+ All Categories
Home > Documents > COMBINATORIAL THEORY OF Q,T-SCHRODER¨ …COMBINATORIAL THEORY OF Q,T-SCHRODER¨ POLYNOMIALS,...

COMBINATORIAL THEORY OF Q,T-SCHRODER¨ …COMBINATORIAL THEORY OF Q,T-SCHRODER¨ POLYNOMIALS,...

Date post: 30-Jun-2020
Category:
Upload: others
View: 3 times
Download: 0 times
Share this document with a friend
130
COMBINATORIAL THEORY OF Q,T-SCHR ¨ ODER POLYNOMIALS, PARKING FUNCTIONS AND TREES Chunwei Song A Dissertation in Mathematics Presented to the Faculties of the University of Pennsylvania in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy 2004 Supervisor of Dissertation Graduate Group Chairperson
Transcript

COMBINATORIAL THEORY OF Q,T-SCHRODER

POLYNOMIALS, PARKING FUNCTIONS AND TREES

Chunwei Song

A Dissertation in Mathematics

Presented to the Faculties of the University of Pennsylvania

in Partial Fulfillment of the Requirements for

the Degree of Doctor of Philosophy

2004

Supervisor of Dissertation

Graduate Group Chairperson

COPYRIGHT

Chunwei Song

2004

To Athena (Lingyun),

Jie,

and my parents,

with much love and gratitude.

iii

Acknowledgements

I would like to express my deep gratitude to my advisor, James Haglund, for his encourage-

ment and caring guidance. I tremendously benefited from his sharp insight of the subjects,

in depth knowledge of the field and extraordinary skills of attacking problems.

There have been many people who ever played an important role in my mathematical

development. In special, I must thank Herb Wilf, who attracted me into the field of combi-

natorics through a very first lesson of “generatingfunctionology” [Wil94]; Andre Scedrov,

who instructed me mathematical logic in addition to cryptology; Jennifer Morse, who in

a graceful manner displayed to me Tableau theory and symmetric functions; Amy Myers,

who shared me with the theory of order; Paul Seymour, who brought me into his realm of

graph theory; and Felix Lazebnik, who for a whole year tutored me extremal combinatorics

and probabilistic/algebraic methods. I am also grateful to the following people who lent

their expertise to me during the progress of this dissertation: Doron Zeilberger, Ira Gessel,

Robert Sulanke, Michael Steele and Nick Loehr. They, and the other teachers throughout

my maturation, have shaped my view on MATHEMATICS.

I appreciate the support and training from the University of Pennsylvania, the School

of Arts and Sciences and especially the Department of Mathematics. Top-notch taste of an

education fosters maestros; foremost and broad intellectual pursuit in a University’s ethos is

vital; extraordinary curriculum will have far-reaching impact. In order for a mathematician

to qualify being anintellectual(in Chineseshih), i.e. a member of the vast and complex

iv

array of professionals entrusted with the preservation and perpetuation of certain specific

knowledge or ideas and privileged to be the most indoctrinated members of society, one has

to possess an ultimate concern toward one’s nation, society, and the entire humanity. This

concern is for everything pertinent to the public benefits and must transcend self as well as

coterie interests, which in some sense coincides with the religious spirit of responsibility.

I am very thankful to the math departmental staff, in particular Janet Burns and Monica

Pallanti, for their assistance in many aspects during the past years when I was a graduate

student.

It has been a great pleasure to discuss mathematics at Penn with many of our excellent

fellow students, among whom forgive me to mention only my peers Fred Butler, Irina

Gheorghiciuc and Aaron Jaggard.

My mathematical career started at my age of five due to the enlightenment of my

mother, Ms. Guiying Li, an extremely talented woman who had no chance for the best

education. Athena (Lingyun) Song, a little angel born at the beginning of this dissertation,

has presented much inspiration and joy. Last but not least, I am indebted to the infinite love

and emotional support from my wife, Jie, who has shared every piece of my excitements

and frustrations and provided to me anenriched beautiful life.

Philadelphia, Pennsylvania

April 19, 2004

v

ABSTRACT

COMBINATORIAL THEORY OF Q,T-SCHR ODER

POLYNOMIALS, PARKING FUNCTIONS AND TREES

Chunwei Song

James Haglund

We study various aspects of lattice path combinatorics. A new object,

which has Dyck paths as its subset and is named Permutation paths, is con-

sidered and relative theories are developed. We prove a class of tree enu-

meration theorems and connect them to parking functions. The limit case

of (q, t)-Schroder Theorem is investigated. In the end, we derive a formula

for the number ofm-Schroder paths and study itsq and(q, t)-analogues.

vi

Contents

Acknowledgements iv

Contents vii

List of Figures ix

1 Introduction 1

1.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.2 A Survey of the Literature . . . . . . . . . . . . . . . . . . . . . . . . . . 2

1.3 Summary of New Results . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

2 Dyck Paths and Permutation Paths 17

2.1 On the Symmetry of the (q, t)-Catalan Polynomial . . . . . . . . . . . . . . 17

2.1.1 A Bijection BetweenDn and a Special Set of Permutations . . . . . 18

2.1.2 The Foata-Schutzenberger Involution . . . . . . . . . . . . . . . . 27

2.2 The Theory of Permutation Paths . . . . . . . . . . . . . . . . . . . . . . . 29

2.2.1 Then! Permutation Paths and the Bijectionf betweenSn andPn . 29

2.2.2 Restrictingf to Some Pattern Forbidding Permutations . . . . . . . 33

2.2.3 The Signed Permutation PathsBn . . . . . . . . . . . . . . . . . . 40

3 Tree Enumeration Theorems and the (q, t)-Parking Function Polynomial 43

vii

3.1 Haglund and Loehr’s (q, t)-Parking Function Polynomial . . . . . . . . . . 43

3.2 Counting Special Families of Labelled Trees . . . . . . . . . . . . . . . . . 45

3.3 Major Sequences, Tree Inversion Enumerator and the Tutte Polynomial . . 64

4 The Limit Case of the (q, t)-Schroder Theorem 70

4.1 Background and Basic Properties . . . . . . . . . . . . . . . . . . . . . . . 70

4.2 Investigation of Zeros . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

4.3 Whenq = t . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83

4.4 Agreement oftk Coefficients . . . . . . . . . . . . . . . . . . . . . . . . . 91

5 Higher Dimensional Schroder Theory 106

5.1 m-Schroder Paths andm-Schroder Number . . . . . . . . . . . . . . . . . 106

5.2 q-m-Schroder Polynomials . . . . . . . . . . . . . . . . . . . . . . . . . . 109

5.3 (q, t)-m-Schroder Statistics and the Shuffle Conjecture . . . . . . . . . . . 112

Bibliography 115

viii

List of Figures

1.1 A Dyck pathΠ ∈ D6 with area(Π)=3. . . . . . . . . . . . . . . . . . . . . 4

1.2 A Dyck pathΠ and its bounce pathB. . . . . . . . . . . . . . . . . . . . . 5

1.3 A Schroder pathΠ ∈ S8,4 with area(Π) = 5 andmaj(Π) = 30. . . . . . . 8

1.4 A 2-Schroder path of order6 and with2 diagonal steps. . . . . . . . . . . . 15

2.1 A balanced Dyck path of 4 blocks . . . . . . . . . . . . . . . . . . . . . . 19

2.2 Add strips toΠ(1) to obtainΠ(2) . . . . . . . . . . . . . . . . . . . . . . . 23

2.3 Correspondence betweenP3 andS3 underf . . . . . . . . . . . . . . . . . 31

2.4 A Permutation pathΠ ∈ P8 with 4 caves and a 2-triangle . . . . . . . . . . 38

2.5 σ = 57836241 ∈ S8(4123) butΠ is not 2-triangle-forbidding. . . . . . . . . 39

2.6 A Signed Permutation path. . . . . . . . . . . . . . . . . . . . . . . . . . . 40

3.1 A parking functionP ∈ P8 with area(D(P )) = 6. . . . . . . . . . . . . . 44

3.2 The number of “Least-Child-Being-Monk” trees on{0, 1, 2, 3} is 22 = 4. . 47

3.3 A primary parking functionP ∈ P∗8 , with area(D(P )) = 12. . . . . . . . . 49

3.4 The3 members ofT3,1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

4.1 A Dyck path (on the left) and its inverse (on the right). . . . . . . . . . . . 89

5.1 A 2-Dyck path inD26. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107

5.2 A 2-Schroder path inS26,4. . . . . . . . . . . . . . . . . . . . . . . . . . . 108

ix

5.3 Anm-parking function whosem-reading word is 231. . . . . . . . . . . . 112

5.4 The natural expansion of anm-parking function. . . . . . . . . . . . . . . 113

x

Chapter 1

Introduction

1.1 Overview

The primary focus of this dissertation is on various properties of lattice path enumeration

and their (q, t)-analogues, a rapidly developing topic in the frontier of combinatorics, which

as a domain of mathematics is itself going through a profound revolution.

More specifically, we study combinatorial statistics on lattice paths, multi-variable

analogs of Catalan and Schroder numbers, and related more generalized issues. In intro-

ducing our work, the current chapter provides an introduction of the existing literature as

well as a summary of our new theorems, while the following chapters describe our results

in more details.

1

1.2 A Survey of the Literature

There are two noteworthy sequences of numbers that are of central interests to combinato-

rial mathematicians. Thenth Catalan numberis

Cn =1

n + 1

(2n

n

).

[Sta99, Ex.6.19, pages 219–229] gives 66 combinatorial interpretations of these numbers,

and updated additions are provided in [Sta].

Thenth Schroder numberis

Sn =n∑

d=0

1

n− d + 1

(2n− d

d, n− d, n− d

).

[Sta99, Ex.6.39, pages 239–240] provides 19 combinatorial interpretations of these num-

bers.

Among the interpretations, we are primarily interested in those associated to the lattice

paths.

Definition 1.2.1. A Dyck pathof ordern is a lattice path from(0, 0) to (n, n) that never

goes below the main diagonal{(i, i), 0 ≤ i ≤ n}, with steps(0, 1) (or NORTH, for brevity

N) and(1, 0) (or EAST, for brevityE). LetDn denote the set of all Dyck paths of ordern.

Definition 1.2.2. A Schroder pathof ordern is a lattice path from(0, 0) to (n, n) that never

goes below the main diagonal{(i, i), 0 ≤ i ≤ n}, with steps(0, 1) (or NORTH, for brevity

N), (1, 0) (or EAST, for brevityE) and (1,1) (or Diagonal, for brevityD). LetSn denote the

set of all Schroder paths of ordern.

The Catalan numberCn = 1, 2, 5, 14, 42, ..., counts the number of Dyck paths of order

n, while the Schroder numberSn = 2, 6, 22, 90, 394, ..., counts the number of Schroder

2

paths of ordern. In this dissertation, we are often more concerned with theSchroder paths

with d diagonal steps.

Definition 1.2.3. A Schroder pathof ordern and withd diagonal steps is a lattice path

from (0, 0) to (n, n) that never goes below the main diagonal{(i, i), 0 ≤ i ≤ n}, with

(0, 1) (or NORTH, for brevityN), (1, 0) (or EAST, for brevityE) and exactlyd (1,1) (or

Diagonal, for brevityD) steps. LetSn,d denote the set of all Schroder paths of ordern and

with d diagonal steps.

The number of Schroder paths of ordern and withd diagonal steps is counted by

Sn,d =

(2n− d

d

)Cn

=1

n− d + 1

(2n− d

d, n− d, n− d

).

ClearlySn =∑n

d=0 Sn,d andCn = Sn,0.

By a statisticon a given setS, we mean a combinatorial rule that associates a nonneg-

ative integer to each element ofS. Efforts and progresses have been made by considering

1 variable ([CR64], [FH85], [BSS93], etc) and 2 variable ([GH96], [HL], [EHKK03], etc)

generalizations of these numbers, through studying various invented statistics associated

with these lattice paths. The applications of these works expand from almost every sub-

field of discrete mathematics to other areas such as representation theory and algebraic

geometry.

Two important statistics onDn arearea andbounce.

Given Π ∈ Dn, area(Π) is defined to be the number of complete squares between

Π and the main diagonal liney = x. More specifically, letai(Π) be the number of

complete squares in theith row, from top to bottom, that are belowΠ and above the

main diagonal. The numberai(Π) is also called thelength of the ith row of Π, and

3

(a1(Π), a2(Π), . . . , anΠ) is thearea vectorof Π. Finally, area(Π) =∑n

i=1 ai(Π). An

example of a Dyck path of order 6 with area vector(1, 0, 0, 1, 1, 0) is illustrated in Figure

1.1.

0

1

0

1

1

0

Figure 1.1: A Dyck pathΠ ∈ D6 with area(Π)=3.

Carlitz and Riordan [CR64] defined the following naturalq-analogue ofCn,

Cn(q) =∑

Π∈Dn

qarea(Π),

and showed that

Theorem 1.2.1.

Cn(q) = qk−1

n∑

k=1

Ck−1(q)Cn−k(q), n ≥ 1.

The statisticbounce was introduced by Haglund in [Hag03]. Here we adopt the de-

scription of [HL] to define it: start by placing a ball at the upper corner(n, n) of a Dyck

pathΠ, then push the ball straight left. Once the ball intersects a vertical step of the path, it

“ricochets” straight down until it intersects the diagonal, after which the process is iterated;

the ball goes left until it hits another vertical step of the path, then follows down to the

diagonal, etc. On the way from(n, n) to (0, 0) the ball will strike the diagonal at various

points(ij, ij). We definebounce(Π) to be the sum of theseij. For convenience, we also let

the Dyck path so obtained in this process be thebounce pathof Π and denote it byb(Π).

4

In addition, we sayΠ is balancedif and only if Π = b(Π). In Figure 1.2, a Dyck path

Π is represented by the solid line and its bounce pathb(Π) = B is the dashing line. As

illustrated,bounce(Π) = 2 + 6 = 8.

6

2

Figure 1.2: A Dyck pathΠ and its bounce pathB.

Throughout this dissertation we use the standard notation

[n] := (1− qn)/(1− q), [n]! := [1][2] · · · [n],

[n

k

]:=

[n]!

[k]![n− k]!

for the q-analogue of the integern, the q-factorial, and theq-binomial coefficient and

(a)n := (1 − a)(1 − qa) · · · (1 − qn−1a) for the q-rising factorial. Sometimes it is nec-

essary to write the baseq explicitly as in[n]q, [n]!q,[nk

]q

and(a; q)n, but we often omitq if

it is clear from the context. Occasionally, wheni+ j +k = n, we also use[

ni,j,k

]:= [n]!

[i]![j]![k]!

to represent theq-trinomial coefficient.

We frequently make use of the following “q-binomial theorem” as a tool to prove iden-

tities.

Theorem 1.2.2.[And98, page 36]The “q-binomial theorem”. Forn ∈ N,

n∑

k=0

q(k2)

[n

k

]zk = (−z; q)n,

5

and∞∑

k=0

[n + k − 1

k

]zk =

1

(z; q)n

.

In [GH96], Garsia and Haiman introduced a complicated rational functionCn(q, t)

which they proved has the following properties:

Cn(q, 1) =∑

Π∈Dn

qarea(Π) = Cn(q)

q(n2)Cn(q, 1/q) =

1

[n + 1]

[2n

n

].

In order to interpretCn(q, t), Haglund [Hag03] introduced the distribution function

Fn(q, t) =∑

Π∈Dn

qarea(Π)tbounce(Π)

and conjectured thatFn(q, t) = Cn(q, t). Garsia and Haglund ( [GH02], [GH01]) proved

this by using symmetric function methods, and as a byproduct also the conjecture in [GH96]

thatCn(q, t) is a polynomial with positive integer coefficients. Therefore,Cn(q, t) is now

called the (q, t)-Catalan polynomial.

There is a pair of basic statistics on the symmetric groupSn, inv andmaj. In general,

for any integer word or multiset permutationw = w1w2 · · ·wn, inv andmaj are defined as

inv(w) =∑i<j

wi>wj

1

maj(w) =∑

iwi>wi+1

i.

For later use, we also define thedescent setof a wordw

Des(w) := {i : wi > wi+1},

6

and the number of descents ofw

des(w) := |Des(w)|.

The following result due to MacMahon [Mac60] is now classical.

Theorem 1.2.3.For any fixed integers and any vectorα ∈ Ns, if Mα denotes the set of all

permutations of the multiset{0α01α1 · · · sαs}, then

∑w∈Mα

qinv(w) =

[n

α1, · · · , αs

]=

∑w∈Mα

qmaj(w).

Accordingly we say thatinv andmaj aremultiset Mahonian statistics. If we let s = n,

α0 = 0, α1 = · · · = αn = 1 in the above theorem, thenMα specializes to the symmetric

groupSn,[

nα1,··· ,αs

]= n!, and therefore we say that the two statisticsinv andmaj on Sn

are bothMahonian statistics.

Given a Dyck pathΠ, if we encode eachN step by a 0, and eachE step by a 1, then

from (0, 0) to (n, n) we obtain a wordw(Π) of n 0’s andn 1’s. Thus, the subset ofMn,n

each element of which has at least as many 0’s as 1’s in any initial segment is in bijection

withDn. We call this special subset of 01 words theCatalan wordsof ordern and denote it

by CWn. Hence we may associate with eachΠ the statistics ofinv andmaj by inv(Π) =

inv(w(Π)) andmaj(Π) = maj(w(Π)). It is easy to see that(

n2

) − inv(Π) = area(Π).

The following classical result of MacMahon [Mac60, page 214] has a simple combinatorial

proof in [FH85].

Theorem 1.2.4.∑

Π∈Dn

qmaj(Π) =1

[n + 1]

[2n

n

].

Much of the theory about Dyck paths can be generalized to Schroder paths. In general

7

for a lattice pathΠ that never goes below the diagonal linex = y, definelower triangleto

be a triangle with vertices(i, j), (i + 1, j) and(i + 1, j + 1), and let thearea of Π, denoted

by area(Π), be the number of lower triangles betweenΠ and the main diagonal. This new

definition ofarea agrees with the old one for Dyck paths, and is well defined for Schroder

paths. Similarly, if we mapSn,d to the words ofn− d 0’s, d 1’s andn− d 2’s by replacing

eachN step by a 0, eachD step by a 1 and eachE step by a 2 in a Schroder pathΠ, then

we have themaj statistic for Schroder paths. Bonin, et. al. showed that [BSS93]

Theorem 1.2.5.∑

Π∈Sn,d

qmaj(Π) =1

[n− d + 1]

[2n− d

n− d, n− d, d

].

In Figure 1.3 below, the Schroder pathΠ ∈ S8,4 is encoded by 001221010221, which

implies thatmaj(Π) = 5 + 6 + 8 + 11 = 30, and has area vector (0,1,1,0,0,2,1,0), which

saysarea(Π) = 1 + 1 + 2 + 1 = 5. The length of each row, as computed from the number

of lower triangles, is shown on the right.

0

1

1

0

0

2

1

0

Figure 1.3: A Schroder pathΠ ∈ S8,4 with area(Π) = 5 andmaj(Π) = 30.

Egge, et. al [EHKK03] generalizedbounce to Schroder paths through a decomposition

8

procedure and defined the (q, t)-Schroder polynomial

Sn,d(q, t) =∑

Π∈Sn,d

qarea(Π)tbounce(Π).

They generalized Garsia and Haiman’s result to the following

q(n2)−(d

2)Sn,d(q,1

q) =

1

[n− d + 1]

[2n− d

n− d, n− d, d

],

They also conjectured that the (q, t)-Schroder polynomial is symmetric and made a stronger

conjectural interpretation ofSn,d(q, t) involving a linear operator∇ defined on the modified

Macdonald basis (for details see [EHKK03], [Hag04] or [HL]).

Conjecture 1.2.1.For all integersn, d with d ≤ n,

Sn,d(q, t) =< ∇en, en−dhd > .

This was recently proved in [Hag04] and thus became the (q, t)-Schroder Theorem.

1.3 Summary of New Results

In this section, we list the main theorems in the chapters that follow.

First, in Chapter 2 we obtain some partial results about the symmetry of the (q, t)-

Catalan polynomial and develop the theory of Permutation paths, which is a kind of gener-

alized lattice path that contains Dyck paths as a subset.

Theorem 1.3.1.The (q, t)-Catalan polynomial,Cn(q, t), is equal to the following distribu-

9

tion function defined onTn, whereTn is a subset of the symmetric groupSn.

Cn(q, t) =∑σ∈Tn

qinv(σ)t(n2)−maj(σ).

Definition 1.3.1. A Permutation path of ordern is a lattice path from(0, 0) to (n, n), which

never goes below the main diagonal(i, i), 0 ≤ i ≤ n, or above the liney = n, and consists

of NORTH(0, 1), EAST(1, 0) and SOUTH(0,−1) steps but never repeats (i.e. no NORTH

step followed or preceded by a SOUTH step). LetPn denote the collection of Permutation

paths of ordern.

Theorem 1.3.2.

|Pn| = n! .

Furthermore, there exists a weight-preserving bijectionf betweenSn andPn that maps

the inversion statistic to the area statistic. Namely, for anyσ ∈ Sn, we have

inv(σ) = area(f(σ)).

Next we consider the restriction off to Sn(312), the312-avoidingpermutations, and call

it f ∗. We show thatf ∗ is a bijective map betweenSn(312) and Dyck pathsDn, a subset of

the image set Permutation paths.

Theorem 1.3.3.f ∗ is a weight-preserving bijection betweenSn(312) and Dyck pathsDn

that maps the inversion statistic to the area statistic, and therefore

σ∈Sn(312)

qinv(σ) =∑

Π∈Dn

qarea(Π).

Theorem 1.3.3 can be generalized to thek12 . . . (k−1)-avoidingpermutations in a less

perfect way. For more details see Chapter 2. In the last section of Chapter 2 we introduce

10

Signed Permutation paths, which may be viewed as a generalization of both Permutation

paths and Schroder paths. Some parallel results on Signed Permutation paths are also

included.

In Chapter 3, we prove some graph theory enumeration results while investigating the

parking function polynomialRn(q, t) as introduced in [HL]. We are able to show that

Rn(q, 1) is equivalent to a group of other combinatorial statistics.

Theorem 1.3.4.(“Least-Child-Being-Monk”)DefineTn+1,0 to be the set of labelled trees

on {0, 1, 2, ..., n + 1}, such that the least labelled child of 0 has no children (we say such

trees have the Least-Child-Being-Monk property). Then the cardinality ofTn+1,0, which we

denote bytn+1,0, is equal tonn.

Corollary 1.3.5. Whenn goes to infinity, the probability for a labelled tree to be “Least-

Child-Being-Monk” ise−2.

Theorem 1.3.6.DefineTn+1,p to be the set of labelled trees on{0, 1, 2, ..., n+1}, such that

the total number of descendants of the least labelled child of 0 isp. Then, the cardinality

of Tn+1,p, denoted bytn+1,p, is equal to

(n− p)n−p(p + 1)p−1

(n + 1

p

).

Corollary 1.3.7. Whenn goes to infinity, the probability for a labelled tree on{0, 1, 2, · · · , n}to have the property that the least labelled child of 0 has exactlyp descendants is

(p + 1)p−1

p !e−2−p.

Theorem 1.3.8.(Hereditary-Least-Single Trees Recurrence) A rooted labelled tree is Hereditary-

Least-Single if it has the property that every least child has no children. Let the number

of Hereditary-Least-Single trees (rooted at the least labelled vertex) withn vertices behn.

11

Thenhn satisfies the following recurrence:

hn =(n− 1)hn−1 − 2∑

1≤i≤n−2

hn−ihi+1

(n− 2

i− 1, n− i− 1

)

+∑

1≤i≤n−2

∑1≤j≤n−i−1

ihihjhn+1−i−j

(n− 2

i− 1, j − 1, n− i− j

).

The following list contains{hn}, for n from 1 to 10, which is computed by Maple using

our recurrence: 1, 1, 1, 4, 15, 96, 665, 6028, 60907, 725560 ...

Theorem 1.3.9. Consider the exponential generating functionH(x) =∑

n≥0hn+1

n!xn.

ThenH(x) satisfies the simple functional equation

H2(x)−H(x) + 1 = exH(x).

Let

Rn(q, 1) =∑

P∈Pn

qarea(P ),

wherePn is the set of parking functions [HL], and

Mn(q) =∑

s∈Mn

qarea(s),

whereMn is the set of major sequences [Kre80].

Theorem 1.3.10.

Rn(q, 1) = Mn(q).

Corollary 1.3.11. The 5 combinatorial statistics (see[Bei82], [Bj o92]and [Ste02]) are all

equal, i.e.

Rn(q, 1) = Rn(1, q) = Mn(q) = TKn+1(1, q) = In(q),

12

and they all satisfy the following same recurrence:

Stat1(q) = 1,

Statn(q) =n∑

i=1

(n− 1

i− 1

)[i] Stati−1(q) Statn−i(q).

In Chapter 4, we attack a combinatorial proof of a interesting identity derived from the

limit case of the (q, t)-Schroder theorem. That is,

Theorem 1.3.12.For n ∈ N,

n∑

k=1

∑a1+···+ak=n

ai>0

q∑k

i=1 (ai2 ) t

∑k−1i=1 (k−i)ai

1

(tk; q)a1(q; q)ak

×k−1∏i=1

[ai + ai+1 − 1

ai

]1

(tk−i; q)ai+ai+1

× (q; q)n(t; t)n

= [zn]∏i,j≥0

(1 + qitjz)× (q; q)n(t; t)n

=∑σ∈Sn

qmaj(σ) t(n2)−maj(σ−1),

Above we use[zn]f(z) to denote the coefficient ofzn in f(z), a series in powers ofz.

Sometimes we also use[zn]{f(z)}, especially whenf(z) is a long formula. We analyze

several special cases, make parallels of some results by Carlitz [Car56] and also obtain

some refined results and conjectures relating the (q, t)-Schroder polynomial statistics to

the permutations whose longest increasing subsequence is of a fixed size. One of our

byproducts is Theorem 1.3.13.

Definition 1.3.2. The inverseof a Catalan wordw ∈ CWn is defined to be

w−1 = r(w),

13

wherer denotes the reverse operation and− denotes the complement operation that ex-

changes 0 and 1. We sayw is aninvolution if and only if w = w−1.

Example1.3.1. Whenn=3,

(000111)−1 = 000111,

(001011)−1 = 001011,

(001101)−1 = 010011,

(010011)−1 = 001101,

(010101)−1 = 010101.

So the involution set consists of 000111, 001011 and 010101.

It is easy to see thatw−1 ∈ CWn if and only if w ∈ CWn, so the inverse operation is

closed onCWn. Geometrically, givenw, we may obtainw−1 by finding the Dyck pathΠ

thatw corresponds to under the natural map, reflectingΠ over the NW-SE main diagonal

to obtain a new Dyck pathΠ−1, and then taking the Catalan word that corresponds toΠ−1.

Theorem 1.3.13.

∑w∈CWn:

w is an involution

qmaj(w)−ndes(w) =∑

σ∈Sn(123):σ is an involution

qmaj(σ)−maj(σ−1).

In Chapter 5 we turn to higher dimensional Schroder theory. That is, we study general-

ized Schroder paths inside a rectangle of lengthmn and widthn. We derive a formula for

the number ofm-Schroder paths and study itsq and(q, t)-analogues.

Definition 1.3.3. An m-Dyck pathof ordern is a lattice path from(0, 0) to (mn, n) which

never goes below the main diagonal{(mi, i) : 0 ≤ i ≤ n}, with steps(0, 1) (or NORTH,

14

Figure 1.4: A2-Schroder path of order6 and with2 diagonal steps.

for brevity N) and(1, 0) (or EAST, for brevityE). Let Dmn denote the set of allm-Dyck

paths of ordern.

Definition 1.3.4. An m-Schroder pathof ordern and withd diagonal steps is a lattice path

from (0, 0) to (mn, n), which never goes below the main diagonal{(mi, i) : 0 ≤ i ≤ n},with (0, 1) (or NORTH, for brevityN ), (1, 0) (or EAST, for brevityE) and exactlyd (1,1)

(or Diagonal, for brevityD) steps. LetSmn,d denote the set of allm-Schroder paths of order

n and withd diagonal steps.

Figure 1.3 illustrates a2-Schroder pathΠ ∈ S26,4.

Theorem 1.3.14.The number ofm-Schroder paths of ordern and withd diagonal steps,

denoted bySmn,d, is equal to

1

mn− d + 1

(mn + n− d

mn− d, n− d, d

).

Remark1.3.1. Whenm=1, the theorem above counts the ordinary Schroder paths. When

d = 0, them-Dyck paths are counted. Actually the later result, i.e.|Dmn | = 1

mn+1

(mn+n

n

)is

quite new [GH96] [HPW99], and not a single niceq-version seems to exist.

The following theorem generalizes a result of [BSS93].

15

Definition 1.3.5. Define them-Narayanapolynomialdmn (q) overm-Schroder paths of or-

dern to be

dmn (q) =

∑Π∈Sm

n

qdiag(Π),

where diag(Π) is the number ofD steps in them-Schroder pathΠ.

Theorem 1.3.15.dmn (q) hasq = −1 as a root.

In [FH85], there is a refinedq-identity,

k≥1

∑w∈CWn,k

qmajw =∑

k≥1

1

[n]

[n

k

][n

k − 1

]=

1

[n + 1]

[2n

n

],

whereCWn,k is the set of Catalan words consisting ofn 0’s,n 1’s, withk ascents (i.e.k−1

descents). For the generalized version, Cigler proved that there are exactly

1

n

(n

k

)(mn

k − 1

)

m-Dyck paths withk peaks (consecutive NE pairs) [Cig87]. In order to generalize the

results of [FH85], we prove a generalizedq-identity.

Theorem 1.3.16.

k≥d

[k

d

]1

[n]

[n

k

][mn

k − 1

]q(k−d)(k−1) =

1

[mn− d + 1]

[mn + n− d

mn− d, n− d, d

].

In the last section of Chapter 5, we mention a conjecture of Haglund, Haiman, Loehr,

Remmel and Ulyanov which defines the (q, t)-m-Schroder polynomial and relates it to the

∇ operator.

16

Chapter 2

Dyck Paths and Permutation Paths

2.1 On the Symmetry of the (q, t)-Catalan Polynomial

The (q, t)-Catalan polynomialCn(q, t), introduced in [GH96] as a rational function, is sym-

metric inq andt from its definition. However, the original definition is very complicated

and it is only because of the fact thatFn(q, t) = Cn(q, t), which is proved in [GH02]

[GH01], do we know thatCn(q, t) is a polynomial and has positive coefficients. Here

Fn(q, t) =∑

Π∈Dn

qarea(Π)tbounce(Π),

wherearea andbounce are statistics on Dyck pathsDn as introduced in Chapter 1. There is

no direct proof thatFn(q, t) is symmetric, i.e.,Fn(q, t) = Fn(t, q). Therefore it is desirable

to prove this combinatorially.

In this section we construct a bijectiong between Dyck pathsDn and a special sub-

group ofSn, which we callTn, interchangingarea andinv, andbounce and(

n2

) − maj

simultaneously. Thereby we hope to prove the symmetry of the (q, t)-Catalan number com-

binatorially by working on the new distribution function of the statisticsinv and(

n2

)−maj

17

onTn.

2.1.1 A Bijection BetweenDn and a Special Set of Permutations

Given a Dyck pathΠ ∈ Dn, we construct an injectiong, which mapsΠ to a permutation

σ ∈ Sn, with the properties that

area(Π) = inv(σ),

bounce(Π) =

(n

2

)−maj(σ).

We define this map by a procedure involving two steps.

Step 1: whenΠ ∈ Dn is a balanced path.

First consider the case thatΠ is a balanced path. That is,Π = b(Π). SupposeΠ is

made up ofk blocks, i.e.Π hask right (from NORTH to EAST) turns and hits the diagonal

exactlyk + 1 times including at(0, 0) and at(n, n). To better illustrate, we consider the

casek = 4, as it will be easy to extend this to generaln. As illustrated by Figure 2.1,

let the sizes of the 4 blocks bea, b, c andd, respectively, from bottom to top. Notice that

n = a + b + c + d.

The image permutationσ = g(Π) is defined as follows.

σ =a(a− 1) · · · 1(a + b)(a + b− 1) · · · (a + 1)(a + b + c)

(a + b + c− 1) · · · (a + b + 1)n(n− 1) · · · (a + b + c + 1).

That is,σ is made up of4 descending blocks, while any element in thejth block is smaller

18

a

b

c

d

Figure 2.1: A balanced Dyck path of 4 blocks

than any element in the(j + 1)st block, for1 ≤ j ≤ 3. Apparently,

area(Π) =

(a

2

)+

(b

2

)+

(c

2

)+

(d

2

),

inv(σ) =

(a

2

)+

(b

2

)+

(c

2

)+

(d

2

),

and thereforearea(Π) = inv(σ).

It is also easy to observe thatσ−1 = σ. Becauseσ has descents everywhere except the

last position of each block, we have

maj(σ−1) = maj(σ) =

(n

2

)− a− (a + b)− (a + b + c).

Note that

bounce(Π) = a + (a + b) + (a + b + c).

Therefore,

bounce(Π) =

(n

2

)−maj(σ).

19

For convenience we define the set of “balanced permutations”.

Definition 2.1.1. A permutationσ = σ1 · · ·σn ∈ Sn is said to be balanced if its one line

notation can be partitioned into a number of continuously descending blocks, such that any

element in a preceding block is smaller than any element in a later block, i.e.,σ is of the

form

σ =a1(a1 − 1) · · · 1 (a1 + a2)(a1 + a2 − 1) · · · (a1 + 1) · · ·

n(n− 1) · · · (ak−1 + . . . + a1 + 1),

for some integerk ≥ 1.

Let the number of balanced permutations inSn be bn. Above we have established

a bijectiong between the balanced Dyck paths and balanced permutations. From both

combinatorial structures, it is not hard to observe the following recurrence relation:

bn =n∑

i=0

bi,

b0 = 1.

Thereforebn = 2n−1.

Step 2: for generalΠ ∈ Dn .

Next we turn to the general case. It is useful to introduce the notion ofright balanced

path. Given a Dyck pathΠ, letB = b(Π) be the bounce path ofΠ. ClearlyB is a balanced

path. SupposeB consists ofm + 1 blocks. That is,B hasm left turns (from EAST to

NORTH), which are just the hits at the main diagonal, andm+1 right turns (from NORTH

to EAST). In general,Π has more area squares thanB geometrically. For1 ≤ j ≤ m, if

Π is identical withB from thejth left turn to the end point(n, n), we say thatΠ is j-right

20

balanced. In the caseΠ itself is a balanced path, i.e.B andΠ are identical from the origin

(0, 0), we say thatΠ is 0-right balanced(so 0-right balanced means balanced). Clearly,Π

beingj-right balancedimplies thatΠ is (j+1)-right balanced. Furthermore we allowj to

be any integer by convention.

In order to extend the mapg to the entire setDn, we start withB = b(Π) and let

Π(0) = B. According toStep 1, there is a permutationσ(0) = g(Π(0)) such that

area(Π(0)) = inv(σ(0)),

bounce(Π(0)) =

(n

2

)−maj(σ(0)).

Intuitively, we obtaing(Π) by each time adding squares to the area between two consec-

utive blocks ofΠ(0), so that it gets closer toΠ, and finding the corresponding permutation

that is the image of the modified path. SinceΠ(0) = B hasm + 1 blocks, we will reachΠ

together withg(Π) afterm steps. In other words, each time we modify the path obtained

earlier to become “more”right balanced, until we getΠ:

B = Π(0) → Π(1) → · · · → Π(m) = Π,

where eachΠ(j) is j-right balanced, 0 ≤ j ≤ m.

Inductively, for eachj, 1 ≤ j ≤ m, we modifyσ(j−1) = g(Π(j−1)) to obtainσ(j) =

g(Π(j)), which satisfies

area(Π(j)) = inv(σ(j)),

bounce(Π(j)) =

(n

2

)−maj(σ(j)).

For technical reasons, this modification is realized through a trick: givenσ(j−1), we

21

work on its inverse(σ(j−1))−1, instead ofσ(j−1) itself, to obtain(σ(j))−1, and afterwards

take the inverse again to obtainσ(j). We illustrate this process for the casem + 1 = 4, i.e.,

B consists of 4 blocks, and use the same setup ofΠ in Step 1for B. W.O.L.G., we choose

to show the second stage: assume forj = 2, we have already foundσ(1) = g(Π(1)) which

satisfies the two statistical identities withΠ(1), we go on to constructσ(2) = g(Π(2)). For

technical as well as symbolic convenience, letρ = (σ(1))−1

andτ = (σ(2))−1

. Also, we

make the inductive assumption thatρ is of the particular form

ρ =ρ1 · · · ρa+b(a + b + c)(a + b + c− 1) · · · (a + b + 1)

n(n− 1) · · · (a + b + c + 1),

whereρ1 · · · ρa+b could be any permutation inSa+b. Finally letΠ(2) be obtained fromΠ(1)

by adding a strip ofb1 squares to the top of the second EAST segment (from the bottom)

of B, adding a strip ofb2 squares to the top of the just added strip of lengthb1, . . ., and in

the end adding a strip ofbr squares to the top of the previously added strip of lengthbr−1.

To understand this construction, be aware that

1 ≤ br ≤ · · · ≤ b1 ≤ b,

1 ≤ r ≤ c− 1,

and see Figure 2.2 (ifr = 0 then just letτ = ρ, we are done and pass on; ifr = c, then

Π(2), and thereforeΠ(1) would have a different balanced path thanB which is impossible).

22

a

b

c

d

b1

b2

br

r<c

Figure 2.2: Add strips toΠ(1) to obtainΠ(2)

Now τ is ready.

τ =ρ1 · · · ρa+b−b1(a + b + c)ρa+b−b1+1 · · · ρa+b−b2(a + b + c− 1)

ρa+b−b2+1 · · · · · · ρa+b−br(a + b + c− (r − 1))ρa+b−br+1 · · · ρa+b

(a + b + c− r) · · · (a + b + 1) n(n− 1) · · · (a + b + c + 1).

Intuitively, in the process to findτ , we just move(a + b + c) b1 positions left to corre-

spond to theb1 squares in the newly added first row, and(a + b + c − 1) b2 positions left,

etc, while leave the remaining untouched.

Obviously,area(Π(2)) − area(Π(2)) = b1 + · · · + br. Note that for any permutationς,

we haveinv(ς) = inv(ς−1). Therefore,

inv(σ(2))− inv(σ(1)) = inv(τ)− inv(ρ)

= b1 + · · ·+ br,

23

and hence we have the first statistical identity forσ(2) andΠ(2) by the inductive hypothesis.

SinceB is the balanced path of all theΠ(j)s, bounce(Π(2)) = bounce(Π(1)) by the

definition of the bounce statistic. Next we prove thatmaj(σ(2)) = maj(σ(1)) by showing

that σ(1) = τ−1 andσ(2) = ρ−1 have the same descent set, so that we have the desired

second statistical identity forσ(2) andΠ(2).

Lemma 2.1.1.For any integeri, with 1 ≤ i ≤ n− 1,

i ∈ Des(ρ−1) ⇔ i ∈ Des(τ−1).

Proof. There are three cases fori.

• a + b + c + 1 ≤ i ≤ n. This case is trivial.

• 1 ≤ i ≤ a + b. Observe that by the construction ofτ ,

i ∈ Des(ρ−1) ⇔ ∃x < y, s.t.

1 ≤ x < y ≤ a + b,

ρ(x) = i + 1,

andρ(y) = i.

⇔ ∃x′ < y′, s.t.

1 ≤ x′< y

′ ≤ a + b + r,

τ(x′) = i + 1,

andτ(y′) = i.

⇔ i ∈ Des(τ−1).

• a + b + 1 ≤ i ≤ a + b + c. Thenρ(i) = a + b + 1 + (a + b + c − r). Because

ρ−1(a + b + c + 1) = n, i ∈ Des(ρ−1) iff i 6= a + b + c. On the other hand, this is

24

also the case forτ since it is also true that

τ−1(a + b + c) < τ−1(a + b + c− 1) < · · · < τ−1(a + b),

andτ−1(a + b + c + 1) = n.

So we have found a mapg from Dyck paths to permutations preserving the statistical

identities. The last thing we need to prove is thatg is an injection so that it is reversible.

Theorem 2.1.2.The mapg described in the above algorithm is an injection.

Proof. Given two Dyck pathsΠ1 andΠ2, we prove that they result in different images.

Case 1:supposeΠ1 andΠ2 have the same bounce pathB and they are obtained by the

following procedures, respectively,

B = Π(0)1 → Π

(1)1 → · · · → Π

(m)1 = Π1,

B = Π(0)2 → Π

(1)2 → · · · → Π

(m)2 = Π2.

whereΠ(j) is j-right balanced, 0 ≤ j ≤ m. For eachj, let σ(j)1 = g(Π

(j)1 ) andσ

(j)2 =

g(Π(j)2 ). So,g(Π1) = g(Π

(m)1 ) = σ

(m)1 andg(Π2) = g(Π

(m)2 ) = σ

(m)2 .

Assume the two procedures do not agree with each other for the first time at theith

step, i.e.,Π(j)1 = Π

(j)2 , and thereforeσ(j)

1 = σ(j)2 , 0 ≤ j ≤ i − 1, but Π

(i)1 6= Π

(i)2 . For

convenience, letρ = (σ(i)1 )−1 andτ = (σ

(i)2 )−1. W.O.L.G, assume the first disagreement

of Π1 andΠ2 is thatΠ1 adds a strip of more squares thanΠ2 does to the same row of the

jth block. As a result, while constructingρ there will be an integerx movingr places left,

passingr numbersy1, · · · , yr; but in the process of constructingτ the same integerx only

movess places left, withs < r, passing the verys numbersy1, · · · , ys. Notice that the

25

relevant order ofx andy1, · · · , yr will never change again after that in the later process of

constructingσ(m)1 andσ

(m)2 . So,x is to the right ofyr in (σ

(m)1 )−1, but to the left ofyr in

(σ(m)2 )−1, and thusσ(m)

1 6= σ(m)2 .

Case 2:supposeΠ1 andΠ2 have different bounce paths, respectivelyB1 andB2, and

Π1 andΠ2 are obtained by the following procedures,

B1 = Π(0)1 → Π

(1)1 → · · · → Π

(m1)1 = Π1,

B2 = Π(0)2 → Π

(1)2 → · · · → Π

(m2)2 = Π2.

Also letσ(j)1 = g(Π

(j)1 ) andσ

(j)2 = g(Π

(j)2 ). Sog(Π1) = σ

(m1)1 andg(Π2) = σ

(m2)2 .

W.O.L.G., assumeB1 andB2 do not agree for the first time at theith block (from bottom

to top), and that theith block of B1 is larger in size than theith block of B2. Then, theith

descending block of(σ(0)1 )−1 will have the form(x + r) · · · (x + 1)x and accordingly, the

ith descending block of(σ(0)2 )−1 has the form(x+s) · · · (x+1)x with r ≥ s+1. Be aware

that (x + s + 1) is always to the left of(x + s) in every(σ(j)1 )−1 during our procedure of

reachingΠ1 because we maintain the relative order of each block. Hence(x + s + 1) is to

the left of (x + s) in (σ(m1)1 )−1. On the other hand, observe that the(i + 1)st descending

block of (σ(0)2 )−1 is of the formy(y − 1) · · · (x + s + 1). That is,(x + s + 1) is at the the

end of the next block of(x + s) in (σ(0)2 )−1. By the rule of our algorithm,(x + s + 1) stays

unmoved during the modification from(σ(i)2 )−1 to (σ

(i+1)2 )−1 (again, because otherwiseΠ2

should have a different bounce path fromB2). Therefore(x+s+1) is to the right of(x+s)

in (σ(m2)2 )−1, and soσ(m1)

1 6= σ(m2)2 .

Henceg is a bijection betweenDn and a subset ofSn, which we callTn, and accordingly

we have the following corollary.

Corollary 2.1.3. Cn(q, t), the (q, t)-Catalan sequence is equal to the following distribution

26

function defined onTn.

Cn(q, t) = Fn(q, t) =∑σ∈Tn

qinv(σ)t(n2)−maj(σ).

Example2.1.1. Whenn = 3, T3 = {321, 231, 213, 132, 123}.Whenn = 4, T4 = {4321, 3214, 1324, 2134, 1234, 2143, 1243, 1432, 2314, 3421, 3241, 2431,

3142, 1342}.

2.1.2 The Foata-Schutzenberger Involution

An involution ψ is described in [FS78] that interchangesinv andmaj and preserves the

descent set. More specifically, we have the following theorem.

Theorem 2.1.4. [FS78] There exists an involutionψ : Sn → Sn with the property that

inv(ψ(σ)) = maj(σ) andinv(σ) = maj(ψ(σ)) hold simultaneously.

Lemma 2.1.5.The following polynomial is symmetric inq andt:

Sn(q, t) =∑σ∈Sn

qinv(σ)t(n2)−maj(σ).

Proof. Let c be the “complement” map such thatc(σ) = (n + 1− σ1)(n + 1− σ2) · · · (n +

1 − σn) for σ ∈ Sn. Note thatκ = cψ, whereψ is the Foata-Schutzenberger involution in

Theorem 2.1.4, is a bijection fromSn to Sn. Furthermore,

inv(κ(σ)) =

(n

2

)− inv(ψ(σ))

=

(n

2

)−maj(σ),

27

and

(n

2

)−maj(κ(σ)) =

(n

2

)− (

(n

2

)−maj(ψ(σ)))

= maj(ψ(σ))

= inv(σ).

Therefore,

Sn(q, t) =∑σ∈Sn

qinv(σ)t(n2)−maj(σ)

=∑σ∈Sn

qinv(κ(σ))t(n2)−maj(κ(σ))

=∑σ∈Sn

q(n2)−maj(σ)tinv(σ)

= Sn(t, q).

Notice that if we could replaceSn by Tn in the above theorem, then the symmetry of

the (q, t)-Catalan polynomial would be proved due to the bijectiong betweenDn andTn.

That however, would require a proof forκ = cψ to be closed onTn and would require a

better understanding of the setTn.

28

2.2 The Theory of Permutation Paths

2.2.1 Then! Permutation Paths and the Bijectionf betweenSn and

Pn

As an attempt to extend the idea of Dyck paths, we introduce the notion of Permutation

paths and develop the related theory.

Definition 2.2.1. A Permutation path of ordern is a lattice path from(0, 0) to (n, n) which

never goes below the main diagonal(i, i), 0 ≤ i ≤ n, or above the liney = n, and consists

of NORTH(0, 1), EAST(1, 0) and SOUTH(0,−1) steps but never repeats (i.e. no NORTH

step followed or preceded by a SOUTH step).

Let Pn denote the collection of Permutation paths of ordern. Figure 2.3 is an illustra-

tion of the 6 members inP3.

Theorem 2.2.1.

|Pn| = n! .

Proof. Note that any Permutation pathΠ ∈ Pn consists of some NORTH steps, some

SOUTH steps and exactlyn EAST steps made at different columns. In more detail, forj

from 1 ton−1, thesen EAST steps are in the form of(j−1, hj) → (j, hj) wherehj could

be any integer satisfyingj ≤ hj ≤ n becauseΠ never goes below the main diagonal or

above the liney = n. In factΠ is uniquely decided by these EAST steps or equivalently the

sequence of their heights (y-values)(h1, · · · , hn). Once these EAST steps are fixed, we just

connect them up by continuous NORTH or SOUTH steps, or possibly an empty vertical

move ifhj = hj+1. Since repeats are not allowed, the connection is unique. Therefore the

number of Permutation paths of ordern is equal to the number of sequences(h1, · · · , hn).

For eachj, sincej ≤ hj ≤ n, there aren + 1 − j ways to choosehj. Furthermore, the

choices of differenthj ’s are independent. So we are done.

29

The cardinality ofn! naturally motivates us to give a bijection betweenPn and the

symmetric groupSn. Actually the previous proof already provides hints of this bijection.

Lemma 2.2.2.There exists a bijectionf between the symmetric groupSn and the Permu-

tation pathsPn.

Proof. For anyΠ ∈ Pn, define theheight sequenceof Π to be the sequence of the heights

(y-values) ofΠ’s EAST steps, from left to right, as in the previous proof. Denote it by

h(Π) = (hΠ1 , . . . , hΠ

n ). Clearlyj ≤ hΠj ≤ n for eachj and any integer vector satisfying this

requirement is a height sequence for some uniquely decided Permutation path.

Givenσ = σ1 · · ·σn ∈ Sn, find its “lifted word” l(σ) = lσ1 · · · lσn where for1 ≤ j ≤ n,

lσj is whatσj would become if we map{σj, · · · , σn} to the set{j, · · · , n} keeping the

relative order of each element.

For example, ifn = 6 andσ = 6 2 4 3 5 1, thenl(σ) = 6 3 5 5 6 6: lσ1 = 6 because

σ1 = 6 is the biggest among{6, 2, 4, 3, 5, 1} when the set{6, 2, 4, 3, 5, 1} is mapped to

{1, 2, 3, 4, 5, 6} where 6 is also the biggest;lσ2 = 3 becauseσ2 = 2 is the second smallest

in {2, 4, 3, 5, 1} when the set{2, 4, 3, 5, 1} is mapped to{2, 3, 4, 5, 6} where the second

smallest element is 3, etc. Notice thatlσ1 is always equal toσ1, lσn is alwaysn and that

i ≤ lσi ≤ n for everyi.

So, l(σ) is a height sequence. Find its corresponding Permutation pathΠ and let

f(σ) = Π.

Conversely, given any Permutation pathΠ, locate its height sequenceh(Π). Actually

we will useh(Π) as the “lifted word” to findσ = f−1(Π). Let σ1 = hΠ1 . For i from 2 ton,

let σi be the(hΠi +1− i)th smallest number in the set{1, · · · , n}−{σ1, · · · , σi−1}. Clearly

l(σ) = h(Π) and hence the permutationσ so obtained isf−1(Π).

So we have established the bijectionf as desired.

Sometimes we use the height sequence to represent a Permutation path for convenience.

30

That is, we may writeΠ directly asΠ = (hΠ1 , . . . , hΠ

n ), whereh(Π) is the height sequence

of Π.

Example2.2.1. Whenn = 3, there are3! = 6 Permutation paths, and their correspondence

with the permutations inS3 throughf is indicated in Figure 2.3.

(1,3,3)=f(132) (2,2,3)=f(213)(1,2,3)=f(123)

(2,3,3)=f(231) (3,2,3)=f(312) (3,3,3)=f(321)

Figure 2.3: Correspondence betweenP3 andS3 underf

Thearea statistic, previously defined on Dyck pathsDn, may be extended toPn natu-

rally since a Permutation path never goes below the main diagonal. Simply, we letarea(Π)

be the number of complete squares betweenΠ and the main diagonal liney = x. This

agrees with the old definition ofarea onDn, which is a subset ofPn, if Π is also a Dyck

path.

The bijectionf has the nice property of mappinginv to area.

Theorem 2.2.3.f is a weight-preserving bijection betweenSn andPn that maps the in-

version statistic to the area statistic. Namely, for anyσ ∈ Sn, we have

inv(σ) = area(f(σ)).

31

Proof. Let f(σ) = Π andh(Π) = (hΠ1 , · · · , hΠ

n ). Notice that

area(Π) =n∑

i=1

hΠi − i

and for1 ≤ i ≤ n− 1 (hΠn − n = 0) we have

hΠi − i = |{j : σi > σj andi < j ≤ n}|.

So it is clear.

Corollary 2.2.4.∑

Π∈Pn

qarea(Π) = [n]!.

Proof. Recall that theinv statistic is Mahonian onSn [Mac60],

∑σ∈Sn

qinv(σ) = [n]!.

So the conclusion follows from Theorem 2.2.3. Alternatively, it could also be obtained

easily by induction.

As we did for Dyck paths, we may associate each Permutation path with an appropriate

word. Still encode eachN step by a0, eachE step by a1, and in addition encode eachS

step by a special character0∗.

Definition 2.2.2.A Permutation word of ordern is a permutation of the multiset{0n+s1n0∗s},where1 ≤ s ≤ bn

2cbn−1

2c , with the property that for1 ≤ i ≤ 2n+2s, in the initial subword

w = w1w2 · · ·wi,

• the number of0’s is at least as many as the sum of the numbers of1’s and the number

of 0∗’s;

32

• the number of0’s minus the number of0∗’s is at mostn.

We letAWn denote the set of Permutation words of ordern.

Definition 2.2.3. The inversion statistic of a Permutation wordw = w1w2 · · ·w2n+2s is

defined to be

inv(w) =∑

i: wi=1

n1(i)− n2(i),

wheren1(i) is the number of 0’s afterwi andn2(i) is the number of0∗’s afterwi.

The inversion statistic so defined onAWn is apparently an extension of the inversion

statistic on the Catalan wordsCWn, wheres = 0, i.e. no0∗’s exist.

Corollary 2.2.5.∑

w∈AWn

q(n2)−inv(w) = [n]!.

Proof. For anyw ∈ AWn, it is easy to see that

(n

2

)− inv(w) = area(Π(w)),

whereΠ(w) is the Permutation path thatw corresponds to.

2.2.2 Restrictingf to Some Pattern Forbidding Permutations

Since Dyck pathsDn is a subset ofPn and at least thearea statistic is extended toPn in

a nice way, we studyf−1(Dn) and some other related objects in order to understand the

Catalan phenomena as well as derive more general theories.

First we need some preliminary background on the theory of patterns.

Given permutationsτ ∈ Sk andσ ∈ Sn, we define anoccurrenceof the patternτ in σ

to be a choice ofk slots

1 ≤ i1 < · · · < ik ≤ n

33

such that the sequenceσi1 , . . . , σik is in the same order of relative size as the sequence

τ1, . . . , τk. In other words, for1 ≤ j1 < j2 ≤ k,

σij1< σij2

iff τj1 < τj2 .

Sometimes we also say that{σi1 , · · · , σik} is aτ -occurrence in σ.

Accordingly, if σ does not contain anyτ -occurrences of the pattern , we say thatσ is

τ -avoiding. Denote the set of allτ -avoiding permutations inSn by Sn(τ) [Pri97].

Example2.2.2. Considerσ = 51324 ∈ S5 and τ = 123 ∈ S3. σ is NOT τ -avoiding

because

{σ2, σ3, σ5} = {1, 3, 4}

is aτ -occurence inσ. Notice that

{σ2, σ4, σ5} = {1, 2, 4}

is also aτ -occurence inσ, but finding one occurence is sufficient for our purpose here.

Alternatively, let’s considerσ′

= 32541 ∈ S5 and τ′

= 312 ∈ S3. Thenσ′

is τ′-

avoiding because we can not find any 312-occurence inσ′= 32541. So we can say that

32541 ∈ S5(312).

The theory of pattern avoidance has been studied extensively. It is now well known

(see, for example, [Knu73]) that for anyτ ∈ S3, |Sn(τ)| = Cn. Partly because of this, we

are motivated to considerf−1(Dn), and we find that this pre-image is indeedSn(312). In

fact, there have been two direct bijections betweenSn(312) andDn which have occurred

recently in the literature: [Kra01] and [BK01]. The latter also gives a weight-preserving

34

bijection exchanging theinversion andarea statistics but it is different from ourf .

Let’s consider the restriction off onSn(312), the312-avoidingpermutations, and call

it f ∗. We prove thatf ∗ is a bijective map betweenSn(312) and Dyck pathsDn, a subset of

the image set of Permutation paths.

Theorem 2.2.6.f ∗ is a weight-preserving bijection betweenSn(312) and Dyck pathsDn

that maps the inversion statistic to the area statistic, and therefore

σ∈Sn(312)

qinv(σ) =∑

Π∈Dn

qarea(Π),

where the right hand side is Carlitz-Riordan’sq-Catalan polynomialCn(q) [CR64] which

satisfies the recurrence

Cn(q) =n∑

k=1

qk−1Ck−1(q)Cn−k(q).

Proof. Any Permutation pathΠ ∈ Pn can be uniquely represented by its height vector

h(Π) = (hΠ1 , . . . , hΠ

n ). Notice thatΠ ∈ Dn if and only if for 1 ≤ i ≤ n− 1, so we have

hΠi ≤ hΠ

i+1.

Givenσ ∈ Sn(312), we provef(σ) = Π ∈ Dn. Recall thath(Π) = l(σ). This means for

1 ≤ i ≤ n, hΠi = lσi is whatσi would become if we map{σi, · · · , σn} to the set{i, · · · , n}

keeping the relative order of each element. Revising this a little bit, let(lσi+1)− denote what

σi+1 would be at the “previous stage”, i.e, when we map{σi, · · · , σn} to {i, · · · , n} rather

than replacingi by i + 1 (for which we would getlσi+1 ). Now note that

• If σi+1 > σi, thenlσi+1 = (lσi+1)− > lσi .

35

• If σi+1 < σi, thenlσi+1 = (lσi+1)− + 1. So

lσi+1 ≥ lσi

⇔ (lσi+1)− ≥ lσi − 1

⇔ @k, s.t.i < i + 1 < k andσi+1 < σk < σi.

From the above conditions, it is clear thatσ ∈ Sn(312) impliesΠ ∈ Dn.

Conversely, givenΠ ∈ Dn, consider its pre-imagef−1(Π) = σ. Supposeσ /∈ Sn(312).

Take aminimal312-occurence{σi, σj, σk} in the sense

i < j < k,

σj < σk < σi,

and|j − i|+ |k − i| is minimal.

Π ∈ Dn implies thathΠi ≤ hΠ

i+1, and hencelσi ≤ lσi+1. This requiresσi − σi+1 ≤ 1, and

hencej 6= i+1. Then what aboutσi+1? If σi+1 > σk, then{σi+1, σj, σk} would be another

312-occurence, violating the minimality. Soσi+1 < σk. But then{σi, σi+1, σk} would be a

“less” 312-occurence. This shows thatσ /∈ Sn(312) is impossible.

Thereforef(Sn(312)) = Dn, or we may say that there exists a bijectionf ∗ from

Sn(312) toDn.

Sincef is weight-preserving, its restrictionf ∗ is also weight-preserving.

The above result gives rise to more general questions. Since we now knowf−1(Sn(312)) =

Dn, what isf−1(Sn(k12 . . . k − 1)) for generalk? We answer this question partially by

giving a lower bound.

For convenience, let the restriction off onSn(k12 · · · k− 1), the set ofk12 · · · (k− 1)-

avoidingpermutations, bef (k). To explain our combinatorial interpretation, we need the

36

following definition.

Definition 2.2.4. An m-cave of a Permutation pathΠ, with height sequenceh(Π) =

hΠ1 hΠ

2 · · ·hΠn , is a stepi satisfying that

max{hΠ1 , . . . , hΠ

i−1} − hΠi = m,

wherem ≥ 1. For anm-cavec, sometimes we say thatc is of depthm. An m-triangle

is a sequence of caves,c1, c2, · · · , cm, not necessarily continuous but in order from left to

right, wherecj is of depthat leastm + 1− j for each1 ≤ j ≤ m. If a Permutation pathΠ

does not contain anym-triangle, we say that P ism-triangle-forbidding. Denote the set

of m-triangle-forbidding Permutation paths of ordern byFn,m.

Geometrically, anm-cave ofΠ ∈ Pn is a step which ism squares down compared

with the highest level thatΠ has reached earlier (observe that the highest level thatΠ will

reach later is alwaysy = n). The juxtaposition of the not-necessarily continuous sequence

of caves in anm-triangle contains an isosceles right triangle with leg lengthm. Since we

requirem ≥ 1, a Dyck path has nom-cave orm-triangle. In addition,Fn,1 = Dn. The

following figure illustrates a pathΠ ∈ P8 with four caves:c1 at step 2 is of depth 1,c2 at

step 4 is of depth 1,c3 at step 5 is of depth 3 andc4 at step 6 is of depth 2. The cavec1

itself is a 1-triangle, and so is every other cave. The sequence of caves{c3, c4} forms the

only 2-triangle ofΠ and there is nom-triangle form ≥ 3.

The next theorem provides a combinatorial interpretation of the(m1 . . . m−1)-avoiding

permutations in terms of the(m− 2)-triangle-forbidding paths.

Theorem 2.2.7.For m ≥ 3, if σ ∈ Sn contains anm1 · · ·m− 1 pattern, then the Permu-

tation pathΠ = f(σ) must have an(m− 2)-triangle. That is,

Fn,m−2 ⊆ f(Sn(m1 · · ·m− 1)).

37

Figure 2.4: A Permutation pathΠ ∈ P8 with 4 caves and a 2-triangle

Whenm = 3,⊆ is replaced by equality.

Proof. For anyσ /∈ Sn(m12 . . .m−1), we proveΠ = f(σ) contains some(m−2)-triangle.

Assume{σi1 , · · · , σim} is a(m12 . . . m− 1)-occurence inσ, namely

i1 < i2 < · · · < im,

σi2 < · · · σim < σi1 .

In fact, forj from 2 tom− 1, we show that there is adj-cave at stepij, wheredj ≥ m− j.

Note that

hΠi2

< · · · < hΠim−1

.

Furthermore becauseσim−1 < σim < σi1, we have

hΠim−1

< hΠi1.

So,

max{hΠ1 , . . . , hΠ

ij−1} − hΠij≥ hΠ

i1− hΠ

ij≥ m− j.

Hence at stepij, 2 ≤ ij ≤ m − 1, we have a cave of depth at leastm − j and therefore

38

Π = f(σ) contains an(m− 2)-triangle.

m = 3 is the case of Dyck paths discussed in Theorem 2.2.6.

Remark2.2.1. Conversely, givenσ ∈ Sn(m12 . . .m − 1), Π = f(σ) does not necessarily

forbid (m − 2)-triangles. For example, letn = 8 andm = 4, σ = 57836241 ∈ S8(4123),

but the corresponding Permutation pathΠ is not 2-triangle-forbidding. See Figure 2.5.

Figure 2.5:σ = 57836241 ∈ S8(4123) butΠ is not 2-triangle-forbidding.

Nonetheless, we now know that in terms of cardinality,

|Fn,m−2| ≤ |Sn(m12 . . . m− 1)|.

A recent result of Backelin, West and Xin ([BWG], see [SW02]) implies that for anyk ∈ N,

|Sn(m12 . . .m− 1) = Sn(12 . . . m)|.

So Theorem 2.2.7 provides a lower bound estimate of the number of permutations inSn

whose longest increasing subsequence has lengthm.

39

2.2.3 The Signed Permutation PathsBn

In this section we introduce the notion of Signed Permutation paths. This may be viewed

as a generalization of both Permutation paths and Schroder paths.

Definition 2.2.5. A Signed Permutation path of ordern is a lattice path from(0, 0) to (n, n)

which never goes below the main diagonal(i, i), 0 ≤ i ≤ n, or above the liney = n, and

consists of NORTH(0, 1), EAST (1, 0), SOUTH (0,−1) and Diagonal(1, 1) steps but

never repeats (i.e. no NORTH step followed or preceded by a SOUTH step).

Let Bn denote the collection of Signed Permutation paths of ordern. Figure 2.6 is an

illustration of a Signed Permutation path inB8.

Figure 2.6: A Signed Permutation path.

Similar to the situation before,Bn is closely related to the set of signed permutations.

A signed permutationis a permutationσ ∈ Sn where eachσi has a plus or minus sign

attached to it [HLR]. The set of signed permutations is also called the hyperoctahedral

group, which is denoted byBn and studied by Reiner [Rei93].

Definition 2.2.6. For a signed permutationσ ∈ Bn, the absolute-value inversion statistic

inv is defined to be

inv(σ) = inv(|σ|),

40

where|σ| denotes the ordinary permutation obtained fromσ by removing the plus or minus

sign attached to eachσi.

Theorem 2.2.8.There exists a weight-preserving bijectionϕ betweenBn andBn that maps

the absolute-value inversion statistic to the area statistic. Namely, for anyσ ∈ Bn, we have

inv(σ) = area(ϕ(σ)).

Proof. Given σ ∈ Bn, find f(|σ|) = Π, wheref is our old weight-preserving bijection

betweenSn andPn. Suppose the height sequence ofΠ is h(Π) = (hΠ1 , . . . , hΠ

n ). NotehΠi

means that at theith column, theE step ofΠ goes from(i−1, hΠi ) to (i, hΠ

i ). Now for each

1 ≤ i ≤ n, if σi is positive, leave it untouched; ifσi negative, then change the originalE

step to aD step which goes(i − 1, hΠi − 1) to (i, hΠ

i ). The path so modified fromΠ is a

Signed Permutation path and we let it beϕ(σ).

Conversely, givenΠ ∈ Bn, first raise each of itsD steps to anE step with the same

height of the ending height of theD step. Connect where appropriate. Find the preimage

underf of the Permutation path thus obtained and call itτ . For 1 ≤ i ≤ n, if Π has an

E step at columni, let σi = τi; if Π has aD step at columni, let σi = −τi. The signed

permutationσ so obtained isϕ−1(Π).

In our construction of the mapϕ, the Signed Permutation path has the same area as the

Permutation path modified by changingD steps toE steps, andinv(σ) = inv(|σ|). Hence

by Theorem 2.2.3, our conclusion follows.

In Figure 2.6,ϕ−1(Π) = (−7)6(−4)8(−5)(−2)13. Clearly, it is true thatarea(Π) =

22 = inv(76485213).

Corollary 2.2.9.∑Π∈Bn

qarea(Π) = 2n[n]!.

41

Proof. For eachσ ∈ Sn, there are2n ways to attach plus or minus signs to its entries to

make it a signed permutation. Each of the2n signed permutation has the same absolute-

value inversion statistic asinv(σ). So it is clear from Theorem 2.2.8.

42

Chapter 3

Tree Enumeration Theorems and the

(q, t)-Parking Function Polynomial

3.1 Haglund and Loehr’s (q, t)-Parking Function Polyno-

mial

The standard definition ofparking functionis as follows [Sta99, Ex.5.49, pages 94-95]:

For fixedn, there aren parking places1, 2, . . . , n (in that order) on a one-way street. Cars

C1, . . . , Cn enter that street in that order and try to park. Each carCi has a preferred space

ai. A car will drive to its preferred space and try to park there. If the space is already

occupied, the car will park in the next available space. If the car must leave the street

without parking, then the process fails. IfP = (a1, . . . , an) is a sequence of preferences

that allows every car to park, then we callP a parking function. It is easy to see that

a sequence(a1, . . . , an) is a parking function if and only if the increasing rearrangement

(b1, . . . , bn) of (a1, . . . , an) satisfiesbi ≤ i. It is also known that the number of parking

43

functions of lengthn is given by

Park(n) = (n + 1)n−1,

which is equal to the number of labelled trees on the labeling set{0, 1, 2, . . . , n}.As introduced in [HL], a parking functionP can also be obtained by starting with a

Dyck pathD and placingn “cars”, denoted by the integers 1 throughn, in the squares

immediately to the right of the vertical segments ofD, with the restriction that if cari is

placed immediately on top of carj, theni > j. It is easy to see this definition is in bijection

with the one defined earlier: having carsi1, . . . , ij at columni is equivalent to say that

exactly those carsCi1 , . . . , Cij havei as their preferred space. For any parking functionP ,

let D(P ) be the Dyck path thatP corresponds to, i.e.,D is obtained by removing the cars

from P . LetPn = {P : D(P ) ∈ Dn} be the parking functions onn cars. An example of a

parking function is given in Figure 3.1.

2

4

7

5

1

8

3

6

Figure 3.1: A parking functionP ∈ P8 with area(D(P )) = 6.

Haglund and Loehr introduced a distribution function over the set of parking functions

defined in this manner and made the following conjecture.

44

Conjecture 3.1.1.Define

Rn(q, t) =∑

P :D(P )∈Dn

qarea(D(P ))tdinv(P ),

where the sum is over all parking functions onn cars, anddinv is another statistic on

Dyck paths (which we shall not discuss here, see[HL] ). ThenRn(q, t) = Hn(q, t), where

Hn(q, t) denotes the Hilbert Series of the space of diagonal harmonics (see[Hai94]).

Although Conjecture 3.1.1 has been verified forn ≤ 11, by Garsia, Haglund, Loehr

and Ulyanov using Maple, it remains a conjecture. Even the symmetry ofRn(q, t) in q and

t is not proved. The main obstacle is the lack of a recurrence forRn(q, t) . As a partial

result, Loehr proved thatRn(q, 1) = Rn(1, q) [Loe03].

In the second section of this chapter, we prove some labelled tree counting theorems,

which are meaningful by themselves, and use the “Least-Child-Being-Monk” theorem to

get a recurrence relation onRn(1, 1). In the last section, we prove thatRn(q, 1) = Mn(q),

whereMn(q) is thearea statistic for the major sequence [Kre80], thus establishing the

equivalence between theq = 1 or t = 1 specialization of (q, t)-parking functions and a

group of other combinatorial statistics and thereby getting another recurrence onRn(q, 1) =

Rn(1, q) by the known facts.

3.2 Counting Special Families of Labelled Trees

In the process of looking for a recurrence relation onRn(q, 1), we found some interesting

theorems about enumerating special families of labelled trees. Before proceeding to the

main theorems, we first cite a lemma of L. E. Clarke to be used in our proof.

Lemma 3.2.1. ( [Cla58], see also [Moo70] or [Ber76])The number of forests consisting

45

of k rooted trees onn− j nodes is

(n− j

k

)k(n− j)n−j−k−1.

Throughout this section we consider rooted trees and rooted forests, where the notions

“child” and “descendant” are defined in the standard way: nodei is achild of nodej if i is

exactly one edge further away from the root; nodei is adescendantof nodej if i is one or

more edges further away from the root. But sometimes we may drop the word “rooted” if

there is no confusion. The convention we use is that every free tree corresponds to a rooted

tree naturally by designating the least labelled vertex to be the root. Our main concern is

to count families of labelled trees with some special structures, and the results will have no

difference if we designate a different root, which we will do occasionally.

Definition 3.2.1. A labelled tree rooted at its least labelled vertex isLeast-Child-Being-

Monk if it has the property that the least labelled child of 0 has no children (or equivalently,

is a leaf).

Theorem 3.2.2.(“Least-Child-Being-Monk”) DefineTn+1,0 to be the set of trees labelled

on {0, 1, 2, ..., n + 1} with the Least-Child-Being-Monk property. Then the cardinality of

Tn+1,0, which we denote bytn+1,0, is equal tonn.

Example3.2.1. Whenn = 2, there are altogether 4 trees labelled on{0, 1, 2, 3} having the

Least-Child-Being-Monk property, as illustrated by Figure 3.2.

Proof. Nontrivially, assumen ≥ 2 so that0 has more than one descendants. If the least

child of 0 is i and0 hasj other children, then thesej children are selected randomly from

i + 1, · · · , n, n + 1. Now that we have0 and its childreni, c1, c2,· · · ,cj, we only need to

build the othern− j nodes into a forest ofk rooted trees and attach these roots of the forest

to some or all of the “free children”c1, c2, · · · , cj in jk ways, but that is exactly counted by

46

0

1 2 3 1 2

3

1 3

2

2 3

1

0 0 0

Figure 3.2: The number of “Least-Child-Being-Monk” trees on{0, 1, 2, 3} is 22 = 4.

Lemma 3.2.1. Therefore,

|Tn+1,0| =n+1∑i=1

n+1−i∑j=0

(n + 1− i

j

) n−j∑

k=0

(n− j

k

)k(n− j)n−j−k−1 jk

=n∑

j=0

n−j∑

k=0

(n− j

k

)k(n− j)n−j−k−1jk

n+1∑i=1

(n + 1− i

j

)

=n∑

j=0

n−j∑

k=0

(n− j

k

)k(n− j)n−j−k−1jk

(n + 1

j + 1

)

=n∑

j=0

(n− j)n−j−1

(n + 1

j + 1

)j

n− j

n−j∑

k=0

(n− j

k

)k(

j

n− j)k−1

=n∑

j=0

(n− j)n−j−1

(n + 1

j + 1

)j

n− j

d[(1 + x)n−j]

dx|x= j

n−j

=n∑

j=0

(n− j)n−j−1

(n + 1

j + 1

)j

n− j(n− j)(1 +

j

n− j)n−j−1

=n∑

j=0

(n + 1

j + 1

)jnn−j−1

= nn−1(n∑

j=0

(n + 1

j + 1

)(j + 1)n−j −

n∑j=0

(n + 1

j + 1

)n−j).

47

Since

n∑j=0

(n + 1

j + 1

)(j + 1)n−j

=n+1∑

k=1

(n + 1

k

)(k)(

1

n)k−1

=d[(1 + x)n+1]

dx|x= 1

n

=(n + 1)n+1

nn

and

n∑j=0

(n + 1

j + 1

)n−j

= n

n+1∑

k=1

(n + 1

k

)(1

n)k

=(n + 1)n+1

nn− n,

we haven∑

j=0

(n + 1

j + 1

)(j + 1)n−j −

n∑j=0

(n + 1

j + 1

)n−j = n.

Therefore,

|Tn+1,0| = nn−1 · n = nn.

Corollary 3.2.3.

Rn(1, 1) =n∑

i=1

(i− 1)i−1

(n

i

)Rn−i(1, 1)

Proof. DefineP∗n, primary parking functionsof order n, to be the subset ofPn which

48

touches the main diagonaly = x only at(0, 0) and(n, n), and let

R∗n(q, t) :=

P∈P∗nqarea(P )tdinv(P ).

Figure 3.3 shows an example of a primary parking function, while the parking function

illustrated by Figure 3.1 isnot primary.

2

4

5

3

1

8

6

7

Figure 3.3: A primary parking functionP ∈ P∗8 , with area(D(P )) = 12.

Then, by decomposing the parking functions inPn,

Rn(1, 1) =n∑

i=1

R∗i (1, 1)

(n

i

)Rn−i(1, 1).

So it suffices to show

R∗i (1, 1) = (i− 1)i−1.

R∗i (1, 1) counts the number of primary parking functions of orderi. [HL] provides a

bijection betweenP∗i andTi,0. Therefore it follows from Theorem 3.2.2.

Remark3.2.1. The primary parking functions are also counted by a different way in [Sta99],

where a different definition of primary parking functions is used.

49

Corollary 3.2.4. Whenn goes to infinity, the probability for a labelled tree to be “Least-

Child-Being-Monk” ise−2.

Proof. Using Cayley’s formula and Theorem 3.2.2, the desired probability is

limn→∞

nn

(n + 2)n=

1

e2.

Corollary 3.2.5. Definefn,0 to be the number of rooted forests onn nodes such that each

tree in the forest is rooted at its least labelled vertex and has the “Least-Child-Being-Monk”

property. Then

fn,0 =n∑

k=1

∑n1+···+nk=n

ni>0

(n

n1,··· ,nk

)|(n1 − 2)n1−2 · · · (nk − 2)nk−2|k!

,

where the| | is to ensure its validity when someni takes the value of 1.

Proof. Note thattn+1,0 = nn actually counts labelled trees with the “Least-Child-Being-

Monk” property onn + 2 vertices (and it does not make any difference which vertex is

the root). Givenn vertices, our task is to partition them intok groups and build the nodes

in each group into a tree with the “Least-Child-Being-Monk” property so that the least

labelled vertex is the root in each tree. The conclusion readily follows.

The initial terms of{fn,0}n≥1 are 1, 2, 5, 18, 93, 104, ...

Remark3.2.2. Let t∗n = tn−1,0 so thatt∗n denotes the number of rooted trees onn nodes

with the “Least-Child-Being-Monk” property, and define exponential generating functions

of t∗n andfn,0 by T (x) andF (x), respectively. Then by the Exponential Formula [Wil94],

we have

F (x) = eT (x),

50

which is equivalent to our formula forfn,0 as in the Corollary. Furthermore, letf(k)n,0 be the

number of rooted forests onn nodes that consist ofk rooted trees with “Least-Child-Being-

Monk” property, and introduce the 2-variable generating function

F (x, y) =∑

n,k≥0

f(k)n,0

xn

n!yk.

Again by the the Exponential Formula [Wil94], we have

F (x, y) = eyT (x).

It is not hard to derive from here that

f(k)n,0 = [

xn

n!yk]F (x, y)

= [xn

n!yk]eyT (x)

=∑

n1+···+nk=nni>0

(n

n1,··· ,nk

)(n1 − 2)n1−2 · · · (nk − 2)nk−2

k!.

This is a refinement of Corollary 3.2.5, or as we may say, another proof.

The “Least-Child-Being-Monk” theorem has some nice generalizations. Instead of re-

quiring the “Least-Child-Being-Monk”, we may let the least labelled child of 0 havep

descendants.

Theorem 3.2.6.DefineTn+1,p to be the set of labelled trees on{0, 1, 2, ..., n+1}, such that

the total number of descendants of the least labelled child of 0 isp. Then, the cardinality

of Tn+1,p, denoted bytn+1,p, is equal to

(n− p)n−p(p + 1)p−1

(n + 1

p

).

51

Example3.2.2. The casep = 0 is dealt with in Theorem 3.2.2. Figure 3.4 is an illustration

of T3,1 for the casen = 2 andp = 1. There are a total oft3,1 = 3 trees labelled on

{0, 1, 2, 3} so that the least child of 0 has exactly 1 descendant.

0 0 0

3 21 1 2

2 3 1

3

Figure 3.4: The3 members ofT3,1.

First Proof. The proof is similar to thep = 0 case, though a little more complicated.

tn+1,p =n+1∑i=1

n+1−i−p∑j=0

(n + 1− i

j

)(n− j

p

)(p + 1)p−1

n−j−p∑

k=0

(n− j − p

k

)k(n− j − p)n−j−k−1−p jk

= (p + 1)p−1

n−p∑j=0

n−j−p∑

k=0

(n− j − p

k

)k(n− j − p)n−j−p−k−1jk

(n− j

p

)

n+1∑i=1

(n + 1− i

j

)

= (p + 1)p−1

n−p∑j=0

n−j−p∑

k=0

(n− j − p

k

)k(n− j − p)n−j−p−k−1jk

(n− j

p

)(n + 1

j + 1

)

= (p + 1)p−1

n−p∑j=0

(n− j

p

)(n + 1

j + 1

)(n− j − p)n−j−p−1 j

n− j − p

n−j−p∑

k=0

(n− j − p

k

)k(

j

n− j − p)k−1

52

= (p + 1)p−1

n−p∑j=0

(n− j

p

)(n + 1

j + 1

)(n− j − p)n−j−p−1

j

n− j − p

d[(1 + x)n−j−p]

dx| x= j

n−j−p

= (p + 1)p−1

n−p∑j=0

(n− j

p

)(n + 1

j + 1

)(n− j − p)n−j−p−1

j

n− j − p(n− j − p)(1 +

j

n− j − p)n−j−p−1

= (p + 1)p−1

n−p∑j=0

(n− j

p

)(n + 1

j + 1

)j(n− p)n−j−p−1

= (p + 1)p−1(n− p)n−p−1 (n + 1)!

p! (n− p + 1)!

n−p∑j=0

(n− p + 1

j + 1

)j(n− p)−j.

Analogous to the previous proof, we have

n−p∑j=0

(n− p + 1

j + 1

)(j + 1)(n− p)−j

=

n−p+1∑

k=1

(n− p + 1

k

)(k)(

1

n− p)k−1

=d[(1 + x)n−p+1]

dx|x= 1

n−p

=(n− p + 1)n−p+1

(n− p)(n−p)

53

and

n−p∑j=0

(n− p + 1

j + 1

)(n− p)−j

= (n− p)

n−p+1∑

k=1

(n− p + 1

k

)(

1

n− p)k

=(n− p + 1)n−p+1

(n− p)(n−p)− (n− p).

So,n−p∑j=0

(n− p + 1

j + 1

)j(n− p)−j = n− p,

and hence,

tn+1,p = (p + 1)p−1(n− p)n−p−1

(n + 1

p

)(n− p)

= (n− p)n−p(p + 1)p−1

(n + 1

p

).

¤

Second Proof. Notice that we just need to (1) choosep nodes from{1, · · · , n + 1} as the

descendants of the least child of 0, (2) arrange thep nodes into a rooted forest, (3) build

the remainingn − j + 1 nodes as well as 0 into a tree with the Least-Child-Being-Monk

property, and (4) attach all the roots of the forest obtained in the second step to the least

child of 0 obtained in the third step. Clearly, the numbers of ways to realize the first two

steps are(

n+1p

)and(p + 1)p−1, respectively. By the “Least-Child-Being-Monk” theorem,

there are in total(n− p)n−p ways in the third step. The fourth step is done in a unique way.

So it is clear.

¤

54

Remark3.2.3. If we add up all thep’s, i.e, the total number of descendants of the least

labelled child of 0, then we get the following identity by Cayley’s formula:

n∑p=0

(n− p)n−p(p + 1)p−1

(n + 1

p

)= (n + 2)n,

which is equivalent to

n+1∑p=0

(n− p)n−p(p + 1)p−1

(n + 1

p

)= 0

and further becomes

n∑p=0

(n− 1− p)n−1−p(p + 1)p−1

(n

p

)= 0 (3.2.1)

when we drop the scale fromn + 1 to n.

Eq. ( 3.2.1) reminds us ofAbel’s indentity, a striking generalization of the binomial

theorem.

Theorem3.2.7. (Abel’s identity, see [Abe26], [Com74] or [Str92])For all x, y, z, we have:

n∑

k=0

(n

k

)x(x− kz)k−1(y + kz)n−k = (x + y)n.

If we let k = p, x = 1, z = −1 andy = n− 1, then we have the specialization

n∑p=0

(n− 1− p)n−p(p + 1)p−1

(n

p

)= nn.

This is very similar to ( 3.2.1); however ( 3.2.1) can not be derived from any direct spe-

cialization of Abel’s identity. Nevertheless, our identity is indeed obtainable if we apply

55

Theorem 3.2.7 twice. Since(

n+1p

)=

(np

)+

(n

p−1

),

n∑p=0

(n− p)n−p(p + 1)p−1

(n + 1

p

)

=n∑

p=0

(n− p)n−p(p + 1)p−1

(n

p

)+

n∑p=0

(n− p)n−p(p + 1)p−1

(n

p− 1

).

In Abel’s identity, letk = p, x = 1, z = −1 andy = n, then

n∑p=0

(n− p)n−p(p + 1)p−1

(n

p

)= (n + 1)n. (3.2.2)

On the other hand,

n∑p=0

(n− p)n−p(p + 1)p−1

(n

p− 1

)

=n+1∑p=1

(n− p)n−p(p + 1)p−1

(n

p− 1

)− (−1)−1(n + 2)n

(n

n

)

=n∑

j=0

(n− j − 1)n−j−1(j + 2)j

(n

j

)+ (n + 2)n.

In Abel’s identity, letk = n− j, x = −1, z = −1 andy = n + 2, then

n∑j=0

(n

j

)(−1)(n− j − 1)n−j−1(j + 2)j = (n + 1)n.

So,

n∑p=0

(n− p)n−p(p + 1)p−1

(n

p− 1

)= −(n + 1)n + (n + 2)n. (3.2.3)

56

Sum up (3.2.2) and (3.2.3), again we have

n∑p=0

(n− p)n−p(p + 1)p−1

(n + 1

p

)= (n + 2)n.

Corollary 3.2.8. Whenn goes to infinity, the probability for a labelled tree on{0, 1, 2, · · · , n}to have the property that the least labelled child of 0 has exactlyp descendants is

(p + 1)p−1

p !e−2−p.

Proof.

limn→∞

(n− p)n−p(p + 1)p−1(

n+1p

)

(n + 2)n=

(p + 1)p−1

p !e−2−p.

Let Tn+1 be the collection of all labelled trees on{0, 1, · · · , n + 1}. For anyT ∈Tn+1, let y(T ) denote the size of the “youngest descendant branch”, i.e., the number of

descendants of the least labelled child of0. Furthermore, let

Yn(q) =∑T∈Tn

qy(T ).

A different way to state Theorem 3.2.6 is

Yn+1(q) =n∑

p=0

(n− p)n−p(p + 1)p−1

(n + 1

p

)qp.

Another generating function worth studying is

Zn(q) =∑

T∈Tn+1

qz(T ),

57

wherez(T ) = degT (0) − 1, i.e., z(T ) is the number of children of0 excluding the least

labelled one.

Theorem 3.2.9.

Zn+1(q) =n∑

j=0

n−j∑p=0

(p + 1)p−1j(n− p)n−p−j−1

(n + 1

p, j + 1, n− p− j

)qj.

Proof. Let

Zn+1,p(q) :=∑

T∈Tn+1,p

qz(T ).

It suffices to show that for1 ≤ p ≤ n,

Zn+1,p(q) =

n−p∑j=0

(p + 1)p−1j(n− p)n−p−j−1

(n + 1

p, j + 1, n− p− j

)qj.

Similar to the proof of Theorem 3.2.6 and following exactly the same process as we went

through, we obtain

Zn+1,p(q) =n+1∑i=1

n+1−i−p∑j=0

(n + 1− i

j

)(n− j

p

)(p + 1)p−1qj

n−j−p∑

k=0

(n− j − p

k

)k(n− j − p)n−j−k−1−p jk

= · · ·

= (p + 1)p−1(n− p)n−p−1 (n + 1)!

p! (n− p + 1)!

n−p∑j=0

(n− p + 1

j + 1

)j(

q

n− p)j

.

Now it is not hard to see the coefficient ofqj of the above formula is

(p + 1)p−1j(n− p)n−p−j−1

(n + 1

p, j + 1, n− p− j

).

58

Remark3.2.4. Theorem 3.2.9 covers and refines Theorem 3.2.2 and Theorem 3.2.6. For

instance,

[qj]Zn+1,p(q) = (p + 1)p−1j(n− p)n−p−j−1

(n + 1

p, j + 1, n− p− j

)

gives the number of labelled trees on{0, 1, · · · , n + 1} such that the “least child” of 0 has

exactlyp descendants anddegT (0) = j + 1. In addition, if we specializeq = 1 andp = 0

in Theorem 3.2.9, then we get

Zn+1,0(1) =n∑

j=0

jnn−j−1

(n + 1

j + 1

)

= nn,

which is the value of|Tn+1,0| as in Theorem 3.2.2 ortn+1,0 as in Theorem 3.2.6.

From a different point of view, we can also extend the behavior of the “least child of 0

being single” to a “hereditary” property. That is, we want every “least child” in a rooted

tree, instead of just the least child of 0, to be single. According to some old European

Church tradition, the youngest child of each family will serve as a priest, so as to have

no children. Translating this situation into graph theory, we are motivated to consider the

following counting question: what is the number of labelled trees such that the least labelled

child of any vertexhas no children?

Definition 3.2.2. A labelled tree rooted at its least labelled vertex isHereditary-Least-

Singleif it has the property thatevery least childin this tree has no children (or equivalently,

is a leaf).

59

Lemma 3.2.10.(Hereditary-Least-Single Trees Pre-recurrence) Let the number of Hereditary-

Least-Single trees withn vertices behn. Thenhn satisfies the following recurrence:

hn =n−2∑

k=0

(n− 1

k + 1

) ∑n1+···+nk=n−k−2

ni≥0

(n− k − 2

n1, · · · , nk

)hn1+1 · · ·hnk+1, n ≥ 2,

h1 = 1.

Proof. Firstly, we choosek + 1 children of the root:c0 < c1 < · · · < ck. Secondly, for

each of thek children except the least, let the number of descendants thatci will have be

ni ≥ 0, i = 1 . . . k. Lastly, for eachi, build up a subtree rooted atci together with itsni

descendants. Note that here is a trick: althoughci may not be the least labelled vertex in

this subtree, the number of such subtrees is exactlyhni+1 since the Hereditary-Least-Single

property is essentially only concerned with every vertex except the root.

Theorem 3.2.11.Consider the exponential generating functionH(x) =∑

n≥0hn+1

n!xn.

ThenH(x) satisfies the simple functional equation

H2(x)−H(x) + 1 = exH(x).

Proof. By Lemma 3.2.10, forn ≥ 2,

hn =n−2∑

k=0

(n− 1

k + 1

)(n− k − 2)!

n1+···+nk=n−k−2

hn1+1

n1!· · · hnk+1

nk!

=n−2∑

k=0

(n− 1

k + 1

)(n− k − 2)![xn−k−2]Hk(x)

= (n− 1)!n−2∑

k=0

[xn−k−2]Hk(x)

(k + 1)!.

Here by the symbol[xn]f(x) we mean the coefficients ofxn in the seriesf(x). So on one

60

hand,

[xn−1]H(x) =hn

(n− 1)!=

n−2∑

k=0

[xn−k−2]Hk(x)

(k + 1)!.

On the other hand,

n−2∑

k=0

[xn−k−2]Hk(x)

(k + 1)!=

n−2∑

k=0

[xn−1]xk+1Hk(x)

(k + 1)!

= [xn−1]∞∑

k=0

xk+1Hk(x)

(k + 1)!

= [xn−1]1

H(x)

∞∑

k=0

(xH(x))k+1

(k + 1)!

= [xn−1]exH(x) − 1

H(x),

for anyn ≥ 2. Therefore,

H(x) =exH(x) − 1

H(x)+ C,

whereC is a constant to be determined. By checking the constant term on both sides, we

seeC = 1 and the equation follows.

Theorem 3.2.12.(Hereditary-Least-Single Trees Recurrence)hn satisfies the following

recurrence:

hn =(n− 1)hn−1 − 2∑

1≤i≤n−2

hn−ihi+1

(n− 2

i− 1, n− i− 1

)

+∑

1≤i≤n−2

∑1≤j≤n−i−1

ihihjhn+1−i−j

(n− 2

i− 1, j − 1, n− i− j

).

Proof. For simplicity, let

H(x) =∑n≤0

anxn

61

and

a0 = 1,

s.t.

an =hn+1

n!.

Take the derivative with respect tox on both sides of

xH = ln(H2 −H + 1)

to obtain

xH ′ + H =2HH ′ −H ′

H2 −H + 1.

Therefore

(xH ′ + H)(H2 −H + 1) = 2HH ′ −H ′.

For the above equation, since

[xn] LHS =∑

0≤m≤n

[xm](xH ′ + H)[xn−m](H2 −H + 1)

=∑

0≤m≤n

(mam + am)(∑

0≤r≤n−m

aran−m−r − an−m + δn−m,0)

=∑

0≤m≤n

(m + 1)(∑

0≤r≤n−m

aran−m−ram − aman−m + δn−m,0am),

62

and

[xn] RHS =∑

0≤m≤n

[xm](2H − 1)[xn−m](H ′)

=∑

0≤m≤n

(2am − δm,0)(n−m + 1)an−m+1

=∑

0≤m≤n

(2an−m − δn−m,0)(m + 1)am+1

= (n + 1)an+1 +∑

0≤m≤n−1

2an−m(m + 1)am+1,

we have

an+1

=

[xn] LHS −∑

0≤m≤n−1

2an−m(m + 1)am+1

n + 1

=

(n + 1)an +∑

0≤m≤n−1

(m + 1)(∑

0≤r≤n−m

aran−m−ram − aman−m − 2an−mam+1)

n + 1

= an +

∑0≤m≤n−1

(m + 1)(∑

0≤r≤n−m−1

aran−m−ram − 2an−mam+1)

n + 1.

Hence,

an−1 = an−2 +

∑0≤m≤n−3

(m + 1)(∑

0≤r≤n−m−3

aran−2−m−ram − 2an−2−mam+1)

n− 1,

63

which is equivalent to

hn

(n− 1)!=

hn−1

(n− 2)!+

∑0≤m≤n−3

(m + 1)(∑

0≤r≤n−m−3

hr+1

r!

hn−1−m−r

(n− 2−m− r)!

hm+1

m!− 2

hn−1−m

(n− 2−m)!

hm+2

(m + 1)!)

n− 1.

Therefore,

hn = (n− 1)hn−1 −∑

0≤m≤n−3

2hn−1−mhm+2

(n− 2

n− 2−m, m

)

+∑

0≤m≤n−3

∑0≤r≤n−m−3

(m + 1)hr+1hn−m−r−2+1hm+1

(n− 2

r,m, n− 2−m− r

)

= (n− 1)hn−1 −∑

1≤i≤n−2

2hn−ihi+1

(n− 2

n− i− 1, i− 1

)

+∑

1≤i≤n−2

∑1≤j≤n−1−i

ihihjhn−i−j+1

(n− 2

i− 1, j − 1, n− i− j

).

So we are done.

The following list computed by Maple using our recurrence contains{hn}10n=1: 1, 1, 1,

4, 15, 96, 665, 6028, 60907, 725560 ...

3.3 Major Sequences, Tree Inversion Enumerator and the

Tutte Polynomial

In this section we investigate several interesting combinatorial objects related toRn(q, t),

and establish the equivalence betweenRn(q, 1) and a group of combinatorial formulae: the

area enumerator for major sequences, the inversion enumerator for labelled trees and the

64

Tutte polynomial specialized at(1, q). Loehr [Loe03] recently proved thatRn(q, 1) =

Rn(1, q), so we can also includeRn(1, q) in this equivalence family.

An integer sequences = (s1, · · · , sn) is called amajor sequenceof lengthn [Kre80]

if its non-decreasing rearrangement(z1, · · · , zn) satisfies

i ≤ zi ≤ n, for all 1 ≤ i ≤ n.

The area of a major sequences = (s1, · · · , sn) is defined as

area(s) =n∑

i=1

si −(

n + 1

2

).

If we denote byMn the set of major sequences of lengthn, then we can further define the

area enumerator for major sequences of lengthn to be

Mn(q) =∑

s∈Mn

qarea(s).

Theorem 3.3.1.

Rn(q, 1) = Mn(q)

Proof. Notice that there is a simple bijection betweenMn andPn. Given a major sequence

s = (s1, · · · , sn), it corresponds to its “complement”(n + 1 − s1, · · · , n + 1 − sn) =

(f1, · · · , fn), which corresponds to a standard parking functionP as defined in [Sta99],

which corresponds to a parking functionP′of ours as defined in [HL], and which further

corresponds to its reflection about the NW-SE diagonal.

Next it suffices to show that ifz = (z1, · · · , zn) is the non-decreasing rearrangement of

s, then we havearea(z) = area(P ′). In fact,

65

# of squares counted in theith column ofP ′

= # of squares counted in theith row of P

= (n− i)− (theith biggestfj −1)

= (n− i)− (n + 1− zi − 1)

= zi − i.

So we are done.

To define the Tutte polynomial, denoted byTG(x, y), we make the convention that

G is a connected graph (actuallyTG(x, y) is just the product of the Tutte polynomials of

components ofG, if G is disconnected). Recursively, the Tutte polynomial could be defined

by the following set of rules [Ste02]:

1. If G has no edges, thenTG(x, y) = 1.

2. If e is an edge ofG that is neither a loop nor an isthmus, then

TG(x, y) = TG′e(x, y) + TG′′e

(x, y),

whereG′e is the graphG with the edgee deletedandG

′′e is the graphG with the edge

e contracted.

3. If e is an isthmus, thenTG(x, y) = xTG′e(x, y).

4. If e is a loop, thenTG(x, y) = yTG′′e(x, y).

From a different prospective, which is “graph theoretical” as well but apparently more

“combinatorial statistical”),TG(x, y) is also defined in terms of two graph theory statistics:

the external activity and internal activity [GS96]. Now suppose we are givenG and the

66

lexicographic ordering of its edges, i.e., edgeij with i < j is smaller than edgekl with

k < l iff (i < k) or (i = k, j < l). Consider a spanning treeT of G. An edgee ∈ G − T

is externally activeif it is the largest edge in the uniquecyclecontained inT ∪ e. We let

EAG(T ) = set of external active edges ofT

and

eaG(T ) = |EAG(T )|.

Dually, an edgee ∈ T is internally activeif it is the largest edge in the uniquecocycle

contained in(G− T ) ∪ e. We let

IAG(T ) = set of internal active edges ofT

and

iaG(T ) = |IAG(T )|.

Finally, the Tutte polynomial is defined as [Tut54]

TG(x, y) =∑T⊆G

xiaG(T )yeaG(T ).

Beissinger and Peled [BP97] proved bijectively that

Mn(q) = TKn+1(1, q),

whereKn+1 is the completed graph on{0, 1, · · · , n}. Henceforth, together with Theorem

3.3.1, we get the following corollary.

67

Corollary 3.3.2.

Rn(q, 1) = TKn+1(1, q).

Remark3.3.1. We remark thatTKn+1(q, t) = Rn(q, t) is false in general. In fact, the Tutte

polynomial is not even symmetric. For example,

TK5(q, 1) = q4 + 6q3 + 21q2 + 46q + 51

while

TK5(1, q) = q6 + 4q5 + 10q4 + 20q3 + 30q2 + 36q + 24.

Yet another statistic is thetree inversion enumerator. If T is a tree on{0, 1, · · · , n}, an

inversionof T is a pairi, j such that1 ≤ i < j ≤ n andj lies on the path from 0 toi in T .

We denote byinv(T ) the number of inversions ofT . The inversion enumerator for trees

on{0, 1, · · · , n} is defined as

In(q) =∑

T

qinv(T ),

where the sum is over all trees in{0, 1, · · · , n}.Bjorner [Bjo92] discovered that

In(q) = TKn+1(1, q),

by considering the generating function of the number of connected spanning subgraphs

with k edges of a complete graph and using his results in matroids as well as a result of

Gessel and Wang [GW79]. A bijective proof was given by Beissinger [Bei82].

Kreweras [Kre80] proved that

In(q) = Mn(q),

68

by a bijective method and also by proving that they both satisfy the same recurrence rela-

tion:

Statn(q) =n∑

i=1

(n− 1

i− 1

)[i] Stati−1(q) Statn−i(q).

Recently, Loehr proved in his dissertation [Loe03] that

Rn(q, 1) = Rn(1, q),

by a nontrivial combinatorial argument.

Therefore we have the following conclusion.

Corollary 3.3.3. The 5 combinatorial statistics are all equal, i.e.

Rn(q, 1) = Rn(1, q) = Mn(q) = TKn+1(1, q) = In(q),

and they all satisfy the following recurrence:

Stat1(q) = 1,

Statn(q) =n∑

i=1

(n− 1

i− 1

)[i] Stati−1(q) Statn−i(q).

69

Chapter 4

The Limit Case of the (q, t)-Schroder

Theorem

4.1 Background and Basic Properties

Throughout this chapter we use the standard notation:

(a; q)k := (1− a)(1− qa) · · · (1− qk−1a)

is the “q-rising factorial” and[nk

]=

[nk

]q

is theq-binomial coefficient.

A closed-form expression of the (q, t)-Schroder polynomial, which doesn’t reference

the bounce or area statistic, is obtained in [EHKK03]:

Theorem 4.1.1.For all n, d ∈ N,

Sn+d,d(q, t) =n∑

k=1

∑a1+···+ak=n, ai>0b0+···+bk=d, bi≥0

q∑k

i=1 (ai2 )t

∑k−1i=0 (k−i)bi+

∑k−1i=1 (k−i)ai

[b0 + a1 − 1

b0

][bk + ak

bk

k−1∏i=1

[bi + ai + ai+1 − 1

bi, ai, ai+1 − 1

].

70

Recall that in [Hag04], Haglund proved the (q, t)-Schroder Theorem. A corollary is the

limit case:

Theorem 4.1.2.[Hag04]For n ∈ N,

limd→∞

Sn+d,d(q, t) = [zn]∏i,j≥0

(1 + qitjz).

On the other hand, if we letd go to infinity, then the upper bound restriction ofb0, . . . , bk

on the right hand side of Theorem 4.1.1 is removed. So,

limd→∞

Sn+d,d(q, t) = limd→∞

n∑

k=1

∑a1+···+ak=n, ai>0b0+···+bk=d, bi≥0

q∑k

i=1 (ai2 )t

∑k−1i=0 (k−i)bi+

∑k−1i=1 (k−i)ai

[b0 + a1 − 1

b0

][bk + ak

bk

k−1∏i=1

[bi + ai + ai+1 − 1

bi, ai, ai+1 − 1

].

=n∑

k=1

∑a1+···+ak=n

ai>0

q∑k

i=1 (ai2 ) t

∑k−1i=1 (k−i)ai

∞∑

b0=0

tkb0

[b0 + a1 − 1

b0

]

limbk→∞

[bk + ak

bk

k−1∏i=1

[ai + ai+1 − 1

ai

] ∞∑

bi=0

t(k−i)bi

[bi + ai + ai+1 − 1

ai + ai+1 − 1

].

For fixeda1, . . . , ak, using theq-binomial theorem, we have

∞∑

b0=0

tkb0

[b0 + a1 − 1

b0

]=

1

(tk; q)a1

∞∑

bi=0

t(k−i)bi

[bi + ai + ai+1 − 1

ai + ai+1 − 1

]=

1

(tk−i; q)ai+ai+1

.

71

Furthermore, forq < 1,

limbk→∞

[bk + ak

bk

]= lim

bk→∞(1− qbk+1) · · · (1− qbk+ak)

(1− q) · · · (1− qak)

=1

(q; q)ak

.

Hence, we have derived an identity from the limit case of (q, t)-Schroder theorem.

Theorem 4.1.3.For n ∈ N,

n∑

k=1

∑a1+···+ak=n

ai>0

q∑k

i=1 (ai2 ) t

∑k−1i=1 (k−i)ai

1

(tk; q)a1(q; q)ak

×k−1∏i=1

[ai + ai+1 − 1

ai

]1

(tk−i; q)ai+ai+1

= [zn]∏i,j≥0

(1 + qitjz) .

However, because Haglund’s proof makes heavy use of symmetric function identities

and plethystic machinery, the combinatorics behind it is not understood. Therefore it is

worthwhile to study it combinatorially. This chapter will be devoted to obtaining some

partial results and develop related theorems by looking at some special cases.

The right hand side of Theorem 4.1.3 has been studied in connection with bipartite

partitions. Wright [Wri61] conjectured and Gordon [Gor63] proved that the right hand

side times(q; q)n(t; t)n is a polynomial of degree12n(n − 1) in each variable. In fact, it is

convenient to multiply both sides of Theorem 4.1.3 by(q; q)n(t; t)n, so we will do that and

thereby get the following modified version of Theorem 4.1.3.

72

Theorem 4.1.4.For n ∈ N,

n∑

k=1

∑a1+···+ak=n

ai>0

q∑k

i=1 (ai2 ) t

∑k−1i=1 (k−i)ai

1

(tk; q)a1(q; q)ak

×k−1∏i=1

[ai + ai+1 − 1

ai

]1

(tk−i; q)ai+ai+1

× (q; q)n(t; t)n

= [zn]∏i,j≥0

(1 + qitjz) × (q; q)n(t; t)n

Furthermore, it could be shown that the right hand side of Theorem 4.1.4 is equal to∑

σ∈Snqmaj(σ) t(

n2)−maj(σ−1). In fact, a result in [Sta99, page 385] establishes

∑σ∈Sn

qmaj(σ) tmaj(σ−1) = [zn]∏i,j≥0

1

1− qitjz× (q; q)n(t; t)n. (4.1.1)

Wright [Wri61, page 884] showed that

[zn]∏i,j≥0

1

1− qi(1t)jz

= (−1)ntn[zn]∏i,j≥0

(1 + qitjz). (4.1.2)

Combing (4.1.1) and (4.1.2) together, we have

∑σ∈Sn

qmaj(σ) t(n2)−maj(σ−1)

= t(n2)(q; q)n(

1

t;1

t)n[zn]

∏i,j≥0

1

1− qi(1t)jz

= (q; q)n(t; t)n(−1)n

tn[zn]

∏i,j≥0

1

1− qi(1t)jz

= (q; q)n(t; t)n[zn]∏i,j≥0

(1 + qitjz).

Therefore we have another modified version of Theorem 4.1.3.

73

Theorem 4.1.5.For n ∈ N,

n∑

k=1

∑a1+···+ak=n

ai>0

q∑k

i=1 (ai2 ) t

∑k−1i=1 (k−i)ai

1

(tk; q)a1(q; q)ak

×k−1∏i=1

[ai + ai+1 − 1

ai

]1

(tk−i; q)ai+ai+1

× (q; q)n(t; t)n

=∑σ∈Sn

qmaj(σ) t(n2)−maj(σ−1).

In the rest of this section we analyze some basic properties of Theorem 4.1.4. Let

Fn(q, t) denote the left hand side of Theorem 4.1.4 andGn(q, t) denote the right hand side.

First let’s review some nice facts. Carlitz [Car56] proved the following properties ofGn:

Gn(q, t) = (qt)(n2)Gn(

1

q,1

t), (4.1.3)

Gn(q, 0) = q(n2), (4.1.4)

Gn(q, 1) =1− qn

1− qGn−1(q, 1), (4.1.5)

Gn(q, q) = qn(n−1)Gn(1

q,1

q), (4.1.6)

Gn(1, 1) =n. (4.1.7)

74

We prove that all of the above hold in exactly the same way forFn(q, t). Namely,

Fn(q, t) = (qt)(n2)Fn(

1

q,1

t), (4.1.8)

Fn(q, 0) = q(n2), (4.1.9)

Fn(q, 1) =1− qn

1− qFn−1(q, 1), (4.1.10)

Fn(q, q) = qn(n−1)Fn(1

q,1

q), (4.1.11)

Fn(1, 1) =n. (4.1.12)

Note that ( 4.1.8) implies ( 4.1.11) and ( 4.1.10) implies ( 4.1.12). So we only need to prove

( 4.1.8), ( 4.1.9) and ( 4.1.10).

Lemma 4.1.6.

Fn(q, t) = (qt)(n2)Fn(

1

q,1

t).

Proof. Let fn(q, t) = Fn(q, t)/(q; q)n(t; t)n. Since

(q; q)n(t; t)n

(1q; 1

q)n(1

t; 1

t)n

= qntn(qt)(n2),

it suffices to showfn(1

q, 1

t)

fn(q, t)= qntn.

For any givenk, 1 ≤ k ≤ n, anda1 + · · ·+ ak = n with ai > 0, define

b(q, t) = q∑k

i=1 (ai2 ) t

∑k−1i=1 (k−i)ai ,

r0(q, t) =1

(tk; q)a1(q; q)ak

,

ri(q, t) =

[ai + ai+1 − 1

ai

]1

(tk−i; q)ai+ai+1

, i = 1 . . . k − 1.

75

Now it will be enough to prove

k−1∏i=0

ri(1q, 1

t)

ri(q, t)=

b(q, t)

b(1q, 1

t)qntn.

In fact, for1 ≤ i ≤ k − 1,

[ai+ai+1−1

ai

]1q[

ai+ai+1−1ai

] =q(

ai2 )+(ai+1

2 )+ai

q(ai+ai+1−1

2 ),

and1

(( 1t)k−i; 1

q)ai+ai+1

1(tk−i;q)ai+ai+1

= t(k−i)(ai+ai+1)q(ai+ai+1−1

2 ).

Hence,ri(

1q, 1

t)

ri(q, t)= q(

ai2 )q(

ai+12 )qait(k−i)ait(k−(i+1))ai+1tai+1 .

In addition,

r0(1q, 1

t)

r0(q, t)=

(tk; q)a1(q; q)ak((1

t)k; 1

q

)a1

(1q; 1

q)ak

= q(a12 )q(

ak2 )qakt(k−1)a1ta1 .

Therefore finally

k−1∏i=0

ri(1q, 1

t)

ri(q, t)= q2

∑ki=1 (ai

2 )t2∑k

i=1 (k−i)aiqntn =b(q, t)

b(1q, 1

t)qntn.

Lemma 4.1.7.

Fn(q, 0) = q(n2).

Proof. ConsiderFn(q, 0). The only non-zero term occurs when∑k−1

i=1 (k − i)ai = 0, i.e.,

whenk = 1. The rest is clear.

76

Corollary 4.1.8.

limt→∞

Fn(q, t)

Gn(q, t)= 1.

Proof. By Lemma 4.1.6 and Lemma 4.1.7,

limt→∞

Fn(q, t)

t(n2)

= q(n2) lim

t→∞Fn(

1

q,1

t) = q(

n2)Fn(

1

q, 0) = 1.

Similarly, from ( 4.1.3) and ( 4.1.4), we also have

limt→∞

Gn(q, t)

t(n2)

= 1.

Lemma 4.1.9.

Fn(q, 1) =1− qn

1− qFn−1(q, 1).

Proof. ConsiderFn(q, 1). The multiplicity of zero’s on top isn, generated by(t, t)n. In

order for the bottom to cancel these many zero’s,k must ben. It is easy to see that imme-

diately

Fn(q, 1) =(q)n

(1− q)n.

From Lemma 4.1.7 and Lemma 4.1.9,Fn(q, 0) = Gn(q, 0) andFn(q, 1) = Gn(q, 1) for

any givenq.

4.2 Investigation of Zeros

The following is a theorem by Carlitz.

77

Theorem 4.2.1. [Car56]If ξ is a primitivegth root of unity, whereg|n, then

Gn(q, ξ) =(−1)n−n

g (q)n

(1− qg)ng

.

We are happy to see exactly the same hold forFn.

Theorem 4.2.2.If ξ is a primitivegth root of unity, whereg|n, then

Fn(q, ξ) =(−1)n−n

g (q)n

(1− qg)ng

.

Proof. Sinceg|n, 1− ξn = 0. Hence,

(ξ)n

(ξ)k

6= 0 ⇔ k = n.

So,

Fn(q, ξ) =ξ

∑k−1i=1 (n−i)(q)n(ξ)n

(ξn; q)1 (q; q)1 (ξn−1; q)2 · · · (ξ; q)2

n(n−1)2 (q)n

(1− ξn−1q)(1− ξn−2q) · · · (1− ξq)(1− q)

n(n−1)2 (q)n

((1− ξg−1q)(1− ξg−2q) · · · (1− q))ng

.

Now on the bottom,

(1− ξg−1q)(1− ξg−2q) · · · (1− q)

= qg(x− 1) · · · (x− ξg−1)|x= 1q

= qg(xg − 1)|x= 1q

= 1− qg.

78

On the top, letξ = 2πijg

, wheregcd(j, g) = 1 and supposen = yg, wherey ∈ N. Then,

ξn(n−1)

2 = e2πij

gn(n−1)

2 = eπijy(n−1) = (−1)jy(n−1).

Notice thatjy(n− 1)− (n− y) = jy(yg − 1)− (yg − y) = y(jyg − j − g + 1) must be

even becausej andg should have different parities. Therefore,

ξn(n−1)

2 = (−1)n−y = (−1)n−ng .

So we are done.

Remark4.2.1. Gn is obviously symmetric from its definition. For the case discussed above

we also have the symmetry forFn: Fn(q, ξ) = Fn(ξ, q). To see that, noticeFn(ξ, q) only

has one non-zero term, i.e. the one whereak = n in order to cancel1 − ξn which is

generated by(ξ; ξ)n on the top, and thusk = 1. The rest is similar to the proof of Theorem

4.2.2. In the end we have the same expression

Fn(ξ, q) =(−1)n−n

g (q)n

(1− qg)ng

,

so thatFn(q, ξ) = Fn(ξ, q).

For any fixedq, let dq(t) := Fn(q, t)−Gn(q, t). Theorem 4.2.2 shows thatdq(t) hasn

zero’s, since everynth root of unity is a primitivegth root, for someg|n. By ( 4.1.4) and

Lemma 4.1.7,t = 0 is another zero. So altogether we have foundn + 1 zeros ofdq(t),

whereq is any given real number, which is a partial support forFn(q, t) andGn(q, t) to be

equal.

Next we present a theorem aboutFn, parallel to a result of Gordon aboutGn.

79

Theorem 4.2.3. [Gor63]Assumeω = e2πiu

h andξ = e2πiv

g , wheregcd(u, h) = gcd(v, g) =

1, g < h ≤ n andn = yh + s = xg + r (so that 1 ≤ y ≤ x), then

Gn(ω, ξ) = 0.

Theorem 4.2.4.Assumeω = e2πiu

h andξ = e2πiv

g , wheregcd(u, h) = gcd(v, g) = 1, g <

h ≤ n andn = yh + s = xg + r (so that 1 ≤ y ≤ x), then

Fn(ω, ξ) = 0.

Proof. For each selection ofk anda1, a2, · · · , ak with a1 + a2 + · · ·+ ak = n, clearly the

number of zero’s on the numerator ofFn(ω, ξ) is x + y. We only need to make sure that

the number of zero’s generated by the bottom is something smaller. Leta0 = 0, then

The bottom= (ω, ω)ak

k−1∏j=0

(1− ξk−j) · · · (1− ξk−jωaj+aj+1−1).

For any fixedj and0 ≤ p ≤ aj + aj+1 − 1,

1− ξk−jωp = 0 ⇔ v

g(k − j) +

u

hp ∈ N.

Note that ifpi, i = 1, 2, makes

v

g(k − j) +

u

hpi ∈ N,

thenu

h(p1 − p2) ∈ N⇔ h|(p1 − p2),

which says that the difference between any two “good”p’s is at leasth.

80

Next, letf = gcd(g, h), g = g1f, h = h1f and vg(k − j) + u

hp = m ∈ N. Then,

hv(k − j) + gup = mgh ⇒ h1v(k − j) + g1up = mg1h1f

and therefore,

gcd(g1u, h1) = 1 ⇒ h1|p (4.2.1)

gcd(h1v, g1) = 1 ⇒ g1|(k − j). (4.2.2)

So, only whenj = k − g1, j = k − 2g1, · · · , there could possibly be zero’s on the bottom.

• Case 1:g1 = 1. Theng = f | h ⇒ h ≥ 2g. Therefore,

The number of zero’s on the bottom

≤ bak

hc+

k−1∑j=0

baj + aj+1

hc

≤ b(a0 + a1) + · · ·+ (ak−1 + ak) + ak

hc

= b2nhc

≤ bngc

= x

< x + y.

81

• Case 2:g1 ≥ 2. Then,

The number of zero’s on the bottom

≤ bak

hc+ bak−g1 + ak−g1+1

hc+ · · ·+ b

ak−b kg1cg1

+ ak−b kg1cg1+1

hc

≤ bak + ak−1 + · · ·+ a1 + a0

hc

= bnhc

= y

< x + y.

So, in either case,Fn(ω, ξ) = 0.

By Theorem 4.2.4 and the following result, we also have symmetry ofFn(q, t) for such

values ofq andt.

Theorem 4.2.5.Assumeω = e2πiu

h andξ = e2πiv

g , wheregcd(u, h) = gcd(v, g) = 1, g <

h ≤ n andn = yh + s = xg + r (so that 1 ≤ y ≤ x), then

Fn(ξ, ω) = 0

Proof. The proof is similar to but simpler than the proof of Theorem 4.2.4. If we use

the same notation, then (4.2.1) and (4.2.2) in the proof of Theorem 4.2.4 will become

g1|(p1−p2) andh1|(k−j), respectively. Moreover, only whenj = k−g1, j = k−2g1, · · · ,

82

could there be zero’s on the bottom. Noteh > g ⇒ h1 > g1 ≥ 1 ⇒ h1 ≥ 2. So nicely,

The number of zero’s on the bottom

≤ bak

gc+ bak−h1 + ak−h1+1

gc+ bak−2h1 + ak−2h1+1

gc+ · · ·+ b

ak−b kh1ch1

+ ak−b kh1ch1+1

gc

≤ bak + (ak−h1+ak−h1+1

) + (ak−2h1 + ak−2h1+1) + · · ·+ (ak−b kh1ch1

+ ak−b kh1ch1+1)

gc

≤ bak + ak−1 + · · ·+ a1 + a0

gc

= bngc

= x

< x + y.

4.3 Whenq = t

In this section we refer to the version of Theorem 4.1.5. That is, we use the sameFn(q, t)

as previously but let

Gn(q, t) :=∑σ∈Sn

qmaj(σ) t(n2)−maj(σ−1).

Furthermore, for1 ≤ k ≤ n, define

Fn,k(q, t) :=∑

a1+···+ak=nai>0

q∑k

i=1 (ai2 ) t

∑k−1i=1 (k−i)ai

1

(tk; q)a1(q; q)ak

×k−1∏i=1

[ai + ai+1 − 1

ai

]1

(tk−i; q)ai+ai+1

× (q; q)n(t; t)n,

83

and

Gn,k(q, t) :=∑

σ∈Sn,k

qmaj(σ) t(n2)−maj(σ−1),

whereSn,k = {σ ∈ Sn : isn(σ) = k} and isn(σ) denotes the length of the longest in-

creasing subsequence ofσ ∈ Sn. By definition,Fn(q, t) =∑n

i=1 Fn,i(q, t) andGn(q, t) =

∑ni=1 Gn,i(q, t). For brevity, we also let

fn,k(q, t) =k∑

i=1

Fn,i(q, t)

and

gn,k(q, t) =k∑

i=1

Gn,i(q, t)

=k∑

i=1

∑σ∈Sn,i

qmaj(σ) t(n2)−maj(σ−1).

Here we investigate the interesting case whereq = t. One property of the(q, q)-case

comes from Lemma 4.1.9 and ( 4.1.6):

Fn(q, q) = qn(n−1)Fn(1

q,1

q)

and

Gn(q, q) = qn(n−1)Gn(1

q,1

q).

This property can be refined to the following.

Lemma 4.3.1.

Fn,k(q, q) = qn(n−1)Fn,k(1

q,1

q),

84

and

Gn,k(q, q) = qn(n−1)Gn,k(1

q,1

q)

Proof. The first equality is immediate from the proof of Lemma 4.1.6. Observe that the

second equality will be true as soon as we prove thatSn,k is closed under the inverse

operation of symmetric groups. In fact, it is easy to see that ifisn(σ) = k thenisn(σ−1) ≥k, and vice versa, so it is clear.

The main conjecture regarding the(q, q) case is the following.

Conjecture 4.3.1.For 1 ≤ k ≤ n,

Fn,k(q, q) = Gn,k(q, q).

Equivalently, for1 ≤ k ≤ n,

fn,k(q, q) = gn,k(q, q).

Conjecture 4.3.1 has been verified by using Maple for small values ofn andk, but the

proof appears to be quite hard even fork = 2. It is not even clear thatFn,k(q, q) or fn,k(q, q)

has to be polynomial for generaln andk.

Example4.3.1. Whenk = 1,

fn,1(q, q) = Fn,1(q, q) = q(n2) (q; q)n(q; q)n

(q; q)n(q; q)n

= q(n2),

and

gn,1(q, q) = Gn,1(q, q) = qmaj(n···21)+(n2)−maj((n···21)−1) = q(

n2).

In [FH85], cn(λ; q) is introduced as a kind ofq-Catalan number with parameterλ by

85

means of the expansion formula

z =∞∑

n=1

cn(λ; q)zn

q(n2)(−q−nz)n(−qλz)n

.

Furlinger and Hofbauer showed that

cn(λ; q) =1

[n]

n∑

k=1

qk2+λk

[n

k

][n

k − 1

],

and in particular

cn(1; q) =1

[n + 1]

[2n

n

].

Lemma 4.3.2.Whenk = 2,

fn,2(q, q) = q(n2)cn(1− n, q).

Proof.

Fn,2(q, q) =n−1∑a1=1

q(a12 )+(n−a1

2 )+a1(q; q)2

n

(q2; q)a1(q; q)n−a1

[n− 1

a1

]1

(q; q)n

=n−1∑a1=1

q(a12 )+(n−a1

2 )+a1(q; q)n

(q; q)a1(q; q)n−a1

1

[a1 + 1]

[n− 1

a1

]

=n−1∑a1=1

qa21+a1(1−n)+(n

2)[

n

a1

][n

a1 + 1

]1

[n]

= q(n2) 1

[n]

(n∑

k=1

qk2+(1−n)k

[n

k

][n

k − 1

]− [n]− 0

)

= q(n2)(cn(1− n; q)− 1).

So,

fn,2(q, q) = Fn,1(q, q) + Fn,2(q, q) = q(n2)cn(1− n; q).

86

Remark4.3.1. In the summation above,

1

[n]

[n

k

][n

k − 1

]

is theq-analog of theRunyon numbersrn,k = 1n

(nk

)(n

k−1

)[Rio68, page 17], which counts

the number of Dyck pathsΠ ∈ Dn with k − 1 “valleys” (consecutive EN pair steps) or

equivalentlyk “peaks” (consecutive NE pair steps), i.e.des(w(Π)) = k− 1. Furlinger and

Hofbauer also obtained the following combinatorial interpretation ofcn(λ, q):

cn(λ, q) =∑

w∈CWn

qmaj(w)+(λ−1)des(w).

Corollary 4.3.3. fn,2(q, q) is a polynomial with non negative coefficients.

Proof. For brevity letdes(w) = d. Then,

(n

2

)+ maj(w) + ((1− n)− 1)des(w)

≥(

n

2

)+ (2 + 4 + · · ·+ 2d)− nd

=

(n

2

)+ d2 + d− nd

= (d− n− 1

2)2 +

n2 − 1

4

n2−14

, if n is odd;

n2

4, if n is even.

.

Sofn,2(q, q) is a polynomial of positive coefficients, following Lemma 4.3.2 and Remark

4.3.1.

87

Computationally, we found that

fn,2(q, q) = gn,2(q, q)

for n up to 8. SinceSn,1∪Sn,2 = Sn(123), whereSn(123) stands for the set of123-avoiding

permutations inSn, we can rewrite the conjecturefn,2(q, q) = gn,2(q, q) in the following

way.

Conjecture 4.3.2.

∑w∈CWn

qmaj(w)−ndes(w) =∑

σ∈Sn(123)

qmaj(σ)−maj(σ−1).

Although the conjecture appears to be formidable, it is still interesting to consider some

related situation. If restricted to the set of “involutions” on both sides, then as a partial

result, we prove a similar identity.

Definition 4.3.1. The inverseof a Catalan wordw ∈ CWn is defined to be

w−1 = r(w),

wherer denotes the reverse operation and− denotes the complement operation that ex-

changes 0 and 1. We sayw is aninvolution if and only if w = w−1.

Example4.3.2. Whenn=3,

(000111)−1 = 000111,

(001011)−1 = 001011,

(001101)−1 = 010011,

(010011)−1 = 001101,

88

(010101)−1 = 010101.

So the involution subset ofCW3 consists of 000111, 001011 and 010101.

It is easy to see thatw−1 ∈ CWn if and only if w ∈ CWn, so the inverse operation is

closed onCWn. Geometrically, givenw, we may obtainw−1 by finding the Dyck pathΠ

thatw corresponds to under the natural map, reflectingΠ over the NW-SE main diagonal to

obtain a new Dyck pathP , and then taking the Catalan word that corresponds toP . For this

reason, we also define theinverseof a Dyck pathΠ by Π−1 = P through this reflection,

and sayΠ is an involution if and only if Π = Π−1. Figure 4.1 illustrates a Dyck path

together with its inverse. Note the reflection over the dashed NW-SE diagonal.

Figure 4.1: A Dyck path (on the left) and its inverse (on the right).

Theorem 4.3.4.

∑w∈CWn:

w is an involution

qmaj(w)−ndes(w) =∑

σ∈Sn(123):σ is an involution

qmaj(σ)−maj(σ−1).

89

Proof. Forw ∈ CWn,

maj(w−1) =∑

i∈des(w−1)

i

=∑

2n−i∈des(w)

i

=∑

j∈des(w)

2n− j

= 2ndes(w)−maj(w).

Clearlydes(w−1) = des(w). Thus

maj(w−1)− ndes(w−1) = (2ndes(w)−maj(w))− ndes(w)

= ndes(w)−maj(w)

= −(maj(w)− ndes(w)).

If w is an involution, this impliesmaj(w) − ndes(w) = 0. So it suffices to prove that the

number of involutions inCWn, which is also the number of Dyck path involutions inDn,

is equal to the number of involutions inSn(123).

On one hand, Simion and Schmidt [SS85] showed that the number of involutions in

Sn(123) is equal to (n

bn2c)

.

On the other hand, we can decompose the set of Dyck path involutions according to the

place where they reach the NW-SE diagonal about which they are symmetric. Fork ≥ m,

let S(k, m) be the set of all 01 wordsw = w1w2 · · ·wk+m consisting ofk 0s andm 1s;

let S+(k, m) denote the subset ofS(k, m) consisting of those words, such that no ini-

tial segment contains more 1s than 0s; and letS−(k, m) = S(k,m) \ S+(k,m). Obvi-

90

ously |S(k, m)| = (k+m

k

). Furlinger and Hofbauer [FH85] constructed a bijection between

S−(k, m) andS(k + 1,m− 1). Furthermore, since every Dyck path involution is uniquely

decided by its beginning half (the part from (0, 0) to where it reaches the NW-SE diagonal),

there is an apparent bijection betweenS+(k, n−k) and the subset of Dyck path involutions

in Dn that reach the NW-SE diagonal at(n − k, k). Therefore, the number of Dyck path

involutions inDn is equal to

n∑

k=dn2e|S+(k, n− k)| =

n∑

k=dn2e|S(k, n− k)| − |S−(k, n− k)|

=n∑

k=dn2e

(n

k

)− |S(k + 1, n− k − 1)|

=n∑

k=dn2e

(n

k

)−

(n

k + 1

)

=

(n

dn2e)

.

So we are done.

4.4 Agreement oftk Coefficients

Theorem 4.1.5 would be proved if the coefficients oftk agreed on both sides. It’s trivial

to see that the constant terms, i.e. the coefficients oft0 on the two sides, are both equal to

q(n2).

We will show that this is also true fork = 1, 2, 3 and look for a pattern to generalize.

Lemma 4.4.1.The coefficients oft1 on the two sides of Theorem 4.1.5 are both equal to

q(n−1

2 )([

n

1

]− q(n−1)

).

91

Proof. There are two cases on theleft hand side:

• k = 1 and soa1 = n.

LHS = q(n2) (t; t)n(q; q)n

(t; q)n(q; q)n

.

So we only need to find the coefficient oft in

q(n2)

((1 + tq)(1 + tq2) · · · (1 + tqn−1)

)

which isq(n2)q[n− 1].

• k = 2, soa1 must be1 anda2 = n− 1.

LHS = q(n−1

2 )t(t; t)n(q; q)n

(t2; q)1(q; q)n

[n− 1

1

]1

(t; q)n

.

So the coefficient oft is

q(n−1

2 )(1− qn)

[n− 1

1

].

Summing up the two cases we getq(n−1

2 ) ([n1

] − q(n−1)).

Right hand side:

[t1] RHS=∑

(n2)−maj(σ−1) =1

qmaj(σ)

=∑

(n2)−maj(σ)=1

qmaj(σ−1).

Note that (n

2

)−maj(σ) = 1 ⇔ σ1 < · · · < σn−1 > σn.

92

Soσ2 = n. Let σ1 = i, 1 ≤ i ≤ n− 1, theni will be the only non-descent ofσ−1. Hence

(n2)−maj(σ)=1

qmaj(σ−1) =n−1∑i=1

q(n2)−i

= q(n−1

2 )[n− 1]

= q(n−1

2 )([

n

1

]− q(n−1)

).

Theorem 4.4.2.The coefficients oft2 on the two sides of Theorem 4.1.5 are both equal to

q(n−2

2 )+1

([n

2

]− q2(n−2)

).

Proof. First consider the LHS. Notice that∑k−1

i=1 (k − i)ai will be bigger than2 if k is

bigger than2. Thus in order to generatet2 we only have the following three situations:

• k = 1 anda1 = n

LHS is easily reduced toq(n2)(t; t)n/(t; q)n. So we only need to get the coefficient of

t2 in

q(n2)(1− t2)(1 + tq + (tq)2) · · · (1 + tqn−1 + (tqn−1)2),

which is clearlyq(n2)(−1 + q2[n− 1]q2 + q3

[n−1

2

]).

• k = 2, a1 = 1 anda2 = n− 1

In this case,

93

LHS = q(n−1

2 ) t(t; t)n (q; q)n

(t2; q)1 (q; q)n−1

[n− 1

1

]1

(t; q)n

(t; q)n .

So it suffices to find the coefficient oft in

q(n−1

2 )+1 (1− qn) [n− 1] (1 + tq) · · · (1 + tqn−1).

That is clearlyq(n−1

2 )+1(1− qn) [n− 1]2.

• k = 2, a1 = 2 anda2 = n− 2

Similarly, the coefficient oft2 in

q(n−2

2 )+1 t2(t; t)n (q; q)n

(t2; q)2 (q; q)n−2

[n− 1

2

]1

(t; q)n

(t; q)n

is q(n−2

2 )+1(1− qn)(1− qn−1)[n−1

2

]).

Summing up the contributions from the three cases, we find the coefficient oft2 on the LHS

is equal to

q(n−2

2 )

(1− q) (1− q2)(q − qn − qn+1 − q2n−3 + q2n−2 + q2n−1)

= q(n−2

2 )+1

([n

2

]− q2(n−2)

).

94

Next consider the RHS. The coefficient oft2 on the RHS is equal to

(n2)−maj(σ−1) =2

qmaj(σ) =∑

(n2)−maj(σ)=2

qmaj(σ−1)

Observe that

(n

2

)−maj(σ) = 2

⇔ Des(σ) = {1, 2, · · · , n− 1}\{2}

⇔ σ1 < σ2 < · · · < σn−2 > σn−1 > σn.

There are three possible cases according to the value ofσ3. We call the subsets ofSn in

these casesΠ1, Π2 andΠ3 respectively.

• σ3 = n−1. Then we must haveσ1 = n. Letσ2 = k, where1 ≤ k ≤ n−2. It is easy

to see that the only non-descent ofσ−1 is k whereσ−1(k) = 2 andσ−1(k + 1) =

n− k + 1. Therefore,maj(σ−1) =(

n2

)− k and so we have,

∑σ∈Π1

qmaj(σ−1) =n−2∑

k=1

q(n2)−k

= q(n−1

2 ) (q + . . . + qn−2)

= q(n−1

2 )+1 [n− 2].

• σ3 = n andσ2 = k, σ1 = j, with 1 ≤ k < j − 1 ≤ n. Note that the case of

k = j − 1 will be dealt with differently next. Now,

σ =

1 2 3 · · · n− j + 2 n− j + 3 · · · n− k + 1 n− k + 2 · · ·j k n · · · j + 1 j − 1 · · · k + 1 k − 1 · · ·

95

implies that

σ−1 =

· · · k − 1 k k + 1 · · · j − 1 j j + 1 · · · n

· · · n− k + 2 2 n− k + 1 · · · n− j + 3 1 n− j + 2 · · · 3

.

So,

∑σ∈Π2

qmaj(σ−1) =∑

1≤k<j−1≤n−2

q(n2)−k−j

= q(n−2

2 )∑

1≤k<j−1≤n−2

q(n−2−k)+(n−1−j)

= q(n−2

2 )∑

0≤r<s≤n+s

qr+s

= q(n−2

2 )+1

[n− 2

2

].

• σ3 = n, σ2 = k, andσ1 = k + 1, with 1 ≤ k ≤ n− 2. In this case,

σ−1 =

k + 1 k n · · · k + 2 k − 1 · · ·1 2 3 · · · n− k + 1 n− k + 2 · · ·

=

· · · k − 1 k k + 1 k + 2 · · · n

· · · n− k + 2 2 1 n− k + 1 · · · 3

.

96

So,

∑σ∈Π3

qmaj(σ−1) =∑

1≤k≤n−2

q(n2)−(k+1)

= q(n−1

2 )∑

1≤k≤n−2

qn−1−(k+1)

= q(n−1

2 )∑

0≤r≤n−3

qr

= q(n−1

2 )[n− 2].

Adding together, the coefficient oft2 on the RHS is also

q(n−2

2 )+1

([n

2

]− q2(n−2)

).

Theorem 4.4.3.The coefficients oft3 on the two sides of Theorem 4.1.5 are both equal to

q(n−3

2 )+3

[n− 3

3

]+ q(

n−22 )+1 [3]

[n− 3

2

]+ q(

n−12 ) [3]

[n− 3

1

]+ q(

n−22 )

[n− 1

2

].

Proof. There are5 possible cases on theleft hand side.

97

• k = 1 anda1 = n

[t3]{q(n2)(t; t)n/(t; q)n}

= [t3]{q(n2)(1− t2)(1− t3)(1 + tq + (tq)2 + (tq)3)

· · · (1 + tqn−1 + (tqn−1)2 + (tqn−1)3}

= q(n2)(q6

[n− 1

3

]− q[n− 1] +

n−1∑i=1

q2i(q[n− 1]− qi)

− 1 + q3 + · · ·+ q3(n−1))

= q(n2)(q6

[n− 1

3

]+ q[n− 1](q2[n− 1]q2 − 1)− 1).

• k = 2, a1 = 1 anda2 = n− 1

[t3]{q(n−12 )t

(t; t)n(q; q)n

(t2; q)1(q; q)n−1

[n− 1

1

]1

(t; q)n

}

= [t2]{q(n−12 )[n− 1](1− qn)(1 + tq + t2q2) · · · (1 + tqn−1 + (tqn−1)2)}

= q(n−1

2 )[n− 1](1− qn)(q2[n− 1]q2 + q3

[n− 1

2

]).

• k = 2, a1 = 2 anda2 = n− 2

[t3]{q(n−22 )+1t2

(t; t)n(q; q)n

(t2; q)2(q; q)n−2

[n− 1

2

]1

(t; q)n

}

= [t1]{q(n−22 )+1(1− qn)(1− qn−1)

[n− 1

2

](t; t)n

(t; q)n

}

= [t1]{q(n−22 )+1(1− qn)(1− qn−1)

[n− 1

2

](1 + tq) · · · (1 + tqn−1)}

= q(n−2

2 )+1(1− qn)(1− qn−1)

[n− 1

2

]q[n− 1].

98

• k = 2, a1 = 3 anda2 = n− 3

[t3]{q(n−32 )+3t3

(t; t)n(q; q)n

(t2; q)3(q; q)n−3

[n− 1

3

]1

(t; q)n

}

= q(n−3

2 )+3(1− qn)(1− qn−1)(1− qn−2)

[n− 1

3

].

• k = 3, a1 = 1, a2 = 1 anda3 = n− 2

[t3]{q(n−22 )t3

(t; t)n(q; q)n

(t3; q)1(q; q)n−2

[1

1

]1

(t2; q)2

[n− 2

1

]1

(t; q)n−1

}

= q(n−2

2 )(1− qn)(1− qn−1)[n− 2].

It is not hard to verify that the sum of the results in the 5 cases above is equal to the desired

formula.

Right hand side: there are altogether8 possible cases.

[t3] RHS=∑

(n2)−maj(σ−1)=3

qmaj(σ)

=∑

(n2)−maj(σ)=3

qmaj(σ−1).

(n2

)−maj(σ) = 3 implies either

Des(σ) = {1, 2, · · · , n− 1}\{1, 2},

or

Des(σ) = {1, 2, · · · , n− 1}\{3}.

99

Case 1Des(σ) = {1, 2, · · · , n− 1}\{1, 2}. This means

σ1 < σ2 < σ3 > · · · > σn,

which implies thatσ3 = n. Let σ1 = k andσ2 = j, where1 ≤ k < j ≤ n − 1. Then we

have

σ−1 =

1 2 3 · · · n− j + 2 · · · n− k + 1− 1 · · ·k j n · · · j + 1 · · · k + 1 · · ·

−1

=

· · · k − 1 k · · · j j + 1 · · · n

· · · n− k + 2 1 · · · 2 n− j + 2 · · · 3

.

Therefore,Des(σ−1) = {1, 2, · · · , n− 1}\{k, j}. Hence,

(n2)−maj(σ)=3

qmaj(σ−1) =∑

1≤k<j≤n−1

q(n2)−k−j

= q(n−2

2 )−1∑

1≤k<j≤n−1

q(n−1−k)+(n−1−j)

= q(n−2

2 )−1∑

0≤r<s≤n−2

qr+s

= q(n−2

2 )[n− 1

2

].

Case 2Des(σ) = {1, 2, · · · , n− 1}\{3}. This means

σ1 > · · · > σn−3 < σn−2 > σn−1 > σn.

Apparentlyσ4 ≥ n−2. We explore the three situations according toσ4 = n−2, σ4 = n−1,

100

or σ4 = n.

Subcase 2.1σ4 = n − 2. It must be true thatσ2 = n − 1 andσ1 = n. Let σ3 = k,

where1 ≤ k ≤ n− 3, then

σ−1 =

1 2 3 4 · · · n− k + 1 n− k + 2 · · ·n n− 1 k n− 2 · · · k + 1 k − 1 · · ·

−1

=

· · · k − 1 k k + 1 · · · n− 2 n− 1 n

· · · n− k + 2 3 n− k + 1 · · · 4 2 1

.

Hencemaj(σ−1) =(

n2

)− k and accordingly

(n2)−maj(σ)=3

qmaj(σ−1) =∑

1≤k≤n−3

q(n2)−k

= q(n−1

2 )∑

1≤k≤n−3

qn−1−k

= q(n−1

2 )∑

0≤r≤n−4

qr+2

= q(n−1

2 )+2[n− 3].

Subcase 2.2σ4 = n−1, thenσ3 = k. Letσ3 = k andσ2 = j, where1 ≤ k < j ≤ n−2.

σ−1 =

1 2 3 4 · · · n− j + 2 n− j + 3 · · ·n j k n− 1 · · · j + 1 j − 1 · · ·

−1

=

· · · k j − 1 j j + 1 · · · n− 1 n

· · · 3 n− j + 3 2 n− j + 2 · · · 4 1

.

101

It is a little tricky here because it makes a difference ifk = j − 1 or not. In fact,

k = j − 1 ⇒ maj(σ−1) =

(n

2

)− j,

k < j − 1 ⇒ maj(σ−1) =

(n

2

)− k − j.

Therefore,

(n2)−maj(σ)=3

qmaj(σ−1) =∑

1≤k<j−1≤n−3

q(n2)−k−j +

∑2≤j≤n−2

q(n2)−j

= q(n−2

2 )+2∑

1≤k<j−1≤n−3

q(n−2−j)+(n−3−k) + q(n−1

2 )∑

2≤j≤n−2

qn−1−j

= q(n−2

2 )+2∑

0≤r<s≤n−4

qr+s + q(n−1

2 )∑

0≤r≤n−4

qr+1

= q(n−2

2 )+3

[n− 3

2

]+ q(

n−12 )+1[n− 3].

Subcase 2.3σ4 = n. Let σ3 = k, σ2 = j, andσ1 = i, where1 ≤ k < j < i ≤ n − 1.

This is a complicated case and we will only illustrate the idea.

σ−1 =

1 2 3 4 · · · n− i + 3 · · ·i j k n · · · i + 1 · · ·

−1

=

· · · k · · · j · · · i i + 1 · · · n

· · · 4 · · · 2 · · · 1 n− i + 3 · · · 4

.

The only possible non-descents ofσ−1 are thus spotsk, j andi. There are different situa-

tions dependent on the values ofj − k andi− j.

102

• Sub-subcase 2.3.1: j − k > 1 andi− j > 1. This implies

maj(σ−1) =

(n

2

)− i− j − k.

So,∑

(n2)−maj(σ)=3

qmaj(σ−1)

=∑

1≤k<j−1<i−2≤n−3

q(n2)−i−j−k

= q(n−3

2 )∑

1≤k<j−1<i−2≤n−3

q((n−3)−(i−2))+((n−3)−(j−1))+((n−3)−k)

= q(n−3

2 )∑

0≤r<s<t≤n−4

qr+s+t

= q(n−3

2 )+3

[n− 3

3

].

• Sub-subcase 2.3.2: j − k > 1 while i− j = 1. This implies

maj(σ−1) =

(n

2

)− i− k.

So,∑

(n2)−maj(σ)=3

qmaj(σ−1) =∑

1≤k<i−2≤n−3

q(n2)−i−k

= q(n−2

2 )+1∑

1≤k<i−2≤n−3

q((n−3)−(i−2))+((n−3)−k)

= q(n−2

2 )+1∑

0≤r<s≤n−4

qr+s

= q(n−2

2 )+2

[n− 3

2

].

103

• Sub-subcase 2.3.3: i− j > 1 while j − k = 1. This implies

maj(σ−1) =

(n

2

)− i− j.

So,∑

(n2)−maj(σ)=3

qmaj(σ−1) =∑

2≤j<i−1≤n−2

q(n2)−i−j

= q(n−2

2 )∑

2≤j<i−1≤n−2

q(n−2−(i−1))+(n−2−j)

= q(n−2

2 )∑

0≤r<s≤n−4

qr+s

= q(n−2

2 )+1

[n− 3

2

].

• Sub-subcase 2.3.4: i− j = 1 andj − k = 1. This implies

maj(σ−1) =

(n

2

)− i.

So,∑

(n2)−maj(σ)=3

qmaj(σ−1) =∑

3≤i≤n−1

q(n2)−i

= q(n−1

2 )∑

3≤i≤n−1

qn−1−i

= q(n−1

2 )∑

0≤r≤n−4

qr

= q(n−1

2 )[n− 3].

Adding up the results of Case 1, Subcases 2.1, 2.2, Sub-subcases 2.3.1, 2.3.2, 2.3.3 and

2.3.4, we have the desired sum.

104

105

Chapter 5

Higher Dimensional Schroder Theory

5.1 m-Schroder Paths andm-Schroder Number

While the standard Catalan and Schroder theories both have been extensively studied, peo-

ple have only begun to investigate higher dimensional versions of the Catalan number (see

[HPW99] and [GH96]). In this chapter, we study a yet more general case, namely the higher

dimensional Schroder theory. We introduce and derive results about them-Schroder paths

and words. First let’s introduce the notions of generalized Dyck and Schroder paths.

Definition 5.1.1. An m-Dyck pathof ordern is a lattice path from(0, 0) to (mn, n) which

never goes below the main diagonal{(mi, i) : 0 ≤ i ≤ n}, with steps(0, 1) (or NORTH,

for brevity N) and(1, 0) (or EAST, for brevityE). Let Dmn denote the set of allm-Dyck

paths of ordern.

A 2-Dyck path of order6 is illustrated in Figure 5.1.

As in them = 1 case, givenΠ ∈ Dmn , we encode eachN step by a 0 and eachE step

by a 1 so as to obtain a wordw(Π) of n 0’s andmn 1’s. This clearly provides a bijection

106

Figure 5.1: A2-Dyck path inD26.

betweenDmn andCWm

n , where

CWmn = {w ∈ Mn,mn|at any initial segment ofw, the number of 0’s times

m is at least as many as the number of 1’s.}

We call this special subset of 01 words,CWmn , Catalan words of ordern and dimension

m.

It is shown in [HP91] (see also [HPW99]) that the number ofm-Dyck paths, denoted

by Cmn , is equal to

1

mn + 1

(mn + n

n

),

which we call them-Catalan number. In fact, Cigler [Cig87] proved that the number ofm-

Dyck paths withk peaks, i.e., those with exactlyk consecutive NE pairs, is the generalized

Runyon number,

Rmn,k =

1

n

(n

k

)(mn

k − 1

).

Now we turn to the more generalm-Schroder theory.

Definition 5.1.2. An m-Schroder pathof ordern is a lattice path from(0, 0) to (mn, n)

which never goes below the main diagonal{(mi, i) : 0 ≤ i ≤ n}, with steps(0, 1) (or

NORTH, for brevityN), (1, 0) (or EAST, for brevityE) and (1,1) (or Diagonal, for brevity

D). LetSmn denote the set of allm-Schroder paths of ordern.

107

Definition 5.1.3. An m-Schroder pathof ordern and withd diagonal steps is a lattice path

from (0, 0) to (mn, n) which never goes below the main diagonal{(mi, i) : 0 ≤ i ≤ n},with (0, 1) (or NORTH, for brevityN), (1, 0) (or EAST, for brevityE) and exactlyd (1,1)

(or Diagonal, for brevityD) steps. LetSmn,d denote the set of allm-Schroder paths of order

n and withd diagonal steps.

A 2-Schroder path of order6 and with4 diagonal steps is illustrated in Figure 5.2.

Figure 5.2: A2-Schroder path inS26,4.

Theorem 5.1.1.The number ofm-Schroder paths of ordern and withd diagonal steps,

denoted bySmn,d, is equal to

1

mn− d + 1

(mn + n− d

mn− d, n− d, d

).

Proof. For anm-Dyck pathΠ, let its number of peaks, or consecutive NE pairs, be de-

noted bypeak(Π). Notice that anym-Schroder path withd diagonal steps can be obtained

uniquely by choosingd of the peaks of a uniquely decidedm-Dyck pathΠ of the same

order, and changing each of the chosen consecutive NE pair steps to a Diagonal step. Con-

versely, given anm-Dyck pathΠ of ordern, choosingd of its peaks (if there ared to

choose) and changing them toD steps will give a path inSmn,d. For example, the 2-Schroder

path as illustrated in Figure 5.2 is one of(42

)= 6 paths inS2

6,4 that can be obtained from

108

the 2-Dyck path shown in Figure 5.1. Hence,

Smn,d =

∑Π∈Dm

n

(peak(Π)

d

)

=∑

k≥d

(k

d

)Rm

n,k

=∑

k≥d

(k

d

)1

n

(n

k

)(mn

k − 1

)

=

(nd

)

n

k≥d

(n− d

n− k

)(mn

k − 1

)

=

(nd

)

n

(mn + n− d

n− 1

)

=1

mn− d + 1

(mn + n− d

d, n− d,mn− d

)

Above we used the Vandermonde Convolution (see, say, [Com74, page 44]).

As a generalization of them = 1 case, we name

Smn =

n∑

d=0

1

mn− d + 1

(mn + n− d

mn− d, n− d, d

)

them-Schroder number.

5.2 q-m-Schroder Polynomials

When Bonin, Shapiro and Simion [BSS93] studiedq-analogues of the Schroder numbers,

they obtained several classical results for several single variable analogue cases. Here we

generalize some of them to them case.

Definition 5.2.1. Define them-Narayanapolynomialdmn (q) over them-Schroder paths of

109

ordern to be

dmn (q) =

∑Π∈Sm

n

qdiag(Π),

where diag(Π) is the number ofD steps in the pathΠ.

Theorem 5.2.1.dmn (q) hasq = −1 as a root.

Proof. We use the idea of [BSS93]. The statement is equivalent to say that there are as

manym-Schroder paths of ordern with an even number ofD steps as there are with an

odd number ofD steps. For anyΠ ∈ Smn , there must be some first occurrence of either a

consecutive NE pair of steps, or aD step. According to which occurs first, either replace

the consecutive NE pair by aD, or replace theD with a consecutive NE pair. Notice that

this presents a bijection between the two sets of objects we wish to show have the same

cardinality.

In [FH85], there is a refinedq-analogue identity,

k≥1

∑w∈CWn,k

qmajw =∑

k≥1

1

[n]

[n

k

][n

k − 1

]qk(k−1) =

1

[n + 1]

[2n

n

], (5.2.1)

whereCWn,k is the set of Catalan words consisting ofn 0’s,n 1’s, withk ascents (i.e.k−1

descents or the corresponding Dyck path hask peaks). As for them-version, Cigler proved

there are exactly1

n

(n

k

)(mn

k − 1

)

m-Dyck paths withk peaks [Cig87]. In order to generalize the results of [FH85], we prove

the followingq-identity.

Theorem 5.2.2.

k≥d

[k

d

]1

[n]

[n

k

][mn

k − 1

]q(k−d)(k−1) =

1

[mn− d + 1]

[mn + n− d

d, n− d,mn− d

].

110

Before we proceed to the proof of Theorem 5.2.2, we cite theq-Vandermonde Convo-

lution, which may be obtained as a corollary of theq-binomial theorem.

Lemma 5.2.3. [Hagon]Theq-Vandermonde Convolution.

h∑j=0

q(n−j)(h−j)

[n

j

][m

h− j

]=

[m + n

h

].

Proof. Proof of Theorem 5.2.2.

k≥d

[k

d

]1

[n]

[n

k

][mn

k − 1

]q(k−d)(k−1)

=

[nd

]

[n]

n∑

k=d

[n− d

n− k

][mn

k − 1

]q(k−d)(k−1)

=

[nd

]

[n]

n−d∑j=0

[n− d

j

][mn

n− 1− j

]q(n−d−j)(n−1−j)(q-Vandermonde Convolution)

=

[nd

]

[n]

[mn + n− d

n− 1

]

=1

[mn− d + 1]

[mn + n− d

d, n− d,mn− d

].

Remark5.2.1. It is difficult to find a combinatorial interpretation for the left hand side of

Theorem 5.2.2. As a matter of fact, the most straightforward generalization of (5.2.1) even

fails for the2-Dyck paths:

w∈CW 22

qmajw = 1 + q2 + q3 6= [1]

[5]

[6

2

]= 1 + q2 + q4.

111

5.3 (q, t)-m-Schroder Statistics and the Shuffle Conjecture

Similar to the manner of [HL], for anm-Dyck path of ordern, we may associate it with

m-parking functionsby placing one of then “cars”, denoted by the integers 1 throughn, in

the square immediately to the right of eachN step ofD, with the restriction that if cari is

placed immediately on top of carj, theni > j. LetPmn denote the collection ofm-parking

functions onn cars.

Definition 5.3.1. Given anm-parking function, itsm-reading wordis obtained by reading

from NE to SW line by line, starting from the lines farther from them-diagonalx = my.

Figure 5.3 illustrates anm-parking function with 231 as itsm-reading word. The first

line we look at is the line connecting cars 2 and 3. We read it from NE to SW so that 2 is

before 3. Then the next line is them-diagonalx = my which contains car 1.

3

1

2

Figure 5.3: Anm-parking function whosem-reading word is 231.

Definition 5.3.2. Given anm-parking function, itsnatural expansionis defined as follows:

starting from (0, 0), eachN step, together with the car to its right, is duplicatedm times,

the car within theN step is duplicatedm times and put one to each of them N steps

duplicated; leave eachE step untouched.

Figure 5.4 illustrates the natural expansion of them-parking function shown in Figure

5.3. Note that the natural expansion of anm-parking function is kind of a “non-strict”

standard parking function in the sense that if car placingi immediately on top of carj

implies thati ≥ j instead ofi > j.

112

1

1

3

3

2

2

Figure 5.4: The natural expansion of anm-parking function.

Definition 5.3.3. [Sta99, page 482, Ex. 7.93] For two wordsu = (u1, . . . , uk) ∈ Sk and

v = (v1, . . . , vl) ∈ S(k + 1, k + l), whereS(m + 1,m + l) denotes all the permuted words

of {k + 1, · · · , k + l}, sh(u, v) or sh((u1, . . . , uk), (v1, . . . , vl)) is the set ofshufflesof u

andv, i.e.,sh(u, v) consists of all permutationsw = (w1, . . . , wk+l) ∈ Sk+l such that both

u andv are subsequences ofw.

If the m-reading word of anm-parking functionP is a shuffle of the two words(n−d+

1, · · · , n) and(n− d, · · · , 2, 1), the increasing order of(n− d + 1, · · · , n) will imply that

any singleN segment ofP contains at most 1 of{n− d+1, · · · , n}. Furthermore, each of

{n − d + 1, · · · , n} should occupy the top spot of someN segment. Hence if we change

thesed top N steps all toD steps and remove the cars in them-parking function, we will

get anm-Schroder path withd diagonal steps. Conversely, given a pathΠ ∈ Smn,d, we may

change itsd diagonal steps tod NE pairs; after that place cars{n − d + 1, · · · , n} to the

right of thed newN steps, and place cars{n− d, · · · , 2, 1} to the right of the othern− d

D steps in the uniquely right order so that them-reading word of them-parking function

formed is a shuffle of the two words(n− d + 1, · · · , n) and(n− d, · · · , 2, 1). In this way

everym- Schroder corresponds to anm-parking function of the particular type. Because it

is easier to manipulate when there are noD steps, we define them-Schroder polynomial in

the following way.

113

Definition 5.3.4. Them-Schroder polynomial is defined as

Smn,d(q, t) =

∑Π: Π∈Pmn and them-reading word ofΠ∈sh((n−d+1,··· ,n),(n−d,··· ,1))

qdinvm(Π)tarea(Π),

wheredinvm(Π) = dinv(Π), Π is the natural expansion ofΠ, anddinv is the obvious

generalization of the statistic on parking functions introduced in [HL].

Them-Shuffle Conjecture is due to Haglund, Haiman, Loehr, Remmel and Ulyanov.

Conjecture 5.3.1. [HHL+]

Smn,d(q, t) =< ∇men, en−dhd >,

where∇ is a linear operator defined in terms of the modified Macdonald polynomials (for

details see[HHL+]).

114

Bibliography

[Abe26] N. H. Abel, Beweis eines Ausdrucks, von welchem die Binomial-Formel ein

einzelner Fall ist, Crelle’s J. Reine Angew. Math.1 (1826), 159–160.

[And98] George E. Andrews,The theory of partitions, Cambridge Mathematical Li-

brary, Cambridge University Press, Cambridge, 1998, Reprint of the 1976

original.

[Bei82] Janet Simpson Beissinger,On external activity and inversions in trees, J. Com-

bin. Theory Ser. B33 (1982), no. 1, 87–92.

[Ber76] Claude Berge,Graphs and hypergraphs, revised ed., North-Holland Publish-

ing Co., Amsterdam, 1976, Translated from the French by Edward Minieka,

North-Holland Mathematical Library, Vol. 6.

[Bj o92] Anders Bjorner,The homology and shellability of matroids and geometric lat-

tices, Matroid applications, Encyclopedia Math. Appl., vol. 40, Cambridge

Univ. Press, Cambridge, 1992, pp. 226–283.

[BK01] Jason Bandlow and Kendra Killpatrick,An area-to-inv bijection between Dyck

paths and 312-avoiding permutations, Electron. J. Combin.8 (2001), no. 1,

Research Paper 40, 16 pp. (electronic).

115

[BP97] Janet S. Beissinger and Uri N. Peled,A note on major sequences and exter-

nal activity in trees, Electron. J. Combin.4 (1997), no. 2, Research Paper 4,

approx. 10 pp. (electronic), The Wilf Festschrift (Philadelphia, PA, 1996).

[BSS93] Joseph Bonin, Louis Shapiro, and Rodica Simion,Someq-analogues of the

Schroder numbers arising from combinatorial statistics on lattice paths, J.

Statist. Plann. Inference34 (1993), no. 1, 35–55.

[BWG] J. Backelin, J. West, and G. Xin,Wilf-equivalence for singleton classes,

Preprint.

[Car56] L. Carlitz,The expansion of certain products, Proc. Amer. Math. Soc.7 (1956),

558–564.

[Cig87] J. Cigler,Some remarks on Catalan families, European J. Combin.8 (1987),

no. 3, 261–267.

[Cla58] L. E. Clarke,On Cayley’s formula for counting trees, J. London Math. Soc.33

(1958), 471–474.

[Com74] Louis Comtet,Advanced combinatorics, enlarged ed., D. Reidel Publishing

Co., Dordrecht, 1974.

[CR64] L. Carlitz and J. Riordan,Two element lattice permutation numbers and their

q-generalization, Duke Math. J.31 (1964), 371–388.

[EHKK03] E. S. Egge, J. Haglund, K. Killpatrick, and D. Kremer,A Schroder generaliza-

tion of Haglund’s statistic on Catalan paths, Electron. J. Combin.10 (2003),

no. 1, Research Paper 16, 21 pp. (electronic).

[FH85] J. Furlinger and J. Hofbauer,q-Catalan numbers, J. Combin. Theory Ser. A40

(1985), no. 2, 248–264.

116

[FS78] Dominique Foata and Marcel-Paul Schutzenberger,Major index and inversion

number of permutations, Math. Nachr.83 (1978), 143–159.

[GH96] A. M. Garsia and M. Haiman,A remarkableq,t-Catalan sequence andq-

Lagrange inversion, J. Algebraic Combin.5 (1996), no. 3, 191–244.

[GH01] A. M. Garsia and J. Haglund,A positivity result in the theory of Macdonald

polynomials, Proc. Natl. Acad. Sci. USA98 (2001), no. 8, 4313–4316 (elec-

tronic).

[GH02] , A proof of theq,t-Catalan positivity conjecture, Discrete Math.256

(2002), no. 3, 677–717, LaCIM 2000 Conference on Combinatorics, Computer

Science and Applications (Montreal, QC).

[Gor63] B. Gordon,Two theorems on multipartite partitions, J. London Math. Soc.38

(1963), 459–464.

[GS96] Ira M. Gessel and Bruce E. Sagan,The Tutte polynomial of a graph, depth-

first search, and simplicial complex partitions, Electron. J. Combin.3 (1996),

no. 2, Research Paper 9, approx. 36 pp. (electronic), The Foata Festschrift.

[GW79] Ira Gessel and Da Lun Wang,Depth-first search as a combinatorial correspon-

dence, J. Combin. Theory Ser. A26 (1979), no. 3, 308–313.

[Hag03] J. Haglund,Conjectured statistics for theq,t-Catalan numbers, Adv. Math.175

(2003), no. 2, 319–334.

[Hag04] , A proof of theq, t-Schroder conjecture, Int. Math. Res. Not. (2004),

no. 11, 525–560.

117

[Hagon] , Theq,t-Catalan numbers and the space of diagonal harmonics, Uni-

versity Lecture Series, American Mathematical Society, Providence, RI, in

prepation.

[Hai94] Mark D. Haiman,Conjectures on the quotient ring by diagonal invariants, J.

Algebraic Combin.3 (1994), no. 1, 17–76.

[HHL+] J. Haglund, M. Haiman, N. Loehr, J. Remmel, and A. Ulyanov,A combinato-

rial formula for the character of the diagonal coinvariants, Duke Math. J., to

appear.

[HL] J. Haglund and N. Loehr,A conjectured combinatorial formula for the Hilbert

series for diagonal harmonics, Discrete Math., (Proceedings of the FPSAC

2002 Conference held in Melbourne, Australia), to appear.

[HLR] J. Haglund, N. Loehr, and J. Remmel,Statistics on wreath products, perfect

matchings and signed words, Eur. J. Combin., (Special issue on “Combina-

torics and Representation Theory”), to appear.

[HP91] Peter Hilton and Jean Pedersen,Catalan numbers, their generalization, and

their uses, Math. Intelligencer13 (1991), no. 2, 64–75.

[HPW99] Peter Hilton, Jean Pedersen, and Tamsen Whitehead,On paths on the integral

lattice in the plane, Far East J. Math. Sci. (FJMS) (1999), no. Special Volume,

Part I, 1–23.

[Knu73] Donald E. Knuth,The art of computer programming. Volume 3, Addison-

Wesley Publishing Co., Reading, Mass.-London-Don Mills, Ont., 1973, Sort-

ing and searching, Addison-Wesley Series in Computer Science and Informa-

tion Processing.

118

[Kra01] C. Krattenthaler,Permutations with restricted patterns and Dyck paths, Adv. in

Appl. Math.27 (2001), no. 2-3, 510–530, Special issue in honor of Dominique

Foata’s 65th birthday (Philadelphia, PA, 2000).

[Kre80] G. Kreweras, Une famille de polynomes ayant plusieurs proprietes

enumeratives, Period. Math. Hungar.11 (1980), no. 4, 309–320.

[Loe03] N. Loehr,Multivariate analogues of Catalan numbers, parking functions, and

their extensions, Ph.D. thesis, University of California at San Diego, 2003.

[Mac60] Percy A. MacMahon,Combinatory analysis, Two volumes (bound as one),

Chelsea Publishing Co., New York, 1960.

[Moo70] J. W. Moon,Counting labelled trees, From lectures delivered to the Twelfth

Biennial Seminar of the Canadian Mathematical Congress (Vancouver), vol.

1969, Canadian Mathematical Congress, Montreal, Que., 1970.

[Pri97] A. L. Price,Packing densities of layered patterns, Ph.D. thesis, University of

Pennsylvania, 1997.

[Rei93] Victor Reiner,Signed permutation statistics, European J. Combin.14 (1993),

no. 6, 553–567.

[Rio68] John Riordan,Combinatorial identities, John Wiley & Sons Inc., New York,

1968.

[SS85] Rodica Simion and Frank W. Schmidt,Restricted permutations, European J.

Combin.6 (1985), no. 4, 383–406.

[Sta] Richard P. Stanley, Catalan addendum, Available at www-

math.mit.edu/˜rstan/catadd.pdf.

119

[Sta99] , Enumerative combinatorics. Vol. 2, Cambridge Studies in Advanced

Mathematics, vol. 62, Cambridge University Press, Cambridge, 1999, With a

foreword by Gian-Carlo Rota and appendix 1 by Sergey Fomin.

[Ste02] J. Michael Steele,Minimal spanning trees for graphs with random edge

lengths, Mathematics and computer science, II (Versailles, 2002), Trends

Math., Birkhauser, Basel, 2002, pp. 223–245.

[Str92] Volker Strehl,Identities of Rothe-Abel-Schlafli-Hurwitz-type, Discrete Math.

99 (1992), no. 1-3, 321–340.

[SW02] Zvezdelina Stankova and Julian West,A new class of Wilf-equivalent permu-

tations, J. Algebraic Combin.15 (2002), no. 3, 271–290.

[Tut54] W. T. Tutte,A contribution to the theory of chromatic polynomials, Canadian

J. Math.6 (1954), 80–91.

[Wil94] Herbert S. Wilf,generatingfunctionology, second ed., Academic Press Inc.,

Boston, MA, 1994.

[Wri61] E. M. Wright, Partition of multipartite numbers into a fixed number of parts,

Proc. London Math. Soc. (3)11 (1961), 499–510.

120


Recommended