Valuation Rings: 3.1 Extension Theorems
Valuation Rings: 3.1 Extension Theorems
Valuation Rings: 3.1 Extension Theorems
Valuation Rings
The results of this chapter come into play when analyzing the behavior of a rational
function defined in the neighborhood of a point on an algebraic curve.
3.1.1 Theorem
Let R be a subring of the field K, and h : R → C a ring homomorphism from R into an
algebraically closed field C. If α is a nonzero element of K, then either h can be extended
to a ring homomorphism h : R[α] → C, or h can be extended to a ring homomorphism
h : R[α−1 ] → C.
Proof. Without loss of generality, we may assume that R is a local ring and F = h(R) is
a subfield of C. To see this, let P be the kernel of h. Then P is a prime ideal, and we can
extend h to g : RP → C via g(a/b) = h(a)/h(b), h(b) = 0. The kernel of g is P RP , so
by the first isomorphism theorem, g(RP ) ∼ = RP /P RP , a field (because P RP is a maximal
ideal). Thus we may replace (R, h) by (RP , g).
Our first step is to extend h to
a homomorphism of polynomial rings. If f ∈ R[x] with
f (x) = ai xi , we take h(f ) = h(ai )xi ∈ F [x]. Let I = {f ∈ R[x] : f (α) = 0}. Then
J = h(I) is an ideal of F [x], necessarily principal. Say J = (j(x)). If j is nonconstant,
it must have a root β in the algebraically closed field C. We can then extend h to
h : R[α] → C via h(α) = β, as desired. To verify that h is well-defined, suppose f ∈ I, so
that f (α) = 0. Then h(f ) ∈ J, hence h(f ) is a multiple of j, and therefore h(f )(β) = 0.
Thus we may assume that j is constant. If the constant is zero, then we may extend h
exactly as above, with β arbitrary. So we can assume that j = 0, and it follows that
1 ∈ J. Consequently, there exists f ∈ I such that h(f ) = 1.
1
2 CHAPTER 3. VALUATION RINGS
Choose r as small as possible. We then carry out the same analysis with α replaced by
α−1 . Assuming that h has no extension to R[α−1 ], we have
s
1, i = 0
−i
bi α = 0 with bi ∈ R and bi = h(bi ) = (2)
i=0
0, i > 0
Take s minimal, and assume (without loss of generality) that r ≥ s. Since h(b0 ) = 1 =
h(1), it follows that b0 − 1 ∈ ker h ⊆ M, the unique maximal ideal of the local ring R.
Thus b0 ∈ / M (else 1 ∈ M), so b0 is a unit. It is therefore legal to multiply (2) by b−1
0 α
s
to get
αs + b−1
0 b1 α
s−1
+ · · · + b−1
0 bs = 0 (3)
Finally, we multiply (3) by ar αr−s and subtract the result from (1) to contradict the
minimality of r. (The result of multiplying (3) by ar αr−s cannot be a copy of (1). If so,
r = s (hence αr−s = 1)and a0 = ar b−1 −1
0 bs . But h(a0 ) = 1 and h(ar b0 bs ) = 0.) ♣
It is natural to try to extend h to a larger domain, and this is where valuation rings
enter the picture.
3.1.2 Definition
A subring R of a field K is a valuation ring of K if for every nonzero α ∈ K, either α or
α−1 belongs to R.
3.1.3 Examples
The field K is a valuation ring of K, but there are more interesting examples.
1. Let K = Q, with p a fixed prime. Take R to be the set of all rationals of the form
pr m/n, where r ≥ 0 and p divides neither m nor n.
2. Let K = k(x), where k is any field. Take R to be the set of all rational functions
pr m/n, where r ≥ 0, p is a fixed polynomial that is irreducible over k and m and n
are arbitrary polynomials in k[x] not divisible by p. This is essentially the same as the
previous example.
3. Let K = k(x), and let R be the set of all rational functions f /g ∈ k(x) such that
deg f ≤ deg g.
4. Let K be the field of formal Laurent series over k. Thus a nonzero element of K looks
∞
like f = i=r ai xi with ai ∈ k, r ∈ Z, and ar = 0. We may write f = ar xr g, where
g belongs to the ring R = k[[x]] of formal power series over k. Moreover, the constant
term of g is 1, and therefore g, hence f , can be inverted (by long division). Thus R is a
valuation ring of K.
We now return to the extension problem.
3.2. PROPERTIES OF VALUATION RINGS 3
3.1.4 Theorem
Let R be a subring of the field K, and h : R → C a ring homomorphism from R into an
algebraically closed field C. Then h has a maximal extension (V, h). In other words, V is
a subring of K containing R, h is an extension of h, and there is no extension to a strictly
larger subring. In addition, for any maximal extension, V is a valuation ring of K.
Proof. Let S be the set of all (Ri , hi ), where Ri is a subring of K containing R and hi
is an extension of h to Ri . Partially order S by (Ri , hi ) ≤ (Rj , hj ) if and only if Ri is a
subring of Rj and hj restricted to Ri coincides with hi . A standard application of Zorn’s
lemma produces a maximal extension (V, h). If α is a nonzero element of K, then by
(3.1.1), h has an extension to either V [α] or V [α−1 ]. By maximality, either V [α] = V or
V [α−1 ] = V . Therefore α ∈ V or α−1 ∈ V . ♣
αn + cn−1 αn−1 + · · · + c1 α + c0 = 0
with the ci in V . We must show that α ∈ V . If not, then α−1 ∈ V , and if we multiply
the above equation of integral dependence by α−(n−1) , we get
5. If I and J are ideals of V , then either I ⊆ J or J ⊆ I. Thus the ideals of V are totally
ordered by inclusion.
Suppose that I is not contained in J, and pick a ∈ I \ J (hence a = 0). If b ∈ J, we
must show that b ∈ I. If b = 0 we are finished, so assume b = 0. We have b/a ∈ V (else
a/b ∈ V , so a = (a/b)b ∈ J, a contradiction). Therefore b = (b/a)a ∈ I.
4 CHAPTER 3. VALUATION RINGS
6. Conversely, let V be an integral domain with fraction field K. If the ideals of V are
totally ordered by inclusion, then V is a valuation ring of K.
If α is a nonzero element of K, then α = a/b with a and b nonzero elements of V . By
hypothesis, either (a) ⊆ (b), in which case a/b ∈ V , or (b) ⊆ (a), in which case b/a ∈ V .
7. If P is a prime ideal of the valuation ring V , then VP and V /P are valuation rings.
First note that if K is the fraction field of V , it is also the fraction field of VP . Also, V /P
is an integral domain, hence has a fraction field. Now by Property 5, the ideals of V are
totally ordered by inclusion, so the same is true of VP and V /P . The result follows from
Property 6.
8. If V is a Noetherian valuation ring, then V is a PID. Moreover, for some prime p ∈ V ,
every ideal is of the form (pm ), m ≥ 0. For any such p, ∩∞ m
m=1 (p ) = 0.
3.3.2 Examples
We can place a discrete valuation on all of the fields of Subsection 3.1.3. In Examples
∞ 2, wei take v(p m/n) = r. In Example 3, v(f /g) = deg g − deg f . In Example 4,
r
1 and
v( i=r ai x ) = r (if ar = 0).
3.3.3 Proposition
If v is a discrete valuation on the field K, then V = {a ∈ K : v(a) ≥ 0} is a valuation
ring with maximal ideal M = {a ∈ K : v(a) ≥ 1}.
Proof. The defining properties (a), (b) and (c) of (3.3.1) show that V is a ring. If a ∈
/ V,
then v(a) < 0, so v(a−1 ) = v(1) − v(a) = 0 − v(a) > 0, so a−1 ∈ V , proving that V is a
valuation ring. Since a is a unit of V iff both a and a−1 belong to V iff v(a) = 0, M is
the ideal of nonunits and is therefore the maximal ideal of the valuation ring V . ♣
3.3.5 Proposition
Let t be a uniformizer in the discrete valuation ring V . Then t generates the maximal
ideal M of V , in particular, M is principal. Conversely, if t is any generator of M, then
t is a uniformizer.
Proof. Since M is the unique maximal ideal, (t) ⊆ M. If a ∈ M, then v(a) ≥ 1, so
v(at−1 ) = v(a) − v(t) ≥ 1 − 1 = 0, so at−1 ∈ V , and consequently a ∈ (t). Now suppose
M = (t ). Since t ∈ M, we have t = ct for some c ∈ V . Thus
3.3.6 Proposition
If t is a uniformizer, then every nonzero element a ∈ K can be expressed uniquely as
a = utn where u is a unit of V and n ∈ Z. Also, K = Vt , that is, K = S −1 V where
S = {1, t, t2 , . . . }.
Proof. Let n = v(a), so that v(at−n ) = 0 and therefore at−n is a unit u. To prove
uniqueness, note that if a = utn , then v(a) = v(u) + nv(t) = 0 + n = n, so that n, and
hence u, is determined by a. The last statement follows by Property 1 of Section 3.2 and
the observation that the elements of V are those with valuation n ≥ 0. ♣
A similar result holds for ideals.
3.3.7 Proposition
Every nonzero proper ideal I of the DVR V is of the form Mn , where M is the maximal
ideal of V and n is a unique nonnegative integer. We write v(I) = n; by convention,
M0 = V .
Proof. Choose a ∈ I such that n = v(a) is as small as possible. By (3.3.6), a = utn , so
tn = u−1 a ∈ I. By (3.3.5), M = (t), and therefore Mn ⊆ I. Conversely, let b ∈ I, with
v(b) = k ≥ n by minimality of n. As in the proof of (3.3.6), bt−k is a unit u , so b = u tk .
Since k ≥ n we have b ∈ (tn ) = Mn , proving that I ⊆ Mn . The uniqueness of n is a
consequence of Nakayama’s lemma. If Mr = Ms with r < s, then Mr = Mr+1 = MMr .
Thus Mr , hence M, is 0, contradicting the hypothesis that I is nonzero. ♣
We may interpret v(I) as the length of a composition series.
3.3.8 Proposition
Let I be a nonzero proper ideal of the discrete valuation ring R. Then v(I) = lR (R/I),
the composition length of the R-module R/I.
3.3. DISCRETE VALUATION RINGS 7
R/I ⊃ M/I ⊃ M2 /I ⊃ · · · ⊃ Mn /I = 0.
R/I M/I
l(R/I) = l( ) + l(M/I) = l(R/M) + l( 2 ) + l(M2 /I).
M/I M /I
n−1
l(R/I) = l(Mi /Mi+1 ).
i=0
3.3.9 Proposition
Let
√ I be an ideal of the Noetherian ring R. Then for some positive integer m, we have
( I)m ⊆ I. In particular (take I = 0), the nilradical of R is nilpotent.
√
Proof. √Since R is Noetherian, I is finitely generated, say by a1 , . . . , at , with ani i ∈ I.
t
Then ( I)m is generated by all products ar11 · · · art t with i=1 ri = m. Our choice of m
is
t
m=1+ (ni − 1).
i=1
t
t
m= ri < 1 + (ni − 1) = m,
i=1 i=1
√
a contradiction. But then each product ar11 · · · art t is in I, hence ( I)m ⊆ I. ♣
3.3.10 Proposition
Let M be a maximal ideal of the Noetherian ring R, and let Q be any ideal of R. The
following conditions are equivalent:
√ is M-primary.
1. Q
2. Q = M.
3. For some positive integer n, we have Mn ⊆ Q ⊆ M.
8 CHAPTER 3. VALUATION RINGS
Proof. We have (1) implies (2) by definition of M-primary; see (1.1.1). The implication
(2) ⇒ (1) follows from (1.1.2). To prove that (2) implies (3), apply (3.3.9) with I = Q to
get, for some positive integer n,
Mn ⊆ Q ⊆ Q = M.
To prove that (3) implies (2), observe that by (1.1.1),
√ √
M = Mn ⊆ Q ⊆ M = M. ♣
Now we can characterize discrete valuation rings.
3.3.11 Theorem
Let R be a Noetherian local domain with fraction field K and unique maximal ideal
M = 0. (Thus R is not a field.) The following conditions are equivalent:
1. R is a discrete valuation ring.
2. R is a principal ideal domain.
3. M is principal.
4. R is integrally closed and every nonzero prime ideal is maximal.
5. Every nonzero ideal is a power of M.
6. The dimension of M/M2 as a vector space over R/M is 1.
Proof.
(1) ⇒ (2): This follows from (3.3.7) and (3.3.5).
(2) ⇒ (4): This holds because a PID is integrally closed, and a PID is a UFD in which
every nonzero prime ideal is maximal.
(4) ⇒ (3): Let t be a nonzero element of M. By hypothesis, M is the only nonzero
prime ideal, so the radical of (t), which is the intersection of all prime ideals containing
t, coincides with M. By (3.3.10), for some n ≥ 1 we have Mn ⊆ (t) ⊆ M, and we
may assume that (t) ⊂ M, for otherwise we are finished. Thus for some n ≥ 2 we have
Mn ⊆ (t) but Mn−1 ⊆ (t). Choose a ∈ Mn−1 with a ∈ / (t), and let β = t/a ∈ K. If
β −1 = a/t ∈ R, then a ∈ Rt = (t), contradicting the choice of a. Therefore β −1 ∈ / R.
Since R is integrally closed, β −1 is not integral over R. But then β −1 M ⊆ M, for if
β −1 M ⊆ M, then β −1 stabilizes a finitely generated R-module, and we conclude from
the implication (4) ⇒ (1) in (2.1.4) that β −1 is integral over R, a contradiction.
Now β −1 M ⊆ R, because β −1 M = (a/t)M ⊆ (1/t)Mn ⊆ R. (Note that a ∈ Mn−1
and Mn ⊆ (t).) Thus β −1 M is an ideal of R, and if it were proper, it would be contained
in M, contradicting β −1 M ⊆ M. Consequently, β −1 M = R, hence M is the principal
ideal (β).
(3) ⇒ (2): By hypothesis, M is a principal ideal (t), and we claim that ∩∞ n=0 M = 0.
n
bn+1 tn+1 , hence bn = bn+1 t. Thus (bn ) ⊆ (bn+1 ) for all n, and in fact (bn ) = (bn+1 ) for
sufficiently large n because R is Noetherian. Therefore bn = bn+1 t = ctbn for some c ∈ R,
so (1 − ct)bn = 0. But t ∈ M, so t is not a unit, and consequently ct = 1. Thus bn must
be 0, and we have a = bn tn = 0, proving the claim.
Now let I be any nonzero proper ideal of R. Then I ⊆ M, but by the above claim
we have I ⊆ ∩∞ n=0 M . Thus there exists n ≥ 0 such that I ⊆ M and I ⊆ M
n n n+1
.
3.3. DISCRETE VALUATION RINGS 9
3.3.12 Corollary
The ring R is a discrete valuation ring if and only if R is a local PID that is not a field.
In particular, since R is a PID, it is Noetherian.
Proof. The “if” part follows from (2) implies (1) in (3.3.11). For the “only if” part, note
that a discrete valuation ring R is a PID by (1) implies (2) of (3.3.11); the Noetherian
hypothesis is not used here. Moreover, R is a local ring by Property 3 of Section 3.2. If R
is a field, then every nonzero element a ∈ R is a unit, hence v(a) = 0. Thus the valuation
v is trivial, a contradiction. ♣
3.3.13 Corollary
Let R be a DVR with maximal ideal M. If t ∈ M \ M2 , then t is a uniformizer.
Proof. This follows from the proof of (5) implies (3) in (3.3.11). ♣