Abstract Algebra
Abstract Algebra
Abstract Algebra
A COURSE IN
ABSTRACT ALGEBRA
D R A F T : J U LY 3 , 2 0 1 7
To Abigail and Emilie
We may always depend upon it that
algebra which cannot be translated
into good English and sound common
sense is bad algebra.
William Kingdon Clifford
(18451879),
The Common Sense of the Exact Sciences
(1886) 21
Preface
Mathematics is written for mathemati-
cians
Nicolaus Copernicus (14731543),
bstract algebra is a fascinating, versatile and powerful subject
A with many important applications not just throughout the wider
preface to De Revolutionibus Orbium
Clestium (1543)
world of mathematics, but also in several other disciplines as well.
Group theory, the focus of the first seven chapters of this book, is par-
ticularly good at describing symmetry. It thereby enables molecular
chemists and crystallographers to study and understand the structure
and properties of molecules and crystals, and it gives particle physi-
cists valuable insights into the fundamental particles and forces that
form our universe. Finite fields and sporadic groups have important
applications in cryptography and information theory, in the construc-
tion of strong public-key cryptosystems and error-correcting codes for
secure and reliable transmission or storage of data.
But its abstract nature, the thing that makes it so useful and versatile,
can also sometimes render it opaque, dry and difficult to understand
for students meeting the subject for the first time. I struggled consider-
ably with much of this material when I was a student, and there were
certainly several occasions when I lost sight of why we were learning
a particular topic and how it joined up with everything else. The effort
ultimately paid off but I was left wondering whether it needed to be
so difficult.
In some sense, the answer is yes: mathematics requires a level of
formality and rigour greater than probably any other subject, and
sooner or later we must put in the work needed to fully understand
and internalise the concepts and theorems under discussion. To borrow
Euclids famous attributed comment to the Egyptian pharaoh Ptolemy
I Soter (c.367c.283BC), there is no royal road to algebra.
In the introduction to Herman Hesses novel The Glass Bead Game, the
narrator remarks that no textbook of the Game will ever be written,
because nobody who has devoted the long years of study necessary to
master it would have any interest in making it easier for anyone else
do so. There were times when I almost wondered if something similar
was occasionally the case with mathematics.
vi a course in abstract algebra
The main aim of this book is to provide a clear, detailed and com-
prehensive account of the topics covered in a typical undergraduate
course on abstract algebra. I hope that Ive mostly succeeded, but if
not, then to echo Hardy and Wright I hope that my incompetence at
least hasnt been too extravagant.
1. Groups 2. Subgroups
1.1 Numbers 2.1 Groups within groups
1.4 Permutations
8. Rings
3. Normal subgroups 8.1 Numbers
3.1 Cosets and conjugacy classes 8.2 Matrices
3.2 Quotient groups 8.3 Polynomials
4. Homomorphisms
4.1 Structure-preserving maps 9. Ideals
4.2 Kernels and images 9.1 Subrings
7.5 Extensions
An assertion whose original source Ive been unable to trace, but which
is often, plausibly, attributed to George Plya (18871985) states that
mathematics is not a spectator sport. To this end I have included
a selection of exercises at the end of each chapter. Some are fairly
routine, slight variations on examples in the text, while others are
more involved and require further thought.
Credits
Nicholas Jackson,
Coventry, July 2017
Contents
Preface v
How to read this book vii
Acknowledgements x
Credits x
1 Groups 1
1.1 Numbers 1
1.2 Matrices 17
1.3 Symmetries 20
1.4 Permutations 24
2 Subgroups 39
2.1 Groups within groups 39
2.2 Cosets and Lagranges Theorem 49
2.3 Eulers Theorem and Fermats Little Theorem 62
3 Normal subgroups 71
3.1 Cosets and conjugacy classes 71
3.2 Quotient groups 82
3.A Simple groups 93
4 Homomorphisms 105
4.1 Structure-preserving maps 105
4.2 Kernels and images 112
4.3 The Isomorphism Theorems 118
5 Presentations 135
5.1 Free groups 136
5.2 Generators and relations 146
5.3 Finitely generated abelian groups 167
5.A Coset enumeration 180
xiv a course in abstract algebra
But having got as far as the invention of zero, its not much further
a step to invent negative numbers.4 With a bit of thought, we can 4
The earliest known treatment of nega-
formulate perfectly reasonable questions that cant be answered in tive numbers occurs in the ancient Chi-
nese text Jiuzhang suanshu (Nine Chap-
either N or N0 , such as what number, when added to 3, gives the ters on the Mathematical Art), which
answer 2? dates from the Han dynasty (202BC
220AD) in which positive numbers are
Attempts to answer such questions, where the need for a consistent represented by black counting rods and
answer is pitted against the apparent lack of a physical interpreta- negative numbers by red ones.
tion for the concept, led in this case to the introduction of negative
numbers. This sort of paradigm shift occurs many times throughout
the history of mathematics: we run up against a question which is
unanswerable within our existing context, and then ask But what if
this question had an answer after all? This process is often a slow
and painful one, but ultimately leads to an expanded understanding
of the subject at hand. It took somewhere in excess of a thousand
years for the concept of negative numbers to fully catch on. The Greek
mathematician Diophantus, writing in the third century AD, rejected
negative solutions to linear or quadratic equations as absurd. Even as
late as the 16th century, the Italian mathematician Girolamo Cardano
(15011576) referred to such numbers as fict, or fictitious, although his
Italian predecessor Leonardo of Pisa, better known as Fibonacci, had
interpreted them in a financial context as a loss or debit. Meanwhile, Wikimedia Commons
the Indian mathematicians Brahmagupta (598668) and Mahavira (9th Figure 1.1: Page from Nine Chapters on
the Mathematical Art
century) had made the necessary intuitive leap and developed rules
for multiplying negative numbers (although even Mahavira baulked
at considering their square roots). Adjoining negative numbers to N0
yields the set of integers
Z = {. . . , 3, 2, 1, 0, 1, 2, 3, . . .}.
Later on, we will examine how Z was extended to construct more
sophisticated number systems, in particular the rational numbers Q,
the real numbers R and the complex numbers C, but for the moment
this will suffice.
The operation of addition can be extended to the negative integers
in an obvious and consistent way, and we find that all of those tricky
questions involving subtraction can now be solved. More importantly,
it turns out that for any integer n N0 there is a unique negative
integer n Z such that
n + (n) = 0 = (n) + n. (1.3)
But this also works for the negative integers themselves, as long as we
define
(n) = n
for any integer n Z. So, we now have a set Z equipped with
6 a course in abstract algebra
The first few groups we will meet are all abelian, although in a short
while we will study some examples of nonabelian groups as well.
Wikimedia Commons / Johan Grbitz (17821853)
Our first abelian example is a slight modification of Z, but instead of Abelian groups are named after the
taking the infinite set of integers, we take a finite subset, and instead Norwegian mathematician Niels Hen-
rik Abel (18021829), whose brilliant
of using the usual addition operation, we use modular arithmetic: career (as well as his life) was cut trag-
Example 1.12 (Cyclic groups) Let ically short by tuberculosis at the age
of 26. At the age of 19 he proved, in-
Zn = {0, . . . , n1} dependently of his similarly tragic con-
temporary variste Galois (18111832),
be the set consisting of the first n non-negative integers, and let that the general quintic equation
+ : Zn Zn Zn be addition modulo n. That is, for any two ax5 + bx4 + cx3 + dx2 + ex + f = 0
a, b Zn we define a+b to be the remainder of the integer a+b Z cannot be solved by radicals. His mono-
after division by n. graph on elliptic functions was only
discovered after his death, which oc-
This is the cyclic group of order n. curred two days before the arrival of a
letter appointing him to an academic
We can regard the elements of this group geometrically as n equally- post in Berlin.
spaced points around the circumference of a circle, and obtain a+b by In 2002, to commemorate his bicente-
starting at point a and counting b positions clockwise round the circle nary (and approximately a century af-
ter the idea had originally been pro-
to see which number we end up with. See Figure 1.2 for a geometric posed) the Norwegian Academy of Sci-
depiction of 5 + 9 = 2 in Z12 . ences and Letters founded an annual
prize in his honour, to recognise stellar
Its natural, when we meet a new mathematical construct, to ask what achievement in mathematical research.
the simplest possible example of that construct is. In the case of
7
groups, the following example answers this question. 6 8
Example 1.13 Let G = {0} be the set consisting of a single element.
5 9
There is only one possible binary operation that can be defined on 11 0 1
this set, namely the one given by 0 0 = 0. A routine verification 10 2
shows that this operation satisfies all of the group axioms: 0 is the 4 9 3
identity element, its its own inverse, and the operation is trivially 8 4
associative and commutative. This is the trivial group. 7 6 5
3
We could denote the trivial group as Z1 , although nobody usually
2 0
does: depending on the context we typically use 1 or 0 instead. Note 1
that although we can define a binary operation of sorts (the empty Figure 1.2: Addition in Z12 . Here 5 +
operation) on the empty set , we dont get a group structure because 9 = 14 2 mod 12
8 a course in abstract algebra
1 2
1 1 2
2 1
2 2 1
( h 1 g 1 ) ( g h ) = h 1 ( g 1 g ) h = h 1 h = e 4
3 1 Re z
( g h ) ( h 1 g 1 ) = g ( h h 1 ) g 1 = g g 1 = e
2
Another important observation concerning the group C3 in Exam-
ple 1.14 is that its multiplication table has essentially the same structure
Figure 1.4: Cube roots of unity
as the multiplication table for the cyclic group Z3 .
+ 0 1 2 1 2
0 0 1 2 1 1 2
1 1 2 0 2 1
2 2 0 1 2 2 1
g h = g 1 h 1 = ( h g ) 1 = h g
as required.
Given that an isomorphism preserves at least some aspects of the
structure of a group, it is reasonable to ask how it affects the order of
a given element. The answer is that it leaves it unchanged:
Proposition 1.21 If : G H is an isomorphism, then |( g)| = | g|
for any g G.
( g)n = ( g) ( g) = ( g g) = ( gn ) = (eG ) = e H .
We must now check that n is the smallest positive integer such that
( g)n = e H . Suppose that there exists some 1 6 k < n such that
( g)k = e H . Then
( g ) k = ( g ) ( g ) = ( g g ) = ( g k ).
0 1
satisfies the identity condition. Not every square matrix has an inverse,
but as long as we restrict our attention to those that do, we should be
able to find some interesting examples of groups.
Example 1.33 The matrices This matrix group { I, A, B} is an ex-
" # " # ample of quite a powerful and impor-
1 23 12 3
1 0 tant concept called a representation:
I= , A = 32 , B= 2 . broadly speaking, a formulation of a
0 1 2 21 23 12 given group as a collection of matrices
forms a group with multiplication table which behave in the same way. We will
study this idea in a little more detail in
Section 8.A.
I A B
I I A B
A A B I
B B I A
which is clearly isomorphic to the table for the group Z3 that we met
a little while ago, under the isomorphism
I 7 0, A 7 1, B 7 2.
We now use the fact that matrix multiplication is, in general, noncom-
mutative to give us our first examples of nonabelian groups.
Example 1.34 Let GLn (R) denote the set of invertible, nn matrices
with real entries. This set forms a group (the general linear group)
under the usual matrix multiplication operation.
The set GLn (R) is closed under multiplication (since any two invertible
matrices A and B form a product AB with inverse B1 A1 ), matrix
multiplication is associative, the nn identity matrix In GLn (R),
and each matrix A GLn (R) has an inverse A1 GLn (R). We can
specialise this example to give three other (also nonabelian) groups
which are contained within GLn (R):
Example 1.35 The special linear group SLn (R) is the group of nn
real matrices with determinant 1.
18 a course in abstract algebra
Example 1.37 The special orthogonal group SOn (R) is the group
of nn orthogonal real matrices with determinant 1.
These groups GLn (R), SLn (R), On (R), SOn (R), Sp2n (R), Un and SUn
are, for n > 1 at least, all infinite and nonabelian. But they also have
a continuous structure inherited from R, which means that we can
regard them as being locally similar to Rm for some value of m. Such
an object is called a manifold, and a group which has a compatible
manifold structure in this way is known as a Lie group, after the
Norwegian mathematician Marius Sophus Lie (18421899).
Example 1.41 Let I, A and B be as in Example 1.33, and let
" # " #
12 23 1 3
1 0 2 2
C= , D= , E = .
0 1 23 1
2 2
3 1
2
Then the set { I, A, B, C, D, E} forms a group with multiplication table
I A B C D E
1933 CERN / Wolfgang Pauli Archive
I I A B C D E The Austrian theoretical physicist Wolf-
A A B I D E C gang Pauli (19001958) was one of the
pioneers of quantum mechanics and
B B I A E C D particle physics. His contributions in-
C C E D I B A clude the Pauli Exclusion Principle,
which states that no two electrons
D D C E A I B
(more generally, any two fermions) can
E E D C B A I exist in the same quantum state at the
same time, for which he was awarded
This group can be seen to be nonabelian: its multiplication table is the Nobel Prize for Physics in 1945. In
asymmetric in the leading (top-left to bottom-right) diagonal. 1930, while working on the problem
of beta decay, he postulated the exis-
Example 1.42 The three unitary matrices tence of a new particle (the neutrino)
whose existence was confirmed experi-
mentally 26 years later.
0 i
0 1 1 0
x = y = z = A notorious perfectionist, he would of-
1 0 i 0 0 1
ten dismiss work he considered sub-
are known as Paulis matrices, and are particularly relevent in par- standard as falsch (wrong) or
ganz falsch (completely wrong); fa-
ticle physics (where they represent observables relating to the spin mously he once remarked of one re-
of spin 12 particles such as protons, neutrons and electrons) and search paper es ist nicht einmal falsch
quantum computing (where they represent an important class of (it is not even wrong).
e r r2 m1 m2 m3
latter convention, but you should be aware that some books follow the
e e r r2 m1 m2 m3
other one. Its multiplication table is shown in Table 1.5. r r r2 e m3 m1 m2
r2 r2 e r m2 m3 m1
More generally, we can do the same thing with a regular ngon. This m1 m1 m2 m3 e r r2
has ordern rotational symmetry, and n axes of symmetry. In the case m2 m2 m3 m1 r2 e r
where n is odd, each of these axes passes through the midpoint of one m3 m3 m1 m2 r r2 e
side and the opposite vertex. If n is even, then half of these axes pass Table 1.5: The multiplication table for
the dihedral group D3
through pairs of opposite vertices, and the other half pass through the
midpoints of opposite sides. m2
Dn = {e, r, . . . , r n1 , m1 , . . . , mn }
is the symmetry group of the regular plane ngon. Here, e is the m4
identity, r represents an anticlockwise rotation through an angle of
2
n , and mk denotes a reflection in the line which makes an angle of
k
n with the horizontal.
Heres another example: the dihedral group D4 , which describes the Figure 1.7: Axes of symmetry of the
symmetry of a square. square
2 3
Example 1.44 The dihedral group D4 = {e, r, r2 , r3 , m1 , m2 , m3 , m4 } e r r r m1 m2 m3 m4
e e r r 2 r 3 m1 m2 m3 m4
consists of the eight possible symmetry transformations on the square. r r r 2 r 3 e m4 m1 m2 m3
Here, e denotes the identity transformation, r is an anticlockwise r 2 r 2 r 3 e r m3 m4 m1 m2
rotation through an angle of 2 , r2 is therefore a rotation through r 3 r 3 e r r 2 m2 m3 m4 m1
m 1 m1 m2 m3 m4 e r r2 r3
an angle of and r3 a clockwise rotation through an angle of 2 , m2 m2 m3 m4 m1 r 3 e r r 2
m1 and m3 are reflections in the squares diagonals, and m2 and m4 m3 m3 m4 m1 m2 r 2 r 3 e r
m4 m4 m1 m2 m3 r r 2 r 3 e
are reflections in, respectively, the vertical and horizontal axes. The
Table 1.6: The multiplication table for
multiplication table is shown in Table 1.6.
the dihedral group D4
A rectangle has fewer symmetries than a square, and as we might v
expect, its symmetry group is simpler:
Example 1.45 Let V4 = {e, r, h, v} be the symmetry group of a (non-
square) rectangle, where e is as usual the identity, r denotes a rotation
h
through an angle of about the centre point, and h and v are,
respectively, reflections in the horizontal and vertical axes. Then
we obtain the group with multiplication table shown in Table 1.7
This group is isomorphic to Z2 Z2 from Example 1.30, via the Figure 1.8: Axes of symmetry of a rect-
isomorphism angle
e r h v
e (0, 0) r (1, 1) h (1, 0) v (0, 1)
e e r h v
and is hence the Klein 4group in a very superficial disguise. r r e v h
h h v e r
We can also think of this group as the dihedral group D2 , the sym- v v h r e
Table 1.7: Multiplication table for sym-
metry group of a 2sided polygon or bigon, the lens-shaped figure
metry group of a rectangle
depicted in Figure 1.9.
Example 1.33, as well as the orthogonal groups On (R) and SOn (R)
are examples of isometry groups: groups of transformations which
preserve lengths and distances.
Definition 1.46 Given some set S Rn , define Isom+ (S) to be the
group of all direct isometries of S: length-preserving transforma-
tions which also preserve the orientation of S and the underlying
Figure 1.9: Symmetries of a bigon
space Rn . Define Isom(S) to be the full isometry group of S, consist-
ing of the direct isometries together with the opposite isometries:
those which reverse the orientation of S and Rn .
In this terminology, if we let Pn denote the regular nsided polygon,
then Isom+ ( Pn )
= Zn and Isom( Pn ) = Dn .
Definition 1.47 Let E2+ = Isom+ (R2 ) be the group of all orientation-
preserving isometries of the plane R2 . Then E2+ consists of trans-
lations and rotations. The full isometry group E2 = Isom(R2 ) also
includes reflections and glide reflections (a reflection in some line,
followed by a translation parallel to the same line). This group E2 is
called the (two-dimensional) Euclidean group.
In this table, the operation in row a and column b is the product ab;
that is, the isometry a followed by the isometry b.
The full isometry group Isom(3 ) also includes twelve opposite isome-
tries: six order2 reflections m12 , m13 , m14 , m23 , m24 and m34 (in planes
passing along one edge and through the midpoint of the opposite
Figure 1.11: Direct isometries of the
edge) and six other roto-reflection isometries tA , tB and tC formed by tetrahedron 3 : axes for order3 rota-
rotating the tetrahedron through an angle of 2 about one of the tions (top) and order2 rotations (bot-
tom)
three axes s A , s B or sC and then reflecting in a plane perpendicular to
that axis. (The typet isometries may also be generated by pairs of
typem reflections.)
24 a course in abstract algebra
Given a set X, we can form the set Sym( X ) of all possible permutations
of X. In order to turn this into a group, we need to impose a suitable
binary operation on it, and the one we choose is that of composition.
Viewing two permutations , Sym( X ) as bijections , : X X,
their composites : X X and : X X are certainly well-
defined. These composites, happily, are also permutations in their
own right, since the composite of two bijections is itself a bijection. If
we view and as (possibly different) ways of shuffling or reordering
the elements of X, then if we do and then do to the result, the
combined effect is also a (possibly different) way of shuffling the
elements of X.
So, we have a binary operation defined on Sym( X ). This operation
is associative, since composition of functions is associative. There is
an identity element: the identity permutation : X X which maps
every element to itself. Also, every permutation has an inverse 1 ,
which can be regarded as either the inverse function 1 : X X or
as the permutation which puts all of the elements of X back to how
they were before we applied . Hence Sym( X ) forms a group:
Definition 1.51 Let X be a (possibly infinite) set. The group Sym( X )
of all permutations : X X is called the symmetric group on X.
If X = {1, . . . , n} is a finite set consisting of n elements, then we call
Sym( X ) the symmetric group on n objects and denote it Sn .
choices for where 1 maps to, then n1 choices for where 2 goes (since
we can map 2 to any of the remaining numbers except for the one we
mapped 1 to), n2 choices for where 3 maps to, and so on. This gives
us n(n1)(n2) . . . 1 = n! possible permutations of n objects, and so
|Sn | = n!.
It so happens that S2
= Z2 and S3 = D3 . The symmetric group S4
is isomorphic to the symmetry group of the tetrahedron. Later on
we will meet Cayleys Theorem, which states that any group can be
regarded as a group of permutations (although not, in general, the full
symmetric group Sn for some value of n).
To investigate these permutation groups, we need a coherent and
consistent notation, at least for permutations on finite sets.
One method, given that a permutation : X X is determined
completely by its action on the elements of the set X, is to represent it
in the form of an array:
x1 x2 xn
h
...
i
( x1 ) ( x2 ) . . . ( x n )
The first row lists the elements of X, and the second lists their images
under the action of . So, suppose that S5 maps 1 7 1, 2 7 3,
3 7 5, 4 7 4 and 5 7 2. Then we can represent as
h
1 2 3 4 5 .
i
1 3 5 4 2
The second row of this array shows the effect of applying to the first
row. Suppose we have another permutation S5 such that 1 7 2,
2 7 4, 3 7 5, 4 7 1 and 5 7 3. This can be represented as
h
1 2 3 4 5 .
i
2 4 5 1 3
Composition of permutations can be represented fairly simply using
this notation:
h
1 2 3 4 5
i
1 3 5 4 2
i = 1
h
2 3 4 5
i
= h 2 5 3 1 4
1 2 3 4 5
2 4 5 1 3
Note Because we regard permutations as bijections, and the product
operation as being composition, we write products from right to left,
rather than left to right. So means followed by . Stacking the
arrays vertically, we read down the page, so means followed by .
In general, composition of permutations isnt commutative, so we have
to be careful of the order. For example,
h
1 2 3 4 5
i
2 4 5 1 3
i = 1
h
2 3 4 5 6= .
i
= h 3 4 2 1 5
1 2 3 4 5
1 3 5 4 2
26 a course in abstract algebra
This notation is quite clear, and certainly makes it easy to work out the
1
composite of two permutations, but unfortunately it becomes some-
5 2 what unwieldy with large numbers of permuting elements. Also, it
doesnt tell us very much about the actual structure of the permutation.
4 3
For example: the permutation maps 2 7 3, 3 7 5 and 5 7 2,
but leaves 1 and 4 unchanged. Repeated applications of cause the
elements 2, 3 and 5 to cycle amongst themselves, while 1 and 4 stay
1
where they are. Furthermore, if we apply three times, we get the
5 2 identity permutation . Hence 3 = ; equivalently, has order 3 in S5 .
The permutation , on the other hand, maps 1 7 2, 2 7 4 and
4 3 4 7 1, but additionally swaps 3 5. So is doing two different,
1 independent things at the same time: cycling 1 7 2 7 4 7 1 and
transposing 3 5. Neither of these two operations interfere with each
5 2 other, because they are acting on disjoint subsets of the set {1, 2, 3, 4, 5}.
If we apply three times, then the subset {1, 2, 4} will be back where it
4 3 started, but meanwhile 3 and 5 will have swapped places three times,
and will hence be the other way round from where they started. But if
1
we apply another three times, {1, 2, 4} will go round another period-
three cycle, and 3 and 5 will have swapped back to their original
5 2 places. More concisely, 6 = .
We can depict this situation graphically, as shown in Figure 1.12, but
4 3 this also isnt very compact, and tends to become more involved for
Figure 1.12: Graphical depictions of larger numbers of permuting objects.
permutations , , and in S5
Ideally, we would like a new notation which is at the same time
more compact, and also makes it easier to see at a glance the internal
structure of the permutation. The key is to split the permutation into
disjoint cyclic subpermutations, and write them as parenthesised lists.
Fortunately, there is a relatively straightforward procedure for decom-
posing a permutation as a product of disjoint cycles:
Algorithm 1.53
Open parenthesis (
Start with the first element 1 and write it down.
Next write down the image (1) of 1.
Next write down the image ( (1)) = 2 (1).
..
.
When we get back to 1, close the parentheses ).
Now repeat this process, starting with the smallest number not yet
seen, and so on until you have a list of parenthesised lists of numbers.
In practice, we delete any single-number lists, because they just tell
us that the number in question isnt changed by the permutation.
groups 27
= (2 3 5)(1 2 4)(3 5)
Start with 1, and read through the list of cycles from right to left,
applying each one in turn until youve done them all:
(3 5) (1 2 4) (2 3 5)
1 7 1 7 2 7 3
Now do the same process, but starting with the number (in this case 3)
28 a course in abstract algebra
Cycles of length 2 will play an important rle in the next part of the
discussion, so we give them a special name:
Definition 1.56 Let = ( x1 x2 ) be a period2 cyclic permutation in
some (possibly infinite) symmetric group Sym( X ). Then we call a
transposition.
gives one valid way, but since any cyclic permutation ( x1 x2 . . . xk ) can
be rewritten as, for example, ( xk x1 x2 . . . xk1 ) (or any other cyclic
permutation of that list), the decomposition will not be unique.
Proposition 1.60 The symmetric group Sn is generated by any of the
following sets of cyclic permutations:
(i) {(1 2), (1 3), . . . , (1 n)}
(ii) {(1 2), (2 3), . . . , (n1 n)}
(iii) {(1 2), (1 2 . . . n)}
= ( x1 x2 )( x1 x3 ) . . . ( x1 xn )( x2 x3 ) . . . ( xn1 xn ).
Given a permutation Sn , we define
( P)( x1 , . . . , xn ) = P( x(1) , . . . , x(n) ).
The permuted polynomial ( P)( x1 , . . . , xn ) has exactly the same fac-
tors ( xi x j ) as P( x1 , . . . , xn ), except that some of the variables xi have
been permuted, and some of the factors will change sign. We may
therefore define the sign of to be the quotient
( P)( x1 , . . . , xn )
sign( ) = = 1.
P ( x1 , . . . , x n )
In general, for any two permutations , Sn ,
groups 31
Summary
28
Definition 1.22, page 12. many of them, say that the group is finitely generated.28 In the case of
the integers Z or the finite cyclic groups Zn , we can recover the whole
of the group by taking finite products (or sums) of a single nontrivial
generator; groups which are generated by a single element are said
29
Definition 1.23, page 12. to be cyclic.29 There is only one cyclic group (up to isomorphism) of
30
Proposition 1.24, page 13. a given (finite or infinite) order.30 The order of a (finite or infinite)
31
Proposition 1.25, page 13. cyclic group is equal to the order of any of its generators.31 For the
finite cyclic groups Zn , a nonzero integer k Zn generates the whole
32
Proposition 1.26, page 13. group if and only if k and n are coprime.32
Given two groups G and H we can form a new group by taking the
cartesian product G H and defining the obvious combined group
structure on it; we call this the direct product G H (or, if were using
33
Definition 1.27, page 14. additive notation, the direct sum G H).33 , 34 In particular, the Klein
34
Proposition 1.28, page 14. 4group V4 is isomorphic to the direct sum Z2 Z2 ,35 and in general
35
Example 1.30, page 15. Zm Zn = Zmn if and only if m and n are coprime.36
36
Proposition 1.32, page 16. Matrices yield an important and rich class of groups, many of which
are nonabelian. Among these are the general linear groups GLn (K)
37
Example 1.34, page 17. of invertible matrices,37 the orthogonal groups On (K) of matrices
38
Example 1.36, page 18. whose inverse is equal to their transpose,38 and the unitary groups of
39
Example 1.38, page 18. complex matrices whose inverse is equal to their conjugate transpose.39
(Here K can be either Q, R or C, and later we will consider the case of
a finite field F p .) Each of these groups has a special version consisting
40
Example 1.35, page 17. of just the matrices of the given type with determinant 1,40 , 41 , 42 and
41
Example 1.37, page 18. in addition we have the symplectic groups Sp2n (K) of symplectic
42
Example 1.39, page 18. matrices.43 An interesting and important finite example is given
43
Example 1.40, page 18. by Paulis spin matrices; this is closely related to the eight-element
44
Example 1.42, page 19. quaternion group Q8 .44
Matrix groups led naturally to a discussion of geometric operations on
vector spaces, and in particular we considered isometries: operations
which preserve distance. Isometries which preserve orientation are
said to be direct, while those which reverse orientation are said to be
45
Definition 1.46, page 22. opposite.45 Considering those isometries which map a particular geo-
metric object to itself leads to the notion of symmetry groups such as
the dihedral groups Dn : these comprise the 2n symmetry operations
46
Definition 1.43, page 21. (reflections and rotations) we can perform on a regular ngon.46 The
Klein 4group V4 can similarly be viewed as the symmetry group of a
47
Example 1.45, page 21. rectangle.47 More generally, the Euclidean groups En+ and En are de-
fined as, respectively, the direct and full isometry groups Isom+ (Rn )
and Isom(Rn ) of Rn , and in particular we characterised the various
48
Definition 1.47, page 22. types of elements in the two- and three-dimensional case.48 , 49
49
Definition 1.48, page 22. In turn, symmetry groups led to a discussion of permutations: bijec-
50
Definition 1.50, page 24. tions : X X from a set to itself.50 Composition of permutations is
groups 35
Exercises
1.1 Let be a binary operation defined on R such that x y = x + y + xy.
(a) Is associative?
(b) Is commutative?
(c) Is there an element e R which serves as an identity with respect to ?
(d) Does every element x R have an inverse x 1 with respect to ?
1.2 Let S = { a, b} be a set with two elements. How many possible binary operations can be defined
on this set? How many of these determine semigroups, that is, how many are associative? How
many of those determine monoids? How many determine a group structure?
1.3 Which of the following binary operations (defined on the set R of real numbers) are associative,
and which are commutative?
(a) x y = x + 2y
(b) x y = x + y + xy + 1
(c) x y = x
1.4 Show that the multiplication table for a finite group G satisfies the Latin square property. That
is, show that each element of the group occurs exactly once in each row and column of the
table.
1.5 Let : RR R by x y = xy + 1. Is this operation commutative? Does it determine a group
structure on R? If so, prove it by verifying the necessary criteria; if not, which axioms fail?
1.6 Which of the following are groups? For any that are not, say which of the axioms G0G3 fail.
(a) (Z, ) (e) (Q, ) where ba dc = ba+
+c , for a, b, c, d Z and b, d > 0
d
(b) (Z3 , 3 ) (f) (Zeven , +)
(c) (Z4 , 4 ) (g) (Zodd , +)
(d) (Zn , n ) if n isnt prime (h) (Z, )
Here, represents the usual multiplication operation defined on Z, and n represents multi-
plication modulo n.
1.7 Given a group G, and some fixed element g G, show that the function g : x 7 g1 xg is an
isomorphism G G.
1.8 Write out the group multiplication table for Z2 Z5 .
1.9 Let G be a group. Show that if ( ab)2 = a2 b2 for all a, b G, then G must be abelian.
1.10 Let G be a nonempty set and let : G G G be an associative binary operation. Suppose
that there is a distinguished element e G such that:
(a) e x = x for all x G. (That is, e is a left identity.)
(b) For any x G, there exists y G such that y x = e. (That is, G has left inverses.)
Then show that G = ( G, ) is a group. (This demonstrates that the conditions in Definition 1.9
are slightly stronger than necessary.)
groups 37
1.11 Let G be a group, and suppose that g G and m, n Z. Prove that ( gm )n = gmn and
gm gn = gm+n .
1.12 Show that | a b| = |b a| for any elements a, b of some group G = ( G, ).
1.13 Show that any finite group of even order has an odd number of elements of order 2.
1.14 Show that Isom(3 ), the full isometry group of the tetrahedron, is isomorphic to the symmetric
group S4 , and that the direct isometry group Isom+ (3 ) is isomorphic to the alternating group
A4 .
1.15 Let = 12 23 35 41 54 66 be an element of S6 . Write this permutation as a product of disjoint cycles,
to be the rotation group R3 = Isom+ (2 ) = Z3 . So in some sense, D3 Table 2.1: The multiplication table for
the dihedral group D3 with the sub-
has a copy of R3 (or, for that matter, Z3 ) inside it. group {e, r, r2 } highlighted
Similarly, we could throw away everything except the identity and
e r r 2 m1 m2 m3
a single reflection (m1 for example) and still be left with an order2 e e r r 2 m1 m2 m3
group isomorphic to Z2 . So, in the same sense, D3 also has a copy r r r 2 e m3 m1 m2
r 2 r 2 e r m2 m3 m1
(more accurately, three copies) of Z2 embedded in it. m m m2 m3 e r r 2
1 1
Similarly, looking at the multiplication table for Z4 reveals an order2 m2 m2 m3 m1 r 2 e r
m3 m3 m1 m2 r r 2 e
group {0, 2} = Z2 inside it.
Table 2.2: The multiplication table for
In this chapter, we will make this concept more precise, and develop the dihedral group D3 with the sub-
these ideas to learn more about the internal structure of groups. Along group {e, m1 } highlighted
the way, we will prove Lagranges Theorem, which imposes an im- + 0 1 2 3
portant necessary (but not sufficient) condition on the order of these 0 0 1 2 3
1 1 2 3 0
groups-within-groups, and use it to prove Eulers Theorem and Fer- 2 2 3 0 1
mats Little Theorem, two important and useful number-theoretic 3 3 0 1 2
results about the factorisation of integers. Table 2.3: The multiplication table for
the cyclic group Z4 with the subgroup
{0, 2} highlighted
2.1 Groups within groups On the other side, the general opinion
has been and is that it is indeed by
experience that we arrive at the mathe-
The main concept we will explore in this chapter is that of matics, but that experience is not their
a subgroup: a smaller group neatly embedded inside a larger one. proper foundation: the mind itself con-
tributes something.
More precisely, given some group G = ( G, ), we are interested in
Arthur Cayley (18211895),
finding a subset H G which is a group in its own right under the Presidential Address to the British
same binary operation as G. Not just any subset will do, however, Association for the Advancement of
Science (1883)
so we need to decide what criteria we want a given subset to satisfy.
40 a course in abstract algebra
h1i
h5i h3i
h2i
h25i h15i h9i
h10i h6i
h4i h75i h45i
h50i h30i h18i
h20i h12i h225i
h150i h90i
h100i h60i h36i
h450i
h300i h180i
h900i
Again, all of the subgroups are cyclic, generated by some single ele-
ment of Z. (The trivial subgroup isnt, and cant be, shown: it appears
as the infinitely distant bottom vertex of the lattice, contained in all of
the other subgroups.) We will often write the set
hni = {. . . , 3n, 2n, n, 0, n, 2n, 3n, . . .}
of integer multiples of n as nZ.
These observations lead us to ask whether all subgroups of (finite or
44 a course in abstract algebra
infinite) cyclic groups are themselves cyclic, and it turns out that the
answer is yes:
Proposition 2.8 Let G = h gi be a (finite or infinite) cyclic group. If
H < G then H is also cyclic.
D4
{e}
The subgroups on the second level are all of order 4; two of them
({e, r2 , m2 , m4 } and {e, r2 , m1 , m3 }) are isomorphic to the Klein group
V4 , while the other (the rotation subgroup {e, r, r2 , r3 }) is isomorphic
to Z4 . The subgroups on the third level are all isomorphic to Z2 .
Q8 For any dihedral group Dn , the rotation subgroup {e, r, r2 , . . . , r n1 }
is a cyclic subgroup since it can be generated by the single element
{ E, I } { E, J } { E, K }
r. And since it has order n, then by Proposition 1.24, it must be
isomorphic to Zn .
{ E}
The quaternion group Q8 (see Example 1.42) has the lattice diagram
{ E} shown in Figure 2.6.
Figure 2.6: Lattice diagram for the This diagram is obviously different in structure to that for D4 , and
quaternion group Q8
hence D4 6
= Q8 .
subgroups 45
Proof Since both H and K are subgroups of G, they both contain the
identity e of G, and hence their intersection H K does as well, and
is therefore not empty. Given elements g1 , g2 H K, it follows that
both g1 , g2 H and g1 , g2 K. Therefore their product g1 g2 H
and also g1 g2 K, hence g1 g2 H K. Thus H K is closed
under the induced multiplication operation. Finally, we need to show
that H K contains all the necessary inverses. But if g H K then
g H and so g1 H; also g K and so g1 K as well. Therefore
g1 H K as required, and hence H K is a subgroup of G.
Since H K H, and we have just seen that it forms a group under
the induced multiplication operation, it must therefore be a subgroup
of H. By an almost identical argument, H K is a subgroup of K.
We now introduce a related concept and proposition that will come in
useful in the next section when we classify groups of order 6.
Definition 2.10 Let H, K < G be two subgroups of a group G. Then
denote by HK the set
HK = {h k : h H, k K }.
If we are using additive notation, we may choose to denote this as
H + K = {h + k : h H, k K }.
Note that if G is nonabelian, it neednt be the case that HK = KH.
If, however, the elements of H commute with the elements of K then
HK = KH is a subgroup of G:
Proposition 2.11 Let H and K be subgroups of a group G. Then HK is
a subgroup of G if and only if HK = KH.
Isom+ (3 )
{e, s A , s B , sC }
The key idea of the rest of this proof is to use these permutations g
to construct a subgroup of Sym( G ) which happens to be isomorphic
to G itself.
Let S = { g : g G } be the set of these left-multiplication permuta-
tions. We now need to show that this subset S Sym( G ) is actually a
subgroup of Sym( G ), and furthermore that its isomorphic to G.
The group operation on S is the induced operation from Sym( G ),
which is just composition of permutations. So we need to show that S
is closed under composition, and also that 1
g S for all g G.
To check closure we need to examine how composition works in S and
how this relates to multiplication in G. Given g, h, k G we have
( g h )(k) = g (h (k)) = g (h k) = g h k = gh (k).
So g h = gh , and since, for any g, h G, the product g h is also
in G, it follows that gh is in S. Hence S is closed under composition.
Next we have to check the existence of inverses. As noted earlier,
the function e is the identity map idG , which is just the identity
permutation Sym( G ). So the inverse 1
g of g is the permutation
h such that h g = e = g h . But we know from the previous
paragraph that g h = gh , so what were really looking for is
h G such that h g = e, which rle is clearly served by h = g1 , and
hence 1
g = g 1 .
So S is closed under composition, and contains both an identity ele-
ment and a full set of inverses, and is therefore a subgroup of Sym( G ).
All thats left is to show S
= G, which requires us to find a bijection
f : G S satisfying the structural condition in Definition 1.17.
An obvious candidate for this isomorphism f is the map that takes an
element g to its corresponding permutation g S; that is, f : g 7 g .
Then f is injective, because if f ( g) = g = h = f (h) for some
g, h G, it follows that g (k) = g k = h k = h (k) for every
element k G, and by the right cancellation law (1.5) it follows that
g = h and hence f is injective. The definition of S immediately
confirms that f is surjective: g S precisely when g G.
All that remains is to show that f respects the group structure:
f ( g ) f ( h ) = g h = gh = f ( g h )
This completes the proof.
To make sure we understand this, lets try a concrete example.
Example 2.14 Let G = Z3 . Cayleys Theorem says we can find a
permutation group isomorphic to Z3 = {0, 1, 2}, and by the construc-
tion in the above proof we know that were actually looking for a
subgroups 49
subgroup of Sym(Z3 ) = S3 .
The permutation 0 is obviously the identity permutation , since
0 (n) = 0 + n = n.
The permutation 1 is defined by 1 (n) = 1 + n (mod 3), which can
be written as 01 12 20 , or (0 1 2) in cycle notation.
So Z3
= {0 , 1 , 2 } = {, (0 1 2), (0 2 1)} < S3 .
This example concerns a finite group, but theres nothing in our proof
of Cayleys Theorem that relies on the groups G, Sym( G ) or S being
finite, and in fact the theorem works for infinite groups too:
Example 2.15 Let G = Z, which is countably infinite. Then for any
integer n Z, the function n : Z Z is defined by n (m) = n + m
for all m Z. This is clearly a bijection on Z; it permutes the integers
by shifting everything along by n places in the appropriate (positive
or negative) direction. The group S = {n : n Z} is a subgroup of
the full permutation group Sym(Z), and the isomorphism f : Z S
that we want is precisely the function defined by mapping an integer
n to the shift everything along by n places permutation n .
2.2 Cosets and Lagranges Theorem The reader will find no figures in this
work. The methods which I set forth
do not require either constructions or
In the proof of Cayleys Theorem we introduced permutations geometrical or mechanical reasonings:
g : G G which we used to show that any group can be viewed as but only algebraic operations, subject
to a regular and uniform rule of proce-
a permutation group. These permutations are defined on the whole dure.
of G, but since part of our motivation for introducing the concept of Joseph-Louis Lagrange (17361813),
a subgroup was to investigate the internal structure of groups, its preface to Mcanique Analytique (1788)
r ( R3 ) = {r, r2 , e} = R3 , m2 ( R3 ) = {m2 , m3 , m1 } = D3 \ R3 ,
r2 ( R3 ) = {r2 , e, r } = R3 , m3 ( R3 ) = {m3 , m1 , m2 } = D3 \ R3 .
Immediately, we notice that (apart from some inconsequential reorder-
ing) the image g ( R3 ) has one of two different forms. If g R3 then
g ( R3 ) = R3 . (This is exactly what wed expect, since R3 , being a
subgroup, is closed under the induced multiplication operation.) If,
on the other hand, g 6 R3 then g ( R3 ) = {m1 , m2 , m3 } = D3 \ R3 .
So the action of the various permutations on the subgroup R3 neatly
partitions D3 into two different subsets. These images g ( R3 ) look
like they might be a useful concept, so we give them a special name:
Definition 2.16 Let G be a group, and suppose H < G is a subgroup
of G. Then, for a given element g G, the left coset gH is defined
to be the subset g ( H ) = { g h : h H } G.
Table 2.8: Multiplication table for the In this case as well, the corresponding left and right cosets coincide.
quaternion group Q8
All the examples weve seen so far are of finite groups. These concepts
carry across to the case where the group under investigation is infinite.
Example 2.22 Consider the (infinite) cyclic subgroup 3Z = h3i of
the (infinite) cyclic group Z. This has three distinct left cosets (which,
since Z is abelian, are each identical to the corresponding right coset):
0+3Z = {. . . , 6, 3, 0, 3, 6, . . .} = 3Z+0,
1+3Z = {. . . , 5, 2, 1, 4, 7, . . .} = 3Z+1,
2+3Z = {. . . , 4, 1, 2, 5, 11, . . .} = 3Z+2.
Example 2.23 Recall that the orthogonal group O3 (R) consists of all
33 real orthogonal matrices; that is, matrices A such that A1 = A T .
These must have determinant 1 since if A T = A1 then det A T =
det( A1 ) = (det A)1 . But also det A T = det A in general. By
putting these together we have det A = (det A)1 , or (det A)2 = 1
which implies that det A = 1.
The special orthogonal group SO3 (R) is the subgroup of orthogonal
matrices which have determinant +1. It has two cosets, namely
SO3 (R) itself and the complement O3 (R) \ SO3 (R), comprising all
orthogonal matrices with determinant 1.
For any special orthogonal matrix A, we see that
A SO3 (R) = SO3 (R) = SO3 (R) A
since SO3 (R) is closed under matrix multiplication.
If, on the other hand, B is orthogonal with determinant 1, then BA
is also orthogonal, because O3 (R) is itself a group and hence closed
subgroups 53
One thing all of these examples have in common is that the cosets
partition the group: their union is the whole group, and no two cosets
intersect, so the group neatly splits into a (possibly infinite) collection
of non-overlapping cosets. The next proposition confirms this is true Figure 2.8: Cosets of R < C
in general; more importantly it will help us prove Lagranges Theorem,
an important result about the internal structure of finite groups.
Proposition 2.26 Let H < G be a subgroup of a (possibly infinite) group
G, and let g, k G be two arbitrary elements of G. Then the following
three statements are equivalent:
(i) k Hg,
(ii) Hg = Hk, and
(iii) kg1 H.
Proof First we prove that (i) = (ii). Suppose that k Hg. Then
there exists a unique h H such that h g = k. Multiplying on
the left by h1 gives g = h1 k. As usual, to prove Hg = Hk we
need to show that Hg Hk and Hk Hg. So, choose some element
a Hg. Then a = b g for some b H, and so a = b h1 k, which
means that a Hk (since b h1 H as both b, h1 H). Hence
Hg Hk. Similarly, choose c Hk. Then c = d k for some d H,
so c = d h g, which means that c Hg (because d h H as both
d, h H). Hence Hk Hg and therefore Hg = Hk.
The converse (ii) = (i) holds as well: if Hg = Hk then since k Hk,
it follows that k Hg as well.
Next we show that (i) = (iii). If k Hg then, as noted above, there
is a unique h H such that h g = k. Multiplying on the right by g1
Wikimedia Commons / Robert Hart (fl.1830s)
Joseph-Louis Lagrange (17361813),
born Giuseppe Lodovico Lagrangia in
gives k g1 = h H as required.
Turin, was an Italian and French math- Finally, we show the converse (iii) = (i). If h = k g1 H then it
ematician and physicist who made
many important contributions to alge- follows almost immediately that k = hg Hg.
bra, analysis, number theory and clas-
sical mechanics. His Mcanique Analy-
This proposition has a couple of important corollaries:
tique (1788) was the most comprehen- Corollary 2.27 Two right cosets Hg and Hk are either equal or disjoint.
sive treatment of classical mechanics
since Newtons celebrated Philosophi
Proof Suppose a Hg Hk. Then, by Proposition 2.26, since a Hg,
Naturalis Principia Mathematica (1687).
In 1766 he succeeded Leonhard Euler it follows that Ha = Hg; also since a Hk we have Ha = Hk and
as director of mathematics at the Prus- hence Hg = Hk. The alternative is that no such element a exists, in
sian Academy of Sciences in Berlin and
which case Hg Hk = .
remained there for twenty years, mov-
ing to Paris in 1786 at the invitation of Corollary 2.28 The right cosets of H in G partition G.
Louis XVI. His first few years in Paris
were marked by a prolonged episode The next proposition tells us an important fact about the relative sizes
of depression and the political insta-
bility of the Revolution; he survived right cosets of a finite subgroup H < G.
both and was appointed to a chair in
Proposition 2.29 Let H be a subgroup of a (possibly infinite) group G.
analysis at the cole Polytechnique in
1794, and a chair in mathematics at the Then | Hg| = | H | for any element g G.
cole Normale in 1795. Contemporary
accounts indicate that he was consid- Proof By the right cancellation law in Proposition 1.15, if h1 g =
erably less gifted at teaching than at
h2 g for some h1 , h2 H and g G, then h1 = h2 . This means
research.
He was elected a Fellow of the Royal that the function f : H Hg given by h 7 h g is a bijection, so
Society of London in 1806, appointed | H | = | Hg| as required.
a Grand Officier of the Lgion dHonneur,
and a Comte de lEmpire by Napoleon These last two facts, Corollary 2.28 and Proposition 2.29, give the
in 1808, who later awarded him the following elegant result.
Grand Croix of the Ordre Imprial de la
Runion a week before his death in 1813. Theorem 2.30 (Lagranges Theorem) Let H be a subgroup of a finite
He was granted the honour of a tomb group G. Then the order | H | of H divides the order | G | of G.
in the crypt of the Panthon, and he
is one of 72 eminent French scientists We can use this fact, that the order of a subgroup H is a factor of the
commemorated on the Eiffel Tower.
whole group G, to devise some measure of the relative sizes of G and
G, simply by looking at the quotient | G |/| H |. Lagranges Theorem
tells us that this will always be a positive integer.
subgroups 55
For example, the rotation subgroup R3 < D3 is clearly half the size of
D3 itself, so this quotient | D3 |/| R3 | = 2. Similarly, |Sn |/| An | = 2 for
any n N. The subgroup h6i < Z12 is obviously smaller relative to
the whole group Z12 , and indeed the quotient |Z12 |/|h6i| = 6.
So, the larger this quotient is, the smaller the subgroup in question
is relative to the full group. We might wonder if this concept can be
extended to the case of infinite groups, and indeed it can. The quotient
| G |/| H | is really just the number of distinct right cosets of H in G, so
by looking at this way we can formulate the following definition:
Definition 2.31 The index | G:H | of a subgroup H < G is the num-
ber of right cosets of H in G. If G is finite then | G:H | = | G |/| H |.
none of them can have more than p elements since there are at most p
distinct cyclic permutations of a given ptuple, and if all of them had
p elements, the total number of ptuples in X would be of the form
mp+1 for some integer m; we know that it cant because we showed a
little earlier that | X | has to be a multiple of p.
So there must be at least one equivalence class with fewer than p
elements.
If this class consists of just one ptuple then weve found the orderp
element were looking for: any such ptuple must have g1 = g p
otherwise cyclic permutation would yield at least one other distinct
ptuple; given such an element ( g, . . . , g) we immediately have g p = e
as required.
So, suppose we have an equivalence class comprising fewer than p but
Wikimedia Commons / Zphirin Belliard (17981861)
more than one distinct ptuples. Then two of them must be equal: after Jean Roller (17981866)
Augustin-Louis Cauchy (17891857)
( g r + 1 , . . . , g p , g1 , . . . , g r ) = ( g s + 1 , . . . , g p , g1 , . . . , g s ) made numerous contributions to many
areas of mathematics: during his career
Suppose, without loss of generality, that r < s and then cycle back r he published 789 papers and several
times (or forward ( pr ) times) to get books on mechanics, real and complex
analysis, number theory, algebra and
( g1 , . . . , g p ) = ( g k + 1 , . . . , g p , g1 , . . . , g k ) geometry, and has more concepts and
theorems named after him than any
where k = s r. Then we find that gi = gk+i for all 1 6 i 6 p and so other mathematician.
After a short career as a civil engineer
g1 = gk+1 = g2k+1 (mod p) = = g ( p 1) k +1 (mod p) . (2.1) in Cherbourg, he returned to Paris in
1812 and undertook research in mathe-
Now suppose ak + 1 bk + 1 (mod p) where 0 6 a < b 6 p1. Then matics, publishing papers on geometry
p divides (b a)k. But this cant happen, because p is prime and both and algebra. Over the next few years
he was appointed to chairs at the cole
(b a) and k are strictly less than p. Polytechnique, the Collge de France
and the University of Paris, and elected
So the numbers 1, k+1, 2k+1, . . . , ( p1)k+1 are all different modulo
to the Acadmie des Sciences.
p. Furthermore, there are p of them, and so (modulo p) we have some Following the July Revolution in 1830
permutation of the numbers 1, . . . , p. Substituting these into (2.1) and he refused to swear allegiance to the
new king, Louis Philippe I, and was
rearranging, we get
dismissed from his academic posts. He
g1 = g2 = = g p . subsequently became tutor to the ex-
iled Henri dArtois, nephew and heir to
Hence our ptuple ( g1 , . . . , g p ) = ( g1 , . . . , g1 ), and multiplying ever- the previous king Charles X. This was
p a disaster: Henri acquired a lifelong ha-
thing out we get g1 = e as required.
tred of mathematics and Cauchy was
We can use this theorem to help classify (that is, make a complete list unable to do much research, although
he was granted the title of Baron in
of) groups of order 4, 6 and 8. (Groups of orders 1, 2, 3, 5 and 7 have recognition of his service.
already been classified: the trivial group is the only possible group of Returning to Paris in 1838, he contin-
order 1, and by Proposition 2.35 there is only a single group, up to ued his research but didnt regain his
academic posts until Louis Philippe
isomorphism, of a given prime order p, namely Z p .) was deposed ten years later. He re-
We start with groups of order 4: mained a professor at the University
until his death at the age of 67. He
Proposition 2.38 A group of order 4 must be isomorphic either to the is one of 72 eminent scientists whose
names are inscribed on the Eiffel Tower.
cyclic group Z4 or the Klein 4group V4 .
60 a course in abstract algebra
Mathematics is the queen of sciences 2.3 Eulers Theorem and Fermats Little Theorem
and arithmetic the queen of mathemat-
ics. She often condescends to render
service to astronomy and other natural In this section, we examine two applications of Lagranges Theorem
sciences, but in all relations she is enti- to number theory, and revisit Eulers totient function .
tled to the first rank.
Carl Friedrich Gauss (17771855), First we introduce some finite multiplicative groups.
quoted by Wolfgang Sartorius von
Walterhausen (18091876), in Gauss
Definition 2.41 Let Z n consist of those integers 0, . . . , n 1 which
zum Gedchtnis (1856) form a group under modulon multiplication. The identity element
is clearly 1, and all the other elements are those integers 0 6 k 6 n1
for which there exists a modulon multiplicative inverse; that is,
an integer 0 6 h 6 n1 where hk 1 (mod n). Clearly 0 is not
invertible, so 0 6 Z n.
In fact, Z
n consists of exactly those integers 0 6 k 6 n 1 for which
gcd(k, n) = 1.
If, on the other hand, k and n are coprime, then by the argument
in Proposition 1.26 and the following paragraph, k generates Zn .
Hence there exists m Z such that mn = 1 Zn . Or, equivalently,
mk = an + 1 Z. Now let l = [m]n , the remainder or residue of m
modulo n. Then lk = bn + 1 Z for some b Z, hence lk = 1 Z n
and so k is an invertible element with inverse l, and is thus in Z
n.
x n + yn = zn (iii) The only integers between 1 and pn that arent coprime to pn are
n n1 of these, and hence ( pn ) =
for n>2, remained unproven until 1, p, 2p, . . . , p . There are exactly p
n
1994, when Andrew Wiles proved the p p n 1 n 1
= p ( p 1).
TaniyamaShimura Conjecture on ratio-
nal elliptic curves and modular forms, The next proposition is a clever application of Lagranges Theorem.
which was known to imply Fermats
conjecture as a corollary. Proposition 2.47 Suppose n > 3. Then (n) is even.
The original statement of the conjec-
ture occurs in Fermats hand-annotated Proof Let G = Z n and let H = {1, n 1}. It is simple to show that H
copy of the treatise Arithmetica, written is a subgroup of G: it is closed under multiplication modulo n, with
in the third century AD by the Greek
mathematician Diophantus of Alexan- the only nonimmediate case being
dria. Fermat notes that he has discov-
ered a marvellous proof (demonstra- (n1)(n1) = n2 2n + 1 1 (mod n).
tionem mirabilem) of the insolubility of
the general case, but that the margin is By Lagranges Theorem, | H | = 2 must be a factor of | G | = |Zn| =
too small to contain it (hanc marginis (n), and so (n) is even. (If n = 2 this fails, because n1 = 1 and
exiguitas non caperet).
hence the group H = {1} has only one element; indeed, we saw in
Born in Gascony, the son of a merchant
named Dominique Fermat, he stud- Example 2.42 that |Z2 | = (2) = 1.)
ied law at the University of Orlans,
and then embarked on a legal career. Corollary 2.48 (Fermats Little Theorem) Let p N be prime. Then
In 1631, he was appointed as a coun- for any k Z,
cillor and judge at the Parlement de
Toulouse, thereby entitling him to add kp k (mod p)
the honorific de to his surname.
or, equivalently, p|(k p k ).
Fluent in six languages (French, Latin,
Greek, Occitan, Spanish and Italian), he
was also an accomplished writer and
Proof Proposition 2.46(i) tells us that ( p) = p1, and then by Eulers
poet. He communicated most of his Theorem we have
mathematical discoveries informally, in
letters to friends, usually omitting the k p = k ( k p 1 ) = k ( k ( p ) ) k (mod p)
proofs, which were reconstructed by
other mathematicians after his death. as required.
subgroups 65
Summary
If you can find a copy, Singhs 1996 BBC Horizon documentary Fermats Last Theorem is also worth
watching: at the time of writing, it is available to watch online via the BBCs website.
An interesting account of the history and subsequent development of Cauchys Theorem can be found
in the following article:
M Meo, The mathematical life of Cauchys group theorem, Historia Mathematica 31 (2004) 196221
Cauchys original exposition and proof ran to about nine pages at the end of a long article on
permutation groups published in 1845.
A L Cauchy, Mmoire sur les arrangements que lon peut former avec des lettres donnes, et sur les permu-
tations et substitutions laide desquelles on passe dun arrangement un autre, Exercises danalyse et de
physique mathmatique 3 (1845) 151252
Meo also notes that Cauchys proof contains a subtle error: at one point he assumes that the direct
product of two subgroups is also a subgroup. This is true if both subgroups are normal, as we will see
in Corollary 3.23 in the next chapter, but not in general.
An elegant and concise proof, just ten lines long, was published in 1959 by the American mathematician
James McKay (19282012):
J H McKay, Another proof of Cauchys group theorem, American Mathematical Monthly 66.2 (1959) 119
Exercises
2.1 For each of the following groups G and subsets H G, say whether or not H is a subgroup of G;
if not, explain briefly why not.
2.2 Repeat Example 2.14 with the cyclic group Z4 and the Klein group V4 = Z2 Z2 .
2.3 Let m, n N such that n > 2 and 0 6 m 6 n. Let H be the subgroup of the symmetric group
Sn consisting of those permutations which permute the first m elements among themselves. Use
Lagranges Theorem in a manner analogous to Proposition 2.47 to show that the binomial coefficent
(mn ) = m!(nn!m)! is an integer.
2.4 Find the orders of Z
n for n = 10, 12, 21, 60. In the cases where |Zn | = 4, determine whether Zn is
isomorphic to Z4 or the Klein group V4 .
2.5 Find the subgroups of Z14 .
2.6 Let p be prime, and show that ( p1) is the only element of Z p of order 2. By considering the
product of all elements of Z p , prove Wilsons Theorem: that ( p1)! 1 (mod p) if and only if
p is prime.
subgroups 69
2.7 Let H be a subgroup of a group G, and suppose that g G is some arbitrary element. Show that
there exists some element k G such that gH = Hk.
2.8 (a) Show that a group of even order has at least one element of order 2.
(b) Show that a group of odd order has no elements of order 2.
2.9 (a) If G is abelian, show that H = { g G : g2 = e} is a subgroup of G.
(b) Find an example of a nonabelian group G such that H is not a subgroup of G.
2.10 (a) If G is abelian, show that H = { g2 : g G } is a subgroup of G.
(b) Find an example of a nonabelian group G for which H is a subgroup.
(c) Find a nonabelian group G for which H is not a subgroup.
Up and after ordering some things to-
wards my wifes going into the coun-
try, to the office, where I spent the
morning upon my measuring rules
very pleasantly till noon, and then
comes Creed and he and I talked about
mathematiques, and he tells me of a
way found out by Mr. Jonas Moore
which he calls duodecimal arithme-
3 Normal subgroups tique, which is properly applied to
measuring, where all is ordered by
inches, which are 12 in a foot, which I
have a mind to learn.
Samuel Pepys (16331703),
diary entry for Tuesday 9 June 1663
n the last chapter we introduced the concept of a coset of a sub-
I group H of a group G. In the case where G is abelian, obviously
each left coset gH is equal to the corresponding right coset Hg, but if G
is nonabelian this need not be the case, and often isnt. In this chapter
we will examine this topic further, along the way deriving a condition
on H that forces the left cosets to coincide with their right coset twins.
Subgroups H which have this property are called normal subgroups
and have an important rle to play: given a normal subgroup H of a
group G we can construct the quotient group G/H. This method of
decomposing larger groups into products of smaller ones gives us new
techniques for understanding their internal structure. This is (very
loosely) analogous to the number-theoretic study of the multiplicative
properties of integers by factorising them into smaller ones, and just
as we then find a class of integers, namely the prime numbers, which
cant be factorised into smaller ones, we will also investigate the class
of simple groups: those which have no normal subgroups (apart from
themselves and the trivial subgroup) and hence cant be factorised
into smaller ones.
3.1 Cosets and conjugacy classes The species and genus are always the
work of nature; the variety often that of
culture; and the class and order are the
It is an interesting question to ask when the left and right cosets work of nature and art.
of a subgroup H 6 G coincide: is there some property that H or G Carl Linnaeus (17071778),
Philosophia Botanica (1751) 162
might have that then forces gH = Hg for any element g G?
An obvious sufficient condition is that G (and hence, by Proposition 2.2,
H) be abelian, but we know alread that this isnt a necessary condition:
at the beginning of Section 2.2 we saw that the rotation subgroup
R3 6 D3 has this property, but D3 isnt abelian, although R3 = Z3 is.
Perhaps the subgroup H has to be abelian in order to have this coinci-
dence property. This is a reasonable suggestion, but the next example
72 a course in abstract algebra
It turns out that the answer is a bit more complicated, and in order to
understand it, we need to introduce a new concept: that of conjugacy.
Impatient or alert readers may wonder why it matters whether the left
and right cosets are the same. This question will be answered in the
next section, where well find that subgroups having this property are
useful for understanding the internal structure of the larger group.
Definition 3.2 An element g G is conjugate to an element h G
if there exists some element k G such that k g k1 = h. We say
also that h is the result of conjugating g with k.
This may seem like a strange criterion to impose on elements of a
group, but if youve studied a bit of linear algebra you should have
normal subgroups 73
one of the other axes. Thinking about it, we can get the same result
e r r2 m1 m2 m3 as applying m2 by rotating the whole triangle by a third of a rotation,
e e e e e e e doing m1 and then rotating the triangle back again. So m1 and m2 are
r r r r r2 r2 r2
conjugate to each other. On the other hand, m1 and r arent conjugate
r2 r2 r2 r2 r r r
m1 m1 m2 m3 m1 m3 m2 to each other, since there exists no g D3 such that m1 = grg1 .
m2 m2 m3 m1 m3 m2 m1
m3 m3 m1 m2 m2 m1 m3
Table 3.1 shows this more clearly. The group element at the intersection
of row g with column h is that obtained by conjugating g with h, that
Table 3.1: Conjugation in D3
is, hgh1 . Here we see that e is conjugate only to itself, r is conjugate to
r2 and vice versa, and the reflections m1 , m2 and m3 are all conjugate
The subtables to each other.
are examples of another interesting al- Proof To show that conjugacy is an equivalence relation we need to
gebraic structure: a quandle. These
show that it is reflexive, symmetric and transitive.
objects (and a slightly more general
structure called a rack) occur naturally Conjugacy is reflexive, since for any element g G we have e g
in the study of knot theory; the third,
for example, represents the nontrivial
e1 = g and hence g g.
3colouring of the trefoil knot. Conjugacy is symmetric, since if g h there exists some x G such
that h = x g x 1 . Multiplying both sides of this equation on the left
by x 1 and on the right by x yields
x 1 h x = x x 1 g x 1 x = e g e = g.
k = y h y 1 = y x g x 1 y 1 = ( y x ) g ( y x ) 1 .
Example 3.6 We know from linear algebra that some, but not all,
matrices in GLn (R) are similar to (or, in our more general terminol-
ogy, conjugate to) a diagonal matrix. More generally, every matrix
M GLn (R) is conjugate to one in Jordan canonical form. So, the
conjugacy classes of GLn (R) are indexed by the various types of nn
Jordan matrices.
A Jordan block with eigenvalue R of degree k is a kk matrix
J,k = [ ai j ] with diagonal elements ai i = , the supradiagonal ele-
ments ai i+1 = 1 for 1 6 i < k and all other elements ai j = 0 if j 6= i
or i +1.
A matrix is in Jordan canonical form if it can be decomposed as a
block diagonal matrix where the diagonal blocks are Jordan blocks,
and all the off-diagonal blocks are zero.
The conjugacy classes of matrices in GLn (R) are in bijective corre-
spondence with the different Jordan matrices.
So, considering GL2 (R) we have three different types:
diag( J,1 , J,1 ) = 0 0 , diag( J,1 , J,1 ) = 0 0 , J,2 = 0 1
This is true for any abelian group, as the following proposition shows.
Proposition 3.8 Let G be an abelian group. Then for any element g G,
the conjugacy class of g is { g}.
h g h1 = h h1 g = e g = g.
While were on the subject of conjugation, its worth noting that the
conjugation operation can be used to construct an isomorphism from
a group to itself:
76 a course in abstract algebra
f g ( g1 h g) = g g1 h g g1 = e h e = h.
f g ( h k ) = g h k g 1 = g h e k g 1 =
g h g 1 g k g 1 = f g ( h ) f g ( k )
as required.
An isomorphism from a group to itself is called an automorphism,
and automorphisms of the form discussed in Proposition 3.9 are called
inner automorphisms.
We are now in a position to study precisely what conditions a subgroup
H 6 G must satisfy in order to ensure that each of its left cosets is
equal to the corresponding right coset.
Looking again at the dihedral group D3 we recall that the rotation
subgroup R3 = {e, r, r2 } has this property gR3 = R3 g for any g D3 .
The subgroups {e, m1 }, {e, m2 } and {e, m3 } dont, however, although
the trivial subgroup {e} and the full group D3 do.
The clue lies in the observation that D3 = {e} {r, r2 } {m1 , m2 , m3 },
R3 = {e} {r, r2 } and the trivial subgroup {e} are all unions of
conjugacy classes of elements of D3 , while the reflection subgroups
{e, m1 }, {e, m2 } and {e, m3 } arent.
Why does this matter? Well, if a subgroup H < G is a union of
conjugacy classes (which not every subgroup need be) then it ensures
that for any element h H, all elements of the form g h g1 are
also in H. That is, H is closed under conjugation by elements of G.
The following proposition confirms that were on the right track:
Proposition 3.10 A subgroup H < G is a union of conjugacy classes if
and only if Hg = gH for all g G.
This proposition provides the answer we were looking for: the sub-
groups whose corresponding left and right cosets gH and Hg coincide,
are exactly those subgroups which happen to be unions of conjugacy
classes. In the case of an abelian group, we saw that the conjugacy
classes are just sets consisting of individual elements. So any subgroup
of an abelian group is therefore a union of conjugacy classes, and by
Proposition 3.10 its left and right cosets must be the same. (In addition
to the more obvious reason that all the elements commute.)
Equivalently, the left and right cosets of H in G coincide if and only if
g h g1 H for all g G and h H. We give a subgroup of this
type a special name:
Definition 3.11 A subgroup H < G is said to be a normal subgroup
of G or normal in G if its left and right cosets coincide, in the sense
that gH = Hg for all g G.
Equivalently, by Proposition 3.10, H is normal in G if it is a union of
conjugacy classes of elements of G. We denote normal subgroups by
H C G or H P G.
In the examples earlier on, we noticed that R3 C D3 and A4 C S4 . In
both of these cases, the subgroups have index 2 in the larger group:
| D3 : R3 | = |S4 : A4 | = 2. It is certainly not the case that a normal
subgroup must have index 2, since 3Z C Z and |Z : 3Z| = 3. The
converse, however, is true in general, as the following argument shows:
Proposition 3.12 Let H < G be an index2 subgroup of G. Then H is
normal in G.
Proof Suppose that | G : H | = 2. Then there are only two distinct right
cosets: H itself, and the complement G \ H. Similarly, there are only
two distinct right cosets: H and G \ H again. Clearly, each left coset is
then equal to the corresponding right coset, and hence H C G.
Example 3.15 SLn (R) C GLn (R) since every conjugate of a matrix
with determinant 1 also has determinant 1, by the multiplicative
property of matrix determinants. Suppose A SLn (R) (so det A = 1)
and that B GLn (R) with det B = d R. Then det( B1 ) = 1d and
so det( BAB1 ) = det( B) det( A) det( B1 ) = 1d 1 d = 1. Hence
SLn (R) C GLn (R).
By a similar argument, SOn (R) C On (R) and SUn C Un .
The centre of a group isnt just a subset, its actually a normal subgroup,
as the following proposition shows. (Well make use of this fact later
on when we study quotient groups.)
Proposition 3.17 The centre Z ( G ) of a group G is a normal subgroup.
Z/nZ
= Zn and D3 /R3
= Z2 .
To decide whether this construction works for any subgroup H of
any group G, well try another example. Let M denote the subgroup
{e, m1 } of D3 . This has three left cosets
M = {e, m1 } rM = {r, m3 } r 2 M = {r 2 , m2 }
from which we can see immediately that M is, unlike R3 , not a normal
subgroup of D3 . In particular, this means that whereas with the
previous examples we could cheerfully ignore whether the resulting
groups consisted of left or right cosets, we have to be more careful
in this case. For the moment, lets consider the set of left cosets of
M in D3 and see how they interact. Following on from the previous
examples, given two elements g, h D3 , we require that the product of
any element of the left coset gM with any element of the left coset hM
be an element of the left coset ( g h) M. Unfortunately, this doesnt
always work. For example,
Whats going wrong here? What precisely is stopping the left cosets
of M in D3 from behaving like elements of a group?
We want to be able to define a group structure on the set of left cosets
of M such that
( gM)(hM) = ( g h) M
for all g, h D3 . But the problem is that the set { M, rM, r2 M } isnt
closed under this multiplication operation, because the multiplication
operation in D3 doesnt line up neatly with the left cosets of M in D3 .
It works in some cases. For example, (rM)(eM) = {r, m3 } = rM as
required. But if we try (eM)(rM) we find that it doesnt. In particular,
m1 r = ( e m1 ) (r e )
= e ( r r 1 ) m 1 r e
= (e r ) (r 1 m1 r ) e.
In the last line, the first parenthesised expression (e r ) is equal to
r, as we want it to be. But in order for this whole expresion to be
an element of rM, we also need the second parenthesised expression
(r 1 m1 r ) to be an element of M, which it isnt since
r 1 m 1 r = r 2 m 1 r = m 3 .
Notice that (r 1 m1 r ) is a conjugate of an element of M. In fact,
if we tried any other combination of cosets of M (rM and r2 M, for
example) wed end up with a similar requirement: that a particular
conjugate of an element of M be itself an element of M.
In other words, the reason this particular example doesnt work is
because M isnt closed under conjugation. In the last section, we intro-
duced a special class of subgroups which are closed under conjugation,
namely normal subgroups. So the reason we cant form a group from
the left cosets of M is because M isnt a normal subgroup of D3 . The
rotation subgroup is normal in D3 , and the subgroup nZ is normal in
Z, which is why this construction worked properly in those cases.
The following proposition states this more generally:
Proposition 3.24 If H is a normal subgroup of G, the set of all left cosets
of H in G forms a group under the multiplication operation
( gH )(kH ) = ( g k) H
for all g, k G.
that ( G H )/({e} H )
= G? Yes, as the next proposition confirms. Table 3.9: Z6 /h3i
= Z3
Proposition 3.29 The direct product G H of any two groups G and H
has a normal subgroup isomorphic to H. Factoring out by this subgroup
yields a group isomorphic to G.
obtaining a new shape (see Figure 3.2). This new shape, similar to a
tif
y
ble 3.10) or by observing that the cosets rZ, m1 Z and m2 Z all have Table 3.10: Coset multiplication table
for D4 /Z ( D4 )
= V4
order 2 under the coset multiplication operation.
Well come back to this later in Section 7.2 when we use the quotient
G/Z ( G ) to form the upper central series of a group, but for the
moment just think of it as giving an indication of the nonabelianness
of a group.
The commutator subgroup [ G, G ] of a group G is also a measure
(albeit a slightly different one) of the nonabelianness of G: if [ G, G ] is
trivial then G is abelian, and the more elements [ G, G ] has, the more
nonabelian G is. Proposition 3.21 confirms that [ G, G ] is normal in
G, so we can form the quotient group G/[ G, G ]. In doing this, we
are killing off all of the noncommutativity in G. More precisely, we
are making every commutator gh g1 h1 equal to the identity e,
forcing every element of G to commute with every other element of G.
Proposition 3.31 For any group G, the quotient G/[ G, G ] is abelian.
Furthermore, if N P G and G/N is abelian, then [ G, G ] 6 N.
[ aG 0 , bG 0 ] = ( aG 0 )(bG 0 )( aG 0 )1 (bG 0 )1 =
( aba1 b1 ) G 0 = [ a, b] G 0 = G 0 .
(The last step follows because [ a, b] G 0 .) This means that the com-
mutator of [ aG 0 , bG 0 ] of any two cosets in G/G 0 is trivial, and hence
G/G 0 is abelian.
90 a course in abstract algebra
Geometrically, we can regard C /R+ as the group of rotations in
the plane R2 . That is, the matrix group
cos sin
: [0, 2 ) .
sin cos
This is exactly the special orthogonal group SO2 (R). So C
/R =
+i
SO2 (R). This is also isomorphic to the unitary group U1 = e :
[0, 2 ) .
Alternatively, we can view the cosets as infinite radii in the complex
plane; when we factor by R + , we collapse each of these radii onto its
intersection point with the unit circle.
In fact, there are relatively few types of simple groups. The only other
ones weve met so far, apart from the prime order cyclic groups Z p
are the alternating groups An for n 6= 4. The groups A1 and A2 are
trivial (and hence simple). The group A3 is isomorphic to Z3 and is
hence simple too. The group A4 , however, is not simple: it has a single
proper, nontrivial normal subgroup
Proof (Omitted.) 3
The proof was declared complete in
The proof of this classification theorem took about a hundred mathe- 1983, but a subtle gap (concerning qu-
asithin groups) was discovered, which
maticians approximately five decades to complete, between 1955 and took a further thousand or so pages to
2004,3 and the details are spread through somewhere in excess of ten patch.
96 a course in abstract algebra
4
It has been suggested that no sin- thousand pages of journal articles and research monographs.4
gle living person fully understands all
the details of the proof: the American
A full discussion of the third and fourth classes is beyond the scope of
group theorist Daniel Gorenstein (1923 this book, but we can look at some examples of finite groups of Lie
1992), one of the main driving forces type now. Table 3.11 lists the nonabelian simple groups with less than
behind the classification, is generally
regarded as having had the best overall 10 000 elements.
perspective on the project. Finite groups of Lie type are related to the matrix groups we met in
Group Order Section 1.2, but we replace R and C with finite analogues. Well study
A5
= A1 (4) = PSL2 (4) these objects, finite fields, in more detail in Chapters 811, but the
= A1 (5) = PSL2 (5) 60
following (incomplete) definition will do for the moment.
A1 (7) = PSL2 (7)
= A2 (2) = PSL3 (2) 168 Definition 3.45 Let p be prime, and let F p = {0, . . . , p1} be the
A6 = A1 (9) = PSL2 (9) 360 set consisting of the first p non-negative integers, equipped with two
A1 (8) = PSL2 (8) 504 binary operations + and representing, respectively, addition and
A1 (11) = PSL2 (11) 660 multiplication modulo p. Then F p is the finite field of order p.
A1 (13) = PSL2 (13) 1092
A1 (17) = PSL2 (17) 2448
More generally, given any prime power q = pn for some n N, we
A7 2520 can define the finite field Fq , but the general construction is more
A1 (19) = PSL2 (19) 3420 complicated, and well postpone it until Section 11.3.
A1 (16) = PSL2 (16) 4080 Here we run into a slight difficulty with notation. The finite groups of
A2 (3) = PSL3 (3) 5616 Lie type classify neatly into a number of infinite families, many having
2 A (33 ) = PSU (3)
2 3 6048 more than one name. One naming convention, used for some of these
A1 (23) = PSL2 (23) 6072 groups, intentionally echoes the names used for the corresponding
A1 (25) = PSL2 (25) 7800 infinite matrix groups. Another widely-used convention (used also for
M11 7920 the Coxeter groups well meet in Section 5.C) uses the capital letters
A1 (27) = PSL2 (27) 9828 AG with numeric subscripts and superscripts; if were not careful,
this could cause confusion with the alternating groups An and the
Table 3.11: Nonabelian simple groups
of order less than 10 000
dihedral groups Dn . To resolve this, well use a slightly different
(sans-serif) font in this context so that An is the alternating group on
n objects, An (q) is the finite simple group described next, and An is
the finite Coxeter group discussed later (and which, confusingly, is
isomorphic to the symmetric group Sn+1 ).
In the following examples, unless otherwise stated, q will be a positive
power of a prime integer, to ensure that Fq is indeed a finite field.
Example 3.46 For n > 1, let An (q) or PSLn+1 (q) denote the projec-
tive special linear group over Fq . That is, the special linear group
SLn+1 (Fq ) of (n+1)(n+1) square matrices, with entries from the
orderq finite field Fq and determinant 1 (calculated in Fq ), factored
by its centre as described in Example 3.38.
In general,
n
|An (q)| = |PSLn+1 (q)| = 1
gcd(n+1,q1)
qn(n+1)/2 ( q i +1 1 ).
i =1
normal subgroups 97
Group Order
An (q) = PSLn+1 (q) 1
gcd(n+1,q1)
qn(n+1)/2 in=1 (qi+1 1)
2
Bn (q) = O2n+1 (q) n>1 1
gcd(2,q1)
qn in=1 (q2i 1)
2
Cn (q) = PSp2n (q) n>2 1
gcd(2,q1)
qn in=1 (q2i 1)
+
Dn (q) = O2n (q) n>3 1
gcd(4,qn 1)
qn(n1) (qn 1) in=11 (q2i 1)
1
E6 ( q ) gcd(3,q1)
q36 (q12 1)(q9 1)(q8 1)(q6 1)(q5 1)(q2 1)
1
E7 ( q ) gcd(2,q1)
q63 (q18 1)(q14 1)(q12 1)(q10 1)(q8 1)(q6 1)(q2 1)
G2 ( q ) q6 (q6 1)(q2 1)
2A
n (q
2) = PSUn+1 (q) n>1 1
gcd(n+1,q+1)
qn(n+1)/2 in=1 (qi+1 (1)i+1 )
2D
n (q
2) = O2n (q) n>3 1
gcd(4,qn +1)
q n ( n 1) ( q n + 1) in=11 (q2i 1)
2E 2) 1
6 (q gcd(3,q+1)
q36 (q12 1)(q9 + 1)(q8 1)(q6 1)(q5 + 1)(q2 1)
2F 0 17971200 = 211 33 52 13
4 (2)
in Table 3.12, all of which are simple except for the following:
A1 (2) = PSL2 (2), A1 (3) = PSL2 (3), B2 (2) = O5 (2),
2 2 2
G2 (2), A2 (2 ) = PSU3 (2), B2 (2),
2 2
F4 ( 2 ) , G2 (3).
The Chevalley groups An (q), Bn (q), Cn (q), Dn (q), E6 (q), E7 (q), F4 (q)
and G2 (q) are named after the French mathematician Claude Chevalley
(19091984); the groups 2 An (q2 ), 2 Dn (q2 ), 2 E6 (q2 ) and 3 D4 (q3 ) are
called Steinberg groups after the Moldovan/American mathematician
Robert Steinberg; the groups 2 B2 (22n+1 ) are called Suzuki groups
after the Japanese mathematician Michio Suzuki (19261998); the
groups 2 F4 (22n+1 ) and 2 G2 (32n+1 ) are called Ree groups after the
normal subgroups 99
Group Order
M11 7 920
M12 95 040
M22 443 520
M23 10 200 960
M24 244 823 040
J1 175 560
J2 604 800
J3 50 232 960
J4 86 775 571 046 077 562 880
Co1 4 157 776 806 543 360 000
Co2 42 305 421 312 000
Co3 495 766 656 000
Fi22 64 561 751 654 400
Fi23 4 089 470 473 293 004 800
0
Fi24 1 255 205 709 190 661 721 292 800
McL 898 128 000
HS 44 352 000
Suz 448 345 497 600
He 4 030 387 200
Ru 145 926 144 000
HN 273 030 912 000 000
O0 N 460 815 505 920
Ly 51 765 179 004 000 000
Th 90 745 943 887 872 000
B 4 154 781 481 226 426 191 177 580 544 000 000
M 808 017 424 794 512 875 886 459 904 961 710 757 005 754 368 000 000 000
Table 3.13: Sporadic simple groups
systems. The groups M23 and M24 also happen to be the symmetry
groups of the 23 and 24bit binary error-correcting codes discovered
by the Swiss mathematician and information-theorist Marcel Golay
(19021989); these codes were used by NASA to reliably transmit im-
age data back to earth from the Voyager 1 and 2 space probes during
their 19791981 flyby missions to Jupiter and Saturn.
The existence of the four Janko groups J1 , J2 , J3 and J4 was predicted
by the Croatian mathematician Zvonimir Janko; he constructed J1 in
1965 and the other three were explicitly constructed (and their exis-
tence confirmed) over the next fifteen years by other mathematicians.
The three Conway groups Co1 , Co2 and Co3 were discovered by the
British mathematician John Conway. They arise as subgroups of the
symmetry group Co0 of a 24-dimensional object called the Leech
lattice 24 ; Conway constructed this group during a single twelve-
hour period one Saturday in 1964.
The group Co1 is the quotient of Co0 by its two-element centre, while
the groups Co2 and Co3 are the subgroups which leave unchanged
certain types of lattice vectors in 24 . Four other sporadic groups
exist as subgroups or quotients of Co1 : the McLaughlin group McL,
the HigmanSims group HS, the Suzuki group Suz and the second
Janko group J2 (sometimes called the HallJanko group).
0
The three Fischer groups Fi22 , Fi23 and Fi24 all arise as groups of 3
transpositions, particular types of order2 permutations where the
product of any two has order at most 3. They were discovered in 1971
by the German mathematician Bernd Fischer. The group Fi24 is not
0
itself simple, but its commutator subgroup Fi24 is.
The remaining eleven sporadic groups are the Held group He, the
Rudvalis group Ru, the ONan group O0 N, the HaradaNorton group
HN, the Lyons group Ly, the Thompson group Th, the Baby Monster
group B and the Monster group M.
The last of these contains nineteen of the other twenty-five sporadic
groups as subgroups; these nineteen are sometimes called the happy
family, while the remaining six (J1 , J3 , J4 , Ly, O0 N and Ru) are some-
times referred to as pariahs.
normal subgroups 101
Summary
20
Definition 3.16, page 78. of all elements of G which commute with all other elements of G;20
21
Proposition 3.17, page 78. this is not only a subgroup of G but happens to be normal as well.21
Amongst other things, it gives us a measure of how abelian a particular
group is: Z ( G ) = G exactly when G is abelian.
Similarly, we define the commutator subgroup or derived group
[ G, G ] or G 0 to be the subgroup generated by all elements (called
22
Definition 3.18, page 79. commutators) of the form [ g, h] = gh g1 h1 .22 , 23 Again, this is a
23
Definition 3.19, page 80. normal subgroup of G24 and gives an indication of how abelian G is:
24
Proposition 3.21, page 81. [ G, G ] is trivial if and only if G is abelian, and the larger [ G, G ] is, the
less abelian G is. In particular, for n > 3, the commutator subgroup
25
Proposition 3.20, page 80. [Sn , Sn ] is the alternating group An .25
The reason we care about normal subgroups isnt just because their
left and right cosets line up neatly, its because their cosets interact
26
Proposition 3.24, page 85. in a well-defined way and form a group in their own right.26 A
key example of this construction is the formation of the finite cyclic
group Zn from the cosets of nZ = hni in Z. We say that Zn is the
quotient group Z/nZ. More generally, the (left or right) cosets of a
normal subgroup H P G form a quotient group G/H in which the
27
Definition 3.25, page 86. product ( gH )(kH ) is defined to be the coset ( gk) H.27 We also saw
28
Example 3.28, page 87. that D3 /R3
= Z2 and Z6 /h3i = Z3 ,28 amongst others.
Factoring a group G by its centre Z ( G ) often yields useful information
about the nonabelianness of a group: the smaller the centre, the
larger the quotient G/Z ( G ); if G is abelian then Z ( G ) = G and hence
G/Z ( G ) is trivial. On the other hand, D3 /Z ( D3 ) = D3 , so the dihedral
group D3 is in some sense about as nonabelian as we might reasonably
29
Example 3.30, page 89. expect it to be.29 Factoring the general linear group GLn (R) by its
centre yields the projective general linear group PGLn (R); analogous
projective versions of On (R), SOn (R), Un , SUn and Sp2n (R) exist
30
Example 3.38, page 92. too.30
The commutator subgroup is another measure of the nonabelianness
of a given group: if its trivial then the group is abelian. Factoring out
by this subgroup (which, as with the centre, we can do because its
a normal subgroup) effectively kills all the noncommutativity in the
group, leaving an abelian group, which well denote Gab . This process
31
Example 3.32, page 90. is called abelianisation.31
32
Definition 3.39, page 93. Certain groups have no normal subgroups apart from themselves
33
Example 3.40, page 93. and the trivial subgroup. Another way of looking at this is that they
34
Proposition 3.43, page 94. have no proper and nontrivial quotients. We call groups of this type
35
Example 3.46, page 96. simple,32 and they fall into four different classes: cyclic groups Z p
36
Example 3.47, page 97. where p is prime,33 alternating groups An for n > 4,34 finite groups
37
Table 3.12, page 98. of Lie type,35 , 36 , 37 and twenty-six sporadic groups.38
38
Table 3.13, page 99.
normal subgroups 103
Exercises
Show that [ D2n+1 , D2n+1 ] = R2n+1
= Z2n+1 , and [ D2n , D2n ]
= Zn .
You have shown me a strange image,
and they are strange prisoners.
Like ourselves, I replied; and they
see only their own shadows, or the
shadows of one another, which the
fire throws on the opposite wall of the
cave?
True, he said; how could they see
anything but the shadows if they were
4 Homomorphisms never allowed to move their heads?
And of the objects which are being car-
ried in like manner they would only
see the shadows?
Yes, he said.
And if they were able to converse with
o far, weve met many different types of groups: cyclic, dihedral one another, would they not suppose
Example 4.4 The function f : GLn (R) GLn (R) where A 7 det1 A A
is a homomorphism, since
f ( AB) = det1AB AB = det A1det B AB = det1 A A det1 B B = f ( A) f ( B)
Example 4.9 illustrates another important case, which the next example
generalises.
Example 4.14 For any subgroup H < G there is a unique inclusion
homomorphism i : H G which maps every element of H to itself
in G. (Sometimes we use a hooked arrow for inclusions instead:
i : G , H.) This homomorphism is necessarily injective.
For example:
i1 : Z2 Z2 Z3 i2 : Z3 Z2 Z3 m2 m1
0 7 (0, 0) 0 7 (0, 0)
1 7 (1, 0) 1 7 (0, 1)
m3
2 7 (0, 2)
The homomorphism f ab is called the induced homomorphism, and We can represent the induced homo-
is best illustrated by means of another example. morphism f ab by means of the follow-
ing square of groups and homomor-
Example 4.23 We saw in Example 3.32 that D3ab = Z2 , and it so phisms:
happens that D6ab Z
= 2 Z2 . Example 4.17 gave us an inclusion
G
f
/H
homomorphism i : D3 D6 , so putting all these together should
yield an induced homomorphism iab : Z2 Z2 Z2 . G H
0 / (0, 0) 0 / (0, 0) 1 / (1, 0)
From this, we can see that the induced homomorphism iab is the
inclusion homomorphism i1 : Z2 Z2 Z2 mapping D3ab = Z2 into
the first coordinate of D6ab = Z2 Z2 :
iab : Z2 Z2 Z2
0 7 (0, 0)
1 7 (1, 0)
When you want to see if your picture 4.2 Kernels and images
corresponds throughout with the ob-
jects you have drawn from nature, take
a mirror and look in that at the reflec- Recall that for any function f : X Y we can define the image
tion of the real things, and compare im( f ) of f . This is a subset of the codomain Y and is defined to consist
the reflected image with your picture,
and consider whether the subject of the of those elements of Y that are mapped to by at least one element of
two images duly corresponds in both, the domain X. That is,
particularly studying the mirror.
Leonardo da Vinci (14521519), im( f ) = {y Y : y = f ( x ) for some x X }.
The Practice of Painting
Its natural to ask whether the image of a homomorphism f : G H
has any interesting properties. We know that from Proposition 4.24
that im( f ) must contain the identity e H as well as a full complement
of inverses; that is, for any h im( f ) we know that h1 im( f )
too. We now find ourselves moving inexorably towards the following
proposition.
Proposition 4.25 Let f : G H be a group homomorphism. Then the
image im( f ) is a subgroup of H.
Proof We know from Proposition 4.24 that e H im( f ) and also that
h1 im( f ) for any h im( f ). All we need to do now is show that
im( f ) is closed under the group multiplication operation in H.
Suppose, therefore, that h1 = f ( g1 ) and h2 = f ( g2 ) for some elements
g1 , g2 G. Then
h1 h2 = f ( g1 ) f ( g2 ) = f ( g1 g2 ) im( f ).
homomorphisms 113
1
We are tacitly assuming that (4.2) actu-
ally has a solution; in this case it does.
terise the vectors in R4 which get mapped to the vector
h5i
Homogeneous systems are always solv- 2
able; the original inhomogeneous sys- 4
tem (4.2) is solvable if the vector in R3 by the linear map represented by the matrix
h5i
2 1 3 4 2
4 A= 0 2 5 1 .
0 1 3 0
belongs to the image of A, which it
does. Suppose that v is a solution of (4.2).1 Then any other solution of (4.2)
can be expressed as a sum u + v where u is a solution of the associated
homogeneous system
x h i
1 3 4 2 y 0
0 2 5 1 z = 0 (4.3)
0 1 3 0 w 0
The solutions of the homogeneous sytem form a subspace of the vector
space R4 which we call the null space or kernel of (the linear map
represented by) the matrix A. In this case, the general solution of (4.2)
takes the form " #
5 17
4 +t 1 3
0
6 11
where t R is a real parameter. Here we can see that the kernel of A
is a one-dimensional subspace of R4 .
So, in linear algebra we can derive some useful information about
a linear map (and thereby, perhaps, solve the actual problem under
investigation) by characterising which elements of the domain get
mapped to the zero vector in the codomain; that is, by studying the
kernel of the map.
We now define the analogous concept for group homomorphisms:
Definition 4.28 Let f : G H be a group homomorphism. Then
the kernel ker( f ) of f is defined to be
ker( f ) = { g G : f ( g) = e H }.
That is, the elements of G that are mapped by f to the identity e H in
H.
Looking again at the homomorphism f : D3 D3 from Example 4.26,
we can see that the kernel of f is the set {e, r, r2 }, which happens to
be the rotation subgroup R3 of D3 .
Just as the image of a homomorphism is a subgroup of the codomain,
and the kernel and image of a linear map are vector subspaces of,
respectively, the domain and codomain, it seems reasonable to ask
whether the kernel of a group homomorphism is necessarily a sub-
group of the domain. But the kernel of the homomorphism in Exam-
ple 4.26 isnt just a subgroup of the domain, its a normal subgroup.
Might it be that the kernel is always a normal subgroup of the domain?
The answer to this question is yes, as the following proposition shows.
homomorphisms 115
elements of H.
Definition 4.36 Let f : G H be a group homomorphism, and let
h be some element of H. We define the inverse image or preimage
of h to be the set
f 1 ( h ) = { g G : f ( g ) = h } G
of elements of G which f maps to the element h.
If K is a subset of H, then the inverse image or preimage of K is the
set
f 1 ( K ) = { g G : f ( g ) K } G
of elements of G which f maps to some element of K.
In this case, merely knowing that such a homomorphism exists, with- 12 Z3 oZ8
13 Z3 o D4 with kernel Z2 Z2
out knowing anything else about it, was enough to help us figure out
14 Dic6
what the group was. This isnt always possible, but we can usually
15 Dic3 Z2
derive at least some information about the chosen groups internal
structure, as the following example demonstrates. Table 4.1: The fifteen groups of order 24
What this means is that the set gSg1 consists of all possible conjugates
of elements of S by some fixed element g G. So the normaliser NG (S)
is the set of all elements g G for which conjugation by g either leaves
S unchanged or permutes its elements among themselves.
Example 4.47 Consider the subset S = {1, 1} Z. For some
arbitrary integer n Z, the conjugate set is
n + S n = {n 1 n} {n + 1 n} = {1, 1} = S.
In other words, the set S is fixed by conjugation by any integer n Z.
What this means is that NZ (S) = Z: the normaliser of S in Z is the
whole of Z.
Similarly, considering the subgroup 3Z 6 Z, we see that
n + 3Z n = {n + 3m n : m Z} = {3m : m Z} = 3Z
and hence
NZ (3Z) = {n Z : n + 3Z n = 3Z} = {n Z : 3Z = 3Z} = Z.
Thus NZ (3Z) = Z.
More generally, NG (S) = G for any abelian group G and any subset
S G. At first sight the normaliser doesnt look particularly useful:
in the examples above its just equal to the full group. But as usual,
things become more interesting when we consider nonabelian groups.
Example 4.48 Let G = D3 and consider the four subsets
S1 = { m 1 , m 2 , m 3 } , S2 = {e, r, r2 }, S3 = {e, m1 }, S4 = { r } .
Clearly the identity e is in ND3 (S) for any of these subsets, since
e s e1 = s for any s S.
The normaliser ND3 (S1 ) contains the rotations r and r2 since
r m1 r 2 = m2 , r 2 m1 r = m3 ,
r m2 r 2 = m3 , r 2 m2 r = m1 ,
r m3 r 2 = m1 , r 2 m3 r = m2 .
Observe that not only is g s g1 in S1 for each of these elements
g = r, r2 , but that every element of S1 is represented in this way. That
is, gS1 g1 is actually equal to S1 rather than just being a subset of it.
homomorphisms 123
Sg = {s g : s S} = { g s : s S} = gS
and therefore
g1 Sg = { g1 s g : s S} = S
so g1 NG (S) as required. Thus NG (S) is a subgroup of G.
In Example 4.48 we considered two cases where the subset S happened
to be a subgroup of the larger group D3 . In the first of these cases,
the subset S was actually the rotation subgroup R3 = {e, r, r2 }, which
124 a course in abstract algebra
V4
H1 is not normal in S4 , but it is normal in NS4 ( H1 ) =
H2 is not normal in S4 , but it is normal in NS4 ( H2 )
= S3
H3 is normal in both S4 and its normaliser NS4 ( H3 ) = S4
H2 is not normal in A4 , but it is normal in NA4 ( H2 ) = H2
= A3
H3 is normal in both A4 and its normaliser NA4 ( H3 ) = A4
Proposition 4.55 Let G be a group, and let Aut( G ) be the set of all
automorphisms f : G G. Then Aut( G ) forms a group under the usual
composition operation for homomorphisms.
We now have the ingredients we need for the following elegant result,
an application of the First Isomorphism Theorem that describes the
connection between normalisers, centralisers and automorphisms.
Theorem 4.57 (The NormaliserCentraliser Theorem) Let H be a
subgroup of a group G. The quotient NG ( H )/ZG ( H ) is isomorphic to a
subgroup of the automorphism group Aut( H ).
= ( g1 g2 ) h ( g21 g11 )
= g1 ( g2 h g21 ) g11
= g1 (g2 (h))
= (g1 g2 )(h)
and hence f ( g1 g2 ) = f ( g1 ) f ( g2 ) as required.
We want to apply the First Isomorphism Theorem to this homomor-
phism f , so we need to figure out what its kernel is. The kernel
ker( f ) will consist of all elements of NG ( H ) which map to the iden-
tity element in Aut( H ). This identity element is exactly the identity
homomorphism id : H H. So were looking for the elements of
NG ( H ) that map to some inner automorphism g that happens to be
the identity.
In other words, we want the elements g NG ( H ) for which g (h) = h
for all elements h H. But these are exactly those g NG ( H ) for
which g h g1 = h and hence ker( f ) = ZG ( H ), the centraliser of H
in G.
By the First Isomorphism Theorem, then,
NG ( H )/ ker( f ) = NG ( H )/ZG ( H )
= im( f ) 6 Aut( H )
as required.
We can now use this result to tell us what it means if we take the quo-
tient of a group G by its centre Z ( G ): we get the inner automorphism
group of G.
Corollary 4.58 Let G be a group. Then G/Z ( G ) = Inn( G ).
Were now ready to state and prove the Second Isomorphism Theorem:
Theorem 4.62 (Second Isomorphism Theorem) Let G be a group,
let H be a (not necessarily normal) subgroup of G, and let N be a normal
subgroup of G. Then
HN/N
= H/( H N ).
for some k H.
With this in mind, we define a function f : HN H/( H N ) mapping
the element h n HN to the coset h( H N ) in H/( H N ).
We have to show four things: that f is well-defined, that its surjective
onto H/( H N ), that its a group homomorphism, and that its kernel
is the normal subgroup N.
Dealing with the first of these, let h1 , h2 H and n1 , n2 N such that
h1 n1 = h2 n2
in HN. Rearranging this, we get
h21 h1 = n2 n11 .
On the left-hand side, h21 h1 is clearly an element of H, while on
the right-hand side, n2 n11 is clearly an element of N. Since theyre
equal, it must be the case that both belong to H and N. So both
sides, and specifically h21 h1 , are in the intersection H N. Hence
by Proposition 2.26, since h21 h1 H N, the cosets h1 ( H N ) and
h2 ( H N ) are equal, and so f (h1 n1 ) = h1 ( H N ) = h2 ( H N ) =
f (h2 n2 ). Thus f is well-defined.
To show that f is a group homomorphism, we need the following
observation: given any h H and n N there exists some element
k N such that h n = k h. This is because N, being a normal
subgroup, is closed under conjugation and so there is some element
k N such that h n h1 = k. Rearranging this, we get the desired
fact. Then, for any h1 , h2 H and n1 , n2 N we have
f ((h1 n1 )(h2 n2 )) = f (h1 (n1 h2 )n2 )
= f ( h1 h2 n3 n2 ) (for n3 = h21 n1 h2 )
= (h1 h2 )( H N )
= h1 ( H N ) h2 ( H N )
= f ( h1 n1 ) f ( h2 n2 )
and hence f is a homomorphism.
To see that f is surjective, consider any h H and form the coset
h( H N ). We need to find some element h n of HN which maps to
this coset. Setting n = e N to get the product h e HN suffices
admirably, since f (h e) = h( H N ) as required.
Finally, we need to show that ker( f ) = N. This kernel consists of all
elements h n HN such that f (h n) = H N. From this, we see that
the only elements h H for which h n 7 H N are those satisfying
h( H N ) = H N, and Proposition 2.26 tells us that these are exactly
the elements of H which also belong to N. So ker( f ) = ( H N ) N. But
this is exactly the same as N, so ker( f ) = N.
130 a course in abstract algebra
Summary
10
Definition 4.1, page 106. In this chapter we studied homomorphisms,10 functions f : G H
from one group to another that preserve some aspects of the group
structure, in the sense that f ( g1 g2 ) = f ( g1 ) f ( g2 ) for any ele-
ments g1 , g2 G. Weve already met the concept of a bijective ho-
11
Definition 1.17, page 11. momorphism, an isomorphism,11 and in this chapter we also saw
several examples of injective homomorphisms or monomorphisms
(sometimes denoted with a tailed arrow f : G H) and surjective
homomorphisms or epimorphisms (sometimes denoted with a two-
headed arrow f : G H). In particular, the identity isomorphism
idG : G G is always defined for any group G, but more generally
an isomorphism from a group to itself is called an automorphism,
and a non-bijective homomorphism from a group to itself is called
an endomorphism. For any groups G and H there is a unique zero
homomorphism z : G H which maps every element of G to the
identity element e H in H.
Reassuringly, composing two homomorphisms yields another homo-
12
Proposition 4.11, page 107. morphism.12
Important examples of monomorphisms are given by inclusion ho-
momorphisms i : H , G mapping a subgroup H 6 G into a larger
13
Example 4.14, page 108. group.13 , 14 Important examples of epimorphisms include projection
14
Example 4.16, page 108. homomorphisms.15 Examples of both can often be constructed by
15
Example 4.15, page 108. considering the particular context of the groups in question: in the
case of dihedral or other isometry groups, we can sometimes use our
16
Example 4.17, page 109. geometric intuition to define homomorphisms.16
Another class of homomorphisms derive from the quotient groups
we learned about in the last chapter: given any normal subgroup
H of a group G we can define a canonical, surjective quotient ho-
17
Example 4.18, page 109. momorphism q : G G/H.17 In particular, factoring a group G by
its commutator subgroup [ G, G ] yields the abelianisation homomor-
18
Example 4.21, page 110. phism : G Gab = G/[ G, G ].18 These quotient homomorphisms,
in particular the abelianisation homomorphism, have an interesting
and useful property: given a homomorphism f : G H there is a
19
Proposition 4.22, page 110. well-defined induced homomorphism f ab : Gab H ab .19 , 20
20
Example 4.23, page 111. Homomorphisms map identity elements to identity elements, and
inverses to inverses, in the sense that if f : G H is a homomorphism,
21
Proposition 4.24, page 112. then f (eG ) = e H and f ( g1 ) = f ( g)1 for any g G.21
Since homomorphisms are functions, we can discuss their images;
we denote the image of a homomorphism f : G H by im( f ). It
turns out that the image of a homomorphism is a subgroup of its
22
Proposition 4.25, page 112. codomain,22 but in general not always a normal subgroup.23 If a
23
Example 4.26, page 113.
homomorphisms 133
40
Proposition 4.49, page 123. The normaliser NG (S) is always a subgroup of G40 , and in the case
where the subset S is a subgroup of G it so happens that S is a nor-
mal subgroup of NG (S); in fact, we can regard NG (S) as the largest
41
Exercise 4.1, page 134. subgroup of G in which S is normal.41 The centraliser ZG (S) is a gen-
42
Definition 3.16, page 78. eralisation of the concept of the centre Z ( G ) of a group G,42 and also
a slightly stricter object than the normaliser NG (S). It turns out that
ZG (S) is always a normal subgroup of NG (S), and the Normaliser
Centraliser Theorem tells us that if H is a subgroup of G, then the
quotient NG ( H )/ZG ( H ) is isomorphic to a subgroup of the automor-
phism group Aut( H ). Considering the special case where H = G we
find that the centraliser ZG ( G ) is just the centre Z ( G ), and the nor-
maliser NG ( G ) is G itself; this leads to the discovery that the quotient
G/Z ( G ) is isomorphic to the inner automorphism group Inn( G ).
43
Theorem 4.62, page 128. The Second Isomorphism Theorem43 can be proved by applying the
First Isomorphism Theorem to a suitable homomorphism. Suppose
that G is a group, H is a subgroup of G and N is a normal subgroup
of G. A short lemma confirms that the intersection H N is not just
44
Proposition 2.9, page 45. a subgroup of G,44 but also a normal subgroup of H.45 . The Second
45
Lemma 4.61, page 128. Isomorphism Theorem states that the quotient HN/N (where HN
denotes the subgroup consisting of all products of an element of H
with an element of N) is isomorphic to the quotient H/( H N ).
46
Theorem 4.65, page 130. The Third Isomorphism Theorem46 can also be proved by applying
the First Isomorphism Theorem to an appropriate homomorphism. It
states that if we have three groups G, H and K such that H and K are
normal subgroups of G, and K is also a normal subgroup of H, then
the quotient ( G/K )/( H/K ) is isomorphic to the quotient G/H.
Exercises
4.1 Show that if H is a subgroup of G then H is a normal subgroup of NG ( H ). Show also that
NG ( H ) is the largest subgroup of G in which H is normal.
4.2 Let G be a group. Show that the set Inn( G ) of inner automorphisms of G is a group. Show that
it is a subgroup of Aut( G ).
I am not yet so lost in lexicography, as
to forget that words are the daughters of
the earth, and that things are the sons of
heaven. Language is only the instru-
ment of science, and words are but the
signs of ideas: I wish, however, that
the instrument might be less apt to de-
cay, and that signs might be permanent,
like the things which they denote.
5 Presentations Samuel Johnson (17091784),
preface to A Dictionary of the English
Language (1755)
e w = w = w e.
If we let W be the set of all finite, reduced words in the two symbols
x and y, then the concatenationreduction operation is an associative
binary operation defined on this set, and the empty word e behaves
like an identity. What weve just shown is the following:
Proposition 5.1 Let W be the set of finite, reduced words in two symbols
x and y, and let : W W W be the concatenationreduction operation.
Then (W, ) is a monoid.
In fact, this is the most general possible monoid we can get from two
generating symbols. Monoids are all very well, and as remarked in
Chapter 1 they have numerous applications in parts of mathematics
and computer science, but at the moment were primarily interested in
groups. Is there some way we can modify or extend this construction
to get the most general possible group generated by two symbols x
and y? As it happens, yes there is, and the key is to think about whats
missing from a monoid that stops it being a group, namely inverses.
138 a course in abstract algebra
To see this, concatenate w and w1 either way round, reduce, and see
that you end up with the empty word e:
n n nk n1
w w1 = xi 1 xin2 . . . xi k xi . . . xi =e
1 2 k k 1
nk n1 n
w 1 w = x i . . . xi
1
= e xin11 xin22 . . . xi k
k k
W0 ( X ) = {w W ( X ) : l (w) = 0} = {e}
W1 ( X ) = {w W ( X ) : l (w) = 1} = X
W2 ( X ) = {w W ( X ) : l (w) = 2}
..
.
Wn ( X ) = {w W ( X ) : l (w) = n}
The other definition is more abstract, and its not immediately obvious
that its equivalent (or even that the objects it describes actually exist).
Nevertheless, there are certain technical advantages to this definition as
well, and some books present this as the main definition and relegate
the previous one to a concrete construction.
Definition 5.5 Let X be a set. The free group with basis X is the
group satisfying the property that for any other group G and any
function f : X G, there is a unique homomorphism f : F ( X )
G which extends f . That is, f( x ) = f ( x ) G for any x X.
Equivalently, the diagram
f
X G
i
f
F(X )
Let U ( G ) be the underlying set of G,4 without any group structure. 4 We can regard U as a device that turns
Then Definition 5.5 says that functions f : X U ( G ) are in bijective groups into sets, and we can view the
free group construction F as a machine
correspondence with homomorphisms f : F ( X ) G. that turns sets into groups. More than
If we denote by HomSet ( X, Y ) the set of all possible functions from that, U turns group homomorphisms
into functions, and F turns functions
the set X to the set Y, and by HomGroup ( G, H ) the set of all possible into group homomorphisms. So U and
homomorphisms from the group G to the group H, then we can write F translate neatly between the world of
sets and the world of groups.
this as
Maps of this type are called functors.
HomSet ( X, U ( G ))
= HomGroup ( F ( X ), G ).6 The free group construction F : Set
Group is a functor from the category of
Well make use of this fact a little later, but now we can show that sets to the category of groups. The un-
Definitions 5.4 and 5.5 are equivalent. derlying set map U : Group Set is a
functor from the category of groups to
Proposition 5.6 Let X = { x1 , x2 , . . .} be a (possibly infinite) set, and let the category of sets; it is an example of
F ( X ) be the free group of reduced words in the alphabet X . For any group something called a forgetful functor.
A full discussion of categories and func-
G and function f : X G there is a unique homomorphism f : F ( X ) G tors is beyond the scope of this book,
which extends f . but the branch of mathematics in which
they play a fundamental part, called
Conversely, let F be a group and let X F be some subset of F. If, for category theory (but sometimes flip-
any group G and function f : X G there exists a unique homomorphism pantly referred to as abstract non-
sense) is at the same time perhaps the
f : F G which extends f , then F is isomorphic to the free group F ( X )
most abstract subject in modern math-
of reduced words in the alphabet X . ematics and one of the most powerful:
it provides a suitably precise language
Proof Every element w of F ( X ) has a unique expression as a reduced and framework within which to dis-
cuss many different sub-branches of
word in the elements of the alphabet X : mathematics in generality.
n n
w = xi 1 xin2 . . . xi k Much of algebraic topology, for exam-
1 2 k ple, one of the highlights of twentieth-
where n1 , . . . , nk Z and xi1 , . . . , xik X are a collection of elements and twenty-first century mathematics,
is concerned with the study of functors
of X (some of which might be the same as each other). from categories of topological objects
to categories of algebraic objects such
Then, given a function f : X G, the only way we can extend this to
as groups and rings. Category theory
a homomorphism f : F ( X ) G is by mapping the word also has important applications in the-
n n oretical computer science.
w = xi 1 xin2 . . . xi k Interested readers are directed to the
1 2 k
classic work Categories for the Working
to the product Mathematician by Saunders Mac Lane
(19092005),5 one of the main pioneers
f ( x i1 ) n1 f ( x i2 ) n2 f ( x i k ) n k of the field. Many years after its origi-
nal publication in 1971, it remains one
in G, and mapping the empty word e in F ( X ) to the identity eG in G. of the most comprehensive and lucid
Nothing else satisfies the homomorphism condition. So, the required treatments of the subject.
homomorphism f exists and is unique. 5
S Mac Lane, Categories for the Working
Mathematician, second edition, Gradu-
To prove the converse, suppose that for any f : X G there exists a ate Texts in Mathematics 5, Springer
unique homomorphism f : F G extending f . Let : F ( X ) F be (1998)
the unique homomorphism which maps any reduced word in F ( X ) to 6
This is an example of something cat-
the element of F obtained by regarding this word as a product of the egory theorists call an adjunction or
adjoint relationship. The functor F is
elements of F. That is, suppose ( F, ) is the hypothesised group. Then said to be a left adjoint of U, and U is
n n
for any w = xi 1 xin2 . . . xi k in F ( X ) we set a right adjoint of F.
1 2 k
n n n n
( xi 1 xin2 . . . xi k ) = xi 1 xin2 xi k .
1 2 k 1 2 k
142 a course in abstract algebra
/ X2 i 2 / F ( X2 ) f 1
/ X1 i 1 / F ( X1 )
f
X1 X2
_ llll
ll5 _ ll5
l llll
ll lll
lll lll
i1 i2
lll lll
F ( X1 ) F ( X2 )
(Compare these diagrams with the one in Definition 5.5.)
presentations 143
For the relations of words are in pairs 5.2 Generators and relations
first.
For the relations of words are some-
times in oppositions. In this section we will develop the machinery of group presenta-
For the relations of words are accord- tions. Our first ingredient is the next proposition, which shows that
ing to their distances from the pair.
any group can be regarded as a quotient of a free group.
Christopher Smart (17221771),
Jubilate Agno (17591763) Proposition 5.13 Let G be a group, and let X be a set of generators for
Fragment B, part 4
G. Then G is isomorphic to a quotient of the free group F ( X ).
presentations 147
when its understood implicitly that the word or words to the right of
the colon are equal to the identity.
The group ZZ can be constructed by taking a suitable free group, in
this case F2 = h x, yi and introducing some appropriate relations that
force everything to commute. This is in some senses fairly straightfor-
ward, because all we need to do is to ensure that the generators x and
y commute, and since every other element of our desired group can
be regarded as a reduced word in these generators and their inverses,
all such elements will commute.
This can be effected by imposing the relation xy = yx, to get the
presentation
ZZ
= h x, y : xy = yx i.
Alternatively, replacing the relation xy = yx with the relator xyx 1 y1
(which is also the commutator [ x, y]) gives us
ZZ
= h x, y : xyx 1 y1 i.
The main drawback with this method is that although it does provide
a presentation for a given group, that presentation will usually not
be the simplest one. And for an infinite group, this method becomes
completely impractical for obvious reasons. There will almost always
be a simpler presentation for the groups were likely to meet. For
example, the above presentation for V4 has four generators and sixteen
relations, but we really only need two generators and three relations:
Example 5.20 The Klein group V4 is isomorphic to the direct sum
Z2 Z2 . We already know how to write down a presentation for
ZZ
= h x, y : xy = yx i,
and we can use this to make a presentation for Z2 Z2 . We do this
by introducing two more relations x2 = e and y2 = e, thus forcing
each generator to have order 2. This gives us the presentation
V4
= Z2 Z2
= h x, y : x2 = y2 = e, xy = yx i.
In general we will want a suitably compact presentation of a given
group, or at least one that can be written down in a compact way.
Example 5.21 Let G = Z3 Z2 . Then
G
= h x, y : x3 = y2 = e, xy = yx i.
Examining the presentations in these two examples, we see that each
group is a direct sum of finite cyclic groups, and each has a pre-
sentation comprising the combined generators and relations for the
cyclic group summands together with another relation that forces the
generator of one to commute with the generator of the other. That is,
h x : x2 = ei hy : y2 = ei = h x, y : x2 = e, y2 = e, xy = yx i,
h x : x3 = ei hy : y2 = ei = h x, y : x3 = e, y2 = e, xy = yx i.
Proof Suppose
K = h X Y : R S [ X, Y ]i.
Define a function f : X Y G H by f ( x ) = ( x, e H ) and f (y) =
(eG , y) for all x X and y Y.
Let w = a1 . . . ak R S [ X, Y ] be a relator. We need to check that
f (w) = (eG , e H ), and there are three cases to consider.
If w R then w F ( X ) and so f (w) = (w, e H ) = (eG , e H ) since
w = eG in G.
Similarly, if w S then w F (Y ) and so f (w) = (eG , w) = (eG , e H )
since w = e H in H.
If, on the other hand, w [ X, Y ] then w = xyx 1 y1 for some x X
and y Y. Hence
F ( X ) F (Y ) = h X Y : i = F ( X Y ) .
g = e g e = k 2 g k 2 = k ( k g k 1 ) k 1 = k g m k 1
2
= ( k g k 1 ) m = ( g m ) m = g m
2
and so gm 1 = e, and therefore p must divide m2 1 = (m+1)(m1).
Hence p must divide either m+1 or m1.
If p divides m1 then k g k1 = gm = g, which means that k g =
g k and so G is abelian. If p is odd then |k g| = 2p, which contradicts
our assumption that G has no elements of order 2p. The only other
possibility is that p = 2, in which case G = V4 = D2 = Z2 Z2 .
If, however, p divides m+1 then k g k1 = gm = g1 , which means
that k g = g1 k, and hence G
= D p by Proposition 5.28.
We can now write presentations for all groups with up to seven
elements:
0=h:i Z2
= h x : x2 = ei
h x : x3 = ei
Z3 = h x : x4 = ei
Z4 =
V4
= h x, y : x2 = y2 = e, xy = yx i Z5
= h x : x5 = ei
D3
= h x, y : x3 = y2 = e, yx = x 1 yi Z6
= h x : x6 = ei
Z7
= h x : x7 = ei
There are five groups of order 8 (see Proposition 2.40), and we can
presentations 157
The third follows from the fact that if |i j| > 1 then the transpositions
xi = (i i +1) and x j = ( j j+1) are disjoint, and therefore commute:
(i i +1)( j j+1) = ( j j+1)(i i +1).
We can thus define a surjective homomorphism f : Gn Sn by map-
ping each generator xi in Gn to the transposition (i i +1) in Sn . If we
can also show this homomorphism is injective, then weve finished.
To do this, we can use the fact that |Sn | = n! and show by induction
that | Gn | 6 n!, which will then imply that f must be injective.
When n = 1 we have the trivial group G1 = h : i, which has order
| G1 | = 1 6 1! as required. For n = 2 we have G2 = h x1 : x12 = ei
= Z2
and hence | G2 | = 2 6 2! as well.
Now suppose n > 3 and | Gn1 | 6 (n1)! for the induction hypothesis.
Let H 6 Gn be the subgroup generated by the first n2 generators
x1 , . . . , xn2 and define
y0 = e and y i = x n 1 x n 2 . . . x n i for 1 6 i 6 n1.
So
y0 = e, y1 = xn1 , y2 = xn1 xn2 , . . . , yn1 = xn1 xn2 . . . x1 .
Now let
A = { hyi : h H and 0 6 i 6 n1}.
This is certainly a subset of Gn , but we want to show that its equal to
the whole of Gn . To do this, we consider products of the form hyi x j ,
where h is some element of H. There are six cases to consider:
Case 1 (i = 0, j < n1) hyi x j = hy0 x j = hx j Hy0 A.
Case 2 (i = 0, j = n1) hyi x j = hy0 xn1 = hxn1 = hy1 Hy1 A.
Case 3 (i > 0, j > ni)
hyi x j = h( xn1 . . . x j x j1 . . . xni ) x j
= h ( x n 1 . . . x j x j 1 x j x j 2 . . . x n i )
= h ( x n 1 . . . x j +1 x j 1 x j x j 1 x j 2 . . . x n i )
= h ( x j 1 x n 1 . . . x j +1 x j x j 1 . . . x n i )
= hx j1 yi Hyi A.
Case 4 (i > 0, j = ni)
hyi x j = hxn1 . . . xni xni
= hxn1 . . . xni+1
= hyi1 Hyi1 A.
Case 5 (i > 0, j = ni 1)
hyi x j = hxn1 . . . xni xni1
= hyi+1 Hyi+1 A.
presentations 161
| Gn | 6 | A| 6 n| H | 6 n(n1)! = n!
Since | Gn | 6 | A| and f : Gn Sn is surjective, and since we know
from Proposition 1.52 that |Sn | = n! this means that the only possibility
is that f is also injective, and hence an isomorphism.
Thus Gn = Sn and the given presentation is a valid one for Sn .
The given presentation for Sn is a bit complicated in general, so its
illuminating to look at a specific example.
Example 5.34 The symmetric group S3 has presentation
h x, y : x2 = y2 = ( xy)3 = ei.
The symmetric group S4 has presentation
x, y, z : x2 = y2 = z2 = ( xy)3 = (yz)3 = e,
xz = zx .
Therefore
We can take this a little further and replace any relator with a conjugate
of itself, in particular a cyclic permutation:
Example 5.38 Let Gm,n = h X : Ri = h x, y : x m yn i. Since x m yn = e
we can take the conjugate with x to get x 1 x m yn x = x 1 ex = e
and hence x m1 yn x = e. So x m1 yn x is also in the normal closure
h R F(X ) i and hence we can perform an R operation to get
Gm,n = h x, y : x m1 yn x i.
The next example leads on from Example 5.34 and Proposition 5.28 to
show that the two presentations for D3
= S3 are indeed equivalent.
Example 5.40 From Proposition 5.28 we know that
D3
= h a, b : a3 , b2 , ( ab)2 i,
and from Proposition 5.33 and Example 5.34 we know that
D3
= S3
= h x, y : x2 , y2 , ( xy)3 i.
First we apply an X+ move to the first presentation, introducing the
generator x and the relator x 1 ab (equivalent to x = ab). This yields
h a, b, x : a3 , b2 , ( ab)2 , x 1 abi.
Next, we perform another X+ move to introduce another generator
y and relator by1 (which is equivalent to the relation y = b). This
gives
h a, b, x, y : a3 , b2 , ( ab)2 , x 1 ab, by1 i.
We can now use an R+ move to add a new relator y2 and follow it
with an R move to delete the relator b2 ; this results in the presenta-
tion
h a, b, x, y : a3 , ( ab)2 , x 1 ab, by1 , y2 i.
Similarly, we can replace the relator x 1 ab with x 1 ay, and replace
the relator ( ab)2 with ( ay)2 to obtain the presentation
h a, b, x, y : a3 , ( ay)2 , x 1 ay, by1 , y2 i.
We can now perform an X move to delete the generator b and the
relator by1 , in order to get the presentation
h a, x, y : a3 , ( ay)2 , x 1 ay, y2 i.
Another R transformation replaces the relator x 1 ay with its cyclic
conjugate ayx 1 , and then we perform two more R moves to replace
presentations 165
5.3 Finitely generated abelian groups It is the peculiar beauty of this method,
gentlemen, and one which endears it to
the really scientific mind, that under no
In Section 5.1 we studied free groups, the most general possible circumstance can it be of the smallest
groups we can generate from a particular set of generators. Its worth possible utility.
Henry Smith (18261883)
asking what is the most general possible abelian group we can con-
struct from a given set of generators. We could go through the same
construction as for free groups, in addition requiring that all the gen-
erators commute with each other. This definitely results in an abelian
group whose elements are commuting words in the generators. Unlike
with the more general case, we allow arbitrary arbitrary reordering
of generators, which in particular means that we can group together
generators of the same type in a nice, neat and straightforward way.
For example, suppose X = { x, y, z} is a set of generators, and that
w1 = x2 y1 xz3 y2 and w2 = y 1 z 1 x 2 z 3 y 2 .
Allowing arbitrary reordering means we can reduce these two words
to get
w1 = x 3 y 3 z 3 and w2 = x 2 y 3 z 2 .
Concatenation, reduction and reordering of these words gives
w1 w2 = x 5 y 6 z 5 = w2 w1 .
As before, this concatenationreorderingreduction operation is an
associative binary operation on the set of commuting words in X ,
the empty word serves as an identity element, and each word has a
unique, well-defined inverse; for example
w11 = x 3 y3 z3 and w21 = x 2 y3 z2 .
We therefore have all the necessary ingredients for an abelian group:
Definition 5.42 The group just described is the free abelian group
FA( X ) generated by the set X. If X is finite and | X | = n, we will
often denote this as FAn .
Free abelian groups FA( X ) satisfy a similar fundamental property as
the free group F ( X ) except we restrict our attention to abelian groups:
Definition 5.43 Let X F be a set, and A be some arbitrary abelian
group. Then an abelian group F is free abelian on X if any function
f : X A extends uniquely to a homomorphism f : F A.
xnx
n
w = x1 1 x2n2 . . . = or w = n1 x1 + n2 x2 + = nx x
xX xX
where n1 , . . . Z. Most of the time, well use additive notation for
free abelian groups and their quotients.
Proposition 5.44 Let X = { x1 , . . . , xn } be a finite set of generators.
Then the free abelian group FAn = FA( X ) is isomorphic to the nfold
direct sum Zn = Z Z.
Wikimedia Commons
(Julius Wilhelm) Richard Dedekind
(18311916) was born in Braunschweig,
Proof Written additively, an element of FAn is a word of the form
Germany, the fourth and youngest
w = m1 x1 + + m n x n
child of Julius Dedekind, a professor
at the Collegium Carolinum (now the where m1 , . . . , mn Z.
Technische Universitt Braunschweig).
He entered the University of Gttingen The function f : FAn Zn that maps a word w to the ntuple
in 1850, where he studied with Carl (m1 , . . . , mn ) in Zn is clearly bijective. Its also a homomorphism,
Friedrich Gauss (17771855). Dedekind
became Gauss last student, and in 1852 since for any
was awarded a doctorate for a thesis en-
u = a1 x1 + + a n x n and v = b1 x1 + + bn xn
titled ber die Theorie der Eulerschen In-
tegrale (On the theory of Eulerian inte- we have uv = ( a1 + b1 ) x1 + ( an + bn ) xn and so
grals). As is often the case, this thesis
was adequate but did not display the f (uv) = ( a1 +b1 , . . . , an +bn ) = ( a1 , . . . , an )+(b1 , . . . , bn ) = f (u)+ f (v)
talent evident in his later work. Over
the next two years he studied the lat- as required. Hence f is an isomorphism.
est developments in mathematics, and
in 1854 was awarded his habilitation (a The following proposition is the free abelian analogue of a very similar
postdoctoral qualification required to fact in linear algebra.
teach at a German university).
During his career he made many im- Proposition 5.45 Let F be an abelian group, and let X = { x1 , . . . , xn }
portant advances in analysis, algebra F. Then F is free abelian of rank n and the elements of X are Zlinearly
and number theory. In particular,
he devised the notion of a Dedekind independent if and only if any element of F can be written as a unique
cut, used in one construction of the Zlinear combination of the elements x1 , . . . , xn .
real numbers R. He became an early
admirer of the work of Georg Can- Proof Suppose that any element g F can be written uniquely as the
tor (18451918) on transfinite numbers,
and was a staunch ally in his disputes Zlinear combination
with Leopold Kronecker (18231891).
g = a1 x1 + + a n x n
He also devised the concept of an ideal
and used it to provide a purely alge- where a1 , . . . , an Z. Clearly X generates F, but we want to show that
braic proof of the RiemannRoch The-
orem, an important result in algebraic F is actually the free abelian group FA( X ) generated by X, and which
geometry. consists of all possible formal Zlinear combinations of x1 , . . . , xn .
He retired in 1894 and returned to
Braunschweig, but continued his re-
Let f : X , F be the inclusion function. Then by the fundamental
search and taught occasionally for mapping property of free abelian groups, there exists a unique homo-
some years. He died in 1916 aged 84. morphism f : FA( X ) F which extends f , and which maps a given
Zlinear combination in FA( X ) to the corresponding sum in F. We
presentations 169
Zn1 Zn k Z Z =
n
h x1 : x1n1 i h xk : xk k i h xk+1 : i h xm : i =
n
h x1 , . . . , xm : x1n1 = = xk k = e, [ xi , x j ] = e for 1 6 i, j 6 mi
Example 5.52 Lets list all the abelian groups of order 48. The integer
48 factorises as 22223, so we have the following possibilities:
Z48 , Z2 Z24 , Z2 Z2 Z12
Z4 Z12 , Z2 Z4 Z6 , Z2 Z2 Z2 Z6 ,
Z8 Z6 , Z8 Z6 , Z2 Z2 Z2 Z2 Z3 ,
Z16 Z3 , Z2 Z8 Z3 , and Z4 Z4 Z3 .
From Proposition 1.32 we know that
Z48
= Z16 Z3 ,
Z2 Z24
= Z8 Z6
= Z2 Z8 Z3 ,
Z2 Z2 Z12
= Z2 Z4 Z6
Z2 Z2 Z4 Z3 ,
=
Z4 Z12
= Z4 Z4 Z3 ,
and Z2 Z2 Z2 Z6
= Z2 Z2 Z2 Z2 Z3 .
There are therefore five different abelian groups of order 48, whose
invariant factor decompositions are as follows:
Z48 , Z2 Z24 , Z2 Z2 Z12 , Z4 Z12 and Z2 Z2 Z2 Z6 .
Well split the proof of Theorem 5.48 into two parts: existence (every
finitely generated abelian group G has an invariant factor decomposi-
tion) and uniqueness (isomorphic finitely generated abelian groups
have the same invariant factor decomposition).
In order to prove the first part, well need the following lemma, which
is really just a consequence of Proposition 5.35 in an abelian setting.
Lemma 5.53 Let X = { x1 , . . . , xn }, and let
Y = { x1 , . . . , x j1 , x j + mxi , x j+1 , . . . , xn }
where 1 6 i 6= j 6 n and m Z.
Then if X is a basis for some abelian group G, so is Y.
presentations 173
So, for an abelian group G we can replace any generator x j with a new
generator of the form x j + mxi where i 6= j and m Z.
Proof This is just an application of Tietze transformations. Replacing
the generator x j with a new generator y = x j + mxi amounts to an X+
move introducing a generator y and a relator y x j mxi , followed by
an X move removing the generator x j and the relator y x j mxi .
Proof of Theorem 5.48 (existence) Our abelian group G has genera-
tors x1 , . . . , xs , and relators r1 , . . . , rt , each of which are of the form
ri = ni,1 x1 + + ni,s xs .
If t = 0 (that is, if there are no relators) then we have the free abelian
group FAs of rank s, which is isomorphic to Zs by Proposition 5.44.
Suppose instead that t > 1. Then there exists at least one nontrivial
relator, and amongst those relators there will be a smallest positive
coefficient ni,j attached to the generator x j in relator ri .
We can relabel the generators and relators15 so that this coefficient is 15 This wont change the group un-
n1,1 and were working with the relator der investigation, except by an isomor-
phism, the details of which dont mat-
r1 = n1,1 x1 + + n1,s xs . ter to this discussion.
choices for k1 , together with gcd(n2 , q) valid choices for k2 , all the way
up to kr , for which there are gcd(nr , q) possibilities. Hence the number
of elements k = (k1 , . . . , kr ) satisfying qk = 0 is equal to the product of
the greatest common divisors, namely gcd(n1 , q) gcd(nr , q).
Proof of Theorem 5.48 (uniqueness) We will prove uniqueness in
two parts, first by considering the infinite-order component, and then
by considering the torsion coefficients m1 , . . . , mk .
Let Gfin =Zm1 Zmk and Ginf =Zr , and let Hfin =Zn1 Znl
and Hinf =Zs , so that G = Gfin Ginf and H = Hfin Hinf .
There is a surjective projection homomorphism f 1 : G Ginf which
maps ( g1 , . . . , gk , gk+1 , . . . , gr ) 7 ( gk+1 , . . . , gr ). The kernel ker( f 1 ) of
this homomorphism is obviously the finite-order component Gfin , and
its image is equally obviously the infinite-order component Ginf = Zr .
By the First Isomorphism Theorem16 we see that 16
Theorem 4.40, page 119.
G/ ker( f 1 ) = G/Gfin
= im( f 1 ) = Ginf = Zr .
We can define a similar homomorphism f 2 : H Hinf which maps
(h1 , . . . , hl , hl +1 , . . . , hs ) 7 (hl +1 , . . . , hs ), and use this to show that
H/ ker( f 2 ) = H/Hfin
= im( f 2 ) = Hinf = Zs .
Any isomorphism : G H must map the elements of finite order
in G to the elements of finite order in H, by Proposition 1.21, so
Gfin
= Hfin , and hence there must be an induced isomorphism
: G/Gfin = Ginf H/Hfin = Hinf ,
which means that Zr
= Zs and so r = s as claimed.
All that remains is to prove uniqueness of the torsion coefficients. As
just noted, Gfin
= Hfin . Suppose, without loss of generality, that k > l.
Then we apply Lemma 5.54 to Gfin and Hfin with q = m1 to see that
gcd(m1 , m1 ) gcd(m1 , mk ) = gcd(m1 , n1 ) gcd(m1 , nl )
and since m1 |m2 | . . . |mk we have
gcd(m1 , m1 ) = gcd(m1 , m2 ) = = gcd(m1 , mk ) = m1
and therefore
m1k = gcd(m1 , n1 ) gcd(m1 , nl ).
Each of the factors gcd(m1 , n1 ), . . . , gcd(m1 , nl ) on the right hand side
can be at most m1 , which forces k = l. This in turn means that
gcd(m1 , n1 ) = m1 , and hence m1 must be a factor of n1 .
Now apply Lemma 5.54 with q = n1 to get
gcd(m1 , n1 ) gcd(mk , n1 ) = n1k
so gcd(m1 , n1 ) = n1 and thus n1 divides m1 . Therefore m1 = n1 .
176 a course in abstract algebra
0 am2 amn
and we can apply steps 15 to the (m1) (n1) submatrix
a22 a2n
. ..
.
. .
am2 amn
to get an mn matrix where the only nonzero elements in the first
two rows and columns are on the diagonal. We repeat this until
we get a matrix whose only nonzero entries are on the diagonal.
These diagonal entries wont necessarily satisfy the divisibility
criterion a11 | a22 | . . . | arr . If they dont, we proceed to the next step.
6 For 1 6 i 6 r 1, compare aii with each a jj for i < j 6 r. Let i and
j be the lowest integers for which aii doesnt divide a jj . Use a row
operation of type E3 to add row j to row i, and then reduce this
new mn matrix using steps 15.
x x x
definition deduction
1 2 3 1
1x = 2
2 3 1 2
2x = 3
3 3
At this point we see that the first row is full. In particular, this yields
the expression 3x = 1 for free: we write this in the deduction column
of the coset table, and mark it in the relator table by an equals sign
= across the appropriate dividing line:
x x x
definition deduction
1 2 3 = 1
1x = 2
2 3 1 2
2x = 3 3x = 1
3 3
x x x
definition deduction
1 2 3 = 1
1x = 2
2 3 1 2
2x = 3 3x = 1
3 1 2 3
the relator tables we can see where to draw the directed edges.
1 2
For the group G = h x : x3 i
= Z3 studied in Example 5.65 we get the x
diagram shown in Figure 5.2. Figure 5.2: Cayley graph for the group
h x : x3 i
= Z3 , constructed from the
This procedure extends to presentations with more than one relator: relator tables produced by the Todd
Coxeter algorithm. Here the node 1
we have a relator and coset table for each relator, and must keep all
corresponds to the identity element e,
tables up to date at each step. The following example illustrates this. the node 2 corresponds to the element
x, and the node 3 corresponds to the
Example 5.67 Let G = h x, y : x2 , y2 , xyx 1 y1 i. This is isomorphic element x2 .
to the Klein group V4
= Z2 Z2 .
x x y y x y x 1 y 1
1 1 1 1 1 1
We begin with the first relator, writing a 2 in the available space in the
relator table, and adding the definition 1x = 2 to the first column in
the coset table. This yields the deduction 2x = 1. We then add a new
row to the other relator tables and fill in 1s and 2s as appropriate.
x x y y x y x 1 y 1
1 2 = 1 1 1 1 2 1
2 1 2 2 2 2 2
definition deduction
1x = 2 2x = 1
Next, add a 3 to the first empty space in the second table, add a new
definition 1y = 3, and fill in the relator tables as far as possible.
x x y y x y x 1 y 1
1 2 = 1 1 3 = 1 1 2 3 1
2 1 2 2 2 2 2
3 3 3 1 3 3 1 3
definition deduction
1x = 2 2x = 1
1y = 3 3y = 1
Now we write a 4 in the empty cell in the second row of the second
184 a course in abstract algebra
x x y y x y x 1 y 1
1 2 = 1 1 3 = 1 1 2 4 = 3 1
2 1 2 2 4 = 2 2 1 3 4 2
3 4 3 3 1 3 3 4 2 1 3
4 3 4 4 2 4 4 3 1 2 4
definition deduction
1x = 2 2x = 1
1y = 3 3y = 1
2y = 4 4y = 2
3x = 4
4x = 3
x x y x x x y y definition deduction
123=1 1 2 34=3 1 1x = 2
2 3 4=2 2 3 4 2 2x = 3 3y = 1
3 4 3 3 4 4 3 3x = 4 4y = 3
4 4 4 4 4y = 2
x x y x x x y y definition deduction
122=1 1 2 23=2 1 1x = 2
2 2 3=2 2 2 3 2 2x = 2 2y = 1
3 3 3 3 2x = 3 3y = 2
definition deduction
x x y x x x y y
1x = 2
122=1 1 2 22=2 1
2x = 2 2y = 1
2 2 2=2 2 2 2 2
2y = 2
x x y x x x y y definition deduction
111=1 1 1 11=1 1 1x = 1 1y = 1
This leaves us with just a single row in both relator tables, from
which we conclude that | G | = 1. Thus G must be the trivial group.
With a little bit of work, we can amend the basic ToddCoxeter algo-
rithm to work relative to some finitely-generated subgroup H of the
group G were interested in. In this case, instead of obtaining the order
| G | of the main group, we get the index | G : H | of H in G. And rather
than a permutation representation for G, we get a representation of
the way G permutes the (right) cosets of H.
The procedure is very similar to that in Examples 5.65, 5.67 and 5.68,
but we also have relator tables for the words which generate H. These
subgroup tables are similar to relator tables in the basic version of the
algorithm, but have a single row beginning and ending with 1.
Example 5.69 Let G = h x : x4 i and let H = h x2 i. We can see
immediately that G = Z4 and that H is an index2 subgroup which
is isomorphic to Z2 . We write down the relator table, the subgroup
table and the coset table:
x x x x x x definition deduction
1 1 1 1
empty cells in the relator table and the subgroup table, and adding
the definition 1x = 2 to the coset table. We add a second row to the
relator table, but not to the subgroup table. From the subgroup table
we deduce that 2x = 1, and use this to complete all the tables:
x x x x x x definition deduction
12 12 1 1 2=1 1x = 2 2x = 1
2 1 21 2
x x x y y x y x y y
123=1 1 1 1 1 2 3 1 1 1=1
23=1 2 2 3=2 2 3 2 3 1
3 1 2 3 3 2=3 3 1 1 2=1
definition deduction
1y = 1
1x = 2
2x = 3 3x = 1
2y = 3
3y = 2
We know from Proposition 2.29 that the cosets of H in G are all the
same size, so one way of picturing a transversal is to take all of the
cosets, line them up next to each other, and slice through all of them
in such a way that we hit a single element in each one.
The ToddCoxeter algorithm, in addition to counting the number of
cosets of H in G, also yields a right transversal for the H in G.
In Example 5.65 we applied the ToddCoxeter algorithm to the group
h x : x3 i
= Z3 and found, amongst other things, that | G | = 3 as
188 a course in abstract algebra
5.B Transversals
Perhaps the best-known examples of orders are the usual ones < and
6 defined on the real numbers R, and indeed these are the models on
which the above definition is based.
For our purposes we need a slight refinement of the above concepts:
Definition 5.75 Let 6 be a total ordering defined on a set S, and
let < be the corresponding strict total ordering. If every nonempty
subset of S has a least element with respect to 6 and <, then we say
that S is well-ordered, and that both 6 and < are well-orderings.
With a bit of thought it should be obvious that any finite set can be Axiom of Choice Given any collec-
well-ordered: arrange the elements of the set in a list and define a tion X of nonempty sets, there exists a
strict ordering corresponding to this arrangement. Then the least function f defined on X (a choice func-
tion) such that f (S) S.
element is simply the first element on the list. This method also
extends to countably infinite sets, but the assertion that any set can be Zorns Lemma Let S be a partially-
well-ordered is equivalent to Zorns Lemma or the Axiom of Choice. ordered set such that every totally-
ordered, nonempty subset of S has an
A discussion of the details of axiomatic set theory is beyond the scope upper bound. Then S contains at least
of this book, but the statement of both the Axiom and the Lemma are one maximal element.
given here for completeness, without further comment.
In particular, well use the following strict well-ordering, called vari-
ously the shortlex, radix, lenlex or length-lexicographical order.
190 a course in abstract algebra
In this proof we used the fact (left as an exercise) that for the shortlex
ordering < if v, w F and x X , with v < w, then vx < wx and
xv < xw. In fact, any ordering of F satisfying this property will suffice.
Now that we know every subgroup of a free group has a Schreier
transversal, lets look at a couple of examples.
Example 5.78 Let F = F2 = h x, y : i and let H = h x2 , y2 , xyx 1 y1 i.
In particular, F/H = h x, y : x2 , y2 , xyx 1 y1 i
= V4 , so | F : H | = 4
and thus any transversal of H in F must have four elements. It so
happens that
U = {e, x, y, xy}
is a Schreier transversal for H in F. Furthermore, U is the Schreier
transversal constructed in Proposition 5.77 since each element of U
is the least element in its coset relative to the shortlex ordering.
word e x y x 1 y 1 x2 xy xy1 yx
coset H Hx Hy Hx 1 Hy Hx 1 Hxy Hxy Hyx
Theorem is to find generators for the subgroup H, for which we use the
shortlex-minimal Schreier transversal provided by Proposition 5.77.
First, though, we introduce another little piece of notation. Suppose
that w F is some word in a free group F, and that U is a Schreier
transversal for a subgroup H 6 F. We form the coset Hw and take
the intersection Hw U. This will consist of a single element which
might (but wont necessarily) be equal to w. Denote this element by w.
This process yields a function : F U, given by w 7 w for all w F.
This bar function has a number of properties which we state here
without further proof. Its idempotent, in the sense that w = w for all
w F. Also, Hw = Hw for all w F, and w = w if and only if w U.
Let F be a free group generated by some basis X, let H be a subgroup
of F, and let U be a Schreier transversal for H in F. In the next part of
this discussion well need the set
Y = {uxux 1 : u U, x X }.
Its worth thinking a little bit about what this set Y actually means.
Suppose that u is an element of the Schreier transversal U, and that
x is either a generator, or the inverse of a generator, in X. Then ux is
clearly a word in F, and ux is the corresponding transversal element in
U, and may or may not be equal to ux. If ux = ux, which will happen
exactly when ux U, then the element uxux 1 will be equal to the
identity e F. So uxux 1 tells us whether ux is also in U.
Now consider the element uxx 1 . This is in the same coset as uxx 1 ,
which is in the same coset as uxx 1 = u. Hence
uxx 1 = u. (5.2)
Well use this fact a couple of times later, but first lets construct the
set Y in a specific case to see how it works.
y e y2 yx 1 y1 x 1 e
x 1 e 1
x yx y 1 1 x 3 x 1 y 1 x 1 y 1
xy xyx 1 xy2 x 1 xyx 2 y1 e
yx 2
yx y x 1 1 yxyx e yxy1 x 1
presentations 193
yi = ui xi ui xi 1 = ui xi ui+11 .
Then
T = {uru1 : u U, r R}
5.C Triangles, braids and reflections A bas Euclide! Mort aux triangles!
(Down with Euclid! Death to triangles!)
Jean Dieudonn (19061992)
In this section we will look briefly at a few classes of groups that
can be defined by presentations of particular forms, but which have Although it has been proved that ev-
interesting geometric interpretations. ery braid can be deformed into a sim-
ilar normal form the writer is con-
We start with the dihedral group Dn , which can be expressed by the vinced that any attempt to carry this
presentation out on a living person would only lead
to violent protests and discrimination
h x, y : x2 = y2 = ( xy)n = ei. against mathematics. He would there-
fore discourage such an experiment.
This group has an interpretation as the symmetry group of the regular Emil Artin (18981962),
ngon. Both generators x and y represent reflections, all of which Theory of braids, Annals of
Mathematics 48 (1947) 101126
have order 2, and the product xy represents a 2 n rotation, which has
order n. If we choose x and y to be the right reflections, then we can
generate the entirety of Dn just using them. In general there are many
valid choices, but setting x = m1 and y = m2 works.
The symmetric group Sn has a presentation
xi2 = e 1 6 i 6 n 1
* +
x 1 , . . . , x n 1 : ( x i x i +1 )3 = e 1 6 i 6 n 2 .
xi x j = x j xi 1 6 i < j 6 n1, ji > 1
or
for D3 , and
for D2 = V4
= Z2 Z2 .
4
Bn n>2 2n n! ncube,
ncross
4
F4 1 152 24cell
6
G2 12 regular hexagon
5
H3 120 dodecahedron,
icosahedron
5
H4 14 400 120cell,
600cell
m
I2 (m) m>2 2m regular mgon
A
en n>2 nsimplex
e2 4 4
B =Ce2 square tiling
4
B
en n>3 ndemicubic
4 4
C
en n>3 ncubic
D
en n>4 ndemicubic
E
e6 222
E
e7 331 and 133
E
e8 521 , 251 and 152
4
F
e4 16cell and 24cell
6
G
e2 hexagonal and triangular
complicated reasons that need not concern us here (but which relate
to Lie groups and Lie algebras) the case m = 6 is sometimes treated
as a special case, and denoted G2 ; also I2 (3)
= A2 and I2 (4)
= B2 .
Related to these are the affine Coxeter groups, listed in Table 5.4,
which describe the symmetries of certain regular and semiregular
tilings (or honeycombs) of Rn . They all have infinite order. The
names of each of these Coxeter graphs and their associated groups
indicate that an affine group X
e n is obtained from the corresponding
presentations 203
e n contains n+1
finite group Xn by adjoining an extra vertex; so X
vertices, not n.
Another class of Coxeter groups relate to reflections in hyperbolic
space; these will not concern us here, but a full list can be found in
Humphreys book.23 23
Humphreys, Reflection Groups and
Coxeter Groups, Sections 6.8 and 6.9.
The mathematical study of knots is a vibrant and varied area of
research dating back to the middle of the 19th century. An important
and related concept is that of a braid: an arrangement of finitely
many parallel strings, usually drawn vertically, where adjacent strands
are allowed to cross over each other, but must always point strictly
downwards (we dont allow them to be horizontal or double back on
themselves at any point). See Figure 5.4 for an example. Considered
as topological objects, we say two braids are equivalent if one can be
rearranged to look like the other if the top and bottom end of each
string is kept fixed in place. (Technically, we say that two braids are
equivalent if they are ambient isotopic relative to their endpoints.)
It transpires that any such rearrangement can be decomposed as
a finite sequence of simple moves of three basic types. Type one
allows us to move crossings on non-adjacent strings up and down Figure 5.4: A braid
(see Figure 5.5). A move of the second type allows the introduction
or removal of a pair of cancelling crossings (see Figure 5.6). And the
third move involves the interaction of three crossings in three adjacent
strands, where a strand is allowed to pass over or under a crossing Figure 5.5: A braid move of type 1
in the other two strands (see Figure 5.7). The second and third of
these are essentially braid-theoretic versions of the second and third
1 i 1 i i +1 i +2 n denotes the braid where string i +1 crosses over string i (see Figure 5.8)
i : while i1 is the braid where string i crosses over string i +1.
We can use this viewpoint to construct the following presentation for
1 i 1 i i +1 i +2 n
Bn , first proved in 1925 by the Austrian mathematician Emil Artin
i1 :
(18981962).
Figure 5.8: Elementary braids
Proposition 5.92 Let Bn be the nstring braid group. Then
* +
i j = j i for |i j| > 1
Bn = 1 , . . . , n1 : .
i i+1 i = i+1 i i+1 for 1 6 i 6 n2
The kernel of this homomorphism consists of all braids that map to the
identity permutation . These are exactly the ones in which each string
begins and ends in the same position; they are called pure braids, and
(4) (3) (2) (1) (5)
the subgroup PBn = ker is the nstring pure braid group.
Figure 5.9: Reading a permutation
from a braid This connection between the braid group Bn and the symmetric group
Sn , which is isomorphic to the Coxeter group An1 , yields an entire
family of groups that are related in a similar way to the other Coxeter
groups. First we introduce some notation. Let
h x yim = xyxy . . .
| {z }
m terms
i and vertex i +1 in the An1 Coxeter diagram. We can also write the
other relation in this form too:
i j = hi j i2 = hj i i2 = j i ,
Summary
J E Humphreys, Reflection Groups and Coxeter Groups, Cambridge Studies in Advanced Mathematics
29, Cambridge University Press (1990)
A readable textbook on Coxeter groups, suitable for graduate students and final-year undergraduates.
Exercises
5.1 As in Exercise 1.23, let S = R \ {1, 0, 1}, let f : S S such that f ( x ) = 11+ x
x , and denote by f
n
6.1 Symmetries and transformations The chief forms of beauty are order and
symmetry and definiteness, which the
mathematical sciences demonstrate in
Recall that in Chapter 1 we met a class of groups consisting of a special degree.
symmetry operations defined on some geometric object. Aristotle (384322 BC),
Metaphysics XIII:3
The dihedral groups Dn , for example, consist of rotations and reflec-
tions of a regular nsided polygon. More precisely, we define a subset
m2
Pn of R2 consisting of the regular polygon with vertices at the points
(2k1) (2k1) m3 m1
cos n , sin n for 1 6 k 6 n. 2
1
The reflection m1 determines a particular bijection Pn Pn which
leaves every point along the line y = x tan n fixed where it is, and
m4
swaps every other point with its mirror image in that line. Similarly,
the rotation r fixes the origin and cycles all other points among them-
3
selves in groups of n, and thus also determines a bijection Pn Pn . 4
The first says that the identity e must act as the identity operation on
X, while the second is an associativity condition.
What weve just defined here is more properly called a left action of
G on X. There is a corresponding notion of a right action, which is a
function : X G X satisfying
( x, e) = x, (6.5)
( ( x, g), h) = ( x, g h), (6.6)
or, alternatively,
x e = x, (6.7)
( x g ) h = x ( g h ), (6.8)
for all x X and g G.
It so happens that there is a bijective mapping between left and right
Gactions on a given set X, in the sense that for any left action : G
X X there is a unique right action : X G X and vice versa.
This correspondence doesnt work in quite the way we might expect,
however. The tempting but nave approach is to take a left action
: G X X and define a right action : X G X by setting
( x, g) = ( g, x ) for all g G and x X.
But this doesnt quite work. It certainly satisfies condition (6.7) since
( x, e) = (e, x ) = x.
Condition (6.8), however, fails in general when G is nonabelian:
( ( x, g), h) = (h, ( g, x )) = (h g, x ) = ( x, h g) 6= ( x, g h).
We can make this work, however, if instead of defining ( x, g) =
( g, x ) we define ( x, g) = ( g1 , x ) for all g G and x X. Here,
( x, e) = (e1 , x ) = (e, x ) = x
as before, but
( ( x, g), h) = (h1 , ( g1 , x )) =
(h1 g1 , x ) = (( g h)1 , x ) = ( x, g h).
In the three motivating examples earlier, we interpreted the action
of Dn on Pn , the action of GLn (R) on Rn and the action of Sn on
Xn as functions of the form G X X satisfying the properties in
Definition 6.1. But in all three cases we noted that the action of a
particular group element on the set determined a bijective symmetry
map, invertible linear transformation or a permutation of that set.
So we could also regard those group actions as ways of assigning
a permutation of X to each element of the group G. This is true in
general, as the following proposition shows.
214 a course in abstract algebra
( g h)( x ) = f gh ( x ) = ( g h) x = g (h x ) = f g ( f h ( x ))
= ( f g f h )( x ) = (( g) (h))( x )
as required.
Similarly, any homomorphism : G Sym( X ) determines a unique
left action of G on X, and since left Gactions on X are in bijective
correspondence with right Gactions on X, any such homomorphism
determines a unique right action too.
So Gactions on a set X are essentially the same as homomorphisms
G Sym( X ). We studied homomorphisms quite extensively in
Chapter 4, so by applying some of the things we learned, we should
be able to get some important insights into how group actions work.
Example 6.3 By examining how the elements of D4 permute the
vertices of the square P4 , as shown in Figure 6.3, we can obtain
an action of D4 on the set X4 = {1, 2, 3, 4}. This determines a
homomorphism : D4 S4 such that
2 1 e 7 , r 7 (1 2 3 4),
2 3
r 7 (1 3)(2 4), r 7 (1 4 3 2),
m1 7 (2 4), m2 7 (1 2)(3 4),
m3 7 (1 3), m4 7 (1 4)(2 3).
The next most obvious group action is the one where nothing happens:
Example 6.6 For any group G and any set X we can define the
trivial action of G on X by setting g x = x for all g G and x X.
This certainly satisfies conditions (6.3) and (6.4) and is hence a valid
example of a group action. From the alternative point of view of a
homomorphism G Sym( X ) this is exactly the homomorphism that
maps every element g G to the identity permutation Sym( X ).
This action is not faithful in general, except when the group G is
trivial.
216 a course in abstract algebra
We can take this viewpoint a little further. Suppose that a group G acts
on some set X, and suppose also that H is some subgroup of G. Then
we can define an action of H on X in a straightforward way: every
element of H is also an element of G after all, so for h H and x X
we can just define h x to be whatever it is in the action of G on X.
Another way of looking at this is to define the homomorphism : H
Sym( X ) by composing the inclusion homomorphism i : H , G with
the action homomorphism : G Sym( X ).
Example 6.8 A concrete example of this is given by the inclusion
of the special orthogonal group SOn (R) in the general linear group
GLn (R). Once weve chosen a basis for Rn , that gives us an action
of GLn (R) on Rn : a well-defined method of combining an invertible
nn real matrix with an ncomponent real vector to get another
ncomponent real vector, in such a way that the usual Rlinearity
conditions are satisfied. Whatever other properties (unit determinant,
orthogonality) elements of SOn (R) might have, they are still nn
invertible real matrices, so we can define an SOn (R)action on Rn
by reusing the GLn (R) action we already had.
4
Theorem 2.13, page 47. By Cayleys Theorem4 , any group G can be regarded as a subgroup of
the symmetric group Sym( G ) of permutations of its underlying set.
More precisely, we construct an isomorphism f : G H < Sym( G )
by mapping each element g G to the bijection g : G G defined
by g (h) = g h for all h G. This yields an isomorphism
f : G H = { g : g G } 6 Sym( G ).
We can compose this isomorphism f with the inclusion homomor-
phism i : H , Sym( G ) to get the required action homomorphism
= i f : G Sym( G ).
actions 217
One of the standard group actions we met in the last section was the
trivial action. Its orbits have a particularly simple form:
Example 6.19 Let a group G act trivially on some set X, in the sense
that g x = x for all g G and x X. Then for some arbitrary x X,
the orbit OrbG ( x ) = { g x : g G } = { x }. So the trivial action
partitions a set X into singleton subsets consisting of the individual
elements of X.
Another important action is the conjugation action of a group G acting
6
Definition 6.11, page 217. on itself.6 We met the orbits of this action in a slightly different context
7
Proposition 3.4, page 74. in Chapter 3.7
actions 221
Example 6.20 Let a group G act on its underlying set via the conju-
gation action. Then the orbit of an element h G is the set
OrbG (h) = { g h : g G } = { g h g1 : g G }.
But this is exactly the conjugacy class of h in G. Two elements
h, k G are therefore in the same orbit if they are conjugate: if there
exists some element g G such that k = g h g1 .
Sometimes, all elements of a Gset X lie in the same orbit. The action
of Sn on Xn (or, more generally, of Sym( X ) on a set X) has this property.
That is, for any two elements x, y X, there is at least one element
g G for which g x = y. Another concrete example is the action of
GLn (R) on R3 = R3 \ {0}: for any nonzero vectors v, w R3 we can
find an invertible nn real matrix A such that Av = w.
Definition 6.21 An action of a group G on a set X is transitive if it
partitions X into a single orbit. Equivalently, for any x, y X there
exists at least one element g G such that g x = y.
So, the triangles central point is fixed by all of D3 , any other point
on an axis of symmetry is fixed by just a subset of the form {e, m1 },
and a point which isnt on an axis of symmetry is fixed only by the
identity element e. We give subsets of this type a special name:
Definition 6.23 Let a group G act on a set X, and suppose x is some
element of X. The stabiliser of x in G is the set
StabG ( x ) = { g G : g x = x }
of elements in G which act trivially on x.
ker() xX StabG ( x ).
T
T
Therefore xX StabG ( x ) = ker() as claimed.
In light of the above result, the stabiliser StabG ( x ) is sometimes called
the isotropy subgroup.
Returning yet again to the D3 example, we notice that the orbit of the
triangle P3 s central point consists of a single element, and that its
stabiliser consists of all six elements of D3 . If we choose some other
point on a symmetry axis of P3 then we find that the orbit of this point
contains three points and its stabiliser contains two elements of D3 .
actions 223
| OrbD3 ( x )| | StabD3 ( x )| = 6 = | D3 |.
The next theorem confirms that this is true for any (finite) group acting
on a (finite or infinite) set.
Theorem 6.25 (OrbitStabiliser Theorem) Let G be a finite group
acting on a (not necessarily finite) set X. Then
| OrbG ( x )| | StabG ( x )| = | G |
for all x X.
Proof Let y be some element of the orbit OrbG ( x ). Then there exists
at least one element g G such that g x = y. Now suppose that there
exists some other h G such that h x = y. Then we have h x = g x,
and hence ( g1 h) x = x. This means that g1 h is in the stabiliser
StabG ( x ). Proposition 6.24 tells us that StabG ( x ) is a subgroup of G,
and hence by Proposition 2.26 it follows that if g1 h is in StabG ( x )
then h must be in the left coset g StabG ( x ).
So all group elements which map x to y lie in the same coset of
StabG ( x ). Conversely, if an element k G lies in the coset g StabG ( x )
then k x = g x. Thus a coset of StabG ( x ) consists of exactly those
group elements that map x to a given element of the orbit OrbG ( x ).
By Proposition 2.29 we know that | g StabG ( x )| = | StabG ( x )|. So for
any element y OrbG ( x ) there are exactly | StabG ( x )| group elements
that map x to y. By Corollary 2.28, these left cosets completely partition
G, and so there must be exactly | G |/| StabG ( x )| distinct elements in
the orbit OrbG ( x ). Therefore | G | = | OrbG ( x )| | StabG ( x )| for any
x X, as claimed.
Heres another simple example.
Example 6.26 Let Z7 act on the circle C = {ei : 0 6 < 2 }
by n ei = ei( +2/7) . Then the stabiliser StabZ7 ( x ) of some point
x = ei C is just the trivial subgroup {0}, so | StabZ7 ( x )| = 1. Since
|Z7 | = 7, by OrbitStabiliser Theorem, the orbit OrbZ7 ( x ) should
consist of seven points.
And this is exactly what happens. The orbit OrbZ7 ( x ) = OrbZ7 (ei )
is equal to the set {ei+2k/7 : 0 6 k < 7}, which consists of the
point x = ei together with six other points spaced at intervals of 2
7 Figure 6.6: An orbit of the Z7 action
around the unit circle C. on C = {ei : 0 6 < 2 }
224 a course in abstract algebra
c1 = { }
on its own.
Consider two permutations = ( x1 x2 )( x3 x4 ) and = (y1 y2 )(y3 y4 )
in A5 . Let , : X5 X5 be bijections (and hence permutations in S5 )
such that
: x1 7 y1 , x2 7 y2 , x3 7 y3 , x4 7 y4 , x5 7 y5 ;
: x1 7 y2 , x2 7 y1 , x3 7 y3 x4 , 7 y4 , x5 7 y5 .
The identity e obviously fixes all 46656 possible cubes, the six typet
rotations fix 63 = 216 cubes, the three typer2 rotations fix 64 = 1296
cubes, the eight types rotations fix 62 = 36 cubes, and the six typer
rotations fix 63 = 216 cubes.
So, Burnsides Lemma tells us that there are
1 53424
24 (46656 + 6216 + 31296 + 836 + 6216) = 24 = 2226
distinct ways of numbering the faces of the die.
But a conventional die uses each number only once, so how many
distinct dice can we construct with this property? Well, instead of
66 = 46656 possible configurations we have only 6! = 720: we have
six possible choices for the top face, five remaining for the front face,
four for the left face and so on. The identity fixes all 720 of these, but
none of the other 23 direct isometries fix any. So Burnsides Lemma
tells us that there are 720
24 = 30 distinct dice of this type.
Actually, most conventional six-sided dice also satisfy the condition
that opposite faces add up to seven: the 6 face is opposite the 1 face,
the 5 opposite the 2 and the 4 opposite the 3. How many different
dice are there with this property too?
We have six choices of number for the top face, and when weve made
that choice we know what the number on the bottom face has to be.
Next we choose one of the four remaining numbers for the front face,
230 a course in abstract algebra
Fix 2 ( X ) = 22 = 4, Fix 3 ( X ) = 21 = 2.
r r
What if we have three colours at our disposal: black, white and grey?
There are 34 = 81 possible colourings, of which all are fixed by the
identity e, 3 are fixed by r and r2 , and 9 are fixed by r2 , giving the
following values for | Fixg ( X )|:
Fixe ( X ) = 34 = 81, Fixr ( X ) = 31 = 3,
Fix 2 ( X ) = 32 = 9, Fix 3 ( X ) = 31 = 3.
r r
There is an obvious pattern here, and with a little thought we can see
that for k colours the answer is going to be
Fixe ( X ) = k4 , Fixr ( X ) = k1 = k,
Fix 2 ( X ) = k2 , Fix 3 ( X ) = k1 = k.
r r
1 4
Therefore there are 4 (k + k2
+ 2k) rotationally distinct ways of colour-
ing our square picture frame with k different colours.
The key is to study how the isometry group acts on the object were
colouring. In this case, the rotation group Isom+ (2 ) = R4 = Z4
permutes the four sides of the square: this yields a homomorphism
: R4 S4 which we can write down in cycle form as follows:
e 7 = (1)(2)(3)(4), r 7 (1 2 3 4),
2 3
r 7 (1 3)(2 4), r 7 (1 4 3 2).
actions 231
Here weve numbered the edges of the frame anticlockwise with the
numbers 14. For what follows its helpful to also include cycles of
length 1, which for conciseness we dont usually bother doing.
To each cycle we associate a symbol, its type, as follows.
Definition 6.31 Suppose that a permutation Sn can be decom-
posed as a product of disjoint cycles, of which a1 have length 1, a2
have length 2, and so on, where a1 , a2 , . . . are non-negative integers.
Then the type of the permutation is the partition 1a1 , 2a2 , . . . , n an
g monomials csg
0 w6 +6w5 b+15w4 b2 +20w3 b3 +15w2 b4 +6wb5 +b6 = ( w + b )6 x16
1 w6 + b6 = ( w6 + b6 ) x6
2 w +2w3 b3 +b6
6 = ( w3 + b3 )2 x32
3 w6 +3w4 b2 +3w2 b4 +b6 = ( w2 + b2 )3 x23
4 w6 +2w3 b3 +b6 = ( w3 + b3 )2 x32
5 w6 + b6 = ( w6 + b6 ) x6
generating function for g is equal to the product of the corresponding He stayed in Zrich for a number of
years, publishing prolifically on a wide
generating functions. range of mathematical topics. With
his fellow Hungarian mathematician
Therefore the generating function for the element g, which describes Gbor Szego (18951985) he wrote an
the kcolourings of X which are invariant under the action of g, can influential two-volume analysis prob-
be obtained by substituting xi = c1i + + cik in the cycle symbol lem book: Aufgaben und Lehrstze aus
der Analysis, published in 1925.
csg ( x1 , . . . , xn ). In 1940 he moved to the USA, and after
In order to find the generating function that takes account of the whole two years at Brown University was ap-
pointed to a post at Stanford, where he
Gaction on X, we need to do a little more work. remained for the rest of his life. Retir-
m m ing in 1953, he was appointed Professor
For a given colour distribution c1 1 . . . ck k , we obtain the number of
Emeritus, and continued teaching and
equivalent colourings by summing the appropriate term from the researching into his nineties.
generating function for each g G and then divide by | G |. His work spanned a wide range of top-
ics: analysis, algebra, combinatorics,
To get the generating function for all possible colourings, then, we geometry, probability and geometry in
have to take the generating function for all g G, sum them and particular. His classic book on mathe-
matical problem solving, How to Solve
divide by | G |. This gives the expression
It, first published in 1945, has sold over
1 a million copies and been translated
| G | g
csg (c1 + + ck , c21 + + c2k , . . . , c1n + + cnk ) into seventeen languages, and remains
G recommended reading for any student
as claimed. of mathematics.
236 a course in abstract algebra
Summary
14
Definition 1.43, page 21. Motivated by several groups discussed in Chapter 1, particularly the
15
Example 1.34, page 17. dihedral groups Dn ,14 the general linear groups GLn (R),15 and the
16
Definition 1.51, page 24. symmetric groups Sn ,16 we formulated the concept of an action of
17
Definition 6.1, page 212. a group G on a set X:17 a function : G X X satisfying certain
simple criteria. We usually denote the image of a given element x X
under the action of a specific element g G by g x rather than ( g, x ).
More precisely, this is a left action of G on X; there is a corresponding
notion of a right action defined by a function : X G X, in which
we denote the images x g. There is a bijective mapping between left
and right actions, so in principle we need only develop the theory for
left actions, and the analogous results should hold for right actions as
well.
A group action : G X X determines a unique homomorphism
: G Sym( X ), where maps a given group element g G to
18
Proposition 6.2, page 214. the permutation f g of X defined by f g ( x ) = g x for all x X.18 If
this homomorphism is injective, we say that the action is faithful;
equivalently an action is faithful if, for any distinct g1 , g2 G there is
19
Definition 6.4, page 215. at least one x X for which g1 x 6= g2 x.19 Also, an action is faithful
20
Proposition 4.35, page 116. exactly when ker is trivial.20
The trivial action is defined by g x = x for all g G and x X;
this corresponds to the homomorphism G Sym( X ) where every
21
Example 6.6, page 215. element of G maps to the identity homomorphism Sym( X ).21 The
trivial action is not faithful except when G is the trivial group {e}.
The action of the full symmetric group Sym( X ) on a set X is faith-
ful, and corresponds to the identity homomorphism id : Sym( X )
22
Example 6.5, page 215. Sym( X ).22 For any subgroup G 6 Sym( X ) we can construct a faithful
action of G on X by composing the canonical inclusion map i : G ,
Sym( X ) with this symmetric group action. In particular, this process
yields a faithful action of the alternating group Alt( X ) < Sym( X ) on
23
Example 6.7, page 216. X, of An on the finite set Xn = {1, . . . , n},23 and a faithful action of
24
Example 6.8, page 216. the special orthogonal group SOn (R) on Rn .24
25
Theorem 2.13, page 47. By Cayleys Theorem25 any group G can be regarded as a subgroup of
Sym( G ), the permutation group of its underlying set. This viewpoint
26
Definition 6.9, page 217. yields a faithful action of G on itself: the left or right regular action.26
The group conjugation operation introduced in Chapter 3 gives another
action of a group on itself: the conjugation action in which g h =
27
Definition 6.11, page 217. h g = ghg1 for any g, h G.27 This action is not in general faithful. If
28
Proposition 6.13, page 218. G is abelian then the conjugation action is trivial.28
Unfaithful actions have nontrivial kernels, and thereby determine
actions 237
a
45
Definition 6.31, page 231. cs is the formal monomial x11 x2a2 . . . xnan .45
Given a finite group G acting on a finite set X via an action homo-
morphism : G Sym( X ) = Sn , the cycle index of the action is
46
Definition 6.32, page 231. the polynomial CG ( x1 , . . . , xn ) = |G1 | gG csg ( x1 , . . . , xn ).46 Now con-
sider an object with n components, each of which is assigned one of
k colours. Then the Cycle Index Theorem, a corollary to Burnsides
Lemma, states that the number of distinct kcolourings of X modulo a
47
Corollary 6.33, page 231. given Gaction is equal to CG (k, . . . , k ).47
For example, we can use the Cycle Index Theorem to count the number
48
Example 6.34, page 232. of different kcolourings of an nbead necklace.48 , 49
49
Proposition 6.35, page 232. We can obtain more detailed information about these colourings by
50
Theorem 6.36, page 235. constructing a generating function. Plyas Enumeration Theorem50
says that the generating function for the kcolouring problem on an
ncomponent object X, modulo an action by a finite group G, is given
by the cycle index CG (c1 + + ck , c21 + + c2k , . . . , c1n + + cnk ).
Exercises
6.1 Show that CG H = CG CH for any groups G and H.
6.2 Suppose that the necklaces in Example 6.34 can also be flipped over. That is, replace the Z6 action
with an appropriate D6 action. Configurations of this type are sometimes called bracelets. How
many distinct 2coloured 6bead bracelets are there?
6.3 Modify or extend the proof of Proposition 6.35, replacing Zn with Dn , to calculate the number of
kcoloured nbead bracelets.
6.4 Convince yourself that A4 is 2transitive, and that A5 is 3transitive.
6.5 Try to construct a Steiner system of type S(2, 3, 4) and convince yourself that no such system exists.
The known is finite, the unknown infi-
nite; intellectually we stand on an islet
in the midst of an illimitable ocean of
inexplicability. Our business in every
generation is to reclaim a little more
land, to add something to the extent
and the solidity of our possessions.
Thomas Huxley (18251895),
On the Reception of the Origin of Species,
7 Finite groups in: Francis Darwin, The Life and Letters
of Charles Darwin (1887) II 204
Next we will study how subgroups can be nested inside each other
(like algebraic matryoshka dolls), and prove some important results
about these series of subgroups, such as Schreiers Refinement Theo-
rem4 and the JordanHlder Theorem.5 Some of this work will lead 4
Theorem 7.27, page 254.
into our study of Galois Theory in Chapter 11. 5
Theorem 7.34, page 260.
After that, we will introduce the semidirect product G = H oK of two
groups H and K, a generalisation of the direct product H K, which
will lead us to the study of group extensions: a general method of
constructing larger groups from smaller ones.
Finally, we will use some of these techniques to classify, up to isomor-
phism, all the groups with order less than 32.
1 1 2 0
1 0 2 0
0 1 2 0
0 2 2 0
01 , 02 , 11 , 02 , 22 , 02 , 12 , 02
of order 6. By Lagranges Theorem, it could also have a subgroup
of order 12, but it turns out not to. Again, however, it does have
psubgroups of all possible orders.
This happens to be true in general, and to prove it we first need to take
a little detour into the realm of group actions. Let G be a (possibly
infinite) group acting on a finite set X. Then there are finitely many
orbits in X under this action, and these orbits partition X:
r
|X| = | OrbG (xi )| (7.1)
i =1
where x1 , . . . , xr are representative elements from each orbit. Denote
by FixG ( X ) the set of elements of X that are fixed by every element
of G. Each of these elements must therefore be a single-element orbit
of the Gaction, so FixG ( X ) is the union of these singleton orbits.
Suppose there are s of these, with 0 6 s 6 r. Then | FixG ( X )| = s and
we can rewrite (7.1) as
r
| X | = | FixG ( X )| + | OrbG ( xi )| (7.2)
i = s +1
In the special case where G is a pgroup, we obtain the following
useful fact.
Proposition 7.3 Let G be a pgroup of order pk for some prime p and
positive integer k, and suppose that G acts on some finite set X. Then
| X | | FixG ( X )| (mod p).
What this tells us is that the modulop congruence class of the number
of elements in X is determined solely by the single-element orbits; that
is, the points that are fixed by every element of G.
We are now ready to prove the fact we mentioned earlier: that a finite
group G has psubgroups of all possible orders for any prime factor p
7
7
This is actually a slightly stronger of | G |. We will call this Sylows First Theorem.
version of what most books call Sy-
lows First Theorem. Also, there isnt
Theorem 7.4 (Sylows First Theorem) Let G be a finite group of order
a complete consensus on the number- | G | = pk m, where p Z is prime, k Z is non-negative, and m N
ing of Sylows Theorems: some books with p 6 |m. Then G contains a subgroup of order pi for all 0 6 i 6 k, and
combine them into one large theorem
with several parts, while others divide each subgroup of order pi is a normal subgroup of some subgroup of order
them into three or sometimes four sep- pi+1 for 0 6 i < k.
arate theorems; neither is there com-
plete agreement on the order in which Proof We will proceed by induction on i. The case i = 0 is trivial, and
they should be presented. Here we will
state and prove them as three separate the case i = 1 holds by Cauchys Theorem8 . Suppose, then, that H is
but closely-related theorems, in what a psubgroup of G of order | H | = pi for some i > 1. Let S be the set
seems these days to be the most com-
of left cosets of H in G, and define a left action of H on S by
mon order.
8
Theorem 2.37, page 57. h ( gH ) = (hg) H.
By Proposition 7.3 the number of cosets in S is congruent modulo p to
the number that are fixed by the action of H:
|S| | Fix H (S)| (mod p).
For a coset gH to be fixed by this action means that h ( gH ) = (hg) H =
gH for all h H. Proposition 2.26 tells us that this is equivalent to
saying that hg gH for all h H, which is the same as requiring that
g1 hg H for all h H. This, in turn, is equivalent to saying that g
lies in the normaliser NG ( H ) of H in G.
So, the cosets gH fixed by this action are exactly those for which
g NG ( H ), and hence Fix H (S) = { gH : g NG ( H )} we find that
| Fix H (S)| = | NG ( H ):H |.
Looking at this another way, the left cosets of H in G are all the same
size and partition G, so |S| = | G:H |. Putting all this together we get
| G:H | = |S| | Fix H (S)| = | NG ( H ):H | (mod p). (7.3)
9
Definition 4.46, page 122. Recall9 that the normaliser NG ( H ) is the largest subgroup of G in
which H is normal, so H P NG ( H ). Therefore it makes sense to talk
about the quotient group NG ( H )/H.
Also, H is a psubgroup of G, so | H | = pi for some i > 1. So if
i < k the index | G:H | is a multiple of p, and by (7.3) it follows that
| NG ( H ):H | is also a multiple of p. Thus | NG ( H )/H | is a multiple of p
too, and by Cauchys Theorem NG ( H )/H has a subgroup of order p.
All subgroups of NG ( H )/H are of the form K/H for some subgroup
K such that H 6 K 6 NG ( H ). So there exists some subgroup K/H
finite groups 243
We want to know more about these subgroups: how are they related
and how many are there? Sylows other theorems give answers to
these questions, so well look carefully at some small examples.
Example 7.7 The dihedral group D3 has order 6 = 2 3, and hence
by Cauchys Theorem (and also by Sylows First Theorem) it must
have at least one subgroup of order 2 and at least one of order 3. In
fact, we know from earlier discussions that it has three subgroups
M1 = {e, m1 }, M2 = {e, m2 } and M3 = {e, m3 }
of order 2. These represent the reflections in the three axes of the
equilateral triangle. We might ask how these are related, and the
answer is given by Example 3.5: the three reflections m1 , m2 and m3
are all conjugate to each other, and the identity element e forms a
singleton conjugacy class on its own.
Upon further inspection, then, we can reconstruct the subgroup
M2 = {e, m2 } from the subgroup M1 = {e, m1 } by conjugating each
element by the 2
3 rotation r:
e = rer 1 , m2 = rm1 r 1 .
244 a course in abstract algebra
e = r 1 er, m3 = r 1 m1 r.
So, these subgroups are conjugate to each other:
M2 = rM2 r 1 and M3 = r 1 M1 r.
There is only one subgroup of order 3: the rotation subgroup R3 =
{e, r, r2 }. This is a normal subgroup and is hence conjugate to itself.
Lets look at a slightly larger example: the symmetric group S4 .
Example 7.8 The symmetric group S4 has order 24 = 22 3, and
hence by Sylows First Theorem it must have at least one subgroup
each of orders 2, 3, 4 and 8. In fact, it has three subgroups of order 8,
each isomorphic to the dihedral group D4 :
h(1 2 3 4), (1 3)i, h(1 2 4 3), (1 4)i and h(1 3 2 4), (1 2)i.
These are all conjugate to each other; to see this, we only have to
check the two generators for each subgroup:
(3 4)(1 2 3 4)(3 4) = (1 2 4 3), (3 4)(1 3)(3 4) = (1 4),
(2 3)(1 2 3 4)(2 3) = (1 3 2 4), (2 3)(1 3)(2 3) = (1 2).
There should also be at least one subgroup of order 3; in fact there
are four:
h(1 2 3)i, h(1 2 4)i, h(1 3 4)i and h(2 3 4)i.
These are also conjugate to each other:
(3 4)(1 2 3)(3 4) = (1 2 4), (2 3)(1 2 4)(2 3) = (1 3 4),
(1 2)(1 3 4)(1 2) = (2 3 4), (1 4)(2 3 4)(1 4) = (1 2 3).
The pattern we see from the above examples is that for any prime factor
p of | G |, the Sylow psubgroups are conjugate to each other. This is
true in general, a result that we will call Sylows Second Theorem:
Theorem 7.11 (Sylows Second Theorem) Let G be a finite group with
order | G | = pk m for some prime p, non-negative integer k and positive
integer m not divisible by p. If H and K are Sylow psubgroups of G, then
H and K are conjugate. That is, there exists some element g G such that
gHg1 = K.
G |G| p np | G:H | In order to explore this question, well introduce a couple of useful bits
D3 6 2 3 3 of notation. Suppose that G is a finite group, and p is a prime integer
3 1 2 that divides | G |. Let Syl p ( G ) denote the set of Sylow psubgroups
A4 12 2 1 3 contained in G, and let n p = | Syl p ( G )|.
3 4 4
For example, if G = D3 then n2 = 3 and n3 = 1 since D3 has three
Z12 12 2 1 3
3 1 4
Sylow 2subgroups and one Sylow 3subgroup. Table 7.1 lists some
S4 24 2 3 3 examples of small finite groups.
3 4 8 What patterns can we see in these examples? All of them have an odd
SL2 (3) 24 2 1 3 number of Sylow 2subgroups. The first five examples all have either
3 4 8 one or four Sylow 3subgroups, and the last two examples both have
Z10 10 2 1 5
a single Sylow 5subgroup.
5 1 2
D5 10 2 5 5 The key observation here is to count modulo p. Given that crucial
5 1 2 steps in the proofs of Sylows First and Second Theorems required us
Table 7.1: The number of Sylow p
to count cosets modulo p, in retrospect this shouldnt be a colossal
subgroups H contained in a group G surprise. Counting modulo p is apparently a fundamental aspect of
working with psubgroups. Looking at these examples through a
modulop lens, we find that in every case the number n p of Sylow
psubgroups is congruent to 1.
Also, in each case n p is a factor of the order of the group in question.
But more than that, it is a factor of the index of the given Sylow
psubgroup. Both of these are true in general, by a result we will call
Sylows Third Theorem.
Theorem 7.13 (Sylows Third Theorem) Let G be a finite group of
order | G | = pk m where p is prime, k is a non-negative integer, and m is a
positive integer not divisible by p. Then n p 1 (mod p) and n p |m.
Proof As with Sylows first two Theorems, we will study the action
of a particular group on a suitably chosen set, and then apply (7.2).
To prove the first statement, suppose H is a Sylow psubgroup of G
and let it act on Syl p ( G ) by conjugation. For any Sylow psubgroup K
to be fixed under this action, we require hKh1 = K for all h H; the
subgroup H itself is certainly fixed in this way.
Suppose that K is any Sylow psubgroup fixed under conjugation by
H. Then H is contained in the normaliser NG (K ) of K in G; clearly
also K NG (K ), so both H and K are Sylow psubgroups in NG (K ),
and by Sylows Second Theorem they must be conjugate in NG (K ).
But K is normal in, and hence conjgate to itself in, its own normaliser
NG (K ). So K = H and thus the only Sylow psubgroup in Syl p ( G )
fixed by conjugation with H is H itself. Hence | Fix H (Syl p ( G ))| = 1
and by (7.2) we have, as required,
n7 |178, whence n7 = 1, and that n89 1 (mod 89) and n89 |14, so
n89 = 1.
By Corollary 7.12, the subgroup of order 7 and the subgroup of
order 89 must be normal in G, and so no group of order 1246 can be
simple.
Proof Let G act on itself by conjugation. Then the set FixG ( G ) consists
of all elements h G for which ghg1 = h for all g G. This is
equivalent to requiring gh = hg for all g G, hence FixG ( G ) = Z ( G ).
Applying 7.2, we get
and since g and all its powers are in Z ( G ), so they commute with
everything, in particular any power of h.
250 a course in abstract algebra
we take a subgroup lattice for a finite group G, then any path through V
By this definition, both subnormal series (7.4) and (7.5) for D4 are iso-
morphic, and both normal series (7.6) and (7.7) for Z6 are isomorphic
too. We can take this idea a little further. Consider the normal (and
subnormal) series
{0} C 60Z C 12Z C 2Z C Z (7.8)
for Z. There is still room to fit more subgroups in, and we can do this
in different ways. For example,
{0} C 120Z C 60Z C 12Z C 6Z C 2Z C Z (7.9)
and
{0} C 180Z C 60Z C 12Z C 4Z C 2Z C Z (7.10)
are both valid normal series for Z, obtained by fitting extra groups
Otto Schreier (19011929)
into the series (7.8). They arent isomorphic: the quotient groups
of (7.9) are
120Z
= Z, Z2 , Z5 , Z2 , Z3 and Z2 ,
while the quotient groups of (7.10) are
180Z
= Z, Z3 , Z5 , Z3 , Z2 and Z2 .
These series arent isomorphic to each other, and neither are they
isomorphic to the original series (7.8), but this is a useful idea thats
worth studying further, and to that end we introduce the following
definition.
Definition 7.26 Let H = { H0 , . . . , Hm } and {K0 , . . . , Kn } be be two
normal or subnormal series of a group G. Then we say that K is a
refinement of H if we can obtain K by inserting additional subgroups
into H. More precisely, there exists an injective, strictly increasing
function
f : {0, . . . , m} , {0, . . . , n}
such that Hi = K f (i) for 0 6 i 6 m.
These series are clearly not isomorphic, but if were careful we can find
refinements of each that are isomorphic. Each of these refinements
should, at the very least, have quotient groups Z, Z2 , Z3 and Z4 . We
can refine (7.11) to get
{0} C 24Z C 6Z C 3Z C Z
which has quotient groups
24Z
= Z, Z4 , Z2 and Z3 .
Similarly, we can refine (7.12) to get
the various subgroups involved in the proof (see Figure 7.2). Zassen- AK HB
haus formulated and proved his Lemma at the age of 21, while a
graduate student at the University of Hamburg, in order to provide a
A ( AK )( H B) B
neater proof of Schreiers Refinement Theorem and the JordanHlder
Theorem.
A( H B) B( AK )
Theorem 7.29 (Zassenhaus Lemma) Let G be a group, and let H, K,
A and B be subgroups of G such that A P H and B P K. Then H K
(i) A ( H B ) P A ( H K ), A( H K ) B( H K )
(ii) B( A K ) P B( H K ), and
A( H K ) HK B( H K ) H K
(iii) = = .
A( H B) ( A K )( H B) B( A K ) Figure 7.2: Subgroup lattice for Zassen-
haus Lemma
Figure 7.2 shows the subgroup lattice formed by the subgroups in-
volved in the Lemma. For aesthetic reasons this lattice is drawn upside
down relative to our usual convention: larger subgroups are towards
the bottom of the diagram, while smaller subgroups are towards the
top. Bold lines indicate that the upper subgroup is normal in the lower
one. In general, a subgroup at the apex of two upward lines is the
intersection of the subgroups at the other ends. Similarly, two down-
ward lines meet at a subgroup that is the product of the subgroups at
their other ends.
Proof To prove part (i), we must show first that A( H B) is a subgroup
of A( H K ), and then that it is closed under conjugation by elements
of A( H K ). By Proposition 2.11, A( H B) 6 A( H K ) if and only if
A( H B) = ( H B) A. But we know that A P H, so Ah = hA for any
element h H, and hence this also follows for any h H B H.
Therefore A( H B) = ( H B) A and thus A( H B) 6 A( H K ).
Now we must show that A( H B) is normal in A( H K ), which we
do by showing that it is closed under conjugation by any element
of A( H K ). Let x = ab and y = ck where a, c A, b H B and
k H K, so that x A( H B) and y A( H K ). We want to show
that yxy1 A( H B).
First, observe that yay1 A since y A( H K ) AH H,
and A P H. Next, kbk1 H B since k H K and H B P
H K. Hence yby1 = (ck)b(ck)1 = ckbk1 c1 A( H B) A, and
A( H B) A = A( H B) since A P H. Finally, yxy1 = y( ab)y1 =
(yay1 )(yby1 ) A( H B), and therefore A( H B) P A( H K ) as
claimed.
Part (ii) follows by a very similar argument.
To prove part (iii), we use Dedekinds Modular Law14 and the Second 14
Lemma 7.28, page 254.
Isomorphism Theorem.15 Recall that the latter says that if a group G 15
Theorem 4.62, page 128.
contains a subgroup M and a normal subgroup N, then MN/N =
256 a course in abstract algebra
( H K ) A( H B) = ( H K A)( H B) = ( AK )( H B),
and hence (7.13) becomes
A( H K ) HK
= . (7.14)
A( H B) ( AK )( H B)
By a very similar argument we find also that
B( H K ) HK
= . (7.15)
B( A K ) ( AK )( H B)
Putting (7.14) and (7.15) together yields the required isomorphism.
We are now ready to prove Schreiers Refinement Theorem. The
key idea of this proof is to insert between each group Hi and Hi+1
in the series H a chain of groups of the form Hi ( Hi+1 K j ), and
then to perform the analogous construction with the series K. We
then apply Zassenhaus Lemma to show the existence of a bijective
correspondence between the relevant quotient groups.
Proof of Schreiers Refinement Theorem We first prove the theorem
for subnormal series. Suppose that H and K are two subnormal series
for a group G, with
{e} = H0 C H1 C C Hm = G
and {e} = K0 C K1 C C Kn = G.
Now consider the chain
Proof Let
{e} = H0 C H1 C C Hm = G
and { e } = K0 C K1 C C K n = G
be two composition (or principal) series for G. Then by Schreiers
Refinement Theorem both of these series have isomorphic refinements.
But since all the composition factors are already simple, each group
Hi is maximal normal in Hi+1 and each group K j is maximal normal
in K j+1 for 0 6 i < m and 0 6 j < n, and hence neither series
can be refined any further. Therefore both series must already be
isomorphic.
Also in Example 7.30 we found another composition series for D4 ,
namely
{e} C {e, r2 } C {e, r, r2 , r3 } C D4 .
This series contains the rotation subgroup R4 = {e, r, r2 , r4 }, which is
known to be normal in D4 . In fact, if a group G has a composition
series at all, then we can find one that contains any given normal
subgroup N C G:
finite groups 261
{e} = H0 C H1 C C Hn = G
is a composition (or principal) series for G. The series
{e} C N C G
is both subnormal and normal. By Schreiers Refinement Theorem
there is a refinement of this series that is isomorphic to the given
composition (or principal) series, and hence itself a composition (or
principal) series for G. This refinement will necessarily contain N.
7.3 Soluble and nilpotent groups The world is devoted to physical sci-
ence, because it believes these discover-
ies will increase its capacity of luxury
Particularly important is the case where the composition factors and self-indulgence. But the pursuit
are abelian. This will become especially relevant in Chapter 11 when of science only leads to the insoluble.
When we arrive at that barren term, the
we study the solubility by radicals of polynomial equations. The only Divine voice summons man, as it sum-
abelian finite simple groups are the cyclic groups Z p where p is prime, moned Samuel;
so this condition is equivalent to requiring the composition factors to Benjamin Disraeli, 1st Earl of
Beaconsfield (18041881),
be finite cyclic groups of prime order. Lothair (1870) 70
Definition 7.36 A group G is soluble or solvable if it has a composi-
tion series whose composition factors are all abelian, or equivalently
are cyclic groups Z p of prime order.
{ } C A 5 C S5 ,
which has composition factors A5 and Z2 . The alternating group A5 is
simple but not abelian, so S5 is not soluble. An important consequence
of this is the fact, originally discovered independently in the early
262 a course in abstract algebra
Proof Since N and G/N are soluble, there exist composition series
{e} = H0 C H1 C C Hm = N
and
{e} = K0 C K1 C C Kn = G/N.
In the latter series, each group Ki can be written as Gi /N for some
subgroup Gi , and N corresponds to the identity element in the quotient
group G/N, so this can be written as
N C G1 /N C C Gn /N = G/N.
Each of the composition factors of both of these series are cyclic
groups of prime order, and by the Third Isomorphism Theorem we
have ( Gi+1 /N )/( Gi /N )
= Gi+1 /Gi , which must also be of prime
order. We can therefore construct a composition series
{e} = H0 C H1 C Hm = N C G1 C C Gn = G
for G, all of whose composition factors are cyclic groups of prime
order, and hence abelian. Thus G is soluble.
An important corollary of this last proposition is that direct products
of soluble groups are themselves soluble:
Corollary 7.41 If H and K are soluble groups, then so is H K.
{0} C Z p
with composition factor isomorphic to Z p itself.
Otherwise, we proceed by induction on n = | G |, which we assume
to be composite. Suppose that all abelian groups of order less than n
have already been shown to be soluble. If p is a prime factor of | G |,
17
Theorem 2.37, page 57. then by Cauchys Theorem17 G must have a cyclic subgroup H = Zp.
This subgroup H is normal in G, and its quotient G/H is abelian with
order | G/H | = n/p < n. Therefore both H and G/H are soluble, and
by Proposition 7.40, so is G.
having shown that G (k) is soluble, and since G (k1) /G (k) is abelian
and therefore soluble, Proposition 7.40 implies that G (k1) must also
be soluble. Therefore, by induction, G = G (0) is soluble.
Conversely, suppose G is soluble. Then there is a composition series
{e} = Gm C Gm1 C C G0 = G
for G, with each composition factor Gi /Gi+1 an abelian simple group,
for 0 6 i 6 m. This composition series will be at least as long as the
derived series, and we claim that G (i) 6 Gi for 0 6 i 6 m.
Meanwhile, the derived groups eventually stabilise, in the sense that
there exists a positive integer n such that G (i+1) = G (i) for all i > n.
We proceed by induction on i: for i = 0 we have G (0) = G = G0 ,
and since G0 /G1 is abelian, G (1) = [ G, G ] 6 G1 by Proposition 3.31.
Suppose that G (i) 6 Gi for some i. The quotient Gi /Gi+1 is abelian, so
[ Gi , Gi ] 6 Gi+1 by Proposition 3.31, and hence
G (i+1) = [ G (i) , G (i) ] 6 [ Gi , Gi ] 6 Gi+1
as claimed. Therefore G (n) 6 Gn = {e} and thus the derived series
terminates at the trivial subgroup.
This gives another way of testing for solubility: construct the de-
rived series and see if it eventually terminates at the trivial subgroup.
Moreover, the derived series of a group is the shortest possible subnor-
mal series with abelian quotients, and its length thus gives us some
potentially useful information about the complexity of the group:
Definition 7.48 Let G be a soluble group. Then the length of the
derived series for G is the derived length of the group.
The trivial group {e} is the only group with derived length 0. The
groups of derived length 1 are exactly the nontrivial abelian groups:
if A is abelian and soluble, then A(0) = A, while A(1) = [ A, A] = {e};
conversely if A(1) = [ A, A] = {e} then this means that every element
of A commutes with every other element of A, and therefore A is
abelian. The groups of derived length 2 are precisely those with
nontrivial abelian commutator groups G (1) = [ G, G ]; such groups are
called metabelian, and we will briefly return to them later.
Another important normal subgroup we met in Chapter 3 is the centre
Z ( G ) of a group G. We can use it to construct normal or subnormal
series but in a slightly more complicated way than we did with the
commutator subgroup [ G, G ]. The obvious method, where each group
is the centre of its predecessor, doesnt work, or at least not in a very
interesting way: Z ( G ) is abelian, so its centre is the same; that is,
Z ( Z ( G )) = Z ( G ). Hence any such series will stabilise after at most
one step.
finite groups 267
the case that Gi+1 /Gi is a subgroup of G/Gi , but the requirement that
Gi+1 /Gi 6 Z ( G/Gi ) is stronger, and will not always hold. A central
series will always have abelian quotients Gi+1 /Gi but not every series
with abelian quotients need be central.
There are two ways to form a central series from a group G: start
at {e} and work upwards to construct an ascending series, or start
at G itself and recursively form a descending series. These methods
need not result in the same series; also the ascending series need not
terminate at G, and the descending series might not reach {e}.
Well try both of these approaches in turn. First well build an ascend-
ing series starting at G0 = {e}. We want the second group G1 to have
the property that G1 /G0 6 Z ( G/G0 ), and one way of ensuring this is
to choose G1 such that G1 /G0 = Z ( G/G0 ). That is, we want G1 to be
the subgroup of G whose quotient by G0 is exactly the centre of G/G0 ,
and we can achieve this by setting G1 = Z ( G ).
Next we want G2 to be a normal subgroup of G such that G2 /G1 6
Z ( G/G1 ) and again well choose G2 so that G2 /G1 = Z ( G/G1 ). To
do this, we need to set G2 to be the subgroup of G whose quotient by
G1 is exactly the centre of the quotient G/G1 . Continuing this process
we obtain an ascending central series of groups:
Definition 7.51 For a group G, form an ascending normal series
G0 = {e} C G1 C C Gn
such that G1 = Z ( G ) and Gi is the higher centre Zi ( G ) of G defined
such that Z0 ( G ) = {e}, and
Zi ( G )/Zi1 ( G ) = Z ( G/Zi1 ( G ))
for 0 < i 6 n. This series is the ascending central series or upper
central series of G.
As an example, well do this for the dihedral group D4 :
Example 7.52 Let G = D4 . Then G0 = Z0 ( G ) = {e} and G1 =
Z1 ( G ) = Z ( G ) = {e, r2 }. Next we want G2 = Z2 ( G ) to be the
subgroup of G such that G2 /G1 = Z ( G/G1 ). Now G/G1 consists of
the cosets
eG1 = r2 G1 = {e, r2 }, rG1 = r3 G1 = {r, r3 },
m1 G1 = m3 G1 = {m1 , m3 }, m2 G1 = m4 G1 = {m2 , m4 }.
This quotient group is of order 4 and hence abelian, so Z ( G/G1 ) =
G/G1 , and hence G2 = Z2 ( G ) must be G itself. Therefore the upper
central series of D4 is
{e} C {e, r2 } C D4 .
finite groups 269
A 3 C S3 .
The upper and lower central series for D4 happen to be the same, but
those for S3 arent even the same length.
The case where a group G has a finite-length central series which
reaches {e} at one end, and G at the other, is particularly interesting:
Definition 7.55 Suppose that a group G has a central series
{e} = G0 < G1 < < Gn = G
connecting G with its trivial subgroup {e}. Then we say that G is
nilpotent. The smallest possible length n over all such central series
is called the nilpotency class of G.
The lower central series is in some sense the optimal descending
central series: if it reaches {e} then it does so faster than any other
descending central series.
Proposition 7.56 Suppose that a group G has a finite descending central
series
{e} = Gn+1 < Gn < < G1 = G.
Then i ( G ) 6 Gi for 0 < i 6 n+1. Furthermore, if G has nilpotency
class c, then n > c+1.
To prove this, we first need the following lemma.
Lemma 7.57 Let G be a group, and suppose that K P G and K P H 6 G.
Then H/K 6 Z ( G/K ) if and only if [ G, H ] 6 K.
Proof Suppose that H/K Z ( G/K ). This means that every coset in
H/K commutes with every coset in G/K. That is, for every h H and
g G we have (hK )( gK ) = ( gK )(hK ). Moreover, (hK )( gK ) = (hg)K
and ( gK )(hK ) = ( gh)K, so
Proposition 7.63 If G and H are nilpotent groups, then their direct prod-
uct G H is nilpotent as well.
i+1 ( G H ) = [i ( G H ), G H ] 6 [i ( G )i ( H ), G H ]
6 [i ( G ), G ][i ( H ), H ] = i+1 ( G )i+1 ( H ),
and so the induction holds for all i > 1. Since G and H are nilpotent,
there exists m, n N such that m ( G ) = n ( H ) = {e}. Let k =
max(m, n). Then
k ( G H ) 6 k ( G )k ( H ) = {e}{e} = {(e, e)}.
Hence the lower central series for G H terminates at the trivial sub-
group {(e, e)} after finitely many steps, so G H is nilpotent.
We can extend this by induction to show that direct products of finitely
many nilpotent groups are nilpotent:
Corollary 7.64 Let G1 , . . . , Gn be nilpotent. Then the direct product
G1 Gn is nilpotent too.
Finally, we show that images of nilpotent groups are nilpotent, and
consequently so are their quotients.
Proposition 7.65 Let f : G H be a homomorphism from a nilpotent
group G to some group H. Then the image f ( G ) is nilpotent too.
Now, since G is nilpotent, there exists some positive integer n such that
n ( G ) = {e}, and therefore n ( f ( G )) = f (n ( G )) = f ({e}) = {e},
which means that f ( G ) is nilpotent.
Corollary 7.66 If G is a nilpotent group, and N P G, then the quotient
G/N is also nilpotent.
In the middle of this proof we also derived the following useful fact:
Corollary 7.72 A finite group G is nilpotent if and only if all of its Sylow
subgroups are normal.
We will now briefly look at a third class of groups defined in terms
of subgroup series. By Definition 7.49, a group G is nilpotent if
there exists a finite-length central series stretching between {e} and G.
Recall that for a series to be central, we require Gi+1 /Gi 6 Z ( G/Gi ),
and also that the series is normal; that is, Gi P G for all i.
More generally, a subnormal series is abelian if each quotient Gi+1 /Gi
is abelian (but not necessarily contained in the centre of G/Gi ); a
group has a finite-length abelian series if and only if it is soluble.
Between these two, we can consider the case where G has a finite-
length normal series where the quotients Gi+1 /Gi are cyclic.
finite groups 277
If you have built castles in the air, your 7.4 Semidirect products
work need not be lost; that is where
they should be. Now put the founda-
tions under them. The internal direct product enables us to decompose a group as
Henry Thoreau (18171862), the direct product of two normal subgroups. More precisely, if H and
Walden (1854) 346
K are both normal subgroups of a group G with trivial intersection
H K = {e}, and if HK = G, then G is isomorphic to the direct product
finite groups 279
k1 (k2 h) = k1 (k2 hk 1 1 1
2 ) = k 1 k 2 hk 2 k 1 = ( k 1 k 2 ) h ( k 1 k 2 )
1
= (k1 k2 )h,
Example 7.85 Let Rn denote the additive group of (Rn , +); that is,
the set of ordered ntuples of real numbers, or ncomponent real
vectors. The group GLn (R) acts on Rn by matrix multiplication, and
hence we can form the semidirect product Rn oGLn (R).
The underlying set of this group is Rn GLn (R): the set of ordered
pairs (u, A), with u a vector in Rn and A a nonsingular nn matrix
in GLn (R). The product of two elements (u, A) and (v, B) is
(u, A) (v, B) = (u + Av, AB).
This is isomorphic to the affine general linear group AGLn (R) of
transformations w 7 Aw + u for A GLn (R) and u, w Rn .
For example, suppose that u, v, w Rn and A, B GLn (R). Then
(v, B) acts on w by w 7 Bw + v and (u, A) acts on w by w 7
Aw + u. By composition, we find that
((u, A)(v, B))w = (u, A)( Bw+v) = ABw+ Av+u = ( Av+u, AB)w
and hence in AGLn (R) the multiplication operation is given by
(u, A)(v, B) = (u + Av, AB)
which is exactly that given by the semidirect product construction.
i1
E1
q1
A i2
f
q2
G
E1
if and only if
(eG , g) = ( g, eG ) = 0
finite groups 287
Proof The first of these follows from the fact that s(eG ) = eE . For,
s(eG g) = s( g) = eE s( g) = s(eG )s( g)
and s( geG ) = s( g) = s( g)eE = s( g)s(eG ).
The second follows from the associativity condition: for a1 , a2 , a3 A
and g1 , g2 , g3 G we require that
(( a1 , g1 ) ( a2 , g2 )) ( a3 , g3 ) = ( a1 , g1 ) (( a2 , g2 ) ( a3 , g3 )).
The left hand side yields
Now, given a group G and Gmodule A, we can form the set FS( G, A)
of factor sets : G G A, and the subset IFS( G, A) of inner factor
sets. More interestingly, these sets are closed under a fairly straight-
forward addition operation and thereby form abelian groups:
Proposition 7.96 Let G be a group and A a Gmodule. Suppose that
1 , 2 FS( G, A) are factor sets. Then the function 1 +2 given by
(1 +2 )( g, h) = 1 ( g, h) + 2 ( g, h) is also a factor set. Furthermore, if
1 , 2 IFS( G, A) then 1 +2 IFS( G, A) too.
Proof To see that FS( G, A) and IFS( G, A) are closed under this point-
wise addition operation, we must verify that 1 +2 satisfies the con-
ditions in Proposition 7.92, and that 1 +2 satisfies Definition 7.95.
Both of these tasks are entirely routine, and well omit the details here,
although the interested reader is encouraged to check them.
Lets summarise where we are at the moment. For any extension
i q
A E G
Proof Suppose that the extensions are equivalent; that is, there exists
an isomorphism : E1 E2 such that i1 = i2 and q1 = q2 .
Choose a section s1 : G E1 that realises the Gaction on A. Then s1
finite groups 291
i2 ( g a) = (i1 ( g a)) =
(s1 ( g))(i1 ( a))(s1 ( g))1 = s2 ( g)i2 ( a)s2 ( g)1
and so s2 does indeed realise the correct Gaction.
The factor set is determined by the section s1 as follows:
i1 ( ( g, h)) = s1 ( g)s1 (h)s1 ( gh)1
for all g, h G. Applying to both sides, we get
i2 ( ( g, h)) = s2 ( g)s2 (h)s2 ( gh)1
so the factor set determined by the section s2 is the same as that
determined by s1 . If is some other factor set for E2 determined by a
different section t : G E2 , then by (7.24) and are cohomologous.
Conversely, suppose that E1 and E2 are extensions of G by A relative
to equivalent factor sets and respectively. Then there exists a
function : G A such that
( g, h) ( g, h) = ( g) + g (h) ( gh)
for all g, h G.
Let s1 : G E1 and s2 : G E2 be sections that realise the Gmodule
structure on A and yield the factor sets and respectively. Then
every element in E1 can be written uniquely as i1 ( a)s1 ( g) for some
a A and g G, and every element in E2 can be similarly written in
the form i2 ( a)s2 ( g). Recall that the multiplication operation in E1 is
given by
i1 ( a)s1 ( g) i1 (b)s1 (h) = i1 ( a + g b + ( g, h))s1 ( gh)
and that in E2 is given by
i2 ( a)s2 ( g) i2 (b)s2 (h) = i2 ( a + g b + ( g, h))s2 ( gh).
Now define : E1 E2 by (i1 ( a)s2 ( g)) = i2 ( a + ( g))s2 ( g). This is
clearly a bijection. Furthermore
(i1 ( a)s1 ( g) i1 (b)s1 (h)) = (i1 ( a + g b + ( g, h))s1 ( gh))
292 a course in abstract algebra
= i2 ( a + g b + ( g, h) + ( gh))
and
Proof Suppose that is a factor set in FS( G, A); recall that this is a
function : G G A, and hence its values lie in A. By Proposi-
tion 2.34 these values must all have orders that divide | A| = n, and
so n ( g, h) = 0 for all g, h G. Therefore n = 0 in FS( G, A), and n
also lies in the subgroup IFS( G, A) of inner factor sets.
Now we want to show that m is also an inner factor set. To do this,
we define a function : G A by
( g) = ( g, h).
h G
In particular,
(eG ) = (eG , h) = 0 = 0.
h G h G
finite groups 293
e a b c
e 0 0 0 0
a 0 1 0 1
b 0 1 1 0
c 0 0 1 1
Another interesting case is the one where the group G is cyclic. These
are called cyclic extensions, and the following discussion will be
useful when we classify groups of order 16 in Section 7.A.
If we have a group G with a normal subgroup H P G, such that
G/H = Zn then G is obviously a cyclic extension of Zn by some
(possibly nonabelian) group H. Choose some element g G \ H such
that the coset gH generates the quotient G/H. Let v = gn . Then
vH = gn H = ( gH )n = H in G/H, and so v H. As a normal
subgroup, H is closed under conjugation by any element of G, so let
Aut( H ) be the inner automorphism : h 7 ghg1 . Furthermore,
(v) = gvg1 = ggn g1 = gn = v
so fixes this distinguished element v. Also, for any h H,
n (h) = gn hgn = vhv1
so n is conjugation by v.
We can now discard G and g, and just keep v. Well now see that the
data ( H, n, , v) is all we need to reconstruct and uniquely determine
(up to equivalence) the extension
i q
H G Zn .
Every cyclic extension type determines some group G, and all cyclic
extensions arise in this way:
Theorem 7.101 (Cyclic Extension Theorem) Given a cyclic extension
type ( H, n, , v) there is some group G with H P G and G/H = Zn .
Furthermore, all extensions of Zn by H are determined by some cyclic
extension type.
Well use cyclic extension types in the next section, when we classify
groups of order 16. Finally, well mention a couple of classes of groups
that can be described in terms of extensions.
finite groups 297
Metacyclic groups are metabelian, and include the cyclic groups them-
selves, the dicyclic groups Dicn and the dihedral groups Dn . As
remarked earlier, metabelian groups are exactly those with derived
length at most two.
7.A Classification of small finite groups I tried to make out the names of plants,
and collected all sorts of things, shells,
seals, franks, coins and minerals. The
In this section we will classify, up to isomorphism, groups passion for collecting, which leads a
of order less than 32. The reason well stop at order 31 is that the man to be a systematic naturalist, a
virtuoso, or a miser, was very strong
classification is relatively manageable up to that point. The classi- in me, and was clearly innate, as none
fication of groups of order n = 2k is, in general, quite complicated of my sisters or brother ever had this
taste.
and involves the consideration of a number of cases. As can be seen
Charles Darwin (18091882),
from Table 7.3, the only really involved cases are n = 16 = 24 , with 14 in: Francis Darwin, The Life and Letters
non-isomorphic groups, and n = 24 = 323 , with 15. Everything of Charles Darwin (1887) I 2728
else is comparatively straightforward; however there are 51 groups of
order 32, which is why well stop just before then. n total abelian nonabelian
Before we start classifying, well introduce a few results that will prove 1 1 1 0
2 1 1 0
useful. The first relates to groups that decompose as a semidirect 3 1 1 0
product of two cyclic groups. 4 2 2 0
5 1 1 0
Proposition 7.107 Suppose that a finite group G decomposes as an in- 6 2 1 1
ternal semidirect product HK, where H = h x : x m = 1i
= Zm is normal 7 1 1 0
8 5 3 2
in G, and K = hy : y = 1i = Zn . Then the possible actions of K on H
n
9 2 2 0
are determined by yxy1 = xi where 0 < i < m and in 1 (mod m). 10 2 1 1
11 1 1 0
Proof The action of K on H is determined entirely by the behaviour 12 5 2 3
13 1 1 0
of the generators x and y, specifically by the value of yxy1 H. 14 2 1 1
Since H is cyclic, this must be a nontrivial power of the generator x, 15 1 1 0
16 14 5 9
so yxy1 = xi with 0 < i < m. Furthermore, since K is a finite cyclic 17 1 1 0
n
group of order n, we have yn = 1, so x = yn xyn = xi and therefore 18 5 2 3
in 1 (mod n). 19 1 1 0
20 5 2 3
21 2 1 1
The next result, which we will use particularly in the classification of
22 2 1 1
groups of order 16 and 27, relates to factoring a group by its centre. 23 1 1 0
24 15 3 12
25 2 2 0
26 2 1 1
27 5 3 2
298 a course in abstract algebra
There is a single group with one element: the trivial group {e}.
(i) the cyclic group Z8 , The groups of order 8 were classified in Proposition 2.40, and there
(ii) the abelian groups Z2 Z4
and
are five of them, listed in Table 7.6. Of these, (i)(iii) are abelian, while
(iii) Z2 Z2 Z2 , (iv) and (v) are nonabelian. They are all soluble by Proposition 7.43.
(iv) the dihedral group D4 and
(v) the quaternion group Q8 ,
which is isomorphic to the
dicyclic group Dic2 .
finite groups 299
Groups of order 12
This is the first new case that cant be immediately resolved by existing
classification results. Groups of order 12 that weve met already
include the cyclic group Z12 = Z3 Z4 , the abelian group Z2 Z6 =
Z2 Z2 Z3 , the dihedral group D6 , the alternating group A4 and the
dicyclic group Dic3 . We now prove that these five are the only ones.
Let G be a group of order | G | = 12. First we apply Sylows Third
Theorem32 to calculate the possible numbers n p of Sylow psubgroups. 32
Theorem 7.13, page 246.
Thus n2 1 (mod 2) and n2 |3, which means that either n2 = 1 or 3.
Similarly, n3 1 (mod 3) and n3 |4, so either n3 = 1 or 4.
Case 1 (n3 = 1) Let H be the (unique) Sylow 3subgroup; this must
be isomorphic to Z3 . Furthermore, H C G by Corollary 7.12. Now
let K be a Sylow 2subgroup: this has order 4 and must therefore be
isomorphic to either Z4 or Z2 .
By Proposition 2.34, the order of any element of H K must divide
both | H | = 3 and |K | = 4, which are coprime, so H K = {e}. Then
G = HK must be a semidirect product of H and K. Let H = h x : x3 =
1i and consider the two possibilities for K separately:
Case 1a (K = Z4 ) Let K = hy : y4 = 1i. By Proposition 7.107 the
Kaction is determined by the value of yxy1 = xi , where 0 < i < 3
and i4 1 (mod 3). There are two ways this can go: either yxy1 = x
or yxy1 = x2 = x 1 .
The first of these gives the presentation
G = h x, y : x3 = y4 = 1, xy = yx i
= Z4 Z3
= Z12 .
The other possibility is that yxy1 = x2 , which gives the presentation
G = Z3 oZ4 = h x, y : x3 = y4 = 1, x2 y = yx i.
We claim that this group is isomorphic to the dicyclic group
Dic3 = h a, b : a6 = e, a3 = b2 , ab = ba1 i.
To see this, we note first that
y2 x = y(yx ) = y( x2 y) = (yx ) xy = x2 (yx )y = x4 y2 = xy2
and then that (y2 x )2 = (y2 x )y2 x = xy4 x = x2 . Since x2 has order 3,
it follows that y2 x has order 6. Now let a = y2 x and b = y. Then
b2 a = y4 x = x, so a and b generate all of the group G = Z3 oZ4 .
Furthermore,
a3 = ( y2 x )3 = x 2 y2 x = x 3 y2 = y2 = b2
and aba = y2 xy3 x = xy5 x = xyx = x3 y = y = b.
So G
= Dic3 .
300 a course in abstract algebra
Groups of order 15
Groups of order 16
This is the first of the two complicated cases we will consider (the
other being 24). The following exposition is based heavily on the very
readable article by Marcel Wild,34 and makes use of the theory of 34 M Wild, The groups of order sixteen
cyclic extensions developed in the previous section. made easy, The American Mathematical
Monthly 112 (2005) 2031.
First we prove the following result:
Proposition 7.109 Let G be a group of order 16. If G is not isomorphic
to the direct sum Z2 Z2 Z2 Z2 , then G contains a normal subgroup
isomorphic either to Z8 or Z2 Z4 .
so | x2 g| = 2 which contradicts the hypothesis that g has minimal order This is one of a family of groups SD2n
in G \ H. If = 2 then xg has order 2 and if = 3 then x2 g again or QD2n of the form
n 1 n 2
has order 2. So the only new cyclic extension type is (Z8 , 2, 4 , x2 ), h x, y : x2 =y2 =1, yxy1 = x2 1 i
which is realised by the dicyclic group Dic4 . for n > 2. The case n = 4 is the
first nonabelian group in the sequence,
Case 3 (| g| = 8) In this case there are two possible choices for v, and the first that we havent met in
namely x2 and x6 . Both of these elements are fixed by 1 and 3 , but other guises, since SD4
= Z2 Z2 and
SD8
= Z4 Z2 .
not 2 or 4 , and so we have four potentially new cyclic extension
types
(Z8 , 2, 1 , x2 ), (Z8 , 2, 1 , x6 ), (Z8 , 2, 3 , x2 ) and (Z8 , 2, 3 , x6 ).
The automorphism 4 maps x2 to x6 and commutes with both 1 and
3 , so the first and second of the above types are equivalent, as are the
third and fourth.
Considering (Z8 , 2, 1 , x2 ) we note that ( x3 g)( x3 g) = x3 1 ( x3 ) g2 =
x8 = e. Therefore | x3 g| = 2 and so this extension type doesnt yield a
group we havent already seen. We can discard (Z8 , 2, 3 , x2 ) as well,
because ( xg)( xg) = x3 ( x ) g2 = x8 = e and so | xg| = 2. ( x ) (y)
1 = id x y
Case 4 (| g| = 16) If g has order 16 then the cyclic subgroup h gi must 2 x3 y x2 y
x3
be the whole of G, so G = Z16 . This group realises, amongst others, 3 y
4 xy x2 y
the cyclic extension type (Z8 , 2, 1 , x ). 5 xy y
6 x3 x2 y
We have now found six groups of order 16 that have a normal subgroup 7 x3 y y
isomorphic to Z8 . Next we have to study the cyclic extension types 8 x x2 y
of the form (Z2 Z4 , 2, , v). The automorphism group of Z2 Z4 = Table 7.9: Automorphisms of Z2 Z4
h x, y : x4 = y2 = 1, xy = yx i is isomorphic to the dihedral group
D8 . Table 7.9 lists the eight possible automorphisms and their actions
on the generators x and y. With Proposition 7.104 in mind, we note
that 5 and 7 are conjugate in Aut( H ) and so are 6 and 8 . We can
therefore discard 7 and 8 , as they will give rise to groups isomorphic
to those obtained from, respectively, 5 and 6 .
Also, since H is abelian, conjugation by any element of H is the identity
map, so we can also rule out 2 and 4 , neither of which satisfy the
requirement 2 = id.
We assume now that G has no elements of order 8 or higher, otherwise
it would have a normal subgroup isomorphic to Z8 and have been
covered already in one of the previous four cases. Let H = h x, y :
x4 = y2 = 1, xy = yx i = Z4 Z2 and suppose that g G \ H is a
non-identity element of minimal order.
Case 5 (| g| = 2) In this case v = g2 = e, which is fixed by any of the
remaining four automorphisms under consideration. We thus obtain
304 a course in abstract algebra
(Z2 Z4 , 2, 1 , e)
= Z2 Z2 Z4 , (Z2 Z4 , 2, 3 , e)
= D4 Z2 ,
(Z2 Z4 , 2, 5 , e)
= V4 oZ4 , (Z2 Z4 , 2, 6 , e)
= Q8 oZ2 .
(Z2 Z4 , 2, 3 , x2 )
= Q8 Z2 , (Z2 Z4 , 2, 5 , x2 )
= Z4 oZ4 .
(Z2 Z4 , 2, 1 , y)
= Z4 Z4 .
Groups of order 18
G = h x : x3 = 1i hy, z : y3 = z2 = 1, zyz1 = y1 i
= Z3 D3 .
Case 2c (zxz1 = x 1 and zyz1 = y1 ) This gives the group with
presentation
G = h x, y, z : x3 = y3 = z2 = 1, zxz1 = x 1 , zyz1 = y1 i.
Groups of order 20
G = h x, y : x5 = y4 = 1, yx = xyi
= Z5 Z4
= Z20 .
Case 1b (i = 2) In this case, yxy1 = x2 , giving the presentation
G = h x, y : x5 = y4 = 1, yx = x2 yi.
finite groups 307
This is isomorphic to the affine general linear group AGL1 (5) of affine
linear transformations in the finite field F5 .
Case 1c (i = 4) This time, yxy1 = x4 = x 1 , which gives us the
presentation
G = h x, y : x5 = y4 = 1, yxy1 = x 1 i.
Groups of order 21
G = h x, y : x7 = y3 = 1, xy = yx i
= Z7 Z3
= Z21 .
Case 2 (i = 2) Here,
Groups of order 24
and one for K, so this case yields five groups for our list:
Z8 Z3
= Z24 , D4 Z3 ,
Z2 Z4 Z3
= Z2 Z12 , Q8 Z3 ,
Z2 Z2 Z2 Z3
= Z2 Z2 Z6 .
Case 2 (n2 = 1, n3 = 4) Here we have a normal Sylow 2subgroup
H of order 8 and four conjugate Sylow 3subgroups. Let K be one of
these subgroups. Then H K = {e} and we have a semidirect product
G = HK. We have five choices for H and must consider the possible
Z3 actions. The trivial action yields the five direct products already
listed in Case 1, so at this point were only interested in nontrivial
Z3 actions.
Case 2a (H = Z8 ) Let H = h x : x8 = 1i and K = hw : w3 = 1i. By
Proposition 7.107 we have wxw1 = xi with 0 < i < 8 and i3 1
(mod 8). The only value of i satisfying these criteria is i = 1, which
yields the trivial action.
Case 2b (H = Z2 Z4 ) We want a nontrivial homomorphism : K
Aut( H ) mapping the generator of K to a nontrivial element of Aut( H ).
The generator of K has order 3, so its image in Aut( H ) must also
have order 3; however by Proposition 2.34 the order of any element
of Aut( H ) must divide | Aut( H )| = 8, hence no such homomorphism
can exist. This case also only gives the trivial action and one of the
direct products weve already seen.
Case 2c (H = Z2 Z2 Z2 ) Again, we want a nontrivial homomor-
phism : K Aut( H ). This time | Aut( H )| = 168 (it happens to
be isomorphic to the groups GL3 (2) and PSL2 (7)) which is divisible
by 3. By Cauchys Theorem, it therefore has at least one subgroup
isomorphic to Z3 , whose generator has order 3. In fact, all such ele-
ments are conjugate to each other, so we have only a single nontrivial
action at our disposal: the map that cyclically permutes the factors of
Z2 Z2 Z2 . This action yields a new group, a nontrivial semidirect
product G = (Z2 Z2 Z2 )oZ3 , which is isomorphic to the direct
product Z2 A4 .
Case 2d (H = D4 ) There is no nontrivial Z3 action on D4 since
| Aut( D4 )| = 8, so we dont get any new groups in this case.
Case 2e (H = Q8 ) In this case Aut( Q8 ) =
S4 and thus | Aut( Q8 )| =
24, which is divisible by 3, so we can define a nontrivial Z3 action
on Q8 . All of the 3element subgroups of S4 are conjugate to each
other, so we have only one nontrivial action, which yields a new
group for our list: the nontrivial semidirect product G = Q8 oZ3 .
This is isomorphic to the group SL2 (3) of 22 matrices over F3 with
determinant 1.
310 a course in abstract algebra
Groups of order 27
Groups of order 28
h x, y, z : x7 = y2 = z2 = 1, xy = yx 1 , xz = zx, yz = zyi
(i) The cyclic group
Z28 = Z4 Z7 ,
= D7 Z2
= D14 .
(ii) the abelian group
Z2 Z2 Z7 = Z2 Z14 , If both y and z map to the inversion automorphism then we can
(iii) the dicyclic group Dic7 , and replace either y or z with yz, which then maps to the identity, and this
(iv) the dihedral group D14 .
reduces to the previous case.
Table 7.16: Groups of order 28 Hence there are four groups of order 28, listed in Table 7.16.
Groups of order 30
G = h x, y : x15 = y2 = 1, yxy1 = x 1 i
= D15 . Table 7.17: Groups of order 30
There are thus four different groups of order 30, listed in Table 7.17. order groups
This completes our classification of groups of order less than 32. We 20 = 1 1
stop here because things get considerably more complicated with 21 = 2 1
22 = 4 2
groups of order 2n , and in particular there are 51 groups of order 32. 23 = 8 5
Table 7.18 lists the number of groups of order 2n . To put this in 24 = 16 14
25 = 32 51
perspective, up to isomorphism there are 49 910 529 484 groups with
26 = 64 267
order at most 2000, and 49 487 365 422 of those (just over 99.15%) have 27 = 128 2 328
order 1024. 28 = 256 56 092
29 = 512 10 494 213
210 = 1024 49 487 365 422
Summary
41
Definition 7.2, page 240. finite) group G is a psubgroup if it is itself a pgroup.41
Suppose that G is a pgroup acting on some finite set X. Then the
number of elements in X is equal to the number of elements fixed by
42
Proposition 7.3, page 241. the Gaction, modulo p. That is, | X | | FixG ( X )| (mod p).42
43
Theorem 7.4, page 242. We can use this fact to prove Sylows First Theorem:43 any finite group
G contains at least one psubgroup of each possible order for every
prime factor p of | G |, and each of these subgroups is normal in the
psubgroup that contains it. That is, if G is a finite group of order
pk m, where p is prime, k > 0 and p 6 |m, then it has at least one
subgroup of order pi for all 0 6 i 6 k. Furthermore, each subgroup
of order pi is a normal subgroup of some subgroup of order pi+1 . A
44
Definition 7.10, page 245. maximal psubgroup is called a Sylow psubgroup,44 and Sylows
45
Theorem 7.11, page 245. Second Theorem45 says that for any prime factor p of | G |, the Sylow
psubgroups are conjugate to each other. Therefore, G has exactly one
46
Corollary 7.12, page 245. Sylow psubgroup H if and only if H is normal.46
We denote by Syl p ( G ) the set of Sylow psubgroups of G, and let
n p = | Syl p ( G )| denote the number of Sylow psubgroups. Sylows
47
Theorem 7.13, page 246. Third Theorem47 says that for any finite group G of order pk m, we
have n p 1 (mod p) and n p |m. We can use this to help classify
finite groups of a particular order: for example, if | G | = 15 we find
that n3 = 1 and n5 = 1, so G must have a single normal subgroup
of order 3 and another of order 5. From this we can deduce that
48
Example 7.14, page 247. G = Z15 .48 More generally, if | G | = pq where p and q are distinct
49
Proposition 7.15, page 247. primes, p < q and q 6 1 (mod p) then G = Z pq .49 Any such group
50 must therefore be abelian. 50
Corollary 7.16, page 248.
We can also use Sylows Second and Third Theorems to prove the
nonexistence of finite simple groups of certain orders, by showing that
a group of the given order must have a single Sylow psubgroup for
some p, and this subgroup must necessarily be normal. For example,
there is no simple group of order 1246, since 1246 = 2789 and
51
Example 7.18, page 248. n7 = n89 = 1.51
Another useful fact about pgroups is that they have nontrivial cen-
52
Proposition 7.19, page 249. tre.52 We can use this to show that any group of order p2 is isomorphic
53
Proposition 7.20, page 249. to either Z p2 or Z p Z p ,53 and must in either case be abelian.54
54
Corollary 7.21, page 250. More generally, if | G | = p2 q where p and q are both prime, p < q and
55
Proposition 7.22, page 250. q 6 1 (mod p), then G must be abelian.55
We next discussed the concept of subgroup series: a finite nested
sequence {e} = H0 < H1 < < Hn = G of subgroups of a group
56
Definition 7.23, page 251. G.56 We sometimes make a distinction between an ascending series
and a descending series depending on what order we number the
subgroups in, and we will also sometimes relax the requirement that
finite groups 315
73
Proposition 7.40, page 263. soluble: if H C G and G/H are soluble, then so is G.73 Consequently,
74
Corollary 7.41, page 263. finite direct products of soluble groups are also soluble,74 as are finite
75
Proposition 7.42, page 264. abelian groups75 and finite pgroups.76
76
Proposition 7.43, page 264. Two important results whose proofs are beyond the scope of this
77
Theorem 7.44, page 264. book are Burnsides Theorem,77 which says that any group of order
pm qn , for p and q both prime and m, n > 0, is soluble, and the Feit
78
Theorem 7.45, page 264. Thompson Theorem,78 which says that any group of odd order is
soluble.
The derived series of a group G is defined recursively by using com-
mutator subgroups. We set G (0) = G, and let G (i+1) be the commutator
79
Definition 7.46, page 265. [ G (i) , G (i) ] for i > 0.79 A group is soluble if and only if its derived
series terminates at the trivial subgroup; that is, if G (n) = {e} for some
80
Proposition 7.47, page 265. n > 0.80 The derived length of a soluble group G is the length of
81
Definition 7.48, page 266. its derived series.81 The trivial group is the only group with derived
length 0, the groups of derived length 1 are exactly the abelian groups,
and groups of derived length 2 are called metabelian.
Another important series is the central series, constructed using the
centre. A normal series is said to be central if Gi+1 /Gi 6 Z ( G/Gi )
82
Definition 7.49, page 267. for all 0 6 i < n;82 that is, if each quotient Gi+1 /Gi is contained in
the centre of the corresponding quotient G/Gi of the full group. This
construction only works with normal series, not subnormal ones, since
we need each subgroup Gi to be normal in the full group G for the
quotient G/Gi to be defined.
83
Definition 7.51, page 268. There are two special central series: the upper central series83 and
84
Definition 7.53, page 269. the lower central series.84 The first of these is an ascending series,
and is defined in terms of the higher centres Zi ( G ) of G: we set G0 =
Z0 ( G ) = {e} and recursively define Zi ( G ) such that Zi ( G )/Zi1 ( G ) =
Z ( G/Zi1 ( G )). The second is a descending series, and is defined by
setting G0 = 1 ( G ) = G and recursively defining Gi = i+1 ( G ) =
[ G, i ( G )].
The upper central series need not terminate at G, and the lower central
series need not terminate at the trivial subgroup {e}. The lower
central series is the optimal descending central series: if it reaches the
trivial group, then it does so in fewer steps than any other descending
85
Proposition 7.56, page 270. central series.85 . Analogously, the upper central series is the optimal
ascending central series: if it reaches G then it does so faster than any
86
Proposition 7.58, page 271. other ascending central series.86
A group G with a finite-length central series connecting G with its
trivial subgroup {e} is said to be nilpotent, and the length of the
shortest possible such central series is called the nilpotency class of
87
Definition 7.55, page 270. G.87 The nilpotency class of G will therefore be equal to the length
of the upper or lower central series, whichever is smaller. A group G
finite groups 317
i1
E1
q1
A i2
f
q2
G
E1
commutes.122 A factor set or cocycle is a function : G G A such 122
Definition 7.91, page 285.
that
(eG , g) = ( g, eG ) = 0
for any g G, and
( g1 , g2 ) + ( g1 g2 , g3 ) = g1 ( g2 , g3 ) + ( g1 , g2 g3 )
for all g1 , g2 , g3 G.123 A factor set is determined by a section of the 123
Proposition 7.92, page 286.
extension, and different sections of the same extension yield factor sets
that differ by a coboundary or inner factor set: a function : G G
A such that
( g, h) = ( g) ( gh) + g (h)
for all g, h G, and some function : G A satisfying (eG ) =
0. Two factor sets related in this way are said to be equivalent or
cohomologous.124 Let FS( G, A) be the set of factor sets of extensions 124
Definition 7.95, page 289.
of a group G by a Gmodule A, and let IFS( G, A) be the corresponding
subset of inner factor sets. These both form abelian groups under
the canonical pointwise addition operation.125 We define the group 125
Proposition 7.96, page 290.
Ext( G, A) to be the quotient FS( G, A)/ IFS( G, A); the elements of this
group are equivalence classes of factor sets that differ only by an inner
factor set. Furthermore, equivalent factor sets in this sense correspond
to equivalent extensions,126 so the group Ext( G, A) classifies abelian 126
Proposition 7.97, page 290.
extensions of G by A. The zero element in Ext( G, A) corresponds to
(the equivalence class of) the split extension A , AoG G. The
SchurZassenhaus Theorem says that if G is a group of order n and A
is a Gmodule of order m, where m and n are coprime, then Ext( G, A)
is trivial; that is, the only extension of G by A is the split extension.127 127
Theorem 7.98, page 292.
An extension
i q
A E G
is central if i ( A) Z ( E); that is, if the image of A lies in the centre of
E, so that every element of i ( A) commutes with every element of E.
We can, for example, construct the quaternion group Q8 as a central
extension of the Klein group V4 by Z2 .128 128
Example 7.99, page 293.
Another interesting case is that of a cyclic extension: an extension
of a cyclic group G. Such an extension is determined entirely by a
cyclic extension type, a quadruple ( H, n, , v) where H is a (possibly
nonabelian) group, n N, Aut( H ) and v H such that (v) = v
320 a course in abstract algebra
|G| number G
1 1 {e}
2 1 Z2
3 1 Z3
4 2 Z4 , V4
= Z2 Z2
5 1 Z5
6 2 Z6 = Z2 Z3 , D3
7 1 Z7
8 5 Z8 , Z2 Z4 , Z2 Z2 Z2 , D4 , Q8
9 2 Z9 , Z3 Z3
10 2 Z10 , D5
11 1 Z11
12 5 Z12 = Z4 Z3 , Z2 Z6 = Z2 Z2 Z3 , D6 , Dic3 , A4
13 1 Z13
14 2 Z14 , D7
15 1 Z15 = Z3 Z5
16 14 Z16 , Z2 Z8 , Z4 Z4 , Z2 Z2 Z2 Z2 , D8 , Dic4 , SD16 , D4 Z2 ,
Q8 Z2 , Z8 o3 Z2 , V4 oZ4 , Q8 oZ2 , Z4 oZ4
17 1 Z17
18 5 Z18 = Z9 Z2 , Z3 Z6 = Z3 Z3 Z2 , D9 , Z3 D3 , (Z3 Z3 )oZ2
19 1 Z19
20 5 Z20 = Z4 Z5 , Z2 Z10 = Z2 Z2 Z5 , D10 , Dic5 , AGL1 (5)
21 2 Z21 = Z7 Z3 , Z7 oZ3
22 2 Z22 , D11
23 1 Z23
24 15 Z24 = Z3 Z8 , Z2 Z12 = Z2 Z4 Z3 , Z2 Z2 Z6 = Z2 Z2 Z2 Z6 , S4 ,
D12 , Z3 D4 , Z3 Q8 , Z2 A4 , Z4 D3 , Z2 Dic3 , Z2 D6
= V4 D3 ,
Q8 oZ3 = SL2 (3), Z3 oZ8 , Z3 o D4 , Z3 oQ8
25 2 Z25 , Z5 Z5
26 2 Z26 , D13
27 5 Z27 , Z3 Z9 , Heis(3) = U (3, 3), Z9 oZ3
28 4 Z28 = Z4 Z7 , Z2 Z2 Z7 = Z2 Z14 , Dic7 , D14
29 1 Z29
30 4 Z30 = Z2 Z3 Z5 , D15 , D3 Z5 , D5 Z3
31 1 Z31
279290
In the spirit of James McKays concise proof of Cauchys Theorem, cited at the end of Chapter 2,
Mchel Searcid formulated shorter proofs of Sylows Theorems:
M Searcid, A reordering of the Sylow Theorems, The American Mathematical Monthly 94.2 (1987)
165168
An accessible introduction to semidirect products, motivated by affine transformations (discussed in
Example 7.85), can be found in the following article:
S S Abhyankar and C Christensen, Semidirect products: x 7 ax +b as a first example, Mathematics
Magazine 75.4 (2002) 284289
For more detailed discussion of homological algebra, the following books are good places to start:
K S Brown, Cohomology of Groups, Graduate Texts in Mathematics 87, Springer (1994)
P J Hilton and U Stammbach, A Course in Homological Algebra, second edition, Graduate Texts in
Mathematics 4, Springer (1996)
J J Rotman, An Introduction to Homological Algebra, first edition, Pure and Applied Mathematics 85,
Academic Press (1979)
J J Rotman, An Introduction to Homological Algebra, second edition, Universitext, Springer (2008)
C A Weibel, An Introduction to Homological Algebra, Cambridge Studies in Advanced Mathematics 38,
Cambridge University Press (1994)
All are aimed at graduate or advanced undergraduate students; Weibel and the second edition of
Rotman are pitched at a higher level than the others. Browns book focuses mostly on applications to
group theory, while the others take a more general viewpoint concerning modules over arbitrary rings.
As noted earlier, the classification of groups of order 16 in Section 7.A closely follows the exposition in
the very clear and readable article by Marcel Wild:
M Wild, The groups of order sixteen made easy, The American Mathematical Monthly 112 (2005) 2031
Another classification of groups of order less than 32 can be found in the textbook by John Moody:
J A Moody, Groups for Undergraduates, World Scientific (1994)
Exercises
8.1 Construct the dihedral group D4 as a central extension of V4 by Z2 .
Fafner: New mischief will the Ni-
belung plot against us if the gold gives
him power. You there, Loge! Say with-
out lies: of what great value is the gold
then, that it satisfies the Nibelung?
Loge: It is a toy in the depths of the
water, to give pleasure to laughing chil-
dren; but if it were fashioned into a
round ring it would bestow supreme
8 Rings power and win its master the world.
Wotan: I have heard talk of the Rhines
gold: its glittering glow hides runes of
riches; a ring would give unbounded
power and wealth.
Richard Wagner (18131883),
n Chapter 1 we carefully studied the set Z of integers, giving
I particular attention to its additive structure. By deconstructing and
Das Rheingold (1869)
The following two examples are formed by taking the ring Z and Table 8.1: Addition and multiplication
carefully attaching a specific non-integer element. Well come back to tables for Z2 , Z3 and Z4
326 a course in abstract algebra
this idea later on when we study field extensions, but for the moment
its an interesting and useful way of constructing new rings.
Example 8.5 The set Z[ 2] = { a + b 2 : a, b Z} is a commutative
unital ring under the usual addition and multiplication operations.
It is an abelian group under addition, so R1 holds. Multiplication is
associative and commutative, and also distributive over addition, so
R2, R3 and R5 hold. Finally, 1 = 1 + 0 2 Z[ 2] so R4 holds too.
0 1 1+ i i
The addition and multiplication tables for these rings are shown in
0 0 0 0 0 Tables 8.2 and 8.3. (Note that + 1 = 2 , but since were working
1 0 1 1+ i i over Z2 we have +1 = 1 and hence + 1 = 2 .)
1+ i 0 1+ i 0 1+ i
i 0 i 1+ i 1 Now weve met a few examples of rings, its time to start formulating
0 1 2 1+ 2 some general results about them. The following basic properties are
0 0 0 0 0 almost immediate consequences of Definition 8.1.
1 0 1 2 1+
2
2 0 2 0 2 Proposition 8.8 Let R = ( R, +, ) be a (unital or nonunital) ring. Then,
1+ 2 0 1+ 2 2 1 for any a, b, c R,
0 1 2 (i) 0 a = a 0 = 0,
0 0 0 0 0 (ii) a (b) = ( a) b = ( a b), and
1 0 1 2
0 2 1 (iii) ( a) (b) = ( a b).
2 0 2 1 If R is unital, then
Table 8.3: Multiplication
tables for the
(iv) (1) a = a = a (1) for any a R, and
rings Z2 [i ], Z2 [ 2] and Z2 [ ]
(v) (1) (1) = 1.
rings 327
Proof All of these follow from the ring axioms R1, R2 and R3, and R5
in the case of (iv) and (v). For (i),
0 + a 0 = a 0 = a (0 + 0) = a 0 + a 0
+ 0 1 2 3 0 1 2 3
0 0 1 2 3 0 0 0 0 0
1 1 2 3 0 1 0 0 0 0
2 2 3 0 1 2 0 0 0 0
3 3 0 1 2 3 0 0 0 0
In all of the nontrivial examples of unital rings weve seen so far, the There is a popular story that an em-
multiplicative identity element is different to the additive identity. The inent mathematician once spent sev-
eral hundred pages formally proving
following proposition confirms this in general. that 1+1=2. Unlike many such apoc-
Proposition 8.13 Let R be a unital ring in which 1 = 0; that is, the ryphal legends, this one is true. The
proof occurs on page 379 of the first vol-
additive and multiplicative identities are the same. Then R is trivial. ume of Principia Mathematica, Bertrand
Russell and Alfred North Whiteheads
Proof Suppose that 1 = 0. Then for any a R we have notationally dense three-volume work
on the axiomatic foundations of mathe-
a = 1a = 0a = 0 matics.1
Here we have taken only 329 pages to
and hence R is trivial. confirm that 1 6= 0.
The rings in Example 8.7 all have addition tables structurally identical 1 B A W Russell and A N Whitehead,
Principia Mathematica, Cambridge Uni-
to that of the Klein group V4 = Z2 Z2 . This leads us to ask whether versity Press (19101913)
any of them are structurally identical as rings to Z2 Z2 . But to answer
this question we first need to decide what it means for two rings to be
isomorphic, and also what we mean by the direct sum of two rings.
For the first one, recall that a group isomorphism is a bijection that
respects the group structure in a precisely-defined way.2 Such a 2
Definition 1.17, page 11.
bijection must respect the additive structure of an abelian group, and
since a ring is to some extent just a special kind of abelian group, this
is a good place to start. But a ring also has a multiplicative structure,
so a ring isomorphism should respect that too.
Rather than define ring isomorphisms and then introduce homomor-
phisms as the non-bijective case (as we did with groups) we might as
well introduce both concepts at the same time, although well leave
the main discussion of ring homomorphisms until Chapter 9.
Definition 8.14 Let R = ( R, +, ) and S = (S, , ) be rings, and
suppose that f : R S is a function satisfying
f ( a + b) = f ( a) f (b) and f ( a b) = f ( a) f (b)
for all a, b R. Then f is a (ring) homomorphism.
If f is injective, we call it a monomorphism; if f is surjective, we call
it a epimorphism; and if f is bijective, we call it a isomorphism. A
homomorphism f : R R is called an endomorphism.
The second item on our list is the direct sum of two rings. Again, we
define this by extending the corresponding concept for abelian groups:
Definition 8.15 Let R and S be rings. The direct sum RS is
RS = {(r, s) : r R, s S}
with addition and multiplication operations
( a, b) + (c, d) = ( a + c, b + d) ( a, b) (c, d) = ( a c, b d)
for all a, c R and b, d S.
330 a course in abstract algebra
Proof The first statement follows from Proposition 4.24, since any
ring R is an abelian group, and any ring homomorphism f satisfies
the usual conditions for an abelian group homomorphism.
For the second statement, we want to show that f (1) s = s f (1) = s
for any s S. Since f is surjective, we can find at least one r R such
rings 331
R[ x ] = { a0 + a1 x + a2 x2 + + an x n : a0 , . . . , an R, n Z>0 }.
You have probably met power series before, in the context of Taylors
rings 337
m g g + n g g = (m g + n g ) g
g G g G g G
That is, we multiply two sums in the usual way, combining terms
by using the ordinary multiplication operation in Z and the group
multiplication operation accordingly: n g g mh h = (n g mh )( g h).
We can generalise this by replacing the coefficient ring Z with some
other ring R to get RG.
The units of a ring have some interesting properties: they are closed
under multiplication and include the identity, leading to the following.
Proposition 8.34 Let U ( R) be the set of units of a (unital) ring R. Then
U ( R) forms a group, the group of units, under multiplication.
7
Some books denote U ( R) by R for Example 8.35 The ring Z of integers has only two invertible ele-
this reason.
ments: 1 and 1. Hence U (Z) = {1}
= Z2 .
Example 8.36 The polynomial ring Z[t] also has only two invertible
elements: 1 and 1. Thus U (Z[t]) = {1}
= Z2
The ring Z[t1 ] of Laurent polynomials over Z has infinitely many
units: any element of the form tn , where n Z, is invertible (its
inverse is tn ). Hence U (Z[t1 ]) = {tn : n Z}
= Z2 Z.
rings 339
Example 8.37 The invertible elements in the ring Mn (R) are ex-
actly the nn real matrices with nonzero determinant. Hence
U ( Mn (R)) = GLn (R).
This last example displays the best case scenario: everything is a unit
except for 0 (which definitely cant be). Rings satisfying this criterion
are particularly versatile to work with: we can divide arbitrary nonzero
elements by each other as well as adding, subtracting and multiplying
them. We can legitimately ask (and answer) a wider range of questions
in a ring with this property, such as what gives 1 when multiplied by
3?. It therefore makes sense to give this concept a specific name.
Definition 8.39 A division ring is a ring R such that U ( R) = R =
R \ {0}. That is, every nonzero element r R is a unit.
This isnt to say that rings without this property arent interesting. On
the contrary, such rings often display interestingly awkward behaviour.
Even nicer is the case where such a ring is commutative as well:
Definition 8.40 A field is a commutative division ring. A noncom-
mutative division ring is called a skew field.
Many of the number systems were most familiar with are fields. In
fact, in some sense the point of the rational numbers Q is that they are
the integers Z made as invertible as possible. Well come back to this
idea in a little while when we meet the concept of a field of quotients,
but for the moment here are some examples of fields and skew fields.
Example 8.41 The rational numbers Q, the real numbers R and the
complex numbers C are all fields.
The quaternions H form a skew field.
Using this terminology, we see from Table 8.6 that 2, 3, 4 Z6 are all
zero divisors, but 1 and 5 arent (and by convention neither is 0).
More generally, there is a simple criterion governing whether a partic-
ular nonzero element of a ring Zn is a zero divisor:
Proposition 8.44 An element m Zn is a zero divisor exactly when it
is not coprime to n; that is, when gcd(m, n) > 1.
two fractions together, and also figure out how the ring Z embeds in
this field.
Addition and multiplication is given by
a c ad + bc a c ac
+ = and = ,
b d bd b d bd
and the integers Z embed via the homomorphism i : Z , Q given by
f (n) = n1 .
All of this works fine in any integral domain, leading to the following.
Definition 8.51 Let R be an integral domain, and let
Q( R) = ( R R )/
be the set of equivalence classes of the relation defined by ( a, b)
(c, d) when ad = bc. We represent the equivalence class [( a, b)] by
a/b or ba . Then Q( R) forms a field (the field of fractions or field of
quotients) with the addition and multiplication operations
a c ad + bc a c ac
+ = and =
b d bd b d bd
for any a, b, c, d R such that b, d 6= 0.
Furthermore, there is an injective ring homomorphism i : R Q( R)
mapping r 7 r/1 for all r R.
So, the field of fractions Q(Z) of the integers Z is the field Q of rational
numbers. This shouldnt be surprising, because we formulated the
definition of Q( R) precisely so that this would be the case.
Example 8.52 Lets calculate the field of fractions Q(Z3 ) of the ring
Z3 . First, we take the set
Z3 (Z3 \{0}) = {(0, 1), (1, 1), (2, 1), (0, 2), (1, 2), (2, 2)}.
Now we mod out by the equivalence relation to find that
(0, 1) (0, 2), (1, 1) (2, 2), (1, 2) (2, 1).
Hence
Q(Z3 ) =
0 1 2
1, 1, 1
which is just the image of Z3 under the inclusion homomorphism
i : Z3 , Q(Z3 ). Hence Q(Z3 )
= Z3 .
This makes sense, since Z3 is already a field under the usual modulo3
addition and multiplication operations. More generally:
Proposition 8.53 Let F be a field. Then Q(F) = F.
The main point of the field of fractions is that its the smallest field
containing the given integral domain.
Our knowledge springs from two fun- 8.A Modules and representations
damental sources of the mind; the first
is the capacity of receiving representa-
In the study of linear algebra, we are largely concerned with
tions (receptivity for impressions), the
second is the power of knowing an vector spaces and linear maps between them. Recall that a vector
object through these representations
(spontaneity of concepts). space consists of a set V (of vectors) and a field F (of scalars), together
Immanuel Kant (17241804), with a binary operation + : V V V (called vector addition) and a
Critique of Pure Reason (1781)
function : F V V (called scalr multiplication). We require (V, +)
to form an abelian group, and for the scalar multiplication operation
Nowadays group theoretical methods
especially those involving charac-
to satisfy some other basic operations on its own, and when interacting
ters and representations pervade all with vector addition.
branches of quantum mechanics.
An obvious generalisation of this arrangement is to allow the scalars to
George Whitelaw Mackey
(19162006), Group Theory and its form a weaker structure than a field, say a ring. In doing this, we have
Significance for Mathematics and Physics, to be a little bit careful. There are certain properties that a field has
Proceedings of the American
Philosophical Society 117 (1973) that not all rings have, and which makes this more general scenario
374380 more complicated (and more interesting).
rings 345
for any vector v and scalars h and k. That is, multiplying some vector
v by a scalar k, and then by another scalar h, gives the same result
as multiplying v by h first and then by k. But if our scalars arent
commutative, this wont always be the case.
Another way of looking at this is to define two scalar multiplication
operations: one where we do the multiplication on the left (the usual
notation) and one where we multiply on the right. The associativity
conditions would then look like
Example 8.64 Let G be a group, and let R be its integral group ring
ZG as introduced in Example 8.31. Then a ZGmodule is exactly
the same as a Gmodule in the sense of Definition 7.90.
Given two modules over the same ring, we can combine them to form
a new, larger module in a familiar way:
348 a course in abstract algebra
g ( g) g ( g)
the permutation representation for S3 has the form shown in Ta-
1 0
(1 2)
1 1 ble 8.10. Each of these matrices comprises a 11 matrix [1] in the
01 0 1
0 1
1 0
top left corner, and a 22 submatrix in the bottom right corner.
(1 3) 1 0 (2 3) 1 1
1 1 0 1 The permutation representation thus decomposes into the direct
(1 2 3) 1 0 (1 3 2) 1 1
sum of the 1dimensional trivial representation and another 2
Table 8.11: The standard representation dimensional representation, shown in Table 8.11, called the standard
of S3
representation of S3 . This representation acts on the 2dimensional
subspace {( x, y, z) : x + y + z = 0} of F3 .
Modules Representations
FGmodule representation of G over F
FGmodule homomorphism intertwiner between representations of G over F
direct sum of FGmodules direct sum of representations of G over F
irreducible FGmodule irreducible representation of G over F
Table 8.12: A lexicon of module and
representation terminology and con-
The following important result concerns the reducibility or otherwise cepts
of FGmodules (and representations).
Theorem 8.76 (Maschkes Theorem) Let G be a finite group, F a field
of characteristic zero, and V an FGmodule. If U is an FGsubmodule of
V then there exists another FGsubmodule W of V such that V = U W.
Wikimedia Commons
for all v V. The idea is that in some sense q is the average of p over
Heinrich Maschke (18531908) was the group G.
born in Breslau, Prussia (now Wrocaw,
Poland). He studied mathematics at We need q to be a surjective FGhomomorphism from V onto U. To
the Universities of Heidelberg and check surjectivity, consider any element v V. Then for any g G it
Berlin, graduating with high distinc-
tion from the latter in 1878, and subse- follows that g v lies in V, because V is an FGmodule, then p( g v)
quently undertood graduate research lies in U, because im p = U, and finally g1 p( g v) lies in U as well,
at the University of Gttingen, where
he was awarded his doctorate in 1880
because U is an FGsubmodule of V, and hence closed under the
for a thesis entitled On a triple orthog- action of G.
onal surface system formed from sur-
faces of the third order. Next we must check that q is an FGhomomorphism: an Flinear
After this he worked as a mathemat- map from V to U that commutes with the Gaction. The Flinearity
ics teacher at a gymnasium in Berlin. requirement is satisfied since q is a linear combination of Flinear
But he found the work increasingly un-
fulfilling, especially after a years sab- transformations. And for any h G and v V, we have:
batical undertaking research with Felix 1
Klein (18491925) at Gttingen in 1887. q(h v) =
|G| g1 p( g (h v))
Inspired by two friends who had g G
quickly found academic posts in the 1
USA, he resolved to emigrate. To im- =
|G| g1 p(( gh) v)
prove his chances of finding suitable g G
work, he studied electrical engineering.
1
Arriving in New York in April 1891, he
soon got a job as an electrician with the
=
|G| hh1 g1 p(( gh) v)
g G
Weston Electrical Instrument Company
1
of Newark, New Jersey.
In 1892 he was appointed as an assis-
= h
|G| ( gh)1 p(( gh) v)
gh G
tant professor at the newly founded
1
University of Chicago. Here, he was
able to fully exercise his mathemati-
= h
|G| g 1 p ( g v )
g G
cal abilities: his teaching experience
served him and his students well, and = h q ( v ),
he made many contributions to math-
ematical research, particularly on fi- so q is an FGhomomorphism.
nite groups and quadratic differential
forms. He was an active founder mem- We now set W = ker q, which is therefore an FGsubmodule of V,
ber of the American Mathematical So- and as noted earlier U = im q. Furthermore, the intersection U W
ciety, serving as Vice President in 1907.
consists only of the zero vector 0, and any vector v in V can be written
Despite his generally excellent health,
he died in March 1908 of complications uniquely as a sum v = u+w, for some u U and w W. Hence
from emergency surgery. V = U W as claimed.
rings 353
Summary
28
Example 8.21, page 332. ring of nn matrices with entries from some ring R,28 as well as the
rings UT n ( R), LT n ( R) and Dn ( R) of triangular and diagonal matri-
29
Example 8.22, page 333. ces.29 The ring Mn ( R) is noncommutative except when n = 1 and R is
commutative, and its elements can be regarded as endomorphisms of
R; more generally we can form the ring End( R) of endomorphisms of
30
Example 8.23, page 334. R,30 and also generalise this to endomorphism rings of other objects
31
Example 8.24, page 334. such as abelian groups.31
The next class of rings we studied were polynomial rings such as
32
Example 8.25, page 335. Z[ x ].32 These are formed by adjoining a formal variable x to an ex-
isting ring, in a similar way to the construction of rings such as Z[i ]
and Z2 [ 2]. This process can be generalised in various ways: we can
use an arbitrary coefficient ring to get R[ x ], allowing negative powers
33
Example 8.27, page 336. yields the ring Z[ x 1 ] of Laurent polynomials,33 we can define mul-
tivariate polynomial rings like Z[ x, y] or R[ x1 , . . . , xk ] by using more
34
Example 8.29, page 336. than one variable,34 and by allowing infinitely many terms we obtain
35
Example 8.30, page 336. formal power series rings like ZJxK.35 A related example is that of
the group ring ZG (or RG) of a group G, whose elements are formal
36
Example 8.31, page 337. Zlinear (or Rlinear) combinations of elements of G.36
Next we considered the multiplicative structure of rings in more
37
Definition 8.33, page 338. detail, introducing the notion of an invertible element, or unit.37 The
units in a ring R are closed under multiplication, include the unity
element 1 itself, and all have unique inverses. They therefore form a
38
Proposition 8.34, page 338. multiplicative group, the group of units U ( R).38 The rings Z and Z[t]
have only two units, namely 1, and hence U (Z) = U (Z[t]) = Z2 .
The ring Z[t ] of Laurent polynomials has infinitely many units of
1
that is, those for which gcd(m, n) > 1.43 A ring such as Z or Z[t] that 43
Proposition 8.44, page 340.
contains no zero divisors is said to be an integral domain.44 All fields 44
Definition 8.46, page 341.
are integral domains,45 and all finite integral domains are fields,46 but 45
Proposition 8.47, page 341.
not all infinite integral domains are fields. Although integral domains 46
Proposition 8.49, page 341.
dont necessarily contain a full complement of units, they do satisfy
multiplicative cancellation laws.47 47
Proposition 8.48, page 341.
356 a course in abstract algebra
Exercises
8.1 Show that the prime subring h1i of a ring R is equal to the intersection of all other subrings of R.
8.2 Write down the addition and multiplication tables for the rings Z2 [], Z3 [i ], Z3 [ 2], Z3 [ ] and
Z3 [].
8.3 Show that if f : R S is a homomorphism of unital rings, then there is a unique induced
homomorphism U ( f ) : U ( R) U (S), between the corresponding groups of units, such that
U ( f )(r ) = f (r ) for all r U ( R).
8.4 A ring element r R is idempotent if r2 = r. Show that if R is an integral domain, then 0 and 1
are the only idempotent elements.
8.5 Let R be an integral domain. Show that the field of fractions Q( R) from Definition 8.51 is indeed a
field.
8.6 Let
H = a+bi +cj+dk : a, b, c, d Z or a, b, c, d Z+ 21
little while later that kernels of group homomorphisms are normal, (0, 0) (0, 1) (1, 0) (1, 1)
and conversely that all normal subgroups arise in this way. (0, 0) (0, 0) (0, 0) (0, 0) (0, 0)
(0, 1) (0, 0) (0, 1) (0, 0) (0, 1)
Now we want to study the analogous concepts in ring theory: subrings, (1, 0) (0, 0) (0, 0) (1, 0) (1, 0)
cosets, kernels, homomorphisms, normal subrings and quotient rings. (1, 1) (0, 0) (0, 0) (1, 0) (1, 1)
This will keep us busy for the rest of this chapter. Table 9.1: Addition and multiplication
tables for the ring Z2 Z2
and from the second of these we can see that 4 acts as a unity element.
Similarly, the subset {0, 3} forms an abelian subgroup under ad-
dition, is closed under multiplication, and the element 3 acts as a
multiplicative identity.
In neither of these cases is the unity element in the subset the same
as the unity element 1 Z6 .
too much about exactly which element fulfils this rle? There is, again,
no universally adopted convention, and after more than a century of
intensive study of ring theory, it seems unlikely that a consensus will
form any time soon.
In this book, we will adopt the conventions in the following defini-
tion, but you should be aware that many books use slightly different
definitions or terms.
Definition 9.2 Let R = ( R, +, ) be a ring, and let S be a subset of R.
Then S is a subring of R if (S, +S , S ) is a ring in its own right, where
+S and S are the restrictions to S of the addition and multiplication
operations in R.
If S does not contain a unity element, then we say it is a nonunital
subring or a subring without unity.
If the unity element 1S in S is the same as the unity element 1R in R,
then we call S a unital subring or subring with unity.
In other words, unless otherwise specified we require a subring to have
a unity element, but we dont necessarily require that unity element
to be the same as the one from the larger ring. (We could perhaps
refer to this situation as a subring with different unity, but thats
a bit unwieldy and isnt in common usage.) We reserve the terms
unital subring and subring with unity for the case where the two
unity elements are the same. This is a somewhat complicated way of
doing things, but seems to be the least worst option available.
The following proposition provides criteria for a subset S R to form
a unital subring of R. Essentially, it says that we really just have to
check whether S forms a subgroup of R under addition, is closed
under multiplication, and contains the unity element 1 R.
Proposition 9.3 Let R = ( R, +, ) be a ring, and let S R be a subset
of R. Then S is a subring of R if and only if the following conditions hold:
SR1 (S, +) is a subgroup of ( R, +),
SR2 a b S for all a, b S, and
SR3 there exists a unity element 1S S.
Furthermore, S is a unital subring of R if and only if
SR4 1S = 1R .
Proof This is all just a consequence of Definitions 8.1 and 9.2. Since
( R, +) is an abelian group, condition SR1 is equivalent to requiring
that (S, +) is an abelian group itself, which in turn is equivalent to
condition R1 from Definition 8.1.
The closure condition SR2 is equivalent to requiring the restricted
multiplication operation S to be well-defined on S. If it is, then it
362 a course in abstract algebra
condition s2 = s.
Such an element is said to be idempotent. It is always the case that
the zero element and (if it exists) the unity element are idempotent.
Sometimes, however, there may be other elements satisfying this
condition as well.
This leads us to the following clarification.
Definition 9.17 Let f : R S be a ring homomorphism. Then f is
unital if f (1R ) = 1S , and nonunital if f (1R ) 6= 1S , or if either R or S
are nonunital rings.
f ( a b) = f ( a) f (b) = 0 0 = 0,
Then ker( f ) and ker( g) are both subrings of Z6 . Each has a unity
element (4 in the case of ker( f ), and 3 in ker( g)) but neither is equal
to 1, the unity element of the ambient ring Z6 .
Example 9.26 shows that the kernel of the modulon residue map
f n = []n : Z Zn is the nonunital subring nZ Z consisting of all
integer multiples of n. Any multiple of n times an arbitrary integer is
also a multiple of n, since kn m = (km)n nZ.
Similarly, Example 9.27 shows that the kernel of the evaluation ho-
momorphism eva : Z[ x ] Z consists of all polynomials p that have
( x a) as a factor (or, equivalently, for which p( a) = 0). Any other
polynomial q Z[ x ] multiplied by one of these polynomials will also
have ( x a) as a factor, and therefore belong to ker(eva ).
More precisely, for any k ker( f n ) and any m Z both m k and
k m lie in ker( f n ). Similarly, for any polynomial p ker(eva ) and
any polynomial q Z[ x ] both p q and q p lie in ker(eva ).
In fact, this is true for the kernel of any ring homomorphism:
Proposition 9.28 Let f : R S be a ring homomorphism. Then for any
k ker( f ) and a R we have k a ker( f ) and a k ker( f ).
A (left, right or two-sided) principal ideal is the largest ideal that can
be generated from a single given element of the ring in question.
Example 9.31 The previous example can be extended to finite sets of
elements of a commutative unital ring R. Suppose that a1 , . . . , ak R.
Then
h a1 , . . . , a k i = {r1 a1 + + r k a k : r1 , . . . , r k R }
is the ideal of R generated by a1 , . . . , ak .
Example 9.34 For any ring R, the trivial subring {0} is an ideal: it
satisfies the left and right absorption conditions by Proposition 8.8 (i).
Example 9.35 Any ring R is an ideal of itself, with the left and right
absorption conditions reducing to the requirement that R be closed
under multiplication, which it is.
Definition 9.29 also includes two slightly more general objects: left and
right ideals. If a ring R is commutative, then clearly any left ideal will
also be a right ideal, and vice versa. However, if R isnt commutative
then we have the possibility of left ideals that arent right ideals, and
right ideals that arent left ideals. The following example explores this
idea.
Example 9.36 Let
0 x
I= : x, y R .
0 y
This is a left ideal but not a right ideal of the ring M2 (R), since it
forms an abelian group under addition, is closed under addition and
multiplication, and satisfies the left absorption condition I3L from
Definition 9.29. Similarly, the set
0 0
J= : x, y R
x y
is a right ideal but not a left ideal in M2 (R).
The only one of these ideals of Z12 that contains the unity element 1
is the full ring Z12 = h1i. Furthermore, the only ideal of M2 (Z) that
contains the identity 10 01 is M2 (Z) itself, and R[ x ] is the only ideal
Proof Suppose that u I is a unit. Then the left and right absorption
conditions imply that u r I and r u I for any r R. Setting
r = u1 we see in particular that u u1 = 1 = u1 u I as well, and
hence by Proposition 9.38 it follows that I = R.
Proof First we must prove that R/I forms an abelian group with the
specified addition operation, and this essentially follows from Propo-
sition 3.24. The addition operation is well-defined and associative, the
additive identity is the coset I = 0+ I, and any coset a+ I has inverse
a+ I.
The multiplication operation in R/I is associative since the one in R
is: for any a, b, c R we have
( a + i ) j = a j + i j I, j (a + i) = j a + j i I
and
f ( a b) = ( a b)+ I = ( a+ I ) (b+ I ) = f ( a) f (b).
ideals 383
The prime ideal is a princess of the 9.4 Prime and maximal ideals
world of ideals. Her father is the prince
Point in the world of geometry. Her
mother is the princess Prime Num- We saw earlier that nZ = hni is an ideal of Z. Indeed, all ideals of
bers in the world of numbers. She Z are of this form. Also, we have a series of inclusions h9i h3i Z.
inherits the purity from her parents.
Is there another ideal that fits between h3i and Z? That is, can we find
Kazuya Kato,
lecture, 11 February 2013 an ideal I such that h3i I Z?
The answer is no. Suppose that I strictly contains h3i, and therefore
that there exists some nonzero integer a I \ h3i. This means that a
is not of the form 3n for any integer n. Suppose that a = 3n+1 for
some n Z. Then we need I to satisfy the conditions for an additive
abelian subgroup of Z, and hence I must also contain all elements of
the form 3k+1 for k Z, because it has to be closed under addition. It
must contain all elements of the form 3k+2 as well, since the additive
inverse of 3n+1 is (3n+1) = 3n 1 = 3(n+1)+2. Thus I = Z.
If, on the other hand, a is of the form 3n+2, then I = Z by a very
similar argument. We give ideals of this type a special name:
Definition 9.56 Let R be a (possibly nonunital) ring, and let I be a
proper ideal of R. Then I is a maximal ideal if there exists no other
ideal J such that I J R.
Let a R be some element of R. Then the principal ideal h ai is a
maximal principal ideal if there exists no other principal ideal hbi
such that h ai hbi R.
Here are some examples of maximal and non-maximal ideals. First
lets look at a matrix ring.
Example 9.57 The ideal 20 03 in M2 (Z) consists of all matrices of
All of the ideals of Z12 are principal. In the next chapter we will study
a class of rings with this property, called principal ideal domains.
Example 9.59 The ideal ht2 i in Z[t] consists of all polynomials with
zero constant and linear terms; that is, those of the form
a n t n + + a2 t2 .
This is not a maximal ideal, because it is properly contained in hti,
which is also not a maximal ideal. However, the ideal
ht, 2i = { an tn + + a1 t : a1 , . . . , an Z}
{bn tn + + b1 t + b0 : b0 , . . . , bn 2Z}
is maximal, as is any other ideal of the form ht, pi for p prime. The
quotients are
Z[ t ] / h t i
= Z, Z[t]/ht, 2i
= Z2 .
Of these, Z2 is a field, while Z is just an integral domain. The
quotient Z[t]/ht2 i is not even that, since t2 = 0 means that t is a zero
divisor.
However, if we consider the ring R[t] of polynomials with real co-
efficients, we find that the ideal h x i is maximal: it consists of the
polynomials with zero constant term, and no other proper ideal of
R[t] contains it. The quotient R[t]/hti = R, which is a field.
In fact, not only is ht2 i not maximal in Z[t], no ideal of the form h f i
for f Z[t] is either:
Proposition 9.60 No principal ideal in Z[t] is maximal.
Summary
28
Definition 9.17, page 368. element, or whose domain or codomain are nonunital rings.28 Any
nontrivial ring R admits at least one unital homomorphism, the iden-
tity homomorphism idR : R R that maps everything to itself, and
one nonunital homomorphism, the zero homomorphism z R : R R
that maps everything to 0.
An interesting class of homomorphisms are the evaluation homomor-
29
Example 9.20, page 369. phisms, defined on polynomial rings.29 For some fixed element a R,
the homomorphism eva : R[ x ] R maps a polynomial p R[ x ] to its
evaluation p( a) obtained by setting x = a.
Next we examined the properties of images and kernels of ring homo-
morphisms. Images of group homomorphisms are always subrings
of their codomains, but the situation is slightly more complicated
with ring homomorphisms. In general, images of nonunital homo-
morphisms are (possibly nonunital) subrings, while images of unital
30
Proposition 9.21, page 369. homomorphisms are unital subrings.30 The kernel of a ring homomor-
phism f : R S is the set of elements of R that are mapped to the zero
31
Definition 9.24, page 371. element in S.31 Kernels of ring homomorphisms are at least nonunital
subrings of their domains: they are additive abelian subgroups and
closed under multiplication, but need not have a unity element. They
32
Proposition 9.28, page 372. do, however, satisfy absorption conditions;32 that is, for any r R
and k ker( f ), the products r k and k r lie in ker( f ).
Motivated by this, we defined an ideal I of a ring R to be a subset
of R that satisfies the first two subring conditions and the left and
33
Definition 9.29, page 372. right absorption conditions.33 Ideals are the ring-theoretic analogues
of normal subgroups. In particular, for a commutative unital ring R
and some element a R we defined the principal ideal h ai generated
34
Example 9.30, page 373. by a to be the set {ra : r R}.34 Just as every group has at least two
normal subgroups, itself and the trivial group, every ring has at least
35
Example 9.34, page 374. two ideals: itself and {0}.35 , 36 If R is a unital ring, then the only ideal
36
Example 9.35, page 374. of R that contains 1 is R itself;37 more generally, R is the only ideal
37
Proposition 9.38, page 374. containing a unit.38 That is, no proper ideals contain any invertible
38
Proposition 9.39, page 375. elements or the unity element 1. As anticipated, not only is every ring
homomorphism kernel an ideal, every ideal can be regarded as the
39
Proposition 9.48, page 379. kernel of some ring homomorphism.39
Having obtained a definition for the ring-theoretic analogue of a nor-
mal subgroup, we then turned our attention to cosets. There were
two approaches we could have taken, the additive or multiplicative
viewpoint, but for various reasons the multiplicative one was unsuit-
able for our purposes. So, given a ring R, an ideal I and an element
a R, we defined the coset of I corresponding to a to be the set
40
Definition 9.40, page 375. a+ I = { a + i : i I }.40 This enabled us to define the quotient ring
41
Proposition 9.41, page 376. R/I,41 whose elements are the cosets of I in R, with addition and
ideals 391
Exercises
9.1 Show that the quotient ring Z[ x, y]/h x2 +1, y2 +1, xy+yx i is isomorphic to the ring L of Lipschitz
integers described in Example 8.12.
9.2 Suppose that 1 < k < n and that gcd(k, n) > 1. Show that hki is a nonunital subring of Zn , and
that all nonunital subrings of Zn are of this form.
9.3 Let I denote the 33 identity matrix, and let A be a matrix of the form
0 a b
" #
A= 0 0 c
0 0 0
for some nonzero a, b, c R. Then A3 = 0 but neither A nor A2 are zero. We say that A is nilpotent;
it has finite multiplicative order. Show that I and A generate a commutative subring of M3 ( R).
9.4 Show that the evaluation homomorphism eva : R[ x ] R is indeed a ring homomorphism for any
a R.
9.5 Show that, for a commutative unital ring R and element a R, the principal ideal h ai is indeed an
ideal of R.
392 a course in abstract algebra
Show that I is an ideal of UT3 (R) and that UT3 (R)/I = RRR.
9.10 Let I = 24Z, J = 60Z and K = 105Z. What are the following ideals of Z?
(a) I J (e) I + J
(b) I K (f) I +K
(c) J K (g) J +K
(d) I J K (h) I + J +K
9.11 Let X be a (possibly infinite) set, and let R = { f : X R} be the set of all possible real-valued
functions on X, equipped with the following addition and multiplication operations:
( f + g)( x ) = f ( x ) + g( x ) and ( f g)( x ) = f ( x ) g( x )
for all f , g R and x X. Then R is a commutative ring. For any fixed a R, we define c a R to
be the constant function given by c a ( x ) = a for all x X.
(a) Show that c0 is the zero of R, and c1 is the multiplicative identity.
(b) For any a R let eva : R R be the evaluation homomorphism eva ( f ) = f ( a). Show that
eva is a surjective homomorphism.
(c) Show that Ia = { f R : f ( a) = 0} is a maximal ideal of R.
9.12 Let R be a unital ring, and let I be an ideal of R. Define
U I = { a U ( R ) : ( a 1) I }.
Show that U I P U ( R).
9.13 Let R be a commutative ring, and choose some element a R. Define
Ia = {r R : ar = 0}.
(a) Show that Ia is an ideal of R.
(b) Find I4 and I9 in Z12 .
(c) Suppose that a2 = a. Show that R/Ia
= h ai
There are several requirements that
must be met to establish a domain. In
general it must be responsibly man-
aged. There must be a responsible per-
son to serve as a coordinator for do-
main related questions, there must be
a robust name service, it must be of at
least a minimum size, and the domain
must be registered with the central do-
10 Domains main administrator.
Jon Postel (19431998),
The Domain Names Plan and Schedule,
IETF RFC 881, November 1983
If our ring R has one of these functions that enables us to state and
prove an analogue of the Division Theorem, then that means we can
happily do long division with remainders in R. But the statement of
the Division Theorem really depends on our choice of norm function,
and there may be more than one of these. (Indeed, the function n 7 n2
suffices in Z as well as the usual | |.) So it makes more sense to fold
the Division Theorem condition into the criteria for our norm function,
and thus we arrive at the following definition:
Definition 10.2 Let R be an integral domain. A Euclidean norm or
Euclidean valuation is a function v : R N0 such that
v( ab) > v( a) (10.3)
for all a, b R , and for any a, b R with b 6= 0 there exist q, r R
such that a = qb + r and either r = 0 or v(r ) < v(b).
The term Euclidean norm is also often used to refer to the Euclidean
distance function in Rn determined by Pythagoras Theorem, and
so from now on we will use the term Euclidean valuation to avoid
confusion.
The integral domains were interested in at the moment are those for
which we can find a Euclidean valuation:
Definition 10.3 An integral domain R that admits a Euclidean valu-
ation is said to be a Euclidean domain.
Note that for a given integral domain R to be a Euclidean domain,
we only need to be able to define a Euclidean valuation v : R N0 :
any suitable one will do, and the valuation v isnt itself considered
part of the structure for the domain. That is, the mere existence of a
Euclidean valuation is enough.
So, the ring Z of integers is a Euclidean domain via the absolute value
function v(n) = |n|, and also via v(n) = n2 .
The significance of the condition (10.3) is given by the next proposition:
Proposition 10.4 Let R be a Euclidean domain, and let v : R N0
be a Euclidean valuation for R. Then v(1) 6 v( a) for any a R , and
v( a) = v(1) if and only if a is a unit.
for any a R .
If a R is a unit, then
x4 + 0x3 + x2
5x3 x2 2x
5x3 + 0x2 + 5x
x2 7x 3
x2 + 0x 1
7x 2
The quotient is the polynomial left on the top line, namely q =
x2 + 5x 1, while the remainder is the polynomial left at the end, in
this case r = 7x 2. We can easily check this:
qg + r = ( x2 + 5x 1)( x2 + 1) + (7x 2)
= x4 + 5x3 + 5x 1 7x 2
= x4 + 5x3 2x 3
= f
The next corollary is a special case, which we will use in Section 10.A
to help characterise the prime elements in the ring Z[i ] of Gaussian
integers.
Corollary 10.10 Let F p denote the finite field of prime order p. Then
U (F p )
= Z p 1 .
the significance of the end result of this process? Ideally, wed hope
that its a suitable analogue of a greatest common divisor in whatever
Euclidean domain we happen to be working in.
10.2 Divisors, primes and irreducible elements Prime numbers are those which have
no other factor but one, as three only
has a third, and five only has a fifth,
In order to explore this idea, we first need to decide what we and seven only has a seventh; that is,
mean by a divisor in some arbitrary integral domain. This is fairly they have only one factor. Composite
numbers are those which not only di-
straightforward, and we arrive quickly at the following definition. vide by one, but are also produced by
Definition 10.13 Let R be an integral domain, and let a, b R. Then another number, such as nine, twenty-
one, fifteen and twenty-five. That is, we
a divides b, or is a divisor or factor of b if there exists some element say three times three, and seven times
r R such that b = ra. We denote this by a|b. three, three times five, and five times
five.
This can be extended to noncommutative division rings by the in- St Isidore of Seville (c.560636AD),
troduction of left and right divisors. We say that a is a left divisor Etymologi III:7
of b if there exists some r R such that b = ar, and a right divisor
of b if there exists some s R for which b = sa. Furthermore, a is a
(two-sided) divisor of b if it is both a left and right divisor.
Proof Suppose that a|b. Then there exists some r R such that b = ra.
But the principal ideal h ai consists of all elements of R of the form ra
for some r R. Hence b h ai.
Now suppose b h ai. Then b = ra for some r R, and any other
multiple of b (that is, any element of the form sb for some s R) can
be written as sb = s(ra) = (sr ) a h ai. But hbi consists of all elements
of the form sb, so any element of hbi must also lie in h ai, so hbi h ai.
Finally, suppose hbi h ai. This means that every element of the form
402 a course in abstract algebra
Proof The equivalence of (i) and (ii) follows from Proposition 10.14.
If a b then a|b, which means that hbi h ai. Also b| a, which means
that h ai hbi. Hence h ai = hbi. Each step in this argument is
reversible, so the converse holds too: if h ai = hbi then a b.
Now suppose a b. If a = 0 then b = 0, so suppose both a and b are
nonzero. Then there exist r, s R such that a = rb and b = sa. Then
a = r (sa), so (rs) a a = 0 and hence (rs 1) a = 0. Since R is an
integral domain, it has no zero divisors, and therefore rs = 1. Thus r
and s are units, with s = r 1 . The converse is straightforward: if there
exists a unit q R such that a = qb then b| a, and also b = q1 a so a|b.
Thus a b.
Proposition 10.17 Association is an equivalence relation.
These two definitions give rise to subtly different concepts. The sub-
tlety is exacerbated by the fact that in familiar number systems like
Z, prime elements are irreducible and irreducible elements are prime.
More generally, only the first of these is true in a given integral domain.
Proposition 10.23 Any prime element of an integral domain R is irre-
ducible.
Proof Let r R be prime, and suppose that r = ab. Clearly r |r, so
r | ab, and since r is prime we must have either r | a or r |b. Suppose that
r | a, without loss of generality. But a|r as well since r = ab, so r a,
which means that r = ua for some unit u R. By the cancellation law
for integral domains (Proposition 8.48) we have u = b, so b is a unit
and hence r is irreducible.
In a little while, well investigate the exact circumstances in which the
converse fails to be true, but for the moment here is an example.
Example 10.24 Let Z[ 3] = { a + b 3 : a, b Z}. This is an
integral domain, and we claim that in it, 2 is irreducible but not prime.
To show this, we need to use the norm function N : Z[ 3] N
given by N ( a + b 3) = a2 + 3b2 .
In particular, this function has the following properties:
(i) N (r ) = 1 if and only if r is a unit; and
(ii) N (rs) = N (r ) N (s) for any nonzero r, s Z[ 3].
Now observe that (1+ 3)(1 3) = 4 = 22. Suppose that
2 = rs where r = u1 +v1 3 and s = u2 +v2 3. Then
4 = N (2) = N (rs) = N (r ) N (s) = (u21 + 3v21 )(u22 + 3v22 ).
If N (r ) = 1 then r is a unit and r 1 = u1 v1 3.
406 a course in abstract algebra
Similarly, if N (s) = 1 then s is a unit, with s1 = u2 v2 3.
The only other option is that N (r ) = N (s) = 2 6= 1. This ensures that
neither r nor s are units, but also requires u21 + 3v21 = 2, and there
are no integers u1 and v1 satisfying this equation, so N (r ) cant be 2.
Thus, the only factorisations 2 = rs in Z[ 3] require either r or s to
be a unit. Also, N (2) = 4 6= 1, so 2 itself isnt a unit, and is therefore
irreducible.
Now recall that 2|4 and 4 = (1+ 3)(1 3). We know that 2
isnt a unit, so for it to be prime we would require either 2|(1+ 3)
or 2|(1 3). For this to be the case, we would need an element
t Z[ 3] for which 2t = 1 3. But this requires t = 12 23 ,
neither of which are elements of Z[ 3]. So 2 is not prime in
Z[ 3].
Something else we notice in this example is that 4 factorises in two
different ways 22 = (1+ 3)(1 3) as a product of irreducible
elements in Z[ 3]. In Z, however, were used to irreducible fac-
torisations being unique (at least up to permutation of factors and
multiplication by units u = 1). Later well return to this idea and
investigate when factorisations are unique in this way.
Recall that we defined an ideal I R of a ring R to be prime if, for
5
Definition 9.63, page 387. any a, b R with ab I, either a I or b I.5 This looks very similar
to the definition of a prime element in an integral domain R: p R is
6
Definition 10.22, page 405. is prime if, whenever p| ab for some a, b R, either p| a or p|b.6
In one definition we are concerned with whether the given ideal I
contains one or other element a or b, and in the other we are interested
in whether the given element p is a multiple of a or b.
Is this just a coincidental analogy, or is there a deeper connection?
And what can we say about irreducible elements and maximal ideals?
Proposition 10.25 Let p be a nonzero element of an integral domain R.
Then
(i) p is irreducible if and only if h pi is a maximal principal ideal, and
(ii) p is prime if and only if h pi is a prime ideal.
10.3 Principal ideal domains Given two numbers which are rela-
tively prime, find the least multiple of
each such that one multiple exceeds the
Proposition 10.25 and Corollary 10.26 provide a neat connection be- other by unity.
tween prime elements, prime ideals and integral domains as quotients, Claude Gaspard Bachet de Mziriac
(15811638),
and a somewhat weaker and untidier link between irreducible ele-
Problemes plaisans et delectables, qui se
ments, maximal ideals and fields as quotients. sont par les nombres (1624) 18
As remarked just now, the problem arises when the integral domain
under investigation has maximal ideals that arent principal, and as
we saw in Example 9.59 and Proposition 9.60 there are some by now
quite familiar examples of integral domains that fall into this category.
There are, however, some integral domains that have only principal
408 a course in abstract algebra
ideals, and when were dealing with one of these, the second part of
both Proposition 10.25 and Corollary 10.26 become much neater.
Definition 10.27 An integral domain R is said to be a principal
ideal domain or PID if all of its ideals are principal; that is, of the
form h ai for some a R.
Our model integral domain so far has been Z, and it turns out that it
all of its ideals are principal; that is, of the form hni for some n Z.
Proposition 10.28 The ring Z is a PID.
So, every Euclidean domain is a PID. But the converse doesnt neces-
sarily hold: there are PIDs that arent Euclidean. In particular, the ring
Z[ ], where = 21 (1 + 19), is a PID but not a Euclidean domain.
The proof of this is a little involved and would be a bit of a digression
at this point, so we postpone it to Section 10.A.
PIDs dont satisfy all the properties that Euclidean domains do: were
not guaranteed that some version of the Division Theorem will hold.
Nevertheless, they are still quite civilised rings in which to work. In
particular, were guaranteed to have well-defined greatest common
divisors and least common multiples:
Wikimedia Commons / unknown artist
Proposition 10.30 If R is a PID then for any a, b R there exists ele- Claude Gaspard Bachet, Sieur de
ments d, l R such that d = gcd( a, b) and l = lcm( a, b). Mziriac (15811638) was born to a no-
ble family in Bourg-en-Bresse, at the
Also, there exist r, s R such that ra + sb = d. time part of the Duchy of Savoy. From
early childhood, he lived and studied
The second part of this proposition is known as Bzouts Identity, af- with the Jesuit Order, both his parents
ter the French mathematician Etienne Bzout (17301783), who proved having died by the time he was six
years old. He joined the order in 1601
it for polynomials in his 1779 work Thorie gnrale des quations al- but left the following year due to ill
gbriques. The corresponding result for integers had been proved ear- health and returned to his family es-
lier by the French mathematician, poet and classicist Claude Gaspard tate at Bourg-en-Bresse, where he lived
in comparative leisure for most of the
Bachet de Mziriac (15811638) in his 1612 book Problmes plaisants et rest of his life.
dlectables qui se sont par les nombres. Bachet was an accomplished poet and
writer, publishing poems in French,
Proof Let I = h ai + hbi = {ra + sb : r, s R}. This is an ideal of R, Italian and Latin, and translations of
the psalms and other religious works,
and since R is a PID, I must be principal, and hence equal to hdi for as well as translations of some of the
some d R. This element f is a greatest common divisor of a and works of Ovid (43BCc.17AD). He was
b: indeed, h ai hdi and hbi hdi, so by Proposition 10.14 we have a member of the literary salon that in
1634 became the Acadmie Franaise,
d| a and d|b. Furthermore, for any e R with h ai hei and hbi hei, although recurring health problems
it follows that hdi = h ai + hbi hei, so e|d, and hence d is a greatest prevented him from attending the in-
auguration ceremony and he was for-
common divisor of a and b. mally elected to membership the fol-
Since h ai + hbi = hdi, there exists some element of h ai + hbi equal to lowing year.
In addition to his literary works, he
d. That is, there exist elements r, s R such that ra + sb = d.
also wrote a number of books on math-
Similarly, h ai hbi is an ideal of R, and thus equal to hl i for some ematics, in particular Problmes plaisans
et delectables qui se sont par les nombres
l R since R is a PID. This l is a least common multiple of a and b: (1612). This book consisted of a collec-
indeed, hl i h ai and hl i hbi, so by Proposition 10.14 we have a|l tion of mathematical puzzles, mostly
and b|l. Furthermore, for any m R with hmi h ai and hmi hbi, in the form of arithmetical problems,
and was a forerunner of modern recre-
it follows that hmi h ai hbi = hl i and thus l |m. Hence l is a least ational mathematics books. He also
common multiple of a and b. published a Latin translation of the clas-
sic number theory text Arithmetica by
This can be extended to finite collections of elements: the Greek mathematician Diophantus
of Alexandria (c.200c.285AD). It was
Corollary 10.31 Let R be a PID. Then for any a1 , . . . , am R there exist in the margin of this edition that Pierre
elements d, l R with d = gcd( a1 , . . . , am ) and l = lcm( a1 , . . . , am ). de Fermat (16011665) wrote his cele-
brated Last Theorem.
Also, there exist elements r1 , . . . , rm R such that r1 a1 + + rm am = d.
410 a course in abstract algebra
What this says is not only that any two (or, actually, any finite collection
of) elements of a PID have a greatest common divisor, but also that
it can be written as a linear combination of those elements. We
call integral domains with greatest common divisors GCD domains,
and those that also satisfy the linearity condition are called Bzout
domains, but we will not study them further here.
In Proposition 10.23 we saw that in an integral domain, every prime
element is irreducible, but Example 10.24 showed that the converse
doesnt necessarily hold. However, it does hold in a PID, as the next
proposition shows.
Proposition 10.32 In a PID R, every irreducible element is prime.
Wikimedia Commons / unknown artist
The son and grandson of lawyers, Eti-
enne Bzout (17301783) instead pur- Proof Suppose that r R is irreducible, and that r | ab for some a, b
sued a career in mathematics, inspired R. Then by Proposition 10.30 there exists an element d R that is a
by the work of the Swiss mathemati-
greatest common divisor of r and a, and hence r = ds for some s R.
cian Leonhard Euler (17071783).
In the late 1750s he published memoirs Since r is irreducible, either d or s is a unit. Suppose that d is a unit.
on dynamics and integration, and in By Proposition 10.30 we can find x, y R such that d = xa + yr.
1758 was elected an adjoint member of
the Acadmie des Sciences. Over the Multiplying this by b we get db = xab + yrb. We know that r | ab and
next ten years he was appointed exam- obviously r |yrb, so r |db as well, and hence db = tr for some t R.
iner to the Gardes de la Marine and
the Corps dArtillerie, with responsibil-
But d is a unit, so we can multiply both sides by d1 to get b = d1 tr,
ity for the mathematical education of which means that r |b.
naval and army officer cadets.
If, on the other hand, s is a unit, we have r d and hence r |d. We
In addition to his contributions to
analysis, mechanics and algebra, he know already that d| a, since its a greatest common divisor of r and a,
also wrote a number of influential so therefore r | a.
textbooks, including the four-volume
Cours de mathmatiques lusage des Hence, if r is irreducible, either r | a or r |b, and thus r is prime.
Gardes du Pavillon et de la Marine (1764
1767) and the six-volume Cours com- Combining this with Proposition 10.23 we immediately get the follow-
plet de mathmatiques lusage de la ma-
rine et de lartillerie (17701782) which ing fact.
were used not only by officer cadets but Corollary 10.33 In a PID, an element is irreducible if and only if it is
also by students at the cole Polytech-
nique, and in translation by Harvard prime.
and other American universities.
Bzouts approach to research, which
And putting it together with Proposition 10.25 we find that in PIDs
he called the method of simplifying as- there is a stronger link between prime and maximal ideals.
sumptions was to attack special cases
of difficult problems, gradually devel- Corollary 10.34 A nontrivial ideal of a PID is prime if and only if it is
oping greater insight that often enabled maximal.
him to find a general solution.
Proof Let R be a PID and let I be some nontrivial ideal of R. Then
I = h ai for some a R because R is a PID.
Then I = h ai is prime if and only if a is prime in R (by Proposi-
tion 10.25), a is prime in R exactly when its irreducible in R (by
Proposition 10.32), and a is irreducible in R if and only if h ai is a
maximal principal ideal in R (by Proposition 10.25). But R is a PID, so
maximal principal ideals and maximal ideals are the same thing.
domains 411
So, in a PID irreducible elements and prime elements are the same,
and prime ideals and maximal ideals are the same as well.
13
But n is finite, so this sequence cant be infinite and must eventually
The earliest known proof appears
as Proposition 32 of Book VII of Eu- terminate. Thus n must be factorisable as a finite product of primes.
clids Elements; a more modern proof
appears in Gauss 1801 book Disquisi-
The following corollary is known as the Fundamental Theorem of
tiones Arithmetic. Arithmetic.13
domains 413
Many of the rings weve met so far are Noetherian. Examples of rings
that fail to satisfy the ACC include F [ x1 , x2 , . . .], the polynomial ring
with infinitely many unknowns, which contains infinite chains of the
form h x1 i h x1 , x2 i h x1 , x2 , x3 i , and the ring of algebraic
integers (that is, those complex numbers that are roots of polynomials
414 a course in abstract algebra
Proof Weve just shown that every PID satisfies the ACC for principal
ideals, and is hence a factorisation domain. By Proposition 10.32, in
a PID all irreducible elements are prime, and by Proposition 10.36
every factorisation domain where all irreducible elements are prime is
a UFD. Therefore every PID is a UFD.
In Proposition 10.30 we saw that greatest common divisors and least
common multiples exist in PIDs. This is also true in UFDs:
Proposition 10.44 Let R be a UFD, and let a and b be two nonzero
elements of R. Then there exist d, l R such that d is a greatest common
divisor of a and b, and l is a least common multiple of a and b.
f = p 1 . . . p k g1 . . . g m = q 1 . . . q l h 1 . . . h n
up1 . . . pk = q1 . . . ql
v( a + bi ) = | a + bi |2 = a2 + b2 .
By Proposition 10.4, the units in Z[i ] are exactly those elements a+bi
such that v( a+bi ) = a2 +b2 = 1 = v(1), namely 1 and i. This,
together with the multiplicativity of v, allows us to prove the following
fact:
Proposition 10.60 If r = a + bi Z[i ] and v(r ) is prime in Z, then r
is also prime in Z[i ].
domains 425
Proof Suppose that s|r in Z[i ]. Then v(s)|v(r ) in Z. But v(r ) is prime
in Z, so v(s) must equal either 1 or p. If v(s) = 1 then s is a unit
in Z[i ] by Proposition 10.4 and the above discussion. Otherwise, if
v(s) = p then s r. In either case, p must be prime.
The ring Z[i ] has a copy of Z embedded in it as a unital subring, so
its natural to ask what happens to the primes in Z when we consider
them as elements of Z[i ]. Are they still prime in Z[i ] or not? The
answer is: sometimes, but not always. For example 2 is prime in Z, but
we can factorise it as 2 = (1+i )(1i ) in Z[i ]. Similarly, 5 and 13 are
prime in Z, but factorise as 5 = (1+2i )(12i ) and 13 = (2+3i )(23i )
in Z[i ]. On the other hand, 3, 7 and 11 are prime in both Z and Z[i ].
There are a couple of patterns here: one is that the prime integers that
fail to be prime in Z[i ] seem to factorise as a product of a Gaussian
integer q with its conjugate q; the other is that (apart from 2, which is
often a special case) they are of the form 4n+1.
Well leave the second of these observations for a little while, but the
first leads to the following fact.
Proposition 10.61 Suppose p Z is prime in Z. Then either p is also
prime in Z[i ], or it factorises as p = qq, where q is prime in Z[i ].
Proof If p Z is not prime in Z[i ], then there must exist two Gaussian
integers q, r Z[i ] such that p = qr, with at least one of q and r prime
in Z[i ], and neither of them are units.
Suppose that q is a Gaussian prime. Then by the multiplicativity of v,
p2 = v( p) = v(qr ) = v(q)v(r ).
Then v(q) = v(r ) = p, so by Proposition 10.60 it follows that both q
and r must be prime in Z[i ]. Also, v(q) = qq = p = qr, so it must be
the case that r = q.
Now lets look at what happens to v(q) when q is prime in Z[i ]. We
know by Proposition 10.60 that if v(q) is prime in Z, then q must
be prime in Z[i ], but the converse doesnt necessarily hold: some
Gaussian primes q might yield non-prime values of v(q). Applying v
to some small Gaussian primes, we see the following:
v(1+i ) = 2, v(1+2i ) = 5, v(3) = 9, v(2+3i ) = 13,
v(1+4i ) = 17, v(2+5i ) = 29, v(7) = 49, v(11) = 121.
Again, various patterns are starting to form: v(q) is either a prime
integer, or the square of a prime integer. Also, in the cases when v(q)
is the square of a prime integer, q is an integer of the form 4n1, and
in the cases where q is a Gaussian prime with nonzero imaginary part,
v(q) is of the form 4n+1. Except, that is, for the special case where
v(q) = 2, but 2 is often a special case when discussing prime numbers.
426 a course in abstract algebra
quadratic integer rings that are Euclidean. Figure 10.2: Even and odd integers
Our motivating example of a Euclidean domain, as well as a lot of
the other algebraic structures weve investigated so far, is the ring
6 5 4 3 2 1 0 1 2 3 4 5 6
Z of integers. We can partition this into even and odd numbers: Figure 10.3: Multiples of 3
every element of Z is either even or odd, having the form 2n or 2n+1.
One way of looking at this is to say, as we did in the last chapter, that
Z/2Z = Z2 . But now were interested in the existence or nonexistence 6 5 4 3 2 1 0 1 2 3 4 5 6
v( a+bi ) = a2 +b2 .
Alternatively, every element in Z[i ] can be expressed in the form z+ a
where a h1+i i and z {0, 1, i }. Similarly, every element of Z
can be written as m+ a where a h2i and m = 0, 1; or in the form
m+b where b h3i and m = 0, 1.
This, well see in a moment, is a property shared by every Euclidean
domain R. Given a Euclidean valuation defined on R, we can find a
nonzero element r R such that the cosets of R/hr i are of the form
a+hr i where a is either 0 or a unit of R. Time for some terminology:
Definition 10.69 Let R be an integral domain, and let r R be such
that every coset in R/hr i can be represented as a+hr i where either
a = 0, or a is a unit of R. Then r is a universal side divisor of R.
This is the key concept we need: it turns out that a Euclidean domain
must have at least one universal side divisor.
Proposition 10.70 Let R be a Euclidean domain, and let v be a Euclidean
valuation defined on R. Then let a R be a nonunit element of R for which
v( a) is minimal over all nonunits in R. Then a is a universal side divisor.
Proof First we want to find the units in Z[], and for this we use
the norm N ( a + b) = a2 + ab + 5b2 . This is positive definite; that
is, N ( a + b) > 0 for all a, b Z[], and N ( a + b) = 0 only when
a = b = 0. We can show this by completing the square, which yields
the two expressions
2 2 19 2
2
a + 12 b + 19
4 b and 20 a +
1
a + 5b
2 5
for N ( a + b). The first of these is neater than the second, but both
are clearly positive definite. If b 6= 0 then
2 2 19 2
N ( a + b) = a + 12 b + 19 19
4 b > 4 b > 4 > 4.
domains 431
r2 + 19
= .
c2
r2 +19 1 19 7
If c > 5 then c2
6 4 + c2
6 9 < 1, so N is a DedekindHasse
norm.
r2 +19 23
If c = 5 then |r | 6 2 and so c2
6 25 < 1, and again N is a
DedekindHasse norm.
434 a course in abstract algebra
Summary
Exercises
10.1 List all the primitive polynomials of degree 3 in Z5 [ x ].
10.2 Show that the ring Z[ ] of Eisenstein integers is a Euclidean domain via the valuation v( a + b ) =
a2 ab + b2 .
10.3 Let F be a field, and suppose that f F [ x ] is irreducible. Show that F [ x ]/h f i is a field.
Im very well acquainted, too,
with matters mathematical,
I understand equations,
both the simple and quadratical,
About binomial theorem
Im teeming with a lot o news,
With many cheerful facts
about the square of the hypotenuse.
W S Gilbert (18361911) and
11 Polynomials Arthur Sullivan (18421900),
The Major Generals Song from
The Pirates of Penzance (1879)
a m m = a m 1 m 1 + + a 1 + a 0
and so
||m 1
m m 1
|| 6 N (|| + + | | + 1) = N .
|| 1
440 a course in abstract algebra
g = bk x k + + b1 x + b0 and h = c l x l + + c1 x + c0
Example 11.9 Let f = x4 + 6x2 + 16. Cohns original, base10 test Proposition 11.8 is a partial converse of
Bunyakovskys Conjecture, originally
is inapplicable, because the constant term 16 > 10. But we can stated in 1857 by the Russian math-
apply the generalised test and search for an integer b > 16 for which ematician Viktor Bunyakovsky (1804
1889), and which is still an open ques-
f (b) = b4 +6b2 +16 is prime. Again, this can be done very quickly
tion.
with a short computer program, and we find that b = 31 works, since Conjecture 11.10 (Bunyakovskys
f (31) = 929303 is prime. Hence f is irreducible over Z. Conjecture) Let f Z[ x ] be an irre-
ducible, primitive polynomial of positive
We can also use Cohns Criterion to find irreducible polynomials: degree, with positive leading coefficient.
Then there are infinitely many values of
Example 11.11 Expressing the prime p = 97 in different numerical n Z for which f (n) is prime.
bases b > 1, we obtain the following list of representations, which
correspond to irreducible polynomials in Z[ x ].
11000012 x6 + x5 + 1, 101213 x4 + x2 + 2x + 1,
12014 x3 + 2x2 + 1, 3425 3x2 + 4x + 2,
2416 2x2 + 4x + 1, 1667 x2 + 6x + 6,
1418 x2 + 4x + 1, 1179 x2 + x + 7,
9710 9x + 7, 8911 8x + 9, ...
the set of primitive nth roots of unity. We define the nth cyclotomic
polynomial
n = Pn ( x ) = gcd(k,n)=1 x e2ki/n .
(iv) Every nth root of unity is a primitive dth root of unity for some
unique d|n. So all the roots of x n 1 are roots of a cyclotomic polyno-
mial d for some d|n, and hence x n 1 = d|n d as claimed.
Not only is n monic, its coefficients are all integers. To prove this we
need the following fact.
Proposition 11.18 Let E and F be fields with F E, and suppose f , g
E[ x ] with f , f g F [ x ]. Then g F [ x ] too.
For example, Q(i ) has degree 2 over Q. The reason for this is that, as
noted earlier,
Q(i ) = { a + bi : a, b Q}.
It is not difficult to see that Q(i ) is a 2dimensional rational vector
space: the set {1, i } is a suitable basis. Similarly, Q( 2) also has
degree 2 over Q.
Example 11.29 The extension Q( 3, i ) : Q has degree 4. Recall that
Q( 3, i ) = Q( 3)(i ) = { p(i )/q(i ) : p, q Q( 3)[ x ]}.
By a rather tedious calculation it can be shown that
Q( 3, i ) = { a + b 3 + ci + di 3 : a, b, c, d Q}. Wikimedia Commons
Carl Louis Ferdinand von Linde-
The obvious basis for this is {1, 3, i, i 3}, which has four elements, mann (18521939) was born in
and hence [Q( 3, i ) : Q] = dimQ Q( 3, i ) = 4. Hanover, the son of a schoolteacher.
He studied mathematics at Gttingen,
Also, note that Q( 3, i ) = {( a+b 3) + (c+d 3)i : a, b, c, d Q}, Erlangen and Munich, completing his
and so, considering Q( 3, i ) as a vector space over Q( 3) we see doctorate on non-Euclidean geometry
in 1873, under the supervision of Felix
that [Q( 3, i ) : Q( 3)] = dimQ(3) Q( 3, i ) = 2. Klein (18491925).
His main contributions to mathematics
In the above example it turns out that were in geometry and analysis, and
perhaps his best known achievement is
[Q( 3, i ) : Q] = 4 = 22 = [Q( 3, i ) : Q( 3)][Q( 3) : Q]. his proof, published in 1882, that is
transcendental. This result was based
This is true in general: the degree is multiplicative, in much the same on Charles Hermites proof that e is
transcendental, together with Eulers
way that the index of a subgroup is.8 The formal statement of this fact identity ei = 1.
is sometimes called the Tower Law. In 1883 he was appointed to a chair
Proposition 11.30 (The Tower Law) Let E : K and K : F be finite field at the University of Knigsberg, and
while there he supervised the doctoral
extensions. Then dissertation of David Hilbert (1862
1943).
[ E : F ] = [ E : K ][K : F ].
8
Proposition 2.32, page 55.
Proof Suppose that [ E : K ] = m and [K : F ] = n. Then K is an n
dimensional vector space over F, and so we can find a basis
A = { a1 , . . . , a n }.
for 1 6 i 6 n. Hence
n m
c= i,j (ai bj )
i =1 j =1
and therefore A B spans E over F.
To show linear independence, suppose that
n m
i j ( a i b j ) = 0
i =1 j =1
where i j F. Then
n m
i j b j ai = 0
i =1 j =1
It turns out that all simple transcendental field extensions are of this
form, although to properly state and prove that fact we first need to
formulate a concept of equivalence of field extensions. Recall that an
extension of a field F is an embedding of F into some larger field E.
Given two such inclusions i1 : F , E1 and i2 : F , E2 , what were
looking for is an isomorphism : E1 E2 that respects the embedded
images of F.
450 a course in abstract algebra
9
Example 9.20, page 369. Proof Let : F ( x ) F () be the evaluation homomorphism ev ;9
that is, ( f /g) = f ()/g() for any rational function f /g F ( x ).
Since g is a nonzero polynomial in F [ x ] and is transcendental over
F, it follows that g() 6= 0, so is a well-defined homomorphism.
Furthermore, if f ()/g() = 0 then we must have f () = 0, which
can only happen if f () = 0, again because is transcendental over
F. Therefore ker = {0} and so is injective; it is clearly surjective
as well, and is hence an isomorphism : F ( x ) F (). The restriction
| F is the identity, so is an isomorphism of extensions.
where the first factor has roots 2i and the second factor has roots
polynomials 451
3 i. So if were only interested in recovering , we dont need to
bother with the first factor x2 + 4. In fact, f = x4 4x2 + 16 happens
to be the lowest-degree polynomial in Q[ x ] with as a root. We saw
in Example 11.7 that this polynomial is irreducible over Z and Q.
Proposition 11.38 If is algebraic over a field F, then there is a unique
monic, irreducible polynomial (the minimal polynomial) of minimal de-
gree in F [ x ] such that f () = 0, and f | g for any other g F [ x ] with
g() = 0.
(Recall that a polynomial is monic if its highest-degree coefficient is 1.)
F [ x ]/h f i = F [ x ]/ ker
= im = F ()
as claimed. This isomorphism is the induced map b: F [ x ]/h f i F (),
which maps x +h f i to and any constant element a+h f i to a F ().
Its inverse b1 : F () F [ x ]/h f i is the required isomorphism.
The next example is slightly more complicated, and is one that well
return to later when we discuss the Galois correspondence.
Example 11.50 Lets investigate the
splitting field for f = x3 5
3
over Q. Set = 5 and = 2 + 2 i. Then the roots of f are ,
3 1
We can use this to give the following test for distinct roots:
Proposition 11.59 Let F be a field, and suppose that f F [ x ]. Let L be
a splitting field for f over F. Then the roots of f in L are distinct if and
only if f and d f have no non-constant common factor.
n
Proof Let L be the splitting field of f = x p x over F p . This field has
n
characteristic p, and hence the formal derivative d f = pn x p 1 1 =
1, which has no non-constant factors in common with f . So, by
Proposition 11.59, f has no repeated roots.
Let K be the set consisting of these distinct roots; there are deg( f ) = pn
of them. We claim that K is a subfield of L. Certainly it contains 0 and
polynomials 461
n n
1, since 0 p 0 = 0 and 1 p 1 = 0. Also, for any a, b K we have
n n n
( a + b) p = a p + b p = a + b
by the Freshmans Binomial Theorem,14 and 14
Proposition 8.20, page 332.
pn pn pn
( ab) =a b = ab
n n
and ( a 1 ) p = ( a p ) 1 = a 1 ,
so K is closed under addition, multiplication and inversion. This
subfield K is itself the splitting field of f over F p : it contains all the
roots of f , but no proper subfield of K does.
Hence, for any prime p and positive integer n there exists a field of
order pn with characteristic p. By Proposition 11.57, this field is the
n
splitting field of f = x p x over F p , and by Proposition 11.52 any
other such field is isomorphic to K.
Definition 11.61 For any prime p and positive integer n, the finite
field of characteristic p and order q = pn discussed above is the
Galois field or finite field of order q. We denote it Fq or GF(q).
I have no doubt that an author never 11.4 Field automorphisms and the Galois group
harms his readers more than when he
hides a difficulty.
variste Galois (18111832),
It should now be clear that there is a strong connection between
Deux mmoires dAnalyse pure (1831) solving polynomial equations and studying field extensions. In this
section we will work out the details of this connection, using tech-
niques pioneered by the French mathematician variste Galois (1811
1832) in the few years before his death in a duel at the age of 20.
The approach well take is one that has proved very fruitful on many
occasions before: we will use group theory to study the symmetry of
the objects under investigation, which in this case are field extensions.
We must first decide what the correct notion of symmetry is.
The answer is given by our earlier discussion of simple algebraic field
extensions. Proposition 11.44 in particular says that if and have
the same minimal polynomial over some field F, then the extensions
F () : F and F ( ) : F are Fisomorphic. When studying the internal
structure of a field extension E : F, we want to see how many ways we
can get E by extending F. Equivalently, we want to find all the different
ways of mapping some field E to itself that both respects the field
structure of E and keeps the embedded copy of F fixed. The concept
were heading inexorably towards is that of an Fautomorphism: an
Fisomorphism from E to itself.
It is straightforward to show that the automorphisms of a field E form
a group Aut( E), and it is almost as straightforward to show that the
Wikimedia Commons
Fautomorphisms of E also form a group, a subgroup of Aut( E).
variste Galois (18111832)
Definition 11.63 Let E : F be a field extension. The Galois group
Gal( E : F ) is the group of Fautomorphisms of E; it is a subgroup of
the full automorphism group Aut( E).
other roots from 1 and the required splitting field is Q(1 ). The
polynomial f is irreducible (by Eisensteins Criterion) and monic,
and is hence the minimal polynomial for the extension Q(1 ) : Q,
which must therefore have degree 4.
Suppose that G, and that (1 ) = 2 . Then
( 5) = 2(1 )2 5 = 222 5 = 5,
and hence
(2 ) = ( 5/1 ) = 5/2 = 1 = 3 .
Furthermore,
(3 ) = (1 ) = 2 = 4 , (4 ) = (2 ) = 3 = 1 .
So , and for that matter any other Qautomorphism that maps 1
to 2 , cyclicly permutes the four roots. Therefore hi
= Z4 . Any
Qautomorphism mapping 1 to 4 must similarly be equal to 1 .
Now consider a Qautomorphism that maps 1 to 3 = 1 . Then
( 5) = 2(1 )2 5 = 221 5 = 5
and so
(2 ) = ( 5/1 ) = 5/1 = 2 = 4 .
Furthermore, (3 ) = 1 and (4 ) = 2 , so = 2 .
These are the only possibilities, and so G
= hi
= Z4 .
Its people like that who make you re- 11.5 The Galois Correspondence
alise how little youve accomplished. It
is a sobering thought, for example, that
when Mozart was my age, he had been Its illuminating to look at subgroups of the Galois group of a field
dead for two years. extension E : F, and see how the various Fautomorphisms act on the
Tom Lehrer,
various subfields of E. In Example 11.66 we saw that
introduction to Alma,
from: That Was The Year That Was
Gal(Q( 2, 3) : Q) = {1 = id, 2 , 3 , 4 }
= Z2 Z2 .
(1965)
Every element of this group fixes the base field Q. Every element of the
subgroup {1 , 2 } fixes the subfield Q( 2), but 2 acts nontrivially
on Q( 3) so the subgroup doesnt fix this subfield. Similarly, {1 , 3 }
{1 , 2 , 3 , 4 } leaves Q( 3) invariant but not Q( 2). Neither Q( 2) nor Q( 3)
are fixed by {1 , 4 }, but 6 = 2 3 is mapped to itself by both 1
{1 , 2 } {1 , 3 } {1 , 4 } and 4 , so the subfield Q( 6) is fixed by this subgroup. Finally, the
trivial subgroup {1 } leaves Q( 2, 3) invariant.
{1 } Figure 11.1 shows the subgroup lattice for Gal(Q( 2, 3) : Q), with
Figure 11.1:Subgroup
lattice diagram inclusions
going up the page. Figure 11.2 shows the subfield lattice
for Gal(Q( 2, 3) : Q) = Z2 Z2 for Q( 2, 3), this time drawn upside down, with inclusions going
polynomials 467
down the page. Well come to the reason for this reversal soon,
but for the moment observe the striking and suggestive similarity Q
16 Proposition 8.20, page 332. by the Freshmans Binomial Theorem,16 and so x p has only one
root in F p ( ), and indeed in any other extension of F. Therefore
Gal(F p ( ) : F p ()) is trivial, and hence G(F p ()), the group of all
F p ()automorphisms of F p ( ), must be trivial. Applying F to
this we get the field consisting of every element of F p ( ) fixed
by the identity F p ()automorphism, which is all of F p ( ). So
F (G(F p ())) = F p ( ) and hence F and G arent inverses of each
other.
The problem here is that x p only has a single root in the splitting
field F p ( ); moreover we can adjoin as many other elements to F p ( )
as we like, and it will still only have a single root, because x p =
( x ) p over F p . We have a single root, repeated p times.
We want to find a well-defined characterisation of nice field extensions
or polynomials where neither of these awkward situations arise. We
want to know what criteria an extension E : F must satisfy to ensure
F and G are inverses of each other, so that there is a neat and tidy
correspondence between intermediate subfields of E : F and subgroups
of Gal( E : F ).
Before we get to that, we have a few useful facts about F and G to
prove. The first of these says that while F and G arent necessarily
inverses, we do at least have K F (G(K )) and H 6 G(F ( H )) for any
intermediate subfield K and subgroup H.
Proposition 11.77 Let K be an intermediate subfield of an extension E : F
and suppose H 6 Gal( E : F ). Then K F (G(K )) and H 6 G(F ( H )).
rank( A) + nullity( A) = n.
h1 ( x1 ) y1 + + h m ( x1 ) y m = 0
.. (11.6)
.
h1 ( x n ) y1 + + h m ( x n ) y m = 0
a = a1 x1 + + a m x m . (11.7)
Multiplying the equations in (11.6) by the coefficients a1 , . . . , am re-
spectively, we get:
a1 h1 ( x1 ) y1 + + a1 h n ( x1 ) y n = 0
.. (11.8)
.
a m h1 ( x m ) y1 + + a m h n ( x m ) y n = 0
h1 ( a1 ) h1 ( x1 ) y1 + + h n ( a1 ) h n ( x1 ) y n = 0
.. (11.9)
.
h1 ( a m ) h1 ( x m ) y1 + + h n ( a m ) h n ( x m ) y n = 0
polynomials 471
and then as
h1 ( a1 x1 ) y1 + + h n ( a1 x1 ) y n = 0
.. (11.10)
.
h1 ( a m x m ) y1 + + h n ( a m x m ) y n = 0
Adding all of these equations together and using (11.7) we get
y1 h1 ( a) + + yn hn ( a) = 0. (11.11)
This is true for any a E, and since the elements y1 , . . . , yn arent
all zero, this means that the automorphisms h1 , . . . , hn are linearly
dependent, contrary to Lemma 11.79. Hence our original assertion,
that | H | > [ E : F ( H )], cant be true, so instead | H | 6 [ E : F ( H )].
Now suppose | H | = n < m = [ E : F ( H )]. Let T = { x1 , . . . , xn+1 } S
be a set of n+1 linearly independent vectors in E regarded as an m
dimensional vector space over F ( H ). Consider the m(n+1) matrix
h1 ( x1 ) ... h1 ( xn+1 )
B= .. ..
. .
hn ( x1 ) ... hn ( xn+1 )
First well consider the failure mode seen in Example 11.75. The issue
there was that x3 5 is the minimal polynomial over Q for = 3 5,
but it doesnt split in Q(). We can avoid this situation if we restrict
our attention to extensions in which every polynomial splits:
Definition 11.80 A field extension E : F is normal if and only if
every irreducible polynomial in F [ x ] that has at least one root in E
splits completely over E.
This highlights one of the recurring themes in this chapter: the close
interrelationship between polynomials, algebraic elements and field
extensions. Well now look at some examples illustrating the concept
of separability in each of these contexts.
Example 11.90 The polynomial x2 +1 Q[ x ] is separable over Q,
since it has no repeated roots in its splitting field Q(i ): it factorises
as ( x +i )( x i ).
polynomials 477
Example 11.91 The element 2 Q( 2) is separable over Q, since
its minimal polynomial x2 2 Q[ x ] has no repeated roots in its
splitting field Q( 2): it factorises as ( x + 2)( x 2).
p
is an element bi F such that bi = ai . Hence
p p p
f = b0 + b1 x p + + bn x np
= (b0 + b1 x + + bn x n ) p
which is reducible. So if ( F ) = F then F [ x ] contains no inseparable
polynomials, and hence F is perfect.
Conversely, suppose that ( F ) 6= F. Then choose some element
a F \ ( F ), and consider the polynomial f = x p a F [ x ]. Suppose
that p = a. Then x p a = x p p = ( x ) p by the Freshmans
Binomial Theorem again, and so f has only a single repeated root
in any splitting field over F. All that remains is to show that this
polynomial is irreducible over F. If not, then any proper, nontrivial
monic factor of f must have the form ( x )m for some m such that
0 < m < p. The coefficient of x m1 in ( x )m is m, so for x p a
to be reducible over F we require m F. But m F p F, so
this means that F too, and hence a = p ( F ), which is a
contradiction. So if ( F ) 6= F, there exists at least one irreducible
polynomial x p a that is inseparable over F, and therefore F isnt
perfect.
Corollary 11.100 Finite fields are perfect.
f () = ()|K = |K |K = f () f (),
Gal(K : F ) = im( f )
= Gal( E : F )/ ker( f ) = Gal( E : F )/G(K )
as claimed.
484 a course in abstract algebra
which is very nearly what we want: its just the constant term that
needs addressing. So
b2
x2 + ba x + c
a = (x + b 2
2a ) 4a2
+ c
a = 0,
which we can then rearrange to get
b 2 b2 4ac
(x + 2a ) = 4a2
.
Taking square roots and solving for x we get the familiar quadratic
formula
b b2 4ac
x= .
2a
Explicit formulae for solving cubic and quartic equations were devel-
oped in Italy during the 16th century, amid an environment of fiercely
counterproductive competition. These formulae are rather cumber-
some, but they all only require basic arithmetic operations addition,
subtraction, multiplication and division) together with finite roots.
This isnt true for all polynomial equations, and to understand why
will involve the algebraic machinery developed in this chapter together
with some of the advanced group theory we studied in Chapter 7.
We start by identifying a special class of algebraic field extensions.
Definition 11.109 A field extension E : F is a radical extension if
there exists a sequence
F = K0 , K1 , . . . , K m = E
of fields such that Ki+1 = Ki (i ) where i is a root of a polynomial
f i of the form x ni ai in Ki [ x ], with 0 6 i < m and ni N.
polynomials 485
What this means in practice is that radical extensions are exactly those
of the form F () : F where can be constructed from elements of F
using just the four basic arithmetical operations and finite roots.
Example 11.110 Let = (1+ 3)1/5 3 5( 12 + 2)1/7 . Then Q() : Q
is a radical extension, since
K0 = Q; K1 = K0 ( 0 ) , f 0 = 20 3;
K2 = K1 ( 1 ) , f 1 = 51 (1+0 ); K3 = K2 ( 2 ) , f 2 = 32 5;
K4 = K3 ( 3 ) , f 3 = 23 2; K5 = K4 ( 4 ) , f 4 = 74 (2+3 ).
By the Tower Law, [Q() : Q] = 4i=0 ni = 25327 = 420.
for all a E.
The first important result we will prove is due to David Hilbert (1862
486 a course in abstract algebra
26
This result is often referred to as 1943) and Ernst Kummer (18101893).26
Hilberts Theorem 90, because it is the
Proposition 11.115 (Hilberts Theorem 90) Let E : F be a cyclic ex-
ninetieth theorem stated in Zahlbericht,
his 1897 report on algebraic number tension, and a generator of Gal( E : F ). Then for any a E, the norm
theory; it had, however, been proved in NE : F ( a) = 1 if and only if there exists some b E with a = b/ (b).
1855 by Ernst Kummer.
Proof Figure 11.4 depicts the subfield lattice for this situation.
E Since F L and L E it follows that E is an extension of F. By
Definition 11.47, K is the smallest possible field over which f F [ x ]
L K splits completely, so any other field over which f splits completely
must contain a subfield isomorphic to K. Hence we can regard E as
LK an extension of K.
Let Gal( E : L) be some Lautomorphism of E. The restriction
F = |K is a field monomorphism from K to E. As the restriction of
Figure 11.4: Subfield lattice for Propo- an Lautomorphism, must necessarily fix all the elements of L that
sition 11.118 lie in K, so it is a (K L)monomorphism. We want to show that this
is actually a (K L)automorphism of K.
By Proposition 11.69, , as an Lautomorphism of E, permutes the
roots 1 , . . . , n of f in E = L(1 , . . . , n ). Its restriction also per-
mutes all of these roots in K = F (1 , . . . , n ), while fixing all the
elements of F. Hence is an automorphism of K that, as previously
noted, fixes not just F but K L, so Gal(K : K L).
Restriction to K yields a map : Gal( E : L) Gal(K : K L) such that
() = |K , and we now want to show that this is an isomorphism. It
is a homomorphism, since for any 1 , 2 Gal( E : L) we have
(1 2 ) = (1 2 )|K = (1 |K )(2 |K ) = (1 ) (2 ).
{id} = G0 C G1 C C Gk = G
ing sequence
F = Fk Fk1 F1 F0 = E
of subfields of E, with G( Fi ) = Gal( E : Fi ) = Gi and Gal( Fi1 : Fi ) =
Gi /Gi1 = Zmi , so Fi1 : Fi is a cyclic extension of degree mi for
0 < i 6 k.
Furthermore, Fi contains Fk = E, which contains all mth roots of unity.
Since mi |m, it follows that E, and hence Fi , also contains all mi th roots
of unity, since these are also powers of . By Proposition 11.117 there
exists some element i Fi such that Fi = Fi+1 (i ) is the splitting field
m
over Fi+1 of some polynomial gi = x mi ai Fi+1 [ x ] where ai = i i .
Therefore f is soluble by radicals.
Case 2 Now suppose that F doesnt contain a primitive mth root of
unity. The polynomial x m 1 F [ x ] is obviously soluble by radicals.
Let be a primitive mth root of unity, and adjoin it to F to get
L = F ( ), the splitting field of x m 1 over F. Now let E be the
splitting field of f over L. By Proposition 11.118, E is an extension of
K, and Gal( E : L) = Gal(K : LK ), which is a subgroup of Gal(K : F ),
which is soluble. Proposition 7.38 says that subgroups of soluble
groups are soluble, and hence Gal( E : L) is soluble. This now reduces
to Case 1, and therefore f is again soluble by radicals.
The converse is also true: polynomials that are soluble by radicals
have soluble Galois groups.
Proposition 11.119 Let f F [ x ] be a polynomial over a field F of char-
acteristic zero. If f is soluble by radicals, then its Galois group Gal( f , F )
is soluble.
To prove this we first need the following result.
Proposition 11.120 Let F be a field of characteristic zero, and let f =
x n a F [ x ]. The Galois group Gal( f , F ) is soluble.
n
Now suppose k > 1. Again, let a = 1 1 , and let K be the splitting field
of g = x n1 a over F. Let L be the splitting field of g over E. Then
F E K L. Furthermore, L is the splitting field of g f over F, and
of f over K. Since F (1 ) K it follows that f splits in K (2 , . . . , k ).
That is, since f F [ x ] and F K, we can regard f as an element
of K [ x ]. Its splitting field over K is contained in K (2 , . . . , k ), a rad-
ical extension of K by (k1) adjoined elements. By the induction
hypothesis, this means that Gal( f , K ) = Gal( L : K ) is soluble.
By Proposition 11.120, Gal( g, F ) = Gal(K : F ) is soluble. By part (iv)
of the Fundamental Theorem of Galois Theory,
Gal( g, F ) = Gal(K : F )
= Gal( L : F )/ Gal( L : K )
and since Gal( L : K ) is soluble, Proposition 7.40 says that Gal( L : F ) is
soluble, as an extension of soluble groups.
Then by the Fundamental Theorem of Galois Theory again,
Gal( f , F ) = Gal( E : F )
= Gal( L : F )/ Gal( L : E)
and by Proposition 7.39 this is soluble as a quotient of a soluble group.
This completes the induction step, and thereby the proof.
Combining Propositions 11.112 and 11.119 we get the following:
Theorem 11.121 Let F be a field of characteristic zero. Then a polynomial
f F [ x ] is soluble by radicals if and only if Gal( f , F ) is soluble.
This doesnt mean that f has no roots, nor that it doesnt have a
polynomials 493
11.A Geometric constructions The Academy has taken, this year, the
decision to no longer examine any solu-
tion of the problems of the doubling of
In this section we will look at some applications of Galois Theory the cube, of the trisection of the angle,
to classical geometry. In particular, we will identify what geometric or of the squaring of the circle, nor any
announcement of a perpetual motion
constructions are possible using just an unmarked ruler and a pair machine.
of compasses. As ancient Greek mathematicians discovered, we can Histoire de lAcadmie Royale des
achieve a surprisingly wide range of geometric tasks with just these Sciences (1775) 61
tools.
Example 11.125 (Bisecting an angle) Given an arbitrary angle , we B
can construct an angle 2 with straightedge and compasses as shown Q
in Figure 11.5. R
In the middle of the 19th century, all three of these were finally proved
36
36
Impossible with just straightedge to be impossible, and we will study the reasons why now.
and compasses, that is: the latter two
are possible using slightly more sophis-
Well do this by developing a characterisation of those points in the
ticated techniques. Solutions involv- plane R2 that can be obtained just by using an unmarked ruler and a
ing the intersection of conic curves or pair of compasses, and then using this to construct a hierarchy of field
marked rulers were known to Greek
mathematicians in the fourth century extensions of Q.
BCE. More recently, solutions have First we choose a coordinate system for R2 : an origin O and two lin-
been devised using origami, linkages
and similar mechanisms. early independent vectors e1 and e2 , which determine the coordinate
axes. Then any other vector v in R2 can be expressed uniquely as a
linear combination v = ae1 + be2 , and this corresponds to the point
with coordinates ( a, b).
Suppose that P is a set of points in R2 . We consider the following two
operations:
(L) Draw a line through any two distinct points in P.
(C) Draw a circle centred on some point in P, with radius equal to
the distance between any two distinct points in P.
Let C ( P) be the set of points in R2 that can be obtained as the inter-
section of two lines, two circles, or a line and a circle, resulting from
applying operations (L) and (C) to the points in P. The points in C ( P)
are said to be constructible in one step from P.
Now let A0 = {(0, 0), (1, 0)}. We recursively define a sequence of sets
A0 , A1 , A2 , . . . of points in R2 such that
A i = A i 1 C ( A i 1 ).
That is, each set Ai consists of the points in Ai1 together with all
the points which are constructible in one step from Ai1 . We say any
point in R2 is constructible if it lies in An for some n; that is, if it is
constructible in finitely many steps from A0 . Similarly, we say that an
angle is constructible if the point (cos , sin ) is constructible.
The set A1 is constructed from A0 by means of the operations (L) and
(C). The only possible line at this stage is the horizontal axis, passing
polynomials 495
through (0, 0) and (1, 0). There are two possible circles that can be
drawn, one centred at (0, 0) and the other at (1, 0), both with radius 1.
These intersect at the points ( 2 , 23 ), so
1
3 3
A1 = {(0, 0), (1, 0), ( 21 , 1
2 ), ( 2 , 2 )}.
The circles and lines passing through these four points yield several
more, the elements of A2 , and by continuing this process indefinitely
To translate this into an algebraic context and use the machinery
developed in this and preceding chapters, we define a tower of fields
F0 F1 Fn R
by setting each Fn to be the subfield of R generated by the coordinates
of the points in An . For each point ( x, y) in An \ An1 we adjoin both
x and y to the field Fn1 . So F0 is the field generated by the elements
0 and 1, the coordinates of the points in A0 = {(0, 0), (1, 0)}; thus
F0 = Q. To get F1 we adjoin the elements 12 and 23 , the coordinates of
the points in A1 \ A0 , which yields F1 = Q( 3).
This new field F1 is an algebraic extension of F0 = Q with degree
[Q( 3) : Q] = 2. In fact, each successive extension Fn : Fn1 (and, by
the Tower Law,37 any extension Fn : Fm ) has degree 2k for some k N: 37
Proposition 11.30, page 447.
Proposition 11.126 Let Fn be the subfield of R generated by the coor-
dinates of the constructible points in the set An discussed above. Then
[ Fn : Q] is a power of 2 for all n N.
Corollary 11.128 Not all angles can be trisected using just ruler and Pierre Laurent Wantzel (18141848)
demonstrated a strong aptitude in
compasses. mathematics from an early age. At
the age of 14, he entered the Col-
Proof As a counterexample, let = 3 . To trisect this angle requires lge Charlemagne, having been tutored
us to construct = /3 = 9 , or equivalently the length y = cos . in Latin and Greek by a M. Lievyns,
whose daughter he later married.
Using the trigonometric identity While there, he edited a new edition
of the Trait dArithmetique of Antoine-
cos 3 = 4 cos3 3 cos Andr-Louis Reynaud (17711884) and
1 Etienne Bezout (17301783), and won
and the fact that cos 3 = cos 3 = 2 this gives first prize in French and Latin.
8y3 6y 1 = 0. In 1832 he won first place in the en-
trance exams for both the cole Poly-
Setting x = 2y this yields technique and the science section of the
cole Normale Superieure, which no-
x3 3x 1 = 0 body had done before. He excelled in
his studies and in 1834 began training
and so this particular trisection problem is equivalent to constructing as an engineer at the cole des Ponts
a length satisfying this equation. But the polynomial x3 3x 1 et Chausses. He concluded that he
would be at best a mediocre engineer,
is irreducible over Q, again by Eisensteins Criterion, the extension and much preferred teaching mathe-
Q() : Q has degree. [Q() : Q] = 3, which is not divisible by 2. By matics.
Proposition 11.126 no such construction exists. In 1838 he was appointed Professor
of analysis at the cole Polytechnique,
Some angles can be trisected in this way, however. If = 2 then and three years later Professor of ap-
plied mechanics at the cole des Ponts
the question concerns the polynomial 4x3 3x which factorises as et Chausses. In addition, he over-
x (4x2 3) and so the corresponding field extension is Q( 3) : Q. This saw the entrance exams at the cole
Polytechnique and taught mathematics
has degree 2, and 2 can be trisected in the required manner.
and physics in various schools around
The other classical problem, that of squaring the circle, requires the Paris.
fact that is not algebraic over Q, a fact originally proved in 1882 by His mathematical contributions in-
clude important work on the solution
the German mathematician Ferdinand Lindemann (18521939). of equations by radicals, and the pos-
sibility of ruler and compass construc-
Corollary 11.129 A circle cannot be squared using just ruler and com-
tions. In addition, with his friend and
passes. colleague Jean Claude de Saint-Venant
(17971886), he published three papers
Proof Without loss of generality, consider a circle of unit radius. This on air flow.
has area units, and so a square of the same area must have side He died at 33, apparently from over-
work. In his obituary, Saint-Venant re-
. To construct such a square requires the point (0, ) to be marked: He usually worked in the
constructible from A0 = {(0, 0), (1, 0)} in finitely many steps. By evenings, not going to bed until late;
then he read, sleeping restlessly for
Proposition 11.126 this requires the extension Q( ) : Q to be alge-
only a few hours, alternately abus-
braic with degree a power of 2. But Q( ) Q( ), and is known ing coffee and opium and, until his
not to be algebraic over Q, so no such construction exists. marriage, taking his meals at irregular
times. He trusted without measure in
Proposition 11.126 gives a necessary condition for constructibility. With his constutition, which was naturally
very strong, and which he subjected
a little more work we can show that this condition is also sufficient. with pleasure to all kinds of abuse. He
First we need to prove some preliminary results on constructible points. brought sadness to those who mourn
We begin with some standard ruler and compass constructions. his untimely death.
P Proof Draw circles centred at A and B with radius AB. These circles
intersect at two points P and Q. The line through P and Q is perpen-
dicular to AB, and intersects it at the midpoint C. (See Figure 11.6.)
Proposition 11.131 Given points A, B and C, we can construct a line
through C perpendicular to AB.
A C B
Proof Draw a circle centred at C that intersects AB at points P and Q.
Draw the perpendicular bisector of PQ as in Proposition 11.130. (See
Figure 11.7.)
(Although Figure 11.7 shows the point C off the line AB, this construc-
tion also works if C happens to lie on AB.)
Q Proposition 11.132 Given points A, B and C, we can draw a line parallel
to AB, passing through C.
Figure 11.6: The perpendicular bisector
Proof Draw the line through A and B. Drop a perpendicular from C
C to this line, as in Proposition 11.131, meeting it at a point D. Draw
the line through C and D, and construct the perpendicular line to CD
passing through C. This line is parallel to AB. (See Figure 11.8.)
A B Proposition 11.133 Suppose that a, b R. Then
P Q (i) the point ( a, 0) is constructible if and only if (0, a) is;
(ii) the point ( a, b) is constructible if and only if ( a, 0) and (b, 0) are;
Figure 11.7: Dropping a perpendicular
from C to AB (iii) if ( a, 0) and (b, 0) are constructible, then so are ( a+b, 0), ( ab, 0),
( ab, 0) and, if b 6= 0, ( a/b, 0); and
C
(iv) if a, b Q then ( a, b) is constructible.
Proof Given A0 = {(0, 0), (1, 0)}, we can construct the xaxis by
drawing the line through (0, 0) and (1, 0). By Proposition 11.131 we
A B can construct the line passing through (0, 0) perpendicular to this line;
D that is, the yaxis.
Figure 11.8: Constructing a parallel (i) Given A = ( a, 0), the circle with radius a centred on (0, 0) inter-
sects the yaxis at P = (0, a) (and (0, a)). (See Figure 11.9.)
P Conversely, given (0, a), the same circle intersects the xaxis at ( a, 0)
and ( a, 0).
(ii) Given B = (b, 0), construct (0, b) using the method in part (i).
Draw the line passing through A = ( a, 0) parallel to the yaxis and the
line through P = (0, b) parallel to the xaxis, as in Proposition 11.132.
These lines intersect at Q = ( a, b).
A
Figure 11.9: Constructing (0, a) from Conversely, given Q = ( a, b), draw the line through ( a, b) parallel to
( a, 0) and vice-versa the xaxis; this intersects the yaxis at P = (0, b), and from this we
can construct B = (b, 0) via the method in part (i). Drawing a line
through ( a, b) parallel to the yaxis yields the point A = ( a, 0) as its
intersection with the xaxis. (See Figure 11.10.)
polynomials 499
Draw the line through I = (0, 1) and B = (b, 0), and then construct Figure 11.10: Constructing ( a, b)
the parallel line passing through A = ( a, 0): this intersects the yaxis
at P = (0, a/b), and by part (i) we can then construct Q = ( a/b, 0).
(See Figure 11.13.)
(iv) By part (iii) we can construct (n, 0) for any n Z, and then
(m/n, 0) for any m, n Z with n 6= 0. We can thus construct ( a, 0)
and (0, b) for any a, b Q, and hence ( a, b) by part (ii). Q B A P
Figure 11.11: Constructing ( a+b, 0)
Proposition 11.134 Suppose that a R. If the point ( a, 0) is con-
and ( ab, 0)
structible, then so is ( a, 0).
Summary
N G de Bruijn, A solitaire game and its relation to a finite field, Journal of Recreational Mathematics 5.2
(1972) 133137
D A James, Magic circles, Mathematics Magazine 54.3 (1981) 122125
We saw in Section 11.A that the classical problems of trisecting an arbitrary angle, doubling the cube,
and squaring the circle are impossible with just ruler and compasses. This has been known since the
middle of the nineteenth century. Nevertheless, well-meaning enthusiasts still persist in trying to solve
them, with sometimes exasperating results. The following articles and books give fascinating accounts
of many such attempts.
A De Morgan, A Budget of Paradoxes, Longman, London (1872)
U Dudley, Mathematical Cranks, Mathematical Association of America, Washington, DC (1992)
U Dudley, The Trisectors, Mathematical Association of America, Washington, DC (1994)
A E Hallerberg, Indianas Squared Circle, Mathematics Magazine 50.3 (1977) 136140
D Singmaster, The Legal Values of Pi, Mathematical Intelligencer 7.2 (1985) 6972
Exercises
12.1 Use Proposition 11.5 together with the prime p = 29, in a manner analogous to the method used
in Example 11.11, to find a list of quadratic and higher-degree irreducible polynomials in Z[ x ].
12.2 Show that the extension Q( 2, 3) is simple.
12.3 Adapt the argument in Example 11.92 to show that if [ E : F ] = 2 and char( F ) 6= 2, then E : F is
separable. What happens if char( F ) = 2?
We have offended against thy holy
lawes: We have left undone those
thinges whiche we ought to have done,
and we have done those thinges which
we ought not to have done, and there is
no health in us, but thou, O Lorde, have
mercy upon us miserable offendours.
The Book of Common Prayer (1559)
wider audience, to provide a quick reminder for those who might need
one, and also to standardise notation and terminology. In particular,
we cover sets, infinite cardinals, functions, and relations.
We will often want to combine or compare two or more sets, and this
508 a course in abstract algebra
A B = B A A( BC ) = ( A B)C
Commutative Laws Associative Laws
A B = B A A( BC ) = ( A B)C
A( BC ) = ( A B)( AC ) A = A
Distributive Laws Identity Laws
A( BC ) = ( A B)( AC ) AE = A
A A0 = E
A( A B) = A
Complement Laws A A0 = Absorption Laws
A( A B) = A
0 0
(A ) = A
( A B)( A B0 ) = A ( A B)0 = A0 B0
If x A then its certainly the case that x A B and x A C, Figure A.3: Detail from a memorial
window to John Venn (18341923) in
because each of these sets are just A with the elements of, respectively, the dining hall of Gonville and Caius
B or C included. And since x lies in both A B and A C, then x College, Cambridge
510 a course in abstract algebra
A.2 Functions
little while, but the realisation that not every element of the codomain
need be mapped to inspires the following definition.
Definition A.10 Given a function f : A B, the image or range of
f is the subset
im( f ) = { f ( a) : a A}
of the codomain B.
the functions in the reverse order need not give the same composite
function; indeed the composite f g need not even be defined.
More generally, for two functions f : A B and g : C D, we can
define g f if and only if im( f ) C. That is, if the image of f is a
subset of the domain of g.
The reason we write g f to denote f followed by g is because it
makes the link with g( f ( a)) clearer. In a little while we will see
that this agrees with multiplication of matrices representing linear
transformations.
In some respects its less intuitive: we write the functions in the
reverse order to how theyre applied. Some books sidestep this issue
by applying functions on the right, writing x f instead of f ( x ). This
is a more intuitive ordering for function composition, since x ( f g)
is the same as x f g, but the cost is that this notation is less intuitive
in general. As a result, this convention tends to be rare now, but
sometimes appears in older books.
Sometimes we can use an explicit definition for f and g to obtain an
explicit definition for the composite g f .
Example A.15 Let f : R R such that f ( x ) = x2 + 5x + 1 and
g : R R such that g( x ) = 2x + 3.
Then we can define both composites g f and f g. To obtain explicit
expressions for these, we substitute the image of x under the action
of one function into the definition of the other. So
( g f )( x ) = g( f ( x ))
= g( x2 + 5x + 1)
= 2( x2 + 5x + 1) + 3
= 2x2 + 10x + 5,
( f g)( x ) = f ( g( x ))
= f (2x + 3)
= (2x + 3)2 + 5(2x + 3) + 1
= 4x2 + 22x + 25.
As we can see, f g 6= g f .
The last line of this example quietly introduces the idea of two func-
tions being equal to each other or not: two functions f : A B and
g : C D are equal if and only if A = C, B = D and f ( a) = g( a) for
all a A.
Just as we declared two sets to be equal if they contained exactly the
same elements, we regard two functions as equal if they have the same
domains and codomains, and act exactly the same on all elements of
514 a course in abstract algebra
their domain.
One special function, albeit one that seems trivial when we first meet
it, is the identity function.
Definition A.16 For a given set A, the identity map or identity
function id : A A is defined by id( a) = a for all a A. To avoid
ambiguity we might sometimes denote this function as id A .
Exercise A.1 Show that inverse func- Example A.20 Let g : R>0 R>0 map x x2 , and let h : R>0
tions, if they exist, are unique. That is,
for some function f : A B with two
R>0 map x x. Here we follow the usual convention that x
inverses g : B A and h : B A, it denotes the positive square root of x. Then g and h are the inverse of
must be the case that g = h. Hence it each other.
is appropriate to refer to the inverse
f 1 . Its not always the case that a function is invertible. The functions in
Show also that if g : B A is the the above example are invertible because we chose the domain and
inverse of f : A B, then f is also
the inverse of g. Hence ( f 1 )1 = f . codomain carefully to ensure that everything worked. The closely-
related function f : R R that maps x 7 x2 fails to be invertible in
background 515
Its useful to have a specific word to describe sets for which there
exists a bijective function to the natural numbers N. Since the natural
numbers are the counting numbers that we meet quite early in
primary school, a bijection to N is really just a way of counting the
elements of our chosen set.
Wikimedia Commons
Definition A.25 Let A be a set, and suppose either that A is finite, or Georg Ferdinand Ludwig Philipp
that there exists a bijection f : A N. Then we say A is countable. Cantor (18451918) was born in St Pe-
tersburg but in 1860, the family moved
If A is finite, then the cardinality or order | A| of A is the number to Frankfurt. Here Georg, already
of elements in A. If A is infinite and countable then we say it has an accomplished violinist, excelled
in mathematics and subsequently at-
cardinality | A| = 0 (pronounced aleph null or aleph zero). tended the Universities of Zrich,
Berlin and Gttingen. After completing
We will often use the leminiscate or analemma symbol to denote a doctoral dissertation in number the-
infinity in general, but 0 specifically denotes the countable infinity, ory at Berlin in 1867, he moved to the
University of Halle and was promoted
the cardinality of the natural numbers N.
to Professor in 1879.
Even more surprisingly, it turns out that the rational numbers Q are Much of his most important work was
countable. The proof of this celebrated result is due to the German achieved between 1874 and 1884, a pe-
riod overshadowed by bitter disputes
mathematician Georg Cantor (18451918). with the influential German mathemati-
cian Leopold Kronecker (18231891),
Proposition A.26 The set Q of rational numbers is countable.
a founder of the constructivist school
in mathematics, who had fundamen-
tal objections to Cantors work. Kro-
Proof We start by proving that the positive rationals Q+ are countable.
necker, head of department at Berlin,
To do this, we display them in a grid as follows: also blocked Cantors attempts to se-
1 2 3 4 5 6 cure an academic post there.
1 1 1 1 1 1
The stress of this situation took its toll,
and Cantor suffered the first of several
1 2 3 4 5 6 nervous breakdowns, spending time in
2 2 2 2 2 2 a sanatorium with severe depression.
He moved away from mathematics for
1 2 3 4 5 6 a time, focusing instead on philosophy
3 3 3 3 3 3
and English literature. He researched
the authorship of the plays of William
1 2 3 4 5 6 Shakespeare (15641616), and argued
4 4 4 4 4 4
in favour of Francis Bacon (15611626).
1 2 3 4 5 6 He suffered from depression for most
5 5 5 5 5 5
of the rest of his life, a condition that
was exacerbated by the death of his
1 2 3 4 5 6 youngest son Rudolph in 1899 and oc-
6 6 6 6 6 6
casionally fierce criticism of his work
on transfinite set theory. He died in
.. .. .. .. .. .. .. a sanatorium in 1918, roughly two
. . . . . . . months before his 73rd birthday.
518 a course in abstract algebra
0.y1 y2 . . . yk 9999 . . .
(where yk 6= 9) as
0.y1 y2 . . . (yk +1)0000 . . .
instead, and we pad any finite-length decimal expansion to an infinite
one by appending an infinite sequence of trailing zeros after the last
nonzero digit.
Our sequence x of all real numbers from (0, 1) can be written out as
infinite decimal expansions as follows:
A.4 Relations
equality relation =, but there are others that arise naturally in certain
contexts. In Section 2.2 we formulate a particularly important class of
equivalence relations on groups, and use it to prove Lagranges Theo-
rem.9 Probably the best-known example of an equivalence relation, 9
Theorem 2.30, page 54.
apart from = itself, is that of congruence modulo n.
Definition A.34 Two integers a, b Z are congruent modulo n for
some positive integer n N if ( ab)|n. Or, equivalently, if they both
have the same remainder when divided by n.
We write a n b or a b (mod n).
This is an equivalence relation on the set of integers.
A.5 Number theory God created the integers; all else is the
work of man.
Leopold Kronecker (18231891),
Much of abstract algebra is concerned with extending and gen- lecture at Gttingen (1886),
eralising facts about the integers to wider classes of objects, such as quoted by Heinrich Weber (18421913),
Jahresbericht der Deutschen
polynomials, symmetry operations, matrices and so forth. In this
Mathematiker-Vereinigung 2 (1892) 19
section we review a few elementary number-theoretic results used
elsewhere in the book. First we introduce a couple of basic definitions.
Definition A.35 Let a, b Z. Then a is a factor or divisor of, or
divides b, if b = ka for some integer k Z. We write a|b if this is so,
and a 6 |b if not.
The greatest common divisor or highest common factor of two pos-
itive integers m, n N is the largest positive integer d N that
divides both m and n. We write this as gcd(m, n). If gcd(m, n) = 1
then m and n are said to be coprime or relatively prime.
The ancient Greek mathematician Euclid of Alexandria (fl. 300BC)
gives an algorithm for finding the greatest common divisor of two
positive integers. It appears as Propositions 1 and 2 in Book VII (and 10 This proof makes use of the method
later, in a slightly different form, as Propositions 2 and 3 in Book X) of of Proof by Induction, which relies on
his celebrated treatise (Elements), although it almost certainly (indeed, its validity is formally equiva-
lent to) the Well-Ordering Principle:
dates back even earlier than that.
Axiom A.36 (The Well-Ordering
In order to prove the validity of this algorithm, we need a basic fact Principle) Let S N such that S 6=
about natural numbers, namely that the process of division with re- . That is, let S be a nonempty set of
natural numbers. Then S has a least ele-
mainders, which most of us learn at primary school, actually works.10 ment.
Theorem A.37 (The Division Theorem) Let a, b N. Then there Depending on exactly which version
exist unique integers q, r Z such that a = bq + r and 0 6 r < b. of axiomatic set theory one is working
with, this statement is either a basic
axiom or a provable proposition. We
Proof First, we prove the existence of q and r. If a = b then obviously will cheerfully ignore such concerns in
a = 1b + 0, so q = 1 and r = 0 suffice. the rest of this book.
524 a course in abstract algebra
Proof From the latter part of the above proof, we know that Euclids
Algorithm ensures the existence of integers q1 , . . . , qn+1 and b > r1 >
> rn+1 > 0 such that
r1 = a q1 b,
r2 = b q2 r1 ,
526 a course in abstract algebra
r3 = r1 q3 r2 ,
..
.
r n 1 = r n 3 q n 1 r n 2 ,
d = r n = r n 2 q n r n 1 .
Substituting the penultimate expression into the last one, we obtain
d = r n = r n 2 q n ( r n 3 q n 1 r n 2 )
= (1 + q n q n 1 ) r n 2 q n r n 3 .
Hence d = rn can be expressed as a Zlinear combination of rn3
and rn2 . Now we use the antepenultimate expression on the list to
substitute for rn2 and thereby express d as a Zlinear combination of
rn4 and rn3 .
Continuing this process, we eventually obtain an expression for d as a
Zlinear combination of a and b, as required.
The following important fact about coprime integers is easy to prove,
and will be used elsewhere in the book.
Proposition A.40 Let a and b be coprime integers. Then if a divides bq
then a must divide q.