Ring Notes
Ring Notes
important definitions
This section is incomplete. It does not contain all important definitions introduced in my lectures.
Please see the class note.
Definition 1.1. A commutative ring R with 1 , 0 is called an integral domain if whenever ab = 0
for a, b ∈ R, then either a = 0 or b = 0.
A commutative ring R with 1 , 0 is called a field if for every nonzero element a ∈ R, there exists
an element b ∈ R such that ab = 1.
Definition 1.2. Let R be a ring.
• A nonzero element a ∈ R is called a zero divisor if there exists a nonzero element b in R
such that either ab = 0 or ba = 0.
• An element a ∈ R is called a nonzero divisor if a , 0 and a is not a zero divisor.
• An element a ∈ R is called a nilpotent element if an = 0 for some positive integer n.
• An element a ∈ R is called a unit if there exists another element b ∈ R such that ab = ba = 1.
Definition 1.3. Let R be a commutative ring with 1 , 0. A polynomial in an indeterminate x with
coefficients in R is a formal sum of the form a0 xn +a1 xn−1 +. . .+an . As per convention the polynomial
1.x is simply denoted by x. We define addition and multiplication of polynomials as follows:
(a0 xn + a1 xn−1 + . . . + an ) + (b0 xm + b1 xm−1 + . . . + bm )
(a0 + b0 )xn + (a1 + b1 )xn−1 + . . . + (an + bn ) if n = m
= a0 x + a1 xn−1 + . . . + an−m−1 xm+1 + (an−m + b0 )xm + (an−m+1 + b1 )xm−1 + . . . + (a0 + b0 ) if n > m
n
b0 xm + b1 xm−1 + . . . + bm−n−1 xn+1 + (bm−n + a0 )xn + (bm−n+1 + a1 )xn−1 + . . . + (a0 + b0 ) if n < m
The set of polynomials is denoted by R[x]. It forms a commutative ring with identity under
addition and multiplication.
Definition 1.4. Let R be a commutative ring with 1 , 0. A power series in an indeterminate x with
coefficients in R is a formal infinite sum of the form a0 + a1 x + . . . + an xn + . . . . The addition and
multiplication of power series are defined as follows:
(a0 + a1 x + . . . an xn + . . .) + (b0 + b1 x + . . . + bn xn + . . .) = (a0 + b0 ) + (a1 + b1 )x + . . .
+ (an + bn )xn + . . .
(a0 + a1 x + . . . an xn + . . .)(b0 + b1 x + . . . + bn xn + . . .) = a0 b0 + (a0 b1 + a1 b0 )x + . . .
X
+ ai b j x n + . . .
i+ j=n,0≤i≤n,0≤ j≤n
The set of power series forms a commutative ring with identity under addition and multiplication
and it is denoted by R[|x|]. As before the element 1.x is denoted by x.
1
Definition 1.5. Let R be a ring and Mn (R) be the set of all n × n matrices with entries in R. Addition
and multiplication of matrices are defined as follows:
(ai j )n×n + (bi j )n×n = (ai j + bi j )n×n
n
X
(ai j )n×n (bi j )n×n = (ci j )n×n where ci j = aik bk j .
k=1
The set Mn (R) forms a ring under addition and multiplication of matrices.
2. Important problems
In this section, all rings are commutative and contain nonzero identities.
Question 2.1. Show that a polynomial f ∈ R[x] is a unit in R[x] if and only if f (0) is a unit in R
and all other coefficients of f are nilpotent in R.
Proof. Let f = a0 + a1 x + . . . + an xn where ai ∈ R for i = 0, 1, 2, . . . , n. First we assume that
f (0) = a0 is a unit in R and rest of the coefficients a1 , . . . , an are nilpotent elements of R. The
polynomial g = a1 x + . . . + an xn is a nilpotent element of R[x] since it is a finite sum of nilpotent
elements a1 x, a2 x, . . . , an x. This implies that f = a0 + g is a sum of a unit and a nilpotent element
of R[x]. Therefore f is a unit of R[x].
Conversely, we assume that f is a unit in R[x]. We prove the converse by induction on deg( f ). If
deg( f ) = 0, then f is a constant and there is nothing to prove. We assume that deg( f ) > 0. Since f
is a unit, there exists another polynomial g ∈ R[x] such that f g = g f = 1. Evaluating at x = 0, we
conclude that f (0), g(0) are units in R. Let g = b0 + b1 x + . . . + bm xm . We set ai = bi = 0 for i < 0
(Why is it required?). We have
X
f g = an bm xm+n + (an bm−1 + an−1 bm )xm+n−1 + . . . + ( ak bl )xi + . . . (a1 b0 + a0 b1 )x + a0 b0 .
k+l=i
3. Zorn’s Lemma
All rings in this section are nonzero, not necessarily commutative and may not contain identities.
Definition 3.1. A partial order on a nonempty set A is a relation ≺ satisfying the following :
1 x ≺ x for all x ∈ A,
2 if x ≺ y and y ≺ x, then x = y for any x, y ∈ A,
3 if x ≺ y and y ≺ z, then x ≺ z for any x, y, z ∈ A.
The set A is called a partially ordered set. It is denoted by (A, ≺).
Definition 3.2. (Maximal elements, upper bonds, chains) Let (A, ≺) be a partially ordered set and
B be a nonempty subset of A. We consider B partially ordered with the partial order ≺. Then
(1) The set B is called a chain if for any two elements x, y ∈ B, we have either x ≺ y or y ≺ x.
(2) An element u ∈ A is called an upper bound of B if x ≺ u for all x ∈ B.
(3) An element m ∈ B is called a maximal element of B if there does not exist another element
x ∈ B such that m ≺ x.
Examples 3.3. Let A = {{1, 2, 3}, {1, 2}, {2, 3}, {1, 3}, {1}, {2}, {3}, } be partially ordered by inclusion.
Let B = {{1, 2}, {2, 3}, {1, 3}, {1}, {2}, {3}}, then each of {1, 2}, {2, 3}, {1, 3} is a maximal element of B.
The element {{1, 2, 3} is an upper bound of B. The set B does not contain an upper bound.
Lemma 3.4. (Zorn’s Lemma) If A is a nonempty partially ordered set such that every chain has an
upper bound in A, then A has a maximal element.
Theorem 3.5. Let R be a ring, I be an ideal of R and S be a multiplicatively closed subset of R
such that S ∩ I = ∅. Then the following set has a maximal element.
S = {J : J is an ideal of R, I ⊂ J and J ∩ S = ∅}
If R is commutative, each maximal element is a prime ideal of R.
Proof. The set S is a partially ordered set ordered by inclusion. It is nonempty since I ∈ S. We
show that every chain in S has an upper bound in S. Then by Zorn’s lemma, we conclude that S
has a maximal element.
Let C be a chain in S. To show that C has an upper bound in S, we consider the set T = ∪ J∈C J.
Let x, y ∈ T and r ∈ R. There exist J1 , J2 ∈ C such that x ∈ J1 and y ∈ J2 . Since C is a chain,
we have either J1 ⊂ J2 or else J2 ⊂ J1 . In either cases, we obtain that x − y ∈ T . We also have
rx ∈ J1 ⊂ T . It follows that T is an ideal of R. Moreover T ∩ S = ∅ and I ⊂ T since each J ∈ C
is disjoint from S and contains I. Therefore T ∈ C. It follows that T is an upper bound of C. This
completes the first part of the proof.
5
Now we assume that R is a commutative ring. Let M be a maximal element of S. To show that
M is a prime ideal, we choose a, b ∈ R such that ab ∈ M. Assume on the contrary that a < M and
b < M. We observe that I ⊂ M ⊂ M + (a) and I ⊂ M ⊂ M + (b). Since M is a maximal element of
S, we must have [M + (a)] ∩ S , ∅, [M + (b)] ∩ S , ∅. This implies that there exist s1 , s2 ∈ S such
that s1 = m1 + r1 a and s2 = m2 + r2 b for m1 , m2 ∈ M and r1 , r2 ∈ R. Consequently we have
s1 s2 = (m1 + r1 a)(m2 + r2 b)
= m1 m2 + (m1 r2 b + r1 am2 + r1 ar2 b)
∈ M, since R is commutative and m1 , m2 , ab ∈ M.
This is a contradiction because M ∩ S = ∅. Therefore either a ∈ M or b ∈ M. Hence M is a prime
ideal. □
Corollary 3.6. Let R be a ring with 1 , 0. Then every proper ideal of R is contained in a maximal
ideal of R. If R is a commutative ring, then maximal ideals are prime ideals.
Proof. Let S = {1}. Then S is multiplicatively closed. Let I be a proper ideal of R. Clearly
S ∩ I = ∅. In this situation the set S described in Theorem 3.5 reduces to
S = {J : J is a proper ideal of R contatining I}.
Clearly every maximal element of S is a maximal ideal of R containing I. By Theorem 3.5, S
has a maximal element. Therefore I is contained in a maximal ideal.
To show that maximal ideals are prime ideals, we take I = 0. In this case S is given by
S = {J : J is a proper ideal of R}.
Any maximal ideal of R is a maximal element of S and therefore a prime ideal of R by Theorem
3.5. □
Remark 3.7. Consider the set Q of rational numbers as a ring with usual addition and trivial
multiplication (ab = 0 for a, b ∈ Q). Then Q does not have a maximal ideal. The ring Q does not
have the identity.
Definition 3.8. Let I be an ideal of R. A prime ideal P containing I is called a minimal prime over
I if there does not exist another prime ideal Q such that I ⫋ Q ⫋ P.
Question 3.9. Let I be a proper ideal of a ring R with 1 , 0. Show that I is contained in a minimal
prime ideal over I.
Proof. We consider the set
S = {P : P is a prime ideal containing I}.
We define a relation ≺ on S as follows:
P ≺ Q for P, Q ∈ S whenever Q ⊂ P.
It is easy to check that ≺ defines a partial order on S. We claim that any maximal element M of
S is a minimal prime over I. If it were not true, we would find a prime ideal M ′ containing I such
that I ⫋ M ′ ⫋ M. This implies that M ′ ∈ S and M ≺ M ′ contradicting the maximality of M in S.
6
We use Zorn’s lemma to show that S admits a maximal element. Since I is a proper ideal, I is
contained in a maximal ideal M0 of R. Clearly M0 ∈ S. Therefore S is a nonempty set.
Let C be a chain in S. We show that C has an upper bound contained in S. We define J = ∩P∈C P.
We have I ⊂ J since each prime in C contains I. If we can show that J is a prime ideal, then J ∈ S
and J is an upper bound of C.
Let ab ∈ J for a, b ∈ R. If possible, we assume that a, b < J. Then there exist Q, Q′ in C such
that a < Q and b < Q′ . Since C is a totally ordered set, we have either Q ⊂ Q′ or else Q′ ⊂ Q. If
Q ⊂ Q′ , we have a, b < Q. This implies that ab < Q since Q is a prime ideal. Similarly if Q′ ⊂ Q,
we conclude that ab < Q′ . Thus in either case we have ab < J which is a contradiction. Therefore
either a ∈ J or b ∈ J. It follows that J is a prime ideal.
Thus we establish that every nonempty chain of S has an upper bound in S. By Zorn’s lemma
S has an upper bound and the result follows. □
4. Rings of fractions
Let R be a nonzero commutative ring which does not necessarily contain identity. Let S be a
multiplicatively closed subset of R such that S does not contain the zero element and zero divisors.
We recall the construction of S −1 R. We define a relation ∼ on R × S by setting (r, s) ∼ (r′ , s′ )
if rs′ = sr′ . It was shown that ∼ defines an equivalence relation. As a set S −1 R is the set of
equivalence classes under ∼, i.e.
R×S
S −1 R = = {[(r, s)] : (r, s) ∈ R × S }.
∼
The set S −1 R is a commutative ring with identity [(s, s)] for s ∈ S under the following well
defined operations.
Addition : [(r, s)] + [(r′ , s′ )] = [(rs′ + sr′ , ss′ )],
Multiplication : [(r, s)][(r′ , s′ )] = [(rr′ , ss′ )].
The map i : R → S −1 R given by i(x) = [(xs, s)], s ∈ S is an injective ring homomorphism and
each element of S −1 R can be written as i(r)/i(s) = i(r)i(s)−1 for r ∈ R, s ∈ S . The map i is called
the inclusion map.
= QO
ϕ
ϕ̃
R / S −1 R
i
7
Proof. We know that each element of S −1 R can be written as i(r)/i(s) for r ∈ R, s ∈ S . We define a
map ϕ̃ : S −1 R → Q by ϕ̃(i(r)/i(s)) = ϕ(r)ϕ(s)−1 .
First we show that ϕ̃ is well defined. Assume that i(r)/i(s) = i(r′ )/i(s′ ). Then i(r)i(s′ ) = i(r′ )i(s).
This implies that i(rs′ ) = i(r′ s). Since the map i is injective, we have rs′ = r′ s. It follows that
ϕ(rs′ ) = ϕ(r′ s) which yields ϕ(r)ϕ(s′ ) = ϕ(r′ )ϕ(s). Therefore,
Lemma 4.2. Let R be a commutative ring with 1 , 0 and S = R \ {0}. Then R is an integral domain
if and only if S is multiplicatively closed.
Definition 4.3. (Fields of fractions) Let R be an integral domain and S = R \ {0}. The set S is a
multiplicatively closed set. The ring of fractions S −1 R is a field and contains R as a subring. It is
called the field of fractions of R and is denoted by Q(R).
r
Q(R) = { : r, s ∈ R, s , 0}
s
8
Definition 4.4. (Fields of rational functions) Let R be an integral domain and P = R[x1 , . . . , xn ] be
a polynomial ring in n indeterminate x1 , . . . , xn with coefficients in R. Let S = R[x1 , . . . , xn ] \ {0}
which is multiplicatively closed since P is an integral domain. The field of fractions S −1 P is called
the field of rational functions in indeterminate x1 , . . . , xn with coefficients in R and is denoted by
R(x1 , . . . , xn ).
f (x1 , . . . , xn )
R(x1 , . . . , xn ) = { : f (x1 , . . . , xn ), g(x1 , . . . , xn ) ∈ P, g(x1 , . . . , xn ) , 0}
g(x1 , . . . , xn )
The following result was discussed in the class.
Proposition 4.5. Let R be an integral domain and K = Q(R) be the field of fractions of R. Let
R[x1 , x2 , . . . , xn ], K[x1 , x2 , . . . , xn ] be polynomial rings in n indeterminate with coefficients in R, K
respectively. Then the fields of fractions of R[x1 , x2 , . . . , xn ] and K[x1 , x2 , . . . , xn ] are isomorphic.
In other words
R(x1 , x2 , . . . , xn ) K(x1 , x2 , . . . , xn ).
Proof. The inclusion R → K induces inclusion R[x1 , . . . , xn ] ⊂ K[x1 , . . . , xn ] ⊂ K(x1 , . . . , xn ).
Let j : R[x1 , . . . , xn ] → K(x1 , . . . , xn ) denote the inclusion. The set S = R[x1 , . . . , xn ] \ {0} is
multiplicatively closed. We find that j( f ) = f is a unit in K(x1 , . . . , xn ) for all f ∈ S since 1f
exists in K(x1 , . . . , xn ). Therefore by Theorem 4.1, there exists a unique ring homomorphism ψ :
R(x1 , . . . , xn ) → K(x1 , . . . , xn ) such that the diagram below commutes, i.e. ψ ◦ i = j.
K(x , . . . , xn )
6 1 O
j
ψ
R[x1 , . . . , xn ] / R(x1 , . . . , xn )
i
S 1−1 R1 × . . . × S n−1 Rn
7 O
ϕ
ψ
R / S −1 R
i
Lemma 5.4. Let I be an ideal of R[x] which contains a monic polynomial f of degree n. Then
every nonzero element of the quotient ring Q = R[x] I
can be written as the class of a polynomial g
such that deg(g) < n. If I = ( f ), then g is unique.
Lemma 5.5. Let R be a commutative ring with identity 1 , 0 and a ∈ R. Then a polynomial
f ∈ R[x] in divisible by (x − a) if and only if f (a) = 0.
Proof. By division algorithm, we may write f as f = q(x)(x − a) + r where q(x) ∈ R[x] and r ∈ R.
Evaluating both sides at x = a, we have f (a) = r. This implies that f = q(x)(x − a) + f (a). The
lemma follows obviously. □
The following generalises the above lemma to the case of many variables.
Proof. We recall that R[x1 , . . . , xi+1 ] = R[x1 , . . . , xi ][xi+1 ]. By repeated application of the above
lemma we have
Lemma 5.7. Let R be a commutative ring with 1 , 0. Let s be a nonzero divisor of R and S denote
the multiplicatively closed subset {1, s, s2 , . . .}. Then S −1 R (sx−1)
R[x]
.
11
6. isomorphism theorems
In this section rings are not necessarily commutative and do not necessarily contain identities.
Definition 6.1. Let ϕ : R → S be a ring homomorphism. Then kernel of ϕ denoted by ker ϕ is
defined as
ker ϕ = {x ∈ R : ϕ(x) = 0}.
The following lemma is easy and was discussed in the class.
Lemma 6.2. Let ϕ : R → S be a ring homomorphism. Then ker ϕ is a two sided ideal of R. The
map ϕ is injective if and only if ker(ϕ) = 0.
Notation : Unless there is a scope of confusion, we always denote the additive coset x + I of an
ideal I of a ring R containing an element x ∈ R by x̄, i.e. throughout this note x̄ = x + I.
Theorem 6.3. (Existence of quotient rings) Let R be a ring and I be an ideal of R. Let x̄ denote the
additive coset of I containing x. Then the set of additive cosets
R/I = { x̄ : x ∈ R}
form a ring under the following operations :
Addition : x̄ + ȳ = x + y
Multiplication : x̄ȳ = xy.
Proof. We know that (R, +) is an abelian group, so I is a normal subgroup of R. It follows from
results in Group theory that R/I is also an abelian group under addition as defined above.
We show that multiplication is well defined on R/I. Let x̄ = x̄′ and ȳ = y¯′ . Then x − x′ , y − y′ ∈ I.
This implies that x = x′ + i and y = y′ + j for i, j ∈ I. Therefore we have
xy = (x′ + i)(y′ + j)
= x′ y′ + x′ j + iy′ + y j
= x′ y′ + (x′ j + iy′ + y j)
The second summand of the last line belongs to I since I is a two sided ideal of R. Consequently
xy = x′ y′ establishing that the multiplication is well defined. Checking that R/I is a ring under
above operations is straightforward and therefore left to the reader. □
Theorem 6.4. (The first isomorphism theorem) Let ϕ : R → S be a surjective ring homomorphism.
Then ker ϕ is an ideal of R and the induced map ψ : R/ ker ϕ → S given by ψ( x̄) = ϕ(x) is an
isomorphism. Here x̄ denotes the quotient class of x in R/ ker ϕ.
Proof. Let x, y ∈ ker ϕ and r ∈ R. Then ϕ(x) = ϕ(y) = 0. We have ϕ(x − y) = ϕ(x) − ϕ(y) = 0,
ϕ(rx) = ϕ(r)ϕ(x) = 0 and ϕ(xr) = ϕ(x)ϕ(r) = 0. This shows that x − y, rx, xr ∈ ker ϕ. Therefore,
ker ϕ is an ideal of R.
The map ψ : R/I → S is given by ψ( x̄) = ϕ(x). First we show that ψ is a well defined map.
Let x̄ = ȳ in R/ ker ϕ. Then x − y ∈ ker ϕ. This implies that ϕ(x − y) = 0 giving ϕ(x) = ϕ(y).
Consequently ψ( x̄) = ψ(ȳ). Therefore ψ is well defined.
12
It is fairly easy to see that ψ is a ring homomorphism and ψ is surjective. Both properties follow
from the corresponding properties of ϕ. We leave it to the reader to work out details. To show that
ψ is injective, we prove that ker(ψ) = 0. Let x̄ ∈ ker ψ. Then ψ( x̄) = ϕ(x) = 0. This implies that
x ∈ ker ϕ. Consequently x̄ = 0 in R/ ker ϕ. Therefore ker ψ = 0. Hence ψ is an isomorphism and
the result is established. □
Theorem 6.5. (The second isomorphism theorem) Let R be a ring, S be a subring of R and I be an
ideal of R. Then the following hold true.
1 S + I = {s + i : s ∈ S , i ∈ I} is a subring of R,
2 I is an ideal of S + I,
3 S ∩ I is an ideal of S ,
4 There is an isomorphism
S S +I
given by s̄ 7→ s + i.
S ∩I I
Proof. Statements 1, 2, 3 are fairly easy to establish. We skip details. We only prove the last
assertion. We define a map ϕ : S → S +I I
given by ϕ(x) = x̄. One easily checks that ϕ is a ring
homomorphism. Elements of S +I I
are given by classes of elements of S , so ϕ is surjective. We
observe that ϕ(x) = 0 if and only if x ∈ I proving ker ϕ = I. The isomorphism now follows from
the first isomorphism theorem. □
Theorem 6.6. (The third isomorphism theorem) Let I, J be ideals of R such that I ⊂ J ⊂ R. Then
the map R/I
J/I
→ RJ given by x̄ + JI 7→ x̄ is an isomorphism.
Proof. We define a map ϕ : R/I → R/J given by ϕ(x + I) = x + J. Since I ⊂ J, one can easily check
that ϕ is well defined. As before, it is routine to check that ϕ is a surjective ring homomorphism
with kernel J/I = { x̄ : x ∈ J}. The result follows from the first isomorphism theorem. □
Notation Let F : S → T be a map between sets S , T . Let A ⊂ S and B ⊂ T . We define
F(A) = {y ∈ T : y = F(x) for some x ∈ A} and F −1 (B) = {x ∈ S : F(x) ∈ B}. If F is a ring
homomorphism, then both are subrings of ambient rings.
Theorem 6.7. (The fourth isomorphism theorem) Let I be an ideal of a ring R. The correspondence
A ←→ AI for A being a subring of R containing I establishes an inclusion preserving bijection
between the following sets:
{A : A is a subring of R containing I} ←→ {B is a subring of R/I}.
Proof. We set S = {A : A is a subring of R containing I} and T = {B is a subring of R/I}. Let
q : R → R/I be the quotient map given by q(x) = x̄.
Given a subring A in S, we associate with it the subring q(A) = A/I in T. Conversely, given a
subring B in T, we associate with it the subring q−1 (B) in S. Here q−1 (B) is in S because it contains
q−1 (0) = I.
The proof falls naturally into three steps. Steps 1, 2 together establish that both sets are in one-
one correspondence. Step 3 shows that the correspondence is inclusion preserving.
Step - 1 : q−1 (q(A)) = A for any A ∈ S.
13
Let x ∈ q−1 (q(A)). This implies that q(x) ∈ q(A). It follows that q(x) = q(a) for some a ∈ A.
Consequently q(x − a) = 0. Therefore x − a ∈ ker q = I ⊂ A which implies that x ∈ A. Thus we
show that q−1 (q(A)) ⊂ A. The reverse inclusion is obvious, so the assertion follows.
Step - 2 : q(q−1 (B)) = B for any B ∈ T.
The inclusion q(q−1 (B)) ⊂ B follows obviously. To prove the reverse inclusion, we choose y ∈ B.
The map q : R → R/I is surjective and B ⊂ R/I. It follows that there exists x ∈ R such that y = q(x).
Indeed x ∈ q−1 (B) since y ∈ B. This shows that y = q(x) ∈ q(q−1 (B)). Therefore B ⊂ q(q−1 (B)).
Thus the equality is established.
Step - 3 : q(A1 ) ⊂ q(A2 ) for A1 ⊂ A1 , A1 , A2 ∈ S and q−1 (B1 ) ⊂ q−1 (B2 ) for B1 ⊂ B2 , B1 , B2 ∈ T.
This step follows trivially. □
Z[x]
Question 6.8. Find all proper nonzero subrings of (2,x2 +x+1)
.
Proof. We have the following isomorphisms by the third isomorphism theorem:
Z[x] Z/2Z[x]
2 .
(2, + x + 1) (x + x + 1)
x2
The later ring is a field containing only four elements 1, 0, x, 1 + x. Its only non-trivial proper
subring is {0, 1} which corresponds to the subset {0̄, 1̄} under the above isomorphism. Therefore
Z[x]
(2,x2 +x+1)
has only one nontrivial proper subring, viz. {0̄, 1̄}. □
Theorem 7.6. (Chinese remainder theorem) Let R be a commutative ring with 1 , 0 and I1 , . . . , In
be ideals in R. Then the map f : R → R/I1 × . . . R/In given by f (x) = (x + I1 , . . . , x + In ) is a ring
homomorphism and ker f = I1 ∩ . . . ∩ In .
Moreover if I1 , . . . , In are pairwise co-maximal ideals, then f is surjective. In particular f in-
duces an isomorphism
R/I1 ∩ . . . ∩ In → R/I1 × . . . × R/In .
The ideal Ii is co-maximal to Ik for each k , i. By Lemma 7.5, the ideals Ii and I1 I2 . . . Ii−1 Ii+1 . . . In
are co-maximal ideals, so we have x ∈ Ii and y ∈ I1 I2 . . . Ii−1 Ii+1 . . . In such that x + y = 1. One easily
checks that y ≡ 0 (mod Ik ) for k , i and y ≡ 1 (mod Ii ). This shows that f (y) = ei . Therefore each
ei is in the image of f and the proof follows. □
Remark 7.7. It is well known that the class of m in Z/(n) is a unit if and only if m and n are
co-primes integers. Let n = pi11 . . . pikk be the factorisation of a positive integer n > 1 into power
of distinct primes p1 , . . . , pn . The ideals (pi11 ), (pi22 ), . . . , (pikk ) are pair wise co-maximal ideals. By
Chinese remainder theorem one has Z/(n) Z/(pi11 ) × . . . × Z/(pikk ). For a given prime p, there are
only pl−1 (p − 1) integers which are co-prime to pl and less that pi . This shows that each Z/(pill ) has
exactly pill −1 (pl − 1) units. Therefore Z/(n) has exactly pi11 −1 . . . pikk −1 (p1 − 1) . . . (pk − 1) units.
of k[x, y]. If (x, y) is a principal ideal of R[x, y], then its image (x, y) is also a principal ideal of
k[x, y], which is not true by the argument above since k is an integral domain (actually a field).
Therefore (x, y) is not a principal ideal of R[x, y].
If R is a commutative ring with 1 , 0, one can similarly show that the polynomial ring R[x1 , x2 , . . . , xn ]
is not a PID.
Proof. Let x be an irreducible element of R. Then x is nonzero and not a unit. The ideal (x) is a
proper ideal of R. Let M be a maximal ideal of R containing I, c.f. Corollary 3.6. The ideal M
is principal since R is a PID, so we have m ∈ R such that M = (m). Indeed m is a prime element
because M is also a prime ideal.
Now we have (x) ⊂ (m). This implies that x = cm for c ∈ R. The element x is irreducible and m
is not a unit (actually a prime element), so c is a unit in R. It follows that (x) = (m) = M is a prime
ideal. Therefore x is a prime element. □
The above lemma provides us a method to check whether a given integral domain is a PID.
√
Example 10.4. The ring Z[ −3] is not a PID.
√
Proof. The ring Z[ −3] is an integral domain √ with standard addition and multiplication of complex
numbers. We define a function N : Z[ −3] → N ∪ {0} √ by N(α) = αᾱ. One easily checks
that N is multiplicative,
√ √ i.e. N(αβ) = αβ for α, β ∈ Z[ −3]. Moreover N(α) = a 2
+ 3b2
for
α = a + b −3 ∈ Z[ −3]. √ √
We consider the ring homomorphism F : Z[x] → Z[ −3] given by F Z = idZ and F(x) = −3.
Clearly F is surjective. We compute its kernel. It is easy to see that x2 +3 ∈ ker F which implies that
(x2 + 3) ⊂ ker F. Let g ∈ ker F. By division algorithm (see Theorem 5.3) we have q(x), r(x) ∈ Z[x]
such that g = q(x)(x2 +3)+r(x) with r(x) √ = 0 or deg(r(x)) < deg(x +3) = 2. We write r(x) = ax+b
2
Thus Z[ω] is closed under multiplication. Therefore Z[ω] is a subring of C. Since 1 ∈ Z[ω] and C
is a field, we conclude that Z[ω] is an integral domain.
We define a function N : Z[ω] → N ∪ {0} by N(α) = αᾱ, α ∈ Z[ω]. One easily checks that N is
multiplicative, i.e. N(αβ) = N(α)N(β) for α, β ∈ Z[ω]. If α = a+bω, then N(α) = (a+bω)(a+bω̄) =
a2 + ab(ω + ω̄) + b2 ωω̄ = a2 + ab + b2 ∈ N ∪ {0}.
αβ̄
Let α, β ∈ Z[ω] and β , 0. We have αβ = N(β) = a + bω for a, b ∈ Q. We write a = c + e,
b = d + f where c, d ∈ Z and 0 ≤ |e|, | f | ≤ 2 . Let q = c + dω and r′ = e + f ω. Then q ∈ Z[ω] and
1
N(r′ ) = e2 + e f + f 2 ≤ 43 .
We have αβ = q + r′ which gives α = qβ + r where r = r′ β. If r , 0, then β , 0 and N(r) =
N(r′ )N(β) ≤ 34 N(β) < N(β). Moreover r = α − qβ ∈ Z[ω]. Thus N Z[ω]\{0} : Z[ω] \ {0} → N ∪ {0}
has the property that for any α, β ∈ Z[ω], β , 0, there exist q, r ∈ Z[ω] such that α = βq + r where
r = 0 or N(r) < N(β). Therefore Z[ω] is a Euclidean domain. □
Remark 10.7. Examples 10.4, 10.6 together show that a subring of a Euclidean domain may not
be a Euclidean domain. In fact every integral domain R is a subring or its field of fractions Q(R)
(see Definition 4.3) which is a Euclidean domain. Therefore any non-Euclidean integral domain
provides an example.
k[x,y,z]
Example 10.8. Let k be a field. Then the ring (xz−y2 )
is not a PID.
Proof. We prove that x̄ is an irreducible element but not a prime element. The result then follows
by Lemma 10.3.
20
Let F : k[x, y, z] → k[s, t] be the ring homomorphism given by F k = idk , F(x) = s2 , F(y) = st
and F(z) = t2 . Here k[s, t] is the polynomial ring in variables s, t. Image of F is the subring
k[s2 , st, t2 ], viz. namely set of polynomials in s2 , st, t2 .
First, we show that ker F = (xz − y2 ). We observe that s2 t2 − (st)2 = 0. This implies that
xz − y2 ∈ ker F. Let g ∈ ker F. Now k[x, y, z] = k[x, z][y]. By division algorithm (see Theorem
5.3), we have q, r ∈ k[x, z][y] such that g = q(xz − y2 ) + r where r = 0 or the degree of r in y is at
most one. Therefore r = hy + j for h, j ∈ k[x, z]. Now F(g) = 0 yields F(r) = 0. It follows that
h(s2 , t2 )st + j(s2 , t2 ) = 0. The monomials appearing in h(s2 , t2 )st are of the form s2m+1 t2n+1 , m, n ∈
N ∪ {0} whereas those in j(s2 , t2 ) are of the form s2m t2n , m, n ∈ N ∪ {0}. Consequently no term
of h(s2 , t2 )st can cancel a term in j(s2 , t2 ). Therefore h(s2 , t2 )st = j(s2 , t2 ) = 0 which implies that
h(s2 , t2 ) = j(s2 , t2 ) = 0 in k[s, t]. It follows that h(x, z) = j(x, z) = 0 in k[x, z]. This gives r = 0
and consequently g = q(xz − y2 ) ∈ (xz − y2 ). Therefore ker F = (xz − y2 ). By the first isomorphism
2 ) k[s , st, t ].
k[x,y,z] 2 2
theorem, F induces an isomorphism (xz−y
2 ) . The element x̄ corresponds to s in k[s , st, t ] which
k[x,y,z] 2 2 2
We show that x̄ is irreducible in (xz−y
is not a unit. It is enough to prove that s2 is irreducible in k[s2 , st, t2 ]. Let s2 = l(s, t)m(s, t) for
l(s, t), m(s, t) ∈ k[s2 , st, t2 ]. Comparing degrees in t of both sides of the equation, we conclude
that l(s, t), m(s, t) are free from t. If none of l(s, t), m(s, t) is a unit, then both of them are linear
polynomials of the form as + b ∈ k[s], a , 0, which is a contradiction since k[s2 , st, t2 ] does not
contain a linear polynomial in s. Therefore at least one of l(s, t), m(s, t) is a unit. It follows that s2
is irreducible in k[s2 , st, t2 ]. Therefore x̄ is irreducible in (xz−y
k[x,y,z]
2) .
k[x,y,z]
But x̄ is not a prime element of (xz−y2 )
. This is true because
k[x,y,z] k[x,y,z]
(xz−y2 ) (xz−y2 ) k[x, y, z] k[x, y, z] k[y, z]
(x,xz−y2 )
2
( x̄) (x, xz − y ) (x, y2 ) (y2 )
(xz−y2 )
Definition 11.1. Let R be a commutative ring with 1 , 0. Two elements a, b ∈ R are said to be
associate in R if there exists a unit u ∈ R such that a = ub.
Lemma 11.2. Let R be an integral domain. Two nonzero elements a, b ∈ R are associate in R if
and only if (a) = (b).
Proof. First we assume that a, b are associate in R. Then there is a unit u ∈ R such that a = ub. We
observe that a ∈ (b) which yields (a) ⊂ (b). Since u is a unit, we have b = u−1 a. This gives b ∈ (a)
and consequently (b) ⊂ (a). Therefore (a) = (b).
21
Conversely we assume that (a) = (b). Since a ∈ (b), we have a = cb for some c ∈ R. Similarly
b ∈ (a) gives that b = c′ a for some c′ ∈ R. It follows that a = cb = cc′ a. Since R is an integral
domain and a , 0, we get 1 = cc′ which shows that c is a unit. Therefore a, b are associate in
R. □
I1 ⫋ I2 ⫋ I3 ⫋ . . . .
This contradicts our assumption (1). Therefore S has a maximal element and (2) follows.
(2) =⇒ (3) :
We assume (2). Let I be an ideal of R. We consider the set of ideals :
S = {J : J ⊂ I and J is a finitely generated ideal of R}.
We observe that S , ∅ because (0) ∈ S. By our hypothesis, S has a maximal element. Let J0 be
a maximal element of S. Since J0 ∈ S, the ideal J0 is finitely generated and J0 ⊂ I. We show that
J0 = I.
If J0 ⫋ J, we choose x ∈ I \ J0 . The ideal J0 + (x) is finitely generated because J0 is so. We also
have J0 + (x) ⊂ I. It follows that J0 + (x) ∈ S and J0 ⫋ J0 + (x). This is a contradiction since J0 is a
maximal element of S. Therefore J0 = I. Consequently I is finitely generated. Hence (3) follows.
(3) =⇒ (1) We assume (3). We consider the ascending chain of ideals : I1 ⫋ I2 ⫋ I3 ⫋ . . .. Let
J = ∪ j≥1 I j . Since the union is taken over an ascending chain of ideals, one can easily see that J
is an ideal of R. By our hypothesis J is finitely generated, i.e. there exist x1 , x2 , . . . , xr such that
J = (x1 , x2 , . . . , xr ). We choose N large enough such that x j ∈ IN for j = 1, . . . , r. Then for any
i ≥ 1, we have
J = (x1 , x2 , . . . , xr ) ⊂ IN ⊂ IN+i ⊂ J.
Therefore IN = IN+i for all i ≥ 1 and (1) follows. □
Definition 11.4. A ring R is called a Noetherian ring if R satisfies any of the equivalent statements
of Proposition 11.3.
First, we observe that xn+1 < (x1 , . . . , xn ). This is true because otherwise we have xn+1 =
i=1 xi fi for some fi ∈ R[x1 , x2 , . . .] which leads to 1 = 0, a contradiction by putting x1 =
Pn
. . . = xn = 0 and xn+1 = 1. Therefore we get a strictly ascending chain of ideals :
I = ( f1 , f2 , . . . , fr ) ⊂ (x1 , x2 , . . . , xN ) ⊂ I.
Therefore, I = (x1 , x2 , . . . , xN ) which is a contradiction since xN+1 ∈ I but xN+1 < (x1 , x2 , . . . , xN ).
Hence, I = (x1 , x2 , . . . , xn , . . .) is not a finitely generated ideal.
Lemma 11.6. Let R be a PID and x ∈ R not a unit. Then x is divisible by a prime element.
Proof. We consider the ideal I = (x). Since x is not a unit, ideal I is a proper ideal of R. It follows
that I is contained in a maximal ideal, say M, c.f. Corollary 3.6. The ideal M is a principal ideal
because R is a PID. Therefore M = (p) for some p ∈ R. The ideal M is also a prime ideal, so p is a
prime element. Now x ∈ I ⊂ (p) yields x = cp for c ∈ R. This completes the proof. □
Proposition 11.7. Let R be a PID. Every nonzero, non-unit element of R can be written as a product
of finitely many prime elements of R.
Proof. We begin with an element x ∈ R which is nonzero and not a unit. By Lemma 11.6, we find a
prime element p1 ∈ R such that x = p1 x1 for x1 ∈ R. Since x , 0, we have x1 , 0. We get (x) ⫋ (x1 )
because p1 is not a unit.
If x1 is a unit, then (x) = (p1 ) is a prime ideal, i.e. x is a prime element. We are done. Otherwise
by Lemma 11.6, we find a prime element p2 ∈ R such that x1 = p2 x2 for x2 ∈ R. The inequality
x1 , 0 yields x2 , 0. We observe that (x1 ) ⫋ (x2 ) because p2 is not a unit.
If x2 is a unit, then (x1 ) = (p2 ) which implies that x1 is a prime element. It follows that x = p1 x1
is a prime factorization. In this case the proof completes here. If x2 is not a unit, we get a prime
element p3 ∈ R such that x2 = p3 x3 by Lemma 11.6. Since p3 is not a unit, we have (x2 ) ⫋ (x3 ).
23
Now we check whether x3 is a unit and repeat the argument as before. The following diagram
explains the steps in the argument.
(x)
{ #
(p1 ) (x1 )
{ "
(p2 ) (x2 )
|
(p3 )
We must have xN to be a unit for some N since otherwise we would obtain a strictly ascending
chain of ideals
(x) ⫋ (x1 ) ⫋ (x2 ) ⫋ . . .
which contradicts that R is a Noetherian ring. It follows that xN−1 = pN xN is a prime element.
Therefore x = p1 p2 . . . pN−1 xN−1 is a factorization of x into prime elements. This completes the
proof. □
Proof. It is easy to see that n = 0 if and only if m = 0. Without loss of generality we may assume
that 1 ≤ n ≤ m. We have u−1 p1 p2 . . . pn = q1 q2 . . . qm , so p1 divides q1 q2 . . . qm . Since p1 is a prime
element, p1 must divide some qi1 . This implies that qi1 = p1 u1 for u1 ∈ R. Now qi1 is irreducible
and p1 is not a unit. Therefore u1 is a unit in R. Thus we have p1 p2 . . . pn = uu1 p1 j,i1 q j . Since
Q
R is an integral domain, we may cancel out p1 from both sides. We get p2 . . . pn = uu1 j,i1 q j .
Q
Repeating this argument and cancelling out p2 , . . . , pn one by one, we obtain
n
Y Y
1=u uj qj
i=1 j∈I
A careful analysis of the proof of the lemma above actually gives the following.
24
Lemma 11.9. Let R be an integral domain. Let p1 , . . . , pn be prime elements and q1 , . . . , qm be
irreducible elements of R such that
p1 p2 . . . pn = uq1 q2 . . . qm
for some unit u in R. Then m = n and there exists a pertmutation σ ∈ S n such that pi and qσ(i) are
associate in R for i = 1, . . . , n.
Can you prove it yourself? You cannot say n ≤ m without loss of generality in your proof.
Remark 12.6. In an integral domain R, g.c.d. of two elements a, b is unique upto associate. Let
d, d′ both be g.c.d. of a, b. Then d divides d′ and d′ divides d. This implies that (d) = (d′ ). Therefore
d and d′ are associate in R (see Lemma 11.2). The g.c.d. of a and b is denoted by g. c. d.(a, b).
Proposition 12.7. Let a, b be elements of an integral domain R such that at least one of a, b is
nonzero. Then the g.c.d. of a, b exists if and only if the ideal
I = ∩(a,b)⊂(x) (x)
is a principal ideal. Moreover if I = (d), then d is the g.c.d. of a and b.
Proof. First we assume that g. c. d.(a, b) exists and equals d. Clearly (a, b) ⊂ (d). If (a, b) ⊂ (x),
then x a and x b. It follows that x d which gives (d) ⊂ (x). Therefore I = ∩(a,b)⊂(x) (x) = (d).
Conversely we assume that I is principal and I = (d) for d ∈ R. Since I , 0, the element d , 0.
We have (a, b) ⊂ I = (d). Therefore d a and d b. If x ∈ R\{0} such that x a and x b, then (a, b) ⊂ (x).
This implies that (d) = I ⊂ (x) and consequently x d. Therefore d is the g.c.d. of a and b. □
Example 12.8. 1) If R is a PID, then the g.c.d. of any two elements of R exists. Let a, b ∈ R.
The ideal (a, b) is a principal ideal. If (a, b) = (d), we have ∩(a,b)⊂(x) (x) = (d). Therefore d is
the g.c.d. of a and b.
2) Let R = Z[x]. Then g. c. d.(4, 2x) = 2, but (4, 2x) ⫋ (2). Thus in general the principal ideal
generated by g. c. d.(a, b) may not be equal to the ideal generated by a and b.
The following result is easy and can be proved using Lemmas 11.9, 12.3.
Theorem 12.9. Let R be an integral domain. Then R is a UFD if and only if every nonzero, non-unit
can be expressed as a finite product of prime elements.
we conclude that p1 and pi are associate in R which contradicts the hypothesis. Therefore, our
assumption is wrong and consequently b ≤ a1 . □
Theorem 13.3. Let R be a UFD, then the g.c.d. of any two nonzero elements exists.
Proof. Let a, b be nonzero elements of R. Now each of a, b can be factorized into a finite product
of irreducible elements as follows :
a = upa11 . . . pann qb11 qb22 . . . qbmm ,
a′ a′
b = u′ p11 . . . pnn r1c1 . . . rkck ,
Where u, u′ are units in R, no two different prime elements in each of the sets {p1 , . . . , pn , q1 , . . . , qm },
{p1 , . . . , pn , r1 , . . . , rk }, {q1 , q2 , . . . , qm , r1 , . . . , rk } are associate in R and ai , bi , a′i , ci ∈ N. We see at
min{a ,a′ } min{a ,a′ }
once that d = p1 1 1 . . . pn n n divides both a and b.
Let c a and c b. Any prime factor p of c divides both of a and b. This implies that p is associate
to one of the prime factors of a as well as b. If p is associated to qi , then p is not associate to any
prime factor of b since qi is so. Therefore p is associated to some pi . In other words, each prime
factor of c is associate to one of {p1 , . . . , pn }. Therefore c can be written as c = u′′ pd11 . . . pdnn for
some unit u′′ in R and nonnegative integers d1 , . . . , dn . Each pdi i divides both a and b. By Lemma
13.2, di ≤ min{ai , a′i } for i = 1, . . . , n . Therefore c divides d. Hence g. c. d.(a, b) exists and is equal
to d. □
The following lemma follows from the formula of g.c.d. of two elements in a UFD as described
in the proof of the above lemma.
Lemma 14.1. Let R be an integral domain and r ∈ R. Then r is an irreducible (prime or unit)
element of R if and only if r is an irreducible (prime or unit) element of R[x].
Definition 14.3. Let R be a UFD and f ∈ R[x]. The content of f is defined to be the g.c.d. of the
coefficients of f . It is denoted by c( f ).
Lemma 14.4. Let R be a UFD, f ∈ R[x] and d ∈ R. Then c(d f ) = ud c( f ) for some unit u in R.
Moreover f = c( f ) f1 where f1 ∈ R[x] such that c( f1 ) = 1.
Proposition 14.5. Let R be a UFD and f, g ∈ R[x], then c( f g) = u c( f ) c(g) for some unit u in R.
Proof. First, we show that c( f g) c( f ) c(g). We prove this by induction on the number n of prime
factors of c( f g). If n = 0, then c( f g) is a unit, so c( f g) c( f ) c(g) trivially. We assume n ≥ 1. Let p
be a prime factor of c( f g). It follows that p f g. Since p is a prime element of R, it is also a prime
element of R[x]. Therefore either p f or p g. Without loss of generality we assume that p f . Then
f = p f ′ for some f ′ ∈ R[x]. Clearly c( f ) = pu c( f ′ ) and c( f g) = c(p f ′ g) = u′ p c( f ′ g) for some
units u, u′ ∈ R. One observes that the number of prime factors of c( f ′ g) is less than that of c( f g).
Therefore by induction hypothesis, we get c( f ′ g) c( f ′ ) c(g). This implies that p c( f ′ g) p c( f ′ ) c(g).
Therefore c( f g) c( f ) c(g).
Any coefficient of f g is of the form ai b j where ai is a coefficient of f and b j is that of g. Now
c( f ) ai and c(g) b j together imply that c( f ) c(g) ai b j . Consequently c( f ) c(g) c( f g).
Since R is an integral domain, c( f g) c( f ) c(g) and c( f ) c(g) c( f g) together imply that c( f g) =
u c( f ) c(g) for some unit u in R. □
29
Lemma 14.6. (Gauss’ Lemma) Let R be a UFD and F be the field of fraction of R. Assume that
p(x) ∈ R[x]. If p(x) is reducible in F[x], then p(x) is reducible in R[x]. More precisely if p(x) = f g
for f, g ∈ F[x] \ F, then p = FG for some polynomials F, G ∈ R[x] \ R such that F = r f and G = sg
for some nonzero r, s ∈ F.
Proof. We are given that p(x) = f g for f, g ∈ F[x] \ F. We find nonzero r0 , s0 ∈ R such that
F0 = r0 f ∈ R[x] and G0 = s0 g ∈ R[x]. Then r0 s0 p(x) = F0G0 . Comparing contents of both sides
we get that r0 s0 c(p) = u c(F0 ) c(G0 ) for some unit u in R.
Now F0 = c(F0 )F1 and G0 = c(G0 )G1 for some polynomials F1 , G1 ∈ R[x] such that c(F1 ) =
c(G1 ) = 1. It follows that r0 s0 p(x) = F0G0 = c(F0 ) c(G0 )F1G1 = u−1 r0 s0 c(p)F1G1 . Now R is an
integral domain and r0 , s0 are nonzero elements of R, so cancelling out r0 s0 from both sides we get
p(x) = u−1 c(p)F1G1 = FG,
where F = u−1 c(p)F1 and G = G1 .
Clearly both F, G ∈ R[x]. One observes that
F = u−1 c(p)F1 = u−1 c(p) c(F0 )−1 F0 = r0 u−1 c(p) c(F0 )−1 f and
G = G1 = c(G0 )−1G0 = s0 c(G0 )−1 g.
Taking r = r0 u−1 c(p) c(F0 )−1 and s = s0 c(G0 )−1 , we get F = r f and G = sg. Moreover r, s are
nonzero elements of F, therefore the result follows. □
Lemma 14.7. Let R be a UFD and c( f ) = 1. If f is reducible in R[x], then f is also reducible in
F[x] where F is the field of fraction of R.
Proof. Since f is reducible in R[x], we get nonzero, non-unit polynomials g, h ∈ R[x] such that
f = gh. If either one of g, h were a constant, we would find that a nonzero, non-unit element (viz.
g or h) of R divides c( f ), a contradiction since c( f ) = 1. Therefore deg(g), deg(h) ≥ 1. Hence f is
reducible in F[x]. □
Remark 14.8. Let R and F be as in Lemma 14.6. It shows that if f is irreducible in R[x], then f is
so in F[x]. Lemma 14.7 shows that the converse is true when c( f ) = 1. The assumption c( f ) = 1 is
necessary in Lemma 14.7, e.g. 2x is reducible in Z[x] but not reducible in Q[x] (Why?).
Lemma 14.9. Let R be a UFD and f, g ∈ R[x] be such that c( f ) = c(g) = 1. Let F be the field of
fraction of R. Then f, g are associate in R[x] if and only if they are so in F[x].
Proof. Let f , g are associate in R[x]. Then f = ug for some unit u in R[x]. Since R is an integral
domain u is actually a unit in R and therefore in F. Hence f, g are associate in F[x].
Conversely we assume that f and g are associate in F[x], i.e. f = ug for some unit u in F[x].
Since F is a field, u is actually a nonzero element of F. Let u = ba . It follows that b f = ag. Since
c( f ) = c(g) = 1, we have c(b f ) = b and c(ag) = a. Therefore a = u′ b for some unit u′ of R. Thus
u = u′ and so u is a unit in R. Hence f, g are associate in R[x]. □
Theorem 14.10. If R is a UFD, then R[x] is a UFD.
30
Proof. Let f be a nonzero, non-unit element of R[x]. We show that f can be written as a finite
product of irreducible elements and the decomposition is unique upto associates. Let F be the field
of fraction of R. We consider two cases.
Case - 1 : c( f ) = 1. We view f as an element of F[x]. Since F[x] is a PID, it is a UFD. Therefore
f can be written as f = p1 , . . . , pm for some irreducible polynomials p1 , . . . , pm ∈ F[x]. By Gauss’
lemma (see Lemma 14.6) we have f = q1 , . . . qm where each qi is a polynomial in R[x] and is
associated to pi in F[x]. Since pi is irreducible in F[x], the polynomial qi is irreducible in F[x].
By Gauss’ Lemma 14.6, qi is also irreducible in R[x]. Thus we find a decomposition of f into a
product of finitely many irreducible polynomials in R[x].
Now we assume that f = q1 q2 . . . qm , f = r1 r2 . . . rn are decompositions of f into products of two
different sets {q1 , . . . , qm } and {r1 , r2 , . . . , rn } of irreducible polynomials in R[x]. Since c( f ) = 1,
we conclude by Proposition 14.5 that c(qi ) = c(r j ) = 1 for all i and j. Since each qi , r j are
irreducible polynomials in R[x], they are so in F[x]. Now F[x] is a PID and therefore is a UFD.
It follows that m = n and after suitable reordering of r j ’s, the polynomial qi is associated to ri in
F[x] for i = 1, . . . , n. By Lemma 14.9, the polynomial qi is associated to ri in R[x] for i = 1, . . . , n.
Therefore decomposition of f is unique upto associates.
Case - 2 : c( f ) is not a unit.
We write f = c( f ) f1 for some polynomial f1 in R[x] such that c( f1 ) = 1. Now c( f ) is an element
of R and R is a UFD, so it is a finite product of irreducible elements of R. By Case - 1, f1 is also
a finite product of irreducible elements of R[x]. Each irreducible element in R is also irreducible
in R[x] (see Lemma 14.1). Thus we find a decomposition f into a finite product of irreducible
elements of R[x].
Now we assume that f = q1 q2 . . . qm , f = r1 r2 . . . rn are two different decompositions into finite
products of irreducible elements in R[x]. Let q1 , . . . , q s , r1 , . . . , rt be elements in R and the rest do
not belong to R. The content of a polynomial divides itself. Each of q s+1 , . . . , qm , rt+1 , . . . , rn are
irreducible elements of R[x] of degree at least one, so each of them has content 1. Therefore we get
c( f ) = q1 , . . . , q s and also c( f ) = r1 r2 . . . rn . It follows that
q1 , . . . , q s = ur1 r2 . . . rt
uq s+1 , . . . , qm = rt+1 , . . . , rn
since q1 q2 . . . qm = r1 r2 . . . rn .
The ring R is a UFD and q1 , . . . , q s , r1 , . . . , rt are irreducible elements in R. Therefore the first
equation gives that s = t and qi is associated to ri in R (so in R[x]) for i = 1, . . . , s after some
reordering of ri ’s. The product in either sides of the second equation has content 1. By Case - 1,
m − s = n − t which yields m = n and after some reordering of r j , s + 1 ≤ j ≤ m the polynomial q j
is associated to r j for j = s + 1, . . . , m. Therefore decomposition of f into a product of irreducible
elements is unique upto associates. □
i j
R
If we regard R-algebras A, B as modules over R as explained above, then an R-algebra homomor-
phism f : A → B is a ring homomorphism sending 1A to 1B , which is an R-module homomorphism
as well.
Remark 16.11. Let R be a commutative ring and A be an algebra over R. Then an R-algebra
homomorphism f : R[x1 , . . . , xn ] → A is uniquely determined by the images of x1 , . . . , xn under f .
35
17. 19th March 2019
Let M be a module over a ring R and X be a nonempty subset of M. We define
RX = {r1 x1 + r2 x2 + . . . + rn xn : ri ∈ R, xi ∈ X for i = 1, 2, . . . , n and n ∈ N}.
One can check that RX is a submodule of M. The module M is said to be generated by X if
M = RX. M is called a finitely generated module if M = RX for some finite subset X of M.
Exercise 1. Let R = k[x, y] and m = (x, y). Find a generating set of mn containing n + 1 elements.
Show that mn cannot be generated by less than n + 1 elements.
One readily checks that α∈I Mα is a submodule of M. If I = {1, 2, . . . , n}, then ni=1 Mi is the set
P P
of all elements of the form ni=1 xi where xi ∈ Mi .
P
Definition 17.2. Let Mα , α ∈ I be a family of modules over a ring R. We define the product of
modules as
Y
Mα = {(xα ) : xα ∈ Mα for α ∈ I}.
α∈I
Here (xα ) denotes a set map I → ⊔α∈I Mα (disjoint union) such that α 7→ xα ∈ Mα . Clearly
Πα∈I Mα is a module over R under component wise addition and multiplication by elements of R. If
I = {1, 2, . . . , n}, then we have
M1 × . . . × Mn = {(x1 , . . . , xn ) : xi ∈ Mi for i = 1, 2, . . . , n}.
Definition 17.3. Let Mα , α ∈ I be a family of modules over a ring R. The direct sum of modules is
the submodule of Πα∈I Mα defined as
M Y
Mα = {(xα ) ∈ Mα : xα = 0 for all but finitely many α ∈ I}.
α∈I α∈I
If I = N, then M = {(x
Li ) : xi = 0 for all i ≥ N for some N ∈ N}. It is easy to see that if I
L
Qα∈I α
is a finite set, then α∈I Mα = α∈I
Mα .
Proposition 17.4. Let Ni , i ∈ I be a family of submodules of a module M over a ring R. Then the
following are equivalent.
(1) The map π : ⊕i∈I Ni → i∈I Ni given by π({ni }i∈I ) = i∈I ni is an isomorphism.
P P
(2) Ni ∩ j∈I\{i} N j = 0 for all i ∈ I.
P
(3) Every element x ∈ i∈I Ni can be uniquely written as x = xi1 + . . . + xik where i1 , . . . , ik are k
P
distinct elements of I and xi j ∈ Ni j , j = 1, . . . , k.
36
Proof. (1) =⇒ (2) : Let x ∈ Ni ∩ j∈I\{i} N j . Then we get x = ni = lk=1 n jk where ni ∈ Ni ,
P P
n jk ∈ N jk for k = 1, . . . , l and { j1 , j2 , . . . , jl } are l distinct elements of I \ {i}. We define an element
(xα ) ∈ ⊕i∈I Ni as follows :
ni if α = i,
xα = −n jk if α = jk , k = 1, . . . , l,
0 otherwise.
Definition 17.8. Let X be a non empty subset of an LR-module M. Then M is called free on X, if M
is the direct sum of submodules Rx, x ∈ X, i.e. M = x∈X
Rx.
If M is free on X, we often say in short that M is a free R-module. In this case the set X is called
a basis of M. Every element of M is uniquely written as a finite linear combination of elements of
X with coefficients in R.
Lemma 17.9. Let R be a ring and X be a non empty set. Then there exists a module free on X.
Proof. Let x ∈ X. First, we construct a free module on {x}. We consider the set of symbols
Rx = {rx : r ∈ R}.
The set Rx forms an R-module under addition and multiplication by elements of R given by
We identify x with 1x. Then x ∈ Rx and Rx is free on {x} since every element of Rx is uniquely
written as rx = r(1x).
We consider the R-module
M
Rx.
x∈X
L
We identify x ∈ X with the map (xα ) : X → ⊔ x∈X Rx L ∈ x∈X
Rx whose value
L is 1x at x and
zero otherwise. Thus we L have an inclusion map X → x∈X
Rx. The module x∈X
Rx is free on
X, since every element in x∈X
Rx has a unique presentation of the form r1 x1 + . . . + rk xk for ri ∈ R
and xi ∈ X. □
Example 17.10. (1) Let X = {x1 , x2 }. Then the free module on X is the module
We identify x1 with (1x1 , 0) and x2 with (0, 1x1 ). Thus X ⊂ Rx1 ⊕ Rx2 .
If an R-module M is free on a non empty set X then any R-module homomorphism from M to
another R-module N is uniquely determined by images of elements of X. This is because each ele-
ment of M can be uniquely written as a finite linear combination of elements of X with coefficients
in R. In fancy language, this fact is presented in the form of the following theorem.
38
Theorem 17.11. Let X be a non empty set and M be an R-module free on X. Let i : X → M denote
the inclusion map. Let N be another R-module and j : X → N be a set map. Then there exists a
unique R-module homomorphism ϕ : M → N such that the following diagram commutes.
> NO
j
ϕ
X
i / M
Theorem 20.1. Let R be a PID and M = Re1 ⊕. . . Ren be a free module of rank n. Let N be a nonzero
submodule of M. Then N is a free module and there exist a basis { f1 , . . . , fn } of M and nonzero
elements a1 , a2 , . . . , am ∈ R satisfying (a1 ) ⊃ (a2 ) ⊃ . . . ⊃ (am ) such that {a1 f1 , a2 f2 , . . . , am fm } is a
basis of N.
Proof. We prove the result by induction on rank(M). If rank(M) = 1, the proof is clear. Let
πi : M → R be the R-module homomorphism defined by πi (ei ) = 1 and πi (e j ) = 0 for i , j. We
consider the set of ideals
S = {ϕ(N) : ϕ ∈ Hom(M, R)}.
The ring R is Noetherian, so the set S has a maximal element. Let ψ(N) be a maximal element of
S for ψ ∈ Hom(M, R).
Claim - 1 : ψ(N) , 0
Since ψ(N) is a maximal element of S, no ideal belonging to S properly contains ψ(N). The zero
ideal is properly contained in every nonzero ideal. Therefore if we show that S contains a nonzero
ideal, then ψ(N) , 0. It is given that N is a nonzero module. Let x = c1 e1 + . . . + cn en be a nonzero
element of N. Now at least one coefficient ci must be nonzero. We have ci = πi (x) ∈ πi (N). This
shows that πi (N) is a nonzero ideal contained in S. Therefore ψ(N) , 0.
Since the ring R is a PID, we have a nonzero element a1 ∈ R such that ψ(N) = (a1 ). We choose
f ∈ N such that ψ( f ) = a1 . Clearly f , 0. Let f = b1 e1 + . . . + bn en .
Claim - 2 : The element a1 divides each coefficient bi .
Let (a1 , bi ) = (a′1 ), a′1 ∈ R. Then there exist p, q ∈ R such that a′1 = pa1 + qbi . We consider
the homomorphism pψ + qπi ∈ Hom(M, R). Then (pψ + qπi )( f ) = pa1 + qbi = a′1 . Therefore
a′1 ∈ (pψ + qπi )(N). It follows that ψ(N) = (a1 ) ⊂ (a′1 ) ⊂ (pψ + qπi )(N). The maximality of ψ(N)
forces all inclusions to be equality. Therefore (a1 ) = (a′1 ) = (a1 , bi ) which implies that a1 |bi .
Since each coefficient bi of f is divisible by a1 , we write bi = a1 hi , hi ∈ R for i = 1, . . . , n. Let
f1 = h1 e1 + . . . + hn en . Then f1 ∈ M and f = a1 f1 . We have a1 ψ( f1 ) = ψ(a1 f1 ) = ψ( f ) = a1 . Since
a1 , 0 and R is an integral domain, we get ψ( f1 ) = 1.
We have the following short exact sequence
i ψ
0 → ker ψ →
− M→
− R → 0.
The module R is a free on {1} and ψ( f1 ) = 1. Therefore the sequence is a split short exact sequence.
The R-module homomorphism τ : R → M defined by τ(r) = r f1 satisfies ψ ◦ τ = id. By Remark
19.6 we obtain M = ker ψ ⊕ R f1 . By Theorem 19.9, ker ψ is a free module of rank (n − 1).
We also have the following short exact sequence
i ψ
0 → ker ψ ∩ N →
− N→
− Ra1 → 0.
43
The module Ra1 is free on {a1 } and ψ( f ) = ψ(a1 f1 ) = a1 . Therefore this sequence is also a split
short exact sequence and N = (ker ψ ∩ N) ⊕ Ra1 f1 (see Remark 19.6). If ker ψ ∩ N = 0, then
N = Ra1 f1 . Any basis of ker ψ together with f1 give a basis of M. The proof completes here.
We assume that ker ψ ∩ N , 0. The module (ker ψ ∩ N) is a nonzero submodule of the free
module ker ψ and rank(ker ψ) = n−1. By induction hypothesis we get a basis f2 , . . . , fn and nonzero
elements a2 , . . . , am satisfying (a2 ) ⊃ . . . ⊃ (am ) such that {a2 f2 , . . . , am fm } is a basis of ker ψ ∩ N. It
follows that M = ker ψ ⊕ R f1 = R f1 ⊕ . . . ⊕ R fn and N = (ker ψ ∩ N) ⊕ Ra1 f1 = Ra1 f1 ⊕ . . . ⊕ Ram fm .
We already have (a2 ) ⊃ . . . ⊃ (am ). We require to show that (a1 ) ⊃ (a2 ).
Let γ : M → R be defined by γ( fi ) = 1 for i = 1, 2 and 0 otherwise. Then γ(a1 f1 ) = a1 and
γ(a2 f2 ) = a2 . Since a1 f1 , a2 f2 ∈ N, we obtain (a1 , a2 ) ⊂ γ(N). It follows that ψ(N) = (a1 ) ⊂
(a1 , a2 ) ⊂ γ(N). The maximality of ψ(N) in S forces ψ(N) = γ(N), so (a1 ) = (a1 , a2 ). Therefore
(a1 ) ⊃ (a2 ). The proof completes here. □
Theorem 20.2. (Invariant factor form) Let M be a finitely generated module over a PID R. Then
M Rr ⊕ R/(b1 ) ⊕ . . . ⊕ R/(bl )
such that R ⫌ (b1 ) ⊃ . . . ⊃ (bl ) ⫌ (0).
Proof. Since M is finitely generated module over R, we get a surjective R-module homomorphism
ϕ : Rn → M. Let K = ker ϕ. Then M Rn /K. It follows from Theorem 20.1 that Rn = R f1 ⊕. . .⊕R fn
and K = Ra1 f1 ⊕ . . . ⊕ Ram fn such that (a1 ) ⊃ (a2 ) ⊃ . . . ⊃ (am ). Then
M Rn /K R/(a1 ) ⊕ . . . ⊕ R/(am ) ⊕ Rn−m .
If (ai ) = R, then R/(ai ) = 0 and if (ai ) = 0, then R/(ai ) = R. Let (a1 ) = . . . = (ar ) = R and
(ar+l+1 ) = . . . = (am ) = 0. Then we get
M = R/(ar+1 ) ⊕ . . . ⊕ R/(ar+l ) ⊕ Rm−r−l ⊕ Rn−m
= R/(ar+1 ) ⊕ . . . ⊕ R/(ar+l ) ⊕ Rn−r−l
= Rn−r−l ⊕ R/(ar+1 ) ⊕ . . . ⊕ R/(ar+l )
We observe that R ⫌ (ar+1 ) ⊃ . . . ⊃ (ar+l ) ⫌ (0). Therefore the result follows. □
Let p1 , p2 , . . . , pk be prime elements in a PID R such that no two distinct pi ’s are associate in R.
Let a = pl11 pl22 . . . plkk . By Chinese remainder theorem, it follows that as rings
R/(a) R/(pl11 ) × . . . × R/(plkk ).
The above isomorphism is also an R-module homomorphism. Therefore we obtain an R-module
isomorphism
R/(a) R/(pl11 ) ⊕ . . . ⊕ R/(plkk ).
The following result follows from Theorem 20.2 by the above observation.
44
Theorem 20.3. (Elementary divisor form) Let R be a PID and M be a finitely generated R-module.
Then there are prime elements p1 , . . . , pt such that
M Rr ⊕ R/(pα1 ) ⊕ . . . ⊕ R/(pαt )
1 t
45
22. assignments
22.1. Assignment I. Please solve the following problems from the book : Abstract Algebra, Dum-
mit and Foote.
(1) Page no 230. Problems 3, 5, 6, 7, 13, 14, 17, 20, 26, 27.
(2) Page no 237. Problems 2, 3, 4, 5.
(3) Page no 247. Problems 1, 2, 3, 4, 17, 18, 24, 25, 26, 28, 29, 30, 32, 33, 34, 35, 36.
(4) Page no 256. Problems 2, 3, 5, 7, 8, 10, , 12, 13, 14, 19, 21, 25, 28, 33, 34, 36, 40.
(5) Page no 264. Problems 3, 4, 5.
(6) Page no 267. Problems 1, 3, 4, 5, 6.
22.2. Assignment II. Please solve the following problems from the book : Abstract Algebra,
Dummit and Foote.
(1) Page no 277. Problems 1(b), 2(a), 3, 4(a), 10, 11
(2) Page no 282. Problems 1, 2, 3, 5, 6, 8.
Problem 8 is slightly more difficult. One first shows that if I is an ideal of R, then
a
D−1 I = { : a ∈ I, b ∈ D}
b
is an ideal of D R. Then, one shows that every ideal of D−1 R is of the form D−1 I for some
−1
ideal I of R. From this one concludes that all ideals of D−1 R are principal.
(3) Page no 292. Problems 2, 4, 5, 8
(4) Page no 298. Problems 1, 3, 4, 5, 6, 7, 8, 9, 12, 13, 14, 15, 16, 17
In problem 8, show that the ideal (x, x2 y, x3 y2 , . . . , xn yn−1 , . . .) is not finitely generated.
In problem 13, show that (yF[x,y] F[x,y]
2 −x) is a PID but (y2 −x2 ) is a product of two fields.
46