Download as pdf or txt
Download as pdf or txt
You are on page 1of 53

Groups Simon Wadsley

Visit to download the full and correct content document:


https://textbookfull.com/product/groups-simon-wadsley/
More products digital (pdf, epub, mobi) instant
download maybe you interests ...

Theory of Groups and Symmetries Finite groups Lie


groups and Lie Algebras 1st Edition Alexey P. Isaev

https://textbookfull.com/product/theory-of-groups-and-symmetries-
finite-groups-lie-groups-and-lie-algebras-1st-edition-alexey-p-
isaev/

Dragonmancer 1st Edition Simon Archer Archer Simon

https://textbookfull.com/product/dragonmancer-1st-edition-simon-
archer-archer-simon/

Matrix Groups for Undergraduates Kristopher Tapp

https://textbookfull.com/product/matrix-groups-for-
undergraduates-kristopher-tapp/

Hero of Hybrids 1st Edition Simon Archer Archer Simon

https://textbookfull.com/product/hero-of-hybrids-1st-edition-
simon-archer-archer-simon/
After the Virus 1st Edition Simon Archer Archer Simon

https://textbookfull.com/product/after-the-virus-1st-edition-
simon-archer-archer-simon/

Level Up Super 1st Edition Simon Archer Archer Simon

https://textbookfull.com/product/level-up-super-1st-edition-
simon-archer-archer-simon/

Level Up Checkpoint 1st Edition Simon Archer Archer


Simon

https://textbookfull.com/product/level-up-checkpoint-1st-edition-
simon-archer-archer-simon/

Theory of Groups and Symmetries: Representations of


Groups and Lie Algebras, Applications 1st Edition
Alexey P. Isaev

https://textbookfull.com/product/theory-of-groups-and-symmetries-
representations-of-groups-and-lie-algebras-applications-1st-
edition-alexey-p-isaev/

Level Up Press Start 1st Edition Simon Archer Archer


Simon

https://textbookfull.com/product/level-up-press-start-1st-
edition-simon-archer-archer-simon/
GROUPS

SIMON WADSLEY

Contents
1. Examples of groups 2
1.1. A motivating example 2
1.2. Some initial definitions 4
1.3. Further geometric examples 7
1.4. Subgroups and homomorphisms 9
1.5. The Möbius Group 13
2. Lagrange’s Theorem 17
2.1. Cosets 17
2.2. Lagrange’s Theorem 18
2.3. Groups of order at most 8 19
2.4. The Quaternions 22
2.5. Fermat–Euler theorem 22
3. Group Actions 23
3.1. Definitions and examples 23
3.2. Orbits and Stabilisers 24
3.3. Conjugacy classes 27
3.4. Cayley’s Theorem 28
3.5. Cauchy’s Theorem 29
4. Quotient Groups 29
4.1. Normal subgroups 29
4.2. The isomorphism theorem 32
5. Matrix groups 33
5.1. The general and special linear groups 34
5.2. Möbius maps as projective linear transformations 35
5.3. Change of basis 37
5.4. The orthogonal and special orthogonal groups 39
5.5. Reflections 41
6. Permutations 44
6.1. Permutations as products of disjoint cycles 44
6.2. Permuations as products of transpositions 47
6.3. Conjugacy in Sn and in An 48
6.4. Simple groups 50

1
2 SIMON WADSLEY

Purpose of notes. Please note that these are not notes of the lectures but notes
made by the lecturer in preparation for the lectures. This means they may not
exactly correspond to what was said and/or written during the lectures.

Lecture 1
1. Examples of groups
Groups are fundamentally about symmetry. More precisely they are an algebraic
tool designed to abstract the notion of symmetry. Symmetry arises all over mathe-
matics; which is to say that groups arise all over mathematics. Roughly speaking a
symmetry is a transformation of an object that preserves certain properties of the
object.
As I understand it, the purpose of this course is two-fold. First to introduce
groups so that those who follow the course will be familiar with them and be better
equipped to study symmetry in any mathematical context that they encounter.
Second as an introduction to abstraction in mathematics, and to proving things
about abstract mathematical objects.
It is perfectly possible to study groups in a purely abstract manner without
geometric motivation. But this seems to both miss the point of why groups are
interesting and make getting used to reasoning about abstract objects more difficult.
So we will try to keep remembering that groups are about symmetry. So what do
we mean by that?

1.1. A motivating example.


Question. What are the distance preserving functions from the integers to the
integers? That is what are the members of the set
Isom(Z) := {f : Z → Z such that |f (n) − f (m)| = |n − m| for all n, m ∈ Z}?
These functions might reasonably be called the symmetries of the integers; they
describe all the ways of ‘rearranging’ the integers that preserve the distance between
any pair.
Let’s begin to answer our question by giving some examples of such functions.
Suppose that a ∈ Z is an integer. We can define the function ‘translation by a’ by
ta : n 7→ n + a for n ∈ Z.
For any choice of m, n ∈ Z
|ta (n) − ta (m)| = |(n + a) − (m + a)| = |n − m|.
Thus ta is an element of Isom(Z). We might observe that if a and b are both integers
then
(ta ◦ tb )(n) = ta (b + n) = a + b + n = ta+b (n)
for every integer n, that is that ta+b = ta ◦ tb 1. Moreover t0 is the identity or ‘do
nothing’ function id : Z → Z that maps every integer n to itself. Thus for every
a ∈ Z, t−a is the inverse of ta , that is ta ◦ t−a = id = t−a ◦ ta .

1This is because two functions f, g : X → Y are the same function, i.e. f = g, precisely if
they have the same effect on every element of the set they are defined on, i.e. f (x) = g(x) for all
x ∈ X.
GROUPS 3

Suppose now that f ∈ Isom(Z) is a symmetry of the integers. Consider the


function g := t−f (0) ◦ f . Then for n, m ∈ Z,
|g(n) − g(m)| = |(f (n) − f (0)) − (f (m) − f (0))| = |f (n) − f (m)| = |n − m|,
so g ∈ Isom(Z) is also a symmetry of the integers.
Moreover g(0) = t−f (0) (f (0)) = f (0) − f (0) = 0, i.e. g fixes the integer 0. What
does this tell us about g? For example, what does it tell us about g(1)? Since g is
a symmetry and g(0) = 0 it must be the case that
|g(1)| = |g(1) − 0| = |g(1) − g(0)| = |1 − 0| = 1.
That is g(1) = ±1.
If g(1) = 1, what else can we say? For any n ∈ Z,
|g(n)| = |g(n) − g(0)| = |n − 0| = |n|
i.e. g(n) = ±n. But also,
|g(n) − 1| = |g(n) − g(1)| = |n − 1|
i.e. g(n) = 1 ± (n − 1). These two conditions together force g(n) = n and so g = id.
Now in this case
tf (0) = tf (0) ◦ id = tf (0) ◦ (t−f (0) ◦ f ) = (tf (0) ◦ t−f (0) ) ◦ f = id ◦f = f.
Thus f is translation by f (0) in this case.
What about the case when g(1) = −1? In this case we still must have g(n) = ±n
for every integer n but now also
|g(n) + 1| = |g(n) − g(1)| = |n − 1|
i.e. g(n) = −1 ± (n − 1). These two conditions together force g(n) = −n and so g
is the ‘reflection about 0’-function
s : n 7→ −n for all n ∈ Z.
Now we’ve seen that s = g = t−f (0) ◦ f in this case. It follows that f = tf (0) ◦ s.
We’ve now proven that every element of Isom(Z) is either a translation ta or of
the form ta ◦ s (with a ∈ Z in either case). That is all symmtries of Z are of the
form n 7→ n + a or of the form n 7→ a − n.
It is worth reflecting at this point on some key facts we’ve used in the argument
above which is sometimes known as a ‘nailing to the wall argument’.
(1) We’ve used that the composition of two symmetries of the integers is itself
a symmetry of the integers. In fact, we’ve only used this for some special
cases but it is true in general since if f, g ∈ Isom(Z) and n, m ∈ Z then
|f (g(n)) − f (g(m))| = |g(n) − g(m)| = |n − m|.
We might note that for a, n ∈ Z,
s(ta (n)) = s(n + a) = −a − n = t−a (s(n))
and so s ◦ ta = t−a ◦ s. Thus order of composition matters.
(2) We’ve used that there is a ‘do nothing’ symmetry of the integers id and
that for any other symmetry f , f ◦ id = f = id ◦f .
(3) We’ve used that symmetries are ‘undo-able’, that is that given any sym-
metry f there is a symmetry g such that g ◦ f = id = f ◦ g (in fact we’ve
only used this for f = ta and f = s and only that there is a g such that
g ◦ f = id but again it is true as stated. (Why?).
4 SIMON WADSLEY

(4) We’ve used that composition of symmetries is associative, that is that for
symmetries f, g and h, (f ◦ g) ◦ h = f ◦ (g ◦ h).
We’ll see that these properties say precisely that Isom(Z) is a group.

1.2. Some initial definitions. First we need to make some definitions.


Definition. Suppose that S is a set. A binary operation on S is a function
◦ : S × S → S; (x, y) 7→ x ◦ y.
This defintion means that a binary operation is something that takes an ordered
pair of elements of S and uses them to produce an element of S. If x ◦ y = y ◦ x
then we say that x and y commute (with respect to ◦). We say ◦ is commutative
if every pair of elements of S commute.
Examples.
(1) Composition of functions is a non-commutative binary operation on Isom(Z).
(2) Addition, multiplication, and subtraction are all binary operations on Z.
Note that addition and multiplication are both commutative operations on
Z but distinct integers never commute with respect to subtraction.
(3) Addition and multiplication are also binary operations on N := {1, 2, 3, . . .}.
Subtraction is not a binary operation on N since 2 − 3 6∈ N.
(4) Exponentiation: (a, b) 7→ ba is a binary operation on N.
(5) If X is any set and S = {f : X → X} is the set of all functions from X to
itself then composition of functions is a binary operation on S.
Definition. A binary operation ◦ on a set S is associative if (x ◦ y) ◦ z = x ◦ (y ◦ z)
for all x, y, z ∈ S.
This means that when ◦ is associative there is a well-defined element x ◦ y ◦ z ∈ S
i.e. it doesn’t matter which of the two ◦ we use first. It will be instructive to
convince yourself that if ◦ is an associative binary operation on S and w, x, y, z ∈ S
then
w ◦ (x ◦ y ◦ z) = (w ◦ x) ◦ (y ◦ z) = (w ◦ x ◦ y) ◦ z.
Having done this you should also convince yourself that there is nothing special
about four and the obvious generalisation holds for any (finite) number of elements
of S whenever ◦ is associative. This means that whenever ◦ is an associative binary
operation we may (and will!) omit brackets, writing for example w ◦x◦y ◦z without
ambiguity. If it is clear what operation we have in mind we will often omit it too,
writing wxyz, for example.
Examples.
(1) Addition and multiplication are associative when viewed as binary opera-
tions on Z or N. Subtraction is not associative on Z since ((0 − 1) − 2) = −3
but 0 − (1 − 2) = 1 6= −3.
2
(2) Exponentiation (a, b) 7→ ba is not associative on N since 23 = 29 but
3 2 6 9
(2 ) = 2 6= 2 .
(3) Composition is always an associative operation on the set of functions from
X to X since if f, g and h are three such functions and x ∈ X then
((f ◦ g) ◦ h)(x) = f (g(h(x))) = (f ◦ (g ◦ h))(x).
GROUPS 5

Definition. A binary operation ◦ on a set S has an identity if there is some element


e ∈ S such that for all x ∈ S, e ◦ x = x = x ◦ e.
Examples.
(1) 0 is an identity for addition on Z but addition has no identity on N. 1 is
an identity for multiplication on both these sets. Subtraction on Z does
not have an identity since if e − x = x for all x ∈ Z then e = 2x for all
x ∈ Z and this is absurd. Note however that x − 0 = x for all x ∈ Z. We
sometimes say that 0 is a right identity for subtraction to d9escribe this.
(2) (a, b) 7→ ba does not have an identity but 1 is a left identity in the obvious
sense.
(3) If X is any set then the identity function id : X → X; s 7→ s is an identity
for composition of functions from X to X.
Lemma. If a binary operation ◦ on a set S has an identity then it is unique.
Proof. Suppose that e and e0 are both identities on S. Then e ◦ e0 = e0 = e0 ◦ e
since e is an identity. But also e0 ◦ e = e = e ◦ e0 since e0 is an identity. Thus e = e0
as required. 

Lecture 2
Definition. If a binary operation ◦ on a set S has an identity e then we say that
it has inverses if for every x ∈ S there is some y ∈ S such that x ◦ y = e = y ◦ x.
Examples.
(1) + on Z has inverses since for every n ∈ Z, n + (−n) = 0 = (−n) + n.
Multiplication does not have inverses on N or Z since there is no integer
(and therefore no natural number) n such that 2n = 1.
(2) Multiplication defines an associative binary operation on the rationals Q
with an identity (1) but it still does not have inverses. Although for every
non-zero rational q, 1/q is also rational and q · 1/q = 1 = 1/q · q, 0 is also
rational and there is no rational r such that r·0 = 1. However multiplication
does have inverses on the set Q\{0}.
(3) In general composition on the set of functions X → X does not have in-
verses. For example the function f : Z → Z; n 7→ 0 has no inverse since if
g : Z → Z were an inverse then we’d have f (g(n)) = n for all n ∈ Z but in
fact however g is defined f (g(1)) = 0. This idea can be adapted to show
that whenever |X| > 1 there is a function f : X → X that has no inverse.
Definition. A set G equipped with a binary operation ◦ is a group if
(i) the operation ◦ is associative;
(ii) the operation ◦ has an identity;
(iii) the operation ◦ has inverses.
Examples.
(1) Isom(Z) is a group (under composition).
(2) (Z, +) is a group since + is associative and has an identity and inverses.
(3) (N, +) is not a group since it does not have an identity.
6 SIMON WADSLEY

(4) (Z, −) is not a group since − is not associative.2


(5) (Z, ·) is not a group since it does not have inverses but (Q\{0}, ·) is a group.
(6) If X is a set with more than one element then the set of functions X → X
is not a group under composition of functions since not all such functions
have inverses.
We will sometimes say that G is a group without specifying the operation ◦.
This is laziness and the operation will always be implicit and either clear what it is
(in concrete settings) or unimportant what it is (in abstract settings). We’ll nearly
always call the identity of a group e (or eG if we want to be clear which group it is
the identity for) if we don’t know it by some other name.
Definition. We say that a group G is abelian if any pair of elements of G commute.
Definition. We say that a group G is finite if it has finitely many elements as a
set. We call the number of elements of a finite group G the order of G written |G|.
Example. For every integer n > 1 we can define a group that is the set Zn :=
{0, 1, . . . , n − 1} equipped with the operation +n where x +n y is the remainder
after dividing x + y by n3. It is straightforward to see that Zn is an abelian group
of order n.
Lemma. Suppose that G is a group.
(i) inverses are unique i.e. if g ∈ G there is precisely one element g −1 in G such
that g −1 g = e = gg −1 ;
(ii) for all g ∈ G, (g −1 )−1 = g;
(iii) for all g, h ∈ G, (gh)−1 = h−1 g −1 (the shoes and socks lemma).
Proof. (i) By assumption every element g ∈ G has at least one inverse. We must
show that there is at most one. Suppose that h, k ∈ G such that gh = e = hg and
gk = e = kg. Then
h = eh = kgh = ke = k
i.e. h and k are the same element of G.
(ii) Given g ∈ G, g satisfies the equations for the inverse of g −1 i.e.
gg −1 = e = gg −1
so by (i), (g −1 )−1 = g.
(iii) Given g, h ∈ G,
(gh)(h−1 g −1 ) = ghh−1 g −1 = geg −1 = gg −1 = e
and similarly (h−1 g −1 )(gh) = e. 
Notation. For each element g in a group G and natural number n we define g n
recursively by g 1 := g and g n := gg n−1 for n > 1. We’ll also write g 0 := e and
g n := (g −1 )−n for integers n < 0. It follows that g a g b = g a+b for all a, b ∈ Z.
Definition. If G is a group then we say that g ∈ G has finite order if there is
a natural number n such that g n = e. If g has finite order, we call the smallest
natural number n such that g n = e the order of g and write o(g) = n.
2recall that it also does not have an identity but to see that it is not a group it suffices to see
that any one of the three properties fails.
3Z
12 is familiar from everyday life. When is it used?
GROUPS 7

1.3. Further geometric examples.

1.3.1. Symmetry groups of regular polygons. Suppose we want to consider the set
D2n of all symmetries of a regular polygon P with n vertices (for n > 3) living
in the complex plane C. By symmetry of P we will mean a distance preserving
transformation of the plane that maps P to itself. We might as well assume that
the centre of P is at the origin 0 and that one of the vertices is the point 1 = 1+0i4.
Proposition. D2n is a group of order 2n under composition.

Lecture 3
Proof. First we observe that if g, h ∈ D2n then for z, w ∈ C,
|g(h(z) − g(h(w))| = |h(z) − h(w)| = |z − w|,
and if p ∈ P then g(h(p)) ∈ P since h(p) ∈ P . Thus composition of functions defines
a binary operation on D2n . Since composition of functions is always associative this
operation on D2n is associative. Moreover the identity function id is obviously a
symmetry of P . Thus to see that D2n is a group it remains to show that every
element of D2n is invertible. We could do this directly but instead we’ll use a nailing
to the wall argument as we did for Isom(Z).
Let r : C → C; z 7→ e2πi/n z denote rotation anticlockwise about 0 by 2π/n. Then
r does preserve distances in C as
|r(z) − r(w)| = |e2πi/n ||z − w| = |z − w|
and r(P ) = P i.e. r ∈ D2n . Moreover rm ∈ D2n for all m ∈ N and rn = id; i.e.
r−1 = rn−1 .
Similarly let s : C → C; z 7→ z̄ denote complex conjugation; i.e. reflection in the
real axis. Once again s preserves distances in C and s(P ) = P . Moreover s2 = id;
i.e. s−1 = s.
We will try to write down all the elements of D2n using r and s.
Consider the set V := {e2πik/n | k = 0, 1, . . . n − 1} ⊂ C of nth roots of unity in
the complex plane, the set of vertices of our regular n-gon. Any element of D2n will
preserve V ; that is to say if g ∈ D2n then g(V ) = V 5. So if we pick any g ∈ D2n
then g(1) = e2πik/n for some k = 0, 1, . . . , n − 1. Thus
rn−k g(1) = e2πi(n−k) e2πik/n = 1.
Now e2πi/n and e−2πi/n are the only two elements of V of distance |1 − e2πi/n |
from 1. Thus rn−k g(e2πi/n ) = e±2πi/n .
If rn−k g(e2πi/n ) = e2πi/n then rn−k g is a distance preserving transformation of
C fixing 0, 1 and e2πi/n . Thus rn−k g = id and g = rk .
If rn−k g(e2πi/n ) = e−2πi/n then srn−k g is a distance preserving transformation
of C fixing 0, 1 and e2πi/n . Thus srn−k g = id and g = rk s.
So we have computed that
D2n = {rk , rk s|k = 0, . . . , n − 1}.

4we’ll be able to make precise why this assumption is reasonable later but it should at least
seem reasonable already.
5Recall that g(V ) := {g(v) | v ∈ V }.
8 SIMON WADSLEY

Thus it has 2n elements as required. We can compute for z ∈ C,


srk (z) = s(e2πik/n z) = e−2πik/n z̄ = r−k s(z).
Thus rk srk s = rk r−k ss = e and rk s is self-inverse. Geometrically rk s denote
reflection in the line through 0 and the point eπik/2n . To show this we can compute
rk s(0) = 0,
k
r s(1) = e2πik/n and
rk s(eπik/n) = eπik/n .
More generally if we wish to multiply elements in this standard form,
rk · rl = rk+n l ,
rk · rl s = rk+n l s,
rk s · rl = rk+n (−l) s and
rk s · rl s = rk+n (−l) .
In particular sr = rn−1 s = r−1 s. It would be instructive to reflect on the geometric
meaning of these equations. 
1.3.2. Symmetry groups of regular solids. Suppose that X is a regular solid in R3 .
We can consider Sym(X), the group of distance preserving transformations ρ of
R3 such that ρ(X) = X. These form a group. We will consider the cases X a
tetrahedron and X a cube later in the course.
1.3.3. The Symmetric group. We might hope that given any set X the set of in-
vertible functions from X to X forms a group under composition; that is the set of
functions f : X → X such that there is some g : X → X such that f ◦ g = id = g ◦ f .
This is true but not immediate: we need to check that composition of functions is
a binary operation on this set; that is that the composition of two invertible func-
tions is invertible. Some people would say that we need to check that the binary
operation is closed but ‘closure’ is built into our definition of binary operation.
Lemma. Suppose that f1 , f2 : X → X are invertible. Then f1 ◦ f2 : X → X is
invertible.
Proof. There are functions g1 , g2 : X → X such that f1 ◦ g1 = id = g1 ◦ f1 and
f2 ◦ g2 = id = g2 ◦ f2 . Then since ◦ is associative
(f1 ◦ f2 ) ◦ (g2 ◦ g1 ) = f1 ◦ id ◦g1 = id
and
(g2 ◦ g1 ) ◦ (f1 ◦ f2 ) = g2 ◦ id ◦f2 = id .

It follows that for every set X, the set S(X) = {f : X → X | f is invertible} is a
group under the composition of functions. It is called the symmetric group on X 6.
We call elements of the symmetric group permutations of X. If X = {1, . . . , n} we
write Sn instead on S(X). We will return to the groups Sn later in the course.

6The name comes from the fact that it can be viewed as the set of symmetries of the set X.
This is quite a subtle idea but you might like to think further about it when you come to revise
the course
GROUPS 9

1.4. Subgroups and homomorphisms. Sometimes when considering the sym-


metries of an object we want to restrict ourselves to considering symmetries that
preserve certain additional properties of the object. In fact we’ve already seen this,
the sets of distance preserving transformations of C and of R3 are both groups of
symmetries under composition. The groups D2n and Sym(X) for X a regular solid
are defined to consist of those symmetries that preserve a certain subset of the
whole space. Similary, instead of considering D2n , the group of all symmetries of
a regular n-gon we might want to restrict only to those symmetries that preserve
orientation, that is the rotations. This idea leads us to the notion of subgroup.
Definition. If (G, ◦) is a group then a subset H ⊂ G is a subgroup if ◦ restricts to
a binary operation on H 7 and (H, ◦) is a group. We write H 6 G to denote that
H is a subgroup of G.
Examples.
(1) Isom(Z) 6 S(Z).
(2) D2n 6 Isom(C) 6 S(C).
(3) Isom+ (Z) := {f : Z → Z | f (n) − f (m) = n − m for all n, m ∈ Z} 6 Isom(Z).
(4) Z is a subgroup of (Q, +).
(5) If H ⊂ D2n consists of all rotations of the n-gon then H is a subgroup.
(6) For any n ∈ Z, nZ := {an ∈ Z | a ∈ Z} is a subgroup of (Z, +).
(7) For every group G, {e} 6 G (the trivial subgroup) and G 6 G (we call a
subgroup H of G with H 6= G a proper subgroup).

Lecture 4
Lemma (Subgroup criteria). A subset H of a group G is a subgroup if and only if
the following conditions hold
(i) for every pair of elements h1 , h2 ∈ H, h1 h2 ∈ H;
(ii) the identity e ∈ H;
(iii) for every h ∈ H, h−1 ∈ H.
Proof. We must show that if conditions (i)-(iii) hold then H is a group. Condition
(i) tells us that the binary operation on G restricts to one on H. That this oper-
ation is associative follows immediately from the associativity of its extension to
G. Condition (ii) tells us that it has an identity8 and condition (iii) tells us H has
inverses9.
We must also show that if H is a subgroup of G then conditions (i)-(iii) hold.
That condition (i) holds is immediate from the definition of subgroup. If eH is
the identity in H then eH e = eH = eeH since e is the identity in G. But also
eH eH = eH since eH is the identity in H. Thus eH e = eH eH . But eH has an
inverse in G and we see that e−1 −1
H eH e = eH eH eH ie e = eH ∈ H i.e. condition (ii)
holds. That condition (iii) holds follows from uniqueness of inverses in G since for
h ∈ H the inverse of h in H will also be an inverse of h in G. 
Remark. Our subgroup criteria contain no mention of associativity since as noted
in the proof it is immediate from the associativity of the operation on G.
7precisely h ◦ h ∈ H for all h , h ∈ H
1 2 1 2
8the same identity as G since if h ∈ H then h ∈ G so eh = h = he
9the inverse h−1 of h ∈ H ⊂ G in G is also an inverse in H since hh−1 = e = h−1 h.
10 SIMON WADSLEY

There is an even shorter set of criteria for a subset to be a subgroup.


Corollary. A subset H of G is a subgroup precisely if it is non-empty and h−1
1 h2 ∈
H for all h1 , h2 ∈ H.
Proof. Suppose that conditions (i)-(iii) of the last lemma hold. Condition (ii) tells
us that H is non-empty and conditions (iii) and (i) combined give that h−1 1 h2 ∈ H
for all h1 , h2 ∈ H.
For the converse, since H is non-empty, there is some h ∈ H. By assumption e =
h−1 h ∈ H so condition (ii) holds. Now for h ∈ H, h−1 e = h−1 ∈ H by assumption
i.e. condition (iii) holds. Finally, for h1 , h2 ∈ H, h1 h2 = (h−1
1 )
−1
h2 ∈ H. 
Example. The set H = {f ∈ Isom(Z)|f (0) = 0} is a subgroup of Isom(Z). We can
see this using the corollary. Certainly id(0) = 0 so H 6= ∅. Moreover if h1 , h2 ∈ H
then
h−1 −1 −1
1 h2 (0) = h1 (0) = h1 h1 (0) = id(0) = 0.
Note that this argument isn’t much simpler than verifying conditions (i)-(iii) of the
lemma in practice.
We will also be interested in maps between groups. However we won’t typically
be interested an arbitary functions between two groups but only those that respect
the structure of the two groups. More precisely we make the following definition.
Definition. If (G, ◦) and (H, ∗) are two groups then θ : H → G is a group homo-
morphism (or just homomorphism) precisely if θ(h1 ∗ h2 ) = θ(h1 ) ◦ θ(h2 ) for all
h1 , h2 ∈ H.
Definition. A group homomorphism θ : H → G is an isomorphism if θ is invertible
as a function; ie if there is a function θ−1 : G → H such that θ ◦ θ−1 = idG and
θ−1 ◦ θ = idH .
Examples.
(1) If H 6 G then the inclusion map ι : H → G; h 7→ h is a group homomorphism.
It is not an isomorphism unless H = G.
(2) The function θ : Z → Zn such that θ(a) is the remainder after dividing a by n
is always a homomorphism from (Z, +) to (Zn , +n ) but never an isomorphism.
(3) If G is any group and g ∈ G is any element then θ : Z → G; n 7→ g n is a
homomorphism from (Z, +) to G. Indeed every homomorphism from (Z, +) to
G arises in this way.
(4) θ : Z → Isom+ (Z); n 7→ tn 10 is an isomorphism.
(5) The exponential function defines an isomorphism
exp : (R, +) → ({r ∈ R | r > 0}, ·); a 7→ ea .
The inverse map is given by log = loge .
If you are alert you will be asking why we don’t require homomorphisms θ : H →
G to satisfy θ(eH ) = eG and θ(h−1 ) = θ(h)−1 for all h ∈ H. The following lemma
shows that this is because these properties follow from our definition.
Lemma. Suppose that θ : H → G is a group homomorphism.
(i) θ(eH ) = eG .
(ii) For all h ∈ H, θ(h−1 ) = θ(h)−1 .
10recall t denotes translation by n
n
GROUPS 11

Proof. (i) Since θ is a homomorphism θ(eH ) = θ(eH eH ) = θ(eH )θ(eH ). Thus


eG = θ(eH )−1 θ(eH ) = θ(eH )−1 θ(eH )θ(eH ) = θ(eH ) as required.
(ii) Pick h ∈ H. Then θ(h)θ(h−1 ) = θ(hh−1 ) = θ(eH ) = eG (by (i)). Similarly
θ(h−1 )θ(h) = θ(h−1 h) = θ(eH ) = eG . Thus θ(h−1 ) = θ(h)−1 as required. 
Definition. If θ : H → G is a group homomorphism then the kernel of θ is defined
by
ker θ := {h ∈ H | θ(h) = eG }
and the image of θ is defined by
Im θ := θ(H).
Proposition. If θ : H → G is a homomorphism then ker θ is a subgroup of H and
Im θ is a subgroup of G.
Proof. We’ve seen θ(eH ) = eG so eH ∈ ker θ and eG ∈ Im θ.
Suppose that k1 , k2 ∈ ker θ. Then θ(k1 k2 ) = θ(k1 )θ(k2 ) = eG eG = eG , i.e.
k1 k2 ∈ ker θ. Similarly θ(k1−1 ) = θ(k1 )−1 = e−1 G = eG , i.e. k1
−1
∈ ker θ . So
ker θ 6 H.
Suppose now that θ(h1 ), θ(h2 ) ∈ Im θ. Then θ(h1 )θ(h2 ) = θ(h1 h2 ) so θ(h1 )θ(h2 ) ∈
Im θ. Similarly θ(h1 )−1 = θ(h−1
1 ), so θ(h1 )
−1
∈ Im θ. So Im θ 6 G. 

Lecture 5
Theorem (Special case of the isomorphism theorem). A group homomorphism
θ : H → G is an isomorphism if and only if ker θ = {eH } and Im θ = G. In this
case, θ−1 : G → H is a group homomorphism (and so also an isomorphism).
Proof. Suppose that θ has an inverse. Certainly eH ∈ ker θ and so θ−1 (eG ) =
θ−1 θ(eH ) = eH . Thus if k ∈ ker θ then k = θ−1 θ(k) = θ−1 (eG ) = eH and so
ker θ = {eH }. Moreover for each g ∈ G, θ−1 (g) ∈ H and g = θ(θ−1 (g)). Thus
G = Im θ.
Conversely, suppose that ker θ = eH and Im θ = G. By the latter property
for each g ∈ G we may choose an element hg ∈ H such that θ(hg ) = g. Now if
s : G → H is the function g 7→ hg then θs(g) = θ(hg ) = g for all g ∈ G. It follows
that for all h ∈ H,
θ(s(θ(h))h−1 ) = θ(s(θ(h)))θ(h−1 ) = θ(h)θ(h−1 ) = θ(hh−1 ) = θ(eH ) = eG ,
i.e. s(θ(h))h−1 ∈ ker θ = {eH }. It follows that sθ(h) = h for all h ∈ H and so s is
an inverse of θ.
Finally, suppose that θ is an isomorphism and g1 , g2 ∈ G. Then
θ(θ−1 (g1 g2 )) = g1 g2
and
θ(θ−1 (g1 )θ−1 (g2 )) = θ(θ−1 (g1 ))θ(θ−1 (g2 )) = g1 g2
since θ is a homomorphism. Thus
θ−1 (g1 g2 ) = θ−1 (θ(θ−1 (g1 g2 ))) = θ−1 (θ(θ−1 (g1 )θ(θ−1 (g2 )))) = θ−1 (g1 )θ−1 (g2 )
as required. 
Lemma. The composite of two group homomorphisms is a group homomorphism.
In particular the composite of two isomorphisms is an isomorphism.
12 SIMON WADSLEY

Proof. Suppose θ1 : H → G and θ2 : K → H are homomorpisms and k1 , k2 ∈ K


then θ1 (θ2 (k1 k2 )) = θ1 (θ2 (k1 )θ2 (k2 )) = θ1 θ2 (k1 )θ1 θ2 (k2 ) ie θ1 θ2 is a homomor-
phism. Finally, if θ1 and θ2 are invertible then θ2−1 θ1−1 is an inverse of θ1 θ2 . 

Definition. We say that a group G is cyclic if there is a homomorphism f : Z → G


such that Im f = G. Given such a homomorphism f we call f (1) a generator of G.

Note that G is cyclic with generator g if and only if every element of G is of the
form g i with i ∈ Z. More generally we say that a subset S of G generates G if
every element of G is a product of elements of S and their inverses — that is if G
the unique smallest subgroup of G containing S 11.

Examples.
(1) The identity map id : Z → Z; n 7→ n and the ‘reflection about 0’ map s : Z → Z;
n 7→ −n are both homomorphisms with image Z. Thus Z is cyclic and both 1
and −1 are generators. No other element generates Z.
(2) Zn is cyclic. In Numbers and Sets it is proven that an element of {0, 1, . . . , n−1}
generates Zn if and only if it is coprime to n12. The ‘if’ part is a conseqeunce
from Euclid’s algorithm; the only if part is elementary.

Lemma. Suppose that G is a group containing an element g with g n = e. There is


a unique group homomorphism f : Zn → G such that f (1) = g. In particular every
group of order n with an element of order n is isomorphic to Zn .

Proof. Suppose that f : Zn → G is a homomorphism such that f (1) = g. Then for


a = 0, 1, . . . n − 1, f (a +n 1) = f (a)f (1) = f (a)g. Thus we can see inductively that
f (a) = g a for all a ∈ Zn . Thus if f exists then it is unique.
We now see how to construct f . We define f (a) = g a for all a ∈ Zn and we
must prove that this defines a homomorphism. That is we must show that g a+n b =
f (a+n b) and g a+b = f (a)f (b) are equal for all a, b ∈ Zn . Since a+b−(a+n b) = kn
for some integer k and g n = e, we see that

g a+b g −(a+n b) = (g n )k = ek = e.

Thus g a+b = g a+n b as claimed.


Suppose now that G has order n and g ∈ G has order n. By the previous part
there is a homomorphism f : Zn → G such that f (a) = g a for each a ∈ Zn . Suppose
that f (a) = f (b) for a, b ∈ Zn . Then f (b −n a) = g a−n b = e thus a = b else g would
have order strictly smaller than n. It follows that ker f = {e} and | Im f | has n
elements and so must be the whole of G. Thus f is an isomorphism. 

Notation. We’ll write Cn for any group that is cyclic of order n. We’ve verified
that any two such groups are isomorphic.

11The curious will be reflecting on why G should have a unique smallest subgroup containing
S. Their reflections will do them good
12Recall that non-negative integers a, b are coprime if and only if their only common factor is
1.
GROUPS 13

Recall that we showed that D2n = {ri , ri s | i = 0, 1, . . . , n − 1} where r denotes


a rotation by 2π/n and s denotes a reflection. And that
rk · rl = rk+n l ,
rk · rl s = rk+n l s,
rk s · rl = rk+n (−l) s and
rk s · rl s = rk+n (−l) .
Lemma. Let n > 2 and suppose that G is a group containing elements g, h such
that g n = e, h2 = e and hgh−1 = g −1 . There is a unique group homomorphism
f : D2n → G such that f (r) = g and f (s) = h. Moreover if o(g) = n and |G| = 2n
then f is an isomorphism.
Proof. This is similar to the last proof. Any homomorphsim f : D2n → G such
that f (r) = g and f (s) = h must satisfy f (ri ) = g i and f (ri s) = g i h. So we must
show that this does define a homomorphism i.e. show that f (xy) = f (x)f (y) when
x, y ∈ {ri , ri s} = D2n . We’ll do the most difficult case and leave the rest as an
exercise. We can compute
f (rk s)f (rl s) = g k hg l h = g k (hgh−1 )l = g k g −l = g k+n (−l) = f (rk srl s)
as required.
For the last part suppose that o(g) = n and |G| = 2n. If ri ∈ ker f with
i ∈ {0, 1, . . . n − 1} then g i = e so as o(g) = n, i = 0. If ri s ∈ ker f then
g i h = e so h = g i . Then g −1 = hgh−1 = g i gg −i = g and so g 2 = e contradicting
o(g) = n. Thus ker f = {e}. This means that f can’t take the same value twice:
if f (x) = f (y) then f (xy −1 ) = e so x = y. That Im f = G now follows from the
pigeonhole principle. 

Lecture 6
1.5. The Möbius Group. Informally, a Möbius transformation is a function
f : C → C of the form
az + b
f : z 7→
cz + d
with a, b, c, d ∈ C and ad − bc 6= 0. The reason for the condition ad − bc 6= 0 is that
for such a function if z, w ∈ C then
az + b aw + b (z − w)
f (z) − f (w) = − = (ad − bc)
cz + d cw + d (cz + d)(cw + d)
so f would be constant if ad − bc were 0.13
Unfortunately the function is not well defined if z = −d/c since we may not
divide by zero in the complex numbers. This makes composition of Möbius trans-
formations problematic since the image of one Möbius transformation may not
coincide with the domain of defintion of another. We will fix this by adjoining an
additional point to C called ∞.14
Notation. Let C∞ := C ∪ {∞} which we call the extended complex plane.

13This is bad because we’re really interested in invertible functions


14And pronounced ‘infinity’.
14 SIMON WADSLEY

Definition. Given (a, b, c, d) ∈ C4 such that ad − bc 6= 0 we can define a function


f : C∞ → C∞ as follows:
if c 6= 0 then  az+b
 cz+d if z ∈ C\{−d/c};
f (z) := ∞ if z = −d/c;
a/c if z = ∞;

if c = 0 then  az+b
cz+d if z ∈ C
f (z) :=
∞ if z = ∞.
We call all functions from C∞ to C∞ that arise in this way Möbius transformations
and let
M := {f : C∞ → C∞ | f is a Möbius transformation}.
We’ll see another way to interpret this definition later in the course involving
projective geometry. But for now we’ll work with it as it stands and also take for
granted a result that will will prove later.15
Theorem. The set M defines a subgroup of S(C∞ ).
Lemma. Suppose that f ∈ M such that f (0) = 0, f (1) = 1 and f (∞) = ∞. Then
f = id.
Proof. If f = az+b
cz+d then f (∞) = ∞ forces c = 0 and f (0) = 0 forces b = 0 and then
f (1) = 1 forces a/d = 1 and so f : z 7→ az+0
0z+a = z for all z ∈ C as required. 
Theorem (Strict triple transitivity of Möbius transformations). If (z1 , z2 , z3 ) and
(w1 , w2 , w3 ) are two sets of three distinct points then there is a unique f ∈ M such
that f (zi ) = wi for i = 1, 2 and 3.
Proof. First we consider the case w1 = 0, w2 = 1 and w3 = ∞.
(z2 −z3 )(z−z1 )
If none of z1 , z2 , z3 are ∞ then f (z) = (z2 −z1 )(z−z3 )
does the job.
1
If some zi = ∞ then choose z4 ∈ C∞ \{z1 , z2 , z3 } and let s(z) = z−z 4
then
(s(z1 ), s(z2 ), s(z3 )) does not contain ∞ so by the above we may find g ∈ M such
that g(s(z1 )) = 0, g(s(z2 )) = 1 and g(s(z3 )) = ∞. Then f = gs does the job.
In the general case we can find g, h ∈ M such that g(z1 ) = 0, g(z2 ) = 1,
g(z3 ) = ∞ and h(w1 ) = 0, h(w2 ) = 1 and h(w3 ) = ∞. Then f = h−1 g(zi ) = wi for
i = 1, 2, 3 and we’ve shown existence in general. Moreover if k is another such map
then hkg −1 fixes 0, 1 and ∞ so by the lemma is the identity. Thus k = h−1 g = f
and we’ve shown uniqueness. 
Definition. Given distinct points z1 , z2 , z3 , z4 ∈ C∞ the cross-ratio of z1 , z2 , z3 , z4
written [z1 , z2 , z3 , z4 ] := f (z4 ) where f is the unique Möbius transformation such
that f (z1 ) = 0, f (z2 ) = 1 and f (z3 ) = ∞.16
(z4 −z1 )(z2 −z3 ) 17
Lemma. If z1 , z2 , z3 , z4 ∈ C then [z1 , z2 , z3 , z4 ] = (z2 −z1 )(z4 −z3 ) .

15There is a straight-forward if slightly fiddly way to prove it directly that demands care with
the point ∞. We will give a slightly more sophisticated but less fiddly proof.
16There are 6 essentially different definitions of cross-ratio depending on how we order 0, 1
and ∞ in this definition. It doesn’t really matter which we choose as long as we are consistent.
17We could’ve defined the cross-ratio by this formula but we’d need to be more careful when
some zi = ∞.
GROUPS 15

2 −z3 )(z−z1 )
Proof. We’ve already computed that f (z) = (z (z2 −z1 )(z−z3 ) is the unique Möbius
transformation such that f (z1 ) = 0, f (z2 ) = 1 and f (z3 ) = ∞ and so
[z1 , z2 , z3 , z4 ] = f (z4 )
is given by the required formula. 

Theorem (Invariance of Cross-Ratio). For all z1 , z2 , z3 , z4 ∈ C∞ and g ∈ M,


[g(z1 ), g(z2 ), g(z3 ), g(z4 )] = [z1 , z2 , z3 , z4 ].
Proof. Suppose that f (z1 ) = 0, f (z2 ) = 1 and f (z3 ) = ∞. Then (f g −1 )(g(z1 )) = 0,
(f g −1 )(g(z2 )) = 1 and (f g −1 )(g(z3 )) = ∞. Thus
[g(z1 ), g(z2 ), g(z3 ), g(z4 )] = f g −1 (g(z4 )) = f (z4 ) = [z1 , z2 , z3 , z4 ]
as required. 

Proposition. Every element of M is a composite of Möbius transformations of


the following forms.
(a) Da : z 7→ az = az+0
0z+1 with a ∈ C\{0} (rotation/dilations);
1z+b
(b) Tb : z 7→ z + b = 0z+1 with b ∈ C (translations);
0z+1
(c) S : z 7→ 1/z = 1z+0 (inversion).

Lecture 7
Proof. We’ll give two proofs. First a ‘nailing to the wall’ type argument.
Suppose f ∈ M. If f (∞) ∈ C then ST−f (∞) f (∞) = S(0) = ∞. So, by replacing
f by ST−f (∞) f if necessary, without loss of generality f (∞) = ∞18.
Now since f (∞) = ∞, f (0) 6= ∞. Thus T−f (0) exists, T−f (0) f (0) = 0 and
T−f (0) (∞) = ∞ so we may assume f fixes both 0 and ∞.
Now since f (∞) = ∞ and f (0) = 0, f (1) ∈ C\{0}. Thus D1/f (1) exists and
D1/f (1) (1) = 1, D1/f (1) (0) = 0 and D1/f (1) (∞) = ∞. Thus f = Df (1) and we’re
done.
(bc−ad)
Alternatively if c 6= 0 then az+b a
cz+d = c + c(cz+d) so f = Ta/c D(bc−ad)/c STd Dc . If
c = 0 then f = Tb/d Da/d . 

Definition. A circle in C∞ is a subset that is either of the form {z ∈ C | |z−a| = r}


for some a ∈ C and r > 019 or of the form {z ∈ C | aRe(z) + bIm(z) = c} ∪ {∞}
for some a, b, c ∈ R with (a, b) 6= (0, 0)20
It follows that any three distinct points in C∞ determine a unique circle in C∞ .
Lemma. The general equation of a circle in C∞ is
Az z̄ + B z̄ + B̄z + C = 0
with A, C ∈ R, B ∈ C and AC < |B|2 .21

18Since if we can prove that ST


−f (∞) f is a product of the special maps then f can be obtained
by postcomposing this product with Tf (∞) S
19i.e. a usual circle in C
20i.e. a line in C together with ∞
21where ∞ is understood to be a solution of this equation precisely if A = 0
16 SIMON WADSLEY

Proof. First consider the case A 6= 0. Then


Az z̄ + B z̄ + B̄z + C = 0 ⇐⇒ |z + B/A|2 = |B|2 /A2 − C/A.
This latter defines a usual circle if and only if |B|2 /A2 > C/A i.e. |B|2 > AC.
Next consider the case A = 0. Then
B z̄ + B̄z + C = 0 ⇐⇒ 2Re(B)Re(z) + 2Im(B)Im(z) = −C.
The latter defines a line in C if and only if B 6= 0 ie |B|2 > AC = 0. 
Theorem (Preservation of circles). If f ∈ M and C is a circle in C∞ then f (C)
is a circle in C∞ .
Proof. If f is a rotation/dilation or translation the result is easy. Thus in light of
the last proposition we may assume that f : z 7→ 1/z.
But z is a solution to Az z̄ + B z̄ + B̄z + C if and only if w = 1/z is a solution to
Cww̄ + Bw + B̄ w̄ + A. 
Corollary. Four distinct points z1 , z2 , z3 and z4 in C∞ lie on a circle if and only
if [z1 , z2 , z3 , z4 ] ∈ R.
Proof. Using triple transitivity we can find f ∈ M such that f (z1 ) = 0, f (z2 ) = 1
and f (z3 ) = ∞. By preservation of circles z1 , z2 , z3 and z4 lie on a circle if and only
if f (z1 ), f (z2 ), f (z3 ) and f (z4 ) lie on a circle i.e. if and only if f (z4 ) is real. But
[z1 , z2 , z3 , z4 ] = f (z4 ). 
Remark. It is possible to prove the corollary directly and then use a similar argu-
ment to deduce that Möbius transformations preserve of circles from it.
Definition. Given two elements x, y of a group G we say y is conjugate to x if
there is some g ∈ G such that y = gxg −1 .
Note that if y = gxg −1 then x = g −1 y(g −1 )−1 so the notion of being conjugate
is symmetric in x and y. Morever if also z = hyh−1 then z = (gh)x(gh)−1 so if z
is conjugate to y and y is conjugate to x then z is conjugate to x.
Proposition. Every Möbius transformation f except the identity has precisely one
or two fixed points. If f has precisely one fixed point it is conjugate to the translation
z 7→ z + 1. If f has precisely two fixed points it is conjugate to a map of the form
z 7→ az with a ∈ C\{0}.
Proof. Let f be the map extending z 7→ az+b cz+d and assume that f 6= id.
Suppose first that c = 0. In this case ∞ is a fixed point. Moreover for z ∈ C, f
fixes z if and only if dz = az + b. The possible sets of solutions to this equation are
{b/(d − a)} (when d 6= a), C (when d = a and b = 0), or ∅ (when d = a and b 6= 0).
Thus when c = 0 and f has either one or two fixed points.
Next suppose that c 6= 0 then f does not fix ∞. For z ∈ C, z = f (z) if and only
if cz 2 + dz − az − b = 0. This is a quadratic equation and the coefficient of z 2 is
non-zero so it has one or two solutions (two unless there is a repeated root). Thus
again f has 1 or 2 fixed points in this case.
Now suppose that f has precisely one fixed point w and pick x ∈ C that is not
fixed by f . Then f (x) 6∈ {x, w}.22 Let g be the unique Möbius transformation such
that g(w) = ∞, g(x) = 0 and g(f (x)) = 1. Now for z ∈ C∞ , gf g −1 (z) = z if and
22If f (x) = w then x = f −1 (w) = w
GROUPS 17

only if f g −1 (z) = g −1 (z), i.e. if and only if g −1 (z) is fixed by f . Since w is the only
fixed point of f , ∞ = g(w) is the only fixed point of gf g −1 . Thus by the discussion
above gf g −1 extends z 7→ 0z+a az+b
= z + b/a for some a, b ∈ C\{0}. Since

gf g −1 (0) = gf (x) = 1
we see that b/a = 1 as required.
Suppose instead that f has two distinct fixed points z1 and z2 . Let g be any
Möbius transformation such that g(z1 ) = 0 and g(z2 ) = ∞. Then
gf g −1 (0) = gf (z1 ) = g(z1 ) = 0
and
gf g −1 (∞) = gf (z2 ) = g(z2 ) = ∞.
Thus gf g −1 (z) = az for some a ∈ C∞ . 

Remark. Suppose that f ∈ M. If gf g −1 : z 7→ z + 1 then for each n > 0,


gf n g −1 = (gf g −1 )n : z 7→ z + n
so f n (z) = g −1 (g(z)+n) for z ∈ C∞ not fixed by f . Similarly if gf g −1 : z 7→ az then
for n > 0, f n (z) = g −1 (an g(z)) for z not fixed by f . Thus we can use conjugation
to compute iterates of f ∈ M in a simple manner.

Lecture 8
2. Lagrange’s Theorem
2.1. Cosets.
Definition. Suppose that (G, ◦) is a group and H is a subgroup. A left coset of
H in G is a set of the form g ◦ H := {g ◦ h | h ∈ H} for some g ∈ G. Similarly a
right coset of H in G is a set of the form H ◦ g := {hg | h ∈ H} for some g ∈ G.
We write G/H to denote the set of left cosets of H in G and H\G to denote the
set of right cosets of H in G23.
As usual we will often suppress the ◦ and write gH or Hg.
Examples.
(1) Suppose n ∈ Z, so that nZ := {an | a ∈ Z} is a subgroup of (Z, +). Then
0 + nZ = nZ = n + nZ. 1 + nZ = {1 + an | a ∈ Z} = (1 − n) + nZ. More
generally, b + nZ is the set of integers x such that x − b is a multiple of n.24
(2) Suppose that G = D6 = {e, r, r2 , s, rs, r2 s} and H = {e, s} then
eH = {e, s} = sH
rH = {r, rs} = rsH
r2 H = {r2 , r2 s} = r2 sH.

23This latter is a little unfortunate in the \ is normally used to denote set-theoretic difference
but this should not cause confusion. Why?
24Note that because the operation on Z is addition we don’t suppress it when we name cosets,
i.e. we write a + nZ rather than anZ because the latter would create confusion.
18 SIMON WADSLEY

However,
He = {e, s} = Hs
Hr 2
= {r, r s} = Hr2 s
Hr2 = {r2 , rs} = Hrs.
Thus left cosets and right cosets need not agree when the group is not abelian.
However if K = {e, r, r2 } 6 D6 then K = eK = rK = r2 K = Ke = Kr = Kr2
and {s, rs, r2 s} = sK = rsK = r2 sK = Ks = Krs = Kr2 s. So in this case
the left and right cosets are the same.
(3) Suppose that M is the Möbius group and H = {f ∈ M | f (0) = 0}. Then, for
g ∈ M,
gH = {f ∈ M | f (0) = g(0)} whereas
Hg = {f ∈ M | f −1 (0) = g −1 (0)}.
We’ll return to this idea later in the course.
2.2. Lagrange’s Theorem.
Theorem (Lagrange’s Theorem). Suppose that G is a group and H is a subgroup
of G then the left cosets of H in G partition G. In particular if G is finite then |H|
divides |G|.
Proof. To show that the left cosets of H in G partition G we must show (i) that
every element of G lives inside some left coset; and (ii) if two left cosets have a
common element then they are equal.
For (i) notice that for all g ∈ G, g ∈ gH since e ∈ H and ge = g.
For (ii) suppose that g ∈ g1 H ∩ g2 H. Then we may find h1 , h2 ∈ H such that
g = g1 h1 = g2 h2 . Then for all h ∈ H, g1 h = gh−1 −1 −1
1 h = g2 h2 h1 h. Since h2 h1 h ∈
H, we see that g1 H ⊂ g2 H. By symmetry we may conclude that g2 H ⊂ g1 H and
so g1 H = g2 H as required.
Now for the last part we observe that when PG is finite, H must also be finite
since it is a subset of G. Moreover, |G| = gH∈G/H |gH|. Now for each left
coset gH there is an invertible function lg : H → gH; h 7→ gh — the function
lg−1 is given by gh 7→ g −1 gh = h. Thus |gH| = |H| for every left coset gH and
|G| = |G/H| · |H|. 
Remark. By a very similar argument the right cosets of H in G also partition G.
Corollary. If G is a finite group, then every element of G has order dividing |G|.
Proof. Suppose g ∈ G and let f : Z → G be the homomorphism defined by f (n) =
g n . We claim that the order of g is precisely | Im f | which divides |G| by Lagrange.
To prove the claim we first observe that Im f = {e, g, g 2 , . . . , g o(g)−1 } since if n ∈ Z
there are integers q, r with 0 6 r < o(g) and n = qo(g) + r and then g n =
(g o(g) )q g r = g r . Moreover the elements e, g, . . . , g o(g)−1 are distinct since if g i = g j
with 0 6 i < j < o(g) then g j−i = e contradicting the definition of o(g). 
Proposition. Suppose that p is prime. Then every group of order p is isomorphic
to Cp .
Proof. Let G be a group of order p and g ∈ G\{e}. Since |G| = p and o(g) divides
|G| and is not 1 we must have o(g) = p. Thus g generates G by counting and an
earlier lemma tells us that G is isomorphic to Cp . 
GROUPS 19

2.3. Groups of order at most 8. In this section we will classify all groups of
order at most 8 under the perspective that two groups are the same precisely if
they are isomorphic. We’ve already seen that every group of order 2, 3, 5 or 7 is
isomorphic to the cyclic group of the same order. It is evident that the trivial group
is the only group of order 1 up to isomorphism.
Before we begin this we’ll need the following construction that enables us to
build new groups from old ones.
Example. Suppose that G and H are groups. We can define a binary operation on
G × H via (g1 , h1 )(g2 , h2 ) = (g1 g2 , h1 h2 ) for g1 , g2 ∈ G and h1 , h2 ∈ H. We claim
that this makes G × H into a group.

Proof of claim. Since


((g1 , h1 )(g2 , h2 ))(g3 , h3 ) = (g1 g2 g3 , h1 h2 h3 ) = (g1 , h1 )((g2 , h2 )(g3 , h3 ))
for all g1 , g2 , g3 ∈ G and h1 , h2 , h3 ∈ H, the operation on G × H is associative.
Since (eG , eH )(g, h) = (g, h) = (g, h)(eG , eH ), (eG , eH ) is an identity for the
operation on G × H.
Finally since (g −1 , h−1 )(g, h) = (eG , eH ) = (g, h)(g −1 , h−1 ) the operation on
G × H has inverses. 

Exercise. Show that if G1 , G2 and G3 are groups then G1 × G2 is isomorphic to


G2 × G1 and (G1 × G2 ) × G3 is isomorphic to G1 × (G2 × G3 ).

Lecture 9
Theorem (Direct Product Theorem). Suppose that H1 , H2 6 G such that
(i) H1 ∩ H2 = {e};
(ii) if h1 ∈ H1 and h2 ∈ H2 then h1 and h2 commute;
(iii) for all g ∈ G there are h1 ∈ H1 and h2 ∈ H2 such that g = h1 h2 .
Then there is an isomorphism H1 × H2 → G.
Proof. Let f : H1 × H2 → G be given by f (h1 , h2 ) = h1 h2 . Now if h1 , k1 ∈ H1 and
h2 , k2 ∈ H2 then
f ((h1 , k1 )(h2 , k2 )) = f ((h1 h2 , k1 k2 )) = h1 h2 k1 k2
and
f ((h1 , k1 ))f ((h2 , k2 )) = h1 k1 h2 k2 .
Since k1 and h2 commute (by (ii)), we see that f is a homomorphism. Now if
(h1 , h2 ) ∈ ker f then h1 h2 = e so h1 = h−1
2 ∈ H1 ∩ H2 = {e} (by (i)). Thus
ker f = {e}. Finally Im f = {h1 h2 | h1 ∈ H1 , h2 ∈ H2 } = G (by (iii)) and so f is
the required isomorphism. 

We also need the following result that also appeared on the first example sheet.
Lemma. If G is a group such that every non-identity element has order two25 then
G is abelian.

25Of course in any group the identity has order 1


20 SIMON WADSLEY

Proof. Suppose that x, y ∈ G. We must show that xy = yx.


Note that for all g ∈ G, g 2 = e. So by uniqueness of inverses every element of
G is self-inverse, i.e. g = g −1 . In particular (xy)−1 = xy. By the shoes and socks
lemma it follows that y −1 x−1 = xy. Since x−1 = x and y −1 = y, yx = xy as
required. 
We also recall that every a group of order n with an element of order n is
isomorphism to Cn and that every group of order 2n that has an element g of order
n and an element h of order 2 such that hg = g −1 h is isomorphic to D2n .
Proposition. Every group of order 4 is isomorphic to precisely one of C4 and
C2 × C2 .
Proof. First we’ll show that C4 and C2 × C2 are not isomorphic. Suppose (g, h) ∈
C2 × C2 . Then (g, h)2 = (e, e). So every element of C2 × C2 has order 2 but C4 has
an element of order 4 thus they cannot be isomorphic.26
Now suppose that G is any group of order 4. If G contains an element of order 4
then it is isomorphic to C4 so suppose not. By Lagrange every non-identity element
of G has order 2. Let g, h be two distinct such elements. Let H1 = {e, g} ' C2 and
H2 = {e, h} ' C2 . It suffices to show that the three conditions of the direct product
theorem hold. Condition (i) is immediate since g 6= h. Condition (ii) follows from
the last lemma since it tells us that G must be abelian. Finally, since g 6∈ H2 ,
gH2 6= H2 and G is partitioned by H2 and gH2 by Lagrange. Thus condition (iii)
holds. 
Proposition. Every group of order 6 is isomorphic to precisely one of C6 or D6 .
Proof. We can easily see that C6 and D6 are not isomorphic since C6 is abelian
and D6 is not27.
Now suppose that G is any group of order 6. If G contains an element of order 6
then it is isomorphic to C6 so suppose not. By Lagrange every non-identity element
of G has order 2 or 3. If there were no element of order 3 then any two non-identity
elements would generate a subgroup of order 4 contradicting Lagrange. Thus G
contains an element g of order 3. Let K = {e, g, g 2 } be the subgroup of G generated
by g which is isomorphic to C3 . By Lagrange for any h 6∈ K, G is partitioned by
K and hK i.e. G = {e, g, g 2 , h, hg, hg 2 }. Now consider h2 : h2 6∈ hK else h ∈ K 28.
Thus h2 ∈ K. If h2 is g or g 2 then h has order 6 contradicting our assumption that
G has no elements of order 6. Thus h2 = e.
By the right coset version of Lagrange, G is also partitioned by K and Kh thus
Kh = hK and {h, gh, g 2 h} = {h, hg, hg 2 }. So gh ∈ {hg, hg 2 }. If gh = hg then
(gh)2 = g 2 and (gh)3 = h. Thus gh does not have order 1, 2 or 3, a contradiction29.
Thus gh = hg −1 And we can deduce that G ' D6 . 
Example. The following set of matrices form an non-abelian group Q8 of order 8
        
1 0 i 0 0 1 0 i
± ,± ,± ,± .
0 1 0 −i −1 0 i 0
26If f : Z → C ×C were an isomorphism then f (2) = f (1)2 = e giving f a non-trivial kernel.
4 2 2
27If f : D → C were an isomorphism then for all g , g ∈ D , f (g g ) = f (g )f (g ) =
6 6 1 2 6 1 2 1 2
f (g2 )f (g1 ) = f (g2 g1 ) so g2 g1 = g1 g2 .
28if h2 = hk then h = k
29had we not made the assumption about no elements of order 6 then in this case we’d get
that gk has order 6 so G ' C6
GROUPS 21

It is common to write
       
1 0 i 0 0 1 0 i
1= ,i = ,j = and k = .
0 1 0 −i −1 0 i 0
Then Q8 = {±1, ±i, ±j, ±k}. Then we can compute that 1 is an identity, −1 has
order 2 commutes with everything and multiplies as you’d expect given the notation.
Morever i, j and k all have order 4 (all of them square to −1), ij = k = −ji,
ki = j = ik and jk = i = −kj.
Exercise. Verify that Q8 is a group.

Lecture 10
Proposition. Every group of order 8 is isomorphic to precisely one of C8 , C4 ×C2 ,
C2 × C2 × C2 , D8 or Q8 .
Proof. Let G be a group of order 8.
If G has an element of order 8 then G ' C8 .
If every non-identity element of G has order 2 then G is abelian. Moreover any
two non-identity elements g1 , g2 generate a subgroup H1 of G of order 4 necessarily
isomorphic to C2 × C2 .30 If g3 ∈ G\H1 then H2 = {e, g3 } is a subgroup of order 2.
One can easily verify that G ' H1 × H2 ' C2 × C2 × C2 in a similar fashion to the
argument above for C2 × C2 .31
So we’re reduced to classifying groups G of order 8 with no element of order 8
and at least one element g of order 4. In this setting the set K = {e, g, g 2 , g 3 } is a
subgroup of G isomorphic to C4 .
Let h ∈ G\K. Then G = K∪hK by Lagrange so G = {e, g, g 2 , g 3 , h, hg, hg 2 , hg 3 }.
Moreover h2 ∈ K else h2 ∈ hK whence h ∈ K.
If h2 = g or h2 = g −1 then h has order 8 contradicting our assumption that G
has no such elements. So there are two cases remaining: case A where h2 = e and
case B where h2 = g 2 .
Suppose that we’re in case A and h2 = e. Then consider the product gh. By
Lagrange again, gh ∈ hK = {h, hg, hg 2 , hg 3 }. If gh = hg i then h−1 gh = g i and
2
g = h−2 gh2 = h−1 (h−1 gh)h = h−1 g i h = (h−1 gh)i = (g i )i = g i .32
Thus i = 1 or 3. If i = 1 then G is abelian and we can use the direct product
theorem to show that G ' K × {e, h} ' C4 × C2 . If instead i = 3 then G has an
element g of order 4 and an element h of order 2 such that gh = hg −1 so as G has
order 8 we know that G ' D8 .
Finally suppose we’re in case B and h2 = g 2 . Again consider the product gh ∈
{h, gh, g 2 h, g 3 h} = Kh = hK = {h, hg, hg 2 , hg 3 }. Since g 2 h = h3 = hg 2 , gh = hg
or gh = hg 3 . If gh = hg then
(gh)2 = g 2 h2 = g 4 = e
so {e, gh} 6 G and G ' K × {e, gh} ' C4 × C2 by the direct product theorem.

30the elements are e, g , g and g g it is easy to check that these are distinct and form a
1 2 1 2
subgroup.
31See also Example Sheet 1 Q14
32since h2 = e
22 SIMON WADSLEY

If gh = hg 3 then do some renaming. Write 1 := e, −1 := h2 = g 2 , i := g, j := h,


k := gh. Then
G = {±1, ±i, ±j, ±k}
and we can verify that multiplication is as for Q8 .
Exercise. Complete the proof by showing that no two of the listed groups are
isomorphic. 
2.4. The Quaternions.
Definition. The quaternions are the set of matrices
H := {a1 + bi + cj + dk | a, b, c, d ∈ R} ⊂ Mat2 (C)
where as before
       
1 0 i 0 0 1 0 i
1= ,i = ,j = and k = .
0 1 0 −i −1 0 i 0
The sum or product of two elements of H lives in H and + and · obey the
same associativity and distributivity laws as Q, R and C with identities 0 and 1
respectively. Although the multiplication in H is not commutative (since ij = −ji),
(H, +) is an abelian group and (H\{0}, ·) is a (non-abelian) group.33 To see the
latter we can define ‘quaternionic conjugation’ by
(a1 + bi + cj + dk)∗ = a1 − bi − cj − dk
and then verifying that if x = a1 + bi + cj + dk then
xx∗ = x∗ x = (a2 + b2 + c2 + d2 )1
so x−1 = 1
a2 +b2 +c2 +d2 x

for x 6= 0.
2.5. Fermat–Euler theorem. We can define a multiplication operation on the
set Zn = {0, 1, . . . , n − 1} by setting a ·n b to be the remainder after dividing ab by
n.
Definition. Let Un := {a ∈ Zn | ∃b ∈ Zn s.t. a ·n b = 1} be the set of invertible
elements of Zn with respect to ·n .
It is a result from Numbers and Sets that follows from Euclid’s algorithm that
|Un | = ϕ(n) where ϕ(n) denotes the number of elements of Zn coprime to n. Indeed
Un = {a ∈ Zn | (a, n) = 1}.
Lemma. (Un , ·n ) is an abelian group.
Proof. ·n defines an associative and commutative operation on Zn with an identity
1.34 If a1 , a2 ∈ Un then there are b1 , b2 ∈ Un such that ai ·n bi = 1. Then
(a1 ·n a2 ) ·n (b1 ·n b2 ) = (a1 ·n b1 ) ·n (a2 ·n b2 ) = 1 ·n 1 = 1
so ·n restricts to an associative binary operation on Un . 1 ∈ Un is an identity and
if a ∈ Un then there is b ∈ Zn with a ·n b = b ·n a = 1. Then b ∈ Un and b = a−1 .
So Un has inverses. 
Theorem (Fermat–Euler Theorem). If (a, n) = 1 then aϕ(n) ≡ 1 mod n.
33We say that H is a division algebra. R, C and H are the only division algebras that are finite
dimensional as vector spaces over R.
34for a, b, c ∈ Z , a · (b · c) ≡ abc ≡ (a · b) · c mod n
n n n n n
GROUPS 23

Proof. Since (Un , ·n ) is a group of order ϕ(n) it is consequence of Lagrange’s the-


orem that every element of (Un , ·n ) has order dividing ϕ(n). The result follows
immediately. 

Lecture 11
3. Group Actions
We started the course by saying that groups are fundamentally about symmetry
but the connection has been opaque for the last three lectures. In this chapter we
will discuss how to recover the notion of symmetry from the group axioms.
3.1. Definitions and examples.
Definition. An action of a group G on a set X is a function
· : G × X → X; (g, x) 7→ g · x
such that for all x ∈ X
(i) e · x = x;
(ii) g · (h · x) = (gh) · x for all g, h ∈ G.
Examples.
(1) Isom(Z) acts on Z via f · n = f (n).
(2) The Möbius group M acts on the extended complex plane C∞ via f · z = f (z).
(3) Generalising both the examples above, if H 6 S(X) then H acts on X via
h · x = h(x). We call this the natural action of H on X.
(4) M also acts on the set of circles in C∞ . We proved in §1.5 that if f ∈ M
and C ⊂ C∞ is a circle then f (C) ⊂ C∞ is also a circle so (f, C) 7→ f (C) is a
function. Moreover for all circles C the conditions id(C) = C and f (g(C)) =
(f g)(C) for f, g ∈ M are both clear.
(5) D2n acts on the set of points a regular n-gon. D2n also acts on the set of
vertices of a regular n-gon and on the set of edges of a regular n-gon.
(6) If X is a regular solid then Sym(X) acts on the set of points (and on the sets
of vertices/edges/faces) of X.
(7) If H 6 G then G acts on G/H, the set of left cosets of G in H via g · kH = gkH
for g, k ∈ G. To see this we need to check that if kH = k 0 H then gkH = gk 0 H.
But if kH = k 0 H then k 0 = kh for some h ∈ H so gk 0 = gkh ∈ gk 0 H ∩ gkH and
gkH = gk 0 H by Lagrange. Given this we see that for all k ∈ G, ekH = kH
and g1 (g2 kH) = (g1 g2 )kH for all g1 , g2 ∈ G.
(8) For any group G and set X we can define the trivial action via g · x = x for all
g ∈ G and x ∈ X.
Theorem. For every group G and set X there is a 1 − 1 correspondence
{actions of G on X} ←→ {θ : G → S(X) | θ is a homomorphism}
such that an action · : G × X → X corresponds to the homomorphism θ : G → S(X)
given by θ(g)(x) = g · x.
Proof. First we must show that if · : G × X → X is a action then the formula
θ(g)(x) = g · x does define a homomorphism G → S(X).
For each g ∈ G, define θ(g) : X → X via θ(g)(x) = g · x. So that
θ : G → {f : X → X}.
24 SIMON WADSLEY

Then for g, h ∈ G we can compute


θ(gh)(x) = (gh) · x = g · (h · x) = θ(g) ◦ θ(h)(x)
ie θ(gh) = θ(g)θ(h). So to show that θ defines a homomorphism G → S(X) it
suffices to show that θ(g) ∈ S(X) for each g ∈ G; i.e. that each θ(g) is invertible.
But
θ(g)θ(g −1 ) = θ(e) = θ(g −1 )θ(g)
so to show this it suffices to show that θ(e) is idX . But θ(e)(x) = e·x = x = idX (x).
So we’re done.
To finish we must show that every homomorphism θ : G → S(X) arises like this
in precisely one way. So suppose that θ is any such homomorphism, θ corresponds
to an action · : G × X → X precisely if g · x = θ(g)(x) defines an action. So it
remains to check that e · x = x and g · (h · x) = (gh) · x for all x ∈ X and g, h ∈ G.
But
e · x = θ(e)(x) = idX (x) = x35
and
g · (h · x) = θ(g)(θ(h)(x)) = θ(gh)(x) = (gh) · x
as required. 
Definition. We say that an action of G on X is faithful if the kernel of the corre-
sponding homomorphism G → S(X) is the trivial group.
3.2. Orbits and Stabilisers.
Definition. Suppose a group G acts on a set X and that x ∈ X. The orbit of x
under the action is given by
OrbG (x) := {g · x | g ∈ G} ⊂ X.
The stabiliser of x under the action is given by
StabG (x) := {g ∈ G | g · x = x} ⊂ G.
T
Thus an action is faithful precisely if x∈X StabG (x) = {e}.
Examples.
(1) Under the natural action of Isom(Z) on Z, for all n ∈ Z
OrbIsom(Z) (n) = Z
and
StabIsom(Z) = {id, m 7→ 2n − m}
(2) Under the natural action of M on C∞ , for all z ∈ C∞
OrbM (z) = C∞
and  
az + b
StabM (∞) = z 7→| ad 6= 0 .
0c + d
(3) Under the action of D2n on the set of points of a regular n-gon the orbit of a
vertex of the n-gon is the set of all vertices of the n-gon and the stabliser of a
vertex consists of the identity and reflection in the line through the centre of
the n-gon and the vertex.36
35Since θ is a homomorphism it must send the identity of G to the identity of S(X).
36It would be instructive to think about what the orbits and stabilisers of other points of n-gon
are under the action of D2n .
GROUPS 25

(4) For the left coset action of G on G/H defined earlier


OrbG (eH) = G/H
and
StabG (eH) = {g ∈ G | gH = eH} = H.
More generally
StabG (kH) = {g ∈ G | gkH = kH} = {g ∈ G | k −1 gkH = H} = kHk −1 .
(5) For the trivial action of G on X and any x ∈ X,
OrbG (x) = {x} and StabG (x) = G.
Lemma. Suppose that G is a group acting on a set X.
(i) Each stabiliser StabG (x) is a subgroup of G.
(ii) The orbits OrbG (x) partition X. In particular if X is finite and the distinct
orbits are O1 , . . . , Om then
m
X
|X| = |Oi |
i=1

Lecture 12
Proof. (i) Let x ∈ X. Since e · x = x, e ∈ StabG (x). Suppose that g, h ∈ StabG (x).
Then
(gh) · x = g · (h · x) = g · x = x
so gh ∈ StabG (x) and
x = e · x = (g −1 g) · x = g −1 · (g · x) = g −1 · x
so g −1 ∈ StabG (x). Thus StabG (x) 6 G as required.37
(ii) Since for any x ∈ X, x = e · x so x ∈ OrbG (x), it suffices to prove that
for any x, y ∈ X either OrbG (x) = OrbG (y) or OrbG (x) ∩ OrbG (y) = ∅. Suppose
that z ∈ OrbG (x) ∩ OrbG (y). Then there are some elements g, h ∈ G such that
g · x = z = h · y. Thus x = (g −1 h) · y. Now if w ∈ OrbG (x) then there is some
f ∈ G such that w = f · x. Whence we can compute w = (f g −1 h) · y and so
OrbG (x) ⊂ OrbG (y). By symmetry OrbG (y) ⊂ OrbG (x). Now we can see that
OrbG (x) = OrbG (y) as required.38 
Definition. We say that an action of G on X is transitive if there is only one orbit
i.e. if X = OrbG (x) for any x ∈ X.
Theorem (Orbit-Stabiliser Theorem). Suppose a group G acts on a set X and
x ∈ X. There is a (natural) invertible function
G/ StabG (x) → OrbG (x).
In particular if G is finite
|G| = | OrbG (x)| · | StabG (x)|.
37We used the same argument for a special case of this in Lecture 4 when we showed that
StabIsom(Z) (0) 6 Isom(Z).
38You should spend some time thinking about the relationship between this proof and the
proof of Lagrange’s Theorem. Can you find an action of H on G making Lagrange a special case
of this result?
26 SIMON WADSLEY

Proof. For y ∈ OrbG (x), {g ∈ G | g · x = y} is a left coset of StabG (x) in G since


it is non-empty and if g · x = y then for h ∈ G
h·x=y ⇐⇒ g −1 · (h · x) = x
⇐⇒ g −1 h ∈ StabG (x)
⇐⇒ h ∈ g StabG (x).
Thus there is a function G/ StabG (x) → OrbG (x) given by h StabG (x) 7→ h · x
with inverse OrbG (x) → G/ StabG (x) given by y 7→ {g ∈ G | g · x = y}.
Now if G is finite then | OrbG (x)| = |G/ StabG (x)| = |G|/| StabG (x)| by La-
grange.39 

Examples.
(1) For the natural action of Isom(Z) on Z the set of left cosets of StabIsom(Z) (0) =
{e, n 7→ −n} in Isom(Z) is in bijection with Z. We secretly used this fact when
we computed Isom(Z) in the first lecture.
(2) For the usual action of D2n on the vertices of the n-gon and v such a vertex we
see that |D2n | = | StabD2n (v)|| OrbD2n (v)| = 2n. Again we secretly used this
when we computed |D2n | = 2n.
(3) The symmetric group Sn acts on X = {1, 2, . . . , n} via the natural action
f ·x = f (x). Then OrbSn (n) = X since for each i ∈ X the function fi : X → X;
fi (i) = n, fi (n) = i, fi (x) = x for x 6∈ {i, n} is an element of Sn . Thus
|Sn | = n StabSn (n). But StabSn (n) is isomorphic to Sn−1 by restricting f ∈ Sn
that fixes n to a permutation of {1, . . . , n − 1}. Thus |Sn | = n|Sn−1 |. Since
|S1 | = 140 we deduce that |Sn | = n!.

Lecture 13
Fact. If f : R3 → R3 is an isometry that fixes 4 non-coplanar points then f is
the identity.
(4) Let X be a regular tetrahedron. Then Sym(X) acts transitively on the set of
4 vertices of X and the stabiliser of a vertex v ∈ X consists of three rotations
and three reflections. Thus | Sym(X)| = 6 · 4 = 24.
This calculation enables us to prove that Sym(X) ' S4 : if we label the
vertices by the numbers 1, 2, 3, 4 then the action of Sym(X) on the vertices
defines a homomorphism θ : Sym(X) → S4 . Since any isometry of R3 fixing all
four vertices is the identity we can conclude that ker θ = {id}. By counting we
can deduce Im θ = S4 .
(5) Let X be a cube. Then Sym(X) acts transitively on the set of 6 faces of X and
the stabiliser H := StabSym(X) (F ) of a face F acts transitively on the set of 4
vertices contained in it41. If v is one of these vertices and w is the diagonally
opposite vertex in F then
StabH (v) = {e, reflection in plane containing v, w and the centre of X}42.

39You might like to think about whether you can do this last part without recourse to Lagrange.
40Or if you prefer |S | = 1
0
41This can be seen by considering rotations about an axis through the centre of F and the
centre of its opposite face.
42Since if an isometry of R3 fixes all vertices of F and the centre of X then it is the identity.
GROUPS 27

Thus
Sym(X) = 6|H| = 6 · | OrbH (v)|| StabH (v)| = 6 · 4 · 2 = 48.
3.3. Conjugacy classes.
Definition. If G is a group then the conjugation action of G on itself is given by
· : G × G → G; g · x = gxg −1 .
Note that the conjugation action is indeed an action since for g, h, x ∈ G,
e · x = exe−1 = x
and
g · (h · x) = g(hxh−1 )g −1 = (gh)x(gh)−1 = (gh) · x.
Definition. The orbits of G on itself under the conjugation action are called the
conjugacy classes of G: the orbit of x ∈ G will be denoted ccl(x); i.e.
ccl(x) = {gxg −1 | g ∈ G}.43
The stabliser of x ∈ G under this action is called the centraliser of x and will be
denoted CG (x).
Examples.
(1) Suppose G = Isom(Z) = {ta : n 7→ a + n, sa : n 7→ a − n | a ∈ Z}. Let
H := {ta | a ∈ Z} denote the subgroup of translations We know that for any
a, b ∈ Z,
tb ta t−1
b = ta
and for n ∈ Z,
sb ta s−1
b (n) = sb ta (b − n) = sb (a + b − n) = n − a

ie
sb ta s−1
b = t−a
so for a 6= 0
CG (ta ) = H and ccl(ta ) = {ta , t−a }44.
Similarly for n ∈ Z
tb sa t−1
b (n) = tb sa (n − b) = tb (a − (n − b)) = (2b + a − n) = s2b+a (n)

and
sb sa s−1
b (n) = sb sa (b − n) = sb (a − (b − n)) = b − (a + n − b) = 2b − a − n = s2b−a (n)

so as a ≡ −a mod 2
CG (sa ) = {t0 , sa } and ccl(sa ) = {sa+2b | b ∈ Z}.
That is there are two conjugacy classes of reflections ccl(s0 ) and ccl(s1 ).45

43Since conjugacy classes are orbits of an action they partition G; that is every element of G
lies in precisely one conjugacy class.
44Of course C (t ) = G and ccl(t ) = {t }.
G 0 0 0
45What is the geometric meaning of this?
28 SIMON WADSLEY

(2) We saw in section 1.5 that in the Möbius group M the conjugacy class of
z 7→ z + 1 consists of all Möbius transformations with precisely one fixed point
i.e.
ccl(z 7→ z + 1) = {f ∈ M | f has precisely one fixed point}
and that every Möbius transformation with precisely two fixed points is in the
same conjugacy class as a Möbius transformation of the form z 7→ az. We will
return later to the question of when ccl(z 7→ az) = ccl(z 7→ bz) and what the
centralisers of these elements are.46 Of course ccl(id) = {id} and CM (id) = M.
Definition. The kernel of the homomorphism G → S(G) given by the conjugation
action of G on itself is called the centre of G and written Z(G).
Lemma. Suppose that G is a group.
(a) For x ∈ G, CG (x) = {g ∈ G | xg = gx}. T
(b) Z(G) = {g ∈ G | gx = xg for all x ∈ G} = x∈G CG (x).
(c) Z(G) = {g ∈ G | |ccl(g)| = 1}.
Proof. (a) For g, x ∈ G, g ∈ CG (x) if and only if gxg −1 = x that is if and only if
gx = xg.
(b) For g ∈ G, g ∈ Z(G) if and only if gxg −1 = x for all x ∈ G i.e. if and only if
gx = xg for all x ∈ G. The other equality follows immediately from (a).
(c) This follows easily from (b): g ∈ Z(G) if and only if gx = xg for all x ∈ G
i.e. if and only if xgx−1 = g for all x ∈ G. 

3.4. Cayley’s Theorem. Cayley’s Theorem will tell us that every group is iso-
morphic to a subgroup of a symmetric group.
Definition. If G is a group then the left regular action of G on itself is given by
the function · : G × G → G; g · x = gx.
Example. The left regular action of Z on itself is by translations. i.e. the corre-
sponding homomorphism Z → S(Z) is given by n 7→ tn .47
Lemma. The left regular action of G on G is an action that is both transitive and
faithful.
Proof. First we should check that the left regular action is indeed an action: for all
g, h, x ∈ G, e · x = ex = x and g · (h · x) = ghx = (gh) · x.
Next we observe that OrbG (e) = G and StabG (e) = {e} since for all g ∈ G,
g · e = g. Thus the left regular action is transitive and faithful. 

Lecture 14
Theorem (Cayley’s Theorem). If G is a group then G is isomorphic to a subgroup
of S(G).

46Spoiler: ccl(z 7→ az) = ccl(z 7→ bz) if and only if b ∈ {a, 1/a}, C (z 7→ z + 1) =


G
{translations in M} and, for a 6= 1, CG (z 7→ az) = {dilations/rotations in M}. Can you prove
these facts now? Hint: if g(z 7→ az)g −1 = z 7→ bz for g ∈ M what can you say about g(0) and
g(∞)?
47recall t denotes translation by n
n
GROUPS 29

Proof. The left-regular action defines a homomorphism θ : G → S(G). Since Im θ


is a subgroup of G we may view this as a homomorphism G → Im θ whose image is
still Im θ. Since the action is faithful, ker θ = {e} and we see that G is isomorphic
to Im θ. 
It perhaps should be said that this theorem is simultaneously deep and almost
useless. Deep because it tells us that anything satisfying our abstract definition of a
group can be viewed as symmetries of something. Almost useless because knowing
this doesn’t really help prove things about groups.
3.5. Cauchy’s Theorem.
Theorem (Cauchy’s Theorem). Supppose that p is a prime and G is a finite group
whose order is a multiple of p. Then G contains an element of order p.
Proof. Consider the action of Zp on Gp via
i · (g0 , g1 , . . . , gp−1 ) = (gi , g1+p i , . . . , g(p−1)+p i )
i.e. by cyclic permutation. This is an action since 0 · x = x for all x ∈ Gp and
(i +p j) · (g0 , g1 , . . . , gp−1 ) = (gi+p j , g1+p i+p j , . . . , g(p−1)+p i+p j )
= i · (j · (g0 , g1 , . . . , gp−1 ))
for all i, j ∈ Zp and g0 , . . . , gp−1 ∈ G.
Let X = {(g0 , . . . , gp−1 ) ∈ Gp | g0 g1 · · · gp−1 = e} ⊂ Gp . Then |X| = |G|p−1
since however we choose g0 , . . . , gp−2 ∈ G there is precisely one choice of gp−1 ∈ G
such g0 g1 · · · gp−1 = e48. Moreover the ‘constant tuple’ (g, g, . . . , g) is in X if and
only g p = e — and if g 6= e this is equivalent to o(g) = p.
Now the cyclic permutation action of Zp on Gp restricts to an action on X since
if g0 g1 · · · gp−1 = e and i ∈ Zp then
gi g1+p i · · · g(p−1)+p i = (g0 · · · gi−1 )−1 (g0 · · · gp−1 )(g0 · · · gi−1 ) = e.
Since Zp has prime order the orbit-stabiliser theorem tells us that every orbit
OrbZp (x) has order 1 or p for x ∈ X. Since the orbits partition X and p is a factor
of |X| it follows that the number of orbits of size 1 in X is a multiple of p.
Since the constant tuple (e, e, . . . , e) is an orbit in X of size 1 there must be
at least p − 1 other such orbits. If (g0 , . . . , gp−1 ) is one such orbit then, since
i · (g0 , . . . , gp−1 ) = (g0 , . . . , gp−1 ) for all i ∈ Zp , we can conclude that gi = g0 for all
i ∈ Zp and so g0p = e. 

4. Quotient Groups
4.1. Normal subgroups. Suppose that G is a group. Let P(G) denote the set of
subsets of G, i.e. the power set of G. There is a natural binary operation on P(G)
given by
AB := {ab | a ∈ A, b ∈ B}.
Examples.
(1) If A ∈ P(G) then A∅ = ∅ = ∅A. If A is non-empty then AG = G = GA.
(2) If H 6 G then the binary operation on P(G) restricts to a binary operation
on P(H).
(3) If H 6 G then the sets {g}H are precisely the left cosets gH of H in G.
48That is g −1 .
p−1 = (g0 g1 · · · gp−2 )
Another random document with
no related content on Scribd:
XIII.
SPECIFICATIONS.

Specifications should cover three principal points:


Physical properties: Elastic limit; ultimate tensile strength;
elongation; reduction of area.
Chemical constituents: Limiting silicon, phosphorus, sulphur,
manganese, and copper; all other elements to be absent or mere
traces in quantity, except carbon.
Finish and general condition: Fixing limit of variation in size from
a given standard; conditions as to pipes, seams, laps, uniformity of
grain, and other defects; no red-shortness.

PHYSICAL PROPERTIES.
It has been shown in Chap. V that tensile strength may be had
from 46,800 lbs. per square inch to 248,700 lbs. per square inch.
There are published in many transactions and technical
periodicals thousands of tests giving elastic and ultimate strength,
ductility, etc., so that every engineer can find easily what has been
done to guide him as to what he can get.
In almost every case the engineer must be the judge as to the
requirements in each; therefore it would be useless to attempt to lay
down any fixed rules or limits.
Many engineers adhere to low tenacity and high ductility in the
belief that they are securing that material which will be safest against
sudden shocks and violent accidental strains.
Theoretically this appears to be correct, but if the statements
made in the preceding chapters are credible it is plain that the limit to
such safety can be passed, and that in insisting upon too low
tenacity and high ductility the engineer may be getting simply a
rotten, microscopically unsound material, through no fault of the
manufacturer, who has been compelled to overmelt or overblow his
steel to meet the requirements, and so reducing the quality of
otherwise good material at no saving in cost to himself, and at a
considerable cost in quality to the consumer.
Any manufacturer would rather check his melt between 10 and
15 carbon, or stop his blow so as to be sure not to overblow, if he
were asked to do so, because it would save him time and expense,
and it would yield sounder, better, and easier working steel.
It may not be wise yet for an engineer to fix limits as to blowing or
melting, for the reason that neither he nor his assistants would know
how to insure compliance, and in attempting to do it they might
interfere too far with manufacturing operations and so involve
themselves in responsibilities which they ought not to assume.
On the other hand, if they will let the carbon and tensile strength
run up a little and reduce ductility slightly, it is safe to say that any
manufacturer will be glad of the chance to help them to get the best
results, which involve no extra cost.
Boiler-steel and rivet-steel usually suffer the most in this respect.
A boiler should be tough, yet it is the belief of the author that boilers
made of the 46,800-lb. steel of which the analysis is given in Chap. V
would not last half as long as boilers made of 65,000-lb. to 70,000-lb.
steel when the increased strength was gained by added carbon and
no overmelting was allowed.
In the same table the “Crucible-sheet” column gives a mean of 24
tests, and a mean analysis, of boiler-steel which has been in use in
12 boilers for nearly 16 years. The boilers are in perfectly good
condition; they have been subjected to severe and very irregular
usage, and they have been in every way satisfactory. Only one test-
piece of the 24 was mild enough to stand the ordinary bending test
after quenching.
That 46,800-lb. steel is remarkably pure chemically; it is
unusually red-short. It would appear to some to be an ideal rivet-
steel; it would stand a very high heat, it would head well and finish
beautifully under a button-set. There is every probability that the
majority of rivets driven of that steel would be cracked on the under
side of the head, where the cracks would never be discovered until
in service the heads flew off.
Rails are usually made of 40 to 45 carbon, tires from .65 up to 80
carbon, crank-pins as high as 70 carbon, with 85,000 lbs. to 95,000
lbs. tensile strength and 12% to 15% elongation.
It is difficult to see how a bridge or a boiler is to be subjected to
any such violent usage as these receive daily; and while it is not
advised that even 40 carbon should be used in boilers or bridges,
although it would be perfectly safe, it does seem to be unreasonable
to run to the other extreme to the injury of the material.
For steel for springs, and for all sorts of tools that are to be
tempered, there is no need of a specification of physical properties
as they are indicated by testing-machines.
The requirement that they shall harden safely and do good work
afterwards involves necessarily, high steel of suitable quality.

CHEMICAL CONSTITUENTS.
No engineer should, unless he be an expert steel-maker, attempt
to specify an exact chemical formula and a corresponding physical
requirement; in doing so he would probably make two requirements
which could not be obtained in one piece of steel, and so subject
himself to a back down or to ridicule, or both.
On the other hand, he may properly, and he should fix, a limit
beyond which the hurtful elements would not be tolerated.
Notwithstanding satisfactory machine tests, successful shop-work,
and a liberal margin of safety, no steel can be relied upon that is
overloaded with phosphorus, sulphur, manganese, oxygen,
antimony, arsenic, or nitrogen.
In regard to silicon, it is common to have as much as 20 to 25
points in tires, with 55 to 80 carbon; such tires are made by the best
manufacturers, and they endure well. But it is certain that good,
sound steel can be made for any purpose with silicon not exceeding
10.
Structural steel can be made cheaply within the following limits:

Silicon < .10


Phosphorus < .05
Sulphur < .02
Manganese < .50 or even < .30
Copper < .03
Carbon to meet the physical requirements.
Steel made within these limits and not overblown or overmelted
must be better in every way than steel of

Silicon > .20


Phosphorus > .08
Sulphur > .05
Manganese > .60
Carbon to meet the same requirements
A steel of the latter composition, or with no fixed limits, may be
made cheaper than the first by a dollar or two a ton; but for any large
lot it is believed that the first specification would be bid to at as low a
price as if there were no specification; competition among
manufacturers would fix that. At any rate there is no reason why an
engineer should refuse to demand fairly pure material when he can
do so at little or no extra cost.
Arsenic, antimony, or any other elements should be absent, or <
.005.
FINISH AND GENERAL CONDITIONS.
As there can be no such thing as exact work done, there must be
some tolerance as to variation in size. In standard sections, sheets,
and plates this is usually covered by a percentage of weight; in
forgings or any pieces that are to be machined the consumer should
allow enough to insure a clean, sound surface. But it would be
unwise to lay down any rule here, because conditions vary; a rolled
round bar may finish nicely by a cut of from ¹/₃₂ to ¹/₁₆ of an inch, and
so also a neatly dropped forging; an ordinary hammered forging
might require a cut of ¼ or ⅜ of an inch; such a forging might be
made closer to size at a cost for extra time at the hammer far
exceeding the saving of cost in the lathe. These are cases where
common-sense and good judgment must govern.
Pipes should not be tolerated if they can be discovered; because
a pipe appears small in the end of a bar it is no evidence that it is not
larger farther in.
Seams should not be allowed in any steel that is to be hardened;
they should be a minimum in any steel, as they are of no possible
use; small seams when not too numerous may do no harm in
structural or machinery steel, and consumers should be reasonable
in regard to them, or else they may have too high prices put upon
their work, or too high heat used in efforts to close the last few
harmless seams.
Burns, rough, ragged holes in the faces or on the corners, are
inexcusable and should be rejected; the steel has been abused, or it
is red-short; in either case the ragged breaks are good starting-
points for final rupture.
Laps should not be permitted; they are evidences of
carelessness; there can be no excuse for them.
Fins are sometimes unavoidable in a difficult shape; for instance,
if a trapezoid is wanted, it may be rolled in this form:
or in this:

The consumer must decide which; if he wants sharp angles he


must accept the fin and cut it off, or have it cut off by the
manufacturer.
Rivet-steel should be tested rigidly for red-shortness, because
red-short steel may crack under the head as the steel cools.
Emphasis is laid upon this because engineers will insist upon
excessive ductility in rivet-steel, not realizing that they may be
requiring the manufacturer to overdose his steel with oxygen to its
serious injury.
No sharp re-entrant angles should be allowed under any
circumstances where there is a possibility of vibrations running
through the mass. All re-entrant angles should be filleted neatly.
No deep tool-marks should be allowed; a fine line scored around
a piece by a lathe-tool, or a sharp line cut in a surface by a planing-
tool will fix a line of fracture as neatly as a diamond-scratch will do it
on a piece of glass.
Indentations by hammers or sledges should be avoided; they
may not be as dangerous as lathe-cuts, but they can do no good,
and therefore they are of no use.
XIV.
HUMBUGS.

Steel is of such universal use and interest in all of the arts that it
attracts the attention of would-be inventors perhaps more than any
other one material.
Half-informed, or wholly uninformed, men get a smattering of
knowledge of some one or more of the well-known properties of
steel, make an experiment which produces a result that is new and
startling to them, and at once imagine that they have made a
discovery; this they proceed to patent and then offer it to the world
with a great flourish of trumpets.
Many steel-workers, even men of skill, who know something of
the difficulties that follow irregular work, or who are not quite fully
informed as to the properties of steel, seize upon these discoveries
in the hope that they have found a royal road to success where all
old pitfalls are removed and their path is made easy.
Not wishing to discourage pioneers in legitimate efforts to
improve, it is the object of this chapter to warn them against being
too ready to spend their money because of flaming circulars or glib
tongues. It is the duty and the interest of a steel-maker to examine
and test every apparently new suggestion, for the reason that there
is still room for improvement, and he should let no opportunity for a
betterment slip past him.
As a rule the steel-maker does test every claim that is laid before
him, unless it be a repetition of some old plan long since tried and
found worthless. This is the bane of the steel-maker’s life, and yet he
must keep at this work so that he may know for himself whether
anything of value has been discovered, and also that he may advise
his clientage properly.
Inventions relating to the manufacture of steel have no interest
for steel-users except as lively manufacturers may adopt the
mistaken plan of flourishing trumpets to attract trade, not always
giving a corresponding benefit to the consumer.
Examples of this sort of thing may be illustrated by so-called
phosphorus steel, silicon steel, and aluminum steel; also the case
mentioned before of parties recommending seams as evidences of
excellence in high steel. Such efforts are sometimes costly to
consumers until active competitive manufacturers expose the
humbug.
Among the most absurd of such claims are those where a
nostrum is used to convert ordinary Bessemer or open-hearth steel
into the finest of tool-steel, equal to the best crucible-steel; for
example, a patent to convert mild Bessemer steel into the finest tool-
steel by merely carbonizing it by the old cementation process; this
takes no account of the silicon, manganese, oxygen, and nitrogen in
the mild Bessemer, makes no provision for their removal, and
involves a costly method of putting carbon into poor stock in face of
the fact that a Bessemer steel maker can put the same amount of
carbon there at practically no cost, and so produce a better material.
Among the humbugs that do not involve the manufacturer, the pet
one is a nostrum for restoring burnt steel; these have been evolved
by the dozen, in face of the fact that burned steel cannot be restored
except by smelting, and that overheated steel, coarse-grained steel,
can be restored by merely heating it to the right temperature, a
process which has been explained fully in Chapter VI.
Another pet is some greasy compound for toughening high steel
so as to make it do more work. This is done by heating the steel to
about recalescence and plunging it into the grease, perhaps once, or
possibly two or three times; then working it into a tool and
proceeding in the ordinary way. This will make a good tool; it is the
partial annealing plan explained in a previous chapter. Now take a
similar piece of steel, heat it the same way, lay it down in a warm,
dry place alongside the forge-fire, and let it cool; then heat it and
work it into a tool and it will beat the greased tool.
When all of these operations of restoring, partial annealing,
annealing, etc., depend merely upon temperature and rate of
cooling, why spend money for nostrums that add no possible
benefit?
There is room for improvement in steel, great room for great
improvements; they will come in time as science and knowledge
advance, and great benefits to the consumers will come with them.
This chapter is not written to place difficulties in the way of
legitimate improvement, but to warn unsuspecting people against
quackery. Some of the humbugs are honest productions of well-
meaning ignorance, and some that come from designing
manufacturers are not entitled to such charitable designation. A
knowledge of the simplest properties of steel will enable a thoughtful
man to judge as to whether a proposed improvement is likely to be of
any value or not, and the warnings given are intended as a
protection to the unsuspecting and credulous.
XV.
CONCLUSIONS.

After perusal of the preceding chapters the reader may form a


hasty conclusion that if steel be so sensitive as it is stated to be its
use may be difficult and precarious, and that it must be handled in
fear and trembling, lest the result should be a dangerous structure,
and the builder must be in doubt as to its safety.
The conveyance of any such impressions is not intended at all;
emphasis has been laid upon practices that are hurtful in order that
every steel-user may know what to avoid, solely that he may then be
sure that he has the best, the most reliable, and most useful material
that is known to man.

WHAT TO AVOID.
He should avoid uneven heat, excessive heat, or too low heat.
The range between orange red and the heat that will granulate is so
great that no one who is not a bungler or indifferent need ever get
outside of it.
The uniformity of temperature that is insisted upon is so easily
seen that any person who is not color-blind should have no trouble in
securing it by the simplest manipulations of the furnace.
Practical uniformity of the work put on a piece is readily secured
by any mechanic of ordinary skill.
Red-short, cold-short, or honeycombed steel are easily detected,
and, under reasonable specifications, the steel-makers can as easily
avoid them.
Steel a little higher than most engineers favor in their
specifications is certainly as safe as, and likely to be sounder than,
extremely ductile steel.
Wild steel, resulting almost certainly in micro-honeycombs, if not
worse, can only be avoided by the co-operation of the manufacturer,
and engineers should impress this point with energy.
Such micro-unsoundness as is shown in Mr. Andrews’s report
upon a broken rail and propeller-shaft can be reduced to a minimum
by insisting upon reasonably pure steel.
If sulphur, phosphorus, silicon, and oxygen are kept at a
reasonable minimum, sulphides, phosphides, silicides or silicates,
and oxides must be at a corresponding minimum.
That there is much room for improvement in the manufacture of
steel is evident, and when means of getting rid of oxygen, nitrogen,
and all other undesirable elements have been found the steel of the
future will be very different in kindliness of working and in endurance
of strains than that with which we are familiar.
It is believed, however, that no matter how perfect the
manufacture may become, nor what the final theories of hardening,
etc., may be, the properties stated in these pages will remain the
same as long as steel continues to be essentially a union of iron and
carbon.
Some other alloy or compound may displace carbon steel, and
present an entirely new set of properties, but there is nothing of the
kind in sight now, and engineers need have no fear of having a new
art to learn very soon.
To one who has spent an ordinary business lifetime in making
steel, studying it, and working with it it becomes a subject of
absorbing interest, if not of love; and steel when handled reasonably
is so true that “true as steel” ceases to be a metaphor, it is then a
fact which fills him with the most entire confidence.
Once more, steel highly charged with sulphur, phosphorus,
arsenic, oxygen, and nitrogen is certainly highly charged with so
many elements of disintegration; it takes more serious harm from
ordinary deviations from good practice, such little irregularities as
occur inevitably in daily working, than steel does which is more free
from these elements.
Reasonably pure, sound, reliable steel can be had at moderate
cost, and all consumers should insist upon having it.
Regular, uniform, reliable working can be had where it is
required, and there should be no excuse for irregular grain,
overheated work, uneven work, or any other bungling. Where skill is
required and reasonable discipline is enforced, good work will not
cost any more than bad work.
Many people still hold to the idea that there are many mysteries
connected with steel, and that many unaccountable breaks occur
which make it an unreliable material. It is hoped that what has been
set down in these pages will go far to dissipate these supposed
mysteries, and to give confidence to steel-users.
Many breaks are unaccounted for, but it is not within the author’s
experience that any fracture ever occurred that could not have been
explained if it had been examined thoroughly in the light of what we
know now. There is much to be learned, but there are no mysteries.
GLOSSARY.
There are many shop terms used in this book which may not be
familiar to all steel-users.
They are in common use in steel-manufactories, and definitions
of them will enable a steel-user to understand more clearly the
common talk he will hear in the shops.
Blow-holes.—Blow-holes are the small cavities, usually
spherical, which are formed in ingots as the steel congeals by
bubbles of gas which cannot escape through the already frozen
surface.
Burned.—Burned steel is steel that is reduced to oxide in part by
excessive heating.
Check.—A check is a small rupture caused by water; it may run
in any direction; it is usually not visible until steel is ruptured.
Chemical Numeration.—Chemical quantities are almost
universally expressed in hundredths of one per cent, as explained in
the body of the work. It is a very convenient numeration; any steel-
worker, melter, hammerman, etc., will talk of 20, or 50, or 130
carbon; or 8 phosphorus; or 10, 15, or 25 silicon, etc.; and will talk
intelligently, although he may not know the exact mathematical value
of these points.
Dead-melting; synonym, killing.—Dead-melting—killing—
means melting steel in the crucible or open hearth until it ceases to
boil or evolve gases; it is then dead, it lies quiet in the furnace, and
killed properly it will set in the moulds without rising or boiling.
Dry.—Steel is called dry when its fracture is sandy-looking,
without lustre or sheen, and without a proper blue cast. There is
more of a shade of yellowish sandstone. It is an evidence of impurity
and weakness.
Fiery.—Fiery steel has a brilliant lustre; it is an evidence of high
heat.
If the grain be fairly fine and of bluish cast, it is not
necessarily bad in mild steel; in high steel or in tool-
steel it should not be tolerated.
If the grain be large and of brassy cast, it is sure
evidence of bad condition; the grain should be restored
before the steel is used.
In hardened steel it is always bad, except in dies to
be used under the impact of drop-hammers; in this
case steel must be so hard as to be slightly fiery.
Grade.—Grade applies to quality, as crucible, Bessemer, or
open-hearth grade. Or in the crucible, common, spring, machinery,
tool, special tool, etc., etc. It does not indicate temper or relative
hardness.
Honeycombed.—Unsound from many blow-holes. Usually
applied to ingots. It is a bad condition.
Lap.—A lap is caused by careless hammering, or by badly
proportioned grooves in rolls, or by careless rolling. A portion of the
steel is folded over on itself, the walls are oxidized and cannot unite.
A lap generally runs clear along a bar, practically parallel with its
axis; it may be seen by a novice. Lapped steel should be rejected
always.
Overblown.—Steel that has been blown in a Bessemer
converter after the carbon is all burned; then there is nothing but
steel to burn, and the result is bad.
Overheated.—Steel that has been heated too hot, and not quite
burned; its fiery fracture exposes it. The grain of overheated steel
may be restored, but restored steel is never as reliable as steel that
has not been overheated. Overheating is a disintegrating operation.
Overmelted.—Steel that has been kept too long in fusion. The
finest material may be ruined in a crucible by being kept in the
furnace any considerable time after it has been killed. Open-hearth
steel may be injured seriously in the same way. Prompt teeming after
killing should be the rule.
Pipe.—A pipe is the cavity formed in an ingot when it cools; the
walls chill first and nearly to the full size of the mould, then the
shrinking mass separates in the middle, forming a pipe. A pipe
should be at the top of the ingot; it may occur anywhere by bad
teeming.
Point.—One hundredth of one per cent of any element. You have
say 10 points of carbon, or 10 carbon; you want it raised a few points
to 15 or 18 carbon.
Recalescence.—When a piece of steel is heated above medium
orange color and cools slowly, at about medium orange—1100° to
1200° F.—the change of color ceases, then the color rises
sometimes to bright orange, and afterwards the cooling goes on; this
phenomenon is called recalescence. This is not yet a common shop
term.
Restoring.—When a piece of overheated steel is re-heated to
recalescence, kept there a few minutes, and then cooled slowly, its
grain becomes fine and its fiery lustre disappears; this is called
restoring. No nostrums are necessary.
Sappy.—Well-worked, good steel has a bluish cast, a fine grain,
and a silky sheen. It is sappy; it is as good as it can be made.
Seam.—A seam is a longer or shorter defect, caused by a blow-
hole which working has brought out to the surface and not
eliminated. It usually, or always, runs in the direction of working.
Seams are distinguished from laps by not being continuous; they are
usually only an inch or two in length.
Short (Cold, Red, Hot).—Cold-short steel is weak and brittle
when cold.
Red-short steel is brittle at dark orange or medium
orange heat or at the common cherry red heat. It may
forge well at a lemon heat, and be reasonably tough
when cold.
Hot-short steel is brittle and friable above a medium
orange color; it may forge well from medium orange
down to black heat.
Star.—A brilliant spot in mid-section showing that the pipe is not
all cut away. It should be removed from tool-steel especially, as it
may have considerable depth. It is of no use in any steel.
Temper.—Used by the steel-maker it means the quantity of
carbon present. It is low temper, medium, or high; or number so and
so by his shop numbers.
Used by the steel-user or the temperer it means the
color to which hardened steel is drawn: straw, brown,
pigeon-wing, blue, etc., etc.
Or, it is the steel-maker’s measure of initial
hardness; and it is the steel-user’s measure of final
hardness.
Water-crack.—A crack caused in hardening; it may run in any
direction governed by lines of stress in the mass. It is distinguished
from a check by being larger, and usually plainly visible.
Wild Steel.—Steel in fusion that boils violently, and acts in the
moulds like lively soda-water or beer does when poured into a glass.
Transcriber’s Notes:

New original cover art included with this eBook is granted to the public domain.
The illustrations have been moved so that they do not break up paragraphs and so
that they are next to the text they illustrate.
Typographical and punctuation errors have been silently corrected.
*** END OF THE PROJECT GUTENBERG EBOOK STEEL ***

Updated editions will replace the previous one—the old editions will
be renamed.

Creating the works from print editions not protected by U.S.


copyright law means that no one owns a United States copyright in
these works, so the Foundation (and you!) can copy and distribute it
in the United States without permission and without paying copyright
royalties. Special rules, set forth in the General Terms of Use part of
this license, apply to copying and distributing Project Gutenberg™
electronic works to protect the PROJECT GUTENBERG™ concept
and trademark. Project Gutenberg is a registered trademark, and
may not be used if you charge for an eBook, except by following the
terms of the trademark license, including paying royalties for use of
the Project Gutenberg trademark. If you do not charge anything for
copies of this eBook, complying with the trademark license is very
easy. You may use this eBook for nearly any purpose such as
creation of derivative works, reports, performances and research.
Project Gutenberg eBooks may be modified and printed and given
away—you may do practically ANYTHING in the United States with
eBooks not protected by U.S. copyright law. Redistribution is subject
to the trademark license, especially commercial redistribution.

START: FULL LICENSE


THE FULL PROJECT GUTENBERG LICENSE

You might also like