You are on page 1of 29

1.

HISTORY OF MAHEMATICS

The history of mathematics is an ever-growing series of abstractions.


Evolutionarily speaking, the first abstraction to ever be discovered, one
shared by many animals,[73] was probably that of numbers: the realization
that, for example, a collection of two apples and a collection of two oranges
(say) have something in common, namely that there are two of them. As
evidenced by tallies found on bone, in addition to recognizing how
to count physical objects, prehistoric peoples may have also known how to
count abstract quantities, like time—days, seasons, or years.[74][75]

The Babylonian mathematical tablet Plimpton 322, dated to 1800 BC

Evidence for more complex mathematics does not appear until around
3000 BC, when the Babylonians and Egyptians began using arithmetic,
algebra, and geometry for taxation and other financial calculations, for
building and construction, and for astronomy. [76] The oldest mathematical
texts from Mesopotamia and Egypt are from 2000 to 1800 BC. Many early
texts mention Pythagorean triples and so, by inference, the Pythagorean
theorem seems to be the most ancient and widespread mathematical
concept after basic arithmetic and geometry. It is in Babylonian
mathematics that elementary
arithmetic (addition, subtraction, multiplication, and division) first appear
in the archaeological record. The Babylonians also possessed a place-value
system and used a sexagesimal numeral system which is still in use today
for measuring angles and time.[77]

In the 6th century BC, Greek mathematics began to emerge as a distinct


discipline and some Ancient Greeks such as the Pythagoreans appeared to
have considered it a subject in its own right. [78] Around 300 BC, Euclid
organized mathematical knowledge by way of postulates and first
principles, which evolved into the axiomatic method that is used in
mathematics today, consisting of definition, axiom, theorem, and
proof.[79] His book, Elements, is widely considered the most successful and
influential textbook of all time. [80] The greatest mathematician of antiquity
is often held to be Archimedes (c. 287 – c. 212 BC) of Syracuse.[81] He
developed formulas for calculating the surface area and volume of solids of
revolution and used the method of exhaustion to calculate the area under
the arc of a parabola with the summation of an infinite series, in a manner
not too dissimilar from modern calculus. [82] Other notable achievements of
Greek mathematics are conic sections (Apollonius of Perga, 3rd century
BC),[83] trigonometry (Hipparchus of Nicaea, 2nd century BC),[84] and the
beginnings of algebra (Diophantus, 3rd century AD). [85]
The numerals used in the Bakhshali manuscript, dated between the 2nd
century BC and the 2nd century AD

The Hindu–Arabic numeral system and the rules for the use of its
operations, in use throughout the world today, evolved over the course of
the first millennium AD in India and were transmitted to the Western
world via Islamic mathematics.
2. HISTORY OF GROUP THEORY

Group theory has three main historical sources: number theory, the theory
of algebraic equations, and geometry. The number-theoretic strand was
begun by Leonhard Euler, and developed by Gauss's work on modular
arithmetic and additive and multiplicative groups related to quadratic
fields. Early results about permutation groups were obtained
by Lagrange, Ruffini, and Abel in their quest for general solutions of
polynomial equations of high degree. Évariste Galois coined the term
"group" and established a connection, now known as Galois theory,
between the nascent theory of groups and field theory. In geometry,
groups first became important in projective geometry and, later, non-
Euclidean geometry. Felix Klein's Erlangen program proclaimed group
theory to be the organizing principle of geometry.

Galois, in the 1830s, was the first to employ groups to determine the
solvability of polynomial equations. Arthur Cayley and Augustin Louis
Cauchy pushed these investigations further by creating the theory of
permutation groups. The second historical source for groups stems
from geometrical situations. In an attempt to come to grips with possible
geometries (such as euclidean, hyperbolic or projective geometry) using
group theory, Felix Klein initiated the Erlangen programme. Sophus Lie, in
1884, started using groups (now called Lie groups) attached
to analytic problems. Thirdly, groups were, at first implicitly and later
explicitly, used in algebraic number theory.
The different scope of these early sources resulted in different notions of
groups. The theory of groups was unified starting around 1880. Since then,
the impact of group theory has been ever growing, giving rise to the birth
of abstract algebra in the early 20th century, representation theory, and
many more influential spin-off domains. The classification of finite simple
groups is a vast body of work from the mid 20th century, classifying all
the finite simple groups.
3. INTRODUCTION TO GROUP THEORY

Before we discuss group actions, we will need to know something about


groups. In this section, we explore basic concepts in group theory that will
be of use to us later. Our survey will not be exhaustive, and interested
readers are invited to peruse [1], [2], or [3] for more results.

Definition 1.1. A group is an ordered pair (G, ·), where G is a set and · is a
binary operation (often referred to as multiplication) such that:

(i) For all a, b, c ∈ G, (a · (b · c) = (a · b) · c)


(ii) There exists 1 ∈ G such that for all g ∈ G, 1 · g = g
(iii) For each g ∈ G, there exists g −1 ∈ G such that g · g −1 = g −1 · g = 1

Examples 1.2. (R, +),(C, +),(Q, +) are all groups.

Other interesting examples of groups occur in geometry. An example is D3,


the dihedral group on three elements (the vertices of the triangle), which
contains all of the symmetries of an equilateral triangle. Another is the
points on a circle with the operation of rotation, which is easy to draw and
which obviously satisfies the above axioms. We also have a natural notion
of the size of a group.

Definition 1.3. Let (G, ·) be a group. The order of G, denoted |G|, is the
number of elements in the set G.
Remark 1.4. When there is no risk of confusion regarding the operation
with which G is a group, I will often refer to G as the group.

Note that our operation need not be commutative. For instance, many sets
of matrices form groups with respect to matrix multiplication, which is
generally noncommutative. This leads to our next definition.

Definition 1.5. A group (G, ·) is Abelian if · is commutative.

Examples 1.6. The groups listed in Example 1.2 are all Abelian.

Looking at the examples listed, the reader might notice something; Q ⊂R ⊂


C and all of these sets are groups with respect to addition. This notion is
encapsulated in the following defintion.

Definition 1.7. A subgroup H 6 G is a set H ⊆ G that is a group under the


same operation under which G is a group.

We will prove several results in this paper that help us determine


subgroups of a group. The following is the first of these results, though it
only works of one knows that a group contains certain subgroups. Other
results will not need us to have this knowledge.

Proposition 1.8. If {Hi} are subgroups of G, then their intersection is a


subgroup of G.

Proof. The identity is obviously in the intersection of Hi . Since associativity


and inverses hold by hypothesis for elements in any Hi , they hold for
elements in the intersection.
Proposition 1.9 enables us to make the following definiton:

Definition 1.9. Let A be a subset of a group G and let Hi be subgroups of G


such that A ⊆Hi . hAi = ∩iHi is called the subgroup of G generated by A.

An important special case follows:

Definition 1.10. A group generated by a single element, call it x, is called a


cyclic group and is denoted (x).

For the intuition behing the term cylcic, one need only look at the rotations
of a circle. Cyclic groups allow us to assign an order to an element of a
group if the element generates a cyclic group. The order is simply the order
of the cyclic group the element generates.

We have now examined groups. It is natural to next ask what maps are of
interest. The answer is the maps that preserve group structure.

Definition 1.11. Let G and H be groups. A map ϕ : G → H is a


homomorphism if, given g1, g2 ∈ G, ϕ(g1g2) = ϕ(g1)ϕ(g2).

Remark 1.12. : There are really two operations in the above definition. The
first is the binary operation in G and the second is the one in H.

There are some special homomorphisims to which we will refer by name.

Definitions 1.13. : A homomorphism that is injective is a monomorphism. A


homomorphism that is surjective is an epimorphism. A homomorphism that
is bijective is an isomorphism.
The most important of the above to recognize is isomorphism. If two
groups have an isomorphism between them, they are, as far as algebraists
care, the same. They do not need to have the same elements; algebraists
seldom care about the elements of groups. Rather, the groups have the
same structure - namely binary operations that satisfy the axioms in
Definition 1.1. Two groups with an isomorphism between them are said to
be isomorphic
4. GROUP ACTION AND ITS APPLICATION

4.1 GROUP :

Definition 2.1. A group is a set G together with a binary operation : G G G


such that the following conditions hold:

Closure: For all g, h ∈ G, the element g ◦ h is a uniquely defined element of G.

Associativity: For all f, g, h ∈ G, we have

(f ◦ g) ◦ h = f ◦ (g ◦ h).

Identity: There exists an identity element e ∈ G such that

e ◦ g = g and g ◦ e = g

for all g ∈ G.

Inverse: For each g ∈ G. there exists an inverse element g−1 ∈ G such that

g ◦ g−1 = e and g−1 ◦ g = e.

We will usually simply write gh for the composition g ◦ h.

Definition 2.2. The order of a group G is its cardinality. In other words, it is the
number of elements in G. The order of G is denoted |G|.

Example 2.3. The dihedral group, Dn, is the group of symmetries of a regular
n-gon. The dihedral group of a square is denoted D4, and defined as such:

D4 = {e, r1, r2, r3, s1, s2, s3, s4},

1
Where
r1 : clockwise rotation by 90◦
r2 : clockwise rotation by 180◦
r3 : clockwise rotation by 270◦
e : clockwise rotation by 360◦
s2 : reflection across y-a
s3 : reflection across y = x or BD
s4 : reflection across y = −x or AC.

A B

D C

4.2 . SUBGROUP:

Definition 2.4. A subgroup of a group G is a subset of G such that it itself is a


group under the operation in G.

Definition 2.5. A group G acts on a set S when there is a map G × S → S


(written (g, s) ›→ gs) such that the following conditions hold for all s ∈ S.

(i) Associativity: For g, h ∈ G and s ∈ S,

g(hs) = (gh)s

(ii) Identity: The action of the identity of G on every s ∈ S gives s, in

2
other words,
es = s

Definition 2.7 (Set Multiplication). Given set S and element x, we define


multiplication of S and x
by Sx = {sx : s ∈ S}.

Definition 2.8. Let H be a subgroup of a group G. The right cosets of H are the sets Hg
= {hg : h ∈ H}
for each g ∈ G.
For any element h ∈ H, we have Hh = H by the closure property of groups and
subgroups.
Example 2.10. Let S = G and let G act on itself by conjugation. For g G and s S,
the group G

acts upon S by conjugation by (g, s)→gsg−1 .


We can see that conjugation satisfies the two requirements for a group action:

The group action is associative. For g, h ∈ G and s ∈ S,

g ◦ (h ◦ s) = g(h ◦ s)g −1 = g · (h · s · h−1 ) · g−1 = (g · h) · s · (g · h)−1 = (g · h) ◦ s ,

The action of the identity of G on s gives s.

(e, s) ›→ ese−1 = s.

3
4.3 GROUP ACTION:

1.LEFT GROUP ACTION

Definition 1.1. Suppose that G is a group and S is a set. A left (group) action of G
on S is a rule for combining elements g ∈ G and elements x ∈ S, denoted by g.x.
We additionally require the following 3 axioms.
(0) g.x ∈ S for all x ∈ S and g ∈ G
(1) e.x = x for all x ∈ S.

(2) g1.(g2.x) = (g1g2).x for all x ∈ S and g1, g2 ∈ G. It is critical to note that in (2)
above, g1.(g2.x) has two actions of G on elements of S. On the other hand (g1g2).x
has one multiplication in the group g1g2 and then one action of an element of G
on S
. Now we present an example
Example 1.2. Suppose that S is any set and consider G to be any subgroup of A(S).
Then G acts on S as follows. For each f ∈ G ⊆ A(S) and each x ∈ S, we define
f.x = f(x) ∈ S
Note that f : S → S is a (bijective) function since f ∈ A(S). Thus this makes sense
and condition (0) is verified. For condition (1), note that e ∈ G is the identity
function id : S → S, and id .x = id(x) = x as desired. Finally we verify (2), note that if
f, g ∈ G, then f.(g.x) = f(g.x) = f(g(x)) = (f ◦ g)(x) = (fg).x as desired
We now consider the same example in a slightly more explicit framework

4
2. Right group actions

Definition 2.1. Suppose that G is a group and S is a set. A right (group) action of
G on S is a rule for combining elements g ∈ G and elements x ∈ S, denoted by x.g.
We additionally require the following 3 axioms.
(0) x.g ∈ S for all x ∈ S and g ∈ G
(1) x.e = x for all x ∈ S.
(2) (x.g2).g1 = x.(g2g1) for all x ∈ S and g1, g2 ∈ G
Example 2.1. Suppose that G is the group of 2×2 matrices, ( cos θ sin θ − sin θ cos
θ θ ∈ R ) , under matrix multiplication. Set S = R 2 to be the set of row vectors x
y . We have a right action of G on S defined by matrix multiplication times a
vector:
v.A = Va
for v ∈ R 2 and A ∈ G. If A corresponds to θ, then this action rotates a vector v by
θ radians counter-clockwise. For example, set θ = π/2 and set v = 1 0 . For
example, consider
[ 1 0 ].[ cos θ sin θ − sin θ cos θ]= 0 1 = [01]
-1 0
Just as before it is easy to see that this is a right group action.

4.4 ORBIT AND STABILIZER

Fix an action of a group G on a set X. For each point x of X, we have two important
concepts: DEFINITION: The orbit of x ∈ X is the subset of X O(x) := {g · x | g ∈ G} ⊂
X.

5
DEFINITION: The stabilizer of x is the subgroup of G Stab(x) = {g ∈ G | g · x = x}

4.5 Orbit-Stabilizer Theorem

Throughout this section we fix a group G and a set S with an action of the
group G. In this section, the group action will be denoted by both g · s and gs.

Definition 4.5.1. The orbit of an element s ∈ S is the set

orb(s) = {g · s | g ∈ G} ⊂ S.

Theorem 4.5.2. For y ∈ orb(x), the orbit of y is equal to the orbit of x

Proof. For y ∈ orb(x), there exists some g1 ∈ G such g1 x = y. We can also write
this as x = g1−1 y by left-multiplication with g −1 . For z ∈ orb(y), there exists some h ∈
G such that hy = z. By substituting gx = y into the equation, we get ghx = z. By
closure, gh ∈ G, so z ∈ orb(x) and orb(x) ⊆ orb(y).
Similarly, for w ∈ orb(x) where w /= y, there exists some g2 ∈ G such that g2x = w.
Substituting
X= g1−1 y, we see that g2 g1−1 y = w. By closure, g2 g1−1 ∈ G so w ∈ orb(y) and orb(x) ⊆
orb(y).
Since orb(x) ⊆ orb(y) and orb(y) ⊆ orb(x), orb(y) = orb(x).

Theorem 4.5.3. For any element s ∈ S, the stabilizer stab(s) ⊂


G is a subgroup of G. Proof. We check all the group properties
(i)Closure: Let g, h ∈ stab(s). Then,

g · h · s = g · (h · s) = g · s = s.

6
Therefore, g · h ∈ stab(s).

(ii)Identity: For identity e ∈ G, we have e · s = s. Therefore, e ∈ stab(s).

(iii)Inverse: Let g ∈ stab(s). Then g · s = s. Left-multiplication by g −1


gives e · s = s = g−1 · s. Therefore, g−1 ∈ stab(s).

4.6 BURNSIDE’S LEMMA

Once again, let G be a group acting on a set S.


Definition 4.6.1. For an element g G, a fixed point of a set S is an
element s S such that gs = s. In other words, s is unchanged by the group
operation. We denote the set of points in S fixed by g as fix(g).

Definition 4.6.2. For a group G that acts on set S, let S/G be the set of orbits of
S.

Theorem 4.6.3 (Burnside’s Lemma). For a finite group G that acts on a finite set
S, we have

S/G =|G |g=G 1 fix(g) .

History: the lemma that is not Burnside's

William Burnside stated and proved this lemma in his 1897 book on finite groups,
attributing it to Frobenius 1887. But even prior to Frobenius, the formula was
known to Cauchy in 1845. Consequently, this lemma is sometimes referred to as
the lemma that is not Burnside's.[4] Misnaming scientific discoveries is referred to
as Stigler's law of eponymy

7
Applications of Burnside’s Lemma

Burnside’s Lemma can be applied to a variety of counting problems.

Coloring Problems

Burnside’s Lemma can be used to solve problems about the number of ways
to color various geometric patterns.

.Example 4.6.4 (triangles with n colors). We compute the number of distinct


colorings of the regions of equilateral triangular tile below, up to rotation and
reflection, with n colors. The group of symmetries of the triangle is D3. The order
of D3 is 6.

8
We investigate fix(g), the number of colorings fixed under each
group element g D3. The identity fixes every coloring. Since
there are 7 regions and n colors, fix(e) = n7.

For a coloring to be fixed under rot 120◦ , each of the regions An and
Bn must be the same color, while C can be any color. Thus, fix(rot 120◦ ) =
n3 . Similarly, fix(rot( 120◦ )) = n3 .
There are three group elements that reflect across a line of symmetry of
the triangle. Without loss of generality, let us consider the group element
that reflects across the line of symmetry from the top most vertex to the
midpoint of the side opposite to it. We call this group element ref. For the
coloring to be fixed under ref, regions A2 and A3 must be the same, as well as
regions B1 and B3. The other regions can be any color. Thus, fix(ref) = n5.
Since the number of colorings is equal to the number of orbits of D3, we
can use Burnside’s Lemma to get
Number of colourings = 1 > |f(x,g)| = 1/6 (n7+3n5+2n3

|D3|gE D3

Burnside’s Lemma can also be used to solve problems about the number
of unique necklaces that can be made using various colors and numbers of
beads. In the following, we let C be a set of colors that we are allowed to
use for our necklaces.

9
Definition 4.6.5. Let C be a set. A necklace with colors in C is a finite
sequence x1 . . . xn of ele- ments of C. We say two necklaces are the same if
one is a rotation xixi+1 . . . xn+i−1 or reflection x2ix2i−1x2i−2 . . . x2i−n of the other,
taking subscripts mod n.

Definition 4.6.6. A bead is a color on a necklace

Example 4.6.7 (necklaces with n colors). We compute the number of


distinct necklaces with 6 beads and n possible colors for each bead. The
group of symmetries of a 6-bead necklace is D6. The order of D6 is 12.
We investigate fix(g), the number of necklaces fixed under each group
element g D6. For a necklace to be fixed under the group element rot 60◦ ,
the second bead must be the same color as the first bead, the third the
same as the second, the fourth the same as the third, and so on. Thus, all
of the beads on the necklace must be the same color, so fix(rot 60◦ ) = n.
Using similar methods to find fix(g) for the rest of the group elements, we
have

e: fix(e) = n6

rot ± 60◦ : 2 fix(rot 60◦ ) = 2n

rot ±120◦ : 2 fix(rot 120◦ ) = 2n2

rot 180◦ : fix(rot 180◦ ) = n3

ref(3) : 3 · fix(ref) = 3n4

10
ref(3) : 3 · fix(ref) = 3n3

Since the number of necklaces is equal to the number of orbits of D6, we


can use Burnside’s Lemma to get
Number of necklaces = 1 ≥ |f(x,g)| =

|D3|gE D3

4.7 PERMUTATION REPRESENTATION ASSOCIATE WITH G

In mathematics, the term permutation representation of a (typically

finite) group can refer to either of two closely related notions:

a representation of as a group of permutations, or as a group of permutation


matrices. The term also refers to the combination of the two.

Abstract permutation representation

A permutation representation of a group G on a set X is a homomorphism


from G to the symmetric group of X:

ρ:G Sym (X)

The image ρ(G) Ϲ Sym (X) is a permutation group and the elements of G
are represented as permutations of X. A permutation representation is
equivalent to an action of G on the set X: G x X X

11
Linear permutation representation

If G is a permutation group of degree n, then the permutation


representation of G is the linear representation of G

ρ:G GLn(K)
which maps g € G to the corresponding permutation matrix (here K is an
arbitrary field).[2] That is, G acts on Kn permuting the standard basis vectors.

This notion of a permutation representation can, of course, be composed with the

previous one to represent an arbitrary abstract G group as a group of


permutation matrices. One first represents G as a permutation group and then
maps each permutation to the corresponding matrix. Representing G as a
permutation group acting on itself by translation, one obtains the regular
representation.

4.8 APPLICATIONS OF GROUP ACTIONS

4.8.1 CAYLEY’S THEOREM

The statement of Cayley’s Theorem


For a set S, we will denote by (Bij(S), ◦) the group of bijections f : S → S under
composition.
Theorem 1.1 (Cayley’s Theorem). Let (G, ·) be a group. There is an injective group
homomorphism
Φ : (G, ·) −→ (Bij(G), ◦)
defined by the rule that for all g, h ∈ G, we have Φ(g)(h) = gh.

12
We will prove Cayley’s Theorem below. Before we do that, we mention here that
Cayley’s
Theorem is often stated for finite groups in the following form:
Corollary 1.2. Let G be a finite group of order n. Then G is isomorphic to a
subgroup of
Sn, the symmetric group on n letters.

Proof. Cayley’s Theorem gives an injective homomorphism of groups Φ : G ,→


Bij(G).
Then since there is a bijection of sets G → {1, . . . , n}, we have an isomorphism of
groups
Ψ : Bij(G) → Bij({1, . . . , n}) =: Sn.

2. The group of bijections of a set


Given a set S, we recall that (Bij(S), ◦), the set of bijections f : S → S under
composition,
form a group. Namely, we have a map
◦ : Bij(S) × Bij(S) → Bij(S)
(f, g) 7→ f ◦ g.
The identity element of Bij(S) is IdS, since
IdS ◦f = f
for all f ∈ Bij(S). If f ∈ Bij(S), then the inverse element of f under the group law is
given
by the inverse map f −1 , since
f −1 ◦ f = IdS .

13
Composition of maps is associative. Thus (Bij(S), ◦) is a group

3. Proof of Cayley’s Theorem


Given a group G, there is a map of sets
Φ : G → Map(G, G)
Φ(g)(h) = gh,
for all g, h ∈ G.
Part of the assertion of Cayley’s Theorem is that Im(Φ) ⊆ Bij(G, G) ⊆ Map(G, G). In
other
words, given g ∈ G, the claim is that the map Φ(g) : G → G is a bijection. To show
that
Φ(g) is a bijection, we need to show that it is injective and surjective.
First let us show that if g ∈ G, then Φ(g) is injective. This means, that given h1, h2
∈ G,
then if Φ(g)(h1) = Φ(g)(h2), we need to show that h1 = h2. Well,
gh1 =: Φ(g)(h1) = Φ(g)(h2) := gh2.
Then composing with g −1 on the left, we have that
h1 = g −1 gh1 = g −1 gh2 = h2.
Thus Φ(g) is injective.
Let us now show that Φ(g) is surjective. This means that given h ∈ G, we need to
exhibit
h 0 ∈ G such that Φ(g)(h 0 ) = h. Well, given h ∈ G, if we set h 0 = g −1h, then
h = gh0 = Φ(g)(h 0 ).
Thus Φ(g) is surjective. We have now succeeded in showing that if g ∈ G, then
Φ(g) ∈

14
Bij(G, G).
The next claim of Cayley’s Theorem is that
Φ : G → Bij(G, G)
is a group homomorphism. In other words, given g1, g2 ∈ G, the claim is that
Φ(g1g2) = Φ(g1) ◦ Φ(g2).
It is enough to check this holds when the maps are applied to each h ∈ G. In other
words,
for h ∈ G, we have
Φ(g1g2)(h) := (g1g2)h = g1(g2h) = (Φ(g1) ◦ Φ(g2))(h).
Thus Φ(g1g2) = Φ(g1) ◦ Φ(g2).
The last claim of Cayley’s Theorem is that Φ is injective. In other words, given g1,
g2 ∈ G,
if Φ(g1) = Φ(g2), then the claim is that g1 = g2. To prove this, apply Φ(g1) and
Φ(g2) to the
identity element of G:
g1 = Φ(g1)(e) = Φ(g2)(e) = g2.
This shows that Φ is injective, and completes the proof of Cayley’s Theorem.

4.9 REAL LIFE APPLICATIONS OF GROUP THEORY:

 Group theory algorithms are used to solve Rubik’s cube.


 Many laws of Physics, Chemistry use symmetry and hence, uses group theory
as it is symmetric.

15
 Group theory may be used to investigate any object or system attribute that
is invariant under change because of its symmetry.
 Group theory is also used in harmonic analysis, combinatorics, algebraic
topology, algebraic number theory, algebraic geometry, and cryptography.
 Cryptography: Group Theory plays a crucial role in
encryption algorithms and secure communication.
 Physics: Understanding symmetry in particle physics and quantum
mechanics often involves the use of Group Theory.
 Chemistry: Crystallographic groups, used to describe the symmetry of
crystal structures, are based on Group Theory.

16
5.CONCLUSION

Studying group actions is important for several reasons. Some of them are:

1. - Group actions can help us understand the structure and properties of


groups by relating them to the symmetries of sets and mathematical
objects. For example, we can classify finite groups by their actions on
finite sets, or we can study the representation theory of groups by their
actions on vector spaces.

2. - Group actions can also help us understand the structure and properties
of sets and mathematical objects by relating them to the abstract
algebraic properties of groups. For example, we can use group actions to
define and compute invariants, such as the Euler characteristic, the
homology groups, or the fundamental group of a topological space.

3. - Group actions can also help us discover and prove new results in various
branches of mathematics by providing useful tools and techniques. For
example, we can use group actions to count combinatorial objects, such
as partitions, permutations, or graphs, by applying the Burnside's lemma
or the Polya enumeration theorem.

4. You can find more information about group actions and their applications
from the web search results³. I hope this helps you appreciate the
importance of studying group actions. 😊

"In summary, group action theory serves as a foundational concept in BSc


mathematics, offering a powerful framework to study symmetry, permutation

17
groups, and other abstract algebraic structures. Through its applications in fields
such as combinatorics, geometry, and cryptography, group action not only
deepens our understanding of mathematical structures but also provides practical
tools for solving real-world problems. Its interdisciplinary nature underscores its
significance in shaping modern mathematical discourse and its relevance to
various fields beyond pure mathematics."

Certainly! In conclusion, the study of group actions within the context of BSc
Mathematics offers students a deep understanding of symmetry and
transformational properties across various mathematical structures. By exploring
group actions, students gain insights into fundamental concepts such as orbits,
stabilizers, and permutation groups, which are crucial in fields like algebra,
topology, and geometry. Moreover, group actions serve as a unifying framework
for diverse mathematical theories, enabling students to make connections
between different branches of mathematics and apply their knowledge to solve
real-world problems. Overall, a thorough comprehension of group actions
enriches the mathematical toolkit of BSc Mathematics students, equipping them
with valuable skills for both theoretical exploration and practical applications in
various scientific and engineering disciplines.

18
6.REFERENCES

[1] Jenny Jin. Analysis and Applications of Burnside’s Lemma. 2018.

[2] Rahbar Virk. The Orbit-Stabilizer Theorem.

[3] Camilla and David Jordan. Groups. Modular Mathematics. Newnes, 1994.

[4] Micheal Artin. Algebra. 2011.

[5]Applications of Group Actions: Cauchy’s Theorem and Sylow’s Theorems R. C.


Daileda
[6]Group -- from Wolfram MathWorld. Group -- from Wolfram MathWorld.
[7]Group action - Wikipedia. Group action - Wikipedia.

[8] Group actions | Brilliant Math & Science Wiki. Group actions | Brilliant Math &
Science Wiki.

[9]. John B. Fraleigh, A First Course in Abstract Algebra, Narosa Publishing House,
New Delhi.

[10] Joseph A. Gallian Contemporary Abstract Algebra (4th Edition), Narosa


Publishing House,

New Delhi.
[11] M. Artin, Abstract Algebra, 2nd Ed., Pearson, 2011.

[12] David S. Dummit and Richard M. Foote, Abstract Algebra, 3rd Ed., John Wiley

19
and Sons
(Asia) Pvt. Ltd., Singapore, 2004.

[13] J.R. Durbin, Modern Algebra, John Wiley & Sons, New York Inc., 2000.

[14] B. Fein, W. M. Kantor, M. Schacher, Relative Brauer groups II, J. Reine Angew.
Math. 328 (1981), 39–57.

[15] B. R. Gelbaum and J. M. H. Olmstead, Theorems and Counterexamples in


Mathematics, Springer, New York, 1990.

[16] I. M. Isaacs and M. R. Pournaki, “Generalizations of Fermat’s Little Theorem


Using Group Theory,” Amer. Math. Monthly 112 (2005), 734–740.

20
21

You might also like