41
MA3A6 ALGEBRAIC NUMBER THEORY SAMIR SIKSEK Abstract. This is an incomplete set of lecture notes for for Algebraic Number Theory. I do not know yet if it will be completed, so you are advised to continue taking notes in the lectures. Please send comments, misprints and corrections to [email protected] Contents 1. Orientation 2 1.1. Local Arguments 2 1.2. Infinite Descent 2 1.3. Descent 3 2. Beginnings of Algebraic Number Theory 4 3. Preliminaries on Rings and Ideals 5 3.1. Rings 5 3.2. Ideals 6 3.3. Quotient Rings 6 3.4. Units 6 3.5. Fields 7 3.6. Fields of Fractions 8 3.7. Prime and Maximal Ideals 8 3.8. Unique Factorization Domains 8 4. Algebraic Numbers and Algebraic Integers 9 5. Minimal Polynomials 10 6. Conjugates 12 7. Factorization of Polynomials 13 8. Eisenstein’s Criterion For Irreducibility 14 9. Symmetric Polynomials 15 10. Algebraic Integers form a Ring 17 11. Algebraic Numbers form a Field 19 12. Number Fields 19 13. Fields Generated by Conjugate Elements 21 14. Embeddings 22 15. Field Polynomial 23 16. Ring of Integers 25 17. Determinants and Discriminants 28 18. Ideals 30 18.1. Quotient Rings 31 Date : September 26, 2006. 1

MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

  • Upload
    others

  • View
    9

  • Download
    1

Embed Size (px)

Citation preview

Page 1: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

MA3A6ALGEBRAIC NUMBER THEORY

SAMIR SIKSEK

Abstract. This is an incomplete set of lecture notes for for Algebraic Number

Theory. I do not know yet if it will be completed, so you are advised to continuetaking notes in the lectures. Please send comments, misprints and corrections

to [email protected]

Contents

1. Orientation 21.1. Local Arguments 21.2. Infinite Descent 21.3. Descent 32. Beginnings of Algebraic Number Theory 43. Preliminaries on Rings and Ideals 53.1. Rings 53.2. Ideals 63.3. Quotient Rings 63.4. Units 63.5. Fields 73.6. Fields of Fractions 83.7. Prime and Maximal Ideals 83.8. Unique Factorization Domains 84. Algebraic Numbers and Algebraic Integers 95. Minimal Polynomials 106. Conjugates 127. Factorization of Polynomials 138. Eisenstein’s Criterion For Irreducibility 149. Symmetric Polynomials 1510. Algebraic Integers form a Ring 1711. Algebraic Numbers form a Field 1912. Number Fields 1913. Fields Generated by Conjugate Elements 2114. Embeddings 2215. Field Polynomial 2316. Ring of Integers 2517. Determinants and Discriminants 2818. Ideals 3018.1. Quotient Rings 31

Date: September 26, 2006.

1

Page 2: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

2 SAMIR SIKSEK

19. Prime and Maximal Ideals 3420. Towards Unique Factorization for Ideals I 3521. Towards Unique Factorization for Ideals II 3622. Unique Factorization Proof—A Summary so Far 3723. A Special Case of the Cancellation Lemma 3724. Ideal Classes 3825. Unique Factorization Proof—Summary So Far (again) 4026. What are the prime ideals of K? 40

1. Orientation

Algebraic number theory arose out of the study of Diophantine equations. ADiophantine equation is a polynomial equation in several variables with integercoefficients and one desires the solutions in integers. The study of Diophantineequations seems as old as human civilization itself; they are however named afterDiophantus of Alexandria who lived in the 3rd century BC. Soon we will give ourmotivating example for algebraic number theory. Before that we look at threemethods for attacking Diophantine equations which should be close to the heart ofevery student of number theory: Local Methods, Infinite Descent, Descent.

1.1. Local Arguments. This means that we look at the equation modulo somepositive integer m and try to get information about integral solutions.

Example 1.1. Show that the equation

x2 + 1 = 4y2

does not have solutions in integers.Answer: Suppose x, y is an integer solution. Reducing the equation modulo 4 weget

x2 + 1 ≡ 0 (mod 4).However, the squares modulo 4 are x2 ≡ 0 or 1 (mod 4). So

x2 + 1 ≡ 1 or 2 (mod 4),

giving a contradiction.

Included in ‘local arguments’ is the idea that if an equation does not have realsolutions then it does not have integeral solutions. For example, the equationx2 + 5 = −2z2 does not have integral solutions because it does not have realsolutions.

1.2. Infinite Descent. ‘infinite descent’ is used to show that equations do not havesolutions, or that they do not have non-trivial solutions. The idea is to supposethat a Diophantine equation has solutions, take the smallest one and show thatthere must be a smaller one, giving a contradiction.

Example 1.2. Show that the equation X2 = 2Y 2 does not have non-trivial solu-tions in integers.

You will recognize this as essentially the proof that√

2 is irrational. Supposethat we have non-trivial solutions. Let (x, y) be a non-trivial solution with thevalue of |x| minimal. Since x2 = 2y2 we get that x is even. So x = 2x1 for some

Page 3: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

ALGEBRAIC NUMBER THEORY 3

x1 ∈ Z. So 2x21 = y2 and we show that y = 2y1. Now (x1, y1) is a non-trivial

solution to X2 = 2Y 2 and |x1| < |x|, giving a contradiction.The name ‘infinite descent’comes from the original way in which the method

is applied. We start with a non-trivial solution (x, y) and construct another one(x1, y1), and then from (x1, y1) we get another one (x2, y2) and so on. We get aninfinite list of non-trivial solutions satisfying

|x| > |x1| > |x2| > · · · > 0;

but we cannot squeeze infinitely many integers between |x| and 0 and we have acontradiction.

Here is another example due to Euler.

Example 1.3. Show that the equation

(1) X2 + 2Y 3 + 4Z3 = 0

has no non-trivial solutions.Answer: Suppose it has non-trivial solutions. Let (x, y, z) be a non-trivial solutionwith the value of |x| minimal. We see that x3 is even and so x is even. Writex = 2x1. Then

4x31 + y3 + 2z3 = 0.

Thus y = 2y1 and applying the same trick again z = 2z1. We see that (x1, y1, z1)is a non-trivial solution to equation (1) with |x1| < |x| giving a contradiction.

In both the above examples, we could have obtained a contradition by explainingthat we may assume that x, y are coprime. Then the argument shows that x, y arenot coprime. This is not always the case with infinite descent examples.

1.3. Descent. Descent is not the same as infinite descent. It is a technique inventedby Fermat. It uses variants of the following very simple Lemma:

Lemma 1.1. Suppose that U , V , W are non-zero integers, with U , V coprime.Suppose also that UV = Wn where n is a positive integer. Then U = ±Wn

1 andV = ±Wn

2 for some integers W1, W2. Moreover, if n is odd, we can take U = Wn1

and V = Wn2 .

Proof. Let p1, . . . pr be the distinct prime divisors of U and q1, . . . , qs the distinctprime divisors of V . Since U , V are coprime, these lists are disjoint (this is a veryimportant point). Notice that

p1, . . . , pr, q1, . . . , qs

are the distinct prime divisors of W . By the Fundamental Theorem of Arithmetic,we can write

U = ±pa11 . . . par

r , V = ±qb11 . . . qbs

s , W = ±pc11 . . . pcr

r qd11 . . . qds

r .

Substituting in UV = Wn we get

pa11 . . . par

r qb11 . . . qbs

s = ±pnc11 . . . pncr

r qnd11 . . . qnds

r .

From the uniqueness of factorization we deduce that

ai = nci, bj = ndj ;

it is here that we need the fact that the ps and qs are distinct. Hence

U = ± (pc11 . . . pcr

r )n, V = ±

(qd11 . . . qds

s

)n

.

Page 4: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

4 SAMIR SIKSEK

This completes the proof for n even. For n odd, simply absorb the ± inside then-th power. �

We are ready to give a first example of descent.

Example 1.4. Show that the equation

4X2 − 1 = Y 3

has no solutions in integers apart from (X,Y ) = (0,−1).Answer: First notice that if (X,Y ) is a solution then so is (−X,Y ). So we maysuppose that we have a solution with X ≥ 0. Write

(2X + 1)(2X − 1) = Y 3.

Let d be the gcd of 2X+1 and 2X−1. Since d divides them both, it must dividetheir difference; thus d | 2 and so d = 1 or d = 2. However, d divides 2X + 1 whichis odd, so d is odd and so d = 1. In other words, 2X + 1 and 2X − 1 are coprime.Applying Lemma 1.1 we see that

(2) 2X + 1 = Y 31 , 2X − 1 = Y 3

2 ,

where Y = Y1Y2 (this step is called ‘descent’). Subtracting we get

2 = Y 31 − Y 3

2 = (Y1 − Y2)(Y 21 + Y1Y2 + Y 2

2 ).

Recall our assumption that X ≥ 0. Hence 2X + 1 > 2X − 1 which gives Y1 > Y2;i.e. Y1 − Y2 > 0. We deduce that either

Y1 − Y2 = 1, Y 21 + Y1Y2 + Y 2

2 = 2,

orY1 − Y2 = 2, Y 2

1 + Y1Y2 + Y 22 = 1.

But, from (2) we know that Y1, Y2 are both odd, and so Y1 − Y2 is even. Hencewe have only to deal with the last case: Y1 − Y2 = 2. Write Y1 = Y2 + 2 andsubstitute in to Y 2

1 + Y1Y2 + Y 22 = 1. We get Y 2

2 + 2Y2 + 1 = 0. Thus Y2 = −1 andY1 = Y2 + 2 = 1. Hence Y = Y1Y2 = −1 and we see that X = 0 as required.

2. Beginnings of Algebraic Number Theory

The following example is the beginning of algebraic number theory. Fermatclaimed that he has shown that the only solutions to the equation

X2 + 2 = Y 3

is (X,Y ) = (±5, 3). Euler ‘proved’Fermat’s assertion in 1770. The ‘proof’ mimicsthe standard factorization or descent argument used in the above example. Wesay ‘proof’ in quotes because what Euler wrote down was not rigorous, but can bemade rigorous. Let us see Eulers argument: simply factor the left-hand side to get:

(X +√−2)(X −

√−2) = Y 3.

We leave the usual integers Z and work with Z[√−2]. Now the ‘integers’ X+

√−2,

X −√−2 are ‘coprime’and so each must be a cube. So

X +√−2 = (a+ b

√−2)3,

where a, b are in Z. Once we get past the dodgy step above, the rest of the argumentis respectable. Expand the brackets to get

X +√−2 = (a3 − 6ab2) + (3a2b− 2b3)

√−2.

Page 5: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

ALGEBRAIC NUMBER THEORY 5

Comparing the coefficients of√−2 we get

X = a3 − 6ab2, 1 = 3a2b− 2b3.

Hence b | 1 and so b = ±1 which gives 3a2 − 2 = ±1. Therefore a = ±1 and we getX = a3 − 6ab2 = ±5. Hence (X,Y ) = (±5, 3) as required.

Is this argument respectable? It turns out that it is because Z[√−2] is a unique

factorization domain, and so we have an analogue of the Fundamental Theorem ofArithmetic and can prove the needed analogue of Lemma 1.1.

This and other Diophantine equations have lead us to reconsider what integersare. In Q we have rational (or usual) integers Z. But in other fields such as Q(

√d)

we have an extension of the concept of integer. Unfortunately unique factorisationdoes not always hold. For example, in Q(

√−5), the ‘integers’is the ring Z[

√−5].

Here we do not have unique factorization (i.e. there is no analogue of the Funda-mental Theorem of Arithmetic); for example, 6 can be factorized as a product ofirreducibles in two different ways,

6 = 2× 3 = (1 +√−5)(1−

√−5).

Thus the analogue of our Lemma 1.1 does not hold for this ring.In this course we cover the following ideas:

(1) The correct generalization of the concept of integer.(2) Whilst uniqueness of factorization fails for elements (as above), it holds for

ideals; every ideal can be expressed as a product of powers of distinct primeideals in a unique way.

(3) Minkowski’s Theorem. Essentially this tells us that whilst unique factor-ization fails, it is not by too much.

(4) Dirichelet’s Unit Theorem.

3. Preliminaries on Rings and Ideals

We begin by revising some ideas that you have met in previous algebra courses.

3.1. Rings. By a ring we shall always mean a commutative ring with a unit element1. Examples of rings that you are familiar with are Q, Z, Z[i] (the Gaussianintegers), Q[x], Z[x]. Another important ring is Z/nZ (the integers modulo n).You probably called this ring Zn in your earlier courses, but we will stick with thenotatin Z/nZ.

Definition. Let R be a ring. A non-zero element x ∈ R is called a zero-divisorif there is some other non-zero element y such that xy = 0. A ring that does nothave zero-divisors is called an integeral domain.

Example 3.1. Z, Q[x] do not have zero-divisors and are therefore integral domains.

Example 3.2. In the ring Z/6Z (integers modulo 6) the elements 2 and 3 arezero-divisors, because 2 × 3 ≡ 0 (mod 6) but 2 6≡ 0 (mod 6) and 3 6≡ 0 (mod 6).Thus Z/6Z is not an integral domain.

Exercise 3.3. Suppose n > 1 is an integer. Show that Z/nZ is an integral domainif and only if n is prime.

Page 6: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

6 SAMIR SIKSEK

3.2. Ideals.

Definition. Let R be a ring. A non-empty subset I is an ideal if• a− b ∈ I for every a, b ∈ I (this really says that I is an additive subgroup

of R);• if a ∈ I and r ∈ R then ra ∈ I.

Example 3.4. R is an ideal of R. Every other ideal of R is called a proper ideal.

Example 3.5. Suppose R is an ring and a ∈ R. We define

aR = {ar : r ∈ R} .It is easy to show that aR is an ideal of R (just check the definition). We call aRthe principal ideal generated by a. Another common notation for aR is (a).

The ideal (0) = {0} is called the zero ideal.

3.3. Quotient Rings. Let I be an ideal of the ring R. A coset of I is of the form

x+ I = {x+ a : a ∈ I}.Recall that two cosets are equal x+ I = y + I if and only if x− y = I. We definethe quotient

R/I = {x+ I : x ∈ R}.A priori R/I is just the set of cosets of R, but we can make it into a ring by definingaddition and multiplication as follows:

(x+ I) + (y + I) = (x+ y) + I, (x+ I)(y + I) = xy + I.

Exercise 3.6. Prove that these operations are well-defined and that they do giveus a ring structure on R/I.

3.4. Units.

Definition. An element u of a ring R is called a unit (or an invertible element)if there is some other element v ∈ R such that uv = 1. If R is an integral domain,then v is unique and we call it the inverse of u and write v = u−1.

The set of units in R is denoted by R∗ or U(R).

Example 3.7. The units of Z are ±1. Thus Z∗ = {1,−1}.

Example 3.8. The units of Z[i] = {a+ bi : a, b ∈ Z} are ±1, ±i. It is clear thatthese are units. Let us prove that they are the only ones. Suppose u ∈ Z[i] is aunit and let v = u−1. Then u = a+ bi and v = c+di for some integers a,. . . ,d suchthat

(a+ bi)(c+ di) = 1.Conjugating we get

(a− bi)(c− di) = 1.Multiplying the last two equalities

(a2 + b2)(c2 + d2) = 1.

Now noting that a2 + b2, c2 + d2 are in Z and non-negative we deduce

a2 + b2 = c2 + d2 = 1.

Hence (a, b) = (±1, 0) or (0,±1) giving the u = ±1 or ±i.

Page 7: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

ALGEBRAIC NUMBER THEORY 7

Exercise 3.9. Determine the units of Z[√−2] = {a+ b

√−2 : a, b ∈ Z} and of

Z[1 +

√−3

2

]={a+ b

(1 +

√−3

2

): a, b ∈ Z

}.

Exercise 3.10. If R is a ring then R∗ is a group under multiplication (it is calledthe unit group of R).

Example 3.11. Note that (√

2 + 1)(√

2 − 1) = 1. Thus√

2 + 1 is a unit in thering Z[

√2]. Since the units form a group under multiplication, we see immediately

that (√

2 + 1)n is a unit for all integers n. In fact, it can be shown that

Z[√

2]∗ = {±(√

2 + 1)n : n ∈ Z}.

Dirichelet’s Units Theorem, which we will hopefully meet at the end of thiscourse, describes unit groups R∗ for certain rings called rings of integers of numberfields.

Exercise 3.12. If R is a ring then R[x]∗ = R∗.

3.5. Fields.

Definition. A field is a non-zero ring where every non-zero element is a unit.

Example 3.13. Q, R and C are fields.

Theorem 1. Every finite integral domain is a field.

Proof. Suppose R is a finite integral domain. We want to show that every non-zeroelement is invertible. Suppose x is a non-zero element, and consider the map

φx : R→ R, φx(a) = x a.

We want to show first that φx is one-to-one. So suppose that a, b ∈ R and φx(a) =φx(b). Thus x a = x b, or in other words

x (a− b) = 0.

But R is an integral domain, and so x is not a zero-divisor. Hence a− b = 0, whichgives a = b showing indeed that φx is one-to-one.

Here is where we use the fact that R is finite: recall that a one-to-one map froma finite set to itself must be onto. Hence φx is onto. In particular, there is someelement y ∈ R such that φx(y) = 1. We can re-write this as xy = 1, clearly showingthat x is invertible.

This show that R is a field. �

Example 3.14. Suppose p is a prime. We denote Z/pZ (the integers modulo p)by Fp. We know from Exercise 3.3 that Fp is an integral domain. Since Fp is finite,Theorem 1 tells us that it is a field.

Note that the proof of Theorem 1 is non-constructive. It shows that if x is anon-zero element then it has an inverse, but it does not tell us how to find it. ForFp we can actually give a constructive proof. Suppose x is an integer satisfyingx 6≡ 0 (mod p). This means that p - x and (as p is prime), the integers x and p arecoprime. By Euclid’s algorithm we know that there are integers y, z such that

y x+ z p = 1.

Recuding modulo p we gety x ≡ 1 (mod p),

Page 8: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

8 SAMIR SIKSEK

showing that x is invertible. Euclid’s algorithm is actually a recipe that will writedown for us y (and z). Thus we can calculate the inverse of any non-zero element.

Exercise 3.15. Use Euclid’s algorithm to find the inverse of 14 in F101.

3.6. Fields of Fractions.

Definition. Let R be an integral domain. A field K is said to be the field offractions if

• K contains R as a subring;• every element α ∈ K is expressible as a/b for some a, b ∈ R and b 6= 0.

Example 3.16. Q is the field of fractions of Z. Q(x) is the field of fractions ofQ[x] and of Z[x].

Theorem 2. Every integral domain has a field of fractions (that is unique up toisomorphism).

3.7. Prime and Maximal Ideals.

Definition. Let R be a ring. An ideal ℘ of R is said to be a prime ideal if, ab ∈ ℘implies a ∈ ℘ or b ∈ ℘ for all a, b ∈ R.

An ideal m of R is said to be maximal if it is a proper ideal and is not containedin any other proper ideal.

Exercise 3.17. Show that the zero ideal (0) is prime if and only R is an integraldomain.

Exercise 3.18. Suppose p, q are distinct prime numbers. Let R = Z/pqZ (theintegers modulo pq). Show that pR and qR are prime ideals.

The proof of the following theorem is an easy exercise.

Theorem 3. Let R be a ring. An ideal ℘ is a prime ideal if and only if R/℘ is anintegral domain. An ideal m is maximal if and only if R/m is a field.

3.8. Unique Factorization Domains.

Definition. Let R be a ring. A non-zero element x is called irreducible if x is nota unit and whenever x = ab with a, b ∈ R, one of a, b must be a unit.

Example 3.19. In Z the irreducible elements are of the form ±p where p is prime.The composite elements are of the form ±n where n is composite.

Example 3.20. 2 is irreducible in Z but not in Z[i], since 2 = (1 + i)(1 − i) andneither (1 + i) nor (1− i) is a unit.

Definition. Two elements x, y ∈ R are called associates (written x ∼ y) if thereis a unit u ∈ R such that x = uy.

Definition. A ring R is a unique factorization domain if it satisfies the followingthree conditions:

• R is an integral domain;• Every non-zero, non-unit x ∈ R can be written as a product x = q1 . . . qr of

irreducible elements;• The decomposition of x into irreducibles is unique up to units and permu-

tation of factors. This means that if x = q′1 . . . q′s is another factorization

into irreducibles then r = s and after possibly relabeling we have qi ∼ q′i forall i = 1, . . . , r.

Page 9: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

ALGEBRAIC NUMBER THEORY 9

4. Algebraic Numbers and Algebraic Integers

We now introduce the main objects of study of algebraic number theory.

Definition. Let α ∈ C. We say that α is an algebraic number if there is some non-zero polynomial f(x) ∈ Q[x] (i.e. it has rational coefficients) such that f(α) = 0.

We say that α ∈ C is an algebraic integer if there is some monic polynomialf(x) ∈ Z[x] (i.e. with integral coefficients) such that f(α) = 0.

Clearly every algebraic integer is an algebraic number. We will see that theconverse need not hold. We denote the set of algebraic numbers by Q and the setof algebraic integers by O. Thus

O ⊂ Q ⊂ C.

Lemma 4.1. Z ⊂ O and Q ⊂ Q.

Proof. If α ∈ Z then α is the root of X − α which is a monic polynomial in Z[X],and so α ∈ O. Thus Z ⊂ O, and similarly Q ⊂ Q. �

Example 4.1.√−2 is an algebraic integer because it is a root of the monic poly-

nomial with integral coefficients X2 + 2.Let α =

√2 +

√3. We will show that α is also an algebraic integer. Note

α2 = 5 + 2√

6.

Hence(α2 − 5) = 24.

In other words, α =√

2 +√

3 is a root of

f(X) = (X2 − 5)2 − 24 = X4 − 10X2 + 1.

Since f(X) is monic with integral coefficients, it follows that α is an algebraicinteger.

Example 4.2. Not every complex number is algebraic. Complex numbers thatare not algebraic numbers are called transcendental. Examples of transcendentalnumbers are e and π. We shall not prove this as we do not need it. For proofs, seeStewart’s Galois Theory.

The elements of Z are called the rational integers. The reason is found in thefollowing theorem which says that any algebraic integer which is also a rationalnumber must belong to Z.

Theorem 4. (Gauss) O ∩Q = Z.

Proof. We know that Z ⊂ Q and Z ⊂ O so Z ⊂ O ∩Q. It is enough to prove thatO ∩ Q ⊂ Z. Suppose α ∈ O ∩ Q; we want to show that α ∈ Z. Thus there somepolynomial f(X), monic with coefficients in Z such that f(α) = 0. Write

f(X) = Xn + cn−1Xn−1 + · · ·+ c0, ci ∈ Z.

Since α is rational, we may write α = a/b where a, b are integers, b > 0 andgcd(a, b) = 1. Substituting we get(a

b

)n

+ cn−1

(ab

)n−1

+ · · · c0 = f(α) = 0.

Page 10: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

10 SAMIR SIKSEK

Multiplying by bn we obtain

an + cn−1an−1b+ · · ·+ c0b

n = 0.

Rearranging we getb(−cn−1b

n−2 − · · · − c0)︸ ︷︷ ︸

in Z

= an.

We deduce that b | an. Recall b > 0. We want to show that α = a/b is in Z andfor this it is enought to show that b = 1. Suppose that b 6= 1 and we will derivea contradiction. Then some prime p must divide b. So p | an which implies p | a.This contradicts the assumption that gcd(a, b) = 1. Hence b = 1 and α ∈ Z asrequired. �

Example 4.3. All the examples of algebraic numbers that we have seen so far havebeen algebraic integers. Thanks to the above theorem, we can now give examplesof algebraic numbers that are not algebraic integers: 1/2, −3/4, etc. Any rationalnumber that is not an integer is an algebraic number that is not an algebraic integer.

Exercise 4.4. Show that the following are algebraic numbers:√

3,√

3 +√−3,

e2πi/7, cos(2πi/3).

Eventually we will show that Q is a subfield of C and that O is a subring of C.For this we need symmetric polynomials which we will cover soon. In the meantimewe content with proving the following Lemma.

Lemma 4.2. Suppose α ∈ Q and α 6= 0. Then α−1 ∈ Q.

Proof. Suppose α ∈ Q and α 6= 0. Then α is a root of some non-zero polynomialf(X) with rational coefficients. Write

f(X) = cnXn + cn−1X

n−1 + · · ·+ c0.

Thencnα

n + cn−1αn−1 + · · ·+ c0 = 0.

Dividing by αn we get

cn + cn−1α−1 + · · ·+ c0α

−n = 0.

Hence α−1 is the root of the non-zero polynomial

g(X) = c0Xn + · · ·+ cn−1X + cn

which has rational coefficients; implying α−1 ∈ Q. �

5. Minimal Polynomials

We recall the definition of an algebraic number: α ∈ C is said to be algebraic ifthere is some non-zero polynomial f ∈ Q[x] such that f(α) = 0. For any algebraicα there are of course infinitely many such f . For example, if α = i, we may take fto be any of

x2 + 1, (x2 + 1)(x− 7), (x2 + 1)3, . . . .

It turns out that the ‘best’ choice for f is the one with least degree. By ‘best’ herewe mean the choice which most closely reflects the properties of α. To such an fwe give a special name:

Page 11: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

ALGEBRAIC NUMBER THEORY 11

Definition. Suppose α ∈ C is an algebraic number. We define the minimalpolynomial of α, which we denote by fα to be the monic polynomial with rationalcoefficients and least possible degree satisfying fα(α) = 0.

As with any other definition in Mathematics, we must be concerned with ex-istence and uniqueness. If α is an algebraic number, there is by definition somenon-zero polynomial f with rational coefficients satisfying f(α) = 0. We can makef monic simply by dividing f by the leading coefficient. Out of all such polyno-mials we take fα to be the one of least possible degree, which will be the minimalpolynomial. Of course here we run into the problem of uniqueness, for there maybe two or more such polynomials of minimal degree. We prove below that minimalpolynomials are unique, so until then read ‘let fα be a minimal polynomial for α’when you see the phrase ‘let fα be the minimal polynomial for α’.

Lemma 5.1. Let α be an algebraic number and fα ∈ Q[x] be its minimal polyno-mial. Then

(i) fα is irreducible 1;(ii) if g ∈ Q[x] satisfies g(α) = 0 then g | fα.

Proof. For (i) suppose otherwise. Then fα(x) = g(x)h(x) where g, h are polynomi-als with rational coefficients and deg(g), deg(h) < deg(fα); as fα is monic we cansuppose that g and h are monic (check this). Since fα(α) = 0 we see that eitherg(α) = 0 or h(α) = 0. This contradicts that fact that fα is the monic polynomialsof least degree satisfying fα(α) = 0. This prove (i).

We now turn to (ii). Suppose that g ∈ Q[x] satisfies g(α) = 0. By Euclid’salgorithm we know that

g(x) = q(x)fα(x) + r(x), q(x), r(x) ∈ Q[x],

where either r = 0, or otherwise deg(r) < deg(f). We want to show that r = 0; inthis case g(x) = q(x)fα(x) showing that fα | g and this is what we desire to prove.So let us assume that r 6= 0, and hence r is a non-zero polynomial with rationalcoefficients of degree strictly less than f . Substituting α for x in the above andrecalling that g(α) = fα(α) = 0 leads us at once to conclude that r(α) = 0. Wenow divide r by its leading coefficient to obtain a monic polynomial s ∈ Q[x] withdegree strictly less than fα and satisfying s(α) = 0. This is a contradiction andproves (ii).

There is a slightly different way of proving (ii). We know from (i) that fα isirreducible. Suppose fα - g. Then the polynomials fα and g are coprime. ByEuclid there are polynomials u, v ∈ Q[x] such that

u(x)fα(x) + v(x)g(x) = 1.

Substituting α for x we obtain 0 = 1 which is a contradiction. �

We now prove the uniqueness of the minimal polynomial.

Corollary 5.2. Suppose α is an algebraic number. The minimal polynomial fα isunique.

1Of course we mean here that fα is irreducible over Q. Over C we know by the Fundamental

Theorem of Algebra that we may factorize it as a product of linear factors.

Page 12: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

12 SAMIR SIKSEK

Proof. Suppose there are two minimal polynomials f , g for α. By part (ii) of theabove Lemma we know that f | g and g | f . Hence f = λg for some non-zerorational number λ. But f and g are both monic, and so λ = 1. �

Theorem 5. Suppose α is an algebraic number. A polynomial f ∈ Q[x] is theminimal polynomial of α if and only if f is monic and irreducible and satisfiesf(α) = 0.

Proof. If f is the minimal polynomial of α then f is monic and satisfies f(α) = 0by definition, and f is irreducible by part (i) of Lemma 5.1.

Conversely suppose that f is monic and irreducible and satisfies f(α) = 0. Bypart (ii) Lemma 5.1 we see that fα | f . As f is irreducible, fα is a constant orfα = λf for some non-zero rational λ. Since fα is monic and fα(α) = 0 we see thatfα is non-constant. Hence fα = λf for some non-zero rational λ. But fα and f areboth monic, so λ = 1 proving that f = fα as desired. �

Example 5.1. Let f(X) = X2 − 2; it is monic, has rational coefficients, is irre-ducible and satisfies f(

√2) = 0. By the above theorem, f is the minimal polynomial

of√

2. Notice that it would not have been straightforward to deduce this fact fromthe definition of minimal polynomial.

6. Conjugates

Definition. Suppose that α is an algebraic number and let fα be its minimal polyno-mial. We define the degree of α to be the degree of fα. We define the conjugatesof α to be the roots fα.

By the Fundamental Theorem of Algebra we may factorize fα over C into aproduct of linear factors

fα(X) = (X − α1)(X − α2) · · · (X − αn).

Then α1,. . . ,αn are the conjugates of α. Notice that α is one of them. Here n is thenumber of conjugates of α, and at the same time the degree f (which by definitionis the degree of α). Thus an algebraic number with degree n has n conjugates. Thereader probably expects us to say that an algebraic number with degree n has nconjugates up to repetition, since polynomials can have repeated roots. Howeverthere is no repetition involved here since minimal polynomials (being irreducible)do not have repeated roots, thanks to the following theorem.

Theorem 6. Suppose f ∈ Q[x] is an irreducible polynomial. Then f does not haverepeated roots in C.

In particular, this applies to minimal polynomials of algebraic numbers. Thusan algebraic number has distinct conjugates.

Proof. Suppose f ∈ Q[x] is irreducible, but has repeated root β ∈ C. Consider thederivative f ′ of f . Clearly f ′ ∈ Q[x]. Since f is irreducible and f ′ has degree lessthan f we know that f and f ′ are coprime. By Euclid there are some polynomialsu, v ∈ Q[x] such that

(3) u(x)f(x) + v(x)f ′(x) = 1.

However, β ∈ C is a repeated root of f . Thus

f(x) = (x− β)2g(x)

Page 13: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

ALGEBRAIC NUMBER THEORY 13

where g(x) is a polynomial with coefficients in C. Differentiating we see that

f ′(x) = (x− β)2g′(x) + 2(x− β)g(x).

It is clear that f(β) = f ′(β) = 0. Substituting β for x in (3) gives 0 = 1 which is acontradiction. �

Example 6.1. By Theorem 5, f = X2− 3 is the minimal polynomial of√

3. Thusthe conjugates of

√3 are

√3 and −

√3.

7. Factorization of Polynomials

So far we have concerned ourselves with minimal polynomials of algebraic inte-gers and have neglected algebraic integers. To recall, we say α ∈ C is an algebraicinteger if there exists some monic polynomial f with coefficients in Z such thatf(α) = 0. Of course, algebraic integers are algebraic numbers and so anything thatapplies to algebraic numbers must apply to algebraic integers. It is however naturalto ask, if the minimal polynomial of an algebraic integer must have coefficients inZ. Notice, that there is no reason a priori (on the basis of the above definition) toexpect this. We do not know that the polynomial f is minimal. We know that α hassome minimal polynomial fα which is monic, has rational coefficients, is irreducibleand satisfies fα(α) = 0. As to the relationship between fα and the polynomial fin the above definition, all we can conclude is that fα | f . Since f has integralcoefficients, should we expect that fα has rational coefficients. It turns out thatthe answer is yes. For this we need the following theorem of Gauss.

Theorem 7. (Gauss) Suppose f is a polynomial with coefficients in Z. Supposethat f(x) = g(x)h(x) where g, h are polynomials with coefficients in Q. Then thereis some non-zero rational number λ such that g∗ = λg and h∗ = λ−1 both havecoefficients in Z, and therefore we may factorize f over Z as f = g∗h∗.

Before proving Gauss’ Theorem we need the following Lemma.

Lemma 7.1. Suppose R is an integral domain. Then R[x] is also an integraldomain.

Proof. Suppose R is an integral domain. Suppose f , g are non-zero elements ofR[x]. We would like to show that fg is also non-zero. For this write

f(x) = amxm + am−1x

m−1 + · · ·+ a0, a0, . . . , am ∈ R,

andg(x) = bnx

n + bn−1xn−1 + · · ·+ b0 b0, . . . , bn ∈ R,

with am 6= 0, bn 6= 0. Thus

f(x)g(x) = ambnxm+n + lower terms.

As R is an integral domain and am, bn are non-zero we see that ambn 6= 0. Thusfg is non-zero as required. �

We may now return to prove Gauss’ Theorem.

Proof of Theorem 7. By clearing denominators we may write

(4) nf(x) = g1(x)h1(x)

Page 14: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

14 SAMIR SIKSEK

where n is a positive integer and g1, h1 are polynomials with coefficients in Z whichare multiples of g, h. Write

g1(x) = amxm + am−1x

m−1 + · · ·+ a0, h1(x) = bnxn + bn−1x

n−1 + · · ·+ b0,

where the coefficients are in Z.We will eliminate the prime factors of n, one at a time, until we reach the desired

factorization f = g∗h∗ where g∗ and h∗ have coefficients in Z.If n = 1 then of course there is nothing to prove. Suppose n > 1 and let p be

any prime factor n. Reducing (4) modulo p we obtain

(5) 0 = g1(x)h1(x)

where

g1(x) = amxm + am−1x

m−1 + · · ·+ a0, h1(x) = bnxn + bn−1x

n−1 + · · ·+ b0.

The equality (5) takes place in Fp[x]. But Fp is an integral domain (indeed it is afield), and so by the above Lemma Fp[x] is an integral domain. Hence either g1 = 0or h1 = 0. Without loss of generality, let us say that g1 = 0. This means that allthe coefficients of g1 are divisible by p. Let g2 = g1/p; this has integral coefficients.Let h2 = h1. Thus

n′f = g2h2

where n′ = n/p and g2, h2 have coefficients in Z. We repeat this until we haveeliminated all prime factors of n. �

We now deduce our desired result.

Theorem 8. Suppose α is an algebraic number. Then alpha is an algebraic integerif and only if its minimal polynomial fα has coefficients in Z.

Proof. If fα has integral coefficients, then α is an algebraic integer by the verydefinition of algebraic integers.

Conversely suppose that α is an algraic integer. By definition, h(α) = 0 for somemonic polynomial h with coefficients in Z. By Lemma 5.1 we know that fα | h.Thus we may write

h(x) = fα(x)g(x)

where g(x) ∈ Q[x]. By Gauss’ Theorem above, there is some non-zero rational λsuch that λfα and λ−1g(x) both have integral coefficients. But both h and fα aremonic, and therefore g is monic. Examining the leading coefficients of λfα andλ−1g(x) we see that both λ and λ−1 are integers. Thus λ = ±1. Hence ±fα(x) hasintegral coefficients which implies that fα(x) has integral coefficients, as desired. �

8. Eisenstein’s Criterion For Irreducibility

Theorem 9. Letf(x) = anx

n + an−1xn−1 + · · ·+ a0,

be a polynomials with coefficients in Z, satisfying the following three conditions:(i) p - an;(ii) p | ai for 1 ≤ i ≤ n− 1;(iii) p2 - a0.

Then f is irreducible over Q.

Page 15: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

ALGEBRAIC NUMBER THEORY 15

Proof. We prove this by contradiction. Suppose that f is not irreducible over Q.By Gauss’ Theorem we know that f(x) = g(x)h(x) where g, h are polynomials withcoefficients in Z and degrees strictly less than that of f . Write

g(x) = brxr + · · ·+ b0, h(x) = csx

s + · · ·+ c0.

Note that n = r + s and an = brcs. By assumption (i), an is not divisible by pand so neither br nor cs are. Reducing the identity f(x) = g(x)h(x) modulo p weobtain (using (ii))

g(x)h(x) = anxn;

here the equality takes place in Fp[x]. The only possible way to factorize xn as aproduct of two factors is xuxv for some positive u, v satisfying u+v = n. However,g(x), h(x) are polynomials of degrees r, s with leading coefficients br, cs respectively.We deduce that

g(x) = brxr, h(x) = csx

s.

Comparing the coefficients of g, h with those of g, h we see that p | b0 and p | c0.However, a0 = b0c0. Thus p2 | a0 contradicting (iii). �

Example 8.1. Let f(x) = x7 − 9x+ 3. Letting p = 3 in Eisenstein’s criterion, weimmediately see that f is irreducible.

Many polynomials do not immediatly satisfy the conditions of Eisenstein’s cri-terion, but do satisfy them after making an appropriate substitution. For example,take g(X) = 8x3 − 6x + 1. Eisenstein’s criterion does not apply to g regardless ofthe prime p chosen. Now let

h(x) = g(x+ 1) = 8(x+ 1)3 − 6(x+ 1) + 1 = 8x3 + 24x2 + 18x+ 3.

We see that Eisenstein’s criterion applies to h with p = 3. Thus h is irreducible.But if g was reducible then h would also be reducible because of the relation h(x) =g(x+ 1). Hence g is irreducible.

9. Symmetric Polynomials

This section is based on the corresponding section in Stewart and Tall.Let Z[t1, . . . , tn] be the ring of polynomials in indeterminants t1,. . . ,tn with coeffi-

cients in Z. Let Sn be the symmetric group of permutations on the set {1, 2, . . . , n}.For any permutation π ∈ Sn and any polynomial f ∈ Z[t1, . . . , tn] we define thepolynomial fπ by

fπ(t1, . . . , tn) = f(tπ(1), . . . , tπ(n)).

For example, if n = 5, f = t1 + t2t3 + t24 − t5 and π = (132)(45) then fπ =t3 + t1t2 + t25 − t4.

Definition. We call a polynomial f ∈ Z[t1, . . . , tn] symmetric if fπ = f for allπ ∈ Sn.

For example, if n = 3, then t1 + t2 + t3 and t1t2t3 are symmetric. Perhaps lessobvious is the fact that t1t2 + t2t3 + t3t1 is also symmetric. For general n, we define

s1 =n∑

i=1

ti, s2 =∑

1≤i<j≤n

titj , s3 =∑

1≤i<j<k≤n

titjtk, . . . , sn = t1t2 . . . tn.

Page 16: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

16 SAMIR SIKSEK

The polynomials s1, . . . , sn are called the elementary symetric polynomials in t1, . . . , tn.Note the identity

(X − t1)(X − t2) . . . (X − tn) = Xn − s1Xn−1 + s2X

n−2 − · · ·+ (−1)nsn.

If we act on t1, . . . , tn by any permutation π ∈ Sn then we will leave the left-handside of this identity unchanged, and so the coefficients on the right-hand must beunchanged. This shows that the si are indeed symmetric functions.

It is easy to see that any polynomial in s1, . . . , sn with coefficients in Z is asymmetric function of t1, . . . , tn. We would like to prove the converse.

Theorem 10. (Newton) Every symmetric polynomial in Z[t1, . . . , tn] is expressibleas a polynomial in s1, . . . , sn with coefficients in Z.

Before proving the theorem, we will give an example to illustrate what we mean.

Example 9.1. Let n = 2. The polynomial

f = t21 + t22

is obviously symmetric. We can rewrite it as

f = t21 + t22 = (t1 + t2)2 − 2t1t2 = s21 − 2s2.

Hence f can be written as a polynomial in s1, s2 with coefficients in Z.

Proof of Theorem 10. We give a constructive argument for writing a symmetricpolynomial in terms of s1, . . . , sn. First we define an ordering on the monomialsta11 · · · tan

n . We will order the monomials ‘lexicographically’:

ta11 · · · tan

n < tb11 · · · tbnn iff

a1 > b1,

or a1 = b1 and a2 > b2,

or a1 = b1 and a2 = b2 and a3 > b3,

etc.

Suppose that f is a symmetric polynomial and let cta11 . . . tan

n be the monomialappearing in f which is biggest according to the lexicographic ordering. We claima1 ≥ a2; suppose otherwise that a1 < a2. Since f is symmetric, f also has themonomial cta2

1 ta12 t

a33 . . . tan

n and this is bigger than cta11 . . . tan

n giving a contradiction.Therefore a1 ≥ a2 and similarly a2 ≥ a3 and so on. In other words

a1 ≥ a2 ≥ a3 ≥ · · · ≥ an.

Now write

k1 = a1 − a2, k2 = a2 − a3, . . . , kn−1 = an−1 − an, kn = an.

These are all non-negative since a1 ≥ a2 ≥ a3 ≥ · · · an. Consider

sk11 s

k22 . . . skn

n = (t1 + · · ·+ tn)k1(t1t2 + . . . )k2 . . . (t1 . . . tn)kn .

The biggest monomial in this expression is

tk11 (t1t2)k2(t1t2t3)k3 · · · = tk1+k2+···+kn

1 tk2+···+kn2 . . . tkn

n = ta11 t

a22 · · · tan

n .

Therefore, the biggest monomials in f and in csk11 s

k22 . . . skn

n are the same, and thepolynomial

f2 = f − csk11 s

k22 . . . skn

n

will contain only smaller monomials. Our objective was to show that f can bewritten as a polynomial in s1, . . . sn with coefficients in Z. It is sufficient to do that

Page 17: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

ALGEBRAIC NUMBER THEORY 17

for f2 which has a smaller leading monomial. We apply the argument recursivelyto obtain a sequence of polynomails f1 = f, f2, f3, . . . each with a smaller leadingmonomial; moreover the leading monomial will always be of the shape

tu11 tu2

2 . . . tunn , u1 ≥ u2 ≥ · · · ≥ un ≥ 0.

Clearly this process will stop eventually, leaving us with a a polynomial in s1, . . . sn

with integer coefficients. �

Example 9.2. Express f = t21t22 + t22t

23 + t23t

21 in terms of elementary symmetric

functions.Answer: We follow the recipe in the above proof. The biggest monomial of f ist21t

22. Hence a1 = 2, a2 = 2, a3 = 0. Thus k1 = 2 − 2 = 0, k2 = 2 − 0 = 2, k3 = 0.

The above recipe suggests that we subtract s01s22s

03 = s22:

f − s22 = t21t22 + t22t

23 + t23t

21 − (t1t2 + t2t3 + t3t1)2 = −2t21t2t3 − 2t22t1t3 − 2t23t1t2.

The biggest monomial now is −2t21t2t3. Here a1 = 2, a2 = 1, a3 = 1. Thusk1 = 2− 1 = 1, k2 = 1− 1 = 0 and k3 = 1. We subtract −2s11s

02s

13. In other words,

we add 2s1s3:f − s22 + 2s1s3 = 0.

It follows that f = s22 + 2s1s3.

10. Algebraic Integers form a Ring

We defined O to be the set of algebraic integers. In this section we show thatO is a subring of C. We recall first the elementary symmetric polynomials inindeterminants t1, . . . , tn:

s1 =n∑

i=1

ti, s2 =∑

1≤i<j≤n

titj , s3 =∑

1≤i<j<k≤n

titjtk, . . . , sn = t1t2 . . . tn.

Theorem 11. Suppose α is an algebraic integer of degree n and let α1, . . . , αn beits conjugates. Let h(t1, . . . , tn) ∈ Z[t1, . . . , tn] be a symmetric polynomial. Thenh(α1, . . . , αn) ∈ Z.

For the proof of the theorem we need the following lemma, which in fact is aspecial case of the theorem.

Lemma 10.1. Suppose α is an algebraic integer of degree n and let α1, . . . , αn beits conjugates. Let s1, . . . , sn be the elementary symmetric polynomials in t1, . . . , tnas above. Then

s1(α1, . . . , αn), s2(α1, . . . , αn), . . . , sn(α1, . . . , αn) ∈ Z.

Proof. Recall the identity

(X − t1)(X − t2) . . . (X − tn) = Xn − s1Xn−1 + s2X

n−2 − · · ·+ (−1)nsn.

Now substitute α1, . . . , αn for t1, . . . , tn. We obtain

(X − α1)(X − α2) . . . (X − αn) =

Xn − s1(α1, . . . , αn)Xn−1 + s2(α1, . . . , αn)Xn−2 − · · ·+ (−1)nsn(α1, . . . , αn).

Since α1, . . . , αn are the conjugates of α, we recognize the polynomial on the left asthe minimal polynomial of α. As α is an algebraic integer, its minimal polynomialhas integral coefficients, and the lemma follows. �

Page 18: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

18 SAMIR SIKSEK

Proof of Theorem 11. Suppose h(t1, . . . , tn) ∈ Z[t1, . . . , tn] is symmetric. By New-ton’s Theorem (Theorem 10), h can be expressed as a polynomial in s1, . . . , sn withcoefficients in Z. Lemma 10.1 tells us that

s1(α1, . . . , αn), s2(α1, . . . , αn), . . . , sn(α1, . . . , αn) ∈ Z.

Hence h(α1, . . . , αn) ∈ Z. �

We are now ready to prove our goal.

Theorem 12. The set of algebraic integers O is a subring of C.

Proof. We know that Z ⊂ O, therefore O 6= ∅. To show that O is a subring, it isenought to show that if α, β ∈ O then α+ β, −α, αβ ∈ O.

Suppose α, β ∈ O and let us prove that α + β ∈ O. To do this we construct amonic polynomial h with coefficients in Z such that h(α+ β) = 0. First let

fα(x) = xm + am−1xm−1 + · · ·+ a0 ∈ Z[x]

be the minimal polynomial of α. Suppose β has degree n and let β1, . . . , βn be itsconjugates where we take β1 = β. Consider the product(6)fα(x−t1)fα(x−t2) . . . fα(x−tn) = xmn+umn−1(t1, . . . , tn)xmn−1+· · ·+u0(t1, . . . , tn).

As fα has coefficients in Z, it is clear that ui(t1, . . . , tn) ∈ Z[t1, . . . , tn]. Note thatif we permute the tj then we permute the factors of the left-hand side of (6), andso the product is unchanged. Hence the coefficients ui(t1, . . . , tn) are unchanged onpermuting the tj ; in other words they are symmetric polynomials. By Theorem 11we deduce that

(7) ui(β1, . . . , βn) ∈ Z.

Now substitute β1, . . . , βn for t1, . . . , tn in (6), and let h(x) be the resulting poly-nomial. We obtain

h(x) = fα(x− β1)fα(x− β2) . . . fα(x− βn)

= xmn + umn−1(β1, . . . , βn)xmn−1 + · · ·+ u0(β1, . . . , βn).

Now h(x) is monic. Moreover, h(x) ∈ Z[x] by (7). Finally,

h(α+ β) = fα(α+ β − β1)fα(α+ β − β2) . . . fα(α+ β − βn)

= fα(α)fα(α+ β − β2) . . . fα(α+ β − βn) since β1 = β

= 0 · fα(α+ β − β2) . . . fα(α+ β − βn) = 0.

This shows indeed that α+ β ∈ O.The proofs that −α and αβ ∈ O are left as exercises. �

Example 10.1. Let α be a root of the polynomial x3+4x+2. What is the minimalpolynomial of α+

√2?

Answer: Note that the polynomial f(x) = x3+4x+2 is irreducible by Eisenstein’scriterion (take p = 2). Hence it is the minimal polynomial of α. We follow the stepsin the above proof taking β =

√2. The conjugates of β are β1 =

√2 and β2 = −

√2.

Page 19: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

ALGEBRAIC NUMBER THEORY 19

Then α+ β is a root of

h(x) = f(x− β1)f(x− β2)

= f(x−√

2)f(x+√

2)

=((x3 + 10x+ 2)− (3x2 + 6)

√2)(

(x3 + 10x+ 2) + (3x2 + 6)√

2)

= (x3 + 10x+ 2)2 − 2(3x2 + 6)2

= x6 + 2x4 + 4x3 + 28x2 + 40x− 68.

We haven’t proved that h is the minimal polynomial for α +√

2. To do this wemust show that h is irreducible, which we leave as an exercise for the reader.

11. Algebraic Numbers form a Field

Theorem 13. The set of algebraic numbers Q is in fact a subfield of C.

Proof. We know that Q ⊂ Q. Hence Q is non-empty. We need to show that if α, βare in Q, then so are α+ β, −α, αβ and α−1 (for non-zero α). The last we showedin Lemma 4.2. The first three are similar to the corresponding proofs for algebraicintegers and so we leave them to the reader. �

12. Number Fields

Suppose L is a field and K a subfield. It is easy to see that L is a vector spaceover K (you can add and subtract elements of L, and you can multiply them byelements of K—check that the vector space axioms follow from the field axioms).We denote the dimension of L regarded as a vector space over K by [L : K], andwe say that L is a field extension of K of degree [L : K]. If the degree [L : K] isfinite, we say that L is a finite extension of K. Notice that this phrase does notmean that the field L is finite (like Fp) but that L, as a vector space over K, hasfinite dimension.

We finally come to the main object of study of algebraic number theory.

Definition. A number field K is a subfield of C that is a finite extension of Q.The degree of K is [K : Q].

A quadratic number field is a number field of degree 2, a cubic numberfield is a number field of degree 3 and so on.

Notice that any subfield K of C will contain contain the rationals. To see thisnote first that 1 ∈ K because K is a subfield. Now repeated addition of 1 willshow that the natural numbers are in K, and ‘minusing’ that the integers are inK. Taking ratios shows that the rationals are contained in K. However we do notcall K a number field unless it is a finite extension of Q.

Example 12.1. In previous courses you probably defined

Q(√

2) = {a+ b√

2 : a, b ∈ Q},

and proved that this is a field. We see that 1,√

2 is a basis for Q(√

2) as a Q-vectorspace. Hence Q(

√2) is a number field of degree 2; in other words it is a quadratic

number field.

Page 20: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

20 SAMIR SIKSEK

Definition. Suppose α is an algebraic number. Define

Q(α) ={g(α)h(α)

: g, h ∈ Q[x], h(α) 6= 0}.

We call Q(α) the field generated by α.

Exercise 12.2. Check that Q(α) is indeed a field.

Theorem 14. Suppose α is an algebraic number of degree n. Then Q(α) is anumber field of degree n. The set 1, α, α2, . . . , αn−1 is a Q-basis for Q(α).

In other words, every element of Q(α) can be expressed uniquely as a linearcombination of 1, α, α2, . . . , αn−1. This shows that our old definition of Q(

√2) is

consistent with the new one.

Proof of Theorem 14. It is sufficient to show that the set 1, α, α2, . . . , αn−1 is a Q-basis for Q(α). Suppose β ∈ Q(α) we would like to show that β can be written asa Q-linear combination of 1, α, α2, . . . , αn−1. By definition of Q(α), we know thatβ = g(α)/h(α) where g, h ∈ Q[x] and h(α) 6= 0.

Our first step is to invert h(α). Let fα be the minimal polynomial of α. Thenh(α) 6= 0 implies that fα - h; since fα is irreducible we deduce that fα and h arecoprime. By Euclid there exists polynomials u, v ∈ Q[x] such that

u(x)h(x) + v(x)fα(x) = 1.

Substituting α for x we obtain u(α) = 1/h(α). Hence β = g(α)/h(α) = g(α)u(α).Let w(x) = g(x)u(x) ∈ Q[x]. Thus β = w(α). Thus we have written β as apolynomial in α. In other words, we have written β as a linear combination of1, α, α2, . . . αm where m is the degree of w. Our problem is that m might begreater than n− 1.

Recall that n is the degree of α. We know that this is also the degree of fα. ByEuclid again we know that there are two polynomials q, r ∈ Q[x] such that

w(x) = q(x)fα(x) + r(x), deg(r) < deg(f).

Since deg(r) < deg(f) = n we can write

r(x) = a0 + a1x+ · · · an−1xn−1, ai ∈ Q.

Substitute α for x to obtain

β = w(α) = r(α) = a0 + a1α+ · · · an−1αn−1.

Thus every element of Q(α) can be written as a Q-linear combination of 1, α, . . . , αn−1.To complete the proof that 1, α, . . . , αn−1 is a basis, we must of course show that

it is linearly independent. We leave this to the reader. �

To make sure you understood the proof of the above theorem, do the followingexercise.

Exercise 12.3. Let f(X) = X3 +X2 + 1.(i) Show that f is irreducible.(ii) Let θ be a root of f and K = Q(θ). Write the following elements as Q-linear

combinations of 1, θ, θ2:

(θ + 1)4,1

θ2 − 1,

θ + 1θ2 + 1

.

Page 21: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

ALGEBRAIC NUMBER THEORY 21

Theorem 15. (Primitive Element Theorem) Suppose K is an number field of de-gree n. Then K = Q(α) for some algebraic number α of degree n.

Proof. This will be proved in the Galois Theory course. �

13. Fields Generated by Conjugate Elements

Theorem 16. Suppose that α is an algebraic number and fα(x) is its minimalpolynomial. Then we have an isomorphism of fields

Q[x]/(fα(x)) ∼= Q(α),

explicitly given by

φ : Q[x]/(fα(x)) → Q(α), φ (x+ (fα(x))) = α.

Proof. Define the map

ψ : Q[x] → Q(α), ψ(g(x)) = g(α).

The map ψ is an homomorphism of rings (check). By the First Isomorphism The-orem we have that

Q[x]/Ker(ψ) ∼= Im(ψ).We need to calculate the kernel and the image.

Now g(x) ∈ Ker(ψ) if and only if g(α) = 0. This is equivalent to fα | g which inturn is equivalent to g ∈ (fα). Hence

Ker(ψ) = (fα(x)).

We claim that Im(ψ) = Q(α) (in other words the map ψ is surjective). To seethis, suppose that β ∈ Q(α). By Theorem 14 we can write

β = a0 + a1α+ · · ·+ an−1αn−1, ai ∈ Q.

Let g(x) = a0 + a1x+ · · ·+ an−1xn−1. Then

ψ(g(x)) = g(α) = β.

In other words β ∈ Im(ψ), showing that ψ is surjective.We now deduce that

Q[x]/(fα(x)) ∼= Q(α).To complete the proof, we must determine the isomorphism explicitely. The explicitisomorphism we gave in the statement of the theorem comes from the proof of theFirst Isomorphism Theorem. �

Corollary 13.1. Suppose that α, α′ are conjugates (i.e. algebraic numbers havingthe same minimal polynomial). Then the fields Q(α) and Q(α′) are isomorphic;indeed there is a unique isomorphism

Q(α) → Q(α′)

fixing Q and satisfying α 7→ α′.

Proof. From Theorem 16 we have isomorphisms

φ : Q[x]/(fα(x)) → Q(α), φ (x+ (fα(x))) = α.

andφ′ : Q[x]/(fα(x)) → Q(α′), φ′ (x+ (fα(x))) = α′.

Then φ′ ◦ φ−1 is the required isomorphism. �

Page 22: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

22 SAMIR SIKSEK

Example 13.1. Let K = Q(√

2). Then√

2 has minimal polynomial x2−2. Hencethe conjugates of

√2 are

√2 and −

√2. Now Q(

√2) = Q(−

√2). We know from

the above that the map

σ : K → K, a+ b√

2 7→ a− b√

2 for all a, b ∈ Q

is an isomorphism of fields. The reader might want to check this directly. Noticealso that σ(a) = a for all a ∈ Q. Hence σ fixes the rationals.

14. Embeddings

Lemma 14.1. Let f ∈ R[x] and let γ ∈ C be a root of f . The γ (the complexconjugate of γ) is also a root of f .

Proof. Write

f(x) = anxn + an−1x

n−1 + · · ·+ a0, ai ∈ R.

But f(γ) = 0, so0 = f(γ)

= anγn + · · ·+ a0

= an · γn + · · · a0

= anγn + · · ·+ a0 since ai ∈ R

= f(γ).Hence γ is a root of f . �

Suppose α is an algebraic number and let α1, . . . , αn be its conjugates. Bydefinition, the αi are the roots of fα(x) ∈ Q[x]. By the above Lemma, αi is againof one α1, . . . , αn. This means that either αi is real, or if it is not real then itscomplex conjugate must be included in the list α1, . . . , αn. It is traditional toreorder α1, . . . , αn as follows

α1, . . . , αr, αr+1, . . . , αr+s, αr+s+1, . . . , αr+2s,

where α1, . . . , αr are real, the others are non-real complexes, and

αr+1 = αr+s+1, αr+2 = αr+s+2, αr+s = αr+2s.

Note that n = r + 2s where r is the number of real conjugates of α and s is thenumber of pairs of complex conjugates of α.

Now Corollary 13.1 shows that Q(α) ∼= Q(αi) for i = 1, . . . , n. However, fori = 1, . . . , r we have Q(αi) ⊂ R. Thus composing the isomorphism Q(α) ∼= Q(αi)with the inclusion Q(αi) ⊂ R we obtain an embedding

σi : Q(α) ↪→ R, σi(α) = αi.

An embedding (denoted by the hook arrow ↪→) is an injective homomorphism. Wecall σ1, . . . , σr the real embeddings of K = Q(α).

For i = r + 1, . . . , r + 2s we define the complex embeddings

σi : Q(α) ↪→ C, σi(α) = αi.

Note thatσi(β) = σi+s(β)

for all i = r + 1, . . . , r + s and all β ∈ K.

Page 23: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

ALGEBRAIC NUMBER THEORY 23

Example 14.1. Let K = Q(√

2). Since√

2 has exactly two real conjugates,√

2,−√

2, we have r = 2, s = 0. The real embeddings are

σ1 : K ↪→ R, σ2 : K ↪→ R

whereσ1(a+ b

√2) = a+ b

√2, σ2(a+ b

√2) = a− b

√2

for all a, b ∈ Q.

Example 14.2. Let K = Q(√−3). Since

√−3 has exactly two complex, non-real,

conjugates,√−3, −

√−3, we have r = 0, s = 1. The complex embeddings are

σ1 : K ↪→ C, σ2 : K ↪→ C

whereσ1(a+ b

√−3) = a+ b

√−3, σ2(a+ b

√−3) = a− b

√−3

for all a, b ∈ Q.

Example 14.3. Let K = Q( 3√

2). The minimal polynomial of 3√

2 is f(x) = x3−2.Write ω = exp(2πi/3). Then f has three roots

α1 = 3√

2, α2 = ω3√

2, α3 = ω2 3√

2,

and so these are the conjugates of 3√

2. The first is real and the other two arecomplex conjugates. Hence r = 1, s = 1. We have embeddings

σ1 : K ↪→ R, σ2 : K ↪→ C, σ3 : K ↪→ C

where

σ1

(a+ b

3√

2 + c( 3√

2)2)

= a+ b3√

2 + c( 3√

2)2,

σ2

(a+ b

3√

2 + c( 3√

2)2)

= a+ bω3√

2 + cω2( 3√

2)2,

σ3

(a+ b

3√

2 + c( 3√

2)2)

= a+ bω2 3√

2 + cω( 3√

2)2,

for all a, b, c ∈ Q.

15. Field Polynomial

We continue with the notation of the previous section: K = Q(α) is a numberfield of degree n and σ1, . . . , σn are the embeddings of K.

Definition. Let β ∈ K. We call σ1(β), . . . , σn(β) the K-conjugates of β.We define the field polynomial Fβ(x) by

Fβ(x) =n∏

i=1

(x− σi(β)) .

We define the K-norm by

NK(β) =n∏

i=1

σi(β)

and the K-trace of β by

TrK(β) =n∑

i=1

σi(β)

Page 24: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

24 SAMIR SIKSEK

Example 15.1. Suppose d is a square-free integer, d 6= 0, 1. Let K = Q(√d). By

Theorem 14 any element β ∈ K can expressed uniquely in the form β = a + b√d

for some a, b ∈ Q. The embeddings σ1, σ2 satisfy

σ1(a+ b√d) = a+ b

√d, σ2(a+ b

√d) = a− b

√d.

Note that σ1, σ2 are both real if d > 0 and both complex if d < 0. Thus theK-conjugates of β = a+ b

√d are a+ b

√d and a− b

√d.

The field polynomial of β = a+ b√d is

Fβ(x) = (x− σ1(β)) (x− σ2(β))

= x2 − 2a x+ (a2 − db2).

MoreoverTr(β) = 2a, NK(β) = a2 − db2.

We note in the above example that the for elements of quadratic fields, theK-norms, K-traces are rational, and that the field polynomials have rational coef-ficients. This isn’t a coincidence.

Lemma 15.1. Suppose K = Q(α) is a number field and β ∈ K. Then(i) Fβ(x) ∈ Q[x];(ii) Tr(β) ∈ Q and NK(β) ∈ Q.

Proof. Suppose K = Q(α) is a number field of degree n with embeddings σ1, . . . , σn

corresponding to the conjugates α1, . . . , αn of α.It is clear from the definitions that

Fβ(x) = xn − Tr(β)xn−1 + · · ·+ (−1)nNK(β).

Hence (ii) follows immediately from (i).Let us prove (i). By Theorem 14 we can write β = a0 + a1α + · · · + an−1α

n−1

with ai ∈ Q. Let h(x) = a0 + a1x+ · · ·+ an−1xn−1 ∈ Q[x]. Then β = h(α) and

σi(β) = h (σi(α)) = h(αi).

By definition

Fβ(x) =n∏

i=1

(x− σi(β))

and hence

Fβ(x) =n∏

i=1

(x− h(αi)) .

The polynomial on the right-hand side is unchanged by permutations of α1, . . . , αn.Thus the coefficients of Fβ(x) are symmetric polynomials in α1, . . . , αn with coef-ficients in Q. This shows indeed that the coefficients of Fβ(x) are in Q. �

Theorem 17. Suppose K is a number field of degree n. Suppose β ∈ K has degreem. Then

(i) m | n;(ii) Writing l = n/m, we have

Fβ(x) = fβ(x)l;

(iii) The K-conjugates of β are the conjugates of β repeated l times.

Page 25: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

ALGEBRAIC NUMBER THEORY 25

Here as usual, Fβ is the field polynomial of β and fβ is the minimal polynomial ofβ.

Before proving the theorem we make an important remark. When we definedthe field polynomial Fβ(x), the definition depended not only on β and K but on achoice of a generator α for K. The above theorem shows that Fβ(x) depends onlyon β and the field K.

Proof of Theorem 17. As f(β) = 0 we see that

f (σi(β)) = σi (f(β)) = 0.

Thus σi(β) is a root of f and so by definition a conjugate of β. Our first observationis: every K-conjugate of β (equivalently every root of Fβ) is a conjugate of β.

Recall that the minimal polynomial fβ is irreducible. Thus we can write

(8) Fβ(x) = fβ(x)lg(x)

for some non-negative integer l and some g ∈ Q[x] such that fβ - g. We note that Fβ

and fβ are both monic, hence g is monic. We would like to show that g = 1. To dothis it is enough to show that g has no roots. We argue by contradiction. Supposeγ is a root of g. From (8) we see that γ is a root of Fβ . By our observation above, γis a root of fβ . Thus γ is a common root of fβ and g. As fβ is irreducible and fβ - gwe see that fβ and g are coprime. Hence there are polynomials u(x), v(x) ∈ Q[x]such that

u(x)fβ(x) + v(x)g(x) = 1.Substituting γ for x we get 0 = 1 which is a contradition. Hence g = 1 and thisshows that Fβ(x) = fβ(x)l.

Now Fβ has degree n (the same degree as K). However fβ has degree m (thesame degree as β). Comparing degrees on both sides of Fβ(X) = fβ(x)l we obtainn = lm. Thus m | n. This proves (i) and (ii) simultaneously.

(iii) follows immediately from the equality Fβ(x) = fβ(x)l. �

16. Ring of Integers

We recall that an algebraic integer β is a complex number satisfying f(β) = 0for some monic polynomial with coefficients in Z. We showed that an algebraicnumber β is an algebraic integer if and only if its minimal polynomial fβ ∈ Z[x].We denoted the set of algebraic integers by O and showed that it is a subring of C.

Definition. Let K be a number field. We define the ring of integers OK of K tothe set OK = K ∩ O.

We note that OK is the intersection of two subrings of C and hence must be asubring. Note also that Z ⊆ OK .

Definition. Let K be a number field of degree n. An integral basis for OK is a setω1, . . . , ωn of elements in OK such that every β ∈ OK can be written uniquely as

β = a1ω1 + a2ω2 + · · ·+ anωn

with a1, . . . , an ∈ Z.

Theorem 18. Suppose K is a number field of degree n. Then OK has an integralbasis ω1, . . . , ωn.

Page 26: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

26 SAMIR SIKSEK

We omit the proof of this theorem. This is essentially a theorem about torsion-free abelian groups.

Example 16.1. In Theorem 4 we showed that O∩Q = Z. Thus the ring of integersof the number field Q is OQ = Z. An integral basis for this is {1}.

The following lemma is helpful in deciding if an algebraic number is an algebraicinteger.

Lemma 16.1. Suppose K is a number field and β ∈ K. Then β ∈ OK if and onlyif Fβ ∈ Z[x].

Proof. We know from Theorem 17 that

Fβ(x) = fβ(x)l

for some positive integer l. If β ∈ OK , that is β is an algebraic integer, thenfβ ∈ Z[x] and hence Fβ ∈ Z[x].

Conversely, suppose that Fβ ∈ Z[x]. It follows from Theorem 7 that fβ ∈ Z[x].Thus β is an algebraic integers; i.e. β ∈ OK . �

Example 16.2. Let K = Q(i). We would like to compute OK and an integral basisfor it. We know that Z ⊆ OK . Moreover i is the root of the monic x2 + 1 ∈ Z[x]and so i ∈ OK .

Thus we see that Z[i] ⊆ OK . Here

Z[i] = {a+ bi : a, b ∈ Z} .We would like to determine whether or not OK = Z[i].

Suppose β ∈ K. Since 1, i is a Q-basis for K we may write β = u+ vi for someu, v ∈ Q. Write

u = u′ + u′′, v = v′ + v′′

where u′, v′ ∈ Z and0 ≤ u′′ < 1, 0 ≤ v′′ < 1.

Letβ′ = u′ + v′i, β′′ = u′′ + v′′i.

Now β′ ∈ Z[i] ⊆ OK . Hence

β ∈ OK ⇐⇒ β − β′ ∈ OK

⇐⇒ β′′ ∈ OK

⇐⇒ Fβ′′ ∈ Z[x]

⇐⇒ x2 − 2u′′x+ u′′2 + v′′

2 ∈ Z[x]

⇐⇒ 2u′′ ∈ Z and u′′2 + v′′

2 ∈ Z.

However 0 ≤ u′′ < 1, and so 2u′′ ∈ Z gives us two possibilities u′′ = 0 or 1/2. Ifu′′ = 0 then v′′2 ∈ Z and 0 ≤ v′′ < 1 so v′′ = 0. Thus we see that β = β′ ∈ Z[i].

If however u′′ = 1/2 then 1/4 + v′′2 ∈ Z. But 0 ≤ v′′ < 1 so

14≤ 1

4+ v′′

2<

54.

Hence 1/4 + v′′2 = 1 and so v′′2 = 3/4 giving us a contradiction.

Thus OK = Z[i]. The set 1, i is an integral basis for OK .

Page 27: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

ALGEBRAIC NUMBER THEORY 27

Example 16.3. Let K = Q(√

5). We follow the same steps as in the previousexample to determine OK . Again Z[

√5] ⊆ OK , where

Z[√

5] ={a+ b

√5 : a, b ∈ Z

}.

Suppose β ∈ K. Since 1,√

5 is a Q-basis for K we may write β = u + v√

5 forsome u, v ∈ Q. Write

u = u′ + u′′, v = v′ + v′′

where u′, v′ ∈ Z and0 ≤ u′′ < 1, 0 ≤ v′′ < 1.

Letβ′ = u′ + v′

√5, β′′ = u′′ + v′′

√5.

Now β′ ∈ Z[√

5] ⊆ OK . Hence

β ∈ OK ⇐⇒ β − β′ ∈ OK

⇐⇒ β′′ ∈ OK

⇐⇒ Fβ′′ ∈ Z[x]

⇐⇒ x2 − 2u′′x+ u′′2 − 5v′′2 ∈ Z[x]

⇐⇒ 2u′′ ∈ Z and u′′2 − 5v′′2 ∈ Z.

However 0 ≤ u′′ < 1, and so 2u′′ ∈ Z gives us two possibilities u′′ = 0 or 1/2. Ifu′′ = 0 then v′′2 ∈ Z and 0 ≤ v′′ < 1 so v′′ = 0. Thus we see that β = β′ ∈ Z[

√5].

If however u′′ = 1/2 then 1/4− 5v′′2 ∈ Z. But 0 ≤ v′′ < 1 so−194

<14− 5v′′2 ≤ 1/4.

Hence 1/4− 5v′′2 = −4,−3,−2,−1, 0. Examining all the possibilities we find thatv′′ = 1/2. Thus

β′′ = 0, or β′′ =1 +

√5

2and both of these are in OK . It is now easy to see that

OK = Z

[1 +

√5

2

]=

{a+ b

(1 +

√5

2

): a, b ∈ Z

}.

It follows that the set 1, (1 +√

5)/2 is an integral basis for OK .

We can generalize the above two examples to quadratic number fields.

Theorem 19. Let K = Q(√d) where d is a square-free integer, and d 6= 0, 1

(1) If d 6≡ 1 (mod 4) then

OK = Z[√d] =

{a+ b

√d : a, b ∈ Z

}.

In particular, 1,√d is an integral basis for OK .

(2) If d ≡ 1 (mod 4) then

OK = Z

[1 +

√d

2

]=

{a+ b

(1 +

√d

2

): a, b ∈ Z

}.

In particular 1, (1 +√d)/2 is an integral basis for OK .

Page 28: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

28 SAMIR SIKSEK

We leave the proof of this theorem as an exercise.

17. Determinants and Discriminants

Let K be a number field of degree n with embeddings σ1, . . . , σn. Let ω1, . . . , ωn

be a basis for K over Q. Consider the matrixσ1(ω1) σ1(ω2) . . . σ1(ωn)σ2(ω1) σ2(ω2) . . . σ2(ωn)

......

...σn(ω1) σn(ω2) . . . σn(ωn)

which we denote by (σi(ωj)) for short. We define the determinant of the basisω1, . . . , ωn to be the determinant of this matrix and denote by D(ω1, . . . , ωn). Thus

D(ω1, . . . , ωn) = det(σi(ωj)).

We define the discriminant of the basis ω1, . . . , ωn to be the square of the determi-nant and denote by ∆(ω1, . . . , ωn). Hence

∆(ω1, . . . , ωn) = D(ω1, . . . , ωn)2 = det(σi(ωj))2.

Suppose now that β1, . . . , βn is another basis for K over Q. Then

βi =n∑

j=1

cijωj

for some cij ∈ Q satisfyingdet(cij) 6= 0.

It is an easy linear algebra exercise to show that

D(β1, . . . , βn) = det(cij)D(ω1, . . . , ωn),

and hence∆(β1, . . . , βn) = det(cij)2∆(ω1, . . . , ωn).

Theorem 20. Suppose K = Q(α) is a number field of degree n. Then(i) The discriminant of the basis 1, α, . . . , αn−1 is given by

∆(1, α, . . . , αn−1) =∏

1≤i<j≤n

(αi − αj)2.

where α1, . . . , αn are the conjugates of α.(ii) If β1, . . . , βn is any basis for K over Q then ∆(β1, . . . , βn) 6= 0 and ∆(β1, . . . , βn) ∈

Q.(iii) If β1, . . . , βn is an integral basis then ∆(β1, . . . , βn) 6= 0 and ∆(β1, . . . , βn) ∈

Z.(iv) If β1, . . . , βn and γ1, . . . , γn are both integral bases for OK then

∆(β1, . . . , βn) = ∆(γ1, . . . , γn).

Proof. We know that σi(α) = αi. Hence the determinant of the basis 1, α, . . . , αn−1

is

(9) D(1, α, . . . , αn−1) =

∣∣∣∣∣∣∣∣∣1 α1 α2

1 . . . αn−11

1 α2 α22 . . . αn−1

2...

......

1 αn α2n . . . αn−1

n

∣∣∣∣∣∣∣∣∣ .

Page 29: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

ALGEBRAIC NUMBER THEORY 29

You will instantaneously recognize this as a Vandermonde determinant, and recallthat

D(1, α, . . . , αn−1) =∏

1≤i<j≤n

(αi − αj).

This proves (i).Before proving (ii) in complete generality, we prove it first for the basis 1, α, . . . , αn−1.

We recall that the conjugates α1, . . . , αn are distinct. This proves that ∆(1, α, . . . , αn−1) 6=0. Now permuting α1, . . . , αn permutes the rows of the determinant in (9) and soaffects only its sign. But ∆(1, α, . . . , αn−1) is the square of D(1, α, . . . , αn−1) andso it is a polynomial in α1, . . . , αn that is unaffected by permutations of α1, . . . , αn.We see that ∆(1, α, . . . , αn−1) ∈ Q. This proves (ii) for the basis 1, α, . . . , αn−1.

Now if β1, . . . , βn is some other basis for K over Q then

βi =n∑

j=1

cijαj−1

for some cij ∈ Q satisfyingdet(cij) 6= 0,

and (ii) follows immediately from the relationship

∆(β1, . . . , βn) = det(cij)2∆(1, α, . . . , αn−1).

Finally we turn to (iii). Suppose β1, . . . , βn is an integral basis. From (ii) thediscriminant of this basis is non-zero. We know that σi(βj) is a conjugate of βj .As βj ∈ O, so σi(βj). Hence from the definition of ∆ and the fact that O is a ringwe see that ∆(β1, . . . , βn) ∈ O. But (ii) tells us that ∆(β1, . . . , βn) ∈ Q. FromTheorem 4 we know that O ∩Q = Z. Thus ∆(β1, . . . , βn) ∈ Z as required.

We sketch a proof of (iv). Suppose that β1, . . . , βn and γ1, . . . , γn are bothintegral basis for OK . It follows from the theory of abelian groups that βi =∑n

j=1 cijγj where cij ∈ Z and det(cij) = ±1. So

∆(β1, . . . , βn) = det(cij)2∆(γ1, . . . , γn) = ∆(γ1, . . . , γn).

Definition. Let K a number field. We define the discriminant of K, denoted by∆K , to be the discriminant of any integral basis β1, . . . , βn for OK .

We note, by Theorem 20 that ∆K is independent of the choice of basis andmoreover it is a non-zero rational integer.

Corollary 17.1. Let K = Q(√d) where d is a square-free integer, and d 6= 0, 1

∆K =

{4d if d 6≡ 1 (mod 4)d if d ≡ 1 (mod 4).

Proof. In Theorem 19 we wrote down integral bases for OK . We merely have towrite down their discriminants. We will do the case d ≡ 1 (mod 4) and leave theother case as an exercise for the reader. Here we have 1, (1 +

√d)/2 as an integral

basis for OK . Now

D(1, (1 +

√d)/2

)=∣∣∣∣1 (1 +

√d)/2

1 (1−√d)/2

∣∣∣∣ = −√d.

Page 30: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

30 SAMIR SIKSEK

Hence

∆K = ∆(1, (1 +

√d)/2

)= D

(1, (1 +

√d)/2

)2

= d,

as required. �

Exercise 17.1. Let K = Q( 3√

2). Compute ∆K given that 1, 3√

2, ( 3√

2)2 is anintegral basis.

18. Ideals

We know that unique factorization fails for rings of integers of number fields. Forexample, in Z[

√−5] we can factorize 6 as a product of irreducibles in two different

ways:6 = 2 · 3 = (1 +

√−5)(1−

√−5).

It turns out that we can recover unique factorization for rings of integers of numberfields if we look at ideals instead of elements. What this means is that we will showthat every ideal can be written as a product of powers of prime ideals in a uniqueway.

We mostly use calligraphic letters for ideals of number fields: A, B, C, P, Q.Many books use gothic letters a, b, etc.

Let K be a number field and OK be its ring of integers. A non-empty subset Aof OK is an ideal if it satisfies the following conditions:

• A 6= ∅;• α−β ∈ A for every α, β ∈ A (this really says that A is an additive subgroup

of OK);• if α ∈ A and β ∈ OK then βα ∈ A.

The zero ideal is just {0}. Of course OK is an ideal of OK ; ideals that areproperly contained in OK are called proper ideals.

If 0 6= α ∈ OK we define the principal ideal generated by α be

αOK = {αr : r ∈ OK} .

Another common notation for αOK is (α). Of course (1) = OK . When we thinkof OK as an ideal it is usual to write it as (1).

In more generality, if α1, α2, . . . , αn are non-zero elements OK we define the idealgenerated by α1, . . . , αn to be

(α1, . . . , αn) =

{n∑

i=1

βiαi : β1, . . . , βn ∈ OK

}.

If A, B are ideals the so is the set

(A,B) = {α+ β : α ∈ A, β ∈ B} .

We sometimes write A+B for (A,B). We say that A, B are coprime if A+B = (1).We define the ideal product

AB =

{r∑

i=1

αiβi : αi ∈ A, βi ∈ B

}.

It is an easy exercise to show that AB is again an ideal.

Page 31: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

ALGEBRAIC NUMBER THEORY 31

18.1. Quotient Rings. Let A be an ideal of the ring OK . A coset of A is of theform

x+A = {x+ α : α ∈ A}.Recall that two cosets are equal x+A = y+A if and only if x− y = A. We definethe quotient

OK/A = {x+A : x ∈ A}.A priori OK/A is just the set of cosets of A, but we can make it into a ring bydefining addition and multiplication as follows:

(x+A) + (y +A) = (x+ y) +A, (x+A)(y +A) = xy +A.It is an easy exercise to show that these operations are well-defined and that theydo give a ring structure on OK/A.

We would like to prove that if A is a non-zero ideal then OK/A is finite. Beforewe can do this we need the following lemma.

Lemma 18.1. Suppose that A is a non-zero ideal of OK . Then A ∩ Z 6= {0}.

Proof. We want to show that A contains some non-zero element of Z. As A is anon-zero ideal, we can choose some α ∈ A ⊆ OK such that α 6= 0. Since α is analgebraic integer, its minimal polynomial fα(x) has integer coefficients; say

fα(x) = xn + an−1xn−1 + · · ·+ a0, ai ∈ Z.

Recall that fα is irreducible. If a0 = 0 then x | fα(x) and since fα(x) is irreduciblewe see that fα(x) = x. But this implies that α = 0 contradicting our choice of α.Hence a0 6= 0. Now fα(α) = 0, so

a0 = −αn − an−1αn−1 − · · · − a1α ∈ A,

showing that a0 ∈ A ∩ Z. �

Theorem 21. Let A be a non-zero ideal of OK . Then the quotient ring OK/A isfinite.

Proof. We know that OK has some Z-basis ω1, . . . , ωn. From Lemma 18.1 there issome non-zero m ∈ A ∩ Z. We would like to show that the quotient ring OK/A isfinite. We do this by constructing a map

φ : (Z/mZ)n −→ OK/A.Here Z/mZ are the integers modulo m (which you denoted by Zm in some othercourses). We also show that our map φ is surjective. This will be enough tocomplete the proof: since the domain is finite, the co-domain of this surjective mapmust be finite.

Defineφ(a1, . . . , an) = (a1ω1 + · · ·+ anωn) +A.

There is an issue here, which is to show that map is well-defined. This meansthat if a1, . . . , an and b1, . . . , bn are integers satisfying a1 = b1, . . . , an = bn then(a1ω1 + · · ·+ anωn) +A = (b1ω1 + · · ·+ bnωn) +A. Let’s do that. The conditiona1 = b1, . . . , an = bn means that

a1 − b1 = mc1, . . . , an − bn = mcn, for some c1, . . . , cn ∈ Z.But m ∈ A. Hence ai − bi ∈ A for i = 1, . . . , n and so is (ai − bi)ωi. Thus

(a1ω1 + · · ·+ anωn)− (b1ω1 + · · ·+ bnωn) = (a1 − b1)ω1 + · · ·+ (an − bn)ωn ∈ A,

Page 32: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

32 SAMIR SIKSEK

proving that (a1ω1 + · · ·+ anωn) +A = (b1ω1 + · · ·+ bnωn) +A. This shows thatφ is well-defined.

Since any x ∈ OK can be written as x = a1ω1 + · · ·+ anωn for some ai ∈ Z, wesee that x+A = φ(a1, . . . , an). Hence φ is surjective. �

Definition. Suppose that A is a non-zero ideal of OK . We define the norm of theideal A, denoted by N(A) to be the number of elements of the quotient ring OK/A.

Theorem 22. Let K be a number field of degree n and A a non-zero ideal of OK .Then A has a Z-basis consisting of n elements. Moreover, if δ1, . . . , δn is a Z-basisfor A and ω1, . . . , ωn is an integral basis for OK then

N(A) =∣∣∣∣ D(δ1, . . . , δn)D(ω1, . . . , ωn)

∣∣∣∣ .By a Z-basis consisting of n elements we mean some δ1, . . . , δn ∈ A such that

any element of α ∈ A can be written uniquely as a linear combination

α = a1δ1 + a2δ2 + · · ·+ anδn.

The proof of the theorem is essentially abelian group theory and we omit it.It is natural to ask what is the relationship between the norms of ideals and the

norms of elements of the field. Recall that if β ∈ K we defined the K-norm of β by

NK(β) =n∏

i=1

σi(β)

where σi are the embeddings of K.

Proposition 18.2. If β ∈ OK is non-zero and B = (β) is the principal idealgenerated by β then

N(B) = |NK(β)|.Proof. Suppose that ω1, . . . , ωn is an integral basis forOK . It is clear that βω1, . . . , βωn

is a Z-basis for B = (β) = βOK . Hence by Theorem 22 we have

N(B) =∣∣∣∣D(βω1, . . . , βωn)D(ω1, . . . , ωn)

∣∣∣∣ .But

D(βω1, . . . , βωn) =

∣∣∣∣∣∣∣∣∣σ1(βω1) σ1(βω2) . . . σ1(βωn)σ2(βω1) σ2(βω2) . . . σ2(βωn)

......

...σn(βω1) σn(βω2) . . . σn(βωn)

∣∣∣∣∣∣∣∣∣=

∣∣∣∣∣∣∣∣∣σ1(β)σ1(ω1) σ1(β)σ1(ω2) . . . σ1(β)σ1(ωn)σ2(β)σ2(ω1) σ2(β)σ2(ω2) . . . σ2(β)σ2(ωn)

......

...σn(β)σn(ω1) σn(β)σn(ω2) . . . σn(β)σn(ωn)

∣∣∣∣∣∣∣∣∣= σ1(β) · · ·σn(β)

∣∣∣∣∣∣∣∣∣σ1(ω1) σ1(ω2) . . . σ1(ωn)σ2(ω1) σ2(ω2) . . . σ2(ωn)

......

...σn(ω1) σn(ω2) . . . σn(ωn)

∣∣∣∣∣∣∣∣∣= NK(β)D(ω1, . . . , ωn)

Page 33: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

ALGEBRAIC NUMBER THEORY 33

proving that N(B) = |NK(β)|. �

Eventually we will show that norms are multiplicative. Meaning N(AB) =N(A)N(B). For now we have be content with a special case.

Theorem 23. Suppose that A, B are coprime non-zero ideals. Then

(i) A ∩ B = AB;(ii) (Chinese Remainder Theorem) we have an isomorphism

OK/AB ∼= OK/A×OK/B;

(iii) N(AB) = N(A)N(B).

Proof. We are supposing that A, B are coprime. This means that A + B = (1) orequivalently a+ b = 1 for some a ∈ A and b ∈ B. We know that AB ⊆ A ∩ B. Toprove (i) we must show that A ∩ B ⊆ AB. Equivalently, we must show that anyc ∈ A ∩ B satisfies c ∈ AB. Thus suppose that c ∈ A ∩ B. Remember that thereare elements a ∈ A and b ∈ B such that a + b = 1. Thus c = ac + bc. Now a ∈ Aand c ∈ A∩B ⊆ B, so ac ∈ AB. Similarly bc ∈ AB. Hence c = ac+ bc ∈ AB. Thisproves (i).

To prove (ii) consider the map

φ : OK −→ OK/A×OK/B, φ(c) = (c+A, c+ B).

It is easy to show that φ is a homomorphism of rings. Thus by the First IsomorphismTheorem

OK/Ker(φ) = Im(φ).

Let us calculate the kernel and image. Now

c ∈ Ker(φ) ⇐⇒ φ(c) = (0 +A, 0 + B)

⇐⇒ (c+A, c+ B) = (0 +A, 0 + B)⇐⇒ c ∈ A and c ∈ B⇐⇒ c ∈ A ∩ B.

Thus Ker(φ) = A ∩ B. By (i) A ∩ B = AB so we have Ker(φ) = AB. We wouldlike to show that φ is surjective. Suppose that (c1 +A, c2 + B) ∈ OK/A×OK/B.Let c = c1b + c2a where as above a + b = 1 and a ∈ A, b ∈ B. We claim thatφ(c) = (c1 +A, c2 + B). Note that

c− c1 = c1(b− 1) + c2a = c1(−a) + c2a = a(c2 − c1) ∈ A,

and similarly c− c2 ∈ B. Hence

φ(c) = (c+A, c+ B) = (c1 +A, c2 + B).

This shows that φ is surjective and so Im(φ) = OK/A × OK/B. Putting all thistogether proves (ii).

Now (iii) follows immediately from (ii) and the definition of norms of ideals. �

Page 34: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

34 SAMIR SIKSEK

19. Prime and Maximal Ideals

Definition. We call a proper ideal P prime, if for all α, β ∈ OK we have

αβ ∈ P =⇒ α ∈ P or β ∈ P.We call a proper ideal M maximal if there isn’t any ideal A satisfying

M ( A ( OK .

In words, a proper ideal is maximal if and only if it is not properly contained insome other proper ideal.

Theorem 24. Every maximal ideal is prime. An ideal P is prime if and only ifOK/P is an integral domain. An ideal M is maximal if and only if OK/M is afield. Every maximal ideal is prime.

Proof. The statements about OK/P are reformulations of the definitions of primeand maximal ideals. For the last statement, suppose that M is maximal. ThenOK/M is a field. But fields are integral domains, so M is prime. �

The above theorem tells us that maximal ideals are prime. The converse is nottrue in general, but it is true for non-zero prime ideals of the ring of integers ofa number field.

Theorem 25. Let K be a number field and OK be its ring of integers. A non-zeroideal of OK is prime if and only if it is maximal.

Proof. By Theorem 24 we know that every maximal ideal is prime. Let us prove theconverse. Suppose that P is a prime ideal of OK . Theorem 24 tells us that OK/Pis an integral domain. However, we know by Theorem 21 that OK/P is finite.In other words OK/P is a finite integral domain. But we recall that every finiteintegral domain is a field (if you’ve forgotten how this is proved, see Part I of thesenotes, or do the exercise below). Hence OK/P is a field. We apply Theorem 24again to deduce that P is maximal. �

Exercise 19.1. Let R be a finite integral domain. Prove that R is a field as follows:let x ∈ R\{0}. Consider that list x, x2, x3, . . .. Why must there be repetition inthis list? Use this to show that x must have an inverse. [This proof is in the styleof elementary number theory. The other proof you saw before is in the style ofcombinatorics.]

Exercise 19.2. Here is an alternative way of showing the maximal ideals areprime. Suppose that M is maximal and suppose that αβ ∈ M but α /∈ M. LetM′ = (α) +M. Show that M′ = (1). Deduce that β ∈M. Hence M is prime.

Before proceeding we need one more property of prime ideals.

Lemma 19.1. Suppose that A, B and P are ideals such that P is prime andP ⊇ AB. Then either P ⊇ A or P ⊇ B.

Perhaps you would like to prove this for yourself before looking at the proof.

Proof of Lemma 19.1. Suppose that AB ⊆ P but A 6⊆ P. Then there is someα ∈ A such that α /∈ P. We want to show that B ⊆ P. Consider β ∈ B. Thenαβ ∈ AB ⊆ P. By the definition of a prime ideal, α ∈ P or β ∈ P. But we knowalready that α /∈ P and so β ∈ P. Since this is true for all β ∈ P we deduce thatB ⊆ P as required. �

Page 35: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

ALGEBRAIC NUMBER THEORY 35

20. Towards Unique Factorization for Ideals I

Our objective is to prove the following theorem.

Theorem 26. (Unique Factorization Theorem for Ideals) Let K be a number fieldand OK be its ring of integers. Then every non-zero ideal A can be written as aproduct of of finitely prime ideals

A =n∏

i=1

Pi.

Moreover this factorization is unique up to re-ordering.

We note an important convention, which is that the ideal (1) is regarded as theproduct of zero many prime ideals:

(1) =0∏

i=1

Pi.

The proof of the Unique Factorization Theorem for Ideals will require many steps.One of these is the Cancellation Lemma which we prove later.

Lemma 20.1. (Cancellation Lemma) Let K be is a number field and OK be itsring of integers. Suppose that BA = CA for some non-zero ideals A, B and C.Then B = C.

The proof the Cancellation Lemma is highly non-trivial. You can’t just say“divide”, because we are talking about ideals (which are sets) and we haven’tdefined what division of ideals means.

However the Cancellation Lemma is enough to imply the uniqueness part of theUnique Factorization Theorem.

Lemma 20.2. (Uniqueness Part of the Unique Factorization Theorem) Let K bea number field and suppose the Cancellation Lemma holds for OK . Supposethat P1, . . . ,Pm and Q1, . . . ,Qn are prime ideals of OK . If

m∏i=1

Pi =n∏

j=1

Qj

then n = m and the P1, . . . ,Pm and Q1, . . . ,Qn are the same up to re-ordering.

Proof. We prove the lemma by induction on min(m,n). Suppose first that min(m,n) =0. Without loss of generality suppose that m = 0. If n = 0 then there is nothingto prove. So suppose that n > 0. Hence we have

(1) = Q1Q2 . . .Qn.

But Q1Q2 . . .Qn ⊆ Qi for i = 1, . . . , n, so Qi = (1). As prime ideals are proper bydefinition, we have a contradiction. Hence if min(m,n) = 0 then m = n = 0 andthe lemma is true.

We now come to the inductive step. Suppose min(m,n) ≥ 1. Note that

Pm ⊇m∏

i=1

Pi =n∏

j=1

Qj .

By Lemma 19.1 we see that Pm ⊇ Qj for some j. After re-labeling we can supposethat Pm ⊇ Qn. Now we recall that prime ideals are maximal (in our present

Page 36: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

36 SAMIR SIKSEK

context, not in general). Hence Pm = Qn. Now we apply the Cancellation Lemmato cancel Pm = Qn from

P1P2 . . .Pm = Q1Q2 . . .Qn

to obtainP1P2 . . .Pm−1 = Q1Q2 . . .Qn−1.

Now we can apply the inductive hypothesis to complete the proof. �

21. Towards Unique Factorization for Ideals II

In the previous section we showed that if we can factorize an ideal as a product ofprimes then such a factorization is unique up to re-ordering (of course we assumedthe Cancellation Lemma). To prove the Unique Factorization Theorem we must alsoprove existence: that any non-zero ideal can be written as a product of prime ideals.This is again a long-term goal. For now we introduce the notion of an irreducibleideal and prove the existence of prime factorization under the assumption thatirreducible ideals are prime.

Definition. We say a proper ideal is irreducible if it is not the product of twostrictly larger ideals.

Lemma 21.1. (Irreducibles are Primes) A non-zero ideal is prime if and only ifit is irreducible.

Showing that a prime ideal is irreducible is easy. Suppose P is prime and supposeP = AB. By Lemma 19.1 we see that P ⊇ A or P ⊇ B. Let’s say P ⊇ A. ThenP ⊇ A ⊇ AB = P. Hence P = A. Thus we see that A is not strictly bigger thanP. So we have shown that every prime ideal is irreducible. Showing the conversewill require much effort.

Lemma 21.2. (Existence of Factorization into Prime Ideals) Let K be a numberfield and OK be its ring of integers. Assume that irreducible ideals of OK areprime. Then every non-zero ideal can be written as a product of prime ideals.

Proof. Since we are assuming that irreducible ideals are prime, we need only showthat every non-zero ideal A can be written as a product of irreducible ideals. Wedo this by induction on the norm N(A).

Note first thatN(A) = 1 if and only ifOK/A = 0 which is equivalent toA = OK .But OK = (1) is the product of zero many irreducible ideals.

Now suppose that N(A) > 1 and we want to show that A can be written asa product of irreducible ideals. If A is irreducible then there is nothing to prove.Thus we may suppose that A is reducible, which means that A = BC where B andC are strictly larger ideals (strictly larger means A ( B and A ( C). We will showthat N(B) < N(A) and N(C) < N(A). After this we simply apply the inductivehypothesis to deduce that B and C can written as products of irreducibles and henceso can A = BC.

It remains to show that N(B) < N(A) and N(C) < N(A). To do this define themap

φ : OK/A → OK/B, α+A 7→ α+ Band show that this is well-defined and surjective using B ⊃ A (exercise). Moreovershow that the map is not injective using B ) A (exercise). Hence the cardinality

Page 37: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

ALGEBRAIC NUMBER THEORY 37

(i.e. number of elements) of OK/A is strictly larger than that of OK/B. Recallingthe definition of norm we see that N(A) > N(B) and similarly N(A) > N(C). �

22. Unique Factorization Proof—A Summary so Far

Our objective was (and is) to prove the Unique Factorization Theorem for Ideals(Theorem 26). We made two assumptions:

• The Cancellation Lemma;• Irreducible ideals are prime.

From these two assumptions we deduced the existence (Lemma 21.2) and uniquess(Lemma 20.2) parts of the Unique Factorization Theorem. In other words to com-plete the proof of the Unique Factorization Theorem ‘all’ we have to do is to proveour two assumptions. To do this we need to introduce ideal classes and show thatthere are finitely many ideal classes.

23. A Special Case of the Cancellation Lemma

The Cancellation Lemma 20.1 is one of our two ingredients needed to prove theUnique Factorization Theorem for ideals. In this section we content ourselves withproving a special case of the Cancellation Lemma. In turns out that this specialcase is needed in the proof of the full Cancellation Lemma.

Later on we prove the following statement for ideals of OK : to contain is todivide. What this means is the following: if A, B are non-zero ideals and B ⊇ A(i.e. B contains A) then BD = A for some non-zero ideal D (i.e. B divides A). Wecannot prove this yet, but we can prove a useful special case of this statement.

Lemma 23.1. Let C be a non-zero ideal and β be a non-zero element of OK . IfC ⊆ (β) then there is a non-zero ideal D such that C = (β)D.

Proof. Suppose C ⊆ (β). Define

D = {δ ∈ OK : δβ ∈ C}.

We want to show that D is an ideal that does the required job, namely C = (β)D.First we show that D is an ideal. Note that 0β ∈ C and so 0 ∈ D. Suppose δ1,

δ2 ∈ D. Then δ1β, δ2β ∈ C. Hence

(δ1 + δ2)β ∈ C,

showing that δ1 + δ2 ∈ D.Moreover, if δ ∈ D and α ∈ OK then δβ ∈ C and hence

(αδ)β = α(δβ) ∈ C.

Thus αδ ∈ D. This shows that D is an ideal.From the definition of D it is clear that (β)D ⊆ C. Suppose γ ∈ C. Since C ⊆ (β)

we can write γ = βδ for some δ ∈ OK . Again, from the definition of D we see thatδ ∈ D. Hence γ ∈ (β)D. This shows that C = (β)D. �

Exercise 23.1. This is an exercise on the theme “to contain is to divide”. Recallthat any ideal of Z is principal. Suppose that m, n ∈ Z\{0}. Show that (m) ⊇ (n)if and only if m | n.

Now for our special case of the Cancellation Lemma.

Page 38: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

38 SAMIR SIKSEK

Lemma 23.2. Suppose that β ∈ OK is non-zero and A, C are non-zero ideals ofOK . If (β)A = CA then (β) = C.

Proof. Let ω1, . . . , ωn be a Z-basis for the ideal A. Then βω1, . . . , βωn is a Z-basisfor βA = (β)A. Suppose that γ ∈ C. Then γωi ∈ CA = (β)A. Hence there areaij ∈ Z such that

γωi =n∑

j=1

aijβωj .

This system of equalties for i = 1, . . . , n can be rewritten in matrix notation as

β−1γw = Aw

where A = (aij) and w is the column vector with entries ωi. Hence β−1γ is aneigenvalue of the matrix A. In other words, β−1γ is a root of χA(x) = det(xIn−A).But A has entries in Z and so χA is monic with coefficients in Z. Thus β−1γ is analgebraic integer, and so β−1γ ∈ OK .

We deduce that γ ∈ βOK = (β). This is true for all γ ∈ C. Thus C ⊆ (β).Lemma 23.1 tells us that C = (β)D for some ideal D.

We return to our original equality: (β)A = CA and substitute C = (β)D to obtain(β)A = (β)DA or equivalently βA = βDA. Dividing by β we obtain A = DA.Recall that ωi was a Z-basis for A. It is easy to see that (Exercise)

ωi =∑

δijωj ,

where δij ∈ D. Thus 1 is an eigenvalue of the matrix (δij) and so is a root of thepolynomial

det(xIn − (δij)) = xn + dn−1xn−1 + · · ·+ d0,

where di ∈ D. From this we obtain

1 = −dn−1 − dn−2 − · · · − d0 ∈ D,

so D = (1). Substituting D = (1) in C = (β)D we obtain C = (β) as desired. �

24. Ideal Classes

Definition. Let K be a number field and OK be its ring of integers. We define thefollowing relation on the non-zero ideals of OK : A ∼ B if and only if (β)A = (α)Bfor some non-zero elements α, β ∈ OK . It is easy to show that this is an equivalencerelation and so when A ∼ B we say that A and B are in the same ideal class.We write [A] for the class represented by A. The principal class is the class of(1).

Exercise 24.1. Verify that ∼ really gives an equivalence relation on the non-zeroideals. Show that the principal class consists precisely of principal ideals.

Define multiplication on the set of ideal classes by [A][B] = [AB]. Show thatthis operation is well-defined. Eventually we will show that this operation turnsthe ideal classes into an abelian group. For now, which of the group axioms canyou prove straightforwardly from the definitions? What is the identity element?

Theorem 27. (Finiteness of Ideal Classes) Let K be a number field and OK be itsring of integers. There are only finitely many classes of ideals of OK .

Page 39: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

ALGEBRAIC NUMBER THEORY 39

We delay proving this theorem till later. For now we show that it implies theCancellation Lemma and that irreducible ideals are prime. In other words, we showthat this theorem implies the Unique Factorization Theorem for Ideals.

Lemma 24.1. Assume that OK has only finitely many ideal classes. Givenany non-zero ideal A, there is some exponent h > 0 such that Ah is principal.

Proof. First consider that sequence of ideals A,A2,A3, . . . . Since there are onlyfinitely many ideal classes, there must be some exponents r > s > 0 such thatAr ∼ As. Hence there are non-zero integers α, β such that

(α)Ar = (β)As.

Write C for the ideal αAr−s. Then

CAs = (β)As.

By our special case of the Cancellation Lemma 23.2 we have that C = (β). Hence(α)Ar−s = (β). Hence Ar−s is principal as required. �

We can now prove the Cancellation Lemma itself, assuming finiteness of idealclasses.

Lemma 24.2. (Cancellation Lemma) Let K be is a number field and OK be its ringof integers. Assume that OK has only finitely many ideal classes. Supposethat BA = CA for some non-zero ideals A, B and C. Then B = C.

Proof. As we are assuming that there are only finitely many ideal classes, we knowfrom Lemma 24.1 that Ah = (α) for some non-zero α ∈ OK and exponent h > 0.Multiplying both sides of BA = CA by Ah−1 we obtain B(α) = C(α), which we canrewrite as αB = αC. Dividing by α we obtain B = C. �

Lemma 24.3. (To Contain is to Divide) Let K be is a number field and OK beits ring of integers. Assume that OK has only finitely many ideal classes.If A ⊆ B are non-zero ideals, then there is a non-zero ideal D such that A = BD.

Proof. As we are assuming that there are only finitely many ideal classes, we knowfrom Lemma 24.1 that Bh = (β) for some non-zero β ∈ OK and exponent h > 0.From A ⊆ B we obtain ABh−1 ⊆ Bh and so ABh−1 ⊆ (β). By Lemma 23.1 thereis some non-zero ideal D such that

ABh−1 = (β)D.

Substituting back (β) = Bh we get

ABh−1 = DBh.

By the Cancellation Lemma 24.2 we can cancel Bh−1 to obtain A = DB as required.�

Lemma 24.4. (Irreducibles are Primes) Let K be is a number field and OK be itsring of integers. Assume that OK has only finitely many ideal classes. Anon-zero ideal is prime if and only if it is irreducible.

Proof. We proved before that prime ideals are irreducible. Suppose that P is anon-zero irreducible ideal, we would like to show that it is prime. It is enough

Page 40: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

40 SAMIR SIKSEK

to show that P is maximal (since maximal ideals are prime). Suppose P is notmaximal. Then there is some ideal A such that

P ( A ( OK .

By Lemma 24.3 we see that P = AD for some non-zero ideal D. We know thatP ( A and also P = AD ⊆ D.

Now P = AD is irreducible. By definition of irreducible, P is not properlycontained in A or not properly contained in D. However P ( A. Hence P isnot properly contained in D. Thus P = D. Cancelling from P = AD we getA = (1) = OK contradicting A ( OK . �

25. Unique Factorization Proof—Summary So Far (again)

Recall the conclusion of our previous summary in Section 22: to complete theproof of the Unique Factorization Theorem for Ideals (Theorem 26) we needed toprove two results:

• The Cancellation Lemma;• Irreducible ideals are prime.

In the previous section we proved both of these results under the assumption thatthere ring of integers of a number field has only finitely many ideal classes. Thusto complete the proof of the Unique Factorization Theorem for Idealsall we have to do is to prove the finiteness of ideal classes.

26. What are the prime ideals of K?

In this section we take a well-earned rest from the proof of the proof of theUnique Factorization Theorem to ask ourselves, “What do the prime ideals of ringsof integers of number fields look-like?”. We assume the Unique Factoriza-tion Theorem, Cancellation Lemma, to Contain is to Divide, IrreducibleIdeals are Prime throughout this section. Later on we return to complete theproof of all of these once we had our rest.

Lemma 26.1. Suppose P is a non-zero prime ideal of OK . Then P ∩ Z = pZfor some prime number p. Moreover, P divides pOK (the principal ideal of OK

generated by p).

Proof. It is easy to see that P ∩Z is an ideal of Z. Moreover, we proved previouslythat for any non-zero ideal A we have A∩Z 6= 0. Hence P ∩Z is a non-zero ideal ofZ. We recall that Z is a principal ideal domain. So P ∩ Z = pZ for some non-zerointeger p, which we may take to be positive. We want to show that p is prime.Suppose p | ab where a, b are positive integers. Then ab ∈ pZ ⊆ P ∩ Z ⊆ P. AsP is prime, we see that a ∈ P or b ∈ P. But a, b ∈ Z. Hence, a ∈ P ∩ Z = pZ orb ∈ P ∩ Z = pZ. In otherwords p | a or p | b. This shows that p is prime.

Thus we have shown P∩Z = pZ for some prime p. Now p ∈ P. Hence pOK ⊆ P.By “to contain is to divide” we see that P | pOK . �

We see from the above lemma that any non-zero prime ideal of OK divides pOK

for some prime number p. To get the prime ideals of OK all we have to do isfactorize pOK as a product of prime ideals. The following theorem tells us how,provided OK = Z[α] for some α ∈ OK .

Page 41: MA3A6 ALGEBRAIC NUMBER THEORY - Warwick Insite › ~maseap › teaching › ant › antnot… · Algebraic number theory arose out of the study of Diophantine equations. A Diophantine

ALGEBRAIC NUMBER THEORY 41

Theorem 28. Suppose that OK = Z[α] where α has minimal polynomial fα(X) ∈Z[X] of degree n. Suppose that

(10) fα(X) ≡∏

gi(X)ei (mod p)

where each gi is a monic polynomial in Z[X] which is irreducible modulo p. WritePi = (p, gi(α)). Then each ideal Pi is prime and

pOK =∏

Peii .

Example 26.1. Let us take K = Q(i) and see the factorizations of some smallprimes in OK (the Gaussian integers). We know OK = Z[i], and i has minimalpolynomial f(X) = X2 + 1. Let us factorize the ideals 2OK , 3OK , 5OK in OK .We note

X2 + 1 ≡ (X + 1)2 (mod 2).Hence 2OK = P2 where P = (2, 1 + i) is prime. But we remember that Z[i] is aprincipal ideal domain. So we should be able to write P as a principal ideal. Nownotice that 2 = (1 + i)(1− i). Hence 2 ∈ (1 + i)OK . Thus

(1 + i)OK ⊆ (2, 1 + i) ⊆ (1 + i)OK .

Hence P = (1 + i)OK and 2OK = P2.To factorize 3, we note that X2 + 1 is irreducible modulo 3. Hence 3OK =

(3, i2 + 1) = (3) is a prime ideal.To factorize 5, note that

X2 + 1 ≡ X2 − 4 ≡ (X + 2)(X − 2) (mod 5).

Hence 5OK = PQ where P = (5, 2+i) and Q = (5, 2−i). However 5 = (2+i)(2−i)so P = (2 + i)OK and Q = (2− i)OK .

Samir Siksek, Department of Mathematics, University of Warwick, Coventry, CV4

7AL, United Kingdom

E-mail address: [email protected]