You are on page 1of 80

POLYNOMIAL METHODS: RECENT ADVANCEMENTS IN COMBINATORICS

by

Thomas Fleming

A Thesis

Submitted in Partial Fulfillment of the

Requirements for the Degree of

Master of Science

Major: Mathematical Sciences

The University of Memphis

May 2022
ACKNOWLEDGMENTS
First and foremost, I would like to thank my advisor and professor, Dr. Grynkiewicz. He
supplied me with many of the ideas for this thesis, clarified seemingly mystical results,
and helped me through many rounds of revisions. I would also like to thank the attending
professors for my Thesis defense, Dr. Wierdl and Dr. Nikiforov, who were kind enough to
supervise my defense only days before spring break. Finally, I would like to thank my
family and my fiancee, Amber.

ii
ABSTRACT
Fleming, Thomas Rexford. M.Sc. The University of Memphis. May 2022.
Polynomial Methods: Recent Advancements in Combinatorics. Major Professor: Dr.
David Grynkiewicz.

In this Master’s Thesis, we showcase the use of an array of results collectively


known as the polynomial method. First, we lay groundwork, giving some basic
definitions, notation, and prerequisites. Then, we introduce three related theorems, the
Combinatorial Nullstellensatz, Generalized Combinatorial Nullstellensatz, and Punctured
Combinatorial Nullstellensatz, each concerning properties of multivariate polynomials
with specific constraints on their degree. With these results proven, we then showcase
some simple combinatorial and graph theoretic results which have very simple proofs
making use of the Nullstellensatz theorems. In chapter 4 we give the proof of the q-Dyson
theorem as published by Zeilberger and Bressoud. Finally, in chapters 5, 6 and 7 we
introduce zero-sum theory, prove the Davenport constant of finite abelian p-groups
without making use of group rings, and then state and prove the Chevalley-Warning
theorem and use it to prove the Kemnitz conjecture on zero-sums in cyclic groups.

iii
TABLE OF CONTENTS
Contents Pages

1. Introduction and Preliminary Results 1


Introduction 1
Definitions, Notation, and Preliminary Results 3

2. Nullstellensatz 8
Hilbert’s Nullstellensatz 8
Alon’s Combinatorial Nullstellensatz 9
Ball and Serra’s Punctured Nullstellensatz 14

3. Simple Combinatorial Proofs 20


Sumsets 20
A Nice Result for Graphs 30

4. Zeilberger-Bressoud q-Dyson Theorem 32


The Dyson Conjecture and Lagrange Interpolation 33
The q-Dyson Theorem 34

5. Zero-Sum Theory 43
Definitions and Notation 43
Basic Proofs and Prerequisites 45

6. Davenport constant of finite abelian p-groups 51


Map Functors 51
Combinatorial Identities 56
The Davenport Constant 58
Remarks 60

7. The Chevalley-Warning Theorem and Reiher’s Proof of the Kemnitz Conjec-


ture 62
The Chevalley-Warning Theorem 62
Reiher’s Proof of the Kemnitz Conjecture 64

8. Conclusion 73

References 74

iv
Chapter 1

Introduction and Preliminary Results

Introduction

One of the earliest objects of algebraic study, real (specifically integral) univariate
polynomials generalize the most simple algebraic objects, linear functions, and have many
significant applications. Multivariate polynomials offer further generalization. Expanding
even further, multivariate polynomials over arbitrary rings (and fields) prove surprisingly
useful. In this thesis, we will showcase a variety of methods and applications involving
polynomials. These applications will prove some surprising results, both abstract and
concrete. To understand these proofs, one need only basic knowledge of algebraic tools
and structures, and some experience with methods of combinatorial proof.

Merging the fields of Combinatorics and Number Theory yields Combinatorial


Number Theory and Additive Combinatorics, two overlapping fields answering questions
pertaining to zero-sums over sequences and zeroes of arbitrary polynomials, as well as
other problems at the intersection of Combinatorics, Algebra, and Number Theory. We
begin Chapter 1 by presenting many common objects which we will make use of
throughout the thesis. This will include some objects of common study, such as
polynomials, as well as some notation and conventions common to the field, and the
statements of a few common theorems which the reader will likely be familiar with. After
this, in chapter 2, we introduce the Nullstellensatz theorems, literally translating as
zero-locus theorem, which characterize a polynomial based on its number of zeroes.
These theorems, while extremely powerful, will have relatively simple statements and
proofs culminating in the Punctured Nullstellensatz. Moving on, in chapter 3 we
demonstrate some applications of these Nullstellensatz theorems. These will mostly make
use of the simplest version of the Nullstellensatz which we present, Alon’s Combinatorial
Nullstellensatz II [1]. The applications will come in two main varieties, sumset proofs and

1
graph theoretic proofs. This chapter will provide some of the most concrete applications
of the polynomial method. In chapter 4 we present Dyson’s Theorem [9] on laurent
polynomials and Zeilberger and Bressoud’s q-Dyson variant [8], with a proof based on a
concrete version of the Combinatorial Nullstellensatz. In chapter 5 we provide a short
introduction to zero-sum theory along with a few basic results. Then, in chapter 6, we
prove the davenport constant D of finite abelian p-groups. This proof, though similar to
one used by Christian Elsholtz in [5], is hitherto unpublished. Finally, in chapter 7, we
prove the Kemnitz conjecture concerning zero-sum sequences in Cn2 . First, though, we
must state some prerequisites which we will make use of throughout this thesis.

2
Definitions, Notation, and Preliminary Results

Before we continue on, we define a few basic objects which we will make regular
use of and 3 algebraic results which will go unproven throughout the thesis. Readers are
assumed to have knowledge equivalent to a first course in group and ring theory and a first
course in combinatorics.

Definition 1.1 (Polynomial Ring). Given a commutative ring R, we define R [x] to


be the ring generated by all formal sums of the form f = ∞ i
P
i=0 ci x , ci ∈ R where

ci ̸= 0 for only finitely many i ∈ N0 . Given f = ∞


P i
P∞ i
i=0 ai x , g = i=0 bi x we define

f +g = ∞
P i
P∞ P∞ i+j
i=0 (ai + bi ) x and f g = i=0 j=0 ai bj x . It is routinely checked that
these operations, together with the set of all such polynomials f ∈ R [x], form a ring.
We call R [x] equipped with + and · the polynomial ring in 1 variable over R. We
call a polynomial monic if the coefficient of its highest power term is 1.

Definition 1.2 (Multivariate Polynomial Ring). Given a ring R, we define a polyno-


mial ring over n variables, R [x1 , . . . , xn ] inductively by the rules R [x1 ] = R [x] and
R [x1 , . . . , xi ] = (R [x1 , . . . , xi−1 ]) [xi ].

3
Definition 1.3 (Laurent Polynomial Ring). Given a field F , we define a univariate
laurent polynomial to be a formal sum of the form: f = ∞ i
P
i=−∞ ai x with ai ∈ F

and ai ̸= 0 for only finitely many i ∈ Z. Multiplication and addition are defined just
as with standard polynomial rings, except with the added relation x · x−1 = 1.
We define a multivariate laurent polynomial over n variables to be a formal sum of
P i1 in
the form: f = i1 ,...,in ∈Z ci1 ,...,in x1 . . . xn with ci1 ,...,in ̸= 0 for only finitely many

tuples
(i1 , . . . , in ) ∈ Zn .
Once again, it is routinely checked that the set of all laurent polynomials in n-variables
with coefficients in F , together with + and · and the relation xi · x−1
i = 1 1 ≤ i ≤ n,
forms a ring for each n ∈ N.

Definition 1.4 (Degree). We define the degree of a polynomial f ∈ R [x] as

deg (f ) = min{N ∈ N0 : ci = 0 ∀ i > N }.

For multivariate polynomial rings there are two natural generalizations of degree, total
degree and projected degree. For a polynomial

X
f= ci1 ,...,in xi11 . . . xinn ∈ R [x1 , . . . , xn ]
i1 ,...,in ∈N0

we define the total degree as

n
X
deg (f ) = min{N ∈ N0 : ci1 ,...,in = 0 ∀ i1 , . . . , in ∈ N0 so that ij > N }
j=1

and we define the projected degree in variable xk as the degree of f when interpreted
as a polynomial in the ring (R [x1 , . . . , xk−1 , xk+1 , . . . , xn ]) [xk ]. Essentially, this is the
degree in only the variable xk and it is denoted by degxk (f ).

4
Example.

deg x2 = deg x2 + x + 1 = deg x1 x2 + x21 + x22 + 1 = degx1 x21 x32 x43 = 2


   

Definition 1.5 (Module). Given a ring R, we define a (left) R-module M to be a set


endowed with two operations, addition and multiplication, such that + : M × M →
M and · : R × M → M with + satisfying the axioms of an abelian group and ·
being associative, having an identity, and having both left and right distributive laws:
(rs) (x) = r (sx), 1x = x, ((r + s) x = rx + sx, r (x + y) = rx + ry for r, s ∈ R and
x, y ∈ M ). If R is a field, we instead call M a vector space.
In a vector space M , we call a set B a basis of M if it is linearly independent and spans
M.

Definition 1.6 (Characteristic


 Function). Given a vector x, the characteristic function
 1,

x=y
χx has χx (y) =
 0,
 x ̸= y

Notation (Common Symbols). Throughout this thesis we will make use of the following
symbols, Z for the integers, N for the positive integers (not including 0), N0 for the
nonnegative integers including 0, and Q for the rationals. Since we do not introduce any
proofs using R, we will use [a, b] = {x ∈ Z : a ≤ x ≤ b}. Furthermore, G will be a
group, R a ring, and F a field.

Notation (Polynomial Coefficient). For a laurent polynomial

X
f= ci1 ,...,in xi11 . . . xinn
i1 ,...,in ∈Z

5
over n variables, we denote the coefficient of a particular term by xi11 . . . xinn f := ci1 ,...,in
 

when it is more convenient.

Notation (Vector Polynomials). For a (laurent) polynomial f (x1 , . . . , xn ) evaluated at


x1 , . . . , xn , we implicitly denote x = (x1 , . . . , xn ) and f (x) = f (x1 , . . . , xn ) when it is
unambiguous and convenient.

Notation (Congruence). For a congruence statement a ≡ b ( mod c) we sometimes


shorten it to a ≡ b when the modulus c is unambiguous.

Notation (Cyclic Groups). For a given n ∈ N, we denote the cyclic group of order n as
Cn , i.e., Z/nZ = Cn . Since Z/nZ is also a field for prime p, we use it in this context as
well.

The next theorem has many equivalent statements as well as a few stronger (and weaker)
ones. For the rest of this thesis, we will assume the Fundamental Theorem of Algebra
refers to the following statement.

Theorem 1.1 (Fundamental Theorem of Algebra). Given an integral domain R and


f ∈ R [x] with f nonzero, let X = {x ∈ R : f (x) = 0}. Then, |X| ≤ n.

Next, we provide the binomial theorem, or binomial coefficient theorem.

Theorem 1.2. Let p be prime, f ∈ Cp [x, y], 0 ≤ a < b ≤ p, then xa y b−a (x + y)b =
 

b
. So, (x + y)b = ba=0 ab xa y b−a .
 P 
a

This theorem is easily extended from the integers to the finite field Cp , so we neglect to
prove it.
The next two theorems are routinely checked and commonly proved in a first
introduction to ring theory.

6
Theorem 1.3. Let R be a ring. If R is a unique factorization domain, then R [x1 , . . . , xn ]
is a unique factorization domain for all n ∈ N.

Theorem 1.4. Let R be an integral domain and F its quotient field. If x (x1 ) , y (x) ∈
R [x1 , . . . , xn ] with x (x1 ) being monic and x (x1 ) | y (x) over F then, x (x1 ) | y (x)
over R.

This last prerequisite is a direct consequence of the pigeonhole principle

Theorem 1.5. Let G be a finite abelian group and let A, B ⊆ G be nonempty subsets.
If |A| + |B| ≥ |G| + 2, then all elements x ∈ G have at least 2 representations of the
form a1 + b1 = a2 + b2 = x where a1 , a2 ∈ A, b1 , b2 ∈ B and a1 ̸= a2 , b1 ̸= b2 .

Indeed, this theorem can be generalized to hold for r representations ai + bi , 1 ≤ i ≤ r. A


good proof is found in [7].

7
Chapter 2

Nullstellensatz

Hilbert’s Nullstellensatz

Before we can approach the Combinatorial Nullstellensatz, or any other tools of


the polynomial method, it is important to see a closely related theorem, Hilbert’s
Nullstellensatz. This theorem is more algebraic in nature, though its influence in the
Combinatorial Nullstellensatz will be clear. One statement of this theorem is as follows,

Theorem 2.1 (Hilbert’s Nullstellensatz). Let F be an algebraically closed field and


f, g1 , . . . , gm ∈ F [x1 , . . . , xn ]. Let V (g1 , . . . , gm ) = {x ∈ F n : gi (x) = 0, 1 ≤
i ≤ m}. If f (v) = 0 for all v ∈ V (g1 , g2 , . . . , gm ), then there is an integer k and
polynomials h1 , . . . , hm ∈ F [x1 , . . . , xm ] such that f k = ni=1 hi gi .
P

For a proof of this, see [13].


This theorem, though closely related to the Combinatorial Nullstellensatz, will see
no use in this thesis, though their historical and mathematical relation are good to keep in
mind. With this, we now introduce the main tool of this thesis, Alon’s Combinatorial
Nullstellensatz.

8
Alon’s Combinatorial Nullstellensatz

Alon’s theorem actually falls under a special case of Hilbert’s Nullstellensatz,


Q
when n = m and each gi is univariate and of the form s∈Si (xi − s). However, these
extra conditions allow a stronger conclusion to be drawn, namely that f can split into
polynomials of lesser collective degree than itself. Though this statement is non-rigorous,
it is immediately clear that turning one polynomial into a sum of simpler polynomials can
have useful corollaries. It is in fact one of these corollaries that we will make use of
throughout much of this thesis. Before we can prove such a theorem, though, we need to
state and prove the following lemma:

Lemma 2.1. [1] Let R be an integral domain. Let n ≥ 1, f (x) ∈ R[x1 , x2 , . . . , xn ],


and A1 , A2 , . . . , An ⊆ R be finite and nonempty.
Qn
Suppose degxi f < |Ai | for each 1 ≤ i ≤ n and f (a) = 0 for all a ∈ i=1 Ai .
Then f is the zero polynomial.

Proof. Let F be the quotient field of R. Clearly, any polynomial in R is a polynomial in


F , and as f = 0 in F implies f = 0 in R, we need only consider the case F = R. We will
proceed by induction on n. We already know the case n = 1 holds by the Fundamental
Theorem of Algebra. Now, we prove the case n. First, let us define fi (x1 , x2 , . . . , xn−1 ) to
be the coefficient of xin when considering f as a polynomial in the ring
(F [x1 , . . . , xn−1 ]) [xn ]. Then,

degxn f
X
f (x) = fi (x1 , x2 , . . . , xn−1 ) xin .
i=0

We see each fi ∈ F [x1 , x2 , . . . , xn−1 ] and degxj fi ≤ degxj f < |Ai | for all 1 ≤ j ≤ n − 1
and 0 ≤ i ≤ degxn f . Let ai ∈ Ai be an arbitrary element for each 0 ≤ i ≤ n − 1. As
f (a1 , a2 , . . . , an−1 , a) = 0 for all a ∈ An and degxn f < |An |, the n = 1 case implies that
f (a1 , a2 , . . . , an−1 , xn ) = 0 must be the zero polynomial. Hence fi (a1 , . . . , an−1 ) = 0 for

9
Qn−1
each 1 ≤ i ≤ n and each (a1 , . . . , an−1 ) ∈ i=1 Ai . Then, by the inductive hypothesis
each fi is the zero polynomial, so f is the zero polynomial.

This lemma comprises the majority of the intellectual heavy lifting for our main theorem,
and thus its proof will appear very simple in comparison to its powerful statement:

Theorem 2.2 (Generalized Combinatorial Nullstellensatz [7]). Let R be an integral


domain, n ≥ 1, and let A1 , A2 , . . . , An ⊆ R be finite and nonempty.
Q
Let f (x) ∈ R[x1 , x2 , . . . , xn ] be a polynomial and define gi (xi ) = a∈Ai (xi − a) for
1 ≤ i ≤ n. Then, f (a1 , a2 , . . . , an ) = 0 for all a ∈ ni=1 Ai if and only if there are
Q

polynomials h1 (x) , h2 (x) , . . . , hn (x) ∈ R[x1 , x2 , . . . , xn ] such that

n
X
f (x) = gi (xi ) hi (x)
i=1

and
deg (gi hi ) ≤ deg (f ) and degxj (gi hi ) ≤ degxj (fi )

for all 1 ≤ i, j ≤ n.

Proof. First, assume there exist polynomials h1 , . . . , hn ∈ R [x1 , . . . , xn ] so that


f (x) = ni=1 gi (xi ) hi (x) and the constraints on the degree of gi hi hold. Then, it is clear
P

f (a) = 0 for each a ∈ ni=1 Ai .


Q

Conversely, suppose f (a1 , . . . , an ) = 0 for all a ∈ ni=1 Ai . We wish to show that


Q

there are polynomials h1 , . . . , hn , such that f (x) = ni=1 gi (xi ) hi (x) together with the
P

constraint on the degree of gi hi . For each 1 ≤ j ≤ n, let gji ∈ R so that:

|Aj |−1
|A |
Y X
gj (xj ) = (xj − a) = xj j − gji xij . (2.1)
a∈Aj i=0

|Aj | P|Aj |−1


Hence, as aj ∈ Aj implies that gj (aj ) = 0 we must have that aj = i=0 gij aij . Now,

10
we define a new polynomial f , such that f is the result of starting with the polynomialf
|A | P|A |−1
and applying a substitution of the form xj j 7→ i=0j gij xij for each 1 ≤ j ≤ neuntil
|A |
this is no longer possible (there are no powers of xj j remaining). We see that

degxi f < |Ai | for each i (as any monomial term of degree |Ai | can be substituted for
terms of lesser degree), and as f (a) = f (a) = 0 for each a ∈ ni=1 Ai (this fact is trivial
Q

by the construction of f and f ), then by the preceding lemma we have that f = 0. Next,
consider the polynomial f (x) − f (x). Define f0 = f and fi to be f after the i’th such
substitution. Furthermore, let N ∈ N be such that fN = f . Then, the polynomial fi is
obtained by examining a term of fi−1 with a degree in xj of at least |Aj | and replacing an
|Aj |
individual factor of xj by the sum as we defined earlier. For simplicity let us take the
|Aj |
term cxk11 xk22 . . . xknn · xj (where kj ≥ 0 by our construction) and apply such a
substitution. Then, we see that the difference
 
|Aj |−1
|Aj |
X
fi−1 − fi = cxk11 xk22 . . . xknn xj − gji xij  = cxk11 xk22 . . . xknn · gj (xj )
i=0

by 2.1. Furthermore, as fN = f we have that

f − f = (f0 − f1 ) + (f1 − f2 ) + . . . + (fN −1 − fN ) .

Pn
Hence, f = f − 0 = f − f = j=1 gj (xj ) hj where each hj ∈ R[x1 , x2 , . . . , xn ] is simply
the sum of all of these preceding terms cxk11 . . . xknn as a result of these substitutions.
P 
|Aj |−1 i
Furthermore, we had by construction that deg gj = |Aj | > deg i=0 g x
ji j , so we see

that such a substitution will never increase the degree of our polynomial and hence
deg fi ≤ deg f . Then, as deg gj = |Aj |, we see that

deg cxk11 . . . xknn gj (xj ) ≤ deg fi−1 ≤ deg f.




11
As each hj is simply the sum of such terms, cxk11 . . . xknn , we may substitute its
degree to yield deg (gj hj ) ≤ deg f .
Lastly, we show this holds for projected degree as well. Note that as we are
P|A |−1
replacing a term gj with degxk (gj ) = 0, and as each hj with i=0j gji xij also has
projected degree 0, the overall degree in xk will not increase after i substitutions. Hence
for k ̸= j, degxk cxk11 . . . xknn gj (xj ) ≤ degxk (fi−1 ) ≤ degxk (f ). Moreover, if k = j, the


same argument shows the projected degrees decrease. These facts yield
degxk cxk11 . . . xknn gj ≤ degxk (f ). And, as each hj is simply the sum of these remaining


terms, we see degxk (hj ) + degxk (gj ) ≤ degxk (f ) for each 1 ≤ j, k ≤ n.

This theorem is perhaps too general and too powerful for use in proving our simpler
results. As a corollary, many authors include a weaker second statement of the theorem,
where the splitting of a polynomial is replaced by a constraint on the cardinality of the set
on which f (x) = 0.

Theorem 2.3 (Combinatorial Nullstellensatz [7]). Let R be an integral domain, n ≥ 1,


and A1 , A2 , . . . , An ⊆ R be finite and nonempty, with f (x) ∈ R[x1 , x2 , . . . , xn ] being
a polynomial. Suppose
[xd11 xd22 . . . xdnn ]f (x) ̸= 0

and that deg f = ni=1 di with 0 ≤ di ≤ |Ai | − 1. Then, there exists an element
P

a ∈ ni=1 Ai such that f (a) ̸= 0. That is, given a nontrivial "maximal degree" mono-
Q

mial of f such that the degree of this monomial in each variable, xi , is strictly less than
the size of the corresponding set Ai , f cannot be zero on the whole of the product of
the Ai .

This is in essence a generalization of the Fundamental theorem of Algebra and it is


precisely this simple statement about the number of zeroes of a particular polynomial

12
which provides our combinatorial power. The proof follows rather directly from the
generalized version as follows:

Q
Proof. Let gi = a∈Ai (xi − a) for 1 ≤ i ≤ n and suppose indirectly that f (a) = 0 for all
a ∈ ni=1 Ai with deg (f ) = ni=1 di and xd11 . . . xdnn f (x) ̸= 0 . Now, xd11 . . . xdnn is a
Q P  

maximal degree monomial. Then, theorem 2.1 yields polynomials, which we will denote
as h1 , h2 , . . . , hn ∈ R[x1 , x2 , . . . , xn ] such that f (x) = ni=1 gi (xi ) hi (x) and
P

deg (gi hi ) ≤ deg (f ) for each 1 ≤ i ≤ n. Now let us examine the nature of
[xd11 . . . xdnn ]f (x). As this is a maximal degree monomial of f , then only maximal degree
monomials of hi (x) gi (xi ) can contribute to its coefficient in f as deg (hi gi ) ≤ deg f .
Since deg (g) = |Ai | > di , we find degxi (hi gi ) ≥ |Ai | > di . So, all maximal degree
monomials, M , of f must have degxj (M ) ≥ |Ai | for some 1 ≤ j ≤ n. So, xd11 . . . xdnn
cannot be a maximal degree monomial since di < |Ai | for all 1 ≤ i ≤ n, a contradiction.
Thus, we must have some a ∈ ni=1 Ai so that f (a) ̸= 0.
Q

Our last version of the Combinatorial Nullstellensatz will examine what happens
when f vanishes over not all, but most (in a certain set-theoretic perspective) of the
members of the product. This is known as the Punctured Combinatorial Nullstellensatz
and it builds on the original generalized version. The proof will involve division of
polynomials of many variables, so let us first examine the nature of such an operation.

13
Ball and Serra’s Punctured Nullstellensatz

Sometimes we will not have polynomials which are zero over the whole of a set,
but just on part. These conditions yield the Punctured Nullstellensatz. But, before we may
state the theorem itself, we must state a result about the well-behavedness of multivariate
polynomial division.

Lemma 2.2 (Multivariate Polynomial Division [7]). Let R be an integral domain and
let g1 (x1 ) , . . . , gk (xk ) ∈ R[x1 , x2 , . . . , xn ] be polynomials of one variable with lead-
ing coefficient 1 for 1 ≤ k ≤ n. Then,
Pk
1. If f ∈ i=1 R [x1 , . . . , xn ]·gi (xi ) is a nonzero polynomial in the ideal generated
by g1 (x1 ) , . . . , gk (xk ) over R [x1 , . . . , xn ], then degxi (f ) ≥ deg (gi ) for some
1 ≤ i ≤ k.

2. For a polynomial f ∈ R [x1 , . . . , xn ] there are


w (x) , h1 (x) , . . . , hk (x) ∈ R [x1 , . . . , xn ] such that

k
X
f (x) = hi (x) gi (xi ) + w (x) .
i=1

with degxj (w) < deg (gj ), degxj (gi hi ) ≤ degxj (f ), deg (gi hi ) ≤ deg (f ),
degxj (w) ≤ degxj (f ), and deg (w) ≤ deg (f ) for all 1 ≤ i ≤ k, 1 ≤ j ≤ n.

Furthermore, the polynomial w (x) is unique. We call the polynomial w (x) the mini-
mal coset representative. .

Proof. First, we show the existence of the polynomials hi and w given in (2). As we did
deg(g ) Pdeg(g )−1
before, we let gji ∈ R so that gj (xj ) = xj j − i=0 j gji xij for each j ∈ [1, k].
First, consider g1 . It is clear by normal univariate polynomial division that we can
deg(g )
write f (x) = A1 (x) x1 1 + B1 (x) where degx1 (B1 ) < deg (g1 ),
 
deg(g )
deg (A1 g1 ) = deg A1 x1 1 ≤ deg (f ), deg (B1 ) ≤ deg (f ),

14
 
deg(g )
degxi A1 x1 1 ≤ degxi (f ) degxi (B1 ) ≤ degxi (f ) for all 1 ≤ i ≤ n. Then,
deg(g )
substituting g1 (x1 ) + deg(g1 )−1
g1i xi1 in place of x1 1 , we obtain
P
i=0

deg(g1 )−1
X
g1i xi + B1 (x) .

f (x) = A1 (x) g1 (x1 ) + A1 (x)
i=0

Denote the second term to be f1 (x), so f (x) = A1 (x) g1 (x1 ) + f1 (x) and note that the
previous constraints on degree imply degxi (f1 ) ≤ degxi (f ) for all 1 ≤ i ≤ n and
deg (f1 ) ≤ deg (f ). We repeat this process inductively, assuming we have already
constructed fj (x), 1 ≤ j < k, so that f = ji=1 Ai (x) gi (xi ) + fj (x) we find
P

Aj+1 , Bj+1 ∈ R [x1 , . . . , xn ] so that


 Pdeg(gj+1 )−1 
f (x) = j+1 i
P
i=1 A (x) g i (x i ) + Aj+1 (x) i=0 g(j+1)i x j+1 + Bj+1 (x) with
degxj+1 (Bj+1 ) < deg (gj+1 ), deg (Aj+1 gj+1 ) ≤ deg (fj ), deg (B) ≤ deg (fj ),
 
deg(gj+1 )
degxi Axj+1 ≤ degxi (fj ), and degxi (B) ≤ degxi (fj+1 ) for all 1 ≤ i ≤ n. Then,
Pdeg(g )−1
we denote fj+1 = Aj+1 (x) i=0 j+1 g(j+1)i xij+1 + Bj+1 (x). Hence,
f = j+1
P
i=1 Ai (x) gi (xi ) + fj+1 (x).

Repeating until the case j = k yields f (x) = ki=1 Ai (x) gi (xi ) + fk (x).
P

Labeling Ai = hi and fk = w, we obtain the desired result of (2) with all of the desired
constraints on degree holding true since they held true at each step of the construction.
Next, we show (1) implies fn (x) = w (x) is unique. Suppose w (x) is not unique,
that is,
k
X k
X
f (x) = hi (x) gi (xi ) + w (x) = hi (x) gi (xi ) + v (x)
i=1 i=1

for polynomials h1 , . . . , hk , h1 , . . . , hk , w, v ∈ R [x1 , . . . , xn ] with w (x) ̸= v (x) and the


degree requirements of (2) satisfied. Denote f = f + (g1 (x1 ) , . . . , gk (xk )), where
(g1 (x1 ) . . . gk (xk )) is the ideal generated by the gi ’s. Then, w, v ∈ f , that is w and v are
coset representatives of f . Since w ̸= v, and ideals form an abelian group under addition,
we have w − v ∈ (g1 (x1 ) , . . . , gk (xk )) = ki=1 R [x1 , . . . , xn ] gi (xi ) a nonzero
P

polynomial. Then, (1) implies degxi (w − v) ≥ deg (gi ) for some 1 ≤ i ≤ k, but since

15
degxi (w − v) ≤ max{degxi (w) , degxi (v)} < degxi (gi ) by the first degree constraint of
(2), we have a contradiction. Thus, (1) and (2) imply the polynomial w is unique.
Finally, we show that (1) is indeed true. If k = 1 then the claim is obvious since
f ̸= 0, and f = h1 (x) g1 (x1 ) for some h1 ∈ R [x1 , . . . , xn ] nonzero. So, we proceed by
induction. Assuming the claim is already shown for k − 1 polynomials gi , we prove the
claim for k polynomials gi .
Let f (x) = ki=1 vi (x) gi (xi ), with vi (x) ∈ R [x1 , . . . , xn ]. We can assume that
P

vj (x) ̸∈ (g1 (x1 ) , . . . , gj−1 (xj−1 )), for any 1 ≤ j ≤ k. Otherwise, if


vj gj = j−1
P
i=1 ui gi (xi ) for some polynomials ui ∈ R [x1 , . . . , xn ] we can replace vi with

vi + ui for each 1 ≤ i ≤ k. We can then apply the inductive hypothesis to see the claim is
true.
So, assuming all vj ̸= 0 and each vj is not contained in the ideal
(g1 (x1 ) , . . . , gj−1 (xj−1 )), we find vk ̸≡ 0 mod (g1 (x1 ) , . . . , gk−1 (xk−1 )). Moreover,
letting h ∈ h + (g1 (x1 ) , . . . , gk−1 (xk−1 ))denote the minimal coset representative for any
polynomials h ∈ R [x1 , . . . , xn ], we find vi gi = 0 for all 1 ≤ i ≤ k − 1, so

f = vk (x) + gk (xk ).


Next, note that degxk f ≤ degxk (f ), by applying the definition of f . If we assume
indirectly that degxk (f ) < degxk (gk (xk )), we attain the inequality,

  
degxk vk (x) gk (xk ) = degxk f ≤ degxk (f ) < degxk (gk ) . (2.2)

But, since gj is a polynomial in only the variable xj for each 1 ≤ j ≤ k, it follows that
vk (x) gk (xk ) = vk (x) · gk (xk ) since vk is not in the ideal (g1 (x1 ) , . . . , gk−1 (xk−1 )) as
we noted earlier. Finally, we note gk (xk ) = gk (xk ) so, since we assumed earlier vk ̸= 0,
 
we have degxk vk (x)gk (xk ) ≥ degxk (gk (xk )) by the increasing nature of polynomial

16
degree under nonzero multiplication. This is a contradiction with (2.2), so (1) holds.
Thus, uniqueness of w holds and we see the lemma is thus proven.

Theorem 2.4 (Punctured Combinatorial Nullstellensatz [7]). Let R be an integral do-


main and let A1 , A2 , . . . , An ⊂ R be finite and nonempty. Furthermore, for each
1 ≤ i ≤ n, let Bi ⊆ Ai be nonempty.
Q
Let f (x) ∈ R[x1 , x2 , . . . , xn ] be a polynomial and set gi (xi ) = a∈Ai (xi − a) and
Q
li (xi ) = b∈Bi (xi − b). If

n
! n
!
Y Y
f (a) = 0 for all a ∈ Ai \ Bi (2.3)
i=1 i=1

Qn
but there exists a b ∈ i=1 Bi such that f (b) ̸= 0, then there are polynomial hi ∈
R[x1 , x2 , . . . , xn ] with 1 ≤ i ≤ n such that

n n
X Y gi (xi )
f (x) = gi (xi ) hi (x) + w (x) with w (x) = u (x) . (2.4)
i=1 i=1
li (xi )

with the following conditions holding:


deg (w) ≤ deg (f ) , degxj (w) ≤ degxj (f ) for all 1 ≤ j ≤ n, degxi (w) < |Ai |,
deg (gi hi ) ≤ deg (f ) for 1 ≤ i ≤ n and degxj (gi hi ) ≤ degxj (f ) for 1 ≤ i, j ≤ n.
Consequently ni=1 (|Ai | − |Bi |) ≤ deg (w) ≤ deg (f ).
P

Proof. Lemma 2.2 guarantees there to be polynomials

w (x) , h1 (x) , . . . , hn (x) ∈ R[x1 , x2 , . . . , xn ]

of the desired form and for which the proper conditions hold (the degree restrictions from
implication (2) of the lemma). Our first task, then, is to ensure that the form of w (x) is
that of (2.4). First, let 1 ≤ i ≤ n and without loss of generality select i = 1. We consider
w (x) l1 (x1 ). Then as f (a) = ni=1 gi (ai ) hi (a) = 0 for all
P

17
Qn Qn
a = (a1 , . . . , an ) ∈ i=1 Ai \
Bi , we see w (a) = 0 for all the same a’s. Since
i=1

l1 (a1 ) = 0 for all a1 ∈ B1 , we then have w (a) l1 (a1 ) = 0 for all a ∈ ni=1 Ai . Applying
Q

Theorem 2.2 yields polynomials v1 , . . . , vn ∈ R[x1 , . . . , xn ] such that

n
X
w (x) l1 (x1 ) = gi (xi ) vi (x) . (2.5)
i=1

Now, letting (v1 (x1 ) , . . . , vk (xk )) denote the ideal generated by the vi ’s, 1 ≤ i ≤ k, we
see, by the same construction as in the previous lemma, we can assume without loss of
generality that each vi (x) ̸∈ (v1 (x) , . . . , vi−1 (x)) for all 1 ≤ i ≤ n where vi ̸= 0. Next,
since degxn (w) < deg (gn ) = |An | by lemma 2.2 we have deg (w (x) l1 (x1 )) < |An | for
all n > 1.
Now, once again we let h ∈ h + (g1 (x1 ) , . . . , gn−1 (xn−1 )) be the minimal coset
representative for a polynomial h ∈ R [x1 , . . . , xn ]. Then, applying the same arguments as
in lemma 2.2, we find w (x) l1 (x1 ) = gn (xn ) vn (x). Hence, since w (x) l1 (x1 ) is the
 
minimal coset representative, we have degxn w (x)l1 (x1 ) ≤ degxn (w (x) l1 (x1 )). So,
for all n ≥ 2

   
degxn vn (x) gn (xn ) = degxn w (x) l1 (x1 ) ≤ degxn (w (x) l1 (x1 )) < |An | = degxn (gn ) .
(2.6)
Moreover, since gn (xn ) = gn (xn ), we find vn (x) gn (xn ) = vn (x)gn (xn ), so if
 
vn (x) ̸= 0 then degxn vn (x) gn (xn ) ≥ degxn (gn (xn )), a contradiction with (2.6).
Thus, vn (x) = 0 and, since vn (x) ̸∈ (v1 (x1 ) , . . . , vn−1 (xn−1 )) we must have vn (x) = 0.
So, w (x) l1 (x1 ) = n−1
P
i=1 gi (xi ) vi (x). Then, the same argument will hold for all j > 1,

so we may repeat it until j = 2 to see vi (x) = 0, 2 ≤ i ≤ n. Hence,


w (x) l1 (x1 ) = g1 (x1 ) v1 (x), so g1 (x1 ) | w (x) l1 (x1 ). And, since B1 ⊆ A1 , we know
g1 (x1 )
l1 (x1 ) | g1 (x1 ), so l1 (x1 )
| w (x).
Since we chose i = 1 arbitrarily, we have li (xi ) | gi (xi ) for all 1 ≤ i ≤ n and
gi (xi )
li (xi )
| w (x). Next, let F be the quotient field of R, so theorem 1.3 implies F [x1 , . . . , xn ]

18
is a unique factorization domain. Since each glii(x
(xi )
i)
is in a separate variable we see they can
have no overlap in terms. Hence ni=1 glii(x
(xi )
Q
i)
| w (x) over F [x1 , . . . , xn ]. That is, there is a
u (x) ∈ F [x1 , . . . , xn ] so that

n
Y gi (xi )
w (x) = u (x) . (2.7)
i=1
li (xi )

gi (xi )
Since each li (xi )
is monic and univariate, and w (x) , gi (xi ) , li (xi ) ∈ R [x1 , . . . , xn ], then
Theorem 1.4 implies u (x) ∈ R [x1 , . . . , xn ]. This shows that w is indeed of the form
desired in (2.4), so all that remains to be shown is ni=1 (|Ai | − |Bi |) ≤ deg (w).
P

If w = 0 then (2.4) implies f (a) = 0 for all a ∈ ni=1 Ai since this is the property
Q

of the gi ’s. This is a contradiction as there was atleast one b ∈ ni=1 Bi so that f (b) ̸= 0.
Q

So, we can assume w ̸= 0, and since each gi ̸= 0, 1 ≤ i ≤ n, we have u ̸= 0. So,

n
! n
! n
Y gi (xi ) Y gi (xi ) X
deg (w (x)) = deg (u (x)) + deg ≥ deg = (|Ai | − |Bi |) .
i=1
li (xi ) i=1
li (xi ) i=1

This inequality follows from the definitions of gi and li , namely the fact that
gi (xi ) Q
li (xi )
= a∈Ai \Bi (xi − a) and this completes the proof.

With all three Nullstellensatz now stated and proven, we will move on to an examination
of several smaller combinatorial proofs and a larger proof, the q-dyson conjecture, which
rely on these results.

19
Chapter 3

Simple Combinatorial Proofs

Sumsets

Now, with our the Nullstellensatz stated and proven, let us examine a few simple
results concerning sumsets. First though, we must define what a sumset is.

Definition 3.1 (Sumsets). Let A, B ⊆ G be subsets of an abelian group G.

• We construct a new set A + B, called the sumset of A and B, by A + B = {a +


b : a ∈ A, b ∈ B}

• We construct another new set A ⊕ B, called the restricted sumset of A and B,


by A ⊕ B = {a + b : a ∈ A, b ∈ B, a ̸= b}

• For a polynomial h ∈ Cp [x1 , x2 ] we construct a third new set A h ⊕B, called the
h-restricted sumset of A and B, by

A h ⊕B = {a + b : a ∈ A, b ∈ B, h (a, b) ̸= 0}

• Let A1 , . . . , An ⊆ G be subsets, and h ∈ Cp [x1 , . . . , xn ] a polynomial. We


construct the h-restricted sumset in the same way and denote it

M n
X
n
i=1 Ai ={ ai : ai ∈ Ai , 1 ≤ i ≤ n, h (a1 , . . . , an ) ̸= 0}.
h
i=1

It is worth noting that this notation overlaps heavily with that of the direct sum. For this
reason all uses of ⊕ in this chapter will refer exclusively to the sumset, and all uses
outside of this chapter (namely in chapter 6) will refer to the direct sum.
Our first Theorem, the Cauchy-Davenport Theorem, has been known for quite
some time as the name would suggest. This proof invoking the combinatorial

20
nullstellensatz, however, is relatively new, first appearing in Alon’s original paper on the
subject.

Theorem 3.1 (Cauchy-Davenport Theorem [1]). Given a prime p and nonempty A, B ⊆


Cp , then |A + B| ≥ min {p, |A| + |B| − 1}.

Proof. Suppose |A| + |B| > p. Then, for any element x ∈ Cp , (A) ∩ (x − B) ̸= ∅.
Hence, A + B = Cp (since, if a = x − b for a ∈ A and b ∈ B, we have x = a + b, so
x ∈ A + B ).. Thus, let us assume |A| + |B| ≤ p and suppose indirectly that
|A + B| ≤ |A| + |B| − 2. Let C ⊆ Cp so that A + B ⊆ C with |C| = |A| + |B| − 2.
Q
Next, define f (x, y) = c∈C (x + y − c) and note that we must have f (a, b) = 0 for all
(a, b) ∈ A × B as A + B ⊆ C. Now, note that deg (f ) = |C| = |A| + |B| − 2, and hence
[x|A|−1 y |B|−1 ]f (x, y) = |A|+|B|−2

|A|−1
̸= 0 as |A| + |B| − 2 < p. Hence, by theorem 2.3, we
must have a pair (a, b) ∈ A × B such that f (a, b) ̸= 0, a contradiction. So, we see the
theorem is true.

The next result naturally generalizes the Cauchy-Davenport theorem to h-restricted


sumsets.

Proposition 3.1 ([1]). Let p be a prime, let h (x1 .., xk ) ∈ Cp [x1 , . . . , xk ] be a polyno-
mial, and let A1 , . . . , Ak ⊆ Cp be nonempty. Define m = ki=1 |Ai | − deg (h) − k.
P

|A |−1 |A |−1
If [x1 1 . . . xk k ](( ki=1 xi )m · h (x)) ̸= 0, then
P

M
k
A ≥ m + 1.

i=1 i
h

Consequently, m < p.

Proof. Let A0 , . . . , Ak ⊆ Cp and h (x0 , . . . , xk ) ∈ Cp [x0 , . . . , xk ]. We will prove the


claim for k + 1 sets, thus proving it for k sets since k is arbitrary. Suppose indirectly that
the inequality does not hold. Then choose elements e1 , . . . , em , which are not necessarily

21
distinct, so that h ki=0 Ai ⊆ {e1 , . . . , em }. Let ci = |Ai | − 1 for each 0 ≤ i ≤ k and let
L
Q  Pk 
Q (x) = h (x) m j=1 x
i=0 i − e j . By our construction we must have that Q (x) = 0 for

all x ∈ ki=0 Ai as either h (x) = 0 or ki=0 xi ∈ h ki=0 Ai ⊆ {e1 , . . . , em }. Furthermore


Q P L

deg (Q) = m + deg (h) = ki=0 ci by construction. Next, we see [xc00 . . . cckk ]Q ̸= 0 holds
P

by hypothesis.
Therefore, applying theorem 2.3 yields an a ∈ A such at Q (a) ̸= 0, a
L
contradiction. So, p ≥ h ki=0 Ai ≥ m + 1. Thus, we have m < p.

With this result on h-restricted sumsets proven, we now prove superior bounds on the
standard restricted sumsets. Note that the standard restricted sumset is equivalent to an
Q
h-restricted sumset with the polynomial h (x) = 1≤i,j≤n:i̸=j (ai − aj ).

Theorem 3.2 (Restricted Sumset Theorem[1]). Let p be prime and let A1 , . . . , Ak ⊆


̸ |Aj | for any i ̸= j. If ki=1 |Ai | ≤ p + k+2
P 
Cp be nonempty with |Ai | = 2
− 1, then

k k  
M X k+2
Ai ≥ |Ai | − + 1.

2


i=1 i=1

In order to prove this theorem, let us first state and prove two lemmas concerning the
coefficient of a particular polynomial.

22
Lemma 3.1 (Vandermonde determinant). [10] Let F be a field, and A be the k × k
matrix over F [x1 , . . . , xk ] with

 
1 x1 x1
2
... xk−1
1 
 
1 x2 x2
2 . . . xk−1
2 

A = . . .

 .. .. .. ... .. 
 . . 

 
k−1
1 xk x2k . . . xk

σ(i)
(−1)sgn(σ)
Q P Qk
Then, det (A) = k≥i>j≥1 (xi − xj ) = σ∈Sk i=1 xi , where Sk is the
permutation group on k elements. This polynomial is called the Vandermonde deter-
minant.

Proof. First, note that det (A) ∈ F [x1 , . . . , xk ]. Next, recall that one summand of the
determinant will be produce of the diagonal of A, which is a polynomial of degree
Pk−1 k(k−1)
i=0 i = 2
, and since each xji is unique as a polynomial, we see there is no chance
this summand will cancel with any other. Moreover, all other summands of det (A) will
also consist of the product of one element from each column. Thus, it is clear
k(k−1)
deg (det (A)) = 2
. Next, if xi = xj for some i ̸= j we would have the rows of A not
being linearly independent as there would be duplicates, hence det (A) = 0 in this case.
So, we must have (xi − xj ) | det (A). Since F [x1 , . . . , xk ] is a unique factorization
domain by theorem 1.3 and the factors (xj − xi ) will have no overlap, we see
Q Q
1≤i<j≤k (xj − xi ) | det (A). So, det (A) = F 1≤i<j≤n (xj − xi ) for some
Q 
F ∈ F [x1 , . . . , xk ]. Since deg 1≤i<j≤k (x j − x i ) = k(k−1)
2
, we see deg (F ) = 0 and
as we noted earlier, each product over the diagonals will be distinct with each other and
with coefficient equal to 1, so all terms in det (A) will have coefficient 1. Since all terms
Q
in 1≤i<j≤n (xj − xi ) also have coefficient 1, we see F = 1, so the first equality is shown.
From here it is routinely checked that the second equality follows.

23
Pk k

Lemma 3.2. [3] Let c1 , . . . , ck ∈ Z be nonnegative and define m = 1=0 ci − 2
(it
is trivial that m is nonnegative). Then,

k
!m !
X Y m! Y
C = [xc11 . . . xckk ] xi (xi − xj ) = (ci − cj ) .
i=1 k≥i>j≥1
c1 !c1 ! . . . ck ! k≥i>j≥1

Proof. Once again, we prove the case for k + 1 nonnegative integers, implying the case
for k. Let c0 , . . . , ck ∈ Z and m = ki=0 ci − k+1
P  Q
2
. Then, 1≤j<i≤k (xi − xj ) is the
Vandermonde determinant, which lemma 3.1 prove is equal to
P sgn(σ) Qk σ(i)
σ∈Sk+1 (−1) i=0 xi .
It follows that the coefficient C = σ∈Sk+1 (−1)sgn(σ) (c0 −σ(0))!...(c
m!
P
k −σ(k))!
. Now,
for r ≥ 1 denote φr (s) = s (s − 1) . . . (s − (r − 1)) and φ0 (s) = 1. Next, note that
Q
k≥i>j≥0 (ci − cj ) is precisely the vandermonde determinant on the ci rather than the xi .
 
Finally, let A = (ci )j be the k × k matrix with (ci )j in the i, j’th position for
0 ≤ i, j ≤ k. Sufficient subtraction of linear combinations from the vandermonde matrix
on the ci yields A, so their determinants are equal. Hence, it follows that

m! Y m!
(ci − cj ) = det (A)
c0 ! . . . ck ! k≥i>j≥0 c0 ! . . . ck !
m! X
= (−1)sgn(σ) (c0 )σ(0) . . . (ck )σ(k)
c0 ! . . . ck ! σ∈S
k+1
X m!
= (−1)sgn(σ)
σ∈Sk+1
(c0 − σ (0))! . . . (ck − σ (k))!

= C.

With this out of the way, we now prove theorem 3.2:

24
Proof of Theorem 3.2. For this proof we will take the aforementioned

Y
h (x) = (xi − xj ) .
0≤i<j≤k

Now, let us define ci = |Ai | − 1 and m = ki=0 ci − k+1


P 
2
. Rearranging the assumptions
Pk
of this theorem yields i=0 |Ai | − k+2

2
+ 1 ≤ p and, applying the trivial combinatorial
identity k+2 = k+1
 
2 2
+ (k + 1) yields
Pk k+1

i=0 ci − 2
+ 1 = m + 1 ≤ p (hence m < p). Then

k
! !
X m! Y
[xc00 . . . xckk ] xi h = (ci − cj ) .
i=0
c0 ! . . . ck ! 0≤i<j≤k

We know this product to be nonzero modulo p as ci ̸= cj for i ̸= j by construction and


m < p. Finally, as the coefficient is nonzero and as deg (h) = k+2

2
(as there are k + 2
possible xi ’s and each term of the product will contain two distinct xi ’s so there are k+2

2
Pk
terms each of degree 1), we have m = i=0 ci − deg (h). Hence, applying theorem 3.1
P
yields ⊕h i=0 Ai ≥ m + 1 = ki=0 |Ai | − k+2
k P 
+ 1.

2

A special case of theorem 3.2, when it is taken over only 2 sets, yields an even
tighter bound. This result was originally conjectured by Erdős and Heilbronn and is
named as such. First, though, we must prove another lemma

Lemma 3.3 ([7]). Let F be a field, n ≥ 1, and A1 , . . . , An ⊆ F be finite nonempty


subsets. Let f (x) ∈ F [x1 , . . . , xn ] be a polynomial such that deg (f ) ≤ ni=1 |Ai | − n
P

and  !Pni=1 |Ai |−n−deg(f ) 


h i n
|A |−1
X
x1 1 . . . xn|An |−1 f (x) xi  ̸= 0. (3.1)
i=1

Then,
M X n
n
i=1 Ai ≥ |Ai | − n − deg (f ) + 1.


f
i=1

25
Proof. First of all, we can assume without loss of generality that |F | = ∞, else we can
just replace F with an infinite field extension. Denote f ni=1 Ai = B ′ . Then, suppose
L

indirectly that B ≤ ni=1 |Ai | − n − deg (f ) (in contradiction to the lemma). Since F is
P

infinite, we can always append new elements to B ′ so that equality holds. Denote this new
set B, so B ′ ⊆ B and B = ni=1 |Ai | − n − deg (f ). Let g ∈ F [x1 , . . . , xn ] be the
P

following polynomial !
n
Y X
g (x) = f (x) xi − b .
b∈B i=1

First, by adding degrees we see deg (g) = deg (f ) + |B|, substituting the definition of B
yields deg (g) = ni=1 |Ai | − n. Moreover, for all a ∈ ni=1 Ai we have either f (a) = 0
P Q

or a ∈ B ′ ⊆ B. So, in either case

n
Y
g (a) = 0 for all a ∈ Ai . (3.2)
i=1

Then, observe that any maximal degree monomial in g will be the result of a maximal
|B|
degree monomial of f , times ( ni=1 xi ) , the maximal degree monomials in
P
Q Pn
b∈B ( i=1 xi − b). Moreover,

n
!|B|
h i h i
|A |−1 |A |−1
X
x1 1 . . . xn|An |−1 g (x) = x1 1 . . . x|A
n
n |−1
f (x) · xi .
i=1

Applying the definition of B, we see this is precisely the same coefficient we assumed was
h i
|A |−1 |A |−1
g (x) ̸= 0. So, since deg (g) ≤ ni=1 |Ai | − n,
P
nonzero in (3.1). So, x1 1 . . . xn n
we see equality holds, that is deg (g) = ni=1 (|Ai | − 1), so g satisfies the conditions for
P

theorem 2.3, implying there is a a ∈ ni=1 Ai so that g (a) ̸= 0, a contradiction with (3.2).
Q
L P
So, B = f ni=1 Ai > ni=1 |Ai | − n − deg (f ). So, the lemma holds true.

26
Theorem 3.3 (Erdős-Heilbronn Conjecture [1]). For a prime p and nonempty A, B ⊆
Cp , then |A ⊕ B| ≥ min {p, |A| + |B| − δ} where δ = 3 for the case A = B and
δ = 2 in all other cases.

Proof. First of all, if without loss of generality A = {a}, then


A ⊕ B = {a + b : b ∈ B, a ̸= b}, and since at most one b = a we have
|A ⊕ B| ≥ |B| − 1 = |A| + |B| − 2, so the claim holds.
So, we can assume |A| , |B| ≥ 2. If p = 2, then |A| , |B| = 2, so
A = B = {a1 , a2 } and |A| + |B| − 3 = 1. Of course, by assumption a1 ̸= a2 , so
a1 + a2 ∈ A ⊕ B proving the case for p = 2 since there is at least 1 element in the sumset.
So, we can also assume p ≥ 3.
If |A| + |B| − 1 > p, so |A| + |B| ≥ p + 2, then Theorem 1.5 implies there are two
representations x1 + y1 and x2 + y2 of every x ∈ Cp with x1 , y1 ∈ A, x2 , y2 ∈ B. Suppose
x1 = y1 and x2 = y2 . Then we have 2x1 = 2x2 = x. Since p ≥ 3 is prime, hence odd, we
see 2x1 = 2x2 implies x1 = x2 , a contradiction, so at least one representation will have
two distinct elements, i.e., x1 ̸= y1 , so x1 + y1 = x ∈ A ⊕ B. Since x was arbitrary in Cp ,
we see |A ⊕ B| = p ≤ |A| + |B| − 1, so this case holds as well. Thus, it remains only to
show the case |A| + |B| − 1 ≤ p.
Since |A| , |B| ≥ 2, we must also have |A| , |B| ≤ p − 1 else |A| + |B| − 1 ≤ p
would be violated. Suppose |A| =
̸ |B|, then A ⊕ B = A f ⊕B where
f = x − y ∈ Cp [x, y]. Since deg (f ) = 1 ≤ |A| + |B| − 2 by assumption, we need only
show
 |A|−1 |B|−1
 |A|+|B|−3

x y f (x, y) · (x + y) ̸= 0 (3.3)

in order to apply lemma 3.3 and obtain our result: |A ⊕ B| ≥ |A| + |B| − 2.
Substiting (x − y) in place of f , we wish to examine
(x − y) (x + y)|A|+|B|−3 . Distributing, we see this is simply
 |A|−1 |B|−1 
x y
(x + y)|A|+|B|−3 − x|A|−1 y |B|−2 (x + y)|A|+|B|−3 . Next, applying Theorem
 |A|−2 |B|−1   
x y

27
1.2, we see

   
 |A|−1 |B|−1    |A| + |B| − 3 |A| + |B| − 3
x y ((x − y) · (x + y)|A|+|B|−3 = −
|A| − 2 |A| − 1
(3.4)
(|A| + |B| − 3) (|A| + |B| − 4) . . . (|B|)
= (|A| − |B|) .
(|A| − 1)!
(3.5)

Since |A| =
̸ |B| with both sets being of cardinality at most p − 1, we have |A| − |B| ̸≡ 0
mod p. Similarly, since |A| + |B| − 1 ≤ p, we have all factors of the numerator in (3.5)
being nonzero mod p, so their product is nonzero mod p. Thus, we see for |A| =
̸ |B|,
we have
 |A|−1 |B|−1   
x y f (x, y) · (x + y)|A|+|B|−3 ̸= 0 ∈ Cp .

Next, if A = B, then we may simply subtract one b ∈ B to produce the set B ′ = B \ {b}
with |B ′ | =
̸ |B|, so the previous case implies the claim holds for A ⊕ B ′ , so since
A ⊕ B ′ ⊆ A ⊕ B, the claim holds for A ⊕ B.
Finally, if |A| = |B|, but A ̸= B with A ∩ B = ∅, then A ⊕ B = A + B and
Theorem 3.1 proves the claim true. If A ∩ B ̸= ∅, then A ∩ B ̸= A ∪ B as A ̸= B and
|A ∩ B| =
̸ |A ∪ B|, so we can apply the earlier case to find

|(A ∩ B) ⊕ (A ∪ B)| ≥ |A ∩ B| + |A ∪ B| − 2 = |A| + |B| − 2. (3.6)

Let a + b ∈ (A ∩ B) ⊕ (A ∪ B) and without loss of generality suppose a ∈ A ∩ B and


b ∈ A ∪ B. Then, if b ∈ B, we have a ∈ A and a ̸= b, so a + b ∈ A ⊕ B. If b ∈ A, then
a ∈ B, so a + b ∈ A ⊕ B. So, (A ∩ B) ⊕ (A ∪ B) ⊆ A ⊕ B. Hence (3.6) implies
|A ⊕ B| ≥ |A| + |B| − 2, so the claim holds. Since the claim held for all possible
A, B ⊆ Cp this completes the proof.

28
It is of note that in the last part of the proof, equality actually holds for the sumsets, that is

(A ∩ B) ⊕ (A ∪ B) = A ⊕ B.

This observation may be of use to the reader who chooses to study sumsets in further
detail, but for this theorem inclusion, not equality, is all that was required.

29
A Nice Result for Graphs

Finally, to showcase the versatility of these tools we give two short results from
graph theory, proved using the Combinatorial Nullstellensatz. For these proofs, graphs
will be finite and will not contain loops.

Notation. We define the following standard graph objects. Let G be a graph. Then

• E (G) is the set of all edges in G, and e (G) := |E (G)|.

• V (G) is the set of all vertices in G, and v (G) := |V (G)|.

• For a v ∈ V (G), d (v) is the degree of v, or the number of edges it is part of.

• The incidence matrix of G has rows corresponding to vertices and columns


corresponding to edges. An entry is 1 if its vertex is contained in its edge and it is 0
otherwise.

• A graph G is said to be p-regular if d (v) = p for all v ∈ V (G).

1
P
• d (G) = v(G) v∈V (G) d (v) is the average degree over all v ∈ V (G).

P
Lemma 3.4. Let G be a graph. Then, v∈V d (v) = 2e (G). Consequently d (G) v (G) =
2e (G)

Proof. First, it is clear the claim holds if G contains only 1 edge. Now, we proceed by
induction. If G contains k edges, we may remove one to induce the graph G′ , since the
edge joined two vertices, we see we have removed 2 from the total degree. Hence,
′ ′
P P
v∈V (G) d (v) = v∈V (G′ ) d (v) + 2 = 2e (G ) + 2 = 2 (e (G ) + 1) = 2e (G) by

applying the inductive hypothesis. So the claim is shown.


For the second claim, simply apply the definition of average degree.

30
Theorem 3.4 ([1]). Let p be an odd prime and G a graph with d (G) > 2p − 2 and
maximum degree at most 2p − 1. Then there is a nontrivial p-regular subgraph.

Proof. Let B = (bv,e ) be the incidence matrix of G. For each e ∈ E (G), let xe be an
h i
associated variable taking on values 0 or 1 and define F ∈ Fp (xe )e(G) to be the
polynomial
 !p−1 
Y X Y
F ((xe )e∈E(G) ) = 1 − bv,e xe − (1 − xe ) . (3.7)
v∈V e∈E e∈E

2p−2
The degree of the first product is (p − 1) v (G) = 2
v (G) < 21 d (G) v (G) = e (G). The
degree of the second product is clearly e (G), simply taking the term consisting of the
product over all xe , e ∈ E (G). So, deg (F ) = |E|. Then, since the coefficient of the
maximum order term is nonzero with each xe being to the first power, we can apply
theorem 2.3 with each Ai = {0, 1} (di = 1 < 2 = |Ai | clearly holds) to see there is a
 
(xe )e∈E(G) = x ∈ {0, 1}e(G) such that F (x) ̸= 0, and as F (0)e∈E(G) = 0, we see
x ̸= (0)e∈E(G) .
Lastly, since the second term of (3.7) is clearly zero when evaluated at x, we must
Q  P  P
have v∈V 1 − e∈E bv;e xe = 1. So, e∈E bv;e xe = 0 for every v ∈ V (G).
So, letting H be the subgraph induced by e (H) = {e ∈ E (G) : xe = 1}, we see
P
every v ∈ V (H) has p | e∈E(H) bv;e xe (since we are working mod p) and, since the
maximum degree was at most 2p − 2 and H cannot contain isolated points (since it was
P
generated by edges), we must have p = e∈E bv;e xe . Since xe = 1 for all e ∈ E (H), we
P
see e∈E bb;e xe = dH (v) where dH is the degree in H. So, H is p-regular.

31
Chapter 4

Zeilberger-Bressoud q-Dyson Theorem

The previous results have all been relatively simple applications of the
Combinatorial Nullstellensatz. Now, we show its use in a more complex proof. We will
construct a generalized laurent polynomial based on the g and prove the exact value of its
constant term. To begin, we define a new operation, the q-falling factorial and draw an
analog to the traditional

Definition 4.1 (q-Falling factorial). For a ring R, a q ∈ R, and a k ∈ N0 we define

(x)k;q = (1 − x) (1 − xq) 1 − xq 2 . . . 1 − xq k−1 .


 

For convenience (x)0;q := 1.


As it is unnecessary in the following proofs, we will omit the q from notation so that
(x)k;q = (x)k for an assumed independent variable q.
For simplicity, when referring to this object we will call it the q-falling factorial,
though this covention is not found in the literature and is simply adopted for conve-
nience.

32
The Dyson Conjecture and Lagrange Interpolation

In pursuit of a proof in statistical physics, Dyson himself originally made the


following conjecture in 1963, [9].

Theorem 4.1 (Dyson’s Conjecture). Let a1 , . . . , an ∈ N0 . Then,

ai !
( ni=1 ai )!
 P
0 0
Y x i
[x1 . . . xn ] 1− = Qn .
1≤i<j≤n
x j i=1 (a i !)

Many proofs were later discovered of this theorem buy Gunson (a result which remained
unpublished), Wilson [14], and Good [6]. Good’s proof using univariate Lagrange
interpolation was remarkably simple.
Lagrange interpolation, of course, is another fundamental tool of the polynomial
method. In this thesis we do not prove it (as it is a standard result of calculus) but we will
state the following multivariate form:

Theorem 4.2 (Multivariate Lagrange Interpolation). Let A1 , . . . , An ⊆ F, a field.


Then a laurent polynomial F (x) in n variables with deg (F ) > ni=1 |Ai | has a unique
P

interpolation polynomial of the form

n
∼ X Y Y xi − γ
F (x) = F (c)
ci − γ
c∈ n i=1 γ∈Ai \{ci }
Q
i=1 Ai

for c = (c1 , . . . , cn ).

This is a slightly weaker variation of the theorem presented in [12] (in fact less points are
required to guarantee interpolation).

33
The q-Dyson Theorem

A direct generalization of this, the Zeilberger-Bressoud q-Dyson theorem was


originally posed by Andrews in 1975 [4] and this proof using lagrangian interpolation was
published by Károlyi and Nagy.

Theorem 4.3 (q-Dyson Theorem[8]). First, let a1 , . . . , an ∈ N0 . Then define the fol-
lowing polynomial:

   
Y xi qxj
fq (x) = fq (x1 , x2 , . . . , xn ) := .
1≤i<j≤n
xj ai xi aj

Then,
(q)Pn ai
[x01 . . . x0n ]fq (x) = n i=1 .
Q
j=1 (q)aj

Before we can prove this result, we must examine a related function and prove a result
about its coefficients. Note that this lemma makes implicit use of the Combinatorial
Nullstellenstatz when using lagrangian interpolation.

Lemma 4.1. Let F be a field and F ∈ F [x1 , . . . , xn ] to be a polynomial of degree


d ≤ ni=1 di . Let A1 , . . . , An ⊆ F be subsets with |Ai | = di + 1 for each 1 ≤ i ≤ n,
P

then " #
n
Y X F (c1 , c2 , . . . , cn )
xdi i F =
i=1 c1 ∈A1 ,c2 ∈A2 ,...,cn ∈An
φ′1 (c1 ) φ′2 (c2 ) . . . φ′n (cn )

(z − a) and φ′i being its derivative.


Q
with φi (z) = a∈Ai

Proof. We construct a sequence of polynomials. Let F0 = F and define Fi as follows. Let


Fi (x) = φFii−1
(xi )
as a laurent polynomial in n variables. We see this construction guarantees
Qn di  Qn di 
Fi and for all c ∈ ni=1 Ai we see Fn (c) = F (c) and
Q
x
i=1 i F i−1 = i=1 xi

degxi (Fi ) ≤ di for all 1 ≤ i ≤ n, hence as ni=1 |Ai | > deg (F ), we see lagrangian
P

34
interpolation yields a unique polynomial

n Y xi − γ
∼ X Y
Fn (x) = F (c) .
Qn
i=1
c i − γ
c∈ i=1 Ai γ∈Ai \ci


so that Fn (x) = Fn (x) for all x ∈ Fn . Given φi (z) = a∈Ai (z − a), we have
Q

φ′i (z) = ak ∈Ai i=a∈Ai \ak (z − a), and evaluating at a ci ∈ Ai yields


P Q

φ′i (ci ) = a∈Ai \ci (ci − a). We note that


Q
Qn di 
Fn (x) = c∈Qn Ai Qn Q F (c) (ci −γ) = Qn F (c)
P
i=1 xi i=1 i=1 γ∈Ai \ci φ′ (ci )
. This completes the
i=1 i

lemma.

Now, we prove theorem 4.3.

Proof of theorem 4.3. First, we note that if ai = 0, then (x)ai = (1 − x), hence we may
omit all of these terms as they will not affect the constant. Hence, we know each (relevant)
ai is a positive integer. Let

i −1 aj
aY
!
Y Y
xj − xi q t · x i − xj q t .
 
F (x) =
1≤i<j≤n i=0 t=1

35
1
Denote κ = a a . Then, we see
xi j xj i
Q
1≤i<j≤n

ai   aj  
Y Y xi k−1 Y xj k
fq (x1 , x2 , . . . , xn ) = [ 1− q 1− q ]
1≤i<j≤n k=1
xj k=1
xi
ai aj
Y1 Y k−1
Y k

= a [ x j − x i q x i − x j q ]
xai x j
1≤i<j≤n j i k=1 i=1
ai ja
1 Y Y
k−1
Y
xi − xj q k ]

=Q aj ai [ xj − xi q
1≤i<j≤n xi xj 1≤i<j≤n k=1 k=1
| {z }
=F (x1 ,x2 ,...,xn )

a
Y
xaj i + βi
 
=κ xi j + β j
1≤i<j≤n

a
Y
=κ (xaj i xi j + βij )
1≤i<j≤n

a
Y
= cκ( xaj i xi j ) + β,
1≤i<j≤n

where βi , βj , βij , β are all of the lower order terms from their respective product. We see
the first term is simply a constant, and we see as the exponent of xi will take on all values,
ak , except ai and similarly, the exponent of xj will take on all values except aj . Hence, we
hQ Pn i
n k=1 ak −ai
find [1] fq (x) = x
i=1 i F (x).
Now, we define ni=1 ai = σ, Ai = {1, q, . . . , q σ−ai } and a sequence σ0 = 0 ,
P

σi = i−1
P
j=1 aj for 2 ≤ i ≤ n + 1. It is clear |Ai | = σ − ai + 1 and σn+1 = σ. Now, we aim

to prove that for c = (c1 , . . . , cn ) ∈ ni=1 Ai , we have F (c) = 0 unless ci = q σi for each
Q

1 ≤ i ≤ n.
We will suppose a contradiction, that is there exists a
(c1 , c2 , . . . , cn ) = c ∈ ni=1 Ai such that F (c) ̸= 0 for ci = q αi where 0 ≤ αi ≤ σ − ai .
Q

Then, we see for a pair j > i, we have αj − αi ≥ ai , in particular every pair αj , αi with
i ̸= j is distinct. Then, we see their is a unique ordering π such that
απ(1) < απ(2) < . . . < απ(n) . Then, summing over the inequality απ(i+1) − απ(i) ≥ aπ(i)

36
for 1 ≤ i ≤ n − 1 yields

  
απ(n ) − απ(n−1) + απ(n−1) − απ(n−2) + . . . + απ(2) − απ(1) = απ(n) − απ(1)
n−1
X
≥ aπ(i)
i=1

= σ − aπ(n) .

Then, as απ(1) ≥ 0 and απ(n) ≤ σ − aπ(n) by the original inequality, we find equality must
hold, hence
απ(n) − απ(1) = σ − aπ(n) = σn .

We see this summation works over k < n − 1 terms as well, so απ(k) = σk hence π must
be the identity permutation and we see αk = k−1
P
i=1 ai = σk for all 1 ≤ k ≤ n. . Hence,

we see F (c) = 0 unless each ci = q σi , so the constant

F (q σ1 , . . . , q σn )
[1] fq =
φ′1 (q σ1 ) . . . φ′n (q σn )

by the lemma.
σi

Now, denote τi = 2
+ σi (σ − σi + 1) and we derive a q-falling factorial form of
φ′i (q σi ).

37
Then we see

Y
φ′i (q σi ) = q σi − q t

0≤t≤σ−ai
t̸=σi

i −1
σY
σi t
 σ−a
Yi
q σi − q t

= q −q
t=0 t=σi +1
i −1
σY i −ai
 σ−σ
Y
(−q)t 1 − q σi −t q σi 1 − q t

=
t=0 t=1
i −1
σY i −1
σY σ−σi+1 σ−σi+1
σi
Y Y
t σi −t σi
1 − qt

= (−1) · q · (1 − q )· q ·
i=1 i=0 i=1 i=1
| {z } | {z } | {z } | {z }
σi (σi −1) =(q)σ =q σi (σ−σi+1 ) =(q)σ−σ
Pσi −1 σi
=q i=1 1 =q 2 =q 2 ( ) i i+1

σi
= (−1)σi q ( 2 )+σi (σ−σi+1 ) (q)σi (q)σ−σi+1

= (−1)σi q τi (q)σi (q)σ−σi+1 .

Similarly by double counting the terms and performing simple manipulations, we find
F (q σ1 , . . . , q σn ) in a g form. First, note that we can write a partial sequence from
(σj − σi − ai ) = (σj − σi+1 ) to (σj − σi ) as

i −1
aY Qσj −σi Qσj −σi
σj −σi −t t=1 (1 − q )
t
(1 − q t ) (q)σj −σi
= Qσt=1

1−q = Qσj −σi −ai j −σi+1
= .
t=0 t=1 (1 − q t ) t=1 (1 − q t ) (q)σj −σi+1

Similarly, we find a partial sequence from (σj − σi ) to (σj + aj − σi ) = (σj+1 − σi ) to be

aj Qσj +aj −σi Qσj+1 −σi


Y
σj −σi +t t=1 (1 − q t ) (1 − q t ) (q)σj+1 −σi
= Qt=1

1−q = Qσj −σi σj −σi = .
t=1 t=1 (1 − q t ) t=1 (1 − q t ) (q)σj −σi

Pn Pn  ai
 
Now, let u = i=1 (n − i) ai and v = i=1 (n − i) ai + σi + 2
+ σi (σ − σi+1 )

38
and we see

i −1 aj
aY
!
Y  Y
F (q σ1 , q σ2 , . . . , q σn ) = q σj σi t
q σi − q σj q t

−q q ·
1≤i<j≤n t=0 t=1
 
 
a a
 a −1 a −1 j j

Y Y i Yi Y Y 
(−1)ai  q σi +t (1 − q σj −σi −t ) · q σi 1 − q σj −σi +t 

=
 
 
1≤i<j≤n  |t=0 {z } |t=0 {z }
t=1
| {z }
t=1
| {z } 
 ai

(q)σ −σ (q)
=q ai σi q ( 2 ) = j i =q i (
σ σ−σ i+1 )
=
σj+1 −σi
(q)σ −σ (q)σ −σ
j i+1 j i

Pn (q)σj+1 −σi
[(n−i)(ai σi +(a2i ))+σi (σ−σi+1 )]
Pn Y
i=1 (n−i)ai
= (−1) q i=1

1≤i<j≤n
(q)σj −σi+1
Qn−1 Qn
u i=1 j=i+1 (q)σj+1 −σi
= (−1) q v Qn−1 Qn
i=1 j=i+1 (q)σj −σi+1
Qn−1 Qn+1
i=1 j=i+2 (q)σj −σi
= (−1)u q v Qn Qn
i=2 (q)σj −σi
j=i
Qn+1 Qn−1 Qn Qn−1
j=3 (q) σ −σ ( i=2 j=i+2 (q)σ −σ )( i=2 (q)σn+1 −σi )
= (−1)u q v j
Qn−1 Qn
1 j
Qn−1 Qi+1
i

( i=2 j=i+2 (q)σj −σi )( i=2 j=i (q)σj −σi )


Qn−1 Qn+1
i=2 (q) σ−σ j=3 (q)σj −σ1 (q)σ2 −σ1 (q)σ−σn
= (−1)u q v Qn−1
i
· · (q)σ1 −σ1
i=2 (q) σi+1 −σi
(q)σ 2 −σ 1
(q)σ−σ n
| {z }
1
Qn Qn
(q) i · (q)σ−σ1 · i=1 (q)σi
= (−1)u q v i=2 σ−σ Qn
i=1 (q)σi+1 −σi
n
Y (q)σi (q)σ−σi
= (−1)u q v .
i=1
(q) σi+1 −σ i

39
From here, we see

n
X
(n − i) ai = (n − 1) a1 + (n − 2) a2 + . . . + 0an
i=1
n−1
X n−2
X 1
X 0
X
= ai + ai + . . . + ai + ai
i=1 i=1 i=1 i=1
| {z }
0

= σn + σn−1 + . . . + σ2 + σ1
|{z}
0
n
X
= σi .
i=1

Pn
Hence, u = i=1 σi and the powers of −1 will vanish. Moreover, we show by induction
that the powers of q vanish, that is ni=1 (n − i) ai σi + a2i = ni=1 σ2i . Note that for
P  P 

j = 0, we have 0i=1 (0 − i) ai σi + a2i = 0i=1 σ2i = 0. Similarly, the case n = 1


P  P 

holds. Then, for the general case, we simply expand the right hand side as follows:

n      
X σi σ2 σn
= + ... + (4.1)
i=2
2 2 2
"
1
= (a1 − 1) a1 + (a1 + a2 − 1) (a1 + a2 ) + . . . (4.2)
2
#
+ (a1 + . . . + an−1 − 1) (a1 + . . . + an−1 ) . (4.3)

40
First, grouping the squares and the terms involving −1 and then grouping the
remaining products yields

(a1 + . . . + an−1 − 1) (a1 + . . . + an−1 )

= a2n−1 − an−1 + a2n−2 − an−2 + . . . + a21 − a1 +

+ an−1 (a1 + . . . + an−2 ) + an−2 (a1 + . . . + an−3 ) + an−2 an−1

+ an−3 (a1 + . . . + an−4 ) + an−3 (an−2 + an−1 ) + . . .

+ a3 (a1 + a2 ) + a3 (a4 + . . . + an−1 ) + a2 a1 + a2 (a3 + . . . + an−1 )

+ a1 (a2 + . . . + an−1 ) .

Collecting terms yields

= (an−1 − 1) an−1 + . . . + (a1 − 1) a1 + 2an−1 σn−1 + . . . + 2a2 σ2 .

Applying this identity to (4.3) and adding zero terms where necessary gives the result:

n       
X σi an−1 an−2
= an−1 σn−1 + + 2 an−2 σn−2 +
i=2
2 2 2
  
a2
+ . . . + (n − 2) a2 σ2 +
2
n−i   
X ai
= (n − i) ai σi + .
i=1
2

41
Assembling all results yields

n
F (q σ1 , . . . , q σn ) Y (q)σi (q)σ−σi
[1] fq (x) = ′
=
φ1 (q σ1 ) . . . φ′n (q σn ) i=1 (q)σi (q)σ−σi+1 (q)σi+1 −σi
(q)σ−σ1
= Qn
i=1 (q)σi+1 −σi
(q)
= Qn σ
i=1 (q)ai
(q)a1 +a2 +...+an
= .
(q)a1 . . . (q)an

42
Chapter 5

Zero-Sum Theory

Definitions and Notation

As its name would imply, zero-sum theory is a branch of combinatorial number


theory aiming to categorize exactly when a sequence contains a subsequence which sums
to zero (mod some p). First, though, we must define a sequence and give a few basic
results,

Definition 5.1. For an alphabet A (an arbitrary set of symbols), we define a sequence
to be the string formed by the finite formal commutative multiplication of elements
a ∈ A. That is, if S is a sequence in A, then S = a1 · a2 · . . . · an = ni=1 ai for
Q

some ai ∈ A where the ai need not be distinct for different i. For the alphabet Z,
S0 = 1 · 2 · 3 · 3 · 3 · 4 = 4 · 3 · 2 · 3 · 1 · 3 is a sequence of length 6.
Since the multiplication is commutative, we generally fix an indexing of the sets where
all duplicate terms have adjacent indices. Under this convention, we then will denote
repeated terms by superscript, i.e. if S = s1 ·. . .·sn and si = si+1 , then we may instead
write S = s1 · . . . · s2i · si+2 · . . . · sn . This of course can be repeated to produce larger
superscripts. For example, the sum S0 from the previous example is equal to 1 · 2 · 33 · 4
under this notation.
For a sequence S we denote |S| to be the length of S.
For a sequence S = ni=1 si and an indexing set I ⊆ [1, n] we let S (I) = i∈I si be
Q Q

the subsequence indexed by I. If T is a subsequence of S, we write T | S.


For a sequence S = ni=1 si and an indexed set I ⊆ [1, n] let S (I) = T . Then, we
Q

denote the sequence S ([1, n] \ I) = ST −1 . This sequence contains all the terms which
were in S, but not T . For an abelian group (G, +), we say a sequence S = ni=1 si in
Q

G is zero-sum if ni=1 si = 0G . Instead of writing ni=1 si , we often denote the sum


P P

over S to be σ (S), so S is zero sum if σ (S) = 0.

43
There are alternative formulations of all of the following theorems using multi-sets, free
abelian monoids, or other algebraic objects which encode the same behavior, but the
sequence approach is the most standard.
Next, we define a few terms relating to abelian groups.

Definition 5.2. Recall a finite abelian group G has a unique decomposition into cyclic
groups G = Cq1 ⊕ Cq2 ⊕ . . . ⊕ Cqr . where 1 < q1 < . . . < qr and q1 | q2 | . . . | qr . In
this case, we of course know |G| = ri=1 qi .
Q

We define Rank (G) = r and exp (G) = qr to be the rank and exponent of the group
respectively. These two constants, as well as |G|, give three different notions of the
size of a group.

As it turns out, most theorems in zero-sum theory will be concerned with the exponent of
the abelian group G. As we alluded to earlier, our primary interest with sequences is when
they contain subsequences which are zero-sum.

44
Definition 5.3. Let G be an abelian group, let S be a sequence in G, and let p, q ∈ N.
We define the following constants:

• Np (S) is the number of subsequences, T | S, where |T | = p and σ (T ) = 0.

• Similarly, Np,q (S) is the number of pairs of disjoint subsequences T1 , T2 | S,


where |T1 | = p, |T2 | = q, and σ (T1 ) = σ (T2 ) = 0.

• Let T1 | S be an arbitrary subsequence with |T1 | = q and σ (T1 ) = 0. Then,


Npq (S) is the subsequences T2 | T1 where |T2 | = p and σ (T2 ) = 0.

• D (G) is the minimal number such that any sequence, S in G, of length atleast
D (G) will contain a nonempty zero-sum subsequence. We call this the Daven-
port constant for the group G.

• η (G) is the minimal number such that any sequence S in G with |S| ≥ η (G)
will contain a nonempty susbequence T | S with |T | ≤ exp (G) and σ (T ) = 0.
We sometimes call such zero-sum subsequences T with |T | ≤ exp (G) short
zero-sum subsequences.

• s Cpk is the minimal number such that any sequence, S in G = Cpk with |S| ≥

s Cpk will contain a subsequence T | S with |T | = p = exp (G) and σ (T ) = 0.
 
• sn Cpk is the number such that any sequence S in G = Cpk with |S| ≥ sn Cpk
will contain a subsequence T | S with |T | = n and σ (T ) = 0.
Pr
• If G = Cq1 ⊕ . . . ⊕ Cqr , then D∗ (G) = 1 + i=1 (qi − 1).

Basic Proofs and Prerequisites

With these definitions out of the way, we now prove a few simple theorems
concerning zero-sum sequences before showcasing two larger results. First, it is quite
trivial to prove a loose upper bound on D (G), namely the order of G itself.

45
Proposition 5.1. Let G be a finite abelian group. Then, D (G) ≤ |G|.

Qk
Proof. Let S = s1 . . . sn be a sequence in G with |S| ≥ |G|. Let Sk = i=1 si for each
k ∈ [1, n], so Sn = S and S1 = s1 . Then, if σ (Sk ) = 0 for some k, we are done.
Otherwise, the pigeonhole principle proves there are two indices, n1 < n2 so that
σ (Sn1 ) = σ (Sn2 ) since there are only |G| − 1 nonzero terms in G but at least |G|
subsequences Sk . Then, T = ni=1+n
Q 2
s has σ (T ) = σ (Sn2 ) − σ (Sn1 ) = 0, so T is
1 i

zero-sum and the claim is shown.

It is also quite easy to prove that D∗ (G) is a lower bound.

Proposition 5.2. Let G be a finite abelian group. Then, D (G) ≥ D∗ (G).

Proof. Suppose G = ri=1 Cqi , with ei being a generator for each Cqi . Then
L

T = ri=1 eqi i −1 is a sequence of length ri=1 qi − 1 with no nontrivial zero-sum


Q P

subsequences. To see this, note that a subsequence can only be zero-sum if each of its
coordinates within the direct sum is zero-sum. Since ord (ei ) = qi , this cannot happen, so
T can have no nontrivial zero-sum subsequences. Since |T | + 1 = D∗ (G), we see D (G)
is at least D∗ (G), so the claim is shown.

These two propositions of course imply D (Cp ) = p. The next chapter will generalize this
notion to arbitrary p-groups.
There is one more theorem in basic zero-sum theory which we will not prove, but
is deserving of mention, the davenport constant of a rank 2 abelian group.

Theorem 5.1. Let G be a finite abelian group of rank 2, that is G = Cn1 ⊕ Cn2 with
n1 | n2 . Then, D (G) = D∗ (G).

The proof of this theorem is relatively elementary, though it does require some
homological algebra, so for this reason we exclude it from the current survey.

46
Finally, we provide proofs of the multiplicity of η, D, and s,

Theorem 5.2. Let m, n ∈ N and suppose it is already known that η (Cn2 ) = 3n − 2 and
2 2
η (Cm ) = 3m − 2. Then η (Cnm ) = 3nm − 2.

Note that it suffices only to prove this is an upper bound as a counterexample of length
3mn − 3 is easily produced.

2
Proof. First, let G = Cmn and note that 3mn − 2 = m (3n − 3) + 3m − 2. Let
2 2
φ : Cmn → Cm , x 7→ φ(x) = nx be the left multiplication by n map. Now, let
Q3mn−2
S = i=1 gi be a sequence in G of length 3mn − 2. Define
φ (S) := 3mn−2 φ (gi ) = 3mn−2 2 2
Q Q
i=1 i=1 ngi ∈ Cm . Since φ (S) is a sequence in Cm of length
2
greater than 3m − 2 = η (Cm ), we find a zero-sum sequence T1 | φ (S) with
) = m. So, φ (S) T1−1 ≥ 3mn − 2 − m = m (3n − 2) + 3m − 2. We
2

|T1 | ≤ exp (Cm
repeat this process inductively, assuming we have found a subsequence
Tk | φ (S) T1−1 T2−1 . . . Tk−1
−1
with
φ (S) T1−1 . . . T −1 ≥ m (3n − 3 − k) + 3m − 2 ≥ η (Cm
2
k ) we find another zero-sum
subsequence Tk+1 | φ (S) T1−1 . . . Tk−1 with
φ (S) T1−1 . . . T −1 ≥ m (3n − 3 − k − 1) + 3m − 2 for 1 ≤ k ≤ 3n − 4. So, in total we

k+1

find 3n − 3 disjoint zero-sum subsequences T1 . . . T3n−3 | φ (S). Finally, since


φ (S) T1−1 . . . T3n−3
−1

≥ 3m − 2, we find one more zero-sum subsequence
T3n−2 | ST1−1 . . . T3n−3
−1

Qℓi (j) (j)


Now, note that Ti = j=1 ngi where each ℓi ≤ m and gi | S, 1 ≤ j ≤ ℓi , is a
unique element in S for each 1 ≤ i ≤ 3n − 2 (because the Ti were disjoint). Then, note
that each σ(φ−1 (Ti )) = αi m for some αi ∈ Cn2 . So, we can associate each σ (φ−1 (Ti ))
with an element αi ∈ Cn2 . Then,

X = α1 . . . α3n−2

47
is a sequence in Cn2 with |X| = 3n − 2 > η (Cn2 ) so we find a I ⊆ [1, 3n − 2] with |I| ≤ n
and σ (X (I)) = 0. So finally we see

Y
Y = φ−1 (Ti )
i∈I

is a zero-sum subsequence with |Y | ≤ |I| · max{|Ti | : 1 ≤ i ≤ 3n − 2} ≤ nm. To see the


sequence is zero-sum, recall σ (φ−1 (T )) = αi m, so
−1
Q  P
σ i∈I φ (Ti ) = i∈I αi m = nm ≡ 0 mod nm. So the claim is shown.

While this result is nice, we actually wish to use the result for the s constant. The
argument follows the same line of reasoning, with inequalities replaced with equalities.

2
Theorem 5.3. Let m, n ∈ N and suppose it is already known s (Cm ) = 4m − 3 and
2
s (Cn2 ) = 4n − 3. Then s (Cmn ) = 4mn − 3.

Once again, it suffices only to show 4mn − 3 is an upper bound as


(0, 0)mn−1 (0, 1)mn−1 (1, 0)mn−1 (1, 1)mn−1 is a counterexample of length 4mn − 4.

2
Proof. The proof follows the same structure. Let G = Cmn and note
2 2
4mn − 3 = m (4n − 4) + 4m − 3. Let φ : Cmn → Cm , x 7→ nx be the left multiplication
by n map. Let S = 4mn−3 gi be a sequence in G and let φ (S) = 4mn−3
Q Q
i=1 i=1 ngi . Since
2 2
φ (S) is a sequence in Cm of length greater than s (Cm ) we find a zero-sum sequence
T1 | φ (S) with |T1 | = m so |φ (S) T −1 | = m (4n − 4) + 4m − 3. Then, we repeat this
process inductively, given a zero-sum subsequence Tk | φ (S) T1−1 . . . Tk−1
−1
with |Tk | = m
we find a new zero-sum subsequence Tk+1 | φ (S) T1−1 . . . Tk−1 so that |Tk+1 | = m, so
φ (S) T1−1 . . . T −1 = m (4n − 3 − k − 1) + 4m − 3 for 1 ≤ k ≤ 4n − 5. At the end we

k+1

have 4n − 4 disjoint subsequences T1 , . . . , T4n−4 | φ (S). Since


2
|φ (S) T1 . . . T4n−4 | = 4m − 3 = s (Cm ), we find one more zero-sum subsequence
T4n−3 | φ (S) T1−1 . . . T4n−4
−1
.

48
Qm (j) (j)
Then, noting that each Ti = j=1 ngi where gi | S, 1 ≤ j ≤ m are unique
terms in S for each 1 ≤ i ≤ 4n − 3. Note that σ (φ−1 (Ti )) = αi m for some αi ∈ Cn2 . So,
the sequence
X = α1 . . . α4n−3

is a sequence in Cn2 with |X| = s (Cn )2 , so we find I ⊆ [1, 4n − 3] so that |I| = n and
|σ (X (I))| = 0. Finally, we see

Y
Y = φ−1 (Ti )
i∈I

is a zero-sum subsequence with |Y | = |I| · m = nm. To see this sequence is zero-sum


recall each σ (φ−1 (Ti )) = αi m, so σ −1
Q  P
i∈I φ (Ti ) = i∈I αi m = nm ≡ 0 mod nm.
So, the claim is shown.

There is one more theorem in basic zero-sum theory which we will not prove, but
is deserving of mention, the davenport constant of an arbitrary rank 2 abelian group.

Theorem 5.4. Let G be a finite abelian group of rank 2, that is G = Cn1 ⊕ Cn2 with
n1 | n2 . Then, D (G) = D∗ (G).

The proof of this theorem is relatively elementary, following a similar trajectory to


theorem 5.2, but it is outside the scope of this thesis so we omit it.
Finally, we now prove the most important result in preparation for our two final
theorems, that both η and s are multiplicative. A similar result can be produced for the
davenport constant D. Indeed a mirror theorem to theorems 5.2 and 5.3 provides the
result. For now, though, we are primarily concerned with the s constant and secondarily
concerned with the η constant.


Corollary 5.1. If η Cp2 = 3p − 2 for all primes p, then η (Cn2 ) = 3n − 2 for all n ∈ N.

Similarly, if s Cp2 = 4p − 3 for all primes p, then s (Cn2 ) = 4n − 3 for all n ∈ N.

49
Qℓ
Proof. In either case, n = i=1 pαi i for some primes pi and αi ∈ N, so applying theorems
5.3 or 5.2 repeatedly will yield the claims.

With this short introduction to zero-sum theory, we now aim to prove two larger results.
The first will show that for all finite abelian p-groups G, we have D (G) = D∗ (G). The

second will show that s Cp2 = 4p − 3 for all p ∈ N.

50
Chapter 6

Davenport constant of finite abelian p-groups

Advancing the theme of zero-sum sequences we now aim to determine the


davenport constant for finite abelian p-groups. The formal statement is as follows.

Theorem 6.1. Let p be prime, G = Cq1 ⊕ Cq2 ⊕ . . . ⊕ Cqd with qi = pmi for a prime p
and integers mi > 0. Suppose S = ℓi=1 gi is a sequence in G. If ℓ ≥ D∗ (G) = 1 +
Q
Pd
i=1 (qi − 1), then there is a non-empty subset I ⊆ [1, ℓ] so that σ (S (I)) = 0.

The proof of this theorem will once again employ polynomial methods. This time, we do
not make use of the Combinatorial Nullstellensatz or any of its variants. Instead, we find
polynomials corresponding to which indices get included in I and show that they have
sufficiently small degree. Once again, we find it suffices to show that D∗ is merely a lower
bound of D (G) as (1, 0, . . . , 0)q1 −1 (0, 1, 0, . . . , 0)q2 −1 . . . (0, 0, . . . , 1)qd −1 is a sequence of
length di=1 (qi − 1) which necessarily has no zero-sum subsequences.
P

Map Functors

For the remainder of this proof, F will be a field, ℓ ≥ 1 will be an integer and
{0, 1}ℓ will denote the set of sequences consisting of all 0’s and 1’s taken from the vector
space F ℓ .


Definition 6.1. For a field F and ℓ ≥ 0, define Map {0, 1}ℓ , F = {f : {0, 1}ℓ →
F |f is a function}.


Proposition 6.1. Map {0, 1}ℓ , F is an F -vector space with pointwise addition and

scalar multiplication. That is, for f, g ∈ Map {0, 1}ℓ , F and α ∈ F we define
(f + g) (x) = f (x) + g (x) and (αg) (x) = α (g (x)).

Proof. This statement is routinely checked. First we show the map space to be an abelian

51
group, then we verify composition produces a F -module and subsequently an F -vector
space. Associativity of the map space follows from associativity of F :

(f + (g + h)) (x) = f (x) + (g + h) (x)

= f (x) + g (x) + h (x)

= (f + g) (x) + h (x)

= ((f + g) (h)) (x) .

We choose the identity to be the zero map, denoted 0.



Additive inverses are satisfied by negation. For f ∈ Map {0, 1}ℓ , F , we find

−f = −1 · f ∈ Map {0, 1}ℓ , F and f + (−f ) = 0.
Finally, commutativity is also inherited from F :

(f + g) (x) = f (x) + g (x) = g (x) + f (x) = (g + f ) (x) .


Now, we verify Map {0, 1}ℓ , F is an F -module. Let r, s ∈ F and 1 ∈ F to be the
multiplicative identity of F . We find, for all x ∈ {0, 1}ℓ , (1f ) (x) = 1f (x) = f (x).
Furthermore, left associativity holds :

((rs) f ) (x) = (rs) f (x) = r ((sf ) (x)) = (r (sf )) (x) .

52
Lastly, left distributivity also holds:

((r + s) f ) (x) = (r + s) f (x) = rf (x) + sf (x)

= (rf ) (x) + (sf ) (x)

= (rf + sf ) (x)

and (r (f + g)) (x) = r ((f + g) (x))

= r (f (x) + g (x))

= rf (x) + rg (x)

= (rf ) (x) + (rg) (x)

= (rf + rg) (x)

As these properties hold for composition (multiplication) and the addition operation forms

an abelian group, we have thus verifies Map {0, 1}ℓ , F is an F -module. Since F is a
field, it is then a F -vector space.

Q
Definition 6.2. For an indexed set I ⊆ [1, ℓ] we call a monomial of the form i∈I xi a
special monomial.

Proposition 6.2. The set of all special monomials on F [x1 , . . . , xn ], when interpreted

as functions {0, 1}ℓ → F , forms a basis for Map {0, 1}ℓ , F .

Proof. Recall |P ([1, ℓ])| = 2ℓ where P (X) denotes the power set of X. Thus,
enumerate all possible subsets of [1, ℓ] as J0 , J1 , . . . , J2ℓ −1 where J0 = ∅, J1 , . . . , Jℓ are
singletons, Jℓ+1 , . . . , Jℓ+(ℓ) are 2-tuples, and so on. In general, JPk−1 (ℓ)+1 , . . . , JPk (ℓ)
2 i=1 i i=1 i

are k-tuples for 1 ≤ k ≤ ℓ. We see in this way, all Ji are nonempty and ordered by
 
increasing cardinality for i ∈ 1, 2l − 1 . Next denote all possible special monomials as

53
Q
h0 , h1 , . . . , h2ℓ−1 ∈ F [x1 , . . . , xn ] such that hk = i∈Jk xi for each 0 ≤ k ≤ ℓ. Finally,
note that when interpreted as a function, hk (x) ∈ {0, 1}ℓ for x ∈ {0, 1}ℓ .
First, we must show {hk : 1 ≤ k ≤ 2m − 1} is linearly independent. Let ai ∈ F ,
1 ≤ i ≤ 2ℓ − 1, so that
ℓ −1
2X
a0 h0 + ai hi = 0.
i=1

xi = 1 and hi (0) = 0, 1 ≤ i ≤ 2ℓ − 1, we see a0 = 0. Now, let en


Q
Noting that h0 = i∈∅

be the vector having 1 in the n’th position and 0 elsewhere. Then, hi (en ) = 1 if and only
if hi = xn or i = 0. So, we see
 
ℓ −1
2X
0= ai hi  (en ) = an
i=1

P 
2ℓ −1
where an = [xn ] i=1 ai hi . So, for all 0 ≤ i ≤ ℓ, we see ai = 0, since these are all the
special monomials having at most one variable.
Now, for n1 , n2 ≤ ℓ let en1 ,n2 be the vector having 1 in the n1 ’th and n2 ’th
positions and 0 in the others. Similarly to before, we see hi (en1 ,n2 ) = 1 if and only if
hi ∈ {xn1 , xn2 , xn1 xn2 }. Since an1 = an2 = 0 (as n1 , n2 ≤ ℓ) then

 
ℓ −1
2X
0= ai hi  (en1 ,n2 ) = an1 ,n2
i=1

P

2ℓ −1
ai hi . So, since there are 2ℓ possible

where once again, an1 ,n2 = [xn1 xn2 ] i=1

monomials with 2 variables, we have ai = 0 for ℓ + 1 ≤ i ≤ m + 2ℓ . Combining our




results yields ai = 0 for 0 ≤ i ≤ ℓ + 2ℓ .




Continuing inductively, assume for all monomials of at most j variables, we find


 n1 n 
x1 . . . xj j = an1 ,...,nj = 0. Then, let en1 ,n2 ,...,nj+1 , where n1 < n2 < . . . < nj+1 ≤ ℓ, be
the vector with 1 in the n1 , n2 , . . . , nj+1 coordinates and 0 elsewhere. Then, applying the
previous induction set, we see only the special monomials hi on j + 1 variables will have

54

the property that hi en1 ,n2 ,...,nj+1 = 1 for some n1 , n2 , . . . , nj+1 . So,
P ℓ 
2 −1 
0= i=1 a h
i i en1,n2 ,...,nj+1 = an1 ,n2 ,...,nj+1 where
  P2m−1 
an1 ,...,nj+1 = xn1 xn2 . . . xnj+1 i=1 a i i . So, ak = 0 for all k with
h
Pj ℓ
 Pj+1 ℓ
i=1 i + 1 ≤ i ≤ i=1 i .

Since ℓi=1 ℓi = 2ℓ − 1, we see this process will terminate after ℓ iterations


P 

having proven all ai = 0, 0 ≤ i ≤ 2m − 1. Thus, we see {hi : 0 ≤ i ≤ 2ℓ − 1} is linearly


independent.

Next, we show the set spans Map {0, 1}ℓ , F . To do this we generalize the prior
argument, let eα be the vector having a 1 in the n’th position if n ∈ Jα and 0s elsewhere.
Then, we note that hβ (eα ) = 1 if and only if Jβ ⊆ Jα . Similarly,
 if eα is the vector
 1, Ji ⊆ Jα

with 1 in the i’th position if i ∈ Jα and 0’s elsewhere, then hi (eα ) =
 0,
 Ji ̸⊆ Jα
We use this fact to build a polynomial p which agrees with f for all x ∈ {0, 1}ℓ . Note that,
 
by construction, every x ∈ {0, 1}ℓ has a unique α ∈ 0, 2ℓ − 1 so that eα = x. First, we
construct p0 which agrees with f for all eα with |Jα | ≤ 0. This, of course, is simply
p0 (x) = f (0) h0 (x). Next, we construct p1 which agrees with f for all eα with |Jα | ≤ 1.
This will be p1 (x) = f (0) h0 (x) + ℓα=1 (f (eα − f (0)) hα (x)). Since hi (eα ) = 0
P

unless α = i, we see p1 (eα ) = f (0) + (f (eα − f (0))) = f (eα ).


Next, we construct p2 which will agree with f on all eα with |Jα | ≤ 2. For this
purpose, let en1 ,n2 be the vector having 1 in the n1 and n2 positions and 0 elsewhere,
Jn1 ,n2 = {n1 , n2 }, and hn1 ,n2 = xn1 xn2 . Then, the polynomial will be


X
p2 (x) = f (0) + (f (eα ) − f (0)) hα (x)
n1 =1
X
+ (f (en1 ,n2 ) − f (en1 ) − f (en2 ) + f (0)) hn1 ,n2 (x) .
1≤n1 <n2 ≤ℓ

It is clear that for all vectors, x, with at most 2 nonzero coordinates, we have
p2 (x) = f (x). We continue on in this manner, inductively construction polynomials over

55
all sets of size ≤ 3, ≤ 4, and eventually all sets Jα with |Jα | ≤ ℓ. This, of course, covers
all the Jα and our polynomial will be:
  
ℓ −1
2X |Ji |
X X
(−1)i−j
 
p (x) := pℓ (x) = f (0) h0 (x) + 
  f (ek )
 hi (x) .

i=1 j=0 Jk ⊆Ji
|Jk |=j

This polynomial has the property that p (x) = f (x) for all x ∈ {0, 1}ℓ .

So, p = f and we see {hi : 0 ≤ i ≤ 2ℓ − 1} spans Map {0, 1}ℓ , F .
Finally, we conclude {hi : 0 ≤ i ≤ 2ℓ − 1} forms a basis for the map space.

Proposition 6.3. Let f ∈ Q [x1 , . . . , xℓ ] be a polynomial taking on integer values when


 
evaluated at x ∈ {0, 1}ℓ and let hi , i ∈ 0, 2ℓ − 1 be all possible special monomials
P
ordered we did previously. Then, there is a g ∈ Z [x1 , . . . , xℓ ] such that g = i∈I bi hi
 
for some I ⊆ 0, 2ℓ − 1 and bi ∈ Z so that, for all x ∈ {0, 1}m , f (x) = g (x) when
evaluated as functions.

Proof. Applying proposition 6.2 yields a g having the desired property that g (x) = f (x)
for x ∈ {0, 1}m where

 
2ℓ −1
X X |Ji |
X
i−j

g (x) = f (0) h0 (x) + 
 (−1) f (e k )  hi (x) .

i=1 j=0 Jk ⊂Jα
|Jk |=j

P|Ji |
(−1)i−j
P
Moreover, since f (0) ∈ Z and j=0 Jk ⊂Jα f (xk ) ∈ Z we see g has integer
|Jk |=j
coefficients. Finally, each hi is a monomial, so g is the desired sum of monomials.

Combinatorial Identities

Finally, before we may prove the main theorem, we must show a few binomial
identities which we will make use of later.

56
Proposition 6.4. Let q = pm for p ≥ 2 being prime, m ≥ 1, and y ∈ Z. Then,

 0 mod p, y ̸≡ 0 mod q
  
y−1
≡ . (6.1)
q−1  1 mod p,
 y≡0 mod q

n

Additionally, for n ∈ Z, k
∈Z

Proof. First, recall


 
n n (n − 1) . . . (n − (k − 1))
= .
k k (k − 1) . . . (2) (1)

Hence, if k ≤ n, we find nk ∈ Z. Else, if 0 ≤ n ≤ k − 1, then nk = 0 ∈ Z. So, for all


 

k ∈ N, n ∈ Z, we have nk ∈ Z.


Next, note that

 
y−1 (y − 1) (y − 2) . . . (y − (q − 2)) (y − (q − 1))
= . (6.2)
q−1 (q − 1) (q − 2) . . . (2) (1)

Now, we prove equation (6.1). First, suppose y ≡ 0 mod q, so y ≡ q ≡ 0 mod p, then


y − 1 ≡ q − 1 mod p, y − 2 ≡ q − 2 mod p and so on. In general, y − k ≡ q − k
mod p for 1 ≤ k ≤ q − 1. So, reducing mod p we find the all factors between the
numerator and denominator cancel, yielding y−1

q−1
≡ 1 mod p.
Next, we prove the case y ̸≡ 0 mod q. Note there are q − 1 factors in the
numerator and denominator of equation (6.2). Since q − 1 ≡ p − 1 mod p, and Cp is
cyclic, we find both the numerator and denominator will range over all possible values k
mod p, k ∈ [0, p − 1] atleast p1 (pm − p) = pm−1 − 1 times. This, however, only counts
pm − p terms so there must be p − 1 terms remaining. These terms will once again range
over all possible values k mod p except one. So, in total there is one value k mod p
which is taken on pm−1 − 1 time and all other values k mod p will be taken on pm−1
times. Since we only care about factors of p, we need only check which case the value 0

57
mod p falls into. Notice that if y ̸≡ 0 mod q, we find a unique representative
z ∈ [1, q − 1] so that y ≡ z mod q. Then, we see y − 1 ≡ z − 1 mod p, y − 2 ≡ z − 2
mod p, . . . , y − (q − 1) ≡ z − q + 1 ≡ z + 1 mod p, so the value which is taken on only
pm−1 − 1 times will be z mod p. Since the denominator has q − 1 ≡ p − 1
mod p, q − 2 ≡ p − 2 mod p, . . . , 1 ≡ 1 mod p, we see 0 mod p will be the value
which is taken on only pm−1 − 1 times. So, since z ̸= 0, we have pm−1 terms in the
numerator which are 0 mod p and only pm−1 − 1 terms in the denominator which are 0
mod p. So, after cancelling out like terms, we see p | y−1 , so y−1
 
q−1 q−1
≡ 0 mod p.

The Davenport Constant

Now, with the use of all the prior lemmas and propositions, we will state and prove
the davenport constant for finite p-groups.

Theorem 6.2 (Davenport Constant of Finite p-group). Let p be prime, G = Cq1 ⊕


Cq2 ⊕ . . . ⊕ Cqd with qi = pmi for a prime p and integers mi > 0. Suppose S = ℓi=1 gi
Q

is a sequence in G. If ℓ ≥ D∗ (G) = 1 + di=1 (qi − 1), then there is a non-empty


P

subset I ⊆ [1, ℓ] so that σ (S (I)) = 0.

Proof. To begin, assume indirectly there is a sequence S = g1 . . . gℓ with ℓ ≥ D∗ (G) and


no nontrivial T | S with σ (T ) = 0.
 
(1) (2) (d) (j)
Then, we note each gi = gi , gi , . . . , gi with each gi a representative in
Cqj , 1 ≤ i ≤ ℓ. Let x = (x1 , . . . , xℓ ) and define P ∗ : {0, 1}ℓ → Z by
P 
ℓ (j)
j=1 gi xi −1

Qd
. Note that we can also interpret P ∗ as a polynomial in

P (x) = j=1 q −1
j

Q [x1 , . . . , xℓ ]. We will denote this as the polynomial P ∈ Q [x1 , . . . , xℓ ]. In general, for a


polynomial f , we will denote the function created by evaluating this polynomial as f ∗ .
(j)
By construction, ℓi=1 ai xi ∈ Z [x1 , . . . , xℓ ] when interpreted as polynomials
P

over Z. Moreover, all such monomials are integer valued on {0, 1}ℓ . Next, we identify
each x = (x1 , . . . , xℓ ) ∈ {0, 1}ℓ with a subsequence T (x) | S where
Q
T (x) = 1≤i≤ℓ;xi =1 gi .

58
Fix y ∈ {0, 1}ℓ . If T (x) is a nontrivial zero-sum susbequence, then we have
Pℓ (j)
i=1 gi xi ≡ 0 mod qj for each j ∈ [1, d]. This is because an element is zero in G,
precisely when each of its coordinates is zero in its respective Cqi . So we see proposition
6.3 implies  
d  Pm (j)
i=1 gi xi − 1
Y 
P ∗ (x) = ≡1 mod p.
i=1
qj − 1
Pℓ (j0 )
On the other hand, if T (x) is not zero-sum, we find a j0 ∈ [1, d] so that i=1 gi xi ̸≡ 0
mod qj0 . So, we may applying proposition 6.3 once again to see

Pℓ (j0 )

i=1 gi xi − 1

≡0 mod p.
qj − 1

Thus, P ∗ (x) ≡ 0 mod p as P ∗ was simply the product over all such binomial
coefficients.
So, if T (x) is zero-sum, we have P ∗ (x) ≡ 1 mod p, else if T (x) is not zero-sum
we find P ∗ (x) ≡ 0 mod p. Next, we remark that χ0 (x) = ℓi=1 (1 − xi ) is the
Q

characteristic function for the 0 vector.


Then, we see P ∗ = χ∗0 + pQ∗1 for some polynomial Q1 ∈ Q [x1 , . . . , xℓ ]. Since P
is integer valued on {0, 1}ℓ we can identify it with a polynomial Q ∈ Z [x1 , . . . , xℓ ] by
proposition 6.3 such that P ∗ (x) = Q∗ (x) for all x ∈ {0, 1}ℓ where Q∗ : {0, 1}ℓ → Z is
the function induced by Q. Moreover, the proposition guarantees Q is the sum of special
monomials.
Now, note that [x1 . . . xℓ ] Q1 (x). Then, expanding products, we see
[x1 . . . xℓ ] (Q (x)) = (−1)ℓ + pc0 . So we see (−1)m + pc0 ̸≡ 0 mod p. Thus this
coefficient is nonzero. So, since the coefficient of x1 . . . xℓ is nonzero, we find
deg (Q) ≥ deg (x1 . . . xℓ ) = ℓ.
On the other hand, since xi ∈ {0, 1}, we note xki = xi for all 1 ≤ i ≤ ℓ and every
k ∈ Z. So, replacing every occurence of xki with xi , in P , we see we obtain a new

polynomial P , with rational coefficients and all terms special monomials which agrees

59
with P . Since the special monomials form a basis, we see this augmented polynomial has

P = Q as polynomials. Then, since
P   P  
Pℓ ℓ (j) ℓ (j)

(j)
−1 ai x i − 1 ... i=1 ai xi − (qj − 1)
i=1 ai xi

i=1
=
qj − 1 (qj − 1)!

has each of the q − 1 factors of the numerator containing terms of degree at most 1, we see
 Pℓ  ∼ P ∼
(j)xi )−1
deg ( i=1qaji−1 ≤ qj − 1. Thus, ℓ ≤ deg (Q) = deg P ≤ di=1 (qi − 1) as P
was simply the product over such binomials. So, ℓ ≤ di=1 (qi − 1) < D∗ (G), a
P

contradiction by the initial assumptions. Thus, their must in fact be a nontrivial zero-sum
subsequence T (x) | S, so the claim is shown.

Remarks

The astute reader may have noticed there are no citations or attributions in this
chapter, this is because this proof has gone largely unpublished and without much fanfair.
Its arguments mirror that of a paper by Christian Elsholtz in 2007, Zero-Sum Problems in
Finite Abelian Groups and Affine Caps [5], concerning a similar constant. These methods
eventually made their way to David Grynkiewicz, Grynkiewicz and Elsholtz working
together at TU Graz, who worked out the details but left the result unpublished.
Grynkiewicz, having only reconstructed the arguments, but not writing up all the claims,
then gave this proof to a graduate student at the University of Memphis, John Ebert, who
would flesh out the proof. Ebert once again left the result unpublished, making this thesis
the first time this proof of the upper bound is known to have been published anywhere.
While a different proof of the upper bound making use of group rings has been known for
quite some time, these more elementary methods provide some insight into how one may
approach the general davenport constant problem. Chiefly, the idea to exploit properties
and theorems concerning polynomials, perhaps over the product of boolean domains,
{0, 1}m , corresponding to inclusion of a particular term proves useful in other

60
applications. Problems naturally arise with this method, namely the lack of commutativity
and the p-group structure inherent in our groups, but a general proof may aim to alleviate
these problems rather than forge brand new methods.

61
Chapter 7

The Chevalley-Warning Theorem and Reiher’s Proof of the Kemnitz Conjecture

The Chevalley-Warning Theorem

Our first major result of this section concerns the theorem of Chevalley and
Warning which gives the conditions under which a certain nontrivial solution to a
polynomial in a finite field of characteristic p can exist:

Theorem 7.1 (Chevalley-Warning Theorem [7]). Let F be a finite field of character-


istic p and let f1 , f2 , . . . , fk ∈ F [x1 , x2 , . . . , xn ] be polynomials and N the number
of points x ∈ F n such that f1 (x) = . . . = fk (x) = 0. If ki=1 deg (fi ) < n, then
P

N ≡ 0(mod p).

In order to provide a proof of this statement, let us first state and prove the following
lemma:

Lemma 7.1 (Lemma). Let F be a finite field and k1 , k2 , . . . , kn ≥ 0 such that

min ki ≤ |F | − 2.
1≤i≤n

xk11 xk22 . . . xknn = 0. (Note: if a 00 occurs in the expressions it will


P
Then, x1 ,x2 ,...,xn ∈F

be treated as a 1).

Proof. Assume without loss of generality that k1 < |F | − 1. Then, by factoring out a xk11
from each term of the sum and grouping all such x1 ’s, we have
k1 k2 k1 P k2
P kn
P  kn
x1 ,x2 ,...,xn ∈F x1 x2 . . . xn = x1 ∈F x1 x2 ,...,xn ∈F x2 . . . xn . Hence, we must only

show that x1 ∈F xk11 = 0. Suppose k1 = 0. Then x1 ∈F xk11 = |F | and, since p divides


P P

|F |, we see the case k1 = 0 is trivially true. Now, let ω ∈ F × be a generator of F × . Then,

62
we have that

X X X X X
xk11 = xk11 = (ωx1 )k1 = ω k1 xk11 = ω k1 xk11 .
x1 ∈F x1 ∈F × x1 ∈F × x1 ∈F × x1 ∈F

Taking the difference of the first and last terms of the above equality yields
k1
 P 
ω k1 − 1 x
x1 ∈F 1 = 0, so either the sum is 0 or ω k1 − 1 = 0. As ω is a generator of
the cyclic group F × , we have ω k1 = 1 only if k1 ≡ −1 (mod |F |). But, as
0 < k1 < |F | − 1 this case cannot occur, hence we see x1 ∈F xk11 = 0, so the lemma is
P

proven.

Proof of Chevalley-Warning Theorem. First, recall x|F |−1 = x. Then, define

k 
XY 
M= 1 − fi (x)|F |−1 .
x∈F n i=1

We see, by the earlier proposition, that a term of the sum will be 1 if and only if x is a
solution to the system f1 , f2 , . . . , fk , else it will be 0. Furthermore, it is clear by the
construction that M will be exactly equal to the number of solutions to our system
f1 , . . . , fk , and so M = N .
Qk  
Now, let us define the polynomial g (x) = i=1 1 − fi (x)|F |−1 . Then,
|F |
repeatedly applying the substitution xj → xj to g yields a polynomial g = g for all
x ∈ F n . Furthermore, degxj (g) ≤ |F | − 1 for 1 ≤ j ≤ n (this is clear as, if it were not we
would be able to apply the substitution once again). Then, substituting g in place of g
yields
X
N= g (x) .
x ∈F n

Then, applying our lemma, we see that all monomials with degree |F | − 2 or less will
|F |−1
equal 0 and hence the only possible nonzero terms of g are those of the form ni=1 xi
Q
.
Expanding the product, we see that such a monomial would be of degree n (|F | − 1),
|F |−1
however as deg fi = (|F | − 1) deg (fi ), we see that

63
Pk
deg (g) ≤ (|F | − 1) i=1 deg (fi ) < n (|F | − 1) by construction. Consequently, any such
monomial of g (and hence g) will have a zero coefficient, and thus N ≡ 0 (mod p).

Reiher’s Proof of the Kemnitz Conjecture

The Chevalley-Warning theorem has found many uses in combinatorics and


number theory. One of its most potent results is Christian Reiher’s proof of the Kemnitz
conjecture [11], one which was produced when Reiher was just and undergraduate. This
proof makes heavy use of the Chevalley-Warning Thoerem, as we will see.
Erdos, Ginzburg, and Ziv proved that that s (Cn ) = 2n − 1. That is, given any
sequence of integers of size greater or equal to 2n − 1, we always find a subsequence
whose sum is congruent to n. As it turns out s (Cn2 ) is a much more challenging result.
As we have already shown in theorem 5.3, a simple counterxample exist to prove
4n − 3 is a lower bound, so we need only show it is an upper bound. Moreover, corollary
5.1 proves that if the claim holds for n prime, it holds for all n. So, we proceed by
showing the claim is true for an arbitrary prime p.
Before we can prove this, however, we must prove several congruencies using
theorem 7.1. These lemmas will all require taking three polynomials associated with a
given sequence. The first polynomial will ensure sequences of length congruent to p must
be taken. The second polynomial will ensure the first coordinates sum to 0, and the third
polynomial will ensure the second coordinates sum to 0.
For the remainder of this chapter, G = Cp2 for a prime p.

Lemma 7.2 ([11]). Let S be a sequence in G, with |S| = 3p − 3, then

1 − Np−1 − Np + N2p−1 + N2p ≡ 0 mod p.

Proof. First, let Fp be the finite field of characteristic p. Then, denote the terms of S to be
(ai , bi ) for 1 ≤ i ≤ 3p − 3. Then, define the following three polynomials

64
f1 , f2 , f3 ∈ F [x1 , . . . , x3p−2 ] by

|S|
X
f1 = xp−1
n + xp−1
3p−2
i=1
|S|
X
f2 = an xnp−1
i=1
|S|
X
f3 = bn xp−1
n .
i=1

Suppose x = (x1 , . . . , x3p−2 ) is a common zero of f1 , f2 , f3 . Then note that xp−1


i = 1 if
xi ̸= 0, so there are two cases. First, if x3p−2 = 0, we see f1 (x) = 0 implies
P|S| p−1
i=1 xi ≡ 0 mod p, so precisely 0, p, or 2p coordinates xi ̸= 0. x = 0 is an obvious
common zero. Next, let I ⊆ [1, 3p − 2] be the set of indices for the nonzero coordinates,
so xi ̸= 0 if and only if i ∈ I. Then, we see that f1 (x) = i∈I xp−1
P
i = |I| = 0 ≡ 0
mod p, f2 (x) i∈I ai xp−1
P P
i = i∈I ai ≡ 0 mod p, and
f3 (x) = i∈I bi xp−1
P P
i = i∈I bi ≡ 0 mod p. So, we need only count over all admissible
solutions to f2 , f3 of size 0, p, and 2p in order to count all possible x which are common
zeroes. The common solutions to f2 , f3 are simply sequences S (I) | S with
σ (S (I)) = 0, so there are N0 , Np , N2p of them respectively. Moreover, there are p − 1
possible nonzero xi ∈ Cp which can be permuted across the 0, p, 2p positions which are
nonzero in a common solution. Putting all this together, we obtain
(p − 1)0 N0 + (p − 1)p Np + (p − 1)2p N2p ≡ 1 − Np + N2p mod p solutions if x3p−2 = 0.
Similarly, if x3p−2 ̸= 0, we see there must be p − 1 or 2p − 1 coordinates nonzero
to satisfy f1 . Moreover, letting I be the set of indices which are nonzero (excluding
P P
3p − 2), f2 (x) = i∈I ai ≡ 0 mod p and f3 (x) = i∈I bi ≡ 0 mod p represent
zero-sum subsequences of length p − 1 and 2p − 1 respectively, of which there are Np−1
and N2p−1 respectively. So permuting over all nonzero xi ∈ Cp and all placments of these
elements in the, p or 2p nonzero indices (now including index 3p − 2) yields
(p − 1)p Np−1 + (p − 1)2p N2p−1 ≡ −Np−1 + N2p−1 mod p possibilities.

65
Totaling the cumulative number of solutions, we find

1 − Np−1 − Np + N2p−1 + N2p

solutions to f1 = f2 = f3 = 0. Since we have 3 polynomials over 3p − 2 ≥ 4 variables,


with deg (f1 ) + deg (f2 ) + deg (f3 ) = 3 (p − 1) < 3p − 2 we find the number of solutions
congruent to 0 mod p by applying theorem 7.1.

The following lemma all follow from essentially the same argument using the given set of
polynomials:

Lemma 7.3 ([11]).

If |S| = 3p − 2 or |S| = 3p − 1, then

1 − Np + N2p ≡ 0 mod p. (7.1)

Moreover, Np ≡ 0 implies N2p ≡ −1.

Proof. The second assertion follows directly from the first, so we need only show
equation (7.1). Again, we define the following polynomials

|S|
X
f1 = xp−1
n
i=1
|S|
X
f2 = an xp−1
n
i=1
|S|
X
f3 = bn xnp−1 .
i=1

Since we have no excess variable like in the last proof, we apply the same arguments as
the first case of the previous lemma. Let x = (x1 , . . . , x3p−2 ) be a common zero of

66
f1 , f2 , f3 . f1 (x) = 0 requires the number of xi ̸= 0 to be congruent to 0 mod p, so a
possible common zero must have 0, p, 2p nonzero coordinates. Next, f2 (x) = f3 (x) = 0
requires the subsequence T (x) induced by the nonzero xi to be zero-sum. All together, for
the case of 0 nonzero coordinates we find one possibility, for the case of p nonzero
coordinates we find (p − 1)p Np possibilities and for the case of 2p nonzero coordinates
we find (p − 1)2p N2p possibilities. Summing, we find 1 − Np + N2p possible solutions.
Since deg (f1 ) + deg (f2 ) + deg (f3 ) = 3p − 3 < 3p − 2 we can apply theorem 7.1 to see
1 − Np + N2p ≡ 0 mod p. Applying the same methods as lemma 7.2 yields

1 + (p − 1)p Np + (p − 1)2p N2p

solutions. Evaluating the coefficients and applying Chevalley-Warning yields the


result.

Lemma 7.4 ([11]). If |S| = 4p − 3, then

−1 + Np − N2p + N3p ≡ 0 mod p and, (7.2)

Np−1 − N2p−1 + N3p−1 ≡ 0 (7.3)

Proof. This is proved analogously to the previous lemmas. First, we show (7.2) by taking
the following polynomials

|S|
X
f1 = xp−1
n
i=1
|S|
X
f2 = an xp−1
n
i=1
|S|
X
f3 = bn xnp−1 .
i=1

67
Applying the same argument as the previous lemmas yields

1 + (p − 1)p Np + (p − 1)2p N2p + (p − 1)p N3p

solutions. Evaluating the coefficients and applying Chevalley-Warning yields

1 − Np + N2p − N3p ≡ 0 mod p.

Finally, negating this yields result (7.2).


Similarly, for (7.3) we use the polynomials

|S|
X
f1 = xp−1
n + xp−1
4p−2
i=1
|S|
X
f2 = an xnp−1
i=1
|S|
X
f3 = bn xp−1
n .
i=1

Counting solutions yields

1 + (p − 1)p Np + (p − 1)2p N2p + (p − 1)3p N3p

+ (p − 1)p Np−1 + (p − 1)2p N2p−1 + (p − 1)3p N3p−1

solutions. Evaluating coefficients and applying Chevalley-Warning theorem yields

1 − Np + N2p − N3p − Np−1 + N2p−1 − N3p−1 ≡ 0.

As we already know the first 4 terms to be congruent to 0 negating what is left yields
equation (7.3)

These lemmas make full use of the Chevalley-Warning theorem and show its power in

68
proving these sorts of combinatorial results. Before we can present the proof of the main
theorem (the Kemnitz conjecture) we need to prove a few more lemmas.

Lemma 7.5 ([11]). If |S| = 4p − 3, then

3 − 2Np−1 − 2Np + N2p−1 + N2p ≡ 0 mod p. (7.4)

Proof. Denoting I to be the set of all subsequences I | S with |I| = 3p − 3, we can


apply 7.2 to each I ∈ I to see

X
[1 − Np−1 (I) − Np (I) + N2p−1 (I) + N2p (I)] ≡ 0. (7.5)
I∈I

4p−3

We find there are 3p−3
possible sets I. Then, for a given zero sum subsequence of length
p − 1 in X, we see we can form a set I ⊆ I by simply appending 2p − 2 elements to the
3p−2

subsequence. We find there are 2p−2 possibilities for these appendices to form I.
3p−3

Similarly, given a zero sum subsequence of length p we find 2p−3 possible sets I
containing the sequence. For a sequence of length 2p − 1 we find 2p−2

p−2
possibilities and
for a sequence of length 2p, we find 2p−3

p−3
possibilities.
Permuting over all such zero-sum sequences in J yields

     
4p − 3 3p − 2 3p − 3
Equation (7.5) = − Np−1 − Np
3p − 3 2p − 2 2p − 3
   
2p − 2 2p − 3
+ N2p−1 + N2p ≡ 0 mod p.
p−2 p−3

Finally, expanding and reducing the binomial coefficients modulo p much as in


proposition 6.4 yields equation (7.4).

Before we may prove the main result we need just a few more lemmas:

69
Lemma 7.6 ([2]). Let S be a sequence in G and suppose there is a zero-sum subse-
quence T | S with |T | = 3p. Then, Np (S) > 0

Proof. Let t | T , since |T t−1 | = 3p − 1, we may apply lemma 7.3 to see Np (T t−1 ) ≡ 0
mod p implies N2p (T t−1 ) ≡ −1 mod p. Assume Np (T t−1 ) = 0, then since
N2p (T t−1 ) ̸= 0, we find a zero-sum subsequence T1 | T t−1 with |T1 | = 2p. But then,
σ (T ) = σ (T1 ) + σ T T1−1 , so σ T T1−1 = 0. Since T T1−1 = p this is a contradiction,
 

so Np (T t−1 ) ̸= 0 and since any zero-sum subsequence in T t−1 is also a subsequence in S,


we see Np (S) > 0.

We prove one final lemma:

Lemma 7.7 ([11]). If |S| = 4p − 3 and Np = 0, then Np−1 ≡ N3p−1 mod p. Conse-
quently,
Np−1 − N3p−1 ≡ 0 mod p. (7.6)

Proof. Let N denote the number of partitions of S into 3 disjoint subsequences, A, B, C


with
|A| = p − 1, |B| = p − 2, and |C| = 2p (7.7)

so that

σ (A) = σ (C) ≡ 0 mod p and consequently σ (B) ≡ σ (X) mod p. (7.8)

First, we note that


X
N2p SA−1

N≡
A
P
where A denotes the sum over all sets A ⊆ X so that (7.7) and (7.8) hold. Note that this
simply counts the number of admissible C for each possible set A, leaving whatever
points remain to belong to B, thereby counting the number of possible partitions (modulo

70
p). Next, since |SA−1 | = 3p − 2, we may apply 7.3 to see N ≡
P
A −1 ≡ −Np−1 (S)
mod p.
N2p (SB −1 ), by counting
P
Applying the same method, but over B, we see N ≡ B

all possible sets C for each admissible B. As |SB −1 | = 3p − 1, we may once again apply
P
7.3 to find N ≡ X−B −1 ≡ −N3p−1 mod p.
By method of double counting we attain Np−1 ≡ N3p−1 , the desired result.

With many small parts proven, we need simply combine them to yield the Kemnitz
conjecture.

Theorem 7.2.
s Cn2 = 4n − 3.


Proof of the Kemnitz Conjecture [11]. Adding the equations we have obtained thus far,
we see for |S| = 4p − 3 we have

(7.2) + (7.3) + (7.4) + (7.6) yields

−1 + Np − N2p + N3p + Np−1 − N2p−1 + N3p−1 + 3 − 2Np−1

− 2Np + N2p−1 + N2p + Np−1 − N3p−1

= 2 − Np + N3p ≡ 0 mod p

If p is an odd prime, we see the statement Np ≡ N3p ≡ 0 mod p cannot be true (as
2 ̸≡ 0), Hence, either Np ̸≡ 0 in which case the claim is shown, or N3p ̸≡ 0 implying S
contains a subsequence of size 3p whose sum is zero. Applying lemma 7.6 then yields that
Np > 0, so the claim is shown in this case as well. Else, if p = 2 note that if a sequence
contains any repeated element, it has a zero-sum subsequence of length 2 (taking both
copies of the repeated element). Since |C22 | = 4, any sequence of 5 = 4 · 2 − 3 elements
will contain a repeated element, and thus a zero-sum subsequence of length 2, so this case

71
holds as well. Since the claim holds for all primes p, we see corollary 5.1 implies
Cn2 = 4n − 3, so the Kemnitz conjecture it true.

72
Chapter 8

Conclusion

Ultimately, the polynomial method proves quite useful. It’s many variations are
able to solve problems from number theory, combinatorics, zero-sum theory, graph theory,
and many other domains. While we only touched on some of the basic tools and methods
in this Thesis, many of these ideas can be expanded further to prove even stronger results.
For example, there are generalized Chevalley-Warning theorems. Other ideas prove harder
to generalize. For instance, determining the value of s (Cn3 ) is still an open problem, as is
D (G) for an arbitrary abelian group G. In these cases existing polynomial methods fall
short. In the end, though, the breadth and complexity of the problems that can be solved
with polynomial methods is quite remarkable and deserving of further study. We hope the
general reader has picked up at least some ideas concerning polynomials and their
usefulness which they can introduce into their mathematical toolkit.

73
REFERENCES

[1] Noga Alon. Combinatorial nullstellensatz. Combinatorics, Probability and


Computing, 8(1-2):7–29, 1999.

[2] Noga Alon and Moshe Dubiner. A lattice point problem and additive number theory.
Combinatorica, 15:301–309, 1995.

[3] Noga Alon, Melvyn B. Nathanson, and Imre Ruzsa. The polynomial method and
restricted sums of congruence classes. Journal of Number Theory, 56(2):404–417,
February 1996.

[4] George Andrews. Proc. advanced sem., math. res. center, univ. wisconsin, madison,
wis., 1975. In Richard Askey, editor, Problems and Prospects for Basic
Hypergeometric Functions, pages 191–224. Academic Press, Math Res. Center,
Univ. Wisconsin, 1975.

[5] Yves Edel, Christian Elsholtz, Alfred Geroldinger, Silke Kubertin, and Laurence
Rackham. Zero-Sum Problems in Finite Abelian Groups and Affine Caps. The
Quarterly Journal of Mathematics, 58(2):159–186, 05 2007.

[6] I. J. Good. Short proof of a conjecture by dyson. Journal of Mathematical Physics,


11(6):1884–1884, 1970.

[7] David J. Grynkiewicz. Structural additive theory, chapter The Polynomial Method:
The Erdos:Heilbronn Conjecture. Springer, 2013.

[8] Gyula Károlyi and Zoltán Nagy. A simple proof of the zeilberger–bressoud q-dyson
theorem. Proceedings of the American Mathematical Society, 142(9):3007–3011,
Sep 2014.

[9] Madan Lal Mehta and Freeman J. Dyson. Statistical theory of the energy levels of
complex systems. v. Journal of Mathematical Physics, 4(5):713–719, 1963.

74
[10] D. Nelson. The Penguin Dictionary of Mathematics. Penguin reference library.
Penguin Books Limited, 4 edition, 2008.

[11] Christian Reiher. On kemnitz’ conjecture concerning lattice-points in the plane. The
Ramanujan Journal, 13, Jun 2007.

[12] Thomas Sauer and Yuan Xu. On multivariate lagrange interpolation. Mathematics of
Computation, 64(211):1147–1170, 1995.

[13] B.L. Van der Waerden. Modern Algebra. Julius Springer, 1931.

[14] Kenneth G. Wilson. Proof of a conjecture by dyson. Journal of Mathematical


Physics, 3(5):1040–1043, 1962.

75
ProQuest Number: 29064664

INFORMATION TO ALL USERS


The quality and completeness of this reproduction is dependent on the quality
and completeness of the copy made available to ProQuest.

Distributed by ProQuest LLC ( 2022 ).


Copyright of the Dissertation is held by the Author unless otherwise noted.

This work may be used in accordance with the terms of the Creative Commons license
or other rights statement, as indicated in the copyright statement or in the metadata
associated with this work. Unless otherwise specified in the copyright statement
or the metadata, all rights are reserved by the copyright holder.

This work is protected against unauthorized copying under Title 17,


United States Code and other applicable copyright laws.

Microform Edition where available © ProQuest LLC. No reproduction or digitization


of the Microform Edition is authorized without permission of ProQuest LLC.

ProQuest LLC
789 East Eisenhower Parkway
P.O. Box 1346
Ann Arbor, MI 48106 - 1346 USA

You might also like