Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

2 Basic group theory

2.1 The definition of a group

A group is a set G together with a function ∗ : G × G → G, assigning to each pair (a, b) of elements
of G another element a ∗ b ∈ G, satisfying the following three axioms:

G1 (associativity) We have a ∗ (b ∗ c) = (a ∗ b) ∗ c, for all a, b, c ∈ G.

G2 (existence of identity) There exists an element e ∈ G such that e ∗ a = a ∗ e = a for all a ∈ G.

G3 (existence of inverses) For all a ∈ G there exists an element a0 ∈ G such that a ∗ a0 = a0 ∗ a = e,


where e is an identity element as in axiom G2.

There is no requirement that a ∗ b = b ∗ a for all a, b ∈ G. If this property does hold, we say that G is
abelian.

The element e of axiom G2 is unique: for if e0 is another identity element, then e0 = e0 ∗ e = e by


applying axiom G2 first to e, then to e0 . For each a ∈ G, the inverse element a0 in axiom G3 is unique:
for if a00 is another inverse element, we have

a00 = a00 ∗ e = a00 ∗ (a ∗ a0 ) = (a00 ∗ a) ∗ a0 = e ∗ a0 = a0 ,

by applying successively axioms G2, G3 (for a0 ), G1, G3 (for a00 ) and finally G2 again.

We usually use multiplicative notation and abbreviate a ∗ b as ab or a · b, and write 1 or 1G instead of


e for the identity element of G, and a−1 instead of a0 for the inverse element of a. For any positive
integer n, we write an for the product of a with itself n times, and a−n for the product of a−1 with itself
n times. Finally, we put a0 = 1.

If G is abelian, and only in this case, we sometimes use additive notation, writing a∗b = a+b, denoting
the identity element by 0, and the inverse element of a by −a.

5
The order of G is the cardinality |G|, either a positive integer or ∞. Any group G of order one consists
of the identity element only and is called the trivial group.

We can extend the product of elements in G to subsets: If S and T are subsets of a group G, we define

ST = {st : s ∈ S, t ∈ T }.

2.2 Group homomorphisms

The structure of a group G is revealed by its subgroups and homomorphisms into other groups.

A homomorphism of groups G, G0 is a function f : G → G0 satisfying

f (ab) = f (a)f (b) ∀ a, b ∈ G.

This implies that f (1G ) = 1G0 and that f (g −1 ) = f (g)−1 for all g ∈ G.

The kernel of a homomorphism f : G → G0 is the subset of G defined by

ker f := {g ∈ G : f (g) = 1G0 }.

For all a, b ∈ G we have f (a) = f (b) if and only if ab−1 ∈ ker f . Hence f is injective iff ker f = {1G }.
The image of f is the set theoretic image, defined as above by

im f := f (G) = {g 0 ∈ G : f −1 (g 0 ) 6= ∅}.

There may be many homomorphisms between two given groups. We set

Hom(G, G0 ) = {homomorphisms f : G → G0 }.

An isomorphism f : G → G0 is a bijective group homomorphism. Thus f is an isomorphism if and



only if ker f = {1G } and im f = G0 . We sometimes write f : G −→ G0 to indicate that f is an

isomorphism. Two groups G, G0 are isomorphic if there exists an isomorphism f : G −→ G0 . We
write G ' G0 to indicate that G and G0 are isomorphic, without specifying any particular isomorphism
between them.

We sometimes abuse terminology and say that G is or is a copy of G0 , when we really mean only that
G ' G0 . For example, any two trivial groups are isomorphic, so we allow ourselves to say that the
trivial group is the unique group with one element.

Continuing in this vein, we say that a homomorphism f : G → G0 is trivial if im f = {1G0 }. This is


equivalent to having ker f = G, so being a trivial homomorphism is the opposite of being an isomor-
phism. Hence trivial homomorphisms are just as important as isomorphisms.

An isomorphism from a group to itself is called an automorphism. We set

Aut(G) = {automorphisms of G}.

6
The set Aut(G) forms a group under composition, whose identity element is the identity automor-
phism, which sends g 7→ g for all g ∈ G.

Many automorphisms arise from within the group itself as follows. For each element g in a group G,
the map
cg : G → G given by cg (x) = gxg −1 ∀x ∈ G
is an automorphism of G, called conjugation by g. Automorphisms of this form are called inner
automorphisms. The function c : G → Aut(G) sending g 7→ cg is a homomorphism.

2.3 Subgroups

A subgroup of G is a subset H ⊆ G with the following three properties:

SG1 (closure) ab ∈ H for all a, b ∈ H.


SG2 (identity) The identity element of G is contained in H.
SG3 (inverses) For all a ∈ H we have a−1 ∈ H.

The subsets {1} and G are subgroups of G. All other subgroups of G, if any, are called proper
subgroups. We write H ≤ G to indicate that H is a subgroup of G which is possibly equal to G itself.
We write H < G for a subgroup which is not equal to G.

Lemma 2.1 Let G be a group and let H be a nonempty finite subset of G. Then H ≤ G if and only if
SG1 holds.

Proof: Let h be an element of the nonempty set H. Since H is finite, the powers h, h2 , h3 , . . . must
eventually repeat, so we have hi = hj for some positive integers i < j. It follows that hj−i = 1, so
SG2 holds, and h · hj−i−1 , so SG3 holds. Hence H is a subgroup of G. 

This proof moved from H to a particular kind of subgroup of G. A group C is cyclic if there exists an
element c ∈ C such that C = {cn : n ∈ Z}. In this case we write C = hci.

In an arbitrary group G, any element g ∈ G is contained in the cyclic subgroup


hgi = {g n : n ∈ Z}.
The order of g is the order of the group hgi. The order of g is the smallest positive power m such that
g m = 1, if such an m exists. In this case, the order can be characterized by the useful property that for
any integer d, we have g d = 1 iff m | d. If g d 6= 1 for any nonzero integer d, we say the order of g is
infinite.

More generally, if S is any subset of G, the subgroup generated by S is the smallest subgroup hSi of
G containing S. That is, \
hSi = H
S⊆H≤G
is the intersection of all subgroups H of G which contain the subset S.

7
2.4 Cosets and quotient spaces

A left coset of a subgroup H < G is a subset of G of the form gH = {gh : h ∈ H}. Two left cosets
are either equal or disjoint; we have

gH = g 0 H ⇔ g −1 g 0 ∈ H.

In particular, we have gH = H if and only if g ∈ H. The set of left cosets of H in G is denoted G/H,
and is called the quotient of G by H.

A right coset of H in G is a subset of the form Hg = {hg : h ∈ H}. Two right cosets are either equal
or disjoint; we have
Hg = Hg 0 ⇔ g 0 g −1 ∈ H.
In particular, we have Hg = H if and only if g ∈ H. The set of right cosets of H in G is denoted
H\G.

A coset is a left or right coset. Any element of a coset is called a representative of that coset. We have
canonical bijections H → gH and H → Hg, sending h 7→ gh and h 7→ hg, respectively. Hence if H
is finite, all cosets have cardinality |H|.

There are an equal number (including infinity) of left and right cosets in G. We denote this number by

[G : H] = |G/H| = |H\G|,

and call it the index of H in G. If G is finite, then G is partitioned into [G : H] cosets, each of
cardinality |H|. It follows that [G : H] = |G|/|H|. In particular we have

Lagrange’s Theorem: If H is a subgroup of a finite group G, then |H| divides |G|.

Example 1: If |G| = p a prime, then G has no proper subgroups. Hence for any nonidentity element
g ∈ G we have G = hgi, so G is cyclic.

Example 2: The order of any element in a finite group G divides |G|. in particular, we have g |G| = 1
for all g ∈ G. The smallest positive integer e such that g e = 1 for all g ∈ G is called the exponent of
G. If g ∈ G has order m, then m divides e, which in turn divides |G|.

The converse of Lagrange’s theorem is false. The smallest counterexample is the group A4 of order 12,
which has no subgroup of order 6. However, the converse of Lagrange’s theorem is true for subgroups
H of prime power order. This is part of the Sylow theorems, which we will prove later. However, one
special case is easy:

Proposition 2.2 Any group of even order contains an element of order two.

Proof: Suppose G has even order |G| = 2m. Pair the nonidentity elements with their inverses. Since
there are 2m − 1 such elements, at least one of them is paired with itself. This is a nonidentity element
g ∈ G such that g = g −1 . Thus, g is an element of G of order two. 

8
2.5 Normal subgroups and quotient groups

Let G be a group and let H ≤ G be a subgroup. One attempts to define a group structure on the set
G/H by the rule:
gH ∗ g 0 H = gg 0 H, ∀g, g 0 ∈ G. (1)
However, this rule is only well-defined when every left coset of H in G is also a right coset of H in G.
The subgroup H is said to be normal in G if gH = Hg for all g ∈ H. On the level of elements, this
means that ghg −1 ∈ H for all g ∈ G and h ∈ H. If G is abelian, then ghg −1 = h, so every subgroup
is normal in G. Thus, being normal in G is a weakening, with respect to H, of the abelian condition.
We write H / G or H E G to indicate that H is a normal subgroup of G.

When, and only when H E G, the set G/H = H\G becomes a group under the operation given by
(1). We call G/H the quotient of G by H. It is a group of order equal to the index of H in G:
|G/H| = [G : H].

Example: The center of a group G is the subgroup


Z(G) = {z ∈ G : zg = gz ∀g ∈ G}.
This is clearly a normal subgroup of G. We will see the quotient group G/Z(G) appearing in several
contexts. One useful fact is

Proposition 2.3 If G/Z(G) is cyclic then G is abelian.

Proof: Exercise. 

2.6 The first isomorphism theorem

Any group homomorphism f : G → G0 induces an isomorphism from a quotient of G to a subgroup


of G0 . More precisely, we have the following.

Theorem 2.4 (First isomorphism theorem) Let f : G → G0 be a group homomorphism with kernel
K = ker f . Then K is a normal subgroup of G, and there is an isomorphism

f¯ : G/K −→ im f, given by f¯(gK) = f (g). (2)

Proof: It is a good exercise to check that f¯ is well-defined and bijective. 

It is useful to have this result in slightly more general form:

Theorem 2.5 Let f : G → G0 be a group homomorphism with kernel K = ker f . Let H be a normal
subgroup of G contained in K. Then there is a surjective homomorphism
f¯ : G/H  im f, given by f¯(gH) = f (g), (3)
with ker f¯ = K/H.

9
Proof: Again, this is a good exercise. 

We often say that f¯ is induced by f or that f factors through G/H.

Conversely, every normal subgroup H E G is the kernel of a surjective homomorphism from G into
another group. Namely, the canonical homomorphism

πH : G −→ G/H, given by πH (g) = gH

is surjective with ker πH = H.

Example: If x, y are two elements of G, the commutator

[x, y] = xyx−1 y −1

is an element of G that measures the failure of x, y to commute. The Commutator Subgroup

[G, G] = h[x, y] : x, y ∈ Gi

is the subgroup generated by all commutators. For g ∈ G we have g[x, y]g −1 = [gxg −1 , gxg −1 ]. It
follows that [G, G] E G. The quotient G/[G, G] is called the abelianization of G, and is often denoted
Gab , because of the following result.

Proposition 2.6 The quotient G/[G, G] of a group G by its commutator subgroup is abelian. More-
over, G/[G, G] is the largest abelian quotient of G, in the following sense: If f : G → A is a homo-
morphism from G to an abelian group A, then [G, G] < ker f , so f factors through a homomorphism

f¯ : G/[G, G] −→ A.

Proof: Exercise. 

2.6.1 Exact sequences

A composition of group homomorphisms


f1 f2
G1 −→ G2 −→ G3

is exact at G2 if im f1 = ker f2 . A sequence of group homomorphisms


fi−1 fi
. . . −→ Gi−1 −→ Gi −→ Gi+1 −→ . . .

is an exact sequence if it is exact at Gi for all i. A short exact sequence is a sequence


f1 f2
1 −→ G1 −→ G2 −→ G3 −→ 1

with
ker f1 = {1}, im f1 = ker f2 ' G1 , im f2 = G3 ' G2 /G1 .

10
2.7 The second isomorphism theorem (correspondence theorem)

The subgroups of a quotient group G/H are related to subgroups of G as follows.

Theorem 2.7 Let G be a group with normal subgroup H E G, and let πH : G → G/H be the
canonical homomorphism.

1. If K is any subgroup of G containing H, then H E K and K/H = πH (K) is a subgroup of


G/H.
−1
2. Conversely, if J is any subgroup of G/H, then πH (J) is a subgroup of G containing H as a
−1
normal subgroup and J = πH (J)/H.

3. K/H E G/H if and only if K E G, in which case (G/H)/(K/H) ' G/K.

Thus, we have a one-to-one correspondence

{subgroups of G containing H} ↔ {subgroups of G/H}


K → K/H
−1
πH (J) ← J,

and this correspondence preserves normal subgroups.

There is also a correspondence theorem for homomorphisms, the first part of which is just Thm. 2.5
above.

Theorem 2.8 Let G be a group with normal subgroup H E G.

1. If f : G → G0 is a group homomorphism with H ≤ ker f then f induces a well-defined homo-


morphism f¯ : G/H → G0 , given by f¯(gH) = f (g).

2. If ϕ : G/H → G0 is any homomorphism, then ϕ ◦ πH : G → G0 is a homomorphism whose


kernel contains H.

Thus, we have a one-to-one correspondence

Hom(G/H, G0 ) ↔ {f ∈ Hom(G, G0 ) : H ≤ ker f }


ϕ → ϕ ◦ πH
¯
f ← f

11
2.8 The third isomorphism theorem

The final isomorphism theorem concerns products of subgroups. If H and K are subgroups of a group
G, the product HK = {hk : h ∈ H, k ∈ K} contains the identity, but it need not be closed under
the operation in G, hence HK need not be a subgroup of G. However, it will be so under an additional
condition.

The normalizer of H in G is the subgroup


NG (H) = {g ∈ G : gHg −1 = H}.
We have NG (H) = G if and only if H E G. In general, NG (H) is the largest subgroup of G containing
H as a normal subgroup.

Returning to our two subgroups H, K in G, let us assume that


K ≤ NG (H).
(This assumption holds automatically if H E G.) Then, for h, h0 ∈ H and k, k 0 ∈ K, we have
(hk)(h0 k 0 ) = h(kh0 k −1 ) · kk 0
(where we insert () and · to help parse the product), and kh0 k −1 belongs to H since k ∈ NG (H), so
(hk)(h0 k 0 ) ∈ HK. Similarly, (hk)−1 = k −1 h−1 = (k −1 h−1 k) · k −1 ∈ HK. Hence HK is a subgroup
of G. Since both H and K are contained in NG (H), it follows that HK is also contained in NG (H).
In other words, H is normal in HK.

This proves the first part of the following

Theorem 2.9 (Third Isomorphism Theorem) Let H and K be subgroups of a group G and assume
that K ≤ NG (H). Then

1. HK is a subgroup of G and H is a normal subgroup of HK.


2. H ∩ K is a normal subgroup of K.
3. We have a group isomorphism K/(K ∩ H) ' HK/H, induced by the map f : K → HK/H
given by k 7→ kH.

Proof: We have already proved the first part, and the second part is easy. As for the third part, it is
clear that f is surjective, so it remains to determine ker f . Let k ∈ K. Then f (k) = 1 in HK/H iff
ι(k) ∈ H, which means that k ∈ H. But k ∈ K, so we have f (k) = 1 iff k ∈ H ∩ K, as claimed. 

2.9 Direct products

Let H and K be groups with identity elements 1H and 1K . Then the direct product
H × K = {(h, k) : h ∈ H, k ∈ K}

12
is a group under the operation
(h, k)(h0 , k 0 ) = (hh0 , kk 0 )
and identity element (1H , 1K ). The direct product of finitely many groups G1 , . . . , Gn is defined simi-
larly; we confine our discussion to the case n = 2.

Let G = H × K, and write 1 = (1H , 1K ). The maps φ : H → G and ψ : K → G given by


φ(h) = (h, 1K ) and ψ(k) = (1H , k) are injective homomorphisms. Their images H 0 = φ(H) ' H
and K 0 = ψ(K) ' K are normal subgroups of G such that

H 0 ∩ K 0 = {1} and H 0 K 0 = G.

Conversely, this can be used to recognize direct products as follows.

Proposition 2.10 Let G be a group with subgroups H and K. Assume that

1. H and K are both normal in G;

2. H ∩ K = {1};

3. G = HK.

Then G ' H × K, via the map f : H × K → G given by f (h, k) = hk .

Proof: Let h ∈ H, k ∈ K and parse the commutator [h, k] = hkh−1 k −1 in two ways. On the one
hand, [h, k] = (hkh−1 )k −1 ∈ K since K E G. On the other hand, [h, k] = h(kh−1 k −1 ) ∈ H, since
H E G. But H ∩ K is trivial, so [h, k] = 1. Hence h and k commute for all h ∈ H and k ∈ K. It is
now immediate that f is a homomorphism, which is surjective by assumption 3. Finally, if f (h, k) = 1,
we have h = k −1 ∈ H ∩ K = {1}, so h = k = 1. Therefore f is an isomorphism, as claimed. 

If H and K are abelian groups then we often write H ⊕ K instead of H × K, in accordance with our
use of additive notation for abelian groups.

2.10 Semidirect products (internal view)

Recall that a group G with two normal subgroups H, K E G is the direct product G = H × K iff
HK = G and H ∩ K = {1}. This situation often occurs with the variation that only one of the
subgroups is normal.

Definition 2.11 A group G is a semidirect product of two subgroups H, K ≤ G if the following


conditions hold.

1. One of the subgroups H and/or K is normal in G.

13
2. H ∩ K = {1}.

3. HK = G.

Suppose G is the semidirect product of H and K and (say) H is normal in G. On the set H × K, define
a group law as follows:
(h, k)(h0 , k 0 ) = (hkh0 k −1 , kk 0 ).
Let H o K denote the set H × K with this group law.

Proposition 2.12 If G is a semidirect product of two subgroups H and K with H E G, then the map
(h, k) 7→ hk is a group isomorphism

H o K −→ G.

Proof: exercise. 

2.11 Conjugacy

For g, x ∈ G, let us set


g
x = gxg −1 .
Two elements x, y ∈ G are conjugate in G if y = g x for some g ∈ G. Conjugacy is an equivalence
relation on G, whose equivalence classes are the conjugacy classes of G. Thus any group is partitioned
into conjugacy classes. We write
G
x = {g x : g ∈ G}
for the conjugacy class of x in G.

Some of our earlier notions can be expressed in terms of conjugacy. For example, we have x ∈ Z(G)
if and only if G x = {x}. And H E G if and only if H is a union of conjugacy classes in G.

The centralizer of a given element x ∈ G is the subgroup CG (x) = {h ∈ G : h x = x} consisting of


all elements in G which commute with x. This is generally not a normal subgroup of G, so the quotient
space G/CG (x) is not a group. However, we have

Proposition 2.13 For every x ∈ G, the map g 7→ g x induces a well-defined bijection



G/CG (x) −→ G x.

In particular, if G is finite, we have

|G|
|G x| = [G : CG (x)] = .
|CG (x)|

14
Proof: Exercise 

The last formula in Prop. 2.13 is very useful for computing the sizes of conjugacy classes and central-
izers in finite groups. It implies for example that |G x| divides |G|. Since G is the disjoint union of its
conjugacy classes, we have

Corollary 2.14 (The class equation) Let G be a finite group, let X1 , . . . , Xk be its conjugacy classes,
choose xi ∈ Xi for 1 ≤ i ≤ k, and let Gi = CG (xi ). Then we have
k
X
[G : Gi ] = |G|.
i=1

Alternatively,
1 1 1
+ + ··· + = 1.
|G1 | |G2 | |Gk |

Corollary 2.15 If |G| is a power of a prime p then G has nontrivial center.

Proof: We have G partitioned as G = Z(G) ∪ {noncentral elements}. Every conjugacy class has
size a power of p. This power is zero precisely for those classes consisting of a single element
in Z(G) and every conjugacy class of noncentral elements has size a positive power of p. Hence
|{noncentral elements}| is divisible by p. As |G| is also divisible by p, it follows that p divides |Z(G)|.
Since |Z(G)| ≥ 1, it follows that a positive power of p divides |Z(G)|. 

3 The Symmetric Group

Groups usually arise as symmetries of mathematical structure. The most basic structure is just a set.

For any set X, the symmetric group on X is the group SX of bijections σ : X → X from X to itself,
where the group operation is composition of functions: (στ )(x) = σ(τ (x)). The identity element is
the identity function e(x) ≡ x. The elements of SX are usually called permutations.

If β : X → Y is a bijection between two sets X and Y , then we get a group isomorphism Sβ : SX −→
SY , defined by
Sβ (σ) = β ◦ σ ◦ β −1 .
If X = Y then β ∈ SX and Sβ is conjugation by β.

3.1 Cycle decomposition and conjugacy classes

Assume now that X is finite, say |X| = n. Then SX is finite and one checks that

|SX | = n! = n(n − 1) · · · 2 · 1.

15
Given σ ∈ SX , define an equivalence relation ∼ on X by the rule:
σ

x∼y ⇔ y = σ j (x) for some j ∈ Z.


σ

The equivalence classes are called σ-orbits. If O is a σ-orbit and x ∈ O, the distinct elements of O are

x, σ(x), σ 2 (x), . . . , σ λ−1 (x),

where λ = |O|. We note that σ λ (x) = x.

Number the σ-orbits in X as O1 , . . . Oq , where |O1 | ≥ |O2 | ≥ · · · ≥ |Oq |, and set λi = |Oi | for each
i. Thus, we have a set partition
q
G
X= Oi
i=1

and a corresponding numerical partition


q
X
|X| = λi .
i=1

Let us choose xi ∈ Oi , for each σ-orbit Oi , and set xji = σ j−1 (xi ), so that

Oi = {xji : 1 ≤ λi }.

Let σi ∈ SX be the permutation sending

xi = x1i 7→ x2i 7→ · · · 7→ xλi i 7→ xi ,

and acting as the identity on Ok for k 6= i. We express σi by the symbol

σi = (x1i x2i · · · xλi i ).

Then σi σj = σj σi for i 6= j and


σ = σ1 σ2 · · · σq , (4)
where the product may be taken in any order. This is the disjoint cycle decomposition of σ. The
partition
λ(σ) := [λ1 , λ2 , . . . , λq ]
is called the cycle type of σ. The cycle type determines the conjugacy class, as follows.

Proposition 3.1 Two elements σ, σ 0 ∈ SX are conjugate if and only if λ(σ) = λ(σ 0 ).

Proof: Let O1 , . . . , Oq be the σ-orbits in X. If σ 0 = τ στ −1 for some τ ∈ SX , one checks that


τ (O1 ), . . . , τ (Oq ) are the σ 0 -orbits in X. Since τ is a permutatation, we have |τ (Oi )| = |Oi | for each
i, which implies that λ(σ 0 ) = λ(σ).

Conversely, suppose λ(σ 0 ) = λ(σ). Let O1 , . . . , Oq be the σ-orbits and let O10 , . . . , Oq0 0 be the σ 0 -orbits.
Since λ(σ) = λ(σ 0 ) we have q = q 0 and |Oi | = |Oi0 | for each i. Since the orbits are disjoint there exists

16
a permutation τ ∈ SX such that τ (Oi ) = Oi0 for each i. Replacing σ by τ στ −1 , we may assume that
Oi0 = Oi for each i.

Choose xi ∈ Oi , and set xji = σ j−1 (xi ), yij = (σ 0 )j−1 (xi ), so that x1i = yi1 = xi and
Oi = {x1i , x2i , . . . , xλi i } = {yi1 , yi2 , . . . , yiλi }.
Every element of X may be uniquely expressed in the form xji or in the form yij , where 1 ≤ i ≤ q and
1 ≤ j ≤ λi . Let π ∈ SX be the permutation sending xji 7→ yij for each such i, j. For x = xji , we have
πσπ −1 (x) = πσσ j (xi ) = πσ j+1 (xi ) = (σ 0 )j+1 (xi ) = σ 0 (x).
It follows that πσπ −1 = σ 0 , so σ and σ 0 are conjugate. 

3.2 The group Sn

When X = {1, 2, . . . , n} we write Sn = SX . If X is any set with |X| = n then by labelling the

elements of X we obtain a bijection f : X → {1, 2, . . . , n}, whence a group isomorphism Sf : SX −→
Sn . This isomorphism is noncanonical, because it depends on the chosen labelling f . Moreover, the set
X need not be ordered, as {1, 2, . . . , n} is. The ordering gives additional structure to Sn that is missing
in SX when X is unordered

As in (4), an element in Sn is a product of cycles, but now the symbols in the cycle are numbers in
{1, 2, . . . , n}. For example, the element σ ∈ S6 sending
1 7→ 4, 2 7→ 6, 3 7→ 1, 4 7→ 3, 5 7→ 5, 6 7→ 2
may be written as
σ = (1 4 3)(2 6),
and has cycle type λ(σ) = [3, 2, 1]. Note that 5, which is fixed, is omitted in the cycle decomposition
of σ, but is counted in the cycle type of σ.

We multiply using the cycle decomposition by following the path of each number, starting with the
right-most cycle. For example, we have
(1 4 3)(2 6)(4 6 5) = (1 4 2 6 5 3).
Note that the cycle type of τ = (1 4 3)(2 6)(4 6 5) is not [3, 3, 2], because the cycles are not disjoint.
We first have to make them disjoint, as we have done above. We obtained a single 6-cycle (1 4 2 6 5 3),
so the cycle type of τ in S6 is λ(τ ) = [6].

We illustrate the cycle types for all elements of S3 and S4 , as follows.


λ σ
λ σ [4] (1 2 3 4), (1 3 2 4), (1 2 4 3), (1 4 2 3), (1 3 4 2), (1 4 3 2)
[3] (1 2 3), (3 2 1) [3, 1] (1 2 3), (3 2 1), (1 2 4), (1 4 2), (1 3 4), (1 4 3), (2 3 4), (4 3 2)
[2, 1] (1 2), (2 3), (1 3) [2, 2] (1 2)(3 4), (1 3)(2 4), (1 4)(2 3)
[1, 1, 1] e [2, 1, 1] (1 2), (1 3), (1 4), (2 3), (2 4), (3 4)
[1, 1, 1, 1] e

17

You might also like