Introductory Algebraic Number Theory
Introductory Algebraic Number Theory
ii
CB609-driver CB609/Alaca & Williams August 27, 2003 17:1 Char Count= 0
Algebraic number theory is a subject that came into being through the attempts of mathe-
maticians to try to prove Fermat’s last theorem and that now has a wealth of applications
to Diophantine equations, cryptography, factoring, primality testing, and public-key cryp-
tosystems.
This book provides an introduction to the subject suitable for senior under-
graduate and beginning graduate students in mathematics. The material is presented in
a straightforward, clear, and elementary fashion, and the approach is hands on, with an
explicit computational flavor. Prerequisites are kept to a minimum, and numerous examples
illustrating the material occur throughout the text. References to suggested readings and to
the biographies of mathematicians who have contributed to the development of algebraic
number theory are given at the end of each chapter. There are more than 320 exercises, an
extensive index, and helpful location guides to theorems and lemmas in the text.
Şaban Alaca is Lecturer in Mathematics at Carleton University, where he has been honored by three
teaching awards: Faculty of Science Teaching Award, Professional Achievement Award, and Students
Choice Award. His main research interest is in algebraic number theory.
i
CB609-driver CB609/Alaca & Williams August 27, 2003 17:1 Char Count= 0
ii
CB609-driver CB609/Alaca & Williams August 27, 2003 17:1 Char Count= 0
ŞABAN ALACA
Carleton University, Ottawa
KENNETH S. WILLIAMS
Carleton University, Ottawa
iii
cambridge university press
Cambridge, New York, Melbourne, Madrid, Cape Town, Singapore, São Paulo
Cambridge University Press has no responsibility for the persistence or accuracy of urls
for external or third-party internet websites referred to in this publication, and does not
guarantee that any content on such websites is, or will remain, accurate or appropriate.
CB609-driver CB609/Alaca & Williams August 27, 2003 17:1 Char Count= 0
To our wives
Ayşe and Carole
v
CB609-driver CB609/Alaca & Williams August 27, 2003 17:1 Char Count= 0
vi
CB609-driver CB609/Alaca & Williams August 27, 2003 17:1 Char Count= 0
Contents
vii
CB609-driver CB609/Alaca & Williams August 27, 2003 17:1 Char Count= 0
viii Contents
Suggested Reading 72
Biographies 73
4 Elements Integral over a Domain 74
4.1 Elements Integral over a Domain 74
4.2 Integral Closure 81
Exercises 86
Suggested Reading 87
Biographies 87
5 Algebraic Extensions of a Field 88
5.1 Minimal Polynomial of an Element Algebraic over a Field 88
5.2 Conjugates of α over K 90
5.3 Conjugates of an Algebraic Integer 91
5.4 Algebraic Integers in a Quadratic Field 94
5.5 Simple Extensions 98
5.6 Multiple Extensions 102
Exercises 106
Suggested Reading 108
Biographies 108
6 Algebraic Number Fields 109
6.1 Algebraic Number Fields 109
6.2 Conjugate Fields of an Algebraic Number Field 112
6.3 The Field Polynomial of an Element of an Algebraic Number
Field 116
6.4 The Discriminant of a Set of Elements in an Algebraic Number
Field 123
6.5 Basis of an Ideal 129
6.6 Prime Ideals in Rings of Integers 137
Exercises 138
Suggested Reading 140
Biographies 140
7 Integral Bases 141
7.1 Integral Basis of an Algebraic Number Field 141
7.2 Minimal Integers 160
7.3 Some Integral Bases in Cubic Fields 170
7.4 Index and Minimal Index of an Algebraic Number Field 178
7.5 Integral Basis of a Cyclotomic Field 186
Exercises 189
Suggested Reading 191
Biographies 193
8 Dedekind Domains 194
8.1 Dedekind Domains 194
8.2 Ideals in a Dedekind Domain 195
CB609-driver CB609/Alaca & Williams August 27, 2003 17:1 Char Count= 0
Contents ix
x Contents
List of Tables
√
1 Integral bases and discriminants for Q( 3 k), 2 ≤ k ≤ 20,
k cubefree. √ page 177
2 Integral bases and discriminants for Q( 4 k),
x 4 − k irreducible in Q[x], 2 ≤ k ≤ √10. 177
3 Integral bases and discriminants for Q( 4 −k),
x 4 + k irreducible in Q[x], 1 ≤ k ≤ 10. 178
4 Fundamental units of OQ(√m) , 2 ≤ √ m < 40, m squarefree. 280
5 Nontrivial ideal class groups H (Q( k)), − 30 < k < 0,
k squarefree. √ 322
6 Nontrivial ideal class groups H (Q( k)), 2 ≤ k < 100,
k squarefree. √ 323
7 Class numbers of imaginary quadratic fields K = Q( k),
−195 ≤ k < 0, k squarefree. √ 325
8 Class numbers of real quadratic fields K = Q( k),
0 < k ≤ 197, k squarefree.
√ 326
9 Class numbers of Q( 3 k), 2 ≤ k ≤ 101, k cubefree. 329
10 Class numbers of cyclotomic fields K m , 3 ≤ m ≤ 45,
m ≡ 2 (mod 4). 331
√
11 Fundamental unit (> 1) of Q( 3 m) for a few values
of m ∈ N. 375
12 Fundamental unit of cubic fields K with exactly one real
embedding and −268 ≤ d(K ) < 0. 376
13 Units of totally real cubic fields K with 0 < d(K ) ≤ 1101. 377
14 Fundamental unit of some pure quartic
√
fields Q( 4 −m). 378
15 Solutions (x, y) ∈ Z2 of y 2 = x 3 + k, − 20 ≤ k < 0. 402
16 Solutions (x, y) ∈ Z2 of y 2 = x 3 + k, 0 < k ≤ 20. 403
xi
CB609-driver CB609/Alaca & Williams August 27, 2003 17:1 Char Count= 0
xii
CB609-driver CB609/Alaca & Williams August 27, 2003 17:1 Char Count= 0
Notation
N = {1, 2, 3, . . .}
Z = {0, ±1, ±2, . . .}
Q = field of rational numbers
R = field of real numbers
C = field of complex numbers
φ = empty set
1, if p m and x ≡ m (mod p) is solvable,
2
m
= Legendre symbol = −1, if p m and x 2 ≡ m (mod p) is insolvable,
p
0, if p | m,
where m ∈ Z and p is a prime
[x]
= greatest integer less than or equal to the real number x
m m!
= binomial coefficient = , where m and n are integers such that 0 ≤ n ≤ m
n (m − n)!n!
∗
If A is a set containing 0 then A = A \ {0}
Zn = cyclic group of order n
card(S) = cardinality of the set S
On = n × n zero matrix
In = n × n identity matrix
Or,s = r × s zero matrix
xiii
CB609-driver CB609/Alaca & Williams August 27, 2003 17:1 Char Count= 0
xiv
CB609-driver CB609/Alaca & Williams August 27, 2003 17:1 Char Count= 0
Introduction
This book is intended as an introductory text for senior undergraduate and beginning
graduate students wishing to learn the fundamentals of algebraic number theory. It
is based upon a course in algebraic number theory given by the second author at
Carleton University for more than thirty years. Keeping in mind that this is an intro-
ductory text, the authors have strived to present the material in as straightforward,
clear, and elementary fashion as possible. Throughout the text many numerical ex-
amples are given to illustrate the theory. Each chapter closes with a set of exercises
on the material covered in the chapter, as well as some suggested further reading.
References cited in each chapter are listed under suggested reading. Biographical
references for some of the mathematicians mentioned in the text are also given at
the end of each chapter. For the convenience of the reader, the book concludes with
page references for the definitions, theorems, and lemmas in the text. In addition
an extensive bibliography of books on algebraic number theory is provided.
The main aim of the book is to present to the reader a detailed self-contained
development of the classical theory of algebraic numbers. This theory is one of
the crowning achievements of nineteenth-century mathematics. It came into being
through the attempts of mathematicians of that century to prove Fermat’s last the-
orem, namely, that the equation x n + y n = z n has no solutions in nonzero integers
x, y, z, where n is an integer ≥ 3. A wonderful achievement of the twentieth century
was the proof of Fermat’s last theorem by Andrew Wiles of Princeton University.
Although the proof of Fermat’s last theorem is beyond the scope of this book, we
will show how algebraic number theory can be used to find the solutions in integers
(if any) of other equations.
The contents of the book are divided into fourteen chapters. Chapter 1 serves as
an introduction to the basic properties of integral domains. Chapters 2 and 3 are
devoted to Euclidean domains and Noetherian domains respectively. In Chapter 4
the reader is introduced to algebraic numbers and algebraic integers. Algebraic
number fields are introduced in Chapter 6 after a discussion of algebraic extensions
of fields in Chapter 5. Chapter 7 is devoted to the study of integral bases. Minimal
integers are introduced as a tool for finding integral bases and many numerical
xv
CB609-driver CB609/Alaca & Williams August 27, 2003 17:1 Char Count= 0
xvi Introduction
examples are given. Chapter 8 is concerned with Dedekind domains. The ring
of integers of an algebraic number field is the prototype of a Dedekind domain.
Chapters 9 and 10 discuss the factorization of ideals into prime ideals. The structure
of the unit group of a real quadratic field is determined in Chapter 11. In Chapter
12 the classic theorems of Minkowski in the geometry of numbers are proved and
are used to show that the ideal class group is finite. Dirichlet’s determination of the
units in an arbitrary algebraic number field is presented in Chapter 13 using the
approach given by van der Waerden. Finally, in Chapter 14, the algebraic number-
theoretic tools developed in earlier chapters are used to discuss the solvability of
certain equations in integers.
The prerequisites for this book are a basic course in linear algebra (systems
of linear equations, vector spaces over a field), a basic course in modern algebra
(groups, rings, and fields including Eisenstein’s irreducibility criterion), and a basic
course in elementary number theory (the Legendre symbol, quadratic residues, and
the law of quadratic reciprocity.) No Galois theory is needed.
A possible outline for a one-semester course (three hours of lectures per week
for twelve weeks) together with an approximate breakdown of lecture time is as
follows:
Chapter 1 (excluding Theorem 1.2.2) 2 hours
Chapter 2 (excluding Sections 2.3, 2.4) 2 hours
Chapter 3 3 hours
Chapter 4 3 hours
Chapter 5 3 hours
Chapter 6 5 hours
Chapter 7 (Section 7.1 only) 3 hours
Chapter 8 3 hours
Chapter 9 3 hours
Chapter 10 (excluding Sections 10.4, 10.5, 10.6) 2 hours
Chapter 11 3 hours
Chapter 12 (excluding Section 12.7) 2 hours
Chapter 14 (Section 14.2 only) 2 hours
The authors would like to thank their colleagues John D. Dixon, James G. Huard,
Pierre Kaplan, Blair K. Spearman, and P. Gary Walsh for helpful suggestions in
connection with the writing of this book. The second author would like to thank the
many students who have taken the course Mathematics 70.436*/70.536 Algebraic
Number Theory with him at Carleton University over the years. Special thanks go
CB609-driver CB609/Alaca & Williams August 27, 2003 17:1 Char Count= 0
Introduction xvii
xviii
CB609-01 CB609/Alaca & Williams August 7, 2003 17:16 Char Count= 0
1
Integral Domains
Example 1.1.1 The ring Z = {0, ±1, ±2, . . .} of all integers is an integral domain.
Z + Z 1+ 2 −3 .
√ √
Example 1.1.4 Z + Z m = {a + b m | a, b ∈ Z}, where m is a positive or
√
negative integer that is not a perfect square, is an integral domain. As m is a
√
root of an irreducible quadratic polynomial (namely x 2 − m), Z + Z m is called
1
CB609-01 CB609/Alaca & Williams August 7, 2003 17:16 Char Count= 0
2 Integral Domains
Example 1.1.6 F[x] = the ring of polynomials in the indeterminate x with coef-
ficients from a field F is an integral domain.
Example 1.1.7 Z[x] = the ring of polynomials in the indeterminate x with integral
coefficients is an integral domain.
Example 1.1.8 D[x] = the ring of polynomials in the indeterminate x with coef-
ficients from the integral domain D is an integral domain.
Example 1.1.9 F[x, y] = the ring of polynomials in the two indeterminates x and
y with coefficients from the field F is an integral domain.
ab = ac, a = 0 =⇒ b = c (a, b, c ∈ D)
(c) It is well known that if D is an integral domain then there exists a field F, called the
field of quotients of D or the quotient field of D, that contains an isomorphic copy D
of D (see, for example, Fraleigh [3]). In practice it is usual to identify D with D and so
consider D as a subdomain of F. The quotient field of Z is the field of rational numbers
Q. The quotient field of the polynomial domain F[X ] (where F is a field) is the field
F(X ) of rational functions in X .
Definition 1.1.2 (Divisor) Let a and b belong to the integral domain D. The element
a is said to be a divisor of b (or a divides b) if there exists an element c of D such
that b = ac. If a is a divisor of b, we write a | b. If a is not a divisor of b, we write
a b.
Example 1.1.15 1 + θ − θ 2 | − θ − 2θ 2 in Z + Zθ + Zθ 2 as −θ − 2θ 2 = (1 +
θ − θ 2 )(1 − θ) (see Example 1.1.10).
√ √ √ √ √
Example 1.1.16 2 + 2 3 in Z + Z 2 as 3/(2 + 2) = 3 − 32 2 ∈
Z + Z 2.
Properties of Divisors
Let a, b, c ∈ D, where D is an integral domain. Then the following properties hold.
(a) a | a (reflexive property).
(b) a | b and b | c implies a | c (transitive property).
CB609-01 CB609/Alaca & Williams August 7, 2003 17:16 Char Count= 0
4 Integral Domains
Properties of Units
Let D be an integral domain. Then U (D) has the following properties.
(a) ±1 ∈ U (D).
(b) If a ∈ U (D) then −a ∈ U (D).
(c) If a ∈ U (D) then a −1 ∈ U (D).
(d) If a ∈ U (D) and b ∈ U (D) then ab ∈ U (D).
(e) If a ∈ U (D) then ±a n ∈ U (D) for any n ∈ Z.
Example 1.1.17
(a) i ∈ U (Z + Zi).
(b) ω ∈ U (Z + Zω) (see Example 1.1.3).
(c) θ ∈ U (Z + Zθ + Zθ 2 ) as 1 = θ (−1 − θ 2 ) (see Example 1.1.10).
Abelian groups are named after the Norwegian mathematician Niels Henrik Abel
(1802–1829), who proved in 1824 the impossibility of solving the general quintic
equation by means of radicals.
√= n{±1} Z2 . √
(d) U (Z[x])
(e) ±(1√ + 2)√ ∈ U (Z + Z 2), for all n ∈ Z.
(f) 12 2 + 12 i 2 ∈ U (D), where D is defined in Example 1.1.11.
√ √
U (Z + Z 2) = {±(1 + 2)n | n ∈ Z} Z2 × Z.
Properties of Associates
Let a, b, c ∈ D ∗ = D \ {0}, where D is an integral domain. The following proper-
ties hold.
(a) a ∼ a (reflexive property).
(b) a ∼ b implies b ∼ a (symmetric property).
(c) a ∼ b and b ∼ c imply a ∼ c (transitive property).
(d) a ∼ b if and only if ab−1 ∈ U (D).
(e) a ∼ 1 if and only if a is a unit.
Properties (a), (b), and (c) show that ∼ is an equivalence relation. The equivalence
class containing a ∈ D is just the set {ua | u ∈ U (D)}.
Example 1.1.19
(a) In Z, a ∼ b if and only if a = ±b, equivalently |a| = |b|.
(b) In Z + Zi we have 1 + i ∼ 1 − i as 1+i = i ∈ U (Z + Zi).
√ √ 1−i
√ √ √ √
(c) In Z + Z 2 we have 1 + 3 2 ∼ 5 − 2 2 as 1+3 √2 = 1 +
5−2 2
2 ∈ U (Z + Z 2).
6 Integral Domains
a 2 + 5b2 = 1, 2, or 4.
so that
√
a + b −5 = ±1 or ± 2.
√ √
In the former case a + b −5 is a unit of Z + Z −5. In the latter case
√ 2 2
c + d −5 = √ = = ±1
a + b −5 ±2
√ √
is a unit of Z + Z −5. Hence 2 is irreducible in Z + Z −5.
√ √
Example 1.2.3 7 + −5 is reducible in Z + Z −5 because
√ √ √
7 + −5 = (1 + −5)(2 − −5)
√ √ √
and neither 1 + −5 nor 2 − −5 is a unit of Z + Z −5.
Our next definition generalizes property (1.2.2) to an arbitrary integral domain, and
an element with this property is called a prime element.
(a 2 + b2 )(c2 + d 2 ) = 2(x 2 + y 2 ).
Theorem 1.2.2 Let D be an integral domain that has the following property:
8 Integral Domains
Proof: Let p be an irreducible element in D, which is not prime. Then there exist
a, b ∈ D such that
p | ab, p a, p b.
f (X ) = p X 2 − (a + b)X + r.
In F[X ] we have
We show that f (X ) does not factor into linear factors in D[X ]. Indeed, suppose on
the contrary that
f (X ) = (cX + s)(d X + t)
1.3 Ideals
Subsets of an integral domain D that are closed under addition and under multipli-
cation by elements of D play a special role and are called ideals.
a ∈ I, b ∈ I =⇒ a + b ∈ I,
a ∈ I, r ∈ D =⇒ ra ∈ I.
ri ai | r1 , . . . , rn ∈ D
i=1
1.3 Ideals 9
Proof: If a/b ∈ U (D) then a = bu for some u ∈ U (D). Let x ∈ a. Then x = ac
for some c ∈ D. Hence x = buc with uc ∈ D. Thus x ∈ b. We have shown that
a ⊆ b. As a/b ∈ U (D) and U (D) is a group with respect to multiplication, we
have b/a = (a/b)−1 ∈ U (D). Then, proceeding exactly as before with the roles of
a and b interchanged, we find that b ⊆ a. Thus a = b.
Conversely, suppose that a = b. Then a = bc for some c ∈ D and b = ad for
some d ∈ D. Hence b = bcd. As b = 0 we deduce that 1 = cd so that c ∈ U (D).
Thus a/b = c ∈ U (D).
CB609-01 CB609/Alaca & Williams August 7, 2003 17:16 Char Count= 0
10 Integral Domains
Theorems 1.3.1 and 1.4.1 show that the set of ideals of Z is {kZ | k ∈
{0, 1, 2, . . .}}. Moreover, if I is an ideal of Z then it is generated by the least
positive integer in I .
Other examples of principal ideal domains will be given in Chapter 2 where we
discuss Euclidean domains.
Example √ 1.4.1 It was noted in Section 1.2 that 2 is irreducible but not √ prime
in Z + Z −5. Hence, by Theorem 1.4.3, the integral domain √ Z + Z −5
√ is
not a principal ideal domain. Indeed the ideal 2, 1 + −5 of Z + Z −5
is not principal. This can be shown √ directly as follows. Suppose,√on the
contrary, that the ideal√2, 1 + −5 is principal, that √ is, 2, 1 + −5 =
α for some√α ∈ Z + Z −5. Hence 2 ∈ α and 1 + −5 ∈ α √ so that α | 2
and α | 1 + −5. From the first of these, as 2 is irreducible
√ in Z + Z −5, it must
be the√ case that α √∼ 1 or α ∼ 2. √If α ∼ 2 then 2 | 1 + −5, which √ is impossible
1+ −5
as 2 = 2 + 2 −5 ∈ Z + Z −5. Hence α ∼ 1, and
1 1
√ so 2, 1 + −5 = 1.
This shows
√ that 1 is a linear combination of 2 and 1 + −5 with coefficients from
Z + Z −5; that is, there exist x, y, z, w ∈ Z such that
√ √ √
1 = (x + y −5)2 + (z + w −5)(1 + −5).
√
Equating coefficients of 1 and −5, we obtain
1 = 2x + z − 5w, 0 = 2y + z + w.
The difference of these equations yields
1 = 2(x − y − 3w),
which is clearly impossible as the left-hand side is an odd
√ integer and the right-
√ is an even integer. Hence the ideal 2, 1 + −5 is not principal in
hand side
Z + Z −5.
12 Integral Domains
so that
a = r 1 a1 + · · · + r n an
for some r1 , . . . , rn ∈ D. Thus if c ∈ D is such that
c | a j ( j = 1, 2, . . . , n)
then
c | a.
Moreover, for j = 1, 2, . . . , n, we have
a j ∈ a1 , . . . , an = a
so that
a | aj.
This justifies calling a “a greatest common divisor” of a1 , . . . , an . The elements
a1 , . . . , an are called relatively prime if (a1 , . . . , an ) is a unit, that is,
a1 , . . . , an = 1 = D.
It is easy to verify that
(a1 , . . . , an−1 , an ) = ((a1 , . . . , an−1 ), an ),
so that a greatest common divisor can be obtained by finding a succession of greatest
common divisors of pairs of elements, that is, if (a1 , a2 ) = b then (a1 , a2 , a3 ) =
(b, a3 ), etc.
In the next theorem we use our knowledge of primes and irreducibles in a principal
ideal domain to give conditions under which a prime p can be expressed as u 2 − mv 2
or mv 2 − u 2 for some integers u and v, where m is a given nonsquare integer.
√
Theorem 1.4.4 Let m be a nonsquare integer such that Z + Z m is a principal
ideal domain. Let p be an odd prime for which the Legendre symbol
m
= 1.
p
Then there exist integers u and v such that
p = u 2 − mv 2 if m < 0, or if m > 0,
and there are integers T, U such that T 2 − mU 2 = −1,
p = u 2 − mv 2 or mv 2 − u 2 , if m > 0,
and there are no integers T, U with T 2 − mU 2 = −1.
CB609-01 CB609/Alaca & Williams August 7, 2003 17:16 Char Count= 0
√
In Chapter 2 we give some nonsquare values of m for which Z + Z m is a
principal ideal
domain.
Then, by Theorem 1.4.4, we know that for those odd primes
p for which p = 1 there are integers u and v such that p = u 2 − mv 2 or mv 2 −
m
14 Integral Domains
Proof: As m
p
= 1 there exists an integer z such that z 2 ≡ −m (mod p). Set
z, if z is odd,
y=
p − z, if z is even,
Examples illustrating Theorems 1.4.4 and 1.4.5 are given in Section 2.5.
CB609-01 CB609/Alaca & Williams August 7, 2003 17:16 Char Count= 0
16 Integral Domains
The next example shows that the converse of Theorem 1.5.1 is not true in general.
Example 1.5.3
(a) x is an irreducible element of Z[x] but x is not a maximal ideal of Z[x] as
x ⊂ 2, x ⊂ Z[x].
√ √ √
(b) 1 + −5 is an irreducible element of Z + Z −5 but 1 + −5 is not a maximal
√
ideal of Z + Z −5 as
√ √ √
1 + −5 ⊂ 2, 1 + −5 ⊂ Z + Z −5.
18 Integral Domains
I ⊂ J ⊆ D.
(b + I )(c + I ) = 1 + I.
Thus
bc + I = 1 + I
and so
bc − 1 ∈ I ⊂ J.
bc ∈ J.
Hence
1 = bc − (bc − 1) ∈ J,
(b + I )(y + I ) = by + I = 1 − w + I = 1 + I
a, b ∈ D and ab ∈ P implies a ∈ P or b ∈ P.
CB609-01 CB609/Alaca & Williams August 7, 2003 17:16 Char Count= 0
20 Integral Domains
Proof: Suppose first that D/I is an integral domain and that a, b ∈ D are such that
ab ∈ I . Then (a + I )(b + I ) = ab + I = 0 + I , the zero element of the integral
domain D/I . Because an integral domain has no divisors of zero, we have a + I =
0 + I or b + I = 0 + I ; that is, we have either a ∈ I or b ∈ I , so that I is prime.
Now suppose that I is a prime ideal of D. As I is a proper ideal of D, D/I is a
commutative ring with identity 1 + I . Thus we have only to check that when I is
prime, D/I has no divisors of zero. Suppose that a + I ∈ D/I and b + I ∈ D/I
are such that (a + I )(b + I ) = 0 + I . Then ab + I = I , so that ab ∈ I . As I is
prime, either a ∈ I or b ∈ I ; that is, a + I = 0 + I or b + I = 0 + I , so D/I has
no zero divisors.
The next example shows that the converse of Theorem 1.5.6 is not true in general.
Example 1.5.6 x is a prime ideal of Z[x], but it is not a maximal ideal of Z[x].
Proof: In view of Theorem 1.5.6 we have only to show that if I is a prime ideal of
D then I is a maximal ideal.
Suppose that I is a prime ideal of D that is not maximal. Then there exists an
ideal J of D such that
I ⊂ J ⊂ D.
As D is a principal ideal domain, we have I = a and J = b for some a, b ∈ D.
As a ⊂ b we have a = bc for some c ∈ D. Now bc = a ∈ a = I , and I is
prime, so that either b ∈ I or c ∈ I . If b ∈ I then J = b ⊆ I ⊂ J , which is a
contradiction. Hence c ∈ I . Thus c = ad for some d ∈ D, and so a = bda. But
CB609-01 CB609/Alaca & Williams August 7, 2003 17:16 Char Count= 0
I + J = {i + j | i ∈ I, j ∈ J }.
It is readily checked that I + J is also an ideal and that it is the minimal ideal
containing both I and J . The following properties are also easily checked: For
ideals I, J, K of the integral domain D
I + J = J + I,
(I + J ) + K = I + (J + K ),
I + 0 = I,
I + 1 = 1.
Further, if I = i and J = j are principal ideals, then I + J = i, j. It is easy
to extend Definition 1.6.1 to the sum of a finite number of ideals.
Next we define the product of two ideals.
I J = {x ∈ D | x = i 1 j1 + · · · + ir jr for some r ∈ N,
some i 1 , . . . , ir ∈ I, and some j1 , . . . , jr ∈ J }.
Clearly I J is the set of all finite sums of products of elements of I and J , and it is
easily checked that I J is an ideal. The following properties are also easily verified:
For ideals I, J, K of the integral domain D
I J = J I,
(I J )K = I (J K ),
I 0 = 0,
I 1 = I.
22 Integral Domains
(I + J )K = I K + J K .
Example 1.6.1 Let m and n be integers that are not both zero. Set d = (m, n), the
greatest common divisor of m and n. We show that
Let a ∈ m + n. Then there exist integers r and s such that a = r m + sn. As
d = (m, n) there exist coprime integers m 1 and n 1 such that m = dm 1 , n = dn 1 .
Thus a = r dm 1 + sdn 1 = (r m 1 + sn 1 )d ∈ d. This shows that m + n ⊆ d.
Now let a ∈ d, so that there exists an integer e such that a = de. As d = (m, n)
there exist integers x and y such that d = xm + yn. Hence a = (xm + yn)e =
(xe)m + (ye)n ∈ m + n. This proves that d ⊆ m + n.
The two inclusions show that m + n = d.
Next we give another necessary and sufficient condition for a proper ideal to be
a prime ideal.
A ⊆ P, B ⊆ P, AB ⊆ P.
Our final theorem of this chapter shows that a prime ideal P of an integral domain
D1 remains prime when restricted to a subdomain D of D1 .
Exercises 23
Exercises
1. Prove that U (Z + Zi) = {±1, ±i}.
2. Prove that U (Z + Zω) = {±1, ±ω, ±ω2 }.
3. Let m be an integer with m < −1. Prove that
√
U (Z + Z m) = {±1}.
4. Let m be an integer with m ≡ 1 (mod 4) and m < −3. Prove that
√
1+ m
U Z+Z = {±1}.
2
5. Let
24 Integral Domains
a 2 − mb2 = ±1
√
then α ∈ U (Z + Z m).
18. Let m be a√ positive
integer
with
√
m ≡ 1 (mod 4) that is not a perfect square. Let α =
a + b 1+2 m ∈ Z + Z 1+2 m . Prove that if
1−m
a 2 + ab + b2 = ±1
4
√
then α ∈ U Z + Z 1+2 m .
19. Prove that 1 − 3i, 3 − i is a principal ideal in Z + Zi by finding a generator for this
ideal.
√ √ √ √
20. Prove that 2, 1 + −5 = 2, 1 − −5, 3, 1 + −5 = 3, 1 − −5, 2, 1 +
√ √ √ √ √
−5 = 3, 1 + −5, and 2, 1 + −5 = 3, 1 − −5 in Z + Z −5.
√ √ √
21. Prove that 2, 1 + −5, 3, 1 + −5, and 3, 1 − −5 are prime ideals of Z +
√ √ √ √
Z −5. Determine 2, 1 + −5 ∩ Z, 3, 1 + −5 ∩ Z, and 3, 1 − −5 ∩ Z.
22. Let D be an integral domain. Let a, b, c ∈ D be such that a, c = D. Prove that
a, bc = a, b. √ √ √
23. Prove that 17 − 3 3 ∼ 83 + 47 3 in Z + Z 3.
24. Give an example of an integral domain satisfying (1.2.3).
√ √
25. Express 2 + 8 −5 as a product of irreducibles in Z + Z −5. In how many ways can
this be done?
√ √
26. Prove that −6 is not a prime in Z + Z −6.
√ √
27. Prove that −6 is an irreducible in Z + Z −6.
√
28. Prove that Z + Z −6 is not a principal ideal domain.
√
29. Give an example
√ of an ideal in Z + Z √−6 that is not principal.
30. Prove that √10 is not a prime in Z + Z 10. √
31. Prove that 10 is√an irreducible in Z + Z 10.
32. Prove that Z + Z 10 is not a principal √ideal domain.
33. Give an example of an ideal in Z + Z 10 that is not principal.
34. Let P be a prime ideal of an integral domain D. Let A1 , . . . , Ak be ideals of D such
that P ⊇ A1 · · · Ak . Prove that P ⊇ Ai for some i ∈ {1, 2, . . . , k}.
35. Let r ∈ Z \ {−2, 0}. Prove that
D = {a + bθ + cθ 2 | a, b, c ∈ Z},
where
θ3 + rθ + 1 = 0
Biographies 25
Suggested Reading
1. P. M. Cohn, Rings of fractions, American Mathematical Monthly 78 (1971), 596–615.
The author won the Lester R. Ford award for expository writing for this paper. The paper reviews
Ore’s work on embedding certain non-commutative rings in skew fields, a generalization of the
corresponding standard result for integral domains mentioned in Section 1.1.
2. D. A. Cox, Primes of the Form x 2 + my 2 , Wiley, New York, 1989.
The main theorem of the book (Theorem 9.2, p. 180) asserts (with some details omitted) that if m
is a positive integer then there is a polynomial f m (x)∈ Z[x] (of a certain degree depending only
on m) such that if p is an odd prime satisfying −m p
= 1 then p = u 2 + mv 2 for integers u and
v if and only if the congruence f m (x) ≡ 0 (mod
p) is solvable.
For example if p is an odd prime such that −36 p
= 1, that is, p ≡ 1 (mod 4), then
Biographies
1. E. T. Bell, Men of Mathematics, Simon and Schuster, New York, 1937.
Chapters 14 and 17 are devoted to Gauss and Abel respectively.
2. W. K. Bühler, Gauss: A Biographical Study, Springer-Verlag, Berlin, Heildelberg,
New York, 1981.
This book provides a comprehensive discussion of Gauss’s life and work.
3. G. Eisenstein, Mathematische Werke, Bände I, II, Chelsea Publishing Co., New York,
1989.
The foreword to Eisenstein’s Collected Papers comprises an interesting discussion of Eisenstein’s
work by André Weil.
CB609-01 CB609/Alaca & Williams August 7, 2003 17:16 Char Count= 0
26 Integral Domains
http://www-groups.dcs.st-and.ac.uk/˜history/
2
Euclidean Domains
In the proof of Theorem 1.4.1 we made use of the following property of Z: Given
a, b ∈ Z with b > 0 then there exist q, r ∈ Z such that
a = qb + r, 0 ≤ r < b. (2.0.1)
In fact the integers q and r are uniquely determined by a and b. We have
q = [a/b], r = a − b[a/b], (2.0.2)
where [x] denotes the greatest integer less than or equal to the real number x.
The integer q is called the quotient and the integer r the remainder. An important
class of integral domains are those possessing a property analogous to (2.0.1). Such
domains are called Euclidean domains. In Theorem 2.1.2 we show that Euclidean
domains are principal ideal domains.
27
CB609-02 CB609/Alaca & Williams August 27, 2003 16:51 Char Count= 0
28 Euclidean Domains
Proof: (a) As a ∼ b there exists u ∈ U (D) such that a = ub. Then by (2.1.1) we
have φ(a) = φ(ub) ≥ φ(b). As u ∈ U (D), we have u −1 ∈ U (D) and b = u −1 a, so
again by (2.1.1) we have φ(b) = φ(u −1 a) ≥ φ(a). From these two inequalities, we
deduce that φ(a) = φ(b).
(b) By (2.1.2) there exist q, r ∈ D such that a = qb + r and φ(r ) < φ(b) =
φ(a). Now a | b so that we have a | r . Suppose r = 0. Then by (2.1.1) we have
φ(r ) ≥ φ(a), which is a contradiction. Hence r = 0. Thus a = qb. But a | b so
q ∈ U (D) and thus a ∼ b.
(c) First we have
a ∈ U (D) =⇒ a ∼ 1 =⇒ φ(a) = φ(1)
by part (a). Second, we have
1 | a, φ(1) = φ(a) =⇒ 1 ∼ a =⇒ a ∈ U (D)
by part (b).
(d) By (2.1.2) there exist q, r ∈ D such that
0 = qa + r, φ(r ) < φ(a).
Suppose r = 0. Then q = 0 and by (2.1.1) we have
φ(r ) = φ((−q)a) ≥ φ(a),
which is a contradiction. Hence r = 0 and φ(0) < φ(a).
of Euclidean domains in the next section, we prove the fundamental theorem that
every Euclidean domain is a principal ideal domain.
S = {φ(x) | x ∈ I, x = 0}.
r−1 = a, r0 = b, (2.1.3)
and
rk+2 = 0.
Then
(a, b) = rk+1 .
CB609-02 CB609/Alaca & Williams August 27, 2003 16:51 Char Count= 0
30 Euclidean Domains
Proof: By property (2.1.2) of the Euclidean function φ, the relations (2.1.3) and
(2.1.4) define q1 , q2 , . . . , qk+2 and r−1 , r0 , r1 , . . . , rk+2 , and since the sequence
φ(r1 ), φ(r2 ), . . . is a decreasing sequence of integers bounded below by φ(0) (The-
orem 2.1.1(d)) it must terminate after a finite number of steps (say k + 2 steps) so
that rk+2 = 0. From (2.1.4) we deduce that
r j , r j+1 = q j+2r j+1 + r j+2 , r j+1 = r j+2 , r j+1 = r j+1 , r j+2
for j = −1, 0, 1, 2, . . . , k. Hence
a, b = r−1 , r0 = r0 , r1 = · · · = rk , rk+1
= rk+1 , rk+2 = rk+1 , 0 = rk+1
so that
(a, b) = rk+1 .
Theorem 2.2.1
(a) Z is a Euclidean domain.
(b) Let F be a field. Then F[x] is a Euclidean domain.
From Theorems 2.1.2 and 2.2.1 we see that Z and F[x] are principal ideal
domains. In the remainder of this section we
investigate
√
when the integral domains
√ 1+ m
Z + Z m (m ≡ 2, 3 (mod 4)) and Z + Z (m ≡ 1 (mod 4)) are Euclidean
√2
with respect to the function that maps r + s m to |r 2 − ms 2 |. In this section we
denote this function by φm . Later in Section 9.2 we recognize φm as the absolute
√
value of the norm of the element r + s m. Integral domains that are Euclidean
with respect to the absolute value of the norm are called norm-Euclidean.
√ √
Proof: (a) Let α ∈ Z + Z m so that α = x + y m for some x, y ∈ Z. Then
x 2 − my 2 ∈ Z and |x 2 − my 2 | ≥ 0 so that
√
φm (α) = φm (x + y m) = |x 2 − my 2 | ∈ N ∪ {0}.
√
(b) If m ≡ 1 (mod 4) then Z + Z 1+2 m is an integral domain (Example 1.1.5).
√ √ √
Let α ∈ Z + Z 1+2 m so that α = x + y 1+2 m = x + 2y + 2y m for some
x, y ∈ Z. Then
y y√
φm (α) = φm x + + m
2 2
y y
= |(x + )2 − m( )2 |
2 2
1 1
= |x + x y + (1 − m)y 2 | ∈ N ∪ {0}, as (1 − m) ∈ Z.
2
4 4
√ √
(c) Let α ∈ Q( m) so that α = r + s m for some r, s ∈ Q. Then, as m is
squarefree, we have
√
φm (α) = 0 ⇐⇒ φm (r + s m) = 0
⇐⇒ |r 2 − ms 2 | = 0
⇐⇒ r 2 = ms 2
⇐⇒ r = s = 0
√
⇐⇒ r + s m = 0
⇐⇒ α = 0.
√ √ √
(d) Let α, β ∈ Q( m). Then α = x + y m and β = u + v m for some
x, y, u, v ∈ Z. Thus
√ √
φm (αβ) = φm ((x + y m)(u + v m))
√
= φm ((xu + myv) + (xv + yu) m)
= |(xu + myv)2 − m(xv + yu)2 |
= |x 2 u 2 + m 2 y 2 v 2 − mx 2 v 2 − my 2 u 2 |
= |(x 2 − my 2 )(u 2 − mv 2 )|
= |x 2 − my 2 | |u 2 − mv 2 |
= φm (α)φm (β).
√
(e) Let α, β ∈ Z + Z m with β = 0. By part (c), we have φm (β) = 0. Then, by
part (a), we deduce that φm (α) ≥ 0 and φm (β) ≥ 1. Thus, by part (d), we have
32 Euclidean Domains
(f) This follows in exactly the same way as part (e) except that we use part (b) in
place of part (a).
Our next theorem uses the properties of φm given in Lemma 2.2.1 to give a
√
convenient necessary and sufficient condition for Z + Z m to be Euclidean with
respect to φm , that is, norm-Euclidean.
√
Theorem 2.2.2 Let m be a squarefree integer. Then the integral domain Z + Z m
is Euclidean with respect to φm if and only if for all x, y ∈ Q there exist a, b ∈ Z
such that
√ √
φm ((x + y m) − (a + b m)) < 1. (2.2.1)
√
Proof: Suppose first that Z + Z m is Euclidean with respect to φm . Let x, y ∈ Q.
√ √
Then x + y m = (r + s m)/t for integers r, s, t with t = 0. As φm is a Euclidean
√ √ √ √
function on Z + Z m there exist a + b m, c + d m ∈ Z + Z m such that
√ √ √ √
r + s m = t(a + b m) + (c + d m), φm (c + d m) < φm (t).
Hence
√
√ √ r +s m √
φm ((x + y m) − (a + b m)) = φm − (a + b m)
t
√ √
r + s m − t(a + b m)
= φm
t
√
c+d m
= φm
t
√
φm (c + d m)
= < 1,
φm (t)
by Lemma 2.2.1(d).
√
Now suppose that (2.2.1) holds. To show that Z + Z m is Euclidean with respect
to φm , we must show that (2.1.1) and (2.1.2) hold. The inequality (2.1.1) holds in
√
view of Lemma 2.2.1(e). We now show that (2.1.2) holds. Let r + s m, t +
√ √ √
u m ∈ Z + Z m with t + u m = 0. Then
√
r +s m √
√ = x + y m,
t +u m
where
r t − msu st − r u
x= ∈ Q, y = 2 ∈ Q.
t − mu
2 2 t − mu 2
CB609-02 CB609/Alaca & Williams August 27, 2003 16:51 Char Count= 0
Hence
√ √ √ √
r + s m = (a + b m)(t + u m) + (c + d m)
and
√ √ √ √
φm (c + d m) = φm ((r + s m) − (a + b m)(t + u m))
√ √ √ √
= φm ((x + y m)(t + u m) − (a + b m)(t + u m))
√ √ √
= φm ((t + u m)((x + y m) − (a + b m)))
√ √ √
= φm (t + u m)φm ((x + y m) − (a + b m))
√
< φm (t + u m),
by Lemma 2.2.1(d), which completes the proof of (2.1.2).
Theorem 2.2.3 Let m be a negative squarefree integer. Then the integral domain
√
Z + Z m is Euclidean with respect to φm if and only if m = −1, −2.
√
Proof: First we show that Z + Z m is Euclidean with respect to φm for m = −1
and m = −2. Let x, y ∈ Q. We can choose a, b ∈ Z such that
1 1
|x − a| ≤ , |y − b| ≤ .
2 2
Then
√ √ √
φm ((x + y m) − (a + b m)) = φm ((x − a) + (y − b) m))
= |(x − a)2 − m(y − b)2 |
≤ |x − a|2 + |m||y − b|2
1 1
≤ +2·
4 4
3
= <1
4
√
and, appealing to Theorem 2.2.2, we deduce that Z + Z m is Euclidean with
respect to φm for m = −1 and m = −2.
CB609-02 CB609/Alaca & Williams August 27, 2003 16:51 Char Count= 0
34 Euclidean Domains
√
Now suppose that Z + Z m is Euclidean with respect to φm . Then, by Theorem
2.2.2, there exist a, b ∈ Z such that
1 1√ √
φm ( + m) − (a + b m) < 1;
2 2
that is (as −m = |m|),
2 2
1 1
−a + |m| −b < 1.
2 2
But for any integer x, we have
2
1 1 1 1
| − x| ≥ , −x ≥ ,
2 2 2 4
so
1 |m|
+ < 1;
4 4
that is, |m| < 3. Hence m = −1 and m = −2 are the only possibilities.
In an exactly similar way to the proof of Theorem 2.2.2, we can prove the
following result.
Theorem 2.2.4Let√m be a squarefree integer with m ≡ 1 (mod 4). Then the integral
1+ m
domain Z + Z 2
is Euclidean with respect to φm if and only if for all x, y ∈ Q
there exist a, b ∈ Z such that
√
√ 1+ m
φm (x + y m) − a + b < 1.
2
Theorem 2.2.6 Let m be a positive squarefree integer with m ≡ 2, 3 (mod 4). Then
√
the integral domain Z + Z m is Euclidean with respect to φm if and only if m =
2, 3, 6, 7, 11, 19, 57.
We will not prove these two theorems here. We will just prove the following
result.
√
Theorem 2.2.8 The integral domain Z + Z m is Euclidean with respect to φm for
m = 2, 3, 6.
that is,
36 Euclidean Domains
5
(i) (1 − r1 )2 − 6s12 = 1 or (ii) − ≤ (1 − r1 )2 − 6s12 ≤ −1, (2.2.7)
4
9
1 ≤ (1 + r1 )2 − 6s12 ≤ . (2.2.8)
4
From (2.2.6) and (2.2.8), we obtain
1 ≤ 1 + 2r1 + (r12 − 6s12 ) ≤ 2r1 ,
CB609-02 CB609/Alaca & Williams August 27, 2003 16:51 Char Count= 0
Theorem 2.3.1 Let m be a positive squarefree integer. If there exist distinct odd
primes p and q such that
m m
= = −1,
p q
and positive integers t and u such that
pt + qu = m, p t, q u,
r 2 ≡ pt (mod m),
√
then Z + Z m is not Euclidean with respect to φm .
√
Proof: Suppose that Z + Z m is Euclidean with respect to φm . Then there exist
√
γ , δ ∈ Z + Z m such that
√
r m = mγ + δ, φm (δ) < φm (m).
√
Setting γ = x + y m (x, y ∈ Z) we obtain
√ √
φm (r m − m(x + y m)) < φm (m);
that is
so that
Since
mx 2 − (my − r )2 ≡ −r 2 ≡ − pt (mod m)
CB609-02 CB609/Alaca & Williams August 27, 2003 16:51 Char Count= 0
38 Euclidean Domains
and
0 < pt < pt + qu = m,
we must have
mx 2 − (my − r )2 = − pt or m − pt;
that is
m X 2 − Y 2 = − pt or qu
for integers X (= x) and Y (= my − r ). Suppose that m X 2 − Y 2 = − pt. As mp =
−1 we have p m. Also, as p t we have p || − pt. Hence p X and p Y . Thus
2
m m X2 Y
= = = 1,
p p p
contradicting mp = −1. Now suppose that m X 2 − Y 2 = qu. As mq = −1 we
have q m. Also, as q u we have q || qu. Hence q X and q Y . Thus
2
m m X2 Y
= = = 1,
q q q
√
contradicting mq = −1. This proves that Z + Z m is not Euclidean with respect
to φm .
We next use Theorem 2.3.1 to give some explicit, small, positive, squarefree
√
values of m for which Z + Z m is not Euclidean with respect to φm .
√
Theorem 2.3.2 Z + Z m is not Euclidean with respect to φm for m = 23, 47,
59, 83.
Proof: This follows immediately from Theorem 2.3.1 and the following table.
m p q t u r
23 3 5 1 4 7
47 3 5 4 7 23
59 3 7 15 2 24
83 3 5 1 16 13
√
1+ m
The corresponding result to Theorem 2.3.1 for Z + Z 2
(m ≡ 1 (mod 4))
is not quite so elegant.
CB609-02 CB609/Alaca & Williams August 27, 2003 16:51 Char Count= 0
Theorem 2.3.3 Let m be a positive squarefree integer with m ≡ 1 (mod 4). If there
exist distinct odd primes p and q such that
m m
= = −1
p q
and an odd integer r such that
(m − 1)r 2
p || (m − 1)r − 4m
2
,
4m
(m − 1)r 2
q || (m − 1)r 2 − 4m − 4m,
4m
√
1+ m
then Z + Z 2
is not Euclidean with respect to φm .
|m X 2 − Y 2 | < 4m.
40 Euclidean Domains
Also
m X 2 − Y 2 ≡ −Y 2 ≡ −r 2 (mod m).
Hence
m X 2 − Y 2 ≡ (m − 1)r 2 (mod 4m).
Thus
(m − 1)r 2
m X − Y = (m − 1)r − 4m
2 2 2
4m
or
(m − 1)r 2
m X − Y = (m − 1)r − 4m
2 2 2
− 4m.
4m
In the first case we have p || m X 2 − Y 2 . As mp = −1 we have p m. Thus p X
and p Y . Then
2
m m X2 Y
= = = 1,
p p p
contradicting mp = −1.
The second case can be treated similarly.
√
We now use Theorem 2.3.3 to show that Z + Z 1+2 53 is not Euclidean with
respect to φ53 .
√
Theorem 2.3.4 Z + Z 1+ 53
2
is not Euclidean with respect to φ53 .
Proof: We choose
m = 53, p = 5, q = 19, r = 29.
Clearly
m 53 3
= = = −1,
p 5 5
m 53 −4 −1
= = = = −1,
q 19 19 19
(m − 1)r 2 52 · 292
(m − 1)r − 4m
2
= 52 · 29 − 4 · 53 ·
2
4m 4 · 53
= 43732 − 212 · 206
= 43732 − 43672
= 60 = 5 · 22 · 3,
CB609-02 CB609/Alaca & Williams August 27, 2003 16:51 Char Count= 0
and
(m − 1)r 2
(m − 1)r 2 − 4m − 4m = 60 − 212 = −152 = −19 · 23 ,
4m
so the result follows from Theorem 2.3.3
√
In the next theorem we show that Z + Z m (m ≡ 2, 3 (mod 4)) is not Euclidean
with
respect
√
to φm if m is sufficiently large. The same result is also true for Z +
1+ m
Z 2
(m ≡ 1 (mod 4)) but the proof is more complicated and we will not
give it here.
√ √ √ 2 √
Proof:
√ (a)
√ As m ≥ 42 we have m > 20 + 8 6 = 4( 3 + 2) so that m >
2( 3 + 2) and thus
√ √ √ √
3m − 1 2m − 1 3− 2 √
− = m
2 2 2
√ √
( 3 − 2) √ √
> 2( 3 + 2) = 1.
2
Hence there exists an integer u satisfying
√ √
2m − 1 3m − 1
<u< .
2 2
Set t = 2u + 1 so that t is an odd integer satisfying
that is,
42 Euclidean Domains
and thus
|X 2 − mY 2 | < m
and
X 2 − mY 2 ≡ X 2 ≡ t 2 (mod m).
X 2 − mY 2 = t 2 − 2m
or
X 2 − mY 2 = t 2 − 3m.
In the first case, as t 2 ≡ 1 (mod 8) (since t is odd) and m ≡ 2 (mod 4), we have
X 2 − mY 2 ≡ 5 (mod 8).
mY 2 ≡ 4 (mod 8).
X 2 − mY 2 ≡ 1 − 3m (mod 8).
Hence, as 2 || m, we have
Y 2 ≡ 3 (mod 4),
which is impossible.
CB609-02 CB609/Alaca & Williams August 27, 2003 16:51 Char Count= 0
44 Euclidean Domains
In the first case, as t 2 ≡ 1 (mod 8) (since t is odd) and m ≡ 3 (mod 4), we have
X 2 − mY 2 = t 2 − 5m ≡ 1 − 15 = −14 ≡ 2 (mod 4)
so that
X ≡ Y ≡ 1 (mod 2).
Thus X 2 ≡ Y 2 ≡ 1 (mod 8) so that
1 − 5m ≡ t 2 − 5m = X 2 − mY 2 ≡ 1 − m (mod 8),
giving 4m ≡ 0 (mod 8), which is clearly impossible. In the second case, as
t 2 ≡ 1 (mod 8) and m ≡ 3 (mod 4), we have
X 2 − mY 2 = t 2 − 6m ≡ 1 − 18 = −17 ≡ 7 (mod 8).
If X is odd, so X 2 ≡ 1 (mod 8), then
mY 2 ≡ 2 (mod 8),
which is impossible. If X is even, so X 2 ≡ 0 (mod 4), then −3Y 2 ≡ 3 (mod 4), so
that Y 2 ≡ 3 (mod 4), which is impossible.
√
It is a consequence of Theorem 2.2.5 that the domain Z + Z 1+ 2−19 is not
Euclidean with respect to φ−19 . But could it be Euclidean with respect to some
other function? In fact it is not. How do we see this? One way of showing that
an integral domain is not Euclidean with respect to any function is to show that
it does not possess certain distinguished elements called universal side divisors,
since a domain that has no universal side divisors
√ isnot Euclidean with respect to
any function. Indeed as we shall see Z + Z 1+ 2−19 has no universal divisors and
therefore is not Euclidean with respect to any function.
We now define a universal side divisor. For any integral domain D it is convenient
to set
D̃ = U (D) ∪ {0}
so that
D − D̃ = φ if and only if D is a field.
Definition 2.3.1 (Universal side divisor) Let D be an integral domain that is not a
field so that D − D̃ = φ. An element u ∈ D − D̃ is called a universal side divisor
if for any x ∈ D there exists some z ∈ D̃ such that u|x − z.
Theorem 2.3.6 Let D be an integral domain that is not a field. If D has no universal
side divisors then D is not Euclidean.
CB609-02 CB609/Alaca & Williams August 27, 2003 16:51 Char Count= 0
Proof: Suppose that D is Euclidean with respect to the Euclidean function φ and
has no universal side divisors. Consider the set of integers defined by
S = {φ(v) | v ∈ D − D̃}.
√
If m is a negative squarefree integer with m ≡ 2, 3 (mod 4) then Z + Z m is
Euclidean with respect to φm for m = −1 and m = −2 and is not Euclidean with
respect to φm for m < −2 (Theorem 2.2.3). We now use Theorem 2.3.6 to show
√
that Z + Z m is not Euclidean with respect to any function for m < −2.
√
Proof: Let D = Z + Z m. As m = −1, U (D) = {1, −1} (see Exercise 3 of Chap-
ter 1) so that D̃ = {0, 1, −1}. Suppose that u is a universal side divisor in D. Then
u must divide one of 2 − 1, 2 + 0, or 2 + 1, that is, one of 1, 2, or 3. But u
being a universal side divisor is not a unit, so u 1. Hence u | 2 or u | 3. Since
m ≡ 2, 3 (mod 4) and m < −2 we have m ≤ −5 so that both 2 and 3 are irre-
ducible in D (Exercise 36 of Chapter 1). Hence the only possible universal side
divisors are 2, −2, 3, and −3. However, none of these divides any of the three
√
elements of Z + Z m :
√ √ √
m − 1, m, m + 1,
so that no such universal side divisor can exist. Hence, by Theorem 2.3.6, D is not
Euclidean.
√
If m is a negative squarefree integer with m ≡ 1 (mod 4) then Z + Z 1+2 m
is Euclidean with respect to φm for m = −3, −7, −11 and is not Euclidean with
respectto φ√m
for m < −11 (Theorem 2.2.5). We use Theorem 2.3.6 to show that
1+ m
Z+Z 2
is not Euclidean with respect to any function for m < −11.
46 Euclidean Domains
√
1+ m
Proof: Let D = Z + Z 2
. As m = −3 we have U (D) = {1, −1} (Exercise
4 of Chapter 1) so that D̃ = {0, 1, −1}. Suppose that u is a universal side divisor in
D. Then u must divide one of 2 − 1, 2 + 0, or 2 + 1, that is, one of 1, 2, or 3. As
u is not a unit,√u must divide 2 or 3. Since m ≤ −15, both 2 and 3 are irreducible
in Z + Z 1+2 m (Exercise 37 of Chapter 1). Therefore the only possible side
divisors are 2, −2, 3, and−3. However, none of these divides any of the following
√
1+ m
three elements of Z + Z 2
,
1 √ 1 √ 1 √ 1 √ 1 √
(−1 + m) = (1 + m) − 1, (1 + m), (3 + m) = (1 + m) + 1,
2 2 2 2 2
so that no such universal side divisor can exist. Hence, by Theorem 2.3.6, D is not
Euclidean.
When m is a √
positive
squarefree integer very little is known. Clark [5] has shown
that Z + Z 1+ 69
2
is Euclidean with respect to the function
√
1+ 69 |a 2 + ab − 17b2 |, if (a, b) = (10, 3),
φ a+b =
2 26, if (a, b) = (10, 3).
√
By Theorem 2.2.7 we know that Z + Z 1+2 69 is not Euclidean with respect
to φ69 . This is the first example of a real quadratic domain that is Euclidean but
not norm-Euclidean. Since the 26√in the definition of φ can be replaced by any
integer greater than 25, Z + Z 1+2 69 is Euclidean with respect to infinitely many
√
different functions. Samuel [16] suggests that Z + Z 14 may be Euclidean with
respect to some function different from φ14 , and this has recently been proved by
Harper [9].
48 Euclidean Domains
Theorem 2.5.1 Let p be a prime such that p ≡ 1 (mod 4). Then there exist integers
x and y such that p = x 2 + y 2 .
√
Proof: As p ≡ 1 (mod 4), by (2.5.1) we have −1 p
= 1. Since Z + Z −1 is a
Euclidean domain, by Theorem 2.1.2 it is a principal ideal domain. Thus by Theorem
1.4.4, there are integers x and y such that p = x 2 + y 2 .
Theorem 2.5.2 Let p be a prime such that p ≡ 1, 3 (mod 8). Then there exist
integers x and y such that p = x 2 + 2y 2 .
Jackson [11] has given a short proof of Theorem 2.5.2 when p ≡ 3 (mod 8).
Similarly using (2.5.3)–(2.5.5) and Theorem 1.4.5, we obtain the following three
theorems.
Theorem 2.5.3 Let p be a prime such that p ≡ 1 (mod 3). Then there exist integers
x and y such that p = x 2 + x y + y 2 .
Theorem 2.5.4 Let p be a prime such that p ≡ 1, 2, 4 (mod 7). Then there exist
integers x and y such that p = x 2 + x y + 2y 2 .
CB609-02 CB609/Alaca & Williams August 27, 2003 16:51 Char Count= 0
Exercises 49
Theorem 2.5.5 Let p be a prime such that p ≡ 1, 3, 4, 5, 9 (mod 11). Then there
exist integers x and y such that p = x 2 + x y + 3y 2 .
√
In Theorem 2.2.8 we showed that Z + Z m is Euclidean for m = 2, 3, 6. Recall
from elementary number theory that for an odd prime p
2
= 1 ⇐⇒ p ≡ 1, 7 (mod 8),
p
3
= 1 ⇐⇒ p ≡ 1, 11 (mod 12),
p
6
= 1 ⇐⇒ p ≡ 1, 5, 19, 23 (mod 24).
p
Theorem 2.5.6 Let p be a prime such that p ≡ 1, 7 (mod 8). Then there exist
integers x and y such that p = x 2 − 2y 2 .
Theorem 2.5.7 Let p be a prime such that p ≡ 1, 11 (mod 12). Then there exist
integers x and y such that either p = x 2 − 3y 2 or p = 3y 2 − x 2 .
Theorem 2.5.8 Let p be a prime such that p ≡ 1, 5, 19, 23 (mod 24). Then there
exist integers x and y such that either p = x 2 − 6y 2 or p = 6y 2 − x 2 .
Exercises
1. Let D be an integral domain possessing a Euclidean function φ. Give an example to
show that
50 Euclidean Domains
√
6. Prove that Z + Z 1+2 5 is Euclidean with respect to φ5 using the method of the proof
of Theorem 2.2.8. √
7. Use Theorem 2.3.1 to show that Z + Z 26 is not Euclidean with respect to φ26 .
8. Prove a modification of Theorem 2.3.1 that allows one of the primes p and q to be the
prime 2.
9. Prove an extension of Theorem 2.3.1 that replaces p and q in the equation pt + qu = m
by powers of p and q with odd exponents. √
10. Use Theorem 2.3.3 to prove that Z + Z 1+2 77 is not Euclidean with respect to φ77 .
11. Prove that if p is a prime with p ≡ 3 (mod 4) then there do not exist integers x and y
such that p = x 2 + y 2 .
12. Let p be a prime. Use Theorem 2.5.1 and Exercise 11 to deduce that
p = x 2 + y 2 ⇐⇒ p = 2 or p ≡ 1 (mod 4).
13. Prove that if p is a prime with p ≡ 5, 7 (mod 8) then there do not exist integers x and
y such that p = x 2 + 2y 2 .
14. Let p be a prime. Use Theorem 2.5.2 and Exercise 13 to deduce that
p = x 2 + 2y 2 ⇐⇒ p = 2 or p ≡ 1, 3 (mod 8).
15. Prove that if p is a prime with p ≡ 2 (mod 3) then there do not exist integers x and y
such that p = x 2 + x y + y 2 .
16. Let p be a prime. Use Theorem 2.5.3 and Exercise 15 to deduce that
p = x 2 + x y + y 2 ⇐⇒ p = 3 or p ≡ 1 (mod 3).
17. Prove that if p is a prime with p ≡ 3, 5, 6 (mod 7) then there do not exist integers x
and y such that p = x 2 + x y + 2y 2 .
18. Let p be a prime. Use Theorem 2.5.4 and Exercise 17 to deduce that
p = x 2 + x y + 2y 2 ⇐⇒ p = 7 or p ≡ 1, 2, 4 (mod 7).
19. Prove that if m is a positive integer possessing a prime divisor q ≡ 3 (mod 4) then there
are no integers T and U such that T 2 − mU 2 = −1.
20. Let p be a prime with p ≡ 1, 11 (mod 12). Deduce from Theorem 2.5.7 that
p = x 2 − 3y 2 , if p ≡ 1 (mod 12),
p = 3y 2 − x 2 , if p ≡ 11 (mod 12),
p = x 2 − 6y 2 , if p ≡ 1, 19 (mod 24),
p = 6y 2 − x 2 , if p ≡ 5, 23 (mod 24),
Suggested Reading 51
√ √
22. Prove that the subdomain Z + 3Z −2 of the Euclidean domain Z + Z −2 is not
Euclidean. √ √
23. Prove that the subdomain Z + 7Z 2 of the Euclidean domain Z + Z 2 is not
Euclidean. √ √
24. Prove that the subdomain Z + 2Z 3 of the Euclidean domain Z + Z 3 is not
Euclidean. √ √
25. Prove that the subdomain Z + 5Z 6 of the Euclidean domain Z + Z 6 is not
Euclidean.
26. Let m be a positive integer with m ≡ 1 (mod 4). Show that the solvability of the equation
T 2 + T U + 14 (1 − m)U 2 = −1 in integers T and U (see Theorem 1.4.5) is equivalent
to the solvability of the equation X 2 − mY 2 = −4 in integers X and Y .
27. Let m be an integer with m ≡ 1 (mod 4) that possesses a prime divisor q ≡ 3 (mod 4).
Prove that there are no integers T and U such that T 2 + T U + 14 (1 − m)U 2 = −1.
28. Prove that if p is a prime with p ≡ 1, 4 (mod 5) then there are integers x and y such
that p = x 2 + x y − y 2 . [Hint: Use Theorems 1.4.5 and 2.2.7.]
29. Use Exercise 12 to show that the irreducibles in Z + Zi are 1 + i and its asso-
ciates, x ± i y, where x 2 + y 2 = p (prime) ≡ 1 (mod 4), and their associates, and
q (prime) ≡ 3 (mod 4) and its associates.
√
30. Use Exercise 14 to determine the irreducibles in Z + Z −2.
Suggested Reading
1. P. J. Arpaia, A note on quadratic Euclidean domains, American Mathematical Monthly
75 (1968), 864–865.
Examples of quadratic Euclidean domains that√possess subdomains that are not Euclidean
√ are
given. For example the Gaussian domain Z + Z −1 is Euclidean but its subdomain Z + 2Z −1
is not Euclidean.
2. O. A. Cámpoli, A principal ideal domain that is not a Euclidean domain, American
Mathematical Monthly 95 (1988), 868–871.
√
It is shown in an elementary fashion that Z + Z( 1+ 2−19 ) is a principal ideal domain but not a
Euclidean domain. The idea of a domain being almost Euclidean is introduced (p. 870).
3. H. Chatland, On the Euclidean algorithm in quadratic number fields, Bulletin of the
American Mathematical Society 55 (1949), 948–953.
This paper is a valuable source of references
√
to work on the Euclidean algorithm in quadratic
domains. It should be noted that Z + Z( 1+2 97 ) is not Euclidean, contrary to the claim by Rédei.
This was established by Barnes and Swinnerton-Dyer in 1952.
4. H. Chatland and H. Davenport, Euclid’s algorithm in real quadratic fields, Canadian
Journal of Mathematics 2 (1950), 289–296.
This is where the final steps in the proofs of Theorems 2.2.6 and 2.2.7 are given.
5. D. A. Clark, A quadratic field which is Euclidean but not norm-Euclidean, Manuscripta
Mathematica 83 (1994),√ 327–330.
It is shown that Z + Z( 1+2 69 ) is Euclidean but not norm-Euclidean.
6. D. A. Cox, Primes of the form x 2 + ny 2 , Wiley, New York, 1989.
This book presents a comprehensive treatment of the problem of deciding which primes are
represented by x 2 + ny 2 .
CB609-02 CB609/Alaca & Williams August 27, 2003 16:51 Char Count= 0
52 Euclidean Domains
Biographies 53
Biographies
1. K. Barner, Pierre de Fermat (1601?–1665)—His life beside mathematics, Canadian
Mathematical Society Notes 34 (2002), 3–4, 26–30.
The author relates an interesting account of the nonmathematical life of Fermat.
2. P. Hoffman, The Man Who Loved Only Numbers, Hyperion, New York, 1998.
The story of Paul Erdös, one of the most prolific and eccentric mathematicians of the twentieth
century, is told.
3. M. S. Mahoney, The Mathematical Career of Pierre de Fermat (1601–1605), Princeton
University Press, Princeton, New Jersey, 1973.
For two completely different reviews of this book, see Isis 65 (1974), 398–400 and Bulletin of the
American Mathematical Society 79 (1973), 1138–1149.
4. C. A. Rogers, Harold Davenport, Bulletin of the London Mathematical Society 4 (1972),
66–99.
A memoir on the life and mathematics of Davenport is presented.
5. The website
http://www-groups.dcs.st-and.ac.uk/˜history/
3
Noetherian Domains
I1 ⊂ I2 ⊂ I3 ⊂ I4 ⊂ . . . . (3.1.1)
The importance of domains such as Z that do not contain infinite ascending chains
of ideals of the type (3.1.1) was first recognized by the German mathematician
Emmy Noether (1882–1935). Such domains are now called Noetherian domains in
her honor. We note that some domains do contain infinite chains of ideals of the
type (3.1.1). For example, if F is a field, the domain F[X 1 , X 2 , . . .] contains the
infinite chain of ideals
X 1 ⊂ X 1 , X 2 ⊂ X 1 , X 2 , X 3 ⊂ . . . .
54
CB609-03 CB609/Alaca & Williams August 7, 2003 17:19 Char Count= 0
In = In 0 for all n ≥ n 0 .
Definition 3.1.4 (Noetherian domain) An integral domain that satisfies the as-
cending chain condition is called a Noetherian domain.
More generally we define a Noetherian ring to be a ring R in which every
ascending chain of (two-sided) ideals in R terminates.
From the remarks preceding the definitions, we have the following two examples.
The next theorem gives a necessary and sufficient condition for an integral domain
to be a Noetherian domain.
Proof: Let D be a Noetherian domain. Suppose that not every ideal of D is finitely
generated. Let I be an ideal of D that is not finitely generated. Thus I = 0, and
so there exists a1 ∈ I with a1 = 0. Let A1 be the ideal given by A1 = a1 . Clearly
A1 ⊆ I . Moreover, I = A1 as A1 is finitely generated and I is not. Hence A1 ⊂ I .
Take a2 ∈ I, a2 ∈ A1 , and let A2 be the ideal given by A2 = a1 , a2 . Clearly
A1 ⊂ A2 ⊂ I . Continuing in this way, we obtain an infinite strictly increasing
sequence of ideals A1 ⊂ A2 ⊂ . . . , contradicting that D is a Noetherian domain.
Hence every ideal of a Noetherian domain must be finitely generated.
Now let D be an integral domain in which every ideal is finitely generated.
Let
I1 ⊆ I2 ⊆ I3 ⊆ . . .
be an ascending chain of ideals in D. It is easy to check that ∞ n=1 In is an ideal
of D. Hence ∞ n=1 I n is finitely generated, so there exist finitely many elements
CB609-03 CB609/Alaca & Williams August 7, 2003 17:19 Char Count= 0
56 Noetherian Domains
a1 , a2 , . . . , am of D such that
∞
In = a1 , a2 , . . . , am .
n=1
For each i = 1, 2, . . . , m, ai ∈ ∞ n=1 In , say, ai ∈ Ini . Set l = max (n 1 , n 2 , . . . , n m ).
∞
Clearly Il ⊆ n=1 In . As n i ≤ l we have Ini ⊆ Il , and thus ai ∈ Il for i =
∞
1, 2, . . . , m. Hence a1 , . . . , am ⊆ Il so that n=1 In ⊆ Il . This proves that
∞
n=1 In = Il , and thus In = Il for n ≥ l. Hence D is Noetherian.
From Theorem 3.1.1 we see that principal ideal domains are Noetherian.
Example 3.1.3 By Theorems 2.1.2 and 3.1.2 a Euclidean domain is always Noethe-
rian. Thus
Z (Theorem 2.2.1(a)),
√ √
Z + Z −1, Z + Z −2 (Theorem
2.2.3),
√ √
1 + −3 1 + −7
Z+Z , Z+Z ,
2 2
√
1 + −11
Z+Z (Theorem 2.2.5),
2
√ √ √
Z + Z 2, Z + Z 3, Z + Z 6 (Theorem 2.2.8)
are all examples of Noetherian domains.
Our next objective is to give another condition (called the maximal condition)
that allows us to recognize when an integral domain is Noetherian.
We show that satisfying the maximal condition is equivalent to the domain being
Noetherian.
Proof: Suppose that D is a Noetherian domain that does not satisfy the maximal
condition. Then D possesses a nonempty set S of ideals with the property that for
every ideal I of S there exists an ideal J of S with I ⊂ J . This property enables
us to construct inductively an infinite strictly ascending chain of ideals in S, which
contradicts D being a Noetherian domain. Hence every Noetherian domain must
satisfy the maximal condition.
Now let D be an integral domain that satisfies the maximal condition. Let I1 ⊆
I2 ⊆ I3 ⊆ . . . be an ascending chain of ideals of D. Set S = {In | n = 1, 2, 3, . . .}.
As D satisfies the maximal condition, S contains an ideal Im , which is not properly
contained in any other ideal of S. As Im ⊆ I j for j ≥ m we must have I j = Im
for j ≥ m. Hence the ascending chain I1 ⊆ I2 ⊆ I3 ⊆ . . . terminates and D is
Noetherian.
D = {a1 x r1 + · · · + an x rn | n ∈ N, a1 , . . . , an ∈ C, r1 , . . . , rn ∈ Q,
0 ≤ r1 < · · · < rn }.
CB609-03 CB609/Alaca & Williams August 7, 2003 17:19 Char Count= 0
58 Noetherian Domains
Clearly U (D) = C∗ . We show that D does not possesss any irreducible elements.
Suppose that
f (x) = a1 x r1 + · · · + an x rn
Then
f (x t ) = a1 x r1 t + · · · + an x rn t ∈ C[x].
a1 x r1 t + · · · + an x rn t = an (x − b1 ) · · · (x − brn t ).
Thus
Since
1/2 1/2
x 1/t − b1 = (x 1/2t − b1 )(x 1/2t + b1 ),
1/2
where x 1/2t ± b1 are nonzero, nonunit elements of D, f (x) is reducible, a con-
tradiction.
Thus D does not possess any irreducible elements.
Proof: Suppose that the integral domain D does not contain any irreducibles. As
we are assuming that D is not a field, D has nonzero, nonunit elements. Let a be
one of these. Then a is not an irreducible. Hence a is reducible. Thus there exists a
nonzero, nonunit element a1 of D such that a1 | a and a1 ∼ a. Clearly a ⊂ a1 .
As a1 is not an irreducible, a1 is reducible, and we can repeat the preceding argument
CB609-03 CB609/Alaca & Williams August 7, 2003 17:19 Char Count= 0
Clearly for this domain it is not possible to express each nonzero, nonunit element
as a finite product of irreducibles. Domains in which this is possible are called
factorization domains. The main result of this section is that a Noetherian domain
is always a factorization domain; that is, in a Noetherian domain every nonzero,
nonunit element can be expressed as a finite product of irreducibles. The converse
of this result however is not true as we demonstrate in Example 3.2.2.
Our next result shows that a Noetherian domain is always a factorization domain.
S = {a | a ∈ A}.
60 Noetherian Domains
The next example shows that a factorization domain is not always a Notherian
domain.
S = {a | a ∈ A}.
As l1 is not a unit we have b ⊂ b/l1 . Hence, by the maximality of b, we have
after suitable rearrangement of the h’s
k1 − 1 = j1 − 1, k2 = j2 , . . . , km = jm , m = n,
l1 ∼ h 1 , l2 ∼ h 2 , . . . , lm ∼ h m .
CB609-03 CB609/Alaca & Williams August 7, 2003 17:19 Char Count= 0
62 Noetherian Domains
This contradicts the assumption that b has two essentially different factorizations.
This completes the proof that a principal ideal domain is always a unique factor-
ization domain.
Clearly from Theorems 2.1.2 and 3.3.1 we see that a Euclidean domain is always
a unique factorization domain. Thus the domains listed in Example 3.1.3 are all
unique factorization domains.
√ √ √ √
Example 3.3.1 Z, Z + Z −1, Z + Z −2, Z + Z 1+ 2 −3 , Z + Z 1+ 2 −7 ,
√ √ √ √
Z + Z 1+ 2−11 , Z + Z 2, Z + Z 3, and Z + Z 6 are unique factorization
domains.
The next example shows that the converse of Theorem 3.3.1 is not true.
Example 3.3.6 The unique factorization domain Z[X ] is not a principal ideal
domain as it contains the nonprincipal ideal 2, X .
We remark that if a1 , . . . , an are all units then {π1 , . . . , πk } = φ. From (i), (ii), and
(iii), we see that
k
ai = i π j ei j , i = 1, 2, . . . , n,
j=1
where i ∈ U (D) and the ei j are nonnegative integers. (ei j is positive if and only if
π j | ai .) Set
e j = min ei j , j = 1, 2, . . . , k,
1≤i≤n
and
k
e
a= π j j ∈ D.
j=1
Clearly
a | ai , i = 1, 2, . . . , n,
and if b ∈ D is such that
b | ai , i = 1, 2, . . . , n,
then
b | a.
CB609-03 CB609/Alaca & Williams August 7, 2003 17:19 Char Count= 0
64 Noetherian Domains
3.4 Modules
Analogous to the concept of a vector space over a field is that of a module over a
ring. All rings are assumed to possess an identity.
CB609-03 CB609/Alaca & Williams August 7, 2003 17:19 Char Count= 0
3.4 Modules 65
Definition 3.4.1 (R-action) Let R be a ring with identity and M an additive Abelian
group. A function α : R × M → M is called an R-action on M if α has the following
properties:
α(r + s, m) = α(r, m) + α(s, m), (3.4.1)
α(r, m + n) = α(r, m) + α(r, n), (3.4.2)
α(r, α(s, m)) = α(r s, m), (3.4.3)
α(1, m) = m, (3.4.4)
for all r, s ∈ R and all m, n ∈ M.
It would be more accurate to call what we have just defined a left R-module.
There is a similar definition of a right R-module in which the elements of R are
written on the right. In this book we will keep to left modules throughout.
If M is an R-module with R-action α on M we write α(r, m) as r m to keep the
notation as simple as possible. With this convention (3.4.1)–(3.4.4) become
(r + s)m = r m + sm, (3.4.5)
r (m + n) = r m + r n, (3.4.6)
r (sm) = (r s)m, (3.4.7)
1m = m, (3.4.8)
valid for all r, s ∈ R, m, n ∈ M. Taking n = 0 in (3.4.6) and s = 0 in (3.4.5), we
deduce that r 0 = 0 (r ∈ R) and 0m = 0 (m ∈ M). The reader can easily check
from the axioms that
(−r )m = −(r m) = r (−m)
for all r ∈ R and all m ∈ M.
Example 3.4.1 If F is a field then an F-module is the same thing as a vector space
over F.
Example 3.4.3 Any ring R with identity can be thought of as a module over itself
in a natural way. We just take M to be the additive group R, + of R and define a
map R × R → R by (r, s) → r s (the product of r and s in R).
66 Noetherian Domains
Abelian group operations + and −, and the operation of “multiplying on the left”
by elements of R.
This definition is a valid one because the intersection of all the submodules of
M containing X is such a submodule and is thus the smallest such module. Indeed
if X is a nonempty subset of M then it is not difficult to show that the set
n
ri xi | ri ∈ R, xi ∈ X, n ≥ 1
i=1
of all finite sums of elements of the form r x with r ∈ R and x ∈ X is the smallest
submodule of M containing X , and so it is the submodule of M generated by X .
θ (m 1 + m 2 ) = θ(m 1 ) + θ (m 2 ),
θ (r m) = r θ (m),
Theorem 3.5.1 Let R be a ring with identity. Let M be an R-module and let N
be a submodule of M. Then M is Noetherian if and only if both N and M/N are
Noetherian.
N1 ⊆ N2 ⊆ . . .
M1 ⊆ M2 ⊆ . . .
CB609-03 CB609/Alaca & Williams August 7, 2003 17:19 Char Count= 0
68 Noetherian Domains
M = Rm 1 + Rm 2 + · · · + Rm n .
Each Rm i is an R-module.
For k = 1, 2, . . . , n we define the R-module Mk by
Mk = Rm 1 + · · · + Rm k
so that Mn = M.
We first show that each R-module Rm i (i = 1, . . . , n) is Noetherian. Let
Ni = {r ∈ R | r m i = 0}.
Rm k /Rm k ∩ Mk−1
is Noetherian. Hence
70 Noetherian Domains
√
In Example 3.1.3 we saw that the integral domain Z + Z m is Noetherian for
m = −1, −2, 2, 3, and 6. In fact this is true for an arbitrary integer m that is not a
perfect square.
√
Theorem 3.5.4 Let m be a nonsquare integer. Then Z + Z m is a Noetherian
domain (and thus a factorization domain by Theorem 3.2.2).
√
Proof: We take D = Z and E = Z + Z m in Theorem 3.5.3. As Z is Noetherian
√
(Example 3.1.3) and Z + Z m is a finitely generated Z-module (generated by 1
√
and m) the theorem follows from Theorem 3.5.3.
√
Similarly, taking D = Z and E = Z + Z 1+2 m , where m is a nonsquare inte-
ger with m ≡ 1 (mod 4), in Theorem 3.5.3, we obtain
Exercises 71
√
10 = ( 10)2 = 2 · 5,
√ √
where 2, 5, and 10 are nonassociated
√ irreducibles in Z + Z 10. We just show
that 2 is an irreducible in Z + Z 10. Suppose that
√ √
2 = (a + b 10)(c + d 10)
for some a, b, c, d ∈ Z. Then
4 = (a 2 − 10b2 )(c2 − 10d 2 ).
Hence, as a 2 − 10b2 ∈ Z, we deduce that
a 2 − 10b2 = −4, −2, −1, 1, 2, or 4.
√ √
If a 2 − 10b2 = ±1 then a +√b 10 is a unit of Z + Z√ 10. If a 2 − 10b2 = ±4 then
c2 − 10d 2 = ±1 and c + d 10 is a unit of Z + Z 10. If a 2 − 10b2 = ±2 then
a 2 ≡ ±2 (mod 5), which is a contradiction as a square is congruent to√0, 1, or
√ 2 is irreducible in Z + Z 10.√We
4 (mod 5). Thus this case cannot occur. Hence
√ to show that 5 and 10 are also irreducible in Z + Z 10
leave it to the reader
and that 2, 5, and 10 are not associates of one another (Exercise 13).
In Example 3.5.2 we showed that the equation x 2 − 10y 2 = 2 (or −2) has no
solutions in integers x and y. Here this was very easy to do: We just considered
the equation modulo 5 and got a contradiction. In general one cannot show that
an equation of the type x 2 − my 2 = N has no solutions in integers x and y by
congruence considerations alone. We show how to determine the solvability or in-
solvability of the equation x 2 − my 2 = N (m, n ∈ Z with m positive and nonsquare
√
and 0 < |N | < m) in Section 11.7.
Exercises
1. Let F be a field. If M is an F-module prove that M is a vector space over F. Conversely
show that if M is a vector space over F then M is an F-module.
2. Considering Z as a Z-module, where the Z-action on Z is just multiplication, determine
all the Z-submodules of Z.
3. Let I1 ⊆ I2 ⊆ . . . be an ascending chain of ideals in an integral domain D. Prove that
∞
n=1 In is an ideal in D.
4. Let F be a field. Is the domain F[X ] Noetherian?
5. Prove that the ideal 2, X in Z[X ] is not principal (Example 3.3.6).
6. Prove that a subset N of an R-module M is a submodule of M if and only if
(i) 0 ∈ N ,
(ii) n 1 , n 2 ∈ N =⇒ n 1 − n 2 ∈ N , and
(iii) n ∈ N , r ∈ R =⇒ r n ∈ N .
7. Prove that the intersection of any nonempty collection of submodules of an R-module
is itself a submodule of M.
CB609-03 CB609/Alaca & Williams August 7, 2003 17:19 Char Count= 0
72 Noetherian Domains
M1 + · · · + Mn = {m 1 + · · · + m n | m i ∈ Mi }.
11. Suppose that D is a unique factorization domain and a(= 0) and b(= 0) are coprime
nonunits in D. Prove that if ab = cn for some c ∈ D and some n ∈ N then there is a
unit e ∈ D such that ea and e−1 b are nth powers in D.
12. Let D be a unique factorization domain. Give an example to show that the following
assertion is not true in general: If a is an irreducible element of D then a is a maximal
ideal of D. √ √ √
13. Prove that 5 and 10 are irreducible elements of Z + Z 10 and that 2, 5, and 10 are
not associates of one another, as asserted in Example 3.5.2.
14. Let
pbe a prime and m be a positive nonsquare integer such that the Legendre symbol
±p
q
= −1 for some odd prime factor q of m. Prove that the equation x 2 − my 2 = ± p
√
has no solution in integers x and y. Deduce that p is an irreducible element of Z + Z m.
√
15. Prove that Z + Z −6 is not a unique factorization domain by exhibiting an element
√
of Z + Z −6 that has two different factorizations into irreducibles.
√
16. Prove that Z + Z√−10 is not a unique factorization domain.
17. Prove that Z + Z 15 is not a unique factorization domain.
Suggested Reading
1. P. M. Cohn, Unique factorization domains, American Mathematical Monthly 80 (1973),
1–18 (correction, American Mathematical Monthly 80 (1973), 1115).
A brief survey of unique factorization domains is given in the first six sections of the article.
2. D. S. Dummit and R. M. Foote, Abstract Algebra, second edition, Prentice Hall, Upper
Saddle River, New Jersey, 1999.
In Section 9.3 it is shown that D is a unique factorization domain if and only if D[X ] is a unique
factorization domain, and that if D is a unique factorization domain so is D[X 1 , X 2 , . . .].
3. J. Greene, Principal ideal domains are almost Euclidean, American Mathematical
Monthly 104 (1997), 154–156.
This paper is where Theorem 3.3.4 was first proved.
CB609-03 CB609/Alaca & Williams August 7, 2003 17:19 Char Count= 0
Biographies 73
4. B. Hartley and T. O. Hawkes, Rings, Modules and Linear Algebra, Chapman and Hall,
London, New York, 1974.
Chapters 5 and 6 give a very nice introduction to modules.
5. N. Jacobson, Lectures in Abstract Algebra, Volume I, van Nostrand, Princeton,
New Jersey, 1955.
Chapter VI contains a proof of the Hilbert basis theorem.
6. W. Rudin, Unique factorization of Gaussian integers, American Mathematical Monthly
68 (1961), 907–908.
A very simple and short proof is given that the Gaussian domain is a unique factorization domain.
7. P. Samuel, Unique factorization, American Mathematical Monthly 75 (1968), 945–952.
A classic overview of unique factorization is given.
8. O. Zariski and P. Samuel, Commutative Algebra, Volume 1, van Nostrand Company,
Princeton, New Jersey, 1967.
Chapter 4 gives a proof of the Hilbert basis theorem.
Biographies
1. A. Dick, Emmy Noether, 1882–1935, Birkháüser, Boston, Massachusetts, 1981.
This biography of Emny Noether includes obituaries by B. L. van der Waerden, H. Weyl, and
P. S. Alexandrov as well as a list of her publications.
2. C. H. Kimberling, Emmy Noether, American Mathematical Monthly 79 (1972), 136–149
(addendum, American Mathematical Monthly 79 (1972), 755).
This personal biography has excerpts from articles on Noether by P. S. Aleksandrov and H. Weyl.
3. C. Reid, Hilbert, Springer-Verlag, Berlin, Heidelberg, New York, 1970.
This wonderful book covers the life of David Hilbert, a man deeply devoted to the world of logic
and mathematics.
4. The website
http://www-groups.dcs.st-and.ac.uk/˜history/
4
Elements Integral over a Domain
Definition 4.1.1 (Element integral over a domain) Let A and B be integral do-
mains with A ⊆ B. The element b ∈ B is said to be integral over A if it satisfies a
polynomial equation
x n + an−1 x n−1 + · · · + a1 x + a0 = 0,
where a0 , a1 , . . . , an−1 ∈ A.
74
CB609-04 CB609/Alaca & Williams August 7, 2003 16:37 Char Count= 0
Definition 4.1.3 (Element algebraic over a field) Let A and B be integral domains
with A ⊆ B. Suppose that A is a field and b ∈ B is integral over A; then b is said
to be algebraic over A.
x 2 − 2ux + u 2 − mv 2 = 0,
bn + an−1 bn−1 + · · · + a1 b + a0 = 0.
cn + an−1 cn−1 + · · · + a1 c + a0 = 0.
Hence
Also
bk ∈ Abn−1 + · · · + Ab + A
for all nonnegative integers k. This shows that the integral domain A[b] of polyno-
mials in b with coefficients in A is a finitely generated A-module.
Conversely suppose that A[b] is a finitely generated A-module. Then there exist
u 1 , u 2 , . . . , u n ∈ A[b] such that
A[b] = Au 1 + · · · + Au n .
Clearly u 1 , . . . , u n are not all zero. Now each u i ∈ A[b] and so bu i ∈ A[b] for
i = 1, 2, . . . , n. Thus there exist ai j ∈ A (i, j = 1, 2, . . . , n) such that
bu 1 = a11 u 1 + · · · + a1n u n ,
···
bu n = an1 u 1 + · · · + ann u n .
bn + an−1 bn−1 + · · · + a1 b + a0 = 0,
The proof of our next theorem follows closely that of the previous theorem.
A[b] ⊆ C ⊆ B
C = Ac1 + · · · + Acn .
Clearly c1 , . . . , cn are not all zero. Now b ∈ A[b] and A[b] ⊆ C so that b ∈ C. But
C is an integral domain so that bc1 , . . . , bcn ∈ C. Hence there exist ai j ∈ A (i, j =
1, 2, . . . , n) such that
bc1 = a11 c1 + · · · + a1n cn ,
···
bcn = an1 c1 + · · · + ann cn .
that is,
b − a11 −a12 ··· −a1n
−a21 b − a22 ··· −a2n
.. .. .. = 0.
. . .
−an1 −an2 ··· b − ann
Expanding this determinant we obtain an equation
bn + an−1 bn−1 + · · · + a1 b + a0 = 0,
where a0 , a1 , . . . , an−1 ∈ A. Hence b is integral over A and, by Theorem 4.1.3,
A[b] is a finitely generated A-module.
The special case C = B in Theorem 4.1.4 shows that if A and B are integral
domains with A ⊆ B and B is a finitely generated A-module then B is integral over
A.
so that
C = Ab1 c1 + · · · + Abm cn
is a finitely generated A-module.
CB609-04 CB609/Alaca & Williams August 7, 2003 16:37 Char Count= 0
A ⊆ A[λ] ⊆ A[b1 , b2 ] ⊆ B,
Theorem 4.1.7 Let A and B be integral domains with A ⊆ B. Then the set of all
elements of B that are integral over A is a subdomain of B containing A.
Suppose first that b1 ∈ B is integral over A. Then, by Theorem 4.1.6 and Ex-
ample 4.1.12, we deduce that a0 + a1 b1 + · · · + an b1n is integral over A for all
a0 , a1 , . . . , an ∈ A. This proves that A[b1 ] is integral over A.
Next let b1 , . . . , bn−1 ∈ B be integral over A and suppose that A[b1 , . . . , bn−1 ]
is integral over A. Let bn ∈ B be integral over A. Let f be any element of
A[b1 , . . . , bn ]. Then
f = f 0 + f 1 bn + · · · + f m bnm ,
where f 0 , f 1 , . . . , f m ∈ A[b1 , . . . , bn−1 ]. By the inductive hypothesis
f 0 , f 1 , . . . , f m are all integral over A. Then, as bn is integral over A, we
deduce by Theorem 4.1.6 that f 0 + f 1 bn + · · · + f m bnm is integral over A. Hence
every element f of A[b1 , . . . , bn ] is integral over A, proving that A[b1 , . . . , bn ] is
integral over A.
Theorem 4.2.1 Let A and B be integral domains with A ⊆ B. Then the integral
closure A B of A in B is an integral domain satisfying
A ⊆ A B ⊆ B.
A B = Z + Zi.
x 2 − 2mx + (m 2 + n 2 ) ∈ Z[x].
f (x) = q(x)g(x) + r0 + r1 x.
r0 + r1 α = 0
so that
r0 + r1r = r1 s = 0.
f (x) = q(x)g(x),
where q(x), g(x) ∈ Q[x]. Let a be the least common multiple of the denominators
of the coefficients of q(x) and b the least common multiple of the denominators
of g(x). Then ab f (x) = aq(x)bg(x), where aq(x) and bg(x) ∈ Z[x]. Let c be the
content of aq(x) and d the content of bg(x). (Recall that the content of a nonzero
polynomial an x n + · · · + a1 x + a0 ∈ Z[x] is the greatest common divisor of the
integers an , . . . , a1 , a0 and that a primitive polynomial is a polynomial of Z[x] with
content 1.) Then we have aq(x) = cq1 (x) and bg(x) = dg1 (x), where q1 (x) ∈ Z[x]
and g1 (x) ∈ Z[x] are both primitive polynomials. Also ab f (x) = cq1 (x)dg1 (x).
Since f (x) ∈ Z[x] is monic the content of ab f (x) is ab. By a theorem of Gauss,
the product of two primitive polynomials is primitive. Hence q1 (x)g1 (x) is primitive
and the content of cq1 (x)dg1 (x) is cd. Thus ab = cd and f (x) = q1 (x)g1 (x), where
a b
q1 (x) = q(x) ∈ Z[x], g1 (x) = g(x) ∈ Z[x].
c d
Suppose that
Theorem 4.2.4 Q ∩ = Z.
Theorem 4.2.4 tells us that a rational algebraic integer must be an ordinary integer.
We will use this result on a number of occasions.
If D is an integral domain and F its field of quotients then it may happen that
the integral closure D F of D in F is equal to D. If this happens we say that D
is integrally closed. Apparently the term “integrally closed” was first defined by
Ernst Steinitz (1871–1928) in 1912, but the importance of the concept was already
known to Richard Dedekind (1831–1916).
Theorem 4.2.6 Every algebraic number is of the form a/b, where a is an algebraic
integer and b is a nonzero ordinary integer.
Exercises
1. Prove that
1
1 + 101/3 + 102/3
3
is an algebraic integer.
2. Prove that
102/3 − 1
√
−3
is an algebraic integer.
3. Let m and n be distinct squarefree integers such that m ≡ n ≡ 3 (mod 4). Let l = (m, n)
and set m = lm 1 , n = ln 1 so that (m 1 , n 1 ) = 1. If x0 , x1 , x2 , x3 are integers such that
prove that
1
√ √ √ √
x0 + x1 m + x2 n + x3 m 1 n 1
2
is an algebraic integer.
4. Express the algebraic number
√
1/3 √
1/3
1+ 2 1− 2
+
9 9
Biographies 87
√ √ √
12. Prove that Z + Z 2 + Z 5 + Z 10 is not integrally closed in its quotient field. [Hint:
√ 2
Consider 1+√2 5 .]
√ √ √
√ that the integral closure of Z + Z 5 in the field Q( 5, i) = {a + bi + c 5 +
13. Prove
di 5 | a, b, c, d ∈ Q} is
√
1+ 5
α + iβ | α, β ∈ Z + Z .
2
√ √ √
14. Prove
√ that the integral closure of Z + Z 5 in the field Q( 5, ω) = {a + bω + c 5+
dω 5 | a, b, c, d ∈ Q}, where ω is a primitive cube root of unity, is
√
1+ 5
α + βω | α, β ∈ Z + Z .
2
15. Let A and B be integral domains with A ⊆ B and B integral over A. If I is a nonzero
ideal of B, prove that I ∩ A is a nonzero ideal of A.
Suggested Reading
1. T. W. Atterton, A note on certain subsets of algebraic integers, Bulletin of the Australian
Mathematical Society 1 (1969), 345–352.
√ √
√ author√shows for example that the integral closure of Z + Z 5 in Q( 5, i) = {a + bi +
The
c 5 + di 5 | a, b, c, d ∈ Q} is
√
1+ 5
α + iβ | α, β ∈ Z + Z .
2
On page 141 it is mentioned that Steinitz showed how a small number of abstract ideas, such
as an irreducible ideal, chain conditions, and an integrally closed ring, could be used to prove
general results characterizing Dedekind rings and that the last two of these ideas had already been
introduced by Dedekind.
Biographies
1. The website
http://www-groups.dcs.st-and.ac.uk/˜history/
has biographies of both Richard Dedekind (1831–1916) and Ernst Steinitz (1871–1928).
CB609-05 CB609/Alaca & Williams August 7, 2003 16:40 Char Count= 0
5
Algebraic Extensions of a Field
Clearly the set I K (α) contains the zero polynomial. It is easy to check that I K (α) is
an ideal of K [x]. Moreover, I K (α) = 0 as g(x) ∈ I K (α).
As K is a field, by Theorem 2.2.1(b) we know that K [x] is a Euclidean domain
and thus, by Theorem 2.1.2, a principal ideal domain. Hence there exists p(x) ∈
K [x] such that
Suppose p1 (x) ∈ K [x] is another polynomial that generates I K (α), that is,
I K (α) = p1 (x) .
Then
p(x) = p1 (x)
U (K [x]) = K ∗ ,
so that
u(x) ∈ K ∗ .
88
CB609-05 CB609/Alaca & Williams August 7, 2003 16:40 Char Count= 0
This shows that we may take the polynomial p(x) to be monic, in which case p(x)
is uniquely determined by (5.1.2).
I K (α) = p(x)
Proof: Suppose that irr K (α) is reducible in K [x]. Then there exist nonzero poly-
nomials r (x) ∈ K [x] and s(x) ∈ K [x] such that
with r (x) ∈
/ U (K [x]) and s(x) ∈
/ U (K [x]). Hence r (x) ∈
/ K and s(x) ∈
/ K so that
deg r (x) ≥ 1 and deg s(x) ≥ 1. Thus
deg(irr K (α)) = deg r (x) + deg s(x) > max(deg r (x), deg s(x)). (5.1.4)
so that
and thus
Proof: Suppose that α has two conjugates over K that are the same. Then irr K (α)
has a root of order at least 2. Let β ∈ C be such a multiple root. Then
Thus β is a root of the derivative irr K (α) of irr K (α). As irr K (α) ∈ K [x] we have
so that
and thus
Theorem 5.3.1 If α is an algebraic integer then its conjugates over Q are also
algebraic integers.
Since h(x) ∈ Q[x] and h(α) = 0 we have h(x) ∈ IQ (α) = irrQ (α) so that
for some q(x) ∈ Q[x]. Let β be a conjugate of α over Q. Then β is also a root of
irrQ (α). Hence h(β) = 0 and so β is also an algebraic integer.
p | a1 , . . . , p | an−1 , p | an , p 2 an .
irrQ (α) = x 3 + 6x + 2.
where ω is a complex cube root of unity. Thus α and α are also algebraic integers.
α1 + · · · + αn ∈ Q,
α1 α2 + · · · + αn−1 αn ∈ Q,
···
α1 α2 · · · αn ∈ Q.
CB609-05 CB609/Alaca & Williams August 7, 2003 16:40 Char Count= 0
α1 + · · · + αn , α1 α2 + · · · + αn−1 αn , . . . , α1 α2 · · · αn
are all algebraic integers. Since they are all rational, by Theorem 4.2.4 they must
in fact be ordinary integers. Hence irrQ (α) ∈ Z[x].
then the monic polynomial of least degree in Q[x] of which α is a root belongs to
Z[x].
We use Theorem 5.3.2 to prove the following result (compare Theorem 4.2.2).
√
Theorem
√ √ integral closure of A = Z + Z −3 in the field B =
5.3.3 The
Q( −3) = {a + b −3 | a, b ∈ Q} is
√
1 + −3
A =Z+Z
B
.
2
√ √
Proof: Let α ∈ Z + Z 1+ 2 −3 . Then α = m + n 1+ 2 −3 for some m, n ∈ Z.
Clearly α ∈ B. As α is a root of the monic polynomial
x n + α1 x n−1 + · · · + αn ∈ A[x].
CB609-05 CB609/Alaca & Williams August 7, 2003 16:40 Char Count= 0
Hence α is an algebraic integer and so, by Theorem 5.3.2, irrQ (α) ∈ Z[x], that is
a ∈ Z, if b = 0,
2a, a 2 + 3b2 ∈ Z, if b = 0.
√ √ √
In the former case α = a + b −3 = a = a + 0 1+ 2 −3 ∈ Z + Z 1+ 2 −3 . In
the latter case we have a = m/2 for some m ∈ Z. If m ∈ 2Z then a ∈ Z and b ∈ Z.
If m ∈ 2Z + 1 then 2b ∈ 2Z + 1. Hence, in both cases, we see that a = m/2 and
b = m/2 + n, where m, n ∈ Z. Thus
√ √
m m √ 1 + −3 1 + −3
α= + +n −3 = −n + (m + 2n) ∈Z+Z .
2 2 2 2
√
Hence A B ⊆ Z + Z 1+ 2 −3 .
This completes the proof that
√
1 + −3
AB = Z + Z .
2
we deduce that
Q(α) = {x + yα | x, y ∈ Q},
Theorem 5.4.1 Let K be a quadratic field. Then there exists a unique squarefree
√
integer m such that K = Q( m).
α1 + α2 = −a ∈ Q
we have
Q(α1 ) = Q(α2 )
so that
√
−a + a 2 − 4b √
K = Q(α) = Q(α1 ) = Q = Q( c),
2
c = p/q,
q > 0, ( p, q) = 1.
Let r 2 denote the largest square dividing pq. Then pq = r 2 m, where m is a square-
free integer (= 1) and
√
√
p √ √ √
K =Q c =Q = Q( pq) = Q( r 2 m) = Q(r m) = Q( m).
q
√
Now let n be another squarefree integer such that K = Q( n). Hence
√ √
Q( m) = Q( n)
CB609-05 CB609/Alaca & Williams August 7, 2003 16:40 Char Count= 0
and so
√ √
m=x+y n
If x y = 0 then
√ m − x 2 − ny 2
n= ,
2x y
√
contradicting that n∈
/ Q as n is squarefree. Hence x y = 0. If y = 0 then
√
m = x,
√ √ √
contradicting that m ∈
/ Q as m is squarefree. Thus x = 0 and m = y n so that
m = y 2 n.
√
We next determine the algebraic integers in the quadratic field K = Q( m) =
√
{a + b m | a, b ∈ Q}, where m is a squarefree integer. The set of algebraic integers
in K is denoted by O K .
Theorem 5.4.2 Let K be a quadratic field. Let m be the unique squarefree integer
√
such that K = Q( m). Then the set O K of algebraic integers in K is given by
√
Z + Z m, √ if m ≡ 1 (mod 4),
OK = 1+ m
Z + Z , if m ≡ 1 (mod 4).
2
√
checked that the elements of Z + Z m if m ≡ 1 √
Proof: It is√easily (mod 4) and of
1+ m
Z+Z 2
if m ≡ 1 (mod 4) are algebraic integers in K = Q( m). Thus
√
Z + Z m, √ if m ≡ 1 (mod 4),
OK ⊇ 1+ m
Z + Z , if m ≡ 1 (mod 4).
2
We complete the proof by showing the inclusion in the reverse direction. Let α ∈
√
O K . Then α ∈ K and so α = a + b m for some a, b ∈ Q. Thus α is a root of the
monic polynomial
√
If b = 0 we have α = a ∈ Z ⊂ Z + Z m. Now suppose that b = 0. If 2a ∈ 2Z
then a ∈ Z and so mb2 ∈ Z. Since m is squarefree we see that b ∈ Z. In this case
√ √
α = a + b m ∈ Z + Z m. If 2a ∈ 2Z + 1 then as 4(a 2 − mb2 ) ∈ Z we deduce
that 4mb2 ∈ Z. As m is squarefree we have 2b ∈ Z. If 2b ∈ 2Z then b ∈ Z and so
a 2 = (a 2 − mb2 ) + mb2 ∈ Z,
√
The quadratic field K = Q( m), where m is a squarefree integer, is said to be
real if K ⊆ R and imaginary if K ⊆ R. Clearly K is real if m > 0 and imaginary
if m < 0. We close this section by determining the unit group U (O K ) when K is
an imaginary quadratic field.
CB609-05 CB609/Alaca & Williams August 7, 2003 16:40 Char Count= 0
where the intersection is taken over all subfields F of C, which contain both K and
α. The intersection is nonempty as C itself is such a field. Since the intersection of
subfields of C is again a subfield of C, K (α) is the smallest field containing both
K and α. We say that K (α) is formed from K by adjoining a single element α. A
subfield L of C for which there exists α ∈ C such that L = K (α) is called a simple
extension of K .
f (x) = b0 + b1 x + · · · + bk x k ∈ K [x],
g(x) = c0 + c1 x + · · · + ch x h ∈ K [x],
and
g(α) = 0.
This implies that irr K (α) g(x) and, since irr K (α) is irreducible in K [x], that
n(α)g(α) = 1
so that
1
= n(α)
g(α)
and thus
f (α)
β= = f (α)n(α).
g(α)
Hence each element β of K (α) can be expressed as a polynomial in α with coeffi-
cients in K , say
β = d0 + d1 α + · · · + dl αl ,
h(x) = d0 + d1 x + · · · + dl x l ∈ K [x],
so that β = h(α). As K is a field we can divide h(x) by irr K (α) to obtain polynomials
u(x) ∈ K [x] and v(x) ∈ K [x] such that
Theorem 5.5.1 shows that K (α) can be viewed as an n-dimensional vector space
over K with basis {1, α, . . . , α n−1 }. The dimension n is called the degree of the
extension K (α) over K .
√
Example 5.5.1 Let m be a squarefree integer. Then m ∈ C is a root of the
√ √ √
polynomial x 2 − m ∈ Q[x]. Now x 2 − m = (x − m)(x + m), where ± m ∈ /
Q as m is squarefree, so that x − m is irreducible in Q[x], and thus
2
√
irrQ ( m) = x 2 − m.
By Theorem 5.5.1 we have
√ √
Q( m) = {a0 + a1 m | a0 , a1 ∈ Q}
and
√
[Q( m) : Q] = 2,
√
so that Q( m) is a quadratic extension of Q.
Then
√ √ √
α 3 = (5 + 17) + 3(5 + 17)2/3 (5 − 17)1/3
√ √ √
+ 3(5 + 17)1/3 (5 − 17)2/3 + (5 − 17)
√ √ √ √
= 10 + 3(5 + 17)1/3 (5 − 17)1/3 ((5 + 17)1/3 + (5 − 17)1/3 )
√ √
= 10 + 3((5 + 17)(5 − 17))1/3 α
= 10 + 3 · 81/3 α
= 10 + 6α,
Q(α) = {a0 + a1 α + a2 α 2 | a0 , a1 , a2 ∈ Q}
and
and
deg(irrQ (ω)) = p − 1.
CB609-05 CB609/Alaca & Williams August 7, 2003 16:40 Char Count= 0
√ √
Example 5.5.4 The√quadratic field Q( −3) = {a + b −3 | a, b ∈ Q} is a cy-
clotomic field as Q( −3) = Q(ω), where ω is a complex cube root of unity.
Proof: Let
p(x) = irr K (α), q(x) = irr K (β).
Then
p(x) = (x − α1 ) · · · (x − αm ) ∈ K [x],
where
α1 = α, α2 , . . . , αm
CB609-05 CB609/Alaca & Williams August 7, 2003 16:40 Char Count= 0
contradicting the choice of c. Now let h(x) = irr K 1 (β). Then h(x) | p1 (x) and h(x) |
q(x). Since p1 (x) and q(x) have exactly one common root in C, we must have
deg h(x) = 1. Thus h(x) = x + δ for some δ ∈ K 1 . Now 0 = h(β) = β + δ so that
β = −δ ∈ K 1 . Then α = γ − cβ ∈ K 1 . This shows that
K (α, β) ⊆ K 1 = K (γ ).
Since γ = α + cβ ∈ K (α, β) we have
K (γ ) ⊆ K (α, β)
and thus
K (α, β) = K (γ ).
Set
√ √
α= 2+ 3 ∈ R.
Squaring α we obtain
√
α 2 = 5 + 2 6,
so that
√
α 2 − 5 = 2 6.
Squaring α 2 − 5 we get
α 4 − 10α 2 + 25 = 24.
Thus α is a root of the monic quartic polynomial
f (x) = x 4 − 10x 2 + 1 ∈ Z[x].
This shows that α is an algebraic integer.
We now show that f (x) is irreducible in Z[x] and thus in Q[x]. Since f (±1) =
−8 = 0, f (x) has no linear factors in Z[x]. Thus if f (x) factors in Z[x], it must
factor as a product of two quadratic polynomials in Z[x], say,
x 4 − 10x 2 + 1 = (x 2 + ax + b)(x 2 + cx + d),
where a, b, c, d ∈ Z. Equating coefficients of x 3 , x 2 , x, and 1, we obtain
a + c = 0,
b + ac + d = −10,
bc + ad = 0,
bd = 1.
From the first equation we have c = −a, so the second equation becomes
b + d + 10 = a 2 .
From the last equation we have b = d = ±1, so that b + d = ±2. Hence a 2 = 8
or 12, which is impossible. This proves that x 4 − 10x 2 + 1 is irreducible in Q[x]
and so
√ √
irrQ ( 2 + 3) = x 4 − 10x 2 + 1
and
√ √ √ √
[Q( 2, 3) : Q] = [Q( 2 + 3) : Q] = 4.
√ √
Example
√ √ 5.6.2 We
√ express Q( 3, 3 2) as
√ a simple
√ extension.
√ The√conjugates of
3 are 3 and − 3. The conjugates of 3 2 are 3 2, ω 3 2, and ω2 3 2, where ω is
CB609-05 CB609/Alaca & Williams August 7, 2003 16:40 Char Count= 0
√ √ √ √ √ √
− 3 + ω 2, 3 + ω2 2, − 3 + ω2 2
3 3 3
We conclude this chapter by proving the very important fact that every element
of a simple extension K (α) of a subfield K of C, where α is algebraic over K , is
algebraic over K .
Proof: Let β ∈ K (α), where α is algebraic over K . Let n = deg(irr K (α)). By The-
orem 5.5.1 each of the powers β j , j = 0, 1, . . . , n, of β can be written in the
form
n−1
βj = a jk α k ,
k=0
has a solution (x0 , x1 , . . . , xn ) ∈ K n+1 with not all of the x j equal to zero, as the
number of unknowns is greater than the number of equations. Then
n
n
n−1
n−1
n
xjβ j = xj a jk α k = αk a jk x j = 0,
j=0 j=0 k=0 k=0 j=0
proving that β is algebraic over K and that the degree of β over K is less than or
equal to the degree of α over K .
Exercises
1. Prove that the set I K (α) defined in Section 5.1 is an ideal.
2. Prove that x 4 + 1
√is irreducible in Q[x] (see Example
√ 5.1.1).
3. Prove that x 2 − 2x + 1 is irreducible in Q( 2)[x] (see Example 5.1.2).
CB609-05 CB609/Alaca & Williams August 7, 2003 16:40 Char Count= 0
Exercises 107
4. Determine
1+i
irrQ(i) √
2
and
1+i
irrQ(√−2) √ .
2
√ √
5. Prove that [Q(√ 3 + 3 2) :√Q] = 6 (see Example 5.6.2).
6. Prove that Q(√2, i)√= Q( 2√ + i). √
7. Prove that Q( 2, i 2) = Q( 2 + i 2).
8. Find the minimal polynomial of 21/3 + ω over Q(21/3 ), where ω is a complex cube root
of unity.
9. Determine α ∈ C such that
√ √ √
Q( 2, 3, 5) = Q(α).
10. Prove that
√ √ √
[Q( 2, 3, 5) : Q] = 8.
11. Determine the conjugates of 31/3 − 32/3 .
12. Let θ ∈ C be a root of x 3 + 11x + 4 = 0. Prove that [Q(θ ) : Q] = 3.
13. Prove that (−θ + θ 2 )/2 is an algebraic integer in K = Q(θ ), where θ 3 + 11θ − 4 = 0.
√
14. Let θ ∈ C be a root of x 5 + x + 1 = 0. If θ ∈ Q( −3), what is irrQ θ ?
15. Let ω = e2πi/5 . Prove that
1 √ √
ω= 5 − 1 + i 10 + 2 5 .
4
√ √
16. Let ω = e2πi/5 . Show that 5 ∈ Q(ω) by expressing 5 in the form
√
5 = aω + bω2 + cω3 + dω4
for suitable integers a, b, c, d.
17. Prove that
1 √ √
i 10 + 2 5 + 2i 10 − 2 5
2
is an algebraic integer in Q(e2πi/5 ).
18. Determine the conjugates of
121/5 + 541/5 − 1441/5 + 6481/5
over Q.
√
19. Let m be a squarefree integer ≡ 1 (mod 4). Let A =Z+Z m and
√
B = Q( m). Prove that
√
1+ m
A =Z+Z
B
.
2
20. Let θ be a nonreal algebraic number. Prove that the complex conjugate θ̄ of θ is one of
the conjugates of θ over Q.
CB609-05 CB609/Alaca & Williams August 7, 2003 16:40 Char Count= 0
Suggested Reading
1. E. R. Scheinerman, When close enough is close enough, American Mathematical
Monthly 107 (2000), 489–499.
A technique is presented for proving identities involving algebraic integers numerically. For ex-
ample Shanks’s identity [2]
√ √ √ √ √
5 + 22 + 2 5 = 11 + 2 29 + 16 − 2 29 + 2 55 − 10 29
can be proved using this technique.
2. D. Shanks, Incredible identities, Fibonacci Quarterly 12 (1974), 271, 280.
Biographies
1. H. C. Williams, Daniel Shanks (1917–1996), Notices of the American Mathematical
Society 44 (1997), 813–816.
2. H. C. Williams, Daniel Shanks (1917–1996), Mathematics of Computation 66 (1997),
929–934.
CB609-06 CB609/Alaca & Williams August 27, 2003 16:53 Char Count= 0
6
Algebraic Number Fields
109
CB609-06 CB609/Alaca & Williams August 27, 2003 16:53 Char Count= 0
The form of the elements in an algebraic number field follows immediately from
Theorem 5.5.1. We have
Definition 6.1.2 (The set O K ) The set of all algebraic integers that lie in the
algebraic number field K is denoted by O K ; that is,
OK = ∩ K .
and
√
1+ −3
OQ(√−3) = Z + Z .
2
√
Thus the Gaussian domain √ Z + Z −1 (Example 1.1.2) is the ring of inte-
gers √Q( −1) and the Eisenstein domain Z + Zω = Z +
of √the quadratic field
−1+ −3
Z 2
= Z + Z 1+ 2 −3 (Example 1.1.3) is the ring of integers of the
√
quadratic field Q( −3).
When K is not a quadratic field, it is a more difficult problem to determine O K .
We determine O K for some algebraic number fields of degree > 2 in Chapter 7.
Indeed it is an area of current research to determine O K explicitly for certain
classes of algebraic number fields K . See the references at the end of Chapter 7 in
this connection.
Proof: Let F denote the quotient field of O K , and let α ∈ F. Then α = b/c, where
b ∈ O K and c ∈ O K with c
= 0. As O K ⊆ K we have b ∈ K and c ∈ K so that,
as K is a field, α = b/c ∈ K . Hence F ⊆ K .
Now let α ∈ K . By Theorem 4.2.6 we have α = b/c, where b is an algebraic
integer and c is a nonzero rational integer. Clearly b = αc ∈ K so b ∈ O K . Thus
α = b/c ∈ F. Hence K ⊆ F.
This proves that F = K , so the quotient field of O K is K .
Theorem 6.1.7 Let K be an algebraic number field. Then every nonzero ideal in
O K contains a nonzero rational integer.
CB609-06 CB609/Alaca & Williams August 27, 2003 16:53 Char Count= 0
Proof: Let I
= {0} be an ideal in O K . Choose α ∈ I with α
= 0. As α ∈ I ⊆ O K , α
is an algebraic integer. Let irrQ (α) = x n + b1 x n−1 + · · · + bn . We show that bn
= 0.
If n = 1 then irrQ (α) = x + b1 so that α + b1 = 0. Hence b1 = −α
= 0. If n ≥ 2
then bn
= 0 as irrQ (α) is irreducible in Q[x] by Theorem 5.1.1. By Theorem 5.3.2
we know that irrQ (α) ∈ Z[x] as α is an algebraic integer. Hence b1 , . . . , bn ∈ Z.
Thus bn = −α n − b1 α n−1 − · · · − b1 α ∈ I . Hence bn is a nonzero rational integer
in I .
Theorem 6.2.1 Let K be an algebraic number field of degree n over Q. Then there
are exactly n distinct monomorphisms σk : K → C (k = 1, . . . , n).
by
σk (a0 + a1 θ + · · · + an−1 θ n−1 ) = a0 + a1 θk + · · · + an−1 θkn−1 .
We show that σk (k = 1, 2, . . . , n) is a field homomorphism.
First we show that σk (k = 1, 2, . . . , n) is additive. Let α, β ∈ K . Then
α = a0 + a1 θ + · · · + an−1 θ n−1
and
β = b0 + b1 θ + · · · + bn−1 θ n−1 ,
where a0 , a1 , . . . , an−1 , b0 , b1 , . . . , bn−1 ∈ Q. Hence
α + β = (a0 + b0 ) + (a1 + b1 )θ + · · · + (an−1 + bn−1 )θ n−1
and so
σk (α + β) = (a0 + b0 ) + (a1 + b1 )θk + · · · + (an−1 + bn−1 )θkn−1
= (a0 + a1 θk + · · · + an−1 θkn−1 ) + (b0 + b1 θk + · · · + bn−1 θkn−1 )
= σk (α) + σk (β).
Thus σk is additive.
Next we show that σk (k = 1, 2, . . . , n) is multiplicative. With the same notation,
we let
f (x) = a0 + a1 x + · · · + an−1 x n−1 ∈ Q[x]
and
g(x) = b0 + b1 x + · · · + bn−1 x n−1 ∈ Q[x]
so that
f (θ) = α, g(θ) = β.
Dividing f (x)g(x) by p(x) in Q[x], we obtain a quotient q(x) ∈ Q[x] and a re-
mainder r (x) ∈ Q[x] such that
f (x)g(x) = p(x)q(x) + r (x), deg r (x) < deg p(x) = n.
Hence, as p(θ) = 0, we have
αβ = f (θ )g(θ ) = p(θ )q(θ) + r (θ) = r (θ).
Thus, as p(θk ) = 0, we have
σk (αβ) = σk (r (θ)) = r (θk ) = p(θk )q(θk ) + r (θk ) = f (θk )g(θk ) = σk (α)σk (β),
so that σk is multiplicative.
Hence we have shown that σk (k = 1, 2, . . . , n) is a homomorphism.
CB609-06 CB609/Alaca & Williams August 27, 2003 16:53 Char Count= 0
For k = 1, 2, . . . , n, we have
range σk = σk (K )
= {σk (a0 + a1 θ + · · · + an−1 θ n−1 ) | a0 , a1 , . . . , an−1 ∈ Q}
= {a0 + a1 θk + · · · + an−1 θkn−1 | a0 , a1 , . . . , an−1 ∈ Q}
= Q(θk )
CB609-06 CB609/Alaca & Williams August 27, 2003 16:53 Char Count= 0
so that
σk : Q(θ) → Q(θk )
is an isomorphism. Hence all the fields Q(θk ) (k = 1, 2, . . . , n) are isomorphic.
Proof: Let
n
n
f (x) = (x − φk ) = (x − (c0 + c1 θk + · · · + cn−1 θkn−1 )) ∈ K 1 [x],
k=1 k=1
(up to sign) the coefficients of irrQ (θ) ∈ Q[x], they are all rational numbers, and so
the coefficients of f (x) are all rational. Hence f (x) ∈ Q[x]. As f (φ) = 0 we have
irrQ (φ) | f (x), say f (x) = irrQ (φ)g(x), where g(x) ∈ Q[x]. Then
and
[Q(φk ) : Q] = [Q(θk ) : Q] (= n)
so that
Q(φk ) = Q(θk ), k = 1, 2, . . . , n.
α = c0 + c1 θ + · · · + cn−1 θ n−1
As
√ √ √ √ √ √ √ √
x 4 − 10x 2 + 1 = (x − 2 − 3)(x − 2 + 3)(x + 2 − 3)(x + 2 + 3)
CB609-06 CB609/Alaca & Williams August 27, 2003 16:53 Char Count= 0
Hence
a + 5c = 0, b + 11d = 0, b + 9d = 2, 2c = 0,
so that
a = 0, b = 11, c = 0, d = −1,
giving
α = 11θ − θ 3 .
Clearly fld K (α) ∈ C[x]. However, much more is true as the next theorem shows.
CB609-06 CB609/Alaca & Williams August 27, 2003 16:53 Char Count= 0
α = c0 + c1 θ + · · · + cn−1 θ n−1 .
αk = c0 + c1 θk + · · · + cn−1 θkn−1 , k = 1, 2, . . . , n.
n
n
fld K (α) = (x − αk ) = (x − (c0 + c1 θk + · · · + cn−1 θkn−1 )).
k=1 k=1
Example 6.3.2 The cubic polynomial x 3 + 11x + 4 ∈ Z[x] is irreducible. Let its
three roots be θ1 = θ, θ2 , and θ3 . One of these roots is real and the other two
are nonreal and complex conjugates of one another (Exercise 2 of this Chapter).
Let K = Q(θ ) so that [K : Q] = deg(irrQ (θ)) = deg(x 3 + 11x + 4) = 3. Let α =
(θ + θ 2 )/2 ∈ K . We determine fld K (α). We have
(θ1 + θ12 ) (θ2 + θ22 ) (θ3 + θ32 )
fld K (α) = x − x− x−
2 2 2
= x + a2 x + a1 x + a0 ,
3 2
CB609-06 CB609/Alaca & Williams August 27, 2003 16:53 Char Count= 0
where
(θ1 + θ12 ) (θ2 + θ22 ) (θ3 + θ32 )
a2 = − − −
2 2 2
1 1 2
= − (θ1 + θ2 + θ3 ) − (θ1 + θ22 + θ32 ),
2 2
1
a1 = ((θ1 + θ1 )(θ2 + θ2 ) + (θ2 + θ22 )(θ3 + θ32 ) + (θ3 + θ32 )(θ1 + θ12 ))
2 2
4
1
= ((θ1 θ2 + θ2 θ3 + θ3 θ1 ) + (θ1 θ22 + θ12 θ2 + θ2 θ32 + θ22 θ3 + θ3 θ12 + θ32 θ1 )
4
+ (θ12 θ22 + θ22 θ32 + θ32 θ12 )),
1
a0 = − (θ1 + θ12 )(θ2 + θ22 )(θ3 + θ32 )
8
1
= − θ1 θ2 θ3 (1 + θ1 )(1 + θ2 )(1 + θ3 )
8
1
= − θ1 θ2 θ3 (1 + (θ1 + θ2 + θ3 ) + (θ1 θ2 + θ2 θ3 + θ3 θ1 ) + θ1 θ2 θ3 ).
8
Now
x 3 + 11x + 4 = (x − θ1 )(x − θ2 )(x − θ3 ),
so that
θ1 + θ2 + θ3 = 0,
θ1 θ2 + θ2 θ3 + θ3 θ1 = 11,
θ1 θ2 θ3 = −4.
Hence
θ12 + θ22 + θ32 = (θ1 + θ2 + θ3 )2 − 2(θ1 θ2 + θ2 θ3 + θ3 θ1 ) = −22,
θ12 θ22 + θ22 θ32 + θ32 θ12 = (θ1 θ2 + θ2 θ3 + θ3 θ1 )2 − 2θ1 θ2 θ3 (θ1 + θ2 + θ3 ) = 121,
θ1 θ22 + θ12 θ2 + θ2 θ32 + θ22 θ3 + θ3 θ12 + θ32 θ1
= θ1 θ2 (θ2 + θ1 ) + θ2 θ3 (θ3 + θ2 ) + θ3 θ1 (θ1 + θ3 )
= −3θ1 θ2 θ3 = 12, as θ1 + θ2 + θ3 = 0,
so that
a2 = 11, a1 = 36, a0 = 4.
Hence
fld K (α) = x 3 + 11x 2 + 36x + 4,
showing that α ∈ O K .
In the next theorem we relate the field polynomial of α over K to the minimal
polynomial of α over Q.
CB609-06 CB609/Alaca & Williams August 27, 2003 16:53 Char Count= 0
where h(x) is a monic polynomial of Q[x], which is not divisible by the irreducible
polynomial irrQ (α), and s is a positive integer. Suppose that h(x) is a nonconstant
polynomial. Then h(αk ) = 0 for some k ∈ {1, 2, . . . , n}.
Now choose θ ∈ K such that K = Q(θ ). Let θ1 = θ, θ2 , . . . , θn be the conjugates
of θ over Q. As α ∈ K there exists a polynomial
Then g(θk ) = h(r (θk )) = h(αk ) = 0. Thus g(x) is a multiple of irrQ (θk ) = irrQ (θ) ∈
Q[x]. Hence g(θ j ) = 0 for j = 1, 2, . . . , n. In particular g(θ) = 0. Thus h(α) =
h(r (θ)) = g(θ) = 0. Hence h(x) is a multiple of irrQ (α) in Q[x], contradicting that
h(x) is not divisible by irrQ (α).
We have shown that h(x) is a constant polynomial; that is, h(x) = c, c ∈ Q. But
h(x) is monic so c = 1. Thus
so that
n
s= .
deg(irrQ (α))
Theorem 6.3.2 tells us that the conjugates of α with respect to K are the roots of
irrQ (α) in C each repeated s = n/deg(irrQ (α)) times.
Suppose α ∈ Q, then
αk = σk (α) = α, k = 1, 2, . . . , n,
so all of the K -conjugates of α are equal. Conversely, if all the K -conjugates of α
are equal then
fld K (α) = (x − α)n .
Hence, by Theorem 6.3.2, we have
(irrQ (α))s = (x − α)n .
But the roots of irrQ (α) are all distinct (Theorem 5.2.1) so that
irrQ (α) = x − α, s = n.
As irrQ (α) ∈ Q[x] we deduce that α ∈ Q. Hence we have shown the following
result.
Theorem 6.3.4 Let K be an algebraic number field. Let α ∈ K . Then all the
K -conjugates of α are equal if and only if α ∈ Q.
If the K -conjugates of α are all distinct then fld K (α) is a product of distinct linear
factors and so by Theorem 6.3.2 we have s = 1 and irrQ (α) = fld K (α). Hence
[Q(α) : Q] = deg(irrQ (α)) = deg(fld K (α)) = n = [K : Q].
CB609-06 CB609/Alaca & Williams August 27, 2003 16:53 Char Count= 0
Theorem 6.3.5 Let K be an algebraic number field. Let α ∈ K . Then all the K -
conjugates of α are distinct if and only if K = Q(α).
where ω and ω2 are the two complex cube roots of unity, since
√ √ √ √
irrQ ( 2) = x 3 − 2 = (x − 2)(x − ω 2)(x − ω2 2).
3 3 3 3
so that
√ √ √
2 = −b 2 + c( 2)2 .
3 3 3
Hence
√
1 + b = c 2.
3
√
Since 3
2∈
/ Q we must have 1 + b = c = 0, so that
√ √
ω2 2 = a − ω 2.
3 3
Thus
√
(ω2 + ω) 2 = a;
3
a contradiction. √ √
Hence all the conjugate fields of Q( 3 2) are distinct, and Q( 3 2) is not a normal
field.
√ √
Example 6.3.5 Let K = Q( 4 2) so that K ⊆ R. The conjugates of 4 2 are
√ √ √ √
2, i 2, − 2, − i 2,
4 4 4 4
as
√ √ √ √ √
irrQ ( 2) = x 4 − 2 = (x − 2)(x − i 2)(x + 2)(x + i 2).
4 4 4 4 4
√
Q(i 2) = L (say),
4
√ √
Q(− 2) = Q( 2) = K ,
4 4
√ √
Q(−i 2) = Q(i 2) = L .
4 4
Clearly K
= L as K is a real field and L is a nonreal
√ field.
Hence there are two distinct conjugate fields. Q( 2) is not a normal field.
4
√ √
Example 6.4.1 Let K = Q( 2, 3) and choose
√ √ √ √
ω1 = 1, ω2 = 2, ω3 = 3, ω4 = 2 + 3.
By Example 5.6.1 we know that K is a quartic field. The four monomor-
phisms : K −→ C are given by
√ √ √ √ √ √
σ1 (a + b 2 + c 3 + d 6) = a + b 2 + c 3 + d 6,
√ √ √ √ √ √
σ2 (a + b 2 + c 3 + d 6) = a + b 2 − c 3 − d 6,
√ √ √ √ √ √
σ3 (a + b 2 + c 3 + d 6) = a − b 2 + c 3 − d 6,
√ √ √ √ √ √
σ4 (a + b 2 + c 3 + d 6) = a − b 2 − c 3 + d 6,
where a, b, c, d ∈ Q. Hence
√ √ √ √ 2
1 +
√2 √3 √ 2 √3
√ √ √ √
1 √2 −√3 √2 − √3 .
D(1, 2, 3, 2 + 3) =
1 −√2 −√2 + √3
√3
1 − 2 − 3 − 2 − 3
As the fourth column of the determinant is the sum of the second and third columns,
we deduce that
√ √ √ √
D(1, 2, 3, 2 + 3) = 0.
α by
n−1
z1 z 1n−2 ··· z1 1
n−1
z z 2n−2 ··· z2 1
2
..
. .. .. .. = (z i − z j )
. ··· . . 1≤i< j≤n
z n−1 z nn−2 ··· zn 1
n
(see, for example, [3, pp. 17–18]). Interchanging columns 1 and n, columns 2 and
n − 1, etc., we obtain the evaluation of the Vandermonde determinant
1 z 1 · · · z 1n−2 z 1n−1
1 z 2 · · · z n−2 z n−1
2 2
.. .. . . = (−1)[n/2] (z i − z j )
. . ··· .. ..
1≤i< j≤n
1 z n · · · z n−2 z n−1
n n
n(n−1)
= (−1) 2 (z i − z j )
1≤i< j≤n
= (z j − z i ),
1≤i< j≤n
as
n n(n − 1)
≡ (mod 2).
2 2
Hence for any α ∈ K we have
1 α α2 ··· α n−1
1 α (2) (α (2) )
2
··· (α (2) )
n−1
.. .. .. .. = (α ( j) − α (i) )
. . . ··· .
1≤i< j≤n
1 α (n) (n)
(α )
2
··· (α (n) )n−1
CB609-06 CB609/Alaca & Williams August 27, 2003 16:53 Char Count= 0
so that
2
D(α) = (α ( j) − α (i) ) =
2
(α (i) − α ( j) ) .
1≤i< j≤n 1≤i< j≤n
where n ∈ N and an
= 0. Let x1 , . . . , xn ∈ C be the roots of f (x). The discriminant
of f (x) is the quantity
disc( f (x)) = an2n−2 (xi − x j )2 ∈ C.
1≤i< j≤n
Clearly f (x) has a repeated root if and only if disc( f (x)) = 0. The discriminant is
a 2n−2
times a symmetric polynomial in x1 , . . . , xn . The degree of disc( f (x)) in each
xi is 2(n − 1). Thus when disc( f (x)) is expressed as a function of a0 , a1 , . . . , an ,
it consists of terms
k n
an−1 k1 an−2 k2 a0
Can 2n−2
··· ,
an an an
where
k1 + k2 + · · · + kn ≤ 2n − 2,
Proof: Let α (1) = α, α (2) , . . . , α (n) be the K -conjugates of α. Then the roots of
fld K (α) are α (1) , . . . , α (n) . Hence, by Definition 6.4.3 and Theorem 6.4.1, we have
2
disc(fld K (α)) = (α (i) − α ( j) ) = D(α).
1≤i< j≤n
√ √
Example 6.4.2 Let K = Q( 3 2) and choose α = 3 2. Then the conjugates of α
are
√ √ √
α1 = 2, α2 = ω 2, α3 = ω2 2,
3 3 3
CB609-06 CB609/Alaca & Williams August 27, 2003 16:53 Char Count= 0
√
α2 − α3 = ω(1 − ω) 2,
3
√
α3 − α1 = ω2 (1 − ω) 2,
3
so that
(α1 − α2 )(α2 − α3 )(α3 − α1 ) = (1 − ω)3 2.
Now
(1 − ω)3 = 1 − 3ω + 3ω2 − ω3 = −3(ω − ω2 ),
so that
(1 − ω)6 = 32 (ω − ω2 )2 = 32 (ω2 + ω − 2) = −33 .
Hence
D(α) = ((α1 − α2 )(α2 − α3 )(α3 − α1 ))2 = (1 − ω)6 22 = −22 · 33 .
(b) If ω1 , . . . , ωn ∈ O K then
D(ω1 , . . . , ωn ) ∈ Z.
(c) If ω1 , . . . , ωn ∈ K then
D(ω1 , . . . , ωn )
= 0 if and only if ω1 , . . . , ωn are linearly independent over Q.
Proof: (a) By Theorem 6.1.1 we have K = Q(θ) for some θ ∈ K . Then, for i =
1, 2, . . . , n, we have
ωi = c0 i + c1 i θ + · · · + cn−1 i θ n−1 ,
where c0 i , . . . , cn−1 i ∈ Q. Hence, for j = 1, 2, . . . , n, we have
( j)
ωi = c0 i + c1 i θ j + · · · + cn−1 i θ n−1
j , (6.4.1)
CB609-06 CB609/Alaca & Williams August 27, 2003 16:53 Char Count= 0
where θ1 = θ, θ2 , . . . , θn are the conjugates of θ over Q and ωi(1) , . . . , ωi(n) are the
K -conjugates of ωi (i = 1, 2, . . . , n). Using the expressions (6.4.1) in Definition
6.4.1, we see that any permutation of the conjugates of θ leaves D(ω1 , . . . , ωn )
invariant as it merely causes a permutation of the rows of the matrix of which
D(ω1 , . . . , ωn ) is the square of the determinant. Hence D(ω1 , . . . , ωn ) is a sym-
metric function of the roots of the polynomial
(x − θ1 )(x − θ2 ) · · · (x − θn ) = x n + an−1 x n−1 + · · · + a0 ,
Now suppose that the set {ω1 , . . . , ωn } is linearly independent over Q. Then
{ω1 , . . . , ωn } is a basis for the vector space K over the field Q. In particular as
1, θ, . . . , θ n−1 ∈ K there exist rational numbers ci j (i, j = 1, . . . , n) such that
1 = c11 ω1 + · · · + c1n ωn ,
θ = c21 ω1 + · · · + c2n ωn ,
···
n−1
θ = cn1 ω1 + · · · + cnn ωn .
Hence
D(θ ) = D(1, θ, . . . , θ n−1 ) = |det(ci j )|2 D(ω1 , . . . , ωn ).
D(ω1 , . . . , ωn )
= 0.
CB609-06 CB609/Alaca & Williams August 27, 2003 16:53 Char Count= 0
D(η1 , . . . , ηn )
= 0.
η1 = c, η2 = c θ, . . . , ηn = c θ n−1 ∈ I
and
We are now in a position to prove that every ideal of the ring of integers of an
algebraic number field has a finite basis.
α = x1 η1 + · · · + xn ηn ,
where x1 , . . . , xn ∈ Z.
S = {|D(η1 , . . . , ηn )| : η1 , . . . , ηn ∈ I, D(η1 , . . . , ηn )
= 0}.
Clearly S is a nonempty set of positive integers and thus contains a least member,
say |D(η1 , . . . , ηn )|, η1 , . . . , ηn ∈ I. As D(η1 , . . . , ηn )
= 0, by Theorem 6.4.4(c)
{η1 , . . . , ηn } is a basis for the vector space K over Q. Let α ∈ I . Then there exist
unique rational numbers x1 , . . . , xn such that
α = x1 η1 + · · · + xn ηn .
CB609-06 CB609/Alaca & Williams August 27, 2003 16:53 Char Count= 0
l < x1 < l + 1.
Set
γ = α − lη1 .
γ = (x1 − l)η1 + x2 η2 + · · · + xn ηn .
Hence
D(γ , η2 , . . . , ηn )
(x1 − l)2 =
D(η1 , η2 , . . . , ηn )
so that
This contradicts the minimality of |D(η1 , η2 , . . . , ηn )|. Hence all the xi are integers
and each element α ∈ I can be expressed uniquely in the form α = x1 η1 + · · · +
xn ηn .
CB609-06 CB609/Alaca & Williams August 27, 2003 16:53 Char Count= 0
I ⊆ Zη1 + · · · + Zηn
so that
I = Zη1 + · · · + Zηn ,
α = x1 η1 + · · · + xn ηn (x1 , . . . , xn ∈ Z)
and
n
ηi = ci j λ j , i = 1, 2, . . . , n,
j=1
I = Zλ1 + · · · + Zλn .
I = Zη1 + · · · + Zηn .
C = [ci j ] , D = [di j ],
CB609-06 CB609/Alaca & Williams August 27, 2003 16:53 Char Count= 0
C D = In ,
so that
Hence
(det(di j ))2 = 1
so that
det(di j ) = ±1.
Thus the matrix D = (di j ) has an inverse D −1 = C = (ci j ) all of whose entries are
integers, and
n
ηj = ci j λ j (i = 1, 2, . . . , n).
j=1
Hence
n
n
n
n
α= ai ci j λ j = ai ci j λj,
i=1 j=1 j=1 i=1
n
where each i=1 ai ci j ∈ Z ( j = 1, 2, . . . , n). This proves that every element α of
I can be expressed in the form
α = b1 λ1 + · · · + bn λn
α = b1 λ1 + · · · + bn λn = b1 λ1 + · · · + bn λn ,
e1 λ1 + · · · + en λn = 0,
D(λ1 , . . . , λn ) = 0.
Hence
D(η1 , . . . , ηn ) = 0,
√ √ √
Example 6.5.1 Let K = Q( 7) so that O K = Z + Z 7 = {a + b 7 |√a, b ∈ Z}
by Theorem 5.4.2. Let I be the principal ideal of O K generated by 2 + 7. Then
√ √
I = {(a + b 7)(2 + 7) | a, b ∈ Z}
√ √
= {a(2 + 7) + b(7 + 2 7) | a, b ∈ Z}
√ √
= (2 + 7)Z + (7 + 2 7)Z,
√ √
so that {2 + 7, 7 + 2 7} is a basis for I . However, a little more effort yields
a“simpler” basis, that is, one having a rational integer as one of the basis elements.
CB609-06 CB609/Alaca & Williams August 27, 2003 16:53 Char Count= 0
We have
√ √
I = {(a + b 7)(2 + 7) | a, b ∈ Z}
√
= {(2a + 7b) + (a + 2b) 7 | a, b ∈ Z}
√
= {(2(c − 2b) + 7b) + c 7 | b, c ∈ Z}
√
= {3b + c(2 + 7) | b, c ∈ Z}
√
= 3Z + (2 + 7)Z,
√
showing that {3, 2 + 7} is a basis for I .
If {η1 , . . . , ηn } and {λ1 , . . . , λn } are two bases for the same nonzero ideal of the
ring of integers of an algebraic number field then we know by Theorem 6.5.4 that
D(η1 , . . . , ηn ) = D(λ1 , . . . , λn ).
D(I ) = D(η1 , . . . , ηn ).
Theorem 6.5.5 Let K be a quadratic field. Let m be the unique squarefree integer
√
such that K = Q( m).
(a) m
≡ 1 (mod 4). Let a, b, c ∈ Z with a
= 0 and c
= 0. Then
√ √
{a, b + c m} is a basis for the ideal a, b + c m
if and only if
c | a, c | b, ac | b2 − mc2 . (6.5.2)
CB609-06 CB609/Alaca & Williams August 27, 2003 16:53 Char Count= 0
Proof: (a) Suppose first that (6.5.2) holds. Then there are integers x, y, z such that
Hence
bx − ay = 0, by − az = mc.
√
Let α ∈ I = a, b + c m. Then there exist θ ∈ O K and φ ∈ O K such that
√
α = θa + φ(b + c m).
As θ, φ ∈ O K and m
≡ 1 (mod 4), by Theorem 5.4.2 there exist integers r, s, t, u
such that
√ √
θ = r + s m, φ = t + u m.
Hence
√ √ √
α = (r + s m)a + (t + u m)(b + c m)
√
= (ra + tb + umc) + (sa + tc + ub) m
√
= s(bx − ay) + (ra + tb + u(by − az)) + (scx + tc + ucy) m
√
= (r − sy − uz)a + (t + sx + uy)(b + c m),
√
proving that {a, b + c m} is a basis for I .
√ √
Conversely, suppose that {a, b + c m} is a basis for the ideal I = a, b + c m.
√ √ √
As ma ∈ I and m(b + c m) ∈ I there exist integers x, y, u, v such that
√ √
ma = xa + y(b + c m),
√ √ √
m(b + c m) = ua + v(b + c m).
√
Equating coefficients of 1 and m, we obtain
xa + yb = 0,
yc = a,
ua + vb = cm,
vc = b.
CB609-06 CB609/Alaca & Williams August 27, 2003 16:53 Char Count= 0
From the second and fourth equations, we see that c | a and c | b respectively. From
the third and fourth equations, we obtain
uac + b2 = c2 m
so that
ac | b2 − mc2 .
Proof: Suppose that the assertion of the theorem is false. Then there exists a prime
ideal P1 of O K that is not a maximal ideal. Let S be the set of all proper ideals of
O K that strictly contain P1 . As P1 is not a maximal ideal, S is a nonempty set. By
Theorem 6.5.3 O K is a Noetherian domain. Hence, by Theorem 3.1.3, S contains
a maximal element; that is, there is a maximal ideal P2 such that
P1 ⊂ P2 ⊂ O K .
p = P1 ∩ Z ⊆ P2 ∩ Z ⊆ Z.
Now P2 ∩ Z
= Z as 1 ∈
/ P2 , so as p is a maximal ideal of Z (Theorem 1.5.7), we
have
P1 ∩ Z = P2 ∩ Z = p.
α k + ak−1 α k−1 + · · · + a1 α + a0 = 0,
CB609-06 CB609/Alaca & Williams August 27, 2003 16:53 Char Count= 0
and so
α k + ak−1 α k−1 + · · · + a1 α + a0 ∈ P1 .
Let l be the least positive integer for which there exist b0 , . . . , bl−1 ∈ Z such that
Now, as α ∈ P2 , we have
b0 = (αl + · · · + b1 α + b0 ) − (αl + · · · + b1 α) ∈ P2 .
But b0 ∈ Z so
b0 ∈ P2 ∩ Z = P1 ∩ Z
αl + bl−1 αl−1 + · · · + b1 α ∈ P1 .
If l = 1 then α ∈ P1 , contradicting α ∈
/ P1 . Hence l ≥ 2 and
α(αl−1 + · · · + b1 ) ∈ P1 .
αl−1 + · · · + b1 ∈ P1 ,
Exercises
1. Let K be an algebraic number field of degree n. Let θ ∈ K be such that K = Q(θ ).
Let θ1 = θ, θ2 , . . . , θn be the conjugates of θ over Q. Let α ∈ K so there exist unique
rational numbers c0 , c1 , . . . , cn−1 such that
α = c0 + c1 θ + · · · + cn−1 θ n−1 .
For k = 1, 2, . . . , n let
αk = c0 + c1 θk + · · · + cn−1 θkn−1
Exercises 139
21. Let K = Q(θ ), where θ 4 − 4θ 2 + 8 = 0. Find a rational number c such that Q(θ +
cθ 3 )
= K .
22. Prove that the discriminant of the trinomial polynomial x n + ax + b ∈ Z[x], where n
is an integer ≥ 2, is
(−1)(n−1)(n−2)/2 (n − 1)n−1 a n + (−1)n(n−1)/2 n n bn−1 .
23. Prove that the discriminant of the trinomial polynomial x n + ax r + b ∈ Z[x], where n
and r are integers satisfying n > r ≥ 1 and (n, r ) = 1, is
(−1)(n−1)(n−2)/2 (n − r )n−r r r a n br −1 + (−1)n(n−1)/2 n n bn−1 .
Suggested Reading
1. E. T. Bell, Gauss and the early development of algebraic numbers, National Mathematics
Magazine 18 (1944), 188–204, 219–233.
Bell provides a very readable account of the early development of algebraic numbers.
2. R. L. Goodstein, The discriminant of a certain polynomial, Mathematical Gazette 53
(1969), 60–61.
The formula for the discriminant of x n + ax r + b is derived
3. L. Mirsky, An Introduction to Linear Algebra, Oxford University Press, London 1972.
The evaluation of the Vandermonde determinant is carried out on pages 17 and 18.
4. D. W. Masser, The discriminants of special equations, Mathematical Gazette 50 (1966),
158–160.
The formula for the discriminant of x n + ax + b is derived.
Biographies
1. The website
http://www-groups.dcs.st-and.ac.uk/˜history/
7
Integral Bases
In view of this definition the following theorem, which gives an integral basis
for a quadratic field, is just a restatement of Theorem 5.4.2.
Theorem 7.1.1 Let K be a quadratic field. Let m be the unique squarefree integer
√ √
suchthat K√=Q( m). Then {1, m} is an integral basis for K if m ≡ 1 (mod 4)
and 1, 1+2 m is an integral basis for K if m ≡ 1 (mod 4).
If {η1 , . . . , ηn } and {λ1 , . . . , λn } are two integral bases for an algebraic num-
ber field K then Theorem 6.5.4 shows that D(η1 , . . . , ηn ) = D(λ1 , . . . , λn ), and
that if {η1 , . . . , ηn } is an integral basis for K and λ1 , . . . , λn ∈ O K are such that
D(λ1 , . . . , λn ) = D(η1 , . . . , ηn ) then {λ1 , . . . , λn } is also an integral basis for K .
We can therefore make the following definition.
141
CB609-07 CB609/Alaca & Williams August 7, 2003 16:45 Char Count= 0
Theorem 7.1.2 Let K be a quadratic field. Let m be the unique squarefree integer
√
such that K = Q( m). Then the discriminant d(K ) of K is given by
4m, if m ≡ 1 (mod 4),
d(K ) =
m, if m ≡ 1 (mod 4).
We note that the quadratic field K is a real field if and only if d(K ) > 0.
Next we define the norm of an ideal in the ring of integers of an algebraic number
field.
Proof: Let {η1 , . . . , ηn } be a basis for I and let {ω1 , . . . , ωn } be an integral basis
for K . As η1 , . . . , ηn ∈ O K there exist ci j (i, j = 1, . . . , n) ∈ Z such that
n
ηi = ci j ω j , i = 1, . . . , n.
j=1
CB609-07 CB609/Alaca & Williams August 7, 2003 16:45 Char Count= 0
Hence
so that
is a positive integer.
√
Example
√ 7.1.1 Let K = Q( −5).
√ Let I be the ideal of O K generated by 2 and
1 + −5; that is, I = 2, 1 + −5
. We determine √ the norm N (I ) of the ideal I .
First we find a basis for I . As O K = {x + y −5 | x, y ∈ Z} we have
√ √ √
I = {2(a + b −5) + (1 + −5)(c + d −5) | a, b, c, d ∈ Z}
√
= {(2a + c − 5d) + (2b + c + d) −5 | a, b, c, d ∈ Z}
√
= {(2a + (y − 2b − d) − 5d) + y −5 | a, b, d, y ∈ Z}
√
= {2(a − b − 3d) + y + y −5 | a, b, d, y ∈ Z}
√
= {2x + y + y −5 | x, y ∈ Z}
√
= {2x + (1 + −5)y | x, y ∈ Z},
√
so that {2, 1 + −5} is a basis for I . Hence
√ 2
√ 2 1 + √−5 √
D(I ) = D(2, 1 + −5) = = (−4 −5)2 = −80.
2 1 − −5
Hence
D(I ) −80 √
N (I ) = = = 4 = 2.
d(K ) −20
√
Example 7.1.2 Let K = Q( m), where m is a squarefree integer with m ≡
√ √
1 (mod 4), so that O K = {x + y m | x, y ∈ Z}. Let α = a + b m ∈ O K . We
CB609-07 CB609/Alaca & Williams August 7, 2003 16:45 Char Count= 0
we see that
D(α ) = 22 m(a 2 − mb2 )2 .
We now use Theorem 6.5.5 to determine the norms of a wide class of ideals in a
quadratic field.
Theorem 7.1.5 Let K be a quadratic field. Let m be the unique squarefree integer
√
such that K = Q( m).
a = 0, c = 0, c | a, c | b, ac | b2 − mc2 .
Then
√
N (a, b + c m
) = |ac|.
Then
√
b+c m
N (a,
) = |ac|.
2
CB609-07 CB609/Alaca & Williams August 7, 2003 16:45 Char Count= 0
As m ≡ 1 (mod 4) we have
d(K ) = 4m,
d(K ) = m,
N (c
) = |c|n .
CB609-07 CB609/Alaca & Williams August 7, 2003 16:45 Char Count= 0
α ∈ c
⇐⇒ α = cβ for a unique β ∈ O K
⇐⇒ α = c(x1 ω1 + · · · + xn ωn ) for unique x1 , . . . , xn ∈ Z
⇐⇒ α = x1 (cω1 ) + · · · + xn (cωn ) for unique x1 , . . . , xn ∈ Z.
This shows that {cω1 , . . . , cωn } is a basis for the principal ideal c . Hence
so that
D(c
) √ 2n
N (c
) = = c = |c|n .
d(K )
Hence
showing that
D(θ) = m 2 d(K )
for some positive integer m(= |det(ci j )|). The positive integer m is called the index
of θ.
Proof: We have
{1, θ, θ 2 , . . . , θ n−1 } is an integral basis for K
n
D(θ ) = (−1)n(n−1)/2 f (θi ).
i=1
Proof: We have
n
f (x) = irrQ (θ) = (x − θi ).
i=1
n
n
f (x) = (x − θ j )
i=1 j = 1
j = i
so that
n
f (θi ) = (θi − θ j ), i = 1, 2, . . . , n.
j =1
j = i
CB609-07 CB609/Alaca & Williams August 7, 2003 16:45 Char Count= 0
Hence
n
n
n
f (θi ) = (θi − θ j )
i=1 i=1 j = 1
j = i
= (θi − θ j ) (θi − θ j )
1≤i< j≤n 1≤ j<i≤n
= (−1)n(n−1)/2 (θi − θ j ) (θ j − θi )
1≤i< j≤n 1≤ j<i≤n
= (−1)n(n−1)/2 (θi − θ j ) (θi − θ j )
1≤i< j≤n 1≤i< j≤n
= (−1)n(n−1)/2 (θi − θ j )2
1≤i< j≤n
= (−1) n(n−1)/2
D(θ),
by Theorem 6.4.1.
so that
so
as asserted.
In the next example we find an integral basis for a particular cubic field using
Theorems 7.1.8 and 7.1.10.
(a, b) = (−1, −1), (2, 1), (4, 1), (−1, 3), and (5, 3).
Similarly to Theorem 7.1.10 we can use Theorem 7.1.9 to prove the following
result.
CB609-07 CB609/Alaca & Williams August 7, 2003 16:45 Char Count= 0
Proof: Let {ω1 , ω2 , . . . , ωn } be an integral basis for K . Let ωi(1) = ωi , ωi(2) , . . . , ωi(n)
be the K -conjugates of ωi (i = 1, 2, . . . , n). In the expansion of the determinant
(1)
ω1 (1) (1)
(2) ω2(2) · · · ωn
ω ω2 · · · ωn(2)
1
. . ..
.. .. ··· .
ω(n) ω(n) · · · ω(n)
1 2 n
there are n! terms, half of which occur with positive signs and half with negative
signs. Let the sum of those with positive signs be λ and those with negative signs
CB609-07 CB609/Alaca & Williams August 7, 2003 16:45 Char Count= 0
µ so that
( j)
det(ωi ) = λ − µ.
Set
A = λ + µ, B = λµ.
Then
( j)
d(K ) = (det(ωi ))2 = (λ − µ)2 = (λ + µ)2 − 4λµ = A2 − 4B.
( j)
As ωi ∈ O K (i = 1, 2, . . . , n), by Theorem 6.3.3 each ωi (i, j = 1, 2, . . . , n) ∈ .
Hence λ, µ ∈ so
A ∈ , B ∈ .
A ∈ Q.
Hence A ∈ ∩ Q = Z. Then
A2 − d(K )
B= ∈ Q,
4
so that B ∈ ∩ Q = Z. Finally, as A, B ∈ Z, we have
The next example illustrates the use of Theorem 7.1.14 to determine the discrim-
inant of an algebraic number field K = Q(θ) when D(θ ) is not squarefree.
In the next example we give a cubic field Q(θ ) for which {1, θ, θ 2 } is not an
integral basis.
The first of these is ≡ 0 (mod 4) and the second is ≡ 1 (mod 4), so we cannot
use Theorem 7.1.14 to distinguish between them. We recall from Example 6.3.2 that
(θ + θ 2 )/2 is an integer of K as it is a root of the polynomial x 3 + 11x 2 + 36x + 4 ∈
Z[x]. Hence {1, θ, θ 2 } is not an integral basis for K . Thus
In the next example Theorem 7.1.14 is not sufficient to distinguish between the
possible values of the discriminant and we have to carry out a more detailed analysis.
√ √
Example 7.1.6 Let θ = 3 2 and set K = Q(θ) = Q( 3 2). Since θ is a root of the
irreducible polynomial x 3 − 2 ∈ Z[x], we have irrQ (θ) = x 3 − 2 and [K : Q] =
deg(irrQ (θ )) = 3. By Theorem 7.1.10 we have
α = x1 + x2 θ + x3 θ 2 .
As α ∈ O K , by Theorem6.3.3
√
we have α, α , α ∈ ; as θ ∈ O K we have θ, θ 2 ∈ ;
and as ω, ω2 ∈ Z + Z 1+ 2 −3 = OQ(√−3) we have ω, ω2 ∈ . Thus
α + α + α , θ 2 (α + ω2 α + ωα ), θ(α + ωα + ω2 α ) ∈
so that
Set
yi = 6xi ∈ Z, i = 1, 2, 3,
so that
6α = y1 + y2 θ + y3 θ 2 . (7.1.1)
θ | n in O K =⇒ 2 | n in Z.
yi = 2z i , i = 1, 2, 3,
so that
3α = z 1 + z 2 θ + z 3 θ 2 , z 1 , z 2 , z 3 ∈ Z.
x 3 + c1 x 2 + c2 x + c3 ∈ Q[x],
where
c1 = −(α + α + α ) = −z 1 ,
1
c2 = αα + α α + α α = (z 12 − 2z 2 z 3 ),
3
−1 3
c3 = −αα α = (z + 2z 23 + 4z 33 − 6z 1 z 2 z 3 ).
27 1
Hence
irrQ (α) = x 3 + c1 x 2 + c2 x + c3 .
CB609-07 CB609/Alaca & Williams August 7, 2003 16:45 Char Count= 0
Suppose that at least one of z 1 , z 2 , z 3 is not divisible by 3. Then (7.1.2) and (7.1.3)
show that 3 does not divide any of z 1 , z 2 , z 3 . From (7.1.3) we have
z 13 + 2z 23 + 4z 33 ≡ 0 (mod 3).
z 1 + 2z 2 + z 3 ≡ 0 (mod 3).
so that
z 2 = 2z 1 + 3t, z 3 = z 1 + 3u.
Then
z 13 + 2z 23 + 4z 33 − 6z 1 z 2 z 3
= z 13 + 2(2z 1 + 3t)3 + 4(z 1 + 3u)3 − 6z 1 (2z 1 + 3t)(z 1 + 3u)
= 9z 13 + 54(t z 12 + 2t 2 z 1 + t 3 + 2u 2 z 1 + 2u 3 − tuz 1 )
≡ 9z 13 (mod 27)
≡ 0 (mod 27),
z i = 3wi , i = 1, 2, 3.
Then
α = w1 + w2 θ + w3 θ 2 ∈ Z + Zθ + Zθ 2 ,
proving O K ⊆ Z + Zθ + Zθ 2 as required.
CB609-07 CB609/Alaca & Williams August 7, 2003 16:45 Char Count= 0
1 ω2 3 2 ω( 3 2)2
Richard Dedekind (1831–1916) [5] determined an integral basis for the cubic
√
field Q( 3 m) (with m a cubefree integer) in 1900 (see Theorem 7.3.2).
We conclude
√ this
√ section with the determination of an integral basis for the quartic
field Q( −1 + 2).
√ √ √ √
Example 7.1.7 Let K be the quartic field Q( −1 + 2) = Q( −1, 2). We
show that
√ √ 1 √ √
O K = Z + Z −1 + Z 2 + Z ( 2 + −2)
2
and
d(K ) = 256.
c = b02 + 2b32 ∈ Z,
d = b02 + b12 − 2b22 − 2b32 ∈ Z,
e = b0 b3 − b1 b2 ∈ Z,
so that θ is a root of
d b0 d d 2 + 8e2
f (x) = x − 2b0 x + c +
4 3
x + −2b3 e −
2
x+ ∈ Q[x].
2 2 16
As θ is of degree 4 over Q (since it does not belong to any of the subfields of K ) the
polynomial f (x) must be the minimal polynomial of θ over Q. Hence, as θ ∈ O K ,
we have f (x) ∈ Z[x], and so d/2 ∈ Z and (d 2 + 8e2 )/16 ∈ Z. Hence
Hence
and
If b0 or b1 is odd from (7.1.4) we see that the other is odd as well. Then
from (7.1.5) we deduce that b2 ≡ b3 (mod 2), and (7.1.4) gives the contradiction
2 ≡ 0 (mod 4). Thus
Then
1 √ √ √
θ = (b0 + b1 −1 + b2 2 + b3 −2)
2
√ √ 1 √ √
= c0 + c1 −1 + c2 2 + c3 ( 2 + −2)
2
√ √ 1 √ √
∈ Z + Z −1 + Z 2 + Z ( 2 + −2)
2
as required.√ √ √ √
Thus {1, −1, 2, 12 ( 2 + −2)} is an integral basis for K and
2
√ √ 1 √ √
1 −1 2 ( 2 + −2)
2
√ √ 1 √ √
1 − −1 2 ( 2 − −2)
d(K ) = 2
√ √
= 256.
1 √ √ 1
−1 − 2 (− 2 − −2)
2
√ √ √ √
1
1 − −1 − 2 (− 2 + −2)
2
√ √ √ √
An integral basis for the quartic field Q( m + n) = Q( m, n), where m
and n are distinct squarefree integers, was determined by K. S. Williams [19] in
1970.
Clearly every quadratic field is monogenic. The cubic fields in Examples 7.1.4
and 7.1.6 are monogenic. Dedekind showed in 1878 that not every algebraic number
field is monogenic by proving that the cubic field
K = Q(θ), θ 3 − θ 2 − 2θ − 8 = 0,
so that
1 √ √
θ 2 = i, θ 3 = (− 2 + i 2), θ 4 = −1.
2
Then, by Example 7.1.7, we have
√ 1 √ √
O K = Z + Zi + Z 2 + Z ( 2 + i 2)
2
= Z + Zθ 2 + Z(θ − θ 3 ) + Zθ
= Z + Zθ + Zθ 2 + Zθ 3 ,
so that K is monogenic with power basis {1, θ, θ 2 , θ 3 }.
We conclude this section with a simple upper bound for the absolute value of the
discriminant of an algebraic number field as well as a theorem giving the sign of
the discriminant.
Proof: Let {η1 , . . . , ηn } be an integral basis for K . Then there exist ci j (i, j =
1, 2, . . . , n) ∈ Z such that
λ1 = c11 η1 + · · · + c1n ηn ,
λ2 = c21 η1 + · · · + c2n ηn ,
···
λn = cn1 η1 + · · · + cnn ηn .
Hence
D(λ1 , . . . , λn ) = (det ci j )2 D(η1 , . . . , ηn ) = (det ci j )2 d(K ).
As D(λ1 , . . . , λn ) = 0 we see that det (ci j ) = 0. Thus, as det (ci j ) ∈ Z, we have
(det ci j )2 ≥ 1 and so
|D(λ1 , . . . , λn )| ≥ |d(K )|.
If D(λ1 , . . . , λn ) is squarefree then from
D(λ1 , . . . , λn ) = (det ci j )2 d(K ),
we deduce that det ci j = ±1. Hence D(λ1 , . . . , λn ) = d(K ), proving that
{λ1 , . . . , λn } is an integral basis for K .
ak = 0, ak+1 = · · · = an−1 = 0
so that
α = a0 + a1 θ + · · · + ak θ k
D(ω1 , . . . , ωn ) = 0.
CB609-07 CB609/Alaca & Williams August 7, 2003 16:45 Char Count= 0
Then for each α ∈ O K there exist unique rational integers x1 , . . . , xn such that
n
xj
α= ωj
j=1
D(ω1 , . . . , ωn )
and
D(ω1 , . . . , ωn ) | x 2j , j = 1, 2, . . . , n.
s 2j | D(ω1 , . . . , ωn ), j = 1, 2, . . . , n.
CB609-07 CB609/Alaca & Williams August 7, 2003 16:45 Char Count= 0
Let
rj
x j = y j D(ω1 , . . . , ωn ) = D(ω1 , . . . , ωn ) ∈ Z, j = 1, 2, . . . , n.
sj
Then, from (7.2.2), we obtain
n
xj
α= ωj.
j=1
D(ω1 , . . . , ωn )
proving that {ω1 , . . . , ωn } is an integral basis for K , a result that we have seen
before in Theorem 7.1.16.
We now use Theorem 7.2.1 to bound the denominators of the a j in (7.2.1).
and
1 ≤ s j ≤ |D(θ )|, s 2j | D(θ).
Proof: As θ ∈ O K we have
1, θ, θ 2 , . . . , θ n−1 ∈ O K ,
CB609-07 CB609/Alaca & Williams August 7, 2003 16:45 Char Count= 0
so by Theorem 6.4.4(b)
D(θ) = 0.
Then by Theorem 7.2.1 there exist unique rational integers x1 , . . . , xn such that
n
x j j−1
α= θ
j=1
D(θ)
and
D(θ) | x 2j , j = 1, 2, . . . , n.
and
1 ≤ s j ≤ |D(θ )|.
s 2j | D(θ).
Sk = {ak ∈ Q | a0 + a1 θ + · · · + ak θ k ∈ O K
for some a0 , a1 , . . . , ak−1 ∈ Q}. (7.2.4)
Clearly
S0 = Z
CB609-07 CB609/Alaca & Williams August 7, 2003 16:45 Char Count= 0
and
Sk ⊇ Z, k = 1, 2, . . . , n − 1.
Hence
r r
a = qak∗ + , 0≤ < ak∗ .
m m
As a ∈ Sk there exist b0 , b1 , . . . , bk−1 ∈ Q such that
b0 + b1 θ + · · · + bk−1 θ k−1 + aθ k ∈ O K .
Then
r k
(b0 − qc0 ) + (b1 − qc1 )θ + · · · + (bk−1 − qck−1 )θ k−1 + θ ∈ OK ,
m
so that
r
∈ Sk .
m
If 0 < r/m < ak∗ this contradicts the minimality of ak∗ . Hence r/m = 0 and a =
qak∗ , proving Sk = ak∗ Z.
The next theorem shows that each dk−1 (k = 1, 2, . . . , n − 1) divides its succes-
sor dk .
so that
∗
ak−1 = mak∗
dk = mdk−1 ,
proving
dk−1 | dk , k = 1, 2, . . . , n − 1.
a0 + a1 θ + · · · + ak−1 θ k−1 + θ k
α= .
dk
dl dl r j − al s j−1
l−1
r0 + θj
dl−1 j=1
dl−1
dl dl r j − al s j−1
l−1 cj l−1
r0 + θj = θ j.
dl−1 j=1
dl−1 j=0
dl−1
a0 = c0 , a j = al s j−1 + c j , j = 1, 2, . . . , l − 1.
Then
aj
rj = , j = 0, 1, . . . , l − 1,
dl
and
a0 + a1 θ + · · · + al−1 θ l−1 + al θ l
α= .
dl
This completes the inductive step and the theorem follows by the principle of
mathematical induction and (for the second part) Theorem 7.2.4.
Proof: In any integral basis for K = Q(θ) at least one of the basis elements must
be of the form a0 + a1 θ + · · · + an−1 θ n−1 (a0 , a1 , . . . , an−1 ∈ Q) with an−1 = 0;
otherwise the integral basis could not represent θ n−1 . Replacing the basis element
by its negative, if necessary, we may suppose that an−1 > 0. We choose an integral
basis {ω1 , . . . , ωn } for K with
ωk = ωk − mωn−1 .
Then {ω1 , . . . , ωk−1 , ωk , ωk+1 , . . . , ωn } is an integral basis for K . This contradicts
the minimality of an−1 as the coefficient of θ n−1 in ωk is bn−1 − man−1 , which
is positive and strictly less than an−1 . Hence bn−1 − man−1 = 0, so that bn−1 is a
rational integral multiple of an−1 . Thus there exist rational integers m 1 , . . . , m n−2
such that ω1 = ω1 − m 1 ωn , ω2 = ω1 − m 2 ωn , . . . , ωn−1
= ωn−1 − m n−1 ωn are
integers of degrees at most n − 2 in θ. Moreover, {ω1 , . . . , ωn−1 , ωn } is an integral
basis for K . Among all integral bases {ω1 , . . . , ωn } for which ω1 , . . . , ωn−1 are
integers of degree at most n − 2 in θ, we choose one for which the coefficient of
θ n−2 is positive and minimal, and we continue our construction until we arrive at
an integral basis α0 , α1 , α2 , . . . , αn−1 , where each αi is of degree i in θ. Let
i
αi = aik θ k , ai k ∈ Q.
k=0
Then
and thus
a00 a11 an−1 n−1
∗ · ∗ ··· ∗ ≤ 1.
a0 a1 an−1
As aii ∈ Si (i = 0, 1, . . . , n − 1), by Theorem 7.2.3 each aii /ai∗ (i =
0, 1, . . . , n − 1) is a positive integer and so
aii = ai∗ , i = 0, 1, . . . , n − 1.
Thus each αi (i = 0, 1, . . . , n − 1) is a minimal integer of degree i in θ .
Theorem 7.2.7 gives a method of finding an integral basis for an algebraic number
field of degree n. We have only to find a minimal integer of each degree up to n − 1.
This is illustrated for some cubic fields in the next section.
Our final theorem of this section gives some further useful information about the
denominators of minimal integers.
Further
D(θ)
= d(K ) ∈ Z
(d0 · · · dn−1 )2
so that
(d0 · · · dn−1 )2 | D(θ).
di | di+1 | · · · | dn−1
so that
di2(n−i) | D(θ).
Then
2 2
θ 2 + θ + θ = (θ + θ + θ )2 − 2(θθ + θ θ + θ θ) = −2a,
2 2 2 2
θ 2 θ + θ θ + θ θ 2 = (θθ + θ θ + θ θ)2 − 2θθ θ (θ + θ + θ ) = a 2 ,
CB609-07 CB609/Alaca & Williams August 7, 2003 16:45 Char Count= 0
2 2 2 2
αα + α α + α α = 3y02 + y12 (θ θ + θ θ + θ θ) + y22 (θ 2 θ + θ θ + θ θ 2 )
2 2
+ 2y0 y1 (θ + θ + θ ) + 2y0 y2 (θ 2 + θ + θ )
2 2 2 2 2
+ y1 y2 (θ θ + θ 2 θ + θθ + θ 2 θ + θ θ + θ θ )
= 3y02 + ay12 + a 2 y22 − 4ay0 y2 + 3by1 y2
and
αα α = y03 + y13 θ θ θ + y23 (θ θ θ )2 + y0 y12 (θθ + θ θ + θ θ)
2 2 2 2
+ y0 y22 (θ 2 θ + θ θ + θ θ 2 ) + y02 y1 (θ + θ + θ )
2 2
+ y02 y2 (θ 2 + θ + θ ) + y12 y2 θθ θ (θ + θ + θ )
+ y1 y22 θ θ θ (θθ + θ θ + θ θ)
2 2 2 2
+ y0 y1 y2 (θ θ + θ 2 θ + θθ + θ 2 θ + θ θ + θ θ )
= y03 − by13 + b2 y23 + ay0 y12 + a 2 y0 y22 − 2ay02 y2 − aby1 y22 + 3by0 y1 y2 .
The result now follows as α = y0 + y1 θ + y2 θ 2 is a root of
(x − α)(x − α )(x − α ) = x 3 − (α + α + α )x 2
+ (αα + αα + α α )x − αα α .
so that α 3 − 2α 2 + α − 3 = 0.
1
A = 2 · 11 · = 11,
2
1 1 1 1
B = 11 · 2
+ 112 · 2 + 3 · 4 · · = 36,
2 2 2 2
1 1 1
C = 4 · 3 − 4 · 3 + 4 · 11 · 3 = 4,
2
2 2 2
so that α is a root of the cubic equation x 3 + 11x 2 + 36x + 4 = 0 and thus an
integer of Q(θ) (see Example 6.3.2).
Example 7.3.4 Let θ be a root of the cubic equation x 3 − 21x − 236 = 0. Here
a = −21, b = −236. We consider α = (1 + θ)/3, so that y0 = y1 = 1/3, y2 = 0.
By Theorem 7.3.1 we obtain
1
A = −3 = −1,
3
1 1
B = 3 · 2 − 21 · 2 = −2,
3 3
1 1 1 −216
C = − 3 − 236 3 + 21 3 = = −8,
3 3 3 27
so that α is a root of the equation x 3 − x 2 − 2x − 8 = 0. Hence (1 + θ)/3 is an
integer of the cubic field Q(θ).
CB609-07 CB609/Alaca & Williams August 7, 2003 16:45 Char Count= 0
1 −36
A= (6 − 42) = = −2,
18 18
1 972
B = 2 (12 − 21 + 441 − 168 + 708) = = 3,
18 324
1
C = 3 (8 + 236 − 55696 − 42 + 882 − 168 − 4956 + 1416)
18
−58320
= = −10,
5832
so that α is a root of the cubic equation x 3 − 2x 2 + 3x − 10 = 0. Hence (−2 −
θ + θ 2 )/18 is an integer of the cubic field Q(θ).
In the next four examples we use Theorems 7.2.7 and 7.2.8 to give integral bases
for the following cubic fields:
Definition 7.3.1 (Pure cubic field) A cubic field K is said to be pure if there exists
√
a rational integer m, which is not a perfect cube, such that K = Q( 3 m).
√
In Example 7.1.6 we found an integral basis for the pure cubic field Q( 3 2).
CB609-07 CB609/Alaca & Williams August 7, 2003 16:45 Char Count= 0
As
4 + θ + θ 2 2/3 4 − 2θ + θ 2
21/3 = , 2 = ,
3 3
and {1, 21/3 , 22/3 } is an integral basis for K = Q(21/3 ), we see that
4 + θ + θ 2 4 − 2θ + θ 2
1, ,
3 3
is an integral basis for K = Q(θ). Since
4 + θ + θ 2 4 − 2θ + θ 2
− =θ
3 3
and
4 + θ + θ2 1 + θ + θ2
−1= ,
3 3
a simpler integral basis is
1 + θ + θ2
1, θ, .
3
√
We now give an integral basis for the pure cubic field Q( 3 m). As we have already
mentioned this basis was first given by Dedekind [5] in 1900.
We leave the proof of Theorem 7.3.2 as an exercise (Exercise 6). From Theorem
7.3.2 we obtain Table 1.
If K is a pure cubic field given in the form K = Q(θ), θ 3 + aθ + b = 0, a, b ∈
Z, it is known that −4a 3 − 27b2 = −3c2 for some positive integer c (in Exam-
ple 7.3.10 we have a = 6, b = 2, c = 18), and an integral basis for K has been
given by Spearman and Williams [15].
CB609-07 CB609/Alaca & Williams August 7, 2003 16:45 Char Count= 0
√
Table 2. Integral bases and discriminants for Q( 4 k),
x 4 − k irreducible in Q[x], 2 ≤ k ≤ 10
√
k Integral basis (θ = 4
k) Discriminant
1 {1, θ, θ 2 , θ 3 } 256 = 28
2 {1, θ, θ 2 , θ 3 } 2048 = 211
3 {1, θ, (1 + θ 2 )/2, (θ + θ 3 )/2} 432 = 24 · 33
5 {1, θ, θ 2 , θ 3 } 32000 = 28 · 53
6 {1, θ, θ 2 , θ 3 } 55296 = 211 · 33
7 {1, θ, (1 + θ 2 )/2, (1 + θ + θ 2 + θ 3 )/4} 1372 = 22 · 73
9 {1, θ, θ 2 /3, θ 3 /3} 2304 = 28 · 32
10 {1, θ, θ 2 , θ 3 } 256000 = 211 · 53
√ √ √
Note: Q( 4 −4) = Q(i) and Q( 4 −8) = Q( 4 −2).
Clearly
i(K ) | m(K ). (7.4.1)
Theorem 7.4.1 Let K be an algebraic number field. Then m(K ) = 1 if and only if
K possesses a power basis.
Theorem 7.4.2 Let K be an algebraic number field such that K possesses a power
basis. Then i(K ) = 1.
In Example 7.4.4 we give an algebraic number field K for which i(K ) = 1 but
K does not possess a power basis. This shows that the converse of Theorem 7.4.2
is not true. Theorem 7.4.2 gives a convenient way of establishing that an algebraic
number field does not have a power basis; all we have to do is to show that i(K ) ≥ 2.
In the next theorem we determine the index and minimal index of a quadratic
field directly from their definitions.
so that
and
so that
so that
and
Of course we could have argued that a quadratic field clearly has a power basis
so that by Theorem 7.4.1 m(K ) = 1 and then by (7.4.1) i(K ) = 1.
In the next four examples we determine i(K ) and m(K ) for some cubic fields K .
Example 7.4.1 We determine the index i(K ) and the minimal index m(K ) of the
cubic field K = Q(θ), where θ is a root of f (x) = x 3 − 3x + 9. Let θ and θ
be the other two roots of f (x), so that x 3 −
3x + 9 = (x − θ )(x − θ )(x − θ ).
By Example 7.3.6 we know that 1, θ, θ 2 /3 is an integral basis for K , D(θ) =
−33 · 7 · 11, and d(K ) = −3 · 7 · 11. Let α ∈ O K . Then α = a + bθ + cθ 2 /3 for
CB609-07 CB609/Alaca & Williams August 7, 2003 16:45 Char Count= 0
θ
α = a + bθ + c ,
3
2
θ
α = a + bθ + c .
3
Hence, as θ + θ + θ = 0, we have
c c
α − α = (θ − θ ) b + (θ + θ ) = (θ − θ ) b − θ ,
3 3c
c
α − α = (θ − θ ) b + (θ + θ ) = (θ − θ ) b − θ ,
3 3c
c
α − α = (θ − θ ) b + (θ + θ ) = (θ − θ ) b − θ .
3 3
Thus, by Theorem 6.4.1,
D(α) = (α − α )2 (α − α )2 (α − α )2
c c c
= (θ − θ )2 (θ − θ )2 (θ − θ )2 (b − θ)2 (b − θ )2 (b − θ )2
3 3 3
2
c 3 3b
= D(θ) f
3 c
2
bc2 c3
= −3 · 7 · 11 b −
3 3
+
3 3
= −3 · 7 · 11 (3b − bc + c ) .
3 2 3 2
Then
D(α) −3 · 7 · 11(3b3 − bc2 + c3 )2
ind α = = = |3b3 − bc2 + c3 |.
d(K ) −3 · 7 · 11
Hence
as
so
2
θ2 θ2
1, ,
3 3
Example 7.4.2 We show that m(K ) = i(K ) = 1 for the cubic field K = Q(θ ),
where θ is a root of f (x) = x 3 − x + 4.
Let θ and θ be the other two roots of f (x) so that
−107. Let α ∈ O K . Then α = a + bθ + c θ +2 θ for some a, b, c ∈ Z. Exactly as
2
Then
D(α) −107(2b3 + 3b2 c + bc2 + c3 )2
ind α = = = |2b3 + 3b2 c + bc2 + c3 |.
d(K ) −107
Hence
as
so
2
θ + θ2 θ + θ2
1, ,
2 2
Let θ and θ be the other two roots of f (x) = x 3 − 21x − 236 so that
and
2
(1 + θ ) −2 − θ + θ
α =a+b +c .
3 18
and we obtain
b c c
α − α = (θ − θ ) − − θ ,
3 18 18
and similarly
b c c b c c
α − α = (θ − θ ) − − θ , α − α = (θ − θ ) − − θ .
3 18 18 3 18 18
CB609-07 CB609/Alaca & Williams August 7, 2003 16:45 Char Count= 0
Dedekind [4] gave in 1878 the first example of an algebraic number field without
a power basis, namely, the cubic field L given by
L = Q(φ), φ 3 − φ 2 − 2φ − 8 = 0.
The field L is in fact the same field as the field K = Q(θ), θ 3 − 21θ − 236 = 0,
in Example 7.4.3, as θ and φ are related by θ = 3φ − 1.
CB609-07 CB609/Alaca & Williams August 7, 2003 16:45 Char Count= 0
which shows that the converse of Theorem 7.4.2 does not hold.
√
Example 7.4.4 Let K = Q( 3 175). An integral basis for K is given by
{1, 1751/3 , 2451/3 } and d(K ) = −33 · 52 · 72 (see Theorem 7.3.2). Let α ∈ O K .
Then there exist a, b, c ∈ Z such that
α = a + b1751/3 + c2451/3 .
1 + ω + ω2 = 0, ω3 = 1,
we obtain
Hence
D(α) = (α − α )2 (α − α )2 (α − α )2
= {(1 − ω)(1 − ω2 )(ω − ω2 )}2 (175b3 − 245c3 )2
= −27(175b3 − 245c3 )2 = −33 · 52 · 72 (5b3 − 7c3 )2 .
Then
D(α) −33 · 52 · 72 (5b3 − 7c3 )2
ind α = = = |5b3 − 7c3 |.
d(K ) −33 · 52 · 72
Thus
and
Llorente and Nart [12, Theorem 4, p. 585] have given a necessary and sufficient
condition for a cubic field to have index 2.
[K m : Q] = φ(m).
CB609-07 CB609/Alaca & Williams August 7, 2003 16:45 Char Count= 0
The smallest field containing both K m and K n is K [m,n] , where [m, n] denotes the
least common multiple of m and n. Also, K m ∩ K n = K (m,n) . If m ≡ 2 (mod 4) then
K m ⊆ K n holds if and only if m | n. Thus if m and n are distinct and not congruent
to 2 (mod 4) the cyclotomic fields K m and K n are distinct.
The next theorem gives an integral basis for K m as well as a formula for the
discriminant d(K m ).
We refer the reader to Narkiewicz [13, Theorem 4.10, p. 169] for a proof of this
theorem.
Taking m = 3, 4, 5, 8 in Theorem 7.5.1, we obtain d(K 3 ) = −3, d(K 4 ) =
−4, d(K√ 5 ) = 125, d(K 8 ) =√256. The first two of these are familar to us as
K 3 = Q( −3) and √ K4 = √ Q( −1) are quadratic fields. The fourth equality is
√also
known to us as 2 ( 2 + −2) is a primitive eighth root of unity, so K 8 = Q( 2 +
1
√ √ √ √ √
−2) = Q( 2, −2) = Q( 2, i) and we showed that d(Q( 2, i)) = 256 in
Example 7.1.7.
As s = 0 and c2 + s 2 = 1 we obtain
5c4 − 10c2 (1 − c2 ) + (1 − c2 )2 = 0,
that is,
16c4 − 12c2 + 1 = 0,
so that
√
3± 5
c = . 2
8
√
Now c ≈ 0.3, c2 ≈ 0.09, (3 − 5)/8 ≈ 0.09, so
√
3− 5
c =
2
8
and
√ √ √
3− 5 6−2 5 5−1
c= = = .
8 16 4
Hence
√ 2 √
5 − 1 10 + 2 5
s =1−c =1−
2 2
= ,
4 16
so
√
10 + 2 5
s= .
4
We have shown that
2πi 1 √ √
β=e 5 = 5 − 1 + i 10 + 2 5 .
4
Squaring we obtain
4πi 1 √ √
β =e
2 5 = − 5−1+i 10 − 2 5 ,
4
as
√
√ 5−1 √
10 − 2 5 = 10 + 2 5.
2
Further,
12 √ √
β = β̄ =
3
− 5 − 1 − i 10 − 2 5 ,
4
1 √ √
β 4 = β̄ = 5 − 1 − i 10 + 2 5 .
4
CB609-07 CB609/Alaca & Williams August 7, 2003 16:45 Char Count= 0
Exercises 189
Hence
1 √
β − β4 = i 10 + 2 5,
2
so
√
Q i 10 + 2 5 = Q(2β − 2β 4 ) ⊆ Q(β).
Also,
1 √ √ 2
β= −12 + 2 i 10 + 2 5 − i 10 + 2 5 ,
8
so
√
Q(β) ⊆ Q i 10 + 2 5 .
and
√
d Q i 10 + 2 5 = d(K 5 ) = 125.
√
Integral bases for quartic fields like Q(i 10 + 2 5), which contain a quadratic
subfield, have been given by Huard, Spearman, and Williams [10].
The final theorem of this chapter is immediate from Definition 7.1.5 and Theorem
7.5.1.
Theorem 7.5.2 The cyclotomic field K m = Q(ζm ) is monogenic for every positive
integer m.
Exercises
1. Let D denote the discriminant of
Prove that
D ≡ 0 or 1 (mod 4).
√ √
2. Using
√ the method of Example 7.1.6, prove
√ that {1, 3 3, ( 3 3)2 } is an integral basis for
Q( 3 3). What is the discriminant of Q( 3 3)?
CB609-07 CB609/Alaca & Williams August 7, 2003 16:45 Char Count= 0
3. Prove that
√ √ √ √ √
1 + 5 1 + 13 1 + 5 + 13 + 65
1, , ,
2 2 4
√ √
√ for K = Q( 5, 13). What is d(K )?
is an integral√basis
4. Let K = Q( 5, 13). Use Exercise 3 to prove that
1 √ √ √
O K = { (x + y 5 + z 13 + w 65) | x, y, z, w ∈ Z,
4
x ≡ y ≡ z ≡ w (mod 2), x − y − z + w ≡ 0 (mod 4)}.
−22 · 29.
9. Let K 1 = Q(θ1 ), where θ13 + 27θ1 + 240 = 0, and K 2 = Q(θ2 ), where θ23 + 27θ2 +
72 = 0. Prove that d(K 1 ) = d(K 2 ) = −35 . Is K 1 = K 2 ?
10. Let K = Q(θ),√ where θ 4 − 17θ 2 − 34θ − 17 = 0. Prove that d(K ) = 173 .
11. Let K = Q( 2). Prove that d(K ) = −211 .
4
12. Prove from first principles that K = Q(θ ), θ 3 + 30θ + 90 = 0, θ ∈ R, is a pure cubic
field, and express K in the form K = Q(m 1/3 ) for some cubefree integer m.
13. Let K = Q(θ ), where θ 3 − 4θ + 2 = 0. Prove that {1, θ, θ 2 } is an integral basis for K
and that d(K ) = 22 ·37. √
14. Prove that K 5 = Q(i 5 + 2 5).
15. If p is an odd prime prove that
p−1
[Q(e2πi/ p + e−2πi/ p ) : Q] = .
2
16. Suppose that x 3 + ax + b ∈ Z[x] is irreducible. Prove that K = Q(θ ), θ 3 + aθ + b =
0, θ ∈ R, is a pure cubic field if and only if −4a 3 − 27b2 = −3c2 for some positive
integer c.
17. Let K be an algebraic number field. Let L be a conjugate field of K . Prove that
d(K ) = d(L).
18. Let K be an algebraic number field. Let σ be a monomorphism : K −→ C. Let L be the
conjugate field σ (K ). Let {ω1 , . . . , ωn } be an integral basis for K . Is {σ (ω1 ), . . . , σ (ωn )}
an integral basis for L?
19. Determine an integral basis for
K = Q(θ), θ 3 + 30θ + 15 = 0.
20. If K and L are algebraic number fields with K ⊆ L prove that d(K ) | d(L).
CB609-07 CB609/Alaca & Williams August 7, 2003 16:45 Char Count= 0
21. Let K be an algebraic number field of degree n over Q. Let θ ∈ O K be such that
K = Q(θ). Let α ∈ O K . Express α in the form
n−1
α= yjθ j,
j=0
y j D(θ ) ∈ Z, j = 0, 1, . . . , n − 1.
28. Let m be a positive integer. Let ζm be a primitive mth root of unity. What is sgn(d(Q(ζm +
ζm−1 )))?
Suggested Reading
1. Ş. Alaca, p-integral bases of a cubic field, Proceedings of the American Mathematical
Society 126 (1998), 1949–1953.
A p-integral basis of a cubic field K is determined for each rational prime p, and then an integral
basis of K and the discriminant d(K ) of K are obtained from its p-integral bases.
2. Ş. Alaca, p-integral bases of algebraic number fields, Utilitas Mathematica 56 (1999),
97–106.
The properties of p-integral bases of an algebraic number field K are developed and used to
show how an integral basis of K can be obtained from its p-integral bases.
3. A. Brill, Ueber die Discriminante, Mathematische Annalen 12 (1877), 87–89.
This is the original paper of Brill giving the sign of the discriminant of an algebraic number field.
CB609-07 CB609/Alaca & Williams August 7, 2003 16:45 Char Count= 0
4. R. Dedekind, Über den Zusammenhang zwischen der Theorie der Ideale und der Theo-
rie der höheren Kongruenzen, Abh. Kgl. Ges. Wiss. Göttingen 23 (1878), 1–23. (Gesam-
melte Mathematische Werke I, pp. 202–232, Vieweg, Wiesbaden, 1930.)
It is shown that the cubic field K = Q(θ), θ 3 − θ 2 − 2θ − 8 = 0, does not have a power
basis.
5. R. Dedekind, Über die Anzahl der Idealklassen in reinen kubischen Zahlkörpern, Jour-
nal für die reine und angewandte Mathematik 121 (1900), 40–123. (Gesammelte Math-
ematische Werke II, pp. 148–233, Vieweg, Wiesbaden, 1931.)
√
An integral basis is given for the pure cubic field Q( 3 m).
6. D. S. Dummit and H. Kisilevsky, Indices in cyclic cubic fields, in Zassenhaus, H. (ed.),
Number Theory and Algebra, Collected Papers Dedicated to Henry B. Mann, Arnold
E. Ross and Olga Taussky-Todd, pp. 29–42, Academic Press, New York, 1979.
It is shown that infinitely many cyclic cubic fields have a power basis.
7. H. T. Engstrom, On the common index divisors of an algebraic field, Transactions of
the American Mathematical Society 32 (1930), 223–237.
The basic properties of the index of an algebraic number field are given.
8. T. Funakura, On integral bases of pure quartic fields, Mathematical Journal of Okayama
University 26 (1984), 27–41.
An explicit integral basis is given for a pure quartic field.
9. M.-N. Gras, Sur les corps cubiques cycliques dont l’anneau des entiers est monogène,
Annales Scientifiques de l’Université de Besançon, Mathematics, 1973, 26 pp.
Necessary and sufficient conditions are given for a cyclic cubic field to have a power basis.
It can be deduced from these that infinitely many cyclic cubic fields do not have a power
basis.
10. J. G. Huard, B. K. Spearman, and K. S. Williams, Integral bases for quartic fields with
quadratic subfields, Journal of Number Theory 51 (1995), 87–102.
√
Let L be√ quartic
field √ with quadratic √subfield Q( c), where √
c is a squarefree integer. Then
L = Q( c, a + b c), where a + b c is not a square in Q( c). The discriminant of L and
an integral basis for L are determined explicitly.
11. R. H. Hudson and K. S. Williams, The integers of a cyclic quartic field, Rocky Mountain
Journal of Mathematics 20 (1990), 145–150.
An explicit integral basis is given for a quartic field with Galois group Z4 .
12. P. Llorente and E. Nart, Effective determination of the rational primes in a cubic field,
Proceedings of the American Mathematical Society 87 (1983), 579–585.
A necessary and sufficient condition is given for a cubic field to have index 2.
13. W. Narkiewicz, Elementary and Analytic Theory of Algebraic Numbers, Springer-
Verlag, Berlin, 1990.
The principal properties of cyclotomic fields are summarized in Theorem 4.10, p. 169.
14. B. K. Spearman and K. S. Williams, The conductor of a cyclic quartic field, Publica-
tiones Mathematicae 48 (1996), 13–43.
An explicit formula is given for the discriminant of a cyclic quartic field Q(θ), where θ 4 + Aθ 2 +
Bθ + C = 0.
15. B. K. Spearman and K. S. Williams, An explicit integral basis for a pure cubic field,
Far East Journal of Mathematical Sciences 6 (1998), 1–14.
An explicit integral basis is given for a pure cubic field K = Q(θ), θ 3 + aθ + b = 0.
CB609-07 CB609/Alaca & Williams August 7, 2003 16:45 Char Count= 0
Biographies 193
16. B. K. Spearman and K. S. Williams, Cubic fields with a power basis, Rocky Mountain
Journal of Mathematics 31 (2001), 1103–1109.
It is shown that there exist infinitely many cubic fields L with a power basis such that the splitting
field M of L contains a given quadratic field K .
17. B. K. Spearman and K. S. Williams, Cubic fields with index 2, Monatshefte für Math-
ematik 134 (2002), 331–336.
Let d be a squarefree integer with d = 1 allowed. If d ≡ 1 (mod 8) it√is shown that there do
not exist any cubic fields with index 2 whose splitting field contains Q( d). If d ≡ 1 (mod 8) it
√ exist infinitely many cubic fields K with i(K ) = m(K ) = 2 whose splitting
is shown that there
field contains Q( d).
18. L. Stickelberger, Über eine neue Eigenschaft der Diskriminanten algebraischer
Zahlkörper, International Mathematische Kongress, Zürich, 1897, 182–193.
It is shown that d(K ) ≡ 0 or 1 (mod 4) for an algebraic number field K .
19. K. S. Williams, Integers of biquadratic fields, Canadian Mathematical Bulletin 13
(1970), 519–526.
√ √ √ √
This paper gives an explicit integral basis for the quartic field Q( m + n) = Q( m, n),
where m and n are distinct squarefree integers.
20. E. Zyliński, Zur Theorie der ausserwesentlichen Discriminantenteiler algebraischer
Körper, Mathematische Annalen 73 (1913), 273–274.
It is shown that a prime p can only be an inessential discriminant divisor of an algebraic number
field of degree n if p < n.
Biographies
1. E. T. Bell, Men of Mathematics, Simon and Schuster, New York, 1937.
Chapter 27 is devoted to Ernst Kummer (1810–1893) and Richard Dedekind (1831–1916).
2. R. A. Mollin, Algebraic Number Theory, Chapman and Hall/CRC Press, London/Boca
Raton, Florida, 1999.
A brief biography of Ludwig Stickelberger (1850–1936) is given on page 43.
3. The website
http://www-groups.dcs.st-and.ac.uk/˜history/
8
Dedekind Domains
An integral domain with these properties is called a Dedekind domain after Richard
Dedekind, the creator of the modern theory of ideals.
Definition 8.1.1 (Dedekind domain) An integral domain D that satisfies the fol-
lowing three properties:
D is a Noetherian domain, (8.1.1)
D is integrally closed, and (8.1.2)
each prime ideal of D is a maximal ideal, (8.1.3)
is called a Dedekind domain.
Theorem 8.1.1 Let K be an algebraic number field. Let O K be the ring of integers
of K . Then O K is a Dedekind domain.
The next theorem gives another class of integral domains that are Dedekind
domains.
194
CB609-08 CB609/Alaca & Williams August 7, 2003 16:49 Char Count= 0
Our main objective in this chapter is to show that every ideal I (= 0, 1) of a
Dedekind domain can be expressed uniquely as a product of prime ideals. We also
show that every ideal of a Dedekind domain is generated by at most two elements.
Proof: Suppose that D is a Noetherian domain that possesses at least one nonzero
ideal that does not contain a product of one or more prime ideals. Let S be the set
of all such ideals. By assumption S is not empty. As D is Noetherian, by Theorem
3.1.3 S contains a (nonzero) ideal A maximal with respect to the property of not
containing a product of one or more prime ideals. Clearly A itself is not a prime
ideal. Hence, by Theorem 1.6.1, there exist ideals B and C such that
BC ⊆ A, B ⊆ A, C ⊆ A.
Define the ideals B1 and C1 of D by
B1 = A + B, C1 = A + C.
Clearly
A ⊂ B1 , A ⊂ C1 ,
so that B1 ∈ S, C1 ∈ S. Hence there exist prime ideals P1 , . . . , Pk such that
B1 ⊇ P1 · · · Ph , C1 ⊇ Ph+1 · · · Pk .
But
B1 C1 = (A + B)(A + C) ⊆ A,
so
A ⊇ P1 · · · Pk ,
contradicting that A ∈ S.
CB609-08 CB609/Alaca & Williams August 7, 2003 16:49 Char Count= 0
Condition (iii) means that the elements of a fractional ideal have γ as a “common
denominator.”
I = α1 , . . . , αk ,
Definition 8.2.2 (The set P̃ for a prime ideal P) Let D be an integral domain
and let K be the quotient field of D. For each prime ideal P of D we define the set
P̃ by
P̃ = {α ∈ K : α P ⊆ D}.
Theorem 8.2.3 Let D be an integral domain and let P be a prime ideal of D. Then
P̃ is a fractional ideal of D.
P P̃ = D or P P̃ = P.
CB609-08 CB609/Alaca & Williams August 7, 2003 16:49 Char Count= 0
We choose k to be the least positive integer for which such an inclusion holds. Since
P1 · · · Pk ⊆ β ⊆ P
so
Pγ = Pδ/β ⊆ D
and thus γ ∈ P̃. Hence γ ∈ P̃ \ D in this case. This completes the proof that
D ⊂ P̃.
CB609-08 CB609/Alaca & Williams August 7, 2003 16:49 Char Count= 0
showing that αβ ∈ P̃. Hence P̃ is closed under multiplication. This proves that P̃ is
an integral domain, which strictly contains D. As D is a Noetherian domain, all its
ideals (integral or fractional) are finitely generated. Hence P̃ is a finitely generated
fractional ideal of D. Thus P̃ is a finitely generated D-module. Hence, by the
remark following Theorem 4.1.4, P̃ is integral over D. However, D is integrally
closed in its quotient field (since D is a Dedekind domain) so that D = P̃. This
contradicts that P̃ ⊃ D. Hence P P̃ = D.
Proof: Suppose there exist integral ideals (= 0, D) of D that are not products of
prime ideals. As D is a Dedekind domain, it is Noetherian, and so by the maximal
principle (Theorem 3.1.3) there is an ideal A(= 0, D) of D maximal with respect
to the property of not being a product of prime ideals. By Theorem 8.2.1 there exist
prime ideals P1 , . . . , Pk (k ≥ 1) of D such that
P1 · · · Pk ⊆ A.
Let k be the smallest positive integer for which such a product exists. If k = 1 then
P1 ⊆ A ⊂ D. As P1 is a prime ideal, it is a maximal ideal since D is a Dedekind
domain. Thus A = P1 . This is impossible as A is not a product of prime ideals.
Hence k ≥ 2. By Theorem 8.2.4 we have P̃ 1 P1 = D so that
P̃ 1 P1 P2 · · · Pk = D P2 · · · Pk .
Hence
P̃ 1 A ⊇ P̃ 1 P1 · · · Pk = P2 · · · Pk .
From the proof of Theorem 8.2.4 we have D ⊂ P̃ 1 so that A ⊆ P̃ 1 A. If A = P̃ 1 A
then
A ⊇ P2 · · · Pk ,
which contradicts the minimality of k as k − 1 ≥ 1. Hence A ⊂ P̃ 1 A. Since P̃ 1 A
is an ideal of D, by the maximality property of A, we have
P̃ 1 A = Q 2 · · · Q h
for prime ideals Q 2 , . . . , Q h . Then
A = AD = A P̃ 1 P1 = P1 Q 2 · · · Q h
is also a product of prime ideals, which contradicts the way A was chosen. Hence
every ideal (= 0, D) of D is a product of prime ideals.
CB609-08 CB609/Alaca & Williams August 7, 2003 16:49 Char Count= 0
Suppose now that factorization of ideals as products of prime ideals is not always
unique. By the maximal principle we may choose B to be an ideal (= 0, D)
maximal with respect to the property of having at least two distinct factorizations
as the product of prime ideals, say,
B = P1 · · · Pk = Q 1 · · · Q l ,
where P1 , . . . , Pk , Q 1 , . . . Q l are prime ideals. Then, as
P1 · · · Pk ⊆ Q 1 ,
and Q 1 is a prime ideal, by Theorem 1.6.1 we have
Pi ⊆ Q 1
for some i ∈ {1, 2, . . . , k}. Relabeling P1 as Pi and vice versa, we may suppose
that
P1 ⊆ Q 1 .
Since P1 is a prime ideal, it is a maximal ideal as D is a Dedekind domain, and thus
P1 = Q 1 .
Therefore
B P̃ 1 = P̃ 1 P1 P2 · · · Pk = P2 · · · Pk
and
B P̃ 1 = B Q̃ 1 = Q̃ 1 Q 1 · · · Q h = Q 2 · · · Q h .
If B P̃ 1 = B then B P̃ 1 P1 = B P1 , so B = B P1 . Define the fractional ideal B̃ of D
by
B̃ = P̃ 1 · · · P̃ k .
Then
B B̃ = P1 · · · Pk P̃ 1 · · · P̃ k = P1 P̃ 1 · · · Pk P̃ k = D
so that
D = B B̃ = B P̃ 1 B̃ = P1 ,
which is false as P1 (being a prime ideal) is a proper ideal of D. Hence B P̃ 1 = B.
As D ⊂ P̃ 1 we have B ⊆ B P̃ 1 . But B P̃ 1 = B, so we must have
B ⊂ B P̃ 1 .
Since B P̃ 1 is an ideal of D strictly containing B, by the maximality of B,
B P̃ 1 has exactly one factorization as a product of prime ideals. Thus from
B P̃ 1 = P2 · · · Pk = Q 2 · · · Q h we deduce that k − 1 = h − 1 (that is, k = h) and
CB609-08 CB609/Alaca & Williams August 7, 2003 16:49 Char Count= 0
after relabeling, we obtain Pi = Q i (i = 2, . . . , k). This implies that the two fac-
torizations of B into prime ideals are the same, which is a contradiction.
This completes the proof of the theorem.
Theorem 8.3.2 Let K be an algebraic number field. Then every proper integral
ideal of O K can be expressed uniquely up to order as a product of prime ideals.
Then
√ √ √
P = 2, 1 + −5 = 2, 1 + −5, 2 − (1 + −5)
√ √ √ √
= 2, 1 + −5, 1 − −5 = 2, 2 − (1 − −5), 1 − −5
√
= 2, 1 − −5,
√ √ √
P 2 = 2, 1 + −52 = 2, 1 + −52, 1 − −5
√ √
= 4, 2(1 + −5), 2(1 − −5), 6
√ √
= 22, 1 + −5, 1 − −5, 3
= 21 = 2,
√ √
P1 P2 = 3, 1 + −53, 1 − −5
√ √
= 9, 3(1 + −5), 3(1 − −5), 6
√ √
= 33, 1 + −5, 1 − −5, 2
= 31 = 3,
√ √
P P1 = 2, 1 + −53, 1 + −5
√ √ √
= 6, 2(1 + −5), 3(1 + −5), (1 + −5)2
√ √ √
= 1 + −51 − −5, 2, 3, 1 + −5
√ √
= 1 + −51 = 1 + −5,
CB609-08 CB609/Alaca & Williams August 7, 2003 16:49 Char Count= 0
Hence
√ √
2 = P 2 , 3 = P1 P2 , 1 + −5 = P P1 , 1 − −5 = P P2 ,
and
√ √
6 = 23 = 1 + −51 − −5 = P 2 P1 P2 .
A = Q1 · · · Qh ,
where Q 1 , . . . , Q h are prime ideals. Let P1 , . . . , Pn denote the distinct prime ide-
als among Q 1 , . . . , Q h . Suppose that Pi (i = 1, 2, . . . , n) occurs ai times among
Q 1 , . . . , Q h so that each ai ≥ 1 (i = 1, 2, . . . , n) and a1 + a2 + · · · + an = h. Then
A = P1a1 · · · Pnan ,
n
n
n
A= Piai , B= Pibi , AB = Pici ,
i=1 i=1 i=1
CB609-08 CB609/Alaca & Williams August 7, 2003 16:49 Char Count= 0
where we have grouped together all equal prime ideal factors so that ai , bi , ci (i =
1, 2, . . . , n) are nonnegative integers. Hence
n
n
n
n
Pici = AB = Piai Pibi = Piai +bi ,
i=1 i=1 i=1 i=1
we have
n
n
Piri +u i = Pisi +ti ,
i=1 i=1
Hence we can define the prime ideal factorization of the fractional ideal A to be
n
A= Pisi −ri
i=1
and this definition is a valid one since it is independent of the choice of common
denominator of A. With this notation, as P P̃ = 1 for any prime ideal P of D, we
have
P̃ = P −1 .
If P1 , . . . , Pn are prime ideals such that
n
n
Piai = Pibi ,
i=1 i=1
Theorem 8.3.3 The set of all nonzero integral and fractional ideals of a Dedekind
domain D forms an Abelian group with respect to multiplication. The identity
n
element of the group is 1 = D and the inverse of A = i=1 Piai , where P1 , . . . , Pn
are distinct prime ideals and a1 , . . . , an are integers (positive, negative, or zero),
is
n
A−1 = Pi−ai .
i=1
Theorem 8.3.4 Let K be an algebraic number field. Let O K be the ring of integers
of K . Then the set of all nonzero integral and fractional ideals of O K forms an
Abelian group I (K ) with respect to multiplication.
Thus
1
P −1 = P.
2
We check this directly. We have
1 1 1 √
P P = P 2 = 2, 62
2 2 2
1 √ √
= 4, 2 6, 6 = 2, 6, 3 = 1,
2
√
as 1 = 3 − 2 ∈ 2, 6, 3. This shows that P −1 = 12 P.
√
Example
√ 8.3.3 Let D = Z + Z 6. We determine the inverse A−1 of the ideal
A = 6 of D, illustrating the ideas of this section. Let
√ √
P = 2, 6, Q = 3, 6.
A−1 = P −1 Q −1 ,
where
1
P −1 = P̃ = P (Example 8.3.2)
2
and
1
Q −1 = Q̃ = Q.
3
Thus
−1 1 1 1 1
A = P Q = P Q = A.
2 3 6 6
This is clear as A( 16 A) = 16 A2 = 16 6 = 1.
where the Pi are distinct prime ideals and the ai are integers (positive, negative, or
zero).
Definition 8.4.1 (Order of an ideal with respect to a prime ideal) With the
preceding notation, the order of the nonzero ideal A of the Dedekind domain D
with respect to the prime ideal Pi (i = 1, 2, . . . , n), written ord Pi (A), is defined
by
ord Pi (A) = ai .
ord P (A) = 0.
so that
We now extend the concept of divisibility from integral ideals (Definition 8.3.1)
to fractional ideals.
The next theorem gives a necessary and sufficient condition for an ideal A to
divide an ideal B. It is usually remembered as “To contain is to divide.”
A | B if and only if A ⊇ B.
A ⊇ B ⇐⇒ A A−1 ⊇ B A−1
⇐⇒ D ⊇ B A−1
⇐⇒ B A−1 is an integral ideal of D
⇐⇒ B A−1 = C for some integral ideal C of D
⇐⇒ B = AC for some integral ideal C of D
⇐⇒ A | B.
The two basic properties of the function ord P (A) are given in the next theorem.
where the products are taken over all prime ideals P of D, so that
P ord P (AB) = AB = P ord P (A)+ord P (B) .
P P
Of course only finitely many of the exponents in the products are nonzero. Hence,
by the uniqueness property, we have
AC −1 + BC −1 = (A + B)C −1 = CC −1 = D.
Hence AC −1 ⊆ AC −1 + BC −1 = D and BC −1 ⊆ AC −1 + BC −1 = D. So AC −1
and BC −1 are both integral ideals of D. Suppose AC −1 ⊆ P and BC −1 ⊆ P. Then
D = AC −1 + BC −1 ⊆ P + P = P,
We next define the order of a nonzero element with respect to a prime ideal.
Definition 8.4.3 (Order of a nonzero element with respect to a prime ideal) Let
D be a Dedekind domain with quotient field K . For α ∈ K , α = 0, we define
α ∈ A if and only if ord P (α) ≥ ord P (A) for all prime ideals P of D.
CB609-08 CB609/Alaca & Williams August 7, 2003 16:49 Char Count= 0
Proof: We have
α ∈ A ⇐⇒ α ⊆ A
⇐⇒ A | α (by Theorem 8.4.1)
⇐⇒ ord P (A) ≤ ord P (α) for all prime ideals P of D
⇐⇒ ord P (α) ≥ ord P (A) for all prime ideals P of D.
The next theorem gives the basic properties of the order of an element with
respect to a prime ideal.
(b) For α, β, α + β ∈ K ∗ ,
ord P (α + β) ≥ min(ord P (α), ord P (β)).
Thus
Hence
Example 8.4.3 We give a simple example to show that if ord P (α) = ord P (β)
then ord P (α + β) may actually be larger than ord P (α). Take D = Z, α = 1, β =
4, P = 5. Then
Theorem 8.4.5 Let D be a Dedekind domain with quotient field K . Given any finite
set of prime ideals P1 , . . . , Pk of D and a corresponding set of integers a1 , . . . , ak
then there exists α ∈ K such that
ord Pi (α) = ai , i = 1, 2, . . . , k,
and
Proof: As
k
k
P1a1 Piai +1 | P1a1 +1 Piai +1 ,
i=2 i=2
so that
k
k
P1a1 Piai +1 ⊃ P1a1 +1 Piai +1 .
i=2 i=2
CB609-08 CB609/Alaca & Williams August 7, 2003 16:49 Char Count= 0
Thus
ord P1 (α1 ) = a1
and
ord Pi (α1 ) ≥ ai + 1 for i = 1.
Similarly we can define α j ∈ K for j = 2, . . . , k such that
ord P j (α j ) = a j
and
ord Pi (α j ) ≥ ai + 1 for i = j.
Now set
α = α1 + α2 + · · · + αk ∈ K .
Then, by Theorem 8.4.4(b), we have
ord P1 (α2 + · · · + αk ) ≥ min(ord P1 (α2 ), . . . , ord P1 (αk )) ≥ a1 + 1 > ord P1 (α1 ),
so that
ord P1 (α) = ord P1 (α1 + (α2 + · · · + αk ))
= min(ord P1 (α1 ), ord P1 (α2 + · · · + αk ))
= ord P1 (α1 ),
that is,
ord P1 (α) = a1 .
Similarly,
ord P j (α) = a j , j = 2, . . . , k.
Finally, for P = P1 , . . . , Pk we have
ord P (αi ) ≥ 0, i = 1, 2, . . . , k,
so that
ord P (α) ≥ 0.
We observe that
A | a − b ⇐⇒ a − b ⊆ A ⇐⇒ a − b ∈ A ⇐⇒ a + A = b + A.
The properties
a ≡ a (mod A),
a ≡ b (mod A) =⇒ b ≡ a (mod A),
a ≡ b (mod A), b ≡ c (mod A) =⇒ a ≡ c (mod A),
a ≡ b (mod A) =⇒ ac ≡ bc (mod A)
P | Q1.
Now
Hence
P1 | P2a2 · · · Pkak ,
CB609-08 CB609/Alaca & Williams August 7, 2003 16:49 Char Count= 0
x1 + y1 = 1.
Thus
Now let
α = α1 y1 + · · · + αk yk ∈ D.
α ≡ α1 y1 ≡ α1 (mod P1a1 ).
Similarly,
a
α ≡ α j (mod P j j ), j = 2, . . . , k.
Example 8.4.4 Let D = Z[x]. D is not a Dedekind domain as the prime ideal x
is not a maximal ideal (Example 1.5.6). Consider the pair of congruences
α ≡ 0 (mod 2),
α ≡ 1 (mod x).
The moduli 2 and x are distinct prime ideals. However, the congruences are not
simultaneously solvable in D, since any solution of α ≡ 0 (mod 2) has an even
constant term, whereas any solution of α ≡ 1 (mod x) has a constant term equal
to 1. This shows that the Chinese remainder theorem does not necessarily hold in
an integral domain that is not a Dedekind domain.
CB609-08 CB609/Alaca & Williams August 7, 2003 16:49 Char Count= 0
Hence
A = α + AB.
Finally,
Exercises
1. Let D be a Dedekind domain. Let A and B be integral ideals of D with A = D, B = D.
Prove from first principles that AB = D.
2. Let D be a Dedekind domain. Let A be a nonzero integral ideal of D. Let P be a prime
ideal of D. If P does not divide A prove that ord P (A) = 0.
√
3. Determine all fractional ideals of Z + Z −1.
√
4. Find all ideals in Z + Z −6 that contain 6.
5. Let D be a principal ideal domain with quotient field K . Prove that every fractional
ideal of D is of the form {dα | d ∈ D} for some α ∈ K .
6. Let K be an algebraic number field of degree n. Let a be a nonzero rational integer.
Prove that a belongs to at most a n integral ideals of O K .
√ √ √
7. Show that 3, 1 + 2 −5 | 1 + 2 −5 in O K , where K = Q( −5). Determine an
integral ideal A such that
√ √
1 + 2 −5 = 3, 1 + 2 −5A.
√ √
8. Determine the fractional ideal 3, 1 + 2 −5−1 of O K , where K = Q( −5).
9. Let K be an algebraic number field. Let I be an integral ideal of O K . Let a ∈ I . Prove
that there exists an integral ideal I of O K such that a = I I .
10. Let K be an algebraic number field. Let I be a nonzero integral ideal of O K . Let α ∈ K
have the following property:
a ∈ I =⇒ aα ∈ I.
Prove that α ∈ O K .
√ √ √ √
11. Let I be the ideal of Z + Z −5 generated by 1 + −5, 3 + −5, and 19 + 9 −5.
√
Determine α, β ∈ Z + Z −5 such that I = α, β.
12. Let I and J be nonzero integral ideals of a Dedekind domain D. Let P1 , . . . , Pk be the
distinct prime ideals dividing either I or J (or both) so that
I = P1a1 · · · Pkak , J = P1b1 · · · Pkbk ,
for nonnegative integers a1 , . . . , ak , b1 , . . . , bk . The greatest common divisor gcd (I, J )
and the least common multiple lcm (I, J ) of I and J are defined by
gcd(I, J ) = P1min(a1 ,b1 ) · · · Pkmin(ak ,bk ) ,
lcm(I, J ) = P1max(a1 ,b1 ) · · · Pkmax(ak ,bk ) .
Prove that
gcd (I, J ) = I + J
CB609-08 CB609/Alaca & Williams August 7, 2003 16:49 Char Count= 0
and
lcm (I, J ) = I ∩ J.
13. Prove that a Dedekind domain is a unique factorization domain if and only if it is a
principal ideal domain.
14. Let D be a Dedekind domain. Let A, B, C be ideals of D with A = 0 and AB = AC.
Prove that B = C.
15. Let D be a Dedekind domain. Let A and B be nonzero integral ideals of D. Prove that
there exists a ∈ A such that gcd (AB, a) = A.
16. Let D be a Dedekind domain. Let A and B be nonzero integral ideals of D. Prove that
there is an integral ideal C of D such that AC is a principal ideal and gcd (B, C) = D.
17. Let D be a Dedekind domain. Let A be a nonzero integral ideal of D. Prove that there
exist only finitely many integral ideals of D that divide A.
18. Let D be a Dedekind domain. A nonzero integral ideal I of D is said to be primary if
the following condition holds:
a, b ∈ D, ab ∈ I, a ∈ I =⇒ bm ∈ I for some m ∈ N.
Prove that a primary ideal must be a power of a prime ideal.
19. Let K be an algebraic number field. Prove that O K contains infinitely many prime
ideals.
√
20. Determine the prime ideal factorization of 54 in Z + Z −6.
21. Let D be a Dedekind domain. Let I be an ideal of D with I = 0, 1. Prove that
D/I D/P1a1 × · · · × D/Prar ,
where
I = P1a1 · · · Prar
is the factorization of I into distinct prime ideals P1 , . . . , Pr .
Suggested Reading
1. F. T. Howard, A generalized Chinese remainder theorem, The College Mathematics
Journal 33 (2002), 279–282.
An extension of the Chinese remainder theorem that allows the moduli of the linear congruences
to have common factors is proved.
2. O. Zariski and P. Samuel, Commutative Algebra, Volume 1, van Nostrand, Princeton,
New Jersey, 1958.
Chapter 5 of this classic book on algebra is devoted to Dedekind domains and the classical theory
of ideals.
CB609-09 CB609/Alaca & Williams August 7, 2003 16:51 Char Count= 0
9
Norms of Ideals
218
CB609-09 CB609/Alaca & Williams August 7, 2003 16:51 Char Count= 0
Proof: The proof is by induction on the size n of the matrix C. If n = 1 the result
is clearly true. Now suppose that it is true for all (n − 1) × (n − 1) matrices with
rational integer entries. Let C be a given n × n matrix with rational integer entries.
If C is the zero matrix there is nothing to prove, so we may suppose that C = On .
Let k denote any one of the nonzero entries in C. By means of elementary operations
of type 1 we can transform C into a matrix B = (bi j ) in which b11 = k. If k does
not divide all the remaining entries in the first row and first column we can find an
integer j (2 ≤ j ≤ n) such that b1 j or b j1 = qk + r with 0 < r < |k|; then by an
elementary operation of type 2, subtracting q times the first row or column from the
jth row or column, we obtain a matrix with an entry r < |k|. Applying elementary
operations of type 1 we can move r to the (1, 1) position and repeat the process.
After a finite number of operations we obtain a matrix in which the (1, 1) entry
divides all the entries in the first row and column. Thus by means of a finite number
of operations of types 1 and 2 we can transform the matrix into one of the form
d1 0 ··· 0
0
. (n − 1) × (n − 1) .
. . submatrix with
integer entries
0
Applying the inductive hypothesis to the (n − 1) × (n − 1) submatrix we finally
get a matrix of the required diagonal type.
Clearly, elementary operations of types 1 and 2 at most change the sign of the
determinant so that
|det C| = |d1 | · · · |dn |.
Proof: We wish to transform the matrix C into the form given in Theorem 9.1.1 by
means of elementary operations of types 1 and 2.
An elementary operation of type 1, that is, interchanging rows or columns of C,
corresponds to rearranging the order of the generators η1 , . . . , ηn of H or ω1 , . . . , ωn
of G, and it so leaves [G : H ] unchanged.
The elementary operation of type 2, which adds k times the ith row to the lth row,
corresponds to replacing ci j by ci j + kcl j ( j = 1, 2, . . . , n) and hence replaces
n
ηi = ci j ω j
j=1
by
n
ηi + kηl = (ci j + kcl j )ω j .
j=1
n
= ci h ωh + (ci j + kcil )ω j + cil (ωl − kω j )
h=1
h = j, l
n
= ci h ωh ,
h=1
where
ci h , h = j,
ci h =
ci j + kcil , h = j,
CB609-09 CB609/Alaca & Williams August 7, 2003 16:51 Char Count= 0
and
ωh , h= l,
ωh =
ωl − kω j , h = l.
Thus [G : H ] remains unchanged.
Hence transforming C into diag(d1 , . . . , dn ), where |d1 | · · · |dn | = |det C|, by
elementary operations as in Theorem 9.1.1, we obtain a set of generators for G,
namely,
G = ω1 , . . . , ωn ,
such that
H = d1 ω1 , . . . , dn ωn .
Clearly
x1 ω1 + · · · + xn ωn ∈ H
⇐⇒ x1 ω1 + · · · + xn ωn = y1 d1 ω1 + · · · + yn dn ωn
for some yi ∈ Z (i = 1, 2, . . . , n)
⇐⇒ xi = yi di (i = 1, 2, . . . , n)
⇐⇒ di | xi (i = 1, 2, . . . , n).
Suppose now that det C = 0. Hence d1 · · · dn = 0 so that each di = 0 (i =
1, 2, . . . , n). Then a complete set of coset representatives for G modulo H is
{x1 ω1 + · · · + xn ωn | x1 = 0, 1, . . . , |d1 | − 1; . . . ; xn = 0, 1, . . . , |dn | − 1},
and thus
[G : H ] = |d1 | · · · |dn | = |det C|.
Finally, suppose that det C = 0. Hence d1 · · · dn = 0 so that di = 0 for some
i ∈ {1, 2, . . . , n}. Then kωi + H (k = 0, 1, 2, . . .) are distinct cosets of H in G so
that [G : H ] = ∞.
Proof: Let {η1 , . . . , ηn } be a basis for A and {ω1 , . . . , ωn } an integral basis for K .
Then
n
ηi = ci j ω j , i = 1, 2, . . . , n,
j=1
CB609-09 CB609/Alaca & Williams August 7, 2003 16:51 Char Count= 0
If α ∈ Q then by Theorem 6.3.4 we know that all the K -conjugates of α are all
equal to α. Hence for α ∈ Q we have
tr(α) = α + · · · + α = nα
and
N (α) = α · · · α = α n .
√
If K is a quadratic field then K = Q( m) for some squarefree integer m.
√
Let α ∈ K . Then α = r + s m for some r, s ∈ Q. The K -conjugates of α are
√ √
α = r + s m and α = r − s m. The trace of α is
tr(α) = α + α = 2r
and the norm of α is
N (α) = αα = r 2 − s 2 m.
Recalling Definition 2.2.1 we observe that φm (α) = |N (α)|.
CB609-09 CB609/Alaca & Williams August 7, 2003 16:51 Char Count= 0
n
N (αβ) = σk (αβ) = (σk (α)σk (β))
k=1
k=1
n
n
= σk (α) σk (β) = N (α)N (β).
k=1 k=1
n
σk (α) = N (α) = ±1.
k=1
Set
n
β=± σk (α),
k=2
so that
αβ = 1.
N (α) = ± p,
α = βγ .
N (β) or N (γ ) = ±1.
CB609-09 CB609/Alaca & Williams August 7, 2003 16:51 Char Count= 0
We emphasize
√ that the converse
√ of Theorem 9.2.3 is not true. To see this take
K =√ Q( −5) and √ α = 1 + −5. Then α is an irreducible in O K but N (α) =
(1 + −5)(1 − −5) = 6 is not prime.
In the next theorem we use Theorems 9.2.2 and 9.2.3 to give a condition that,
when satisfied by an element α of a cubic field, guarantees that α is a unit, and we
give a similar condition for α to be an irreducible.
Theorem 9.2.4 Let a and b be integers such that the cubic polynomial x 3 + ax + b
is irreducible in Z[x]. Let
K = Q(θ ), where θ 3 + aθ + b = 0,
then r + sθ + tθ 2 is a unit in O K .
(b) If r, s, t are integers such that
θ + θ + θ = 0,
θ θ + θ θ + θ θ = a,
θθ θ = −b.
Then
2 2
θ 2 + θ + θ = (θ + θ + θ )2 − 2(θθ + θ θ + θ θ) = −2a,
2 2 2 2
θ 2 θ + θ θ + θ θ 2 = (θ θ + θ θ + θ θ)2 − 2θθ θ (θ + θ + θ ) = a 2 ,
2 2 2 2
θθ + θ 2 θ + θ θ + θ 2 θ + θ θ + θ θ
= (θ + θ )θ θ + (θ + θ )θθ + (θ + θ )θ θ
= −θ θ θ − θ θ θ − θθ θ = −3θθ θ = 3b.
CB609-09 CB609/Alaca & Williams August 7, 2003 16:51 Char Count= 0
Hence
2 2
N (r + sθ + tθ 2 ) = (r + sθ + tθ 2 )(r + sθ + tθ )(r + sθ + tθ )
= r 3 + s 3 θθ θ + t 3 (θθ θ )2 + r s 2 (θθ + θ θ + θ θ)
2 2 2 2
+ r t 2 (θ 2 θ + θ θ + θ θ 2 ) + r 2 s(θ + θ + θ )
2 2
+ r 2 t(θ 2 + θ + θ ) + s 2 tθθ θ (θ + θ + θ )
+ st 2 θθ θ (θθ + θ θ + θ θ)
2 2 2 2
+ r st(θθ + θ 2 θ + θθ + θ 2 θ + θ θ + θ θ )
= r 3 − bs 3 + b2 t 3 + ar s 2 + a 2r t 2 − 2ar 2 t − abst 2 + 3br st.
The assertions of the theorem now follow from Theorems 9.2.2 and 9.2.3.
a = −4, b = 2, r = −1, s = 1, t = 0
of Theorem 9.2.4(a) as
a = −4, b = 2, r = −1, s = 2, t = 0
and
Theorem 9.2.5 Let K be an algebraic number field of degree n. Let O K be the ring
of integers of K . Let α ∈ O K . Then
N (α) = |N (α)|.
2
σ1 (ω1 ) ··· σ1 (ωn )
σ (ω ) ··· σ2 (ωn )
2
2 1
= (σ1 (α)σ2 (α) · · · σn (α))
. ..
.. ··· .
σn (ω1 ) ··· σn (ωn )
= N (α)2 d(K ),
so that
D(α)
N (α) = = N (α)2 = |N (α)|.
d(K )
We see that Example 7.1.2 and Theorem 7.1.6 are special cases of Theorem 9.2.5.
We also observe that if α ∈ O K , where K is an algebraic number field, then
Next we determine the norm of the principal ideal α in terms of the constant term
of the minimal polynomial of α over Q.
Then
and thus
b0 = (−1)m α1 α2 · · · αm .
Again by Theorem 6.3.2 a complete set of conjugates of α is
α1 , . . . , α 1 , α2 , . . . , α 2 , . . . , α m , . . . , α m ,
where each αi is repeated n/m times. Hence
n/m n/m
N (α) = α1 α2 · · · αmn/m = (α1 α2 · · · αm )n/m
n/m
= ((−1)m b0 )n/m = (−1)n b0
and thus by Theorem 9.2.5
n/m
N (α) = |N (α)| = |(−1)n b0 | = |b0 |n/m .
√ √ √
Example 9.2.2
√ Let K = Q( 2 + 3). We determine N ( 2). The minimal poly-
nomial of 2 over Q is x 2 − 2. Hence in the notation of Theorem 9.2.6 we have
n = 4, m = 2, b1 = 0, b0 = −2. Thus by Theorem 9.2.6 we obtain
√
N ( 2) = | − 2|4/2 = 22 = 4.
Proof: Let P1 , . . . , Pn be the set of distinct prime ideals for which either
ord Pi (A) = 0 or ord Pi (AB) = 0 (or both).
This set is nonempty as A = D. By Theorem 8.4.5 we can find an element γ of the
quotient field of D such that
ord Pi (γ ) = ord Pi (A), i = 1, 2, . . . , n,
ord P (γ ) ≥ 0, P = P1 , . . . , Pn .
Thus
ord P (γ ) ≥ ord P (A) for all prime ideals P,
CB609-09 CB609/Alaca & Williams August 7, 2003 16:51 Char Count= 0
and so
γ ∈ A.
Now for i = 1, 2, . . . , n we have
ord Pi (γ + AB) = min(ord Pi (γ ), ord Pi (AB))
= min(ord Pi (γ ), ord Pi (AB))
= min(ord Pi (A), ord Pi (AB))
= ord Pi (A),
as B is an integral ideal. For a prime ideal P = P1 , . . . , Pn we have ord P (A) =
ord P (AB) = 0 so that
ord P (γ + AB) = min(ord P (γ ), ord P (AB))
= min(ord P (γ ), 0)
=0
= ord P (A).
Hence
ord P (γ + AB) = ord P (A)
for all prime ideals P, and so
A = γ + AB.
Clearly,
δ − αi ∈ A = γ + AB
so there exist σ ∈ D and τ ∈ AB such that
δ − αi = σ γ + τ.
Similarly, there is a unique integer j (1 ≤ j ≤ l) such that
σ ≡ β j (mod B),
that is,
σ − β j ∈ B.
As γ ∈ A we have
(σ − β j )γ ∈ AB.
Hence
δ = αi + σ γ + τ = αi + β j γ + (σ − β j )γ + τ ≡ αi + β j γ (mod AB).
This shows that the set of kl elements αi + β j γ + AB (i = 1, . . . , k; j = 1, . . . , l)
is a complete set of representatives of D/AB. We must still show that they are
distinct. Suppose
αi + β j γ + AB = α p + βq γ + AB.
Then
αi + β j γ ≡ α p + βq γ (mod AB)
and thus
αi − α p ≡ (βq − β j )γ (mod AB).
But γ ∈ A so
αi − α p ∈ A.
Thus i = p and
β j γ ≡ βq γ (mod AB).
Hence
(β j − βq )γ ∈ AB.
h bi
Now let B = i=1 Pi (bi > 0) be the prime ideal decomposition of B. Then
ord Pi (A) = ord Pi (γ + AB)
= min(ord Pi (γ ), ord Pi (AB))
= min(ord Pi (γ ), ord Pi (A) + ord Pi (B))
CB609-09 CB609/Alaca & Williams August 7, 2003 16:51 Char Count= 0
Definition 9.4.1 is valid for if I and J are nonzero integral ideals of O K and α
and β are nonzero elements of O K such that
1 1
A= I = J
α β
CB609-09 CB609/Alaca & Williams August 7, 2003 16:51 Char Count= 0
then
β I = α J,
so that we have the equal products of integral ideals
βI = αJ,
and thus by Theorem 9.3.2
N (β)N (I ) = N (βI ) = N (αJ ) = N (α)N (J ),
so that
N (I ) N (J )
= .
N (α) N (β)
When the fractional ideal A of Definition 9.4.1 is actually an integral ideal the two
definitions of the norm coincide.
The next theorem shows that the multiplicative property of the norm carries over
to fractional ideals.
Theorem 9.4.1 Let K be an algebraic number field. Let O K be its ring of integers.
Let A and B be nonzero fractional ideals of O K . Then
N (AB) = N (A)N (B).
Proof: As A and B are nonzero fractional ideals of O K there exist nonzero integral
ideals I and J of O K and nonzero elements α and β of O K such that
1 1
A= I, B = J.
α β
Then
1
AB = IJ
αβ
so that
N (I J )
N (AB) = (Definition 9.4.1)
N (αβ)
N (I J )
= (Definition 1.6.2)
N (αβ)
N (I )N (J )
= (Theorem 9.3.2)
N (α)N (β)
N (I ) N (J )
= ·
N (α) N (β)
= N (A)N (B).
√ √
Example 9.4.1 Let K = Q( 6) so that O K = {a + b 6 | a, b ∈ Z}. Let A be the
fractional ideal of O K given by
CB609-09 CB609/Alaca & Williams August 7, 2003 16:51 Char Count= 0
Exercises 233
1√
A = 1, 6.
2
Then
1
A= I,
2
√
where I is the integral ideal 2, 6. Now
√ √ √
I 2 = 2, 62 = 4, 2 6, 6 = 22, 6, 3 = 2
√
(as 1 = 3 − 2 ∈ 2, 6, 3) so that
N (I )2 = N (I 2 ) = N (2) = 22
and thus
N (I ) = 2.
Hence
N (I ) 2 1
N (A) = N ( 12 I ) = = 2 = .
N (2) 2 2
Exercises
1. Let p be a prime such that p ≡ 3 or 5 (mod 8). Prove that there does not exist an element
α ∈ OQ(√ p) such that N (α) = 2.
2. Let p be a prime such that p ≡ 5 or 7 (mod 8). Prove that there does not exist an element
α ∈ OQ(√ p) such that N (α) = −2.
3. Let K be an algebraic number field and O K its ring of integers. If I is an integral ideal
of O K such that N (I ) is a prime, then I is a prime ideal.
4. Let K be an algebraic number field and O K its ring of integers. If I is a nonzero integral
ideal of O K , prove that I | N (I ).
5. Let K be an algebraic number field. Let n be a given positive integer. Prove that there
are only finitely many integral ideals I of O K such that N (I ) = n.
6. Let K = Q(θ ), where θ 3 − θ − 1 = 0. Prove that 23, 3 − θ is a prime ideal in O K .
√ √
7. Let K = Q( −23). Let I = 2, 12 (1 + −23).
(a) Prove that N (I ) = 2.
√
(b) Prove that I 3 = (−3 + −23)/2.
(c) Use each of (a) and (b) to prove that I is not a principal ideal.
8. Let K be an algebraic number field and O K its ring of integers. Let I be an integral
ideal of O K such that N (I ) = |N (a)| for some a ∈ I . Prove that I = a.
9. Let K be an algebraic number field and O K its ring of integers. Let P be a prime ideal
of O K . Prove that G = {a + P | a ∈ O K , a ∈ P} is a cyclic group with respect to
multiplication. What is the order of G?
10. Let K be an algebraic number field and O K its ring of integers. Let P be a prime ideal
of O K . Prove that P ∩ Z = p for some prime p.
CB609-09 CB609/Alaca & Williams August 7, 2003 16:51 Char Count= 0
11. Let K be an algebraic number field and O K its ring of integers. Show that
a p ≡ a (mod P) ⇐⇒ a ≡ m (mod P) for some m ∈ Z,
where P ∩ Z = p (see Exercise 10).
12. Determine the fractional ideals of Z + Zi.
13. Let K be an algebraic number field and O K its ring of integers. Let m ∈ Z \ {0}. Prove
that there exist only finitely many integral ideals of O K to which m belongs.
√
14. Find all the ideals of Z + Z√−5 that contain 6.
15. Find all the ideals of Z + Z 2 with norm 12.
16. Determine the set of positive integers that are not norms of ideals of Z + Zi.
17. Let K be an algebraic number field. Let O K be its ring of integers. Let P be a prime
ideal of O K . Let a ∈ O K be such that P a. Prove that
a N (P) − 1 ≡ 0 (mod P).
18. Give an example of an algebraic number field K and an integral ideal I = a, b, c of
O K such that I = a, b, a, c, b, c.
19. Determine all complex quadratic fields K for which O K possesses elements of norm
38 and trace 11.
20. Let K be an algebraic number field. Let α, β ∈ O K \ {0}. Prove that N (α)tr(β/α) ∈ Z.
√ √ √
21. Solve 3x ≡ 5 (mod A) in Z + Z −5, where A = 3 −5, 10 + 10 −5.
22. Let K be an algebraic number field. Let I be an integral ideal of O K . Let m be the least
positive integer in I . Prove that m | N (I ).
23. Let K be an algebraic number field. Let I be an integral ideal of O K such that pq | N (I ),
where p and q are distinct primes. Prove that I is not a prime ideal.
24. Let K be a quadratic field. Let α ∈ O K be such that |N (α)| = ab, where a and b are
coprime positive integers. Prove that
a, αb, α = α.
25. Prove that d1 , d2 , . . . , dn in Theorem 9.1.1 can be arranged to satisfy d1 | d2 | · · · | dn .
26. Let K be an algebraic number field. Prove that O K is a principal ideal domain if and
only if for every pair (α, β) ∈ O K × O K such that
α = 0, β = 0, β α, |N (α)| ≥ |N (β)|,
there exist γ ∈ O K and δ ∈ O K such that
0 < |N (αγ − βδ)| < |N (β)|.
Suggested Reading
1. P. B. Bhattacharya, S. K. Jain, and S. R. Nagpaul, Basic Abstract Algebra, second edition,
Cambridge University Press, Cambridge, United Kingdom, 1994.
Chapter 20 is devoted to the Smith normal form of a matrix over a principal ideal domain.
2. C. C. MacDuffee, An Introduction to Abstract Algebra, Wiley, New York, 1956.
Section 105 discusses the Smith normal form of a matrix and mentions that this form is named for
H. J. S. Smith.
CB609-09 CB609/Alaca & Williams August 7, 2003 16:51 Char Count= 0
Biographies 235
3. I. N. Stewart and D. O. Tall, Algebraic Number Theory, second edition, Chapman and
Hall, London, 1987.
Section 1.6 contains a very readable discussion of free Abelian groups.
Biographies
1. H. J. S. Smith, Report on the Theory of Numbers, Chelsea, New York, 1964.
This book contains a biographical sketch of the Irish mathematician Henry John Stephen Smith
(1826–1883).
2. The website
http://www-groups.dcs.st-and.ac.uk/˜history/
10
Factoring Primes in a Number Field
P | p.
P ∩ Z = p
P ⊇ p
P | p.
P | q.
P ⊇ p, q.
As p and q are distinct primes we have gcd( p, q) = 1 so that there are integers
a and b such that ap + bq = 1. Hence 1 ∈ p, q ⊆ P. Thus O K ⊆ P, which is
impossible.
Hence the prime p is uniquely determined by P | p.
236
CB609-10 CB609/Alaca & Williams August 7, 2003 16:53 Char Count= 0
The rational prime p in Theorem 10.1.1 is called the prime lying below P as
P ⊇ p. Given a rational prime p, any prime ideal P such that P | p is said to
be a prime ideal lying over p.
Next we relate the norm of the prime ideal P to the prime p lying below P.
a − a = P Q
a + P = a + P ⇐⇒ a ≡ a (mod p).
Hence the cosets a + P (a ∈ {0, 1, . . . , p − 1}) are distinct and the prime field of
O K /P is
F = {a + P | a = 0, 1, . . . , p − 1} Z/ p.
O K /P = {a + P | a = 0, 1, . . . , p − 1}.
e1 f 1 + · · · + eg f g = n.
where
ei = e K (Pi ), i = 1, 2, . . . , g.
The following theorem of Dedekind, which we shall not prove here, enables us
to recognize when a rational prime p ramifies in an algebraic number field K .
Theorem 10.1.5 (Dedekind) Let K be an algebraic number field. Then the rational
prime p ramifies in K if and only if p | d(K ).
In the next section we examine the factorization of a rational prime p into prime
ideals in O K , when K is a quadratic field.
We conclude this section by proving the following simple but useful result.
Theorem 10.1.6 Let K be an algebraic number field. Let I (= 0) be an ideal of
OK .
which contradicts that p is a rational prime as N (P), N (Q), and N (A) are positive
integers with N (P) > 1 and N (Q) > 1. Thus I is a prime ideal.
(b) By Theorem 9.1.3 we have
N (I ) = card (O K /I ).
Hence
that is,
N (I )x ∈ I, for all x ∈ O K .
Taking x = 1 ∈ O K we obtain
N (I ) ∈ I
as asserted.
In other words,
(i) p = P1 P2 , N (P1 ) = N (P2 ) = p, P1 = P2 ,
(ii) p = P 2 , N (P) = p,
(iii) p = P, N (P) = p 2 ,
where P1 , P2 , P denote prime ideals of O K . In case (i) we say that p splits in K ,
in case (ii) that p ramifies in K , and in case (iii) that p is inert (or remains prime)
in K . In cases (i) and (iii) p is unramified in K .
Our next theorem gives necessary and sufficient conditions for each of (i), (ii),
(iii) to occur. As usual mp denotes the Legendre symbol of the integer m modulo
the odd prime p.
Theorem 10.2.1 Let K be a quadratic field so that there exists a squarefree integer
√
m such that K = Q( m). Let p be a rational prime.
(a) If p > 2, m
p
= 1 or p = 2, m ≡ 1 (mod 8) then
p = P1 P2 ,
where P1 and P2 are distinct prime ideals with N (P1 ) = N (P2 ) = p.
(b) If p > 2, p | m or p = 2, m ≡ 2 or 3 (mod 4) then
p = P 2 ,
where P isa prime
ideal with N (P) = p.
(c) If p > 2, p = −1 or p = 2, m ≡ 5 (mod 8) then
m
p is a prime ideal of O K .
p = P1 P2 (P1 = P2 ) or P12 .
√
In each case we have N (P1 ) = p so that f K (P1 ) = 1. Hence, as m ∈ O K , there
exists a ∈ Z such that
√
m ≡ a (mod P1 )
and so
m ≡ a 2 (mod P1 ).
But, as m ∈ Z, a 2 ∈ Z and P1 | p, we must have
m ≡ a 2 (mod p),
contradicting m
p
= −1. Thus p is a prime ideal in O K , and this case falls
under (c).
√
(iii): p > 2, p | m. Set P = p, m. Then
√ √ √
P 2 = p, m p, m = p 2 , p m, m = pI,
where I is the ideal
√
I = p, m, m/ p.
As m is squarefree, we have gcd( p, m/ p) = 1, so that there exist integers x and y
such that
x p + y(m/ p) = 1.
Hence
1 = x p + y(m/ p) ∈ I,
so I = 1, that is, p = P 2 , and this case falls under (b).
√
(iv): p = 2, m ≡ 2 (mod 4). Set P = 2, m. Then
√ √ √
P 2 = 2, m2, m = 4, 2 m, m = 2I,
CB609-10 CB609/Alaca & Williams August 7, 2003 16:53 Char Count= 0
From the proof of Theorem 10.2.1 we see that we can express the factorizations
of the principal ideal p (with p prime) into prime ideals of O K , where K is the
√
quadratic field Q( m) (m squarefree), as follows:
2, √ √ if m ≡ 5 (mod 8),
2, 2 (1 + m)2, 2 (1 − m), if m ≡ 1 (mod 8),
1 1
2 = √
2, 1 + m2 , if m ≡ 3 (mod 4),
√ 2
2, m , if m ≡ 2 (mod 4),
and for p > 2
p, if p m and x 2 ≡ m (mod p)
√ √
is insolvable,
p = p, x + m p, x − m if p m and x 2 ≡ m (mod p)
is solvable,
√
p, m2 , if p | m.
Recalling that for a squarefree integer m
√ m, if m ≡ 1 (mod 4),
d(Q( m)) =
4m, if m ≡ 2, 3 (mod 4),
CB609-10 CB609/Alaca & Williams August 7, 2003 16:53 Char Count= 0
Example 10.2.1
√
(a) 11 is inert in Q( −163) as
−163 2
= = −1.
11 11
√
(b) 23 is inert in Q( 37) as
37 14 2 7
= = = (+1)(−1) = −1.
23 23 23 23
√
(c) 2 ramifies in Q( 7) as
28
= 0.
2
√
Indeed 2 = 2, 1 + 72 .
CB609-10 CB609/Alaca & Williams August 7, 2003 16:53 Char Count= 0
√
Let K be the quadratic field Q( m), where m is a squarefree integer. There are
exactly two monomorphisms : K −→ C, namely, 1 and σ given by
√ √
1(a + b m) = a + b m,
√ √
σ (a + b m) = a − b m,
for all a, b ∈ Q. Let I be an ideal of O K . By Theorem 8.5.1 we know that I is
generated by at most two elements. Hence I = α or I = α, β and we define the
conjugate ideal σ (I ) of I by
σ (I ) = σ (α) or σ (α), σ (β)
respectively. It is customary to write α
for σ (α) and
√ I for √σ (I ). Clearly
(α ) =√ σ (α) =√α so that (I ) = I . Thus if I = 2 + 3, 1 − 2 3 then I =
2
Theorem 10.2.3 Let I and J be ideals of the ring of integers of a quadratic field
K . Then
(I J )
= I
J
.
Proof: As α = α, α we may suppose that both I and J are generated by two
elements, say,
I = α, β, J = γ , δ.
Then
I J = α, βγ , δ = αγ , βγ , αδ, βδ
CB609-10 CB609/Alaca & Williams August 7, 2003 16:53 Char Count= 0
so that
(I J )
= (αγ )
, (βγ )
, (αδ)
, (βδ)
= α
γ
, β
γ
, α
δ
, β
δ
= α
, β
γ
, δ
= I
J
as asserted.
Proof: The assertion of the theorem is trivial if I = 0 or 1, so we may suppose
that I = 0, 1. Let
b1 br
I = P1a1 P1
· · · Prar Pr
Q c11 · · · Q cs s R1d1 · · · Rtdt
be the prime ideal decomposition of I , where P1 , . . . , Pr are distinct prime ideals
such that
P P
= p, N (P) = N (P
) = p, P = P
,
Q 1 , . . . , Q s are distinct prime ideals such that
Q = Q
= q, N (Q) = q 2 ,
R1 , . . . , Rt are distinct prime ideals such that
R = R
, R 2 = r , N (R) = r,
and p, q, r denote rational primes. Then, by Theorem 10.2.3, we have
a1 ar
I
= P1
P1b1 · · · Pr
Prbr Q c11 · · · Q cs s R1d1 · · · Rtdt .
Hence
a1 +b1 ar +br
I I
= P1a1 +b1 P1
· · · Prar +br Pr
Q 2c 1 2cs 2d1
1 · · · Q s R1 · · · Rt
2dt
= p1 a1 +b1 · · · pr ar +br q1 2c1 · · · qs 2cs r1 d1 · · · rt dt
= p1a1 +b1 · · · prar +br q12a · · · qs2cs r1d1 · · · rtdt .
Further, by Theorem 9.3.2, we have
N (I ) = N (P1 )a1 N (P1
)b1 · · · N (Pr )ar N (Pr
)br N (Q 1 )c1 · · ·
N (Q s )cs N (R1 )d1 · · · N (Rt )dt
= p1a1 p1b1 · · · prar prbr q12c1 · · · qs2cs r1d1 · · · rtdt .
CB609-10 CB609/Alaca & Williams August 7, 2003 16:53 Char Count= 0
Thus
I I
= N (I )
as asserted.
Theorem 10.3.1 Let K = Q(θ ) be an algebraic number field of degree n such that
O K = Z + Zθ + · · · + Zθ n−1 .
Let p be a rational prime. Let
f (x) = irrQ θ ∈ Z[x].
−
Let denote the natural map : Z[x] → Z p [x], where Z p = Z/ pZ. Let
f¯ (x) = g1 (x)e1 · · · gr (x)er ,
CB609-10 CB609/Alaca & Williams August 7, 2003 16:53 Char Count= 0
where g1 (x), . . . , gr (x) are distinct monic irreducible polynomials in Z p [x] and
e1 , . . . , er are positive integers. For i = 1, 2, . . . , r let f i (x) be any monic polyno-
mial of Z[x] such that f¯ i = gi . Set
Pi = p, f i (θ), i = 1, 2, . . . , r.
p = P1e1 · · · Prer
and
N (Pi ) = p deg fi , i = 1, 2, . . . , r.
νi (h(θ)) = h̄(θi ).
Then
so that
and thus
p, f i (θ) ⊆ ker νi .
ḡ(θi ) = νi (g(θ )) = 0
proving
ker νi ⊆ p, f i (θ).
Pi = p, f i (θ ) = ker νi , i = 1, 2, . . . , r.
f j (θ ) = pg(θ) + f i (θ)h(θ)
g j (x) = gi (x)l(x)
for some l(x) ∈ Z p [x]. As gi (x) and g j (x) are both monic polynomials, which are
irreducible in Z p [x], we have l(x) = 1 so that gi (x) = g j (x) and thus i = j.
We show next that
p = P1e1 · · · Prer .
(A + B1 )(A + B2 ) ⊆ A + B1 B2 ,
so that
and so
p | P1e1 · · · Prer .
Now Pi = p, f i (θ ) ⊇ p, so
Pi | p, i = 1, 2, . . . , r.
CB609-10 CB609/Alaca & Williams August 7, 2003 16:53 Char Count= 0
Hence
p = P1k1 · · · Prkr ,
where
ki ∈ {1, 2, . . . , ei }, i = 1, 2, . . . , r. (10.3.1)
Now
so that
where
Hence we have
so that
d1 k1 + · · · + dr kr = n. (10.3.2)
Comparing degrees in
we obtain
d1 e1 + · · · + dr er = n. (10.3.3)
ki = ei , i = 1, 2, . . . , r,
so that
p = P1e1 · · · Prer ,
as asserted.
Finally, we observe that
f¯i
N (Pi ) = p di = p deg = p deg fi , i = 1, 2, . . . , r.
CB609-10 CB609/Alaca & Williams August 7, 2003 16:53 Char Count= 0
The set P( f ) is discussed in the beautiful article by Gerst and Brillhart [4].
If f (x) factors modulo p into a product of distinct linear factors, we say that
f (x) splits completely modulo p. The set of all primes p such that f (x) splits
completely modulo p is denoted by Spl( f ). This set is discussed by Wyman in his
classic article [7]. Thus for example
Spl(x 3 − 31x + 62) = { p (prime > 2) ≡ 1, 2, 4, 8, 15, 16, 23, 27, 29, 30
(mod 31)}
(see [5]).
The next section will be devoted to numerical examples illustrating Theorem
10.3.1.
Example
√ 10.4.1 √We factor 5 as a product of prime ideals in O K , where K =
Q( 2). Set θ = 3 2. We have seen in Example 7.1.6 that {1, θ, θ 2 } is an integral
3
5 = P Q,
where
We have
P Q = 5, θ + 25, θ 2 + 3θ + 4
= 25, 5(θ + 2), 5(θ 2 + 3θ + 4), θ 3 + 5θ 2 + 10θ + 8
= 25, 5(θ + 2), 5(θ 2 + 3θ + 4), 5θ 2 + 10θ + 10
= 55, θ + 2, θ 2 + 3θ + 4, θ 2 + 2θ + 2
= 5
as
2 = P Q 2 ,
where
are distinct prime ideals with N (P) = N (Q) = 2. In fact P and Q are both prin-
cipal ideals as we now show. From
θ 3 − 9θ − 6 = 0,
we deduce that
so that
θ + 1 | 2.
Hence
Q = 2, θ + 1 = θ + 1
CB609-10 CB609/Alaca & Williams August 7, 2003 16:53 Char Count= 0
and
P = 2Q −2
2
=
(θ + 1)2
6(θ + 1) + 3(θ + 1)2 − (θ + 1)3
=
(θ + 1)2
2
= 3 + 3 − (θ + 1)
θ +1
= 3(6 + 3(θ + 1) − (θ + 1)2 ) + 2 − θ
= 26 + 2θ − 3θ 2 .
as θ 3 = 9θ + 6 and θ 4 = 9θ 2 + 6θ.
Turning to the prime 3 we have
x 3 − 9x − 6 ≡ x 3 (mod 3),
3 = R 3 ,
so that
R = θP −1
= θ Q 2 (P Q 2 )−1
θ (θ + 1)2
=
2
θ 3 + 2θ 2 + θ 2θ 2 + 10θ + 6
= =
2 2
= 3 + 5θ + θ 2 .
CB609-10 CB609/Alaca & Williams August 7, 2003 16:53 Char Count= 0
To complete the verification of R 3 = 3 we must show that 721 + 1299θ + 393θ 2
is a unit of O K = Z + Zθ + Zθ 2 so that
so that 721 + 1299θ + 393θ 2 is a unit. Alternatively, we could have used Theorem
9.2.4.
N (θ + 1) = (θ + 1)(θ
+ 1)(θ
+ 1)
= θ θ
θ
+ (θθ
+ θ
θ
+ θ
θ) + (θ + θ
+ θ
) + 1
=6−9+0+1
= −2,
CB609-10 CB609/Alaca & Williams August 7, 2003 16:53 Char Count= 0
so that
107 = P 2 Q,
and
θ = −2 + α − α 2 .
Hence
θ + θ2
O K = Z + Zθ + Z = Z + Z(−2 + α − α 2 ) + Zα = Z + Zα + Zα 2 ,
2
proving that {1, α, α 2 } is an integral basis for K . This basis is clearly a power
basis, so K is monogenic.
As K is monogenic we can apply Theorem 10.3.1 to factor the prime 2 in O K .
By Example 7.3.2 we know that α = (θ + θ 2 )/2 is a root of x 3 − x 2 + 3x − 2 = 0.
Thus K = Q(α), where α 3 − α 2 + 3α − 2 = 0. Now
2 = P1 Q 1 ,
If we had in error applied Theorem 10.3.1 directly to the prime 2, we would have
obtained the incorrect factorization
2 = 2, θ 2, 1 + θ2 ,
as
x 3 − x + 4 ≡ x(x + 1)2 (mod 2),
showing that the condition p ind(θ) is essential in Theorem 10.5.1.
We refer the reader to Mann’s book [6] for a proof of this theorem.
Exercises 261
Theorem 10.6.1
2 = P1 P2 ,
where P1 and P2 are distinct prime ideals with
N (P1 ) = N (P2 ) = 23 .
By Theorem 7.5.2 the cyclotomic field K 7 is monogenic so we can apply Theorem
10.3.1 to obtain P1 and P2 explicitly. We have
x7 − 1
irrQ (ζ7 ) = = x6 + x5 + x4 + x3 + x2 + x + 1
x −1
and
x 6 + x 5 + x 4 + x 3 + x 2 + x + 1 ≡ (x 3 + x + 1)(x 3 + x 2 + 1) (mod 2),
so that
P1 = 2, 1 + ζ7 + ζ73 , P2 = 2, 1 + ζ72 + ζ73 .
Exercises
1. In Example 10.4.2 show that 721 + 1299θ + 393θ 2 is a unit by finding its norm.
2. Factor 2 into prime ideals in OQ(√47) .
3. Factor 6 into prime ideals in OQ(√366) .
4. Factor 2 into prime ideals in OQ( √3 2) .
5. Factor 2 into prime ideals in OQ( √3 3) .
6. Factor 2 into prime ideals in OQ(√2+√−1) .
7. Prove√(10.1.2) and (10.1.3).
8. Is Q( 3 10) monogenic?
9. Modify the proof
√ of Theorem 10.3.1 to prove Theorem 10.5.1. √
10. Let K = Q( 3). In O K we have 3 = P 3 , where P = 3 3 is a prime ideal of norm
3
3. Are there any rational primes p = 3 such that p = Q 3 in O K for some prime ideal
Q? √
11. Determine all rational primes p that ramify in Q( 3 6) together with their prime ideal
factorizations. √ √
12. Determine the prime ideal decomposition of the prime 47 in Q( 2, 3).
13. Let K = Q(θ), where θ 3 − θ + 4 = 0. The ideal I = 2, θ is principal in O K . Find a
generator of I .
14. Factor 5 into prime ideals in O K 5 .
15. Factor 3 into prime ideals in O K 7 .
16. As ζm is a unit of OQ(ζm ) , we know that N (ζm ) = ±1. Show that the + sign holds.
17. Prove that 1 + ζm + ζm2 + · · · + ζmk−1 is a unit of OQ(ζm ) if k is a positive integer coprime
with m.
18. Let K 1 and K 2 be algebraic number fields. Suppose that the prime p is totally ramified
in O K 1 and unramified in O K 2 . Prove that K 1 ∩ K 2 = Q.
CB609-10 CB609/Alaca & Williams August 7, 2003 16:53 Char Count= 0
21. If [Q(θ1 ) : Q] and [Q(θ2 ) : Q] are coprime, prove that equality holds in the inequality
in Exercise 20.
22. Let p be an odd prime. Prove that
j −j
(1 − ζ p )(1 − ζ p )
, j = 1, 2, . . . , p − 1,
(1 − ζ p )(1 − ζ p−1 )
are real units of OQ(ζ p ) .
23. Let
α = 1 + ζ23
2
+ ζ23
4
+ ζ23
5
+ ζ23
6
+ ζ23
10
+ ζ23
11
and
β = 1 + ζ23 + ζ23
5
+ ζ23
6
+ ζ23
7
+ ζ23
9
+ ζ23
11
.
Suggested Reading
1. G. Bachman, The decomposition of a rational prime ideal in cyclotomic fields, American
Mathematical Monthly 73 (1966), 494–497.
An alternate proof of the way a rational prime ideal decomposes in the ring of integers of a
cyclotomic field is given
2. Z. I. Borevich and I. R. Shafarevich, Number Theory, Academic Press, New York and
London, 1966.
Example 10.4.2 is based upon Example 2, p. 230.
3. R. Dedekind, Über den Zusammenhang zwischen der Theorie der Ideale und der Theorie
der höheren Kongruenzen, Abh. Kgl. Ges. Wiss. Göttingen 23 (1878), 1–23. (Gesam-
melte Mathematische Werke I, pp. 202–232, Vieweg, Wiesbaden, 1930.)
Theorem 10.3.1 is Theorem 1 on pages 212 and 213 of Dedekind’s Collected Papers.
4. I. Gerst and J. Brillhart, On the prime divisors of polynomials, American Mathematical
Monthly 78 (1971), 250–260.
The set of all primes for which an irreducible polynomial has at least one linear factor (mod p) is
considered.
5. J. G. Huard, B. K. Spearman, and K. S. Williams, The primes for which an abelian cubic
polynomial splits, Tokyo Journal of Mathematics 17 (1994), 467–478.
Let x 3 + ax + b ∈ Z[x] be an irreducible abelian cubic polynomial. Explicit integers a1 , . . . , an , m
are determined such that x 3 + ax + b ≡ 0 (mod p) has three solutions ⇐⇒ p ≡ a1 , . . . , an
(mod m) except for finitely many primes p.
CB609-10 CB609/Alaca & Williams August 7, 2003 16:53 Char Count= 0
11
Units in Real Quadratic Fields
√
11.1 The Units of Z + Z 2
In Theorem 5.4.3 we determined the unit group U (O K ) for an imaginary quadratic
field K . The objective of this chapter is to determine the structure of the unit group
U (O K ) for an arbitrary real quadratic field K . We show that
U (O K ) Z2 × Z
(see Theorems 11.5.1 and 11.5.2). This is accomplished by showing that there exists
a unit in O K such that every unit is of the form ± n (n ∈ Z). We show further
that there exists a unique unit > 1 of O K with this property. This unit is called the
fundamental unit of O K (or of K ). In Section 11.6 we show how continued fractions
can be used to determine the fundamental unit. In Chapter 13 we prove Dirichlet’s
unit theorem, which gives the structure of U (O K ) for an arbitrary algebraic number
field K .
To illustrate some of the ideas that will be involved, we begin by determining
√
U (OQ(√2) ) = U (Z + Z 2).
√ √
Theorem 11.1.1 All the units of Z + Z 2 are given by ±(1 + 2)n (n ∈ Z), so
that
√
U (Z + Z 2) Z2 × Z.
√
Proof: We begin by showing that there does not exist a unit λ of Z + Z 2 satisfying
√
1 < λ < 1 + 2. (11.1.1)
Suppose on the contrary that such a unit λ exists having property (11.1.1).
√ By
Theorem 6.2.1 there are exactly two monomorphisms σ1 and σ2 : Q( 2) → C.
These monomorphisms are given by
√ √ √ √
σ1 (x + y 2) = x + y 2, σ2 (x + y 2) = x − y 2,
264
CB609-11 CB609/Alaca & Williams August 7, 2003 16:58 Char Count= 0
√
11.1 The Units of Z + Z 2 265
for all x, y ∈ Q. As λ is a unit we have λ | 1 so that
√
1 = λµ for some µ ∈ Z + Z 2. (11.1.2)
Hence
1 = (λλ )(µµ ).
λλ = ±1.
and thus
1 λ + λ 1
0.7 < √ < < 1 + √ < 1.8.
2 2 2
As (λ + λ )/2 ∈ Z we must have (λ + λ )/2 = 1. From λλ = 1 and λ + λ = 2
we deduce that λ = λ = 1, contradicting λ > 1.
so that
0 < λ + λ < 2;
that is,
λ + λ
0< < 1.
2
This is a contradiction as (λ + λ )/2 ∈ Z. √
This
√ completes the proof that there are no units of Z + Z 2 between 1 and
1 + 2.
CB609-11 CB609/Alaca & Williams August 7, 2003 16:58 Char Count= 0
Thus
√ −n √
1 ≤ η(1 + 2) < 1 + 2.
√ √
As η(1 + −2)−n is a unit of Z + Z 2, we have
√
η = (1 + 2)n , n ∈ N. (11.1.3)
If η is a unit with 0 < η < 1 then 1/η is a unit with 1/η > 1. Hence, from (11.1.3)
we have
1 √
= (1 + 2)n
η
for some n ∈ N, so that
√
η = (1 + 2)−n , n ∈ N.
If η is a unit with −1 < η < 0 then −1/η is a unit with −1/η > 1. Hence, by
(11.1.3), there exists n ∈ N such that
−1 √
= (1 + 2)n ,
η
so that
√ −n
η = −(1 + 2) , n ∈ N.
so that
√ n
η = −(1 + 2) , n ∈ N.
Clearly
√
±1 = ±(1 + 2)0 .
Theorem 11.2.1 Let m be a positive integer that is not a perfect square. Then there
exist integers x and y with (x, y) = (±1, 0) such that
x 2 − my 2 = 1.
Proof: Let N be a positive integer. We show first that there exist integers x and y
such that
√ 1
0 < |x − y m| < , 0 < y ≤ N . (11.2.1)
N
We divide the interval 0 < x ≤ 1 into N subintervals r/N < x ≤ (r + 1)/N , r =
0, 1, . . . , N − 1, each of the same length 1/N . For i = 0, 1, . . . , N we define the
integers xi and yi by
√
xi = [i m] + 1, yi = i.
Now
√ √ √
[i m] ≤ i m < [i m] + 1
so that
√
xi − 1 ≤ yi m < xi ;
that is,
√
0 < xi − yi m ≤ 1, i = 0, 1, . . . , N .
√
Thus we have N + 1 numbers xi − yi m lying in the interval 0 < x ≤ 1. Hence
at least two of these numbers lie in the same subinterval (r/N , (r + 1)/N ]; that is,
CB609-11 CB609/Alaca & Williams August 7, 2003 16:58 Char Count= 0
and
√
r/N < x j − y j m ≤ (r + 1)/N .
so that
√ √ 1
0 ≤ (xi − yi m) − (x j − y j m) < .
N
We note that yi − y j = i − j = 0. We define the integers x and y by
(xi − x j , yi − y j ), if yi − y j > 0,
(x, y) =
(x j − xi , y j − yi ), if yi − y j < 0.
Thus
−1 √ 1
< x − y m < , y > 0,
N N
and
y = |y| = |yi − y j | = |i − j| ≤ N .
Let N1 be any positive integer. By (11.2.1) there exist integers x1 and y1 such that
√ 1
0 < |x1 − y1 m| < , 0 < y1 ≤ N1 .
N1
√
Now let N2 be any positive integer > 1/|x1 − y1 m|. By (11.2.1) there exist inte-
gers x2 and y2 such that
√ 1
0 < |x2 − y2 m| < , 0 < y2 ≤ N2 .
N2
Nr , xr , yr (r = 1, 2, . . . , k − 1), we choose
Continuing in this way, after obtaining √
Nk to be any integer > 1/|xk−1 − yk−1 m| and integers xk and yk (> 0) such that
√ 1
0 < |xk − yk m| < , 0 < yk ≤ Nk .
Nk
CB609-11 CB609/Alaca & Williams August 7, 2003 16:58 Char Count= 0
Since there are |t| residue classes modulo t, we can find two such solutions, say
(x1 , y1 ) and (x2 , y2 ), such that
u 1 x1 ≡ u 2 x2 (mod t).
Then we have
√ √ √
x1 + m y1 (x1 + m y1 )(x2 − m y2 )
√ =
x2 + m y2 x22 − my22
√
(x1 x2 − my1 y2 ) + m(x2 y1 − x1 y2 )
=
√ t
= x + m y,
where
Now (u 1 u 2 , t) = 1 so that
x2 y1 − y2 x1 ≡ 0 (mod t),
so that as (u 1 u 2 , t) = 1 we have
and so
so that
x 2 − my 2 = 1.
so that
(x1 , y1 ) = ±(x2 , y2 ).
But x1 > 0 and x2 > 0, so (x1 , y1 ) = (x2 , y2 ), contradicting that (x1 , y1 ) and (x2 , y2 )
are distinct solutions of (11.2.3).
Hence we have shown the existence of a pair of integers (x, y) = (±1, 0) such
that x 2 − my 2 = 1.
Theorem 11.3.1 Let m be a positive squarefree integer. Let x and y both be integers
or both halves of odd integers such that x 2 − my 2 = 1. Then
√
x + y m > 1 ⇐⇒ x > 0, y > 0, (11.3.1)
√
0 < x + y m < 1 ⇐⇒ x > 0, y < 0, (11.3.2)
√
−1 < x + y m < 0 ⇐⇒ x < 0, y > 0, (11.3.3)
√
x + y m < −1 ⇐⇒ x < 0, y < 0. (11.3.4)
CB609-11 CB609/Alaca & Williams August 7, 2003 16:58 Char Count= 0
and
x y
Sm = {( , ) | x ∈ N, y ∈ N, x ≡ y (mod 2)}, if m ≡ 1 (mod 4).
2 2
Let (a, b) ∈ Sm be the solution of a 2 − mb2 = 1 for which a is least. (Theorem
√
11.2.1 guarantees that (a, b) exists.) Let = a + b m so that is a unit of OQ(√m)
of norm 1. The unit is called the fundamental unit of norm 1 of OQ(√m) .
CB609-11 CB609/Alaca & Williams August 7, 2003 16:58 Char Count= 0
We note that
√ √
≥ 1 + m ≥ 1 + 2, if m ≡ 2, 3 (mod 4),
√ √
1+ m 1+ 5
≥ ≥ , if m ≡ 1 (mod 4),
2 2
so that
> 1.
Our next theorem shows how the units of norm 1 in OQ(√m) are related to the
fundamental unit of norm 1.
Suppose that 1 is a unit of OQ(√m) of norm 1 with 1 < 1 < . Then, by Theorems
5.4.2 and 11.3.1, we have
√
1 = a1 + b1 m, (a1 , b1 ) ∈ Sm , a12 − mb12 = 1.
a < a1
so that
a2 − 1 a2 − 1
b2 = < 1 = b12 ,
m m
and thus
b < b1 .
Hence
√ √
= a + b m < a1 + b1 m = 1 ,
contradicting 1 < . Thus no such unit 1 exists, proving that is the smallest unit
of OQ(√m) of norm 1 that is greater than 1.
CB609-11 CB609/Alaca & Williams August 7, 2003 16:58 Char Count= 0
(b) Let η be a unit of OQ(√m) of norm 1. Let η∗ be the unit of OQ(√m) of norm 1
defined by
η, if η ≥ 1,
1/η, if 0 < η < 1,
η∗ = (11.3.5)
−1/η, if − 1 < η < 0,
−η, if η ≤ −1,
so that
η∗ ≥ 1.
Let k be the unique nonnegative integer such that
k ≤ η∗ < k+1 .
Then η∗ −k is a unit of OQ(√m) of norm 1 satisfying
1 ≤ η∗ −k < .
By part (a) there is no unit in OQ(√m) of norm 1 strictly between 1 and . Hence
η∗ −k = 1
and so
η∗ = k .
Then, from (11.3.5), we obtain
η = ± n
for some choice of sign and some integer n.
(c) By assumption we have
= ±τ l ,
for some integer l, and by part (b) we have
τ = ± n ,
for some integer n. Hence
= ±(± n )l = ± ln
so that
ln−1 = ±1
and thus
2(ln−1) = 1.
CB609-11 CB609/Alaca & Williams August 7, 2003 16:58 Char Count= 0
Theorem 11.4.1 Let m be a positive squarefree integer. Suppose that OQ(√m) con-
tains units of norm −1. Then there exists a unique unit σ > 1 of norm −1 in OQ(√m)
such that every unit in OQ(√m) is of the form ±σ n for some integer n.
Proof: Let ρ be a unit in OQ(√m) of norm −1. Let ρ denote its conjugate. Then
ρρ = N (ρ) = −1
so that
ρ 2 ρ = 1.
2
so that
ρ = ± k .
Hence
N (ρ) = N (± k ) = N ()k = 1,
contradicting N (ρ) = −1. Thus n must be odd, say n = 2l + 1, and so
ρ 2 = 2l+1 .
Hence
= (ρ −l )2 .
Set
σ = ρ −l
so that σ is a unit of norm −1 such that
= σ 2.
If µ is a unit of OQ(√m) of norm −1 then µρ −1 is a unit of OQ(√m) of norm 1 and
thus by Theorem 11.3.2(b)
µρ −1 = ± k
for some k ∈ Z. Hence, as ρ = l σ and = σ 2 , we deduce that
µ = ± k ρ = ± k+l σ = ±σ 2(k+l)+1 .
However, if µ is a unit of OQ(√m) of norm 1 then by Theorem 11.3.2(b)
µ = ± k
for some k ∈ Z. Hence, as = σ 2 , we deduce that
µ = ± k = ±σ 2k .
Thus every unit of OQ(√m) is of the form
±σ n (n ∈ Z).
Note that n even gives the units of norm 1 and n odd the units of norm −1.
Replacing σ by 1/σ if 0 < σ < 1, by −1/σ if −1 < σ < 0, and by −σ if
σ < −1, we may suppose that σ > 1. We show that σ is uniquely determined: For
suppose σ and τ are two units > 1 of norm −1 in OQ(√m) such that every unit is
of the form ±σ n (n ∈ Z) and of the form ±τ q (q ∈ Z). Then there exist integers k
and l such that
σ = ±τ k , τ = ±σ l .
CB609-11 CB609/Alaca & Williams August 7, 2003 16:58 Char Count= 0
Hence
σ = ±σ kl
and so
σ 2 = σ 2kl ,
giving
σ 2(kl−1) = 1.
Suppose kl − 1 = 0. Then σ is a root of unity. But OQ(√m) being a real field contains
no roots of unity except ±1. Thus σ = ±1, a contradiction. Hence kl − 1 = 0 and
so
k = l = ±1.
Thus
σ = ±τ or ± τ −1 .
But σ > 1 and τ > 1 so
σ = τ,
proving that σ is unique.
We next relate the fundamental unit of norm 1 and the fundamental unit σ of
norm −1 when OQ(√m) contains units of norm −1.
Theorem 11.4.2 Let m be a positive squarefree integer such that OQ(√m) contains
units of norm −1. Then the fundamental unit of norm 1 and the fundamental unit
σ of norm −1 are related by
= σ 2.
Theorem 11.5.1 Let m be a positive squarefree integer. Then every unit of OQ(√m)
is of the form ±ηn (n ∈ Z), where η is the fundamental unit of OQ(√m) . If OQ(√m)
contains units of norm −1 these are given by ±ηn with n odd and the ones of norm
1 by ±ηn with n even.
U (O K ) Z2 × Z.
Proof: Let η be the fundamental unit of O K and suppose that there exists a unit θ
of O K with
1 < θ < η.
θ = ±ηn
for some n ∈ Z. As θ and η are both positive, the positive sign must hold and we
have
θ = ηn .
If n ≥ 1 then
θ = ηn ≥ η,
θ = ηn ≤ 1,
contradicting θ > 1. Hence no such θ can exist, proving that η is the smallest unit
greater than 1.
Before proceeding to find the norm of the fundamental unit η of OQ(√m) for
certain special values of m, we present in Table 4 the values of , σ , and η for
squarefree positive integers m < 40.
We next determine the norm of the fundamental unit of OQ(√m) when m is an
odd prime p. First we consider the case p ≡ 1 (mod 4).
Theorem 11.5.4 Let p be a prime with p ≡ 1 (mod 4). Then the fundamental unit
of OQ(√ p) has norm −1.
We give two proofs of this theorem, the first due to Hilbert and the second due
to Peter Gustav Lejeune Dirichlet (1805–1859).
CB609-11 CB609/Alaca & Williams August 7, 2003 16:58 Char Count= 0
First proof: Suppose that the fundamental unit η of OQ(√ p) has norm 1. Then
N (η) = ηη = 1.
1+η
η= .
1 + η
m | 1 + η, m | 1 + η
CB609-11 CB609/Alaca & Williams August 7, 2003 16:58 Char Count= 0
in OQ(√ p) . Set
1+η
γ = ∈ OQ(√ p)
m
and
1 + η
γ = ∈ OQ(√ p)
m
so that
γ
η=
γ
and
k | γ , k | γ , k ∈ N =⇒ k = 1. (11.5.1)
Now
γ = ηγ ,
γ = γ . (11.5.2)
Q | γ . (11.5.3)
Q | γ . (11.5.4)
Q | γ . (11.5.5)
In the first case, from (11.5.3) and (11.5.4), we deduce that q | γ , q | γ , con-
tradicting (11.5.1).
In the second case we have by (11.5.3) and (11.5.5) as Q and Q are distinct
prime ideals
q = Q Q | γ = γ .
γ = 1
and so
γ =λ
γ λ λ2
η= = = = ±λ2 ,
γ λ λλ
contradicting that η is the fundamental unit of OQ(√ p) . If j = 1 then
√
γ = p
so that
√
γ =λ p
Second proof: Suppose that the fundamental unit η of OQ(√ p) has norm 1. Then
√
x+y p
η= ,
2
CB609-11 CB609/Alaca & Williams August 7, 2003 16:58 Char Count= 0
Proof: Suppose that the fundamental unit η of OQ(√m) has norm −1. Then
√
x+y m
η= ,
2
where x and y are integers such that
x ≡ y ≡ 0 (mod 2), if m ≡ 2, 3 (mod 4),
x ≡ y (mod 2), if m ≡ 1 (mod 4),
and
x 2 − my 2
= N (η) = −1.
4
Hence
x 2 − my 2 = −4.
As q | m we deduce that
x 2 ≡ − 4 (mod q).
CB609-11 CB609/Alaca & Williams August 7, 2003 16:58 Char Count= 0
Thus
−1 −4
= =1
q q
so that q ≡ 1 (mod 4), contradicting that q ≡ 3 (mod 4). This proves that η must
have norm 1.
The following two theorems can be proved using Dirichlet’s method in a similar
manner to the second proof of Theorem 11.5.4.
Theorem 11.5.6 Let p be a prime with p ≡ 5 (mod 8). Then the fundamental unit
of OQ(√2 p) has norm −1.
Up to this point we have said almost nothing about calculating the fundamental
unit η of OQ(√m) for a particular value of m. We address this problem in the next
section.
and
1
αn+1 = , n = 0, 1, 2, . . . . (11.6.2)
αn − [αn ]
CB609-11 CB609/Alaca & Williams August 7, 2003 16:58 Char Count= 0
an = [αn ], n = 0, 1, 2, . . . , (11.6.4)
CB609-11 CB609/Alaca & Williams August 7, 2003 16:58 Char Count= 0
so that {an }n≥1 is a sequence of positive integers. Since {αn }n≥1 is a purely periodic
sequence so is the sequence {an }n≥1 .
Next we define two further sequences of integers {h n }n≥−1 and {kn }n≥−1 by
and
{h n }n≥−1 = {1, 5, 6, 11, 39, 206, 657, 863, 1520, 16063, 17583, . . .},
{kn }n≥−1 = {0, 1, 1, 2, 7, 37, 118, 155, 273, 2885, 3158, . . .}.
hn 1
= a0 + , n = 0, 1, 2, . . . . (11.6.7)
kn 1
a1 +
. 1
a2 + . . +
1
an−1 +
an
To save space we abbreviate the fraction on the right-hand side of (11.6.7) by the
space-saving flat notation [a0 , a1 , a2 , . . . , an ] so that (11.6.7) becomes
hn
= [a0 , a1 , a2 , . . . , an ], n = 0, 1, 2, . . . . (11.6.8)
kn
√
It is known that limn→∞ h n /kn exists and is equal to m.
CB609-11 CB609/Alaca & Williams August 7, 2003 16:58 Char Count= 0
and so
A3 + 3AB 2 m = 8h l−1 ,
3A2 B + B 3 m = 8kl−1 .
Hence
1/3
A | h l−1 , 1 ≤ A < 2h l−1
and
1/3
kl−1
B | kl−1 , 1 ≤ B < 2 .
m
This gives the following algorithm for determining the fundamental unit η of
OQ(√m) for a positive squarefree integer m.
Step 1: h −1 = 1, k−1 = 0.
√ √
P0 = 0, Q 0 = 1, a0 = [ m], h 0 = [ m], k0 = 1.
Step 3: l = N − 1.
Step 5: If m ≡ 5 (mod 8) determine all positive odd divisors A of h l−1 less than
1/3
2h l−1 and all positive odd divisors B of kl−1 less than 2 (kl−1 /m)1/3 . If for some
pair (A, B) we have
A3 + 3AB 2 m = 8h l−1 , 3A2 B + B 3 m = 8kl−1 ,
then
√
A+B m
η= , N (η) = (−1)l ;
2
CB609-11 CB609/Alaca & Williams August 7, 2003 16:58 Char Count= 0
otherwise
√
η = h l−1 + kl−1 m, N (η) = (−1)l .
We present several examples.
As
P4 = P1 = 6, Q 4 = Q 1 = 5,
we have
N = 4, l = N − 1 = 3, h l−1 = h 2 = 32, kl−1 = k2 = 5,
√ √
η = h l−1 + kl−1 m = 32 + 5 41, N (η) = (−1)l = −1.
√
The fundamental unit of O Q(√41) is 32 + 5 41 of norm −1.
Next,
1/3
A odd, A | h l−1 , 1 ≤ A < 2h l−1 =⇒ A | 9, 1 ≤ A < 5.3 =⇒ A = 1 or 3,
1/3
kl−1
B odd, B | kl−1 , 1 ≤ B < 2 =⇒ B | 5, 1 ≤ B < 1.5 =⇒ B = 1.
m
Of the pairs (A, B) = (1, 1), (3, 1) only the latter satisfies the pair of equations
A3 + 39AB 2 = 144, 3A2 B + 13B 3 = 40.
Hence the fundamental unit η(>1) of O Q(√13) is
√
3 + 13
η= , N (η) = −1.
2
Theorem 11.7.1 Let m be a positive nonsquare integer. Let {h n }n≥−1 and {kn }n≥−1
be defined as in (11.6.5) and (11.6.6). Let gn = h 2n − mkn2 , n = −1, 0, 1, . . . . Let
CB609-11 CB609/Alaca & Williams August 7, 2003 16:58 Char Count= 0
n hn kn gn = h 2n − 31kn2
0 5 1 −6
1 6 1 5
2 11 2 −3
3 39 7 2
4 206 37 −3
5 657 118 5
6 863 155 −6
7 1520 273 1
√
By Theorem 11.7.1 the equation x 2 − 31y 2 = N (0 < |N | < 31) is solvable in
coprime integers x and y for
N = −3, 1, 2, 5
It is easily checked that these are solutions of x 2 − 31y 2 = −3. In particular the
solution (x, y) = (206, 37) is given by
√ √
−(1520 + 273 31)(11 − 2 31).
Example 11.7.3 We answer the question “Is the equation x 2 − 41y 2 = 2 solvable
in integers√x and y?”
As 2 < 41 we can apply Theorem 11.7.1. From Example 11.6.2 we have l = 3,
so that s = 2 and sl − 1 = 5, and the values of gn (n = 0, 1, 2, 3, 4, 5) are as
follows:
n hn kn gn
0 6 1 −5
1 13 2 5
2 32 5 −1
3 397 62 5
4 826 129 −5
5 2049 320 1
Example 11.7.4 We determine all the solutions in integers x and y of the equation
x 2 − 10y 2 = 10. In this case we cannot apply Theorem 11.7.1 directly as m = N =
√
10 and |N | < m. Thus we proceed differently.
Let (x, y) ∈ Z2 be a solution of x 2 − 10y 2 = 10. Then 10 | x 2 and as 10 is
squarefree we deduce that 10 | x. Setting x = 10z in the equation we obtain
CB609-11 CB609/Alaca & Williams August 7, 2003 16:58 Char Count= 0
Exercises 297
√
y 2 − 10z 2 = −1. Thus y + z 10 is a unit of OQ(√10) of norm −1. As the fun-
√
damental unit of OQ(√10) is 3 + 10 (of norm −1), we have by Theorem 11.4.1
√ √
y + z 10 = ±(3 + 10)2n+1
Exercises
1. Prove Theorem 11.5.6.
2. Prove Theorem 11.5.7.
3. Determine the fundamental unit of OQ(√11) .
4. Determine the fundamental unit of OQ(√17) .
√
5. Prove that the fundamental unit of OQ(√94) is η = 2143295 + 221064 94.
√
6. Prove that the fundamental unit of OQ(√163) is η = 64080026 + 5019135 163.
√
7. Prove that the fundamental unit of OQ(√165) is η = (13 + 165)/2.
√
8. Show that √1790 = [42, 3, 4, 8, 4, 3, 84].
9. Show that 925 = [30, 2, 2, 2, 2, 60]. √
10. Determine the length of the continued fraction expansion of 850.
11. Determine the norm of the fundamental unit of OQ(√1378) .
12. Let m be a positive squarefree integer such that OQ(√m) contains units of norm −1. Let
σ be the fundamental unit of OQ(√m) of norm −1. Prove that σ is the smallest unit > 1
of norm −1 in OQ(√m) .
13. Let η be the fundamental unit of OQ(√134) . Determine α ∈ OQ(√134) such that 2η = α 2 .
√
14. Let p be a prime ≡ 3 (mod 8). Let t + u p be the fundamental unit of OQ(√ p) , which
necessarily is of norm 1. Starting from t 2 − pu 2 = 1, and using Dirichlet’s method of
proving Theorem 11.5.4, prove that the equation x 2 − py 2 = −2 is solvable in integers
x and y.
15. Let p be a prime ≡ 7 (mod 8). Prove that the equation x 2 − py 2 = 2 is solvable in
integers x and y.
16. Let p be a prime ≡ 9 (mod 16) for which the congruence x 4 ≡ 2 (mod p) is insolvable.
Prove that the norm of the fundamental unit of OQ(√2 p) is −1.
17. Let p and q be distinct primes with p ≡ q ≡ 1 (mod 4) and qp = 1 (so that qp = 1
by the law of quadratic reciprocity). Suppose that the congruences x 4 ≡ p (mod q) and
y 4 ≡ q (mod p) are insolvable. Prove that the norm of the fundamental unit of OQ(√ pq)
is −1.
18. Is the equation x 2 − 82y 2 = 2 solvable in integers x and y?
19. Determine all solutions of x 2 − 96y 2 = 161.
CB609-11 CB609/Alaca & Williams August 7, 2003 16:58 Char Count= 0
20. Prove that all solutions of x 2 − 10y 2 = 10 (see Example 11.7.1) are given recursively
by ±(xk , yk ), where
xk+1 = 19xk + 60yk , yk+1 = 6xk + 19yk , k = 0, ±1, ±2, . . . ,
and x0 = 10, y0 = 3.
21. Let m be a positive integer such that m − 1 and m are not perfect squares but 4m + 1
is a perfect square. Prove that the equation x 2 − my 2 = −1 is insolvable in integers x
and y.
Suggested Reading
1. H. W. Lenstra Jr., Solving the Pell equation, Notices of the American Mathematical
Society 49 (2002), 182–192.
This up-to-date article describes the use of smooth numbers to solve Pell’s equation x 2 − my 2 = 1.
2. I. Niven, H. S. Zuckerman, and H. L. Montgomery, An Introduction to the Theory of
Numbers, fifth edition, Wiley, New York, 1991.
Chapter 7 contains a comprehensive treatment of continued fractions.
3. W. Patz, Tafel der Regelmässigen Kettenbrüche und ihrer Vollständigen Quotienten für
die Quadratwurzeln aus der Natürlichen Zahlen von 1–10000, Akademie-Verlag, Berlin,
1955.
√
This book gives the continued fraction expansions of m for all nonsquares m up to 10,000.
Biographies
1. E. T. Bell, Men of Mathematics, Simon and Schuster, New York, 1937.
Chapters 9 and 10 are devoted to Leonhard Euler (1707–1783) and Joseph-Louis Lagrange (1736–
1813) respectively.
2. J. J. Burckhardt, Leonhard Euler, 1707–1783, Mathematics Magazine 56 (1983), 262–
273.
A brief overview of the work of Euler is given.
3. H. Davenport, Dirichlet, Mathematical Gazette 43 (1959), 268–269.
A short biography of Dirichlet is given.
4. H. Koch, Gustav Peter Lejeune Dirichlet, in H. G. W. Begehr, H. Koch, J. Kramer,
N. Schappacher, and E. J. Thiele (Eds.), Mathematics in Berlin, 33–40, Birkhäuser
Verlag, Berlin, 1998.
The book in which this biography of Dirichlet is included describes the many facets of Berlin’s role
in mathematics. It includes biographies of many famous mathematicians connected with Berlin.
5. The website
http://www-groups.dcs.st-and.ac.uk/˜history/
12
The Ideal Class Group
Definition 12.1.1 (Ideal class group) Let K be an algebraic number field. Let
I (K ) be the group of nonzero fractional and integral ideals of O K . Let P(K ) be the
subgroup of principal ideals of I (K ). Then the factor group I (K )/P(K ) is called
the ideal class group of K and is denoted by H (K ).
Definition 12.1.2 (Class number) Let K be an algebraic number field. The order
of the ideal class group H (K ) is called the class number of K and is denoted by
h(K ).
299
CB609-12 CB609/Alaca & Williams August 7, 2003 17:0 Char Count= 0
as
√ √
32, 1 − −5 = 6, 3(1 − −5)
√ √ √
= (1 + −5)(1 − −5), 3(1 − −5)
√ √
= 1 − −51 + −5, 3.
Leonard Carlitz (1907–1999) has shown that h(K ) = 1 or 2 if and only if when-
ever a nonzero nonunit α ∈ O K can be written α = uπ1 · · · πs = u π1 · · · πt with
u, u units and π1 , . . . , πs , π1 , . . . , πt prime elements of O K then s = t [2].
In the next three sections we prove three theorems of Minkowski from which we
can deduce that the class number is always finite.
Zn = {(x1 , . . . , xn ) ∈ Rn | x1 , . . . , xn ∈ Z}.
The elements of Zn are called lattice points and Zn is called a lattice. Clearly Zn is
a group under addition. For α = (a1 , . . . , an ) ∈ Rn we set
Sα = {α + β | β ∈ S}.
aS = {aβ | β ∈ S}.
Clearly S is convex if it contains the line segment joining β and γ for all points
β and γ in S.
The volume V (S) of a convex body S is defined by means of the multiple integral
V (S) = ··· d x1 · · · d xn .
S
It is easy to check that the hypercube Ht is a convex body. Its volume is given by
t t t n
V (Ht ) = ··· dβ1 · · · dβn = dβ = (2t)n .
β1 =−t βn =−t β=−t
Proof: We first treat the case V (S) > 1. Suppose on the contrary that
S ∩ Sα = φ for all α ∈ Zn \ {0}.
First we prove that
Sβ ∩ Sγ = φ for all β, γ ∈ Zn , β = γ . (12.2.1)
Let x ∈ Sβ ∩ Sγ . Then x ∈ Sβ and x ∈ Sγ . Hence x − γ ∈ Sβ−γ and x − γ ∈ S.
Thus x − γ ∈ Sβ−γ ∩ S, contradicting our assumption. This proves (12.2.1).
Now let N be an arbitrary positive integer, and let
T = {α ∈ Zn | ||α|| ≤ N }.
Clearly
card T = (2N + 1)n .
Our second step is to show that
V Sα = (2N + 1)n V (S). (12.2.2)
α∈T
CB609-12 CB609/Alaca & Williams August 7, 2003 17:0 Char Count= 0
We have
V Sα = V (Sα ) (by (12.2.1))
α∈T α∈T
= V (S)
α∈T
= V (S)card T
= (2N + 1)n V (S),
as asserted.
As S is a bounded set, we can define the diameter d ∈ R+ of S by
d = max ||s1 − s2 ||.
s1 ,s2 ∈S
Hence
||β|| = ||α + s|| ≤ ||α|| + ||s|| ≤ N + d
so that β ∈ H . Hence Sα ⊆ H for α ∈ T .
We are now in a position to complete the proof. From (12.2.3) we have
Sα ⊆ H
α∈T
so that
V Sα ≤ V (H )
α∈T
and thus
(2N + 1)n V (S) ≤ (2N + 2d)n .
CB609-12 CB609/Alaca & Williams August 7, 2003 17:0 Char Count= 0
Therefore
d n
n 1+
2N + 2d N
V (S) ≤ = .
2N + 1 1
1+
2N
As d and n are fixed, letting N → ∞ we obtain V (S) ≤ 1, contradicting V (S) > 1.
Hence there is at least one α ∈ Zn \ {0} with S ∩ Sα = φ.
We now turn to the case V (S) = 1. Let k be a positive integer. We consider the
convex body
1
S(k) = 1 + S
k
obtained by magnifying the convex body S by a factor 1 + 1/k. As S contains 0 so
does S(k). The volume of S(k) satisfies
1 1 n 1 n
V (S(k)) = V 1+ S = 1+ V (S) = 1 + > 1.
k k k
Thus the first part of the theorem applies to the convex body S(k). Hence there
exists a translate (S(k))αk (αk ∈ Zn \ {0}) of S(k) such that
S(k) ∩ (S(k))αk = φ.
xk ∈ S(k) ⊆ S(1)
and
βk ∈ H2d1 ,
CB609-12 CB609/Alaca & Williams August 7, 2003 17:0 Char Count= 0
so that
Ak ⊆ H2d1 .
But
Ak ⊆ Zn
so
Ak ⊆ H2d1 ∩ Zn .
Thus every Ak is contained in the finite set H2d1 ∩ Zn . Hence we can find a subse-
quence of the sequence {xk } for which the corresponding βk ∈ Ak is constant, say,
equal to α. After relabeling we may therefore assume that
xk ∈ S(k) ∩ (S(k))α .
As each xk ∈ S(1), the infinite sequence {xk } is bounded. Hence, by the Bolzano–
Weierstrass theorem, the sequence {xk } has at least one limit point, say x. Let l be
an arbitrary positive integer. We have
xl ∈ S(l),
xl+1 ∈ S(l + 1) ⊂ S(l)
xl+2 ∈ S(l + 2) ⊂ S(l + 1) ⊂ S(l),
···
so that the infinite sequence {xk }k≥l lies in S(l). As x is a limit point of {xk }k≥l and
S(l) is closed, we deduce that x ∈ S(l) for every positive integer l. Thus
∞ ∞
1
x∈ S(l) = 1+ S = S.
l=1 l=1
l
Similarly,
∞
x∈ (S(l))α = Sα .
l=1
Hence
x ∈ S ∩ Sα
must contain at least one lattice point different from the origin. We begin by making
the notion “symmetrical about the origin” precise.
−α ∈ S for all α ∈ S.
We note that a subset S of Rn , which is both centrally symmetric and convex, must
contain the origin 0 = (0, 0, . . . , 0). To see this take any α ∈ S. As S is centrally
symmetric, we have −α ∈ S. Then, as S is convex, we have 12 (α) + 12 (−α) ∈ S,
that is, 0 ∈ S.
We now use Minkowski’s translate theorem to prove his convex body theorem.
V (T ) = V ( 12 S) = 1
2n
V (S) ≥ 1,
T ∩ Tα = φ.
a jk ∈ R for j = 1, 2, . . . , r ; k = 1, 2, . . . , n, (12.4.1)
CB609-12 CB609/Alaca & Williams August 7, 2003 17:0 Char Count= 0
and
δ j = δ j+s , j = r + 1, . . . , r + s. (12.4.6)
Os,r Os −1 I
2i s
where Os = Os,s . The determinant of a triangular matrix is the product of its diag-
onal entries so that
s
−1 s i
detA = = .
2i 2
As the elementary column operations used to obtain A do not change the value of
the determinant of the matrix, we have
s
∂Mj i
det = detA = .
∂ Lk 2
∂M
The quantity det ∂ L kj is the Jacobian of the M j with respect to the L k so that
s
∂(M1 , . . . , Mn ) i
= .
∂(L 1 , . . . , L n ) 2
Hence
s
∂(M1 , . . . , Mn ) ∂(M1 , . . . , Mn ) ∂(L 1 , . . . , L n ) i
= = det(a jk ),
∂(x1 , . . . , xn ) ∂(L 1 , . . . , L n ) ∂(x1 , . . . , xn ) 2
so that
∂(x1 , . . . , xn ) 2s
=
∂(M , . . . , M ) |det(a )| . (12.4.8)
1 n jk
S = {x ∈ Rn | |L j (x)| ≤ δ j , j = 1, 2, . . . , r + s}.
CB609-12 CB609/Alaca & Williams August 7, 2003 17:0 Char Count= 0
V (S) = ··· d x1 · · · d xr +s .
|L 1 (x)| ≤ δ1
.
.
.
|L r +s (x)| ≤ δr +s .
Now, as
|L r + j (x)| = |L r + j (x)| = |L r +s+ j (x)|, j = 1, 2, . . . , s,
and
δr + j = δr +s+ j , j = 1, 2, . . . , s,
we deduce that
|L r + j (x)| ≤ δr + j ⇐⇒ |L r + j (x)|2 ≤ δr2+ j
⇐⇒ |L r + j (x)L r +s+ j (x)|2 ≤ δr2+ j
⇐⇒ |(Mr + j (x) + i Mr +s+ j (x))(Mr + j (x) − i Mr +s+ j (x))|2 ≤ δr2+ j
⇐⇒ Mr + j (x)2 + Mr +s+ j (x)2 ≤ δr2+ j
for j = 1, 2, . . . , s. Hence
V (S) = ··· d x1 · · · d xn .
|M1 (x)| ≤ δ1
.
.
.
|Mr (x)| ≤ δr
Mr +1 (x)2 + Mr +s+1 (x)2 ≤ δr2+1
.
.
.
Mr +s (x)2 + Mn (x)2 ≤ δr2+s
∂(x1 , . . . , xn )
V (S) = ···
∂(M , . . . , M ) d M1 · · · d Mn .
1 n
|M1 | ≤ δ1
.
.
.
|Mr | ≤ δr
Mr +1 + Mr2+s+1 ≤ δr2+1
2
.
.
.
Mr2+s + Mn2 ≤ δr2+s
CB609-12 CB609/Alaca & Williams August 7, 2003 17:0 Char Count= 0
.
.
.
Mr2+s + Mn2 ≤ δr2+s
Now
δ
d M = 2δ
−δ
and
··· d M d N = area of a circle of radius δ = πδ 2 ,
M 2 + N 2 ≤ δ2
so that
2s r s
V (S) = 2δ j πδr2+k
|det(a jk )| j=1 k=1
2r +s π s δ1 · · · δr (δr +1 · · · δr +s )2
=
|det(a jk )|
2r +s π s δ1 · · · δn
= (by (12.4.6))
|det(a jk )|
s
2r +s π s 2
≥ |det(a jk )| (by (12.4.5))
|det(a jk )| π
= 2r +2s = 2n .
from which the asserted result follows by (12.4.3), (12.4.4), and (12.4.6).
CB609-12 CB609/Alaca & Williams August 7, 2003 17:0 Char Count= 0
n
L j (x) = σ j (αk )xk .
k=1
Let
s/n
2
δj = N (A)1/n |d(K )|1/2n , j = 1, 2, . . . , n.
π
Then
s s
2 2
δ 1 · · · δn = N (A)|d(K )| =
1/2
|det(a jk )|
π π
CB609-12 CB609/Alaca & Williams August 7, 2003 17:0 Char Count= 0
so, by Minkowski’s linear forms theorem (Theorem 12.4.1), there exist integers
y1 , . . . , yn , not all zero, such that
s/n
2
|L j (y)| ≤ N (A)1/n |d(K )|1/2n , j = 1, 2, . . . , n.
π
Choose m ∈ {1, 2, . . . , n} such that σm = 1, where 1 denotes the identity monomor-
phism from K to K . Set
n
n
α = L m (y) = σm (αk )yk = αk yk ,
k=1 k=1
Hence
s/n
2
|σ j (α)| ≤ N (A)1/n |d(K )|1/2n , j = 1, 2, . . . , n,
π
and so
s
2
|N (α)| = |σ1 (α) · · · σn (α)| ≤ N (A)|d(K )|1/2
π
as asserted.
We next establish that there are only finitely many integral ideals in the ring of
integers of an algebraic number field having a given norm.
Hence
k + A = k(1 + A) = 0 + A
so that k ∈ A. Thus k ⊆ A and so A | k. By Theorem 8.3.1 there exist distinct
prime ideals P1 , . . . , Pr and nonnegative integers a1 , . . . , ar such that
A = P1c1 · · · Prcr ,
where
ci ∈ {0, 1, . . . , ai } for i = 1, 2, . . . , r.
possibilities for A.
We now use Theorems 12.5.2 and 12.5.3 to show that the ideal class group of an
algebraic number field is finite.
Theorem 12.5.4 Let K be an algebraic number field. Then the ideal class group
H (K ) of K is a finite group (so that the class number h(K ) = card H (K ) is finite).
Proof: By Theorem
s √12.5.3 there are only finitely many integral ideals B of O K
with N (B) ≤ π2 |d(K )|. It follows from Theorem 12.5.2 that each ideal class
is represented by an integral ideal of O K from a finite set. Thus there are only
finitely many ideal classes. Hence H (K ) is a finite group and h(K ) is finite.
The quantity on the right-hand side of the inequality in Theorem 12.5.2 is called
the Minkowski bound for the number field K .
CB609-12 CB609/Alaca & Williams August 7, 2003 17:0 Char Count= 0
We illustrate this algorithm by finding the ideal class group of several algebraic
number fields.
We denote the class containing the ideal A by [A] and the class of principal ideals
by 1.
√
Example 12.6.1 We show that K = Q( −19) has class number h(K ) = 1. Here
n = 2, r = 0, s = 1, d(K ) = −19.
the principal ideals 2 and 3 are both prime ideals in O
√K . This is the situation
described just before the algorithm
√ and so h(K ) = h(Q( −19)) = 1. Hence the
1+ −19
√
ring of integers Z + Z 2
of Q( −19) is a principal ideal domain and thus
a unique factorization domain.
√
Example 12.6.2 We show that K = Q( −163) has class number h(K ) = 1. Here
n = 2, r = 0, s = 1, d(K ) = −163.
n = 2, r = 2, s = 0, d(K ) = 92.
Finally, as
92 1
= = 1,
7 7
7 splits in O K . We have
√ √
7 = 7, 3 + 237, 3 − 23.
CB609-12 CB609/Alaca & Williams August 7, 2003 17:0 Char Count= 0
Thus
√
[P][P1 ]2 = [P P12 ] = [2 − −14] = 1.
As [P]2 = [P 2 ] = 1 we deduce that [P] = [P1 ]2 and [P1 ]4 = 1. Also, [P1 ][P2 ] =
[P1 P2 ] = 1 = [P1 ]4 so that [P2 ] = [P1 ]3 . Thus 1, [P1 ], [P1 ]2 , and [P1 ]3 comprise
all the ideal classes. We show that these four ideal classes are in fact distinct. We
√ that [P1 ] = 1. Suppose that [P1 ] √= 1. Then [P] = 1 so that the
2 2
do this by proving
ideal P = 2, −14 is principal, say P = x + y −14, where x, y ∈ Z. Then
√ √
2 = N (2, −14) = N (x + y −14) = x 2 + 14y 2 ,
√
which is impossible. This proves that H (Q( −14)) is a cyclic group of order 4
generated by the class of P1 .
It is often useful when determining the ideal class group of an algebraic number
field K = Q(θ ) to calculate N (k + θ) for k = 0, 1, 2, . . . and use those values that
only involve the primes p ≤ M K to find relations among the ideal classes.
This is illustrated in the next example.
where
√ √
P1 = 2, 1 + −65, Q 1 = 3, 1 + −65,
√ √
Q 2 = 3, 1 − −65, P2 = 5, −65
We have
√ √
4+ −65 = 3 + (1 + −65) ∈ Q 1 ,
so that
√
4 + −65 ⊆ Q 1
and thus
√
Q 1 | 4 + −65.
√
Suppose that Q 2 | 4 + −65; then
√
3 = Q 1 Q 2 | 4 + −65,
√
which is impossible. Hence Q 2 4 + −65. Let r be the unique positive integer
such that
√
Q r1 || 4 + −65.
CB609-12 CB609/Alaca & Williams August 7, 2003 17:0 Char Count= 0
Then
√
4 + −65 = Q r1 B
showing that
[Q 1 ]4 = 1.
[Q 2 ] = [Q 1 ]3 .
We have
√ √
5+ −65 = 2(2) + 1(1 + −65) ∈ P1 ,
√ √
5 + −65 = 2(3) − 1(1 − −65) ∈ Q 2 ,
√ √
5 + −65 = 1(5) + 1( −65) ∈ P2 ,
so that
√ √ √
P1 | 5 + −65, Q 2 | 5 + −65, P2 | 5 + −65.
Hence
√
5 + −65 = P1r Q s2 P2t B,
y. Then
18 = N (P1 Q 21 ) = x 2 + 65y 2 ,
which is impossible. This proves that
√
H (Q( −65)) = {1, [Q 1 ], [Q 1 ]2 , [Q 1 ]3 , [P1 ], [P1 ][Q 1 ], [P1 ][Q 1 ]2 , [P1 ][Q 1 ]3 },
where
[Q 1 ]4 = [P1 ]2 = 1,
CB609-12 CB609/Alaca & Williams August 7, 2003 17:0 Char Count= 0
Table √
5. Nontrivial ideal class groups
H (Q( k)), −30 < k < 0, k squarefree
√
k H (Q( k))
√
−5 {1, A} Z2 , A = [2, 1√+ −5], A2 = 1
−6 {1, A} Z2 , A = [2, √−6], A2 = 1
−10 {1, A} Z2 , A = [2, −10], √ A2 = 1
−13 {1, A} Z2 , A = [2, 1 + −13],√ A2 = 1
−14 {1, A, A2 , √
A3 } Z4 , A = [3, 1 +√ −14],
A = [2, −14], A3 = [3,√1 − −14], A4 = 1
2
−17 {1, A, A , A } √
2 3
Z4 , A = [3, 1 + −17], √
A2 = [2, 1 + −17], A3 = [3, 1 − √ −17], A4 = 1
−21 {1, A, B, √AB} Z2 × Z2 , A = [2, √1 + −21],
B = [3, −21], AB = [5, 3 + −21],
A2 = B 2 = (AB)2 = 1 √
−22 {1, A} Z2 , A = [2, −22], A√2 = 1
−23 {1, A, A2 } Z3 ,√A = [2, 12 (1 + −23)],
A2 = [2, 12 (1 − −23)], A3 = 1 √
−26 {1, A, A2 , A3 , √
A4 , A5 } Z6 , A = √[5, 2 + −26],
A2 = [3, 1 − √−26], A3 = [2, −26], √
A4 = [3, 1 + −26], A5 = [5, 2 − −26], √ A =1
6
−29 {1, A, A , A , √
2 3
A , A } Z6 , A = [3, √
4 5
1 + −29],
A2 = [5, 1 + √−29], A3 = [2, 1 +√ −29],
A4 = [5, 1 − −29], A5 = [3, 1 −√ −29], A6 = 1
−30 {1, A, B, √AB} Z2 × Z2 , A √ = [2, −30],
B = [3, −30], AB = [5, −30],
A2 = B 2 = (AB)2 = 1
so that
√
H (Q( −65)) Z4 × Z2 .
Table √
6. Nontrivial ideal class groups
H (Q( k)), 2 ≤ k < 100, k squarefree
√
k H (Q( k))
√
10 {1, A} Z2 , A = [2, 10], √ A = 21
2
and
−1 d πr
h(K ) = log sin , if d > 0.
log η 1≤r <d/2 r d
Here w(d) denotes the number of roots of unity in OQ(√d) (d < 0) so that
6, if d = −3,
w(d) = 4, if d = −4,
2, if d < −4,
where dn (n ∈ N) is the Kronecker symbol and η is the fundamental unit of
OQ(√d) (d > 0).
CB609-12 CB609/Alaca & Williams August 7, 2003 17:0 Char Count= 0
14
√ 1 −15
h(Q( −15)) = − r
15 r =1 r
−1
= (1(1) + 2(1) + 3(0) + 4(1) + 5(0) + 6(0) + 7(−1) + 8(1)
15
+ 9(0) + 10(0) + 11(−1) + 12(0) + 13(−1) + 14(−1))
−1
= (1 + 2 + 4 − 7 + 8 − 11 − 13 − 14)
15
−1
= (−30) = 2.
15
√
√ Dirichlet’s formula to show that h(Q( 5)) = 1. In this
Example 12.6.7 We use
case d = 5, η = (1 + 5)/2, and Theorem 12.6.1 gives
2
√ −1 5 πr
h(Q( 5)) = √ log sin
1+ 5 r =1
r 5
log
2
−1 π 2π
= √ log sin − log sin
1+ 5 5 5
log
2
1 2π π
= √ log sin − log sin
1+ 5 5 5
log
2
√ √
1 10 + 2 5 10 − 2 5
= √ log − log
1+ 5 4 4
log
2
√
3+ 5
√ log
1 10 + 2 5 2
= √ log √ = √ = 1.
1+ 5 10 − 2 5 1+ 5
2 log 2 log
2 2
√
Tables 7 and 8, which give the class numbers of quadratic fields Q( k) with k
squarefree between −195 and 197, can be constructed using Dirichlet’s formula.
We conclude this section by determining the ideal class group for two cubic fields
and a quartic field.
CB609-12 CB609/Alaca & Williams August 7, 2003 17:0 Char Count= 0
Table 7. Class
√ numbers of imaginary quadratic fields
K = Q( k), − 195 ≤ k < 0, k squarefree
Example 12.6.8 We show that H (K ) is trivial for the cubic field K = Q(θ), where
θ 3 + θ + 1 = 0 (see Example 7.1.3). Thus O K is a principal ideal domain and so
is a unique factorization domain. Here
D(θ) = −4 · 13 − 27 · 12 = −31
is negative so that x 3 + x + 1 = 0 has one real root and two nonreal roots. Thus
r = 1, s = 1.
Table 8. Class
√ numbers of real quadratic fields
K = Q( k), 0 < k ≤ 197, k squarefree
irreducible (mod 2), so that the principal ideal 2 is prime in O K . The factor-
ization of x 3 + x + 1 into irreducibles (mod 3) is
x 3 + x + 1 ≡ (x − 1)(x 2 + x − 1) (mod 3),
so that by Theorem 10.3.1 the factorization of 3 into prime ideals in O K is
3 = P Q,
where
P = 3, θ − 1, N (P) = 3,
Q = 3, θ 2 + θ − 1, N (Q) = 32 .
Now
(θ − 1)3 + 3(θ − 1)2 + 4(θ − 1) + 3 = 0,
CB609-12 CB609/Alaca & Williams August 7, 2003 17:0 Char Count= 0
so that
θ − 1 | 3,
and thus
P = 3, θ − 1 = θ − 1.
Further,
3
= −4 − 3(θ − 1) − (θ − 1)2 = −2 − θ − θ 2
θ −1
so that
3
Q = 3P −1 = 3θ − 1−1 =
θ −1
= −2 − θ − θ 2 = 2 + θ + θ 2 .
Hence all the prime ideals dividing the principal ideals p ( p (prime) ≤ M K ) are
principal so that the ideal class group H (K ) is trivial.
√ √
Example
√ 12.6.9 We show that H (Q( 3
2)) is trivial. Let θ = 3
2 and K = Q(θ) =
Q( 3 2). Clearly, irrQ θ = x 3 − 2, which has one real root (namely θ) and two
nonreal roots (namely ωθ and ω2 θ, where ω is a complex cube root of unity). Thus
r = 1, s = 1.
It was shown in Example 7.1.6 that {1, θ, θ 2 } is an integral basis for K and d(K ) =
−108. The Minkowski bound is
s
2 2√ 2 21
MK = |d(K )| = 108 < · = 7.
π π 3 2
Thus the primes p ≤ M K are p = 2, 3, and 5. Clearly,
2 = P 3 ,
3 = Q 3 ,
since 1 + θ + θ 2 is a unit of O K as
(1 + θ + θ 2 )(−1 + θ ) = −1 + θ 3 = 1.
CB609-12 CB609/Alaca & Williams August 7, 2003 17:0 Char Count= 0
Finally, as
x 3 − 2 = (x + 2)(x 2 − 2x − 1) (mod 5),
by Theorem 10.3.1 we have
5 = P Q,
where P and Q are distinct prime ideals with
P = 5, 2 + θ , N (P) = 5,
Q = 5, −1 − 2θ + θ 2 , N (Q) = 52 .
Now
5 = 4 + 1 = θ 6 + 1 = (θ 2 + 1)(θ 4 − θ 2 + 1) = (θ 2 + 1)(1 + 2θ − θ 2 ),
so that 1 + 2θ − θ 2 | 5 and thus
Q = 1 + 2θ − θ 2
and
P = 5Q −1 = 51 + 2θ − θ 2 −1
= 5(1 + 2θ − θ 2 )−1 = 5(1 + 2θ − θ 2 )−1
5
= 1 + θ 2 .
1 + 2θ − θ 2
√
Since√all the prime factors of 2, 3, and 5 are principal, H (Q( 3 2)) is trivial. Hence
h(Q( 3 2)) = 1.
√
Q( 3 k) for cubefree
The class numbers of √ √ positive
√ integers
√k up to 101 are given
in Table 9. Note that Q( 3 −k) = Q( 3 k) and Q( k 2 ) = Q( 3 k).
3
Example√12.6.10 We show that the ideal class group H (K ) of the quartic field
K = Q( 2 + i) is trivial. We have already observed that K is a cyclotomic field,
namely, K = Q(ζ8 ) = K 8 . Thus, by Theorem 7.5.2, K is a monogenic
√ field.
√ Indeed
O K = Z + Zζ8 + Zζ8 + Zζ8 , by Theorem 7.5.1. Set θ = ζ8 = ( 2 + −2)/2.
2 3
so that by Theorem 10.3.1 the factorization of the principal ideal 2 into prime
ideals in O K is
2 = P14 ,
where
P1 = 2, 1 + θ , N (P1 ) = 2.
Now
so that 1 + θ | 2. Hence
P1 = 2, 1 + θ = 1 + θ
is principal.
CB609-12 CB609/Alaca & Williams August 7, 2003 17:0 Char Count= 0
so that by Theorem 10.3.1 the factorization of the principal ideal 3 into prime
ideals in O K is
3 = P2 P3 ,
where
(1 − θ − θ 2 )(1 + θ − θ 2 ) = −3θ 2 ,
we deduce that
1 − θ − θ 2 | 3 and 1 + θ − θ 2 | 3,
P2 = 1 − θ − θ 2 , P3 = 1 + θ − θ 2
are principal.
The factorization of x 4 + 1 into irreducible polynomials modulo 5 is given by
so that by Theorem 10.3.1 the factorization of the principal ideal 5 into prime
ideals in O K is
5 = P4 P5 ,
where
Now
(2 + θ 2 )(−2 + θ 2 ) = −4 + θ 4 = −5,
so that
2 + θ 2 | 5, − 2 + θ 2 | 5.
P4 = 2 + θ 2 , P5 = −2 + θ 2
are principal.
CB609-12 CB609/Alaca & Williams August 7, 2003 17:0 Char Count= 0
We have shown that all the prime ideals dividing the principal ideals
p, where p is a prime ≤ M K , are principal so that the ideal class group H (K ) is
trivial. Hence h(K 8 ) = 1.
We conclude this section with a short table of class numbers of cyclotomic fields
(Table 10).
where
!
0, if m ≡ 1 (mod 4),
δ= (12.7.2)
1, if m ≡ 2 or 3 (mod 4).
Let p be an odd prime such that
d(K )
= 1. (12.7.3)
p
We observe that (12.7.3) is equivalent to m
p
= 1 as p = 2. Then
p = P1 P2 ,
CB609-12 CB609/Alaca & Williams August 7, 2003 17:0 Char Count= 0
where P1 and P2 are distinct conjugate prime ideals of O K . Let h denote the class
number h(K ). Then
[P1h ] = [P1 ]h = 1.
showing that 41−δ p h is represented by one or both of the binary quadratic forms
x 2 − my 2 and −x 2 + my 2 .
We prove that
!
1, if m ≡ 2 or 3 (mod 4),
(x, y) = (12.7.5)
1 or 2, if m ≡ 1 (mod 4).
Suppose that q is an odd prime with q | (x, y). Then x = q x1 and y = qy1 for
integers x1 and y1 with x1 ≡ y1 (mod 2) if m ≡ 1 (mod 4). From (12.7.4) we deduce
that q 2 | 41−δ p h . As q = 2 we must have q = p. Thus
√ √
x1 + y1 m x1 + y1 m
P1 = p
h
= P1 P2 ,
21−δ 21−δ
CB609-12 CB609/Alaca & Williams August 7, 2003 17:0 Char Count= 0
so that P2 | P1h , contradicting that P1 and P2 are distinct prime ideals. Hence there
are no odd primes dividing (x, y) and so
(x, y) = 2w
for some nonnegative integer w. From (12.7.4) we deduce that 22w | 22(1−δ) so that
0 ≤ w ≤ 1 − δ. If m ≡ 2 or 3 (mod 4) then δ = 1 and w = 0. If m ≡ 1 (mod 4)
then δ = 0 and w = 0 or 1.
If m is negative then x 2 − my 2 > 0, so the plus sign holds in (12.7.4). If m is
positive and there exist integers T and U such that T 2 − mU 2 = −1 then
x 2 − y 2 ≡ x 2 − my 2 = ±4 p h ≡ 4 (mod 8),
p h = ±(u 2 − mv 2 ).
x = v + 2u, y = v,
so that
p h = ±(u 2 + uv + ( 1−m
4
)v 2 ).
so that (u, v) = 1 or 2. But (u, v)2 | p h , so that (u, v) = p t for some nonnegative
integer t. Hence t = 0 and (u, v) = 1. We have proved the following result.
Theorem
12.7.1 Let m be a squarefree integer. Let p be an odd prime with
√
m
p
= 1. Let h denote the class number of the quadratic field Q( m).
CB609-12 CB609/Alaca & Williams August 7, 2003 17:0 Char Count= 0
The reader should compare this theorem with Theorems 1.4.4 and 1.4.5.
In the opposite direction to Theorem 12.7.1 we have the following simple result.
p k = au 2 + buv + cv 2 . (12.7.6)
Proof: Suppose on the contrary that d
p
= −1 and there are coprime integers u
and v satisfying (12.7.6). Then, as d = b − 4ac, we have
2
In the next three examples we apply Theorems 12.7.1 and 12.7.2 in the cases
m = −1, −2, and −3. We recover Theorems 2.5.1, 2.5.2, and 2.5.3 respectively
(see Exercises 12, 14, and 16 of Chapter 2).
√
Example
12.7.1 m = −1. Here h = h(Q( −1)) = 1. If p is an odd prime with
−1
p
= 1, by Theorem 12.7.1 there exist (coprime) integers u and v such that
p = u 2 + v 2 . Conversely, by Theorem 12.7.2, if there exist integers u and v such
CB609-12 CB609/Alaca & Williams August 7, 2003 17:0 Char Count= 0
p = 2. Since −1p
= 1 ⇐⇒ p ≡ 1 (mod 4), and 2 = 12 + 12 , we deduce that for
a prime p
p = u 2 + v 2 ⇐⇒ p = 2 or p ≡ 1 (mod 4)
p = u 2 + 2v 2 ⇐⇒ p = 2 or p ≡ 1, 3 (mod 8)
Since −3p
= 1 ⇐⇒ p ≡ 1 (mod 3), 2 = u 2 + uv + v 2 , and 3 = 12 + 1 · 1 + 12 ,
we have for a prime p
p = u 2 + uv + v 2 ⇐⇒ p = 3 or p ≡ 1 (mod 3)
√
We now give an example with h(Q( m)) > 1.
√
Example
12.7.4 m = −5. Here h = h(Q( −5)) = 2. If p is an odd prime with
−5
p
= 1, by Theorem 12.7.1 there exist coprime integers u and v such that p 2 =
u 2 + 5v 2 . Conversely, by Theorem 12.7.2, if there exist coprime
integers u and v
such that p = u + 5v , where p is an odd prime, then −5
2 2 2
p
= 0 or 1. Hence
p = 5 or p ≡ 1, 3, 7, or 9 (mod 20). Clearly 22 , 52 = u 2 + 5v 2 with (u, v) = 1, so
that for a prime p
We note that
p = P1 P2 ,
2 p = x 2 + 5y 2 =⇒ 2 p ≡ x 2 (mod 5) =⇒ 2 p ≡ 1 or 4 (mod 5)
p
=⇒ p ≡ 2, 3 (mod 5) =⇒ = −1.
5
By the law of quadratic reciprocity, we have
−1 p −1 5 −5
= = = 1.
p 5 p p p
Hence, for a prime p = 2, 5, we have proved that
−1 p
p = x + 5y ⇐⇒
2 2
= = 1,
p 5
−1 p
2 p = x + 5y ⇐⇒
2 2
= = −1.
p 5
Then there exist integers y and z such that 2 p = z 2 + 5y 2 . Clearly z ≡ y (mod 2).
Thus we can define an integer x by z = y + 2x. Then
2 p = (y + 2x)2 + 5y 2 = 4x 2 + 4x y + 6y 2
so that
p = 2x 2 + 2x y + 3y 2 .
p = x 2 + 5y 2 =⇒ p 2 = u 2 + 5v 2 , (u, v) = 1,
with u = x 2 − 5y 2 , v = 2x y,
and
p = 2x 2 + 2x y + 3y 2 =⇒ p 2 = u 2 + 5v 2 , (u, v) = 1,
with u = 2x 2 + 2x y − 2y 2 , v = 2x y + y 2 .
and
√
[A][B] = [AB] = [5, 3 + −21], [A]2 = [B]2 = [AB]2 = 1
p = P1 P2 ,
Similarly, if [P1 ] = [B] we find that 3 p = x 2 + 21y 2 (x, y ∈ Z) and if [P1 ] = [AB]
then 5 p = x 2 + 21y 2 (x, y ∈ Z).
Next if p = x 2 + 21y 2 then we have
2 2
p x + 21y 2 x
= = = 1,
3 3 3
2
p x + 21y 2 x2
= = = 1,
7 7 7
−1
= 1, as p ≡ x 2 + y 2 ≡ 1 (mod 4).
p
CB609-12 CB609/Alaca & Williams August 7, 2003 17:0 Char Count= 0
Similarly,
p p −1
2 p = x + 21y =⇒
2 2
= −1, = 1, = −1,
3 7 p
p p −1
3 p = x 2 + 21y 2 =⇒ = 1, = −1, = −1,
3 7 p
p p −1
5 p = x + 21y =⇒
2 2
= −1, = −1, = 1.
3 7 p
−21
However, if p = 2, 3, 7 is a prime such that p
= −1 then p, 2 p, 3 p, 5 p =
x 2 + 21y 2 , so we have proved the following: If p = 2, 3, 7 is a prime then
p p −1
p = x 2 + 21y 2 ⇐⇒ = 1, = 1, = 1,
3 7 p
p p −1
2 p = x + 21y ⇐⇒
2 2
= −1, = 1, = −1,
3 7 p
p p −1
3 p = x + 21y ⇐⇒
2 2
= 1, = −1, = −1,
3 7 p
p p −1
5 p = x + 21y =⇒
2 2
= −1, = −1, = 1.
3 7 p
We leave it to the reader to determine which forms represent the primes 2, 3, and 7.
Exercises 341
However, such a result does not hold for every form ax 2 + bx y + cy 2 . This is proved
in [6], where it is shown that every arithmetic progression either contains no primes
of the form x 2 + 14y 2 or contains primes of both forms x 2 + 14y 2 and 2x 2 + 7y 2 ,
proving that congruences cannot be used to distinguish the representability of a
prime by x 2 + 14y 2 from that by 2x 2 + 7y 2 . By the methods used in Examples
12.7.5 and 12.7.6 we can prove that for a prime p = 2, 7
p 2
p = x + 14y or 2x + 7y ⇐⇒
2 2 2 2
= =1
7 p
⇐⇒ p ≡ 1, 9, 15, 23, 25, 39 (mod 56)
and
p 2
p = 3x + 2x y + 5y ⇐⇒
2 2
= = −1
7 p
⇐⇒ p ≡ 3, 5, 13, 19, 27, 45 (mod 56).
Muskat [5] has shown how to distinguish the representations p = x 2 + 14y 2
and p = 2x 2 + 7y 2 as follows. Let p be a prime with p ≡ 1, 9, 15, 23, 25,
39 (mod 56). Then, as in Example 12.7.1, we can show that p = u 2 + 7v 2 for some
integers u and v. If p ≡ 1 (mod 8) then u is odd and v ≡ 0 (mod 4), and replacing
u by −u if necessary we may suppose that u ≡ 1 (mod 4); if p ≡ 7 (mod 8) then
u ≡ 0 (mod 4) and v is odd, and replacing v by −v if necessary we may suppose
that v ≡ 1 (mod 4). Thus in both cases we have
2 p + u + v ≡ 3 (mod 4)
and Muskat has proved that
p = x 2 + 14y 2 ⇐⇒ 2 p + u + v ≡ 3 (mod 8),
p = 2x 2 + 7y 2 ⇐⇒ 2 p + u + v ≡ 7 (mod 8).
Exercises
√ √
1. Prove that H (Q( −6)) = {1, [2, −6]} Z2 .
√
2. Prove that h(Q( −7)) = 1.
√
3. Prove that h(Q( −11)) = 1.
√
4. Prove that H (Q( −13)) Z2 .
√
5. Prove that H (Q( −15)) Z2 .
√
6. Prove that H (Q( −17)) Z4 .
√
7. Prove that H (Q( −23)) Z3 .
√
8. Prove that H (Q( −26)) Z6 .
√
9. Prove that H (Q( −30)) Z2 × Z2 .
√
10. Prove that H (Q(√ −47)) Z5 .
11. Prove that h(Q(√6)) = 1.
12. Prove that h(Q( 10)) = 2.
CB609-12 CB609/Alaca & Williams August 7, 2003 17:0 Char Count= 0
16. 3
Prove that h(Q( √ 5)) = 1.
17. Determine h(Q( 4 2)).
18. Let p be a prime = 2, 5. Prove that
Biographies 343
√
28. Let p be a prime with p ≡ 3 (mod 4). It is known that h(Q( p)) is odd. Use this fact
to prove that there exist integers a and b such that
a 2 − pb2 = (−1)( p+1)/4 2.
√
[Hint: Consider the ideal 2, 1 + p.]
Suggested Reading
1. Z. I. Borevich and I. R. Shafarevich, Number Theory, Academic Press, New York and
London, 1966.
Dirichlet’s formula for the class number of a quadratic field is proved in Chapter 5.
2. L. Carlitz, A characterization of algebraic number fields with class number two, Pro-
ceedings of the American Mathematical Society 11 (1960), 391–392.
It is proved that an algebraic number field K has class number h(K ) ≤ 2 if and only if whenever
a nonzero, nonunit α ∈ O K can be written as α = uπ1 · · · πs = u π1 · · · πt with u, u units and
π1 , . . . , πs , π1 , . . . , πt are primes in O K then s = t.
3. D. A. Marcus, Number Fields, Springer-Verlag, New York, Heidelberg, Berlin, 1977.
Chapter 5 contains a proof of (12.5.1).
4. J. M. Masley and H. L. Montgomery, Cyclotomic fields with unique factorization, Journal
für die reine und angewandte Mathematik 286/287 (1976), 248–256.
The authors prove that there are precisely 29 distinct cyclotomic fields K m (m ≡ 2 (mod 4)) with
h(K m ) = 1, namely those given by
m = 3, 4, 5, 7, 8, 9, 11, 12, 13, 15, 16, 17, 19, 20, 21, 24, 25, 27, 28, 32, 33, 35, 36,
40, 44, 45, 48, 60, 84.
Biographies
1. J. V. Brawley, In memoriam: Leonard Carlitz (1907–1999), Finite Fields and Applica-
tions 6 (2000), 203–206.
A brief biography of Carlitz is given.
2. F. T. Howard, In memoriam—Leonard Carlitz, Fibonacci Quarterly 38 (2000), 316.
Another brief biography of Carlitz is given.
CB609-13 CB609/Alaca & Williams August 7, 2003 17:4 Char Count= 0
13
Dirichlet’s Unit Theorem
n = r + 2s (13.1.1)
and
1 n
r + s ≥ (r + 2s) = ≥ 1.
2 2
If s = 0 then all the conjugate fields of K are real and K is said to be a totally
real field. If r = 0 then K and all its conjugate fields are nonreal and K is said to be
a totally complex or totally imaginary field. If K is a normal field then K is either
totally real or totally complex, since all the conjugate fields of K coincide.
344
CB609-13 CB609/Alaca & Williams August 7, 2003 17:4 Char Count= 0
K = Q(θ), where θ 3 − 6θ + 2 = 0,
√
Example
√ 13.1.2
√ The√field Q( √2 + i) is totally
√ complex as the conjugates of
2 + i are 2 + i, 2 − i, − 2 + i, − 2 − i.
√
Example 13.1.3 Let K = Q( 3 2). Here n = 3, r = 1, s = 1. The three monomor-
phisms : K −→ C are given by
√ √ √ √
σ1 (a + b 2 + c( 2)2 ) = a + b 2 + c( 2)2 ,
3 3 3 3
√ √ √ √
σ2 (a + b 2 + c( 2)2 ) = a + bω 2 + cω2 ( 2)2 ,
3 3 3 3
√ √ √ √
σ3 (a + b 2 + c( 2)2 ) = a + bω2 2 + cω( 2)2 ,
3 3 3 3
√
where ω = e2πi/3 = (−1 + i 3)/2 is a complex cube root of unity, so that σ3 = σ2 .
Then
√ √ √ √ √ √ √ √
β1 ( 2 + ( 2)2 ) = |σ1 ( 2 + ( 2)2 )| = | 2 + ( 2)2 | = 2 + ( 2)2
3 3 3 3 3 3 3 3
CB609-13 CB609/Alaca & Williams August 7, 2003 17:4 Char Count= 0
and
√ √ √ √ √ √ 2
β2 ( 2 + ( 2)2 ) = |σ + = |ω + ω 2) |
3 3 3 3 2 3 2 3
( 2 ( 2) )| 2 (
2
√
−1 + i √3 √ −1 − i 3 √
= 2+
3
( 2)2
3
2 2
√
−1 √ √ 3 √ √
= ( 2 + ( 2)2 ) + i ( 2 − ( 2)2 )
3 3 3 3
2 2
√ √ √ √
( 3 2 + ( 3 2)2 )2 + 3( 3 2 − ( 3 2)2 )2
=
4
√ √
= −2 + 2 2 + ( 2)2 .
3 3
a = c 1 ω1 + · · · + c n ωn .
We set
and
D 2 = d(K )
CB609-13 CB609/Alaca & Williams August 7, 2003 17:4 Char Count= 0
so that
|D| = |d(K )|1/2 (13.2.3)
and
D = 0. (13.2.4)
Proof: We have
a = c 1 ω1 + · · · + c n ωn
so that for i = 1, 2, . . . , n
σi (a) = c1 σi (ω1 ) + · · · + cn σi (ωn )
and thus for i = 1, 2, . . . , r + s we have
βi (a) = |σi (a)| = |c1 σi (ω1 ) + · · · + cn σi (ωn )|
≤ |c1 ||σi (ω1 )| + · · · + |cn ||σi (ωn )|
≤ CM + ··· + CM
= nC M.
Lemma 13.2.3 tells us that the integers of K with bounded coordinates have
bounded valuations.
CB609-13 CB609/Alaca & Williams August 7, 2003 17:4 Char Count= 0
Proof: We have
a = c1 ω1 + · · · + cn ωn
so that
σi (a) = c1 σi (ω1 ) + · · · + cn σi (ωn ), i = 1, 2, . . . , n.
Hence, by Cramer’s rule, we have
Ni
ci = , i = 1, 2, . . . , n, (13.2.5)
D
where the determinant D is defined in (13.2.2) and the determinant Ni is
formed from D by replacing the ith column by the column consisting of
σ1 (a), σ2 (a), . . . , σn (a). Expanding Ni by its ith column we obtain
n
Ni = σk (a)(−1)k+i k ,
k=1
Lemma 13.2.4 tells us that the integers of K with bounded valuations have
bounded coordinates. The next lemma is an immediate consequence of this fact.
Lemma 13.2.5 There are only finitely many a ∈ O K , all of whose valuations
βi (a) (i = 1, 2, . . . , r + s) lie below a given limit.
where each ci ∈ Z and |ci | ≤ n!L M n−1 /|d(K )|1/2 . The number of possible choices
for each ci is
n!L M n−1
2 + 1,
|d(K )|1/2
so the number of a ∈ O K with βi (a) ≤ L (i = 1, 2, . . . , r + s) is at most
n
n!L M n−1
2 +1 .
|d(K )|1/2
Lemma 13.2.6 Let a ∈ K . Then
r +s
N (a) = βi (a)di ,
i=1
where
1, i = 1, . . . , r
di = (13.2.7)
2, i = r + 1, . . . , r + s.
Proof: We have
N (a) = |N (a)|
n
=| σi (a)|
ri=1
+s +2s
r
= σi (a) σi (a)
i=r +s+1
ri=1
+s r +s
= σi (a) σi+s (a)
i=r +1
ri=1
+s r +s
= σi (a) σ̄i (a)
i=r +1
i=1
r r +s
= σi (a) σi (a)σi (a)
i=r +1
i=1
r r +s
= σi (a) |σi (a)|2
i=1 i=r +1
r r +s
= |σi (a)| |σi (a)|2
i=1 i=r +1
r +s
= βi (a)di ,
i=1
0 ≤ pi < qi , i = 1, 2, . . . , r + s.
h i = h i−s .
All the constant terms in this system are real. The determinant of the coefficient
matrix of this system is D = 0 (see (13.2.2)), so that the system has a unique
solution (b1 , . . . , bn ) ∈ Cn . The first r equations in the system (13.2.8) have real
coefficients and the last n − r = 2s equations occur in complex conjugate pairs.
Hence (b1 , . . . , bn ) ∈ Rn .
Now let
qi − pi
δ = min ,
1≤i≤r +s 2Mn
so that
qi − pi
0<δ≤ , i = 1, 2, . . . , r + s. (13.2.9)
2Mn
Next choose ci ∈ Q such that
|bi − ci | < δ, i = 1, 2, . . . , n.
CB609-13 CB609/Alaca & Williams August 7, 2003 17:4 Char Count= 0
Set
a = c 1 ω1 + · · · + c n ωn ∈ K .
Then
σi (a) = c1 σi (ω1 ) + · · · + cn σi (ωn ), i = 1, 2, . . . , n,
so that
σi (a) − h i = (c1 − b1 )σi (ω1 ) + · · · + (cn − bn )σi (ωn ), i = 1, 2, . . . , n.
Hence for i = 1, 2, . . . , r + s we have
|σi (a) − h i | ≤ |c1 − b1 ||σi (ω1 )| + · · · + |cn − bn ||σi (ωn )|
≤ M(|c1 − b1 | + · · · + |cn − bn |)
< Mnδ
qi − pi
≤ ,
2
so that
(qi − pi ) (qi − pi )
hi − < |σi (a)| < h i + , i = 1, 2, . . . , r + s,
2 2
that is,
pi < βi (a) < qi , i = 1, 2, . . . , r + s.
βi (b) ≤ mkn M, i = 1, 2, . . . , r + s.
b βi (b) mkn M
βi (a) = βi = ≤ = kn M, i = 1, 2, . . . , r + s,
m βi (m) m
by Lemmas 13.2.1 and 13.2.2.
Lemma 13.2.10 There exists a fixed bound B > 0 such that for each number a ∈ K
with 12 < N (a) ≤ 1 there exists a unit ∈ O K such that
β j (a) ≤ B, j = 1, 2, . . . , r + s.
Proof: Let a ∈ K satisfy 12 < N (a) ≤ 1. Set I = a so that 12 < N (I ) ≤ 1. Let
S be the set of all such distinct principal ideals I . By Lemma 13.2.9, for each
I = a in S there exists b(= 0) ∈ I such that
βi (b) ≤ n M, i = 1, 2, . . . , r + s.
so that
Hence among the principal ideals q there are only finitely many different ones,
say,
q1 , . . . , qt .
CB609-13 CB609/Alaca & Williams August 7, 2003 17:4 Char Count= 0
Thus each q = q j for some unit of O K and some j ∈ {1, 2, . . . , t}. Set
l= max βi (q −1
j ).
i = 1, . . . , r + s
j = 1, . . . , t
Thus
βi (a) ≤ lβi (q j )βi (a) = lβi (q j a) = lβi (qa) = lβi (b) ≤ ln M,
that is βi (a) ≤ B (i = 1, 2, . . . , r + s) with B = ln M.
If j ∈ {r + 1, . . . , r + s − 1} then
2
2 s−1 1
β1 (a) · · · βr (a)βr +1 (a) · · · βr +s (a) > B
2 2 r
B n−2
21/2 B 2
B r +2s−2 1
= n−2
=
2B 2
and
2
r
2/n 2 s−1 1
β1 (a) · · · βr (a)βr +1 (a)2 · · · βr +s (a)2 < 21/n B 2 B
22−n B
1 1 n−2
2
r +2s−2
2 n B r +2s−2
= 2 = 1,
21− n B n−2
so that by Lemma 13.2.6
1
< N (a) < 1.
2
Lemma 13.2.12 For each j ∈ {1, 2, . . . , r + s − 1} there exists a unit j ∈ O K
such that
βi ( j ) < 1, i = 1, 2, . . . , r + s, i = j,
β j ( j ) > 1.
so that β j ( j ) > 1.
if and only if
1r1 · · · krk = 1 (r1 , . . . , rk ∈ Z) =⇒ r1 = · · · = rk = 0.
Our next lemma shows that the units 1 , 2 , . . . , r +s−1 constructed in Lemma
13.2.12 are independent.
As
r +s−1
−ρ j
j =1
j=1
we can replace (ρ1 , . . . , ρr +s−1 ) by (−ρ1 , . . . , −ρr +s−1 ), if necessary, to ensure that
at least one of ρ1 , . . . , ρr +s−1 is positive. Relabeling 1 , . . . , r +s−1 , if necessary,
we may suppose that ρ1 , . . . , ρk (k ≥ 1) are positive and ρk+1 , . . . , ρr +s−1 are
nonpositive. From the valuations β1 , . . . , βk we form the product
β = β1d1 · · · βkdk
and from the remaining valuations we form
β = βk+1
d d
k+1
· · · βr +s
r +s
.
By Lemmas 13.2.1 and 13.2.2 we have
β(1) = β (1) = 1, β(a)β(b) = β(ab), β (a)β (b) = β (ab), (13.2.11)
for all a and b in K . By Lemma 13.2.7 we obtain
β()β () = 1,
and thus
β () = β −1 () (13.2.12)
for every unit of O K . For j = 1, 2, . . . , k we have
β ( j ) = βk+1
d d
k+1
· · · βr +s
r +s
( j ) = βk+1 ( j )dk+1 · · · βr +s ( j )dr +s ,
so that
β ( j ) < 1, j = 1, 2, . . . , k,
CB609-13 CB609/Alaca & Williams August 7, 2003 17:4 Char Count= 0
so that
β( j ) < 1, j = k + 1, . . . , r + s − 1,
so that by (13.2.12)
k r +s−1
β ( j )ρ j β( j )−ρ j = 1.
j=1 j=k+1
However, all of the factors on the left-hand side are ≤ 1, and the first k of them are
< 1. This is the required contradiction.
For j = 1, 2, . . . , r + s − 1 we set
a j = β j ( j ), (13.2.13)
a j > 1, j = 1, 2, . . . , r + s − 1. (13.2.14)
βν () ≤ 1, ν = 1, 2, . . . , r + s − 1,
satisfies
βr +s (η)dr +s = 1
βr +s (η) = 1,
CB609-13 CB609/Alaca & Williams August 7, 2003 17:4 Char Count= 0
and this is all we require in this case. Thus we may suppose that r + s ≥ 2. Let
be a fixed unit of O K satisfying
βν () ≤ 1, ν = 1, 2, . . . , r + s − 1.
with
kν ≥ 0, ν = 1, 2, . . . , r + s − 1,
and
βν (η) ≤ aν , ν = 1, 2, . . . , r + s − 1.
We note that is such a unit in view of (13.2.14). For these units we have by Lemma
13.2.12
k
βr +s (η) = βr +s (1k1 · · · r +s−1
r +s−1
)
r +s−1
= βr +s () βr +s (i )ki
i=1
≤ βr +s (),
1 < βν (η), ν = 1, 2, . . . , r + s − 1.
1 ≥ βν0 (η).
for ν = ν0 we have
Lemma 13.2.16 For each unit ∈ O K there exist integers τ1 , . . . , τr +s−1 such that
the unit
τ
η = 1τ1 · · · r +s−1
r +s−1
satisfies
1 < βν (η) ≤ aν , ν = 1, 2, . . . , r + s − 1,
and
βr +s (η) ≤ 1.
Proof: The case r + s = 1 follows as in the proof of Lemma 13.2.14. Thus we may
suppose that r + s ≥ 2. Let be a unit of O K . Set
X= max βν ().
1≤ν≤r +s−1
σ
By Lemma 13.2.15 there exists a unit 0 = 1σ1 · · · r +s−1
r +s−1
of O K satisfying
1 < βν (0 ), ν = 1, 2, . . . , r + s − 1.
Set
Y = min βν (0 ),
1≤ν≤r +s−1
so that
Y > 1.
We may choose k ∈ N so that
Y k ≥ X.
Then
βν (0 )k ≥ βν (), ν = 1, 2, . . . , r + s − 1.
Hence the unit λ = 0−k of O K satisfies
βν ()
βν (λ) = βν (0−k ) = ≤ 1, ν = 1, 2, . . . , r + s − 1.
βν (0 )k
CB609-13 CB609/Alaca & Williams August 7, 2003 17:4 Char Count= 0
Thus, by Lemma 13.2.14, there exist integers ρ1 , . . . , ρr +s−1 such that the unit
ρ ρ
η = λ1 1 · · · r +s−1
r +s−1
satisfies
1 < βν (η) ≤ aν , ν = 1, 2, . . . , r + s − 1.
We observe that
ρ ρ τ
η = 0−k 1 1 · · · r +s−1
r +s−1
= 1τ1 · · · r +s−1
r +s−1
with
τ j = ρ j − kσ j , j = 1, 2, . . . , r + s − 1.
Lemma 13.2.17 There exist finitely many units η1 , . . . , ηh of O K such that every
unit of O K is of the form
ρ ρ
= η j 1 1 · · · r +s−1
r +s−1
Hence, by Lemma 13.2.5, there are only finitely many such η, say η1 , . . . , ηh . Thus
ρ ρ
= η j 1 1 · · · r +s−1
r +s−1
We are now in a position to complete the proof of Dirichlet’s unit theorem in the
next section.
1 , . . . , r +s−1 , η1 , . . . , ηh ,
CB609-13 CB609/Alaca & Williams August 7, 2003 17:4 Char Count= 0
that is,
U (O K ) = 1 , . . . , r +s−1 , η1 , . . . , ηh .
H = 1 , . . . , r +s−1 .
( H )h = H,
so that
h ∈ H.
a11 x1 + · · · + am1 xm = 0,
···
a1 r +s−1 x1 + · · · + am r +s−1 xm = 0.
= 10 · · · r0+s−1
= 1.
This proves that any m units of O K with m ≥ r + s are not independent. Therefore
there are no more than r + s − 1 independent units in O K . But by Lemma 13.2.13
the r + s − 1 units 1 , . . . , r +s−1 are independent. Hence, by the main theorem of
CB609-13 CB609/Alaca & Williams August 7, 2003 17:4 Char Count= 0
As η, θ are roots of unity so is ηθ −1 and thus there is a positive integer k such that
−1 k
ηθ = 1. Hence
k(y1 −x1 ) k(y −xr +s−1 )
1 · · · r +s−1
r +s−1
= 1.
so that
x1 = y1 , . . . , xr +s−1 = yr +s−1 ,
and thus η = θ.
and
⇐⇒ r + s = 1
⇐⇒ r = 1, s = 0 or r = 0, s = 1
⇐⇒ K = Q or K = imaginary quadratic field.
In the next six lemmas we prove elementary results about φ(k) that will be used
in the proofs of our theorems giving information about the roots of unity in the ring
of integers of an algebraic number field (Theorems 13.5.1–13.5.4).
so that
φ( p k ) ≥ pk .
and if α ≥ 2
√ √
φ(n) = 2α−1 φ(N ) ≥ 2α/2 φ(N ) ≥ 2α/2 N = n
so that
n
φ(n) ≥
2
for all positive integers n.
as α ≥ 2.
We now use Lemma 13.5.2 to show that the ring of integers of an algebraic
number field can only contain finitely many roots of unity.
and thus
[Q(ζk ) : Q] ≤ [K : Q],
that is,
φ(k) ≤ n.
If K has odd degree then we can say exactly which roots of unity are in O K .
Theorem 13.5.2 Let K be an algebraic number field of odd degree n. Then the
only roots of unity in O K are ±1.
Theorem 13.5.3 The only roots of unity in the ring of integers of a cubic field are
±1.
The situation is much more complicated when n is even. We just determine the
roots of a unity in the ring of integers of a quartic field.
Theorem 13.5.4 Let K be a quartic field. Then the only possible roots of
unity = ±1 in O K are
ζ3 , ζ4 , ζ5 , ζ6 , ζ8 , ζ10 , ζ12 ,
Proof: Let ζk be a primitive kth root of unity in O K . Then Q(ζk ) ⊆ K and thus
[Q(ζk ) : Q] | [K : Q], that is, φ(k) | 4. Hence φ(k) = 1, 2, or 4. Thus, by Lem-
mas 13.5.3–13.5.5, we have k = 1, 2, 3, 4, 5, 6, 8, 10, or 12. Hence the only possi-
ble roots of unity in O K are ζ1 = 1, ζ2 = −1, ζ3 , ζ4 , ζ5 , ζ6 , ζ8 , ζ10 , ζ12 , and their
powers. √ √
As e2πi/3 = (−1 + i 3)/2, we have Q(ζ3 ) = Q(e2πi/3 ) = Q( −3), so that
√
ζ3 ∈ O K ⇐⇒ K ⊇ Q(ζ3 ) ⇐⇒ K ⊇ Q( −3).
√
Similarly,
√ we can show that ζ4 ∈ O K ⇐⇒ K ⊇ Q( −1) and ζ6 ∈ O K ⇐⇒ K ⊇
Q( −3).
Next, as
1 √ √
e 2πi/5
= ( 5−1+i 10 + 2 5),
4
we have
√
Q(ζ5 ) = Q(e 2πi/5
) = Q( −10 − 2 5),
CB609-13 CB609/Alaca & Williams August 7, 2003 17:4 Char Count= 0
so that
ζ5 ∈ O K ⇐⇒ K ⊇ Q(ζ5 )
⇐⇒ K = Q(ζ5 ) (as [K : Q] = [Q(ζ5 ) : Q] = 4)
√
⇐⇒ K = Q( −10 − 2 5).
Similarly, we can show that
√
ζ10 ∈ O K ⇐⇒ K = Q( −10 − 2 5).
Finally,
1+i
e2πi/8 = √ ,
2
so that
1+i √ √
Q(ζ8 ) = Q(e 2πi/8
)=Q √ = Q( 2, −1).
2
Thus
ζ8 ∈ O K ⇐⇒ K ⊇ Q(ζ8 )
⇐⇒ K = Q(ζ8 ) (as [K : Q] = [Q(ζ8 ) : Q] = 4)
√ √
⇐⇒ K = Q( 2, −1).
The corresponding result for ζ12 can be shown similarly.
Example 13.5.1 There are infinitely many quartic fields K such that ζ4 ∈ O K .
Let
P = {2, 3, 5, 7, 11, 13, 17, . . .}
be the set of prime numbers. It is a theorem going back to Euclid that P is an infinite
set. For p ∈ P let
√ √
K p = Q( −1, p).
An easy calculation shows that
[K p : Q] = 4 for all p ∈ P.
Moreover, the only quadratic subfields of K p ( p ∈ P) are
√ √ √
Q( −1), Q( p), Q( − p).
√
Let p, q ∈ P with p√ = q. Then Q( q) is a quadratic subfield of K q but not of
√ √ √
K p as Q( q) = Q( −1), Q( p), Q( − p). Hence K p = K q . This shows that
{K p | p ∈ P} is an infinite set of distinct quartic fields. The ring of integers of each
K p ( p ∈ P) contains ζ4 by Theorem 13.5.4.
CB609-13 CB609/Alaca & Williams August 7, 2003 17:4 Char Count= 0
Example 13.5.2 There are infinitely many quartic fields K such that the only roots
of unity in their rings of integers O K are ±1. Let
f p (x) = x 4 − 2x 2 + (1 − p) ∈ Z[x].
[K p : Q] = 4.
√
If q ∈ P3,4 is such that q = p then an easy calculation shows that q ∈ / K p . But
√
q ∈ K q so that K p = K q . Hence, as there are infinitely many primes p ≡ 3
(mod 4), {K p | p ∈ P3,4 } is an infinite set of distinct quartic fields. As K p ⊆ R,
none of the roots of unity ζk , k ∈ {3, 4, 5, 6, 8, 10, 12}, belongs to O K p . Thus, by
Theorem 13.5.4, the only roots of unity in O K p are ±1.
with equality if and only if θ = (2k + 1)π/2, k ∈ Z. Thus, for all x ∈ R and all
θ ∈ R, we have
as
so that
Thus
Proof: We have
x 2 x
15 2 3 x 27
3 33 27
9
−3 − − = − ≥ − = > 1,
8 8 4 4 4 4 4 4 4 8
so that
x 2
x 15 2
−3 −1> −
8 8 4
and thus
2
x x 15
−3 −1> − .
8 8 4
Theorem 13.6.1 Let K be a real cubic field with two complex embeddings. Let
η > 1 be the fundamental unit of O K . If |d(K )| ≥ 33 then
|d(K )| − 27
η3 > .
4
CB609-13 CB609/Alaca & Williams August 7, 2003 17:4 Char Count= 0
Hence
|d(K )| < 4(η3 + η−3 + 6).
Thus
|d(K )|
η3 + η−3 > −6
4
and so
|d(K )|
η −
6
− 6 η3 + 1 > 0.
4
Completing the square, we obtain, as |d(K )| ≥ 33,
2
2
|d(K )|
η − 3
−3 <0
8
then
|d(K )| |d(K )| 15
− 3 − η3 > −
8 8 4
so that
15 3
η3 < −3= ,
4 4
contradicting η > 1. Hence
|d(K )|
η − 3
−3 ≥0
8
so that
|d(K )| |d(K )| 15
η −
3
−3 > − ,
8 8 4
which gives
|d(K )| − 27
η3 > .
4
We next use Theorem
√ 13.6.1 to determine the fundamental unit > 1 of the ring
of integers of Q( 3 2).
CB609-13 CB609/Alaca & Williams August 7, 2003 17:4 Char Count= 0
243 − 27
η3 > = 54,
4
CB609-13 CB609/Alaca & Williams August 7, 2003 17:4 Char Count= 0
so that
and thus
η2 > 14.
1 < u < η2 .
√
Table 11 gives the fundamental unit > 1 for a few pure cubic fields Q( 3 m), m ∈
N.
Table 12 gives the fundamental unit of the first thirty cubic fields K with exactly
one real embedding arranged in order of increasing |d(K )|.
−23 x3 + x2 − 1 θ + θ2
−31 x3 − x2 − 1 θ
−44 x − x2 − x − 1
3
θ
−59 x 3 + 2x − 1 2 + θ2
−76 x 3 − 2x − 2 1+θ
−83 x3 − x2 + x − 2 1 + θ2
−87 x 3 + x 2 + 2x − 1 2 + θ + θ2
−104 x3 − x − 2 1 + θ + θ2
−107 x − x 2 + 3x − 2
3
3 + θ2
−108 x3 − 2 1 + θ + θ2
−116 x − x2 − 2
3
1 + θ + θ2
−135 x 3 + 3x − 1 3 + θ2
−139 x + x2 + x − 2
3
3 + 2θ + θ 2
−140 x 3 + 2x − 2 3 + θ + θ2
−152 x 3 − x 2 − 2x − 2 1 + θ + θ2
−172 x3 + x2 − x − 3 2 + 2θ + θ 2
−175 x 3 − x 2 + 2x − 3 2 + θ2
−199 x 3 − x 2 + 4x − 1 4 − θ + θ2
−200 x 3 + x 2 + 2x − 2 9 + 5θ + 3θ 2
−204 x3 − x2 + x − 3 4 + θ + 2θ 2
−211 x 3 − 2x − 3 2 + 2θ + θ 2
−212 x − x 2 + 4x − 2
3
15 − 2θ + 4θ 2
−216 x 3 + 3x − 2 17 + 3θ + 5θ 2
−231 x3 + x2 − 3 2 + 2θ + θ 2
−239 x3 − x − 3 2 + 2θ + θ 2
−243 x3 − 3 4 + 3θ + 2θ 2
−244 x + x 2 − 4x − 6
3
5 + 6θ + 2θ 2
−247 x3 + x − 3 2 + θ + θ2
−255 x3 − x2 − 3 2 + θ + θ2
−268 x + x 2 − 3x − 5
3
3 + 3θ + θ 2
discriminant of x 3 − 4x + 2 is
But the smallest discriminant of a cubic field with three real embeddings is 49
(see Table 13). Hence d(K ) = 37 and so d(K ) = 148. Thus K must be the third
field listed in Table 13. Hence K = Q(φ), where φ 3 + φ 2 − 3φ − 1 = 0. The
CB609-13 CB609/Alaca & Williams August 7, 2003 17:4 Char Count= 0
Table 13. Units of totally real cubic fields K with 0 < d(K) ≤ 1101
d(K ) K = Q(θ) 1 2
49 x 3 + x 2 − 2x − 1 −1 + θ + θ 2 2 − θ2
81 x 3 − 3x − 1 2 + θ − θ2 −θ
148 x + x 2 − 3x − 1
3
θ 2 − θ2
169 x 3 − x 2 − 4x − 1 2 + 2θ − θ 2 −θ
229 x 3 − 4x − 1 θ 2+θ
257 x 3 − 5x − 3 4 + θ − θ2 5 + θ − θ2
316 x + x 2 − 4x − 2
3
−3 + θ + θ 2 −5 + θ + θ 2
321 x 3 + x 2 − 4x − 1 −θ −1 + 2θ + θ 2
361 x 3 + x 2 − 6x − 7 4 + θ − θ2 5 − θ2
404 x 3 − x 2 − 5x − 1 −θ 1 − θ − θ2
469 x 3 + x 2 − 5x − 4 −1 − θ −1 + 2θ + θ 2
473 x 3 − 5x − 1 −θ −2 − θ
564 x + x 2 − 5x − 3
3
−2 + θ −1 − θ + θ 2
568 x 3 − x 2 − 6x − 2 −5 − θ + θ 2 −7 − 4θ + 2θ 2
621 x 3 − 6x − 3 −2 − θ 1 + 2θ
697 x − x 2 − 8x − 5
3
6 + 2θ − θ 2 7 + 2θ − θ 2
733 x 3 + x 2 − 7x − 8 1+θ −5 − 2θ
756 x 3 − 6x − 2 5 − θ2 11 + θ − 2θ 2
761 x 3 − x 2 − 6x − 1 θ 2+θ
785 x 3 + x 2 − 6x − 5 1+θ −4 + θ + θ 2
788 x 3 − x 2 − 7x − 3 2+θ −1 − 2θ
837 x 3 − 6x − 1 −θ −3 − 6θ − 2θ 2
892 x + x 2 − 8x − 10
3
3 + θ − θ2 1 + 3θ + θ 2
940 x 3 − 7x − 4 −11 − 2θ + 2θ 2 −3 + θ + θ 2
961 x + x 2 − 10x − 8
3
−1 + 2θ + 2θ 2 3 + 4θ − 2θ 2
985 x 3 + x 2 − 6x − 1 θ −2 + θ
993 x 3 + x 2 − 6x − 3 5 − θ − θ2 5 − θ2
1016 x 3 + x 2 − 6x − 2 7 − θ − θ2 −11 − θ + θ 2
1076 x 3 − 8x − 6 1+θ −7 − 3θ
1101 x 3 + x 2 − 9x − 12 5 + 2θ − θ 2 −7 − 4θ + 2θ 2
1
φ= .
θ −1
{φ, 2 − φ 2 }.
{θ − 1, 2θ 2 − 4θ + 1}
CB609-13 CB609/Alaca & Williams August 7, 2003 17:4 Char Count= 0
1 θ − θ2 + θ3
2 −1 + θ 2 − θ 3
5 −2 + 2θ − θ 2
6 1 + 4θ − 4θ 2 + 2θ 3
10 −27 + 12θ − θ 2 − 3θ 3
we see that
{θ − 1, 2θ − 1}
√
Table 14 gives a fundamental unit for a few pure quartic fields Q( 4 −m), m ∈ N.
Such fields are totally imaginary quartic fields.
13.7 Regulator
Let {1 , . . . , r +s−1 } and {1 , . . . , r +s−1 } be any two fundamental systems of units
for the ring of integers O K of an algebraic number field K . As {1 , . . . , r +s−1 } is
a fundamental system of units, we have
j = ζ b j 1 1 j · · · r +s−1
a a
r +s−1 j
, j = 1, 2, . . . , r + s − 1, (13.7.1)
a1 j a
j = ρ b j 1 · · · r +s−1
r +s−1 j
, j = 1, 2, . . . , r + s − 1, (13.7.2)
CB609-13 CB609/Alaca & Williams August 7, 2003 17:4 Char Count= 0
A = [ai j ], A = [ai j ].
A A = Ir +s−1 .
Thus
As the matrices A and A have integral entries, both det A and det A are integers
so that det A = det A = ±1 and hence
so that
r +s−1
r +s−1
r +s−1
|σk ( j )| = σk (ζ )b j σk (l )al j = |σk (ζ )|b j |σk (l )|al j = |σk (l )|al j
l=1 l=1 l=1
and thus
+s−1
r
log |σk ( j )| = al j log |σk (l )|. (13.7.5)
l=1
CB609-13 CB609/Alaca & Williams August 7, 2003 17:4 Char Count= 0
E = AE,
and so
Finally,
by (13.7.4).
We have shown that the nonnegative real number
R(K ) = |det E|
R(K ) = log η,
√ √
Example 13.7.1 The fundamental
√ unit of O K , where K = Q( 2), is η = 1 + 2
so that R(K ) = log(1 + 2).
K = Q(θ), θ 3 − 4θ + 2 = 0.
n = r = 3, s = 0, r + s − 1 = 2.
Thus a fundamental system of units of O K comprises two units, and it was shown
in Example 13.6.1 that these can be taken to be θ − 1 and 2θ − 1.
We choose the roots θ, θ , θ of x 3 − 4x + 2 = 0 so that θ < θ < θ . Thus
Then
|θ − 1| 3.2143, log |θ − 1| 1.1676,
|2θ − 1| 5.4286, log |2θ − 1| 1.6916,
|θ − 1| 0.4609, log |θ − 1| −0.7745,
|2θ − 1| 0.0782, log |2θ − 1| −2.5484,
|θ − 1| 0.6751, log |θ − 1| −0.3928,
|2θ − 1| 2.3502, log |2θ − 1| 0.8545.
Hence
log |θ − 1| log |θ − 1|
R(K ) = |det |
log |2θ − 1| log |2θ − 1|
= |log |θ − 1|log |2θ − 1| − log |θ − 1|log |2θ − 1||
|(1.1676)(−2.5484) + (0.7745)(1.6916)|
| − 2.9755 + 1.3101|
= | − 1.6654| = 1.6654.
CB609-13 CB609/Alaca & Williams August 7, 2003 17:4 Char Count= 0
Exercises
1. Prove that there do not exist a, b, c ∈ Z such that
√ √ √ √
−2 + 2 2 + ( 2)2 = a + b 2 + c( 2)2
3 3 3 3
√ √
7. Prove that 109 + 60 3 6 + 33( 3 6)2 is the fundamental unit (> 1) of OQ( √3 6) .
√ √
8. Prove that 4 + 2 3 7 + ( 3 7)2 is the fundamental unit (> 1) of OQ( √3 7) .
CB609-13 CB609/Alaca & Williams August 7, 2003 17:4 Char Count= 0
9. Prove that
K = Q(θ), θ 3 + θ 2 − 2θ − 1 = 0,
√
are independent units in Q( 4 34).
11. Prove that
√ √ √ 1 √ √
1 + 2, 2 + 3, 2+ 6
2
√ √
is a fundamental system of units for O K , where K = Q( 2, 3).
12. Let K be an algebraic number field such that U (O K ) contains a nonreal root of unity.
Prove that N (α) > 0 for every α ∈ K \ {0}.
13. Let K = Q(ζm ), where ζm is a primitive mth root of unity, m ≥ 3. Determine r and s
for K .
14. Let K be a totally imaginary quartic field containing a real quadratic field k. By Theorem
13.4.2 K possesses a fundamental unit. Give conditions under which this fundamental
unit can be taken to be the fundamental unit of k.
15. Prove that
√ √
{1 + 2, 1 + ( 2)2 }
4 4
√
is a fundamental system of units for O K , where K = Q( 4 2).
Suggested Reading
1. H. Cohen, A Course in Computational Number Theory, Springer-Verlag, Berlin,
Heidelberg, New York, 1996.
This book describes 148 algorithms that are fundamental for number theoretic computations.
Algorithms 4.9.9 and 4.9.10 calculate the roots of unity in the ring of integers of an arbitrary
algebraic number field. Algorithm 6.5.8 computes a fundamental system of units.
2. C. Levesque, Systemes fondamentaux d’unites de certains composes de deux corps
quadratiques, I, Canadian Journal of Mathematics 33 (1981), 937–945.
√ √
The author determines a fundamental system of units for certain quartic fields Q( m, n), where
m and n are positive integers.
3. I. Niven, H. S. Zuckerman, and H. L. Montgomery, An Introduction to the Theory of
Numbers, fifth edition, Wiley, New York, 1991.
A proof that Euler’s phi function φ(n) is multiplicative can be found on page 69.
4. B. L. van der Waerden, Ein Logarithmenfreier Beweis des Dirichletschen Einheiten-
satzes, Abhandlungen aus dem Mathematischen Seminar der Universität Hamburg 6
(1928), 259–262.
The proof of Dirichlet’s unit theorem given in this chapter is based upon the approach in this paper.
CB609-13 CB609/Alaca & Williams August 7, 2003 17:4 Char Count= 0
Biographies
1. G. Frei, Bartel Leendert van der Waerden, Historia Mathematica 20 (1993), 5–11.
A brief biography of van der Waerden (1903–1996) is given.
2. G. Frei, J. Top, and L. Walling, A short biography of B. L. van der Waerden, Nieuw
Archief voor Wiskunde 12 (1994), 137–144.
A well-written biography of van der Waerden is given.
CB609-14 CB609/Alaca & Williams August 7, 2003 17:6 Char Count= 0
14
Applications to Diophantine Equations
385
CB609-14 CB609/Alaca & Williams August 7, 2003 17:6 Char Count= 0
Set
k = M 3 − N 2.
y 2 + N 2 = x 3 + M 3 = (x + M)(x 2 − M x + M 2 ). (14.1.2)
Hence x 2 − M x + M 2 is odd and (14.1.3) shows that it has at least one prime factor
p ≡ 3 (mod 4). Thus y 2 ≡ −N 2 (mod p). By assumption p N . Hence
2
−1 −N 2 y
= = = 1,
p p p
The following table gives some values of k (with |k| < 100) covered by Theorem
14.1.1.
M −1 15 3 3
N 2 58 2 10
k −5 11 23 −73
The following table gives some values of k (|k| < 100) for which Theorem
14.1.2 applies.
M 2 −2 2 −2 6 14 6
N 1 1 5 5 13 53 17
k 7 −9 −17 −33 47 −65 −73
If x ≡ 3 (mod 8) then
x 2 − M x + M 2 ≡ 1 − 3M + M 2 ≡ ±3 (mod 8).
Hence x 2 − M x + M 2 is odd and at least one of its prime factors p is ≡ ±3 (mod 8).
Thus p N , y 2 ≡ 2N 2 (mod p), and so
2 2
2 2N y
= = = 1,
p p p
contradicting p ≡ ±3 (mod 8).
If x ≡ 7 (mod 8) then
x + M ≡ 7 + M ≡ ±3 (mod 8).
Hence x + M is odd and at least one of its prime factors p is ≡ ±3 (mod 8). Thus
p N , y 2 ≡ 2N 2 (mod p), and so
2 2
2 2N y
= = = 1,
p p p
contradicting p ≡ ±3 (mod 8).
This proves the insolvability of y 2 = x 3 + k in integers x and y.
The following table gives some values of k (|k| < 100) for which Theorem 14.1.3
applies.
M −2 −4 −10 −4 4 −2
N 1 7 23 1 1 7
k −6 34 58 −62 66 90
Set
k = M 3 − 2N 2 .
y 2 ≡ x 3 + 2 (mod 4).
CB609-14 CB609/Alaca & Williams August 7, 2003 17:6 Char Count= 0
y 2 ≡ x 3 − 2 (mod 8)
so that x
≡ 7 (mod 8). Hence x ≡ 3 (mod 8). Next
y 2 + 2N 2 = x 3 + M 3 = (x + M)(x 2 − M x + M 2 ).
The following table gives some values of k (|k| < 100) for which Theorem 14.1.4
applies.
M 4 4 −4 −4 4
N 3 1 1 3 9
k 46 62 −66 −82 −98
AB = C n ,
where n is a positive integer. Then there exist ideals A1 and B1 of D such that
A = An1 , B = B1n , C = A1 B1 .
C = P1a1 · · · Prar ,
CB609-14 CB609/Alaca & Williams August 7, 2003 17:6 Char Count= 0
Theorem 14.2.2 Let K be an algebraic number field. Let h denote the class number
of K . Let A be an integral ideal of O K such that Ak is a principal ideal for some
positive integer k coprime with h. Then A is a principal ideal.
Proof: Let [A] denote the class of A. As the order of H (K ) is h, we have [A]h = I
so that Ah is a principal ideal. Since (h, k) = 1 there exist integers r and s such that
r h + sk = 1. Then, as Ak is a principal ideal, so is
r k s
A = Ar h+sk = Ah A
as asserted.
We now sketch the ideas involved in using Theorems 14.2.1 and 14.2.2 to obtain
classes of rational integers k for which we can find the solutions (if any) of the
Diophantine equation y 2 = x 3 + k.
We begin by supposing that the equation y 2 = x 3 + k has a solution in integers
x and y, so that
√ √
x 3 = (y + k)(y − k),
√ √ √
where y + k and y − k are integers of the quadratic field K = Q( k). √ We
assume that k is squarefree and that k ≡ 2 or 3 (mod 4) so that O K = Z + Z √ k.
(The latter condition avoids 2’s in the denominators of the integers of K = Q( k).)
Passing to ideals, we obtain
√ √
x3 = y + ky − k.
√ √
If the values of k are chosen so that the principal ideals y + k and y − k
are coprime, then we can deduce from Theorem 14.2.1 that
√
y + k = A3
CB609-14 CB609/Alaca & Williams August 7, 2003 17:6 Char Count= 0
so that b = ±1. It is now an easy matter to determine the possibilities for a, and
then the solutions x, y in integers of y 2 = x 3 + k (see Theorem 14.2.3). If k is
√there are infinitely many possibilities for . Indeed = ±η , where
l
positive then
η = T + U k (> 1) is the fundamental unit of O K and l ∈ Z√(Theorem 11.5.1).
Absorbing the cubes −1 = (−1)3 and ζ 3m = (ζ m )3 into (a + b k)3 we see that we
have only to examine the three equations
√ √
y + k = (a + b k)3 ,
√ √
y + k = η(a + b k)3 ,
and
√ √
y+ k = η2 (a + b k)3 .
The first of these equations can be treated as in the case k < 0. For the other two
equations it is convenient to impose congruence conditions on k, T , and U to
ensure that they do not have any solutions. This is illustrated in Theorem 14.2.4. It
should be noted that absorbed cubes must be taken into account when seeking all
solutions of y 2 = x 3 + k.
CB609-14 CB609/Alaca & Williams August 7, 2003 17:6 Char Count= 0
and so
√ √
2k = k(2 k) ∈ P. (14.2.2)
Now
√ √
y + ky − k = y 2 − k = x 3 = x3
so that
P | x3 .
P | x.
Thus
x ∈ P. (14.2.3)
domain by Theorem 8.1.1, and thus by Theorem 14.2.1 there exists an ideal A of
O K such that
√
y + k = A3 .
√
Thus A3 is a principal ideal and, as h(Q( k))
= 0 (mod 3), by Theorem 14.2.2, A
is a principal ideal, say,
√
A = a + b k,
where a, b ∈ Z. Hence
√ √ √
y + k = a + b k3 = (a + b k)3 .
Thus
√ √ √ √
x 3 = y 2 − k = (y + k)(y − k) = (a + b k)3 (a − b k)3
√ √ 3
= ((a + b k)(a − b k))3 = a 2 − kb2
so that
x = a 2 − kb2 . (14.2.6)
Adding and subtracting (14.2.4) and (14.2.5), we obtain
√ √
2y = ((a + b k)3 + (a − b k)3 )
and
√ √ √
2 k = ((a + b k)3 − (a − b k)3 ),
so that
y = (a 3 + 3kab2 ), 1 = (3a 2 b + kb3 ).
From 1 = b(3a 2 + kb2 ) we see that b = ±1, so that b = ±. If b = then
x = a 2 − k, y = (a 3 + 3ka), 1 = 3a 2 + k,
so that
k = 1 − 3a 2
and
x = 4a 2 − 1, y = ±(3a − 8a 3 ).
Clearly
2
x 3 + k = (4a 2 − 1)3 + (1 − 3a 2 ) = 64a 6 − 48a 4 + 9a 2 = 8a 3 − 3a = y 2 .
If b = − then
x = a 2 − k, y = (a 3 + 3ka), 1 = −3a 2 − k,
so that
k = −1 − 3a 2
and
x = 4a 2 + 1, y = ±(3a + 8a 3 ).
Clearly
3 2
x 3 + k = 4a 2 + 1 − 1 − 3a 2 = 64a 6 + 48a 4 + 9a 2 = 8a 3 + 3a = y 2 .
This completes the proof of the theorem
CB609-14 CB609/Alaca & Williams August 7, 2003 17:6 Char Count= 0
The values of k in the range −200 < k < −2 that√ satisfy the conditions of The-
√ k = −74 = 1 − 3 · 5 (h(Q( −74)) = 10) and k = −146 =
2
orem 14.2.3(a) are
1 − 3 · 72 (h(Q( −146)) = 16). Hence, by Theorem 14.2.3(a), the only solutions
in integers of y 2 = x 3 − 74 are (x, y) = (99, ±985) and the only solutions in inte-
gers to y 2 = x 3 − 146 are (x, y) = (195, ±2723).
In the next theorem we find a result similar to that of Theorem 14.2.3(c) in the
case when k is positive.
k > 0,
k is squarefree,
k ≡ 2, 3 (mod 4),
√
h(Q( k))
≡ 0 (mod 3).
√ √
Let T + U k be the fundamental unit of K = Q( k) of norm 1. If
or
or
= ±ηl
for some l ∈√Z. As the cubes −1√= (−1)3 and η√3m = (ηm )3 can be absorbed into the
cube (a + b k)3 , we have y + k = (a + b k)3 , where = 1, η, or η2 . Further,
as η = η3 /η2 and η2 = η3 /η, we have
√ √ 1 1
y+ k = (a + b k)3 , where ∈ {1, η, } or {1, 2 , η2 }.
η η
We choose ∈ {1, η, 1/η} if η has norm 1 and ∈ {1, 1/η2 , η2 } if η has norm −1.
Thus in both cases
√ √
∈ {1, T + U k, T − U k},
√
where T + U k is√ the fundamental unit (> 1) of O K of norm 1. If = 1, equating
the coefficients of k we obtain 1 = 3a 2 b + kb3 , so that b | 1 and thus b = ±1.
√ ±1 = b = 3a b + kb = 3a + k ≥ k > 1, a contradiction. Thus = T ±
2 2 4 2
Hence
U k. Then
√ √ √
y + k = (T ± U k)(a + b k)3
√ √
= (T ± U k)((a 3 + 3kab2 ) + (3a 2 b + kb3 ) k)
= (T (a 3 + 3kab2 ) ± U k(3a 2 b + kb3 ))
√
+ (T (3a 2 b + kb3 ) ± U (a 3 + 3kab2 )) k
so that
so that
√ √
x 3 = y 2 − k = (y + k)(y − k)
√ √ √ √
= (T ± U k)(a + b k)3 (T ∓ U k)(a − b k)3
3
= (T 2 − kU 2 ) a 2 − kb2
3
= a 2 − kb2
and hence
x = a 2 − kb2 .
Now
x 3 ≡ 0, 1, 6 (mod 7)
and
y 2 ≡ 0, 1, 2, 4 (mod 7),
so
y 2 − x 3 = k ≡ 4 (mod 7)
gives
y 2 ≡ 4 (mod 7), x 3 ≡ 0 (mod 7).
Thus
x ≡ 0 (mod 7)
and so
a 2 − 4b2 ≡ 0 (mod 7);
that is,
a ≡ ± 2b (mod 7).
From U ≡ 0 (mod 7) and T 2 − kU 2 = 1 we deduce that
T 2 ≡ 1 (mod 49)
so that
T ≡ ± 1 (mod 49).
Then from (14.2.7) we obtain
1 ≡ ± 2b3 (mod 7),
which is impossible.
CB609-14 CB609/Alaca & Williams August 7, 2003 17:6 Char Count= 0
so that
y 2 = x 3 − 31 (14.2.9)
Thus
√
P | −31.
√
But P and −31 are both prime ideals so that
√
P = −31.
Hence
√
√ y + −31
−31 |
2
y+
√
−31 √
so that 2
∈ −31. This shows that there exist integers u and v such
that
√ √
y+ −31 √ u + v −31
= −31 .
2 2
√ u =
1 and y =√−31v,
Hence
contradicting 31 y. This proves that the ideals
y + −31 y − −31
2
and 2
are coprime. Thus, replacing y by −y if necessary,
we see from (14.2.11) that there exists an ideal A of O K such that
√ √
y + −31 3 + −31 3
= 2, A ,
2 2
√ √
y − −31 3 − −31 3 (14.2.12)
= 2, Ā ,
2 2
x = A Ā,
2
√
where Ā denotes the conjugate ideal of A. Since h(Q( −31)) = 3 the ideal A3
is principal.
√ Then, from the first equality in (14.2.12), we deduce that the ideal
3 + −31
2, 2
is principal, a contradiction. This completes the proof that the equation
y 2 = x 3 − 31 has no solutions in integers x and y.
We conclude this section by giving two short tables (Tables 15 and 16) of solutions
of y 2 = x 3 + k.
K = Q(θ), θ 3 − 4θ + 2 = 0, (14.3.1)
CB609-14 CB609/Alaca & Williams August 7, 2003 17:6 Char Count= 0
k Solutions (x, y) of y 2 = x 3 + k
−1 (1, 0)
−2 (3, ±5)
−3 insolvable
−4 (2, ±2), (5, ±11)
−5 insolvable
−6 insolvable
−7 (2, ±1), (32, ±181)
−8 (2, 0)
−9 insolvable
−10 insolvable
−11 (3, ±4), (15, ±58)
−12 insolvable
−13 (17, ±70)
−14 insolvable
−15 (4, ±7)
−16 insolvable
−17 insolvable
−18 (3, ±3)
−19 (7, ±18)
−20 (6, ±14)
that is, we determine all those integers that are simultaneously a product of two
consecutive integers and a product of three consecutive integers. This problem was
proposed by Edgar Emerson to Burton W. Jones (1902–1983) and was first solved
by Louis J. Mordell (1888–1972) in 1963. We follow the solution given by Mordell
in his paper [6].
We need the following facts about the field K and its ring of integers O K :
O K = Z + Zθ + Zθ 2 , (14.3.3)
O K is a unique factorization domain, (14.3.4)
a fundamental system of units of O K is{, η},
where = θ − 1 and η = 2θ − 1. (14.3.5)
k Solutions (x, y) of y 2 = x 3 + k
1 (−1, 0), (0, ±1), (2, ±3)
2 (−1, ±1)
3 (1, ±2)
4 (0, ±2)
5 (−1, ±2)
6 insolvable
7 insolvable
8 (−2, 0), (1, ±3), (2, ±4), (46, ±312)
9 (−2, ±1), (0, ±3), (3, ±6), (6, ±15), (40, ±253)
10 (−1, ±3)
11 insolvable
12 (−2, ±2), (13, ±47)
13 insolvable
14 insolvable
15 (1, ±4), (109, ±1138)
16 (0, ±4)
17 (−2, ±3), (−1, ±4), (2, ±5), (4, ±9), (8, ±23),
(43, ±282), (52, ±375), (5234, ±378661)
18 (7, ±19)
19 (5, ±12)
20 insolvable
If we set
X = 2x + 2, Y = 2y + 1, (14.3.6)
2Y 2 = X 3 − 4X + 2. (14.3.7)
Clearly any solution of (14.3.7) must have X even and Y odd. We will show that
the only solutions of (14.3.7) are
(X, Y ) = (−2, ±1), (0, ±1), (2, ±1), (4, ±5), (12, ±29).
(x, y) = (0, 0), (0, −1), (−1, 0), (−1, −1), (−2, 0), (−2, −1), (1, 2),
(1, −3), (5, 14), (5, −15).
This proves that the only integers that are simultaneously a product of two consec-
utive integers as well as a product of three consecutive integers are 0, 6, and 210.
CB609-14 CB609/Alaca & Williams August 7, 2003 17:6 Char Count= 0
Thus
2
∈ U (O K )
θ3
and so
2
=ρ
θ3
for some ρ ∈ U (O K ).
(X − θ)(X 2 + θ X + (θ 2 − 4)) = ρθ 3 Y 2 .
(X − θ)(X 2 + θ X + (θ 2 − 4)) = X 3 − 4X − θ 3 + 4θ
= X 3 − 4X + 2
= 2Y 2
= ρθ 3 Y 2 .
(X 2 + θ X + (θ 2 − 4)) − (X + 2θ)(X − θ)
3θ 3 − 4θ 3(4θ − 2) − 4θ 8θ − 6
= 3θ 2 − 4 = = =
θ θ θ
2
= (4θ − 3) = ρ(4θ − 3)θ 2 ,
θ
by (14.3.1) and Lemma 14.3.3. As ρ is a unit this shows that the only possibilities
for π are π = θ and π = 4θ − 3.
Taking the Eq. (14.3.19) modulo 4, we see that the plus sign holds. Then (14.3.18)–
(14.3.20) can be written as
X = 4AC + 2B 2 + 8C 2 , (14.3.21)
(A + 4C)2 + 4B 2 − 4BC = 1, (14.3.22)
B(A + 4C) = C 2 . (14.3.23)
Hence
x 4 ≥ 4B 3 x − 3B 4
and so
x4
+ 4B 2 − 4Bx ≥ B 2 .
B2
Taking x = C in this inequality, and appealing to (14.3.24), we deduce that
1 ≥ B 2,
C 4 + 4 ∓ 4C = 1,
X = 2 − 12 + 8 = −2.
CB609-14 CB609/Alaca & Williams August 7, 2003 17:6 Char Count= 0
The first equation shows that A is odd and the second that A is even. This case
cannot occur.
(iv): (L , M) = (1, 1). In this case (14.3.15) becomes
2
±(X − θ) = θ(1 − θ )(1 − 2θ) A + Bθ + Cθ 2 .
The first equation shows that A is even and then that B is even since
6B 2 ≡ 0 (mod 4). The second equation shows that B is odd. This case cannot
occur.
This completes the proof of the theorem.
are
(x, y) = (0, 0), (0, −1), (−1, 0), (−1, −1), (−2, 0), (−2, −1), (1, 2), (1, −3),
(5, 14), (5, −15).
CB609-14 CB609/Alaca & Williams August 7, 2003 17:6 Char Count= 0
Proof: This follows immediately from Theorem 14.3.1 by using the transformation
(14.3.6).
Exercises
1. Determine all integers k in the range |k| < 200 to which Theorem 14.1.1 applies.
2. Determine all integers k in the range |k| < 200 to which Theorem 14.1.2 applies.
3. Determine all integers k in the range |k| < 200 to which Theorem 14.1.3 applies.
4. Determine all integers k in the range |k| < 200 to which Theorem 14.1.4 applies.
5. Let M and N be integers such that
M ≡ 4 (mod 8), N ≡ 1 (mod 2),
p (prime) | N =⇒ p ≡ 1 or 3 (mod 8).
Set
k = M 3 − 2N 2 .
Prove that the equation y 2 = x 3 + k has no solutions in integers x and y.
6. Determine all integers k in the range |k| < 200 to which the result of Exercise 5 applies.
7. Let M and N be integers such that
M ≡ 3 (mod 4), N ≡ ±2 (mod 6),
p (prime) | N =⇒ p ≡ ±1 (mod 12).
Set
k = M 3 + 3N 2 .
Prove that the equation y 2 = x 3 + k has no solutions in integers x and y.
8. Determine all integers k in the range |k| < 200 to which the result of Exercise 7 applies.
9. Formulate and prove a result analogous to that of Exercise 7 when k has the form
M 3 − 3N 2 .
10. Determine all integers k in the range |k| < 200 to which the result of Exercise 9 applies.
11. Prove that the equation
y 2 = x 3 + 45
has no solutions in integers x and y.
12. Determine a class of integers k containing k = 45 for which the equation
y2 = x 3 + k
has no solutions in integers x and y.
13. Let M and N be integers such that
M ≡ 2 (mod 6), N ≡ ±1 (mod 6),
p (prime) | M =⇒ p ≡ 2 (mod 3).
Set
k = 4M 3 − 3N 2 .
Prove that the equation y 2 = x 3 + k has no solutions in integers x and y.
CB609-14 CB609/Alaca & Williams August 7, 2003 17:6 Char Count= 0
Biographies 411
14. Determine all integers k in the range |k| < 200 to which the result of Exercise 13
applies.
15. Show that the condition M ≡ 2 (mod 6) can be replaced by M ≡ 0 (mod 6), M
= 0
in Exercise 13 without affecting the result.
16. Determine all integers k in the range |k| < 200 to which the result of Exercise 15
applies.
17. Formulate and prove an analogous result to that of Exercise 13 for k of the form
4M 3 + 3N 2 .
18. Prove that y 2 = x 3 + 13 has no solutions in integers.
19. Prove that y 2 = x 3 + 51 has no solutions in integers.
Suggested Reading
1. W. W. Rouse Ball, A Short Account of the History of Mathematics, Dover, New York,
1960.
A brief discussion of Bachet’s work is given on pages 305 and 306.
2. E. Brown and B. T. Myers, Elliptic curves from Mordell to Diophantus and back, Amer-
ican Mathematical Monthly 109 (2002), 639–649.
In this beautifully written article the authors discuss the number of integer points (x, y) on the
elliptic curve y 2 = x 3 − x + m 2 , where m is a nonnegative integer.
3. H. M. Edgar, Classes of equations of the type y 2 = x 3 + k having no rational solutions,
Nagoya Mathematical Journal 28 (1966), 49–58.
Conditions are given for the equation y 2 = x 3 + k to have no rational solutions.
4. T. Heath, A History of Greek Mathematics, Volume 1: From Thales to Euclid, Volume 2:
From Aristarchus to Diophantus, Dover, New York, 1981.
Chapter 20 in Volume 2 contains an interesting discussion of the work of Diophantus including
his methods for finding solutions in integers of equations of degrees 1, 2, and 3.
5. H. London and R. Finkelstein, On Mordell’s Equation y 2 − k = x 3 , Bowling Green State
University Press, Bowling Green, Ohio, 1973.
The authors provide a comprehensive treatment of the equation y 2 = x 3 + k with many references.
6. L. J. Mordell, On the integer solutions of y(y + 1) = x(x + 1)(x + 2), Pacific Journal
of Mathematics 13 (1963), 1347–1351.
Section 13.3 is based on this beautifully written paper of Mordell.
7. L. J. Mordell, Diophantine Equations, Academic Press, London and New York, 1969.
Mordell’s book is a very readable standard reference text on Diophantine equations.
Biographies
1. V. Bjerknes, Axel Thue, Nordisk Matematisk Tidskrift 4 (1922), 33–46.
A biography of Axel Thue is given.
2. J. W. S. Cassels, L. J. Mordell, Bulletin of the London Mathematical Society 6 (1974),
69–96.
A biography of L. J. Mordell (1888–1972) is given.
3. H. Davenport, L. J. Mordell, Acta Arithmetica 9 (1964), 3–12.
Another biography of L. J. Mordell is presented.
CB609-14 CB609/Alaca & Williams August 7, 2003 17:6 Char Count= 0
http://www-groups.dcs.st-and.ac.uk/˜history/
List of Definitions
413
CB609-15 CB609/Alaca & Williams August 7, 2003 17:11 Char Count= 0
416
CB609-15 CB609/Alaca & Williams August 7, 2003 17:11 Char Count= 0
Location of Theorems
417
CB609-15 CB609/Alaca & Williams August 7, 2003 17:11 Char Count= 0
420
CB609-15 CB609/Alaca & Williams August 7, 2003 17:11 Char Count= 0
Location of Lemmas
421
CB609-15 CB609/Alaca & Williams August 7, 2003 17:11 Char Count= 0
422
CB609-15 CB609/Alaca & Williams August 7, 2003 17:11 Char Count= 0
Bibliography
[1] Z. I. Borevich and I. R. Shafarevich, Number Theory, Academic Press, New York and
London, 1966.
[2] H. Cohen, A Course in Computational Algebraic Number Theory, Springer-Verlag,
Berlin/Heidelberg/New York, 1996.
[3] H. Cohn, Advanced Number Theory, Dover, New York, 1980.
[4] H. Cohn, A Classical Invitation to Algebraic Numbers and Class Fields,
Springer-Verlag, New York/Heidelberg/Berlin, 1978.
[5] D. A. Cox, Primes of the Form x 2 + ny 2 , Wiley, New York, 1989.
[6] R. Dedekind, Theory of Algebraic Integers, Cambridge University Press, Cambridge,
UK, 1996.
[7] J. Esmonde and M. Ram Murty, Problems in Algebraic Number Theory,
Springer-Verlag, New York, 1999.
[8] G. J. Janusz, Algebraic Number Fields, Second Edition, Graduate Studies in
Mathematics Volume 7, American Mathematical Society, Providence, Rhode Island,
1996.
[9] H. Koch, Number Theory: Algebraic Numbers and Functions, Graduate Studies in
Mathematics Volume 24, American Mathematical Society, Providence, Rhode Island,
2000.
[10] S. Lang, Algebraic Number Theory, Springer-Verlag, New York, 1986.
[11] R. L. Long, Algebraic Number Theory, Dekker, New York, 1977.
[12] H. B. Mann, Introduction to Algebraic Number Theory, Ohio State University Press,
Columbus, Ohio, 1955.
[13] D. A. Marcus, Number Fields, Springer-Verlag, New York/Heidelberg/Berlin,
1977.
[14] R. A. Mollin, Algebraic Number Theory, Chapman and Hall/CRC Press,
London/Boca Raton, Florida, 1999.
[15] R. Narasimhan, S. Raghavan, S. S. Rangachari, and S. Lal, Algebraic Number
Theory, Tata Institute of Fundamental Research, Bombay, India, 1966.
[16] W. Narkiewicz, Elementary and Analytic Theory of Algebraic Numbers,
Springer-Verlag, Berlin/Heidelberg/New York, 1989.
[17] T. Ono, An Introduction to Algebraic Number Theory, Plenum, New York, 1990.
[18] M. E. Pohst, Computational Algebraic Number Theory, Birkhäuser Verlag,
Basel/Boston/Berlin, 1993.
[19] M. Pohst and H. Zassenhaus, Algorithmic Algebraic Number Theory, Cambridge
University Press, Cambridge, UK, 1989.
[20] H. Pollard and H. G. Diamond, The Theory of Algebraic Numbers, Mathematical
Association of America, Washington, DC, 1975.
423
CB609-15 CB609/Alaca & Williams August 7, 2003 17:11 Char Count= 0
424 Bibliography
Index
425
CB609-IND CB609/Alaca & Williams August 7, 2003 17:14 Char Count= 0
426 Index
Index 427
428 Index