Math 120A Introduction to Group Theory
Neil Donaldson
December 1, 2025
1 Introduction: what is abstract algebra and why study groups?
To abstract something means to remove context and application. Modern mathematics largely in-
volves studying patterns and symmetries (often observed in the real world) abstractly so as to ob-
serve commonalities between structures in seemingly distinct places.
One reason to study groups is that they are relatively simple: a set and a single operation which
together satisfy a few basic properties. Indeed you’ve been using this structure since Kindergarten!
Example 1.1. The integers Z = {. . . , 1, 0, 1, 2, 3, . . .} together with the operation + form a group.
We’ll see a formal definition shortly, at which point we’ll be able to verify that (Z, +) really is a
group. The simplicity of the group structure means that it is often used as a building block for
more complicated structures.
1
Other reasons to study groups are their ubiquity and multitudinous
applications. Here are just a few of the places where the language of group theory is essential.
Permutations In mathematics, the word group was first used to describe the ways in which a set
could be reordered, or permuted. Understanding permutations is of crucial importance to many
areas of mathematics, particularly combinatorics, probability and Galois theory: this last, the
crown jewel of undergraduate algebra, develops a deep relationship between the solvability of
a polynomial and the permutation group of its set of roots.
Geometry Figures in Euclidean geometry (e.g. triangles) are congruent if one may be transformed
to the other by an element of the Euclidean group (a translation, rotation or reflection). More gen-
eral geometries may also be described by their groups of symmetries. Groups may also be
employed to describe geometric properties: for example, the number of holes in an object (a
sphere has none, a torus one, etc.) is related to the structure of its fundamental group.
Chemistry Group Theory may be applied to describe the symmetries of molecules and of crystalline
substances.
Physics Materials science sees group theory similarly to chemistry. Modern theories of the nature of
the universe and fundamental particles/forces (e.g. gauge/string theories) also rely heavily on
groups.
Of course, the best reason to study groups is simply that they’re fun!
1
For instance, the set of integers Z together with the two basic operations of addition and multiplication is a ring, as
you’ll study in a later course.
1
2 Group Axioms and Basic Examples
In this chapter we define our main objects of study and introduce some of the vocabulary and stan-
dard examples used throughout the course. The “Key concepts/definitions” listed at the start of each
Exercise set summarize these.
2.1 The Axioms of a Group
Definition 2.1 (Closure). Let G be a set and a function : G ×G G. We describe this arrange-
ment in four different ways, though all mean exactly the same thing:
(a) x, y G, x y G (b) G is closed under .
(c) is a binary operation on G. (d) (G, ) is a binary structure.
In abstract situations (including most theorems) we typically drop the symbol and use juxtaposition
(x y = xy). In explicit examples this might be a bad idea, say if is addition. . .
Examples 2.2. 1. Addition (+) is a binary operation on the set of integers Z:
Given x, y Z, we know that x + y Z
This isn’t a claim you can prove, since it is really part of the definition of integer addition.
2. Subtraction () is not a binary operation on the positive integers N = {1, 2, 3, 4, . . .}. This you
can prove; to show that condition (a) is false, we exhibit a counter-example
1 7 = 6 N (x, y N such that x y N)
On the integers, however, subtraction is a binary operation: Z is closed under .
3. On a small set, it can be convenient to represent a binary operation in tabular
form. The given table describes an operation on a set of three elements
G = {e, a, b}. Read the left column first, then the top row: for instance,
a b = e, or simply ab = e
e a b
e e a b
a a e e
b b e a
We’ll continue checking these examples for each of the remaining axioms.
Definition 2.3 (Associativity). A binary structure (G, ) is associative if
x, y, z G, x(yz) = (xy)z
If is associative, then the expression xyz has unambiguous meaning, as does exponential notation:
x
n
= x ···x (n factors).
Examples (2.2, ver. II). 1. Addition is associative: x + (y + z) = (x + y) + z for any integers.
2. ( Z, ) is non-associative: e.g. (1 3) 2 = 4, but 1 (3 2) = 0.
3.
{e, a, b},
, described in the table, is non-associative: e.g. a(bb) = aa = e, but (ab)b = eb = b.
2
Definition 2.4 (Identity). A binary structure (G, ) has an identity element e G if
x G, ex = xe = x
Examples (2.2, ver. III). 1. Addition on Z has identity 0, since 0 + x = x + 0 = x for any integer x.
2. ( Z, ) does not have an identity: if e x = x, then e = 2x would depend on x!
3.
{e, a, b},
has identity e: observe the first row and column of the table.
If G is finite and has an identity (e.g. Example 2.2.3), convention dictates that we list it first. Indeed,
we can always list it first, since. . .
Lemma 2.5 (Uniqueness of identity). A binary structure (G, ) has at most one identity.
It is now legitimate to refer to the identity e. Uniqueness proofs in mathematics often follow a stan-
dard pattern: suppose there are two such objects and show that they are identical.
Proof. Suppose e, f G are identities. Then
e f =
(
f since e is an identity
e since f is an identity
Since f = e, there is only one identity.
We used almost nothing about (G, ); in particular it need not be associative (e.g. Example 2.2.3).
Definition 2.6 (Inverse). Suppose a binary structure (G, ) has identity e. An element x G has an
inverse y G if
xy = yx = e
Examples (2.2, ver. IV). 1. Every integer x has an inverse under addition: x + (x) = (x) + x = 0.
2. Since ( Z, ) has no identity, the question of inverses is irrelevant.
3. Since ee = aa = ab = ba = e, we see that every element has an
inverse; indeed a has two inverses!
Element e a b
Inverses e a, b a
Lemma 2.7 (Uniqueness of inverses). Suppose a binary structure (G, ) is associative and has an
identity. If an element x G has an inverse, then said inverse is unique.
Proof. Suppose x has inverses y, z G. By associativity,
z(xy) = (zx)y = ze = ey = z = y
In such a situation it is legitimate to write x
1
(or x) for the inverse of x. Example 2.2.3 shows that
associativity is necessary: a non-associative binary structure can have non-unique inverses.
3
Definition 2.8 (Commutativity). Let (G, ) be a binary structure. Elements x, y G commute if
xy = yx. We say that is commutative if all elements commute:
x, y G, xy = yx
Examples (2.2, ver.V). 1. Addition of integers is commutative: x, y Z, x + y = y + x.
2. Subtraction of integers is non-commutative: e.g. 2 3 = 3 2.
3. The relation is commutative since its table is symmetric across the main diagonal.
To obtain our main definition simply assemble the pieces!
Definition 2.9 (Group axioms). A group G is a binary structure (G, ) satisfying the associativity and
identity axioms, and for which all elements have inverses. This is summarized by the mnemonic
Closure, Associativity, Identity, Inverse
The order of a group is the cardinality (size)
|
G
|
of the underlying set.
2
In addition, a group G is said to be abelian if the operation is commutative.
In a multiplicative group, the operation is written multiplicatively or using juxtaposition (includes
composition of functions). A group is additive
3
if the operation is addition. Abstract groups are
almost always written multiplicatively.
Examples (2.2, ver.VI). 1. (Z, +) is an infinite, abelian, additive group. Precisely the same observa-
tions show that ( Q, +), (R, +) and (C, +) are also.
2. ( Z, ) is not a group since subtraction is neither associative, nor has an identity (nor inverses).
3. This binary relation is non-associative and so does not define a group.
While it is common practice to refer to a set G as a group, you should do so only if the operation is
obvious to everyone. Writing Z is a group under addition,” is safer than Z is a group:” it might be a
group under many different operations!
Examples 2.10. 1. The non-zero real numbers R
×
form an abelian group under multiplication.
Closure If x, y = 0, then xy = 0
Associativity x, y, z, x(yz) = (xy)z
Identity 1 R
×
is the identity since, for any x = 0, we have 1 · x = x ·1 = x
Inverse Given x = 0, observe that x
1
=
1
x
is an inverse: x ·
1
x
=
1
x
· x = 1
Commutativity If x, y = 0, then xy = yx
As with addition of integers, we cannot prove these claims since they are part of the definition
of multiplication. Similarly, (Q
×
, ·) and (C
×
, ·) are abelian groups.
2
A finite group has finite order, while an infinite group has infinite order; unless absolutely necessary, it is rare to be
specific about infinite cardinalities (countable, uncountable, etc.).
3
These are distinctions only of notation. For instance x + x + x = 3x in an additive group corresponds to xxx = x
3
in a
multiplicative group.
4
2. The set of even integers 2Z = {2z : z Z} forms an abelian group under addition.
3. The set of odd integers 1 + 2Z = {1 + 2n : n Z} does not form a group under addition since
they are not closed. For instance, 1 + 1 = 2 ∈ 1 + 2Z.
4. Every vector space is an abelian group under addition.
5. ( R, ·) is not a group since 0 has no multiplicative inverse. Similarly (Q, ·), (C, ·) are not groups.
6. A Cayley table
4
is a tabular representation of a (small)
group. Groups of orders 1, 2 and 3 are shown. The one-
element group {e} is often called the trivial group.
Note the magic square (sudoku) property: each row/column
contains every element exactly once (see Exercise 11).
e
e e
e a
e e a
a a e
e a b
e e a b
a a b e
b b e a
Theorem 2.11 (Cancellation laws & inverses). Suppose G is a group and that x, y, z G. Then
1. xy = xz = y = z 2. xz = yz = x = y 3. (xy)
1
= y
1
x
1
Part 3 should remind you of matrix multiplication.
Proof. Parts 1 & 2 are exercises. For part 3, multiply out using associativity:
(y
1
x
1
)(xy) = y
1
(x
1
x)y = y
1
ey = y
1
y = e
Similarly (xy)(y
1
x
1
) = e. Thus y
1
x
1
is the inverse of xy (unique by Lemma 2.7).
Exercises 2.1. Key concepts/definitions/examples: make sure you can state the formal definitions.
Group (closure, associativity, identity, inverse) Commutativity/abelian Cayley table
1. Given the binary operation table, calculate
(a) c d (b) a (c b)
(c) (c b) a (d) (d c) (b a)
a b c d
a c d c b
b d c b a
c a d c d
d b a b c
2. The table for a binary operation on the set {a, b, c}is given. Compute
a (b c) and (a b) c. Does the expression a b c make sense?
Why/why not?
a b c
a b c b
b c a a
c b a c
3. Are the binary operations in the previous questions commutative? Explain.
4. (a) Describe (without writing them all out!) all possible binary operation tables on a set of two
elements {a, b}. Of these, how many are commutative?
(b) How many commutative/non-commutative operations are there on a set of n elements?
(Hint: a commutative table has what sort of symmetry?)
4
Englishman Arthur Cayley (1821–95) was an early group theorist. Similarly abelian honors the Norwegian Niels Abel
(1802–29), after whom the Abel Prize is named (often considered the Nobel Prize in Mathematics).
5
5. Which are binary structures? For those that are, which are commutative and which associative?
Give brief arguments in each case.
(a) ( Z, ), a b = a + b + 1 (b) (R, ), a b = 2(a + b)
(c) ( R, ), a b = 2a + b (d) ( R, ), a b =
a
b
(e) ( N, ), a b = a
b
(f) ( Q
+
, ), a b = a
b
, where Q
+
= {x Q : x > 0}
(g) ( N, ), a b = product of the distinct prime factors of ab. Also define 1 1 = 1.
(e.g. 42 10 = (2 ·3 ·7) (2 ·5) = 2 ·3 · 5 ·7 = 210)
6. Verify the axioms of an abelian group; if any are false, provide a counter-example.
(a) N under addition. (b) Q under multiplication.
(c) X = {a, b, c} with x y := y. (d) R
3
with the cross/vector product ×.
(e) For each n R, the set nZ = {nz : z Z} of multiples of n under addition.
7. (a) Prove the cancellation laws (Theorem 2.11 parts 1 & 2).
(b) True or false? In a group, if xy = e, then y = x
1
.
(c) In a multiplicative group G, we can unambiguously write (x
1
)
n
= x
1
···x
1
| {z }
n times
.
For any n N and x G, prove that (x
1
)
n
= (x
n
)
1
. By convention, this object is
denoted x
n
. How would we write this in an additive group (Footnote 3)?
8. Let G be a group. Prove:
(a) x, y G, (xyx
1
)
2
= xy
2
x
1
(b) x G, (x
1
)
1
= x
(c) G is abelian x, y G, (xy)
1
= x
1
y
1
9. Prove or disprove:
R \ {1},
is an abelian group, where x y := x + y xy.
10. Let U be a set and P(U) its power set (the set of subsets of X).
(a) Which of the group axioms are satisfied by the union operator on P(U)?
(b) Repeat part (a) for the intersection operator.
(c) The symmetric difference of sets A, B U is the set
AB := (A B) \(A B)
i. Use Venn diagrams to give a sketch argument that is associative on P(U).
ii. Is
P(U),
a group? Explain your answer.
11. (Magic Square) Suppose (G, ) is associative and that G is finite.
Prove that (G, ) is a group if and only if its (multiplication) table satisfies two conditions:
i. One row and column (by convention the first) is a perfect copy of G itself.
ii. Every element of G appears exactly once in each row and column.
6
2.2 Subgroups
The prefix sub- in mathematics usually indicates a subset that retains the indicated structure.
Definition 2.12 (Subgroup). Let G be a group. A subgroup of G is a non-empty subset H G which
remains a group with respect to the same binary operation. We write H G.
A subgroup H is a proper subgroup if H = G. This is written H < G.
The trivial subgroup of G is the 1-element set {e}; all other subgroups are non-trivial.
Examples 2.13. The following should be immediate from the definition: all you need is a non-empty
subset that remains a group!
1. {e} G and G G for any group G 2. (Z, +) < (Q, +) < ( R, +) < (C, +)
3. ( Q
×
, ·) < (R
×
, ·) < (C
×
, ·) 4. (R
n
, +) < (C
n
, +) 5. (2Z, +) < (Z, +)
4. ( R
m
, +) (R
n
, +) if m n. For instance, with respect to the standard basis, R
m
consists of all
column vectors in R
n
whose last n m entries are zero.
5. (C(R), +) < (C
1
(R ), +) . Think back to calculus: the sub of any two continuous functions is
continuous; every differentiable function is continuous; etc., etc.
Thankfully one doesn’t have to check all the group axioms to see that a subset is a subgroup.
Theorem 2.14 (Subgroup criterion). Let G be a group. A non-empty subset H G is a subgroup if
and only if it is closed and has inverses in H (with respect to the group operation on G):
h, k H, hk H and h
1
H ()
Proof. () H is a group and therefore satisfies all the axioms, including closure and inverse.
() By assumption, H satisfies the closure axiom. Moreover, the group operation on G is automati-
cally associative on any subset,
5
including H. It remains to verify that the identity element e (of
G) lies in H, for then our assumption (h
1
H) says that that inverse axiom is also satisfied.
Since H = , we may choose some (any!) h H. By (), h
1
H. A second application of ()
finishes things off:
e = hh
1
H
Examples 2.15. 1. All of Examples 2.13 can be confirmed using the theorem. For instance, part 5:
Non-empty subset: plainly 2Z = {. . . , 2, 0, 2, 4, . . .} = {2z : z Z} is such (of Z).
Closure: 2m, 2n 2Z = 2m + 2n = 2(m + n) 2Z.
Inverses: 2m 2Z has inverse ( 2m) = 2(m) 2Z.
2. The positive integers N are closed under addition but do not satisfy the inverse axiom (for
instance, no x N satisfies x + 2 = 0). Thus (N, +) is not a subgroup of (Z, +).
5
Associativity does not care where x(yz) = (xy)z lives: G does not appear in Definition 2.3!
7
3. Denote by 1 + 3Z the set of integers with remainder 1 when divided by 3:
1 + 3Z =
1 + 3n : n Z
=
1, 4, 7, 10, 13, . . . , 2, 5, 8, . . .
Since 1 1 + 3Z but 1 + 1 = 2 1 + 3Z, we see that 1 + 3Z is not a subgroup of (Z, +).
4. The circle group S
1
:=
e
iθ
: θ [0, 2π)
is plainly a non-empty subset of (C
×
, ·). The standard
exponential laws and the fact that e
2πi
= 1 verify that S
1
is in fact a subgroup.
Closure: e
iθ
e
iψ
= e
i(θ+ψ)
S
1
(equals e
(θ+ψ2π)i
if you feel it necessary).
Inverses: (e
iθ
)
1
= e
iθ
= e
(2πθ)i
.
Subgroup Diagrams It can be helpful to represent subgroup relations pictorially,
using descending lines. For instance, the diagram on the right summarizes the sub-
group relations
6Z < 2Z < Z, 6Z < 3Z < Z, 6Z < Z
Z
2Z 3Z
6Z
where all four groups under addition. If G has only finitely many subgroups, then its subgroup
diagram is the complete depiction of all subgroups.
Exercises 2.2. Key concepts: (Proper/trivial/non-trivial) Subgroup
Subgroup criterion (non-empty subset, closure, inverses) Subgroup diagram
1. Use the subgroup criterion to verify that Q
×
is a subgroup of R
×
under multiplication.
2. Give two reasons why the non-zero integers do not form a subgroup of Z under addition.
3. Describe/explain the relationship between positive integers m and n if (mZ, +) (nZ, + ).
4. Prove or disprove: the set H = {
a
2
n
: a Z, n N
0
} forms a group under addition.
5. Briefly explain why “subgroup” is transitive: that is, if K H and H G, then K G.
6. Suppose H and K are subgroups of G. Prove that H K is also a subgroup of G.
7. Let H be a non-empty subset of a group G. Prove that H is a subgroup of G if and only if
x, y H, xy
1
H
8. (Hard) On an abstract set Q
8
= {±1, ±i, ±j, ±k} of eight elements, we define an operation
(’multiplication’) using several properties:
1 is the identity.
1 commutes with everything in the expected way: e.g. i = (1)i = i(1), etc.
(1)
2
= 1, i
2
= j
2
= k
2
= 1 and ij = k.
Multiplication is associative.
(a) Prove that (Q
8
, ·) is a non-abelian group by completing its Cayley table.
(Hint: You should easily be able to fill in 44 of 64 entries; now use associativity. . . )
(b) Find all subgroups of Q
8
and draw its subgroup diagram.
8
2.3 Modular Arithmetic
Many commonly encountered examples in abstract algebra make use of modular arithmetic: the ad-
dition and multiplication of remainders. Such arithmetic should be at least somewhat familiar so we
offer only a brief refresher. At present, these groups are very informal and are introduced primarily
to supply examples; more rigorous discussions will be given in Chapters 3 & 5.
Definition 2.16. Let n be a positive integer. We denote by Z
n
the set of equivalence classes of integers
modulo n. These are typically written as remainders (i.e., as integers),
Z
n
= {0, 1, . . . , n 1}
where x = y Z
n
means that the integers x, y have the same remainder on division by n.
More formally,
6
x = y Z
n
means x y (mod n), or equivalently x = y + λn for some integer λ.
Example 2.17. In the most commonly used notation, we write Z
5
= {0, 1, 2, 3, 4}. Several other
notations are available. For instance, here is a calculation written in four different ways (note that
6 = 1 in Z
5
because they both have the same remainder 1 on division by 5):
(a) Group/Number Theory style: 4 + 2 = 6 = 1 in Z
5
.
(b) Modular arithmetic: 4 + 2 6 1 (mod 5).
(c) Decorated operation: 4 +
5
2 = 6 = 1.
(d) Equivalence classes: [4] +
5
[2] = [ 6] = [1].
For reasons of brevity we mostly use notation (a), though feel
free to use another if it makes you more comfortable. Regard-
less of notation, you must make it clear in which Z
n
you are
working: 4 + 2 = 1 is not acceptable on its own!
[0]
= {. . . , 0, 5, 10, . . .}
[1]
= {
. . . ,
4, 1, 6, . . .
}
[2]
= {
. . . ,
3, 2, 7, . . .
}
[3]
= {
. . . ,
2, 3, 8, . . .
}
[4]
= {
. . . ,
1, 4, 9, . . .
}
+1
+1
+1
+1
+1
Theorem 2.18. Z
n
forms an abelian group of order n under addition modulo n.
A rigorous proof (in the language of Footnote 6) is tedious, but will come for free in Chapter 5 when
Z
n
is properly defined as a factor group. These groups are so common that we usually just state “The
group Z
n
,” rather than (Z
n
, +
n
). In the exercises, we’ll also consider how multiplication modulo n
can be used to create groups of remainders.
6
An element of Z
n
is strictly an equivalence class: for instance, [x] Z
n
denotes the class of all integers with the same
remainder as the representative x Z:
[x] =
z Z : x z (mod n)
=
. . . , x n, x, x + n, x + 2n . . .
= {x + kn : k Z} = x + nZ
Addition of equivalence classes is well-defined (multiplication similarly): if [ x] = [w] and [y] = [z], then w = x + κn and
z = y + λn, from which
[w] +
n
[z] = [w + z] =
(x + κn) + (y + λn)
=
x + y + n(κ + λ)
= [x + y] = [x] +
n
[y]
While it is important to appreciate that the elements of Z
n
are not really numbers, the tediousness of this formal language
means that it is usually avoided. Equivalence classes and well-definition are not critical right now, but will become so later.
9
Examples 2.19. Here are the Cayley tables for Z
1
, Z
2
, Z
3
and Z
4
.
+
1
0
0 0
+
2
0 1
0 0 1
1 1 0
+
3
0 1 2
0 0 1 2
1 1 2 0
2 2 0 1
+
4
0 1 2 3
0 0 1 2 3
1 1 2 3 0
2 2 3 0 1
3 3 0 1 2
In each case 0 is the identity element. If you compare these to the the tables in Example 2.10.6, the
patterns should look familiar (we will explore this further in Section 2.5).
Subgroups of Z
n
It is easy to spot certain subgroups of Z
n
just think about divisors of n!
Example 2.20. Plainly 2 is a divisor of 4. By covering up the rows/columns corresponding to 1 and
3, we obtain the Cayley table for the subgroup H = {0, 2} of Z
4
.
+
4
0 1 2 3
0 0 1 2 3
1 1 2 3 0
2 2 3 0 1
3 3 0 1 2
+
4
0 2
0 0 2
2 2 0
Hopefully it is obvious why H is a subgroup (think about the subgroup criterion Theorem 2.14!).
Now suppose 1 were in some subgroup K Z
4
and consider what the group axioms tell us:
Closure = 2 = 1 + 1 K
Inverse = 3 = 1 K
Identity = 0 K
= K = {0, 1, 2, 3} = Z
4
The same thing happens if a subgroup contains 3. The upshot is that Z
4
has precisely three
subgroups: itself, {0, 2} and the trivial subgroup {0}. The full subgroup diagram is drawn.
Z
4
{0, 2}
{0}
Here is a more general version. Suppose d is a divisor of n, write n = dk, and consider the subset of
multiples of d in Z
n
:
d
=
0, d, 2d, . . . , (k 1)d
In the language of the subgroup criterion (Theorem 2.14), this set is:
Non-empty: Plainly 0
d
.
Closed under addition: κd + λd = (κ + λ)d
d
.
Closed under inverses: The inverse of λd is λd = (k λ)d
d
.
We have therefore proved:
Lemma 2.21. If d is a divisor of n, then the set of multiples
d
in Z
n
is a subgroup of order
n
d
.
As in Example 2.20, in Chapter 3 we’ll see that these are in fact the only subgroups of Z
n
.
10
Exercises 2.3. Key concepts: Z
n
, Multiples as subgroups: d | n =
d
Z
n
1. Refresh your memory of modular arithmetic by evaluating the following:
(a) 17 + 22 in Z
30
(b) 31 ·4 in Z
12
(c) 5
6
in Z
14
, (d) 19
2
42 ·13 in Z
17
2. State the Cayley tables for the groups Z
5
and Z
6
(more formally (Z
5
, +
5
) and (Z
6
, +
6
)).
3. State the Cayley tables for all proper subgroups of Z
6
. Now draw the full subgroup diagram
for Z
6
.
(Hint: consider Lemma 2.21 and the remark that follows)
4. Draw the subgroup diagram for Z
12
.
5. Suppose n is a positive integer 2.
(a) Explain why Z
n
is not a group under multiplication. If you’re unsure what to do, consider
an example: what is the multiplication table for (Z
3
, ·)?
(b) Explain why {1, 2, 3, 4, 5} isn’t a group under multiplication modulo 6.
(c) Hypothesize for which integers n 2 the set {1, 2, 3, . . . , n 1} forms a group under
multiplication modulo n. If you want a challenge, try to prove your assertion.
6. The set Z
×
n
denotes the units in Z
n
, those elements which are relatively prime to n:
Z
×
n
=
x Z
n
: gcd(x, n) = 1
In part (d), we verify that Z
×
n
is an abelian group under multiplication modulo n.
(a) Construct the Cayley tables for the groups Z
×
3
= {1, 2} and Z
×
4
= {1, 3}.
(b) Construct the Cayley table for the group Z
×
5
= {1, 2, 3, 4}. Now identify its subgroups.
(c) Construct the Cayley tables for Z
×
8
and Z
×
9
. What is the order of each group?
(d) (Hard) Prove that Z
×
n
forms an abelian group under multiplication modulo n by verifying
the group axioms.
(Hint: Recall B´ezout’s identity gcd(x, n) = 1 κ, λ Z such that κx + λn = 1)
(e) i. Compare the orders of the groups Z
×
3
, Z
×
4
and Z
×
12
. What do you observe?
ii. What about the orders of Z
×
2
, Z
×
6
and Z
×
12
? What is going on?
iii. The order of Z
×
n
is the value of Eulers totient function ϕ(n). Research some of the prop-
erties of this function: can you find something that helps explain your observations in
parts i and ii? Better still, take a course in Number Theory!
11
2.4 Geometric Symmetries & Matrix Groups
Geometric symmetries an matrices provide further large families of groups.
Now consider any geometric figure that has some symmetry (typically viewed as rotational or re-
flective), such as a triangle, square, or tetrahedron. Each symmetry corresponds to a function that
transforms the original figure in such a way that the result occupies the same location as the original.
Example 2.22 (Klein four-group). The pictured rectangle has three obvious symmetries:
(a) Rotation by 180°.
(b) Vertical reflection.
(c) Horizontal reflection.
a
c
b
Each symmetry may be viewed as a function transforming the rectangle (or permuting its ver-
tices/edges if you prefer). Group Theorists also consider the identity function e as a symmetry: it
simply leaves the rectangle alone.
7
It should be clear that the set V := {e, a, b, c} comprises every symmetry of the rectangle. We claim
that V forms a group whose binary operation is composition of functions.
Closure After applying any two symmetries in sequence, the rectangle still occu-
pies the same location on the page, the result of applying a single symmetry.
The composition table is shown and is easily be verified by, for instance,
drawing a smiley face on one side of a sheet of paper.
e a b c
e e a b c
a a e c b
b b c e a
c c b a e
Original Face on back
Face up
c
b
a
The pictures confirm b c = a: remember that the right side comes first when composing
functions! The diagonal symmetry of the table shows that the operation is commutative.
Associativity Composition of functions is always associative (Exercise 9).
Identity The function e leaves the rectangle alone. Plainly e f = f e = f for any symmetry f .
Inverse To find the inverse of a symmetry, simply undo what you just did! In the case of the rectan-
gle, every symmetry is its own inverse.
The symmetries of the rectangle thus form an abelian group of order 4. This is named the Klein
four-group in honor of Felix Klein, a 19
th
century German mathematician whose application of group
theory transformed modern geometry. The letter V comes from the original German: vierergruppe.
7
There is little benefit to being explicit, but if you choose co-ordinates with the origin at the center of the rectangle, these
functions can be written formulaically:
e(x, y) = (x, y), a(x, y) = (x, y), b(x, y) = (x, y), c(x, y) = (x, y)
12
A similar discussion applies to any geometric figure.
Theorem 2.23. The symmetries of any geometric figure form a group under composition.
The orientation-preserving symmetries form a subgroup, often called the rotation group.
8
Example (2.15.4, cont.). Since the complex function z 7 e
iθ
z rotates counter-clockwise by θ around
the origin, the circle group S
1
may be viewed as the group of rotations of the plane (or the circle).
Definition 2.24. A regular n-gon has two commonly associated symmetry groups.
Dihedral Group The full symmetry group D
n
has order 2n. It splits into two subsets of size n:
Rotations Labelled e, ρ
1
, . . . , ρ
n1
where ρ
k
rotates counter-clockwise by
2πk
n
radians (
360k
n
°). The
identity e = ρ
0
is considered a rotation (by 0°).
Reflections These are typically labelled µ
k
or δ
k
.
Rotation group Denoted R
n
= {e, ρ
1
, . . . , ρ
n1
}. This group is abelian, which follows because com-
position of rotations simply sums angles:
ρ
j
ρ
k
= ρ
j+k (mod n)
= ρ
k+j ( mod n)
= ρ
k
ρ
j
Example 2.25. Denote the elements of the dihedral group D
3
as in the picture, where labeling the
vertices 1, 2, 3 helps to keep track of things:
D
3
=
e, ρ
1
, ρ
2
, µ
1
, µ
2
, µ
3
Its Cayley table is below. Constructing the table from scratch is a lot
of work and is not worth memorizing!
The highlighted computation is µ
1
ρ
2
= µ
3
(remember to ’do’ ρ
2
to the triangle first!). To verify, we could again try the smiley face
trick, or alternatively consider the movement of a vertex:
ρ
2
moves vertex 1 to vertex 3; µ
1
moves this to vertex 2.
Since the composition of a rotation and a reflection is a
reflection (the triangle has been flipped over once!), the
result must be the reflection mapping 1 7 2, namely µ
3
.
The lack of symmetry in the Cayley table shows that D
3
is a non-
abelian group: indeed
ρ
2
µ
1
= µ
2
= µ
3
= µ
1
ρ
2
1
2
3
µ
1
µ
2
µ
3
ρ
1
ρ
2
e ρ
1
ρ
2
µ
1
µ
2
µ
3
e e ρ
1
ρ
2
µ
1
µ
2
µ
3
ρ
1
ρ
1
ρ
2
e µ
3
µ
1
µ
2
ρ
2
ρ
2
e ρ
1
µ
2
µ
3
µ
1
µ
1
µ
1
µ
2
µ
3
e ρ
1
ρ
2
µ
2
µ
2
µ
3
µ
1
ρ
2
e ρ
1
µ
3
µ
3
µ
1
µ
2
ρ
1
ρ
2
e
The Cayley table for the (abelian) rotation group R
3
= {e, ρ
1
, ρ
2
} is visible in the top left corner.
8
In low dimensions, orientation-preserving means that a transformation doesn’t change the usual right-hand rule (e.g.,
cross products). In two dimensions these are precisely the planar rotations. In three dimensions a general orientation-
preserving symmetry is the composition of two pure rotations (recall spherical polar co-ordinates).
13
Geometric Subgroup Relations
These are often straightforward to observe by drawing two shapes in such a way that all the symme-
tries of one are also symmetries of the other. Since the symmetries of both shapes form a group, this
pictorial approach justifies the only necessary condition in Definition 2.12: the subset property.
Examples 2.26. 1. For any n, we see that R
n
< S
1
: every rotation of a regular n-gon is also a rotation
of a circle (with the same center).
2. Consider a regular hexagon, inside which have been drawn two equilateral triangles. Every
symmetry of either triangle is also a symmetry the hexagon. We conclude:
µ
1
µ
3
µ
5
ρ
2
ρ
4
µ
2
µ
4
µ
6
ρ
2
ρ
4
(a) R
3
< R
6
. Be careful with notation! With respect to the hexagon, ρ
k
is rotation counter-
clockwise by 60k°, whence the subgroup relation should be written
R
3
= {e, ρ
2
, ρ
4
} < R
6
= {e, ρ
1
, ρ
2
, ρ
3
, ρ
4
, ρ
5
}
Also note that both triangles have the same rotation group.
(b) D
3
< D
6
. This is a little more complicated. Labeling the reflections µ
1
, . . . , µ
6
as in the
picture, we see that the two triangles actually have different (full) symmetry groups:
D
I
3
= {e, ρ
2
, ρ
4
, µ
1
, µ
3
, µ
5
} < D
6
and D
II
3
= {e, ρ
2
, ρ
4
, µ
2
, µ
4
, µ
6
} < D
6
Otherwise said, D
6
has two distinct subgroups that look like D
3
.
Matrix Groups
As observed in any elementary linear algebra course (see also Exercise 10), matrix multiplication is
associative. This quickly yields several examples.
Example 2.27. The general linear group comprises the invertible n × n matrices under multiplication.
For this course, only such matrices with real number entries will be encountered:
GL
n
(R ) =
A M
n
(R ) : det A = 0
(non-abelian when n 2)
Since associativity holds in general, we need only verify the other three axioms.
Closure follows from the familiar result det AB = det A det B.
The identity (drum roll. . . ) is the identity matrix I.
Finally, invertibility is assumed. Part 3 of Theorem 2.11 should now seem very
familiar: (xy)
1
= y
1
x
1
.
I =
1 0
0 1
.
.
.
.
.
.
.
.
.
0
0 1
14
Matrix subgroups
The general linear group GL
n
(R ) has many subgroups. Here is one; some others are in the Exercises.
Example 2.28. The orthogonal group O
n
(R ) = {A M
n
(R ) : A
T
A = I} consists of those matrices
whose inverse equals their transpose (A
1
= A
T
). We verify that this is a subgroup of GL
n
(R ) using
the subgroup criterion (Theorem 2.14) and simple matrix properties.
Non-empty subset Every orthogonal matrix is invertible, whence O
n
(R ) GL
n
(R ). This is imme-
diate in two ways: if A O
n
(R ), then,
A
T
A = I = A
1
= A
T
exists, or,
1 = det I = det A det A
T
= (det A)
2
= det A = 0
Moreover, I
T
I = I
2
= I, so I O
n
(R ): we have non-emptiness.
Closure Suppose A, B O
n
(R ). Then
(AB)
T
(AB) = B
T
A
T
AB = B
T
IB = B
T
B = I = AB O
n
(R )
Inverses Suppose A O
n
(R ). Then
(A
1
)
T
A
1
= (A
T
)
T
A
T
= (AA
T
)
T
= I
T
= I = A
1
O
n
(R )
The 2 ×2 orthogonal matrices can be interpreted as rotations and reflections.
9
For instance, the matrix
1
2
1 1
1 1
O
2
(R ) rotates the plane counter-clockwise by 45°.
Exercises 2.4. Key concepts: Klein 4-group V Dihedral group D
n
Rotation group R
n
Geometric subgroup relations General linear group GL
n
(R )
1. Use Theorem 2.14 to explain why the set of rotations of a planar figure is a subgroup of its full
symmetry group (rotations and reflections).
2. Explicitly state the Cayley table for the rotation group R
4
of a square.
3. Find the subgroup diagram of the Klein four-group. Explain how you know you are correct.
4. Repeat the previous question for the rotation group R
6
.
5. (a) Find all subgroups and the subgroup diagram for the group D
3
.
(Don’t worry about being rigorous as to how you know you’ve found them all.)
(b) Describe the symmetry group and Cayley table of a non-equilateral isosceles triangle. What
about a scalene triangle?
9
You might have seen this in another course. Left-multiplication by:
cos θ sin θ
sin θ cos θ
rotates vectors counter-clockwise by θ radians.
cos θ sin θ
sin θ cos θ
reflects across the line making angle θ/2 with the positive real axis.
This interpretation allows us to view D
n
as a subgroup of O
2
(R).
15
6. (a) Represent the elements of the Klein four-group V (as in Footnote 7) using matrix notation
(i.e. a is left-multiplication by what matrix). As such, identify V as a subgroup of O
2
(R ).
(b) Modeling Example 2.26, draw three pictures which describe different ways in which V
may be viewed as a subgroup of D
6
.
7. Determine whether each of the following sets of matrices is a group under multiplication.
(a) K = {A M
2
(R ) : det A = ±1} (b) L = {A M
2
(R ) : det A = 7}
(c) N =
a b
0 d
M
2
(R ) : ad = 0
8. Prove that each set of matrices forms a group under multiplication (don’t memorize these un-
less you really love matrices. . . ).
(a) Special linear group: SL
n
(R ) =
A M
n
(R ) : det A = 1
(b) Special orthogonal group: SO
n
(R ) = {A M
n
(R ) : A
T
A = I and det A = 1}
(c) Q
n
=
A M
n
(R ) : det A Q
×
(d) (Hard) SL
n
(Z ) =
A M
n
(Z ) : det A = 1
: all entries in these matrices are integers.
(Hint: look up the classical adjoint adj A of a square matrix)
Now construct a diagram showing the subgroup relationships between the groups
GL
n
(R ), SL
n
(R ), O
n
(R ), SO
n
(R ), Q
n
, SL
n
(Z )
9. (a) Let X be any set. Prove that composition of functions f : X X is associative.
(Hint: ( f g) h = f (g h) means that both functions do the same thing to the same input. . . )
(b) Suppose X contains at least two distinct elements a = b. Prove that there exist functions
f , g : X X for which f g = g f .
10. (a) Prove that matrix multiplication of (real square) matrices is associative.
(Hint: If A has entries (a
ij
), etc., what are the pq
th
entries of the matrices A(BC) and (AB)C?)
(b) Show that multiplication of (invertible) n ×n matrices is non-commutative when n 2.
11. Prove that D
n
is non-abelian (n 3).
(Hint: label vertices and proceed as is Example 2.25)
12. Consider rotating (in 3D) a regular tetrahedron. Any face (equilat-
eral triangle) may be rotated to its desired location (four options),
in which it has three possible orientations. The rotation group of
the tetrahedron therefore has 4 ×3 = 12 elements.
(a) Find the order of the rotation group of a cube.
(b) Repeat for a regular octahedron. Give a geometric reason
why your answer is the same as part (a).
(Hint: Try joining the midpoints of each face. . . ).
(c) What about the dodecahedron and the icosahedron?!
Common polyhedral dice
In case you don’t recognize it, the pictured red die is not one of the five Platonic solids: it has
ten kite-shaped faces, and its rotation group has order 10. We’ll return to these examples in
later sections.
16
2.5 Homomorphisms & Isomorphisms
In the previous sections, you should have felt like you were encountering similar examples in dif-
ferent contexts. A key goal of abstract mathematics is the comparison of similar/identical structures
with outwardly different appearances. The standard approach to such comparison is uses functions.
Example 2.29. Compare the rotation group R
3
of an equilateral triangle to the modular arithmetic
group Z
3
. Their Cayley tables look almost identical, particularly if we write ρ
0
for the identity in R
3
.
To a mathematician, the groups have the same structure; they are merely labelled differently.
Relabelling means defining an invertible function µ : R
3
Z
3
:
the obvious choice from looking at the tables is µ(ρ
k
) = k.
Since the tables describe all possible interactions between the
elements of each group, it is clear that µ satisfies
ρ
j
, ρ
k
R
3
, µ(ρ
j
ρ
k
) = µ(ρ
j
) +
3
µ(ρ
k
)
ρ
0
ρ
1
ρ
2
ρ
0
ρ
0
ρ
1
ρ
2
ρ
1
ρ
1
ρ
2
ρ
0
ρ
2
ρ
2
ρ
0
ρ
1
+
3
0 1 2
0 0 1 2
1 1 2 0
2 2 0 1
(R
3
, ) (Z
3
, +
3
)
Indeed, both sides simply equal j +
3
k! This is a critical formula. To see why, suppose we are given
remainders x, y Z
3
and consider two courses of action:
1. First combine the remainders x +
3
y in Z
3
, then map to R
3
using the function to obtain µ(x +
3
y).
2. First map both remainders to R
3
using µ, then combine in R
3
to obtain µ(x) µ(y).
Regardless of the order (combine/map or map/combine) we always obtain the same result! Such
structure-preserving functions are at the heart of abstract algebra.
Definition 2.30 (Homo- & Isomorphisms). Suppose (G, ) and (H, ) are binary structures and
ϕ : G H a function. We say that ϕ is a homomorphism of binary structures if
x, y G, ϕ (x y) = ϕ(x) ϕ(y)
An isomorphism
10
of binary structures G, H is a bijective/invertible homomorphism µ : G H.
Binary structures G, H are isomorphic, written G
=
H, if there exists some isomorphism µ : G H.
The notation is typical: µ (rather than ϕ) is often used when we know we have an isomorphism. For
most of these notes (certainly after this chapter), all binary structures will be groups.
Examples 2.31. 1. (2.29 cont.) We have isomorphic groups R
3
=
Z
3
. Indeed the function µ : R
3
Z
3
is an isomorphism of groups (or group isomorphism).
2. The function ϕ : (N, +) (R, +) defined by ϕ(x) =
2x is a homomorphism (of binary
structures: (N, +) is not a group!),
ϕ(x + y) =
2(x + y) =
2x +
2y = ϕ(x) + ϕ(y)
As before, it is worth spelling this out:
Sum then map ϕ(x + y) gives the same result as map then sum ϕ(x) + ϕ(y).
10
These terms come from ancient Greek: homo- (similar, alike), iso- (equal, identical), and morph(e) (shape, structure).
17
3. If V, W are vector spaces then every linear map T : V W is a group homomorphism:
11
v
1
, v
2
V, T(v
1
+ v
2
) = T(v
1
) + T(v
2
)
You’ve been encountering homomorphisms your entire mathematical career: for instance, both
the calculus identity
d
dx
( f + g) =
d f
dx
+
dg
dx
and the distributive law of matrix multiplication
A(x + y) = Ax + Ay are homomorphism properties!
4. (Trivial homomorphism) If G and L are any groups, then the function ϕ : G L defined by
ϕ(g) = e
L
(the identity in L) is a homomorphism:
x, y G, ϕ (xy) = e
L
= (e
L
)
2
= ϕ(x)ϕ (y)
5. (Inclusion map) If H is a subgroup of G, then ϕ(x) = x defines a homomorphism ϕ : H G:
x, y G, ϕ (xy) = xy = ϕ(x)ϕ(y)
6. The function ϕ(ρ
k
) = ρ
2k (mod 4)
is a homomorphism ϕ : R
4
R
4
:
ϕ(ρ
j
ρ
k
) = ϕ(ρ
j+k (mod 4)
) = ρ
2(j+k) (mod 4)
= ρ
2j (mod 4)
ρ
2k (mod 4)
= ϕ(ρ
j
) ϕ(ρ
k
)
Establishing Isomorphicity
We must do four things if we suspect binary structures (G, ) and (H, ) to be isomorphic:
Definition: Define µ : G H and, if necessary, verify that it is a function.
12
Homomorphism: Verify that µ(x y) = µ(x) µ (y) for all x, y G.
Injectivity/1–1: Check that µ(x) = µ(y) = x = y.
Surjectivity/onto: Check range µ = H. Equivalently h H, g G such that h = µ(g).
The last three steps can be done in any order, and injectivity/surjectivity can be combined if you
manage to exhibit an explicit inverse function µ
1
: H G. If you are unsure how to start, often the
best thing is to play with the homomorphism property itself.
Examples 2.32. 1. We show that ( 2Z, +) and (3Z, +) are isomorphic groups.
Definition: The obvious candidate
13
is ϕ(x) =
3
2
x. Plainly ϕ(2n) = 3n whence ϕ : 2Z 3Z.
Homomorphism: ϕ(x + y) =
3
2
(x + y) =
3
2
x +
3
2
y = ϕ(x) + ϕ(y)
Injectivity: ϕ(x) = ϕ(y) =
3
2
x =
3
2
y = x = y.
Surjectivity: If z = 3n 3Z, then z =
3
2
·
2
3
z =
3
2
(2n) = ϕ(2n) range ϕ.
The last steps are essentially the observation that ϕ
1
(z) =
2
3
z.
More generally, the groups (mZ, +) and (nZ, +) are isomorphic whenever m, n = 0.
11
The scalar multiplication condition T(λv) = λT(v) of a linear map is not relevant here.
12
If G is a set of equivalence classes, we also need to check that ϕ is well-defined. This subtlety is why we haven’t (yet)
had an example where Z
n
is the domain of a homomorphism. We will do so later (Example 3.5.2, Theorem 3.7, etc.).
18
2. We prove that (R, +)
=
(R
+
, ·): these are isomorphic abelian groups (recall that R
+
= (0, )
is the set of positive real numbers).
Definition/Homomorphism: We need a bijective function µ : R R
+
which converts addi-
tion to multiplication µ(x + y) = µ(x)µ(y). But exponentiation does exactly this: defining
µ(x) = e
x
, we see that the homomorphism property is the familiar exponential law!
µ(x + y) = e
x+y
= e
x
e
y
= µ(x)µ (y)
Bijectivity: µ
1
(z) = ln z is the inverse function of µ.
Other exponential functions also provide suitable isomorphisms: e.g. 2
x
, 10
x
, etc.
Demonstrating Non-Isomorphicity (Structural Properties)
Suppose we suspect that binary structures (G, ) and (H, ) are non-isomorphic. Unless G, H are very
small, individually verifying that every possible function µ : G H is a non-isomorphism would
be unrealistic! Instead we consider structural properties: properties that isomorphic structures must
share. If any such is held by one structure but not the other, then the structures are non-isomorphic.
Here is a non-exhaustive list of structural properties; we’ll check some in Exercise 11. Throughout,
we assume that µ : (G, ) (H, ) is an isomorphism.
Cardinality/order: Since G and H are bijectively paired, their cardinalities are the same.
Commutativity & Associativity: Suppose (G, ) is commutative and let X, Y H. Since µ is surjec-
tive, we may write X = µ(x) and Y = µ(y) for some x, y G. The homomorphism property
now shows that (H, ) is commutative:
X Y = µ(x) µ(y) = µ(x y) = µ(y x) = µ(y) µ(x) = Y X
The argument for associativity is similar, though more tedious.
Identities & Inverses: If (G, ) has identity e, then µ(e) is the identity for (H, ). Similarly µ maps
inverses to inverses.
Solutions to equations: Related equations have the same number of solutions. For instance,
x x = x µ(x) µ(x) = µ(x)
says that the equations x x = x (in G) and z z = z (in H) have the same number of solutions.
14
Being a group If G is a group, so also is H.
Examples 2.33. 1. Recall that N
0
= {0, 1, 2, 3, . . .}. Since (N
0
, +) contains the identity element 0
whereas (N, +) has no identity, we conclude that these binary structures are non-isomorphic.
2. Z
5
is not isomorphic to D
3
since the two groups have different orders (5 and 6).
13
You might think, “How can I turn an even number into a multiple of three?” Of perhaps you start by thinking about
the homomorphism property: multiplication by a constant certainly satisfies ϕ(x + y) = ϕ(x) + ϕ(y).
14
Such solutions are called idempotents; thus existence of idempotents is itself a structural property.
19
3. The binary structures defined by the two tables are non-isomorphic.
For instance, the first is commutative while the second is not.
4. GL
2
(R ) and (R, +) are non-isomorphic for the same reason: the first
is non-abelian and the second abelian.
a b
a a b
b b a
c d
c c d
d c d
5. To see that (Q, +) and (R, +) are non-isomorphic groups, it is enough to recall that the sets
have different cardinalities: Q is countably infinite while R is uncountable.
6. The groups (Z, +) and (Q, +) have the same (countably infinite) order, and are both abelian. To
see that they are non-isomorphic, consider the equation x + x = 1 which has no solutions in Z.
If µ : Z Q were an isomorphism, then the equation µ(x) + µ(x) = µ(1) does have a solution
y = µ(x) =
1
2
µ(1) in Q. But then x = µ
1
(y) solves the original equation: contradiction!
7. (S
1
, ·) and (R, +) are non-isomorphic: consider the equations x x = e. . .
Many properties are non-structural and therefore cannot be used to show non-isomorphicity: the type
of element (number, matrix, etc.), the type of binary operation (addition, multiplication, etc.).
Transferring a Binary Structure
Suppose µ : G H is a bijection of sets, where one of G, H has a binary structure. A binary structure
may be defined on the other by insisting that µ be an isomorphism.
Example 2.34. The function µ(x) = x
3
+ 8 is a bijection R R. Starting with the binary (group)
structure (R, +) and treating µ as an isomorphism, we may create a new isomorphic structure. There
are two ways to do this:
Pull-back: Suppose µ : (R, ) (R, +). Since µ(x y) = µ(x) + µ(y), the new operation must be
x y := µ
1
µ(x) + µ(y)
= µ
1
(x
3
+ y
3
+ 16) =
3
q
x
3
+ y
3
+ 8
All structural properties transfer from (R, +) to (R, ): for instance, (R, ) is an abelian group
with identity element
µ
1
(0) =
3
8 = 2
As a sanity check, observe that we really do have x (2) =
3
p
x
3
+ (2)
3
+ 8 = x!
Push-forward: View µ : (R, +) (R, ) as an isomorphism. Computation of is an exercise.
“Up to Isomorphism”
This phrase is ubiquitous in group theory. To illustrate, consider that if
{e, a},
is a group with identity e, then its Cayley table must be as shown (Example 2.10.6).
Otherwise said: there may be infinitely many distinct groups of order two, but all are
isomorphic to each other. This is too wordy, so a mathematician might instead say:
e a
e e a
a a e
Up to isomorphism, there is a unique group of order two.
Make sure you include the snippet “up to isomorphism,” for otherwise the sentence is false!
20
The start of group theory can feel very challenging. With its focus on functions and its unfamiliar
words, this last introductory section likely seems particularly so. Complete fluency with the vocab-
ulary is not required at this stage. The remaining chapters provide plenty opportunity to reinforce the
language introduced in this chapter.
For the same reason, several of the following Exercises (particularly number 11 onwards) will likely
seem difficult. Try these (and discuss them) now, even if you aren’t sure what to do; return later
when you feel more comfortable. Learning abstract concepts isn’t quick; give the ideas a chance to
sink in. By the end of the course, these Exercises should seem much easier!
Exercises 2.5. Key concepts: Homomorphism Isomorphism Injective/surjective/bijective
Structural property ‘Up to isomorphism’
1. Which of the following are homomorphisms/isomorphisms of binary structures? Explain.
(a) ϕ : (Z, +) (Z, +), ϕ(n) = n (b) ϕ : (Z, +) (Z, +), ϕ(n) = n + 1
(c) ϕ : (Q, +) (Q, +), ϕ(x) =
4
3
x (d) ϕ : (Q, ·) (Q, ·), ϕ(x) = x
2
(e) ϕ : (R, ·) (R, ·), ϕ(x) = x
5
(f) ϕ : (R, +) (R, ·), ϕ(x) = 2
x
(g) ϕ : (M
2
(R ), ·) (R, ·), ϕ(A) = det A
(h) ϕ : (M
n
(R ), +) (R, +), ϕ(A) = tr A (trace: add the entries on the main diagonal)
2. Show that ( Z, +)
=
(nZ, +) for any non-zero constant n.
3. Prove or disprove: (R
3
, +)
=
(R
3
, ×) (cross product).
4. µ(n) = 2 n is a bijection of Z with itself. For each of the following, define a binary relation
on Z such that µ is an isomorphism.
(a) µ : (Z, ) (Z, +) (b) µ : (Z, ) (Z, ·) (c) µ : (Z, ) (Z, max(a, b))
5. Finish Example 2.34 by computing the push-forward X Y for any X, Y R.
6. µ(x) = x
2
is a bijection µ : R
+
R
+
. Find x y if µ is to be an isomorphism.
(a) µ : (R
+
, ) (R
+
, +) (b) µ : (R
+
, +) (R
+
, ) (c) µ : (R
+
, ) (R
+
, ·)
7. Show that x y = x + y xy is the pull-back of (R
×
, ·) to R \ {1} by µ(x) = 1 x. Use this to
provide an alternative quick argument for Exercise 2.1.9.
8. Recall Exercise 2.3.6c. Prove that the Klein four-group and Z
×
8
are isomorphic.
9. (a) Prove that S :=
a b
b a
M
2
(R )
forms a group under matrix addition.
(b) Prove that T = S \{0} (S without the zero matrix) forms a group under matrix multipli-
cation.
(c) Define ϕ
a b
b a
= a + ib. Prove that ϕ : S C and ϕ
T
: T C
×
are both isomorphisms
ϕ : (S, +)
=
(C, +), ϕ
|
T
: (T, ·)
=
(C
×
, ·)
(In a future class, ϕ will be described as an isomorphism of rings/fields)
10. (Recall Exercise 2.4.8 and Footnote 9) Prove that S
1
=
SO
2
(R ) via an isomorphism
µ(e
iθ
) =
cos θ sin θ
sin θ cos θ
21
11. Suppose µ : (G, ) (H, ) is an isomorphism of binary structures. Prove:
(a) If e is the identity for G, then µ(e) is the identity for H.
(b) If x G has an inverse y, then µ(y) is an inverse to µ(x) (in H).
(c) Suppose ϕ : G H is a group homomorphism. Show that parts (a), (b) still hold: ϕ(e
G
) = e
H
and ϕ(x
1
) =
ϕ(x)
1
.
For a challenge: What happens if ϕ is merely a homomorphism of binary structures?
12. Given a group homomorphism ϕ : G H , define the image ϕ(G) and kernel ker ϕ as follows:
ϕ(G) = Im ϕ =
ϕ(x) : x G
, ker ϕ :=
x G : ϕ(x) = e
(a) Compute the image and kernel of ϕ : (R
×
, ·) (R
×
, ·) where ϕ(x) = x
2
.
(b) Prove that ϕ(G) is a subgroup of H (in general, not just for the example in (a)!).
(c) Prove that ker ϕ is a subgroup of G.
(We’ll return to these important concepts later)
13. The groups (Q, +) and (Q
+
, ·) are both abelian and both have the same cardinality: nonethe-
less, we prove that they are non-isomorphic.
Assume, for contradiction, that µ : Q Q
+
is an isomorphism.
(a) If c Q is constant, what equation in Q
+
corresponds to x + x = c?
(b) By considering the number of solutions to the equations in part (a), obtain a contradiction
and hence conclude that ( Q, +) (Q
+
, ·).
(Extra challenge) Suppose ϕ : (Q, +) (R, ·) is a homomorphism (of binary structures) and that
ϕ(1) = a: find a formula for ϕ(x).
14. Recall the magic square property (Exercise 2.1.11).
(a) Up to isomorphism, explain why there is a unique group of order three.
(This is another reason the groups in Example 2.29 must be isomorphic!)
(b) Show that, up to isomorphism, there are precisely two groups of order four.
(Hint: If G = {e, a, b, c}, why may we assume, without loss of generality, that b
2
= e? Your
answers should look like the Klein four-group V and the group Z
4
.)
(c) (Hard) What happens for order five?
15. Prove that isomorphic is an equivalence relation on any collection of groups. That is, for all
groups G, H, K:
Reflexivity: G
=
G.
Symmetry: G
=
H = H
=
G.
Transitivity: G
=
H and H
=
K = G
=
K.
22
3 Cyclic & Finite Abelian Groups
In this chapter we consider a general family of groups and see how to combine these to describe any
finite abelian group.
3.1 Definitions and Basic Examples
The foundational idea of a cyclic group is that it may be generated from a single element.
Examples 3.1. 1. The integers (Z, +) are generated by the element 1: all integers may be produced
by repeatedly combining 1 using only the group operation (+) and inverses (). For instance,
4 = (1 + 1 + 1 + 1)
2. The modular arithmetic groups (Z
n
, +
n
) (Section 2.3) are also generated by (the remainder) 1.
Since the group is finite, inverses are not necessary. For instance,
Z
4
= {0, 1, 2, 3} = {0, 1, 1 + 1, 1 + 1 + 1}
3. The group R
n
of rotations of a regular n-gon (Definition 2.24) is generated by the ‘1-step’ rota-
tion ρ
1
: that is, ρ
k
= ρ
k
1
.
We formalize this idea by considering the subset of a group that may be produced from a single
element, the group operation, and inverses.
Lemma 3.2 (Cyclic subgroup). Let G be a group and g G. The set
g
:= {g
n
: n Z} = {. . . , g
1
, e, g, g
2
, . . .}
is a subgroup of G. We call this the cyclic subgroup
15
generated by g.
Proof. We follow the subgroup criterion (Theorem 2.14).
Non-emptiness: Plainly g
g
.
Closure: Every element of
g
has the form g
k
for some k Z. The required condition
follows from standard exponential notation (Definition 2.3): g
k
· g
l
= g
k+l
g
.
Inverses: This is Exercise 2.1.7c: (g
k
)
1
= g
k
g
.
Definition 3.3 (Cyclic group). A group G is cyclic if it has a generator: g G such that G =
g
.
In any group G, the order of an element g is the order (cardinality) of the cyclic subgroup
g
G.
Warning! Don’t confuse the order of a group G with the order of an element g G. Cyclic groups are
precisely those containing elements (generators) whose order equals that of the group!
15
Since this is an abstract result, the lemma is written multiplicatively. If G is an additive group, then cyclic subgroups
are written
g
= {ng : n Z} = {. . . , 2g, g, 0, g, 2g, 3g, . . .}. As in Example 3.1.2, for finite cyclic groups convention
dictates that the identity element is written first, e.g.
g
= {e, g, g
2
, . . . , g
n1
}.
23
Examples (3.1 cont). 1. Z =
1
=
1
is generated by either 1 or 1. The cyclic subgroup
generated by 2 is the group of even numbers under addition
2
= {. . . , 2, 0, 2, 4, . . .} = {2m : m Z} = 2Z
2. Z
n
is generated by both 1 and 1 = n 1, but may have other generators (we’ll consider how
to find them all shortly). For instance, Z
5
is generated also by 2:
2
= {0, 2, 2 + 2, 2 + 2 + 2, . . .} = {0, 2, 4, 1, 3} = Z
5
3. R
n
=
ρ
1
= {e, ρ
1
, ρ
2
1
, . . . , ρ
n1
1
}. As with Z
n
, this group typically has other generators.
Another commonly encountered family of cyclic groups arise as subgroups of (C
×
, ·) (or (S
1
, ·)).
Definition 3.4 (Roots of Unity). Let n N. The group of n
th
roots of unity U
n
is the cyclic subgroup
of (S
1
, ·) generated by ζ := e
2πi
n
:
U
n
:=
ζ
=
1, ζ, ζ
2
, ··· , ζ
n1
These are precisely the n complex solutions to the equation z
n
= 1. To emphasize n, write ζ
n
= e
2πi
n
.
For instance U
2
=
1
= {1, 1} and U
4
=
i
= {1, i, 1, i}. In general,
the n
th
roots are the vertices of a regular n-gon centered at 0 with radius 1:
ζ
k
=
|
ζ
|
k
= 1 and arg ζ
k
= arg e
2πk
n
=
2πk
n
= k arg ζ
We stop listing the elements at ζ
n1
since ζ
n
= e
2πi
= 1. The periodicity of
the complex exponential (e
iθ
= 1 θ 2πZ) results in a simple tie-in
with modular arithmetic:
ζ
k
= ζ
l
1 = ζ
kl
= e
2πi(kl)
n
k l (mod n)
1 1
ζ
ζ
2
ζ
3
ζ
4
ζ
5
ζ
6
i
i
Seventh roots: ζ
7
= e
2πi
7
Examples 3.5. 1. Observe that ζ
2
6
= (e
2πi
6
)
2
= e
2πi
3
= ζ
3
.
This produces a subgroup relationship: writing ζ = ζ
6
, we have
U
3
=
1, ζ
2
, ζ
4
< U
6
=
1, ζ, ζ
2
, ζ
3
, ζ
4
, ζ
5
The picture makes this geometrically trivial (compare Example 2.26.2).
S
1
ζζ
2
ζ
3
ζ
4
ζ
5
1
2. (Example 2.29, cont.) Below is the Cayley table for U
3
. Writing 1 = ζ
0
and ζ = ζ
1
makes the
isomorphic relationship with (Z
3
, +
3
) and (R
3
, ) obvious: (U
3
, ·)
=
(Z
3
, +
3
)
=
(R
3
, ).
· 1 ζ ζ
2
1 1 ζ ζ
2
ζ ζ ζ
2
1
ζ
2
ζ
2
1 ζ
· ζ
0
ζ
1
ζ
2
ζ
0
ζ
0
ζ
1
ζ
2
ζ
1
ζ
1
ζ
2
ζ
0
ζ
2
ζ
2
ζ
0
ζ
1
+
3
0 1 2
0 0 1 2
1 1 2 0
2 2 0 1
ρ
0
ρ
1
ρ
2
ρ
0
ρ
0
ρ
1
ρ
2
ρ
1
ρ
1
ρ
2
ρ
0
ρ
2
ρ
2
ρ
0
ρ
1
24
For a little practice here is a formal argument that Z
3
=
U
3
: we show explicitly that
µ : Z
3
U
3
: x 7 ζ
x
is an isomorphism. Since the domain Z
3
consists of equivalence classes, this requires a little care.
Well-definition: We must prove that if x = y in Z
3
, then µ(x) = µ(y).
Given x = y Z
3
, then (as integers) x = y + 3k for some integer k. But then
µ(x) = ζ
x
= ζ
x+3k
= ζ
y
(ζ
3
)
k
= ζ
y
= µ(y)
Homomorphism: µ(x + y) = ζ
x+y
= ζ
x
ζ
y
= µ(x)µ (y)
Injectivity: µ(x) = µ(y) = ζ
x
= ζ
y
= ζ
xy
= 1 = x y (mod 3) = x = y in Z
3
.
(Notice how injectivity is the converse of well-definition!)
Surjectivity: range µ = {ζ
x
: x Z} = {1, ζ, ζ
2
} = U
3
, since ζ
x+3k
= ζ
x
.
In the next section we’ll essentially repeat this discussion in the abstract, so make sure this
example makes sense before moving on.
Exercises 3.1. Key concepts: Generator Order of an element Cyclic (sub)group Roots of unity
1. Compute the cyclic subgroup
12
of Z
20
(write the elements in the order generated).
2. Find/describe all the generators of each cyclic group.
(a) ( Z, +) (b) {2
n
3
n
: n Z} under multiplication
(c) ( Z
5
, +
5
) (d)
a 0
0 a
,
0 b
b 0
: a, b = ±1
under multiplication
3. State all cyclic subgroups of Z
9
. What is the order of each element?
4. Recall Example 2.25. What is the cyclic subgroup of D
3
generated by ρ
1
? Generated by µ
1
?
5. (a) Find all cyclic subgroups of the Klein four-group V. What is the order of each element?
(b) V is a finite non-cyclic group. Give an example of an infinite non-cyclic group, and explain
how you know you are correct.
6. Compute the cyclic subgroup
ζ
5
8
of U
8
, listing its elements in the order generated.
7. (a) Prove that (U
3
, ·) is a subgroup of (U
9
, ·).
(b) Complete the sentence and prove your assertion:
U
m
U
n
if and only if (relationship between m and n)
8. (a) Show that Z
×
5
= {1, 2, 3, 4} forms a cyclic group under multiplication modulo 5.
(b) What about Z
×
8
= {1, 3, 5, 7} under multiplication modulo 8? To what well-known group
is this isomorphic?
9. Suppose that a cyclic group G has order
|
G
|
3. Explain why it has at least two generators.
10. Modeling Example 3.5.2, prove explicitly that Z
n
=
U
n
for any n N.
11. In contrast to the real case (Example 2.32.2), verify that ϕ : C C
×
: z 7 e
z
is a homomorphism
(C, +)
=
(C
×
, ·) but not an isomorphism.
25
3.2 The Classification and Structure of Cyclic Groups
This section is significantly more abstract that what has come before: take your time, read carefully,
and use the examples to help. Our goal is to describe general properties of all cyclic groups, including
their generators and subgroup structures. The first result is, mercifully, very simple.
Lemma 3.6. Every cyclic group is abelian.
Proof. Let G =
g
. Any two elements of G can be written g
k
, g
l
for some k, l Z. The result follows
from standard exponential notation and the fact that addition of integers is commutative:
g
k
g
l
= g
k+l
= g
l+k
= g
l
g
k
Note that the converse is false: the Klein four-group V is abelian but not cyclic (Exercise 3.1.5).
Our major classification theorem proves a pattern you’ve likely already guessed.
Theorem 3.7 (Isomorphs). Every cyclic group G =
g
is isomorphic either to Z or to some Z
n
:
If G is infinite, then G
=
Z.
If G is finite with order n, then G
=
Z
n
.
In either case, µ : Z
(n)
G : x 7 g
x
is an explicit isomorphism.
To help understand the theorem and set up the proof, some remarks and examples are helpful.
Map Generator to Generator The isomorphism maps the generator 1 Z
(n)
to the generator g of G.
Together with the homomorphism property, this idea defines µ
µ(1) = g = µ(x) = µ(x + ···+ x) =
µ(1)
x
= g
x
This observation makes it easy to find suitable isomorphisms in examples.
The Cyclic Group of Order n? The theorem says that, up to isomorphism, there is a unique cyclic
group of order n. For this reason, many algebraists use the symbol Z
n
(or C
n
for ‘cyclic’) to
refer to any example of a cyclic group of order n (e.g. R
n
, U
n
, etc.). For clarity, in these notes,
Z
n
will always be the explicit group of remainders under addition modulo n.
The Order of G =
g
It is helpful to introduce a set of positive integers
S := {m N : g
m
= e} (mg = e if G is additive)
This set plays a crucial role in the proof, distinguishing the finite/infinite cases and detecting
the order of G. . .
Examples 3.8. 1. Z
4
=
1
is additive, so S = {m N : m = 0 Z
4
} = {4, 8, 12, . . .}. The minimal
element 4 is plainly the order of
|
Z
4
|
.
2. (Example 3.5.2) In U
3
=
ζ
, we have ζ
m
= 1 3 | m, whence S = {3, 6, 9, . . .}. Plainly 3 is
the order of U
3
. Moreover µ(x) = ζ
x
is the isomorphism µ : Z
3
U
3
seen previously!
26
3. 5Z =
5
is an infinite cyclic group. In this case, S = {m N : 5m = 0} = is empty. We have
an isomorphism µ : Z 5Z : x 7 5x (map the generator 1 of Z to the generator 5 of 5Z).
Proof. We first establish that µ is a bijection. The generic cases depend on the minimal element of S.
Case 1: S = . Suppose x > y and that g
x
= g
y
. Then g
xy
= e = x y S: contradiction. The
elements . . . , g
2
, g
1
, e, g, g
2
, . . . are distinct, and so µ : Z G : x 7 g
x
is bijective.
Case 2: min S = n. We first check that µ : Z
n
G : x 7 g
x
is well-defined:
y = x Z
n
= y = x + kn for some k Z (as integers)
= µ(y) = g
y
= g
x+kn
= g
x
(g
n
)
k
= g
x
= µ(x) (n S, so g
n
= e)
This moreover tells us that G is finite (there are at most n distinct elements of G)
G =
g
= {. . . , g
2
, g
1
, e, g, g
2
, . . .} = {e, g, . . . , g
n1
}
Now suppose two of these terms were equal; if 0 y x n 1, then
g
x
= g
y
= g
xy
= e = x = y (0 x y n 1 < n = min S)
We conclude that G = {e, g, . . . , g
n1
} has order n, and that µ is a bijection.
Finally, note that the homomorphism property in both cases is merely standard exponential notation
µ(x + y) = g
x+y
= g
x
g
y
= µ(x)µ (y)
As advertised, the use of S in the proof yields a useful alternative notion for the order of an element.
Corollary 3.9 (Order of an element). If finite, the order of g equals the minimal positive integer n
for which g
n
= e. Moreover g
m
= e n | m.
Otherwise said, if g has order n, then S = {m N : g
m
= e} = {kn : k N} is the set of positive
multiples of n.
Examples 3.10. 1. The group of 7
th
roots of unity U
7
is isomorphic to Z
7
via µ : Z
7
U
7
: k 7 ζ
k
7
.
As a sanity check, observe that 7 = min{m N : ζ
m
7
= 1} is indeed the order of ζ
7
= e
2πi
7
.
2. ( R, +) is non-cyclic since its (uncountable) cardinality is larger than that of the integers. This is
also straightforward directly: if R =
x
were cyclic (x = 0), then we obtain the contradiction
x
2
/ {. . . , 2x, x, 0, x, 2x, 3x . . .} =
x
= R
x
2
The same argument shows that, for instance, that (Q, +) is non-cyclic.
3. Let ξ = e
2πi
2
and consider the cyclic subgroup G :=
ξ
< (C
×
, ·). For integers m, observe that
ξ
m
= e
2πim
2
= 1
m
2
Z m = 0
We conclude that G is an infinite cyclic group and that µ : Z G : z 7 ξ
z
is an isomorphism.
Multiplication by ξ essentially performs an irrational fraction (
1
2
) of a full rotation.
27
Subgroups of Cyclic Groups are also Cyclic!
For the remainder of this section we describe the complete subgroup structure of all cyclic groups.
Our approach is motivated by a simple example.
Example 3.11. The cyclic subgroup 2Z Z is generated by 2: the minimal positive integer in the
subgroup.
For an abstract subgroup H of a cyclic group G =
g
, our goal is to identify a suitable ‘minimal
element’ of H and then demonstrate that this generates H.
Theorem 3.12 (Subgroups of Cyclic Groups). Every subgroup of a cyclic group is cyclic.
Proof. Suppose H is a subgroup of G =
g
. Let s be the smallest positive integer
16
such that g
s
H.
We prove that H is generated by g
s
by establishing the set equality H =
g
s
.
() Since H is a group, the closure and inverse axioms force (g
s
)
k
H for all k Z.
() Let g
m
H. We must prove that g
m
g
s
. By the division algorithm, there exist unique
integers q, r such that
m = qs + r and 0 r < s
Since H satisfies the closure and inverse axioms,
g
m
= g
qs+r
= (g
s
)
q
g
r
= g
r
= (g
s
)
q
g
m
H ()
The minimality of s forces r = 0, from which we conclude that g
m
= (g
s
)
q
g
s
.
If the proof seems hard, try rewriting it for our motivational example, where G = Z, H = 2Z and
s = 2; remember that G is additive, so () is simply r = 2s + m 2Z!
We finish by considering the finite and infinite cases separately. The latter is very simple.
Corollary 3.13 (Subgroups of infinite cyclic groups). If G is an infinite cyclic group and H G,
then either H = {e} is trivial, or H
=
G.
The proof as an exercise—just generalize the following generic example!
Example 3.14. We write things out explicitly in additive notation when G = Z. By Theorem 3.12,
every subgroup has the form
s
= sZ (the multiples of s Z). There are two generic situations:
If s = 0 we have the trivial subgroup:
0
= {0}.
If s = 0, then sZ is isomorphic to Z via the isomorphism µ : Z sZ : x 7 sx.
16
This exists for two reasons:
1. H is a non-empty subset of G =
g
and so must contain some element g
k
. We may assume k 1 since the subgroup
also contains g
k
(inverse axiom).
2. The set of natural numbers {k N : g
k
H} is non-empty and so (well-ordering) has a minimal element s.
Note that s = 1 H =
e
is the trivial subgroup (g
1
= e
1
= e H). In this special case the division algorithm argument
is trivial: every m = q · 1 + 0 is immediately a multiple of 1, and g
m
= e for all m.
28
Finite cyclic groups are more complicated, so we first illustrate the main result with an example.
Example 3.15. Consider the 6
th
roots of unity U
6
= {1, ζ, ζ
2
, ζ
3
, ζ
4
, ζ
5
}. Since every subgroup of U
6
is
necessarily cyclic (Theorem 3.12), it is enough to consider each cyclic subgroup
z
in turn (for each
z U
6
). We obtain three proper subgroups, as arranged in the subgroup diagram below.
z subgroup
z
1 {1}
ζ {1, ζ, ζ
2
, ζ
3
, ζ
4
, ζ
5
}
ζ
2
{1, ζ
2
, ζ
4
}
ζ
3
{1, ζ
3
}
ζ
4
{1, ζ
4
, ζ
2
}
ζ
5
{1, ζ
5
, ζ
4
, ζ
3
, ζ
2
, ζ}
ζ
= U
6
ζ
2
= U
3
ζ
3
= U
2
1
= U
1
Observe the repetitions:
ζ
=
ζ
5
= U
6
and
ζ
2
=
ζ
4
= U
3
. Note also that we really do have
equality of groups here: for instance, ζ
2
= (e
2πi
6
)
2
= e
2πi
3
is a generator of U
3
.
For comparison, here is the same data for subgroups of the additive group (Z
6
, +
6
).
x subgroup
x
0 {0}
1 {0, 1, 2, 3, 4, 5}
2 {0, 2, 4}
3 {0, 3}
4 {0, 4, 2}
5 {0, 5, 4, 3, 2, 1}
1
= Z
6
2
=
Z
3
3
=
Z
2
0
=
Z
1
Since U
6
=
Z
6
, differences are entirely notational. One subtle distinction is that we don’t use equals in
the second subgroup diagram: for instance,
2
= {0, 2, 4} is isomorphic but not equal to Z
3
= {0, 1, 2}.
The Example should suggest a pattern (previewed in Lemma 2.21): the subgroups of Z
n
are precisely
those generated by the divisors of n, with one subgroup for each divisor:
d | n =
d
=
Z
n
d
, moreover gcd(s, n) = d =
s
=
d
Our final result merely asserts this for general finite cyclic groups.
Corollary 3.16 (Subgroups of finite cyclic groups). Let G =
g
have order n. For each divisor of n,
G has a unique subgroup with this order; these are moreover the only subgroups of G.
More precisely,
d = gcd(s, n) =
g
s
= g
d
, where this subgroup has order
n
d
(isomorphic to Z
n
d
)
In particular: g
s
has order
n
gcd(s,n)
and generates G if and only if gcd(s, n) = 1.
As with Theorem 3.12, if the proof seems hard, try rewriting it when G = Z
n
.
29
Proof. Suppose d = gcd(s, n). We prove set inclusion in both directions.
() Since d divides s, we have s = kd for some k Z. But then
g
s
= (g
d
)
k
g
d
= (g
s
)
t
= (g
d
)
kt
=
g
s
g
d
() Apply B
´
ezout’s identity (extended Euclidean alg.): d = κs + λn for some κ, λ Z, whence
g
d
= (g
s
)
κ
(g
n
)
λ
= (g
s
)
κ
g
s
=
g
d
g
s
To finish, note that since d | n, there are precisely
n
d
elements of
g
d
:
g
d
=
e, g
d
, g
2d
, . . . , g
nd
Example 3.17. 1. Z
8
is generated by 1, 3, 5 and 7: precisely the elements for which gcd(s, 8) = 1. For
example,
5
= {5, 2, 7, 4, 1, 6, 3, 0} = Z
8
Similarly, the subgroup
6
= {6, 4, 2, 0} has order 4 =
8
gcd(6,8)
. The complete
collection of subgroups is in the table: the first column lists each divisor d of 8 (the
possible values of gcd(x, 8)), while the second column has the explicit subgroup
generated by each x, and the group isomorph Z
8
d
. The ‘smallest’ generator is used
for each subgroup in the subgroup diagram.
d = gcd(x, 8) Subgroup
x
=
Z
8
d
1 {0, 1, 2, 3, 4, 5, 6, 7}= Z
8
2 {0, 2, 4, 6}
=
Z
4
4 {0, 4}
=
Z
2
8 {0}
=
Z
1
1
= Z
8
2
=
Z
4
4
=
Z
2
0
=
Z
1
2. We repeat the discussion for Z
30
.
d = gcd(x, 30) Subgroup
x
=
Z
30
d
1 {0, 1, 2, 3, . . . , 7, . . . , 11, 12, 13, . . .
17, 18, 19, . . . , 23, . . . , 29}= Z
30
2 {0, 2, 4, 6, 8, 10, 12, 14, 16,
18, 20, 22, 24, 26, 28}
=
Z
15
3 {0, 3, 6, 9, 12, 15, 18, 21, 24, 27}
=
Z
10
5 {0, 5, 10, 15, 20, 25}
=
Z
6
6
{0, 6, 12, 18, 24}
=
Z
5
10 {0, 10, 20}
=
Z
3
15 {0, 15}
=
Z
2
30 {0}
=
Z
1
1
= Z
30
2
=
Z
15
3
=
Z
10
5
=
Z
6
6
=
Z
5
10
=
Z
3
15
=
Z
2
0
=
Z
1
You should consider how the shape of the subgroup diagram for Z
n
depends on the prime decomposi-
tion of n: for instance, each prime appears exactly once in 30 = 2 · 3 ·5.
30
Exercises 3.2. Key concepts: Every cyclic group is isomorphic to Z or Z
n
g
m
= e (ord g) | m
g
=
Z
n
=
g
s
=
Z
n
gcd(s,n)
Subgroup diagrams for finite cyclic groups
1. Construct the subgroup diagram and give a generator of each subgroup:
(a) ( Z
10
, +
10
) (b) (Z
42
, +
42
).
2. A generator of the cyclic group U
n
group is known as a primitive n
th
root of unity. For instance,
the primitive 4
th
roots are ±i. Find all the primitive roots when:
(a) n = 5 (b) n = 6 (c) n = 8 (d) n = 15
3. Find the complete subgroup diagram of U
p
2
q
where p, q are distinct primes.
(Hint: Try U
12
first if this seems too difficult)
4. If r N and p is prime, find all subgroups of (Z
p
r
, +
p
r
) and give a generator for each.
5. (a) Suppose µ : G H is an isomorphism of cyclic groups. If g is a generator of G, prove that
µ(g) is a generator of H. Do you really need µ to be an isomorphism here?
(b) If G is an infinite cyclic group, how many generators has it got?
(c) Recall Exercise 3.1.6b. Describe an isomorphism ϕ : Z
4
Z
×
5
.
6. True or false: In any group G, if g has order n, then g
s
has order
n
gcd(s,n)
. Explain.
7. Suppose G =
g
is infinite and H =
g
s
is an infinite subgroup. Prove Corollary 3.13 by
describing an isomorphism µ : G H.
8. Prove Corollary 3.9: you’ll need the division algorithm for the second part!
9. Let x, y be elements of a group G. If xy has finite order n, prove that yx also has order n.
(Hint: (xy)
m
= x(yx)
m1
y)
10. For which real numbers θ is the multiplicative cyclic group G =
e
2πiθ
C
×
finite? Describe
the order of G in terms of θ.
11. Let G be a group and X a non-empty subset of G. The subgroup generated by X is the subgroup
created by making all possible combinations of elements and inverses of elements in X.
(a) Explain why ( Z, +) is generated by the set X = {2, 3}.
(b) If m, n (Z, +), show X = {m, n} generates dZ, where d = gcd( m, n).
(c) The Klein four-group V is not cyclic, so it cannot be generated by a singleton set. Find a
set of two elements which generates V.
(d) Describe the subgroup of (Q, +) generated by X = {
1
2
,
1
3
}.
(e) (Hard) (Q, +) is plainly generated by the infinite set {
1
n
: n N}. Explain why (Q, +) is
not finitely generated: i.e. there exists no finite set X generating Q.
(Hint: Think about the prime factors of the denominators of elements of X)
31
3.3 Direct Products & Finite Abelian Groups
In this section we discuss a straightforward way to create new groups from old using the Cartesian
product. In the abstract, this discussion applies to any groups, though the ingredients in most of our
examples will be cyclic.
Example 3.18. Given Z
2
= {0, 1}, the Cartesian product
Z
2
×Z
2
=
(0, 0), (0, 1), (1, 0), (1, 1)
has four elements. This set inherits a binary structure via addition of co-ordinates
(x, y) + (v, w) := (x + v, y + w)
where x + v and y + w are both computed in (Z
2
, +
2
). This binary operation has an addition table
that should looks very familiar: it has exactly the same structure as the Klein four-group!
+ (0, 0) (0, 1) (1, 0) (1, 1)
(0, 0) (0, 0) (0, 1) (1, 0) (1, 1)
(0, 1) (0, 1) (0, 0) (1, 1) (1, 0)
(1, 0) (1, 0) (1, 1) (0, 0) (0, 1)
(1, 1) (1, 1) (1, 0) (0, 1) (0, 0)
e a b c
e e a b c
a a e c b
b b c e a
c c b a e
We conclude that Z
2
×Z
2
=
V is indeed a group.
This type of construction works in general.
Theorem 3.19 (Direct product). The natural component-wise operation on the Cartesian product
n
k=1
G
k
= G
1
×···× G
n
, (x
1
, . . . , x
n
) ·(y
1
, . . . , y
n
) := (x
1
y
1
, . . . , x
n
y
n
)
defines a group structure: the direct product. This group is abelian if and only if each G
k
is abelian.
The proof is a simple exercise. Being a Cartesian product, a direct product has order equal to the
product of the orders of its components
n
k=1
G
k
=
n
k=1
|
G
k
|
Examples 3.20. 1. The direct product of the groups (Z
2
, +
2
) and (Z
3
, +
3
) is
Z
2
×Z
3
=
(0, 0), (0, 1), (0, 2), (1, 0), (1, 1), (1, 2)
This is abelian of order 6, so we might guess that it is isomorphic to (Z
6
, +
6
) and thus cyclic.
This is indeed the case: simply observe that (1, 1) is a generator,
(1, 1)
=
(0, 0), (1, 1), (0, 2), (1, 0), (0, 1), (1, 2)
= Z
2
×Z
3
In accordance with Theorem 3.7, µ(x) = (x, x) defines an isomorphism µ : Z
6
=
Z
2
×Z
3
.
32
2. If each G
k
is abelian and written additively, the direct product is sometimes called a direct sum,
17
n
M
k=1
G
k
= G
1
··· G
n
We won’t use this notation, though you’ve likely encountered it in linear algebra: for instance,
the direct sum of n copies of the real line R is the familiar vector space
R
n
=
n
M
i=1
R = R ···R
Orders of Elements in a Direct Product
In Example 3.20.1, we saw that the element (1, 1) Z
2
× Z
3
had order 6 and thus generated the
group. There is another general pattern here; to help spot it, consider another example.
Example 3.21. We find the order of the element (10, 2) Z
12
×Z
8
? Recall Corollary 3.16:
10 Z
12
has order 6 =
12
gcd(10,12)
2 Z
8
has order 4 =
8
gcd(2,8)
If we repeatedly add (10, 2) to itself, then the first co-ordinate resets after 6 summations, whereas
the second resets after 4. For both to reset simultaneously, we need a common multiple of 6 and 4
summands. We can check this explicitly:
(10, 2)
=
(10, 2), (8, 4), (6, 6), (4, 0), (2, 2), (0, 4), (10, 6), (8, 0), (6, 2), (4, 4), (2, 6), (0, 0)
The order of the element (10, 2) is indeed the least common multiple 12 = lcm(6, 4).
Theorem 3.22. Suppose x
k
G
k
has order r
k
. Then (x
1
, . . . , x
n
)
n
k=1
G
k
has order lcm(r
1
, . . . , r
n
).
Proof. We appeal to Corollary 3.9:
(x
1
, . . . , x
n
)
m
= (x
m
1
, . . . , x
m
n
) = (e
1
, e
2
, . . . , e
n
) k, x
m
k
= e
k
k, r
k
| m
The order is the minimal positive integer m satisfying this, namely m = lcm(r
1
, . . . , r
n
).
Example 3.23. Find the order of (1, 3, 2, 6) Z
4
×Z
7
×Z
5
×Z
20
.
Again with reference to Corollary 3.16, the element has order
lcm
4
gcd(1,4)
,
7
gcd(3,7)
,
5
gcd(2,5)
,
20
gcd(6,20)
= lcm(4, 7, 5, 10) = 140
17
In these notes a direct product/sum will only ever have finitely many factors, in which case the concepts are identical.
The slight difference in the concepts when there are infinitely many factors is not worth discussing here.
33
When is a direct product of finite cyclic groups cyclic?
Recall that Z
2
×Z
3
=
Z
6
is cyclic, whereas Z
2
×Z
2
=
V is non-cyclic. It is reasonable to hypothesize
that the issue is whether the orders of the components are relatively prime.
Corollary 3.24. Z
m
×Z
n
is cyclic gcd(m, n) = 1. In such a case Z
m
×Z
n
=
Z
mn
.
More generally:
Z
m
1
×···× Z
m
k
=
Z
m
1
···m
k
i = j, gcd(m
i
, m
j
) = 1.
If n = p
r
1
1
··· p
r
k
k
is written in its prime factorization, then Z
n
=
Z
p
r
1
1
×···× Z
p
r
k
k
Proof. We prove the first part; the generalization follows by induction.
() Suppose gcd(m, n) = 1. We claim that (1, 1) is a generator of Z
m
× Z
n
. But this element has
order lcm(m, n) =
mn
gcd(m,n)
= mn =
|
Z
m
×Z
n
|
, so we’re done.
() This is Exercise 11.
Examples 3.25. 1. (Example 3.23) The group Z
4
×Z
7
×Z
5
×Z
20
is non-cyclic since, for instance,
gcd( 4, 20) = 1. The maximum order of an element in this group is
lcm( 4, 7, 5, 20) = 140 < 2800 =
|
Z
4
×Z
7
×Z
5
×Z
20
|
2. Is Z
5
×Z
7
×Z
12
cyclic? The Corollary says yes, since no pair of 5, 7, 12 have common factors.
It is ghastly to write, but there are 12 different ways (up to reordering) of expressing this group
as a direct product!
Z
420
=
Z
3
×Z
140
=
Z
4
×Z
105
=
Z
5
×Z
84
=
Z
7
×Z
60
=
Z
3
×Z
4
×Z
35
=
Z
3
×Z
5
×Z
28
=
Z
3
×Z
7
×Z
20
=
Z
4
×Z
5
×Z
21
=
Z
4
×Z
7
×Z
15
=
Z
5
×Z
7
×Z
12
=
Z
3
×Z
4
×Z
5
×Z
7
We may combine/permute the factors of 420 = 2
2
·3 ·5 ·7, provided we don’t separate 2
2
= 4.
The Fundamental Theorem of Finite(ly Generated) Abelian Groups
Direct products create finite abelian groups from cyclic building blocks. Our final result provides a
powerful converse, though we don’t have the technology to prove it (yet—though see Section 7.3).
Theorem 3.26. Every finite abelian group is isomorphic to a group of the form
Z
p
r
1
1
×···× Z
p
r
k
k
where each r
j
N and the p
i
are primes (not necessarily distinct). More generally, every finitely
generated abelian group (see Exercise 3.2.11) is isomorphic to some
Z
p
r
1
1
×···× Z
p
r
k
k
×Z ×···× Z
34
Examples 3.27. 1. Up to isomorphism, there are five distinct abelian groups of order 81 = 3
4
:
Z
81
, Z
3
×Z
27
, Z
9
×Z
9
, Z
3
×Z
3
×Z
9
, Z
3
×Z
3
×Z
3
×Z
3
Such groups can often be distinguished by considering the orders of elements. For instance:
G is abelian of order 81 and,
G has an element of order 27 and,
All elements of G have order 27
= G
=
Z
3
×Z
27
2. Since 450 = 2 ·3
2
·5
2
is a prime factorization, the fundamental theorem says that every abelian
group of order 450 is isomorphic to one of four groups:
(a) Z
2
×Z
3
2
×Z
5
2
=
Z
450
(cyclic, maximum order of an element 450)
(b) Z
2
×Z
3
×Z
3
×Z
5
2
(non-cyclic, maximum order 150 = 2 ·3 ·5
2
)
(c) Z
2
×Z
3
2
×Z
5
×Z
5
(non-cyclic, maximum order 90 = 2 ·3
2
·5)
(d) Z
2
×Z
3
×Z
3
×Z
5
×Z
5
(non-cyclic, maximum order 30 = 2 ·3 ·5)
As previously, there are multiple isomorphic ways to express each group as a direct product.
We finish by listing, up to isomorphism, all groups of order 15 and all abelian groups of order 16.
order abelian non-abelian
1 Z
1
2 Z
2
3 Z
3
4 Z
4
, V
=
Z
2
×Z
2
5 Z
5
6 Z
6
=
Z
2
×Z
3
D
3
=
S
3
7 Z
7
8 Z
8
, Z
2
×Z
4
, Z
2
×Z
2
×Z
2
D
4
, Q
8
9 Z
9
, Z
3
×Z
3
10 Z
10
=
Z
2
×Z
5
D
5
11 Z
11
12 Z
12
=
Z
3
×Z
4
, Z
2
×Z
6
=
Z
2
×Z
2
×Z
3
D
6
, A
4
, Q
12
13 Z
13
14 Z
14
=
Z
2
×Z
7
D
7
15 Z
15
=
Z
3
×Z
5
16 Z
16
, Z
4
×Z
4
, Z
2
×Z
8
, Z
2
×Z
2
×Z
4
, Z
2
×Z
2
×Z
2
×Z
2
(nine)
The Fundamental Theorem & Corollary 3.24 supply the abelian groups. In the non-abelian column:
The dihedral groups D
n
are the familiar symmetries of a regular n-gon (Definition 2.24).
S
3
and A
4
will be described in Chapter 4 (symmetric and alternating groups).
Q
8
is the quaternion group (Exercise 2.2.8). The generalized quaternion group Q
12
is related.
There are nine non-isomorphic non-abelian groups of order 16: D
8
and the direct product Z
2
× Q
8
are explicit examples. You might suspect from the table that all non-abelian groups have even order:
this is not so, though the smallest counter-example has order 21.
35
Exercises 3.3. Key concepts: Direct product Order of an element via lcm Cyclic/gcd criteria
Fundamental theorem of finitely generated abelian groups
1. List the elements of the following direct product groups:
(a) Z
2
×Z
4
(b) Z
3
×Z
3
(c) Z
2
×Z
2
×Z
2
2. Prove Theorem 3.19 by checking each of the axioms of a group.
3. Prove that G × H
=
H ×G.
4. Prove that a direct product
G
k
is abelian if and only if its components G
k
are all abelian.
5. Find the orders of the following elements and write down the cyclic subgroups generated by
each (list all of the elements explicitly):
(a) ( 1, 3) Z
2
×Z
4
(b) ( 4, 2, 1) Z
6
×Z
4
×Z
3
6. Is the group Z
12
×Z
27
×Z
125
cyclic? Explain.
7. Find a generator of the group Z
3
×Z
4
and hence define an isomorphism µ : Z
12
=
Z
3
×Z
4
.
(Hint: read the proof of Corollary 3.24)
8. State three non-isomorphic groups of order 50.
9. Let p, q be distinct primes. Up to isomorphism, how many abelian groups have order p
2
q
2
?
10. Give a simple explanation for why D
8
is not isomorphic to Z
2
× Q
8
.
11. Complete the proof of Corollary 3.24: if Z
m
×Z
n
is cyclic, then gcd(m, n) = 1.
(Hint: if gcd(m, n) 2, what is the maximum order of an element in Z
m
×Z
n
?)
12. Suppose G is an abelian group of order m, where m is a square-free positive integer (k Z
2
such that k
2
|m). Prove that G is cyclic.
13. (a) Let G be a finitely generated abelian group and let H be the subset of G consisting of the
identity e together with all the elements of order 2 in G. Prove that H is a subgroup of G.
(b) In the language of the Fundamental Theorem, to which direct product is H isomorphic?
14. Suppose G is finite abelian and that m divides
|
G
|
. Prove that G has a subgroup of order m.
(Hint: use the prime decomposition of m and the Fundamental Theorem)
15. Suppose G is an abelian group and let T G be the subset of elements with finite order.
(a) Prove that T is a subgroup of G.
(Your proof shouldn’t use the Fundamental Theorem—why not?)
(b) Compute T when:
i. G = (R
×
, ·) ii. G = (S
1
, ·)
36
4 Permutations and Orbits
In this chapter we return to the origins of group theory by considering the ways in which a set may
be reordered.
4.1 The Symmetric Group, Cycle Notation & Cayley’s Theorem
Definition 4.1. Let A be a set. A permutation of A is a bijective/invertible function σ : A A.
The symmetric group S
A
is the set of all permutations of A under functional composition.
The symmetric group on n-letters
18
S
n
is the group S
A
when A = {1, 2, . . . , n}.
Examples 4.2. 1. If A = {1}, there is only one (bijective) function A A, namely the identity
function e : 1 7→ 1. Thus S
1
has only one element and is isomorphic to Z
1
.
2. If A = {1, 2}, then there are two bijections e, µ : A A:
e(1) = 1 and e(2) = 2 defines the identity function.
σ(1) = 2 and σ(2) = 1 swaps the elements of A.
e σ
e e σ
σ σ e
The Cayley table is immediate: plainly S
2
is isomorphic to Z
2
.
3. S
3
= S
{1,2,3}
has six elements and is non-abelian. We have met this group before: note how every
rotation/reflection of the triangle in Example 2.25 corresponds to a permutation of the three
vertices 1, 2, 3. Otherwise said, S
3
is isomorphic to the dihedral group D
3
.
Lemma 4.3. 1. S
A
is a group under composition of functions.
2. If A has at least three elements, then S
A
is non-abelian.
3. The order of S
n
is n! (Warning! The order of S
n
is not the subscript n)
4. S
m
S
n
whenever m n (Strictly, S
n
contains a subgroup isomorphic to S
m
)
Proof. 1. The axioms follow from familiar function properties. Write out the details if you’re unsure.
Closure: If σ, τ : A A are bijective, so is the composition σ τ.
Associativity: Composition of functions is associative (Exercise 2.4.9).
Identity: The identity function e
A
: a 7 a for all a A is certainly bijective.
Inverse: If σ is a bijection, then its inverse function σ
1
is also bijective.
The remaining parts are exercises.
From now on we use juxtaposition: στ := σ τ. Exponentiation therefore means self-composition:
e.g. σ
3
= σσσ = σ σ σ. Remember also that στ is a function A A, and that evaluation means
that we act with τ first:
a A, στ(a) = σ
τ(a)
18
This choice of A makes S
n
explicit. In practice, any set with n elements will do and any group isomorphic to this is
usually also called S
n
(see Exercise 8 and the remark regarding “up to isomorphism” on Page 20).
37
Cycle Notation
Efficient computation in S
n
is facilitated by some new notation.
Definition 4.4. Suppose {a
1
, . . . , a
k
} {1, . . . , n}. The k-cycle σ = (a
1
a
2
···a
k
) S
n
is the function
σ :
a
j
7 a
j+1
if j < k
a
k
7 a
1
x 7 x if x {a
1
, . . . , a
k
}
a
1
a
2
a
3
···
a
k
all other x
Cycles (a
1
···a
k
) and (b
1
···b
l
) are disjoint if {a
1
, . . . , a
k
}{b
1
, . . . , b
l
} = .
1-cycles and the 0-cycle () can be helpful in calculations, though both are simply the identity e.
Example 4.5. A 4-cycle σ = (1 3 4 2) and a 2-cycle τ = (1 4) in S
4
are defined in the table:
x 1 2 3 4
σ(x) 3 1 4 2
τ(x) 4 2 3 1
σ : 1
//
3
//
4
//
2
ww
τ : 1
>>
4
||
2
zz
3
zz
To compose cycles, remember that both are functions and you won’t go wrong!
x 1 2 3 4
τ(x) 4 2 3 1
στ(x) 2 1 4 3
στ : 1
~
??
2
||
3
??
4
||
The result is a product of disjoint 2-cycles στ = (1 2) (3 4).
Algorithmic Cycle Composition Computation using tables is impractically slow. Here is an al-
gorithmic approach that, with practice, should prove more efficient. We illustrate by verifying the
previous calculation: at each stage we write only a single number or bracket, building up the right
column below.
Open a bracket and write 1. στ = (1
Since 1
τ
7 4
σ
7 2, we next write 2. = (1 2
2
τ
7 2
σ
7 1 restarts the cycle; close it and start another with an unused value. = (1 2)(3
3
τ
7 3
σ
7 4, so next write 4. = (1 2)(3 4
4
τ
7 1
σ
7 3 restarts the current cycle, so close it. = (1 2)(3 4)
All values 1, 2, 3, 4 have appeared so the algorithm terminates.
It should be clear how to extend the algorithm when composing more cycles. Any 1-cycles obtain
should be deleted. Shortly we’ll prove that the algorithm always terminates in a product of disjoint
cycles (Theorem 4.13). For now, practice it by verifying the following:
Examples 4.6. 1. ( 1 4)(1 3 4 2) = (1 3)(2 4) 2. (1 3 5 4)(2 3 4) = (1 3)(2 5 4)
3. ( 1 2 3 4)(1 2 3)(1 2) = (1 4)(2 3) 4. ( 1 2 3 4 5 6)
3
= (1 4)(2 5)(3 6)
38
Geometric Symmetry Groups (Section 2.4 revisited)
We’ve already seen (Example 4.2.3) how the symmetry group D
3
of an equilateral triangle is iso-
morphic to the symmetric group S
3
. The same trick applies more generally: label the vertices (or
edges/faces) of a figure with numbers 1, 2, 3, . . . and represent each rotation/reflection by how it
permutes these values. Cycle notation makes calculating compositions easy!
Examples 4.7. 1. Label the edges of a rectangle to view the Klein four-group V as a subgroup of S
4
:
the 2-cycles ( 1 3) and (2 4) are reflections, and their composition is rotation by 180°.
1
2
3
4
V
=
e, (1 3), (2 4), (1 3) (2 4)
Alternative descriptions of V can be obtained by using different labellings (Exercise 3).
2. Label the vertices of a regular hexagon 1 through 6.
The 2,2-cycle (1 5)( 2 4) represents reflection across the line through
3 and 6.
The 6-cycle ( 1 2 3 4 5 6) represents 60° counter-clockwise rotation.
Both functions describe elements of the dihedral group D
6
. By extension, D
6
may be identified
as (is isomorphic to) a subgroup of S
6
.
3. By labelling the vertices of a square, we may identify D
4
with a subgroup
of S
4
. All elements and the complete subgroup diagram are given be-
low. By convention we denote reflections across diagonals (δ
j
) and the
midpoints of sides (µ
j
) differently.
Note the ease of computations: e.g.,
(2 4)(1 2)(3 4) = (1 4 3 2) = δ
1
µ
1
= ρ
3
1
2
3
4
Element Cycle notation
ρ
0
e = ()
Rotations
ρ
1
(1234)
ρ
2
(13)(24)
ρ
3
(1432)
µ
1
(12)(34)
Reflections
µ
2
(14)(23)
δ
1
(24)
δ
2
(13)
Subgroup Isomorph
{ρ
0
} Z
1
{ρ
0
, µ
i
} Z
2
{ρ
0
, δ
i
} Z
2
{ρ
0
, ρ
2
} Z
2
{ρ
0
, ρ
1
, ρ
2
, ρ
3
} Z
4
{ρ
0
, µ
1
, µ
2
, ρ
2
} V
{ρ
0
, δ
1
, δ
2
, ρ
2
} V
D
4
V Z
4
V
Z
2
Z
2
Z
2
Z
2
Z
2
Z
1
You should be able to recognize these subgroups geometrically; e.g. the blue copy of V is pre-
cisely that in the first example! Try to convince yourself why there are no other subgroups.
The same thing can be done for 3D figures like the tetrahedron (Example 4.26.3).
39
Cayley’s Theorem
The word group, at least in mathematics, originally referred to a set of permutations. We finish this
section with a foundational result that links to the original meaning of the word: every element of a
group may be viewed as a permutation—indeed of the group itself!
Theorem 4.8 (Cayley). Every group is isomorphic to a group of permutations.
The proof of Cayley’s Theorem is merely the abstraction of a simple example.
Example 4.9. To each integer g, we may associate the function “add g.” For instance, 1 corresponds
to the function +1,” etc. The function “add g is a bijection of the integers: its inverse function is
“add g.” Each integer is therefore naturally associated to a permutation, an element of the group S
Z
.
Proof. Let G be a group. For each g G, define left-multiplication by g to be the function L
g
: G G
where,
x G, L
g
(x) = gx
This is bijective (inverse function (L
g
)
1
= L
g
1
), and therefore a permutation of G: that is L
g
S
G
.
Now define a function ϕ : G S
G
by ϕ(g) = L
g
. We prove that ϕ is an injective homomorphism.
Injectivity ϕ(g) = ϕ (h) = L
g
= L
h
= L
g
(e) = L
h
(e) = ge = he = g = h
Homomorphism We show that ϕ(gh) = ϕ(g)ϕ(h) as functions by evaluating on all x G:
ϕ(gh)
(x) = L
gh
(x) = (gh)x = g( hx) = L
g
L
h
(x)
=
ϕ(g)ϕ(h)
(x)
By Exercise 2.5.12, ϕ is an isomorphism onto its image/range. Otherwise said, G is isomorphic to the
subgroup ϕ(G) of the symmetric group S
G
.
Be careful! Cayley’s Theorem does not say that every group is isomorphic to some symmetric group.
It says that that every group G is isomorphic to some subgroup of S
G
.
Exercises 4.1. Key concepts: Permutation Symmetric group Cycle notation
1. Which of the following functions are permutations? Explain.
(a) f : Z Z such that f (x) = x 7.
(b) f : Z Z such that f (x) = 3x + 4.
(c) f : R R such that f (x) = x
3
x.
(d) f : R R such that f (x) = x
3
+ x.
(e) f : {fish, horse, dog, cat} {fish, horse, dog, cat} where
f (fish) = horse, f (horse) = cat, f (dog) = dog, f ( cat) = fish
2. Compute the following products (compositions) of permutations in cycle notation.
(a) ( 1 2)(3 4) (1 2 3) S
4
(b) ( 1 4)(2 3)(3 4)(1 4) S
4
(c) ( 1 2 3)(2 3 4)(3 4 1)(4 1 2) S
4
(d) ( 1 2 4 5)
2
(2 4 5)
2
S
5
40
3. (Example 4.7.1, cont.) Use different labellings of the edges of the rectangle to find another two
subgroups of S
4
isomorphic to V. What happens if you instead label the rectangle’s vertices?
4. By labelling vertices, view the dihedral group D
7
of symmetries
of the regular heptagon as a subgroup of S
7
. Each µ
i
is reflec-
tion across the indicated dashed line, while ρ
j
is rotation j steps
counter-clockwise.
(a) State µ
4
in cycle notation.
(b) Compute µ
3
ρ
1
using cycle notation. What element of D
7
does
this represent?
(c) Calculate (ρ
2
µ
3
ρ
1
)
666
.
µ
1
µ
2
µ
3
µ
4
µ
5
µ
6
µ
7
1
2
3
4
5
6
7
ρ
1
= (1 2 3 4 5 6 7), ρ
j
= ρ
j
1
5. State the elements of the rotation group R
5
in cycle notation when viewed as a subgroup of S
5
.
6. Prove parts 2, 3, and 4 of Lemma 4.3.
7. How many distinct subgroups of S
4
are isomorphic to S
3
. Describe them.
8. Suppose sets A and B have the same cardinality: that is, µ : A B bijective.
(a) If σ S
A
is a permutation, show that µσµ
1
S
B
.
(b) Hence prove that S
A
and S
B
are isomorphic.
9. Cayley’s Theorem says that G is isomorphic to a subgroup of S
G
. What can you say about a
(finite) group G if G
=
S
G
?
10. In Cayley’s Theorem we defined ϕ(g) = L
g
: G G via left-multiplication.
(a) Does the argument still work if ϕ(g) = R
g
: G G is right-multiplication R
g
(x) = xg?
(b) (Harder) Suppose we take ϕ (g) = C
g
: G G to be the function C
g
(z) = gxg
1
. Does
the proof of Cayley’s Theorem work this time? Why/why not?
11. Show that the group S
3
is indecomposable: there are no groups G, H of order less than
|
S
3
|
= 6
for which S
3
=
G × H.
(Hint: If S
3
is decomposable, there is only one possibility; why does this decomposition make no sense?)
12. Let n 3. Prove that if σ S
n
commutes with every other element of S
n
(i.e. σρ = ρσ, ρ S
n
)
then σ is the identity.
(Hint: suppose σ(a) = b = a and consider the cases σ(b) = a and σ(b) = a separately)
13. (a) Let µ D
n
be a reflection and ρ the one-step counter-clockwise rotation. Prove that
µρ = ρ
n1
µ and more generally that µρ
k
= ρ
nk
µ.
(b) Define a group G abstractly as the elements generated by objects ρ, µ with the properties:
ρ has order n, µ
ρ
has order 2, and µρ = ρ
n1
µ
Prove that G = {e, . . . , ρ
n1
, µ, . . . , ρ
n1
µ} has order 2n, and that each ρ
k
µ has order 2.
(In many textbooks this is the definition of D
n
.)
41
4.2 Orbits
Group theory is often applied by allowing the elements of a group to transform a set.
19
We’ve already
seen several examples: for instance, how rotations transform a geometric figure. The simplest general
example is built into the definition of the symmetric group and appears naturally in cycle notation.
Definition 4.10. The orbit of σ S
n
containing x {1, 2, . . . , n} is the set
orb
x
(σ) = {σ
k
(x) : k Z} {1, 2, . . . , n}
Warning! Each orbit is a subset of {1, 2, . . . , n}, not of the group S
n
.
Observe also that orb
σ
k
(x)
(σ) = orb
x
(σ) for any k Z.
Examples 4.11. If σ S
n
is written as a product of disjoint cycles, then the cycles are the orbits.
1. The orbits of ( 1 3 4) S
4
are the disjoint sets {1, 3, 4}, {2}. Note the singleton orbit {2}!
2. The orbits of ( 1 2)(4 5) S
5
are {1, 2}, {3}, {4, 5}.
3. Non-disjoint cycles are not orbits. For instance, σ = (1 3)(2 3 4) S
4
maps
1 7 3 7 4 7 2 7→ 1
so there is only one orbit: orb
x
(σ) = {1, 2, 3, 4}for any x. This comports with the result obtained
by multiplying cycles via the usual algorithm: σ = (1 2 3 4).
Given that disjoint cycle notation is so useful for reading orbits, it is natural to ask if any permutation
can be written in such a manner. The answer is yes, as we demonstrate in the next two results.
Lemma 4.12. The orbits of σ S
n
partition X = {1, 2, . . . , n}.
Proof. Define a relation on X = {1, 2, . . . , n} by x y y orb
x
(σ). We claim that this is an
equivalence relation.
20
Reflexivity x x since x = σ
0
(x).
Symmetry x y = y = σ
k
(x) for some k Z. But then x = σ
k
(y) = y x.
Transitivity Suppose that x y and y z. Then y = σ
k
(x) and z = σ
l
(y) for some
k, l Z. But then z = σ
k+l
(x) and so x z.
The equivalence classes of are clearly the orbits of σ, which therefore partition X.
19
The formal definition of such group actions is postponed until Chapter 7.
20
The relationship between equivalence relations and partitions underpins several upcoming ideas, and should be famil-
iar from a previous course. Here is a very quick review:
Given x X and a relation on X, define the set [x] := {y X : y x}.
Theorem: The sets [x] partition X (every y X lies in precisely one such subset [x]) if and only if is an equivalence
relation (reflexive, symmetric, transitive). In such a case we call [x] an equivalence class.
In our situation, [x] = orb
x
(σ); the Lemma simply proves that the orbits of σ are equivalence classes.
42
Theorem 4.13. Every permutation can be written as a product of disjoint cycles.
Proof. We formalize the algorithm from page 38. Suppose σ S
n
is given.
1. List the elements of orb
1
(σ) in the order they appear within the orbit:
orb
1
(σ) = {1, σ(1), σ
2
(1), . . .}
If this all of X = {1, . . . , n}, we are finished: σ = (1 σ(1) σ
2
(1) . . . σ
n1
(1)
is an n-cycle.
2. Otherwise, let x
2
= min{x X : x orb
1
(σ)} and construct its orbit:
orb
x
2
(σ) = {x
2
, σ(x
2
), σ
2
(x
2
), . . .}
By Lemma 4.12, orb
x
2
(σ) is disjoint with orb
1
(σ). If orb
1
(σ) orb
x
2
(σ) = X then we are done,
and σ is the product of two disjoint cycles:
σ =
1 σ(1) σ
2
(1) ···
x
2
σ(x
2
) σ
2
(x
2
) ···
3. Otherwise, repeat. At stage k, let x
k
= min{x X : x orb
1
(σ) ··· orb
k1
(σ)}. By
the Lemma, orb
x
k
(σ) is disjoint with orb
1
(σ) ··· orb
k1
(σ). The process continues until
orb
1
(σ) ···orb
k
(σ) = X, which must happen eventually since X is a finite set. The result is
a product of disjoint cycles:
σ =
1 σ(1) σ
2
(1) ···
| {z }
orb
1
(σ)
x
2
σ(x
2
) σ
2
(x
2
) ···
| {z }
orb
x
2
(σ)
··· ···
| {z }
orb
x
3
(σ)
···
··· ···
| {z }
orb
x
k
(σ)
The Theorem explains why our cycle algorithm always produces a product of disjoint cycles! By
convention, 1 < x
2
< ··· < x
k
, though there is no need to do this: disjoint cycles can be listed in any
order and may start with any element, e.g.,
(1 3)(2 5 4) = (5 4 2)(3 1)
Also, by convention, we delete any orbits of size 1 (1-cycles).
Orders of Elements in S
n
Recall (Corollary 3.9) that the order of an element σ S
n
is the smallest positive integer k for which
σ
k
= e.
Example 4.14. If σ = (1 2 3 4 5 6) S
6
, then
σ
2
= (1 3 5)(2 4 6) σ
3
= (1 4)(2 5)(3 6) σ
4
= (1 5 3)(2 6 4)
σ
5
= (1 6 5 4 3 2) σ
6
= e
whence the order of σ is 6. This is intuitive if we identify σ with a counter-
clockwise rotation of a regular hexagon: indeed
σ
= {e, σ, . . . , σ
5
}
=
R
6
.
1
2
3
4
5 6
By similarly considering a regular k-gon, it should be clear that every k-cycle has order k.
43
Things are trickier when you don’t have a single cycle. However, we are again saved by the discus-
sion of disjoint cycles, since disjoint cycles commute.
Examples 4.15. 1. Since (1 2 3) and (4 5) are disjoint cycles, we know that (1 2 3)(4 5) = (4 5)(1 2 3).
We therefore easily compute
(1 2 3)(4 5)
3
= (1 2 3)(4 5)(1 2 3)(4 5)(1 2 3)(4 5)
= (1 2 3)
3
(4 5)
3
= e(4 5) = (4 5)
2. Given σ = (2 5 3)(1 5 4 3) S
5
, we find σ
8
. It is tempting to write
σ
8
?
= (2 5 3)
8
(1 5 4 3)
8
=
(2 5 3)
3
2
(2 5 3)
2
(1 5 4 3)
4
2
= e
3
(2 3 5)e
2
= (2 3 5)
but this would be incorrect: the cycles don’t commute (2 5 3)(1 5 4 3) = (1 5 4 3)(2 5 3) so we can-
not distribute the exponent (8). If instead we first find σ as a product of disjoint cycles, then the
exponent does distribute and the computation is easy
σ = (1 3)(2 5 4) = σ
8
= (1 3)
8
(2 5 4)
8
= (2 5 4)
2
= (2 4 5)
This approach also tells us the order of σ. Observe that
e = σ
k
= (1 3)
k
(2 5 4)
k
k is divisible by both 2 and 3
The order if σ is therefore 6.
Corollary 4.16. The order of a permutation σ is the least common multiple of the lengths of its
disjoint cycles.
Proof. Write σ = σ
1
···σ
m
as a product of disjoint cycles, where the cycle σ
j
has length α
j
. Since
disjoint cycles commute,
σ
k
= σ
k
1
···σ
k
m
Since each factor σ
k
j
permutes disjoint sets, it follows that
σ
k
= e j, σ
k
j
= e j, α
j
| k (the order of an α
j
-cycle is α
j
)
The order of σ is the smallest k satisfying this condition, namely lcm(α
1
, . . . , α
m
).
Example 4.17. Since σ = (1 4 5)(3 6 2 7)(8 9) S
9
is written as a product of disjoint cycles, its order
is lcm( 3, 4, 2) = 12.
To compute σ
3465
, first observe that 3465 = 12 ·288 + 9 (division algorithm). From this,
σ
3465
= (σ
12
)
288
σ
9
= σ
9
= (1 4 5)
9
(3 6 2 7)
9
(8 9)
9
= (3 6 2 7)(8 9)
since ( 1 4 5), (3 6 2 7) and (8 9) have orders 3, 4 and 2 respectively.
44
Exercises 4.2. Key concepts: Orbit Partition Disjoint cycles Order of element via lcm
1. Find the orbits of the following permutations, and their orders:
(a) ρ = (1 4 5)(2 3 4 5) S
5
.
(b) σ = (1 5 4)(2 5 4)(1 2 3 4) S
5
.
(c) τ = (1 5 7 4)( 3 2 4)(3 2 5 6) S
7
.
2. If σ S
A
is any permutation, we may define its orbits similarly: orb
a
(σ) = {σ
j
(a) : j Z}.
What are the orbits of the permutation σ : Z Z : n 7→ n + 3?
3. Given σ = (1 3)(2 4 5) S
5
, find the elements of the cyclic group
σ
S
5
generated by σ.
4. What is the largest possible order of an element of the group S
3
×Z
4
×V? Exhibit one.
5. What is the maximum order of an element in each of the groups S
4
, S
5
, S
6
, S
7
, S
8
? Exhibit a
maximum order element in each case.
6. For which integers n does there exist a subgroup C
n
S
8
where C
n
is cyclic of order n? Explain
your answer.
7. Let σ S
n
. For each k > 0, prove that each orbit of σ
k
is a subset of an orbit of σ.
8. Consider the permutations σ = (1 3 5)(2 7 4 9 6) and τ = (1 5 3 2)(6 9) in S
9
.
(a) Compute στ and τσ in cycle notation.
(b) Find the orders of σ, τ, στ and τσ.
(c) Compute (στ)
432
σ
43
as a product of disjoint cycles.
(d) Construct the subgroup diagram of
σ
and give a generator for each subgroup.
45
4.3 Transpositions & the Alternating Group
In the previous sections we viewed a permutation in terms of its orbits. An alternative approach
involves the construction of permutation from the very simplest bijections.
Definition 4.18. A 2-cycle (a
1
a
2
) is also known as a transposition, since it swaps two elements of
{1, 2, . . . , n} and leaves the rest untouched.
Theorem 4.19. Every σ S
n
(n 2) may be written as a product of transpositions.
Proof. There are many, many ways to do this. One approach is first to write σ as a product of disjoint
cycles, before decomposing each cycle as follows:
(a
1
··· a
k
) = (a
1
a
k
)(a
1
a
k1
) ···(a
1
a
2
)
Just read carefully and you should be convinced this works!
Example 4.20. The method in the proof results in the decomposition
(1 7 6 4 5) = (1 5)(1 4)(1 6)(1 7)
Other decompositions are possible, for instance (1 7)(3 6)(5 7)(4 7)(3 6)( 6 7).
While representations as a product of transpositions are non-unique, a simple commonality may be
observed via matrix notation. Consider, for instance,
1 0 0 0
0 0 0 1
0 0 1 0
0 1 0 0
a
b
c
d
=
a
d
c
b
and
0 0 1 0
0 0 0 1
0 1 0 0
1 0 0 0
a
b
c
d
=
c
d
b
a
Each of these 4 × 4 matrices permutes the rows of any 4 × k matrix following it. These matrices
correspond naturally to two elements of S
4
:
The transposition ( 2 4) (swap rows 2 and 4).
The 4-cycle ( 1 4 2 3) (map row 1 to row 4, row 4 to row 2, etc.)
Definition 4.21. A permutation matrix is square matrix which is zero except for a single 1 in each row
and column. Equivalently, it is obtained by permuting the rows of the identity matrix.
Lemma 4.22. The set of n ×n permutation matrices forms a multiplicative group isomorphic to S
n
.
We omit a formal proof, though the rough idea is hopefully clear: replace each σ S
n
with its corre-
sponding permutation matrix and observe that composition corresponds to matrix multiplication.
What does this have to do with transpositions? Since a transposition swaps two elements, it corre-
sponds to multiplication by a row-swapping elementary matrix; any such matrix has determinant 1.
Suppose that a permutation is written as a product of transpositions:
σ = σ
1
···σ
m
46
and view this as a product of matrices, taking determinant of both sides shows that
det σ = (1)
m
The value of the determinant plainly depends only on whether m is even or odd. . .
Definition 4.23. A permutation σ S
n
is even/odd if it can be written as the product of an even/odd
number of transpositions.
By our matrix/determinant discussion, the parity of a permutation is well-defined: every permuta-
tion is either even or odd; it cannot be both!
Plainly the composition of even permutations remains even, as does the inverse of such. We may
therefore define a new subgroup of S
n
.
Definition 4.24. The alternating group A
n
(n 2) is the group of even permutations in S
n
.
Theorem 4.25. A
n
contains exactly half the elements of S
n
: otherwise said,
|
A
n
|
=
n!
2
.
Proof. Since n 2, we have (1 2) S
n
. Define ϕ : S
n
S
n
by ϕ(σ ) = (1 2)σ. Since
(1 2)(1 2)σ = σ
we see that ϕ is invertible (ϕ is its own inverse!). Moreover, ϕ maps even permutations to odd and
vice versa. It follows that there are exactly the same number of odd and even permutations.
Examples 4.26. We describe the smallest three alternating groups A
4
.
1. A
2
= {e}
=
Z
1
is trivial.
2. A
3
= {e, (1 3)(1 2) , (1 2)(1 3)} = {e, (1 2 3), (1 3 2)}
=
Z
3
is a cyclic group.
3. When n = 4 we obtain the first ‘new’ group in the alternating family; a group of order 12.
A
4
= {e, (1 2 3), (1 3 2), (1 2 4), (1 4 2), (1 3 4), (1 4 3), (2 3 4), (2 4 3),
(1 2)(3 4), (1 3)(2 4), (1 4)(2 3)}
A
4
is non-abelian: for example,
(1 2 3)(1 2 4) = (1 3)(2 4) = (1 4)(2 3) = ( 1 2 4)(1 2 3)
While we’ve already countered one non-abelian group of order
12, the dihedral group D
6
, we quickly see that A
4
D
6
: all ele-
ments of A
4
have orders 1, 2 or 3, while D
6
contains a rotation (by
60°) of order 6.
By labelling faces (or vertices), A
4
may be visualized the 3D-
rotation group of a regular tetrahedron: can you see how each
element acts?
47
Exercises 4.3. Key concepts: Transposition (representation of elements) Odd/even permutations A
n
1. Write (1 3 4 6)(2 4 6) as a product of transpositions in two different ways.
2. State σ = (1 3) and τ = (1 3 2) as 3 ×3 permutation matrices S and T. Compute the matrix
product ST and verify that it is the permutation matrix corresponding to στ S
3
.
3. Give examples of two non-isomorphic non-abelian groups of order 360.
4. Explain why every finite group is isomorphic to a group of matrices under multiplication.
5. S
4
has four distinct subgroups isomorphic to the Klein four-group V; state them. Only one of
these is a subgroup of A
4
; which?
6. We just saw that the rotation group of a regular tetrahedron is isomorphic to A
4
.
(a) What is the order of the rotation group of a cube?
(Hint: each face may be rotated to any of six faces, and then rotated in place. . . )
(b) Repeat the calculation for the remaining three platonic solids (octahedron, dodecahedron,
icosahedron).
(c) By placing a vertex at the center of each face of a cube, argue that the rotation group of an
octahedron is also isomorphic to S
4
.
What happens when you do this for a dodecahedron? A tetrahedron?
(d) Label the four diagonals of a cube 1, 2, 3, 4. Describe geometrically the effect of the per-
mutation (2 3 4) on the cube. What about (2 3)? Hence conclude that the rotation group of
a cube is isomorphic to S
4
.
(The dodecahedral and icosahedral rotation groups are both isomorphic to the alternating group A
5
,
though this is harder to visualize than the cube situation—try researching a proof. . . )
7. (Hard) Find the entire subgroup diagram of A
4
.
8. (Hard) Prove that D
n
is a subgroup of A
n
n 1 (mod 4)
(Do this in one shot if you like; otherwise use the following steps to guide your thinking)
(a) Label the corners of a regular n-gon 1 through n counter-clockwise so that every element
of D
n
may be written as a permutation of {1, 2, . . . , n}. Write in a sentence what you are
required to prove: what condition characterizes being in the group A
n
?
(b) Consider the rotation ρ
1
= (1 2 3 ··· n) of the n-gon one step counter-clockwise. Is ρ
1
odd
or even, and how does this depend on n?
(c) Show that every rotation ρ
i
D
n
is generated by ρ
1
. When is the set of rotations in D
n
a
subgroup of A
n
?
(d) A reflection µ D
n
permutes corners of the n-gon by swapping pairs. How many pairs
of corners does µ swap when n 1 (mod 4)? Is µ an odd or even permutation? You may
use a picture, provided it is sufficiently general.
(e) Summarize parts (a–d) to argue the direction of the theorem.
(f) Prove the direction of the theorem by exhibiting an element of D
n
which is not in A
n
whenever n ≡ 1 (mod 4).
48
5 Cosets & Factor Groups
In this chapter we partition a group into subsets in such a way that the set of subsets inherits a natural
group structure. This is a long and abstract story, though the essentials aren’t new; this is precisely
what happens with modular arithmetic (itself a generalization of even and odd).
Example 5.1. In Z
3
= {0, 1, 2} the elements are really subsets [0], [1], [2] of the integers Z: that is,
[0] = {x Z : x 0 (mod 3)} = {. . . , 3, 0, 3, 6, . . .}
[1] = {x Z : x 1 (mod 3)} = {. . . , 2, 1, 4, 7, . . .}
[2] = {x Z : x 2 (mod 3)} = {. . . , 1, 2, 5, 8, . . .}
When we write 1 +
3
2 = 0 Z
3
, what we really mean is
x [1], y [2] we have x + y [0]
The group operation (addition) on Z naturally induces the group operation (addition modulo 3) on
the set of subsets Z
3
= {[0], [1], [2]}.
5.1 Cosets & Normal Subgroups
Our main goal is to generalize Example 5.1. Start by observing that the identity element [0] Z
3
is
in fact a subgroup (3Z) of Z from which the other subsets [1], [2] are obtained by translation.
Definition 5.2. Let H be a subgroup of G and g G. The left coset of H containing g is
gH := {gh : h H} (x gH h H such that x = gh)
This is a subset of G. The right coset of H containing g is defined similarly:
Hg := {hg : h H}
The identity coset H = eH = He is the left & right coset of H containing the identity e.
H is a normal subgroup of G, written H G, if the left and right cosets containing g are always equal
H G g G, gH = Hg
Example (5.1 cont). Since G = Z is written additively, the left and right cosets of H = [0] = 3Z
containing g are written
g + H := {g + h : h H} H + g := {h + g : h H}
These cosets are precisely the elements of Z
3
!
3Z = 0 + 3Z = 3Z + 0 = [0] = {. . . , 3, 0, 3, 6, . . .}
1 + 3Z = 3Z + 1 = [1] = {. . . , 2, 1, 4, 7, . . .}
2 + 3Z = 3Z + 2 = [2] = {. . . , 1, 2, 5, 8, . . .}
Since the left and right cosets are equal, H = 3Z is a normal subgroup of Z.
49
The last observation is in fact general—the proof is an exercise.
Lemma 5.3. Every subgroup of an abelian group G is normal.
A subgroup of a non-abelian group might be normal, but is more likely not to be (see Example 5.4.2).
Examples 5.4. 1. The subgroup H =
2
= {0, 2, 4} Z
6
has two distinct cosets (left = right since
Z
6
is abelian):
H = {0, 2, 4}
= 2 + H = 4 + H
1 + H = {1, 3, 5}
= 3 + H = 5 + H
Observe how the cosets partition Z
6
into equal-sized subsets.
2. The left and right cosets of the subgroup H = {e, µ
1
} D
3
are as follows:
Left cosets Right cosets
H = µ
1
H = {e, µ
1
} H = Hµ
1
= {e, µ
1
}
ρ
1
H = µ
3
H = {ρ
1
, µ
3
} Hρ
1
= Hµ
2
= {ρ
1
, µ
2
}
ρ
2
H = µ
2
H = {ρ
2
, µ
2
} Hρ
2
= Hµ
3
= {ρ
1
, µ
3
}
To verify this, either revisit the multiplication table for D
3
(Example 2.25) or use cycle notation
(e.g. Example 4.7). This time the left and right cosets of H are not the same: H is not a normal
subgroup of D
3
. The partitioning observation still holds: the left cosets partition D
3
into three
equal-sized subsets; the right cosets also partition D
3
into equal-sized (but different) subsets.
3. Consider a 1-dimensional subspace W R
2
. Each coset
u + W = {u + w : w W}
is a line parallel to W. Once again, the cosets (all lines parallel
to W) partition R
2
.
If this feels too abstract, consider the special case where W is
the y-axis: each coset is a vertical line (same x-co-ordinate).
More generally, if W is a subspace of some vector space, then
the cosets u + W are the sets parallel to W. Only the zero coset
W = 0 + W is a subspace.
u
W
u + W
4. Consider the alternating group A
n
as a subgroup of S
n
. Generalizing the argument from Theo-
rem 4.25, we see that for any α A
n
and σ S
n
,
ασ even σ even σα even
Otherwise said, for any σ S
n
the cosets of A
n
containing σ are
σA
n
= A
n
σ =
(
A
n
if σ even
B
n
if σ odd
where B
n
is the set of odd permutations in S
n
. In particular, A
n
is a normal subgroup of S
n
.
50
As observed in the examples, the cosets of any subgroup H G partition G.
Theorem 5.5. Let H be a subgroup of G. Then the left cosets of H partition G. Moreover,
y xH x
1
y H xH = yH
The right cosets also partition G:
y Hx yx
1
H Hx = Hy
The blue criterion is often very easy to check. Before reading the proof, convince yourself that each
of our previous examples satisfies the result. When H is non-normal, the right cosets partition G
differently to the left cosets (e.g. Example 5.4.2).
Example 5.6. This should seem familiar when G = Z and H = nZ. Written additively,
a + H = b + H b a H n | b a a b (mod n)
The cosets of H are precisely the equivalence classes modulo n. Indeed you’ve likely encountered
the main proof in the context of modular arithmetic.
Proof. We start by verifying the first connective.
y xH h H such that y = xh x
1
y = h H
Now define a relation on G via x y x
1
y H. We claim that this is an equivalence relation:
Reflexivity: x x since x
1
x = e H.
Symmetry: x y = x
1
y H = y
1
x = (x
1
y)
1
H = y x.
Transitivity: If x y and y z then x
1
y H and y
1
z H. But H is closed, whence
x
1
z = (x
1
y)(y
1
z) H = x z
The equivalence classes therefore partition G. Since x y y xH, the equivalence class of x is
indeed the left coset xH.
The subgroup status of H is precisely what guarantees a partition (compare Theorem 2.14):
Reflexivity: H contains the identity (and is thus non-empty).
Symmetry: H satisfies the inverse axiom.
Transitivity: H is closed under the group operation.
When H is not a subgroup, the coset construction need not produce a partition.
Example 5.7. The subset H = {0, 1} Z
3
is not a subgroup. Its left ‘cosets’ fail to partition Z
3
:
H = {0, 1}, 1 + H = {1, 2}, 2 + H = {2, 1}
51
By combining the criteria in Theorem 5.5, we obtain a useful result for identifying when a subgroup
is normal without explicitly having to compute its cosets. The proof is an exercise.
Corollary 5.8. Normal subgroups are precisely those which are closed under conjugation:
H G H G and g G, h H, ghg
1
H
Since this holds for all g, we may equivalently observe g
1
hg H.
Exercises 5.1. Key concepts: left/right cosets & partitioning normal subgroup kernels are normal
1. Find the cosets of each subgroup: since the groups are abelian, left and right cosets are equal.
(a) 2Z Z (b) 4Z 2Z (c)
4
Z
10
(d)
6
Z
30
(e)
20
Z
30
2. Find the cosets of H =
(0, 0), (2, 0), (0, 2), (2, 2)
Z
4
×Z
4
3. Find the left and right cosets of {ρ
0
, ρ
1
, ρ
2
} D
3
. Is the subgroup normal?
4. (a) Find the left and right cosets of H := {e, (1 2 3), (1 3 2)} A
4
. Is the subgroup normal?
(b) Repeat the question for the subgroup V := {e, (1 2)(3 4), (1 3)(2 4), (1 4)(2 3)} of A
4
.
5. Revisit Examples 4.7, particularly its use of cycle notation.
(a) Find the left and right cosets of H = {ρ
0
, δ
1
} D
4
. Is H normal?
(b) Repeat for the subgroup K = {ρ
0
, ρ
2
}.
6. Prove Lemma 5.3: every subgroup of an abelian group is normal.
7. Suppose H is a subset of G, but not necessarily a subgroup.
(a) If H has only one element, show that the sets gH = {gh : h H} do partition G.
(b) Show that the ‘cosets’ of H = {1, 3} also partition Z
4
, even though H is not a subgroup.
8. Let H = {σ S
4
: σ(4) = 4}. Show that H is a subgroup of S
4
. Is it normal?
9. Prove Corollary 5.8.
10. (a) Suppose G = H ×K, J H and let
e
J = J × {e
K
}. Prove that
e
J G.
(When J = H this is often written H (H ×K). A similar result holds for K.)
(b) Explain how Example 5.4.3 fits into part (a).
11. Let H, K be subgroups of G. Define on G by
a b h H, k K such that a = hbk
(a) Prove that is an equivalence relation on G and describe the elements of the equivalence
class of a G; this is called a double coset.
(b) Compute the double cosets of H = {e, (1 2)} and K = {e, (1 3)} as subgroups of S
3
.
52
5.2 Lagrange’s Theorem & Indices
We’ve been inching towards a powerful result; hopefully you’ve hypothesized this already!
Theorem 5.9 (Lagrange). In a finite group, the order of a subgroup divides the order of the group:
21
H G =
|
H
|
|
G
|
Note that the converse is false: e.g. A
4
has order 12, but no subgroup of order 6 (Exercise 4.3.7). The
argument is merely a generalized version of the proof of Theorem 4.25 (
|
A
n
|
=
1
2
|
S
n
|
).
Proof. Suppose H G and fix g G. The function left-multiplication by g
L
g
: H gH : h 7 gh
is a bijection (inverse function (L
g
)
1
= L
1
g
). Every left coset of H therefore has the same cardinality
as H. Since the left cosets partition G (Theorem 5.5), we conclude that
|
G
|
= (number of left cosets of H) ·
|
H
|
=
|
H
|
|
G
|
We could instead have used the right coset partition. Here is an example of Lagrange’s power.
Corollary 5.10. Up to isomorphism, there is a unique group of prime order p, namely Z
p
.
Proof. Suppose G is a group with prime order p. Since p 2, we may choose some element g = e.
The order of the cyclic subgroup
g
G satisfies:
|
g
|
2 since g = e.
|
g
|
= 1 or p by Lagrange, since p is prime.
We conclude that
|
g
|
= p. But then G =
g
is cyclic and so isomorphic to Z
p
(Theorem 3.7).
Example 5.11. G = Z
4
× Z
2
has order 8 so its non-trivial proper subgroups can only have orders
2 or 4 and are thus isomorphic to Z
2
, Z
4
or V. These can be identified by thinking about all pos-
sible generators; V requires three elements of order 2 which we indeed have! Here is the subgroup
diagram: all proper subgroups are cyclic except V = {(0, 0), (2, 0), (0, 1), (2, 1)}.
generator order subgroup
(1, 0) or (3, 0) 4 {(0, 0), (1, 0), (2, 0), (3, 0)}
(1, 1) or (3, 1) 4 {(0, 0), (1, 1), (2, 0), (3, 1)}
(2, 0) 2 {(0, 0), (2, 0)}
(0, 1) 2 {(0, 0), (0, 1)}
(2, 1) 2 {(0, 0), (2, 1)}
(0, 0) 1 {(0, 0)}
Z
4
×Z
2
(1, 0)
V
(1, 1)
(0, 1)
(2, 0)
(2, 1)
(0, 0)
21
This is often misremembered as ‘the order of an element divides the order of the group,’ which is the special case when
H is a cyclic subgroup of G. Corollary 3.16 is the even more special case when G is cyclic:
s
Z
n
has order
n
gcd( s,n)
.
53
The proof of Lagrange tells us that the number of left and right cosets of H G is identical: both equal
the quotient
|
G
|
|
H
|
. This motivates a new concept.
Definition 5.12. The index (G : H) of a subgroup H G is the cardinality of the set of (left) cosets:
(G : H) =
|
{gH : g G}
|
The index is also the cardinality of the set of right cosets (Exercise 9). If G is finite, then (G : H) =
|
G
|
|
H
|
.
Examples 5.13. 1. If G = Z
20
and H =
2
= {0, 2, 4, . . . , 18}, then there are (G : H) =
20
10
=
|
G
|
|
H
|
= 2
cosets (left and right are equal here):
H =
2
= {0, 2, 4, . . . , 18} and 1 + H = {1, 3, 5, . . . , 19}
2. Recall the orthogonal and special orthogonal groups (Example 2.28 & Exercise 2.2.8):
O
n
(R ) = {A M
n
(R ) : A
T
A = I}, SO
n
(R ) = {A O
n
(R ) : det A = 1}
Since every orthogonal matrix has determinant ±1, it feels as if SO
n
(R ) should be ‘half of
O
2
(R ). Since both groups are infinite (indeed uncountable), we need the index to confirm this
intuition. Recall Theorem 5.5: given A, B O
n
(R ),
A SO
n
= B SO
n
(R ) B
1
A SO
n
(R )
det(B
1
A) = 1
det B = det A
Since determinant can take only two possible values in O
n
(R ), we conclude that there are pre-
cisely two cosets
O
n
(R ) : SO
n
(R )
= 2.
3. The integers form an (additive) subgroup of the real numbers R. Observe that
xZ = yZ y x Z
It follows that there is exactly one coset for every real number in the half-open interval [0, 1);
otherwise said, ψ(x) = xZ defines a bijection of the interval [0, 1) with the set of cosets {xZ}.
We conclude that there are uncountably many cosets!
Theorem 5.14. If K H G is a sequence of subgroups, then
(G : K) = (G : H)(H : K)
If G is a finite group then the result is essentially trivial:
(G : K) =
|
G
|
|
K
|
=
|
G
|
|
H
|
·
|
H
|
|
K
|
= (G : H)(H : K)
Our proof also covers infinite groups (and even infinite indices, if multiplication of such makes sense
to you!). To help understand how it works, first consider a straightforward example.
54
Example (5.13.1, cont.). If G = Z
20
, H =
2
and K =
10
, then
K = {0, 10} H = {0, 2, 4, 6, 8, 10, 12, 14, 16, 18} G = {0, 1, 2, 3, . . . , 19}
so we have the required subgroup relationship. Here are the indices and (left) cosets in each case:
(G : H) = 2 with cosets H and 1 + H. Each coset has the form g
i
+ H with either g
0
= 0 or
g
1
= 1: these values are representatives of the two cosets.
(H : K) =
10
2
= 5 cosets, with representatives h
0
= 0, h
1
= 2, h
2
= 4, h
3
= 6, h
4
= 8:
K = {0, 10}, 2 + K = {2, 12}, 4 + K = {4, 14}, 6 + K = {6, 16}, 8 + K = {8, 18}
(G : K) =
20
2
= 10 = (G : H)(H : K). The cosets this time are
K = {0, 10}, 1 + K = {1, 11}, 2 + K = {2, 12}, . . . , 9 + K = {9, 19}
with representatives 0, 1, . . . , 9. The crucial observation for the proof is that representatives are
formed by summing those above: the cosets of K in G are precisely (g
i
+ h
j
) + K.
Proof. Choose an element g
i
from each left coset of H in G and an element h
j
from each left coset of
K in H. Plainly
(G : H) =
|
{g
i
}
|
and (H : K) =
{h
j
}
We claim that the left cosets of K in G are precisely the sets (g
i
h
j
)K. Certainly each of these is a coset;
we show that these partition G, whence the collection {(g
i
h
j
)K} must comprise all left cosets.
Every g G lies in some left coset of H, so g
i
G with g g
i
H.
Now g
i
1
g H lies in some left coset of K in H, so h
j
H with g
i
1
g h
j
K.
But then g (g
i
h
j
)K, whence every g G lies in at least one set (g
i
h
j
)K.
Suppose y g
i
h
j
K g
α
h
β
K. Since K H and the left cosets of H partition G, we have
y g
i
H g
α
H = g
α
= g
i
But then g
i
1
y h
j
K h
β
K = h
β
= h
j
similarly, since the left cosets of K in H partition H. It
follows that the sets (g
i
h
j
)K are disjoint.
Since the left cosets of K in G are given by {(g
i
h
j
)K}, it is immediate that
(G : K) =
{g
i
h
j
}
=
|
{g
i
}
|
{h
j
}
= (G : H)(H : K)
Exercises 5.2. Key concepts: Lagrange’s Theorem index of a subgroup (counting cosets)
1. Find the indices of the following subgroups:
(a)
9
Z
12
(b) 6Z 2Z (c) ( Q
+
, ·) (Q
×
, ·)
55
2. Let G = Z
8
, H =
2
= {0, 2, 4, 6} and K =
4
= {0, 4}. Write out all the cosets for the three
subgroup relations K H, H G and K G, and verify the index multiplication formula
(Theorem 5.14).
3. Let G = S
4
and consider the subgroup tower K H G where
K = {e, (1 2 3), (1 3 2)}
=
Z
3
and H = {σ S
4
: σ(4) = 4}
=
S
3
(a) Write out the elements of the two left cosets K H, namely K = eK and (1 2)K with
representatives h
0
= e and h
1
= (1 2).
(b) Repeat part (a) for the four cosets of H in G = S
4
: observe that g
0
= e, g
1
= (1 4),
g
2
= (2 4), g
3
= (3 4) are representatives.
(c) Compute the eight left cosets of K in S
4
and verify that they are (g
i
h
j
)K in accordance with
the proof of Theorem 5.14.
4. Let G have order pq where p, q are both prime. Show that every proper subgroup of G is cyclic.
5. Use Lagrange’s Theorem to prove that all proper subgroups of Z
3
×Z
3
are cyclic. Hence con-
struct its subgroup diagram.
6. Find the subgroups of Z
6
×Z
2
and draw its subgroup diagram.
(Hint: At least one proper subgroup is non-cyclic!)
7. Suppose (G : H) = 2. Prove that H is a normal subgroup of G.
8. Prove that {e} and G are both normal subgroups of G: what are the cosets and the indices in
each case?
(Remember that G could be infinite!)
9. For each left coset gH of H in G, choose a representative g
j
. Prove that the function
Φ : g
j
H 7 Hg
1
j
is injective from the set of left cosets to the set of right cosets.
(With the reverse argument this shows that the sets of left and right cosets have the same cardinality)
10. Let G = {a + b
2 : a, b Z}.
(a) Prove that G is a group under addition.
(b) Prove that H = {3m + 2n
2 : m, n Z} is a subgroup of index six in G.
(Hint: what does it mean for a + b
2 and c + d
2 to lie in the same coset of H?)
11. We modify Example 5.13.3.
(a) The sets Q and Z are both groups under addition. Show that there is exactly one coset of
Z in Q for each rational q [0, 1). Hence conclude that (Q : Z) =
0
is countably infinite.
(b) Describe the index of U
n
(n
th
roots of unity) as a subgroup of the circle group S
1
.
56
5.3 Factor Groups
Given H G, we ask whether the set of left cosets {gH : g G} has a natural group structure
inherited from that of G. In our motivating Example (5.1) this is precisely how Z
3
was created from
the integers, by ‘squashing’ all elements of each coset down to a single object. We simply want to do
this in general. To see how this might (or might not) work, recall some previous examples.
Examples (5.4, cont). 1. The set of (left) cosets for H =
2
= {0, 2, 4} Z
6
is
H, 1 + H
=
n
{0, 2, 4}, {1, 3, 5}
o
We use addition in Z
6
to define addition of cosets via
(a + H) (b + H) := (a + b) + H
This seems nice, though consider the steps of the computation
more carefully:
(a) First choose representatives a and b of the two cosets.
(b) Then add within the original group a + b Z
6
.
(c) Finally, take the left coset (a + b) + H.
If is to make sense, the outcome (a + b) + H must be independent
of the choices in step (a). For instance, to properly conclude that
H (1 + H) = 1 + H we must check nine possibilities:
a {0, 2, 4}, b {1, 3, 5} = a + b {1, 3, 5} (modulo 6)
0
0 0
1
2
3
4
5
1
1
1
2
3
4
5
0
2
2
2
3
4
5
0
1
3
3
3
4
5
0
1
2
4
4
4
5
0
1
2
3
5
5
5
0
1
2
3
4
+
6
Addition in Z
6
H
HH
H
1+H
1+H
1+H1+H
Addition in {H, 1 + H}
To save time, we verify all possibilities simultaneously: If x a + H and y b + H, then
(x + y) (a + b) = (x a) + (y b) H = (x + y) + H = (a + b) + H
Addition of cosets is therefore well-defined. The second table above suggests that the set of
cosets forms a group under ; indeed ψ(x) = x + H defines an isomorphism of Z
2
with this
so-called factor group.
2. We repeat the process for the subgroup H = {e, µ
1
} D
3
. The left cosets are
H = µ
1
H = {e, µ
1
}, ρ
1
H = µ
3
H = {ρ
1
, µ
3
}, ρ
2
H = µ
2
H = {ρ
2
, µ
2
}
As above, we attempt to define the ‘natural’ operation on the set of left cosets
aH bH := (ab)H (ab is composition/multiplication within D
3
)
This time there is a serious problem: note that ρ
1
H = µ
3
H, however,
ρ
1
H ρ
1
H = (ρ
1
ρ
1
)H = ρ
2
H µ
3
H µ
3
H = (µ
3
µ
3
)H = eH = H
This is a contradiction: ρ
2
H = H, but the result of both calculations should be the same. The
freedom of choice (part (a)) in the definition of leads to different outcomes: the natural
operation does not exist (is not well-defined), and thus cannot produce a group structure.
57
Well-definition of the Factor Group Structure
The examples indicate that only some subgroups H G behave nicely when trying to view the set
of left cosets as a group. But which subgroups? To answer this question, we repeat some of our
discussion abstractly. Let H be a subgroup of G and define the natural multiplication of left cosets
22
aH ·bH := (ab)H ()
This is well-defined precisely when
xH = aH, yH = bH = (xy)H = (ab)H (†)
If the natural multiplication of cosets is well-defined, the fact that hH = H and gH = gH for any
h H, g G, tells us that
(hg)H = gH, or equivalently g
1
hg H (Theorem 5.5)
Since this holds for all g G and h H, Corollary 5.8 says that H is a normal subgroup of G. Not
only is the converse also true, but the resulting structure forms a group under this operation.
Theorem 5.15. Suppose H G and consider the natural operation () on the set of (left) cosets.
1. () is well-defined if and only if H is a normal subgroup of G.
2. In such cases, () defines a group structure on the set of cosets.
Since this process only works when H is a normal subgroup, the prefix ‘left’ is irrelevant.
Definition 5.16. Suppose H G. The group of cosets
G
H
:=
gH : g G
(read G mod H’) under
the natural operation () is termed a factor group (or quotient group).
Factor group notation looks like division in part because if G is finite, then
G
H
= (G : H) =
|
G
|
|
H
|
.
Proof. 1. () The above discussion shows that well-definition of () implies H G.
() Assume H G, and suppose xH = aH and yH = bH. By Theorem 5.5, h := x
1
a and
˜
h := y
1
b are both in H. But then,
(xy)
1
(ab) = y
1
(x
1
a)b = y
1
hb = (y
1
hy)
˜
h
which lies in H by Corollary 5.8 (y
1
hy H). We conclude that (xy)H = (ab)H (†).
2. Since the natural operation is well-defined, we need only verify the group axioms for
G
H
, ·
.
Closure: Given aH, bH
G
H
, we see that aH · bH = (ab)H is also coset.
Associativity: aH ·(bH ·cH) = aH ·(bc)H = a(bc)H. Similarly (aH ·bH) ·cH = (ab)cH. By the
associativity of (G, ·), these cosets are identical.
Identity: The identity coset H = eH does the job: eH · aH = (ea)H = aH = (ae)H = aH · eH.
Inverse: a
1
H · aH = (a
1
a)H = eH = H, etc., therefore (aH)
1
= a
1
H.
22
Since this operation arises naturally from that on G, we use the same notation (here multiplication).
58
’Identifying’ Factor Groups
The first goal when faced with a factor group
G
H
is often to identify it by recognizing some well-
understood group to which it is isomorphic. In Section 6.2 we’ll develop the main piece of abstract
machinery for doing this. Since this upcoming approach can be difficult to apply, it is worth first
spending a little time with some basic examples. In particular, we can straightforwardly describe the
factor groups of every cyclic group: by Theorem 3.7, we need only do this for Z and Z
n
. . .
Factor Groups of Z (modular arithmetic done right) If n is a positive integer, its integer multiples
nZ =
n
form a (normal) subgroup of Z. The coset of nZ containing x Z is plainly
x + nZ = {x + kn : k Z} =
y Z : y x (mod n)
This coset is what we have been calling x in Z
n
! This provides the formal definition of Z
n
(super-
seding Definition 2.16) and trivially demonstrating that Z
n
is an abelian group.
Definition 5.17. Let n N. The group Z
n
is the factor group
Z
nZ
.
We typically drop the repeated nZ terms when calculating, recovering our familiar notation: e.g.
4 + 5 = 2 Z
7
means (4 + 7Z) + (5 + 7Z) = 2 + 7Z
Z
7Z
Factor Groups of Finite Cyclic Groups The first example on page 57 shows that
Z
6
2
=
Z
2
. This
generalizes straightforwardly.
Example 5.18.
5
= {0, 5, 10, 15} Z
20
has factor group
Z
20
5
=
5
, 1 +
5
, 2 +
5
, 3 +
5
, 4 +
5
which is isomorphic to Z
5
via the isomorphism
ψ : Z
5
Z
20
5
: x 7 x +
5
Theorem 5.19. If d | n, then
Z
n
d
=
Z
d
. More generally,
Z
n
s
=
Z
gcd(s,n)
.
Proof. Define ψ : Z
d
Z
n
d
: x 7 x +
d
. We prove that this is an isomorphism.
Well-definition/injectivity: For any x, y Z
d
,
x = y ( Z
d
) x y
d
x +
d
= y +
d
ψ(x) = ψ(y)
Surjectivity: Any coset x +
d
(being ψ(x)) lies in range(ψ).
Homomorphism: For any x, y Z
d
,
ψ(x + y) = (x + y) +
d
=
x +
d
+
y +
d
= ψ(x) + ψ(y)
The general case follows by Corollary 3.16: if d = gcd(s, n), then
s
=
d
.
59
Further Examples Na
¨
ıve identification of factor groups often boils down to a two-step hack.
1. Find the order of the factor group
G
H
by computing the index (G : H).
2. Determine which group of order ( G : H) is correct.
If
G
H
is abelian, the Fundamental Theorem (3.26) might supply candidates. Step 2 can often be
accomplished by considering orders of elements (cosets). The simple observation can help with this.
Lemma 5.20. Let
G
H
be a factor group. Then (gH)
m
= H g
m
H (mg H if G additive).
By Corollary 3.9, the order of the element gH
G
H
is the smallest m N for which g
m
H.
Moreover, this number min{m N : g
m
H} divides the order of g (in G).
Examples 5.21. Let G = Z
4
×Z
8
. We identify the factor group
G
H
for three subgroups H.
1. The subgroup H =
(0, 1)
=
(0, 0), (0, 1), . . . , ( 0, 7)
has 8 elements, so the factor group
G
H
has order
|
G
|
|
H
|
=
4·8
8
= 4. By the Fundamental Theorem,
G
H
is isomorphic to Z
4
or Z
2
×Z
2
.
We can decide which by considering the orders of elements in
G
H
:
Z
2
×Z
2
: Every element has order at most 2.
Z
4
: There exists a generator with order 4.
Start playing with elements (cosets)! It doesn’t take long to observe that
k(1, 0) = (k, 0) H 4 | k (†)
Otherwise said, (1, 0) + H
G
H
has order 4: we conclude that
G
H
=
Z
4
. Since (1, 0) + H is a
generator, this approach provides an explicit isomorphism ψ : Z
4
=
G
H
:
ψ(x) = (x, 0) + H
=
(x, 0), (x, 1), . . . , (x, 7)
2. The subgroup H =
(0, 2)
=
(0, 0), (0, 2), (0, 4), (0, 6)
has 4 elements, so
G
H
=
32
4
= 8.
The factor group is isomorphic to one of Z
8
, Z
4
×Z
2
or Z
2
×Z
2
×Z
2
.
Exactly as in (), (1, 0) + H
G
H
has order 4. This rules out Z
2
×Z
2
×Z
2
as a candidate.
Z
8
is ruled out since every (x, y) + H has order at most 4: for any (x, y),
4(x, y) = (4x, 4y) = (0, 4y) = 2y(0, 2) H
By process of elimination, we conclude that
G
H
=
Z
4
×Z
2
.
3. The subgroup H =
(2, 4)
=
(0, 0), (2, 4)
produces a factor group of order
G
H
=
32
2
= 16,
so we must consider five non-isomorphic possibilities:
Z
16
, Z
2
×Z
8
, Z
4
×Z
4
, Z
2
×Z
2
×Z
4
, Z
2
×Z
2
×Z
2
×Z
2
The coset ( 0, 1) + H
G
H
has order 8, since
k
(0, 1) + H
= H (0, k) H 8 | k
Every (x, y) + H has order dividing 8, since
8(x, y) = (8x, 8y) = (0, 0) H
The only candidate satisfying both properties is Z
2
×Z
8
.
60
For factor groups of infinite or non-abelian groups, more creative strategies might be required.
Examples 5.22. 1. As seen in Exercise 5.2.7, every index-2 subgroup H G is normal. Since there is
only one group of order 2 up to isomorphism, in such a case
G
H
=
Z
2
.
For instance, R
n
D
n
has index (D
n
: R
n
) = 2. The two cosets are,
Rotations: R
n
=
ρ
0
, . . . , ρ
n1
Reflections: µR
n
=
µ
1
, . . . , µ
n
, where µ is any reflection.
R
n
µR
n
R
n
R
n
µR
n
µR
n
µR
n
R
n
The Cayley table for the factor group
D
n
R
n
= {R
n
, µR
n
} is shown. Consider how the calcula-
tion (µR
n
)(µR
n
) = R
n
says that the composition of two reflections is a rotation!
2. In Exercise 5.1.4, we saw that the alternating group A
4
has the Klein four-group identified as
V =
e, (1 2)(3 4), (1 3)(2 4), (1 4)(2 3)
as a normal subgroup. Since the factor group has order (A
4
: V) =
12
4
= 3 and there is only one
group of order 3 up to isomorphism, we conclude that
A
4
V
=
Z
3
.
3. (Example 5.4.3, simplified) Given W = {(0, y) : y R} R
2
, each coset (vertical line)
(x, y) : y R
= (x, 0) + W
intersects the x-axis at a unique point (x, 0). Otherwise said, we have a bijection,
ψ : R
R
2
W
: x 7 (x, 0) + W
This is moreover an isomorphism, since
ψ(x
1
+ x
2
) = (x
1
+ x
2
) + W = (x
1
+ W) + (x
2
+ W) = ψ(x
1
) + ψ(x
2
)
4. (Hard) Let G = Z ×Z
4
and consider the subgroup
H =
(2, 1)
=
. . . , (4, 2), (2, 3), (0, 0), (2, 1), (4, 2), (6, 3), (8, 0), (10, 1), . . .
We cannot count cosets using the index formula. Instead we find a simple representative of
each coset:
If x = 2n is even: (x, y) + H = (2n, y) + H = (0, y n) + H
If x = 2n + 1 is odd: (x, y) + H = (2n + 1, y) + H = (1, y n) + H
In any coset, there is a unique representative either of the form (0, z) or (1, z), where z Z
4
: it
follows that there are 8 cosets. To finish things off, observe that
k(1, 0) = (k, 0) H 8 | k
whence ( 1, 0) + H
G
H
has order 8: we conclude that
G
H
=
Z
8
.
We’ll see more examples, and revisit others, in Section 6.2 once we’ve developed more machinery.
61
Exercises 5.3. Key concepts:
Factor group
G
H
a group H G Z
n
:=
Z
nZ
identifying
G
H
1. List the cosets of the subgroup H =
3
in G = Z
15
. By mimicking the proof of Theorem 5.19,
show that ψ : Z
3
7
G
H
: x 7 x + H is a well-defined isomorphism.
2. (Exercise 5.1.2 cont.) If H =
(0, 0), (0, 2), (2, 0), (2, 2)
, identify the factor group
Z
4
×Z
4
H
.
3. Identify the factor group
G
H
where H =
(2, 4)
G = Z
4
×Z
6
.
4. Let G = Z
9
×Z
9
.
(a) State the elements of the subgroup H =
(3, 6)
. Now identify
G
H
by showing that every
element of the factor group has order at most 9 and that it contains an element of order 9.
(b) Repeat with H =
3
×
6
(this isn’t a trick question).
5. Let G be any group. To what groups are
G
{e}
and
G
G
isomorphic?
6. (a) If G is abelian and H G, prove that
G
H
is abelian.
(b) If
G
H
is abelian, can we conclude that G and/or H is abelian? Explain.
7. (a) Let G be a cyclic group with subgroup H. Prove that
G
H
is cyclic.
(b) If
G
H
is cyclic, does it follow that G is cyclic? Explain.
8. (Exercise 5.2.10, cont.) The factor group
G
H
is abelian and of order 6, whence it is cyclic. Prove
this explicitly by finding a generator and thus an isomorphism ψ : Z
6
G
H
.
9. (Example 5.22.2, cont.) Find an explicit isomorphism ψ : Z
3
=
A
4
V
.
10. Exercise 5.1.5 showed that {ρ
0
, ρ
2
} is a normal subgroup of D
4
. To what well-known group is
the factor group
D
4
{ρ
0
, ρ
2
}
isomorphic? Prove your assertion.
11. If H G has index (G : H) = 2, then the factor group
G
H
is necessarily isomorphic to Z
2
.
(a) Prove that ψ : Z
2
S
n
A
n
defined by ψ(x) = (1 2)
x
A
n
is an isomorphism.
(b) Find an explicit isomorphism ψ : Z
2
O
n
(R)
SO
n
(R)
.
12. (Exercise 5.1.10, cont.) Suppose G = H ×K, J H and
b
J = J × {e
K
}. Prove that
G
e
J
=
H
J
×K.
Can you quickly verify Examples 5.21.1, 2, and 5.22.3 in this language?
13. Complete the proof of Lemma 5.20 by showing that the order of the element/coset gH
G
H
divides the order of the element g G.
(Hint: Use the division algorithm similarly to the proof of Theorem 3.12)
14. (Hard) Let H =
(2, 3)
G = Z
5
×Z. Prove that
G
H
=
Z
15
.
15. (Hard!!) If G = Z
10
×Z
6
×Z and H =
(4, 2, 3)
, identify
G
H
as a direct product Z
m
×Z
n
.
(Hint: show that exactly one representative of the coset (x, y, z) + H has z = 0, 1 or 2)
62
6 Homomorphisms & The First Isomorphism Theorem
The main goal of this chapter is to discuss the link between homomorphisms, normal subgroups and
factor groups. The key result is the subject of Section 6.2: the First Isomorphism Theorem.
6.1 Kernels, Images and Normal Subgroups
In Section 2.5 we established several important facts that are worth refreshing and summarizing here.
Definition 6.1. The kernel and image of a homomorphism
23
ϕ : G L are the sets
ker ϕ =
g G : ϕ(g) = e
, Im ϕ =
ϕ(g) : g G
Lemma 6.2. Suppose ϕ : G L is a homomorphism.
1. (Exercise 2.5.11c) ϕ maps identity to identity and inverse to inverse:
ϕ(e
G
) = ϕ(e
L
) and g G,
ϕ(g)
1
= ϕ(g
1
)
2. (Exercise 2.5.12) ker ϕ is a subgroup of G and Im ϕ is a subgroup of L.
Do these exercises now, even if you did them earlier!
Examples 6.3. 1. The function ϕ(x) = 2x (mod 4) defines a homomorphism ϕ : Z Z
4
with
ker ϕ = {x Z : 2x 0 (mod 4)} = 2Z, Im ϕ = {0, 2}
Plainly ker ϕ is a subgroup of Z and Im ϕ is a subgroup of Z
4
.
2. Every linear map T : V W between vector spaces is a homomorphism. In linear algebra its
kernel is better known as its nullspace
ker T = N(T) =
v V : T(v) = 0
If T = L
A
: R
n
R
m
is left-multiplication by a matrix A, then Im T is the column space of A.
Since the groups in these examples are abelian, all subgroups are automatically normal. In fact the
kernel of every homomorphism is a normal subgroup, though the same cannot be said for images.
Lemma 6.4. If ϕ : G L is a homomorphism, then ker ϕ G.
Proof. The kernel is a subgroup by part 2 of Lemma 6.2, so we need only check normality. For this
we appeal to the conjugation criterion (Corollary 5.8). If g G and k ker ϕ, then
ϕ
gkg
1
= ϕ(g)ϕ(k)ϕ(g)
1
= ϕ(g)ϕ(g)
1
= e
L
= gkg
1
ker ϕ
Note how the first equality used the homomorphism property and part 1 of the Lemma.
23
All homomorphisms in this chapter (indeed outside of Section 2.5) are between groups.
63
Examples 6.5. 1. (Recall Exercises 2.4.8a & 2.5.1g) We’ve seen that det : GL
n
(R ) R
×
is a homo-
morphism. We therefore obtain a normal subgroup
ker det = {A GL
n
(R ) : det A = 1} GL
n
(R )
Otherwise said, SL
n
(R ) is a normal subgroup of GL
n
(R ).
2. (Example 5.4.2, cont.) ϕ : Z
2
D
3
defined by ϕ(x) = µ
x
1
is a homomorphism. Certainly
ker ϕ = {0} Z
2
, however Im ϕ = {e, µ
1
} is not a normal subgroup of S
3
.
Kernels, Images and Describing Homomorphisms
Since every kernel is a normal subgroup, it makes sense to ask how its cosets may be counted and
distinguished.
Lemma 6.6. Let ϕ : G L be a homomorphism. There is precisely one coset of ker ϕ for each
element of Im ϕ:
g
1
ker ϕ = g
2
ker ϕ ϕ(g
1
) = ϕ(g
2
)
Otherwise said, the index equals the order of the image: (G : ker ϕ) =
|
Im ϕ
|
.
The proof is an abstraction of Example 5.13.2 (where ϕ = det : O
2
(R ) R
×
).
Proof. For all g
1
, g
2
G, we have
g
1
ker ϕ = g
2
ker ϕ g
1
2
g
1
ker ϕ (Theorem 5.5)
ϕ(g
1
2
g
1
) = e
L
(Definition of ker ϕ)
ϕ(g
2
)
1
ϕ(g
1
) = e
L
(Homomorphism properties)
ϕ(g
1
) = ϕ(g
2
)
The Lemma is really part of the upcoming punchline in Section 6.2. For the moment we apply it to
help describe and count homomorphisms.
Theorem 6.7. Let ϕ : G L be a homomorphism. If G is finite, then Im ϕ is a finite group whose
order divides that of G. The same holds for L. Otherwise said,
|
G
|
< =
|
Im ϕ
|
|
G
|
and
|
L
|
< =
|
Im ϕ
|
|
L
|
If both groups are finite, then
|
Im ϕ
|
gcd
|
G
|
,
|
L
|
).
Proof. If G is a finite group, then ker ϕ G is finite. Apply Lemma 6.6 to see that
|
Im ϕ
|
= (G : ker ϕ) =
|
G
|
|
ker ϕ
|
is a divisor of
|
G
|
. The second situation
|
Im ϕ
|
|
L
|
is Lagrange’s Theorem (5.9).
64
Examples 6.8. 1. If ϕ : G L is a homomorphism and gcd
|
G
|
,
|
L
|
= 1, then we have
|
Im ϕ
|
= 1.
The image is necessarily the trivial subgroup Im ϕ = {e
L
} and there is exactly one homomor-
phism, namely the trivial homomorphism (g G, ϕ(g) = e
L
).
2. Consider all homomorphisms ϕ : Z
4
S
3
. Since the domain is cyclic, describing ϕ( 1) is
enough to obtain ϕ(x) =
ϕ(1)
x
. Since gcd(
|
Z
4
|
,
|
S
3
|
) = 2, we have
|
Im ϕ
|
= 1 or 2. Either:
Im ϕ = {e} and we obtain the trivial homomorphism ϕ
0
(x) = e.
Im ϕ is a subgroup of order 2, of which S
3
has exactly three: {e, (2 3)}, {e, (1 3)}, {e, (1 2) }.
This results in three further homomorphisms (for a total of four)
ϕ
1
(x) = (2 3)
x
, ϕ
2
(x) = (1 3)
x
, ϕ
3
(x) = (1 2)
x
We now turn to the general question of homomorphisms between finite cyclic groups ϕ : Z
m
Z
n
.
Two facts make this relatively simple:
1. As above, choosing ϕ(1) defines the homomorphism: ϕ(x) = ϕ(1) + ··· + ϕ(1) = ϕ(1) · x.
2.
|
Im ϕ
|
must divide d = gcd(m, n). Since Z
n
has exactly one subgroup of each order dividing n
(Corollary 3.16), Im ϕ must be a subgroup of the unique subgroup
n
d
Z
n
of order d:
Im ϕ
D
n
d
E
=
0,
n
d
,
2n
d
, . . . ,
(d 1)n
d
It is enough to see what happens if we let ϕ(1) be each element of this subgroup in turn. . .
Corollary 6.9. There are d = gcd(m, n) distinct homomorphisms ϕ : Z
m
Z
n
, namely
ϕ
k
(x) =
kn
d
x where k = 0, . . . , d 1
Proof. We check that each ϕ
k
is well-defined; the calculation uses the fact that
m
d
is an integer.
x = y Z
m
= y = x + λm for some m Z
= ϕ
k
(y) = ϕ
k
(x + λm) =
kn
d
(x + λm) =
kn
d
x + λk
m
d
n =
kn
d
x
= ϕ
k
(x)
(in Z
n
)
By observation 1, each ϕ
k
is a homomorphism. Observation 2 says there are no other candidates.
Example 6.10. We describe all homomorphisms ϕ : Z
12
Z
20
.
Since gcd( 12, 20) = 4, we see that Im ϕ
5
= {0, 5, 10, 15} Z
20
. There are four choices:
ϕ
0
(x) = 0, ϕ
1
(x) = 5x, ϕ
2
(x) = 10x, ϕ
3
(x) = 15x (mod 20)
We similarly see that there are four distinct homomorphisms ψ : Z
20
Z
12
:
ψ
0
(x) = 0, ψ
1
(x) = 3x, ψ
2
(x) = 6x, ψ
3
(x) = 9x (mod 12)
Most of these examples rely on the domain of a homomorphism being cyclic (observation 1). It is
typically much harder to find and describe homomorphisms when the domain is non-cyclic.
65
Exercises 6.1. Key concepts: (G : ker ϕ) =
|
Im ϕ
| |
Im ϕ
|
gcd
|
G
|
,
|
L
|
1. Check that you have a homomorphism (use Corollary 6.9) and compute its kernel and image.
(a) ϕ : Z
8
Z
14
defined by ϕ(x) = 7x (mod 14).
(b) ϕ : Z
36
Z
20
defined by ϕ(x) = 5x (mod 20).
2. Describe all homomorphisms between the groups:
(a) ϕ : Z
15
Z
80
(b) ϕ : Z Z
3
(c) ϕ : Z
6
D
4
(d) ϕ : Z
15
A
4
3. State a linear map T : R
2
R
2
whose kernel is the y-axis (a normal subgroup of R
2
). What
type of linear map is the function T?
4. Find the kernel and image of each homomorphism and verify that ker ϕ is normal:
(a) The trace of a matrix tr : M
2
(R ) R :
a b
c d
7 a + d.
(b) T : R
3
R
4
: x 7
1 1 1
0 3 1
1 4 2
2 5 3
!
x
5. Explain why the map ϕ is a homomorphism and find ker ϕ:
ϕ : S
n
{1, 1}, ·
: σ 7
(
1 if σ even
1 if σ odd
6. Suppose ϕ : G L is a homomorphism, and that H G and K L are subgroups.
(a) Prove that ϕ(H) := {ϕ(h) : h H} is a subgroup of Im ϕ.
(b) Give an example to show that Im ϕ need not be a normal subgroup of L.
(c) Prove that the inverse image ϕ
1
(K) = {g G : ϕ(g) K} is a subgroup of G.
7. (Exercise 3.2.6, cont.) Prove that the number of distinct isomorphisms ϕ : Z
n
Z
n
equals the
order of the multiplicative group of units (Z
×
n
, ·
n
).
8. (a) Suppose ϕ : Z
(m)
× Z
(n)
L is a well-defined homomorphism.
24
Prove that ϕ(x, y) =
ax + by, where a = ϕ(1, 0) and b = ϕ(0, 1).
(b) Prove that ϕ : Z
m
×Z
n
Z
m
×Z
n
is a well-defined homomorphism if and only if there
exist integers a, b, c, d for which
ϕ(x, y) =
ax + by, cx + dy
, m | bn and n | cm
9. Find all homomorphisms ϕ : Z
2
×Z
7
Z
2
×Z
5
. How do you know that there are no more?
10. Consider ϕ : D
4
D
4
: σ 7 σ
2
. Explain why ϕ is not a homomorphism.
24
As in Theorem 3.7, Z
(m)
is a generic additive cyclic group: either Z or Z
m
.
66
6.2 The First Isomorphism Theorem
Lemma 6.4 says that every kernel is a normal subgroup. In fact all normal subgroups arise this way.
Theorem 6.11 (Canonical Homomorphism). Let G be a group and H G. The function
γ : G
G
H
defined by γ(g) = gH
is a homomorphism with ker γ = H.
Proof. Since H is normal,
G
H
is a factor group. By the definition of multiplication in
G
H
,
γ(g
1
)γ(g
2
) = g
1
H · g
2
H = (g
1
g
2
)H = γ(g
1
g
2
)
whence γ is a group homomorphism. Moreover, the identity in the factor group is H, whence
g ker γ γ(g) = H g H
This might feel a little sneaky, and we’d have preferred a codomain that wasn’t a factor group! Our
hope is vain however, for the next result—arguably the most important in elementary group theory—
shows that every homomorphism with the same kernel is essentially γ in disguise.
Theorem 6.12 (1
st
Isomorphism). Let ϕ : G L be a homomorphism with kernel H. Then
µ :
G
H
Im ϕ defined by µ(gH) = ϕ(g)
is an isomorphism. Otherwise said: ϕ = µ γ and
G
ker ϕ
=
Im ϕ.
A commutative diagram provides a helpful summary: ϕ = µ γ means
we get the same result by following arrows from G to Im ϕ regardless
of the path.
The 1
st
isomorphism theorem has analogues in several other parts of
mathematics: at the very least you should have met its close kin from
linear algebra, the rank–nullity theorem.
G
ϕ
//
γ
!!
Im ϕ
G
ker ϕ
µ
;;
Proof. The factor group exists since ker ϕ G (Lemma 6.4). We check the isomorphism properties:
Well-definition and Bijectivity: These are immediate after writing H = ker ϕ:
g
1
H = g
2
H
Lemma
6.6
ϕ(g
1
) = ϕ(g
2
) µ(g
1
H) = µ(g
2
H)
Homomorphism: For all g
1
H, g
2
H
G
H
,
µ(g
1
H · g
2
H) = µ(g
1
g
2
H) = ϕ(g
1
g
2
) (coset multiplication & definition of µ)
= ϕ(g
1
)ϕ(g
2
) (ϕ is a homomorphism)
= µ(g
1
H)µ(g
2
H)
67
Examples 6.13. 1. Let ϕ : Z
12
Z
20
be the homomorphism ϕ(x) = 5x (mod 20) (Example 6.10). Its
kernel and image are
ker ϕ =
x Z
12
: 5x 0 (mod 20)
= {0, 4, 8} =
4
Z
12
Im ϕ = {5x Z
20
: x Z
12
} = {0, 5, 10, 15} =
5
Z
20
The relevant factor group is
Z
12
ker ϕ
=
n
{0, 4, 8}, {1, 5, 9}, {2, 6, 10}, {3, 7, 11}
o
=
4
, 1 +
4
, 2 +
4
, 3 +
4
The canonical homomorphism γ and the isomorphism µ are
γ(x) = x +
4
Z
12
γ
//
ϕ
++
Z
12
4
µ
//
Im ϕ
µ
x +
4
= 5x x
//
x +
4
//
5x
2. The homomorphism ϕ : R (C
×
, ·) : x 7 e
2πix
has
ker ϕ =
x R : e
2πix
= 1
= Z and Im ϕ = S
1
(S
1
is the circle group)
The canonical homomorphism γ and the isomorphism µ from the theorem are
γ : R
R
Z
: x 7 x + Z and µ :
R
Z
S
1
: x + Z 7→ e
2πix
Think about how this relates to Example 5.13.3.
3. The homomorphism ϕ : Z ×Z Z : (x, y) 7 3x 2y satisfies
ϕ(x, y) = (0, 0) 3x = 2y (x, y) = (2n, 3n) for some n Z
We conclude that ker ϕ =
(2, 3)
. The canonical homomorphism is therefore
γ : Z ×Z
Z ×Z
(2, 3)
: (x, y) 7 (x, y) +
(2, 3)
Since n = ϕ(n, n), the homomorphism is surjective: Im ϕ = Z. By the 1
st
isomorphism theorem,
Z ×Z
(2, 3)
=
Z via µ
(x, y) +
(2, 3)
= 3x 2y
4. ϕ(x, y) = (x, y x) is a well-defined homomorphism ϕ : Z ×Z
6
Z
3
×Z
6
. Moreover,
ϕ(x, y) = (0, 0)
(
x = 3k, and
y = x = 3k Z
6
whence ker ϕ =
(3, 3)
=
. . . , (6, 0), (3, 3), (0, 0), (3, 3), (6, 0), . . .
Moreover, ϕ is surjective: e.g. (m, n) = ϕ(m, n + m). We conclude that
Z ×Z
6
(3, 3)
=
Z
3
×Z
6
via µ
(x, y) +
(3, 3)
) = (x, y x)
68
Identifying Factor Groups (revisited)
The 1
st
isomorphism theorem may be applied to the identification of factor groups. Given H G, we
cook up a homomorphism ϕ : G L with ker ϕ = H and conclude that
G
H
=
Im ϕ. This typically
requires some creativity: there are many options, and both L and a formula for ϕ need to be found
simultaneously! We first revisit some examples from the previous section in this context.
Examples (5.21, mk.II). For each subgroup H of G = Z
4
× Z
8
, we hunt for a suitable homomor-
phism with ker ϕ = H.
1. Given H =
(0, 1)
, we need a homomorphism for which ϕ(0, 1) is the identity. This is very
simple: just ignore y! Define
ϕ : Z
4
×Z
8
Z
4
: (x, y) 7 x
This indeed has kernel ker ϕ = {(0, y) : y Z
8
} = H. By the 1
st
isomorphism theorem,
G
H
=
Im ϕ = Z
4
via the isomorphism µ
(x, y) + H
= x. Note that (x, y) + H = (x, 0) + H, whence µ is the
inverse of the isomorphism ψ : x 7 (x, 0) + H stated previously!
2. Given H =
(0, 2)
we require ϕ(0, 2) to be the identity. This may be achieved by taking y
modulo 2 and defining
ϕ : Z
4
×Z
8
Z
4
×Z
2
: (x, y) 7 (x, y)
This is well-defined
ϕ(x + 4m, y + 8n) = (x + 4m, y + 8n) = (x, y) = ϕ (x, y) (since 2 | 8)
and a homomorphism with required kernel H. Moreover, ϕ is surjective, whence
G
H
=
Im ϕ = Z
4
×Z
2
via the isomorphism µ
(x, y) + H
= (x, y). Once again µ is the inverse of ψ(x, y) = (x, y) + H
in the original example.
3. Finding a homomorphism with kernel H =
(2, 4)
=
(0, 0), (2, 4)
is somewhat trickier. One
approach is to observe that
(x, y) H x 0 ( mod 2) and y 2x 0 (mod 8)
This suggests
ϕ : Z
4
×Z
8
Z
2
×Z
8
: (x, y) 7
x, y 2x
It is worth checking that this is well-defined: the 2x in the second co-ordinate is crucial! Cer-
tainly ϕ has the correct kernel. It is moreover surjective: (m, n) = ϕ(m, n + 2m). In conclusion,
G
H
=
Im ϕ = Z
2
×Z
8
via the isomorphism µ
(x, y) + H
= (x, y 2x).
69
Examples 6.14. Two final examples follow a similar theme, though with infinite groups. We’ll
generalize these in Exercise 10.
1. We identify the factor group
G
H
=
Z ×Z
(1, 1)
.
We require a homomorphism where (1, 1) ker ϕ. An obvious candidate is
ϕ : Z ×Z Z : (x, y) 7 x + y
Certainly (x, y) ker ϕ x + y = 0 y = x (x, y)
(1, 1)
. Moreover,
x = ϕ(x, 0) for any x Z, so this is a surjective homomorphism. We conclude that
Z ×Z
(1, 1)
=
Im ϕ = Z
2. We identify the factor group
G
H
=
Z ×Z
(4, 6)
.
It is tempting to try a homomorphism ψ(x, y) = 6x 4y or ψ(x, y) = 3x 2y. These certainly
satisfy ψ(4, 6) = 0, but they also have ψ(2, 3) = 0 which we don’t want!
As a hint for how to proceed, note that the element/coset (2, 3) + H
G
H
has order 2:
2( 2, 3) = (4, 6) H
We’d therefore like a homomorphism that maps (2, 3) to an element of order 2: the simple
function ξ : Z ×Z Z
2
: (x, y) 7 y x certainly does this! We try combining these functions:
ϕ : Z ×Z Z ×Z
2
: (x, y) 7 (3x 2y, y x)
Let’s compute the kernel:
ϕ(3x 2y, y x) = (0, 0) = 3x = 2y Z = (x, y) = (2m, 3m) : m Z
Substitute this into the second factor: y x = m = 0 Z
2
= m = 2n is even, from which
(x, y) = (4n, 6n). It is trivially verified that all such pairs lie in the kernel, whence ker ϕ =
(4, 6)
, as required. Finally,
ϕ(1, 1) = (1, 0), ϕ(2, 3) = (0, 1) = (u, v) = ϕ(u + 2v, u + 3v)
says that ϕ is surjective. We conclude that
G
H
=
Z ×Z
2
.
Exercises 6.2. Key concepts:
Canonical homomorphism γ : G
G
H
1
st
isomorphism theorem µ :
G
ker ϕ
=
Im ϕ
1. Let ϕ : Z
18
Z
12
be the homomorphism ϕ(x) = 10x.
(a) Find the kernel of and image of ϕ.
(b) List the elements of the factor group
Z
18
ker ϕ
.
(c) State an explicit isomorphism µ :
Z
18
ker ϕ
Im ϕ.
(d) To what basic group Z
n
is the factor group isomorphic?
70
2. Repeat the previous question for the homomorphism ϕ : Z Z
20
: x 7 8x.
3. For each function ϕ : Z ×Z Z, find the kernel and identify the factor group
Z ×Z
ker ϕ
.
(a) ϕ(x, y) = 3x + y (b) ϕ(x, y) = 2x 4y
4. (a) If a subgroup H of G = Z
15
×Z
3
has order 5, find its elements.
(b) Show that ϕ(x, y) = (x, y) is a homomorphism ϕ : G Z
3
×Z
3
with ker ϕ = H.
(c) What does the 1
st
isomorphism theorem tell us about the factor group
G
H
?
5. Suppose G is a finite group with normal subgroup H and that ϕ : G L is a homomorphism
with ker ϕ = H. Prove that (G : H)
|
L
|
with equality if and only if ϕ is surjective.
6. Consider the map ϕ : Z ×Z
12
Z
3
×Z
6
defined by
ϕ(x, y) =
2x + y, y
(a) Verify that ϕ is a well-defined homomorphism.
(b) Compute ker ϕ and identify the factor group
Z ×Z
12
ker ϕ
7. Let H =
(3, 1)
G = Z
9
×Z
3
. Find an explicit homomorphism ϕ : G Z
9
whose kernel is
H, and thus identify the factor group
G
H
.
(Hint: (x, y) H = {(0, 0), (3, 1), (6, 2)} . . .)
8. Consider H =
(3, 3)
G = Z
9
× Z
9
. Find a surjective homomorphism ϕ : G Z
3
× Z
9
whose kernel is H and hence prove that
G
H
=
Z
3
×Z
9
.
9. Let ϕ : S
1
S
1
: z 7 z
2
.
(a) Find the kernel of ϕ and describe the canonical homomorphism γ : S
1
S
1
ker ϕ
.
(b) What does the first isomorphism theorem say about the factor group
S
1
ker ϕ
.
(c) For each n, identify the factor group
S
1
U
n
, where U
n
is the group of n
th
roots of unity.
10. We identify
G
H
=
Z ×Z
(a, b)
. Suppose d = gcd(a, b) where a, b are not both zero.
(a) Let ψ(x, y) =
b
d
x
a
d
y be the homomorphism ψ : Z ×Z Z. Find ker ψ.
(b) Show that
G
H
contains an element of order d, namely (
a
d
,
b
d
) + H.
(c) B
´
ezout’s identity says κ, λ Z for which κa + λb = d. The function
ξ : Z ×Z Z
d
: (x, y) 7 κx + λy
plainly maps (
a
d
,
b
d
) to the generator 1 Z
d
. Compute ker ϕ for the combined map
ϕ : Z ×Z Z ×Z
d
: (x, y) 7
b
d
x
a
d
y, κx + λy
(d) Identify the factor group
Z ×Z
(a, b)
.
(e) Find an explicit isomorphism µ :
Z ×Z
(21, 35)
Z ×Z
7
.
71
6.3 Conjugation, Cycle Types, Centers and Automorphisms
In this section we consider an important type of homomorphism and some its consequences.
Definition 6.15. Let G be a group and x, y G. We say that y is conjugate to x if
g G such that y = gxg
1
If g G is fixed, then conjugation by g is the map c
g
: G G : x 7 gxg
1
.
We’ve met this notion before: recall that a subgroup H is normal if and only if c
g
(h) H for all g G
(Corollary 5.8). It should also be familiar from linear algebra, in the context of similarity. Recall that
square matrices A, B are similar if B = MAM
1
for some invertible M. Such matrices have the same
eigenvalues and, essentially, ‘do the same thing’ with respect to different bases. An explicit group
theory analogue of this is Theorem 6.19 below.
Lemma 6.16. Conjugation by g is a isomorphism c
g
: G
=
G.
Proof. Conjugation by g
1
is the inverse function of c
1
g
:
c
g
1
c
g
(x)
= g
1
gxg
1
(g
1
)
1
= x, etc.
We moreover have a homomorphism:
c
g
(xy) = g(xy)g
1
=
gxg
1
gyg
1
= c
g
(x)c
g
(y)
Lemma 6.17. Conjugacy is an equivalence relation (x y g G such that y = gxg
1
).
The proof is an exercise. The equivalence classes under conjugacy are termed conjugacy classes.
Examples 6.18. 1. If G is abelian then every conjugacy class contains only one element:
x y g G such that y = gxg
1
= xgg
1
= x
2. The smallest non-abelian group is S
3
has conjugacy classes
{e}, {(1 2), (1 3), (2 3)}, {(1 2 3), (1 3 2)}
This can be computed directly, but it follows immediately from. . .
Theorem 6.19. The conjugacy classes of S
n
are the cycle types: elements are conjugate if and only if
they have the same cycle-type.
If an element σ S
n
is written as a product of disjoint cycles, then its cycle-type is clear. For instance:
( 1 2 3)( 4 5) has the same cycle-type as (1 5 6)(2 3): we might call these 3,2-cycles.
( 1 2)(3 4) has cycle-type 2,2.
72
Before seeing a proof it is beneficial to try an example.
Example 6.20. If ρ = (2 4 3) and σ = (1 2)(3 4) in S
4
, then
ρσρ
1
= (2 4 3)(1 2)(3 4) (2 3 4) = (1 4)(2 3)
This plainly has the same cycle-type as σ. Moreover, it may be obtained by applying ρ to the entries
of σ!
ρσρ
1
= (1 4)(2 3) =
ρ(1) ρ(2)
ρ(3) ρ(4)
()
This tells us how to reverse the process: given 2,2-cycles σ = (1 2)(3 4) and
τ = (1 4)(2 3) simply place σ on top of τ in a table to define a suitable
ρ = (2 4 3) for which ρσρ
1
= τ.
x 1 2 3 4
ρ(x) 1 4 2 3
The proof is a notational horror, but it amounts to nothing more than the example done abstractly:
our goal is to recognize ρσρ
1
in exactly the form (), taking ρ of each entry in each cycle.
Proof. () First consider conjugation of a k-cycle σ = (a
1
···a
k
) by ρ S
n
. We claim that
ρσρ
1
=
ρ(a
1
) ···ρ(a
k
)
is also a k-cycle. Since this is an equality of functions, we verify by evaluating on all elements
of X = {1, 2, . . . , n}. There are two cases, which correspond to partitioning X = A (X \ A)
where A = {a
1
, . . . , a
k
}. First observe that ρ bijectively maps A A and X \ A X \ A.
1. Suppose x A. Since A = ρ(A), we may uniquely write x = ρ(a
j
). But then
ρσρ
1
(x) = ρσρ
1
ρ(a
j
)
= ρσ(a
j
) = ρ(a
j+1
)
where a
k+1
is understood to equal a
1
.
2. Suppose x A. We know that ρ
1
(x) A is unmoved by σ, whence
ρσρ
1
(x) = ρσ(ρ
1
(x)) = ρρ
1
(x) = x
We conclude that ρσρ
1
=
ρ(a
1
) ···ρ(a
k
)
is as claimed. More generally, if σ = σ
1
···σ
l
is a
product of disjoint cycles, then
ρσρ
1
= (ρσ
1
ρ
1
)(ρσ
2
ρ
1
) ···( ρσ
l
ρ
1
)
has the same cycle-type as σ.
() Suppose σ = σ
1
···σ
l
and τ = τ
1
···τ
l
S
n
have the same cycle-type, written so that corre-
sponding orbits have the same length. Assume we’ve all single-element orbits are included so
that
S
σ
i
= {1, . . . , n} =
S
τ
i
. Define ρ S
n
by writing the orbits of σ and τ on top each other
x σ
1
σ
2
··· σ
l
ρ(x) τ
1
τ
2
··· τ
l
Since the orbits of σ, & τ partition X = {1, . . . , n}, this is plainly defines a permutation. Finally,
if s
i,j
and t
i,j
are the j
th
elements of the orbits σ
i
and τ
i
, then
ρσρ
1
(t
i,j
) = ρσ(s
i,j
) = ρ
s
i,j+1
= t
i,j+1
= τ
t
i,j
We conclude that ρσρ
1
= τ, as required.
73
Examples 6.21. 1. The permutations σ = (1 4 5)(2 7 6) and τ = (1 6 5)(3 4 7) in S
7
have the same
cycle-type and are thus conjugate. The table defines a suitable ρ.
x 1 4 5 2 7 6 3
ρ(x) 1 6 5 3 4 7 2
= ρ = (2 3)( 4 6 7)
It is worth making a sanity check:
ρσρ
1
= (2 3)(4 6 7)(1 4 5)(2 7 6)(2 3) (4 7 6) = (1 6 5)(3 4 7) = τ
Other choices of ρ are available: just write the orbits of σ, τ in different orders. Can you figure
out how many distinct choice of ρ will work?
2. (Example 5.3.2) We checked previously that V =
e, (1 2)(3 4), (1 3)(2 4), (1 4)(2 3)
is a normal
subgroup of A
4
. Since V contains the identity and all 2,2-cycles it is closed under conjugacy and
thus a normal subgroup of both A
4
and S
4
.
Automorphisms
We’ve already seen that conjugation c
g
: G G by a fixed element g G is an isomorphism. We
now consider all such maps.
Definition 6.22. An automorphism of a group G is an isomorphism of G with itself. The set of such is
denoted Aut G. The inner automorphisms are the conjugations
Inn G = {c
g
: G G where c
g
(x) = gxg
1
}
Example 6.23. There are four homomorphisms ϕ
k
: Z
4
Z
4
(Corollary 6.9);
ϕ
0
(x) = 0, ϕ
1
(x) = x, ϕ
2
(x) = 2x, ϕ
3
(x) = 3x
of which two are automorphisms: Aut Z
4
= {ϕ
1
, ϕ
3
}. Observe that ϕ
1
is the identity function and that
ϕ
3
ϕ
3
= ϕ
1
. The automorphisms therefore comprise a group (isomorphic to Z
2
) under composition
of functions.
As for conjugations, since Z
4
is abelian these are uninteresting:
g, x Z
4
= c
g
(x) = g + x + (g) = x = c
g
= ϕ
1
There is only one inner automorphism of Z
4
, namely the identity function ϕ
1
. This indeed holds in
general: if G is abelian, then Inn(G)
=
Z
1
is trivial.
In general, hunting for automorphisms is difficult. Here is a simple observation that helps to narrow
things down. The proof is an exercise.
Lemma 6.24. If ϕ Aut G and x G, then the orders of x and ϕ(x) are identical.
This helps to streamline the previous example: ϕ(1) must have the same order (four) as 1 and so our
only possibilities are ϕ(1) = 1 or ϕ(1) = 3. These generate the two observed automorphisms.
74
Example 6.25. We describe all automorphisms ϕ of S
3
. Consider σ = (1 2) and τ = (1 2 3). Since the
order of an element is preserved by ϕ, we conclude that
ϕ(e) = e, ϕ(σ)
(1 2), (1 3), (2 3)
, ϕ(τ)
(1 2 3), (1 3 2)
We therefore have a maximum of six possible automorphisms; it is tedious to check, but all in fact
define automorphisms! Indeed all are conjugations, from which Aut S
3
= Inn S
3
. Here is the data;
verify some of it for yourself:
element g c
g
(e) c
g
(1 2) c
g
(1 3) c
g
(2 3) c
g
(1 2 3) c
g
(1 3 2)
e e (1 2) (1 3) (2 3) (1 2 3) (1 3 2)
(1 2) e (1 2) (2 3) (1 3) (1 3 2) (1 2 3)
(1 3) e (2 3) (1 3) (1 2) (1 3 2) (1 2 3)
(2 3) e (1 3) (1 2) (2 3) (1 3 2) (1 2 3)
(1 2 3) e (2 3) (1 2) (1 3) (1 2 3) (1 3 2)
(1 3 2) e (1 3) (2 3) (1 2) (1 2 3) (1 3 2)
As the next result shows, the automorphisms always form a group under composition; in this case
Aut S
3
is a group of order 6 which is easily seen to be non-abelian,
c
(1 2)
c
(1 3)
= c
(1 3 2)
= c
(1 2 3)
= c
(1 3)
c
(1 2)
By process of elimination, we conclude that Aut S
3
=
S
3
.
Theorem 6.26. Aut G and Inn G are groups under composition. Moreover Inn G Aut G.
Proof. That Aut G is a group is simply the fact that composition and inverses of isomorphisms are
isomorphisms: you should already have made this argument when answering Exercise 2.5.15. By
Lemma 6.16, every conjugation is an isomorphism, and it is simple to check that c
g
c
h
= c
gh
and
c
1
g
= c
g
1
. We conclude that Inn G Aut G.
For normality, we check that Inn G is itself closed under conjugation! Let τ Aut G and c
g
Inn G.
For any x G, we have
25
(τc
g
τ
1
)(x) = τ
c
g
τ
1
(x)
= τ
g
τ
1
(x)
g
1
(definition of c
g
)
=
τ(g)
τ
τ
1
(x)
τ(g
1
)
(since τ is a homomorphism)
=
τ(g)
x
τ(g)
1
(again since τ is an homomorphism)
= c
τ(g)
(x)
We conclude that τc
g
τ
1
= c
τ(g)
Inn G, from which Inn G Aut G.
25
The challenge in reading the proof is keeping track of where everything lives. To help, the inverse symbol is colored:
τ
1
(in red) means the inverse function, whereas g
1
(in blue) means the inverse of an element in G.
75
Centers
We say that an element g G commutes with another element x G if the order of multiplication
is irrelevant: gx = xg. Otherwise said, g conjugates x to itself: c
g
(x) = x. It natural to ask whether
there are any elements which commute with all others. There are two very simple cases:
If G is abelian, then every element commutes with everything!
The identity e commutes with everything, regardless of G.
In general, the set of such elements falls somewhere between these extremes. This subset will turn
out to be yet another normal subgroup of G.
Definition 6.27. The center of a group G is the subset of G which commutes with everything in G:
Z(G) := {g G : h G, gh = hg}
We will prove that Z(G) G shortly. First we give a few examples; unless G is abelian, the center is
typically difficult to compute, so we omit more of the details.
Examples 6.28. 1. Z(G) = G G is abelian.
2. Z(S
n
) = {e} if n 3. This is Exercise 4.1.12.
3. Z(D
2n+1
) = {e} and Z(D
2n
) = {e, ρ
n/2
}, where ρ
n/2
is rotation by 180°. Part of this is in
Exercise 12: more generally, this follows from Exercise 4.1.13.
4. Z
GL
n
(R )
= {λI
n
: λ R
×
}. If you’ve done enough linear algebra to be familiar with
eigenvectors, an argument is reasonably straightforward (Exercise 11)
Theorem 6.29. Let G be any group. Then:
1. Z(G) G 2.
G
Z(G)
=
Inn G
Proof. Everything follows from defining a suitable homomorphism with the correct kernel!
Define ϕ : G Inn G by ϕ(g) = c
g
. This is indeed a homomorphism (ϕ(gh) = ϕ(g)ϕ(h)):
ϕ(gh)
(x) = c
gh
(x) = (gh)x(gh)
1
= g(hxh
1
)g
1
= c
g
(c
h
(x))
=
ϕ(g)ϕ(h)
(x)
Now observe that
g ker ϕ x G, c
g
(x) = gxg
1
= x
g Z(G)
By Lemma 6.4, ker ϕ = Z(G) is a normal subgroup of G.
To finish, observe that ϕ is surjective (definition of Inn G). The 1
st
isomorphism theorem tells us that
G
Z(G)
=
Im ϕ = Inn G
76
Exercises 6.3. Key concepts: Conjugation Conjugacy Classes Cycle Types are Conjugacy Classes (S
n
)
(Inner) Automorphism Center of a Group
1. Either find some ρ G such that ρσρ
1
= τ, or explain why no such element exists:
(a) σ = (1 2 3), τ = (1 3 2) where G = S
3
.
(b) σ = (1 4 5 6)(2 3)(5 6), τ = (1 2 3 4)(5 6)(2 6) where G = S
6
.
(c) σ = (1 4 5 6)(2 3)(5 6), τ = (1 2)(3 5 6) where G = S
6
.
2. Recall Example 6.21.1. Find another element ν = ρ for which νσν
1
= τ. Now determine how
many distinct such ν exist: that is, how many ways can we conjugate σ to get τ?
3. Prove Lemma 6.17. Prove that the relation
x y y is conjugate to x
is an equivalence relation on any group G.
4. (a) Suppose y is conjugate to x in a group G. Prove that the orders of x and y are identical.
(b) Show that the converse to part (a) is false by exhibiting two non-conjugate elements of the
same order in some group.
5. Let H G, fix a G and define the conjugate subgroup K = c
a
(H) = {aha
1
: h H}.
(a) Prove that K is indeed a subgroup of G.
(b) Prove that the function ψ : H K : h 7 aha
1
is an isomorphism of groups.
(c) If H G, what can you say about c
a
(H)?
(d) Let H = {e, (1 2)} S
3
and a = (1 2 3). Compute the conjugate subgroup K = c
a
(H).
6. In Example 6.21.2 we saw that V = {e, (1 2)(3 4), (1 3)(2 4), (1 4)(2 3)} S
4
.
(a) Show that normal subgroup is not transitive by giving an example of a normal subgroup
K V which is not normal in S
4
.
(b) How many other subgroups does S
4
have which are isomorphic to V? Why are none of
them normal in S
4
?
(c) Explain why
S
4
V
is a group of order six. Prove that
(1 2)V(1 3)V = (1 3)V( 1 2) V
Hence conclude that
S
4
V
=
S
3
.
(d) Why is it obvious that the following six left cosets are distinct.
V, (1 2)V, (1 3)V, (2 3)V, (1 2 3)V, (1 3 2)V
(Hint: Think about how none of the representatives a of the above cosets move the number 4 and
consider aV = bV b
1
a V . . .)
(e) Define an isomorphism µ :
S
4
V
S
3
and prove that it is an isomorphism.
77
7. Prove Lemma 6.24: if ϕ Aut G and x G, then ϕ(x) has the same order as x.
8. Describe all automorphisms of the Klein four-group V.
(Hint: use the previous question!)
9. Recall Exercise 6.1.7. Explain why Aut Z
n
=
Z
×
n
.
(Hint: consider ϕ
k
(x) = kx where gcd(k, n) = 1 and map ψ : k 7 ϕ
k
)
10. Let G be a group. Prove directly that Z(G) G, without using Theorem 6.29. That is:
(a) Prove that Z(G) is closed under the group operation and inverses.
(b) Prove that gZ(G) = Z(G)g for all g G.
11. We identify the center of the general linear group.
The n ×n matrix A =
0 1 0 ··· ··· 0
0 0 1 0
0 0 0 0
.
.
.
.
.
.
.
.
.
0 0 0 0 1
0 0 0 ··· 0
has a single one-dimensional eigenspace: Ae
1
= 0.
(a) Let B Z
GL
n
(R )
. Use the fact that AB = BA to prove that Be
1
= λe
1
for some λ = 0.
(b) Let x R
n
be non-zero and X an invertible matrix for which Xe
1
= x (i.e. x is the 1
st
column of X). Prove that Bx = λx.
(c) Since the observation in part (b) holds for any x R
n
, what can we conclude about B?
What is the group Z
GL
n
(R )
?
12. (a) Prove that D
4
has center Z(D
4
) = {e, ρ
2
}, where ρ
2
is rotation by 180
.
(b) State the cosets of Z(D
4
). What is the order of each? Determine whether
D
4
Z(D
4
)
is
isomorphic to Z
4
or to the Klein four-group V.
(c) (Hard) Can you find a homomorphism ϕ : D
4
D
4
whose kernel is Z(D
4
)?
(Hint: draw a picture and think about doubling angles of rotation and reflection!)
13. The centralizer of an element x G is the set of elements which conjugate x to itself:
C(x) =
g G : gxg
1
= x
(a) Prove directly that C(x) is a subgroup of G.
(b) We compute a few centralizers directly. The centralizer of σ = (1 2 3) S
3
is the set of all
ρ for which
(1 2 3) = ρσρ
1
=
ρ(1) ρ(2) ρ(3)
Otherwise said, ρ must permute the elements {1, 2, 3} in order. We conclude that
C(σ) =
e, (1 2 3), (1 3 2)
= {e, σ, σ
2
}
=
Z
3
In similar fashion, compute the centralizers of the following elements and identify the
group C(σ):
i. σ = (1 2) S
4
ii. σ = (1 2 3 4 5) in both S
5
and A
5
.
iii. σ = (1 2)(3 4) in both S
4
and A
4
.
14. Prove that if
G
Z(G)
is cyclic, then G is abelian. What is the group
G
Z(G)
in such a situation?
78
7 Group Actions
7.1 Group Actions, Fixed Sets and Isotropy Subgroups
You may feel by now that groups are worthy of study purely in their own right—if so great! For
many, however, the fundamental reason to care about groups is because of how they transform sets.
Recall how the symmetric group S
n
(Section 4) was defined in terms of what its elements do to the
set {1, . . . , n}. This is an example of a general situation.
Definition 7.1. An action
26
of a group G on a set X is a function · : G × X X for which,
(a) x X, e · x = x, and,
(b) x X, g, h G, g ·(h · x) = (gh) · x.
Part (b) says g 7 g· is a homomorphism of binary structures (G, ·)
{f : X X},
.
Examples 7.2. 1. The symmetric S
n
group acts on X = {1, 2, . . . , n}. As a sanity check:
(a) e(x) = x for all x {1, . . . , n}.
(b) σ
τ(x)
= (στ)(x) is simply composition of functions!
2. If X is the set of orientations of a regular n-gon centered at the origin and with one vertex at
(1, 0), then D
n
acts on X by rotations and reflections. This is essentially our definition of D
n
!
3. Any group G acts on itself by left multiplication (X = G). This is essentially the proof of
Cayley’s Theorem (4.8). G also acts on itself by conjugation (c
g
c
h
= c
gh
is Theorem 6.26).
4. Matrix groups act on vector spaces by matrix multiplication. For example the orthogonal group
O
2
(R ) transforms vectors via rotations and reflections:
O
2
(R ) ×R
2
R
2
: (A, v) 7 Av
5. A group can act on many different sets. Here are three further actions of the orthogonal group:
i. O
2
(R ) acts on X = {1, 1} via A ·x := (det A)x.
ii. O
2
(R ) acts on X = R
3
via A ·v := A(v
1
i + v
2
j) + v
3
k.
iii. O
2
(R ) acts on the unit circle X = S
1
R
2
via matrix multiplication A ·v := Av.
We often use actions to help visualize a group (or even define it!); in this context, some actions are
better than others. Consider the three actions of O
2
(R ) in part 5 above.
i. X = {1, 1} is too small to provide a detailed picture of the group since many matrices act in
the same way: e.g.
1 0
0 1
and
1 0
0 1
both “multiply by 1 (= det A)”.
ii. X = R
3
is unnecessarily large. The action leaves any vertical vector untouched.
iii. X = S
1
is large enough that the action of distinct matrices can be distinguished without being
inefficiently large (a Goldilocks action, perhaps?).
26
This is strictly a left action. There is an analogous definition of a right action. In these notes, all actions will be left.
79
These notions can be formalized.
Definition 7.3. Let G × X X be an action.
1. The fixed set of g G is the set (subset of X)
Fix(g) := {x X : g ·x = x} (also written X
g
)
2. The isotropy subgroup or stabilizer of x X is the set (subset of G)
Stab(x) := {g G : g ·x = x} (also written G
x
)
3. The action is faithful (or effective) if the only element of G fixing everything is e. Equivalently:
(a) Fix(g) = X g = e (b)
T
xX
Stab(x) = {e}
4. The action is transitive if any element of X may be transformed to any other:
x, y X, g G such that y = g · x
Very loosely, an action that is both faithful and transitive is likely reasonable for visualizing a group.
Examples (7.2 cont). 1. The action of S
n
on {1, 2, . . . , n} is both faithful and transitive:
Faithful: if σ(x) = x for all x {1, 2, . . . , n}, then σ = e.
Transitive: if x = y, then the 2-cycle (x y) maps x 7 y.
2. D
n
acts faithfully and transitively on the orientations of the n-gon.
3. The action of a group on itself by left multiplication is both faithful and transitive. Conjugation
is more complex: in this context, the stabilizer of x G is called its centralizer C(x) = Stab(x).
4. The action of O
2
(R ) on R
2
is faithful but not transitive: for instance the zero vector cannot be
transformed into any other vector so Stab(0) = O
2
(R ).
5. We leave these as exercises.
Lemma 7.4. For each x X, the stabilizer Stab(x) is indeed a subgroup of G.
Proof. We use the subgroup criterion:
Non-emptiness Part (a) of Definition 7.1 says that e Stab(x).
Closure This is part (b) of the Definition. Let g, h Stab(x), then
(gh) · x = g ·(h ·x) = g ·x = x = gh Stab(x)
Closure This relies on both parts of the Definition. If g Stab(x), then
x = g · x = g
1
· x = g
1
·(g · x) = (g
1
g) · x = e · x = x
= g
1
Stab(x)
80
Example 7.5. The dihedral group D
3
= {e, ρ
1
, ρ
2
, µ
1
, µ
2
, µ
3
} acts on the set X of vertices of an equi-
lateral triangle.
27
The fixed sets and stabilizers for this action are as follows:
Element g Fix(g)
e {1, 2, 3}
ρ
1
ρ
2
µ
1
{1}
µ
2
{2}
µ
3
{3}
Vertex x Stab(x)
1 {e, µ
1
}
2 {e, µ
2
}
3 {e, µ
3
}
1
2
3
D
3
also acts on the set of edges of the triangle E =
{1, 2}, {1, 3}, {2, 3}
. You needn’t write all these
out since stabilizing an edge is equivalent to stabilizing its opposite vertex. Still, here is the data:
Element g
Fix(g)
e E
ρ
1
ρ
2
µ
1
{2, 3}
µ
2
{1, 3}
µ
3
{1, 2}
Edge {x, y} Stab({x, y})
{1, 2} {e, µ
3
}
{1, 3} {e, µ
2
}
{2, 3} {e, µ
1
}
Exercises 7.1. Key concepts: (left) action Fix(g) Stab(x) G faithful & transitive actions
1. For part 5 of Example 7.2, determine whether each action is faithful and/or transitive.
2. Consider the cyclic subgroup G =
σ
of S
6
generated by the 6-cycle σ = (1 2 3 4 5 6).
(a) State the fixed sets and stabilizers for the natural action of G on the set X = {1, 2, 3, 4, 5, 6}.
(b) Is the action of G faithful? Transitive?
3. Repeat the previous question when σ = (1 3)(2 4 6).
4. Mimic Example 7.5 for the actions of D
4
on X = {vertices} and E = {edges} of the square.
(Use whatever notation you like; ρ, µ, δ or cycle notation)
5. Prove that left multiplication of G on itself g ·x = gx is:
(a) Free: x G, g · x = x = g = e (this is stronger than faithfulness [ versus ]).
(b) Transitive.
6. Suppose that G acts on itself (X = G) by conjugation g · x = gxg
1
.
(a) Prove that conjugation is faithful if and only if the center is trivial: Z(G) = {e}.
(b) If G is abelian, is conjugation faithful? Transitive? Explain.
(c) If G = S
n
, is conjugation faithful? Transitive? Explain.
7. Suppose G has a left action on X. Prove that G acts faithfully on X if and only if no two distinct
elements of G have the same action on every element.
27
Recall that ρ
1
rotates 120° counter-clockwise, that ρ
2
= ρ
2
1
and that µ
i
reflects across the altitude through vertex i.
81
7.2 Orbits & Burnside’s Formula
We first met orbits in the context of the symmetric groups S
n
. The same idea applies to any action.
Definition 7.6. Let G × X X be an action. The orbit of x X under G is the set of elements into
which x may be transformed:
Gx = {g · x : g G} X (also written G · x)
Examples 7.7. 1. If X = {1, 2, . . . n} and G =
σ
S
n
, then
Gx = {σ
k
(x) : k Z} = orb
x
(σ)
The definition of orbits therefore coincides with that in Section 4.2.
2. A transitive action
28
has only one orbit.
3. If O
2
(R ) acts on R
2
by matrix multiplication, then the orbits are circles centered at the origin!
Lemma 7.8. The orbits of an action partition X.
We omit the proof: compare the special case where S
n
acts on X = {1, . . . , n} (Lemma 4.12).
Our next result is analogous to Lemma 6.6 where we counted the number of (left) cosets of ker ϕ.
Lemma 7.9. The cardinality of the orbit Gx is the index of the isotropy subgroup Stab(x):
|
Gx
|
=
G : Stab(x)
For a finite group
|
G
|
, the size of the orbit necessarily divides the order of the group
|
G
|
.
Proof. Observe that
g · x = h ·x h
1
g · x = x h
1
g Stab(x)
g Stab(x) = h Stab(x)
Otherwise said (contrapositive) distinct elements of Gx correspond to distinct left cosets.
Examples 7.10. 1. Let σ = (1 4)(2 7 3) S
7
. Consider X = {1, 2, 3, 4, 5, 6, 7} under the action of
the cyclic group G =
σ
. The orbits are precisely the disjoint cycles: {1, 4}, {2, 3, 7}, {5}, {6}.
Observe that G has six elements:
e, σ = (1 4)(2 7 3), σ
2
= (2 3 7), σ
3
= (1 4), σ
4
= (2 7 3), σ
5
= (1 4)(2 3 7)
The Lemma is easily verifiable: for instance,
Stab( 3) = {τ G : τ(3) = 3} = {σ
k
: σ
k
(3) = 3} = {e, σ
3
}
=
G : Stab(3)
=
6
2
= 3 =
|
{2, 3, 7}
|
=
|
G3
|
28
We now have two meanings of transitive; one for equivalence relations and one for actions. Be careful!
82
2. In Exercise 6.3.13 we computed several centralizers (stabilizers under conjugation). Notice
what the Lemma says about the size of the orbit of the 3-cycle σ = (1 2 3) under conjugation:
|
S
3
(σ)
|
=
|
S
3
|
|
C(σ)
|
=
3!
3
= 2
As it should be, this is precisely the size of the conjugacy class of 2-cycles!
Burnside’s Formula
It can be useful to count the number of orbits of an action. For finite actions, this can be done in two
different ways, which leads to an interesting formula. We start by observing that
S =
(g, x) G ×X : g · x = x
has cardinality
|
S
|
=
xX
|
Stab(x)
|
=
gG
|
Fix(g)
|
We now count the elements of S in a different way. Since (G : Stab(x)) =
|
G
|
|
Stab(x)
|
, we see that
|
S
|
|
G
|
=
xX
|
Stab(x)
|
|
G
|
=
xX
1
(G : Stab(x))
=
xX
1
|
Gx
|
()
where we used Lemma 7.9 for the last equality. Consider a fixed orbit Gy. Since
|
Gx
|
=
|
Gy
|
is
constant for each x Gy, we conclude that
xGy
1
|
Gx
|
=
|
Gy
|
|
Gy
|
= 1
The summation () therefore counts 1 for each distinct orbit. In concludsion:
Theorem 7.11 (Burnside’s formula). Let G be a finite group acting on a finite set X. Then
# orbits =
1
|
G
|
xX
|
Stab(x)
|
=
1
|
G
|
gG
|
Fix(g)
|
Example (7.10.1 cont). Here is the data when G =
σ
=
(1 4)(2 7 3)
acts on X = {1, 2, 3, 4, 5, 6, 7}:
x X Stab(x)
1, 4 {e, σ
2
, σ
4
}
2, 3, 7 {e, σ
3
}
5, 6 G = {e, σ, σ
2
, σ
3
, σ
4
, σ
5
}
g G Fix(g)
e X = {1, 2, 3, 4, 5, 6, 7}
σ, σ
5
{5, 6}
σ
2
, σ
4
{1, 4, 5, 6}
σ
3
{2, 3, 5, 6, 7}
Burnside’s formula merely sums the cardinalities of all the subsets in the right column of each table:
4 = # orbits =
1
|
G
|
xX
|
Stab(x)
|
=
1
6
(3 + 2 + 2 + 3 + 6 + 6 + 2)
=
1
|
G
|
gG
|
Fix(g)
|
=
1
6
(7 + 2 + 4 + 5 + 4 + 2)
83
One reason to count the number of orbits of an action is that we often want to consider objects as
equivalent if they differ by the action of some group.
Example 7.12. A child’s toy consists of a wooden equilateral triangle whose edges are painted using
any choice of colors of the rainbow. How many distinct toys could we create?
This might feel like an imprecisely-posed problem, but think for a moment like a child: wouldn’t
they likely consider the two pictured triangles below to be the same? The language of group actions
can help make this precise.
If we orient each triangle the same way, then a single toy may be considered as an element of
X = {ordered color triples}. Since there are 7 choices for the color of each edge, we see that
|
X
|
= 7
3
= 343 (a large set!).
Two toys are equivalent if they differ only by a ro-
tation in 3-dimensions. This amounts to the natural
action of D
3
on X: for instance, taking ρ
1
to rotate
120° counter-clockwise,
ρ
1
·(red,green,violet) = (violet,red,green)
ρ
1
The number of distinct toys is therefore the number of orbits of D
3
on X, which we may compute
using Burnside. Since it would be time consuming to find the stabilizer of each element of X, we use
the fixed set approach.
Identity e: For any action Fix(e) = X.
Rotations ρ
1
, ρ
2
: If a color-scheme is fixed by ρ
j
, then all pairs of adjacent edges must be the same
color. Since there are seven colors in the rainbow, we see that
|
Fix(ρ
i
)
|
= 7.
Reflections µ
1
, µ
2
, µ
3
: Since µ
j
swaps two edges, any color-scheme in its fixed set must have these
edges the same color. We have 7 choices for the color of the switched edges, and an independent
choice of 7 colors for the other edge. It follows that
Fix(µ
j
)
= 7
2
= 49.
The number of distinct toys is therefore
# orbits =
1
|
D
3
|
σD
3
|
Fix(σ)
|
=
1
6
(7
3
+ 7 + 7
|{z}
rotations
+ 7
2
+ 7
2
+ 7
2
| {z }
reflections
)
=
7
6
(49 + 1 + 1 + 7 + 7 + 7) = 84
For simpler version of the problem, consider the situation where all sides must be different colors. In
this case D
3
acts on a set of color-schemes with cardinality
|
Y
|
= 7 · 6 ·5 = 210. Moreover, only the
identity element has a non-empty fixed set. The number of distinct three-colored toys is therefore
# orbits =
1
|
D
3
|
σD
3
|
Fix(g)
|
=
1
6
(210 + 0 + ··· + 0) =
210
6
= 35
Of course you are welcome to answer questions like these using pure combinatorics without any
resort to group theory!
84
Example 7.13 (Dice-rolling for Geeks). Various games (such as Dungeons & Dragons) make use of
polyhedral dice beyond merely the cube (see, for instance, Exercise 2.4.12).
Since dice are for rolling, we consider two such to be equivalent if one can be rotated into the other. It
shouldn’t be hard to convince yourself that the pictured tetrahedral dice are distinct (non-equivalent):
the first cannot be rotated to make the second.
Indeed, it is not difficult to see that these are the only tetrahe-
dral dice: if you place 4 on the table, then the remaining faces
must be numbered 1, 2, 3 either counter-clockwise or clockwise.
For larger dice, such a counting approach is impractical.
However, with a little thinking about symmetry groups,
Burnside’s formula will ride to the rescue.
Suppose a regular polyhedron has f faces, each with n sides.
The faces may be labelled 1 thorough f in f ! distinct ways. Denote the set of distinct labellings
by X.
We may rotate the polyhedron so that any face is mapped to any other, in any orientation. For
instance, the face labelled 1 may be rotated to any of the f faces, before rotating the polyhedron
around that face in any of n orientations. The rotation group G of the polyhedron therefore has
f n elements.
Each non-identity element of the rotation group G moves at least one face, whence
|
Fix(g)
|
=
(
X if g = e
if g = e
By Burnside’s formula, the number of distinct dice for a regular polyhedron is therefore
# orbits =
1
|
G
|
|
Fix(e)
|
=
|
X
|
|
G
|
=
f !
f n
=
( f 1)!
n
Here is the complete data for all the regular platonic solids.
Polyhedron f n Rotation Group # distinct dice (orbits)
Tetrahedron 4 3 A
4
2
Cube 6 4 S
4
30
Octahedron 8 3 S
4
1,680
Dodecahedron 12 5 A
5
7,983,360
Icosahedron 20 3 A
5
40,548,366,802,944,000
Note that we don’t need an explicit description of the rotation group, only its order
|
G
|
. We saw
that the rotation group of regular tetrahedron was A
4
in Example 4.26.3. Proving that the remaining
groups are as claimed is somewhat trickier. . .
85
Exercises 7.2. Key concepts: Orbits partition X
|
Gx
|
= (G : Stab(x)) Burnside’s formula
1. Determine the orbits of G =
σ
on X = {1, 2, 3, 4, 5, 6} for each of Exercises 7.1.2 and 3. In both
cases verify Burnside’s formula.
2. Revisit Example 7.12. How may distinct toys may be created if:
(a) A maximum of two colors can be used?
(b) Exactly two colors must be used?
3. Prove Lemma 7.8: the orbits of a (left) action partition X.
4. Revisit Exercise 6.3.13 and Example 7.10.2. Prove that there are two distinct conjugacy classes
of 5-cycles in A
5
. Provide an example of two 5-cycles which are not conjugate in A
5
.
5. A 10-sided die (see Exercise 2.4.12) is shaped so that all faces are congruent kites: five faces are
arranged around the north pole and five around the south.
(a) Argue that the group of rotational symmetries of such a die has ten elements, and that it
is isomorphic to D
5
.
(b) Use Burnside’s formula to determine how many distinct 10-sided dice (faces numbered 1
to 10) may be produced.
6. A soccer ball is constructed from 20 regular hexagons and 12 regular
pentagons as in the picture.
Suppose the 20 hexagonal patches are all to have different colors, as are
the 12 pentagonal patches. How many distinct balls may be produced?
7. The faces of a cuboid measuring 1 ×1 ×2 in are to be painted using (at
most) two colors. Up to equivalence by rotations, how many ways can
this be done?
8. Repeat the previous question for a regular tetrahedron.
9. (Hard) A cube has four diagonals: a, b, c, d, which are permuted by
its rotation group. For instance, we might rotate by 180° around
an axis through the midpoints of the sides a
1
b
1
and a
2
b
2
. This
switches diagonals a and b, but leaves c and d alone and therefore
acts as the 2-cycle (a b) on {a, b, c, d}. As a function on vertices,
(a b) :
a
1
b
1
c
1
d
1
a
2
b
2
c
2
d
2
7
b
1
a
1
c
2
d
2
b
2
a
2
c
1
d
1
(a) Describe as best you can how to rotate the cube in a way that corresponds to each of the
following rotations (written in cycle notation):
(a b c), (a b c d), (a b) (c d)
Hence argue that the rotation group of the cube is isomorphic to S
4
.
(b) A cube can operate as a “3-sided” die by labelling two each of its faces using the numbers
1, 2, 3. In how many distinct ways can this be done?
86
7.3 The Class Equation, p-groups, and the Theorems of Cauchy and Sylow
The toolkit provided by group actions is central to an enormous topic: the classification of groups
and their internal structure. This final section offers a small taste of this discussion.
Suppose G acts on a finite set X. Denote the set of 1-element orbits by
X
G
=
\
gG
Fix(g) = {x X : g G, g ·x = x}
Let x
1
, . . . , x
r
be representatives of the remaining (larger) orbits. Since the orbits partition X,
|
X
|
=
|
X
G
|
+
r
j=1
Gx
j
=
|
X
G
|
+
r
j=1
G : Stab(x
j
)
=
|
X
G
|
+
r
j=1
|
G
|
Stab(x
j
)
()
We focus on the special case when G acts on itself (X = G) by conjugation g · x = gxg
1
.
The 1-element orbits comprise the group center: X
G
= Z(G).
In this context the stabilizer of an element x G is known as its centralizer:
C(x) = Stab(x) = {g G : gxg
1
= x} = {g G : gx = xg}
Equation () is called the class equation:
|
G
|
=
|
Z(G)
|
+
r
j=1
conjugacy class of x
j
=
|
Z(G)
|
+
r
j=1
|
G
|
C(x
j
)
Example 7.14. The conjugacy classes in S
4
are the cycle-types, so the class equation is easily verified:
24 =
|
{e}
|
+ #
2-cycles
+ #
3-cycles
+ #
4-cycles
+ #
2,2-cycles
= 1 + 6 + 8 + 6 + 3
Our next result applies the class equation to obtain a partial converse to Lagrange’s Theorem.
Theorem 7.15 (Cauchy). If a prime p divides
|
G
|
, then G contains a subgroup/element of order p.
Proof. 1. Exercise 9 supplies an inductive proof that abelian G have such subgroups.
2. If G is non-abelian, then the center Z(G) is a proper (abelian) subgroup.
Let x Z(G). Plainly C(x) is a proper subgroup of G. If p does not divide
|
C(x)
|
, then p divides
|
Gx
|
=
G : C(x)
=
|
G
|
|
C(x)
|
. If this holds for all non-trivial orbits, the class equation says that
|
Z(G)
|
is divisible by p. Either way, G has a proper subgroup H whose order is divisible by p.
If H is abelian, apply case 1. Otherwise, repeat starting with H to find an even smaller subgroup
whose order is divisible by p. If this process never reached an abelian subgroup, then we’d have
an infinite descending sequence of proper subgroups: contradiction.
The resulting subgroup is cyclic (Corollary 5.10), whence an element of order p also exists.
87
Example 7.16. If G has order 60 = 2
2
· 3 ·5, then it has subgroups of orders 2, 3 and 5. It also has
subgroups of other orders (at the very least 1, 60 and 4 this last by the 1
st
Sylow Theorem below).
Haven’t we done this already? Exercise 3.3.14 appears to cover abelian groups, but this depends on
the Fundamental Theorem (3.26), the standard proof of which in fact relies on Cauchy (Exercise 10)!
Definition 7.17. Let p be a prime. A group G is a p-group if all its elements have order a power of p.
Examples 7.18. 1. G = Z
8
is a 2-group.
2. G =
2
(
=
Z
9
) is a 3-subgroup of Z
18
(this last is not a 3-group).
3. Z
3
×
2
is a 3-subgroup of Z
3
×Z
6
.
4. V =
e, (1 2)(3 4), (1 3)(2 4), (1 4)(2 3)
is a 2-subgroup of S
4
.
The proofs of the next result are exercises.
Theorem 7.19. 1. A finite group G is a p-group if and only if
|
G
|
= p
n
for some n.
2. If G is a p-group, then its center Z(G) is non-trivial.
Corollary 7.20. Let p be prime. If G has order p
2
, then it is abelian.
Proof. By the Theorem, we know that
|
Z(G)
|
= p or p
2
. In the second case we are done: Z(G) = G
says that G is abelian. In the first case the factor group is cyclic:
G
Z(G)
= p =
G
Z(G)
=
Z
p
Exercise 6.3.14 says G is abelian, though really it’s a contradiction: Z(G) = G =
G
Z(G)
=
Z
1
.
Our final results go some way towards establishing the existence and number of certain subgroups.
We omit the proofs: a typical approach uses induction with Cauchy’s Theorem as the base case, and
the application of various ingenious group actions. If you are interested, look them up!
Theorem 7.21 (Sylow). Let G be finite and write
|
G
|
= p
n
m, where p is prime and gcd(p, m) = 1.
1. G contains a subgroup H
p
i
of order p
i
for each i = 1, . . . , n. Moreover, any such subgroup with
i < n is a normal subgroup of some subgroup of order p
i+1
:
{e} H
p
H
p
2
··· H
p
n
G (†)
2. Any two maximal p-subgroups H
p
n
(called Sylow p-subgroups) are conjugate.
3. The number of Sylow p-subgroups divides
|
G
|
and is congruent to 1 modulo p.
Only the last inclusion in () can be non-normal. For a given example, the 3
rd
Theorem might show
that there is a unique Sylow p-subgroup H
p
n
: by Exercise 6.3.5, such a subgroup is necessarily normal.
88
Examples 7.22. 1. Suppose
|
G
|
= 15 = 3 ·5. By the 1
st
Sylow Theorem, G has at least one Sylow
3-subgroup (isomorphic to Z
3
) and at least one Sylow 5-subgroup (isomorphic to Z
5
).
The divisors of 15 are listed, along with whether each is con-
gruent to 1 modulo 3 or 5. By the 2
nd
and 3
rd
Theorems, G has
exactly one subgroup isomorphic to Z
3
and one to Z
5
: being
self-conjugate, both subgroups are normal.
Divisor d 1 3 5 15
d 1 (mod 2)?
d 1 (mod 3)?
By Lagrange, the orders of elements in G can only be 1, 3, 5 or 15. Since an order 3 element
generates a subgroup isomorphic to Z
3
(the unique Sylow 3-subgroup), only two elements in
G have order 3. Similarly only four elements in G have order 5 (each being a generator of the
unique Sylow 5-subgroup). Since only the identity has order 1, the remaining 8 = 15 1 2 4
elements must have order 15, and thus generate G.
In conclusion: if
|
G
|
= 15, then G
=
Z
15
is cyclic. Exercise 5 extends this analysis to when
|
G
|
= pq, where p < q are primes for which q 1 is not divisible by p: in such a case G
=
Z
pq
.
2. Suppose
|
G
|
= 100 = 2
2
·5
2
. By the 1
st
Sylow Theorem, G has at least one Sylow 5-subgroup of
order 5
2
= 25, and at least one Sylow 2-subgroup of order 4 = 2
2
. We can be more precise by
applying the other results.
The full list of divisors of
|
G
|
= 100 is: 1, 2, 4, 5, 10, 20, 25, 50, 100.
The only divisor congruent to 1 modulo 5 is 1 itself! By the 3
rd
Theorem, there is precisely
one Sylow 5-subgroup H
25
of G. Since this is the only subgroup of order 25, it must be
normal: H
25
G. By Corollary 7.20, H
25
is abelian. The Fundamental Theorem says that
either H
25
=
Z
25
or H
25
=
Z
5
×Z
5
.
The divisors 1, 5 and 25 are congruent to 1 modulo 2, so there might be 1, 5 or 25 distinct
Sylow 2-subgroups of G.
We give several sub-examples where the Sylow p-subgroups can be seen explicitly.
(a) G = Z
100
has one Sylow 5-subgroup
4
=
Z
25
, and one Sylow 2-subgroup
25
=
Z
4
.
(b) G = Z
50
× Z
2
has one Sylow 5-subgroup
(2, 0)
=
Z
25
, and one Sylow 2-subgroup
25
×Z
2
=
Z
2
×Z
2
.
(c) G = Z
5
×Z
20
has one Sylow 5-subgroup Z
5
×
4
=
Z
5
×Z
5
, and one Sylow 2-subgroup
(0, 5)
=
Z
4
.
(d) G = D
50
has one Sylow 5-subgroup consisting of half the 50 rotations:
ρ
2
= {ρ
0
, ρ
2
, . . . , ρ
48
}
=
Z
25
There are 25 distinct Sylow 2-subgroups, each of which is isomorphic to the Klein 4-group
V. These may be described explicitly if we label the reflections µ
1
, . . . , µ
50
:
V
i
= {ρ
0
, ρ
25
, µ
i
, µ
i+25
}, i = 1, . . . , 25
Note that D
50
has no subgroups isomorphic to Z
4
(reflections have order 2 and rotations
have orders 1, 5 or 25). This is also clear from the 2
nd
Theorem: all Sylow 2-subgroups are
conjugate and thus isomorphic (to V).
89
Exercises 7.3. 1. State all the 3-subgroups of Z
36
.
2. Let p be a prime of your choice. In contrast to Corollary 7.20, give an example of a non-abelian
group of order p
3
.
3. Suppose
|
G
|
= 225 = 3
2
·5
2
. Prove that G has a normal subgroup of order 25.
4. Suppose
|
G
|
= 20 = 2
2
·5.
(a) Determine how many distinct Sylow 2- and Sylow 5-subgroups G can have. Must they be
normal subgroups?
(b) Similarly to Example 7.22.2, state the Sylow subgroups when
i. G = Z
20
, ii. G = D
10
5. Let G have order pq, where p < q are distinct primes ( 3). Then all proper subgroups of G are
cyclic (isomorphic to Z
p
, Z
q
or Z
1
).
(a) If pq = 2 ·3 = 6, then G
=
Z
6
or
=
S
3
. State all the Sylow 2- and 3-subgroups in each case.
(b) Returning to the general case, prove that there is a unique Sylow q-subgroup H
=
Z
q
(which is therefore normal) by showing that p 1 (mod q) is a contradiction.
(c) Suppose moreover that q 1 is not divisible by p. Prove that there is a unique Sylow p-
subgroup K
=
Z
p
(which is also normal). Extending the analysis of Example 7.22.1, prove
that G
=
Z
pq
.
(d) If
|
G
|
= 21 (p = 3, q = 7), show that there are either one or seven Sylow 3-subgroups. In
the former case, argue that G
=
Z
21
.
(The latter case [seven 3-subgroups] results in a new non-abelian group not seen in these notes.)
6. Prove Theorem 7.19, part 1.
(Hints: one direction uses Lagrange’s Theorem, the other Cauchy)
7. Suppose G is a p-group which acts on a finite set X.
(a) Prove that
|
X
|
|
X
G
|
(mod p).
(Hint: what can you say about
|
Gx
|
if x lies in a non-trivial orbit?)
(b) Prove part 2 or Theorem 7.19, that p divides the order of the center Z(G).
8. The alternating group A
5
has order 60 = 2
2
· 3 · 5 and comprises the identity, all 3-cycles, 5-
cycles, and 2,2-cycles in S
5
.
(a) Describe all the Sylow 2-, 3-, and 5-subgroups of A
5
and verify that the number of each is
in line with the 3
rd
Theorem.
(b) Extending Exercise 7.2.4, find the sizes of all conjugacy classes in A
5
. Hence prove that A
5
is a simple group: if N is a normal subgroup of A
5
, then N = {e} or A
5
.
(Simple groups are very important and have applications throughout mathematics. Their full
classification was completed in 2004—an enormous undertaking: look it up. . . )
90
9. We prove the abelian part of Cauchy’s Theorem by induction on the order of G.
(a) Explain why the base case
|
G
|
= 2 is true.
Fix n 3, let G be abelian of order n and assume p divides n. For the induction hypothesis,
assume that if K is any abelian group of order
|
K
|
< n where p divides
|
K
|
, then K has a
subgroup of order p.
Let x G be a non-identity element with order m =
|
x
|
(necessarily m 2).
Choose a prime q dividing m, define y := x
m/q
and let H :=
y
.
(b) What is the order of H? Explain why are we done if q = p.
(c) If q = p, use the induction hypothesis to explain why there exists a coset zH
G
H
of
order p. Now prove that z
q
has order p in G.
10. We sketch part of the proof of the Fundamental Theorem of Finite Abelian Groups (3.26).
Suppose G is abelian and that
|
G
|
= p
n
m where p does not divide m. Define
H = {x G : x
p
n
= e}, K = {x G : x
m
= e}
(a) Show that H is a p-group.
(b) Since gcd(m, p
n
) = 1, we may write 1 = κm + λp
n
for some κ, λ Z.
i. For any x G, prove that x
κm
H and x
λp
n
K.
ii. Prove that the following function is an isomorphism:
ψ : G H × K : x 7
x
κm
, x
λp
n
(Hint: try evaluating ψ(hk). . . ).
(c) Prove that
|
H
|
= p
n
by applying Cauchy’s Theorem to show that p does not divide
|
K
|
.
(Inducting on this shows that G
=
H
1
×···× H
k
where each H
j
is a p
j
-group of maximal order. A
little more work is needed to show that each H
j
is itself a direct product of cyclic groups and thus
complete the proof.)
91