412notes6 PDF
412notes6 PDF
Ideals.
Finally we are ready to study kernels and images of ring homomorphisms. We have
seen two major examples in which congruence gave us ring homomorphisms: Z → Zn and
F [x] → F [x]/(p(x)). We shall generalize this to congruence in arbitrary rings and then
see that it brings us very close to a complete understanding of kernels and images of ring
homomorphisms.
Recall the definition of a ring. For congruence, we need a special subring that will
behave like nZ or like p(x)F [x] = { p(x)f (x) | f (x) ∈ F [x] }.
Before we look at examples, recall that to be a subring means I is closed under multi-
plication and subtraction. Thus we get
Examples 1, 2 and 3 above were all of a special type which we can generalize.
Theorem 6.2. Let R be a commutative ring with identity. Let c ∈ R. The set
I = {rc | r ∈ R } is an ideal of R.
We call the ideal in Theorem 6.2 the principal ideal generated by c and denote it
by (c) or by Rc. The ideal in Example 4 was not principal. To see this, note that n ≥ 2
and x both lie in I. If I were generated by some polynomial p(x), then both n and x must
be multiples of p(x). But then n = p(x)q(x) implies that p(x) is a constant c. Note that
c 6= ±1, for that would make I the whole ring, which it is not since 1 ∈ / I. Now we also
have x = cr(x) for some r(x), which is impossible since c does not divide x (i.e., c has no
inverse in Z[x] since the only units are ±1). In fact it is easy to see that I is generated by
the two elements n and x in the sense of the next theorem.
We call the ideal I of Theorem 6.3 the ideal generated by c1 , . . . , cn and denote it
by (c1 , c2 , . . . , cn ).
In analogy to congruence in Z and F [x] we now will build a ring R/I for any ideal I in
any ring R. For a, b ∈ R, we say a is congruent to b modulo I [and write
a ≡ b (mod I)] if a − b ∈ I. Note that when I = (n) ⊂ Z is the principal ideal generated
by n, then a − b ∈ I ⇐⇒ n|(a − b), so this is our old notion of congruence.
Theorem 6.5. Let I be an ideal of a ring R. If a ≡ b (mod I) and c ≡ d (mod I), then
(1) a + c ≡ b + d (mod I);
(2) ac ≡ bd (mod I).
Looking at this proof, we see that it is multiplication that fails if we have only a left
or right ideal that is not 2-sided. The equivalence classes for this relation, are commonly
called cosets. What do they look like? The congruence class of a modulo I is
{ b ∈ R | b ≡ a (mod I) } = { b ∈ R | b − a ∈ I }
= { b ∈ R | b − a = i for some i ∈ I }
= { b ∈ R | b = a + i for some i ∈ I }
= { a + i | i ∈ I }.
14/35. A commutative ring R with 1 is a field iff its only two ideals are (0) and R.
Proof. ( =⇒ ) Any nonzero ideal I contains some nonzero element, which is a unit since
R is a field. By #13, I = R.
( ⇐= ) Let 0 6= a ∈ R and let I = (a). By hypothesis, I = R, so I contains the identity 1.
Therefore 1 = ra for some r ∈ R, so that r is the inverse of a. Therefore R is a field.
Proof. Assume I 6= (0) (which is principal). Let c be the smallest positive element in I
(exists by the well-ordering axiom). Then (c) ⊆ I. Conversely, let a ∈ I. By the division
algorithm, we can write a = cq + r with 0 ≤ r < c. Then r = a − cq ∈ I. By our choice of
c, we must have r = 0, as otherwise it is a smaller positive element of I. Therefore a ∈ (c),
so I = (c) is principal.
39. (a) S = { m
n | m, n ∈ Z, n odd } is a subring of Q.
m
(b) I = { n ∈ S | m even } is an ideal in S.
(c) S/I has exactly two cosets.
Theorem 6.5 gives the fact that addition and multiplication are well-defined on congru-
ence classes. Translating this into the language of cosets gives
Proof. The fact that π preserves addition and multiplication follows from the definition of
addition and multiplication in R/I. It is surjective since any coset r + I is the image of
r ∈ R. Finally, the kernel is the set of all r ∈ R such that π(r) = 0 + I, the zero element
of R/I. But r + I = 0 + I iff r ≡ 0 (mod I) iff r ∈ I. Thus the kernel is just I.
We can generalize this idea of ideals and kernels to any ring homomorphism.
Theorem 6.10. Let f : R → S be any homomorphism of rings and let K = ker f . Then
K is an ideal in R.
And furthermore, the kernel tests for injectivity just as it does for linear transformations.
For any s ∈ S, we know there is some r ∈ R with f (r) = s, and therefore φ(r + K) = s
showing that φ is surjective. To show that φ is injective, we show that ker φ is zero in R/K:
if φ(r + K) = 0, then f (r) = 0, so r ∈ K, hence r + K = 0 + K. Therefore φ : R/K → S
is an isomorphism.
The 2nd and 3rd isomorphism theorems are left to Math 413.
2. We saw that the ideals in Z all look like (n) for some integer n ≥ 0. Thus the
homomorphic images are either isomorphic to Z itself (when n = 0) or to Z/(n) = Zn
(when n > 0).
Exercise 8(a), p. 151. Let I = { 0, 3 } in Z6 . I is an ideal since it is the principal ideal (3).
Z6 /I ∼
= Z3 via the mapping n + I 7→ [n]3 as this is the only possible homomorphism since
1 must map to 1; there are several details to check. We define a function φ : Z6 → Z3 by
φ([n]6 ) = [n]3 and check that it is well-defined. It is clearly surjective and one can check
that it is a homomorphism (see Example 1, p. 7 of Chapter 3 notes). Now check that
I = ker φ and use the First Isomorphism Theorem. This exercise is a special case of the
Third Isomorphism Theorem (see Exercise 33).
Exercise 19, p. 152. Let I, J be ideals in R and define f : R → R/I × R/J by f (a) =
(a + I, a + J). [For an example, again see Example 1, p. 7 of Chapter 3 notes.]
(a) f is a homomorphism: f (a + b) = (a + b + I, a + b + J) = (a + I, a + J) + (b + I, b + J) =
f (a) + f (b) and f (ab) = (ab + I, ab + J) = (a + I, a + J)(b + I, b + J) = f (a)f (b).
(b) It was surjective in the example mentioned above: Z6 → Z6 /(3) × Z6 /(2) ∼ = Z2 × Z3 .
On the other hand, for Z → Z/(2) × Z/(4), nothing maps onto (1, 0) since n ≡ 1 (mod 2)
implies n 6≡ 0 (mod 4).
(c) Check that ker f = I ∩ J.
The Chinese Remainder Theorem in Chapter 13 deals with the issue of when the mapping
is surjective.
Prime ideals.
What is special about an ideal I for which R/I is an integral domain or a field? For this
section we assume that R is a commutative ring with identity, since these are necessary
conditions to hope to have an integral domain. We want a condition on the ideal to make
sure there are no zero divisors in the quotient ring. We have seen several examples in the
case of principal ideals—what we needed was that the generator was irreducible (called
prime in the case of Z). The appropriate generalization is one of the equivalent forms we
had for an irreducible polynomial or integer.
Examples. 1. Theorem 4.11 says (p(x)) in F [x] is a prime ideal iff p(x) is irreducible.
And we have seen that F [x]/(p(x)) is a field iff p(x) is irreducible (Theorem 5.10).
2. Theorem 1.8 says (p) in Z is a prime ideal iff p is prime. And we have seen that Zp
is a field iff p is prime (Theorem 2.8).
4. For a nonprincipal
P consider the ideal P = (p, x) in Z[x] where p is prime.
ideal, P
i
Assume f (x) = ai x , g(x) = bj xj ∈ Z[x] with f (x)g(x) ∈ P . This says the constant
term a0 b0 is divisible by p. But then either p|a0 (and so f (x) ∈ P ) or p|b0 (and so
g(x) ∈ P ). Therefore P is a prime ideal. In this case, the quotient ring is Zp , a field.
Theorem 6.14. Let P be an ideal in R. P is a prime ideal iff R/P is an integral domain.
How much more do we need to assume to have R/P be a field? Our main example was
(4) and (5) above: Z[x] modulo (x) was an integral domain, but modulo the larger ideal
(p, x) it was a field. So it helps to have big ideals.
The same definition is used in noncommutative rings with 1. The examples (1), (2),
and (4) above were maximal. And we saw that in each case, the quotient ring was a field.
19. The proof that every ideal I 6= R is contained in a maximal ideal uses Zorn’s lemma
(a fact equivalent to the axiom of choice in set theory).
Assume R has a unique maximal ideal. Let I be the set of nonunits in R. For r ∈ R,
a ∈ I, the products ar and ra are also nonunits, so they lie in I. Assume a, b ∈ I and that
a + b is not in I. Then a + b is a unit, hence there exists u ∈ R with (a + b)u = 1. But
10
(a) and (b) are ideals smaller than R, hence are contained in maximal ideals. Since there
is only one maximal ideal, say M , we have a ∈ M , b ∈ M and therefore 1 = u(a + b) ∈ M ,
a contradiction. Therefore a + b is in I and I is an ideal. Conversely, assume the set I of
all nonunits is an ideal. Then it is clearly the unique maximal ideal of R since no other
element can be in a maximal ideal other than a nonunit.
Example: { m n ∈ Q | n is odd }. (2) is the unique maximal ideal.