Sei sulla pagina 1di 12

MAT3200-ABSTRACT ALGEBRA 2017

Lecture 2: December
Lecturer: Dr C. Chileshe Distance Class

2.1 Introduction To Group Theory

Towards the end of the previous Lecture 1, we touched on the 4 properties that a set can satisfy to be
classified as a special kind of set. We formally study such structures endowed with the 4 properties and
possibly more. Since we already studied properties around Sn . We first study further results about Sn . In
this way, we begin with the concept of equivalence relation on a non-empty set X.

Definition 2.1 A relation ∼ on a non-empty set X that is reflexive, symmetric and transitive is called an
equivalence relation. Thus for all x, y, z ∈ X, we obtain the following

(a) x ∼ x, called the reflexive property,


(b) if x ∼ y then y ∼ x called the symmetric property,
(c) if x ∼ y and y ∼ z, then x ∼ z, called the transitive property.

Definition 2.2 Suppose that ∼ is an equivalence relation on X, then the equivalence class of an element
x ∈ X is the set of all elements that are equivalent to x and is denoted by

[x] = {y ∈ X | y ∼ x}.

An even more important result is the following Theorem 2.3 that stresses that an equivalence relation on a
set X defines a partition of X by means of equivalence classes.

Theorem 2.3 Suppose ∼ is an equivalence relation on a set X, then we have that

(a) if x ∼ y, then [x] = [y],


(b) if x 6∼ y then [x] ∩ [y] = ∅,
(c) The collection of (distinct) equivalence classes is a partition of X into non-empty sets.

Proof: Exercise for the student.

Definition 2.4 Consider the set Sn . We say that α, β ∈ Sn are conjugate if there exists another permutation
γ ∈ Sn such that β = γαγ −1 .

Remark 2.5 The relation x ∼ y in Definition 2.4 is an equivalence relation on Sn . The equivalence classes
are called conjugacy classes of Sn and are denoted by [x] for x ∈ Sn . That is [x] = {gxg −1 |g ∈ Sn }.

Since to this point we are conversant with the concept of permutations, we get into it and see how the set
breaks into conjugacy classes.

2-1
2-2 Lecture 2: December

Example 2.6 If β = (4 6 2)(3 1) and α = (2 3 1)(4 5) are in S6 . Use Definition 2.4 to find an element
g ∈ S6 such that gαg −1 = β .
Working
Set g = (1 2 4 3 6 5) and notice that gαg −1 = (g(2) g(3) g(1))(g(4) g(5)) = (4 6 2)(3 1) = β.

Remark 2.7 The number of conjugacy classes of Sn equals the number of partitions of n. And from Lecture
n!
1, students are aware that the number of conjugates of an r − cycle, say ρ ∈ Sn , is given by
(n − r)r
and generally if ρ ∈ Sn can be written in terms of ρ1 1 − cycles, ρ2 2 − cycles · · · ρk k − cycles such that
ρ1 + 2ρ2 + 3ρ3 + · · · kρk = n. That is ρ = 1ρ1 2ρ2 · · · k ρk . Then the number of elements conjugate to ρ is
n!
|[ρ]| = Qk
i=1 iρi (ρi !)

Example 2.8 Consider the group S4 (recall that this is so much one example of a group). Let α = (1 2 3)
and β = (1 2)(3 4). Find

(a) number of conjugates of α in S4 and list them,


(b) number of conjugates of β in S4 and list them,
(c) denote by CS4 (β), the elements in S4 that commute with β. Compute this set.

Working

4! 24
(a) The number of conjugates of α in S4 is given by |[α]| = = = 8 and the set could be obtained as
1!3 3
[α] = {(1 2 3), (1 3 2), (1 2 4), (1 4 2), (2 4 3), (2 3 4), (1 4 3), (1 3 4)}. (Check).

Parts (b) and (c) are left as exercises for the student.

Exercise 2.9 Compute the conjugacy class representatives of S5 and obtain their respective sizes.

Much as Sn is an important group (every other group can be embedded in this group, a point that will
become clearer in the course of our lectures), it is not the only group in the whole universe. We had to play
inside it since in a way it is tangible. We next generalise results on any structure that is qualified to be
called a group. Formally

Definition 2.10 A group is a non-empty set G, together with an operation ∗ : G × G → G, that assigns to
each ordered pair (a, b) of elements in G an element a ∗ b ∈ G, that satisfies the following properties

(a) Associativity: (a ∗ b) ∗ c = a ∗ (b ∗ c) for all a, b, c ∈ G,


(b) Identity: There is an element e ∈ G (called the identity) such that a ∗ e = e ∗ a = a for all a ∈ G.
(c) Inverses: For each element a ∈ G, there exists an element b ∈ G such that a ∗ b = b ∗ a = e. The
element b is denoted by a−1 .

By juxtaposition, we will use ab for a ∗ b.


Lecture 2: December 2-3

Definition 2.11 The number of elements of a group (finite or infinite) is called the order of the group. We
denote this by |G|.

Theorem 2.12 For each element a in a group G, there is a unique element b ∈ G such that ab = ba = e.

Proof: Suppose b and c are both inverses of an element a. Then we have that b = be = b(ac) = (ba)c =
ec = c. Hence b = c, thus inverses are unique.
The following cancellation property is very useful in the study of groups. We go on and introduce more
properties about permutations and the next one (Theorem 2.13 ), in a way guarantees a distinguished
property of the identity permutation. That is 1Sn is the only permutation such that 1Sn α = α1Sn = α for
all α ∈ Sn .

Theorem 2.13 In a group G, if ba = ca, then b = c. Similarly, if ab = ac, then b = c.

Proof: Suppose ba = ca, then

(ba)a−1 = (ca)a−1
=⇒ b(aa−1 ) = c(aa−1 )
=⇒ be = ce
=⇒ b = c.

Left cancellation can be proved in a similar manner.

Definition 2.14 For an element a of a group G, the smallest positive integer m such that am = e is called
the order of a. If no such integer exists then a is said to have infinite order. We will denote order of a by
ord(a).

Theorem 2.15 Let a be an element of a group G and k an integer such that ak = e, then ord(a) divides k.

Proof: Let m = ord(a), then we use the Division Algorithm (students may benefit by checking what this
algorithm says even if they are promised to meet it in the study of integers) to conclude that there exist
q, r ∈ Z such that k = qm + r, where 0 ≤ r < m. Thus e = ak = aqm+r = (aqm )ar and since r < ord(a), we
have that r = 0.

Definition 2.16 The Cayley table of a finite group G is a tabulation of the values of the operation ∗. Let
G = {g1 , g2 , · · · , gn }. The Cayley table of G is given in Table 2.1

As a consequence of Definition 2.10, we get the following conclusions from Table 2.1.

(a) Each gk ∈ G occurs exactly once in each row of the table.

(b) Each gk ∈ G occurs exactly once in each column of the table.

(c) If the (i, j)th entry of the table is equal to the (j, i)th entry then gi ∗ gj = gj ∗ gi .

(d) If the table is symmetric about the diagonal & then g ∗ h = h ∗ g for all g, h ∈ G.
2-4 Lecture 2: December

∗ g1 g2 ··· gj ··· gn
g1
g2
..
.
gi gi ∗ gj
..
.
gn

Table 2.1: Cayley Table for any Finite Group G

o (1) (1 2) (1 3) (2 3) (1 2 3) (1 3 2)
(1) (1) (1 2) (1 3) (2 3) (1 2 3) (1 3 2)
(1 2) (1 2) (1) (1 3 2) (1 2 3) (2 3) (1 3)
(1 3) (1 3) (1 2 3) (1) (1 3 2) (1 2) (2 3)
(2 3) (2 3) (1 3 2) (1 2 3) (1) (1 3) (1 2)
(1 2 3) (1 2 3) (1 3) (2 3) (1 2) (1 3 2) (1)
(1 3 2) (1 3 2) (2 3) (1 2) (1 3) (1) (1 2 3)

Table 2.2: Cayley Table for S3

Example 2.17 Obtain the Cayley table for S3 .


Working
We know that S3 = {(1), (1 2), (1 3), (2 3), (1 2 3), (1 3 2)}. Thus we obtain the following Cayley Table 2.3.
Note that S3 is not abelian. In general Sn is not abelian for n ≥ 3.

We can now boldly say that for the examples of groups we will discuss in the rest of the course and those
others we will encounter during the rest of our lives, the following stands true and is very vital: if the set
satisfies the properties of a group then

(1) inverses are unique and,


(2) cancellation property holds.

Example 2.18 (Further Examples of Groups)

(a) The set of integers Z, the set of rational numbers Q, and the set of real numbers R are all groups under
addition. The identity element in these groups is 0. The inverse of an arbitrary element say a in any
of these groups is −a.
(b) The set Q∗ = {s ∈ Q|s 6= 0} is a group under ordinary multiplication. The identity element is 1 and
1
the inverse of an arbitrary element s is .
s
(c) For each (xo , yo ) ∈ R2 , define T(xo ,yo ) : R2 → R2 by (x, y) → (x + xo , y + yo ). Then one can show that
the set T (R2 ) = {T(xo ,yo ) | xo , yo ∈ R} is a group under function composition. (Check)
(d) Let {f1 , f2 , f3 , f4 , f5 , f6 } be a set of functions R − {0, 1} → R − {0, 1} under composition. Where
1 1
f1 (x) = x, f2 (x) = 1 − x, f3 (x) = , f4 (x) = 1 − ,
x x
1 x
f5 (x) = , f6 (x) = .
1−x x−1
Lecture 2: December 2-5

This set forms a group (Check).

2.1.1 Subgroups

Definition 2.19 If G is a group and H is a subset of G which is also a group under same operation, then
we say H is a subgroup of G, and we write H < G.

Theorem 2.20 Let G be a group and H a non-empty subset of G. If

(a) for every a, b ∈ H, ab ∈ H (closed under multiplication),

(b) for every a ∈ H, a−1 ∈ H (closed under inverses),

then H is a subgroup of G.

Proof: We need to show that with the given conditions, all the 4 properties in a group are satisfied. Note
that the operation is binary on the set H as given in the statement of the theorem and that H is closed
under inverses. We are left to show existence of identity element in H as well as associativity. Since H is
non-empty, it contains atleast one element say h and h−1 ∈ H, so that by close h−1 h which is the identity
in H. Now let h1 , h2 , h3 ∈ H. Then (h1 h2 )h3 = h1 (h2 h3 ) in G and therefore also in H. Hence all the four
properties are satisfied. Thus H is a subgroup of G.

Example 2.21 Consider S3 = {(1, (1 2), (1 3), (2 3), (1 2 3), (1 3 2)}. Let H =< (1 2 3) >=
{(1), (1 2 3), (1 3 2)}. Check whether this is indeed a subgroup of S3 .

Working
Notice that closure holds in H since we obtain the following Cayley table Furthermore, inverses exist and

o (1) (1 2 3) (1 3 2)
(1) (1) (1 2 3) (1 3 2)
(1 2 3) (1 2 3) (1 3 2) (1)
(1 3 2) (1 3 2) (1) (1 2 3)

Table 2.3: Cayley Table for H

they are in H. This concludes (by Theorem 2.20) that H is a subgroup of S3 .

Definition 2.22 The centre of a group G is the subset Z(G) of G consisting all elements that commute with
every element of G. That is Z(G) = {a ∈ G | ag = ga, for all g ∈ G}.

Exercise 2.23 Prove that for any group G, the centre Z(G) is a subgroup of G.

2.1.2 Cyclic Groups

Definition 2.24 A group is said to be cyclic if there is an element in G, say g, such that every other element
of G is a power of g. That is G = {g k | k ∈ Z}. We write G =< g > and say g is a generator for G. If g
has order n then G = {e, g, g 2 , g 3 , · · · , g (n−1) } and we say G is a cyclic group of order n.
2-6 Lecture 2: December

Exercise 2.25 Prove that any cyclic group G is abelian. (By abelian group, we mean a group G in which
elements commute).

Definition 2.26 Let n > 1 be an integer. define an operation on the set Zn = {0, 1, 2, 3, · · · , n − 1} (this
set is called the set of integers modulo n) as follows: For a, b ∈ Zn , let a +n b be the remainder of a + b when
divided by n. This way Zn forms a group called the group of integers modulo n. Since it is cyclic, it is also
called the additive group of order n.

The group in Definition 2.26 is also denoted by Z/nZ read as Z mod nZ.

Remark 2.27 The group Zn is cyclic.

Exercise 2.28 Construct the Cayley table for Z7 . Obtain the inverse and order of every element in Z7 .

2.1.3 Dihedral groups

Let G denote the set of ways in which a square can be picked, moved in some way and put back in the
original space it occupied without changing the first position. The following illustration shows this number
of ways:

a b

1 2

4 3
d

In the figure above consider first the clockwise rotations; R0 , that is rotation of 0 degrees (do nothing)
equivalent to (1). R90 , a rotation of 90 degrees equivalent to (1 2 3 4), R180 , a rotation of 180 degrees
equivalent to (1 3)(2 4), R270 , a rotation of 270 degrees equivalent to (1 4 3 2). Secondly, let us consider the
following reflections: reflection A in the line a equivalent to (1 2)(3 4), reflection B in the line b equivalent
to (1 3), reflection C in the line c equivalent to (1 4)(2 3), reflection D in the line d equivalent to (2 4). It is
routine to check that the set D8 = {R0 , R90 , R180 , R270 , A, B, C, D} forms a group. This group is called the
dihedral group of order 8. This analysis done for a square could be extended to an n − gon, where n ≥ 3.
Dihedral groups are frequently used by mineralogists to study crystals. (Though we study them here for the
beauty of mathematics!).
In general, write D2n =< a, b : an = b2 = e, b−1 ab = a−1 >, to define the dihedral group of order 2n.

2.1.4 Conjugacy Classes of Dihedral Groups

Theorem 2.3 could be used to get the conjugacy classes of D2n , we leave out the technicalities of the process
and refer students to [GM2001] for more details. The result is stated as below:
Lecture 2: December 2-7

Theorem 2.29 Let D2n =< a, b : an = b2 = e, b−1 ab = a−1 >. If n is odd, D2n has exactly 1
2 (n + 3)
(n−1) (n−1)
conjugacy classes namely, {e}, {a, a−1 }, · · · , {a 2 , a − 2
}, {b, ab, · · · , an−1 b}.
If n is even, that is n = 2m, then D2n has exactly m+3 conjugacy classes namely, {e}, {am }, {a, a−1 }, · · · ,
{am−1 , a−m+1 }, {a2j b : 0 ≤ j ≤ m − 1}, {a2j+1 b : 0 ≤ j ≤ m − 1}.

Proof: See [GM2001].

Example 2.30 Compute the conjugacy classes of D8 .

Working
Note that 8 = 2 × 4. Thus m = 2. By Theorem 2.29, we seek m + 3 = 2 + 3 = 5 conjugacy classes namely,
{e}, {a2 }, {a, a−1 }, {b, a2 b}, {ab, a3 b}.

Exercise 2.31 (a) Let G = {a, b, c, d} have an operation ∗ with corresponding Cayley table

∗ a b c d
a a b c d
b b a d c
c c d a b
d d c b a

Table 2.4: Cayley Table for G

Justify whether G is a group or not under this operation.

(b) Write out a complete multiplication table for D6 .

(c) For any elements a and b of a group G prove that (b−1 ab)n = b−1 an b where n is any integer.

(c) Let a and b be elements of an abelian group G and let n be any integer. Show that (ab)n = an bn . Does
this hold, in general, for non abelian groups.

(d) Prove that if G is a group with property that the square of every element is the identity, then G is
abelian.

We will in the following Section 2.1.5 consider a result that compares the order of the group and the order
of respective subgroups. This result has several other applications as will be noted in a moment.

2.1.5 Lagrange’s Theorem

We begin by discussing the concept of cosets as this is required in the proof of Lagrange’s theorem. The
concept of cosets emanates right from the process of defining an equivalence relation on a group say G with
reference to a certain subgroup say H.
Let H < G. Define a relation on G as follows: Two elements a, b ∈ G are related ( a ∼H b) if and only if
a−1 b ∈ H. Equivalently, say a ∼H b if and only if b = ah for some h ∈ H. Please convince yourself that this
relation forms an equivalence relation on G and that [a] = {ah | h ∈ H}. We therefore have the following
Definition 2.32 as a consequence.
2-8 Lecture 2: December

Definition 2.32 Let G be a group and H a subgroup of G. For any a ∈ G, the set aH = {ah | h ∈ H} is
called the left coset of H in G containing a. The left coset of H in G is given by Ha = {ha | h ∈ H}. The
element a is called the coset representative.

Remark 2.33 We observe that the relation a ∼H b if and only if a−1 b ∈ H breaks G into cosets thus for
any two left cosets aH and bH, we either have aH = bH or aH ∩ bH = ∅.

Example 2.34 Let S3 = {(1), (1 2), (1 3), (2 3), (1 2 3), (1 3 2)} and H =< (1 2) >= {(1), (1 2)}. Find
the left cosets of H in G.

Working
The left cosets of H in G are (1)H = H = {(1), (1 2)}, (1 3)H = {(1 3), (1 2 3)}, (2 3)H = {(2 3), (1 3 2)}.
That is the set {H, (1 3)H, (2 3)H}. Coset representatives are (1), (1 3) and (2 3).

Exercise 2.35 Consider G = Z12 and H =< 3 >. Find the right cosets of H in G.

The following Corollary 2.36 gives properties around cosets.

Corollary 2.36 Let G be a group and H a subgroup of G. For any a, b ∈ G, we have that

(a) a ∈ aH,
(b) aH = H if and only if a ∈ H,
(c) either aH = bH or aH ∩ bH = ∅,
(d) aH = bH if and only if a−1 b ∈ H if and only if b−1 a ∈ H,
(e) if H is finite, then |aH| = |H|,
(f ) aH = Ha if and only a−1 Ha = H.

Proof:

(a) By definition, we have that aH = {ah | h ∈ H}. Thus take h = e, then we have that ah = ae = a ∈ aH.
(b) Suppose that aH = H, then by (a), a ∈ aH = H and we are done. Conversely, suppose that a ∈ H,
then aH ⊂ H. Furthermore, if b ∈ H then by closure a−1 b ∈ H, thus b ∈ aH. Hence H ⊂ aH. This
establishes that aH = H.
(c) This follows when the student is convinced that the relation a ∼H b if and only if b = ah for some
h ∈ H forms an equivalence relation on G and the equivalence classes are given by [a] = {ah | h ∈ H}.
(d) Suppose that aH = bH, then we have that a−1 bH = H and using part (b) conclude that a−1 b ∈ H
and b−1 a ∈ H, the latter following since (a−1 b)−1 = b−1 a ∈ H.
(e) Define φ : H → aH by φ(h) = ah. This map is well defined and a bijection for the following reasons;
for h1 , h2 ∈ H, φ(h1 ) = φ(h2 ) if and only if ah1 = ah2 for a ∈ G if and only h1 = h2 . This shows the
map is well defined and 1 − 1. For the assertion that it is surjective, let b ∈ aH. By the equivalence
relation on G, there is an h in H such that b = ah. Thus a−1 b ∈ H and φ(a−1 b) = b. This proves
that φ is surjective and finally we have that φ is a bijection. We conclude therefore that the two sets
H and aH have same cardinality. That is |H| = |aH|.
Lecture 2: December 2-9

(f ) Suppose that aH = Ha, then for any h ∈ H there is an element x ∈ H such that ax = ha, thus
a−1 ha ∈ H. Therefore a−1 Ha ⊂ H. Moreover, for any h ∈ H, there is a y ∈ H such that ah = ya
so that h = a−1 ya ∈ a−1 Ha. Therefore H ⊂ a−1 Ha. Hence H = a−1 Ha. Conversely, suppose
a−1 Ha = H, then for any h ∈ H there is an element x ∈ H such that a−1 xa = h, so that ah = xa ∈ Ha.
Therefore aH ⊂ Ha. A similar argument yields that Ha ⊂ aH. We therefore have that aH = Ha.

Having been able to define an equivalence relation on a group G to break it up into cosets, we state Lagrange’s
theorem and carefully understand its proof.

Theorem 2.37 (Lagrange’s Theorem)


If G is a finite group and H is a subgroup of G, then |H|||G|.

Proof: Let ∼H be an equivalence relation on G. The equivalence classes are the left cosets [a] = aH. Let
a1 H, a2 H, · · · , ak H denote distinct left cosets of H in G. Since |aH| = |H| for all a ∈ G, we have that
G = a1 H ∪˙ a2 H ∪˙ · · · ∪˙ ak H, (disjoint union). Therefore |G| = |a1 H| + |a2 H| + · · · + |ak H| = k|H|. Thus
|H|||G|.

Remark 2.38 Note that the converse of Lagrange’s theorem is not necessarily true. For instance, the order
of A4 is 12. We know that 6 divides 12 however A4 has no subgroup of order 6.

A nice application on the structure of the group is the next Corollary 2.39 which is a consequence of
Lagrange’s theorem

Corollary 2.39 Let G be a group. Let a ∈ G. The order of a divides the order of G. Furthermore, a|G| = e
identity in G.

Proof: Suppose that the order of a is k, then take H = {e, a2 , a3 , · · · , ak−1 } =< a >. Then |H| = k
and by Lagrange’s theorem, k | |G| since H < G. Furthermore, |G| = rk for some positive integer r. Hence
a|G| = ark = (ak )r = er = e, the identity element in G.
The above result shows more pattern in a group. Students could test this result on most of the groups they
have learnt so far just to convince themselves and be happy!

2.1.6 Homomorphisms

In group theory it is a more natural question as to which groups share same properties. If G and H are groups
and there is a map between them that preserves their properties then we would say the two groups are in a
way the same except for notation of the elements and operations in them. The word homomorphism has its
trace from the Greek word homo meaning ”same” and morph meaning ”shape” or ”form”. see also Rotman
[JJR03]. Thus homorphisms allow us to have a comparison of groups and in way also have a preferable choice
of study among groups. This opportunity paves way to representation theory which students will encounter
later in their study of algebra.

Definition 2.40 If (G, o) and (H, ∗) are groups, then a function ρ : G → H is a homomorphism if

ρ(xoy) = ρ(x) ∗ ρ(y) (notice the preserving of structure here) (2.1)

for all x, y ∈ G.
2-10 Lecture 2: December

Remark 2.41 If further ρ is also a bijection then we say that it is an isomorphism and in this case we write
G∼
= H to read G is isomorphic to H.

Example 2.42 Let R[x] denote the group of all polynomials with real coefficients and the operation usual
dg
addition of polynomials. Define ρ : R[x] → R[x] by ρ(g) = for all g ∈ R[x]. Show that ρ is a homomor-
dx
phism.
working
d(g + h) dg dh
Let g, h ∈ R[x]. Then ρ(g + h) = = + = ρ(g) + ρ(h).
dx dx dx

Exercise 2.43 Find all elements h ∈ R[x] such that ρ(h) = 0, the zero polynomial.

The above exercise motivates the following Definition 2.44 in general.

Definition 2.44 If ρ : G → H is a homomorphism, then the Kernel of ρ is given by Ker (ρ) = {g ∈


G | ρ(g) = eH } and the image of ρ is given by im (ρ) = {h ∈ H | ρ(g) = h}.

Exercise 2.45 (a) Let ρ : G → H be a homomorphism. Prove that Ker (ρ) is a subgroup of G and im (ρ)
is a subgroup of H (hint: use Theorem 2.20 ).
(b) Prove that ρ is one to one (injective) if and only if Ker(ρ) = {eG }.

(c) Prove that for all g ∈ G and k ∈ Ker(ρ) we have that gkg −1 ∈ Ker(ρ).

We have so far encountered subgroups, that is subsets that are in their own right also groups under same
operation. At this point, we introduce a special type of subgroup called a normal subgroup and they arise
as kernels of homomorphisms. Formally we have

Definition 2.46 Let G be a group. A subgroup H of G is said to be a normal subgroup of G if h ∈ H and


g ∈ G means that ghg −1 ∈ H. We usually write H E G.

The subgroup given in Definition 2.46 comes with its beauty in constructing other subgroups of G as will be
noticed in a moment. but first note that the kernel of any homomorphism is a normal subgroup. Students
should ensure that they fully understand why this is the case. They can simply use the definitions to prove
this fact.

Example 2.47 Let us consider S3 and H =< (123) >. Show that H E S3 .
working
We need to first show that H ≤ G, but this was done in Example 2.21 (whenever we are asked to show a
subset forms a normal subgroup of the main group, the starting point is to show that it is no mere subset but
a subgroup!). Now for the normality part, let x ∈ S3 = {(1), (1 2), (1 3), (1 2 3), (1 3 2)}. Then we show
x(1 2 3)x−1 ∈ H = {(1), (1 2 3), (1 3 2)}. This is easily seen in the following Table 2.5.
Lecture 2: December 2-11

x x(1 2 3)x−1
(1) (1 2 3)
(1 2) (1 3 2)
(1 3) (1 3 2)
(2 3) (1 3 2)
(1 2 3) (1 2 3)
(1 3 2) (1 2 3)

Table 2.5: x(1 2 3)x−1 elements

The algorithm to check normality was quite daunting, was it not? Well the fact is we have to learn the
definitions very well and then be quite smart to use the results. Here is a result that we could be smart to
use in some cases.

Corollary 2.48 (a) If G is a group and H is a subgroup of index 2 in G, then g 2 ∈ H for every g ∈ G.
(a) If G is a group and H is a subgroup of index 2 in G, then H is normal in G.

Proof:

(a) Note that since H is of index 2 in G, then G = H ∪˙ bH where b ∈ / H. Now let g ∈ G with
/ H. Then g = ah for some h ∈ H and g 2 = ah0 for some h0 ∈ H. Assume that g 2 ∈
g ∈ / H. But
g = g −1 g 2 = (ah)−1 (ah0 ) = h−1 a−1 ah0 = h−1 h0 ∈ H a contradiction. Hence g 2 ∈ H for all g ∈ G.
(a) We now exploit the definition of normailty here. Thus we are to show that for all g ∈ G and h ∈ H,
ghg −1 ∈ H. Now since H is of index 2 in G, we have that there are two cosets, namely H and bH with
b∈/ H. If g ∈ H, then ghg −1 ∈ H since H is a group in its own right. Assume that g ∈/ H, then g = bz
for some z ∈ H. Now ghg −1 = b(zhz −1 )b−1 = bh0 b−1 . Thus if ghg −1 ∈
/ H, then ghg −1 = bh0 b−1 ∈ bH.
This means that bh0 b−1 = bm for some m ∈ H. Thus bh0 = m and we have that b = m−1 h0 ∈ H, a
contradiction. Hence ghg −1 ∈ H and we have that H E G.

Remark 2.49 Using the result in Corollary 2.48, we can easily see that H =< (1 2 3) > has index 2 in S3
therefore it is normal.

Example 2.50 The alternating group An is normal in Sn by Corollary 2.48. Also note that An is the kernel
of the homomorphism sgn : Sn → {1, −1} which is such that sgn(α) = 1 if α is even and sgn(α) = −1 if α
is odd. That An is normal by this argument is exactly what the student has proved in Exercise 2.45 part (c).

Example 2.51 The group ofquaternions  is given by Q= {I, A, A2 , A3 , B, BA, BA2 , BA3 }, where I
0 1 0 i
is the identity element, A = and B = . The cyclic subgroups of Q are given by < A >
−1 0 i 0
and < −I > and students can verify they are normal in Q. In fact this is the only group that has all its
subgroups normal in it and it is called a Hamiltonian group due to W. R. Hamilton.

Exercise 2.52 (a) Let V = {(1), (12)(34), (13)(24), (14)(23)} be the Klein 4 − group and H =<
(12)(34) >. Show that H E V but H is not normal in S4 . (This gives a counter example that
normality is not transitive).
2-12 Lecture 2: December

(b) Prove that the intersection of any family of normal subgroups of a group say G is itself normal in G.

(c) Let G, H and K be groups. Let ρ1 : G → H and ρ2 : H → K be two homomorphisms. Prove that
ρ1 oρ2 is a homomorphism.
(d) Prove that a group G is abelian if and only ρ : G → G, given by ρ(a) = a−1 , is a homomorphism.
(e) Let ρ : G → H be a homomorphisms. Let a ∈ G such that order of a is m. Prove that ρ(a) has order
m. (Order is a group invariant: it is preserved by a group homomorphism).
(f ) Obtain the center of the quaternion group of order 8.
(g) Prove that the quaternion group of order 8 and the dihedral group of order 8 are non-isomorphic.
(h) An Automomorphism of a group G is an isomorphism G → G.

(1) Prove that Aut(G), the set of all automorphisms of a group G is a group under function compo-
sition.
(2) Prove that τ : G → Aut(G), defined by g 7→ τg (where τg : G → G is given by τg (h) = ghg −1 ), is
a homomorphism.
(3) Prove that ker(τ ) = Z(G).
(4) Prove that im(τ ) E Aut(G).
(i) Prove that if H is a subgroup such that aH = Ha for all a ∈ G, then H is normal in G.

Next we will discuss a construction of certain subgroups of a group G by considering its normal subgroups
and also exploiting Lagrange’s Theorem.

2.1.7 Factor Groups

References
[JJR95] J. J. Rotman, “An Introduction to the Theory of Groups,” Springer-Verlag New York, Inc,
1995.
[GM2001] G. James and M. Liebeck, “Representations and Characters of Groups,” Cambridge Uni-
versity Press 1993, 2001.
[JJR03] J. J. Rotman, “A First Course in Abstract Algebra,” Prentice Hall, 2003.

Potrebbero piacerti anche