asymptotic enumeration in pattern avoidance and in...

91
ASYMPTOTIC ENUMERATION IN PATTERN AVOIDANCE AND IN THE THEORY OF SET PARTITIONS AND ASYMPTOTIC UNIFORMITY By MICAH SPENCER COLEMAN A DISSERTATION PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY UNIVERSITY OF FLORIDA 2008 1

Upload: nguyendiep

Post on 24-Aug-2019

213 views

Category:

Documents


0 download

TRANSCRIPT

ASYMPTOTIC ENUMERATION IN PATTERN AVOIDANCE AND IN THE THEORYOF SET PARTITIONS AND ASYMPTOTIC UNIFORMITY

By

MICAH SPENCER COLEMAN

A DISSERTATION PRESENTED TO THE GRADUATE SCHOOLOF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT

OF THE REQUIREMENTS FOR THE DEGREE OFDOCTOR OF PHILOSOPHY

UNIVERSITY OF FLORIDA

2008

1

c© 2008 Micah Spencer Coleman

2

I dedicate this dissertation with love and pride to the memory of Juliana Cole, whose

curiosity and lifelong devotion to learning have defined my family and instilled in me what

was critical to survive my graduate career. Enjoy the bagpipes, Grandmother.

3

ACKNOWLEDGMENTS

My deepest love and admiration to my wife Hiroko. While studying in a foreign

language in a foreign land, she took up two of the hardest imaginable roles, that of

military spouse and that of mathematician caretaker. Daisuki! I thank my parents Bob

and Bobbi Coleman, my brother Matt, and mi vecino Abby for their patience, humor, and

integrity. Great thanks go to Professors Julie Miller, Tina Carter, and Norm Levin, for

first introducing me to “real math”, to our Graduate Coordinator Paul Robinson, and to

the greatest advisory committee ever assembled, Professors David Drake, Meera Sitharam,

Andrew Vince, and Neil White. I am honored and humbled to be associated with each of

them. Finally, my deepest respect and appreciation are held for my advisor, Bona Miklos.

4

TABLE OF CONTENTS

page

ACKNOWLEDGMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

ABSTRACT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

CHAPTER

1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

1.1 Asymptotic Enumeration . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101.2 Notation for Asymptotic Growth Rates . . . . . . . . . . . . . . . . . . . . 101.3 Generating Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

2 PATTERN AVOIDANCE IN PERMUTATIONS AVOIDING A MONOTONEPATTERN . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2.1 Permutations and Permutation Patterns . . . . . . . . . . . . . . . . . . . 122.2 An Open Problem by M. Atkinson . . . . . . . . . . . . . . . . . . . . . . 242.3 Generating Trees . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262.4 “Hat” Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 312.5 Monotone Increasing Patterns q . . . . . . . . . . . . . . . . . . . . . . . . 332.6 The Pattern q = 123 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

3 PATTERN PACKING . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

3.1 General Pattern Packing . . . . . . . . . . . . . . . . . . . . . . . . . . . . 533.2 Pattern Packing in 123-avoiding Permutations . . . . . . . . . . . . . . . . 573.3 Pattern Packing in q-avoiding Permutations . . . . . . . . . . . . . . . . . 593.4 Packing Density and Further Directions . . . . . . . . . . . . . . . . . . . . 61

4 ASYMPTOTIC NORMALITY AND UNIFORMITY . . . . . . . . . . . . . . . 63

4.1 Probability Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 634.2 Triangular Arrays . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 644.3 Asymptotic Normality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 654.4 Asymptotic Uniformity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 664.5 Generating Polynomials with Real, Non-Positive Roots . . . . . . . . . . . 684.6 Asymptotic Normality Implies Asymptotic Uniformity . . . . . . . . . . . 71

5 ON THE ROOTS OF THE BELL POLYNOMIALS . . . . . . . . . . . . . . . . 74

5.1 Stirling Numbers of the Second Kind . . . . . . . . . . . . . . . . . . . . . 745.2 Bell Polynomials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 775.3 Bounds on the Roots of the Bell Polynomials . . . . . . . . . . . . . . . . . 805.4 Asymptotics of the Roots of the Bell Polynomials . . . . . . . . . . . . . . 83

5

REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88

BIOGRAPHICAL SKETCH . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

6

LIST OF FIGURES

Figure page

2-1 The permutation 3142. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

2-2 The permutation 532614. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

2-3 The permutation 865321947. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

2-4 A rooted tree. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

2-5 The complete binary tree. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

2-6 The Fibonacci tree. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

2-7 T(123,132). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

2-8 Tree in W(123,231) rooted at 42153 . . . . . . . . . . . . . . . . . . . . . . . . . 40

2-9 The layered permutation 213654 with layers 21, 3, and 654 . . . . . . . . . . . . 41

3-1 The permutation W (6) = 342516 . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

7

Abstract of Dissertation Presented to the Graduate Schoolof the University of Florida in Partial Fulfillment of theRequirements for the Degree of Doctor of Philosophy

ASYMPTOTIC ENUMERATION IN PATTERN AVOIDANCE AND IN THE THEORYOF SET PARTITIONS AND ASYMPTOTIC UNIFORMITY

By

Micah Spencer Coleman

May 2008

Chair: Miklos BonaMajor: Mathematics

We demonstrate asymptotic properties of some popular combinatorial objects,

including a partial answer to an open problem posed by Michael Atkinson and a general

result on conditions for the coincidence of asymptotic normality and uniformity.

For a permutation π ∈ Sn written in one line notation as π = π1π2 · · · πn, we say

π contains the pattern σ ∈ Sm if there exists a subsequence πi1 · · · πim such that for all

1 ≤ j, k ≤ m it holds that πij < πik if and only if σj < σk. Denote by sn(τ, σ) the number

of n-permutations which do not contain either of the patterns σ and τ . Letting σ′ denote

the pattern (m + 1)σ, we construct classes of patterns for which the limit supremum of

sn(123 · · · r, σ)1/n agrees with the limit supremum of sn(123 · · · r, σ′)1/n for several classes

of patterns σ. We also construct classes of permutations which avoid 123 · · · r and contain

many patterns.

Many combinatorial sequences are of the form (an,k) where n ranges over the

non-negative integers N and, for each n, there exists m = m(n) such that k ranges

from 1 to m. We call such a sequence a combinatorial distribution. Many combinatorial

distributions, upon rescaling, approach in distribution the normal distribution as n grows

to infinity, a phenomenon we call asymptotic normality. A combinatorial distribution is

said to be asymptotically uniform if, for each positive integer q and each residue class

modulo q, the sum of coefficients an,k with k ≡ r (mod q) approaches 1/q as n grows to

infinity. We call this asymptotic uniformity. We prove that if the generating polynomials

8

for a combinatorial distribution have real, nonnegative zeros, asymptotic normality implies

asymptotic uniformity. We apply this result to several sequences from the literature.

Finally, we present original results on the zeros of the Bell polynomials which were

first attained in proving the asymptotic uniformity of the Stirling numbers of the second

kind Sn,k.

9

CHAPTER 1INTRODUCTION

1.1 Asymptotic Enumeration

Enumerative combinatorics involves counting discrete objects, determining the

cardinalities of sets which are indexed by one or more integers. In many cases exact

formulae are known. However, some formulae are so convoluted as to obscure their most

telling information. In either of these cases asymptotics provide powerful methods for

understanding the classes under question.

With that said, we could define asymptotic combinatorics as the study of the growth

of combinatorial sequences indexed by an integer n as n grows without bound, n → ∞.

For a survey of these techniques, the reader is reffered to the chapter by Odlyzko [26] in

the Handbook of Combinatorics.

Definition 1. We introduce some notation to be used throughout. Let [n] denote the set

1, 2, . . . , n, and [a, b] denotes the set a, a + 1, . . . , b. N denotes the natural numbers

n ≥ 0, while P denotes the positive integers n ≥ 1.

1.2 Notation for Asymptotic Growth Rates

Let f : P → P be a function with some sort of predictable behavior. What does it

mean to understand the asymptotic behavior or growth of f? If it is anything like most

functions we encounter, as n grows arbitrarily large, the growth of f probably follows some

pattern, for example a straight line or logarithmic curve. Perhaps f acts erratically on

every small interval but has a smooth overall growth which we can mimic with some other

function g which is easier to understand.

Example 1. Define the function φ(n) = n2 + (−1)n. This function is easy to write but

not so easy to graph. However, as n grows very large we see that φ(n) is very close to the

polynomial n2.

10

Such a situation motivates Big O notation. Given two functions f and g defined on

the positive integers, we write

f(n) = O(g(n))

if there exists a nonzero constant M such that |f(n)| < |Mg(n)| for all n ≥ 1.

Similarly,

f(n) = Ω(g(n))

if there exists a constant M such that |f(n)| > |Mg(n)| for all n ≥ 1.

Finally,

f(n) ∼ g(n)

if

limn→∞

f(n)

g(n)= 1

holds.

1.3 Generating Functions

A powerful area of combinatorics is Generatingfunctionology. For surveys of the field,

the reader is referred to the texts by Wilf [40], Stanley [33], [34], and Bona [8].

Definition 2. Given a sequence (an)n≥0, the associated ordinary generating function is

the power series

f(x) =∑n≥0

anxn.

Similarly, for a sequence (an,k) defined for all n ≥ 0 and 0 ≤ k ≤ m(n) for some function

m(n), the associated generating polynomial for each n is the polynomial

gn(x) =

m(n)∑

k≥0

an,kxk.

The power is in handling these power series and polynomials to reveal information

about the sequence in question. Generating polynomials will play a fundamental role in

the last two chapters.

11

CHAPTER 2PATTERN AVOIDANCE IN PERMUTATIONS AVOIDING A MONOTONE PATTERN

2.1 Permutations and Permutation Patterns

One aspect of enumerative combinatorics, one of the simple beauties that make

it such an attractive discipline, is the ease with which it can be explained to the

non-specialist or even non-mathematician / non-scientist. We seek to count the objects in

some class of size n. For example, we may estimate the number of objects in some class

which are of size n and have statistics l, k as n grows arbitrarily large. In fact, there are

some combinatorial objects which are more easily explained to a non-mathematician than

to some of our mathematical colleagues. The case in mind is the permutation. In the

field of permutation patterns, one views a permutation in one-line notation; i.e., as an

arrangement of the numbers 1 through n for some n. That is it. While we readily admire

and appreciate the wisdom and complexity of our friends the algebraists and topologists,

there is a certain frustration and loss of momentum when trying to describe the simple

properties of permutations that we are dealing with here to such an audience. There is no

consideration of cycle structure, what set is acted upon, etc. From such a rough definition

we can pose many of the fundamental questions dealt with in this field of research.

In the first three chapters of this dissertation we introduce the concept of a

permutation pattern, survey some of the major results and trends in the field of

permutation patterns, develop a foundation for the work contained herein, and pose

further questions to be explored.

Definition 3 (rough). A permutation of length n or a permutation on [n] or an

n-permutation is an arrangement of the numbers 1 through n.

Definition 4 (rigorous). An n-permutation π is an injective mapping from the set

[n] = 1, 2, . . . , n onto itself. We denote π(i) by πi and write π as the concatenation

π1π2 · · · πn, calling the numbers π1, . . . , πn the entries of π. It should be noted that by

12

convention we allow there to be one permutation on the empty set. Finally, for each n ≥ 0

we denote by Sn the set of n-permutations.

We use the convention that an ascent in a permutation π is the index of an entry πi

with πi < πi+1. Likewise a descent in π is the index of an entry πi with πi > πi+1. We

define an ascendee (resp. descendee) to be the index of an entry πi with πi−1 < πi (resp.

πi−1 > πi).

Example 2. For π = 3142, 2 is an ascent; 1 and 3 are descents. 3 is an ascendee; 2 and

4 are descendees. As these are geometric as opposed to algebraic concepts, they may best be

understood visually as in Figure 2-1

x3

x1

x4

x2

AAAAAAAAAAAAAAAAAA£

£££££££££££££££££££££££££A

AAAAAAAAAAAAAAAAA

Figure 2-1. The permutation 3142.

In all contexts considered here, a permutation pattern or simply a pattern is itself a

permutation but there are subtle differences which we shall exploit. Given a permutation

π = π1 · · · πn, a subsequence of π is an ordered subset of the entries of π, (πi1 , . . . , πik) for

some k, which we write in the same order as they appear in π, so i1 < i2 < · · · < ik.

13

Example 3. Consider the 5-permutation π = 13425. If we want to construct a subse-

quence, we have two choices of what to do with each entry πi, whether to include πi in our

subsequence or not. Therefore, we have 25 = 32 subsequences of π. We list them here in an

obvious but meaningless ordering:

Length 0 : ∅

Length 1 : 1 3 4 2 5

Length 2 : 13 14 12 15 34 32 35 42 45 25

Length 3 : 134 132 135 142 145 125 342 345 325 425

Length 4 : 1342 1345 1325 1425 3425

Length 5 : 13425

Roughly speaking, each such subsequence can be associated with some pattern, and

from this we generate vast fields of research. With such tasking, let us formalize the

concept of the permutation pattern.

Definition 5. We call a finite sequence X of distinct positive integers reduced if the

elements of the sequence form the set [n] for some n, i.e. if X is a permutation of [n].

Let X = X1X2 . . . Xn be a finite sequence of n distinct positive integers. Then,

there exists a unique permutation π ∈ Sn such that for all pairs of indices i and j with

1 ≤ i < j ≤ n,

Xi < Xj if and only if πi < πj.

So, X and π are in the same “order”, or order isomorphic. As we assumed π to be a

permutation in Sn, π is reduced, and we call π the reduction of X.

Definition 6. Let γ ∈ Sn. We say that γ contains the pattern σ ∈ Sm (m ≤ n) if there

exists a subsequence γi1γi2 · · · γim (i1 < i2 < · · · < im) whose reduction is σ. If there is no

such subsequence, we say that γ avoids the pattern σ.

14

One more note is in order on the above definition, which implicitly defines patterns.

A pattern is itself simply a permutation, but we use the two words to describe distinct

sets. We ask questions such as does the permutation π contain the pattern σ, etc. These

meanings will be clear in the context. Please also note that we will use the convention

wherever possible of letting m denote the length of a pattern σ and letting n denote the

length of a permutation π for which we want to determine σ-avoidance, etc. Much of

this dissertation will be devoted to questions surrounding the avoidance of a set of two

patterns, one of which will be monotone increasing. We will conventionally let q denote

the monotone increasing pattern in question and let r denote its length.

Example 4. The permutation 34521 contains the pattern 123, as the first three entries of

34521 are 345, which reduces to 123.

Example 5. We exhaust all patterns contained in the permutation 13425 from Example 3

by reducing all subsequences, in the same ordering as above:

Length 0 : ∅

Length 1 : 1 1 1 1 1

Length 2 : 12 12 12 12 12 21 12 21 12 12

Length 3 : 123 132 123 132 123 123 231 123 213 213

Length 4 : 1342 1234 1324 1324 2314

Length 5 : 13425

Of course, we could continue in this fashion. It is a great exercise for the beginner

in this area to exhaust the subsequences of a permutation to determine what patterns

the permutation contains or avoids and pose some conjectures. This is how one learns

anything in combinatorics, by “getting our hands dirty”, doing enough manual labor on

our combinatorial objects to get a feel for their growth and other properties.

15

Definition 7. Let σ ∈ Sm be a pattern. For each n ≥ 0 we denote by Sn(σ) the set

of all n-permutations which avoid σ, and write sn(σ) = |Sn(σ)|, i.e. the number of

n-permutations which avoid σ.

More generally, let Σ be a set of patterns. For each n ≥ 0 we denote by Sn(Σ)

the set of all n-permutations which avoid all patterns in Σ, enumerated by sn(Σ). If

Σ = σ1, σ2, . . . , σk, we write Sn(Σ) as Sn(σ1, σ2, . . . , σk) and likewise write sn(Σ) as

sn(σ1, σ2, . . . , σk), dropping the brackets.

On a historical note, the Sn and sn above may have been coined in honor of Simion

and Schmidt, whose 1985 paper [31] launched pattern avoidance and contained some

results which are still hallmarks of the field.

As the name would imply, enumerative combinatorialists most enjoy enumerating

sets, that is, determining a precise formula for the cardinality of each set which depends

only on the index or indices of that set. Unfortunately, we often find quite interesting

discrete objects whose nature is complex enough to elude precise formulae. Alas, we will

see that for most patterns σ, the sequence sn(σ) falls into the latter category. However,

all is not lost. As was briefly discussed in the introductory chapter, great information

can still be had by the asymptotics of a sequence, and many of the current results in the

field of permutation patterns involve bounds and limits which are not as strong as precise

formulae, but carry power and beauty of their own.

Let us first see some examples of patterns for which we can give an exact formula. We

will treat all patterns in Sm for 0 ≤ m ≤ 3. We will make use of the Kronecker delta δi,j,

defined by

δi,j =

1 if i = j,

0 if i 6= j.

16

Proposition 6. We have the following exact formulae for σ ∈ Sm, m = 0, 1, 2:

sn(∅) = 0,

sn(1) = δn,0,

sn(12) = 1,

sn(21) = 1.

Proof. As we can take an empty subsequence of any permutation, including the empty

permutation itself, no permutation can avoid the pattern ∅, and sn(∅) = 0.

Similarly, every non-empty permutation has a non-empty subsequence of length

1, and sn(1) = 0 unless n = 0, in which case we have s0(1) = 1, counting the sole

permutation in S0.

Note that a permutation π ∈ Sn avoiding the pattern 12 is equivalent to π having

no ascents, implying π is the unique monotone descending sequence of length n, so

sn(12) = 1. The dual statement is that avoiding the pattern 21 is equivalent to avoiding

descents, thus sn(21) = 1.

The reader may (should) have been amazed by the fact that sn(12) = sn(21) and their

dual proofs. In fact, the duality involved was that 12 and 21 are reverses of each other,

and for each n the unique permutation in Sn(12), the monotone decreasing permutation,

is the reverse of the monotone increasing permutation, the unique element of Sn(21). Of

course, one could also prove the equality with the fact that 12 and 21 are complements of

each other. Perhaps these statements also apply to longer, more interesting patterns?

17

Definition 8. Let π = π1π2 · · · πn be a permutation. We define the reverse, complement,

and algebraic inverse, resp., of π as

πR = πnπn−1 · · · π1,

πC = (n− π1 + 1)(n− π2 + 1) · · · (n− πn + 1),

π−1 = φ1φ2 · · ·φn,

where, for all 1 ≤ i ≤ n, πφi= i.

Example 7. Let π = 24315. Then, we have the eight related permutations

π = 24315,

πR = 51342,

πC = 42351,

π−1 = 41325,

πRC(= πCR) = 15324,

(πR)−1 = 25341,

(πC)−1 = 52314,

(πRC)−1 = 14352.

Of course, the algebraic inverse π−1 is just what the name would imply, the inverse

of π considered as an element of the group Sn. Such a definition as the one above may

seem crude, but we had promised to only consider these permutations as linear orders, not

group elements.

In our example of π = 24315, we made easy use of the obvious fact that R and

C commute as operators. We leave it as an exercise that they each commute with the

algebraic inverse as well.

The definition of reverse should be clear. We note that we can write each entry of πR

as πRi = πn−i+1. We can think of complement as “flipping” the permutation in a vertical

18

sense. Now that we have these definitions, we return to the question of equivalences

among patterns.

Suppose a permutation π = π1 · · · πn contains the pattern σ = σ1 · · · σm. Then, we

have a set of indices 1 ≤ p1 ≤ · · · pm ≤ n such that the reduction of the subsequence

πp1 · · · πpm is precisely σ, i.e. for all i < j,

πpi< πpj

if and only if σi < σj.

By the remarks following Definition 8, this is equivalent to the statement that, for each

pair i < j,

πRn−pi+1 > πR

n−pj+1 if and only if σRn−i+1 > σR

n−i+1.

As we run through all pairs i < j, we see that this collection of statements is equivalent to

πRpi

> πRpj

if and only if σRi > σR

j .

Altogether, π containing σ is equivalent to πR containing σR. Likewise, π containing σ

is equivalent to πC containing σC and π−1 containing σ−1. Of course, we can replace the

word “containing” by the word “avoiding” in each statement.

Critically, as the three maps ·R, ·C , and ·−1 are bijections Sn → Sn, it follows that

the number of n-permutations which avoid σ is the same as the number of n-permutations

avoiding σR, σC , σ−1. We have achieved the following result.

Lemma 1. Let σ be a pattern. Then, for all n ≥ 0,

sn(σ) = sn(σR) = sn(σC) = sn(σ−1).

In fact, we can say that each σ ∈ Sm belongs to an equivalence class of m-patterns

whose sequences sn are equal.

Definition 9. We say the patterns σ and τ are Wilf-equivalent if sn(σ) = sn(τ) holds for

all n ≥ 0. This leads naturally to the definition of the Wilf-equivalence class of a pattern σ

as the set of all patterns which are Wilf-equivalent to σ.

19

For our first non-trivial results, we move to m = 3. S3 consists of the six patterns

123, 132, 213, 231, 312, and 321. One quickly recognizes that there are at most two Wilf

equivalence classes in S3 by noting

123R = 321,

132R = 231,

132C = 312,

132RC = 213.

In fact, there is only one Wilf equivalence class in S3 as is seen in the following

lemma.

Lemma 2. (Simion and Schmidt [31]) For all n ≥ 0,

sn(123) = sn(132).

Proof. We call a permutation entry which is less than all entries preceding it a left-to-right

minimum. The remaining entries we call remaining entries.

Let π be a 132-avoiding n-permutation. Form a new n-permutation π′ by fixing the

left-to-right minima and transposing (swapping) pairs of remaining entries until they are

in decreasing order. They will still all be remaining entries, because at each transposition

the left, smaller entry is preceded by some left-to-right minimum which will still precede

the larger remaining entry after the transposition, and the smaller remaining entry moves

to the right, still not a left-to-right minimum. Note that the left-to-right minima are in

decreasing order, as πi < πj with i < j would imply that πj is not a left-to-right minimum.

Therefore, π′ is composed of two decreasing sequences, so by the Pigeon-Hole Principle

there cannot be an increasing subsequence of three entries (three pigeons cannot fit into

two pigeon holes). So, we have mapped our 132-avoiding π to the 123-avoiding π′ and can

apply this process to all π ∈ Sn(132).

20

To see that this process is reversible, we fix the left-to-right minima of π′, leaving

blanks at the other indices and placing the remaining entries in a sack. Moving left to

right, we fill each blank with the least entry still in our sack which is larger than the

rightmost left-to-right minimum to the left of said blank. Now we have a permutation π,

which we claim to be 132-avoiding. Indeed, if there were a copy of 132, then there would

be a copy of 132 beginning with a left-to-right minimum, however this is a contradiction as

the entries following and larger than each left-to-right minimum are increasing. Thus, we

have a bijection.

So, in fact we see that for all σ, τ ∈ S3 and n ≥ 0, sn(σ) = sn(τ). One may be

tempted to suspect such a statement holds for patterns of every length m. However, with

a computer check or a few pages of scribbling, one obtains

s6(1342) = 512 6= s6(1234) = 513.

We now turn to asymptotics. In 1980, Richard Stanley and Herb Wilf independently

conjectured that for each pattern σ there exists a constant cσ such that, for all n ≥ 0, we

have

sn(σ) ≤ cnσ.

In [2], Arratia proved that this was equivalent to the following, long known as the

Stanley-Wilf Conjecture.

Theorem 8 (Marcus-Tardos). Let σ be a pattern. Then, the limit

limn→∞

sn(σ)1/n

exists (and is finite).

The validity of the Stanley-Wilf Conjecture was established by the 2003 proof by

Marcus and Tardos [23] of the Furedi-Hajnal Conjecture on permutation matrices. That

the Furedi-Hajnal Conjecture implies the Stanley-Wilf conjecture was proven by Klazar

21

[21]. For a clear and concise treatment of all, see section 4.5 of [5]. The reader is also

encouraged to see Doron Zeilberger’s alternative rendition [41].

There is still hope for a tighter proof of the Stanley-Wilf Conjecture, as the

Furedi-Hajnal Conjecture only proves there exists a constant, but the constants which

we get from the proof are astronomically larger than the observed constants. It does still

give structure to our work to know that for any pattern σ, sn(σ) is at most exponential,

and additionally the limit

limn→∞

sn(σ)1/n

exists. We denote this limit by L(σ) as it is critical to the sequel.

Now, for a finite set of at least two patterns, we do not have such a strong general

result. For any such set of patterns Σ, it is readily seen that sn(Σ) < sn(σ) for each

pattern σ ∈ Σ, so we have the following corollary to Theorem 8.

Corollary 1. Let Σ be a nonempty, finite set of patterns. Then, there exists a constant cΣ

such that for all n > 0,

sn(Σ) < cnΣ.

So, for any such set of patterns Σ, we see that the sequence sn(Σ) is at most

subexponential, i.e. bounded above by an exponential function, but we did not mention

the existence of a limit of sn(Σ)1/n. For many classes of such sets we cannot prove the

existence of such a limit, an open problem with much interest, discussed in [37].

Our work in this chapter is restricted to sets of the form q, σ, where q is a

monotone increasing pattern, i.e. q = 12 · · · r for some r ≥ 3. We have the following

result on a class of such sets, a generalization of Arratia’s proof [2] that the exponential

bound (i.e. the Marcus-Tardos theorem) implies the existence of the limit for the case of a

single pattern. This generalization was suggested by M. Klazar [22].

Definition 10. We say a permutation π ∈ Sn is decomposable if there exists 1 ≤ k < n

such that for all 1 ≤ i ≤ k < j, πi > πj.

22

Lemma 3. Let q = 12 · · · r and σ ∈ Sm for some r ≥ 3 and m ≥ 3. Then, if σ is not

decomposable, the limit [

L(q, σ) = limn→∞

sn(q, σ)1/n

exists.

Proof. Let m, n ≥ 1. For each pair of permutations α ∈ Sm(q, σ) and β ∈ Sn(q, σ),

construct the (m + n)-permutation

π = (α1 + n) (α2 + n) · · · (αm + n) β1 β2 · · · βn.

By our assumptions on α and β, π avoids q and σ in its first m entries and in its last n

entries. Furthermore, as q and σ are not decomposable, there is no copy of q or σ which

contains entries from these two sets. Thus, π is (q, σ)-avoiding and uniquely determined by

α and β, proving the following inequality.

sm+n(q, σ) ≥ sm(q, σ)sn(q, σ).

We have the following lemma by Fekete [17] for superadditive sequences. The analog of

this lemma for subadditive sequences was used in Arratia’s proof for the case of a single

pattern.

Lemma 4 (Fekete). For every sequence ann∈N satisfying

an+m ≥ anam

for all m,n ≥ 0, the limit

limn→∞

an

n

exists.

Applying Fekete’s lemma to the sequence log sn(q, σ), we thus have that the sequence

sn(q, σ)1/n has a limit. As this sequence is bounded above, this limit is finite.

With these facts in mind, we define L on sets of two patterns as follows.

23

Definition 11. Let σ and τ be patterns. Then, define the following quality.

L(q, σ) = lim supn→∞

sn(σ, τ)1/n.

Of course, for pairs of patterns σ and τ for which the limit exists, such as those in

Lemma 3, the limit and lim sup agree, and L is as we want it to be.

2.2 An Open Problem by M. Atkinson

This work was originated in response to a question posed by Michael Atkinson at

the fifth International Conference on Pattern Avoiding Permutations. We answer the

question in the affirmative for all increasing patterns q on some classes of patterns σ and

specifically for q = 123 for larger classes of patterns σ.

For a pattern σ of length m, define σ′ to be the pattern (m + 1)σ. For example, for

σ = 2431, we have σ′ = 52431. Given a monotone ascending pattern q and a pattern σ,

does it hold that L(q, σ′) = L(q, σ)?

In any permutation the entries preceding the first ascendee form a decreasing

sequence. For a permutation π, if the first ascendee of π is the index i, we call πi the

threshold of π and call the decreasing sequence preceding the threshold the front end. The

set of entries following the threshold we call the back end.

Example 9. For π = 532614, our first ascendee is the index 4, so the threshold of π is the

entry π4 = 6, the front end is 532, and the back end is 14, as shown in Figure 2-2.

Note that in our previous example each entry of the back end is less than the

threshold. Creating our own good luck, we chose 532614 for our example specifically

because it avoids 123. In fact, every 123-avoiding permutation shares this property,

a simple structure of which we shall take great advantage in our handling of these

permutations. To see this property, suppose our threshold is πt and there is an entry

πj in the back end (equivalently j > t) with πj > πt. By definition of ascendee, πt−1 < πt,

and the subsequence πt−1 πt πj forms a 123.

We borrow the next few definitions from the recent paper by Vatter [37].

24

x

x

x

AAAAAA@

@@

Front End

xThreshold

x

x

£££££££££

Back End

Figure 2-2. The permutation 532614.

Definition 12. An interval in a permutation π is a set of consecutive indices

i, i +1, . . . , i+ r such that the set of values πi, πi+1, . . . , πi+r is a set of consecutive

integers. A decreasing interval is then a set of consecutive indices whose values appear

in decreasing order. Finally, a maximal decreasing interval is a decreasing interval

i, i + 1, . . . , i + r such that i − 1, . . . , i + r and i, . . . , i + r + 1 are not decreasing

intervals. Increasing intervals and maximal increasing intervals are defined analogously.

We denote an interval by its set of indices, written with set notation, or by its entries,

written in one-line notation.

Example 10. Let π = 3756124. Then, π contains the interval 2, 3, 4, i.e. 756. Indeed,

2, 3, 4 is a set of consecutive indices, and the set of values π2, π3, π4 = 7, 5, 6 is a set

of consecutive integers. Note that π also contains the maximal increasing interval 5, 6,i.e. 12.

25

Definition 13. Given a permutation π ∈ Sn with interval I = i, i + 1, . . . , i + r, we

define the deflation of π at I to be the reduction of π1 · · · πi−1 πi πi+r+1 · · · πn. Similarly,

the inflation of π at the index i by the permutation σ ∈ Sm is the permutation obtained

from π by increasing by m − 1 each entry greater than πi and replacing the entry πi with

the interval whose reduction is σ.

Example 11. The deflation of the permutation 264513 at the interval 645 is the reduction

is 2413. The inflation of the permutation 3124 at index 3 by the permutation 321 is

514326.

It was remarked above that the front end of a permutation is monotone decreasing.

Thus, the front end has a unique factorization as the concatenation of one or more

maximal decreasing intervals.

Example 12. Let π = 865321947. Then, the front end of π is 865321, the concatenation

of maximal decreasing intervals 8, 65, and 321.

2.3 Generating Trees

A tree is simply a connected, acyclic simple graph, or equivalently a connected simple

graph on n vertices with n − 1 edges. A forest is a collection of trees. A rooted, labeled

tree is a tree with an assignment of labels (typically non-negative integers) to the nodes

and one node designated the root, giving an orientation to the entire graph. We say that

a node y is a child of the node x if the final edge in the unique path from the root to y

is the edge x, y. Likewise, we call x the parent of y and define the depth of y to be the

number of children. The descendants of y is the set of nodes x such that the unique path

from the root to x passes through y. Finally, we call the set of all nodes whose unique

path to the root contains k edges the kth generation of the tree.

In [39] Julian West defines a generating tree as a rooted, labeled tree having the

property that the labels of the children of each node x can be determined from the label of

x itself. This leads to the characterization of a generating tree by the label of its root and

26

x

x

x

x

x

x

x

x

x

@@

@

@@

@@

@@

8

6

5

3

2

1

9

4

7

Figure 2-3. The permutation 865321947.

a set of succession rules which determine the number of children and labels of children for

each node of a given length and label.

The classic task for a combinatorial enumerologist is to determine the number of some

combinatorial objects of size n, perhaps further indexed with respect to some property or

some statistic k. Typically, one is presented with an initial object of some small size and a

recursion rule which says how many objects of each successive generation (objects of size

n + 1) can be created inductively from those of the previous generation (objects of size n).

Define the gth level-number of a tree to be the number of nodes in the gth generation.

Thus the generating tree is easily seen as a tool which lends itself quite readily to

combinatorial enumeration. We consider the nodes of our tree to be the combinatorial

objects themselves. There are many situations when the number of (n + 1)-objects which

27

x

x

x

x x x

x

´´

´´

´´

´´

QQ

QQ

QQ

QQQ´

´´

´´

´´

´

QQ

QQ

QQ

QQQ

root

parent

children

third generation of tree

Figure 2-4. A rooted tree.

can be generated from any n-object is all we need to know, so we might as well label each

node with its depth.

In [39] West begins with a trivial example, the complete binary tree. We begin with a

root with label (2). Our succession rule is that each node with label (2) has two children

also labeled (2).

Example 13 ([39], Example 1). The complete binary tree is determined by the set of rules

Root : (2)

Rule : (2) → (2)(2)

A long celebrated integer sequence is that of the Fibonacci numbers (Fn)n≥0 =

0, 1, 1, 2, 3, 5, . . ., where we have the initial assumptions F0 = 0 and F1 = F2 = 1, and each

additional number Fn is the sum of the preceding two numbers of the sequence. They will

play a role in our studies of pattern avoidance, so we show their generating tree as a less

trivial example of generating trees and a slow introduction to this sequence.

28

(2)

(2) (2)

(2) (2) (2) (2)

%%

%%

%%

%%

ee

ee

ee

ee

¢¢

¢¢

¢¢

¢¢

AAAAAAAA

¢¢

¢¢

¢¢

¢¢

AAAAAAAA

qqq

qqq

qqq

qqq

Figure 2-5. The complete binary tree.

Example 14 ([39], Example 3). The Fibonacci tree is determined by the set of rules

Root : (1)

Rules : (1) → (2)

(2) → (1)(2)

The observant reader undoubtedly noticed that our succession rules are not the same as the

recursion we gave to define our sequence. We verify that these rules are in fact equivalent

to the statement Fn+1 = Fn + Fn−1. For each n ≥ 1 we have a set of G(2)n objects labeled

(2) and a set of G(1)n objects labeled (1). Each object in either set produces an offspring of

size n + 1 with label (2). So, we have

G(2)n+1 = G(1)

n + G(2)n = Fn.

Furthermore, each object in G(1)n produces an offspring of size n + 1 with label (1), so

G(1)n+1 = G(2)

n = Fn−1.

where the second equality follows from our previous statement. We have thus accounted for

all Fn+1 objects, and we have our recurrence.

29

(1)

(2)

(1) (2)

(2) (1) (2)

%%

%%

%%

%%

ee

ee

ee

ee

¢¢

¢¢

¢¢

¢¢

AAAAAAAA

qqq

qqq

qqq

Figure 2-6. The Fibonacci tree.

For a detailed exposition on the use of generating trees in the study of pattern

avoidance, see [39], [12], [36], [10] and [24].

Here we define the generating trees which will be used throughout. These definitions

depend on the patterns σ and q which are being avoided, so we assume the patterns to

be given. This will be clear from context. First we explain the motivations. Recall our

′ notation. For a pattern σ ∈ Sm, the pattern σ′ ∈ Sm+1 is obtained by prepending σ

with the entry (m + 1). The fundamental question here is whether various limits (or

limit suprema) for the number of permutations which avoid some pattern σ are the same

as those which avoid σ′ (assuming for now that the limits exist). It was noted above

that σ-avoidance implies σ′-avoidance, but there are σ′ avoiders which contain σ. So,

our question boils down to just how many of these there are, in particular what are

the asymptotics of these permutations with respect to the set of σ-avoiders. We would

30

like to make use of generating trees to study the set of σ′-avoiders which contain σ, i.e.

π : π avoids σ′ \ π : π avoids σ.For each n ≥ 0, let Tn = Tn(q, σ) be the set of all (q, σ)-avoiding permutations of

length n, enumerated by tn. We construct the generating tree T = T (q, σ) whose nth level

is Tn. A permutation π ∈ Tn+1 is a child of γ ∈ Tn if and only if π can be obtained by

inserting n + 1 at one of the n + 1 open sites in γ. Similarly, for each n ≥ 0, let Un be

the set of all (q, σ′)-avoiding permutations of length n, enumerated by un, and construct

the generating tree U = U(q, σ) with levels Un and succession defined as for T . Atkinson’s

question is thus answered in the affirmative for some pattern by showing that un does

not grow asymptotically faster than tn. We will focus our attention on the sets Wn of

all (q, σ′)-avoiding permutations of length n which contain at least one copy of σ, i.e.

Wn = Un\Tn, enumerated by wn. This motivates the forest W = U\T . Note that each

tree in W is rooted at a q-avoiding permutation which avoids σ′ and contains σ but whose

parent in U avoids σ. For q = 123 and σ = 132, we have the tree shown in Figure 2-7.

An active site in a permutation is a valid insertion point, that is, a site where we

can insert n + 1 and obtain a child which is still in the current generating tree, so for

our purposes an active site is such that the insertion will not cause an occurence of any

pattern which we seek to avoid. The depth of a permutation π is the number of active

sites in π, equivalent to the notion of depth defined above on generating trees. We note

that the depth depends on both the permutation itself and on the tree, specifically the

pattern being avoided which determines the tree.

2.4 “Hat” Notation

Given a permutation π and a pattern σ, we denote by σ any copy of σ in π. For

each index p, we use the notation σp to refer to an entry which serves the role of σp in

some σ. As there may be more than one copy of the pattern σ in some permutation, σp

generally does not refer to any specific entry. For example, with σ = 132, the permutation

π = 24135 contains one σ, namely 243. In this case σ1 refers to the entry π1 = 2, σ2 refers

31

1 (2)%

%%

%%

%%%

21 (3)¡

¡¡

¡¡

¡¡¡

321 (4)¡

¡¡

¡¡

¡¡¡

4321 (5)

££

££

££

££3421 (1)

BBBBBBBB

3241 (1)

@@

@@

@@

@@3214 (1)

BBBBBBBB231 (1)

4231 (2)

@@

@@

@@

@@213 (1)

4213 (2)

ee

ee

ee

ee12 (2)

BBBBBBBB312 (2)£

££

££

£££

4312 (3)

BBBBBBBB3412 (1)

qqq

qqq

qqq

qqq

qqq

qqq

qqq

qqq

Figure 2-7. T(123,132).

to π2 = 4, and σ3 refers to π4 = 3. On the other hand, the permutation 1432 contains

three σ’s, namely 143, 142, and 132. In this case we can refer to the σ1, the entry 1, but

we have several σ2’s and several σ3’s. It should also be noted that one entry could be both

a σi and a σj for some i 6= j.

Example 15. With σ = 132, in the permutation π = 25431, π3 = 4 is the σ2 = 3 of the

subsequence 254 as well as the σ3 = 2 of the subsequence 243.

Often we will determine the depth of a permutation π by the index of the rightmost

σ1 in π, by which we mean the entry with the greatest index of those entries πk such that

there exists a σ which begins at πk.

Example 16. For σ = 231 and π = 43521, the entries π1 and π2 are both σ1’s, and we call

π2 the rightmost σ1.

32

Any confusion over hat notation should fade upon seeing its motivation in the

following proofs. As long as we use the ˆ notation carefully, the meaning should always be

clear.

2.5 Monotone Increasing Patterns q

For any pattern σ we have by construction that σ′ contains σ, so it follows immediately

from our introductory remarks on pattern avoidance that

sn(σ) ≤ sn(σ′).

for all n, and

L(σ) ≤ L(σ′).

Of course, these bounds also extend to

L(σ, Π) ≤ L(σ′, Π).

for any set of patterns Π, etc.

Let q be the increasing pattern 123 · · · r. We will show that the statement L(q, σ′) =

L(q, σ) holds if a pattern σ begins with its greatest entry. Our proof builds on Bona’s

proof for the case of single patterns, found in [4].

Definition 14. A left-to-right maximum is an entry in a permutation which is greater

than each entry to its left. The remaining entries of a permutation are those which are

not left-to-right maxima . A weak class (weak n-class) is a set of permutations (resp.

n-permutations) whose left-to-right maxima are the same and are in the same respective

positions.

Example 17. The permutations 32415 and 31425 both have left-to-right maxima 3, 4, and

5, at the first, third, and fifth entries, so we say that 32415 and 31425 are in the same

weak class or weak 5-class.

33

Example 18. The permutations 3412 and 2413 both have left-to-right maxima in the first

two positions, but as the maxima themselves are not the same, 3412 and 2413 are not in

the same weak 4-class.

Lemma 5. For each r ≥ 1, there exists a polynomial fr(x) such that for all n ≥ 1 the

number of weak n-classes of permutations with exactly r left-to-right maxima is less than

fr(n).

Proof. Fix n. We can easily count the weak n-classes with one or two left-to-right

maxima. An n-permutation with only one left-to-right maximum must begin with n,

so there is only one such weak n-class. Next we claim that there are(

n2

)weak n-classes

with exactly two left-to-right maxima. Indeed, pick two numbers a, b ∈ [n] with a < b.

Place the entry a in the first position, and place the entry n in the (n + 1 − b)th position.

Given such constraints, we can always place the remaining entries to find at least one

permutation in each weak n-class. For r > 2 we take the same approach but allow

overcounting. There are at most(

nr

)2ways to choose the entries and positions (numerical

and geographical values) for the r left-to-right maxima, as this includes all possible

sequences of left-to-right maxima, as well as some sequences which cannot possibly be the

set of left-to-right maxima of a permutation. As(

nr

)2is a polynomial in n for any fixed r,

the statement holds.

Corollary 2. For each r ≥ 1, the number of weak n-classes with less than r left-to-right

maxima is bounded by a polynomial in n.

Proof. For each 1 ≤ k < r, we have a polynomial upper bound fk(n) on the number of

weak n-classes with exactly k left-to-right maxima. The sum over all such k is clearly an

upper bound for the number of weak n-classes with fewer than r left-to-right maxima. As

a sum of (a fixed number of) polynomials is a polynomial, we are finished.

34

Proposition 19. Let q be the ascending pattern 1 2 · · · r, and let σ be a permutation which

begins with its greatest entry. Then,

L(q, σ′) = L(q, σ).

Proof. We first note that a permutation which avoids q has fewer than r left-to-right

maxima. Now, if a permutation avoids σ′, then its remaining entries avoid σ. Indeed, by

definition each remaining entry is preceded by a left-to-right maximum. If the remaining

entries of a permutation contain σ, then we may prepend this σ with any left-to-right

maximum which is to the left of and greater than σ1 to obtain a σ′, as σ1 is itself greater

than all other entries of σ by the fact that σ begins with its largest entry.

Therefore, we can overcount n-permutations which avoid σ′ by multiplying the

number of weak n-classes which have fewer than r left-to-right maxima by the number of

possible (q, σ)-avoiding permutations of the remaining entries. By Corollary 2 the number

of such weak n-classes is at most a polynomial function f(n). Therefore our overcount of

n-permutations is f(n)sn−1(q, σ). We are now in position to take our limits.

L(q, σ′) = lim supn→∞

sn(q, σ′)1/n

≤ lim supn→∞

(f(n)sn−1(q, σ))1/n

= lim supn→∞

f(n)1/nsn−1(q, σ)1/n

= lim supn→∞

1 · sn−1(q, σ)1/n

= L(q, σ).

Combined with the knowledge that L(q, σ′) ≥ L(q, σ), we are finished.

The following lemma from [4] and [5] provides an upper bound on the number of

permutations of length n which avoid the increasing pattern of length r.

35

Lemma 6. sn(1 2 · · · r) ≤ (r − 1)2n for all r, n ≥ 2.

Proof. We define a rank function on the entries of each permutation π ∈ Sn(12 · · · r) by

setting the rank of an entry πi to be the length of any maximal increasing subsequence

πj1πj2 · · · πi. Note that this generalizes the concept of the left-to-right-maxima, which are

precisely those entries with rank 1.

If an entry πi has rank t, then there exists some increasing subsequence πj1πj2 · · · πjt−1πi

of length t, so for any entry πk > πi with k > i, the rank of πk is at least t + 1, as we have

the increasing subsequence πj1πj2 · · · πjt−1πiπk. Therefore, for each 1 ≤ t ≤ n, we see that

the entries of rank t form a decreasing sequence.

There are at most n such decreasing sequences, and as sets of indices they form a

set partition of [n]. As π is assumed to be 12 · · · r-avoiding, π has no entry of rank r or

greater, and there are at most r − 1 blocks of our set partition. Assigning each entry of π

to one of r − 1 blocks, we may overcount and see that there are at most (r − 1)n possible

assignments of the indices to the blocks and at most (r − 1)n possible assignments of the

values to the blocks.

With this lemma in hand, we subtly alter another proof of Bona to achieve:

Proposition 20. Let q be the ascending pattern 1 2 · · · r, let σ be a pattern, and let c be a

constant such that for all n ≥ 1, sn(q, σ) ≤ cn. Then, for all n ≥ 1,

sn(q, σ′) ≤ (c + (r − 2)2)n−1.

Proof. In order for a permutation to avoid σ′, it is necessary that it avoid σ in the region

to the right of n. So, we may overcount the number of (q, σ′)-avoiding permutations

by choosing where to place n, which entries to place to the left of n such that they are

1 2 · · · (r − 1)-avoiding (because any 1 2 · · · ( ˆr − 1) among them could be postpended

with n to create a q), and how to arrange the entries to the right of n such that they

are (q, σ)-avoiding. We let k be the position of n, so there are(

n−1k−1

)possibilities for the

entries preceding n. By Lemma 6 there are at most (r − 1)2(k−1) possible permutations

36

of these entries. Finally, by our original hypothesis on Sn(q, σ), there are at most cn−k

possible permutations of the n − k entries which follow n. Altogether, these work out to

the binomial expansion

sn(q, σ′) ≤n∑

k=1

(n− 1

k − 1

)(r − 2)2(k−1)cn−k

= (c + (r − 2)2)n−1,

concluding the proof.

In particular, for q = 123, we have (r−2)2 = 1, so with the assumption sn(123, σ) ≤ cn

for all n, we find sn(123, σ′) ≤ (c + 1)n−1 for all n.

The Stanley-Wilf Conjecture (Marcus-Tardos Theorem) tells us that for any pattern

σ or set of patterns Σ there is such a constant c as in the above hypothesis. In the case of

avoiding a single pattern σ, Arratia showed in [2] that the sequence sn(σ)1/n is increasing.

However, there are sets of patterns Σ for which the sequence sn(Σ)1/n is not increasing.

Thus, taking c to be the least constant such that sn(Σ) ≤ cn for all n < N for some N , it

may be that there is a constant d < c such that sn(Σ) ≤ dn for all n > N . In particular,

the constant c may be significantly greater than L(Σ), so the new constant d is closer to

our limit and thus a better indicator of the asymptotic behavior of our sequence sn(Σ).

Such a situation motivates a strengthening of the previous proposition.

Proposition 21. Let q be the ascending pattern 1 2 · · · r, let σ be a pattern, and let d be a

constant such that for some N and all n ≥ N, sn(q, σ) ≤ dn. Then, there exists a constant

D such that, for all n ≥ N ,

sn(q, σ′) ≤ D(d + (r − 2)2)n−1.

Proof. The set sn(q, σ) : 1 ≤ n < N is finite, so it is bounded above, and we can choose

a constant D such that sn(q, σ) < Ddn for all 1 ≤ n < N . We retain all machinery from

the proof of the previous proposition, except instead of counting at most cn−k possible

37

permutations of the n − k entries which follow n, we count them by Ddn−k. Then our

expansion becomes

sn(q, σ′) ≤n∑

k=1

(n− 1

k − 1

)(r − 2)2(k−1)Ddn−k

= D

n∑

k=1

(n− 1

k − 1

)(r − 2)2(k−1)dn−k

= D(d + (r − 2)2)n−1.

2.6 The Pattern q = 123

In this section we restrict our attention to the 123-avoiding environment. Some

statements will be generalized to longer ascending patterns q in the following section.

Consider the active sites of a permutation π ∈ W . As any child of π is 123-avoiding,

there can be no active site to the right of the first ascendee. As any child of π is

σ′-avoiding, there can be no active site to the left of a σ1. However, the consecutive

sites satisfying these two criteria are all active. So, our understanding of depth reduces

to understanding where these two bounds lie. Furthermore, if π has depth d, inserting

n + 1 into one of the d active sites will not increase the depth. Indeed, n + 1 is inserted

to the left of the first ascendee of π and itself becomes the first ascendee in the child.

The rightmost σ1 of π remains in place in the child, so the child will have a rightmost σ1

which is at the same position or to the right of that of π. This demonstrates the following

lemma. Recall our definition of the threshold of a permutation as the leftmost ascendee of

the permutation.

Lemma 7. Let π ∈ W for some pattern σ. Then, the depth of π is the distance from the

righmost σ1 to the threshold, i.e. the difference of the index of the threshold and the index

of the rightmost σ1.

38

We proceed with a lemma which will be used extensively as it provides a polynomial

bound for the level-numbers of each tree in the forest W .

Lemma 8. Let σ be a pattern of length m which does not begin with m. Let π be a

(123, σ′)-avoiding permutation with depth d which contains σ but whose parent avoids σ.

Then, the number of descendants of π at the jth generation, i.e. the jth level-number of the

tree in W rooted at π, is bounded above by

(d + j

d

),

which is a polynomial in j of degree d.

Proof. (We are in fact bounded by the lesser polynomial(

d+j−1d−1

), however we dropped the

−1’s for neatness.) Let n be the length of π. First we note that π does not begin with n.

Indeed, as the parent of π has no σ but π has a σ, n is the m in any σ in π. π contains a

copy of σ, which does not begin with its largest entry, so n appearing to the left of any σ

would complete a σ′, contradicting the assumption that π is σ′-avoiding. In fact, by this

argument we see that n is the threshold and σ must contain entries in the front end of π.

We are trying to show that the number of descendants at each generation is bounded

by our polynomial, so we might as well assume the worst case scenario, that inserting at

the kth active site always produces a child of depth k, i.e. that the rightmost σ1 of a child

is always at the same position as the rightmost σ1 of the parent. Then, for all 1 ≤ k ≤ d,

we in fact have the succession rules

Root : (d)

Rule : (k) → (1)(2) · · · (k)

An example of a tree in such a forest is shown in Figure 2-8.

39

42153 (3)´

´´

´´

´´

´´

´´

462153 (1)

4762153qqq

426153 (2)¢

¢¢

¢¢

¢¢¢

4726153qqq

AAAAAAAA

4276153qqq

QQ

QQ

QQ

QQ

QQ

QQ421653 (3)

¢¢

¢¢

¢¢

¢¢4721653

qqq

BBBBBBBB

4271653qqq

@@

@@

@@

@@4217653qqq

Figure 2-8. Tree in W(123,231) rooted at 42153

Denoting by aj,k the number of permutations at the jth level with depth k, 0 ≤ j and

1 ≤ k ≤ d, we have the recursive system

a0,d = 1

a0,k = 0 ∀k 6= d

aj,k =d∑

t=k

aj−1,t ∀j ≥ 1, 1 ≤ k ≤ d.

From this recursive system we see that aj,k =(

d−k+1+jd−k+1

). For each d and j we may

sum over all k to attain the level-number(

d+j−1d−1

). However, a combinatorial proof is

preferable. We know that(

d+j−1d−1

)counts j-multisets of [d], and such a multiset written in

nonincreasing order spells out the order of active sites chosen in the lineage from π to a

permutation of length n + j.

40

Definition 15. ([5] Definition 5.33) A permutation π is called layered if it can be written

as the concatenation q1q2 · · · qk where each qi is a decreasing sequence of consecutive

integers and the leading entry of qi is smaller than the leading entry of qi+1 for 1 ≤ i ≤k − 1.

Example 22. An example of a layered permutation is 213654, which has three layers, as

shown in Figure 2-9.

x2

x1

x3

x6

x5

x4

@@

@@

@@¢

¢¢¢¢¢¢¢¢¢¢¢

££££££££££££££££££@

@@

@@

@@

@@

@@

@

Figure 2-9. The layered permutation 213654 with layers 21, 3, and 654

Many results are known concerning pattern avoidance and pattern packing for layered

permutations. One can see Section 5.2.2 of [5], [27], [6], and [20]. Our next result is on

layered patterns with just two layers, equivalently non-monotone layered patterns which

avoid 123.

Proposition 23. Let σ be a pattern of the form

d (d− 1) · · · 1 m (m− 1) · · · (d + 1)

41

for some d ≥ 1. Then,

L(123, σ′) = L(123, σ).

Proof. Note that any permutation which contains σ has depth at most d. In fact, we will

use a stronger property of the roots of the trees in W in which the depth is exactly d.

Indeed, let the N -permutation ρ be such a root. We know that ρ contains at least one

copy of σ, while the parent of ρ in U avoids σ. As the only change between the parent

of ρ and ρ is the insertion of N , N must act as m in any σ in ρ. This implies that there

is a set of d entries preceding N which serve as the d, ˆ(d− 1), . . . , 1 in a σ in ρ. In fact,

as the front ends of both σ and ρ are decreasing (because σ and ρ are 123-avoiding),

and as every front end entry of σ is less than every back end entry of σ, the d entries

immediately preceding N serve these roles. Therefore, the rightmost σ1 cannot be more

than d positions to the left of N . As N is the only m in ρ, the entry which is d positions

to the left of N is clearly the rightmost σ1.

That there is a fixed depth which applies to every root in W makes our job easy.

For each N , the total number of trees in W whose roots have length N is less than

NuN−1, as each such root in W has a parent in U on the (N − 1)st level of U , and an

(N − 1)-permutation can have depth at most N . Each such tree has level-numbers which

are bounded above by the polynomial(

d+n−N−1d−1

). So, the nth level-number of W , wn, is

bounded by

42

wn

n∑N=m

NuN−1

(d + n−N − 1

d− 1

)

≤ un

n∑N=m

N

(d + n−N − 1

d− 1

)

≤ un

n∑N=m

n

(d + n−m− 1

d− 1

)

≤ unn2

(d + n−m− 1

d− 1

),

un times a fixed polynomial f(n), a polynomial which only depends on σ. With the fact

that sn(123, σ) ≤ sn(123, σ′) for all n, we have

u1/nn ≤ w1/n

n ≤ u1/nn (f(n))1/n .

The nth roots of this fixed polynomial approach 1, so we may apply the Squeeze Theorem

to achieve

lim supn→∞

w1/nn = lim sup

n→∞u1/n

n .

We just saw that, if the front end is composed of entries which are all lower than

any entry in the back end, then there is a polynomial function which depends only on the

pattern and dictates the number of descendants for any permutation. Now consider how

the situation changes if there is more than one maximal decreasing interval in the front

end. Our shortest and prototypical example is 3142. The critical difference here is that

our depth is no longer bounded as it was for the patterns considered in Proposition 23.

For example, for any d > 2, we have the (123, 3142)-avoiding permutation (d + 1) (d −1) (d − 2) · · · 1 (d + 2) d, which has depth d. This means we can are not guaranteed a

fixed bound to the degrees of our generating polynomials of the level-numbers of our trees.

All is not lost, however, as we can show that for any of a class of patterns σ and any k

there is an upper bound to the number of trees in W whose root has depth which is k

43

less than its length. For the pattern 3142 it follows easily from the fact that the depth

of a permutation in W is completely determined by the length of the second maximal

decreasing interval in the front end. First, we calculate L(123, 3142).

As was mentioned in our introduction to generating trees, a famous combinatorial

sequence is the Fibonacci numbers Fn defined for all n ≥ 0 by the recursive system

F0 = 0

F1 = 1

Fn+1 = Fn + Fn−1 ∀ n ≥ 1.

Multiplying both sides of the recursive equation Fn+1 = Fn + Fn−1 by xn, summing

over all n ≥ 1, and solving for the ordinary generating function F (x) =∑

n≥0 Fnxn, we

find

F (x) =x

1− x− x2=

∑n≥0

1√5

((1 +

√5

2

)n

−(

1−√5

2

)n),

from which we see that Fn is asymptotically approximated by the nth power of the

dominant term, the golden ratio,

Fn ∼(

1 +√

5

2

)n

.

Furthermore, as the golden ratio enjoys the property of being a solution to the equation

x2 = x + 1, so that(

1+√

52

)2

= 1+√

52

+ 1 = 3+√

52

, we immediately achieve an approximation

for every other Fibonacci number

F2n+1 ∼(

3 +√

5

2

)n

.

44

By iteratively applying the recursive equation for Fn, we achieve a recursion on F2n−1:

F2n−1 = F2n−2 + F2n−3

= F2n−3 + F2n−3 + F2n−4

= F2n−3 + F2n−3 + F2n−5 + F2n−6

= · · ·

= F2n−3 +n−1∑

k=0

F2k−1

= 2F2n−3 + F2n−5 + F2n−7 + · · ·+ F1.

Proposition 24. For all n ≥ 1,

sn(123, 3142) = F2n−1.

where Fn is the the nth Fibonacci number. Furthermore,

L(123, 3142) =3 +

√5

2.

Proof. As the number of (123, 3142)-avoiding permutations of lengths 1, 2, and 3 are

1,2, and 5, respectively, we see that the boundary conditions are satisfied. Writing sn for

sn(123, 3142), it suffices to show that our sequence satisfies the appropriate recursion,

sn = sn−1 +n−1∑

k=1

sn = 2sn−1 + sn−2 + · · ·+ s1.

Given a (123, 3142)-avoiding permutation of length n − 1, we can always prepend n, as n

can never be the first entry of a 123 pattern nor a 3142 pattern, so the new permutation

will remain (123, 3142)-avoiding. This is clearly an bijection between Sn−1(123, 3142) and

permutations in Sn(123, 3142) which begin with n, accounting for the first term, sn−1.

For a (123, 3142)-avoiding permutation π of length n which does not begin with n,

let j be the index of n, so πj = n, (and j ranges from 2 to n as π ranges over all such

permutations). As π is 123-avoiding, n is in fact the threshold. Therefore, the front end

45

of π consists of one maximal decreasing interval. Indeed, if the front end contains more

than one maximal decreasing interval, then there exist entries in the front end πi and πi+1,

with the entry (πi+1 − 1) in the back end, and the reduction of πi πi+1 n (πi − 1) is 3142,

contradicting our assumption on π.

As the front end of π is an interval, the deflation of π at the front end is a (123, 3142)-avoiding

permutation π of length n − j + 2. The largest entry of π is of course π2 = (n − j + 2).

Removing this entry, we obtain a (123, 3142)-avoiding permutation of length n − j + 1

(which may or may not begin with its largest entry).

We claim that this is a bijection. Indeed, given a permutation γ ∈ Sn−j+1(123, 3142),

the inflation of γ at the first index by the permutation (j − 1) (j − 2) · · · 1 is a

(123, 3142)-avoiding permutation γ of length n − 1. Inserting n at index j regains a

permutation in Sn(123, 3142), and this is an injection. Such a construction over all j

accounts for the terms sn−1 + · · ·+ s1 in our recursion.

The limit L(123, 3142) = 3+√

52

follows immediately from the above discussion of the

asymptotics of the Fibonacci numbers.

Proposition 25. Let σ be the pattern 3142. Then, the limit

limn→∞

sn(123, σ′)1/n

exists and

L(123, σ′) = L(123, σ),

i.e. L(123, 53142) = L(123, 3142) =3 +

√5

2.

Proof. Suppose ρ is the root of a tree in W and has depth d and length n. Recall that

being a root implies ρ contains one copy of σ, but the parent of ρ in U is σ-avoiding. We

will map ρ to a permutation of length n−d+1 by removing certain superfluous entries and

reducing. By superfluous we mean that there is a subset of the entries of ρ which do not

46

add any information to the identity of ρ in the sense that they are completely determined

by the other entries and the fact that ρ has length n and depth d.

Consider the structure of the front end of ρ. We claim that its factorization consists

of exactly two maximal decreasing intervals. Indeed, n is both a 4 and the threshold of

ρ, so there exist 3 and 1 in the front end and a 2 in the back end, so the front end itself

cannot be a decreasing interval and thus contains at least two maximal decreasing

intervals. On the other hand, suppose the front end of ρ contains three maximal

decreasing intervals I1, I2, and I3. Then, we can choose indices i1 ∈ I1, i2 ∈ I2, and

i3 ∈ I3 and indices j < k such that ρi1 > ρj > ρi2 and ρi2 > ρk > ρi3 , and these five entries

form a σ′, contradicting the assumption that ρ is σ′-avoiding.

We will proceed by removing the superfluous entries, reducing accordingly at each

step. We know n contained no information, because its location is determined by the

depth d of ρ and the location of the rightmost σ1. Remove n, ridding the permutation of

any σ. We retain the first (leftmost) maximal decreasing interval, but we do not need all

the entries from the second maximal decreasing interval, as we can detect their presence

(or absence) from d. So, deflate this interval. We now have a permutation of size n− d + 1

which we call the prototype of ρ.

Definition 16. Given a root ρ of W , the permutation obtained by deflating the second

maximal decreasing interval of ρ and removing its largest entry is called its prototype.

By this process, we map our roots of length n and depth d into the set of σ-avoiding

permutations of length n − d + 1 which contain at least two maximal decreasing intervals

in the front end and contain an ascent (we always leave a 1 and a 2). To see that the

mapping is injective, let γ be such an (n − d + 1)-permutation. We will construct the

permutation of length n and depth d of which γ is the prototype. Let p be the lowest

index in the second maximal decreasing interval in the front end of γ. Inflate γ at p by

the decreasing permutation (d − 1) (d − 2) · · · 1, obtaining a permutation of length n − 1.

Finally, insert n before the (p + d − 1)st entry. As this reverse mapping holds for any

47

σ-avoiding permutation which contains an ascent and at least two maximal decreasing

intervals in its front end, we can overcount them as follows. We know that Sk(123, 3142)

is F2k−1, F2k−1 < βk for all k ≥ 1, and F1/k2k−1 → β, where β = (3 +

√5)/2. So, for any

N , the number of roots in W of length N and depth d is less than F2(N−d+1)−1 and, letting

k = N − d + 1 be the length of the prototype of each root, they each have level-number

polynomial

(d + j − 1

d− 1

)=

(N − k + 1 + j − 1

N − k + 1− 1

)

=

(N − k + j

N − k

)

=

(N − k + (n−N)

N − k

)

=

(n− k

N − k

).

We can now overcount all permutations in W by the roots of their trees and the

prototypes of the roots:

wn <

n−1∑

k=3

F2k−1

n∑

N=k+1

(n− k

N − k

)

<

n−1∑

k=3

F2k−12n−k

<

n−1∑

k=3

βk2n−k

<

n−1∑

k=3

βn

= (n− 3)βn.

In the end, we see that indeed the nth root approaches β = L(123, 3142).

48

As was mentioned prior to this proposition, the pattern 3142 makes our work easier

because the front end of the pattern itself is the concatenation two maximal decreasing

intervals. The above proof provided a warmup for the main result of this section. The

following lemma allows us to work with longer front ends.

Lemma 9. Let σ ∈ Sm be a pattern with a decreasing back end. Let ρ ∈ Sn be a root of

W . Then, the front end of ρ has at most m maximal decreasing intervals.

Proof. We first note that if i, i + 1, . . . , i + r is a maximal decreasing interval in the front

end of ρ, with i ≥ 2, then by maximality and the fact that the front end is decreasing,

ρi + 1 is in the back end of ρ.

Assume the front end of ρ has at least m + 1 maximal decreasing intervals. We show

that this implies the parent of ρ in U contains σ, a contradiction.

Label the rightmost m + 1 maximal decreasing intervals in the front end of ρ by

I1, I2, . . . , Im in order of their greatest entries. So, the front end of ρ is the concatenation

ρ1 · · · Im+1ImIm−1 · · · I1 or Im+1Im · · · I1 if ρ ∈ Im+1.

For each 1 ≤ i ≤ m, if the entry i is in the front end of σ, we choose an entry in Ii

to be our i. If the entry i is in the back end of σ, we choose an entry in the back end of ρ

which is greater than every entry of Ii and less than every entry of Ii+1 (if i < m) to be

our i. Doing so, we achieve our σ.

Example 26. If σ = 231, and ρ = 97642531, then our four labeled maximal decreasing

intervals are I1 = 2, I2 = 4, I3 = 76, and I4 = 9. Thus the front end of ρ is I4I3I2I1.

We construct our σ. As the entry 3 is greater than every entry of I1 and less than every

entry of I2, we choose 3 as our 1. Next we choose 4 as our 2. Finally, as 5 is greater than

every entry of I3 and less than every of I4, we choose 5 as our 3, and 453 indeed reduces

to σ = 231.

We now answer M. Atkinson’s question in the affirmative for patterns σ which avoid

123 and contain the entry 1 in the front end, noting that this restriction implies the back

49

end is decreasing. Any such pattern is not decomposable, so we do know that the limit L

exists and the sequence sn(123, σ)1/n is non-decreasing.

Theorem 27. Let σ ∈ Sm be a pattern with the entry 1 in its front end. Then,

L(123, σ′) = L(123, σ).

Proof. We generalize the proof of Proposition 25. Let d0 denote the depth of σ. Note that

any root in W contains in its front end a copy of the front end of σ and therefore has

depth which is at least d0. Let d ≥ d0. We begin with a prototype γ, a permutation in

Sk(123, σ) for some k, and insert d− d0 entries in γ to construct a root in W with depth d.

Let γ ∈ Sk(123, σ) have a child ρ(0) in U which is a root in W and has the same depth

as σ, d0. So, there exists an index t such that inserting k + 1 at t in γ gives us ρ(0), and

γ contains a copy of σ \ m. If d = d0, we take ρ(0) to be our root in W . Otherwise, we

build ρ(d−d0) inductively. Let r be the index of the rightmost σ1 of ρ(0), and note that by

construction t is the index of the threshold of ρ(0), and γ1γ2 · · · γt−1 = ρ(0)1 ρ

(0)2 · · · ρ(0)

t−1.

Besides the threshold, we will insert each entry using one of two methods. One

method is to insert at a maximal decreasing interval, i.e. choose an index in a maximal

decreasing interval in the front end of ρ(j) and inflate the permutation at that index by the

permutation 21.

Example 28. Let ρ(0) = 6431752. We may insert an entry by inflating ρ(0) at index 2 to

obtain 75431862, where the bold-faced 5 is the inserted entry.

The second method may be applied for any pair of entries in the back end of σ which

have consecutive indices and consecutive values. Suppose ρ(0)p and ρ

(0)p+1 correspond to such

a pair in the copy of σ in ρ(0) and are the least such pair. By being the least such pair we

mean that if ρ(0)p+1 = σq, and the greatest entry in the front end of σ which is less than σq

is σi, then the maximal decreasing interval containing σi contains the greatest entry in the

front end of ρ(0) which is less than σq.

50

As the depth of ρ(0) is the depth of σ, there is no entry in the front end of ρ(0) whose

value is between those of ρ(0)p and ρ

(0)p+1, so we have ρ

(0)p = ρ

(0)p+1 + 1. Let ρ

(0)i be the greatest

entry in the front end of ρ(0) which is less than ρ(0)p+1. Then, our second method of insertion

will be to insert the entry ρ(0)p at the index i and increase by one every entry which is at

least as large as ρ(0).

Example 29. Let σ = 41532 and ρ(0) = 521643. We may insert an entry in the front end

of ρ(0) to obtain 6421753, where the bold-faced 4 is the inserted entry.

Of course, in ρ(0) the pair ρ(0)p and ρ

(0)p+1 have consecutive values, and this may not

be the case in ρ(j) for greater values of j, as we may have already inserted at an index

q using the second method and corresponding to the pair ρ(0)p and ρ

(0)p+1. In this case, we

inflate ρ(j) at the index q by the permutation 21.

By Lemma 9, we know that the front end of ρ(0) has at most m maximal decreasing

intervals. In fact, there are exactly M maximal decreasing intervals in the subsequence

γr+1 · · · γt−1 for some M < m. Suppose there are P pairs of entries for which we may

apply our second method above, so altogether we have M + P choices for each insertion,

and we order them as C1, C2, . . . , CM+P . So, let x1, . . . , xd−d0 be a multiset of [M + P ]

with x1 ≥ x2 ≥ · · · ≥ xd−d0 . For each 1 ≤ j ≤ d − d0, inductively define ρ(j) to be

the insertion into ρ(j−1) determined by the choice Cxj. So, we obtain ρ(d−d0) by inserting

entries into some of the maximal decreasing intervals of ρ(0). As these entries were inserted

between the rightmost σ1 and the threshold, we have that ρ(j) has depth d0 + j for each j,

and each one is a root in W . Altogether, we obtain our root of depth d, ρ = ρ(d−d0).

Now, we have our root and we can overcount the (123, σ′)-avoiding permutations of

length n. We are in the same situation as in the proof of Proposition 25 except that we’ve

made(

M+P+d−d0−1M+P−1

)choices of how to insert the entries into our prototype to obtain our

root. Of course, M + P ≤ m and d− d0 < n, so this is bounded above by(

m+nm

). Note that

since σ contains the pattern 132, L(123, σ) ≥ L(123, 132) = 2. Altogether, we have

51

wn <

(m + n

m

) n−1∑

k=3

sk(123, σ)n∑

N=k+1

(n− k

N − k

)

<

(m + n

m

) n−1∑

k=3

sk(123, σ)2n−k

<

(m + n

m

) n−1∑

k=3

sn(123, σ)

=

(m + n

m

)(n− 3)sn(123, σ).

Taking the limit,

limn→∞

w1/nn ≤ lim

n→∞

(m + n

m

)1/n

(n− 3)1/nsn(123, σ)1/n

= limn→∞

sn(123, σ)1/n

= L(123, σ).

Experimental results support the following conjecture of Atkinson.

Conjecture 30. Let σ be any pattern. Then,

L(123, σ′) = L(123, σ).

52

CHAPTER 3PATTERN PACKING

3.1 General Pattern Packing

In the previous section we posed anew some classic questions in the field of pattern

avoidance with the restriction that all permutations considered are given to avoid some

increasing pattern q. Along the same line, we can consider the question of pattern packing

on the set of permutations which avoid q.

We define the function pat on all permutations by letting pat(π) denote the number

of distinct patterns contained in the permutation π. (Note that we say distinct to avoid

confusion with the number of unique patterns, an entirely distinct (and unique) area of

research.)

Example 31. We have pat(1432) = 7, as 1432 contains the patterns ∅, 1, 12, 21, 132, 321, 1432

and no others.

An easy result on the function pat follows.

Proposition 32. Let q be the monotone increasing permutation 12 · · · r. Then,

pat(q) = r + 1.

Proof. As q is increasing, so is every subsequence, so the only patterns we find in q are the

increasing patterns of length 0, 1, . . . , r.

In the previous chapter we discussed equivalence classes for pattern avoidance. If the

permutation π contains the pattern σ, then π−1 contains σ−1. Similarly for πR and πC .

From this we immediately achieve for all permutations π

pat(π) = pat(π−1) = pat(πR) = pat(πC).

So, π, π−1, πR, and πC are in the same equivalence class with respect to the function pat.

Next, we define the function maxpat over all nonnegative integers by

maxpat(n) = max pat(π)|π ∈ Sn.

53

An obvious upper bound to maxpat(n) is the total number of subsequences of a

permutation, 2n. We are therefore interested in the asymptotics of maxpat(n), or the

growth rate of maxpat(n)1/n as n → ∞. This question was originally posed by Herb Wilf

at the Conference on Permutation Patterns, Dunedin, Otago, New Zealand in 2003. He

presented the following class of permutations which give a lower bound for maxpat(n).

To construct Wilf’s class, begin with the empty permutation, which we label W (0),

and the permutation 1, which we label W (1). Inductively assume that we have constructed

W (n−1). If n is even, we postpend W (n−1) with n, i.e. W (n) = W (n−1)n. If n is odd, we

increase each entry by one and postpend W (n−1) with 1. The first few permutations of our

class are evidently

W (1) = 1,

W (2) = 12,

W (3) = 231,

W (4) = 2314,

W (5) = 34251,

W (6) = 342516.

See Figure 3-1 for an example. The claim is that for all n ≥ 1, pat(W (n)) ≥ Fn,

where Fn is the nth Fibonacci number, as defined in the previous section.

Proposition 33. (Wilf) For all n ≥ 0, we have the lower bound

maxpat(n) ≥ Fn.

Furthermore, we have the bounds

1 +√

5

2≤ lim inf

n→∞maxpat(n)1/n ≤ lim sup

n→∞maxpat(n)1/n ≤ 2.

54

x3

x4

x2

x5

x1

x6

¡¡

¡¡

¡¡A

AAAAAAAAAAA£

£££££££££££££££££CCCCCCCCCCCCCCCCCCCCCCC¥¥¥¥¥¥¥¥¥¥¥¥¥¥¥¥¥¥¥¥¥¥¥¥¥¥¥¥¥

Figure 3-1. The permutation W (6) = 342516

Proof. We have pat(1) = 1 = F1 and pat(2) > 1 = F2, so the first statement follows

immediately for n = 1 and n = 2.

Now, for each n ≥ 3, the number of distinct patterns we obtain from subsequences

of length at least 2 which end with W(n)n is equal to the number of distinct patterns in

W (n−1), as any such pair of subsequences are of the form αW(n)n and βW

(n)n , where α

and β are necessarily distinct subsequences of W (n−1). Therefore, the number of distinct

patterns which we find in subsequences ending with the last entry of W (n) is pat(W (n−1)),

which is at least Fn−1 by induction.

Likewise, the number of distinct patterns of length at least 2 in W (n) which end with

W(n)n−1 is at least pat(W (n−2)) ≥ Fn−2 by similar arguments. It remains to show that these

two sets are disjoint. We note that, for n even (odd), each subsequence which ends at

W(n)n ends at its greatest (resp. least) entry, whereas each subsequence which ends at W

(n)n−1

55

ends at its least (resp. greatest) entry. These are mutually exclusive conditions by our

assumption that each pattern has length at least 2.

We see that indeed we have a lower bound which satisfies the Fibonacci recurrence.

From our previous work, we know that

Fn = Ω

((1 +

√5

2

)n).

The upper bound on the limit supremum is trivial for maxpat, which is necessarily

bounded above by the number of subsequences of an n-permutation, 2n.

Based on empirical evidence, (see sequence A088532 in the On-Line Encyclopedia of

Integer Sequences [32]), it seemed this number may “approach” the trivial upper bound of

2n (the number of all subsets of an n-permutation). In [15] this author constructed a class

of permutations over which it was shown that

limn→∞

maxpat(n)1/n = 2.

However, while confirming our suspicion that the nth root approaches 2, this result

still left open the possibility that maxpat(n)2n → 0 as n → ∞, i.e. the possibility that

maxpat(n) = o(2n). Recently Miller [25] and Albert et al. [14] independently proved with

a refinement of the original class from [15] and more delicate counting techniques that

indeed

limn→∞

maxpat(n)

2n= 1, i.e. maxpat(n) ∼ 2n.

In particular, in [25], Miller showed the wonderfully exact bounds

2n −O(n22n−√2n

)≤ maxpat(n) ≤ 2n −O

(n2n−√2n

).

Definition 17. We generalize the function maxpat by defining, for any pattern σ, the

function

maxpatσ(n) = max pat(π) : π ∈ Sn(σ).

56

Our task is to construct a class of permutations which maximize as much as possible

the function maxpatq(n) for increasing patterns q. First, consider what sort of upper

bounds we can find on maxpatσ(n) for a few simple patterns σ.

Proposition 34. For q = 123, we have the upper bound

maxpatq(n) ≤ t(n) :=n∑

m=0

min

(cm,

(n

m

)),

where cm is the mth Catalan number and counts 123-avoiding permutations of length m.

Furthermore, we have the limit

limn→∞

t(n)1/n = 2.

Proof. For each 0 ≤ m ≤ n, the number of m-patterns in any n-permutation π is at

most the number of subsequences of π of length m, i.e. the number of m-subsets of [n],(

nm

). Similarly, there are at most cm possible 123-avoiding m-patterns. As π itself is

123-avoiding, we know that any pattern contained in π is also 123-avoiding. So, for each

m, the number of m-patterns satisfies both these bounds, and the total number of patterns

is at most the sum of the minima over all m.

The significance of the above limit is that one may hope to pack just as many

patterns into 123-avoiding permutations as in the general case. Furthermore, for any

increasing pattern q = 12 · · · r with r > 3, the same upper bound holds, albeit trivial.

3.2 Pattern Packing in 123-avoiding Permutations

Proposition 34 gives us hope that, even with the 123-avoiding restriction, we

may pack as many patterns as we would like in a permutation of sufficient length.

Experimentation supports such a conjecture. Here we give constructions for q-avoiding

permutations with “many” patterns for increasing patterns q. Our constructions are

modeled on those of [15],[16], [14], and [25] and meet or surpass the original lower bound

57

given for the general case by Wilf,(

1+√

52

)n

. We note that Wilf’s original construction is

132-avoiding, so we already have it established that maxpat132(n) >(

1+√

52

)n

.

We begin with a family of 123-avoiding permutations which show that the limit

infimum of maxpat123(n)1/n is at least 1+√

52

as n →∞.

We define our permutations P (n) inductively. Let P (1) = 1 and P (2) = 12. For each

odd n ≥ 3, let P (n) = nP (n−1), i.e. P (n−1) prepended with n. For each even n ≥ 4, let P (n)

be P (n−1) with n inserted immediately after 1. So, we have

P (1) = 1,

P (2) = 12,

P (3) = 312,

P (4) = 3142,

P (5) = 53142,

P (6) = 531642.

So, each P (n) consists of two decreasing subsequences, and thus avoids 123. The next

proposition gives a lower bound on pat(n) with a proof similar to the proof given for

Proposition 33.

Proposition 35. For all n ≥ 1,

pat(P (n)) ≥ Fn.

Furthermore, we have the bounds

1 +√

5

2≤ lim inf

n→∞maxpat123(n)1/n ≤ lim sup

n→∞maxpat123(n)1/n ≤ 2.

Proof. Our induction hypothesis will be the stronger statement that P (n) contains at least

Fn−1 patterns containing the entry 1. We have that P (1) contains the pattern 1, and P (2)

contains the pattern 12, so the statement of the proposition holds for n = 1 and n = 2.

Assume the statement holds up to and including n ≥ 2. Then, P (n+1) contains a set A

58

of at least pat(P (n)) ≥ Fn−1 patterns which contain the entry (n + 1) and the entry 1.

P (n+1) also contains a set B of at least pat(P (n−1) ≥ Fn−2 patterns which do not contain

the entry (n + 1) but do contain the entries n and 1. We claim that the sets A and B are

disjoint. Indeed, if n + 1 is even (odd), then for each pattern in A the largest entry occurs

after (resp. before) the entry 1, while for each pattern in B the largest entry occurs before

(resp. after) the entry 1. Therefore, we have P (n+1) ≥ Fn−1 + Fn−2 = Fn. The bounds

follow immediately.

3.3 Pattern Packing in q-avoiding Permutations

Next we extend our construction to longer increasing patterns q. Suppose q =

12 · · · (r + 1) is the increasing pattern of length r + 1. We define the permutations Q(n)

inductively. Set Q(1) = 1, Q(2) = 12, . . . , Q(r) = 12 · · · r. For n ≥ r, our permutation

Q(n) consists of r maximal decreasing subsequences. If n + 1 ≡ p (mod r) with 1 ≤ p ≤ r,

we construct Q(n+1) by inserting n + 1 at the beginning of the pth maximal decreasing

subsequence.

Example 36. If q = 12345, i.e. r = 4, we have

Q(5) = 51234,

Q(6) = 516234,

Q(7) = 5162734,

Q(8) = 51627384,

Q(9) = 951627384.

Definition 18. For k ≥ 2, the Fibonacci k-step numbers, denoted F(k)n for n ≥ 0, are

defined by

F (k)n = 0 ∀ n ≤ 0,

F(k)1 = · · · = F

(k)k = 1,

59

and, for all n ≥ k + 1,

F (k)n = F

(k)n−k + F

(k)n−k+1 + · · ·+ F

(k)n−1.

The classical Fibonacci numbers are the special case k = 2.

It is well known that the limit

limn→∞

(F (k)

n

)1/n

is the unique real root greater than 1 of the equation

xk = xk−1 + xk−2 + · · ·+ x + 1.

This root, which we denote by αk, is called the k-anacci constant. We have that αk

increases with k and

limk→∞

αk = 2.

Proposition 37. Let q be the increasing pattern of length r + 1. Then,

pat(Q(n)) ≥ F(r)n−r+2.

Furthermore, we have the bounds

αr ≤ lim infn→∞

maxpatq(n)1/n ≤ lim supn→∞

maxpatq(n)1/n ≤ 2.

Proof. For each 1 ≤ n ≤ r, we only count the pattern Q(n) = 1 · · ·n itself and have

pat(Q(n)) ≥ 1 ≥ F(r)n−r+2.

For n ≥ r + 1, we count the patterns of subsequences which contain the entries 1, 2, . . . , r,

i.e. subsequences which contain the lowest entry of each maximal decreasing subsequence

in Q(n). Fix n ≥ r + 1 and assume by induction that, for all N < n, the number of such

patterns is at least F(r)N−r+2. For 0 ≤ i ≤ r − 1, let Ai be the set of subsequences of Q(n)

which contain the entries 1, 2, . . . , r and whose greatest entry is n− i, and let Bi be the set

of patterns of subsequences in Ai. We claim that the Bi’s are pairwise disjoint. Indeed, for

60

i 6= j, suppose σi ∈ Bi and σj ∈ Bj, where n − i ≡ p (mod r) and n − j ≡ q (mod r),

0 ≤ p, q ≤ r − 1. Then, the greatest entry of σi occurs after p ascents, and the greatest

entry of σj occurs after q ascents.

Also, for each i there is a bijection between the patterns in Bi and the patterns

counted for Q(n−i)−1. So, if we count all patterns in the union of the Bi’s, we have

pat(Q(n)) ≥ pat(Q(n−1)) + · · ·+ pat(Q(n−r)).

As we have the recursion for the Fibonacci r-step numbers, the induction follows.

It may be noted that these classes of permutations resemble the classes used in [15],

[16], [14], and [25], except that whereas in each of these papers the construction consisted

of k rows of k entries each, or a stripped down version of that, the permutations we are

using here have restricted row lengths (number of maximal decreasing subsequences) due

to the q-avoiding restriction.

It was noted that αk → ∞. From this fact we see that as we let r → ∞, our

constructions for increasing patterns of length r + 1 “approach” the constructions for the

general case and the lim infs approach 2.

3.4 Packing Density and Further Directions

Pattern packing is in a weak sense a dual concept to that of pattern avoidance. We

say the duality is only weak as pattern avoidance, when compared to the total number

of permutations, can be understood as the probability that a permutation will have the

property that it avoids the given pattern, whereas the question of pattern packing is

extremal, asking what is the maximum of a certain statistic, the number of patterns

contained in a permutation, over all permutations.

Perhaps the proper dual concept to pattern avoidance question is a question also

posed by Herb Wilf, at the 1992 SIAM Conference on Discrete Mathematics. For a given

pattern σ, how many copies of σ can a permutation contain? The answer depends on the

packing density of the pattern σ.

61

Definition 19. The packing density ρ(σ) of a pattern σ ∈ Sm is defined by

ρ(σ) = limn→∞

g(σ, n)(nm

) ,

where g(σ, n) is the maximum number of copies of σ in any n-permutation.

For more on the work in this area, please see the Ph.D. dissertations of Dan Warren

[38] and Alkes Price [27]. Herb Wilf has recently suggested that the notion of packing

density be examined in the q-avoiding environment.

Why do we raise the issue of these dualities? We saw in the section of pattern

avoidance that in the restricted environments of 123-avoiding permutations, or more

generally q-avoiding permutations, it is more likely that a randomly chosen permutation

avoids some pattern τ , or, critically, the chance of avoiding σ approaches or equals the

chance of avoiding σ′, a statement which is usually false in the general case. So, restricting

to q-avoiding permutations makes our life easier in that work.

However, pattern packing becomes more difficult. Indeed, if we consider all approaches

used to prove pattern packing or superpatterns, they take advantage of a lattice or

checkerboard structure in the class of constructions to bound maxpat. If we restrict the

length of the increasing sequences allowed, we lose this structure. So, the meta-question

is whether we lose our asymptotic limits or simply need more delicate techniques to see

them.

62

CHAPTER 4ASYMPTOTIC NORMALITY AND UNIFORMITY

4.1 Probability Theory

In this paper we only deal with discrete probabilities and their limits. For the sake of

simplicity we avoid the measure theoretic foundations of probability theory and define a

smaller class of random variables, giving associated properties which we will need in this

chapter.

Definition 20. A random variable X is a function from the unit interval [0, 1] to the

real line R. For each r ∈ R, we denote by P (X = r) the Lebesgue measure of the set

ω ∈ [0, 1] : X(ω) = r and call this the probability that X = r. The set of values

X(ω) : ω ∈ [0, 1] we call the range of X. A random variable X is called discrete if its

range is countable or if X has the weaker condition that there is a countable subset B of

the reals with P (X ∈ B) = 1. Finally, a 0-1 random variable or indicator random variable

is a random variable whose range is the set 0, 1.Definition 21. Let X be a random variable with finite range R. The mean or expected

value of X is defined by

E(X) =1

|R|∑r∈R

r.

The variance of X is E((X − E(X))2), which, by the linearity of expectation, is also given

by

Var X = E(X2)− (E(X))2.

The square root of the variance is the standard deviation.

We note that for a 0-1 random variable I, I2 = I, a fact that eases computation

of mean and variance for such variables. Indeed, given a 0-1 random variable I with

P (I = 1) = p, we have E(I) = p and Var I = p− p2.

As we will only be dealing (at least before taking limits) with random variables with

finite range, we will not develop all the measure theory behind these definitions. For this,

the reader is referred to the texts by Halmos [18], Taylor [35], and Chung [13]. For a

63

treatment of discrete probabilities, the reader is referred to the texts by Alon and Spencer

[1] and Charalambides [11].

4.2 Triangular Arrays

Definition 22. Given a random variable Y whose range R is a finite set of integers,

define its probability generating polynomial to be

pY (x) =∑r∈R

P (Y = r)xr.

By a triangular array of nonnegative real numbers (an,k) we mean a sequence of

numbers which are indexed by n = 0, 1, 2, . . ., and k = 0, 1, . . . , m = m(n) for some

function m defined on N, so for each fixed n there is a finite number of terms an,k. The

term triangular array comes to us from the probabilists. When the sequence consists

of non-negative integers, we use the term combinatorial distribution found in [19]. In

combinatorial applications an,k counts objects of size n with some statistic k. For example,

in the next chapter we will study the numbers Sn,k, counting set partitions of an n element

set into k blocks.

Given such a sequence an,k, we set sn = an,1 +an,2 + · · ·+an,m for each n and construct

a new sequence bn,k =an,k

sn, interpreting bn,k as the probability that a randomly selected

n-object has statistic k.

Example 38. A fundamental combinatorial distribution is that of the binomial coefficients(

nk

). It is well known that for each n we have the sum

(n0

)+

(n1

)+ · · · + (

nn

)= 2n. The

generating polynomial

(x + 1)n

2n=

n∑

k=0

(nk

)

2nxk

can be interpreted as the probability generating polynomial for a random variable Xn which

is the sum of n coin flips, i.e. Xn is the sum of n independent 0-1 random variables Yn,k,

k = 1, 2, . . . , n, with

P (Yn,k = 1) =1

2

64

for all n and k.

The term triangular array can be seen in the well-known Pascal’s triangle listing the

binomial coefficients:

1

1 1

1 2 1

1 3 3 1

1 4 6 4 1

......

......

......

We will only deal here with triangular arrays whose generating polynomials have

real, non-positive roots only. The motivation for this will become clear when it is seen

what strong statements can be made concerning such sequences and the clever, powerful

techniques used to deal with them.

4.3 Asymptotic Normality

For a sequence of random variables Xn such that, for each n, Xn is the sum of

m = m(n) independent, identical random variables, with m → ∞, the associated

distribution approaches the normal distribution, most easily understood as the limit (in

distribution) of the binomial coefficients. Many significant results in probability theory are

of the form of Central Limit Theorems proving sufficient conditions for convergence to the

normal distribution.

Definition 23. For a random variable X, let X denote the normalized random variable

X−E(X)√Var(X)

. For a sequence of random variables Xn, we write Xn → N(0, 1) to mean that Xn

converges in distribution to the standard normal variable. Finally, we say that a sequence

of random variables Xn is asymptotically normal if Xn → N(0, 1).

We have the following theorem, appearing as Theorem 6.7.11 of [35].

65

Theorem 39 (Central Limit Theorem for Triangular Arrays). Given a triangular

array of random variables Xn,1, . . . , Xn,m with respective variances σn,1, . . . , xn,m, let

σ2n = σ2

n,1 + · · · + σ2n,m and Xn = Xn,1 + · · · + Xn,m. Then, the distribution

an,k

snconverges

weakly to the unit normal distribution if the Lindeberg condition is satisfied:

∀ε > 0m∑

k=1

1

σ2n

|x|≥σnε

x2Qn,k(dx) → 0 as n →∞,

where, for each k, Qn,k is the cumulative distribution function of Xn,k.

4.4 Asymptotic Uniformity

Another asymptotic property of some combinatorial distributions which provides us

with both qualitative and quantitative understanding of their asymptotic nature is the

concept of asymptotic uniformity. We build up to this with Bona’s definition of balance

([7], [9]).

Definition 24. For each positive integer q, we say that the combinatorial distribution an,k

is q-balanced if for each 0 ≤ r < q, we have the limit

limn→∞

∑k≡r an,k∑

k an,k

=1

q,

where all congruences are modulo q.

So, if an,k counts n-objects with statistic k, q-balance implies that there is eventually

equal probability of a randomly selected n-object to have k-statistic in any one of the

residue classes of q.

Example 40. We give a quick proof that the binomial coefficients are 2-balanced. Evalu-

ating their generating polynomial (x + 1)n at x = −1 gives us the difference of binomial

coefficients for even and odd values of k:

(−1 + 1)n =∑

(−1)k

(n

k

)=

k even

(n

k

)−

k odd

(n

k

).

As (−1 + 1)n = 0n = 0, we have that for all n ≥ 1 the set of evens and the set of odds each

account for exactly half of the total.

66

Of course, most cases are not so clean as this and we must examine the limits.

However, we will see that evaluating the generating polynomial at the qth root of unity as

above is integral to our work.

Definition 25. We say the combinatorial distribution an,k is asymptotically balanced if it

is q-balanced for all q ∈ P.

The following lemma and proof are based on those of Bona in [7] and [9]

Lemma 10. Let q ∈ P. Suppose the combinatorial distribution an,k has generating

polynomials pn(x) defined by

pn(x) =m∑

k=0

an,kxk.

Then the sequence an,k is q-balanced if for all 0 < t < q

limn→∞

pn(ζt)

pn(1)= 0, (4–1)

where ζ = exp 2πiq

, a primitive qth root of unity.

Proof. Let 0 ≤ r ≤ q − 1. For now we assume q divides m. Consider the sum

Sn,r(ζ) :=

q−1∑t=0

pn(ζt)ζ−tr (4–2)

=

q−1∑t=0

(m∑

k=1

an,k(ζt)k

)ζ−tr (4–3)

=m∑

k=1

an,k

q−1∑t=0

(ζk)tζ−tr (4–4)

=m∑

k=1

an,k

(q−1∑t=0

(ζk−r)t

). (4–5)

Enclosed in parentheses in (4–5) is a geometric sum, and we take advantage of the

property that the (k − r)-roots of unity add up to 1 to see that

q−1∑t=0

(ζk−r)t =

0 if ζk−r 6= 1.

q if ζk−r = 1.

67

Therefore, we have annihilated all terms in (4–5) for which q does not divide k − r, and

(4–2) reduces to

Sn,r(ζ) = q

m/q∑j=1

an,jq+r, or (4–6)

Sn,r(ζ)

qpn(1)=

∑m/qj=1 an,jq+r

pn(1). (4–7)

We recognize the right hand side of (4–7) as the function from the definition of q-balance.

So, q-balance is equivalent to this expression converging to 1q

for all r, and assuming (4–1)

we can drop our assumption that q divides m, and we have the limit

limn→∞

Sn,r(ζ)

qpn(1)=

1

qlim

n→∞

q−1∑t=0

pn(ζt)

pn(1)ζ−tr

=1

q

q−1∑t=0

ζ−tr limn→∞

pn(ζt)

pn(1)

=1

q(1 + 0 + · · ·+ 0)

=1

q.

4.5 Generating Polynomials with Real, Non-Positive Roots

Throughout this section we will assume that an,k is a triangular array of non-negative

reals whose generating polynomials have real, non-positive roots and set sn = an,0 + an,1 +

· · ·+ an,m. Also, Xn will refer to the random variable defined by

P (Xn = k) =an,k

sn

∀ k ≥ 0,

with probability generating polynomial

pn(x) =m∑

k=0

an,k

sn

xk.

68

The notion of asymptotic normality or normal convergence is that the normalized

sequencean,k

snconverges in distribution to the normal distribution. For each n we will

find a set of 0-1 random variables whose sum is Xn and use these to prove asymptotic

uniformity. We develop here the method attributed to Harper [19] and also used in [30].

By our assumption on the roots of our generating polynomials, we can factor the

probability generating polynomial pn(x) as

pn(x) =an,m

sn

1≤k≤m

(x + λn,k) = an,m

1≤k≤m

x + λn,k

1 + λn,k

,

with 0 ≤ λ1 ≤ λ2 ≤ · · · ≤ λm ∈ R. For each n, define the mutually independent 0-1

random variables In,1, . . . , In,m by

P (In,k = 0) =λn,k

1 + λn,k

, and

P (In,k = 1) =1

1 + λn,k

.

From this the mean and variance of each In,k follow immediately.

E(In,k) =1

1 + λn,k

.

Var In,k = E(I2n,k)− (E(In,k))

2

=1

1 + λn,k

− 1

(1 + λn,k)2

=λn,k

(1 + λn,k)2.

We have the respective probability generating polynomials gn,1(x), . . . , gn,m(x) given by

gn,k(x) = P (In,k = 0) + P (In,k = 1)x =λn,k

1 + λn,k

+1

1 + λn,k

x.

69

Now we can rewrite each probability generating polynomial pn(x) in terms of the

probability generating polynomials gn,k(x).

pn(x) = an,m

1≤k≤m

x + λn,k

1 + λn,k

= an,m

1≤k≤m

gn,k(x).

We have thus expressed Xn as the sum of the 0-1 random variables In,1, . . . , In,m. We

will be interested in the variance of Xn, which we denote σ2n. We are now in position to

state the following theorem of Bender [3] which follows from Theorem 39:

Theorem 41 (Bender, 73). Let the combinatorial distribution an,k be as above. Then, the

sequence of random variables Xn is asymptotically normal if and only if

σn →∞ as n →∞.

We have the following computation, where 1 ≤ j, k ≤ m for all j, k.

σ2n = E(X2

n)− (E(Xn))2

= E((∑

k

In,k)2)− (

k

E(In,k))2

=∑

k

E(In,k) + 2∑

j<k

E(In,jIn,k)− 2∑

k

(E(In,k))2 −

j<k

E(In,j)E(In,k)

=∑

k

E(In,k)−∑

k

(E(In,k))2

=∑

k

Var In,k

=∑

k

λn,k

(1 + λn,k)2.

With Theorem 41 and keeping the above notation, we have shown the following.

Lemma 11. With real, non-positive zeros, asymptotic normality is equivalent to the

divergence

limn→∞

m∑

k=0

λn,k

(1 + λn,k)2= ∞.

70

4.6 Asymptotic Normality Implies Asymptotic Uniformity

With the results of the above two sections, in particular Lemmas 10 and 11, we can

state the main result of this chapter.

Theorem 42. Let the combinatorial distribution an,k have generating polynomials with

real, non-positive roots only. Then, asymptotic normality implies asymptotic uniformity.

Proof. We retain the notation of Section 4.5. Assuming asymptotic normality, Lemma 11

tells us

m∑

k=1

λn,k

(1 + λn,k)2→∞ as n →∞. (4–8)

By Lemma 10, it suffices to show that this divergence implies

pn(ζr)

pn(1)→ 0 (4–9)

as n →∞ for all q and all 0 < r < q, where ζ = exp 2πiq

. If ζr = −1 is a root of pn(x), then

pn(ζr) = 0, so we can assume here that this is not the case in order to avoid pathologies.

Taking logarithms, (4–9) is equivalent to the condition

m∑

k=1

(log (1 + λn,k)− log |ζr + λn,k|) →∞. (4–10)

By Lemma 12 below, there exists a constant α > 0 which depends only on ζr such that for

all n and k we have

log (1 + λn,k)− log |ζr + λn,k| > αλn,k

(1 + λn,k)2.

Thus,m∑

k=1

(log (1 + λn,k)− log |ζr + λn,k|) > α

m∑

k=1

λn,k

(1 + λn,k)2,

and (4–8) implies (4–10) which is itself equivalent to (4–9).

71

Lemma 12. Let z = eiθ for some 0 < θ < 2π. Then, there exists a constant α = α(z) such

that for all x ≥ 0,

f(x) = log (1 + x)− log |z + x| > αx

(x + 1)2.

Proof. Write f(x) as

f(x) =

∫ x+1

1

dt

t−

∫ |x+z|

1

dt

t=

∫ x+1

|x+z|

dt

t.

The length of the domain of integration is the function d(x) = x + 1− |x + z|, and we have

f(x) >d(x)

x + 1.

So, we are interested in a constant α such that

d(x) > αx

x + 1

for all x ≥ 0. We first show that there is a constant α1 such that d(x) > α1 for all x > 1.

We have the formula |x + z| = (x2 + 2x cos θ + 1)1/2. Therefore, for all x > 1,

d′(x) =d

dx(x + 1− (x2 + 2x cos θ + 1)1/2)

= 1− 2x + 2 cos θ

2(x2 + 2x cos θ + 1)1/2

= 1−(

x2 + 2x cos θ + cos2 θ

x2 + 2x cos θ + 1

)1/2

> 0.

Therefore, d(x) is monotonically increasing for positive x, and for all x > 1 it is greater

than d(1), which we denote by α1.

Next we show there exist constants α2 and δ such that d(x) > α2 for all 0 < x < δ.

We have d′(0) = 1− cos θ, so there exists δ > 0 such that for all 0 < x < δ

d(x) > x1− cos θ

2.

72

Moreover, as 1x+1

< 1, for all 0 < x < δ,

d(x) >1− cos θ

2· x

x + 1,

so we let α2 = 1−cos θ2

. Now we have constants for x ∈ [0, δ) ∪ [1,∞). As [δ, 1] is a compact

set, there exists a constant α3 such that d(x) > α3x

x+1for all x ∈ [δ, 1]. Finally, set α =

min α1, α2, α3.

Example 43. The signless Stirling numbers of the first kind cn,k count n-permutations

with k cycles. For each n, we have the generating polynomial

Cn(x) =n∑

k=1

cn,kxk = x(x + 1) · · · (x + n− 1)

So, for each n, Cn(x) has the roots 0,−1,−2, · · · ,−n + 1. We have the following limit.

limn→∞

n∑j=1

j

(1 + j)2=

∑j≥1

j

1 + j· 1

j

>∑j≥1

1

2· 1

j

→∞

Therefore, by Lemma 11 and Theorem 42, the signless Stirling numbers of the first

kind are asymptotically normal and thus uniform.

73

CHAPTER 5ON THE ROOTS OF THE BELL POLYNOMIALS

5.1 Stirling Numbers of the Second Kind

This chapter is concerned with the generating polynomials for a classic combinatorial

distribution, the Stirling numbers of the second kind. Recalling the notation [n] =

1, 2, . . . , n, we begin with the definition of a fundamental combinatorial object.

Definition 26. For n ∈ P, a set partition of [n], or an n-set partition, is a set of

disjoint, nonempty subsets of [n] called blocks whose union is [n].

Example 44. A set partition of [6] into three blocks is 1, 4, , 2, 3, 5, 6.The natural question posed by an enumerologist is how many such objects are there.

In particular, given non-negative integers n and k, how many set partitions are there of [n]

with k blocks? The answer is the Stirling number of the second kind Sn,k. By convention

we set S0,0 = 1 and Sn,k = 0 for all k > n > 0. The results of this chapter rest primarily on

the following recurrence relation.

Lemma 13. For all n ≥ 0 and 1 ≤ k ≤ n + 1,

Sn+1,k = Sn,k−1 + kSn,k. (5–1)

Proof. Let π be a set partition counted by Sn+1,k. If n + 1 is a block in π, we may

remove it, leaving a set partition of [n] with k − 1 blocks, an object counted by Sn,k−1.

On the other hand, if there are other elements in that block of π which contains n + 1,

we simply remove n + 1 from that block and still have k blocks of n elements, an object

counted by Sn,k. This is clearly an injection and thus shows that the left hand side above

is less than or equal to the right hand side. Now, let γ be a set partition counted by the

right hand side above. If γ has k−1 blocks, we append a block containing only n+1. Else,

we choose one of the k blocks in γ into which we insert n + 1. Again we have an injection,

and the equality holds.

74

Lemma 14. For all n ≥ 1 and 1 ≤ k ≤ n, we have the exact formula

Sn,k =1

k!

k∑i=0

(−1)i

(k

i

)(k − i)n. (5–2)

Furthermore, we have the asymptotic formulae

Sn,k =kn

k!−O((k − 1)n) (5–3)

and

Sn,k ∼ kn

k!. (5–4)

Proof. A set partition of an n-set is constructed by placing the elements into k disjoint

sets. Consider instead the case where there are k labeled boxes and n labeled balls which

we place in the boxes such that no box is empty. Call this an (n, k)-placement and let an,k

be the number of (n, k)-placements. Then, an,k = k! · Sn,k, as there are Sn,k ways to place

the balls into unlabeled boxes, and k! ways to label these boxes. So, it suffices to show that

the summation in (5–2) is an,k.

We apply the method of Inclusion-Exclusion. There are kn ways to distribute the

balls to the boxes, as we have k choices of what to do with each of n balls. However, we

may have overcounted distributions in which a box was left empty, so we subtract the(

k1

)(k − 1)n ways to choose a box to be empty and distribute the balls to the remaining

k − 1 boxes. Now, we have overcounted the distributions in which there are two empty

boxes. Indeed, for some 1 ≤ p < q ≤ k, consider the set of distributions which leave boxes

p and q empty. We counted such a distribution once in the first term, but subtracted

twice in the second term, once for leaving box p empty and once for leaving box q empty.

So, we need to add 1 for each such pair of boxes and each way to leave that pair empty

and distribute the balls to the other boxes,(

k2

)(k − 2)n. This process continues until the

number i of empty boxes reaches zero, and we have the desired sum.

75

For the asymptotics, we write (5–2) as

Sn,k =kn

k!− k(k − 1)n

k!+

k(k − 1)(k − 2)n

2k!− · · · .

As n → ∞, the dominant term is kn

k!. Furthermore, the dominant term of Sn,k − kn

k!is

k(k−1)n

k!, a constant multiple of (k − 1)n, demonstrating (5–3). Finally, as n →∞,

Sn,k

kn

k!

= 1− O((k − 1)n)kn

k!

→ 1.

Lemma 15. For all n ≥ 1 and 1 ≤ k ≤ n, we have the bounds

kn−k ≤ Sn,k ≤ kn

k!(5–5)

and

kn

k!−

dk/2e−1∑i=0

(k

2i + 1

)(k − 2i− 1)n < Sn,k <

kn

k!. (5–6)

Proof. The upper bound, Sn,k ≤ kn

k!, follows trivially from (5–2). By the recurrence (5–1),

for all n > k ≥ 1, we have

Sn,k ≥ kSn−1,k

≥ k2Sn−2,k

≥ · · ·

≥ kn−kSk,k

= kn−k,

as Sk,k = 1 for all k, completing (5–5). The lower bound of (5–6) is the difference of the

dominant term and all odd-indexed, hence negative, terms of (5–2).

76

5.2 Bell Polynomials

For all n ≥ 1, let Bn(x) be the generating polynomial for the Stirling numbers of the

second kind:

Bn(x) =n∑

k=1

Sn,kxk.

These are known as the Bell polynomials.

Example 45. The first few Bell polynomials are as follows.

B0(x) = 1,

B1(x) = x,

B2(x) = x + x2,

B3(x) = x + 3x2 + x3,

B4(x) = x + 7x2 + 6x3 + x4.

Lemma 16. The following recurrence relation on the Bell polynomials holds for all n ≥ 0.

Bn+1(x) = xBn(x) + xB ′n (x). (5–7)

Proof. Applying (5–1) to each term of Bn+1(x), we have

Bn+1(x) =∑

Sn+1,kxk

=∑

(Sn,k−1 + kSn,k)xk

=∑

Sn,k−1xk +

∑kSn,kx

k

= x∑

Sn,k−1xk−1 + x

∑kSn,kx

k−1

= xBn(x) + xB ′n (x).

We have the following well-known result on the roots of the Bell polynomials.

77

Lemma 17. For all n ≥ 1, Bn(x) has n distinct, real, non-positive roots, including zero.

Proof. (Wilf, [40])

The statement holds for B1(x). Assume by induction the statement holds for Bn(x).

Multiplying each term in (5–7) by ex,

exBn+1(x) = x(exBn(x))′.

By Rolle’s theorem, B′n(x) has n − 1 roots, one between each consecutive pair of roots

of Bn(x). Multiplication by x guarantees a root at zero. Since exBn(x) approaches zero

as x → −∞, its derivative has one more root to the left of the leftmost root of Bn(x),

accounting for all n + 1 roots of Bn+1(x).

Thus we can factor each Bell polynomial as

Bn(x) = x(x + λn,2) · · · (x + λn,n), (5–8)

where 0 = λn,1 < λn,2 < · · · < λn,n, and n ≥ 1. In this chapter we explore the roots

λn,k, finding bounds and limits for each fixed k. First, we have a theorem on the Stirling

numbers which builds on the results of the previous chapter but was held aside for the

foregoing definitions and discussion.

Theorem 46. The Stirling numbers of the second kind Sn,k are asymptotically uniform.

Proof. Harper [19] proved that the Stirling numbers of the second kind are asymptotically

normal. The above discussion shows that their generating polynomials have real,

non-positive roots only, so by Theorem 42 in the previous chapter, we achieve the desired

result.

Now, let us consider the roots more closely.

78

Example 47. We solve for the roots of the first few Bell polynomials.

B1(x) = x,

B2(x) = x(x + 1),

B3(x) = x(x +−3 +

√5

2)(x +

−3−√5

2).

For higher n, the algebra becomes unwieldy, as can be readily seen in the other Bell

polynomials given in Example 45.

As our work often deals with sums of products of reciprocals of these roots, for

notational convenience we will write δn,k = 1λn,k

for k = 2, 3, . . . , n, or simply δk = 1λk

if n is

clear from context.

Lemma 18. For all n ≥ 1 and 2 ≤ k ≤ n− 1, we have

Sn,k =∑

λn,i1 · · ·λn,in−k,

where the sum is taken over all (n−k)-tuples i1, i2, . . . , in−k with 2 ≤ i1 < · · · < in−k ≤ n.

Proof. Expanding the factorization in (5–8), an xk term is achieved by choosing x from k

of the terms and λn,i from the remaining n−k terms. The sum of all such products is Sn,k,

the coefficient of xk in Bn(x).

We note that, for all n ≥ 1, we have Sn,n = 1. Lemma 18 could be rewritten to reflect

this by allowing the sum over all (n− n)-tuples to be 1.

Lemma 19. For all n ≥ 2 and all nonempty subsets M ⊆ 2, . . . , n,∏i∈M

λn,i =∏

i∈MC

δn,i,

where MC = 2 ≤ i ≤ n| i 6∈ M is the complement of M .

Proof. We first note that the product of the n− 1 non-zero λn,k’s is Sn,1 = 1, so

n∏

k≥2

λn,k = 1 =n∏

k≥2

δn,k.

79

Thus, for any M ⊆ 2, . . . , n,∏

i∈MC

λn,i

∏i∈M

λn,i = 1, so

∏i∈M

λn,i =

( ∏

i∈MC

λn,i

)−1

∏i∈M

λn,i =∏

i∈MC

δn,i.

Lemma 20. For all n ≥ 1 and 2 ≤ k ≤ n− 1, we have

Sn,k =∑

δn,i1 · · · δn,ik−1, (5–9)

where the sum is taken over all (k−1)-tuples i1, i2, . . . , ik−1 with 2 ≤ i1 < · · · < ik−1 ≤ n.

Proof. Apply Lemma 19 to each (n− k)-tuple in Lemma 18.

For more on Stirling numbers of the second kind and Bell polynomials, particularly

their properties and myriad applications, the reader is referred to the classic by Riordan

[28] as well as the excellent texts by Roman [29] and Wilf [40].

5.3 Bounds on the Roots of the Bell Polynomials

Lemma 21. For all n ≥ 2,

2n−1

n≤ δn,2 < 2n−1.

Proof. By Lemmas 14 and 20 and the nonnegativity of the δn,i’s,

δn,2 ≤n∑

k=2

δn,k = Sn,2 < 2n−1.

As δn,2 is the largest term of the sum, it is at least as large as the mean, so δn,2 > 2n−1−1n−1

.

Finally, this last expression is at least 2n−1

nfor all n ≥ 2.

80

Lemma 22. For all n ≥ 3,

3n−1

2n· 4

9n2≤ δn,3 <

3n−1

2n· n.

Proof. We begin with the upper bound. By Lemmas 14 and 20,

∑2≤i<j≤n

δn,iδn,j = Sn,3 =1

23n−1 −O(2n) <

3n−1

2. (5–10)

In particular,

δn,2δn,3 <3n−1

2.

Applying the inequality δn,2 > 2n−1

nfrom Lemma 21,

δn,3 <3n−1

2δn,2

<n3n−1

2n.

For the lower bound, inequality (5–5) of Lemma 15 provides

Sn,3 > 3n−3.

The next inequality follows from Lemma 20, the fact that the largest of the(

n−12

)terms in

(5–10) is δn,2δn,3, and our upper bound δn,2 < 2n−1.

2n−1δn,3 >3n−3

(n2

) =2 · 3n−3

n(n− 1).

Finally, dividing by 2n−1,

δn,3 >3n−3

2n−1· 2

n(n− 1)

>3n−1

2n· 4

9n2.

81

Theorem 48. For each k ≥ 2, there exist polynomials fk(x) and Fk(x) of degrees 2k−2 and

2k−2 − 1, respectively, such that for all n ≥ k

(k

k − 1

)n1

fk(n)≤ δn,k <

(k

k − 1

)n

Fk(n).

Proof. We induct on k, generalizing the proof of Lemma 22. We proved the statement for

δn,2 and δn,3 in Lemmas 21 and 22, respectively. Let k ≥ 4 and assume by induction that

the statement holds for all δn,j with 2 ≤ j < k. By Lemmas 14 and 20, for all n ≥ k,

∑δn,i1 · · · δn,ik−1

= Sn,k <kn

k!, (5–11)

where the sum is taken over all (k−1)-tuples i1, i2, . . . , ik−1 with 2 ≤ i1 < · · · < ik−1 ≤ n.

By induction we have the following lower bound for each 2 ≤ j < k.

δn,j ≥(

j

j − 1

)n

· 1

fj(n).

which holds for all n ≥ k and where fj(x) is a polynomial of degree 2j−2 for each j

(independent of n). Define the degree 2j−1 − 1 polynomial F (x) = 1k!

f2(x)f3(x) · · · fk−1(x).

Then,

δn,2δn,3 · · · δn,k <kn

k!

δn,k <kn

k!δn,2 · · · δn,k−1

<kn

1

(k − 2)n

(k − 1)n

(k − 3)n

(k − 2)n· · · 2

n

1F (n)

=F (n)kn

(k − 1)n.

For the lower bound, inequality (5–5) of Lemma 15 provides

Sn,k > kn−k.

82

The largest of the(

n−1k−1

)terms in (5–11) is δn,2δn,3 · · · δn,k. By our induction

hypothesis, δn,j obeys the following bound for each 2 ≤ j < k.

δn,j <

(j

j − 1

)n

Fj(n)

for polynomials Fj(x). Altogether,

δn,2δn,3 · · · δn,k >kn−k

(n−1k−1

) ,

δn,k >kn−k

1

(k − 2)n

(k − 1)n

(k − 3)n

(k − 2)n· · · 2

n

1

1

fk(n).

where fk(x) =(

n−1k−1

)F2(x)F3(x) · · ·Fk−1(x).

5.4 Asymptotics of the Roots of the Bell Polynomials

The next two lemmas use our bounds on δn,2 and δn,3 to obtain the asymptotic

equivalences δn,2 ∼ 2n−1 and δn,3 ∼ 3n−1

2n .

Lemma 23. We have the following limit.

limn→∞

δn,2

2n−1= 1.

Proof. The inequality δn,2 < 2n−1 from Lemma 21 implies δn,2

2n−1 < 1 for all n ≥ 2, and it

suffices to show that the terms δn,2

2n−1 are bounded below by a sequence converging to 1. By

Lemmas 14 and 20 we know that Sn,2 = 2n−1 − 1 is the sum of δn,2, δn,3, and n − 3 other

terms, each of which is less than δn,3, so for all n > 3 we have the following inequalities

83

(applying the bound n(

32

)n> δn,3)

δn,2 + nδn,3 > 2n−1,

δn,2 + n2

(3

2

)n

> 2n−1,

δn,2 > 2n−1 − 2n2

(3

2

)n

,

δn,2

2n−1> 1− n2

(3

4

)n

→ 1.

The final limit, in essence 2n2(

34

)n → 0 as n → ∞, holds as 2n2 is a polynomial and(

34

)nis a geometric progression with positive ratio less than one.

Lemma 24. We have the following limit.

limn→∞

δn,32n

3n−1= 1

Proof. Lemma 20 expresses Sn,3 as the sum of all products δiδj. We can split this into two

sums, one sum containing those pairs which include δn,2 and one sum containing all other

pairs,

δn,2

∑i≥3

δn,i +∑

3≤i<j≤n

δn,iδn,j = Sn,3 =3n−1

2−O(2n). (5–12)

We proceed by showing that the second summation above is negligible with respect to

the dominant term 3n−1. Lemma 22 gives us the upper bound δn,3 < n3n−1

2n , and by our

ordering on the δ’s we have δn,j < δn,3 for all j > 3, implying δn,iδn,j <(

n3n−1

2n

)2

for each of

the(

n−22

)pairs i, j with 3 ≤ i < j ≤ n. Thus,

∑3≤i<j≤n

δn,iδn,j <

(n− 2

2

)(n3n−1

2n

)2

<

(9

4

)nn4

2, (5–13)

and∑

3≤i<j≤n

δn,iδn,j = O

((9

4

)n

n4

).

84

Combining (5–12) and (5–13), we have

δn,2

∑i≥3

δn,i = Sn,3 −O

((9

4

)n

n4

)

=3n−1

2−O

((9

4

)n

n4

).

Dividing by our asymptotic approximation of δn,2, δn,2 ∼ 2n−1,

∑i≥3

δn,i =3n−1

2n−O

((9

8

)n

n4

).

By Theorem 48,(

43

)nF4(n) > δn,4 > δn,5 > · · · for the fixed polynomial of degree 3 F4(x),

and (9

8

)n

n4 + (n− 3)

(4

3

)n

F4(n) = O

((4

3

)n

n4

),

so

δn,3 =3n−1

2n−O

((4

3

)n

n4

).

Finally, we divide by 3n−1

2n and take the limit as n →∞.

δn,32n

3n−1= 1−O

((8

9

)n

n4

).

limn→∞

δn,32n

3n−1= lim

n→∞1−O

((8

9

)n

n4

)

= 1.

The following theorem generalizes these results to all k, giving the asymptotic

approximation δn,k ∼ kn−1

(k−1)n for each fixed k.

Theorem 49. For each fixed k, we have the following limit.

limn→∞

δn,k(k − 1)n

kn−1= 1.

Proof. We generalize the proof of the previous lemma to all k ≥ 4. We induct on k, the

cases k = 2 and k = 3 given in the previous two lemmas. The induction hypothesis gives

85

us asymptotic bounds which improve upon those given in the previous section. Indeed, for

all ε > 0 there exists N > 0 such that n > N implies that for all 2 ≤ j < k, we have

δn,j(j − 1)n

jn−1∈ (1− ε, 1 + ε).

We can rewrite equation (5–9) of Lemma 20 as

δn,2δn,3 · · · δn,k−1

j≥k

δn,j +∑

δn,j1 · · · δn,jk−1= Sn,k =

kn

k!−O((k − 1)n), (5–14)

where the second sum is taken over all (k − 1)-tuples 2 ≤ j1 < · · · < jk−1 ≤ n containing

at least two elements which are each at least k. We first show that this summation is

negligible with respect to kn. Let 2 ≤ j1 < · · · < jk−1 ≤ n be such a (k − 1)-tuple. Then,

the product of the greatest k − 3 δn,j’s is bounded above by the product δn,2 · · · δn,k−2, and

the product of the remaining two δn,j’s is bounded above by δ2n,k.

δn,j1 · · · δn,jk−1< δn,2δn,3 · · · δn,k−2δn,kδn,k

< 2n 3n

2n· · · (k − 2)n

(k − 3)nF3(n) · · ·Fk−2(n)

kn

(k − 1)n

kn

(k − 1)nFk(n)Fk(n)

=

(k2(k − 2)

(k − 1)2

)n

G(n),

where G(x) = F3(x) · · ·Fk−2(x)F 2k (x). Set α = k2(k−2)

(k−1)2and note that k − 1 < α < k. The

summation was over a subset of all (k − 1)-tuples, so we overcount the terms by setting

H(x) = G(x)(

nk

)and we can therefore rewrite equation (5–14) as

δn,2δn,3 · · · δn,k−1

j≥k

δn,j =kn

k!−O(αnH(n)). (5–15)

Now, by induction we have

δn,2δn,3 · · · δn,k−1 ∼ 2n−1 3n−1

2n· · · (k − 1)n−1

(k − 2)n

=(k − 1)n−1

(k − 2)!.

86

While the above was an asymptotic approximation, we have the useful information

δn,2δn,3 · · · δn,k−1 = O ((k − 1)n) .

As α > k − 1, this implies

δn,2δn,3 · · · δn,k−1 = O(αn).

Dividing equation (5–15) by (k−1)n−1

(k−2)!,

j≥k

δn,j =kn−1

(k − 1)n−O(βnH(n)), (5–16)

where β = αk−1

= k2(k−2)(k−1)3

. Now it suffices to annihilate all terms on the left hand side

except δn,k. By Theorem 48,

(k + 1

k

)n

Fk+1(n) > δn,k+1 > · · · .

So,

δn,k =kn−1

(k − 1)n−

i≥k+1

δn,i −O(βnH(n))

=kn−1

(k − 1)n−O (βn(H(n) + nFk+1(n))) .

Taking the limit as n →∞, we achieve our desired result.

87

REFERENCES

[1] N. Alon and J. Spencer. The Probabilistic Method. Wiley-Interscience, New York, NY,2000.

[2] R. Arratia. On the Stanley-Wilf conjecture for the number of permutations avoiding agiven pattern. Electronic Journal of Combinatorics, 6(1):N1, 1999.

[3] E. A. Bender. Central and local limit theorems applied to asymptotic enumeration.Journal of Combinatorial Theory, Ser. A, 15:91–111, 1973.

[4] M. Bona. Permutations avoiding certain patterns; the case of length 4 andgeneralizations. Discrete Mathematics, 175:55 – 67, 1997.

[5] M. Bona. Combinatorics of Permutations. Chapman & Hall, Boca Raton, FL, 2004.

[6] M. Bona. The limit of a Stanley-Wilf sequence is not always rational and layeredpatterns beat monotone patterns. Journal of Combinatorial Theory, 110:223–235,2005.

[7] M. Bona. On a balanced property of derangements. Electronic Journal of Combina-torics, 13, 2006.

[8] M. Bona. A Walk Through Combinatorics. World Scientific, River Edge, NJ, 2006.

[9] M. Bona. On a balanced property of compositions. Online Journal of AnalyticCombinatorics, 2, 2007.

[10] M. Bousquet-Melou. Four classes of pattern-avoiding permutations under one roof:Generating trees with two labels. Electronic Journal of Combinatorics, 9(2):R19,2003.

[11] C. A. Charalambides. Combinatorial Methods in Discrete Distributions.Wiley-Interscience, Hoboken, NJ, 2005.

[12] F.R.K. Chung, R.L. Graham, V.E. Hoggatt Jr, and M. Kleiman. The number ofBaxter permutations. Journal of Combinatorial Theory (Series A), 24:382–394, 1978.

[13] K. L. Chung. A Course In Probability Theory. Elsevier, San Diego, CA, 2001.

[14] M. Coleman, M. Albert, I. Leader, and R. Flynn. Permutations containing manypatterns. Annals of Combinatorics, to appear.

[15] Micah Coleman. An answer to a question by Wilf on packing distinct patterns in apermutation. Electron. J. Combin., 11(1):Note 8, 4 pp. (electronic), 2004.

[16] H. Eriksson, K. Eriksson, S. Linusson, and J. Wastlund. Dense packing of patternsin a permutation, Proceedings of the 15th Conference on Formal Power Series andAlgebraic Combinatorics (Melbourne, Australia). http://www.i3s.unice.fr/fpsac/

FPSAC02/articles.html, 2002.

88

89

[17] M. Fekete. Uber die Verteilung der Wurzeln bei gewissen algebraischen gleichungenmit ganzzahligen koeffizienten. Mathematische Zeitschrift, 17:228 – 249, 1923.

[18] P. Halmos. Measure Theory. D. Van Nostrand Co., Berlin, 1956.

[19] L. H. Harper. Stirling behavior is asymptotically normal. Ann. Math. Statist.,38:410–414, 1967.

[20] P. A. Hasto. The packing density of other layered permutations. Electronic Journal ofCombinatorics, 9(2):R1, 2002.

[21] M. Klazar. The Furedi-Hajnal conjecture implies the Stanley-Wilf conjecture. InFormal Power Series and Algebraic Combinatorics, pages 250–255, Berlin, Germany,2000. Springer Verlag.

[22] M. Klazar. Personal communication, 2008.

[23] A. Marcus and G. Tardos. Excluded permutation matrices and the Stanley-Wilfconjecture. Journal of Combinatorial Theory Series A, 107:153–160, July 2004.

[24] D. Marinov and R. Radoicic. Counting 1324-avoiding permutations. ElectronicJournal of Combinatorics, 9(2):R13, 2002.

[25] Alison Miller. Asymptotic bounds for permutations containing many differentpatterns, 2006. preprint.

[26] A. M. Odlyzko. Handbook of Combinatorics, volume 2, chapter AsymptoticEnumeration Methods, pages 1063–1229. Elsevier, Cambridge, MA, 1995.

[27] A. Price. Packing Densities of Layered Patterns. PhD thesis, University ofPennsylvania, 1997.

[28] J. Riordan. An Introduction to Combinatorial Analysis. Wiley, New York, NY, 1980.

[29] S. Roman. The Umbral Calculus. Dover, Mineola, NY, 1984.

[30] A. Rucinski. Random Graphs, chapter 2, Proving Normality in Combinatorics, pages215–231. Wiley Interscience, Cambridge, UK, 1992.

[31] R. Simion and F. W. Schmidt. Restricted permutations. European Journal ofCombinatorics, 6:383 – 406, 1985.

[32] N.J.A. Sloane. The on-line encyclopedia of integer sequences. http://www.research.att.com/~njas/sequences/A088532, 2003.

[33] R. Stanley. Enumerative Combinatorics, Volume 1. Cambridge University Press,Cambridge, UK, 1997.

[34] R. Stanley. Enumerative Combinatorics, Volume 2. Cambridge University Press,Cambridge, UK, 1999.

90

[35] J. Taylor. An Introduction to Measure and Probability. Springer, New York, NY,1997.

[36] V. Vatter. Permutations avoiding two patterns of length three. Electronic Journal ofCombinatorics, 9(2):R6, 2003.

[37] V. Vatter. Small permutation classes. http://arxiv.org/abs/0712.4006, 2007.

[38] D. Warren. Optimizing the Packing Behavior of Layered Permutation Patterns. PhDthesis, University of Florida, 2005.

[39] J. West. Generating trees and forbidden subsequences. Discrete Mathematics, 157(1 -3):363 –374, 1996.

[40] H. Wilf. Generatingfunctionology. A K Peters, Wellesley, MA, 2006.

[41] D. Zeilberger. A loving rendition of the Marcus-Tardos amazing proof of theFuredi-Hajnal conjecture. http://www.math.rutgers.edu/~zeilberg/mamarim/

mamarimPDF/martar.pdf, 2003.

BIOGRAPHICAL SKETCH

A native Florida Cracker, I graduated from Seabreeze High School in 1994 and

entered college as a music major. Realizing my lack of the skill and devotion necessary

for a professional musician, I enlisted in the United States Navy, stationed for four years

in Yokosuka, Japan, where I met my wife Hiroko Shinohara. After my active duty period,

I completed an Associate’s degree at Daytona Beach Community College and joined

my brother and his wife here in Gainesville, entering the world of higher math for the

first time at the University of Florida. As a junior I first met my advisor and fell in love

with Combinatorics. I graduated summa cum laude in 2004 and accepted a fellowship at

UF. I have thoroughly enjoyed my graduate career, even my eight-month sabbatical to

Baghdad, Iraq last year, fully funded by the Naval Reserve. I look forward to a long career

of mathematics research and hope to make some small contribution to the Conversation.

91