Graham D. Williams
Graham D. Williams
Williams
Notes
Contents
2 Groups 43
2.1 Operations and Monoids . . . . . . . . . . . . . . . . . . 43
2.2 Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
2.3 Subgroups, Cyclic Groups and
Lagrange’s Theorem . . . . . . . . . . . . . . . . . . . . 66
2.4 Homomorphisms and Isomorphisms . . . . . . . . . . . . 85
2.5 Quotient Groups . . . . . . . . . . . . . . . . . . . . . . 102
2.6 More on Permutations . . . . . . . . . . . . . . . . . . . 121
2.7 Examples: Symmetries and Matrices . . . . . . . . . . . 140
2.7.1 Symmetries . . . . . . . . . . . . . . . . . . . . . 140
2.7.2 Matrix Groups . . . . . . . . . . . . . . . . . . . 146
1
2 CONTENTS
Chapter 1
1.1 Sets
3
4 CHAPTER 1. SETS, RELATIONS AND FUNCTIONS
A and B are counted as equal if they have precisely the same elements:
We have used here the shorthand symbol ⇔, which is short for “if
and only if”. It means that the two statements on either side of it are
equivalent. If either one of them is true, then the other must follow as a
logical consequence. We shall have more to say about logical symbols like
this in the future.
The elements of a set don’t have to be all of the same “type”:
√
C = { 2, Albert Einstein, the planet Mercury} is a perfectly good set. If
a set is large, in particular infinite, we cannot list all the elements explic-
itly. We might write A = {1, 2, . . . , 1000} to denote the set of all integers
from 1 to 1000, although strictly this is imprecise, since the reader is be-
ing asked to guess the pattern involved. A few sets occur so frequently
throughout Mathematics that we reserve special symbols for them:
3
√ √
Thus 4 ∈ Q, 43 ∈
/ Z, 2 ∈ R, 2 ∈
/ Q.
belief among non-experts is that there are lots of empty sets: one that
is empty of apples, one that is empty of bananas, etc. This is not so.
Suppose ∅a and ∅b were two potentially different empty sets. We claim
that they are in fact the same. If this were not so, then by the criterion
above for two sets to be equal, one of them, say ∅a , would contain an
element x not belonging to the other. But this is impossible, since ∅a
hasn’t got any elements. That’s called logic!
The next important idea is that of a subset. Suppose we have two
sets A and B and that B contains all the elements of A and maybe some
more. Then we say that A is a subset of B and write A ⊂ B. This can
also be written B ⊃ A. Formally:
We have used here another logical symbol ⇒. This is short for “im-
plies”. Thus if P and Q are two statements or propositions (which might
or might not be true), then P ⇒ Q is the proposition “P implies Q” or
“if P then Q”. This is not to say that either statement necessarily is true,
merely that if P holds, then Q automatically follows. We can also write
this as Q ⇐ P . Note also that P ⇔ Q amounts to saying that P ⇒ Q
and Q ⇒ P simultaneously.
So to prove that A ⊂ B we have to demonstrate that whenever an
element belongs to A then it also belongs to B. From 1.1.1 and 1.1.2 we
have at once:
1.1.3. A = B ⇔ A ⊂ B and B ⊂ A .
1.1.4. A ⊂ A and ∅ ⊂ A .
We remind the reader that if P, Q are two propositions, then the com-
pound proposition “P or Q” is counted as true as soon as at least one of
the component parts is true, whereas “P and Q” is true only if both parts
8 CHAPTER 1. SETS, RELATIONS AND FUNCTIONS
The above may seem like a lot of tedious nit-picking to establish results
which are in most cases intuitively obvious, especially if you draw Venn
diagrams, but we have included it to give you a chance to start learning
how to go about proving things formally. (i) and (ii) tell us that ∪ and ∩
are commutative processes, and (iii) and (iv) say that they are associative.
10 CHAPTER 1. SETS, RELATIONS AND FUNCTIONS
We round off this section by introducing one more idea, that of the
complement of a set.
Exercises:
(a) {1, 2, 3} (b) {1, 3, 3} (c) {{1, 3}} (d) {1, 3, {3}} (e) N (f ) {N}
(g) {1, N} (h) {N, Z} (i) ∅ (j) {∅} (k) {{∅}} (l) {∅, {∅}}.
(a) {1, 2} (b) {a, b, c} (c) ∅ (d) {∅} (e) {∅, ∅}.
1.2 Relations
Another way to make a new set out of two given sets A and B is to form
their Cartesian product A × B (pronounced “A cross B”). The elements
12 CHAPTER 1. SETS, RELATIONS AND FUNCTIONS
of this are not elements of A or B, but new objects (a, b) called ordered
pairs, where a ∈ A and b ∈ B. We will take (a, b) to be a primitive
concept, although it is possible to give a formal definition of it in terms of
sets. This is, however, artificial and unilluminating, so we won’t do it. In
practice, you will be quite familiar with the idea from ordinary coordinates
of points in the plane, although this is now much more general, since A
and B need not be subsets of R. The point to stress is that the order
matters. For example, if A = B = R, then (1, 2) and (2, 1) are different
ordered pairs. So, the formal definition is this:
(i) R = {(1, 1), (2, 2)} ; (ii) S = {(1, −1), (3, 0), (5, 1)} ;
(iii) T = {(1, −1), (1, 0), (1, 1), (2, 0)} ;
(iv) U = {(2, −2), (4, −2), (6, −2), (1, −1), (2, −1), (3, −1), (4, −1), (5, −1),
(6, −1), (1, 1), (2, 1), (3, 1), (4, 1), (5, 1), (6, 1), (2, 2), (4, 2), (6, 2)}.
Certainly all the above examples are of a “mathematical” type, but we
don’t want to limit our options in advance by specifying that only certain
types of relationship are mathematically “significant”. So we go for broke
and make the following:
that a is related to b (via R), or R-related to b. Two extreme case are the
total relation R = A×B in which everything in A is related to everything in
B, and the empty relation ∅ ⊂ A×B where nothing is related to anything.
If A = B, then a relation R ⊂ A × A is often just called a relation on
A. For example, the set {(x, y) ∈ Z × Z : y 2 = x3 + 17} is a relation
on Z. Actually, by very advanced means(theory of elliptic curves and
algebraic number fields) this last set can be shown to consist of the pairs
(−2, ±3), (−1, ±4), (2, ±5), (4, ±9), (8, ±23), (43, ±282), (52, ±375),
(5234, ±378661), but we won’t go into that here!
If we concentrate for a moment on the familiar set R, then a relation
on R is just a subset of the plane R2 = R × R, which we can visualise.
For instance,the relation given by the equation x2 +y 2 = 1 is the unit circle
around the origin, ie. the graph of the equation. Note that according to
the definition the relation is the graph.
An important (but dull) relation on a set A is that of equality. This is
given by the so-called diagonal subset ∆ of A × A, namely ∆ = {(a, a) :
a ∈ A}. Here a∆b ⇔ a = b. In the case A = R the corresponding
picture (the graph) is the straight line y = x.
For the rest of this section we will restrict our attention to relations on
a single set A (ie. to the case A = B). Many naturally arising relations on
A have one or more especially nice properties, and we want to concentrate
on a few of these now:
We emphasise at once that these are very special properties and most
subsets of A×A will not satisfy any of them. Let us spell out the meaning
of each of these a bit more. To say that R is reflexive means that at the
very least it must contain the diagonal set ∆ above as a subset. Put
another way, we require that every element of A is related to itself. If R
is symmetric, then whenever a is related to b we must also have that b is
related to a. In the case A = R, the corresponding picture (the graph)
of a reflexive relation must contain the diagonal line y = x, and that of a
symmetric relation must have mirror symmetry about this line. Transitivity
is harder to visualise directly, but the requirement for this sort of relation
is that whenever a is related to b and b is related to c, then we must also
have that a is related to c.
1.2.5 Examples. To get used to the ideas here are a few examples of
relations on the set A = {1, 2, 3, 4}:
The first is not symmetric, since although it contains (1, 2) it does not
contain (2, 1). The other three don’t have this asymmetry. The relation
S is not transitive, since although 1S2 and 2S3, we don’t have 1S3. Like-
wise 3U 4 and 4U 3, yet we don’t have 3U 3, so U is not transitive. The
16 CHAPTER 1. SETS, RELATIONS AND FUNCTIONS
fact that the other two relations are transitive is a rather tedious check
through all the possibilities, and we leave you todo this. You have to check
that whenever the relation contains two pairs such that the second coor-
dinate of the first equals the first coordinate of the second, then the pair
consisting of the first coordinate of the first pair followed by the second
coordinate of the second is also present! For example, since T contains
(2, 3) and (3, 2) it had better contain (2, 2) as well, and of course it does.
Note that these two pairs also fit together the other way round, so T needs
to contain (3, 3) too in order to have a chance of being transitive. It is
important to carry out all of the required checks for transitivity, not just
some of them. However, you can save some time by noticing that you
need never check a couple of pairs of the form (a, a) and (a, b). For (a, b)
is present in the relation already. Likewise in the case (a, b) and (b, b). So
R is transitive for rather trivial reasons.
Note that on any set the total relation A × A and the equality (di-
agonal) relation ∆ have all three properties above. Here are a few more
examples:
We will now concentrate on the important and very nice class of rela-
tions having all three properties above:
The total and equality relations on any set are equivalence relations,
as is Example 1.2.6(iv). So also is the relation “is parallel to” on the set
of all straight lines in the plane. We now investigate the behaviour of such
relations:
Proof. Suppose xRy. To prove that [x] = [y] we must show that whenever
an element z belongs to [x] then it also belongs to [y], and vice-versa. So
let z ∈ [x]. By definition we have xRz, and we also have yRx since R is
assumed symmetric. But now transitivity of R gives yRz, so that z ∈ [y].
So far we have shown that z ∈ [x] ⇒ z ∈ [y], and so [x] ⊂ [y]. For
18 CHAPTER 1. SETS, RELATIONS AND FUNCTIONS
Note that in proving the above, we have made use of all three proper-
ties in the definition of an equivalence relation. The upshot is that if we
have an equivalence relation R on a set A, then the various distinct equiv-
alence classes partition up A into a number of disjoint subsets. All the
elements in a given subset are equivalent to each other (under R), and el-
ements in different subsets are inequivalent. Here are two stupid examples:
(i) If R is the relation “=”, then each equivalence class has just one
element: [x] = x.
(ii) If R is the total relation (in which everything is related to everything
else) there is just one equivalence class, namely A.
In other words, if two elements are related to each other both ways
round, they are forced to be equal. Only very boring relations are sym-
metric and antisymmetric at the same time. As an exercise, show that if
this is the case, then in fact R ⊂ ∆.
Observe that in Examples 1.2.5 R is antisymmetric, but the other three
are not. Consider also Examples 1.2.6. Of these, (i) is clearly antisymmet-
ric: it is a basic property of the real numbers that if a ≤ b and b ≤ a, then
a = b. So also is (ii), though this might at first seem a little confusing.
We have to show that any time we simultaneously have a < b and b < a,
then in fact a = b. But the first pair of conditions can never occur, so
20 CHAPTER 1. SETS, RELATIONS AND FUNCTIONS
the check can never go wrong! Put another way, if this implication were
false, there would have to exist two numbers simultaneously satisfying the
conditions a < b, b < a, a 6= b, which is impossible. 1.2.6(iii) is also
antisymmetric, since if a and b are two positive integers and each is a
multiple of the other, then a = b. Examples (iv), (v) and (vi) are not
antisymmetric. Motivated by Example 1.2.6(i) we make the following:
1.2.12 Definition. A relation R on a set A is a partial order relation or
partial ordering if it is reflexive, antisymmetric and transitive.
1.2.13 Examples. (i) The relation ≤ on |R is a partial order; this is Ex-
ample 1.2.6(i) again:
(ii) The relation “a is a multiple of b on N is a partial order:
(iii) If A = {1, 2, 3, 4, 5}, then R = {(1, 1), (2, 2), (3, 3), (4, 4), (5, 5), (1, 2),
(1, 3), (1, 4), (1, 5), (3, 4), (3, 5)} is a partial order on A, as a little tedious
checking shows.
Motivated by the first of these examples, we often use the symbol ≤,
rather than R, for some partial ordering on a general set A, and may write
a ≤ b rather than aRb Of course, if we refer to such a partial order on a
set A, we are not implying that a and b are real numbers or that ≤ means
an ordinary inequality between numbers, but simply that it is a relation
satisfying the following properties:
Exercises:
1.2. RELATIONS 21
9. The following are some relations on the set A = {0, 1, 2, 3, 4}. Fill
out the table with ticks and crosses to indicate whether each relation is
reflexive, symmetric or transitive:
10. Which of the previous relations are equivalence relations? For any
that is, write down the equivalence classes.
11. Here are some more relations on various sets. Fill out the table
with ticks and crosses to indicate which of the properties applies to each:
1.2. RELATIONS 23
12. Are any of the relations of the previous question equivalence rela-
tions? If so, what are the equivalence classes?
1.3 Functions
In this section we return to the set-up of a relation F ⊂ A × B between
any two sets A and B, and we suppose now that F has a very special
property. Namely, suppose that whatever element a ∈ A we start with,
there is exactly one pair (no more and no less) in F beginning with a, ie.
whose first coordinate is a. Then we call F a function from A to B. Let
us repeat this definition formally and extend it:
use small letters such as f, g, h for functions. To get used to the idea, here
are a few examples of relations from A = {1, 2, 3, 4, 5} to B = {0, 2, 4, 6}.
Some are functions, some are not:
1.3.2 Examples.
Relation Function?
f : (1, 0), (2, 4), (3, 6), (4, 0), (5, 2) X
g : (1, 6), (2, 4), (3, 0), (4, 0), (5, 4) X
h : (1, 2), (2, 2), (3, 2), (4, 0), (5, 0) X
R : (1, 4), (2, 6), (4, 0), (5, 2) ×
S : (1, 0), (2, 0), (2, 4), (3, 4), (4, 6), (5, 2) ×
f(1)=0,f(2)=4,f(3)=6,f(4)=0,f(5)=2
g(1)=6,g(2)=4,g(3)=0,g(4)=0,g(5)=4
h(1)=2,h(2)=2,h(3)=2,h(4)=0,h(5)=0.
In the case of functions, we use the notation f : A → B, rather than
f ⊂ A×B, to refer to a function f from A to B. Occasionally we write the
f
f over the arrow, instead of at the front, thus: A → B. Although logically,
according to our definition, f is a subset of A×B, psychologically we don’t
26 CHAPTER 1. SETS, RELATIONS AND FUNCTIONS
This means that whatever input real number we start with, the corre-
sponding output is x2 . Thus f (2) = 4, f (0) = 0 and f (−3) = 9. The
corresponding subset of R × R, ie. the graph of f (which, logically, is just
f ) is the parabola with equation y = x2 . It is important to realise that
the symbol x used for the variable in defining the function is a dummy
variable, and can be replaced by any other letter. We could just as well
define the function by specifying that f (y) = y 2 or f (s) = s2 , but not
f (p) = q 2 !
1.3. FUNCTIONS 27
Here, the second arrow tells us what to send the general element x to.
Note the special form of this arrow; we use this type of arrow between
elements and an ordinary arrow between sets. We can pronounce it as
“x goes to x2 ”. Here are a few more examples of functions defined by
various rules:
1.3.3 Examples.
g : R → R , x 7→ 3x3 − 1
(
1 (if x ∈ Q)
h : R → R , h(x) =
0 (if x ∈
/ Q)
k : R → R , x 7→ x−1
√
l : R → R , x 7→ x
√
γ : R+ → R , x 7→ ± x , where R+ = {x ∈ R : x ≥ 0}
δ : Z → N , x 7→ |x| .
We recall in this last example that |x| is the absolute value or modulus of
a real number x, defined by:
(
x (if x ≥ 0)
|x| = .
−x (if x < 0)
Thus |3| = 3 , | − 7| = 7 .
It is also crucial to realise that specifying the sets A and B is part and
parcel of defining a function f : A → B. Let us define formally what we
mean by two functions being equal:
Let us take our time in getting used to these ideas. Firstly injective
functions. In the older literature these are also known as one-one functions.
The requirement is that the only way two elements x and y can have the
same image (ie. map to the same value) is when, in fact, they were the
same element to start with. Put another way, distinct elements map to
distinct images: x 6= y ⇒ f (x) 6= f (y). Thus the function f : R →
R , x 7→ x2 is not injective, since, for example, f (1) = f (−1). If we
return to Examples 1.3.3 we see that none of h, α, β are injective, since we
have, for instance, h(1) = h(2), α(1) = α(1.5) and β(1) = β(4). However
g is injective, since 3x3 − 1 = 3y 3 − 1 ⇒ 3x3 = 3y 3 ⇒ x3 = y 3 ⇒ x = y.
In the last implication we have used an elementary fact about cube-roots of
real numbers, which you may care to prove. Here are some more examples:
1.3.8 Examples. Let us take A = {1, 2, 3, 4, 5} and B = {0, 2, 4, 6, 8}:
(i) f : A → B given by 1 7→ 0, 2 7→ 6, 3 7→ 4, 4 7→ 2, 5 7→ 8
(ii) g : A → B given by 1 7→ 0, 2 7→ 6, 3 7→ 0, 4 7→ 2, 5 7→ 8
(iii) h : A → B given by 1 7→ 2, 2 7→ 2, 3 7→ 2, 4 7→ 2, 5 7→ 2 .
Clearly f is injective, but the other two are not, since, for example,
g(1) = g(3) and h(1) = h(2).
Next let us consider surjective functions (also known as onto func-
tions). The requirement here is that every element of the codomain B
should be hit, ie. should appear as a value of the function. Putting it
formally:
1.3.19 Example. (i) We have seen that the function f of Examples 1.3.8
is bijective. Its inverse f −1 : B → A is given by 0 7→ 1, 6 7→ 2, 4 7→
3, 2 7→ 4, 8 7→ 5.
(ii) We have also seen that the function g : R → R , x 7→ 3x3 − 1 of
Examples 1.3.3 is bijective, and we q
effectively computed its inverse earlier.
It is given by g −1 : R → R , x 7→ 3 13 (x + 1).
1.3. FUNCTIONS 37
To reverse all this we must undo each of these processes in the opposite
order :
plus 1 divide 3 cube−root
R −→ R −→ R −→ R
1
q .
3 1
x 7−→ x + 1 7−→ 3 (x + 1) 7−→ 3 (x + 1)
We conclude this section by looking at two important constructions of
subsets associated with a function:
Exercises:
√
1. Let A = {1, 2, 3, 4, 5, 6} and {−1, 0, 2, π, e}. Indicate by a tick
or cross whether each of the following is a function from A to B:
Function from A to B?
f : (1, 0), (2, −1), (3, −1), (4, e), (5, π), (6, e)
√
g : (1, π), (2, e), (3, −1), (2, 2), (5, 0), (4, e), (6, 0)
h : (1, π), (2, π), (3, −1), (6, 0), (5, e)
√
k : (1, e), (2, e), (4, 0), (5, −1), (3, π), (6, 2)
√
l : (1, 0), (2, 2), (4, e), (3, π), (6, −1), (5, 2)
Function?
f : Z → Z , n 7→ n2 − 1
g : N → N , n 7→ n2 − 1
h : R 7→ R , x 7→ (x2 + 1)−1/2
α : R 7→ R , x 7→ (x2 − 1)−1/2
β : N → N , n 7→ 1/n
γ : R → Q , γ(x) = the least rational number > x
12. Which of the following functions are bijective? For each one that
is, write down the inverse function:
(a) f : Z → Z , x 7→ −x (b) g : Z → Z , x 7→ 2x
(c) h : R → R , x 7→ 2x (d) α : R → R , x 7→ x2
(e) β : R+ → R+ , x 7→ x2 (f ) γ : R∗ → R∗ , x 7→ x−1
(g) p : R → R , x 7→ 2x5 + 7.
13. Consider a function f : A → B and subsets X, Y ⊂ A. Prove:
(a) f (X ∪ Y ) = f (X) ∪ f (Y )
(b) f (X ∩ Y ) ⊂ f (X) ∪ f (Y ).
Give a concrete example to show that in (b) the two sides may not be
42 CHAPTER 1. SETS, RELATIONS AND FUNCTIONS
equal.
(c) Can you think of a property of the function f which would be enough
to guarantee that we do get equality in (b)?
Groups
43
44 CHAPTER 2. GROUPS
2.1.2 Examples.
(i) R × R → R, (a, b) 7→ a + b (addition)
(ii) R × R → R, (a, b) 7→ a − b (subtraction)
(iii) R × R → R, (a, b) 7→ ab (multiplication).
We can get more examples from these three by replacing R with Z, Q or
C.
(iv) N × N → N, (m, n) 7→ mn (exponentiation)
(v) M × M → M, (A, B) 7→ AB (matrix multiplication), where here
2.1. OPERATIONS AND MONOIDS 45
M stands for the set of all (2×2) matrices with real number entries. More
precisely, the standard symbol for this set is M2 (R). (If you don’t know
about matrices yet, then ignore this example for the moment.)
∗ a b c
a a a b
b a c a
c b c c
For any elements x, y, to work out x ∗ y we read off the element where
the row labelled x meets the column labelled y.
This is just as well: aspects of everyday life such as adding up shopping bills
would be pretty impossible otherwise! Subtraction, however, has neither.
For instance 3 − 5 6= 5 − 3 and (3 − 4) − 5 6= 3 − (4 − 5).
Example (iv) also has neither property (Exercise: give concrete coun-
terexamples). As for (v), it is shown in basic methods courses that matrix
multiplication is associative: (AB)C = A(BC), but not commutative: in
general AB 6= BA.
Example 2.1.3 also has neither property. For example b∗c = a, whereas
c ∗ b = c, so that ∗ is not commutative. It is also not associative, since
(a ∗ b) ∗ b = a ∗ b = a, whereas a ∗ (b ∗ b) = a ∗ c = b.
((x∗y)∗z)∗t, (x∗y)∗(z ∗t), (x∗(y ∗z))∗t, x∗((y ∗z)∗t), x∗(y ∗(z ∗t)).
2.1. OPERATIONS AND MONOIDS 47
In general these could all be different. But if ∗ is associative, they are all
the same. For instance, if we temporarily put a = x∗y, the first expression
equals (a ∗ z) ∗ t = a ∗ 9z ∗ t0 and this is the second expression. I leave
you to check that the other three are also equal to this. For this reason,
we can extend the convention above and simply write x ∗ y ∗ z ∗ t, so long
as ∗ is associative. Although we will not stop to do this now, one can, in
fact, formulate and prove a generalized associativity law which allows us,
for an associative operation, unambiguously to write arbitrarily long ex-
pressions x1 ∗x2 ∗. . .∗xn , and we shall do this without further comment.
There is no identity element for subtraction. For suppose there were, say
e. Then the requirement x − e = x implies that e = 0. But then we would
also have x = e − x = 0 − x = −x, which is not true if x 6= 0.
48 CHAPTER 2. GROUPS
!
1 0
Matrix multiplication has the well-known identity I = .
0 1
As for Example 2.1.3, we see at once that there is no identity. Indeed,
if there were an identity e, then the column labelled e in the table would
have to be the same as the column down the left-hand border, and this
does not occur.
We note next that an operation cannot have more than one identity
element:
Strictly speaking, the specification of the law ∗ is part of the data of the
monoid, which ought therefore to be denoted as the ordered pair (M, ∗)
consisting of both the set M and the operation ∗ on it. If we wished to
make explicit reference to the identity element, we might also denote the
monoid as (M, ∗, e). In practice we usually just denote the monoid by the
underlying set M , provided the operation has been specified. If we need,
2.1. OPERATIONS AND MONOIDS 49
∗ e a b
e e a b
a a b a
b b a b
The first row and column show that e is the identity. This leaves us with
checking associativity. If we now simply write x ∗ y as xy, this amounts to
checking the equation (xy)z = x(yz) for all possible choices of x, y and
z. Since there are three choices for each, thee are in principle 33 = 27
checks to be done. However, if any one of x, y or z equals e the equation
50 CHAPTER 2. GROUPS
(v) The set M = {e, a, b} also becomes a monoid under the operation •
given by the table:
• e a b
e e a b
a a b e
b b e a
The checks are similar, and I leave them to you. Of course, all this checking
is somewhat tedious. We shall see in due course that for many naturally
arising monoids the associativity checks can be cut down to a minimum,
or even avoided altogether.
2.1. OPERATIONS AND MONOIDS 51
Proof. y = y ∗ e = y ∗ (x ∗ z) = (y ∗ x) ∗ z = e ∗ z = z.
Exercises:
a b c d
a d c a b
b c b b a
c a b c d
d b a d a
2.1. OPERATIONS AND MONOIDS 53
e a b e a b e a b
e e a b e e a b e e a b
(a) (b) (c)
a a a a a a a a a a a a
b b b b b b b a b b a b
e a b e a b
e e a b e e a b
(d) (e)
a a a b a a e b
b b b a b b b e
In each case it is clear that e is an identity element. Fill out the fol-
lowing table with ticks and crosses to indicate whether each of the eight
strings aaa, aab, . . . , bbb satisfies the associativity condition, and hence
determine which of the above are monoids:
2.2 Groups
We are at last ready to define and begin our study of one of the most
fundamental structures in Algebra, and indeed Mathematics as a whole,
namely a group. Groups are ubiquitous throughout Mathematics and are
also crucial to our understanding of many other branches of science such as
quantum physics, relativity, crystallography, chemistry and particle physics.
1 −1
1 1 −1
−1 −1 1
(iv) The set M = {e, a, b} becomes a group under the operation • given
by the table:
• e a b
e e a b
a a b e
b b e a
e a b c
e e a b c
a a b c e
b b c e a
c c e a b
e
e e
In Examples 2.2.2 above, groups (i) and (ii) are infinite, (iii) has order
2, (iv) has order 3, (v) has order 4 and the trivial group (vi) is of order
1. In fact we have already used the notation |S| more generally for the
number of elements in any finite set.
The word “abelian” comes from the name of Niels Hendrik Abel (1802-
1829), the great Norwegian mathematician whose ideas, among others, led
to the modern concept of a group. It is a mark of real fame in Mathematics
when your name becomes an adjective in small letters! All of Examples
2.2. GROUPS 57
2.2.2 are abelian groups, but we shall soon meet many that are not. From
the point of view of the multiplication table (at least for a finite group),
the commutative condition amounts to saying that the table has mirror
symmetry in the main diagonal, eg:
e a b c
e a b c
a a c e
b b c a
c c e a
SX × SX → SX , (α, β) 7→ α ◦ β .
2.2.5 Theorem. Let X be a set. Then the set SX of all bijections from
X to X is a group under the operation of composition of functions:
SX × SX → SX , (α, β) 7→ α ◦ β .
i α β ρ σ τ
i i α β ρ σ τ
α α β i σ τ ρ
β β i α τ ρ σ
ρ ρ τ σ i β α
σ σ ρ τ α i β
τ τ σ ρ β α i
Proof. (i) Suppose ac = bc. Multiplying on the right by c−1 gives acc−1 =
bcc−1 . But acc−1 = ae = a, and likewise on the right. Hence a = b. The
proof of (ii) is similar.
Proof. (i) The previous proposition shows that r is injective. Now take
any b ∈ G. Then r(bc−1 ) = bc−1 c = be = b, proving that r is also
surjective. Part (ii) is similar.
i α β ρ σ τ
i i α β ρ σ τ
α α β i σ y x
β β
ρ ρ
σ σ
τ τ
where x, y and all the rest of the table are as yet undetermined. Then
x must be different from all the other elements in the second row and
last column, and this forces that x = ρ. Now that we know this, we can
further deduce that y = τ .
Returning to the discussion of 2.1.5, we remarked there that in a
monoid M (or indeed any set with an associative multiplicative opera-
tion) we may unambiguously write arbitrarily long expressions x1 x2 . . . xn ,
since no matter how sets of brackets are inserted to clarify the order of
operations the result will always be the same. We will not interrupt the
flow of our account to prove this rather tricky technical result known as
generalized associativity. One needs to use the method of mathematical
induction to establish it, and we will consign it to the Exercises. Granted
that, if x ∈ M and n ≥ 1 it is meaningful to consider the element xx . . . x
(n times). Just as in elementary algebra, we denote this by xn . We also
make the convention that x0 = e (the identity of M ). In the case of a
group we may even talk about negative powers:
2.2. GROUPS 63
where we have several times used the special case above. The remaining
cases m ≤ 0, n ≥ 0 and m ≤ 0, n ≤ 0 are similar and left to you.
Those who consider the shuffling argument of the first three lines above
to be a little loose may give a more formal proof using the method of in-
duction.
Z Note that the last Proposition applies in particular to all pairs of el-
ements in an abelian group. But if x and y do not commute, the rule
above is in general quite false. As an exercise, find a counterexample in
the symmetric group S3 .
groups). This is so, for example, for the group (R, +, 0). In such a case we
write nx rather than xn . Thus when n ≥ 1 we have nx = x + x + . . . + x
(n times), and as we saw earlier the inverse of x is written −x.
Exercises:
2. Let G be a group with the property that x2 = e (∀x ∈ G). Prove that
G is abelian.
! !
1 2 3 4 5 1 2 3 4 5
3. For the elements α = and β =
2 3 1 5 4 1 3 4 5 2
of the symmetric group S5 , calculate α , β , αβ, βα, α and β −1 .
2 2 −1
p q r s t p q r s t
p p q r
q r q r
(a) (b)
r r
s s s
t t t
(a) Explain why the first table cannot be completed in any way so as
to define a group.
(b) You are given that the second table can be completed (in exactly one
way) to define a group. Fill out the rest of the entries to give the complete
group table. (Hint: Think what the identity must be first. This enables
you to complete some of the table straight away. Then consider the penul-
timate entry down column 1, bearing in mind Remark 2.2.12, and carry
on like that.)
We see now that conditions (i), (ii) and (iii) above are precisely what
are needed in order that the subset H should be a group in its own right
with respect to the same multiplication operation.
68 CHAPTER 2. GROUPS
2.3.5 Examples. (i) In the group (R, +, 0), the cyclic subgroup h1i equals
Z. For that matter, Z is also generated by −1: Z = h1i = h−1i.
h2i is the set of even integers, and h 21 i is the set of all integers and half-
integers.
(ii) In the group (R∗ , ×, 1) we have h2i = {. . . , 81 , 14 , 21 , 1, 2, 4, 8, . . .}. This
subgroup can also be generated by 21 .
(iii) Consider again the group G = {e, a, b, c} of Example 2.2.2(v). We
compute the powers of a:
Note that the identity element is the only element of order one.
k = qn + r, 0 ≤ r ≤ n − 1.
2.3.11 Examples. (i) All elements of (R, +, 0) apart from 0 have infinite
order.
(ii) The element −1 in (R∗ , ×, 1) has order 2. All other elements apart
from 1 are of infinite order.
(iii) For the group G = {e, a, b, c} of Examples 2.2.2(v) we see that a and
c have order 4, but b has order 2.
(iv) In the group M = {e, a, b} of Examples 2.2.2(iv), the elements a and
b both have order 3.
We are nearly ready to state our first big theorem. First a definition:
H = {e, b}. Then aH = {a, c}, bH = {e, b} and cH = {c, a}. Note that
aH = cH and eH = bH. Since G is abelian, the right cosets are just the
same.
(iv) Consider the subgroup H = {i, ρ} of S3 = {i, α, β, ρ, σ, τ } (see
2.2.9). Then iH = H, αH = {α, σ}, βH = {β, τ }, ρH = {ρ, i}, σH =
{σ, α} and τ H = {τ, β}. Thus there are precisely three distinct left
cosets: H = {i, ρ}, {α, σ} and {β, τ }. As for the right cosets, we calcu-
late in the same way that there are three of these also: H = {i, ρ}, {α, τ }
and {β, σ}. But they are not the same as the left cosets.
Next we note that the left and right cosets of H all have the “same
size”:
2.3.18 Remark. All the above works just as well for right cosets. This
time we would define an equivalence relation by x ∼ y ⇔ y = hx (for
some h ∈ H), or equivalently yx−1 ∈ H.
We can now prove one of the most important results in group theory:
Z Lagrange’s Theorem does not say that there necessarily exists a sub-
group of G of every possible order dividing |G|. It simply limits the possi-
bilities. For example, as we shall later see, there is a group of order 12 (the
alternating group A4 , a certain subgroup of S4 ) which has no subgroup of
order 6. Moreover, a group may have many different subgroups of a given
order. For instance, the symmetric group S3 has three subgroups of order
2.
The next result shows that groups of prime order are very simple:
2.3.23 Proposition. Let G be a cyclic group. Then all its subgroups are
cyclic.
Proof. (i) follows at once from 2.3.24. Now Zm+Zn contains 1m+0n =
m, and so m ∈ Zd. But this just means d|m, and similarly d|n. Since
d ∈ Zd = Zm + Zn, we have d = am + bn, for some integers a, b, and
we have proved (ii) and (iv). Part (iii) follows from (iv), since e|am + bn.
Finally (ii) and (iii) say that d is the greatest common divisor of m and
n.
Thus 4 and 15 are coprime, but 4 and 6 are not. The case d = 1 of
the above gives at once:
2.3. SUBGROUPS, CYCLIC GROUPS ANDLAGRANGE’S THEOREM79
Order n Generators
2 x
3 x, x2
4 x, x3
5 x, x2 , x3 , x4
6 x, x5
7 x, x2 , x3 , x4 , x5 , x6
8 x, x3 , x5 , x7
Proof :(i) We already know that the subgroups all have the stated form.
Now suppose hxm i = hxn i, for some m, n ≥ 0. Then xm = xan and
xn = xbm , with a, b ∈ Z. As we are in the infinite case, m = an and
n = bm. Either m = n = 0 or both are positive, and then again m = n,
each being a divisor of the other.
(ii) Let d be a positive divisor of n and consider the element y = xn/d . Cer-
tainly y d = xn = e. On the other hand, suppose y c = e. Then xnc/d = e,
so by 2.3.12 nc/d is a multiple of n, which in turn means that c/d ∈ Z,
or again that d|c. This establishes that y has order d. Conversely, let H
be any subgroup of order d. As H is cyclic, we may write H = hxm i.
Then xmd = e, and 2.3.12 again tells us that md = an, for some integer
a. Thus m = a(n/d) and so xm ∈ hxn/d i. As xm generates H, it follows
that H ⊂ hxn/d i and finally that the two sides are equal, as they have the
same size.
Exercises:
(b) Give a complete list of all of the subgroups of S3 (Hint: there are
six).
82 CHAPTER 2. GROUPS
(c) For each x ∈ S3 work out the cyclic subgroup hxi and state the order
of x. Hence show that S3 is not cyclic.
5. Check that the set G = {e, a, b, c} becomes a group under the op-
eration given by the table
e a b c
e e a b c
a a e c b
b b c e a
c c b a e
ment of order n.
14. You are given that there is a group Q = {1, −1, i, −i, j, −j, k, −k} of
order 8 (the quaternion group) in which 1 is the identity, −1 behaves ac-
cording to the ordinary laws of algebra, and in addition i2 = j 2 = −1, ij =
k and ji = −k. In other words, you are given this much of the group
table:
1 −1 i −i j −j k −k
1 1 −1 i −i j −j k −k
−1 −1 1 −i i −j j −k k
i i −i −1 k
−i −i i
j j −j −k −1
−j −j j
k k −k
−k −k k
(a) Work your way round systematically, using associativity and so on,
to complete the table. (This can be done in many ways. For example:
i(−i) = i(−1)i = iiii = (−1)(−1) = 1 = and ik = iij = (−1)j = −j.
Now do the rest.)
2.4. HOMOMORPHISMS AND ISOMORPHISMS 85
2.4.1 Examples. (i) Consider the group R = (R, +, 0) and the doubling
function f : R → R , x 7→ 2x. Given two real numbers x and y, we can
combine them together first within the group (add) and then apply f to
the result, getting 2(x + y). Or we can apply f to each one at the start,
and afterwards combine the results to obtain 2x + 2y. We end up with
the same thing.
(ii) For the same group R consider instead the squaring function g : R →
R , x 7→ x2 . If we first add and then apply g, we obtain (x + y)2 . But if
we apply these processes the other way round, this time we get x2 + y 2 , a
different result.
We see that the function f here behaves in a nice way with respect
to the group structure (addition), whereas g does not. The difference is
expressed by saying that f is a homomorphism, whilst g is not. The formal
definition follows:
86 CHAPTER 2. GROUPS
the pairs (a, a),(a, b),(a, c),(b, b),(b, c) and (c, c). For the first of these we
have θ(aa) = θ(b) = 1 and θ(a)θ(a) = (−1)(−1) = 1, and these are
equal. For the second θ(ab) = θ(c) = −1 and θ(a)θ(b) = (−1)1 = −1,
once again equal. I leave you to check the other cases.
All this checking may seem rather painful. Once we have developed
some machinery, we shall be able to avoid a great deal of it, and construct
homomorphisms in a more streamlined manner . Now for a few more:
In the light of all this, the only proper way to deal with the function k of (ii)
above is to produce a concrete numerical counterexample. For instance:
taking x = y = 2 gives k(xy) = 17 25
2 , which is not equal to k(x)k(y) = 4 .
The first three things to observe about homomorphisms are given in:
2.4. HOMOMORPHISMS AND ISOMORPHISMS 89
2.4.9 Example. Let G and H be any two groups, with identities e and
e0 . There is always at least one homomorphism from G to H, namely the
trivial homomorphism f : G → H , x 7→ e0 for all x ∈ G. This just sends
every element of G to the identity.
Proof. The first statement is evident, and the second follows from
(g ◦ f )(xy) = g(f (xy)) = g(f (x)f (y)) = g(f (x))g(f (y))
= (g ◦ f )(x)(g ◦ f )(y).
Recall from 1.3.20 the concepts of the image and inverse image of
subsets under a function. The next result is a generalization of the previous
Proposition, and is left for the reader to prove along the same lines:
Proof. Let f be injective and let x ∈ ker f . Then f (x) = e0 = f (e) and
so x = e, proving that ker f = {e}. Conversely, assume that ker f = {e}
and let f (x) = f (y). Then f (xy −1 ) = f (x)f (y −1 ) = f (x)(f (y))−1 = e0 .
Thus xy −1 ∈ ker f and it follows from our assumption that xy −1 = e, so
that x = y, proving injectivity.
• e a b • i α β
e e a b i i α β
a a b e α α β i
b b e a β β i α
We have spelt all this out in detail. In future we will not labour the
point, as it will usually be clear immediately that the properties we study
pass over from a given group to one isomorphic to it.
The next result shows that isomorphism behaves just like an equiva-
lence relation:
2.4.19 Remark. Why did we say above that isomorphism behaves like an
equivalence relation, rather than is an equivalence relation? The answer is
that an equivalence relation takes place on a set, and for technical reasons
to do with the foundations of set theory the collection of all groups cannot
be regarded as a set. We should call it a class. It turns out that some
classes or collections are so big that the unrestrained use of all the usual
set-theory operations on them leads to logical paradoxes. The problems
only arise with unimaginably vast collections such as the class of all sets,
or (here) the class of all groups. We need not worry any further about
this, as all the classes which arise in practical everyday Mathematics fall
within the scope of set-theory. If two groups are isomorphic, we say that
they lie in the same isomorphism class.
We have seen above that isomorphic groups have the same order. The
converse is very definitely false. The simplest example is this:
e a b c e0 p q r
e e a b c e0 e0 p q r
a a b c e p p e0 r q
b b c e a q q r e0 p
c c e a b r r q p e0
This has one interesting consequence. For any group G, define the
left multiplication map by an element g ∈ G to be the function lg : G →
G , x 7→ gx. Rather as in the proof of 2.4.24 we note at once that
le = i, lg lh = lgh , (lg )−1 = lg−1 , and in particular that lg is a permutation
of the set G. (It is not a homomorphism). We now have:
Exercises:
(a) f : G → G given by e 7→ e, a 7→ b, b 7→ c, c 7→ a
(b) g : G → G given by e 7→ e, a 7→ c, b 7→ e, c 7→ a
(c) h : G → G given by e 7→ a, a 7→ b, b 7→ e, c 7→ c
(d) k : G → G given by e 7→ e, a 7→ b, b 7→ e, c 7→ b .
4. For each of the above that is a homomorphism, find the kernel and
image.
5. For the group G of Ex.3 show that there are precisely four homo-
morphisms from G to G.
ϕ : G → G/H , x 7→ xH.
We now try to turn the set G/H into a group in such a way that ϕ becomes
a homomorphism. If this is to be so, we will have to have ϕ(x)ϕ(y) =
ϕ(xy), or in other words (xH)(yH) = xyH. This last equation tells us
how we are going to have to define the product of two cosets if we are to
have any chance of things working.
But there is a problem here. For a given coset may be represented in
2.5. QUOTIENT GROUPS 103
xH = H ⇒ xyH = yH.
Recall from 2.4.24 that an element of the form ghg −1 is called a con-
104 CHAPTER 2. GROUPS
Proof. We have seen already that (i) and (ii) are equivalent. Suppose
now that (ii) holds. Then replacing g by g −1 gives g −1 Hg ⊂ H too.
Multiplying on the left by g and on the right by g −1 leads to H ⊂ gHg −1 .
Together with gHg −1 ⊂ H this establishes (iii), and clearly (iii) implies
(ii). Finally, multiplying (iii) on the right by g leads to (iv), and similarly
in the reverse direction.
2.5.4 Examples. (i) If G is abelian then all subgroups are normal. This
is clear from any of the criteria above.
2.5. QUOTIENT GROUPS 105
Proof. There are only two distinct left cosets, namely H and one other,
say xH, which has, of course, to be the complementary set G − H. We
have to check that gH = Hg, for all g. This is trivial if g ∈ H, so assume
now that g ∈
/ H. Then gH = xH = G − H. Similarly Hg = G − H, and
we are done.
We can now return to the theme which began this section and prove:
2.5.10 Example. Let Q = {1, −1, i, −i, j, −j, k, −k} be the quaternion
group of Section 1.3, Exercise 14. The full table of this is:
1 −1 i −i j −j k −k
1 1 −1 i −i j −j k −k
−1 −1 1 −i i −j j −k k
i i −i −1 1 k −k −j j
−i −i i 1 −1 −k k j −j
j j −j −k k −1 1 i −i
−j −j j k −k 1 −1 −i i
k k −k j −j −i i −1 1
−k −k k −j j i −i 1 −1
108 CHAPTER 2. GROUPS
One quickly checks that C = {1, −1} is a normal subgroup, and then
clearly Q/C = {C, iC, jC, kC}. We now compute the group structure
on this. For example (iC)(iC) = i2 C = (−1)C = C and (jC)(kC) =
jkC = iC, and I leave you to do the rest. The table for Q/C is:
C iC jC kC
C C iC jC kC
iC iC C kC jC
jC jC kC C iC
kC kC jC iC C
2.5.11 Remarks. (i) One can view the quotient group construction G/H
as being in some sense a counterpoint to the study of the subgroup H.
In the latter, we narrow our attention down to what is going on inside H,
blinkering ourselves to the rest of the ambient group G. But when we
look at the quotient group, it’s as if we have “collapsed” H right out of
the picture, and are now concentrating on what residual structure is left.
Indeed, each coset has effectively been shrunk down to a single point.
(ii) As noted in Section 1.3, the group law in some abelian groups is written
additively, in which case we write the cosets as x + H. In that case we
will continue to use an additive notation in the quotient group and write
(x + H) + (y + H) = x + y + H.
(iii) When working in a quotient group G/H, it rapidly becomes tedious
2.5. QUOTIENT GROUPS 109
We have talked a lot about cyclic groups and their properties in Section
1.3, but how do we know that they actually exist? We can now fill in this
gap:
has order n. This is called the group of integers mod n and will also be
denoted by Zn . It is evidently cyclic, generated by [1].
2.5.14 Theorem. (i) Any two cyclic groups of the same order are iso-
morphic;
(ii) For each n ≥ 1 there is, up to isomorphism, precisely one cyclic group
of order n. A model for this is Zn ;
(iii) There is, up to isomorphism, precisely one cyclic group of infinite
order. A model for this is Z.
Proof. We have now seen in Example 2.5.13 that cyclic groups of all
possible orders exist. It remains to prove uniqueness. Let G be a cyclic
group, generated by an element x. Assume first that G is infinite. Clearly
the map f : Z → G , k 7→ xk is a homomorphism. Since G = hxi it
is surjective. Moreover, as x has infinite order xk = e ⇒ k = 0. Hence
2.5. QUOTIENT GROUPS 111
We can now see how to avoid doing all the tedious associativity checks
to verify that the group G = {e, a, b, c} of Examples 2.2.2 really is a group.
Indeed, we can see at once that it is the cyclic group of order four, and as
such exists.
The method whereby we showed above that Zn is isomorphic to G may
be greatly extended to prove the next Theorem, which is one of the central
planks of group theory. Recall that we already observed in Proposition
2.5.7 that the kernel of a homomorphism is a normal subgroup.
◦ even odd
even even odd
odd odd even
Now let C = {1, −1} be the cyclic group of order two under multiplication,
and define a function f : S3 → C by sending the even permutations to 1,
and the odd ones to −1. This is seen at once to be a surjective homo-
morphism with kernel H = {i, α, β}, and thus it induces an isomorphism
∼
S3 /H → C, as in Example 2.5.9.
We shall see in a later section how the idea of even and odd permuta-
tions extends to all symmetric groups Sn .
2.5.21 Example. In Section 1.4, Exercise 9 you were asked to show that
the set of all functions fa,b : R → R , x 7→ ax + b, with a, b ∈ R , a 6= 0
114 CHAPTER 2. GROUPS
ϕ(H) = {ē}. Thus the reverse map is also well-defined and we just have
to check that the two maps are inverses of each other, or equivalently that
the composition each way round is the identity. This is left as an exercise
for the reader.
2.5.23 Example. Recall the group Zn = Z/nZ, where n ≥ 1. By Propo-
sition 2.3.24 the subgroups of Z are all of the form K = mZ, and one
checks at once that K contains nZ if and only if m|n (m divides n).
Thus the subgroups of Z/nZ are the various quotients mZ/nZ, where
m|n and we may as well take m ≥ 1. For instance Z/12Z has subgroups
Z/12Z, 2Z/12Z, 3Z/12Z 4Z/12Z, 6Z/12Z and 12Z/12Z = {0̄}.
If we reexamine the first two lines in the proof of Theorem 2.5.15,
we see that the argument there also establishes the following very useful
principle:
2.5.24 Proposition. If f : G → H is a homomorphism and K is a
normal subgroup of G contained in ker f , then f induces a homomorphism
f˜ : G/K → H, defined by f˜(ḡ) = f (g). We have f = f˜ ◦ ϕ, where
ϕ : G → G/K is the canonical homomorphism, and we say that f factors
through G/K.
Of course f˜ will not be an isomorphism unless K = ker f and f
happens to be surjective.
There are two further isomorphism theorems, which are actually just
particular cases of 2.5.15, but nevertheless useful to record. First a defi-
nition:
2.5.25 Definition. If X and Y are two subsets of a group (or monoid) G,
define their product to be the set XY = {xy : x ∈ X, y ∈ Y }.
116 CHAPTER 2. GROUPS
∼
H/H ∩ K → HK/K , given by h(H ∩ K) 7→ hK.
2.5. QUOTIENT GROUPS 117
(We could just write h̄ 7→ h̄, provided we remember that the two h̄’s mean
different things).
mZ/eZ ∼
= dZ/nZ .
∼
(G/K)/(H/K) → G/H given by ḡ(H/K) 7→ gH .
Exercises:
13. Consider the direct product G × H of two groups (Section 2.4, ex.7)
and let G0 , H 0 be normal subgroups of G, H respectively. Use Theorem
2.5.15 to show that G0 × H 0 C G × H and that (G × H)/(G0 × H 0 ) ∼ =
G/G0 × H/H 0 .
14. Let G be an abelian group of order pq, where p, q are distinct primes.
We aim to prove that G is cyclic.
(i) If all elements 6= e have order p, choose one such element x and derive
a contradiction by considering the quotient G/ < x > .
(ii) Deduce that either G is cyclic or that it contains an element a of order
q, and likewise an elements b of order p.
(iii) Show that in the latter case ab has order pq, and G is again cyclic.
2.6. MORE ON PERMUTATIONS 121
Note that in cycle notation we can start anywhere we like and proceed
round in cyclic order. Thus (13254) = (32541) = (25413) = (54132) =
(41325). We will adopt the convention that ak+1 = a1 . We can then say
that σ(ai ) = ai+1 for all i such that 1 ≤ i ≤ k. Of course, in general two
permutations do not commute with each other. But there is one situation
in which they do:
2.6.3 Lemma. The size of Orb(a) is the least positive integer k such that
σ k (a) = a. We have Orb(a) = {a, σ(a), σ 2 (a), . . . σ k−1 (a)}.
Z It is important to realise that the gaps between adjacent cycles just rep-
resent composition of functions and that when we multiply two permuta-
tions we must always work from right to left. For example, if σ is as above
and τ = (1 3 8)(5 9 7)(6 4 10), then στ = (2 9)(1 4 7)(3 6 5 10)(1 3 8)(5 9 7)(6 4 10).
to work this out we see what is the effect on each of the integers 1, . . . , 10
in turn, working from right to left. The effect of the successive cycles on
1 is as follows: 1 7→ 1 7→ 1 7→ 3 7→ 6 7→ 6 7→ 6, so that 1 7→ 6. We now
see what happens to 6: 6 7→ 4 7→ 4 7→ 4 7→ 4 7→ 7 7→ 7. Continuing in
this way, we find that στ = (1 6 7 10 5 2 9)(3 8 4).
2.6.5 Proposition. (i) The order of a k-cycle equals the length k of the
cycle;
(ii) If σ = σ1 σ2 · · · σr is a permutation, expressed as a product of disjoint
cycles, then the order of σ is the least common multiple of the lengths of
the individual cycles σi .
Thus (i j) just interchanges i and j and leaves all the other integers
alone. We shall now see that all permutations can be generated as products
of transpositions:
Proof. The formula is easily checked, by working out the effect of the right-
hand side on the ai and on the other elements of I. By combining this
with Theorem 2.6.4 it follows that all permutations can be decomposed
into transpositions.
It is clear that in general σ will shuffle round the brackets in the expression
for ∆(x1 , . . . , xn ), possibly changing some of the signs, so that therefore
σ · ∆ = ±∆. This leads to the:
even odd
even even odd
odd odd even
.
We have seen in Theorem 2.6.7 that every permutation can be ex-
pressed as a product of transpositions. In view of Lemma 2.6.11, this
gives us an effective way of finding the sign of a permutation:
2.6. MORE ON PERMUTATIONS 129
Proof. Write ρ = (σ(a1 )σ(a2 ) . . . σ(ak )). Let us calculate the effect of
each side on the elements m of I = {1, . . . , n}. If m = σ(ai ), for some
i, then στ σ −1 (m) = στ (ai ) = σ(ai+1 ) = ρ(m). On the other hand, if m
is not equal to any σ(ai ), then σ −1 (m) is not equal to any ai , and hence
τ leaves it alone. So στ σ −1 (m) = σσ −1 (m) = m. But also ρ(m) = m,
and we have shown that στ σ −1 = ρ.
As our example shows, the Theorem does more than just tell us when
two permutations are conjugate. It actually provides a method for finding
an element ρ which conjugates the one into the other. By now, you may
well have noticed the following simple fact, the proof of which is left as
an exercise:
Observe at once that since geg −1 = e for all g, the conjugacy class
of e is just {e}. Using Theorem 2.6.19 it is now easy to determine the
conjugacy classes in S3 and S4 . The results are as follows:
{i} ∪ {(• •)(• •)} = {i, (1 2)(3 4), (1 3)(2 4), (1 4)(2 3)} and K = {i} ∪
{(• • •)} ∪ {(• •)(• •)} = {i, (1 2 3), (1 3 2), (1 2 4), (1 4 2), (1 3 4),
(1 4 3), (2 3 4), (2 4 3), (1 2)(3 4), (1 3)(2 4), (1 4)(2 3)} .
Of course, we still have to check that these really are subgroups: all we
know so far is that if they are subgroups, then they will be normal as well.
But we have seen K before: it is the alternating group A4 of Examples
2.6.16. As for H, we check at once that each element is self-inverse,
and the the product of any two non-identity elements is the third. So H
really is a subgroup. It is often denoted V and called Klein’s Vierergruppe
(“fours group”).
In summary, S4 has precisely four normal subgroups: {i}, V, A4 and S4
.
To round off this section let us think some more about the ways in
which a permutation can be written as a product of disjoint cycles. Let
us begin with the example of σ = (1 2)(3 4 5) in S5 . Even if we insist on
keeping the 2-cycle before the 3-cycle, there are still 2×3 = 6 ways of writ-
ing σ, namely (1 2)(3 4 5), (1 2)(4 5 3), (1 2)(5 3 4), (2 1)(3 4 5), (2 1)(4 5 3)
and (2 1)(5 3 4) . This is because a k-cycle can be denoted in k different
ways, by cyclically permuting round the entries. In just the same way, every
permutation of the form (• •)(• • •) can be written in six ways. Suppose
we wish to determine the size of the conjugacy class (• •)(• • •). We can
fill in the five slots with the numbers 1, . . . , 5 in 5! = 120 ways. But in so
doing, we have actually counted each permutation six times over. So the
true size of this class is 120/6 = 20.
Consider now the more substantial example of the permutation τ =
(1)(2)(3 4)(5 6)(7 8)(9 10 11)(12 13 14) in S14 . This has the cycle-structure
(•)(•)(• •)(• •)(• •)(• • •)(• • •), where we have made sure to put in
the 1-cycles. Not only can each cycle in τ be cyclically reordered, but also
the 1-cycles can be placed in either of 2! = 2 orders, and likewise for the
3-cycles. And the 2-cycles may be rearranged in 3! = 6 ways. For example
we can also write τ = (2)(1)(5 6)(7 8)(3 4)(12 13 14)(9 10 11). All in all,
including 1-cycles, τ can be written in 2!×3!×2!×23 ×32 = 1728 different
ways. Now there are 14! ways to fill in the slots of (•)(•)(• •)(• •)(• •)(• •
•)(• • •) with the numbers 1, . . . , 14. So the size of this conjugacy class
is 14!/1728 = 50450400. This method clearly extends to prove the next
result. The reader may supply the formal details.
136 CHAPTER 2. GROUPS
σ (1 4)(2 3 5)
(2 3)(1 4 5)
(2 3)(4 5 1)
(2 3)(5 1 4)
τ
(3 2)(1 4 5)
(3 2)(4 5 1)
(3 2)(5 1 4)
Sending each entry of σ to the one directly beneath it ! in the first ex-
1 4 2 3 5
pression for τ defines a permutation ρ = = (1 2)(3 4)
2 3 1 4 5
with the property that ρσρ−1 = τ . Using instead the other representa-
tions of τ leads to five more elements ρ with the desired property namely
ρ = (1 2 4 3 5), (1 2 5 4 3), (1 3 4 2), (1 3 5)(2 4) and (1 3)(2 5 4) . These six
2.6. MORE ON PERMUTATIONS 137
σ (1 3)(2 7)(4 6 5) ρ
(1 3)(2 7)(4 6 5) i
(1 3)(2 7)(6 5 4) (4 6 5)
(1 3)(2 7)(5 4 6) (4 5 6)
(1 3)(7 2)(4 6 5) (2 7)
(1 3)(7 2)(6 5 4) (2 7)(4 6 5)
(1 3)(7 2)(5 4 6) (2 7)(4 6 5)
(3 1)(2 7)(4 6 5) (1 3)
(3 1)(2 7)(6 5 4) (1 3)(4 6 5)
(3 1)(2 7)(5 4 6) (1 3)(4 5 6)
(3 1)(7 2)(4 6 5) (1 3)(2 7)
(3 1)(7 2)(6 5 4) (1 3)(2 7)(4 6 5)
(3 1)(7 2)(5 4 6) (1 3)(2 7)(4 6 5)
σ
(2 7)(1 3)(4 6 5) (1 2)(3 7)
(2 7)(1 3)(6 5 4) (1 2)(3 7)(4 6 5)
(2 7)(1 3)(5 4 6) (1 2)(3 7)(4 5 6)
(2 7)(3 1)(4 6 5) (1 2 3 7)
(2 7)(3 1)(6 5 4) (1 2 3 7)(4 6 5)
(2 7)(3 1)(5 4 6) (1 2 3 7)(4 5 6)
(7 2)(1 3)(4 6 5) (1 7 3 2)
(7 2)(1 3)(6 5 4) (1 7 3 2)(4 6 5)
(7 2)(1 3)(5 4 6) (1 7 3 2)(4 5 6)
(7 2)(3 1)(4 6 5) (1 7)(2 3)
(7 2)(3 1)(6 5 4) (1 7)(2 3)(4 6 5)
(7 2)(3 1)(5 4 6) (1 7)(2 3)(4 5 6)
2.6. MORE ON PERMUTATIONS 139
Exercises:
4. Show that a group is abelian if and only if all its conjugacy classes
are of size one.
5. Consider the set of permutations S = {i, (1 2 3), (1 3 2), (1 2), (1 3), (2 3)}
in S4 . Show that for distinct g, h ∈ S we cannot have g −1 h ∈ V , the Klein
group, and hence that the six elements of S represent different cosets of
V . Deduce that S4 /V ∼ = S3 .
2.7.1 Symmetries
Consider the set R2 = R × R, which we may visualise as the ordinary
Euclidean plane with its system of x- and y-axes. If P = (x, y) and
P 0 = (x0 , y 0 ) are two points of R2 , we have the usual notion of the
2.7. EXAMPLES: SYMMETRIES AND MATRICES 141
p
distance in the plane d(P, P 0 ) = (x − x0 )2 + (y − y 0 )2 between them.
We will be interested in those functions from R2 to itself that preserve
distances in the following sense:
(i) translations: we slide or translate each point of the plane the same
amount and in the same direction. Thus f (x, y) = (x + a, y + b), for some
fixed a, b ∈ R.
(ii) rotations: all points are rotated through a fixed angle about a given
point of the plane.
(iii) reflections: we fix a particular line in the plane and then send each
point to its reflection, or mirror-image, in this line.
(iv) glide-refections: these are by definition the compositions of a reflec-
tion with a translation.
Under composition, the isometries form a group:
Now for any subset or “figure” F in the plane, we may consider just
those isometries f of the plane which preserve F , in the sense that f (F ) =
F:
Let us introduce some notation for these symmetries. For any θ ∈ R let
rθ be the symmetry of anticlockwise rotation around the origin through
an angle of θ radians, and let s be reflection in the x-axis. We leave you
to check that the formulas for these symmetries are
You are further left to check that the product srθ (in other words, the
composite s ◦ rθ ), which is given by the formula (srθ )(x, y) = (x cos θ −
y sin θ, −x sin θ − y cos θ) represents reflection in the line x sin(θ/2) +
y cos(θ/2)
= 0 obtained by rotating the x-axis clockwise about the origin through
an angle of θ/2. As θ varies we obtain in this way all reflections in lines
through the origin. Hence S(D) = {rθ : θ ∈ R} ∪ {srθ : θ ∈ R}. Note
that there are repetitions here: if θ ≡ ϕ (mod 2π), ie. θ − ϕ is a multiple
of 2π, then rθ = rϕ . So if we wanted we could restrict θ to lie in the range
0 ≤ θ < 2π.
The group structure is determined by the following table, which shows how
144 CHAPTER 2. GROUPS
rϕ srϕ
rθ rθ+ϕ srϕ−θ
srθ srθ+ϕ rϕ−θ
Note that this symmetry group can be described in other ways: S(D) =
S(B) = S(T) = S({(0, 0)}), where B = {(x, y) ∈ R2 : x2 + y 2 < 1}
and T = {(x, y) ∈ R2 : x2 + y 2 = 1}.
and Dn is a group of order 2n. We have srs−1 = r−1 and the set of
rotations H = {i, r, r2 , . . . , rn−1 } is a normal subgroup of index 2. The
remaining elements are reflections in the various axes of symmetry making
angles of kπ/n with the x-axis. Since sr 6= rs, the dihedral group Dn is
nonabelian.
2.7. EXAMPLES: SYMMETRIES AND MATRICES 145
The hook on the arrow is a useful piece of notation, signifying that the
mapping is injective. To all intents and purposes we may identify Dn with
its image under this mapping, and regard Dn as a subgroup of Sn . Let us
see how this works out in a couple of concrete cases:
is an array of real numbers aij . The set of all such matrices is denoted
Mn (R). We can similarly talk about the sets Mn (C), Mn (Q), Mn (Z) of
complex, rational or integral matrices. Two (n×n) matrices A = (aij ) and
P
B = (bij ) have a product AB = C = (cij ), defined by cij = k aik bkj ,
and this defines an operation on Mn (R). There is an identity element
2.7. EXAMPLES: SYMMETRIES AND MATRICES 147
Proof. Write G = GLn (R). If A and B are invertible, then so is AB, and
(AB)−1 = B −1 A−1 . Thus matrix multiplication does indeed restrict to
give an operation on the subset G. We already know that this operation
is associative, and there is an identity I ∈ G. Finally, by very definition,
each element A ∈ G has an inverse A−1 ∈ G.
It is not our purpose now to set up and develop the theory of determinants,
interesting though it is. It will suffice for now to know that determinants
148 CHAPTER 2. GROUPS
exist and have certain nice properties. For completeness, we give the
definition here, and refer you to the Appendix for more information.
2.7.14 Definition. The real special linear group is the group SLn (R) =
{A ∈ GLn (R) : det A = 1}. Likewise for the complex, rational and
integral special linear groups SLn (C), SLn (Q) and SLn (Z).
We have:
Proof. In the first three cases the homomorphism det : GLn (K) → K ∗ is
a 0 ... 0
0 1 . . . 0
surjective. For, given any a ∈ K ∗ , the matrix A =
... ... . . . ... has
0 0 ... 1
determinant a. By definition, the kernel is SLn (K), and the result follows
from Theorem 2.5.15. The argument is similar in the last case.
Exercises:
2. Show that every reflection t = srθ in the disk group S(D) has or-
der two, and that trφ t−1 = r−ϕ and t(srϕ )t−1 = sr2θ−ϕ .
150 CHAPTER 2. GROUPS
151