You are on page 1of 79

Groups

Marco Mackaaij

2022/2023

Contents
1 The integers 2
1.1 Euclidean division . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 The Euclidean algorithm . . . . . . . . . . . . . . . . . . . . . 3
1.3 Prime factorization . . . . . . . . . . . . . . . . . . . . . . . . 7
1.4 The least common multiple . . . . . . . . . . . . . . . . . . . 9
1.5 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

2 Group theory 12
2.1 Groups: definition and examples . . . . . . . . . . . . . . . . . 12
2.1.1 Congruences . . . . . . . . . . . . . . . . . . . . . . . . 16
2.1.2 The symmetric group . . . . . . . . . . . . . . . . . . . 19
2.1.3 The dihedral group . . . . . . . . . . . . . . . . . . . . 25
2.1.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.2 Subgroups, homomorphisms and the direct product . . . . . . 30
2.2.1 Subgroups . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.2.2 Homomorphisms . . . . . . . . . . . . . . . . . . . . . 32
2.2.3 The direct product . . . . . . . . . . . . . . . . . . . . 37
2.2.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . 43
2.3 Generators, cyclic groups, order and cosets . . . . . . . . . . . 45
2.3.1 Generators and cyclic groups . . . . . . . . . . . . . . 45
2.3.2 Order . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
2.3.3 Cosets . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
2.3.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . 54
2.4 Normal subgroups, quotient groups and homomorphism and
isomorphism theorems . . . . . . . . . . . . . . . . . . . . . . 56

1
2.4.1 Normal subgroups . . . . . . . . . . . . . . . . . . . . 57
2.4.2 Quotient groups . . . . . . . . . . . . . . . . . . . . . . 60
2.4.3 Homomorphism and isomorphism theorems . . . . . . . 63
2.4.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . 67
2.5 Group actions, conjugacy classes, semidirect products . . . . . 69
2.5.1 Group actions . . . . . . . . . . . . . . . . . . . . . . . 69
2.5.2 Conjugacy classes . . . . . . . . . . . . . . . . . . . . . 73
2.5.3 Semidirect product . . . . . . . . . . . . . . . . . . . . 76
2.5.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . 78

1 The integers
1.1 Euclidean division
We start with a well-known fundamental result, called the Euclidean division
lemma (aka division with remainder lemma).

Lemma 1.1.1. Let a, b ∈ Z, such that b > 0. Then there are unique q, r ∈ Z
(the quotient and the remainder of the division of a by b, respectively) such
that 0 ≤ r < b and
a = qb + r (1)

Proof. First assume that a ≥ 0 and use induction w.r.t. a. For a = 0, the
unique solution of (1) is q = r = 0, of course. Now assume that a > 0.
By applying the induction hypothesis to a − 1, we see that there are unique
q ′ , r′ ∈ Z such that 0 ≤ r′ < b and

a − 1 = q ′ b + r′ .

To show that this implies the desired result for a, we consider two cases:
r′ = b − 1 and r′ < b − 1. When r′ = b − 1, we get a − 1 = q ′ b + b − 1, so

a = (q ′ + 1)b,

showing that we can take q := q ′ + 1 and r := 0. When r′ < b − 1, we can


take q := q ′ and r := r′ + 1 (note that r′ < b − 1 implies that r < b in this
case) because

a = (a − 1) + 1 = (q ′ b + r′ ) + 1 = q ′ b + (r′ + 1).

2
This proves the lemma for a ≥ 0.
Now assume that a < 0. Then −a > 0, so there are unique q ′ , r′ ∈ Z such
that 0 ≤ r < b and
−a = q ′ b + r′ .
When r′ = 0, we have −a = q ′ b, so a = (−q ′ )b, showing that we can take
q := −q ′ and r := 0. When 0 < r′ < b, we can take q := −q ′ − 1 and
r := b − r′ because

a = −q ′ b − r′ = (−q ′ − 1)b + (b − r′ ).

This proves existence of the pair q, r in the lemma.


To prove uniqueness, suppose that there are q, q ′ , r, r′ ∈ Z such that
0 ≤ r, r′ < b and

a = qb + r;
a = q ′ b + r′ .

We need to show that q = q ′ and r = r′ . If q = q ′ , then r = a − qb =


a − q ′ b = r′ as well, and we are done. Now suppose that q > q ′ . The
equalities qb + r = a = q ′ b + r′ imply that

(q − q ′ )b = r′ − r.

But this leads to a contradiction: On the one hand, there is the inequality
(q − q ′ )b > b because q > q ′ by assumption (and b > 0), and on the other
hand, there is the inequality r′ − r < b because r′ < b and r ≥ 0. Therefore,
the assumption q > q ′ was false. The assumption q < q ′ can not be true
either, because it leads to a similar contradiction. This proves uniqueness of
q, r in the lemma.

1.2 The Euclidean algorithm


The Euclidean algorithm can be used to determine the greatest common
divisor of two integers. Before we explain how it works, let us briefly recall
the relevant definitions.

Defininition 1.2.1. Let a, b ∈ Z. If there is an integer q such that a = qb,


we say that d divides a, or that d is a divisor of a, or that b is a multiple of
a. (Notation: d | a.)

3
Example 1.2.2. 3 | 15, 0 | 0, 1 | −1, 3 ∤ 7
The following properties are immediate:
c | b ∧ b | a =⇒ c | a,
b | a ∧ b | a′ =⇒ b | a ± a′ ,
b | 0, for all b ∈ Z,
1 | a, for all a ∈ Z,
0 | a ⇐⇒ a = 0,
b | a ⇐⇒ |b| | |a|,
b | a ∧ a ̸= 0 ⇐⇒ |b| ≤ |a|.
Note that the last property implies that any non-zero integer has only finitely
many divisors.
Defininition 1.2.3. Suppose that a and b are not both equal to zero. Their
greatest common divisor is the greatest positive integer d such that d | a and
d | b. (Notation: gcd(a, b).) We say that a and b are coprime if gcd(a, b) = 1.
Note that gcd(0, b) = |b|, for all b ∈ Z\{0}, and gcd(a, b) = gcd(|a|, |b|)
for all a, b ∈ Z such that at least one of them is non-zero.
Example 1.2.4. gcd(3, 15) = 3, gcd(8, 12) = 4, gcd(4, 9) = 1.
Let a, b, q, r ∈ Z such that 0 ≤ r < b and a = qb + r, as in Lemma 1.1.1.
Lemma 1.2.5. We have
gcd(a, b) = gcd(b, r).
Proof. If d | a and d | b, then d | a − qb = r. Conversely, if d | b and d | r,
then d | qb + r = a. This shows that the set of common divisors of a and
b coincides with the set of common divisors of b and r, which implies the
lemma.
Given a, b ∈ Z, such that at least one of them is non-zero, one can de-
termine gcd(a, b) by the Euclidean algorithm. Define a sequence of integers
r0 > r1 > r2 . . . ≥ 0 recursively by r0 = |a|, r1 = |b| and, for n ≥ 2,
rn−1 = qn rn + rn+1 ,
where rn+1 is the remainder of division of rn−1 by rn . Since the sequence of
non-negative integers ri is strictly decreasing, there must be an N ∈ N0 such
that rN = 0.

4
Proposition 1.2.6. With the above notation, we have

gcd(a, b) = rN −1 .

Proof. In the first place, note that

gcd(a, b) = gcd(|a|, |b|) = gcd(r0 , r1 ).

By repeatedly applying Lemma 1.2.5, we see that

gcd(r0 , r1 ) = gcd(r1 , r2 ) = . . . = gcd(rN −1 , 0) = rN −1 ,

which proves the proposition.


Example 1.2.7. Let a = r0 = 1057 and b = r1 = 315. Then

1057 = 3 · 315 + 112


315 = 2 · 112 + 91
112 = 1 · 91 + 21
91 = 4 · 21 + 7
21 = 3 · 7 + 0.

Thus gcd(1057, 315) = 7.


Proposition 1.2.8. Let a, b ∈ Z, such that at least one of them is non-zero,
and put d := gcd(a, b). There exist x, y ∈ Z such that

xa + yb = d

Proof. Keeping the notation from above, we define inductively two sequences
of integers x0 , x1 , x2 , . . . and y0 , y1 , y2 , . . . such that

xn a + yn b = rn

for all n. For n = 0, we have r0 = |a| = ±a, so we can take x0 = ±1, y0 = 0.


Similarly, we can take x1 = 0, y1 = ±1 because r1 = |b| = ±b. If n ≥ 1 and
rn ̸= 0, then we can use induction and subtract qn times

xn a + yn b = rn

from
xn−1 a + yn−1 b = rn−1 .

5
Since rn−1 − qn rn = rn+1 , this yields
(xn−1 − qn xn )a + (yn−1 − qn yn )b = rn+1 .
Thus we can take xn+1 := xn−1 − qn xn , yn+1 := yn−1 − qn yn . Recall that
rN −1 = d (with N such that rN = 0), so
xN −1 a + yN −1 b = d.

Example 1.2.9. As in Example 1.2.7, let a = 1057 and b = 315. Then we


get
1 · 1057 + 0 · 315 = 1057
0 · 1057 + 1 · 315 = 315
1 · 1057 + (−3) · 315 = 112
(−2) · 1057 + 7 · 315 = 91
3 · 1057 + (−10) · 315 = 21
(−14) · 1057 + 47 · 315 = 7.
Corollary 1.2.10. Let a, b, d ∈ Z.
a) We have
gcd(a, b) = 1 ⇐⇒ ∃ x, y ∈ Z : xa + yb = 1.

b) If d | ab and gcd(a, d) = 1, then d | b.


c) If d | a and d | b, then d | gcd(a, b).
Proof. a) If gcd(a, b) = 1, then there exist x, y ∈ Z such that xa + yb = 1
by Proposition 1.2.8.
Conversely, suppose that there exist x, y ∈ Z such that xa + yb = 1
and let d := gcd(a, b). The fact that d is a common divisor of a and b
implies that d | xa + yb = 1, thus d = 1.
b) Suppose that d | ab and gcd(a, d) = 1. By Proposition 1.2.8, there exist
x, y ∈ Z such that xa + yd = 1. This implies that
d | xab + ydb = (xa + yd)b = b.

c) By Proposition 1.2.8, there exist x, y ∈ Z such that xa + yb = gcd(a, b).


If d | a and d | b, then d | xa + yb, so d | gcd(a, b).

6
1.3 Prime factorization
Recall the definition of a prime number.
Defininition 1.3.1. A prime number (or just a prime) is an integer p whose
only divisors are 1 and p.
Lemma 1.3.2. Let p be a prime and b1 , b2 , . . . , bm ∈ Z for some m ∈ N. If
p | b1 b2 · · · bm , then there is an index i ∈ {1, 2, . . . , m} such that p | bi .
Proof. We use induction w.r.t. m. For m = 1, the result is immediate.
Now, suppose that m > 1. Since p is prime and gcd(b1 , p) | p, we must have
gcd(b1 , p) = 1 or gcd(b1 , p) = p. If gcd(b1 , p) = 1, then Corollary 1.2.10 b) im-
plies that p | b2 · · · bm and thus, by induction, there is an index i ∈ {2, . . . , m}
such that p | bi . If gcd(b1 , p) = p, then p | b1 , because p is prime.
The following theorem is called the prime factorization theorem.
Theorem 1.3.3. Every a ∈ N>1 can be written as
a = p1 p2 · · · pt ,
for some t ∈ N, where the pi are prime numbers. Moreover, these prime
factors are unique up to reordering (we say that the prime factorization is
essentially unique).
Proof. We first prove existence, using induction w.r.t. a. If a = 2, then t = 1
and a = p1 = 2. Now, let a > 2. If a is prime, then again take t = 1 and
a = p1 . If a is not prime, we can write a = bc, for some divisors 1 < b, c < a.
By induction, each divisor can be factored into prime numbers, so the same
holds for a by multiplying both factorizations.
Uniqueness we also prove with induction on a. When a = 2, uniqueness
is clear. Now, let a > 2 and suppose that it has two prime factorizations:
a = p1 p2 · · · pt = q1 q 2 · · · q u .
By Lemma 1.3.2, there is an i ∈ {1, 2, . . . , u} such that p1 | qi . Since both are
prime, this means that p1 = qi . Dividing both factorizations by this prime,
we get
a
= p2 p3 · · · pt = q1 q2 · · · qi−1 qi+1 · · · qu .
p1
Since a/p1 < a, the induction hypothesis holds for a/p1 , so the two sets of
prime factors {p2 , p3 , . . . , pt } and {q1 , . . . , qi−1 , qi+1 , . . . , qu } coincide, which
proves uniqueness of the factorization of a.

7
We denote the number of times a prime p appears in the prime factor-
ization of a by νp (a) (the multiplicity of p in the prime factorization of a),
where we also use the convention that νp (1) = 0, so that we can write
a= p νp (a) .
Y

p prime

This product runs over the infinite set of all prime numbers, but almost all
factors are equal to 1 because νp (a) = 0 for almost all prime numbers p.
By convention, we put νp (1) = 0 for all prime numbers p (although, strictly
speaking, the number 1 does not have a prime factorization).
Example 1.3.4. The prime factorizations of :
600 = 6 · 100 = 2 · 3 · 22 · 52 = 23 · 3 · 52 .
Thus we see that
3,
if p = 2;




1,
if p = 3;


νp (600) =




2, if p = 5;
0, if p ̸= 2, 3, 5.

Corollary 1.3.5. For p a prime number and a, b ∈ N, we have


νp (ab) = νp (a) + νp (b)
Proof. This is an immediate consequence of Theorem 1.3.3:
ab = p νp (a) p νp (b) = p νp (a)+νp (b) .
Y Y Y

p prime p prime p prime

Corollary 1.3.6. For any c, d ∈ N, we have


d | c ⇐⇒ νp (d) ≤ νp (c) for all prime numbers p.
Proof. If d | c, then c = qd for some q ∈ N. Corollary 1.3.5 then implies that
νp (d) = νp (c) − νp (q) ≤ νp (c) for all prime numbers p.
Conversely, suppose that νp (d) ≤ νp (c) for all prime numbers p. Then
p νp (c) = p νp (c)−νp (d)+νp (d) = p νp (c)−νp (d) p νp (d) ,
Y Y Y Y

p prime p prime p prime p prime

so c = qd, where q := p νp (c)−νp (d) .


Q
p prime

8
Corollary 1.3.7. For any a, b ∈ N, we have
gcd(a, b) = p min{νp (a),νp (b)} .
Y

p prime

Proof. Let d ∈ N. By Corollary 1.3.6, we have


d | a ∧ d | b ⇐⇒ νp (d) ≤ νp (a) ∧ νp (d) ≤ νp (b) for all primes p.
This implies that νp (gcd(a, b)) ≤ min{νp (a), νp (b)} for all prime numbers p,
which, again by Corollary 1.3.6, means that
gcd(a, b) | p min{νp (a),νp (b)} .
Y

p prime

Since the latter product is a common divisor of a and b, yet again by the
same corollary, it must be equal to gcd(a, b).
Example 1.3.8. Compare the following calculation to the one in Exam-
ple 1.2.7:
gcd(315, 1057) = gcd(32 · 5 · 7, 7 · 151) = 7.

1.4 The least common multiple


This is a very small subsection, but its content will be needed in the sequel.
Defininition 1.4.1. Let a, b ∈ Z. The least common multiple of a and b
is the smallest positive integer m such that a | m and b | m. (Notation:
lcm(a, b).)
Proposition 1.4.2. Let a, b ∈ Z. Then
a) any common multiple of a and b is also a multiple of lcm(a, b);
b) if a, b > 0, then
lcm(a, b) = p max{νp (a),νp (b)} .
Y

p prime

Proof. For a starter, suppose that a, b > 0 and let m ∈ N. Corollary 1.3.6
implies that
a | m ∧ b | m ⇐⇒ νp (a) ≤ νp (m) ∧ νp (b) ≤ νp (m) for all primes p
⇐⇒ max{νp (a), νp (b)} ≤ νp (m) for all primes p.

9
By the same corollary, we thus see that

p max{νp (a),νp (b)} | lcm(a, b).


Y

p prime

Since the former product is a common multiple of a and b, yet again by the
same corollary, we see that it must be equal to lcm(a, b). This shows that
both a) and b) hold for a, b > 0.
The general case of a) can be reduced to the previous case by using the
abolute value of a and b, except when ab = 0, in which case the result is
immediate.

Example 1.4.3. By the factorizations given in Example 1.3.8, we see that

lcm(315, 1057) = 32 · 5 · 7 · 151 = 47565.

Corollary 1.4.4. Let a, b ∈ Z, such that at least one of them is non-zero.


Then
|ab| = gcd(a, b) · lcm(a, b).
In particular, if gcd(a, b) = 1, then |ab| = lcm(a, b).

Proof. Without loss of generality, we can assume that a, b > 0. Note that
νp (a) + νp (b) = min{νp (a), νp (b)} + max{νp (a), νp (b)}. By Proposition 1.4.2,
this implies that

ab = p νp (a)+νp (b)
Y

p prime

= p min{νp (a),νp (b)} p max{νp (a),νp (b)}


Y Y

p prime p prime
= gcd(a, b) · lcm(a, b).

Example 1.4.5. By Example 1.3.8 again, we see that


315 · 1057
lcm(315, 1057) = = 47565.
7
Remark 1.4.6. Both the greatest common divisor and the least common
multiple can be defined for any finite number of integers and the results in
these subsections all generalize in a straightforward manner.

10
1.5 Exercises
1. For any non-zero a, b ∈ Z, prove that a/ gcd(a, b) and b/ gcd(a, b) are
coprime.

2. Compute gcd(4511, 1625) and determine x, y ∈ Z such that

x · 4511 + y · 1625 = gcd(4511, 1625).

3. Let a, b, c ∈ Z. Show that

a) gcd(a, b) = gcd(a, c) = 1 =⇒ gcd(a, bc) = 1;


b) gcd(a, b) = 1 ∧ a | c ∧ b | c =⇒ ab | c.

4. Let a, b, c ∈ N0 . Show that

c · gcd(a, b) = gcd(ca, cb).

5. Let a, b ∈ N0 .

a) Let r be the remainder of a divided by b. Prove that 2r − 1 is the


remainder of 2a − 1 divided by 2b − 1.
b) Prove that
2b − 1 | 2a − 1 ⇐⇒ b | a.
c) Prove that  
gcd 2a − 1, 2b − 1 = 2gcd(a,b) − 1.

d) Are a), b) and c) also true when 2 is replaced by an arbitrary


positive integer c > 2.

6. Let p ∈ Z>1 . Prove that p is prime if and only if p has no divisor d



satisfying 1 < d ≤ p.

7. Prove that there are infinitely many prime numbers.

8. Let a, n ∈ N>1 . Prove that

an − 1 is a prime =⇒ a = 2 ∧ n is a prime.

9. Let a, b ∈ Z and n ∈ N0 .

11
a) Prove that a − b | an − bn .
b) Suppose that n is odd. Prove that a + b | an + bn .

10. Let p > 3 be a prime. Show that 24 | p2 − 1.

11. Let x ∈ Q>0 . Prove that for each prime number p there exists a unique
integer np ∈ Z, almost always equal to zero, such that

x= pnp .
Y

p prime

2 Group theory
2.1 Groups: definition and examples
Defininition 2.1.1. A group (G, ◦) consists of a set G together with a func-
tion ◦ : G × G → G (aka a binary operation on G) such that

i) (a ◦ b) ◦ c = a ◦ (b ◦ c), for all a, b, c ∈ G (associativity);

ii) there is an element e ∈ G satisfying e ◦ a = a = a ◦ e, for all a ∈ G


(existence of neutral element);

iii) for every a ∈ G there exists a∗ ∈ G satisfying a ◦ a∗ = e = a∗ ◦ a


(existence of inverses).

Below we will use the notation a∗∗ := (a∗ )∗ , a∗∗∗ := ((a∗ )∗ )∗ etc.

Defininition 2.1.2. Let (G, ◦) be a group.

a) The order of (G, ◦) is the cardinality of the set G;

b) We say that (G, ◦) is finite if it has finite order;

c) We say that (G, ◦) is abelian if a ◦ b = b ◦ a (terminology: a and b


commute), for all a, b ∈ G.

Remark 2.1.3. When the binary operation is clear, we often denote the
group (G, ◦) simply by G and write ab or a + b instead of a ◦ b.

Example 2.1.4. Here are a couple of examples and counter-examples:

12
1. (Z, +), (Q, +), (R, +), (C, +) are all abelian groups, with e = 0 and
a∗ = −a for all a.

2. (N, +) is not a group, because 0 ̸∈ N and a∗ = −a ̸∈ N when a ∈ N.

3. (Q\{0}, ×), (R\{0}, ×), (C\{0}, ×) are all abelian groups, with e = 1
and a∗ = a−1 for all a. These groups are often denote by Q∗ , R∗ , C∗ .

4. (Z\{0}, ×) is not a group, because a−1 ∈ Z\{0} iff a = ±1. Note


that this also implies that ({±1}, ×) is an abelian group. This group is
sometimes denoted by Z∗ .

5. The Klein four-group (V4 , ◦) is the abelian group given by V4 := {e, a, b, c}


and multiplication table:

◦ e a b c
e e a b c
a a e c b (2)
b b c e a
c c b a e

The last two axioms in Definition 2.1.1 can actually be weakened.

Proposition 2.1.5. Suppose that ◦ is an associative binary operation on a


set G satisfying one of the following two pairs of axioms:

i) there exists an element e ∈ G satisfying e ◦ a = a, for all a ∈ G;

ii) for every a ∈ G, there exists a∗ ∈ G satisfying a∗ ◦ a = e,

or

iii) there exists an element e ∈ G satisfying a ◦ e = a, for all a ∈ G;

iv) for every a ∈ G, there exists a∗ ∈ G satisfying a ◦ a∗ = e.

Then (G, ◦) is a group.

Proof. We only prove the case in which ◦ satisfies the first two weaker axioms.
The proof of the other case is analogous.

13
Assuming that i) and ii) hold, let us first show that a ◦ a∗ = e for all
a ∈ G:

a ◦ a∗ = e ◦ (a ◦ a∗ )
= (e ◦ a) ◦ a∗
= ((a∗∗ ◦ a∗ ) ◦ a) ◦ a∗
= (a∗∗ ◦ (a∗ ◦ a)) ◦ a∗
= (a∗∗ ◦ e) ◦ a∗
= a∗∗ ◦ (e ◦ a∗ )
= a∗∗ ◦ a∗
= e.

It then quickly follows that a ◦ e = a as well:

a ◦ e = a ◦ (a∗ ◦ a) = (a ◦ a∗ ) ◦ a = e ◦ a = a

Here are some elementary but important consequences of the group ax-
ioms.

Proposition 2.1.6. Let (G, ◦) be a group.

a) The neutral element in G is unique.

b) The inverse of an element in G is unique.

c) For all a, b ∈ G, we have

(a ◦ b)∗ = b∗ ◦ a∗ (3)

Proof. a) Suppose that e and e′ are two neutral elements in G. On the one
hand, we have e ◦ e′ = e because e′ is a neutral element. On the other
hand, we have e ◦ e′ = e′ because e is a neutral element. Therefore,
e = e′ .

b) Let a ∈ G and suppose that b, c are two inverses of a. Then

b = b ◦ e = b ◦ (a ◦ c) = (b ◦ a) ◦ c = e ◦ c = c.

14
c) Note that

(a◦b)◦(b∗ ◦a∗ ) = ((a◦b)◦b∗ )◦a∗ = (a◦(b◦b∗ ))◦a∗ = (a◦e)◦a∗ = a◦a∗ = e.

Since the inverse of a ◦ b is unique, this shows that (a ◦ b)∗ = b∗ ◦ a∗ .

The last proposition of this subsection is easy but quite useful.

Proposition 2.1.7. Let (G, ◦) be a group and a, b ∈ G.

a) The equation ax = b has a unique solution.

b) The equation xa = b has a unique solution.

Proof. a) Since a∗ a = e, we have

ax = b ⇔ a∗ (ax) = a∗ b ⇔ (a∗ a)x = a∗ b ⇔ ex = a∗ b ⇔ x = a∗ b.

b) Similarly, we have
xa = b ⇔ x = ba∗ .

If a1 , . . . , an ∈ G, for some n ∈ N, we define

a1 ◦ a2 ◦ · · · ◦ an := (((· · · ((a1 ◦ a2 ) ◦ a3 ) ◦ · · · ) ◦ an ) · · · ).

Note that the notation without brackets is justified, because associativity


guarantees that we would get exactly the same element by bracketing the
elements in a different way. It follows immediately from (3) that

(a1 ◦ a2 ◦ · · · ◦ an )∗ = a∗n ◦ · · · ◦ a∗2 ◦ a∗1 .

The next examples of groups are more involved and thus deserve separate
subsubsections.

15
2.1.1 Congruences
For the definition of the next two groups, we introduce the notion of con-
gruence between integers. Let n ∈ N. We say that two integers a and b are
congruent modulo n (notation: a ≡ b mod n) if n | a − b. Note that
n | a − b ⇐⇒ a − b = kn, for some k ∈ Z,
so a and b are congruent modulo n iff they differ by a multiple of n.
Example 2.1.8. Let n = 8. We have 0 ≡ 8 mod 8, because 8 − 0 = 8.
Similarly, 3 ≡ 27 mod 8, because 27 − 3 = 24 = 3 · 8.
Congruence modulo n is an equivalence relation:
a ≡ a mod n (reflexivity);
a ≡ b mod n ⇐⇒ b ≡ a mod n (symmetry);
a ≡ b mod n ∧ b ≡ c mod n =⇒ a ≡ c mod n (transitivity),
for all a, b, c ∈ Z. We can thus divide the integers into equivalence classes,
called congruence classes modulo n (aka residue classes modulo n) in this
case. For any a ∈ Z, its congruence class modulo n, denoted a, is the subset
of integers defined by
a := {a + kn | k ∈ Z} ,
and we denote the set of all congruence classes modulo n by Z/nZ. The
elements of a congruence class are called representatives. Note that b ∈ Z is
a representative of a iff a = b iff a ≡ b mod n, so
Z/nZ = {0, 1, . . . , n − 1} .
We are now going to show that Z/nZ inherits an abelian group structure
from Z. For any a, b ∈ Z, define
a + b := a + b.
This binary operation is well-defined, because the definition of + does not
depend on the choice of representatives. To show this, suppose that a = a′
and b = b′ , i.e., a − a′ = kn and b − b′ = ℓn, for some k, ℓ ∈ Z. Then
(a + b) − (a′ + b′ ) = (a − a′ ) + (b − b′ ) = (k + ℓ)n,
which proves that
a + b = a′ + b ′ .

16
Proposition 2.1.9. The pair (Z/nZ, +) forms an abelian group (the additive
group of integers modulo n).
Proof. Associativity in Z/nZ follows immediately from associativity in Z:

(a + b) + c = a + b + c = (a + b) + c;
a + (b + c) = a + b + c = a + (b + c).

It is also obvious that e = 0 and a∗ = −a:

0 + a = 0 + a = a = a + 0 = a + 0;
a + −a = a − a = 0 = −a + a = −a + a,

for every a ∈ Z.
Example 2.1.10. Let n = 4. The multiplication table of Z/4Z (which per-
haps should be called the addition table) is equal to

+ 0 1 2 3
0 0 1 2 3
1 1 2 3 0
2 2 3 0 1
3 3 0 1 2

In general, we will write −a := −a. This is justified by the fact that


Z/nZ also inherits a product from Z:

a · b := ab,

for any a, b ∈ Z. This definition is also independent of the choice of repre-


sentatives. To show this, suppose a = a′ and b = b′ , i.e., a − a′ = kn and
b − b′ = ℓn, for certain k, ℓ ∈ Z. Then

ab − a′ b′ = (a − a′ )b + a′ (b − b′ ) = (kb + a′ ℓ)n,

so ab = a′ b′ . This product is clearly associative and abelian, and 1 is the


neutral element, but not all elements of Z/nZ are invertible, e.g. 0 · a =
0 · a = 0, for all a ∈ Z, so 0 is not invertible because 0 ̸= 1.
Lemma 2.1.11. The element a ∈ Z/nZ is invertible (w.r.t. the product) if
and only if gcd(a, n) = 1.

17
Proof. Suppose gcd(a, n) = 1. By Corollary 1.2.10, there exist x, y ∈ Z such
that
xa + yn = 1.
This implies that x · a = xa = xa + yn = 1, so a is invertible in Z/nZ.
Now suppose that gcd(a, n) = d > 1. Then a = dx and n = dy for
certain x, y ∈ Z satisfying 1 < x < a and 1 < y < n. This implies that
ay = dxy = dyx = nx, whence a · y = ay = nx = 0. If a were invertible,
with inverse a∗ , then

y = 1y = (a∗ a)y = a∗ (ay) = a∗ 0 = 0.

But that would contradict the fact that 1 < y < n, so a can not be invertible.

Define
(Z/nZ)∗ := {a ∈ Z/nZ | gcd(a, n) = 1} .
We denote the inverse of a ∈ (Z/nZ)∗ by a−1 . Note that a−1 ̸∈ Z in general,
so the notation a−1 would not make sense. Note also that to compute a−1
we can use the algorithm from the proof of Proposition 1.2.8 to find x, y ∈ Z
such that ax + ny = 1, because then x = a−1 .

Example 2.1.12. Let n = 315. Example 1.2.9 implies that


−1
1057 = −14.

The following proposition is an immediate consequence of the above.

Proposition 2.1.13. The pair ((Z/nZ)∗ , ·) forms an abelian group (the mul-
tiplicative group of integers modulo n).

Example 2.1.14. Let n = 8. Then (Z/8Z)∗ = {1, 3, 5, 7}. The multiplica-


tion table of (Z/8Z)∗ is equal to

· 1 3 5 7
1 1 3 5 7
3 3 1 7 5
5 5 7 1 3
7 7 5 3 1

18
Note that this multiplication table matches that of V4 in (2) under the corre-
spondence
e ↔ 1, a ↔ 3, b ↔ 5, c ↔ 7.
This shows that (V4 , ◦) and ((Z/8Z)∗ , ·) are "essentially the same" group. As
we will see below, the above correspondence is an example of what is called
an isomorphism, i.e., both groups are isomorphic.

The order of (Z/nZ)∗ is equal to ϕ(n), where ϕ : N → N is Euler’s ϕ-


function. In one of the next sections, we will show that

1
!
ϕ(n) = n 1−
Y

p|n
p
p prime

In Example 2.1.14 we saw that the order of (Z/8Z)∗ is 4, and indeed


1 8
 
ϕ(8) = 8 1 − = = 4.
2 2
Here is another example:

Example 2.1.15. Suppose that p ∈ N is a prime number. Then all integers


between 1 and p − 1 are coprime with p, so (Z/pZ)∗ = {1, 2, . . . , p − 1}, and
indeed
1
!
ϕ(p) = p 1 − = p − 1.
p

2.1.2 The symmetric group


Let X be a set and S(X) the set of all bijections f : X → X.

Proposition 2.1.16. Composition of bijections defines a group structure on


S(X) with the neutral element being the identity on X (notation: idX ).

Proof. Recall that the composite of f, g ∈ S(X) (in the operator order) is
defined by
g ◦ f (x) = g(f (x))
for all x ∈ X. This is clearly associative: if f, g, h ∈ S(X), then

(h ◦ g) ◦ f (x) = h(g(f (x))) = h ◦ (g ◦ f )(x),

19
for all x ∈ X. It is also clear that idX is the neutral element in S(X):
idX ◦ f (x) = idX (f (x)) = f (x) = f (idX (x)) = f ◦ idX (x).
Finally, each f ∈ S(X) has an inverse, precisely because f is a bijection.
Defininition 2.1.17. When |X| = n ∈ N, we call S(X) the symmetric
group on n letters and denote it by Sn .
Remark 2.1.18. From now on, we will write the composition of permuta-
tions as a product, omitting ◦.
By ordering the elements of X, the elements of Sn can be identified with
the permutations of the elements of {1, . . . , n}. One way to denote a permu-
tation σ ∈ Sn (we typically denote the elements of Sn by Greek lowercase
letters such as σ or τ ) is
1 2 n−1
!
... n
.
σ(1) σ(2) . . . σ(n − 1) σ(n)
Example 2.1.19. Let n = 5. Then
1 2 3 4 5
!

3 2 1 5 4
denotes the permutation of {1, 2, 3, 4, 5} sending 1 to 3, 2 to 2, 3 to 1, 4 to
5, and 5 to 4.
Lemma 2.1.20. The group Sn has order n!.
Proof. We have to count the number of σ ∈ Sn . There are n choices for σ(1).
After chosing σ(1), there remain n − 1 choices for σ(2). After chosing σ(2),
there remain n − 2 choices for σ(3), etc.
We can also draw permutations by two rows of numbers, the top row
being 1, 2, . . . , n and the bottom row σ(1), σ(2), . . . , σ(n), and connecting
equal numbers in the top and bottom row by straight lines.
Example 2.1.21.
1 2 1 2 3 1 2 3

1 2 1 2 3 1 2 3
! ! !
= = =
2 1 3 2 1 2 3 1

2 1 3 2 1 2 3 1

20
Given two permutation σ, τ ∈ Sn , the diagram for στ is obtained by gluing
the diagram for τ on top of the diagram for σ, renumbering the strands of
σ according to the bottom row of τ (which yields the bottom row for the
composite) and straightening the strands of the whole diagram.
Example 2.1.22. If σ and τ are the last two permutations in Example 2.1.21,
then
1 2 3

στ =

1 3 2

A permutation σ ∈ Sn is called a cycle of length k, where 1 < k ≤ n, if


there are k different elements a1 , a2 , . . . , ak ∈ {1, . . . , n} such that
σ(ai ) = ai+1 ∀ 1 ≤ i < k;
σ(ak ) = a1 ;
σ(x) = x ∀ x ̸∈ {a1 , . . . , ak }.
In that case, we denote σ by
(a1 a2 . . . ak ).
This notation is slightly ambiguous, because cyclic permutation of the ai
yields the same cycle:
(a1 a2 . . . ak ) = (ak a1 . . . ak−1 ) = . . . = (a2 a3 . . . a1 ).
By convention, cycles of length 1 are of the form (a), where a ∈ {1, . . . , n},
and they all correspond to the identity permutation.
Example 2.1.23. The permutation
1 2 3 4
!
= (1 3 4) = (4 1 3) = (3 4 1)
3 2 4 1
is a cycle (of length 3). The permutation
1 2 3 4
!
= (2 4)(1 3)
3 4 1 2
is not a cycle, but the product of two cycles (each of length 2).

21
Two cycles (a1 a2 . . . ak ) and (b1 b2 . . . bℓ ) are called disjoint if the sets
{a1 , a2 , . . . , ak } and {b1 , b2 , . . . , bℓ } are disjoint. Note that disjoint cycles
σ, τ ∈ Sn commute, i.e., στ = τ σ.

Proposition 2.1.24. Every element of Sn can be written as a product of


pairwise disjoint cycles. This product is unique up to reordering the cycles
and up to cycles of length 1 (in other words, it is essentially unique).

Proof. We prove this by induction w.r.t. n. When n = 1, the only element


of Sn is the identity, which is the cycle (1). Now, let n > 1 and consider the
elements
1, σ(1), σ 2 (1), σ 3 (1), . . . .
Since {1, . . . , n} has only n elements, it is clear that there are r, s ∈ {1, . . . , n}
with r < s such that σ r (1) = σ s (1). Therefore, we have

σ s−r (1) = 1.

Let k ∈ {1, . . . , n} be the smallest integer such that σ k (1) = 1. Then

σ1 := (1 σ(1) σ 2 (1) . . . σ k−1 (1))

is a cycle of length k and the original permutation σ restricts to a permutation


of the subset {1, . . . , n}\{1, σ(1), . . . , σ k−1 (1)}. By induction, this restricted
permutation can be written as the product of t disjoint cycles σ2 · · · σt of
length at most n − k, for some t ∈ N. This implies that σ can be written as
the product of t + 1 disjoint cycles:

σ = σ1 σ2 · · · σt .

We leave the proof of essential uniqueness to the reader.


From now on we will often omit the cycles of length 1 in these products.

Example 2.1.25.

1 2 3 4 5 6 7 8 9 10
!
= (1 4 5 3)(2 10)(6 7 9),
4 10 1 5 3 7 9 8 6 2

where we have omitted (8).

22
Given some arbitrary cycles, it is not hard to write their product as a
disjoint union of cycles, following the procedure described in the proof of
Proposition 2.1.24, where it is easiest to work from right to left. We just give
two examples to illustrate this:

Example 2.1.26. Let n = 4. Then

(1 2 3)(2 3 4) = (1 2)(3 4);


(1 2)(2 3)(1 2) = (1 3).

A cycle of length 2 is called a transposition.

Corollary 2.1.27. Any non-trivial permutation can be written as a product


of transpositions.

Proof. Let σ ∈ Sn . By Proposition 2.1.24, we can write

σ = σ1 σ2 · · · σt ,

where the σi are disjoint cycles. Now, suppose that σ1 = (a1 a2 . . . ak ). Then
we can write
σ1 = (a1 a2 )(a2 a3 ) · · · (ak−1 ak ).
The same holds the other σi , of course, which proves that σ can be written
as a product of transpositions.
Contrary to the factorization of a permutation σ ∈ Sn into disjoint cycles,
a factorization of σ into transpositions is not unique in general (nor are the
transpositions disjoint), e.g.

(1 2)(2 3)(1 2) = (1 3) = (1 3)(1 2)(1 3)

However, there is one property which is shared by all these factorizations.


Before we explain that, we first have to prepare the ground.
An inversion in a permutation σ ∈ Sn is a pair 1 ≤ i < j ≤ n such that
σ(i) > σ(j). Let N (σ) be the number of inversions in σ.

Defininition 2.1.28. The sign ϵ(σ) of a permutation σ ∈ Sn is defined as

ϵ(σ) := (−1)N (σ) .

23
Proposition 2.1.29. The sign function

ϵ : Sn → {±1}
σ 7→ ϵ(σ)

satisfies
ϵ(στ ) = ϵ(σ)ϵ(τ )
for any σ, τ ∈ Sn .

Proof. We can use the Vandermonde polynomial

P (x1 , . . . , xn ) := (xi − xj )
Y

1≤i<j≤n

to give an alternative definition of the sign of a permutation σ ∈ Sn :

P (xσ(1) , . . . , xσ(n) )
ϵ(σ) = . (4)
P (x1 , . . . , xn )

Note that P (xσ(1) , . . . , xσ(n) ) and P (x1 , . . . , xn ) have the same factors up to
a sign, which implies that the quotient on the right-hand side in (4) is indeed
equal to ±1, depending on the parity of the number of inversions in σ.
Now, let σ, τ ∈ Sn . Then

P (xσ(τ (1)) , . . . , xσ(τ (n)) )


ϵ(στ ) =
P (x1 , . . . , xn )
P (xσ(τ (1)) , . . . , xσ(τ (n)) ) P (xτ (1) , . . . , xτ (n) )
= ·
P (xτ (1) , . . . , xτ (n) ) P (x1 , . . . , xn )
P (xσ(1) , . . . , xσ(n) ) P (xτ (1) , . . . , xτ (n) )
= ·
P (x1 , . . . , xn ) P (x1 , . . . , xn )
= ϵ(σ)ϵ(τ ).

Proposition 2.1.30. Let σ ∈ Sn . All factorizations of σ into transpositions


have the same number of factors modulo 2, i.e., the number of transpositions
is always even or always odd.

24
Proof. Let σ = σ1 · · · σt be any factorization of σ into transpositions. Every
transposition has a negative sign (see Exercise 5), hence

ϵ(σ) = (−1)t . (5)

This implies that the number of transpositions in any factorization is always


congruent to t modulo 2.
As it turns out, permutations can be factored even further. A transpo-
sition (ij) ∈ Sn is called elementary if |i − j| = 1. Thus, the elementary
transpositions in Sn are
σi := (i i + 1),
for i = 1, . . . , n − 1.

2.1.3 The dihedral group


If X is endowed with some geometric structure, like a bilinear form or a
particular shape, then one can define the subset of S(X) consisting of only
those bijections which preserve that geometric structure. Composition in
S(X) restricts to such a subset, inducing a group structure on it.

Defininition 2.1.31. The infinite dihedral group D∞ consists of all isome-


tries of the Euclidean plane which preserve the origin.

By the well-known classification of isometries of the Euclidean plane, an


isometry which preserves the origin is either a rotation ρα around the origin
through an angle α ∈ [0, 2π[ or a reflection σβ in a line (the reflection axis)
through the origin which forms an angle β ∈ [0, π[ with the positive x-axis.
These symmetries are all linear, so they can be represented by matrices:

cos(α) − sin(α) cos(2β) sin(2β)


! !
ρα = , σβ = . (6)
sin(α) cos(α) sin(2β) − cos(2β)

Note that det(ρα ) = 1, whereas det(σβ ) = −1. Further, recall that ρ0 = idR2
and that, for any α, β ∈ [0, 2π[, the composite of two rotations is again a
rotation: 
ρ
α+β , if α + β < 2π;
ρα ρβ =
ρα+β−2π , if α + β ≥ 2π.

25
Recall also that the composite of two reflections is a rotation (which is con-
sistent with the observation about the determinants above). More precisely,
let β, γ ∈ [0, π[ such that β ≥ γ. Using the usual trigonometric formulae
cos(2(β − γ)) = cos(2β) cos(2γ) + sin(2β) sin(2γ)
sin(2(β − γ)) = sin(2β) cos(2γ) − cos(2β) sin(2γ),
we see that
cos(2β) sin(2β) cos(2γ) sin(2γ)
! !
σβ σγ =
sin(2β) − cos(2β) sin(2γ) − cos(2γ)
cos(2(β − γ)) − sin(2(β − γ))
!
=
sin(2(β − γ)) cos(2(β − γ))
= ρ2(β−γ) .
In particular, this shows that σβ σ0 = ρ2β , where σ0 is the reflection in the
x-axis. Multiplying both sides of that equation by σ0 on the right and using
that σ02 = ρ0 = idR2 , yields
σβ = ρ2β σ0 ,
for any β ∈ [0, π[. We thus see that
D∞ = {ρα | 0 ≤ α < 2π} ∪ {ρα σ0 | 0 ≤ α < 2π} . (7)
There are also finite dihedral groups.
Defininition 2.1.32. Let n ∈ N. The dihedral group Dn of order 2n consists
of the isometries of the Euclidean plane which preserve a regular polygon Pn
with n-sides and centered at the origin.
Without loss of generality, we may assume that the coordinates of the ver-
tices of Pn are (cos(2kπ/n), sin(2kπ/n)), for k = 0, 1, . . . , n−1. By definition
Dn ⊂ D∞ , and the description of D∞ in (7) implies that
n o n o
Dn = ρ2kπ/n | k = 0, 1, . . . , n − 1 ∪ ρ2kπ/n σ0 | k = 0, 1, . . . , n − 1 . (8)

When dealing with Dn , we will often use the notation 1 := ρ0 , ρk := ρ2kπ/n


and σ := σ0 . Note that
ρn = 1; (9)
σ 2 = 1; (10)
σρk = ρn−k σ, k = 0, 1, . . . , n − 1, (11)
where the last equality follows from Exercise 8.

26
2.1.4 Exercises
1. Prove that e∗ = e and a∗∗ = a for all a ∈ G.

2. Let G be a group.

i) Show that G is abelian if

(ab)∗ = a∗ b∗ ,

for all a, b ∈ G;
ii) Show that G is abelian if

a∗ = a,

for all a ∈ G.
−1
3. Compute 100 ∈ (Z/257Z)∗ .

4. Prove that Sn is abelian iff n ≤ 2.

5. Prove that

a) ϵ(σ) = −1 if σ is a transposition;
b) ϵ(σ) = (−1)k−1 if σ is a cycle of length k.

6. a) Write the following permutations as products of disjoint cycles


and do the same for their inverses
1 2 3 4 1 2 3 4 5 1 2 3 4 5 6
! ! !
, , .
3 2 1 4 2 4 5 1 3 6 4 5 1 2 3

b) Write the following permutations as products of disjoint cycles e


compute their parity.

(1 4 5 2)(2 3 4) and (2 3 4)(1 4 5 2).

7. Show that

a) for all i, j = 1, . . . , n − 1, we have

σi2 = e, σi σj = σj σi if |i−j| > 1, σi σj σi = σj σi σj if |i−j| = 1;

27
b) any σ ∈ Sn factors as a product of σi , i.e.,
σ = σj1 σj2 · · · σjm ,
for some m ∈ N and j1 , . . . , jm ∈ {1, 2, . . . , n − 1} (which need not
be pairwise distinct).
8. Show that the following relation holds in the infinite dihedral group
D∞ :
σ0 ρα = ρ2π−α σ0 ,
for any α ∈ [0, 2π[.
9. Let σ := σ0 and τ := ρσ in Dn . Prove that
σ 2 = 1, τ 2 = 1, (στ )n = 1.

10. Quaternions (invented by Sir William Rowan Hamilton in 1843) are


generalizations of complex numbers of the form
a + ib + jc + kd,
where a, b, c, d ∈ R and i, j, k satisfy
i2 = j 2 = k 2 = −1,
ij = k, ji = −k,
jk = i, kj = −i,
ki = j, ik = −j.
Denote the set of all quaternions by H and define the sum and the
product of two quarternions by

(a + ib + jc + kd) + (a′ + ib′ + jc′ + kd′ ) =


(a + a′ ) + i(b + b′ ) + j(c + c′ ) + k(d + d′ ),
and
(a + ib + jc + kd) · (a′ + ib′ + jc′ + kd′ ) = (aa′ − bb′ − cc′ − dd′ )
+ i(ab′ + ba′ + cd′ − dc′ )
+ j(ac′ − bd′ + ca′ + db′ )
+ k(ad′ + bc′ − cb′ + da′ ),

28
respectively.
Finally, the conjugate of a quaternion is defined by

a + ib + jc + kd) = a − ib − jc − kd.

a) Show that

(a + ib + jc + kd) · (a + ib + jc + kd) =
(a + ib + jc + kd) · (a + ib + jc + kd) =
a2 + b 2 + c 2 + d 2 .

b) Show that every quaternion can be written as

α + kβ,

where α, β ∈ C. Using this form, write down the sum and the
product of two quaternions and the conjugate of a quaternion.
c) Prove that (H, +) is an abelian additive group.
d) Let H∗ := H\{0}. Prove that (H∗ , ·) is a non-abelian multiplica-
tive group.

11. Let X be a set and P (X) the power set of X, which is the set of all
subsets of X. Define the symmetric difference of A, B ∈ P (X), denoted
A △ B, by
A △ B := (A ∪ B)\(A ∪ B).
Show that (P (X), △) is an abelian group.

12. Given a non-empty set X and a group (G, ◦), let GX be the set of all
functions f : X → G and define the product f1 · f2 of two functions
f1 , f2 ∈ GX by
f1 · f2 (x) := f1 (x) ◦ f2 (x),
for x ∈ X.

a) Show that (GX , ◦) is a group.


b) Show that GX is abelian iff G is abelian.

29
2.2 Subgroups, homomorphisms and the direct prod-
uct
2.2.1 Subgroups
Just as in the previous subsection, we start with a definition.
Defininition 2.2.1. Let (G, ◦) be a group. A subset H ⊆ G is a subgroup
if
i) H ̸= ∅;
ii) for all a, b ∈ H, we have a ◦ b ∈ H;
iii) for all a ∈ H, we have a∗ ∈ H;
Example 2.2.2. 1. For any group (G, ◦), the subsets {e} and G are nat-
urally subgroups.
2. The increasing chain of subsets Z ⊂ Q ⊂ R ⊂ C is a chain of subgroups
w.r.t. addition.
3. The subset R>0 ⊂ R is not a subgroup w.r.t. addition, because −a ̸∈
R>0 when a ∈ R>0 .
4. However, R>0 ⊂ R\{0} is a subgroup of the multiplicative group R∗ .
5. For any n ∈ N, the subset

nZ := {na | a ∈ Z} ⊆ Z

is a subgroup of (Z, +).


Proposition 2.2.3. If H is a subgroup of a group G, with operation ◦, then
(H, ◦) is a group itself.
Proof. Thanks to axiom ii) in Definition 2.2.1, the operation ◦ : G × G → G
restricts to ◦ : H ×H → H. Associativity for (H, ◦) is therefore an immediate
consequence of associativity for (G, ◦).
Thanks to axiom i) in Definition 2.2.1, there exists at least one element
a ∈ H and, thanks to axiom iii) in Definition 2.2.1, we also have a∗ ∈ H. By
axiom ii) in Definition 2.2.1, we therefore see that

e = a ◦ a∗ ∈ H.

30
In other words, (H, ◦) always contains the neutral element of (G, ◦), which
of course is also the neutral element of the subgroup.
Existence of inverses in (H, ◦) is guaranteed by axiom iii) in Defini-
tion 2.2.1.

Proposition 2.2.4. a) A subset H ⊆ Z is a subgroup of (Z, +) iff H =


nZ, for some n ∈ N.

Let n ∈ N. A subset
b) n o H ⊆ Z/nZ is a subgroup of (Z/nZ, +) iff H =
0, d, 2d, . . . , n − d for some d ∈ N such that d | n.

Proof. a) We already know that nZ is a subgroup of (Z, +).


Conversely, assume that H ⊆ Z is a subgroup of (Z, +). If H = {0},
then H = nZ with n = 0. If H ̸= {0}, then H contains an integer
a ̸= 0 and, thanks to axiom iii) in Definition 2.2.1, it also contains −a,
so H contains a positive integer (because either a or −a is positive).
Let n be the smallest positive integer contained in H. We claim that
H = nZ.
The inclusion nZ ⊆ H is an immediate consequence of axioms ii)
and iii) of Definition 2.2.1, because na = n(a−1)+n ∈ H (use induction
w.r.t. a) and n(−a) = −na ∈ H, for any a ∈ N.
For the other inclusion: let a ∈ H. We can write

a = qn + r,

for q ∈ Z and r ∈ {0, 1, . . . , n − 1}. But that implies that r ∈ H


because r = a − qn, and thus r = 0 or else n would not be the smallest
positive integer contained in H.
n o
b) Let d | n. Then 0, d, 2d, . . . , n − d is clearly a subgroup of (Z/nZ, +).
Conversely, let H be a subgroup of (Z/nZ, +). Define the subset K ⊂ Z
as
K := {a ∈ Z | a ∈ H} .
We claim that K is a subgroup of (Z, +). As a subgroup of (Z/nZ, +),
the subset H ⊂ Z/nZ has to contain 0, so K contains 0 and is thus
non-empty. Further, for any a, b ∈ K, we have a − b ∈ K because
a − b = a − b ∈ H. By Exercise 1, this proves the claim.

31
Since K is a subgroup of (Z, +), there must be a positive integer d
such that K = dZ, by Proposition 2.2.4 a). Note that n ∈ K because
n = 0 ∈ H, so n = da for some a ∈ N, which proves that d | n. This
completes the proof that H = {0, d, 2d, . . . , n − d}.

2.2.2 Homomorphisms
Any mathematical category whose objects are sets with some additional struc-
ture also contains functions between these objects, called morphisms, which
are compatible with the additional structure. For example, the morphisms
in the category of vector spaces are the linear maps and the morphisms in
the category of topological spaces are the continous maps. The morphisms
in the category of groups are called homomorphisms.

Defininition 2.2.5. Suppose that (G1 , ◦1 ) and (G2 , ◦2 ) are two groups.

a) A homomorphism between G1 and G2 is a function f : G1 → G2 satis-


fying
f (a ◦1 b) = f (a) ◦2 f (b), for all a, b ∈ G.
The set of all homomorphisms is denoted by Hom(G1 , G2 ).

b) An isomorphism is a bijective homomorphism. If there exists an iso-


morphism f : G1 → G2 , we say that G1 and G2 are isomorphic (nota-
tion: G1 ∼
= G2 ).
c) An endomorphism of (G1 , ◦1 ) is a homomorphism f : G1 → G1 . The
set of all such endomorphisms is denoted by End(G1 ).

d) An automorphism of G1 is a bijective endomorphism of G1 . The set of


all such automorphisms is denoted by Aut(G1 ).

Example 2.2.6. 1. Multiplication by an arbitrary but fixed integer n de-


fines an endomorphism of (Z, +), because

n(a + b) = na + nb, for all a, b, ∈ Z.

Note that this endomorphism is injective for any n ̸= 0 but only sur-
jective for n = ±1.

32
2. The exponential map x 7→ ex defines an isomorphism between (R, +)
and (R>0 , ∗), because

ea+b = ea eb , for all a, b, ∈ R.

3. The determinant defines a homomorphism GL(n, R) → R∗ , because

det(AB) = det(A) det(B), for all A, B ∈ GL(n, R).

4. The sign function ϵ : Sn → {±1} from Definition 2.1.28 is a homomor-


phism by Proposition 2.1.29. Note that it is surjective for any n > 1,
but only injective for n = 2.

5. The map a → e2aπ/n defines a homomorphism between Z/nZ and


GL(1, C), because

e2(a+b)π/n = e2aπ/n e2bπ/n , for all a, b ∈ Z/nZ.

Note that this homomorphism is well-defined:

e2(a+kn)π/n = e2aπ/n e2kπ = e2aπ/n , for all a, k ∈ Z.

Note also that it is injective.

6. The correspondence in Example 2.1.14 defines an isomorphism between


the groups V4 and (Z/8Z)∗ .

7. For any subgroup H ⊆ G of some group (G, ◦), the natural embedding
H ,→ G is an injective homomorphism.

8. Let G be any group and a ∈ G a fixed element. Then n 7→ an defines a


homomorphism Z → G, where a−1 denotes a∗ . This holds because

am+n = am an , for all m, n ∈ Z.

9. Let G be any group and g ∈ G a fixed element. Conjugation by g is the


automorphism of G defined by

x 7→ gxg ∗ , x ∈ G.

33
This is indeed a homomorphism, because the equality g ∗ g = e implies
that
gxyg ∗ = gxeyg ∗ = gxg ∗ gyg ∗ ,
for all x, y ∈ G. It is a bijection because the inverse of conjugation by
g is conjugation by g ∗ :

gxg ∗ = y ⇐⇒ g ∗ gxg ∗ g = g ∗ yg ⇐⇒ x = g ∗ yg,

for all x, y ∈ G.

From Lemma 2.2.7 until Proposition 2.2.13, let G1 and G2 be two groups
with neutral elements e1 and e2 , respectively, and f : G1 → G2 a homomor-
phism.

Lemma 2.2.7. We have

a) f (e1 ) = e2 ;

b) f (a∗ ) = f (a)∗ , for all a ∈ G1 .

Proof. a) On the one hand, f (e1 )f (e1 ) = f (e1 e1 ) = f (e1 ), while on


the other hand, f (e1 )e2 = f (e1 ). By Proposition 2.1.7, the equation
f (e1 )x = f (e1 ) has a unique solution in G2 , so we must have f (e1 ) = e2 .

b) For any a ∈ G, we have

f (a)f (a∗ ) = f (aa∗ ) = f (e1 ) = e2 .

By Proposition 2.1.6, the inverse of f (a) in G2 is unique, so we must


have f (a∗ ) = f (a)∗ .

There are two important subgroups associated to f : its kernel and its
image.

Defininition 2.2.8. The kernel of f is the subset of G1 defined by

ker(f ) := {a ∈ G1 | f (a) = e2 } .

Proposition 2.2.9. ker(f ) is a subgroup of G1 .

34
Proof. By Lemma 2.2.7, ker(f ) ̸= ∅ because it contains e1 . Moreover, for
any a, b ∈ ker(f ) we have ab∗ ∈ ker(f ), because

f (ab∗ ) = f (a)f (b∗ ) = f (a)f (b)∗ = e2 e∗2 = e2 e2 = e2 ,

where we have used Lemma 2.2.7 for the second equality and Exercise 1 for
the fourth equality. By Exercise 1, this implies that ker(f ) is a subgroup of
G1 .
As already remarked above, the sign ϵ : Sn → {±1} is a homomorphism,
see Definition 2.1.28 and Proposition 2.1.28.

Example 2.2.10. Let n ∈ N. The alternating group An is the subgroup of


Sn defined by
An := ker(ϵ).
In other words, the alternating group is the group of even permutations.

Proposition 2.2.11. f is injective if and only if ker(f ) = {e1 }.

Proof. Suppose that f is injective and let x ∈ ker(f ). Then

f (x) = e2 = f (e1 ).

Injectivity implies that x = e1 , so ker(f ) = {e1 }.


Conversely, suppose that ker(f ) = {e1 } and let x, y ∈ G1 such that
f (x) = f (y). Since f is a homomorphism, we get

f (xy ∗ ) = f (x)f (y)∗ = f (y)f (y)∗ = e2 ,

which means that xy ∗ = e1 , whence x = y. Thus f is injective.


Let us now look at the other subgroup associated to a homomorphism.

Defininition 2.2.12. The image of f is the subset of G2 defined by

im(f ) := {f (a) ∈ G2 | a ∈ G1 } .

Proposition 2.2.13. im(f ) is a subgroup of G2 .

35
Proof. By Lemma 2.2.7, im(f ) ̸= ∅ because it contains e2 . Moreover, for any
x, y ∈ im(f ) we have xy ∗ ∈ im(f ): If x = f (a) and y = f (b), for certain
a, b ∈ G1 , then

xy ∗ = f (a)f (b)∗ = f (a)f (b∗ ) = f (ab∗ ),

where the equalities hold for the same reasons as in the proof of Proposi-
tion 2.2.9. By Exercise 1, this implies that im(f ) is a subgroup of G2 .
The following result is called Cayley’s Theorem.

Theorem 2.2.14. Any group G is isomorphic to a subgroup of S(G). In


particular, if |G| = n, then G is isomorphic to a subgroup of Sn .

Proof. For every a ∈ G, the map λa : G → G, defined by

λa (x) := ax, x ∈ G,

is a bijection, since
y = ax ⇔ a∗ y = x,
for all x, y ∈ G. Thus λa ∈ S(G).
Now, define f : G → S(G) by

f (a) := λa , a ∈ G.

We claim that this map is an injective homomorphism. It is a homomorphism


because

(λa ◦ λb )(x) = λa (λb (x)) = λa (bx) = a(bx) = (ab)x = λab (x),

for all a, b, x ∈ G. It is also injective because, if a ∈ ker(f ), then λa = idG ,


which implies that
a = ae = λa (e) = idG (e) = e,
so ker(f ) = {e}.
This shows that f defines an embedding of G into S(G). By Exercise 11,
this implies that G ∼
= im(f ), which is indeed a subgroup of S(G), by Propo-
sition 2.2.13. When |G| = n, then S(G) = Sn by definition, so the final
remark of this theorem is clear.
Finally, let us look at the composition of homomorphisms.

36
Proposition 2.2.15. Let G1 , G2 , G3 be three groups and f : G1 → G2 and
g : G2 → G3 two homomorphisms. Then g ◦ f : G1 → G3 is also a homomor-
phism. If f1 and f2 are isomorphisms, then so is g ◦ f .
Proof. For any a, b ∈ G1 , we have
g ◦ f (ab) = g(f (ab)) = g(f (a)f (b)) = g(f (a))g(f (b)) = (g ◦ f (a))(g ◦ f (b)).
This shows that g ◦ f is indeed a homomorphism. The second claim of the
proposition follows from the fact that the composite of two bijections is again
a bijection.
Proposition 2.2.16. Let G1 , G2 be two groups and f : G1 → G2 an isomor-
phism. The inverse map f −1 : G2 → G1 is also an isomorphism.
Proof. For any x, y ∈ G2 , we have
f (f −1 (xy)) = xy = f (f −1 (x))f (f −1 (y)) = f (f −1 (x)f −1 (y)).
Since f is injective, this implies that
f −1 (xy) = f −1 (x)f −1 (y),
so f −1 is a homomorphism. As is well-known, the inverse of a bijection is
also a bijection, which proves the second claim of the proposition.

2.2.3 The direct product


The direct product of groups is the analog of the cartesian product of sets.
Defininition 2.2.17. The direct product of two groups G1 and G2 is the
group which consists of the set
G1 × G2 = {(a1 , a2 ) | a1 ∈ G, a2 ∈ G2 }
and the binary operation
(a1 , a2 ) ◦ (b1 , b2 ) := (a1 b1 , a2 b2 ), a1 , b1 ∈ G1 , a2 , b2 ∈ G2 .
Note that G1 ×G2 is indeed a group, since the binary operation on G1 ×G2
is defined componentwise, so all group axioms hold because they hold for
each component separately. For example, the neutral element of G1 × G2 is
(e1 , e2 ), where e1 and e2 are the neutral elements of G1 and G2 , respectively.
Similarly, the inverse of (a1 , a2 ) in G1 × G2 is equal to (a∗1 , a∗2 ), where the first
inverse is taken in G1 and the second in G2 .

37
Example 2.2.18. (R2 , +) = (R, +) × (R, +).

Note that G1 × G2 ∼
= G2 × G1 , where the isomorphism is given by

(a1 , a2 ) 7→ (a2 , a1 ), a1 ∈ G1 , a2 ∈ G2 .

Given three groups G1 , G2 , G3 , there is also an isomorphism

(G1 × G2 ) × G3 ∼
= G1 × (G2 × G3 )

given by

((a1 , a2 ), a3 ) 7→ (a1 , (a2 , a3 )), a1 ∈ G1 , a2 ∈ G2 , a3 ∈ G3 .

Thus the direct product of groups is associative and, without ambiguity, we


can write
G1 × G2 × · · · × Gn
for the direct product of n groups, for any n ∈ N.
The direct product extends to homomorphisms. If f1 : G1 → H1 , f2 : G2 →
H2 , . . . , fn : Gn → Hn are n homomorphisms of groups, then

(f1 , f2 , . . . , fn ) : G1 × G2 × · · · × Gn → H1 × H2 × · · · × Hn ,
(f1 , . . . , fn )(a1 , . . . , an ) := (f1 (a1 ), . . . , fn (an ))

defines a homomorphism of groups between the direct product of the Gi and


the direct product of the Hi .
Let G := G1 × G2 be a direct product of two groups with unit element
e := (e1 , e2 ). Note that the subsets H1 := G1 × {e2 } and H2 := {e1 } × G2
are subgroups of G isomorphic to G1 and G2 , with the obvious isomorphisms
being given by (a1 , e2 ) ↔ a1 and (e1 , a2 ) ↔ a2 , respectively. One can easily
check that these subgroups satisfy:

i) h1 h2 = h2 h1 , for all h1 ∈ H1 , h2 ∈ H2 ;

ii) H1 ∩ H2 = {e};

iii) for every g ∈ G, there are h1 ∈ H1 , h2 ∈ H2 such that g = h1 h2 .

As a matter of fact, these three properties caracterize the direct product.

38
Proposition 2.2.19. Let G be a group and suppose that H1 , H2 are subgroups
of G satisfying i), ii) and iii), then there is an isomorphism

H1 × H2 ∼
= G.

Proof. We claim that the map f : H1 × H2 → G, defined by

f ((h1 , h2 )) := h1 h2 ,

is an isomorphism.
First of all, f is a homomorphism, because condition i) implies that, for
all h1 , h′1 ∈ H1 and h2 , h′2 ∈ H2 , the elements

f ((h1 , h2 )(h′1 , h′2 )) = f ((h1 h′1 , h2 h′2 )) = h1 h′1 h2 h′2

and
f ((h1 , h2 ))f ((h′1 , h′2 )) = h1 h2 h′1 h′2
are equal.
Secondly, f is injective. By Proposition 2.2.11, it suffices to show that
ker(f ) = {(e, e)}. If f (h1 , h2 ) = e, for some h1 ∈ H1 and h2 ∈ H2 , then
h1 h2 = e, and thus h1 = h∗2 . But that implies that both h1 and h2 belong to
H1 ∩ H2 , so they both have to be equal to e, by condition ii).
Finally, surjectivity of f is precisely condition iii).
The following result is well-known and goes under the name of Chinese
Remainder Theorem.

Theorem 2.2.20. Let n1 , n2 , . . . , nt be t pairwise coprime positive integers,


for some t ∈ N. Then the map

Z/(n1 n2 · · · nt )Z → Z/n1 Z × Z/n2 Z × · · · × Z/nt Z


a mod (n1 n2 · · · nt ) 7→ (a mod n1 , a mod n2 , . . . , a mod nt )

defines an isomorphism.

Proof. We prove the theorem by induction w.r.t. t ∈ N. When t = 1,


there is nothing to prove. Now, assume that t > 1. For a start, define
f : Z/(n1 n2 · · · nt )Z → Z/n1 Z × Z/(n2 · · · nt )Z by

f (a) := (ã, â).

39
This is well-defined: if a ≡ b mod (n1 n2 · · · nt ), then n1 n2 · · · nt | a − b,
which implies that n1 | a − b and n2 · · · nt | a − b, so a ≡ b mod n1 and
a ≡ b mod (n2 · · · nt ). In other word, if a = b, then ã = b̃ and â = b̂.
The fact that f is a homomorphism is easy to check:

f (a + b) = f (a + b) = (a]
+ b, a[
+ b) = (ã + b̃, â + b̂)
= (ã + b̃, â + b̂) = f (a) + f (b),

for all a, b ∈ Z.
Let us now prove that f is injective. Suppose that f (a) = (0̃, 0̂). Then
ã = 0̃ and â = 0̂, so n1 | a and n2 · · · nt | a, and thus lcm(n1 , (n2 · · · nt )) | a.
The assumption that the ni be pairwise coprime implies that, for every prime
p, we have
min{νp (n1 ), νp (n2 ), . . . , νp (nt )} = 0,
whence

gcd(n1 , (n2 · · · nt )) = pmin{νp (n1 ),νp (n2 ),...,νp (nt )} = 1,


Y

p prime

by the obvious generalization of Corollary 1.3.7. By Corollary 1.4.4, we then


see that lcm(n1 , (n2 · · · nt )) = n1 n2 · · · nt , so n1 n2 · · · nt | a, meaning that
a = 0. This shows that ker(f ) = {0}, so f is injective by Proposition 2.2.11.
Surjectivity of f now follows from the fact that both Z/(n1 n2 · · · nt )Z and
Z/n1 Z × Z/(n2 · · · nt )Z have order n1 n2 · · · nt . Since f is injective, the order
of im(f ) has to be equal to n1 n2 · · · nt as well, by Exercise 11, which implies
that im(f ) = Z/n1 Z × Z/(n2 · · · nt )Z. In other words, our f is surjective.
Finally, by induction, the map

Z/(n2 · · · nt )Z → Z/n2 Z × · · · × Z/nt Z


a mod (n2 · · · nt ) 7→ (a mod n2 , . . . , a mod nt )

defines an isomorphism. Postcomposition of f by this isomorphism yields


the desired result.
The following corollary is in fact equivalent to Theorem 2.2.20, as one
can see immediately.

40
Corollary 2.2.21. Let n1 , n2 , . . . , nt be pairwise coprime positive integers,
for some t ∈ N. For any a1 , a2 , . . . , at ∈ Z, there exists an integer a ∈ Z,
unique modulo n1 n2 · · · nt , such that
a ≡ a1 mod n1 ;
a ≡ a2 mod n2 ;
.. .. ..
. . .
a ≡ at mod nt .
Recall that
(Z/nZ)∗ = {a ∈ Z/nZ | gcd(a, n) = 1} ,
for n ∈ N.
Corollary 2.2.22. Let n1 , n2 , . . . , nt be pairwise coprime positive integers,
for some t ∈ N. Then the map
(Z/(n1 n2 · · · nt )Z)∗ → (Z/n1 Z)∗ × (Z/n2 Z)∗ × · · · × (Z/nt Z)∗
a mod (n1 n2 · · · nt ) 7→ (a mod n1 , a mod n2 , . . . , a mod nt )
defines an isomorphism.
Proof. For a start, note that the isomorphism in Theorem 2.2.20 restricts
and corestricts to a bijection between the subsets of invertible elements on
both sides, because
gcd(a, (n1 n2 · · · nt )) = 1 ⇐⇒ gcd(a, ni ) = 1 for all i = 1, . . . , t. (12)
To see this, recall that
νp (n1 n2 · · · nt ) = νp (n1 ) + νp (n2 ) + . . . + νp (nt ),
by Corollary 1.3.5. Suppose that νp (a) > 0, for some prime p. Then
νp (n1 n2 · · · nt ) = 0 ⇐⇒ νp (ni ) = 0 for all i = 1, . . . , t,
which implies the equivalence in (12).
Secondly, note that the bijection in Corollary 2.2.22 is really a homomor-
phism, because
(ab mod n1 , ab mod n2 , . . . , ab mod nt ) =
(a mod n1 , a mod n2 , . . . , a mod nt ) · (b mod n1 , b mod n2 , . . . , b mod nt ) ,
for all a, b ∈ Z. This shows that this bijection is an isomorphism of groups.

41
Recall that
ϕ(n) = | (Z/nZ)∗ |,
for n ∈ N.
Corollary 2.2.23. For any n ∈ N, we have
1
!
ϕ(n) = n 1−
Y
.
p|n
p
p prime

Proof. Let n ∈ N and consider its prime factorization:


n = pn1 1 pn2 2 · · · pnk k ,
where ni := νpi (n) ∈ N for i = 1, 2, . . . , k. By Corollary 2.2.22, there is an
isomorphism
(Z/nZ)∗ ∼ ∗ ∗ ∗
= (Z/pn1 1 Z) × (Z/pn2 2 Z) × · · · × (Z/pk k Z) ,
n

which implies that


ϕ(n) = ϕ(pn1 1 )ϕ(pn2 2 ) · · · ϕ(pnk k ).
For any i = 1, 2, . . . , k, the order of (Z/pni i Z)∗ is easy to compute. Let
a ∈ {1, 2, . . . , pni i }. Then
gcd(a, pni i ) = 1 ⇔ gcd(a, pi ) = 1.
Since pi | a if and only if a = pi x, for some x = 1, . . . , pni −1 , we see that the
number of elements in {1, 2, . . . , pni i } which are indivisible by pi is given by
1
!
ϕ(pni i ) = pni i − pini −1 = pni i 1− .
pi
Thus
1 1 1
! ! !
ϕ(n) = 1−
pn1 1 pn2 2 1 − · · · pnk k 1 −
p1 p2 pk
1 1 1
! ! !
= pn1 1 pn2 2 · · · pk k 1 −
n
1− ··· 1 −
p1 p2 pk
1 1 1
! ! !
= n 1− 1− ··· 1 − .
p1 p2 pk

42
2.2.4 Exercises
1. Let (G, ◦) be a group and H ⊆ G a subset. Show that H is a subgroup
if the following two axioms are satisfied:

i) H ̸= ∅;
ii) for all a, b ∈ H, we have ab∗ ∈ H.

2. Let G be a group and H ⊆ G a non-empty finite subset satisfying

x, y ∈ H =⇒ xy ∈ H.

a) Show that H is a subgroup of G.


b) Give an example showing that finiteness of H is necessary in a).

3. Let (Hi )i∈I be a collection of subgroups of a group (G, ◦). Prove that
∩i∈I Hi is also a subgroup of (G, ◦).

4. a) Let G be a group and H1 ⊆ G, H2 ⊆ G two subgroups. Show that

G = H1 ∪ H2 ⇐⇒ G = H1 ∨ G = H2 .

b) Show that the Klein four-group V4 (see Example 2.1.4) has three
subgroups H1 , H2 , H3 , all different from G, which satisfy

G = H1 ∪ H2 ∪ H3 .

5. Show that S3 has 6 subgroups.

6. Let k, n ∈ N such that k < n. Define the subset H ⊂ Sn by

H := {σ ∈ Sn | 1 ≤ σ(i) ≤ k, ∀i = 1, 2, . . . , k} .

a) Show that H is a subgroup of Sn .


b) Show that there is an isomorphism H ∼
= Sk × Sn−k .
7. Show that Sn has a subgroup isomorphic to Dn , for any n ∈ N.

8. What is the smallest n ∈ N such that Sn contains a subgroup isomor-


phic to Z/6Z?

43
9. Let G be a group. Show that the map f : G → G, defined by
f (x) := x2 , x ∈ G,
is an endomorphism iff G is abelian.
10. Let G1 and G2 be two groups and assume that G2 is abelian. For any
two f, g ∈ Hom(G1 , G2 ) we define their product f ⋆ g : G1 → G2 by
f ⋆ g(x) := f (x)g(x), x ∈ G1 .
Show that f ⋆ g is a homomorphism.
11. Suppose that f : G1 → G2 is an injective homomorphism. Show that
G1 ∼
= im(f ).

12. a) Prove that


V4 ∼
= Z/2Z × Z/2Z.
b) Prove that
V4 ∼
= D2 ,
where D2 is the dihedral group of order 4.
c) Let X be a set with two elements. Prove that
V4 ∼
= P (X),
where P (X) is the power set of X with the symmetric difference
as group operation (see Exercise 11).
13. Let X be a set. Define
ϕ : (Z/2Z)X → P (X)
by
ϕ(f ) := {x ∈ X | f (x) = 1} ,
for any function f : X → Z/2Z. Prove that ϕ is an isomorphism, with
inverse χ given by

1, if x ∈ A;
χ(A)(x) := 
0, if x ̸∈ A,
for any subset A ⊂ X.

44
14. Let m, n ∈ N. Prove that

Z/mZ × Z/nZ ∼
= Z/ gcd(m, n)Z × Z/lcm(m, n)Z.

Hint: Use the Chinese Remainder Theorem to reduce the proof to the
case in which m and n are both powers of some prime p.

2.3 Generators, cyclic groups, order and cosets


2.3.1 Generators and cyclic groups
Let G be a group (from now on we suppress the binary operation in the
notation and write ab for a ◦ b and a−1 for a∗ in abstract settings, even
though in some concrete cases the binary operation is given by addition) and
X ⊆ G a subset.

Defininition 2.3.1. The subgroup of G generated by X, denoted ⟨X⟩, is


given by the subset of all elements a ∈ G which can be written as

a = x1 x2 · · · xn ,

for some n ∈ N and x1 , x2 , . . . , xn ∈ G such that xi ∈ X or x−1


i ∈ X for all
i = 1, 2, . . . , n.

Example 2.3.2. 1. Every group G is generated by G. However, in gen-


eral we are interested in smaller sets of generators.

2. The additive group Z is generated by 1, because



a · 1
= 1 + 1 + . . . + 1, if a ∈ Z≥0 ;
a=
−a · (−1) = −1 − 1 − . . . − 1, if a ∈ Z≤0 .

3. Similarly, the additive group Z/nZ, for any n ∈ N, is generated by 1.

4. The Klein four-group is generated by {a, b} or {a, c} or {b, c}, but not
by subsets of fewer elements.

5. By Proposition 2.1.24, the symmetric group Sn is generated by all cy-


cles. By Corollary 2.1.27 it is also generated by the transpositions, and
even (see Exercise 7) by the elementary transpositions.

45
Let G be a group and x ∈ G an element. Then

⟨x⟩ = {xn | n ∈ Z} .

Remark 2.3.3. If the group operation in G is written additively, then

⟨x⟩ = {nx | n ∈ Z} .

Defininition 2.3.4. We say that G is cyclic if

G∼
= ⟨x⟩

for some x ∈ G. When this is the case, we call x a generator of G.


Example 2.3.5. We have already seen that Z and Z/nZ, for any n ∈ N, are
cyclic. In Corollary 2.3.9, we will see that these are the only cyclic groups,
up to isomorphism. In particular, V4 and Sn , for n ≥ 3, are not cyclic.
Note that every cyclic group is abelian: if G ∼
= ⟨x⟩, for some x ∈ G, then

xm · xn = xm+n = xn · xm ,

for all m, n ∈ Z. The converse is false in general, e.g., V4 is abelian but not
cyclic.

2.3.2 Order
Defininition 2.3.6. Let G be a group and x ∈ G. The order of x (notation:
ord(x)) is the smallest n ∈ N such that xn = e, provided it exists. If there is
no such n, we define ord(x) = ∞.
Before giving some examples, note that

ord(x) = |⟨x⟩|,

for any x ∈ G.
Example 2.3.7. 1. Let V4 = {e, a, b, c} be the Klein four-group. Then
ord(e) = 1 and ord(a) = ord(b) = ord(c) = 2
2. In Z, we have 
∞, if a ̸= 0;
ord(a) =
1, if a = 0.

46
3. Let n ∈ N. Recall that

Z/nZ = {0, 1, . . . , n − 1} .

For any a ∈ {0, 1, . . . , n − 1}, we claim that


n
ord(a) = . (13)
gcd(a, n)

To prove this claim, define d := gcd(a, n) and note that (13) is equiva-
lent to the claim that
⟨ d ⟩ = ⟨ a ⟩. (14)
By Proposition 1.2.8, there exist x, y ∈ Z such that xa + yn = d. This
implies that xa = d, whence ⟨ a ⟩ ⊇ ⟨ d ⟩. Conversely, note that a = qd
for some q ∈ N, whence ⟨ a ⟩ ⊆ ⟨ d ⟩. This proves the equation in (14)
and, therefore, the equation in (13) as well.

4. Let k ∈ N. By its very definition, a k-cycle in Sn has order k.

Proposition 2.3.8. Let G be a group, x ∈ G, and ord(x) = n < ∞. For


any m ∈ Z, we have
xm = e ⇐⇒ n | m.

Proof. Suppose that m = qn, for some q ∈ Z. Then

xm = xqn = (xn )q = eq = e.

Conversely, suppose that xm = e. By Euclidean division, there are q, r ∈ Z


such that 0 ≤ r < n and
m = qn + r.
Suppose that r > 0. Then

xr = xm−qn = xm (xn )−q = e · e−q = e.

But this implies that ord(x) ≤ r < n, which is a contradiction. Thus we see
that r = 0, which means that n | m.

Corollary 2.3.9. Let G be a cyclic group. If |G| = ∞, then G ∼


= Z. If

|G| = n < ∞, then G = Z/nZ.

47
Proof. Since G is cyclic, there is an element x ∈ G such that G ∼
= ⟨x⟩.
Suppose first that |G| = ∞. Then ord(x) = ∞ and we can define the
homomorphism

f : Z → ⟨x⟩
f (n) = xn .

By definition, this homomorphism is surjective. It is also injective, because


xn ̸= e for all n ̸= 0, so ker(f ) = {0}. Therefore, f is an isomorphism.
Now suppose that ord(x) = n < ∞. Define the homomorphism

f : Z/nZ → ⟨x⟩
f (a) = xa .

Note that f is well-defined: if b = a, then b = a + kn for some k ∈ Z, so

xb = xa (xn )k = xa · ek = xa .

As before, surjectivity of f is immediate. To prove injectivity, note that

f (a) = e ⇔ xa = e ⇔ n | a ⇔ a = 0,

where the second equivalence is a consequence of Proposition 2.3.8. This


shows that ker(f ) = {0}, so f is injective, which completes the proof that f
is an isomorphism.
Remark 2.3.10. a) Note that the binary operation in ⟨x⟩ is written mul-
tiplicatively, whereas in Z or Z/nZ it is written additively. Therefore,
people often use the notation Cn for the multiplicative cyclic group of
order n, where n ∈ N ∪ {∞}. Corollary 2.3.9 implies that C∞ ∼ = Z and

Cn = Z/nZ, for n ∈ N.

b) Note that, by Corollary 2.3.9, all cyclic groups are abelian and that,
by the same corollary and Proposition 2.2.4, the subgroups of a cyclic
group are also cyclic.
Corollary 2.3.11. Let f : G1 → G2 be a homomorphism between two groups.
a) If x ∈ G1 has finite order, then so does f (x) ∈ G2 and, moreover,

ord(f (x)) | ord(x).

48
b) If f is injective, then ord(f (x)) = ord(x) for all x ∈ G1 .

Proof. a) Suppose that ord(x) = n < ∞. Then

f (x)n = f (xn ) = f (e1 ) = e2 .

By Proposition 2.3.8, this implies that

ord(f (x)) | ord(x).

In particular, ord(f (x)) is finite.

b) Suppose, additionally, that f is injective and let m := ord(f (x). Then,


for any x ∈ G1 , we have

f (xm ) = f (x)m = e2 .

Injectivity of f now implies that xm = e1 and, therefore, we have n | m


by Proposition 2.3.8. As we already know that m | n, this shows that
m = n.

2.3.3 Cosets
Let G be a group and H ⊆ G a subgroup. The left cosets of H in G are the
subsets
aH := {ah | h ∈ H} ,
for a ∈ G an arbitrary but fixed element. Similarly, the right cosets of H in
G are the subsets
Ha := {ha | h ∈ H} ,
for a ∈ G an arbitrary but fixed element. When the group operation of G is
written additively, we denote the left and right cosets by a + H and H + a,
respectively. The set of all left cosets is denoted by G/H, and the set of all
right cosets by H\G.

Remark 2.3.12. One way to think about left or right cosets of H in G as


translations of H inside G.

Example 2.3.13. 1. Note that eH = He = H. More generally, when


a ∈ H, we have aH = Ha = H.

49
2. Let G = Sn (symmetric group), H = An (alternating group) and σ1 :=
(1 2) ∈ Sn (elementary transposition). Every σ ∈ σ1 An is an odd
permutation, because

ϵ(σ1 σ) = ϵ(σ1 )ϵ(σ) = −1 · 1 = −1.

On the other hand, any odd permutation τ ∈ Sn belongs to σ1 An , since

τ = σ1 (σ1 τ ),

because σ12 = 1. Note that σ1 τ ∈ An , as ϵ(σ1 τ ) = ϵ(σ1 )ϵ(τ ) = (−1)2 =


1. Thus the left coset σ1 An consists of all odd permutations. Since
every permutation is either even or odd, we see that

Sn = An ∪ σ1 An

is the disjoint union of the cosets of even and odd permutations.


The same holds for the right cosets of An in Sn . Note also that, for
any σ ∈ An , we have

ϵ(σ1 σσ1 ) = ϵ(σ1 )ϵ(σ)ϵ(σ1 ) = −1 · 1 · (−1) = 1,

so σ1 σσ1 ∈ An . This implies that σ1 An = An σ1 , because

σ1 σ = (σ1 σ)(σ1 σ1 ) = (σ1 σσ1 )σ1 ,

for every σ ∈ An .

3. Let G = R2 and H a straight line through the origin. For any a ∈ R2 ,


the left/right coset a + H = H + a is the straight line through a parallel
to H.

4. Let G = Z and H = nZ for a certain n ∈ N. For every a ∈ Z, the


left/right coset a + H = H + a is precisely the residue class modulo n,
which we denoted by a.

5. Let G = D3 . Recall that


n o
D3 = 1, ρ, ρ2 , σ, ρσ, ρ2 σ ,

50
where ρ is the rotation around the origin through 120 degrees and σ is
the reflection in the x-axis. Recall also that
ρ3 = 1 = σ 2 and ρσ = σρ2 .
Now, let H = {1, σ}. This is a subgroup of D3 because σ 2 = 1, and it
is not hard to see that there are three left cosets of H in D3 :
1H = {1, σ}
ρH = {ρ, ρσ}
n o
ρ2 H = ρ2 , ρ2 σ .
There are also three right cosets of H in D3 :
H1 = {1, σ}
n o
Hρ = {ρ, σρ} = ρ, ρ2 σ
n o n o
ρ2 H = ρ2 , σρ2 = ρ2 , ρσ .
In this case, the left and right cosets do not all coincide: ρH ̸= Hρ and
ρ2 H ̸= Hρ2 .
Observe that in all examples the left, resp. right, cosets of H in G are
mutually disjoint. Moreover, every element of G belongs to a left, resp. right,
coset. This is a general feature, as we will show in the following proposition,
which we only formulate and prove for left cosets but equally holds for right
cosets.
Proposition 2.3.14. Let G be a group, H ⊆ G a subgroup and a, b ∈ G.
Then
a) aH = bH ⇐⇒ a−1 b ∈ H;
b) either aH ∩ bH = ∅ or aH = bH;
c) every x ∈ G belongs to a unique left coset of H in G.
Proof. a) Suppose that aH = bH. Then b = be ∈ bH = aH, so b = ah
for some h ∈ H. This implies that a−1 b = h ∈ H.
Conversely, suppose that a−1 b = h, for some h ∈ H. Then, for every
h′ ∈ H, we have
bh′ = (ah)h′ = a(hh′ ) ∈ aH,

51
which shows that bH ⊆ aH. For the other inclusion, note that

ah′ = (bh−1 )h′ = b(h−1 h′ ) ∈ bH,

so aH ⊆ bH. Thus aH = bH.


b) Suppose that aH ∩ bH ̸= ∅. Then there exists an element x ∈ aH ∩ bH,
so there are h, h′ ∈ H such that

x = ah = bh′ .

Hence a−1 b = h(h′ )−1 ∈ H. By a), this implies that aH = bH.


c) Let x ∈ G. Then x = xe ∈ xH. Moreover, this is the only left coset
which contains x, because the left cosets are mutually disjoint by b).

Remark 2.3.15. The left cosets of H in G can also be obtained as the


equivalence classes of the equivalence relation

a ∼ b ⇔ a−1 b ∈ H,

for a, b ∈ G. Mutatis mutandis, the same is true for the right cosets.
Defininition 2.3.16. Let G be a group and H ⊆ G a subgroup such that
there are finitely many left cosets of H in G. The index of H in G, denoted
[G : H], is defined as
[G : H] := |G/H|.
A system of distinct representatives is a subset S of G containing precisely
one element per left coset of H in G. Such a system satisfies

G =
[
sH;
s∈S

|S| = [G : H].

Example 2.3.17. 1. Let G = Sn and H = An . Then [Sn : An ] = 2 and

S = {1, σ1 }

is a system of distinct representatives of the left cosets of An in Sn .


Recall that σ1 = (1 2), but we could have chosen any other odd permu-
tation as well.

52
2. Let G = Z and H = nZ for a certain n ∈ N. Then [Z : nZ] = n and

S = {0, 1, . . . , n − 1}

is a system of distinct representatives.

The next couple of results could also be formulated for infinite groups,
using cardinal numbers, but we will restrict to finite groups.

Lemma 2.3.18. Let G be a finite group and H ⊆ G a subgroup. Then

|aH| = |Ha| = |H|,

for every a ∈ G.

Proof. The map f : H → aH, given by f (h) := ah, is obviously a bijection.


Hence |aH| = |H|. The proof that |Ha| = |H| is analogous.
The following result is called Lagrange’s Theorem.

Theorem 2.3.19. Let G be a finite group and H ⊆ G a subgroup. Then

|G| = [G : H] · |H|.

Proof. This proof is an immediate consequence of Proposition 2.3.14 and


Lemma 2.3.18.

Corollary 2.3.20. Let G be a finite group and H ⊆ G a subgroup. Then


|H| divides |G|.

Proof. This is an immediate consequence of Theorem 2.3.19, since

|G|
= [G : H] ∈ N.
|H|

Corollary 2.3.21. Let G be a finite group and x ∈ G. Then ord(x) divides


|G|.

Proof. This is an immediate consequence of Corollary 2.3.20, since ⟨x⟩ is a


subgroup of G and ord(x) = |⟨x⟩|.

53
Corollary 2.3.22. Let G be a finite group of prime order p. Then G is
cyclic and isomorphic to Z/pZ

Proof. Since p > 1, there is an element x ̸= e ∈ G. By Corollary 2.3.21, the


order of x divides p, so it is equal to 1 or p. Since x ̸= e, its order has to be
equal to p. Therefore,
|⟨x⟩| = ord(x) = p.
But that means that G ∼ = ⟨x⟩ ∼ = Z/pZ.
Corollary 2.3.23. Let G be a finite group of order less than 5. Then G is
cyclic or isomorphic to the Klein four-group.

Proof. If |G| = 1, then g = {e}, which is cyclic.


If |G| has order 2,3 or 5, then G is cyclic by Corollary 2.3.23.
If |G| = 4, then all non-trivial elements of G have order 2,3 or 4, by
Corollary 2.3.21. If one of them, say a, has order 4, then G is cyclic because

G∼
= ⟨a⟩ ∼
= Z/4Z.

If all of them have order 2, then G = {e, a, b, c} with a2 = b2 = c2 = e. This


means that the multiplication table of G looks like

◦ e a b c
e e a b c
a a e (15)
b b e
c c e

It is not hard to see that this implies that G ∼ = V4 (Klein’s four-group),


because the elements in each column and each row have to be distinct, since
left and right multiplication are bijections of G to G.
Corollary 2.3.23 implies that all groups of order less than 5 are abelian.
The smallest non-abelian group is S3 ∼
= D3 , which has order 6.

2.3.4 Exercises
1. Let G be a group and X ⊆ G a subset.

a) Check that ⟨X⟩ is indeed a subgroup of G.

54
b) Prove that ⟨X⟩ is the smallest subgroup of G containing all ele-
ments of X.
c) Conversely, prove that the smallest subgroup of G containing all
elements of X is ⟨X⟩.

2. Let n ∈ N and m ∈ Z. Show that

⟨m⟩ = Z/nZ ⇐⇒ gcd(m, n) = 1.

a) For any prime p, prove that (Z/pZ)∗ is a cyclic group of order


p − 1.
b) Show that (Z/nZ)∗ need not be cyclic if n ∈ N is not a prime.

3. Let n ∈ N≥3 . Prove that the alternating group An from Example 2.2.10
is generated by the 3-cycles in Sn .

4. Let G be a group and a, b ∈ G. Show that

a) ord(a) = ord(a−1 ).
b) ord(aba−1 ) = ord(b).
c) ord(ab) = ord(ba).

5. Let σ ∈ Sn . Write σ as a product of disjoint cycles:

σ = σ1 σ2 · · · σt ,

where σi is a cycle of length ki , for i = 1, 2, . . . , t. Prove that

ord(σ) = lcm(k1 , k2 , . . . , kt ).

6. Compute the order of every element in the dihedral group Dn , for


n ∈ N.

7. Let G be a group such that ord(x) = 2 for all x ̸= e. Prove that G is


abelian.

8. Let G be an abelian group and x, y ∈ G two elements of finite order.

a) Prove that
ord(xy) | lcm(ord(x), ord(y)).

55
b) Show that gcd(ord(x), ord(y)) = 1 implies that

ord(xy) = ord(x)ord(y).

9. Show that any element of D∞ is equal to the product of two elements


of order 2 and that this implies that the results of Exercise 8 can be
false when G is not abelian.

10. Let G be an abelian group and define the subset

T (G) := {x ∈ G | ord(x) < ∞} ⊆ G.

Show that T (G) is a subgroup of G (the torsion subgroup).

11. Let G1 and G2 be two groups and x1 ∈ G1 , x2 ∈ G2 two elements of


finite order. Let x = (x1 , x2 ) ∈ G1 × G2 . Show that

ord(x) = lcm(ord(x1 ), ord(x2 )).

12. Comparing the orders of the elements of the respective groups, show
that
S4 ∼
̸= D12 and A4 ∼ ̸= S3 × C2 .

13. Let G be a group and H ⊆ G a subgroup. Prove that

a) for all a, b ∈ G, aH = bH iff Ha−1 = Hb−1 ;


b) if S is a system of distinct representatives of the left cosets of H
in G, then S −1 := {s−1 | s ∈ S} is a system of distinct represen-
tatives of the right cosets of H in G;
c) there is a bijection between G/H and H\G.

2.4 Normal subgroups, quotient groups and homomor-


phism and isomorphism theorems
Let G be a group and H ⊆ G a subgroup. In general,
i) the left and the right cosets of H in G do not coincide;

ii) the group structure on G does not induce group structures on G/H
and H\G.

56
It is thus natural to ask for which subgroups the left and right cosets coincide
and for which subgroups the sets of left and right cosets inherit group struc-
tures from G. As we will see in the following two subsections, the answer to
both questions is the same: normal subgroups.

2.4.1 Normal subgroups


Defininition 2.4.1. Let G be a group and H ⊆ G a subgroup. We say that
H is a normal subgroup of G (notation: H  G) if

ghg −1 ∈ H

for all g ∈ G and h ∈ H.

Example 2.4.2. 1. Every group G has two natural normal subgroups,


namely {e} and G.

2. If G is abelian, then every subgroup H ⊆ G is normal, because

ghg −1 = gg −1 h = h,

for all g ∈ G and h ∈ H.

3. The alternating group An is a normal subgroup of Sn , for every n ∈ N.


Recall that An = ker(ϵ), where ϵ : Sn → {±1} is the sign homomor-
phism, and note that

ϵ(στ σ −1 ) = ϵ(σ)ϵ(τ )ϵ(σ) = ϵ(τ )(ϵ(σ))2 = ϵ(τ ) = 1,

for all σ ∈ Sn and τ ∈ An . This shows that στ σ −1 ∈ An for all σ ∈ Sn


and τ ∈ An , which means that An  Sn .

4. Let G be a group and define its center Z(G) by

Z(G) := {a ∈ G | ax = xa, ∀x ∈ G} .

Then Z(G)  G, because

ghg −1 = hgg −1 = h,

for all g ∈ G and h ∈ Z(G).

57
5. Let G be a group. Given two elements g, h ∈ G, define their commuta-
tor [g, h] by
[g, h] := ghg −1 h−1 .
Let [G, G] ⊆ G be the subgroup of G generated by all commutators (the
commutator subgroup). This is a normal subgroup, because

ghg −1 = ghg −1 h−1 h = [g, h]h ∈ [G, G]

for all g ∈ G and h ∈ [G, G].

Proposition 2.4.3. Let G be a group and N ⊆ G a subgroup. Then

N G ⇐⇒ ∀a ∈ G : aN = N a.

Proof. Suppose that N  G. For any a ∈ G and n ∈ N , we have

an = (ana−1 )a ∈ N a and na = a(a−1 na) ∈ aN,

which shows that aN = N a.


Conversely, suppose that aN = N a, for all a ∈ G. This implies that, for
any a ∈ G and n ∈ N , we have an = n′ a for some n′ ∈ N , so

ana−1 = n′ aa−1 = n′ ∈ N.

This proves that N  G.

Corollary 2.4.4. Let G be a group and H ⊆ G a subgroup of index 2. Then


H  G.

Proof. Since [G : H] = 2, there are only two left cosets of H in G: the


subgroup H itself and its complement H c := {g ∈ G | g ̸∈ H}. The same
holds for the right cosets, so both H and H c are left and right cosets, i.e.,
satisfy aH = Ha and aH c = H c a for all a ∈ G. By Proposition 2.4.3, this
implies that H  G and H c  G.

Example 2.4.5. 1. Recall that [Sn : An ] = 2 (see Example 2.3.13). Corol-


lary 2.4.4 thus provides an alternative proof of the fact that An  Sn .

2. Let H = {ρi | i = 0 ≤ i < n} ⊂ Dn . Then [Dn : H] = 2 by (8). By


Corollary 2.4.4, this implies that H  Dn .

58
Note that there is no analog of Corollary 2.4.4 for when [G : H] > 2. For
example, consider G = D3 and H = ⟨σ⟩ = {1, σ}. Then [G : H] = 3 and
ρi H ̸= Hρi for i = 1, 2, as we have seen in point 5 of Example 2.3.13. In
particular, this implies that H is not a normal subgroup of G. Of course,
one can also see this directly, using that ρσ = σρ−1 , e.g.

ρσρ−1 = ρρσ = ρ2 σ ̸∈ ⟨σ⟩.

Proposition 2.4.6. Let f : G1 → G2 be a homomorphism between two


groups. Then
ker(f )  G1 .

Proof. For all g ∈ G1 and h ∈ ker(f ), we have

f (ghg −1 ) = f (g)f (h)f (g)−1 = f (g)e2 f (g)−1 = f (g)f (g)−1 = e2 ,

hence ghg −1 ∈ ker(f ). This proves that ker(f )  G1 .

Example 2.4.7. 1. By definition, we have An = ker(ϵ), where ϵ : Sn →


{±1} is the sign homomorphism, which shows once more that An  Sn
(as we had already seen before), thanks to Proposition 2.4.6 this time.

2. Let f : Dn → {±1} be the homomorphism determined by

f (ρ) = 1 and f (σ) = −1.

Then the subgroup H = {ρi | i = 0 ≤ i < n} from Exercise 2.4.5


is equal to ker(f ), which provides an alternative proof of the fact that
H  Dn , thanks to Proposition 2.4.6.

3. The special linear group SL(n, R) is defined by

SL(n, R) := {A ∈ GL(n, R) | det(A) = 1} .

This is a normal subgroup of the general linear group GL(n, R), because
SL(n, R) = ker(det), where det : GL(n, R) → R∗ is the homomorphism
given by the determinant.

59
2.4.2 Quotient groups
Let G be a group and N  G a normal subgroup. As explained in the
introduction to this section, we use the notation a := aN , for a ∈ G.
Proposition 2.4.8. The set of all left cosets G/N inherits a group structure
from G:
ab := ab,
for all a, b ∈ G. The set G/N together with this group structure is called the
quotient group.
Moreover, the canonical projection π : G → G/N , defined by

π(a) = a

for a ∈ G, is a surjective homomorphism and ker(π) = N .


Proof. We already sketched the proof of well-definedness in the introduction.
Suppose that c = a and d = b, then c = an and d = bn′ for certain n, n′ ∈ N .
Therefore,
cd = anbn′ = ab((b−1 nb)n′ ) ∈ abN
because b−1 nb ∈ N , given that N  G. This proves that

ab = cd,

so the binary operation in G/N is independent of the choice of representa-


tives.
The neutral element is given by e, of course. Associativity in G/H follows
immediately from associativity in G and a−1 = a−1 for all a ∈ G. This proves
that G/H satisfies the group axioms.
Surjectivity of the canonical projection π : G → G/N is immediate. Note
further that
π(a)π(b) = ab = ab = π(ab),
for all a, b ∈ G. This shows that π is indeed a homomorphism, and ker(π) =
N because
π(a) = e ⇐⇒ a = e ⇐⇒ a ∈ N.

Remark 2.4.9. The set of all right cosets N \G inherits a similar group
structure from G, but in this section we will concentrate on G/N .

60
Example 2.4.10. 1. When G = Z and N = nZ, for some n ∈ N,
then the group G/N is exactly the additive group Z/nZ from Propo-
sition 2.1.9.

2. Note that the multiplicative group (Z/nZ)∗ from Proposition 2.1.13 is


not a quotient group, e.g., a−1 ̸= a−1 when a ̸= ±1.

3. Let G = R and N = Z (both as additive groups). Since R is abelian, Z


is automatically a normal subgroup and the quotient group

R/Z

is isomorphic to the one-dimensional (multiplicative) unitary group

U (1) := {e2πix | x ∈ R},

where the isomorphism is given by

f : R/Z → U (1)
f (x) = e2πix ,

for x ∈ R. This is well-defined: if x = y, then x = y + m for some


m ∈ Z, so

f (x) = e2πix = e2πi(y+m) = e2πiy e2πim = e2πiy = f (y),

because e2πim = 1. It is also a homomorphism, because

f (x)f (y) = e2πix e2πiy = e2πi(x+y) = f (x + y),

for x, y ∈ R.

4. Let G = S4 , N = {(1), (1 2)(3 4), (1 3)(2 4), (1 4)(2 3)} ∼


= V4 the normal
subgroup from Exercise 1, and note that |S4 /N | = 24/4 = 6. Let us
show that S3 ∼
= S4 /N .
We can embed S3 into S4 , as a (non-normal) subgroup, by the injective
homomorphism ι : S3 ,→ S4 which sends σ ∈ S3 to σ ′ ∈ S4 , defined by

σ(i), for i = 1, 2, 3;
σ ′ (i) := 
4, for i = 4.

61
Now, define the homomorphism f : S3 → S4 /N by

f (σ) := σ ′ ,

for σ ∈ S3 . By Proposition 2.2.15, this is indeed a homomorphism,


because f = π ◦ ι, where π : S4 → S4 /N is the canonical projection.
We claim that f is injective. By Proposition 2.2.11, we have to show
that ker(f ) = {(1)}. Note that, for any σ ∈ S3 , we have

σ ∈ ker(f ) ⇐⇒ σ ′ ∈ N.

By definition, σ ′ (4) = 4, so we see that σ ′ ∈ N if and only if σ ′ = (1) ∈


S4 , because the other permutations in N do not fix 4. This implies that

σ ∈ ker(f ) ⇐⇒ σ = (1) ∈ S3 ,

which proves the claim.


Injectivity of f now implies that |im(S3 )| = |S3 | = 6 (see Exercise 11 in
Section 2.3.3). Since |S4 /N | = 6 as well, we must have im(S3 ) = S4 /N ,
so f is surjective as well. This proves that f is an isomorphism of
groups, so
S3 ∼
= S4 /N.

Finally, let us look at the structure of subquotient groups, i.e., subgroups


of quotient groups.

Proposition 2.4.11. Let G be a group and N a normal subgroup of G. If


H is a subgroup of G such that N ⊆ H, then H/N is a subgroup of G/N .
Conversely, any subgroup of G/N is of that form.

Proof. If H ⊆ G is a subgroup containing N , then H/N is clearly a subgroup


−1
of G/N , because e ∈ H/N and ab = ab−1 ∈ H/N for all a, b ∈ H.
Conversely, let K ⊆ G/N be a subgroup. Define
n o
H := h ∈ G | h ∈ K .

This is a subgroup of G, because e ∈ H (as e ∈ K) and ab−1 ∈ H for all


−1
a, b ∈ H (as ab−1 = ab ∈ K). Since n = e ∈ K for all n ∈ N , we see that
N ⊆ H and K = H/N .

62
2.4.3 Homomorphism and isomorphism theorems
For a start, let us prove a fundamental theorem about homomorphisms.
Theorem 2.4.12 (the homomorphism theorem). Let f : G1 → G2 be a ho-
momorphism between two groups and N  G1 a normal subgroup satisfying
N ⊆ ker(f ). Then there exists a unique homomorphism g : G1 /N → G2 such
that f = g ◦ π, where π : G1 → G1 /N is the canonical projection. Moreover,
we have ker(g) = ker(f )/N ⊆ G1 /N .
f
G1 G2

π g

G1 /N
Proof. As usual, we write π(a) = a, for a ∈ G1 . Define
g : G1 /N → G2
g(a) := f (a).
This is well-defined: if a = b ∈ G1 /N , then a = bn for some n ∈ N , so
f (a) = f (bn) = f (b)f (n) = f (b)e2 = f (b).
Note that f (n) = e2 because N ⊆ ker(f ). By definition,
f (a) = g(a) = g(π(a)) = (g ◦ π)(a)
for all a ∈ G1 , so f = g ◦ π.
If g ′ : G1 /N → G2 is another homomorphism satisfying f = g ′ ◦ π, then
g ′ (a) = g ′ (π(a)) = (g ′ ◦ π)(a) = f (a) = g(a),
so g ′ = g, which proves unicity of g.
Finally, let us compute ker(g). For all a ∈ G1 , we have
a ∈ ker(g) ⇐⇒ g(a) = e2
⇐⇒ f (a) = e2
⇐⇒ a ∈ ker(f )
⇐⇒ a ∈ ker(f )/N,
so ker(g) = ker(f )/N .

63
Next, we are going to prove three fundamental theorems about isomor-
phisms, called the isomorphism theorems.
Theorem 2.4.13 (the first isomorphism theorem). Let f : G1 → G2 be a
homomorphism between two groups. Then there is an isomorphism
G1 / ker(f ) ∼
= im(f ).
Proof. By Theorem 2.4.12, there is a unique homomorphism
g : G1 / ker(f ) → G2
such that f = g ◦ π, where π : G1 → G1 / ker(f ) is the canonical projection.
Moreover, this homomorphism is injective, because
ker(g) = ker(f )/ ker(f ) = {e1 } ⊆ G1 /N.
Since f = g ◦ π, we have im(f ) = im(g) ⊆ G2 . We can thus corestrict g to
its image to get an isomorphism
G1 / ker(f ) ∼
= im(g) = im(f ).

Corollary 2.4.14. Let f : G1 → G2 be a surjective homomorphism between


two groups. Then there is an isomorphism
G1 / ker(f ) ∼
= G2 .
Proof. Apply Theorem 2.4.13, using that im(f ) = G2 because f is surjective.

Example 2.4.15. Let’s revisit two old friends.


1. Recall the surjective homomorphism
f : R → U (1)
f (x) = e2πix
from Example 2.4.10. Its kernel is given by the integers, as
e2πix = 1 ⇐⇒ x ∈ Z.
By Corollary 2.4.14, we conclude (again) that f induces an isomor-
phism
R/Z ∼= U (1).

64
2. Let G be a group and x ∈ G an element of order n, for some n ∈ N.
Define

f : Z → G,
f (m) := xm .

We already know that f is a homomorphism, see Example 2.2.6. By


Proposition 2.3.8, we know that

xm ⇐⇒ n | m,

so ker(f ) = nZ. Note also that im(f ) = ⟨x⟩. By Corollary 2.4.14, this
implies that f induces an isomorphism

Z/nZ ∼
= ⟨x⟩,

as we already knew from the proof of Proposition 2.3.8.


Before we can prove the second isomorphism theorem, we need a lemma.
Lemma 2.4.16. Let G be a group, N  G a normal subgroup and H ⊆ G
any subgroup. Then
i) H ∩ N is a normal subgroup of H;

ii) HN := {hn | h ∈ H, n ∈ N } is a subgroup of G.


Proof. i) This is immediate, because both N and H are normal subgroups
of H (note that H need not be a normal subgroup of G, but it is always
a normal subgroup of itself).

ii) Note that e ∈ H and e ∈ N , so e = ee ∈ N H, which shows that


N H ̸= ∅. Further, let h1 , h2 ∈ H and n1 , n2 ∈ N . Then

h1 n1 (h2 n2 )−1 = h1 n1 n−1


2 h2 = h1 h2 (h2 n1 n2 h2 ) ∈ HN,
−1 −1 −1 −1

because N  G. Thus HN is a subgroup of G.

Theorem 2.4.17 (the second isomorphism theorem). Let G be a group, N 


G a normal subgroup and H ⊆ G any subgroup. There is an isomorphism

H/(H ∩ N ) ∼
= HN/N.

65
Proof. Let π : G → G/N be the canonical projection and ϕ : H → G/N its
restriction to H. Then ker(ϕ) = H ∩ ker(π) = H ∩ N , so ϕ induces an
isomorphism
H/(H ∩ N ) ∼
= im(ϕ), (16)
by Theorem 2.4.13. Remains to compute im(ϕ). If x ∈ G, then

x ∈ im(ϕ) ⇔ ∃h ∈ H : x = h ⇔ ∃ h ∈ H : x ∈ hN.

This shows that im(ϕ) = HN/N and, by (16), that ϕ induces an isomorphism

H/(H ∩ N ) ∼
= HN/N.

Example 2.4.18. Without saying it, we already used the second isomorphism
theorem in the fourth point of Example 2.4.10. Let G = S4 , H = {σ ∈
S4 | σ(4) = 4} ∼
= S3 and N = {(1), (1 2)(3 4), (1 3)(2 4), (1 4)(2 3))}. Then
H ∩N = {(1)} ⊂ S4 , so Theorem 2.4.17 implies that there is an isomorphism

S3 ∼
= S3 N/N,

where S3 N/N ∼ = HN/N is a subgroup of S4 /N . Since |S3 N/N | = |S3 | = 6


and |S4 /N | = 6 as well, we conclude that S3 N/N = S4 /N , whence

S3 ∼
= S4 /N.

Before we prove the third isomorphism theorem, we need another lemma.

Lemma 2.4.19. Let G be a group and N  G a normal subgroup. If N ′  G


is a normal subgroup satisfying N ⊆ N ′ , then N ′ /N  G/N . Conversely,
any normal subgroup of G/N is of that form.

Proof. The proof is analogous to the proof of Proposition 2.4.11.

Theorem 2.4.20 (the third isomorphism theorem). Let G be a group and


N, N ′  G two normal subgroups such that N ⊆ N ′ . Then there is an
isomorphism
(G/N )/(N ′ /N ) ∼
= G/N ′ .

66
Proof. Let π ′ : G → G/N ′ be the canonical projection. Since N ⊆ N ′ , The-
orem 2.4.12 implies that there is a unique homomorphism g : G/N → G/N ′
such that π ′ = g ◦ π, where π : G → G/N is also the canonical projection.
This is illustrated by the following commutative diagram.

π′
G G/N ′

π g

G/N
Moreover, we have ker(g) = ker(π ′ )/N = N ′ /N . Surjectivity of π ′ implies
surjectivity of g (because π ′ = g ◦ π), which implies that g induces an iso-
morphism
(G/N )/(N ′ /N ) ∼
= G/N ′ ,
by Corollary 2.4.14.

2.4.4 Exercises
1. Let G = S4 and H = {(1), (1 2)(3 4), (1 3)(2 4), (1 4)(2 3)}.

a) Show that H is a subgroup of S4 , which is isomorphic to Klein’s


four-group.
b) Show that H  S4 .

2. Let G = Sn . Show that

H := {σ ∈ Sn | σ(1) = 1}

is a non-normal subgroup of Sn , for any n ≥ 3.

3. Let G be a group and Ni  G, i ∈ I, for some indexing set I. Show


that
Ni  G.
\

i∈I

4. Let G be a group. Show that

G is abelian ⇐⇒ Z(G) = G ⇐⇒ [G, G] = {e}.

67
5. Compute Z(D4 ) and [D4 , D4 ].

6. Let G be a group and recall that the commutator subgroup [G, G] is a


normal subgroup of G, see Example 2.4.2.

a) Prove that G/[G, G] is abelian.


b) Let N  G. Show that G/N is abelian if and only if [G, G] ⊆ N .
c) Prove that [Sn , Sn ] = An , for all n ∈ N.
d) Determine all normal subgroups of S3 and A4 .

7. Let G be a group and A ⊆ G a finite subset of order n. Define the


normalizer NG (A) and the centralizer CG (A) of A in G by
n o
NG (A) := g ∈ G | ∀a ∈ A : gag −1 ∈ A ;
CG (A) := {g ∈ G | ∀a ∈ A : ga = ag} .

Show that

a) NG (A) is a subgroup of G and CG (A) a normal subgroup of G;


b) there is a homomorphism f : NG (A) → Sn such that ker(f ) =
CG (A).

8. Let G = GL(2, R) and


1 1
( !)
A= .
0 1
Determine NG (A) and CG (A).

9. Let G be an abelian group and T (G) ⊆ G its torsion subgroup (see


Exercise 10). Show that there is an isomorphism

T (C∗ ) ∼
= Q/Z.

10. Determine the structure of G/N for G = S3 , D4 , A4 and all N  G.

11. Let G be a group and

N := ⟨x2 | x ∈ G⟩ ⊆ G.

Show that

68
a) N  G;
b) [G, G] ⊆ N .
12. Let G be a group such that G/Z(G) is cyclic. Prove that G is abelian.
13. Prove that
Sn ∼
̸= An × C2 ,
for n ≥ 3. Hint: compute the centers of both groups and compare.
14. Determine all homomorphisms f : Sn → C∗ for n ∈ N.
15. Let f : G1 → G2 be a homomorphism between two groups and suppose
that N2  G2 . Define

N1 := f −1 (N2 ) = {x ∈ G1 | f (x) ∈ N2 } .

Show that
a) N1  G1 ;
b) if f is surjective, then

G1 /N1 ∼
= G2 /N2 .

2.5 Group actions, conjugacy classes, semidirect prod-


ucts
2.5.1 Group actions
Let G be a group and X a set.
Defininition 2.5.1. A (left) action of G on X is a homomorphism

ϕ : G → S(X).

We will often simplify the notation for the action of g ∈ G on x ∈ X:

gx := ϕ(g)(x).

In this notation, the fact that ϕ is an action translates into

ex = x,
(gh)x = g(hx),

69
for all g, h ∈ G and x ∈ X.
In these notes we will concentrate on left actions and, therefore, omit the
adjective "left" henceforth. For completeness, though, let us also recall the
definition of a right action.

Remark 2.5.2. A right action of G on X is a homomorphism

ψ : G → S(X)op ,

where S(X)op is defined as S(X) but with composition in the opposite order.
To simplify notation, we can write

xg := ψ(g)(x),

such that ψ being a homomorphism translates into the conditions

xe = x,
x(gh) = (xg)h,

for all g, h ∈ G and x ∈ X.

As usual, first some examples.

Example 2.5.3. 1. By definition, the symmetric group Sn acts on {1, 2, . . . , n}


by permuting the elements.

2. By definition, the dihedral group Dn acts on the set of vertices of a


regular n-gon.

3. Let G be a group and H ⊆ G a subgroup. Then G acts on G/H by left


multiplication:
gx := gx.
(Check that this is well-defined).

4. Let G be a group, then G acts on itself by conjugation.

Let G be a group acting on a set X. We say that two elements x, y ∈ X


are equivalent (notation: x ∼G y) under the G-action if there exists an
element g ∈ G such that gx = y. This defines an equivalence relation:

• (reflexiveness) for all x ∈ X, we have x ∼G x because x = ex;

70
• (symmetry) for all x, y ∈ X, we have x ∼G y if and only if y ∼G x,
because x = gy if and only if y = g −1 x, for g ∈ G;
• (transitivity) for all x, y, z ∈ X, if x ∼G y and y ∼G z, then x ∼G z,
because if x = gy and y = hz for some g, h ∈ G, then x = ghz.
Defininition 2.5.4. The orbits in X under the G-action are the equivalence
classes with respect to the above equivalence relation.
We denote the orbit containing x ∈ X by

Gx := {gx | g ∈ G}

and note that distinct orbits are disjoint:

Gx = Gy or Gx ∩ Gy = ∅.

We say that the G-action on X is transitive if there is only one orbit, which
means that all elements of X are equivalent under the G-action.
Example 2.5.5. All G-actions in Example 2.5.3 are transitive, except pos-
sibly the last one. The equivalence classes in that case are called conjugacy
classes and, in general, there can be more than one conjugacy class. For
example, if G is abelian, then each conjugacy class has only one element, so
the total number of conjugacy classes is equal to |G|.
Defininition 2.5.6. Let G be a group acting on a set X and x ∈ X. The
stabilizer of x under the G-action Gx is defined as

Gx := {g ∈ G | gx = x} ⊆ G.

Before we prove that stabilizers are subgroups, let us recall the notion of
conjugate subgroups. If H is a subgroup of a group G and g ∈ G, then define
the conjugate subgroup gHg −1 by
n o
gHg −1 := ghg −1 | h ∈ H .

By definition, we have H  G if and only if gHg −1 = H for all g ∈ G.


Proposition 2.5.7. Let G be a group acting on a set X. For every x ∈ X,
the stabilizer Gx is a subgroup of G and, for every g ∈ G, we have

Ggx = gGx g −1 .

71
Proof. Let x ∈ X. Note that e ∈ Gx , because ex = x, so Gx ̸= ∅. Further,
let g, h ∈ Gx . Then
gh−1 x = gh−1 hx = gex = gx = x,
whence gh−1 ∈ Gx and thus Gx is a subgroup of G.
To prove the second claim, let g ∈ G. Then
hgx = gx ⇔ g −1 hgx = g −1 gx ⇔ g −1 hgx = ex ⇔ g −1 hgx = x,
for every h ∈ G. This proves that
h ∈ Ggx ⇔ g −1 hg ∈ Gx ⇔ h = g(g −1 hg)g −1 ∈ gGx g −1 .

The length of a finite orbit (i.e., the number of elements it contains) can
be expressed in terms of the stabilizer of any of its elements.
Proposition 2.5.8. Let G be a group acting on a finite set X and x ∈ X.
Then
|Gx| = [G : Gx ]. (17)
Proof. Define the map f : G/Gx → Gx by
f (g) := gx,
for g ∈ G. For all g, h ∈ G, we have
g = h ⇔ h−1 g ∈ Gx ⇔ h−1 gx = x ⇔ gx = hx,
so f is well-defined and injective. By definition of Gx, it is also surjective
and thus bijective, which proves the proposition.
The following corollary is an immediate consequence of Proposition 2.5.8
and the fact that X is the disjoint union of the G-orbits.
Corollary 2.5.9. Let G be a group acting on a finite set X and let Y ⊆ X
be any subset containing precisely one element of each orbit. Then
|X| = [G : Gx ]. (18)
X

x∈Y

By Proposition 2.5.7, different choices of Y yield conjugate stabilizers for


each orbit. Since conjugate stabilizers have the same order and, therefore,
the same index in G, the right-hand side of (18) does indeed not depend on
the choice of Y .

72
2.5.2 Conjugacy classes
Conjugation is an important special example of a group action, deserving a
subsection on its own. In this case G = X and the group action is given by
the homomorphism G → Aut(G) sending g ∈ G to the automorphism

x 7→ gxg −1 ,

for x ∈ G. The orbit of a given x ∈ G under conjugation is called its


conjugacy class, denoted by Cl(x):
n o
Cl(x) := gxg −1 | g ∈ G .

The total number of conjugacy classes of G is called the class number of


G and the stabilizer of an element x ∈ G is called its centralizer, usually
denoted by CG (x):
n o
CG (x) := g ∈ G | gxg −1 = x = {g ∈ G | gx = xg} .

Finally, the center Z(G) of G is equal to the intersection of all centralizers:

Z(G) := {g ∈ G | gx = xg, ∀x ∈ G} = CG (x).


\

x∈G

Finally, recall that each element of Z(G) yields a conjugacy class on its own.
With this in mind, Proposition 2.5.8 and Corollary 2.5.9 translate into:

Proposition 2.5.10. Let G be a finite group. For any x ∈ G, the number


of elements in Cl(x) is equal to [G : CG (x)]. Further, the order of G satisfies
the class equation:

|G| = |Z(G)| + [G : CG (x)],


X

x∈Y ′

where Y ′ ⊆ G is any subset containing precisely one element of each conju-


gacy class not belonging to Z(G).

Corollary 2.5.11. Let G be a finite group and x ∈ G. Then

|Cl(x)| divides |G|.

73
Proof. By Proposition 2.5.10, we have |Cl(x)| = [G : CG (x)] = |G|/|CG (x)|,
whence
|G|
= |CG (x)| ∈ N.
|Cl(x)|

We are now going to determine the conjugacy classes of the symmetric


group Sn , for any n ∈ N. Recall that in Proposition 2.1.24 we proved that
any permutation σ ∈ Sn can be written as the product of disjoint cycles:

σ = σ1 σ2 · · · σt .

In this case, we do include all cycles of length one as well. Let ki be the
length of σi , for i = 1, . . . , t, and assume that k1 ≥ k2 ≥ . . . ≥ kt > 0 (recall
that the σi commute pairwise, so we can always reorder them if necessary).
We say that (k1 , k2 , . . . , kt ) is the cycle type of σ.
Example 2.5.12. The cycle type of (1 2 3)(5 7)(4)(6)(8) ∈ S8 is (3, 2, 1, 1, 1)
By including all cycles of length one as well, we get
t
ki = n.
X

i=1

We say that (k1 , k2 , . . . , kt ) is a partition of n into t parts (notation: (k1 , k2 , . . . , kt ) ⊢


n).
Proposition 2.5.13. Let n ∈ N. Two elements of Sn are conjugate if and
only if they have the same cycle type. Thus there is a bijection between the
conjugacy classes of Sn and the partitions of n.
Proof. For a start, let
σ := (a1 a2 . . . ak )
be a cycle of length k and τ an arbitrary permutation in Sn . We claim that

τ στ −1 = (τ (a1 ) τ (a2 ) . . . τ (ak )).

First note that



τ (a
i+1 ), if i = 1, 2, . . . , k − 1;
τ στ −1 (τ (ai )) = τ σ(ai ) = 
τ (a1 ), if i = k.

74
Secondly, let j ∈ {1, 2, . . . , n}\{a1 , a2 , . . . , ak }. Then σ(j) = j, so
τ στ −1 (τ (j)) = τ σ(j) = τ (j).
This proves the claim.
Now, let σ ∈ Sn be any permutation and write it as a product of disjoint
cycles σ = σ1 σ2 · · · σt as above. Then, for any τ ∈ Sn , we have
τ στ −1 = τ σ1 σ2 · · · σt τ −1 = (τ σ1 τ −1 )(τ σ2 τ −1 ) · · · (τ σt τ −1 ).
By our claim, this implies that τ στ −1 and σ have the same cycle type.
Conversely, suppose that σ, ρ ∈ Sn have the same cycle type and write
both as products of disjoint cycles as above:
σ = σ1 σ2 · · · σt and ρ = ρ1 ρ2 · · · ρt .
For each i = 1, 2, . . . , t, write σi = (ai1 ai1 . . . aiki ) and ρi = (bi1 bi2 . . . biki ).
Then define τ ∈ Sn by
τ (aij ) := bij ,
for i = 1, 2, . . . , t and j = 1, 2, . . . , ki . Our claim shows that
τ σi τ −1 = ρi ,
for all i = 1, 2, . . . , t, whence
τ στ −1 = (τ σ1 τ −1 )(τ σ2 τ −1 ) · · · (τ σi τ −1 ) = ρ1 ρ2 · · · ρt = ρ.

Partitions of n can be represented by Young diagrams. If (k1 , k2 , . . . , kt ) ⊢


n, then the corresponding Young diagram consists of t vertically stacked rows
of boxes, such that the i-th row has ki boxes, for i = 1, 2, . . . , t. When n = 4,
for example, the possible partitions and corresponding Young diagrams are
(4) (3, 1) (2, 2) (2, 1, 1) (1, 1, 1, 1)

In general, the number of partitions of n is given by the partition function


p(n), e.g., p(4) = 5. This function has some nice properties, but unfortu-
nately there is no closed formula for it (yet).

75
2.5.3 Semidirect product
In this subsection, we will introduce a generalization of the direct product of
two groups.

Defininition 2.5.14. Let H and N be two groups and τ : H → Aut(N ) a


homomorphism. The semidirect product of N and H (notation: N ⋊ H) is
the group defined by the set N × H together with the binary operation

(n1 , h1 ) · (n2 , h2 ) := (n1 τ (h1 )(n2 ), h1 h2 ) ,

for n1 , n2 ∈ N and h1 , h2 ∈ H.

Example 2.5.15. 1. If τ is the trivial homomorphism, i.e., τ (h) = idN


for all h ∈ H, then N ⋊ H is simply the direct product N × H.

2. The Weyl group W of any classical Dynkin type is a semidirect product:

• In type An , for n ∈ N, we have

W ∼
= Sn+1 ∼
= {1} ⋊ Sn+1 ;

• In type Bn and Cn , for n ∈ N≥2 , we have

W ∼
= C2×n ⋊ Sn ,

where C2 = {±1} and Sn acts on C2×n by permuting the signs;


• In type Dn , for n ∈ N≥4 , we have

W ∼
×(n−1)
= C2 ⋊ Sn ,
×(n−1)
where C2 is isomorphic to the kernel of the homomorphism

f : C2×n → C2 ;
f (ϵ1 , ϵ2 , . . . , ϵn ) := ϵ1 ϵ2 · · · ϵn ,

with ϵi ∈ {±1} for i = 1, 2, . . . , n. Note that f ◦ σ = f , for all


σ ∈ Sn , so the action of Sn on C2×n restricts to an action of Sn
on ker(f ) ∼
×(n−1)
= C2 .

76
Let G := N ⋊ H be the semidirect product of two groups with homomor-
phism τ : H → Aut(N ) and unit element e = (eN , eH ). Then N ′ := N ×{eH }
and H ′ := {eN }×H are subgroups of G isomorphic to N and H, respectively.
The surjection

G→H
(n, h) 7→ h

is a homomorphism whose kernel is equal to N ′ . This implies that N ′  G,


by Proposition 2.4.6, and
G/N ∼ = H,
by Theorem 2.4.13. It is now easy to see that N ′ and H ′ satisfy:

i) N ′ ∩ H ′ = {e};

ii) for every g ∈ G, there are n′ ∈ N ′ and h′ ∈ H ′ such that

g = n′ h′ .

Finally, note that

(eN , h)(n, eH )(eN , h−1 ) = (τ (h)(n), h)(eN , h−1 )


= (τ (h)(n)τ (h)(eN ), hh−1 )
= (τ (h)(n), eH ),

for all n ∈ N and h ∈ H, where we have used that τ (h) is an automorphism


of N and thus τ (h)(eN ) = eN .

Proposition 2.5.16. Let G be a group with neutral element e, H ⊆ G a


subgroup and N  G a normal subgroup satisfying i) and ii) above. Then
there is an isomorphism
N ⋊H ∼ = G,
where τ : H → Aut(N ) is defined by τ (h)(n) := hnh−1 , for n ∈ N and
h ∈ H.

Proof. The map f : N ⋊ H → G, defined by

f (n, h) := nh,

77
is a homomorphism, because

f (n1 , h1 )f (n2 , h2 ) = n1 h1 n2 h2
= n1 h1 n2 h−1
1 h1 h2
= f (n1 h1 n2 h−1
1 , h1 h2 )
= f ((n1 , h1 ) · (n2 , h2 )),

for all n1 , n2 ∈ N and h1 , h2 ∈ H. As in the proof of Proposition 2.2.19,


bijectiveness of f follows from properties i) and ii). Therefore, f is an iso-
morphism.

2.5.4 Exercises
1. Let G be a group and H ⊆ G a subgroup of finite index. Show that

[G : H] = [G : gHg −1 ],

for every g ∈ G.
2. Let G be a group acting transitively on a set X, with |X| ≥ 2. Prove
that
∃ g ∈ G : ∀x ∈ X : gx =
̸ x.

3. Let G be a finite group acting on a finite set X. For every g ∈ G, define

n(g) := | {x ∈ X | gx = x} |.

Prove that the number of G-orbits in X is equal to


1 X
n(g).
|G| g∈G

(This is called Burnside’s formula.)


4. Let G be a group of order p2 , where p is a prime number. Show that
G is abelian.
5. Determine the conjugacy classes of the dihedral group Dn , for any
n ∈ N.
6. Determine all partitions and the corresponding Young diagrams for
n = 6.

78
7. Let G be a group and x, y ∈ G. Show that x and y are conjugate in G
if and only if ∃ a, b ∈ G : x = ab ∧ y = ba.

8. Suppose that G is a finite group with precisely two conjugacy classes.


Show that |G| = 2.

9. Show that N ⋊ H is indeed a group.

10. Show that the dihedral group Dn , for n ∈ N, is isomorphic to the


semidirect product of N := {1, ρ, . . . , ρn−1 } ∼
= Z/nN and H := {1, σ} ∼
=
Z/2Z.

11. Show that the symmetric group Sn , for n ∈ N≥2 , is isomorphic to the
semidirect product of N := An and H := {(1), (i j)}, for any 1 ≤ i <
j ≤ n.

79

You might also like