FPM Algebra Notes
FPM Algebra Notes
Algebra
Throughout the semester, all course materials will be available from the course
LEARN page.
In the notes there are some paragraphs inside large squares. These correspond
to proofs and computations that we encourage to try to do on your own
before reading. Similary, we encourage you to try to solve the Check Your
Understanding exercises before reading their solutions. This is a very good
way to cement your understanding of the material.
Contents
0 Revision 5
0.1 Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
0.2 Making counting arguments rigorous . . . . . . . . . . . . . . . . . 6
0.3 Permutations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
0.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
4 Groups Acting 69
4.1 Group Actions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
4.2 Orbits and Stabilizers . . . . . . . . . . . . . . . . . . . . . . . . . 72
4.3 The Orbit-Stabilizer Theorem . . . . . . . . . . . . . . . . . . . . . 75
4.4 First applications of the Orbit-Stabilizer Theorem . . . . . . . . . . 77
4.5 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
3
4 CONTENTS
Revision
We need the following first year material, and will use it extensively throughout.
We will revise other first year material as we require it.
0.1 Functions
We need everything in [L, §19], so you are strongly recommended to review this
material. It is also contained in [J, §2.2]. As notation, in this course we will write
a function f from X to Y as f : X → Y . Further, we will write the composition of
functions f : X → Y and g : Y → Z as g ◦ f , so g ◦ f means ‘do f first, then g ’.
Definition 0.1.1. A function f : X → Y is called
injective if f (x1 ) = f (x2 ) implies that x1 = x2 (equivalently, if x1 ̸= x2 then
f (x1 ) ̸= f (x2 )).
surjective if for every y ∈ Y , there exists x ∈ X such that f (x) = y .
bijective if it is both injective and surjective.
If f is bijective, we denote its inverse function by f −1 .
An easy way to remember this is via the pictures
∃x y
X Y X Y X Y
injective surjective bijective
‘don’t lose information’ ‘hits everything’ ‘the same’
5
6 CHAPTER 0. REVISION
Proof. Recall that for any element t ∈ T , the set f −1 ({t}) is the set of elements
s ∈ S such that f (s) = t; it’s a subset of S. Since every element of S must be
carried, under f , to some element of T , it follows that S is equal to the union of
all the subsets f −1 ({t}): [
S= f −1 ({t}) .
t∈T
just as we suspected.
Check Your Understanding. If S and T are finite sets, then how many maps are
there S → T ?
Corollary 0.2.3. Let S and T be two finite sets, and let f : S → T be a map. If
f is an injection, then |S| ≤ |T |. If f is a surjection, then |S| ≥ |T |. If f is an
bijection, then |S| = |T |.
Proof. This is because f is an injection if and only if, for any t ∈ T , one has
|f −1 ({t})| ≤ 1, and f is an surjection if and only if, for any t ∈ T , one has
|f −1 ({t})| ≥ 1.
Example 0.2.4. Let S and T be two finite sets, and assume that |S| > |T |. Let
f : S → T be any map. Then f isn’t injective; in other words, there exist two
elements s1 , s2 ∈ S that have the same value under f , so that f (s1 ) = f (s2 ). This
is sometimes called the pigeonhole principle.
Definition 0.2.5. For any two sets S and T , let us write Σ(S, T ) for the set of
bijections S → T .
The previous corollary implies that if S and T have different cardinality, then
Σ(S, T ) is empty. On the other hand, if S and T have the same cardinality, then
we can use Theorem 0.2.1 to compute the cardinality Σ(S, T ).
Before we give this result and its precise proof, we should give an informal
argument. When |S| = |T | = n, what does it take to specify a bijection f : S → T ?
For each element s ∈ S, you have to give a value f (s) ∈ T , but you have to do so
in such a way that you never repeat any value. So if you start with some element,
call it s1 , of S, then you may choose any value you like of T ; you have n options.
But then for the next element you take, s2 , you can choose any element of T other
than f (s1 ) as your value for f (s2 ); you have n − 1 options. This process continues,
and in the end you’re left with n(n − 1)(n − 2) · · · 1 = n! options for how to specify
your bijection. Now, how do we formalise this? Well, as a rule, any time we use
the phrase ‘and so on’ or ‘this process continues’, what we’re really doing is an
induction argument.
Theorem 0.2.6. Let S and T be two finite sets, and assume that |S| = |T | = n.
Then there are n! bijections S → T .
set e −1 ({t}) is the set of bijections f : S → T such that f (s) = t. For any t ∈ T ,
if g : S − {s} → T − {t} is a bijection, then we can define a bijection G : S → T
by the rule (
g (x) if x ̸= s ;
G (x) =
t if x = s .
This defines a bijection Σ(S − {s}, T − {t}) → e −1 ({t}). By our induction hy-
pothesis, |Σ(S − {s}, T − {t})| = (k − 1)!, so |e −1 ({t})| = (k − 1)! Now Theorem
0.2.1 implies that
X X
|Σ(S, T )| = |e −1 ({t})| = (k − 1)! = k(k − 1)! = k! ,
t∈T t∈T
as claimed.
0.3 Permutations
A permutation is a bijection from a set to itself. In this section, we’re going to
compute the number of permutations of a set X , we’re going to show how to
compose permutations, and we’re going to show that permutations have inverses.
This material is all revision from Proofs and Problem Solving.
Check Your Understanding. Hey, wait a minute. Why is it that if you compose
two permutations, you get another permutation?
Example 0.3.2. Every set X has a special permutation, called the identity e : X →
X . This sends s 7→ s. Please note thet for any other permutation σ : X → X , we
have
es = s and se = s .
Example 0.3.3. Here’s an element σ ∈ S3 :
1 7→ 3 , 2 7→ 2 , 3 7→ 1 .
1 7→ 2 , 2 7→ 3 , 3 7→ 1 .
0.3. PERMUTATIONS 9
1 7→ 2 , 2 7→ 1 , 3 7→ 3 .
Here’s τ σ:
1 7→ 1 , 2 7→ 3 , 3 7→ 2 .
Please note that στ ̸= τ σ. This product is not commutative!
Example 0.3.4. To try to make things a little easier for ourselves, we can use the
following ‘array’ notation for permutations: the ‘array’ notation where an element
σ ∈ Sn is specified by the 2-row array
1 2 ... n
.
σ(1) σ(2) ... σ(n)
Thus the elements of S3 from the previous example can be written
1 2 3 1 2 3 1 2 3 1 2 3
σ= , τ= , στ = , τσ = .
3 2 1 2 3 1 2 1 3 1 3 2
Example 0.3.5. Here’s an element
1 2 3 4 5 6
σ= ∈ S6
3 5 2 4 6 1
Let’s take powers of σ.
2 1 2 3 4 5 6
σ = σσ =
2 6 5 4 1 3
1 2 3 4 5 6
σ 3 = σσσ =
5 1 6 4 3 2
1 2 3 4 5 6
σ4 =
6 3 1 4 2 5
Now σ 5 turns out to be the identity permutation e! In this case, we say that σ has
order 5.
Definition 0.3.6. More generally, if σ : X → X is a permutation, then the order
of σ is the smallest natural number k > 0 such that σ k = e.
Proposition 0.3.7. Every element of Sn has a finite order.
Proof. Let σ ∈ Sn be a permutation. Define a function
N = {1, 2, ... } → Sn
that carries i to σ i . Since N is infinite but Sn is finite, this map cannot be injective.
Hence there exist two natural numbers 0 < i < j such that σ i = σ j . So of course
j − i > 0, and for any x ∈ {1, ... , n}, note that
Since σ i is a bijection, it follows that x = σ j−i (x). But since this is true for every
x ∈ {1, ... , n}, we deduce that σ j−i = e. Now since there exists a positive natural
number k such that σ k = e, there must also exist a smallest such positive natural
number.
(a1 a2 · · · ak )
of k distinct elements of {1, ... , n}. The corresponding element of Sn is the unique
permutation σ of {1, ... , n} such that
ai+1 if x = ai for 1 ≤ i ≤ k − 1 ;
σ(x) = a1 if x = ak ;
x otherwise.
We just write
σ = (a1 a2 · · · ak ) .
Let’s illustrate with examples.
Example 0.3.8. When we write the 3-cycle (1 3 4) ∈ S5 , we mean the permutation
1 2 3 4 5
3 2 4 1 5
This cycle is equal to (4 1 3) and (3 4 1); these are three ways of writing the same
thing.
Similarly, when we write the transposition (2 5) ∈ S5 , we mean the permutation
1 2 3 4 5
1 5 3 4 2
in cycle notation. That is, you want to write this permutation as a product of
cycles. What you do is you follow what’s called the orbits of the elements. (We’ll
0.3. PERMUTATIONS 11
talk more about orbits later.) So σ carries 1 to 3, and it carries 3 to 4, and it carries
4 to 1; so σ involves the cycle (1 3 4). But that’s not all σ does: it also carries 2
to 5 and 5 back to 2; thus σ also involves a 2-cycle (2 5) – aka a transposition.
Since that accounts for all the elements, our σ is written as a product
σ = (1 3 4)(2 5) = (2 5)(1 3 4) .
στ = τ σ .
Why?
Example 0.3.10. You cannot expect cycles to commute unless they are disjoint. To
see this, let’s look at the transpositions (1 2) and (2 3) in S3 . One one hand we
have
(1 2)(2 3) = (1 2 3) ,
but on the other, we have
(2 3)(1 2) = (1 3 2) .
It turns out that any element of Sn can be written as a product of disjoint
cycles. This representation is not quite unique, but it’s close: the only obstacles to
unicity are that you can write the products in a different order, or you can cyclically
permute the entries of any cycle.
For counting purposes, it’s often convenient to require that the first entry of
any cycle is as small as possible, and that the factors in a product of disjoint cycles
are ordered by that first entry, from smallest to largest. So if you take the product
(4 6 2)(1 3 9) ∈ S9 ,
(1 3 9)(2 4 6) .
(123) , (132) .
Example 0.3.12. Let’s write all 24 of the permutations of {1, 2, 3, 4} in cycle nota-
tion.
Example 0.3.13. Now let’s write all 120 of the permutations of {1, 2, 3, 4, 5} in
cycle notation.
Oof . . . actually . . . let’s not. Let’s see if we can account for all of them by
thinking through how we’d go about listing them all if we had to, though.
For any triple of numbers a < b < c, there are two 3-cycles with these
numbers: (abc) and (acb). Therefore, there are 2 53 = 20 of those.
0.4. EXERCISES 13
For any quadruple of numbers a < b < c < d, to pick out a 4-cycle, you start
with a, and then there is a unique 4-cycle with any ordering of b, c, and d.
So there are 6 54 = 30 of those.
σ = (a1 a2 · · · ak ) ,
0.4 Exercises
Exercise on functions (revision)
Exercise 0.1. We often take functions for granted, but this is dangerous. Are the
following functions?
14 CHAPTER 0. REVISION
1. f : Z → Z sending a 7→ a + 2.
a a
2. f : Q → Q sending b 7→ b + 2.
a
3. f : Q → Z sending b 7→ a + b.
4. f : Z × Z → Z sending (a, b) 7→ a + 2.
5. f : Q × Z → Q sending ( ba , z) 7→ a
bz + 2.
6. f : Q × Q → Q sending ( ba , dc ) 7→ a
b + dc .
7. f : Q × Q → Q sending ( ba , dc ) 7→ a+c
b+d .
Remark: You will see that not all of the examples above are functions. In this
situation, where we have written down something that looks like the definition of a
function but that does not actually define a function, mathematicians often say “f
is not well-defined.”
Chapter 1
You are advised to read [J, §1]. Beware that it is an introductory chapter and so
many things will only fully make sense later! The theme of [J, §1], and this first
chapter, is the slogan ‘symmetries give groups’.
(d) (e) (f )
(g) (h)
15
16 CHAPTER 1. SYMMETRIES AND GROUPS
Note that the definition of a graph only refers to the set of vertices, and to
which pairs of vertices are joined by an edge. Thus it is possible to draw the same
graph in many different ways. For instance, if we take graph (c), rotate it, and
stretch the triangles, we get a picture of a graph with the same essential structure:
(c ′ )
The vertices in the two pictures are in one-to-one correspondence with each other,
and so are the edges that join them.
This sort of situation—where two seemingly different objects have the same
structure—is so prevalent in mathematics that there is a special word we use when-
ever the situation arises: that word is isomorphism from the Greek “isos” (equal)
and “morphe” (form). This word appears in many different contexts (we will see
another one later in the course), but in the specific context of graphs, we can
formulate the notion precisely as follows:
f : V1 → V2
such that f (v1 ) and f (v2 ) are joined by an edge if and only if v1 and v2 are joined
by an edge.
We say that Γ1 and Γ2 are isomorphic if there exists an isomorphism f : Γ1 → Γ2 .
Example 1.1.4. Let’s prove mathematically that the graphs (c) and (c’) are really
isomorphic.
To do this, let’s give their vertices names:
1 4
D
A C
B
2 3
(c) (c ′ )
Thus the two sets of vertices are V = {1, 2, 3, 4} and V ′ = {A, B, C , D}. Let us
define a function f : V → V ′ by the following rules:
Then f describes what happens to the vertices under our operation of rotating and
stretching. Let us check that f does indeed satisfy the axioms of an isomorphism.
1.1. SYMMETRIES OF GRAPHS 17
Example 1.1.8. Graphs (c) and (d) are also not isomorphic. They do have the same
number of vertices, so it is possible to find a bijection between their vertex sets.
However, in graph (d) every pair of vertices is connected by an edge, whereas in
graph (c) the top left and bottom right vertices are not. Hence the bijection cannot
preserve all edges.
We will use the notion of a graph isomorphism to define our first key concept:
a symmetry of a graph.
Definition 1.1.9. A symmetry of a graph is an isomorphism from the graph to
itself, i.e. if the set of vertices is V , then a symmetry is a bijection f : V → V that
preserves edges. That is, a symmetry is a bijection f : V → V such that f (v1 ) and
f (v2 ) are joined by an edge if and only if v1 and v2 are joined by an edge.
Remark 1.1.10. The mathematical definition of symmetry is independent of how
we draw the graph. Thus you must use the mathematical definition, and not just
rely on your intuition.
Remark 1.1.11. A symmetry f : V → V must preserve the valency of a vertex (prove
this!). For example, if v1 has valency three, then f (v1 ) must also have valency three.
Example 1.1.12. Describe all symmetries of the graph
1 2 3 4
Thus there are precisely two symmetries of the graph, namely the identity and
the above reflection.
1.2. GROUPS AND EXAMPLES 19
Logic 1.1.13. When faced with a graph, it is often easy to guess some symmetries.
However, when you want to determine all symmetries, you must argue that there
are no more. This is usually more difficult. For example, in Example 1.1.12 we took
an arbitrary symmetry and argued that it must be one of two things. This shows
there are precisely two symmetries.
Example 1.1.14. Find all symmetries of the square
Check Your Understanding. Do Exercises 0.1 and 1.1. Show that there are 6
symmetries of the triangle graph:
S × S := {(a, b) | a, b ∈ S},
The key point is that we can compose two symmetries f and g to obtain another
symmetry (see the proof of G1 in Theorem 1.2.4 below). Thus we define ∗ by the
rule f ∗ g := f ◦ g (composition of functions), then this gives an operation on S.
Definition 1.2.3 (Definition of a group, [J, §4.3]). We say that a nonempty set G
is group under ∗ if
Groups are one of the basic building blocks of pure mathematics. One of the
main reasons they are so important is that they appear often, and in many different
contexts.
Theorem 1.2.4. The set of symmetries of a graph forms a group (under composi-
tion).
1.2. GROUPS AND EXAMPLES 21
5. For any n ∈ N the sets of vectors Qn , Rn and Cn are groups under addition
of vectors, with identity the zero vector. Here we are forgetting the “extra
structure” that one can also multiply a vector by a scalar. More generally
every vector space V is a group under addition of vectors.
7. The non-zero real numbers R∗ form a group under multiplication (by which
we mean x ∗ y := xy ) with identity 1 and the inverse of x being 1/x. Similarly
C∗ and Q∗ are groups under multiplication. These groups have order |R∗ | =
|Q∗ | = |C∗ | = ∞. Similarly, if p ∈ N is a prime number, then the non-zero
elements Z∗p of Zp form a group under multiplication, with identity 1, inverse
x −1 = x p−2 , and order Z∗p = p − 1. (Check this!)
Abelian groups are named after the Norwegian mathematician Niels Henrik Abel
(1802-1829).
Remark 1.2.7. It is very important to remember that not all groups are abelian. In
fact, you already know an example of a non-abelian group. What is it?
Check Your Understanding. Do Exercises 1.7 and 1.8. Find a graph from Sec-
tion 1.1 whose symmetry group is not abelian.
g g ◦g
1 2
1 2 3 1 3 2
24 CHAPTER 1. SYMMETRIES AND GROUPS
Similarly g ◦ h is equal to
D3 = {e, g , g ◦ g , h, g ◦ h, h ◦ g }.
The identity e ∈ Dn .
where some lines don’t pass through any vertices. Regardless of whether n is
even or odd, there are n reflections.
Let us denote by h the reflection in the line through the bottom left vertex, i.e.
h h
n even n odd
Then we have
Lemma 1.3.3. The set Sn is a group under composition of order |Sn | = n!.
...
where there are n vertices and no edges. Thus Sn is a group by Theorem 1.2.4.
26 CHAPTER 1. SYMMETRIES AND GROUPS
B C
C
B
A A A
E
A
E
D
A B B
F
The octahedron:
1.3. SYMMETRIES GIVE GROUPS 27
D D
A
A G
D
B B
E
H
E
The dodecahedron:
B
E
L B
F K
A J D I
L
H D G
C J
G
H
The icosahedron:
E K
F
B R
Q
E
I
F
I
A D N
L
B K Q J
G
M
C
R S
O
D T
H
P
N
Check Your Understanding. The paragraph above states that the rotational sym-
metries of a Platonic solid form a group. What would you need to do to prove this?
The Platonic solids have interesting symmetries, but it is much harder to prove
anything about them (e.g. how many rotational symmetries there are) by arguing
directly as in §1.1. We will come back to these examples throughout the course, as
we learn more theory.
3. (AB)C = A(BC ).
Check Your Understanding. Which of the groups defined in this section are
abelian?
G × H = {(g , h) | g ∈ G , h ∈ H}
2. Consider the graph (e) in Example 1.1.2. Its symmetry group is S2 × S2 , since
any symmetry is given by specifying a permutation of the up–down arms, and
a permutation of the left–right arms, i.e. a pair of permutations.
1.5 Exercises
Suitable questions from Jordan and Jordan
We will come back to the exercises in Chapter 1 once we have done group
actions.
Chapter 2, Questions 1–2, 4, 6–12 and 14–20 are practise with permutations,
and should mainly be revision from Proofs and Problem Solving [L, §20]. It
is important that you can do them.
Chapter 4, Questions 1–11 (all questions).
1.5. EXERCISES 31
Exercise 1.2. Consider the group D3 , which has as elements the identity e and
g ◦h
h h◦g
g g ◦g
Exercise 1.3. Show that in Dn , the identity e, the n − 1 rotations and the n
reflections are precisely the elements
{e, g , g 2 , ... , g n−1 , h, gh, g 2 h, ... , g n−1 h},
2π
where g is counterclockwise rotation by n , h is a reflection, and
Exercise 1.4. We can view the symmetric group Sn as symmetries of the graph
with n vertices and no edges. For every n ≥ 2, produce a different graph which
also has symmetry group Sn . This is part of the general phenomenon that many
different graphs can have the same symmetry group.
32 CHAPTER 1. SYMMETRIES AND GROUPS
Exercise 1.5. We would like an example of a rotational symmetry of the cube which,
if repeated three times, gives the identity. The identity itself has this property. Are
there any others?
Exercise 1.6. How many elements does GL(2, Z2 ) have? How about GL(n, Zp )
with p a prime?
Exercise 1.7. Are the following operations? For the ones that are operations, do
they make the set into a group?
1. Z with a ∗ b := a + 2.
2. Z with a ∗ b := a + b.
a c a
3. Q, with b ∗ d := b + dc .
a c a+c
4. Q, with b ∗ d := b+d .
x 2 = y 2 = z 2 = e, xy = z, yz = x, zx = y
and eg = g for all g ∈ K . (Note: you need to check for the existence of
inverses and for whether the multiplication is associative: try and organize
your reasoning for the latter, exploiting symmetry rather than checking all 64
possibilities.)
Exercise 1.9. Fix n ∈ N and consider multiplication mod n. Let G be the subset
of {1, 2, ... , n − 1} consisting of all those elements that have a multiplicative in-
verse (under multiplication mod n). Show that G is a group under multiplication.
Describe this group when n = 12.
1.5. EXERCISES 33
Exercise 1.11. Prove that a group (G , ∗) is abelian if and only if for all g , h in G ,
(g ∗ h)2 = g 2 ∗ h2 . (1.1)
Exercises on products
Exercise 1.13. Show that if G and H are groups, then G × H is abelian if and only
if G and H are both abelian.
34 CHAPTER 1. SYMMETRIES AND GROUPS
Chapter 2
2.1 Subgroups
Definition 2.1.1 ([J, §5]). Let G be a group. We say that a nonempty subset H
of G is a subgroup of G if H itself is a group (under the operation from G ). We
write H ≤ G if H is a subgroup of G . If also H ̸= G , we write H < G and say that
H is a proper subgroup.
1. eH = eG
S2. If h, k ∈ H then h ∗ k ∈ H
35
36 CHAPTER 2. SUBGROUPS AND LAGRANGE’S THEOREM
S1.
f H is not empty.
f If h, k ∈ H then h ∗ k −1 ∈ H
S2.
Remark 2.1.5. If G is finite, then there is a slightly easier test for a subgroup. See
Exercise 2.1.
Example 2.1.6. The test for a subgroup can be used to establish all the following
examples.
7. There are also many other interesting subgroups of GL(n, R), for example
Check Your Understanding. Show that SL(n, R), O(n, R) and SO(n, R) are indeed
subgroups of GL(n, R).
ab := a ∗ b = a + b.
This would imply that we are writing ab to mean ‘add a and b’. Since you are so
used to writing ab to mean ‘multiply a and b’, this will cause confusion. Hence,
when we are dealing with groups under addition (like Z or Zn ), it is helpful to keep
the ∗ or + in the notation (see for example Example 2.2.5(3) and Example 2.3.5
later). Nevertheless, when discussing a general group G , we will drop the ∗ as much
as possible.
38 CHAPTER 2. SUBGROUPS AND LAGRANGE’S THEOREM
Definition 2.2.3 (Order of a group). A finite group G is one with only a finite
number of elements. The order of a finite group, written |G |, is the number of
elements in G . (Note that if X is a set, we also often write |X | to be the number
of elements in X .)
g n = e.
2. In D3 , if g is defined as before as
then
3 2 1 3
g g g
1 2 3 1 2 3 1 2
1| ∗ {z
... ∗ 1} = n ̸= 0 = e
n
4. See Exercise 2.10 through 2.15 for many examples of finite order.
Theorem 2.2.6. In a finite group, every element has finite order.
3. In Dn , the subgroup H consisting of the identity and all the rotations, i.e.
Remark 2.2.12. If g ∈ G , then by the line above, o(g ) = |⟨g ⟩|, i.e. the order of an
element equals the order of the subgroup that it generates.
This gives another proof that D3 is not cyclic, since it has no element of order
6. The only possible orders are 1 (the identity), 3 (the two rotations) and 2 (the
three reflections).
We next investigate how cyclic groups behave with respect to our previous
constructions in Section 2.1 (subgroups) and Section 1.4.1 (product groups).
Thus by above any subgroup of any cyclic group is also cyclic. However, products
of cyclic groups are not so well behaved.
Theorem 2.2.16. Let m, n ∈ N, let G = ⟨g ⟩ be a cyclic group of order m and
H = ⟨h⟩ be a cyclic group of order n. Then
G × H is cyclic ⇐⇒ m and n are coprime (i.e. gcd(m, n) = 1).
42 CHAPTER 2. SUBGROUPS AND LAGRANGE’S THEOREM
2.3 Cosets
In the next section we will prove Lagrange’s Theorem, the first main theorem of
this course. This theorem is so important we will state it now, although it is not
proved until Theorem 2.4.2.
Theorem 2.3.1 (Lagrange’s Theorem). Let G be a finite group and let H be a
subgroup of G . Then |H| divides |G |.
Note that this theorem does not follow immediately from the definitions! In
fact, you might want to try to prove it yourself before reading further to get an
appreciation of the new theoretical tools we need to develop for this proof.
For various reasons, often we want to partition this set into smaller pieces, which
we draw as
Whenever we need to partition a set into pieces in this way, the tool to use is an
equivalence relation.
Definition 2.3.2 ([L,§18]). Let X be a set, and R a subset of X × X ; thus R
consists of some ordered pairs (s, t) with s, t ∈ X . If (s, t) ∈ R we write s ∼ t
and say “s is related to t”. We call ∼ a relation on X . A relation ∼ is called an
equivalence relation on X if it has satisfies the following three axioms:
2.3. COSETS 43
The key point from Proofs and Problem Solving [L, 18.1] is that equivalence
relations partition sets. If ∼ is an equivalence relation on a set X , then the set X
is partitioned into pieces called the equivalence classes. In our previous picture
the shaded region is the equivalence class containing x, which by definition is just
all elements that are related to x. It is denoted cl(x). In mathematical symbols,
the shaded region
Check Your Understanding. Remind yourself (see [L], Exercise 18.4) of the rela-
tionship between partitions and equivalence relations. That is, suppose X is parti-
Sn if x ∈ X , then
tioned into the subsets X1 , ... , Xn of X . (Recall that this means that
x is an element of exactly one of the subsets Xj : that is, (i) X = j=1 Xj , and (ii),
if i ̸= j then Xi ∩ Xj = ∅.) Prove that the relation:
defines an equivalence relation on X , and that the equivalence classes are just the
Xj .
AB := {ab | a ∈ A, b ∈ B} , gA := {ga | a ∈ A} ,
Definition 2.3.4 ([J, §10.1]). Let H ≤ G and let g ∈ G . Then a left coset of H
in G is a subset of G of the form gH, for some g ∈ G .
Example 2.3.5. Consider Z4 under addition, and let H = {0, 2}. Recall e = 0. Now
the cosets of H in G are
eH =e ∗H = {e ∗ h | h ∈ H} = {0 + h | h ∈ H} = {0, 2}.
1H =1∗H = {1 ∗ h | h ∈ H} = {1 + h | h ∈ H} = {1, 3}.
2H =2∗H = {2 ∗ h | h ∈ H} = {2 + h | h ∈ H} = {0, 2}.
3H =3∗H = {3 ∗ h | h ∈ H} = {3 + h | h ∈ H} = {1, 3}.
(a) g1 H = g2 H.
(c) g2 ∈ g1 H.
Remark 2.3.10. The properties of right cosets are entirely analogous to those of left
cosets. In particular, Theorem 2.3.8 holds equally well if we replace all left cosets
with right cosets.
A3
A1
A2
A4
Now pick one of the A, then by definition of equivalence class A = cl(g ) for some
g ∈ G . By Lemma 2.4.1, the equivalence class containing g (i.e. A) has precisely |H|
members. Since A was arbitrary, every equivalence class has precisely |H| members,
and so
In the proof of Lagrange’s Theorem, equation (2.1) shows that the number of
|G |
equivalence classes is |H| . This then implies:
2.4. LAGRANGE’S THEOREM 47
|G |
Corollary 2.4.3. |G /H| = |H| .
ap ≡ a mod p.
2.5 Exercises
Suitable questions from Jordan and Jordan
Chapter 5, Questions 1–13 (all questions).
Chapter 6, Questions 1–21 (all questions).
Chapter 10, Questions 1, 4–8, and 10–14.
Exercises on subgroups
Exercise 2.1. Suppose that G is a finite group and let H be a non-empty subset
of G . Show that H is a subgroup of G if and only if h, k ∈ H implies that hk ∈ H.
H = {g ∈ G | gv = λv for some λ ∈ R} .
Exercise 2.6. Find a non-trivial symmetry f of the square (that is, an element of
D4 that is not the identity) that commutes with all the other elements. Which of
the groups Dn have such an element?
Exercise 2.8. Show that Z∗7 (the multiplicative group of nonzero integers mod 7)
is a cyclic group.
Exercises on order
Exercise 2.10. In the dihedral group D6 what are the orders of the various sym-
metries? (Think carefully and if necessary cut a regular hexagon out of paper and
experiment.)
Exercise 2.11. What is the order of the various elements of the symmetric group
S3 ?
Exercise 2.12. Consider Zn under addition. Find the orders of all the elements in
the cases n = 3, 4, 5, 6. What is your guess for the possible orders of elements in
Zn ?
Exercise 2.14. Suppose o(g ) = k. What can you say about the order of g 2 ?
Exercise 2.16. Consider the group D3 . Find the left and right cosets of H in D3
where:
1. H = ⟨g ⟩ = {e, g , g 2 }.
2. H = ⟨h⟩ = {e, h}
Exercise 2.17. Show that if G is abelian and H ≤ G , then the left cosets of H in
G are the same as the right cosets. Find the cosets in the case where G = Z9 and
H = {0, 3, 6}.
Note that in the definition, the product xy on the left is formed using the group
operation in G , whilst the product ϕ(x)ϕ(y ) is formed using the group operation in
H.
Remark 3.1.3. An isomorphism thus matches up the two groups and their group
operations perfectly. In other words, if G and H are isomorphic groups then they
are algebraically indistinguishable. In the world of group theory, isomorphism is the
idea of equality; we view two isomorphic groups as ‘the same’. (Recall that when
two graphs are isomorphic, then from the point of view of graph theory, they are
‘the same’.)
Example 3.1.4. Here are some examples of isomorphisms:
1. Consider R under addition and R∗+ (the group of positive real numbers) under
multiplication. The map exp : R → R∗+ is a group homomorphism since
exp(x + y ) = exp(x) exp(y ). It is bijective, hence is an isomorphism.
51
52 CHAPTER 3. GOING BETWEEN GROUPS
ker ϕ := {g ∈ G | ϕ(g ) = eH }.
∃g h e
G H G H
im ϕ ker ϕ
Note that im ϕ is a subgroup of H and that ker ϕ is a subgroup of G (use the test
for a subgroup).
Furthermore, recall from the second algebra handin the definition of a normal
subgroup:
Definition 3.1.7. A subgroup N ≤ G is normal if the left and right cosets of N
are equal: gN = Ng for all g ∈ G .
If N is a normal subgroup of G we write N ◁ G .
It turns out that kernels of homomorphisms are always normal subgroups.
Proposition 3.1.8. Let ϕ : G → H be a group homomorphism. Then ker ϕ ◁ G .
Proof. See Exercises 3.4 and 3.5
54 CHAPTER 3. GOING BETWEEN GROUPS
Check Your Understanding. Do Exercises 3.2 and 3.3 below, as well as Exercise
9.5 from [J].
ST := {st | s ∈ S, t ∈ T } .
Remark 3.2.2. The logic in the above is that H and K start life as given subgroups
of G . However, we can simply regard them as groups in their own right and take
their abstract product to form H × K . Under the assumption that H ∩ K = {e}, the
conclusion of the first claim is that HK is a set which is bijective to H × K . Under
the further assumption that hk = kh for all h ∈ H, k ∈ K , the second claim is that
actually HK is a subgroup of G , and furthermore HK is the same as (=isomorphic
to) H × K as groups, not just as sets.
3.2. PRODUCTS AND ISOMORPHISMS 55
h
g
Consider the set H consisting of those symmetries of the hexagon which are
also symmetries of the triangle. Since H contains precisely the symmetries of the
triangle, H is a subgroup of D6 which is isomorphic to D3 . Explicitly,
H = {e, g 2 , g 4 , h, g 2 h, g 4 h} ∼
= D3 .
which is thus isomorphic to Z2 . Similarly for the reflection in the vertical line
— it generates a subgroup K which is isomorphic to Z2 . These two reflections
commute, hence HK ∼ = Z2 ×Z2 . Since HK ⊆ G and both have four elements,
G = HK and so G ∼ = Z2 × Z2 .
3.3 Exercises
Suitable questions from Jordan and Jordan
Chapter 9, Questions 1–4 and 9–10.
Exercises on homomorphisms
Exercise 3.1. Show that exp is a group homomorphism between R (under addition)
and R∗ := R − {0} (under multiplication). What is its kernel and what is its image?
Now answer the same question for exp : C → C∗ .
Exercise 3.2. Let p, q be different primes. Show that the only homomorphism
ϕ : Cp → Cq is the trivial one (i.e. ϕ(g ) = e for all g ). (Hint: consider kernels and
images.)
Exercise 3.3. Consider the function det : GL(n, R) → R∗ . Show that it is a group
homomorphism. Identify its kernel and image.
Exercise 3.6. Consider the group C under addition. Show that complex conjugation
ϕ : z → z̄ is an isomorphism C → C.
3.3. EXERCISES 57
Exercise 3.12. Show that no two of the three abelian groups of size 8
Z2 × Z2 × Z2 , Z2 × Z4 , Z8
are isomorphic.
Groups Acting
We first met groups as symmetry groups: thus we were given not only a set of
symmetries, but an object which the symmetries are symmetries of. We formalise
this in the definition of a group action. Group actions will be the major topic in the
second half of the course.
Definition 4.1.1. Let G be a group, and let X be a nonempty set. Then a (left)
action of G on X is a map
G × X → X,
The definition is dense, and a little hard to understand first time round. What
it says is that if we take an element g ∈ G and an element x ∈ X (i.e. an element
(g , x) ∈ G ×X ), we can combine them to produce another element of X (the image
of (g , x) under the map G × X → X ), which we denote g · x. We want to think
of elements of G as “symmetries” of X in some sense, although what this means
in practice can be a bit confusing.
In our blob picture of the set X , if we pick x ∈ X , and g ∈ G , then these
combine to produce some element g · x ∈ X . We think of this as “g takes x to
g · x” and draw it
59
60 CHAPTER 4. GROUPS ACTING
g ·x
g·
1. Let G be the symmetry group of a graph, and let V be the set of vertices of
the graph. We claim that G acts on V by g · x := g (x).
(a) The symmetric group Sn acts on the set {1, 2, ... , n}.
(b) The group Dn acts on the set {1, 2, ... , n}, where we think of the numbers
as labelling the vertices of the n-gon.
2. Let G be the symmetry group of a graph, and let E be the set of edges of the
graph. Then G acts on E , since if e ∈ E connects vertices v1 and v2 , define
g · e := the edge connecting g (v1 ) and g (v2 ). The two axioms follow in a
similar way as above.
3. A group can act on many different sets. For example Dn acts on the set
{1, 2, ... , n} as above. Alternatively, if we label the two faces of the n-gon
T , B (“top” and “bottom”), then Dn also acts on the set X := {T , B} where
g ∈ Dn acts by the identity if it leaves the n-gon the same way up (i.e. g is
a rotation), and by swapping T , B if it turns it over (i.e. g is a reflection).
4. Let G be any group and X any (nonempty) set. Then g · x := x for all g ∈ G
and all x ∈ X defines an action. We call this the trivial action. This action
tells us no information, but it does always exist.
5. G acts on itself (i.e. take X = G ), in many different ways. Three of these are
4.1. GROUP ACTIONS 61
Check Your Understanding. Check that the definition of the right action of a
group on itself above does define an action.
N := {g ∈ G | g · x = x for all x ∈ X } .
Then N is a subgroup of G .
Check Your Understanding. Show that the subgroup N defined above is actually
a normal subgroup of G . (Hint: Exercise 3.4.)
Definition 4.1.4. Suppose that G acts on X . Then the subgroup N defined above
in Proposition 4.1.3 is called the kernel of the action. If N = {e} then we say that
the action is faithful. See [J,§15.2].
Remark 4.1.5. Note in [J] the kernel of an action is denoted by ker ·, but this
notation is quite hard to read.
Example 4.1.6. Here are some examples and non-examples of faithful actions:
1. Let G be the symmetry group of a graph acting on the set V of vertices. Only
the identity fixes every vertex, so the action is faithful. (See Exercise 4.2.)
2. Let G be the rotational symmetry group of a Platonic solid acting on the set
of faces. Again, only the identity fixes every face, so the action is faithful.
3. The action of Dn on the set {T , B} in Example 4.1.2(3) is not faithful, since
for example the rotation g fixes both T and B.
Check Your Understanding. Do Exercises 4.1 and 4.2.
g1 · x
g1 ·
x g2 · g2 · x
4.2. ORBITS AND STABILIZERS 63
In contrast, the stabilizer is all the elements of G that don’t ‘move’ x, i.e. all those
g ∈ G for which the picture looks like this:
g·
In particular, note that the orbit is a subset of X , and the stabilizer is a subset (in
fact a subgroup) of G .
Example 4.2.4. Here are some examples of orbits and stabilizers:
1. Consider the symmetry group G of the graph
v1 v2 v3
Thus
OrbG (v1 ) = {g · v1 | g ∈ G } = {v1 , v3 }
OrbG (v2 ) = {g · v2 | g ∈ G } = {v2 }
OrbG (v3 ) = {g · v3 | g ∈ G } = {v3 , v1 }
and
StabG (v1 ) = {g ∈ G | g · v1 = v1 } = {e}
StabG (v2 ) = {g ∈ G | g · v2 = v2 } = {e, h} = G
StabG (v3 ) = {g ∈ G | g · v3 = v3 } = {e}.
The stabilizer of g ∈ G = X is
3. See Problems 4.3 – 4.7 for more examples of orbits and stabilizers.
x ∼ y ⇐⇒ y = g · x for some g ∈ G
Example 4.2.8. 1. For any given graph, as in Example 4.1.2 part 1, the group
of symmetries acts on the set of vertices. This action may or may not be
transitive (see Problem 4.2).
2. The dihedral group acts transitively on the set of vertices V of the n-gon.
Let v1 , v2 be vertices, then certainly there exists some rotation g t for which
v1 = g t · v2 .
sendx (y ) = gH.
Recall from Definition 2.3.6 that if H ≤ G then we write G /H for the set (which
might not be a group!) of left cosets of H in G .
Theorem 4.3.4. [J, p117] Let G act on X , let x ∈ X , and set H := StabG (x)
Then the map
Then |X | = (|G |)p−1 since the first p−1 elements g0 , ... , gp−2 can be arbitrary
and then there is one and only one gp−1 such that (g0 ... gp−2 )gp−1 = e (by
Lemma 1.4.6).
(g0 ... gk−1 )(gk ... gp−1 ) = e ⇐⇒ (gk ... gp−1 )(g0 ... gk−1 ) = e,
3. Note that the orbits of size one are those of the form {(g , g , ... , g )} where
g p = e.
Certainly there is at least one orbit of size one (namely (e, e, ... , e)), so
X
|X | = 1 + |Oi | (4.1)
all other orbits
This shows that o(g ) ≤ p, in fact it shows that p must be a multiple of o(g ) by [J,
Thm 3(iii), p60]. Since p is prime and o(g ) ̸= 1, it follows that o(g ) = p.
See Problem 4.11 for another theoretical application of the orbit–stabilizer the-
orem.
4.5 Exercises
Suitable questions from Jordan and Jordan
Chapter 1, Questions 4–10 are good practice for orbits and stabilizers.
Chapter 7, Questions 1–4, 8–9 and 13–15 and Chapter 11, Questions 1 and
5–8 mainly involve the orbit–stabilizer theorem.
Exercise 4.2. Show that the symmetry group of any graph always acts faithfully
on the set of vertices. Find a graph (with at least two edges) whose symmetry
group acts transitively on the set of vertices. Find one whose symmetry group does
not act transitively.
Exercise 4.3. Let G act on X and let x ∈ X . Prove that the stabilizer of x is a
subgroup of G .
Exercise 4.4. Consider the group G = Sn acting on X = {1, 2, ... , n}. Let x = n.
How many elements does the stabilizer StabG (x) have? To what group is this
stabilizer isomorphic?
Exercise 4.5. Fix a line L through opposite vertices of a cube. Consider the
subgroup H of the symmetries of the cube generated by g , where g is a rotation
by 1/3 of a turn about L (this is the element in Problem 1.5). Then H acts on the
set of vertices of the cube. Describe the orbits.
Exercise 4.6. Consider the rotational symmetry group G of the cube acting on
the set of vertices of the cube. Describe the stablizer of a particular chosen vertex.
How many elements does it have? Now do the same for the action of G on the set
of faces and the set of edges.
Exercise 4.7. Consider the group G = GL(2, Z2 ). Then G acts on the vectors in Z22
in the usual way by multiplication. Let X be the set of 1-dimensional subspaces of
Z22 . Then G acts on X . Show that |G | = 6, |X | = 3, and that every permutation of
the elements of X is realized by an element of G . (Hence this is another realization
of S3 acting by permutation on {1, 2, 3}.)
Exercise 4.8. Check the orbit-stabilizer theorem for the case of the group Sn acting
on a set of n objects by permuting them.
Exercise 4.10. By Problem 4.9 you should know that there are 24 elements in the
group of rotational symmetries of the octahedron. What are they? (You just have
to describe 24 distinct symmetries, then automatically you have them all.)
g· y
g·
x g ·y
We have the following beautiful formula, which is known by several names: the
orbit-counting theorem, Burnside’s lemma, the Pólya counting theorem. We’ll call
the technique from this theorem Pólya counting in this course.
Theorem 5.1.2 ([J, 11.3]). Let G be a finite group acting on a finite set X . Then
1 X
the number of orbits in X = |Fix(g )| .
|G |
g ∈G
71
72 CHAPTER 5. COUNTING AND CAYLEY
and so
1 X X 1
|Fix(g )| = .
|G | |Orb(x)|
g ∈G x∈X
Finally, since orbits always partition X , this means we can split X into pieces
O1 , ... , On . Thus we can write
X 1 X 1 X 1
= + ... +
|Orb(x)| |Orb(x)| |Orb(x)|
x∈X x∈O1 x∈On
1 1 1 1
= + ... + + ... + + ... +
|O1 | |O1 | |On | |On |
| {z } | {z }
|O1 | |On |
= 1 + ... + 1
= the number of orbits in X ,
as desired.
Example 5.1.3. How many essentially different ways are there of colouring the
vertices of a regular 7-gon with three colours? We will say that two colourings are
the same if they can be made to coincide by an element of the dihedral group D7 .
It is not required that every colouring uses all three colours.
Examples include
The colourings
5.1. PÓLYA COUNTING 73
are equivalent.
To solve this, we consider the action of D7 on the set X of all 37 = 2187 possible
colourings. The problem just asks how many orbits there are, so by Theorem 5.1.2
we must analyse the fixed points.
Example 5.1.4. How many essentially different ways can a tetrahedron be coloured
using n colours, each face being a single colour? (Two colourings are regarded as
the same if one can be obtained from the other by a rotational symmetry.)
There are 12 rotational symmetries of the tetrahedron, and further we know their
explicit form. The group of rotational symmetries G acts on the set X of all possible
n4 colourings. The question asks for the number of orbits, so by Theorem 5.1.2 we
must analyse the fixed points.
The identity fixes everything, so |Fix(e)| = n4 .
Consider a vertex–face rotation (of which there are eight).
B
A
74 CHAPTER 5. COUNTING AND CAYLEY
Then for a colouring to be fixed, the face through which the rotation takes
place (A in the above picture) can have an arbitrary colour, whereas the
remaining faces must all have the same colour. Thus the fixed set contains
n2 elements.
Consider an edge-edge rotation (of which there are three)
B
A
For a colouring to be fixed, the face A in the picture must have the same
colour as the back face that touches the bottom line. Similarly, the face B
must have the same colour as the back face that touches the top line. Thus
the fixed set contains n2 elements.
Hence by Pólya counting the number of orbits is equal to
4
1
12 (n + n2 + ... + n2 + n2 + ... + n2 ) = 1 2 2
12 n (n + 11).
| {z } | {z }
8 3
Remark 5.1.5. See [J, §12] for many other types of these problems.
Remark 5.1.6. Read [J, §12.1] for what ‘essentially the same’ should mean in this
context.
Check Your Understanding. Do Exercises 5.1 and 5.2, as well as the problems in
[J, chapter 12].
This action of G on itself is called the conjugation action. The orbits are called
the conjugacy classes of G . We say that g , g ′ ∈ G are conjugate if there exists
h ∈ G such that g ′ = hgh−1 . That is, two elements are conjugate if they lie in the
same conjugacy class.
Orbits of a group action of G on a set X always partition X (Theorem 4.2.5).
Since here X = G , the orbits of the conjugacy action (=the conjugacy classes)
partition G , so we can draw:
where the shaded piece is the conjugacy class containing g . By definition it contains
all the elements that are in the same orbit as g , i.e. all the elements h · g as h runs
through the elements of G . In mathematical symbols, the bold piece is thus
{hgh−1 | h ∈ G }.
1. Two matrices in GL(n, R) are conjugate if and only if they are similar.
Proof.
76 CHAPTER 5. COUNTING AND CAYLEY
C (G ) := {g ∈ G | gh = hg for all h ∈ G } .
Proof. 1. C (g ) =TStabG (g ), and stabilizers are always subgroups (by Lemma 4.2.2).
2. Note C (G ) = g ∈G C (g ), and an intersection of subgroups is always a subgroup
(Exercise 2.3).
3. This is just the orbit-stabilizer theorem (Corollary 4.3.5).
4. If g is conjugate to e, then there exists h ∈ G such that g = heh−1 , so
g = hh−1 = e. Thus the only element conjugate to e is e itself.
5. {g } is a conjugacy class ⇐⇒ the only element conjugate to g is g itself ⇐⇒
g = hgh−1 for all h ∈ G ⇐⇒ gh = hg for all h ∈ G ⇐⇒ g ∈ C (G ).
We adopt the convention that Conj1 = {e}. Let the conjugacy classes have sizes
c1 = | Conj1 |, ... , cn = | Conjn | (so that c1 = 1).
|G |
1. If g ∈ Conjk , then ck = |C (g )| . In particular, ck divides the order of the group.
2. We have
|G | = c1 + c2 + ... + cn ,
and further each of the cj divides |G |. This is called the class equation of G .
Proof. Part 1 is just the orbit-stabilizer theorem (Corollary 4.3.5) applied to the con-
jugacy action. Part 2 is a trivial consequence of G being partitioned into conjugacy
classes. Each cj divides |G | by part 1.
Example 5.2.8. See Problems 5.11 – 5.12 for examples of the class equation.
The class equation has theoretical consequences. The proof of part one in the
following should remind you of the argument in the proof of Cauchy’s Theorem
(Theorem 4.4.2).
Proof.
78 CHAPTER 5. COUNTING AND CAYLEY
For example, (1534) and (27) are disjoint, but (1534) and (57) are not.
Check Your Understanding. Prove that if σ1 and σ2 are disjoint cycles, then
σ1 σ2 = σ2 σ1 .
Theorem 5.3.3 (See [L, 20.3]). Every permutation can be written as a product of
disjoint cycles.
Proof. We will do this in one example, from which you will probably be able to
write down the general proof yourself (if not, consult [L, §20]). Consider
1 2 3 4 5 6 7 8 9
σ= .
5 7 4 1 3 6 2 9 8
5.3. CONJUGACY IN Sn IS DETERMINED BY CYCLE TYPE 79
Remark 5.3.4. Since disjoint cycles commute, above we could equally write
Proof.
Theorem 5.3.8. Two permutations in Sn are conjugate if and only if they have the
same cycle type (up to ordering).
It turns out that we can easily count the number of elements of a given cycle
type in Sn , and thus determine the size of the conjugacy classes.
Theorem 5.3.9. The number of elements of Sn of cycle type 1m1 , 2m2 , ... , nmn is
n!
.
m1 ! ... mn !1m1 2m2 ... nmn
Example 5.3.10. Here are some examples of how to use this theorem:
5.4. ACTIONS, HOMOMORPHISMS, AND CAYLEY’S THEOREM 81
2. The three possible cycle types in S3 are 13 , and 1, 2, and 3. By the formula,
these contain one, three and two elements respectively.
3. By part 2, we can work out all the conjugacy classes in S3 . There are three
conjugacy classes (since there are three cycle types), and so the conjugacy
classes in S3 are described by
bij(X ) := {bijections X → X }.
This is a group under composition of functions (just amend slightly the proof of
Theorem 1.2.4). Note that if X is finite, then bij(X ) = S|X | , the symmetric group
on |X | letters.
In this section we see that an action of the group G on the set X gives exactly
the same information as a group homomorphism from G to bij(X ).
Lemma 5.4.1 ([J, 7.4]). If G acts on a set X , then for all g ∈ G the map
fg : X → X
defined x 7→ g · x is a bijection.
82 CHAPTER 5. COUNTING AND CAYLEY
Proof.
Theorem 5.4.2. [J, 7.4, 9.3] Let G be a group, and let X be a set. Then
This theorem has a remarkable consequence for the structure of finite groups:
on the set
G =X = {1, 2, 3}. Thus in Cayley’s Theorem, g gets sent to the
1 2 3
element of S3 . Hence
2 3 1
1 2 3
C3 = ⟨g ⟩ ∼
= ≤ S3 .
2 3 1
science-of.html
5.5. APPLICATION TO PLATONIC SOLIDS 85
So for example, the cube duals to the octahedron, which then duals back to the
cube. Hence we can draw a cube inside an octahedron inside a cube. Because of
this,
symm of (outer) cube ⊆ symm of octahedron ⊆ symm of cube
and thus we have equality throughout. This shows that the symmetries of the
cube gives the same group as symmetries of the octahedron (its dual). In exactly
the same way, the symmetries of the dodecahedron give the same group as the
symmetries of the icosahedron.
Thus, to determine the rotational symmetry groups of all the Platonic solids,
we need only determine
1. The rotational symmetry group of the tetrahedron.
2. The rotational symmetry group of the cube.
3. The rotational symmetry group of the dodecahedron.
We consider each in turn:
The cube Consider the group G of rotational symmetries of the cube, acting on
the set X of diagonals of the cube. Note that |X | = 4. This gives us a group
homomorphism ϕ : G → S4 = S|X | , as in Theorem 5.4.2. Now no non-identity
element of the cube fixes all the diagonals (we know all 24 rotational symmetries,
so just check each), hence the action is faithful. Again, by Theorem 5.4.2 this
shows that G ∼= im ϕ ≤ S4 . Since we already know that |G | = 24, this means that
im ϕ is a subset of S4 with 24 elements. But |S4 | = 4! = 24, so im ϕ = S4 . Hence
G∼ = S4 (via ϕ).
86 CHAPTER 5. COUNTING AND CAYLEY
The dodecahedron This is harder, but only because it is a little more difficult to
visualize. Pack the dodecahedron with 5 tetrahedra2 , as in
5.6 Exercises
Suitable questions from Jordan and Jordan
Chapter 11, Questions 2–4 and 9–10 involve fixed sets and Pólya counting.
Chapter 13, Q1–21 are good practice for conjugates, centralizers and centres
and cycle types.
Chapter 12, Questions 1–14 cover symmetries of polyhedra, and also review
Pólya counting and actions as homomorphisms.
Exercise 5.1. Dice have the numbers 1, 2, 3, 4, 5, 6 on the faces of a cube in such
a way that opposite faces add up to 7. How many different dice are there? (Two
dice are the same if they differ by a rotational symmetry of the cube. There are 24
such symmetries.) You can solve this by “pure thought” only because the numbers
involved are small; try to find a group-theoretic answer.
Exercise 5.2. How many ways are there of colouring the vertices of a 6-gon with two
colours, where two colourings are regarded as the same if they differ by an element
of D6 ? (Be careful to think about the different sorts of rotation and reflection. )
Exercise 5.4. Let g ∈ G . Prove directly (i.e. without using general properties of
group actions) that C (g ) ≤ G .
1 1
Exercise 5.5. Let A = ∈ G = GL(2, Z3 ). Show that |C (A)| = 4. What is
0 2
|G |? How many conjugates does A have in G ?
Exercise 5.9. What is the centre of the dihedral group Dn ? (Hint: consider n odd
and even separately.)
Exercise 5.10. Find the conjugacy classes in D4 . (If necessary, cut out a square
from paper, label the corners and play with it.)
Exercise 5.12. Write down the class equation for D4 . (You should have done most
of the work in Problem 5.10.)
Exercise 5.13. Let p be a prime. You know that every group of order p is cyclic,
and thus abelian (by Problem 2.4). You also know that every group of order p 2 is
abelian. Is the same true for p 3 , i.e. is every group of order p 3 abelian?
Exercise 5.14. Let σ ∈ Sn have cycle type l1 , ... , lk . What is the order of σ? What
are the possible orders of elements in S7 ?
Exercise 5.15. Let σ ∈ Sn and let H := ⟨σ⟩ act on {1, 2, ... , n} in the obvious way.
Convince yourself that the cycle type of σ is the list of the sizes of all the orbits of
this action.
Exercise 5.17. S4 has five conjugacy classes. Find them, and the number of
elements in each, and hence write down the class equation for S4 .
Exercise 5.18. Show that (12)(34) together with its conjugates and the identity
form a subgroup of S4 isomorphic to C2 × C2 .
Exercise 5.19. Let D3 act on itself by conjugation. Write down the corresponding
homomorphism D3 → S|D3 | = S6 given by Theorem 5.4.2. Is the action faithful?
Exercise 5.20. Identify the subgroup of S4 that arises in Cayley’s theorem applied
to C2 × C2 . Do the same for C4 .
Exercise 5.21. Let σ ∈ Sn . Another notation for writing σ is the following: write
the numbers 1, ... , n in two rows, one above the other. Then for every number i in
the top row, draw a line between i (in the top row) and σ(i) on the bottom row,
for example
1 2 3 4 5
1 2 3 4 5