Download as pdf or txt
Download as pdf or txt
You are on page 1of 38

SUBGROUP LATTICES OF CYCLIC

GROUPS WITH MAPLE

MSc THESIS

MOE HNIN PHYU

DEPARTMENT OF MATHEMATICS
TAUNGGYI UNIVERSITY
MYANMAR

DECEMBER 2020
SUBGROUP LATTICES OF CYCLIC
GROUPS WITH MAPLE

by

MA MOE HNIN PHYU


MSc II Math-11

DEPARTMENT OF MATHEMATICS
TAUNGGYI UNIVERSITY

Thesis submitted in partial fulfillment of the requirements for the


degree of Master of Science in Mathematics in the Taunggyi
University.

MSc Thesis December 2020


Candidate name : Ma Moe Hnin Phyu
Roll No. : MSc II Math-11
University : Taunggyi University
Thesis Title : Subgroup Lattices of Cyclic Groups with Maple
Supervisor : Dr Daw Aye Pyone
: Associate Professor
: Department of Mathematics
: Taunggyi University
Academic Year : 2019-2020
SUBGROUP LATTICES OF CYCLIC GROUPS WITH
MAPLE

MOE HNIN PHYU

This Thesis is Submitted to the Board of Examiners in Mathematics,


Taunggyi University in Partial Fulfillment for the Degree of Master
of Science.

APPROVED

························ ························
EXTERNAL EXAMINER SUPERVISOR
Dr U Zaw Myint Dr Daw Aye Pyone
Professor, Head of Department Associate Professor
Department of Mathematics Department of Mathematics
Monywa University Taunggyi University

························ ························
SECRETARY CHAIRPERSON
Dr Ne Ne Le Dr Khin Maw Maw
Associate Professor Professor, Head of Department
Department of Mathematics Department of Mathematics
Taunggyi University Taunggyi University
Acknowledgements

I am very grateful to Dr Khin Maw Maw, Professor and Head, Department of


Mathematics, Taunggyi University, for her warm encouragement and suggestions
during this thesis.
My deepest gratitude goes to my supervisor, Dr Daw Aye Pyone, Associate
Professor, Department of Mathematics, Taunggyi University, for her advice and
guidance in preparing the materials for this thesis.
I am also thankful to all my teachers throughout my student life. Finally, I
would like to thank my beloved parents for helping in my life.

December, 2020 Moe Hnin Phyu


Contents

Abstract ii

1 Congruence Classes 1
1.1 Some Results on the Integers . . . . . . . . . . . . . . . . . . . . . 1
1.2 On Congruence Classes Modulo n . . . . . . . . . . . . . . . . . . 5
1.3 Programs in Maple 2018 . . . . . . . . . . . . . . . . . . . . . . . 8
1.3.1 Integers . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.3.2 Loops and Statements . . . . . . . . . . . . . . . . . . . . 9
1.3.3 Procedures . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

2 Groups of Integers Modulo n 14


2.1 Additive Groups of Integers Modulo n . . . . . . . . . . . . . . . 14
2.2 Multiplicative Groups of Integers Modulo n . . . . . . . . . . . . 17

3 Cyclic Groups 20
3.1 Subgroup Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.2 Properties of Cyclic Groups . . . . . . . . . . . . . . . . . . . . . 22
3.3 Subgroups of Cyclic Groups . . . . . . . . . . . . . . . . . . . . . 24

Bibliography 31

i
Abstract

We present that the set of all integers can present in terms of the distinct
equivalence classes of integers modulo n. Next, we study an algebraic structure
containing a set G and an operation on it, namely, a group G. In this thesis, we
discuss only additive and multiplicative groups of integers modulo n. We present
that every subgroup of a cyclic group is cyclic and how many subgroups a finite
cyclic group has and how to find them. By using this result the subgroups of
additive group modulo n, Zn is found. Finally, we will show another structure
of Zn , subgroup lattice. By using Maple programming, we draw these lattice
diagrams.

ii
Chapter 1

Congruence Classes

In this chapter we show that the set of all integers, Z, is the family of equiv-
alence classes. That is, Zn = {[0], [1], . . . , [n − 1]} under modulo n. To end this
we collect some results on integers.

1.1 Some Results on the Integers


In this section recall some definitions and results on the integers. Now for
proper understanding of this thesis we need to explain some terminologies used.
Moreover, there are some basic propositions and theorems, which we use to back
our points. We denote Z is the set of all integers and n is a positive integer.
The basic assumption we make about the set of integers is the Well-Ordering
Principle:
Any nonempty set of nonnegative integers has a smallest member.
The following result is the first application of well-ordering principle.

Theorem 1.1.1. (Euclid’s Algorithm) If m and n are integers with n > 0,


then there exist integers q and r, with 0 ≤ r < n, such that m = qn + r.

Proof. Let W = {m − tn| t ∈ Z}. Note that W contains some nonnegative


integers, for if t is large enough and negative, then m − tn > 0. Let
V = {v ∈ W | v ≥ 0}; by the well-ordering principle V has a smallest element,
r. Since r ∈ V , r ≥ 0, and r = m − qn for some q. We claim that r < n. If not,
m − qn ≥ n, hence m − (q + 1)n ≥ 0. But m − (q + 1)n in V , m − (q + 1)n < r,
contradicting the minimal value of r in V . The proof is complete.

Euclid’s Algorithm will give about the notion of divisibility.

Definition 1.1.2. Given integers m 6= 0 and n we say that m divides n write


as m|n, if n = cm for some integer c.

1
Thus for instance, 2|6, (−2)|4. If m|n, we call m a divisor or factor of n, and
n a multiple of m. If m is not a divisor of n, we write m - n; for instance 3 - 5.

Now we will show the basic elementary properties of divisibility.

Lemma 1.1.3. The following are true:

(i) 1|n for all n.

(ii) If m 6= 0, then m|0.

(iii) If m|n and n|q, then m|q.

(iv) If m|n and m|q, then m|(un + vq) for all u, v.

(v) If m|1, then m = 1 or m = −1.

(vi) If m|n and n|m, then m = ±n.

Proof. (i) Since n = n1, 1|n.

(ii) Since 0 = 0m, m|0.

(iii) If m|n and n|q, n = tm and q = sn where t, s ∈ Z. So q = s(tm) = (st)m;


hence m|q.

(iv) Since m|n and m|q, n = tm and q = sm. So,

un + vq = u(tm) + v(sm)
= (ut + vs)m;

and hence, m|(un + vq).

(v) By assumption, m|1 and since ±1 divide 1, m = ±1. Thus m = 1 or m = −1.

(vi) If m|n and n|m then n = tm and m = sn where t, s ∈ Z. Thus m = stm;


st = 1. So, s = t = 1 or s = t = −1. Thus m = ±n.

Now we introduce the concept of a divisor of two (or more) integers, namely,
the greatest common divisor of these integers.

Definition 1.1.4. Given a, b (not both 0), then their greatest common divisor
c is defined by:

(i) c > 0.

(ii) c|a and c|b.

2
(iii) If d|a and d|b, then d|c.
We write this c as c = (a, b). In other words, the greatest common divisor of
a and b is the positive number c which divides a and b and is divisible by every
d which divides a and b.
Example 1.1.5. (3, 9) = 3, (1, 5) = 1 and (5, 8) = 1.
Note that we can easily find the greatest common divisor of any small integers.
But we cannot find easily for any two large integers. So we will give an algorithm
to find the greatest common divisor for these large integers after the following
theorem.
Theorem 1.1.6. If a and b are not both 0, then their greatest common divisor
c = (a, b) exists, is unique, and moreover, c = ma + nb for some suitable m and
n.
Proof. Let A = {ma + nb| m, n ∈ Z}. Since a, b are not both 0, a or b ∈ A. If
x ∈ A and x > 0, then −x is also in A and −x > 0, for if x = m1 a + n1 b, then
−x = (−m1 )a + (−n1 )b, so is in A. Thus, A has positive elements; hence by the
well-ordering principle there is a smallest positive element c, in A. Since c ∈ A,
by the form of the element of A, we have c = ma + nb for some m, n ∈ Z.
Now we claim that c is the greatest common divisor of a and b. If d|a and d|b,
then d|(ma + nb) by Lemma 1.1.3 (iv), that is d|c. So, to verify that c = (a, b)
we need only show that c|a and c|b.
By Euclid’s Algorithm, a = qc+r, where 0 ≤ r < c, that is a = q(ma+nb)+r.
Therefore, r = (1 − qm)a − (qn)b. So, r is in A. But r < c and is in A, so by the
choice of c, r cannot be positive. Hence r = 0; in other words, a = qc and so c|a.
Similarly, c|b.
For uniqueness of c, if t > 0 also satisfied t|a, t|b and d|t for all d such that d|a
and d|b, we would have t|c and c|t. By Lemma 1.1.3 (vi), t = c since both are
positive.
This theorem will illustrate by some examples. Now we find an explicit algo-
rithm to find greatest common divisor of a and b, (a, b). For instance, we direct
examine (24, 9) = 3. Note that
24 = (2 × 9) + 6
9 = (1 × 6) + 3
6 = (2 × 3) + 0
so that (24, 9) = 3 and
3 = 9 − [1 × 6]
= 9 − 1 × [24 − (2 × 9)]
= [3 × 9] + [(−1) × 24];

3
so, 3 = [3 × 9] + [(−1) × 24].
This note forces to construct the procedure to find the greatest common divisor
of a and b. In general, if a and b are large, we can find (a, b) by the following
algorithm:
(i) If b > a, then
b = qa + r1 , 0 ≤ r1 < a
a = q1 r1 + r2 , 0 ≤ r2 < r1
r1 = q2 r2 + r3 , 0 ≤ r3 < r2
..
.
rn−1 = qn rn + rn+1 , 0 ≤ rn+1 < rn .

(ii) If rn+1 = 0, rn = (a, b).


Now we will illustrate Theorem 1.1.6 by using this algorithm.
Example 1.1.7. We find (72, 25) and express (72, 25) as m(72) + n(25) where
m, n ∈ Z. We have
72 = [2 × 25] + 22
25 = [1 × 22] + 3
22 = [7 × 3] + 1
3 = [3 × 1] + 0,
so, (72, 25) = 1 and
1 = 22 + [(−7) × 3]
= 22 + (−7)[25 − 22]
= [8 × 22] + [(−7) × 25]
= 8[72 − (2 × 25)] − [7 × 25]
= 8(72) − 23(25).
Example 1.1.8. We find (116, −84) and express (116, −84) as m(116) + n(−84)
for some m, n ∈ Z. We have
116 = [(−1) × (−84)] + 32
−84 = [(−2) × 32] + (−20)
32 = [(−1) × (−20)] + 12
−20 = [(−1) × 12] + (−8)
12 = [(−1) × (−8)] + 4
−8 = [(−2) × 4] + 0,

4
so, (116, −84) = 4 and

4 = 12 − [(−1) × (−8)]
= 12 − [(−1) × (−20 − (−1) × 12)]
= (2 × 12) + (−20)
= 2[32 − (−1) × (−20)] + [(−84) − (−2) × 32]
= (4 × 32) + 2(−20) + (−84)
= (4 × 32) + 2[−84 − (−2) × 32] + (−84)
= 8(32) + 3(−84)
= 8[116 − (−1) × (−84)] + 3(−84)
= 8(116) + 11(−84).

1.2 On Congruence Classes Modulo n


We have seen that Z is the set of all integers. In this section, we describe
Z is the union of distinct equivalent classes modulo n. So we first consider the
informal definition of congruence.
If a and b are any two integers such that a − b = 0, then any positive integer
n divides a − b. Also, n divides 0, (n|(t · 0) = t(a − b)).
For instance, 5|(3−3 = 0), also, 5|(8·0), 5|(10·0),. . . . If n is a positive integer,
we say that two integers are congruent modulo n if their difference is a multiple
of n. To say that a − b = nk for some integer k means that n divides a − b. So
we have this formal definition:

Definition 1.2.1. Let a, b and n be integers with n > 0. Then a is congruent


to b modulo n [written a ≡ b(mod n)], provided that n divides a − b.

For example, since 4|(8 − 4), 8 ≡ 4(mod 4). Since 5|(8 + 12), 8 ≡ −12(mod 5).
Now we will collect properties on congruence.

Theorem 1.2.2. Let n be a positive integer. For all a, b, c ∈ Z,

(i) a ≡ a(mod n);

(ii) if a ≡ b(mod n), then b ≡ a(mod n);

(iii) if a ≡ b(mod n) and b ≡ c(mod n), then a ≡ c(mod n).

Proof. (i) Since a − a = 0, a ≡ a(mod n).

5
(ii) a ≡ b(mod n) means that a − b = nk for some integer k. Therefore

b − a = −(a − b)
= −nk
= n(−k).

So, n|b − a. Hence b ≡ a(mod n).

(iii) If a ≡ b(mod n) and b ≡ c(mod n), then by the definition of congruence,


there are integers k and t such that a − b = nk and b − c = nt. Therefore

(a − b) + (b − c) = nk + nt
a − c = n(k + t).

Thus n|(a − c) and hence a ≡ c(mod n).

Definition 1.2.3. A relation ∼ on a nonempty set S is called an equivalence


relation if, for all a, b, c ∈ S, it satisfies:

(i) a ∼ a(reflexivity).

(ii) a ∼ b implies that b ∼ a (symmetry).

(iii) a ∼ b, b ∼ c implies that a ∼ c (transitivity).

We define a ∼ b if a ≡ b(mod n). By Theorem 1.2.2, this relation is an equiva-


lence relation.

Theorem 1.2.4. If a ≡ b(mod n) and c ≡ d(mod n), then

(i) a + c ≡ b + d(mod n);

(ii) ac ≡ bd(mod n).

Proof. (i) By the definition of congruence, there are integers k and t such that
a − b = nk and c − d = nt. Therefore

n(k + t) = nk + nt
= (a − b) + (c − d)
= (a + c) − (b + d).

Thus n divides (a + c) − (b + d), so that a + c ≡ b + d(mod n).

6
(ii) Using the fact that −bc + bc = 0, we have
ac − bd = ac + 0 − bd
= ac − bc + bc − bd
= (a − b)c + b(c − d)
= (nk)c + b(nt)
= n(kc + bt).
Therefore n|(ac − bd), and so ac ≡ bd(mod n).

Definition 1.2.5. Let a and n be integers with n > 0. The congruence class
of a modulo n (denoted [a]) is the set of all those integers that are congruent to
a modulo n, that is,
[a] = {b ∈ Z|a ≡ b(mod n)}.
To say that b ≡ a(mod n) means that b − a = nk for some integer k or,
equivalently, that b = a + nk. Thus
[a] = {b| b ≡ a(mod n)}
= {b| b = a + nk, k ∈ Z}
= {a + nk| k ∈ Z}.
Example 1.2.6. In congruence modulo 5, we have
[0] = {0 + 5k| k ∈ Z}
= {..., −15, −10, −5, 0, 5, 10, 15, ...};
[1] = {..., −14, −9, −4, 1, 6, 11, 16, ...};
[2] = {..., −13, −8, −3, 2, 7, 12, ...};
[3] = {..., −12, −7, −2, 3, 8, 13, ...};
[4] = {..., −11, −6, −1, 4, 9, 14, 19, ...};
[5] = {..., −10, −5, 0, 5, 10, 15, ...};
[6] = {..., −9, −4, 1, 6, 11, 16, ...}.
Note that
[0] = [5] = [10] = · · ·
[1] = [6] = [11] = · · ·
[2] = [7] = [12] = · · ·
[3] = [8] = [13] = · · ·
[4] = [9] = [14] = · · · ,
and so
Z = [0] ∪ [1] ∪ [2] ∪ [3] ∪ [4].

7
We will investigate how any two classes are the same.

Theorem 1.2.7. a ≡ c(mod n) if and only if [a] = [c].

Proof. Assume that a ≡ c(mod n). To prove that [a] = [c], we first show that
[a] ⊆ [c]. Let b ∈ [a]. Then by definition of congruence class, b ≡ a(mod n).
By transitivity, b ≡ c(mod n). By definition of congruence class, b ∈ [c]. Hence
[a] ⊆ [c]. To prove [c] ⊆ [a], let d ∈ [c]. Then by definition of congruence class,
d ≡ c(mod n). By transitivity, d ≡ a(mod n). By definition of congruence class,
d ∈ [a](mod n), so that [c] ⊆ [a]. Hence [a] = [c].
Conversely, suppose that [a] = [c]. Since a ≡ a(mod n) by reflexivity,
a ∈ [a] = [c]. By definition of congruence class, a ≡ c(mod n).

Now we recall the definition on set:

Definition 1.2.8. Let A and B be any two sets. Then A and B are disjoint if
A ∩ B = ∅. Then A and B are intersect if A ∩ B 6= ∅.

Corollary 1.2.9. Two congruence classes modulo n are either disjoint or iden-
tical.

Proof. If [a] and [c] are disjoint, there is nothing to prove. Suppose that [a] ∩ [c]
is nonempty. Then there is an integer b with b ∈ [a] and b ∈ [c]. By the definition
of congruence class, b ≡ a(mod n) and b ≡ c(mod n). Therefore by symmetry
and transitivity, a ≡ c(mod n). Hence [a] = [c] by Theorem 1.2.7.

Corollary 1.2.10. There are exactly n distinct congruence classes modulo n,


namely, [0], [1], [2], ..., [n − 1].

Proof. We first claim that no two of 0, 1, 2, ..., n − 1 are congruent modulo n. To


see this, suppose that 0 ≤ s < t < n. Then t − s is a positive integer and less
than n. Thus n does not divide t − s and hence t 6≡ s(mod n). Since no two of
0, 1, 2, ..., n − 1 are congruent, the classes [0], [1], [2], ..., [n − 1] are all distinct, by
Theorem 1.2.7. To complete the proof, we need only show that every congruence
class is one of these n classes. Let a ∈ Z. By the Division Algorithm, a = nq + r,
with 0 ≤ r < n. Thus a − r = nq, so that a ≡ r(mod n). By Theorem 1.2.7,
[a] = [r]. Since 0 ≤ r < n, [a] is one of [0], [1], [2], . . . , [n − 1].

1.3 Programs in Maple 2018


We write a program for Euclid’s algorithm in Maple 2018 software. A Maple
procedure is a program consisting of Maple statement.

8
1.3.1 Integers
We first write some commands in Maple to write procedure for Euclid’s al-
gorithm.
The command irem(a,b) computes the integer remainder of a divided by b.
The command iquo(a,b) computes the integer quotient of a divided by b. For
example
[> a := 20; b := 7;

a := 20
b := 7

[> r := irem(a, b);


r := 6
[> q := iquo(a, b); printf(“Quotient=% d \ n ”,q);

q := 2
Quotient = 2

It should be the case that a = bq + r. We check as follows:


[> a = b · q + r;
20 = 20
The command igcd(a,b) and ilcm(a,b) compute the greatest common divisor
and least common multiple of integers a and b, respectively.
[> igcd(6, 4);
2
[> ilcm(6, 4);
12

1.3.2 Loops and Statements


To do a sequence of calculations it will be handy to know how to use some of
Maple’s looping commands and also the if command. This has the following
forms.

if<condition>then<statements>else<statements>fi

or just

if<condition>then<statements>fi

9
[>if 3 > 1 then print(good) else print(bad) fi;

good

To execute one or more statements zero or more times in a loop use the for
command. It has the following form
for<variable>from<start>to<end>do<statements>od
[> for i from 1 to 5 do i3 ;od;
1
8
27
64
125
To execute some statements while a condition is true use the while loop. It
has the syntax

while< condition > do < statements > od

In the following example we repeatedly divide an integer n by 2 until it is odd.


[> n := 16; while irem(n, 2) = 0 do n := iquo(n, 2);od;

n := 16

n := 8
n := 4
n := 2
n := 1
[> add (i2 , i = 1..5);
55

1.3.3 Procedures
The functions defined by the arrow operator are a special case of a procedure.
The general form of a procedure is as follows.
procedure name : proc (variables)
command sequence
end proc:

10
For example
[> f x :=piecewise(x < 0, 0, x < 1, x2 , 2 ∗ x − 1);



 0, x < 0;

f x := x2 , x<1



2x − 1, otherwise.

We could use the Maple procedure to find the above function.


[> f : = proc(x)
if x < 0 then return 0
elif x < 1 then return x2
else return 2x − 1
end if;
end proc:
[> f (0); f (1/2); f (3);
0
1/4
5
If we wish to use one or more variables only inside the procedure, it is good
practise to declare them to be local variables. For example, we may wish to
use the variable i with a for loop. To do this, place the Maple command local i; at
the top of our procedure definition. For example, a simple function or procedure
to sum the first n integers could be defined in the following way.
[> f : = proc(n)
local i, s;
s := 0;
for i from 1 to n do
s := s + i;
end do;
s;
end proc:
[> f (3); f (5);
6
15
Now we write a procedure for Euclid’s algorithm of all positive values.
[> EuclideanAlgorithm :=proc(a::integer,b::integer) local c, d, r;
(c, d):=(abs(a),abs(b));

11
while d 6= 0 do r := irem(c, d); (c, d) := (d, r);od;
c;
end proc:
[> EuclideanAlgorithm (24, 210);

The simplest debugging tool is to insert print statements in the procedure. For
example
[> EuclideanAlgorithm :=proc(a::integer,b::integer) local c, d, r;
(c, d):=(abs(a),abs(b));
while d 6= 0 do r := irem(c, d); print(r); (c, d) := (d, r);od;
c;
end proc:
[> EuclideanAlgorithm (24, 210);
24
18
6
0
6
The next simplest debugging tool is the trace command.
[>trace(EuclideanAlgorithm);

EuclideanAlgorithm

[>EuclideanAlgorithm (24, 210);

c, d := 24, 210

r := 24
24
c, d := 210, 24
r := 18
18
c, d := 24, 18
r := 6
6

12
c, d := 18, 6
r := 0
0
c, d := 6, 0
6
6
Here we print the quotients in the Euclid’s algorithm. We notice the three
argument version of the iquo command. It computes and returns the quotient
but assigns the third input (a variable) the value of the remainder.
[> EuclideanAlgorithm :=proc(a::integer,b::integer) local c, d, r, q;
(c, d):=(abs(a),abs(b));
while d 6= 0 do
r := irem(c, d, q);
printf (“Quotient=%d \ n”, q); (c, d) := (d, r);od;
c;
end proc:
[> EuclideanAlgorithm (24, 210);

Quotient =0
Quotient =8
Quotient =1
Quotient =3
6

Here is a recursive implementation of Euclid’s algorithm.


[> EuclideanAlgorithm:=proc(a::integer,b::integer)
if a < 0 then EuclideanAlgorithm(−a, b)
elif b < 0 then EuclideanAlgorithm(a, −b)
elif a < b then EuclideanAlgorithm(b, a)
elif b = 0 then a
else EuclideanAlgorithm(b, irem(a, b))
fi;
end:
[>EuclideanAlgorithm (24, 12);
12

13
Chapter 2

Groups of Integers Modulo n

We have seen that [0], [1], [2], ..., [n − 1] are distinct classes under modulo n. So
we define Zn = {[0], [1], [2], . . . , [n − 1]} as analogously Z is the set of all integers.
We discuss the algebraic structure containing a single operation and a set, namely
a group. In this chapter we prove that Zn is a group under the addition of any
two classes. But Zn is not a group under the multiplication of any two classes.
But a piece of Zn is a group under multiplication.

2.1 Additive Groups of Integers Modulo n


In this section, Zn is a group under the addition of two classes. So, we first
show that the operation ⊕n , addition of any two classes is well-defined.

Proposition 2.1.1. Let Zn = {[0], [1], [2], . . . , [n − 1]}.


Define ⊕n : Zn × Zn −→ Zn by

[a] ⊕n [b] = [a + b],

for all [a] and [b] in Zn . Then this ⊕n is well-defined.

Proof. Suppose ([a], [b]) = ([c], [d]). Then [a] = [c] and [b] = [d], and so
a ≡ c(mod n) and b ≡ d(mod n), by Theorem 1.2.7. By Theorem 1.2.4,
a + b ≡ c + d(mod n). By Theorem 1.2.7, [a + b] = [c + d]. Thus we have proved
that ⊕n is well-defined.

Now we construct a Cayley table for Z3 and Z4 as follows:

14
⊕4 [0] [1] [2] [3]
⊕3 [0] [1] [2]
[0] [0] [1] [2] [3]
[0] [0] [1] [2]
[1] [1] [2] [3] [0]
[1] [1] [2] [0]
[2] [2] [3] [0] [1]
[2] [2] [0] [1]
[3] [3] [0] [1] [2]
Table 2.1.1: Addition of elements in Z3 and Z4

From Table 2.1.1, we have seen that the operation ⊕3 is well-defined. This is also
called closure law on Z3 × Z3 . Also, since

([1] ⊕3 [2]) ⊕3 [1] = [0] ⊕3 [1] = [1]

and

[1] ⊕3 ([2] ⊕3 [1]) = [1] ⊕3 [0] = [1],

([1] ⊕3 [2]) ⊕3 [1] = [1] ⊕3 ([2] ⊕3 [1]). This is called associative law on Z3 × Z3 if
⊕3 satisfies for each element in Z3 . We see that [0] is the identity element in Z3 .
For instance, [0] ⊕3 [1] = [1], [1] ⊕3 [0] = [1]. The additive inverse of each element
in Z3 are [0], [2] and [1]. In other words, the inverse of [0], [1] and [2] are [3 − 0],
[3 − 1] and [3 − 2]. So, Z3 is a group under ⊕3 . Similarly, Z4 is a group under
⊕4 .
Now we recall a formal definition of a group.

Definition 2.1.2. A nonempty set G is called a group under the operation ∗


on it, a function ∗ : G × G −→ G satisfied the following axioms (conditions):

(i) a ∗ b ∈ G for all a, b ∈ G(closure law).

(ii) (a ∗ b) ∗ c = a ∗ (b ∗ c) for all a, b, c ∈ G(associative law).

(iii) There exists an element e ∈ G such that e ∗ a = a ∗ e = a for all a ∈ G(e is


called the identity or unit element of G).

(iv) There exists an element b ∈ G such that a ∗ b = b ∗ a = e for each a ∈ G(b


is called the inverse element of a in G).

If for every a, b ∈ G, a ∗ b = b ∗ a then G is called an abelian group.

Theorem 2.1.3. Zn is an abelian group under the operation in Proposition 2.1.1.

15
Proof. Clearly, closure law is satisfied, by Proposition 2.1.1. For any [a], [b], [c]
in Zn , we have

([a] ⊕n [b]) ⊕n [c] = [a + b] ⊕n [c] by definition of ⊕n ,


= [(a + b) + c] by definition of ⊕n ,
= [a + (b + c)] since a, b, c ∈ Z,
= [a] ⊕n [b + c] by definition of ⊕n ,
= [a] ⊕n ([b] ⊕n [c]) by definition of ⊕n .

So, associative law is satisfied. For any [a] ∈ Zn , we have

[a] ⊕n [0] = [a + 0] = [a] = [0 + a] = [0] ⊕n [a];

so, [0] is the identity element in Zn . For any [a] ∈ Zn , we have

[a] ⊕n [n − a] = [a + n − a] = [0] = [n − a] ⊕n [a].

So, [n − a] is an inverse of [a]. For any [a], [b] ∈ Zn , we have

[a] ⊕n [b] = [a + b] by definition of ⊕n ,


= [b + a] since a, b ∈ Z,
= [b] ⊕n [a] by definition of ⊕n .

So, Zn is an abelian group under ⊕n .

Definition 2.1.4. A group G is called cyclic if there is an element a in G such


that G = {an |n ∈ Z}, write G = hai. Such an element a is called a generator
of G.

Example 2.1.5. Z5 is a cyclic group. In fact, Z5 = {[0], [1], [2], [3], [4]}, [1] is a
generator of Z5 since

[2] = [1 + 1] = [1] ⊕5 [1],


[3] = [1] ⊕5 [1] ⊕5 [1],
[4] = [1] ⊕5 [1] ⊕5 [1] ⊕5 [1],
[0] = [5] = [1] ⊕5 [1] ⊕5 [1] ⊕5 [1] ⊕5 [1].

Note that [2], [3] and [4] are generators of Z5 . A cyclic group has one or more
generators.

Now we show that Zn is a cyclic group under ⊕n .

Theorem 2.1.6. Zn is a cyclic group under ⊕n .

16
Proof. Let Zn = {[0], [1], [2], . . . , [n − 1]}. Then we have

[2] = [1] ⊕n [1] = 2[1]


[3] = 3[1]
..
.
[n − 1] = (n − 1)[1]
[0] = n[1];

so, [1] is a generator of Zn . Thus the proof is complete.

2.2 Multiplicative Groups of Integers Modulo n


We have seen that Zn is a group under addition, ⊕n . We have a question, is
Zn is a group under multiplication, ⊗n ? In general, it is not a group. In this
section, we study how to group under this operation. So, we determine whether
or not this operation is well-defined.
Theorem 2.2.1. The function ⊗n : Zn × Zn −→ Zn by

⊗n ([a], [b]) = [a] ⊗n [b] = [ab],

for all [a], [b] ∈ Zn . Then ⊗n is well-defined.


Proof. Suppose ([a], [b]) = ([c], [d]). Then [a] = [c] and [b] = [d]. By Theorem
1.2.7, a ≡ c(mod n) and b ≡ d(mod n). By Theorem 1.2.4, ab ≡ cd(mod n). By
Theorem 1.2.7, [ab] = [cd]. So, ⊗n is well-defined.

Now we will illustrate Zn is not a group under ⊗n .


Example 2.2.2. Zn is not a group under ⊗n . In fact, closure law is satisfied, by
Theorem 2.2.1. [1] is the identity element since

[a] ⊗n [1] = [a1] = [a] = [1a] = [1] ⊗n [a],

for all [a] ∈ Zn . Since

[0] ⊗n [a] = [0a] = [0],

[0] has not an inverse in Zn .


We have seen that zero element [0] has not an inverse under ⊗n . Moreover,
nonzero element in Zn has not inverse.
Example 2.2.3. Z6 is not a group under ⊗n since [2] has not inverse by Table
2.2.1.

17
⊗6 [0] [1] [2] [3] [4] [5]
[0] [0] [0] [0] [0] [0] [0]
[1] [0] [1] [2] [3] [4] [5]
[2] [0] [2] [4] [0] [2] [4]
[3] [0] [3] [0] [3] [0] [3]
[4] [0] [4] [2] [0] [4] [2]
[5] [0] [5] [4] [3] [2] [1]
Table 2.2.1: Multiplication of elements in Z6
Note that Zn is not a group under ⊗n since nonzero element in Zn has not
inverse. So we find a piece of Zn which is a group under ⊗n .
Definition 2.2.4. U (n) = {[a] ∈ Zn |(a, n) = 1}. It is called the set of units in
Zn .
Example 2.2.5. U (2) = {[1]} is a group under ⊗2 since [1] ⊗2 [1] = [1].
U (4) = {[1], [3]} is an abelian group under ⊗4 by Table 2.2.2.
⊗4 [1] [3]
[1] [1] [3]
[3] [3] [1]
Table 2.2.2: Multiplication of elements in U (4)
U (5) = {[1], [2], [3], [4]} is an abelian group under ⊗5 by Table 2.2.3.
⊗5 [1] [2] [3] [4]
[1] [1] [2] [3] [4]
[2] [2] [4] [1] [3]
[3] [3] [1] [4] [2]
[4] [4] [3] [2] [1]
Table 2.2.3: Multiplication of elements in U (5)
In general:
Theorem 2.2.6. U (n) is an abelian group under ⊗n .
Proof. If [a], [b] ∈ U (n) then (n, a) = 1, and (n, b) = 1. Suppose d = (n, ab)
where d > 1. So, d|n and d|ab; hence d|a or d|b; which contradicts to d = 1. Thus
(n, ab) = 1; hence [ab] ∈ U (n). For any [a], [b], [c] ∈ U (n), we have
([a] ⊗n [b]) ⊗n [c] = [ab] ⊗n [c]
= [(ab)c]
= [a(bc)]
= [a] ⊗n [bc]
= [a] ⊗n ([b] ⊗n [c]);

18
so, associative law is satisfied. For any [a] ∈ U (n),

[a] ⊗n [1] = [a1] = [a] = [1a] = [1] ⊗n [a];

[1] is the identity element in U (n). For any [a] ∈ U (n), we have

1 = sn + ta, where s, t ∈ Z.

So,

[1] = [sn + ta] = [s][n] ⊕n [t][a]


= [t][a];

and [t] is an inverse of [a]. Since 1 = sn + ta, (n, t) = 1; so [t] ∈ U (n). For any
[a], [b] ∈ U (n), we have

[a] ⊗n [b] = [ab] by definition of ⊗n ,


= [ba] since a, b ∈ Z,
= [b] ⊗n [a] by definition of ⊗n .

So, U (n) is an abelian group under ⊗n .

This group U (n) is called the group of units in the integers mod n.

Definition 2.2.7. The number of elements of a group G (finite or infinite) is


called its order. The order of G is denoted by |G|.

Example 2.2.8. In Example 2.2.5, |U (2)| = 1, |U (4)| = 2, |U (5)| = 4.

Definition 2.2.9. The order of an element g in a group G is the smallest


positive integer n such that g n = e. If no such integer exists, we say that g has
infinite order. The order of an element g is denoted by |g|.

Example 2.2.10. The order of each element in Z6 in Example 2.2.3 are:

|[0]| = 1, |[1]| = 6, |[2]| = 3, |[3]| = 2, |[4]| = 3, |[5]| = 6.


The order of each element in U (5) in Example 2.2.5 are:

|[1]| = 1, |[2]| = 4, |[3]| = 4, |[4]| = 2.

19
Chapter 3

Cyclic Groups

In this chapter, we discuss a finite group of integers additive modulo n and


multiplicative modulo n. We also present a formula to find subgroups of Zn . We
also present a structure of all subgroups of Zn , namely, subgroup lattice. So, we
collect some definitions.

3.1 Subgroup Tests


In set theory, we know that every set is a subset of itself. So, every group is
a subgroup of itself under a given operation. We know that a nonempty set is a
group under the given operation if every element in a given group satisfies the
four axioms of Definition 2.1.2 under the given operation. This process take a
long time. So, we discuss a test on a nonempty subset of a finite group. So, we
recall the formal definition of subgroup of a group.

Definition 3.1.1. If a subset H of a group G is itself a group under the operation


of G, we say that H is a subgroup of G. We write H ≤ G if H is a subgroup of
G.

Example 3.1.2. {[0], [2]} is a subgroup of Z4 = {[0], [1], [2], [3]}. Also {[1], [2], [4]}
is a subgroup of U (7) = {[1], [2], [3], [4], [5], [6]}.

Theorem 3.1.3. (One-step subgroup test) Let G be a group and H a nonempty


subset of G. If ab−1 is in H whenever a and b are in H, then H is a subgroup of
G. (In additive notation, if a − b is in H whenever a and b are in H, then H is
a subgroup of G).

Proof. If x ∈ H, then x ∈ G since H ⊆ G. Hence every element in H satisfies


associative law. Next, we show that the identity element e ∈ H. Since H is a
nonempty, we may take some x ∈ H. Then, letting a = x and b = x in the

20
hypothesis, we have

e = xx−1 = ab−1 ∈ H.

Next, we show that x−1 ∈ H for every x ∈ H. Then, letting a = e and b = x in


the hypothesis, we have

x−1 = ex−1 = ab−1 ∈ H.

Finally, we show that every element in H satisfies the closure law. For every
x, y ∈ H, we have shown that y −1 ∈ H. Then, letting a = x and b = y −1 in the
hypothesis, we have

xy = x(y −1 )−1 = ab−1 ∈ H.

By Definition 3.1.1, H is a subgroup of G.


Theorem 3.1.4. (Two-step subgroup test) Let G be a group and let H be a
nonempty subset of G. If ab is in H whenever a and b are in H, and a−1 is in
H whenever a is in H, then H is a subgroup of G.
Proof. Suppose a, b ∈ H. Then ab−1 ∈ H by assumption. Hence by Theorem
3.1.3, H is a subgroup of G.
Theorem 3.1.5. (Finite subgroup test) Let H be a nonempty finite subset of a
group G. If H is closed under the operation of G, then H is a subgroup of G.
Proof. To prove theorem, we use Theorem 3.1.4. By assumption, closure law is
satisfied. So, we need only prove that if a ∈ H, then a−1 ∈ H. If a = e, then
a−1 = e−1 = e ∈ H. If a 6= e, consider the sequence a, a2 , a3 , .... By closure
property, all of these elements belong to H. Since H is finite, not all of these
elements are distinct. Let ai = aj and i > j. Then ai−j = e, and since a 6= e,
i−j > 1. Thus, aai−j−1 = ai−j = e, and therefore, ai−j−1 = a−1 . But i−j −1 ≥ 1
implies ai−j−1 ∈ H.
Theorem 3.1.6. Let G be a group, and let a be any element of G. Then, hai is
a subgroup of G.
Proof. Since a ∈ hai, hai is nonempty. Let an , am ∈ hai. Then

an (am )−1 = an−m ∈ hai;

so, by Theorem 3.1.3, hai is a subgroup of G.


Example 3.1.7. In U (10), h[3]i = {[3], [9], [7], [1]} is a subgroup of U (10).
Example 3.1.8. In Z10 , h[0]i = h[10]i,h[1]i = h[3]i = h[7]i = h[9]i,
h[2]i = h[4]i = h[6]i = h[8]i, h[5]i are subgroups of Z10 .

21
3.2 Properties of Cyclic Groups
We have known that if every element in a group G can be expressed as a power
of an element in G, G is called cyclic group. In this section, we discuss how to
find this element for Zn .

Example 3.2.1. The set of all integers Z is cyclic group under ordinary addition.
In fact, every positive integer n is the addition of 1 with n terms and negative
integer n is the addition of −1 with n terms.
We observed that Z has only two generators but Zn may have many generators.

Example 3.2.2. Z9 = h[1]i = h[2]i = h[4]i = h[5]i = h[7]i = h[8]i. Z9 = h[4]i


since h[4]i = {[4], [4]+[4], [4]+[4]+[4], ...} is the set {[4], [8], [3], [7], [2], [6], [1], [5], [0]}.
Thus [4] is a generator of Z9 . On the other hand, [3] is not a generator since
h[3]i = {[3], [6], [0]} =
6 Z9 .

Example 3.2.3. U (5) = {[1], [2], [3], [4]} = h[2]i = h[3]i.So, [2] and [3] are gener-
ators for U (5).

Theorem 3.2.4. Let G be a group, and let a belong to G. If a has infinite


order, then ai = aj if and only if i = j. If a has finite order, say n, then
hai = {e, a, ..., an−1 } and ai = aj if and only if n divides i − j.

Proof. If a has infinite order, there is no nonzero n such that an is the identity.
Since ai = aj implies ai−j = e, we must have i − j = 0.
Now assume that |a| = n. We will prove that hai = {e, a, ..., an−1 }. By
definition of hai, the elements e, a, ..., an−1 ∈ hai.
Now suppose that ak is an arbitrary member of hai. By the Division algorithm,
there exist integers q and r such that

k = qn + r, 0 ≤ r < n.

Then

ak = aqn+r = (anq )ar = (an )q ar = ar ,

so that ak ∈ {e, a, ..., an−1 }. This proves that hai = {e, a, ..., an−1 }.
Next, we assume that ai = aj and prove that n divides i − j. We observe that
a = aj implies ai−j = e. Again, by the Division algorithm,
i

i − j = qn + r, 0 ≤ r < n.

Then

ai−j = aqn+r = (aqn )ar = (an )q ar = ar .

22
Since n is the smallest positive integer such that an = e, r = 0. Thus n divides
i − j.
Conversely, if i − j = nq, then

ai−j = anq = e.

So, ai = aj .

Theorem 3.2.4 reveals order of the group generated by an element in a group and
order of this element.

Corollary 3.2.5. For any group element a, |a| = |hai|.

Corollary 3.2.6. Let G be a group and let a be an element of order n in G. If


ak = e, then n divides k.

Proof. Since ak = e = a0 , n divides (k − 0) by Theorem 3.2.4 .

The following theorem gives to find |ak | if we know hak i = haj i.

Theorem 3.2.7. Let a be an element of order n in a group and let k be a positive


integer. Then hak i = ha(n,k) i and |ak | = n/(n, k).

Proof. Let d = (n, k) and k = dr. Since ak = (ad )r , hak i ⊂ had i. There are
integers s and t such that d = ns + kt. So,

ad = ans+kt = (an )s (ak )t = (ak )t ∈ hak i.

Thus had i ⊂ hak i. So, we have proved that hak i = ha(n,k) i.


To prove the second part, we claim that |ad | = n/d. Since (ad )n/d = an = e,
|a | ≤ n/d. On the other hand, if 0 < i < n/d, then (ad )i 6= e by definition of |a|.
d

So, we have |ak | = |hak i| = |ha(n,k) i| = |a(n,k) | = n/(n, k).

Example 3.2.8. If |a| = 20, then

ha14 i = ha(20,14) i = ha2 i

and

ha8 i = ha(20,8) i = ha4 i.

Hence |a14 | = 20/2 = 10 and |a8 | = 20/4 = 5.

Corollary 3.2.9. Let |a| = n. Then hai i = haj i if and only if (n, i) = (n, j) and
|ai | = |aj | if and only if (n, i) = (n, j).

23
Proof. By Theorem 3.2.7,hai i = ha(n,i) i and haj i = ha(n,j) i. If hai i = haj i, then
ha(n,i) i = ha(n,j) i, so that (n, i) = (n, j).
Conversely, if (n, i) = (n, j), then ha(n,i) i = ha(n,j) i; so hai i = haj i.
The second part is satisfied since

(n, i) = (n, j) ⇐⇒ n/(n, i) = n/(n, j)


⇐⇒ |ai | = |aj |.

Corollary 3.2.10. Let |a| = n. Then hai = haj i if and only if (n, j) = 1 and
|a| = |aj | if and only if (n, j) = 1.

Proof. By Corollary 3.2.9,

hai = haj i ⇐⇒ (n, 1) = (n, j)


⇐⇒ 1 = (n, j)

and

|a| = |aj | ⇐⇒ (n, 1) = (n, j)


⇐⇒ 1 = (n, j).

Corollary 3.2.11. An integer k in Zn is a generator of Zn if and only if


(n, k) = 1.

Proof. The element [1] in Zn has order n and it is the generator of Zn . By


Corollary 3.2.10,

h[1]i = hk([1])i ⇐⇒ (n, 1) = (n, k),


h[1]i = hki ⇐⇒ 1 = (n, k).

Example 3.2.12. The generators of Z6 are [1] and [5].


The generators of Z9 are [1], [2], [4], [5], [7] and [8].

3.3 Subgroups of Cyclic Groups


In this section, we present the theorem which is how many subgroups a finite
cyclic group and how to find them.

24
Theorem 3.3.1. Every subgroup of a cyclic group is cyclic. Moreover, if
|hai| = n , then the order of any subgroup of hai is a divisor of n; and, for each
positive divisor k of n, the group hai has exactly one subgroup of order k, namely,
han/k i.

Proof. Let G = hai and suppose that H is a subgroup of G. We must show that
H is cyclic. If H = {e}, then H is cyclic. So we may assume that H 6= {e}.
We now claim that H contains an element of the form at , where t is positive.
Since G = hai, every element of H has the form at ; and when at belongs to H
with t < 0, then a−t belongs to H also and −t is positive.
Now let m be the least positive integer such that am ∈ H. By closure,

ham i ⊂ H.

We next claim that H = ham i. To prove this claim, it suffices to let b be an


arbitrary member of H and show that b is in ham i. Since b ∈ G = hai, we
have b = ak for some k. Since |a| = m, there are two integers r, s such that
k = ms + r with 0 ≤ r < m. Then ak = ams+r = (am )s ar so that ar = a−ms ak .
Since ak = b ∈ H and a−ms = (a+m )−s is in H also, ar ∈ H. But, m is the least
positive integer such that am ∈ H, and 0 ≤ r < m, so r must be 0. Therefore,
b = ak = ams = (am )s ∈ ham i. Therefore, we have just proved that every
subgroup of a cyclic group is cyclic.
To prove the second part of theorem, suppose that |hai| = n and H is any
subgroup of hai. We have already shown that H = ham i, where m is the least
positive integer such that am ∈ H. Since an = e, n = ms.
To prove the last part of theorem, let k be any positive divisor of n. We will
show that han/k i is the one and only subgroup of hai of order k. By Theorem
3.2.7, |han/k i| = n/(n, n/k) = n/(n/k) = k.
Now, let H be any subgroup of hai of order k. We have shown that H = ham i
where m is a divisor of n. Then

m = (n, m)

and
k = |am | = |a(n,m) | = n/(n, m) = n/m.
Thus m = n/k and H = han/k i.

In general, if hai has order n and k divides n, then han/k i is the unique subgroup
of order k.
By taking a cyclic group in Theorem 3.3.1 to be Zn and a to be [1], we obtain
the following important special case.

Corollary 3.3.2. For each positive divisor k of n, the set hn/ki is the unique
subgroup of Zn of order k; moreover, these are the only subgroups of Zn .

25
Example 3.3.3. We list the elements of the subgroups h[20]i and h[10]i in Z30 .
The elements of the subgroup h[20]i are [20], [10] and [0]. The elements of the
subgroup h10i are [10], [20] and [0]. Let |a| = 30 in Z30 . Also ha20 i = {a20 , a10 , e}
and ha10 i = {a10 , a20 , e}.

Now we illustrate the connection between subgroups of a finite group. So, we


recall first a partial relation on a set S.

Definition 3.3.4. A relation ∼ on a nonempty set S is called a partial order


relation if, for all a, b, c ∈ S, it satisfies:

(i) a ∼ a(reflexivity).

(ii) a ∼ b and b ∼ a implies that a = b (antisymmetry).

(iii) a ∼ b, b ∼ c implies that a ∼ c (transitivity).

(S, ∼) is called a partially order set.


The collection of subgroups of a group forms a complete lattice, that is, a
nonempty partially ordered set in which every subset has a greatest lower bound
and a least upper bound.
Note that the set of all subgroups of a cyclic group under the subgroup relation
is a lattice. This lattice is called the subgroup lattice of the given group.
We write to simplify Zn = {0, 1, 2, . . . , n − 1} instead of
Zn = {[0], [1], [2], . . . , [n − 1]}.

Example 3.3.5. The subgroups of Z40 are:

h1i = {0, 1, 2, ..., 39} order 40,


h2i = {0, 2, 4, ..., 38} order 20,
h4i = {0, 4, 8, ..., 36} order 10,
h5i = {0, 5, 10, 15, 20, 25, 30, 35} order 8,
h8i = {0, 8, 16, 24, 32} order 5,
h10i = {0, 10, 20, 30} order 4,
h20i = {0, 20} order 2,
h40i = {0} order 1.

Example 3.3.6. The elements of U (40) are as follows:


U (40) = {1, 3, 7, 9, 11, 13, 17, 19, 21, 23, 27, 29, 31, 33, 37, 39}.

26
The subgroups of U (40) are:

h1i = {1} order 1,


h3i = {1, 3, 9, 27} order 4,
h7i = {1, 7, 9, 23} order 4,
h9i = {1, 9} order 2,
h11i = {1, 11} order 2,
h13i = {1, 13, 9, 37} order 4,
h17i = {1, 17, 9, 33} order 4,
h19i = {1, 19} order 2,
h21i = {1, 21} order 2,
h23i = {1, 23, 9, 7} order 4,
h27i = {1, 27, 9, 3} order 4,
h29i = {1, 29} order 2,
h31i = {1, 31} order 2,
h33i = {1, 33, 9, 17} order 4,
h37i = {1, 37, 9, 13} order 4,
h39i = {1, 39} order 2.

Now we draw the subgroup lattices of additive groups integer modulo n and
multiplicative group integer modulo n by using Maple.

27
Example 3.3.7.

Figure 3.3.1: Subgroup Lattice of Z40

28
Figure 3.3.2: Subgroup Lattice of Z110

29
Figure 3.3.3: Subgroup Lattice of Z8

30
Bibliography

[1] Bernardin L., Chin P., DeMarco P., Geddes K. O., Hare D. E. G.,Heal K.
M., Labahn G., May I. P., McCarron J., Monagan M. B., Ohashi D., and
Vorkoetter S. M., “Maple 18 Programming Guide - Maplesoft”,
http://www.maplesoft.com > documentation center > maple18

[2] Dummit D. S. and Foote R. M., “Abstract Algebra”, Third Edition, John
Wiley and Sons, Inc., NJ., 2004.

[3] Gallian J.A., “Contemporary Abstract Algebra”, Seventh Edition,


BROOK/COLE Cengage Learning, USA., 2010.

[4] Herstein I.N., “ Abstract Algebra”, Third Edition, Prentice-Hill, Inc., New
Jersey, 1996.

[5] Monagan M.,“MAPLE Notes for Computer Algebra Maple as a ... -


CECM”,
http://www.cecm.sfu.ca >∼mmonagan > teaching > CAS17

[6] Mihailovs A., “Abstract Algebra with Maple - DocPlayer.net”,


http:// www.docplayer.net > 86866811-Abstract-algebra-with-maple

31

You might also like