Notes
Notes
Lecture-1
Definition 2. A group is an ordered pair (G, ∗) where G is a set and ∗ is a binary operation on
G satisfying the following properties
1. x ∗ (y ∗ z) = (x ∗ y) ∗ z for all x, y, z in G.
1. a set, and
Then, one must verify that the binary operation is associative, that there is an identity in the set,
and that every element in the set has an inverse.
Convention If it is clear what the binary operation is, then the group (G, ∗) may be referred to
by its underlying set G alone.
Examples of Groups:
1
6. (Zn , +) is a group with identity 0, where + is the addition modulo n operation, that is
x + y = r if the reminder when x + y is divided by n is r. The inverse of x ∈ Zn is n − x if
x 6= 0, the inverse of 0 is 0.
7. (Rn , +) where + is vector addition. The identity is the zero vector (0, 0, . . . , 0) and the inverse
of the vector x = (x1 , x2 , . . . , xn ) is the vector −x = (−x1 , −x2 , . . . , −xn ).
8. (Zn2 , +) where + is vector addition modulo 2 operation. The identity is the zero vector
(0, 0, . . . , 0) and the inverse of the vector x is the vector itself.
9. (M2 (K), +) where K is any one of Z, Q, R, Zn is a group whose identity is the zero matrix
0 0
0 0
and the inverse of the matrix
a b
A=
c d
is the matrix
−a −b
−A = .
−c −d
Note that the binary operations in the above examples are all commutative. For historical
reasons, there is a special name for such groups:
Definition 3. A group (G, ∗) is said to be abelian if x ∗ y = y ∗ x for all x and y in G. A group
is said to be non-abelian if it is not abelian.
1. For each n ∈ N, the set Sn of all permutations on [n] = {1, 2, . . . , n} is a group under
compositions of functions. This is called the symmetric group of degree n. We discuss
this group in detail in the next lecture. The group Sn is non-abelian if n ≥ 3.
2. Let K be any one of Q, R or Zp , where p is a prime number. Define GL(2, K) to be the set
of all matrices in M2 (K) with non-zero determinant. Then (GL(2, K), ·) is a group. Here ·
represents matrix multiplication. The identity of GL(2, K) is the identity matrix
1 0
0 1
and the inverse of
a b
c d
is d −b
ad−bc ad−bc .
−c a
ad−bc ad−bc
GL(2, K) is called the general linear group of degree 2 over K. These groups are non-
abelian. We discuss them in more detail later.
2
Lemma 1.1. If (G, ∗) is a group then:
(a) The identity of G is unique.
(b) The inverse of each element in G is unique.
The above definition is used when we think of the group’s operation as being a type of mul-
tiplication or product. If instead the operation is denoted by +, we have instead the following
definition.
Definition 5. Let (G, +) be a group. Let a be any element of G. We define −a to be the inverse
of a in the group G.
Remark 1.2. Let (G, ∗) be a group with identity e. It is easy to check that the following hold for
all elements a, b, c, d in G:
3
MTH-204: Abstract Algebra
Lecture-2
1 Symmetric Group
If n is a positive integer, we denote [n] = {1, 2, . . . , n}. A permutation of [n] is a one-to-one, onto
function from [n] to [n] and Sn is the set of all permutations of [n].
Let us discuss the different ways to specify a function from [n] to [n] and how to tell when we
have a permutation. It is traditional (but not compulsory) to use lower case Greek letters such as
σ, τ , α, β, etc., to indicate elements of Sn . To be specific let n = 4. We may define a function
σ : [4] → [4] by specifying its values at the elements 1, 2, 3, and 4. For example, let’s say:
We call this the two line or two row notation. The function σ just defined is one-to-one and onto,
that is, it is a permutation of [4].
For another example, let
1 2 3 4
τ= .
1 3 1 4
The function τ is not one-to-one since 1 6= 3 but τ (1) = τ (3). This problem can always be identified
by the existence of the same element more than once in the second line of the two line notation. τ
is also not onto since the element 2 does not appear in the second line.
Let
1 2 ··· n
σ= .
σ(1) σ(2) · · · σ(n)
be the two line notation of an arbitrary function σ : [n] → [n]. Then:
(1) σ is one-to-one if and only if no element of [n] appears more than once in the second
line.
(2) σ is onto if and only if every element of [n] appears in the second line at least once.
1
Thus σ is a permutation if and only if the second row is just a rearrangement or shuffling of the
numbers 1, 2, . . . , n.
If σ and τ are elements of Sn , then στ is defined to be the composition of the functions σ and τ .
That is, στ is the function whose rule is given by:
We sometimes call στ simply the product of σ and τ . Let’s look at an example to see how this
works. Let σ and τ be defined as follows:
1 2 3 1 2 3
σ= , τ=
2 1 3 2 3 1
It follows that
στ (1) = σ(τ (1)) = σ(2) = 1
στ (2) = σ(τ (2)) = σ(3) = 3
στ (3) = σ(τ (3)) = σ(1) = 2
Thus we have
1 2 3
στ =
1 3 2
One can also find products of permutations directly from the two line notation as follows:
1 2 3 1 2 3 1 2 3
First Step: =
2 1 3 2 3 1 1 − −
1 2 3 1 2 3 1 2 3
Second Step: =
2 1 3 2 3 1 1 3 −
1 2 3 1 2 3 1 2 3
Third Step: =
2 1 3 2 3 1 1 3 2
Whenever we need to prove two functions are equal, we require the following definition:
The identity of Sn :
2
The identity of Sn is the so-called identity function
ι : [n] → [n].
which is defined by the rule:
ι(x) = x, for all x ∈ [n].
In the two line notation ι is described by
1 2 ··· n
ι=
1 2 ··· n
The function ι is clearly one-to-one and onto and satisfies
ισ = σ and σι = σ, for all σ ∈ Sn .
So ι is the identity of Sn with respect to the binary operation of composition.
[Note that we use the Greek letter ι (iota) to indicate the identity of Sn .]
If σ ∈ Sn , then by definition σ : [n] → [n] is one-to-one and onto. Hence the rule
σ −1 (y) = x if and only if σ(x) = y
defines a function σ −1 : [n] → [n]. The function σ −1 is also one-to-one and onto (check this!) and
satisfies
σσ −1 = ι and σ −1 σ = ι,
so it is the inverse of σ in the group sense also.
In terms of the two line description of a permutation, if
··· x ···
σ=
··· y ···
then
−1 ··· y ···
σ =
··· x ···
The inverse of a permutation in the two line notation may be obtained by interchanging the
two lines and then reordering the columns so that the numbers on the top line are in numerical
order. Here’s an example:
1 2 3
σ=
2 3 1
Interchanging the two lines we have:
2 3 1
.
1 2 3
Reordering the columns we obtain
1 2 3
σ −1 = .
3 1 2
3
Lemma 1.1. For any three functions
α : A → B, β : B → C, γ:C→D
we have
(γβ)α = γ(βα).
With this corollary, we complete the proof that Sn under the binary operation of composition is
a group.
1. Draw a dot and number it 1. Let i1 = σ(1). If i1 6= 1 draw another dot and label it i1 .
3. Assume that dots numbered 1, i1 , i2 , . . . , ik have been drawn. Consider two cases:
(i) There is an arrow leaving every dot drawn so far. In this case let ik+1 be the smallest
number in [n] not yet labeling a dot. If there are no such then stop, you have completed
the diagram, otherwise draw a new dot and label it ik+1
(ii) There is a dot numbered j with no arrow leaving it. In this case let ik+1 = σ(j). If there
is no dot labeled ik+1 draw a new dot and label it ik+1 . Draw an arrow from dot j to
dot ik+1 .
4
Definition 2. Let i1 , i2 , . . . , ik be a list of k distinct elements from [n]. Define a permuation σ in
Sn as follows:
σ(i1 ) = i2
σ(i2 ) = i3
σ(i3 ) = i4
.. .. ..
. . .
σ(ik−1 ) = ik
σ(ik ) = i1
and if x ∈
/ {i1 , i2 , . . . , ik } then
σ(x) = x
Such a permutation is called a cycle or a k-cycle and is denoted by
(i1 i2 · · · ik ).
For example, let σ be the 3-cycle defined by σ = (3 2 1). σ may be considered as an element of
S3 in which case in two line notation we have
1 2 3
σ= .
3 1 2
Notice that according to the definition if x ∈/ {3, 2, 1} then σ(x) = x. So we could also consider
(3 2 1) as an element of S4 . In which case we would have:
1 2 3 4
σ= .
3 1 2 4
Similarly, (3 2 1) could be an element of Sn for any n ≥ 3. Note also that we could specify the
same permutation by any of the following
In this case, there are three numbers 1, 2, 3 in the cycle, and we can begin the cycle with any one
of these. In general, there are k different ways to write a k-cycle. One can start with any number
in the cycle.
Definition 3. Two cycles (i1 i2 · · · ik ) and (j1 j2 · · · j` ) are said to be disjoint if the sets
{i1 , i2 , . . . , ik } and {j1 , j2 , . . . , j` } are disjoint.
So, for example, the cycles (1 2 3) and (4 5 8) are disjoint, but the cycles (1 2 3) and (4 2 8)
are not disjoint.
5
Lemma 1.3. If σ and τ are disjoint cycles, then στ = τ σ.
Proof Let σ = (a1 · · · ak ) and τ = (b1 · · · b` ). Let {c1 , · · · , cm } be the elements of [n] that are in
neither {a1 , . . . , ak } nor {b1 , · · · , b` }. Thus
We want to show στ (x) = τ σ(x) for all x ∈ [n]. To do this we consider first the case x = ai for
some i. Then ai ∈ / {b1 , · · · , b` } so τ (ai ) = ai . Also σ(ai ) = aj , where j = i + 1 or j = 1 if i = k. So
also τ (aj ) = aj . Thus
Thus, στ (ai ) = τ σ(ai ). It is left to the reader to show that στ (x) = τ σ(x) if x = bi or x = ci ,
which will complete the proof.
σ = σ1 σ2 · · · σm (1)
where σ1 , σ2 , . . . , σm are pairwise disjoint cycles, that is, for i 6= j, σi and σj are disjoint. If all
1-cycles of σ are included, the factors are unique except for the order.
To save time we omit a formal proof of this theorem. The process of finding the disjoint cycle
decomposition of a permutation is quite similar to finding the cycle diagram of a permutation.
Consider, for example, the permutation α ∈ S15
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
α= .
13 11 7 6 5 4 3 10 2 12 14 1 15 9 8
To obtain this, one starts a cycle with 1, since α(1) = 13 we have the partial cycle (1 13. Next,
we observe that α(13) = 15. This gives the partial cycle (1 13 15. We continue in this way till
we obtain the cycle (1 13 15 8 10 12). Then we pick the smallest number in [15] not used so far,
namely, 2. We start a new cycle with 2: Noting that α(2) = 11 we have the partial cycle (2 11.
Continuing we obtain the cycle (2 11 14 9). And we continue in this way till all the elements of
[15] are in some cycle.
Note that the transposition (i j) interchanges i and j and leaves the other elements of [n] fixed.
It transposes i and j.
6
Lemma 1.5. Every element of Sn can be written as a product of transpositions.
Proof. We see that every cycle can be written as a product of transpositions as follows:
Then, since each permutation is a product of cycles, we can obtain each permutation as a product
of transpositions.
Proof. Let id = t1 t2 · · · tm−1 tm where ti ’s are transpositions. We need to show that m is even. Note
that m 6= 1 as a single transposition is not the identity.
If m = 2 we are done.
We proceed by (strong) induction. Suppose that the theorem is true for any integer less than
m, m ≥ 2. We will show that it holds for m. Let tm = (a, b)
The idea is that we will try to rewrite the permutation in such a way that we shift a as far
left as possible until we eventually remove a from the permutation. The last pair of transpositions
tm−1 tm must be one of these four cases:
(ab)(ab), (bc)(ab), (ac)(ab), (cd)(ab).
If tm−1 tm = (ab)(ab) = id, we are left with m − 2 transpositions and by induction m − 2 is even
and so m is even.
If tm−1 tm = (bc)(ab), then we can replace it by (ac)(bc) since (bc)(ab) = (ac)(bc).
If tm−1 tm = (ac)(ab), then we can replace it by (ab)(bc) since (ac)(ab) = (ab)(bc).
If tm−1 tm = (cd)(ab), then we can replace it by (ab)(cd) since (cd)(ab) = (ab)(cd).
So we have rewritten tm−1 tm in such a way that a no longer occurs in the last transposition.
Successively, we rewrite the pairs tm−1 tm , then tm−2 tm−1 , tm−2 tm−1 , and so on. Eventually, we
will reach the first case above, (ab)(ab), where we can cancel out two transpositions. If we don’t,
then the left most transposition t1 will have the only occurrence of a. This would contradict the
assumption that the permutation is the identity, because if only one transposition contains a, then
the permutation does not fix a.
Once we cancel the two transpositions, then there are only m − 2 transpositions in the permu-
tation, and we can apply our induction hypothesis.
Theorem 1.7. Every element of Sn can be written as a product of transpositions. The factors of
such a product are not unique, however, if σ ∈ Sn can be written as a product of m transpositions
and if the same σ can also be written as a product of n transpositions, then k and ` have the same
parity.
Proof. The first part of this theorem follows from the above lemma. If σ = t1 t2 · · · tm = s1 s2 · · · sn ,
where ti and sj are transpositions, then id = σ.σ −1 = t1 t2 · · · tm sn sn−1 · · · s2 s1 . Since the identity
permutation is even, m + n is even. So m and n are both even or both odd.
7
Definition 5. A permutation is even if it is a product of an even number of transpositions and is
odd if it is a product of an odd number of transpositions. We define the function sign : Sn → {1, −1}
by
1 if σ is even
sign(σ) =
−1 if σ is odd
If n = 1 then there are no transpositions. In this case to be complete we define the identity
permutation ι to be even.
Remark. Let A = [aij ] be an n × n matrix. The determinant of A may be defined by the sum
X
det(A) = sign(σ)a1σ(1) a2σ(2) · · · anσ(n) .
σ∈Sn
For example, if n = 2 we have only two permutations ι and (1 2). Since sign(ι) = 1 and sign((1 2)) =
−1 we obtain
det(A) = a11 a22 − a12 a21 .
8
MTH-204: Abstract Algebra
Lecture-3
1 Subgroups
Definition 2. Let a be an element of the group G. If there exists n ∈ N such that an = e we say
that a has finite order. and we define
o(a) = min{n ∈ N | an = e}
o(a) = ∞.
1
Lemma 1.1. If G is a finite group, then every element of G has finite order.
a1 , a2 , a3 , . . . , ai , . . .
of elements in G. Since G is finite, all the elements in the list cannot be different. So there must
be positive integers i < j such that ai = aj . Since i < j, j − i is a positive integer. Then we have
That is, an = e for the positive integer n = j − i. So a has finite order, which is what we wanted
to prove.
Proof. By the Euclidean algorithm, there exist integers r and s such that d = mr + ns. We have
Lemma 1.3. For G a group, any non empty subset H of G is a subgroup of G if and only if for
every a, b ∈ H we have ab−1 ∈ H.
Proof. If H ≤ G, the two given statements clearly hold as H contains the identity of G and is
closed under inverses and multiplication.
To prove the converse, let x be any element of H (which exists as H 6= ∅). We have xx−1 ∈
H =⇒ 1 ∈ H. As H contains 1, for any element h of H, H contains 1h−1 = h−1 , that is, it is
closed under inverses. For any x and y in H, as y −1 ∈ H, we have that x(y −1 )−1 = xy ∈ H, that
is, H is closed under multiplication.
To prove the second part, we see that x, x2 , x3 , . . . ∈ H for any x ∈ H. Using above Lemma,
we see that x is of finite order n. Then x−1 = xn−1 ∈ H so H is closed under inverses.
Lemma 1.4. Let H = hxi. Then |H| = |x| (where if one side of the inequality is infinite, so is the
other).
2
a, b ∈ Hi for all i. Since Hi is a subgroup, we have ab−1 ∈ Hi for all
T
Proof. If a, b ∈ i∈I Hi , then
i. So ab−1 ∈ i∈I Hi . So i∈I Hi is a subgroup of G. Consider 3Z and 5Z with the operation of
T T
addition: their union is not a subgroup of Z, because for example 8 ∈ / 3Z ∪ 5Z.
Since gag −1 = g if and only if ga = ag, CG (A) is the set of all elements that commute with
every element of A.
Now observe that CG (A) is a subgroup of G as first of all, 1 ∈ CG (A) so CG (A) 6= ∅, and second
of all, if x, y ∈ CG (A), we have xax−1 = a and yay −1 = a, that is, y −1 ay = a for all a ∈ A. We
then have a = xax−1 = x(y −1 ay)x−1 = (xy −1 )a(xy −1 )−1 so xy −1 ∈ CG (A). Thus, CG (A) ≤ G.
Z(G) is the set of all elements that commute with every element of G.
As Z(G) = CG (G), we have Z(G) ≤ G.
Definition 1.8. Let G be a group and A be a subset of G. Define gAg −1 = {gag −1 | a ∈ A}.
Define
NG (A) = {g ∈ G | gAg −1 = A}.
This set is called the normalizer of A in G.
The proof that NG (A) ≤ G is similar to that we used to prove that CG (A) ≤ G.
Note that CG (A) ≤ NG (A).
3
MTH-204: Abstract Algebra
Lecture-4
Lagrange’s Theorem, one of the most important results in finite group theory, states that the order
of a subgroup must divide the order of the group. This theorem provides a powerful tool for
analyzing finite groups; it gives us an idea of exactly what type of subgroups we might expect a
finite group to possess. Central to understanding Lagranges’s Theorem is the notion of a coset.
2 Cosets
Hg = {hg : h ∈ H}.
If left and right cosets coincide or if it is clear from the context to which type of coset that we are
referring, we will use the word coset without specifying left or right.
Example 1. Let H be the subgroup of Z6 consisting of the elements 0 and 3. The cosets are
0 + H = 3 + H = {0, 3}
1 + H = 4 + H = {1, 4}
2 + H = 5 + H = {2, 5}.
We will always write the cosets of subgroups of Z and Zn with the additive notation we have used
for cosets here. In a commutative group, left and right cosets are always identical.
Example 2. Let H be the subgroup of S3 defined by the permutations {(1), (123), (132)}. The
left cosets of H are
1
The right cosets of H are exactly the same as the left cosets:
It is not always the case that a left coset is the same as a right coset. Let K be the subgroup
of S3 defined by the permutations {(1), (12)}. Then the left cosets of K are
The following lemma is quite useful when dealing with cosets and the proof is easy.
Lemma 2.1. Let H be a subgroup of a group G and suppose that g1 , g2 ∈ G. The following
conditions are equivalent.
1. g1 H = g2 H;
2. Hg1−1 = Hg2−1 ;
3. g1 H ⊆ g2 H;
4. g2 ∈ g1 H;
5. g1−1 g2 ∈ H.
In all of our examples the cosets of a subgroup H partition the larger group G. The following
theorem proclaims that this will always be the case.
Theorem 2.2. Let H be a subgroup of a group G. Then the left cosets of H in G partition G.
That is, the group G is the disjoint union of the left cosets of H in G.
2
Remark. The above theorem can also be proved in the following way. Define a relation R on G
by xRy if x−1 y ∈ H. It is easy to check that R is an equivalence relation on G and the equivalence
classes are nothing but the left cosets of H. We have [x] = {y ∈ G : xRy} = {y ∈ G : x−1 y ∈
H} = {y ∈ G : y ∈ xH} = xH. Since the equivalences classes partition the set, we get that G is
the disjoint union of the left cosets of H in G.
Remark. There is nothing special in this theorem about left cosets. Right cosets also partition
G; the proof of this fact is exactly the same as the proof for left cosets except that all group
multiplications are done on the opposite side of H.
Example 4. Suppose that G = S3 , H = {(1), (123), (132)}, and K = {(1), (12)}. Then [G : H] = 2
and [G : K] = 3.
Theorem 2.3. Let H be a subgroup of a group G. The number of left cosets of H in G is the same
as the number of right cosets of H in G.
Proof. Let LH and RH denote the set of left and right cosets of H in G, respectively. If we can define
a bijective map φ : LH → RH , then the theorem will be proved. If gH ∈ LH , let φ(gH) = Hg −1 .
By Lemma 2.1, the map φ is well-defined; that is, if g1 H = g2 H, then Hg1−1 = Hg2−1 . To show
that φ is one-to-one, suppose that
3 Lagrange’s Theorem
Proof. We first show that the map φ is one-to-one. Suppose that φ(h1 ) = φ(h2 ) for elements
h1 , h2 ∈ H. We must show that h1 = h2 , but φ(h1 ) = gh1 and φ(h2 ) = gh2 . So gh1 = gh2 , and by
left cancellation h1 = h2 . To show that φ is onto is easy. By definition every element of gH is of
the form gh for some h ∈ H and φ(h) = gh.
Theorem 3.2 (Lagrange). Let G be a finite group and let H be a subgroup of G. Then |G|/|H| =
[G : H] is the number of distinct left cosets of H in G. In particular, the number of elements in H
must divide the number of elements in G.
Proof. The group G is partitioned into [G : H] distinct left cosets. Each left coset has |H| elements;
therefore, |G| = [G : H]|H|.
3
Corollary 3.3. Suppose that G is a finite group and g ∈ G. Then the order of g must divide the
number of elements in G. In particular a|G| = e for every a ∈ G.
Proof. Let H = hgi. Then |H| = o(g) and by Lagrange’s theorem o(g) divides |G|.
Corollary 3.4. Let |G| = p with p a prime number. Then G is cyclic and any g ∈ G such that
g 6= e is a generator.
Proof. Let g be in G such that g 6= e. Then by Corollary 3.3, the order of g must divide the order
of the group. Since |hgi| > 1, it must be p. Hence, g generates G.
Corollary 3.4 suggests that groups of prime order p must somehow look like Zp .
Corollary 3.5. Let H and K be subgroups of a finite group G such that G ⊃ H ⊃ K. Then
[G : K] = [G : H][H : K].
The converse of Lagrange’s Theorem is false. The group A4 has order 12; however, it can be
shown that it does not possess a subgroup of order 6. According to Lagrange’s Theorem, subgroups
of a group of order 12 can have orders of either 1, 2, 3, 4, or 6. However, we are not guaranteed
that subgroups of every possible order exist. To prove that A4 has no subgroup of order 6, we will
assume that it does have a subgroup H such that |H| = 6 and show that a contradiction must
occur. The group A4 contains eight 3-cycles; hence, H must contain a 3-cycle. We will show that
if H contains one 3-cycle, then it must contain every 3-cycle, contradicting the assumption that H
has only 6 elements.
Theorem 3.6. Two cycles τ and µ in Sn have the same length if and only if there exists a σ ∈ Sn
such that µ = στ σ −1 .
4
Then µ = στ σ −1 .
Conversely, suppose that τ = (a1 , a2 , . . . , ak ) is a k-cycle and σ ∈ Sn . If σ(ai ) = b and
σ(a(i mod k)+1 ) = b0 , then µ(b) = b0 . Hence,
Proof. Since [A4 : H] = 2, there are only two cosets of H in A4 . In as much as one of the cosets is
H itself, right and left cosets must coincide; therefore, gH = Hg or gHg −1 = H for every g ∈ A4 .
By above theorem, if H contains one 3-cycle, then it must contain every 3-cycle, contradicting the
order of H.
The Euler φ-function is the map φ : N → N defined by φ(n) = 1 for n = 1, and, for n > 1, φ(n)
is the number of positive integers m with 1 ≤ m < n and gcd(m, n) = 1.
Recall that Zn = {0, 1, 2, · · · , n − 1} is a group with respect to the binary operation addition
modulo n. Note that multiplication modulo n is also a binary operation on Zn and with respect to
this operation we say that a nonzero element a is a unit (or has an inverse) if there exists b ∈ Zn
such that ab ≡ 1 mod n. If n is not a prime, not all nonzero element in Zn is a unit. For example
in Z9 , 4 is a unit as 4.7 ≡ 1 mod 9 whereas 6 is not a unit.
It is easy to check that the set of all units in Zn is a group under multiplication modulo n.
We will denote this group by U (n). Let r ∈ Zn such that r is coprime to n. Then there exists
integers q and t such that qr + tn = 1. Then qr ≡ 1 mod n and hence r is a unit in Zn . So
U (n) = {r ∈ N : r < n and (r, n) = 1} and hence |U (n)| = φ(n).
So we proved the following theorem:
Theorem 4.1. Let U (n) be the group of units in Zn . Then |U (n)| = φ(n).
For example, |U (12)| = φ(12) = 4 since the numbers that are relatively prime to 12 are 1, 5, 7,
and 11. For any prime p, φ(p) = p − 1.
The following theorem is an important result in number theory, due to Leonhard Euler.
Theorem 4.2 (Euler’s Theorem). Let a and n be integers such that n > 0 and gcd(a, n) = 1. Then
aφ(n) ≡ 1 (mod n).
Proof. The order of the group U (n) is φ(n). Consequently, aφ(n) = 1 for all a ∈ U (n); or aφ(n) − 1
is divisible by n. Therefore, aφ(n) ≡ 1 (mod n).
If we consider the special case of Euler’s Theorem in which n = p is prime and recall that
φ(p) = p − 1, we obtain the following result, due to Pierre de Fermat.
5
Theorem 4.3 (Fermat’s Little Theorem). Let p be any prime number and suppose that p - a. Then
6
MTH-204: Abstract Algebra
Lecture-5
If H is a subgroup of a group G, then right cosets are not always the same as left cosets; that is,
it is not always the case that gH = Hg for all g ∈ G. The subgroups for which this property holds
play a critical role in group theory: they allow for the construction of a new class of groups, called
factor or quotient groups.
Normal Subgroups
Example 2. Let H be the subgroup of S3 consisting of elements (1) and (12). Since
and
H(123) = {(123), (23)},
H cannot be a normal subgroup of S3 . However, the subgroup N , consisting of the permutations
(1), (123), and (132), is normal since the cosets of N are
Theorem 1.1. Let G be a group and N be a subgroup of G. Then the following statements are
equivalent.
1
2. For all g ∈ G, gN g −1 ⊂ N .
3. For all g ∈ G, gN g −1 = N .
Proof. (1) ⇒ (2). Since N is normal in G, gN = N g for all g ∈ G. Hence, for a given g ∈ G and
n ∈ N , there exists an n0 in N such that gn = n0 g. Therefore, gng −1 = n0 ∈ N or gN g −1 ⊂ N .
(2) ⇒ (3). Let g ∈ G. Since gN g −1 ⊂ N , we need only show N ⊂ gN g −1 . For n ∈ N ,
g −1 ng= g −1 n(g −1 )−1 ∈ N . Hence, g −1 ng = n0 for some n0 ∈ N . Therefore, n = gn0 g −1 is in
−1
gN g .
(3) ⇒ (1). Suppose that gN g −1 = N for all g ∈ G. Then for any n ∈ N there exists an n0 ∈ N
such that gng −1 = n0 . Consequently, gn = n0 g or gN ⊂ N g. Similarly, N g ⊂ gN .
Factor Groups
If N is a normal subgroup of a group G, then the cosets of N in G form a group G/N under the
operation (aN )(bN ) = abN . This group is called the quotient group of G and N . Our first task is
to prove that G/N is indeed a group.
Theorem 1.2. Let N be a normal subgroup of a group G. The cosets of N in G form a group
G/N of order [G : N ].
Proof. The group operation on G/N is (aN )(bN ) = abN . This operation must be shown to be
well-defined; that is, group multiplication must be independent of the choice of coset representative.
Let aN = bN and cN = dN . We must show that
(aN )(cN ) = acN = bdN = (bN )(dN ).
Then a = bn1 and c = dn2 for some n1 and n2 in N . Hence,
acN = bn1 dn2 N
= bn1 dN
= bn1 N d
= bN d
= bdN.
The remainder of the theorem is easy: eN = N is the identity and g −1 N is the inverse of gN . The
order of G/N is, of course, the number of cosets of N in G.
It is very important to remember that the elements in a factor group are sets of elements in
the original group.
Example 3. Consider the normal subgroup of S3 , N = {(1), (123), (132)}. The cosets of N in S3
are N and (12)N . The factor group S3 /N has the following multiplication table.
N (12)N
N N (12)N
(12)N (12)N N
2
This group is isomorphic to Z2 . At first, multiplying cosets seems both complicated and strange;
however, notice that S3 /N is a smaller group. The factor group displays a certain amount of infor-
mation about S3 . Actually, N = A3 , the group of even permutations, and (12)N = {(12), (13), (23)}
is the set of odd permutations. The information captured in G/N is parity; that is, multiplying
two even or two odd permutations results in an even permutation, whereas multiplying an odd
permutation by an even permutation yields an odd permutation.
0 + 3Z = {. . . , −3, 0, 3, 6, . . .}
1 + 3Z = {. . . , −2, 1, 4, 7, . . .}
2 + 3Z = {. . . , −1, 2, 5, 8, . . .}.
+ 0 + 3Z 1 + 3Z 2 + 3Z
0 + 3Z 0 + 3Z 1 + 3Z 2 + 3Z
1 + 3Z 1 + 3Z 2 + 3Z 0 + 3Z
2 + 3Z 2 + 3Z 0 + 3Z 1 + 3Z
nZ
1 + nZ
2 + nZ
..
.
(n − 1) + nZ.
The sum of the cosets k + Z and l + Z is k + l + Z. Notice that we have written our cosets additively,
because the group operation is integer addition.
Example 5. Consider the dihedral group Dn , generated by the two elements r and s, satisfying
the relations
rn = id
s2 = id
srs = r−1 .
The element r actually generates the cyclic subgroup of rotations, Rn , of Dn . Since srs−1 = srs =
r−1 ∈ Rn , the group of rotations is a normal subgroup of Dn ; therefore, Dn /Rn is a group. Since
there are exactly two elements in this group, it must be isomorphic to Z2 .
Let G be a group and H,K are two subgroups of G. We define the product of H and K by
HK = {hk : h ∈ H, k ∈ K}. Then HK need not be a subgroup of G. For example let G = S3 ,
H = {e, (1, 2)} and K = {e, (2, 3)}. Then HK has exactly 4 elements and by Lagrange’s theorem
HK can not be a subgroup of S3 .
However, if one of them is a normal subgroup then their product is a subgroup.
3
Theorem 1.3. Let H be a subgroup of a group G (not necessarily normal in G) and N a normal
subgroup of G. Then HN is a subgroup of G and H ∩ N is a normal subgroup of H.
Proof. Certainly the set HK has |H||K| symbols. However, not all symbols need represent distinct
group elements. That is, we may have hk = h0 k 0 although h 6= h0 and k 6= k 0 . We must determine
the extent to which this happens.
For every t ∈ H ∩ K, hk = (ht)(t−1 k), so each group element in HK is represented by at least
|H ∩ K| products in HK.
But hk = h0 k 0 implies t = h−1 h0 = k(k 0 )−1 ∈ H ∩ K so that h0 = ht and k 0 = t−1 k. Thus each
element in HK is represented by exactly |H ∩ K| products. So,
|H||K|
|HK| = .
|H ∩ K|
Remark: Alternate Proof: We will see the following outlined proof when we study about
group actions.
The group H × K acts on the set HK ⊆ G via (h, k)x := hxk −1 . The action is transitive. The
stabilizer of 1 ∈ HK is easily seen to be isomorphic to H ∩ K. Then the orbit-stabilizer theorem
implies that |HK| · |H ∩ K| = |H × K| = |H| · |K|.
This proof also works when H, K are infinite.
4
MTH-204: Abstract Algebra
Lecture-6
1 Group Homomorphisms
even odd
even even odd
odd odd even
We use homomorphisms to study relationships such as the one we have just described.
Example: Recall that the circle group T consists of all complex numbers z such that |z| = 1.
We can define a homomorphism φ from the additive group of real numbers R to T by φ : θ → 7
1
cos θ + i sin θ. Indeed,
φ(α + β) = cos(α + β) + i sin(α + β)
= (cos α cos β − sin α sin β) + i(sin α cos β + cos α sin β)
= (cos α + i sin α) + (cos β + i sin β)
= φ(α)φ(β).
Geometrically, we are simply wrapping the real line around the circle in a group-theoretic fashion.
Example: The map sign : Sn → {1, −1} is a homomorphism, as sign(σ.τ ) = sign(σ).sign(τ ).
Many groups may appear to be different at first glance, but can be shown to be the same by a
simple renaming of the group elements. For example, Z4 and the subgroup of the circle group T
generated by i can be shown to be the same by demonstrating a one-to-one correspondence between
the elements of the two groups and between the group operations. In such a case we say that the
groups are isomorphic.
Example: To show that Z4 ∼ = hii, define a map φ : Z4 → hii by φ(n) = in . We must show that φ
is bijective and preserves the group operation. The map φ is one-to-one and onto because
φ(0) = 1
φ(1) = i
φ(2) = −1
φ(3) = −i.
Since
φ(m + n) = im+n = im in = φ(m)φ(n),
the group operation is preserved.
Example: log : R+ → R where R+ denotes the positive real numbers with the operation mul-
tiplication, is an isomorphism since from calculus we know that log is one to one and onto and
log(xy) = logx + logy for all positive real numbers x and y.
Example: We can define an isomorphism φ from the additive group of real numbers (R, +) to the
multiplicative group of positive real numbers (R+ , ·) with the exponential map; that is,
φ(x + y) = ex+y = ex ey = φ(x)φ(y).
Of course, we must still show that φ is one-to-one and onto, but this can be determined using
calculus.
Example: The integers are isomorphic to the subgroup of Q∗ consisting of elements of the form
2n . Define a map φ : Z → Q∗ by φ(n) = 2n . Then
φ(m + n) = 2m+n = 2m 2n = φ(m)φ(n).
By definition the map φ is onto the subset {2n : n ∈ Z} of Q∗ . To show that the map is injective,
assume that m 6= n. If we can show that φ(m) 6= φ(n), then we are done. Suppose that m > n and
assume that φ(m) = φ(n). Then 2m = 2n or 2m−n = 1, which is impossible since m − n > 0.
2
Example: The groups Z8 and Z12 cannot be isomorphic since they have different orders; however,
it is true that U (8) ∼
= U (12). We know that
U (8) = {1, 3, 5, 7}
U (12) = {1, 5, 7, 11}.
An isomorphism φ : U (8) → U (12) is then given by
1 7→ 1
3 7→ 5
5 7→ 7
7 7→ 11.
The map φ is not the only possible isomorphism between these two groups. We could define another
isomorphism ψ by ψ(1) = 1, ψ(3) = 11, ψ(5) = 5, ψ(7) = 7. In fact, both of these groups are
isomorphic to Z2 × Z2
Example: Even though S3 and Z6 possess the same number of elements, we would suspect that
they are not isomorphic, because Z6 is abelian and S3 is nonabelian. To demonstrate that this is
indeed the case, suppose that φ : Z6 → S3 is an isomorphism. Let a, b ∈ S3 be two elements such
that ab 6= ba. Since φ is an isomorphism, there exist elements m and n in Z6 such that
φ(m) = a
φ(n) = b.
However,
ab = φ(m)φ(n) = φ(m + n) = φ(n + m) = φ(n)φ(m) = ba,
which contradicts the fact that a and b do not commute.
Theorem 1.1. Let φ : G → H be an isomorphism of two groups. Then the following statements
are true.
1. φ−1 : H → G is an isomorphism.
2. |G| = |H|.
3. If G is abelian, then H is abelian.
4. If G is cyclic, then H is cyclic.
5. If G has a subgroup of order n, then H has a subgroup of order n.
Proof. Assertions (1) and (2) follow from the fact that φ is a bijection. We will prove (3) here and
proofs of the others are similar.
(3) Suppose that h1 and h2 are elements of H. Since φ is onto, there exist elements g1 , g2 ∈ G
such that φ(g1 ) = h1 and φ(g2 ) = h2 . Therefore,
h1 h2 = φ(g1 )φ(g2 ) = φ(g1 g2 ) = φ(g2 g1 ) = φ(g2 )φ(g1 ) = h2 h1 .
3
The following proposition lists some basic properties of group homomorphisms.
Proof. (1) Suppose that e and e0 are the identities of G1 and G2 , respectively; then
By cancellation, φ(e) = e0 .
(2) This statement follows from the fact that
(3) The set φ(H1 ) is nonempty since the identity of H2 is in φ(H1 ). Suppose that H1 is a
subgroup of G1 and let x and y be in φ(H1 ). There exist elements a, b ∈ H1 such that φ(a) = x
and φ(b) = y. Since
xy −1 = φ(a)[φ(b)]−1 = φ(ab−1 ) ∈ φ(H1 ),
φ(H1 ) is a subgroup of G2 .
(4) Let H2 be a subgroup of G2 and define H1 to be φ−1 (H2 ); that is, H1 is the set of all
g ∈ G1 such that φ(g) ∈ H2 . The identity is in H1 since φ(e) = e. If a and b are in H1 , then
φ(ab−1 ) = φ(a)[φ(b)]−1 is in H2 since H2 is a subgroup of G2 . Therefore, ab−1 ∈ H1 and H1 is a
subgroup of G1 . If H2 is normal in G2 , we must show that g −1 hg ∈ H1 for h ∈ H1 and g ∈ G1 .
But
φ(g −1 hg) = [φ(g)]−1 φ(h)φ(g) ∈ H2 ,
since H2 is a normal subgroup of G2 . Therefore, g −1 hg ∈ H1 .
4
Example 9. Let us examine the homomorphism φ : GL2 (R) → R∗ defined by A 7→ det(A). Since
1 is the identity of R∗ , the kernel of this homomorphism is all 2 × 2 matrices having determinant
one. That is, ker φ = SL2 (R).
Example 10. The kernel of the group homomorphism φ : R → C∗ defined by φ(θ) = cos θ + i sin θ
is {2πn : n ∈ Z}. Notice that ker φ ∼
= Z.
Example 11. Suppose that we wish to determine all possible homomorphisms φ from Z7 to
Z12 . Since the kernel of φ must be a subgroup of Z7 , there are only two possible kernels, {0}
and all of Z7 . The image of a subgroup of Z7 must be a subgroup of Z12 . Hence, there is no
injective homomorphism; otherwise, Z12 would have a subgroup of order 7, which is impossible.
Consequently, the only possible homomorphism from Z7 to Z12 is the one mapping all elements to
zero.
Example 12. Let G be a group. Suppose that g ∈ G and φ is the homomorphism from Z to G
given by φ(n) = g n . If the order of g is infinite, then the kernel of this homomorphism is {0} since
φ maps Z onto the cyclic subgroup of G generated by g. However, if the order of g is finite, say n,
then the kernel of φ is nZ.
5
MTH-204: Abstract Algebra
Lecture-7
1 Isomorphism Theorems
The main goal in group theory is to classify all groups; however, it makes sense to consider two
groups to be the same if they are isomorphic. For two groups G and H, we say G is related to H
if G and H are isomorphic. The proof of the following theorem is easy.
Theorem 1.1. The isomorphism of groups determines an equivalence relation on the class of all
groups.
Hence, we can modify our goal of classifying all groups to classifying all groups up to isomor-
phism; that is, we will consider two groups to be the same if they are isomorphic.
Though at first it is not evident that factor groups correspond exactly to homomorphic images,
we can use factor groups to study homomorphisms. We already know that with every group
homomorphism φ : G → H we can associate a normal subgroup of G, ker φ; the converse is also
true. Every normal subgroup of a group G gives rise to homomorphism of groups.
Let H be a normal subgroup of G. Define the canonical map
φ : G → G/H
by
φ(g) = gH.
This is indeed a homomorphism, since
φ(g1 g2 ) = g1 g2 H = g1 Hg2 H = φ(g1 )φ(g2 ).
The kernel of this homomorphism is H. The following theorems describe the relationships among
group homomorphisms, normal subgroups, and factor groups.
Theorem 1.2 (First Isomorphism Theorem). If ψ : G → H is a group homomorphism with
K = ker ψ, then K is normal in G. Let φ : G → G/K be the canonical homomorphism. Then there
exists a unique isomorphism η : G/K → ψ(G) such that ψ = ηφ.
Proof. We already know that K is normal in G. Define η : G/K → ψ(G) by η(gK) = ψ(g). We
must first show that this is a well-defined map. Suppose that g1 K = g2 K. For some k ∈ K,
g1 k = g2 ; consequently,
η(g1 K) = ψ(g1 ) = ψ(g1 )ψ(k) = ψ(g1 k) = ψ(g2 ) = η(g2 K).
1
Since η(g1 K) = η(g2 K), η does not depend on the choice of coset representative. Clearly η is onto
ψ(G). To show that η is one-to-one, suppose that η(g1 K) = η(g2 K). Then ψ(g1 ) = ψ(g2 ). This
implies that ψ(g1−1 g2 ) = e, or g1−1 g2 is in the kernel of ψ; hence, g1−1 g2 K = K; that is, g1 K = g2 K.
Finally, we must show that η is a homomorphism, but
Mathematicians often use diagrams called commutative diagrams to describe such theorems.
The following diagram “commutes” since ψ = ηφ.
ψ
G -H
J
φJ η
J^
J
G/K
Example 13. Let G be a cyclic group with generator g. Define a map φ : Z → G by n 7→ g n . This
map is a surjective homomorphism since
Clearly φ is onto. If |g| = m, then g m = e. Hence, ker φ = mZ and Z/ ker φ = Z/mZ ∼ = G. On the
other hand, if the order of g is infinite, then ker φ = 0 and φ is an isomorphism of G and Z. Hence,
two cyclic groups are isomorphic exactly when they have the same order. Up to isomorphism, the
only cyclic groups are Z and Zn . In particular, if G is a group of order p, where p is a prime
number, then G is isomorphic to Zp .
Theorem 1.3 (Second Isomorphism Theorem). Let H be a subgroup of a group G (not necessarily
normal in G) and N a normal subgroup of G. Then
H/H ∩ N ∼
= HN/N.
By the First Isomorphism Theorem, the image of φ is isomorphic to H/ ker φ; that is,
HN/N = φ(H) ∼
= H/ ker φ.
2
Since
ker φ = {h ∈ H : h ∈ N } = H ∩ N,
HN/N = φ(H) ∼
= H/H ∩ N .
Proof. Let H be a subgroup of G containing N . Since N is normal in H, H/N makes sense. Let
aN and bN be elements of H/N . Then (aN )(b−1 N ) = ab−1 N ∈ H/N ; hence, H/N is a subgroup
of G/N .
Let S be a subgroup of G/N . This subgroup is a set of cosets of N . If H = {g ∈ G : gN ∈ S},
then for h1 , h2 ∈ H, we have that (h1 N )(h2 N ) = hh0 N ∈ S and h−1
1 N ∈ S. Therefore, H must be
a subgroup of G. Clearly, H contains N . Therefore, S = H/N . Consequently, the map H 7→ H/H
is onto.
Suppose that H1 and H2 are subgroups of G containing N such that H1 /N = H2 /N . If h1 ∈ H1 ,
then h1 N ∈ H1 /N . Hence, h1 N = h2 N ⊂ H2 for some h2 in H2 . However, since N is contained in
H2 , we know that h1 ∈ H2 or H1 ⊂ H2 . Similarly, H2 ⊂ H1 . Since H1 = H2 , the map H 7→ H/H
is one-to-one.
Suppose that H is normal in G and N is a subgroup of H. Then it is easy to verify that the
map G/N → G/H defined by gN 7→ gH is a homomorphism. The kernel of this homomorphism is
H/N , which proves that H/N is normal in G/N .
Conversely, suppose that H/N is normal in G/N . The homomorphism given by
G/N
G → G/N →
H/N
Notice that in the course of the proof of the above theorem, we have also proved the following
theorem.
Theorem 1.5 (Third Isomorphism Theorem). Let G be a group and N and H be normal subgroups
of G with N ⊂ H. Then
G/N
G/H ∼= .
H/N
Z/mZ ∼
= (Z/mnZ)/(mZ/mnZ).
3
Cayley’s Theorem
Cayley proved that if G is a group, it is isomorphic to a group of permutations on some set; hence,
every group is a permutation group. Cayley’s Theorem is what we call a representation theorem.
The aim of representation theory is to find an isomorphism of some group G that we wish to study
into a group that we know a great deal about, such as a group of permutations or matrices.
+ 0 1 2
0 0 1 2
1 1 2 0
2 2 0 1
The addition table of Z3 suggests that it is the same as the permutation group G = {(0), (012), (021)}.
The isomorphism here is
0 1 2
0 7→ = (0)
0 1 2
0 1 2
1 7→ = (012)
1 2 0
0 1 2
2 7→ = (021).
2 0 1
Proof. Let G be a group. We must find a group of permutations G that is isomorphic to G. For
any g ∈ G, define a function λg : G → G by λg (a) = ga. We claim that λg is a permutation of G.
To show that λg is one-to-one, suppose that λg (a) = λg (b). Then
Hence, a = b. To show that λg is onto, we must prove that for each a ∈ G, there is a b such that
λg (b) = a. Let b = g −1 a.
Now we are ready to define our group G. Let
G = {λg : g ∈ G}.
We must show that G is a group under composition of functions and find an isomorphism between
G and G. We have closure under composition of functions since
Also,
λe (a) = ea = a
4
and
(λg−1 ◦ λg )(a) = λg−1 (ga) = g −1 ga = a = λe (a).
ga = λg a = λh a = ha.
Hence, g = h. That φ is onto follows from the fact that φ(g) = λg for any λg ∈ G.
5
MTH-204: Abstract Algebra
Lecture-8
1 Direct Products
Given two groups G and H, it is possible to construct a new group from the Cartesian product of
G and H, G × H. Conversely, given a large group, it is sometimes possible to decompose the group;
that is, a group is sometimes isomorphic to the direct product of two smaller groups. Rather than
studying a large group G, it is often easier to study the component groups of G.
If (G, ·) and (H, ◦) are groups, then we can make the Cartesian product of G and H into a new
group. As a set, our group is just the ordered pairs (g, h) ∈ G × H where g ∈ G and h ∈ H. We
can define a binary operation on G × H by
that is, we just multiply elements in the first coordinate as we do in G and elements in the second
coordinate as we do in H. We have specified the particular operations · and ◦ in each group here
for the sake of clarity; we usually just write (g1 , h1 )(g2 , h2 ) = (g1 g2 , h1 h2 ).
Proposition 1.1. Let G and H be groups. The set G × H is a group under the operation
(g1 , h1 )(g2 , h2 ) = (g1 g2 , h1 h2 ) where g1 , g2 ∈ G and h1 , h2 ∈ H.
Proof. Clearly the binary operation defined above is closed. If eG and eH are the identities of the
groups G and H respectively, then (eG , eH ) is the identity of G × H. The inverse of (g, h) ∈ G × H
is (g −1 , h−1 ). The fact that the operation is associative follows directly from the associativity of G
and H.
Example 7. Let R be the group of real numbers under addition. The Cartesian product of
R with itself, R × R = R2 , is also a group, in which the group operation is just addition in each
coordinate; that is, (a, b) + (c, d) = (a + c, b + d). The identity is (0, 0) and the inverse of (a, b) is
(−a, −b).
Example Consider
Z2 × Z2 = {(0, 0), (0, 1), (1, 0), (1, 1)}.
1
Although Z2 × Z2 and Z4 both contain four elements, it is easy to see that they are not isomorphic
since for every element (a, b) in Z2 × Z2 , (a, b) + (a, b) = (0, 0), but Z4 is cyclic.
The group G × H is called the external direct product of G and H. Notice that there is
nothing special about the fact that we have used only two groups to build a new group. The direct
product
Yn
Gi = G1 × G2 × · · · × Gn
i=1
of the groups G1 , G2 , . . . , Gn is defined in exactly the same manner. If G = G1 = G2 = · · · = Gn ,
we often write Gn instead of G1 × G2 × · · · × Gn .
Example The group Zn2 , considered as a set, is just the set of all binary n-tuples. The group
operation is the “exclusive or” of two binary n-tuples. For example,
This group is important in coding theory, in cryptography, and in many areas of computer science.
Theorem 1.2. Let (g, h) ∈ G × H. If g and h have finite orders r and s respectively, then the
order of (g, h) in G × H is the least common multiple of r and s.
Proof. Suppose that m is the least common multiple of r and s and let n = |(g, h)|. Then
Hence, n must divide m, and n ≤ m. However, by the second equation, both r and s must divide
n; therefore, n is a common multiple of r and s. Since m is the least common multiple of r and s,
m ≤ n. Consequently, m must be equal to n.
Q
Corollary
Q 1.3. Let (g1 , . . . , g n ) ∈ Gi . If gi has finite order ri in Gi , then the order of (g1 , . . . , gn )
in Gi is the least common multiple of r1 , . . . , rn .
Example Let (8, 56) ∈ Z12 × Z60 . Since gcd(8, 12) = 4, the order of 8 is 12/4 = 3 in Z12 . Similarly,
the order of 56 in Z60 is 15. The least common multiple of 3 and 15 is 15; hence, (8, 56) has order
15 in Z12 × Z60 .
(0, 0), (0, 1), (0, 2), (1, 0), (1, 1), (1, 2).
In this case, unlike that of Z2 × Z2 and Z4 , it is true that Z2 × Z3 ∼ = Z6 . We need only show that
Z2 × Z3 is cyclic. It is easy to see that (1, 1) is a generator for Z2 × Z3 .
The next theorem tells us exactly when the direct product of two cyclic groups is cyclic.
Theorem 1.4. The group Zm × Zn is isomorphic to Zmn if and only if gcd(m, n) = 1.
2
Proof. Assume first that if Zm × Zn ∼= Zmn . We need to show gcd(m, n) = 1. To show this, we will
prove the contrapositive; that is, we will show that if gcd(m, n) = d > 1, then Zm × Zn cannot be
cyclic. Notice that mn/d is divisible by both m and n; hence, for any element (a, b) ∈ Zm × Zn ,
Corollary 1.6. If
m = pe11 · · · pekk ,
where the pi s are distinct primes, then
Zm ∼
= Zpe11 × · · · × Zpek .
k
e
Proof. Since the greatest common divisor of pei i and pj j is 1 for i 6= j, the proof follows from above
corollary.
Later we will prove that all finite abelian groups are isomorphic to direct products of the form
Zpe1 × · · · × Zpek
1 k
The external direct product of two groups builds a large group out of two smaller groups. We
would like to be able to reverse this process and conveniently break down a group into its direct
product components; that is, we would like to be able to say when a group is isomorphic to the
direct product of two of its subgroups.
Let G be a group with subgroups H and K satisfying the following conditions.
• G = HK = {hk : h ∈ H, k ∈ K};
• H ∩ K = {e};
3
• hk = kh for all k ∈ K and h ∈ H.
Example 13. The dihedral group D6 is an internal direct product of its two subgroups
H = {id, r3 }
K = {id, r2 , r4 , s, r2 s, r4 s}.
= S3 ; consequently, D6 ∼
It can easily be shown that K ∼ = Z2 × S3 .
Example 14. Not every group can be written as the internal direct product of two of its proper
subgroups. If the group S3 were an internal direct product of its proper subgroups H and K,
then one of the subgroups, say H, would have to have order 3. In this case H is the subgroup
{(1), (123), (132)}. The subgroup K must have order 2, but no matter which subgroup we choose
for K, the condition that hk = kh will never be satisfied for h ∈ H and k ∈ K.
Theorem 1.7. Let G be the internal direct product of subgroups H and K. Then G is isomorphic
to H × K.
Proof. Since G is an internal direct product, we can write any element g ∈ G as g = hk for some
h ∈ H and some k ∈ K. Define a map φ : G → H × K by φ(g) = (h, k).
The first problem that we must face is to show that φ is a well-defined map; that is, we must show
that h and k are uniquely determined by g. Suppose that g = hk = h0 k 0 . Then h−1 h0 = k(k 0 )−1 is
in both H and K, so it must be the identity. Therefore, h = h0 and k = k 0 , which proves that φ is,
indeed, well-defined.
To show that φ preserves the group operation, let g1 = h1 k1 and g2 = h2 k2 and observe that
φ(g1 g2 ) = φ(h1 k1 h2 k2 )
= φ(h1 h2 k1 k2 )
= (h1 h2 , k1 k2 )
= (h1 , k1 )(h2 , k2 )
= φ(g1 )φ(g2 ).
We will leave the proof that φ is one-to-one and onto as an exercise.
Example 15. The group Z6 is an internal direct product isomorphic to {0, 2, 4} × {0, 3}.
4
• G = H1 H2 · · · Hn = {h1 h2 · · · hn : hi ∈ Hi };
• Hi ∩ h∪j6=i Hj i = {e};
The following theorem is just a generalization of the above theorem and the proof is the same.
Theorem 1.8. LetQ G be the internal direct product of subgroups Hi , where i = 1, 2, . . . , n. Then
G is isomorphic to i Hi .
5
MTH-204: Abstract Algebra
Lecture-9
Proof. It is easy to check that Inn(G) is a subgroup of Aut(G). Let φ ∈ Aut(G) and ig ∈ Inn(G).
Then we must show that φ.ig .φ−1 ∈ Inn(G).
We claim that φ.ig .φ−1 = iφ(g) . This is because φ.ig .φ−1 (x) = φ.ig (φ−1 (x)) = φ(gφ−1 (x)g −1 ) =
φ(g)xφ(g)−1 = iφ(g) (x).
1
Remark: From the above theorem we have G ∼ = Inn(G) if Z(G) = {e}. For example if G = Sn
∼
then since Z(Sn ) = {e} we have Sn = Inn(Sn ).
Definition 1.3. Suppose that H and K are groups and let φ : K → Aut(H) be a homomorphism.
We may define an action of K on H as k · h = φ(k)(h). Then we define the external semi-direct
product H oφ K of H and K with respect to φ as follows. As a set H oφ K = H × K. The group
operation of H oφ K is defined by
3. H E H oφ K.
4. H ∩ K = 1.
5. HK = H oφ K.
The following proposition says that the direct product is same as the semi-direct product if φ
is the trivial homomorphsim.
Proposition 1.5. Let H and K be groups and let φ : K → AutH be a homomorphism. The
following are equivalent.
1. The identity map between H o K and H × K is a group homomorphism (and hence isomor-
phism).
3. K E H o K.
Like in the case of direct product here also we can define internal semi-direct product of two
subgroups H and K of G. The following theorem says that for two subgroups, their internal
semi-direct product is same as their external semi-direct product. The proof is the same.
Theorem 1.6. Suppose that G is a group and H, K ≤ G such that
1. H E G, and
2. H ∩ K = 1.
2
Let φ : K → Aut(H) be the homomorphism defined by φ(k)(h) = khk −1 . Then, HK ∼ = H o K. In
particular, if G = HK with H and K satisfying (1) and (2) above then G is the semidirect product
of H and K.
3
MTH-204: Abstract Algebra
Lecture-10
1 Group Actions
Group actions generalize group multiplication. If G is a group and X is an arbitrary set, a group
action of an element g ∈ G and x ∈ X is a product, gx, living in X. Many problems in algebra
may best be attacked via group actions. For example, the proofs of the Sylow theorems and of
Burnside’s Counting Theorem are most easily understood when they are formulated in terms of
group actions.
1. ex = x for all x ∈ X;
Under these considerations X is called a G-set. Notice that we are not requiring X to be related
to G in any way. It is true that every group G acts on every set X by the trivial action (g, x) 7→ x;
however, group actions are more interesting if the set X is somehow related to the group G.
Example 1. Let G = GL2 (R) and X = R2 . Then G acts on X by left multiplication. If v ∈ R2 and
I is the identity matrix, then Iv = v. If A and B are 2×2 invertible matrices, then (AB)v = A(Bv)
since matrix multiplication is associative.
Example 2. Let G = D4 , the symmetry group of a square. If X = {1, 2, 3, 4} is the set of vertices
of the square, then we can consider D4 to consist of the following permutations:
The elements of D4 act on X as functions. The permutation (13)(24) acts on vertex 1 by sending
it to vertex 3, on vertex 2 by sending it to vertex 4, and so on. It is easy to see that the axioms of
a group action are satisfied.
1
In general, if X is any set and G is a subgroup of SX , the group of all permutations acting on
X, then X is a G-set under the group action
(σ, x) 7→ σ(x)
for σ ∈ G and x ∈ X.
Example 3. If we let X = G, then every group G acts on itself by the left regular representation;
that is, (g, x) 7→ λg (x) = gx, where λg is left multiplication:
e · x = λe x = ex = x
(gh) · x = λgh x = λg λh x = λg (hx) = g · (h · x).
H × G → G,
via
(h, g) 7→ hgh−1
for h ∈ H and g ∈ G. Clearly, the first axiom for a group action holds. Observing that
Example 5. Let H be a subgroup of G and LH the set of left cosets of H. The set LH is a G-set
under the action
(g, xH) 7→ gxH.
Again, it is easy to see that the first axiom is true. Since (gg 0 )xH = g(g 0 xH), the second axiom is
also true.
Proof. The relation ∼ is reflexive since ex = x. Suppose that x ∼ y for x, y ∈ X. Then there exists
a g such that gx = y. In this case g −1 y = x; hence, y ∼ x. To show that the relation is transitive,
suppose that x ∼ y and y ∼ z. Then there must exist group elements g and h such that gx = y
and hy = z. So z = hy = (hg)x, and x is equivalent to z.
2
If X is a G-set, then each partition of X associated with G-equivalence is called an orbit of X
under G. We will denote the orbit that contains an element x of X by Ox . So
Ox = {g.x : g ∈ G}.
Now suppose that G is a group acting on a set X and let g be an element of G. The fixed
point set of g in X, denoted by Xg , is the set of all x ∈ X such that gx = x. We can also study
the group elements g that fix a given x ∈ X. This set is more than a subset of G, it is a subgroup.
This subgroup is called the stabilizer subgroup or isotropy subgroup of x. We will denote the
stabilizer subgroup of x by Gx . Note that
Gx = {g ∈ G : g.x = x}.
Example 7. Let X = {1, 2, 3, 4, 5, 6} and suppose that G is the permutation group given by the
permutations
{(1), (12)(3456), (35)(46), (12)(3654)}.
Then the fixed point sets of X under the action of G are
X(1) = X,
X(35)(46) = {1, 2},
X(12)(3456) = X(12)(3654) = ∅,
and the stabilizer subgroups are
G1 = G2 = {(1), (35)(46)},
G3 = G4 = G5 = G6 = {(1)}.
It is easily seen that Gx is a subgroup of G for each x ∈ X.
Proposition 2.2. Let G be a group acting on a set X and x ∈ X. The stabilizer, Gx , of x is a
subgroup of G.
Proof. Clearly, e ∈ Gx since the identity fixes every element in the set X. Let g, h ∈ Gx . Then
gx = x and hx = x. So (gh)x = g(hx) = gx = x; hence, the product of two elements in Gx is also
in Gx . Finally, if g ∈ Gx , then x = ex = (g −1 g)x = (g −1 )gx = g −1 x. So g −1 is in Gx .
We will denote the number of elements in the fixed point set of an element g ∈ G by |Xg | and
denote the number of elements in the orbit of x of x ∈ X by |Ox |. The next theorem demonstrates
the relationship between orbits of an element x ∈ X and the left cosets of Gx in G.
3
Theorem 2.3 (Orbit-Stabilizer Theorem). Let G be a finite group and X a finite G-set. If x ∈ X,
then |Ox | = [G : Gx ].
Proof. We know that |G|/|Gx | is the number of left cosets of Gx in G by Lagrange’s Theorem. We
will define a bijective map φ between the orbit Ox of X and the set of left cosets LGx of Gx in
G. Let y ∈ Ox . Then there exists a g in G such that gx = y. Define φ by φ(y) = gGx . First we
must show that this map is well-defined and does not depend on our selection of g. Suppose that
h is another element in G such that hx = y. Then gx = hx or x = g −1 hx; hence, g −1 h is in the
stabilizer subgroup of x. Therefore, h ∈ gGx or gGx = hGx . Thus, y gets mapped to the same
coset regardless of the choice of the representative from that coset.
To show that φ is one-to-one, assume that φ(x1 ) = φ(x2 ). Then there exist g1 , g2 ∈ G such that
x1 = g1 x and x2 = g2 x. Since there exists a g ∈ Gx such that g2 = g1 g,
x2 = g2 x = g1 gx = g1 x = x1 ;
consequently, the map φ is one-to-one. Finally, we must show that the map φ is onto. Let gGx be
a left coset. If gx = y, then φ(y) = gGx .
Let X be a finite G-set and XG be the set of fixed points in X; that is,
|G| = |Z(G)| + n1 + · · · + nk .
The stabilizer subgroups of each of the xi ’s, C(xi ) = {g ∈ G : gxi = xi g}, are called the centralizer
subgroups of the xi ’s. From Theorem 12.3, we obtain the class equation:
One of the consequences of the class equation is that the order of each conjugacy class must divide
the order of |G|.
4
Example 8. It is easy to check that the conjugacy classes in S3 are the following:
Example 10. For Sn it takes a bit of work to find the conjugacy classes. We begin with cycles.
Suppose that σ = (a1 , . . . , ak ) is a cycle and let τ ∈ Sn . By Theorem 5.9,
Consequently, any two cycles of the same length are conjugate. Now let σ = σ1 σ2 · · · σr be a cycle
decomposition, where the length of each cycle σi is ri . Then σ is conjugate to every other τ ∈ Sn
whose cycle decomposition has the same lengths.
The number of conjugate classes in Sn is the number of ways in which n can be partitioned
into sums of positive integers. For example, we can partition the integer 3 into the following three
sums:
3 = 1+1+1
3 = 1+2
3 = 3;
Theorem 3.1. Let G be a group of order pn where p is prime. Then G has a nontrivial center.
|G| = |Z(G)| + n1 + · · · + nk .
Since each ni > 1 and ni | G, p must divide each ni . Also, p | |G|; hence, p must divide |Z(G)|.
Since the identity is always in the center of G, |Z(G)| ≥ 1. Therefore, |Z(G)| ≥ p and there exists
some g ∈ Z(G) such that g 6= 1.
Proof. By above theorem we have |Z(G)| = p or p2 . If |Z(G)| = p2 , then we are done. Suppose that
|Z(G)| = p. Then Z(G) and G/Z(G) both have order p and must both be cyclic groups. Choosing
a generator aZ(G) for G/Z(G), we can write any element gZ(G) in the quotient group as am Z(G)
for some integer m; hence, g = am x for some x in the center of G. Similarly, if hZ(G) ∈ G/Z(G),
5
there exists a y in Z(G) such that h = an y for some integer n. Since x and y are in the center of
G, they commute with all other elements of G; therefore,
6
MTH-204: Abstract Algebra
Lecture-11
Suppose that we are to color the vertices of a square with two different colors, say black and white.
We might suspect that there would be 24 = 16 different colorings. However, some of these colorings
are equivalent. If we color the first vertex black and the remaining vertices white, it is the same
as coloring the second vertex black and the remaining ones white since we could obtain the second
coloring simply by rotating the square 90◦ .
B W W B
W W W W
W W W W
B W W B
Proof. Let G act on X by (g, x) 7→ g · x. Since x ∼ y, there exists a g ∈ G such that g · x = y. Let
a ∈ Gx . Since
gag −1 · y = ga · g −1 y = ga · x = g · x = y,
1
we can define a map φ : Gx → Gy by φ(a) = gag −1 . The map φ is a homomorphism since
Suppose that φ(a) = φ(b). Then gag −1 = gbg −1 or a = b; hence, the map is injective. To show
that φ is onto, let b be in Gy ; then g −1 bg is in Gx since
g −1 bg · x = g −1 b · gx = g −1 b · y = g −1 · y = x;
Theorem 1.2 (Burnside). Let G be a finite group acting on a set X and let k denote the number
of orbits of X. Then
1 X
k= |Xg |.
|G|
g∈G
Proof. We look at all the fixed points x of all the elements in g ∈ G; that is, we look at all g’s and
all x’s such that gx = x. If viewed in terms of fixed point sets, the number of all g’s fixing x’s is
X
|Xg |.
g∈G
That is X X
|Xg | = |{(g, x) ∈ G × X : g.x = x}| = |Gx |.
g∈G x∈X
Example 11. Let X = {1, 2, 3, 4, 5} and suppose that G is the permutation group G = {(1), (13), (13)(25), (25)}.
The orbits of X are {1, 3}, {2, 5}, and {4}. The fixed point sets are
X(1) = X
X(13) = {2, 4, 5}
X(13)(25) = {4}
X(25) = {1, 3, 4}.
2
2 The Sylow Theorems
We already know that the converse of Lagrange’s Theorem is false. If G is a group of order m and
n divides m, then G does not necessarily possess a subgroup of order n. For example, A4 has order
12 but does not possess a subgroup of order 6. However, the Sylow Theorems do provide a partial
converse for Lagrange’s Theorem: in certain cases they guarantee us subgroups of specific orders.
These theorems yield a powerful set of tools for the classification of all finite nonabelian groups.
We will use the idea of group actions to prove the Sylow Theorems. A group G acts on itself
by conjugation via the map (g, x) 7→ gxg −1 . Let x1 , . . . , xk be representatives from each of the
distinct conjugacy classes of G that consist of more than one element. Then the class equation can
be written as
|G| = |Z(G)| + [G : C(x1 )] + · · · + [G : C(xk )],
where Z(G) = {g ∈ G : gx = xg for all x ∈ G} is the center of G and C(xi ) = {g ∈ G : gxi = xi g}
is the centralizer subgroup of xi .
Let p be a prime. A group G called a p-group if |G| is a power of p.
Theorem 2.1 (Sylow Theorems). Let G be a group, |G| = pm r, where p is a prime and gcd(r, p) =
1. Then
3
By Orbit-Stabilizer Theorem, we have |OS | = |G|/|P | = pm r/|P |. By choice of the S we picked,
p does not divide |OS |, that is p does not divide pm r/|P | and |P | has to be a multiple of pm , or
equivalently pm divides |P |. So |P | ≥ pm .
We claim that |P | ≤ pm . Let us define the map λx , x ∈ S, by λx : P → S, g 7→ gx. In words,
this map goes from P , which is a subgroup of G, to S, which is an element of X, that is a subset
of G with cardinality pm . Note that this map is well-defined since gx ∈ S for any x ∈ S and any
g ∈ P by definition of P being the stabilizer of S. It is also clearly injective (gx = hx implies g = h
since x is an element of the group G and thus is invertible). If we have an injection from P to S,
that means |P | ≤ |S| = pm . We are done.
Proof of (2): Let P be a Sylow p-subgroup of G and let R be a p-group of G. We will prove
that R (being a p-group in general) is contained in a conjugate of P . Let R act by multiplication
on the set Y of left cosets of P : Y = {gP, g ∈ G}.
We want to prove that there is an orbit of size 1 under this action. By Lagrange’s Theorem,
we know that |Y | = |G|/|P | = pm r/pm = r and thus p does not divide |Y | by assumption on r.
P
We have a partition of Y by its orbits, we get |Y | = |Oy | and there exists one orbit Oy such
that p - |Oy |.
By the Orbit-Stabilizer Theorem, we have |Oy | divides |R|, which has order a power of p, so
there is an orbit of size 1. Let gP ∈ Y be the element whose orbit size is 1.
We have hgP = gP for h ∈ R, since gP belongs to its orbit. Thus g −1 hg ∈ P iff h ∈ gP g −1 for
all h ∈ R.
We have just proved that the p-group R is contained in a conjugate of P . All we needed for the
proof is that R is a p-group, so the same proof holds for the case of a Sylow p-subgroup, for which
we get that R is contained in a conjugate of P , and both have same cardinality, which concludes
the proof. We will use the fact that the proof works for R a p-group in general for proving one
corollary.
Proof of (3): Consider the set X be the set of all Sylow p- subgroups of G. We have |X| = np .
By the 1st Sylow Theorem, this set is non-empty and there exists at least one Sylow p-subgroup P
in X, whose order is pm .
Let P act on X by conjugation, i.e., gQ = gQg −1 , g ∈ P , Q ∈ X. Note that in the case where
P is the only Sylow p-subgroup, then we can take Q = P . By the Orbit-Stabilizer Theorem, we
have |OQ | = |P |/|GQ | = pm /|GQ |. So |OQ | = 1 or a power of p.
|OQ0 + |OQ” |, where Q0 and Q” denote subgroups whose
P P P
Again we have |X| = |OQ | =
orbit has respectively one element or at least two elements. Since p divides the second sum, we
have |X| ≡ number of orbits of size 1 mod p. To conclude the proof, we thus have to show that
there is only one Sylow p-subgroup whose orbit has size 1, namely P itself
Let us assume there is another Sylow p-subgroup Q whose orbit has only one element, namely:
gQg −1 = Q, g ∈ P , which translates into gQ = Qg for all g ∈ P and so P Q = QP . This says that
P Q is a subgroup of G.
We know that |P Q| = |P ||Q|/|P ∩ Q = pm pm /|P ∩ Q|. So |P Q| is a power of p, say pc for some
c which cannot be bigger than m, since |G| = pm r. Hence pm = |P | ≤ |P Q| ≤ pm . So |P | = |P Q|
4
and so P = Q as they have same cardinality.
Corollary 2.2. (Cauchy) Let G be a finite group and p a prime such that p divides the order of
G. Then G contains a subgroup of order p.
Proof. Let P be a Sylow p-subgroup of G (which exists by the 1st Sylow Theorem), and pick x 6= 1
k−1
in P . The order |x| of x is a power of p by definition of a p-group, say |x| = pk . Then xp has
order p.
5
MTH-204: Abstract Algebra
Lecture-12
The Sylow Theorems allow us to prove many useful results about finite groups. By using them,
we can often conclude a great deal about groups of a particular order if certain hypotheses are
satisfied. Recall that a group G is said to be simple if it has no non-trivial normal subgroups.
Example: Using the Sylow Theorems, we can determine that A5 has subgroups of orders 2, 3,
4, and 5. The Sylow p-subgroups of A5 have orders 3, 4, and 5. The Third Sylow Theorem tells
us exactly how many Sylow p-subgroups A5 has. Since the number of Sylow 5-subgroups must
divide 60 and also be congruent to 1 (mod 5), there are either one or six Sylow 5-subgroups in A5 .
All Sylow 5-subgroups are conjugate. If there were only a single Sylow 5-subgroup, it would be
conjugate to itself; that is, it would be a normal subgroup of A5 . Since A5 has no normal subgroups,
this is impossible; hence, we have determined that there are exactly six distinct Sylow 5-subgroups
of A5 .
Theorem 1.1. If p and q are distinct primes with p < q, then every group G of order pq has a single
subgroup of order q and this subgroup is normal in G. Hence, G cannot be simple. Furthermore, if
q 6≡ 1 (mod p), then G is cyclic.
Proof. We know that G contains a subgroup H of order q. The number of conjugates of H divides
pq and is equal to 1 + kq for k = 0, 1, . . .. However, 1 + q is already too large to divide the order of
the group; hence, H can only be conjugate to itself. That is, H must be normal in G.
The group G also has a Sylow p-subgroup, say K. The number of conjugates of K must divide q
and be equal to 1+kp for k = 0, 1, . . .. Since q is prime, either 1+kp = q or 1+kp = 1. If 1+kp = 1,
then K is normal in G. In this case, we can easily show that G is an internal direct product of H
and K. Since H is isomorphic to Zq and K is isomorphic to Zp , G ∼ = Zp × Zq ∼
= Zpq .
Example: Every group of order 15 is cyclic. This is true because 15 = 5 · 3 and 5 6≡ 1 (mod 3).
Example: Let us classify all of the groups of order 99 = 32 · 11 up to isomorphism. First we will
show that every group G of order 99 is abelian. By the Third Sylow Theorem, there are 1 + 3k
Sylow 3-subgroups, each of order 9, for some k = 0, 1, 2, . . .. Also, 1 + 3k must divide 11; hence,
there can only be a single normal Sylow 3-subgroup H in G. Similarly, there are 1 + 11k Sylow
1
11-subgroups and 1 + 11k must divide 9. Consequently, there is only one Sylow 11-subgroup K in
G. By Corollary 12.5, any group of order p2 is abelian for p prime; hence, H is isomorphic either to
Z3 × Z3 or to Z9 . Since K has order 11, it must be isomorphic to Z11 . Therefore, the only possible
groups of order 99 are Z3 × Z3 × Z11 or Z9 × Z11 up to isomorphism.
To determine all of the groups of order 5 · 7 · 47 = 1645, we need the following theorem.
Theorem 1.2. Let G0 = h[a, b] = aba−1 b−1 : a, b ∈ Gi be the subgroup consisting of all finite
products of elements of the form aba−1 b−1 in a group G. Then G0 is a normal subgroup of G and
G/G0 is abelian.
Example: We will now show that every group of order 5 · 7 · 47 = 1645 is abelian, and cyclic
by Corollary 8.11. By the Third Sylow Theorem, G has only one subgroup H1 of order 47. So
G/H1 has order 35 and must be abelian by above theorem. Hence, the commutator subgroup of
G is contained in H which tells us that |G0 | is either 1 or 47. If |G0 | = 1, we are done. Suppose
that |G0 | = 47. The Third Sylow Theorem tells us that G has only one subgroup of order 5 and
one subgroup of order 7. So there exist normal subgroups H2 and H3 in G, where |H2 | = 5 and
|H3 | = 7. In either case the quotient group is abelian; hence, G0 must be a subgroup of Hi , i = 1, 2.
Therefore, the order of G0 is 1, 5, or 7. However, we already have determined that |G0 | = 1 or 47.
So the commutator subgroup of G is trivial, and consequently G is abelian.
Given a finite group, one can ask whether or not that group has any normal subgroups. Recall that
a simple group is one with no proper nontrivial normal subgroups. As in the case of A5 , proving a
group to be simple can be a very difficult task; however, the Sylow Theorems are useful tools for
proving that a group is not simple. Usually some sort of counting argument is involved.
Example: Let us show that no group G of order 20 can be simple. By the Third Sylow Theorem, G
contains one or more Sylow 5-subgroups. The number of such subgroups is congruent to 1 (mod 5)
and must also divide 20. The only possible such number is 1. Since there is only a single Sylow
5-subgroup and all Sylow 5-subgroups are conjugate, this subgroup must be normal.
Example: Let G be a finite group of order pn , n > 1 and p prime. We know that, G has a
nontrivial center. Since the center of any group G is a normal subgroup, G cannot be a simple
group. Therefore, groups of orders 4, 8, 9, 16, 25, 27, 32, 49, 64, and 81 are not simple. In fact,
the groups of order 4, 9, 25, and 49 are abelian as these numbers are of the form p2 for some prime
p.
2
Example: No group of order 56 = 23 · 7 is simple. We have seen that if we can show that there
is only one Sylow p-subgroup for some prime p dividing 56, then this must be a normal subgroup
and we are done. By the Third Sylow Theorem, there are either one or eight Sylow 7-subgroups.
If there is only a single Sylow 7-subgroup, then it must be normal.
On the other hand, suppose that there are eight Sylow 7-subgroups. Then each of these sub-
groups must be cyclic; hence, the intersection of any two of these subgroups contains only the
identity of the group. This leaves 8 · 6 = 48 distinct elements in the group, each of order 7. Now let
us count Sylow 2-subgroups. There are either one or seven Sylow 2-subgroups. Any element of a
Sylow 2-subgroup other than the identity must have as its order a power of 2; and therefore cannot
be one of the 48 elements of order 7 in the Sylow 7-subgroups. Since a Sylow 2-subgroup has order
8, there is only enough room for a single Sylow 2-subgroup in a group of order 56. If there is only
one Sylow 2-subgroup, it must be normal.
For other groups G it is more difficult to prove that G is not simple. Suppose G has order 48.
In this case the technique that we employed in the last example will not work.
Example: To demonstrate that a group G of order 48 is not simple, we will show that G contains
either a normal subgroup of order 8 or a normal subgroup of order 16. By the Third Sylow Theorem,
G has either one or three Sylow 2-subgroups of order 16. If there is only one subgroup, then it
must be a normal subgroup.
Suppose that the other case is true, and two of the three Sylow 2-subgroups are H and K. We
claim that |H ∩ K| = 8. If |H ∩ K| ≤ 4, then,
16 · 16
|HK| = = 64,
4
which is impossible. So H ∩ K is normal in both H and K since it has index 2. The normalizer
of H ∩ K contains both H and K, and |H ∩ K| must both be a multiple of 16 greater than 1 and
divide 48. The only possibility is that |N (H ∩ K)| = 48. Hence, N (H ∩ K) = G.
The following famous conjecture of Burnside was proved in a long and difficult paper by Feit
and Thompson.
Theorem 1.3. (Odd Order Theorem) Every finite simple group of nonprime order must be of
even order.
The proof of this theorem laid the groundwork for a program in the 1960s and 1970s that
classified all finite simple groups. The success of this program is one of the outstanding achievements
of modern mathematics.
3
MTH-204: Abstract Algebra
Lecture-13
The ultimate goal of group theory is to classify all groups up to isomorphism; that is, given a
particular group, we should be able to match it up with a known group via an isomorphism. For
example, we have already proved that any finite cyclic group of order n is isomorphic to Zn ; hence,
we “know” all finite cyclic groups. It is probably not reasonable to expect that we will ever know
all groups; however, we can often classify certain types of groups or distinguish between groups in
special cases. In this lecture we will characterize all finite abelian groups.
Note that Zmn ∼ = Zm × Zn when gcd(m, n) = 1. In fact, much more is true. Every finite abelian
group is isomorphic to a direct product of cyclic groups of prime power order; that is, every finite
abelian group is isomorphic to a group of the type
Zpα1 × · · · × Zpαnn .
1
First, let us examine a slight generalization of finite abelian groups. Suppose that G is a group
and let {gi } be a set of elements in G, where i is in some index set I (not necessarily finite). The
smallest subgroup of G containing all of the gi ’s is the subgroup of G generated by the gi ’s. If
this subgroup of G is in fact all of G, then G is generated by the set {gi : i ∈ I}. In this case the
gi ’s are said to be the generators of G. If there is a finite set {gi : i ∈ I} that generates G, then
G is finitely generated.
Example: Obviously, all finite groups are finitely generated. For example, the group S3 is gen-
erated by the permutations (12) and (123). The group Z × Zn is an infinite group but is finitely
generated by {(1, 0), (0, 1)}.
Example: Not all groups are finitely generated. Consider the rational numbers Q under the
operation of addition. Suppose that Q is finitely generated with generators p1 /q1 , . . . , pn /qn , where
each pi /qi is a fraction expressed in its lowest terms. Let p be some prime that does not divide any
of the denominators q1 , . . . , qn . We claim that 1/p cannot be in the subgroup of Q that is generated
by p1 /q1 , . . . , pn /qn , since p does not divide the denominator of any element in this subgroup. This
fact is easy to see since the sum of any two generators is
pi /qi + pj /qj = (pi qj + pj qi )/(qi qj ).
1
Theorem 1.1. Let H be the subgroup of a group G that is generated by {gi ∈ G : i ∈ I}. Then
h ∈ H exactly when it is a product of the form
h = giα11 · · · giαnn ,
The reason that powers of a fixed gi may occur several times in the product is that we may
have a nonabelian group. However, if the group is abelian, then the gi ’s need occur only once. For
example, a product such as a−3 b5 a7 could always be simplified (in this case, to a4 b5 ).
Proof. Let K be the set of all products of the form giα11 · · · giαnn , where the gik ’s are not necessarily
distinct. Certainly K is a subset of H. We need only show that K is a subgroup of G. If this is
the case, then K = H, since H is the smallest subgroup containing all the gi ’s.
Clearly, the set K is closed under the group operation. Since gi0 = 1, the identity is in K. It
remains to show that the inverse of an element g = g1k1 · · · giknn in K must also be in K. However,
Now let us restrict our attention to finite abelian groups. We can express any finite abelian
group as a finite direct product of cyclic groups. We shall prove that every finite abelian group is
isomorphic to a direct product of cyclic p-groups. Before we state the main theorem concerning
finite abelian groups, we shall consider a special case.
Theorem 1.2. A finite abelian group is isomorphic to the direct product of its distinct Sylow
subgroups.
2
Example: Suppose that we wish to classify all abelian groups of order 540 = 22 · 33 · 5. The
Fundamental Theorem of Finite Abelian Groups tells us that we have the following six possibilities.
• Z2 × Z2 × Z3 × Z3 × Z3 × Z5 ;
• Z2 × Z2 × Z3 × Z9 × Z5 ;
• Z2 × Z2 × Z27 × Z5 ;
• Z4 × Z3 × Z3 × Z3 × Z5 ;
• Z4 × Z3 × Z9 × Z5 ;
• Z4 × Z27 × Z5 .
Proof. Suppose that the order of G is pn . We shall induct on n. If n = 1, then G is cyclic of order
p and must be generated by g. Suppose now that the statement of the lemma holds for all integers
m
k with 1 ≤ k < n and let g be of maximal order in G, say |g| = pm . Then ap = e for all a ∈ G.
Now choose h in G such that h ∈ / hgi, where h has the smallest possible order. Certainly such an
h exists; otherwise, G = hgi and we are done. Let H = hhi.
We claim that hgi ∩ H = {e}. It suffices to show that |H| = p. Since |hp | = |h|/p, the order of
hp is smaller than the order of h and must be in hgi by the minimality of h; that is, hp = g r for
some number r. Hence,
m−1 m−1 m
(g r )p = (hp )p = hp = e,
and the order of g r must be less than or equal to pm−1 . Therefore, g r cannot generate hgi. Notice
that p must occur as a factor of r, say r = ps, and hp = g r = g ps . Define a to be g −s h. Then a
cannot be in hgi; otherwise, h would also have to be in hgi. Also,
ap = g −sp hp = g −r hp = h−p hp = e.
We have now formed an element a with order p such that a ∈ / hgi. Since h was chosen to have the
smallest order of all of the elements that are not in hgi, |H| = p.
Now we will show that the order of gH in the factor group G/H must be the same as the order
of g in G. If |gH| < |g| = pm , then
m−1 m−1
H = (gH)p = gp H;
m−1
hence, g p must be in hgi ∩ H = {e}, which contradicts the fact that the order of g is pm . There-
fore, gH must have maximal order in G/H. By the Correspondence Theorem and our induction
hypothesis,
G/H ∼= hgHi × K/H
3
for some subgroup K of G containing H. We claim that hgi ∩ K = {e}. If b ∈ hgi ∩ K, then
bH ∈ hgHi ∩ K/H = {H} and b ∈ hgi ∩ H = {e}. It follows that G = hgiK implies that
G∼ = hgi × H.
Proof of the Fundamental Theorem: The proof of the Fundamental Theorem of Finite
Abelian Groups follows very quickly from the above lemma. Suppose that G is a finite abelian
group. Then G is isomorphic to direct product of its Sylow pi -subgroups. So it is sufficient to show
that each Sylow pi -subgroup Hi is a product of cyclic pi subgroups. Note that each Hi is of order a
power of pi . Let g be an element of maximal order in Hi . If hgi = Hi , then we are done; otherwise,
Hi ∼
= Z|g| ×Ki for some subgroup Ki contained in Hi by the lemma. Since |Ki | < |Hi |, by induction
Ki is a product of cyclic pi -subgroups and hence Hi is a product of cyclic pi subgroups.
We now state the more general theorem for all finitely generated abelian groups. The proof is
complicated and we skip it.
4
MTH-204: Abstract Algebra
Lecture-14
1 Rings
Definition: Let R be a non-empty set which has two laws of composition defined on it. (we call
these law “addition” and “multiplication” respectively and use the familiar notation). We say that
R is a ring if the following hold:
1. a + b ∈ R and ab ∈ R ∀a, b ∈ R
Thus a ring is an additive Abelian group on which an operation of multiplication is defined; this
operation being associative and distributive with respect to the addition.
R is called a commutative ring if it satisfies in addition ab = ba for all a, b ∈ R . The term
non-commutative ring usually stands for “a not necessarily commutative ring”
Integral Domain: An integral domain is a commutative ring R with identity 1 6= 0 with no
zero divisors; that is, ab = 0 implies that a = 0 or b = 0.
Field: A Field is an integral domain in which every nonzero element has a multiplicative
inverse.
The following can be deduced from the axioms for a ring:
2. Given a ∈ R, −a is uniquely
1
4. a + b = a + c if and only if b = c for a, b, c ∈ R
5. Given a, b ∈ R, the equation x + a = b has a unique solution x = b + (−a)
6. −(a + b) = −a − b for all a, b ∈ R
7. −(a − b) = −a + b for all a, b ∈ R
8. a · 0 = 0 · a = 0 for all a ∈ R
9. a(−b) = (−a)b = −ab for all a, b ∈ R
10. (−a)(−b) = ab for all a, b ∈ R
11. a(b − c) = ab − ac for all a, b, c ∈ R
A subset S of a ring R is called a subring of R if S itself is a ring with respect to the laws of
composition of R.
2
Lemma 1.1. A non-empty subset S of a ring R is a subring of R if and only if a − b ∈ S and
ab ∈ S whenever a, b ∈ S
Proof. If S is a subring then obviously the given condition is satisfied. Conversely, suppose that
the condition holds. Take any a ∈ S. We have a − a ∈ S hence 0 ∈ S. Hence for any x ∈ S we
have 0 − x ∈ S so −x ∈ S. Finally, if a, b ∈ S then by the above −b ∈ S. Therefore a − (−b) ∈ S,
i.e., a + b ∈ S. So S is closed with respect to both addition and multiplication. Thus S is a subring
since all the other axioms are automatically satisfied.
1. I is a subring of R
Proof. Easy.
Let I be an ideal of a ring R and x ∈ R . Then the set of elements {x + i : i ∈ I} is called the coset
of x in R with respect to I. It is denoted by x + I
When dealing with cosets, it is more important to realise that, in general, a given coset can
be represented in more than one way. The next lemma shows how the coset representatives are
related.
Proof. Easy.
We denote the set of all cosets of R with respect to I by R/I. We can give R/I the structure
of a ring as follows: Define (x + I) + (y + I) = (x + y) + I and (x + I)(y + I) = xy + I for x, y ∈ R.
The key point here is that the sum and the product of R/I are well-defined, that is, they are
independent of the coset representatives chosen.
The ring R/I is called the residue class ring of R with respect to I
The zero element of R/I is 0 + I = i + I for any i ∈ I . If S is a subset of R with S ⊇ I we
denote by S/I the subset {s + I : s ∈ S} of R/I.
3
1. Every ideal of the ring R/I is of the form K/I where K is an ideal of R and K ⊇ I. Also
conversely, if K is an ideal of R and K ⊇ I then K/I is an ideal of R/I
2. There is a one to one correspondence between ideals of the ring R/I and the ideals of R
containing I
1. θ(0R ) = 0S
3. K = {x ∈ R : θ(x) = 0S } is an ideal of R
4. θR = {θ(r) : r ∈ R} is a subring of S
Proof. Easy.
K is called the kernel of θ and θR is called the (homomorphic) image of R. The ideal K is
sometimes denoted by ker θ.
Isomorphism: Let θ be a homomorphism of a ring R into a ring S. Then θ is called an
isomorphism if θ is a one to one and onto map. We say that R and S are isomorphic rings and
denote this by R ∼
= S.
Question: Given a ring R, what rings can occur as its homomorphic images?
The importance of the first isomorphism theorem lies in the fact that it shows the answer to
lie with R itself. It tells us that if we know all the ideals of R then we know all the homomorphic
images of R. Only the first isomorphism theorem contains new information. The other two are
simply its application.
4
Theorem 1.6. Let θ be a homomorphism of a ring R into a ring S. Then θR ∼
= R/I where
I = ker θ
Proof. Define σ : R/I → R by σ(x + I) = θ(x) for all x ∈ R. The map σ is well defined since for
x, y ∈ R, x + I = y + I ⇒ x − y ∈ I = ker θ ⇒ θ(x − y) = 0 ⇒ θ(x) = θ(y). The map σ is easily
seen to be the required isomorphism.
Proof. Let σ be the natural homomorphism R → R/I. Restrict σ to the ring L. We have σL =
(L + I)/I. The kernel of σ restricted to L is L ∩ I. Now apply previous theorem.
Proof. K/I is an ideal of R/I and so (R/I)/(K/I) is defined. Define a map γ : R/I → R/K by
γ(x + I) = x + K for all x ∈ R. The map γ is easily seen to be well defined and a homomorphism
onto R/K. Further,
γ(x + I) = K ⇐⇒ x+K =K
⇐⇒ x∈K
⇐⇒ x + I ∈ K/I
5
MTH-204: Abstract Algebra
Lecture-15
(a) a ≤ a
(b) a ≤ b, b ≤ c ⇒ a ≤ c
(c) a ≤ b, b ≤ a⇒ a = b
2. Let S be a partially ordered set. A non-empty subset τ is said to be totally ordered if for
every pair a, b ∈ τ we have either a ≤ b or b ≤ a
4. Let τ be a totally ordered subset of a partially ordered set S. We say that τ has an upper
bound in S if there exists c ∈ S such that x ≤ c for all x ∈ τ.
Theorem 1.1 (Zorn’s Lemma (Axiom)). If a partially ordered set S has the property that every
totally ordered subset of S has an upper bound in S, then S contains a maximal element.
A non-empty set S is said to be well-ordered if it is totally ordered and every non-empty subset
of S has a minimal element.
Theorem 1.2 (The Well ordering Principle). Any non-empty set can be well-ordered.
The Axiom of Choice: Given a class of sets, there exists a “choice function”, i.e., a function
which assigns to each of these sets one of its elements.
It can be shown that Axiom of Choice is logically equivalent to Zorn’s Lemma which is logically
equivalent to the Well-ordering Principle.
Maximal Ideal: An ideal I of a ring R is said to be maximal if I 6= R and I is not properly
contained in any other ideal of R.
1
Ideal Generators: If R is commutative and has a 1, then the ideal of R generated by a subset
A of R is defined by:
Proposition 1.3. Let I 6= R be an ideal of a ring R. Then there exists a maximal ideal M of R
such that M ⊇ I.
Proof. We will prove this by using Zorn’s Lemma. Let S be the set of all proper ideals of R
containing I. Partially order S by inclusion. Let {Tα }α∈Λ be a totally ordered subset of S. Let
T = ∪α∈Λ Tα . Then T Cr R and T ⊇ I. Moreover T is proper since T = R ⇒ 1 ∈ T ⇒ 1 ∈ Tα for
some α ∈ Λ ⇒ Tα = R. Thus T 6= R and so T ∈ S. Thus T 6= R and so T ∈ S. Now T ⊇ Tα for
all α ∈ Λ. Hence Zorn’s Lemma applies and S contains a maximal element, say M . Clearly M is
a maximal ideal and M ⊇ I.
Remark: This is not true for rings without 1. For example take the abelian group (Q, +) and
define the multiplication of any two elements to be zero, i.e., xy = 0 for any x, y. Then this is a
ring without a maximal ideal because (Q, +) doesn’t have a maximal subgroup.
Lemma 1.5. A (non-zero) ring R is a field if and only if its only ideals are {0} and R.
Note that we don’t need elements to define the ideals {0} and R. {0} can be defined as the
ideal that all other ideals contain, and R is the ideal that contains all other ideals. Alternatively,
we can reword this as “R is a field if and only if it has only two ideals” to avoid mentioning explicit
ideals.
This is another reason why fields are special. They have the simplest possible ideal structure.
There is an easy way to recognize if an ideal is maximal.
Proof. R/I is a field if and only if {0} and R/I are the only ideals of R/I. By the ideal corre-
spondence, this is equivalent to saying I and R are the only ideals of R which contains I, i.e. I is
maximal. So done.
2
This is a nice result. This makes a correspondence between properties of ideals I and properties
of the quotient R/I. Here is another one:
Prime Ideal: An ideal I of a ring R is said to be a prime ideal if ab ∈ I for a, b ∈ R then
either a ∈ I or b ∈ I.
Examples: A non-zero ideal nZ of Z is prime if and only if n is a prime.
To show this, first suppose n = p is a prime, and a · b ∈ pZ. So p | a · b. So p | a or p | b, i.e.
a ∈ pZ or b ∈ pZ.
For the other direction, suppose n = pq is a composite number (p, q 6= 1). Then n ∈ nZ but
p 6∈ nZ and q 6∈ nZ, since 0 < p, q < n.
So instead of talking about prime numbers, we can talk about prime ideals instead, because
ideals are better than elements.
We prove a result similar to the above:
Proof. Let I be prime. Let a + I, b + I ∈ R/I, and suppose (a + I)(b + I) = 0R/I . By definition,
(a+I)(b+I) = ab+I. So we must have ab ∈ I. As I is prime, either a ∈ I or b ∈ I. So a+I = 0R/I
or b + I = 0R/I . So R/I is an integral domain.
Conversely, suppose R/I is an integral domain. Let a, b ∈ R be such that ab ∈ I. Then
(a + I)(b + I) = ab + I = 0R/I ∈ R/I. Since R/I is an integral domain, either a + I = 0R/I or
b + I = 0R/i , i.e. a ∈ I or b ∈ I. So I is a prime ideal.
Prime ideals and maximal ideals are the main types of ideals we care about. Note that every
field is an integral domain. So we immediately have the following result:
Proof. An ideal I of R is maximal implies R/I is a field implies R/I is an integral domain implies
I is prime.
The converse is not true. For example, {0} ⊆ Z is prime but not maximal. Less stupidly,
(X) ∈ Z[X, Y ] is prime but not maximal (since Z[X, Y ]/(X) ∼
= Z[Y ]). We can provide a more
explicit proof of this, which is essentially the same.
Alternative proof. Let I be a maximal ideal, and suppose a, b 6∈ I but ab ∈ I. Then by maximality,
I +(a) = I +(b) = R = (1). So we can find some p, q ∈ R and n, m ∈ I such that n+ap = m+bq = 1.
Then
1 = (n + ap)(m + bq) = nm + apm + bqn + abpq ∈ I,
since n, m, ab ∈ I. This is a contradiction.
3
Note that for any ring R, there is a unique ring homomorphism Z → R, given by
ι:Z→R
n ≥ 0 7→ 1R + 1R + · · · + 1R
| {z }
n times
n ≤ 0 7→ −(1R + 1R + · · · + 1R )
| {z }
−n times
Any homomorphism Z → R must be given by this formula, since it must send the unit to the unit,
and we can show this is indeed a homomorphism by distributivity. So the ring homomorphism is
unique.
We then know ker(ι) is an ideal of Z. Thus ker(ι) = nZ for some n.
Characteristic of a ring: Let R be a ring, and ι : Z → R be the unique such map. The
characteristic of R is the unique non-negative n such that ker(ι) = nZ.
Example: The rings Z, Q, R, C all have characteristic 0. The ring Z/nZ has characteristic n.
In particular, all natural numbers can be characteristics.
Lemma 1.9. Let R be an integral domain. Then its characteristic is either 0 or a prime number.
Proof. Consider the unique map φ : Z → R, and ker(φ) = nZ. Then n is the characteristic of R by
definition.
By the first isomorphism theorem, Z/nZ = Im(φ) ≤ R. So Z/nZ is an integral domain. So nZ
is a prime ideal of Z. So n = 0 or a prime number.
4
MTH-204: Abstract Algebra
Lecture-16
Lemma 1.1. Let R be a finite ring which is an integral domain. Then R is a field.
a·−:R→R
b 7→ a · b
We want to show this is injective. For this, it suffices to show the kernel is trivial. If r ∈ ker(a · −),
then a · r = 0. So r = 0 since R is an integral domain. So the kernel is trivial.
Since R is finite, a · − must also be surjective. In particular, there is an element b ∈ R such
that a · b = 1R . So a has an inverse. Since a was arbitrary, R is a field.
So far, we know fields are integral domains, and subrings of integral domains are integral
domains. We have another good source of integral domain as follows:
Lemma 1.2. Let R be an integral domain. Then R[X] is also an integral domain.
Proof. We need to show that the product of two non-zero elements is non-zero. Let f, g ∈ R[X] be
non-zero, say
f = a0 + a1 X + · · · + an X n ∈ R[X]
g = b0 + b1 X + · · · + bm X m ∈ R[X],
Notation 1.3. Write R[X, Y ] for (R[X])[Y ], the polynomial ring of R in two variables. In general,
write R[X1 , · · · , Xn ] = (· · · ((R[X1 ])[X2 ]) · · · )[Xn ].
1
Then if R is an integral domain, so is R[X1 , · · · , Xn ].
We now mimic the familiar construction of Q from Z. For any integral domain R, we want to
construct a field F that consists of “fractions” of elements in R. Recall that a subring of any field
is an integral domain. This says the converse — every integral domain is the subring of some field.
Field of fractions Let R be an integral domain. A field of fractions F of R is a field with the
following properties
1. R ≤ F
2. Every element of F may be written as a · b−1 for a, b ∈ R, where b−1 means the multiplicative
inverse to b 6= 0 in F .
Proof. The construction is exactly how we construct the rationals from the integers — as equiva-
lence classes of pairs of integers. We let
S = {(a, b) ∈ R × R : b 6= 0}.
We need to show this is indeed a equivalence relation. Symmetry and reflexivity are obvious. To
show transitivity, suppose
(a, b) ∼ (c, d), (c, d) ∼ (e, f ),
i.e.
ad = bc, cf = de.
We multiply the first equation by f and the second by b, to obtain
Rearranging, we get
d(af − be) = 0.
Since d is in the denominator, d 6= 0. Since R is an integral domain, we must have af − be = 0, i.e.
af = be. So (a, b) ∼ (e, f ). This is where being an integral domain is important.
Now let
F = S/∼
be the set of equivalence classes. We now want to check this is indeed the field of fractions. We
first want to show it is a field. We write ab = [(a, b)] ∈ F , and define the operations by
a c ad + bc
+ =
b d bd
a c ac
· = .
b d bd
2
These are well-defined, and make (F, +, ·, 01 , 11 ) into a ring. There are many things to check, but
those are straightforward, and we will not waste time doing that here.
a a 0
Finally, we need to show every non-zero element has an inverse. Let b 6= 0F , i.e. b 6= 1, or
a · 1 6= b · 0 ∈ R, i.e. a 6= 0. Then ab ∈ F is defined, and
b a ba
· = = 1F .
a b ba
a
So b has a multiplicative inverse. So F is a field.
We now need to construct a subring of F that is isomorphic to R. To do so, we need to define
an injective isomorphism φ : R → F . This is given by
φ:R→F
r
r 7→ .
1
This is a ring homomorphism, as one can check easily. The kernel is the set of all r ∈ R such that
r
1 = 0, i.e. r = 0. So the kernel is trivial, and φ is injective. Then by the first isomorphism theorem,
R∼= Im(φ) ⊆ F .
Finally, we need to show everything is a quotient of two things in R. We have
−1
a a 1 a b
= · = · ,
b 1 b 1 1
as required.
This gives us a very useful tool. Since this gives us a field from an integral domain, this allows
us to use field techniques to study integral domains. Moreover, we can use this to construct new
interesting fields from integral domains.
Example Consider the integral domain C[X]. Its field of fractions is the field of all rational
functions p(X)
q(X) , where p, q ∈ C[X].
We now move on to tackle the problem of factorization in rings. For sanity, we suppose throughout
the section that R is an integral domain. We start by making some definitions.
Unit: An element a ∈ R is a unit if there is a b ∈ R such that ab = 1R . Equivalently, if the
ideal (a) = R.
Division: For elements a, b ∈ R, we say a divides b, written a | b, if there is a c ∈ R such that
b = ac. Equivalently, if (b) ⊆ (a).
Associates: We say a, b ∈ R are associates if a = bc for some unit c. Equivalently, if (a) = (b).
Equivalently, if a | b and b | a.
In the integers, this can only happen if a and b differ by a sign, but in more interesting rings,
more interesting things can happen.
3
When considering division in rings, we often consider two associates to be “the same”. For
example, in Z, we can factorize 6 as
6 = 2 · 3 = (−2) · (−3),
but this does not violate unique factorization, since 2 and −2 are associates (and so are 3 and −3),
and we consider these two factorizations to be “the same”.
Irreducible We say a ∈ R is irreducible if a 6= 0, a is not a unit, and if a = xy, then x or y is
a unit.
For integers, being irreducible is the same as being a prime number. However, “prime” means
something different in general rings.
Prime: We say a ∈ R is prime if a is non-zero, not a unit, and whenever a | xy, either a | x or
a | y.
It is important to note all these properties depend on the ring, not just the element itself.
Example: 2 ∈ Z is a prime, but 2 ∈ Q is not (since it is a unit).
Similarly, the polynomial 2X ∈ Q[X] is irreducible (since 2 is a unit), but 2X ∈ Z[X] not
irreducible.
We have two things called prime, so they had better be related.
Lemma 1.5. A principal ideal (r) is a prime ideal in R if and only if r = 0 or r is prime.
Proof. (⇒) Let (r) be a prime ideal. If r = 0, then done. Otherwise, as prime ideals are proper,
i.e. not the whole ring, r is not a unit. Now suppose r | a · b. Then a · b ∈ (r). But (r) is prime. So
a ∈ (r) or b ∈ (r). So r | a or r | b. So r is prime.
(⇐) If r = 0, then (0) = {0} C R, which is prime since R is an integral domain. Otherwise, let
r 6= 0 be prime. Suppose a · b ∈ (r). This means r | a · b. So r | a or r | b. So a ∈ (r) and b ∈ (r).
So (r) is prime.
Note that in Z, prime numbers exactly match the irreducibles, but prime numbers are also
prime (surprise!). In general, it is not true that irreducibles are the same as primes. However, one
direction is always true.
Proof. Let r ∈ R be prime, and suppose r = ab. Since r | r = ab, and r is prime, we must have
r | a or r | b. wlog, r | a. So a = rc for some c ∈ R. So r = ab = rcb. Since we are in an integral
domain, we must have 1 = cb. So b is a unit.
4
By definition, it is a subring of a field. So it is an integral domain. What are the units of the ring?
There is a nice trick we can use, when things are lying inside C. Consider the function
N : R → Z≥0
given by √
N (a + b −5) 7→ a2 + 5b2 .
It is convenient to think of this as z 7→ z z̄ = |z|2 . This satisfies N (z · w) = N (z)N (w). This is a
desirable thing to have for a ring, since it immediately implies all units have norm 1 — if r · s = 1,
then 1 = N (1) = N (rs) = N (r)N (s). So N (r) = N (s) = 1.
So to find the units, we need to solve a2 + 5b2 = 1, for a and b units. The only solutions are
±1. So only ±1 ∈ R can be units, and these obviously are units. So these are all the units.
Next, we claim 2 ∈ R is irreducible. We again use the norm. Suppose 2 = ab. Then 4 =
N (2) = N (a)N (b). Now note that nothing has norm 2. a2 + 5b2 can never be 2 for integers
a, b ∈ √
Z. So we √
must have, wlog, N (a) = 4, N (b) = 1. So b must be a unit. Similarly, we see that
3, 1 + −5, 1 − −5 are irreducible (since there is also no element of norm 3).
We have four irreducible elements in this ring. Are they prime? No! Note that
√ √
(1 + −5)(1 − −5) = 6 = 2 · 3.
√ √
We now claim 2 does not divide 1 + −5 or 1 − −5. So 2 is not prime.
√ √ √
+ −5. Then N (2) | N (1 + √
To show this, suppose 2 | 1√ −5). But N (2) = 4 and N (1 + −5) =
6, and 4 - 6. Similarly, N (1 − −5) = 6 as well. So 2 - 1 ± −5.
There are several life lessons here. First is that primes and irreducibles are not the same thing
in general. We’ve always thought they were the same because we’ve been living in the fantasy land
of the integers. But we need to grow up.
The one is that factorization into irreducibles is not necessarily unique, since 2 · 3 =
√ second √
(1 + −5)(1 − −5) are two factorizations into irreducibles.
However, there is one situation when unique factorizations holds. This is when we have a
Euclidean algorithm available.
5
MTH-204: Abstract Algebra
Lecture-17
a = b · q + r,
φ(f ) = deg(f ).
(3) The Gaussian integers R = Z[i] ≤ C is a Euclidean domain with φ(z) = N (z) = |z|2 .
Before we move on to prove unique factorization, we first derive something we’ve previously
mentioned. Recall we showed that every ideal in Z is principal, and we proved this by the Euclidean
algorithm. So we might expect this to be true in an arbitrary Euclidean domain.
Principal ideal domain A ring R is a principal ideal domain (PID) if it is an integral domain,
and every ideal is a principal ideal, i.e. for all ideal I of R, there is some a such that I = (a).
Example: Z is a principal ideal domain.
We have already proved this, just that we did it for a particular Euclidean domain Z. Nonethe-
less, we shall do it again.
1
Proof. Let R have a Euclidean function φ : R \ {0} → Z≥0 . We let I be a non-zero ideal of R, and
let b ∈ I \ {0} be an element with φ(b) minimal. Then for any a ∈ I, we write
a = bq + r,
with r = 0 or φ(r) < φ(b). However, any such r must be in I since r = a − bq ∈ I. So we cannot
have φ(r) < φ(b). So we must have r = 0. So a = bq. So a ∈ (b). Since this is true for all a ∈ I,
we must have I ⊆ (b). On the other hand, since b ∈ I, we must have (b) ⊆ I. So we must have
I = (b).
This is exactly, word by word, the same proof as we gave for the integers, except we replaced
the absolute value with φ.
Example: Z is a Euclidean domain, and hence a principal ideal domain. Also, for any field F,
F[X] is a Euclidean domain, hence a principal ideal domain.
Also, Z[i] is a Euclidean domain, and hence a principal ideal domain.
What is a non-example of principal ideal domains?
In Z[X], the ideal (2, X) is not a principal ideal. Suppose it were. Then (2, X) = (f ). Since
2 ∈ (2, X) = (f ), we know 2 ∈ (f ) , i.e. 2 = f · g for some g. So f has degree zero, and hence
constant. So f = ±1 or ±2.
If f = ±1, since ±1 are units, then (f ) = Z[X]. But (2, X) 6= Z[X], since, say, 1 6∈ (2, X). If
f = ±2, then since X ∈ (2, X) = (f ), we must have ±2 | X, but this is clearly false. So (2, X)
cannot be a principal ideal.
Example: Let A ∈ Mn×n (F) be an n × n matrix over a field F. We consider the following set
2
2. If p1 p2 · · · pn = q1 · · · qm with pi , qj irreducibles, then n = m, and they can be reordered such
that pi is an associate of qi .
This is a really nice property, and here we can do things we are familiar with in number theory.
So how do we know if something is a unique factorization domain?
Our goal is to show that all principal ideal domains are unique factorization domains. To do
so, we are going to prove several lemmas that give us some really nice properties of principal ideal
domains.
√
Recall we saw that every prime is an irreducible, but in Z[ −5], there are some irreducibles
that are not prime. However, this cannot happen in principal ideal domains.
Lemma 1.2. Let R be a principal ideal domain. If p ∈ R is irreducible, then it is prime.
Note that this is also true for general unique factorization domains, which we can prove directly
by unique factorization.
b = rpb + sab.
This is similar to the argument for integers. For integers, we would say if p - a, then p and a are
coprime. Therefore there are some r, s such that 1 = rp + sa. Then we continue the proof as above.
Hence what we did in the middle is to do something similar to showing p and a are “coprime”.
Another nice property of principal ideal domains is the following:
Lemma 1.3. Let R be a principal ideal domain. Let I1 ⊆ I2 ⊆ I3 ⊆ · · · be a chain of ideals. Then
there is some N ∈ N such that In = In+1 for some n ≥ N .
So in a principal ideal domain, we cannot have an infinite chain of bigger and bigger ideals.
Ascending chain condition: A ring satisfies the ascending chain condition (ACC) if there is
no infinite strictly increasing chain of ideals.
Noetherian ring: A ring that satisfies the ascending chain condition is known as a Noetherian
ring.
So we are proving that every principal ideal domain is Noetherian.
3
Proof. The obvious thing to do when we have an infinite chain of ideals is to take the union of
them. We let
∞
[
I= In ,
n≥1
which isSagain an ideal. Since R is a principal ideal domain, I = (a) for some a ∈ R. We know
a∈I= ∞ n=0 In . So a ∈ IN for some N . Then we have
(a) ⊆ IN ⊆ I = (a)
Notice it is not important that I is generated by one element. If, for some reason, we know I
is generated by finitely many elements, then the same argument work. So if every ideal is finitely
generated, then the ring must be Noetherian. It turns out this is an if-and-only-if — if you are
Noetherian, then every ideal is finitely generated. We will prove this later on in the course.
Finally, we have done the setup, and we can prove the proposition promised.
Proposition 1.4. Let R be a principal ideal domain. Then R is a unique factorization domain.
But then we have an ascending chain of ideals. By the ascending chain condition, these are all
eventually equal, i.e. there is some n such that (rn ) = (rn+1 ) = (rn+2 ) = · · · . In particular, since
(rn ) = (rn+1 ), and rn = rn+1 sn+1 , then sn+1 is a unit. But this is a contradiction, since sn+1 is
not a unit. So r must be a product of irreducibles.
To show uniqueness, we let p1 p2 · · · pn = q1 q2 · · · qm , with pi , qi irreducible. So in particular
p1 | q1 · · · qm . Since p1 is irreducible, it is prime. So p1 divides some qi . We reorder and suppose
p1 | q1 . So q1 = p1 · a for some a. But since q1 is irreducible, a must be a unit. So p1 , q1 are
associates. Since R is a principal ideal domain, hence integral domain, we can cancel p1 to obtain
p2 p3 · · · pn = (aq2 )q3 · · · qm .
p2 p3 · · · pn = q2 q3 · · · qm .
We can then continue to show that pi and qi are associates for all i. This also shows that n = m,
or else if n = m + k, saw, then pk+1 · · · pn = 1, which is a contradiction.
4
MTH-204: Abstract Algebra
Lecture-18
Proof. We construct the greatest common divisor using the good-old way of prime factorization.
We let p1 , p2 , · · · , pm be a list of all irreducible factors of ai , such that no two of these are
associates of each other. We now write
m
n
Y
ai = ui pj ij ,
j=1
and choose
m
m
Y
d= pj j .
j=1
Then we must have tj ≤ nij for all i, j. So we must have tj ≤ mj for all j. So d0 | d.
Uniqueness is immediate since any two greatest common divisors have to divide each other.
1
1.1 Factorization in polynomial rings
Since polynomial rings are a bit more special than general integral domains, we can say a bit more
about them.
Recall that for F a field, we know F [X] is a Euclidean domain, hence a principal ideal domain,
hence a unique factorization domain. Therefore we know
3. Let f be irreducible, and suppose (f ) ⊆ J ⊆ F [X]. Then J = (g) for some g. Since (f ) ⊆ (g),
we must have f = gh for some h. But f is irreducible. So either g or h is a unit. If g is a
unit, then (g) = F [X]. If h is a unit, then (f ) = (g). So (f ) is a maximal ideal. Note that
this argument is valid for any PID, not just polynomial rings.
To use the last item, we can first show that F [X]/(f ) is a field, and then use this to deduce that
f is irreducible. But we can also do something more interesting — find an irreducible f , and then
generate an interesting field F [X]/(f ).
So we want to understand reducibility, i.e. we want to know whether we can factorize a poly-
nomial f . Firstly, we want to get rid of the trivial case where we just factor out a scalar, e.g.
2X 2 + 2 = 2(X 2 + 1) ∈ Z[X] is a boring factorization.
Content: Let R be a UFD and f = a0 + a1 X + · · · + an X n ∈ R[X]. The content c(f ) of f is
c(f ) = gcd(a0 , a1 , · · · , an ) ∈ R.
Lemma 1.2 (Gauss’ lemma). Let R be a UFD, and f ∈ R[X] be a primitive polynomial. Then f
is reducible in R[X] if and only if f is reducible F [X], where F is the field of fractions of R.
We can’t do this right away. We first need some preparation. Before that, we do some examples.
Example: Consider X 3 + X + 1 ∈ Z[X]. This has content 1 so is primitive. We show it is not
reducible in Z[X], and hence not reducible in Q[X].
2
Suppose f is reducible in Q[X]. Then by Gauss’ lemma, this is reducible in Z[X]. So we can
write
X 3 + X + 1 = gh,
for some polynomials g, h ∈ Z[X], with g, h not units. But if g and h are not units, then they
cannot be constant, since the coefficients of X 3 + X + 1 are all 1 or 0. So they have degree at least
1. Since the degrees add up to 3, we wlog suppose g has degree 1 and h has degree 2. So suppose
g = b0 + b1 X, h = c0 + c1 X + c2 X 2 .
b0 c0 = 1
c2 b1 = 1
Proof. We let
f = a0 + a1 X + · · · + an X n ,
g = b0 + b1 X + · · · + bm X m ,
where an , bm 6= 0, and f, g are primitive. We want to show that the content of f g is a unit.
Now suppose f g is not primitive. Then c(f g) is not a unit. Since R is a UFD, we can find an
irreducible p which divides c(f g).
By assumption, c(f ) and c(g) are units. So p - c(f ) and p - c(g). So suppose p | a0 , p | a1 , . . . ,
p | ak−1 but p - ak . Note it is possible that k = 0. Similarly, suppose p | b0 , p | b1 , · · · , p | b`−1 , p - b` .
We look at the coefficient of X k+` in f g. It is given by
X
ai bj = ak+` b0 + · · · + ak+1 b`−1 + ak b` + ak−1 b`+1 + · · · + a0 b`+k .
i+j=k+`
However, the terms ak+` b0 + · · · + ak+1 b`−1 , is divisible by p, as p | bj for j < `. Similarly,
ak−1 b`+1 + · · · + a0 b`+k is divisible by p. So we must have p | ak b` . As p is irreducible, and hence
prime, we must have p | ak or p | b` . This is a contradiction. So c(f g) must be a unit.
3
Corollary 1.4. Let R be a UFD. Then for f, g ∈ R[X], we have that c(f g) is an associate of
c(f )c(g).
Again, we cannot say they are equal, since content is only well-defined up to a unit.
Proof. We can write f = c(f )f1 and g = c(g)g1 , with f1 and g1 irreducible. Then
f g = c(f )c(g)f1 g1 .
Since f1 g1 is primitive, so c(f )c(g) is a gcd of the coefficients of f g, and so is c(f g), by definition.
So they are associates.
Proof. We will show that a primitive f ∈ R[X] is reducible in R[X] if and only if f is reducible in
F [X].
One direction is almost immediately obvious. Let f = gh be a product in R[X] with g, h not
units. As f is primitive, so are g and h. So both have degree > 0. So g, h are not units in F [X].
So f is reducible in F [X].
The other direction is less obvious. We let f = gh in F [X], with g, h not units. So g and h
have degree > 0, since F is a field. So we can clear denominators by finding a, b ∈ R such that
(ag), (bh) ∈ R[X] (e.g. let a be the product of denominators of coefficients of g). Then we get
abf = (ag)(bh),
and this is a factorization in R[X]. Here we have to be careful — (ag) is one thing that lives in
R[X], and is not necessarily a product in R[X], since g might not be in R[X]. So we should just
treat it as a single symbol.
We now write
(ag) = c(ag)g1 ,
(bh) = c(bh)h1 ,
So cancelling ab gives
f = u−1 g1 h1 ∈ R[X].
So f is reducible in R[X].
4
We will do another proof performed in a similar manner.
Proposition 1.6. Let R be a UFD, and F be its field of fractions. Let g ∈ R[X] be primitive.
We let J =< g >⊂ R[X] be the ideal generated by g in R[X] and I =< g >⊂ F [X] be the ideal
generated by g in F [X]. Then
J = I ∩ R[X].
In other words, if f ∈ R[X] and we can write it as f = gh, with h ∈ F [X], then in fact h ∈ R[X].
Proof. The strategy is the same — we clear denominators in the equation f = gh, and then use
contents to get that down in R[X].
We certainly have J ⊆ I ∩ R[X]. Now let f ∈ I ∩ R[X]. So we can write
f = gh,
bf = g(bh) ∈ R[X].
We let
(bh) = c(bh)h1 ,
for h1 ∈ R[X] primitive. Thus
bf = c(bh)gh1 .
Since g is primitive, so is gh1 . So c(bh) = uc(bf ) for u a unit. But bf is really a product in R[X].
So we have
c(bf ) = c(b)c(f ) = bc(f ).
So we have
bf = ubc(f )gh1 .
Cancelling b gives
f = g(uc(f )h1 ).
So g | f in R[X]. So f ∈ J.
5
MTH-204: Abstract Algebra
Lecture-19
Proposition 1.1. Let R be a UFD, and F be its field of fractions. Let g ∈ R[X] be primitive.
We let J =< g >⊂ R[X] be the ideal generated by g in R[X] and I =< g >⊂ F [X] be the ideal
generated by g in F [X]. Then
J = I ∩ R[X].
In other words, if f ∈ R[X] and we can write it as f = gh, with h ∈ F [X], then in fact h ∈ R[X].
Proof. We know R[X] has a notion of degree. So we will combine this with the fact that R is a
UFD.
Let f ∈ R[X]. We can write f = c(f )f1 , with f1 primitive. Firstly, as R is a UFD, we may
factor
c(f ) = p1 p2 · · · pn ,
for pi ∈ R irreducible (and also irreducible in R[X]). Now we want to deal with f1 .
If f1 is not irreducible, then we can write
f1 = f2 f3 ,
with f2 , f3 both not units. Since f1 is primitive, f2 , f3 also cannot be constants. So we must have
deg f2 , deg f3 > 0. Also, since deg f2 + deg f3 = deg f1 , we must have deg f2 , deg f3 < deg f1 . If
f2 , f3 are irreducible, then done. Otherwise, keep on going. We will eventually stop since the
degrees have to keep on decreasing. So we can write it as
f1 = q1 · · · qm ,
1
with qi irreducible. So we can write
f = p1 p2 · · · pn q1 q2 · · · qm ,
a product of irreducibles.
For uniqueness, we first deal with the p’s. We note that
c(f ) = p1 p2 · · · pn
f1 = q1 q2 · · · qm = r1 r2 · · · r` .
Note that each qi and each ri is a factor of the primitive polynomial f1 , so are also primitive.
Now we do (maybe) the unexpected thing. We let F be the field of fractions of R, and consider
qi , ri ∈ F [X]. Since F is a field, F is a Euclidean domain, hence principal ideal domain, hence
unique factorization domain.
By Gauss’ lemma, since the qi and ri are irreducible in R[X], they are also irreducible in F [X].
As F [X] is a UFD, we find that ` = m, and after reordering, ri and qi are associates, say
ri = ui qi ,
with ui ∈ F [X] a unit. What we want to say is that ri is a unit times qi in R[X]. Firstly, note that
ui ∈ F as it is a unit. Clearing denominators, we can write
ai ri = bi qi ∈ R[X].
Taking contents, since ri , qi are primitives, we know ai and bi are associates, say
bi = vi ai ,
The key idea is to use Gauss’ lemma to say the reducibility in R[X] is the same as reducibility
in F [X], as long as we are primitive. The first part about contents is just to turn everything into
primitives.
Note that the last part of the proof is just our previous proposition. We could have applied it,
but we decide to spell it out in full for clarity.
Example: We know Z[X] is a UFD, and if R is a UFD, then R[X1 , · · · , Xn ] is also a UFD.
This is a useful thing to know. In particular, it gives us examples of UFDs that are not PIDs.
However, in such rings, we would also like to have an easy to determine whether something is
reducible. Fortunately, we have the following criterion:
2
Proposition 1.3 (Eisenstein’s criterion). Let R be a UFD, and let
f = a0 + a1 X + · · · + an X n ∈ R[X]
1. p - an ;
3. p2 - a0 .
Then f is irreducible in R[X], and hence in F [X] (where F is the field of fractions of F ).
It is important that we work in R[X] all the time, until the end where we apply Gauss’ lemma.
Otherwise, we cannot possibly apply Eisenstein’s criterion since there are no primes in F .
g = r0 + r1 X + · · · + rk X k
h = s0 + s1 X + · · · + s` X ` ,
for rk , s` 6= 0.
We know rk s` = an . Since p - an , so p - rk and p - s` . We can also look at bottom coefficients.
We know r0 s0 = a0 . We know p | a0 and p2 - a0 . So p divides exactly one of r0 and s0 . wlog, p | r0
and p - s0 .
Now let j be such that
p | r0 , p | r1 , · · · , p | rj−1 , p - rj .
aj = r0 sj + r1 sj−1 + · · · + rj−1 s1 + rj s0 .
p | r0 sj + r1 sj−1 + · · · + rj−1 s1 .
Also, since p - rj and p - s0 , we know p - rj s0 , using the fact that p is prime. So p - aj . So we must
have j = n.
We also know that j ≤ k ≤ n. So we must have j = k = n. So deg g = n. Hence ` = n − h = 0.
So h is a constant. But we also know f is primitive. So h must be a unit. So this is not a proper
factorization.
Example: Consider the polynomial X n − p ∈ Z[X] for p a prime. Apply Eisenstein’s criterion
with p, and observe all the conditions hold. This is certainly primitive, since this is monic. So
3
√
X n − p is irreducible in Z[X], hence in Q[X]. In particular, X n − p has no rational roots, i.e. n p
where p is a prime number. If we look at this, we notice Eisenstein’s criteria does not apply. What
should we do? We observe that
Xp − 1
f= .
X −1
So it might be a good idea to let Y = X − 1. Then we get a new polynomial
(Y + 1)p − 1
p p p
fˆ = fˆ(Y ) = = Y p−1 + Y p−2 + Y p−3 + · · · + .
Y 1 2 p−1
When we look at it hard enough, we notice Eisenstein’s criteria can be applied — we know p | pi
p
for 1 ≤ i ≤ p − 1, but p2 - p−1 = p. So fˆ is irreducible in Z[Y ].
then we get
fˆ(Y ) = g(Y + 1)h(Y + 1)
in Z[Y ]. So f is irreducible.
Hence none of the roots of f are rational (but we already know that — they are not even real!).