Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Groups Thm Proof

Download as pdf or txt
Download as pdf or txt
You are on page 1of 32

Part IA — Groups

Theorems with proof

Based on lectures by J. Goedecke


Notes taken by Dexter Chua

Michaelmas 2014

These notes are not endorsed by the lecturers, and I have modified them (often
significantly) after lectures. They are nowhere near accurate representations of what
was actually lectured, and in particular, all errors are almost surely mine.

Examples of groups
Axioms for groups. Examples from geometry: symmetry groups of regular polygons,
cube, tetrahedron. Permutations on a set; the symmetric group. Subgroups and
homomorphisms. Symmetry groups as subgroups of general permutation groups. The
Möbius group; cross-ratios, preservation of circles, the point at infinity. Conjugation.
Fixed points of Möbius maps and iteration. [4]

Lagrange’s theorem
Cosets. Lagrange’s theorem. Groups of small order (up to order 8). Quaternions.
Fermat-Euler theorem from the group-theoretic point of view. [5]

Group actions
Group actions; orbits and stabilizers. Orbit-stabilizer theorem. Cayley’s theorem
(every group is isomorphic to a subgroup of a permutation group). Conjugacy classes.
Cauchy’s theorem. [4]

Quotient groups
Normal subgroups, quotient groups and the isomorphism theorem. [4]

Matrix groups
The general and special linear groups; relation with the Möbius group. The orthogonal
and special orthogonal groups. Proof (in R3 ) that every element of the orthogonal
group is the product of reflections and every rotation in R3 has an axis. Basis change
as an example of conjugation. [3]

Permutations
Permutations, cycles and transpositions. The sign of a permutation. Conjugacy in Sn
and in An . Simple groups; simplicity of A5 . [4]

1
Contents IA Groups (Theorems with proof)

Contents
0 Introduction 4

1 Groups and homomorphisms 5


1.1 Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2 Homomorphisms . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.3 Cyclic groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.4 Dihedral groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.5 Direct products of groups . . . . . . . . . . . . . . . . . . . . . . 8

2 Symmetric group I 9
2.1 Symmetric groups . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2 Sign of permutations . . . . . . . . . . . . . . . . . . . . . . . . . 10

3 Lagrange’s Theorem 12
3.1 Small groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.2 Left and right cosets . . . . . . . . . . . . . . . . . . . . . . . . . 14

4 Quotient groups 15
4.1 Normal subgroups . . . . . . . . . . . . . . . . . . . . . . . . . . 15
4.2 Quotient groups . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
4.3 The Isomorphism Theorem . . . . . . . . . . . . . . . . . . . . . 16

5 Group actions 17
5.1 Group acting on sets . . . . . . . . . . . . . . . . . . . . . . . . . 17
5.2 Orbits and Stabilizers . . . . . . . . . . . . . . . . . . . . . . . . 17
5.3 Important actions . . . . . . . . . . . . . . . . . . . . . . . . . . 18
5.4 Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

6 Symmetric groups II 21
6.1 Conjugacy classes in Sn . . . . . . . . . . . . . . . . . . . . . . . 21
6.2 Conjugacy classes in An . . . . . . . . . . . . . . . . . . . . . . . 21

7 Quaternions 22

8 Matrix groups 23
8.1 General and special linear groups . . . . . . . . . . . . . . . . . . 23
8.2 Actions of GLn (C) . . . . . . . . . . . . . . . . . . . . . . . . . . 23
8.3 Orthogonal groups . . . . . . . . . . . . . . . . . . . . . . . . . . 23
8.4 Rotations and reflections in R2 and R3 . . . . . . . . . . . . . . . 24
8.5 Unitary groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

9 More on regular polyhedra 27


9.1 Symmetries of the cube . . . . . . . . . . . . . . . . . . . . . . . 27
9.2 Symmetries of the tetrahedron . . . . . . . . . . . . . . . . . . . 27

2
Contents IA Groups (Theorems with proof)

10 Möbius group 28
10.1 Möbius maps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
10.2 Fixed points of Möbius maps . . . . . . . . . . . . . . . . . . . . 29
10.3 Permutation properties of Möbius maps . . . . . . . . . . . . . . 30
10.4 Cross-ratios . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

11 Projective line (non-examinable) 32

3
0 Introduction IA Groups (Theorems with proof)

0 Introduction

4
1 Groups and homomorphisms IA Groups (Theorems with proof)

1 Groups and homomorphisms


1.1 Groups
Proposition. Let (G, ∗) be a group. Then
(i) The identity is unique.
(ii) Inverses are unique.
Proof.
(i) Suppose e and e′ are identities. Then we have ee′ = e′ , treating e as an
inverse, and ee′ = e, treating e′ as an inverse. Thus e = e′ .
(ii) Suppose a−1 and b both satisfy the inverse axiom for some a ∈ G. Then
b = be = b(aa−1 ) = (ba)a−1 = ea−1 = a−1 . Thus b = a−1 .
Proposition. Let (G, ∗) be a group and a, b ∈ G. Then
(i) (a−1 )−1 = a
(ii) (ab)−1 = b−1 a−1
Proof.
(i) Given a−1 , both a and (a−1 )−1 satisfy

xa−1 = a−1 x = e.

By uniqueness of inverses, (a−1 )−1 = a.


(ii) We have

(ab)(b−1 a−1 ) = a(bb−1 )a−1


= aea−1
= aa−1
=e

Similarly, (b−1 a−1 )ab = e. So b−1 a−1 is an inverse of ab. By the uniqueness
of inverses, (ab)−1 = b−1 a−1 .
Lemma (Subgroup criteria I). Let (G, ∗) be a group and H ⊆ G. H ≤ G iff
(i) e ∈ H
(ii) (∀a, b ∈ H) ab ∈ H
(iii) (∀a ∈ H) a−1 ∈ H
Proof. The group axioms are satisfied as follows:
0. Closure: (ii)
1. Identity: (i). Note that H and G must have the same identity. Suppose that
eH and eG are the identities of H and G respectively. Then eH eH = eH .
Now eH has an inverse in G. Thus we have eH eH e−1 −1
H = eH eH . So
eH eG = eG . Thus eH = eG .

5
1 Groups and homomorphisms IA Groups (Theorems with proof)

2. Inverse: (iii)
3. Associativity: inherited from G.
Lemma (Subgroup criteria II). A subset H ⊆ G is a subgroup of G iff:
(I) H is non-empty
(II) (∀a, b ∈ H) ab−1 ∈ H
Proof. (I) and (II) follow trivially from (i), (ii) and (iii).
To prove that (I) and (II) imply (i), (ii) and (iii), we have
(i) H must contain at least one element a. Then aa−1 = e ∈ H.
(iii) ea−1 = a−1 ∈ H.
(ii) a(b−1 )−1 = ab ∈ H.

Proposition. The subgroups of (Z, +) are exactly nZ, for n ∈ N (nZ is the
integer multiples of n).
Proof. Firstly, it is trivial to show that for any n ∈ N, nZ is a subgroup. Now
show that any subgroup must be in the form nZ.
Let H ≤ Z. We know 0 ∈ H. If there are no other elements in H, then
H = 0Z. Otherwise, pick the smallest positive integer n in H. Then H = nZ.
Otherwise, suppose (∃a ∈ H) n ∤ a. Let a = pn + q, where 0 < q < n. Since
a − pn ∈ H, q ∈ H. Yet q < n but n is the smallest member of H. Contradiction.
So every a ∈ H is divisible by n. Also, by closure, all multiples of n must be in
H. So H = nZ.

1.2 Homomorphisms
Lemma. The composition of two bijective functions is bijective
Proposition. Suppose that f : G → H is a homomorphism. Then
(i) Homomorphisms send the identity to the identity, i.e.
f (eG ) = eH

(ii) Homomorphisms send inverses to inverses, i.e.


f (a−1 ) = f (a)−1

(iii) The composite of 2 group homomorphisms is a group homomorphism.


(iv) The inverse of an isomorphism is an isomorphism.
Proof.
(i)
f (eG ) = f (e2G ) = f (eG )2
f (eG )−1 f (eG ) = f (eG )−1 f (eG )2
f (eG ) = eH

6
1 Groups and homomorphisms IA Groups (Theorems with proof)

(ii)

eH = f (eG )
= f (aa−1 )
= f (a)f (a−1 )

Since inverses are unique, f (a−1 ) = f (a)−1 .


(iii) Let f : G1 → G2 and g : G2 → G3 . Then g(f (ab)) = g(f (a)f (b)) =
g(f (a))g(f (b)).
(iv) Let f : G → H be an isomorphism. Then
n  o
f −1 (ab) = f −1 f f −1 (a) f f −1 (b)
 
n  o
= f −1 f f −1 (a)f −1 (b)
= f −1 (a)f −1 (b)

So f −1 is a homomorphism. Since it is bijective, f −1 is an isomorphism.

Proposition. Both the image and the kernel are subgroups of the respective
groups, i.e. im f ≤ H and ker f ≤ G.
Proof. Since eH ∈ im f and eG ∈ ker f , im f and ker f are non-empty. Moreover,
suppose b1 , b2 ∈ im f . Now ∃a1 , a2 ∈ G such that f (ai ) = bi . Then b1 b−1 2 =
f (a1 )f (a−1
2 ) = f (a a−1
1 2 ) ∈ im f .
Then consider b1 , b2 ∈ ker f . We have f (b1 b−1
2 ) = f (b1 )f (b2 )
−1
= e2 = e. So
−1
b1 b2 ∈ ker f .
Proposition. Given any homomorphism f : G → H and any a ∈ G, for all
k ∈ ker f , aka−1 ∈ ker f .
Proof. f (aka−1 ) = f (a)f (k)f (a)−1 = f (a)ef (a)−1 = e. So aka−1 ∈ ker f .

Proposition. For all homomorphisms f : G → H, f is


(i) surjective iff im f = H
(ii) injective iff ker f = {e}

Proof.
(i) By definition.
(ii) We know that f (e) = e. So if f is injective, then by definition ker f = {e}. If
ker f = {e}, then given a, b such that f (a) = f (b), f (ab−1 ) = f (a)f (b)−1 =
e. Thus ab−1 ∈ ker f = {e}. Then ab−1 = e and a = b.

7
1 Groups and homomorphisms IA Groups (Theorems with proof)

1.3 Cyclic groups


Lemma. For a in g, ord(a) = |⟨a⟩|.
Proof. If ord(a) = ∞, an ̸= am for all n ̸= m. Otherwise am−n = e. Thus
|⟨a⟩| = ∞ = ord(a).
Otherwise, suppose ord(a) = k. Thus ak = e. We now claim that ⟨a⟩ =
{e, a1 , a2 , · · · ak−1 }. Note that ⟨a⟩ does not contain higher powers of a as ak = e
and higher powers will loop back to existing elements. There are also no repeating
elements in the list provided since am = an ⇒ am−n = e. So done.
Proposition. Cyclic groups are abelian.

1.4 Dihedral groups


1.5 Direct products of groups
Proposition. Cn × Cm ∼
= Cnm iff hcf(m, n) = 1.
Proof. Suppose that hcf(m, n) = 1. Let Cn = ⟨a⟩ and Cm = ⟨b⟩. Let k be the
order of (a, b). Then (a, b)k = (ak , bk ) = e. This is possible only if n | k and
m | k, i.e. k is a common multiple n and m. Since the order is the minimum
nm
value of k that satisfies the above equation, k = lcm(n, m) = hcf(n,m) = nm.
Now consider ⟨(a, b)⟩ ≤ Cn × Cm . Since (a, b) has order nm, ⟨(a, b)⟩ has nm
elements. Since Cn × Cm also has nm elements, ⟨(a, b)⟩ must be the whole of
Cn × Cm . And we know that ⟨(a, b)⟩ ∼ = Cnm . So Cn × Cm ∼ = Cnm .
On the other hand, suppose hcf(m, n) ̸= 1. Then k = lcm(m, n) ̸= mn. Then
for any (a, b) ∈ Cn × Cm ,we have (a, b)k = (ak , bk ) = e. So the order of any (a, b)
is at most k < mn. So there is no element of order mn. So Cn × Cm is not a
cyclic group of order nm.
Proposition (Direct product theorem). Let H1 , H2 ≤ G. Suppose the following
are true:
(i) H1 ∩ H2 = {e}.
(ii) (∀ai ∈ Hi ) a1 a2 = a2 a1 .
(iii) (∀a ∈ G)(∃ai ∈ Hi ) a = a1 a2 . We also write this as G = H1 H2 .
Then G ∼
= H1 × H2 .
Proof. Define f : H1 × H2 → G by f (a1 , a2 ) = a1 a2 . Then it is a homomorphism
since
f ((a1 , a2 ) ∗ (b1 , b2 )) = f (a1 b1 , a2 b2 )
= a1 b1 a2 b2
= a1 a2 b1 b2
= f (a1 , a2 )f (b1 , b2 ).
Surjectivity follows from (iii). We’ll show injectivity by showing that the kernel
is {e}. If f (a1 , a2 ) = e, then we know that a1 a2 = e. Then a1 = a−1 2 . Since
a1 ∈ H1 and a−1 2 ∈ H 2 , we have a1 = a −1
2 ∈ H 1 ∩ H 2 = {e}. Thus a 1 = a2 = e
and ker f = {e}.

8
2 Symmetric group I IA Groups (Theorems with proof)

2 Symmetric group I
2.1 Symmetric groups
Theorem. Sym X with composition forms a group.
Proof. The groups axioms are satisfied as follows:

0. If σ : X → X and τ : X → X, then σ ◦ τ : X → X. If they are both


bijections, then the composite is also bijective. So if σ, τ ∈ Sym X, then
σ ◦ τ ∈ Sym X.
1. The identity 1X : X → X is clearly a permutation, and gives the identity
of the group.

2. Every bijective function has a bijective inverse. So if σ ∈ Sym X, then


σ −1 ∈ Sym X.
3. Composition of functions is associative.
Lemma. Disjoint cycles commute.

Proof. If σ, τ ∈ Sn are disjoint cycles. Consider any n. Show that: σ(τ (a)) =
τ (σ(a)). If a is in neither of σ and τ , then σ(τ (a)) = τ (σ(a)) = a. Otherwise,
wlog assume that a is in τ but not in σ. Then τ (a) ∈ τ and thus τ (a) ̸∈ σ. Thus
σ(a) = a and σ(τ (a)) = τ (a). Therefore we have σ(τ (a)) = τ (σ(a)) = τ (a).
Therefore τ and σ commute.

Theorem. Any permutation in Sn can be written (essentially) uniquely as a


product of disjoint cycles. (Essentially unique means unique up to re-ordering of
cycles and rotation within cycles, e.g. (1 2) and (2 1))
Proof. Let σ ∈ Sn . Start with (1 σ(1) σ 2 (1) σ 3 (1) · · · ). As the set {1, 2, 3 · · · n}
is finite, for some k, we must have σ k (1) already in the list. If σ k (1) = σ l (1),
with l < k, then σ k−l (1) = 1. So all σ i (1) are distinct until we get back to 1.
Thus we have the first cycle (1 σ(1) σ 2 (1) σ 3 (1) · · · σ k−1 (1)).
Now choose the smallest number that is not yet in a cycle, say j. Repeat to
obtain a cycle (j σ(j) σ 2 (j) · · · σ l−1 (j)). Since σ is a bijection, nothing in this
cycle can be in previous cycles as well.
Repeat until all {1, 2, 3 · · · n} are exhausted. This is essentially unique
because every number j completely determines the whole cycle it belongs to,
and whichever number we start with, we’ll end up with the same cycle.
Lemma. For σ ∈ Sn , the order of σ is the least common multiple of cycle
lengths in the disjoint cycle notation. In particular, a k-cycle has order k.
Proof. As disjoint cycles commute, we can group together each cycle when we take
powers. i.e. if σ = τ1 τ2 · · · τl with τi all disjoint cycles, then σ m = τ1m τ2m · · · τlm .
Now if cycle τi has length ki , then τiki = e, and τim = e iff ki | m. To get an
m such that σ m = e, we need all ki to divide m. i.e. m is a common multiple of
ki . Since the order is the least possible m such that σ m = e, the order is the
least common multiple of ki .

9
2 Symmetric group I IA Groups (Theorems with proof)

2.2 Sign of permutations


Proposition. Every permutation is a product of transpositions.

Proof. As each permutation is a product of disjoint cycles, it suffices to prove


that each cycle is a product of transpositions. Consider a cycle (a1 a2 a3 · · · ak ).
This is in fact equal to (a1 a2 )(a2 a3 ) · · · (ak−1 ak ). Thus a k-cycle can be written
as a product of k − 1 transpositions.

Theorem. Writing σ ∈ Sn as a product of transpositions in different ways, σ is


either always composed of an even number of transpositions, or always an odd
number of transpositions.
Proof. Write #(σ) for the number of cycles in disjoint cycle notation, including
singleton cycles. So #(e) = n and #((1 2)) = n − 1. When we multiply σ by a
transposition τ = (c d) (wlog assume c < d),

– If c, d are in the same σ-cycle, say, (c a2 · · · ak−1 d ak+1 · · · ak+l )(c d) =


(c ak+1 ak+2 · · · ak+l )(d a2 a3 · · · ak−1 ). So #(στ ) = #(σ) + 1 .
– If c, d are in different σ-cycles, say
(d a2 a3 · · · ak−1 )(c ak+1 ak+2 · · · ak+l )(c d)
= (c a2 · · · ak−1 d ak+1 · · · ak+l )(c d)(c d)
= (c a2 · · · ak−1 d ak+1 · · · ak+l ) and #(στ ) = #(σ) − 1.
Therefore for any transposition τ , #(στ ) ≡ #(σ) + 1 (mod 2).
Now suppose σ = τ1 · · · τl = τ1′ · · · τk′ . Since disjoint cycle notation is unique,
#(σ) is uniquely determined by σ.
Now we can construct σ by starting with e and multiplying the transpositions
one by one. Each time we add a transposition, we increase #(σ) by 1 (mod 2).
So #(σ) ≡ #(e) + l (mod 2). Similarly, #(σ) ≡ #(e) + k (mod 2). So l ≡ k
(mod 2).
Theorem. For n ≥ 2, sgn : Sn → {±1} is a surjective group homomorphism.

Proof. Suppose σ1 = τ1 · · · τl1 and σ2 = τ1′ · · · τl2 . Then sgn(σ1 σ2 ) = (−1)l1 +l2 =
(−1)l1 (−1)l2 = sgn(σ1 ) sgn(σ2 ). So it is a homomorphism.
It is surjective since sgn(e) = 1 and sgn((1 2)) = −1.
Lemma. σ is an even permutation iff the number of cycles of even length is
even.
Proof. A k-cycle can be written as k − 1 transpositions. Thus an even-length
cycle is odd, vice versa.
Since sgn is a group homomorphism, writing σ in disjoint cycle notation,
σ = σ1 σ2 · · · σl , we get sgn(σ) = sgn(σ1 ) · · · sgn(σl ). Suppose there are m even-
length cycles and n odd-length cycles, then sgn(σ) = (−1)m 1n . This is equal to
1 iff (−1)m = 1, i.e. m is even.
Proposition. Any subgroup of Sn contains either no odd permutations or
exactly half.

10
2 Symmetric group I IA Groups (Theorems with proof)

Proof. If Sn has at least one odd permutation τ , then there exists a bijection
between the odd and even permutations by σ 7→ στ (bijection since σ 7→ στ −1
is a well-defined inverse). So there are as many odd permutations as even
permutations.

11
3 Lagrange’s Theorem IA Groups (Theorems with proof)

3 Lagrange’s Theorem
Proposition. aH = bH ⇔ b−1 a ∈ H.
Proof. (⇒) Since a ∈ aH, a ∈ bH. Then a = bh for some h ∈ H. So b−1 a =
h ∈ H.
(⇐). Let b−1 a = h0 . Then a = bh0 . Then ∀ah ∈ aH, we have ah = b(h0 h) ∈
bH. So aH ⊆ bH. Similarly, bH ⊆ aH. So aH = bH.
Lemma. The left cosets of a subgroup H ≤ G partition G, and every coset has
the same size.

Proof. For each a ∈ G, a ∈ aH. Thus the union of all cosets gives all of G. Now
we have to show that for all a, b ∈ G, the cosets aH and bH are either the same
or disjoint.
Suppose that aH and bH are not disjoint. Let ah1 = bh2 ∈ aH ∩ bH. Then
b−1 a = h2 h−1
1 ∈ H. So aH = bH.
To show that they each coset has the same size, note that f : H → aH with
f (h) = ah is invertible with inverse f −1 (h) = a−1 h. Thus there exists a bijection
between them and they have the same size.
Theorem (Lagrange’s theorem). If G is a finite group and H is a subgroup of
G, then |H| divides |G|. In particular,

|H||G : H| = |G|.

Proof. Suppose that there are |G : H| left cosets in total. Since the left cosets
partition G, and each coset has size |H|, we have

|H||G : H| = |G|.

Corollary. The order of an element divides the order of the group, i.e. for any
finite group G and a ∈ G, ord(a) divides |G|.
Proof. Consider the subgroup generated by a, which has order ord(a). Then by
Lagrange’s theorem, ord(a) divides |G|.
Corollary. The exponent of a group divides the order of the group, i.e. for any
finite group G and a ∈ G, a|G| = e.
Proof. We know that |G| = k ord(a) for some k ∈ N. Then a|G| = (aord(a) )k =
ek = e.
Corollary. Groups of prime order are cyclic and are generated by every non-
identity element.
Proof. Say |G| = p. If a ∈ G is not the identity, the subgroup generated by a
must have order p since it has to divide p. Thus the subgroup generated by a
has the same size as G and they must be equal. Then G must be cyclic since it
is equal to the subgroup generated by a.

Proposition. The equivalence classes form a partition of A.

12
3 Lagrange’s Theorem IA Groups (Theorems with proof)

Proof. By reflexivity, we have a ∈ [a]. Thus the equivalence classes cover the
whole set. We must now show that for all a, b ∈ A, either [a] = [b] or [a] ∩ [b] = ∅.
Suppose [a] ∩ [b] ̸= ∅. Then ∃c ∈ [a] ∩ [b]. So a ∼ c, b ∼ c. By symmetry, c ∼ b.
By transitivity, we have a ∼ b. Now for all b′ ∈ [b], we have b ∼ b′ . Thus by
transitivity, we have a ∼ b′ . Thus [b] ⊆ [a]. Similarly, [a] ⊆ [b] and [a] = [b].
Lemma. Given a group G and a subgroup H, define the equivalence relation
on G with a ∼ b iff b−1 a ∈ H. The equivalence classes are the left cosets of H.

Proof. First show that it is an equivalence relation.


(i) Reflexivity: Since aa−1 = e ∈ H, a ∼ a.
(ii) Symmetry: a ∼ b ⇒ b−1 a ∈ H ⇒ (b−1 a)−1 = a−1 b ∈ H ⇒ b ∼ a.
(iii) Transitivity: If a ∼ b and b ∼ c, we have b−1 a, c−1 b ∈ H. So c−1 bb−1 a =
c−1 a ∈ H. So a ∼ c.
To show that the equivalence classes are the cosets, we have a ∼ b ⇔ b−1 a ∈
H ⇔ aH = bH.
Proposition. Un is a group under multiplication mod n.

Proof. The operation is well-defined as shown above. To check the axioms:


0. Closure: if a, b are coprime to n, then a · b is also coprime to n. So
[a], [b] ∈ Un ⇒ [a] · [b] = [a · b] ∈ Un
1. Identity: [1]

2. Let [a] ∈ Un . Consider the map Un → Un with [c] 7→ [ac]. This is injective:
if [ac1 ] = [ac2 ], then n divides a(c1 − c2 ). Since a is coprime to n, n divides
c1 − c2 , so [c1 ] = [c2 ]. Since Un is finite, any injection (Un → Un ) is also a
surjection. So there exists a c such that [ac] = [a][c] = 1. So [c] = [a]−1 .
3. Associativity (and also commutativity): inherited from Z.

Theorem (Fermat-Euler theorem). Let n ∈ N and a ∈ Z coprime to n. Then

aϕ(n) ≡ 1 (mod n).

In particular, (Fermat’s Little Theorem) if n = p is a prime, then for any a not


a multiple of p.
ap−1 ≡ 1 (mod p).
Proof. As a is coprime with n, [a] ∈ Un . Then [a]|Un | = [1], i.e. aϕ(n) ≡ 1
(mod n).

3.1 Small groups


Proposition. Any group of order 4 is either isomorphic to C4 or C2 × C2 .

13
3 Lagrange’s Theorem IA Groups (Theorems with proof)

Proof. Let |G| = 4. By Lagrange theorem, possible element orders are 1 (e only),
2 and 4. If there is an element a ∈ G of order 4, then G = ⟨a⟩ ∼
= C4 .
Otherwise all non-identity elements have order 2. Then G must be abelian
(For any a, b, (ab)2 = 1 ⇒ ab = (ab)−1 ⇒ ab = b−1 a−1 ⇒ ab = ba). Pick
2 elements of order 2, say b, c ∈ G, then ⟨b⟩ = {e, b} and ⟨c⟩ = {e, c}. So
⟨b⟩ ∩ ⟨c⟩ = {e}. As G is abelian, ⟨b⟩ and ⟨c⟩ commute. We know that bc = cb has
order 2 as well, and is the only element of G left. So G ∼
= ⟨b⟩ × ⟨c⟩ ∼= C2 × C2
by the direct product theorem.

Proposition. A group of order 6 is either cyclic or dihedral (i.e. is isomorphic


to C6 or D6 ). (See proof in next section)

3.2 Left and right cosets

14
4 Quotient groups IA Groups (Theorems with proof)

4 Quotient groups
4.1 Normal subgroups
Lemma.
(i) Every subgroup of index 2 is normal.

(ii) Any subgroup of an abelian group is normal.


Proof.
(i) If K ≤ G has index 2, then there are only two possible cosets K and G \ K.
As eK = Ke and cosets partition G, the other left coset and right coset
must be G \ K. So all left cosets and right cosets are the same.

(ii) For all a ∈ G and k ∈ K, we have aka−1 = aa−1 k = k ∈ K.


Proposition. Every kernel is a normal subgroup.
Proof. Given homomorphism f : G → H and some a ∈ G, for all k ∈ ker f , we
have f (aka−1 ) = f (a)f (k)f (a)−1 = f (a)ef (a)−1 = e. Therefore aka−1 ∈ ker f
by definition of the kernel.
Proposition. A group of order 6 is either cyclic or dihedral (i.e. ∼
= C6 or D6 ).
Proof. Let |G| = 6. By Lagrange theorem, possible element orders are 1, 2, 3 and
6. If there is an a ∈ G of order 6, then G = ⟨a⟩ ∼= C6 . Otherwise, we can only
have elements of orders 2 and 3 other than the identity. If G only has elements
of order 2, the order must be a power of 2 by Sheet 1 Q. 8, which is not the case.
So there must be an element r of order 3. So ⟨r⟩ ◁ G as it has index 2. Now G
must also have an element s of order 2 by Sheet 1 Q. 9.
Since ⟨r⟩ is normal, we know that srs−1 ∈ ⟨r⟩. If srs−1 = e, then r = e,
which is not true. If srs−1 = r, then sr = rs and sr has order 6 (lcm of the
orders of s and r), which was ruled out above. Otherwise if srs−1 = r2 = r−1 ,
then G is dihedral by definition of the dihedral group.

4.2 Quotient groups


Proposition. Let K ◁ G. Then the set of (left) cosets of K in G is a group
under the operation aK ∗ bK = (ab)K.
Proof. First show that the operation is well-defined. If aK = a′ K and bK = b′ K,
we want to show that aK ∗ bK = a′ K ∗ b′ K. We know that a′ = ak1 and b′ = bk2
for some k1 , k2 ∈ K. Then a′ b′ = ak1 bk2 . We know that b−1 k1 b ∈ K. Let
b−1 k1 b = k3 . Then k1 b = bk3 . So a′ b′ = abk3 k2 ∈ (ab)K. So picking a different
representative of the coset gives the same product.
1. Closure: If aK, bK are cosets, then (ab)K is also a coset
2. Identity: The identity is eK = K (clear from definition)
3. Inverse: The inverse of aK is a−1 K (clear from definition)

4. Associativity: Follows from the associativity of G.

15
4 Quotient groups IA Groups (Theorems with proof)

Lemma. Given K ◁ G, the quotient map q : G → G/K with g 7→ gK is a


surjective group homomorphism.

Proof. q(ab) = (ab)K = aKbK = q(a)q(b). So q is a group homomorphism.


Also for all aK ∈ G/K, q(a) = aK. So it is surjective.
Proposition. The quotient of a cyclic group is cyclic.
Proof. Let G = Cn with H ≤ Cn . We know that H is also cyclic. Say Cn = ⟨c⟩
and H = ⟨ck ⟩ ∼
= Cℓ , where kℓ = n. We have Cn /H = {H, cH, c2 H, · · · ck−1 H} =
⟨cH⟩ ∼
= Ck .

4.3 The Isomorphism Theorem


Theorem (The Isomorphism Theorem). Let f : G → H be a group homomor-
phism with kernel K. Then K ◁ G and G/K ∼
= im f .
Proof. We have proved that K ◁ G before. We define a group homomorphism
θ : G/K → im f by θ(aK) = f (a).
First check that this is well-defined: If a1 K = a2 K, then a−1
2 a1 ∈ K. So

f (a2 )−1 f (a1 ) = f (a−1


2 a1 ) = e.

So f (a1 ) = f (a2 ) and θ(a1 K) = θ(a2 K).


Now we check that it is a group homomorphism:

θ(aKbK) = θ(abK) = f (ab) = f (a)f (b) = θ(aK)θ(bK).

To show that it is injective, suppose θ(aK) = θ(bK). Then f (a) = f (b). Hence
f (b)−1 f (a) = e. Hence b−1 a ∈ K. So aK = bK.
By definition, θ is surjective since im θ = im f . So θ gives an isomorphism
G/K ∼ = im f ≤ H.
Lemma. Any cyclic group is isomorphic to either Z or Z/(nZ) for some n ∈ N.

Proof. Let G = ⟨c⟩. Define f : Z → G with m 7→ cm . This is a group


homomorphism since cm1 +m2 = cm1 cm2 . f is surjective since G is by definition
all cm for all m. We know that ker f ◁ Z. We have three possibilities. Either
(i) ker f = {e}, so F is an isomorphism and G ∼
= Z; or
(ii) ker f = Z, then G ∼
= Z/Z = {e} = C1 ; or
(iii) ker f = nZ (since these are the only proper subgroups of Z), then G ∼
=
Z/(nZ).

16
5 Group actions IA Groups (Theorems with proof)

5 Group actions
5.1 Group acting on sets
Proposition. Let X be a set and G be a group. Then φ : G → Sym X is a
homomorphism (i.e. an action) iff θ : G × X → X defined by θ(g, x) = φ(g)(x)
satisfies

0. (∀g ∈ G)(x ∈ X) θ(g, x) ∈ X.


1. (∀x ∈ X) θ(e, x) = x.
2. (∀g, h ∈ G)(∀x ∈ X) θ(g, θ(h, x)) = θ(gh, x).

5.2 Orbits and Stabilizers


Lemma. stab(x) is a subgroup of G.
Proof. We know that e(x) = x by definition. So stab(x) is non-empty. Suppose
g, h ∈ stab(x), then gh−1 (x) = g(h−1 (x)) = g(x) = x. So gh−1 ∈ stab(X). So
stab(x) is a subgroup.
Lemma. The orbits of an action partition X.
Proof. Firstly, (∀x)(x ∈ orb(x)) as e(x) = x. So every x is in some orbit.
Then suppose z ∈ orb(x) and z ∈ orb(y), we have to show that orb(x) =
orb(y). We know that z = g1 (x) and z = g2 (y) for some g1 , g2 . Then g1 (x) =
g2 (y) and y = g2−1 g1 (x).
For any w = g3 (y) ∈ orb(y), we have w = g3 g2−1 g1 (x). So w ∈ orb(x). Thus
orb(y) ⊆ orb(x) and similarly orb(x) ⊆ orb(y). Therefore orb(x) = orb(y).
Theorem (Orbit-stabilizer theorem). Let the group G act on X. Then there
is a bijection between orb(x) and cosets of stab(x) in G. In particular, if G is
finite, then
| orb(x)|| stab(x)| = |G|.
Proof. We biject the cosets of stab(x) with elements in the orbit of x. Recall
that G : stab(x) is the set of cosets of stab(x). We can define

θ : (G : stab(x)) → orb(x)
g stab(x) 7→ g(x).

This is well-defined — if g stab(x) = h stab(x), then h = gk for some k ∈ stab(x).


So h(x) = g(k(x)) = g(x).
This map is surjective since for any y ∈ orb(x), there is some g ∈ G such
that g(x) = y, by definition. Then θ(g stab(x)) = y. It is injective since if
g(x) = h(x), then h−1 g(x) = x. So h−1 g ∈ stab(x). So g stab(x) = h stab(x).
Hence the number of cosets is | orb(x)|. Then the result follows from La-
grange’s theorem.

17
5 Group actions IA Groups (Theorems with proof)

5.3 Important actions


Lemma (Left regular action). Any group G acts on itself by left multiplication.
This action is faithful and transitive.
Proof. We have
1. (∀g ∈ G)(x ∈ G) g(x) = g · x ∈ G by definition of a group.
2. (∀x ∈ G) e · x = x by definition of a group.

3. g(hx) = (gh)x by associativity.


So it is an action.
To show that it is faithful, we want to know that [(∀x ∈ X) gx = x] ⇒ g = e.
This follows directly from the uniqueness of identity.
To show that it is transitive, ∀x, y ∈ G, then (yx−1 )(x) = y. So any x can
be sent to any y.
Theorem (Cayley’s theorem). Every group is isomorphic to some subgroup of
some symmetric group.
Proof. Take the left regular action of G on itself. This gives a group homo-
morphism φ : G → Sym G with ker φ = {e} as the action is faithful. By the
isomorphism theorem, G ∼ = im φ ≤ Sym G.
Lemma (Left coset action). Let H ≤ G. Then G acts on the left cosets of H
by left multiplication transitively.

Proof. First show that it is an action:


0. g(aH) = (ga)H is a coset of H.
1. e(aH) = (ea)H = aH.
2. g1 (g2 (aH)) = g1 ((g2 a)H) = (g1 g2 a)H = (g1 g2 )(aH).

To show that it is transitive, given aH, bH, we know that (ba−1 )(aH) = bH.
So any aH can be mapped to bH.
Lemma (Conjugation action). Any group G acts on itself by conjugation (i.e.
g(x) = gxg −1 ).

Proof. To show that this is an action, we have


0. g(x) = gxg −1 ∈ G for all g, x ∈ G.
1. e(x) = exe−1 = x
2. g(h(x)) = g(hxh−1 ) = ghxh−1 g −1 = (gh)x(gh)−1 = (gh)(x)

Lemma. Let K ◁ G. Then G acts by conjugation on K.


Proof. We only have to prove closure as the other properties follow from the
conjugation action. However, by definition of a normal subgroup, for every
g ∈ G, k ∈ K, we have gkg −1 ∈ K. So it is closed.

18
5 Group actions IA Groups (Theorems with proof)

Proposition. Normal subgroups are exactly those subgroups which are unions
of conjugacy classes.

Proof. Let K ◁ G. If k ∈ K, then by definition for every g ∈ G, we get


gkg −1 ∈ K. So ccl(k) ⊆ K. So K is the union of the conjugacy classes of all its
elements.
Conversely, if K is a union of conjugacy classes and a subgroup of G, then
for all k ∈ K, g ∈ G, we have gkg −1 ∈ K. So K is normal.

Lemma. Let X be the set of subgroups of G. Then G acts by conjugation on


X.
Proof. To show that it is an action, we have
0. If H ≤ G, then we have to show that gHg −1 is also a subgroup. We
know that e ∈ H and thus geg −1 = e ∈ gHg −1 , so gHg −1 is non-empty.
For any two elements gag −1 and gbg −1 ∈ gHg −1 , (gag −1 )(gbg −1 )−1 =
g(ab−1 )g −1 ∈ gHg −1 . So gHg −1 is a subgroup.
1. eHe−1 = H.
2. g1 (g2 Hg2−1 )g1−1 = (g1 g2 )H(g1 g2 )−1 .

Lemma. Stabilizers of the elements in the same orbit are conjugate, i.e. let G
act on X and let g ∈ G, x ∈ X. Then stab(g(x)) = g stab(x)g −1 .

5.4 Applications
Proof. Consider the left coset action of G on H. We get a group homomorphism
φ : G → Sn since there are n cosets of H. Since H = ̸ G, φ is non-trivial and
ker φ ̸= G. Now ker φ ◁ G. Since G is simple, ker φ = {e}. So G ∼ = im φ ⊆ Sn
by the isomorphism theorem. So |G| ≤ |Sn | = n!.
We can further refine this by considering sgn ◦φ : G → {±1}. The kernel
of this composite is normal in G. So K = ker(sgn ◦ϕ) = {e} or G. Since
G/K ∼ = im(sgn ◦ϕ), we know that |G|/|K| = 1 or 2 since im(sgn ◦ϕ) has at most
two elements. Hence for |G| > 2, we cannot have K = {e}, or else |G|/|K| > 2.
So we must have K = G, so sgn(φ(g)) = 1 for all g and im φ ≤ An . So
|G| ≤ n!/2
Theorem (Cauchy’s Theorem). Let G be a finite group and prime p dividing
|G|. Then G has an element of order p (in fact there must be at least p − 1
elements of order p).
Proof. Let G and p be fixed. Consider Gp = G × G × · · · × G, the set of p-tuples
of G. Let X ⊆ Gp be X = {(a1 , a2 , · · · , ap ) ∈ Gp : a1 a2 · · · ap = e}.
In particular, if an element b has order p, then (b, b, · · · , b) ∈ X. In fact, if
(b, b, · · · , b) ∈ X and b ̸= e, then b has order p, since p is prime.
Now let H = ⟨h : hp = e⟩ ∼ = Cp be a cyclic group of order p with generator h
(This h is not related to G in any way). Let H act on X by “rotation”:

h(a1 , a2 , · · · , ap ) = (a2 , a3 , · · · , ap , a1 )

This is an action:

19
5 Group actions IA Groups (Theorems with proof)

0. If a1 · · · ap = e, then a−1 1 = a2 · · · ap . So a2 · · · ap a1 = a−1


1 a1 = e. So
(a2 , a3 , · · · , ap , a1 ) ∈ X.

1. e acts as an identity by construction


2. The “associativity” condition also works by construction.
As orbits partition X, the sum of all orbit sizes must be |X|. We know that
|X| = |G|p−1 since we can freely choose the first p − 1 entries and the last one
must be the inverse of their product. Since p divides |G|, p also divides |X|. We
have | orb(a1 , · · · , ap )|| stabH (a1 , · · · , ap )| = |H| = p. So all orbits have size 1 or
p, and they sum to |X| = p× something. We know that there is one orbit of size
1, namely (e, e, · · · , e). So there must be at least p − 1 other orbits of size 1 for
the sum to be divisible by p.
In order to have an orbit of size 1, they must look like (a, a, · · · , a) for some
a ∈ G, which has order p.

20
6 Symmetric groups II IA Groups (Theorems with proof)

6 Symmetric groups II
6.1 Conjugacy classes in Sn
Proposition. If (a1 a2 · · · ak ) is a k-cycle and ρ ∈ Sn , then ρ(a1 · · · ak )ρ−1 is
the k-cycle (ρ(a1 ) ρ(a2 ) · · · ρ(a3 )).
Proof. Consider any ρ(a1 ) acted on by ρ(a1 · · · ak )ρ−1 . The three permutations
send it to ρ(a1 ) 7→ a1 7→ a2 7→ ρ(a2 ) and similarly for other ai s. Since ρ is
bijective, any b can be written as ρ(a) for some a. So the result is the k-cycle
(ρ(a1 ) ρ(a2 ) · · · ρ(a3 )).
Corollary. Two elements in Sn are conjugate iff they have the same cycle type.
Proof. Suppose σ = σ1 σ2 · · · σℓ , where σi are disjoint cycles. Then ρσρ−1 =
ρσ1 ρ−1 ρσ2 ρ−1 · · · ρσℓ ρ−1 . Since the conjugation of a cycle conserves its length,
ρσρ−1 has the same cycle type.
Conversely, if σ, τ have the same cycle type, say

σ = (a1 a2 · · · ak )(ak+1 · · · ak+ℓ ), τ = (b1 b2 · · · bk )(bk+1 · · · bk+ℓ ),

if we let ρ(ai ) = bi , then ρσρ−1 = τ .

6.2 Conjugacy classes in An


Proposition. For σ ∈ An , the conjugacy class of σ splits in An if and only if
no odd permutation commutes with σ.
Proof. We have the conjugacy classes splitting if and only if the centralizer does
not. So instead we check whether the centralizer splits. Clearly CAn (σ) =
CSn (σ) ∩ An . So splitting of centralizer occurs if and only if an odd permutation
commutes with σ.

Lemma. σ = (1 2 3 4 5) ∈ S5 has CS5 (σ) = ⟨σ⟩.


Proof. | cclSn (σ)| = 24 and |S5 | = 120. So |CS5 (σ)| = 5. Clearly ⟨σ⟩ ⊆ CS5 (σ).
Since they both have size 5, we know that CS5 (σ) = ⟨σ⟩
Theorem. A5 is simple.

Proof. We know that normal subgroups must be unions of the conjugacy classes,
must contain e and their order must divide 60. The possible orders are 1, 2,
3, 4, 5, 6, 10, 12, 15, 20, 30. However, the conjugacy classes 1, 15, 20, 12, 12
cannot add up to any of the possible orders apart from 1 and 60. So we only
have trivial normal subgroups.

21
7 Quaternions IA Groups (Theorems with proof)

7 Quaternions
Lemma. If G has order 8, then either G is abelian (i.e. ∼ = C8 , C4 × C2 or
C2 × C2 × C2 ), or G is not abelian and isomorphic to D8 or Q8 (dihedral or
quaternion).

Proof. Consider the different possible cases:


– If G contains an element of order 8, then G ∼
= C8 .
– If all non-identity elements have order 2, then G is abelian (Sheet 1, Q8).
Let a = ̸ b ∈ G \ {e}. By the direct product theorem, ⟨a, b⟩ = ⟨a⟩ × ⟨b⟩.
Then take c ̸∈ ⟨a, b⟩. By the direct product theorem, we obtain ⟨a, b, c⟩ =
⟨a⟩ × ⟨b⟩ × ⟨c⟩ = C2 × C2 × C2 . Since ⟨a, b, c⟩ ⊆ G and |⟨a, b, c⟩| = |G|,
G = ⟨a, b, c⟩ ∼
= C2 × C2 × C2 .
– G has no element of order 8 but has an order 4 element a ∈ G. Let
H = ⟨a⟩. Since H has index 2, it is normal in G. So G/H ∼ = C2 since
|G/H| = 2. This means that for any b ̸∈ H, bH generates G/H. Then
(bH)2 = b2 H = H. So b2 ∈ H. Since b2 ∈ ⟨a⟩ and ⟨a⟩ is a cyclic group, b2
commutes with a.
If b2 = a or a3 , then b has order 8. Contradiction. So b2 = e or a2 .
We also know that H is normal, so bab−1 ∈ H. Let bab−1 = aℓ . Since
a and b2 commute, we know that a = b2 ab−2 = b(bab−1 )b−1 = baℓ b−1 =
2
(bab−1 )ℓ = aℓ . So ℓ2 ≡ 1 (mod 4). So ℓ ≡ ±1 (mod 4).

◦ When l ≡ 1 (mod 4), bab−1 = a, i.e. ba = ab. So G is abelian.


∗ If b2 = e, then G = ⟨a, b⟩ ∼
= ⟨a⟩ × ⟨b⟩ ∼
= C4 × C2 .
∗ If b = a , then (ba ) = e. So G = ⟨a, ba−1 ⟩ ∼
2 2 −1 2
= C4 × C2 .
◦ If l ≡ −1 (mod 4), then bab−1 = a−1 .
∗ If b2 = e, then G = ⟨a, b : a4 = e = b2 , bab−1 = a−1 ⟩. So G ∼
= D8
by definition.
∗ If b2 = a2 , then we have G ∼= Q8 .

22
8 Matrix groups IA Groups (Theorems with proof)

8 Matrix groups
8.1 General and special linear groups
Proposition. GLn (F ) is a group.
Proof. Identity is I, which is in GLn (F ) by definition (I is its self-inverse). The
composition of invertible matrices is invertible, so is closed. Inverse exist by
definition. Multiplication is associative.
Proposition. det : GLn (F ) → F \ {0} is a surjective group homomorphism.
Proof. det AB = det A det B. If A is invertible, it has non-zero determinant and
det A ∈ F \ {0}.
To show it is surjective, for any x ∈ F \ {0}, if we take the identity matrix
and replace I11 with x, then the determinant is x. So it is surjective.

8.2 Actions of GLn (C)


Proposition. GLn (C) acts faithfully on Cn by left multiplication to the vector,
with two orbits (0 and everything else).
Proof. First show that it is a group action:
1. If A ∈ GLn (C) and v ∈ Cn , then Av ∈ Cn . So it is closed.
2. Iv = v for all v ∈ Cn .
3. A(Bv) = (AB)v.
Now prove that it is faithful: a linear map is determined by what it does
on a basis. Take the standard basis e1 = (1, 0, · · · , 0), · · · en = (0, · · · , 1). Any
matrix which maps each ek to itself must be I (since the columns of a matrix
are the images of the basis vectors)
To show that there are 2 orbits, we know that A0 = 0 for all A. Also, as A
is invertible, Av = 0 ⇔ v = 0. So 0 forms a singleton orbit. Then given any
two vectors v ̸= w ∈ Cn \ {0}, there is a matrix A ∈ GLn (C) such that Av = w
(cf. Vectors and Matrices).
Proposition. GLn (C) acts on Mn×n (C) by conjugation. (Proof is trivial)

8.3 Orthogonal groups


Lemma (Orthogonal matrices are isometries). For any orthogonal A and x, y ∈
Rn , we have
(i) (Ax) · (Ay) = x · y
(ii) |Ax| = |x|
Proof. Treat the dot product as a matrix multiplication. So

(Ax)T (Ay) = xT AT Ay = xT Iy = xT y

Then we have |Ax|2 = (Ax) · (Ax) = x · x = |x|2 . Since both are positive, we
know that |Ax| = |x|.

23
8 Matrix groups IA Groups (Theorems with proof)

Lemma. The orthogonal group is a group.


Proof. We have to check that it is a subgroup of GLn (R): It is non-empty,
since I ∈ O(n). If A, B ∈ O(n), then (AB −1 )(AB −1 )T = AB −1 (B −1 )T AT =
AB −1 BA−1 = I, so AB −1 ∈ O(n) and this is indeed a subgroup.
Proposition. det : O(n) → {±1} is a surjective group homomorphism.
Proof. For A ∈ O(n), we know that AT A = I. So det AT A = (det A)2 = 1. So
det A = ±1. Since det(AB) = det A det B, it is a homomorphism. We have
 
−1 0 · · · 0
 0 1 · · · 0
det I = 1, det  .  = −1,
 
.. . .
 .. . . 0
0 0 ··· 1
so it is surjective.
 
−1 0 ··· 0
0 1 ··· 0
Lemma. O(n) = SO(n) ∪  .  SO(n)
 
.. . .
 .. . . 0
0 0 ··· 1
Proof. Cosets partition the group.

8.4 Rotations and reflections in R2 and R3


Lemma. SO(2) consists of all rotations of R2 around 0.
 
a b
Proof. Let A ∈ SO(2). So AT A = I and det A = 1. Suppose A = .
c d
 
−1 d −b T −1
Then A = . So A = A implies ad − bc = 1, c = −b, d = a.
−c a
Combining these equations we obtain a2 + c2 = 1. Set a = cos θ = d, and
c = sin θ = −b. Then these satisfies all three equations. So
 
cos θ − sin θ
A= .
sin θ cos θ

Note that A maps (1, 0) to (cos θ, sin θ), and maps (0, 1) = (− sin θ, cos θ), which
are rotations by θ counterclockwise. So A represents a rotation by θ.
Corollary. Any matrix in O(2) is either a rotation around 0 or a reflection in a
line through 0.
Proof. If A ∈ SO(2), we’ve show that it is a rotation. Otherwise,
    
1 0 cos θ − sin θ cos θ − sin θ
A= =
0 −1 sin θ cos θ − sin θ − cos θ
 
1 0
since O(2) = SO(2) ∪ SO(2). This has eigenvalues 1, −1. So it is a
0 −1
reflection in the line of the eigenspace E1 . The line goes through 0 since the
eigenspace is a subspace which must include 0.

24
8 Matrix groups IA Groups (Theorems with proof)

Lemma. Every matrix in SO(3) is a rotation around some axis.


Proof. Let A ∈ SO(3). We know that det A = 1 and A is an isometry. The
eigenvalues λ must have |λ| = 1. They also multiply to det A = 1. Since we are
in R, complex eigenvalues come in complex conjugate pairs. If there are complex
eigenvalues λ and λ̄, then λλ̄ = |λ|2 = 1. The third eigenvalue must be real and
has to be +1.
If all eigenvalues are real. Then eigenvalues are either 1 or −1, and must
multiply to 1. The possibilities are 1, 1, 1 and −1, −1, 1, all of which contain an
eigenvalue of 1.
So pick an eigenvector for our eigenvalue 1 as the third basis vector. Then in
some orthonormal basis,  
a b 0
A =  c d 0
0 0 1
Since the third column is the image of the third basis vector, and by orthogonality
the third row is 0, 0, 1. Now let
 
a b
A′ = ∈ GL2 (R)
c d

with det A′ = 1. A′ is still orthogonal, so A′ ∈ SO(2). Therefore A′ is a rotation


and  
cos θ − sin θ 0
A =  sin θ cos θ 0
0 0 1
in some basis, and this is exactly the rotation through an axis.
Lemma. Every matrix in O(3) is the product of at most three reflections in
planes through 0.
 
1 0 0
Proof. Recall O(3) = SO(3) ∪ 0 1 0  SO(3). So if A ∈ SO(3), we know
  0 0 −1
cos θ − sin θ 0
that A =  sin θ cos θ 0 in some basis, which is a composite of two
0 0 1
reflections:   
1 0 0 cos θ sin θ 0
A = 0 −1 0  sin θ − cos θ 0 ,
0 0 1 0 0 1
 
1 0 0
Then if A ∈ 0 1 0  SO(3), then it is automatically a product of three
0 0 −1
reflections.

8.5 Unitary groups


Lemma. det : U(n) → S 1 , where S 1 is the unit circle in the complex plane, is a
surjective group homomorphism.

25
8 Matrix groups IA Groups (Theorems with proof)

Proof. We know that 1 = det I = det A† A = | det A|2 . So | det A| = 1. Since


det AB = det A det B, it is a group
 homomorphism.

λ 0 ··· 0
 0 1 · · · 0
Now given λ ∈ S1 , we have  . . . ∈ U(n). So it is surjective.
 
 .. .. . . 0

0 0 0 1

26
9 More on regular polyhedra IA Groups (Theorems with proof)

9 More on regular polyhedra


9.1 Symmetries of the cube
Proposition. G+ ∼
= S4 , where G+ is the group of all rotations of the cube.
Proof. Consider G+ acting on the 4 diagonals of the cube. This gives a group
homomorphism φ : G+ → S4 . We have (1 2 3 4) ∈ im φ by rotation around
the axis through the top and bottom face. We also (1 2) ∈ im φ by rotation
around the axis through the mid-point of the edge connect 1 and 2. Since (1 2)
and (1 2 3 4) generate S4 (Sheet 2 Q. 5d), im φ = S4 , i.e. ϕ is surjective. Since
|S4 | = |G+ |, φ must be an isomorphism.
Proposition. G ∼
= S4 × C2 , where G is the group of all symmetries of the cube.
Proof. Let τ be “reflection in mid-point” as shown above. This commutes with
everything. (Actually it is enough to check that it commutes with rotations
only)
We have to show that G = G+ ⟨τ ⟩. This can be deduced using sizes: since
+
G and ⟨τ ⟩ intersect at e only, (i) and (ii) of the Direct Product Theorem gives
an injective group homomorphism G+ × ⟨τ ⟩ → G. Since both sides have the
same size, the homomorphism must be surjective as well. So G ∼ = G+ × ⟨τ ⟩ ∼
=
S4 × C2 .

9.2 Symmetries of the tetrahedron

27
10 Möbius group IA Groups (Theorems with proof)

10 Möbius group
10.1 Möbius maps
Lemma. The Möbius maps are bijections C∞ → C∞ .
az+b dz−b
Proof. The inverse of f (z) = cz+d is g(z) = −cz+a , which we can check by
composition both ways.

Proposition. The Möbius maps form a group M under function composition.


(The Möbius group)
Proof. The group axioms are shown as follows:
 
a1 z+b1
a2 c1 z+d1 + b2
a1 z+b1 a2 z+b2
0. If f1 (z) = c1 z+d1 and f2 (z) = c2 z+d2 , then f2 ◦f1 (z) =   =
a1 z+b1
c2 c1 z+d1 + d2
(a1 a2 + b2 c1 )z + (a2 b1 + b2 d1 )
. Now we have to check that ad − bc ̸= 0:
(c2 a1 + d2 c1 )z + (c2 b1 + d1 d2 )
we have (a1 a2 + b2 c1 )(c2 b1 + d1 d2 ) − (a2 b1 + b2 d1 )(c2 a1 + d2 c1 ) = (a1 d1 −
b1 c1 )(a2 d2 − b2 c2 ) ̸= 0.
̸ ∞, − dc11 . We have to manually check the special cases,
(This works for z =
which is simply yet more tedious algebra)
1z+0
1. The identity function is 1(z) = 0+1 which satisfies ad − bc ̸= 0.

2. We have shown above that f −1 (z) = dz−b


−cz+a with da − bc =
̸ 0, which are
also Möbius maps

3. Composition of functions is always associative

 
a b az + b
Proposition. The map θ : GL2 (C) → M sending 7→ is a
c d cz + d
surjective group homomorphism.
Proof. Firstly, since the determinant ad − bc of any matrix in GL2 (C) is non-zero,
it does map to a Möbius map. This also shows that θ is surjective.
We have previously calculated that

(a1 a2 + b2 c1 )z + (a2 b1 + b2 d1 )
θ(A2 ) ◦ θ(A1 ) = = θ(A2 A1 )
(c2 a1 + d2 c1 )z + (c2 b1 + d1 d2 )

So it is a homomorphism.
Proposition. Every Möbius map is a composite of maps of the following form:

(i) Dilation/rotation: f (z) = az, a ̸= 0


(ii) Translation: f (z) = z + b
1
(iii) Inversion: f (z) = z

28
10 Möbius group IA Groups (Theorems with proof)

Proof. Let az+b


cz+d ∈ M .
If c = 0, i.e. g(∞) = ∞, then g(z) = ad z + db , i.e.

a a b
z 7→ z 7→ z + .
d d d
1
If c ̸= 0, let g(∞) = z0 , Let h(z) = z−z 0
. Then hg(∞) = ∞ is of the above form.
We have h (w) = w + z0 being of type (iii) followed by (ii). So g = h−1 (hg) is
−1 1

a composition of maps of the three forms listed above.


Alternatively, with sufficient magic, we have
d 1 ad + bc a ad + bc az + b
z 7→ z + 7→ d
7→ − d
7→ − 2 d
= .
c z+ c c2 (z
+ c) c c (z + c ) cz + d

10.2 Fixed points of Möbius maps


Proposition. Any Möbius map with at least 3 fixed points must be the identity.

Proof. Consider f (z) = az+b


cz+d . This has fixed points at those z which satisfy
az+b 2
cz+d = z ⇔ cz + (d − a)z − b = 0. A quadratic has at most two roots, unless
c = b = 0 and d = a, in which the equation just says 0 = 0.
However, if c = b = 0 and d = a, then f is just the identity.
Proposition. Any Möbius map is conjugate to f (z) = νz for some ν ̸= 0 or to
f (z) = z + 1.
Proof. We have the surjective group homomorphism θ : GL2 (C) → M . The
conjugacy classes of GL2 (C) are of types
 
λ 0 λz + 0 λ
7→ g(z) = = z
0 µ 0z + µ µ
 
λ 0 λz + 0
7→ g(z) = = 1z
0 λ 0z + λ
 
λ 1 λz + 1 1
7→ g(z) = =z+
0 λ λ λ

But the last one is not in the


 form z + 1. We knowthat  the last g(z) can
1 λ1 1 1
also be represented by , which is conjugate to (since that’s its
0 1 0 1
1
Jordan-normal form). So z + λ is also conjugate to z + 1.

Proposition. Every non-identity has exactly 1 or 2 fixed points.


Proof. Given f ∈ M and f ̸= id. So ∃h ∈ M such that hf h−1 (z) = νz. Now
f (w) = w ⇔ hf (w) = h(w) ⇔ hf h−1 (h(w)) = h(w). So h(w) is a fixed point
of hf h−1 . Since h is a bijection, f and hf h−1 have the same number of fixed
points.
So f has exactly 2 fixed points if f is conjugate to νz, and exactly 1 fixed
point if f is conjugate to z + 1.

29
10 Möbius group IA Groups (Theorems with proof)

10.3 Permutation properties of Möbius maps


Proposition. Given f, g ∈ M . If ∃z1 , z2 , z3 ∈ C∞ such that f (zi ) = g(zi ), then
f = g. i.e. every Möbius map is uniquely determined by three points.
Proof. As Möbius maps are invertible, write f (zi ) = g(zi ) as g −1 f (zi ) = zi . So
g −1 f has three fixed points. So g −1 f must be the identity. So f = g.
Proposition. The Möbius group M acts sharply three-transitively on C∞ .

Proof. We want to show that we can send any three points to any other three
points. However, it is easier to show that we can send any three points to 0, 1, ∞.
Suppose we want to send z1 → ∞, z2 7→ 0, z3 7→ 1. Then the following works:

(z − z2 )(z3 − z1 )
f (z) =
(z − z1 )(z3 − z2 )

If any term zi is ∞, we simply remove the terms with zi , e.g. if z1 = ∞, we have


f (z) = zz−z 2
3 −z2
.
So given also w1 , w2 , w3 distinct in C∞ and g ∈ M sending w1 7→ ∞, w2 7→
0, w3 7→ 1, then we have g −1 f (zi ) = wi .
The uniqueness of the map follows from the fact that a Möbius map is
uniquely determined by 3 points.
Lemma. The general equation of a circle or straight line in C is

Az z̄ + B̄z + B z̄ + C = 0,

where A, C ∈ R and |B|2 > AC.


Proof. This comes from noting that |z − B| = r for r ∈ R > 0 is a circle;
|z − a| = |z − b| with a ̸= b is a line. The detailed proof can be found in Vectors
and Matrices.
Proposition. Möbius maps send circles/straight lines to circles/straight lines.
Note that it can send circles to straight lines and vice versa.
Alternatively, Möbius maps send circles on the Riemann sphere to circles on
the Riemann sphere.
Proof. We can either calculate it directly using w = az+b dw−b
cz+d ⇔ z = −cw+a and
substituting z into the circle equation, which gives A′ ww̄ + B̄ ′ w + B ′ w̄ + C ′ = 0
with A′ , C ′ ∈ R.
Alternatively, we know that each Möbius map is a composition of translation,
dilation/rotation and inversion. We can check for each of the three types. Clearly
dilation/rotation and translation maps a circle/line to a circle/line. So we simply
do inversion: if w = z −1

Az z̄ + B̄z + B z̄ + C = 0
⇔ Cww̄ + Bw + B̄ w̄ + A = 0

30
10 Möbius group IA Groups (Theorems with proof)

10.4 Cross-ratios
Lemma. For z1 , z2 , z3 , z4 ∈ C∞ all distinct, then

[z1 , z2 , z3 , z4 ] = [z2 , z1 , z4 , z3 ] = [z3 , z4 , z1 , z2 ] = [z4 , z3 , z2 , z1 ]

i.e. if we perform a double transposition on the entries, the cross-ratio is retained.


Proof. By inspection of the formula.

Proposition. If f ∈ M , then [z1 , z2 , z3 , z4 ] = [f (z1 ), f (z2 ), f (z3 ), f (z4 )].


Proof. Use our original definition of the cross ratio (instead of the formula). Let
g be the unique Möbius map such that [z1 , z2 , z3 , z4 ] = g(z4 ) = λ, i.e.
g
z1 7−
→∞
z2 7→ 0
z3 7→ 1
z4 7→ λ

We know that gf −1 sends

f −1 g
f (z1 ) 7−−→ z1 7−
→∞
f −1 g
f (z2 ) 7−−→ z2 7−
→0
f −1 g
f (z3 ) 7−−→ z3 7−
→1
f −1 g
f (z4 ) 7−−→ z4 7−
→λ

So [f (z1 ), f (z2 ), f (z3 ), f (z4 )] = gf −1 f (z4 ) = g(z4 ) = λ.


Corollary. z1 , z2 , z3 , z4 lie on some circle/straight line iff [z1 , z2 , z3 , z4 ] ∈ R.
Proof. Let C be the circle/line through z1 , z2 , z3 . Let g be the unique Möbius
map with g(z1 ) = ∞, g(z2 ) = 0, g(z3 ) = 1. Then g(z4 ) = [z1 , z2 , z3 , z4 ] by
definition.
Since we know that Möbius maps preserve circle/lines, z4 ∈ C ⇔ g(z4 ) is on
the line through ∞, 0, 1, i.e. g(z4 ) ∈ R.

31
11 Projective line (non-examinable) IA Groups (Theorems with proof)

11 Projective line (non-examinable)

32

You might also like