Group Theory
Group Theory
∗
Lecture Notes
2021/04/07
1 Group Theory
1.1 Sets, Functions and equivalence Relations
Sets
• Set operations:
set (A = {. . . | . . . }), ∈ (an element of), ⊆ (subset), ⊂ (proper subset), ∪ (union), ∩ (intersec-
− B) (difference) , cartesian product (A × B = {(a, b)| a ∈ A, b ∈ B}), cardinality
tion), (A
# elements in A, if A is finite
(|A| = .
∞, Otherwise
• Logical Operations:
∀ (for each), ∃ (there exist), ! (unique), =⇒ (implies, if then), ⇐⇒ (if and only if, iff, necessary and
sufficient).
• Number Systems:
Z = {0, ±1, ±2, . . . } (integral numbers or integers)
Q = {a/b | a, b ∈ Z, b 6= 0} (rational numbers or rationals)
R = {all decimal expressions ± d1 d2 . . . dn .a1 a2 a3 . . . } (real numbers or reals)
C = {a + bi | a, b ∈ R, i2 = −1} (complex numbers)
Z+ , Q+ and R+ denote the positive (nonzero) in Z, Q and R respectively.
Z ⊂ Q ⊂ R ⊂ C (all proper inclusions)
Functions
f
Definition. A function f : A → B (also written A −→ B) between two sets A (domain) and B
(codomain) is defined as:
∗
This lecture is loosely based on the book: D. S. Dummit & R. M. Foote, Abstract Algebra, 3rd Edition, John
Wiley & Sons, Inc.
1
”a rule a 7→ f (a) that associates, to each element a ∈ A, a unique element f (a) ∈ B”.
Or (eqivalently)
”i) f (a) ∈ B, ∀ a ∈ A and ii) a1 = a2 ⇒ f (a1 ) = f (a2 ), f or a1 , a2 ∈ A”.
• Function Examples:
Identity function: IdA : A → A, a 7→ a.
Inclusion of a subset A ⊆ S : A ,→ S, a 7→ a.
Restriction of f : S → T to a subset A ⊆ S : f |A : A → T, a 7→ f (a).
Projections: S ← S × T → T, s ← (s, t) 7→ t.
Constant functions: S → T , s 7→ t0 , where t0 is a fixed element of T .
Composition of functions: Given functions f : A → B and g : B → C, define g ◦ f : A → C by
• More on functions:
Let f : A → B be a function.
f −1 (C) = {a ∈ A | f (a) ∈ C}
consisting of the elements of A mapping into C under f is called the preimage or inverse image of
C under f .
Remark 1. For each b ∈ B, the preimage of {b} under f (i.e., f −1 ({b})) is called the fiber of f
over b. Note that f −1 is not in general a function and that the fibers of f generally contain many
elements since there may be many elements of A mapping to the element b.
Definition. f is surjective or is a surjection if for all b ∈ B there is some a ∈ A such that f (a) = b,
i.e., the image of f is all of B.
Note that since a function always maps onto its range (by definition) it is necessary to specify the
codomain B in order for the question of surjectivity to be meaningful.
Proposition. Let f : A → B.
(1) The map f is injective if and only if f has a left inverse.
(2) The map f is surjective if and only if f has a right inverse.
(3) The map f is a bijection if and only if there exists g : B → A such that f ◦ g is the identity map
on B and g ◦ f is the identity map on A.
(4) If A and B are finite sets with the same number of elements (i.e., |A| = |B|), then f : A → B
is bijective if and only if f is injective if and only if f is surjective.
Proof. Hints: 1.(⇒) For t ∈ im(f ), define `(t) to be the (unique) s ∈ f −1 ({t}), and define `(t)
arbitrary for t 6∈ im(f ).
2. (⇒) Define r(t) = any s ∈ f −1 ({t}).
3. (⇐) directly follows by parts (1) and (2). (Why? Explain.)
(⇒) By parts (1) and (2), there exists g : B → A and h : B → A such that g ◦ f = IA and f ◦ h = IB .
Next it is shown that g = h. Let b ∈ B then f (h(b)) = b whereby g(f (h(b))) = g(b) and hence
h(b) = g(b).
4. It suffices to prove that ”f is injective if and only if f is surjective.” The proof follows by the
definition of injectivity and surjectivity. (Why? Explain.)
Note that in the situation of part (3) of the proposition above the map g is necessarily unique
and we shall say g is the 2-sided inverse (or simply the inverse) of f .
Binary Relation
Let A be a nonempty set.
1. b ∈ a if and only if b ⊆ a.
2. b ∼ a if and only if b = a.
Proof: Hints: 2) Define relation on A as a ∼ b iff a, b ∈ Ai for some i. Complete the proof!!!
1 ) We have to prove two facts:
i. ∪a∈A [a] = A (why?).
ii. Let a, b ∈ A then either [a] = [b] or [a] ∩ [b] = ∅, (suppose [a] ∩ [b] 6= ∅, then ∃ c ∈ A such
that c ∼ a and c ∼ b. By symmetric property, a ∼ c. Now ∀ x ∈ [a], by transitive property,
x ∼ a, a ∼ c, c ∼ b ⇒ x ∼ b, that is, [a] ⊆ [b]. Similarly, [b] ⊆ [a]. Hence [a] = [b] )
Definition. The set of all equivalence classes of an equivalence relation ∼ on A is called the quotient
set of A by ∼
A/ ∼:= {B ⊆ A | B = a for some a ∈ A}
and the map A → A/ ∼, x → x is a surjection from A onto A/ ∼ (called the natural projection of
A onto A/ ∼).
Home Work
1. Complete the proofs of Propositions 1 and 2.
Exercises
In Exercises 1 to 4 let A be the set of 2 × 2 matrices with real number entries. Recall that matrix
multiplication is defined by
! ! !
a b p q ap + br aq + bs
=
c d r s cp + dr cq + ds
!
1 1
Let M = and let
0 1
B = {X ∈ A | M X = XM }
! ! ! ! ! !
1 1 1 1 0 0 1 1 1 0 0 1
, , , , ,
0 1 1 1 0 0 1 0 0 1 1 0
2. Prove that if P, Q ∈ B, then P + Q ∈ B (where + denotes the usual sum of two matrices).
3. Prove that if P, Q ∈ B, then P · Q ∈ B. (where . denotes the usual product of two matrices).
!
p q
4. Find conditions on p, q, r, s which determine precisely when ∈B
r s
6. Determine whether the function f : R+ → Z defined by mapping a real number r to the first
digit to the right of the decimal point in a decimal expansion of r is well defined.
3. Greatest common divisor If a, b ∈ Z − {0}, there is a unique positive integer d, called the
greatest common divisor of a and b (or g.c.d. of a and b), satisfying:
(a) d|a and d|b (so d is a common divisor of a and b), and
(b) if e|a and e|b, then e|d (so d is the greatest such divisor).
The g.c.d. of a and b will be denoted by (a, b). If (a, b) = 1, we say that a and b are relatively
prime.
4. Least common multiple If a, b ∈ Z − {0}, there is a unique positive integer l, called the
least common multiple of a and b (or l.c.m. of a and b), satisfying:
(a) a|l and b|l (so l is a common multiple of a and b), and
(b) if a|m and b|m, then l|m (so l is the least such multiple).
Fact: The connection between the greatest common divisor d and the least common multiple
l of two integers a and b is given by dl = ab.
5. The Division Algorithm If a, b ∈ Z − {0}, then there exist unique q, r ∈ Z such that
where q is the quotient and r the remainder. This is the usual ”long division” familiar from
elementary arithmetic.
a = q0 b + r0 , (0)
b = q1 r0 + r1 , (1)
r0 = q2 r1 + r2 , (2)
r1 = q3 r2 + r3 , (3)
..
.
rn−2 = qn rn−1 + rn , (n)
rn−1 = qn+1 rn , (n + 1)
7. Fact One consequence of the Euclidean Algorithm which we shall use regularly is the follow-
ing:
if a, b ∈ Z − {0}, then there exist x, y ∈ Z such that
(a, b) = ax + by
that is, the g.c.d. of a and b is a Z-linear combination of a and b. This follows by recursively
writing the element rn in the Euclidean Algorithm in terms of the previous remainders (namely,
use equation (n) above to solve for rn = rn−2 + (−qn )rn−1 in terms of the remainders rn−1 and
rn−2 , then use equation (n − 1) to write rn in terms of the remainders rn−2 and rn−3 , etc.,
eventually writing rn in terms of a and b).
17 = (302)57970 − (1691)10353
as can easily be checked directly. Hence the equation ax + by = (a, b) for the greatest common
divisor of a and b in this example has the solution x = 302 and y = −1691. Note that it is
relatively unlikely that this relation would have been found simply by guessing. The integers
x and y in (7) above are not unique. In the example with a = 57970 and b = 10353 we
determined one solution to be x = 302 and y = 1691, for instance, and it is relatively simple
to check that x = 307 and y = 1719 also satisfy 57970x + 10353y = 17.
8. Prime An element p of Z+ is called a prime if p > 1 and the only positive divisors of p are 1
and p (initially, the word prime will refer only to positive integers). An integer n > 1 which
is not prime is called composite. For example, 2, 3, 5, 7, 11, 13, 17, 19, . . . are primes and
4, 6, 8, 9, 10, 12, 14, 15, 16, 18, . . . are composite. An important property of primes (which
in fact can be used to define the primes is the following:
p is a prime if and only if p|ab, for some a, b ∈ Z, implies either p|a or p|b.
This factorization is unique in the sense that if q1 , q2 , . . . , qt are any distinct primes and
β1 , β2 , . . . , βt positive integers such that
then s = t and if we arrange the two sets of primes in increasing order, then qi = pi and
αi = βi , 1 ≤ i ≤ s. For example, n = 1852423848 = 23 32 112 193 31 and this decomposition into
the product of primes is unique.
where p1 , p2 , . . . , ps are distinct and the exponents are ≥ 0 (we allow the exponents to be 0
here so that the products are taken over the same set of primes - the exponent will be 0 if that
prime is not actually a divisor). Then the greatest common divisor of a and b is
(and the least common multiple is obtained by instead taking the maximum of the αi , and βi
instead of the minimum).
Example 3
In the example above, a = 57970 and b = 10353 can be factored as a = 2 · 5 · 11 · 17 · 31 and
b = 3 · 7 · 17 · 29, from which we can immediately conclude that their greatest common divisor
is 17. Note, however, that for large integers it is extremely difficult to determine their prime
factorizations (several common codes in current use are based on this difficulty, in fact), so
that this is not an effective method to determine greatest common divisors in general. The
Euclidean Algorithm will produce greatest common divisors quite rapidly without the need for
the prime factorization of a and b.
10. Euler ϕ-function The Euler ϕ-function is defined as follows: for n ∈ Z+ let ϕ(n) be the
number of positive integers a ≤ n with a relatively prime to n, i.e., (a, n) = 1. For example,
ϕ(12) = 4 since 1, 5, 7 and 11 are the only positive integers less than or equal to 12 which
have no factors in common with 12. Similarly, ϕ(1) = 1, ϕ(2) = 1, ϕ(3) = 2, ϕ(4) = 2, ϕ(5) =
4, ϕ(6) = 2, etc. For primes p, ϕ(p) = p − 1, and, more generally, for all a ≥ 1 we have the
formula
For example, ϕ(12) = ϕ(22 )ϕ(3) = 21 (2 − 1)30 (3 − 1) = 4. The reader should note that we
shall use the letter ϕ for many different functions throughout the text so when we want this
letter to denote Euler’s function we shall be careful to indicate this explicitly.
2. Compute greatest common divisor of 12 and 32 and find x, y ∈ Z, such that (12, 32) =
12x + 32y.
Solution (12,32)=4 and x = 3, y = −1
Exercise
1. For each of the following pairs of integers a and b, , determine their greatest common divisor,
their least common multiple, and write their greatest common divisor in the form ax + by for
some integers x and y.
(a) a = 20, b = 13.
(b) a = 792, b = 275.
(c) a = 1761, b = 1567.
2. Prove that if the integer k divides the integers a and b then k divides as + bt for every pair of
integers s and t.
4. Prove that if d divides n then ϕ(d) divides ϕ(n) where ϕ denotes Euler’s ϕ-function.
determined by the possible remainders after division by n. ( That is, if a ∈ Z is arbitrary then by
Division Algorithm, there exist unique q and r such that a = qn + r, where 0 ≤ r < n. This implies
a ∼ r and hence a = r ∈ {0, 1, 2, . . . , n − 1}, as required.)
Clearly, these residue classes partition the integers Z. The set of equivalence classes under this
equivalence relation will be denoted by Z/nZ and called the integers modulo n (or the integers mod
n).
Warning!! Note that for different n’s the equivalence relation and equivalence classes are different
so we shall always be careful to fix n first before using the bar notation.
• Example: Suppose n = 12 and consider Z/12Z, which consists of the twelve residue classes
¯
0̄, 1̄, 2̄, . . . , 11
determined by the twelve possible remainders of an integer after division by 12. The elements in
the residue class 5̄, for example, are the integers which leave a remainder of 5 when divided by 12
(the integers congruent to 5 mod 12). Any integer congruent to 5 mod 12 (such as 5, 17, 29, . . .
or 7, 19, . . . ) will serve as a representative for the residue class 5̄. Note that Z/12Z consists of the
twelve elements above (and each of these elements of Z/12Z consists of an infinite number of usual
integers).
• Terminology: The process of finding the equivalence class mod n of some integer a is often
referred to as reducing a mod n. This terminology also frequently refers to finding the smallest
nonnegative integer congruent to a mod n (the least residue of a mod n).
• Example: Suppose now that ā = 5̄ and b̄ = 8̄. The most obvious representative for ā is the
integer 5 and similarly 8 is the most obvious representative for b̄. Using these representatives for the
¯ = 1̄ since 13 and 1 lie in the same class modulo n = 12. Had we
residue classes we obtain 5̄ + 8̄ = 13
instead taken the representative 17, say, for ā (note that 5 and 17 do lie in the same residue class
modulo 12) and the representative −28, say, for b̄, we would obtain 5̄ + 8̄ = (17 − 28) = −11 = 1̄ and
as we mentioned the result does not depend on the choice of representatives chosen. The product
¯ = 4̄, also independent of the representatives chosen.
of these two classes is āb̄ = 5.8 = 40
Theorem: The operations of addition and multiplication on Z/nZ defined above are both well
defined, that is, they do not depend on the choices of representatives for the classes involved. More
precisely,
if a1 , a2 ∈ Z and b1 , b2 ∈ Z with a1 = b1 and a2 = b2 , then a1 + a2 = b1 + b2 and a1 a2 = b1 b2 , i.e., if
a1 ≡ b1 (mod n) and a2 ≡ b2 (mod n) then a1 + a2 ≡ b1 + b2 (mod n) and a1 a2 ≡ b1 b2 (mod n).
Proof: . Suppose a1 ≡ b1 (mod n), i.e., a1 − b1 is divisible by n. Then a1 = b1 + sn for some integer
s. Similarly, a2 ≡ b2 (mod n) means a2 = b2 + tn for some integer t.
. Then a1 + a2 = (b1 + b2 ) + (s + t)n so that a1 + a2 ≡ b1 + b2 (mod n), which shows that the sum
of the residue classes is independent of the representatives chosen.
. Similarly, a1 a2 = (b1 + sn)(b2 + tn) = b1 b2 + (bl t + b2 s + stn)n shows that a1 a2 ≡ b1 b2 (mod n) and
so the product of the residue classes is also independent of the representatives chosen, completing
the proof.
Proof : . We claim that the set {a ∈ Z/nZ | (a, n) = 1} is well defined i.e., if a = b and (a, n) = 1
then (b, n) = 1 (The proof follows by contradiction) Suppose a = b and (b, n) = d 6= 1 then b = a+kn,
d|b and d|n this implies a + kn = dl and n = dm for some l, m ∈ Z, this implies a = dl − kn and
n = dm, this implies d|a and d|n, this implies (a, n) = d, a contradiction to the hypothesis.
. ⊆ proof follows from Exc. 8. (Proof Exc. 8: if (a, n) = d 6= 1 then d|a and d|n this implies a = kd
and n = ld for some k, l ∈ Z. Now, if we choose b = l, then 1 ≤ b < n and ab = (kd)( nd ) = kn ≡ 0
(mod n), the desired. Next, we show that there cannot be an integer c such that ac = 1. Suppose,
on contrary that, there exist c such that ac = 1, i.e., 1 = qn + ac. Multiplication by b gives
b=l n=ld
b = bqn + bac = bqn + b(kd)c = bqn + (bd)kc = lqn + (ld)kc = lqn + nkc, a contradiction because
1 ≤ b < n. ).
. ⊇ proof follows from Exc. 9. (Proof Exc. 9 : if (a, n) = 1 then there exist integers c and f such
that ac + nf = 1 this implies ac ≡ 1 (mod n), the desired. )3
Example:
For n = 9 we obtain (Z/9Z)× = {1̄, 2̄, 4̄, 5̄, 7̄, 8̄} from the proposition. The multiplicative inverses of
these elements are {1̄, 5̄, 7̄, 2̄, 4̄, 8̄}, respectively.
• Method of finding multiplicative inverses in Z/nZ:
If a is an integer relatively prime to n then the Euclidean Algorithm produces integers x and y
satisfying ax + ny = 1, hence ax ≡ 1 (mod n), so that x is the multiplicative inverse of a in Z/nZ.
This gives an efficient method for computing multiplicative inverses in Z/nZ.
Example:
Suppose n = 60 and a = 17. Applying the Euclidean Algorithm we obtain
60 = (3)17 + 9
17 = (1)9 + 8
9 = (1)8 + 1
so that a and n are relatively prime, and (−7)I7+(2)60 = 1. Hence −7 = 53 is the multiplicative
inverse of 17 in Z/60Z.
Class Activity
1. Write down all the elements in (Z/12Z)× .
¯
Sol: (Z/12Z)× = {1̄, 5̄, 7̄, 11}
4. 7 + 9 = in Z/12Z ?
Sol: 4
5. 7 · 9 = in Z/12Z ?
Sol: 3?
Home Work
Do the proofs of the proposition and Theorem.
Do the topic ”application of modular arithmetic in number theory” given in this lecture.
Exercise
0.3.1, 0.3.2, 0.3.3, 0.3.4, 0.3.5, 0.3.6, 0.3.10, 0.3.11, 0.3.15
1. Write down explicitly all the elements in the residue classes of Z/18Z.
2. Prove that the distinct equivalence classes in Z/nZ are precisely 0̄, 1̄, 2̄, . . . , n − 1 (use the
Division Algorithm).
6. Prove that the number of elements of (Z/nZ)× is ϕ(n) where ϕ denotes the Euler ϕ- function.
8. Let n ∈ Z, n > 1, and let a ∈ Z with 1 ≤ a ≤ n. Prove if a and n are not relatively prime,
there exists an integer b with 1 ≤ b < n such that ab ≡ 0 (mod n) and deduce that there
cannot be an integer c such that ac ≡ 1 (mod n).
10. For each of the following pairs of integers a and n, show that a is relatively prime to n and
determine the multiplicative inverse of a in Z/nZ.
(a) a = 13, n = 20.
(b) a = 69, n = 89.
(c) a = 1891, n = 3797.
(d) a = 6003722857, n = 77695236973. [The Euclidean Algorithm requires only 3 steps for
these integers.]
1.4 Groups
Examples
(1) + (usual addition) is a binary operation on Z (or on Q, R, or C respectively).
(2) × (usual multiplication) is a binary operation on Z (or on Q, R, or C respectively).
(3) − (usual subtraction) is a binary operation on Z, where −(a, b) = a − b. The map a 7→ −a is
not a binary operation (not binary).
(4) − is not a binary operation on Z+ (nor Q+ , R+ ) because for a, b ∈ Z+ with a < b, a − b ∈
/ Z+ ,
that is, − does not map Z+ × Z+ into Z+ .
Definition. A group is an ordered pair (G, ?) where G is a set and ? is a binary operation on G
satisfying the following axioms:
(i) (a ? b) ? c = a ? (b ? c), ∀ a, b, c ∈ G, i.e., ? is associative,
(ii) ∃ e ∈ G, called an identity of G, such that ∀ a ∈ G we have a ? e = e ? a = a,
(iii) ∀ a ∈ G ∃ x ∈ G, called an inverse of a, such that a ? x = x ? a = e.
Examples:
(1) Z, Q, R and C are groups under + with e = 0 and inverse of a is, −a, for all a.
(2) Q − {0}, R − {0}, C − {0}, Q+ , R+ are groups under × with e = 1 and inverse of a is, a1 , for all
a. Note however that Z − {0} is not a group under × because although × is an associative binary
operation on Z − {0}, the element 2 (for instance) does not have an inverse in Z − {0}.
Glossing Over: We have glossed over the fact that the associative law holds in these familiar
examples. For Z under + this is a consequence of the axiom of associativity for addition of nat-
ural numbers. The associative law for Q under + follows from the associative law for Z a proof
The associative axiom for multiplication may be established via a similar development, starting
first with Z. Since R and C will be used largely for illustrative purposes and we shall not construct
R from Q we shall take the associative laws (under + and × ) for R and C as given.
Examples (continued):
(3) The axioms for a vector space V include those axioms which specify that (V, +) is an abelian
group (the operation + is called vector addition). Thus any vector space such as Rn is, in particular,
an additive group.
(4) For n ∈ Z+ , Z/nZ is an abelian group under the operation + of addition of residue classes as
described in Chapter 0. We shall prove in Chapter 3 (in a more general context) that this binary
operation + is well defined and associative; for now we take this for granted. The identity in this
group is the element 6 and for each ā ∈ Z/nZ, the inverse of ā is −a. Henceforth, when we talk
about the group Z/nZ it will be understood that the group operation is addition of classes mod n.
(5) For n ∈ Z+ , the set (Z/nZ)× of equivalence classes ā which have multiplicative inverses mod n is
an abelian group under multiplication of residue classes as described in Section 1.3. Again, we shall
take for granted (for the moment) that this operation is well defined and associative. The identity of
this group is the element 1̄ and, by definition of (Z/nZ)× , each element has a multiplicative inverse.
Henceforth, when we talk about the group ((Z/nZ)× it will be understood that the group operation
is multiplication of classes mod n.
(6) If (A, ?) and (B, ) are groups, we can form a new group A × B, called their direct product,
whose elements are those in the Cartesian product
A × B = {(a, b) | a ∈ A, b ∈ B}
For example, if we take A = B = R (both operations addition), R×R is the familiar Euclidean plane.
The proof that the direct product of two groups is again a group is left as a straightforward exercise
(later) the proof that each group axiom holds in A × B is a consequence of that axiom holding in
2
D. S. Dummit & R. M. Foote, Abstract Algebra, 3rd Edition, John Wiley & Sons, Inc.
Definition. The group (G, ?) is called abelian (or commutative) if a ? b = b ? a for all a, b ∈ G.
Examples
(1) The groups given in examples 1 and 2 are abelian or commutative groups.
(2) The symmetric group S3 3 is non-abelian or non-commutative group.
Proof
(1) Let f and g be both identities, by axiom (ii) of the definition of a group
f g = g (take a = g and e = f ).
a(bc) = e.
The associative law on the left hand side and the definition of e on the right give
so
e(bc) = a−1
hence
bc = a−1 .
Now multiply both sides on the left by b−1 and simplify similarly:
b−1 (bc) = b−1 a−1 , (b−1 b)c = b−1 a−1 , ec = b−1 a−1 , c = b−1 a−1 ,
as claimed.
Notation:
(3) For an abstract group G (operation ·) we denote the identity of G by 1.
(4) For any group G (operation . implied) and x ∈ G and n ∈ Z+ since the product xx...x (n terms)
does not depend on how it is bracketed, we shall denote it by xn . Denote x−1 x−1 ...x−1 (n terms)
by x−n Let x0 = 1, the identity of G.
Proposition. 2. Let G be a group and let a, b ∈ G. The equations ax = b and ya = b have unique
solutions for x, y ∈ G. In particular, the left and right cancellation laws hold in G, i.e.,
(1) if au = av, then u = v, and
(2) if ub = vb, then u = v.
Proof:
x = a−1 b.
u = v.
Consequences of Proposition 2:
(1) if a is any element of G and for some x ∈ G, ax = e or xa = e, then x = a−1 , i.e., we do not
have to show both equations hold.
(2) Also, if for some b ∈ G, ab = a (or ba = a), then b must be the identity of G, i.e., we do not
have to check bx = xb = x for all x ∈ G.
Definition. For G a group and x ∈ G, define the order of x to be the smallest positive integer n
such that xn = 1, and denote this integer by |x|. In this case x is said to be of order n. If no positive
power of x is the identity, the order of x is defined to be infinity and x is said to be of infinite order.
Or, in other words, for any element a ∈ G,
n, if n ∈ Z+ is the smallest such that xn = e
|a| =
∞, otherwise.
Examples:
(1) An element of a group has order 1 if and only if it is the identity.
(2) In the additive groups Z, Q, R or C every nonzero (i.e., nonidentity) element has infinite order.
(3)In the multiplicative groups R − {0} or Q − {0} the element −1 has order 2 and all other
nonidentity elements have infinite order.
(4) In the additive group Z/9Z the element 6̄ has order 3, since 6̄ 6= 0̄, 6̄ + 6̄ = 12 = 3̄ 6= 0̄, but
6 + 6 + 6 = 18 = 6, the identity in this group. Recall that in an additive group the powers of an
element are the integer multiples of the element. Similarly, the order of the element 5̄ is 9, since 45
is the smallest positive multiple of 5 that is divisible by 9.
(5) In the multiplicative group (Z/7Z)× , the powers of the element 2̄ are 2̄, 4̄, 8̄ = 1̄, the identity in
this group, so 2̄ has order 3. Similarly, the element 3̄ has order 6, since 36 is the smallest positive
power of 3 that is congruent to 1 modulo 7.
2. In the group Z with the operation of addition, what is the inverse of the element 5?
Solution: We have 5 + (−5) = (−5) + 5 = 0. Therefore, by definition of the inverse element
in the group, the inverse of 5 in (Z,+) is −5.
3. Consider the set M2,2 (R) of all 2 × 2 matrices with entries from R, endowed with the standard
operation of matrix multiplication. Is M2,2 (R) a group?
Solution: No, this is not a group. Since for any 3 × 3 matrices A, B, C ∈ M2,2 (R) we have
(AB)C = A(BC), the # axiom of being a group is satisfied for (M2,2 (R), · ).
" first
1 0
Moreover, for E = , we have
0 1
Thus the second axiom of being a group is satisfied for (M2,2 (R), ·).
However, not" every# element in (M2,2 (R), ·) has a multiplicative inverse. For example, for the
0 0
matrix O = there does not exist a matrix B ∈ (M2,2 (R), ·) such that OB = E, since
0 0
we always have OB = O. Thus the third axiom of being a group fails here.
4. Consider the set (R, +) as a group with respect to addition. Compute the 3rd power of 5 ∈ R,
with respect to addition. What is the order of the element 5 in (R, +)?
Solution: The 3rd power of 5 in (R, +) is 5 + 5 + 5 = 15. The order of 5 in (R, +) is ∞
since for every integer n ≥ 1, 5n 6= 0.
5. In the group Z/7Z with the operation of addition, what is the order of the element 2?
Solution: The order of 2 in (Z/7Z, +) is 7. Indeed 7 · 2 = 0 in Z/7Z and a direct check shows
that for every n = 1, 2, . . . , 6, 2n 6= 0 in Z/7Z.
Exercises
Page 20 Compiled on 2021/04/07 at 22:54:52
1. Determine which of the following binary operations are associative:
(a) the operation ∗ on R defined by a ∗ b = a + b + ab
(b) the operation ∗ on Q defined by a ∗ b = (a + b)/5
(c) the operation ∗ on Z × Z defined by (a, b) ∗ (c, d) = (ad + bc, bd)
2. Decide which of the binary operations in the preceding exercise are commutative.
3. Prove for all n > 1 that Z/nZ is not a group under multiplication of residue classes.
5. (a) Find the orders of the following elements of the additive group Z/36Z:
1, 2, 6, 9, 10, 12, −1, −10, −18.
(b) Find the orders of the following elements of the multiplicative group (Z/36Z)× :
1, −1, 5, 13, −13, 17.
6. For x an element in G show that x and x−1 have the same order.
Definition. Let G be a group. The cardinality of G, i.e., |G|, is called the order of the group G.
Eamples: The order of the group Z/nZ is n that is, |Z/nZ| = n . The order of the group Z is ∞,
that is, |Z| = ∞.
Dihedral groups
An important family of examples of groups is the class of groups whose elements are symmetries of
geometric objects. The simplest subclass is when the geometric objects are regular planar figures.
For each n ∈ Z+ , n ≥ 3
where a symmetry4 is any rigid motion of the n-gon which can be effected by taking a copy of the
n-gon, moving this copy in any fashion in 3-space and then placing the copy back on the original
n-gon so it exactly covers it. More precisely, we can describe the symmetries by first choosing a
labelling of the n vertices, for example as shown in the following figure.
Figure 1:
Now make D2n into a group by defining st for s, t ∈ D2n to be the symmetry obtained by first
applying t then s to the n-gon (note that we are viewing symmetries as functions on the n-gon, so
st is just function composition - read as usual from right to left). The binary operation on D2n is
associative since composition of functions is associative. The identity of D2n is the identity symmetry
4
A symmetry of P is a rigid motion (i.e. distance preserving function f : C → C) which carries P onto itself (i.e.
f (P ) = P )
Figure 2:
the lines x = 0 (y-axis), y = 0 (x-axis), y = x and y = −x (note that ”reflection” through the
origin is not a reflection but a rotation of n radians).
We now fix some notation and mention some calculations for future use. Fix a regular n-gon centered
at the origin in an x, y plane and label the vertices consecutively from 1 to n in a clockwise manner.
Let r be the rotation clockwise about the origin through 2π/n radian. Let s be the reflection about
the line of symmetry through vertex 1 and the origin. Observe the following calculations:
i.e., each element can be written uniquely in the form sk ri for some k = 0 or 1 and 0 ≤ i ≤ n − 1.
(5) rs = sr−1 . This shows in particular that r and s do not commute so that every element in D2n
does not have the abelian/commutative property. Hence D2n is non-abelian group.
5
See proof on the page 24 of the Dummit & Footie Book or see the link http://www.math.uconn.edu/ kcon-
rad/blurbs/grouptheory/dihedral.pdf for more rigorous proof.
Having done these calculations, we now observe that the complete multiplication table of D2n
can be written in terms r and s alone, that is, all the elements of D2n have a (unique) representation
in the form sk ri , k = 0 or 1 and 0 ≤ i ≤ n − 1, and any product of two elements in this form can be
reduced to another in the same form using only ”relations” (1), (2) and (6) (reducing all exponents
mod n). For example, if n = 12, (sr9 )(sr6 ) = s(r9 s)r6 = s(sr−9 )r6 = s2 r−9+6 = r−3 = r9 .
Klein 4-Group
Definition. The Klein 4-group, V4 , is defined by
V4 = {1, a, b, c}
1 a b c
1 1 a b c
a a 1 c b
b b c 1 a
c c b a 1
1 -1 i -i j -j k -k
1 1 -1 i -i j -j k -k
-1 -1 1 -i i -j j -k k
-i -i i 1 -1 -k k j -j
i i -i -1 1 k -k -j j
j j -j -k k -1 1 i -i
-j -j j k -k 1 -1 -i i
k k -k j -j -i i -1 1
-k -k k -j j i -i 1 -1
Matrix Groups
Definition. (1) A field is a set F together with two binary operations + and · on F such that (F, +)
is an abelian group (call its identity 0) and (F − {0}, ·) is also an abelian group, and the following
distributive law holds:
a · (b + c) = (a · b) + (a · c), for all a, b, c ∈ F.
GLn (F ) is a group (Exc.) under matrix multiplication, called the general linear group of degree n.
Definition. A subset S of elements of a group G with the property that every element of G can be
written as a (finite) product of elements of S and their inverses is called a set of generators of G.
We shall indicate this notationally by writing G = hSi and say G is generated by S or S generates
G.
V4 = {a, b | a2 = b2 = (ab)2 = 1}
4. Consider the group (Z2 , +) with respect to addition. Give an example of a finite subset S⊆ Z2
such that S generates Z2 .
Solution: For example S = {(1, 0), (0, 1)} generates Z2 .
5. Consider a group given by the presentation G = {a, b | ab = ba}. Is it true that this group is
abelian?
Solution: Yes, it is true that this group is abelian.
Exercises
1. Compute the order of each of the elements in Q8 .
Symmetric/Permutation Group
Set n = {1, 2 . . . , n} Then
Sn := {all bijections from n → n}
is a group under composition of functions. (Exc.) Its elements are called permutations (of n symbols
or letters), and (Sn , ◦) is called the symmetric group of degree n. Note that Sn is a finite, nonabelian
(for n ≥ 3) group of order n!,
|Sn | = n! (why?)
More generally, for any set A (possibly infinite), the set SA of bijections A → A is a group(Exc.)
under composition of functions, called the symmetric group on A.
is the permutation which sends each ij to the next ij+1 in the list, sends ik to i1 , and leaves all
else fixed (draw circular picture). Note that ”cyclic permutations” of the list, e.g. (i2 i3 . . . ik i1 ),
represent the same cycle. A 2-cycle is also called a transposition.
Example:
(1 3 4)(2 5 7) = (2 5 7)(1 3 4)
Fact: Every σ ∈ Sn can be written uniquely (up to rearranging its cycles and cyclically permuting
the numbers within each cycle) as a product of disjoint (i.e. non overlapping) cycles, called its cycle
decomposition. (Often omit 1-cycles in notation)
Examples:
!
1 2 3 4 5
(1) = (1 4)(2 5 3)
4 5 2 1 3
!
1 2 3 4 5
(2) = (1 4)(2)(3 5) = (1 4)( 35) = (3 5)(4 1)
4 2 5 1 3
Multiplication of permutations:
To multiply (i.e. compose) two permutations, first juxtapose their cycle decompositions. What
results is a product of cycles that might not be disjoint. To rewrite this in disjoint cycle form, work
from right to left (as with composition of functions) to see where each number 1, . . . , n maps.
Example:
For example, to compute π = στ where σ = (2 1 4 5 3) and τ = (1 5)(2 3), first see where 1 maps:
we have τ (1) = 5 and σ(5) = 3, and so π(1) = 3. Similarly π(3) = 1 (giving (1 3) as one of the
cycles in π), π(2) = 2 (giving the cycle (2)), π(4) = 5 and π(5) = 4 (giving the cycle (4 5)). Thus
Inverse of a permutation:
For any σ ∈ Sn , the cycle decomposition of σ −1 is obtained by writing the numbers in each cycle of
the cycle decomposition of σ in reverse order.
Order of a permutation:
Note The order (as an element of Sn ) of any k-cycle is k, and in general the order of a permutation
is the least common multiple (lcm) of the orders of the (disjoint) cycles in its cycle decomposition
(why?).
Example:
(not 10 for the latter, since the cycles are not disjoint).
Exercises
Page 29 Compiled on 2021/04/07 at 22:54:52
1. Let σ and τ be the permutations given by: !
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
τ=
13 2 15 14 10 6 12 3 4 1 7 9 5 11 8
!
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
σ=
14 9 10 2 12 6 5 11 15 3 8 7 4 1 13
Find the cycle decompositions of the following permutations: σ, τ, σ 2 , στ, τ σ and τ 2 σ.
2. Compute the order of each of the elements in the following groups: (a) S3 (b) S4 .
7. Prove that the order of an element in Sn is the least common multiple of the lengths of the
cycles in its cycle decomposition.
is called a homomorphism.
Intuitively, a map ϕ is a homomorphism if it respects the group structures of its domain and
codomain.
A homomorphism from a group to itself (i.e. G = H) is called an endomorphism.
The image of ϕ is:
im(ϕ) := {ϕ(x) | x ∈ G} = ϕ(G)
Definition. The map ϕ : G → H is called an isomorphism (and G and H are said to be isomorphic
or of the same isomorphism type), written G ∼
= H, if
(1) ϕ is a homomorphism (i.e., (ϕ(xy) = ϕ(x)ϕ(y)), and
(2) ϕ is a bijection.
Examples:
1. exp : (R, +) → (R − {0}, ·) is an (injective) homomorphism.
2. det : GLn (R) → R − {0} is a surjective homomorphism (where it is understood that the operation
is multiplication in both groups).
3(a). For any abelian group G, the map ψ : G → G; x 7→ x−1 is a homomorphism since ψ(xy) =
(xy)−1 = y −1 x−1 = x−1 y −1 = ψ(x)ψ(y). In fact ψ is an automorphism which is its own inverse!
3(b). If G is nonabelian, then ψ is not a homomorphism: for any a, b ∈ G with ab 6= ba, have
ψ(ab) = (ab)−1 and ψ(a)ψ(b) = a−1 b−1 = (ba)−1 , but (ab)−1 6= (ba)−1 since inverses are unique.
1. f (1) = 1 .
2. f (x−1 ) = (f (x))−1 .
Proof: Do yourself.
2
Proof: If k is finite, then for any i and j we can write i − j = qk + r with 0 ≤ r < k. Thus
ai = aj ⇐⇒ ai−j = 1 ⇐⇒ ar = 1, since aqk+r = (ak )q ar = ar . But ar = 1 ⇐⇒ r = 0, since
r < k = |a|, and r = 0 ⇐⇒ i ≡ j (mod k). The last statement is clear since ai = aj ⇐⇒ ai−j = 1.
2 Exc. If ϕ : G → H is an isomorphism of groups, then show that ϕ−1 : H → G is isomorphism.
Proof: do yourself.
We can use Lemma 2 to prove, two groups are not isomorphic.
We Can also count the number of elements of a given order to distinguish groups.
Examples:
1 D24 S4 , since D24 has 12 elements of order 2 while S4 has 9 only (why?).
1 (R − {0}, ×) (R, +), since in (R − {0}, ×) the element −1 has order 2 whereas (R, +) has no
element of order 2
Facts:
1 There is only one group of any given prime order
2 There exist 2 of order 4 (Z/4Z, V4 ), 2 of order 6 (Z/6Z, S3 ), 5 of order 8 (Z/8Z, Z/4Z ×
Z/2Z, Z/2Z × Z/2Z × Z/2Z, D8 , Q8 ), 2 of order 9 ( Z/9Z, Z/3Z × Z/2Z), 2 of order 10 (Z/10Z, D10 ),
5 of order 12, 14 of order 16 . . . (see page 168 in text)
Homomorphisms and isomorphisms between two groups given by generators and relations:
Facts:
1 Let G be a finite group of order n for which we have a presentation and let S = {s1 , . . . , sm } be
the generators. Let H be another group and r1 , . . . , rm be elements of H. Suppose that any relation
satisfied in G by the si is also satisfied in H when each si is replaced by ri . Then there is a (unique)
homomorphism ϕ : G → H which maps si to ri .
2 If H is generated by the elements r1 , . . . , rm , then ϕ is surjective.
3 If, in addition, H has the same (finite) order as G, then any surjective map is necessarily injective,
i.e., ϕ is an isomorphism: G ∼ = H.
Intuitively, we can map the generators of G to any elements of H and obtain a homomorphism
provided that the relations in G are still satisfied.
Examples:
1 Recall that D2n = {r, s | rn = s2 = 1, sr = rs−1 }. Suppose H is a group containing elements
a and b with an = 1, b2 = 1 and ba = a−1 b. Then there is a homomorphism from D2n to H
mapping r to a and s to b. For instance, let k be an integer dividing n with k ≥ 3 and let
D2k = {r1 , s1 | r1k = s21 = 1, s1 r1 = r1−1 s1 ). Define
2. Suppose that f : G → H is a group homomorphism such that f (G) = H and such that G is
abelian. Does this imply that H is abelian?
Solution: Yes, this does imply that H is abelian. Indeed, let h1 , h2 ∈ H be arbitrary. Then,
since f (G) = H, there exist g1 , g2 ∈ G such that f (g1 ) = h1 and f (g2 ) = h2 . Since f is a
homomorphism, we have f (g1 g2 ) = f (g1 )f (g2 ) = h1 h2 and f (g2 g1 ) = f (g2 )f (g1 ) = h2 h1 . Since
G is abelian, g1 g2 = g2 g1 and hence h1 h2 = f (g1 g2 ) = f (g2 g1 ) = h2 h1 , so that H is abelian.
3. Give an example of two finite groups G and H such that |G| = |H| but G is not isomorphic to
H.
Solution: For example, take G = S3 and H = Z/6Z. We have |G| = |H| = 6. However, H is
abelian but G is non-abelian, and hence G and H are not isomorphic.
Exercises
1. Let G and H be groups. Let ϕ : G → H be a homomorphism.
(a) Prove that ϕ(xn ) = ϕ(x)n for all n ∈ Z+ .
(b) Do part (a) for n = −1 and deduce that ϕ(xn ) = ϕ(x)n for all n ∈ Z.
Subgroups
Definition: A subset H of a group G is a subgroup of G, written H ≤ G,
if it is a group under the operation induced from G, i.e.
(S1) x, y ∈ H =⇒ xy ∈ H (S2) 1 ∈ H (S3) x ∈ H =⇒ x−1 ∈ H.
(Note: associativity is automatic)
Examples:
1 Every group has the trivial subgroups 1 and G. Any other subgroup will be called proper.
2 For k = 0, 1, 2, . . . , the set kZ = {nk | n ∈ Z} of all multiples of k is a subgroup of Z; there are
no others. (Note that + is the operation in Z, so (S2) reads 0 ∈ kZ.) 0Z = {0} and 1Z = Z are
trivial, 2Z = evens, etc.
3 {1, r, . . . , rn−1 } and {1, s} are subgroups of D2n .
4 {1, −1} are {1, i, −1, −i} and are subgroups of Q8 .
Subgroup Criterion:
A subset H of a group G is a subgroup if and only if
a H is nonempty, and b x, y ∈ H =⇒ xy −1 ∈ H.
Proof:( =⇒ ) 1 ∈ H by (S2) so H 6= ∅. If x, y ∈ H, then y −1 ∈ H by (S3) so xy −1 ∈ H by (S1).
Thus b holds.
(⇐=) ∃ x0 ∈ H by a , so 1 = x0 .x−1
0 ∈ H by b =⇒ (S2). Now x ∈ H =⇒ x
−1
= 1.x−1 ∈ H by
b =⇒ (S3). Finally x, y ∈ H =⇒ y −1 ∈ H by (S3) =⇒ xy = x(y −1 )−1 ∈ H by b , so (S1)
holds.
Corollary: Let H be a finite subset of a group G.
NG (A) := {x ∈ G | xA = Ax}
Example: Z(Q8 ) = {1, −1} (verify). For A = {1 − 1, i, −i} ≤ Q8 , have ji 6= ij, ki 6= ik, etc. and
so CQ8 (A) = A. However jA = {j, −j, −k, k} is the same set as Aj = {j, −j, k, −k}, . . . and so
NQ8 (A) = Q8 .
Homework
1. Let f : G → H be a homomorphism of groups. Prove that
2. Let G be a group. Let A ⊆ G. Prove that Z(G), CG (A) and NG (A) are subgroups of G.
Exercises
Do the following selected exercises from Book.
Exercise 2.1: Q.1,2,3,4,6,8,9,10,11,12,14,15,16,17.
Exercise 2.2: Q. 1,2,3,4,5,6,11.
Cyclic Groups
Definition: For any element x in a group G, denote by hxi the set of all powers of x (or multiples of
x if G is additive),
hxi = {xk | k ∈ Z}.
If G = hxi for some x ∈ G we say G is a cyclic group, and any such x is called a generator of G
(there may be many, for example, hxi = hx−1 i).
Examples:
as desired.
Case II: Next suppose |x| = ∞ (i.e., no positive power of x is the identity) so for integers a 6= b,
xa 6= xb (because if xa = xb , for some a and b with, say, a < b, then xb−a = 1, a contradiction).
Hence distinct powers of x are distinct elements of hxi so |hxi| = ∞.
Proof. (a) By the Euclidean Algorithm there exist integers r and s such that d = mr + ns, where d
is the g.c.d. of m and n. Thus
(b) If xm = 1, let n = |x|. If m = 0, certainly n|m, so we may assume m 6= 0. Since some nonzero
power of x is the identity, n < ∞. Let d = (m, n) so by the preceding result xd = 1. Since 0 < d ≤ n
and n is the smallest positive power of x which gives the identity, we must have d = n, that is, n|m,
as asserted.
1. Z/5Z
2. Z/6Z
3. Z/6Z
4. Z × Z
Solution:
Exercises
Do exercises: 2.3.1, 2.3.2, 2.3.3, 2.3.4, 2.3.5, 2.3.6
ϕ : hxi −→ hyi
xk 7−→ y k
ψ : Z −→ hxi
k 7−→ xk
Proof: (1) Suppose hxi and hyi are both cyclic groups of order n.
Let ϕ : hxi −→ hyi be defined by ϕ(xk ) = y k ;
we must first prove ϕ is well defined, that is, if xr = xs , then ϕ(xr ) = ϕ(y s ).
Since xr−s = 1, Proposition (in section 1.9) implies n | r − s.
Write r = tn + s so ϕ(xr ) = ϕ(xtn+s ) = y tn+s = (y n )t y s = y s = ϕ(xs ). This proves ϕ is well defined.
It is immediate from the laws of exponents that ϕ(xa xb ) = ϕ(xa )(ϕ(xb ) (check this), that is, ϕ is a
homomorphism.
Since the element y k of hyi is the image of xk under ϕ, this map is surjective.
Since both groups have the same finite order, any surjection from one to the other is a bijection,
so ϕ is an isomorphism (alternatively, ϕ has an obvious two-sided inverse).
Notation: For each n ∈ Z+ , let Zn be the cyclic group of order n (written multiplicatively).
Up to isomorphism, Zn is the unique cyclic group of order n and Zn ∼ = Z/nZ. On occasion when
we find additive notation advantageous we shall use the latter group as our representative of the
isomorphism class of cyclic groups of order n.
Note: We shall occasionally say ”let hxi be the infinite cyclic group” (written multiplicatively),
however we shall always use Z (additively) to represent the infinite cyclic group.
As noted earlier, a given cyclic group may have more than one generator. The next two propositions
determine precisely which powers of x generate the group hxi.
1 = (xa )m = xam .
Also,
x−am = (xam )−1 = 1−1 = 1.
Now one of am or −am is positive (since neither a nor m is 0) so some positive power of x is
the identity. This contradicts the hypothesis |x| = ∞, so the assumption |xa | < ∞ must be
false, that is, ( 1) holds.
(2) Let d = (n, a) be the g.c.d of a and b. Then ∃ b, c ∈ Z with b > 0 such that n = db, a = dc.
This implies (b, c) = 1 (since d is gcd of a and n). We must show that
n n
|xa | = = = b.
(a, n) d
so by Proposition 3 applied to x, n|ak, i.e., db|dck. Thus b|ck. Since (b, c) = 1, b must divide
k. Thus b | |xa | and hence |xa | ≥ b. Since|xa | ≤ b and |xa | ≥ b, |xa | = b, which proves (2).
if x = 2̄ ∈ (Z/9Z)×
Home Work
Exercises
Do Exercises: 2.3.7, 2.3.8, 2.3.9, 2.3.10, 2.3.11, 2.3.12, 2.3.13, 2.3.14, 2.3.15
Proposition 6.
(1) Assume |x| = ∞. Then hxi = hxa i if and only if a = ±1.
(2) Assume |x| = n < ∞. Then hxi = hxa i if and only if (a, n) = 1.
In particular, the number of generators of hxi is ϕ(n) (where ϕ is Euler’s ϕ -function).
1. Every subgroup of cyclic group is cyclic. More precisely, if K ≤ hxi then either K = {1} or
K = hxd i, where d is the smallest positive integer such that xd ∈ K.
2. Assume |hxi| = n < ∞. If a is any positive integer dividing n, then there is a unique subgroup
n
hx a i of hxi of order a. Furthermore, for every integer m, hxm i = hx(m,n) i, 6 so that the
subgroups of hxi correspond bijectively with the positive divisors of n.
3. If |hxi| = ∞, then for any distinct nonnegative integers a and b, hxa i = 6 hxb i. Furthermore,
for every integer m, hxm i = hx|m| i, where |m| denotes the absolute value of m, so that the
nontrivial subgroups of hxi correspond bijectively with the integers 1, 2, 3, . . .
Proof: 1. Let K ≤ hxi. If K = {1}, the statement is true for this subgroup, so we assume
K 6= {1}. Thus there exists some a 6= 0 such that xa ∈ K. If a < 0 then since K is a group, also
x−a = (xa )−1 ∈ K. Hence K always contains some positive power of x. Let
P = {b | b ∈ Z+ and xb ∈ K}.
By the above, P is a nonempty set of positive integers. By the Well Ordering Principle P has a
minimum element–call it d. Since K is a subgroup and xd ∈ K, hxd i ≤ K. Since K is a subgroup of
hxi, any element of K is of the form xa for some integer a. By the Division Algorithm write
a = qd + r, 0 ≤ r < d.
6
(This implies if hxi has a subgroup of order k, then k|n. So that the subgroups of hxi correspond bijectively with
the positive divisors of n.)
n n
= a = |K| = |xb | = ,
d (n, b)
so d = (n, b). In particular, d|b. Since b is a multiple of d, xb ∈ hxd i, hence K = hxb i ≤ hxd i.
Since|hxd i| = a = |K|, we have K = hxd i.
The final assertion of 2. follows from the observation that hxm i is a subgroup of hx(n,m) i (check this!)
and, it follows from Proposition 5(2) and Proposition 2 that they have the same order. Since (n, m)
is certainly a divisor of n, this shows that every subgroup of H arises from a divisor of n, completing
the proof.
3. Do by yourself.
Example:
We can use Proposition 6 and Theorem 7 to list all the subgroups of Z/nZ for any given n. For
example, the subgroups of Z/12Z are
e. h6i (order 2)
Home Work
Exercises
1. Do exercise 2.3: Q.14, 18, 19, 20, 21, 22, 23, 24
Exercises
1. Do exercise 2.4: Q.5, 13, 14.
Exercises
1. Do exercise 2.5: Q.9, 10.
called respectively a left coset and a right coset of N in G. Any element of a coset is called a
representative for the coset.
Example 1: Let G = D8 = {1, r, r2 , r3 , s, sr, sr2 , sr3 }, and N = {1, r2 }, then all the left cosets of
N in G are:
1N = {1, r2 } = N ,
rN = {1.r, 1.r3 } = {r, r3 },
r2 N = {r2 .1, r2 .r2 } = {r2 , r4 } = {r2 , 1} = {1, r2 } = N ,
r3 N = {r3 .1, r3 .r2 } = {r3 , r5 } = {r3 , r} = {r, r3 } = rN ,
sN = {s.1, s.r2 } = {s, sr2 },
srN = {sr.1, sr.r2 } = {sr, sr3 },
sr2 N = {sr2 .1, sr2 .r2 } = {sr2 , sr4 } = {sr2 , s} = {s, sr2 } = sN ,
sr3 N = {sr3 .1, sr3 .r2 } = {sr3 , sr5 } = {sr3 , sr} = {sr, sr3 } = srN ,
It is clear that there are only four distinct left cosets of N in D8 so the set of left cosets of N in D8
is
{N, rN, sN, srN }
{N, N r, N s, N sr}
Lemma (Coset Decomposition): Let N be any subgroup of the group G. The set of left cosets
of N in G form a partition of G.
Furthermore, for all u, v ∈ G, uN = vN if and only if v −1 u ∈ N and in particular, uN = vN if and
only if u and v are representatives of the same coset (i.e., v ∈ uN ).
Proof. First of all note that since N is a subgroup of G, 1 ∈ N . Thus g = g.1 ∈ gN for all g ∈ G,
S S
i.e., G ⊆ g∈G gN. Furthermore, since gN ⊆ G, for all g ∈ G, Thus g∈G gN ⊆ G.
To show that distinct left cosets have empty intersection, suppose uN ∩vN 6= φ. (We show uN = vN ,
which is contradiction.) Let x ∈ uN ∩ vN . Write
ut = (vm1 )t ∈ vN.
This proves uN ⊆ vN . By interchanging the roles of u and v one obtains similarly that vN ⊆ uN .
Thus two cosets with nonempty intersection coincide.
By the first part of the proposition, uN = vN if and only if u ∈ vN if and only if u = vn, for
some n ∈ N if and only if v −1 u ∈ N , as claimed. Finally, v ∈ uN is equivalent to saying v is a
representative for uN , hence uN = vN if and only if u and v are representatives for the same coset
(namely the coset uN = vN ).
The following is a part of Decomposition Lemma but it is stated explicitly for its importance.
Exc. Prove the Coset Decomposition Lemma for the right cosets of the subgroup H in group G.
1. xH = H ⇐⇒ x ∈ H ⇐⇒ Hx = H.
2. H ≤ NG (H).
Remark: One can verify that in Example 1, the set of left cosets of N in D8 (i.e., {N, rN, sN, srN })
forms a partition of D8 .
Theorem (Lagrange’s Theorem): If G is a finite group and H is a subgroup of G, then the order
|G|
of H divides the order of G (i.e., |H| | |G|) and the number of left cosets of H in G equals |H| .
Proof: Let |H| = n and let the number of distinct left cosets of H in G equal k. By the coset
decomposition lemma, the set of left cosets of H in G partition G. That is,
Therefore,
|G| = |H| + |g2 H| + . . . |gk H|.
f : H → gH defined by f (h) = gh
is a surjection from H to the left coset gH. The left cancellation law implies this map is injective
i.e., gh1 = gh2 implies h1 = h2 . This proves that |gH| = |H|.
Since |gH| = |H| = n therefore |G| = |H| + |H| + · · · + |H| or |G| = kn. Thus k = |G| n
|G|
= |H| ,
completing the proof.
Remark: With the help of Lagrange’s Theorem, we can count at a glance, the number of left
(right) cosets of a subgroup H in G. For instance, if H = h−1i ≤ Q8 , then the number of left (right)
cosets of H in Q8 is:
|Q8 | 8
= = 4.
|H| 2
Index: If G is a group (possibly infinite) and H ≤ G, the number of left cosets of H in G is called
the index of H in G and is denoted by |G : H|.
Example 2: 0 is of infinite index in Z and hni is of index n in Z for every n > 0).
Remark: Example 2 indicates that infinite groups may have subgroups of finite or infinite index.
Question: A question arises whether the index can be defined by the right cosets of H in G.
The answer is affirmative as the number of left cosets of H in G and the number of right cosets of
H in G are same (Exc.)
Corollary 1: If G is a finite group and x ∈ G, then the order of x divides the order of G.
In particular x|G| = 1 for all x ∈ G.
Proof: We know that |x| = |hxi|. The first part of the corollary follows from Lagrange’s Theorem
applied to H = hxi. The second statement is clear since |G| is a multiple of the order of x.
Corollary 2: If G is a group of prime order p, then G is cyclic, hence G ∼ = Zp .
Proof: Since G = p > 1 therefore there exists x ∈ G, x 6= 1. Thus |hxi| > 1 and |hxi| divides |G|.
Since |G| is prime we must have |hxi| = |G| , hence G = hxi is cyclic (with any nonidentity element
x as generator). ”Classification Theorem for Cyclic Groups” completes the proof.
Home Work
1. Let G be a group and H ≤ G. Show that the number of left cosets of H in G is equal to the
number of right cosets of H in G. [Hint: Show that for any g ∈ G there is a biective map
between gH and Hg. ]
Exercises
1. Which of the following are permissible orders for subgroups of a group of order 120: 1,2 5, 7,
9, 15, 60, 240.
4. Use Lagrange’s Theorem in the multiplicative group (Z/pZ)× to prove Fermat’s Little Theo-
rem: if p is a prime then ap ≡ a (mod p) for all a ∈ Z.
5. Use Lagrange’s Theorem in the multiplicative group (Z/pZ)× to prove Euler’s Theorem:
aϕ(n) ≡ 1 (mod n) for every integer a relatively prime to n, where ϕ denotes Euler’s ϕ-
function.
100
6. Determine the last two digits of 33 . [Determine 3100 (mod ϕ(100)) and use the previous
Question 5. ]
uN · vN = (uv)N
2. If the above operation is well defined, then it makes the set of left cosets of N in G into a
group. In particular the identity of this group is the coset 1N and the inverse of gN is the
coset g −1 N i.e., (gN )−1 = g −1 N .
Proof: (1) Assume first that this operation is well defined, that is, for all u, v ∈ G, if
1g −1 N = ng −1 N i.e., g −1 N = ng −1 N.
gN g −1 = {gng −1 | n ∈ N }
G/N := {gN | g ∈ G}
is a group, called the quotient group, G/N (read G modulo N or simply G mod N ) or factor group.
Remark: Note that the structure of G is reflected in the structure of the quotient G/N when N is
a normal subgroup (for example, the associativity of the multiplication in G/N is induced from the
associativity in G and inverses in G/N are induced from inverses in G). We shall see more of the
relationship of G to its quotient G/N when we consider the Isomorphism Theorems.
1. N G
3. gN = N g for all g ∈ G
4. the operation on left cosets of N in G described in Proposition 5 makes the set of left cosets
into a group
5. gN g −1 ⊆ N for all g ∈ G.
Proof: Do yourself.
Remarks: In some cases we can minimize the computations necessary to determine whether a given
subgroup N is normal in a group G. For example
3. if generators for G are also known, then it suffices to check that these generators for G normalize
N . In particular, if generators for both N and G are known, this reduces the calculations to
a small number of conjugations to check.
4. if N is a finite group then it suffices to check that the conjugates of a set of generators for N
by a set of generators for G are again elements of N
Examples :
H ≤ NG (H) ≤ G.
By Lagrange’s Theorem, the order of H divides the order of NG (H) and the order of NG (H)
divides the order of G. Since G has order 6 and H has order 3, the only possibilities for NG (H)
are H or G. A direct computation gives
7
so H is not a normal subgroup of S3 .
7
Example 1 on page 91 of the book shows how the binary operation uH.vH = uvH fails to be well defined when
H is not normal subgroup.
gng −1 = n ∈ N .
Home Work
1. Prove Theorem 6.
2. Let G be group and N a normal subgroup of G. Prove that every subgroup of the quotient
G/N must be of the form A/N , where A ≤ G such that A ⊃ N . [Hint: Use the fact that
π : G → G/H is surjectiove homomorphism and the Excercise 1 below.]
Exercises
1. Let ϕ : G → H be a homomorphism and let E be a subgroup of H. Prove that ϕ−1 (E) ≤ G.
If E H prove that ϕ−1 (E) ≤ G. Deduce that ker ϕ G.
4. Let G and H be groups and let ϕ : G → H be a homomorphism. Then ϕ(g n ) = ϕ(g)n for all
n ∈ Z.
5. (a) Prove that if H and K are normal subgroups of a group G then their intersection H ∩ K
is also a normal subgroup of G.
7. Let A and B be groups. Show that {(a, 1)|a ∈ A} is a normal subgroup of A × B and the
quotient of A × B by this subgroup is isomorphic to B.
8. Let A be an abelian group and let D be the (diagonal) subgroup {(a, a)|a ∈ A} of A × A.
Prove that D is a normal subgroup of A × A and (A × A)/D ∼= A.
2. If G is an abelian group, any subgroup N of G is normal because for all g ∈ G and all n ∈ N ,
gng −1 = gg −1 n = n ∈ N.
and G/N = Z/nZ is a cyclic group with generator 1 = 1 + nZ (note that 1 is a generator
for G).
(b) Suppose now that G is the cyclic group. Let x be a generator of G and let N ≤ G. By
Classification Theorem for Subgroups of Cyclic Groups, N = hxd i, where d is the smallest
power of x which lies in N . Now G/N = {gN | g ∈ G} = {xa N | a ∈ Z} = {(xN )a | a ∈
Z} = hxN i i.e., G/N is cyclic with xN as a generator.
Suppose now that G = Zk is the cyclic group of order k. The order of xN in G/N equals
d (because i. (xN )d = xd N = N and ii. if (xN )l = N for some 0 < l < d, then xl ∈ N
contradicting d is the smallest such that xd ∈ N ). By Proposition 2.2 and Lagrange’s
|G| 8
Theorem, d = |N |
.
In summary,
Proof. Let G = {1, a, b, c}. By Lagrange’s theorem, we must have that the order of a is either 2 or 4. But if it were 4
the group would be cyclic, so the order of a is 2, i.e. a2 = 1. So if you assume ab = 1, you get a = a1 = aab = a2 b = b,
a contradiction (the starting assumption is that G has four different elements, of course). So ab cannot be 1. It
cannot be a either, because ab = a = a1 =⇒ b = 1, and similarly it cannot be b. So it has to be c. The same
argument shows that |a| = |b| = 2, ba = c = ab, ca = b = ac and cb = a = bc.
6. The property ”is a normal subgroup of” is not transitive. For example,
hsi hs, r2 i D8
(each subgroup is of index 2 in the next), however, hsi is not normal in D8 because rsr−1 =
sr2 ∈
/ hsi.
Solution: (a) Since G is abelian so every subgroup of G is abelian. This proves (a).
(b) Note that |S| = 4 (since (0, 1) has order 4) so G/S has order |G/S| = |G|/|S| = 8/2 = 2,
and therefore G/S ∼ = Z2 .
The subgroups H, J and K all have order 2. So, |G/H| = |G/J| = |G/K| = 8/2 = 4.
One can see that G/H ∼ = G/J ∼ = Z4 , but G/K ∼ = Z2 × Z2 One can verify this by computing
the orders of elements in the quotient, e.g. (0, 1) + H has order 4 in G/H.
Exercises
1. Do the following questions in Exercise 3.1.1: 14, 17, 20, 21.
3. Prove that every subgroup of Q8 is normal. For each subgroup find the isomorphism type of
its corresponding quotient.
4. Find all normal subgroups of D8 and for each of these find the isomorphism type of its corre-
sponding quotient.
HK = {hk | h ∈ H, k ∈ K}.
Proposition 14: If H and K are subgroups of a group, HK is a subgroup if and only if HK = KH.
Proof: Assume first that HK = KH and let a, b ∈ HK. We prove ab−1 ∈ HK so HK is a subgroup
by the subgroup criterion. Let
a = h1 k1 and b = h2 k2 for some h1 , h2 ∈ H and k1 , k2 ∈ K. Thus b−1 = k −1 h−1 , so ab−1 =
h1 k1 k2−1 h−1
2 . Let k3 = k1 k2
−1
∈ K and h3 = h−1 2 . Thus ab
−1
= h1 k3 h3 . Since HK = KH,
k3 h3 = h4 k4 , for some h4 ∈ H, k4 ∈ K.
Thus ab−1 = h1 h4 k4 , and since h1 , h4 ∈ H, k4 ∈ K, we obtain ab−1 ∈ HK, as desired.
Conversely, assume that HK is a subgroup of G. Since K ≤ HK and H ≤ HK, by the closure
property of subgroups, KH ⊆ HK. To show the reverse containment let hk ∈ HK. Since HK is
assumed to be a subgroup, write hk = a−1 , for some a ∈ HK. If a = h1 k1 , then
hk = (h1 k1 )−1 = k1−1 h−1
1 ∈ KH, completing the proof.
hk = (hkh−1 )h ∈ KH.
This proves HK ⊆ KH. Similarly, kh = h(hkh−1 ) ∈ HK, proving the reverse containment. The
corollary follows now from the preceding Lemma.
Proof. Exercise.
Since the group operation in AB/B is well defined it is easy to see that ϕ is a homomorphism:
It is clear from the definition of AB that ϕ is surjective. The identity in AB/B is the coset 1B , so
1. The image of HN and H are same since π(hn) = π(h)π(n) = π(h) and since HN includes N , π(H) = π(HN ) =
(HN )/N .
2. Let restriction of π on H is π 0 then π 0 is an homomorphism from H to G/N . What is the kernel of π 0 ?
Ker(π 0 ) = H ∩ N . Then by first ismomorphim theorem π 0 (H) ∼
= H/(H ∩ N ).
From 1 and 2, we have desired result.
http://math.stackexchange.com/questions/722632/interpretation-of-second-isomorphism-theorem?rq=1
The third Isomorphism Theorem considers the question of taking quotient groups of quotient
groups.
(G/H)/(K/H) ∼
= G/K.
Proof. We leave as an easy exercise the verification that K/H G/H. Define
Theorem (20). (The Fourth or Lattice Isomorphism Theorem) Let G be a group and let N be a
normal subgroup of G. Then there is a bijection from the set of subgroups A of G which contain N
onto the set of subgroups A of G/N .
Home Work
12
Indeed, Ψ(A) = π(A) = {π(a) | a ∈ A} = {aN | a ∈ A} = A/N
13
1 = 1N ∈ G/N
Excercise
Home Work
1. Find a group of permutations such that V4 is isomorphic to that group of permutations, where
V4 = ha, b | a2 = b2 = (ab)2 = 1
2.
3.
Excercise
1. Excercise
Examples:
1. Since the identity permutation i of A leaves each element of A fixed, the orbits of i are the
one-element subsets of A.
!
1 2 3 4 5 6 7 8
2. Find the orbits of the permutation in S8 .
3 8 6 7 4 1 5 2
Solution: To find the orbit containing 1, we apply a repeatedly, obtaining symbolically
σ σ σ σ σ σ σ
1 → 3 → 6 → 1 → 3 → 6 → 1 → 3....
Remark: While permutation multiplication in general is not commutative, it is readily seen that
multiplication of disjoint cycles is commutative. Since the orbits of a permutation are unique,
the representation of a permutation as a product of disjoint cycles, none of which is the identity
permutation, is unique up to the order of the factors.
That is, any cycle is a product of transpositions. We then have the following as a corollary to
previous Theorem.
Corollary. Any permutation of a finite set of at least two elements is a product of transpositions.
Examples:
We have seen that every permutation of a finite set with at least two elements is a product of
transpositions. The transpositions may not be disjoint, and a representation of the permutation
in this way is not unique. For example, we can always insert at the beginning the transposition
(1,2) twice, because (1, 2) (1, 2) is the identity permutation. What is true is that the number of
thatis, σ ∈ An is mapped into (1, 2)σ by λτ . Observe that since σ is even, the permutation (1, 2)σ
can be expressed as a product of a (1 + even number), or odd number, of transpositions, so (1, 2)σ
is indeed in Bn .
If for σ and µ in An , it is true that λτ (σ) = λτ (µ), then
(1 2)σ = (1 2)µ
so if ρ ∈ Bn , then
τ −1 ρ ∈ An
and
λτ (τ −1 ρ) = τ (τ −1 ρ) = ρ.
Thus λτ is onto Bn .
Theorem (3). If n ≥ 2, then the collection of all even permutations of {1, 2, 3, . . . , n} forms a
subgroup of order n!/2 of the symmetric group Sn .
Proof. Exercise.14
Definition. The subgroup of Sn consisting of the even permutations of n letters is the alternating
group An on n letters.
Both Sn and An are very important groups. Cayley’s theorem shows that every finite group G
is structurally identical to some subgroup of Sn for n = |G|. It can be shown that there are no
formulas involving just radicals for solution of polynomial equations of degree n for n ≥ 5. This fact
is actually due to the structure of An , surprising as that may seem!
Home Work
Excercise
2. In an abelian group G, every element a ∈ G is related to itself because x−1 ax = a for any
x ∈ G.
Proof. Do by yourself.
[a] = {b ∈ G | b ∼ a}
= {b ∈ G | b = x−1 ax, x ∈ G}
not.
= {x−1 ax | x ∈ G} = : Cl(a)
called the Conjugacy class of a. We know that the set of equivalence classes partition G.
Example:
Cl((1)) = {(1)}, Cl((12)) = {(12), (13), (23)}, and Cl((123)) = {(123), (132)}.
x−1 ax = y −1 ay ⇐⇒ axy −1 = xy −1 a
⇐⇒ xy −1 ∈ CG (a)
⇐⇒ CG (a)x = CG (a)y.
Corollary 3: Let G be a group then |Cl(a)| = |G|/|CG (a)| = |G : CG (a)|, where CG (a) is the
centralizer of a in G.
Theorem 4 (The Class Equation): Let G be a finite group and let Cl(g1 ), Cl(g2 ), . . . , Cl(gr )
be the distinct Conjugacy classes of G not contained in the center Z(G) of G. Then
Pr
|G| = |Z(G)| + |G : CG (gi )|.
i=1
Proof: Let Z(G) = {1, z2 , . . . , zm }, then by Prop. 2(1), the full set of conjugacy classes of G is
given by
{1}, {z2 }, . . . , {zm }, Cl(g1 ), Cl(g2 ), . . . , Cl(gr )
Examples:
1. The class equation for any abelian group of order n is n.
2. In any group G we have hgi ⊆ CG (g); this observation helps to minimize computations of con-
jugacy classes. For example, in the quaternion group Q8 we see that hii ≤ CQ8 (i) ≤ Q8 . Since
i∈/ Z(Q8 ) we must have CQ8 (i) = hii and |Q8 : hii| = 2. Thus i has precisely 2 conjugates in Q8 ,
namely i and −i = kik −1 . The other conjugacy classes in Q8 are determined similarly and are
The first two classes form Z(Q8 ) and the class equation for this group is
|Q8 | = 2 + 2 + 2 + 2.
Theorem 5: If p is a prime and P is a group of prime power order pα for some α ≥ 1, then
P has a nontrivial center i.e., Z(P ) 6= 1.
Proof: By the class equation
r
P
|P | = |Z(P )| + |P : CP (gi )|
i=1
where g1 , . . . , gr are representatives of the distinct non-central conjugacy classes. Since gi ∈
/ Z(P ),
16
CP (gi ) 6= P for i = 1, 2, . . . , r, so p divides |P : CP (gi )| by Lagrange’s Theorem . Since p also
divides |P | it follows that p divides |Z(P )|, hence the center must be nontrivial.
Home Work
2. If p is a prime and P is a group of prime power order pα for some α ≥ 1, then prove that P
has a nontrivial center: Z(P ) 6= 1.
Excercise
1. Find the Conjugacy classes and the Class equation for the group S3 , D6 , D8 , Z/nZ.
if G is a finite group and n divides |G|, then G need not have a subgroup of order n.
4!
Example: let A4 be the alternating group of degree 4. We know that |A4 | = 2
= 12 and
n
A4 = I, (1 2 3), (1 2 4), (1 3 4), (2 3 4), (1 3 2), (1 4 2), (1 4 3), (2 4 3), (1 2)(3 4), (1 3)(2 4), (1 4)(2 3)}
A partial converse which holds for arbitrary finite groups is the following result:
Theorem (Cauchy’s Theorem): If G is a finite group and p is a prime dividing |G|, then G has
an element of order p.
1. In group Z/12Z, the subgroups h3i and h6i are both 2-groups. The subgroup h3i is also Sylow
2-subgroup but the subgroup h6i is not Sylow 2-subgroup. The subgroup h4i is 3-subgroup as
well as Sylow 3-subgroup. The subgroup h2i is neither a 2-subgroup nor a 3-subgroup hence
neither a Sylow 2-subgroup nor a Sylow 3-subgroup.
Theorem (Sylow’s Theorem): Let G be a group of order pa m, where p is a prime not dividing
m.
(1) Sylow p-subgroups of G exist, i.e., Sylp (G) 6= ∅.
(2) If P is a Sylow p-subgroup of G and Q is any p-subgroup of G, then there exists g ∈ G such that
Q ≤ gP g −1 , i.e., Q is contained in some conjugate of P . In particular, any two Sylow p-subgroups
of G are conjugate in G.
(3) The number of Sylow p-subgroups of G is of the form 1 + kp, i.e.,
np ≡ 1(mod p).
Further, np is the index in G of the normalizer NG (P ) for any Sylow p-subgroup P , hence np
divides m.
Corollary 1: Any two Sylow p-subgroups of a group (for the same prime p) are isomorphic.19
Corollary 2: Let P be a Sylow p-subgroup of G. Then the following are equivalent:
(1) P is the unique Sylow p-subgroup of G, i.e., np = 1;
(2) P is normal in G.
18
Wikipedia: In mathematics, specifically in the field of finite group theory, the Sylow theorems are a collection of
theorems named after the Norwegian mathematician Ludwig Sylow (1872) that give detailed information about the
number of subgroups of fixed order that a given finite group contains.
19
Indeed. By Sylow Theorem (2), for any two Sylow p-subgroups P and Q in G there exists g ∈ G such that
Q = gP g −1 . The map x 7→ gxg −1 gives P ∼= gP g −1 = Q.
Home Work
2.
3.
Excercise
1. Excercise
1. Groups of order pq, p and q primes with p < q are not simple.
Proof 1.:Suppose |G| = pq for primes p and q with p < q. let Q ∈ Sylq (G). We show that Q is
normal in G.
Now by Sylow’s Theorem (3), nq = 1 + kq for some k ≥ 0 and nq divides p. Combining this with
p < q, gives k = 0 i.e., nq = 1 and hence Q G, by Cor. 2.
Lemma: If the Sylow p-subgroups have order p, then the intersection of any two distinct Sylow
p-subgroups must be the identity.
Proof: Suppose P and Q be distinct Sylow p-subgroups of order p and P ∩ Q 6= 1. Since P ∩ Q
is a subgroup of P , |P ∩ Q| | |P |, by Lagrange’s theorem. Which implies that |P ∩ Q| = p i.e.,
P ∩ Q = P . By similar arguments, P ∩ Q = Q. But this is a contradiction to the fact that P 6= Q.
Since q is prime, either q|p − 1 or q|p + 1. The former is impossible since q > p so the latter holds.
Since q > p but q|p + 1, we must have q = p + 1. This forces p = 2, q = 3 and |G| = 12. The result
now follows from the fact that ”a group of order 12 is not simple”.21
Exercises
1. Prove that a group of order 56 has a normal Sylow p-subgroup for some prime p dividing its
order.
2. Prove that a group of order 312 has a normal Sylow p-subgroup for some prime p dividing its
order.
3. Prove that a group of order 351 has a normal Sylow p-subgroup for some prime p dividing its
order.
4. Let |G| = pqr, where p, q and r are primes with p < q < r. Prove that G has a normal Sylow
subgroup for either p, q or r.
5. Prove that if |G| = 105 then G has a normal Sylow 5-subgroup and a normal Sylow 7-subgroup.
6. Prove that a group of order 200 has a normal Sylow 5-subgroup.
7. Prove that if |G| = 6545 then G is not simple.
8. Prove that if |G| = 1365 then G is not simple.
9. Prove that if |G| = 2907 then G is not simple.
10. Prove that if |G| = 132 then G is not simple.
11. Prove that if |G| = 462 then G is not simple.
G1 × G2 × · · · × Gn = {(g1 , g2 , . . . , gn ) | gi ∈ Gi }
21
As we assumed n3 = nq > 1, we must have n2 = np = 1.
Proposition 1. If G1 , . . . , Gn are groups, their direct product is a group of order |G1 ||G2 | . . . |Gn |
(if any G is infinite, so is the direct product).
Proof: Note that since gi ∈ G, hi ∈ Gi and Gi is a group, we have gi hi ∈ Gi . Thus G1 ×G2 ×· · ·×Gn
is closed under the binary operation.
The associative law in G1 × G2 × · · · × Gn is thrown back onto the associative law in each component
as follows:
(a1 , a2 , . . . , an )[(b1 , b2 , . . . , bn )(c1 , c2 , . . . , cn )]
= (a1 , a2 , . . . , an )(b1 c1 , b2 c2 , . . . , bn cn )
= ((a1 (b1 c1 ), a2 (b2 c2 ), . . . , an (bn cn ))
= ((a1 b1 )c1 ), (a2 b2 )c2 ), . . . , (an bn )cn ))
= (a1 b1 , a2 b2 ), . . . , an bn )(c1 , c2 , . . . , cn )
= [(a1 , a2 , . . . , an )(b1 , b2 , . . . , bn )](c1 , c2 , . . . , cn )
If g.c.d.(m, n) 6= 1 , every element of Zm × Zn has order at most l, hence has order strictly less than
mn23 , so Zm × Zn cannot be isomorphic to Zmn .
Conversely, if (m, n) = 1, then |(1, 1)| = 24 l.c.m.(|1|, |1|) = l.c.m.(m, n) = mn. Thus, by order
considerations, Zm × Zn = h(1, 1)i is cyclic, completing the proof.
Corollary 3: The group Zm1 × Zm2 × · · · × Zmn , is cyclic and isomorphic to Zm1 m2 ...mn if and only
if the numbers mi for i = 1, . . . , n are such that the g.c.d. of any two of them is 1.
23
since g.c.d.(m, n) l.c.m.(m, n) = mn
24
Here taking a power of (1, 1) in our additive notation will involve adding (1, 1) to itself repeatedly. Under addition
by components, the first component 1 ∈ Zm yields 0 only after m summands, 2m summands, and so on, and the second
component 1 ∈ Zn yields 0 only after n summands, 2n summands, and so on. For them to yield 0 simultaneously, the
number of summands must be a multiple of both m and n. The smallest number that is a multiple of both m and n
will be l.c.m.(1, 1).
then Zn is isomorphic to
Z(p1 )n1 × Z(p2 )n2 × · · · × Z(pr )nr
Example: Find the order of (8, 4, 10) in the group Z12 × Z60 × Z24 ?
Since the gcd of 8 and 12 is 4, we see that 8 is of order 12
4
= 3 in Z12 . Similarly, we find that 4 is of
order 15 in Z60 and 10 is of order 12 in Z24 . The l.c.m. of 3, 15, and 12 is 3.5.4 = 60, so (8, 4, 10) is
of order 60 in the group Z12 × Z60 × Z24 .
Example: The group Z × Z2 is generated by the elements (1, 0) and (0, 1). More generally, the
direct product of n cyclic groups, each of which is either Z or Zm for some positive integer m, is
generated by the n n-tuples
Such a direct product might also be generated by fewer elements. For example, Z3 × Z4 × Z35 is
generated by the single element (1, 1, 1).
that is, the set of all n-tuples with the identity elements in all places but the ith, is a subgroup
of G1 ×G2 ×· · ·×Gn . It is also clear that this subgroup Gi is naturally isomorphic to Gi ; just rename
The group Gi , is mirrored in the ith component of the elements of Gi , and the ej in the other
components just ride along. We consider G1 × G2 × · · · × Gn to be the internal direct product of
2. Is Z8 × Z9 × Z25 is cyclic?
Solution: By using Corollary 3, since 8, 9 and 25 are pairwise disjoint so
Z8 × Z9 × Z25 ∼
= Z8·9·25 and Z8 × Z9 × Z25 = h(1, 1, 1)i
Exercises
1. Show that the center of a direct product is the direct product of the centers: Z(G1 × G2 ×
· · · × Gn ) = Z(G1 ) × Z(G2 ) × · · · × Z(Gn ). Deduce that a direct product of groups is abelian
if and only if each of the factors is abelian.
Home Work
1.
3.
Excercise
1. Excercise