You are on page 1of 26

Contents

1 Binary Relations 2
1.1 Binary Relations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Properties of Relations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Equivalence Relations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.4 Binary Operations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

2 Groups and Subgroups 4


2.1 Groups and Subgroups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.2 Permutations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.3 Order of Elements and Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.4 Homomorphisms and Isomorphisms . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.5 Cosets and Lagrange’s theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.6 Conjugacy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.7 Normal Subgroups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.8 Quotient Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.9 Products of Subgroups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.10 G-Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.11 Orbit-Stabiliser Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.12 Sylow’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.13 Simple Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.14 Direct Products . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.15 Abelian p-groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.16 Nilpotent Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.17 Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.18 Jordan Hoelder Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.19 Soluble Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.20 Group up to 15 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

3 Rings and Subrings 18


3.1 Rings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.2 Units and Zero Divisors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.3 Ideals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.4 Homomorphisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.5 Factorization and Irreducibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.6 Euclidean Domains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

4 Fields 23
4.1 Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
4.2 Vector Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.3 Vector Spaces: Bases, Independence, Dim . . . . . . . . . . . . . . . . . . . . . . 24
4.4 Linear Transformations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

1
1 Binary Relations
1.1 Binary Relations
Definition 1 (Binary Relation). Let A be a non-empty set. A binary relation on A is a subset
R of A × A. (The set R is the set of all pairs for which the relation is true.) If (a, b) ∈ R, then
we say that a is related to b and write aRb. If (a, b) ∈
/ R, then we write a@Rb.

Definition 2 (congruent modulo). Let n be a positive integer. Let a and b be integers. We say
that a is congruent to b modulo n, and write a ≡ b (mod n) whenever a − b is divisible by n.

Adjacency Matrix: If A if finite then a relation R on A can be represented by its adjacency


matrix. Given some ordering a1 , a2 , . . . an of the elements A, then the adjacency matrix of R is
an n × n matrix (rij )n×n where
(
1 if ai R aj
rij =
0 if ai @
R aj
Two different ways to visualise a relation.

• Adjacency Matrix The relation > on the set A = {1, 2, 3} taken in this order has the
following adjacency matrix:  
0 0 0
1 0 0
1 1 0

• A digraph where the ordered relations are shown using nodes and directed arrows.

1.2 Properties of Relations


• reflexive. A relation R on a set A is reflexive if for all a ∈ A, aRa.
• symmetric. A relation R on a set A is symmetric if for all a, b ∈ A, if aRb, then bRa.
• transitive. A relation R on a set A is transitive if for all a, b, c ∈ A, if aRb and bRc, then
aRc.

1.3 Equivalence Relations


Definition 3 (equivalence relation). A relation ∼ on a set A is an equivalence relation on A if
∼ is reflexive, symmetric and transitive on A.

One of the reasons equivalence relations are so useful is that they allow us to partition sets
into parts that share a common property.

Definition 4 (equivalence class). Let ∼ be an equivalence relation on a (nonempty) set A and


let a ∈ A. The equivalence class of a, denoted [a], is the set of elements of A which are related
to a. That is,
[a] = {b ∈ A : b ∼ a}
Note that a ∈ [a] and so [a] 6= ∅.

Lemma 1.1. Let ∼ be an equivalence relation on a set A. Then for all a, b ∈ A

a ∼ b ⇐⇒ [a] = [b]

Lemma 1.2. Let ∼ be an equivalence relation on a set A, and let a, b ∈ A. If [a] 6= [b], then
[a] ∩ [b] = ∅.

Let ∼ be an equivalence relation on a set A. We can choose a collection ai ∈ A where i ∈ I for


some indexing set I, such that every element is in exactly one of the sets [ai ], i ∈ I. {[ai ] : i ∈ I}
then forms a partition of the elements of A. The ai are equivalence class representatives.

2
1.4 Binary Operations
Definition 5 (binary operation). A binary operation * on A is a mapping form A × A to A.
We denote the image of (a, b) by a ∗ b. To show that an operation * is a binary operation, we
need to show that it is defined for each pair a, b ∈ A, and that the product a ∗ b is an element of
A. That is, we need to demonstrate closure.

Theorem 1.3. Let n be a positive integer. Let a, b, c, d ∈ Z. Suppose a ≡ b (mod n) and c ≡ d


(mod n). Then

1. a + c ≡ b + d (mod n)

2. ac ≡ bd (mod n)

Theorem 1.4 (Division Theorem). Let a be an integer, and let n be a positive integer. Then
there are integers q and r, with 0 ≤ r < n such that a = qn + r. Furthermore, q and r are
unique.

Definition 6 (congruence class representatives). Let n be a positive integer. Then

Z = {0, 1, 2, . . . , n − 1}

If r1 , r2 ∈ Zn , then |r1 − r2 |< n. Thus r1 ≡ r2 (mod n) ⇐⇒ r1 = r2 . Consequently, the


elements of Zn are distinct from each other modulo n. The elements of Zn are the congruence
class representatives.

Definition 7 (binary operations on Zn ). Let n be a positive integer, and let a, b ∈ Zn . We


define a ⊕n b to be the unique element of Zn which is congruent to a + b (mod n) and a ⊗n b to
the the element of Zn which is congruent to ab (mod n).

• A binary operation * on a set A is commutative on A is for all a, b ∈ A, a ∗ b = b ∗ a. If


A is finite and * is commutative then the multiplication table of A under * is symmetric
about the main diagonal. We also say that A is commutative under *.
• A binary operation * on a set A is associative on A if for all a, b, c ∈ A, a∗(b∗c) = (a∗b)∗c.
Unfortunately, there is not easy way to verify that a binary operation on A is associative
from its multiplication table.
• A binary operation * on a set A has an identity element e ∈ A if for all a ∈ A, a ∗ e = a =
e ∗ a.
• Let * be a binary operation on a set A which has an identity element e ∈ A. An element
a ∈ A is invertible if there is an element b ∈ A such that

a∗b=e=b∗a

Lemma 1.5. Let ∗ be a binary operation on a set A. If A under ∗ has an identity element,
then this identity element is unique.

Lemma 1.6. Let ∗ be an associative binary operation on a set A with an identity element e. If
a ∈ A under ∗ is invertible, then a has a unique inverse.

3
2 Groups and Subgroups
2.1 Groups and Subgroups
Definition 8 (Group). A group consists of a non-empty set G, with a binary operation ∗
defined on G × G which satisfies the following axioms:
• G0 (Closure). For all x and y in G: x ∗ y is in G
• G1 (Associativity). For all x, y, and z in G: (x ∗ y) ∗ z = x ∗ (y ∗ z)
• G2 (Identity). There is an element e in G such that: e ∗ x = x ∗ e = x for all x in G.
• G3 (Inverse). For all x in G there is an x0 in G such that: x ∗ x0 = x0 ∗ x = e
We denote a group as a pair (G, ∗). An element e (sometimes called 1 but not the number
1) with the property stated in G2 is said to be the identity element for the ∗ operation in
G, and an element x0 in G3 is said to be the inverse element for x. If G is a group and |G|
is finite, then |G| is known as the order of G; a group with infinitely many elements is said to
have infinite order.
Definition 9 (subgroup). Let (G, ∗) be a group with identity e. A non-empty subset H of G is
a subgroup of G if (H, ∗) is a group. We write H ≤ G. We call H a proper subgroup if H 6= G,
and a non-trivial subgroup if H 6= {e}.
Lemma 2.1. Every group G has exactly one identity element, and every element G has exactly
one inverse.
A group G is abelian or commutative if xy = yx for all x, y ∈ G. That is, if the binary
operation is commmutative. A useful test for a group to be abelian is the following:
Lemma 2.2. Let G be a group. If z 2 = e for all z ∈ G, then G is abelian.
The converse of this is false as there are many abelian groups G with elements z 2 6= e.
Let Un = {m ∈ Zn : gcd(m, n) = 1}.
Theorem 2.3. Let n be a positive integer. Then (Un , ⊗n ) is an abelian group.
Proposition 2.4 (The Subgroup Test). Let G be a group and let H be a subset of G. Then H
is a subgroup of G ⇐⇒ H satisfies
1. H is non-empty
2. for all h, k ∈ H, hk ∈ H
3. for all h ∈ H, h−1 ∈ H
Proposition 2.5 (The Finite Subgroup Test). Let G be a group and let H be a finite non-empty
subset of G. Then H is a subgroup of G if and only if H is closed.
The subgroups of GLn (F ) are
SLn (F ) = {A ∈ GLn (F ) : det(A) = 1}
SL±
n (F ) = {A ∈ GLn (F ) : det(A) = ±1}
U Tn (F ) = {A ∈ GLn (F ) : A is upper triangular}
LTn (F ) = {A ∈ GLn (F ) : A is lower triangular}
Diagn (F ) = {A ∈ GLn (F ) : A is diagonal}
On = {A ∈ GLn (F ) : AAT = In } (orthogonal matrices)
Note that the diagonal matrices are the intersection of the upper and the lower triangular ma-
trices.

Geometry. Let Φ ⊆ Rn . Φ is sometimes called a figure. A symmetry of Φ is a bijection


α : Φ → Φ, which is an isometry i.e. distance preserving. The symmetry group of a regular
n-sided polygon is called the dihedral group of order 2n, and is denoted Dih(2n). It consists
of n rotations and n reflections. The set of symmetries of a figure Φ, under composition of
mappings, forms a group, denoted by Sym(Φ).

4
2.2 Permutations
A permutation of a set X is a bijection from X to itself. If X is a finite set of size n, then
there are n! distinct permutations of X. The set of permutations of X, denoted by Sx , forms a
group under composition of mappings.

Definition 10 (cycle). Let X = {1, 2, . . . , n} and let x1 , x2 , . . . , xk be distinct elements of X.


Let α be a permutation of X defined as follows.

α(x1 ) = x2 α(x2 ) = x3 ··· α(xk−1 ) = xk α(xk ) = x1

α(x) = x, ∀x ∈
/ {x1 , x2 , . . . , xk }. Then α is a cycle of length k. We write α = (x1 x2 . . . xk ).

Definition 11 (transposition). A cycle of length 2 is called a transposition.

Definition 12 (disjoint cycles). Two cycles (x1 x2 . . . xk ) and (y1 y2 . . . yl ) of X =


{1, 2, . . . , n} are disjoint if xi 6= yj for all possible values of i and j. Equivalently, {x1 , x2 , . . . , xk } ⊂
{y1 , y2 , . . . , yl } = ∅.

Definition 13 (even,odd permutation). A permutation of a finite set X is even if it can be


expressed as a product of an even number of transpositions and is odd if it can be expressed as
a product of an odd number of transpositions.

Definition 14 (Alternating Group). Let An be the set of even permutations of {1, 2, . . . , n}.
An is called the alternating group of degree n. If n ≥ 2 then it has 12 n! elements.

Lemma 2.6. If α and β are disjoint cycles, then

1. α ◦ β = β ◦ α; and
2. (α ◦ β)n = αn ◦ β n , for all positive integers n.

Theorem 2.7. Every permutation of a finite non-empty set X can be expressed as a product
of pairwise disjoint cycles. This expression is unique except for the order in which the cycles
appear.

Proposition 2.8. Suppose that α is a permutation written as a product of pairwise disjoint


cycles having i1 , i2 , . . . , ik elements. Then the order is

o(α) = lcm(i1 , i2 , . . . , ik )

Lemma 2.9. Every permutation of a finite set can be written as a product of transpositions.

The length of a cycle is simply the number of elements in the cycle. Given a permutation
σ of a finite set, we let c(σ) be the number of cycles when σ is written as a product of disjoint
cycles, including the cycles consisting of a single element.

Lemma 2.10. Suppose σ is a permutation of a finite set X and i and j are distinct elements
of X. Then c(α ◦ (i j)) = c(σ) ± 1.

Theorem 2.11. No permutation of a finite set X can be expressed both as a product of an even
number of transpositions and as a product of an odd number of transpositions.

2.3 Order of Elements and Groups


Definition 15 (Order). If x is an element of a finite group G then the least positive integer m
such that xm = 1 is called the order of x in G. When G is an infinite group the order of x is
defined in the same way, provided that m exists; otherwise x is said to have infinite order.

Definition 16 (Cyclic). A group G is said to be cyclic if it contains an element x such that


every member of G is a power of x. The element x is said to generate G, and we write G = hxi.

5
Definition 17 (generate). Let a be an element of a group G. We call hai the cyclic subgroup
generated by a. A group G is cyclic if G = hxi for some x ∈ G. We say that x generates G.
Sometimes we denote a cyclic group with n elements by Cn .

Theorem 2.12. Let x be an element of order m in a finite group G. Then xs = 1 in G iff s is


a multiple of m.

If x generates G, and all the powers of x are distinct, then

G = {. . . , x−3 , x−2 , x−1 , 1, x, x2 , x3 , . . . }

In this case we say that G is an infinite cyclic group, and we use the symbol C∞ for a typical
group of this kind. In the case of a cyclic group G with generator x such that the powers of x
are not all distinct then, in this case x is an element of finite order m and:

G = {1, x, x2 , . . . , xm−1 }

And G is said to be a cyclic group of order m. The most familiar instance of a group
isomorphic to Cm is the set Zm of integers modulo m with respect to the + operation.
The order of elements is related to the order of the group.

Proposition 2.13. Let G be a group with a ∈ G. Then hai is a group (under the same binary
operation as G) and the order of hai is o(a).

Proposition 2.14. Let G be a group. If G is cyclic, then G is abelian.

Definition 18 (presentation). A presentation of a group in terms of generators and relations


provides enough information to carry out any multiplication in the group.

2.4 Homomorphisms and Isomorphisms


Isomorphisms try to formalise the notion of a change of notation. Isomorphic groups have the
same multiplication tables after a suitable change of notation.

Definition 19 (homomorphism). Let (G, ∗) and (H, ) be groups. A map θ : G → H is said to


be a homomorphism if for all g1 , g2 ∈ G we have

θ(g1 ∗ g2 ) = θ(g1 )  θ(g2 )

If we encounter some group G in a particular context, we usually set out to show that it
is isomorphic to a “standard example” H, whose properties are already worked out. Then the
group-theoretical properties of G are precisely those of H.

Definition 20 (isomorphism). We say that G and H are isomorphic, denoted G ∼


= H, if there
is a map θ : G → H such that:

• θ is a bijection
• θ is a homomorphism

Lemma 2.15. Let θ : G → H be a homomorphism, and let g be any element of G. Then

1. θ(1G ) = 1H
2. θ(g)−1 = θ(g −1 ); and
3. If o(g) = n for some positive integer n, then o(θ(g)) is finite and divides n.

Corollary 2.16. Let θ : G → H be an isomorphism, and let g ∈ G. Then o(θ(g)) = o(g).


Moreover, for each positive integer n, the number of elements of order n in G is equal to the
number of elements of order n in H.

Lemma 2.17. Let G and H be isomorphic groups and θ : G → H be an isomorphism. Then

6
1. G and H have the same order
2. G is abelian if and only if H is abelian
3. G is cyclic if and only if H is cyclic
4. G and H have the same numbers of elements of each order

When trying to construct isomorphisms, it is often possible to make use of the properties
given in Lemma 2.15 and Corollary 2.16. For example, no θ can be an isomorphism unless it
maps the identity of G to the identity of H and preserves the order of the other elements. To
show that two groups are not isomorphic, we must find a structural property that only one of
them has. For example, if one of the groups has no elements of order two, and the other group
has three elements of order two, then they cannot be isomorphic.

Lemma 2.18. Any two cyclic groups of the same order are isomorphic.

For a homomorphism θ : G → H, the kernel of θ, denoted ker(θ), is the set of elements of


G which are mapped to the identity element of H. That is

ker(θ) = {g ∈ G : θ(g) = 1H }

The image of θ, denoted Im(θ) is the set of elements of H that are of the form θ(g) for some
g ∈ G. That is
Im(θ) = {θ(g) : g ∈ G}

Proposition 2.19. Let θ : G → H be a homomorphism.

1. θ is surjective if and only if Im(θ) = H.


2. θ is injective if and only if ker(θ) = 1G .

Proposition 2.20. Let θ : G → H be a homomorphism. Then ker(θ) is a subgroup of G and


Im(θ) is a subgroup of H.

2.5 Cosets and Lagrange’s theorem


Definition 21 (Cosets). Let H be a subgroup of a, not necessarily finite, group G. The left
coset gH of H with respect to an element g in G is defined to be the set obtained by multiplying
each element of H on the left by g, that is

gH = {x ∈ G|x = gh for some h ∈ H}

The right coset of H with respect to g is defined as

Hg = {x ∈ G|x = hg for some h ∈ H}

Definition 22 (index). If G is a finite group and H ≤ G, then |G|/|H| is called the index of
H in G and is denoted by |G : H|. It is the number of right cosets of H whose union is G.
Thus, |G|= |G : H|×|H|.

If H is a finite subgroup, say H = {h1 , h2 , . . . , hm }, then the elements which belong to


the left coset gH are: gh1 , gh2 , . . . , ghm . It is clear that they are all distinct due to the left
cancellation rule. This gives rise to a fundamental property of cosets:

|gH|= |H| (g ∈ G)

Lemma 2.21. Let G be a group H ≤ G. The relation ∼, defined on G by g1 ∼ g2 is an


equivalence relation if and only if g1 ∈ Hg2 . The equivalence classes are the right cosets of H
in G.

Theorem 2.22. Let H be a subgroup of a group G. If g1 and g2 are any elements of G, the left
cosets g1 H and g2 H are either identical or they have no elements in common.

7
Theorem 2.23 (Lagrange’s theorem). If G is a finite group and H is a subgroup of G, then
|H| divides |G|.

By Lagrange’s Theorem |G|= |G : H|×|H|.

Corollary 2.24 (Lagrange Theorem Corollaries).


• If G is a group of prime order p, then G is cyclic and any non-identity element of G
generates G.
• If G is a group of prime order, then any non-identity element of G generates G.
• Let G be a finite group and a ∈ G with o(a) = n. Then n divides |G|.
• Let G be a finite group and a ∈ G. Then a|G| = e.
• Every subgroup of a cyclic group is cyclic.

Theorem 2.25. If G is a finite group of order n ≥ 2, the following statements are equivalent
1. G is a cyclic group
2. For each divisor d of n the number of elements x in G which satisfy xd = 1 is d
3. For each divisor d of n the number of elements x in G which have order d is ϕ(d).

Proposition 2.26. Let H ≤ G, and g, x, y ∈ G. Then g ∈ Hx if and only if Hg = Hx and


g ∈ yH if and only if gH = yH.

Lemma 2.27. Let H ≤ G and g1 , g2 ∈ G. Then g1 H = g2 H ⇐⇒ g1−1 g2 ∈ H and Hg1 =


Hg2 ⇐⇒ g1 g2−1 ∈ H.

2.6 Conjugacy
Conjugacy is a way of finding subgroups. Conjugation maps subgroups to isomorphic subgroups.

Definition 23 (y is conjugate to x in G). Let G be a group and x, y ∈ G. We say that y is


conjugate to x in G if there is an element g of G such that y = gxg −1 . We can write y conj
x for this or xg for gxg −1 . Conjugacy is an equivalence relation on G. The element g is called
the conjugating element.

Definition 24 (conjugacy class). The set of all y which are conjugate to x is called the conju-
gacy class of x in G and denoted xG . Thus

xG = {gxg −1 : g ∈ G}

The definition of conjugacy can be extended to cover subgroups. Then gHg −1 is a subgroup
isomorphic to H. For a subgroup H of a group G and an element g of G, we define the conjugate
of H by G to be the set
gHg −1 = {ghg −1 : h ∈ H}

Definition 25 (centralizer). For x ∈ G the centralizer in G of x, denoted CG (x), is the set of


all g ∈ G which commute with x. That is

CG (x) = {g ∈ G : gx = xg}

Elements of CG (x) are said to centralize x. If G is abelian then for all x, g ∈ G, gxg −1 = x so
each x is only conjugate to itself. And the conjugacy classes all consist of exactly one element.

Definition 26 (normaliser). The normaliser of H in G, denoted NG (H) is the set of group


elements g such that gHg −1 = H. Note that NG (H) always contains H, because H is closed
under multiplication.

Definition 27 (centre). The centre, Z(G), of a group G is the set of all g ∈ G which commute
with every element of G. That is

Z(G) = {g ∈ G : gx = xg, ∀x ∈ G}

8
CG (x) ≤ G for all x ∈ G and Z(G) ≤ G. Centralisers are about an element in a group.
Centres are about the entire group.
Proposition 2.28. Let G be a group, x ∈ G. Then |G|= |xG |×|CG (x)|.
Lemma 2.29. Let G be a group and let x and y be elements of G. If y is conjugate to x, then
x and y have the same order.
Proposition 2.30. Let G = Sn and x, y ∈ G. Then gxg −1 = y, where y is obtained by
replacing each number i in x (written as the product of disjoint cycles) by g(i). For example,
(12)g = (g(1)g(2)).
This then becomes a quick way of calculating gHg −1 .
Theorem 2.31. Elements of Sn are conjugate in Sn ⇐⇒ they have the same cycle type.
Lemma 2.32. If H ≤ G and g is an element of G, then gHg −1 ≤ G and isomorphic to H.
Proposition 2.33. Let H be a subgroup of a group G and let g be an element of G. Now let
K = gHg −1 . Define a map θ : H → K by θ(h) = ghg −1 . Then θ is an isomorphism.
Lemma 2.34. For every group G, Z(G) is an abelian subgroup of G.
Lemma 2.35. For every element x of a group G, CG (x) is a subgroup of G and Z(G) ≤ CG (x).
Proposition 2.36. If x and g are elements of a group G, then gxg −1 = x ⇐⇒ x commutes
with g. Let H = CG (x). Suppose that g1 and g2 are in the same left H-coset of G. That is,
g1 H = g2 H. Then g2 xg2−1 = g1 xg1−1 .

2.7 Normal Subgroups


Definition 28 (normal subgroup). Let G be a group and H a subgroup of G. We say that H
is a normal subgroup of G and write H E G, if gHg −1 = H, ∀g ∈ G. This is equivalent to
saying that G = NG (H), or that Hg = gH, ∀g ∈ G. In other words, H is normal if for every
element g of G, the left H-coset of g consists of the same elements as the right H-coset of g.
Proposition 2.37. Let H ≤ G, and g, x, y ∈ G. Then g ∈ Hx ⇐⇒ Hg = Hx and g ∈
yH ⇐⇒ gH = yH.
Lemma 2.38. Suppose G is a group, H ≤ G and A is a set of coset representatives such that
the distinct right H-cosets are {Hx : x ∈ A}. Then H E G ⇐⇒ Hx = xH ∀x ∈ A.
Proposition 2.39.
• Let G be an abelian group. Let H be any subgroup of G. Then H E G.
• For any group G, {1} E G and G E G.
• Z(G) is a normal subgroup of G, for every group G.
• Let f be a homomorphism from G to a group H. Then the ker(f ) E G.
• For g ∈ G, H ≤ G, gHg −1 is a subgroup of G with the same order as H. H is normal in
G if and only if gHg −1 ∈ H, ∀g ∈ G and h ∈ H.
Lemma 2.40. Let N E G. For any subgroup H of G, N ∩ H is a normal subgroup of H.
Note that N ∩ H is not necessarily a normal subgroup of G, because, given n ∈ N ∩ H, if
g∈/ H there is no reason for gng −1 ∈ H, and hence gng −1 is not necessarily contained in N ∩ H,
although it will be in N .
Lemma 2.41. If |G : H|= 2, then H is normal in G.
Lemma 2.42. Let G be a group, H ≤ G. Then the following are equivalent:
1. Hg = gH, ∀g ∈ G, that is, H E G.
2. ghg −1 ∈ H, ∀h ∈ H, g ∈ G.
Lemma 2.43. Let H be a subgroup of a group G and g be an element of G. Then gHg −1 is a
subgroup of G with the same order as H. Furthermore, H is normal in G ⇐⇒ gHg −1 = H
for all g ∈ G. (Recall that we showed H is normal ⇐⇒ ghg −1 ∈ H ∀ g ∈ G and h ∈ H.)
Corollary 2.44. Let H be a subgroup of G, then H is normal ⇐⇒ it is a union of conjugacy
classes of G.

9
2.8 Quotient Groups
Theorem 2.45 (Quotient Group). Let G be a group and N E G a normal subgroup. The set
of left N -cosets of G denoted by G/N , is a group under the multiplication (gN )(hN ) = (gh)N .
It is known as the quotient of G by N.
Let A be the set of normal subgroups and B be the set of kernels of homomorphisms. We
know that B ⊆ A since the ker(θ) is a normal subgroup. To go in the other direction, let N E G
and define a map θ : G → G/N as θ(g) = gN .
θ(g1 g2 ) = (g1 g2 )N = (g1 N )(g2 N ) = θ(g1 )θ(g2 )
Hence θ is a homomorphism. The kernel of θ is the set
{g ∈ G : θ(g) = N } = {g ∈ G : gN = N } = N
Thus, N is the kernel of a homomorphism. This shows that A ⊆ B. So A = B.
Theorem 2.46 (The Homomorphism Theorem). Let G and H be groups. If θ : G → H is a
homomorphism from G to H, then
Im(θ) ∼
= G/ker(θ)
Theorem 2.47. Let G be a group, H ≤ G and N E G. Then
H ∼ HN
= .
N ∩H N
Theorem 2.48. Suppose H and N are normal subgroups of G, and that N ≤ H. Then H/N
is a normal subgroup of G/N , and
G/N ∼
= G/H
H/N
Definition 29 (extension). We say that a group G is an extension of N by a group H if G
has a normal subgroup isomorphic to N such that the group G/N is isomorphic to H.

2.9 Products of Subgroups


Proposition 2.49. Let G be a group and let H and K be subgroups of G. Then H ∩ K ≤ G.
On the other hand, H ∪ K is not generally a subgroup of G.
Another way to get a group from H and K is to take hH, Ki which is called the subgroup
generated by H and K. It is defined to be the smallest subgroup of G containing both H and
K. Equivalently it is the intersection of all the subgroups of G containing both H and K.
Let H, K ≤ G. Then the product HK is defined as
HK = {hk : h ∈ H, k ∈ K}
Proposition 2.50. Let H and K be subgroups of a group G. If HK is a subgroup of G, then
HK = hH, Ki.
Although HK may or may not equal hH, Ki, in the case where HK is a subgroup of G the
two coincide, so we do not get any new subgroups by considering HK.
Proposition 2.51. For H, K ≤ G, HK is a subgroup of G ⇐⇒ HK = KH.
Proposition 2.52. Let N E G and H be any subgroup of G (not necessarily normal). Then
N H = HN and hN, Hi = N H.
That is, if we have subgroups H and K of a group G, and H is normal, then
HK = {hk : h ∈ H, k ∈ K} = ∪k∈K Hk = ∪k∈K kH = KH
Therefore HK is a subgroup.
Corollary 2.53. If G is abelian, then HK is a subgroup of G for every H, K ≤ G.
Lemma 2.54. Let H and K be finite subgroups of a group G. Then
|H|×|K|
|HK|=
|H ∩ K|

10
2.10 G-Sets
Definition 30. Let G be a group and X be a set. We say X is a G-set if each element g of G
can be thought of as a map g : X → X such that the following conditions hold
• ∀x ∈ X, 1G · x = x.
• ∀x ∈ X and all g1 , g2 ∈ G, (g1 g2 ) · x = g1 · (g2 · x)
We sometimes say that G ‘acts’ on X, or that there is a G-action on X, to mean that X is a
G-set.
Essentially the elements of G are being converted into elements of Sn . We could equivalently
define this idea by saying that X is a G-set if there exists a homomorphism ρ : G → Sx .
Lemma 2.55. Suppose X is a G-set for some group G, and let H be any subgroup of G. Then
(under the same action) X is also a H-set.
Proposition 2.56. A group G can itself be made into a G-set (that is, take X = G) in the
following ways
• by left multiplication. For g, x ∈ G, define g · x = gx.
• by right inverse multiplication. For g, x ∈ G, define g · x = xg −1 .
• by conjugate action. For g, x ∈ G, define g · x = gxg −1 .
Three examples of how the above Proposition might work are as follows:
1. Let G be a group and let X be the collection of all subsets of G. Then G acts on X by
left multiplication, right inverse multiplication and conjugation.
2. Let G be a group and let X be the set of all subgroups of G. Then G acts by conjugation
on X.
3. Let G be a group and H ≤ G. Then G acts by left multiplication on the set of left cosets
of H, and by right inverse multiplication on the set of right cosets of H.

2.11 Orbit-Stabiliser Theorem


• Where can x get sent by elements of G? Orbit.
• What are the elements of G which fix x? Stabiliser.

Definition 31 (orbit). Let X be a G-set and let x ∈ X. The orbit of x, denoted by orb(x), is
the set
orb(x) = {g · x : g ∈ G} ⊆ X
The size of an orbit is its length.
Definition 32 (stabiliser). The stabiliser of x, denoted by Gx , is the set

Gx = {g ∈ G : g · x = x} ⊆ G

Proposition 2.57. Suppose X is a G-set. Let ∼ on X be the binary relation given by x ∼ y if


there exists g ∈ G with x = g · y. Then ∼ is an equivalence relation on X and for each x ∈ X,
the equivalence class [x] of X is orb(x).
Corollary 2.58. For a G-set X, the orbits form a partition of X.
Proposition 2.59. Let X be a G-set. For any x ∈ X, Gx is a subgroup of G.
Theorem 2.60 (Orbit-Stabilizer Theorem). Suppose G is a finite group. Let X be a G-set. For
any x ∈ X,
|orb(x)|×|Gx |= |G|
|G|
Put another way, |orb(x)|= |Gx | = |G : Gx |, the index of the stabilizer Gx in G (provided it is
finite).

11
As a consequence of the Orbit-Stabilizer Theorem, if G is a finite group and x ∈ G, then
|G|
|xG |=
|CG (x)|
The orbit of x is the conjugacy class, and the stabilizer is the centralizer.

Examples
Example 1: Let G act on itself by conjugation. For x ∈ G,
orb(x) = {g · x : g ∈ G} = {gxg −1 : g ∈ G} = xG
So the orbit of x is exactly the conjugacy class of x in this case. Similarly, when G acts on itself
by conjugation then the stabilizer of Gx is the set
{g ∈ G : gxg −1 = x} = {g ∈ G : gx = xg}
Example 2: Take X to be the set of all subgroups of G. Then G acts on X by conjugation.
The orbit of a subgroup H is the set of subgroups conjugate to H, namely {gHg −1 : g ∈ G}.
The stabiliser of H is the normaliser NG (H) of H in G. The orbit stabiliser theorem says that
the number of conjugates of H is the index of NG (H) in G.
Definition 33 (p-groups). p-groups are groups of order pn for some prime p and some positive
integer n.
Proposition 2.61. Let p be prime and suppose G is a group of order pn for some positive n.
Then the center of G, Z(G) is non-trivial.
Proposition 2.62. Let G be a group such that G/Z(G) is cyclic. Then G is abelian.
Proposition 2.63. Suppose G has order p2 for some prime p. Then G is abelian.

2.12 Sylow’s Theorem


Theorem 2.64. Let G be a finite group of order pn m where p is a prime which does not divide
m, and n is a positive integer. Let np be the number of subgroups of G which have order pn
(known as Sylow p-groups). Then
1. G contains at least one Sylow p-subgroup. Also np ≡ 1 (mod p) and np divides |G|.
2. If P is a Sylow p-subgroup, then any subgroup Q of G with |Q|= pr for some r ≥ 0 is
contained in some Sylow p-subgroup conjugate to P . In particular, the Sylow p-subgroups
form a single conjugacy class of subgroups.
n 
Lemma 2.65. Let p be prime and m ∈ Z+ . If m is not divisible by p, then ppnm ≡ m (mod p)
Lemma 2.66. Let G be a group with order pn m. Then G has at least one Sylow p-subgroup.
Lemma 2.67. Let G be a group with order pn m. Then np ≡ 1 (mod p).
Proposition 2.68. Let P be a Sylow p-subgroup of G. Any p-subgroup of NG (P ), the normaliser
of P in G, is contained in P .
Lemma 2.69. Let G be a group of order pn m for some prime p, positive integer n, and positive
integer m coprime to p. Let P be a Sylow p-subgroup of G. Then any subgroup Q of G with
|Q|= pr for some r ≥ 0 is contained in some conjugate of P .
Lemma 2.70. Let G be a group of order pn m for some prime p, positive integer n, and positive
integer m coprime to p. Then the Sylow p-subgroups form a single conjugacy class of subgroups,
and hence np divides the order of |G|.
Theorem 2.71 (Cauchy’s Theorem). Suppose G is a finite group, and p is a prime number
such that p divides |G|. Then G has at least one element of order p.
Proposition 2.72. Suppose G is an abelian group of square-free order. Then G is cyclic.
Lemma 2.73. Any group of order 4 is either cyclic or a Klein 4-group.
Theorem 2.74. Let p be an odd prime. A group G of order 2p is either cylic or dihedral.

12
2.13 Simple Groups
Definition 34 (simple groups). A group G is said to be simple if G has no non-trivial proper
normal subgroups.

There are four types of (non-trivial) finite simple groups


1. Cyclic groups of prime order.
2. Alternating groups An for n ≥ 5.
3. ‘Group of Lie Type’ (these are various collections of groups of matrices)
4. Sporadic Simple Groups (26 finite simple groups that aren’t in any of the other 3 categories)

Proposition 2.75. Let G be a (non-trivial) finite abelian simple group. Then G is cyclic of
prime order.

Lemma 2.76. If n ≥ 3, then An is generated by 3-cycles.

Theorem 2.77. If n ≥ 5, then An is simple.

2.14 Direct Products


Definition 35. Let H and K be groups. The direct product of H and K, denoted H × K, is
the set of ordered pairs (h, k), with h ∈ H, k ∈ K. Multiplication is pointwise, so that

(h1 , k1 )(h2 , k2 ) = (h1 h2 , k1 k2 )

The direct product H × K is a group with this multiplication; the identity element is (1H , 1K ).

• If H, K are abelian then H × K is also abelian.

• If G = Cr × Cs for some positive integers r, s, then every element of G has order at most
lcm(r, s), and this bound is attained. Thus G is cyclic if and only if gcd(r, s) = 1.

• The |H × K|= |H|×|K|.

• There is a natural projection πk ((h, k)) = k, where π is a map H × K → K. π is a


homomorphism. The ker(πk ) = {(h, k) ∈ H × K : πk (h, k) = 1} = {(h, 1) ∈ H × K} =
H.
b Therefore, Hb is a normal subgroup since the kernel of a homomorphism is a normal
subgroup.

• If we define H
b = {(h, 1) : h ∈ H} and K
b = {(1, k) : k ∈ K} then H
b and K
b are normal

subgroups of H × K. Furthermore, H = H ∼
b and K = K. b

Lemma 2.78. Let G be a finite group and suppose that H and K are normal subgroups of G with
the property that H ∩ K = {1}. Then hH, Ki = HK ∼ = H × K. If, in addition, |H|×|K|= |G|,
then G ∼
= H × K.

Proposition 2.79. Let G be a finite group. If H1 , H2 , . . . , Hr are normal subgroups of G with


the property that Hi ∩ Hj = {1} for all i 6= j, and |G|= |H1 |×|H2 |× · · · × |Hr |, then

G∼
= H1 × H2 × · · · × Hr

Lemma 2.80. Suppose H, K are normal subgroups of a group G, with H ∩ K = {1}. If every
element of G can be written hk for some h ∈ H, k ∈ K, then G ∼
= H × K.

Definition 36. Suppose H, K are normal subgroups of a group G, with H ∩ K = {1}. If


G = HK or equivalently if every element of G can be written hk for some h ∈ H, k ∈ K, then
we say that G is the internal direct product of H and K.

Lemma 2.81. Suppose G is the internal direct product H × K. Then every element of G can
be written in a unique way in the form hk for h ∈ H, k ∈ K.

13
Theorem 2.82. Every finite nontrivial abelian group is the internal direct product of its Sylow
subgroups.

Proposition 2.83. Suppose G is abelian and can be decomposed as an internal direct product

G∼
= H1 × H2 × · · · × Hr

Let p be any prime dividing |G|, and let Pi be the Sylow p-subgroup of Hi (where Pi = {1}
if p does not divide |Hi |). Then the Sylow p-subgroup of G ∼
= P1 × · · · × Pr . For example, if
G∼ = C4 × C6 then the Sylow 2-subgroup of G ∼= C4 × C2 .

2.15 Abelian p-groups


Recall that proposition 2.19 states that any group of order p2 is abelian. An abelian group
is said to be elementary abelian if all its non identity elements have the same order. By
Cauchy’s theorem this order must be a prime p. The group is then called an elementary
abelian p-group.

Proposition 2.84. A group of order p2 (p is prime) is either cyclic or elementary abelian (and
isomorphic to Cp × Cp ).

Lemma 2.85. Let s and n be positive integers with s ≤ n and let p be prime. Then the set of
n n−s
elements of order dividing ps in Cpn = hg : g p = 1i is {g dp : d ∈ Z}.

So, for example, if we let pn = 81 and ps = 9 then the set of elements of order dividing
ps in Cpn would be those elements of g which are a multiple of pn−s = 9 times a scalar:
{9, 18, 27, 36, 45, 54, 63, 72, 81}.

Lemma 2.86. Let G be an abelian p-group, let a be an element of maximal order in G and let
H = hai. If bH is an element of G/H having order pm , then bH contains an element of order
pm in G.

Remark: The point here is that for the coset bH to have order pm , all that is required is
m
that bp ∈ H, so b may well have order greater than pm . This lemma establishes that we can
find at least one element in the coset that really does have order pm as an element of G.

Theorem 2.87. Let G be an abelian p-group. Then G is an internal direct product of cyclic
p-groups.

If G is abelian and is the internal direct product of cyclic p-groups P1 · · · Pk then WLOG
(since the group is abelian) we may assume that |P1 |≥ |P2 |≥ |Pk |. So any decomposition of G
as a direct product of cyclic p-groups is of the form Cpλ1 × · · · Cpλr for some λ1 ≥ λ2 · · · λr . We
say that this decomposition has type [λ1 , . . . , λr ].

Theorem 2.88. Let G be a finite abelian p-group. Then every decomposition of G as a direct
product of cyclic p-groups has the same type.

Theorem 2.89. Every finite nontrivial abelian group G is isomorphic to a direct product of
cyclic p-groups. This decomposition is unique up to re-ordering of the factors. That is, the
orders of the cyclic p-groups involved, and their multiplicities, are uniquely determined by G.

Lemma 2.90. Suppose G ∼ = Cn1 × Cn2 × · · · Cnt for some positive integers n1 , . . . nt . Then
G∼ C
= n1 n2 ...nt iff n 1 , . . . nt are coprime.

2.16 Nilpotent Groups


Proposition 2.91. Suppose G is a group and N a normal subgroup. If Q is a subgroup of G/N ,
then the set H = {g : gN ∈ Q} is a subgroup of G. Moreover, N ≤ H and H/N = Q. Finally
Q is normal in G/N iff H is normal in G.

14
Definition 37 (nilpotent). Let G be a group. Define Z1 (G) = Z(G). Now consider Z(G/Z(G)).
Let Z2 (G) be the subgroup of G such that Z2 (G)/Z(G) = Z(G/Z(G)). For each i ≥ 2, define
Zi (G) to be the subgroup of G for which

Zi (G)/Zi−1 (G) = Z(G/Zi−1 (G))

The series of subgroups {1} ≤ Z1 (G) ≤ Z2 (G) ≤ · · · is known as the upper central series of
G. If for some n we have Zn (G) = G, then we say that G is nilpotent. If G is nilpotent and n
is the smallest positive integer for which Zn (G) = G, then we say G is nilpotent of class n.

Lemma 2.92. Every finite p-group is nilpotent.

Theorem 2.93. Let G be a finite group. If G is an internal direct product of its Sylow subgroups,
then G is nilpotent.

2.17 Series
Definition 38 (series). Let G be a group

1. A subnormal series for G is a chain of subgroups

G = G0 ≥ G1 ≥ G2 ≥ · · · ≥ Gr = {1}

where for each i we have Gi E Gi−1 . Note that Gi does not have to be normal in G, just
in Gi−1 .

2. A refinement of a subnormal series G = G0 ≥ G1 ≥ G2 ≥ · · · ≥ Gr = {1} is a subnormal


series G = H0 ≥ H1 ≥ H2 ≥ · · · ≥ Hs = {1} such that each Gi is equal to some Hj .

3. A composition series for G is a subnormal series without repetitions such that all re-
finements have repeating terms.

4. A normal series for G is a subnormal series where each Gi is normal in G, and not just
in Gi−1 .

5. A normal refinement of a normal series G = G0 ≥ G1 ≥ G2 ≥ · · · ≥ Gr = {1} is a


normal series G = H0 ≥ H1 ≥ H2 ≥ · · · ≥ Hs = {1} such that each Gi is equal to some
Hj .

6. A chief series for G is a normal series without repetitions such that all normal refine-
ments have repeating terms.

Definition 39 (isomorphic normal/subnormal). Suppose S and T are subnormal (respectively


normal) series for a group G, so that for some subgroups Gi , Hj of G, the series S is given by

G = G0 ≥ G1 ≥ G2 ≥ · · · ≥ Gr = {1}

and T is given by.


G = H0 ≥ H1 ≥ · · · ≥ Hs = {1}
We say that S and T are isomorphic subnormal(respectively normal) series iff the elements
of the sequence [G0 /G1 , G1 /G2 , . . . , Gr−1 /Gr ] can be reordered so that they are isomorphic to
corresponding elements of the sequence [H0 /H1 , H1 /H2 , . . . , Hr−1 /Hr ].

15
2.18 Jordan Hoelder Theorem
Lemma 2.94 (Zassenhaus ‘butterfly’). Let H and K be subgroups of a group G. Let A be a
normal subgroup of H and B be a normal subgroup of K. Then

(H ∩ K)A ∼ (H ∩ K)B
=
(H ∩ B)A (A ∩ K)B

Theorem 2.95 (Shreier). Given subnormal series S and T for a group G, there exist subnormal
series S 0 and T 0 such that S 0 is a refinement of S, T 0 is a refinement of T and S 0 and T 0 are
isomorphic. If S and T are normal series, then S 0 and T 0 can be chosen to be isomorphic normal
refinements of S and T .

Lemma 2.96. Let N, H, and K be subgroups of a group G, with K ≤ H. Suppose that N is


normal in G and K is normal in H. Then KN is normal in HN .

Theorem 2.97 (Jordan-Hoelder). If a group has a composition series then any two composition
series for that group are isomorphic. If a group has a chief series then any two chief series for
that group are isomorphic.

Definition 40 (composition factors of G). Let G = G0 > G1 > · · · > Gr = {1} be a composition
series for a group G. Then the quotient groups G0 /G1 , G1 /G2 , . . . , Gr−1 /Gr are called the
composition factors of G. Similarly if the series is a chief series, then the associated quotient
groups are called the chief factors of G.

Proposition 2.98. If A is a composition factor of a group, then A is simple.

Corollary 2.99. The only composition factors of non-trivial finite abelian groups are cyclic of
prime order.

Theorem 2.100 (Fundamental Theorem of Arithmetic). Every integer greater than 1 can be
uniquely factorised as a product of prime numbers.

Definition 41 (automorphism). Let G be a group. An automorphism of G is an isomorphism


from G to G.

Definition 42 (characteristic subgroup). Let G be a group. A subgroup N of G is a charac-


teristic subgroup of G if f (N ) = N for all automorphisms f of G.

Definition 43 (characteristically simple). A group G is said to be characteristically simple


if its only characteristic subgroups are {1} and G.

Proposition 2.101. Let G be a finite characteristically simple group. Then G is an internal


direct product of simple groups all of which are isomorphic.

Theorem 2.102. A chief factor of a finite group is characterstically simple, and hence is a
direct product of isomorphic simple groups.

Corollary 2.103. A non-trivial abelian chief factor of a finite group is an elementary abelian
p-group.

Theorem 2.104. The smallest nonabelian simple group is A5 , which has order 60.

Corollary 2.105. Let G be a non-trivial group of order less than 60. Then the composition
factors of G are cyclic of prime order, and the chief factors of G are elementary abelian.

16
2.19 Soluble Groups
Definition 44 (soluble). A group G is soluble if it has a normal series G = G0 ≥ G1 ≥ · · · ≥
Gr = {1} whose factors Gi /Gi+1 are all abelian.

Lemma 2.106. Let G be finite. Then G is soluble iff the chief factors of G are abelian, which
is iff the chief factors of G are elementary abelian.

Proposition 2.107. Suppose G is soluble and H is a subgroup of G. Then H is soluble.

Proposition 2.108. Let G be a finite simple soluble group. Then G is cyclic of prime order.

Definition 45. Let g, h be elements of a group G. Then the commutator of g and h, denoted
[g, h], is the element ghg −1 h−1 . The commutator subgroup or derived group of G, denoted
G0 or [G, G], is the subgroup generated by the commutators of G. So

G0 = [G, G] = h[g, h] : g, h ∈ Gi

Note that if g and h commute then [g, h] = 1. Therefore, if G is abelian, then G0 = [G, G] = {1}.
In fact, G is abelian iff G0 = {1}.

Theorem 2.109. Let G be a group. Then G0 is a characterstic subgroup of G. Moreover, G0


is the smallest normal subgroup with abelian quotient. That is, if N is a normal subgroup of G
such that G/N is abelian, then G0 ≤ N .

Definition 46. For a group G, the derived series is defined as follows:

G(0) = G; G(1) = G0 = [G, G]; G(2) = [G0 , G0 ]; ...; G(r) = [G(r−1) , G(r−1) ]

Theorem 2.110. A group G is soluble iff G(r) = {1} for some r.

2.20 Group up to 15
Proposition 2.111. There are exactly two non-isomorphic non-abelian groups of order 8; they
are Dih(8) and Q8 .

Lemma 2.112. A4 ∼
= G = ha, b, c : a2 = 1, b2 = 1, c3 = 1, ab = ba, ca = bc, cb = abci.

Proposition 2.113. There are exactly three non-isomorphic non-abelian groups of order 12;
they are Dih(12), Q12 and A4 .

17
3 Rings and Subrings
3.1 Rings
When we define a ring we have in mind the set of integers with the two operations of addition
and multiplication.

Definition 47 (ring). A ring is a set R with two binary operations called addition and multi-
plication satisfying axioms R1 to R4. For a, b ∈ R, addition is denoted a + b, and multiplication
is denoted ab.

• R1 (R, +) is an abelian group with identity called 0.


• R2 For all a, b ∈ R, ab ∈ R (multiplicative closure)
• R3 For all a, b, c ∈ R, a(bc) (multiplicative associativity)
• R4 For all a, b, c ∈ R, (a + b)c = ac + bc and c(a + b) = ca + cb (distributive laws)

The first axiom is actually five axioms because it says that R is an abelian group under addition.

Definition 48. Let R be a ring. Then

• R is a ring with identity if there exists an element 1 ∈ R, with 1 6= 0, such that for all
a ∈ R, 1a = a1 = a.
• R is a division ring if it is a ring with identity and for all a ∈ R with a 6= 0, there exists
b ∈ R with ab = ba = 1.
• R is a commutative ring if for all a, b ∈ R, ab = ba.
• R is a field if it is a commutative division ring.

A field is a ring in which multiplication is commutative and every element except 0 has a
multiplicative inverse. Thus, in a field F we have: U (F ) = F \{0}.

Lemma 3.1. Let R be a ring. For all a ∈ R, a0 = 0a = 0. Moreover, if R has at least two
elements, then R is not a group under multiplication.

Lemma 3.2. Let R be a set with addition and multiplication operations defined on it. Then R
is a field if and only if (R, +) is an abelian group with identity 0, (R∗ , ·) is an abelian group,
both 0x and x0 belong to R for all x ∈ R and the distributive axiom holds.

The usefulness of this lemma is that since it is an iff proof, it gives us a way of proving
something is a field.

• the trivial ring has just one element R = {0}.


• the zero ring has the multiplication operation ∀a, b ∈ R, a · b = 0.
• the only subrings of Z are precisely the subrings nZ.

Definition 49 (polynomial ring). Let R be a ring. An R-polynomial (or polynomial over R) in


x is simply a formal sum

P (x) = a0 + a1 x + a2 x2 + · · · + an xn

where each ai ∈ R and n ≥ 0 is an integer, and either ai = 0 for all i or an 6= 0. The zero
element is the zero polynomial 0(x) = 0. We denote the set of all R-polynomials in x by R[x].
If R is a ring with identity 1, we say f (x) is a monic polynomial if an = 1.

Definition 50 (subring). Let R be a ring. A subring of R is a subset S of R which is itself a


ring under the same addition and multiplication as R.

Associativity is inherited from R. Therefore we only need to check the following:

Proposition 3.3 (subring criterion). Let R be a ring. A subset S of R is a subring if S is


non-empty and for all a, b ∈ S we have:

18
1. a + b ∈ S
2. −a ∈ S
3. ab ∈ S

Definition 51 (number ring). A number ring is a subring of the ring C, with the usual addition
and multiplication.

Definition 52 (Gaussian integers). The set of Gaussian integers is the set Z[i] = {a + bi : a, b ∈
Z}.

3.2 Units and Zero Divisors


Definition 53 (unit). Let R be a ring with identity 1, and a ∈ R. Then a is a unit if it has a
multiplicative inverse, that is, there exist a0 ∈ R with aa0 = a0 a = 1.

The multiplicative inverse, if it exists, is unique.

Lemma 3.4. Let R be a division ring. Then every linear equation has a unique solution.

Note that since every field is a division ring, and so linear equations in fields are always
solvable.

Proposition 3.5. Let R be a ring with identity and U = U (R) be the set of units of R. Then
U is a group under the multiplication defined for R.

Lemma 3.6. Let p be prime, then Zp is a field.

In fact, Zn is a field ⇐⇒ n is prime.

Theorem 3.7 (Fermat’s Little Theorem). Let p be prime and a be a positive integer not divisible
by p. Then ap−1 ≡ (mod p).

Lemma 3.8. Every subring of a commutative ring is itself commutative.

Definition 54 (zero divisor). Let R be a ring. An element a ∈ R is known as a zero divisor if


a 6= 0 and there exists a nonzero z ∈ R such that az = 0.

Definition 55 (integral domain). A commutative ring with identity that has no zero divisors is
called an integral domain.

Therefore, all number rings containing 1 are integral domains.

Lemma 3.9. Let R be a ring with identity and a ∈ R. If a is a unit, then a is not a zero
divisor.

Lemma 3.10 (Cancellation). Let R be a ring and a, x, y ∈ R. Suppose ax = ay, where a is


neither 0 nor a zero divisor. Then x = y.

Definition 56 (Field of Fractions). Let R be an integral domain. Then a field F is a field of


fractions of R if R is a subring of F , every element of F can be thought of as ab−1 for some
a, b ∈ R, b 6= 0. Q is the field of fractions of Z.

Theorem 3.11. Any integral domain has a field of fractions.

3.3 Ideals
Definition 57 (ideal). Let R be a ring and S a subring of R. Then S is an ideal of R if for all
r ∈ R, s ∈ S, we have rs ∈ S and sr ∈ S.

Lemma 3.12 (The Ideal Criterion). Let R be a ring. A subset of S of R is an ideal if S is


non-empty and for all x, y ∈ S and r ∈ R we have

19
• x+y ∈S
• −x ∈ S
• xr ∈ S and rx ∈ S

For a ring R and a ∈ R, we define hai to be the smallest ideal containing a, and call it the
ideal generated by a. So, if I is an ideal with a ∈ I then hai ⊂ I and hai is the intersection of
all ideals of R that contain a. By definition, any ideal containing a must at least contain ar for
all r ∈ R. We write hai = {ar : r ∈ R} = aR.

Definition 58 (principal ideal). Let R be a commutative ring with identity. We say that an
ideal I of R is a principal ideal if I = hai = aR for some a ∈ R. If R is an integral domain and
every ideal of R is a principal ideal, then R is called a principal ideal domain.

Lemma 3.13. Let R be a commutative ring with identity. Them the set aR is an ideal which
contains a and hence hai = aR.

3.4 Homomorphisms
Definition 59 (ring homomorphism). Let R and S be rings. A map θ : R → S is a (ring)
homomorphism if for all a, b ∈ R:

θ(a + b) = θ(a) + θ(b) and θ(ab) = θ(a)θ(b)

If θ is also bijective, then we say θ is a (ring) isomorphism, and that R and S are isomorphic
rings. R ∼= S.

Lemma 3.14. Suppose R and S are number rings with identity, and θ : R → S is a homomor-
phism. Then either θ is the zero homomorphism (θ(a) = 0 for all a ∈ R), or θ(1) = 1.

If θ is a ring homomorphism from Z to Z, then

θ(2) = θ(1 + 1) = θ(1) + θ(1) = 2θ(1)

and in general θ(n) = nθ(1), for all integers n. This means the homomorphism is completely
determined θ(1). In fact it is either zero homomorphism or the identity map.

Definition 60 (image and kernel). Let θ : R → S be a ring homomorphism.

1. The image is im(θ) = {θ(a) : a ∈ R}.


2. The kernel is ker(θ) = {a ∈ R : θ(a) = 0}.

Lemma 3.15. If θ : R → S is a ring homomorphism, then im(θ) is a subring of S and ker(θ)


is an ideal of R. Moreover, θ is injective iff ker(θ) = {0} and surjective iff im(θ) = S.

Examples: Unit 7

• The units of Zn are precisely the elements coprime to n.


• Z2 is a field, Z4 is not.
• The group of units of Mn (R) are the matrices GLn (R).
• The units of the Gaussian integers, Z[i] are {1, −1, i, −i}.
• Let R be an integral domain, then R[x] is an integral domain.
• In a ring R both the trivial ring {0} and R itself are ideals.
• Z is a subring of R but it is not an ideal.
• In Z every subring is an ideal.
• The only ideals of a field are the trivial ideal and the field F .
• In the Boolean rings, if S and T are ideals then S ∩ T is also an ideal.
• Let R be a commutative ring with identity, then the set aR is an ideal and that hai = aR.
• Z is a principal ideal domain.

20
3.5 Factorization and Irreducibility
In this section we only work with integral domains.
Theorem 3.16 (The Fundamental Theorem of Arithmetic). Let n be an integer with n ≥ 2.
Then n can be expressed as a product of primes. Futhermore, if n = p1 · · · pr and n = q1 · · · qs ,
where pi , qi are all prime, then r = s and the qi s can be ordered so that pi = qi (1 ≤ i ≤ r).
Definition 61 (associates). Two elements r, s of an integral domain R are said to be associates
if there is a unit u such that s = ru.
Irreducibles are the analogue of prime numbers.
Definition 62 (irreducible). Let R be an integral domain, and p ∈ R with p 6= 0 and p not
a unit. We say that p is irreducible if p = ab (with a, b ∈ R) =⇒ either a or b is a unit.
Otherwise, we say p is irreducible.
Definition 63 (UFD). Let R be an integral domain. We say that R is a Unique Factorization
Domain if all elements other than zero and units can be factorized into irreducibles, and moreover
that if p1 · · · pm = q1 · · · qm where each pi and qj is irreducible, then m = n and the factors can
be reordered so that for 1 ≤ i ≤ m = n, pi and qi are associates.

3.6 Euclidean Domains


Lemma 3.17. Let a and b be integers and let p be a prime number. If p|ab, then p|a or p|b.
Definition 64 (Euclidean norm). Let R be an integral domain. A Euclidean norm on R is a
non-negative integer valued function N on the nonzero elements of R such that for all a, b ∈ R:
• if a, b 6= 0, then N (ab) ≥ N (a)
• if b 6= 0, then there exists q, r ∈ R with a = bq + r and either r = 0 or N (r) < N (b).
If R is equipped with such a function, then we say R is a Euclidean Domain
Theorem 3.18 (The Division Theorem for Polynomials). Let f (x) and g(x) be polynomials
over a field F , and suppose g(x) 6= 0. Then there exists polynomials q(x) and r(x) with f (x) =
q(x)g(x) + r(x) and either r(x) is the zero polynomial or deg(r(x)) < deg(g(x)).
Corollary 3.19. If F is a field, then F [x] is a Euclidean Domain.
Proposition 3.20. Let R be a Euclidean Domain. Then R is a Principal Ideal Domain.
Definition 65. Let R be an integral domain. For a, b ∈ R we say that d|a, or is a divisor of a,
is there is some c ∈ R for which cd = a.
Lemma 3.21. Let R be an integral domain. If two nonzero elements a and b have a greatest
common divisor d, then an element e of R is a gcd of a and b iff it is an associate of d.
Proposition 3.22. Any two nonzero elements a, b of a Principal Ideal Domain R have a greatest
common divisor d. Moreover there exists r, s ∈ R such that d = ar + bs.
Proposition 3.23. Any two nonzero elements a, b of a Euclidean Domain R have a greatest
common divisor d. Moreoever, there exist r, s ∈ R such that d = ar + bs.
Lemma 3.24. Suppose R is a Euclidean Domain. Let p ∈ R be irreducible. If p|ab, then p|a or
p|b.
Theorem 3.25. Suppose that R is a Euclidean Domain, and that, for some a ∈ R we have
a = p1 p2 · · · pm and a = q1 q2 · · · qn for some irreducibles pi and qj . Then m = n and the factors
can be reordered so that for 1 ≤ i ≤ m = n, pi and qi are associates.
Lemma 3.26. Let a, b, c be nonzero elements of a Euclidean Domain. If a = bc and N (b) =
N (a), then c is a unit.

21
Theorem 3.27. Every Euclidean Domain is a UFD.

Theorem 3.28. Let R be a Principal Ideal Domain. Then R is a UFD.

Every Euclidean Domain is a Principal Ideal Domain and every PID is a UFD.

Lemma 3.29. Let F be a field, f (x) a polynomial in F [x], and suppose a ∈ F is a root of f (x).
Then (x − a) divides f (x) in F [x].

Theorem 3.30. Let F be a field and f (x) a nonzero polynomial of degree n in F [x]. Then f
has at most n distinct roots in F .

Examples: Unit 8

• If R is a field, then there are no irreducibles, because every element is either zero or a unit.
• The units of Z are ±1. The irreducibles are p and −p for each prime p.
• In Z, N (a) = |a|.
• The ring of Gaussian integers is a Euclidean domain. N (a) = m2 +n2 for a = m+in ∈ Z[i].
• If R is a commutative ring with a, b ∈ R then the set {ar + bs : r, s ∈ R} is an ideal

22
4 Fields
4.1 Fields
Definition 66 (Field). A field F is a set with two binary operations called addition and mul-
tiplication that satisfy axioms F 1 to F 3 below. For a, b ∈ F addition is denoted a + b and
multiplication is denoted ab or a · b. We let F ∗ denote F − {0}.

F1 (F, +) is an abelian group, with identity called 0.


F2 (F ∗ , ·) is an abelian group, with identity called 1.
F3 For all a, b, c ∈ F , (a + b)c = ac + bc and c(a + b) = ca + cb (distributive laws).

Definition 67 (subfield-extension). Let F be a field and let K ⊆ F . If the field axioms are
satisfied for K, then we say that K is a subfield of F , and that F is an extension of K.

Definition 68 (characteristic). Let F be a field. If there is a positive integer n such that n1 = 0,


then the characteristic of F is defined to be char(F ) = min{n ∈ Z+ : n1 = 0}. Otherwise, we
define char(F ) to be zero.

Lemma 4.1. Let F be a field. Then either char(F ) = 0 or char(F ) is prime.

Proposition 4.2. The order of a finite field F is pk for some prime p and positive integer k
and every non-zero element of F had (additive) order p.

It can be shown that for any prime p and positive integer n, there exists a field of order pn .
It can also be shown that two finite fields of the same order are isomorphic. The unique field of
order pn is called the Galois Field of order pn , and is denoted by GF (pn ).

Lemma 4.3. Every finite integral domain is a field.

Definition 69 (exponent). Let G be a finite group. The exponent exp(G) is the least common
multiple of the orders of the elements of G.

Lemma 4.4. Suppose G is a finite abelian group. Then G contains an element of order exp(G).

Theorem 4.5. Let F be a field. Then any finite subgroup of the multiplicative group of F (F ∗ )
is cyclic.

Corollary 4.6. The multiplicative group of any finite field is cyclic.

Corollary 4.7. Every finite subgroup of (C∗ , ×) is cyclic.

23
4.2 Vector Spaces
Definition 70 (vector space). Let F be a field. A vector space over F is a set of V of elements
called vectors satisfying the following axioms:
(V1) This is a map +, known as addition on V , such that u + v ∈ V for every u, v ∈ V .
Furthermore, (V, +) is an abelian group (with identity called 0).
(V2) For every a ∈ F and v ∈ V , there is an element av ∈ V (closure under scalar multiplica-
tion).
(V3) a(bv) = (ab)v for all a, b ∈ F and v ∈ V (associativity of scalar multiplication).
(V4) 1v = v for all v ∈ V .
(V5) a(u + v) = au + av and (a + b)v = av + bv for all a, b ∈ F and u, v ∈ V (distributivity).
A p-ary word is a string of elements of Zp . Words over Z3 are sometimes called ternary
words.
Lemma 4.8 (Properties of 0 and 0). Let V be a vector space over F . Then for all a ∈ F and
all v ∈ V
(a) 0v = 0
(b) a0 = 0
(c) If av = 0 then either a = 0 of v = 0.
Definition 71 (subspace). Let V be a vector space over F . Then a subset U of V is a subspace
of V if U is a vector space over F (with the same addition and scalar multiplication as V ).
Lemma 4.9 (The subspace criterion). Let V be a vector space over F and U a nonempty set
of V . Then U is a subspace of V if and only if au + bv ∈ U for every a, b ∈ F and u, v ∈ U .
Definition 72 (sum of U + V ). Let U and W be subspaces of a vector space V . The sum of
U and W , denoted U + W , is the set

U + W = {u + w; u ∈ U, w ∈ W }

Write A = U + W . If it happens that U ∩ W = {0} then we say that A is the direct sum of U
and W , and write A = U ⊕ V .
Proposition 4.10. Let U and W be subspaces V . Then U + W is a subspace of V .

4.3 Vector Spaces: Bases, Independence, Dim


Definition 73 (Linear Combination). Let V be a vector space over F , and v1 , v2 , . . . , vn be
elements of V . We say that a vector v in V is a linear combination of {v1 , . . . , vn } if there
exist a1 , . . . , an ∈ F such that v = a1 v1 + · · · + an vn . The set of all linear combinations of
{v1 , . . . , vn } is referred to as the span of {v1 , . . . , vn } and denoted hv1 , . . . , vn i or sphv1 , . . . vn i.
Thus, ( n )
X
hv1 , . . . , vn i = ai vi : a ∈ F
i=1
The set {v1 , . . . , vn } is referred to as a spanning set of hv1 , . . . , vn i. If we have a set S whose
elements
P we don’t wish to list (perhaps it is infinite), then we can talk of linear combinations
a
w∈S w w, and refer to hSi for the span of S.
Proposition 4.11. Let V be a vector space over F , and S a subset of V . Then hSi is a subspace
of V , known as the subspace spanned by S
Definition 74 (linear dependence). Let S be a set of nonzero vectors. Then S is linearly
dependent if for some positive integer n there exist distinct vectors v1 , v2 , . . . , vn ∈ S and
scalars a1 , a2 , . . . , an , not all zero, such that

a1 v1 + a2 v2 + · · · + an vn = 0

24
If, on the other hand, for all n ∈ Z+ , distinct v1 , v2 , . . . , vn ∈ S and scalars a1 , a2 , . . . , an ,
the relationship a1 v1 + a2 v2 + · · · + an vn = 0 implies a1 = · · · = an = 0, then S is linearly
independent. By convention the empty set is linearly independent.

Definition 75. Let V be a vector space, and B a subset of V . Then B is a basis of V is B


is a linearly independent spanning set for V . We say that V is finite dimensional if it has a
finite basis.

Theorem 4.12. Let V be a vector space over a field F and S a finite subset of V . Then there
is a subset B of S which is linearly independent such that hBi = hSi.

Corollary 4.13. Let V be a vector space and S a spanning set of V . Then there is a subset B
of S which is a basis of V .

Theorem 4.14. Let V be a finite dimensional vector space and S = {v1 , . . . , vm } be a linearly
independent subset of V . Then either S is a basis of V or there exist vectors vm+1 , . . . , v, for
some n, such that {v1 , . . . , vm , vm+1 , . . . , vn } is a basis of V .

Definition 76 (dimension). The dimension of a finite dimensional vector space if the size of
any basis for the vector space.

Theorem 4.15. Let V be a finite dimensional vector space. If B and C are bases of V , with
C finite, then B and C have the same size. Hence any two bases of a finite dimensional vector
space have the same size.

Lemma 4.16. Let V be a finite dimensional vector space, and S ⊆ V . If S is linearly indepen-
dent, then |S|≤ dim V . If hSi then |S|≥ dim V .

Proposition 4.17. Let U be a subspace of a finite dimensional vector space V . Then U is finite
dimensional and dim U ≤ dim V , with equality ⇐⇒ U = V .

Theorem 4.18. If U and W are subspaces of a finite dimensional vector space V , then

dim(U + W ) = dim U + dim W − dim(U ∩ W )

4.4 Linear Transformations


Definition 77 (linear transformation). Let V and W be vector spaces over the same field F . A
map f : V → W is called a linear transformation if for all u, v ∈ V and scalars a, b we have
f (au + bv) = af (u) + bf (v). We say that V and W are isomorphic if there exists a linear
transformation between them that is also a bijection.

Lemma 4.19. Let f : V → W be an isomorphism of vector spaces over a field F . Then f


is invertible, and f −1 is also an isomorphism. Hence if V is isomorphic to W , then W is
isomorphic to V .

Definition 78 (image and kernel). Let f : V → W be a linear transformation of vector spaces


over a field F . The image of f is the set im(f ) = {f (v) : v ∈ V }. The kernel of f is the set
ker(f ) : {v ∈ V : f (v) = 0}.

Proposition 4.20. Let f : V → W be a linear transformation of vector spaces over a field F .


Then im(f ) is a subspace of W and ker(f ) is a subspace of V . Furthermore, f is surjective
⇐⇒ im(f ) = W and injective ⇐⇒ ker(f ) = {0}.

Definition 79 (rank, nullity). The rank of a linear transformation f : V → W is the dimension


im(f ) of its image. The nullity of f is the dimension of its kernel ker(f ).

Theorem 4.21. Let V be a finite-dimensional vector space over a field F , and f : V → W be


a linear transformation. Then rank(f ) + nullity(f ) = dim V .

25
Lemma 4.22. Let V be a vector space over a field F with a basis B = {v1 , v2 , . . . , vn }. Then
every element v ∈ V can be expressed in the form
n
X
v= a i vi
i=1

for some ai ∈ F , and this expression is unique. So, if we also have


n
X
v= a0i vi
i=1

then ai = a0i for i = 1, 2, . . . , n.

Corollary 4.23. Let p be prime, and let V be a finite dimensional vector space over Zp . Then
dim V = k ⇐⇒ |V |= pk .

Theorem 4.24. Let V and W be finite dimensional vector spaces over the same field F . Then
V is isomorphic to W ⇐⇒ dim V = dim W .

Theorem 4.25. There is a bijective correspondence between n × m matrices over a field F and
linear transformation from F n to F m .

Theorem 4.26. Let f : U → V and g : V → W be linear transformations, and suppose


dim U = I, dim V = m and dim W = n. Choose bases for U, V and W . Let P be the matrix
representing f and Q the matrix representing g with respect to these bases. Then the matrix
representing g ◦ f is P Q.

26

You might also like