You are on page 1of 69

Algebra I - Lecture Script

Prof. Ozlem
Imamoglu
Mitschrift von Manuela D
ubendorfer
October 19, 2008

ii

Contents
1 Groups
1.1 Basic Definitions and Examples . . . . . . . . . . . . . . . . . . . .
1.1.1 Diedral groups . . . . . . . . . . . . . . . . . . . . . . . . .
1.1.2 Quaternion group . . . . . . . . . . . . . . . . . . . . . . . .
1.1.3 Symmetric group . . . . . . . . . . . . . . . . . . . . . . . .
1.2 Homomorphisms and Isomorphisms . . . . . . . . . . . . . . . . . .
1.3 Subgroups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.3.1 Special subgroups . . . . . . . . . . . . . . . . . . . . . . . .
1.3.2 Cyclic groups an their subgroups . . . . . . . . . . . . . . .
1.4 Group Actions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.5 Normal subgroups and Quotient groups . . . . . . . . . . . . . . . .
1.6 Homomorphisms and Normal subgroups . . . . . . . . . . . . . . .
1.7 Groups acting on themselves by conjugation - class equation . . . .
1.8 Composition series and the Holder program . . . . . . . . . . . . .
1.9 Sylows Theorems . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.10 Direct products and Abelian groups . . . . . . . . . . . . . . . . . .
1.10.1 The fundamental theorem of finitely generated abelian groups
1.10.2 Semidirect products . . . . . . . . . . . . . . . . . . . . . . .
1.11 Free groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1
1
4
5
5
6
7
8
10
13
20
22
27
30
32
37
39
41
42

2 Rings
2.1 Basic Definitions and examples . . . . . . . . .
2.2 Ideals, Ring Homomorphisms and quotient rings
2.3 Properties of ideals . . . . . . . . . . . . . . . .
2.4 Rings of fractions . . . . . . . . . . . . . . . . .
2.5 Polynomials . . . . . . . . . . . . . . . . . . . .
2.5.1 Irreducible polynomials . . . . . . . . . .

45
45
49
55
59
61
64

iii

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

iv

CONTENTS

Chapter 1
Groups
1.1

Basic Definitions and Examples

Definition 1.1.1. 1) A group is an ordered pair (G, ) where G is a set and is


a map
: G G G, (g, h) 7 g h := g h =: gh
satisfying the following axioms
(G1) For all a, b, c G, (a b) c = a (b c),
(G2) There exists an element e G such that a = a e = e a a G, e is called
the identity,
(G3) For all a G, there exists a1 G such that a a1 = a1 a = e, a1 is
called the inverse of a.
2) The group (G, ) is called abelian (or commutative) if a b = b a, for all
a, b G.
Remark. (1) Well drop (G, ) and write just G.
(2) We say G is a finite group if G is a finite set.
(3) Note (G2) ensures that a group is always nonempty.
(4) When the group G is abelian we typically write a + b for ab and a b for ab1 .
Example 1.1.2. Z, Q, R, C groups under +, with e = 0 and a1 = a for all a.
Example 1.1.3. Q\{0}, R\{0}, C\{0}, Q+ , R+ groups under with e =
1, a1 = a1 , a. Z\{0} is NOT a group under . (2 does not have an inverse for
example.)
Example 1.1.4. Any vector space (V, +) is an abelian group (The operation +
is the vector addition).
1

CHAPTER 1. GROUPS

Example 1.1.5. Integers mod n : Z/nZ


Let n be a fixed positive integer. Define a relation on Z by a b if and only if
n|(b a). We write a b ( mod n) if a b. a is congruent to b mod n.
is an equivalence relation a a (reflexive), a b b a (symmetric),
a b, b c a c (transitive).
a
= {a + kn; k Z} = {a, a n, a 2n, . . . }
is the congruence class of a residue class of a. There are precisely n distinct
equivalende classes mod n, namely 0, 1, 2, . . . , n 1. The equivalence classes
are called the integers modulo n, denoted by Z/nZ. We can define a sum and
product by
b = ab.
a
+ b := a + b, a
Theorem 1.1.6. The operation of addition and multiplication on Z/nZ defined
above are both well defined, i.e. they do not depend on the choices of representatives
for the classes involved. More precisely:
If a1 , a2 Z, b1 , b2 Z and a
1 = b1 , a
2 = b2 then a1 + a2 = b1 + b2 and a1 a2 =
b1 b2 , i.e. if a1 b1 mod n and a2 b2 mod n then a1 + a2 b1 + b2 mod n
and a1 a2 b1 b2 mod n.
Proof. Suppose a1 b1 mod n, i.e. a1 b1 = ns for some integer s. Similarly a2
b2 mod n means a2 b2 = nt for some t Z. Then a1 + a2 = (b1 + b2 ) + (s + t)n,
i.e. a1 + a2 b1 + b2 mod n.
An important subset of Z/nZ consists of the collection of residue classes which
have a multiplicative inverse in Z/nZ:
(Z/nZ) = {
a Z/nZ;
c Z/nZ with a
c = 1}
Proposition 1.1.7. (Z/nZ) = {
a Z/nZ; (a, n) = 1}.
Proof. Exercise.
Example 1.1.8. n = 4 : Z/4Z = {0, 1, 2, 3}, (Z/4Z) = {1, 3}.
(Z/nZ, +) and ((Z/nZ) , ) are groups.
Example 1.1.9. If (A, ), (B, ) are two groups, we can form a new group A B,
the direct product whose elements are
A B = {(a, b); a A, b B}.
We define the operation componentwise
(a1 , b1 )(a2 , b2 ) = (a1 a2 , b1 b2 ).

1.1. BASIC DEFINITIONS AND EXAMPLES

Example 1.1.10. GLn (R) denotes the group of invertible n n matrices under
matrix multiplication.
Example 1.1.11. Let X be a set, S(X) Permutations of X
S(X) = {f : X X; f a bijection}
with composition as group oparation. When X = {1, . . . , n} we write Sn , the
Symmetric group on n letters.
1
. . a1}.
For any group G, a G, n Z+ , we denote an by aa
. . . a}, an by a
| .{z
| {z
n times

n times

a0 = 1 is the identity.
Proposition 1.1.12. Let (G, ) be a group then
(1) The identity of G is unique,
(2) For all a G, a1 is unique,
(3) (a1 )1 = a, a G,
(4) (a b)1 = (b1 ) (a1 ),
(5) for any a1 , . . . , an G, a1 an is independent of how the expression is
bracketed.
Proposition 1.1.13. Let G be a group, a, b G. The equations ax = b and
ya = b have unique solutions x, y G. In particular, the left and right cancelation
laws hold.
(1) If au = av then u = v,
(2) if ub = vb then u = v.
Definition 1.1.14. Let G a group, and x G. We define the order of x to be the
smallest positive integer n such that xn = 1. We denote this integer by |x|. If no
positive power of x is 1, x is said to be order.
Example 1.1.15. (1) a G has |a| = 1 if and only if a = 1.
(2) (Z, +), (R; +): every non-zero element has order.
(3) (R\{0}, ): 1 has order 2, all other elements have order.
(4) G = Z/6Z: 2 6= 0, 2 + 2 6= 0, 2 + 2 + 2 = 0, hence 2 has order 3 in Z/6Z.
Definition 1.1.16. Let G = {1 = g1 , . . . , gn } be a finite group. The multiplication
table or group table of G is the n n matrix whose ij-entry is the group element
gi gj .
Example 1.1.17. G = ((Z/3Z), +)
+ 0 1 2
0 0 1 2
1 1 2 0
2 2 0 1

CHAPTER 1. GROUPS

There is another formulation of the axioms for a group, which is equivalent to the
original definition but somewhat simpler
Theorem 1.1.18. Let G be a set, an associative binary operation on G. Assume
that there exists an element e G such that x e = x for all x G and assume
that for any x G there exists an element y G such that x y = e. Then (G, )
is a group (e is called right identity, y is called right inverse)
Proof. We want to show that e is also a left identity and y is also left inverse of
x. Let x G, e e = e since x e = x, x G. Let y be the right inverse of x,
i.e. x y = e. Then x y = e = e e = e (x y) = (e x) y by association.
x y = (e x) y. Now we use right of y to cancel ys. x = e x, therefore e is
also a left identity. The other part is an exercise.
Note the following example
Example 1.1.19. (Z, ) where x y := x. Check: ist associative, 1 is a right
identity x 1 = x, x Z and 1 is a left inverse for every x Z, 1 x = 1. But
(Z, ) is not a group since for example there is no two-sided identity element.
Now we will study some very important examples of groups

1.1.1

Diedral groups

One important family of examples of groups is the class of groups whose elements
are symmertries of geometric objects. The simplest subclass is when the geometric objects are regular polygons. For each n Z+ , n 3, let D2n be the set
of symmetries of a regular n-gon. Fix a regular n-gon centered at the origin in
the xy plane. Label the vertices consecutively from 1 to n in clockwise direction.
and s the reLet r denote the rotation clockwise about the origin through 2
n
flection about the line of symmetry through vertex 1 and origin. Then one can
verify
(a) 1, r, r2 , . . . , rn1 all distinct and rn = 1, i.e. |r| = n,
(b) |s| = 2,
(c) s 6= ri for any i,
(d) sri 6= srj 0 i, j n with i 6= j.
D2n = {1, r, r2 , . . . , rn1 , s, sr, sr2 , . . . , srn1 }
{z
} |
{z
}
|
rotations

reflections

a group of order 2n. Every element can be written in terms of r and s. We say
r, s are generators. Well look at this example in more detail later. (Note also
sr = r1 s and hence D2n is not abelian).

1.1. BASIC DEFINITIONS AND EXAMPLES

1.1.2

Quaternion group

Let Q8 = {1, 1, i, i, j, j, k, k} with the relations 1 a = a 1 = a, for all


a Q8 , (1) (1) = 1, (1) a = a (1) = a and
i i = j j = k k = 1
ij = k, jk = i, ki = j, ji = k, kj = i, ik = j.
Q8 is a non abelian group of order 8.

1.1.3

Symmetric group

Let X 6= be a set, and Sx be the set of all bijections from X to X. Then (Sx , )
is a group, where denotes the composition of functions, the permutations of X.
When X = {1, 2, . . . , n}, then we write Sx = Sn , which is the symmetric group on
n letters.
First note |Sn | = n! If is any permutation in Sn , it can send 1 to any of {1, . . . , n},
hence there are n choices for (1), n 1 choices for (2), n 2 choices for (3), . . . .
Therefore there are n! injective functions from {1, . . . , n} to itself.
An efficient way of writing elements of Sn is through their cycle decomposition.
Definition 1.1.20. A cycle is a string of integers which represents the elements
of Sn which cyclically permutes these integers and fixes all others.
r-cycle Sn when there exists r-element subset {x1 , . . . , xr } {1, . . . , n} such
that
(a) (xi ) = xi+1 , 1 i r,
(b) (xr ) = x1 ,
(c) (y) = y y {1, . . . , n}\{x1 , . . . , xr }
We write (x1 , x2 , . . . , xr ) for the r-cycle .
Example 1.1.21. = (15732) S8 is the permutation


1 2 3 4 5 6 7 8
5 1 2 4 7 6 3 8
In general for every Sn , the numbers from 1 to n will be rearranged and
grouped into k-cycles
(a1 . . . am1 )(am1 +1 am1 +2 . . . am2 ) . . . (amk1 +1 . . . amk ).

CHAPTER 1. GROUPS

Example 1.1.22. (153)(27)(68) S8




1 2 3 4 5 6 7 8
5 6 1 4 3 8 2 6
Well study this group in more detail later.
Consider s = (123) S3 , |S3 | = 6 and r = (23) S3 ,
D6 are the symmetries of a triangle and has order 6, too. r can be represented by
its effect on 3 corners as (123), s is a reflection which fixes 1 and swaps 2, 3. In
fact one can show there is a bijection
: S3 D6 , (123) 7 r, (23) 7 s.
They look alike. Next we make this notion more precise.

1.2

Homomorphisms and Isomorphisms

Definition 1.2.1. Let (G1 , ), (G2 , ) be two groups. A map : G1 G2 is


called a homomorphism if
(x + y) = (x) (y), x, y G1 .
Well typically not write the group operations explicity, (xy) = (x)(y),
is called an isomorphism and G1 , G2 are called isomorphic (notation: G1
= G2 )
if is a bijection and is a homomorphism, i.e. G1 , G2 are isomorphic if there
exists a bijection between them which preserves the group operations. (onto homomorphism = epimorphism, 1-1 homomorphism = monomorphism).
Example 1.2.2. For any group G, G
= G2 , G2
= G3 then G1
= G3 ;
= G; if G1

if G1 = G2 then G2 = G1 . Hence = is an equivalence relation. The equivalence


classes are called isomorphism classes.
Example 1.2.3. exp : R R+ , x 7 ex . ex+y = ex ey is a bijection with inverse
log : R+ R, hence (R, +)
= (R+ , ).
Example 1.2.4. Let x, y be finite sets, non-empty. Then Sx
= Sy |x| = |y|.
Example 1.2.5. Well be able to prove that any non-abelian group of order 6 is
isomorphic to S3 . Hence indeed D6
= S3 .

1.3. SUBGROUPS

Example 1.2.6. : G G, g 7 e g G and : G G, g 7 g are trivial


homomomorphisms.
Example 1.2.7. : (Z, +) (Z/nZ, ), m 7 m
= m mod n is a homomorphism but not a monomorphism (not 1-1), because (0) = (n), but it is an
epimorphism, it is onto.
Example 1.2.8. Sometimes it is easy to see that two groups are not isomorphic.
For example S3
6= Z/6Z since Z/6Z is abelian and S3 is not.
Example 1.2.9. Fix g G, G a group. The map g : G G, x 7 gxg 1 is an
isomorphism of G onto itself (automorphism). The map is called conjugation.

1.3

Subgroups

Definition 1.3.1. Let G be a group. H G, a subset of G. H is called a subgroup


of G, written H G, if H 6= and H is closed under products and inverses, i.e.
a subset 6= h of a group (G, ) is a subgroup of G if the elements of H form a
group under .
Note: e H : if x H then x1 H and x1 x = e H since H is closed
under products and inverses. When we say H is a subgroup, we always mean
that the operation for H is the operation on G restricted to H. For example:
(Q\{0}, ) 6 (R, +) even though both are groups, Q\{0} R as sets, but =
6 +.
Example 1.3.2. (Q, +) (R, +).
Example 1.3.3. (2Z, +) (Z, +).
Example 1.3.4. G a group, g G, H = hgi := {g n ; n Z}, where g 0 := 1, is
a subgroup, called the cyclic subgroup generated by g. If hgi = G, G is called a
cyclic group. For example Z = h1i.
Example 1.3.5. Let G be the group of real-valued functions on the real line,
under addition of functions
G = {f : R R}, (f + g)(x) = f (x) + g(x).
Let H = {continuous functions} ( G. Then H is closed under addition and
inverses (f, g continuous, so is f + g and f ), hence H is a subgroup.
Example 1.3.6. If G = (Z, +), n any integer, then the set nZ = hni of multiples
of n is a subgroup of G. In particular for n = 0 we get the trivial subgroup {0},
for n = 1 we get G itself.

CHAPTER 1. GROUPS

In fact nZ are all the subgroups of (Z, +). For if H is a subgroup of G, other than
{0}, then there exists positive integers in H. Let n = min{m Z+ ; m H}.
Then we claim H = nZ. Clearly since n H, H a subgroup nZ H.
To see that H nZ, note that if h H we can write h = qn + r, 0 r < n
by division algorithm. Then r = h qn H since h H, qn nZ H. But
this contradicts the minimality of n unless r = 0 in which case h = qn nZ. So
H nZ.

1.3.1

Special subgroups

Definition 1.3.7. Let G be any group. The center of G, denoted by Z(G) is the
set of elements that commute with everything in G, i.e.
Z(G) = {u G; zg = gz g G}.
Remark. If G is abelian then G = Z(G). If G is not abelian, then Z(G) $ G.
Further we can show that Z(G) is a subgroup of G:
Proof. Since if z1 , z2 Z(G) then z1 g = gz1 and z2 g = gz2 g G. It follows
(z1 z2 )g = z1 (z2 g) = z 1(gz2 ) = (z1 g)z2 = g(z1 z2 )
hence z1 z2 Z(G). Further
z11 g = (g 1 z1 )1 = (z1 g 1 )1 = gz11 ,
and so z11 Z(G). Clearly e Z(G) and eg = ge = e by definition of e. Therefore
we have Z(G) < G.
na b 
o
Example 1.3.8. Let G = GL2 (R) =
; a, b, c, d R, ad bc 6= 0 . Let
c d




a b
0 1
Z(G), then since
G and
c d
1 0


 



 

a b
0 1
0 1
a b
b a
c d
=
we have
=
.
c d
1 0
1 0
c d
d c
a b

 

a b
a b
and hence we have a = d, b = c and therefore
=
.
c d
b a


1 1
On the other hand since
GL2 (R) then
0 1


 



 

a b
1 1
1 1
a b
a a+b
a+b b
=
, hence
=
b a
0 1
0 1
c d
b b+a
b+a a

1.3. SUBGROUPS



a 0
and therefore b = 0. Hence every element in Z(G) has the form
, a 6= 0.
0 a




a 0
1 0
Converse is easy clearly
=a
Z(G), and therefore
0 a
0 1
na 0
o
Z(GL2 (R)) =
; a 6= 0 .
0 a
Definition 1.3.9. Another imoportant subgroup is the Centralizer of an element
x in G:
CG (x) = {g G; gx = xg} = {g G; gxg 1 = x}.
Remark. Check that CG (x) < G. Further if G is abelian then CG (x) = G, x G.
Definition 1.3.10. Two elements x, y G are called conjugate in G if there exists
g G such that gxg 1 = y.
Remark. This defines an equivalence relation on G: a b if there exists g G
such that a = gbg 1 .
Definition 1.3.11. The equivalence class a is called the conjugacy class of a
a = {gag 1 ; g G}.
One more example of an important subroup
Definition 1.3.12. Let H be a subgroup of a group G. Let x G and x1 Hx
denote {x1 ax; a H}. First note x1 Hx is also a subgroup of G. Then
NG (H) = {g G; g 1 Hg = H} = {g G; Hg = gH}
is called the normalizer of H, where gH = {gh; h H}.
Remark. Then NG (H) < G. Note H < NG (H) trivially.
To verify that a subset H G is a subgroup, we check that it is closed under
multiplication and inverses. These two can be combined to give
Proposition 1.3.13. Let G be a group. A non-empty subset H of G is a subgoup
iff for all x, y H we have xy 1 H. If H is finite, then is suffices to check that
it is closed under muliplication.
Proof. If H G then certainly for all x, y H we have xy 1 H. Assume =
6 H
satisfies xy 1 H x, y h. We apply this with y = x then xy 1 = xx1 = 1 H,
hence the identity is in H. Further 1x1 = x1 H, and hence H is closed under
multiplication and therefore a subgroup of G.
Suppose H is finite and closed under multiplication. Let x H. Then {x, x2 , . . . }
is a finite set, hence xa = xb for some a, b with b > a. If n = b a then xn = 1, so
every element of H has finite order and x1 = xn1 H, hence H is also closed
under inverses.

10

CHAPTER 1. GROUPS

Definition 1.3.14. If : G H is a group homomorphism, then the kernel of


is
ker = {g G; = 1H }.
Proposition 1.3.15. ker is a subgroup of G.
Proof. If g1 , g2 ker then
(g1 g2 ) = (g1 )(g2 ) = 1 1 = 1,
hence g1 g2 ker .
1 = (1) = (g1 g11 ) = (g1 )(g11 ) = 1 (g11 )
and therefore (g11 ) = 1, hence g 11 ker .

1.3.2

Cyclic groups an their subgroups

Definition 1.3.16. A group G is called cyclic if there exists g G such that


G = hgi = {g n ; n Z}. g is called a generator.
If the group operation is additive we write this {ng; n Z}.
Weve seen in the spacial case of the cyclic group Z = h1i that every subgroup is
of the form nZ = hni. Note that in case of Z, 1 is also a generator.
Proposition 1.3.17. Let G be a cycic group, G = hgi. Then |G| = |g|. More
specifically:
(1) If |G| = n < then g n = 1 and 1, g, g 2 , . . . , g n1 are all distinct elements of
G.
(2) If |G| = then g n 6= 1 n 6= 0 and g a = g b a 6= b in Z.
Proof. Let |g| = n. First assume, n < . Then 1, g, . . . , g n1 are all distinct since
if g a = g b then g ab = 1 and a b < n contradicting that n = |g| is tha smallest
power such that g n = 1. Thus G has at least n elements 1, g, . . . , g n1 and these
are in fact all of them.
Since let g t be any element of G. Using division algorithm we can write t = nq + k
with 0 k n then g t = g nq+k = (g n )q g k = g k {1, g, . . . , g n1 }.
If |g| = , then no positive power of g is 1. If g a = g b for some a, b say a < b then
g ba = 1 a contradiction. Thus the distinct powers of g are all distinct elements
of G, |G| = .
Proposition 1.3.18. Let G be a group, g G, m, n Z. If g m = 1 and g n = 1
then g d = 1 where d = gcd(m, n). In particular if g m = 1 for some m Z, then
|g| | m.

1.3. SUBGROUPS

11

Proof. By Euclidean algorithm we can find r, s Z such that d = mr + ns.


Therefore g d = g mr g ns = 1. The second part we leave as exercise.
Proposition 1.3.19. Let G be any group, g G, a Z\{0}
(1) If |g| = then |g a | = ,
(2) if |g| = n < then
n
.
|g a | =
(n, a)
Proof. (1) Assume on the contrary |g| = , but |g a | = m < .
1 = (g a )m = g am = 1, also g am = (g a )m = 1,
so either am, am > 0 and this contradicts |g| = .
(2) Let h = g a , (n, a) = d, n = dn0 , a = da0 for some n0 , a0 with n0 > a0 . We want
to show |h| = nd = n0 .
0

0 0

0 0

hn = g an = g a n = g da n = (g dn )a = 1
Applying proposition 1.3.18 |h| divides n0 . Let |h| = k. Then k|n0 and g ak =
hk = 1. Applying proposition 1.3.18 again to hgi, we have |g| = n | ak and
dn0 | ak = da0 k, so n0 | a0 k, but (n0 , a0 ) = 1, hence n0 | k together with k | n0 gives
k = n0 .
The next proposition gives a criteria to find all generators of a cyclic group.
Proposition 1.3.20. Let G = hgi.
(1) If |g| = then G = hg a i if and only if a = 1.
(2) If |g| = n < then G =< g a > if and only if (a, n) = 1. In particular the
number of generators of G is (n), Eulers -function
(n) := #{k Z; 1 k n, (n, k) = 1}
Proof. Exercise.
Next well see that subgroups of a cyclic subgroup as in the case of Z, have special
stucture.
Theorem 1.3.21. Let G = hgi be a cyclic group.
(1) Every
of G is cyclic, more precisely if H G then either H = {1}

dsubgroup

or H = g where d is the smallest positive integer such that g d H.

b
a
(2) If |G| = then for any
distinct
non-negative
integers
a
and
b,
hg
i
=
6
g .


For any integer m hg m i = g |m| where |m| is the absolute value. The nontrivial
subgroups of G correspond bijectively with positive integers.

12

CHAPTER 1. GROUPS

(3) If |G| = n < . Then for each positive integer a dividing

n there exists a
unique subgroup H of G of order a.
This subgroup H is g d where d = na . For
every integer m, we have hg m i = g (n,m) , i.e. the subgroups of G correspond
bijectively with positive divisors of n.
Proof. (1) Let H G. If H = {1} there is nothing to prove. So assume H 6= {1}.
Then there exist a 6= 0 such that g a H. If a < 0 then since H is a subgroup,
g a H and a > 0. Hence H contains some positive power of g. By well
ordering principle the set
S = {a; Z+ , g a H}
has a minimum, call it d. Since h is a subgroup of G = hgi every element h of H
is of the form g t where t = qd + r. Therefore we have g qd+r = (g d )q g r H, but
(g d )q H
and g r H. This contradicts the minimality of d unless r = 0, t = qd,
hence h g d .
(2) Is similar to the proof of (3).
(3) Assume |G| = n < and a|n. Let d = na . Then
n
|g|
= =a
(|g| , d)
d


by proposition 1.3.19. Hence H = g d is a subgroup of order a.
d
g =

To show

uniqueness, suppose K is any subgroup of order a. Then by part (1)


K = g b where b is the smallest power of g such that g b K.

a = |K| = g b =

|g|
n
=
,
(|g| , b)
(n, b)

n
and hence a = nd = (n,b)
and d = (n, b). In particular d | b, so b = db0 and


0
g b = (g d )b
g d , and therefore K = g b g d . But < g d > = a = |K|,
hence K = g d .

The final assertion of (3) follows because


hg m i g (m,n) , |g m | =



n
n
n
and g (m,n) =
=
,
(n, m)
(n, (m, n))
(n, m)


then hg m i = g (m,n) . Since (m, n) is a divisor of n, every subgroup of G arises
from a divisor of n.

1.4. GROUP ACTIONS

13

Example 1.3.22. G = (Z/12Z, ). Then G =< 1 >, |G| = 12. Positive divisors
of 12 are 1,2,3,4,6,12, hence the distinct subgroups of G are
G = h1i ,
order12

h2i ,
12
=6
(2, 12)

h3i ,

h4i ,

12
=4
(3, 12)

12
=3
4

h6i ,

h0i = {e}

Note h5i = h1i = G since (5, 12) = (1, 12). Similarly h7i = h11i = 1. We have
h8i = h4i since h8, 12i = h4, 12i and h9i = h3i, h10i = h2i. Lattice of subgroups
G = h1i


h2i


h3i


h4i


h6i


h0i

Theorem 1.3.23. (a) Let n be a positive integer, G a cyclic group of order n.


Then G
= (Z/nZ, ), i.e. any cyclic group of order n are isomorphic.
(b) Let G be a cyclic group. Then G
= Z. Any two infinite cyclic groups are
isomorphic.
Proof. Exercise Serie 2.

1.4

Group Actions

Definition 1.4.1. A group action of a group G on a set X is a map from GX


X, written g x, for all g G, x X with the following properties
(1) g1 (g2 x) = (g1 g2 ) x g1 , g2 G, x X,
(2) 1 x = x, x X.
In (1) the product (g1 g2 ) is taken in G.
Note: For each g G, we get a map
g : X X, x 7 g x.
Fact 1: g is a permutation of X, namely g is a bijection with inverse g1 . For,
let x X, then
(g1 g )(x) = g1 (g x) = g 1 (g x) = (g 1 g) x = 1 x = x.
Therefore g1 g : X X is the identity map. Changing the roles of g, g 1
gives that g g1 is also the identity map, hence g is a bijection.

14

CHAPTER 1. GROUPS

Fact 2: The map : G SX is a homomorphism. To see this we need to


show
(g1 g2 ) = (g1 ) (g2 ).
The permutations (g1 g2 ) and (g1 )(g2 ) are equal iff their values agree on every
x X.
(g1 g2 )(x) = g1 g2 (x)
= (g1 g2 ) x
= g1 (g2 x)
= g1 (g2 (x))
= ((g1 ) (g2 ))(x)

by definition of ,
by definition of g1 g2 ,
by property 1. of group actions
definition of g1 , g2 ,
definition of .

Definition 1.4.2. The homomorphism : G Sx is called the permutation


representation associated to the given action.
This process is reversible, in the sense that if : G Sx is any homomorphism
from G to a symmetric group on a set X, then the map from G X X defined
by g x = (g)(x) gives a group action of G on X. (Group action of G on X
means that every element g in G acts as a permutation of X in a way consistent
with the group operations in G.)
Example 1.4.3. Trivial action:
G X X, (g, x) 7 x x X, g G.
G is said to act trivially on X. Note distinct elements of g all induce the same
permutation on X, namely the identity permutation. The associated permutation
representation is
: G SX , g 7 id.
Definition 1.4.4. Any action of G on X is said to be faithful if
g1 6= g2 g1 6= g2 ,
i.e. an action is faithful when the associated permutation representation is injective.
Definition 1.4.5. The kernel of an action of G on X is the set
{g G; g x = x x X}.
These are exactly the elements of G that fix all elements of X.

1.4. GROUP ACTIONS

15

Note that this kernel is the same as ker , : G SX , g 7 g . For trivial


action this kernel is G. Check: The kernel of an action is a subgroup of G.
Example 1.4.6. Every group acts on itself by left (right) multiplication
G G G, (g, a) 7 ga
Each fixed g G permutes elements of G by left multiplication
g : a 7 ga
(if the group is written additively, a 7 g + a, left translation).
Definition 1.4.7. This action is called left regular action of G on itself.
By cancelation law in groups this is a faithful action.
Example 1.4.8. Let G be a group, X = G. The map
G G G, (g, a) 7 gag 1 g, a G
is a (left) group action. This action is called conjugation.
For fixed g, G G, a 7 gag 1 is an isomorphism. The kernel of this action
{g G; gag 1 = a a G} = {g G; ga = ag, a G} = Z(G).
Lets consider the general case, where G acts on X, x X.
Definition 1.4.9. The stabilizer of x in G is defined as
Gx = {g G; g x = x}.
Proposition 1.4.10. Gx is a subgroup of G.
Proof. Exercise.
Proposition 1.4.11. Pick x X, g G and set y = g x. Show that Gy =
gGx g 1 .
Example 1.4.12. Let X = P(G) = {all subsets of G}. Then G acts on X by
conjugation: for each g G, B G
g : B 7 gBg 1 = {gbg 1 ; b B}.
The stabilizer of A G is
GA = {g G; gAg 1 = A} = {g G; gA = Ag} = NG (A),
the normalizer of A in G.

16

CHAPTER 1. GROUPS

Example 1.4.13. The group G = SL(2, R) acts on the UHP (upper half plane)
H = {z = x + iy; y > 0} via
a b  
az + b
.
GHH ,
, z 7
c d
cz + d
Since

 az + b 
im(z)
im
=
>0
cz + d
|az + d|2


1 0
z =z
0 1

acts trivially. Check that



 
a b  A B 
Az + B 
a b
.
z =

c d
C D
c d
Cz + D
Let i H. The stabilizer of i is
na b  ai + b
o
Gi =
;
=i
c d
ci + d
therefore ai + b = c + di and a = d,
n a
Gi =
b

b = c, so

o
b
; a2 + b 2 = 1 ,
a

a compact group! SO(2).


Proposition 1.4.14. Given a group G acting on a set X. The relation on X
defined by x y iff x = g y for some y G, defines an equivalence relation.
Proof. By definition of group action 1 x = x x X, hence x x.
If x y, g G such that x = g y, then g 1 x = g 1 g y = 1 y = y, hence
y x.
If x y, y z then g, h G such that x = gy, y = hz, therefore x = g (hz) =
gh z and x y.
Definition 1.4.15. The equivalence class containing x is called the orbit of x or
coset of x under G
Gx = {gx; g G}.
Example 1.4.16. G = SL(2, R), S = H, x = i. Then
Gi = {g i; g G}.

1.4. GROUP ACTIONS


Note since let

17



y x
GL2 (R),
i 1

since y > 0 and g i = yi + x = z H for any x + iy = z H, Gi = H, so there


is just one orbit.
In General:
Definition 1.4.17. If G acts on X, and there is only one orbit, the action is
called transitive, i.e. for all x, y X, there exists g G such that g x = y.
The orbits of a group action G on X partitions the set X into disjoint sets, i.e.
[
X=
Gx,
xR

where R is a set of representations for the equivalence classes. We write for the
set of orbits G\X if G acts on the left and X/G if G acts on the right.
Example 1.4.18 (important example). We have seen that G acts on itself by
multiplication. One can also make a subgroup H of G act on G by multiplication
on the left or on the right
r : H G G, (h, g) 7 hg (on the right)
l : H G G, (h, g) 7 gh1 (on the left).
The orbit of g G then has the form
Hg = {hg; h H}
for some action r and is called a right coset of H in G (or gH = {gh; h H} for
some action l is the left coset of H in G).
Example 1.4.19. 3Z Z act on Z by translation
3Z |{z}
Z Z, (3n, m) 7 3n + m.
|{z}
=G

=X

Orbits, Cosets:
3Z + 0 = {0, 3, 6, . . . }
3Z + 1 = {. . . , 5, 2, 1, 4, 7, . . . }
3Z + 2 = {. . . , 4, 1, 2, 5, 8, . . . }

= 3Z + 3 = 3Z + 6, . . . ,
= 3Z + 4 = 3Z + 7 = 3Z 2, . . . ,
= 3Z + 5 = 3Z 1 = 3Z + 8, . . . .

Z = 3Z 3Z + 1 3Z + 2 a set of representories: 3Z\Z = {0, 1, 2}. Note in this


case 3Z + 0 = 0 + 3Z, left and right cosets of 3Z in Z are the same.

18

CHAPTER 1. GROUPS

This is not the case in general (see the next series for an example, G = S3 , H =
h1, (23)i).
Example 1.4.20. Z R under addition. Z acts on R by translation.
Z R R, (n, r) 7 n + r.
How does the orbits look like? What is a set of representatives for this action?
For any r R, the orbits or the right cosets are
Z + r = {n + r; n Z}.
Write r =

[r]
|{z}

+{r}, then {r} [0, 1). Then Z + r = Z + {r} and if Z + s =

integer part

Z + {r} then assume analog {s}


S {r} then {s} {r} Z and {s} {r} [0, 1),
hence {s} = {r} and R =
Z + {r}. A set of representatives is the interval
r[0,1)

[0, 1). If one identifies 0 and 1 then we get the circle.


Example 1.4.21. Let = Z + Zi = {m + ni; m, n Z} C. acts on C by
translation
C C, (l, z) 7 l + z.
OrbitsS L + z. A set of repesentatives for C/L is {m + ni; 0 m, n < 1},
C=
L + z.
zP

Example 1.4.22. G = (Z/2Z, ), H =< 4 >= {0, 4, 8}. The right cosets of H
are
H =H 0
=H 1
=H 2
=H 3

= {0, 4, 8}
= {1, 5, 9}
= {6, 10, 2}
= {7, 11, 3}

=H 4
=H 5
=H 6
=H 7

=H 8
=H 9
= H 10
= H 11

G = {0, 4, 8} {1, 5, 9} {6, 10, 2} {7, 11, 1} .


| {z } | {z } | {z } | {z }
0

Note each coset has the same number of elements, i.e. 3 elements. There are 4
different cosets.
Definition 1.4.23. The number of distinct right cosets of H is called the index of
H in G, denoted by [G : H].

1.4. GROUP ACTIONS

19

Example 1.4.24. (1) For H = 3Z, G = Z : [G : H] = 3,


(2) H = Z, G = R then [G : H] = ,
(3) And for H =< 4 >, G = Z/12Z, we have [G : H] = 4, because |G| = 12,
|H| = 3.
Theorem 1.4.25 (Lagranges theorem). Let G be a group, H G. Then
|G| = [G : H] |H| .
In particular the order of H and the order of any element of G divides the order
of G.
Proof. G is a disjoint union of its right cosets, hence it suffices to prove that the
order of any coset is equal to the order of |H| = |He|. Let g G, define a map
: H Hg, h 7 hg.
It suffices to check that is a bijection. is clearly surjective. Suppose (h1 ) =
(h2 ), then h1 g = h2 g. But then h1 = h2 and is injective and we are done.
Corollary 1.4.26. Every group G of prime order is cyclic.
Proof. By assumption |G| > 1. Pick any element g G, g 6= e. Let H = hgi.
Then the order of H divides |G| = p. Since p is prime, |H| = 1 or p. Since
e 6= a H, |H| =
6 1, |H| must be p, so |G| = |H| and G = H = hgi is cyclic.
Warning: Full converse of Lagranges theorem is not true, i.e. if G is a finite
group, and n | |G| then G need not have a subgroup of order n.
Weve seen for a group G, H a subgroup of G, the sets of left (or right) cosets of
H in G is
{Hg; g G} = G/H (or {Hg; g G} = H\G}).
A natural question is: Under what circumstances can we put a group structure on
the set of left cosets of H in G? Suppose we are given two left cosets g1 H, g2 H.
Then how do we define a product? We ought to define
(g1 H)(g2 H) = g1 g2 H.
The problem with this is that to define the product we had to choose a repesentative of the left coset g1 H, namely g1 . However we could just as well chosen any
other element of this left coset as a representative and we could potentially get
another choice of left coset for the product, i.e. if aH = a0 H, bH = b0 H then is it
true that abH = a0 b0 H? In general not!!!

20

CHAPTER 1. GROUPS

Example 1.4.27. G = S3 , H = {1, (23)} = {1, s}, s = (23), r = (123), then


check that
rH = {r, rs} = (rs)H
r2 H = {r2 , r2 s} = (r2 s)H
We define
(rH)(r2 H) = r3 H = H.
On the other hand if we choose rs as the representative for rH, then
((rs)H)(r2 H) = rsr2 H = (13)H = {(13), (13)(23)} =
6 H.
Different representatives give rise to different answers for the procuct, which is not
good!
Solution: We can really only define a product on the set of cosets in a well-defined
manner if the group H is special! That brings us to the next section.First note
that
Lemma 1.4.28. Let H be a subgroup of G and a, b G. Then aH = bH iff
a bH iff b1 a H.
Proof. Exercise.

1.5

Normal subgroups and Quotient groups

Definition 1.5.1. Let G be a group, H be a subgroup. We say that H is normal


in G, and write H / G, if g 1 Hg H for all g G.
Proposition 1.5.2. Let G be a group, H G. Then it is equivalent:
i) for every g G, gHg 1 H,
ii) for every g G, gHg 1 = H,
iii) Ha = aH, a G,
iv) The set of left cosets is equal to the set of right cosets (i.e. for every a
G, aH = Hb for some b G).
Proof. Exercise
The significance of the concept of the normal subgroup lies in the next theorem
Theorem 1.5.3. Let G be a group, H E G. Then the set of left cosets is a group
denoted by G/H, where the multiplication is defined by the rule
(g1 H)(g2 H) := (g1 g2 )H.
The identity element of G/H is eG/H = eH = H. The inverse of an element
gH G/H is given by g 1 H.

1.5. NORMAL SUBGROUPS AND QUOTIENT GROUPS

21

Proof. It suffices to check that this rule of multiplication is well-defined. (The fact
the multiplication is associative is inherited from the associativity of multiplication
in G.) Pick two left cosets g1 H, g2 H. Suppose gi H = gi0 H, i = 1, 2, i.e. gi0 is another
representatives for gi H.We want to show that g1 g2 H = g10 g20 . If gi H = gi0 h then
hi H such that gi0 = gi hi . Then
g10 g20 = (g1 h1 )(g2 h2 ) = g1 (h1 g2 )h2 .
Now since
H / G, g 1 Hg H, g21 h1 g2 g 1 Hg H
there exists h01 such that g21 h1 g2 = h01 , i.e. h1 g2 = g2 h01 for some h01 H. Therefore
we have g10 g20 = g1 (h1 g2 )h2 = g1 g2 h01 h2 and (g10 g20 )H = (g1 g2 )H as wanted.
Example 1.5.4. Check that for G = S3 , H = {1, (23)} H 6C G since right cosets
6= left cosets.
Example 1.5.5. G = S3 , K = h(123)i. Check that K / G.
Example 1.5.6. If : G G0 a group homomorphism. Then weve seen that
ker is a subgroup of G. In fact ker / G. To see this, let H = ker . Take an
aribitrary element g 1 hg in g 1 Hg then
(g 1 hg) = (g 1 )(h)(g) = (g)1 e(g) = e.
Therefore g 1 Hg ker = H. Hence g 1 Hg H and ker / G.
Example 1.5.7. Let G be any group, Z(G) < G. In fact Z(G) / G and any
subgroup of Z(G) is normal in G. To see this, let H Z(G).
To show that H / G we need to show that g 1 hg H for all g G, h H, but
since h H Z(G) we have hg = gh, for all g G and g 1 hg = g 1 gh = h H.
A corollary of this is
Lemma 1.5.8. If G is abelian then every subgroup of G is normal.
Proof. If G is abelian, then Z(G) = G and every subgroup of G is a subgroup of
Z(G), hence normal by the previous result.
Example 1.5.9. It can happen that G is not abelian, yet every subgroup is normal.
For example G = Q8 , the quaternions. The subgroups of Q8 are {1}, {1}, hii , hji , hki.
{1} and {1} are normal because theyre contained in Z(Q8 ). The other subgroups
are normal because they have index 2 in Q8 since
Lemma 1.5.10. Let H G. If [G : H] = 2, then H E G

22

CHAPTER 1. GROUPS

Proof. Exercise.
Example 1.5.11. G = Q8 , H = {1}, H / G and G/H is a group of order
|G|
= 4. Is it Z4 or V4 (kleins 4 group)?
[G : H] = |H|
(Hi)2 = Hi2 = H(1) = H = eG/H .
Similarly (Hj)2 = H, (Hk)2 = H, therefore every element in G/H = {H, Hi, Hj, Hk}
has order 2 and Q8 /{1}
= V4 .
Example 1.5.12. G = (Z, +), H = nZ. Then G/H has n elements, namely
H + 0, H + 1, . . . , H + (n 1), which is Z/nZ.

1.6

Homomorphisms and Normal subgroups

In this section well extablish a connection between the seemingly unrelated concepts homomorphisms and normal subgroups. Weve seen that the kernel of a
homomorphism is a normal subgroup. More precisely
Proposition 1.6.1. Let : G G0 be a group homomorphism. Then
ker = {g G; (g) = eG0 }
is a normal subgroup of G.
In fact, normal subgroups are precisely the same as the kernels of homomorphisms.
Namely
Proposition 1.6.2. A subgroup N of G is normal if and only if it is the kernel
of some homomorphism.
Proof. : is Proposition 1.6.1.
: Let N / G. Then G/N is a group. Define a map
: G G/N, g 7 gN, g G.
Then by definition of operation in G/N
(g1 g2 ) = g1 g2 N = (g1 N )(g2 N ) = (g1 )(g2 )
and is a homomorphism,
ker = {g G; (g) = eG/N = N } = {g G; gN = N }
= {g G; g N } = N,
as wanted.

1.6. HOMOMORPHISMS AND NORMAL SUBGROUPS

23

The group G/N has the following universal property


Theorem 1.6.3. Let : G G0 be a group homomorphism with H ker .
Then there exists a unique homomorphism : G/H G0 such that = ,
i.e. the following diagram commutes
G
&

G0
%

G/H
Moreover if H = ker then is injective.
Proof. We define a map : G/H G0 , gH 7 (g). We first check that it is
well-defined. Suppose gH = g 0 H then g 0 = gh for some h H and
(g 0 ) = (gh) = (g)(h) = (g).
Since by assumption h ker we have (h) = e and therefore is well-defined.
(Note clearly that , has the same image.)
From its definition it is clear that = . The fact that is a homomorphism follows easily since is a homomorphism. Finally if H = ker then is
injective since if (g1 H) = (g2 H), then (g1 ) = (g2 ) and (g1 g21 ) = e, hence
g1 g21 ker = H and g1 H = g2 H.
Corollary 1.6.4 (Fundamental Theorem of Isomorphisms). If : G G0 is a
homomorphism, then
G/ ker
= im(),
the image of .
Proof. The in Theorem 1.6.3 is injective, clearly onto, im() = im().
Example 1.6.5. Let : Z Cn = hgi be a cyclic group of order n with a 7 g a .
(a + b) = g a+b = g a g b = (a)(b).
Clearly is surjective:
ker = {m Z; g m = 1} = {m Z; |g| |m}
= {m Z; n|m} = nZ
hence Z/nZ
= Cn . The fiber of over an element g a of Cn is
1 (g a ) = {m Z; g m = g a } = {m Z; g ma = 1}
= {m Z; n|m a} = {m Z; m a mod n} = a.

24

CHAPTER 1. GROUPS

Fibers of are presicely the residue classes modulo n.


nZ + 1 . . . nZ + a . . . nZ + (n 1)
1
a
n1
1n
an
n1n
1 2n . . . a 2n . . . (n 1) 2n
..
..
..
.
.
.

g0 = 1
g1
ga
g n1


a b
Example 1.6.6. Let : GL(1, R) (R\{0}, ),
7 ad bc. is onto


 c d
r 0
r 0
since for any r 6= 0, g =
GL2 (R) and
=r
0 1
0 1
nZ
0
n
2n
..
.

ker = {g GL(2, R); detg = 1} = SL(2, R)


therefore GL(2, R)/SL2 (R)
= (R\{0}, ).
Example 1.6.7. Let G = (R, +), S 1 = ({z C {0}; |z| = 1}, ), i.e. S 1 is
multiplicative group of complex numbers of absolute value 1. Let
: R S 1 , r 7 e2i = cos r + i sin r
then
ker = {r R; e2ir = 1} = Z,
so R/Z
= S 1.
Remark. (1) N / G. The elements of the quotient group are subsets of the original
group G. These subsets are the fibers or cosets of the kernel N of the homomorphism : G G/N . N and its cosets are projected (or collapsed) onto single
elements in the quotient group.
(2) N / G. Note that the structure of G is reflected in the structure of G/N , for
example the associativity of multiplication in G/N is indeed from associativity of
G inverses in G/N are indeed from inverses in G.
(3) We have seen before G acting on the set of subsets of G by conjugation.
G P(G) P, (g, A) 7 gAg 1 .
Let H G, H P(G). The stabilizer of H, under this action
GH = {g G; gHg 1 H},

1.6. HOMOMORPHISMS AND NORMAL SUBGROUPS

25

was called the normalizer of H in G and denoted by NG (H). Since stabilizers of


any action are subgroups of G we have NG (H) G.
NG (H) gives a criterion which determines precisely when a subgroup H is normal, namely when NG (H) = G. NG (H) is a measure of how close H to being
normal.
(4) Note being normal depends on the relation of H to G, not on the internal
structure of H. The same group H may be normal subgroup of G but not normal
in a larger group containing G.
Example 1.6.8. Take H = {e, (12)(34)}, G = {e, (12)(34), (13)(24), (14)(23)}, K =
S4 . Then H / G, but H C
6 S4 .
Theorem 1.6.9 (2nd Isomorphism Theorem). Suppose H, K are normal subgroups
of G and K H. Then K is normal in H and we have
G/K
= G/H.
H/K
Proof. If K H / G then clearly K / H. Let h H, since h G and K / G.
hKh1 K K / H. Well apply the fundamental theorem. Define a map
: G/K G/H, gK 7 gH.
is well-defined, since if g1 K = g2 K then g11 g2 K H hence g1 H = g2 H, i.e.
(g1 K) = (g2 K). is a homomorphism since
((g1 K)(g2 K)) = (g1 g2 K) = g1 g2 H = (g1 H)(g2 H) = (g1 K)(g2 K)
is onto since if gH G/H then we have gK G/K and (gK) = gH. We
really only have to check that ker = H/K. But
gK ker (gK) = eG/H = H gH = H g H gK H/K,
so ker = H/K and

G/K
H/K

= G/H.

Example 1.6.10. G = (Z, +), H = 2Z, K = 6Z, then 6Z 2Z Z and


Z/6Z
= Z/2Z.
2Z/6Z
Theorem 1.6.11 (3rd Isomorphism theorem). Let H, K be subgroups of G. Assume K / G. Then
H/H K
= HK/K,

26

CHAPTER 1. GROUPS

where HK = {hk; h H, k K}.


G
|
HK


K



H K
|
e

Proof. First note that everything in the conclusion makes sense.


Lemma 1.6.12. If H, K G, K / G then HK G.
Proof. Exercise.
K / HK follows from K / G and H K / H follows from simple.
Lemma 1.6.13. H G, K / G, K / G then H K / H.
Proof. Exercise.
Well prove the theorem by using the fundamental theorem, i.e. well define a
homomorphism : H HK/K and show that its kernel is H K. Let
: H HK/K, h 7 Kh
(note Kh HK/K since h HK). is a homomorphism since
(h1 h2 ) = Kh1 h2 (Kh1 )(Kh2 ) = (h1 )(h2 ).
is onto since any element HK/K has the form K(hk) for some h H, k H
and
K(hk) = KH Kk = Kh K = Kh.
Thus K(hK) = (h), hence the fundamental theorem gives H/ ker
= HK/K.
Now to show ker = H K assume x H then
x ker (x) = eHK/K = K (x) = K Kx = K x K x H K
and were done.
The next theorem describes the relation between the lattice of subgroups of the
quotient group G/N and the lattice of subgroups of G. There is a one-to-one
correspondence between the subgroups K of G containing N and the subgroups
of G/N .

1.7. GROUPS ACTING ON THEMSELVES BY CONJUGATION - CLASS EQUATION27


Theorem 1.6.14. Let G be a group, N / G. Then there is a bijection from {A
G; N A} onto the set of subgroups {A = A/N G/N }. In particular every
subgroup of G/N is of the form A/N for some subgroup A of G containing N . This
bijection has the following properties for all A, B G with N A and N B
(1) A B iff A B,


(2) If A B then |B : A| = B : A ,
(3) A E G iff A / G.
Proof. Exercise.
Example 1.6.15. Let G = Q8 , N =< 1 >= {1}.
Q8
|

hii
hji
hki

|

, Q8 /{1}
= V4 ,
h1i
|
e


V4 = {e, a, b, c}

|

hai
hbi
hci

|

e

The lattice for G/N appears at the top of the lattice for G.
Next we go back to group actions. We have seen that if G acts on a set X. The
relation defined by a b iff a = g b for some g G is an equivalence relation.
Lemma 1.6.16. The number of elements in the equivalence class containing a is
[G : Ga ], index of stabilizer if a G.
Proof. Let Ga = be the orbit of a, G/Ga the set of cosets of Ga in G,
: Ga G/Ga , b = g a 7 gGa.
is surjective since for any g G, g a Ga. is injective since
g a = h a h1 g Ga hGa = gGa ,
hence is a bijection.

1.7

Groups acting on themselves by conjugation


- class equation

Consider a function G G G, (g, a) 7 gag 1 . The equivalence classes under


this action are called conjugacy classes. Let a G be any element of G. Then
|Ga| denotes the number of conjugators of G
|Ga| = #{gag 1 = a} = CG (a)

28

CHAPTER 1. GROUPS

is called the centralizer of a in G. Then by Lemma 1.6.16 we have


|Ga| = [G : CG (a)].
Note if a Z(G) and
CG (a) = {g G; gag 1 = a} = G,
since a commutes with every g G, the conjugacy class of a has [G : CG (a)] =
[G : G] = 1 element.
The converse is also true if
|Ga| = 1 1 = [G : CG (a)] G = CG (a) a commutes g G a Z(G).
Since conjugacy classes, being equivalence classes, are disjoint, we have
X
G = conjugacy classes, |G| =
|conjugacy classes| .
We have
Theorem 1.7.1 (class equation). Let G be a finite group. Let g1 , g2 , . . . gr be
representatives of distinct conjugacy classes of G not contained in the center Z(G)
of G. Then
r
X
|G| = |Z(G)| +
[G : CG (gi )].
i=1

Proof. {a} is conjugacy class of size 1 if and only if a Z(G). Let Z(G) =
{1, z1 , . . . , zm }. Let K1 , K2 , . . . , Kr be conjugacy classes of G not contained in the
center. Let gi be a representative of Ki , i = 1, . . . , r. Then the conjugacy classes
of G are
1, {z1 }, {z2 }, . . . , {zm }, K1 , K2 , . . . , Kr .
This partitions G into disjoint sets and we have
|G| =

m
X
i=1

1+

r
X
i=1

r
X
|Ki | = |Z(G)| +
[G : CG (gi )].
i=1

Note that all the summands on the righthandside of class equation are divisors
of |G|. (Since they are indices of subgroups of G.) This restricts their possible
values.

1.7. GROUPS ACTING ON THEMSELVES BY CONJUGATION - CLASS EQUATION29


Example 1.7.2. S3 = {e, r, r2 , s, rs, r2 s}, r = (123), s = (23). Check that
Z(S3 ) = {e}, class of r : {r, r2 }, class of s : {s, rs, r2 s}
S3 = {e} {r, r2 } {s, rs, r2 s},
G = |S3 | = 1 + 2 + 3, note 2 | 6, 3 | 6, 1 | 6.
Example 1.7.3. G = Q8 . Note: In any group hgi CG (g). This helps to
minimize the computations of conjugacy classes.
In Q8 we see < i > CQ8 (i) Q8 . Since i
/ Z(Q8 ) = {1} and [Q8 :< i >] = 2,
we must have CQ8 (i) =< i > (|CQ8 (i)| |8, so it can be 1,2,4, since i
/ Z(G),
CQ8 (i) 6= Q8 ). Therefore i has precisely two conjugates in Q8 , i, i = kik 1 .
The other conjugacy classes are {1}, {1}, {i}, {j},{k}, note 8 = |Q8 | =
2 + 2 + 2 + 2.
Some applications of class equation
Groups of prime power order have nontrivial centers.
Theorem 1.7.4. Let p be a prime, P a group of prime power order p for some
1. Then P has a nontivial center, i.e. Z(G) 6= {e}.
Proof. Because of the class equation we have
r
X
|P | = |Z(G)| +
[P : CP (gi )],
i=1

where g1 , . . . , gr are representatives of distinct non-central conjugacy classes. By


definition CP (gi ) 6= P, i = 1, . . . , r,
[P : CP (gi )]| |P | = p ,
and it is not 1, hence p | [P : CP (gi )], i = 1, . . . , r, since p | |P | we have p | |Z(G)|
and therefore |Z(G)| =
6 1, so Z(G) 6= {e}.
Corollary 1.7.5. If |P | = p2 , then P is abelian. More precisely P is isomorphic
to Zp2 or Zp Zp .
Proof. Exercise.
In the next section, well prove a partial converse theorem to Lagranges theorem.
Namely if G is a finite group of order pa m, p - m, then G has a subgroup of order
pa .

30

1.8

CHAPTER 1. GROUPS

Composition series and the H


older program

Weve seen Lagrages theorem, which says if G is a finite group, H G, then


|H| | |G|. We already mentioned that the converse of thes statement is false.
There are groups e.g. symmetries of regular tetrahedron which has order 12, but
does not have a subgroup of order 6. Here are partial converses of Lagranges
theorem
Theorem 1.8.1 (Cauchys Theorem). If G is a finite group, p prime dividing |G|
then G has an element of order p.
Strongest converse theorem to Lagranges theorem is
Theorem 1.8.2 (Sylow). If G is a finite group of order p m, where p is prime,
(p, m) = 1, then G has a subgroup of order p .
(Cauchys theorem follows from Sylows theorem.) Well prove Cauchys theorem
for finite abelian groups without using Sylows theorems and in fact then well
use Cauchys theorem for finte abelian groups for the proof of Sylows theorems.
The proof of Cauchys theorem for finite abelian groups is also important in terms
of its structure which demonstates an important technique in finite group theory,
induction. The proof shows how the information on a normal subgroup N and on
G/N can be used to get information about G.
Theorem 1.8.3. Let G be a finite abelian group and p prime dividing |G|. Then
G has an element (hence a subgroup) of order p.
Proof. We use induction on |G|. Assume the theorem is valid for all abelian groups
of order < |G|. Let x 6= e be an element of G and H = hxi. If H = G then G is
cyclic and were done, since we know the theorem holds for cyclic groups.
Otherwise H G and H / G, since G is abelian. |G| = |H| |G/H| and both
factors on the right, |H| , |G/H| < G. Since p | |G| and p is prime then either
p | |H| or p | |G/H|.
If p | |H| and H is a group whose order is strictly less then |G|. By induction
hypohesis it has a subgroup of order p which is of course also a subgroup of G of
order p and were done.
So assume p | |G/H|. Now since G is abelian all factor groups of G are also
abelian. Hence G/H is an Abelian group of order < |G|. By induction hypothesis
G/H has a subgroup of order p. Since any group of order p is cyclic, this subgroup
/ H
must be cyclic, so it is hHgi for some g = Hg G/H of order p. Note g


1.8. COMPOSITION SERIES AND THE HOLDER
PROGRAM

31

since |Hg| 6= 1, but g p H since (Hg)p = eG/H = H, therefore hg p i 6= hgi and


ord(g)
, hence (ord(g), p) > 1.
ord(g p ) < ord(g), but ord(g p ) = (ord(g),p)
Since p is prime, p | ord(g), say (ord(g)) = kp, ord(g k ) =
is the subgroup were looking for.

kp
(kp,k)


= p and g k

The proof worked because both N , G/N are subgroups of order smaller than G,
and we could piece each information together. Clearly a basic obstruction to this
method is the non-existence of a normal subgroup. In the Theorem 1.9.9 we
could find a normal subgroup, because G was abelian. Groups without non-trivial
normal subgroups are fundamental obstruction to this method.
Definition 1.8.4. A group G is called simple if |G| > 1 (can be ) and the only
normal subgroups of G are 1 and G.
We have seen that if |G| is prime then G
= Zp , Zp has no non-trivial subgroups,
therefore they are simple for all primes p. The smallest non-abelian simple group is
A5 of order 60. Simple groups cannot be factored as N , G/N and can be thought of
building blocks as the primes in Z. There is also a unique factorization theorem.
Definition 1.8.5. In a group G, a sequence of subgroups
1 = N0 N1 N2 Nk1 Nk = G
is called a composition series if Ni E Ni+1 and Ni+1 /Ni are simple, 0 i k 1.
The quotient groups Ni+1 /Ni are called composition factors.
Note were not assuming Ni E G, just Ni E Ni+1 .
Example 1.8.6. 1 E hsi E hs, r2 i / D8 or 1 E hr2 i E hri E D8 . 2 different
composition series, 3 compsition factors each of which isomorphic to (simple) Z2 ,
hsi /1
= Z2 , hs, r2 i / hsi
= Z2 , D8 / hs, r2 i
= Z2 .
Theorem 1.8.7 (Jordan-Holder Theorem). Let G 6= 1 be a finite group. Then
(1) G has a composition series (not unique),
(2) The composition factors in a composition series are unique, i.e. if
1 = N0 E N1 E E Nr = G and 1 = M0 E M1 E E Ms = G,
then r = s and there exists some permutation of {1, . . . , r} such that
M(i) /M(i)1
= Ni /Ni1 , 1 i r.
Proof. Exercise.

32

CHAPTER 1. GROUPS

Remark. (1) Every group has a factorization. Even though composition series
is not unique the number and isomorphism type of composition factors are unique.
(2) Note that non-isomorphic groups might have the same (up to isomorphism)
list of composition factors, e.g. 1 E {1}/ < i > /Q8 .
Composition factors: Q8 / hii
= Z2 , hii /{1}
= Z2 , {1}/{1}
= Z2 , the same as

D8 , but Q8 6= D8 .
H
older program is a two part-program for classifying all finite groups up to
isomorphism
(1) Classify all finite simple groups.
(2) Find all ways of putting simple groups together to form other groups.
The classification of all finite simple groups, i.e. part (1) was completed in 1980.
Theorem 1.8.8. There is a list of 18 (infinite) families of simple groups and 26
simple groups not belonging to these families (the sporadic simple groups) such that
every finite simple group is isomorphic to one of the groups in the list.
A big theorem in this classification (whose proof 255 pages).
Theorem 1.8.9 (Feit-Thompson). If G is simple group of odd order, then G
= Zp
for some prime p.
Part (2) is very difficult, sometimes called extention problem. A class of groups
which appear in the theory of polynomial equations is solvable groups.
Definition 1.8.10. G is called solvable if there exists a chain of subgroups
1 = G0 E G1 E G1 E E Gs = G
such that Gi+1 /Gi is abelian for i = 0, 1, . . . , s 1.
(Finite solvable groups are precisely those groups whose composition factors are
all of prime order.) Finite solvable groups satisfy the following generalization of
Sylows Theorem.
Theorem 1.8.11 (Hall). Let G be a finite solvable group, then for every divisor
) = 1, G has a subgroup of order n.
n of |G| such that (n, |G|
n

1.9

Sylows Theorems

Definition 1.9.1. Let G be a group, p prime. A group of order p for some 1


is called a p-group. Subgroups of G which are p-groups are called p-subgroups. If
G is a group of order p m where p - m, then a subgroup of order p is called a
Sylow p-subgroup of G.

1.9. SYLOWS THEOREMS

33

Theorem 1.9.2 (1st Sylow Theorem). Let G be a finite group, p prime. If pk | |G|
then G has a subgroup of order pk . In particular G has a p-Sylow subgroup.
Theorem 1.9.3 (2nd Sylow Theorem). If P is a Sylow p-subgroup of G and Q
is any p-subgroup of G, then there exists g G such thath Q gP g 1 , i.e. Q is
contained in some conjugate of P . In particular, any two Sylow p-subgroups of G
are conjugate in G.
Theorem 1.9.4 (3rd Sylow Theorem). The number of Sylow p-subgroups of G,
called np , is of the form 1 + kp, i.e. np 1 ( mod p). Further np (G) = [G :
NG (P )], hence np | m.
For the proof of Sylows theorems recall following lemma and corollary
Lemma 1.9.5. If G acts on a set X. Let a X, a = Ga = {ga; g G} orbit of
a. Then |a| = [G : Ga ] where Ga = {g G; g a = a} stabilizer of a in G.
Corollary 1.9.6. The number of conjugates of a set S in a group G is the index
of normalizer of S, [G : NG (S)]. In particular the number of conjugates of an
element s of G is [G : CG (S)].
We first prove the existence of Sylow p-subgroups
Proof of Theorem 1.9.2. We proceed by induction on |G|. If |G| = 2, the result
is trivial. Assume the statement is true for all groups of order less than |G|, and
suppose pk | |G|. If G has a proper subgroup H whose index is not divisible by
p, then pk divides |H| so by induction hypothesis H has a subgroup of order pk ,
which is of course also a subgroup of G. Thus we may assume p | [G : H] for every
proper subgroup H of G. Recall the class equation
|G| = |Z(G)| +

r
X

[G : CG (gi )].

i=1

Since p | |G| , p | [G : CG (gi )] we know that p | Z(G). Then by Cauchys Theorem


for abelian groups Z(G) has a subgroup N of order p. Now since N < Z(G), N /G,
G/N is a group of order |G| /p. If pk divides |G| then pk1 | |G/N |, so by the
induction hypothesis G/N has a subgroup P of order pk1 . If we let P be the
subgroup of G containing N such that P/N = P then

P = |P/N | |P | = pk1 |N | = pk1 p = pk .

34

CHAPTER 1. GROUPS

Before proving the 2nd and 3rd Sylow Theorems we overse some facts; By the 1st
Sylow Theorem, we know that there exists P Sylp (G). Let
X = {P1 , P2 , . . . , Pr } = {gP g 1 ; g G},
be the set of all conjugates of P . (Note G acts on X, G has only one orbit.) Let Q
be any p-subgroup of G. Q also acts on X by conjugation. Write X as a disjoint
union of orbits under this action
X = O1 O2 Os , where r = |O1 | + + |Os | .
Note r does not depend on Q, but the number of Q-orbits s does. Renumber
the elements of X if necessary such that the Q orbits = Pi Oi , 1 i s.
|Oi | = [Q : NQ (Pi )] by Lemma 1.9.5. By definition
NQ (Pi ) = NG (Pi ) Q = Pi Q,
because of the following Lemma
Lemma 1.9.7. Let P Sylp (G). If Q is any p-subgroup of G. Then
Q NG (P ) = Q P.
Using these
()

|Oi | = [Q : NG (Pi )] = [Q : NG (Pi ) Q] and |Oi | = [Q : Pi Q], 1 i s.

Since Q was arbitrary, we may take it as Q = P1 , |O1 | = 1. Now, for all i > 1, P1 6=
Pi such that P1 Pi < P1 . By () we have |Oi | = [P1 : P1 Pi ] > 1, 2 i s.
Since P1 is a p-group, [P1 : P1 Pi ] is a power of p, hence P | Oi , 2 i s. Thus
r = |O1 | + (|O2 | + + |Os |) 1 mod p.
Now we prove Theorem 1.9.3 and Theorem 1.9.4.
Proof of Theorem 1.9.3. Let Q be any p-subgroup of G. Suppose Q is not contained in Pi for any i {1, . . . , r}, Q 6E gP g 1 for any g G. Then QPi < Q i,
hence by (*) we have |Oi | = [Q : Q Pi ] > 1, 1 i s. Thus p | |Oi | , i, so
p divides |O1 | + |O2 | + + |Os | = r. But this contradicts r 1 mod p (r does
not depend on choice of Q). This proves Q gP g 1 for some g G.
To see that all Sylow p-subgroups of G are conjugate, let Q be any Sylow psubgroup of G, by preceding argument Q < gP g 1 for some g but |Q| = |P | =
|gP g 1 | = p we must have gP g 1 = Q. This proves Theorem 1.9.3.

1.9. SYLOWS THEOREMS

35

Proof of Theorem 1.9.4. Finally since all sylow p-subgroups are conjugate, X =
{all Sylow p-subgroups}. G acts on X by conjugation, P Sylp (G), orbit of P
under G is all of X. |X| = np = [G : NG (P )] for any P Sylp (G). We have
already seen that np 1 mod p.
Corollary 1.9.8. Let P be a Sylow p-subgroup of G. Then there holds equivalence
(1) P is the unique Sylow p-subgroup of G, i.e. np = 1,
(2) P / G,
(3) p = char(G).
Examples of some applications of Sylow Theorems
Remark. No group of order 20 is simple.
Proof. If |G| = 20 = 4 5, G has Sylow-5 subgroups n5 1 mod 5 and n5 | 20.
Note in general np | [G : NG (P )], hence
np | [G : NG (P )] [NG (P ) : P ] = [G : P ] =

mp
= m.
p

so in fact n5 | 4. This implies n5 = 1, hence the Sylow-p is unique and therefore


normal in G.
The following simple theorem isolates conditions under which a group G is isomorphic to the direct product of two of its subgroups and is very useful in classification
theorems.
Theorem 1.9.9. Let A, B G such that
(1) A / G, B / G,
(2) AB = G,
(3) A B = {e},
then G
= A B.
Observe (1)-(3) imply two more properties of the subgroups A and B
(4) If ab = a1 b1 , where a, a1 A, b, b1 B, then a = a1 , b = b1 ,
(5) If a A, b B then ab = ba.
Observe that (2) says that every element of G can be written as ab for some
a A, b B, (4) says this representation is unique.
Proof. To see (4) note
b1 b1 ,
ab = a1 b1 a1
1 a = |
{z }
|{z}
A

36

CHAPTER 1. GROUPS

1
1
hence a1
A B = {e}, a1
and a1 = a, b1 = b. To see (5)
1 , b1 b
1 a = e = b1 b
let
1
1
1 1
bab1 a1 = ( bab
| {z } )a = b( ab
| {za } ),
A,since A/G

B, since B/G

therefore bab1 a1 A B = {e} and ab = ba.


To prove the Theorem define : A B G, (a, b) 7 ab with
(5)

((a, b)(a1 , b1 )) = (aa1 , bb1 ) = aa1 bb1 = aba1 b1 = (a, b)(a1 , b1 ),


so is a homomorphism. Clearly it is onto by assumtion (2). We have
ker = {(a, b) A B; (a, b) = ab = e},
hence (a, b) ker and a = b1 , a, b A B, therefore a = e, b = e and
(a, b) = (e, e) = eAB , and we have G
= A B.
Theorem 1.9.10. If p, q are prime, p < q then every group G of order pq has a
single subgroup of order q, hence it is normal and G is not simple. Moreover if
q 6 1 mod p then G is abelian (and in fact cyclic).
Proof. G has a Sylow-q subgroup Q. The number of such subgroups nq 1
mod q and nq | p, but p < q hence nq = 1 and Q / G. Hence G is not simple.
Similarly there exists P Sylp (G), np 1 mod p, np | q then np = 1 or q.
Hence if q 6 1 mod p then np = 1, P / G.
Every element 6= e of Q has order q (exept e) and every element 6= e of P has
order p, hence Q P = {e}.
QP is a subgroup of G properly containing Q of order dividing pq, hence QP = G.
By previous theorem G
= QP
= Zp Zp . Hence G is abelian, in fact cyclic.
Example 1.9.11. Let G be a group of order 12. Then G either has a normal
Sylow-3 subgroup or G
= A4 .
Proof. Suppose n3 6= 1 and P Syl3 (G). n3 | 4, n3 1 mod 3, hence n3 = 4.
Since distinct Sylow-3 subgroups intersect in the identity, each Sylow-3 subgroup
contains 2 elements of order 3 and G contains 4 2 = 8 elements of order 3. Since
[G : NG (P )] = n3 = 4 and P NG (P ) we have NG (P ) = P . Let X be the set of
Sylow-3 subgroups of G, |X| = 4.

1.10. DIRECT PRODUCTS AND ABELIAN GROUPS

37

Now G acts on X by conjugation. The permutation representation of this action gives a homomorphism
: G SX = S4 , g 7 (g) = P gP g 1 , P X.
ker = {g G; (g) = identity} = {g G; gPi g 1 = Pi , i = 1, 2, 3, 4}.
In particular ker NG (P ) = P . Since P 6C G, ker / G, ker = 1, hence
G
= (G) S4 .
G contains 8 elements of order 3 and these are precisely 8 elements of order 3
in S4 . All these elements (of order 3) are contained in A4 . Hence (G) A4
has order at least 8. But (G) A4 is a subgroup of A4 , which has order
|A4 | = 12, |(G) A4 | > 8 and divides 12, implies that |(G) A4 | = 12 and
hence (G) = A4 and G
= A4 .

1.10

Direct products and Abelian groups

Weve seen before that given two groups G1 , G2 we can form another group, namely
G1 G2 , the direct product of G1 and G2 by defining the group operation componentwise. In general we have
Definition 1.10.1. The direct product G1 G2 Gn of the groups (G1 , 1 ),
(G2 , 2 ), . . . , (Gn , n ) is the set of n-tuples (g1 , . . . , gn ), gi Gi with operation
defined as
(g1 , g2 , . . . , gn ) (h1 , . . . , hn ) := (g1 1 h1 , g2 2 h2 , . . . , gn n hn ).
Typically even though the operation may be different in each of the factors of a
direct product we drop subscrips in i s and simple write
(g1 , g2 , . . . , gn ) (h1 , . . . , hn ) := (g1 h1 , g2 h2 , . . . , gn hn ).
When the groups are abelian we write the operation additively and call the group
n
L
direct sum of groups G1 , G2 , . . . , Gn ,
Gi .
i=1

Direct products, aside from providing us with an easy way of building new groups
they often also enable us to understand a given group better. This happens when
we are able to realize that the given group is isomorphic to the direct product of
some of its subgroups. In this way, we can break the group down into simpler
components that are easier to deal with. We have already seen in Thm 1.9.9, the
conditions under which a group G is isomorphic to the direct product of two of its

38

CHAPTER 1. GROUPS

subgroups. The clue really comes from H K itself.


Although for given groups H, K, they are not subgroups of H K, there are subgroups H , and K of H K such that H
= K, thus H K
= H K .
= H, K
Namely take H = H {eK }, K = {eH } K. What we can say about H , K ?
(1) H , K are normal in H K
(h, k)(eH , x)(h, k)1 = (heH h1 , kxk 1 ) = (eH , kxk 1 ) {eH } K,
or
1 : H K H, (h, k) 7 h,

2 : H K K, (h, k) 7 k

are surjective homomorphisms whose kernels are


ker 1 = {eH } K = K ,

ker 2 = H {eK } = H

(2) H K = H K because any element (h, k) H K can be written as


(h, eK )(eH , k) with (h, eK ) H , (eH , k) K .
(3) Finally H K = (eH , eK ) = eHK .
And as Theorem 1.9.9 shows these 3 indicated properties captive the essense of
the direct products.
Example 1.10.2. G = V = {e, a, b, c}, H = {e, a}, K = {e, b}. Then H, K /
G, HK = {e, a, b, ab} = {e, a, b, c} = V, H K = {e}, V
=H K
= Z2 Z2 .
Example 1.10.3. Let G = hxi cyclic group of order mn where (m, n) = 1. Let
H = hxn i , K = hxm i. H, K are normal subgroups of order m, n resp.
ord(xn ) =

nm
nm
ord x
=
=
= m.
(ord x, n)
(m, n)
n

The order of any element in H K must divide |H| and |K|, hence must divide
m, n but (m, n) = 1, so H K = {e}.
|HK| =

|H| |K|
mn
=
= mn,
|H K|
1

hence HK = G. Thus G = H K, i.e. Zmn


= Zm Zn .
In general we have
Proposition 1.10.4. Let G1 , G2 , . . . Gn be groups and let G = G1 Gn be
their direct product.
(1) For fixed i, the set {(1, 1, . . . , 1, gi , 1, . . . , 1); gi Gi } is a subgroup of G
isomorphic to Gi , Gi
= {(1, . . . , 1, gi , 1, . . . , 1); gi G}. If we identify Gi with
this subgroup, then Gi E G and G/Gi
= G1 Gi1 Gi+1 Gn .

1.10. DIRECT PRODUCTS AND ABELIAN GROUPS


(2) For each fixed i define i :
surjective homomorphism with

39

G Gi , (g1 , g2 , . . . , gn ) 7 gi , then i is a

ker i = {g1 , . . . , gi1 , 1, gi+1 , . . . , gn ; gj Gj i 6= j}

= G1 Gi1 Gi+1 Gn .
(3) Under the identification in part (1), if x Gi , y Gj for some i 6= j, then
xy = yx.
Example 1.10.5. Let p be a prime, n Z+ . Consider Epn = Zp Zp Zp .
Then Epn is an abelian group of order pn . Every element x of Epn satisfy xp = 1.
Epn is called elementary abelian group of order pn .
If n = 2, Ep2 = Zp Zp is a group of order p2 . It has exactly p + 1 subgroups
of order p (There are more than the two obvious ones). Since every non-identity
element of Ep2 has order p, each of these elements genereate a cyclic subgroup of
order p. Since distinct subgoups of order p intersect trivially, the p2 1 non-identity
elements are partitioned into subsets of size p1, i.e. each of these subsets consists
of the non-identity elements of some subgroup of order p. Hence there must be
p2 1
= p + 1 subgroups of order p.
p1
V4
= Z2 Z2 = {(0, 0), (0, 1), (1, 0), (1, 1)} has 3 subgroups:
{(0, 0), (0, 1)}, {(0, 0), (1, 0)}, {(0, 0), (1, 1)} of order 2.

1.10.1

The fundamental theorem of finitely generated abelian


groups

Definition 1.10.6. A group G is said to be finitely generated if there is a finite


set A of G such that G = hAi = the smallest subgroup of G containing {a; a A}.
0
r
For each r Z, r 0, let Zr = Z
| Z
{z Z} (we take Z = 1). The group Z
r copies

is called the free abelian group of rank r.


Note (1) any finite group is finitely generated. Take A = G,
(2) Zr is finitely generated, generated by A = {(0, 0, . . . , 0, 1, 0, . . . , 0) = ei ; 1
i r}.
Theorem 1.10.7 (Fundamental Theorem of Finitely Generated Abelian groups).
Let G be a finitely generated abelian group. Then
G
= Zr Zd1 Zd2 Zds

40

CHAPTER 1. GROUPS

for some integers r, d1 , d2 , . . . , ds such that r 0, di 2 i and di+1 | di for 1


i s1. Moreover the expression above is unique if G
= Z t Zn1 Zn2 Znk ,
where t, nj satisfy the conditions above (t 0, nj 2 j and nj+1 | nj for
1 j k 1) then t = r, s = k and di = ni i.
Definition 1.10.8. The integer r is called the rank of G, d1 , d2 , . . . ds are called
the invariant factors of G.
The Fundamental Theorem of Finitely Generated Abelian groups provides us with
a way of listing all finite abelian groups of a given order, i.e. we must find all finite
sequences of integers d1 , d2 , . . . , ds such that
(1) dj 2 j {1, . . . , s},
(2) dj+1 | dh , 1 j s 1,
(3) d1 d2 . . . ds = n.
The Theorem says that there exists a bijection between sequences d1 , d2 , . . . , ds
satisfying (1)-(3) and isomorphism classes of finite abelian groups of order n.
Example 1.10.9. If n = p1 p2 . . . pt where pi s are disinct then up to isomorphism
there is only one abelian group of order n, Zn .
Note that for any n, di s satisfy d1 d2 ds , i.e. d1 is the largest invariant factor, each di | n (property (3)). If p is a prime divisor of n, then p | di
for some i, by (2) p | dj j i, hence every prime divisor of n divides d1 .
In particular if n = p1 . . . pt , each pi | d1 , hence n = p1 . . . ps | d1 ,but d1 | n,
so n = d1 . There is only one invariant factor d1 = n.
Example 1.10.10. n = 180 = 22 32 5, 2 3 5 | d1 , hence the possible values of
d1 are d1 = 22 32 5, 22 3 5, 2 32 5, 2 3 5.
For each of these we have to work out all possible d2 . For each pair d1 , d2 we work
out all possible d3 , etc.
Abelian groups
Invariant factors
2
2
d1 = 2 3 5
Z180
2
d1 = 2 3 5, d2 = 3
Z60 Z3
d1 = 2 32 5, d2 = 2
Z90 Z2
d1 = 2 3 5, d2 = 2 3
Z30 Z6
The following theorem gives a much faster way of determining all finite abelian
groups. If n = p1 1 . . . pk k , then all possible lists of invariant factors of abelian
groups of order n can be determined by finding all possible lists for groups of
order pi for each i.
Theorem 1.10.11. Let G be an abelian group of order n. Let n = p1 1 . . . pk k , pi s
are distinct primes. Then

1.10. DIRECT PRODUCTS AND ABELIAN GROUPS

41

(1) G = A1 A2 Ak , where |Ai | = pi i ,


(2) For each A {A1 , . . . , Ak } with |A| = p
A
= Zp1 Zp2 Zpt ,
where 1 2 t 1, 1 + 2 + + t = ,
(3) The decomposition in (1),(2) is unique, i.e. if G = B1 Bm with |Bi | =
pi i , i, then Bi
= Ai and Bi , Ai have the same invariant factors.
Definition 1.10.12. The integers pi are called the elementary divisors of G. The
decomposition in (1) and (2) is called elementary divisor decomposition of G.
Example 1.10.13. 180 = 22 32 5. 22 : Z2 Z2 , Z4 , 32 : Z3 Z3 , Z9 , 5 : Z5
Z2 Z2 Z3 Z3 Z5
Z2 Z2 Z9 Z5
Z4 Z3 Z3 Z5
Z4 Z9 Z5

1.10.2

= Z30 Z6

= Z90 Z2

= Z60 Z3

= Z180

Semidirect products

Semidirect products is a generalization of the notion of direct products of H and


K. We relax the requirement that H and K are both normal. Well build a larger
group G from H and K such that G contains isomorphic copies of H and K, as
in the case of direct products except that H will be normal but not necessarly K.
This way we can construct a non-abelian group from abelian groups H, K.
Suppose we already have a group G containing subgroups H, K such that
(a) H / G (But K not necessarily normal)
(b) H K = 1
We still have HK G a subgroup. Every element of HK can be written uniquely
1
as a product hk, for some h H, k K (h1 k1 = h2 k2 h1
2 h1 = k2 k1 H
K = {e} h2 = h1 and k2 = k1 ), i.e. there is a bijection HK H K, hk 7
(h, k) and
(h1 k1 )(h2 k2 ) = h1 k1 h2 (k11 k1 )k2 = h1 (k1 h2 k11 ) k1 k2 ,
{z
} |{z}
|
h3

k3

since H / G. These calculations were done by the assumption that there already
existed a group G containing H, K with H E G, H K = 1.
The product k1 k2 is obtained from the multiplication in K, hence easy to understand. If we understand how the element k1 h2 k11 arises in terms of H and K

42

CHAPTER 1. GROUPS

without reference to G then the group HK will have been described entirely in
terms of H and K.
Since H / G, K acts on H by conjugation k h := khk 1 , (h1 , k1 )(h2 , k2 ) =
(h1 (k1 h2 ), k1 k2 ). The action of K on H gives a homomorphism : K
Aut(H) = SH , k 7 (k) : h 7 khk 1 . Multiplication in HK depends on
multiplication in K, multiplication in H and the homomorphism .
Theorem 1.10.14. Let H, K be groups, let be a homomorphism from K into
Aut(H). Let G be the set of ordered pairs (h, k) with h H, k K, define the
multiplication on G by
(h1 , k1 )(h2 , k2 ) := (h1 (k)(h), k1 k2 ).
Then G is a group and H
= {(h, 1); h H} = H , K
= {(1, k); k K} = K ,

moreover H E G, H K = {e}.
Proof. Exercise.
Definition 1.10.15. The group G is called the semidirect product of H and K
with respect to denoted by H o K. We mostly drop and write H o K when
there is no danger of confusion.
Example 1.10.16. Let H be any abelian group, K = hxi
= Z2 group of order
2. Define : K Aut(H), 1 7 identity map, x 7 (x) : h h1 . Then
G = H o K contains the subgroup H of index 2 and xhx1 = h1 , h H.
(1, x)(h, 1)(1, x1 ) = (1(x)(h), x1)(1, x1 ) = (h1 , x)(1, x1 )
= (h1 (x)(1), xx1 ) = (h1 11 , 1) = (h1 , 1).
If H = Zn , then G = H o K = Zn o Z2
= D2n . The diedral group of order n is
the group of symmetries of a regular n-gon.

1.11

Free groups

Let S be a set, {si }I , I not necessarily finite. We think S as an alphabet and si


are letters of the alphabet. Any symbol of the form sni , n Z is a syllable and a
finite string w of syllables written in juxtaposition is a word. The empty word 1
has no syllables.
Example 1.11.1. S = {s1 , s2 , s3 } then
1 11
3
3 1
s1 s1
3 , s 2 , s 2 s1 s3 s1 s1

are all words.

1.11. FREE GROUPS

43

There are two natural way of modification of certain words


m+n
(1) replacing an occurence sni sm
,
i by si
0
(2) Replacing the occurence of si by the word 1, i.e. dropping out the word.
These are the two elementary contractions. By a finite number of elementary
contractions every word can be changed to a reduced word, one for which no more
elementary contraction is possible.
5
2 2
2 7
Example 1.11.2. s21 s32 s1
2 s3 s1 s1 s1 s2 s3 s1

Definition 1.11.3. The set of all reduced words from an alphabet S is F (S), the
free group generated by S (S is free of relations).
Theorem 1.11.4. F (S) is a group under juxtaposition.
F (S) has the important universal property that any map from the set S to a
group G can be uniquely extented to a homomorphism from the group F (S) to G.
Theorem 1.11.5. Let G be a group, S a set and f : S G a set map. Then
there exists a unique group homomorphism : F (S) G such that the following
diagram commutes
S
, F (S)
.
f &
G
Definition 1.11.6. The cardinality of S is called the rank of the free group.
Theorem 1.11.7 (Schreier). Subgroups of a free group are free.
Warning: The rank of a subgroup of a free group can be greater than the rank
of the whole group.
Example 1.11.8. Let S = {x, y}, G = F (S), a free group on two letters. The
rank of G is 2. Let yk = xk yxk , k 0. Let H = F ({yk ; k 0}) the free group
generated by yk . Then H G, but rank(H) = !

44

CHAPTER 1. GROUPS

Chapter 2
Rings
2.1

Basic Definitions and examples

Definition 2.1.1. A ring R is a set together with two binary operation + and
(called addition and multiplication) satisfying the following axioms
(1) (R, +) is an abelian group
(2) is associative, i.e. (a b) c = a (b c), a, b, c R,
(3) The Distributive laws hold: (a + b) c = (a c) + (b c) and a (b + c) =
(a b) + (a c), a, b, c R.
R is called commutative if multiplication is commutative. R is said to have identity,
if there exists 1 R such that a 1 = 1 a = a a R. We write ab instead of
a b.
Example 2.1.2. (Z, +, ), (Q, +, ), (R, +, ) are all commutative rings with
1.
Example 2.1.3. (2Z, +, ) commutative ring, but without 1.

Example 2.1.4. R = {a + b 2; a, b Z} with the usual +, is a ring with 1.


Example 2.1.5. (Zn , , )
These are interesting because they already start to show behaviour different from
that of (Z, +, ), e.g. in Z6 , 2 3 = 0 even though 2 6= 0 and 3 6= 0 and in Z8 23 = 0.
Definition 2.1.6. Let R be a ring. An element a R is called a zero-divisor if
there exists b R, b 6= 0 such that ab = 0 or ba = 0.
a is called nilpotent if n such that an = 0.
At the opposite extreme from these badly behavioured elements are units.
45

46

CHAPTER 2. RINGS

Definition 2.1.7. Let R be a ring with 1. An element a R is called a unit if


there exists b R such that ab = ba = 1. The set of units in R is denoted by R .
It is not difficult to see that R is a group under multiplication, and hence is called
the group of units of R.
Definition 2.1.8. A ring R with 1, where 1 6= 0 is called a division ring (skew
field) if every non-zero element a R has a multiplicative inverse. A commutative
division ring is called a field.
Example 2.1.9. Trivial rings: R any commutative group, define a multiplication
on R via ab = 0, for all a, b R. With this multiplication R is a commutative
ring. If R = {0}, trivial rings do not have 1, i.e. R = {0} is the only ring 0 = 1.
Thats why we will mostly write 1 6= 0 to exclude this case. These rings give no
more information that could not be obtained from abelian group theory.
Example 2.1.10. (Real) Hamiltonian Quaternions.
H = {a + bi + cj + dk; a, b, c, d R} with
(a + bi + cj + dk) + (a0 + b0 i + c0 j + d0 k) = (a0 + a) + (b + b0 )i + (c + c0 )j + (d + d0 )k
multiplication is defined by expending using the distributive law and i2 = j 2 =
k 2 = 1 and ij = k, jk = i, ki = j. H is a division ring
(a + bi + cj + dk)1 =

a bi cj dk
.
a2 + b 2 + c 2 + d 2

Example 2.1.11. R = {f : [0, 1] R; f iscontinuous}, R = {f R; f (x) 6=


0 x [0, 1]}, f 1 = f1 . R contains many zero divisors. For example let
(
0,
0 x 12
f (x) =
x 21 , 21 x 1.
Let g(x) = f (1 x), then f g = 0.
On the other hand f (x) = x 12 is neither a unit nor a zero-divisor. It is not
a unit since it has a zero at x = 12 . If there exists g such that gf = 0, then g = 0
for all x 6= 21 , but g is continuous. Therefore g 0. Hence f is not a zero-divisor.
Example2.1.12.
 M2 (R) = 22 matrices with real entries, non-commutative

 ring
1 0
0 0
with 1 =
, then theres exists A, B M2 (R) such that AB =
6= BA,
0 1
0 0
namely




1 0
0 0
A=
, B=
.
1 0
0 1

2.1. BASIC DEFINITIONS AND EXAMPLES


This explains why in the definition 
of zero-divisors
2 
0 1
0
It also contains nilpotent elements
=
0 0
0

47
ithas ab = 0 or ba = 0.
0
.
0

Basic aritmetic holds for general rings. More precisely, we have


Proposition 2.1.13. Let R be a ring, a, b R. Then
(1) a 0 = 0 a = 0,
(2) a(b) = (a)b = (ab),
(3) (a)(b) = ab,
(4) if R has an identity, then identity is unique,
(5) m(ab) = (ma)b = a(mb) for any integer m.
Proof. (1) a 0 + a 0 = a(0 + 0) = a 0 a 0 = 0.
(2) a(b) = (ab) a(b) + ab = 0 a(b + b) = 0 a 0 = 0, etc.
Corollary 2.1.14. Let R be a ring with 1. Let u R be a unit. Then u is not a
zero-divisor in R.
Proof. If there exists r R such that ur = 0 or ru = 0 then we want to show that
r = 0. But if ur = 0 then r = (u1 u)r = u1 (ur) = u1 0 = 0.
Definition 2.1.15. A commutative ring with 1 6= 0 is called an integral domain
if it has no zero divisors.
Example 2.1.16. Z and every field F is an integral domain since every non-zero
element of F is a unit.
Example of fields: Q, R, C, (Zp , , ) p prime since any r {1, 2, . . . , p 1}
satisfies (r, p) = 1 and hence a unit in Zp . Since in general a Zn is a unit if
and only if (a, n) = 1. Every element of Zn is either a unit or a zero-divisor. If
(a, n) = d > 1 let 
b = nd then 0 < b < n. Hence b 6= 0 and by construction n | ab

since ab = a

n
d

a
d

n. Hence ab = 0 in Zn , i.e. a is a zero divisor.

The absence of zero-divisors in integral domains gives a cancellation property.


Proposition 2.1.17. Let a, b, c R, R any ring. Assume a is not a zero-divisor
If ab = ac then either a = 0 or b = c (i.e. if a 6= 0 we can cancel as). In particular
if R is an integral domain and ab = ac, then either a = 0 or b = c.
Proof. If ab = ac then a(b c) = 0 if a is not a zero divisor then either a = 0 or
b c = 0.

48

CHAPTER 2. RINGS

Proposition 2.1.18. Any finite integral domain is a field.


Proof. Let a R be a non-zero element of R. By cancellation law the map x ax
is injective. Since R is finite it is also surjective. Hence a is a unit, R is a field.
Definition 2.1.19. A subring of a ring R is a subgroup of R that is closed under
multiplication.
To show that a subset S of a ring R is a subring it is enough to check that S 6=
and it is closed under subtraction and under multiplication.
Example 2.1.20. 2Z is a subring of Z. Note even though Z is a ring with 1, 2Z
has no unity!
Example 2.1.21. R = R R, S = {(r, 0); r R}. S is a subring of R R, R R
has unity (1, 1)
/ S. S has its own unity, namely (1, 0).
Example 2.1.22. R = {f : R R} all real-valued funtions on R under pointwise addition and multiplication,
S = {f R; f is continuous}, subring of R, T = {f R; f (0) = 0}, subring of R, U = {f R; f (0) = 1}, not a subring of R: If f (0) = g(0) = 1,
(f g)(0) = 0, but f g
/ U.

Important examples
Example 2.1.23. Let R be a ring. Well take R commutative with 1, even though
this is not necessary. Let x be an indeterminate. Let
R[x] = {an xn + an1 xn1 + . . . + a0 x0 ; n 0, ai R}.
R[x] is the ring of polynomials in the variable x with coefficients in R. Addition,
multiplication are given by the familiar operation from elementary algebra.
(an xn + an1 xn1 + . . . + a0 x0 ) + (bn xn + bn1 xn1 + . . . + b0 x0 )
= (an + bn )xn + + (a0 + b0 )
(an xn + an1 xn1 + . . . + a0 x0 ) (bn xn + bn1 xn1 + . . . + b0 x0 )
l
X

= a0 b0 + (a0 b1 + b0 a1 )x + +
ak blk xl + . . . .
k=0

Polynomials g(x) = an xn + an1 xn1 + . . . + a0 x0 is said to be of degree n if


an 6= 0, an xn is called the leading term. H is called monic if an = 1.
R is a subring of R[x] (R= constant poynomials). The ring in which the coefficients are taken makes a big difference in the behaviour of polynomials, e.g.
x2 + 1 is not a perfect square in Z[x], but x2 + 1 = (x + 1)2 in Z/2Z[x].

2.2. IDEALS, RING HOMOMORPHISMS AND QUOTIENT RINGS

49

Proposition 2.1.24. Let R be an integral domain and p(x), q(x) R[x]. Then
(1) degree(p(x)q(x)) = degree(p(x)) + degree(q(x)),
(2) units of R[x] are just the units of R,
(3) R[x] is an integral domain.
Proof. If R has no zero-divisors, if p(x), q(x) has leading coefficient an xn , bm xm
resp. then the leading term of p(x)q(x) = an bm xn+m and an bm 6= 0. Hence R[x] is
an integral domain and (1) holds.
Finally to see (2): If p(x) is a unit then there exists q(x) such that p(x)q(x) =
1 hence degree(p(x)) + degree(q(x)) = 0 so deg(p(x)) = deg(q(x)) = 0 and
p(x), q(x) R, hence p(x), q(x) are units in R.
Example 2.1.25. Group Rings. Let R be a commuative ring with 1 6= 0, G a
finite group. The group ring of G, RG is defined as the set of all formal sums
a1 g1 + a2 g2 + + an gn , ai R, 1 i n, where G = {1 = g1 , . . . , gn }, ai
R, agi = ai RG. Similarly 1g RG, g G. Addition is defined componentwise. To define multiplication we first define (agi )(bgj ) = (ab)gk where gk = gi gj
the product in G. Then extend P
this multiplication
to all formal sums by distribuP
P
tive laws. Coefficients of gk in ( ai gi )( bj gj ) is
= gk ai bj .
gi gj

If RG is commutative then G is also commutative. If |G| > 1, RG has always


zero-devisors. If g G, ord(g) = m then (1 g)(1 + g + + g m1 ) = 1 g m = 0,
hence 1 g is a zero divisor. If S is a subgroup of R, SG is a subring of RG.

2.2

Ideals, Ring Homomorphisms and quotient


rings

In dealing with groups, we have seen that some subgroups are better than
others. For example if H < G, then G/H has a group structure iff H / G. We
encounter a similar situation with rings.
Let (R, +, ) be a ring, S a subring. Since (R, +) is abelian (S, +) is a normal subgroup, and (R/S, +) is a group whose elements are true cosets of S in R,
{a + S; a R} with addition defined as (a + S) + (b + S) = (a + b) + S.
We would like to define a multiplication on the quotient R/S such that (R/S, +, )
becomes a ring. There is a natural choice, namely (a + S)(b + S) := ab + S. We
need to check that this multiplication is well-defined, i.e. if a + S = a0 + S and
b + S = b0 + S, then ab + S = a0 b0 + S, i.e. we want to show if a a0 S, b b0 S

50

CHAPTER 2. RINGS

then ab a0 b0 S for any choice a, b, a0 , b0 R.


For example if we take a S, a0 = 0, b arbitrary and b = b0 then we have
a a0 S, b b0 = 0 S. We want ab 0 b S, i.e. we want if a S, b R,
then ab S. Similary we want if a R, b S then ab S. In fact these two
conditions are enough. To see this assume a a0 S, b b0 S. Then
ab a0 b0 = (a a0 )b + a0 (b b0 )
| {z } | {z }
S

by the two conditions on S, hence ab a0 b0 S and our multiplication is welldefined. Subrings that have the special properties required to make multiplication
of the additive cosets well-defined are called ideals.
Definition 2.2.1. A subring S of a ring R is called an ideal of R if for all s
S, r R, we have rs S, sr S.
Definition 2.2.2. Let R be a ring, I a subset of R. Let r R.
(1) rI = {ra; a I}, Ir = {ar; a I}
(2) A subset I of R is a left ideal of R if
(i) I is a subring of R
(ii) rK I, r R (closed under left multiplication). Similary for left ideals,
if Ir I, r R, i.e. I is an ideal (two-sided ideal) if it is both, a left and
right ideal.
Theorem 2.2.3. Let (R, +, ) be a ring, (I, +, ) be an ideal of R. Then the set
of cosets R/I is a ring under the operation (I + a) + (I + b) := I + (a + b), (I +
a)(I + b) := I + ab.
Proof. Exercise.
Definition 2.2.4. R/I is called the quotient ring or factor ring of R by I.
The next theorem gives an efficient characteritic of ideals.
Theorem 2.2.5. Let R be a ring, S be a subset. Then S is an ideal of R iff the
following holds
1) S is an additive subgroup of R ( S 6= and S is closed under subtraction),
2) For all r R, s S, we have rs S, sr S.
Proof. Exercise.
Example 2.2.6. (Z, +, ), (nZ, +, ) is an ideal since if we multiply an element
of mZ by an element of Z, we get an element of mZ.

2.2. IDEALS, RING HOMOMORPHISMS AND QUOTIENT RINGS

51

Example 2.2.7. (Z, +, ) is a subring of (Q, +, ) but it is not an ideal. For


/ Z.
example 1 Z, 13 Q, but 1 13
In fact if R is a ring with 1 then the only ideal of R that contains 1 is R itself.
Definition 2.2.8. R is always an ideal of R, R is called improper ideal of R, all
other ideals are called proper ideals, {0} is called the trivial ideal.
Example 2.2.9. Let I = {(r, 0); r R} then I is an ideal of R R. Since
(r, 0)(a, b) = (ra, 0) I for any r, a, b R, then (a, b)(r, 0) I. Note I has
a unity, yet I is a proper ideal of R. Why does this not contradict the above
statement? (Note, unity in R is not unity in I!)
With groups we have seen that if : G H is a group homomorphism, then
N = ker is normal in G, and infact, every normal subgroup is the kernel of
a homomorphism, namely the canonical projection : G G/N . We have a
similar situation for rings. But first we define ring homomorphisms.
Definition 2.2.10. Let R, S be rings.
(1) A ring homomorphism is a map : R S such that
(i) (a + b) = (a) + (b), i.e. is a group homomorphism of additive groups
(R, +), (S, +),
(ii) (ab) = (a)(b) a, b R.
(2) The kernel of the ring homomorphism is ker = {r R; (r) = 0S }, (i.e. it
is the kernel of viewed as homomorphism of additive groups).
(3) A bijective ring homomorphism is called an isomorphism.
Example 2.2.11.
: Z Z/2Z,

(
0, if n is even
n 7
1, if n is odd

is multiplicative since the product of even integers is even, the product of an even
and an odd integer is even and the product of two odd integers is odd, ker = 2Z,
which is an ideal of Z.
Example 2.2.12. Fix n Z. The maps n : Z Z, x 7 nx are not ring homomorphisms in general (except for n = 0, 1). Since n (xy) = nxy, n (x)n (y) =
n2 xy, hence n is a ring homomorphism if n2 = n. Hence only when n = 0, 1. But
note
n (x + y) = n(x + y) = nx + ny = n (x) + n (y),
hence n is a homomorphism of groups.

52

CHAPTER 2. RINGS





a b
a b
Example 2.2.13. : (M2 (R), +, ) (R, +, ),
7 det
.
c d
c d
(AB) = (A)(B), but (A + B) 6= (A)(B), hence is not a ring homomorphism.
Example 2.2.14. Let R be the ring of real valued functions on R under pointwise
addition and multiplication. Let R, : R R, f 7 f () evaluation at ,
then is a ring homomrophism, since
(f + g) = (f + g)() = f () + g() = (f ) + (g).
Similarly for (f g) = (f ) (g).
Proposition 2.2.15. Let R, S be rings. : R S be a ring homomorphism.
Then
(1) The image of is a subring of S,
(2) ker is an ideal of R.
Proof. (1) If s1 , s2 im(), then s1 = (r1 ), s2 = (r2 ) for some r1 , r2 R.
Then (r1 r2 ) = s1 s2 , (r1 r2 ) = s1 s2 , hence s1 s2 , s1 s2 im(), so im()
is a subring of S.
(2) If a, b ker then (a) = (b) = 0. Hence (a b) = (a) (b) =
0, (ab) = (a)(b) = 0, so ker is a subring of R. To see that it is an ideal,
let r R, ker then (r) = (r)() = (r) 0 = 0, hence r ker .
Similarly r ker , so ker is an ideal.

The following theorem is the analog of the first isomorphism theorem of groups.
Theorem 2.2.16 (1st isomorphism theorem). (1) If : R T is an onto homomorphism then
R/ ker
= T.
(2) If is the canonical homomorphism from R to R/ ker then there exists an
isomorphism : R/ ker T , such that = .
Proof. Exercise.
If : R T is an onto homomorphism, then we have a one-to-one correspondence between subrings of T and subrings of R which contain ker , with ideals
corresponding to ideals.

2.2. IDEALS, RING HOMOMORPHISMS AND QUOTIENT RINGS

53

Example 2.2.17. Let R = Z[x], I = {p(x) Z[x]; each term of p(x) has degree 2}
{0}, i.e. p(x) I has no constant term and has no x term. Then I is an ideal
since the product of p(x) I and any q Z[x],
(a2 x2 + + an xn )(b0 + bx + + bxm ) = b0 a2 x2 + + an bm xn+m
has no constant term and has no x-term, hence p(x)q(x) I. Two polynomials
p, q are in the same subset p(x) q(x) I, i.e. they differ by a polynomial of
degree at least 2, e.g. 1 + 3x x4 , 1 + 3x x5 are in the same coset.
A complete set of representatives of R/I are given by polynomials of degree at most
1, i.e. {a + bx; a, b Z} = R/I, for example 1 + 3x x4 = 1 + 3x = 1 + 3x x5 .
Addition and multiplication in the quotient are performed by representatives. e.g.
(1 + 3x) + (2 + 5x) = 3 + 8x, (1 + 3x)(2 + 5x) = (2 + 11x + 15x2 ) = 2 + 11x.
Since 15x2 = 0, i.e. 15x2 I = eR/I .
Note in R/I we have zero-divisors for example xx = x2 = 0 even though x = 0,
even though R has no zero-divisors. This example shows that the structure of the
quotient ring may seem worse than the original ring.
This is also the case for
Example 2.2.18. Z, 8Z is an ideal of Z. Z/8Z
= Z8 , Z8 has zero-divisors, Z does
not, but
Example 2.2.19. I = {0, 3} Z6 is an ideal of Z6 , Z6 /I has 3 elements
0 + I = {0, 3}, 1 + I = {1, 4}, 2 + I = {2, 5},

Z6 /I
= Z3

under the correspondence 0 + I 0, 1 + I 1, 2 + I 2. R = Z6 has


zero-divisors, Z6 /I = Z3 is a field.
Example 2.2.20. Reduction Homomorphism. The canonical projection map
: Z Z/nZ,

m 7 m = m mod n.

is a ring homomorphism and has important applications to number theory,


e.g. x2 + y 2 = 3z 2 to be solved in integers (Diophantic equation). Suppose such
solutions exists, i.e. there exist x, y, z Z such that x2 + y 2 = 3z 2 . First we
assume x, y, z have no common factors. Since if they did, we could find solutions
x0 , y 0 , z 0 that are relatively prime. If this equation holds in the ring Z, it must hold
in any quotient ring Z/nZ. Choose n = 4. This is a conveniant choice since the

54

CHAPTER 2. RINGS

only squares mod 4 are 0 and 1 (02 = 0, 12 = 1, 22 = 0, 32 = 1). Therefore the


possible solutions in Z/4Z are
   
   
0
0
0
0
+
3
=
mod 4.
1
1
1
3
Checking all possibilities shows that we must take 0 each time, hence x, y, z are
even integers ((2n + 1)2 1 (4)). But we assumed x, y, z to have no common
factors, contradiction follows. Hence the equation x2 + y 2 = 3z 2 has no integer
solution. (There are equations which have a solution modulo every integer but do
not have integer solutions, e.g. 3x2 + 4y 3 + 5z 3 = 0.)
Example 2.2.21. R = {f : R R} under pointwise addition and multiplication.
Fix c C and define
c : R R,
f 7 f (c)
is an onto homomorphism since for every r R the constant polynomial f (x) = r
has c (f ) = r. Since
ker c = Ic = {f R; f (c) = 0}, then R/ ker c
= R.
Example 2.2.22. Let R be a commutative ring with 1, G = {g1 , . . . , gn } a finite
group and RG the group ring. Define
: RG R,

n
X

ai gi 7

i=1

n
X

ai ,

i=1

the augmentation map. In = ker = augmentation ideal = elements in RG whose


coefficients add to 0, for example gi . . . gj IA . is surjective (ae 7 a), hence
RG/IA
= R.
The remaining isomorphism theorems can be proved as in the case of groups.
Theorem 2.2.23 (2nd Isomorphism Theorem). Let R be a ring, S a subring, I
an ideal of R. Then
S + I = {a + b; a S, b I}
is a subring of R, S I is an ideal of S and
(S + I)/I
= S/(S I).
(
Theorem 2.2.24 (3rd Isomorphism Theorem). Let R be a ring, I, J ideals of R.
Suppose I J. Then J/I is an ideal of R/I and
R/I
= R/J.
J/I

2.3. PROPERTIES OF IDEALS

2.3

55

Properties of ideals

Let R be a ring with 1 6= 0.


Definition 2.3.1. Let A be a subset of R.
(1) (A) denotes the smallest ideal of R containing A, called the ideal generated by
A
\
(A) =
I.
I an ideal, AI

(2) RA denotes the set of finite terms of elements of the form ra, with r R, a
A
RA = {r1 a1 + + rn an ; ri R, ai A, n Z+ }.
Similarly for AR.
RAR = {r1 a1 s1 + + rn an sn ; ri , si R, ai A, n Z+ }.
(3) (a) denotes the ideal generated by a single element and is called the principal
ideal (generated by a).
(4) An ideal generated by a finite set is called finitely generated.
If R is commutative then RA = AR = RAR = (A) and (a) = {ra; r R}. In
a commutative ring, principal ideals is a particularly easy way of forming ideals.
Similar to generating cyclic subgroups of a group. Note for b R, b (a) iff
b = ra for some r R, i.e. b (a) iff b is a multiple of a or iff a devides b
(i.e. divide is to contain). Note also b (a) iff (b) (a).
Commutative rings in which all ideals are principal are among the easiest to study
and have arithmetic properties similar to Z.
Definition 2.3.2. An integral domain with the property that every ideal is principal is called a principal ideal domain (PID).
Example 2.3.3. Z is a principal ideal domain, since every ideal in Z is of the
form nZ for some n.
We have observed that for a ring R and an ideal I the structure of R/I can be
better or worse than that of R.
Example 2.3.4. (a) Z has no zero-divisors, Z/6Z has zero-divisors.
(b) I = {0, 3} Z6 ,
Z6 /I
= Z3 has no zero-divisors.

56

CHAPTER 2. RINGS

Given R a ring, I an ideal we can sensibly ask which properties of R translate into
R/I?
For example if a R, when is I + a a zero-divisor in R/I? I + a is a zerodivisor iff there exists I + b R/I such that I + b 6= 0R/I = I + 0 and either
(I + a)(I + b) = I or (I + b)(I + a) = I. Hence I + a is a zero-divisor iff there
exists b
/ I such that either ab I or ba I. I + a is a (non-zero) zero-divisor in
R/I iff a
/ I and there exists b
/ I such that ab I or ba I.
Example 2.3.5. 2
/ 6Z, 3
/ 6Z but 6 6Z, hence 2 + 6Z is a zero-divisor in
Z/6Z.
To rule out zero-divisors in R/I the condition we need on I is
Definition 2.3.6. Let R be a ring, I an ideal. Then I is called prime if whenever
a, b R and ab I, then at least one of a or b is in I.
Example 2.3.7. p is prime in Z. Then pZ = (p) is a prime ideal because ab (p)
iff p | ab iff p | a or p | b.
Theorem 2.3.8. R/I has no zero-divisors iff I is a prime ideal.
Proof. Clear from the discussion above.
Corollary 2.3.9. If R is a ring with 1, I an ideal in R. Then R/I is an integral
domain iff I is a (proper) prime ideal.
Natural question: When is R/I a field? Let R be a ring with 1. R/I is a field
iff it is non trivial and each of its non-zero elements is a unit. R/I is nontrivial iff
I is a proper ideal.
Every non-zero element of R/I is a unit iff for all a R\I,there exists b R such
that
(I + a)(I + b) = I + 1
iff
()

a
/ I b R such that ab 1 I.

Claim 2.3.10. () holds iff I is maximal.


Definition 2.3.11. An ideal I of R is called maximal if I is proper and there is
no proper ideal J % I.
Proof of the claim. Suppose I is maximal. Take a
/ I. Then the set
J = {ar + x; r R, x I}

2.3. PROPERTIES OF IDEALS

57

is an ideal that properly includes I, hence J = R. In particular there exists


r0 R, x0 I such that ar0 + x0 = 1. Hence ar0 1 = x0 I, i.e. () holds.
Suppose () holds. Let J be an ideal such that I $ J. Take a J\I and
b R such that y := ab 1 I. Then a J, y I J. Hence 1 = ab y J.
But this implies r 1 J for all r R, therefore J = R and I is maximal.
We have shown
Theorem 2.3.12. Let R be a commutative ring with 1, I an ideal in R. Then
R/I is a field iff I is maximal.
Corollary 2.3.13. Let R be a commutative ring with 1, I an ideal. Then
I maximal I prime.
Proof. If I is maximal then R/I is a field, therefore R/I is an integral domain and
hence I is prime.
Warning! I prime 6 I maximal.
Example 2.3.14. R = Z Z, I = {(a, 0); a Z} non-trivial prime ideal.
((a, b), (A, B) R with (aA, bB) I then bB = 0 and hence either b = 0 or
B = 0.) Let J = {(a, 2b); a, b Z} be a proper ideal, I $ J. Hence I is not
maximal.
Note we have a simple lemma
Lemma 2.3.15. Let I be an ideal of R with 1.
(1) I = R iff I contains a unit.
(2) Assume R is commutative. Then R is a field iff its only ideals are 0 and R.
Proof. (1) Weve already seen.
(2) Let R be a field, then every non-zero ideal contains a unit, therefore I = R by
(a). Conversly if 0, R are the only ideals of R. Let u R, u 6= 0. By hypothesis
(u) = R, so 1 (u). This means there exists r R such that 1 = ru, and hence
u is a unit. Since u was arbitrary, R is a field.
Hence using lemma 2.3.15 we could also prove thoerem 2.3.12 using Isomorphism
Theorems
R/I a field R/I has no proper ideals.
By isomorphism theorems the ideals of J containing I correspond bijectively with
the ideals of R/I, therefore there exists no proper ideal of R/I 6 a proper ideal
J such that I $ J R I is maximal. Theorem 2.3.12 tells us how to construct
fields. Take a ring R with 1 (commutative) and take its quotient by a maximal
ideal.

58

CHAPTER 2. RINGS

Proposition 2.3.16. In a ring with 1, every proper ideal is contained in a maximal


ideal.
Proof. We skip the proof. Uses Zorns Lemma. (If S is a partially ordered set such
that every chain in S has an upper bound in S then S has at least one maximal
element.)
Definition 2.3.17. A partial ordering of a set S is given by a relation defined
for certain ordered pairs of elements of S such that the following is true
(1) a a, a S (reflexive),
(2) if a b, b a then a = b (antisymmetric),
(3) if a b, b c then a c (transitive).
Remark. In a partially ordered set, not every two elements need to be comparable,
i.e. for a, b S we need not have either a b or b a.
Definition 2.3.18. A subset T of a partially ordered set S is a chain if every two
elements a, b T are comparable, i.e. either a b or b a.
Example 2.3.19. The ideal (x) in Z[x] is not maximal because x (2, x) Z[x],
where (2, x) is the ideal generated by 2 and x in Z[x]. Observe
(2, x) = {2p(x) + xq(x); p(x), q(x) Z[x]},
the polynomials with integer coefficients whose constant term is even.

(2, x) = ker Z[x] Z/2Z, p(x) 7 p(0) mod 2
= {p(x) Z[x]; p(0) 2Z}.
Note (2, x) is not principal since if it were, then (2, x) = (a(x)) for some a(x)
Z[x]. Since 2 (a(x)), there exists p(x) such that 2 = p(x)a(x).
0 = deg(p(x)a(x)) = deg(p) + deg(a),
hence both p(x), a(x) have degree 0. Since 2 is prime a(x), p(x) {1, 2} if
a(x) = 1} then (a(x)) = Z[x], but this is a contradiction since (2, p(x)) is proper.
If a(x) = 2, then since x (2, x) = (a(x)) = (2) = (2) we have x = 2q(x) for
some q(x) Z[x] and this is clearly impossible since q is an integer polynomial.
Since (x) is not maximal Z[x]/(x) is not a field. In fact Z[x]/(x)
= Z since

(x) = ker : Z[x] Z, p(x) 7 p(0) .
The ideal (2, x) is maximal in Z[x] because Z[x]/(2, x)
= Z/2Z is a field.

2.4. RINGS OF FRACTIONS

59

Let F be a field. Using the homomorphism


: Z F, n 7 |1 + 1 +{z + 1} = n 1
n times

one can show that


Theorem 2.3.20. Let F be a field. Then F has a subfield that is isomorphic to
either Q or some Zp and F has only one such subfield.
Proof. Exercise.
Definition 2.3.21. The subfield of F that is isomorphic to Q or some Zp is called
the prime subfield of F . If F is either Q or Zp , then F is its own prime subfield.
Q and Zp are called prime fields.
If the prime subfield of F is isomorphic to Q equivalently 1F has infinite order in
(F, +), we say F has characteristic 0.
If F s prime subfield is isomorphic to Zp , F is said to have characteristic p.
Every finite field must have characteristic p for some prime p.
The nature of the solutions of equations involving elements of F can be heavily influenced by whether F has characteristic 0 or p. For example 0 = x2 + 1 has
no solutions in R, but has solutions in Z2 : 12 + 1 = 0 in Z2 .

2.4

Rings of fractions

Let R be a commutative ring. In Proposition 2.1.17 weve seen that if a is not


a zero-divisor and a 6= 0 then we have: if ab=ac then b = c. Thus a non zero
element which is not a zero-divisor enjoys some of the properties of a unit without necessarily possesing an inverse. (Recall that we have also seen that if a is a
zero-divisor then it can not be a unit.)
It turns out that a commutative ring R can always be made a subring of a larger
ring Q, in which every non zero element of R which is not a zero-divisor is a unit
in Q. In the case that R is an integral domain Q will be a field, called its field of
fractions or quotient field. The construction of Q from R takes inspiration from
the construction of rationals numbers Q from Z. What are the essential features
of construction of Q from Z?
Any rational number q Q is a quotient of two integers, but may be represented in many different ways as a quotient of two integers 31 = 26 = 93 = . . . .

60

CHAPTER 2. RINGS

These reprsentations are related by


c
a
= ad = bc,
b
d
more precisely each fraction ab is the equivalence class of ordered pairs of integers
(a, b), b 6= 0 under the equivalence relation
(a, b) (c, d) ad = bc.
Addition and multiplication on fractions are given by
a c
ad + bc
+ =
,
b d
bd

a c
ac
= .
b d
bd

Z is identified with the subring { a1 ; a Z} of Q. Every non-zero integer n has an


inverse n1 in Q.
Now turning to general ring R. If we allow arbitrary denominators in constructing Q we run into problems. If b is a zero-divisor in R say bd = 0 and if we
allow b as a denominator then we will have (assume R has 1)
d=

bd
0
d
=
= = 0 in Q,
1
b
b

i.e. if we allow zero-divisors or 0 in the denominators, there will be some collapsing


and we cannot expect R to appear naturally as a subring of ring of fractions.
(1) Hence the set of denominators should not have zero-divisors.
(2) A second restriction is more obviously imposed by laws of addition and multiplication: if b, d are allowed as denominators, then bd should also be allowed, i.e.
the set of denominators must be closed under multiplication.
The next theorem shows these two conditions are sufficient to construct a ring of
fractions of R.
Theorem 2.4.1. Let R be a commutative ring. Let D be any non-empty subset of
R that does not contain 0 or any zero-divisors, and is closed under multiplication,
i.e. for all a, b D we have ab D. Then there exists a commutative ring Q with
1 such that R is a subring of Q and every element of D is a unit in Q. The ring
Q has the properties
(1) Every element of Q is of the form rd1 for some r R, d D. If D = R\{0},
i.e. R has no zero-divisors, then Q is a field.
(2) (Uniqueness of Q) The ring Q is the smallest ring containing R in which all
elements of D become units. Q is the smallest in the following sense:
Let S be any commutative ring with 1. Let : R S be any injective ring

2.5. POLYNOMIALS

61

homomorphism such that (d) is a unit in S for all d D. Then there existis
an injective homomorphism : Q S such that R = . In other words, any
ring containing an isomorphism copy of R in which all elements of D become units
must also contain an isomorphism copy of Q.
Proof. Let Q = {(r, d); r R, d D} and : (r, d) (s, e) re = sd. Then
(r, d) = {(a, b); a R, b D and rb = ad}.
ad + bc a c
ac
a c
+ =
,
= .
b d
bd
b d
bd
Show
1) These operations are well-defined.
2) Q is an abelian group under addition. Additive identity is
inverse of ad is ad .
3) Multiplication is associative, distributive and commutative.
4) Q has an identity (= dd for any d D).

0
d

for all d D,

Definition 2.4.2. The ring Q is called the ring of fractions of D with respect to
R and is denoted by D1 R. If R is an integral domain and D = R\{0} then Q is
called field of fractions or quotient field of R.
Example 2.4.3. (1) If R is a field, its ring of fractions is R.
(2) The ring of fractions of Z is Q.
(3) 2Z also has no zero-divisors (no identity), its field of fractions is also Q.
(4) Let R be an integral domain. R[x] is also an integral domain and its field of
fractions is the field of rational functions in x, R[x].

2.5

Polynomials

Let R be a ring, then R[x] denotes the ring of polynomials in x with coefficients
in R. Weve seen that
Proposition 2.5.1.
Let R be an integral domain, p(x), q(x) R[x], then

1) deg p(x)q(x) = deg p(x) + deg q(x) ,
2) units of R[x] are the units of R,
3) R[x] is an integral domain.
The property being an integral domain is clearly passed from R to R[x]. One
property that will obviously not be passed on is that of being a field. If F is a
field, then F [x] is not a field. Clearly x is not a unit in F [x]. It is still reasonable
to expect that assuming F to be a field will have some beneficial impact on F [x].

62

CHAPTER 2. RINGS

Theorem 2.5.2. Let F be a field and f (x), g(x) F [x]. If g(x) 6= 0 then there
exist q(x), r(x) F [x] such that
f (x) = q(x)g(x) + r(x)
and either f (x) = 0 or deg(r) < deg(g). (Division algorithm for F [x].)
Proof. If f (x) = 0 or if f (x) 6= 0 and deg(f ) < deg(g) we write f (x) = 0g(x)+f (x)
and we are done.
We proceed induction on deg(f ). If deg(f ) = 0 then we are done by above unless
deg(g) = 0. In this case both f, g are constant polynomials, f (x) = a0 , g(x) = b0
and then
f (x) = (a0 b1
0 )g(x) + 0.
Note b1
0 exists because F is a field and b0 6= 0. Assume the result has been proved
for deg(f ) < n and suppose
f (x) = a0 + a1 x + + an xn , g(x) = b0 + b1 x + + bm xm , an , bm 6= 0.
If n < m we are done by the beginning of the proof. If n m we write
nm
f (x) = an b1
g(x) + h(x)
m x

where h(x) = 0 or deg(h) < deg(f ) = n. If h(x) = 0 we are done. If h(x) 6= 0


then by inductive hypothesis we can write
h(x) = q(x)g(x) + r(x)
with r(x) = 0 or deg(r) < deg(g). Then
f (x) = an b1
xnm g(x) + q(x)g(x) + r(x)
 m

1 nm
= an b m x
+ q(x) g(x) + r(x)
and we are done.
This proof really boils down to long division of polynomials. (In fact the
quotient and remainder (q, r) are uniquely determined by f and g.) There are
some important corollaries of the division algorithm in F [x].
Definition 2.5.3. Let F be a field, f (x) = a0 +a1 x+ +an xn F [x]. An element
a F is called a root (or zero) of f (x) if f (a) = 0, i.e. if a0 + a1 a + + an an = 0
in F .

2.5. POLYNOMIALS

63

Theorem 2.5.4 (Factor theorem). Let F be a field, a F , f (x) F [x]. Then


f (a) = 0 iff (x a) divides f (x) in F [x] (i.e. h(x) F [x] such that f (x) =
(x a)h(x)).
Proof. If f (x) = (x a)h(x) then f (a) = 0 h(a) = 0. Let

a (f ) = a (x a)(h(x)) = a (x a)a (h(x)) = 0 h(a)
be the evaluation homomorphism. Conversly suppose f (a) = 0. Apply division
algorithm to f (x), (x a) to get
f (x) = (x a)q(x) + r(x)
where either r(x) = 0 or deg(r) < deg(x a) = 1. Thus deg(r(x)) = 0 or r = 0 in
either case there exists c F such that r(x) = c and f (x) = (x a)q(x) + c.
Applying evaluation homomorphism a again gives
1 = f (a) = a (f (x)) = a ((x a)q(x) + c) = 0 q(a) + c
and therefore c = 0, so (x a) divides f (x).
Corollary 2.5.5. A non-zero polynomial f (x) F [x] of degree n can have at most
n zeroes in a field F .
Proof. Use induction on n. If n = 0 then f (x) is a non-zero constant polynomial,
hence has no roots in F . Suppose n = m + 1, the result holds for polynomials of
degree m. Suppose f has m + 2 roots in F say a1 , a2 , . . . am+2 .
If we write f (x) = (x a1 )h(x) then since F is a field, in particular an integral domain, i.e. has no zero-divisors. Each of a1 , . . . , am+2 must be a root of
either x a1 or h(x) (f (ai ) = 0 either ai a1 = 0 or h(ai ) = 0). In particular a2 . . . am+1 are m + 1 distinct roots of h(x). Since h(x) has degree m this
contradicts the inductive hypothesis and completes the proof.
Corollary 2.5.6. Let F be an infinite field. S an infinite subset of F . If f (x)
F [x] and f (s) = 0, s S. Then f 0.
Corollary 2.5.7. Let F be an infinite field, f (x), g(x) F [x], S an infinite subset.
If f (s) = g(s), for all s S then f (x) = g(x) for all x F .

64

2.5.1

CHAPTER 2. RINGS

Irreducible polynomials

Definition 2.5.8. Let F be a field. A non-constant polynomial f (x) F [x] is


called irreducible in F [x] (or irreducible over F ) if f cannot be written as the
product of two non-constant polynomials in F [x].
In general F [x] may contain irreducible polynomials of degree higher than 2. It
is still true that every non-constant polynomial can be written as a product of
irreducible factors. Note that the concept of irreducibility over F is a concept
that depends on the field. A polynomial f (x) may be irreducible over F but not
irreducible over a larger field containing F .
Example 2.5.9. x2 2 is a polynomial in Q[x] but has no zeroes in Q. Hence
x2 2 is irreducible over Q.Howeverx2 2 is also a polynomial in R[x] where
it is reducible, namely (x 2)(x + 2). x2 = 2 has no solutions in Q, since if
 2
m
m

Q,
(m,
n)
=
1
and
= 2 then m2 = 2n2 , hence we have
n
n
2 | 2n2 2 | m2 2 | m 4 | m2 4 | 2n2
2 | n2 2 | n, but (m, n) = 1,
and hence we have the contradiction.
Theorem 2.5.10. Let F be a field, f (x) a non-constant polynomial in F [x]. Then
there exist irreducible polynomials f1 (x), . . . , fk (x) F [x] such that
f (x) = f1 (x) . . . fk (x).
Proof. Use induction on deg(f ). If deg(f ) = 1 then f (x) itself is irreducible and
were done. Suppose deg(f ) = n, and the theorem is proved for all polynomials
of degree less than n. If f (x) is irreducible we are done. Otherwise we can write
f (x) = g(x)h(x), where g, h have degree at least 1; hence each have degree strictly
less than n. So by inductive hypothesis we can factor g(x) and h(x) into irreducible
factors. This yields the desired factorization.
If f has degree 2 or 3, f F [x], F a field, then f is reducible in F [x] f (x) has a
root in F . Note this is not true for deg(f ) > 3. For example (x2 +1)(x2 +1) R[x]
is clearly reducible but has no root in R. There are special techniques that can be
used in case of checking irreducibility in Q[x]. In fact we can confine our attention
to polynomials with integer coefficients. Since if
a0 a1
an
f (x) =
+ x + + xn Q[x],
b0
b1
bn
then the irreducibility of f (x) is equivalent to the irreducibility of g(x) where
g(x) Z[x] obtained from f (x) by multiplying f (x) by b0 b1 . . . bn . We have

2.5. POLYNOMIALS

65

Lemma 2.5.11. Let f (x) Z[x]. If f (x) can be written as the product of two
non-constant polynomials in Q[x] then f (x) can be written as the product of two
non-constant polynomials in Z[x].
To prove this lemma we first define
Definition 2.5.12. The content of a non-zero polynomial in Z[x] is the greatest
common divisor of its coefficients. A polynomial is called primitive if its content
is 1.
The essential fact about those concepts is
Lemma 2.5.13 (Gauss lemma). The product of two primitive polynomials is
primitive.
Proof. Exercise.
Proof of lemma 2.5.11. Exercise too!

You might also like