairs of tableaux (7.29)

(๐‘ƒ0 , ๐‘„0 ) = (โˆ…, โˆ…), (๐‘ƒ1 , ๐‘„1 ), (๐‘ƒ2 , ๐‘„2 ), . . . , (๐‘ƒ๐‘› , ๐‘„๐‘› ) = (๐‘ƒ, ๐‘„)

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

242

7. Counting with Symmetric Functions

๐‘„= 1 4

2

6

7

๐‘„โ€ฒ = 1 4

5

2

6

7

5

8

8 10

9

9

Figure 7.11. The result ๐‘„โ€ฒ of placing 10 at (3, 2) in ๐‘„

by starting with the empty pair and then letting ๐‘ƒ๐‘˜ ๐‘„๐‘˜

= ๐‘Ÿ๐œ‹๐‘˜ (๐‘ƒ๐‘˜โˆ’1 ), = place ๐‘˜ in ๐‘„ ๐‘˜โˆ’1 at the cell where ๐‘Ÿ๐œ‹๐‘˜ terminates,

for ๐‘˜ = 1, 2, . . . , ๐‘›. We then let (๐‘ƒ, ๐‘„) = (๐‘ƒ๐‘› , ๐‘„๐‘› ). Note that by construction we have sh ๐‘ƒ๐‘˜ = sh ๐‘„ ๐‘˜ for all ๐‘˜. A complete example is worked out in Figure 7.12 where the elements of the lower line of ๐œ‹ as well as their counterparts in the ๐‘ƒ๐‘˜ are set in bold. If RS(๐œ‹) = (๐‘ƒ, ๐‘„), we also write ๐‘ƒ(๐œ‹) = ๐‘ƒ and call ๐‘ƒ the ๐‘ƒ-tableau or insertion tableau of ๐œ‹. Similarly we use the notation ๐‘„(๐œ‹) = ๐‘„ for the ๐‘„-tableau of ๐œ‹ which is also called the recording tableau. We now come to our main theorem about this procedure. Theorem 7.5.2. The map RS

๐œ‹ โ†ฆ (๐‘ƒ, ๐‘„) is a bijection between permutations ๐œ‹ โˆˆ ๐”–๐‘› and pairs (๐‘ƒ, ๐‘„) of SYT of the same shape ๐œ† โŠข ๐‘›. Proof. It suffices to construct the inverse. This will be done by reversing the algorithm step by step. So we will build the sequence (7.29) backwards, starting from (๐‘ƒ๐‘› , ๐‘„๐‘› ) = (๐‘ƒ, ๐‘„) and, in the process, recover ๐œ‹. Assume that we have reached (๐‘ƒ๐‘˜ , ๐‘„ ๐‘˜ ) and let (๐‘–, ๐‘—) be the cell containing ๐‘˜ in ๐‘„ ๐‘˜ . To obtain ๐‘„ ๐‘˜โˆ’1 we merely erase ๐‘˜ from ๐‘„ ๐‘˜ . As for finding ๐‘ƒ๐‘˜โˆ’1 and ๐œ‹๐‘˜ , we note that (๐‘–, ๐‘—) must have been the cell at which the insertion into ๐‘ƒ๐‘˜โˆ’1 terminated. So we use the following deletion procedure to undo this insertion. SR1 Let ๐‘ฅ be the (๐‘–, ๐‘—) entry of ๐‘ƒ๐‘˜ and erase it from ๐‘ƒ๐‘˜ . Set ๐‘… โ‰” the (๐‘– โˆ’ 1)st row of ๐‘ƒ๐‘˜ . SR2 While ๐‘… is not the zeroth row of ๐‘ƒ๐‘˜ , let ๐‘ฆ be the rightmost element of ๐‘… smaller than ๐‘ฅ and replace ๐‘ฆ by ๐‘ฅ in ๐‘ƒ๐‘˜ . Repeat this step with ๐‘… โ‰” the row above ๐‘… and ๐‘ฅ โ‰” ๐‘ฆ. SR3 Now ๐‘… is the zeroth row so let ๐œ‹๐‘˜ = ๐‘ฅ and terminate. It should be clear from the constructions that insertion and deletion are inverses of each other. So we are done. โ–ก In order to generalize (7.27), we will consider two sets of variables ๐ฑ = {๐‘ฅ1 , ๐‘ฅ2 , . . . } and ๐ฒ = {๐‘ฆ1 , ๐‘ฆ2 , . . . }. The next result is called Cauchyโ€™s Identity and it can be found in Littlewoodโ€™s text [58]. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

7.5. The Robinsonโ€“Schenstedโ€“Knuth correspondence

๐œ‹=

1 2 3 ๐Ÿ“ ๐Ÿ ๐Ÿ‘

4 ๐Ÿ”

5 ๐Ÿ’

243

6 ๐Ÿ

7 . ๐Ÿ•

๐๐ค โˆถ โˆ…

๐Ÿ“

๐Ÿ ๐Ÿ“

๐Ÿ ๐Ÿ‘ ๐Ÿ“

๐Ÿ ๐Ÿ‘ ๐Ÿ” ๐Ÿ“

๐Ÿ ๐Ÿ‘ ๐Ÿ’ ๐Ÿ“ ๐Ÿ”

๐Ÿ ๐Ÿ‘ ๐Ÿ’ ๐Ÿ ๐Ÿ” ๐Ÿ“

๐Ÿ ๐Ÿ‘ ๐Ÿ’ ๐Ÿ• ๐Ÿ ๐Ÿ” ๐Ÿ“

=๐

๐‘„ ๐‘˜ โˆถ โˆ…,

1

1 2

1 3 2

1 3 4 2

1 3 4 2 5

1 3 4 2 5 6

1 3 4 7 2 5 6

=๐‘„

Figure 7.12. The Robinsonโ€“Schensted map

Theorem 7.5.3. We have 1 1 โˆ’ ๐‘ฅ๐‘– ๐‘ฆ ๐‘— ๐‘–,๐‘—โ‰ฅ1

โˆ‘ ๐‘ ๐œ† (๐ฑ)๐‘ ๐œ† (๐ฒ) = โˆ

(7.30)

๐œ†

where the sum is over all partitions ๐œ† of any nonnegative integer. To give a bijective proof of this formula, we must interpret each side as a weightgenerating function. On the left, we clearly have the weight ogf for pairs (๐‘‡, ๐‘ˆ) of semistandard tableaux of the same shape ๐œ† where wt(๐‘‡, ๐‘ˆ) = ๐ฑ๐‘ˆ ๐ฒ๐‘‡ . For the right-hand side, consider the set Mat of all infinite matrices ๐‘€ with rows and columns indexed by โ„™, entries in โ„•, and only finitely many entries nonzero. A matrix ๐‘€ โˆˆ Mat is shown in the upper-left corner of Figure 7.13 where all entries not shown are zero. Weight these matrices by wt ๐‘€ = โˆ (๐‘ฅ๐‘– ๐‘ฆ๐‘— )๐‘€๐‘–,๐‘— . ๐‘–,๐‘—โ‰ฅ1

Our example matrix has weight wt ๐‘€ = (๐‘ฅ1 ๐‘ฆ2 )2 (๐‘ฅ1 ๐‘ฆ3 )3 (๐‘ฅ2 ๐‘ฆ1 )(๐‘ฅ2 ๐‘ฆ2 )2 (๐‘ฅ3 ๐‘ฆ1 )(๐‘ฅ3 ๐‘ฆ3 ) = ๐‘ฅ15 ๐‘ฅ23 ๐‘ฅ32 ๐‘ฆ21 ๐‘ฆ42 ๐‘ฆ43 .

โŽก โŽข ๐‘€=โŽข โŽข โŽข โŽฃ

0 2 3 0 1 2 0 0 1 0 1 0 0 0 0 0 โ‹ฎ โ‹ฎ โ‹ฎ โ‹ฎ

โ‹ฏ โ‹ฏ โ‹ฏ โ‹ฏ

โŽค โŽฅ โŽฅโ†ฆ๐œ‹= 1 1 2 2 โŽฅ โŽฅ โŽฆ 1 RSK โ†ฆ (๐‘‡, ๐‘ˆ) = ( 2 3

1 3

1 3

1 3

2 1

2 2

2 2

3 1

3 3

1 2 2 3 3 , 1 1 1 1 1 3 2 3 2 2 2 ) 3

Figure 7.13. The Robinsonโ€“Schenstedโ€“Knuth map

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

244

7. Counting with Symmetric Functions

So, by the Sum and Product Rules for weight ogfs 1 . 1 โˆ’ ๐‘ฅ๐‘– ๐‘ฆ ๐‘— ๐‘–,๐‘—โ‰ฅ1

โˆ‘ wt ๐‘€ = โˆ โˆ‘ (๐‘ฅ๐‘– ๐‘ฆ๐‘— )๐‘˜ = โˆ ๐‘€โˆˆMat

๐‘–,๐‘—โ‰ฅ1 ๐‘˜โ‰ฅ0

Thus we need a weight-preserving bijection RSK

(7.31)

๐‘€ โ†ฆ (๐‘‡, ๐‘ˆ)

between matrices ๐‘€ โˆˆ Mat and pairs (๐‘‡, ๐‘ˆ) โˆˆ SSYT(๐œ†) ร— SSYT(๐œ†) as ๐œ† varies over all partitions. Such a map was given by Knuth [49]. It will be convenient to reinterpret the elements of Mat as two-line arrays. Given ๐‘€ โˆˆ Mat, we create an array ๐œ‹ such that ๐‘€๐‘–,๐‘— = the number of times a column

๐‘– occurs in ๐œ‹, ๐‘—

and the columns are arranged in lexicographic order with the top row taking precedence. The two-line array ๐œ‹ associated with the matrix in Figure 7.13 is displayed at the top right. We can now define the map (7.31). Given ๐‘€, construct its two-line array ๐œ‹. Now, starting with the empty tableau, insert the elements of the lower row of ๐œ‹ sequentially to form ๐‘‡ using exactly the same rules RS1โ€“RS3 as before. Note that this algorithm never used the assumption that ๐‘ฅ โˆ‰ ๐‘ƒ and so it applies equally well to semistandard tableaux. As one does the insertions, one places the corresponding elements of the upper row of ๐œ‹ in ๐‘ˆ so that the two tableaux always have the same shape. Figure 7.13 displays the final output of this algorithm on the bottom line. The reader should now be able to fill in the details of the proof of the following theorem. Theorem 7.5.4. The map RSK

๐‘€ โ†ฆ (๐‘‡, ๐‘ˆ) is a weight-preserving bijection between matrices ๐‘€ โˆˆ Mat and pairs (๐‘‡, ๐‘ˆ) of SSYT such that sh ๐‘‡ = st ๐‘ˆ. โ–ก We will use the notation RSK(๐‘€) = RSK(๐œ‹) = (๐‘‡, ๐‘ˆ) where ๐œ‹ is the two-line array corresponding to ๐‘€.

7.6. Longest increasing and decreasing subsequences One of Schenstedโ€™s motivations [81] for introducing the algorithm which bears his name was to study the lengths of longest increasing and decreasing subsequences of a permutation. He proved that these quantities were given by the length of the first row and the length of the first column, respectively, of the associated tableaux. In this section, we will prove his result. Along the way we will see what effect reversing a sequence has on its insertion tableau. Consider a permutation ๐œ‹ = ๐œ‹1 ๐œ‹2 . . . ๐œ‹๐‘› โˆˆ ๐”–๐‘› . Then an increasing subsequence of ๐œ‹ of length ๐‘™ is ๐œ‹๐‘–1 < ๐œ‹๐‘–2 < โ‹ฏ < ๐œ‹๐‘–๐‘™ where ๐‘–1 < ๐‘–2 < โ‹ฏ < ๐‘–๐‘™ . A decreasing subsequence is defined similarly with the inequalities among the ๐œ‹๐‘–๐‘— reversed. We let lis ๐œ‹ = length of a longest increasing subsequence of ๐œ‹ Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

7.6. Longest increasing and decreasing subsequences

245

and lds ๐œ‹ = length of a longest decreasing subsequence of ๐œ‹. If ๐œ‹ = 5236417 is the permutation in Figure 7.12, then ๐œ‹ has increasing subsequences 2347 and 2367 of length 4 and none longer, so lis(๐œ‹) = 4. Similarly, lds(๐œ‹) = 3 because of the subsequence 531, among others. The reader will notice from Figure 7.12 that the first row of the insertion tableau ๐‘ƒ (or of the recording tableau ๐‘„) has length 4 = lis(๐œ‹) and the length of the first column is 3 = lds ๐œ‹. This is always the case. RS

Theorem 7.6.1. If ๐œ‹ โ†ฆ (๐‘ƒ, ๐‘„) with sh ๐‘ƒ = sh ๐‘„ = ๐œ†, then lis ๐œ‹ = ๐œ†1 . Proof. Let ๐‘ƒ๐‘˜โˆ’1 be the tableau formed after inserting ๐œ‹1 . . . ๐œ‹๐‘˜โˆ’1 . We claim that if ๐œ‹๐‘˜ enters ๐‘ƒ๐‘˜โˆ’1 in column ๐‘—, then the length of a longest increasing subsequence of ๐œ‹ ending with ๐œ‹๐‘˜ is ๐‘—. Note that the claim proves the theorem since after inserting all of ๐œ‹ we will have an increasing sequence of length ๐œ†1 ending at the element ๐‘ƒ1,๐œ†1 . And there is no longer subsequence since there is no element of ๐‘ƒ in cell (1, ๐œ†1 + 1). To prove the claim, we induct on ๐‘˜, where the case ๐‘˜ = 1 is trivial. For the induction step, suppose ๐‘ฅ is the element in cell (1, ๐‘— โˆ’ 1) of ๐‘ƒ๐‘˜โˆ’1 . Then there is an increasing subsequence ๐œŽ of ๐œ‹1 . . . ๐œ‹๐‘˜โˆ’1 of length ๐‘— โˆ’ 1 ending in ๐‘ฅ. Since ๐œ‹๐‘˜ entered ๐‘ƒ๐‘˜โˆ’1 in a column to the right of ๐‘ฅ we must have ๐‘ฅ < ๐œ‹๐‘˜โˆ’1 by RS2 and RS3. It follows that the concatenation ๐œŽ๐œ‹๐‘˜ is an increasing subsequence of ๐œ‹ of length ๐‘— ending in ๐œ‹๐‘˜ . To show that ๐‘— is the length of a longest such subsequence suppose, towards a contradiction, that ๐œ๐œ‹๐‘˜ is increasing of length greater than ๐‘—. Let ๐‘ฆ be the last element of ๐œ. Then, by induction, when ๐‘ฆ was inserted it entered in a column ๐‘—โ€ฒ โ‰ฅ ๐‘—. Since ๐‘ฆ < ๐œ‹๐‘˜ and rows increase, the element in cell (1, ๐‘—) just after ๐‘ฆโ€™s insertion must be less than ๐œ‹๐‘˜ . And since elements only bump elements larger than themselves, the element in cell (1, ๐‘—) in ๐‘ƒ๐‘˜โˆ’1 must still be smaller than ๐œ‹๐‘˜ . But this contradicts the fact that ๐œ‹๐‘˜ enters in column ๐‘— since it must bump an element larger than itself. โ–ก Note that the previous proof did not show that the first row of ๐‘ƒ is actually an increasing subsequence of ๐œ‹. In fact, this assertion is false as can be seen in Figure 7.12. To prove our suspicion about lds ๐œ‹, we need to do insertion by columns. Define column insertion of ๐‘ฅ โˆ‰ ๐‘ƒ, where ๐‘ƒ is a partial tableau, using RS1โ€“RS3 but with โ€œrowโ€ replaced by โ€œcolumnโ€ everywhere and โ€œleftmostโ€ by โ€œuppermostโ€. Denote the result of column insertion by ๐‘๐‘ฅ (๐‘ƒ). Amazingly, the row and column operators commute. Lemma 7.6.2. Let ๐‘ƒ be a partial tableau and ๐‘ฅ, ๐‘ฆ distinct positive integers with ๐‘ฅ, ๐‘ฆ โˆ‰ ๐‘ƒ. Then ๐‘๐‘ฆ ๐‘Ÿ๐‘ฅ (๐‘ƒ) = ๐‘Ÿ๐‘ฅ ๐‘๐‘ฆ (๐‘ƒ). Proof. Let ๐‘š = max({๐‘ฅ, ๐‘ฆ} โŠŽ ๐‘ƒ). Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

246

7. Counting with Symmetric Functions

๐‘ƒ=

๐‘ฅ

๐‘ฅ

๐‘š

๐‘š Figure 7.14. The case ๐‘ฆ = ๐‘š in the proof of Lemma 7.6.2

Note that by RS2 and RS3, ๐‘š cannot bump any element during the insertion process. There are two cases depending on which set ๐‘š comes from. Case 1: ๐‘ฆ = ๐‘š. (The case ๐‘ฅ = ๐‘š is similar.) Represent ๐‘ƒ schematically as on the left in Figure 7.14. Since ๐‘š is the maximum element, ๐‘๐‘š will insert ๐‘š at the end of the first column of whatever tableau to which the operator is applied. Suppose ๐‘ฅ is the last element to be bumped during the insertion ๐‘Ÿ๐‘ฅ (๐‘ƒ), and suppose ๐‘ฅ comes to rest in cell ๐‘ข. If ๐‘ข is at the end of the first column, then it is easy to check that ๐‘๐‘š ๐‘Ÿ๐‘ฅ (๐‘ƒ) and ๐‘Ÿ๐‘ฅ ๐‘๐‘š (๐‘ƒ) are both the middle diagram in Figure 7.14. Similarly, if ๐‘ข is not at the end of the first column, then both insertions result in the diagram on the right in Figure 7.14. Case 2: ๐‘š โˆˆ ๐‘ƒ. We induct on #๐‘ƒ. The case when #๐‘ƒ = 1 is easy to check. Let ๐‘ƒ = ๐‘ƒ โˆ’ {๐‘š}, that is, ๐‘ƒ with ๐‘š erased from its cell. Using the fact that ๐‘š never bumps another element as well as induction gives ๐‘๐‘ฆ ๐‘Ÿ๐‘ฅ (๐‘ƒ) โˆ’ {๐‘š} = ๐‘๐‘ฆ ๐‘Ÿ๐‘ฅ (๐‘ƒ) = ๐‘Ÿ๐‘ฅ ๐‘๐‘ฆ (๐‘ƒ) = ๐‘Ÿ๐‘ฅ ๐‘๐‘ฆ (๐‘ƒ) โˆ’ {๐‘š}. So to finish the proof, we need to show that ๐‘š is in the same position in both ๐‘๐‘ฆ ๐‘Ÿ๐‘ฅ (๐‘ƒ) and ๐‘Ÿ๐‘ฅ ๐‘๐‘ฆ (๐‘ƒ). Let ๐‘ฅ be the last element displaced during ๐‘Ÿ๐‘ฅ (๐‘ƒ) and let ๐‘ข be the cell it occupies at the end of the insertion. Similarly define ๐‘ฆ and ๐‘ฃ for ๐‘๐‘ฅ (๐‘ƒ). We now have two subcases depending on the relative locations of ๐‘ข and ๐‘ฃ. Subcase 2a: ๐‘ข = ๐‘ฃ. The first two schematic diagrams in Figure 7.15 illustrate ๐‘Ÿ๐‘ฅ (๐‘ƒ) and ๐‘๐‘ฆ (๐‘ƒ) in this case. If (1, ๐‘—1 ), (2, ๐‘—2 ), . . . , (๐‘˜, ๐‘— ๐‘˜ ) are the cells whose elements change during a row insertion, then it is easy to prove that ๐‘—1 โ‰ฅ ๐‘—2 โ‰ฅ โ‹ฏ โ‰ฅ ๐‘— ๐‘˜ . So during ๐‘Ÿ๐‘ฅ (๐‘ƒ) the only columns which are disturbed are those weakly right of the column of ๐‘ข. Similarly, the insertion ๐‘๐‘ฆ (๐‘ƒ) only changes rows which are weakly below the row of ๐‘ข. It follows that the insertion paths for ๐‘Ÿ๐‘ฅ and ๐‘๐‘ฆ in either order do not intersect until they come to ๐‘ข. If ๐‘ฅ < ๐‘ฆ (the case ๐‘ฆ < ๐‘ฅ is similar), then ๐‘๐‘ฆ ๐‘Ÿ๐‘ฅ (๐‘ƒ) and ๐‘Ÿ๐‘ฅ ๐‘๐‘ฆ (๐‘ƒ) will both be as in the last diagram in Figure 7.15. Note also that ๐‘ฅ and ๐‘ฆ must be in the same column since this is clearly true for ๐‘๐‘ฆ ๐‘Ÿ๐‘ฅ (๐‘ƒ). Now if ๐‘š was not in cell ๐‘ข in ๐‘ƒ, then it will not be bumped Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

7.6. Longest increasing and decreasing subsequences

๐‘Ÿ๐‘ฅ (๐‘ƒ) =

๐‘๐‘ฆ (๐‘ƒ) =

๐‘ฅ

247

๐‘ฆ

๐‘ฅ ๐‘ฆ

Figure 7.15. The subcase ๐‘ข = ๐‘ฃ in the proof of Lemma 7.6.2

by either insertion and so will remain in its cell. If ๐‘š is in cell ๐‘ข, then it is easy to check that it will be bumped into the column just to the right of ๐‘ข in both orders of insertion. This completes this subcase. Subcase 2b: ๐‘ข โ‰  ๐‘ฃ. This subcase is taken care of using arguments similar to those in the rest of the proof, so it is left as an exercise. โ–ก If ๐œ‹ = ๐œ‹1 ๐œ‹2 . . . ๐œ‹๐‘› โˆˆ ๐”–๐‘› , then its reversal (as defined in Exercise 37(a) of Chapter 1) is ๐œ‹๐‘Ÿ = ๐œ‹๐‘› ๐œ‹๐‘›โˆ’1 . . . ๐œ‹1 . The insertion tableaux of ๐œ‹ and ๐œ‹๐‘Ÿ are intimately related. Theorem 7.6.3. If ๐‘ƒ(๐œ‹) = ๐‘ƒ, then ๐‘ƒ(๐œ‹๐‘Ÿ ) = ๐‘ƒ ๐‘ก where ๐‘ก denotes transpose. Proof. Clearly, inserting a single element into an empty tableau gives the same result whether it be by rows or columns. Using this and the previous lemma repeatedly ๐‘ƒ(๐œ‹๐‘Ÿ ) = ๐‘Ÿ๐œ‹1 โ‹ฏ ๐‘Ÿ๐œ‹๐‘›โˆ’1 ๐‘Ÿ๐œ‹๐‘› (โˆ…) = ๐‘Ÿ๐œ‹1 โ‹ฏ ๐‘Ÿ๐œ‹๐‘›โˆ’1 ๐‘๐œ‹๐‘› (โˆ…) = ๐‘๐œ‹๐‘› ๐‘Ÿ๐œ‹1 โ‹ฏ ๐‘Ÿ๐œ‹๐‘›โˆ’1 (โˆ…) โ‹ฎ = ๐‘๐œ‹๐‘› ๐‘๐œ‹๐‘›โˆ’1 โ‹ฏ ๐‘๐œ‹1 (โˆ…) = ๐‘ƒ๐‘ก โ–ก

which is the conclusion we seek. We can now characterize lds(๐œ‹) in terms of the shape of its output tableaux. RS

Corollary 7.6.4. If ๐œ‹ โ†ฆ (๐‘ƒ, ๐‘„) with sh ๐‘ƒ = sh ๐‘„ = ๐œ†, then lds ๐œ‹ = ๐œ†๐‘ก1 , where ๐œ†๐‘ก is the transpose of ๐œ†. Proof. Reversing a permutation interchanges increasing and decreasing subsequences so that lds ๐œ‹ = lis ๐œ‹๐‘Ÿ . By Theorem 7.6.1, lis ๐œ‹๐‘Ÿ is the length of the first row of ๐‘ƒ(๐œ‹๐‘Ÿ ). And ๐‘ƒ(๐œ‹๐‘Ÿ ) = ๐‘ƒ ๐‘ก by Theorem 7.6.3. So lds ๐œ‹ is the length of the first column of ๐‘ƒ, as desired. โ–ก Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

248

7. Counting with Symmetric Functions

We note that Greene [34] has proved the following extension of Schenstedโ€™s theorem. Theorem 7.6.5. Let lis๐‘˜ (๐œ‹) be the longest length of a subsequence of ๐œ‹ which is a union RS

of ๐‘˜ disjoint increasing subsequences. If ๐œ‹ โ†ฆ (๐‘ƒ, ๐‘„) with sh ๐‘ƒ = sh ๐‘„ = ๐œ†, then lis๐‘˜ (๐œ‹) = ๐œ†1 + ๐œ†2 + โ‹ฏ + ๐œ†๐‘˜ and similarly for decreasing subsequences. Interestingly, there does not seem to be an easy interpretation of the individual ๐œ†๐‘– in the shape of the output tableaux. For example, if we consider ๐œ‹ = 247951368, then ๐‘ƒ(๐œ‹) = 1 2

3

5

4

9

6

8 .

7 So ๐œ†1 + ๐œ†2 = 5 + 3 = 8 and 2479 โŠŽ 1368 is a union of two increasing subsequences of ๐œ‹ and is of length 8. But one can check that there is no length 8 subsequence which is a disjoint union of two increasing subsequences of lengths 5 and 3.

7.7. Differential posets In this section we will give a second proof of (7.27) based on properties of Youngโ€™s lattice, ๐‘Œ . This technique can be generalized to a wider class of posets which were introduced and further studied by Stanley [88, 90]. These posets are called differential because of an identity which they satisfy. To connect the summation side of (7.27) with ๐‘Œ , we will use a simple bijection between standard Young tableaux of shape ๐œ† and saturated โˆ…โ€“๐œ† chains in ๐‘Œ . Specifically, a ๐‘‡ โˆˆ SYT(๐œ†) where ๐œ† โŠข ๐‘› will be associated with the chain ๐ถ โˆถ โˆ… = ๐œ†0 โ‹– ๐œ†1 โ‹– โ‹ฏ โ‹– ๐œ†๐‘› = ๐œ† where ๐œ†๐‘˜ is the shape of the subtableau of ๐‘‡ containing the elements [๐‘˜] for 0 โ‰ค ๐‘˜ โ‰ค ๐‘›. An example will be found in Figure 7.16. To go the other way, given a chain ๐ถ, we define ๐‘‡ to be the tableau which has ๐‘˜ in the unique cell of the skew partition ๐œ†๐‘˜ /๐œ†๐‘˜โˆ’1 where skew partitions were defined in (3.7). It is easy to see that these two maps are inverses of each other. From this discussion, it should be clear that (๐‘“๐œ† )2 is the number of pairs of saturated โˆ…โ€“๐œ† chains in ๐‘Œ . In order to work with this observation, we will use a common technique for turning sets into vector spaces. If ๐‘‹ is a set, then consider the set of finite formal linear combinations โ„‚๐‘‹ = { โˆ‘ ๐‘๐‘ฅ ๐‘ฅ โˆฃ ๐‘๐‘ฅ โˆˆ โ„‚ for all ๐‘ฅ and only finitely many ๐‘๐‘ฅ โ‰  0} .

(7.32)

๐‘ฅโˆˆ๐‘‹

๐‘‡=

1

3

2

4

๐ถโˆถ

โˆ…

โ‹–

โ‹–

โ‹–

โ‹–

โ‹–

5 Figure 7.16. The bijection between SYT and saturated chains in Youngโ€™s lattice

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

7.7. Differential posets

249

+ ๐ท(

)

=

๐‘ˆ(

)

=

+

+

Figure 7.17. The down and up operators in Youngโ€™s lattice

Now โ„‚๐‘‹ is a vector space with vector addition and scalar multiplication given by โˆ‘ ๐‘๐‘ฅ ๐‘ฅ + โˆ‘ ๐‘‘๐‘ฅ ๐‘ฅ = โˆ‘(๐‘๐‘ฅ + ๐‘‘๐‘ฅ )๐‘ฅ, ๐‘ฅ

๐‘ฅ

๐‘ฅ

๐‘ โˆ‘ ๐‘๐‘ฅ ๐‘ฅ = โˆ‘ ๐‘๐‘๐‘ฅ ๐‘ฅ. ๐‘ฅ

๐‘ฅ

Note that ๐‘‹ is a basis for โ„‚๐‘‹. We will define two linear operators on โ„‚๐‘Œ . The down operator ๐ท โˆถ โ„‚๐‘Œ โ†’ โ„‚๐‘Œ is defined by ๐ท(๐œ†) = โˆ‘ ๐œ†โˆ’ ๐œ†โˆ’ โ‹–๐œ†

and linear extension. An example is given in Figure 7.17. It will be useful to think of ๐ท(๐œ†) as the sum of all partitions which can be reached by taking a walk of length one downward from ๐œ† in ๐‘Œ viewed as a graph. Note that ๐ท(โˆ…) is the empty sum so that ๐ท(โˆ…) = 0, the zero vector. Similarly, the up operator ๐‘ˆ โˆถ โ„‚๐‘Œ โ†’ โ„‚๐‘Œ is ๐‘ˆ(๐œ†) = โˆ‘ ๐œ†+ . ๐œ†+ โ‹—๐œ†

Again, Figure 7.17 contains an example and a similar walk interpretation holds. We claim that (7.33)

๐ท๐‘› ๐‘ˆ ๐‘› (โˆ…) = ( โˆ‘ (๐‘“๐œ† )2 )โˆ…. ๐œ†โŠข๐‘›

Indeed, the coefficient of ๐œ† โŠข ๐‘› in ๐‘ˆ ๐‘› (โˆ…) is the number of walks from โˆ… to ๐œ† in ๐‘Œ which always go up. But such a walk is just a saturated โˆ…โ€“๐œ† chain so that, by the previous bijection, ๐‘ˆ ๐‘› (โˆ…) = โˆ‘ ๐‘“๐œ† ๐œ†. ๐œ†โŠข๐‘›

By the same token ๐ท๐‘› ๐œ† = ๐‘“๐œ† โˆ… since walks which always go down also follow saturated chains. So applying ๐ท๐‘› to the expression for ๐‘ˆ ๐‘› (โˆ…) and using linearity gives the desired equality. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

250

7. Counting with Symmetric Functions

To make use of (7.33), we need a closer investigation of the structure of ๐‘Œ . Say that ๐œ† โˆˆ ๐‘Œ covers ๐‘˜ elements if #{๐œ†โˆ’ โˆฃ ๐œ†โˆ’ โ‹– ๐œ†} = ๐‘˜. Similarly define the phrase โ€œis covered by ๐‘˜ elementsโ€. Also say that ๐œ†, ๐œ‡ โˆˆ ๐‘Œ cover ๐‘™ elements if #{๐œˆ โˆฃ ๐œˆ โ‹– ๐œ†, ๐œ‡} = ๐‘™ and ditto for being covered. Proposition 7.7.1. The poset ๐‘Œ has the following two properties for all distinct ๐œ†, ๐œ‡ โˆˆ ๐‘Œ . (a) ๐œ† covers ๐‘˜ elements if and only if it is covered by ๐‘˜ + 1 elements. (b) ๐œ†, ๐œ‡ cover ๐‘™ elements if and only if they are covered by ๐‘™ elements. In this case ๐‘™ โ‰ค 1. Proof. (a) It suffices to prove the forward direction since then the number of elements which cover ๐œ† is uniquely determined by the number which it covers. The elements which ๐œ† covers are precisely those obtained by removing an inner corner of ๐œ†. And those which cover ๐œ† are the partitions obtained by adding an outer corner to ๐œ†. But inner corners and outer corners alternate along the southeast boundary of ๐œ†, beginning with an outer corner at the end of the first row and ending with an outer corner at the end of the first column. The result follows. (b) Again, we only need to prove the forward implication. There are two cases. If there is an element ๐œˆ โ‹– ๐œ†, ๐œ‡, then, since ๐‘Œ is ranked, it must be rk ๐œ† = rk ๐œ‡ = ๐‘› for some ๐‘› and rk ๐œˆ = ๐‘› โˆ’ 1. Since ๐‘Œ is a lattice, it follows that ๐œˆ = ๐œ† โˆง ๐œ‡ = ๐œ† โˆฉ ๐œ‡ is unique and so ๐‘™ = 1. Also |๐œ† โˆฉ ๐œ‡| = |๐œˆ| = ๐‘› โˆ’ 1 so that |๐œ† โˆจ ๐œ‡| = |๐œ† โˆช ๐œ‡| = ๐‘› + 1. It follows that ๐œ†, ๐œ‡ are covered by a unique element, namely ๐œ† โˆจ ๐œ‡. If there is no element covered by both ๐œ†, ๐œ‡, then similar considerations show that no element covers ๐œ†, ๐œ‡. We leave this verification to the reader. โ–ก We can translate this result in terms of the down and up operators. Proposition 7.7.2. The operators ๐ท, ๐‘ˆ on ๐‘Œ satisfy (7.34)

๐ท๐‘ˆ โˆ’ ๐‘ˆ๐ท = ๐ผ

where ๐ผ is the identity map. Proof. By linearity, it suffices to show that this equation is true when applied to a basis element ๐œ† โˆˆ โ„‚๐‘Œ . First consider ๐ท๐‘ˆ(๐œ†). The coefficient of ๐œ‡ in this expression is the number of walks ๐œ† to ๐œ‡ which first go up an edge of ๐‘Œ and then come down an edge (possibly the same one). These are precisely the walks of length 2 going through some element covering both ๐œ† and ๐œ‡. From the previous proposition, we get ๐ท๐‘ˆ(๐œ†) = (๐‘˜ + 1)๐œ† + โˆ‘ ๐œ‡ where ๐‘˜ + 1 elements cover ๐œ† and the sum is over all ๐œ‡ โ‰  ๐œ† such that ๐œ‡, ๐œ† are covered by a common element. In a similar way ๐‘ˆ๐ท(๐œ†) = ๐‘˜๐œ† + โˆ‘ ๐œ‡ where the sum is over the same set of ๐œ‡. Subtracting the two equalities gives (7.34). Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

โ–ก

7.7. Differential posets

251

Note that (7.34) is reminiscent of an identity from calculus. Consider a differentiable function ๐‘“(๐‘ก). Let ๐ท stand for differentiation and let ๐‘ˆ be multiplication by ๐‘ก. Then ๐ท๐‘ˆ(๐‘“(๐‘ก)) = (๐‘ก๐‘“(๐‘ก))โ€ฒ = ๐‘“(๐‘ก) + ๐‘ก๐‘“โ€ฒ (๐‘ก) = ๐ผ(๐‘“(๐‘ก)) + ๐‘ˆ๐ท(๐‘“(๐‘ก)) which is just (7.34) with the negative term moved to the other side of the equation. We will need an extension of (7.34) where ๐‘ˆ is replaced by an arbitrary operator which is a polynomial in ๐‘ˆ. Corollary 7.7.3. For any polynomial ๐‘(๐‘ก) โˆˆ โ„‚[๐‘ก] we have ๐ท๐‘(๐‘ˆ) = ๐‘โ€ฒ (๐‘ˆ) + ๐‘(๐‘ˆ)๐ท where ๐‘โ€ฒ (๐‘ก) is the derivative of ๐‘(๐‘ก). Proof. By linearity it suffices to prove this result for the powers ๐‘ˆ ๐‘› for ๐‘› โ‰ฅ 0. The base case ๐‘› = 0 is easy to check. Assuming the result is true for ๐‘› and then applying (7.34) we obtain ๐ท๐‘ˆ ๐‘›+1 = (๐ท๐‘ˆ ๐‘› )๐‘ˆ = (๐‘›๐‘ˆ ๐‘›โˆ’1 + ๐‘ˆ ๐‘› ๐ท)๐‘ˆ = ๐‘›๐‘ˆ ๐‘› + ๐‘ˆ ๐‘› (๐ผ + ๐‘ˆ๐ท) = (๐‘› + 1)๐‘ˆ ๐‘› + ๐‘ˆ ๐‘›+1 ๐ท โ–ก

as desired. We are now ready to reprove (7.27) which we restate here for ease of reference: โˆ‘ (๐‘“๐œ† )2 = ๐‘›! .

(7.35)

๐œ†โŠข๐‘›

Proof. Because of (7.33), it suffices to show that ๐ท๐‘› ๐‘ˆ ๐‘› (โˆ…) = ๐‘›! โˆ…. We induct on ๐‘›, where the case ๐‘› = 0 is trivial since ๐ท0 ๐‘ˆ 0 = ๐ผ. Applying Corollary 7.7.3, the fact that ๐ทโˆ… = 0, and induction gives ๐ท๐‘› ๐‘ˆ ๐‘› (โˆ…) = ๐ท๐‘›โˆ’1 (๐ท๐‘ˆ ๐‘› )(โˆ…) = ๐ท๐‘›โˆ’1 (๐‘›๐‘ˆ ๐‘›โˆ’1 + ๐‘ˆ ๐‘› ๐ท)(โˆ…) = ๐‘›๐ท๐‘›โˆ’1 ๐‘ˆ ๐‘›โˆ’1 (โˆ…) + 0(โˆ…) = ๐‘›(๐‘› โˆ’ 1)! โˆ… which is what we wished to show.

โ–ก

Stanley generalized these ideas using the following definition. Call a poset ๐‘ƒ differential if it satisfies the following three properties where ๐‘ฅ, ๐‘ฆ are distinct elements of ๐‘ƒ. DP1 ๐‘ƒ is ranked. DP2 If ๐‘ฅ covers ๐‘˜ elements, then it is covered by ๐‘˜ + 1 elements. DP3 If ๐‘ฅ, ๐‘ฆ cover ๐‘™ elements, then they are covered by ๐‘™ elements. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

252

7. Counting with Symmetric Functions

From what we have proved, ๐‘Œ is a differential poset. Another example is given in Exercise 29. In fact, Stanley defined a more general type of poset called ๐‘Ÿ-differential which will be studied in Exercise 30. As with Youngโ€™s lattice, we can show that the parameter ๐‘™ must satisfy ๐‘™ โ‰ค 1. Lemma 7.7.4. If poset ๐‘ƒ satisfies DP1 and DP3, then ๐‘™ โ‰ค 1. Proof. As noted in the proof of Proposition 7.7.1, DP3 implies that its converse is also true. Suppose the lemma is false and pick a pair ๐‘ฅ, ๐‘ฆ with ๐‘™ โ‰ฅ 2. Since ๐‘ƒ is ranked by DP1, we must have rk ๐‘ฅ = rk ๐‘ฆ. Pick the counterexample pair ๐‘ฅ, ๐‘ฆ to be of minimum rank and let ๐‘ฅโ€ฒ , ๐‘ฆโ€ฒ be two of the elements covered by ๐‘ฅ, ๐‘ฆ. But since ๐‘ฅโ€ฒ , ๐‘ฆโ€ฒ are covered by at least two elements, they must cover at least two elements. This contradicts the fact that we took a minimum-rank pair. โ–ก We want to define up and down operators in a differential poset. But to make sure they are well-defined, the sums need to be finite. Lemma 7.7.5. If ๐‘ƒ satisfies DP1 and DP2, then its ๐‘›th rank Rk๐‘› ๐‘ƒ is finite for all ๐‘› โ‰ฅ 0. Proof. We induct on ๐‘›. Since ๐‘ƒ is ranked by DP1, it has a 0ฬ‚ and so the result holds for ๐‘› = 0. Assume the lemma through rank ๐‘›. Now any ๐‘ฅ โˆˆ Rk๐‘› ๐‘ƒ covers at most # Rk๐‘›โˆ’1 ๐‘ƒ elements. So, by DP2, # Rk๐‘›+1 ๐‘ƒ โ‰ค (# Rk๐‘› ๐‘ƒ)(1 + # Rk๐‘›โˆ’1 ๐‘ƒ) โ–ก

which forces Rk๐‘›+1 ๐‘ƒ to be finite. Thus we can define two operators on โ„‚๐‘ƒ by ๐ท(๐‘ฅ) = โˆ‘ ๐‘ฅโˆ’ ๐‘ฅโˆ’ โ‹–๐‘ฅ

and ๐‘ˆ(๐‘ฅ) = โˆ‘ ๐‘ฅ+ . ๐‘ฅ+ โ‹—๐‘ฅ

The proof of the next result is similar enough to that of Proposition 7.7.2 that it is left as an exercise. Proposition 7.7.6. Let ๐‘ƒ be a ranked poset with Rk๐‘› ๐‘ƒ finite for all ๐‘› โ‰ฅ 0. Then ๐‘ƒ is differential โŸบ ๐ท๐‘ˆ โˆ’ ๐‘ˆ๐ท = ๐ผ.

โ–ก

Also, the reader should be able to generalize the operator proof of (7.27) to the setting of differential posets and show the following. Theorem 7.7.7. In any differential poset ๐‘ƒ we have โˆ‘ (๐‘“๐‘ฅ )2 = ๐‘›! ๐‘ฅโˆˆRk๐‘› ๐‘ƒ

ฬ‚ chains. where ๐‘“๐‘ฅ is the number of saturated 0โ€“๐‘ฅ Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

โ–ก

7.8. The chromatic symmetric function

253

The reader might wonder if there is a way to give a bijective proof of this theorem just as the Robinsonโ€“Schensted algorithm provides a bijection for (7.27). This has been done by Fomin as part of his theory of duality of graded graphs [27, 28].

7.8. The chromatic symmetric function Stanley [91] defined a symmetric function associated with graph colorings which generalizes the chromatic polynomial. In this section we will prove some of his results about this function, including expressions for its expansion in the monomial and power sum bases for Sym. Let ๐บ be a graph with vertex set ๐‘‰ = {๐‘ฃ 1 , . . . , ๐‘ฃ ๐‘› }. Consider a coloring ๐‘ โˆถ ๐‘‰ โ†’ โ„™ of ๐บ using the positive integers as color set. Then ๐‘ has monomial ๐ฑ๐‘ = ๐‘ฅ๐‘(๐‘ฃ1 ) ๐‘ฅ๐‘(๐‘ฃ2 ) โ‹ฏ ๐‘ฅ๐‘(๐‘ฃ๐‘› ) .

(7.36)

For example, if ๐บ is the graph on the left in Figure 7.18 and ๐‘ is the coloring on the right, then ๐ฑ๐‘ = ๐‘ฅ๐‘(แต†) ๐‘ฅ๐‘(๐‘ฃ) ๐‘ฅ๐‘(๐‘ค) ๐‘ฅ๐‘(๐‘ง) = ๐‘ฅ1 ๐‘ฅ22 ๐‘ฅ4 . Now define the chromatic symmetric function of ๐บ to be ๐‘‹(๐บ) = ๐‘‹(๐บ; ๐ฑ) =

โˆ‘ ๐ฑ๐‘ ๐‘ โˆถ ๐‘‰ โ†’โ„™

where the sum is over all proper colorings ๐‘ โˆถ ๐‘‰ โ†’ โ„™. To illustrate, let us return to the graph of Figure 7.18. Because ๐บ contains a triangle, we must use three or four colors for a proper coloring. If we use four different colors, then this will give rise to a monomial ๐‘ฅ๐‘– ๐‘ฅ๐‘— ๐‘ฅ๐‘˜ ๐‘ฅ๐‘™ for some distinct ๐‘–, ๐‘—, ๐‘˜, ๐‘™. And any of the 4! = 24 ways of assigning these colors to the four vertices is proper. Since this count is independent of which four colors we use, the contribution of such colorings to ๐‘‹(๐บ) is 24๐‘š14 . If we use three colors, then one of them must be used twice and so correspond to a monomial ๐‘ฅ๐‘–2 ๐‘ฅ๐‘— ๐‘ฅ๐‘˜ for distinct ๐‘–, ๐‘—, ๐‘˜. One copy of color ๐‘– must go on vertex ๐‘ค and the other can be on ๐‘ข or ๐‘ฅ, giving two choices. The other two colors can be distributed among the remaining two vertices in two ways. So these colorings give a term 4๐‘š212 . In total ๐‘‹(๐บ) = 24๐‘š14 + 4๐‘š212 which the reader will note is a symmetric function. We will now prove that this is always the case, as well as showing a connection with the chromatic polynomial of ๐บ. Proposition 7.8.1. Let ๐บ be a graph with vertex set ๐‘‰. (a) ๐‘‹(๐บ) โˆˆ Sym๐‘› where ๐‘› = #๐‘‰. (b) If we set ๐‘ฅ1 = โ‹ฏ = ๐‘ฅ๐‘ก = 1 and ๐‘ฅ๐‘– = 0 for ๐‘– > ๐‘ก, written ๐ฑ = 1๐‘ก , then ๐‘‹(๐บ; 1๐‘ก ) = ๐‘ƒ(๐บ; ๐‘ก). ๐‘ข

๐‘ฃ

2

4

1

2

๐‘=

๐บ= ๐‘ง

๐‘ค

Figure 7.18. A graph and a coloring using โ„™

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

254

7. Counting with Symmetric Functions

Proof. (a) It is clear that ๐ฑ๐‘ has ๐‘› factors for any coloring ๐‘ so that ๐‘‹(๐บ) is homogeneous of degree ๐‘›. To show that it is symmetric, note that any permutation of the colors of a proper coloring is proper. This means that permuting the subscripts in ๐‘‹(๐บ) leaves it invariant; that is, ๐‘‹(๐บ) is symmetric. (b) The given substitution results in ๐ฑ๐‘ = 1 if ๐‘ only uses colors from [๐‘ก] and ๐ฑ๐‘ = 0 otherwise. So ๐‘‹(๐บ; 1๐‘ก ) is just the number of proper colorings ๐‘ โˆถ ๐‘‰ โ†’ [๐‘ก]. But this was the definition of ๐‘ƒ(๐บ; ๐‘ก). โ–ก Since ๐‘‹(๐บ) is symmetric, we can expand it in terms of various bases for the symmetric functions and see if the coefficients have any nice combinatorial interpretation. We start with the monomial basis. To describe the coefficients, we will need some definitions. If ๐บ = (๐‘‰, ๐ธ) is a graph, then ๐‘Š โŠ† ๐‘‰ is independent or stable if there is no edge of ๐บ between any pair of vertices of ๐‘Š. For example, if ๐บ = ๐‘‡1 as in Figure 1.9, then ๐‘Š = {2, 5, 6} is stable but ๐‘Š = {2, 3, 6} is not because of the edge 36. The reason we care about stable sets is that if ๐‘ is a proper coloring of ๐บ, then the set of all vertices with a given color ๐‘Ÿ, in other words the vertices in ๐‘โˆ’1 (๐‘Ÿ), form a stable set. Similarly, call a partition ๐œŒ = ๐ต1 / . . . /๐ต๐‘˜ of ๐‘‰ independent or stable if each block is. Returning to Figure 1.9, the partition 13/256/4 is stable in ๐‘‡1 . The type of a set partition ๐œŒ = ๐ต1 / . . . /๐ต๐‘˜ is the integer partition ๐œ†(๐œŒ) = (๐œ†1 , . . . , ๐œ†๐‘˜ ) obtained by arranging #๐ต1 , . . . , #๐ต๐‘˜ in weakly decreasing order. To illustrate, ๐œ†(1456/27/38) = (4, 2, 2). For a graph ๐บ, let ๐‘–๐œ† (๐บ) = number of independent partitions of ๐‘‰ of type ๐œ†. For the graph in Figure 7.18 we have ๐‘–14 (๐บ) = 1, ๐‘–212 (๐บ) = 2 and all other ๐‘–๐œ† (๐บ) = 0. Any proper coloring ๐‘ of ๐บ induces a stable partition of ๐‘‰ whose blocks are the nonempty ๐‘โˆ’1 (๐‘Ÿ) for the colors ๐‘Ÿ in the color set. Finally, if ๐œ† = (1๐‘š1 , 2๐‘š1 , . . . , ๐‘›๐‘š๐‘› ) is an integer partition in multiplicity notation, then let ๐œ†! = ๐‘š1 ! ๐‘š2 ! โ‹ฏ ๐‘š๐‘› ! . Theorem 7.8.2. If graph ๐บ has #๐‘‰ = ๐‘›, then ๐‘‹(๐บ) = โˆ‘ ๐‘–๐œ† (๐บ)๐œ†! ๐‘š๐œ† . ๐œ†โŠข๐‘› ๐œ†

๐œ†

Proof. If ๐œ† = (๐œ†1 , . . . , ๐œ†๐‘˜ ), then the coefficient of ๐‘ฅ1 1 โ‹ฏ ๐‘ฅ๐‘˜๐‘˜ is the coefficient of ๐‘š๐œ† since ๐‘‹(๐บ) is symmetric. And the coefficient of this monomial is the number of proper colorings ๐‘ โˆถ ๐‘‰ โ†’ [๐‘˜] where ๐‘– gets used ๐œ†๐‘– times for all ๐‘–. By the discussion preceding this theorem, these colorings can be obtained by taking an independent partition ๐œŒ โŠข ๐‘‰ and then deciding which colors get assigned to which blocks of ๐œŒ. The number of choices for ๐œŒ is ๐‘–๐œ† (๐บ). Now any of the ๐‘š๐‘— colors which are used ๐‘— times can be used on any of the ๐‘š๐‘— blocks of ๐œŒ of size ๐‘—. The number of such assignments is ๐‘š๐‘— ! and this is true for all ๐‘—. This gives the factor of ๐œ†! . โ–ก The expansion of ๐‘‹(๐บ) in the power sum basis will be found by Mรถbius inversion. Let ๐บ = (๐‘‰, ๐ธ) be a graph. Then any spanning subgraph ๐ป can be identified with its set of edges ๐ธ(๐ป) since we have ๐‘‰(๐ป) = ๐‘‰(๐บ). To illustrate, for the graph ๐บ in Figure 3.5 we would identify ๐น1 with the edge set {12, 24} and ๐น2 with the edge set {14, 24}. Given Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

7.8. The chromatic symmetric function

255

๐น โŠ† ๐ธ, we get a partition ๐œŒ(๐น) of the vertex set where a block of ๐œŒ is the set of vertices in a component of the corresponding spanning subgraph. Returning to our example, ๐น1 and ๐น2 both have partition ๐œŒ = 124/3. Let ๐œ†(๐น) = type of the partition ๐œŒ(๐น). In our running example ๐œ†(๐น1 ) = ๐œ†(๐น2 ) = (3, 1). Theorem 7.8.3. If ๐บ = (๐‘‰, ๐ธ) is a graph, then ๐‘‹(๐บ) = โˆ‘ (โˆ’1)#๐น ๐‘ ๐œ†(๐น) ๐นโŠ†๐ธ

where the sum is over all subsets ๐น of the edge set of ๐บ and #๐น is the number of edges in ๐น. Proof. Consider the Boolean algebra ๐ต๐ธ of all subsets of ๐ธ ordered by containment. Given ๐น โŠ† ๐ธ, we define the power series ๐›ผ(๐น) = โˆ‘ ๐ฑ๐‘ ๐‘

where the sum is over all colorings ๐‘ โˆถ ๐‘‰ โ†’ โ„™ such that ๐‘(๐‘ข) = ๐‘(๐‘ฃ) for all ๐‘ข๐‘ฃ โˆˆ ๐น. These are the colorings which are monochromatic on each component of ๐นโ€™s spanning subgraph. So if ๐œŒ(๐น) = ๐ต1 / . . . /๐ต๐‘˜ , then each ๐ต๐‘– can get any of the colors in โ„™. It follows that ๐‘˜

(7.37)

#๐ต๐‘–

๐›ผ(๐น) = โˆ(๐‘ฅ1

#๐ต๐‘–

+ ๐‘ฅ2

+ โ‹ฏ) = ๐‘ ๐œ†(๐น) .

๐‘–=1

Also define ๐›ฝ(๐น) = โˆ‘ ๐ฑ๐‘ ๐‘

where the sum is over all colorings ๐‘ โˆถ ๐‘‰ โ†’ โ„™ such that ๐‘(๐‘ข) = ๐‘(๐‘ฃ) for all ๐‘ข๐‘ฃ โˆˆ ๐น and ๐‘(๐‘ข) โ‰  ๐‘(๐‘ฃ) for ๐‘ข๐‘ฃ โˆˆ ๐ธ โˆ’ ๐น. So these colorings are constant on the components of ๐น but also cannot have any other edge of ๐ธ monochromatically colored. Given any spanning subgraph ๐น โ€ฒ and a coloring ๐‘ appearing in ๐›ผ(๐น โ€ฒ ), one can define a unique spanning subgraph ๐น by letting ๐‘ข๐‘ฃ โˆˆ ๐น if and only if ๐‘(๐‘ข) = ๐‘(๐‘ฃ). By the definition of ๐›ผ it is clear that ๐น โŠ‡ ๐น โ€ฒ . Also, the definition of ๐น implies that ๐‘ is a coloring in the sum for ๐›ฝ(๐น). It follows that ๐›ผ(๐น โ€ฒ ) = โˆ‘ ๐›ฝ(๐น) ๐นโŠ‡๐น โ€ฒ

for all ๐น โ€ฒ โŠ† ๐ธ. Applying Mรถbius inversion (Theorem 5.5.5) as well as (5.6) and (7.37) gives ๐›ฝ(โˆ…) = โˆ‘ ๐œ‡(๐น)๐›ผ(๐น) = โˆ‘ (โˆ’1)#๐น ๐‘ ๐œ†(๐น) . ๐นโˆˆ๐ต๐ธ

๐นโŠ†๐ธ

But ๐›ฝ(โˆ…) is the generating function for all colorings ๐‘‘ โˆถ ๐‘‰ โ†’ โ„™ such that ๐‘‘(๐‘ข) โ‰  ๐‘‘(๐‘ฃ) for ๐‘ข๐‘ฃ โˆˆ ๐ธ, and these are exactly the proper colorings. Thus ๐›ฝ(โˆ…) = ๐‘‹(๐บ) and we are done. โ–ก We end this section with an open question about ๐‘‹(๐บ). To appreciate it, we first prove a result about the chromatic polynomial. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

256

7. Counting with Symmetric Functions

Proposition 7.8.4. Let ๐‘‡ be a graph with #๐‘‰ = ๐‘›. We have that ๐‘‡ is a tree if and only if ๐‘ƒ(๐‘‡; ๐‘ก) = ๐‘ก(๐‘ก โˆ’ 1)๐‘›โˆ’1 . Proof. We will prove the forward direction and leave the reverse implication as an exercise. If ๐‘ฃ โˆˆ ๐‘‰, then we color ๐‘‡ by first coloring ๐‘ฃ, then all the neighbors of ๐‘ฃ, then all the uncolored vertices which are neighbors of neighbors of ๐‘ฃ, etc. The number of ways to color ๐‘ฃ is ๐‘ก. Because ๐‘‡ is connected and acyclic, when coloring each vertex ๐‘ค โˆˆ ๐‘‰ โˆ’ {๐‘ฃ} we will have ๐‘ค adjacent to exactly one already colored vertex. So the number of colors available for ๐‘ค is ๐‘ก โˆ’ 1. The result follows. โ–ก This proposition is sometimes summarized by saying that the chromatic polynomial does not distinguish trees since all trees on ๐‘› vertices have the same polynomial. Stanley asked if the opposite was true for his chromatic symmetric function. Question 7.8.5. If ๐‘‡1 and ๐‘‡2 are nonisomorphic trees, is it true that ๐‘‹(๐‘‡1 ) โ‰  ๐‘‹(๐‘‡2 )? It has been checked for trees with up to 23 vertices that the answer to this question is โ€œyesโ€ and there are other partial results in the literature.

7.9. Cyclic sieving redux Cyclic sieving phenomena are often associated to results in representation theory. In this section we will give a second proof of Theorem 6.6.2 using this approach. To do so, we will assume the reader is familiar with the material in the appendix in this book. We start by presenting a general paradigm for proving a CSP by using group actions. Recall that we start with a set ๐‘‹, a cyclic group ๐บ acting on ๐‘‹, and a polynomial ๐‘“(๐‘ž) โˆˆ โ„•[๐‘ž]. Also, for the rest of this section we let ๐œ” = ๐‘’2๐œ‹๐‘–/๐‘› . The triple (๐‘‹, ๐บ, ๐‘“(๐‘ž)) was said to exhibit the cyclic sieving phenomenon if, for all ๐‘” โˆˆ ๐บ, (7.38)

#๐‘‹ ๐‘” = ๐‘“(๐›พ)

where ๐›พ is a root of unity satisfying ๐‘œ(๐›พ) = ๐‘œ(๐‘”). To interpret the left side of (7.38), consider the permutation representation โ„‚๐‘‹ of ๐บ. Since ๐‘” โˆˆ ๐บ takes each basis element in ๐‘‹ to another basis element, the matrix [๐‘”]๐‘‹ consists of zeros and ones. And there is a one on the diagonal precisely when ๐‘”๐‘ฅ = ๐‘ฅ for ๐‘ฅ โˆˆ ๐‘‹. So this representation has character (7.39)

๐œ’(๐‘”) = tr[๐‘”]๐‘‹ = #๐‘‹ ๐‘” .

As for the right-hand side of (7.38), let โ„Ž be a generator of ๐บ where #๐บ = ๐‘› and let ๐œ” = ๐œ”๐‘› . For ๐‘– โ‰ฅ 0, let ๐‘‰ (๐‘–) be the irreducible ๐บ-module such that the matrix of โ„Ž is [๐œ”๐‘– ] and let its character be ๐œ’(๐‘–) . Suppose ๐‘“ = โˆ‘๐‘–โ‰ฅ0 ๐‘š๐‘– ๐‘ž๐‘– where ๐‘š๐‘– โˆˆ โ„• for all ๐‘–. Since the coefficients are nonnegative integers, we can define a corresponding ๐บ-module ๐‘‰๐‘“ =

โจ

๐‘š๐‘– ๐‘‰ (๐‘–)

๐‘–โ‰ฅ0

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

7.9. Cyclic sieving redux

257

with character ๐œ’๐‘“ . Let ๐‘” = โ„Ž๐‘— and ๐›พ = ๐œ”๐‘— . Now using (A.1) and the fact that the ๐‘‰ (๐‘–) are 1-dimensional, ๐œ’๐‘“ (๐‘”) = โˆ‘ ๐‘š๐‘– ๐œ’(๐‘–) (โ„Ž๐‘— ) = โˆ‘ ๐‘š๐‘– ๐œ”๐‘–๐‘— = ๐‘“(๐œ”๐‘— ) = ๐‘“(๐›พ) ๐‘–โ‰ฅ0

๐‘–โ‰ฅ0

which is the right side of (7.38). Now appealing to Theorem A.1.4, we have proved the following result or Reiner, Stanton, and White [72]. Theorem 7.9.1. The triple (๐‘‹, ๐บ, ๐‘“(๐‘ž)) exhibits the cyclic sieving phenomenon if and only if โ„‚๐‘‹ โ‰… ๐‘‰ ๐‘“ as ๐บ-modules. โ–ก We will now give a second proof that the triple (7.40)

(((

[๐‘›] ๐‘›+๐‘˜โˆ’1 ] ) )), โŸจ(1, 2, . . . , ๐‘›)โŸฉ, [ ๐‘˜ ๐‘˜ ๐‘ž

exhibits the CSP. We will write vectors in boldface to distinguish them from scalars. To use the previous theorem, any ๐บ-module isomorphic to โ„‚๐‘‹ will suffice. So we will construct one whose structure is easy to analyze. Let ๐‘‰ be a vector space of dimension ๐‘› and fix a basis ๐ต = {๐›1 , ๐›2 , . . . , ๐›๐‘› } for ๐‘‰. Let ๐‘ƒ ๐‘˜ (๐ต) be the set of all formal polynomials of degree ๐‘˜ using the elements of ๐ต as variables and the complex numbers as coefficients. Clearly ๐‘ƒ ๐‘˜ (๐ต) is a vector space with basis (7.41)

๐ต โ€ฒ = {๐›๐‘–1 ๐›๐‘–2 โ‹ฏ ๐›๐‘–๐‘˜ โˆฃ 1 โ‰ค ๐‘–1 โ‰ค ๐‘–2 โ‰ค โ‹ฏ โ‰ค ๐‘–๐‘˜ โ‰ค ๐‘›}

In particular, if one takes ๐‘‰ = โ„‚[๐‘›] with the basis ๐ต = {๐ข โˆฃ ๐‘– โˆˆ [๐‘›]}, then we will write ๐‘ƒ ๐‘˜ (๐‘›) for ๐‘ƒ ๐‘˜ (๐ต). To illustrate, ๐‘ƒ2 (3) = {๐‘ 1 ๐Ÿ๐Ÿ + ๐‘ 2 ๐Ÿ๐Ÿ + ๐‘ 3 ๐Ÿ‘๐Ÿ‘ + ๐‘ 4 ๐Ÿ๐Ÿ + ๐‘ 5 ๐Ÿ๐Ÿ‘ + ๐‘ 6 ๐Ÿ๐Ÿ‘ โˆฃ ๐‘ ๐‘– โˆˆ โ„‚ for 1 โ‰ค ๐‘– โ‰ค 6}. One can turn ๐‘ƒ ๐‘˜ (๐‘›) into a ๐บ๐‘› -module for ๐บ๐‘› = โŸจ(1, 2, . . . , ๐‘›)โŸฉ by letting (7.42)

๐‘”(๐ข1 ๐ข2 โ‹ฏ ๐ข๐‘˜ ) = ๐‘”(๐ข1 )๐‘”(๐ข2 ) โ‹ฏ ๐‘”(๐ข๐‘˜ )

for ๐‘” โˆˆ ๐บ๐‘› and extending linearly. It should be clear from the definitions that (7.43)

[๐‘›] โ„‚(( )) โ‰… ๐‘ƒ ๐‘˜ (๐‘›) ๐‘˜

as ๐บ๐‘› -modules. So we will use the latter in establishing the CSP. The advantage of using ๐‘ƒ ๐‘˜ (๐‘›) is that it is also a GL๐‘› -module. In particular, we can define the action of ๐‘” โˆˆ GL๐‘› exactly as in (7.42), where ๐ข๐‘— is thought of as the ๐‘—th coordinate vector and the linear combinations ๐‘”(๐ข1 ), . . . , ๐‘”(๐ข๐‘˜ ) are multiplied together formally to get an element of ๐‘ƒ ๐‘˜ (๐‘›). For example, if ๐‘› = 3, ๐‘˜ = 2, and 1 ๐‘”=[ 3 0

2 4 0

0 0 ], โˆ’1

then ๐‘”(๐Ÿ๐Ÿ‘) = ๐‘”(๐Ÿ)๐‘”(๐Ÿ‘) = (๐Ÿ + 3 ๐Ÿ)(โˆ’๐Ÿ‘) = โˆ’๐Ÿ๐Ÿ‘ โˆ’ 3 ๐Ÿ๐Ÿ‘. It is not hard to show that this is an action. We will also need the fact that if ๐ตฬƒ is any other basis for โ„‚[๐‘›], then the monomials defined by (7.41) (where one replaces each ๐›๐‘–๐‘— by the corresponding element of ๐ต)ฬƒ also form a basis for ๐‘ƒ ๐‘˜ (๐‘›). Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

258

7. Counting with Symmetric Functions

We now compute the character of ๐‘ƒ๐‘˜ (๐‘›). Notice from Corollary A.1.2 that, since ๐บ๐‘› is cyclic, there is a basis ๐ต = {๐›1 , ๐›2 , . . . , ๐›๐‘› } for โ„‚[๐‘›] which diagonalizes [๐‘”] for all ๐‘” โˆˆ ๐บ๐‘› ; say [๐‘”]๐ต = diag(๐‘ฅ1 , ๐‘ฅ2 , . . . , ๐‘ฅ๐‘› ). To compute the action of ๐บ๐‘› in ๐‘ƒ ๐‘˜ (๐‘›), we use the basis ๐ตโ€ฒ in (7.41) and see that ๐‘”(๐›๐‘–1 ๐›๐‘–2 โ‹ฏ ๐›๐‘–๐‘˜ ) = ๐‘”(๐›๐‘–1 )๐‘”(๐›๐‘–2 ) โ‹ฏ ๐‘”(๐›๐‘–๐‘˜ ) = ๐‘ฅ๐‘–1 ๐‘ฅ๐‘–2 โ‹ฏ ๐‘ฅ๐‘–๐‘˜ ๐›๐‘–1 ๐›๐‘–2 โ‹ฏ ๐›๐‘–๐‘˜ . So ๐ตโ€ฒ diagonalizes this action with [๐‘”]๐ตโ€ฒ = diag(๐‘ฅ๐‘–1 ๐‘ฅ๐‘–2 โ‹ฏ ๐‘ฅ๐‘–๐‘˜ โˆฃ 1 โ‰ค ๐‘–1 โ‰ค ๐‘–2 โ‰ค โ‹ฏ โ‰ค ๐‘–๐‘˜ โ‰ค ๐‘›). This gives the character ๐œ’โ€ฒ (๐‘”) =

โˆ‘

๐‘ฅ๐‘–1 ๐‘ฅ๐‘–2 โ‹ฏ ๐‘ฅ๐‘–๐‘˜ = โ„Ž๐‘˜ (๐‘ฅ1 , ๐‘ฅ2 , . . . , ๐‘ฅ๐‘› )

1โ‰ค๐‘–1 โ‰ค๐‘–2 โ‰ค. . .โ‰ค๐‘–๐‘˜ โ‰ค๐‘›

which is just a complete homogeneous symmetric polynomial in the eigenvalues. For example, if ๐‘› = 3 and ๐‘˜ = 2, then we would have a basis ๐ต = {๐š, ๐›, ๐œ} such that [๐‘”]๐ต = diag(๐‘ฅ1 , ๐‘ฅ2 , ๐‘ฅ3 ). So in ๐‘ƒ 2 (3) ๐‘”(๐œ๐œ) = ๐‘ฅ32 ๐œ๐œ, ๐‘”(๐š๐š) = ๐‘ฅ12 ๐š๐š, ๐‘”(๐›๐›) = ๐‘ฅ22 ๐›๐›, ๐‘”(๐š๐›) = ๐‘ฅ1 ๐‘ฅ2 ๐š๐›, ๐‘”(๐š๐œ) = ๐‘ฅ1 ๐‘ฅ3 ๐š๐œ, ๐‘”(๐›๐œ) = ๐‘ฅ2 ๐‘ฅ3 ๐›๐œ, which gives ๐œ’โ€ฒ (๐‘”) = ๐‘ฅ12 + ๐‘ฅ22 + ๐‘ฅ32 + ๐‘ฅ1 ๐‘ฅ2 + ๐‘ฅ1 ๐‘ฅ3 + ๐‘ฅ2 ๐‘ฅ3 . To prove the CSP we will need to relate homogeneous symmetric polynomials to ๐‘žbinomial coefficients. This is done via the principal specialization which sets ๐‘ฅ๐‘– = ๐‘ž๐‘–โˆ’1 for ๐‘– โ‰ฅ 1. Proposition 7.9.2. We have the principal specializations ๐‘’ ๐‘˜ (1, ๐‘ž, . . . , ๐‘ž๐‘›โˆ’1 ) = ๐‘ž( 2 ) [ ๐‘˜

๐‘› ] ๐‘˜ ๐‘ž

and โ„Ž๐‘˜ (1, ๐‘ž, . . . , ๐‘ž๐‘›โˆ’1 ) = [

๐‘›+๐‘˜โˆ’1 ] . ๐‘˜ ๐‘ž

Proof. We will prove the identity for โ„Ž๐‘˜ , leaving the one for ๐‘’ ๐‘˜ as an exercise. From the definition of the complete homogeneous symmetric functions we see that โ„Ž๐‘˜ (1, ๐‘ž, . . . , ๐‘ž๐‘›โˆ’1 ) =

โˆ‘

๐‘ž๐‘—1 ๐‘ž๐‘—2 โ‹ฏ ๐‘ž๐‘—๐‘˜ .

0โ‰ค๐‘—1 โ‰ค๐‘—2 โ‰คโ‹ฏโ‰ค๐‘—๐‘˜ โ‰ค๐‘›โˆ’1

But a sequence ๐‘—1 โ‰ค ๐‘—2 โ‰ค โ‹ฏ โ‰ค ๐‘— ๐‘˜ corresponds to an integer partition ๐œ† obtained by listing the nonzero elements of the sequence in weakly decreasing order. Furthermore ๐‘ž๐‘—1 ๐‘ž๐‘—2 โ‹ฏ ๐‘ž๐‘—๐‘˜ = ๐‘ž|๐œ†| and the bounds on the ๐‘—๐‘– imply that ๐œ† โˆˆ โ„›(๐‘˜, ๐‘› โˆ’ 1), the set of partitions contained in a ๐‘˜ ร— (๐‘› โˆ’ 1) rectangle. The equality now follows from Theorem 3.2.5. โ–ก Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Exercises

259

We are now ready to complete the representation-theoretic demonstration that the triple (7.40) exhibits the cyclic sieving phenomenon. Consider [(1, 2, . . . , ๐‘›)] acting as a linear transformation on โ„‚[๐‘›]. Its characteristic polynomial is ๐‘ฅ๐‘› โˆ’ 1 which has roots 1, ๐œ”, ๐œ”2 , . . . , ๐œ”๐‘›โˆ’1 . So, by Corollary A.1.2, there is a diagonalizing basis ๐ต for the action of ๐บ๐‘› = โŸจ(1, 2, . . . , ๐‘›)โŸฉ with [(1, 2, . . . , ๐‘›)]๐ต = diag(1, ๐œ”, ๐œ”2 , . . . , ๐œ”๐‘›โˆ’1 ). Since any ๐‘” โˆˆ ๐บ๐‘› has the form ๐‘” = (1, 2, . . . , ๐‘›)๐‘— for some ๐‘— and since the generator has been diagonalized, we have [๐‘”]๐ต = diag(1๐‘— , ๐œ”๐‘— , ๐œ”2๐‘— , . . . , ๐œ”(๐‘›โˆ’1)๐‘— ) = diag(1, ๐›พ, ๐›พ2 , . . . , ๐›พ๐‘›โˆ’1 ) where ๐›พ = ๐œ”๐‘– is a primitive ๐‘œ(๐‘”)th root of unity. But the previous theorem and the discussion just preceding it show that ๐œ’โ€ฒ (๐‘”) = โ„Ž๐‘˜ (1, ๐›พ, ๐›พ2 , . . . , ๐›พ๐‘›โˆ’1 ) = [

๐‘›+๐‘˜โˆ’1 ] ๐‘˜ ๐›พ

where ๐œ’โ€ฒ is the character of ๐‘ƒ๐‘˜ (๐‘›). But, by (7.43), ๐‘”

[๐‘›] ๐œ’ (๐‘”) = #(( )) . ๐‘˜ โ€ฒ

Equating the last two displayed equations completes the proof.

Exercises (1) (a) Show that (7.2) satisfies the definition of a group action. (b) Show that Sym is an algebra; that is, it is a vector space which is closed under multiplication of symmetric functions. (2) Prove Proposition 7.1.2(b). (3) Prove Theorem 7.1.3(b). (4) (a) Prove that lexicographic order is a total order on partitions. (b) Prove that adding parts of a partition makes it larger in lexicographic order. (c) Prove that lexicographic order on partitions is a linear extension of dominance order as introduced in Section 1.12. (d) Prove that the lexicographic order inequalities in the proof of Theorem 7.1.3(a) and (b) can be strengthened to dominance order inequalities. (e) Show that part (d) is also true in the statement of Theorem 7.2.2. (5) Show that ๐”–๐‘› is generated by the adjacent transpositions (๐‘–, ๐‘– + 1) for 1 โ‰ค ๐‘– < ๐‘›. Hint: Induct on inv ๐œ‹ for ๐œ‹ โˆˆ ๐”–๐‘› . (6) Consider the proof of Theorem 7.2.3. (a) Show in part (a) that every path family whose associated permutation is not the identity must intersect. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

260

7. Counting with Symmetric Functions

(b) In part (a), provide a description of the inverse of the map from lattice path families to SSYT. Be sure to prove that it is well-defined and indeed an inverse to the forward map. (c) Prove part (b). (7) Verify that the real sequence ๐‘Ž0 , . . . , ๐‘Ž๐‘› is log-concave if and only if ๐‘Ž0 /๐‘Ÿ, . . . , ๐‘Ž๐‘› /๐‘Ÿ is, where ๐‘Ÿ โˆˆ โ„ โˆ’ {0}. (8) Use Theorem 7.2.5 to prove log-concavity of the following sequences. ๐‘› ๐‘› ๐‘› (a) ( ), ( ), . . . , ( ). 0 1 ๐‘› (b) ๐‘(๐‘›, 0), ๐‘(๐‘›, 1), . . . , ๐‘(๐‘›, ๐‘›) (signless first kind Stirling numbers). (9) Prove equation (7.9). (10) Prove equation (7.13). (11) A plane partition of shape ๐œ† is a filling ๐‘ƒ of the cells of ๐œ† with positive integers so that rows and columns weakly decrease. Let ๐‘๐‘๐‘› be the number of plane partitions ๐‘ƒ (of any shape) such that |๐‘ƒ| = ๐‘›. Show that โˆ‘ ๐‘๐‘๐‘› ๐‘ฅ๐‘› = โˆ ๐‘›โ‰ฅ0

๐‘–โ‰ฅ1

1 (1 โˆ’ ๐‘ฅ๐‘– )๐‘–

in two ways: by taking a limit in (7.14) and by providing a proof in the spirit of the Hillmanโ€“Grassl algorithm. (12) (a) Prove that the output array of HG1โ€“HG3 is a reverse plane partition. (b) Prove that (7.16) is a total order. (c) Show that in GH3 of the Hillmanโ€“Grassl construction, the reverse path ๐‘ must reach the rightmost cell in row ๐‘Ž. (d) Prove that after adding ones along the reverse path, the result is still a reverse plane partition. (13) (a) Construct the inverse of the map used in the proof of (7.20). (b) Prove (7.21). (c) Prove (7.24). (14) Complete the proof of Lemma 7.4.3. (15) Derive (7.26) from (7.25) using a bijection. (16) (a) Show that the ๐‘ƒ๐œ†โˆ— -partitions are exactly the reverse plane partitions of shape ๐œ†. (b) Show that for any finite poset ๐‘ƒ we have #โ„’(๐‘ƒ) = #โ„’(๐‘ƒ โˆ— ). (17) (a) Let ๐œ be a poset. Call ๐œ a rooted tree if it has a 0ฬ‚ and its Hasse diagram is a tree in the graph-theoretic sense of the term. If #๐œ = ๐‘›, then a natural labeling of ๐œ is an order-preserving bijection ๐œ โ†’ [๐‘›]. See Figure 7.19 for an example of a rooted tree (on the left) and a natural labeling (in the middle). Let ๐‘“๐œ be the number of natural labelings of ๐œ. Define the hooklength of ๐‘ฃ โˆˆ ๐œ to be โ„Ž๐‘ฃ = #๐‘ˆ(๐‘ฃ) Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Exercises

261

5

3

6

2

4

1

1

1 3

2 6

1

Figure 7.19. A rooted tree poset, a natural labeling, and its hooklengths

where ๐‘ˆ(๐‘ฃ) is the upper-order ideal generated by ๐‘ฃ. The right-hand tree in Figure 7.19 lists its hooklengths. Prove that if #๐œ = ๐‘›, then ๐‘“๐œ =

๐‘›! โˆ๐‘ฃโˆˆ๐œ โ„Ž๐‘ฃ

in two ways: probabilistically and using induction on ๐‘›. (b) The comb is the infinite poset on the left in Figure 7.20. Let ๐ฟ๐‘› be the set of lower-order ideals of the comb which have ๐‘› elements. The three elements of ๐ฟ3 are displayed on the right in Figure 7.20. Note that the last two order ideals are considered distinct even though they are isomorphic as posets. Show that #๐ฟ๐‘› = ๐‘“๐‘› , the Fibonacci numbers defined in (1.2). (c) Using the notation of part (a), show that โˆ‘ (๐‘“๐œ )2 = ๐‘›! . ๐œโˆˆ๐ฟ๐‘›

(18) (a) Show that if ๐‘ƒ is a PYT and ๐‘ฅ โˆ‰ ๐‘ƒ, then ๐‘ƒ โ€ฒ = ๐‘Ÿ๐‘ฅ (๐‘ƒ) is still a PYT; that is, the rows and columns of ๐‘ƒ โ€ฒ still increase. (b) Show that the RSK map is well-defined in that ๐‘‡ and ๐‘ˆ are both semistandard. (19) Consider three sets of variables ๐ฑ = {๐‘ฅ๐‘– }๐‘–โ‰ฅ1 , ๐ฒ = {๐‘ฆ๐‘— }๐‘—โ‰ฅ1 , and ๐ฑ๐ฒ = {๐‘ฅ๐‘– ๐‘ฆ๐‘— }๐‘–,๐‘—โ‰ฅ1 . Prove that โˆ‘ ๐‘ ๐œ† (๐ฑ)๐‘ ๐œ† (๐ฒ) = ๐‘ ๐‘› (๐ฑ๐ฒ) ๐œ†โŠข๐‘›

where the variables in ๐ฑ๐ฒ can be substituted into the Schur function in any order since ๐‘ ๐‘› is symmetric. Hint: Use the Cauchy identity, equation (7.30).

โ‹… โ‹…โ‹…

Figure 7.20. The comb and its three lower-order ideals with 3 elements

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

262

7. Counting with Symmetric Functions

(20) Prove Theorem 7.5.4. Hint: First prove that if ๐‘‡ is an SSYT into which one inserts ๐‘ฅ and ๐‘ฆ in that order with ๐‘ฅ โ‰ค ๐‘ฆ, then the box added to the shape by the insertion of ๐‘ฆ is strictly to the right of the box for the insertion of ๐‘ฅ. (21) Show that (7.27) can be derived from (7.30) by taking the coefficient of ๐‘ฅ1 โ‹ฏ ๐‘ฅ๐‘› ๐‘ฆ1 โ‹ฏ ๐‘ฆ๐‘› on both sides. (22) Fill in the details of the proof of Lemma 7.6.2. (23) If ๐œ‹ is a two-line array, then let ๐œ‹ฬ‚ be the upper line and let ๐œ‹ฬŒ be the lower line. (a) Show that any two-line array ๐œ‹ with entries in โ„™ and columns which are lexicographically weakly increasing corresponds to a matrix ๐‘€ โˆˆ Mat. (b) Show that if RSK(๐œ‹) = (๐‘‡, ๐‘ˆ) with sh ๐‘‡ = sh ๐‘ˆ = ๐œ†, then ๐œ†1 = the length of a longest weakly increasing subsequence of ๐œ‹ฬŒ by mimicking the proof of Theorem 7.6.1. (c) Let ๐‘‡ be a semistandard Young tableau and suppose that the content of ๐‘‡ is co ๐‘‡ = [๐›ผ1 , ๐›ผ2 , . . . , ๐›ผ๐‘˜ ]. The standardization of ๐‘‡ is the tableau std ๐‘‡ obtained by replacing the ones in ๐‘‡ by the numbers 1, 2, . . . , ๐›ผ1 from left to right, replacing the twos in ๐‘‡ by the numbers ๐›ผ1 + 1, ๐›ผ1 + 2, . . . , ๐›ผ1 + ๐›ผ2 from left to right, and so on. Show that std ๐‘‡ is a standard Young tableau. (d) Standardize a two-line array ๐œ‹ by using the same left-to-right replacement as in the previous part on ๐œ‹ฬ‚ and then doing so again on ๐œ‹.ฬŒ Clearly std ๐œ‹ is a permutation in two-line form if the columns of ๐œ‹ are lexicographically ordered. Show that in this case RS(std(๐œ‹)) = std(RSK(๐œ‹)). (e) Use part (d) and the result (rather than the proof) of Theorem 7.6.1 to give a second proof of part (b). (24) (a) Extend column insertion to semistandard tableaux ๐‘‡ by having an element ๐‘ฅ bump the uppermost element in a column greater than or equal to ๐‘ฅ. Show that with this definition ๐‘๐‘ฅ (๐‘‡) is semistandard. (b) Give two proofs that for any semistandard Young tableau ๐‘‡ and positive integers ๐‘ฅ, ๐‘ฆ we have ๐‘๐‘ฆ (๐‘Ÿ๐‘ฅ (๐‘‡)) = ๐‘Ÿ๐‘ฅ (๐‘๐‘ฆ (๐‘‡)), one by mimicking the proof of Lemma 7.6.2 and one by using the standardization operator std. (c) Give two proofs that if RSK(๐œ‹) = (๐‘‡, ๐‘ˆ) with sh ๐‘‡ = sh ๐‘ˆ = ๐œ†, then ๐œ†๐‘ก1 = the length of a longest decreasing subsequence of ๐œ‹, one by using part (b) and one using the standardization operator std from the previous exercise. (d) Prove the identity โˆ‘ ๐‘ ๐œ† (๐ฑ)๐‘ ๐œ†๐‘ก (๐ฒ) = โˆ (1 + ๐‘ฅ๐‘– ๐‘ฆ๐‘— ). ๐œ†

๐‘–,๐‘—โ‰ฅ1

Hint: Use column insertion to define a weight-preserving bijection ๐‘€ โ†’ (๐‘‡, ๐‘ˆ) where ๐‘€ โˆˆ Mat is a matrix with all entries zero or one, sh ๐‘‡ = sh ๐‘ˆ, and ๐‘‡, ๐‘ˆ ๐‘ก semistandard. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Exercises

263

โ‹ฎ 1111

โ‹ฎ 211

111

121

โ‹ฎ 112

21 11

22 12

2 1 โˆ…

Figure 7.21. Part of the poset โ„ฑ

(25) Finish the proof of Proposition 7.7.1. (26) Show that the base case holds in Corollary 7.7.3. (27) Prove Proposition 7.7.6. (28) Prove Theorem 7.7.7. (29) (a) Let โ„ฑ๐‘› be the set of all words ๐‘ค of ones and twos having โˆ‘๐‘– ๐‘ค ๐‘– = ๐‘›. Show that for ๐‘› โ‰ฅ 0 we have #โ„ฑ๐‘› = ๐‘“๐‘› , the Fibonacci numbers defined in (1.2). (b) Put a partial order on โ„ฑ = โจ„๐‘›โ‰ฅ0 โ„ฑ๐‘› with covers ๐‘ฃ โ‹– ๐‘ค whenever F1 ๐‘ฃ can be obtained from ๐‘ค by removing the first one in ๐‘ค or F2 ๐‘ค can be obtained from ๐‘ฃ by changing the first one in ๐‘ฃ to a two. The lower ranks of โ„ฑ are shown in Figure 7.21. Show that โ„ฑ is differential. (c) Show that โ„ฑ is a lattice. Hint: Prove that ๐‘ฃโˆง๐‘ค exists by induction on rk ๐‘ฃ+rk ๐‘ค and then use Exercise 12(c) in Chapter 5. (d) Give a second proof that โ„ฑ is a lattice using Exercise 12(d) in Chapter 5. (30) Let ๐‘Ÿ โˆˆ โ„™. Say poset ๐‘ƒ is ๐‘Ÿ-differential if it satisfies DP1, DP3, and the following axiom for any ๐‘ฅ โˆˆ ๐‘ƒ: DP2r If ๐‘ฅ covers ๐‘˜ elements, then it is covered by ๐‘˜ + ๐‘Ÿ elements. Prove the following statements. (a) If ๐‘ƒ is ๐‘Ÿ-differential, then # Rk๐‘› ๐‘ƒ is finite for all ๐‘›. So there are well-defined ๐ท and ๐‘ˆ operators. (b) Let ๐‘ƒ be ranked with # Rk๐‘› ๐‘ƒ finite for all ๐‘›. We have ๐‘ƒ is ๐‘Ÿ-differential โŸบ ๐ท๐‘ˆ โˆ’ ๐‘ˆ๐ท = ๐‘Ÿ๐ผ. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

264

7. Counting with Symmetric Functions

(c) In any ๐‘Ÿ-differential poset we have โˆ‘ (๐‘“๐‘ฅ )2 = ๐‘Ÿ๐‘› ๐‘›! ๐‘ฅโˆˆRk๐‘› ๐‘ƒ

ฬ‚ chains. where ๐‘“ is the number of saturated 0โ€“๐‘ฅ (d) If ๐‘ƒ is ๐‘Ÿ-differential and ๐‘„ is ๐‘ -differential, then ๐‘ƒ ร— ๐‘„ is (๐‘Ÿ + ๐‘ )-differential. In particular, if ๐‘ƒ is differential, then ๐‘ƒ๐‘Ÿ is ๐‘Ÿ-differential. ๐‘ฅ

(31) Show that the backwards implication in Proposition 7.8.4 holds. Hint: Use Theorem 3.8.4. (32) (a) The star ๐‘†๐‘› is the tree with ๐‘‰ = {๐‘ฃ 1 , . . . , ๐‘ฃ ๐‘› } and with edges ๐ธ = {๐‘ฃ 1 ๐‘ฃ 2 , ๐‘ฃ 1 ๐‘ฃ 3 , . . . , ๐‘ฃ 1 ๐‘ฃ ๐‘› }. Find expressions for ๐‘‹(๐‘†๐‘› ) in the monomial basis and in the power sum basis with coefficients which are, up to sign, products of binomial coefficients. (Of course, every integer is a binomial coefficient since (๐‘›1) = ๐‘›. But any such factor should come choosing one thing from ๐‘› things combinatorially.) (b) Prove that for any graph ๐บ ๐‘ƒ(๐บ; ๐‘ก) = โˆ‘ (โˆ’1)#๐น ๐‘กโ„“(๐œ†(๐น)) ๐นโŠ†๐ธ

in three ways: by using deletion-contraction, by using a sign-reversing involution, and by using ๐‘‹(๐บ). (33) (a) Show that (7.42) defines an action of GL๐‘› on ๐‘ƒ๐‘˜ (๐‘›). (b) Show that if ๐ตฬƒ = {๐›ฬƒ 1 , ๐›ฬƒ 2 , . . . , ๐›ฬƒ ๐‘› } is any basis for โ„‚[๐‘›], then {๐›ฬƒ ๐‘– ๐›ฬƒ ๐‘– โ‹ฏ ๐›ฬƒ ๐‘– โˆฃ ๐‘–1 โ‰ค ๐‘–2 โ‰ค โ‹ฏ โ‰ค ๐‘–๐‘˜ } 1

2

๐‘˜

๐‘˜

is a basis for ๐‘ƒ (๐‘›). (34) (a) Let ๐‘’ ๐‘˜ (๐‘›) = ๐‘’ ๐‘˜ (๐‘ฅ1 , ๐‘ฅ2 , . . . , ๐‘ฅ๐‘› ) and similarly for โ„Ž๐‘˜ (๐‘›). Prove that ๐‘’ ๐‘˜ (0) = โ„Ž๐‘˜ (0) = ๐›ฟ ๐‘˜,0 and for ๐‘› โ‰ฅ 1 ๐‘’ ๐‘˜ (๐‘›) = ๐‘’ ๐‘˜ (๐‘› โˆ’ 1) + ๐‘ฅ๐‘› ๐‘’ ๐‘˜โˆ’1 (๐‘› โˆ’ 1), โ„Ž๐‘˜ (๐‘›) = โ„Ž๐‘˜ (๐‘› โˆ’ 1) + ๐‘ฅ๐‘› โ„Ž๐‘˜โˆ’1 (๐‘›). (b) Give three proofs of the first identity in Proposition 7.9.2: one by induction, one using the ๐‘ž-Binomial Theorem (Theorem 3.2.4), and one using Exercise 7(c) in Chapter 3. (c) Give two more proofs of the second identity in Proposition 7.9.2: one by induction and one which uses the ๐‘ž-Binomial Theorem (Theorem 3.2.4). (35) (a) Prove that ๐‘†[๐‘›, ๐‘˜] = โ„Ž๐‘›โˆ’๐‘˜ ([1]๐‘ž , [2]๐‘ž , . . . , [๐‘˜]๐‘ž ), where ๐‘†[๐‘›, ๐‘˜] is the ๐‘ž-Stirling number of the second kind introduced in Chapter 3, Exercise 21. Hint: Use part (a) of Exercise 34. (b) Prove that ๐‘[๐‘›, ๐‘˜] = ๐‘’ ๐‘›โˆ’๐‘˜ ([1]๐‘ž , [2]๐‘ž , . . . , [๐‘› โˆ’ 1]๐‘ž ), where ๐‘[๐‘›, ๐‘˜] is the signless ๐‘ž-Stirling number of the first kind introduced in Chapter 3, Exercise 22. Hint: Use part (a) of Exercise 34. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Exercises

265

(36) Define a relation on the polynomial ring โ„[๐‘ž] by letting ๐‘“(๐‘ž) โ‰ค ๐‘”(๐‘ž) if, for all ๐‘– โˆˆ โ„•, the coefficient of ๐‘ž๐‘– in ๐‘“(๐‘ž) is less than or equal to the coefficient of ๐‘ž๐‘– in ๐‘”(๐‘ž). (a) Prove that this relation is a partial order on โ„[๐‘ž], but not a total order. (b) Define a sequence of polynomials ๐‘“0 (๐‘ž), ๐‘“1 (๐‘ž), . . . to be ๐‘ž-log-concave if ๐‘“๐‘˜ (๐‘ž)2 โ‰ฅ ๐‘“๐‘˜โˆ’1 (๐‘ž)๐‘“๐‘˜+1 (๐‘ž) for all ๐‘˜ โ‰ฅ 1. If the polynomial sequence is finite, we let ๐‘“๐‘˜ (๐‘ž) = 0 for all sufficiently large ๐‘˜. Prove the following. (i) For any ๐‘› โˆˆ โ„• the finite sequence ๐‘ž( 2 ) [ 0

1 ๐‘› ๐‘› ๐‘› ๐‘› ] , ๐‘ž (2) [ ] , . . . , ๐‘ž( 2 ) [ ] 0 1 ๐‘›

is ๐‘ž-log-concave. Hint: Use the first identity in Proposition 7.9.2. (ii) For any ๐‘› โˆˆ โ„• the infinite sequence [

๐‘› ๐‘›+1 ๐‘›+2 ],[ ],[ ] , ... 0 1 2

is ๐‘ž-log-concave. Hint: Use the second identity in Proposition 7.9.2. (iii) For any ๐‘› โˆˆ โ„• the finite sequence ๐‘[๐‘›, 0], ๐‘[๐‘›, 1], . . . , ๐‘[๐‘›, ๐‘›] is ๐‘ž-log-concave. Hint: Use part (b) of the previous exercise.

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

10.1090/gsm/210/08

Chapter 8

Counting with Quasisymmetric Functions

While symmetric functions are invariant under arbitrary permutations of variables, quasisymmetric functions only need to be preserved by order-preserving bijections on the variable subscripts. Quasisymmetric functions are implicit in the work of Stanley on ๐‘ƒ-partitions [84] but were first explicitly defined and studied by Gessel [32]. As we will see in this chapter, these functions also have interesting connections with chain enumeration in posets, pattern avoidance, and graph coloring.

8.1. The algebra of quasisymmetric functions, QSym We start by defining what it means for a power series to be quasisymmetric. We will introduce two important bases for the algebra of quasisymmetric functions and discuss their relationship with symmetric functions. As usual, ๐ฑ = {๐‘ฅ1 , ๐‘ฅ2 , ๐‘ฅ3 , . . . } will be a countably infinite variable set. A power ๐›ผ ๐›ผ ๐›ผ series ๐‘“(๐ฑ) โˆˆ โ„‚[[๐ฑ]] is quasisymmetric if any two monomials of the form ๐‘ฅ๐‘–11 ๐‘ฅ๐‘–22 โ‹ฏ ๐‘ฅ๐‘–๐‘™ ๐‘™ ๐›ผ ๐›ผ ๐›ผ with ๐‘–1 < ๐‘–2 < โ‹ฏ < ๐‘–๐‘™ and ๐‘ฅ๐‘—11 ๐‘ฅ๐‘—22 โ‹ฏ ๐‘ฅ๐‘—๐‘™ ๐‘™ with ๐‘—1 < ๐‘—2 < โ‹ฏ < ๐‘— ๐‘™ have the same coefficient. Note the increasing condition on the subscripts which is not present in the definition of a symmetric function. So this is more a restrictive condition; that is, every symmetric function is quasisymmetric but not conversely. For example (8.1) ๐‘“(๐ฑ) = 5๐‘ฅ14 ๐‘ฅ2 + 5๐‘ฅ14 ๐‘ฅ3 + 5๐‘ฅ24 ๐‘ฅ3 + โ‹ฏ โˆ’ ๐‘ฅ1 ๐‘ฅ22 ๐‘ฅ3 โˆ’ ๐‘ฅ1 ๐‘ฅ22 ๐‘ฅ4 โˆ’ ๐‘ฅ1 ๐‘ฅ32 ๐‘ฅ4 โˆ’ ๐‘ฅ2 ๐‘ฅ32 ๐‘ฅ4 โˆ’ โ‹ฏ is quasisymmetric but not symmetric. An equivalent way to define ๐‘“(๐ฑ) being qua๐›ผ ๐›ผ ๐›ผ sisymmetric is to say that any monomial of the form ๐‘ฅ๐‘–11 ๐‘ฅ๐‘–22 โ‹ฏ ๐‘ฅ๐‘–๐‘™ ๐‘™ with ๐‘–1 < ๐‘–2 < โ‹ฏ < ๐‘–๐‘™ ๐›ผ ๐›ผ ๐›ผ has the same coefficient as ๐‘ฅ1 1 ๐‘ฅ2 2 โ‹ฏ ๐‘ฅ๐‘™ ๐‘™ . 267 Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

268

8. Counting with Quasisymmetric Functions

To set up notation, let QSym๐‘› = QSym๐‘› (๐ฑ) = {๐‘“(๐ฑ) โˆˆ โ„‚[[๐ฑ]] โˆฃ ๐‘“ is quasisymmetric and homogeneous of degree ๐‘›}. Then the algebra of quasisymmetric functions is QSym = QSym(๐ฑ) =

โจ

QSym๐‘› (๐ฑ).

๐‘›โ‰ฅ0

Note that, as with symmetric functions, since QSym is a direct sum the quasisymmetric power series in it are of bounded degree. Note also that, unlike symmetric functions, it is not obvious that this is an algebra, i.e., that QSym is closed under multiplication and not just under taking linear combinations. The proof of this fact will follow from Theorem 8.3.1. As we will see, bases for QSym are indexed by integer compositions. We will be interested in two particular bases. Given a composition ๐›ผ = [๐›ผ1 , ๐›ผ2 , . . . , ๐›ผ๐‘™ ], the associated monomial quasisymmetric function is ๐›ผ ๐›ผ ๐›ผ โˆ‘ ๐‘ฅ๐‘–11 ๐‘ฅ๐‘–22 โ‹ฏ ๐‘ฅ๐‘–๐‘™ ๐‘™ . ๐‘€๐›ผ = ๐‘–1 โ‹ฏ > 0. This term is used since the reversal ๐œ‹๐‘Ÿ is of a form usually called layered. The increasing subsequences ๐‘š, ๐‘š + 1, . . . , ๐‘› and so forth are called its layers and their lengths are the layer lengths. For example ๐œ‹ = 789561234 is reverse layered with layers 789, 56, 1234 and corresponding layer lengths 3, 2, 4. Lemma 8.4.1. We have Av๐‘› (132, 213) = {๐œ‹ โˆˆ ๐”–๐‘› โˆฃ ๐œ‹ is reverse layered}. Proof. Note that the reverse layered permutations ๐œ‹ are exactly the ones such that, for all ๐‘Ž, ๐‘ โˆˆ [๐‘›], if ๐‘Ž < ๐‘ and ๐‘Ž is before ๐‘ in ๐œ‹, then every element of [๐‘Ž, ๐‘] comes between ๐‘Ž, ๐‘ in ๐œ‹. Now ๐œ‹ contains 132 if and only if ๐œ‹ contains a subsequence ๐‘Ž๐‘๐‘ with ๐‘ โˆˆ [๐‘Ž, ๐‘] coming after ๐‘. Similarly ๐œ‹ containing 213 is equivalent to there being a ๐‘ โˆˆ [๐‘Ž, ๐‘] appearing before ๐‘Ž. So ๐œ‹ avoids both patterns precisely when ๐œ‹ is reverse layered. โ–ก To bring in quasisymmetric functions define, for any set ฮ  of permutations, the pattern quasisymmetric function โˆ‘

๐‘„๐‘› (ฮ ) =

๐นDes ๐œ .

๐œโˆˆAv๐‘› (ฮ )

Note that ๐‘„๐‘› (ฮ ) is a sum of fundamental quasisymmetric functions for permutations avoiding ฮ , while ๐นฮ  is a sum over the elements of ฮ  itself. There are times when ๐‘„๐‘› (ฮ ) being symmetric implies that the same is true for ๐‘„๐‘› (ฮ โ€ฒ ) for certain other ฮ โ€ฒ , similar to what happens with Wilf equivalence. Consider the dihedral group ๐ท of the square as defined in (1.11). The following lemma is easy to prove and so its demonstration is left as an exercise. Lemma 8.4.2. For any ๐‘“ โˆˆ ๐ท, any set of permutations ฮ , and any ๐‘› โ‰ฅ 0 we have ๐‘“(Av๐‘› (ฮ )) = Av๐‘› (๐‘“(ฮ )).

โ–ก

Recall that if ๐œ‹ is a permutation, then its complement and reversal (see Exercise 37 of Chapter 1) are denoted ๐œ‹๐‘ and ๐œ‹๐‘Ÿ , respectively. Proposition 8.4.3. Suppose that ๐‘„๐‘› (ฮ ) is symmetric and that it has Schur expansion ๐‘„๐‘› (ฮ ) = โˆ‘ ๐‘ ๐œ† ๐‘ ๐œ† . ๐œ† ๐‘

(a) We have ๐‘„๐‘› (ฮ  ) is symmetric and ๐‘„๐‘› (ฮ ๐‘ ) = โˆ‘ ๐‘ ๐œ† ๐‘ ๐œ†๐‘ก . ๐œ†

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

278

8. Counting with Quasisymmetric Functions

(b) We have ๐‘„๐‘› (ฮ ๐‘Ÿ ) is symmetric and ๐‘„๐‘› (ฮ ๐‘Ÿ ) = โˆ‘ ๐‘ ๐œ† ๐‘ ๐œ†๐‘ก . ๐œ† ๐‘๐‘Ÿ

(c) We have ๐‘„๐‘› (ฮ  ) is symmetric and ๐‘„๐‘› (ฮ ๐‘๐‘Ÿ ) = โˆ‘ ๐‘ ๐œ† ๐‘ ๐œ† . ๐œ†

Proof. (a) Applying the definition of ๐‘„๐‘› (ฮ ) and then Theorem 8.2.5 to the hypothesis we have (8.11)

โˆ‘

๐นDes ๐œ = ๐‘„๐‘› (ฮ ) = โˆ‘ ๐‘ ๐œ† ๐‘ ๐œ† = โˆ‘ ๐‘ ๐œ† ๐œ†

๐œโˆˆ๐”–๐‘› (ฮ )

๐œ†

โˆ‘

๐นDes ๐‘ƒ .

๐‘ƒโˆˆSYT(๐œ†)

In passing from ๐œŽ to ๐œŽ๐‘ , every descent is changed into an ascent and vice versa. So we have Des ๐œŽ๐‘ = [๐‘› โˆ’ 1] โˆ’ Des ๐œŽ.

(8.12)

Also note that in any SYT ๐‘ƒ with ๐‘˜ in cell (๐‘–, ๐‘—) and ๐‘˜ + 1 in cell (๐‘–โ€ฒ , ๐‘—โ€ฒ ), either we have ๐‘– < ๐‘–โ€ฒ and ๐‘— โ‰ฅ ๐‘—โ€ฒ which implies ๐‘˜ โˆˆ Des ๐‘ƒ, or we have ๐‘– โ‰ฅ ๐‘–โ€ฒ and ๐‘— < ๐‘—โ€ฒ which implies ๐‘˜ โˆ‰ Des ๐‘ƒ. It follows that Des ๐‘ƒ ๐‘ก = [๐‘› โˆ’ 1] โˆ’ Des ๐‘ƒ.

(8.13)

Using Lemma 8.4.2, then equations (8.12), (8.11), and (8.13), and then finally Theorem 8.2.5 in that order yields ๐‘„๐‘› (ฮ ๐‘ ) =

โˆ‘

๐น[๐‘›โˆ’1]โˆ’Des ๐œ = โˆ‘ ๐‘ ๐œ†

๐œโˆˆ๐”–๐‘› (ฮ )

๐œ†

โˆ‘

๐น[๐‘›โˆ’1]โˆ’Des ๐‘ƒ = โˆ‘ ๐‘ ๐œ† ๐‘ ๐œ†๐‘ก . ๐œ†

๐‘ƒโˆˆSYT(๐œ†)

(b) The proof is similar to part (a) except that one uses Theorem 7.6.3 in place of equation 8.13. The details are left to the reader. (c) This is implied by the first two parts and Lemma 8.4.2.

โ–ก

We will now use our results to prove that two particular ฮ  โŠ† ๐”–3 have ๐‘„๐‘› (ฮ ) symmetric for all ๐‘› and determine its Schur expansion. For a complete list of all such ฮ , as well as examples which are not subsets of ๐”–3 , see [42]. A partition ๐œ† is a hook if ๐œ† = (๐‘Ž, 1๐‘ ) for ๐‘Ž โ‰ฅ 1 and ๐‘ โ‰ฅ 0. Let โ„‹๐‘› = {๐œ† โŠข ๐‘› โˆฃ ๐œ† is a hook}. Theorem 8.4.4. We have ๐‘„๐‘› (132, 213) = ๐‘„๐‘› (231, 312) = โˆ‘ ๐‘ ๐œ† . ๐œ†โˆˆโ„‹๐‘›

Proof. Since {231, 312} = {132, 213}๐‘ it suffices, by the previous proposition, to prove that ๐‘„๐‘› (132, 213) has the given Schur expansion. First we claim that ๐‘„๐‘› (132, 213) =

โˆ‘

๐น๐‘† .

๐‘†โŠ†[๐‘›โˆ’1]

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

8.5. The chromatic quasisymmetric function

279

To prove the claim, it suffices to show that there is a bijection between subsets ๐‘† โŠ† [๐‘› โˆ’ 1] and permutations ๐œŽ โˆˆ Av๐‘› (132, 213) such that Des ๐œŽ = ๐‘†. Recall from Lemma 8.4.1 that these ๐œŽ are exactly the reverse layered permutations. If ๐›ผ = [๐›ผ1 , . . . , ๐›ผ๐‘˜ ] is the composition of layer lengths of ๐œŽ, then Des ๐œŽ = ๐‘† where ๐‘† = ๐œ™โˆ’1 (๐›ผ) and ๐œ™ is the bijection (1.8). So it suffices to prove that for any ๐›ผ โŠง ๐‘› there is a unique ๐œŽ โˆˆ Av๐‘› (132, 213) having ๐›ผ as its composition of layer lengths. But this is clear since we must put the ๐›ผ1 largest elements of [๐‘›] in the first layer, the next ๐›ผ2 largest elements in the second layer, and so on. From the previous paragraph, we will be done if we can show that โˆ‘ ๐‘ ๐œ† = ๐œ†โˆˆโ„‹๐‘›

โˆ‘

๐น๐‘† .

๐‘†โŠ†[๐‘›โˆ’1]

Because of Theorem 8.2.5, it suffices to show that there is a bijection between subsets ๐‘† โŠ† [๐‘› โˆ’ 1] and hook tableaux ๐ป such that Des ๐ป = ๐‘†. But given any ๐‘† โ€ฒ โŠ† [2, ๐‘›], there is a unique hook tableau ๐ป whose first column is ๐‘† โ€ฒ โˆช {1}. And Des ๐ป = ๐‘† โ€ฒ โˆ’ 1, the set obtained from ๐‘† โ€ฒ by subtracting one from each entry. So the bijection ๐‘† โ€ฒ โ†ฆ ๐‘† = ๐‘† โ€ฒ โˆ’ 1 completes the proof. โ–ก

8.5. The chromatic quasisymmetric function Chromatic quasisymmetric functions were introduced by Shareshian and Wachs [82] in part to study the (๐Ÿ‘ + ๐Ÿ)-Free Conjecture of Stanley and Stembridge [93]. These quasisymmetric functions refine the chromatic symmetric functions from Section 7.8 and have many interesting properties, including a connection with Hessenberg varieties from algebraic geometry. Here, we consider what happens when such a function is symmetric as well as making a connection with reverse ๐‘ƒ-partitions. Throughout this section, ๐บ = (๐‘‰, ๐ธ) will be a graph with ๐‘‰ = [๐‘›]. We let ๐’ซ๐ถ(๐บ) = {๐‘ โˆถ ๐‘‰ โ†’ โ„™ โˆฃ ๐‘ is a proper coloring of ๐บ}. The set of ascents of ๐‘ โˆˆ ๐’ซ๐ถ(๐บ) is Asc ๐‘ = {๐‘–๐‘— โˆˆ ๐ธ โˆฃ ๐‘– < ๐‘— and ๐‘(๐‘–) < ๐‘(๐‘—)} with corresponding ascent number asc ๐‘ = # Asc ๐‘. Figure 8.2 displays three proper colorings of a path with edges 13 and 32. They have Asc ๐‘ 1 = โˆ…, Asc ๐‘ 2 = {13}, and Asc ๐‘ 3 = {13, 23} so that asc ๐‘ 1 = 0, asc ๐‘ 2 = 1, and asc ๐‘ 3 = 2, respectively. Given ๐‘1 =

7

5

7

1

3

2

๐‘2 =

4

5

7

1

3

2

๐‘3 =

Figure 8.2. Three proper colorings of a graph with ๐‘‰ = [3]

Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

4

5

4

1

3

2

280

8. Counting with Quasisymmetric Functions

variable set ๐ฑ = {๐‘ฅ1 , ๐‘ฅ2 , . . . } as well as another parameter ๐‘ž, define the chromatic quasisymmetric function of ๐บ to be ๐‘‹(๐บ; ๐ฑ, ๐‘ž) =

โˆ‘

๐ฑ๐‘ ๐‘žasc ๐‘

๐‘โˆˆ๐’ซ๐ถ(๐บ) ๐‘

where ๐ฑ is defined by (7.36). To illustrate with the graph in Figure 8.2, if we consider the colorings ๐‘ such that ๐‘(1) = ๐‘(2) = ๐‘– and ๐‘(3) = ๐‘— > ๐‘–, then each such map contributes ๐‘ฅ๐‘–2 ๐‘ฅ๐‘— ๐‘ž2 to the sum for a total of ๐‘€21 ๐‘ž2 . If the inequality between ๐‘– and ๐‘— is reversed, then there are no ascents and the contribution is ๐‘€12 . Similar considerations apply to the six ways three distinct colors could be assigned to the vertex set, giving a total of ๐‘‹(๐บ; ๐ฑ, ๐‘ž) = (๐‘€12 + 2๐‘€13 ) + (2๐‘€13 )๐‘ž + (๐‘€21 + 2๐‘€13 )๐‘ž2 . Because of the similarity to our notation ๐‘‹(๐บ; ๐ฑ) for the chromatic symmetric function, we will always include the ๐‘ž when referring to its quasisymmetric cousin. Also, it will be convenient to define ๐‘‹๐‘˜ (๐บ; ๐ฑ) = โˆ‘ ๐ฑ๐‘ ๐‘โˆˆ๐’ซ๐ถ(๐บ) asc ๐‘=๐‘˜

so that ๐‘‹(๐บ; ๐ฑ, ๐‘ž) = โˆ‘ ๐‘‹๐‘˜ (๐บ; ๐ฑ)๐‘ž๐‘˜ . ๐‘˜โ‰ฅ0

We first show that the chromatic quasisymmetric function lives up to its name and is a refinement of the chromatic symmetric function. Proposition 8.5.1. Let ๐บ be a graph with ๐‘‰ = [๐‘›]. We have: (a) ๐‘‹๐‘˜ (๐บ; ๐ฑ) โˆˆ QSym๐‘› for all ๐‘˜ โ‰ฅ 0. (b) ๐‘‹(๐บ; ๐ฑ, ๐‘ž) has degree #๐ธ as a polynomial in ๐‘ž. (c) ๐‘‹(๐บ; ๐ฑ, 1) = ๐‘‹(๐บ; ๐ฑ). Proof. (a) We have ๐‘‹๐‘˜ (๐บ, ๐ฑ) is homogeneous of degree ๐‘› since in ๐ฑ๐‘ there is a factor ๐‘ฅ๐‘– for each vertex ๐‘– โˆˆ ๐‘‰. To show it is quasisymmetric consider a monomial of ๐‘‹๐‘˜ (๐บ, ๐ฑ), ๐›ผ ๐›ผ ๐›ผ say ๐ฑ๐‘ = ๐‘ฅ๐‘–11 ๐‘ฅ๐‘–22 โ‹ฏ ๐‘ฅ๐‘–๐‘™ ๐‘™ with ๐‘–1 < ๐‘–2 < โ‹ฏ < ๐‘–๐‘™ which arose from a proper coloring ๐‘ โˆถ ๐‘‰ โ†’ {๐‘–1 , ๐‘–2 , . . . , ๐‘–๐‘™ }. Take any other set of positive integers ๐‘—1 < ๐‘—2 < โ‹ฏ < ๐‘— ๐‘™ and consider the coloring ๐‘โ€ฒ = ๐‘“ โˆ˜ ๐‘ where ๐‘“(๐‘–๐‘š ) = ๐‘—๐‘š for all ๐‘š. Then ๐‘โ€ฒ is proper since ๐‘“ is a bijection and asc ๐‘โ€ฒ = asc ๐‘ = ๐‘˜ since ๐‘“ is an increasing function. It follows that โ€ฒ ๐›ผ ๐›ผ ๐›ผ ๐‘‹๐‘˜ (๐บ; ๐ฑ) contains a corresponding monomial ๐ฑ๐‘ = ๐‘ฅ๐‘–๐‘—1 ๐‘ฅ๐‘–๐‘—2 โ‹ฏ ๐‘ฅ๐‘–๐‘—๐‘™ with ๐‘—1 < ๐‘—2 < โ‹ฏ < ๐‘— ๐‘™ . Thus ๐‘‹๐‘˜ (๐บ; ๐ฑ) is quasisymmetric. (b) Since asc ๐‘ counts a subset of ๐ธ for any coloring ๐‘, we have that the degree of ๐‘‹(๐บ; ๐ฑ, ๐‘ž) is no greater than #๐ธ. And the coloring ๐‘(๐‘–) = ๐‘– for all ๐‘– โˆˆ ๐‘‰ has asc ๐‘ = #๐ธ, so that is the degree. (c) This follows trivially from the definitions.

โ–ก

To explore what happens if ๐‘‹(๐บ; ๐ฑ, ๐‘ž) is in fact a symmetric function in ๐ฑ, we need to discuss reversal and palindromicity of sequences. Given a sequence ๐‘Ž โˆถ ๐‘Ž0 , ๐‘Ž1 , . . . , ๐‘Ž๐‘› , its reversal is then ๐‘Ž๐‘Ÿ โˆถ ๐‘Ž๐‘› , ๐‘Ž๐‘›โˆ’1 , . . . , ๐‘Ž0 . We say that ๐‘Ž is palindromic with center ๐‘›/2 Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

8.5. The chromatic quasisymmetric function

281

if ๐‘Ž = ๐‘Ž๐‘Ÿ . For example, the sequence 0, 7, 2, 2, 7, 0 is palindromic with center 5/2. We ๐‘› also say that the generating function ๐‘“(๐‘ž) = โˆ‘๐‘–=0 ๐‘Ž๐‘– ๐‘ž๐‘– has one of these properties if the corresponding sequence does. So, in our example, 7๐‘ž + 2๐‘ž2 + 2๐‘ž3 + 7๐‘ž4 is palindromic with center 5/2. Note that if ๐‘“(๐‘ž) is palindromic with center ๐‘›/2, then ๐‘› need not be the degree of ๐‘“(๐‘ž) because of possible initial zeros. Also, since the ๐‘Ž๐‘– may contain other variables, we sometimes say that ๐‘“(๐‘ž) is palindromic in ๐‘ž to be specific. There is a simple algebraic test for being a palindrome. We leave its proof to the reader. Lemma 8.5.2. Let ๐‘Ž โˆถ ๐‘Ž0 , ๐‘Ž1 , . . . , ๐‘Ž๐‘› be a sequence with generating function ๐‘“(๐‘ž). (a) The generating function for ๐‘Ž๐‘Ÿ is ๐‘ž๐‘› ๐‘“(1/๐‘ž). (b) The given sequence is a palindrome with center ๐‘›/2 if and only if โ–ก

๐‘“(๐‘ž) = ๐‘ž๐‘› ๐‘“(1/๐‘ž). Define an involution ๐œŒ on the monomial quasisymmetric functions by letting ๐œŒ(๐‘€๐›ผ ) = ๐‘€๐›ผ๐‘Ÿ .

Extend ๐œŒ by linearity to QSym[๐‘ž], the polynomials in ๐‘ž whose coefficients are quasisymmetric functions, where powers of ๐‘ž are treated as scalars. We will see that this involution reverses ๐‘‹(๐บ; ๐ฑ, ๐‘ž) as a polynomial in ๐‘ž. To express the resulting power series, define the descent set and descent number of a coloring ๐‘ to be Des ๐‘ = {๐‘–๐‘— โˆˆ ๐ธ โˆฃ ๐‘– < ๐‘— and ๐‘(๐‘–) > ๐‘(๐‘—)} and des ๐‘ = # Des ๐‘, respectively. Theorem 8.5.3. Let ๐บ be a graph with ๐‘‰ = [๐‘›] and #๐ธ = ๐‘š. (a) ๐œŒ(๐‘‹(๐บ; ๐ฑ, ๐‘ž)) = ๐‘ž๐‘š ๐‘‹(๐บ; ๐ฑ, ๐‘žโˆ’1 ). (b) ๐œŒ(๐‘‹(๐บ; ๐ฑ, ๐‘ž)) =

โˆ‘

๐ฑ๐‘ ๐‘ždes ๐‘ .

๐‘โˆˆ๐’ซ๐ถ(๐บ) ๐›ผ

๐›ผ

Proof. (a) For a composition ๐›ผ = [๐›ผ1 , . . . , ๐›ผ๐‘™ ] and ๐‘˜ โˆˆ โ„•, the coefficient of ๐‘ฅ1 1 โ‹ฏ ๐‘ฅ๐‘™ ๐‘™ ๐‘ž๐‘˜ on the left side of (a) is ๐›ผ

๐›ผ

๐›ผ

๐›ผ

[๐‘ฅ1 1 โ‹ฏ ๐‘ฅ๐‘™ ๐‘™ ๐‘ž๐‘˜ ]๐œŒ(๐‘‹(๐บ; ๐ฑ, ๐‘ž)) = [๐‘ฅ1 ๐‘™ โ‹ฏ ๐‘ฅ๐‘™ 1 ๐‘ž๐‘˜ ]๐‘‹(๐บ; ๐ฑ, ๐‘ž) ๐›ผ

๐›ผ

= #{proper ๐‘ โˆฃ ๐ฑ๐‘ = ๐‘ฅ1 ๐‘™ โ‹ฏ ๐‘ฅ๐‘™ 1 and asc ๐‘ = ๐‘˜}.

(8.14) Similarly ๐›ผ

๐›ผ

๐›ผ

๐›ผ

[๐‘ฅ1 1 โ‹ฏ ๐‘ฅ๐‘™ ๐‘™ ๐‘ž๐‘˜ ](๐‘ž๐‘š ๐‘‹(๐บ; ๐ฑ, ๐‘žโˆ’1 )) = [๐‘ฅ1 1 โ‹ฏ ๐‘ฅ๐‘™ ๐‘™ ๐‘ž๐‘šโˆ’๐‘˜ ]๐‘‹(๐บ; ๐ฑ, ๐‘ž) (8.15)

โ€ฒ

๐›ผ

๐›ผ

= #{proper ๐‘โ€ฒ โˆฃ ๐ฑ๐‘ = ๐‘ฅ1 1 โ‹ฏ ๐‘ฅ๐‘™ ๐‘™ and asc ๐‘โ€ฒ = ๐‘š โˆ’ ๐‘˜}.

So it suffices to find a bijection between the ๐‘ counted by (8.14) and the ๐‘โ€ฒ counted by (8.15). Define ๐‘“ โˆถ [๐‘™] โ†’ [๐‘™] by ๐‘“(๐‘–) = ๐‘™ โˆ’ ๐‘– + 1 for all ๐‘–. Now given ๐‘ as in (8.14), we let ๐‘โ€ฒ = ๐‘“ โˆ˜ ๐‘. We have that ๐‘โ€ฒ is still proper since ๐‘“ is a bijection. If ๐‘ sends ๐›ผ๐‘– vertices to color ๐‘–, then ๐‘โ€ฒ sends that many vertices to color ๐‘™ โˆ’ ๐‘– + 1. So their monomials are related as desired. And since ๐‘“ is order reversing, asc ๐‘โ€ฒ = ๐‘š โˆ’ asc ๐‘. Finally, ๐‘“ induces a bijection on colorings since ๐‘ = ๐‘“ โˆ˜ ๐‘โ€ฒ . Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

282

8. Counting with Quasisymmetric Functions

1 2 2

4 3

4 ๐‘ƒ(๐บ) =

๐บ= 1

5

3

5

Figure 8.3. A graph and its poset

(b) For any proper coloring we have Des ๐‘ = ๐ธ โˆ’ Asc ๐‘ and so des ๐‘ = ๐‘š โˆ’ asc ๐‘. It follows that ๐‘ž๐‘š ๐‘‹(๐บ; ๐ฑ, ๐‘žโˆ’1 ) =

โˆ‘

๐ฑ๐‘ ๐‘ždes ๐‘ .

๐‘โˆˆ๐’ซ๐ถ(๐บ)

The result now follows from part (a).

โ–ก

Similarly to QSym[๐‘ž], define Sym[๐‘ž] to be the set of polynomials in ๐‘ž whose coefficients are symmetric functions. The reader should find it easy to supply the details of the demonstration of the next result. Corollary 8.5.4. Let ๐บ be a graph with ๐‘‰ = [๐‘›] and #๐ธ = ๐‘š such that ๐‘‹(๐บ; ๐ฑ, ๐‘ž) โˆˆ Sym[๐‘ž]. (a) ๐‘‹(๐บ; ๐ฑ, ๐‘ž) =

โˆ‘

๐ฑ๐‘ ๐‘ždes ๐‘ .

๐‘โˆˆ๐’ซ๐ถ(๐บ)

(b) ๐‘‹(๐บ; ๐ฑ, ๐‘ž) is palindromic in ๐‘ž with center ๐‘š/2.

โ–ก

We end by making a connection with reverse ๐‘ƒ-partitions. Given any graph ๐บ with ๐‘‰ = [๐‘›], there is an associated poset ๐‘ƒ(๐บ) on [๐‘›] defined as follows. Call a path ๐‘–1 , ๐‘–2 , . . . , ๐‘–๐‘™ in ๐บ decreasing if ๐‘–1 > ๐‘–2 > โ‹ฏ > ๐‘–๐‘™ . Now define ๐‘– โŠด ๐‘— in ๐‘ƒ(๐บ) if there is a decreasing path from ๐‘– to ๐‘— in ๐บ. See Figure 8.3 for an example. We must make sure that ๐‘ƒ(๐บ) satisfies the poset axioms. Lemma 8.5.5. If ๐บ is a graph with ๐‘‰ = [๐‘›], then ๐‘ƒ(๐บ) is a poset on [๐‘›]. Proof. We have ๐‘– โŠด ๐‘– for ๐‘– โˆˆ [๐‘›] because of the path of length zero from ๐‘– to itself which is decreasing. If ๐‘– โŠด ๐‘— and ๐‘— โŠด ๐‘–, then the decreasing path from ๐‘– to ๐‘— forces ๐‘– โ‰ฅ ๐‘—. Similarly ๐‘— โ‰ฅ ๐‘– so that ๐‘– = ๐‘—. Finally, if ๐‘– โŠด ๐‘— and ๐‘— โŠด ๐‘˜, then consider the concatenation ๐‘ƒ of the decreasing paths from ๐‘– to ๐‘— and from ๐‘— to ๐‘˜. Now the vertices on ๐‘ƒ form a decreasing sequence since the two individual paths are decreasing and the terminal vertex of the first equals the initial vertex of the second. This also shows that ๐‘ƒ is a path since the decreasing condition makes it impossible to repeat a vertex. So ๐‘– โŠด ๐‘˜ which finishes transitivity and the proof. โ–ก Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Exercises

283

We can now show that the coefficient of ๐‘ž0 in ๐‘‹(๐บ; ๐ฑ, ๐‘ž) is the quasisymmetric generating function for reverse ๐‘ƒ(๐บ)-partitions. Theorem 8.5.6. If ๐บ is a graph with ๐‘‰ = [๐‘›], then ๐‘‹0 (๐บ; ๐ฑ) = rpar(๐‘ƒ(๐บ); ๐ฑ). Proof. It suffices to show that any proper coloring of ๐บ with no ascents is a reverse ๐‘ƒ(๐บ)-partition and conversely. So let ๐‘ โˆถ ๐‘‰ โ†’ โ„™ be a proper coloring with asc ๐‘ = 0. Recall that it suffices to prove that RPP1 and RPP2 hold for covers ๐‘– โŠฒ ๐‘—. But then ๐‘– > ๐‘— must be a decreasing path consisting of a single edge. And since this is a proper coloring with no ascents, ๐‘(๐‘–) < ๐‘(๐‘—). So both of the desired conditions hold. For the other direction, let ๐‘“ โˆถ ๐‘ƒ(๐บ) โ†’ โ„™ be a reverse partition. By definition of ๐‘ƒ(๐บ), any cover ๐‘– โŠฒ ๐‘— comes from an edge ๐‘–๐‘— with ๐‘– > ๐‘—. By RPP2, we have ๐‘“(๐‘–) < ๐‘“(๐‘—). So ๐‘“ is a proper coloring since the second inequality is strict and, by comparing the last two inequalities, has no ascents. This completes the proof, as well as the book. โ–ก

Exercises (1) Prove that QSym is an algebra; that is, show it is closed under linear combinations and products. (2) Prove Proposition 8.1.2. (3) (a) Show that ๐‘€๐›ผ is symmetric if and only if ๐›ผ = [๐‘–๐‘š ] for some ๐‘–, ๐‘š. In addition, show that ๐‘€[๐‘–๐‘š ] = ๐‘š(๐‘–๐‘š ) . (b) Show that ๐น๐›ผ is symmetric if and only if ๐›ผ = [๐‘›] or ๐›ผ = [1๐‘› ] for some ๐‘›. In addition, show that ๐น๐‘› = โ„Ž๐‘› and ๐น1๐‘› = ๐‘’ ๐‘› . (c) Suppose that ๐›ผ, ๐›ฝ โŠง ๐‘› are distinct compositions where ๐‘› > 3. Show that ๐น๐›ผ + ๐น ๐›ฝ is symmetric if and only if {๐›ผ, ๐›ฝ} = {[๐‘›], [1๐‘› ]}. (4) (a) Show that the ๐‘€๐›ผ can be expressed in terms of the ๐น๐›ผ as ๐‘€๐›ผ = (โˆ’1)โ„“(๐›ผ) โˆ‘ (โˆ’1)โ„“(๐›ฝ) ๐น ๐›ฝ ๐›ฝโ‰ค๐›ผ

where โ„“(โ‹…) is the length function in two ways: using Theorem 1.3.3(d) and using Mรถbius inversion. (b) Show that part (a) implies Theorem 8.1.4. (5) Prove Lemma 8.2.1. (6) Complete the proof of equation (8.4). (7) (a) Prove that the map between reverse compatible and compatible functions given in the text is a well-defined bijection and that if ๐‘“ maps to ๐‘”, then |๐‘“| = |๐‘”| + ๐‘› where the permutations come from ๐”–๐‘› . (b) Given a poset ๐‘ƒ on [๐‘›], find a poset ๐‘„ on [๐‘›] such that there is a bijection ๐œ“ โˆถ RPar ๐‘ƒ โ†’ Par ๐‘„ satisfying |๐‘“| = |๐œ“(๐‘“)| + ๐‘› for all ๐‘“ โˆˆ RPar ๐‘ƒ. Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

284

8. Counting with Quasisymmetric Functions

(8) (a) Show that if ๐‘„ is the recording tableau for ๐œ‹ under the Robinsonโ€“Schensted map, then Des ๐‘„ = Des ๐œ‹. (b) Show in the proof of Theorem 8.2.5 that Des ๐œ‹ โŠ† Des ๐‘„. (9) Prove Lemma 8.2.3. (10) Prove Theorem 8.2.4. (11) Prove Proposition 8.2.6. (12) (a) Show that if ๐œ‹ โˆˆ ๐”–๐‘› , then we have the principal specialization โ€ฒ

๐‘žmaj ๐œ‹ (1 โˆ’ ๐‘ž)(1 โˆ’ ๐‘ž2 ) โ‹ฏ (1 โˆ’ ๐‘ž๐‘› ) where ๐œ‹โ€ฒ is the reverse complement of ๐œ‹ as given by (8.3) (b) Use part (a) and Theorem 8.2.4 to rederive equation (7.23). ๐นDes ๐œ‹ (1, ๐‘ž, ๐‘ž2 , . . . ) =

(13) Prove that if ๐œŽ โˆˆ ๐‘ƒ(๐‘ˆ) and ๐œ โˆˆ ๐‘ƒ(๐‘‰) where ๐‘ˆ, ๐‘‰ are disjoint and #๐‘ˆ = ๐‘š, #๐‘‰ = ๐‘›, then ๐‘š+๐‘› โˆ‘ ๐‘žmaj ๐œ‹ = ๐‘žmaj ๐œ+maj ๐œ [ ] . ๐‘š ๐œ‹โˆˆ๐œโงข๐œ ๐‘ž Conclude that maj is a shuffle compatible function on permutations. Hint: Use equation (7.23). (14) Show that the two maps in the proof of Theorem 8.3.1 are inverses. (15) (a) Let ๐‘ƒ be a finite, ranked poset with a 1ฬ‚ having rk 1ฬ‚ = ๐‘›. Associate with any chain ๐ถ โˆถ 0ฬ‚ = ๐‘ฅ0 < ๐‘ฅ1 < โ‹ฏ < ๐‘ฅ๐‘˜ = 1ฬ‚ the set ๐‘†(๐ถ) = {rk ๐‘ฅ1 , rk ๐‘ฅ2 , . . . , rk ๐‘ฅ๐‘˜โˆ’1 } โŠ† [๐‘› โˆ’ 1]. Show that ๐œ™(๐‘†(๐ถ)) = ๐›ผ(๐ถ) where ๐œ™ is the map in (1.8). (b) Complete the proof of Theorem 8.3.2. (16) (a) Show that for the ๐‘›-chain we have ๐‘€(๐ถ๐‘› ) = โ„Ž๐‘› , the complete homogeneous symmetric function of degree ๐‘›. ๐œ† ๐œ† (b) Show that if the prime factorization of ๐‘› is ๐‘› = ๐‘1 1 โ‹ฏ ๐‘๐‘™ ๐‘™ where ๐œ† = (๐œ†1 , . . . , ๐œ†๐‘™ ) is a partition, then for the divisor lattice ๐‘€(๐ท๐‘› ) = โ„Ž๐œ† . (17) Prove that a permutation ๐œ‹ โˆˆ ๐”–๐‘› is reverse layered if and only if we have ๐œ‹๐‘–+1 โ‰ค ๐œ‹๐‘– + 1 for 1 โ‰ค ๐‘– < ๐‘›. (18) Prove Lemma 8.4.2. (19) Prove part (b) of Proposition 8.4.3. (20) Prove that part (c) of Proposition 8.4.3 is implied by parts (a) and (b) and Lemma 8.4.2. (21) Show that if ๐‘– โˆˆ [๐‘›], then ๐‘ (๐‘–,1๐‘›โˆ’๐‘– ) = โˆ‘ ๐น๐‘† ๐‘†

where the sum is over ๐‘† โˆˆ ([๐‘›โˆ’1] ). ๐‘›โˆ’๐‘– (22) (a) The Knuth class corresponding to an SYT ๐‘ƒ is the set RS

๐พ(๐‘ƒ) = {๐œ‹ โˆฃ ๐œ‹ โ†ฆ (๐‘ƒ, ๐‘„) for some SYT ๐‘„}. Prove that if ๐นฮ  is as defined in (8.9) and sh ๐‘ƒ = ๐œ†, then ๐น๐พ(๐‘ƒ) = ๐‘ ๐œ† . Licensed to AMS. License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms

Exercises

285

(b) The Knuth aggregate corresponding to a partition ๐œ† is the set ๐พ(๐œ†) =

โจ„

๐พ(๐‘ƒ).

๐‘ƒโˆˆSYT(๐œ†)

Prove that ๐น๐พ(๐œ†) = ๐‘“๐œ† ๐‘ ๐œ† . (c) Show that ๐‘„๐‘› (โˆ…) = โˆ‘ ๐‘“๐œ† ๐‘ ๐œ† . ๐œ†โŠข๐‘›

(d) If ๐œ„๐‘˜ = 12 . . . ๐‘˜, then show ๐‘„๐‘› (๐œ„๐‘˜ ) = โˆ‘ ๐‘“๐œ† ๐‘ ๐œ† . ๐œ†1