0 Notes2 MemoryCPU
0 Notes2 MemoryCPU
0 Notes2 MemoryCPU
Number Theory
Wissam Raji
2
Preface
These notes serve as course notes for an undergraduate course in number the-
ory. Most if not all universities worldwide offer introductory courses in number
theory for math majors and in many cases as an elective course.
The notes contain a useful introduction to important topics that need to be ad-
dressed in a course in number theory. Proofs of basic theorems are presented in
an interesting and comprehensive way that can be read and understood even by
non-majors with the exception in the last three chapters where a background in
analysis, measure theory and abstract algebra is required. The exercises are care-
fully chosen to broaden the understanding of the concepts. Moreover, these notes
shed light on analytic number theory, a subject that is rarely seen or approached
by undergraduate students. One of the unique characteristics of these notes is the
careful choice of topics and its importance in the theory of numbers. The
freedom is given in the last two chapters because of the advanced nature of the
topics that are presented.
Thanks to professor Pavel Guerzhoy from University of Hawaii for his
contri- bution in chapter 6 on continued fraction and to Professor Ramez
Maalouf from Notre Dame University, Lebanon for his contribution to chapter 8.
Contents
1 Introduction 7
1.1 Algebraic Operations With Integers . . . . . . . . . . . . . . . . 8
1.2 The Well Ordering Principle and Mathematical Induction . . . . . 9
1.2.1 The Well Ordering Principle . . . . . . . . . . . . . . . 10
1.2.2 The Pigeonhole Principle . . . . . . . . . . . . . . . . . 10
1.2.3 The Principle of Mathematical Induction . . . . . . . . 10
1.3 Divisibility and the Division Algorithm . . . . . . . . . . . . . . 13
1.3.1 Integer Divisibility . . . . . . . . . . . . . . . . . . . . . 13
1.3.2 The Division Algorithm . . . . . . . . . . . . . . . . . . 15
1.4 Representations of Integers in Different Bases . . . . . . . . . . . 16
1.5 The Greatest Common Divisor . . . . . . . . . . . . . . . . . . . 20
1.6 The Euclidean Algorithm . . . . . . . . . . . . . . . . . . . . . . 24
1.7 Lame’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2 Prime Numbers 31
2.1 The Sieve of Eratosthenes . . . . . . . . . . . . . . . . . . . . . . 31
2.2 The infinitude of Primes . . . . . . . . . . . . . . . . . . . . . . 34
2.3 The Fundamental Theorem of Arithmetic . . . . . . . . . . . . . 35
2.3.1 The Fundamental Theorem of Arithmetic . . . . . . . . . 36
2.3.2 More on the Infinitude of Primes . . . . . . . . . . . . . . 39
2.4 Least Common Multiple . . . . . . . . . . . . . . . . . . . . . . 41
3
4 CONTENTS
3 Congruences 51
3.1 Introduction to congruences . . . . . . . . . . . . . . . . . . . . 51
3.2 Residue Systems and Euler’s φ-Function . . . . . . . . . . . . . . 57
3.2.1 Residue Systems . . . . . . . . . . . . . . . . . . . . . . 57
3.2.2 Euler’s φ-Function . . . . . . . . . . . . . . . . . . . . . 59
3.3 Linear Congruences . . . . . . . . . . . . . . . . . . . . . . . . . 59
3.4 The Chinese Remainder Theorem . . . . . . . . . . . . . . . . . 62
3.5 Theorems of Fermat, Euler, and Wilson . . . . . . . . . . . . . . 64
Introduction
Integers are the building blocks of the theory of numbers. This chapter contains
somewhat very simple and obvious observations starting with properties of inte-
gers and yet the proofs behind those observations are not as simple. In this
chapter we introduce basic operations on integers and some algebraic definitions
that will be necessary to understand basic concepts in this book. We then
introduce the Well ordering principle which states basically that every set of
positive integers has a smallest element. Proof by induction is also presented as
an efficient method for proving several theorems throughout the book. We
proceed to define the con- cept of divisibility and the division algorithm. We
then introduce the elementary but fundamental concept of a greatest common
divisor (gcd) of two integers, and the Euclidean algorithm for finding the gcd of
two integers. We end this chap- ter with Lame’s Lemma on an estimate of the
number of steps in the Euclidean algorithm needed to find the gcd of two
integers.
7
8 CHAPTER 1. INTRODUCTION
a+ b= b+ a
a· b= b· a
(a + b) + c = a + (b +
c) (a · b) · c = a · (b · c)
a · (b + c) = a · b + a ·
c.
1.2. THE WELL ORDERING PRINCIPLE AND MATHEMATICAL INDUCTION9
In the set Z there are ”identity elements” for the two operations + and ·, and
these are the elements 0 and 1 respectively, that satisfy the basic properties
a+ 0= 0+ a =
aa· 1= 1· a =
a
for every a ∈ Z.
The set Z allows additive inverses for its elements, in the sense that for every
a ∈ Z there exists another integer in Z, denoted by −a, such that
a + (−a) = 0. (1.3)
While for multiplication, only the integer 1 has a multiplicative inverse in the
sense that 1 is the only integer a such that there exists another integer, denoted
by
a−1 or by 1/a, (namely 1 itself in this case) such that
a · a−1 = 1. (1.4)
From the operations of addition and multiplication one can define two other
operations on Z, namely subtraction (denoted by −) and division (denoted by
/). Subtraction is a binary operation on Z, i.e. defined for any two integers in Z,
while division is not a binary operation and thus is defined only for some
specific couple of integers in Z. Subtraction and division are defined as follows:
1. a − b is defined by a + (−b), i.e. a − b = a + (−b) for every a, b ∈ Z
2. a/b is defined by the integer c if and only if a = b · c.
the well ordering principle. We then state what is known as the pigeonhole prin-
ciple, and then we proceed to present an important method called mathematical
induction.
This principle can be taken as an axiom on integers and it will be the key to
proving many theorems. As a result, we see that any set of positive integers is
well ordered while the set of all integers is not well ordered.
Proof. Suppose that none of the boxes contains more than one object. Then there
are at most k objects. This leads to a contradiction with the fact that there are s
objects for s > k.
k + 1, and if this set contains 1 then it must be the set of all positive integers.
More generally, a property concerning the positive integers that is true for n = 1,
and that is true for the integer n + 1 whenever it is true for the integer n, must
be true for all positive integers.
Proof. Let S be the set of positive integers containing the integer 1, and the
integer k + 1 whenever it contains k. Assume also that S is not the set of all
positive integers. As a result, there are some integers that are not contained in S
and thus those integers must have a least element α by the well ordering
principle. Notice that α = 1 since 1 ∈ S. But α − 1 ∈ S and thus using the
property of S, α ∈ S. Thus S must contain all positive integers.
n! ≤ nn
Proof. Let T be a set of integers containing 1 and such that for every positive
integer k, if it contains 1, 2, ..., k, then it contains k + 1. Let S be the set of all
positive integers k such that all the positive integers less than or equal to k are in
T . Then 1 is in S, and we also see that k + 1 is in S. Thus S must be the set of
all positive integers. Thus T must be the set of all positive integers since S is a
subset of T .
1.3. DIVISIBILITY AND THE DIVISION ALGORITHM 13
Exercises
1. Prove using mathematical induction that n < 3n for all positive integers n.
Pn
2. Show that 2 n(n+1)(2n+1) .
j=1 j = 6
Pn j−1 2
3. Use mathematical induction to prove that j=1 (−1) j = (−1)n−1 n(n+
1)/2.
Pn
4. Use mathematical induction to prove that j=1 j 3 = [n(n+1)/2]2 for every
positive integer n.
Pn
5. Use mathematical induction to prove that j=1 (2j − 1) = n2
Example 3. a) Note that any even integer has the form 2k for some integer k,
while any odd integer has the form 2k + 1 for some integer k. Thus 2|n if n is
even, while 2 - n if n is odd.
14 CHAPTER 1. INTRODUCTION
Proof. Since a | b and b | c, then there exist integers k1 and k2 such that b = k1 a
and c = k2 b. As a result, we have c = k1 k2 a and hence a | c.
The following theorem states that if an integer divides two other integers
then it divides any linear combination of these integers.
Proof. Since c | a and c | b, then by definition there exists k1 and k2 such that
a = k1 c and b = k2 c. Thus
a | b1 , a | b2 , ..., a | bn
then
X
n
a| kj bj (1.6)
j=1
Theorem 5. The Division Algorithm If a and b are integers such that b > 0, then
there exist unique integers q and r such that a = bq + r where 0 ≤ r < b.
r > r − b = a − bq − b = a − b(q + 1) =≥
0.
b(q1 − q2 ) + (r1 − r2 ) =
0.
As a result we have
b(q1 − q2 ) = r2 − r1 .
Exercises
1. Show that 5 | 25, 19 | 38 and 2 | 98.
16 CHAPTER 1. INTRODUCTION
2. Use the division algorithm to find the quotient and the remainder when 76
is divided by 13.
3. Use the division algorithm to find the quotient and the remainder when
-100 is divided by 13.
4. Show that if a, b, c and d are integers with a and c nonzero, such that a | b
and c | d, then ac | bd.
6. Prove that the sum of two even integers is even, the sum of two odd
integers is even and the sum of an even integer and an odd integer is odd.
7. Show that the product of two even integers is even, the product of two odd
integers is odd and the product of an even integer and an odd integer is
even.
10. Show that the square of any integer is of the form 3m or 3m + 1 but not of
the form 3m + 2.
decimal notation in daily life is simply better because we have ten fingers which
facilitates all the mathematical operations.
Notation An integer a written in base b expansion is denoted by (a)b .
Theorem 6. Let b be a positive integer with b > 1. Then any positive integer m
can be written uniquely as
m = bq0 + a0 , 0 ≤ a0 < b.
q1 = bq2 + a2 , 0 ≤ a2 < b,
.
.
.
ql−2 = bql−1 + al−1 , 0 ≤ al−1 < b,
ql−1 = b · 0 + al , 0 ≤ al < b.
m = b(bq1 + a1 ) + a0 = b2 q1 + a1 b + a0 ,
18 CHAPTER 1. INTRODUCTION
m = b3 q2 + a2 b2 + a1 b + a0 ,
.
.
.
= bl ql− 1 + al−1 bl−1 + ... + a1 b + a0 ,
= al bl + al−1 bl−1 + ... + a1 b + a0 .
What remains to prove is that the representation is unique. Suppose now that
where if the number of terms is different in one expansion, we add zero coeffi-
cients to make the number of terms agree. Subtracting the two expansions, we
get
If the two expansions are different, then there exists 0 ≤ j ≤ l such that cj = aj .
As a result, we get
We now get
aj − cj = (al − cl )bl−j + ... + (aj+1 − cj+1 )b,
we do the following
214 = 3 · 71 +
71 = 13 · 23 +
2
23 = 3 · 7 + 2
7 = 3· 2+ 1
2 = 3· 0+ 2
Example 7. To find the base 10 expansion, i.e. the decimal expansion, of (364)7 :
We do the following: 4 · 70 + 6 · 71 + 3 · 72 = 4 + 42 + 147 = 193.
In some cases where base b > 10 expansion is needed, we add some
characters to represent numbers greater than 9. It is known to use the
alphabetic letters to denote integers greater than 9 in base b expansion for b
> 10. For example (46BC 29)13 where A = 10, B = 11, C = 12.
To convert from one base to the other, the simplest way is to go through base
10 and then convert to the other base. There are methods that simplify
conversion from one base to the other but it will not be addressed in this book.
Exercises
20 CHAPTER 1. INTRODUCTION
Definition 2. The greatest common divisor of two integers a and b is the greatest
integer that divides both a and b.
We denote the greatest common divisor of two integers a and b by (a, b). We
also define (0, 0) = 0.
There are couples of integers (e.g. 3 and 4, etc...) whose greatest common
divisor is 1 so we call such integers relatively prime integers.
Note that every integer has positive and negative divisors. If a is a positive
divisor of m, then −a is also a divisor of m. Therefore by our definition of the
greatest common divisor, we can see that (a, b) = (| a |, | b |).
We now present a theorem about the greatest common divisor of two
integers. The theorem states that if we divide two integers by their greatest
common divisor, then the outcome is a couple of integers that are relatively
prime.
Proof. We will show that a/d and b/d have no common positive divisors other
than 1. Assume that k is a positive common divisor such that k | a/d and k | b/d.
As a result, there are two positive integers m and n such that
The next theorem shows that the greatest common divisor of two integers
does not change when we add a multiple of one of the two integers to the other.
We now present a theorem which proves that the greatest common divisor of
two integers can be written as a linear combination of the two integers.
Theorem 9. The greatest common divisor of two integers a and b, not both 0 is
the least positive integer such that ma + nb = d for some integers m and n.
Proof. Assume without loss of generality that a and b are positive integers. Con-
sider the set of all positive integer linear combinations of a and b. This set is non
empty since a = 1 · a + 0 · b and b = 0 · a + 1 · b are both in this set. Thus this
set has a least element d by the well-ordering principle. Thus d = ma + nb for
some integers m and n. We have to prove that d divides both a and b and that it
is the greatest divisor of a and b.
By the division algorithm, we have
a = dq + r, 0 ≤ r < d.
Thus we have
r = a − dq = a − q(ma + nb) = (1 − qm)a −
qnb.
Example 11. The integers 3, 6, 7 are mutually relatively prime since (3, 6, 7) =
1
although (3, 6) = 3.
Definition 6. The integers a1 , a2 , ..., an are called pairwise prime if for each i =
j, we have (ai , aj ) = 1.
Example 12. The integers 3, 14, 25 are pairwise relatively prime. Notice also
that these integers are mutually relatively prime.
Notice that if a1 , a2 , ..., an are pairwise relatively prime then they are
mutually relatively prime.
Exercises
6. Show that if m and n are integers such that (m, n) = 1, then (m+n,m-n)=1
or 2.
9. Show that if a1 , a2 , ..., an are integers that are not all 0 and c is a positive
integer, then (ca1 , ca2 , ..., can ) = c(a1 , a2 , ...an ).
Lemma 1. If a and b are two integers and a = bq + r where also q and r are
integers, then (a, b) = (r, b).
The above lemma will lead to a more general version of it. We now present
the Euclidean algorithm in its general form. It states that the greatest common
divisor of two integers is the last non zero remainder of the successive division.
rn+1 = 0.
Then (a, b) = rn .
r0 = r1 q1 + r2 0 ≤ r2 < r 1 ,
r1 = r2 q2 + r3 0 ≤ r3 < r 2 ,
.
.
.
rn−2 = rn−1 qn−1 + rn 0 ≤ rn < rn−1 ,
rn−1 = rn qn .
Notice that, we will have a remainder of 0 eventually since all the remainders
are integers and every remainder in the next step is less than the remainder in the
previous one. By Lemma 1, we see that
Example 13. We will find the greatest common divisor of 4147 and 10672:
26 CHAPTER 1. INTRODUCTION
Note that
10672 = 4147 · 2 +
4147 = 2378,
2378 · 1 +
1769,
2378 = 1769 · 1 + 609,
1769 = 609 · 2 + 551,
609 = 551 · 1 + 58,
551 = 58 · 9 + 29,
58 = 29 · 2,
We now use the steps in the Euclidean algorithm to write the greatest
common divisor of two integers as a linear combination of the two integers. The
following example will actually determine the variables m and n described in
Theorem 9. The following algorithm can be described by a general form but for
the sake of simplicity of expressions we will present an example that shows
the steps for obtaining the greatest common divisor of two integers as a linear
combination of the two integers.
29 = 551 − 9 · 58,
= 551 − 9(609 − 551 · 1),
= 10.551 − 9.609,
= 10 · (1769 − 609 · 2) − 9 · 609,
= 10 · 1769 − 29 · 609,
= 10 · 1769 − 29(2378 − 1769 · 1),
= 39 · 1769 − 29 · 2378,
= 39(4147 − 2378 · 1) − 29 · 2378,
= 39 · 4147 − 68 · 2378,
= 39 · 4147 − 68(10672 − 4147 · 2),
= 175 · 4147 − 68 · 10672,
1. Use the Euclidean algorithm to find the greatest common divisor of 412
and
32 and express it in terms of the two integers.
2. Use the Euclidean algorithm to find the greatest common divisor of 780
and
150 and express it in terms of the two integers.
4. Let a and b be two positive even integers. Prove that (a, b) = 2(a/2, b/2).
5. Show that if a and b are positive integers where a is even and b is odd, then
(a, b) = (a/2, b).
28 CHAPTER 1. INTRODUCTION
Theorem 11. using the Euclidean algorithm to find the greatest common divisor
of two positive integers has number of divisions less than or equal five times the
number of decimal digits in the minimum of the two integers.
Proof. Let a and b be two positive integers where a > b. Applying the Euclidean
algorithm to find the greatest common divisor of two integers with a = r0 and
b = r1 , we get
r0 = r1 q1 + r2 0 ≤ r2 < r 1 ,
r1 = r2 q2 + r3 0 ≤ r3 < r 2 ,
.
.
.
rn−2 = rn−1 qn−1 + rn 0 ≤ rn < rn−1 ,
rn−1 = rn qn .
Notice that each of the quotients q1 , q2 , ..., qn−1 are all greater than 1 and qn ≥ 2
and this is because rn < rn−1 . Thus we have
rn ≥ 1 = f2 ,
rn−1 ≥ 2rn ≥ 2f2 = f3 ,
≥ rn−1 + rn ≥ f3 + f2 = f4 ,
rn−2
≥ rn−2 + rn−1 ≥ f4 + f3 =
rn−3 f5 ,
.
.
.
r2 ≥ r3 + r4 ≥ fn−1 + fn−2 = fn ,
b = r1 ≥ r2 + r3 ≥ fn + fn−1 = fn+1
.
30 CHAPTER 1. INTRODUCTION
Thus notice that b ≥ fn+1 . By Lemma 2, we have fn+1 > αn−1 for n > 2. As a
result, we have b > αn−1 . Now notice since
log10 α > 1 ,
5
we see that
log10 b > (n − 1)/5.
Thus we have
n − 1 < 5log10 b.
Now let b has k decimal digits. As a result, we have b < 10k and thus log10 b <
k. Hence we conclude that n − 1 < 5k. Since k is an integer, we conclude
that n ≤ 5k.
Exercises
1. Find an upper bound for the number of steps in the Euclidean algorithm
that is used to find the greatest common divisor of 38472 and 957748838.
2. Find an upper bound for the number of steps in the Euclidean algorithm
that is used to find the greatest common divisor of 15 and 75. Verify your
result by using the Euclidean algorithm to find the greatest common
divisor of the two integers.
Chapter 2
Prime Numbers
Prime numbers, the building blocks of integers, have been studied extensively
over the centuries. Being able to present an integer uniquely as product of primes
is the main reason behind the whole theory of numbers and behind the interesting
results in this theory. Many interesting theorems, applications and conjectures
have been formulated based on the properties of prime numbers.
In this chapter, we present methods to determine whether a number is prime
or composite using an ancient Greek method invented by Eratosthenes. We also
show that there are infinitely many prime numbers. We then proceed to show
that every integer can be written uniquely as a product of primes.
We introduce as well the concept of diophantine equations where integer so-
lutions from given equations are determined using the greatest common divisor.
We then mention the Prime Number theorem without giving a proof of course in
addition to other conjectures and major results related to prime numbers.
31
32 CHAPTER 2. PRIME NUMBERS
Note that any integer greater than 1 that is not prime is said to be a composite
number.
Proof. We present the proof of this Lemma by contradiction. Suppose that there
is an integer greater than one that has no prime divisors. Since the set of integers
with elements greater than one with no prime divisors is nonempty, then by the
well ordering principle there is a least positive integer n greater than one that has
no prime divisors. Thus n is composite since n divides n. Hence
Notice that a < n and as a result since n is minimal, a must have a prime divisor
which will also be a divisor of n.
Theorem 12. If n is a composite integer, then n has a prime factor not exceeding
√
n.
Proof. Since n is composite, then n = ab, where a and b are integers with 1 <
√
a ≤ b < n. Suppose now that a > n, then
√
n<a≤b
and as a result
√ √
ab > n n = n.
2.1. THE SIEVE OF ERATOSTHENES 33
√
Therefore a ≤ n. Also, by Lemma 3, a must have a prime divisor a1 which is
√
also a prime divisor of n and thus this divisor is less than a1 ≤ a ≤ n.
We now present the algorithm of the Sieve of Eratosthenes that is used to de-
termine prime numbers up to a given integer.
1. Write a list of numbers from 2 to the largest number n you want to test.
Note √that every composite integer less than n must have a prime factor less
than n. Hence you need to strike off the multiples of the primes that are
√
less than n
2. Strike off all multiples of 2 greater than 2 from the list . The first
remaining number in the list is a prime number.
4. Repeat the above steps√until no more multiples are found of the prime inte-
gers that are less than n
Exercises
1. Use the Sieve of Eratosthenes to find all primes less than 100.
2. Use the Sieve of Eratosthenes to find all primes less than 200.
Proof. We present the proof by contradiction. Suppose there are finitely many
primes p1 , p2 , ..., pn , where n is a positive integer. Consider the integer Q
such that
Q = p1 p2 ...pn + 1.
By Lemma 3, Q has at least a prime divisor, say q. If we prove that q is not one
of the primes listed then we obtain a contradiction. Suppose now that q = pi for
1 ≤ i ≤ n. Thus q divides p1 p2 ...pn and as a result q divides Q − p1 p2 ...pn .
Therefore q divides 1. But this is impossible since there is no prime that divides
1 and as a result q is not one of the primes listed.
The following theorem discusses the large gaps between primes. It simply
states that there are arbitrary large gaps in the series of primes and that the
primes are spaced irregularly.
Notice that every integer in the above sequence is composite because k divides
(n + 1)! + k if 2 ≤ k ≤ n + 1 by 4.
2.3. THE FUNDAMENTAL THEOREM OF ARITHMETIC 35
Exercises
Lemma 4. If a,b,c are positive integers such that (a, b) = 1 and a | bc, then a | c.
Proof. We present the proof of this result by induction. For k = 1, the result
is trivial. Assume now that the result is true for k. Consider n1 n2 ...nk+1 that is
divisible by p. Notice that either
We now state the fundamental theorem of arithmetic and present the proof
using Lemma 5.
n = p1 p2 ...ps = q1 q2 ...qr
Cancel out all common primes from the factorizations above to get
Thus all the primes on the left side are different from the primes on the right side.
Since any pjl (l = 1, · · · , n) divides pj1 pj2 ...pju , then pjl must divide qi1 qi2
...qiv , and hence by Lemma 5, pj1 must divide qjk for some 1 ≤ k ≤ v which is
impos- sible. Hence the representation is unique.
where all the pi are distinct for 1 ≤ i ≤ j. One can also write a formal product
Y
α
n= pi i , (2.2)
all primes pi
38 CHAPTER 2. PRIME NUMBERS
a a
a = p 11 p2 2 ...pnan and b = pb11 pb22 ...pnbn ,
where we exclude in these expansions any prime p with power 0 in both a and b
(and thus some of the powers above may be 0 in one expansion but not the
other). Of course, if one prime pi appears in a but not in b, then ai = 0 while bi
= 0, and
vise versa. Then the greatest common divisor is given
by
Proof. Let d1 = (a, d) and d2 = (b, d). Since (a, b) = 1 and writing a and b in
terms of their prime decomposition, it is clear that d = d1 d2 and (d1 , d2 ) = 1.
Note that every prime power in the factorization of d must appear in either d1 or
d2 . Also the prime powers in the factorization of d that are prime powers
dividing a must appear in d1 and that prime powers in the factorization of d that
are prime powers dividing b must appear in d2 .
2.3. THE FUNDAMENTAL THEOREM OF ARITHMETIC 39
is a divisor of ab.
This result had been conjectured by Gauss but was first proved by Dirichlet.
Dirichlet proved this theorem using complex analysis, but the proof is so chal-
lenging. As a result, we will present a special case of this theorem and prove that
there are infinitely many primes in a given arithmetic progression. Before stating
the theorem about the special case of Dirichlet’s theorem, we prove a lemma that
will be used in the proof of the mentioned theorem.
Lemma 7. If a and b are integers both of the form 4n + 1, then their product ab
is of the form 4n +
1
where n3 = 4n1 n2 + n1 +
n2 .
Theorem 17. There are infinitely many primes of the form 4n + 3, where n is a
positive integer.
40 CHAPTER 2. PRIME NUMBERS
Proof. Suppose that there are finitely many primes of the form 4n + 3, say p0 =
3, p1 , p2 , ..., pn . Let
N = 4p1 p2 ...pn + 3.
Notice that any odd prime is of the form 4n + 1 or 4n + 3. Then there is at least
one prime in the prime factorization of N of the form 4n + 3, as otherwise, by
Lemma 7, N will be in the form 4n + 1. We wish to prove that this prime in the
factorization of N is none of p0 = 3, p1 , p2 , ..., pn . Notice that if
3 | N,
pi | (N − 4p1 p2 ...pn ) = 3.
Hence none of the primes p0 , p1 , p2 , ..., pn divides N. Thus there are infinitely
many primes of the form 4n + 3.
Exercises
3. Show that all the powers of in the prime factorization of an integer a are
even if and only if a is a perfect square.
Definition 9. The least common multiple (l.c.m.) of two positive integers is the
smallest positive integer that is a multiple of both.
We denote the least common multiple of two positive integers a an b by ha, bi.
Example 18. h2, 8i = 8, h5, 8i = 40
We can figure out ha, bi once we have the prime factorization of a and b.
To do that, let
a a
a = p 1 p 2 ...pan and b = pb1 pb2 ...pbn ,
1 2 m 1 2 m
where (as above) we exclude any prime with 0 power in both a and b. Then
max(a1 ,b1 ) max(a2 ,b2 )
ha, bi = p1 p2 ...pmmax(an ,bn ), where max(a, b) is the maximum of
the two integers a and b. We now prove a theorem that relates the least common
multiple of two positive integers to their greatest common divisor. In some
books, this theorem is adopted as the definition of the least common multiple.
To prove the theorem we present a lemma
min(a, b) + max(a, b) = a + b
and
ha, bi = pmax(a
1 p2 ...pmax(a
1 ,b1 ) max(a2 ,b2 )
m
n ,bn )
,
then
max(a1 ,b1 ) max(a2 ,b2 )
ha, bi(a, b) = p1 p2 ...pmax(a
m
n ,bn ) min(a1 ,b2 ) min(a2 ,b2 )
p1 p2 ...pmin(a
n
n ,bn )
Note also that we used Lemma 8 in the above equations. For part 3, it would be a
nice exercise to show that ab/(a, b) | m (Exercise 6). Thus ha, bi | m.
Exercises
3. Find the least common multiple and the greatest common divisor of 25 56
72 11
and 23 58 72 13.
4. Show that every common multiple of two positive integers a and b is divis-
ible by the least common multiple of a and b.
2.5. LINEAR DIOPHANTINE EQUATIONS 43
5. Show that if a and b are positive integers then the greatest common divisor
of a and b divides their least common multiple. When are the least
common multiple and the greatest common divisor equal to each other.
x = x0 + (b/d)t y = y0 − (a/d)t
where t is an arbitrary
integer.
Now we have to prove that if d | c, then the equation has integral solution.
Assume that d | c. By theorem 9, there exist integers m and n such that
d = am + bn.
c = dk
x0 = km and y0 = kn.
We have to prove now that x and y are solutions for all integers t. Notice that
We now show that every solution for the equation ax + by = c is of the form
x = x0 + (b/d)tand y = y0 − (a/d)t.
a(x − x0 ) + b(y − y0 ) = 0.
Hence
a(x − x0 ) = b(y − y0 ).
2.5. LINEAR DIOPHANTINE EQUATIONS 45
a/d(x − x0 ) = b/d(y − y0 ).
Notice that (a/d, b/d) = 1 and thus we get by Lemma 4 that a/d | y − y0 . As a
result, there exists an integer t such that y = y0 − (a/d)t. Now substituting y −
y0 in the equation
a(x − x0 ) = b(y − y0 ).
We get
x = x0 + (b/d)t.
Example 19. The equation 3x+6y = 7 has no integer solution because (3, 6) =
3
does not divide 7.
Example 20. There are infinitely many integer solutions for the equation 4x +
6y = 8 because (4, 6) = 2 | 8. We use the Euclidean algorithm to determine m
and n where 4m + 6n = 2. It turns out that 4(−1) + 6(1) = 2. And also 8 =
2.4. Thus x0 = 4.(−1) = −4 and y0 = 4.1 = 4 is a particular solution. The
solutions are given by
x = −4 + 3t y = 4 − 2t
for all integers t.
Exercises
1. Either find all solutions or prove that there are no solutions for the
diophan- tine equation 21x + 7y = 147.
2. Either find all solutions or prove that there are no solutions for the
diophan- tine equation 2x + 13y = 31.
3. Either find all solutions or prove that there are no solutions for the
diophan- tine equation 2x + 14y = 17.
46 CHAPTER 2. PRIME NUMBERS
4. A grocer orders apples and bananas at a total cost of $8.4. If the apples cost
25 cents each and the bananas 5 cents each, how many of each type of fruit
did he order.
2.6 The function [x] , the symbols ”O”, ”o” and ”∼”
We start this section by introducing an important number theoretic function. We
proceed in defining some convenient symbols that will be used in connection
with the growth and behavior of some functions that will be defined in later
chapters.
Definition 11. The function [x] represents the largest integer not exceeding x. In
other words, for real x, [x] is the unique integer such that
1. [x + n] = [x] + n, if n is an integer.
Using the definition of [x], it will be easy to see that the above properties are
direct consequences of the definition.
We now define some symbols that will be used to estimate the growth of number
theoretic functions. These symbols will be not be really appreciated in the
context of this book but these are often used in many analytic proofs.
Now, the relation g(x) = o(f (x)), pronounced ”small-oh” of f (x), is used
to indicate that f (x) grows much faster than g(x). It formally says that
g(x)
lim = 0. (2.4)
x→∞ f (x)
g(x)
lim = 0. (2.5)
x→b f (x)
48 CHAPTER 2. PRIME NUMBERS
Example 22. sin(x) = o(x) at ∞, and xk = o(ex ) also at ∞ for every constant
k.
f (x)
lim = 1. (2.6)
x→∞ g(x)
There are some other properties that we did not mention here, properties that are
rarely used in number theoretic proofs.
Exercises
π(x) ∼ x/logx
50 CHAPTER 2. PRIME NUMBERS
So this theorem says that you do not need to find all the primes less than x to
find out their number, it will be enough to evaluate x/logx for large x to find an
estimate for the number of primes. Notice that I mentioned that x has to be large
enough to be able to use this estimate.
Several other theorems were proved concerning prime numbers. many great
mathematicians approached problems that are related to primes. There are still
many open problems of which we will mention some.
Conjecture 1. Twin Prime Conjecture There are infinitely many pairs primes p
and p + 2.
Conjecture 4. Polignac Conjecture For every even number 2n are there infinitely
many pairs of consecutive primes which differ by 2n.
Congruences
51
52 CHAPTER 3. CONGRUENCES
Theorem 21. Let a, b, c and d denote integers. Let m be a positive integers. Then:
m | ((a + c) − (b + c))
and as a result
a + c ≡ b + c(mod m).
and as a result
a − c ≡ b − c(mod m).
5. If a ≡ b(mod m), then m | (a − b). Thus there exists integer k such that
a − b = mk and as a result ac − bc = m(kc). Thus
m | (ac − bc)
and hence
ac ≡ bc(mod m).
6. If a ≡ b(mod m), then m | (a − b). Thus there exists integer k such that
a − b = mk and as a result
ac − bc = mc(k).
Thus
mc | (ac − bc)
and hence
ac ≡ bc(mod mc).
54 CHAPTER 3. CONGRUENCES
As a result,
m | ((a + c) − (b + d)),
hence
a + c ≡ b + d(mod m).
(a − b) − (c − d) = (a − c) − (b − d) = m(k −
l).
As a result,
m | ((a − c) − (b − d)),
hence
a − c ≡ b − d(mod m).
As a result,
m | (ac − bd),
hence
ac ≡ bd(mod m).
3.1. INTRODUCTION TO CONGRUENCES 55
We now present a theorem that will show one difference between equations
and congruences. In equations, if we divide both sides of the equation by a non-
zero number, equality holds. While in congruences, it is not necessarily true. In
other words, dividing both sides of the congruence by the same integer doesn’t
preserve the congruence.
Theorem 22. 1. If a, b, c and m are integers such that m > 0, d = (m, c) and
ac ≡ bc(mod m), then a ≡ b(mod m/d).
Proof. Part 2 follows immediately from Part 1. For Part 1, if ac ≡ bc(mod m),
then
m | (ac − bc) = c(a − b).
56 CHAPTER 3. CONGRUENCES
Hence there exists k such that c(a − b) = mk. Dividing both sides by d, we get
(c/d)(a − b) = k(m/d). Since (m/d, c/d) = 1, it follows that m/d | (a − b).
Hence a ≡ b(mod m/d).
Example 25. 38 ≡ 10(mod 7). Since (2, 7) = 1 then 19 ≡ 5(mod 7).
Theorem 23. If
a ≡ b(mod m1 ), a ≡ b(mod m2 ), ..., a ≡ b(mod mt )
Exercises
Definition 13. A complete residue system modulo m is a set of integers such that
every integer is congruent modulo m to exactly one integer of the set.
The easiest complete residue system modulo m is the set of integers 0, 1, 2, ..., m−
1. Every integer is congruent to one of these integers modulo m.
Example 27. The set of integers {1, 5} is a reduced residue system modulo
6.
The following lemma will help determine a complete residue system modulo
any positive integer m.
Proof. We will prove this lemma by contradiction. Suppose that the set of m
integers does not form a complete residue system modulo m. Then we can find at
least one integer a that is not congruent to any element in this set. Hence non of
the elements of this set is actually congruent to the remainder when a is divided
by m. Thus dividing by m yields to at most m − 1 remainders. Therefore by the
pigeonhole principle, at least two integers in the set that have the same remainder
modulo m. This is a contradiction since the set of integers is formed of m
integers that are incongruent modulo m.
Proof. Let us prove first that no two elements of the set {ka1 +b, ka2 +b, ..., kam
+
b} are congruent modulo m. Suppose there exists i and j such that
ai ≡ aj (mod m)
Proof. The proof proceeds exactly in the same way as that of Theorem
24.
Exercises
1. Give a reduced residue system modulo 12.
Theorem 26. Let a, b and m be integers such that m > 0 and let c = (a, m). If c
does not divide b, then the congruence ax ≡ b(mod m) has no solutions. If c | b,
then
ax ≡ b(mod m)
x = x0 + (m/c)t
Thus the above values of x are solutions of the congruence ax ≡ b(mod m).
Now we have to determine the number of incongruent solutions that we have.
Suppose that two solutions are congruent, i.e.
Thus we get
(m/c)t1 ≡ (m/c)t2 (mod m).
t1 ≡ t2 (mod c).
3.3. LINEAR CONGRUENCES 61
Example 29. Let us find all the solutions of the congruence 3x ≡ 12(mod 6).
Notice that (3, 6) = 3 and 3 | 12. Thus there are three incongruent solutions
modulo 6. We use the Euclidean algorithm to find the solution of the equation
3x − 6y = 12 as described in chapter 2. As a result, we get x0 = 6. Thus the
three incongruent solutions are given by x1 = 6(mod 6), x1 = 6 + 2 = 2(mod
6) and x2 = 6 + 4 = 4(mod 6).
Exercises
x ≡ b1 (mod n1 ),
x ≡ b2 (mod n2 ),
.
.
.
x ≡ bt (mod nt ),
x0 ≡ x1 (mod N ).
We now present an example that will show how the Chinese remainder theo-
rem is used to determine the solution of a given system of congruences.
x ≡ 1(mod 2)
x ≡ 2(mod 3)
x ≡ 3(mod 5).
y1 ≡ 1(mod 2).
As a result, we get
Exercises
2. Find all integers that leave a remainder of 4 when divided by 11 and leaves
a remainder of 3 when divided by 17.
Theorem 28. Let p be a prime. A positive integer m is its own inverse modulo
p
if and only if p divides m + 1 or p divides m −
1.
Hence p | m2 − 1. As a result,
p | (m − 1)or p | (m + 1).
3.5. THEOREMS OF FERMAT, EULER, AND WILSON 65
Thus
m2 ≡ 1(mod p).
Proof. When p = 2, the congruence holds. Now let p > 2. Using Theorem
26, we see that for each 1 ≤ m ≤ p, there is an inverse 1 ≤ ≤ p such that
m¯
mm ≡ 1(mod p). Thus by Theorem 28, we see that the only two integers that
¯
have their own inverses are 1 and p − 1. Hence after coupling the integers from
2 to p − 2 each with its inverse, we get
Thus we get
1.2.3.....(p − 2)(p − 1) ≡ (p − 1)(mod p)
Note also that the converse of Wilson’s theorem also holds. The converse
tells us whether an integer is prime or not.
(m − 1)! + 1 ≡ 0 (mod m)
then m is prime.
66 CHAPTER 3. CONGRUENCES
we get
c1 | ((m − 1)! + 1).
Thus, we get
ap ≡ a(mod p).
68 CHAPTER 3. CONGRUENCES
Now if p | a, we have
ap ≡ a ≡ 0(mod p).
Theorem 33. If p is a prime number and a is an integer such that p - a, then ap−2
is the inverse of a modulo p.
Hence
ap−2 a ≡ 1(mod p).
Exercises
Later define the Mobius function which investigate integers in terms of their
prime decomposition. The summatory function of a given function takes the sum
of the values of f at the divisors of a given integer n. We then determine the
Mobius inversion of this function which writes the values of f in terms of the
values of its summatory function. We end this chapter by presenting integers
with interesting properties and prove some of their properties.
69
70 CHAPTER 4. MULTIPLICATIVE NUMBER THEORETIC FUNCTIONS
we have
Y
s
ak
f (n) = f (p
k
).
k=1
So we have to prove that if
Y
s+1
a
n= pk k ,
k=1
then
Y
s+1
ak
f (n) = f (p
k
).
k=1
Notice that for
Y
s+1
a
n= pk k ,
k=1
Qs a as+1
we have ( k=1 pk k , ps+1 ) = 1. Thus we have get
Y
s+1 sY
a
ak ak
f (n) = f ( pk ) = f ( kp )f s+1
(p s+1 )
k=1 k=1
From the above theorem, we can see that to evaluate a multiplicative function
at an integer, it will be enough to know the value of the function at the primes
that are in the prime factorization of the number.
We now define summatory functions which represents the sum of the values
of a given function at the divisors of a given number.
This function determines the sum of the values of the arithmetic function at
the divisors of a given integer.
Exercises
1. Determine whether the arithmetic functions f (n) = n! and g(n) = n/2 are
completely multiplicative or not.
4.2. MULTIPLICATIVE NUMBER THEORETIC FUNCTIONS 73
Proof. The first part is obvious since every positive integer less than p is
relatively prime to p. Conversely, suppose that p is not prime. Then p = 1 or p is
a composite number. If p = 1, then φ(p) = p − 1. Now if p is composite, then p
has a positive divisor. Thus φ(p) = p − 1. We have a contradiction and thus p is
prime.
Theorem 38. Let m and n be two relatively prime positive integers. Then φ(mn) =
φ(m)φ(n).
Proof. Denote φ(m) by s and let k1 , k2 , ..., ks be a reduced residue system modulo
m. Similarly, denote φ(n) by t and let k 01, k20 , ..., kt0 be a reduced residue system
modulo n. Notice that if x belongs to a reduced residue system modulo mn, then
(x, m) = (x, n) = 1.
Thus
x ≡ ki (mod m)and x ≡ k0 (mod n)
j
some i, j then (x, mn) = 1 and thus x belongs to a reduced residue system modulo
mn. Thus a reduced residue system modulo mn can be obtained by by determin-
ing all x that are congruent to ki and k0 modulo m and n respectively. By the
j
Chinese remainder theorem, the system of equations
has a unique solution. Thus different i and j will yield different answers. Thus
φ(mn) = st.
1
φ(pai i ) = piai − piai −1 = piai 1 − .
pi
a a
φ(n) = φ(p1 1 p2 2 ...psas )
a a
= φ(p11 )φ(p2 2 )...φ(psas )
1 1 1
= pa11 1 − p2a2 1 − ...pass 1 −
p1 p2 ps
a 1 1 1
= pa11 p2a2 ...pk k 1 − 1− ... 1 −
p1 p2 ps
1 1 1
= n 1− 1− ... 1 − .
p1 p2 ps
1 1
φ(200) = φ(23 52 ) = 200 1 − 1− = 80.
2 5
Theorem 40. Let n be a positive integer greater than 2. Then φ(n) is even.
a
Proof. Let n = p1a1 p2a2 ...pk k . Since φ is multiplicative, then
Y
k
a
φ(n) = φ(pj j ).
j=1
aj aj −1−1
φ(pj ) = pj (pj − 1).
aj
We see then φ(pj )is even if pj is an odd prime. Notice also that if pj = 2, then it
aj
follows that φ(pj ) is even. Hence φ(n) is even.
76 CHAPTER 4. MULTIPLICATIVE NUMBER THEORETIC
FUNCTIONS
Proof. Split the integers from 1 to n into classes. Put an integer m in the class Cd
if the greatest common divisor of m and n is d. Thus the number of integers in
the Cd class is the number of positive integers not exceeding n/d that are
relatively prime to n/d. Thus we have φ(n/d) integers in Cd . Thus we see that
X
n= φ(n/d).
d|n
Theorem 43. Let p be a prime and let n = p1a1 p2a2 ...pt at be a positive integer. Then
pa+1 − 1
σ(pa ) = ,
p− 1
and as a result,
Y
t paj +1 − 1
j
σ(n) =
j=1
pj − 1
24 − 1 53 − 1
Example 38. σ(200) = σ(23 52 ) = 2−1 5−1 = 15.31 = 465.
multiplicative.
Theorem 45. Let p be a prime and let n = pa1 pa2 ...pat be a positive integer.
Then 1 2 t
τ (pa ) = a + 1,
and as a result,
Y
t
τ (n) = (aj + 1).
j=1
τ (pa ) = a +
1
Exercises
7. Find the sum of positive integer divisors and the number of positive integer
divisors of 35
8. Find the sum of positive integer divisors and the number of positive integer
divisors of 25 34 53 73 13.
Note that if n is divisible by a power of a prime higher than one then µ(n) =
0. In connection with the above definition, we have the following
80 CHAPTER 4. MULTIPLICATIVE NUMBER THEORETIC
FUNCTIONS
Example 41. Notice that µ(1) = 1, µ(2) = −1, µ(3) = −1 and µ(4) = 0.
µ(mn) = µ(m)µ(n).
µ(mn) = 0 = µ(m)µ(n).
What remains to prove that if m and n are square-free integers say m = p1 p2 ...ps
where p1 , p2 , ..., ps are distinct primes and n = q1 q2 ...qt where q1 , q2 , ..., qt .
Since (m, n) = 1, then there are no common primes in the prime
decomposition be- tween m and n. Thus
Proof. For n = 1, we have F (1) = µ(1) = 1. Let us now find µ(pk ) for any
integer k > 0. Notice that
Thus by Theorem 36, for any integer n = p1a1 p2a2 ...pt at > 1 we have,
a a
F (n) = F (p1 1 )F (p2 2 )...F (ptat ) = 0
Theorem 48. Suppose that f is an arithmetic function and suppose that F is its
summatory function, then for all positive integers n we have
X
f (n) = µ(d)F (n/d).
d|n
Proof. We have
X X X
µ(d)F (n/d) = µ(d) f (e)
d|n
d|n e|(n/d)
X X
= µ(d)f (e)
d|n e|(n/d)
X X
= µ(d)f (e)
e|n d|(n/e)
X X
= f (e) µ(d)
e|n
d|(n/d)
82 CHAPTER 4. MULTIPLICATIVE NUMBER THEORETIC FUNCTIONS
P
Notice that d |(n/e) µ(d) = 0 unless n/e = 1 and thus e = n. Consequently we
get
X X
f (e) µ(d) = f (n).1 = f (n).
e|n
d|(n/d)
Example 42. A good example of a Mobius inversion formula would be the in-
version of σ(n) and τ (n). These two functions are the summatory functions
of f (n) = n and f (n) = 1 respectively. Thus we get
X
n= µ(n/d)σ(d)
d|n
and
X
1= µ(n/d)τ (d).
d|n
Exercises
2. Find the value of µ(n) for each integer n with 100 ≤ n ≤ 110.
P
3. Use the Mobius inversion formula and the identity n = d |n φ(n/d) to
t t t−1
show that φ(p ) = p − p where p is a prime and t is a positive integer.
In other words, a perfect number is a positive integer which is the sum of its
proper divisors.
Example 43. The first perfect number is 6, since σ(6) = 12. You can also view
this as 6 = 1 + 2 + 3. The second perfect number is 28, since σ(28) = 56 or
28 = 1 + 2 + 4 + 7 +
14.
The following theorem tells us which even positive integers are perfect.
Theorem 49. The positive integer n is an even perfect number if and only
if
σ(n) = 2n.
Notice now that (2r+1 − 1, 2r+1 ) = 1 and thus 2r+1 | σ(s). Therefore there
exists an integer q such that σ(s) = 2r+1 q. As a result, we have
So we get that q | s. We add q to both sides of the above equation and we get
We have to show now that q = 1. Notice that if q = 1, then s will have three
divisors and thus σ(s) ≥ 1 + s + q. Hence q = 1 and as a result s = 2r+1 − 1.
Also notice that σ(s) = s + 1. This shows that s is prime since the only divisors
of s are 1 and s. As a result,
n = 2r (2r+1 − 1),
Notice that the two factors above are both greater than 1. Thus 2m −1 is not
prime. This is a contradiction.
We prove a theorem that help decide whether Mersenne numbers are prime.
p1 = kp + 1 = 2mp +
1.
We now define Fermat numbers and prove some theorems about the
properties of these numbers.
Definition 26. Integers of the form Fn = 22n + 1 are called Fermat numbers.
Fermat conjectured that these integers are primes but it turned out that this is
not true. Notice that F0 = 3, F1 = 5, F2 = 17, F3 = 257 and F4 = 65, 537
while F5 is composite. It turned out the F5 is divisible by 641. We now present a
couple of theorems about the properties of these numbers.
86 CHAPTER 4. MULTIPLICATIVE NUMBER THEORETIC FUNCTIONS
F0 F1 F2 ...Fn−1 = Fn − 2
Proof. We will prove this theorem by induction. For n = 1, the above identity is
true. Suppose now that
F0 F1 F2 ...Fn−1 = Fn − 2
Notice that
n n n+1
F0 F1 F2 ...Fn = (Fn − 2)Fn = (22 − 1)(22 + 1) = 22 − 1 = Fn+1 − 2.
Using Theorem 53, we prove that Fermat numbers are relatively prime.
Proof. Assume without loss of generality that s < t. Thus by Theorem 52, we
have
F0 F1 F2 ...Fs ...Ft−1 = Ft − 2
Assume now that there is a common divisor d of Fs and Ft . thus we see that d
divides
Ft − F0 F1 F2 ...Fs ...Ft−1 = 2.
Exercises
4. We say n is abundant if σ(n) > 2n. Prove that if n = 2m−1 (2m − 1) where
m is a positive integer such that 2m − 1 is composite, then n is abundant.
89
90 CHAPTER 5. PRIMITIVE ROOTS AND QUADRATIC RESIDUES
by the well ordering principle, there is a least positive integer x that satisfies this
congruence ax ≡ 1(mod n).
To find all integers x such that ax ≡ 1(mod b), we need the following
theorem.
Theorem 54. If (a, b) = 1 with b > 0, then the positive integer x is a solution of
the congruence ax ≡ 1(mod b) if and only if ordb a | x.
Proof. Having ordb a | x, then we have that x = k.ordb a for some positive
integer
k. Thus
ax = akordb a = (aordb a )k ≡ 1(mod b).
Now if ax ≡ 1(mod b), we use the division algorithm to write
Now since ax ≡ 1(mod b),we have ar ≡ 1(mod b). Since ordb a, we get r = 0.
Thus x = q.ordb a and hence ordb a | x.
Example 47. Since ord7 2 = 3, then 215 ≡ 1(mod 7) while 10 is not a solution
for 2x ≡ 1(mod 7).
ai ≡ aj (mod b)
5.1. THE ORDER OF INTEGERS AND PRIMITIVE ROOTS 91
i ≡ j(mod ordb a)
ai ≡ aj ai−j ≡ aj (mod
b)
We introduce now primitive roots and discuss their properties. We are inter-
ested in integers whose order modulo another integer is φ(b). In one of the exer-
cises, one is asked to prove that if aand b are relatively prime then ordb a | φ(b).
Example 48. Notice that φ(7) = 6 hence 2 is not a primitive root modulo 7. While
ord7 3 = 6 and thus 3 is a primitive root modulo 7.
Proof. To prove that the set {r 1 , r2 , ...r φ(m) } form a reduced residue set
modulo m we need to show that every two of them are relatively prime and
that no two of them are congruent modulo m. Since (r, m) = 1, it follows that
(r n , m) = 1 for all positive integers n. Hence all the powers of r are relatively
prime to m. To show that no two powers in the above set are equivalent modulo
m, assume that
r i ≡ r j (mod m).
Proof. Let
Thus ordm r u = φ(m) and r u is a primitive root if and only if (u, φ(m)) = 1.
Exercises
1. Determine ord13 10.
2. Determine ord11 3.
Example 50. Notice that x ≡ 3(mod 11) is a root for f (x) = 2x2 + x + 1 since
f (3) = 22 ≡ 0(mod 11).
We now introduce Lagrange’s theorem for primes. This is modulo p, the fun-
damental theorem of algebra. This theorem will be an important tool to prove
that every prime has a primitive root.
m(x) = bn xn + −1 x
n−1
+ ... + b1 x + b0
bn
m(x) = b1 x + b0 and p - b1 .
A root of m(x) is a solution for b1 x+b0 (mod p). Since p - b1 , then this
congruence has exactly one solution by Theorem 26.
Suppose that the theorem is true for polynomials of degree n − 1, and let
m(x) be a polynomial of degree n with integer coefficients and where the leading
coefficient is not divisible by p. Assume now that m(x) has n + 1 incongruent
roots modulo p, say x0 , x1 , ..., xn . Thus
m(xk ) ≡ 0(mod p)
5.2. PRIMITIVE ROOTS FOR PRIMES 95
Thus f (xk ) ≡ 0(mod p) for all 1 ≤ k ≤ n and thus x1 , x2 , ..., xn are roots of
f (x). This is a contradiction since we a have a polynomial of degree n − 1 that
has n distinct roots.
Theorem 60. Consider the prime p and let p − 1 = kn for some integer k. Then
xn − 1 has exactly n incongruent roots modulo
p. Proof. Since p − 1 = kn, we have
We now prove a lemma that gives us how many incongruent integers can
have a given order modulo p.
Lemma 11. Let p be a prime and let m be a positive integer such that p − 1 =
mk
for some integer k.
Then
for all positive integers k. By Theorem 60, we know that xm − 1 has exactly m
incongruent roots modulo p, so that every root is congruent to one of these
powers of a. We also know by Theorem 57 that the powers of ak with (k, m)
= 1 have order m. There are exactly φ(m) such integers with 1 ≤ k ≤ m and
thus if there is one element of order m modulo p, there must be exactly φ(m)
such positive integers less than p. Hence S(m) ≤ φ(m).
Proof. Let p be a prime and let m be a positive integer such that p − 1 = mk for
some integer k. Let F (m) be the number of positive integers of order m modulo
p that are less than p. The order modulo p of an integer not divisible by p divides
p − 1, it follows that X
p− 1= F (m).
m|p−1
5.2. PRIMITIVE ROOTS FOR PRIMES 97
we see that F (m) = φ(m) for each positive divisor m of p − 1. Thus we conclude
that F (m) = φ(m). As a result, we see that there are p − 1 incongruent integers
of order p − 1 modulo p. Thus p has φ(p − 1) primitive roots.
Exercises
9. Show that if p is a prime and p ≡ 1(mod 4), then there is an integer x such
that x2 ≡ −1(mod p).
98 CHAPTER 5. PRIMITIVE ROOTS AND QUADRATIC
RESIDUES
Theorem 62. If p is an odd prime with primitive root r, then one can have either
r or r + p as a primitive root modulo p2
.
ordp r = φ(p) = p − 1.
p − 1 | m.
By Theorem 54, we have
m | φ(p2 ).
r p−1 ≡ 1(mod p2 ).
5.3. THE EXISTENCE OF PRIMITIVE ROOTS 99
Hence
p2 | sp−1 − (1 −
then
p2 | pr p−2 .
Thus we have
p | r p−2
We now show that any power of an odd prime has a primitive root.
Theorem 63. Let p be an odd prime. Then any power of p is a primitive root.
Moreover, if r is a primitive root modulo p2 , then r is a primitive root modulo
pm for all positive integers m.
100 CHAPTER 5. PRIMITIVE ROOTS AND QUADRATIC RESIDUES
Proof. By Theorem 62, we know that any prime p has a primitive root r which
is also a primitive root modulo p2 , thus
for all integers m ≥ 2. Once we prove the above congruence, we show that r is
also a primitive root modulo pm . Let n = ordpm r. By Theorem 54, we know that
n | φ(pm ). Also, we know that φ(pm ) = pm (p − 1). Hence n | pm (p − 1). On
the other hand, because
pm | (r n − 1),
Since φ(p) = p − 1, we see that by Theorem 54, we have n = l(p − 1). also
n | pm−1 (p − 1), we have that n = ps (p − 1), where 0 ≤ s ≤ m − 1. If
n = ps (p − 1) with s ≤ m − 2, then
m−2 (p−1)
pk | r p − 1,
ordpm r = φ(pm ).
We prove now (7.5) by induction. Assume that our assertion is true for all
m ≥ 2. Then
pm - (r pm−2 (p−1) − 1).
Because (r, p) = 1, we see that (r, pm−1 ) = 1. We also know from Euler’s
theorem that
m−2
pm−1 | (r p (p−1)
− 1).
5.3. THE EXISTENCE OF PRIMITIVE ROOTS 101
m−2
rp (p−1)
= 1 + kpm−1 .
m−1
rp (p−1)
= (1 + kpm−1 )p
≡ 1 + kpm (mod pm+1 )
Because p - k, we have
m−1
(p−1)
pm+1 - (r p − 1).
Example 52. Since 3 is a primitive root of 7, then 3 is a primitive root for 7k for
all positive integers k.
≡ 1(mod 2k ).
k−2
m2
m2 = 4n2 + 4n + 1 = 4n(n + 1) +
1.
m2 = 1 + q.2k .
k−2
m2 = 1 + q.2k+1 + q 2 22k
k−1
Thus
2k+1 | (m2
k−1
− 1).
Theorem 65. If m is not pa or 2pa , then m does not have a primitive root.
s
Proof. Let m = ps1 ps2 ...p i . If m has a primitive root r then r and m are
relatively 1 2 i
ps | (rφ(p s) − 1).
Now let
s
L = [φ(ps11 ), φ(p2s2 ), ..., φ(pi i )].
We know that
s
r L ≡ 1(mod pkk )
for all 1 ≤ k ≤ m. Thus using the Chinese Remainder Theorem, we get
m | (r L − 1),
are not relatively prime unless m = ps or m = 2ps where p is an odd prime and t
is any positive integer.
We now show that all integers of the form m = 2ps have primitive
roots.
Theorem 66. Consider a prime p = 2 and let s is a positive integer, then 2ps
has a primitive root. In fact, if r is an odd primitive root modulo ps , then it is
also a primitive root modulo 2ps but if r is even, r + ps is a primitive root
modulo 2ps .
ps | (r φ(p s) − 1)
and no positive exponent smaller than φ(ps ) has this property. Note also that
φ(2ps ) = φ(ps ),
so that
ps | (rφ(2p s ) − 1).
If r is odd, then
2 | (rφ(2p s) − 1).
ps | ((r + ps )φ(2p
s )
−
1).
As a result, we see that 2ps | ((r + ps )φ(2ps ) − 1) and since for no smaller power
of
r + ps is congruent to 1 modulo 2ps , we see that r + ps is a primitive root
modulo
2ps .
Theorem 67. The positive integer m has a primitive root if and only if n = 2, 4, ps
or 2ps
Exercises
1. Which of the following integers 4, 12, 28, 36, 125 have a primitive root.
4. Show that there are the same number of primitive roots modulo 2ps as
there are modulo ps , where p is an odd prime and s is a positive integer.
6. Show that the integer n has a primitive root if and only if the only solutions
of the congruence x2 ≡ 1(modn) are x ≡ ±1(mod n).
5.4. INTRODUCTION TO QUADRATIC RESIDUES AND
NONRESIDUES105
x2 ≡ a(mod p)
Hence
x0 ≡ x00 (mod p) or x0 ≡ −x00 (mod p).
106 CHAPTER 5. PRIMITIVE ROOTS AND QUADRATIC
RESIDUES
The following theorem determines the number of integers that are quadratic
residues modulo an odd prime.
Proof. To find all the quadratic residues of p among all the integers 1, 2, ..., p −
1, we determine the least positive residue modulo p of 12 , 22 , ..., (p − 1)2 .
Consider- ing the p − 1 congruences and because each congruence has either no
solution or two incongruent solutions, there must be exactly (p − 1)/2 quadratic
residues of p among 1, 2, ..., p − 1. Thus the remaining are (p − 1)/2 quadratic
nonresidues of p.
Exercises
4. Show that if p is prime and p ≥ 7, then there are always two consecutive
quadratic residues of p. Hint: Show that at least one of 2, 5 or 10 is a
quadratic residue of p.
5. Show that if p is prime and p ≥ 7, then there are always two quadratic
residues of p that differ by 3.
Example 54. Notice that using the previous example, we see that
1 2
= = 4 =1
7 7
3 7
= 5 = −1
7 =
7 6
7
a
Proof. Assume that p
= 1. Then the congruence x2 ≡ a(mod p) has a
solution say x = x0 . According to Fermat’s theorem, we see that
a
Now if p
= −1, then x2 ≡ a(mod p) is not solvable. Thus by Theorem 26,
we have that for each integer k with (k, p) = 1 there is an integer l such that
kl ≡ a(mod p). Notice that i = j since x2 ≡ a(mod p) has no solutions. Thus
we can couple the integers 1, 2, ..., p − 1 into (p − 1)/2 pairs, each has product
a. Multiplying these pairs together, we find out that
3
Example 55. Let p = 13 and a = 3. Then 13
= −1 ≡ 36 (mod 13).
a ≡ aφ(p)/2 (mod p)
p
and
φ(p)/2
b
≡b (mod p).
p
Thus we get
a b ab
≡ (ab)φ(p)/2 ≡ (mod p).
p p p
We now show when is −1 a quadratic residue of a prime p .
5.5. LEGENDRE SYMBOL 109
Corollary 3. If p = 2 is a, then
(
−1 1 if p ≡ 1(mod 4)
= −1 if p ≡ −1(mod 4).
p
Proof. By Euler’s criterion, we know that
a
= (−1)φ(p)/2 (mod p)
p
If 4 | (p − 1), then p = 4m + 1 for some integer m and thus we get
(−1)φ(p)/2 = (−1)2m = 1.
and if 4 | (p − 3), then p = 4m + 3 for some integer m and we also get
(−1)φ(p)/2 = (−1)2m+1 = −1.
p− 1
Now notice that 2 ! 6≡ 0(mod p) and thus we get
2
2(p−1)/2 ≡ (−1)(p −1)/8
(mod p).
Proof. Let m1 , m2 , ..., ms be those integers greater than p/2 in the set of the
least positive residues of the integers a, 2a, ..., ((p − 1)/2)a and let n1 , n2 , ...,
nt be
those less than p/2. We now show that
p − m1 , p − m2 , ..., p − mk , p − n1 , p − n2 , ..., p −
nt
Y
k Y
t
p− 1
(p − mi ) ni !(mod p),
2
≡
i=1 i=1
which implies
p− 1
(−1)s m1 m2 ...(p − mk )n1 n2 ...nt ≡ !(mod p),
2
Simplifying, we get
Thus we get
a(p−1)/2 ≡ (−1)k (mod p).
135
Example 56. To find using Gauss’s lemma, we calculate
X
6
[5i/13] = [5/13] + [10/13] + [15/13] + [20/13] + [25/13] + [30/13] = 5
i=1
5
Thus we get 13 = (−1)5 = −1.
Exercises
7j
3. Find the value of Legendre symbol for j = 1, 2, 3, 4, 5, 6.
11
7
4. Evaluate the Legendre symbol by using Euler’s criterion.
5. Let a and b be integers not divisible by p. Show that either one or all three
of the integers a, b and ab are quadratic residues of p.
6. Let p be a prime and a be a quadratic residue of p. Show that if p ≡
1(mod 4), then −a is also a quadratic residue of p, whereas if p ≡ 3(mod
4), then −a is a quadratic nonresidue of p.
ap2
= 1.
P (p−1)/2
a = (−1) i=1 [ia/p]
.
p
Proof. Consider the least positive residues of the integers a, 2a, ..., ((p − 1)/2)a;
let m1 , m2 , ..., ms be integers of this set such that mi > p/2 for all i and let
n1 , n2 , ..., nt be those integers where ni < p/2. Using the division algorithm, we
see that
ia = p[ia/p] + r
where r is one of the mi or ni . By adding the (p − 1)/2 equations, we obtain
(p−1)/2
X X
(p−1)/2 X
s X
t
ia = p[ia/p] + mi + ni . (5.3)
i=1 i=1 i=1 i=1
p − m1 , p − m2 , ..., p − ms , p − n1 , p − n2 , ..., p − nt
are precisely the integers 1, 2, ..., (p − 1)/2, in the same order. Now we obtain
(p−1)/2
X X
s X
t sX t X
i= (p − mi ) + ni = ps − mi + ni . (5.4)
i=1 i=1 i=1 i=1 i=1
(p−1)/2
X X
(p−1)/2 (p−1)/2
X X
s
ia − i= p[ia/p] − ps + 2 mi .
i=1 i=1 i=1 i=1
Now since we are taking the following as exponents for −1, it suffice to look at
them modulo 2. Thus
X
(p−1)/2
(p−1)/2
X
[ia/p] ≡ s(mod 2)
i=1
Using Gauss’s lemma, we get
a P(p−1)/2
p = (−1)s = (−1) i=1 [ia/p]
.
Theorem 73. The Law of Quadratic Reciprocity Let p and q be distinct odd
primes. Then
p q p− 1
. q−21
= (−1) 2
q p
Proof. We consider now the pairs of integers also known as lattice points (x, y)
with
1 ≤ x ≤ (p − 1)/2and 1 ≤ y ≤ (q −
1)/2.
p− 1
The number of such pairs is . q− 1 . We divide these pairs into two groups
2 2 de-
pending on the sizes of qx and py. Note that qx = py for all pairs because p and
q are distinct primes.
We now count the pairs of integers (x, y) with
1 ≤ x ≤ (p − 1)/2, 1 ≤ y ≤ (q − 1)/2and qx >
py.
is
(p−1)/2
X
[qi/p].
i=1
5.6. THE LAW OF QUADRATIC RECIPROCITY 115
p p p− 1 . q−21
= (−1) 2
q q
Exercises
53
3
1. Evaluate .
641
31
2. Evaluate .
3. Using the law of quadratic reciprocity, show that if p is an odd prime, then
(
3 1 if p ≡ ±1(mod 12)
=
p −1 if p ≡ ±5(mod 12).
a Ym
a ci
= .
n i=1
pi
Example 57. Notice that from the prime factorization of 45, we get that
2
= 2 2 = (−1)(−1) = 1
55 5 11
We now prove some properties for Jacobi symbol that are similar to the prop-
erties of Legendre symbol.
Theorem 74. Let n be an odd positive integer and let a and b be integers such
that(a, n) = 1 and (b, n) = 1. Then
1. if n | (a − b), then
a b
= .
n n
2.
ab a
= .
n n nb
a b
= .
p p
5.7. JACOBI SYMBOL 117
As a result, we have
a Y
m
a ci Y
m
b ci
= =
n i=1
pi i=1
pi
ab a b
Proof of 2: Note that by Theorem 71, we have p = p p for any prime p
appearing in the prime factorization of n. As a result, we have
ab Y
m
= ab
ci
n
i=1 pi
m
Y a ci Y
m
b ci
=
i=1
pi i=1
pi
a b
= .
n n
−1 2
In the following theorem, we determine n and n .
1.
−1
= (−1)(n−1)/2 .
n
2.
2 2
n = (−1)(n −1)/8 .
−1 Y
m
−1
ci
=
n i=1
pi
P
= (−1)m i=1 ci (pi −1)/2 .
118 CHAPTER 5. PRIMITIVE ROOTS AND QUADRATIC RESIDUES
a
pi i = (1 + (pi − 1))ci ≡ 1 + ci (pi − 1)(mod 4)
As a result, we have
X
m
(n − 1)/2 ≡ ci (pi − 1)/2 (mod 2).
i=1
2
2 = (−1)(p −1)/8
.
p
Hence
2 Pm
= (−1) i=1 ci (pi −1)/8 .
2
n
2
Because 8 | p − 1, we see similarly that
i
(1 + (pi2 − 1))ci ≡ 1 + ci (p
i
2
− 1)(mod 64)
and thus
X
m
n2 ≡ 1 + ci (p
i
2
− 1)(mod 64),
i=1
We now show that the reciprocity law holds for Jacobi symbol.
5.7. JACOBI SYMBOL 119
b a Y
n Y
m
pj
cj di
qi
=
a b i=1 j=1
qi pj
and
n
−1 b− 1
X qi
di ≡ (mod 2).
i=1
2 2
Thus we conclude that
X m i −1 b− 1
c j pj − 1 X
n
≡a− 1.
qi 2 2 2 (mod 2).
j=1 2
di=1
Exercises
4520
258
1. Evaluate .
2307
1008
2. Evaluate .
3. For which positive integers n that are relatively prime to 15 does the Jacobi
n
15
symbol equal 1?
4. Let n be an odd square free positive integer. Show that there is an integer a
In this chapter, we introduce continued fractions, prove their basic properties and
apply these properties to solve some problems. Being a very natural object, con-
tinued fractions appear in many areas of Mathematics, sometimes in an unex-
pected way. The Dutch mathematician and astronomer, Christian Huygens
(1629-
1695), made the first practical application of the theory of ”anthyphaeiretic
ratios” (the old name of continued fractions) in 1687. He wrote a paper
explaining how to use convergents to find the best rational approximations for
gear ratios. These approximations enabled him to pick the gears with the best
numbers of teeth. His work was motivated by his desire to build a mechanical
planetarium. Further continued fractions attracted attention of most prominent
mathematicians. Euler, Jacobi, Cauchy, Gauss and many others worked with the
subject. Continued frac- tions find their applications in some areas of
contemporary Mathematics. There are mathematicians who continue to
develop the theory of continued fractions nowadays, The Australian
mathematician A.J. van der Poorten is, probably, the most prominent among
them.
121
122 CHAPTER 6. INTRODUCTION TO CONTINUED FRACTIONS
Notation 1. We write
1
[a0 ; a1 , a2 , . . . , an ] = a0
+ a1 + + . 1. .
a2
+ an
1
Still, in the case of infinite number of terms a certain amount of work must
be carried out in order to make the above formula meaningful. At the same time,
for the finite number of terms the formula makes sense.
Example 58.
we have
Example 59. Consider the continued fraction expansion for 42/31. We obtain
a0 = [42/31] = 1, δ = 42/31 − 1 = 11/31. Now r1 = 1/δ = 31/11 and
a1 = [α1 ] = [31/11] = 2. The new δ = 31/11 − 2 = 9/11. Now r2 = 1/δ =
11/9 and a2 = [α2 ] = [11/9] = 1. It follows that δ = 11/9 − 1 = 2/9.
Now r3 = 1/δ = 9/2 and a3 = [α3 ] = [9/2] = 4. It follows that δ = 9/2 − 4
= 1/2. Now r4 = 1/δ = 2 and a4 = [α4 ] = [2] = 2. It follows that δ = 2 − 2 =
0 and we are done.
124 CHAPTER 6. INTRODUCTION TO CONTINUED FRACTIONS
The above example shows that the algorithm stops after finitely many steps.
This is in fact quite a general phenomenon. In order to practice with the
introduced notations let us prove a simple but important proposition.
Theorem 77. An infinite continued fraction converges and defines a real number.
There is a one-to-one correspondence between
• all (finite and infinite) continued fractions [a0 ; a1 , a2 , . . .] with an integer a0
and positive integers ak for k > 0 (and the last term an > 1 in the case of finite
continued fractions)
and
• real numbers.
Note that the algorithm we developed above can be applied to any real
number and provides the corresponding continued fraction.
Theorem 77 has certain theoretical significance. L.Kronecker (1823-1891)
said, ”God created the integers; the rest is work of man”. Several ways to
represent real numbers out of integers are well-known. Theorem 77 provides
yet another way to fulfill this task. This way is constructive and at the same time
is not tied to any particular base (say to decimal or binary decomposition).
We will discuss some examples later.
Exercises
1. Prove that under the assumption an > 1 the continued fraction representa-
tion given in Proposition 1 is unique. In other words, the correspondence
126 CHAPTER 6. INTRODUCTION TO CONTINUED
FRACTIONS
between
• finite continued fractions [a0 ; a1 , a2 , . . . an ] with an integer a0 ,
positive integers ak for k > 0 and an > 1
and
• rational numbers
is one-to-one.
pk
sk = (6.4)
qk
pk = ak pk−1 + pk−2
(6.5)
qk = ak qk−1 + qk−2 .
Remark. It does not matter here whether we deal with finite or infinite con-
tinued fractions: the convergents are finite anyway. Proof. We use the induction
argument on k. For k = 2 the statement is true.
Now, assume (6.5) for 2 ≤ k < l. Let
pl
α = [a0 ; a1 , a2 , . . . al ] =
ql
6.2. MAIN TECHNICAL TOOL 127
β = [a1 ; a2 , . . . , al
]
and denote by p0r /qr0 its r-th convergent. We have α = a0 + 1/β which translates
as
0 0
pl = a0 pl−1 + l−1
0 q (6.6)
ql = pl−1 .
Also, by the induction assumption,
0 = al p0l−2 + p0l−3
pl−1
(6.7)
0 0
ql−1 = al ql−2 + q0l−3
and
0
ql = al pl−2 + 0l−3 = al ql−1 + ql−2 ,
p
which complete the induction step. We have thus proved that
pk
sk = ,
qk
where pk and qk are defined by the recursive formulas (6.5). We still have to
check that these are the quantities defined by (6.4), namely that qk > 0 and that
qk and pk are relatively prime. The former assertion follows from (6.5) since ak
> 0 for k > 0. To prove the latter assertion, multiply the equations (6.5) by qk−1
and pk−1 respectively and subtract them. We obtain
pk−1 pk ( −1)k
− = (6.9)
qk−1 qk qk qk−1
and
pk−2 pk ( −1) k ak
− = .
qk−2 qk qk qk−2
Since all the numbers qk and ak are positive, the above formulas imply the
follow- ing.
Lemma 15. If
a c
≤
b d
then
a a+c c
≤ ≤ .
b b+ d d
Consider now the sequence of fractions
pk pk + pk+1 pk + 2pk+1 pk + ak pk+2
, , ,..., = , (6.10)
pk+1
qk qk + qk+1 qk + 2qk+1 qk + ak qk+1 qk+2
where the last equality follows from (6.5).
It follows that the sequence (6.10) is increasing if k is even and is decreasing
if k is odd. Thus, in particular, the fraction
pk + pk+1
qk + qk+1 (6.11)
is between the quantities pk /qk and α. Therefore the distance between pk /qk
and the fraction (6.11) is smaller than the distance between pk /qk and α:
pk p + pk+1 1
α− ≥ k = .
qk qk + qk+1 q k k + qk+1 )
(q
The second (right) inequality in Theorem 79 is now proved. This finishes the
proof of Theorem 79.
Exercises
4. Prove Proposition 2
pk 1
α− ≤ .
qk qk qk+1
6. Prove Lemma 15
7. Use (6.5) to show that the sign of the difference between two consecutive
fractions in (6.10) depends only on the parity of k.
Also
a pk+1 a 1
α− ≥ − ≥ ,
b qk+1 b bqk+1
which implies
1
|bα − a| ≥ .
qk+1
At the same time Theorem 79 (it right inequality multiplied by qk ) reads
1
|qk α − pk | ≤ .
qk+1
It follows that
|qk α − pk | ≤ |bα − a| ,
132 CHAPTER 6. INTRODUCTION TO CONTINUED
FRACTIONS
and the latter inequality together with (6.12) show that a/b is not a ”good” ap-
proximation of α in this case.
This finishes the proof of Theorem 80.
Exercises
2. Show that if a/b > p1 /q1 then a/b is not a ”good” approximation to α.
6.4 An Application
Consider the following problem which may be of certain practical interest. As-
sume that we calculate certain quantity using a computer. Also assume that we
know in advance that the quantity in question is a rational number. The com-
puter returns a decimal which has high accuracy and is pretty close to our
desired answer. How to guess the exact answer?
To be more specific consider an example.
123456
121169
α = 123456/121169 + 10−15 =
1.018874464590779169333740478175110795665558022266421279370135
92
58556231379313190667579991582005298384900428327377464533007617
45
99113634675535821868629765038912593155014896549447465936006734395761292
07
with some two hundred digits of accuracy which, of course come short to help
in guessing the period and the exact denominator of 121169.
6.5. A FORMULA OF GAUSS, A THEOREM OF KUZMIN AND LE´ VI AND A PROBLEM OF
ARNOLD1
We are not going to check all convergents, because we notice the irregularity:
one element, 68110 is far more than the others. In order to explain this we use
the left inequality from Theorem 79 together with the formula (6.5). Indeed, we
have an approximation of α which is unexpectedly good: |α − pk /qk | is very
small (it is around 10−15 ) and with a modest qk too. We have
and
pk 1
α− .
qk ≥ q 2
k (ak+1 + qk−1 /qk )
It follows that 1/qk2 (ak+1 + qk −1 /qk ) is small (smaller than 10−15 ) and therefore,
ak+1 should be big. This is exactly what we see. Of course, our guess is correct:
123456
= [1, 52, 1, 53, 2, 4, 1, 2, 1].
121169
In this way we conclude that in general an unexpectedly big element allows
to cut the continued fraction (right before this element) and to guess the exact
rational quantity. There is probably no need (although this is, of course,
possible) to quantify this procedure. I prefer to use it just for guessing the correct
quantities on the spot from the first glance.
Theorem 81. For almost every real α the probability for a number k to appear as
an element in the continued fraction expansion of α is
1 1
ck = ln 1 + . (6.13)
ln 2 k(k + 2)
Remarks. 1. The words ”for almost every α” mean that the measure of the set
of exceptions is zero.
2. Even the existence of pk (defined as a limit) is highly non-trivial.
Theorem 81 may (and probably should) be considered as a result from
ergodic theory rather than number theory. This constructs a bridge between
these two ar- eas of Mathematics and explains the recent attention to continued
fractions of the mathematicians who study dynamical systems. In particular,
V.I.Arnold formu- lated the following open problem. Consider the set of pairs of
integers (a, b) such that the corresponding points on the plane are contained in a
quarter of a circle of radii N :
a 2 + b2 ≤ N
2
.
Expand the numbers p/q into continued fractions and compute the frequencies
sk for the appearance of k in these fractions. Do these frequencies have limits
as N → ∞? If so, do these limits have anything to do with the probabilities,
given by (6.13)? These questions demand nothing but experimental computer
investigation, and such an experiment may be undertaken by a student. Of
course, it would be extremely challenging to find a phenomena experimentally in
this way and to prove it after that theoretically.
Of course, one can consider more general kinds of continued fractions. In
particular, one may ease the assumption that the elements are positive integers
6.5. A FORMULA OF GAUSS, A THEOREM OF KUZMIN AND LE´ VI AND A PROBLEM OF
ARNOLD1
and consider, allowing arbitrary reals as the elements (the question of conver-
gence may usually be solved). The following identities were discovered inde-
pendently by three prominent mathematicians. The English mathematician R.J.
Rogers found and proved these identities in 1894, Ramanujan found the iden-
tities (without proof) and formulated them in his letter to Hardy from India in
1913. Independently, being separated from England by the war, I. J. Schur found
the identities and published two different proofs in 1917. We refer an interested
reader to [2] for a detailed discussion and just state the amazing identities here.
s
√ √
5+ 5 5 + 1 2π/5
[0; e−2π , e−4π , e−6π , e−8π , . . .] = − e
2 2
s √ √
5− 5 5− 1
[1; e−π , e−2π , e−3π , e−4π , . . .] = − eπ/5
2 2
Exercises
The distribution of prime numbers has been the object of intense study by many
modern mathematicians. Gauss and Legendre conjectured the prime number the-
orem which states that the number of primes less than a positive number x is
asymptotic to x/logx as x approaches infinity. This conjecture was later proved
by Hadamard and Poisson. Their proof and many other proofs lead to the what is
known as Analytic Number theory.
In this chapter we demonstrate elementary theorems on primes and prove el-
ementary properties and results that will lead to the proof of the prime number
theorem.
7.1 Introduction
P∞
It is well known that the harmonic series n=1 n1 diverges. We therefore deter-
P
mine some asymptotic formulas that determines the growth of the n ≤x n1 . We
start by introducing Euler’s summation formula that will help us determine the
asymptotic formula.
137
138 CHAPTER 7. INTRODUCTION TO ANALYTIC NUMBER THEORY
We might ask the following question. What if the sum is taken over all the
primes. In this section, we show that the sum over the primes diverges as well.
We also show that an interesting product will also diverge. From the following
theorem, we can actually deduce that there are infinitely many primes.
For the proof of Euler’s summation formula see [3, Chapter 3].
Proof. We use Euler’s summation formula by taking f (t) = 1/t. We then get
X1 Zx Z x
1 {t} 1
= dt − dt + 1 + O
n≤x
n 1 t 1 t 2 x
Z∞ Z∞
{t} {t}
= log x + 1 − 2 dt + dt + O 1
1 t x t2
x
Notice now that {t} ≤ t and hence the two improper integrals exist since they
are
dominated by integrals that converge. We therefore have
Z∞
{t} 1
0≤ 2 dt ≤ ,
x t x
we also let Z∞
{t}
γ = 1− dt
1 t2
7.1. INTRODUCTION 139
and we get the asymptotic formula. Notice that γ is called Euler’s constant.
Notice also that similar steps can be followed to find an asymptotic formulas
for other sums involving powers of n.
We now proceed to show that if we sum over the primes instead, we still get
a divergent series.
Pp p
Qp p
Theorem 82. Both 1 and 1 (1 − )
diverge.
Y 2X
1 X 1
m−1 [x/2]
1 1
1+ + ... + > >
p≤x
p pm n=1
n n=1
n
140 CHAPTER 7. INTRODUCTION TO ANALYTIC NUMBER THEORY
u2
log(1/u − 1) < u + (1/1 − u), 0 < u < 1.
2
We now let u = 1/p for each p ≤ x, then
1 1
log 1 − <
p 2p(p − 1)
1 − 1/p
Thus
X
log P (x) = log(1/1 − p).
p≤x
Thus we have
∞
1X 1 1X 1
log P (x) − S(x) < < n(n − 1)
2 p≤x p(p − 1) 2 n=1
where A(x) = 0 for x < 1. Assume also that g has a continuous derivative on the
interval [y, x], where 0 < y < x. Then we have
X Zx
f (n)g(n) = A(x)g(x) − A(y)g(y) − A(t)g 0 (t)dt.
y<n≤x y
7.2. CHEBYSHEV’S FUNCTIONS 141
1. Show that one gets every 1 , n ∈ Z+ where each prime factor of n is less
n
than or equal to x in the proof of Theorem 1.
We define also the following functions, the last two functions are called
Cheby- shev’s functions.
P
1. π(x) = p ≤x 1.
P
2. θ(x) = p≤x logp
P
3. ψ(x) = n ≤x Ω(n)
Notice that
X X
∞ X X
∞ X
ψ(x) = Ω(n) Ω(pm ) = logp.
=
n≤x
m=1, pm ≤x p m=1 p≤x1/m
Notice that the above sum will be a finite sum since for some m, we have that
x1/m < 2 and thus θ(x1/m ) = 0.
We use Abel’s summation formula now to express the two functions π(x) and
θ(x) in terms of integrals.
Theorem 84. For x ≥ 2, we have
Zx
θ(x) = π(x) log x − π(t)
dt
2 t
and Zx
θ(x) θ(t)
π(x) = + dt.
log x 2 t log2 t
Proof. We define the characteristic function χ(n) to be 1 if n is prime and 0 oth-
erwise. As a result, we can see from the definition of π(x) and θ(x) that they can
be represented in terms of the characteristic function χ(n). This representation
will enable use to apply Abel’s summation formula where f (n) = χ(n) for θ(x)
and where f (n) = χ(n) log n for π(x). So we have,
X X
π(x) = χ(n) and θ(x) = χ(n) log n
1≤n/leqx 1≤n≤x
Now let g(x) = log x in Theorem 84 with y = 1 and we get the desired result for
the integral representation of θ(x). Similarly we let g(x) = 1/ log x with y =
3/2
and we obtain the desired result for π(x) since θ(t) = 0 for t < 2.
We now prove a theorem that relates the two Chebyshev’s functions θ(x) and
ψ(x). The following theorem states that if the limit of one of the two functions
θ(x)/x or ψ(x)/x exists then the limit of the other exists as well and the two
limits are equal.
7.3. GETTING CLOSER TO THE PROOF OF THE PRIME NUMBER THEOREM143
Exercises
1. Show that
where m ≤ log2 x.
√ √
2. Show that 0 ≤ ψ(x) − θ(x) ≤ (log2 (x)) x log x and thus the result of
Theorem 86 follows.
θ(x) = x + O x
log x
π(x) log x
lim =1 (7.1)
x→∞ x
θ(x)
lim =1 (7.2)
x→∞ x
ψ(x)
lim = 1. (7.3)
x→∞ x
Proof. We have proved in Theorem 86 that (7.2) and (7.3) are equivalent, so if
we show that (7.1) and (7.2) are equivalent, the proof will follow. Notice that
using the integral representations of the functions in Theorem 85, we obtain
Z
θ(x) π(x) log x 1 x π(t)
= − dt
x x x 2 t
and Z x
π(x) log x θ(x) log x θ(t)
= + dt.
x x x 2 t log2 t
Now to prove that (7.1) implies (7.2), we need to prove that
Z
1 x π(t)
lim dt = 0.
x→∞ x 2 t
π(t) 1
Notice also that (7.1) implies that t
=O log t
for t ≥ 2 and thus we have
Z x Z x
1 π(t) 1 dt
dt = O
x 2 t x 2 log t
Now once you show that (Exercise 1)
Zx √ √
dt x x− x
≤ + √ ,
2 log t log 2 log x
then (7.1) implies (7.2) will follow. We still need to show that (7.2) implies (7.1)
and thus we have to show that
Z x
log x θ(t)dt
lim = 0.
x→∞ x 2 t log2 t
7.3. GETTING CLOSER TO THE PROOF OF THE PRIME NUMBER THEOREM145
π(x) π(x)
l1 = lim inf , L1 = lim sup ,
x→∞ x/logx x→∞ x/logx
θ(x) θ(x)
l2 = lim inf x , L2 = lim sup x ,
x→∞ x→∞
and
ψ(x) ψ(x)
l3 = lim inf , L3 = lim sup ,
x→∞ x x→∞ x
then l1 = l2 = l3 and L1 = L2 = L3
where m ≤ log2 x
Also,
X log x X log x
ψ(x) = log p≤ log p = log xπ(x).
p≤x
log p p≤x
log p
Thus we have
θ(x) ≤ ψ(x) ≤ π(x) log x
146 CHAPTER 7. INTRODUCTION TO ANALYTIC NUMBER THEORY
As a result, we have
θ(x) ψ(x) π(x)
≤ ≤
x x x/ log x
and we get that L2 ≤ L3 ≤ L1 . We still need to prove that L1 ≤ L2 .
Let α be a real number where 0 < α < 1, we have
X X
θ(x) = log p ≥ log p
p≤x
X xα ≤p≤x
> α log x (log p > α log x)
xα ≤p≤x
= αlogx{π(x) − π(xα )}
As a result,
θ(x) απ(x) − αxα−1 log x
>
x x/ log x
Since limx→∞ α log x/x1−α = 0, then
π(x)
L2 ≥ α lim sup
x→∞ x/ log x
As α → 1, we get L2 ≥ L1 .
Proving that l1 = l2 = l3 is left as an exercise.
We now present an inequality due to Chebyshev about π(x).
Proof. Put
π(x) π(x)
l = lim inf , L = lim sup ,
x→∞ x/ log x x→∞ x/ log x
It will be sufficient to prove that L ≤ 4 log 2 and l ≥ log 2. Thus by Theorem
2, we have to prove that
θ(x) ≤ 4 log 2 (7.4)
lim sup
x→∞ x
and
ψ(x) ≥ log 2 (7.5)
lim inf
x→∞ x
To prove (7.4), notice that
(n + 1)(n + 2)...(n + n)
N = C (2n, n) = < 22n < (2n + 1)N
n!
Suppose now that p is a prime such that n < p < 2n and hence p | N . As a
result,
Q
we have N ≥ n<p<2n p. We get
N ≥ θ(2n) − θ(n).
Since N < 22n , we get that θ(2n) − θ(n) < 2n log 2. Put n = 1, 2, 22 , ..., 2m−1
where m is a positive integer. We get that
X
µp
2n n
sp = i
−2 i .
i=11
p p
148 CHAPTER 7. INTRODUCTION TO ANALYTIC NUMBER THEORY
h i
log 2n
where µp = log . Thus we have N = p≤2n psp . If x is a positive integer then
Q p
Hence we get
log N ≤ ψ(2n).
Using the fact that 22n < (2n + 1)N , we can see that
ψ(2n) > 2n log 2 − log(2n + 1).
x
Let x > 2 and put n = 2
≥ 1. Thus 2x − 1 < n < 2 x and we get 2n ≤ x. So we
get
As a result, we get
ψ(x) ≥ log 2.
lim inf
x→∞ x
Exercises
2. Show that Zx √ √
dt x x− x
≤ + √ ,
2 log t log 2 log x
7.3. GETTING CLOSER TO THE PROOF OF THE PRIME NUMBER THEOREM149
3. Show that Zx √ √
dt x x− x
2
≤ 2 + √
2 log t log 2 log2 x
4. Show that
(n + 1)(n + 2)...(n + n)
N = C (2n, n) = < 22n < (2n + 1)N
n!
2n 2n
2√
5. Show that 2 n < N = C (2n, n) < √22n .
Hint: For one side of the inequality, write
This chapter discusses various topics that are of profound interest in number the-
ory. Section 1 on cryptography is on an application of number theory in the field
of message decoding, while the other sections on elliptic curves and the Riemann
zeta function are deeply connected with number theory. The section on Fermat’s
last theorem is related, through Wile’s proof of Fermat’s conjecture on the non-
existence of integer solutions to xn + y n = z n for n > 2, to the field of elliptic
curves (and thus to section 2).
8.1 Cryptography
In this section we discuss some elementary aspects of cryptography, which con-
cerns the coding and decoding of messages. In cryptography, a (word) message
is transformed into a sequence a of integers, by replacing each letter in the
message by a specific and known set of integers that represent this letter, and
thus forming a large integer a by concatenation. Then this integer a is
transformed (i.e. coded) into another integer b by using a congruence of the
form b = ak (mod m) for some chosen k and m, as described below, with k
unknown except to the sender and receiver. b is then sent to the receiver who
decodes it into a again by using
151
152 CHAPTER 8. OTHER TOPICS IN NUMBER
THEORY
¯
a congruence of the form a = bk (mod m), where k¯ is related to k and is
itself only known to the sender and receiver, and then simply transforms the
integers in a back to letters and reveals the message again. In this procedure,
if a third party intercepts the integer b, the chance of transforming this into a,
even if m and the integers that represent the letters of the alphabet are exactly
known, is almost impossible to do (i.e. has a fantastically small probability of
being achieved) if k is not known, that practically the transformed message will
not be revealed except to the intended receiver.
The basic results on congruences to allow for the above procedure are in the
following two lemmata, where φ in the statements is Euler’s φ-function.
Lemma 16. Let a and m be two integers, with m positive and (a, m) = 1. If k
¯
and k¯ are positive integers with kk¯ = 1(mod φ(m)), then akk = a(mod
m).
Before giving the proof, one has to note that the above lemma is in fact an
if-and-only-if statement, i.e. (k, φ(m)) = 1 if and only if r k , r k , · · · , r k forms
a 1 2 n
reduced residue system modulo m. However we only need the if part, as in the
lemma.
8.1. CRYPTOGRAPHY 153
Proof. Assume first that (k, φ(m)) = 1. We show that r1k , 2r k , · · · n, r k is a reduced
residue system modulo m. Assume otherwise, i.e. assume that ∃i, j such that
rki = rkj (mod m), in which case rki and krj would belong to the same class and thus
rk1 , rk2 , · · · ,krn would not form a reduced residue system. Then, since (k, φ(m)) =
1, ∃k¯ with kk¯ = 1(mod φ(m)), and so
¯ ¯
r kki = ri (mod m) and rjkk = rj (mod m) (8.2)
¯ ¯
by the previous lemma. But if rik = rjk (mod m) then (rik )k = (r k k
j ) (mod m), and
¯ ¯
since r kk = r (mod m) and r kk = r (mod m), then = r (mod m) giving that r
r i i j j i j i
1. Transform S into a (large) integer a by replacing each letter and each space
between words by a certain representative integer (e.g. three or four digit
integers for each letter). a is formed by concatenating the representative
integers that are produced.
has a reduced residue system that contains a very large number of integers
of the order of m itself. Hence almost every integer smaller than m, with
a probability of the order 1 − 1/10100 (almost 1), is in a reduced residue
system r1 , r2 , · · · , rφ(m) of m. Thus almost every positive integer
smaller than m is relatively prime with m, with probability of the order 1 −
1/10100 .
3. Now given that almost every positive integer smaller than m is relatively
prime with m, the integer a itself is almost certainly relatively prime with
m, and hence is in a reduced residue system for m. Hence, by lemma 17
above, if k is a (large) integer such that (k, φ(m)) = 1, then ak belongs to
a reduced residue system for m, and there exists a unique positive b
smaller than m with b = ak (mod m).
4. Send b to the destination where φ(m) and k are known. The destination can
determine a k¯ such that kk¯ = 1(mod φ(m)), and then finds the unique c
¯
such that c = bk (mod m). Now since, almost certainly, (a, m) = 1, then
¯ ¯ ¯
almost certainly c = a since c = bk (mod m) = (ak )k (mod m) = akk
(mod m), and which by lemma 16, is given by a(mod m) almost
certainly since (a, m) = 1 almost certainly. Now the destination
translates a back to letters and spaces to reveal the sentence S. Note
that if any third party intercepts b, they almost certainly cannot reveal
the integer a since the chance of them knowing φ(m) = p1 p2 is almost
zero, even if they know m and k. In this case they practically won’t be
able to determine a k¯ with kk¯ = 1(mod φ(m)), to retrieve a and
transform it to S.
in analytic number theory. More generally, one can define similar curves over
arbitrary algebraic fields as follows. Let f (x, y) be a polynomial of any degree
in two variables x and y, with coefficients in an algebraic field F . We define
the algebraic curve Cf (F ) over the field F by
Of course one can also similarly define the algebraic curve Cf (Q) over a field Q,
where Q is either a subfield of the field F where the coefficients of f exist, or is
an extension field of F . Thus if f ∈ F [x, y], and if Q is either an extension or
a subfield of F , then one can define Cf (Q) = {(x, y) ∈ Q × Q : f (x, y) = 0}.
Our main interest in this section will be in third order polynomials (cubic curves)
with coefficients in R, with the associated curves Cf (Q) over the field of rational
numbers Q ⊂ R. Thus, basically, we will be interested in points (x, y) ∈ R2
that have rational coordinates x and y, and called rational points, that satisfy
f (x, y) = 0. Of course one can first imagine the curve f (x, y) = 0 in R2 , i.e.
the curve Cf (R) over R, and then choosing the points on this curve that have ra-
tional coordinates. This can simply be expressed by writing that Cf (Q) ⊂ Cf (R).
It has to be mentioned that ”rational curves” Cf (Q) are related to diophantine
equations. This is in the sense that rational solutions to equations f (x, y) = 0
produce integer solutions to equations f 0 (x, y) = 0, where the polynomial f 0
is very closely related to the polynomial f , if not the same one in many cases.
For example every point in Cf (Q), where f (x, y) = xn + y n , i.e. every rational
solu- tion to f (x, y) = xn + y n = 0, produces an integer solution to xn + y n =
0. Thus algebraic curves Cf (Q) can be of genuine interest in this sense.
In a possible procedure to construct the curve Cf (Q) for a polynomial f (x, y)
∈ R[x, y] with real coefficients, one considers the possibility that, given one
ratio- nal point (x, y) ∈ Cf (Q) ⊂ Cf (R), a straight line with a rational slope m
might
156 CHAPTER 8. OTHER TOPICS IN NUMBER
THEORY
intersect the curve Cf (R) in a point (x0 , y 0 ) that is also in Cf (Q). This possibility
comes from the simple fact that if (x, y), (x0 , y 0 ) ∈ Cf (Q), then the slope of the
straight line that joins (x, y) and (x0 , y 0 ) is a rational number. This technique, of
determining one point in Cf (Q) from another by using straight lines as
mentioned, works very well in some cases of polynomials, especially those of
second degree, and works reasonably well for third order polynomials.
Two aspects of this technique of using straight lines to determine points in
Cf (Q), and which will be needed for defining elliptic curves, are the following.
The first is illustrated by the following example.
Consider the polynomial f (x, y) = y 2 − x2 + y = (y − x + 1)(y + x).
The curve Cf (R) contains the two straight lines y = x − 1 and y = −x. The
point (2, 1) ∈ Cf (Q), and if one tries to find the intersection of the particular line
y = x − 1 that passes through (2, 1) with Cf (R), one finds that this includes the
whole line y = x − 1 itself, and not just one or two other points (for example).
This result is due to the fact that f is a reducible polynomial, i.e. that can be
factored in the form f = f 0 f 00 with f and f 00 not just real numbers.
In this direction one has the following general theorem concerning the
number of intersection points between a straight line L and an algebraic curve Cf
(R):
line, will also pass through a unique third point (x3 , y3 ). By the above theorem,
if a line intersects the curve Cf (R) associated with the third order polynomial f
in more than three points, then the line itself is a subset of Cf (R). This will be
excluded for the kind of third degree polynomials f whose associated algebraic
curves shall be called elliptic curves.
One other thing to be excluded, to have third order curves characterized as
elliptic curves, is the existence of singular points on the curve, where a singular
point is one where the curve does not admit a unique tangent.
It has to be mentioned that in the previous discussion, the points on the curve
Cf (R) may lie at infinity. To deal with this situation we assume that the curve is
in fact a curve in the real projective plane P2 (R). We now can define an elliptic
curve Cf (R) as being such that f (x, y) is an irreducible third order polynomial
with Cf (R) having no singular points in P2 (R).
The main idea behind the above definition for elliptic curves is to have a
curve whereby any two points A and B on the curve can determine a unique
third point, to be denoted by AB, using a straight line joining A and B. The
possibilities are as follows: If the line joining A and B is not tangent to the
curve Cf (R) at any point, then the line intersects the curve in exactly three
different points two of which are A and B while the third is AB. If the line
joining A and B is tangent to the curve at some point p then either this line
intersects Cf (R) in exactly two
points, p and some other point p0 , or intersects the curve in only one point p. If
the
line intersects Cf (R) in two points p and p0 , then either p = A = B in which
case AB = p0 , or A = B in which case (irrespective of whether p = A and p0 =
B or vice-versa) one would have p = AB. While if the line intersects Cf (R) in
only one point p then p = A = B = AB.
The above discussion establishes a binary operation on elliptic curves that
pro- duces, for any two points A and B a uniquely defined third point AB. This
binary operation in turn produces, as will be described next, another binary
operation, denoted by +, that defines a group structure on Cf (R) that is
associated with the
158 CHAPTER 8. OTHER TOPICS IN NUMBER
THEORY
A + B = 0(AB), (8.5)
meaning that we first determine the point AB as above, then we determine the
point 0(AB) corresponding to 0 and AB. Irrespective of the choice of the point
0, one has the following theorem on a group structure determined by + on Cf
(R).
Theorem 90. Let Cf (R) be an elliptic curve, and let 0 be any point on Cf (R).
Then the above binary operation + defines an Abelian group structure on Cf (R),
with 0 being the identity element and −A = A(00) for every point A.
The proof is very lengthy and can be found in [18]. We first note that if 0 and
0
0 are two different points on an elliptic curve with associated binary operations
+ and +0 , then one can easily show that for any two points A and
B
A +0 B = A + B − 00 . (8.6)
This shows that the various group structures that can be defined on an elliptic
curve
by considering all possible points 0 and associated operations +, are essentially
the same, up to a ”translation”.
Lemma 18. Consider the group structure on an elliptic curve Cf (R), correspond-
ing to an operation + with identity element 0. If the cubic polynomial f has
rational coefficients, then the subset Cf (Q) ⊂ Cf (R) of rational solutions to
f (x, y) = 0 forms a subgroup of Cf (R) if and only if 0 is itself a rational point
(i.e. a rational solution).
Proof. If Cf (Q) is a subgroup of Cf (R), then it must contain the identity 0, and
thus 0 would be a rational point. Conversely, assume that 0 is a rational point.
8.2. ELLIPTIC CURVES 159
First, since f has rational coefficients, then for any two rational points A and B
in Cf (Q) one must have that AB is also rational, and thus (since 0 is assumed
rational) that 0(AB) is rational, making A + B = 0(AB) rational. Thus Cf (Q)
would be closed under +. Moreover, since for every A ∈ Cf (Q) one has that
−A = A(00), then −A is also rational, which makes Cf (Q) closed under inver-
sion. Hence Cf (Q) is a subgroup.
Thus by lemma 18, the set of all rational points on an elliptic curve form a
subgroup of the group determined by the curve and a point 0, if and only if
the identity element 0 is itself a rational point. In other words, one finds that if
the elliptic curve Cf (R) contains one rational point p, then there exists a group
structure on Cf (R), with 0 = p and the corresponding binary operation +, such
that the set Cf (Q) of all rational points on Cf (R) is a group.
One thing to note about rational solutions to general polynomial functions
f (x, y), is that they correspond to integer solution to a corresponding homoge-
neous polynomial h(X, Y, Z ) in three variables, and vice-verse, where homoge-
neous practically means that this function is a linear sum of terms each of which
has the same power when adding the powers of the variables involved in this
term. For example X Y 2 − 2X 3 + X Y Z + Z 3 is homogeneous.
In fact a rational solution x = a/b and y = c/d for f (x, y) = 0, where
a, b, c, d are integers, can first be written as x = ad/bd and y = cb/bd, and thus
one can always have this solution in the form x = X/Z and y = Y /Z ,
where X = ad, Y = cb and Z = bd. If x = X/Z and y = Y /Z are
replaced in f (x, y) = 0, one obtains a new version h(X, Y, Z ) = 0 of this
equation written in terms of the new variables X, Y, Z . One can immediately
see that this new polynomial function h(X, Y, Z ) is homogeneous in X, Y, Z .
The homogeneous function h(X, Y, Z ) in X, Y, Z is the form that f (x, y) takes
in projective space, where in this case the transformations x = X/Z and y = Y
/Z define the projec- tive transformation that take f (x, y) to h(X, Y, Z ).
If we now go back to cubic equation f (x, y) = 0, one can transform this
160 CHAPTER 8. OTHER TOPICS IN NUMBER THEORY
h(X, Y, Z ) = aX 3 + bX 2 Y + cX Y 2 + dY 3 + eX 2 Z
+ f X Y Z + gY 2 Z + hX Z 2 + iY Z 2 + jZ 3 ,
(8.7)
h(X, Y, Z ) = cX Y 2 + eX 2 Z + f X Y Z + hX Z 2 + iY Z 2 + jZ 3 .
(8.8)
Which, by using the projective transformation again, and using new coefficients,
gives that points on the curve Cf (R) are precisely those on the curve Ch (R),
where
i.e. that Cf (R) = Cg (R). The equation g(x, y) = 0, where g is given in (8.10),
is said to be the Weierstrass normal form of the equation f (x, y) = 0. Thus, in
particular, any elliptic curve defined by a cubic f , is birationally equivalent to
an elliptic curve defined by a polynomial g(x, y) as above. Birational
equivalence between curves is defined here as being a rational transformation,
together with its inverse transformation, that takes the points on one curve to
another, and vice- versa.
8.3. THE RIEMANN ZETA FUNCTION 161
The Riemann zeta function ζ (z) is an analytic function that is a very important
function in analytic number theory. It is (initially) defined in some domain in the
complex plane by the special type of Dirichlet series given by
X∞
1
ζ (z) = z
, (8.11)
n=1
n
where Re(z) > 1. It can be readily verified that the given series converges
locally uniformly, and thus that ζ (z) is indeed analytic in the domain in the
complex plane C defined by Re(z) > 1, and that this function does not have a
zero in this domain.
We first prove the following result which is called the Euler Product Formula.
Theorem 91. ζ (z), as defined by the series above, can be written in the
form
Y
∞
1
ζ (z) = , (8.12)
n=1 1 1
− pzn
1 X
∞
xk ,
1− x = (8.13)
k=0
1 X
∞
1
= , (8.14)
1 1pnz pnkz
k=0
−
162 CHAPTER 8. OTHER TOPICS IN NUMBER THEORY
since every |1/pz | < 1 if Re(z) > 1. This gives that for any integer N
n
Y
N
1
N
Y 1
1
= 1+ + 2z + · · ·
n=1 1 1
pn pn
z
n=1
− pzn
X 1
= k z
(8.15)
pkn11z · · · pni
j
X 1
=
nz
where i ranges over 1, · · · , N , and j ranges from 0 to ∞, and thus the integers
n in the third line above range over all integers whose prime number
factorization consist of a product of powers of the primes p1 = 2, · · · , pN . Also
note that each such integer n appears only once in the sum above.
Now since the series in the definition of ζ (z) converges absolutely and the
order of the terms in the sum does not matter for the limit, and since, eventu-
ally, every integer n appears on the right hand side of 8.15 as N −→ ∞, then
P
limN →∞ z N = ζ (z). Moreover, limN →∞ n=1 exists, and the re-
1
Q N 1
n 1− p1z
sult follows. n
The Riemann zeta function ζ (z) as defined through the special Dirichlet
series above, can be continued analytically to an analytic function through out
the com- plex plane C except to the point z = 1, where the continued function
has a pole of order 1. Thus the continuation of ζ (z) produces a meromorphic
function in C with a simple pole at 1. The following theorem gives this result.
Given this continuation of ζ (z), and also given the functional equation that
is satisfied by this continued function, and which is
πz
ζ (z) = 2z π z−1 sin 2 Γ(1 − z)ζ (1 − z), (8.16)
8.3. THE RIEMANN ZETA FUNCTION 163
(see a proof in [3]), where Γ is the complex gamma function, one can deduce that
the continued ζ (z) has zeros at the points z = −2, −4, −6, · · · on the
negative real axis. This follows as such: The complex gamma function Γ(z)
has poles at the points z = −1, −2, −3, · · · on the negative real line, and thus
Γ(1 − z) must have poles at z = 2, 3, · · · on the positive real axis. And since ζ
πz
(z) is analytic at these points, then it must be2 that either sin or ζ (1 − z)
must have zeros at the points z = 2, 3, · · · to cancel out the poles of Γ(1 − z),
πz
and thus make ζ (z) analytic at these points.
2
And since sin has zeros at z
= 2, 4, · · · , but not at z = 3, 5, · · · , then it must be that ζ (1 − z) has zeros at
z = 3, 5, · · · . This gives that ζ (z) has zeros at z = −2, −4, −6 · · · .
It also follows from the above functional equation, and from the above men-
tioned fact that ζ (z) has no zeros in the domain where Re(z) > 1, that these
zeros at z = −2, −4, −6 · · · of ζ (z) are the only zeros that have real parts
either less that 0, or greater than 1. It was conjectured by Riemann, The
Riemann Hypothe- sis, that every other zero of ζ (z) in the remaining strip 0 ≤
Re(z) ≤ 1, all exist on the vertical line Re(z) = 1/2. This hypothesis was
checked for zeros in this strip with very large modulus, but remains without a
general proof. It is thought that the consequence of the Riemann hypothesis on
number theory, provided it turns out to be true, is immense.
164 CHAPTER 8. OTHER TOPICS IN NUMBER THEORY
Bibliography
165
166 BIBLIOGRAPHY
[8] H.M. Edwards, Riemann’s Zeta Function, Dover, New York, 2001.
[9] E. Grosswald, Topics from the Theory of Numbers. New York: The
Macmil- lan Co. (1966).
[10] G.H. Hardy and E.M. Wright, An Introduction to the Theory of Numbers,
5th ed. Oxford University Press, Oxford, 1979.
[15] W.J. Leveque, Elementary Theory of Numbers, Dover, New York, 1990.
BIBLIOGRAPHY 167
[16] W.J. Leveque, Fundamentals of Number Theory, Dover, New York, 1996.
[21] Kenneth H. Rosen, Elementary Number Theory and its Applications. Fifth
Edition. Pearson, Addison Wesley, USA, 2005.
168 BIBLIOGRAPHY
Index
169
170 INDEX