MATH225 Abstract Algebra Notes Part I
MATH225 Abstract Algebra Notes Part I
Abstract Algebra
Course Notes∗
Jan E. Grabowski
Department of Mathematics and Statistics
Lancaster University
[email protected]
http://www.maths.lancs.ac.uk/~grabowsj
∗
These are course notes, rather than lecture notes. That is, they contain material you
are expected to study that might not be explained in detail in lectures, as well as material
for further reading that is not examinable.
1
From your previous mathematical studies, you will be aware of examples
of different types of symmetries. As well as geometric symmetries (rota-
tions, reflections, translations), you encountered permutations, which can
be thought of as symmetries of finite sets of objects. You have also studied
linear transformations, as structure-preserving functions of vector spaces -
these too are a type of symmetry.
The notion of a group is designed precisely to capture the common ele-
ments of all of these manifestations of symmetries. It provides an abstract
model, where we can free ourselves from thinking about unimportant inform-
ation that is specific to one particular example, and where we can prove very
general results that we can apply to lots of examples.
In this module, we will look at lots of different examples of groups.
We will begin to study abstract groups, looking at their internal structure
(subgroups) and how groups can be related by structure-preserving functions
(homomorphisms). We will prove many basic results about groups as well as
several more major theorems, which we will apply to our various examples.
You have also encountered several different generalizations of “number
systems” - the integers, the integers modulo n, polynomials and fields. Just
as groups are abstract models of symmetries, so rings are abstract models
for number systems. They have an addition and a multiplication, satisfying
most (but not all) of the familiar properties of the examples we have just
mentioned.
Ring theory also does the same job: stripping away unnecessary com-
plications, making it easier to prove general results that we can apply to
many examples. Again we will study the internal structure of rings, their
structure-preserving maps and some basic results and fundamental theor-
ems.
The examples and results studied here are used and developed further
throughout later pure mathematics modules in algebra, analysis, combinat-
orics and geometry.
2
Syllabus
This module will cover the following topics:
3
Aims and Outcomes
The aims of this module are to
• Give the definitions of a group, a ring and a field (including how these
are related) and give several examples of each of them;
• Prove basic results in group theory and ring theory and identify, recall
and apply significant theorems from those theories, including Lag-
range’s Theorem and the First Isomorphism Theorems.
4
Contents
1 Introduction to Abstract Algebra 7
1.1 Binary operations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2 Algebraic structures 19
Index 193
5
References
[AF] Marlow Anderson and Todd Feil. A first course in abstract algebra.
CRC Press, Boca Raton, FL, third edition, 2015. ISBN 978-1-4822-
4552-3.
[Alu] Paolo Aluffi. Algebra: chapter 0, volume 104 of Graduate Studies in
Mathematics. American Mathematical Society, Providence, RI, 2009.
ISBN 978-0-8218-4781-7.
[Cam] Peter J. Cameron. Sets, logic and categories. Springer Undergraduate
Mathematics Series. Springer-Verlag London, Ltd., London, 1999.
ISBN 1-85233-056-2.
[Der] John Derbyshire. Unknown Quantity: A Real and Imaginary History
of Algebra. Atlantic Books, 2008. ISBN 978-1-84354-570-5.
[dS] Marcus du Sautoy. Finding Moonshine: A Mathematician’s Journey
through Symmetry. Fourth Estate, 2008. ISBN 978-0007214617.
[Neu] Peter M. Neumann. A Hundred Years of Finite Group Theory. The
Mathematical Gazette, 80(487):106–118, March 1996.
[NST] Peter M. Neumann, Gabrielle A. Stoy, and Edward C. Thompson.
Groups and geometry. Oxford Science Publications. Oxford Univer-
sity Press, New York, 1994. ISBN 0-19-853452-3; 0-19-853451-5.
[Ron] Mark Ronan. Symmetry and the monster. Oxford University Press,
Oxford, 2006. ISBN 978-0-19-280722-9; 0-19-280722-6.
[ST] Geoff Smith and Olga Tabachnikova. Topics in group theory. Springer
Undergraduate Mathematics Series. Springer-Verlag London Ltd.,
London, 2000. ISBN 1-85233-235-2.
Above, we list the books referred to in these notes and some others
covering the topics herein. There are many books on groups and rings
available in the library or to buy. The books above will provide an alternative
perspective to the recommended texts and you will find others not on this
list that are equally suitable. However, these notes contain all of the material
that you need for this course.
Acknowledgements
This module and its notes owe a great deal to all those who have lectured it
in previous forms, most recently as Groups and Rings. Some parts closely
follow previous iterations of the course notes or are extracted from them;
we thank colleagues for their permission to use these as well as for many
conversations that fed into the development of this module.
6
Lecture plan
1. Overview: what is abstract algebra? Binary operations (pp. 2-17)
abstract
adjective
1. existing in thought or as an idea but not having a physical
or concrete existence.
“abstract concepts such as love or beauty”
synonyms: theoretical, conceptual, notional, intellectual, meta-
physical, philosophical, academic;
antonyms: actual, concrete
• dealing with ideas rather than events.
• not based on a particular instance; theoretical.
“we have been discussing the problem in a very abstract
manner”
• (of a noun) denoting an idea, quality, or state rather
than a concrete object.
2. . . .
verb
1. consider something theoretically or separately from (some-
thing else).
2. extract or remove (something).
“applications to abstract more water from streams”
3. make a written summary of (an article or book).
synonyms: summarize, write a summary of, precis, abridge,
condense, compress, shorten, cut down, abbreviate, synop-
size;
algebra
noun
the part of mathematics in which letters and other general sym-
bols are used to represent numbers and quantities in formulae
and equations.
“courses in algebra, geometry, and Newtonian physics”
• a system of algebra based on given axioms.
Origin
late Middle English: from Italian, Spanish, and medieval Latin,
from Arabic al-jabr ‘the reunion of broken parts’, ‘bone-setting’,
from jabara ‘reunite, restore’. The original sense, ‘the surgical
treatment of fractures’, probably came via Spanish, in which
it survives; the mathematical sense comes from the title of a
book, ’ilm al-jabr wa’l-muqābala ‘the science of restoring what
is missing and equating like with like’, by the mathematician
al-Kwārizmı̄ (see algorithm).
¯
Oxford English Dictionary
7
The landscape
Before we embark on our journey into abstract algebra, let us try to get
a sense of the wider landscape. The following diagram over-simplifies, of
course, but it is a starting point.
Pure
Algebra Geometry
Arithmetic Analysis
Here we see that the four major topics in pure mathematics all have overlap
with each other. Since we want to understand the place of algebra, let us
zoom in a little further.
8
Pure
Algebra
Geometry
Groups
Rings
3
2
1
Arithmetic
Analysis
Now we also include two major areas within algebra: groups and rings.
According to this point of view, much of abstract algebra is concerned with
groups, and rings are a particular type of group. This last statement is
technically correct but rather misleading: we will see why, as the module
develops.
Before we move on, we highlight some of the topics where algebra has
particularly important intersection with the other major areas; these are
(2) Cryptography
9
doing mathematics, a place where the sorts of questions we ask are different,
and unsurprisingly the answers are different too. It will help you to recognise
that the newness is on these two levels: it is not just the objects in question
that are new, but the way we approach we are doing is too, and that can
cause you to feel disoriented.
The formal definitions will come soon enough, but first let’s take a bit
of time to try to appreciate what it is that abstract algebra was invented to
study. We will try to make as many links to mathematics you already know
as we can, showing how to build on that, but we will also at various points
look forwards to what the theory we begin to construct in this module can
eventually do, though details will have to be left to future modules.
First and foremost, abstract1 algebra is about mathematical objects,
defined by axioms, and their relationships. The principal examples are
groups, rings, fields and vector spaces, though there are others.
Fundamentally, groups come from considering symmetries. Think of a
square drawn in the plane. What are its symmetries? We would natur-
ally think of the various rotations and reflections, so there is some set of
symmetries. This set includes the (seemingly!) uninteresting symmetry of
“doing nothing”, the identity symmetry. It also, very importantly, has the
property that if we take two symmetries and do one after the other, we get
another symmetry: this puts an operation (of composition, in this case) on
the set. We can also undo each symmetry, and these properties (plus one
other slightly less obvious one) give us the properties of what we now call
groups. Many of the other examples of groups we will meet soon arise in
a similar way, as symmetries of something, but we will make an abstract
definition of a group as a set with an operation satisfying some properties.
Rings are a little more complicated but equally have their origins in
natural and basic mathematical questions. Rings are designed to be a gen-
eralisation of the integers and give us new and different types of “number
systems”. So rings are sets with two operations, satisfying axioms corres-
ponding to natural properties of addition and multiplication. If we just look
at the addition and forget the multiplication, we have a group. So from this
perspective, rings are special sorts of groups.
Fields are special sorts of rings: they are rings—so have addition and
multiplication—with the extra property that they also have division. The
prototype example here is the rationals, where every (non-zero) rational a
has a multiplicative inverse 1/a. Just as subtraction is the inverse operation
for addition, rather than a completely new operation, so fields are sets with
two operations, addition and multiplication, but where the multiplication
1
You might be wondering when you missed out on the “Concrete Algebra” module.
Well, you didn’t, and this is one of those occasions where the adjective—abstract, in
this case—is being used for emphasis, rather than as a qualifier. That is, in modern
mathematics, essentially all algebra is abstract, but we say “abstract algebra” to remind
us that we’re approaching algebra in a particular way.
10
has the extra special property of admitting multiplicative inverses. The reals
and the complex numbers are fields but both of these have extra structure
that is better understood in the context of analysis; from an algebraic point
of view, we consider them as (slightly special) fields.
Vector spaces are also defined in a similar way: they generalise R2 , R3
and so on, and are also sets with two operations. One of these is addition,
so that like rings and fields, vector spaces have a group structure, but what
makes them different is their second operation. Vector spaces do not have
a multiplication like a ring—we can’t multiply a vector by a vector like we
would integers or real numbers—but instead they have scalar multiplication,
where an element of the base field and a vector can be multiplied.
So this is one way to understand the different parts of algebra, namely
group theory, ring theory, linear algebra and so on. But there is another
way to view things, namely by the types of question one asks.
Abstract algebra has several major themes in these terms:
11
Abstract algebra also has its own style among the various areas of pure
mathematics. Here are some words commonly associated with algebra:
By the end of this module, hopefully you will appreciate why some of
these words are a good description of algebra, but also hopefully you’ll
disagree with one of them!
The final remarks to be made before we begin on formal definitions are
about how algebra works. The first main theme is “proceeding from the
specific to the general”. That is, take some (relatively) concrete mathemat-
ical situation, such as the symmetries of a square or the arithmetic of the
integers, and try to abstract 2 the situation to give a definition that captures
just the essentials. Then we can study objects that satisfy this definition, in
general, and see what we can learn. Often it is helpful to strip away some of
the clutter of a specific case and by seeing things abstractly we get a better
sense of the essential nature of the question.
The other major theme is “new from old”, taking one or two objects and
producing more related ones. This might be “sub” objects, smaller ones
found inside a given object, or ways to combine two objects to give a larger
one.
In trying to make sense of what will be introduced over the next few
weeks, come back to this introduction and try to see how what we are doing
fits with this philosophy.
2
In the sense of the definition of “abstract” as a verb given at the start of this section.
12
1.1 Binary operations
The starting point for defining any of the algebraic objects referred to above
is the notion of a binary operation. “Binary” means “two”, and “operation”
indicates that we take (two) elements and operate on them, producing an-
other. Formally, however, binary operations are just functions, so we remind
ourselves of the definition of a function first. Before we do, recall that given
sets X and Y , we may form the Cartesian product X × Y of ordered pairs
{(x, y) | x ∈ X, y ∈ Y }.
We write f (x) for the unique element of Y such that (x, f (x)) ∈ Γ and
call f (x) the image of x under f . The set X is called the domain of f , the
set Y is called the codomain of f .
Then the definition of a binary operation is simply the following.
13
Examples 1.3.
You can surely think of many more binary operations, probably mostly
using addition or multiplication. However binary operations can take any
set S as input, not necessarily a set of “numbers”. For example, composition
of functions can give rise to a binary operation: given a non-empty set X
and function f : X → X and g : X → X, their composition g ◦ f : X → X
is again a function. This gives us a binary operation on F(X, X), the set of
functions with domain and codomain X. (Note that if Y �= X, composition
is not a binary operation on F(X, Y ), since we cannot compose two functions
f : X → Y and g : X → Y .)
As a particular example of this, think of the set of rotations and reflec-
tions of a square. More precisely, let S be the unit square centered at the
origin in R2 (thought of as a subset of R2 ) and let Sym(S) be the set of
invertible linear transformations of R2 preserving S (as a set). Composition
is also a binary operation on Sym(S).
In the following diagram, we illustrate the unit square (top left), the
square after applying a rotation clockwise by π about its centre (top right),
after applying a reflection in the line y = x (bottom left) and after first
applying the rotation and then the reflection:
14
a b c d
d c b a
c b a d
a c
d b
Plain binary operations are not usually very interesting. They become more
so when they have particular properties, such as those that addition or mul-
tiplication of integers satisfy. The most important of these are the following;
from this point onwards, we will use the notation a ∗ b rather than ∗(a, b).
Definition 1.4. Let S be a non-empty set and ∗ : S × S → S be a binary
operation on S. We say ∗ is associative if
(a ∗ b) ∗ c = a ∗ (b ∗ c)
for all a, b, c ∈ S.
This property should be familiar, even if the name is not. Many of
the binary operations you would think of, including all those above, are
associative. There are interesting non-associative binary operations3 but we
won’t encounter them in this course.
There are several different ways to think about associativity, from the
very formal (there are two natural ways to build ternary operations on S
3
The operation [ , ] : Mn (C) × Mn (C) → Mn (C), [A, B] = AB − BA is a very important
example.
15
from ∗ and these two functions should be equal) to the more conceptual
(“we don’t need brackets”).
It is very easy to overlook the extremely important last few words of this
definition: the equation (a ∗ b) ∗ c = a ∗ (b ∗ c) is required to hold for all
a, b, c ∈ S. The operation ∗ is not associative if this equation fails for even a
single triple a, b, c. Hence, to show that an operation is not associative, we
simply need to find a, b, c ∈ S such that the equation fails; we will reiterate
this several times below, but it is crucial that we do so with an explicit
choice of a, b and c.
Class Exercise 1.5. Which of these binary operations is not associative?
(a) S = Z and a ∗ b = a + b;
(b) S = Z and a ∗ b = ab;
(c) S = Z and a ∗ b = a − b.
Find integers a, b, c that show this.
Answer. The answer is (c), since 1 − (2 − 3) = 1 − (−1) = 2 but (1 − 2) − 3 =
−1 − 3 = −4.
16
Two important things to notice here are firstly that again we insist that
this equation holds for all a ∈ S, and secondly that both e ∗ a and a ∗ e are
required to be equal to a. There are examples where we might have e ∗ a = a
for all a but where a ∗ e �= a, though it is fair to say that you have to go
looking quite hard in order to find them.
For some binary operations, knowing a “left-handed” statement implies
a “right-handed” one, and vice versa. This can happen if the operation is
rather special4 , in that the order we combine the elements does not matter.
Definition 1.8. Let S be a non-empty set and let ∗ : S × S → S be a binary
operation on S. We say that ∗ is commutative if
a∗b=b∗a
for all a, b ∈ S.
For example, addition and multiplication of integers are both commut-
ative operations but composition of functions is in general not.
Lastly, we need one extra property, but this is of a slightly different
nature. It is a property that elements of S might (or might not) have with
respect to a binary operation on S if that operation has an identity.
Definition 1.9. Let S be a non-empty set and ∗ : S × S → S a binary
operation on S. Assume that there exists e ∈ S such that e is an identity
for ∗. We say that an element a ∈ S has an inverse with respect to ∗ if there
exists b ∈ S such that
a ∗ b = e = b ∗ a.
The prototype example here is the integers: given a ∈ Z, there exists
−a ∈ Z such that a + (−a) = 0 = (−a) + a. In fact, every element of Z has
an inverse with respect to addition, making the integers quite special (hold
this thought for later!).
However this need not always be the case: for example, with respect to
multiplication on the integers, only 1 and −1 have inverses (1 · 1 = 1 = 1 · 1
and (−1)·(−1) = 1 = (−1)·(−1), and in fact these elements are self-inverse).
The number 2 has no inverse in Z because 12 ∈ / Z.
As another (non-)example, most functions are not invertible5 . However,
invertible functions have inverses(!), so by definition the elements of Sym(S)
have inverses.
4
Here your intuition from your previous experience will lead you badly astray: you
might think what follows is typical behaviour, but this is very far from true. And this is
just as well, or else the lecturer’s research area would be even less interesting than most
people already think it is.
5
Recall from MATH112 Discrete Mathematics that a function f : X → X is invertible
if there exists g : X → X such that g ◦ f = IX = f ◦ g. Here IX is the identity function on
X, which is the identity for composition. For composition to be a binary operation, we
need f : X → X rather than f : X → Y , which is why this statement is a slightly special
case of the general definition we gave in MATH112.
17
You might possibly be wondering about whether a binary operation can
have more than one identity, or if an element can have more than one inverse.
If you are, then congratulations, you are well on the way to becoming an
algebraist. If you are not, then now is a good time to start adopting this
type of healthy questioning when you meet a new definition. Algebraists are
by training sceptical when they see a definition: is the concept well-defined?
is there uniqueness? and so on.
You already know the answer to the second question, at least in the case
of functions (we proved that the inverse of a function is unique if it exists
in MATH112 Discrete Mathematics). It should not therefore be a surprise
that we can prove the following proposition.
(ii) We have
b1 = b1 ∗ e
= b1 ∗ (a ∗ b2 )
= (b1 ∗ a) ∗ b2
= e ∗ b2
= b2 .
That is, subject to the assumptions given, both identity elements and
inverses are unique if they exist. The assumptions are necessary, as shown
by the following exercise.
Exercise 1.11. Let S = {a, b, e} and define a binary operation ∗ by the
following table (known as a Cayley table):
∗ e a b
e e a b
a a e e
b b e e
18
Prove the following:
(a) ∗ has an identity element;
(b) ∗ is not associative;
(c) there exists an element of S having two different inverses.
Notice that we can prove the proposition in complete generality, for
binary operations on arbitrary sets. We now never need to prove either
of these again for particular examples: why would we bother? As we said
before, this idea—proving results in the most general setting possible—is
the “abstract” in Abstract Algebra.
2 Algebraic structures
An algebraic structure is just a set with some operations. We might ask that
the operations have some of the properties discussed above, or that different
operations interact in a particular way. In principle there are many different
types of algebraic structure but in practice some have much richer theories
than others and more (interesting) examples “in nature”. In this section,
we will introduce the main ones we will study in this course.
An algebraic structure consisting of a set S and n operations will be
denoted by (S, ∗1 , ∗2 , . . . , ∗n ). Sometimes we will use different letters for
the set S, to help us remember which type of algebraic structure we are
thinking about, and we might use established or more suggestive notation
for the operations, such as + or ×.
Definition 2.1. A group is an algebraic structure (G, ∗) with ∗ an associ-
ative binary operation, having an identity and such that every element of G
has an inverse with respect to ∗.
a × (b + c) = (a × b) + (a × c) and (a + b) × c = (a × c) + (b × c)
for all a, b, c ∈ R. The identity element for the group (R, +) will be denoted
0R .
The two equations given here are called the first and second (or left and
right) distributive laws. We could be more general and say that × distributes
over + from the left if a × (b + c) = (a × b) + (a × c), and correspondingly
for the right-handed version. Notice that these are not symmetric in + and
×.
19
Definition 2.4. A field is an algebraic structure (F, +, ×) such that (F, +, ×)
is a ring, × is commutative and has an identity 1F �= 0F and every element
of F \ {0F } has an inverse with respect to ×.
Number systems
• Q, R and C are fields.
• Z is a ring (but not a field); in many ways, Z is the prototype ring and
many ring-theoretic questions are inspired by the structure of Z.
• For each n, the set of integers modulo n, Zn (with addition and mul-
tiplication modulo n) is a ring.
• Since Q, R and C are fields, as we said in the remarks after the defin-
ition of a field, we have associated groups Q \ {0}, R \ {0} and C \ {0}
with respect to multiplication.
More generally, the set of invertible elements in a ring R forms a group
under multiplication (called the group of units, R∗ ). For a field F , we
have F ∗ = F \ {0F }, because every non-zero element is invertible.
Symmetry groups
• Natural examples of symmetry groups are the set of rotations and
reflections of a square, an n-gon, a circle or other 2-dimensional figures,
as well as rotations and reflections of a cube, tetrahedron, etc.
20
• The “symmetries of a set” S are given by the set of bijections
def
Bij(S) = {f : S → S | f is a bijection}. We saw in MATH112 Discrete
Mathematics that this set satisfies all the conditions to be a group with
respect to composition of functions.
We also went on to discuss permutations, which are re-orderings of a
set of n objects: a little thought and one realises that permutations
are exactly bijections from the set of objects to itself. These groups
are called the symmetric groups.
They (and the symmetry groups of geometric objects) are usually not
Abelian, in contrast to all the previous examples.
Matrices
• We can add m×n matrices and multiply n×n matrices, we know what
the zero and identity matrices are and we know that sometimes we can
invert matrices. So it should not be a surprise that sets of matrices
can have either group or ring structures, though they are almost never
fields.
More precisely, Mn (Z), Mn (Q), Mn (R) and Mn (C) are all rings; in fact
Mn (R) is one for any ring R. We call these matrix rings. (In linear
algebra, we usually work over a field, because we want to be able to
divide by scalars as well as multiply by them.)
• Inside any matrix ring over a field F , the subset of invertible matrices
(those with non-zero determinant) forms a group, the group of units,
as defined above. We have a special name and notation for the group
of units of matrix rings: we call them general linear groups and write
GLn (F ). The group operation is matrix multiplication.
Sets of functions
• The set of functions from R to R, denoted F(R, R), is a ring with
respect to pointwise addition and multiplication:
This list might give the (erroneous) impression that there are lot more
rings than groups. This is not the case: since every ring is a group and not
21
every group is a ring, there are definitely more groups than rings. It just so
happens that the examples you are familiar with are mostly rings or fields.
One important observation about the above list is that lots of these ex-
amples, notably the symmetric groups, the matrix rings and general linear
groups, are naturally symmetries of something. The description of sym-
metry groups we gave says this explicitly. One of the main results in Lin-
ear Algebra is that matrices are the same thing as linear transformations
of vector spaces, and hence GLn (F ) is the symmetry group of F n , the n-
dimensional vector space over F . This idea leads to a major area of study
in algebra, called representation theory.
Having mentioned vector spaces, let us briefly say where they fit in terms
of the definitions so far. Every vector space V over a field F is an Abelian
group under addition, (V, +). Vector spaces are not rings (or fields): we do
not multiply vectors by vectors.
However vector spaces have more structure than just addition, namely
they have scalar multiplication. Scalar multiplication is a slightly different
type of operation: it is not a binary operation on V . It takes two inputs, a
scalar λ ∈ F and a vector v ∈ V , and produces a new vector λv ∈ V . We
can formalise this as a function · : F × V → V , λ · v = λv, and say that a
vector space is an algebraic structure (V, +, ·) where + and · satisfy some
compatibility conditions, namely those that give linearity.
This concludes our overview of abstract algebra and we will now concen-
trate on groups and rings in more detail, beginning with groups.
22
3 The internal structure of groups
In this chapter, we will look at the internal structure of groups. That is, we
will have a group—either a specific example or just “some group G”—and
we will ask about properties it has or that its elements or its subsets have.
3.1 Permutations
Firstly, however, we will look in more detail at the symmetric groups, as we
will use these as one of our main classes of examples to illustrate the theory
that follows.
One justification for this is that every finite group can be found inside
a symmetric group. This is called Cayley’s theorem and we will state and
prove it later (Theorem 4.28).
Recall that the set of bijections from a set S to itself is a group with
respect to composition of functions, (Bij(S), ◦). Let us fix S = {1, 2, . . . , n}
for some n ∈ N.
23
The first approach is called “two-line notation”. Remember that any
function π : {1, 2, . . . , n} → {1, 2, . . . , n} is completely described
� by a list of�
1 2 3 ··· n
its values, π(i) for i ∈ {1, 2, . . . , n}. We can write this as π(1) π(2) π(3) ··· π(n) .
For example, the bijection π sending 1 to 2 (so π(1) = 2), 2 to 3 and 3
to 1 would be written as ( 12 23 31 ).
This helps us see why |Sn | = n!, since there are n choices for π(1), n − 1
for π(2) (π is a bijection so we cannot have π(2) = π(1)) and so on, down
to only one choice left for π(n). That is, n! choices in total.
In fact, the order of the columns does not matter: ( 12 23 31 ) and ( 31 23 12 )
hold the same information.
We can see that ι = ( 11 22 ··· n
··· n ) is the two-line notation for the identity
permutation, which is the identity bijection and so is the identity element
for the composition of permutations.
We can express the composition of permutations as
� a1 a2 ··· an � � 1 2 ··· n � � 1 2 ··· n �
b1 b2 ··· bn ◦ a1 a2 ··· an = b1 b2 ··· bn .
Notice that the order here is in keeping with our convention for composing
functions, writing g ◦ f for first f and then g. So we do the right-hand most
permutation first, and work right to left.
The example in the above paragraph explaining that most symmetric
groups are not Abelian can be written in two-line notation as the results of
the compositions
( 23 11 32 ) ◦ ( 12 21 33 ) = ( 13 21 32 )
and
( 12 33 21 ) ◦ ( 11 23 32 ) = ( 12 23 31 )
not being equal. � �
Finally, the inverse of a11 a22 ··· ··· n
an is ( a11 a22 ··· an
··· n ). For example, the
inverse of ( 12 23 31 ) is ( 21 32 13 ) = ( 13 21 32 ).
However there is a lot of redundancy in two-line notation, and the rule
for composition can be awkward to apply. We will try to simplify things by
using cycle notation; next, we define what we mean by a cycle.
24
For example, the permutation that in two-line notation is ( 12 23 31 ) is a 3-
cycle, namely (2 3 1). Similarly, the permutation ( 11 24 35 43 52 ) is a 4-cycle,
namely (2 4 3 5); notice that the numbers in a cycle do not have to come in
increasing order and that 1 is fixed so does not appear in this cycle.
Remark 3.5. We note the following properties of cycles:
(c) We can compose two cycles by ‘feeding in’ elements from the right
in turn; for example, if π = (1 2 5) and σ = (2 3 4) then π ◦ σ =
(1 2 5)(2 3 4) maps
1 �→ 1 �→ 2, 2 �→ 3 �→ 3, 3 �→ 4 �→ 4, 4 �→ 2 �→ 5, 5 �→ 5 �→ 1.
We often say this as “1 goes to 1, 1 goes to 2; 2 goes to 3, 3 goes to
3” and so on. That is, we start with 1 and write “(1 ”. Then we see
where the first and then the second permutation send it, and write
the partial permutation “(1 2 ”; next it would be “(1 2 3 ” and so on,
until we reach an element that is sent to 1, at which point we close
the cycle. So π ◦ σ = (1 2 3 4 5).
(e) If we compose an r-cycle with itself r times, the result will be the
identity permutation. So every cycle has the property that some power
of it is the identity; we will formalize this idea later.
25
π ◦ σ = (1 2 3)(4 5 6 7)
= ( 12 23 31 45 56 67 74 88 )
= (4 5 6 7)(1 2 3) = σ ◦ π
Now not all permutations are cycles. The permutation ( 12 23 31 45 56 67 74 88 ) we
just obtained is not a cycle, for example. However, it is a product of two
cycles which moved disjoint sets of points; this suggests how we may extend
the idea of cycles to cover all permutations.
Definition 3.6. Let π = (a1 a2 · · · ar ) and σ = (b1 b2 · · · bs ) be cycles in
Sn . If the sets {a1 , a2 , . . . , ar } and {b1 , b2 , . . . , bs } are disjoint, we say that π
and σ are disjoint cycles. This terminology is extended in the obvious way
to sets of more than two cycles.
For the permutation α = ( 12 23 31 45 56 67 74 88 ) above, it is easy to recover
the cycles (1 2 3) and (4 5 6 7) that it is a product of. Based on the above
remark, we can write it equally well as any of the following:
α = ( 12 23 31 45 56 67 74 88 )
= (1 2 3)(4 5 6 7)
= (4 5 6 7)(1 2 3)
= (1 2 3)(4 5 6 7)(8)
= (2 3 1)(4 5 6 7)(8)
or in indeed various other equivalent ways. The following theorem shows
that the behaviour observed in this example is typical.
Theorem 3.7. Every permutation in Sn can be written as a product of
disjoint cycles; moreover this expression is unique up to
(i) the order in which the cycles occur,
(ii) the different ways of writing each cycle (rotating the cycle), and
(iii) the presence or absence of 1-cycles.
Proof. Let π ∈ Sn , and choose any a1 ∈ {1, 2, . . . , n}. Consider the sequence
a1 , a2 , a3 , . . . defined by ai+1 = π(ai ) for i ∈ N. Since this sequence is
infinite, and {1, 2, . . . , n} is a finite set, not all the ai can be distinct. Let r
be minimal such that ar+1 = as for some 1 ≤ s ≤ r. We claim that s = 1, in
other words, that the sequence is a1 , a2 , . . . , ar , a1 , . . . where a1 , . . . , ar are
distinct.
To see this, suppose s > 1. Then π(ar ) = ar+1 = as = π(as−1 ). As π is
injective this would force ar = as−1 , contrary to the minimality of r. Thus
we must have s = 1, and we have the cycle γ1 = (a1 a2 · · · ar ).
If r = n then π = γ1 as required, so now assume r < n. We choose
b1 ∈ {1, 2, . . . , n} \ {a1 , a2 , . . . , ar } and proceed as above. This gives another
26
cycle γ2 = (b1 b2 · · · bt ). The cycles γ1 and γ2 are disjoint, because if ai = bj
then
b1 = π t+1−j (bj ) = π t+1−j (ai ) = ai+t+1−j ,
contrary to the choice of b1 . Continuing in this way we obtain π as a product
of disjoint cycles γ1 γ2 · · · γk .
It is clear that two expressions which differ only in ways (i), (ii) and (iii)
represent the same permutation. Conversely, two expressions which differ
in any other way represent different permutations.
From this point onwards, we will often not write π ◦ σ for composition of
permutations, but will just write πσ, as this is more consistent with disjoint
cycle notation. We will also often talk of the product of two permutations,
rather than their composition, although they mean the same thing. Note
however that the order we write the permutations is still important in gen-
eral; only disjoint cycles commute with each other (as in (f) above).
Definition 3.8. A permutation which is written as a product of disjoint
cycles is said to be in disjoint cycle notation; if all 1-cycles are included, it
is said to be in full cycle notation.
In disjoint cycle notation, we often leave out the 1-cycles. Then the
identity permutation ι in disjoint cycle notation has no cycles at all; this is
why we give it a notation, ι, of its own.
Example 3.9. Let π = ( 13 28 36 41 57 64 79 82 95 ) and σ = ( 11 25 36 49 57 62 73 88 94 ); then
the expressions for π, σ, πσ and π −1 in full cycle notation are as fol-
lows:
π = (1 3 6 4)(2 8)(5 7 9),
Recall that composing an r-cycle with itself r times gives the identity
permutation ι. We can extend this idea to all permutations (and later we
will extend it to all elements of a group) as follows.
Definition 3.10. Let π ∈ Sn . The order o(π) of π is the minimal r ∈ N
r times
r
� �� �
such that π = ππ · · · π = ι.
27
Then one particular advantage of disjoint cycle notation is that it is easy
to calculate the order of a permutation written this way.
Proposition 3.11. If π = γ1 γ2 · · · γk , where the γi are disjoint cycles, then
the order of π is the lowest common multiple of the lengths of the cycles γi .
Proof. Since the γi commute, for n ∈ N we have
π n = (γ1 · · · γk )n = γ1 n · · · γk n .
28
Notice that in this example the expression (1 4)(4 8)(8 7)(1 2)(1 5)(1 2)(5 3)
gives the same permutation. In general there will be many ways of expressing
a given permutation as a product of transpositions; the essential uniqueness
of Theorem 3.7 occurs because the cycles there are required to be disjoint. In
the next section we will see what can be said about uniqueness of expressions
in terms of transpositions.
where |γi | denotes the length of the cycle γi . Note that ν(π) is well-defined
because of the uniqueness in Theorem 3.7 (in particular, cycles of length 1
contribute 0 to the sum, and so may be ignored).
Class Exercise 3.16. If π = ( 14 25 32 48 53 66 71 87 ) = (1 4 8 7)(2 5 3)(6), then
ν(π) = (4 − 1) + (3 − 1) + (1 − 1) = 3 + 2 + 0 = 5.
Note that π −1 has the same cycle lengths as π, so we have ν(π −1 ) = ν(π).
Example 3.17. If π is as above, π −1 = (7 8 4 1)(3 5 2)(6), so
ν(π −1 ) = 3 + 2 + 0 = 5 = ν(π).
Clearly ν(π) = 1 if and only if π is a transposition. In fact we may
interpret ν(π) as follows:�if all possible 1-cycles are included in the expression
k
π = γ1 γ2 . . . γk , then i=1 |γi | = n; thus ν(π) = n − k, i.e. ν(π) is the
difference between n and the number of cycles in the full cycle notation.
Example 3.18. If π is as above, then n = 8 and k = 3, so ν(π) = 8 − 3 = 5.
Now ν(π) on its own is not particularly important, but it enables us to
make the following definition.
sign(π) = (−1)ν(π) .
29
Remark 3.21. It is possible to define the sign of a permutation π in another
way, as (−1)n(π) where n(π) is the number of ‘inversions’ of π, i.e. the
number of pairs (i, j) with i < j and π(i) > π(j); we shall not go into this
here.
We now consider how the sign of a permutation is affected by multiplic-
ation by a transposition.
Lemma 3.22. Let π ∈ Sn and let τ ∈ Sn be a transposition. Then
sign(πτ ) = (−1) · sign(π) = − sign(π).
Proof. Put τ = (a1 a2 ) and consider π as a product of disjoint cycles. We
begin by comparing ν(πτ ) and ν(π) and we deal separately with the two
cases where the cycles in π containing a1 and a2 are different or the same.
If a1 and a2 are in different cycles of π, we have
π = γ1 · · · γk−1 (a1 b1 · · · br a2 c1 · · · cs )
�� �
k−1
with r, s ≥ 0. Then ν(π) = i=1 (|γi | − 1) + r + s + 1. This time
30
So ν(πτ ) = 4 and sign(πτ ) = (−1)4 = 1 = − sign(π).
Corollary 3.24. If τ1 , . . . , τr ∈ Sn are transpositions, then
sign(τ1 · · · τr ) = (−1)r .
Proof. We may use induction on r. The result is true for r = 1, since we have
ν(τ1 ) = 1 and so sign(τ1 ) = (−1)1 . The induction step follows immediately
from Lemma 3.22.
This shows what can be said about an expression for a given permuta-
tion as a product of transpositions: an even (respectively odd) permutation
can only be written as a product of an even (respectively odd) number of
transpositions.
Theorem 3.25. Let π, σ ∈ Sn . Then sign(π) sign(σ) = sign(πσ).
Proof. By Theorem 3.14 we may write
π = τ1 · · · τr , σ = τ1 � · · · τ s �
Here
ν(π) = 1 + 2 + 1 = 4 so sign(π) = 1;
ν(σ) = 2 + 2 + 1 = 5 so sign(σ) = −1;
ν(πσ) = 1 + 1 + 3 = 5 so sign(πσ) = −1.
This result shows that if we define a function Σ : Sn → {1, −1} by setting
Σ(π) = sign(π) for all π ∈ Sn , then Σ respects products, in the sense that
Σ(πσ) = Σ(π)Σ(σ) (where the product on the right is the usual product of
integers). Such maps will be very important later on.
Notice too that if we take the product of two even permutations, we
obtain another even permutation (1 · 1 = 1!), so the subset of Sn consisting
of all the even permutations is closed with respect to the group operation.
Also, the identity permutation is even and the inverse of an even permutation
is even. When we move to the general theory shortly, such a subset will be
called a subgroup; a subgroup is in particular a group in its own right.
This particular subgroup of Sn is important enough to have its own
name.
31
Definition 3.27. The subgroup of Sn consisting of the even permutations
is called the alternating group of degree n and is denoted An .
0 0 1 0 0 0 0 1
0 0 0 1
0 1 0 0
Aστ = A(1 4) =
0
and
0 1 0
1 0 0 0
0 0 0 1 0 1 0 0 0 0 0 1
1 0 0
0 0 0 1
0 0 1 0 0
Aσ Aτ =
0 = = Aστ .
1 0 0 1 0 0 0 0 0 1 0
0 0 1 0 0 0 0 1 1 0 0 0
32
Via permutations, we have actually encountered many of the basic ideas
of group theory: groups, subgroups, functions that preserve the group op-
eration. What we will do next is to turn these into abstract definitions that
can apply to any group.
In doing so, we are following in the footsteps of the founders of the
subject: Galois, Cauchy, Lagrange, Cayley, Abel and others. For groups first
arose as groups of permutations, in the work of Galois on the solvability of
polynomial equations around 1830, and an abstract definition was not given
until work of von Dyck in 1882.
33
3.3 Elementary properties, or, algebraists like to be fussy
We are now going to embark on our systematic study of groups, starting
from the definition and working up. Along the way, there will of course be
many examples drawn from our big list above, but for a while at least things
will become more abstract.
Let us remind ourselves of the definition of a group.
(iv) For a, b ∈ G, the equation a ∗x = b has the unique solution x = a−1 ∗b.
34
Corollary 3.30. Let (G, ∗) be a group and let a ∈ G. Then there is a
bijection between the set G and the sets {a ∗ g | g ∈ G} and {g ∗ a | g ∈ G}.
Proof. First note that since G is a group, the sets {a ∗ g | g ∈ G} and
{g ∗ a | g ∈ G} are subsets of G.
Define a function La : G → G, La (g) = a ∗ g. Then the image of La is
by definition equal to the set {a ∗ g | g ∈ G} and hence it suffices to show
that La is injective.
To show this, assume that La (b) = La (c). Then
a ∗ b = La (b) = La (c) = a ∗ c.
35
product of sets X and Y is the set X × Y = {(x, y) | x ∈ X, y ∈ Y } of
ordered pairs of elements from X and Y . If X and Y have more structure,
such as being groups or rings, we may “extend” those structures to X × Y .
Proposition 3.33. Let (G, ∗) and (H, ◦) be groups. Then the Cartesian
product G × H is a group with respect to the binary operation • defined by
for all g1 , g2 ∈ G, h1 , h2 ∈ H.
Exercise 3.34. Prove Proposition 3.33.
Subsequently, if we refer to the Cartesian product of groups, G × H, we
will mean (G × H, •) with • as given in this Proposition.
3.4 Subgroups
We begin with a comment on notation. If we are working in an arbitrary
group (G, ∗) then we will write e for the identity of the group (sometimes
eG if more than one group is being discussed) and also we will sometimes
suppress the symbol for the binary operation, ∗, and write simply ab for
a ∗ b. This “multiplicative” notation (more properly called concatenation,
meaning “placing side by side”) is used for convenience and does not mean
that the binary operation is necessarily multiplication. We will also talk
about “the group G” rather than writing (G, ∗) every time, although if we
are discussing more than one group or more than one binary operation, or
if we need to explicitly talk about properties of the operation, we might be
more precise again.
If we are working in a particular group and there is an appropriate
symbol for either the identity or the binary operation, such as “0” or “+”,
then we use it.
As we saw with permutations, it is often natural to ask about subsets
of a group that are closed under the binary operation. That is, for any
two elements a, b belonging to the subset, we have a ∗ b also in the subset.
However what we really want is for the subset to itself be a group: then we
can apply all the theory of groups to it too. It turns out that just being
closed under the operation is not enough (shortly we will see what else one
should ask for) and the definition we make next is morally the right one.
Definition 3.35. Let (G, ∗) be a group. A subset H of G is said to be a
subgroup if (H, ∗) is a group. If this is the case, we write H ≤ G.
Remarks 3.36.
(a) From the definition, we see that “H is a subgroup of G” means not
just that “H is a group with respect to some operation”, but it means
that H is a group with respect to the same operation as G.
36
(b) It is immediate from this definition that every subgroup of an Abelian
group is Abelian: if a∗b = b∗a for all a, b ∈ G, then certainly a∗b = b∗a
for all a, b ∈ H ⊆ G.
(c) If you prefer to be a little bit more formal, remember that ∗ : G × G →
G is a function sending pairs of elements of G to an element in G. To
say that (H, ∗) is a group means that when we restrict to looking at
pairs of elements of H (i.e. looking at the function ∗|H×H : H ×H → G
given by restricting ∗ to H × H), we have Im ∗|H×H ⊆ H so that
there is a well-defined function ∗|H H×H : H × H → H, restricting the
domain to H × H and the codomain to H, and (H, ∗|H H×H ) is a group.
Unsurprisingly, we tend to brush this under the carpet a bit and just
write (H, ∗).
Examples 3.37.
(a) (Z, +) is a subgroup of (Q, +), and both are subgroups of (R, +).
(b) (Q∗ , ×) is a subgroup of (R∗ , ×).
(c) (Q∗ , ×) is not a subgroup of (Q, +), since the two binary operations
are different.
(d) For n ∈ N let ζ = e2πi/n ∈ C. The set Cn = {ζ k | k ∈ Z} is a subgroup
of (C∗ , ×). Notice that |Cn | = n (since e2πi = 1), so that Cn is a finite
group, while C∗ is certainly not.
(e) In the group GLn (R) (where the operation is matrix multiplication),
the set of matrices with determinant 1 is a subgroup; this follows
from properties of determinants you proved in Linear Algebra. This
subgroup has a special name: it is known as the special linear group,
SLn (R). (As before, one may replace R by any field F and obtain
SLn (F ).)
(f) We can consider Sn−1 as a subgroup of Sn : if σ ∈ Sn−1 then we think
of σ as a permutation of {1, 2, . . . , n}, by setting σ(n) = n. More
generally, the set of permutations fixing every element of some subset
of {1, 2, . . . , n} is a subgroup: if two permutations fix some i so does
their product. Also if a permutation fixes i, so does its inverse.
(g) We claimed that the set of even permutations An is a subgroup of Sn ;
this claim will be justified more rigorously very shortly.
(h) In any group G, {e} and G are subgroups.
This last pair of examples is “uninteresting” (in the sense that they don’t
tell us anything we didn’t already know) and on occasion we will want to
exclude these from claims about subgroups. So we introduce the following
terminology.
37
Definition 3.38. Let H be a subgroup of G.
The subgroup {e} is called the trivial subgroup of G. Considered as a
group in its own right, it is called the trivial group. If H �= {e}, we say that
H is a non-trivial subgroup of G.
If H �= G (so that H is a proper subset of G), we say H is a proper
subgroup of G and write H � G or more simply H < G.
In principle, to show that a subset H of a group (G, ∗) is itself a group—
that is, to check the definition of a subgroup above—one should show that
H has a well-defined binary operation that is associative, has an identity
element and where every element has an inverse. However this is a lot to
check and in fact one can check rather less6 .
We will give two lists of properties to check and we will refer to these as
the “subgroup tests”.
Proposition 3.39 (Subgroup test I). Let H be a subset of a group (G, ∗)
and denote the identity element of G by eG . Then H is a subgroup of G if
and only if the following three conditions are satisfied:
(SG1) eG ∈ H;
(SG2) for all h1 , h2 ∈ H, we have h1 ∗ h2 ∈ H;
(SG3) for all h ∈ H, we have h−1 ∈ H.
6
The corresponding situation for rings makes this even more desirable: to check that a
set with two binary operations is a ring involves checking nine properties. As we will see,
to show that a subset is a subring, it is possible to get away with checking only four.
38
It remains to show that every element of H has an inverse in H: we
know that every element of H has an inverse in G but for H to be a group
in its own right, that inverse has to be an element of H. However (SG3) tells
us precisely that the element h−1 inverse to h in G is actually an element of
H, as we need.
Be alert here: it is not enough for a subset to be closed under the group
operation for it to be a subgroup. It is possible to find examples of sub-
sets satisfying (SG2) but not one or other of (SG1) and (SG3), as we will
see shortly. But first, let us start off positively and see some examples of
subgroups and the subgroup test in action.
Examples 3.40.
(a) The set of even integers 2Z = {2r | r ∈ Z} is a subgroup of (Z, +), the
integers under addition.
(SG1) The identity of (Z, +) is 0, which is even.
(SG2) The sum of two even integers is even: 2r +2s = 2(r +s) ∈ 2Z.
(SG3) The negative of an even integer is even: if m = 2r then
−m = −2r = 2(−r) ∈ 2Z.
Similarly, for any m ∈ Z, mZ is a subgroup of (Z, +). If m = 0,
mZ = 0Z is trivial; if m = ±1, mZ = ±1Z = Z so mZ is proper
provided m �= ±1.
39
For n = 1, S1 = A1 = {ι} so A1 is trivial and not proper. For n = 2,
A2 = {ι} < S2 is trivial and proper. For n > 2, An is both proper and
non-trivial.
(d) The special linear group SLn (R) (matrices of determinant 1) is a sub-
group of GLn (R) (invertible matrices, with operation matrix multi-
plication).
(SG1) The identity matrix In has determinant 1.
(SG2) We know that det(AB) = det(A) det(B), so if A, B have
determinant 1, so does AB.
(SG3) We also know that det(A−1 ) = det(A)−1 , so if A has determ-
inant 1, so does A−1 .
For n = 1, GL1 (R) = R∗ = R\{0}; SL1 (R) = {1} is trivial and proper.
For any n > 1, SLn (R) is a proper non-trivial subgroup of GLn (R).
You might have been expecting an example in which we have (SG2) and
(SG3) but not (SG1). However this is not possible, since h ∗ h−1 = eG would
then be an element of H. So there is some redundancy in the subgroup test
above and we can give a more “efficient” version, as follows.
40
Proposition 3.42 (Subgroup test II). Let H be a subset of a group (G, ∗).
Then H is a subgroup of G if and only if the following two conditions are
satisfied:
(SG1� ) H �= ∅;
Exercise 3.43. Apply the second subgroup test (Proposition 3.42) to the
examples in Examples 3.40 and 3.41 to confirm what we found out from
using the first subgroup test (Proposition 3.39).
Exercise 3.44. Let (G, ∗) be a group and consider the set
41
Proof. Let H be an I-indexed family of subgroups of (G, ∗), with the ith
subgroup being denoted Hi . That is, for all i ∈ I, Hi is a subgroup of G.
We use the first subgroup test.
Exercise 3.46 (Not essential, but good practice with abstract arguments).
� the subgroups: for any collection H of subgroups
We do not need to index �
of G, we may define H = {x | (∀ H ∈ H)(x ∈ H)}. Prove that H is a
subgroup of G.
Remarks 3.47.
(a) We used the first subgroup test rather than the second, as�it is not
enough to know that each Hi is non-empty to conclude that i∈I Hi is
non-empty. (Two non-empty but disjoint sets have empty intersection,
e.g. {1}∩{2} = ∅.) We need to know that there is a particular element
that belongs to every Hi , which there is, namely eG .
(b) The union of subgroups need not be a subgroup: let H1 = 2Z, H2 = 3Z
be subgroups of (Z, +). Then 2, 3 ∈ H1 ∪ H2 but 2 + 3 = 5 ∈
/ H 1 ∪ H2 .
Example 3.48. In (Z, +), the intersection of 2Z and 3Z is the set of integers
that are divisible by 2 and divisible by 3, and these are precisely the integers
divisible by 6. So 2Z ∩ 3Z = 6Z.
More generally, mZ ∩ nZ = lcm(m, n)Z and
m1 Z ∩ m2 Z ∩ · · · ∩ mr Z = lcm(m1 , m2 , . . . , mr )Z.
The collection H �= {mZ | m ∈ N} is a N-indexed family of subgroups of
Z. The intersection m∈N mZ consists of all � integers that are divisible by
every m ∈ N. Only 0 has this property, so m∈N mZ = {0} is trivial.
Sometimes we have a subset S of a group G that is not a subgroup, but
where we would like to find a subgroup of G that contains S. Obviously we
could take all of G but this won’t tell us much: we really want the smallest
subgroup of G containing S.
“Smallest” here might mean “smallest size”, but for infinite sets it is
better to think of smallest as “smallest under inclusions of sets”. But we
know how to find the smallest set: it is the intersection. (The intersection is
a subset of every set in the collection and is the unique smallest such under
inclusion.) This leads us to make the following definition.
42
Definition 3.49. Let G be a group and let S ⊆ G be a subset of G. Let
S = {H ≤ G�| S ⊆ H} be the family of subgroups of G that contain S. The
intersection S is a subgroup of�G containing S, called the subgroup of G
generated by S. We write �S� = S.
Notice that by Proposition 3.45 (or rather the slightly more general
version in the exercise that follows) �S� is a subgroup of G. It is also the
unique subgroup having the desired property, namely that it contains S. If
S is actually a subgroup already, then �S� = S.
Example 3.50. In (Z, +), we have �m� = mZ (and in particular �1� = Z)
and �{m, n}� = hcf(m, n)Z, where hcf denotes the highest common factor.
We will discuss this in more detail later.
To simplify notation, if S = {g1 , . . . , gr }, we will write �g1 , . . . , gr � rather
than �{g1 , . . . , gr }� for the subgroup generated by S. So for example, we
would write �m, n� = hcf(m, n)Z.
The above description of the subgroup generated by S as an intersection
is good from the point of view of proving that it is actually a subgroup, but
bad from the point of view of knowing what �S� looks like as a subset of G.
So we are now going to try to tie down a more precise description of
the subgroup generated by S, as the set of elements of G which can be
“expressed in terms of” the elements of S. For example, if x, y ∈ S (and
hence in �S� since S ⊆ �S�) then xy 2 x−1 y 2 x3 y −1 x is expressed in terms of
x and y and is in �S�, by repeated application of the subgroup test.
The following proposition expresses this idea in more formal language.
43
We have �S� ⊆ H since H is a subgroup of G which contains S. But
any subgroup of G which contains S must contain each element of H by
(repeated) application of the subgroup test. Thus H ⊆ �S� and so H = �S�.
(a) The integers modulo n under addition, (Zn , +), is a cyclic group.
44
(e) The subset C12 of C∗ comprising the roots of x12 −1 is a cyclic subgroup.
More explicitly, C12 = {e2kπi/12 | 0 ≤ k < 12}. Setting ζ = e2πi/12 , we
observe that ζ, ζ 5 , ζ 7 and ζ 11 each �generate
� C12 but no other powers
of ζ do. For example, ζ 6 = −1 and ζ 6 = {ζ 0 , ζ 6 } = {−1, 1}.
ζ3
ζ4 • ζ2
• •
ζ5 ζ1
• •
ζ6
• •
ζ 0 = 1 = ζ 12
• •
ζ7 ζ 11
• •
ζ8 • ζ 10
ζ9
Even without many tools to hand, we can completely describe the struc-
ture of cyclic groups, which we shall do in the following results.
First, we generalise Definition 3.10 by giving a definition of the order of
an element of an arbitrary group.
Definition 3.54. Let G be a group and let g ∈ G. Define o(g), the order
of g, to be |�g�|.
The phrasing of this definition may surprise you a little: you might have
been expecting o(g) to be defined to be the least positive integer n such that
g n = e, as for permutations. However for general groups, it is possible that
no such n exists. Rather than have to separate the two cases, this definition
handles them both.
Proposition 3.55. Let G = �g� be a cyclic group. Then exactly one of the
following two statements is true:
45
(ii) G is (countably) infinite, in which case the elements g i , i ∈ Z, are
distinct. This group will be denoted C∞ .
g p = g qm+r = (g m )q g r = eq g r = g r
since g m = e. Hence every element of �g� is of the form g r for some 0 ≤ r < m
and �g� = {g r | 0 ≤ r < m}.
To complete the proof that |�g�| = m, it remains to show that the
elements g r , 0 ≤ r < m are distinct. But if we had g s = g t for some
0 ≤ s < t < m, then g t−s = e and t − s < m, contradicting the minimality
of m. So |�g�| = m.
Now suppose �g� is finite. The powers of g cannot all be distinct (else
�g� would be infinite) and g s = g t for some s �= t. Then the above proves
all the claims of (i).
Since we have also shown that if the powers of g are not all distinct then
�g� is finite, we conclude that if (i) does not hold, (ii) must.
46
Remark 3.58. This is a very special case of a general problem, to understand
the subobjects of an algebraic structure. Without any assumptions, it is
hopeless to expect to classify all subgroups of all groups in any meaningful
way and even with assumptions, it is often extremely hard.
However, there are some particular situations where we can give a de-
scription and the case of cyclic groups is one such. Later we will see a closely
related problem for rings, namely the classification of prime ideals of Z.
Proof.
(i) If H = {e}, then H is generated by e and certainly cyclic, so assume
H is non-trivial. Then there exists at least one 0 �= k ∈ Z such that
g k ∈ H; without loss of generality, we may assume k > 0 since g k ∈ H
if and only if g −k ∈ H (by (SG3)). Hence we may let n be the smallest
natural number such that g n ∈ H, so that �g n � ≤ H.
We claim that H = �g n �. Let h ∈ H. Since H ≤ G = �g�, we may
write h = g m for some m. Then there exist integers q, r such that
m = qn + r with 0 ≤ r < n. Thus
47
(iii) By (i), if G = �g� is of infinite order and H is a subgroup then H =
�g m � for some m ≥ 0. If G is of infinite order then the powers of g are
distinct, so the same is true for the powers of g m , whence H is infinite
by Proposition 3.55.
48
(countably infinitely many) but 8Z ⊆ 4Z so 8Z is smaller. But how would
we compare 4Z and 7Z?
The answer is that we can indeed make mathematical sense of the natural
claims that 4Z covers a quarter of the elements of Z and 7Z a seventh, but
by counting how many “copies” or “translations” of 4Z we need to cover Z
rather than counting the number of elements.
This leads us to the penultimate topic in our study of the internal struc-
ture of groups, namely cosets.
3.6 Cosets
Throughout this section, we will consider a group G and a subgroup H
of G. As previously, we will mostly write the group operation in G by
concatenation: if g, h ∈ G, gh will denote the element obtained from g and
h (in that order) under the binary operation of the group. The exception
will be in groups where the binary operation is addition (such as (Z, +)),
where we will write g + h.
{ (1 2)ι = (1 2),
(1 2)(1 2 3) = (1)(2 3) = (2 3),
(1 2)(1 3 2) = (1 3)(2) = (1 3) }
First of all, we see that gH ⊆ G. Since e ∈ H, we have g = ge ∈ gH.
Indeed, every element of gH is, by definition, a “multiple” of g by an element
of H; the left coset gH is precisely all the elements of G that can be written
as gh for some h ∈ H.
If H = {e, h1 , h2 , . . . , hr } is finite, this is clear: gH = {g, gh1 , gh2 , . . . , ghr }
and by Proposition 3.29(iii) (the cancellation property for groups) ghi = ghj
would imply hi = hj , so the elements g, gh1 , . . . , ghr are distinct. In partic-
ular, |gH| = |H|. It is natural to think of gH as the “translations by g” of
the elements of H.
In fact, the claims of the above paragraph still hold if H is infinite, if we
work a little more abstractly.
49
Proof. This follows immediately from Corollary 3.30 and Remark 3.31, but
it is more instructive to write out the proof directly, so we shall.
We need to show that the function Lg is injective and surjective. As-
sume first that Lg (h) = Lg (k) for some h, k ∈ H. Then gh = gk and by
Proposition 3.29(iii), h = k so Lg is injective.
It is surjective since for gh ∈ gH, Lg (h) = gh.
So any two cosets of H have the same size. Notice too that eH = H so
H is a coset of itself (the “translation” of H that does nothing).
However it is very much not true that if g1 �= g2 then g1 H �= g2 H;
shortly we will see the correct condition for when two cosets are equal. This
condition will also tell us when a coset can be a subgroup (sneak preview:
only the “trivial” coset eH = H is).
First, let us look at a very important example.
Example 3.65. We consider (Z, +) and its (cyclic) subgroup 4Z. Since we
are working in an additive group, we write the group operation in our coset
notation, so the cosets of 4Z are r + 4Z = {r + m | m ∈ 4Z} = {r + 4n | n ∈
Z}.
Explicitly,
50
and
3 + 4Z -17 -13 -9 -5 -1 3 7 11 15
2 + 4Z -14 -10 -6 -2 2 6 10 14
1 + 4Z -15 -11 -7 -3 1 5 9 13 17
0 + 4Z -16 -12 -8 -4 0 4 8 12 16
We see that the left cosets of 4Z are precisely the congruence classes of
integers modulo 4. The congruence class modulo 4 containing 3, for example,
is the set of integers {. . . , −9, −5, −1, 3, 7, 11, . . . }, which is exactly the set
of integers of the form 3 + 4n, namely {3 + 4n | n ∈ Z} = 3 + 4Z. So cosets
in particular describe congruence classes of integers.
Notice that every integer belongs to some left coset and that no integer
belongs to more than one left coset. This is because the congruence classes
modulo n for some fixed n partition Z, since congruence modulo n is an
equivalence relation. The following theorem is a vast generalisation of this
result, which corresponds to the specific group (Z, +).
g RH k ⇐⇒ g −1 k ∈ H
is an equivalence relation.
(ii) The equivalence classes for RH are precisely the left cosets gH, g ∈ G.
51
Proof.
g RH k ⇐⇒ g −1 k ∈ H
⇐⇒ g −1 k = h for some h ∈ H
⇐⇒ k = gh for some h ∈ H
⇐⇒ k ∈ gH.
{k | g RH k} = {k | k ∈ gH} = gH.
Examples 3.70.
(a) In S3 , for
H = �(1 2 3)� = {ι, (1 2 3), (1 3 2)}
we saw that the coset (1 2)H in S3 is
52
Notice too that (1 2) RH (1 3) since
(1 2)−1 (1 3) = (1 2)(1 3) = (1 3 2) ∈ H
(b) Recall that in (Z, +), nZ is a subgroup and its cosets are the sets of
the form r + nZ = {r + mn | m ∈ Z}. The corollary tells us that
r + nZ = s + nZ ⇐⇒ r − s ∈ nZ
�
⇐⇒ n � r − s
⇐⇒ r ≡ s mod n,
3 + 4Z -17 -13 -9 -5 -1 3 7 11 15
2 + 4Z -14 -10 -6 -2 2 6 10 14
1 + 4Z -15 -11 -7 -3 1 5 9 13 17
0 + 4Z -16 -12 -8 -4 0 4 8 12 16
53
We see that (1, 2) + H = (7, 6) + H since (7, 6) − (1, 2) = (6, 4) =
2(3, 2) ∈ H. (In the diagram, H = �(3, 2)� is indicated by red circles
and (1, 2) + H is indicated in blue.)
This answers our question about the relative size of 4Z and 7Z in (Z, +).
We see from the example above that 4Z has index 4 and a similar set of
calculations shows that 7Z has index 7. So 4Z and its three translations
1 + 4Z, 2 + 4Z and 3 + 4Z cover Z; 7Z needs seven translations to cover Z.
54
Example 3.72. For n ≥ 2, the set of even permutations An in Sn has index
|Sn : An | = 2. By Theorem 3.25, if π ∈
/ An and τ = (1 2) then
55
Analogously to Definition 3.71, define the right index of H in G to be
the number of right cosets of H in G and denote it by |H : G|.
Example 3.75. In S4 , let
{ ι(1 2) = (1 2),
(1 2 3 4)(1 2) = (1 3 4),
(1 3)(2 4)(1 2) = (1 4 2 3)
(1 4 3 2)(1 2) = (2 4 3) }
Notice that this is not the same example as before: this is an example
in S4 and we have used K for the subgroup to emphasise that this is differ-
ent. The reason is that the previous example was too small to illustrate an
important point about left and right cosets.
Namely, in general the left cosets of a subgroup are not the same as the
right cosets; certain special conditions have to hold for this to be the case.
In the example, the left coset (1 2)K is {(1 2), (2 3 4), (1 3 2 4), (1 4 3)}
which we see is not equal to K(1 2).
So we must be careful to specify whether we mean left cosets or right
cosets. Fortunately, while we need to make a choice, some key properties
are independent of which choice we make. The size of a left or a right coset
is the same: every coset, whether left or right, has the same size as H (there
is a natural right-handed version of Lemma 3.64).
A little more work also shows that the number of left cosets of H in G is
equal to the number of right cosets of H in G; that is, |G : H| = |H : G|. In
fact, this is why we just say “index” and not “left index” or “right index”.
This is justified by the following exercise.
Exercise 3.76. Let H be a subgroup of G. Let LH = {gH | g ∈ G} and
RH = {Hg | g ∈ G} be the sets of left and right cosets of G respectively.
Prove that the function ψ : LH → RH , ψ(gH) = Hg −1 is a bijection.
This function ψ is a function between two sets: in general, the set of
cosets of a subgroup has no algebraic structure.
Remark 3.77. Subgroups such that gH = Hg for all g ∈ G are very im-
portant, so much so that they have a special name. They are called normal
subgroups and we will examine them and their importance in detail in the
following sections.
But, as a glimpse of what is to come, we will see that the set of (left)
cosets of a normal subgroup can be given an algebraic structure. Specifically,
56
the group operation descends7 to a group operation on the set of cosets: if
gH = Hg for all g ∈ G, we can define a binary operation on cosets by
57
Our mental picture should be the elements of G laid out in a grid, with
each row being a list of the elements of H or a coset of it, and the columns
indexed by (some fixed choice of) representatives for the cosets � (the30 trans-
�
versal� TH
� ). As a specific example, let us take G = C 30 = g|g =e ,
H = g 5 and TH = {g 0 = e, g, g 2 , g 3 , g 4 }, as follows.
TH
g4H g4 g9 g 14 g 19 g 24 g 29
g3H g3 g8 g 13 g 18 g 23 g 28
g2H g2 g7 g 12 g 17 g 22 g 27
g1H g1 g6 g 11 g 16 g 21 g 26
g0H g0 g5 g 10 g 15 g 20 g 25
Corollary 3.84. A group of prime order is cyclic and has no proper non-
trivial subgroups. Any non-identity element generates the group.
Proof. If |G| = p with p prime, then any subgroup must have order either
1 or p. Thus if e �= g ∈ G, then |�g�| = p and �g� = G. Hence G is cyclic
and generated by any non-identity element. Any subgroup of such a group
is either trivial (order 1) or it contains �g� for some g �= e and is equal to G.
Groups with no proper non-trivial normal subgroups are called simple groups.
At the end of our work on groups, we will see in what sense these groups are
“simple” (they can be extremely large and complicated by other measures)
and why they are the building blocks for all groups. Cyclic groups of prime
order are very simple: they have no proper non-trivial subgroups at all!
58
Proposition 3.85. If H, K ≤ G and hcf(|H|, |K|) = 1, then H ∩ K = {e}.
Proof. Let |H ∩ K| = r. Then r must divide both |H| and |K|, as H ∩ K is
a subgroup of both H and K, and so r = 1.
and
as required.
59
Example 3.89. If G = S5 , the elements g = (1 2 3) and h = (4 5) satisfy
gh = hg (they are disjoint cycles), and their orders are 3 and 2, which are
coprime; we have gh = (1 2 3)(4 5) = hg, whose order is 6 since
(gh)1 = (1 2 3)(4 5) �= ι
(gh)2 = g 2 h2 = g 2 = (1 3 2) �= ι
(gh)3 = g 3 h3 = h3 = h = (4 5) �= ι
(gh)4 = g 4 h4 = g = (1 2 3) �= ι
(gh)5 = g 5 h5 = g 2 h = (1 3 2)(4 5) �= ι
(gh)6 = (g 3 )2 (h2 )3 = ι
If gh = hg we say that the elements g and h commute. This condition is
certainly needed for the previous result: if we take G = S3 with g = (1 2 3),
h = (2 3), then o(g) = 3 and o(h) = 2, but gh = (1 2) so o(gh) = 2 �= 2 · 3.
(You can easily check for yourself that gh �= hg.)
This concludes our examination of the “internal” structure of groups.
What we are missing is a way to compare two different groups, by some
suitable functions between them that respect the group structure. This is
the topic of the next chapter.
60
4 Relating different groups
4.1 Homomorphisms
When we have two groups G and H and we want to relate the group struc-
ture on G to that on H, to compare properties between them, the right
mathematical thing to do is to start with a function ϕ : G → H. This gives
us a relationship between the underlying sets but this need not relate the
group structures: we need ϕ to be compatible with the binary operations
on G and H.
This leads us to the following definition.
Definition 4.1. Let (G, ∗) and (H, ◦) be groups. A function ϕ : G → H is
called a group homomorphism if
ϕ(g1 ∗ g2 ) = ϕ(g1 ) ◦ ϕ(g2 )
for all g1 , g2 ∈ G.
Notice in particular that the operation in G, ∗, is being used on the left-hand
side of this equation (on g1 , g2 ∈ G), and that the operation in H, ◦ is being
used on the right-hand side (on ϕ(g1 ), ϕ(g2 ) ∈ H). This is an instance when
it is helpful to write the group operations explicitly.
Remark 4.2. This idea of a structure-preserving function is one you have
seen before, in MATH112 Discrete Mathematics. There we discussed iso-
morphisms of graphs, as bijective functions between the sets of vertices that
preserve the edges.
We will come to isomorphisms of groups shortly, but just for now, ignore
the “bijective” bit and concentrate on the “function preserving structure”
bit. In fact, this same idea is used to define homomorphisms for any algebraic
structure, as we will see again later when we talk more about rings.
First, let us deal with some elementary properties of homomorphisms.
Proposition 4.3. Let ϕ : G → H be a group homomorphism. Then
(i) ϕ(eG ) = eH ;
(ii) ϕ(g −1 ) = ϕ(g)−1 for all g ∈ G;
(iii) ϕ(g n ) = ϕ(g)n for all g ∈ G, n ∈ Z and
(iv) o(ϕ(g)) | o(g) for all g ∈ G.
Proof.
61
(ii) Since gg −1 = eG = g −1 g, we have
(σ ◦ ϕ)(g1 g2 ) = σ (ϕ(g1 g2 ))
= σ (ϕ(g1 )ϕ(g2 ))
= σ (ϕ(g1 )) σ (ϕ(g2 ))
= (σ ◦ ϕ)(g1 )(σ ◦ ϕ)(g2 )
Examples 4.5.
62
(b) The function ϕ : R∗ → R∗ , ϕ(x) = x2 for all x ∈ R∗ = (R \ {0}, ×) is
a group homomorphism since
(d) Recall from Remark 3.28 that we have a function ϕ : Sn → GLn (R),
ϕ(σ) = Aσ where �
1 if σ(j) = i
(Aσ )ij =
0 otherwise
is the permutation matrix associated to σ.
Then Aστ = Aσ Aτ , where the left-hand side involves the composition
of permutations and the right-hand side is the usual multiplication of
matrices. (The definition of Aσ , with a 1 when σ(j) = i might seem
to be the “wrong way round”; however if we made the more natural-
looking definition with σ(i) = j, we would have an order reversal in
the products.) It follows that ϕ : Sn → GLn (R), ϕ(σ) = Aσ is a group
homomorphism. For an explicit example, look back to Remark 3.28.
Exercise 4.6. Let (G, ∗) and (H, ◦) be groups with identity elements eG and
eH respectively. Show that the function ϕ : G → H defined by ϕ(g) = eH
for all g ∈ G is a group homomorphism. We call this the trivial group
homomorphism.
Recall that certain sorts of functions are special, namely injective, sur-
jective and bijective functions; also, a function is bijective if and only if
it is invertible. If a group homomorphism has one of these properties, we
(sometimes) use a special name for it. So an injective group homomorphism
is also known as a monomorphism and a surjective group homomorphism is
called an epimorphism.
63
Examples 4.7.
but 3 ∈
/ Im ϕ.
and Im ϕ = {z ∈ C | �z� = 1} � C∗ .
64
Definition 4.9. We say that two groups G and H are isomorphic if and only
if there exists a group isomorphism ϕ : G → H. If G and H are isomorphic,
we often write G ∼ = H, for short.
Lemma 4.10. Let G and H be groups. Then the following are are equival-
ent:
(i) G and H are isomorphic;
65
property is preserved under isomorphism, then if one group has the property
but the other does not, they cannot be isomorphic.
Examples of invariants are the order of the group (as above), the set of
natural numbers that are the orders of the elements of the group (see the
next example), being Abelian and being cyclic.
Example 4.11. The groups C6 and S3 are not isomorphic. They have the
same order (|C6 | = 6 = 3! = |S3 |) but C6 is Abelian and S3 is not. Indeed,
C6 is cyclic and S3 is not (if it were, every permutation in S3 would have to
be an r-cycle for some r, but S3 has both transpositions and 3-cycles).
However, sharing some properties is not (usually) enough to prove that
the groups are isomorphic: almost always you will need to find an explicit
isomorphism between them. That said, if two groups share lots of properties,
this might reasonably lead you to conjecture that they are isomorphic; this
would not be a proof, though.
Example 4.12. Recall that the subset C12 = {e2kπi/12 | 0 ≤ k < 12} of C∗ is
a cyclic subgroup, so is a cyclic group in its own right. We may write it as
C12 = �ζ� for ζ = e2πi/12 .
Let G = �g� be the (generic) cyclic group of order 12, as defined in
Proposition 3.55. So G = {e, g, g 2 , . . . , g 11 } with group operation g r g s =
g r+s . We claim that C12 is isomorphic to G.
Let ϕ : C12 → G be defined by ζ k = g k for all 0 ≤ k < 12. Clearly ϕ is a
bijection and
so ϕ is a homomorphism.
Now consider (Z12 , +), the integers modulo 12, with addition (modulo
12). So Z12 = {�
0, � a + �b = �
� with �
1, . . . , 11} c if and only if a + b ≡ c mod 12.
We claim that this group is also isomorphic to G, and hence is isomorphic
to C12 too.
The function we need is ψ : Z12 → G, ψ(� a) = g a . Then ψ is a bijec-
a b
tion: g = g if and only if a ≡ b mod 12 by Exercise 3.60. It is a group
homomorphism since
So ψ is an isomorphism.
The observant among you will have noticed that the number 12 is irrel-
evant in this example: everything works for a general n ∈ N. We can extend
it further, as follows, obtaining a more formal version of the discussion in
Remark 3.56.
Proposition 4.13. Let G and H be cyclic groups. Then G is isomorphic
to H if and only if they have the same order, |G| = |H|.
66
Proof. As discussed above, if G is isomorphic to H, then G and H have the
same order.
Now assume that |G| = |H|. Let G = �g� and H = �h� and define a
function ϕ : G → H by ϕ(g r ) = hr for all r ∈ Z. By the same argument as
in the previous example, ϕ is a homomorphism.
That ϕ is a bijection follows from Proposition 3.55.
Note that neither the statement or the proof assume that the cyclic
groups are finite.
As we said before, two groups having the same order is far from being
enough to tell us that the groups are isomorphic. The message we take from
this proposition is that being cyclic is a very strong condition: same order
plus cyclic is enough to give us isomorphic.
Previously we noted that the identity map provides an isomorphism of
a group with itself. In fact, a group can have other self-isomorphisms and
it is often an important question to know how many. In some sense, these
are the symmetries of the group (which itself may be the symmetries of
something!).
Proof. Since the composition of bijections is a bijection and since the com-
position of homomorphisms is a homomorphism (Proposition 4.4), we see
that the composition of two automorphisms of G is again an automorphism
of G.
Composition of functions is associative and the identity function is an
automorphism. The inverse of an isomorphism is also an isomorphism
(Lemma 4.10 and the discussion after it) so in particular this is true for
an automorphism. Hence we have a group.
Example 4.17. One can show that Aut(Cn ) ∼ = ((Zn )∗ , ×), the group of units
of the ring of integers modulo n. In particular, we have Aut(C2 ) = {e} and
Aut(C3 ) ∼= C2 .
We will not justify these claims here: finding the automorphism group of
even a small group usually requires more sophisticated technology than we
67
currently have to hand. We will simply note that the above has an important
relationship with number theory: n �→ |Aut(Cn )| is called the Euler totient
function10 and is a very important function about which much more is said
in MATH328 Number Theory.
Instead, we will move on and look at some important subsets of a group
associated to homomorphisms.
10
Also known as Euler’s phi function, since it is often denoted ϕ(n).
68
Ker ϕ Im ϕ
eG eH
ϕ
G H
Very shortly we will come to examples but first we will prove some basic
facts about kernels and images that will be very helpful in working out the
examples.
69
Proposition 4.20. Let ϕ : G → H be a group homomorphism.
Proof.
and so g −1 ∈ Ker ϕ.
(ii) Similarly:
and so h−1 ∈ Im ϕ.
70
(iii) Assume that ϕ is injective and let g ∈ Ker ϕ, so ϕ(g) = eH . By (i),
eG ∈ Ker ϕ also. Then ϕ(g) = ϕ(eG ), but ϕ is injective, so we must
have g = eG . That is, if ϕ is injective, Ker ϕ = {eG }.
Conversely, assume that Ker ϕ = {eG } and let g1 , g2 ∈ G be such
that ϕ(g1 ) = ϕ(g2 ). Then since ϕ is a homomorphism, using Proposi-
tion 4.3(ii),
(v) This follows immediately from the previous two parts, since an iso-
morphism is required to be a bijection, i.e. injective and surjective.
Notice the similarities between the proofs of parts (i) and (ii), and what
ingredients we need for these.
Remark 4.21. Let ϕ : G → H be an injective homomorphism. Then we can
consider the function ψ = ϕ|Im ϕ : G → Im ϕ, the codomain restriction to
Im ϕ of ϕ. (This is the function that takes the same values as ϕ but where
we “throw away” any elements of H not in the image of ϕ, and so just take
Im ϕ as the codomain.) Since Ker ψ = Ker ϕ = {eG } and Im ψ = Im ϕ, ψ is
both injective and surjective. Hence ψ is an isomorphism of G with Im ϕ.
Since Im ϕ is a subgroup of H, in this situation we say that G is iso-
morphic to a subgroup of H.
That is, to show that a group G is isomorphic to a subgroup of a group
H, we show that there exists an injective homomorphism from G to H. If
that homomorphism is also surjective, we have that G is isomorphic to H
itself.
Let us revisit the examples from Examples 4.5 and 4.7.
Examples 4.22.
71
square is equal to 1 (which is the identity element in R∗ = (R\{0}, ×)).
So Ker ϕ = {1, −1}.
The image of ϕ is the set of all non-zero real numbers that are squares;
there is no “nice” description of this, so we simply have
Im ϕ = {y ∈ R∗ | y = x2 for some x ∈ R∗ }.
Im ϕ = {z ∈ C | �z� = 1} � C∗ .
Ker Σ = {σ ∈ Sn | sign(σ) = 1}
since 1 is the identity element in ({1, −1}, ×). By definition, these are
the even permutations, An , so Ker Σ = An . Since there exist both
even and odd permutations, Σ is surjective and Im Σ = {1, −1}.
gK = {l ∈ G | g Rϕ l} = {l ∈ G | ϕ(g) = ϕ(l)}.
72
Proof. Given Theorem 3.66 and the above discussion, it suffices to see that
(similarly to the proof of Proposition 4.20(iii))
g 1 Rϕ g 2 ⇐⇒ ϕ(g1 ) = ϕ(g2 )
⇐⇒ eH = ϕ(g1−1 g2 )
⇐⇒ g1−1 g2 ∈ Ker ϕ = K
⇐⇒ g 1 RK g 2 .
Since by Theorem 3.66(ii), the equivalence classes for RK are precisely the
left cosets of K = Ker ϕ, we have the first assertion.
So g1 K = g2 K if and only if ϕ(g1 ) = ϕ(g2 ), so ψ is well-defined and
injective. Since if h ∈ Im ϕ, there exists g ∈ G such that ϕ(g) = h, we have
ψ(gK) = ϕ(g) = h and ψ is surjective and hence a bijection.
This is the analogous result for groups to the Dimension Theorem for vector
spaces, which asserts that for a linear transformation T : V → W we have
dim V = dim Ker T + dim Im T .
In fact, both of these are “numerical shadows” of stronger statements:
shortly we will introduce quotient groups and prove the stronger version for
groups.
Since Ker ϕ ≤ G and Im ϕ ≤ H, we already knew that |Ker ϕ| divides |G|
and |Im ϕ| divides |H|, by Lagrange’s theorem. However this corollary tells
us that |Im ϕ| also divides |G|, which can be useful to know, in applications
such as the following.
Example 4.25. Let G be a group of order 16 and H a group of order 9.
Then the only homomorphism ϕ : G → H is the trivial homomorphism with
ϕ(g) = eH for all g ∈ G. For if ϕ : G → H is a homomorphism, Im ϕ must
divide |G| and |H|, but hcf(16, 9) = 1 so |Im ϕ| = 1 and so Im ϕ = {eH }.
To conclude this section, we recall that properties such as being cyclic
are preserved by isomorphisms and also that (as in Proposition 4.13) if we
have a cyclic group around, we can often say much more than for arbitrary
groups.
In this spirit, we have the following proposition, which tells us that being
cyclic is preserved by surjective homomorphisms (not just isomorphisms,
which are also required to be injective).
73
Proposition 4.26. Let G = �g� be a cyclic group and let H be any group.
If ϕ : G → H is a homomorphism, then Im ϕ is cyclic.
Proof. For any h ∈ Im ϕ, there exists r ∈ Z such that g r ∈ G = �g� and
ϕ(g r ) = h. But by Proposition 4.3(iii), ϕ(g r ) = ϕ(g)r so h ∈ �ϕ(g)� and
hence Im ϕ ⊆ �ϕ(g)�. Since ϕ(g) ∈ Im ϕ, �ϕ(g)� ⊆ Im ϕ and hence Im ϕ =
�ϕ(g)� is cyclic, generated by ϕ(g).
Note again that we make no distinction between finite and infinite cyclic
groups, although if G = �g� is a finite cyclic group, Im ϕ must be finite also.
However, we can certainly have G infinite cyclic and Im ϕ finite, as we can
see from the case G = Z and ϕ : Z → Zn , ϕ(m) = m mod n (which we will
study in more detail soon).
so a = b and λ is injective.
74
Corollary 4.29. Every group G is isomorphic to a subgroup of SG .
In particular, every finite group G of order n is isomorphic to a subgroup
of the symmetric group of degree n, Sn .
Proof. Since Sn is a finite set (of cardinality n!), it has only finitely many
subsets (|P(Sn )| = 2n! ). Therefore Sn has only finitely many subgroups. By
Corollary 4.29, every group of order n is isomorphic to a subgroup of Sn , so
there can be only finitely many groups of order n up to isomorphism.
The phrase “up to isomorphism” at the end here is very important: there
are infinitely many groups with one element, for example, but they are all
isomorphic11 .
Actually, the estimate in this proof is not very good: 2n! is much larger
than the number of subgroups of Sn . This is not surprising: there are lots
of subsets of Sn that are not subgroups (all those that do not contain ι, for
starters).
We do not know exact formulæ for the number of subgroups of Sn , either
up to equality or up to isomorphism, though we do have the following:
11
For each x ∈ R, let Ex = ({x}, ∗) be the group with binary operation x ∗ x = x (note
that ∗ is neither + or ×!). Then if x �= y, Ex �= Ey but Ex ∼ = Ey for all x, y ∈ R via
x �→ y.
12
Online Encyclopedia of Integer Sequences, A005432, https://oeis.org/A005432.
13
Online Encyclopedia of Integer Sequences, A174511, https://oeis.org/A174511.
75
Example 4.31. Let G = C6 , generated by g such that g 6 = e. Let a = g 2 .
Then � �
e g g2 g3 g4 g5
La = .
g2 g3 g4 g5 e g
We could then put all of the La ’s together, into a table:
Example 4.32. For G = C6 as above, we have
e g g2 g3 g4 g5
e e g g2 g3 g4 g5
g g g2 g3 g4 g5 e
g2 g2 g3 g4 g5 e g
g3 g3 g4 g5 e g g2
g4 g4 g5 e g g2 g3
g5 g5 e g g2 g3 g4
A table such as this is known as a Cayley table. Although they were used
as ways of explicitly describing groups around the time that group theory
was being developed, the more abstract approaches we have been taking
soon meant that Cayley tables fell out of fashion (except in primary school
teaching, where they are better known as addition and multiplication tables,
or “times tables”). For while Cayley tables are very explicit, it is not hard
to see that for groups larger than about ten elements, they quickly become
unwieldy. Even checking basic properties such as associativity is not always
straightforward, for example.
Cayley’s theorem also gives us the first of our methods for calculating in
groups. Namely, we can take our group and represent it as a group of per-
mutations: G is isomorphic to its image under the injective homomorphism
λ : G → SG . Since computers can compose and invert permutations quickly,
this is effective, even for very large finite groups.
From a theoretical point of view, Cayley’s theorem tells us that per-
mutation representations exist, that is that every group can be thought of
as acting on a set S of cardinality |G| by permuting S. Other permuta-
tion representations might exist too, i.e. G might act on other sets. Group
actions will be studied in more detail in MATH321 Groups and Symmetry.
We can also join two previous results together: Cayley’s theorem and the
homomorphism from the symmetric group to the general linear group given
by taking permutation matrices. This gives us, for free, a matrix version of
Cayley’s theorem.
76
Proof. Firstly, by Cayley’s theorem, there is an injective homomorphism
λ : G → SG . For G finite, there is an isomorphism κ : SG → S|G| so that
composing these, we have an injective homomorphism λ� = κ ◦ λ : G → S|G| .
(Remember from MATH112 Discrete Mathematics that the composition of
two injective functions is again injective.)
Recall from Example 4.5(d) and Example 4.7(d) that there is an injective
homomorphism ϕ : Sn → GLn (R), given by ϕ(σ) = Aσ , the permutation
matrix associated to σ.
Setting n = |G|, we obtain ϕ ◦ λ� : G → GL|G| (R) an injective homo-
morphism as required.
77
Proof. Let G and H be groups of order p. By Corollary 3.84, G and H are
cyclic groups (of the same order). Then by Proposition 4.13, G and H are
isomorphic.
78
B A C
A α α C α α B
C B A
β β β
C B A
β β β
A α2 α2 C α2 α2 B
B A C
Lemma 4.37. Let D2n be the dihedral group of order 2n. Then D2n is
generated by two elements α and β subject to the following relations:
• αn = e,
• β 2 = e, and
We are quietly overlooking a few minor issues, such as when we say “the”
dihedral group. It is not hard to show that the dihedral group is unique to
up to isomorphism, with suitable technical preparation, but we will omit
this. We have also not carefully defined what we mean by relations but we
need not worry about this either.
Notice that the dihedral groups are not Abelian in general: α and β do
not commute (unless n = 2; see below), since αβ = βα−1 �= βα.
The “base case” D4 needs a little extra comment, as the “regular 2-gon”
does not quite make sense. Rather, D4 is the group of isometries of R2
79
preserving a rectangle. (A rectangle is less symmetric than a square (P4 ),
which is why it has a smaller symmetry group.)
Looking at the explicit description above, D4 = {e, α, β, αβ = βα} and
in fact D4 ∼= C2 × C2 , where the first cyclic group of order 2 is {e1 , a} and
the second is {e2 , b} and α �→ (a, e2 ), β �→ (e1 , b).
Here C2 × C2 is the Cartesian product (Proposition 3.33) of two iso-
morphic copies of C2 . This group is also known as the Viergruppe (German
for “four group”) and notice that D4 is Abelian, since C2 is.
In fact, this and the cyclic group of order 4 are the only possibilities.
Proof. Let H be the subgroup of G generated by {a, b}. Since a and b have
order 2 and commute, any product of a and b will be equal to one of the
elements of the set {e, a, b, ab}. (For example, abbaababa = a5 b4 = a.) This
set is also closed under inverses (indeed, every element is self-inverse) so
H = {e, a, b, ab}. Since by Lagrange’s theorem, the order of a subgroup
divides the order of the group, we are done.
Proof. Let |G| = 6, and assume G is not cyclic; then the possible orders of
elements of G are 1, 2 and 3 (as it cannot have an element of order 6).
Since |G| is even, there must be some element of order 2 (else G would
consist of distinct pairs of elements {x, x−1 } together with e, which makes
an odd number).
If all elements of G had order 1 or 2, then choose x, y ∈ G with x �= y,
x, y �= e. Then xyxy = (xy)2 = e = x2 y 2 = xxyy. Multiplying the equation
xyxy = xxyy on the left by x−1 and on the right by y −1 , we obtain yx = xy,
80
but then G contains two distinct commuting elements of order 2, which
contradicts Lemma 4.39 as 4 � 6.
Thus there must be x, y ∈ G with o(x) = 3, o(y) = 2; then the elements
e, x, x2 , y, yx, yx2
are all distinct, because if y i xj = y k x� then xj−� = y k−i ∈ �x� ∩ �y� = {e},
and so xj = x� and y i = y k .
We now consider the element xy; this must be equal to yx or yx2 . How-
ever, if xy = yx then o(xy) = 6 by Proposition 3.88 but we are assuming
that G is not cyclic, so this is a contradiction.
So we must have xy = yx2 . Thus we see that x and y satisfy the relations
x = y 2 = e and xy = yx−1 , which are the same as the relations in D6 . One
3
81
4.5 Normal subgroups
In Section 3.6, we defined the left and rights cosets of a subgroup H in a
group G:
gH = {gh | h ∈ H}
Hg = {hg | h ∈ H}
We also saw that in general gH �= Hg, but commented that the class of
subgroups for which this is true for all g ∈ G is important. In this section,
we will begin to see why.
82
automatically closed under conjugation by elements of the subgroup, since
subgroups are closed under taking inverses and products.)
You need not worry too much about the apparent asymmetry in (NSG2):
notice that gng −1 = (g −1 )−1 n(g −1 ). Since N is normal and g −1 ∈ G and
n ∈ N , we must have gng −1 = (g −1 )−1 n(g −1 ) ∈ N . So it is equally valid to
show gng −1 ∈ N for all g ∈ G and n ∈ N .
Examples 4.46.
(a) The trivial subgroup {e} and G itself are normal subgroups of G.
To see this for {e},
or equivalently,
(NSG2) for all g ∈ G and e ∈ {e}, g −1 eg = g −1 g = e ∈ {e}.
For G itself,
or equivalently,
(NSG2) for all g ∈ G and h ∈ G, g −1 hg ∈ G, since G is a group.
or equivalently,
(NSG2) for all g ∈ G and n ∈ N , g −1 ng = ng −1 g = ne = n ∈ N .
You should look carefully to see where the assumption that G is
Abelian is used (specifically, that every pair of elements of G com-
mute).
� �
(c) In D6 = α, β | α3 = β 2 = e, αβ = βα2 , the subgroup generated by α
is normal.
The elements of D6 are {e, α, α2 , α3 , β, αβ, α2 β} and the elements of
�α� are {e, α, α2 , α3 }. We use (NSG2) and notice that we need only
check g ∈ D6 \ �α� and n ∈ �α�, since if g ∈ �α�, g −1 ng ∈ �α� automat-
ically because �α� is a subgroup. Similarly we do not need to check
n = e, since g −1 eg = e ∈ �α� for any g.
83
So first let β ∈ D6 \ �α�. We have
β −1 αβ = β(αβ) = β(βα2 ) = β 2 α2 = α2
and
β −1 α2 β = β(α(αβ)) = β(α(βα2 )) = β 2 α4 = α.
A similar check for αβ and α2 β confirms that �α� is normal.
In fact, there is a much quicker, more general and more theoretical
way to prove this, as we will see shortly.
in S4 is not normal.
Then we saw that the left coset (1 2)K is
Exercise 4.47. Recall from Exercise 3.44 that the centre of a group G,
Proof. Recall from Theorem 3.66 that the cosets of N are the equivalence
classes for the equivalence relation RN defined in that theorem. In particu-
lar, the cosets are a partition of G.
Now since N is a subgroup of index 2, its left cosets must be eN = N
and everything else, i.e. G \ N , so that G = N ∪ (G \ N ). But its right cosets
must also be N e = N and G \ N . So we see that N has the same left and
right cosets and hence is normal, by (NSG1).
84
Examples 4.49.
(a) Recall that |D2n | = 2n and notice that |�α�| = n, so that by Lagrange’s
theorem |D2n : �α�| = 2. Hence �α� is normal, as we saw explicitly for
n = 3 in (c) above.
Ker Σ = {σ ∈ Sn | sign(σ) = 1}
85
Definition 4.53. A group G is said to be simple if it has no non-trivial
proper normal subgroups.
Examples 4.54. The cyclic groups of prime order, Cp , p prime, are simple
(in fact, by Proposition 3.59(ii), they have no non-trivial proper subgroups
at all). The alternating group A5 is also simple and is in fact the smallest
non-Abelian simple group.
Before we prove this, let us say something about what the issue is here.
It is that the formula we have written down—•(gN, hN ) = (g ∗ h)N —might
not be a well-defined function. That is, for any pair (gN, hN ) there should
exist a unique coset kN in LN such that •(gN, hN ) = kN . However, this
could depend on the choices of representatives g and h for gN and hN
respectively.
Here is an illustration of what can go wrong. Let G = S4 and let
K = �(1 2 3 4)�.
86
Consider the following cosets of K:
Then we see that if we tried to define (1 2)K • (1 3)K = ((1 2)(1 3))K,
on the one hand this would equal (1 3 2)K (from the penultimate line).
On the other hand, (1 2)K = (2 3 4)K, so we should get the same
answer if we calculated ((2 3 4)(1 3))K. But from the last line, this is equal
to K which is certainly not equal to (1 3 2)K. So we cannot define the
operation consistently: we get a different answer for the output depending
on which representative of the coset we choose, so this fails the uniqueness
requirement for a function.
The claim of the proposition is that everything is OK if N is a normal
subgroup, as we prove now.
Proof of Proposition 4.55. We need to prove that •(gN, hN ) = (g ∗ h)N is
a function. For each pair (gN, hN ) ∈ LN × LN , we certainly have a coset
(g ∗ h)N ∈ LN , so (as indicated by the counterexample) the issue lies in the
uniqueness.
Assume that g1 N = g2 N and h1 N = h2 N . We need to prove that
(g1 ∗ h1 )N = (g2 ∗ h2 )N . By Corollary 3.69, this is equivalent to saying
that we have that g1−1 ∗ g2 ∈ N and h−1 1 ∗ h2 ∈ N and want to show that
−1
(g1 ∗ h1 ) ∗ (g2 ∗ h2 ) ∈ N .
Now
as required.
Indeed, a careful examination of this proof and a little more work would
show that • is a binary operation if and only if N is a normal subgroup.
This is how the above counterexample of K = �(1 2 3 4)� in S4 was chosen:
we showed that K is not normal.
87
Definition 4.56. Let (G, ∗) be a group and let N be a normal subgroup
of G. The quotient group of G by N , denoted G/N , is the group (LN , •)
where LN is the set {gN | g ∈ G} of left cosets of N in G and • is the binary
operation gN • hN = (g ∗ h)N .
Strictly speaking, this definition contains some claims we have not veri-
fied, so we had better do so now.
gN • (hN • kN ) = gN • (h ∗ k)N
= (g ∗ (h ∗ k))N
= ((g ∗ h) ∗ k)N
= (g ∗ h)N • kN
= (gN • hN ) • kN,
as required.
There is a natural candidate for the identity element of G/N , namely
eN = N . Since for all g ∈ G we have
gN • g −1 N = (g ∗ g −1 )N = eN = (g −1 ∗ g)N = g −1 N • gN
From now on, we will use the notation G/N in preference to (LN , •), so
that when we write G/N it is understood that G/N as a set is the set of
left cosets of a normal subgroup N of G and that the binary operation is •.
88
Example 4.59. We will now look carefully at the quotient group construction
for an example we discussed earlier (Example 3.65), that of (G, ∗) = (Z, +)
and N = 4Z = {4m | m ∈ Z}. Note that N = 4Z is normal because G is
Abelian (Example 4.46(b)).
We saw that a natural choice of transversal for 4Z in Z is T4Z = {0, 1, 2, 3},
so that we can write
This gives us the set underlying the quotient group Z/4Z, so the next step
is to understand how the operation • works.
The results above tell us that
That is, when we apply the binary operation to two cosets r +4Z and s+4Z,
represented by r and s respectively, the result is the coset represented by
r + s. Let us compute this explicitly in a Cayley table for •:
• 0 + 4Z 1 + 4Z 2 + 4Z 3 + 4Z
0 + 4Z 0 + 4Z 1 + 4Z 2 + 4Z 3 + 4Z
1 + 4Z 1 + 4Z 2 + 4Z 3 + 4Z 4 + 4Z
2 + 4Z 2 + 4Z 3 + 4Z 4 + 4Z 5 + 4Z
3 + 4Z 3 + 4Z 4 + 4Z 5 + 4Z 6 + 4Z
But 4+4Z, 5+4Z and 6+4Z, while they are valid cosets, are not written
as elements of L4Z using the transversal we chose. Each of them is a coset
in L4Z and is equal to one of 0 + 4Z, 1 + 4Z, 2 + 4Z and 3 + 4Z, but we
should work out which one. (This is so that we can understand • as a binary
operation on L4Z more clearly, by rewriting our cosets to use our preferred
set of representatives, although strictly speaking it is unnecessary.)
Fortunately, we already saw how to tell when two cosets of 4Z are equal
(Example 3.70(b)): r + 4Z = s + 4Z if and only if r ≡ s mod 4. For example,
(2 + 4Z) • (3 + 4Z) = 5 + 4Z = 1 + 4Z
89
Indeed, if we look carefully, we deduce that Z/4Z is a cyclic group,
generated by 1 + 4Z. For Z/4Z has 4 elements (|L4Z | = |{0, 1, 2, 3}| = 4)
and 1 + 4Z is a generator, since it has order 4:
(r + 4Z) +4 (s + 4Z) = t + 4Z
for t ≡ r + s mod 4.
Notice that nowhere in this example did we use any special properties
of 4. Indeed the above discussion works perfectly well for any n ∈ Z. The
definition below formalises our somewhat ad-hoc construction of the integers
modulo n, and has the advantage that the integers modulo n are a group
by definition (so we do not need to prove this).
• α4 = e,
• β 2 = e, and
90
Notice that by Lagrange’s theorem, |G : N | = |G|/|N | = 8/2 = 4 �= 2 so
that we cannot simply use Proposition 4.48.
The condition (NSG2) requires that for all g ∈ G and n ∈ N , we have
g −1 ng ∈ N . Now since (gh)−1 n(gh) = h−1 (g −1 ng)h, we see that it is enough
to check (NSG2) for a generating set of G, since then any element of G is
a product of generators and if N is closed under conjugation by generators,
the aforementioned equation shows that it is closed under conjugation by
any product of generators.
In this case, we have α−1 α2 α = α2 and
eN = N = {e, α2 }
αN = {α, α3
βN = {β, βα2 = α2 β}
αβN = {αβ, α3 β}
• eN αN βN αβN
eN eN αN βN αβN
αN αN eN αβN βN
βN βN αβN eN αN
αβN αβN βN αN eN
91
identified, breaking the fourfold symmetry. Then a square whose opposite
pairs of edges are no longer “the same” is “like” having a rectangle, whose
isometry group is D4 .
There is also a more algebraic approach. For we have from Exercise 4.47
that the centre of a group is always a normal subgroup. Then we may show
that Z(D4n ) = �αn � and then argue as above to see that D4n /Z(D4n ) ∼
= D2n .
so that π is a homomorphism.
Ker π = {g ∈ G | π(g) = eN }
= {g ∈ G | gN = eN }
= {g ∈ G | g ∈ N }
= N.
16
In previous years, this has been known as the Fundamental Isomorphism Theorem
and you will see that name in previous exams.
92
Then we have shown the following.
Corollary 4.65. Let G be a group and N a normal subgroup of G. Then N
is the kernel of a group homomorphism with domain G, namely the quotient
homomorphism π : G → G/N .
Examples 4.66.
(a) Consider G = (Z, +) and N = nZ for n ∈ Z. Recall that N is
normal since G is Abelian. Then nZ is the kernel of the quotient
homomorphism π : Z → Z/nZ given by π(r) = r + nZ.
But the cosets a + nZ are precisely the sets of integers congruent to
a modulo n. So π is the map that takes an integer to its congruence
class modulo n.
This is a homomorphism, as you already know in a disguised fashion:
in the notation of MATH111 Numbers and Relations, writing � a for
a + nZ,
a�+b=� a + �b.
(On the right hand side here, it would be consistent with the above to
write +n to emphasise that this is addition modulo n in Zn .)
(b) We saw in Example 4.49(b) that the subgroup of even permutations,
An , is a normal subgroup of Sn (since it has index 2). So An is the
kernel of π : Sn → Sn /An . Since An has index 2, there are precisely
two cosets in Sn /An and, as in the proof of Proposition 4.48, Sn =
An ∪ Sn \ An . Now (1 2) is odd so (1 2) ∈
/ An and hence a transversal
for An in Sn is {ι, (1 2)}.
We see that the two cosets are precisely the even elements and the odd
elements in Sn . Then via the isomorphism sending the coset ιAn to 1
and (1 2)An to −1, we see that π in this case is essentially the same
as the sign homomorphism Σ : Sn → {1, −1}, Σ(τ ) = sign(τ ) for all
τ ∈ Sn from Example 4.7(e) and Example 4.52.
Next, we make some preparations for the First Isomorphism Theorem.
Most likely, it will not be obvious to you why we do what we are about
to, but do not worry about this. It is part of a much larger picture that
would eventually become clearer after more advanced (undergraduate and
postgraduate) study in algebra; here it can simply be seen as a technical
result that we need to be able to prove what we want.
Theorem 4.67 (Universal property of the quotient group). Let G and H
be groups. Let N be a normal subgroup of G and let π : G → G/N be the
associated quotient homomorphism.
Then for every homomorphism ϕ : G → H such that N ⊆ Ker ϕ, there
exists a unique homomorphism ϕ̄ : G/N → H such that ϕ = ϕ̄ ◦ π.
Furthermore, Ker ϕ̄ = (Ker ϕ)/N and Im ϕ̄ = Im ϕ.
93
We can illustrate the maps in this theorem via the following diagram,
known as a “commutative diagram” (the “commutative” refers to the fact
that following along the arrows in either possible way gives the same result).
ϕ
G H
π
∃!ϕ̄
G/N
The kernel of ϕ̄ is
where we remark that the quotient (Ker ϕ)/N makes sense because N being
normal in G implies that N is normal in any subgroup of G containing N ,
such as Ker ϕ in this case.
94
The image of ϕ̄ is
G/ Ker ϕ ∼
= Im ϕ
Proof. We apply Theorem 4.67 to G, Im ϕ (instead of H), ϕ|Im ϕ : G → Im ϕ
(instead of ϕ) and N = Ker ϕ, all of which satisfy the necessary assumptions.
Then from that theorem, we conclude that there exists a homomorphism
ϕ̄ : G/ Ker ϕ → Im ϕ with Ker ϕ̄ = (Ker ϕ)/ Ker ϕ = {e Ker ϕ} trivial and
Im ϕ̄ = Im ϕ. That is, ϕ̄ is an injective and surjective homomorphism, hence
an isomorphism ϕ̄ : G/ Ker ϕ → Im ϕ.
Recall from Corollary 4.58 that |G/N | = |G|/|N |, so it follows from the
First Isomorphism Theorem that |G| = |Ker ϕ||Im ϕ|. In fact, we already
knew this, and showed it directly (without the need for quotient groups) in
Corollary 4.24. So at first sight, all we have done is provided a much more
complicated proof.
It would be a mistake to think this way, though. Actually, the First
Isomorphism Theorem is a much stronger claim: it asserts that the group
structures of the groups G/ Ker ϕ and Im ϕ are the same, not just the car-
dinalities of their underlying sets. The latter is an immediate consequence of
95
the former, but we already know that there are two non-isomorphic groups
of order 4, so to say that the groups are isomorphic is a much more signific-
ant claim. This is why the proof needed is more complex: stronger claims
need more justification.
When we originally proved Corollary 4.24, we remarked that the formula
|G| = |Ker ϕ||Im ϕ| is a group-theoretic analogue of the Dimension Theorem
from Linear Algebra, dim V = dim Ker T + dim Im T for a linear transform-
ation T : V → W . With just a little more work than was done in MATH220
Linear Algebra II, one can define quotient vector spaces and prove a First
Isomorphism Theorem for Vector Spaces, which in turn implies the Dimen-
sion Theorem. The existence of bases for vector spaces simplifies matters
somewhat, and allows us to think of quotients in terms of complementary
subspaces (orthogonal subspaces, for real and complex vector spaces).
We conclude this section with some examples of applications of the First
Isomorphism Theorem. Largely, they fall into the class “Trying To Avoid
Using Quotients”.
Examples 4.69.
(g s N )r = g rs N = eN
96
ζ3
ζ4 ζ2
ζ5 ζ1
ζ6
ζ 0 = 1 = ζ 12
ζ7 ζ 11
ζ8 ζ 10
ζ9
and
Im ϕ = {(g i )s | 0 ≤ i ≤ n − 1} = {(g s )i | 0 ≤ i ≤ n − 1} = �g s � ∼
= Cr .
97
where S 1 is common notation for the circle of radius 1 centered at 0
in the complex plane. The image of ϕ is
Im ϕ = {r ∈ R∗ | (∃ z ∈ C∗ )(�z� = r} = {r ∈ R∗ | r > 0} = R+
since every non-zero positive real number is the modulus of some non-
zero complex number: �r + 0i� = r for r ∈ R+ , for example. It is also
straightforward to check directly that R+ is a subgroup of R∗ (remem-
bering that the operation being used is multiplication), although we
get this automatically from it being the image of a homomorphism.
Hence C∗ /S 1 ∼= R+ . Geometrically, this corresponds to viewing a
point in the complex plane (without the origin) as being determined
completely by an angle θ and a positive real number r: this is precisely
the polar form z = reiθ . Below we give an illustration of this.
S 1 = {z ∈ C∗ | �z� = 1}
θ = 2π/10
R+ = {r ∈ R∗ | r > 0} r=3
98
Let us construct a homomorphism ϕ : Z2 → Z such that �(3, 2)� =
Ker ϕ. (One might reasonably ask why I am “guessing” that I can
put Z as the codomain here. A heuristic would be that Z2 is “2-
dimensional”, �(3, 2)� is a line and hence “1-dimensional”, so that we
would expect the quotient to be “1-dimensional”, so look like Z. Fortu-
nately, this somewhat casual reasoning turns out to be correct. Indeed,
we see that there are |Z| cosets, indicated by the dotted lines parallel
to the line spanned by (3, 2).)
We want ϕ to respect addition in Z2 so it had better be linear in
both arguments: ϕ(a, b) = am + bn for some m, n. Since we want
ϕ(3, 2) = 0, choosing m = 2, n = −3 looks plausible.
99
We double-check that this does define a homomorphism:
100
n Isomorphism types of groups of order n
1 {e}
2 C2
3 C3
4 C4 , C2 × C2
5 C5
6 C6 , D 6
7 C7
8 C8 , C2 × C4 , C2 × C2 × C2 , D 8 , H∗
9 C9 , C3 × C3
This has been a huge endeavour for group theorists and it is now com-
monly held to have been completed. The whole classification is well beyond
our reach—it would take long enough just to define all of the groups ap-
pearing in the classification—but we can make a start.
One might legitimately ask why we single out the finite simple groups for
classification. This is because if we knew all finite simple groups and if we
knew all the possible ways to “glue” simple groups together to make bigger
groups, then we would know all finite groups. As hard as this approach may
appear, it is considerably easier than the order-by-order approach.
{e} = G0 � G1 � G2 � · · · � Gn = G
{e} � G1 � · · · � Gn−1 � Gn = G
101
with Gi a maximal proper normal subgroup of Gi+1 (normal in Gi+1 but
not necessarily in G). That is, for all i, there does not exist Ni � Gi+1 with
Gi � Ni � Gi+1 .
The factors Gi+1 /Gi are called composition factors.
{e} � C2 � C2 × C2 � A4
Once one has proved this theorem, we can talk about the composition
factors of G, rather than of a particular composition series for G. What
this means from a classification point of view is that it justifies our focus
on simple groups: the composition factors of a group are simple, and every
group is iteratively built up out of these factors.
A vector space analogue says that if V is an n-dimensional vector space
then a descending sequence of subspaces V � V � � V �� � · · · , each of
dimension exactly one less than the last, terminates in the zero subspace
and has length n + 1. The factors are all copies of the field (so here there is
one factor, up to isomorphism, but it occurs with maximum multiplicity).
102
4.8.2 Finite Abelian groups
Abelian groups are considerably easier to understand than groups in general:
here we will sketch the (full!) classification of finite Abelian groups, up to
isomorphism. By “classify”, we mean that we will state a theorem that says
that every finite Abelian group is isomorphic to one of a list of explicitly
described groups. The list will be infinite but this is not surprising—there
is a cyclic group having any order you care to name and all of these are
non-isomorphic. On the other hand, the explicit description is really very
explicit and given an Abelian group, one can find out which group on the
list it is isomorphic to relatively easily.
The strategy is the following:
(a) show that a finite Abelian group A has a Cartesian product decom-
position into subgroups having prime-power order, for different primes
dividing |A|, then
(b) show that each of these subgroups is a direct product of cyclic groups.
Hence A will be a direct product of cyclic groups. We see here a hint of how
important prime numbers are in finite group theory.
We begin with a general definition, even though we will only be interested
in the Abelian case for now.
Definition 4.73. Let G be a group and p a prime. We say G is a p-group
if the order of every element of G is a power of p. That is, for all g ∈ G
o(g) = pk for some k depending on g.
Clearly a non-trivial group G can be a p-group for at most one p but lots
of groups are not p-groups (e.g. C2 × C3 , S3 ). A very useful characterisation
of finite Abelian p-groups is the following.
Lemma 4.74. Let A be a non-trivial finite Abelian group and let p be a
prime. Then A is a p-group if and only if A has p-power order, i.e. |A| = pr
for some r ≥ 1.
Remark 4.75. This result is true for non-Abelian groups but is much harder
to prove in generality (it relies in part on a result known as Cauchy’s theorem
[ST, Theorem 3.33]).
Another name for Abelian p-groups is primary groups, and so the follow-
ing theorem is often called the Primary Decomposition Theorem for finite
Abelian groups.
Theorem 4.76. Let A be a finite Abelian group and let P = {p1 , . . . , pn } be
the set of primes occurring as orders of elements of A. Then the elements
of A of pi -power order form a subgroup A(pi ) and
A∼
= A(p1 ) × A(p2 ) × · · · × A(pn ).
103
This establishes part (a) of the strategy, the reduction to the case of
finite Abelian p-groups. Now we look for part (b).
Theorem 4.77. A finite Abelian p-group A is isomorphic to a direct product
of cyclic p-groups.
The two theorems of this section then yield the classification of finite
Abelian groups.
Theorem 4.78. Every finite Abelian group A is a direct product of cyclic
groups.
Proof. Theorem 4.76 tells us that a finite Abelian group is the direct product
of its primary components, each of which is a finite Abelian p-group (for some
p) and thus a direct product of cyclic groups by Theorem 4.77.
A∼
= C d1 × C d2 × · · · × C dt
such that 1 < di | di+1 for 1 ≤ i < t and t and the sequence (di ) are uniquely
determined by A.
This standard decomposition might not be exactly what you expected:
C2 × C3 would have the associated data t = 1, d1 = 6 (corresponding to its
isomorphic form C6 ), rather than t = 2, (d1 , d2 ) = (2, 3), since 2 � 3.
Example 4.80. If A is Abelian of order 8 then A is isomorphic to exactly
one of C8 , C2 × C4 and C2 × C2 × C2 .
Exercise 4.81. Classify the finite Abelian groups having order 100, order 128
and order 74,088.
104
(ii) pq with p �= q primes, or
By proving results like this, one obtains theorems such as the following.
Using this as the base case for an induction, one can then tackle the
inductive step and hence prove the following theorem.
One may also show that there is a unique non-Abelian simple group of
order 168, and there are none having intermediate order. This simple group
is not A6 , of course: |A6 | = 360. We will show that it is a group called
PSL(2, 7); however, the next simple group after PSL(2, 7) has order 360 and
is indeed A6 .
(a) The general linear group of degree n over F , GLn (F ), is the group of
n × n matrices with entries from F and non-zero determinant.
(b) The special linear group of degree n over F , SLn (F ), is the subgroup
of GLn (F ) of matrices of determinant one.
105
(d) The projective special linear group of degree n over F , PSL(n, F ), is
the quotient group of SLn (F ) by its centre Z(SLn (F )).
106
Lemma 4.87. Let F be a field.
Proof. The first part is linear algebra; the second follows immediately.
Lemma 4.88. Let F be a field of order q and let t be the number of roots
of the polynomial xn − 1 over F . Then t = (n, q − 1), the highest common
factor of n and q − 1. Furthermore,
�
(i) |GLn (q)| = n−1 n i
i=0 (q − q ),
These groups are more familiar to you than they might appear at first
sight.
Examples 4.89.
(a) GL1 (F ) ∼
= F ∗ = F \ {0} and so
= SL1 (F ) ∼
PGL(1, F ) ∼ = PSL(1, F ) ∼
= {e}.
(b) Z(SL2 (C)) = {I2 , −I2 }, since the only two square roots of unity in
C (or indeed, any field) are 1 and −1. In fact, one can show that
PSL(2, C) ∼= PGL(2, C). This group is known by a special name: it
is the group of Möbius transformations of the Riemann sphere (the
extended complex plane).
107
(c) Taking F = F2 to be the field of two elements (i.e. Z2 with addition
and multiplication modulo 2), we have the finite group PSL(2, 2) of
order ((22 − 1)(22 − 21 )/(2 − 1))/(1) = 6. One can easily show that
this group is not Abelian and hence, by our previous classification, is
isomorphic to D6 (∼= S3 ).
(e) Taking F = F3 , the field of three elements, |PSL(2, 3)| = 12 and one
may show that PSL(2, 3) is isomorphic to A4 .
The terminology “projective . . . ” refers to the fact that these groups act
on projective spaces. Spaces such as Cn and Rn are called affine spaces and
projective spaces are quotients of these. One major geometric reason for
studying projective space rather than affine space is that projective spaces
are compact topological spaces and are thus better behaved. The general
definition of projective n-space is as follows.
108
(0, 1, 0)
(0, 1, 1) (1, 1, 0)
(1, 1, 1)
As noted at the end of the previous section, the list of non-Abelian finite
simple groups begins A5 , PSL(2, 7), A6 (where we prefer to label the first
and third of these by their alternating group name, rather than (one of)
their projective special linear group name(s)). Their orders are 60, 168 and
360 respectively. We know that the alternating groups of degree at least
5 continue to contribute to this sequence, but what about the projective
special linear groups? Do these stop being simple at some point, or are
they all alternating groups? Well, no and no (the latter by looking at their
orders).
Theorem 4.92. For any n ≥ 2 and q a prime power, the group PSL(n, q) is
simple, of order given by the formula in Lemma 4.88, unless (n, q) = (2, 2)
or (2, 3).
We remark that the proof does not proceed in the way we did for
PSL(2, 7). Instead, one identifies certain elements with nice properties and
uses linear algebra and geometrical arguments.
109
The projective special linear groups are part of a family called the groups
of Lie type and the finite simple members of this family make up the third
and final infinite class of finite simple groups. Groups of Lie type are matrix
groups over finite fields (including orthogonal and unitary groups, among
others) or “twisted” versions of these. We conclude this discussion by looking
at the result we have been aiming towards, the classification of finite simple
groups.
110
• the O’Nan group, O� N;
The group M11 is the smallest sporadic simple group, of order 7920, and the
group M, the largest of the sporadic simple groups, has order
The finite simple groups still have much to tell us, particularly with
regard to their representation theory. The study of finite simple groups is
by no means over, and certainly the study of finite groups is not. However,
our study of them is at an end.
For further reading, a good starting place is [Neu]. Recent books on the
subject include [Ron] and [dS], both “popular” books but with the “profes-
sional” reader in mind, also. Those interested in the developmental history
of modern abstract algebra more generally (as sketched in the introduction
to these notes), i.e. encompassing fields, rings, algebraic geometry and more,
should consider [Der].
Now, however, we turn our attention to rings.
111