You are on page 1of 20

10

Chapter 1

Groups

1.1 Isomorphism theorems


Throughout the chapter, well be studying the category of groups. Let G, H be groups. Recall
that a homomorphism f : G H means a function such that f (g1 g2 ) = f (g1 )f (g2 ) for all
g1 , g2 G. A homomorphism is called a monomorphism if it is injective, i.e. there exists a set
function f 0 : H G such that f 0 f = idG . Its an epimorphism if its surjective, i.e. there
exists a set function f 0 : H G such that f f 0 = idH . Its an isomorphism if its bijective, i.e.
both injective and surjective. Note this definition of isomorphism of groups is equivalent to the
definition of isomorphism given earlier for an arbitrary category: luckily for groups a morphism is
an isomorphism if and only if it is a bijection. We also talk about endomorphisms (homomorphisms
from G to itself) and automorphisms (isomorphisms from G to itself).
Given a homomorphism f : G H, we set

im f = f (G) = {f (g) | g G}, ker f = f 1 (1H ) = {g G | f (g) = 1H },

its image and kernel. A subgroup of a group G is a non-empty subset K of G such that k1 k21 K
for every k1 , k2 K. This condition implies that K is itself a group in its own right, the operation
being the restriction of the operation on G to K. We usually write K G to indicate that K
is a subgroup of G. For example, im f H (and f is an epimorphism if and only if im f = H);
ker f G (and f is a monomorphism if and only if ker f = {1G }).
Actually, ker f is more than just a subgroup of G: it is a normal subgroup. By definition, a
subgroup N G is normal if any of the following equivalent conditions hold:

(N1) gN g 1 = {gng 1 | n N } = N for all g G;


(N2) gN g 1 = {gng 1 | n N } N for all g G;
(N3) gng 1 N for all g G, n N ;
(N4) N g = gN for all g G.

We usually write N E G to indicate that N is a normal subgroup of G.


The sets N g = {ng | n N } and gN = {gn | n N } appearing in (N4) are the right cosets
and left cosets of N in G. Lets make this notion more precise. Let K be any subgroup of G.
Define an equivalence relation on G called left congruence modulo K by g1 g2 if g11 g2 K. The
equivalence classes are called the left cosets of K in G: each equivalence class has the form gK for
some (any) representative g G of the equivalence class. The set of all left cosets of K in G is
denoted G/K.
One also defines right cosets to be the equivalence classes of a relation called right congruence
modulo K, namely, g1 g2 if g1 g21 K. Right cosets look like Kg for g G, and the set of all

11
12 CHAPTER 1. GROUPS

right cosets is denoted K\G. Note the map g 7 g 1 induces a bijection between G/K and K\G,
so it usually doesnt matter whether you choose work with left or right cosets.
In case G is a finite group and K G, we let [G : K] = |G/K| denote the number of left cosets
of K in G, the index of K in G. Now, all the left cosets have the same size, namely, |K|. Since
equivalence classes partition G into disjoint subsets, we deduce:

Lagranges theorem. |G| = |K|[G : K].

Lagranges theorem shows in particular that the order of any subgroup of a finite group G divides
the order of the group.
Now return to the case that N is a normal subgroup of G. Then, (N4) says that G/N = N \G,
i.e. left cosets are the same as right cosets so we can just call them simply cosets. Define a
multiplication on the set of cosets G/N by

(g1 N )(g2 N ) = (g1 g2 )N.

The fact that this is well-defined depends on N being a normal subgroup. This multiplication
gives G/N the structure of a group in its own right, called the quotient group of G by the normal
subgroup N . Note that there is a canonical homomorphism : G G/N , g 7 gN , which is
an epimorphism of G onto G/N with kernel exactly N . The group G/N together with the map
: G G/N has the following universal property:

Universal property of quotients. Let N E G and : G G/N be the canonical epimorphism.


Given any homomorphism f : G H with N ker f , there exists a unique homomorphism
f : G/N H such that f = f .

Exercise. Let G be a group and H, K be subgroups. We can consider the subset HK = {hk | h
H, k K} of G.
(1) HK is a subgroup of G if and only if HK = KH as subsets of G.
(2) If either of H or K is a normal subgroup of G, then HK = KH, hence is a subgroup of G.
G.
(3) If in fact both H and K are normal subgroups of G, then HK is a normal subgroup of G.

Now let me state the isomorphism theorems for groups. These are all really consequences of
the universal property of quotients: you should be able to prove them for yourselves.

First isomorphism theorem. Let f : G H and N = ker f . Then, N E G and f factors


through the quotient G/N to induce an isomorphism f : G/N im f .

Second isomorphism theorem. Let K G, N EG. Then, N EKN, K N EK and K/K N


=
KN/N.

Third isomorphism theorem. Let K N G with K E G, N E G. Then, N/K E G/K and


G/N
= (G/K)/(N/K).

There is one other important result traditionally included with the isomorphism theorems: the
lattice isomorphism theorem. First, recall that a relation on a set X is called a partial ordering
(and X is called a partially ordered set) if for all x, y, z X
(R1) x x;
(R2) x y and y x implies y = x;

(R3) x y and y z implies x z.


1.2. CYCLIC AND DIHEDRAL GROUPS 13

If in addition we have that for all x, y X one of x < y, x = y, x > y holds, then < is called a
total order or a linear order (and X is called a totally ordered set).
Now let X be a partially ordered set and A X. An element a X is called the join or least
upper bound of A if a is the unique minimal element of {x X | a x a A}. The join A may or
may not exist: for instance there could be many such minimal elements or no minimal element at
all. Similarly, an element a X is called the meet or greatest lower bound of A if a is the unique
maximal element of {x X | a x a A}. A partially ordered set is called a lattice if every pair
of elements of X has both a join and a meet in X. A partially ordered set is called a complete
lattice if every non-empty subset of X has both a join and a meet in X.
Let G be any group and X be the set of all subgroups of G, partially ordered by inclusion.
Then, X is a complete lattice. The meet of a set of subgroups of G is simply their intersection, also
a subgroup of G. The join of a set of subgroups of G is the subgroup generated by the subgroups,
that is, the intersection of all subgroups of G that contain all the given subgroups. More generally,
given any subset A of G, we write hAi for the subgroup of G generated by A, namely, the intersection
of all subgroups of G that contain A. In the special case A = {a} for some a G, we write simply
hai for the subgroup of G generated by the element a.
As a variation, let G be any group and X be the set of all normal subgroups of G, partially
ordered by inclusion. One checks that the intersection of a family normal subgroups is normal,
and that the group generated by a family of normal subgroups is normal. Hence, X in this case is
again a complete lattice, with join and meet being as for subgroups.

Lattice isomorphism theorem. Let f : G G0 be an epimorphism with kernel K. Then, the


map H 7 f (H) gives an isomorphism between the lattice of subgroups (resp. normal subgroups)
of G containing K and the lattice of subgroups (resp. normal subgroups) of G0 .

(The inverse map sends a subgroup H 0 of G0 to its preimage f 1 (H 0 ) in G.)

1.2 Cyclic and dihedral groups


Its nearly time for some examples of groups. But first, let G be any group and g G an element.
Recall hgi denotes the subgroup of G generated by the element g. So, hgi = {1G , g 1 , g 2 , g 3 , . . . },
and is called the cyclic subgroup generated by g. It may or may not be a finite group; if it is finite,
its order is denoted o(g), the order of the element g. Note Lagranges theorem implies that in a
finite group, the order of every element divides the order of the group.
A group G is called cyclic if G = hgi for some element g G. We then say that g is a generator
of the cyclic group G. For example, the group (Z, +) of integers under addition is an infinite cyclic
group generated by the integer 1 (the identity element of Z is the integer 0 of course we use the
additive notation in this Abelian group!). Actually, any infinite cyclic group is isomorphic to the
group (Z, +). We can classify all subgroups of the infinite cyclic group:

1.2.1. Lemma. If H (Z, +), then H = hni for some integer n 0.

Proof. Take n > 0 minimal subject to the condition that n H (if no such n exists, then H = {0}
and there is nothing to prove). Then, hni H. Now take any m H, and write m = qn + r for
integers q, r with 0 r < n. Then, r = m qn H, so by minimality of n we have that r = 0.
Hence, m = qn, so that m hni. This shows that H = hni.
Note that hni hmi if and only if m|n. It follows that the lattice of subgroups of Z is
isomorphic to the opposite of the lattice of non-negative integers partially ordered by divides.
The non-negative integer n corresponds under the isomorphism to the subgroup hni.
Given n N, we define the group (Zn , +), the group of integers modulo n under addition, to be
the quotient group Z/hni. Thus, the elements of the group Zn are the cosets of hni in Z, namely,
{[0], [1], . . . , [n 1]} where [i] = {i + jn | j Z}. We have that [i] + [j] = [i + j]. The group Zn is
the cyclic group of order n: it is generated by the element [1] which has order n in Zn . Sometimes
14 CHAPTER 1. GROUPS

we denote the cyclic group of order n instead by Cn , to indicate that were writing the operation
multiplicatively instead. So Cn = {1, x, x2 , . . . , xn1 } where x is any element of Cn of order n.
By the lattice isomorphism theorem, the lattice of subgroups of Zn is isomorphic to the opposite
of the lattice of divisors of n. The divisor d of n corresponds under the isomorphism to the cyclic
subgroup of Zn generated by [d], which is a subgroup of order n/d. In other words:

1.2.2. Lemma. The group Zn has a unique subgroup of order d for each divisor d of n, namely,
the cyclic subgroup generated by [n/d].

Given groups G1 , G2 , their (external) direct product is the set G1 G2 (Cartesian product) with
coordinatewise multiplication. Actually, G1 G2 , with the obvious projections i : G1 G2 Gi
is the product of G1 and Q G2 in the categorical sense. More generally given a family Gi (i I)
of groups, their product iI Gi is simply their Cartesian product as sets with coordinatewise
multiplication. In other words, the category of groups possesses arbitrary products (see (0.3.2)).

1.2.3. Lemma. If n = st with (s, t) = 1, then Zn


= Zs Zt .

Proof. The element ([1]s , [1]t ) Zs Zt has order st as (s, t) = 1. Hence it generates a cyclic
subgroup of Zs Zt of order st. Hence, it generates all of Zs Zt , which is therefore isomorphic
to the cyclic group of order n.
The Euler function is defined by

(n) = #{x Zn | o(x) = n} = #{1 k < n | (k, n) = 1}.

By the lemma, if n = st with (s, t) = 1, then

(n) = (s)(t).

It follows immediately that to compute (n) it suffices to know (pn ) for each prime power pn . In
this special case, it is an exercise to show that
 
n n 1
(p ) = p 1 .
p

Heres an important number theoretic fact about the Euler function:

1.2.4. Lemma. For n N, X


n= (d).
d|n

Proof. We have from Lagranges theorem that


X
n = |Zn | = (the number of elements of Zn of order d).
d|n

Now if x Zn has order d, it generates a subgroup of order d, isomorphic to Zd . By Lemma 1.2.2,


Zn has a unique such subgroup of order d. Hence, the number of elements of Zn of order d equals
the number of elements of Zd of order d, namely (d).
I end the section by introducing one other basic example of a finite group: the dihedral group Dn
of order 2n. This is probably easiest left as an exercise. So, start with the group O2 of orthogonal
linear transformations of the plane R2 . So O2 consists of all rotations (det = +1) around the
origin and all reflections (det = 1) in axes passing through the origin. Consider the elements
f, g O2 where g is reflection in the x-axis and f is counterclockwise rotation through the angle
2/n. Obviously, f n = 1, g 2 = 1. Let Dn be the subgroup of O2 generated by f and g.
1.3. THE SYMMETRIC GROUP 15

(1) Show that the product f i g is reflection in the line at angle i/n to the x-axis. Hence, the
transformations 1, f, f 2 , . . . , f n1 , g, f g, . . . , f n1 g are all distinct.
(2) Now show that the elements {1, f, f 2 , . . . , f n1 , g, f g, . . . , f n1 g} form a subgroup of O2 .
(3) Deduce that
Dn = {1, f, f 2 , . . . , f n1 , g, f g, . . . , f n1 g}
is a group of order 2n. Picture: Dn is the group of symmetries of a regular n-gon. It contains the
cyclic group Cn as a subgroup, namely, the subgroup {1, f, f 2 , . . . , f n1 } consisting of all rotations.
(4) Find all normal subgroups of Dn .
(5) For any group G, its centre is defined as
Z(G) = {x G | xy = yx for all y G}.
Determine Z(Dn ).

1.3 The symmetric group


We come now to one of the most important examples of a group. Let X be any set. Denote
by A(X) the set of all bijective (invertible) functions from X to itself. Then, A(X) is a group
with multiplication being composition of functions. The identity element of the group A(X) is the
identity function idX : X X, x 7 x. The group A(X) is called the symmetric group on X. In
the special case that X is the set {1, 2, . . . , n}, we denote the group A(X) instead by Sn and call
it the symmetric group on n letters (even though theyre numbers!). Obviously, |Sn | = n! so Sn is
a finite group.
Now let G be any group. Then, A(G) is also a group. Define : G A(G), g 7 g where
g A(G) is the function x 7 gx (i.e. g is left multiplication by g). Note is a group
homomorphism. Moreover, its injective, for if g = idG , then g (1G ) = 1G = g so ker = {1G }.
Hence, defines an isomorphism between G and im A(G). Weve proved:
Cayleys theorem. Every group is isomorphic to a subgroup of the symmetric group A(X) for
some set X. Every finite group, is isomorphic to a subgroup of Sn for some n.
By the way, you can also define : G A(G), g 7 g , where g is the function right
multiplication by g. But satisfies (gh) = (h)(g), the wrong way round to be a homorphism.
In fact, is an antihomomorphism from G to A(G).
Now lets discuss the finite group Sn in more detail. So its the group of all permutations of
{1, . . . , n}. Let x1 , . . . , xa be distinct elements of {1, . . . , n}. Denote the permutation which maps
xa 7 x1 , xi 7 xi+1 for each i = 1, . . . , a 1 and fixes all other points, by
(x1 x2 . . . xa ).
This is called an a-cycle. A 2-cycle is called a transposition.
Disjoint cycle notation. Every permutation f Sn can be written as a product of disjoint cycles
f = (x1 x2 . . . xa )(y1 y2 . . . yb ). . . . (z1 z2 . . . za ).
Moreover, this representation of the permutation f is unique up to deleting 1-cycles and reordering
the product (disjoint cycles commute!).
(You should be quite familiar with using the disjoint cycle notation.) Now define a function
sgn : Sn {1}
as follows. Take g Sn . Write g = c1 . . . ca as a product of disjoint cycles, where ci is an oi -cycle.
Set
Ya
sgn(g) = (1)oi 1 .
i=1
So for example, sgn(1) = 1, sgn(t) = 1 for t a transposition,...
16 CHAPTER 1. GROUPS

1.3.1. Lemma. Every element g Sn can be expressed as a product of transpositions.

Proof. Using the disjoint cycle decomposition, it suffices to show that any a-cycle can be written as
a product of transpositions. Then, for example, the cycle (12 . . . n) equals (12)(23)(34) . . . (n1 n).

1.3.2. Theorem. sgn is a group homomorphism.

Proof. We first prove:


Claim. If s is a transposition and w Sn is arbitrary, then sgn(sw) = sgn(w). To see
this, say s = (a b) and w = c1 . . . cm as a product of disjoint cycles. If s and w are disjoint, the
conclusion follows by definition. Else, without loss of generality, we may assume a appears in the
cycle c1 . If b appears in c1 too, then c1 = (a x1 . . . xk b y1 . . . yl ) and

(a b)c1 = (b y1 . . . yl )(a x1 . . . xk )

which is a product of disjoint cycles. In then follows by definition of sgn that sgn((a b)w) =
sgn(w). The other possibility is if b appears in another of the cycles, say c2 (allowing 1-cycles).
Now another similar explicit calculation with cycle notation expresses (a b)c1 c2 as a product of
disjoint cycles, and the conclusion again follows using the definition of sgn.
Now we can prove the theorem. We need to show that sgn(xw) = sgn(x) sgn(w) for any
x, w Sn . Write x = s1 . . . sm as a product of transpositions, applying Lemma 1.3.1, and proceed
by induction on m, the case m = 1 being the claim. For m > 1, set y = s1 x = s2 . . . sm .
Then, by the claim, sgn(x) = sgn(y); by the induction hypothesis, sgn(yw) = sgn(y) sgn(w) =
sgn(x) sgn(w). Hence, using the claim once more, sgn(xw) = sgn(s1 (yw)) = sgn(yw) =
sgn(x) sgn(w).

1.3.3. Corollary. If w Sn is written as a product of m transpositions, then sgn(w) = (1)m .

We define the alternating group An to be the kernel of the homomorphism sgn : Sn {1}.
Providing n > 1, An is a normal subgroup of Sn of index 2. The elements of An are called even
permutations. A group G is called simple if it has no proper normal subgroups. For instance,
the cyclic group Cp where p is a prime is a simple group for Lagranges theorem implies that
Cp has no proper subgroups at all. The goal in the remainder of the section is to prove that
An is simple for n 5. (On the other hand, A4 is not simple, for it contains the Klein 4 group
V4 = {1, (12)(34), (13)(24), (14)(23)} as a normal subgroup of index 3.)
As a first step to the goal, we need to understand the conjugacy classes of the group Sn . In
any group G, the conjugate of an element x G by g G is defined to be
g
x := gxg 1 .

The conjugacy class of x is the set Gx := {gx | g G}. The conjugacy classes of G partition it into
disjoint subsets (because conjugacy classes are the equivalence classes of the equivalence relation
is conjugate to). Observe that a subgroup N G is a normal subgroup if and only if it is a
union of conjugacy classes of G: so once we understand conjugacy classes we can quite easily test
subgroups for normality.

1.3.4. Lemma. Take g, x Sn and write

x = (a1 . . . as )(b1 . . . bt ) . . .

as a product of disjoint cycles. Then,

gxg 1 = (ga1 . . . gas )(gb1 . . . gbs ) . . . .


1.3. THE SYMMETRIC GROUP 17

Proof. Calculate what each side does to an integer i.


Define the cycle-type of x Sn to be the ordered tuple (o1 , o2 , . . . , ol ) consisting of the orders
of the disjoint cycles of x in cycle notation, arranged so Plthat o1 o2 ol > 0. You
should include all the trivial 1-cycles so that in addition i=1 oi = n. For instance (1 2) S5 has
cycle-type (2, 1, 1, 1). The lemma immediately implies the description of the conjugacy classes of
Sn :

1.3.5. Theorem. The conjugacy classes of Sn are precisely the {x Sn | x has cycle-type } as
runs over all conceivable cycle-types.

Now we can prove that An is simple for n 5. We proceed with a series of lemmas.

1.3.6. Lemma. The group A5 is simple.

Proof. Let us list the conjugacy classes in S5 .

Cycle-type Size sgn Splits in A5 ?


(5) 24 + yes
(4, 1) 30
(3, 2) 20
(3, 1, 1) 20 + no
(2, 2, 1) 15 + no
(2, 1, 1, 1) 10
(1, 1, 1, 1, 1) 1 + no

Now an Sn -conjugacy class that is contained in An can either be equal to a single An -conjugacy
class, or else it can be a union of two An -conjugacy classes of the same size. We have listed in the
table which of the S5 -classes split as A5 -classes in this way. Therefore, the conjugacy classes of
A5 have sizes 12, 12, 20, 15, 1.
Now, a normal subgroup of A5 must be a union of these conjugacy classes, must contain 1 and
must be of order dividing 60. Theres no way to do this by elementary arithmetic based on the
orders of the classes in A5 !

1.3.7. Lemma. If H E An for n 5 and H contains a 3-cycle, then H = An .

Proof. Show that An is generated by 3-cycles and all 3-cycles are conjugate in An .

1.3.8. Lemma. The group A6 is simple.

Proof. Let {1} = 6 H C A6 . Suppose 1 6= g H has a fixed point. Without loss of generality,
suppose g6 = 6. Then, g lies in the naturally embedded subgroup A5 < A6 . So, {1} =6 H A5 EA5 ,
so H A5 = A5 by simplicity of A5 . Hence, H contains a 3-cycle, so H = A6 by the previous
lemma.
This reduces to the case that all elements 1 6= g H move all 6 points. Now consider the two
possible cycle types one by one and get a contradiction. For instance, if (12)(3456) H, then its
square is a non-identity element with a fixed point. Otherwise, (123)(456) H (or an element of
the same cycle-type). In this case, conjugating gives that

(123)(456)(234)(123)(456)(243) H.

But this is a non-identity element that fixes 1.

1.3.9. Theorem. For n 5, An is simple.


18 CHAPTER 1. GROUPS

Proof. We may assume that n 7. Take {1} 6= H C An . Using induction on n, the argument in
the proof of the previous lemma reduces to the case when all 1 6= g H move all of 1, . . . , n. Take
such a g such that, without loss of generality, g1 = 2. Consider

(234)g(243)g 1 H.

Its a non-trivial element since it sends 2 to 3. But its a product of two 3 cycles, so it can move at
most 6 points, so as n 7 it must have a fixed point.

1.4 Free groups


Let X be a set. A group G is said to be free on X if
(1) X G;
(2) for every set function f : X H to a group H, there exists a unique group homomorphism
f : G H whose restriction to X G equals f .
(It is sometimes convenient to weaken condition (1) to say just that there is a given distinguished
injective function i : X , G. In this formulation, (2) would say that there exists a unique
f : G H such that f = f i.)
This definition is our first detailed example of a definition by a universal property. It will
always be the case with such definitions that if such an object exists, it is unique up to canonical
isomorphism. The proof always follows the same pattern, and the one below is the only one Ill
ever give:

1.4.1. Lemma. If G and G0 are both free on X then there is a unique isomorphism j : G G0
such that j(x) = x for all x X.

Proof. Consider the inclusion maps f : X , G0 and g : X , G. Since G is free on X, there


exists a unique f : G G0 which is the identity on X. Since G0 is free on X, there exists a unique
g : G0 G which is the identity on X. Now, f g : G0 G0 is a group homomorphism that is the
identity on X. But since G0 is free on X, there exists a unique such homomorphism, namely, the
identity on G0 . Hence, f g = idG0 . Similarly, g f = idG .
In view of the lemma, we will henceforth abuse notation and call G the free group on X if it
satisfies the properties (1) and (2) above. There is now a problem: it is not at all clear whether
such a group even exists at all! This is always the case with univeral property definitions: we now
need to give an explicit construction to prove existence.

1.4.2. Theorem. For every set X, there exists a group G that is free on X.

Proof. We need a construction. Let X 1 = {x1 | x X} be a set of distinct elements disjoint


from X. By convention, we set (x1 )1 = x for x X. Let 1 be an element disjoint from X X 1 .
A word means an ordered tuple (a1 , a2 , . . . , an ) for n 1 with each ai {1} X X 1 . A word
is called reduced if
(1) a2 , . . . , an 6= 1, and a1 6= 1 in case n > 1;
(2) ai 6= a1 i+1 for i = 1, . . . , n 1.
Given two reduced words, we define their product by concatenation followed by replacement of
xx1 s by 1s and removal of 1s whenever possible. Then we would like to check that the set
of reduced words on X, with this operation, forms a group G. But this is hard: the problem is
proving associativity. So instead, we proceed indirectly via a trick.
Let W be the set of reduced words. For each x X, define functions |x| and |x1 | from W to
W by
x x1 . . . xnn if x 6= x
 1
1 ,
1
|x |(x11 . . . xnn ) = 2 n 1
x2 . . . xn if x = x1 .
1.4. FREE GROUPS 19

Note |x| is the inverse function to |x1 |. Hence, each |x| is an element of the symmetric group
A(W ). Now define G to be the subgroup of A(W ) generated by the elements {|x| | x X}. We
claim that G, together with the function i : X G, x 7 |x|, is free on X.
We first need to observe that i is injective. To see this, note any element of G can be expressed
as a reduced word |x11 | . . . |xnn | in the |x|, |x1 |. Moreover this reduced representation is unique,
for applying this function to 1 yields x11 . . . xnn , a reduced word in W , and words in W have unique
reduced spelling by definition. In particular, since each |x| for x X is reduced, we get from this
that i is injective.
Now we verify that G, together with the map i : X , G, really is free on X. Take f : X H
any function to a group H. Define f : G H on a reduced word |x11 | . . . |xnn | by
f(|x11 | . . . |xnn |) = f (x1 )1 . . . f (xn )n .
This is well-defined since we showed that reduced words have unique spelling. It just remains to
check that f is a group homomorphism. But this follows since cancellation in G implies cancellation
in H.

1.4.3. Corollary (of proof). If G is free on X, then G is generated by X.


Proof. We showed in the construction that every element of G can be written as a reduced word
in X.
Now suppose that X is a set and R is a set of reduced words in X. Define the group hX|Ri,
the group given by generators X and relations R, to be the quotient group F/N , where F is the
free group on X and N is the normal subgroup of F generated by R (the intersection of all normal
subgroups of F containing R). Informally, hX|Ri is the largest group that can be generated by
the elements X in which the relations R hold. Formally:
1.4.4. Theorem. Given any group G generated by (not necessarily distinct) elements {tx | x X}
satisfying the relations R, there is a unique epimorphism f : hX|Ri G such that f (x) = tx for
each x X.
Proof. Let F be the free group on X, and : F hX|Ri be the canonical quotient map. By
the universal property of free groups, there exists a unique homomorphism h : F G such that
h(x) = tx for each x X. Since the relations are satisfied by the tx , R ker h. Hence, the normal
subgroup of F generated by R lies in ker h. Hence, by the universal property of quotients, h factors
through the quotient hX|Ri of F to induce a unique f : hX|Ri G such that f(x) = tx for each
x X.
Now for some examples of groups given by generators and relations (which can often be an
entirely intractable way of defining a group!).
Consider G = ha, b | an = b2 = 1, ab = ban1 i. Using the relations, one easily shows that every
element of G can be reduced to one of the words ai bj , i = 0, 1, . . . , n 1, j = 0, 1. Hence, |G| 2n.
To prove that |G| = 2n is hard from the point of view of generators and relations. However, we
have already seen the dihedral group Dn which is generated by a rotation f of order n and a
reflection g, and these satisfy the relation f g = gf n1 . Hence, applying Theorem 1.4.4, the map
a 7 f, b 7 g extends to a unique epimorphism G Dn . This shows that |G| |Dn | = 2n. Hence,
|G| = 2n and G = Dn .
Heres a harder example giving generators and relations for the symmetric group Sn . Let Gn
be the group with generators {s1 , s2 , . . . , sn1 } subject to the relations s2i = 1, si sj = sj si for
|i j| > 1 and si si+1 si = si+1 si si+1 . Let Sn denote the symmetric group, and ti denote the basic
transposition (i i + 1) in Sn .
(1) Prove that the ti satisfy the same relations as the si .
(2) Embed Sn1 into Sn as the subgroup consisting of all permutations fixing n. Prove that
{1, tn1 , tn2 tn1 , . . . , t1 t2 . . . tn1 } is a set of Sn /Sn1 -coset representatives.
(3) By considering the subgroup Gn1 of Gn generated by s1 , . . . , sn2 only and using induction,
prove that Gn = Sn .
20 CHAPTER 1. GROUPS

1.5 Group actions


We say a group G acts (on the left) on a set X if there is a given map G X X, (g, x) 7 gx
such that
(A1) 1x = x for all x X;
(A2) (gh)x = g(hx) for all x X, g, h G.
We call X a (left) G-set if G acts on X.
Equivalently, X is a G-set if the map : G A(X), g 7 g , where g (x) = gx for x X,
is a group homomorphism. Then, the G-set X is called faithful if this map is injective. So the
action is faithful if and only if gx = x for all x X implies g = 1.
The action of G on X is called transitive if for every x, y X, there exists at least one g G
such that gx = y. More generally, let X be a G-set and define an equivalence relation on X by
x y if there exists a g G such that gx = y. The equivalence classes are called the orbits of G
on X; the orbit of x X is denoted Gx = {gx | g G}. So the action is transitive if and only if
there exists just one orbit.
Elements x X with gx = x for all g G (i.e. the orbit of x is just {x}) are called fixed points,
and we write F ix(X) for the set of all fixed points. Given x X, its stabilizer Gx , or the isotropy
group of x, is defined to be Gx = {g G | gx = x}. Note Gx is a subgroup of G. So, x is a fixed
point if and only if Gx = G.
Given x X, the resulting orbit map is the map f : G X, g 7 gx. Clearly, f (g) = f (g 0 ) if
and only if g 1 g 0 Gx , i.e. gGx = g 0 Gx , i.e. g and g 0 lie in the same left Gx -coset. Hence, the
orbit map induces a well-defined map

G/Gx Gx, gGx 7 gx.

This map is clearly surjective, and we checked above that it is injective. In other words, the orbit
map induces a bijection between the cosets G/Gx of Gx in G and the orbit Gx of x. In particular,
if G is finite, we get that
|Gx| = [G : Gx ],
so the size of an orbit divides the order of G.

1.5.1. Lemma. If x and y lie in the same orbit, then Gx and Gy are conjugate in G, i.e. there
exists g G such that gGx g 1 = Gy .

Proof. Pick g G such that gx = y, i.e. x = g 1 y. Then, for h Gx , we have that ghg 1 y =
ghx = gx = y, so gGx g 1 Gy . Similarly, g 1 Gy g Gx . Hence, Gy gGx g 1 .
Now we have the language, we are ready for examples.

1.5.2. Given any G-set X, a subset Y X is called G-stable if gy Y for each g G, y Y . So,
Y is G-stable if and only if it is a union of orbits. In that case, Y is itself a G-set in its own right
via the restriction of the action of G on X to Y .

1.5.3. If H < G is any subgroup, then H acts on G by left multiplication, i,e, (h, g) 7 hg. The
orbits are the right cosets H\G.

1.5.4. If H < G is any subgroup, then H acts on G by (h, g) 7 gh1 . The orbits are the left cosets
G/H.

1.5.5. If H, K < G, then H K acts on G by ((h, k), g) 7 hgk 1 . The orbits are the double cosets
K\G/H, subsets of the form HgK = {hgk | h H, k K}.
1.5. GROUP ACTIONS 21

1.5.6. G acts on itself by conjugation, (g, x) 7 gxg 1 . The orbits are just the conjugacy classes
of G. In this case, given x G, the stabilizer Gx = {g G | gxg 1 = x} is called the centralizer of
x in G. We often write CG (x) for the centralizer of x in G. The order of the conjugacy class Gx is
the index of CG (x) in G:
|Gx| = [G : CG (x)].
The fixed points for this action are precisely the elements in the center Z(G). An element x G
lies in Z(G) if and only if CG (x) = G.

1.5.7. More generally, G acts on the set of subsets of G by conjugation. So for a subset S G, the
action is by (g, S) 7 gSg 1 = {gsg 1 | s S}. The stablizer GS of S under this action is usually
called the normalizer of S in G, denoted NG (S). So, NG (S) = {g G | gSg 1 = S}. In particular,
if H is a subgroup of G, we have its normalizer NG (H), and H E NG (H). In particular, H E G if
and only if G = NG (H). For an arbitrary subgroup H G, the number of distinct conjugates of
H in G is given by the formula
|GH| = [G : NG (H)].

1.5.8. The symmetric group Sn acts faithfully on the numbers {1, . . . , n} by the very definition of
Sn ! This action is transitive so that there is a single orbit, and each point stabilizer is isomorphic
to the symmetric group Sn1 .

1.5.9. The dihedral group D4 of symmetries of the square acts naturally on the set of vertices
(labelled {1, 2, 3, 4} say) of the square. This yields a homomorphism : D4 S4 which is in fact
injective, hence D4 embeds as a subgroup of S4 . In other words, the action of D4 on the vertices
of the square is faithful.

Given a finite G-set X, we have the class equation


X
|X| = |F ix(X)| + [G : Gx ]
x

where the sum is over a set of representatives x of the non-trivial orbits. Applying this in particular
to the conjugation action of G on itself, we obtain
X
|G| = |Z(G)| + [G : Gx ]
x

summing over a set of representatives for the conjugacy classes of non-central elements.
Let p be a prime. We call a finite group G a p-group if the order of G has order a power of p.
(More generally, a possibly infinite group G is called a p-group if every element of G has order a
power of p). Now if X is a G-set and x X is not a fixed point, then Gx is a proper subgroup of
G, so [G : Gx ] is a divisor of |G| strictly greater than 1. So if G is a finite p-group, [G : Gx ] is a
proper power of p, so [G : Gx ] 0 (mod p). Taking both sides of the class equation above modulo
p, we deduce:

1.5.10. Lemma. If G is a finite p-group and X is a finite G-set, then

|F ix(X)| |X| (mod p).

Applying this to the conjugation action of G on itself, this shows in particular:

1.5.11. Corollary. If G is a finite p-group, then Z(G) 6= {1}.

Proof. By the lemma, |Z(G)| |G| 0 (mod p). But 1 Z(G), so |Z(G)| 1. But 1 6 0
(mod p) so we cannot have |Z(G)| = 1.
You should be able to prove now that if |G| = p2 for a prime p, then G is Abelian.
22 CHAPTER 1. GROUPS

1.5.12. Lemma. Let G be a finite Abelian group and p be a prime dividing the order of G. Then,
there exists x G with o(x) = p.

Proof. Say |G| = pm. If m = 1, then G = Cp and the result is obvious. If m > 1, take 1 6= x G.
If o(x) = pk for some k, then xk has order p and were done. So we may assume that p - o(x).
Then, p||G/hxi|, so by induction there exists an element y G such that y / hxi but y p hxi.
Then, y = 1 for some h coprime to p (indeed, h divides the order of x). So the order of y h
ph

divides p, so either equals p, as required, or 1. But in the latter case, we have that y h = 1. Write
1 = ah + bp for integers a, b. Then, y = y ah+bp = y bp which lies in hxi as y p hxi. Hence, y hxi,
a contradiction.

Cauchys theorem. If G is a finite group and p is a prime dividing the order of G, then there
exists an element x G with o(x) = p.

Proof. Let |G| = pm. If m = 1, the conclusion is obvious. Otherwise, we have the class equation:
X
|G| = |Z(G)| + [G : CG (x)].
x

For each x in the sum, |CG (x)| < |G|, so if p||CG (x)| for any such x we get the conclusion by
induction. Hence, p - |CG (x)| for any such x, in other words, p divides each [G : CG (x)]. Since p
also divides |G|, we deduce that p divides |Z(G)|. Then Z(G) contains an element of order p by
the lemma.

1.6 The Sylow theorems


Now suppose that G is a finite group of order pn k for a prime p with (p, k) = 1. A subgroup
H G with |H| = pn is called a Sylow p-subgroup. Equivalently, H is a Sylow p-subgroup of G if
H is a p-group and p - [G : H].

1.6.1. Lemma. If H is a p-subgroup of G, then [NG (H) : H] [G : H] (mod p). In particular, if


p | [G : H] (i.e. H is not a Sylow p-subgroup of G) then NG (H) 6= H.

Proof. Let H act on G/H by left multiplication. Then, aH F ix(G/H) if and only if HaH = aH
which is if and only if a1 Ha = H which is if and only if a NG (H). Hence, [NG (H) : H] =
|F ix(G/H)|. Now apply Lemma 1.5.10.

First Sylow theorem. Let H < G be a p-subgroup of G that is not a Sylow p-subgroup. Then,
there exists a p-subgroup K G with H C K and |K| = p|H|. In particular, H can be embedded
into a Sylow p-subgroup of G.

Proof. Since p|[G : H], we have by the preceeding lemma that H 6= NG (H), H C NG (H) and
NG (H)/H is a group of order divisible by p. Pick any x NG (H)/H of order p. Then, the
pre-image in NG (H) of hxi is a subgroup of order p|H| which contains H as a normal subgroup.

Second Sylow theorem. All Sylow p-subgroups of G are conjugate.

Proof. Let H, K be two Sylow p-subgroups of G. Let H act on G/K by left multiplication. Then,
aK F ix(G/K) if and only if HaK = aK which is if and only if a1 Ha K, i.e. a1 Ha = K
since they have the same order. So to prove that H and K are conjugate in G, it suffices to show
that |F ix(G/K)| =6 0. But |F ix(G/K)| |G/K| 6 0 (mod p) by Lemma 1.5.10.
1.6. THE SYLOW THEOREMS 23

Third Sylow theorem. Let m be the number of distinct Sylow p-subgroups of G. Then, m 1
(mod p) and m||G|.
Proof. Let H be a Sylow p-subgroup (exists by the first Sylow theorem). Applying the second
Sylow theorem, m is the number of conjugates of H, which is [G : NG (H)]. This gives that m||G|.
Now let X be the set of all Sylow p-subgroups of G and let H act on X by conjugation. Then,
|F ix(X)| m (mod p). It therefore suffices to show that |F ix(X)| = 1.
Now, K F ix(X) if and only if hKh1 = K for all h H, i.e. H NG (K). But H NG (K)
if and only if H is a Sylow p-subgroup of NG (K). By the second Sylow theorem, K is the unique
Sylow p-subgroup of NG (K), so this is if and only if H = K. Hence, F ix(X) = {H}.
We give some examples to illustrate applying Sylow theorems.
1.6.2. Let |G| = pq with p > q prime and G not Abelian. Then, q|p 1 and G = ha, b | aq = bp =
1, aba1 = br i for some 1 < r < p with rq 1 (mod p).
Proof. Let K = hbi be a Sylow p-subgroup of G. We have that np , the number of Sylow p-
subgroups, divides pq and is congruent to 1 mod p. Since p > q, this means that np = 1. Hence,
K must be normal in G.
Now let H = hai be a Sylow q-subgroup of G. Since G is not Abelian, we cannot have that
H C G. (If it was then for any h H, k K we would have that hkh1 k 1 H K = {1}. Hence
hk = kh for all h H, k K which implies that G = H K is Abelian.) One deduces that the
number of Sylow q-subgroups of G must be p, and that q|p 1 by the third Sylow theorem.
Now, since K C G, aba1 K so equals br for some 1 < r < p (cannot have r = 1 since G is not
q
Abelian). Next, aq baq = br = b implies that rq 1 (mod p). Weve now shown that G satisfies
the given relations. Hence, there exists an epimorphism from the group with the given relations to
G, which is injective since you easily check that any group satisfying the given relations has order
at most pq.

1.6.3. As a special case of the previous example, we get that a group of order 2p for p an odd
prime is either isomorphic to C2p or to Dp .
1.6.4. There is no simple group of order 12.
Proof. Let G be a simple group of order 12. Consider n2 , the number of Sylow 2-subgroups. It is
either 1 or 3. In the former case, G has a normal Sylow 2-subgroup so is not simple. Hence, n2 = 3.
Now, G acts on the Sylow 2-subgroups by conjugation, giving a non-constant homomorphism
G S3 . The kernel is a proper normal subgroup of G since |G| = 12, |S3 | = 6.
Here is a table listing all groups of order 12:
Order G
1 {1}
2 C2
3 C3
4 V4 = C2 C2 , C4
5 C5
6 C2 C3 = C6 , D3
= S3
7 C7
8 C8 , C4 C2 , C2 C2 C2 , D4 , Q8
9 C9 , C3 C3
10 C2 C5 , D5
11 C11
12 C12 , C4 C3 , C2 C2 C3 , D6 , A4 , T
For the definitions of the groups T of order 12 and the quaternion group Q8 of order 8, see
Hungerford (or section 2.1 for Q8 ).
24 CHAPTER 1. GROUPS

1.7 Semidirect products


In (1.6.2), we showed that if G is a non-Abelian group of order |G| = pq with p > q prime, then
q|p 1 and
G = ha, b | aq = bp = 1, aba1 = br i
for some 1 < r < p with rq 1 (mod p). We stopped short of showing that there actually exists
such a group G for all possible choices of p, q, r. However, we showed in the proof that G contained
a normal subgroup K = hbi = Cp . The quotient group G/K is isomorphic to Cq . So you can think
of G as being built by glueing the group Cq on top of the group Cp . In other words, G is an
extension of Cq by Cp .
For the general definition, let H, K be two groups. We call a group G an extension of H by
K if G contains a copy of K as a normal subgroup and G/K = H. We can represent this by the
following exact sequence:
1KGH1
(the precise meaning of this will be given later when studying modules.) It is important to change
the point of view and ask: given H and K what possible extensions G of H by K can we form?
For instance, G = H K is an example of an extension of H by K: the kernel of the canonical
projection G H is isomorphic to K. (Equally well, you could call G = H K an extension of K
by H.) But this extension H K is not really very interesting! How can we build more complicated
extensions? We discuss here one important class of extensions which are not too complicated to
understand, known as semidirect products. This is really a generalization of the construction of
the direct product there are in general other extensions which are not semidirect products.
So now let H and K be given groups and suppose we also have a group homomorphism

: H Aut K, h 7 h

(where Aut K is the group of all automorphisms of the group K). Define G = K o H, the
external semidirect product, to be the group equal to the Cartesian product K H as a set, with
multiplication defined by
(k, h)(k 0 , h0 ) = (kh (k 0 ), hh0 ).
You need to check that this really is a multiplication making K H into a group! For instance,
(k, h)1 = (h1 (k 1 ), h1 ).
Now, (k, 1)(k 0 , 1) = (kk 0 , 1), so the set K 0 = {(k, 1) | k K} is a subgroup of G isomorphic to
K. The map G H determined by projection onto the second coordinate is a surjective group
homomorphism, with kernel K 0 . This verifies that G has a copy of K as a normal subgroup and
the quotient group is isomorphic to H. Hence: the semidirect product G is an extension of H by
K. Observe moreover that:

(1, h)(k, 1)(1, h1 ) = (h (k), h)(1, h1 ) = (h (k), 1).

This shows that the homomorphism : H Aut K can be recovered from the semidirect product:
h is precisely the homomorphism determined by conjugating by the element (1, h) of G.
Actually, K o H is a very special sort of extension: the map : G H determined by
projection onto the second coordinate is split. This means that there is a group homomorphism
: H G called a splitting such that = idH , namely, the map h 7 (1, h). In fact an
extension of H by K is a semidirect product if and only if such a splitting map exists; more general
extensions need not have a splitting map (i.e. need not be semidirect products).
This last observation is the key to recognizing whether or not a given group G is isomorphic to
a semidirect product of groups H and K. Indeed, given a group G and subgroups H and K, we
say that G is the internal semidirect product of H and K if
(1) K C G;
(2) H K = {1};
(3) G = KH = {kh | k K, h H}.
1.8. SOLVABLE AND NILPOTENT GROUPS 25

If G is the internal semidirect product of H and K, we define a map : H Aut K by setting


h (k) = hkh1 for all k K. Then, G is isomorphic to the external semidirect product K o H.
Conversely, any external semidirect product K o H is the internal semidirect product of {1} H
and K {1}.
Take the very special case of external semidirect product where h = idK for each h H. Then,
K o H is exactly the same as the definition of the direct product K H: so semidirect products
are a generalization of direct products. The previous paragraph gives in this special case that a
group G is isomorphic to the direct product of subgroups H, K G if H C G, K C G, H K = {1}
and G = HK.
We return to the example (1.6.2). So we have primes p > q with q|p 1 and 1 < r < p with
rq 1 (mod p). We wish to prove that there exists a non-Abelian group G of order pq. Let
K = Zp , H = Zq (both written additively) and define a map

: H Aut K
q
as follows. Define 1 : Zq Zq , [i] 7 [ir ]. Note (1 )q ([i]) = [ir ] = [i1 ] = [i]. This shows
in particular that 1 is an automorphism of Zq . Now we obtain a well-defined map : H
Aut K, [n] 7 (1 )n . Finally, define G = K o H. Its an extension of H by K, you just need to
check that it is non-Abelian to complete the proof of the existence of such a group.

1.8 Solvable and nilpotent groups


In the previous section, we started thinking of extensions of groups. Now we take the idea further.
Define a subnormal series of a group G to be a chain of subgroups

G = G0 G1 Gn Gn = {1}

such that Gi E Gi1 for each i = 1, . . . , n. The factors of the subnormal series are the groups
Gi1 /Gi for i = 1, . . . , n. The length of the series is the number of non-trivial factors. A normal
series is a subnormal series in which in addition each Gi is a normal subgroup of G.
The basic idea to keep in mind is that the structure of the factors in a subnormal series should
tell you something about the structure of the group G. In this section, we focus on two special
classes of group defined in terms of certain subnormal series: solvable and nilpotent groups.
So now let G be a group and H, K be subgroups. Define

[H, K] = h[h, k] | h H, k Ki

where [h, k] denotes the commutator hkh1 k 1 . Note that [H, K] = [K, H], and that [H, K] = {1}
if and only if every element of H commutes with every element of K.

1.8.1. Lemma. If H E G then [H, G] E G, [H, G] H and H/[H, G] is Abelian.

Proof. Exercise.
The subgroup G0 := [G, G] of G is called the commutator subgroup of G. By the lemma, it is
a normal subgroup of G and G/G0 is Abelian. Indeed, given any normal subgroup N E G such
that G/N is Abelian, we have for any g1 , g2 G that [g1 N, g2 N ] = N , hence [g1 , g2 ] N , hence
G0 N . This shows that G0 is the unique smallest normal subgroup of G whose associated quotient
group is Abelian. The group G/G0 is sometimes called the Abelianization of G (Abelianization is
a functor!).
Now, define G(0) = G and inductively set G(i) = [G(i1) , G(i1) ], the commutator subgroup of
(i1)
G . So, G(i) is the unique smallest normal subgroup of G(i1) with Abelian quotient. It follows
by induction on i that in fact G(i) is a normal subgroup even of G. Proof: by induction, G(i1) is
normal in G and so gG(i) g 1 is the unique smallest normal subgroup of gG(i1) g 1 = G(i1) with
Abelian quotient hence coincides with G(i) .
26 CHAPTER 1. GROUPS

We therefore obtain a descending chain of normal subgroups of G:

G = G(0) G(1) G(2) . . .

called the derived series of G. The group G is called solvable if for some n >> 1 we have that
G(n) = {1}. Then, the derived series is in fact a normal series of G. The basic properties of
solvable groups are given in the following lemma:

Properties of solvable groups. (1) If G is solvable, then so is every subgroup.


(2) Given N E G, G is solvable if and only if both N and G/N are solvable.
(3) G is solvable if and only if G has a subnormal series with Abelian factors.
(4) Finite direct products of solvable groups are solvable.
(5) A semidirect product of two solvable groups is solvable.

Proof. (1) Let H G with G solvable. Show by induction on n that H (n) G(n) . Since
G(n) = {1} for some n we get that H (n) = {1} for some n, so H is solvable.
(2) Induction shows that (G/N )(n) = G(n) N/N . It follows easily using (1) too that if G is
solvable, so are both N and G/N . Conversely, if N and G/N are solvable, there exists n such that
G(n) N and m such that N (m) = {1}. Then, G(n+m) N (m) = {1} so G is solvable too.
(3) If G is solvable, the derived series is a subnormal series with Abelian quotients. Conversely,
suppose G has a subnormal series with Abelian quotients; we prove G is solvable by induction on
the length of the subnormal series. If the series has length 1, the result is clear. Else, consider the
first subgroup G1 C G in the subnormal series different from G. It has a subnormal series with
Abelian quotients of length one less than that of G, so is solvable by induction. Also, G/G1 is
Abelian so solvable. Hence by (2), G is solvable.
(4) Apply (2) and induction on the number of direct factors.
(5) Apply (2).
For example, the dihedral group Dn is an internal semidirect product of its rotation subgroup
Cn (which is normal of index two) and any subgroup generated by a reflection. So (5) implies that
Dn is solvable. You should also easily be able to prove directly that the symmetric group S4 is
solvable. On the other hand, no non-Abelian simple group is solvable, so no An (n 5) is solvable.
In particular, this shows that Sn for n 5 is not solvable.

We turn now to nilpotent groups. The definition is very similar to the definition of a solvable
group, but involves a different descending chain of subgroups : set G0 = G and inductively define
Gi = [G, Gi1 ]. By Lemma 1.8.1, Gi is a normal subgroup of G and Gi1 /Gi is Abelian. We
obtain a descending chain of normal subgroups of G

G = G 0 G1 . . .

called the descending central series. The group G is nilpotent if Gn = {1} for some n >> 1, in
which case the descending central series is in fact a normal series of G with Abelian factors. By
property (3) of solvable groups, we have at once that nilpotent groups are solvable.

Properties of nilpotent groups. (1) If G is nilpotent, so is every subgroup.


(2) If G is nilpotent and N E G, then G/N is nilpotent.
(3) A non-trivial group G is nilpotent if and only if Z(G) 6= {1} and G/Z(G) is nilpotent.
(4) Direct products of finitely many nilpotent groups are nilpotent.
(5) If G is nilpotent and H < G is a proper subgroup, then H 6= NG (H).

Proof. (1) Show by induction on n that H n Gn for a subgroup H G.


(2) Show by induction on n that (G/N )n = Gn N/N .
(3) If G is nilpotent, consider the last non-trivial subgroup C in the descending central series
of G. Then, [G, C] = {1} so that {1} < C Z(G), giving that Z(G) 6= {1}. Then by (2), G/Z(G)
1.9. JORDAN-HOLDER THEOREM 27

is nilpotent. Conversely, suppose that Z(G) 6= {1} and that G/Z(G) is nilpotent. Then, using the
result proved in (2), for some n we have that Gn Z(G). Then, Gn+1 [G, Z(G)] = {1} so G is
nilpotent.
(4) Let me prove this just for the direct product of two nilpotent groups (the general case
following by induction from this). Consider G H with G, H nilpotent. Then, we can find some
n >> 1 such that both Gn = {1} and H n = {1} Then, (G H)n = Gn H n = {1} too.
(5) Choose i so that Gi+1 H but Gi 6 H. Then, [Gi , H] [Gi , G] = Gi+1 H. In other
words, Gi NG (H). Since Gi 6 H, this shows that NG (H) is strictly larger than H.
Now we can give an example of a solvable group that is not nilpotent: the dihedral group Dn .
You should be able to prove that Z(Dn ) = {1} if n is odd or {1, t} if n is even, where t is the 180
degree rotation. So if n is odd, Dn is certainly not nilpotent by (3), but Dn is always solvable as
its a semidirect product of two cyclic groups. If n is even, Dn /Z(Dn ) = Dn/2 , and using this and
(3) you show that in fact Dn is nilpotent if and only if n is a power of 2.
We end with an important characterization of finite nilpotent groups in terms of Sylow p-
subgroups. (There is a similar, but more complicated, such characterization of finite solvable
groups called P. Halls theorem, which is really a generalization of the Sylow theorems for solvable
groups.)

Characterization of finite nilpotent groups. Let G be a finite group. Then, G is nilpotent if


and only if G is isomorphic to the direct product of its Sylow p-subgroups for all p.

Proof. Suppose G is the direct product of its Sylow p-subgroups for all p. Using property (4), to
show that G is nilpotent, we just need to show that every finite p-group is nilpotent. This follows
by induction on the order using property (3) and Corollary 1.5.11.
Conversely, suppose that G is nilpotent and take a prime p dividing the order of G. Let P be
a Sylow p-subgroup. If P = G, we are done: there is nothing to prove.
We claim that NG (NG (P )) = NG (P ). Indeed, P is the unique Sylow p-subgroup of NG (P ).
Take any x NG (NG (P )). Then, xP x1 is the unique Sylow p-subgroup of xNG (P )x1 = NG (P ).
Hence, xP x1 = P . Hence, x NG (P ). This proves the claim.
It follows from property (5) that in fact, NG (P ) = G, so that P is normal in G. This means
that G has a unique Sylow p-subgroup for each prime p dividing its order.
Now let p1 , . . . , pn be the primes dividing |G| and let Pi be the unique Sylow pi -subgroup. Define
a map : P1 Pn G, (g1 , . . . , gn ) 7 g1 . . . gn . To see that this is a group homomorphism,
it suffices to see that [Pi , Pj ] = {1} for i 6= j, so that elements of Pi and Pj pairwise commute. But
[Pi , Pj ] is a subgroup of both Pi and Pj , hence of Pi Pj which is trivial by Lagranges theorem.
Now suppose that we have (g1 , . . . , gn ) ker . Then, we have that g1 = g21 g31 . . . gn1 in G.
But g1 has order a power of p1 , whereas the order of the element on the right hand side is coprime
to p1 . This shows that g1 = 1. Similarly, each gi = 1. Hence, is injective.
Finally, is a surjection by considering orders. Hence, G is isomorphic to the direct product
of its Sylow p-subgroups for all p.

1.9 Jordan-Holder theorem


Recall the definition of a subnormal series of a group G from the previous section. Let

G = G0 G1 Gn = {1}

be a subnormal series. A one-step refinement means a subnormal series

G = G0 G1 Gi H Gi+1 Gn = {1}

for some subgroup H E Gi . A refinement of a subnormal series means a subnormal series obtained
by finitely many one-step refinements.
28 CHAPTER 1. GROUPS

Zassenhaus butterfly lemma. Fix a group G and subgroups A E A, B E B in G. Then,


A (A B ) E A (A B) and B (A B) E B (A B). Moreover,

A (A B)/A (A B )
= B (A B)/B (A B).

Proof. Set D = (A B)(A B ). Let : A A/A be the quotient map. The restriction of
to A B has kernel A B, so A B E A B. Similarly, A B E A B. Hence, D E A B.
Now define : A (A B) (A B)/D by setting (ax) = xD for any a A , x A B. We
need to check this is well-defined first. Suppose ax = a0 x0 for another a0 A , x0 A B. Then,
x0 x1 = a(a0 )1 A B D. Hence, x0 x1 D = D, i.e. x0 D = xD as required. Moreover, is a
group homomorphism, which you can check using just that A E A, and it is clearly surjective.
Now we calculate ker . For a A , x A B, we have that ax ker if and only if x D
if and only if ax A D = A (A B ). Hence, ker = A (A B ). So, A (A B ) E A (A B)
and A (A B)/A (A B ) = A B/D.
Similarly, B (A B) E B (A B) and B (A B)/B (A B)
= A B/D. The result follows.

We call two subnormal series G = G0 G1 Gn = {1} and G = H0 H1 Hm =


{1} equivalent if m = n and there exists a permutation w Sn such that Gi /Gi+1
= Hwi /Hwi+1
for each i = 0, . . . , n 1.

Schreier refinement lemma. Any pair of subnormal series for a group G have equivalent re-
finements.

Proof. Consider the subnormal series G = G0 G1 Gn = {1} and H = H0 H1


Hm = {1}. For any 1 i n and 1 j m, take A = Hj , A = Hj1 , B = Gi , B = Gi1
in the butterfly lemma. You deduce that there are subnormal series

Gi1 = Gi (Gi1 H0 ) Gi (Gi1 H1 ) Gi (Gi1 Hm ) = Gi . . .

and

Hj1 = Hj (G0 Hj1 ) Hj (G1 Hj1 ) Hj (Gn Hj1 ) = Hj . . .

These are both series with mn terms and

Gi (Gi1 Hj1 )/Gi (Gi1 Hj )


= Hj (Gi1 Hj1 )/Hj (Gi Hj1 ).

Hence we have constructed equivalent refinements.


A subnormal series G = G0 G1 Gn = {1} is called a composition series if for each
i = 1, . . . , n, Gi1 /Gi is a non-trivial simple group. Warning: a group may or may not possess a
composition series! Moreover, even if it possesses a composition series, it can have many different
composition series.
For example, consider the group G = S4 . Any group G1 starting a composition series must
have G/G1 simple, in other words, G1 must be a maximal normal subgroup of G. Now, S4 has
a unique maximal normal subgroup, namely, A4 . Hence, G1 = A4 . Now G2 must be a maximal
normal subgroup of A4 . The only possibility is V4 . Then, G3 must be a maximal normal subgroup
of V4 = C2 C2 . But this is Abelian and has three subgroups of order 2, each isomorphic to C2
which is simple. We deduce that S4 has exactly three different composition series, G = S4 > A4 >
V4 > C > {1} where C is one of h(12)(34)i, h(13)(24)i or h(14)(23)i. In all three cases, the factors
are C2 , C3 , C2 , C2 .
An easy induction argument shows that all finite groups have a composition series. (Pick a
maximal normal subgroup, possible since G has finitely many subgroups, then find a composition
series for that subgroup by induction.)
1.9. JORDAN-HOLDER THEOREM 29

Jordan-Holder theorem. If G is a group possessing a composition series, then all composition


series of G are equivalent.

Proof. Take two composition series. By the Schreier refinement lemma, they have equivalent
refinements. But refining a composition series simply adds trivial composition factors...
If G has a composition series, its length is called the composition length of G. This num-
ber, together with the isomorphism types of the factors appearing in any composition series, are
important invariants of a group G.
Its amusing to note a special case of the Jordan-Holder theorem: the fundamental theorem
of arithmetic asserting that every integer has a unique factorization into primes. This follows
from the Jordan-Holder theorem applied to the group Zn , because the composition factors of a
composition series of Zn are exactly the prime divisors of n. This is an example of cracking a nut
with a sledgehammer!

You might also like