You are on page 1of 81

Contents

Notation iii
Chapter 1. Introduction 1
§1.1. What is discrete mathematics? 1
§1.2. Applications 2
§1.3. Famous theorems 2
§1.4. Combinatorial objects 2
§1.5. Combinatorial ideas 3
§1.6. Prizes and journals in discrete mathematics 3
§1.7. People in discrete mathematics 4
§1.8. Conjectures in discrete mathematics 4
§1.9. Computer programming 6
§1.10. Internet sources 6
§1.11. Recommended books 6
Chapter 2. Basic mathematical logic 7
§2.1. Propositions and truth tables 8
§2.2. Mathematical proof 12
§2.3. Paradoxes 14
Exercises 2 14
Chapter 3. Enumeration 15
§3.1. How to count 15
§3.2. Set partitions and multisets 17
§3.3. Integer partitions and compositions 18
§3.4. Permutations 24
§3.5. Words 31

i
ii Contents

§3.6. Young tableaux, the Robinson-Schensted correspondence, and beyond 39


§3.7. Topics for an undergraduate thesis 42
Exercises 3 43
Chapter 4. Combinatorial numbers and identities 47
§4.1. Binomial coefficients 48
§4.2. Catalan numbers 52
§4.3. Fibonacci numbers 56
§4.4. Bell numbers 57
§4.5. Topics for an undergraduate thesis 58
Exercises 4 59
Select solutions to Exercises 4 61
Bibliography 63
Author Index 73
Subject Index 75
Notation

A⇤ The set of finite words over an alphabet A, 31


A+ The set of nonempty finite words over an alphabet A, 31
f The number of standard tableaux of shape ` n, 39
`n A partition of the integer n, 18
µ |= n A composition of the integer n, 19
[n] The set {1, 2, . . . , n}, 15
N The set of natural integers, 15
¬p It is not the case that p, 8
p^q The proposition p and the proposition q, 9
Q
(q) Euler’s function (q) = n 1 (1 q n ), 21
p,q The proposition p if and only if the proposition q, 9
p)q The proposition p implies the proposition q, 9
p_q The proposition p or the proposition q, 9
[q n ]f (q) The coefficient of q n in the Taylor’s expansion of the series f (q) at q = 0, 17
c(n, r) The signless Stirling numbers of the first kind, 25
S(n, r) The Stirling numbers of the second kind, 17
s(n, r) The Stirling numbers of the first kind, 25
sh( ) The shape of the Ferrers/Young diagram , 19
Sn The set of permutations on the set [n], 24
Z The set of integers, 15
Z+ The set of positive integers, 15

iii
Chapter 1

Introduction

Discrete Mathematics appeared in university curricula in the 1980s, initially as a computer


science support course. The curriculum has thereafter developed in conjunction with e↵orts
by ACM and MAA into a course that is basically intended to develop mathematical maturity
for first-year students.

1.1. What is discrete mathematics?


(1) Theoretical computer science: graph theory, mathematical logic, algorithm, data
structure, automato theory, formal language theory, . . .

(2) Information theory and coding theory, . . .

(3) Logic: valid reasoning and inference, directed graphs, fuzzy logic, . . .

(4) Combinatorics: enumerative combinatorics, analytic combinatorics, algebraic com-


binatorics, probabilistic combinatorics, topological combinatorics, arithmetic combi-
natorics, partition theory (q-series, special functions, orthogonal polynomials, . . . ),
graph theory (structural, algebraic, topological, . . . ), design theory, finite geometry,
matroid theory, extremal combinatorics, . . .

(5) Order theory: posets, lattices, . . .

(6) Combinatorics on words

(7) Discrete and computational geometry: tiling, . . .

(8) Number theory: cryptography, modular arithmetic, transcendental numbers, p-adic


analysis, Diophantine approximation

(9) Topology: combinatorial topology, topological graph theory, topological combina-


torics, computational topology, finite topological space, . . .

1
(10) Operations research: Linear programming, optimization, queuing theory, scheduling
theory, . . .
(11) Decision theory, utility theory, social theory, game theory, . . .
(12) Discretization: Numerical analysis, . . .
(13) Discrete analogues of continuous mathemtics: discrete Fourier transforms, discrete
geometry, di↵erence equations, . . .
(14) . . .

1.2. Applications
(1) Computer algorithms
(2) Programming
(3) Cryptography: Tutte at World War II and the public-key cryptography in the Cold
War.
(4) Automated theorem proving
(5) Software development

1.3. Famous theorems


(1) The four color theorem solved by Appel and Haken in 1976.
(2) The marriage theorem.
(3) The friendship theorem: Suppose in a group of people we have the situation that
any pair of persons have precisely one common friend. Then there is always a person
(the “politician”) who is everybody’s friend.
(4) The art gallary theorem.
(5) Bu↵on’s needle theorem: If a short needle of length l is dropped on paper that is
ruled with equally spaced lines of distance d l, then the probability that the needle
comes to lie in a position where it crosses one of the lines is exactly (2l)/(⇡d).
(6) Monsky’s theorem: It is impossible to dissect a square into an odd number of trian-
gles of equal area.
(7) The Sylvester-Gallai theorem: In any configuration of n points in the plane, not all
on a line, there is a line which contains exactly two of the points.
(8) . . .

1.4. Combinatorial objects


(1) Sets, sequences, permutations, partitions
(2) Graphs
(3) Lattice paths
(4) Young tableaux, symmetric functions
(5) All other kinds of combinatorial configurations, . . .

1.5. Combinatorial ideas


(1) A = B: Design an autoproof, try double counting, . . .
(2) A nonstandard method of counting trees: Put a cat into each tree, walk your dog,
and count how often he barks.
(3) Make something extreme (cf. consider a limit)
(4) The pigeon-hole principle.
(5) Find recurrence
(6) . . .

1.6. Prizes and journals in discrete mathematics

The Fulkerson Prize is awarded for outstanding papers in discrete mathematics. See also
Kirkman Medal.

(1) Annals of Combinatorics.


(2) Combinatorial Theory.
(3) Combinatorica.
(4) Combinatorics, Probability and Computing.
(5) Designs, Codes and Cryptography.
(6) Discrete & Computing Geometry.
(7) Discrete Applied Mathematics.
(8) Discrete Mathematics.
(9) Discrete Mathematics & Theoretical Computer Science.
(10) Electronic Jounral of Combinatorics.
(11) European Jounral on Combinatorics.
(12) Geometriae Dedicata.
(13) Journal of Algebraic Combinatorics.
(14) Journal of Algorithms.
(15) Graphs and Combinatorics.

(16) Journal of Combinatorial Designs.

(17) Journal of Combinatorial Optimization.

(18) Journal of Combinatorics.

(19) Journal of Combinatorics Series A.

(20) Journal of Combinatorics Series B.

(21) Journal of Cryptology.

(22) Journal of Discrete Algorithms.

(23) Journal of Graph Theory.

(24) Random Structures and Algorithms.

(25) Séminaire Lostharingien de Combinatoire.

(26) SIAM Journal on Discrete Mathematics.

(27) The Ramanujan Journal.

(28) . . .

1.7. People in discrete mathematics


(1) Paul Erdős

(2) Richard Stanley

(3) Lauren Williams

(4) Laszlo Lovász

(5) . . .

1.8. Conjectures in discrete mathematics

Conjecture 1.1 (Erdős-Turán conjecture). Let A be a set of positive integers such that
X1
= 1.
n
n2A

Then A contains an arithmetic progression of any given length.

Green and Tao [52] solved the case the A consists of all positive primes.
P
Conjecture 1.2 (Kurepa [70]). Define !n = nj=01 j!. Then no odd prime p divide !p.
Conjecture 1.3 (Collatz conjecture, the 3n + 1 problem). Define the function f by
(
n/2, if n is even;
f (n) =
3n + 1, if n is odd.

Then for any positive integer n, there exists a number j such that the jth iteration f (j) (n)
equals 1.

Paul Erdős said about the Collatz conjecture: “Mathematics may not be ready for such
problems.” He also o↵ered 500 US dollars for its solution. Lagarias in 2010 claimed that
based only on known information about this problem, “this is an extraordinarily difficult
problem, completely out of reach of present day mathematics.”

Conjecture 1.4 (Singmaster, 1971). The multiplicities of entries in Pascal’s triangle that is
not 1 have a finite upper bound. In other words, the number N (a) of times the number a > 1
appears in Pascal’s triangle satisfies N (a) = O(1).

For instance,
✓ ◆ ✓ ◆ ✓ ◆ ✓ ◆ ✓ ◆ ✓ ◆
120 16 10 10 16 120
120 = = = = = = ,
1 2 3 7 14 119
✓ ◆ ✓ ◆ ✓ ◆ ✓ ◆ ✓ ◆ ✓ ◆ ✓ ◆ ✓ ◆
3003 78 15 14 14 15 78 3003
3003 = = = = = = = = ,
1 2 5 6 8 10 76 3002
It can be checked that
N (120) = 6 and N (3003) = 8.
It is not known whether there is a number a 6= 3003 such that N (a) = 8. It is also not known
whether there is a number a such that N (a) = 5 or N (a) = 7. See Kane [64], Singmaster
[109].

Conjecture 1.5 (Ringel, 1964). For any tree T with m edges, the complete graph K2m+1
decomposes into 2m + 1 copies of T .

Conjecture 1.6 (Kotzig and Ringel, 1964). Every tree has an elegant labeling, where an
elegant labeling of a tree with m edges is an injection f : V (G) ! {0, 1, . . . , m} such that
{|f (u) f (v)| : uv 2 E(G)} = {1, 2, . . . , m}.

Conjecture 1.7 (Erdős-Straus). For any integer n 2, the number 4/n is the sum of three
positive unit fractions, i.e., the Diophantine equation
4 1 1 1
= + +
n x y z
has a positive integer solution (x, y, z).

Conjecture 1.7 is confirmed for n up to 1017 , see arxiv:1406.6307 and Erdős [37].

Conjecture 1.8 (Erdős–Gyárfás conjecture, 1995). Every graph with minimum degree 3
contains a simple cycle whose length is a power of 2.
Erdős o↵ered a prize of $100 for proving this conjecture, or $50 for a counterexample.
Daniel and Shauger (A result on the Erdős–Gyárfás conjecture in planar graphs, 2001) showed
its truth for planar claw-free graphs.
Conjecture 1.9 (Thomassen). Every 4-connected line graph is Hamiltonian.

1.9. Computer programming


(1) Mathematical softwares: C, Maple, Mathematica, Matlab, Sage, . . . .
(2) Operating systems: Linus, MacOS, Unix, Windows, . . .

1.10. Internet sources


(1) Wiki in English.
(2) OEIS.
(3) Links on my homepage.
(4) . . . .

1.11. Recommended books

Aigner and Ziegler [1].


Chapter 2

Basic mathematical
logic

When dealing with people,


remember you are not dealing
with creatures of logic...
Dale Carnegie

Three classical paradoxes:

(1) Is the answer to this question ‘No’ ?


(2) A barber shaves all those, and those only, who do not shave themselves.
(3) Impossible is not in my vocabulary.

7
Much material of this chapter comes from Byer, Smeltzer, and Wantz [13], Epp [34]. The
next paragraph is almost copied from Epp’s book.
The first great treatises on logic were written by the Greek philosopher Aristotle. They
were a collection of rules for deductive reasoning that were intended to serve as a basis for
the study of every branch of knowledge. In the 17th century, Leibniz conceived the idea
of using symbols to mechanize the process of deductive reasoning in much the same way
that algebraic notation had mechanized the process of reasoning about numbers and their
relationships. Leibniz’s idea was realized in the 19th century by Boole and De Morgan, who
founded the modern subject of symbolic logic.
In 1977, Barwise [7] makes a rough division of contemporary mathematical logic into four
areas:

(1) set theory,


(2) model theory,
(3) recursion theory,
(4) proof theory and constructive mathematics.

For instance, basic questions addressed by recursion theory include:

• What does it mean for a function on the natural numbers to be computable?


• How can noncomputable functions be classified into a hierarchy based on their level
of noncomputability?

We do not use the definition enviroments for most “definitions” in this chapter, for most
notion involved are informally described. Rather, to be inspiring are they expected.

2.1. Propositions and truth tables

A statement or proposition is a declarative sentence that is true of false but not both. Every
proposition can be assigned an associated truth value, either true or false, denoted by T
and F , respectively. A conjecture is a statement that is believed to be true and has not been
proven.
The symbol ¬ is used in front of a proposition to denote the negation of the proposition.
The navigation statement ¬p is read “not p” when symbols are alone used. While the
statements are written out, the negation is usually written using typical sentence structures.
For example, if
p : Earth is a planet,
then the negation is
¬p : It is not the case that Earth is a planet,
or, alternatively,
¬p : Earth is not a planet.
Let p and q be propositions. We introduce four logical connectors, each of which is to create
a compound proposition based on p and q.

(1) The symbol p ^ q reads “p and q”, and called a conjunction.


(2) The symbol p _ q reads “p or q”, and called a disjunction.
(3) The symbol p ) q, reads
• p implies q, or
• If p, then q, or
• q if p, or
• p is sufficient for q, or
• q is necessary for p,
and called an implication, or a conditional statement. In the proposition p ) q,
the proposition p is called the hypothesis or antecedent, and q the conclusion or
consequent.
(4) The symbol p , q, reads
• p if and only if q, or
• p is necessary and sufficient for q, or
• If p, then q, and conversely,
and called a biconditional statement.

The statement “A unless B” means if B is false then A is true. A vacuous truth is a conditional
statement that is true because the antecedent cannot be satisfied.
Table 2.1 summarizes succinctly the rules for determining the truth value of the compound
proposition of two simple statements p and q, formed by negation, conjuction, disjunction,
conditional and biconditional.
Table 2.1. A truth table example.

p q ¬p p ^ q p _ q p ) q p , q
T T F T T T T
T F F F T F F
F T T F T T F
F F T F F T T

For a given compound proposition C, we determine the truth value of C incrementally,


using the order of operations as

(1) parentheses,
(2) ¬,
(3) ^,
(4) _,
(5) ),
(6) ,.

For example, Table 2.2 is the truth table for the proposition (p ) q) ) r.

Table 2.2. The truth table for the proposition (p ) q) ) r.

p q r p ) q (p ) q) ) r
T T T T T
T T F T F
T F T F T
T F F F T
F T T T T
F T F T F
F F T T T
F F F T F

A compound proposition is a tautology (resp., contradiction) if its truth value is true


(resp., false) for all truth value combinations of the simple propositions. For example,
(p ^ q) ) p and (p ^ q) , (q ^ p)
are tautologies; p ^ ¬p is a contradiction.
Theorem 2.1. Let p, q and r be propositions. Then the following laws hold.

(1) Commutative laws:


• p ^ q () q ^ p and
• p _ q () q _ p.
(2) Associative laws:
• (p ^ q) ^ r () p ^ (q ^ r) and
• (p _ q) _ r () p _ (q _ r).
(3) Distributive laws:
• (p ^ q) _ r () (p _ r) ^ (q _ r) and
• (p _ q) ^ r () (p ^ r) _ (q ^ r).
(4) the double negation law: ¬(¬p) () p.
(5) De Morgan’s laws:
• The negation of a disjunction is the conjunction of the negations, namely
¬(p ^ q) () (¬p) _ (¬q).
• The negation of a conjunction is the disjunction of the negations, namely
¬(p _ q) () (¬p) ^ (¬q).
(6) Idempotent laws:
• (p ^ p) () p and
• (p _ p) () p.
Augustus De Morgan (1806–1871) introduced a formal version of the laws to classical
propositional logic. De Morgan’s formulation was influenced by algebraization of logic un-
dertaken by George Boole. Nevertheless, a similar observation was made by Aristotle, and
was known to Greek and Medieval logicians. Still, De Morgan is given credit for stating the
laws in the terms of modern formal logic, and incorporating them into the language of logic.
An instructor who do not know de Morgan’s law may ask the students that “do you like
my teaching and would you join my class next semester?” He may feel confused, or not, if a
student replies “No.”
Let p ) q be an implication. Then we have the following terminologies.

(1) The converse of p ) q is


q ) p.

(2) The inverse of p ) q is


¬p ) ¬q.

(3) The contrapositive of p ) q is


¬q ) ¬p.

A predicate is a statement that may be true of false depending on the values of its
variables. For example, x2 > 0 is a predicate, while “for each real number x, there exists a
real number y such that x2 + y > 0” is a proposition.

(1) The phrase “for each” (alternatively, for all, for every, for any) is called universal
quantifiers and denoted by the symbol 8.

(2) The phrase “there exists” (alternatively, there is/are) is called existential quantifiers
and denoted by the symbol 9.

Both universal quantifiers and existential quantifiers are called quantifiers. Incorporating
more than one variable using nested quantifiers can render predicates significantly more
complicated. Two common propositions are

(1) 8 x, 9 y, P (x, y).

(2) 9 x, 8 y, P (x, y).

Consider the following dialogue which is adapted from [13].

• Duadua: I believe all boys in your class admires some girl.

• David: She must be quite busy.

The sentence implies a possibility that all boys admires the same girl. To minimize ambiguity,
the universal quantifier in a statement of the form
8 a 9 b, P (a, b),
is often translated as “For each” or “For any”, rather than “For all” or “For every”.
2.2. Mathematical proof

A mathematical proof that consists of a sequence of statements that begins with hypotheses
or premises and ends with a conclusion is also called a formal argument. The final conclusion
of a proof establishes a theorem, a statement that has been proven to be true using other
mathematical statements and acceptable rules of reasoning.
A syllogism is a kind of logical argument that applies deductive reasoning to arrive at a
conclusion based on two or more propositions that are asserted or assumed to be true. Some
basic forms of valid arguments are called rules of inference. Here are some basic rules of
inference.

(1) Modus ponens (implication elimination):


p ^ (p ) q) =) q.
(2) Modus tollens (denying the consequent):
¬q ^ (p ) q) =) ¬p.
(3) Conjunctive simplification:
p ^ q =) p.
(4) Disjunctive syllogism:
(p _ q) ^ ¬p =) q.
(5) Hypothetical syllogism:
(p ) q) ^ (q ) r) =) p ) r.
(6) Universal modus ponens:
8 x 2 S, P (x) ) Q(x) ^ P (s) =) Q(s).
(7) Universal modus tollens:
8 x 2 S, P (x) ) Q(x) ^ ¬Q(s) =) ¬P (s).

A fallacy is the use of invalid or otherwise faulty reasoning in the construction of an


argument.

(1) Appeal to probability is the logical fallacy of taking something for granted because
it would probably be the case or might possibly be the case.
For example, do you think every problem in the final examination of our class
should be lectured by David before the examination?
(2) The fallacy fallacy:
(p ) q) ^ ¬p =) ¬q.
For example, students go to class when there is no COVID-19. Now, along with
COVID-19, do students go to class or not?
(3) A false exclusionary disjunct:
(p _ q) ^ p =) ¬q.
For example, David writes a recommendation letter for a student if the student
obtained the full score in the class of Algebra or the student passed the examination
of Discrete Mathematics. Would David recognize a student with the full score in
Algebra as failed in the Discrete Mathematics examination?
(4) Fallcy of the converse:
(p ) q) ^ q =) p.
For example, a student has basic knowledge on Discrete Mathematics if he has
joined David’s class. Has every student who has basic knowledge on Discrete Math-
emaics joined David’s class?
(5) Fallacy of the inverse:
(p ) q) ^ ¬p =) ¬q.
A modified version of David’s class example works as an example of this fallacy.
(6) The existential fallacy:
(p ) q) =) 9 p.
For example, a girl claims that she will marry the first man who proposed to her
riding a colorful Xiangyun.
(7) Affirmative conclusion from a negative premise:
(p ) ¬q) ^ (q ) ¬r) =) (p ) r).
For example, no dogs are cats, and no cats can fly, therefore all dogs can fly.
(8) Illicit major (the major term is undistributed in the major premise but distributed
in the conclusion):
(p ) q) ^ (r ) ¬p) =) (r ) ¬q).
An example hint is that cats are animals and dogs are not cats.
(9) Illicit minor (the minor term is undistributed in the minor premise but distributed
in the conclusion):
(p ) q) ^ (p ) r) =) (q ) r).
An example hint is that dogs are animals and dogs weigh less than 1 ton.
(10) A fallacy of necessity is a fallacy whereby a degree of unwarranted necessity is placed
in the conclusion.
For example, bachelors are necessarily unmarried and Duadua is a bachelor.
Does these imply that Duadua cannot marry?

There are several methods of proving something that we have been already very familiar
with:

(1) direct proof,


(2) proof by contradiction,
(3) proof by contrapositive,
(4) constructive proof,
(5) counterexample,
(6) probablistic proof for existence,
(7) vacuous proof,
(8) . . .

2.3. Paradoxes

A paradox is a logically self-contradictory statement or a statement that runs contrary to


one’s expectation.

(1) The liar paradox: I am lying.


(2) The barber paradox: one who shaves all those, and those only, who do not shave
themselves.
(3) Is the answer to this question ‘No’ ?
(4) Impossible is not in my vocabulary.
(5) My wish is to make my wish fail.

Exercises 2

1) Consider the following inference: Since time is money and money is no object, time is no
object. How do you think?

2) Check Zeno’s paradox on Achilles and the tortoise, and find out a solution for the paradox.

3) Please tell some examples of the logic fallacies in your real life.

4) In philosophical logic, the masked-man fallacy is commited when one makes an illicit
use of Leibniz’s law in an argument, which states that if A and B are the same object,
then A and B are indiscernible. The masked-man fallacy can be seen from the argument
that duadua thinks David is smart but the instructor of Discrete Mathematics is not, and
obtain the conclusion that David is not the instructor. However, we know in mathematics
that if x = y and y 6= z then x 6= z. What di↵erences lie in there if we compare these
two inferences?

Homework: Ex. 2 1). Due time: 11.59 pm, Sept. 26th, 2021.
Chapter 3

Enumeration

For general and extensive intrdouction for all kinds of enumerative problems arising from
combinatorics, see Stanley’s bible book [114].
Throughout this note, we use Z to denote the set of integers, N the set of natural integers,
and Z+ the set of positive integers.

3.1. How to count


(1) The number of subsets of the set
[n] = {1, 2, . . . , n}
is 2n , since every element belongs to or does not belong to a given subset.
(2) The number of permutations on [n] is n!, since there are n choices for the element
at the first position, n 1 choices for the element at the second position, . . . , and 1
choice for the element at the last position.
(3) The number of derangements on [n]:
n
X ( 1)i
dn = n! .
i!
i=0
This formula can be obtained by using the principle of inclusion-exclusion, see Sec-
tion 3.4.
(4) Let f (n) be the number of n ⇥ n 01-matrices M such that every row and every
column of M has three ones. For example, f (0) = f (1) = f (2) = 0, f (3) = 1. In
fact,
1 X ( 1) n!2 ( + 3 )!2↵ 3 (3n)!
f (n) = n 2
⇠ n 2,
6 ↵! ! ! 6 36 e
where ↵, , and run over nonnegative integers such that
↵+ + = n.

15
(5) Let f (n) be the number of subsets of [n] that do not contain two consecutive integers.
For example, f (4) = 8. It is easy to see that
f (n) = f (n 1) + f (n 2)
for n 2, which implies that
1
f (n) = p ⌧ n+2 ⌧¯n+2 ,
5
p p
where ⌧ = (1 + 5)/2 and ⌧¯ = (1 5)/2.
Example 3.1. Find the sequence a0 = 1, a1 , a2 , . . . of real numbers satisfying
Xn
(3.1) ak an k = 1
k=0
for all n 2 N.

Solution. Consider the ordinary generating function


X
F (x) = an xn .
n 0

Then Eq. (3.1) is to say that


X
F (x)2 = xn
n 0
and ✓ ◆
1 X 1/2
F (x) = p = ( x)n .
1 x n 0 n
It follows that
✓ ◆
n 1/2 ( 1/2)( 3/2)( 5/2) · · · ((1 2n)/2) (2n 1)!!
an = ( 1) = ( 1)n = .
n n! (2n)!!

The double factorial n!! is the product of all positive integers at most n that have the
same parity as n.
For general ideas of analysing methods and more examples for generating functions,
see Flajolet and Sedgewick [41] and Wilf [126].
Example 3.2. Suppose that a0 = a1 = 1 and
(3.2) an = an 1 + (n 1)an 2

for n 2. Consider the exponential generating function


X xn
F (x) = an .
n!
n 0

By Eq. (3.2), we deduce that


X xn X xn
F (x) = 1 + x + an =1+x+ an 1 + (n 1)an 2 .
n! n!
n 2 n 2
Taking di↵erentiation of both sides yields
F 0 (x) = 1 + (F (x) 1) + xF (x).
Solving it we obtain ✓◆
x2
F (x) = exp x + .
2
With the generating function in hand, we can deduce for an an explicit formula, a recurrence,
an asymptotic formula and so on. First,
! !
X xn 2
X xn X x2n
an = ex ex /2 = .
n! n! (2n)!!
n 0 n 0 n 0

Extracting the coefficient of xn /n!yields


X ✓n ◆
an = (i 1)!!.
0in
i
i even

Second, by di↵erentiation, we obtain


X ✓ ◆ X xn
xn 1 x2
an = (1 + x) exp x + = (1 + x) an .
(n 1)! 2 n!
n 0 n 0

Equating the coefficients of xn /n!


yields the recurrence Eq. (3.2). Thirdly, regarding as a
function of a complex variable, exp x + x2 /2 is an entire function so that standard technique
from the theory of asymptotic estimates can be used to estimate that
✓ ◆
1 n/2 n p 1
f (n) ⇠ p n exp + n .
2 2 4

Conclusion.

(1) Explicit formulas.


(2) Recurrences.
(3) Generating functions: Ordinary generating functions and exponential generating
functions.
(4) Asymptotic formulas.
(5) . . .

We use the notation [q n ]f (q) to denote the coefficient of q n in the Taylor’s expansion of
the series f (q) at q = 0.

3.2. Set partitions and multisets

Definition 3.3. A partition of the set [n] is a distribution of the elements of [n] into several
disjoint nonempty sets B1 , B2 , . . . , Br , so that each element is placed into exactly one set Bi .
The sets Bi are called blocks. The Stirling number S(n, k) of the second kind is the number
of ways of partitioning [n] into k nonempty indistinguishable blocks. A matching of [n] is a
partition of [n] whose every block contains at most 2 elements.
Let n = 7 and r = 4. Then
{1, 2, 4}, {3, 6}, {5}, {7}
is a partition of the set [7] into 4 blocks.

n=0 1
n=1 0 1
n=2 0 1 1
n=3 0 1 3 1
n=4 0 1 7 6 1
n=5 0 1 15 25 10 1

Figure 3.1. The Stirling numbers S(n, k) of the second kind for 0  n  5.

Proposition 3.4. For all n, k 1,


k ✓ ◆
1 X i k
S(n, k) = ( 1) (k i)n .
k! i
i=0

Hint. Consider the surjections from [n] to [k], and use the principle of inclusion-exclusion. ⇤
Conjecture 3.5 (Wegner, 1973). For any n 3, there is only one index rn such that
S(n, rn ) = max S(n, k).
1kn

It is checked for n  106 , see Canfield and Pomerance [15, 16], Wegner [123].
A multiset is a set allowing repeated elements. We write
n o
M = am 1
1
, a m2
2 , . . . , a mr
r

to denote a multiset, where the elements ai are distinct, and mi is the multiplicity of ai .

3.3. Integer partitions and compositions

A bible for the theory of integer partitions is Andrews and Eriksson [4], see also Stanley [114,
Chapter 1].
Definition 3.6. A partition of a positive integer n is a sequence of positive integers,
commonly written as
= ( 1 , 2 , . . . , r ) ` n,
such that
r
X
1 2 ··· r 1 and i = n.
i=1
Let p(0) = 1 and let p(n) be the number of integer partitions of n for n 1. The function
p(n) is called the partition function. A composition of n is a finite sequence µ of positive
integers, commonly written as
µ = (µ1 , µ2 , . . . , µr ) |= n,
such that
r
X
µi = n.
i=1
The numbers i and µi are said to be the parts of and µ, respectively. A weak composition
of n is a list of nonnegative integers whose sum is n.

Set p(0) = 1. The values of the partition function p(n) for small n, see Table 3.1. The
Table 3.1. The values of the partition function p(n) for n = 0, 1, . . . , 16.

n 0 1 2 3 4 5 5 7 8 9 10 11 12 13 14 15 16
p(n) 1 1 2 3 5 7 11 15 22 30 42 56 77 101 135 176 231

only known direct formula for p(n) is


⇣ q ⌘
p ✓ k 1 ⇣ ⌫ ◆ ⇡ 2 1
X k X 2⇡inh X j hj hj 1⌘ d sinh k 3 z 24
p exp + ⇡i · q .
⇡ 2 0<h<k k k k k 2 dz z 24 1
k 1 j=1 z=n
(h,k)=1

From definition, we see that a partition ` n is essentially a set of integers whose sum
is n. Thus it corresponds to a multiset 1m1 2m2 . . . nmn bijectively, where mi is the number
of occurrences of the part i in . Commonly the term imi is omitted if mi = 0. In this way,
the partition is denoted compactly as
mk
= am 1 m2
1 a2 · · · ak ,
where n a 1 > a 2 > · · · > ak 1 are positive integers, and mj 1 is the multiplicity of
aj for all j 2 [k]. The multiplicities mj that equal to 1 are also omitted. For instance, the
partition (4, 4, 2, 1, 1) ` 12 is alternatively denoted by 42 212 ` 12.
By considering the last part of a composition, we know that the number of compositions
of n into ones and twos equals the Fibonacci number Fn+1 , which is defined recursively by
Fn = Fn 1 + Fn 2 and F1 = F2 = 1.
Proposition 3.7. The di↵erence p(n) p(n 1) is the number of partitions whose every
part is at least 2. As a consequence, p(n) > p(n 1).

Proof. By adding a 1-part to a partition of n 1, one obtains the combinatorial interpretation


of p(n) p(n 1). ⇤
Definition 3.8. For any partition = ( 1 , . . . , r ) ` n, we draw a left-justified array of n
dots with i dots in the ith row. This array is called the Ferrers diagram of . If we replace
the dots by juxtaposed squares, then we call the resulting diagram the Young diagram. The
squares are also called boxes or cells. The partition that is represented by such a diagram
is said to be the shape of the diagram, denoted sh . For any cell c = (i, j) in a Ferrers
diagram , the set of dots that lie to the right of c and c itself is said to be the arm of c, and
the set of dots that lie below c and c itself is said to be the leg of c. The hook with respect
to c is the union of c and its arm and leg. Flipping the diagram of a partition of n along
its main diagonal gives another partition of n. We call the resulting partition the conjugate
of . A partition is self-conjugate if it is its own conjugate. The Durfee square of a Young
diagram is the largest square that fits inside the board.

See Fig. 3.2 for these graphs corresponding to the partition 42 211 , whose conjugate is
5322 .The self-conjugate partitions of 12 are
6214 , 53212 , and 42 22 .
If we turn the Young diagram up-side-down, then we call the resulting diagram the Young

Figure 3.2. The Ferrers diagram and Young board corresponding to the partition 42 211 .

diagram in the French notation.


Proposition 3.9. The number of partitions of n with m parts equals the number of partitions
of n with the largest part m.

Proof. Direct by considering the conjugate of a partition. ⇤


Proposition 3.10. The number of self-conjugate partitions of n equals the number of parti-
tions of n into distinct odd parts.

Proof. The hooks of the cells on the diagonal of a self-conjugat partition of n form a partition
of n into distinct odd parts. For example, the self-conjugate partition 53212 ` 12 corresponds
to the partition 93 ` 12. It is obvious that this correspondence is a bijection. ⇤
Proposition 3.11. The number of partitions of n with Durfee side j is
X
p(m | parts  j) · p(n j 2 m | parts  j).
m

where p(t | parts  j) is the number of partitions of t whose every part is at most j.

Proof. Let be a partition with Durfee side j. The Durfee square of divides into a right
part and a lower part. Count the number of such partitions with respect to the number of
boxes in the right part yields the desired formula. ⇤
Proposition 3.12. For any set S of positive integers, let
mk
Pn (S) = am 1 m2
1 a2 · · · ak ` n : ai 2 S for all i

be the set of partitions of n whose parts are in S, and


mk
Pn (S, d) = am 1 m2
1 a2 · · · ak 2 Pn (S) : mi  d for all i

the set of partitions in Pn (S) whose every part appears at most d times. Then
X Y 1
(3.3) |Pn (S)|q n = , and
1 qn
n 0 n2S
X Y1 q (d+1)n
(3.4) |Pn (S, d)|q n = .
1 qn
n 0 n2S

Proof. Suppose that S = {n1 , n2 , . . . , ns }. Then


Y 1
= 1 + q n1 + q 2n1 + · · · · 1 + q n2 + q 2n2 + · · · · · · · · 1 + q nk + q 2nk + · · · .
1 qn
n2S

Extracting the coefficient of q n , we find


Y 1
[q n ] = (a1 , a2 , . . . , ak ) 2 Nk : a1 n1 + a2 n2 + · · · + ak nk = n ,
1 qn
n2S

which equals Pn (S) by definition. Similarly, since


Y1 q (d+1)n
= 1 + q n1 + · · · + q dn1 1 + q n2 + · · · + q dn2 · · · 1 + q nk + · · · + q dnk ,
1 qn
n2S

Eq. (3.4) is to say that


⇢ k
X
k
|Pn (S, d)| = (a1 , a2 , . . . , ak ) 2 N : ai ni = n and ai  d for all i 2 [k] ,
i=1

which is true by the definition of Pn (S, d). ⇤


Remark 3.13. Eq. (3.3) can be obtained from Eq. (3.4) by taking d ! 1.
Definition 3.14. The Euler’s function is
Y
(q) = (1 q n ).
n 1

Euler’s function is a model example of a q-series, a modular form, and provides the
prototypical example of a relation between combinatorics and complex analysis.
Corollary 3.15. The coefficient in the formal power series expansion for 1/ (q) gives the
number of partitions of n. In other words,
X 1
p(n)q n = .
(q)
n 0

Proof. Immediate by taking S in Proposition 3.12 to be the set of positive integers. ⇤


Corollary 3.16. For any set S of positive integers, let p0n (S) be the number of partitions
of n whose parts are distinct and are in S. Then
X Y
p0n (S)q n = (1 + q n ).
n 0 n2S

Proof. Immediate by taking d = 1 in Proposition 3.12. ⇤

Corollary 3.17. Let pn (m) be the number of partitions of n with at most m parts. Then
X 1
pn (m)q n = .
(1 q)(1 q 2 ) · · · (1 q m )
n 0

Proof. Using the conjugation transformation, we know that the number of partitions of n
whose every part is at most m equals the number of partitions of n with at most m parts.
Taking S = [m] in Eq. (3.3), we obtain the desired identity. ⇤

Corollary 3.18. The number pn (3) is the integer nearest to


(n + 3)2
.
12

Proof. Considering the generating function, we can deduce that


X 1
pn (3)q n =
(1 q)(1 q 2 )(1 q 3 )
n 0
1 1 17 1 q+2
= + + + +
6(1 q)3
4(1 q)2
72(1 q) 8(1 + q) 9(1 + q + q 2 )
1 1 1 1
= 3
+ 2
+ +
6(1 q) 4(1 q) 4(1 q ) 3(1 q 3 )
2
✓ ◆
1X n+2 n 1X 1 X 2n 1 X 3n
= q + (n + 1)q n + q + q
6 2 4 4 3
n 0 n 0 n 0 n 0
X✓ (n + 3)2 1 n 1 2n 1 3n

n
= q q + q + q
12 3 4 3
n 0

X (n + 3)2 ◆
= + ✏(n) q n ,
12
n 0

where ⇢
1 1 1
✏(n) 2 , , 0, .
3 12 4
The desired formula then follows from the fact that |✏(n)| < 1/2. ⇤

Theorem 3.19 is a classical result due to Euler.

Theorem 3.19. Every number has as many integer partitions into odd parts as into distinct
parts.
For instance, the integer partitions of 6 into odd parts are

16 , 313 , 32 , and 51,

and the integer partitions of 5 into distinct parts are

6, 51, 42, and 321.

Proof. We construct a bijection as follows.


For any partition ` n into odd parts, we describe a procedure of repeated merging of
pairs of equal parts: whenever one gets a partition with two equal parts i , merge them and
get a new part 2 i . For example,

3 + 3 + 3 + 1 + 1 + 1 + 1 ! (3 + 3) + 3 + (1 + 1) + (1 + 1) = 6 + 3 + (2 + 2) ! 6 + 3 + 4.

Conversely, we describe a procedure of repeated splitting of even parts: whenever one gets a
partition with an even part 2k, split them and get 2 new parts k + k. For example,

6 + 4 + 3 ! (3 + 3) + (2 + 2) + 3 = 3 + 3 + 3 + 2 + 2 ! 3 + 3 + 3 + (1 + 1) + (1 + 1).

It is easy to check that the above two procedures do not depend on the order of merging and
spliting, and that they form a bijection. ⇤

An algebraic proof to Theorem 3.19. Let an be the number of partitions of n into odd
parts, and bn the number of partitions of n into distinct parts. Then
X Y Y1 q 2n Y 1 X
bn q n = (1 + q n ) = = = an q n .
1 qn 1 qn
n 0 n 1 n 1 n odd n 0

Extracting the coefficient of q n from both sides, one obtains an = bn . ⇤

Theorem 3.20 (Rogers-Ramanujan identity). The number of partitions of n whose every


part equals 1 or 4 modulo 5 equals the number of partitions of n whose any two parts have
di↵erence at least 2.

For example, the partitions of 9 whose every part equals 1 or 4 modulo 5 are

9, 613 , 42 1, 415 , and 19 ,

while the partitions of 9 whose any two parts have di↵erence at least 2 are

9, 81, 72, 63, and 531.

McCoy (in a joint paper with Berkovich (1998)) speaking to the International Congress
of Mathematicians gave a survey of applications of Roger-Ramanujan identities in physics.
In addition, the fruitful interaction of partition identities with other combinatorial models is
beautifully outlined by Alladi (1995).
3.4. Permutations

There is a huge number of books and papers concentrated on permutation combinatorics.


See Bóna [11], Kitaev [66] for instance.
Definition 3.21. A permutation is a bijection from a set onto itself. The set of permutitions
on the set [n] is commonly denoted Sn . The complement of ⇡ is the permutation 1 2 · · · n ,
where i = n + 1 ⇡i .

There are several commonly notation for permutations in Sn :

(1) One-line notation: ⇡ = ⇡1 ⇡2 · · · ⇡n 2 Sn . Eg., ⇡ = 25431.


(2) Two-line notation:
✓ ◆ ✓ ◆
1 2 3 4 5 3 2 5 1 4
⇡= =
2 5 4 3 1 4 5 1 2 3
(3) Cycle notation: ⇡ = (125)(34).
(4) Matrix representation:
0 1
0 1 0 0 0
B 0 0 0 0 1 C
B C
⇡=B
B 0 0 0 1 0 C
C
@ 0 0 1 0 0 A
1 0 0 0 0

Let ⇡ 2 Sn . Let ci be the number of cycles of length i in ⇡. Then the total number
c(⇡) = c1 + c2 + · · · + cn
is the number of cycles of ⇡. We call the integer partition 1c1 2c2 · · · ncn the type of ⇡. Then
c1 + 2c2 + 3c3 + · · · + ncn = n.
As the notation of a multiset, the number ci is commonly omitted if ci = 1, and the whole
part ici is omitted if ci = 0. For instance, the type of the permutation ⇡ = (125)(34) is
10 21 31 40 50 ` 5,
which can be alternatively written as 32 to denote that ⇡ consists of a cycle of length 3 and
a cycle of length 2. Since all information contained in the notation 1c1 2c2 · · · ncn is contained
in the simpler notation
(c1 , c2 , . . . , cn ),
one may find in some literature that the sequence (c1 , c2 , . . . , cn ) is used to denote the type
of ⇡, or used together with the notation 1c1 2c2 · · · ncn . For clarification, we adopt the notation
1c1 2c2 · · · ncn only.
All permutations of [n] form a symmetric group, denoted Sn , where the group operation
is function composition. In general, the composition of two permutations of the same length
is not commutative, that is,
⇡ 6= ⇡.
A permutation 2 Sn is a conjugate of ⇡ if there is ⌧ 2 Sn such that
1
=⌧ ⇡⌧.
Two permutations in Sn are conjugate if and only if they are of the same type, see Sagan
[104, Page 3].
Proposition 3.22. The number of ⇡ 2 Sn of type c = 1c1 2c2 · · · ncn is n!/zc , where
zc = 1c1 c1 ! 2c2 c2 ! · · · ncn cn !.

Proof. Let Sn (c) be the set of permutations of type c in Sn . Define a map


: Sn ! Sn (c)
for ⇡ = ⇡1 ⇡2 · · · ⇡n 2 Sn by defining (⇡) to be the permutaiton
(⇡1 )(⇡2 ) · · · (⇡c1 )(⇡c1 +1 ⇡c1 +2 )(⇡c1 +3 ⇡c1 +4 ) · · · (⇡c1 +2c2 1 ⇡c1 +2c2 ) · · · ,
i.e., each of the first c1 elements of ⇡ forms a cycle of length 1 in (⇡), the next 2c2 elements
of ⇡ form c2 cycles of length 2, . . . .
For any 2 Sn (c), we claim that there are zc ways to write it in disjoint cycle notation
so that the cycle lengths are non-decreasing from left to right. In fact, one may order the
cycles of length i in ci ! ways, and choose the first element of each of these cycles in i ways.
These choices are all independent, and the claim is proved.
Hence the map is a zc -to-1 map, i.e.,
1
| ( )| = zc
for each 2 Sn (c). The proof follows since |Sn | = n!. ⇤
Definition 3.23. The numbers
s(n, k) = ( 1)n k
c(n, k)
are the Stirling numbers of the first kind , where c(n, k) is the number of ⇡ 2 Sn with exactly k
cycles, called a signless Stirling number of the first kind .
Proposition 3.24. The numbers c(n, k) satisfy the recurrence
c(n, k) = (n 1)c(n 1, k) + c(n 1, k 1),
for n, k 1, with the initial conditions c(n, k) = 0 if n  0 or k  0, except c(0, 0) = 1.
Moreover,
Xn
c(n, k)xk = x(x + 1)(x + 2) · · · (x + n 1).
k=0

Proof. Choose a permutation ⇡ 2 Sn 1 and we use its cycle notation. In order to obtain a
permutation in Sn with k cycles, we can either add a cycle (n) to ⇡ if ⇡ has k 1 cycles,
or place the letter n after any of the letters in some cycle of ⇡ if ⇡ has k cycles. These two
cases give the summands c(n 1, k 1) and (n 1)c(n 1, k) respectively. This proves the
desired recurrence. Let
Xn
Fn (x) = x(x + 1)(x + 2) · · · (x + n 1) = b(n, k)xk .
k=0
Then
n
X n
X1
Fn (x) = (x + n 1)Fn 1 (x) = b(n 1, k 1)xk + (n 1) b(n 1, k)xk .
k=1 k=0
Extracting the coefficients of Fn (x), we find
b(n, k) = (n 1)b(n 1, k) + b(n 1, k 1).
It is clear that b(0, 0) = 1 and b(n, k) = 0 if n < 0 or k < 0. Hence b(n, k) satisfies the same
recurrence and initial conditions as c(n, k) and they agree. ⇤

The number dn of derangements on [n] can be found by using the principle of inclusion-
exclusion:
Xn ✓ ◆ ✓ ◆
n n i 1 1 1 ( 1)n
dn = ( 1) i! = n! 1 + + ··· + ,
i 1! 2! 3! n!
i=0
which is the integer nearest to n!/e. As another example of using the principle of inclusion-
exclusion, let h(n) be the number of permutations of the multiset Mn = {12 , 22 , . . . , n2 } with
no consecutive terms equal. Then h(1) = 0, h(2) = 2, and
Xn ✓ ◆
n (n + i)!
h(n) = ( 1)n i .
i 2i
i=0

Permutation statistics is an active and popular research field. We introduce some basic
statistics on permutations.
Definition 3.25 (Permutation statistics). Let ⇡ = ⇡1 ⇡2 · · · ⇡n 2 Sn .

(1) A descent of ⇡ is a number i 2 [n 1] such that


⇡i > ⇡i+1 .
An ascent of ⇡ is a number i 2 [n 1] that is not a descent, i.e., a number i 2 [n 1]
such that
⇡i < ⇡i+1 .
The major index maj(⇡) is the sum of descents of ⇡, i.e.,
X
maj(⇡) = i.
⇡i >⇡i+1

(2) An inversion of ⇡ is a pair (⇡i , ⇡j ) such that


i<j and ⇡i > ⇡j .
(3) A fixed point of ⇡ is a number i 2 [n] such that ⇡i = i. An excedance of ⇡ is a
number i 2 [n] such that ⇡i > i, see Ehrenborg and Steingrı́msson [31]. A drop of ⇡
is a number i 2 [n] such that ⇡i < i.
(4) A peak of ⇡ is a number i 2 {2, 3, . . . , n 1} such that
⇡i 1 < ⇡i and ⇡i > ⇡i+1 .
A valley of ⇡ is a number i 2 {2, 3, . . . , n 1} such that
⇡i 1 > ⇡i and ⇡i < ⇡i+1 .
(5) A left-to-right maxima of ⇡ is a number ⇡i such that
⇡i > ⇡j
for all j < i. Similarly one may defines left-to-right minima, right-to-left maxima,
right-to-left minima.
(6) The number of cycles of ⇡ is denoted
c(⇡) = c1 + c2 + · · · + cn ,
where ci is the number of cycles of ⇡ of length i.
(7) . . . .

Two important kinds of permutations are involutions and derangements:

• An involution is a permutation whose every cycle has length at most 2.


• A derangement is a permutation ⇡ 2 Sn without fixed points.

The number of descents of ⇡ is denoted by des(⇡). Simliar notation contains asc(⇡),


inv(⇡), maj(⇡), . . . . For example, the permutation ⇡ = 25431 2 S5 has descent set {2, 3, 4},
ascent set {1}, inversion set
{(5, 4), (5, 3), (4, 3), (2, 1), (5, 1), (4, 1), (3, 1)},
excedance set {1, 2, 3}, and thus
des(⇡) = 3, maj(⇡) = 9, asc(⇡) = 1, inv(⇡) = 7, and exc(⇡) = 3.
The set of left-to-right maxima of ⇡ is {2, 5}, The set of left-to-right minima is {2, 1}, The
set of right-to-left maxima is {1, 3, 4, 5}, and The set of right-to-left minima is {1}.
The major index is named after Major MacMahon who showed Corollary 3.34 in 1913,
see MacMahon [82].
Definition 3.26. The nth Eulerian polynomial is
X n
X
An (x) = x1+des(⇡) = A(n, k)xk ,
⇡2Sn k=1
where the coefficient
A(n, k) = |{⇡ 2 Sn : des(⇡) = k 1}|
is called an Eulerian number.

See Petersen [97] for Eulerian numbers.


Proposition 3.27. For 1  k  n,
A(n, k + 1) = (k + 1)A(n 1, k + 1) + (n k)A(n 1, k).

Proof. Choose a permutation ⇡ = ⇡1 ⇡2 · · · ⇡n 1 2 Sn 1. In order to obtain a permutation


in Sn with k descents, we can place the letter n

(1) to the left of ⇡1 if ⇡ has k 1 descents, or


(2) between ⇡i and ⇡i+1 if ⇡i < ⇡i+1 and if ⇡ has k 1 descents, or

(3) to the right of ⇡n 1 if ⇡ has k descents, or

(4) between ⇡i and ⇡i+1 if ⇡i > ⇡i+1 and if ⇡ has k descents.

By definition, there are A(n 1, k) permutations in Sn 1 with k 1 descents, which are


exactly those with n k 1 ascents. For each of them, the first two possibilites above give n k
permutations in Sn with k descents. Similarly, for each of the A(n 1, k 1) permutations
in Sn 1 with k descents, the last two possibilities above give k + 1 permutations in Sn with k
descents. Combining them together, we obtain the desired recurrence. ⇤
Definition 3.28. Any statistic on permutations that has the same distribution with the
descent numbers is called an Eulerian statistic. Any statistic on permutations that has the
same distribution with the inversion numbers is called a Mahonian statistic.
Theorem 3.29. Excedance is an Eulerian statistic, i.e., for any n 1,
X X
q des(⇡) = q exc(⇡) .
⇡2Sn ⇡2Sn

Proof. Define a bijection : Sn ! Sn such that des(⇡) = exc( (⇡)). For example,
(931682745) = (139)(286)(47)(5).
Precisely speaking, form a cycle (⇡j ⇡j 1 · · · ⇡1 ), where j = ⇡ 1 (1). Continue this process
iteratively with the next smallest integer in what remains in ⇡, building up (⇡) cycle by
cycle. It is easy to show that is bijective. ⇤
Theorem 3.30. For n, r 1,
r ✓ ◆
1 X n k
S(n, r) = A(n, k) .
r! r k
k=0

Proof. We shall present a bijection between ordered partitions of [n] into r blocks and
permutations of [n] with at most r 1 descents. Since the former is counted by the left hand
side, and the latter the right hand side, we obtain a bijective proof.
In fact, the number of ordered partitions of [n] into r blocks is r!S(n, r). Given such an
ordered partition, we write the elements in every block in the increasing order, concatenate
the blocks, and remove the parentheses for distinguishing blocks. In this way, we obtain a
permutation of [n] with at most r 1 descents.
Conversely, let 0  k  r. Let ⇡ be a permutation of [n] with k 1 descents. Then the
descents partition ⇡ into k ascending runs. We are going to construct a partition of [n] into
r blocks based on the k ascending runs. For this, we need to partition some of the ascending
runs into small pieces so that the total number of pieces is r. Thus we need to “cut” r k
times on the ascending runs. Note that for any ascending run of t letters, there are t 1
candidate places to cut. Thus there are in total n k candidate places for ⇡ to do the cutting.
Therefore, we have nr kk choices to cut the runs and obtain exactly r ascending runs in a
sequence. ⇤
Definition 3.31. Let n 2 Z+ be a positive integer. The q-analogue of n is the polynomial
1 qn
[n]q = 1 + q + q 2 + · · · + q n 1 = ,
1 q
where q is an indeterminate. The q-analogue of n! is
[n]q ! = [n]q [n 1]q · · · [1]q .
The q-analogue of the binomial coefficient nk is
✓ ◆
n [n]q !
= .
k q [k]q ! [n k]q !

The q-analogue of n is a generalization of n, since one recovers n by taking the limit as


q ! 1. The indeterminate q is simply a device to track the operations performed on n.
Theorem 3.32 (Rodriguez [99]). For n 1,
X
q inv(⇡) = [n]q !.
⇡2Sn

Proof. By considering the e↵ect of inserting the letter n into a permutation on [n 1], we
obtain X X
q inv(⇡) = [n]q q inv(⇡) .
⇡2Sn ⇡2Sn 1

By induction, the desired formula holds. ⇤


Theorem 3.33. The inversion number and the major index are equidistributed, namely,
X X
xinv(⇡) y maj(⇡) = xmaj(⇡) y inv(⇡) .
⇡2Sn ⇡2Sn

Corollary 3.34 (MacMahon [83]). The major index is a Mahonian statistic, i.e., for any
n 1, X X
q inv(⇡) = q maj(⇡) .
⇡2Sn ⇡2Sn

Proof. Taking x = q and y = 1 in Theorem 3.33. ⇤


Theorem 3.35. p
X X xn y 1
y peak(⇡) =p p .
n! y 1 tan(x y 1)
n 0 ⇡2Sn

See Elizalde and Noy [32], Kitaev [65], Mendes and Remmel [87].
Lin and Kim [71] gave a joint distribution identity on 12 statistics.
Definition 3.36. Let L = (a0 , a1 , . . . , an ) be a sequence of positive numbers. We say that L
is unimodal if there exists an index 1  k  n such that
a0  a1  · · ·  ak ak+1 ··· an .
We say that L is log-concave if
ak 1 ak+1  a2k
P
for all k. We say that L is real-rooted if every zero of the polynomial nj=0 aj xj is real.
n
For example, the sequence k for k = 0, 1, . . . , n is unimodal, log-concave, and real-
rooted.
Theorem 3.37. Let L = (a0 , a1 , . . . , an ) be a sequence of positive numbers. Then we have
the following implications.

(1) If L is real-rooted, then L is log-concave.

(2) If L is log-concave, then L is unimodal.

(3) If L is real-rooted, then it has either 1 or 2 maximal elements.

In particular monotonic sequences of positive numbers are unimodal. One may find
examples to show that the converses of Theorem 3.37 are false.
Theorem 3.38. The sequence {A(n, k)}k of Eulerian numbers is real-rooted, i.e., all zeros
of the Eulerian polynomial are real.

Proof. Note that A0 (x) = x. By Proposition 3.27,


An (x) = (x x2 )A0n 1 (x) + nxAn 1 (x)

for n 1, which implies


d⇣ ⌘
(3.5) An (x) = x(1 x)n+1(1 x) n An 1 (x) .
dx
Now every polynomial An (x) has a zero x = 0. Except this zero, Eq. (3.5) implies, by
Rolle’s theorem, that An (x) has a zero between each pair of consecutive zeros of An 1 (x).
This accounts for n 1 of the zeros of An (x). Since we have accounted all but one zero,
the remaining last zero must be real since complex zeros of a polynomial come in conjugate
pairs. ⇤

A parallel study on permutations relates patterns.


Definition 3.39. The reduced form of a permutation on a set {j1 , j2 , . . . , jr } is the permu-
tation obtained by renaming the letters of so that ji is renamed i for all i 2 [r]. For r  n,
we say that a permutation ⇡ 2 Sn has an occurrence of the pattern ⌧ 2 Sr if there exist
1  i1 < i2 < · · · < ir  n such that ⌧ is the reduced form of the permutation ⇡i1 ⇡i2 · · · ⇡ir .
A permutation ⇡ 2 Sn is called ⌧ -avoiding if it has no occurrences of the pattern ⌧ .

For instance, the reduced form of the permutation 3174 is 2143, and the permutation
25431 is 213-avoiding.
Theorem 3.40 (Marcus and Tardos [84]). For any pattern ⌧ , the limit
p
lim n sn (⌧ )
n!1

exists and is finite, where sn (⌧ ) is the number of ⌧ -avoiding permutations ⇡ 2 Sn .


p
n
The limit limn!1 sn (⌧ ) is called a Stanley-Wilf limit.
3.5. Words

The readers are referred to Berstel and Perrin [10] for the history of combinatorics on words.
From Wiki, one may find the following introduction.
“Combinatorics on words is a fairly new field of mathematics, branching from combina-
torics, which focuses on the study of words and formal languages. The subject looks at letters
or symbols, and the sequences they form. Combinatorics on words a↵ects various areas of
mathematical study, including algebra and computer science. There have been a wide range
of contributions to the field. Some of the first work was on square-free words by Axel Thue
in the early 1900s. He and colleagues observed patterns within words and tried to explain
them. As time went on, combinatorics on words became useful in the study of algorithms
and coding. It led to developments in abstract algebra and answering open questions.”
“The first books on combinatorics on words that summarize the origins of the subject
were written by a group of mathematicians, many of whom were students of Marcel-Paul
Schützenberger, that collectively went by the name of M. Lothaire. Their first book was
published in 1983, when combinatorics on words became more widespread.”
Definition 3.41. Let A be a set that we shall call an alphabet. Its elements will be called
letters. A finite word over A is a finite sequence of elements of A. The set of all finite words
over A is denoted by A⇤ . It is equipped with a binary operation obtained by concatenating
two words.

The concatenation is obviously associative, which allows writing a word as a1 · · · an .


Definition 3.42. Let w = a1 · · · an be a word. The length of w is the number n, denoted |w|.
The empty word is the word of length 0, denoted ✏. Denote the set of nonempty finite words
over A by
A+ = A⇤ \ {✏}.
The reversal of w is the word an · · · a1 . A word is a palindrome if it equals its reversal. A
word v 2 A⇤ is a factor of a word x 2 A⇤ if there exists words u, w 2 A⇤ such that x = uvw.
For any factor v of x, we say that v is

• proper if v 6= x,
• a left factor or prefix if x = vw for some word w 2 A⇤ ,
• a right factor or suffix if x = uv for some word u 2 A⇤ .

A word x 2 A⇤ is primitive if it is not a power of another word, that is, if x 6= 1 and if x 2 z ⇤


for z 2 A⇤ implies x = z. Two words x and y are conjugate if there exist u, v 2 A⇤ such that
x = uv and y = vu.

Here are some palindromes:

• A man, a plan, a canal: Panama!


• Rise to vote sir.
• Was it a cat I saw?
• Madam, I’m adam.


The conjugate relation on A⇤ is an equivalence.


Theorem 3.43. Two words x, y 2 A+ are conjugate if and only if there exists z 2 A⇤ such
that
(3.6) xz = zy.
More precisely, Eq. (3.6) holds if and only if there exists u, v 2 A⇤ such that
x = uv, y = vu, z 2 u(vu)⇤ .

Proof. If x and y are conjugate, then there exists u, v 2 A⇤ such that x = uv and y = vu.
Hence xz = zy holds for z = u. This proves the forward direction. For the backward
direction, suppose that xz = zy for some z 2 A⇤ . It follows that for any n 1,
(3.7) xn z = xn 1
zy = xn 2
zy 2 = · · · = zy n .
On the other hand, there exists a unique integer n 1 such that
(3.8) (n 1)|x|  |z| < n|x|.
Consider this n in Eq. (3.7), we find that there exists u 2 A⇤ such that
z = xn 1
u.
Substituting it into Eq. (3.7), we obtain x2n 1u = xn 1 uy n , which reduces to
n n
(3.9) x u = uy .
Substituting it into Ineq. (3.8), we obtain 0  |u| < |x|. Therefore, there exists v 2 A+
such that x = uv. Substituting it into Eq. (3.9), we obtain (uv)n u = uy n , which reduces to
y = vu. It follows that x and y are conjugate. On the other hand,
z = xn 1
u = (uv)n u = u(vu)n 2 u(vu)⇤ .
This completes the proof. ⇤
Definition 3.44. Let A be an ordered alphabet. The lexicographic order on the set A+ is
as follows: for any words u, v 2 A+ , u < v if and only if either v = ux with x 2 A+ , or
u = ras, v = rbt, with a, b 2 A, a < b, and r, s, t 2 A⇤ .
A Lyndon word is a primitive word that is minimal in its conjugate class.

Equivalently speaking, a word w 2 A+ is Lyndon if and only if w < vu for any factor-
ization w = uv. For the enumeration of Lyndon words, see Exercise 2.38). Lyndon words
are named after mathematician Roger Conant Lyndon [78]. Lyndon’s original motivation of
introducing these words was to describe free Lie algebras.
Below we introduce two characterizations of Lyndon words. Denote by L the set of
Lyndon words.
Theorem 3.45. A word w 2 A+ is Lyndon if and only if it is strictly smaller than any of
its proper suffixes.

Proof. Let w 2 A+ . If w < v for all proper suffixes v of w = uv, then


w < v < vu,
which implies that w 2 L by definition.
Conversely, suppose that w 2 L. We claim that any proper suffix of w is not a prefix of
w. Suppose not. Then
w = uv = vt
for some u, v, t 2 A+ . By Theorem 3.43,
v = x(yx)k , t = yx, u = xy and w = x(yx)k+1 = (xy)k+1 x
for some x, y 2 A⇤ and k 0. Since w is Lyndon, it is smaller than its conjugate x(xy)k+1 :
x(yx)k+1 < x(xy)k+1 .
Cancelling the first common letter x, we find
(3.10) (yx)k+1 < (xy)k+1 .
Concatenating the letter x at the ends of the words on both sides of Ineq. (3.10), we obtain
(yx)k+1 x < (xy)k+1 x = w.
In other words, w is larger than its conjugate (yx)k+1 x, contradicting that w 2 L. This
proves the claim.
Now, suppose that w = uv with u, v 2 A+ . If v < w, then vu < w by the claim,
contradicting to that w 2 L. Hence w < v. ⇤
Theorem 3.46. A word w 2 A+ is Lyndon if and only if either

• w 2 A, or
• w = uv with Lyndon words u and v such that u < v.

More precisely, if v is the longest proper Lyndon suffix of a Lyndon word w = uv, then u is
Lyndon.

Proof. For the backwards direction, suppose that w = uv with u, v 2 L such that u < v.
We claim that
uv < v.
In fact, if u is a prefix of v, then uv < v by applying Theorem 3.45 for v 2 L; otherwise, the
premise u < v implies uv < v directly. Now, for any suffix v 0 of v, we have
w = uv < v  v 0
by Theorem 3.45; for any proper suffix u0 of u, we have u < u0 by applying Theorem 3.45 for
u 2 L and thus uv < u0 v. Therefore, w is smaller than any of its proper suffixes and thus
Lyndon.
For the forwards direction, suppose that w = uv 2 \A, where v is the longest proper
Lyndon suffix of w. Indeed v exists since the last letter of w is a proper Lyndon suffix. If
u 2 A, then u 2 L and u < uv < v by applying Theorem 3.45 for w 2 L, as desired.
Below we can suppose that u 62 A, i.e., |u| 2. We shall show that u 2 L. Let u0 be a
proper suffix of u. By the definition of v, we have u0 v 62 L. By Theorem 3.45, u0 v has a proper
suffix t such that u0 v > t. If u0 < t, then u0 < t < u0 v and t = u0 t0 for some t0 2 A+ such
that t0 < v. Now, the word v has a proper suffix t0 that is smaller than v. By Theorem 3.45,
v 62 L, a contradiction. Therefore, t  u0 . It follows that
u < uv < t  u0 .
By Theorem 3.45, we obtain u 2 L. This completes the proof. ⇤
Theorem 3.47. Any word w 2 A+ may be written uniquely as a nonincreasing product of
Lyndon words:
w = l1 · · · lk , li 2 L and l1 · · · lk .
Moreover, lk is the smallest suffix of w.

Proof. Any word w 2 A+ can be written as the product of Lyndon words since every letter
is Lyndon. Let
w = l1 · · · lk
be a factorization of w into Lyndon words with minimal number k of Lyndon words. If
li < li+1 for some i, then li li+1 2 L by Theorem 3.46, contradicting the minimality of k. This
proves that w has a factorization into nonincreasing Lyndon words. Suppose that
w = l1 · · · lk = l10 · · · lh0
be two distinct factorizations of w into nonincreasing Lyndon words. Without loss of gener-
ality, we can suppose that
l1 = l10 · · · li0 u,
0 . By Theorem 3.45, we deduce that
where u is a prefix of li+1
0
l1 < u  li+1  l10 < l1 ,
a contradition! This proves the uniqueness of the factorization.
Let v be the smallest suffix of w and set w = uv. Then v 2 L by Theorem 3.45. If u = ✏,
then the result is proved. Otherwise, let s be the smallest suffix of u with w = rsv. If s < v,
then sv 2 L by Theorem 3.45. It follows that sv < v by Theorem 3.45, contradicting the
choice of v. Hence s v. Iterating this process gives the factorization of w. ⇤

It is possible to obtain the factorization from left to right. See Duval [28].
From Lothaire [73, Page 18], one may find the following point of view for the study
of combinatorics on words. “The investigation of words includes a series of combinatorial
studies with rather surprising conclusions that can be summarized roughly by the following
statement: Each sufficiently long word over a finite alphabet behaves locally in a regular
fashion... The discovery and the analysis of these unavoidable regularities constitute a major
topic in the combinatorics of words.” See also Lothaire [74, 75].
Definition 3.48. Let A be an alphabet. A square is a word of the form uu over A, where
u 2 A+ . A square-free word is a word that does not contain a square. A cube is a word of
the form uuu, where u 2 A+ . A cube-free word is a word that does not contain a cube.

The only square-free nonempty words over two letters a and b are
a, b, ab, ba, aba, bab.
We will show that there is an infinite number of finite square-free words over a 3-letter
alphabet, see Proposition 3.52 and Theorem 3.58. For this purpose, we introduce the concept
of overlapping factors, hereditary properties of words, and the infinite word of Thue-Morse.
Theorem 3.49. Let A be an alphabet. A word w 2 A⇤ contains two overlapping occurrences
of a word u 2 A+ if and only if w contains a factor of the form avava, where a is a letter
and v is a word.

Proof. Suppose that


(3.11) w = xuy = x0 uy 0
for some words x, u, y, x0 , y 0 , where the occurrences of u overlap. We can suppose that
|x| < |x0 | < |xu| < |x0 u|.
Then it follows from Eq. (3.11) that
(3.12) x0 = xs,
(3.13) xu = x0 z, and
0
(3.14) x u = xut
for some words s, z, t 2 A+ . Substituting Eq. (3.12) into Eqs. (3.13) and (3.14), we obtain
u = sz and su = ut
respectively. Substituting Eq. (3.13) into Eq. (3.14), we obtain
u = zt.
Let a be the first letter of s, and therefore also of u and of z. Set s = av and z = az 0 . Then
u = sz = avaz 0 , and
w = (x0 u)y 0 = (xut)y 0 = x(su)y 0 = x(av)(avaz 0 )y 0
contains a factor avava, in which the factor ava has two overlapping occurrences. ⇤
Corollary 3.50. Any infinite word that has no overlapping factors is cube-free.

Proof. Assume that an infinite word ! over an alphabet A contains a cube uuu. Suppose
that u = av where a 2 A and v 2 A⇤ . Then the factor avava is contained in uuu, and thus
in !. ⇤
Definition 3.51. An infinite word over an alphabet A is a function
! : N ! A,
denoted ! = a0 a1 · · · , where ai 2 A are letters. For any k 2 N, the prefix of length k of ! is
the word
! [k] = a0 a1 · · · ak 1 .
A factor of ! is a factor of ! [k] for some k. For any property P of words over A, we say
that ! satisfies P if so does every factor of !. A property for words is hereditary if it carries
over from a word to its factor; in other words, if a word w satisfies this property, then so does
every factor of w.

For instance, the property of being square-free is hereditary. Infinite words are useful in
dealing with hereditary properties, see Proposition 3.52.
Proposition 3.52. Let A be a finite alphabet and P a hereditary property. Then the set of
finite words over A that satisfies P is infinite if and only if there is an infinite word over A
that satisfies P .

Proof. Let L be the set of finite words over A that satisfies P . If there is an infinite word
over A satisfies P , then L contains all factors of that word and is infinite. This proves the
backwards direction.
For the forward direction, suppose that L is infinite. Since A is finite, there exists a
letter a0 2 A that starts an infinite number of words in L. Let L0 be the set of words in L
that starts with a0 . Since L0 is infinite, there exists a letter a1 2 A such that a0 a1 starts
an infinite number of words in L0 . Continuing in this way, one may obtain an infinite word
! = a0 a1 · · · whose every prefix starts an infinite number of words in L. We claim that !
satisfies P . In fact, for any factor u = ai · · · aj of !, every word in the infinite set Lj starts
with a0 · · · ai 1 u and satisfies P . Since P is hereditary, u satisfies P . ⇤

Note that Proposition 3.52 gives a way of producing infinite words from finite ones. We
now introduce another way of doing this.
Let w0 , w1 , · · · 2 A⇤ such that wn 1 is a prefix of wn for all n. This allows us to define
an infinite word ! over A by
! [|wn |] = wn for all n.
The word ! is called the limit of the sequence {wn }n 0 , denoted ! = limn!1 wn or simply
! = lim wn .

Below is an important application of this way of defining definite words.


Proposition 3.53. Let ↵ : A⇤ ! A⇤ be a morphism such that

• for any letter a 2 A, the word ↵(a) 6= ✏, and


• there exists a letter a0 2 A such that ↵(a0 ) = a0 u for some word u 2 A+ .

Then ↵n (a0 ) is a prefix of ↵n+1 (a0 ) for all n, and the limit lim ↵n (a0 ) exists.

Proof. It is direct to show that


↵n (a0 u) = ↵n (a0 )↵n (u).
In fact, this is true for n = 1 since ↵ is a morphism, and true by induction for n 2 since
↵n (a0 u) = ↵ ↵n 1
(a0 u) = ↵ ↵n 1
(a0 )↵n 1
(u) = ↵n (a0 )↵n (u).
It follows that
↵n+1 (a0 ) = ↵n (a0 u) = ↵n (a0 )↵n (u).
Thus ↵n (a0 ) is a prefix of ↵n+1 (a0 ), and therefore the limit lim ↵n (a0 ) exists. ⇤

We say that the limit lim ↵n (a0 ) is obtained by iterating ↵ on a0 . Note that any map
f : A⇤ ! A⇤ can be extended to a map on all words over A by defining
f (b0 b1 · · · ) = f (b0 )f (b1 ) · · · , where bi 2 A.
Since ↵(a) 6= ✏ for any a 2 A, the image f (b0 )f (b1 ) · · · is an infinite word.

Theorem 3.54. The limit !0 = lim ↵n (a0 ) is a fixed point of ↵, namely


↵(!0 ) = !0 .

Proof. Suppose that !0 = a0 a1 · · · , where ai 2 A. For any prefix u = a0 · · · ak of !0 ,


↵(!0 ) = ↵(a0 a1 · · · )
= ↵(a0 )↵(a1 ) · · ·
= ↵(a0 · · · ak )↵(ak+1 ak+2 · · · )
= ↵(u)↵(ak+1 ak+2 · · · ).
Thus the word ↵(u) is a prefix of ↵(!0 ). By definition, any prefix u must be of the form
↵n (a0 ) for some n 1. As a result, ↵(!0 ) = !0 . ⇤

With Proposition 3.53 in hand, one may define the infinite word of Thue-Morse as follows.

Definition 3.55. Let A = {a, b}. Let µ : A⇤ ! A⇤ be the morphism defined by


µ(a) = ab and µ(b) = ba.
The infinite word of Thue-Morse is the limit lim µn (a).

For instance, the first few words in the sequence {µn (a)}n 1 are as follows:
µ(a) = ab,
µ2 (a) = abba,
µ3 (a) = abbabaab,
µ4 (a) = abbabaabbaababba.

Proposition 3.56. The infinite word lim µn (a) = t0 t1 · · · of Thue-Morse satisfies


(
a, if the number of digits 1 in the binary expansion of n is even,
tn =
b, otherwise.
Proof. Let t = lim µn (a) = t0 t1 · · · . By Theorem 3.54,
t = µ(t) = µ(t0 )µ(t1 ) · · · .
Thus µ(tn ) = t2n t2n+1 . By the definition of µ, this implies that
t2n = tn and t2n+1 = t¯n ,
where ā = b and b̄ = a. If n is even, then tn = tn/2 satisfies the desired formula. Suppose
that n is odd. For any odd n 1, there is a unique nonnegative integer r such that
2r < n < 2r+1 .
We shall show by induction on r. It is direct to check the desired truth for all integers n with
r = 1. Since 2r 1 < (n 1)/2 < 2r , we have
tn = t̄(n 1)/2
(
ā, if the number of digits 1 in the binary expansion of (n 1)/2 is even
=
b̄, otherwise
(
b, if the number of digits 1 in the binary expansion of n is odd
=
a, otherwise,
as desired. This completes the proof. ⇤
Theorem 3.57. The infinite word of Thue-Morse has no overlapping factor.

Proof. Let A = {a, b} and X = {ab, ba}.


First of all, we claim that axa, bxb 62 X ⇤ for any x 2 X ⇤ . In fact, this can be shown by
induction on |x|. It is true if |x| = 0 since aa, bb 62 X ⇤ . Suppose that axa 2 X ⇤ . Then
axa = (ab)y(ba)
for some word y 2 X ⇤ . By induction, we know that x = byb 62 X ⇤ , a contradiction. This
proves axa 62 X ⇤ . Similarly one may show that bxb 62 X ⇤ . This confirms the claim.
Second, we claim that the morphism µ preserves the property of having no overlapping
factor. In fact, suppose that µ(w) has an overlapping factor for some w 2 A⇤ . We shall show
that w has an overlapping factor. By Theorem 3.49, there are words x, v, y 2 A⇤ and a letter
c 2 {a, b} such that
µ(w) = xcvcvcy.
Note that |cvcv| is even. Since µ(w) 2 X ⇤ , we find |µ(w)| is even, and |xy| is odd. Therefore,
either

• |x| is even, and x, cvcv, cy 2 X ⇤ , or


• |x| is odd, and xc, vcvc, y 2 X ⇤ .

If |v| is even, then cvc, v 2 X ⇤ , contradicting the first claim. Thus |v| is odd.
If |x| is even, then cv 2 X ⇤ . Thus w = rsst where r, s, t 2 A⇤ are determined by
µ(r) = x, µ(s) = cv and µ(t) = cy.
Then both s and t start with the letter c. Suppose that s = cs0 . Then w contains the factor
ssc = czczc. By Theorem 3.49, w has an overlapping factor.
If |x| is odd, then vc 2 X ⇤ . Thus w = rsst such that
µ(r) = xc, µ(s) = vc and µ(t) = y.
Then both r and s end with the letter c̄, and w contains the overlapping factor c̄ss. This
confirms the second claim.
Now, if the infinite word ! of Thue-Morse has an overlapping factor avava, then avava is
contained in a prefix µk (a) for some k. On the other hand, since a has no overlapping factor,
by iterated application of the second claim, the prefix µn (a) for any n has no overlapping
factor, a contradiction. This completes the proof. ⇤
Theorem 3.58. There exists a square-free infinite word over the alphabet B = {a, b, c}.

Proof. Let A = {a, b} and B = {a, b, c}. Consider the morphism : B ⇤ ! A⇤ defined by
(a) = abb, (b) = ab and (c) = a.
Let ! be an infinite word over A = {a, b} starting with a, and without overlapping factor.
By Theorem 3.57, such a word ! exists. By algorithmic arguments, we see that ! can be
factored uniquely as
! = y0 y1 · · · ,
where yi 2 {abb, ab, a}. In other words, there exists a unique infinite word ⌫ 2 B ⇤ such that
(⌫) = !. If ⌫ contains a square uu, then
! = (⌫) = · · · (u) (u)a · · ·
contains an overlapping factor avava, where v is the suffix of (u) such that (u) = av, a
contradiction. Hence ⌫ is square-free infinite word over B. ⇤

3.6. Young tableaux, the Robinson-Schensted correspondence, and beyond

Definition 3.59. A Young tableau consists of a Young board with each board cell filled with
an integer. A Young tableau is said to be standard if the union of the integers is [n], such
that the integers increasing in every row and in every column, where n is the number of cells.
The number of standard tableaux of shape ` n is denoted by f .

Rich structures of Young tableaux would be clearer when one studies the symmetric
functions and Schur bases, wich is closely related to the representation theory, see Fulton
[45], Mendes and Remmel [88], Sagan [104].
Theorem 3.60 is the famous “hook-length formula” due to Frame, Robinson, and Thrall,
see also Franzblau and Zeilberger [44] for a combinatorial proof.
Theorem 3.60 (Frame et al. [43]). The number of standard tableaux of shape ` n is
n!
f =Q ,
(i,j) h(i,j)

where h(i,j) is the hook length of the cell (i, j) in .


In 1938, Robinson discovered a correspondence between the set Sn of permutations and
the set of pairs of standard tableaux of n. It was then found independently in quite a di↵erent
form by Schensted. Now we present the latter version of the algorithm, which is called the
Robinson-Schensted correspondence.
Definition 3.61. Let ⇡ = ⇡1 ⇡2 · · · ⇡n 2 Sn . We construct a sequence of tableaux pairs
(P0 , Q0 ) = (;, ;), (P1 , Q1 ), (P2 , Q2 ), ..., (Pn , Qn ) = (P, Q),
where ⇡1 , ⇡2 , . . . , ⇡n are inserted into the tableaux P and the numbers 1, 2, . . . , n are placed
in the tableaux Q. The operations of insertion and placement are now described. Let P be
a partial tableau, i.e., an array with distinct entries whose rows and columns increase. Let x
be an element that is not in P . To row insert x into P , we proceed as follows:

RS1: Set R to be the first row of P .


RS2: While x is less than some element in Row R, do
RSa: Let y be the smallest element of R that is larger than x and replace y by x
in R, denoted R x.
RSb: Set x to be y and set R to be the next row down.
RS3: Now x is larger than every element of R. Put x at the end of Row R and stop.

Below we describe the procedure of forming the tableau Q. At the step of putting the
element ⇡k into P , we add a new cell to Qk 1 that is contained in Pk but not in Pk 1 , and
place the number k into that cell. So the shape of Qk coincide with the shape of Pk for each k.
We call P = Pn the P -tableau or insertion tableau, of ⇡, and write P = P (⇡). Similarly,
Q = Qn is called the Q-tableau or recording tableau, and denoted Q = Q(⇡). The above
procedure defines a map ⇡ 7! (P, Q).

For instance, the procedure of row-inserting the element 3 into the partial tableau is as

1 2 5 8
P = 4 7
6
9

follows:

1 2 5 8 3 1 2 3 8 1 2 3 8 1 2 3 8
4 7 4 7 5 4 5 4 5 .
6 6 6 7 6 7
9 9 9 9

Theorem 3.62 (Robinson [98], Schensted [105]). The map ⇡ 7! (P, Q) defined above is a
bijection between Sn and the set of pairs of standard tableaux (P, Q) of partitions of n, where
P and Q have the same shape.

The proof below can be found from [104, Page 94].


Proof. To show that the map ⇡ 7! (P, Q) is a bijection, it suffices to construct the inverse
(P, Q) 7! ⇡. We merely reverse the map ⇡ 7! (P, Q) step by step.
We begin by defining (Pn , Qn ) = (P, Q). Suppose that (Pk , Qk ) has been constructed.
We will find the kth element ⇡k for ⇡ and the tableau pair (Pk 1 , Qk 1 ). To avoid double
subscripting in what follows, we use Pij to stand for the (i, j) entry of Pk .
Find the cell (i, j) in Qk that contains the number k. Since k is the largest entry in Qk ,
the cell Pij must have been the last element to be displaced in the construction of Pk . We
can then use the following procedure to delete Pij from P .

SR1: Set x = Pij and erase Pij . Set R to be the (i 1)th row of Pk .

SR2: While R is not the zeroth row of Pk , do


SRa: Let y be the largest entry of R that is smaller than x and replace y by x in R.
SRb: Set x to be y and set R to be the next row up.

SR3: Now x has been removed from the first row. Set ⇡k = x.

It is easy to see that Pk 1 is Pk after the deletion process is complete, and that Qk 1 is Qk
with the number k erased. Continuing in this way, we eventually recover all elements of ⇡ in
reverse order. ⇤
Corollary 3.63. For n 1,
X
n! = (f )2 .
`n

Proof. By Theorem 3.62, the number n! of permutations on [n] equals the number of pairs
of standard tableaux of n, which is counted by the right hand side of the desired formula. ⇤
Theorem 3.64 (Schützenberger [106]). For any permutation ⇡, the insertion tableau of ⇡ 1

coincides with the recording tableau of ⇡, namely


1
P (⇡ ) = Q(⇡).

Proof. See [106] or [104, Page 111]. ⇤


P
Corollary 3.65. The number of involutions on [n] is the number `n f of standard
tableaux of n.

Proof. Note that involutions are exactly permutations ⇡ such that ⇡ 1 = ⇡. By Theo-
rem 3.64, the pair (P, Q) of tableaux corresponding to an involution satisfies P = Q. The
number of such pairs is exactly the number of standard tableaux of n. Hence the desired
result follows from Theorem 3.62. ⇤

For example, there are 4 involutions on [3]:


123, 132, 213, and 321.
The 4 standard tableaux of 3 are illustrated in Fig. 3.3.
1 2 3 1 2 1 3 1
3 2 2
3

Figure 3.3. The 4 standard tableaux of 3.

Theorem 3.66 (Greene [53]). Given ⇡ 2 Sn . Let shP (⇡) = ( 1 , . . . , l ) with conjugate
( 01 , . . . , 0m ). Then for any k, the sum 1 + · · · + k is the largest total length of k disjoint
increasing subsequences of ⇡, and the sum 01 + · · · + 0k is the largest total length of k disjoint
decreasing subsequences of ⇡.

Proof. See [104, Page 104]. ⇤

More interesting structure includes the Knuth relations, Viennot’s geometric construction,
Schützenberger’s jeu de taquin, . . . .
Stanley [113] talked about the length of a longest inceasing subsequence in a permutation
on ICM 2006. Let isn (⇡) be the length of a longest increasing subsequence of ⇡. The expected
value isn (⇡) of isn (⇡), where ⇡ ranges uniformly over Sn , is
1 X
En = isn (⇡).
n!
⇡2Sn
It is shown that p
n p
 En  e n.
2
Hammersley [59] showed in 1972, using subadditive ergodic theory, that the limit
En
lim p
n!1 n
exists. This limit was then showed to be 2. One of the proofs is based on the identity
1 X 2
En = 1· f ,
n!
`n
which is an immediate consequence of the Robinson-Schensted correspondence.

3.7. Topics for an undergraduate thesis


(1) Combinatorial proof for identities involving integer partitions.
(2) Eulerian and Mahonian statistics over permutations.
(3) Young tableaux in enumerative combinatorics.
(4) Counting with symmetric functions.
Exercises 3

2.1) Is an “element of a set” a special case of a “subset of a set”?

2.2) Define the union of several sets.

2.3) A nonempty set has the same number of odd subsets (i.e., subsets with an odd number
of elements) as even subsets.

2.4) What is the number of integers with at most / exactly n digits?

2.5) We have 20 kinds of presents and we want to distribute them to 12 children. We have a
large supply of each kind. If no children get two copies of the same present, in how many
ways can we give presents?

2.6) n boys and n girls go out to dance. Assume that only couples of mixed gender dance
with each other. In how many ways can they all dance simultaneously?

2.7) How many bits does 10100 have if written in base 2?

2.8) What is the number of 20-digit integers in which no two consecutive digits are the same?

2.9) Alice has 10 distinct balls. First she splits them into 2 piles. Then she picks one of the
piles with at least 2 balls, and split it into 2. She repeats this until each pile has only 1
ball.
(a) How many steps does this take?
Q
(b) The number of di↵erent ways in which she can carry out this procedure is 10 n
n=3 2 .

2.10) What is the number of ways to color n objects with 3 colors if every color must be used
at least once?

2.11) Find all positive integers a, b, and c for which

✓ ◆✓ ◆ ✓ ◆
a b a
=2 .
b c c

2.12) 20 persons are sitting around a table. How many ways can we choose 3 persons, no 2 of
whom are neighbors?

2.13) In a mathematical contest, three problems, A, B, C were posed. Among the participants
ther were 25 students who solved at least one problem each. Of all the contestants who
did not solve problem A, the number who solved B was twice the number who solved C.
The number of students who solved only problem A was one more than the number of
students who solved A and at least one other problem. Of all students who solved just
one problem, half did not solve problem A. How many students solved only problem B?

2.14) Find the ordinary generating functions of each of the following sequences, in simple,
closed form. In each case the sequence is defined for all n 0.
(a) an = ↵n2 + n + .
(b) an = a · bn + c · dn 1.

2.15) Find the conjugate partition to


(a) 33 22 12 ,
(b) 71,
(c) 5422 1.

2.16) A partition is said to be long rectangular if its Ferrers graph is rectangular with length
at least as great as height. Show that the number of partitions of n whose consecutive
parts di↵er by 2 is equal to the number of long rectangular partitions of n.

2.17) The partition function p(n) is odd if and only if the number of partitions of n into distinct
odd parts is odd.

2.18) What is the number of standard tableaux of shape two rows of equal length m?

2.19) The parts in a partition is super-distinct if every di↵erence is at least 2. The number of
partitions of n into super-distinct parts equals the number of partitions of n into distinct
parts such that every part is not 2.

2.20) The order of a permutation ⇡ is the smallest positive integer m such that ⇡ m is the
identity. Show that
(a) a permutation is of odd order if and only if each of its cycles is of odd length, and
(b) find the number of odd order permutations of [n].

2.21) For any set S of positive integers, let Pn (m, S) be the set of partitions of n into m parts
such that each part is in S, and Pn (m, S, d) the set of partitions in Pn (m, S) such that
each part appear at most d times. Then

XX Y 1
|Pn (m, S)|z m q n = ,
1 zq n
n 0m 0 n2S
XX Y1 z d+1 q (d+1)n
|Pn (m, S, d)|z m q n = .
1 zq n
n 0m 0 n2S

2.22) For any set S of positive integers, let p0n (m, S) be the number of partitions of n into m
distinct parts in S. Then

XX Y
p0n (m, S)z m q n = (1 + zq n ).
n 0m 0 n2S

2.23) Show that

2n + 3 + ( 1)n
pn (2) = .
4
2.24) The number of permutations of [n] with k 1 excedances is the Eulerian number A(n, k).
2.25) Suppose that A(0, 0) = 1 and A(n, 0) = 0 for n > 0. Then for all nonnegative integers n
and real numbers x,
Xn ✓ ◆
n x+n k
x = A(n, k) ,
n
k=1
Xn ✓ ◆
n x+k 1
x = A(n, k) .
n
k=0

2.26) Prove that the nth Eulerian polynomial has the alternative description
X
An (x) = (1 x)n+1 in xi .
i 0

2.27) Prove that


XX un 1 t
A(n, k)tk = .
n 0k 0
n! 1 teu(1 t)

2.28) How may ways of changing n fens into jiaos, yuans, and 100-yuans?

2.29) What statistics on integer partitions or compositions can you define? What statistics
on permutations can you define? From what perspective do you think they might be
interesting?

2.30) Enjoy the movie The Man Who Knew Infinity.

2.31) Show that the number dn of derangements on the set [n] satisfies

dn = ndn 1 + ( 1)n ,
dn = (n 1) dn 1 + dn 2 , and
X xn 1
dn = .
n! (1 x)ex
n 0

2.32) Let n, k 2 Z+ . For any partition ` n, let mk ( ) be the number of times k appears as a
part of , and let gk ( ) be the number of distinct parts of that appear at least k times.
For example, if = 323 12 , then m2 ( ) = 3 and g2 ( ) = 2. Show that
X X
mk ( ) = gk ( ).
`n `n

2.33) For each nonempty word w 2 A⇤ , there exists a unique primitive word x such that w 2 x⇤ .

2.34) Two nonempty word commute if and only if they are powers of the same word.

2.35) Let x, y 2 A⇤ . If two powers xp and y q have a common prefix of length at least |x| + |y|
gcd |x|, |y|, then x and y are powers of the same word.

2.36) In the definition of the infinite word of Thue-Morse, the word µ2n (a) are palindromes for
all n 1.
2.37) With the notations in the proof of Theorem 3.58: ⌫ is a square-free word such that neither
aba nor acbca is a factor of ⌫ if and only if (⌫) has no overlapping factor.

2.38) The number of Lyndon words of length n over an ordered alphabet A is


1X
µ(d)|A|n/d ,
n
d|n

where µ is the Möbius function defined on Z+ by


8
>
<1, if n = 1,
i
µ(n) = ( 1) , if n is the product of i distinct primes,
>
:
0, otherwise.
Chapter 4

Combinatorial numbers
and identities

This section is devoted to combinatorial numbers which appear frequently in modern research
in combinatorics.

Figure 4.1. The On-Line Encyclopedia of Integer Sequences (OEIS), also cited simply as
Sloane’s, is an online database of integer sequences.
1

1 1

1 2 1

1 3 3 1

1 4 6 4 1

1 5 10 10 5 1

1 6 15 20 15 6 1

1 7 21 35 35 21 7 1

Figure 4.2. The first lines of Pascal’s triangle.

4.1. Binomial coefficients

The binomial expansions of small degrees were known in the 13th century mathematical
works of Yang Hui and also Chu Shih-Chieh. Yang attributes the method to a much earlier
11th century text of Jia Xian, although those writings are now lost.
Definition 4.1. For n 2 R and k 2 Z, a binomial coefficient is a number of the form
✓ ◆
n nk
= ,
k k!
where
nk = n(n 1) · · · (n k + 1)
is called the Pochhammer symbol .

Note that for n, k 2 Z such that n 1, we have


✓ ◆
n
= 0, if k < 0 or k > n.
k
This fact allows us to simplify the sum notation as in the following occasion:
X n ✓ ◆ X ✓n ◆ X ✓n ◆
n
f (n, k) = f (n, k) = f (n, k),
k k k
k=0 k2Z k

where the range for k is Z by default, and the sum has essentially a finite number of nonzero
terms. We will adopt this convenience.

47
Figure 4.3. Yang Hui triangle (Pascal’s triangle) using rod numerals, as depicted in a
publication of Zhu Shijie in 1303 AD. Stolen from Wiki.

Theorem 4.2 (Binomial theorem). Let x, y 2 R and n 2 N. Then


X ✓n ◆
(4.1) (x + y)n = xk y n k .
k
k 0

Proof. The binomial coefficient nk is the number of ways of chooing k elements from [n].
This coincides with the interpretation that in the product

(x + y)n = (x + y)(x + y) · · · (x + y),

according to the distributive law, one chooses either x or y from each of the n factors x + y
and obtains a term of the form xk y n k by choosing exactly k terms x. ⇤

Around 1665, Newton generalized Theorem 4.2 to allow n 2 R, restricting that |x| < |y|
to gaurantee the convergence. The same generalization also applies to complex numbers n.
For example, taking n = 1, Eq. (4.1) reduces to the geometric series formula, valid for
|x| < 1:
1
= 1 x + x2 x3 + x4 x5 + · · · .
1+x
Taking n = 1/2, Eq. (4.1) reduces to
1 1 3 5 3 35 4 63 5
p =1 x + x2 x + x x + ··· .
1+x 2 8 16 128 256
Note that the infinite series in Eq. (4.1) reduces to a finite sum if n 2 N.
Theorem 4.3 (Chu-Vandermonde identity). Let n, m 2 C and k 2 N. Then
X ✓n◆✓ m ◆ ✓n + m◆
= .
j k j k
j

Proof. For positive integers n and m, one may show combinatorially as follows. Choosing
several elements from [n] and another several elements from the set {n + 1, n + 2, . . . , n + m}
so that in total k elements are chosen, is equivalent to choosing k elements from the set [n+m].
The two sides of the desired identity count the above two number of choices, respectively.
For n, m 2 C, algebraic proof works:
X ✓m + n ◆
xr = (1 + x)m+n
r
r 0
! !
X ✓m◆ X ✓n ◆
= xi xj
i j
i 0 j 0
X ✓m◆✓ n ◆
= xr .
k r k
r,k 0

Extracting the coefficient of xr , one obtains the desired identity. ⇤


Proposition 4.4. The binomail coefficients satisfy:
✓ ◆ ✓ ◆
n n
= ,
k n k
✓ ◆ ✓ ◆ ✓ ◆
n n 1 n 1
= + ,
k k k 1
✓ ◆ ✓ ◆
n n n 1
= ,
k k k 1
✓ ◆✓ ◆ ✓ ◆✓ ◆
a b a a c
= ,
b c c b c
X ✓n ◆
= 2n ,
k
k
X ✓n◆2 ✓2n◆
= ,
j n
j
X ✓ ◆
j n
( 1) = 0,
j
j
X ✓n ◆
2j = 3n ,
j
j
X ✓n◆✓ k ◆ ✓ n ◆
= 2n m
.
k m m
k

Proof. Omitted. But please consider a combinatorical proof and an algebraic proof for each
of these identities. ⇤
Proposition 4.5 (Hockey-stick identity).
✓ ◆ ✓ ◆ ✓ ◆ ✓ ◆
n n 1 r n+1
+ + ··· + = .
r r r r+1

Hint. Consider an algebraic proof and a combinatorial proof. ⇤


Proposition 4.6. For any 0  k  n,
k
X ✓ ◆ ✓ ◆
j n k n 1
( 1) = ( 1) .
j k
j=0

Proof. We show it by using the telescoping method .


✓ ◆ ✓ ◆ ✓ ◆ ✓ ◆ ✓ ◆
n n n n k n
+ + · · · + ( 1)
0 1 2 3 k
✓ ◆ ✓ ◆ ✓ ◆ ✓ ◆ ✓ ◆
n 1 n 1 n 1 n 1 n 1
= + + +
0 0 1 1 2
✓ ◆ ✓ ◆
n 1 n 1
· · · + ( 1)k +
k 1 k
✓ ◆
n 1
= ( 1)k .
k

Definition 4.7. Let n 2 N. For any nonnegative integers k1 , k2 , . . . , km whose sum is n, the
integer ✓ ◆
n n!
=
k 1 , k 2 , . . . , km k1 ! k2 ! · · · km !
is called a multinomial coefficient. Denote by
✓✓ ◆◆
n
k
the number of multisets {a1 , . . . , ak } such that ai 2 [n]. It is called a multiset coefficient in
some literature.

Replacing n by n in Eq. (4.1), one obtains the generating function for the multiset
coefficients:
1 X ⇣ ✓n ◆ ⌘
= xk .
(1 x)n k
k 0
Theorem 4.8 (Multinomial theorem). Let n 2 N. Then
m
!n ✓ ◆Y
m
X X n k
xj = xj j .
k 1 , k 2 , . . . , km
j=1 k +k +···+k =n
1 2 m j=1
k1 ,...,km 2N

Proposition 4.9. The number of ways to distribute n identical pennies to k children is


✓✓ ◆◆
n+1
.
k 1
If we require additionally that each child gets at least one penny, then the number of ways is
✓ ◆
n 1
.
k 1

Hint. Use the method of stars and bars. ⇤


Theorem 4.10. For n 2 Z and k 2 N,
✓✓ ◆◆ ✓ ◆
n n+k 1
= .
k k

Proof. By Proposition 4.9, or by using the method of stars and bars, the number of positive
integer solutions (x1 , x2 , . . . , xk ) to the equation
x1 + x2 + · · · + xk = n + k
is n+k 1
k 1 . Let yi = xi 1. Then a positive integer solution (x1 , . . . , xk ) is in a bijection with
a nonnegative integer solution (y1 , . . . , yk ) to the equation
y1 + y2 + · · · + yk = n.
Therefore, by Proposition 4.9,
✓✓ ◆◆ ✓ ◆
n+1 n+k 1
= .
k 1 k 1
Replacing n by n 1 and replacing k by k + 1 in the above equation, one obtains the desired
formula. ⇤

4.2. Catalan numbers

Definition 4.11. The nth Catalan number is


✓ ◆
1 2n
Cn = .
n+1 n

Eugène Charles Catalan studied Catalan numbers in [17] in the attempt of figuring out
the solution of the equation
Pn+1 = Pn + Pn 1 P3 + Pn 2 P4 + · · · + P4 Pn 2 + P3 Pn 1 + Pn .
Most material of Section 4.2 come from Stanley [115]. We start from the definition as the
historically first combinatorial interpretation.
Let Pn denote a convex n-gon in the plane. A triangulation of Pn is a set of n 3 diagonals
of Pn which do not cross in their interiors. It follows easily that these diagonals partition the
interior of Pn into n 2 triangles. Define the nth Catalan number Cn to be the number of
triangulations of Pn+2 . See Fig. 4.4 for the triangulations for small n.
The first Catalan numbers (OEIS A000108) are listed in Table 4.1. Combinatorially, it
Table 4.1. The Catalan numbers Cn for n = 0, 1, . . . , 10.

n 0 1 2 3 4 5 5 7 8 9 10
Cn 1 1 2 5 14 42 132 429 1430 4862 16796

Figure 4.4. The triangulations of the triangle, square, pentagon and hexagon. Stolen
from [115, Page 2].

is easy to show that for n 0,


n
X
(4.2) Cn+1 = Ci Cn i .
i=0
Multiplying Eq. (4.2) by z n and sum over n 0, we find the generating function
X
C(z) = Cn z n
n 0
satisfies the equation
C(z) 1
= C(z)2 ,
z
from which we can solve out p
1 1 4z
C(z) = .
2z
Using Theorem 4.2, we can expand
✓ X ✓1/2◆ ◆
1
Cn = [z n ]C(z) = [z n ] 1 ( 4z)n
2z n
n 0
✓ ◆ ✓ ◆
1 1/2 1 2n
= ( 4)n+1 = .
2 n+1 n+1 n
Asymptotically, the Catalan numbers grow as
4n
Cn ⇠ p .
n3/2 ⇡

We now introduce more combinatorial interpretations for Catalan numbers, which are
considered most basic.
Theorem 4.12. The Catalan number Cn counts the following:

(1) Triangulations of a convex polygon with n + 2 vertices.


(2) Binary trees with n vertices.
(3) Plane trees with n + 1 vertices.
(4) Ballot sequences of length 2n.
(5) Parenthesizations (or bracketings) of a string of n + 1 x’s subject to a nonassociative
binary operation.
(6) Dyck paths of length 2n.
Definition 4.13. A binary tree is defined recursively as follows. The empty set ; is a binary
tree. Otherwise a binary tree has a root vertex v, a left subtree T1 , and a right subtree T2 ,
both of which are binary trees. We also call the root of T1 (if T1 is nonempty) the left child
and the root of T2 (if T2 is nonempty) the right child of the root v.

We draw a binary tree by putting the root v at the top, the left subtree T1 below and to
the left of v, and the right subtree T2 below and to the right of v, with an edge drawn from
v to the root of each nonempty Ti . Fig. 4.5 shows the five binary trees with 3 vertices.

Figure 4.5. The five binary trees with 3 vertices. Stolen from [115, Page 6].

Definition 4.14. A plane tree (also called an ordered tree) P may be defined recursively as
follows. One specially designated vertex v is called the root of P . Thus plane trees cannot be
empty. Then either P consists of the single vertex v, or else it has a sequence (P1 , . . . , Pm )
of subtrees Pi , 1  i  m, each of which is a plane tree. Thus the subtrees of each vertex are
linearly ordered.

When drawing such trees, these subtrees are written in the order left-to-right. The root v
is written on the top, with an edge drawn from v to the root of each of its subtrees. Fig. 4.6
shows the five plane trees with 4 vertices.
Definition 4.15. A ballot sequence of length 2n is a sequence with n each of 1’s and 1’s
such that every partial sum is nonnegative.
Figure 4.6. The five plane trees with 4 vertices. Stolen from [115, Page 6].

The five ballot sequences of length 6 (abbreviating 1 by just ) are given by


111 , 11 1 , 11 1 , 1 11 , 1 1 1 .
Definition 4.16. A parenthesization or bracketing of a string of n + 1 x’s consists of the
insertion of n left parentheses and n right parentheses that define n binary operations on the
string.

The five ways to parenthesize a string of 4 x’s are as follows:


x x(xx) , x (xx)x , (xx)(xx), x(xx) x, and (xx)x x.

Definition 4.17. Let S ✓ Zd . A lattice path in Zd of length k with steps in S is a sequence


v0 , v1 , . . . , vk 2 Zd such that each consecutive di↵erence vi vi 1 lies in S. Common steps
include the east step E = (1, 0), the northeast step N E = (1, 1), the southeast step SE =
(1, 1), and their variates. Below are some special lattice paths.

(1) A Dyck path of length 2n is a lattice path in Z2 from (0, 0) to (2n, 0) with steps N E
and SE, with the additional condition that the path never passes below the x-axis.

(2) A Motzkin path of length 2n is a lattice path in Z2 from (0, 0) to (2n, 0) with steps
N E, SE, and E, with the additional condition that the path never passes below the
x-axis.

(3) A Schröder path of length 2n is a lattice path in Z2 from (0, 0) to (2n, 0) with steps
N E, SE, and E 0 = (2, 0), with the additional condition that the path never passes
below the x-axis.

Fig. 4.7(a) shows the five Dyck paths of length 6. A trivial but useful variant of Dyck
paths (sometimes also called a Dyck path) is obtained by replacing the step (1, 1) with (1, 0)
and (1, 1) with (0, 1). In this case we obtain lattice paths from (0, 0) to (n, n) with steps
(1, 0) and (0, 1), such that the path never rises above the line y = x. See Fig. 4.7(b) for the
case n = 3.
Moreover, the Catalan number Cn counts the number of standard tableaux of shape
(n, n). Stanley’s book [115] contains 214 combinatorial interpretations of Catalan numbers.
Pak contributed an extensive history on Catalan numbers as its Appendix B, from where
one may found Catalan numbers have historical connections with Euler, Goldbach, Segner,
Kotelnikow, Fuß, Liouville, Pfa↵, Kirkman, Cayley, Taylor, Schröder, Lucas and so on, as
well as the following quotes.
Figure 4.7. The five Dyck paths of length 6. Stolen from [115, Page 8].

It is not exaggerated to say that the Catalan numbers are the most promi-
nent sequence in combinatorics. —Kauers and Paule (2011)

The Catalan numbers are one of the most important sequences of combi-
natorial numbers, with a large range of occurrences in apparently di↵erent
counting problems. —Cameron (2013)

Catalan numbers are probably the most frequently occurring combinatorial


numbers after the binomial coefficients. —Neil Sloane and Simon Plou↵e
(1995)
Theorem 4.18. The n ⇥ n Hankel matrix whose (i, j)-entry is the Catalan number Ci+j 2
has determinant 1. The n ⇥ n Hankel matrix whose (i, j)-entry is the Catalan number Ci+j 1
has determinant 1.

For example,
0 1 0 1
1 1 2 5 1 2 5 14
B 1 2 5 14 C B 2 5 14 42 C
det B C B
@ 2 5 14 42 A = det @
C = 1.
5 14 42 132 A
5 14 42 132 14 42 132 429

4.3. Fibonacci numbers

The Fibonacci numbers (OEIS A000045) are named after the Italian mathematician Fi-
bonacci, whose name was made up in 1838 by the Franco-Italian historian Guillaume Libri
and is short for filius Bonacci (“son of Bonacci”). Fibonacci was considered to be “the most
talented Western mathematician of the Middle Ages”.
Definition 4.19. The Fibonacci numbers Fn are defined by F0 = 0, F1 = 1 and
(4.3) Fn = Fn 1 + Fn 2 for n 2.

The first Fibonacci numbers are listed in Table 4.2. The generating function of the
Fibonacci numbers is X z
Fn z n = .
1 z z2
n 0
Table 4.2. The Fibonacci numbers Fn for n = 0, 1, . . . , 16.

n 0 1 2 3 4 5 5 7 8 9 10 11 12 13 14 15 16
Fn 0 1 1 2 3 5 8 13 21 34 55 89 144 233 377 610 987

Binet’s formula expresses the nth Fibonacci number as


'n n 'n n
Fn = = p ,
5
where p
1+ 5
'= ⇡ 1.6180339887...
2
is the golden ratio, and
p
1 5 1
= =1 '= ⇡ 0.6180339887....
2 '
Another formula with the aid of the floor function is

'n 1
Fn = p + .
5 2
We also have the Cassini’s identity
Fn2 Fn+1 Fn 1 = ( 1)n 1
.

4.4. Bell numbers

The Bell numbers (OEIS A000110) are named after Eric Temple Bell.
Definition 4.20. The nth Bell number Bn is the number of partitioning the set [n].

From definition, we know that Bn is the sum of the Stirling numbers S(n, k) of the second
kind, namely
Xn
Bn = S(n, k)
k=0
The first few Bell numbers are listed in Table 4.3. Choosing a subset from [n 1] to form a
Table 4.3. The Bell numbers Bn for n = 0, 1, . . . , 9.

n 0 1 2 3 4 5 5 7 8 9
Bn 1 1 2 5 15 52 203 877 4140 21147

block together with the element n, one finds


X ✓n 1

Bn = Bk .
k
k
The exponential generating function of the Bell numbers is
X xn x
Bn = ee 1 .
n!
n 0
Extracting the coefficient of xn from the right hand side, we obtain Dobiński’s formula, see
Rota [102] for another proof by considering moments in Poisson distribution.
Theorem 4.21 (Dobiński’s formula).
1 X kn
Bn = .
e k!
k 0

Asymptotically, the Bell numbers grow as


✓ ◆n+1/2 ✓ ◆
1 n n
Bn ⇠ p exp n 1 ,
n W (n) W (n)
where W (n) is the Lambert W function which is defined by
W (n)eW (n) = n.

There is a huge number of combinatorial numbers that are not mentioned in our class, such
as Motzkin numbers, Schröder numbers, Padovan numbers, Narayana numbers, Genocchi
numbers, . . . .

4.5. Topics for an undergraduate thesis


(1) Survey on Catalan numbers — A combinatorial view of point.
Exercises 4

3.1) Find 3 more kinds of combinatorial numbers and try understand their history and back-
groud.

3.2) Fix n. For which value of k, the di↵erence


✓ ◆ ✓ ◆
n n
k+1 k
is the largest?

3.3) In a city, the North-South streets are called Street 1, Street 2, ..., Street 20, and the
East-West streets are called Avenue 1, Avenue 2, ..., Avenue 10. What is the minimum
number of blocks you have to walk to get from the corner of Street 1 and Avenue 1 to
the corner of Street 20 and Avenue 10? In how many ways can you get there walking
this minimum number of blocks?

3.4) Prove that


✓ ◆ r
n 2 n
⇠ 2 .
n/2 ⇡n
n n
What about n/3 and n/k ?

3.5) Suppose that you want to choose a (2k + 1)-element subset of the set [n]. You decide to
do this by choosing first the middle element, then the k elements to its left, then the k
elements to its right. Formulate the combinatorial identity you get from this.

3.6) The Catalan numbers grow as


4n
Cn ⇠ p .
n3/2 ⇡
3.7) The Catalan numbers have the integral representation
Z 4 nr
x 4 x
Cn = .
0 2⇡ x
3.8) How many ways can you place n rooks on a chessboard so that no two attack each other?
How many ways can you do this if you have 4 wooden and 4 marble rooks? How many
ways can you do this if all the 8 rooks are distinct?

3.9) What if the chessboard is replaced by the Chinese chessboard in Exercise 3.8)?

3.10) In how many ways can you distribute n pennies to k children if each child is supposed to
get at least 2?

3.11) Two candidates A and B are running in an election. There are 2n voters who vote
sequentially for one of the two candidates. At the end each candidate receives n votes.
What is the probability that A never trails B in the voting and both candidates receive
n votes?
3.12) Two candidates A and B are running in an election. There are n voters who vote sequen-
tially for one of the two candidates. What is the probability that A never trails B in the
voting?

3.13) Show Theorem 4.12.

3.14) The Fibonacci numbers satisfy the following identities:


(a) Catalan’s identity
Fn2 Fn+r Fn r = ( 1)n r Fr2 ,
(b) Vajda’s identity
Fn+i Fn+j Fn Fn+i+j = ( 1)n Fi Fj ,
(c) d’Ocagne’s identity
Fm Fn+1 Fm+1 Fn = ( 1)n Fm n.

3.15) The Fibonacci numbers satisfy the divisibility property


gcd(Fm , Fn ) = Fgcd(m,n) .

3.16) Any 3 consecutive Fibonacci numbers are pairwise coprime.

3.17) Evaluate the determinant


C0 C1 C2 C3
C1 C0 C1 C2
,
C2 C1 C0 C1
C3 C2 C1 0
where Cn is the nth Catalan number. Do you have an answer for the general determinants
of square matrix of size n? How about change Cn to the nth Fibonacci number Fn ?

3.18) How many subsets does the set [n] have that contain no two consecutive integers?

3.19) Show that for even integer n, the Fibonacci numbers Fn satisfy
Fn = Fn 1 + Fn 3 + · · · + F1 .

3.20) Show that if n is a multiple of 4, then the Fibonacci number Fn is a multiple of 3.

3.21) The Bell number Bn the Bell number counts the number of partitions of the set [n + 1]
into blocks of nonconsecutive integers.

3.22) No Bell number is divisible by 8.


Select solutions to
Exercises 4

3.12) The probability is ✓ ◆


n
2 n.
bn/2c
This problem relates to the reflection princple in the theory of lattice paths.
Consider the set Wn of length-n words over the alphabet {A, B}, which correpond to
the voting sequences in a natrual way. It is clear that
|Wn | = 2n .
Let w 2 W , where [
W = Wn .
n 0
Denote by a(w) the number of occurrences of the letter A in w, and by b(w) the number
of occurrences of B in w. Define the height of w to be
h(w) = a(w) b(w).
It suffices to count the words whose every prefix has a nonnegative height.
Let Wn,k be the set of words w 2 Wn such that b(w) = k. Then
✓ ◆
n
|Wn,k | = .
k
Let Xn,k be the set of words in Wn,k that has a prefix of negative height. We shall
establish a bijection
: Xn,k ! Wn, k 1 .
If this could be done, then
|Xn,k | = |Wn, k 1 |,
and the number of voting sequences in which A never trials B would be
bn/2c
X bn/2c
X ✓ ◆
n
|Wn,k | |Xn,k | = |Wn,k | |Wn, k 1 | = Wn,bn/2c = .
bn/2c
k=0 k=0
In fact, let x 2 Xn,k . Let ↵ be the shortest prefix of x with height 1, and write
x=↵ .
Define
(x) = ↵ ,
where ↵ the word obtained from ↵ by exchanging its letters A and B. Since
b( (x)) = a(↵) + b( ) = b(↵) 1 + b( ) = b(x) 1=k 1,
we find (x) 2 Wn, k 1 . Conversely, let w 2 Wn, k 1. Denote by ↵0 the shortest prefix
of w with height 1. Then
w = ↵0 0
for some word 0 2 W . Define
(w) = ↵0 0 .
It is easy to verify that (w) 2 Xn,k . Moreover, the maps and are inverses each
other, namely,
( (x)) = x and ( (w)) = w.
This completes the solution.
Bibliography

[1] M. Aigner and G. M. Ziegler. Proofs from THE BOOK. Springer-Verlag Berlin Heidel-
berg, 6th edition, 2018.

[2] R. B. Allan and R. Laskar. On domination and independent domination numbers of a


graph. Discrete mathematics, 23(2):73–76, 1978.

[3] G. E. Andrews. Euler’s pentagonal number theorem. Math. Mag., 56(5):279–284,


1983. ISSN 0025-570X. doi: 10.2307/2690367. URL https://mathscinet.ams.org/
mathscinet-getitem?mr=720648.

[4] G. E. Andrews and K. Eriksson. Integer Partitions. Cambridge University Press, Cam-
bridge, 2004. ISBN 0-521-84118-6; 0-521-60090-1. doi: 10.1017/CBO9781139167239.
URL https://mathscinet.ams.org/mathscinet-getitem?mr=2122332.

[5] W. Anglin. The square pyramid puzzle. The American Mathematical Monthly, 97(2):
120–124, 1990.

[6] A. Avron and N. Dershowitz. Cayley’s formula: A page from the book. The American
Mathematical Monthly, 123(7):699–700, 2016.

[7] J. Barwise. Handbook of Mathematical Logic. 1977.

[8] G. Baxter. On fixed points of the composite of commuting functions. Proc. Amer.
Math. Soc., 15:851–855, 1964. ISSN 0002-9939. doi: 10.2307/2034894. URL https:
//mathscinet.ams.org/mathscinet-getitem?mr=184217.

[9] E. A. Bender and J. R. Goldman. On the applications of Möbius inversion in combinato-


rial analysis. Amer. Math. Monthly, 82(8):789–803, 1975. ISSN 0002-9890. doi: 10.2307/
2319793. URL https://mathscinet.ams.org/mathscinet-getitem?mr=376360.

[10] J. Berstel and D. Perrin. The origins of combinatorics on words. European J. Combin.,
28(3):996–1022, 2007. URL https://www.sciencedirect.com/science/article/
pii/S0195669805001629.

63
[11] M. Bóna. Combinatorics of Permutations. Discrete Mathematics and its Applications
(Boca Raton). CRC Press, Boca Raton, FL, 2nd edition, 2012. ISBN 978-1-4398-5051-
0. doi: 10.1201/b12210. URL https://mathscinet.ams.org/mathscinet-getitem?
mr=2919720. With a foreword by Richard Stanley.

[12] N. Bonichon, M. Bousquet-Mélou, and E. Fusy. Baxter permutations and plane bipolar
orientations. In The International Conference on Topological and Geometric Graph
Theory, volume 31 of Electron. Notes Discrete Math., pages 69–74. Elsevier Sci. B. V.,
Amsterdam, 2008. doi: 10.1016/j.endm.2008.06.011. URL https://mathscinet.ams.
org/mathscinet-getitem?mr=2571108.

[13] O. D. Byer, D. L. Smeltzer, and K. L. Wantz. Journey Into Discrete Mathematics,


volume 41. Amer. Math. Soc., 2018.

[14] O. D. Byer, D. L. Smeltzer, and K. L. Wantz. Journey into Discrete Mathematics,


volume 41 of AMS/MAA TEXTBOOKS. MAA Press, 2018. ISBN 9781470449124.

[15] E. R. Canfield and C. Pomerance. On the problem of uniqueness for the maximum
Stirling number(s) of the second kind. Integers, 2:A1, 13, 2002. ISSN 1553-1732. URL
https://mathscinet.ams.org/mathscinet-getitem?mr=1876176.

[16] E. R. Canfield and C. Pomerance. Corrigendum to: “On the problem of uniqueness
for the maximum Stirling number(s) of the second kind” [Integers 2 (2002), A1, 13 pp.
(electronic); mr1876176]. Integers, 5(1):A9, 1, 2005. ISSN 1553-1732. URL https:
//mathscinet.ams.org/mathscinet-getitem?mr=2192227.

[17] E. Catalan. Note sur une équation aux di↵érences finies. Journal de mathématiques
pures et appliquées, 3:508–516, 1838.

[18] G. Chartrand and P. Zhang. Chromatic Graph Theory. Discrete Mathematics and its
Applications (Boca Raton). CRC Press, Boca Raton, FL, 2009. ISBN 978-1-58488-800-
0. URL https://mathscinet.ams.org/mathscinet-getitem?mr=2450569.

[19] T. Y. Chow. The path-cycle symmetric function of a digraph. Advances in Mathematics,


118(1):71–98, 1996. ISSN 0001-8708. doi: https://doi.org/10.1006/aima.1996.0018.
URL http://www.sciencedirect.com/science/article/pii/S0001870896900183.

[20] F. Chung and R. Graham. Erdős on Graphs. A K Peters, Ltd., Wellesley, MA,
1998. ISBN 1-56881-079-2; 1-56881-111-X. URL https://mathscinet.ams.org/
mathscinet-getitem?mr=1601954. His legacy of unsolved problems.

[21] F. R. K. Chung, R. L. Graham, V. E. Hoggatt, Jr., and M. Kleiman. The num-


ber of Baxter permutations. J. Combin. Theory Ser. A, 24(3):382–394, 1978. ISSN
0097-3165. doi: 10.1016/0097-3165(78)90068-7. URL https://mathscinet.ams.org/
mathscinet-getitem?mr=491652.

[22] V. Chvátal. A combinatorial theorem in plane geometry. Journal of Combinatorial


Theory, Series B, 18(1):39–41, 1975.

[23] F. N. David and M. G. Kendall. Tables of symmetric functions. I. Biometrika, 36:


431–449, 1949. ISSN 0006-3444. doi: 10.1093/biomet/36.3-4.431. URL https://
mathscinet.ams.org/mathscinet-getitem?mr=33788.

[24] de la Vallée-Poussin. Recherches analytiques sur la théorie des nombres premiers, i–iii.
Ann. Soc. Sci. Bruxelles, 20:183–256, 1896.

[25] R. Diestel. Graph theory. Graduate Texts in Mathematics, 2017. ISSN 2197-5612. doi:
10.1007/978-3-662-53622-3. URL http://dx.doi.org/10.1007/978-3-662-53622-3.

[26] P. L. Dirichlet. Beweis des satzes, dass jede unbegrenzte arithmetische progression,
deren erstes glied und di↵erenz ganze zahlen ohne gemeinschaftlichen factor sind, un-
endlich viele primzahlen enthält. Abhandlungen der Königlich Preussischen Akademie
der Wissenschaften, 45:81, 1837.

[27] P. Doubilet. On the foundations of combinatorial theory. VII: Symmetric functions


through the theory of distribution and occupancy. Studies in Applied Mathematics,
51(4):377–396, 1972. doi: https://doi.org/10.1002/sapm1972514377. URL https://
onlinelibrary.wiley.com/doi/abs/10.1002/sapm1972514377.

[28] J.-P. Duval. Mots de Lyndon et périodicité. RAIRO - Theoretical Informatics and
Applications - Informatique Théorique et Applications, 14(2):181–191, 1980. URL http:
//www.numdam.org/item/ITA_1980__14_2_181_0/.

[29] T. E. Easterfield. A combinatorial algorithm. Journal of the London Mathematical


Society, 1(3):219–226, 1946.

[30] Ő. Eǧecioǧlu and J. Remmel. The monomial symmetric functions and the frobe-
nius map. Journal of Combinatorial Theory, Series A, 54(2):272–295, 1990. ISSN
0097-3165. doi: https://doi.org/10.1016/0097-3165(90)90035-U. URL https://www.
sciencedirect.com/science/article/pii/009731659090035U.

[31] R. Ehrenborg and E. Steingrı́msson. The excedance set of a permutation. Adv. in


Appl. Math., 24(3):284–299, 2000. ISSN 0196-8858. doi: 10.1006/aama.1999.0671.
URL https://mathscinet.ams.org/mathscinet-getitem?mr=1751147.

[32] S. Elizalde and M. Noy. Consecutive patterns in permutations. volume 30, pages 110–
125. 2003. doi: 10.1016/S0196-8858(02)00527-4. URL https://mathscinet.ams.org/
mathscinet-getitem?mr=1979785. Formal power series and algebraic combinatorics
(Scottsdale, AZ, 2001).

[33] N. D. Elkies. On A4 + B 4 + C 4 = D4 . Mathematics of Computation, pages 825–835,


1988.

[34] S. S. Epp. Discrete Mathematics with Applications. Cengage learning, 2010.

[35] S. S. Epp. Discrete Mathematics: Introduction to Mathematical Reasoning. Nelson


Education, 2014.

[36] P. Erdős. On a new method in elementary number theory which leads to an elementary
proof of the prime number theorem. Proceedings of the National Academy of Sciences
of the United States of America, 35(7):374, 1949.
[37] P. Erdős. Az 1/x1 + 1/x2 + ... + 1/xn = a/b egyenlet egész számú megoldásairól (on a
diophantine equation). Matematiki Lapok.(in Hungarian), 1:192–210, 1950.

[38] P. Erdős. On a problem in graph theory. The Mathematical Gazette, 47(361):220–223,


1963.

[39] L. Euler. Solutio problematis ad geometriam situs pertinentis. Commentarii academiae


scientiarum Petropolitanae, pages 128–140, 1741.

[40] S. Fisk. A short proof of chvátal’s watchman theorem. Journal of Combinatorial


Theory, Series B, 24(3):374, 1978. ISSN 0095-8956. doi: https://doi.org/10.1016/
0095-8956(78)90059-X. URL http://www.sciencedirect.com/science/article/
pii/009589567890059X.

[41] P. Flajolet and R. Sedgewick. Analytic Combinatorics. Cambridge University Press,


Cambridge, 2009. ISBN 978-0-521-89806-5. doi: 10.1017/CBO9780511801655. URL
https://mathscinet.ams.org/mathscinet-getitem?mr=2483235.

[42] S. Fomin and J.-Y. Thibon. Inverting the Frobenius map. Algebra i Analiz,
10(3):183–192, 1998. ISSN 0234-0852. URL https://mathscinet.ams.org/
mathscinet-getitem?mr=1628046.

[43] J. S. Frame, G. d. B. Robinson, and R. M. Thrall. The hook graphs of the


symmetric groups. Canad. J. Math., 6:316–324, 1954. ISSN 0008-414X. doi:
10.4153/cjm-1954-030-1. URL https://mathscinet.ams.org/mathscinet-getitem?
mr=62127.

[44] D. S. Franzblau and D. Zeilberger. A bijective proof of the hook-length formula. J.


Algorithms, 3(4):317–343, 1982. ISSN 0196-6774. doi: 10.1016/0196-6774(82)90029-3.
URL https://mathscinet.ams.org/mathscinet-getitem?mr=681218.

[45] W. Fulton. Young Tableaux, volume 35 of London Mathematical Society Student Texts.
Cambridge University Press, Cambridge, 1997. ISBN 0-521-56144-2; 0-521-56724-6.
URL https://mathscinet.ams.org/mathscinet-getitem?mr=1464693. With appli-
cations to representation theory and geometry.

[46] H. Furstenberg. On the infinitude of primes. American Mathematical Monthly, 62:353,


1955.

[47] J. Gallier. Mathematical Reasoning, Proof Principles, and Logic. Universitext. Springer,
2011.

[48] V. Gasharov. A short proof of the littlewood–richardson rule. European Jour-


nal of Combinatorics, 19(4):451–453, 1998. ISSN 0195-6698. doi: https://doi.org/
10.1006/eujc.1998.0212. URL https://www.sciencedirect.com/science/article/
pii/S0195669898902128.

[49] C. F. Gauß. Arithmeticae, Disquisitiones. 1801.

[50] D. Goldfeld. Beyond the last theorem. Math Horizons, 4(1):26–34, 1996.
[51] F. Göring. Short proof of menger’s theorem. Discrete Mathematics, 219(1-3):295–296,
2000.

[52] B. Green and T. Tao. The primes contain arbitrarily long arithmetic progressions. Ann.
of Math. (2), 167(2):481–547, 2008. ISSN 0003-486X. doi: 10.4007/annals.2008.167.481.
URL https://mathscinet.ams.org/mathscinet-getitem?mr=2415379.

[53] C. Greene. An extension of Schensted’s theorem. Advances in Math., 14:254–265, 1974.


ISSN 0001-8708. doi: 10.1016/0001-8708(74)90031-0. URL https://mathscinet.ams.
org/mathscinet-getitem?mr=354395.

[54] J. Hadamard. Sur la distribution des zéros de la fonction ⇣(s) et ses conséquences
arithmétiques. Bull. Soc. Math. France, 24:199–220, 1896. ISSN 0037-9484. URL
https://mathscinet.ams.org/mathscinet-getitem?mr=1504264.

[55] P. Hall. On representatives of subsets. J. London Math. Soc, 10(1):26–30, 1935.

[56] P. Hall. The eulerian functions of a group. The Quarterly Journal of Mathematics, (1):
134–151, 1936.

[57] P. Hall. On representatives of subsets. In Classic Papers in Combinatorics, pages 58–62.


Springer, 2009.

[58] P. R. Halmos and H. E. Vaughan. The marriage problem. In Classic Papers in Com-
binatorics, pages 146–147. Springer, 2009.

[59] J. M. Hammersley. A few seedlings of research. In Proceedings of the Sixth Berke-


ley Symposium on Mathematical Statistics and Probability (Univ. California, Berke-
ley, Calif., 1970/1971), Vol. I: Theory of statistics, pages 345–394, 1972. URL
https://mathscinet.ams.org/mathscinet-getitem?mr=0405665.

[60] F. Harary. On the reconstruction of a graph from a collection of subgraphs. In Theory


of Graphs and its Applications (Proc. Sympos. Smolenice, 1963), pages 47–52. Publ.
House Czechoslovak Acad. Sci., Prague, 1964. URL https://mathscinet.ams.org/
mathscinet-getitem?mr=0175111.

[61] G. H. Hardy and E. M. Wright. An introduction to the theory of numbers. Oxford


University Press, Oxford, 6th edition, 2008. ISBN 978-0-19-921986-5. URL https:
//mathscinet.ams.org/mathscinet-getitem?mr=2445243. Revised by D. R. Heath-
Brown and J. H. Silverman, With a foreword by Andrew Wiles.

[62] J. Huh. Milnor numbers of projective hypersurfaces and the chromatic polynomial of
graphs. Journal of the American Mathematical Society, 25(3):907–927, 2012.

[63] C. Jongsma. Introduction to Discrete Mathematics via Logic and Proof. Undergraduate
Texts in Mathematics. Springer Nature, 2019.

[64] D. M. Kane. Improved bounds on the number of ways of expressing t as a binomial


coefficient. Integers, 7:A53, 7, 2007. ISSN 1553-1732. doi: 10.5391/ijfis.2007.7.1.007.
URL https://mathscinet.ams.org/mathscinet-getitem?mr=2373115.
[65] S. Kitaev. Introduction to partially ordered patterns. Discrete Appl. Math., 155(8):
929–944, 2007. ISSN 0166-218X. doi: 10.1016/j.dam.2006.09.011. URL https://
mathscinet.ams.org/mathscinet-getitem?mr=2319869.

[66] S. Kitaev. Patterns in Permutations and Words. Monographs in Theoretical Computer


Science. An EATCS Series. Springer, Heidelberg, 2011. ISBN 978-3-642-17332-5; 978-
3-642-17333-2. doi: 10.1007/978-3-642-17333-2. URL https://mathscinet.ams.org/
mathscinet-getitem?mr=3012380. With a foreword by Je↵rey B. Remmel.

[67] C. Kostka. Über den zusammenhang zwischen einigen formen von symmetrischen funk-
tionen. Crelle’s Journal, 93:89–123, 1882. URL http://gdz.sub.uni-goettingen.
de/no_cache/dms/load/img/?IDDOC=259790.

[68] A. Kulikauskas and J. Remmel. Lyndon words and transition matrices between ele-
mentary, homogeneous and monomial symmetric functions. The Electronic Journal of
Combinatorics, 13(1), Feb. 2006. URL https://doi.org/10.37236%2F1044.

[69] C. Kuratowski. Sur le problème des courbes gauches en topologie. Fundamenta Math-
ematicae, 15(1):271–283, 1930. URL http://eudml.org/doc/212352.

[70] D. Kurepa. On the left factorial function !n. Math. Balkanica, 1:147–153, 1971. ISSN
0350-2007. URL https://mathscinet.ams.org/mathscinet-getitem?mr=286736.

[71] Z. Lin and D. Kim. A sextuple equidistribution arising in pattern avoidance. J. Combin.
Theory Ser. A, 155:267–286, 2018. ISSN 0097-3165. doi: 10.1016/j.jcta.2017.11.009.
URL https://mathscinet.ams.org/mathscinet-getitem?mr=3741429.

[72] D. E. Littlewood and A. R. Richardson. Group characters and algebra. Philo-


sophical Transactions of the Royal Society of London. Series A, Containing Papers
of a Mathematical or Physical Character, 233:99–141, 1934. ISSN 02643952. URL
http://www.jstor.org/stable/91293.

[73] M. Lothaire. Combinatorics on words. Cambridge Mathematical Library. Cam-


bridge University Press, Cambridge, 1997. ISBN 0-521-59924-5. doi: 10.1017/
CBO9780511566097. URL https://mathscinet.ams.org/mathscinet-getitem?mr=
1475463. With a foreword by Roger Lyndon and a preface by Dominique Perrin, Cor-
rected reprint of the 1983 original, with a new preface by Perrin.

[74] M. Lothaire. Algebraic combinatorics on words, volume 90 of Encyclopedia of Math-


ematics and its Applications. Cambridge University Press, Cambridge, 2002. ISBN
0-521-81220-8. doi: 10.1017/CBO9781107326019. URL https://mathscinet.ams.
org/mathscinet-getitem?mr=1905123. A collective work by Jean Berstel, Dominique
Perrin, Patrice Seebold, Julien Cassaigne, Aldo De Luca, Ste↵ano Varricchio, Alain
Lascoux, Bernard Leclerc, Jean-Yves Thibon, Veronique Bruyere, Christiane Frougny,
Filippo Mignosi, Antonio Restivo, Christophe Reutenauer, Dominique Foata, Guo-Niu
Han, Jacques Desarmenien, Volker Diekert, Tero Harju, Juhani Karhumaki and Woj-
ciech Plandowski, With a preface by Berstel and Perrin.

[75] M. Lothaire. Applied combinatorics on words, volume 105 of Encyclopedia of Math-


ematics and its Applications. Cambridge University Press, Cambridge, 2005. ISBN
978-0-521-84802-2; 0-521-84802-4. doi: 10.1017/CBO9781107341005. URL https:
//mathscinet.ams.org/mathscinet-getitem?mr=2165687. A collective work by Jean
Berstel, Dominique Perrin, Maxime Crochemore, Eric Laporte, Mehryar Mohri, Na-
dia Pisanti, Marie-France Sagot, Gesine Reinert, Sophie Schbath, Michael Waterman,
Philippe Jacquet, Wojciech Szpankowski, Dominique Poulalhon, Gilles Schae↵er, Ro-
man Kolpakov, Gregory Koucherov, Jean-Paul Allouche and Valérie Berthé, With a
preface by Berstel and Perrin.

[76] L. Lovász and M. D. Plummer. Matching Theory. AMS Chelsea Publishing, Providence,
RI, 2009. ISBN 978-0-8218-4759-6. doi: 10.1090/chel/367. URL https://mathscinet.
ams.org/mathscinet-getitem?mr=2536865. Corrected reprint of the 1986 original
[MR0859549].

[77] L. Lovász, P. József, and K. Vsztergombi. Discrete Mathematics: Elementary and


Beyond. Undergraduate Texts in Mathematics. Springer, 2003.

[78] R. C. Lyndon. On Burnside’s problem. Trans. Amer. Math. Soc., 77(2):202–215, 1954.

[79] D. G. Ma. An elementary proof of the solutions to the diophantine equation 6y 2= x


(x+ 1)(2x+ 1). Sichuan Daxue Xuebao, 4:107–116, 1985.

[80] I. Macdonald. Symmetric Functions and Hall Polynomials. Oxford Mathematical Mono-
graphs. Clarendon Press, 1979. ISBN 9780198535300. URL https://books.google.
com/books?id=XoMpAQAAMAAJ.

[81] I. G. Macdonald. Symmetric Functions and Hall Polynomials. Oxford Mathe-


matical Monographs. The Clarendon Press, Oxford University Press, New York,
second edition, 1995. ISBN 0-19-853489-2. URL https://mathscinet.ams.org/
mathscinet-getitem?mr=1354144. With contributions by A. Zelevinsky, Oxford Sci-
ence Publications.

[82] P. A. MacMahon. The Indices of Permutations and the Derivation Therefrom of Func-
tions of a Single Variable Associated with the Permutations of any Assemblage of Ob-
jects. Amer. J. Math., 35(3):281–322, 1913. ISSN 0002-9327. doi: 10.2307/2370312.
URL https://mathscinet.ams.org/mathscinet-getitem?mr=1506186.

[83] P. A. MacMahon. Combinatory Analysis, Volumes I and II, volume 137. American
Mathematical Soc., 2001.

[84] A. Marcus and G. Tardos. Excluded permutation matrices and the Stanley-Wilf con-
jecture. J. Combin. Theory Ser. A, 107(1):153–160, 2004. ISSN 0097-3165. doi: 10.
1016/j.jcta.2004.04.002. URL https://mathscinet.ams.org/mathscinet-getitem?
mr=2063960.

[85] Y. V. Matiyasevich. The diophantineness of enumerable sets. In Doklady Akademii


Nauk, volume 191, pages 279–282. Russian Academy of Sciences, 1970.

[86] E. Mendelson. Introduction to Mathematical Logic. Textbooks in Mathematics. CRC


Press, 6th edition, 2015.
[87] A. Mendes and J. Remmel. Permutations and words counted by consecutive patterns.
Adv. in Appl. Math., 37(4):443–480, 2006. ISSN 0196-8858. doi: 10.1016/j.aam.2005.
09.005. URL https://mathscinet.ams.org/mathscinet-getitem?mr=2266893.
[88] A. Mendes and J. Remmel. Counting with Symmetric Functions. Number 43 in Develop-
ments in Mathematics. Springer International Publishing, 2015. ISBN 9783319236186.
[89] P. Mihailescu. Primary units and a proof of catalan’s conjecture. Submitted. Available
online from http://www-math. unipaderborn. de/preda, 2002.
[90] F. D. Murnaghan. The characters of the symmetric group. American Journal of Math-
ematics, 59(4):739–753, 1937. ISSN 00029327, 10806377. URL http://www.jstor.
org/stable/2371341.
[91] F. D. Murnaghan. The characters of the symmetric group. Proceedings of the National
Academy of Sciences (PNAS), 37(1):55–58, 1951.
[92] T. Nakayama. On some modular properties of irreducible representations of a symmetric
group, i. Japanese Journal of Mathematics, 17:165–184, 1940. doi: 10.4099/jjm1924.
17.0 165. URL https://doi.org/10.4099%2Fjjm1924.17.0_165.
[93] T. Nakayama. On some modular properties of irreducible representations of symmetric
groups, II. Japanese Journal of Mathematics, 17:411–423, 1940. doi: 10.4099/jjm1924.
17.0 411. URL https://doi.org/10.4099%2Fjjm1924.17.0_411.
[94] D. J. Newman. Simple analytic proof of the prime number theorem. The American
Mathematical Monthly, 87(9):693–696, 1980.
[95] J. Oesterlé. Nouvelles approches du “théoreme” de fermat. Astérisque, 161(162):165–
186, 1988.
[96] R. Otter. The number of trees. Annals of Mathematics, pages 583–599, 1948.
[97] T. K. Petersen. Eulerian Numbers. Birkhäuser Advanced Texts: Basler Lehrbücher.
[Birkhäuser Advanced Texts: Basel Textbooks]. Birkhäuser/Springer, New York, 2015.
ISBN 978-1-4939-3090-6; 978-1-4939-3091-3. doi: 10.1007/978-1-4939-3091-3. URL
https://mathscinet.ams.org/mathscinet-getitem?mr=3408615. With a foreword
by Richard Stanley.
[98] G. d. B. Robinson. On the Representations of the Symmetric Group. Amer. J. Math., 60
(3):745–760, 1938. ISSN 0002-9327. doi: 10.2307/2371609. URL https://mathscinet.
ams.org/mathscinet-getitem?mr=1507943.
[99] O. Rodriguez. Note sur les inversions, ou dérangements produits dans les permutations.
J. de Math, 4(1839):236–240, 1839.
[100] K. H. Rosen. Discrete Mathematics and Its Applications. McGraw-Hill Education, 8th
edition, 2019.
[101] K. H. Rosen, D. R. Shier, and W. Goddard, editors. Handbook of Discrete and Com-
binatorial Mathematics. Discrete Mathematics and Its Applications. CRC Press, 2nd
edition, 2018.
[102] G.-C. Rota. The number of partitions of a set. Amer. Math. Monthly, 71:498–504,
1964. ISSN 0002-9890. doi: 10.2307/2312585. URL https://mathscinet.ams.org/
mathscinet-getitem?mr=161805.

[103] G.-C. Rota. On the foundations of combinatorial theory. I. Theory of Möbius functions.
Z. Wahrscheinlichkeitstheorie und Verw. Gebiete, 2:340–368 (1964), 1964. doi: 10.1007/
BF00531932. URL https://mathscinet.ams.org/mathscinet-getitem?mr=174487.

[104] B. E. Sagan. The Symmetric Group: Representations, Combinatorial Algorithms, and


Symmetric Functions, volume 203 of Graduate Texts in Mathematics. Springer-Verlag,
New York, 2nd edition, 2001. ISBN 0-387-95067-2. doi: 10.1007/978-1-4757-6804-6.
URL https://mathscinet.ams.org/mathscinet-getitem?mr=1824028.

[105] C. Schensted. Longest increasing and decreasing subsequences. Canadian J. Math.,


13:179–191, 1961. ISSN 0008-414X. doi: 10.4153/CJM-1961-015-3. URL https:
//mathscinet.ams.org/mathscinet-getitem?mr=121305.

[106] M. P. Schützenberger. Quelques remarques sur une construction de Schensted. Math.


Scand., 12:117–128, 1963. ISSN 0025-5521. doi: 10.7146/math.scand.a-10676. URL
https://mathscinet.ams.org/mathscinet-getitem?mr=190017.

[107] M.-P. Schützenberger. La correspondance de robinson. In Combinatoire et


Représentation du Groupe Symétrique, pages 59–113. Springer Berlin Heidelberg, 1977.
doi: 10.1007/bfb0090012. URL https://doi.org/10.1007%2Fbfb0090012.

[108] A. Selberg. An elementary proof of the prime-number theorem. Annals of Mathematics,


pages 305–313, 1949.

[109] D. Singmaster. Research problems: How often does an integer occur as a binomial
coefficient? Amer. Math. Monthly, 78(4):385–386, 1971. ISSN 0002-9890. doi: 10.2307/
2316907. URL https://mathscinet.ams.org/mathscinet-getitem?mr=1536288.

[110] R. P. Stanley. A symmetric function generalization of the chromatic polynomial of a


graph. Advances in Mathematics, 111(1):166–194, 1995.

[111] R. P. Stanley. Enumerative Combinatorics, Vol. 1, volume 49 of Cambridge Studies in


Advanced Mathematics. Cambridge University Press, Cambridge, 1st edition, 1997.

[112] R. P. Stanley. Enumerative Combinatorics, Vol. 2, volume 62 of Cambridge Studies in


Advanced Mathematics. Cambridge University Press, Cambridge, 1999.

[113] R. P. Stanley. Recent progress in algebraic combinatorics. volume 40, pages 55–
68. 2003. doi: 10.1090/S0273-0979-02-00966-7. URL https://mathscinet.ams.org/
mathscinet-getitem?mr=1943133. Mathematical challenges of the 21st century (Los
Angeles, CA, 2000).

[114] R. P. Stanley. Enumerative Combinatorics, Vol. 1, volume 49 of Cambridge Studies in


Advanced Mathematics. Cambridge University Press, Cambridge, 2 edition, 2011.

[115] R. P. Stanley. Catalan Numbers. Cambridge University Press, 2015.


[116] J. R. Stembridge. A concise proof of the littlewood-richardson rule. The Electronic
Journal of Combinatorics, 9(1), Mar. 2002. doi: 10.37236/1666. URL https://doi.
org/10.37236%2F1666.
[117] G. P. Thomas. Baxter algebras and Schur functions. PhD thesis, University College of
Swansea, 1974.
[118] P. Turán. On an external problem in graph theory. Mat. Fiz. Lapok, 48:436–452, 1941.
[119] B. Veklych. A minimalist proof of the law of quadratic reciprocity. The American
Mathematical Monthly, 126(10):928–928, 2019.
[120] K. Wagner. Über eine eigenschaft der ebenen komplexe. Mathematische Annalen, 114
(1):570–590, 1937.
[121] W. D. Wallis. A Beginner’s Guide to Discrete Mathematics. Birkhäuser/Springer,
New York, 2012. ISBN 978-0-8176-8285-9. URL https://mathscinet.ams.org/
mathscinet-getitem?mr=2866528. Second edition [of MR1935109].
[122] E. Waring. Meditationes Algebraicae: An English Translation of the Work of Edward
Waring; Ed. and Transl. from the Latin by Dennis Weeks. American Mathematical
Society, 1991.
[123] H. Wegner. Über das Maximum bei Stirlingschen Zahlen zweiter Art. J. Reine Angew.
Math., 262(263):134–143, 1973. ISSN 0075-4102. doi: 10.1515/crll.1973.262-263.134.
URL https://mathscinet.ams.org/mathscinet-getitem?mr=335282.
[124] L. Weisner. Abstract theory of inversion of finite series. Transactions of the American
Mathematical Society, 38(3):474–484, 1935. ISSN 00029947. URL http://www.jstor.
org/stable/1989808.
[125] A. Wiles. Modular elliptic curves and fermat’s last theorem. Annals of mathematics,
141(3):443–551, 1995.
[126] H. S. Wilf. generatingfunctionology. A K Peters, Ltd., Wellesley, MA, 3rd edition,
2006. ISBN 978-1-56881-279-3; 1-56881-279-5. URL https://mathscinet.ams.org/
mathscinet-getitem?mr=2172781.
Author Index

Aristotle (384-322 BC), 8 Lagarias, Je↵rey Clark (1949–), 5


Lambert, Johann Heinrich (1728–1777), 58
Bell, Eric Temple (1883–1960), 57 Leibniz, Gottfried Wilhelm (1646–1716), 8
Binet, Jacques Philippe Marie (1786–1856), 57 Lyndon, Roger Conant (1917–1988), 32
Boole, George (1815–1864), 8
Newton, Isaac (1643–1727), 49
Carnegie, Dale (1888–1955), 7
Cassini, Giovanni Domenico (1625–1712), 57
Catalan, Eugène Charles (1814–1894), 52 Pochhammer, Leo August (1841–1920), 48

De Morgan, Augustus (1806–1871), 8 Schützenberger, Marcel-Paul (1920–1996), 31


Durfee, William Pitt (1855–1941), 20 Singmaster, David Breyer (1938–), 5
Stirling, James (1692–1770), 17, 25
Erdős, Paul (1913–1996), 6
Euler, Leonhard (1707–1783), 21, 22
Thue, Axel (1863–1922), 31
Ferrers, Norman Macleod (1829–1903), 19
Fibonacci (1170 – 1240–50), 56 Young, Alfred (1873–1940), 19

73
Subject Index

q-analogue, 29 derangement, 27
descent, 26
alphabet, see word diagram
antecedent, 9 arm, 20
appeal to probability, 12 box, 19
argument, 12 cell, 19
arm, see diagram Ferrers, see Ferrers diagram
ascent, 26 French notation, 20
hook, 20
leg, 20
ballot sequence, 54
shape, 19
Binet’s formula, 57
square, 19
binomial coefficient, 48
Young, see Young diagram, 19
binomial theorem, the, 49
disjunction, 9
block, see partition (of a set)
double factorial, 16
box, see diagram
drop, 26
bracketing, 55
Durfee square, 20

Cassini’s identity, 57 Erdős-Turán conjecture, 4


cell, see diagram Euler’s function (q), the, 21
child excedance, 26
left, 54 existential fallacy, the, 13
right, 54
Chu-Vandermonde identity, the, 50 factor, see word
Collatz conjecture, 5 left, 31
complement proper, 31
(of a permutation), 24 right, 31
composition, 19 word, 31
conclusion, 9, 12 fallacy, 12
conjecture, 8 masked-man, the, 14
conjugate fallacy fallacy, the, 12
(of a permutation), 25 fallacy of necessity, 13
(of an integer partition), the, 20 Fallacy of the inverse, 13
self-, 20 fallcy of the converse, 13
word, 31 false exclusionary disjunct, 12
conjunction, 9 Ferrers diagram, 19
Conjunctive simplification, 12 fixed point, 26
consequent, 9 floor function, the, 57
contradiction, 10
contrapositive, 11 generating function
converse, 11 exponential, 16
cube (word), 35 ordinary, 16
cube-free (word), 35 geometric series formula, the, 49
cycle notation, 24 golden ratio, 57

75
hereditary (property), 36 path
hockey-stick identity, the, 51 Dyck, 55
hook, see diagram lattice, 55
hook-length formula, the, 39 Motzkin, 55
hypothese, 12 Schröder, 55
hypothesis, 9 pattern, 30
pattern avoiding, 30
Illicit major, 13 peak, 26
Illicit minor, 13 permutation, 24
implication, 9 Pochhammer symbol, 48
insertion tableau, the, 40 predicate, 11
inverse prefix, 31
(of an implication), 11 premise, 12
inversion, 26 primitive, see word
involution, 27 proposition, 8

lattice path, 61 Q-tableau, the, 40


left-to-right maxima, 27 quantifier, 11
left-to-right minima, 27 existential, 11
leg, see diagram nested, 11
length universal, 11
word, 31
letter, see word recording tableau, the, 40
limit, see word reduced form, 30
logic connector, 9 reflection principle, the, 61
Lyndon word, 32 reversal, 31
right-to-left maxima, 27
Möbius function, the, 46 right-to-left minima, 27
major index, the, 26 Robinson-Schensted correspondence, the, 40, 42
matrix representation, 24 root, 54
Modus ponens, 12 row insert, 40
modus ponens rules of inference, 12
universal, 12
Modus tollens, 12 shape, see diagram
modus tollens Singmaster’s conjecture, 5
universal, 12 square, see diagram
multinomial coefficient, 51 square (word), 35
multinomial theorem, the, 52 square-free (word), 35
multiset, 18 Stanley-Wilf limit, 30
multiset coefficient, 51 stars and bars, 52
statement, 8
necessary, 9 biconditional, 9
negation, 8 conditional, 9
number Stirling number
Bell, 57 of the first kind, 25
Catalan, 52 of the first kind, signless, 25
Fibonacci, 56 of the second kind, 17
Genocchi, 58 sufficent, 9
Narayana, 58 suffix, 31
syllogism, 12
one-line notation, 24 disjunctive, 12
order:lexicographic, the, 32 hypothetical, 12

P -tableau, the, 40 tableau


palindrome, 31 P -, the, 40, see also RSK algorithm, the
paradox, 14 Q-, the, 40, see also RSK algorithm, the
parenthesization, 55 insertion, 40, see also RSK algorithm, the
part, 19 partial, 40
partition recording, 40, see also RSK algorithm, the
(of a set), 17 Young, see Young tableau
block, 17 tautology, 10
(of an integer), 18 telescoping method, the, 51
theorem, 12 conjugate, 31
tree empty, the, 31
binary, 54 infinite, 35
ordered, 54 infinite word of Thue-Morse, the, 37
plane, 54 letter, 31
triangulation, 53 limit, 36
truth value, 8 Lyndon, 32
two-line notation, 24 primitive, 31
type of a permutation, the, 24 reversal, 31

valley, 26
Young diagram, 19
word Young tableau, 39
alphabet, 31 standard, 39

You might also like