Aata Original
Aata Original
Aata Original
Thomas W. Judson
Stephen F. Austin State University
c 1997 by Thomas W. Judson.
vii
viii PREFACE
Chapters 0–5
Chapter 9
Chapter 15 Chapter 13
Chapter 19
Chapter 20
Chapter 21
appears at the end of each chapter. The nature of the exercises ranges over
several categories; computational, conceptual, and theoretical problems are
included. A section presenting hints and solutions to many of the exercises
appears at the end of the text. Often in the solutions a proof is only sketched,
and it is up to the student to provide the details. The exercises range in
difficulty from very easy to very challenging. Many of the more substantial
problems require careful thought, so the student should not be discouraged
if the solution is not forthcoming after a few minutes of work. A complete
solutions manual is available for the instructor’s use.
There are additional exercises or computer projects at the ends of many
of the chapters. The computer projects usually require a knowledge of pro-
gramming. All of these exercises and projects are more substantial in nature
and allow the exploration of new results and theory.
Acknowledgements
I would like to acknowledge the following reviewers for their helpful com-
ments and suggestions.
• David Anderson, University of Tennessee, Knoxville
• Robert Beezer, University of Puget Sound
• Myron Hood, California Polytechnic State University
• Herbert Kasube, Bradley University
• John Kurtzke, University of Portland
• Inessa Levi, University of Louisville
• Geoffrey Mason, University of California, Santa Cruz
• Bruce Mericle, Mankato State University
• Kimmo Rosenthal, Union College
• Mark Teply, University of Wisconsin
I would also like to thank Steve Quigley, Marnie Pommett, Cathie Griffin,
Kelle Karshick, and the rest of the staff at PWS for their guidance through-
out this project. It has been a pleasure to work with them.
Thomas W. Judson
Contents
Preface vii
0 Preliminaries 1
0.1 A Short Note on Proofs . . . . . . . . . . . . . . . . . . . . . 1
0.2 Sets and Equivalence Relations . . . . . . . . . . . . . . . . . 4
1 The Integers 22
1.1 Mathematical Induction . . . . . . . . . . . . . . . . . . . . . 22
1.2 The Division Algorithm . . . . . . . . . . . . . . . . . . . . . 26
2 Groups 35
2.1 The Integers mod n and Symmetries . . . . . . . . . . . . . . 35
2.2 Definitions and Examples . . . . . . . . . . . . . . . . . . . . 40
2.3 Subgroups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3 Cyclic Groups 56
3.1 Cyclic Subgroups . . . . . . . . . . . . . . . . . . . . . . . . . 56
3.2 The Group C∗ . . . . . . . . . . . . . . . . . . . . . . . . . . 60
3.3 The Method of Repeated Squares . . . . . . . . . . . . . . . . 64
4 Permutation Groups 72
4.1 Definitions and Notation . . . . . . . . . . . . . . . . . . . . . 73
4.2 The Dihedral Groups . . . . . . . . . . . . . . . . . . . . . . . 81
x
CONTENTS xi
6 Introduction to Cryptography 97
6.1 Private Key Cryptography . . . . . . . . . . . . . . . . . . . . 98
6.2 Public Key Cryptography . . . . . . . . . . . . . . . . . . . . 101
8 Isomorphisms 138
8.1 Definition and Examples . . . . . . . . . . . . . . . . . . . . . 138
8.2 Direct Products . . . . . . . . . . . . . . . . . . . . . . . . . . 143
14 Rings 232
14.1 Rings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232
14.2 Integral Domains and Fields . . . . . . . . . . . . . . . . . . . 237
14.3 Ring Homomorphisms and Ideals . . . . . . . . . . . . . . . . 239
14.4 Maximal and Prime Ideals . . . . . . . . . . . . . . . . . . . . 243
xii CONTENTS
15 Polynomials 256
15.1 Polynomial Rings . . . . . . . . . . . . . . . . . . . . . . . . . 257
15.2 The Division Algorithm . . . . . . . . . . . . . . . . . . . . . 261
15.3 Irreducible Polynomials . . . . . . . . . . . . . . . . . . . . . 265
19 Fields 322
19.1 Extension Fields . . . . . . . . . . . . . . . . . . . . . . . . . 322
19.2 Splitting Fields . . . . . . . . . . . . . . . . . . . . . . . . . . 333
19.3 Geometric Constructions . . . . . . . . . . . . . . . . . . . . . 336
Notation 387
• 3 + 56 − 13 + 8/2.
• 2 + 3 = 5.
1
2 CHAPTER 0 PRELIMINARIES
• 2x = 6 exactly when x = 4.
• x3 − 4x2 + 5x − 6.
All but the first and last examples are statements, and must be either true
or false.
A mathematical proof is nothing more than a convincing argument
about the accuracy of a statement. Such an argument should contain enough
detail to convince the audience; for instance, we can see that the statement
“2x = 6 exactly when x = 4” is false by evaluating 2 · 4 and noting that
6 6= 8, an argument that would satisfy anyone. Of course, audiences may
vary widely: proofs can be addressed to another student, to a professor,
or to the reader of a text. If more detail than needed is presented in the
proof, then the explanation will be either long-winded or poorly written. If
too much detail is omitted, then the proof may not be convincing. Again
it is important to keep the audience in mind. High school students require
much more detail than do graduate students. A good rule of thumb for an
argument in an introductory abstract algebra course is that it should be
written to convince one’s peers, whether those peers be other students or
other readers of the text.
Let us examine different types of statements. A statement could be as
simple as “10/5 = 2”; however, mathematicians are usually interested in
more complex statements such as “If p, then q,” where p and q are both
statements. If certain statements are known or assumed to be true, we
wish to know what we can say about other statements. Here p is called
the hypothesis and q is known as the conclusion. Consider the following
statement: If ax2 + bx + c = 0 and a 6= 0, then
√
−b ± b2 − 4ac
x= .
2a
The hypothesis is ax2 + bx + c = 0 and a 6= 0; the conclusion is
√
−b ± b2 − 4ac
x= .
2a
Notice that the statement says nothing about whether or not the hypothesis
is true. However, if this entire statement is true and we can show that
0.1 A SHORT NOTE ON PROOFS 3
ax2 + bx + c = 0
b c
x2 + x = −
a a
2 2
b b b c
x2 + x + = −
a 2a 2a a
b 2 b2 − 4ac
x+ =
2a 4a2
√
b ± b2 − 4ac
x+ =
2a 2a
√
−b ± b2 − 4ac
x = .
2a
If we can prove a statement true, then that statement is called a propo-
sition. A proposition of major importance is called a theorem. Sometimes
instead of proving a theorem or proposition all at once, we break the proof
down into modules; that is, we prove several supporting propositions, which
are called lemmas, and use the results of these propositions to prove the
main result. If we can prove a proposition or a theorem, we will often,
with very little effort, be able to derive other related propositions called
corollaries.
• Quantifiers are important. Words and phrases such as only, for all,
for every, and for some possess different meanings.
4 CHAPTER 0 PRELIMINARIES
• Never assume any hypothesis that is not explicitly stated in the theo-
rem. You cannot take things for granted.
• Suppose you wish to show that an object exists and is unique. First
show that there actually is such an object. To show that it is unique,
assume that there are two such objects, say r and s, and then show
that r = s.
• Sometimes it is easier to prove the contrapositive of a statement. Prov-
ing the statement “If p, then q” is exactly the same as proving the
statement “If not q, then not p.”
• Although it is usually better to find a direct proof of a theorem, this
task can sometimes be difficult. It may be easier to assume that the
theorem that you are trying to prove is false, and to hope that in the
course of your argument you are forced to make some statement that
cannot possibly be true.
Remember that one of the main objectives of higher mathematics is
proving theorems. Theorems are tools that make new and productive ap-
plications of mathematics possible. We use examples to give insight into
existing theorems and to foster intuitions as to what new theorems might
be true. Applications, examples, and proofs are tightly interconnected—
much more so than they may seem at first appearance.
X = {x1 , x2 , . . . , xn }
X = {x : x satisfies P}
0.2 SETS AND EQUIVALENCE RELATIONS 5
E = {2, 4, 6, . . .}
or
E = {x : x is an even integer and x > 0}.
We write 2 ∈ E when we want to say that 2 is in the set E, and −3 ∈ / E to
say that −3 is not in the set E.
Some of the more important sets that we will consider are the following:
{4, 5, 8} ⊂ {2, 3, 4, 5, 6, 7, 8, 9}
and
N ⊂ Z ⊂ Q ⊂ R ⊂ C.
Trivially, every set is a subset of itself. A set B is a proper subset of a
set A if B ⊂ A but B 6= A. If A is not a subset of B, we write A 6⊂ B; for
example, {4, 7, 9} 6⊂ {2, 4, 5, 8, 9}. Two sets are equal, written A = B, if we
can show that A ⊂ B and B ⊂ A.
It is convenient to have a set with no elements in it. This set is called
the empty set and is denoted by ∅. Note that the empty set is a subset of
every set.
To construct new sets out of old sets, we can perform certain operations:
the union A ∪ B of two sets A and B is defined as
A ∪ B = {x : x ∈ A or x ∈ B};
A ∩ B = {x : x ∈ A and x ∈ B}.
6 CHAPTER 0 PRELIMINARIES
A ∪ B = {1, 2, 3, 5, 9}
and
A ∩ B = {1, 3}.
We can consider the union and the intersection of more than two sets. In
this case we write
n
[
Ai = A1 ∪ . . . ∪ An
i=1
and
n
\
Ai = A1 ∩ . . . ∩ An
i=1
for the union and intersection, respectively, of the collection of sets A1 , . . . An .
When two sets have no elements in common, they are said to be disjoint;
for example, if E is the set of even integers and O is the set of odd integers,
then E and O are disjoint. Two sets A and B are disjoint exactly when
A ∩ B = ∅.
Sometimes we will work within one fixed set U , called the universal
set. For any set A ⊂ U , we define the complement of A, denoted by A0 ,
to be the set
A0 = {x : x ∈ U and x ∈ / A}.
We define the difference of two sets A and B to be
A \ B = A ∩ B 0 = {x : x ∈ A and x ∈
/ B}.
A = {x ∈ R : 0 < x ≤ 3}
and
B = {x ∈ R : 2 ≤ x < 4}.
Then
A ∩ B = {x ∈ R : 2 ≤ x ≤ 3}
A ∪ B = {x ∈ R : 0 < x < 4}
A \ B = {x ∈ R : 0 < x < 2}
A0 = {x ∈ R : x ≤ 0 or x > 3 }.
0.2 SETS AND EQUIVALENCE RELATIONS 7
1. A ∪ A = A, A ∩ A = A, and A \ A = ∅;
2. A ∪ ∅ = A and A ∩ ∅ = ∅;
3. A ∪ (B ∪ C) = (A ∪ B) ∪ C and A ∩ (B ∩ C) = (A ∩ B) ∩ C;
4. A ∪ B = B ∪ A and A ∩ B = B ∩ A;
5. A ∪ (B ∩ C) = (A ∪ B) ∩ (A ∪ C);
6. A ∩ (B ∪ C) = (A ∩ B) ∪ (A ∩ C).
Proof. We will prove (1) and (3) and leave the remaining results to be
proven in the exercises.
(1) Observe that
A ∪ A = {x : x ∈ A or x ∈ A}
= {x : x ∈ A}
= A
and
A ∩ A = {x : x ∈ A and x ∈ A}
= {x : x ∈ A}
= A.
Also, A \ A = A ∩ A0 = ∅.
(3) For sets A, B, and C,
A ∪ (B ∪ C) = A ∪ {x : x ∈ B or x ∈ C}
= {x : x ∈ A or x ∈ B, or x ∈ C}
= {x : x ∈ A or x ∈ B} ∪ C
= (A ∪ B) ∪ C.
1. (A ∪ B)0 = A0 ∩ B 0 ;
2. (A ∩ B)0 = A0 ∪ B 0 .
8 CHAPTER 0 PRELIMINARIES
(A \ B) ∩ (B \ A) = ∅.
(A \ B) ∩ (B \ A) = (A ∩ B 0 ) ∩ (B ∩ A0 )
= A ∩ A0 ∩ B ∩ B 0
= ∅.
{(x, 1), (x, 2), (x, 3), (y, 1), (y, 2), (y, 3)}
and
A × C = ∅.
We define the Cartesian product of n sets to be
A B
f
1 a
2 b
3 c
g
A B
1 a
2 b
3 c
Figure 1. Mappings
not all functions can be described in this manner. For example, the function
f : R → R that sends each real number to its cube is a mapping that must
be described by writing f (x) = x3 or f : x 7→ x3 .
Consider the relation f : Q → Z given by f (p/q) = p. We know that
1/2 = 2/4, but is f (1/2) = 1 or 2? This relation cannot be a mapping
because it is not well-defined. A relation is well-defined if each element in
the domain is assigned to a unique element in the range.
If f : A → B is a map and the image of f is B, i.e., f (A) = B, then
f is said to be onto or surjective. A map is one-to-one or injective
if a1 6= a2 implies f (a1 ) 6= f (a2 ). Equivalently, a function is one-to-one if
f (a1 ) = f (a2 ) implies a1 = a2 . A map that is both one-to-one and onto is
called bijective.
Example 5. Let f : Z → Q be defined by f (n) = n/1. Then f is one-to-one
but not onto. Define g : Q → Z by g(p/q) = p where p/q is a rational number
expressed in its lowest terms with a positive denominator. The function g
is onto but not one-to-one.
Given two functions, we can construct a new function by using the range
of the first function as the domain of the second function. Let f : A → B
and g : B → C be mappings. Define a new map, the composition of f and
g from A to C, by (g ◦ f )(x) = g(f (x)).
A B C (a)
f g
1 a X
2 b Y
3 c Z
A C (b)
gof
1 X
2 Y
3 Z
and
(g ◦ f )(x) = g(f (x)) = 2x2 + 5.
In general, order makes a difference; that is, in most cases f ◦ g 6= g ◦ f .
Example 8. Sometimes it is the case that f ◦ g = g ◦ f . Let f (x) = x3 and
√
g(x) = 3 x. Then
√ √
(f ◦ g)(x) = f (g(x)) = f ( 3 x ) = ( 3 x )3 = x
and √
3
(g ◦ f )(x) = g(f (x)) = g(x3 ) = x3 = x.
Example 9. Given a 2 × 2 matrix
a b
A= ,
c d
π(1) = 2
π(2) = 1
π(3) = 3.
12 CHAPTER 0 PRELIMINARIES
Proof. We will prove (1) and (3). Part (2) is left as an exercise. Part (4)
follows directly from (2) and (3).
(1) We must show that
h ◦ (g ◦ f ) = (h ◦ g) ◦ f.
For a ∈ A we have
(3) Assume that f and g are both onto functions. Given c ∈ C, we must
show that there exists an a ∈ A such that (g ◦f )(a) = g(f (a)) = c. However,
since g is onto, there is a b ∈ B such that g(b) = c. Similarly, there is an
a ∈ A such that f (a) = b. Accordingly,
If S is any set, we will use idS or id to denote the identity mapping
from S to itself. Define this map by id(s) = s for all s ∈ S. A map g : B → A
is an inverse mapping of f : A → B if g ◦f = idA and f ◦g = idB ; in other
0.2 SETS AND EQUIVALENCE RELATIONS 13
f (f −1 (x)) = f (ex ) = ln ex = x
and
f −1 (f (x)) = f −1 (ln x) = eln x = x
whenever composition makes sense.
Example 13. Suppose that
3 1
A= .
5 2
We can find an inverse map of TA by simply inverting the matrix A; that is,
TA−1 = TA−1 . In this example,
−1 2 −1
A = ;
−5 3
TB (x, y) = (3x, 0)
14 CHAPTER 0 PRELIMINARIES
and
(x, y) = T ◦ TB−1 (x, y) = (3ax + 3by, 0)
and
−18 33
B= ,
−11 20
then A ∼ B since P AP −1 = B for
2 5
P = .
1 3
A = P −1 BP = P −1 B(P −1 )−1 .
the relation is transitive. Two matrices that are equivalent in this manner
are said to be similar.
A partition P of a set X is a collection
S of nonempty sets X1 , X2 , . . .
such that Xi ∩ Xj = ∅ for i 6= j and k Xk = X. Let ∼ be an equivalence
relation on a set X and let x ∈ X. Then [x] = {y ∈ X : y ∼ x} is called the
equivalence class of x. We will see that an equivalence relation gives rise
to a partition via equivalence classes. Also, whenever a partition of a set
exists, there is some natural underlying equivalence relation, as the following
theorem demonstrates.
[x] = [y] or [x] ∩ [y] = ∅. Suppose that the intersection of [x] and [y] is not
empty and that z ∈ [x] ∩ [y]. Then z ∼ x and z ∼ y. By symmetry and
transitivity x ∼ y; hence, [x] ⊂ [y]. Similarly, [y] ⊂ [x] and so [x] = [y].
Therefore, any two equivalence classes are either disjoint or exactly the same.
Conversely, suppose that P = {Xi } is a partition of a set X. Let two
elements be equivalent if they are in the same partition. Clearly, the relation
is reflexive. If x is in the same partition as y, then y is in the same partition
as x, so x ∼ y implies y ∼ x. Finally, if x is in the same partition as y and
y is in the same partition as z, then x must be in the same partition as z,
and transitivity holds.
r − t = r − s + s − t = kn + ln = (k + l)n,
and so r − t is divisible by n.
18 CHAPTER 0 PRELIMINARIES
Notice that [0] ∪ [1] ∪ [2] = Z and also that the sets are disjoint. The sets
[0], [1], and [2] form a partition of the integers.
The integers modulo n are a very important example in the study of
abstract algebra and will become quite useful in our investigation of vari-
ous algebraic structures such as groups and rings. In our discussion of the
integers modulo n we have actually assumed a result known as the division
algorithm, which will be stated and proved in Chapter 1.
Exercises
1. Suppose that
A = {x : x ∈ N and x is even},
B = {x : x ∈ N and x is prime},
C = {x : x ∈ N and x is a multiple of 5}.
(a) A ∩ B (c) A ∪ B
(b) B ∩ C (d) A ∩ (B ∪ C)
2. If A = {a, b, c}, B = {1, 2, 3}, C = {x}, and D = ∅, list all of the elements in
each of the following sets.
(a) A × B (c) A × B × C
(b) B × A (d) A × D
7. Prove A ∩ (B ∪ C) = (A ∩ B) ∪ (A ∩ C).
8. Prove A ⊂ B if and only if A ∩ B = A.
9. Prove (A ∩ B)0 = A0 ∪ B 0 .
10. Prove A ∪ B = (A ∩ B) ∪ (A \ B) ∪ (B \ A).
11. Prove (A ∪ B) × C = (A × C) ∪ (B × C).
12. Prove (A ∩ B) \ B = ∅.
13. Prove (A ∪ B) \ B = A \ B.
14. Prove A \ (B ∪ C) = (A \ B) ∩ (A \ C).
15. Prove A ∩ (B \ C) = (A ∩ B) \ (A ∩ C).
16. Prove (A \ B) ∪ (B \ C) = (A ∪ B) \ (A ∩ B).
17. Which of the following relations f : Q → Q define a mapping? In each case,
supply a reason why f is or is not a mapping.
p+1 p+q
(a) f (p/q) = (c) f (p/q) =
p−2 q2
3p 3p2 p
(b) f (p/q) = (d) f (p/q) = 2
−
3q 7q q
18. Determine which of the following functions are one-to-one and which are
onto. If the function is not onto, determine its range.
(a) f : R → R defined by f (x) = ex
(b) f : Z → Z defined by f (n) = n2 + 3
(c) f : R → R defined by f (x) = sin x
(d) f : Z → Z defined by f (x) = x2
19. Let f : A → B and g : B → C be invertible mappings; that is, mappings
such that f −1 and g −1 exist. Show that (g ◦ f )−1 = f −1 ◦ g −1 .
20. (a) Define a function f : N → N that is one-to-one but not onto.
(b) Define a function f : N → N that is onto but not one-to-one.
21. Prove the relation defined on R2 by (x1 , y1 ) ∼ (x2 , y2 ) if x21 + y12 = x22 + y22 is
an equivalence relation.
22. Let f : A → B and g : B → C be maps.
(a) If f and g are both one-to-one functions, show that g ◦ f is one-to-one.
(b) If g ◦ f is onto, show that g is onto.
(c) If g ◦ f is one-to-one, show that f is one-to-one.
(d) If g ◦ f is one-to-one and f is onto, show that g is one-to-one.
20 CHAPTER 0 PRELIMINARIES
x+1
f (x) = .
x−1
What are the domain and range of f ? What is the inverse of f ? Compute
f ◦ f −1 and f −1 ◦ f .
25. Determine whether or not the following relations are equivalence relations on
the given set. If the relation is an equivalence relation, describe the partition
given by it. If the relation is not an equivalence relation, state why it fails to
be one.
(a) x ∼ y in R if x ≥ y (c) x ∼ y in R if |x − y| ≤ 4
(b) m ∼ n in Z if mn > 0 (d) m ∼ n in Z if m ≡ n (mod 6)
26. Define a relation ∼ on R2 by stating that (a, b) ∼ (c, d) if and only if a2 +b2 ≤
c2 + d2 . Show that ∼ is reflexive and transitive but not symmetric.
22
1.1 MATHEMATICAL INDUCTION 23
8 = 23 > 3 + 4 = 7,
2(k + 4) = 2k + 8 > k + 5 = (k + 1) + 4
since k is positive. Hence, by induction, the statement holds for all integers
n ≥ 3.
Example 2. Every integer 10n+1 + 3 · 10n + 5 is divisible by 9 for n ∈ N.
For n = 1,
101+1 + 3 · 10 + 5 = 135 = 9 · 15
is divisible by 9. Suppose that 10k+1 + 3 · 10k + 5 is divisible by 9 for k ≥ 1.
Then
is divisible by 9.
Example 3. We will prove the binomial theorem using mathematical in-
duction; that is,
n
n
X n
(a + b) = ak bn−k ,
k
k=0
24 CHAPTER 1 THE INTEGERS
If n = 1, the binomial theorem is easy to verify. Now assume that the result
is true for n greater than or equal to 1. Then
1.1 MATHEMATICAL INDUCTION 25
a = bq + r
where 0 ≤ r < b.
S = {a − bk : k ∈ Z and a − bk ≥ 0}.
a − b(q + 1) = a − bq − b = r − b > 0.
In this case we would have a − b(q + 1) in the set S. But then a − b(q + 1) <
a−bq, which would contradict the fact that r = a−bq is the smallest member
of S. So r ≤ b. Since 0 ∈
/ S, r 6= b and so r < b.
Uniqueness of q and r. Suppose there exist integers r, r0 , q, and q 0 such
that
a = bq + r, 0 ≤ r < b
and
a = bq 0 + r0 , 0 ≤ r0 < b.
1.2 THE DIVISION ALGORITHM 27
Theorem 1.4 Let a and b be nonzero integers. Then there exist integers r
and s such that
gcd(a, b) = ar + bs.
Furthermore, the greatest common divisor of a and b is unique.
Proof. Let
S = {am + bn : m, n ∈ Z and am + bn > 0}.
Clearly, the set S is nonempty; hence, by the Well-Ordering Principle S
must have a smallest member, say d = ar + bs. We claim that d = gcd(a, b).
Write a = dq + r where 0 ≤ r < d . If r > 0, then
r = a − dq
= a − (ar + bs)q
= a − arq − bsq
= a(1 − rq) + b(−sq),
which is in S. But this would contradict the fact that d is the smallest
member of S. Hence, r = 0 and d divides a. A similar argument shows that
d divides b. Therefore, d is a common divisor of a and b.
Suppose that d0 is another common divisor of a and b, and we want to
show that d0 | d. If we let a = d0 h and b = d0 k, then
d = ar + bs = d0 hr + d0 ks = d0 (hr + ks).
So d0 must divide d. Hence, d must be the unique greatest common divisor
of a and b.
Corollary 1.5 Let a and b be two integers that are relatively prime. Then
there exist integers r and s such that ar + bs = 1.
28 CHAPTER 1 THE INTEGERS
Example 4. Let us compute the greatest common divisor of 945 and 2415.
First observe that
Reversing our steps, 105 divides 420, 105 divides 525, 105 divides 945, and
105 divides 2415. Hence, 105 divides both 945 and 2415. If d were another
common divisor of 945 and 2415, then d would also have to divide 105.
Therefore, gcd(945, 2415) = 105.
If we work backward through the above sequence of equations, we can
also obtain numbers r and s such that 945r + 2415s = 105. Observe that
b = aq1 + r1
a = r1 q2 + r2
r1 = r2 q3 + r3
..
.
rn−2 = rn−1 qn + rn
rn−1 = rn qn+1 .
1.2 THE DIVISION ALGORITHM 29
To find r and s such that ar + bs = d, we begin with this last equation and
substitute results obtained from the previous equations:
d = rn
= rn−2 − rn−1 qn
= rn−2 − qn (rn−3 − qn−1 rn−2 )
= −qn rn−3 + (1 + qn qn−1 )rn−2
..
.
= ra + sb.
The algorithm that we have just used to find the greatest common divisor
d of two integers a and b and to write d as the linear combination of a and
b is known as the Euclidean algorithm.
Prime Numbers
Let p be an integer such that p > 1. We say that p is a prime number, or
simply p is prime, if the only positive numbers that divide p are 1 and p
itself. An integer n > 1 that is not prime is said to be composite.
Proof. Suppose that p does not divide a. We must show that p | b. Since
gcd(a, p) = 1, there exist integers r and s such that ar + ps = 1. So
n = p1 p2 · · · pk ,
n = q1 q2 · · · ql ,
n = p1 p2 · · · pk = q1 q2 · · · ql ,
n0 = p2 · · · pk = q2 · · · ql
a1 = p 1 · · · p r
a2 = q1 · · · qs .
Therefore,
a = a1 a2 = p1 · · · pr q1 · · · qs .
So a ∈
/ S, which is a contradiction.
Historical Note
EXERCISES 31
Prime numbers were first studied by the ancient Greeks. Two important results
from antiquity are Euclid’s proof that an infinite number of primes exist and the
Sieve of Eratosthenes, a method of computing all of the prime numbers less than a
fixed positive integer n. One problem in number theory is to find a function f such
that f (n) is prime for each integer n. Pierre Fermat (1601?–1665) conjectured that
n
22 + 1 was prime for all n, but later it was shown by Leonhard Euler (1707–1783)
that 5
22 + 1 = 4,294,967,297
is a composite number. One of the many unproven conjectures about prime numbers
is Goldbach’s Conjecture. In a letter to Euler in 1742, Christian Goldbach stated
the conjecture that every even integer with the exception of 2 seemed to be the sum
of two primes: 4 = 2 + 2, 6 = 3 + 3, 8 = 3 + 5, . . .. Although the conjecture has been
verified for the numbers up through 100 million, it has yet to be proven in general.
Since prime numbers play an important role in public key cryptography, there is
currently a great deal of interest in determining whether or not a large number is
prime.
Exercises
1. Prove that
n(n + 1)(2n + 1)
12 + 22 + · · · + n2 =
6
for n ∈ N.
2. Prove that
n2 (n + 1)2
1 3 + 2 3 + · · · + n3 =
4
for n ∈ N.
3. Prove that n! > 2n for n ≥ 4.
4. Prove that
n(3n − 1)x
x + 4x + 7x + · · · + (3n − 2)x =
2
for n ∈ N.
5. Prove that 10n+1 + 10n + 1 is divisible by 3 for n ∈ N.
6. Prove that 4 · 102n + 9 · 102n−1 + 5 is divisible by 99 for n ∈ N.
7. Show that
n
√ 1X
n
a1 a2 · · · an ≤ ak .
n
k=1
(n) (n)
8. Prove the Leibniz rule for f (x), where f is the nth derivative of f ; that
is, show that
n
X n
(f g)(n) (x) = f (k) (x)g (n−k) (x).
k
k=0
32 CHAPTER 1 THE INTEGERS
For every positive integer n, show that a set with exactly n elements has a
power set with exactly 2n elements.
13. Prove that the two principles of mathematical induction stated in Section 1.1
are equivalent.
14. Show that the Principle of Well-Ordering for the natural numbers implies
that 1 is the smallest natural number. Use this result to show that the
Principle of Well-Ordering implies the Principle of Mathematical Induction;
that is, show that if S ⊂ N such that 1 ∈ S and n + 1 ∈ S whenever n ∈ S,
then S = N.
15. For each of the following pairs of numbers a and b, calculate gcd(a, b) and
find integers r and s such that gcd(a, b) = ra + sb.
16. Let a and b be nonzero integers. If there exist integers r and s such that
ar + bs = 1, show that a and b are relatively prime.
17. Fibonacci Numbers. The Fibonacci numbers are
1, 1, 2, 3, 5, 8, 13, 21, . . . .
18. Let a and b be integers such that gcd(a, b) = 1. Let r and s be integers such
that ar + bs = 1. Prove that
20. Using the division algorithm, show that every perfect square is of the form
4k or 4k + 1 for some nonnegative integer k.
a2 + b2 = r2
2 2
a −b = s2 .
22. Let n ∈ N. Use the division algorithm to prove that every integer is congruent
mod n to precisely one of the integers 0, 1, . . . , n − 1. Conclude that if r is
an integer, then there is exactly one s in Z such that 0 ≤ s < n and [r] = [s].
Hence, the integers are indeed partitioned by congruence mod n.
23. Define the least common multiple of two nonzero integers a and b,
denoted by lcm(a, b), to be the nonnegative integer m such that both a and
b divide m, and if a and b divide any other integer n, then m also divides n.
Prove that any two integers a and b have a unique least common multiple.
26. Prove that gcd(a, c) = gcd(b, c) = 1 if and only if gcd(ab, c) = 1 for integers
a, b, and c.
29. Prove that there are an infinite number of primes of the form 6n + 1.
30. Prove that there are an infinite number of primes of the form 4n − 1.
Programming Exercises
1. The Sieve of Eratosthenes. One method of computing all of the prime
numbers less than a certain fixed positive integer N is to list all of the numbers
n such that 1 < n < N . Begin by eliminating all of the multiples of 2. Next
eliminate all of the multiples of 3. Now eliminate all of the multiples of 5.
Notice that 4 has already been crossed out. Continue in this manner, √ noticing
that we do not have to go all the way to N ; it suffices to stop at N . Using
this method, compute all of the prime numbers less than N = 250. We
can also use this method to find all of the integers that are relatively prime
to an integer N . Simply eliminate the prime factors of N and all of their
multiples. Using this method, find all of the numbers that are relatively
prime to N = 120. Using the Sieve of Eratosthenes, write a program that
will compute all of the primes less than an integer N .
2. Let N0 = N ∪ {0}. Ackermann’s function is the function A : N0 × N0 → N0
defined by the equations
A(0, y) = y + 1,
A(x + 1, 0) = A(x, 1),
A(x + 1, y + 1) = A(x, A(x + 1, y)).
Use this definition to compute A(3, 1). Write a program to evaluate Ack-
ermann’s function. Modify the program to count the number of statements
executed in the program when Ackermann’s function is evaluated. How many
statements are executed in the evaluation of A(4, 1)? What about A(5, 1)?
3. Write a computer program that will implement the Euclidean algorithm.
The program should accept two positive integers a and b as input and should
output gcd(a, b) as well as integers r and s such that
gcd(a, b) = ra + sb.
35
36 CHAPTER 2 GROUPS
7 + 4 ≡ 1 (mod 5) 7 · 3 ≡ 1 (mod 5)
3 + 5 ≡ 0 (mod 8) 3 · 5 ≡ 7 (mod 8)
3 + 4 ≡ 7 (mod 12) 3 · 4 ≡ 0 (mod 12).
Example 2. Most, but not all, of the usual laws of arithmetic hold for
addition and multiplication in Zn . For instance, it is not necessarily true
that there is a multiplicative inverse. Consider the multiplication table for
Z8 in Table 2.1. Notice that 2, 4, and 6 do not have multiplicative inverses;
that is, for n = 2, 4, or 6, there is no integer k such that kn ≡ 1 (mod 8).
2.1 THE INTEGERS MOD N AND SYMMETRIES 37
Proof. We will prove (1) and (6) and leave the remaining properties to be
proven in the exercises.
(1) Addition and multiplication are commutative modulo n since the
remainder of a + b divided by n is the same as the remainder of b + a divided
by n.
(6) Suppose that gcd(a, n) = 1. Then there exist integers r and s such
that ar + ns = 1. Since ns = 1 − ar, ra ≡ 1 (mod n). Letting b be the
equivalence class of r, ab ≡ 1 (mod n).
Conversely, suppose that there exists a b such that ab ≡ 1 (mod n).
Then n divides ab − 1, so there is an integer k such that ab − nk = 1. Let
d = gcd(a, n). Since d divides ab − nk, d must also divide 1; hence, d = 1.
38 CHAPTER 2 GROUPS
Symmetries
A symmetry of a geometric figure is a rearrangement of the figure preserv-
ing the arrangement of its sides and vertices as well as its distances and
angles. A map from the plane to itself preserving the symmetry of an object
is called a rigid motion. For example, if we look at the rectangle in Fig-
ure 2.1, it is easy to see that a rotation of 180◦ or 360◦ returns a rectangle in
the plane with the same orientation as the original rectangle and the same
relationship among the vertices. A reflection of the rectangle across either
the vertical axis or the horizontal axis can also be seen to be a symmetry.
However, a 90◦ rotation in either direction cannot be a symmetry unless the
rectangle is a square.
A B A B
identity
-
D C D C
A B C D
180◦
-
rotation
D C B A
A B B A
reflection
-
vertical
D C axis C D
A B D C
reflection
-
horizontal
D C axis A B
B B
T identity
- T
A B C
T T id =
T T A B C
A B C A A C
T rotation
- T
A B C
T T ρ1 =
T T B C A
A B C C C B
T rotation
- T
A B C
T T ρ2 =
T T C A B
A B C B C A
T reflection
- T
A B C
T T µ1 =
T T A C B
A B C A B B
T reflection
- T
A B C
T T µ2 =
T T C B A
A B C C A A
T reflection
- T
A B C
T T µ3 =
T T B A C
A C B C
(a ◦ b) ◦ c = a ◦ (b ◦ c)
for a, b, c ∈ G.
2.2 DEFINITIONS AND EXAMPLES 41
e ◦ a = a ◦ e = a.
a ◦ a−1 = a−1 ◦ a = e.
0·2 = 0 1·2 = 2
2·2 = 4 3·2 = 0
4·2 = 2 5 · 2 = 4.
matrix
a b
A=
c d
is in GL2 (R) if there exists a matrix A−1 such that AA−1 = A−1 A = I,
where I is the 2 × 2 identity matrix. For A to have an inverse is equivalent
to requiring that the determinant of A be nonzero; that is, det A = ad−bc 6=
0. The set of invertible matrices forms a group called the general linear
group. The identity of the group is the identity matrix
1 0
I= .
0 1
The inverse of A ∈ GL2 (R) is
−1 1 d −b
A = .
ad − bc −c a
The product of two invertible matrices is again invertible. Matrix multipli-
cation is associative, satisfying the other group axiom. For matrices it is
not true in general that AB 6= BA; hence, GL2 (R) is another example of a
nonabelian group.
Example 8. Let
1 0
1 =
0 1
0 1
I =
−1 0
0 i
J =
i 0
i 0
K = ,
0 −i
where i2 = −1. Then the relations I 2 = J 2 = K 2 = −1, IJ = K, JK = I,
KI = J, JI = −K, KJ = −I, and IK = −J hold. The set Q8 =
{±1, ±I, ±J, ±K} is a group called the quaternion group. Notice that Q8
is noncommutative.
Example 9. Let C∗ be the set of nonzero complex numbers. Under the
operation of multiplication C∗ forms a group. The identity is 1. If z = a + bi
is a nonzero complex number, then
a − bi
z −1 =
a2 + b2
44 CHAPTER 2 GROUPS
is the inverse of z. It is easy to see that the remaining group axioms hold.
A group is finite, or has finite order, if it contains a finite number of
elements; otherwise, the group is said to be infinite or to have infinite
order. The order of a finite group is the number of elements that it con-
tains. If G is a group containing n elements, we write |G| = n. The group
Z5 is a finite group of order 5; the integers Z form an infinite group under
addition, and we sometimes write |Z| = ∞.
2.2 DEFINITIONS AND EXAMPLES 45
It makes sense to write equations with group elements and group opera-
tions. If a and b are two elements in a group G, does there exist an element
x ∈ G such that ax = b? If such an x does exist, is it unique? The following
proposition answers both of these questions positively.
This proposition tells us that the right and left cancellation laws
are true in groups. We leave the proof as an exercise.
We can use exponential notation for groups just as we do in ordinary
algebra. If G is a group and g ∈ G, then we define g 0 = e. For n ∈ N, we
define
gn = g · g · · · g
| {z }
n times
and
g −n = g −1 · g −1 · · · g −1 .
| {z }
n times
Theorem 2.8 In a group, the usual laws of exponents hold; that is, for all
g, h ∈ G,
1. g m g n = g m+n for all m, n ∈ Z;
2. (g m )n = g mn for all m, n ∈ Z;
3. (gh)n = (h−1 g −1 )−n for all n ∈ Z. Furthermore, if G is abelian, then
(gh)n = g n hn .
It is important to realize that the last statement can be made only because
Z and Zn are commutative groups.
Historical Note
Although the first clear axiomatic definition of a group was not given until the
late 1800s, group-theoretic methods had been employed before this time in the
development of many areas of mathematics, including geometry and the theory of
algebraic equations.
Joseph-Louis Lagrange used group-theoretic methods in a 1770–1771 memoir to
study methods of solving polynomial equations. Later, Évariste Galois (1811–1832)
succeeded in developing the mathematics necessary to determine exactly which
polynomial equations could be solved in terms of the polynomials’ coefficients.
Galois’ primary tool was group theory.
The study of geometry was revolutionized in 1872 when Felix Klein proposed
that geometric spaces should be studied by examining those properties that are
invariant under a transformation of the space. Sophus Lie, a contemporary of
Klein, used group theory to study solutions of partial differential equations. One of
the first modern treatments of group theory appeared in William Burnside’s The
Theory of Groups of Finite Order [1], first published in 1897.
2.3 Subgroups
Definitions and Examples
Sometimes we wish to investigate smaller groups sitting inside a larger group.
The set of even integers 2Z = {. . . , −2, 0, 2, 4, . . .} is a group under the
operation of addition. This smaller group sits naturally inside of the group
of integers under addition. We define a subgroup H of a group G to be a
subset H of G such that when the group operation of G is restricted to H,
H is a group in its own right. Observe that every group G with at least two
elements will always have at least two subgroups, the subgroup consisting of
the identity element alone and the entire group itself. The subgroup H = {e}
of a group G is called the trivial subgroup. A subgroup that is a proper
subset of G is called a proper subgroup. In many of the examples that we
2.3 SUBGROUPS 47
have investigated up to this point, there exist other subgroups besides the
trivial and improper subgroups.
Example 10. Consider the set of nonzero real numbers, R∗ , with the group
operation of multiplication. The identity of this group is 1 and the inverse
of any element a ∈ R∗ is just 1/a. We will show that
1. The identity e of G is in H.
2. If h1 , h2 ∈ H, then h1 h2 ∈ H.
3. If h ∈ H, then h−1 ∈ H.
Exercises
1. Find all x ∈ Z satisfying each of the following equations.
(a) ◦ a b c d (b) ◦ a b c d
a a c d a a a b c d
b b b c d b b a d c
c c d a b c c d a b
d d a b c d d c b a
(c) ◦ a b c d (d) ◦ a b c d
a a b c d a a b c d
b b c d a b b a c d
c c d a b c c b a d
d d a b c d d d b c
50 CHAPTER 2 GROUPS
3. Write out Cayley tables for groups formed by the symmetries of a rectangle
and for (Z4 , +). How many elements are in each group? Are the groups the
same? Why or why not?
4. Describe the symmetries of a rhombus and prove that the set of symmetries
forms a group. Give Cayley tables for both the symmetries of a rectangle
and the symmetries of a rhombus. Are the symmetries of a rectangle and
those of a rhombus the same?
5. Describe the symmetries of a square and prove that the set of symmetries
is a group. Give a Cayley table for the symmetries. How many ways can
the vertices of a square be permuted? Is each permutation necessarily a
symmetry of the square? The symmetry group of the square is denoted by
D4 .
6. Give a multiplication table for the group U (12).
7. Let S = R \ {−1} and define a binary operation on S by a ∗ b = a + b + ab.
Prove that (S, ∗) is an abelian group.
8. Give an example of two elements A and B in GL2 (R) with AB 6= BA.
9. Prove that the product of two matrices in SL2 (R) has determinant one.
10. Prove that the set of matrices of the form
1 x y
0 1 z
0 0 1
1 x0 y 0 1 x + x0 y + y 0 + xz 0
1 x y
0 1 z 0 1 z0 = 0 1 z + z0 .
0 0 1 0 0 1 0 0 1
11. Prove that det(AB) = det(A) det(B) in GL2 (R). Use this result to show
that the binary operation in the group GL2 (R) is closed; that is, if A and B
are in GL2 (R), then AB ∈ GL2 (R).
12. Let Zn2 = {(a1 , a2 , . . . , an ) : ai ∈ Z2 }. Define a binary operation on Zn2 by
Prove that Zn2 is a group under this operation. This group is important in
algebraic coding theory.
13. Show that R∗ = R \ {0} is a group under the operation of multiplication.
EXERCISES 51
22. Show that addition and multiplication mod n are associative operations.
23. Show that multiplication distributes over addition modulo n:
24. Let a and b be elements in a group G. Prove that abn a−1 = (aba−1 )n .
25. Let U (n) be the group of units in Zn . If n > 2, prove that there is an element
k ∈ U (n) such that k 2 = 1 and k 6= 1.
−1
26. Prove that the inverse of g1 g2 · · · gn is gn−1 gn−1 · · · g1−1 .
27. Prove Theorem 2.6: if G is a group and a, b ∈ G, then the equations ax = b
and xa = b have unique solutions in G.
28. Prove the right and left cancellation laws for a group G; that is, show that
in the group G, ba = ca implies b = c and ab = ac implies b = c for elements
a, b, c ∈ G.
29. Show that if a2 = e for all a ∈ G, then G must be an abelian group.
30. Show that if G is a finite group of even order, then there is an a ∈ G such
that a is not the identity and a2 = e.
31. Let G be a group and suppose that (ab)2 = a2 b2 for all a and b in G. Prove
that G is an abelian group.
52 CHAPTER 2 GROUPS
32. Find all the subgroups of Z3 × Z3 . Use this information to show that Z3 × Z3
is not the same group as Z9 .
33. Find all the subgroups of the symmetry group of an equilateral triangle.
34. Compute the subgroups of the symmetry group of a square.
35. Let H = {2k : k ∈ Z}. Show that H is a subgroup of Q∗ .
36. Let n = 0, 1, 2, . . . and nZ = {nk : k ∈ Z}. Prove that nZ is a subgroup of
Z. Show that these subgroups are the only subgroups of Z.
37. Let T = {z ∈ C∗ : |z| = 1}. Prove that T is a subgroup of C∗ .
38. Let G consist of the 2 × 2 matrices of the form
cos θ − sin θ
sin θ cos θ
(a) Show that the UPC number 0-50000-30042-6, which appears in Fig-
ure 2.3, is a valid UPC number.
(b) Show that the number 0-50000-30043-6 is not a valid UPC number.
(c) Write a formula to calculate the check digit, d12 , in the UPC number.
(d) The UPC error detection scheme can detect most transposition errors;
that is, it can determine if two digits have been interchanged. Show
that the transposition error 0-05000-30042-6 is detected. Find a trans-
position error that is not detected.
(e) Write a program that will determine whether or not a UPC number is
valid.
54 CHAPTER 2 GROUPS
0 50000 30042 6
2. It is often useful to use an inner product notation for this type of error
detection scheme; hence, we will use the notion
to mean
d1 w1 + d2 w2 + · · · + dk wk ≡ 0 (mod n).
One problem is that d10 might have to be a 10 to make the inner product zero;
in this case, 11 digits would be needed to make this scheme work. Therefore,
the character X is used for the eleventh digit. So ISBN 3-540-96035-X is a
valid ISBN code.
(a) Is ISBN 0-534-91500-0 a valid ISBN code? What about ISBN 0-534-
91700-0 and ISBN 0-534-19500-0?
(b) Does this method detect all single-digit errors? What about all trans-
position errors?
EXERCISES 55
The groups Z and Zn , which are among the most familiar and easily under-
stood groups, are both examples of what are called cyclic groups. In this
chapter we will study the properties of cyclic groups and cyclic subgroups,
which play a fundamental part in the classification of all abelian groups.
56
3.1 CYCLIC SUBGROUPS 57
Proof. The identity is in hai since a0 = e. If g and h are any two elements
in hai, then by the definition of hai we can write g = am and h = an for some
integers m and n. So gh = am an = am+n is again in hai. Finally, if g = an
in hai, then the inverse g −1 = a−n is also in hai. Clearly, any subgroup H
of G containing a must contain all the powers of a by closure; hence, H
contains hai. Therefore, hai is the smallest subgroup of G containing a.
Remark. If we are using the “+” notation, as in the case of the integers
under addition, we write hai = {na : n ∈ Z}.
Example 3. Notice that a cyclic group can have more than a single gen-
erator. Both 1 and 5 generate Z6 ; hence, Z6 is a cyclic group. Not every
element in a cyclic group is necessarily a generator of the group. The order
of 2 ∈ Z6 is 3. The cyclic subgroup generated by 2 is h2i = {0, 2, 4}.
The groups Z and Zn are cyclic groups. The elements 1 and −1 are
generators for Z. We can certainly generate Zn with 1 although there may
be other generators of Zn , as in the case of Z6 .
21 = 2 22 = 4
23 = 8 24 = 7
25 = 5 26 = 1.
S3
!! a
a
! !! S aa
aa
! S
!! S aa
{id, ρ1 , ρ2 } {id, µ1 } {id, µ2 } {id, µ3 }
aa !
aa S
! !!
aa S !!
aaS !
!
{id}
Proof. The main tools used in this proof are the division algorithm and
the Principle of Well-Ordering. Let G be a cyclic group generated by a and
suppose that H is a subgroup of G. If H = {e}, then trivially H is cyclic.
Suppose that H contains some other element g distinct from the identity.
Then g can be written as an for some integer n. We can assume that n > 0.
Let m be the smallest natural number such that am ∈ H. Such an m exists
by the Principle of Well-Ordering.
We claim that h = am is a generator for H. We must show that every
h0 ∈ H can be written as a power of h. Since h0 ∈ H and H is a subgroup
of G, h0 = ak for some positive integer k. Using the division algorithm, we
can find numbers q and r such that k = mq + r where 0 ≤ r < m; hence,
ak = amq+r = (am )q ar = hq ar .
3.1 CYCLIC SUBGROUPS 59
C = {a + bi : a, b ∈ R},
z + w = (2 + 3i) + (1 − 2i) = 3 + i
and
zw = (2 + 3i)(1 − 2i) = 8 − i.
Also,
2 3
z −1 = − i
13 13
√
|z| = 13
z = 2 − 3i.
y
z 1 = 2 + 3i
z 3 = –3 + 2i
0 x
z 2 = 1 – 2i
a + bi
θ
0 x
plane where a is the x (or real) coordinate and b is the y (or imaginary)
coordinate. This is called the rectangular or Cartesian representation.
The rectangular representations of z1 = 2 + 3i, z2 = 1 − 2i, and z3 = −3 + 2i
are depicted in Figure 3.2.
Nonzero complex numbers can also be represented using polar coordi-
nates. To specify any nonzero point on the plane, it suffices to give an angle
θ from the positive x axis in the counterclockwise direction and a distance
r from the origin, as in Figure 3.3. We can see that
Hence,
p
r = |z| = a2 + b2
62 CHAPTER 3 CYCLIC GROUPS
and
a = r cos θ
b = r sin θ.
a = 2 cos 60◦ = 1
and √
b = 2 sin 60◦ = 3.
√
Hence, the rectangular representation is z = 1 + 3 i.
Conversely, if we are given a rectangular representation of a complex
number,
√ it is√ often useful to know the number’s polar representation. If
z = 3 2 − 3 2 i, then
p √
r = a2 + b2 = 36 = 6
and
b
θ = arctan = arctan(−1) = 315◦ ,
a
√ √
so 3 2 − 3 2 i = 6 cis 315◦ .
Although the circle group has infinite order, it has many interesting finite
subgroups. Suppose that H = {1, −1, i, −i}. Then H is a subgroup of the
circle group. Also, 1, −1, i, and −i are exactly those complex numbers that
satisfy the equation z 4 = 1. The complex numbers satisfying the equation
z n = 1 are called the nth roots of unity.
64 CHAPTER 3 CYCLIC GROUPS
The z’s are distinct since the numbers 2kπ/n are all distinct and are greater
than or equal to 0 but less than 2π. The fact that these are all of the roots
of the equation z n = 1 follows from the Fundamental Theorem of Algebra
(Theorem 19.16), which states that a polynomial of degree n can have at
most n roots. We will leave the proof that the nth roots of unity form a
cyclic subgroup of T as an exercise.
A generator for the group of the nth roots of unity is called a primitive
nth root of unity.
Example 11. The 8th roots of unity can be represented as eight equally
spaced points on the unit circle (Figure 3.4). The primitive 8th roots of
unity are
√ √
2 2
ω = + i
2√ 2√
2 2
ω3 = − + i
√2 √2
2 2
ω5 = − − i
√ 2 √ 2
2 2
ω7 = − i.
2 2
i y
ω3 ω
–1 0 1 x
ω5 ω7
–i
1000000
22 .
However, such numbers are so large that we do not want to attempt the
calculations; moreover,past a certain point the computations would not be
feasible even if we had every computer in the world at our disposal. Even
writing down the decimal representation of a very large number may not be
reasonable. It could be thousands or even millions of digits long. However,
if we could compute something like 237398332 (mod 46389), we could very
easily write the result down since it would be a number between 0 and
46,388. If we want to compute powers modulo n quickly and efficiently, we
will have to be clever.
The first thing to notice is that any number a can be written as the sum
of distinct powers of 2; that is, we can write
k multiplications by computing
0
a2 (mod n)
1
a2 (mod n)
..
.
k
a2 (mod n).
Each step involves squaring the answer obtained in the previous step, divid-
ing by n, and taking the remainder.
Example 12. We will compute 271321 (mod 481). Notice that
321 = 20 + 26 + 28 ;
hence, computing 271321 (mod 481) is the same as computing
0 +26 +28 0 6 8
2712 ≡ 2712 · 2712 · 2712 (mod 481).
So it will suffice to compute 271 2i (mod 481) where i = 0, 6, 8. It is very
easy to see that
1
2712 ≡ 73, 441 (mod 481)
≡ 329 (mod 481).
2
We can square this result to obtain a value for 2712 (mod 481):
2 1
2712 ≡ (2712 )2 (mod 481)
2
≡ (329) (mod 481)
≡ 1, 082, 411 (mod 481)
≡ 16 (mod 481).
n n n+1
We are using the fact that (a2 )2 ≡ a2·2 ≡ a2 (mod n). Continuing, we
can calculate
6
2712 ≡ 419 (mod 481)
and
8
2712 ≡ 16 (mod 481).
Therefore,
0 +26 +28
271321 ≡ 2712 (mod 481)
20 26 8
≡ 271 · 271 · 2712 (mod 481)
≡ 271 · 419 · 16 (mod 481)
≡ 1, 816, 784 (mod 481)
≡ 47 (mod 481).
EXERCISES 67
The method of repeated squares will prove to be a very useful tool when
we explore RSA cryptography in Chapter 6. To encode and decode messages
in a reasonable manner under this scheme, it is necessary to be able to
quickly compute large powers of integers mod n.
Exercises
1. Prove or disprove each of the following statements.
4. Find the subgroups of GL2 (R) generated by each of the following matrices.
(a) 0 1 (b) 0 1/3
−1 0 3 0
(c) 1 −1 (d) 1 −1
1 0 0 1
√
1 −1 3/2
(e) (f ) √1/2
−1 0 −1/2 3/2
27. If g and h have orders 15 and 16 respectively in a group G, what is the order
of hgi ∩ hhi?
28. Let a be an element in a group G. What is a generator for the subgroup
ham i ∩ han i?
29. Prove that Zn has an even number of generators for n > 2.
30. Suppose that G is a group and let a, b ∈ G. Prove that if |a| = m and |b| = n
with gcd(m, n) = 1, then hai ∩ hbi = {e}.
31. Let G be an abelian group. Show that the elements of finite order in G form
a subgroup. This subgroup is called the torsion subgroup of G.
32. Let G be a finite cyclic group of order n generated by x. Show that if y = xk
where gcd(k, n) = 1, then y must be a generator of G.
33. If G is an abelian group that contains a pair of cyclic subgroups of order 2,
show that G must contain a subgroup of order 4. Does this subgroup have
to be cyclic?
34. Let G be an abelian group of order pq where gcd(p, q) = 1. If G contains
elements a and b of order p and q respectively, then show that G is cyclic.
35. Prove that the subgroups of Z are exactly nZ for n = 0, 1, 2, . . ..
36. Prove that the generators of Zn are the integers r such that 1 ≤ r < n and
gcd(r, n) = 1.
37. Prove that if G has no proper nontrivial subgroups, then G is a cyclic group.
38. Prove that the order of an element in a cyclic group G must divide the order
of the group.
39. For what integers n is −1 an nth root of unity?
40. If z = r(cos θ + i sin θ) and w = s(cos φ + i sin φ) are two nonzero complex
numbers, show that
Programming Exercises
1. Write a computer program that will write any decimal number as the sum
of distinct powers of 2. What is the largest integer that your program will
handle?
2. Write a computer program to calculate ax (mod n) by the method of re-
peated squares. What are the largest values of n and x that your program
will accept?
A B C A B C A B C
A C B C B A B A C
A 7→ B
B 7→ C
C 7→ A.
72
4.1 DEFINITIONS AND NOTATION 73
Then
1 2 3 4
στ = ,
1 4 3 2
but
1 2 3 4
τσ = .
3 2 1 4
Cycle Notation
The notation that we have used to represent permutations up to this point is
cumbersome, to say the least. To work effectively with permutation groups,
we need a more streamlined method of writing down and manipulating per-
mutations.
A permutation σ ∈ SX is a cycle of length k if there exist elements
a1 , a2 , . . . , ak ∈ X such that
σ(a1 ) = a2
σ(a2 ) = a3
..
.
σ(ak ) = a1
is a cycle of length 3.
Not every permutation is a cycle. Consider the permutation
1 2 3 4 5 6
= (1243)(56).
2 4 1 3 6 5
σ = (1352)
τ = (256).
We can think of σ as
1 7→ 3
3 7→ 5
5 7→ 2
2 7→ 1
and τ as
2 7→ 5
5 7→ 6
6 7→ 2
Example 5. The cycles (135) and (27) are disjoint; however, the cycles
(135) and (347) are not. Calculating their products, we find that
(135)(27) = (135)(27)
(135)(347) = (13475).
The product of two cycles that are not disjoint may reduce to something
less complicated; the product of disjoint cycles cannot be simplified.
a1 7→ a2
a2 7→ a3
..
.
ak−1 7→ ak
ak 7→ a1 .
στ (ai ) = σ(τ (ai )) = σ(ai ) = a(i mod k)+1 = τ (a(i mod k)+1 ) = τ (σ(ai )) = τ σ(ai ).
Example 6. Let
1 2 3 4 5 6
σ =
6 4 3 1 5 2
1 2 3 4 5 6
τ = .
3 2 1 5 6 4
σ = (1624)
τ = (13)(456)
στ = (136)(245)
τ σ = (143)(256).
Remark. From this point forward we will find it convenient to use cycle
notation to represent permutations. When using cycle notation, we often
denote the identity permutation by (1).
Transpositions
The simplest permutation is a cycle of length 2. Such cycles are called
transpositions. Since
(23)(16)(23)
or by
(35)(16)(13)(16)(13)(35)(56),
but (16) will always be the product of an odd number of transpositions.
id = τ1 τ2 · · · τr ,
(ab)(ab) = id
(bc)(ab) = (ab)(ac)
(cd)(ab) = (ab)(cd)
(bc)(ac) = (ab)(bc).
The first equation simply says that a transposition is its own inverse. If
this case occurs, delete τr−1 τr from the product to obtain
id = τ1 τ2 · · · τr−3 τr−2 .
σ = σ1 σ2 · · · σm = τ1 τ2 · · · τn ,
where m is even. We must show that n is also an even number. The inverse
of σ −1 is σm · · · σ1 . Since
id = σσm · · · σ1 = τ1 · · · τn σm · · · σ1 ,
n must be even by Lemma 4.5. The proof for the case in which σ can be
expressed as an odd number of transpositions is left as an exercise.
Proof. Since the product of two even permutations must also be an even
permutation, An is closed. The identity is an even permutation and therefore
is in An . If σ is an even permutation, then
σ = σ1 σ2 · · · σr ,
λ σ : An → B n
by
λσ (τ ) = στ.
Suppose that λσ (τ ) = λσ (µ). Then στ = σµ and so
τ = σ −1 στ = σ −1 σµ = µ.
Historical Note
4.2 THE DIHEDRAL GROUPS 81
Lagrange first thought of permutations as functions from a set to itself, but it was
Cauchy who developed the basic theorems and notation for permutations. He was
the first to use cycle notation. Augustin-Louis Cauchy (1789–1857) was born in
Paris at the height of the French Revolution. His family soon left Paris for the
village of Arcueil to escape the Reign of Terror. One of the family’s neighbors there
was Pierre-Simon Laplace (1749–1827), who encouraged him to seek a career in
mathematics. Cauchy began his career as a mathematician by solving a problem
in geometry given to him by Lagrange. Over 800 papers were written by Cauchy
on such diverse topics as differential equations, finite groups, applied mathematics,
and complex analysis. He was one of the mathematicians responsible for making
calculus rigorous. Perhaps more theorems and concepts in mathematics have the
name Cauchy attached to them than that of any other mathematician.
1
n 2
n-1 3
1 8
8 2 7 1
rotation
7 3 6 2
6 4 5 3
5 4
1 1
8 2 2 8
reflection
7 3 3 7
6 4 4 6
5 5
1 1
6 2 2 6
5 3 3 5
4 4
1 1
5 2 2 5
4 3 3 4
We will denote the rotation 360◦ /n by r. The rotation r generates all of the
other rotations. That is,
360◦
rk = k · .
n
Label the n reflections s1 , s2 , . . . , sn , where sk is the reflection that leaves
vertex k fixed. There are two cases of reflection, depending on whether n
is even or odd. If there are an even number of vertices, then 2 vertices are
left fixed by a reflection. If there are an odd number of vertices, then only
a single vertex is left fixed by a reflection (Figure 4.3). Hence, if n = 2m
for some integer m, then si = si+m for 1 ≤ i < m. The order of sk is two.
Let s = s1 . Then s2 = id and rn = id. Since any rigid motion t of the
n-gon replaces the first vertex by the vertex k, the second vertex must be
replaced by either k + 1 or by k − 1. If it is replaced by k + 1, then t = rk .
If it is replaced by k − 1, then t = rk s. Hence, r and s generate Dn ; that
is, Dn consists of all finite products of r and s. We will leave the proof that
srs = r−1 as an exercise.
1 2
4 3
s1 = (24)
s2 = (13).
rs1 = (12)(34)
r3 s1 = (14)(23).
1 2
4 3
4
3
2 1
1 2 2 1
4 3 4 3
4 4
3 3
2 1 1 2
Proof. From Proposition 4.11, we already know that the motion group of
the cube has 24 elements, the same number of elements as there are in S4 .
There are exactly four diagonals in the cube. If we label these diagonals 1,
2, 3, and 4, we must show that the motion group of the cube will give us
any permutation of the diagonals (Figure 4.5). If we can obtain all of these
permutations, then S4 and the group of rigid motions of the cube must be
the same. To obtain a transposition we can rotate the cube 180◦ about the
axis joining the midpoints of opposite edges (Figure 4.6). There are six such
axes, giving all transpositions in S4 . Since every element in S4 is the product
of a finite number of transpositions, the motion group of a cube must be S4 .
Exercises
(a) 1 2 3 4 5 (b) 1 2 3 4 5
2 4 1 5 3 4 2 5 1 3
(c) 1 2 3 4 5 (d) 1 2 3 4 5
3 5 1 4 2 1 4 3 2 5
(a) {σ ∈ S4 : σ(1) = 3}
(b) {σ ∈ S4 : σ(2) = 2}
(c) {σ ∈ S4 : σ(1) = 3 and σ(2) = 2}
15. If the diagonals of a cube are labeled as Figure 4.5, to which motion of
the cube does the permutation (12)(34) correspond? What about the other
permutations of the diagonals?
16. Find the group of rigid motions of a tetrahedron. Show that this is the same
group as A4 .
17. Prove that Sn is nonabelian for n ≥ 3.
18. Show that An is nonabelian for n ≥ 4.
19. Prove that Dn is nonabelian for n ≥ 3.
20. Let σ ∈ Sn . Prove that σ can be written as the product of at most n − 1
transpositions.
21. Let σ ∈ Sn . If σ is not a cycle, prove that σ can be written as the product
of at most n − 2 transpositions.
22. If σ can be expressed as an odd number of transpositions, show that any
other product of transpositions equaling σ must also be odd.
23. If σ is a cycle of odd length, prove that σ 2 is also a cycle.
24. Show that a 3-cycle is an even permutation.
25. Prove that in An with n ≥ 3, any permutation is a product of cycles of
length 3.
26. Prove that any element in Sn can be written as a finite product of the fol-
lowing permutations.
(a) (12), (13), . . . , (1n)
(b) (12), (23), . . . , (n − 1, n)
(c) (12), (12 . . . n)
27. Let G be a group and define a map λg : G → G by λg (a) = ga. Prove that
λg is a permutation of G.
28. Prove that there exist n! permutations of a set containing n elements.
29. Recall that the center of a group G is
Find the center of D8 . What about the center of D10 ? What is the center of
Dn ?
30. Let τ = (a1 , a2 , . . . , ak ) be a cycle of length k.
(a) Prove that if σ is any permutation, then
is a cycle of length k.
88 CHAPTER 4 PERMUTATION GROUPS
Ox,σ = {y : x ∼ y}.
α = (1254)
β = (123)(45)
γ = (13)(25).
(d) If Ox,σ ∩ Oy,σ 6= ∅, prove that Ox,σ = Oy,σ . The orbits under a permu-
tation σ are the equivalence classes corresponding to the equivalence
relation ∼.
(e) A subgroup H of SX is transitive if for every x, y ∈ X, there exists
a σ ∈ H such that σ(x) = y. Prove that hσi is transitive if and only if
Ox,σ = X for some x ∈ X.
33. Let α ∈ Sn for n ≥ 3. If αβ = βα for all β ∈ Sn , prove that α must be the
identity permutation; hence, the center of Sn is the trivial subgroup.
34. If α is even, prove that α−1 is also even. Does a corresponding result hold if
α is odd?
35. Show that α−1 β −1 αβ is even for α, β ∈ Sn .
36. Let r and s be the elements in Dn described in Theorem 4.10.
(a) Show that srs = r−1 .
(b) Show that rk s = sr−k in Dn .
(c) Prove that the order of rk ∈ Dn is n/ gcd(k, n).
5
Cosets and Lagrange’s
Theorem
Lagrange’s Theorem, one of the most important results in finite group the-
ory, states that the order of a subgroup must divide the order of the group.
This theorem provides a powerful tool for analyzing finite groups; it gives
us an idea of exactly what type of subgroups we might expect a finite group
to possess. Central to understanding Lagranges’s Theorem is the notion of
a coset.
5.1 Cosets
Let G be a group and H a subgroup of G. Define a left coset of H with
representative g ∈ G to be the set
gH = {gh : h ∈ H}.
Right cosets can be defined similarly by
Hg = {hg : h ∈ H}.
If left and right cosets coincide or if it is clear from the context to which type
of coset that we are referring, we will use the word coset without specifying
left or right.
Example 1. Let H be the subgroup of Z6 consisting of the elements 0 and
3. The cosets are
0 + H = 3 + H = {0, 3}
1 + H = 4 + H = {1, 4}
2 + H = 5 + H = {2, 5}.
89
90 CHAPTER 5 COSETS AND LAGRANGE’S THEOREM
We will always write the cosets of subgroups of Z and Zn with the additive
notation we have used for cosets here. In a commutative group, left and
right cosets are always identical.
Example 2. Let H be the subgroup of S3 defined by the permutations
{(1), (123), (132)}. The left cosets of H are
The right cosets of H are exactly the same as the left cosets:
It is not always the case that a left coset is the same as a right coset.
Let K be the subgroup of S3 defined by the permutations {(1), (12)}. Then
the left cosets of K are
The following lemma is quite useful when dealing with cosets. (We leave
its proof as an exercise.)
1. g1 H = g2 H;
2. Hg1−1 = Hg2−1 ;
3. g1 H ⊆ g2 H;
5.1 COSETS 91
4. g2 ∈ g1 H;
5. g1−1 g2 ∈ H.
Proof. Let LH and RH denote the set of left and right cosets of H in
G, respectively. If we can define a bijective map φ : LH → RH , then the
theorem will be proved. If gH ∈ LH , let φ(gH) = Hg −1 . By Lemma 5.1,
the map φ is well-defined; that is, if g1 H = g2 H, then Hg1−1 = Hg2−1 . To
show that φ is one-to-one, suppose that
Proof. We first show that the map φ is one-to-one. Suppose that φ(h1 ) =
φ(h2 ) for elements h1 , h2 ∈ H. We must show that h1 = h2 , but φ(h1 ) = gh1
and φ(h2 ) = gh2 . So gh1 = gh2 , and by left cancellation h1 = h2 . To show
that φ is onto is easy. By definition every element of gH is of the form gh
for some h ∈ H and φ(h) = gh.
Corollary 5.6 Suppose that G is a finite group and g ∈ G. Then the order
of g must divide the number of elements in G.
Corollary 5.7 Let |G| = p with p a prime number. Then G is cyclic and
any g ∈ G such that g 6= e is a generator.
Theorem 5.9 Two cycles τ and µ in Sn have the same length if and only
if there exists a σ ∈ Sn such that µ = στ σ −1 .
τ = (a1 , a2 , . . . , ak )
µ = (b1 , b2 , . . . , bk ).
σ(a1 ) = b1
σ(a2 ) = b2
..
.
σ(ak ) = bk .
Then µ = στ σ −1 .
Conversely, suppose that τ = (a1 , a2 , . . . , ak ) is a k-cycle and σ ∈ Sn . If
σ(ai ) = b and σ(a(i mod k)+1 ) = b0 , then µ(b) = b0 . Hence,
Theorem 5.11 Let U (n) be the group of units in Zn . Then |U (n)| = φ(n).
Theorem 5.12 (Euler’s Theorem) Let a and n be integers such that n >
0 and gcd(a, n) = 1. Then aφ(n) ≡ 1 (mod n).
Historical Note
Joseph-Louis Lagrange (1736–1813), born in Turin, Italy, was of French and Italian
descent. His talent for mathematics became apparent at an early age. Leonhard
Euler recognized Lagrange’s abilities when Lagrange, who was only 19, communi-
cated to Euler some work that he had done in the calculus of variations. That year
he was also named a professor at the Royal Artillery School in Turin. At the age
of 23 he joined the Berlin Academy. Frederick the Great had written to Lagrange
EXERCISES 95
proclaiming that the “greatest king in Europe” should have the “greatest mathe-
matician in Europe” at his court. For 20 years Lagrange held the position vacated
by his mentor, Euler. His works include contributions to number theory, group
theory, physics and mechanics, the calculus of variations, the theory of equations,
and differential equations. Along with Laplace and Lavoisier, Lagrange was one of
the people responsible for designing the metric system. During his life Lagrange
profoundly influenced the development of mathematics, leaving much to the next
generation of mathematicians in the form of examples and new problems to be
solved.
Exercises
1. Suppose that G is a finite group with an element g of order 5 and an element
h of order 7. Why must |G| ≥ 35?
2. Suppose that G is a finite group with 60 elements. What are the orders of
possible subgroups of G?
3. Prove or disprove: Every subgroup of the integers has finite index.
4. Prove or disprove: Every subgroup of the integers has finite order.
5. List the left and right cosets of the subgroups in each of the following.
6. Describe the left cosets of SL2 (R) in GL2 (R). What is the index of SL2 (R)
in GL2 (R)?
7. Verify Euler’s Theorem for n = 15 and a = 4.
8. Use Fermat’s Little Theorem to show that if p = 4n + 3 is prime, there is no
solution to the equation x2 ≡ −1 (mod p).
9. Show that the integers have infinite index in the additive group of rational
numbers.
10. Show that the additive group of real numbers has infinite index in the additive
group of the complex numbers.
11. Let H be a subgroup of a group G and suppose that g1 , g2 ∈ G. Prove that
the following conditions are equivalent.
(a) g1 H = g2 H
96 CHAPTER 5 COSETS AND LAGRANGE’S THEOREM
Cryptography is the study of sending and receiving secret messages. The aim
of cryptography is to send messages across a channel so only the intended
recipient of the message can read it. In addition, when a message is received,
the recipient usually requires some assurance that the message is authentic;
that is, that it has not been sent by someone who is trying to deceive the
recipient. Modern cryptography is heavily dependent on abstract algebra
and number theory.
The message to be sent is called the plaintext message. The disguised
message is called the ciphertext. The plaintext and the ciphertext are both
written in an alphabet, consisting of letters or characters. Characters can
include not only the familiar alphabetic characters A, . . ., Z and a, . . ., z but
also digits, punctuation marks, and blanks. A cryptosystem, or cipher,
has two parts: encryption, the process of transforming a plaintext message
to a ciphertext message, and decryption, the reverse transformation of
changing a ciphertext message into a plaintext message.
There are many different families of cryptosystems, each distinguished
by a particular encryption algorithm. Cryptosystems in a specified cryp-
tographic family are distinguished from one another by a parameter to the
encryption function called a key. A classical cryptosystem has a single key,
which must be kept secret, known only to the sender and the receiver of
the message. If person A wishes to send secret messages to two different
people B and C, and does not wish to have B understand C’s messages or
vice versa, A must use two separate keys, so one cryptosystem is used for
exchanging messages with B, and another is used for exchanging messages
with C.
97
98 CHAPTER 6 INTRODUCTION TO CRYPTOGRAPHY
Systems that use two separate keys, one for encoding and another for
decoding, are called public key cryptosystems. Since knowledge of the
encoding key does not allow anyone to guess at the decoding key, the en-
coding key can be made public. A public key cryptosystem allows A and B
to send messages to C using the same encoding key. Anyone is capable of
encoding a message to be sent to C, but only C knows how to decode such
a message.
c = ap + b mod 26
for p. By Proposition 2.1, this is possible exactly when a has an inverse or,
equivalently, when gcd(a, 26) = 1. In this case
This is only possible if gcd(a, 26) = 1. Recognizing this fact, we will let
a = 5 since gcd(5, 26) = 1. It is easy to see that a−1 = 21. Therefore,
we can take our encryption function to be f (p) = 5p + 3 mod 26. Thus,
ALGEBRA is encoded as 3, 6, 7, 23, 8, 10, 3, or DGHXIKD. The decryption
function will be
f −1 (p) = 21p − 21 · 3 mod 26 = 21p + 15 mod 26.
A cryptosystem would be more secure if a ciphertext letter could rep-
resent more than one plaintext letter. To give an example of this type of
cryptosystem, called a polyalphabetic cryptosystem, we will generalize
affine codes by using matrices. The idea works roughly the same as before;
however, instead of encrypting one letter at a time we will encrypt pairs of
letters. We can store a pair of letters p1 and p2 in a vector
p1
p= .
p2
Let A be a 2 × 2 invertible matrix with entries in Z26 . We can define an
encoding function by
f (p) = Ap + b,
where b is a fixed column vector and matrix operations are performed in
Z26 . The decoding function must be
f −1 (p) = A−1 p − A−1 b.
Example 4. Suppose that we wish to encode the word HELP. The corre-
sponding digit string is 7, 4, 11, 15. If
3 5
A= ,
1 2
then
−1 2 21
A = .
25 3
If b = (2, 2)t , then our message is encrypted as RRCR. The encrypted letter
R represents more than one plaintext letter.
Frequency analysis can still be performed on a polyalphabetic cryptosys-
tem, because we have a good understanding of how pairs of letters appear
in the English language. The pair th appears quite often; the pair qz never
appears. To avoid decryption by a third party, we must use a larger matrix
than the one we used in Example 4.
6.2 PUBLIC KEY CRYPTOGRAPHY 101
n = pq = 667
and
φ(n) = m = (p − 1)(q − 1) = 616.
We can let E = 487, since gcd(616, 487) = 1. The encoded message is
computed to be
23487 mod 667 = 368.
This computation can be reasonably done by using the method of repeated
squares as described in Chapter 3. Using the Euclidean algorithm, we de-
termine that 191E = 1 + 151m; therefore, the decrypting key is (n, D) =
(667, 191). We can recover the original message by calculating
Now let us examine why the RSA cryptosystem works. We know that
DE ≡ 1 (mod m); hence, there exists a k such that
DE = km + 1 = kφ(n) + 1.
By Theorem 5.12,
We can now ask how one would go about breaking the RSA cryptosys-
tem. To find D given n and E, we simply need to factor n and solve for D
by using the Euclidean algorithm. If we had known that 667 = 23 · 29 in
Example 5, we could have recovered D.
Message Verification
There is a problem of message verification in public key cryptosystems.
Since the encoding key is public knowledge, anyone has the ability to send
an encoded message. If Alice receives a message from Bob, she would like
to be able to verify that it was Bob who actually sent the message. Sup-
pose that Bob’s encrypting key is (n0 , E 0 ) and his decrypting key is (n0 , D0 ).
6.2 PUBLIC KEY CRYPTOGRAPHY 103
Also, suppose that Alice’s encrypting key is (n, E) and her decrypting key
is (n, D). Since encryption keys are public information, they can exchange
coded messages at their convenience. Bob wishes to assure Alice that the
message he is sending is authentic. Before Bob sends the message x to Alice,
he decrypts x with his own key:
0
x0 = xD mod n0 .
Anyone can change x0 back to x just by encryption, but only Bob has the
ability to form x0 . Now Bob encrypts x0 with Alice’s encryption key to form
E
y 0 = x0 mod n,
a message that only Alice can decode. Alice decodes the message and then
encodes the result with Bob’s key to read the original message, a message
that could have only been sent by Bob.
Historical Note
Encrypting secret messages goes as far back as ancient Greece and Rome. As we
know, Julius Caesar used a simple shift code to send and receive messages. However,
the formal study of encoding and decoding messages probably began with the Arabs
in the 1400s. In the fifteenth and sixteenth centuries mathematicians such as Alberti
and Viete discovered that monoalphabetic cryptosystems offered no real security.
In the 1800s, F. W. Kasiski established methods for breaking ciphers in which
a ciphertext letter can represent more than one plaintext letter, if the same key
was used several times. This discovery led to the use of cryptosystems with keys
that were used only a single time. Cryptography was placed on firm mathematical
foundations by such people as W. Friedman and L. Hill in the early part of the
twentieth century.
During World War II mathematicians were very active in cryptography. Efforts
to penetrate the cryptosystems of the Axis nations were organized in England and in
the United States by such notable mathematicians as Alan Turing and A. A. Albert.
The period after World War I saw the development of special-purpose machines for
encrypting and decrypting messages. The Allies gained a tremendous advantage in
World War II by breaking the ciphers produced by the German Enigma machine
and the Japanese Purple ciphers.
By the 1970s, interest in commercial cryptography had begun to take hold.
There was a growing need to protect banking transactions, computer data, and
electronic mail. In the early 1970s, IBM developed and implemented LUZIFER,
the forerunner of the National Bureau of Standards’ Data Encryption Standard
(DES).
The concept of a public key cryptosystem, due to Diffie and Hellman, is very
recent (1976). It was further developed by Rivest, Shamir, and Adleman with the
104 CHAPTER 6 INTRODUCTION TO CRYPTOGRAPHY
RSA cryptosystem (1978). It is not known how secure any of these systems are.
The trapdoor knapsack cryptosystem, developed by Merkle and Hellman, has been
broken. It is still an open question whether or not the RSA system can be broken.
At the time of the writing of this book, the largest number factored is 135 digits
long, and at the present moment a code is considered secure if the key is about
400 digits long and is the product of two 200-digit primes. There has been a great
deal of controversy about research in cryptography in recent times: the National
Security Agency would like to keep information about cryptography secret, whereas
the academic community has fought for the right to publish basic research.
Modern cryptography has come a long way since 1929, when Henry Stimson,
Secretary of State under Herbert Hoover, dismissed the Black Chamber (the State
Department’s cryptography division) in 1929 on the ethical grounds that “gentle-
men do not read each other’s mail.”
Exercises
1. Encode IXLOVEXMATH using the cryptosystem in Example 1.
2. Decode ZLOOA WKLVA EHARQ WKHA ILQDO, which was encoded using
the cryptosystem in Example 1.
3. Assuming that monoalphabetic code was used to encode the following secret
message, what was the original message?
XE ≡ X (mod n).
n = x2 − y 2 = (x − y)(x + y).
1: while x2 − y 2 > n do
y ←y+1
if x2 − y 2 < n then
x←x+1
y←1
goto 1
else if x2 − y 2 = 0 then
a←x−y
b←x+y
write n = a ∗ b
√
The expression d n e means the smallest integer greater than or equal
to the square root of n. Write another program to do factorization using
trial division and compare the speed of the two algorithms. Which
algorithm is faster and why?
Which of the following numbers are primes and which are pseudoprimes?
(a) 342 (b) 811
(c) 601 (d) 561
(e) 771 (f ) 631
3. Let n be an odd composite number and b be a positive integer such that
gcd(b, n) = 1. If bn−1 ≡ 1 (mod n), then n is a pseudoprime base b.
Show that 341 is a pseudoprime base 2 but not a pseudoprime base 3.
EXERCISES 107
4. Write a program to determine all primes less than 2000 using trial division.
Write a second program that will determine all numbers less than 2000 that
are either primes or pseudoprimes. Compare the speed of the two programs.
How many pseudoprimes are there below 2000?
There exist composite numbers that are pseudoprimes for all bases to which
they are relatively prime. These numbers are called Carmichael num-
bers. The first Carmichael number is 561 = 3 · 11 · 17. In 1992, Al-
ford, Granville, and Pomerance proved that there are an infinite number of
Carmichael numbers [4]. However, Carmichael numbers are very rare. There
are only 2163 Carmichael numbers less than 25 × 109 . For more sophisticated
primality tests, see [1], [6], or [7].
108
7.1 ERROR-DETECTING AND CORRECTING CODES 109
m-digit message
?
Encoder
n-digit codeword
?
Transmitter
Noise
?
Receiver
(x1 , x2 , . . . , xn ) 7→ (x1 , x2 , . . . , xn , x1 , x2 , . . . , xn , x1 , x2 , . . . , xn ).
To decode the message, we choose as the ith digit the one that appears in the
ith place in at least two of the three transmissions. For example, if the origi-
nal message is (0110), then the transmitted message will be (0110 0110 0110).
If there is a transmission error in the fifth digit, then the received codeword
will be (0110 1110 0110), which will be correctly decoded as (0110).1 This
1
We will adopt the convention that bits are numbered left to right in binary n-tuples.
110 CHAPTER 7 ALGEBRAIC CODING THEORY
triple-repetition method will automatically detect and correct all single er-
rors, but it is slow and inefficient: to send a message consisting of n bits,
2n extra bits are required, and we can only detect and correct single errors.
We will see that it is possible to find an encoding scheme that will encode a
message of n bits into m bits with m much smaller than 3n.
Example 2. Even parity, a commonly used coding scheme, is much more
efficient than the simple repetition scheme. The ASCII (American Standard
Code for Information Interchange) coding system uses binary 8-tuples, yield-
ing 28 = 256 possible 8-tuples. However, only seven bits are needed since
there are only 27 = 128 ASCII characters. What can or should be done with
the extra bit? Using the full eight bits, we can detect single transmission
errors. For example, the ASCII codes for A, B, and C are
A = 6510 = 010000012 ,
B = 6610 = 010000102 ,
C = 6710 = 010000112 .
Notice that the leftmost bit is always set to 0; that is, the 128 ASCII char-
acters have codes
000000002 = 010 ,
..
.
011111112 = 12710 .
The bit can be used for error checking on the other seven bits. It is set to
either 0 or 1 so that the total number of 1 bits in the representation of a
character is even. Using even parity, the codes for A, B, and C now become
A = 010000012 ,
B = 010000102 ,
C = 110000112 .
One bit in the word is set aside as the parity check bit, and is not used to
store information. This bit is set to either 0 or 1, depending on the number
of 1’s in the word.
Adding a parity check bit allows the detection of all single errors because
changing a single bit either increases or decreases the number of 1’s by one,
and in either case the parity has been changed from even to odd, so the new
word is not a codeword. (We could also construct an error detection scheme
based on odd parity; that is, we could set the parity check bit so that a
codeword always has an odd number of 1’s.)
The even parity system is easy to implement, but has two drawbacks.
First, multiple errors are not detectable. Suppose an A is sent and the first
and seventh bits are changed from 0 to 1. The received word is a codeword,
but will be decoded into a C instead of an A. Second, we do not have the
ability to correct errors. If the 8-tuple (10011000) is received, we know that
an error has occurred, but we have no idea which bit has been changed. We
will now investigate a coding scheme that will not only allow us to detect
transmission errors but will actually correct the errors.
In Table 7.1, we present all possible words that might be received for the
transmitted codewords (000) and (111). Table 7.1 also shows the number of
bits by which each received 3-tuple differs from each original codeword.
112 CHAPTER 7 ALGEBRAIC CODING THEORY
Maximum-Likelihood Decoding2
The coding scheme presented in Example 3 is not a complete solution to the
problem because it does not account for the possibility of multiple errors.
For example, either a (000) or a (111) could be sent and a (001) received.
We have no means of deciding from the received word whether there was a
single error in the third bit or two errors, one in the first bit and one in the
second. No matter what coding scheme is used, an incorrect message could
be received: we could transmit a (000), have errors in all three bits, and
receive the codeword (111). It is important to make explicit assumptions
about the likelihood and distribution of transmission errors so that, in a
particular application, it will be known whether a given error detection
scheme is appropriate. We will assume that transmission errors are rare,
and, that when they do occur, they occur independently in each bit; that
is, if p is the probability of an error in one bit and q is the probability of
an error in a different bit, then the probability of errors occurring in both
of these bits at the same time is pq. We will also assume that a received
n-tuple is decoded into a codeword that is closest to it; that is, we assume
that the receiver uses maximum-likelihood decoding.
p
0 - 0
HH *
HHq
qHH
Hj
H
1 p
- 1
pn = (0.995)500 ≈ 0.082.
114 CHAPTER 7 ALGEBRAIC CODING THEORY
Block Codes
If we are to develop efficient error-detecting and error-correcting codes, we
will need more sophisticated mathematical tools. Group theory will allow
faster methods of encoding and decoding messages. A code is an (n, m)-
block code if the information that is to be coded can be divided into blocks
of m binary digits, each of which can be encoded into n binary digits. More
specifically, an (n, m)-block code consists of an encoding function
E : Zm n
2 → Z2
D : Zn2 → Zm
2 .
E(x7 , x6 , . . . , x1 ) = (x8 , x7 , . . . , x1 ),
d(x, y) = 4,
d(x, z) = 3,
d(y, z) = 3.
7.1 ERROR-DETECTING AND CORRECTING CODES 115
The minimum distance for this code is 3. We also have the following weights:
w(x) = 3,
w(y) = 3,
w(z) = 2.
The following proposition lists some basic properties about the weight
of a codeword and the distance between two codewords. The proof is left as
an exercise.
2. d(x, y) ≥ 0;
Then the decoder cannot decide between x and y. Even though we are
aware that an error has occurred, we do not know what the error is.
Suppose dmin ≥ 3. Then the maximum-likelihood decoding scheme cor-
rects all single errors. Starting with a codeword x, an error in the transmis-
sion of a single bit gives y with d(x, y) = 1, but d(z, y) ≥ 2 for any other
codeword z 6= x. If we do not require the correction of errors, then we can
detect multiple errors when a code has a minimum distance that is greater
than 3.
Proof. Suppose that a codeword x is sent and the word y is received with
at most n errors. Then d(x, y) ≤ n. If z is any codeword other than x, then
Historical Note
Lemma 7.4 Let x and y be binary n-tuples. Then w(x + y) = d(x, y).
Proof. Suppose that x and y are binary n-tuples. Then the distance
between x and y is exactly the number of places in which x and y differ.
But x and y differ in a particular coordinate exactly when the sum in the
coordinate is 1, since
1+1 = 0
0+0 = 0
1+0 = 1
0 + 1 = 1.
Consequently, the weight of the sum must be the distance between the two
codewords.
Theorem 7.5 Let dmin be the minimum distance for a group code C. Then
dmin is the minimum of all the nonzero weights of the nonzero codewords in
C. That is,
dmin = min{w(x) : x 6= 0}.
7.2 LINEAR CODES 119
dmin = min{d(x, y) : x 6= y}
= min{d(x, y) : x + y 6= 0}
= min{w(x + y) : x + y 6= 0}
= min{w(z) : z 6= 0}.
Linear Codes
From Example 8, it is now easy to check that the minimum nonzero weight
is 3; hence, the code does indeed detect and correct all single errors. We
have now reduced the problem of finding “good” codes to that of generating
group codes. One easy way to generate group codes is to employ a bit of
matrix theory.
Define the inner product of two binary n-tuples to be
x · y = x1 y1 + · · · + xn yn ,
x · y = xt y
y1
y2
= x1 x2 · · · xn
..
.
yn
= x1 y1 + x2 y2 + · · · + xn yn .
1
This example leads us to hope that there is a connection between matrices
and coding theory.
Let Mm×n (Z2 ) denote the set of all m×n matrices with entries in Z2 . We
do matrix operations as usual except that all our addition and multiplication
operations occur in Z2 . Define the null space of a matrix H ∈ Mm×n (Z2 )
to be the set of all binary n-tuples x such that Hx = 0. We denote the null
space of a matrix H by Null(H).
x2 + x4 = 0
x1 + x2 + x3 + x4 = 0
x3 + x4 + x5 = 0.
Proof. Since each element of Zn2 is its own inverse, the only thing that
really needs to be checked here is closure. Let x, y ∈ Null(H) for some
matrix H in Mm×n (Z2 ). Then Hx = 0 and Hy = 0. So
H(x + y) = H(x + y) = Hx + Hy = 0 + 0 = 0.
7.3 PARITY-CHECK AND GENERATOR MATRICES 121
the received word is not a codeword. We must either attempt to correct the
word or request that it be transmitted again.
Example 12. Suppose that we have the following eight words to be en-
coded:
(000), (001), (010), . . . , (111).
For
0 1 1
A = 1 1 0 ,
1 0 1
the associated standard generator and canonical parity-check matrices are
1 0 0
0 1 0
0 0 1
G=
0 1 1
1 1 0
1 0 1
and
0 1 1 1 0 0
H = 1 1 0 0 1 0 ,
1 0 1 0 0 1
respectively.
Observe that the rows in H represent the parity checks on certain bit
positions in a 6-tuple. The 1’s in the identity matrix serve as parity checks
7.3 PARITY-CHECK AND GENERATOR MATRICES 123
x2 + x3 + x4 = 0
x1 + x2 + x5 = 0
x1 + x3 + x6 = 0.
Here x4 serves as a check bit for x2 and x3 ; x5 is a check bit for x1 and x2 ;
and x6 is a check bit for x1 and x3 . The identity matrix keeps x4 , x5 , and x6
from having to check on each other. Hence, x1 , x2 , and x3 can be arbitrary
but x4 , x5 , and x6 must be chosen to ensure parity. The null space of H is
easily computed to be
(000000) (001101) (010110) (011011)
(100011) (101110) (110101) (111000).
An even easier way to compute the null space is with the generator matrix
G (Table 7.4).
Gx − Gy = G(x − y) = 0.
where
1 i=j
δij =
6 j
0 i=
is the Kronecker delta.
e1 = (100 · · · 00)t
e2 = (010 · · · 00)t
..
.
en = (000 · · · 01)t
126 CHAPTER 7 ALGEBRAIC CODING THEORY
We state this result in the following proposition and leave the proof as
an exercise.
Proposition 7.11 Let ei be the binary n-tuple with a 1 in the ith coordinate
and 0’s elsewhere and suppose that H ∈ Mm×n (Z2 ). Then Hei is the ith
column of the matrix H.
Theorem 7.12 Let H be an m × n binary matrix. Then the null space of
H is a single error-detecting code if and only if no column of H consists
entirely of zeros.
Proof. Suppose that Null(H) is a single error-detecting code. Then the
minimum distance of the code must be at least 2. Since the null space is a
group code, it is sufficient to require that the code contain no codewords of
less than weight 2 other than the zero codeword. That is, ei must not be a
codeword for i = 1, . . . , n. Since Hei is the ith column of H, the only way
in which ei could be in the null space of H would be if the ith column were
all zeros, which is impossible; hence, the code must have the capability to
detect at least single errors.
Conversely, suppose that no column of H is the zero column. By Propo-
sition 7.11, Hei 6= 0.
then the null space of H1 is a single error-detecting code and the null space
of H2 is not.
We can even do better than Theorem 7.12. This theorem gives us con-
ditions on a matrix H that tell us when the minimum weight of the code
formed by the null space of H is 2. We can also determine when the mini-
mum distance of a linear code is 3 by examining the corresponding matrix.
Proof. The n-tuple ei + ej has 1’s in the ith and jth entries and 0’s
elsewhere, and w(ei + ej ) = 2 for i 6= j. Since
0 = H(ei + ej ) = Hei + Hej
can only occur if the ith and jth columns are identical, the null space of H
is a single error-correcting code.
Theorem 7.15 Let H ∈ Mm×n (Z2 ) and suppose that the linear code cor-
responding to H is single error-correcting. Let r be a received n-tuple that
was transmitted with at most one error. If the syndrome of r is 0, then no
error has occurred; otherwise, if the syndrome of r is equal to some column
of H, say the ith column, then the error has occurred in the ith bit.
Hence, x has an error in the third bit and z has an error in the fourth bit. The
transmitted codewords for x and z must have been (110110) and (010011),
respectively. The syndrome of y does not occur in any of the columns of the
matrix H, so multiple errors must have occurred to produce y.
Coset Decoding
We can use group theory to obtain another way of decoding messages. A
linear code C is a subgroup of Zn2 . Coset or standard decoding uses the
cosets of C in Zn2 to implement maximum-likelihood decoding. Suppose that
C is an (n, m)-linear code. A coset of C in Zn2 is written in the form x + C,
130 CHAPTER 7 ALGEBRAIC CODING THEORY
Example 18. Let C be the (5, 3)-linear code given by the parity-check
matrix
0 1 1 0 0
H = 1 0 0 1 0 .
1 1 0 0 1
The code consists of the codewords
(00000) (01101) (10011) (11110).
There are 25−2 = 23 cosets of C in Z52 , each with order 22 = 4. These cosets
are listed in Table 7.5.
Our task is to find out how knowing the cosets might help us to decode
a message. Suppose that x was the original codeword sent and that r is
the n-tuple received. If e is the transmission error, then r = e + x or,
equivalently, x = e + r. However, this is exactly the statement that r is an
element in the coset e + C. In maximum-likelihood decoding we expect the
error e to be as small as possible; that is, e will have the least weight. An
n-tuple of least weight in a coset is called a coset leader. Once we have
determined a coset leader for each coset, the decoding process becomes a
task of calculating r + e to obtain x.
Proposition 7.16 Let C be an (n, k)-linear code given by the matrix H and
suppose that x and y are in Zn2 . Then x and y are in the same coset of C
if and only if Hx = Hy. That is, two n-tuples are in the same coset if and
only if their syndromes are the same.
Proof. Two n-tuples x and y are in the same coset of C exactly when
x − y ∈ C; however, this is equivalent to H(x − y) = 0 or Hx = Hy.
Example 20. Table 7.6 is a decoding table for the code C given in Exam-
ple 18. If x = (01111) is received, then its syndrome can be computed to
be
0
Hx = 1 .
1
Examining the decoding table, we determine that the coset leader is (00010).
It is now easy to decode the received codeword.
Given an (n, k)-block code, the question arises of whether or not coset
decoding is a manageable scheme. A decoding table requires a list of cosets
and syndromes, one for each of the 2n−k cosets of C. Suppose that we have
a (32, 24)-block code. We have a huge number of codewords, 224 , yet there
are only 232−24 = 28 = 256 cosets.
Exercises
1. Why is the following encoding scheme not acceptable?
Information: 0 1 2 3 4 5 6 7 8
Codeword: 000 001 010 011 101 110 111 000 001
2. Without doing any addition, explain why the following set of 4-tuples in Z42
cannot be a group code.
1 0 0 1 1
(c) (d) 0 0 0 1 1 1 1
0 1 0 1 1 0 1 1 0 0 1 1
1 0 1 0 1 0 1
0 1 1 0 0 1 1
if possible.
10. Suppose that a 1000-bit binary message is transmitted. Assume that the
probability of a single error is p and that the errors occurring in different
bits are independent of one another. If p = 0.01, what is the probability of
more than one error occurring? What is the probability of exactly two errors
occurring? Repeat this problem for p = 0.0001.
11. Which matrices are canonical parity-check matrices? For those matrices that
are canonical parity-check matrices, what are the corresponding standard
generator matrices? What are the error-detection and error-correction capa-
bilities of the code generated by each of these matrices?
1 1 0 0 0
(a) (b) 0 1 1 0 0 0
0 0 1 0 0 1
1 0 1 0 0
0 0 0 1 0
0 1 0 0 1
0
1 0 0 0 1 1 1 0 0 0 1
(c) 1 1 1 0
1 0 0 1
0 0 0 1 0 0 0
(d)
0 1 1 0 1 0 0
1 0 1 0 0 1 0
0 1 1 0 0 0 1
12. List all possible syndromes for the codes generated by each of the matrices
in the previous exercise.
134 CHAPTER 7 ALGEBRAIC CODING THEORY
13. Let
0 1 1 1 1
H= 0 0 0 1 1 .
1 0 1 0 1
Compute the syndrome caused by each of the following transmission errors.
14. Let C be the group code in Z32 defined by the codewords (000) and (111).
Compute the cosets of H in Z32 . Why was there no need to specify right or
left cosets? Give the single transmission error, if any, to which each coset
corresponds.
15. For each of the following matrices, find the cosets of the corresponding code
C. Give a decoding table for each code if possible.
(a) 0 1 0 0 0 (b) 0 0 1 0 0
1 0 1 0 1 1 1 0 1 0
1 0 0 1 0 0 1 0 1 0
1 1 0 0 1
1 0 0 1 1
(c) (d) 1 0 0 1 1 1 1
0 1 0 1 1 1 1 1 0 0 1 1
1 0 1 0 1 0 1
1 1 1 0 0 1 0
16. Let x, y, and z be binary n-tuples. Prove each of the following statements.
Programming Exercises
Write a program to implement a (16, 12)-linear code. Your program should be
able to encode and decode messages using coset decoding. Once your program is
written, write a program to simulate a binary symmetric channel with transmission
noise. Compare the results of your simulation with the theoretically predicted error
probability.
Many groups may appear to be different at first glance, but can be shown
to be the same by a simple renaming of the group elements. For example,
Z4 and the subgroup of the circle group T generated by i can be shown
to be the same by demonstrating a one-to-one correspondence between the
elements of the two groups and between the group operations. In such a
case we say that the groups are isomorphic.
φ(0) = 1
φ(1) = i
φ(2) = −1
φ(3) = −i.
Since
φ(m + n) = im+n = im in = φ(m)φ(n),
138
8.1 DEFINITION AND EXAMPLES 139
Of course, we must still show that φ is one-to-one and onto, but this can be
determined using calculus.
Example 3. The integers are isomorphic to the subgroup of Q∗ consisting
of elements of the form 2n . Define a map φ : Z → Q∗ by φ(n) = 2n . Then
U (8) = {1, 3, 5, 7}
U (12) = {1, 5, 7, 11}.
1 7→ 1
3 7→ 5
5 7→ 7
7 7→ 11.
The map φ is not the only possible isomorphism between these two groups.
We could define another isomorphism ψ by ψ(1) = 1, ψ(3) = 11, ψ(5) = 5,
ψ(7) = 7. In fact, both of these groups are isomorphic to Z2 × Z2 (see
Example 14 in Chapter 2).
Example 5. Even though S3 and Z6 possess the same number of elements,
we would suspect that they are not isomorphic, because Z6 is abelian and
S3 is nonabelian. To demonstrate that this is indeed the case, suppose that
φ : Z6 → S3 is an isomorphism. Let a, b ∈ S3 be two elements such that
140 CHAPTER 8 ISOMORPHISMS
φ(m) = a
φ(n) = b.
However,
1. φ−1 : H → G is an isomorphism.
2. |G| = |H|.
Proof. Assertions (1) and (2) follow from the fact that φ is a bijection.
We will prove (3) here and leave the remainder of the theorem to be proved
in the exercises.
(3) Suppose that h1 and h2 are elements of H. Since φ is onto, there
exist elements g1 , g2 ∈ G such that φ(g1 ) = h1 and φ(g2 ) = h2 . Therefore,
We are now in a position to characterize all cyclic groups.
Proof. Let G be a cyclic group with infinite order and suppose that a is a
generator of G. Define a map φ : Z → G by φ : n 7→ an . Then
Hence, we can modify our goal of classifying all groups to classifying all
groups up to isomorphism; that is, we will consider two groups to be the
same if they are isomorphic.
Cayley’s Theorem
Cayley proved that if G is a group, it is isomorphic to a group of permu-
tations on some set; hence, every group is a permutation group. Cayley’s
Theorem is what we call a representation theorem. The aim of represen-
tation theory is to find an isomorphism of some group G that we wish to
study into a group that we know a great deal about, such as a group of
permutations or matrices.
Example 6. Consider the group Z3 . The Cayley table for Z3 is as follows.
+ 0 1 2
0 0 1 2
1 1 2 0
2 2 0 1
142 CHAPTER 8 ISOMORPHISMS
G = {λg : g ∈ G}.
Also,
λe (a) = ea = a
and
(λg−1 ◦ λg )(a) = λg−1 (ga) = g −1 ga = a = λe (a).
We can define an isomorphism from G to G by φ : g 7→ λg . The group
operation is preserved since
ga = λg a = λh a = ha.
Hence, g = h. That φ is onto follows from the fact that φ(g) = λg for any
λg ∈ G.
The isomorphism g 7→ λg is known as the left regular representation
of G.
Historical Note
Arthur Cayley was born in England in 1821, though he spent much of the first
part of his life in Russia, where his father was a merchant. Cayley was educated
at Cambridge, where he took the first Smith’s Prize in mathematics. A lawyer
for much of his adult life, he wrote several papers in his early twenties before
entering the legal profession at the age of 25. While practicing law he continued his
mathematical research, writing more than 300 papers during this period of his life.
These included some of his best work. In 1863 he left law to become a professor
at Cambridge. Cayley wrote more than 900 papers in fields such as group theory,
geometry, and linear algebra. His legal knowledge was very valuable to Cambridge;
he participated in the writing of many of the university’s statutes. Cayley was also
one of the people responsible for the admission of women to Cambridge.
Proof. Suppose that m is the least common multiple of r and s and let
n = |(g, h)|. Then
Example 10. Let (8, 56) ∈ Z12 × Z60 . Since gcd(8, 12) = 4, the order of 8
is 12/4 = 3 in Z12 . Similarly, the order of 56 in Z60 is 15. The least common
multiple of 3 and 15 is 15; hence, (8, 56) has order 15 in Z12 × Z60 .
Example 11. The group Z2 × Z3 consists of the pairs
(0, 0), (0, 1), (0, 2), (1, 0), (1, 1), (1, 2).
Corollary 8.12 If
m = pe11 · · · pekk ,
where the pi s are distinct primes, then
Zm ∼
= Zpe11 × · · · × Zpek .
k
e
Proof. Since the greatest common divisor of pei i and pj j is 1 for i 6= j, the
proof follows from Corollary 8.11.
In Chapter 11, we will prove that all finite abelian groups are isomorphic
to direct products of the form
Zpe1 × · · · × Zpek
1 k
• G = HK = {hk : h ∈ H, k ∈ K};
• H ∩ K = {e};
• Hi ∩ h∪j6=i Hj i = {e};
Exercises
1. Prove that Z ∼
= nZ for n 6= 0.
2. Prove that C∗ is isomorphic to the subgroup of GL2 (R) consisting of matrices
of the form
a b
.
−b a
a ∗ b = a + b + ab.
Prove that G is a group under this operation. Show that (G, ∗) is isomorphic
to the multiplicative group of nonzero real numbers.
EXERCISES 149
and
0 1
B=
1 0
form a multiplicative group isomorphic to Dn .
14. Show that the set of all matrices of the form
±1 n
B= ,
0 1
If H is a subgroup of a group G, then right cosets are not always the same as
left cosets; that is, it is not always the case that gH = Hg for all g ∈ G. The
subgroups for which this property holds play a critical role in group theory:
they allow for the construction of a new class of groups, called factor or
quotient groups. Factor groups may be studied by using homomorphisms,
a generalization of isomorphisms.
152
9.1 FACTOR GROUPS AND NORMAL SUBGROUPS 153
are
N = {(1), (123), (132)}
(12)N = N (12) = {(12), (13), (23)}.
The following theorem is fundamental to our understanding of normal
subgroups.
2. For all g ∈ G, gN g −1 ⊂ N .
3. For all g ∈ G, gN g −1 = N .
Factor Groups
If N is a normal subgroup of a group G, then the cosets of N in G form
a group G/N under the operation (aN )(bN ) = abN . This group is called
the factor or quotient group of G and N . Our first task is to prove that
G/N is indeed a group.
Proof. The group operation on G/N is (aN )(bN ) = abN . This operation
must be shown to be well-defined; that is, group multiplication must be
independent of the choice of coset representative. Let aN = bN and cN =
dN . We must show that
Geometrically, we are simply wrapping the real line around the circle in a
group-theoretic fashion.
The following proposition lists some basic properties of group homomor-
phisms.
Proof. (1) Suppose that e and e0 are the identities of G1 and G2 , respec-
tively; then
e0 φ(e) = φ(e) = φ(ee) = φ(e)φ(e).
By cancellation, φ(e) = e0 .
(2) This statement follows from the fact that
Simplicity of An
Of special interest are groups with no nontrivial normal subgroups. Such
groups are called simple groups. Of course, we already have a whole
class of examples of simple groups, Zp , where p is prime. These groups are
trivially simple since they have no proper subgroups other than the subgroup
consisting solely of the identity. Other examples of simple groups are not
so easily found. We can, however, show that the alternating group, An , is
simple for n ≥ 5. The proof of this result requires several lemmas.
Proof. To show that the 3-cycles generate An , we need only show that
any pair of transpositions can be written as the product of 3-cycles. Since
(ab) = (ba), every pair of transpositions must be one of the following:
(ab)(ab) = id
(ab)(cd) = (acb)(acd)
(ab)(ac) = (acb).
(iaj) = (ija)2
(iab) = (ijb)(ija)2
(jab) = (ijb)2 (ija)
(abc) = (ija)2 (ijc)(ijb)2 (ija).
• σ is a 3-cycle.
is also in N . Since
σ = τ (a1 a2 a3 )(a4 a5 a6 ).
Then
σ −1 (a1 a2 a4 )σ(a1 a2 a4 )−1 ∈ N
since
(a1 a2 a4 )σ(a1 a2 a4 )−1 ∈ N.
So
σ 2 = τ (a1 a2 a3 )τ (a1 a2 a3 )
= (a1 a3 a2 ).
So N contains a 3-cycle.
The only remaining possible case is a disjoint product of the form
σ = τ (a1 a2 )(a3 a4 ),
Historical Note
One of the foremost problems of group theory has been to classify all simple finite
groups. This problem is over a century old and has been solved only in the last
few years. In a sense, finite simple groups are the building blocks of all finite
groups. The first nonabelian simple groups to be discovered were the alternating
groups. Galois was the first to prove that A5 was simple. Later mathematicians,
such as C. Jordan and L. E. Dickson, found several infinite families of matrix
groups that were simple. Other families of simple groups were discovered in the
1950s. At the turn of the century, William Burnside conjectured that all nonabelian
simple groups must have even order. In 1963, W. Feit and J. Thompson proved
Burnside’s conjecture and published their results in the paper “Solvability of Groups
of Odd Order,” which appeared in the Pacific Journal of Mathematics. Their
proof, running over 250 pages, gave impetus to a program in the 1960s and 1970s
to classify all finite simple groups. Daniel Gorenstein was the organizer of this
remarkable effort. One of the last simple groups was the “Monster,” discovered
by R. Greiss. The Monster, a 196,833 × 196,833 matrix group, is one of the 26
sporadic, or special, simple groups. These sporadic simple groups are groups that
fit into no infinite family of simple groups.
162 CHAPTER 9 HOMOMORPHISMS AND FACTOR GROUPS
Since η(g1 K) = η(g2 K), η does not depend on the choice of coset represen-
tative. Clearly η is onto ψ(G). To show that η is one-to-one, suppose that
η(g1 K) = η(g2 K). Then ψ(g1 ) = ψ(g2 ). This implies that ψ(g1−1 g2 ) = e,
or g1−1 g2 is in the kernel of ψ; hence, g1−1 g2 K = K; that is, g1 K = g2 K.
Finally, we must show that η is a homomorphism, but
ψ
G -H
J
φJ
η
J^
J
G/K
H/H ∩ N ∼
= HN/N.
Notice that in the course of the proof of Theorem 9.11, we have also
proved the following theorem.
Z/mZ ∼
= (Z/mnZ)/(mZ/mnZ).
Exercises
1. For each of the following groups G, determine whether H is a normal sub-
group of G. If H is a normal subgroup, write out a Cayley table for the
factor group G/H.
(a) G = S4 and H = A4
(b) G = A5 and H = {(1), (123), (132)}
(c) G = S4 and H = D4
(d) G = Q8 and H = {1, −1, i, −i}
(e) G = Z and H = 5Z
2. Find all the subgroups of D4 . Which subgroups are normal? What are all
the factor groups of D4 up to isomorphism?
3. Find all the subgroups of the quaternion group, Q8 . Which subgroups are
normal? What are all the factor groups of Q4 up to isomorphism?
4. Prove that det(AB) = det(A) det(B) for A, B ∈ GL2 (R). This shows that
the determinant is a homomorphism from GL2 (R) to R∗ .
5. Which of the following maps are homomorphisms? If the map is a homomor-
phism, what is the kernel?
(a) φ : R∗ → GL2 (R) defined by
1 0
φ(a) =
0 a
166 CHAPTER 9 HOMOMORPHISMS AND FACTOR GROUPS
where x ∈ R.
(a) Show that U is a subgroup of T .
(b) Prove that U is abelian.
(c) Prove that U is normal in T .
(d) Show that T /U is abelian.
(e) Is T normal in GL2 (R)?
7. Let A be an m × n matrix. Show that matrix multiplication, x 7→ Ax, defines
a homomorphism φ : Rn → Rm .
8. Let φ : Z → Z be given by φ(n) = 7n. Prove that φ is a group homomor-
phism. Find the kernel and the image of φ.
9. Describe all of the homomorphisms from Z24 to Z18 .
EXERCISES 167
(a) List the elements in HN (we usually write H + N for these additive
groups) and H ∩ N .
(b) List the cosets in HN/N , showing the elements in each coset.
(c) List the cosets in H/(H ∩ N ), showing the elements in each coset.
(d) Give the correspondence between HN/N and H/(H ∩ N ) described in
the proof of the Second Isomorphism Theorem.
C(g) = {x ∈ G : xg = gx}.
ig : G → G,
3. The set of all inner automorphisms is denoted by Inn(G). Show that Inn(G)
is a subgroup of Aut(G).
4. Find an automorphism of a group G that is not an inner automorphism.
5. Let G be a group and ig be an inner automorphism of G, and define a map
G → Aut(G)
by
g 7→ ig .
Prove that this map is a homomorphism with image Inn(G) and kernel Z(G).
Use this result to conclude that
G/Z(G) ∼
= Inn(G).
170
10.1 MATRIX GROUPS 171
A(x + y) = Ax + Ay
αAx = A(αx),
where
x1
x2
x= .
..
.
xn
We will often abbreviate the matrix A by writing (aij ).
Conversely, if T : Rn → Rm is a linear map, we can associate a matrix
A with T by considering what T does to the vectors
e1 = (1, 0, . . . , 0)t
e2 = (0, 1, . . . , 0)t
..
.
en = (0, 0, . . . , 1)t .
x1 e1 + x2 e2 + · · · + xn en .
Consequently, if
then
T (x) = T (x1 e1 + x2 e2 + · · · + xn en )
= x1 T (e1 ) + x2 T (e2 ) + · · · + xn T (en )
n n
!t
X X
= a1k xk , . . . , amk xk
k=1 k=1
= Ax.
Geometrically, SL2 (R) is the group that preserves the areas of parallelo-
grams. Let
1 1
A=
0 1
be in SL2 (R). In Figure 10.1, the unit square corresponding to the vectors
x = (1, 0)t and y = (0, 1)t is taken by A to the parallelogram with sides
(1, 0)t and (1, 1)t ; that is, Ax = (1, 0)t and Ay = (1, 1)t . Notice that these
two parallelograms have the same area.
(1, 1)
(0, 1)
(1, 0) (1, 0)
There is a more geometric way of viewing the group O(n). The orthog-
onal matrices are exactly those matrices that preserve the length of vectors.
10.1 MATRIX GROUPS 175
We can define the length of a vector using the Euclidean inner product,
or dot product, of two vectors. The Euclidean inner product of two vectors
x = (x1 , . . . , xn )t and y = (y1 , . . . , yn )t is
y1
y2
hx, yi = xt y = (x1 , x2 , . . . , xn ) . = x1 y1 + · · · + xn yn .
. .
yn
We define the length of a vector x = (x1 , . . . , xn )t to be
p q
kxk = hx, xi = x21 + · · · + x2n .
Associated with the notion of the length of a vector is the idea of the distance
between two vectors. We define the distance between two vectors x and y
to be kx − yk. We leave as an exercise the proof of the following proposition
about the properties of Euclidean inner products.
Proposition 10.1 Let x, y, and w be vectors in Rn and α ∈ R. Then
1. hx, yi = hy, xi.
2. hx, y + wi = hx, yi + hx, wi.
3. hαx, yi = hx, αyi = αhx, yi.
4. hx, xi ≥ 0 with equality exactly when x = 0.
5. If hx, yi = 0 for all x in Rn , then y = 0.
√
Example 5. The vector x = (3, 4)t has length 32 + 42 = 5. We can also
see that the orthogonal matrix
3/5 −4/5
A=
4/5 3/5
preserves the length of this vector. The vector Ax = (−7/5, 24/5)t also has
length 5.
Since det(AAt ) = det(I) = 1 and det(A) = det(At ), the determinant of
any orthogonal matrix is either 1 or −1. Consider the column vectors
a1j
a2j
aj = .
.
.
anj
176 CHAPTER 10 MATRIX GROUPS AND SYMMETRY
2. A−1 = At .
(sin θ, – cos θ)
(cos θ, sin θ)
(a, b)
θ
(a, –b)
The identity of the group is (I, 0); the inverse of (A, x) is (A−1 , −A−1 x). In
Exercise 6, you are asked to check that E(n) is indeed a group under this
operation.
x+y
10.2 Symmetry
An isometry or rigid motion in Rn is a distance-preserving function f
from Rn to Rn . This means that f must satisfy
kf (x) − f (y)k = kx − yk
T (x)
kxk2 − 2hf (x), f (y)i + kyk2 = kf (x)k2 − 2hf (x), f (y)i + kf (y)k2
= hf (x) − f (y), f (x) − f (y)i
= kf (x) − f (y)k2
= kx − yk2
= hx − y, x − yi
= kxk2 − 2hx, yi + kyk2 .
180 CHAPTER 10 MATRIX GROUPS AND SYMMETRY
Consequently,
hf (x), f (y)i = hx, yi.
Now let e1 and e2 be (1, 0)t and (0, 1)t , respectively. If
x = (x1 , x2 ) = x1 e1 + x2 e2 ,
then
f (x) = hf (x), f (e1 )if (e1 ) + hf (x), f (e2 )if (e2 ) = x1 f (e1 ) + x2 f (e2 ).
f (y) = Ay + x1
g(y) = By + x2 ,
their composition is
Notice that det(Rθ ) = 1, det(Tθ ) = −1, and Tθ2 = I. We can divide the
proof up into two cases. In the first case, all of the elements in G have
determinant one. In the second case, there exists at least one element in G
with determinant −1.
Case 1. The determinant of every element in G is one. In this case every
element in G must be a rotation. Since G is finite, there is a smallest angle,
say θ0 , such that the corresponding element Rθ0 is the smallest rotation in
the positive direction. We claim that Rθ0 generates G. If not, then for some
positive integer n there is an angle θ1 between nθ0 and (n + 1)θ0 . If so, then
(n + 1)θ0 − θ1 corresponds to a rotation smaller than θ0 , which contradicts
the minimality of θ0 .
Case 2. The group G contains a reflection Tθ . The kernel of the ho-
momorphism φ : G → {−1, 1} given by A 7→ det(A) consists of elements
whose determinant is 1. Therefore, |G/ ker φ| = 2. We know that the kernel
is cyclic by the first case and is a subgroup of G of, say, order n. Hence,
|G| = 2n. The elements of G are
Rθ , . . . , Rθn−1 , T Rθ , . . . , T Rθn−1 .
T Rθ T = Rθ−1 .
(–1, 1) (1, 1)
(2, 0)
(–1, –1)
Notice that a lattice can have several bases. For example, the vectors
(1, 1)t and (2, 0)t have the same lattice as the vectors (−1, 1)t and (−1, −1)t
(Figure 10.7). However, any lattice is completely determined by a basis.
10.2 SYMMETRY 183
Given two bases for the same lattice, say {x1 , x2 } and {y1 , y2 }, we can
write
y1 = α1 x1 + α2 x2
y2 = β1 x1 + β2 x2 ,
Parallelogram Hexagonal
To answer the question of how the point groups and the translation
groups can be combined, we must look at the different types of lattices.
Lattices can be classified by the structure of a single lattice cell. The possible
cell shapes are parallelogram, rectangular, square, rhombic, and hexagonal
10.2 SYMMETRY 185
(Figure 10.8). The wallpaper groups can now be classified according to the
types of reflections that occur in each group: these are ordinarily reflections,
glide reflections, both, or none.
p4m p4g
The 17 wallpaper groups are listed in Table 10.1. The groups p3m1 and
p31m can be distinguished by whether or not all of their threefold centers
lie on the reflection axes: those of p3m1 must, whereas those of p31m may
not. Similarly, the fourfold centers of p4m must lie on the reflection axes
whereas those of p4g need not (Figure 10.9). The complete proof of this
186 CHAPTER 10 MATRIX GROUPS AND SYMMETRY
theorem can be found in several of the references at the end of this chapter,
including [5], [6], [10], and [11].
Historical Note
Exercises
1. Prove the identity
1
kx + yk2 − kxk2 − kyk2 .
hx, yi =
2
√ √
(c) 4/ √5 0 3/√5 (d) 1/3 2/3 −2/3
−3/ 5 0 4/ 5 −2/3 2/3 1/3
0 −1 0 −2/3 1/3 2/3
(a) (c)
(b)
Figure 10.10.
Figure 10.11.
17. Determine which of the 17 wallpaper groups preserves the symmetry of the
pattern in Figure 10.11.
18. Find the rotation group of a dodecahedron.
19. For each of the 17 wallpaper groups, draw a wallpaper pattern having that
group as a symmetry group.
G = Hn ⊃ Hn−1 ⊃ · · · ⊃ H1 ⊃ H0 = {e},
where each subgroup Hi is normal in Hi+1 and each of the factor groups
Hi+1 /Hi is abelian, then G is a solvable group. In addition to allowing us
to distinguish between certain classes of groups, solvable groups turn out to
be central to the study of solutions to polynomial equations.
190
11.1 FINITE ABELIAN GROUPS 191
Example 1. Obviously, all finite groups are finitely generated. For example,
the group S3 is generated by the permutations (12) and (123). The group
Z × Zn is an infinite group but is finitely generated by {(1, 0), (0, 1)}.
Example 2. Not all groups are finitely generated. Consider the rational
numbers Q under the operation of addition. Suppose that Q is finitely
generated with generators p1 /q1 , . . . , pn /qn , where each pi /qi is a fraction
expressed in its lowest terms. Let p be some prime that does not divide
any of the denominators q1 , . . . , qn . We claim that 1/p cannot be in the
subgroup of Q that is generated by p1 /q1 , . . . , pn /qn , since p does not divide
the denominator of any element in this subgroup. This fact is easy to see
since the sum of any two generators is
h = giα11 · · · giαnn ,
The reason that powers of a fixed gi may occur several times in the
product is that we may have a nonabelian group. However, if the group is
abelian, then the gi ’s need occur only once. For example, a product such as
a−3 b5 a7 could always be simplified (in this case, to a4 b5 ).
Proof. Let K be the set of all products of the form giα11 · · · giαnn , where the
gik ’s are not necessarily distinct. Certainly K is a subset of H. We need
only show that K is a subgroup of G. If this is the case, then K = H, since
H is the smallest subgroup containing all the gi ’s.
192 CHAPTER 11 THE STRUCTURE OF GROUPS
Clearly, the set K is closed under the group operation. Since gi0 = 1,
the identity is in K. It remains to show that the inverse of an element
g = g1k1 · · · giknn in K must also be in K. However,
Now let us restrict our attention to finite abelian groups. We can express
any finite abelian group as a finite direct product of cyclic groups. More
specifically, letting p be prime, we define a group G to be a p-group if every
element in G has as its order a power of p. For example, both Z2 × Z2 and
Z4 are 2-groups, whereas Z27 is a 3-group. We shall prove that every finite
abelian group is isomorphic to a direct product of cyclic p-groups. Before we
state the main theorem concerning finite abelian groups, we shall consider
a special case.
Proof. If |G| = 1, then the theorem is trivial. Suppose that the order of
G is greater than 1, say
|G| = pα1 1 · · · pαnn ,
where p1 , . . . , pn are all prime, and define Gi to be the set of elements in G of
order pki for some integer k. Since G is an abelian group, we are guaranteed
that Gi is a subgroup of G for i = 1, . . . , n. We must show that
G = G 1 × · · · × Gn .
That is, we must be able to write every g ∈ G as a unique product gp1 · · · gpn
where gpi is of the order of some power of pi . Since the order of g divides
the order of G, we know that
g = g1 · · · gn = h1 · · · hn
with hi ∈ Gi . Then
• Z2 × Z2 × Z3 × Z9 × Z5 ;
• Z2 × Z2 × Z27 × Z5 ;
• Z4 × Z3 × Z3 × Z3 × Z5 ;
• Z4 × Z3 × Z9 × Z5 ;
• Z4 × Z27 × Z5 .
Lemma 11.4 Let G be a finite abelian p-group and suppose that g ∈ G has
maximal order. Then G can be written as hgi×H for some subgroup H of G.
194 CHAPTER 11 THE STRUCTURE OF GROUPS
and the order of g r must be less than or equal to pm−1 . Therefore, g r cannot
generate hgi. Notice that p must occur as a factor of r, say r = ps, and
hp = g r = g ps . Define a to be g −s h. Then a cannot be in hgi; otherwise, h
would also have to be in hgi. Also,
ap = g −sp hp = g −r hp = h−p hp = e.
We have now formed an element a with order p such that a ∈ / hgi. Since h
was chosen to have the smallest order of all of the elements that are not in
hgi, |H| = p.
Now we will show that the order of gH in the factor group G/H must
be the same as the order of g in G. If |gH| < |g| = pm , then
m−1 m−1
H = (gH)p = gp H;
m−1
hence, g p must be in hgi ∩ H = {e}, which contradicts the fact that the
order of g is pm . Therefore, gH must have maximal order in G/H. By the
Correspondence Theorem and our induction hypothesis,
G/H ∼
= hgHi × K/H
otherwise, G ∼
= Z|g| × H for some subgroup H contained in G by the lemma.
Since |H| < |G|, we can apply mathematical induction.
We now state the more general theorem for all finitely generated abelian
groups. The proof of this theorem can be found in any of the references at
the end of this chapter.
G = Hn ⊃ Hn−1 ⊃ · · · ⊃ H1 ⊃ H0 = {e},
Z60 /h3i ∼
= h20i/{0} ∼
= Z3
h3i/h15i ∼
= h4i/h20i ∼
= Z5
h15i/{0} ∼
= Z60 /h4i ∼
= Z4 .
Z60 /h3i ∼
= Z3
∼
h3i/h15i = Z5
h15i/h30i ∼
= Z2
∼ Z2 .
h30i/{0} =
11.2 SOLVABLE GROUPS 197
Example 10. Not every group has a composition series or a principal series.
Suppose that
{0} = H0 ⊂ H1 ⊂ · · · ⊂ Hn−1 ⊂ Hn = Z
is a subnormal series for the integers under addition. Then H1 must be of
the form nZ for some n ∈ N. In this case H1 /H0 ∼ = nZ is an infinite cyclic
group with many nontrivial proper normal subgroups.
we have a composition series for Hn−1 . Our induction hypothesis says that
this series must be equivalent to the composition series
Hn−1 ⊃ · · · ⊃ H1 ⊃ H0 = {e}.
G = Hn ⊃ Hn−1 ⊃ · · · ⊃ H1 ⊃ H0 = {e}
and
are equivalent. If Hn−1 = Km−1 , then the composition series {Hi } and {Kj }
are equivalent and we are done; otherwise, Hn−1 Km−1 is a normal subgroup
of G properly containing Hn−1 . In this case Hn−1 Km−1 = G and we can
apply the Second Isomorphism Theorem once again; that is,
Therefore,
and
Sn ⊃ An ⊃ {(1)}
Exercises
1. Find all of the abelian groups of order less than or equal to 40 up to isomor-
phism.
2. Find all of the abelian groups of order 200 up to isomorphism.
3. Find all of the abelian groups of order 720 up to isomorphism.
4. Find all of the composition series for each of the following groups.
(a) Z12 (b) Z48
(c) The quaternions, Q8 (d) D4
(e) S3 × Z4 (f ) S4
(g) Sn , n ≥ 5 (h) Q
5. Show that the infinite direct product G = Z2 × Z2 × · · · is not finitely
generated.
6. Let G be an abelian group of order m. If n divides m, prove that G has a
subgroup of order n.
7. A group G is a torsion group if every element of G has finite order. Prove
that a finitely generated torsion group must be finite.
8. Let G, H, and K be finitely generated abelian groups. Show that if G × H ∼
=
G × K, then H ∼ = K. Give a counterexample to show that this cannot be
true in general.
9. Let G and H be solvable groups. Show that G × H is also solvable.
200 CHAPTER 11 THE STRUCTURE OF GROUPS
G = Pn ⊃ Pn−1 ⊃ · · · ⊃ P1 ⊃ P0 = {e}
21. Suppose that G is a solvable group with order n ≥ 2. Show that G contains
a normal nontrivial abelian factor group.
22. Zassenhaus Lemma. Let H and K be subgroups of a group G. Suppose
also that H ∗ and K ∗ are normal subgroups of H and K respectively. Then
(c) H ∗ (H ∩ K)/H ∗ (H ∩ K ∗ ) ∼
= K ∗ (H ∩ K)/K ∗ (H ∗ ∩ K)
∼
= (H ∩ K)/(H ∗ ∩ K)(H ∩ K ∗ ).
[Hint: Use the diagram in Figure 11.1. The Zassenhaus Lemma is often
referred to as the Butterfly Lemma because of this diagram.]
H K
QQ
B
Q B
Q B
Q
H ∗ (H ∩ K) Q H ∩ K K ∗ (H ∩ K)B
Q B
Q
B
∗ ∗ ∗ ∗ B
H (H ∩ K )
Q H
K (H ∩ K)B
A HH B
QQ A HH
Q B
A HH B
∗
HB K ∗
Q
H Q A
@ Q A
Q
@ ∗ ∗Q
@ (H ∩ K)(H ∩ K ) QQA
A
@ QA
H∗ ∩ K H ∩ K∗
23. Schreier’s Theorem. Use the Zassenhaus Lemma to prove that two sub-
normal (normal) series of a group G have isomorphic refinements.
24. Use Schreier’s Theorem to prove the Jordan-Hölder Theorem.
Programming Exercises
Write a program that will compute all possible abelian groups of order n. What is
the largest n for which your program will work?
203
204 CHAPTER 12 GROUP ACTIONS
(σ, x) 7→ σ(x)
for σ ∈ G and x ∈ X.
Example 3. If we let X = G, then every group G acts on itself by the
left regular representation; that is, (g, x) 7→ λg (x) = gx, where λg is left
multiplication:
e · x = λe x = ex = x
(gh) · x = λgh x = λg λh x = λg (hx) = g · (h · x).
If H is a subgroup of G, then G is an H-set under left multiplication by
elements of H.
Example 4. Let G be a group and suppose that X = G. If H is a subgroup
of G, then G is an H-set under conjugation; that is, we can define an action
of H on G,
H × G → G,
via
(h, g) 7→ hgh−1
for h ∈ H and g ∈ G. Clearly, the first axiom for a group action holds.
Observing that
Again, it is easy to see that the first axiom is true. Since (gg 0 )xH = g(g 0 xH),
the second axiom is also true.
12.1 GROUPS ACTING ON SETS 205
X(1) = X,
X(35)(46) = {1, 2},
X(12)(3456) = X(12)(3654) = ∅,
206 CHAPTER 12 GROUP ACTIONS
G1 = G2 = {(1), (35)(46)},
G3 = G4 = G5 = G6 = {(1)}.
Proof. Clearly, e ∈ Gx since the identity fixes every element in the set X.
Let g, h ∈ Gx . Then gx = x and hx = x. So (gh)x = g(hx) = gx = x;
hence, the product of two elements in Gx is also in Gx . Finally, if g ∈ Gx ,
then x = ex = (g −1 g)x = (g −1 )gx = g −1 x. So g −1 is in Gx .
We will denote the number of elements in the fixed point set of an element
g ∈ G by |Xg | and denote the number of elements in the orbit of x of x ∈ X
by |Ox |. The next theorem demonstrates the relationship between orbits of
an element x ∈ X and the left cosets of Gx in G.
x2 = g2 x = g1 gx = g1 x = x1 ;
consequently, the map φ is one-to-one. Finally, we must show that the map
φ is onto. Let gGx be a left coset. If gx = y, then φ(y) = gGx .
12.2 THE CLASS EQUATION 207
is the set of points that are fixed by conjugation. The nontrivial orbits of
the action are called the conjugacy classes of G. If x1 , . . . , xk are repre-
sentatives from each of the nontrivial conjugacy classes of G and |Ox1 | =
n1 , . . . , |Oxk | = nk , then
|G| = |Z(G)| + n1 + · · · + nk .
One of the consequences of the class equation is that the order of each
conjugacy class must divide the order of |G|.
Example 8. It is easy to check that the conjugacy classes in S3 are the
following:
{(1)}, {(123), (132)}, {(12), (13), (23)}.
The class equation is 6 = 1 + 2 + 3.
Example 9. The conjugacy classes for D4 are
Example 10. For Sn it takes a bit of work to find the conjugacy classes. We
begin with cycles. Suppose that σ = (a1 , . . . , ak ) is a cycle and let τ ∈ Sn .
By Theorem 5.9,
τ στ −1 = (τ (a1 ), . . . , τ (ak )).
Consequently, any two cycles of the same length are conjugate. Now let
σ = σ1 σ2 · · · σr be a cycle decomposition, where the length of each cycle σi
is ri . Then σ is conjugate to every other τ ∈ Sn whose cycle decomposition
has the same lengths.
The number of conjugate classes in Sn is the number of ways in which
n can be partitioned into sums of positive integers. For example, we can
partition the integer 3 into the following three sums:
3 = 1+1+1
3 = 1+2
3 = 3;
therefore, there are three conjugacy classes. The problem of finding the
number of such partitions for any positive integer n is what computer scien-
tists call NP-complete. This effectively means that the problem cannot be
solved for a large n because the computations would be too time-consuming
for even the largest computer.
|G| = |Z(G)| + n1 + · · · + nk .
Since each ni > 1 and ni | G, p must divide each ni . Also, p | |G|; hence, p
must divide |Z(G)|. Since the identity is always in the center of G, |Z(G)| ≥
1. Therefore, |Z(G)| ≥ p and there exists some g ∈ Z(G) such that g 6= 1.
G/Z(G), we can write any element gZ(G) in the quotient group as am Z(G)
for some integer m; hence, g = am x for some x in the center of G. Similarly,
if hZ(G) ∈ G/Z(G), there exists a y in Z(G) such that h = an y for some
integer n. Since x and y are in the center of G, they commute with all other
elements of G; therefore,
B W W B
W W W W
W W W W
B W W B
gag −1 · y = ga · g −1 y = ga · x = g · x = y,
Suppose that φ(a) = φ(b). Then gag −1 = gbg −1 or a = b; hence, the map is
injective. To show that φ is onto, let b be in Gy ; then g −1 bg is in Gx since
g −1 bg · x = g −1 b · gx = g −1 b · y = g −1 · y = x;
Proof. We look at all the fixed points x of all the elements in g ∈ G; that
is, we look at all g’s and all x’s such that gx = x. If viewed in terms of fixed
point sets, the number of all g’s fixing x’s is
X
|Xg |.
g∈G
Example 11. Let X = {1, 2, 3, 4, 5} and suppose that G is the permutation
group G = {(1), (13), (13)(25), (25)}. The orbits of X are {1, 3}, {2, 5}, and
{4}. The fixed point sets are
X(1) = X
X(13) = {2, 4, 5}
X(13)(25) = {4}
X(25) = {1, 3, 4}.
A Geometric Example
Before we apply Burnside’s Theorem to switching-theory problems, let us
examine the number of ways in which the vertices of a square can be colored
black or white. Notice that we can sometimes obtain equivalent colorings
by simply applying a rigid motion to the square. For instance, as we have
pointed out, if we color one of the vertices black and the remaining three
white, it does not matter which vertex was colored black since a rotation
will give an equivalent coloring.
The symmetry group of a square, D4 , is given by the following permu-
tations:
(1) (13) (24) (1432)
(1234) (12)(34) (14)(23) (13)(24)
The group G acts on the set of vertices {1, 2, 3, 4} in the usual manner. We
can describe the different colorings by mappings from X into Y = {B, W }
where B and W represent the colors black and white, respectively. Each map
f : X → Y describes a way to color the corners of the square. Every σ ∈ D4
212 CHAPTER 12 GROUP ACTIONS
f (1) = B
f (2) = W
f (3) = W
f (4) = W
3. |X
e(1432) | = 2.
e(13)(24) | = 22 = 4.
e(13)(24) , f (1) = f (3) and f (2) = f (4). Thus, |X
4. For X
5. |X
e(12)(34) | = 4.
6. |X
e(14)(23) | = 4.
7. For X
e(13) , f (1) = f (3) and the other corners can be of any color;
e(13) | = 23 = 8.
hence, |X
8. |X
e(24) | = 8.
1 4
(2 + 21 + 22 + 21 + 22 + 22 + 23 + 23 ) = 6
8
ways to color the vertices of the square.
12.3 BURNSIDE’S COUNTING THEOREM 213
Example 13. Suppose that we wish to color the vertices of a square using
four different colors. By Proposition 12.8, we can immediately decide that
there are
1 4
(4 + 41 + 42 + 41 + 42 + 42 + 43 + 43 ) = 55
8
possible ways.
x1 -
x2 -
.. f - f (x1 , x2 , . . . , xn )
.
xn -
Switching Functions
In switching theory we are concerned with the design of electronic circuits
with binary inputs and outputs. The simplest of these circuits is a switching
function that has n inputs and a single output (Figure 12.2). Large electronic
circuits can often be constructed by combining smaller modules of this kind.
The inherent problem here is that even for a simple circuit a large number
of different switching functions can be constructed. With only four inputs
and a single output, we can construct 65, 536 different switching functions.
However, we can often replace one switching function with another merely
by permuting the input leads to the circuit (Figure 12.3).
a - a
f - f (a, b) A f - f (b, a) = g(a, b)
b - b AU
f2 ∼ f4
f3 ∼ f5
f10 ∼ f12
f11 ∼ f13 .
3
For three input variables there are 22 = 256 possible switching func-
4
tions; in the case of four variables there are 22 = 65,536. The number of
equivalence classes is too large to reasonably calculate directly. It is neces-
sary to employ Burnside’s Theorem.
12.3 BURNSIDE’S COUNTING THEOREM 215
Now let us consider a circuit with four input variables and a single out-
put. Suppose that we can permute the leads of any circuit according to the
following permutation group:
The permutations of the four possible input variables induce the permuta-
tions of the output values in Table 12.2.
Hence, there are
1 16
(2 + 2 · 212 + 2 · 26 + 3 · 210 ) = 9616
8
possible switching functions of four variables under this group of permuta-
tions. This number will be even smaller if we consider the full symmetric
group on four letters.
Historical Note
William Burnside was born in London in 1852. He attended Cambridge University
from 1871 to 1875 and won the Smith’s Prize in his last year. After his graduation
he lectured at Cambridge. He was made a member of the Royal Society in 1893.
Burnside wrote approximately 150 papers on topics in applied mathematics, differ-
ential geometry, and probability, but his most famous contributions were in group
theory. Several of Burnside’s conjectures have stimulated research to this day. One
such conjecture was that every group of odd order is solvable; that is, for a group
G of odd order, there exists a sequence of subgroups
G = Hn ⊃ Hn−1 ⊃ · · · ⊃ H1 ⊃ H0 = {e}
EXERCISES 217
such that Hi is normal in Hi+1 and Hi+1 /Hi is abelian. This conjecture was finally
proven by W. Feit and J. Thompson in 1963. Burnside’s The Theory of Groups
of Finite Order, published in 1897, was one of the first books to treat groups in a
modern context as opposed to permutation groups. The second edition, published
in 1911, is still a classic.
Exercises
1. Compute the G-equivalence classes for Examples 1–5 in the first section.
2. Compute all Xg and all Gx for each of the following permutation groups.
(a) X = {1, 2, 3},
G = S3 = {(1), (12), (13), (23), (123), (132)}
(b) X = {1, 2, 3, 4, 5, 6},
G = {(1), (12), (345), (354), (12)(345), (12)(354)}
3. Compute the G-equivalence classes of X for each of the G-sets in Exercise 2.
For each x ∈ X verify that |G| = |Ox | · |Gx |.
4. Let G be the additive group of real numbers. Let the action of θ ∈ G on
the real plane R2 be given by rotating the plane counterclockwise about the
origin through θ radians. Let P be a point on the plane other than the origin.
(a) Show that R2 is a G-set.
(b) Describe geometrically the orbit containing P .
(c) Find the group GP .
5. Let G = A4 and suppose that G acts on itself by conjugation; that is,
(g, h) 7→ ghg −1 .
(a) Determine the conjugacy classes (orbits) of each element of G.
(b) Determine all of the isotropy subgroups for each element of G.
6. Find the conjugacy classes and the class equation for each of the following
groups.
(a) S4 (b) D5
(c) Z9 (d) Q8
7. Write the class equation for S5 and for A5 .
8. If a square remains fixed in the plane, how many different ways can the
corners of the square be colored if three colors are used?
9. How many ways can the vertices of an equilateral triangle be colored using
three different colors?
218 CHAPTER 12 GROUP ACTIONS
10. Find the number of ways a six-sided die can be constructed if each side is
marked differently with 1, . . . , 6 dots.
11. Up to a rotation, how many ways can the faces of a cube be colored with
three different colors?
12. Consider 12 straight wires of equal lengths with their ends soldered together
to form the edges of a cube. Either silver or copper wire can be used for each
edge. How many different ways can the cube be constructed?
13. Suppose that we color each of the eight corners of a cube. Using three
different colors, how many ways can the corners be colored up to a rotation
of the cube?
14. Each of the faces of a regular tetrahedron can be painted either red or white.
Up to a rotation, how many different ways can the tetrahedron be painted?
15. Suppose that the vertices of a regular hexagon are to be colored either red or
white. How many ways can this be done up to a symmetry of the hexagon?
16. A molecule of benzene is made up of six carbon atoms and six hydrogen
atoms, linked together in a hexagonal shape as in Figure 12.4.
(a) How many different compounds can be formed by replacing one or more
of the hydrogen atoms with a chlorine atom?
(b) Find the number of different chemical compounds that can be formed
by replacing three of the six hydrogen atoms in a benzene ring with a
CH3 radical.
H @ H
@ @@
@ @
H @@ H
17. How many equivalence classes of switching functions are there if the input
variables x1 , x2 , and x3 can be permuted by any permutation in S3 ? What if
the input variables x1 , x2 , x3 , and x4 can be permuted by any permutation
in S4 ?
EXERCISES 219
18. How many equivalence classes of switching functions are there if the input
variables x1 , x2 , x3 , and x4 can be permuted by any permutation in the
subgroup of S4 generated by the permutation (x1 x2 x3 x4 )?
19. A striped necktie has 12 bands of color. Each band can be colored by one of
four possible colors. How many possible different-colored neckties are there?
20. A group acts faithfully on a G-set X if the identity is the only element of
G that leaves every element of X fixed. Show that G acts faithfully on X
if and only if no two distinct elements of G have the same action on each
element of X.
21. Let p be prime. Show that the number of different abelian groups of order pn
(up to isomorphism) is the same as the number of conjugacy classes in Sn .
22. Let a ∈ G. Show that for any g ∈ G, gC(a)g −1 = C(gag −1 ).
23. Let |G| = pn and suppose that |Z(G)| = pn−1 for p prime. Prove that G is
abelian.
24. Let G be a group with order pn where p is prime and X a finite G-set. If
XG = {x ∈ X : gx = x for all g ∈ G} is the set of elements in X fixed by the
group action, then prove that |X| ≡ |XG | (mod p).
Programming Exercise
Write a program to compute the number of conjugacy classes in Sn . What is the
largest n for which your program will work?
Theorem 13.1 (Cauchy) Let G be a finite group and p a prime such that
p divides the order of G. Then G contains a subgroup of order p.
220
13.1 THE SYLOW THEOREMS 221
First suppose that p does not divide [G : C(xi )] for some i. Then
pr | |C(xi )|, since pr divides |G| = |C(xi )| · [G : C(xi )]. Now we can ap-
ply the induction hypothesis to C(xi ).
222 CHAPTER 13 THE SYLOW THEOREMS
Hence, we may assume that p divides [G : C(xi )] for all i. Since p divides
|G|, the class equation says that p must divide |Z(G)|; hence, by Cauchy’s
Theorem, Z(G) has an element of order p, say g. Let N be the group
generated by g. Clearly, N is a normal subgroup of Z(G) since Z(G) is
abelian; therefore, N is normal in G since every element in Z(G) commutes
with every element in G. Now consider the factor group G/N of order |G|/p.
By the induction hypothesis, G/N contains a subgroup H of order pr−1 . The
inverse image of H under the canonical homomorphism φ : G → G/N is a
subgroup of order pr in G.
A Sylow p-subgroup P of a group G is a maximal p-subgroup of G.
To prove the other two Sylow Theorems, we need to consider conjugate
subgroups as opposed to conjugate elements in a group. For a group G, let
S be the collection of all subgroups of G. For any subgroup H, S is a H-set,
where H acts on S by conjugation. That is, we have an action
H ×S →S
defined by
h · K 7→ hKh−1
for K in S.
The set
N (H) = {g ∈ G : gHg −1 = H}
is a subgroup of G. Notice that H is a normal subgroup of N (H). In fact,
N (H) is the largest subgroup of G in which H is normal. We call N (H) the
normalizer of H in G.
|G| = pr m = |N (P )| · [G : N (P )] = |N (P )| · k.
Theorem 13.7 (Third Sylow Theorem) Let G be a finite group and let
p be a prime dividing the order of G. Then the number of Sylow p-subgroups
is congruent to 1 (mod p) and divides |G|.
P = {P = P1 , P2 , . . . , Pk },
by conjugation. From the proof of the Second Sylow Theorem, the only
P -conjugate of P is itself and the order of the other P -conjugacy classes is a
224 CHAPTER 13 THE SYLOW THEOREMS
Historical Note
Peter Ludvig Mejdell Sylow was born in 1832 in Christiania, Norway (now Oslo).
After attending Christiania University, Sylow taught high school. In 1862 he ob-
tained a temporary appointment at Christiania University. Even though his ap-
pointment was relatively brief, he influenced students such as Sophus Lie (1842–
1899). Sylow had a chance at a permanent chair in 1869, but failed to obtain the
appointment. In 1872, he published a 10-page paper presenting the theorems that
now bear his name. Later Lie and Sylow collaborated on a new edition of Abel’s
works. In 1898, a chair at Christiania University was finally created for Sylow
through the efforts of his student and colleague Lie. Sylow died in 1918.
element of a Sylow 2-subgroup other than the identity must have as its order
a power of 2; and therefore cannot be one of the 48 elements of order 7 in
the Sylow 7-subgroups. Since a Sylow 2-subgroup has order 8, there is only
enough room for a single Sylow 2-subgroup in a group of order 56. If there
is only one Sylow 2-subgroup, it must be normal.
For other groups G it is more difficult to prove that G is not simple.
Suppose G has order 48. In this case the technique that we employed in the
last example will not work. We need the following lemma to prove that no
group of order 48 is simple.
|H| · |K|
|HK| = .
|H ∩ K|
HK = {hk : h ∈ H, k ∈ K}.
h2 = h1 a−1
k2 = ak1 .
Suppose that the other case is true, and two of the three Sylow 2-
subgroups are H and K. We claim that |H ∩ K| = 8. If |H ∩ K| ≤ 4,
then by Lemma 13.10,
16 · 16
|HK| = = 64,
4
which is impossible. So H ∩ K is normal in both H and K since it has index
2. The normalizer of H ∩ K contains both H and K, and |H ∩ K| must
both be a multiple of 16 greater than 1 and divide 48. The only possibility
is that |N (H ∩ K)| = 48. Hence, N (H ∩ K) = G.
The following famous conjecture of Burnside was proved in a long and
difficult paper by Feit and Thompson [2].
The proof of this theorem laid the groundwork for a program in the
1960s and 1970s that classified all finite simple groups. The success of this
program is one of the outstanding achievements of modern mathematics.
Exercises
1. What are the orders of all Sylow p-subgroups where G has order 18, 24, 54,
72, and 80?
2. Find all the Sylow 3-subgroups of S4 and show that they are all conjugate.
3. Show that every group of order 45 has a normal subgroup of order 9.
4. Let H be a Sylow p-subgroup of G. Prove that H is the only Sylow p-subgroup
of G contained in N (H).
5. Prove that no group of order 96 is simple.
6. Prove that no group of order 160 is simple.
7. If H is a normal subgroup of a finite group G and |H| = pk for some prime
p, show that H is contained in every Sylow p-subgroup of G.
8. Let G be a group of order p2 q 2 , where p and q are distinct primes such that
q6 | p2 − 1 and p6 | q 2 − 1. Prove that G must be abelian. List three pairs of
primes satisfying these conditions.
9. Show directly that a group of order 33 has only one Sylow 3-subgroup.
10. Let H be a subgroup of a group G. Prove or disprove that the normalizer of
H is normal in G.
EXERCISES 229
11. Let G be a finite group divisible by a prime p. Prove that if there is only one
Sylow p-subgroup in G, it must be a normal subgroup of G.
12. Let G be a group of order pr , p prime. Prove that G contains a normal
subgroup of order pr−1 .
13. Suppose that G is a finite group of order pn k, where k < p. Show that G
must contain a normal subgroup.
14. Let H be a subgroup of a finite group G. Prove that gN (H)g −1 = N (gHg −1 )
for any g ∈ G.
15. Prove that a group of order 108 must have a normal subgroup.
16. Classify all the groups of order 175 up to isomorphism.
17. Show that every group of order 255 is cyclic.
18. Let G have order pe11 · · · penn and suppose that G has n Sylow p-subgroups
P1 , . . . , Pn where |Pi | = pei i . Prove that G is isomorphic to P1 × · · · × Pn .
19. Let P be a normal Sylow p-subgroup of G. Prove that every inner automor-
phism of G fixes P .
20. What is the smallest possible order of a group G such that G is nonabelian
and |G| is odd? Can you find such a group?
21. The Frattini Lemma. If H is a normal subgroup of a finite group G and
P is a Sylow p-subgroup of H, for each g ∈ G show that there is an h in H
such that gP g −1 = hP h−1 . Also, show that if N is the normalizer of P , then
G = HN .
22. Show that if the order of G is pn q, where p and q are primes and p > q, then
G contains a normal subgroup.
23. Prove that the number of distinct conjugates of a subgroup H of a finite
group G is [G : N (H)].
24. Prove that a Sylow 2-subgroup of S5 is isomorphic to D4 .
25. Another Proof of the Sylow Theorems.
(a) Suppose p is prime and p does not divide m. Show that
k
p m
p6 | .
pk
(b) Let S denote the set of all pk element subsets of G. Show that p does
not divide |S|.
(c) Define an action of G on S by left multiplication, aT = {at : t ∈ T } for
a ∈ G and T ∈ S. Prove that this is a group action.
(d) Prove p6 | |OT | for some T ∈ S.
230 CHAPTER 13 THE SYLOW THEOREMS
A Project
The main objective of finite group theory is to classify all possible finite groups up
to isomorphism. This problem is very difficult even if we try to classify the groups
of order less than or equal to 60. However, we can break the problem down into
several intermediate problems.
1. Find all simple groups G ( |G| ≤ 60). Do not use the Odd Order Theorem
unless you are prepared to prove it.
2. Find the number of distinct groups G, where the order of G is n for n =
1, . . . , 60.
3. Find the actual groups (up to isomorphism) for each n.
This is a challenging project that requires a working knowledge of the group theory
you have learned up to this point. Even if you do not complete it, it will teach you
a great deal about finite groups. You can use Table 13.1 as a guide.
EXERCISES 231
Up to this point we have studied sets with a single binary operation satis-
fying certain axioms, but often we are more interested in working with sets
that have two binary operations. For example, one of the most natural alge-
braic structures to study is the integers with the operations of addition and
multiplication. These operations are related to one another by the distribu-
tive property. If we consider a set with two such related binary operations
satisfying certain axioms, we have an algebraic structure called a ring. In a
ring we add and multiply such elements as real numbers, complex numbers,
matrices, and functions.
14.1 Rings
A nonempty set R is a ring if it has two closed binary operations, addition
and multiplication, satisfying the following conditions.
1. a + b = b + a for a, b ∈ R.
2. (a + b) + c = a + (b + c) for a, b, c ∈ R.
6. For a, b, c ∈ R,
a(b + c) = ab + ac
(a + b)c = ac + bc.
232
14.1 RINGS 233
This last condition, the distributive axiom, relates the binary operations of
addition and multiplication. Notice that the first four axioms simply require
that a ring be an abelian group under addition, so we could also have defined
a ring to be an abelian group (R, +) together with a second binary operation
satisfying the fifth and sixth conditions given above.
If there is an element 1 ∈ R such that 1 6= 0 and 1a = a1 = a for
each element a ∈ R, we say that R is a ring with unity or identity . A
ring R for which ab = ba for all a, b in R is called a commutative ring.
A commutative ring R with identity is called an integral domain if, for
every a, b ∈ R such that ab = 0, either a = 0 or b = 0. A division ring
is a ring R, with an identity, in which every nonzero element in R is a
unit; that is, for each a ∈ R with a 6= 0, there exists a unique element a−1
such that a−1 a = aa−1 = 1. A commutative division ring is called a field.
The relationship among rings, integral domains, division rings, and fields is
shown in Figure 14.1.
Rings
" b
" b
Commutative Rings with
Rings Identity
"
"
"
"
Integral Division
Domains Rings
b "
b "
Fields
i2 = j2 = k2 = −1
ij = k
jk = i
ki = j
ji = −k
kj = −i
ik = −j.
(a1 + b1 i + c1 j + d1 k) + (a2 + b2 i + c2 j + d2 k) =
(a1 + a2 ) + (b1 + b2 )i + (c1 + c2 )j + (d1 + d2 )k
and
where
α = a1 a2 − b1 b2 − c1 c2 − d1 d2
β = a1 b2 + a1 b1 + c1 d2 − d1 c2
γ = a1 c2 − b1 d2 + c1 a2 − d1 b2
δ = a1 d2 + b1 c2 − c1 b2 − d1 a2 .
(a + bi + cj + dk)(a − bi − cj − dk) = a2 + b2 + c2 + d2 .
1. a0 = 0a = 0;
3. (−a)(−b) = ab.
a0 = a(0 + 0) = a0 + a0;
Z ⊂ Q ⊂ R ⊂ C.
2. rs ∈ S for all r, s ∈ S.
3. r − s ∈ S for all r, s ∈ S.
then T is a subring of R. If
0 0
a b a b
A= and B =
0 c 0 c0
14.2 INTEGRAL DOMAINS AND FIELDS 237
1 < b < n. Since 0 = n1 = (ab)1 = (a1)(b1) and there are no zero divisors
in D, either a1 = 0 or b1 = 0. Hence, the characteristic of D must be less
than n, which is a contradiction. Therefore, n must be prime.
φ(a + b) = (a + b) (mod n)
= a (mod n) + b (mod n)
= φ(a) + φ(b)
and
φ(ab) = ab (mod n)
= a (mod n) · b (mod n)
= φ(a)φ(b).
2. φ(0) = 0.
Example 15. Every ring R has at least two ideals, {0} and R. These ideals
are called the trivial ideals.
Theorem 14.9 Let I be an ideal of R. The factor group R/I is a ring with
multiplication defined by
(r + I)(s + I) = rs + I.
Proof. We already know that R/I is an abelian group under addition. Let
r + I and s + I be in R/I. We must show that the product (r + I)(s + I) =
rs+I is independent of the choice of coset; that is, if r0 ∈ r +I and s0 ∈ s+I,
then r0 s0 must be in rs + I. Since r0 ∈ r + I, there exists an element a in
I such that r0 = r + a. Similarly, there exists a b ∈ I such that s0 = s + b.
Notice that
r0 s0 = (r + a)(s + b) = rs + as + rb + ab
and as + rb + ab ∈ I since I is an ideal; consequently, r0 s0 ∈ rs + I. We will
leave as an exercise the verification of the associative law for multiplication
and the distributive laws.
The ring R/I in Theorem 14.9 is called the factor or quotient ring.
Just as with group homomorphisms and normal subgroups, there is a rela-
tionship between ring homomorphisms and ideals.
I/I ∩ J ∼
= (I + J)/J.
R/J
R/I ∼
= .
I/J
(a + P )(b + P ) = ab + P = 0 + P = P.
Example 20. Every ideal in Z is of the form nZ. The factor ring Z/nZ ∼ = Zn
is an integral domain only when n is prime. It is actually a field. Hence, the
nonzero prime ideals in Z are the ideals pZ, where p is prime. This example
really justifies the use of the word “prime” in our definition of prime ideals.
Historical Note
Amalie Emmy Noether, one of the outstanding mathematicians of this century, was
born in Erlangen, Germany in 1882. She was the daughter of Max Noether (1844–
1921), a distinguished mathematician at the University of Erlangen. Together with
Paul Gordon (1837–1912), Emmy Noether’s father strongly influenced her early
education. She entered the University of Erlangen at the age of 18. Although
women had been admitted to universities in England, France, and Italy for decades,
there was great resistance to their presence at universities in Germany. Noether
was one of only two women among the university’s 986 students. After completing
her doctorate under Gordon in 1907, she continued to do research at Erlangen,
occasionally lecturing when her father was ill.
Noether went to Göttingen to study in 1916. David Hilbert and Felix Klein
tried unsuccessfully to secure her an appointment at Göttingen. Some of the faculty
objected to women lecturers, saying, “What will our soldiers think when they return
to the university and are expected to learn at the feet of a woman?” Hilbert,
annoyed at the question, responded, “Meine Herren, I do not see that the sex of
a candidate is an argument against her admission as a Privatdozent. After all,
246 CHAPTER 14 RINGS
the Senate is not a bathhouse.” At the end of World War I, attitudes changed
and conditions greatly improved for women. After Noether passed her habilitation
examination in 1919, she was given a title and was paid a small sum for her lectures.
In 1922, Noether became a Privatdozent at Göttingen. Over the next 11 years
she used axiomatic methods to develop an abstract theory of rings and ideals.
Though she was not good at lecturing, Noether was an inspiring teacher. One of her
many students was B. L. van der Waerden, author of the first text treating abstract
algebra from a modern point of view. Some of the other mathematicians Noether
influenced or closely worked with were Alexandroff, Artin, Brauer, Courant, Hasse,
Hopf, Pontryagin, von Neumann, and Weyl. One of the high points of her career
was an invitation to address the International Congress of Mathematicians in Zurich
in 1932. In spite of all the recognition she received from her colleagues, Noether’s
abilities were never recognized as they should have been during her lifetime. She
was never promoted to full professor by the Prussian academic bureaucracy.
In 1933, Noether, a Jew, was banned from participation in all academic activi-
ties in Germany. She emigrated to the United States, took a position at Bryn Mawr
College, and became a member of the Institute for Advanced Study at Princeton.
Noether died suddenly on April 14, 1935. After her death she was eulogized by
such notable scientists as Albert Einstein.
x ≡ a (mod m)
x ≡ b (mod n)
has a solution. Furthermore, any two solutions of the system are congruent
modulo n1 n2 · · · nk .
x ≡ a1 (mod n1 )
x ≡ a2 (mod n2 )
..
.
x ≡ ak+1 (mod nk+1 ).
Considering the first k equations, there exists a solution that is unique mod-
ulo n1 · · · nk , say a. Since n1 · · · nk and nk+1 are relatively prime, the system
x ≡ a (mod n1 · · · nk )
x ≡ ak+1 (mod nk+1 )
x ≡ 3 (mod 4)
x ≡ 4 (mod 5)
x ≡ 1 (mod 9)
x ≡ 5 (mod 7).
x ≡ 19 (mod 20)
x ≡ 1 (mod 9)
x ≡ 5 (mod 7).
x ≡ 19 (mod 180)
x ≡ 5 (mod 7).
14.5 AN APPLICATION TO SOFTWARE DESIGN 249
Solving this last system, we find that 19 is a solution for the system that is
unique up to modulo 1260.
Example 23. Suppose that we wish to multiply 2134 by 1531. We will use
the integers 95, 97, 98, and 99 because they are relatively prime. We can
break down each integer into four parts:
and
x ≡ 9 (mod 95)
x ≡ 0 (mod 97)
x ≡ 30 (mod 98)
x ≡ 55 (mod 99).
Exercises
1. Which of the following sets are rings with respect to the usual operations of
addition and multiplication? If the set is a ring, is it also a field?
(a) 7Z
(b) Z18
√ √
(c) Q( 2 ) = {a + b 2 : a, b ∈ Q}
√ √ √ √ √
(d) Q( 2, 3 ) = {a + b 2 + c 3 + d 6 : a, b, c, d ∈ Q}
√ √
(e) Z[ 3 ] = {a + b 3 : a, b ∈ Z}
√
(f ) R = {a + b 3 3 : a, b ∈ Q}
(g) Z[i] = {a + bi : a, b ∈ Z and i2 = −1 }
√ √ √
(h) Q( 3 3 ) = {a + b 3 3 + c 3 9 : a, b, c ∈ Q}
EXERCISES 251
11. Prove that the Gaussian integers, Z[i], are an integral domain.
√ √
12. Prove that Z[ 3 i] = {a + b 3 i : a, b ∈ Z} is an integral domain.
13. Solve each of the following systems of congruences.
14. Use the method of parallel computation outlined in the text to calculate
2234 + 4121 by dividing the calculation into four separate additions modulo
95, 97, 98, and 99.
15. Explain why the method of parallel computation outlined in the text fails
for 2134 · 1531 if we attempt to break the calculation down into two smaller
calculations modulo 98 and 99.
16. If R is a field, show that the only two ideals of R are {0} and R itself.
17. Let a be any element in a ring R with identity. Show that (−1)a = −a.
18. Prove that (−a)(−b) = ab for any elements a and b in a ring R.
19. Let φ : R → S be a ring homomorphism. Prove each of the following state-
ments.
(a) If R is a commutative ring, then φ(R) is a commutative ring.
(b) φ(0) = 0.
(c) Let 1R and 1S be the identities for R and S, respectively. If φ is onto,
then φ(1R ) = 1S .
(d) If R is a field and φ(R) 6= 0, then φ(R) is a field.
20. Prove that the associative law for multiplication and the distributive laws
hold in R/I.
21. Prove the Second Isomorphism Theorem for rings: Let I be a subring of a
ring R and J an ideal in R. Then I ∩ J is an ideal in I and
I/I ∩ J ∼
= I + J/J.
EXERCISES 253
22. Prove the Third Isomorphism Theorem for rings: Let R be a ring and I and
J be ideals of R, where J ⊂ I. Then
R/J
R/I ∼
= .
I/J
(a) S 6= ∅.
(b) rs ∈ S for all r, s ∈ S.
(c) r − s ∈ S for all r, s ∈ S.
T
25. Let R be a ring with a collection of subrings {Rα }. Prove that Rα is a
subring of R. Give an example to show that the union of two subrings cannot
be a subring.
T
26. Let {Iα }α∈A be a collection of ideals in a ring R. Prove that α∈A Iα is also
an ideal in R. Give an example to show that if I1 and I2 are ideals in R,
then I1 ∪ I2 may not be an ideal.
27. Let R be an integral domain. Show that if the only ideals in R are {0} and
R itself, R must be a field.
28. Let R be a commutative ring. An element a in R is nilpotent if an = 0 for
some positive integer n. Show that the set of all nilpotent elements forms an
ideal in R.
29. A ring R is a Boolean ring if for every a ∈ R, a2 = a. Show that every
Boolean ring is a commutative ring.
30. Let R be a ring, where a3 = a for all a ∈ R. Prove that R must be a
commutative ring.
31. Let R be a ring with identity 1R and S a subring of R with identity 1S .
Prove or disprove that 1R = 1S .
32. If we do not require the identity of a ring to be distinct from 0, we will not
have a very interesting mathematical structure. Let R be a ring such that
1 = 0. Prove that R = {0}.
33. Let S be a subset of a ring R. Prove that there is a subring R0 of R that
contains S.
254 CHAPTER 14 RINGS
φ : U (R) → Inn(R)
x ≡ r (mod I)
x ≡ s (mod J)
has a solution.
EXERCISES 255
(b) In addition, prove that any two solutions of the system are congruent
modulo I ∩ J.
(c) Let I and J be ideals in a ring R such that I + J = R. Show that there
exists a ring isomorphism
R/(I ∩ J) ∼
= R/I × R/J.
Programming Exercise
Write a computer program to simulate fast addition and multiplication using
the Chinese Remainder Theorem and the method outlined in the text.
Most people are fairly familiar with polynomials by the time they begin to
study abstract algebra. When we examine polynomial expressions such as
p(x) = x3 − 3x + 2
q(x) = 3x2 − 6x + 5,
we have a pretty good idea of what p(x) + q(x) and p(x)q(x) mean. We just
add and multiply polynomials as functions; that is,
and
(pq)(x) = p(x)q(x)
= (x3 − 3x + 2)(3x2 − 6x + 5)
= 3x5 − 6x4 − 4x3 + 24x2 − 27x + 10.
256
15.1 POLYNOMIAL RINGS 257
p(x) = a0 + a1 x + · · · + an xn
q(x) = b0 + b1 x + · · · + bm xm ,
p(x) = a0 + a1 x + · · · + an xn
q(x) = b0 + b1 x + · · · + bm xm .
p(x) + q(x) = c0 + c1 x + · · · + ck xk ,
where
i
X
ci = ak bi−k = a0 bi + a1 bi−1 + · · · + ai−1 b1 + ai b0
k=0
for each i. Notice that in each case some of the coefficients may be zero.
258 CHAPTER 15 POLYNOMIALS
and
q(x) = 2 + 0x − x2 + 0x3 + 4x4
The product,
Example 2. Let
p(x) = 3 + 3x3
and
q(x) = 4 + 4x2 + 4x4
be polynomials in Z12 [x]. The sum of p(x) and q(x) is 7 + 4x2 + 3x3 + 4x4 .
The product of the two polynomials is the zero polynomial. This example
tells us that R[x] cannot be an integral domain if R is not an integral domain.
Proof. Our first task is to show that R[x] is an abelian group under
polynomial addition. The zero
Pnpolynomial, f (x) = 0, is the additive identity.
Given a polynomial p(x) = a x i , the inverse of p(x) is easily verified to
P i
i=0
be −p(x) = ni=0 (−ai )xi = − ni=0 ai xi . Commutativity and associativity
P
follow immediately from the definition of polynomial addition and from the
fact that addition in R is both commutative and associative.
15.1 POLYNOMIAL RINGS 259
m
! n+p i
X X X
= ai xi bj ci−j xi
i=0 i=0 j=0
m n p
!" ! !#
X X X
i i i
= ai x bi x ci x
i=0 i=0 i=0
= p(x)[q(x)r(x)]
The commutativity and distribution properties of polynomial multiplication
are proved in a similar manner. We shall leave the proofs of these properties
as an exercise.
Proposition 15.2 Let p(x) and q(x) be polynomials in R[x], where R is an
integral domain. Then deg p(x) + deg q(x) = deg(p(x)q(x)). Furthermore,
R[x] is an integral domain.
260 CHAPTER 15 POLYNOMIALS
Theorem 15.4 (Division Algorithm) Let f (x) and g(x) be two nonzero
polynomials in F [x], where F is a field and g(x) is a nonconstant polynomial.
Then there exist unique polynomials q(x), r(x) ∈ F [x] such that
where either deg r(x) < deg g(x) or r(x) is the zero polynomial.
Proof. We will first consider the existence of q(x) and r(x). Let S =
{f (x) − g(x)h(x) : h(x) ∈ F [x]} and assume that
g(x) = a0 + a1 x + · · · + an xn
0 = f (x) = 0 · g(x) + 0;
or
f (x) = g(x)q(x) + r(x).
We need to show that the degree of r(x) is less than the degree of g(x).
Assume that deg g(x) ≤ deg r(x). Say r(x) = b0 + b1 x + · · · + bm xm and
262 CHAPTER 15 POLYNOMIALS
m ≥ n. Then
f (x) − g(x)[q(x) − (bm /an )xm−n ] = f (x) − g(x)q(x)
+ (bm /an )xm−n g(x)
= r(x) + (bm /an )xm−n g(x)
= r(x) + bm xm
+ terms of lower degree
is in S. This is a polynomial of lower degree than r(x), which contradicts
the fact that r(x) is a polynomial of smallest degree in S; hence, deg r(x) <
deg g(x).
To show that q(x) and r(x) are unique, suppose that there exist two
other polynomials q 0 (x) and r0 (x) such that f (x) = g(x)q 0 (x) + r0 (x) and
deg r0 (x) < deg g(x) or r0 (x) = 0, so that
f (x) = g(x)q(x) + r(x) = g(x)q 0 (x) + r0 (x),
and
g(x)[q(x) − q 0 (x)] = r0 (x) − r(x).
If g is not the zero polynomial, then
deg(g(x)[q(x) − q 0 (x)]) = deg(r0 (x) − r(x)) ≥ deg g(x).
However, the degrees of both r(x) and r0 (x) are strictly less than the degree
of g(x); therefore, r(x) = r0 (x) and q(x) = q 0 (x).
Example 3. The division algorithm merely formalizes long division of poly-
nomials, a task we have been familiar with since high school. For example,
suppose that we divide x3 − x2 + 2x − 3 by x − 2.
x2 + x + 4
x − 2 x3 − x2 + 2x − 3
x3 − 2x2
x2 + 2x − 3
x2 − 2x
4x − 3
4x − 8
5
Hence, x3 − x2 + 2x − 3 = (x − 2)(x2 + x + 4) + 5.
Let p(x) be a polynomial in F [x] and α ∈ F . We say that α is a zero
or root of p(x) if p(x) is in the kernel of the evaluation homomorphism φα .
All we are really saying here is that α is a zero of p(x) if p(α) = 0.
15.2 THE DIVISION ALGORITHM 263
and the degree of r(x) must be less than the degree of x − α. Since the
degree of r(x) is less than 1, r(x) = a for a ∈ F ; therefore,
p(x) = (x − α)q(x) + a.
But
0 = p(α) = 0 · q(x) + a = a;
Proof. We will use induction on the degree of p(x). If deg p(x) = 0, then
p(x) is a constant polynomial and has no zeros. Let deg p(x) = 1. Then
p(x) = ax + b for some a and b in F . If α1 and α2 are zeros of p(x), then
aα1 + b = aα2 + b or α1 = α2 .
Now assume that deg p(x) > 1. If p(x) does not have a zero in F , then we
are done. On the other hand, if α is a zero of p(x), then p(x) = (x − α)q(x)
for some q(x) ∈ F [x] by Corollary 15.5. The degree of q(x) is n − 1 by
Proposition 15.2. Let β be some other zero of p(x) that is distinct from α.
Then p(β) = (β − α)q(β) = 0. Since α 6= β and F is a field, q(β) = 0. By
our induction hypothesis, p(x) can have at most n − 1 zeros in F that are
distinct from α. Therefore, p(x) has at most n distinct zeros in F .
Let F be a field. A monic polynomial d(x) is a greatest common
divisor of polynomials p(x), q(x) ∈ F [x] if d(x) evenly divides both p(x)
and q(x); and, if for any other polynomial d0 (x) dividing both p(x) and q(x),
d0 (x) | d(x). We write d(x) = gcd(p(x), q(x)). Two polynomials p(x) and
q(x) are relatively prime if gcd(p(x), q(x)) = 1.
264 CHAPTER 15 POLYNOMIALS
Proposition 15.7 Let F be a field and suppose that d(x) is the greatest
common divisor of two polynomials p(x) and q(x) in F [x]. Then there exist
polynomials r(x) and s(x) such that
Proof. Let d(x) be the monic polynomial of smallest degree in the set
We can write d(x) = r(x)p(x) + s(x)q(x) for two polynomials r(x) and s(x)
in F [x]. We need to show that d(x) divides both p(x) and q(x). We shall
first show that d(x) divides p(x). By the division algorithm, there exist
polynomials a(x) and b(x) such that p(x) = a(x)d(x) + b(x), where b(x) is
either the zero polynomial or deg b(x) < deg d(x). Therefore,
Since d0 (x) | d(x), d(x) is a greatest common divisor of p(x) and q(x).
Finally, we must show that the greatest common divisor of p(x) and
q(x)) is unique. Suppose that d0 (x) is another greatest common divisor of
15.3 IRREDUCIBLE POLYNOMIALS 265
p(x) and q(x). We have just shown that there exist polynomials u(x) and
v(x) in F [x] such that d(x) = d0 (x)[r(x)u(x) + s(x)v(x)]. Since
and d(x) and d0 (x) are both greatest common divisors, deg d(x) = deg d0 (x).
Since d(x) and d0 (x) are both monic polynomials of the same degree, it must
be the case that d(x) = d0 (x).
Notice the similarity between the proof of Proposition 15.7 and the proof
of Theorem 1.4.
p(0) = 2
p(1) = 1
p(2) = 2.
b0 b1 bn
p(x) = + x + · · · + xn ,
c0 c1 cn
1
p(x) = (d0 + d1 x + · · · + dn xn ),
c0 · · · cn
where d0 , . . . , dn are integers. Let d be the greatest common divisor of
d0 , . . . , dn . Then
d
p(x) = (a0 + a1 x + · · · + an xn ),
c0 · · · cn
where c/d is the product of c1 /d1 and c2 /d2 expressed in lowest terms.
Hence, dp(x) = cα1 (x)β1 (x).
15.3 IRREDUCIBLE POLYNOMIALS 267
Proof. Let p(x) have a zero a ∈ Q. Then p(x) must have a linear factor
x − a. By Gauss’s Lemma, p(x) has a factorization with a linear factor in
Z[x]. Hence, for some α ∈ Z
Thus a0 /α ∈ Z and so α | a0 .
Example 6. Let p(x) = x4 − 2x3 + x + 1. We shall show that p(x) is
irreducible over Q[x]. Assume that p(x) is reducible. Then either p(x) has
a linear factor, say p(x) = (x − α)q(x), where q(x) is a polynomial of degree
three, or p(x) has two quadratic factors.
If p(x) has a linear factor in Q[x], then it has a zero in Z. By Corol-
lary 15.10, any zero must divide 1 and therefore must be ±1; however,
p(1) = 1 and p(−1) = 3. Consequently, we have eliminated the possibility
that p(x) has any linear factors.
Therefore, if p(x) is reducible it must factor into two quadratic polyno-
mials, say
a + c = −2
ac + b + d = 0
ad + bc = 1
bd = 1.
ad + bc = b(a + c) = 1.
Proof. By Gauss’s Lemma, we need only show that f (x) does not factor
into polynomials of lower degree in Z[x]. Let
am = b0 cm + b1 cm−1 + · · · + bm c0
is not divisible by p, since each term on the right-hand side of the equation
is divisible by p except for b0 cm . Therefore, m = n since ai is divisible by p
for m < n. Hence, f (x) cannot be factored into polynomials of lower degree
and therefore must be irreducible.
Example 7. The polynomial
Ideals in F [x]
Let F be a field. Recall that a principal ideal in F [x] is an ideal hp(x)i
generated by some polynomial p(x); that is,
Theorem 15.13 Let F be a field and suppose that p(x) ∈ F [x]. Then the
ideal generated by p(x) is maximal if and only if p(x) is irreducible.
270 CHAPTER 15 POLYNOMIALS
Proof. Suppose that p(x) generates a maximal ideal of F [x]. Then hp(x)i is
also a prime ideal of F [x]. Since a maximal ideal must be properly contained
inside F [x], p(x) cannot be a constant polynomial. Let us assume that p(x)
factors into two polynomials of lesser degree, say p(x) = f (x)g(x). Since
hp(x)i is a prime ideal one of these factors, say f (x), is in hp(x)i and therefore
be a multiple of p(x). But this would imply that hp(x)i ⊂ hf (x)i, which is
impossible since hp(x)i is maximal.
Conversely, suppose that p(x) is irreducible over F [x]. Let I be an ideal
in F [x] containing hp(x)i. By Theorem 15.12, I is a principal ideal; hence,
I = hf (x)i for some f (x) ∈ F [x]. Since p(x) ∈ I, it must be the case
that p(x) = f (x)g(x) for some g(x) ∈ F [x]. However, p(x) is irreducible;
hence, either f (x) or g(x) is a constant polynomial. If f (x) is constant,
then I = F [x] and we are done. If g(x) is constant, then f (x) is a constant
multiple of I and I = hp(x)i. Thus, there are no proper ideals of F [x] that
properly contain hp(x)i.
Historical Note
He kept his solution an absolute secret. This may seem surprising today, when
mathematicians are usually very eager to publish their results, but in the days
of the Italian Renaissance secrecy was customary. Academic appointments were
not easy to secure and depended on the ability to prevail in public contests. Such
challenges could be issued at any time. Consequently, any major new discovery was
a valuable weapon in such a contest. If an opponent presented a list of problems
to be solved, del Ferro could in turn present a list of depressed cubics. He kept the
secret of his discovery throughout his life, passing it on only on his deathbed to his
student Antonio Fior (ca. 1506–?).
Although Fior was not the equal of his teacher, he immediately issued a chal-
lenge to Niccolo Fontana (1499–1557). Fontana was known as Tartaglia (the Stam-
merer). As a youth he had suffered a blow from the sword of a French soldier during
an attack on his village. He survived the savage wound, but his speech was perma-
EXERCISES 271
ax3 + bx2 + cx + d = 0.
Cardano shared the secret with his student, Ludovico Ferrari (1522–1565), who
solved the general quartic equation,
In 1543, Cardano and Ferrari examined del Ferro’s papers and discovered that he
had also solved the depressed cubic. Cardano felt that this relieved him of his
obligation to Tartaglia, so he proceeded to publish the solutions in Ars Magna
(1545), in which he gave credit to del Ferro for solving the special case of the cubic.
This resulted in a bitter dispute between Cardano and Tartaglia, who published
the story of the oath a year later.
Exercises
1. List all of the polynomials of degree 3 or less in Z2 [x].
2. Compute each of the following.
(a) (5x2 + 3x − 4) + (4x2 − x + 9) in Z12
(b) (5x2 + 3x − 4)(4x2 − x + 9) in Z12
(c) (7x3 + 3x2 − x) + (6x2 − 8x + 4) in Z9
(d) (3x2 + 2x − 4) + (4x2 + 2) in Z5
(e) (3x2 + 2x − 4)(4x2 + 2) in Z5
(f) (5x2 + 3x − 2)2 in Z12
3. Use the division algorithm to find q(x) and r(x) such that a(x) = q(x)b(x) +
r(x) with deg r(x) < deg b(x) for each of the following pairs of polynomials.
(a) p(x) = 5x3 + 6x2 − 3x + 4 and q(x) = x − 2 in Z7 [x]
(b) p(x) = 6x4 − 2x3 + x2 − 3x + 1 and q(x) = x2 + x − 2 in Z7 [x]
272 CHAPTER 15 POLYNOMIALS
ax2 + bx + c = 0
to obtain √
−b ± b2 − 4ac
x= .
2a
The discriminant of the quadratic equation ∆ = b2 − 4ac determines the
nature of the solutions of the equation. If ∆ > 0, the equation has two
distinct real solutions. If ∆ = 0, the equation has a single repeated real root.
If ∆ < 0, there are two distinct imaginary solutions.
2. Show that any cubic equation of the form
x3 + bx2 + cx + d = 0
and use this result to show that the three possible solutions for y are
s r s r
q p3 q 2 q p3 q2
i 3 i 3
ω − + + +ω − − + ,
2 27 4 2 27 4
where i = 0, 1, 2.
EXERCISES 275
p3 q2
∆= + .
27 4
Show that y 3 + py + q = 0
(a) has three real roots, at least two of which are equal, if ∆ = 0.
(b) has one real root and two conjugate imaginary roots if ∆ > 0.
(c) has three distinct real roots if ∆ < 0.
8. Solve the following cubic equations.
x4 + ax3 + bx2 + cx + d = 0
can be reduced to
y 4 + py 2 + qy + r = 0
by using the substitution x = y − a/4.
10. Show that 2
1 1 2
y2 + z = (z − p)y 2 − qy + z −r .
2 4
11. Show that the right-hand side of (10) can be put in the form (my + k)2 if
and only if
1 2
q 2 − 4(z − p) z − r = 0.
4
z 3 − pz 2 − 4rz + (4pr − q 2 ) = 0.
Solving the resolvent cubic equation, put the equation found in (10) in the
form 2
2 1
y + z = (my + k)2
2
to obtain the solution of the quartic equation.
13. Use this method to solve the following quartic equations.
276 CHAPTER 15 POLYNOMIALS
One of the most important rings we study is the ring of integers. It was our
first example of an algebraic structure: the first polynomial ring that we
examined was Z[x]. We also know that the integers sit naturally inside the
field of rational numbers, Q. The ring of integers is the model for all integral
domains. In this chapter we will examine integral domains in general, an-
swering questions about the ideal structure of integral domains, polynomial
rings over integral domains, and whether or not an integral domain can be
embedded in a field.
277
278 CHAPTER 16 INTEGRAL DOMAINS
a c ad + bc
+ = ;
b d bd
a c ac
· = .
b d bd
It seems reasonable to define the operations of addition and multiplication
on FD in a similar manner. If we denote the equivalence class of (a, b) ∈ S by
[a, b], then we are led to define the operations of addition and multiplication
on FD by
[a, b] + [c, d] = [ad + bc, bd]
16.1 FIELDS OF FRACTIONS 279
and
[a, b] · [c, d] = [ac, bd],
respectively. The next lemma demonstrates that these operations are inde-
pendent of the choice of representatives from each equivalence class.
[a1 d1 + b1 c1 , b1 d1 ] = [a2 d2 + b2 c2 , b2 d2 ]
(a1 d1 + b1 c1 )(b2 d2 ) = a1 d1 b2 d2 + b1 c1 b2 d2
= a1 b2 d1 d2 + b1 b2 c1 d2
= b1 a2 d1 d2 + b1 b2 d1 c2
= (b1 d1 )(a2 d2 + b2 c2 ).
is a field.
Proof. The additive and multiplicative identities are [0, 1] and [1, 1], re-
spectively. To show that [0, 1] is the additive identity, observe that
and
φ(ab) = [ab, 1] = [a, 1][b, 1] = φ(a)φ(b);
hence, φ is a homomorphism. To show that φ is one-to-one, suppose that
φ(a) = φ(b). Then [a, 1] = [b, 1], or a = a1 = 1b = b. Finally, any element
of FD can expressed as the quotient of two elements in D, since
and
Therefore, ψ is a homomorphism.
To complete the proof of the theorem, we need to show that ψ is one-to-
one. Suppose that ψ([a, b]) = ab−1 = 0. Then a = 0b = 0 and [a, b] = [0, b].
Therefore, the kernel of ψ is the zero element [0, b] in FD , and ψ is injective.
Example 1. Since Q is a field, Q[x] is an integral domain. The field
of fractions of Q[x] is the set of all rational expressions p(x)/q(x), where
p(x) and q(x) are polynomials over the rationals and q(x) is not the zero
polynomial. We will denote this field by Q(x).
We will leave the proofs of the following corollaries of Theorem 16.4 as
exercises.
1. a | b ⇔ hbi ⊂ hai.
3. a is a unit in D ⇔ hai = D.
Theorem 16.8 Let D be a PID and hpi be a nonzero ideal in D. Then hpi
is a maximal ideal if and only if p is irreducible.
Proof. Let p be irreducible and suppose that p | ab. Then habi ⊂ hpi. By
Corollary 14.17, since hpi is a maximal ideal, hpi must also be a prime ideal.
Thus, either a ∈ hpi or b ∈ hpi. Hence, either p | a or p | b.
By Lemma 16.10, there exists a positive integer N such that han i = haN i
for all n ≥ N . Consequently, aN must be irreducible. We have now shown
that a is the product of two elements, one of which must be irreducible.
Now suppose that a = c1 p1 , where p1 is irreducible. If c1 is not a unit,
we can repeat the preceding argument to conclude that hai ⊂ hc1 i. Either
c1 is irreducible or c1 = c2 p2 , where p2 is irreducible and c2 is not a unit.
Continuing in this manner, we obtain another chain of ideals
hai ⊂ hc1 i ⊂ hc2 i ⊂ · · · .
This chain must satisfy the ascending chain condition; therefore,
a = p1 p2 · · · pr
for irreducible elements p1 , . . . , pr .
Uniqueness of the factorization. To show uniqueness, let
a = p1 p2 · · · pr = q1 q2 · · · qs ,
where each pi and each qi is irreducible. Without loss of generality, we can
assume that r < s. Since p1 divides q1 q2 · · · qs , by Corollary 16.9 it must
divide some qi . By rearranging the qi ’s, we can assume that p1 | q1 ; hence,
q1 = u1 p1 for some unit u1 in D. Therefore,
a = p1 p2 · · · pr = u1 p1 q2 · · · qs
or
p2 · · · pr = u1 q2 · · · qs .
Continuing in this manner, we can arrange the qi ’s such that p2 = q2 , p3 =
q3 , . . . , pr = qr , to obtain
u1 u2 · · · ur qr+1 · · · qs = 1.
In this case qr+1 · · · qs is a unit, which contradicts the fact that qr+1 , . . . , qs
are irreducibles. Therefore, r = s and the factorization of a is unique.
Corollary 16.12 Let F be a field. Then F [x] is a UFD.
Example 5. Every PID is a UFD, but it is not the case that every UFD
is a PID. In Corollary 16.22, we will prove that Z[x] is a UFD. However,
Z[x] is not a PID. Let I = {5f (x) + xg(x) : f (x), g(x) ∈ Z[x]}. We can
easily show that I is an ideal of Z[x]. Suppose that I = hp(x)i. Since 5 ∈ I,
5 = f (x)p(x). In this case p(x) = p must be a constant. Since x ∈ I,
x = pg(x); consequently, p = ±1. However, it follows from this fact that
hp(x)i = Z[x]. But this would mean that 3 is in I. Therefore, we can write
3 = 5f (x) + xg(x) for some f (x) and g(x) in Z[x]. Examining the constant
term of this polynomial, we see that 3 = 5f (x), which is impossible.
286 CHAPTER 16 INTEGRAL DOMAINS
Euclidean Domains
We have repeatedly used the division algorithm when proving results about
either Z or F [x], where F is a field. We should now ask when a division
algorithm is available for an integral domain.
Let D be an integral domain such that for each a ∈ D there is a non-
negative integer ν(a) satisfying the following conditions.
1. If a and b are nonzero elements in D, then ν(a) ≤ ν(ab).
2. Let a, b ∈ D and suppose that b 6= 0. Then there exist elements
q, r ∈ D such that a = bq + r and either r = 0 or ν(r) < ν(b).
Then D is called a Euclidean domain and ν is called a Euclidean val-
uation.
Example 6. Absolute value on Z is a Euclidean valuation.
Example 7. Let F be a field. Then the degree of a polynomial in F [x] is
a Euclidean valuation.
Example 8. Recall that the Gaussian integers in Example 9 of Chapter 14
are defined by
Z[i] = {a + bi : a, b ∈ Z}.
We usually measure√ the size of a complex
√ number a + bi by its absolute
value, |a + bi| = a + b ; however, a + b2 may not be an integer. For
2 2 2
our valuation we will let ν(a+bi) = a2 +b2 to ensure that we have an integer.
We claim that ν(a + bi) = a2 + b2 is a Euclidean valuation on Z[i]. Let
z, w ∈ Z[i]. Then ν(zw) = |zw|2 = |z|2 |w|2 = ν(z)ν(w). Since ν(z) ≥ 1 for
every nonzero z ∈ Z[i], ν(z) = ν(z)ν(w).
Next, we must show that for any z = a + bi and w = c + di in Z[i]
with w 6= 0, there exist elements q and r in Z[i] such that z = qw + r
with either r = 0 or ν(r) < ν(w). We can view z and w as elements in
Q(i) = {p + qi : p, q ∈ Q}, the field of fractions of Z[i]. Observe that
c − di
zw−1 = (a + bi)
c2 + d2
ac + bd bc − ad
= + 2 i
c2 + d2
c +d2
n1 n2
= m1 + 2 + m2 + 2 i
c + d2 c + d2
n1 n2
= (m1 + m2 i) + + i
c2 + d2 c2 + d2
= (m1 + m2 i) + (s + ti)
16.2 FACTORIZATION IN INTEGRAL DOMAINS 287
in Q(i). In the last steps we are writing the real and imaginary parts as an
integer plus a proper fraction. That is, we take the closest integer mi such
that the fractional part satisfies |ni /(a2 + b2 )| ≤ 1/2. For example, we write
9 1
= 1+
8 8
15 1
= 2− .
8 8
Thus, s and t are the “fractional parts” of zw−1 = (m1 + m2 i) + (s + ti).
We also know that s2 + t2 ≤ 1/4 + 1/4 = 1/2. Multiplying by w, we have
Factorization in D[x]
One of the most important polynomial rings is Z[x]. One of the first ques-
tions that come to mind about Z[x] is whether or not it is a UFD. We will
prove a more general statement here. Our first task is to obtain a more
general version of Gauss’s Lemma (Theorem 15.9).
Let D be a unique factorization domain and suppose that
p(x) = an xn + · · · + a1 x + a0
288 CHAPTER 16 INTEGRAL DOMAINS
Theorem 16.15 (Gauss’s Lemma) Let D be a UFD and let f (x) and
g(x) be primitive polynomials in D[x]. Then f (x)g(x) is primitive.
Lemma 16.16 Let D be a UFD, and let p(x) and q(x) be in D[x]. Then the
content of p(x)q(x) is equal to the product of the contents of p(x) and q(x).
Proof. Let p(x) = cp1 (x) and q(x) = dq1 (x), where c and d are the
contents of p(x) and q(x), respectively. Then p1 (x) and q1 (x) are primitive.
We can now write p(x)q(x) = cdp1 (x)q1 (x). Since p1 (x)q1 (x) is primitive,
the content of p(x)q(x) must be cd.
Proof. Let a and b be nonzero elements of D such that af (x), bg(x) are
in D[x]. We can find a1 , b2 ∈ D such that af (x) = a1 f1 (x) and bg(x) =
b1 g1 (x), where f1 (x) and g1 (x) are primitive polynomials in D[x]. Therefore,
abp(x) = (a1 f1 (x))(b1 g1 (x)). Since f1 (x) and g1 (x) are primitive polynomi-
als, it must be the case that ab | a1 b1 by Gauss’s Lemma. Thus there exists
a c ∈ D such that p(x) = cf1 (x)g1 (x). Clearly, deg f (x) = deg f1 (x) and
deg g(x) = deg g1 (x).
The following corollaries are direct consequences of Lemma 16.17.
16.2 FACTORIZATION IN INTEGRAL DOMAINS 289
Historical Note
Gauss was clearly the most prominent mathematician in the world in the
early nineteenth century. His status naturally made his discoveries subject
to intense scrutiny. Gauss’s cold and distant personality many times led him
to ignore the work of his contemporaries, making him many enemies. He
did not enjoy teaching very much, and young mathematicians who sought
him out for encouragement were often rebuffed. Nevertheless, he had many
outstanding students, including Eisenstein, Riemann, Kummer, Dirichlet,
and Dedekind. Gauss also offered a great deal of encouragement to Sophie
Germain (1776–1831), who overcame the many obstacles facing women in
her day to become a very prominent mathematician. Gauss died at the age
of 78 in Göttingen on February 23, 1855.
Exercises
√ √ 2
1. Let z = a + b 3 i be in Z[ 3 i].
√ If a + 3b = 1, show that z must be a unit.
Show that the only units of Z[ 3 i] are 1 and −1.
2. The Gaussian integers, Z[i], are a UFD. Factor each of the following elements
in Z[i] into a product of irreducibles.
(a) 5 (c) 6 + 8i
(b) 1 + 3i (d) 2
7. Let p be prime and denote the field of fractions of Zp [x] by Zp (x). Prove
that Zp (x) is an infinite field of characteristic p.
8. Prove that the field of fractions of the Gaussian integers, Z[i], is
Q(i) = {p + qi : p, q ∈ Q}.
(a) If D is a PID and a and b are both nonzero elements of D, prove there
exists a unique greatest common divisor of a and b. We write gcd(a, b)
for the greatest common divisor of a and b.
(b) Let D be a PID and a and b be nonzero elements of D. Prove that
there exist elements s and t in D such that gcd(a, b) = as + bt.
The axioms of a ring give structure to the operations of addition and multi-
plication on a set. However, we can construct algebraic structures, known as
lattices and Boolean algebras, that generalize other types of operations. For
example, the important operations on sets are inclusion, union, and intersec-
tion. Lattices are generalizations of order relations on algebraic spaces, such
as set inclusion in set theory and inequality in the familiar number systems
N, Z, Q, and R. Boolean algebras generalize the operations of intersection
and union. Lattices and Boolean algebras have found applications in logic,
circuit theory, and probability.
17.1 Lattices
Partially Ordered Sets
We begin by the study of lattices and Boolean algebras by generalizing the
idea of inequality. Recall that a relation on a set X is a subset of X × X.
A relation P on X is called a partial order of X if it satisfies the following
axioms.
294
17.1 LATTICES 295
{a, b, c}
! aa
!!
! a
a
{a, b} {a, c} {b, c}
a
aa!! a
aa!!
!! a !! a
{a} {b} {c}
a
aa !
!
a !!
∅
24
"" bb
8 12
"
"
"
"
4 6
"
"
"
"
2 3
bb ""
1
Proof. By the Principle of Duality, we need only prove the first statement
in each part.
(1) By definition a ∨ b is the least upper bound of {a, b}, and b ∨ a is the
least upper bound of {b, a}; however, {a, b} = {b, a}.
298 CHAPTER 17 LATTICES AND BOOLEAN ALGEBRAS
Theorem 17.3 Let L be a nonempty set with two binary operations ∨ and
∧ satisfying the commutative, associative, idempotent, and absorption laws.
We can define a partial order on L by a b if a ∨ b = b. Furthermore, L is
a lattice with respect to if for all a, b ∈ L, we define the least upper bound
and greatest lower bound of a and b by a ∨ b and a ∧ b, respectively.
a ∨ c = a ∨ (b ∨ c) = (a ∨ b) ∨ c = b ∨ c = c,
or a c.
To show that L is a lattice, we must prove that a ∨ b and a ∧ b are,
respectively, the least upper and greatest lower bounds of a and b. Since
a = (a ∨ b) ∧ a = a ∧ (a ∨ b), it follows that a a ∨ b. Similarly, b a ∨ b.
Therefore, a ∨ b is an upper bound for a and b. Let u be any other upper
bound of both a and b. Then a u and b u. But a ∨ b u since
(a ∨ b) ∨ u = a ∨ (b ∨ u) = a ∨ u = u.
17.2 BOOLEAN ALGEBRAS 299
A0 = X \ A = {x : x ∈ X and x ∈
/ A}.
a ∧ (b ∨ c) = (a ∧ b) ∨ (a ∧ c);
A ∩ (B ∪ C) = (A ∩ B) ∪ (A ∩ C)
a ∧ (b ∨ c) = (a ∧ b) ∨ (a ∧ c)
for all a, b, c ∈ L.
a ∨ (b ∧ c) = (a ∨ b) ∧ (a ∨ c)
for all a, b, c ∈ L.
300 CHAPTER 17 LATTICES AND BOOLEAN ALGEBRAS
Observe that
I ∨ b = (I ∨ b) ∧ I = (I ∧ I) ∨ (b ∧ I) = I ∨ I = I.
a ∨ (a ∧ b) = (a ∧ I) ∨ (a ∧ b)
= a ∧ (I ∨ b)
= a∧I
= a.
The other idempotent and absorption laws are proven similarly. Since B
also satisfies (1)–(3), the conditions of Theorem 17.3 are met; therefore, B
must be a lattice. Condition (4) tells us that B is a distributive lattice.
For a ∈ B, O ∨ a = a; hence, O a and O is the smallest element in B.
To show that I is the largest element in B, we will first show that a ∨ b = b
is equivalent to a ∧ b = a. Since a ∨ I = a for all a ∈ B, using the absorption
laws we can determine that
a ∨ I = (a ∧ I) ∨ I = I ∨ (I ∧ a) = I
3. If a ∨ b = I and a ∧ b = O, then b = a0 .
5. I 0 = O and O0 = I.
Proof. We will prove only (2). The rest of the identities are left as exercises.
For a ∨ b = a ∨ c and a ∧ b = a ∧ c, we have
b = b ∨ (b ∧ a)
= b ∨ (a ∧ b)
= b ∨ (a ∧ c)
= (b ∨ a) ∧ (b ∨ c)
= (a ∨ b) ∧ (b ∨ c)
= (a ∨ c) ∧ (b ∨ c)
= (c ∨ a) ∧ (c ∨ b)
= c ∨ (a ∧ b)
= c ∨ (a ∧ c)
= c ∨ (c ∧ a)
= c.
Theorem 17.12 Let B be a finite Boolean algebra. Then there exists a set
X such that B is isomorphic to P(X).
Corollary 17.13 The order of any finite Boolean algebra must be 2n for
some positive integer n.
A a b B
Figure 17.3. a ∧ b
306 CHAPTER 17 LATTICES AND BOOLEAN ALGEBRAS
a
A B
b
Figure 17.4. a ∨ b
We can build more complicated electrical circuits out of series and par-
allel circuits by replacing any switch in the circuit with one of these two
fundamental types of circuits. Circuits constructed in this manner are called
series-parallel circuits.
We will consider two circuits equivalent if they act the same. That is,
if we set the switches in equivalent circuits exactly the same we will obtain
the same result. For example, in a series circuit a ∧ b is exactly the same as
b ∧ a. Notice that this is exactly the commutative law for Boolean algebras.
In fact, the set of all series-parallel circuits forms a Boolean algebra under the
operations of ∨ and ∧. We can use diagrams to verify the different axioms
of a Boolean algebra. The distributive law, a ∧ (b ∨ c) = (a ∧ b) ∨ (a ∧ c),
is illustrated in Figure 17.5. If a is a switch, then a0 is the switch that is
always open when a is closed and always closed when a is open. A circuit
that is always closed is I in our algebra; a circuit that is always open is O.
The laws for a ∧ a0 = O and a ∨ a0 = I are shown in Figure 17.6.
b a b
a
c a c
Figure 17.5. a ∧ (b ∨ c) = (a ∧ b) ∨ (a ∧ c)
a
a a0
a0
a a a
b b0 b
Figure 17.7. (a ∨ b) ∧ (a ∨ b0 ) ∧ (a ∨ b)
Historical Note
George Boole (1815–1864) was the first person to study lattices. In 1847, he pub-
lished The Investigation of the Laws of Thought, a book in which he used lattices to
formalize logic and the calculus of propositions. Boole believed that mathematics
was the study of form rather than of content; that is, he was not so much concerned
with what he was calculating as with how he was calculating it. Boole’s work was
carried on by his friend Augustus De Morgan (1806–1871). De Morgan observed
that the principle of duality often held in set theory, as is illustrated by De Morgan’s
laws for set theory. He believed, as did Boole, that mathematics was the study of
symbols and abstract operations.
308 CHAPTER 17 LATTICES AND BOOLEAN ALGEBRAS
Set theory and logic were further advanced by such mathematicians as Alfred
North Whitehead (1861–1947), Bertrand Russell (1872–1970), and David Hilbert
(1862–1943). In Principia Mathematica, Whitehead and Russell attempted to show
the connection between mathematics and logic by the deduction of the natural
number system from the rules of formal logic. If the natural numbers could be
determined from logic itself, then so could much of the rest of existing mathematics.
Hilbert attempted to build up mathematics by using symbolic logic in a way that
would prove the consistency of mathematics. His approach was dealt a mortal blow
by Kurt Gödel (1906–1978), who proved that there will always be “undecidable”
problems in any sufficiently rich axiomatic system; that is, that in any mathematical
system of any consequence, there will always be statements that can never be proven
either true or false.
As often occurs, this basic research in pure mathematics later became indis-
pensable in a wide variety of applications. Boolean algebras and logic have become
essential in the design of the large-scale integrated circuitry found on today’s com-
puter chips. Sociologists have used lattices and Boolean algebras to model social
hierarchies; biologists have used them to describe biosystems.
Exercises
1. Draw the lattice diagram for the power set of X = {a, b, c, d} with the set
inclusion relation, ⊆.
2. Draw the diagram for the set of positive integers that are divisors of 30. Is
this poset a Boolean algebra?
4. Let B be the set of positive integers that are divisors of 36. Define an order
on B by a b if a | b. Prove that B is a Boolean algebra. Find a set X such
that B is isomorphic to P(X).
6. Draw the switching circuit for each of the following Boolean expressions.
7. Draw a circuit that will be closed exactly when only one of three switches a,
b, and c are closed.
a b c a b
a0 b
a c0 a c0
a b0
a0
a a b
a0
b a0 b
a b c
a0 b0 c
a b0 c0
11. Prove or disprove: The set of all nonzero integers is a lattice, where a b is
defined by a | b.
12. Prove that a ∧ b is the greatest lower bound of a and b in Theorem 17.3.
13. Let L be a nonempty set with two binary operations ∨ and ∧ satisfying the
commutative, associative, idempotent, and absorption laws. We can define a
partial order on L, as in Theorem 17.3, by a b if a ∨ b = b. Prove that the
greatest lower bound of a and b is a ∧ b.
14. Let G be a group and X be the set of subgroups of G ordered by set-theoretic
inclusion. If H and K are subgroups of G, show that the least upper bound
of H and K is the subgroup generated by H ∪ K.
310 CHAPTER 17 LATTICES AND BOOLEAN ALGEBRAS
15. Let R be a ring and suppose that X is the set of ideals of R. Show that X is
a poset ordered by set-theoretic inclusion, ⊆. Define the meet of two ideals
I and J in X by I ∩ J and the join of I and J by I + J. Prove that the set
of ideals of R is a lattice under these operations.
16. Let B be a Boolean algebra. Prove each of the following identities.
17. By drawing the appropriate diagrams, complete the proof of Theorem 17.14
to show that the switching functions form a Boolean algebra.
18. Let B be a Boolean algebra. Define binary operations + and · on B by
a+b = (a ∧ b0 ) ∨ (a0 ∧ b)
a · b = a ∧ b.
Programming Exercises
A Boolean or switching function on n variables is a map f : {O, I}n →
{0, I}. A Boolean polynomial is a special type of Boolean function: it is any
type of Boolean expression formed from a finite combination of variables x1 , . . . , xn
together with O and I, using the operations ∨, ∧, and 0 . The values of the functions
are defined in Table 17.1. Write a program to evaluate Boolean polynomials.
• (α + β)v = αv + βv;
• α(u + v) = αu + αv;
• 1v = v;
where α, β ∈ F and u, v ∈ V .
The elements of V are called vectors; the elements of F are called
scalars. It is important to notice that in most cases two vectors cannot be
312
18.1 DEFINITIONS AND EXAMPLES 313
Proposition 18.1 Let V be a vector space over F . Then each of the fol-
lowing statements is true.
1. 0v = 0 for all v ∈ V .
2. α0 = 0 for all α ∈ F .
3. If αv = 0, then either α = 0 or v = 0.
0v = (0 + 0)v = 0v + 0v;
18.2 Subspaces
Just as groups have subgroups and rings have subrings, vector spaces also
have substructures. Let V be a vector space over a field F , and W a subset
of V . Then W is a subspace of V if it is closed under vector addition and
scalar multiplication; that is, if u, v ∈ W and α ∈ F , it will always be the
case that u + v and αv are also in W .
v = α1 v1 + α2 v2 + · · · + αn vn = β1 v1 + β2 v2 + · · · + βn vn .
Then α1 = β1 , α2 = β2 , . . . , αn = βn .
Proof. If
v = α1 v1 + α2 v2 + · · · + αn vn = β1 v1 + β2 v2 + · · · + βn vn ,
then
(α1 − β1 )v1 + (α2 − β2 )v2 + · · · + (αn − βn )vn = 0.
Since v1 , . . . , vn are linearly independent, αi − βi = 0 for i = 1, . . . , n.
α1 v1 + α2 v2 + · · · + αn vn = 0,
with at least one of the αi ’s not equal to zero. Suppose that αk 6= 0. Then
α1 αk−1 αk+1 αn
vk = − v1 − · · · − vk−1 − vk+1 − · · · − vn .
αk αk αk αk
Conversely, suppose that
Then
From the last two examples it should be clear that a given vector space
has several bases. In fact, there are an infinite number of bases for both
of these examples. In general, there is no unique basis for a vector space.
However, 3
√ every basis of R consists of exactly three vectors, and every basis
of Q( 2 ) consists of exactly two vectors. This is a consequence of the next
proposition.
{v1 , . . . , vk , vk+1 , . . . , vn }
is a basis for V .
Exercises
1. If F is a field, show that F [x] is a vector space over F , where the vectors
in F [x] are polynomials. Vector addition is polynomial addition, and scalar
multiplication is defined by αp(x) for α ∈ F .
√
2. Prove that Q( 2 ) is a vector space.
√ √ √ √
3. Let Q( 2, 3 ) be the field generated√ by elements
√ of the form a + b 2 + c 3,
where a, b, c are in Q. Prove that
√ √Q( 2, 3 ) is a vector space of dimension
4 over Q. Find a basis for Q( 2, 3 ).
4. Prove that the complex numbers are a vector space of dimension 2 over R.
5. Prove that the set Pn of all polynomials of degree less than n form a subspace
of the vector space F [x]. Find a basis for Pn and compute the dimension
of Pn .
8. Show that the set of all possible solutions (x, y, z) ∈ R3 of the equations
Ax + By + Cz = 0
Dx + Ey + Cz = 0
forms a subspace of R3 .
9. Let W be the subset of continuous functions on [0, 1] such that f (0) = 0.
Prove that W is a subspace of C[0, 1].
10. Let V be a vector space over F . Prove that −(αv) = (−α)v = α(−v) for all
α ∈ F and all v ∈ V .
11. Let V be a vector space of dimension n. Prove each of the following state-
ments.
(a) If S = {v1 , . . . , vn } is a set of linearly independent vectors for V , then
S is a basis for V .
(b) If S = {v1 , . . . , vn } spans V , then S is a basis for V .
(c) If S = {v1 , . . . , vk } is a set of linearly independent vectors for V with
k < n, then there exist vectors vk+1 , . . . , vn such that
{v1 , . . . , vk , vk+1 , . . . , vn }
is a basis for V .
12. Prove that any set of vectors containing 0 is linearly dependent.
13. Let V be a vector space. Show that {0} is a subspace of V of dimension zero.
14. If a vector space V is spanned by n vectors, show that any set of m vectors
in V must be linearly dependent for m > n.
15. Linear Transformations. Let V and W be vector spaces over a field F , of
dimensions m and n, respectively. If T : V → W is a map satisfying
T (u + v) = T (u) + T (v)
T (αv) = αT (v)
(d) Let {v1 , . . . , vk } be a basis for the null space of T . We can extend this
basis to be a basis {v1 , . . . , vk , vk+1 , . . . , vm } of V . Why? Prove that
{T (vk+1 ), . . . , T (vm )} is a basis for the range of T . Conclude that the
range of T has dimension m − k.
(e) Let dim V = dim W . Show that a linear transformation T : V → W is
injective if and only if it is surjective.
16. Let V and W be finite dimensional vector spaces of dimension n over a field
F . Suppose that T : V → W is a vector space isomorphism. If {v1 , . . . , vn }
is a basis of V , show that {T (v1 ), . . . , T (vn )} is a basis of W . Conclude that
any vector space over a field F of dimension n is isomorphic to F n .
17. Direct Sums. Let U and V be subspaces of a vector space W . The sum of
U and V , denoted U + V , is defined to be the set of all vectors of the form
u + v, where u ∈ U and v ∈ V .
18. Dual Spaces. Let V and W be finite dimensional vector spaces over a
field F .
(a) Show that the set of all linear transformations from V into W , denoted
by Hom(V, W ), is a vector space over F , where we define vector addition
as follows:
(c) Consider the basis {(3, 1), (2, −2)} for R2 . What is the dual basis for
(R2 )∗ ?
(d) Let V be a vector space of dimension n over a field F and let V ∗∗ be the
dual space V ∗ . Show that each element v ∈ V gives rise to an element
λv in V ∗∗ and that the map v 7→ λv is an isomorphism of V with V ∗∗ .
We wish to be able to compute and study such fields for arbitrary polyno-
mials over a field F .
322
19.1 EXTENSION FIELDS 323
· 0 1 α 1+α
0 0 0 0 0
1 0 1 α 1+α
α 0 α 1+α 1
1+α 0 1+α 1 α
The following theorem, due to Kronecker, is so important and so basic
to our understanding of fields that it is often known as the Fundamental
Theorem of Field Theory.
Proof. To prove this theorem, we will employ the method that we used
to construct Example 2. Clearly, we can assume that p(x) is an irreducible
polynomial. We wish to find an extension field E of F containing an element
α such that p(α) = 0. The ideal hp(x)i generated by p(x) is a maximal ideal
in F [x] by Theorem 15.13; hence, F [x]/hp(x)i is a field. We claim that
E = F [x]/hp(x)i is the desired field.
We first show that E is a field extension of F . We can define a homo-
morphism of commutative rings by the map ψ : F → F [x]/h p(x) i, where
ψ(a) = a + hp(x)i for a ∈ F . It is easy to check that ψ is indeed a ring
homomorphism. Observe that
and
ψ(a)ψ(b) = (a + hp(x)i)(b + hp(x)i) = ab + hp(x)i = ψ(ab).
To prove that ψ is one-to-one, assume that
Algebraic Elements
An element α in an extension field E over F is algebraic over F if f (α) = 0
for some nonzero polynomial f (x) ∈ F [x]. An element in E that is not
algebraic over F is transcendental over F . An extension field E of a field
F is an algebraic extension of F if every element in E is algebraic over
F . If E is a field extension of F and α1 , . . . , αn are contained in E, we
denote the smallest field containing F and α1 , . . . , αn by F (α1 , . . . , αn ). If
E = F (α) for some α ∈ E, then E is a simple extension of F .
√
Example 4. Both 2 and i are algebraic over Q since they are zeros
of the polynomials x2 − 2 and x2 + 1, respectively. Clearly π and e are
algebraic over the real numbers; however, it is a nontrivial fact that they
are transcendental over Q. Numbers in R that are algebraic over Q are in fact
quite rare. Almost all real numbers are transcendental over Q.1 (In many
cases we do not know whether or not a particular number is transcendental;
for example, it is not known whether π + e is transcendental or algebraic.)
A complex number that is algebraic over Q is an algebraic number. A
transcendental number is an element of C that is transcendental over Q.
1
If we choose a number in R, then there is a probability of 1 that the number will be
transcendental over Q.
326 CHAPTER 19 FIELDS
p √
Example 5. We will show that 2 + 3 is algebraic over Q. If α =
p √ √ √
2 + 3, then α2 = 2 + 3. Hence, α2 − 2 = 3 and (α2 − 2)2 = 3.
Since α4 − 4α2 + 1 = 0, it must be true that α is a zero of the polynomial
x4 − 4x2 + 1 ∈ Q[x].
It is very easy to give an example of an extension field E over a field F ,
where E contains an element transcendental over F . The following theorem
characterizes transcendental extensions.
β = b0 + b1 α + · · · + bn−1 αn−1
for bi ∈ F .
αn = −an−1 αn−1 − · · · − a0 .
Similarly,
αn+1 = ααn
= −an−1 αn − an−2 αn−1 − · · · − a0 α
= −an−1 (−an−1 αn−1 − · · · − a0 ) − an−2 αn−1 − · · · − a0 α.
is in F [x] and g(α) = 0. Since the degree of g(x) is less than the degree
of p(x), the irreducible polynomial of α, g(x) must be the zero polynomial.
Consequently,
b0 − c0 = b1 − c1 = · · · = bn−1 − cn−1 = 0,
1, α, . . . , αn
cannot be linearly independent. Hence, there exist ai ∈ F , not all zero, such
that
an αn + an−1 αn−1 + · · · + a1 α + a0 = 0.
Therefore,
p(x) = an xn + · · · + a0 ∈ F [x]
is a nonzero polynomial with p(α) = 0.
Remark. Theorem 19.6 says that every finite extension of a field F is an
algebraic extension. The converse is false, however. We will leave it as an
exercise to show that the set of all elements in R that are algebraic over Q
forms an infinite field extension of Q.
The next theorem is a counting theorem, similar to Lagrange’s Theorem
in group theory. Theorem 19.6 will prove to be an extremely useful tool in
our investigation of finite field extensions.
[K : F ] = [K : E][E : F ].
c1 v1 + c2 v2 + · · · + cn vn = 0
implies that
c1 = c2 = · · · = cn = 0.
330 CHAPTER 19 FIELDS
Let X
u= cij (αi βj ) = 0
i,j
for cij ∈ F . We need to prove that all of the cij ’s are zero. We can rewrite
u as
m n
!
X X
cij αi βj = 0,
j=1 i=1
P
where i cij αi ∈ E. Since the βj ’s are linearly independent over E, it must
be the case that
Xn
cij αi = 0
i=1
for all j. However, the αj are also linearly independent over F . Therefore,
cij = 0 for all i and j, which completes the proof.
The following corollary is easily proved using mathematical induction.
Proof. We know that deg p(x) = [F (α) : F ] and deg q(x) = [F (β) : F ].
Since F ⊂ F (β) ⊂ F (α),
√ √
Example 8. Let us determine an extension field of Q√containing
√ 3+ 5. It
is easy to determine that the minimal polynomial of 3 + 5 is x4 − 16x + 4.
It follows that √ √
[Q( 3 + 5 ) : Q] = 4.
√ √ √ √
We know that √ {1, 3 } is a basis for
√ Q( 3 ) over Q. Hence,
√ 3 + 5 can-
not be in√Q( 3 ). It follows that
√ √5 cannot be √ in Q(
√ 3 ) either.√ There-
fore,√ {1,√ 5√
} is
√ a basis
√ for Q( 3, 5 ) = √ (Q(√3 ))( 5 )√over √Q( 3 ) and
{1, 3, 5, 3 5 = 15 } is a basis for Q( 3, 5 ) = Q( 3 + 5 ) over Q.
19.1 EXTENSION FIELDS 331
√
Notice that 6 5 i is a zero of x6 + 5. We can show that this polynomial is
irreducible over Q using Eisenstein’s Criterion, where we let p = 5. Conse-
quently, √ √ √
6 3
Q ⊂ Q( 5 ) ⊂ Q( 5, 5 i).
√ √ √
But it must be the case that Q( 6 5 i) = Q( 3 5, 5 i), since the degree of
both of these extensions is 6.
1. E is a finite extension of F .
Algebraic Closure
Given a field F , the question arises as to whether or not we can find a field
E such that every polynomial p(x) has a root in E. This leads us to the
following theorem.
Corollary 19.12 The set of all algebraic numbers forms a field; that is, the
set of all complex numbers that are algebraic over Q makes up a field.
where deg q2 (x) = deg p(x) − 2. The process must eventually stop since the
degree of p(x) is finite.
Conversely, suppose that every nonconstant polynomial p(x) in F [x] fac-
tors into linear factors. Let ax − b be such a factor. Then p(b/a) = 0.
Consequently, F is algebraically closed.
It is a nontrivial fact that every field has a unique algebraic closure. The
proof is not extremely difficult, but requires some rather sophisticated set
theory. We refer the reader to [3], [4], or [7] for a proof of this result.
We now state the Fundamental Theorem of Algebra, first proven by
Gauss at the age of 22 in his doctoral thesis. This theorem states that
every polynomial with coefficients in the complex numbers has a root in the
complex numbers. The proof of this theorem will be given in Chapter 21.
p(x) = (x − α1 )(x − α2 ) · · · (x − αn ).
Proof. If p(x) has degree n, then by Theorem 19.5 we can write any
element in E(α) as a linear combination of 1, α, . . . , αn−1 . Therefore, the
isomorphism that we are seeking must be
where
a0 + a1 α + · · · + an−1 αn−1
1. Given an arbitrary angle, can one trisect the angle into three equal
subangles using only a straightedge and compass?
2. Given an arbitrary circle, can one construct a square with the same
area using only a straightedge and compass?
3. Given a cube, can one construct the edge of another cube having
twice the volume of the original? Again, we are only allowed to use a
straightedge and compass to do the construction.
After puzzling mathematicians for over two thousand years, each of these
constructions was finally shown to be impossible. We will use the theory of
fields to provide a proof that the solutions do not exist. It is quite remarkable
that the long-sought solution to each of these three geometric problems came
from abstract algebra.
First, let us determine more specifically what we mean by a straightedge
and compass, and also examine the nature of these problems in a bit more
depth. To begin with, a straightedge is not a ruler. We cannot measure
arbitrary lengths with a straightedge. It is merely a tool for drawing a line
through two points. The statement that the trisection of an arbitrary angle is
impossible means that there is at least one angle that is impossible to trisect
with a straightedge-and-compass construction. Certainly it is possible to
trisect an angle in special cases. We can construct a 30◦ angle; hence, it is
possible to trisect a 90◦ angle. However, we will show that it is impossible
to construct a 20◦ angle. Therefore, we cannot trisect a 60◦ angle.
Constructible Numbers
A real number α is constructible if we can construct a line segment of
length |α| in a finite number of steps from a segment of unit length by using
a straightedge and compass.
Theorem 19.21 The set of all constructible real numbers forms a subfield
F of the field of real numbers.
β
B
1
α C
A E
x
√
Lemma 19.22 If α is a constructible number, then α is a constructible
number.
Proof. In Figure 19.2 the triangles 4ABD, 4BCD, and 4ABC are
similar; hence, 1/x = x/α, or x2 = α.
1 α
A D C
By Theorem 19.21, we can locate in the plane any point P = (p, q) that
has rational coordinates p and q. We need to know what other points can
be constructed with a compass and straightedge from points with rational
coordinates.
Proof. Let (x1 , y1 ) and (x2 , y2 ) be points on a line whose coordinates are
in F . If x1 = x2 , then the equation of the line through the two points is
x−x1 = 0, which has the form ax+by +c = 0. If x1 6= x2 , then the equation
of the line through the two points is given by
y2 − y1
y − y1 = (x − x1 ),
x2 − x1
(x − x1 )2 + (y − y1 )2 − r2 = 0.
2. The intersection of a line that passes through two points that have
coordinates in F and a circle whose center has coordinates in F with
radius of a length in F will give new points in R.
The first case gives no new points in R, since the solution of two equations
of the form ax + by + c = 0 having coefficients in F will always be in F . The
third case can be reduced to the second case. Let
x2 + y 2 + d1 x + e1 x + f1 = 0
x2 + y 2 + d2 x + e2 x + f2 = 0
340 CHAPTER 19 FIELDS
x2 + y 2 + d1 x + e1 x + f1 = 0
ax + by + c = 0
2 2
x + y + dx + ey + f = 0.
Q = F0 ⊂ F1 ⊂ · · · ⊂ Fk
√
such that Fi = Fi−1 ( αi ) with α ∈ Fk . In particular, there exists an integer
k > 0 such that [Q(α) : Q] = 2k .
19.3 GEOMETRIC CONSTRUCTIONS 341
Trisecting an Angle
Trisecting an arbitrary angle is impossible. We will show that it is impossible
to construct a 20◦ angle. Consequently, a 60◦ angle cannot be trisected. We
first need to calculate the triple-angle formula for the cosine:
cos 3θ = cos(2θ + θ)
= cos 2θ cos θ − sin 2θ sin θ
= (2 cos2 θ − 1) cos θ − 2 sin2 θ cos θ
= (2 cos2 θ − 1) cos θ − 2(1 − cos2 θ) cos θ
= 4 cos3 θ − 3 cos θ.
cosine,
1
4α3 − 3α = .
2
Therefore, α is a zero of 8x3 −6x−1. This polynomial has no factors in Z[x],
and hence is irreducible over Q[x]. Thus, [Q(α) : Q] = 3. Consequently, α
cannot be a constructible number.
Historical Note
Algebraic number theory uses the tools of algebra to solve problems in number
theory. Modern algebraic number theory began with Pierre de Fermat (1601–1665).
Certainly we can find many positive integers that satisfy the equation x2 + y 2 = z 2 ;
Fermat conjectured that the equation xn +y n = z n has no positive integer solutions
for n ≥ 3. He stated in the margin of his copy of the Latin translation of Diophantus’
Arithmetica that he had found a marvelous proof of this theorem, but that the
margin of the book was too narrow to contain it. To date, no one has been able to
construct a proof, although the statement has been verified for all n less than or
equal to 4 million. This conjecture is known as Fermat’s Last Theorem.
Attempts to prove Fermat’s Last Theorem have led to important contribu-
tions to algebraic number theory by such notable mathematicians as Leonhard
Euler (1707–1783). Significant advances in the understanding of Fermat’s Last
Theorem were made by Ernst Kummer (1810–1893). Kummer’s student, Leopold
Kronecker (1823–1891), became one of the leading algebraists of the nineteenth
century. Kronecker’s theory of ideals and his study of algebraic number theory
added much to the understanding of fields.
David Hilbert (1862–1943) and Hermann Minkowski (1864–1909) were among
the mathematicians who led the way in this subject at the beginning of the twentieth
century. Hilbert and Minkowski were both mathematicians at Göttingen University
in Germany. Göttingen was truly one the most important centers of mathematical
research during the last two centuries. The large number of exceptional mathemati-
cians who studied there included Gauss, Dirichlet, Riemann, Dedekind, Noether,
and Weyl.
André Weil answered questions in number theory using algebraic geometry, a
field of mathematics that studies geometry by studying commutative rings. From
about 1955 to 1970, A. Grothendieck dominated the field of algebraic geometry.
Pierre Deligne, a student of Grothendieck, solved several of Weil’s number-theoretic
conjectures. One of the most recent contributions to algebra and number theory is
Gerd Falting’s proof of the Mordell-Weil conjecture. This conjecture of Mordell and
Weil essentially says that certain polynomials p(x, y) in Z[x, y] have only a finite
number of integral solutions.
Exercises
EXERCISES 343
1. Show that each of the following numbers is algebraic over Q by finding the
minimal polynomial of the number over Q.
q √
(a) 1/3 + 7
√ √
(b) 3 + 3 5
√ √
(c) 3 + 2 i
(d) cos θ + i sin θ for θ = 2π/n with n ∈ N
p√
3
(e) 2−i
2. Find a basis for each of the following field extensions. What is the degree of
each extension?
√ √
(a) Q( 3, 6 ) over Q
√ √
(b) Q( 3 2, 3 3 ) over Q
√
(c) Q( 2, i) over Q
√ √ √
(d) Q( 3, 5, 7 ) over Q
√ √
(e) Q( 2, 3 2 ) over Q
√ √
(f ) Q( 8 ) over Q( 2 )
√ √
(g) Q(i, 2 + i, 3 + i) over Q
√ √ √
(h) Q( 2 + 5 ) over Q( 5 )
√ √ √ √ √
(i) Q( 2, 6 + 10 ) over Q( 3 + 5 )
11. Let p(x) be a nonconstant polynomial of degree n in F [x]. Prove that there
exists a splitting field E for p(x) such that [E : F ] ≤ n!.
√ √
12. Prove or disprove: Q( 2 ) ∼ = Q( 3 ).
√ √
13. Prove that the fields Q( 4 3 ) and Q( 4 3 i) are isomorphic but not equal.
17. Let E be the algebraic closure of a field F . Prove that every polynomial p(x)
in F [x] splits in E.
19. Prove that if α and β are constructible numbers such that β 6= 0, then so is
α/β.
20. Show that the set of all elements in R that are algebraic over Q form a field
extension of Q that is not finite.
346
20.1 STRUCTURE OF A FINITE FIELD 347
α = a1 α1 + · · · + an αn ,
Fortunately, we have an easy test to determine the separability of any
polynomial. Let
f (x) = a0 + a1 x + · · · + an xn
be any polynomial in F [x]. Define the derivative of f (x) to be
Lemma 20.4 Let F be a field and f (x) ∈ F [x]. Then f (x) is separable if
and only if f (x) and f 0 (x) are relatively prime.
Proof. Let f (x) be separable. Then f (x) factors over some extension field
of F as f (x) = (x − α1 )(x − α2 ) · · · (x − αn ), where αi 6= αj for i 6= j. Taking
the derivative of f (x), we see that
f 0 (x) = (x − α2 ) · · · (x − αn )
+ (x − α1 )(x − α3 ) · · · (x − αn )
+ · · · + (x − α1 ) · · · (x − αn−1 ).
Theorem 20.5 For every prime p and every positive integer n, there exists
a finite field F with pn elements. Furthermore, any field of order pn is
n
isomorphic to the splitting field of xp − x over Zp .
20.1 STRUCTURE OF A FINITE FIELD 349
n
Proof. Let f (x) = xp − x and let F be the splitting field of f (x). Then by
n
Lemma 20.4, f (x) has pn distinct zeros in F , since f 0 (x) = pn xp −1 −1 = −1
is relatively prime to f (x). We claim that the roots of f (x) form a subfield
of F . Certainly 0 and 1 are zeros of f (x). If α and β are zeros of f (x),
n n n
then α + β and αβ are also zeros of f (x), since αp + β p = (α + β)p
n n n
and αp β p = (αβ)p . We also need to show that the additive inverse and
the multiplicative inverse of each root of f (x) are roots of f (x). For any
zero α of f (x), −α = (p − 1)α is also a zero of f (x). If α 6= 0, then
n n
(α−1 )p = (αp )−1 = α−1 . Since the zeros of f (x) form a subfield of F and
f (x) splits in this subfield, the subfield must be all of F .
Let E be any other field of order pn . To show that E is isomorphic
to F , we must show that every element in E is a root of f (x). Certainly
0 is a root of f (x). Let α be a nonzero element of E. The order of the
n
multiplicative group of nonzero elements of E is pn − 1; hence, αp −1 = 1
n
or αp − α = 0. Since E contains pn elements, E must be a splitting field
of f (x); however, by Corollary 19.20, the splitting field of any polynomial is
unique up to isomorphism.
The unique finite field with pn elements is called the Galois field of
order pn . We will denote this field by GF(pn ).
Theorem 20.6 Every subfield of the Galois field GF(pn ) has pm elements,
where m divides n. Conversely, if m | n for m > 0, then there exists a
unique subfield of GF(pn ) isomorphic to GF(pm ).
GF(p24 )
" b
" b
GF(p8 ) GF(p12 )
"
"
"
"
GF(p4 ) GF(p6 )
"
"
"
"
GF(p2 ) GF(p3 )
b "
b "
GF(p)
G∼
= Zpe1 × · · · × Zpek .
1 k
α1 = α α6 = α2 + α3 α11 = α + α2 + α3
α2 = α2 α7 = 1 + α + α3 α12 = 1 + α + α2 + α3
α3 = α3 α8 = 1 + α2 α13 = 1 + α2 + α3
α4 = 1+α α9 = α + α3 α14 = 1 + α3
α5 = α + α2 α10 = 1 + α + α2 α15 = 1.
Example 4. Consider the (6, 3)-linear codes generated by the two matrices
1 0 0 1 0 0
0 1 0
1 1 0
0 0 1 1 1 1
G1 = and G2 = .
1 0 0
1 1 1
0 1 0 0 1 1
0 0 1 0 0 1
352 CHAPTER 20 FINITE FIELDS
This code cannot be cyclic, since (101101) is a codeword but (011011) is not
a codeword.
Polynomial Codes
We would like to find an easy method of obtaining cyclic linear codes. To
accomplish this, we can use our knowledge of finite fields and polynomial
rings over Z2 . Any binary n-tuple can be interpreted as a polynomial in
Z2 [x]. Stated another way, the n-tuple (a0 , a1 , . . . , an−1 ) corresponds to the
polynomial
f (x) = a0 + a1 x + · · · + an−1 xn−1 ,
where the degree of f (x) is at most n − 1. For example, the polynomial
corresponding to the 5-tuple (10011) is
Conversely, with any polynomial f (x) ∈ Z2 [x] with deg f (x) < n we can
associate a binary n-tuple. The polynomial x + x2 + x4 corresponds to the
5-tuple (01101).
Let us fix a nonconstant polynomial g(x) in Z2 [x] of degree n − k. We can
define an (n, k)-code C in the following manner. If (a0 , . . . , ak−1 ) is a k-tuple
to be encoded, then f (x) = a0 + a1 x + · · · + ak−1 xk−1 is the corresponding
polynomial in Z2 [x]. To encode f (x), we multiply by g(x). The codewords
in C are all those polynomials in Z2 [x] of degree less than n that are divisible
by g(x). Codes obtained in this manner are called polynomial codes.
Example 5. If we let g(x) = 1+x3 , we can define a (6, 3)-code C as follows.
To encode a 3-tuple (a0 , a1 , a2 ), we multiply the corresponding polynomial
20.2 POLYNOMIAL CODES 353
(1 + x3 ) · 1 = 1 + x3
(1 + x3 )x = x + x4
(1 + x3 )x3 = x2 + x5 .
Since the smallest weight of any nonzero codeword is 2, this code has the
ability to detect all single errors.
Rings of polynomials have a great deal of structure; therefore, our imme-
diate goal is to establish a link between polynomial codes and ring theory.
Recall that xn − 1 = (x − 1)(xn−1 + · · · + x + 1). The factor ring
Rn = Z2 [x]/hxn − 1i
that satisfy the condition tn = 1. It is an easy exercise to show that Zn2 and
Rn are isomorphic as vector spaces. We will often identify elements in Zn2
with elements in Z[x]/hxn − 1i. In this manner we can interpret a linear
code as a subset of Z[x]/hxn − 1i.
354 CHAPTER 20 FINITE FIELDS
x7 − 1 = (1 + x)(1 + x + x3 )(1 + x2 + x3 ).
20.2 POLYNOMIAL CODES 355
is a generator matrix for the code C with generator polynomial g(t). The
parity-check matrix for C is the (n − k) × n matrix
0 ··· 0 0 hk ··· h0
0 ··· 0 hk ··· h0 0
H=
···
.
··· ··· ··· ··· ··· ···
hk ··· h0 0 0 ··· 0
Example 7. In Example 6,
x7 − 1 = g(x)h(x) = (1 + x + x3 )(1 + x + x2 + x4 ).
where
1 1 ··· 1
α1 α2 ··· αn−1
β = (−1)n+n det
α12 α22 ··· 2
αn−1 .
.. .. .. ..
. . . .
α1n−2 α2n−2 · · · n−2
αn−1
By our induction hypothesis,
Y
β = (−1)n+n (αi − αj ).
1≤j<i≤n−1
f (ω r ) = f (ω r+1 ) = · · · = f (ω r+s−1 ) = 0.
(ω i0 )r x0 + (ω i1 )r x1 + · · · + (ω is−1 )r xn−1 = 0
(ω i0 )r+1 x0 + (ω i1 )r+1 x1 + · · · + (ω is−1 )r+1 xn−1 = 0
..
.
(ω i0 )r+s−1 x0 + (ω i1 )r+s−1 x1 + · · · + (ω is−1 )r+s−1 xn−1 = 0.
However, this system has a unique solution, since the determinant of the
matrix
(ω i0 )r (ω i1 )r (ω is−1 )r
···
(ω i0 )r+1 (ω i1 )r+1 · · · (ω is−1 )r+1
.. .. . . ..
. . . .
i
(ω )0 r+s−1 i
(ω )1 r+s−1 · · · (ω i s−1 ) r+s−1
can be shown to be nonzero using Lemma 20.12 and the basic properties of
determinants (Exercise). Therefore, this solution must be ai0 = ai1 = · · · =
ais−1 = 0.
BCH Codes
Some of the most important codes, discovered independently by A. Hoc-
quenghem in 1959 and by R. C. Bose and D. V. Ray-Chaudhuri in 1960, are
BCH codes. The European and transatlantic communication systems both
use BCH codes. Information words to be encoded are of length 231, and
a polynomial of degree 24 is used to generate the code. Since 231 + 24 =
255 = 28 − 1, we are dealing with a (255, 231)-block code. This BCH code
will detect six errors and has a failure rate of 1 in 16 million. One advantage
of BCH codes is that efficient error correction algorithms exist for them.
20.2 POLYNOMIAL CODES 359
then the cyclic code hg(t)i in Rn is called the BCH code of length n and
distance d. By Theorem 20.13, the minimum distance of C is at least d.
3. The matrix
ω2 ··· ω n−1
1 ω
1 ω2 ω4 ··· ω (n−1)(2)
H=
1 ω3 ω6 ··· ω (n−1)(3)
.. .. .. .. ..
. . . . .
1 ω 2r ω 4r · · · ω (n−1)(2r)
Proof. (1) ⇒ (2). If f (t) is in C, then g(x) | f (x) in Z2 [x]. Hence, for
i = 1, . . . , 2r, f (ω i ) = 0 since g(ω i ) = 0. Conversely, suppose that f (ω i ) = 0
for 1 ≤ i ≤ d. Then f (x) is divisible by each mi (x), since mi (x) is the
minimal polynomial of ω i . Therefore, g(x) | f (x) by the definition of g(x).
Consequently, f (x) is a codeword.
(2) ⇒ (3). Let f (t) = a0 + a1 t + · · · + an−1 vtn−1 be in Rn . The corre-
sponding n-tuple in Zn2 is x = (a0 a1 · · · an−1 )t . By (2),
a0 + a1 ω + · · · + an−1 ω n−1
f (ω)
a0 + a1 ω 2 + · · · + an−1 (ω 2 )n−1 f (ω 2 )
Hx = = =0
.. ..
. .
a0 + a1 ω 2r + · · · + an−1 (ω 2r )n−1 f (ω 2r )
Exercises
1. Calculate each of the following.
(a) [GF(36 ) : GF(33 )] (b) [GF(128) : GF(16)]
(c) [GF(625) : GF(25)] (d) [GF(p12 ) : GF(p2 )]
2. Calculate [GF(pm ) : GF(pn )], where n | m.
3. What is the lattice of subfields for GF(p30 )?
EXERCISES 361
22. Show that every element in GF(pn ) can be written in the form ap for some
unique a ∈ GF(pn ).
24. Wilson’s Theorem. Let p be prime. Prove that (p − 1)! ≡ −1 (mod p).
25. If g(t) is the minimal generator polynomial for a cyclic code C in Rn , prove
that the constant term of g(x) is 1.
26. Often it is conceivable that a burst of errors might occur during transmission,
as in the case of a power surge. Such a momentary burst of interference
might alter several consecutive bits in a codeword. Cyclic codes permit the
detection of such error bursts. Let C be an (n, k)-cyclic code. Prove that
any error burst up to n − k digits can be detected.
27. Prove that the rings Rn and Zn2 are isomorphic as vector spaces.
s(x) = (x + ω a1 )(x + ω a2 ) · · · (x + ω ak ).
3. Recall the (15, 7)-block BCH code in Example 7. By Theorem 7.3, this code
is capable of correcting two errors. Suppose that these errors occur in bits
a1 and a2 . The error-locator polynomial is s(x) = (x + ω a1 )(x + ω a2 ). Show
that
2 2 s3
s(x) = x + s1 x + s1 + .
s1
Finally, at the beginning of the nineteenth century, Ruffini and Abel both
found quintics that could not be solved with any formula. It was Galois,
however, who provided the full explanation by showing which polynomials
could and could not be solved by formulas. He discovered the connection
between groups and field extensions. Galois theory demonstrates the strong
interdependence of group and field theory, and has had far-reaching impli-
cations beyond its original purpose.
In this chapter we will prove the Fundamental Theorem of Galois Theory.
This result will be used to establish the insolvability of the quintic and to
prove the Fundamental Theorem of Algebra.
364
21.1 FIELD AUTOMORPHISMS 365
Proof. We need only show that the set of automorphisms of E that fix F
elementwise is a subgroup of the group of all automorphisms of E. Let σ
and τ be two automorphisms of E such that σ(α) = α and τ (α) = α for all
α ∈ F . Then στ (α) = σ(α) = α and σ −1 (α) = α. Since the identity fixes
every element of E, the set of automorphisms of E that leave elements of F
fixed is a subgroup of the entire group of automorphisms of E.
Proof. Let
f (x) = a0 + a1 x + a2 x2 + · · · + an xn
and suppose that α ∈ E is a zero of f (x). Then for σ ∈ G(E/F ),
0 = σ(0)
= σ(f (α))
= σ(a0 + a1 α + a2 α2 + · · · + an αn )
= a0 + a1 σ(α) + a2 [σ(α)]2 + · · · + an [σ(α)]n ;
Theorem 21.5 Let f (x) be a polynomial in F [x] and suppose that E is the
splitting field for f (x) over F . If f (x) has no repeated roots, then
|G(E/F )| = [E : F ].
21.1 FIELD AUTOMORPHISMS 367
Proof. The proof is similar to the proof of Theorem 19.19. We will use
mathematical induction on the degree of f (x). If the degree of f (x) is 0 or
1, then E = F and there is nothing to show. Assume that the result holds
for all polynomials of degree k with 0 ≤ k < n. Let p(x) be an irreducible
factor of f (x) of degree r. Since all of the roots of p(x) are in E, we can
choose one of these roots, say α, so that F ⊂ F (α) ⊂ E. If β is any other
root of p(x), then F ⊂ F (β) ⊂ E. By Lemma 19.18, there exists a unique
isomorphism σ : F (α) → F (β) for each such β that fixes F elementwise.
Since E is a splitting field of F (β), there are exactly r such isomorphisms.
We can factor p(x) in F (α) as p(x) = (x − α)p1 (x). The degrees of p1 (x)
and q1 (x) are both less than r. Since we know that E is the splitting field
of p1 (x) over F (α), we can apply the induction hypothesis to conclude that
[E : F ] = [E : F (α)][F (α) : F ]
Corollary 21.6 Let F be a finite field with a finite extension E such that
[E : F ] = k. T hen G(E/F ) is cyclic.
Proof. Let p be the characteristic of E and F and assume that the orders
of E and F are pm and pn , respectively. Then nk = m. We can also assume
m
that E is the splitting field of xp − x over a subfield of order p. Therefore,
m
E must also be the splitting field of xp −x over F . Applying Theorem 21.5,
we find that |G(E/F )| = k.
To prove that G(E/F ) is cyclic, we must find a generator for G(E/F ).
n
Let σ : E → E be defined by σ(α) = αp . We claim that σ is the element
in G(E/F ) that we are seeking. We first need to show that σ is in Aut(E).
If α and β are in E,
n n n
σ(α + β) = (α + β)p = αp + β p = σ(α) + σ(β)
Example 4. Let us compute the Galois group of
f (x) = x4 + x3 + x2 + x + 1
ω = cos(2π/5) + i sin(2π/5).
Hence, the splitting field of f (x) must be Q(ω). We can define automor-
phisms σi of Q(ω) by σi (ω) = ω i for i = 1, . . . , 4. It is easy to check that
these are indeed distinct automorphisms in G(Q(ω)/Q). Since
[Q(ω) : Q] = |G(Q(ω)/Q)| = 4,
Separable Extensions
Many of the results that we have just proven depend on the fact that a
polynomial f (x) in F [x] has no repeated roots in its splitting field. It is
evident that we need to know exactly when a polynomial factors into distinct
linear factors in its splitting field. Let E be the splitting field of a polynomial
f (x) in F [x]. Suppose that f (x) factors over E as
r
Y
n1 n2 nr
f (x) = (x − α1 ) (x − α2 ) · · · (x − αr ) = (x − αi )ni .
i=1
Proof. First assume that charF = 0. Since deg f 0 (x) < deg f (x) and
f (x) is irreducible, the only way gcd(f (x), f 0 (x)) 6= 1 is if f 0 (x) is the zero
polynomial; however, this is impossible in a field of characteristic zero. If
charF = p, then f 0 (x) can be the zero polynomial if every coefficient of f (x)
is a multiple of p. This can happen only if we have a polynomial of the form
f (x) = a0 + a1 xp + a2 x2p + · · · + an xnp .
Certainly extensions of a field F of the form F (α) are some of the easiest
to study and understand. Given a field extension E of F , the obvious
question to ask is when it is possible to find an element α ∈ E such that
E = F (α). In this case, α is called a primitive element. We already know
that primitive elements exist for certain extensions. For example,
√ √ √ √
Q( 3, 5 ) = Q( 3 + 5 )
and √ √ √
3 6
Q( 5, 5 i) = Q( 5 i).
Corollary 20.9 tells us that there exists a primitive element for any finite
extension of a finite field. The next theorem tells us that we can often find
a primitive element.
σi (a ± b) = σi (a) ± σi (b) = a ± b
and
σi (ab) = σi (a)σi (b) = ab.
If a 6= 0, then σi (a−1 ) = [σi (a)]−1 = a−1 . Finally, σi (0) = 0 and σi (1) = 1
since σi is an automorphism.
is a subfield of F .
21.2 THE FUNDAMENTAL THEOREM 371
The subfield F{σi } of F is called the fixed field of {σi }. The field fixed
for a subgroup G of Aut(F ) will be denoted by FG .
√ √ √ √
Example√ 5. Let√ σ : Q( 3, √ 5 ) → Q( 3, 5 ) be the√automorphism
√ that
maps 3 to − 3. Then Q( 5 ) is the subfield of Q( 3, 5 ) left fixed by
σ.
|G| = [E : EG ] = [E : F ].
a1 α1 + a2 α2 + · · · + an+1 αn+1 = 0.
has more equations than unknowns. From linear algebra we know that this
system has a nontrivial solution, say xi = ai for i = 1, 2, . . . , n + 1. Since σ1
is the identity, the first equation translates to
a1 α1 + a2 α2 + · · · + an+1 αn+1 = 0.
372 CHAPTER 21 GALOIS THEORY
x1 = 1 − 1 = 0
x2 = a2 − σi (a2 )
..
.
xn+1 = an+1 − σi (an+1 )
Let f (x) be the minimal polynomial of α over F . The field E must contain
all of the roots of f (x) since it is a normal extension F ; hence, E is a splitting
field for f (x).
(2) ⇒ (3). Let E be the splitting field over F of a separable polynomial.
By Proposition 21.11, EG(E/F ) = F . Since |G(E/F )| = [E : F ], this is a
finite group.
(3) ⇒ (1). Let F = EG for some finite group of automorphisms G of E.
Since [E : F ] ≤ |G|, E is a finite extension of F . To show that E is a finite,
normal extension of F , let f (x) ∈ F [x] be an irreducible monic polynomial
that has a root α in E. We must show that f (x) is the product of distinct
linear factors in E[x]. By Proposition 21.3, automorphisms in G permute
the roots of f (x) lying in E. Hence, if we let G Q act on α, we can obtain
distinct roots α1 = α, α2 , . . . , αn in E. Let g(x) = ni=1 (x − αi ). Then g(x)
is separable over F and g(α) = 0. Any automorphism σ in G permutes the
factors of g(x) since it permutes these roots; hence, when σ acts on g(x), it
must fix the coefficients of g(x). Therefore, the coefficients of g(x) must be
in F . Since deg g(x) ≤ deg f (x) and f (x) is the minimal polynomial of α,
f (x) = g(x).
[K : F ] ≤ |G| ≤ |G(K/F )| = [K : F ].
It follows that G = G(K/F ), since they must have the same order.
Before we determine the exact correspondence between field extensions
and automorphisms of fields, let us return to a familiar example.
√ √
Example 6. In Example 2 we examined the automorphisms of Q( 3, 5 )
fixing Q. Figure 21.1 compares
√ the√ lattice of field extensions of Q with the
lattice of subgroups of G(Q( 3, 5 )/Q). The Fundamental Theorem of
Galois Theory tells us what the relationship is between the two lattices.
We are now ready to state and prove the Fundamental Theorem of Galois
Theory.
√ √
{id, σ, τ, µ} Q( 3, 5 )
@ @
@ @
@ @
√ √ √
{id, σ} {id, τ } {id, µ} Q( 3 ) Q( 5 ) Q( 15 )
@ @
@ @
@ @
{id} Q
√ √
Figure 21.1. G(Q( 3, 5 )/Q)
2. If F ⊂ K ⊂ E, then
G(K/F ) ∼
= G(E/F )/G(E/K).
E - {id}
L - G(E/L)
K - G(E/K)
F - G(E/F )
hence, σ(α) must be in the fixed field of G(E/K). Let σ be the restriction
of σ to K. Then σ is an automorphism of K fixing F , since σ(α) ∈ K for
all α ∈ K; hence, σ ∈ G(K/F ). Next, we will show that the fixed field of
G(K/F ) is F . Let β be an element in K that is fixed by all automorphisms
in G(K/F ). In particular, σ(β) = β for all σ ∈ G(E/F ). Therefore, β
belongs to the fixed field F of G(E/F ).
Finally, we must show that when K is a normal extension of F ,
G(K/F ) ∼
= G(E/F )/G(E/K).
Hence, the image of φ is G(K/F ) and φ is onto. Applying the First Isomor-
phism Theorem, we have
G(K/F ) ∼
= G(E/F )/G(E/K).
Example 7. In this example we will illustrate the Fundamental Theorem of
Galois Theory by determining the lattice of subgroups of the Galois group of
f (x) = x4 − 2. We will compare this lattice to the lattice of field extensions
of Q that √ are contained in the splitting field of x4 − 2. The splitting √ field
√ of
f (x) is Q( 4 2, i). To see this, notice that f (x) factors as (x 2 + 2 )(x2 − 2 );
√ √ √
hence, the roots of f (x) are ± 4 2 and ± 4 2 i. We √ first adjoin the root 4 2
to Q and then adjoin√ the root
√ i of x2 + 1 to Q( 4 2 ). The splitting field of
4
f (x) is then Q(
√ 2 )(i) = Q( 4 2, i). √
4 4
Since
√ [Q( √ 2 ) : Q] = 4 and i is not
√ in Q( 2 ), it must be the case that
4 4 4
[Q( 2, i) : Q( 2 )] = 2. Hence, [Q( 2, i) : Q] = 8. The set
√ √ √ √ √ √
{1, 2, ( 2 )2 , ( 2 )3 , i, i 2, i( 2 )2 , i( 2 )3 }
4 4 4 4 4 4
√
4
is a basis
√ of Q( 2, i) over Q. The lattice of field extensions of Q contained
4
in Q( 2, i) is illustrated in Figure 21.3(a).
The Galois group√G of f (x) √ must be of order 8. Let σ be the automor-
4 4
phism defined by σ( 2 ) = 2 and σ(i) = i, and τ be the automorphism
defined by complex conjugation; that is, τ (i) = −i. Then G has an ele-
ment of order 4 and an element of order 2. It is easy to verify by direct
computation that the elements of G are {id, σ, σ 2 , σ 3 , τ, στ, σ 2 τ, σ 3 τ } and
that the relations τ 2 = id, σ 4 = id, and τ στ = σ −1 are satisfied; hence, G
must be isomorphic to D4 . The lattice of subgroups of G is illustrated in
Figure 21.3(b).
Historical Note
Solutions for the cubic and quartic equations were discovered in the 1500s. At-
tempts to find solutions for the quintic equations puzzled some of history’s best
mathematicians. In 1798, P. Ruffini submitted a paper that claimed no such so-
lution could be found; however, the paper was not well received. In 1826, Niels
Henrik Abel (1802–1829) finally offered the first correct proof that quintics are not
always solvable by radicals.
21.2 THE FUNDAMENTAL THEOREM 377
√
Q( 4 2 )
HH PPP
H
H PP
√
4
√
4
√ √ P √
Q( 2 ) Q( 2 i) Q( 2, i) Q((1 + i) 2 ) Q((1 − i) 4 2 )
4
HHH H
HH
H √ H √
Q( 2 ) Q(i) Q( 2 i)
H HH
H
Q (a)
D4
H
HH
H
{id, σ 2 , τ, σ 2 τ } {id, σ, σ 2 , σ 3 } {id, σ 2 , στ, σ 3 τ }
H HH
HH
H HH
{id, τ } {id, σ 2 τ } {id, σ 2 } {id, στ } {id, σ 3 τ }
PP H
PP H
PP HH
{id} (b)
Abel inspired the work of Évariste Galois. Born in 1811, Galois began to display
extraordinary mathematical talent at the age of 14. He applied for entrance to the
École Polytechnique several times; however, he had great difficulty meeting the for-
mal entrance requirements, and the examiners failed to recognize his mathematical
genius. He was finally accepted at the École Normale in 1829.
Galois worked to develop a theory of solvability for polynomials. In 1829, at
the age of 17, Galois presented two papers on the solution of algebraic equations
to the Académie des Sciences de Paris. These papers were sent to Cauchy, who
subsequently lost them. A third paper was submitted to Fourier, who died before
he could read the paper. Another paper was presented, but was not published
until 1846.
Galois’s democratic sympathies led him into the Revolution of 1830. He was
expelled from school and sent to prison for his part in the turmoil. After his release
in 1832, he was drawn into a duel over a love affair. Certain that he would be
killed, he spent the evening before his death outlining his work and his basic ideas
378 CHAPTER 21 GALOIS THEORY
for research in a long letter to his friend Chevalier. He was indeed dead the next
day, at the age of 21.
21.3 Applications
Solvability by Radicals
Throughout this section we shall assume that all fields have characteristic
zero to ensure that irreducible polynomials do not have multiple roots. The
immediate goal of this section is to determine when the roots of a polynomial
f (x) can be computed in a finite number of operations on the coefficients
of f (x). The allowable operations are addition, subtraction, multiplication,
division, and the extraction of nth roots. Certainly the solution to the
quadratic equation, ax2 + bx + c = 0, illustrates this process:
√
−b ± b2 − 4ac
x= .
2a
The only one of these operations that might demand a larger field is the
taking of nth roots. We are led to the following definition.
An extension field E of a field F is an extension by radicals if there
are elements α1 , . . . , αr ∈ K and positive integers n1 , . . . , nr such that
E = F (α1 , . . . , αr ),
Proof. First suppose that F contains all of its nth roots of unity. The roots
√ √ √
of xn −a are n a, ω n a, . . . , ω n−1 n a, where ω is a primitive nth root of unity.
If ζ is one of these roots, then distinct roots of xn − 1 are ζ, ωζ, . . . , ω n−1 ζ,
and E = F (ζ). Since G(E/F ) permutes the roots xn − 1, the elements in
G(E/F ) must be determined by their action on these roots. Let σ and τ be
in G(E/F ) and suppose that σ(ζ) = ω i ζ and τ (ζ) = ω j ζ. If F contains the
roots of unity, then
150
100
50
-3 -2 -1 1 2 3
-50
-100
-150
and once again between 0 and 4 (Figure 21.4). Therefore, f (x) has exactly
three distinct real roots. The remaining two roots of f (x) must be complex
conjugates. Let K be the splitting field of f (x). Since f (x) has five distinct
roots in K and every automorphism of K fixing Q is determined by the
way it permutes the roots of f (x), we know that G(K/Q) is a subgroup of
S5 . Since f is irreducible, there is an element in σ ∈ G(K/Q) such that
σ(a) = b for two roots a and b of f (x). The automorphism of C that takes
a + bi 7→ a − bi leaves the real roots fixed and interchanges the complex
roots; consequently, G(K/Q) ⊂ S5 . By Lemma 21.19, S5 is generated by
a transposition and an element of order 5; therefore, G(K/F ) must be all
of S5 . By Theorem 9.8, S5 is not solvable. Consequently, f (x) cannot be
solved by radicals.
For our proof we shall assume two facts from calculus. We need the
results that every polynomial of odd degree over R has a real root and that
every positive real number has a square root.
Proof. Suppose that E is a proper finite field extension of the complex
numbers. Since any finite extension of a field of characteristic zero is a
simple extension, there exists an α ∈ E such that E = C(α) with α the root
of an irreducible polynomial f (x) in C[x]. The splitting field L of f (x) is a
finite normal separable extension of C that contains E. We must show that
it is impossible for L to be a proper extension of C.
Suppose that L is a proper extension of C. Since L is the splitting field
of f (x)(x2 + 1) over R, L is a finite normal separable extension of R. Let
K be the fixed field of a Sylow 2-subgroup G of G(L/R). Then L ⊃ K ⊃ R
and |G(L/K)| = [L : K]. Since [L : R] = [L : K][K : R], we know that
EXERCISES 383
Exercises
1. Compute each of the following Galois groups. Which of these field extensions
are normal field extensions? If the extension is not normal, find a normal
extension of Q in which the extension field is contained.
√ √
(a) G(Q( 30 )/Q) (b) G(Q( 4 5 )/Q)
√ √ √ √ √
(c) G(Q( 2, 3, 5 )/Q) (d) G(Q( 2, 3 2, i)/Q)
√
(e) G(Q( 6, i)/Q)
3. Give the order and describe a generator of the Galois group of GF(729)
over GF(9).
18. Prove or disprove: Two different subgroups of a Galois group will have dif-
ferent fixed fields.
19. Let K be the splitting field of a polynomial over F . If E is a field extension
of F contained in K and [E : F ] = 2, then E is the splitting field of some
polynomial in F [x].
20. We know that the cyclotomic polynomial
xp − 1
Φp (x) = = xp−1 + xp−2 + · · · + x + 1
x−1
is irreducible over Q for every prime p. Let ω be a zero of Φp (x), and consider
the field Q(ω).
(a) Show that ω, ω 2 , . . . , ω p−1 are distinct zeros of Φp (x), and conclude that
they are all the zeros of Φp (x).
(b) Show that G(Q(ω)/Q) is abelian of order p − 1.
(c) Show that the fixed field of G(Q(ω)/Q) is Q.
21. Let F be a finite field or a field of characteristic zero. Let E be a finite normal
extension of F with Galois group G(E/F ). Prove that F ⊂ K ⊂ L ⊂ E if
and only if {id} ⊂ G(E/L) ⊂ G(E/K) ⊂ G(E/F ).
22. Let F be a field of characteristic zero and let f (x) ∈ F [x] be a separable
polynomial of degree n. If E isQthe splitting field of f (x), let α1 , . . . , αn be the
roots of f (x) in E. Let ∆ = i6=j (αi − αj ). We define the discriminant
of f (x) to be ∆2 .
(a) If f (x) = ax2 + bx + c, show that ∆2 = b2 − 4ac.
(b) If f (x) = x3 + px + q, show that ∆2 = −4p3 − 27q 2 .
(c) Prove that ∆2 is in F .
(d) If σ ∈ G(E/F ) is a transposition of two roots of f (x), show that
σ(∆) = − ∆.
(e) If σ ∈ G(E/F ) is an even permutation of the roots of f (x), show that
σ(∆) = ∆.
(f ) Prove that G(E/F ) is isomorphic to a subgroup of An if and only if
∆ ∈ F.
(g) Determine the Galois groups of x3 + 2x − 4 and x3 + x − 3.
The following table defines the notation used in this book. Page numbers
refer to the first appearance of each symbol.
387
388 NOTATION
Chapter 0. Preliminaries
1. (a) {2}. (b) {5}.
2. (a) {(a, 1), (a, 2), (a, 3), (b, 1), (b, 2), (b, 3), (c, 1), (c, 2), (c, 3)}.
(d) ∅.
6. If x ∈ A ∪ (B ∩ C), then either x ∈ A or x ∈ B ∩ C ⇒ x ∈ A ∪ B and A ∪ C ⇒
x ∈ (A ∪ B) ∩ (A ∪ C) ⇒ A ∪ (B ∩ C) ⊂ (A ∪ B) ∩ (A ∪ C).
Conversely, x ∈ (A ∪ B) ∩ (A ∪ C) ⇒ x ∈ A ∪ B and A ∪ C ⇒ x ∈
A or x is in both B and C ⇒ x ∈ A ∪ (B ∩ C) ⇒ (A ∪ B) ∩ (A ∪ C) ⊂
A ∪ (B ∩ C). Hence, A ∪ (B ∩ C) = (A ∪ B) ∩ (A ∪ C).
10. (A ∩ B) ∪ (A \ B) ∪ (B \ A) = (A ∩ B) ∪ (A ∩ B 0 ) ∪ (B ∩ A0 ) = [A ∩ (B ∪
B 0 )] ∪ (B ∩ A0 ) = A ∪ (B ∩ A0 ) = (A ∪ B) ∩ (A ∪ A0 ) = A ∪ B.
14. A \ (B ∪ C) = A ∩ (B ∪ C)0 = (A ∩ A) ∩ (B 0 ∩ C 0 ) = (A ∩ B 0 ) ∩ (A ∩ C 0 ) =
(A \ B) ∩ (A \ C).
17. (a) Not a map. f (2/3) is undefined.
(c) Not a map. f (1/2) = 3/4 and f (2/4) = 3/8.
18. (a) One-to-one but not onto. f (R) = {x ∈ R : x > 0}.
(c) Neither one-to-one nor onto.
20. (a) f (n) = n + 1.
22. (a) Let x, y ∈ A. Then g(f (x)) = (g ◦ f )(x) = (g ◦ f )(y) = g(f (y)) ⇒ f (x) =
f (y) ⇒ x = y, so g ◦ f is one-to-one.
(b) Let c ∈ C, then c = (g ◦f )(x) = g(f (x)) for some x ∈ A. Since f (x) ∈ B,
g is onto.
23. f −1 (x) = (x + 1)/(x − 1).
24. (a) Let y ∈ f (A1 ∪ A2 ) ⇒ there exists an x ∈ A1 ∪ A2 such that f (x) = y ⇒
y ∈ f (A1 ) or f (A2 ) ⇒ y ∈ f (A1 ) ∪ f (A2 ) ⇒ f (A1 ∪ A2 ) ⊂ f (A1 ) ∪ f (A2 ).
391
392 HINTS AND SOLUTIONS
17. (b) Use mathematical induction. (c) Show that f1 = 1, f2 = 1, and fn+2 =
fn+1 + fn . (d) Use part (c). (e) Use part (b) and Problem 16.
Chapter 2. Groups
1. (a) {. . . , −4, 3, 10, . . .}. (c) {. . . , −8, 18, 44, . . .}. (e) {. . . , −1, 5, 11, . . .}.
2. (a) Not a group. (c) A group.
6. · 1 5 7 11
1 1 5 7 11
5 5 1 11 7
7 7 11 1 5
11 11 7 5 1
(c) 3Z = {. . . , −3, 0, 3, 6, . . .}
1 + 3Z = {. . . , −2, 1, 4, 7, . . .}
2 + 3Z = {. . . , −1, 2, 5, 8, . . .}.
1 0
Cosets
C (00000) (00101) (10011) (10110)
(10000) + C (10000) (10101) (00011) (00110)
(01000) + C (01000) (01101) (11011) (11110)
(00100) + C (00100) (00001) (10111) (10010)
(00010) + C (00010) (00111) (10001) (10100)
(11000) + C (11000) (11101) (01011) (01110)
(01100) + C (01100) (01001) (11111) (11010)
(01010) + C (01010) (01111) (11001) (11100)
HINTS AND SOLUTIONS 397
A decoding table does not exist for C since it is only single error-detecting.
19. Let x ∈ C have odd weight and define a map from the set of odd codewords
to the set of even codewords by y 7→ x+y. Show that this map is a bijection.
23. For 20 information positions, at least six check bits are needed to ensure an
error-correcting code.
Chapter 8. Isomorphisms
1. The group nZ is an infinite cyclic group generated by n. Every infinite cyclic
group is isomorphic to Z.
2. Define φ : C∗ → GL2 (R) by
a b
φ(a + bi) = .
−b a
3. False.
6. Define a map from Zn into the nth roots of unity by k 7→ cis(2kπ/n).
8. Assume that Q is cyclic and try to find a generator.
11. D4 , Q8 , Z8 , Z2 × Z4 , Z2 × Z2 × Z2 .
16. (a) 12. (c) 5.
20. True.
25. Z2 × Z2 × Z13 is not cyclic.
27. Let a be a generator for G. If φ : G → H is an isomorphism, show that φ(a)
is a generator for H.
38. Any automorphism of Z6 must send 1 to another generator of Z6 .
45. To show that φ is one-to-one, let g1 = h1 k1 and g2 = h2 k2 . Then φ(g1 ) =
φ(g2 ) ⇒ φ(h1 k1 ) = φ(h2 k2 ) ⇒ (h1 , k1 ) = (h2 , k2 ) ⇒ h1 = h2 , k1 = k2 ⇒
g1 = g2 .
1 1
kx + yk2 + kxk2 − kyk2 = hx + y, x + yi − kxk2 − kyk2
1.
2 2
1
kxk2 + 2hx, yi + kyk2 − kxk2 − kyk2
=
2
= hx, yi.
10. Show that the kernel of the map det : O(n) → R∗ is SO(n).
13. True.
17. p6m.
N = Nn ⊃ Nn−1 ⊃ · · · ⊃ N1 ⊃ N0 = {e}
G/N = Gn /N ⊃ Gn−1 /N ⊃ · · · G1 /N ⊃ G0 /N = {N }.
The series
is a subnormal series. The factors of this series are abelian since Gi+1 /Gi ∼
=
(Gi+1 /N )/(Gi /N ).
16. Use the fact that Dn has a cyclic subgroup of index 2.
21. G/G0 is abelian.
4. (a) {0}, {0, 9}, {0, 6, 12}, {0, 3, 6, 9, 12, 15}, {0, 2, 4, 6, 8, 10, 12, 14, 16}.
(c) There are no nontrivial ideals.
7. Assume there is an isomorphism φ : C → R with φ(i) = a.
√ √
8. False.
√ Assume there is an isomorphism φ : Q( 2 ) → Q( 3 ) such that
φ( 2 ) = a.
HINTS AND SOLUTIONS 401
30
@
@
10 15
2 5 3
@
@
1
5. False.
6. (a) (a ∨ b ∨ a0 ) ∧ a.
b a
a0
HINTS AND SOLUTIONS 403
(c) a ∨ (a ∧ b).
a b
8. Not equivalent.
10. a0 ∧ [(a ∧ b0 ) ∨ b] = a ∧ (a ∨ b).
15. Let I, J be ideals in R. We need to show that I +J = {r+s : r ∈ I and s ∈ J}
is the smallest ideal in R containing both I and J. If r1 , r2 ∈ I and s1 , s2 ∈ J,
then (r1 + s1 ) + (r2 + s2 ) = (r1 + r2 ) + (s1 + s2 ) is in I + J. For a ∈ R,
a(r1 + s1 ) = ar1 + as1 ∈ I + J; hence, I + J is an ideal in R.
19. (a) No.
21. (⇒). a = b ⇒ (a ∧ b0 ) ∨ (a0 ∧ b) = (a ∧ a0 ) ∨ (a0 ∧ a) = O ∨ O = O.
(⇐). (a∧b0 )∨(a0 ∧b) = O ⇒ a∨b = (a∨a)∨b = a∨(a∨b) = a∨[I ∧(a∨b)] =
a∨[(a∨a0 )∧(a∨b)] = [a∨(a∧b0 )]∨[a∨(a0 ∧b)] = a∨[(a∧b0 )∨(a0 ∧b)] = a∨0 = a.
A symmetric argument shows that a ∨ b = b.
T (u + v) = T (u) + T (v) = 0
T (αv) = αT (v) = α0 = 0.
(u + v) + (u0 + v 0 ) = (u + u0 ) + (v + v 0 ) ∈ U + V
α(u + v) = αu + αv ∈ U + V.
404 HINTS AND SOLUTIONS
Now multiply both sides by q(α)n to show that there is a polynomial in F [x]
that has α as a zero.
18. Since α is algebraic over F of degree n, we can write any element β ∈ F (α)
uniquely as β = a0 +a1 α+· · ·+an−1 αn−1 with ai ∈ F . There are q n possible
n-tuples (a0 , a1 , . . . , an−1 ).
24. Factor xp−1 − 1 over Zp .
Preamble
The purpose of this License is to make a manual, textbook, or other functional
and useful document “free” in the sense of freedom: to assure everyone the effective
freedom to copy and redistribute it, with or without modifying it, either commer-
cially or noncommercially. Secondarily, this License preserves for the author and
publisher a way to get credit for their work, while not being considered responsible
for modifications made by others.
This License is a kind of “copyleft”, which means that derivative works of the
document must themselves be free in the same sense. It complements the GNU
General Public License, which is a copyleft license designed for free software.
We have designed this License in order to use it for manuals for free software,
because free software needs free documentation: a free program should come with
manuals providing the same freedoms that the software does. But this License is
not limited to software manuals; it can be used for any textual work, regardless of
subject matter or whether it is published as a printed book. We recommend this
License principally for works whose purpose is instruction or reference.
406
GFDL LICENSE 407
below, refers to any such manual or work. Any member of the public is a licensee,
and is addressed as “you”. You accept the license if you copy, modify or distribute
the work in a way requiring permission under copyright law.
A “Modified Version” of the Document means any work containing the Doc-
ument or a portion of it, either copied verbatim, or with modifications and/or
translated into another language.
A “Secondary Section” is a named appendix or a front-matter section of
the Document that deals exclusively with the relationship of the publishers or au-
thors of the Document to the Document’s overall subject (or to related matters)
and contains nothing that could fall directly within that overall subject. (Thus, if
the Document is in part a textbook of mathematics, a Secondary Section may not
explain any mathematics.) The relationship could be a matter of historical connec-
tion with the subject or with related matters, or of legal, commercial, philosophical,
ethical or political position regarding them.
The “Invariant Sections” are certain Secondary Sections whose titles are
designated, as being those of Invariant Sections, in the notice that says that the
Document is released under this License. If a section does not fit the above def-
inition of Secondary then it is not allowed to be designated as Invariant. The
Document may contain zero Invariant Sections. If the Document does not identify
any Invariant Sections then there are none.
The “Cover Texts” are certain short passages of text that are listed, as Front-
Cover Texts or Back-Cover Texts, in the notice that says that the Document is
released under this License. A Front-Cover Text may be at most 5 words, and a
Back-Cover Text may be at most 25 words.
A “Transparent” copy of the Document means a machine-readable copy, rep-
resented in a format whose specification is available to the general public, that is
suitable for revising the document straightforwardly with generic text editors or (for
images composed of pixels) generic paint programs or (for drawings) some widely
available drawing editor, and that is suitable for input to text formatters or for
automatic translation to a variety of formats suitable for input to text formatters.
A copy made in an otherwise Transparent file format whose markup, or absence
of markup, has been arranged to thwart or discourage subsequent modification by
readers is not Transparent. An image format is not Transparent if used for any
substantial amount of text. A copy that is not “Transparent” is called “Opaque”.
Examples of suitable formats for Transparent copies include plain ASCII with-
out markup, Texinfo input format, LaTeX input format, SGML or XML using
a publicly available DTD, and standard-conforming simple HTML, PostScript or
PDF designed for human modification. Examples of transparent image formats in-
clude PNG, XCF and JPG. Opaque formats include proprietary formats that can be
read and edited only by proprietary word processors, SGML or XML for which the
DTD and/or processing tools are not generally available, and the machine-generated
HTML, PostScript or PDF produced by some word processors for output purposes
only.
The “Title Page” means, for a printed book, the title page itself, plus such
following pages as are needed to hold, legibly, the material this License requires to
408 GFDL LICENSE
appear in the title page. For works in formats which do not have any title page
as such, “Title Page” means the text near the most prominent appearance of the
work’s title, preceding the beginning of the body of the text.
A section “Entitled XYZ” means a named subunit of the Document whose ti-
tle either is precisely XYZ or contains XYZ in parentheses following text that trans-
lates XYZ in another language. (Here XYZ stands for a specific section name men-
tioned below, such as “Acknowledgements”, “Dedications”, “Endorsements”,
or “History”.) To “Preserve the Title” of such a section when you modify the
Document means that it remains a section “Entitled XYZ” according to this defi-
nition.
The Document may include Warranty Disclaimers next to the notice which
states that this License applies to the Document. These Warranty Disclaimers
are considered to be included by reference in this License, but only as regards
disclaiming warranties: any other implication that these Warranty Disclaimers may
have is void and has no effect on the meaning of this License.
2. Verbatim Copying
You may copy and distribute the Document in any medium, either commercially or
noncommercially, provided that this License, the copyright notices, and the license
notice saying this License applies to the Document are reproduced in all copies, and
that you add no other conditions whatsoever to those of this License. You may not
use technical measures to obstruct or control the reading or further copying of the
copies you make or distribute. However, you may accept compensation in exchange
for copies. If you distribute a large enough number of copies you must also follow
the conditions in section 3.
You may also lend copies, under the same conditions stated above, and you
may publicly display copies.
3. Copying In Quantity
If you publish printed copies (or copies in media that commonly have printed covers)
of the Document, numbering more than 100, and the Document’s license notice
requires Cover Texts, you must enclose the copies in covers that carry, clearly and
legibly, all these Cover Texts: Front-Cover Texts on the front cover, and Back-Cover
Texts on the back cover. Both covers must also clearly and legibly identify you as
the publisher of these copies. The front cover must present the full title with all
words of the title equally prominent and visible. You may add other material on
the covers in addition. Copying with changes limited to the covers, as long as they
preserve the title of the Document and satisfy these conditions, can be treated as
verbatim copying in other respects.
If the required texts for either cover are too voluminous to fit legibly, you should
put the first ones listed (as many as fit reasonably) on the actual cover, and continue
the rest onto adjacent pages.
GFDL LICENSE 409
4. Modifications
You may copy and distribute a Modified Version of the Document under the con-
ditions of sections 2 and 3 above, provided that you release the Modified Version
under precisely this License, with the Modified Version filling the role of the Docu-
ment, thus licensing distribution and modification of the Modified Version to who-
ever possesses a copy of it. In addition, you must do these things in the Modified
Version:
A. Use in the Title Page (and on the covers, if any) a title distinct from that of
the Document, and from those of previous versions (which should, if there
were any, be listed in the History section of the Document). You may use
the same title as a previous version if the original publisher of that version
gives permission.
B. List on the Title Page, as authors, one or more persons or entities responsible
for authorship of the modifications in the Modified Version, together with at
least five of the principal authors of the Document (all of its principal authors,
if it has fewer than five), unless they release you from this requirement.
C. State on the Title page the name of the publisher of the Modified Version,
as the publisher.
D. Preserve all the copyright notices of the Document.
E. Add an appropriate copyright notice for your modifications adjacent to the
other copyright notices.
F. Include, immediately after the copyright notices, a license notice giving the
public permission to use the Modified Version under the terms of this License,
in the form shown in the Addendum below.
G. Preserve in that license notice the full lists of Invariant Sections and required
Cover Texts given in the Document’s license notice.
410 GFDL LICENSE
The author(s) and publisher(s) of the Document do not by this License give
permission to use their names for publicity for or to assert or imply endorsement
of any Modified Version.
5. Combining Documents
You may combine the Document with other documents released under this License,
under the terms defined in section 4 above for modified versions, provided that
you include in the combination all of the Invariant Sections of all of the original
documents, unmodified, and list them all as Invariant Sections of your combined
work in its license notice, and that you preserve all their Warranty Disclaimers.
The combined work need only contain one copy of this License, and multiple
identical Invariant Sections may be replaced with a single copy. If there are multiple
Invariant Sections with the same name but different contents, make the title of
each such section unique by adding at the end of it, in parentheses, the name of
the original author or publisher of that section if known, or else a unique number.
Make the same adjustment to the section titles in the list of Invariant Sections in
the license notice of the combined work.
In the combination, you must combine any sections Entitled “History” in the
various original documents, forming one section Entitled “History”; likewise com-
bine any sections Entitled “Acknowledgements”, and any sections Entitled “Dedi-
cations”. You must delete all sections Entitled “Endorsements”.
6. Collections Of Documents
You may make a collection consisting of the Document and other documents re-
leased under this License, and replace the individual copies of this License in the
various documents with a single copy that is included in the collection, provided
that you follow the rules of this License for verbatim copying of each of the docu-
ments in all other respects.
You may extract a single document from such a collection, and distribute it
individually under this License, provided you insert a copy of this License into the
extracted document, and follow this License in all other respects regarding verbatim
copying of that document.
ply to the other works in the aggregate which are not themselves derivative works
of the Document.
If the Cover Text requirement of section 3 is applicable to these copies of the
Document, then if the Document is less than one half of the entire aggregate,
the Document’s Cover Texts may be placed on covers that bracket the Document
within the aggregate, or the electronic equivalent of covers if the Document is in
electronic form. Otherwise they must appear on printed covers that bracket the
whole aggregate.
8. Translation
Translation is considered a kind of modification, so you may distribute translations
of the Document under the terms of section 4. Replacing Invariant Sections with
translations requires special permission from their copyright holders, but you may
include translations of some or all Invariant Sections in addition to the original
versions of these Invariant Sections. You may include a translation of this License,
and all the license notices in the Document, and any Warranty Disclaimers, provided
that you also include the original English version of this License and the original
versions of those notices and disclaimers. In case of a disagreement between the
translation and the original version of this License or a notice or disclaimer, the
original version will prevail.
If a section in the Document is Entitled “Acknowledgements”, “Dedications”, or
“History”, the requirement (section 4) to Preserve its Title (section 1) will typically
require changing the actual title.
9. Termination
You may not copy, modify, sublicense, or distribute the Document except as ex-
pressly provided for under this License. Any other attempt to copy, modify, sub-
license or distribute the Document is void, and will automatically terminate your
rights under this License. However, parties who have received copies, or rights,
from you under this License will not have their licenses terminated so long as such
parties remain in full compliance.
either of that specified version or of any later version that has been published (not
as a draft) by the Free Software Foundation. If the Document does not specify a
version number of this License, you may choose any version ever published (not as
a draft) by the Free Software Foundation.
If you have Invariant Sections, Front-Cover Texts and Back-Cover Texts, replace
the “with . . . Texts.” line with this:
with the Invariant Sections being LIST THEIR TITLES, with the
Front-Cover Texts being LIST, and with the Back-Cover Texts being
LIST.
If you have Invariant Sections without Cover Texts, or some other combination
of the three, merge those two alternatives to suit the situation.
If your document contains nontrivial examples of program code, we recommend
releasing these examples in parallel under your choice of free software license, such
as the GNU General Public License, to permit their use in free software.
Index
414
INDEX 415
Tartaglia, 270
Third Isomorphism Theorem
for groups, 165
for rings, 243
Thompson, J., 161, 217
Totally ordered set, 310
Transcendental element, 325
Transcendental number, 325
Transposition, 77
Trisection of an angle, 341