Dedekind Domains
In this chapter we shall study the domains in which proper non-zero ideals can be factored in one and only one way as products of prime ideals. The most notable examples are the rings of -integral elements of number fields, that is, finite dimensional extension fields of the field (see BAI, pp. 278281), These are the objects of study of algebraic number theory. Another important class of examples are the rings that occur in the study of algebraic curves. Here we begin with a field F(x,y) where F is a base field (usually algebraically closed), x is transcendental over F, and y is algebraic over F(x). Then the subring of elements that are integral over F[x] has the factorization property stated above.
There are many equivalent ways of defining the class of domains, called Dedekind domains, in which the fundamental factorization theorem for ideals into prime ideals holds (see section 10.2). We shall take as our point of departure a definition based on the concepts of fractional ideals and of invertibility. The latter is equivalent to projectivity as a module for the given ring.
The result that the domains mentioned above are Dedekind can be deduced from a general theorem stating that if D is Dedekind with field of fractions F and E is a finite dimensional extension field of F, then the subring D′ of D-integral elements of E is a Dedekind domain. This is proved in section 10.3. In sections 10.4 and 10.5 we consider the central problem of studying the factorization in D′ of extensions of prime ideals of D. This is closely related to the problem of extension of valuations from F to E that was considered in chapter 9.
Besides the study of ideal theory, we consider the structure of finitely generated modules over Dedekind domains. The special case of torsion-free modules is a classical one that was first treated by E. Steinitz (see section 10.6).
Finally, we consider the class group of a Dedekind domain as defined in section 7.9 and we show in section 10.6 that this group has a concrete realization as a classical group defined by the fractional ideals.
10.1 FRACTIONAL IDEALS. DEDEKIND DOMAINS
Let D be a domain, F its field of fractions. It is useful to extend the concept of an ideal in D to certain submodules of F given in the following
DEFINITION 10.1. If D is a domain and F is its field of fractions, a (D –) fractional ideal I is a non-zero D-submodule of F such that there exists a non-zero a in D such that aI ⊂ D(or, equivalently, aI ⊂ I ∩ D).
The fractional ideals contained in D are the non-zero ideals of D. These are called integral ideals. If I is a fractional ideal and a is a non-zero element of D such that aI ⊂ D, then aI is an integral ideal. If b is any non-zero element of F, then Db = {db\d ∈ D} is a fractional ideal since it is clearly a D-submodule of F and if b = ac− 1, a, c ∈ D, then c(Db) ⊂ D. A fractional ideal of the form Db is called a principal ideal.
If I1 and I2 are fractional ideals, then so is the module sum I1 + I2 since this is a submodule, and if aiIi ⊂ D for ai ≠ 0 in D, then a1a2(I1 + I2) ⊂ D. The intersection I1 ∩ I2 is a fractional ideal also since this is a D-submodule and I1 ∩ I2 ≠ 0 since if bi ≠ 0 is in Ii we have an ai ≠ 0 in D such that aibi ∈ D ∩ Ii. Then (a1 b1)(a2 b2) ≠ 0 is in I1 ∩ I2. Moreover, if a ≠ 0 satisfies aI1 ⊂ D, then a(I1 ∩ I2) ⊂ D. We define I1 I2 = {∑ b1i b2i|b1i ∈ I1, b2i ∈ I2}. Evidently this is a D-submodule of F and the argument used for I1 ∩ I2 shows that I1 I2 is a fractional ideal.
We shall need to know what the homomorphisms of a fractional ideal into D look like. This information is given in
PROPOSITION 10.1. Let f be a D-homomorphism of the fractional ideal I into D and let b be any non-zero element of I. Then f has the form
a ∈ I.
Proof. Choose a d ≠ 0 in D such that dI ⊂ D. Then db ∈ D and da ∈ D for any a ∈ I. We have
Hence af(b) = bf(a) and so f(a) = b − 1 f(b)a. Thus f has the form given in (1).
If I is a fractional ideal, we define
It is clear that I − l is a D-submodule of F. Since there exist a ≠ 0 in D such that aI ⊂ D, it is clear that I − 1 ≠ 0. Moreover, if b ≠ 0 is in I ∩ D, then I − lb ⊂ D. Hence I − 1 is a fractional ideal. If I is the principal ideal Db, b ≠ 0, then it is clear that I−1 = Db−1. It is clear also that I1 ⊂ I2 I1−1 ⊃ I2−1 .
If I is a fractional ideal, then I− 1 I ⊂ D so I− 1I is an integral ideal. We shall now call I invertible if
If I = bD ≠ 0, then I − 1 = b − lD. Evidently I − lI = (b − 1D)(bD) = D. Hence every principal ideal is invertible. The fractional ideals constitute a commutative monoid under multiplication with D as the unit. If I is invertible, then I − 1I = D so I − 1 is the inverse of I in the monoid of fractional ideals. Conversely, let I have the inverse J in this monoid. Then JI = D implies that J ⊂ I − 1 as defined by (2). Since I − 1 I ⊂ D, we have I − 1I = D. Since the inverse of an element of a monoid is unique, we have I − 1 = J. Thus the invertible ideals constitute the group of units of the monoid of fractional ideals. We prove next
PROPOSITION 10.2. Any invertible ideal is finitely generated.
Proof. Let I be invertible. Then I − 1I = D implies the existence of elements cι ∈ I − 1 , bi ∈ I such that ∑ n1 cibi = 1. Now let b ∈ I. Then b = bl = ∑(bci)bi and ai = bci ∈ D. Thus I = ∑ Dbi and (bl ,…, bn) is a set of generators for I.
Next we establish the equivalence of the conditions of invertibility and projectivity for fractional ideals:
PROPOSITION 10.3. A fractional ideal I is invertible if and only if it is projective.
Proof. First suppose that I is invertible. Then 1 = ∑ m1 cibi, ci ∈ I − 1, bi ∈ I. The map fi : a aci, a ∈ I, is in homD(I,D) and a = ∑(aci)bi = ∑fi(a)bi. Hence I is D-projectiveby the “dual basis lemma” (Proposition 3.11, p. 152). Conversely, suppose that I is projective. Then we have a set of generators {bα} of I and corresponding maps fα ∈ hom D(I, D) such that for any a ∈ I, fα(a) = 0 for all but a finite number of α and a = ∑ fα(a)bα. We have shown in Proposition 10.1 that if b is a non-zero element of I, then fα has the form a b −1 fα(b)a. Since fα(b) = 0 for all α except, say, α = 1, 2, …, m, we have fα = 0 for α ≠ 1, 2, …, m and a = ∑m1 b −1 fi(b)abi. Then 1 = ∑ b − 1 fi(b)bi and since b −1 fi(b)a ∈ D for all a ∈ I, ci = b −1 fi(b)∈ I −1. Thus 1 = ∑ cibi, ci ∈ I −1, bi ∈ I, and hence I −1I = D. Thus I is invertible.
We say that a fractional ideal I is a divisor of the fractional ideal J if there exists an integral ideal K such that J = IK. Since K ⊂ D and Dl ⊂ I, this implies that J ⊂ I. If I is invertible we have the important fact that the converse holds: If I ⊃ J, then I is a divisor of J. For, D = I − 1I ⊃ I − 1J, so k = I − 1J is integral. Moreover, J = I(I − 1J) = IK.
We are interested in existence and uniqueness of factorization of ideals into prime ideals. For invertible integral ideals we have the following uniqueness property.
PROPOSITION 10.4. Let I be integral and invertible and suppose that I = P1 P2 … Pm where the Pi are prime ideals of D. Then this is the only factorization of I as a product of prime ideals.
Proof. We remark first that if M is a commutative monoid and a is a unit in M, then any factor of a is a unit. Applying this remark to the multiplicative monoid of fractional ideals we see that the P i are invertible. Now let I = Q1 Q2 … Qn where the Qj are prime ideals. Then P1 ⊃ Q1 … Qn and since P1 is prime, we may assume that P1 ⊃ Q1. Since P1 is invertible, we have Q1 = P1 R1 where R1 is integral. Then R1 ⊃ Q1. Since Q1 is prime, either P1 ⊂ Q1 or R1 ⊂ Q1, so either P1 = Q1 or R1 = Q1. In the latter case R1 = P1 R1 and since I = Q1Q2 … Qn = P1 R1 Q2 … Qn, R1 is invertible. Then D = R1 − 1 R1 = R1 − 1 R1 P1 = P1, contrary to the hypothesis that P1 is prime (hence proper). Thus Q = P1, so we have P1 P2 … Pm = P1 Q2 … Qn and multiplication by P−11 gives D = Q2 … Qn if m = 1 and P2 … Pm = Q2 … Qn if m > 1. In the first case we have n = 1 also, and in the second n > 1 and induction on m can be applied since P2 … Pm is invertible.
We shall now define the class of rings that will concern us in the chapter.
DEFINITION 10.2. A domain D is called a Dedekind domain if every D-fractional ideal of F (the field of fractions of D) is invertible.
By Proposition 10.3, this is equivalent to assuming that every fractional ideal is projective. It is clear from Proposition 10.2 that any Dedekind domain is noetherian. We have also the following important property:
PROPOSITION 10.5. Any prime ideal ≠ 0 in a Dedekind domain is maximal.
Proof. Let I be a non-zero prime ideal ( = integral prime ideal) of the Dedekind domain D. If I is not maximal, we have an ideal J of D such that D J I. Then I = JK where K is integral. Since J ≠ I, we have I J. Also I K since KK − 1 = DK ⊃ JK = I, so I ⊃ K implies I = K and K = JK. Then D = KK− l = JKK − 1 = J, contrary to D ≠ J. Thus we have I = JK with J I and K I. This contradicts the assumption that I is prime.
We have the following fundamental factorization theorem for Dedekind domains:
THEOREM 10.1. Every proper integral ideal of a Dedekind domain can be written in one and only one way as a product of prime ideals.
Proof. In view of Proposition 10.4, all we have to do is prove that if I is a proper integral ideal in the Dedekind domain D then I is a product of prime ideals ≠ 0. Suppose that this is not the case, so the set of proper integral ideals that are not products of prime ideals is not vacuous. Then by the noetherian property of D this set contains a maximal element I. Then I is not prime and hence I is not maximal. Then there exists an ideal I1 in D such that D I1 I. Then I = I1 I2 where I2 is a proper integral ideal. We have I2 ⊃ I, and I2 = I implies that D = II−1 = I1 I2 I2 − 1 = I1. Thus D Ii I for i = 1, 2. By the maximality of I, Ii is a product of prime ideals. This gives the contradiction that I = I1 I2 is a product of prime ideals.
An immediate consequence of this result is
COROLLARY 1. Suppose that I = P1 P2 … Pm where the Pi are primes. Then the integral ideals ≠ D containing I have the form Pi1 Pi2 … Pir where 1 ≤ i1 < i2 < … < ir ≤ m.
The proof is clear.
It is clear also that we have the following consequence of Corollary 1.
COROLLARY 2. If I is an integral ideal in a Dedekind domain D, then D/I is artinian.
It is evident that the fractional ideals of a Dedekind domain constitute a group under multiplication with D as the unit and I − 1 as defined by (2) as the inverse of I. The fundamental factorization theorem gives the structure of this group:
THEOREM 10.2. The group of fractional ideals of a Dedekind domain is a direct product of the cyclic subgroups generated by the prime ideals. These are infinite.
Proof. Let I be a fractional ideal ≠ D and let a be a non-zero element of D such that aI ⊂ D. Then aI is an integral ideal, so either aI = D or this is a product of prime ideals. Also aD is an integral ideal, so either aD = D or aD is a product of prime ideals. Since I(aD) = aI, I = (aI)(aD)− 1. Then it is clear that we can write I = P1k1 P2k2 … Prkr where the Pi are prime ideals and the ki are non-zero integers. It follows from Theorem 10.1 that, conversely, if the ki are non-zero integers and the Pi are prime ideals, then P1k1 P2k2 … Prkr ≠ D. This implies that the representation of a fractional ideal ≠ D as a product P1k1 P2k2 … Prkr is unique. Hence the group of fractional ideals is the direct product of the cyclic subgroups generated by the primes. It is clear also that these are infinite groups.
EXERCISES
1. Show that any finitely generated D-submodule of F ≠ 0 is a fractional ideal.
2. Show that if D is Dedekind, then the only fractional ideal I satisfying I 2 = I is I = D.
3. Use Theorem 10.1 to prove that any p.i.d. is factorial.
4. Give an example of a factorial domain that is not Dedekind.
5. Let D = [], the subring of F = () of elements a + b , a, b ∈ . Let I be the fractional D-ideal D + Dw where w = − + (a primitive cube root of 1). Compute I − 1 and II − 1. Is I invertible?
6. Let D = [] with quotient field F = []. Show that I = D3 + D is a projective D-module that is not free.
7. Show that any Dedekind domain that is factorial is a p.i.d.
In the remaining exercises, D is a Dedekind domain.
8. Let I1 and I2 be D-integral ideals and write where Pi are distinct prime ideals and the ejk ≥ 0. Show that I1 + I2 = p1min(el1, e21) … Pgmin(e1g, e2g) and I1 ∩ I2 = P1max(e11, e21)… Pgmax(e1g,e2g).
9. Show that the lattice of D-integral ideals is distributive (cf. BAI, p. 463).
10. (Chinese remainder theorem for Dedekind domains.) Let Il. …, In be integral ideals in a Dedekind domain D, al, … ,an elements of D. Show that the system of congruences x ≡ aj (mod Ij), 1 ≤ j ≤ n, has a solution x = a in D if and only if for any j, k, 1 ≤ j, k ≤ n, we have aj ≡ ak (mod Ij + Ik). (Hint: The necessity of the conditions is clear. The sufficiency is proved by induction on n. For n = 2 we have a1 − a2 = b1 − b2, bj ∈ Ij. Then a = a1 − b1 = a2 − b2 satisfies a ≡ aj(mod Ij), j = 1, 2. For the inductive step one uses the distributive law n−1j = 1 = nj = 1 Ij + In. See exercise 9 and p. 461f of BAI.)
11. Let I be a fractional ideal, J an integral ideal. Show that there exists an element a ∈ I such that I− 1 a + J = D. (Hint: Let P1, …, Pr be the prime ideals dividing J. For 1 ≤ i ≤ r, choose ai ∈ IP1 … PrPi −1, IP1 … Pr and put a = ∑ 1rai. Then Pi aI − 1, 1 ≤ i ≤ r, and aI − 1 + J = D.)
12. Let I be a fractional ideal and a any non-zero element of I. Show that there exists a b ∈ I such that I = Da + Db. (Hint: Apply exercise 10 to the given I and J = aI− 1.)
13. Let I be a fractional ideal, J an integral ideal. Show that I/IJ and D/J are isomorphic D-modules.
14. Let P be a prime ideal ≠ 0 in D. Show that if e ≥ 1 then D/Pe ⊃ P/Pe ⊃ … ⊃ P e − 1/Pe ⊃ 0 is a composition series for D/Pe as D-module. Show that Pi/Pi + 1, i ≥ 0, is a one-dimensional vector space over the field D/P.
15. Show that if D has only a finite number of ideals Pl, …, PS, then D is a p.i.d. (Hint: Let I = P1e1 … Pses, where the ei ≥ 0. For each i choose ai ∈ IP1 … PsPi − 1, Pie + 1. Put a = ∑si ai. Then a ∈ Pie,, Pie, + 1 and I = Da.)
10.2 CHARACTERIZATIONS OF DEDEKIND DOMAINS
We shall give a number of characterizations of Dedekind domains. The first one involves the important concept of integral closedness that we introduced in section 7.7.
It is easy to see that any factorial domain D is integrally closed (exercise 1, below). We now prove
PROPOSITION 10.6. Any Dedekind domain is integrally closed.
Proof. Let D be a Dedekind domain, F its field of fractions, and let u be an element of F that is D-integral. Let f(x) = xn + a1xn − 1 + … + an be a monic polynomial with coefficients in D such that f(u) = 0 and put M = D 1 + Du + … + Dun − 1. If u = bc − 1 where b, c ∈ D, then cn − 1 M ⊂ D, so M is a fractional ideal. Evidently M2 = M, which implies M = D (exercise 2, p. 604). Then u ∈ M = D and hence D is integrally closed.
We have now established the following three properties of Dedekind domains D:
(1) D is noetherian.
(2) Every non-zero prime ideal in D is maximal.
(3) D is integrally closed.
We proceed to show that these properties characterize Dedekind domains. The proof is based on some lemmas that are of independent interest.
LEMMA 1. Any fractional ideal for a noetherian domain D is finitely generated as D-module.
Proof. If I is a fractional D-ideal, we have an a ≠ 0 in D such that I ′ = aI ⊂ D. Then I ′ is an ideal in D and since D is noetherian, I ′ is a finitely generated D-module. Then I = a − 1I ′ is finitely generated.
LEMMA 2. Any non-zero ideal of a noetherian domain contains a product of non-zero prime ideals.
Proof. As in the proof of Theorem 10.1, if the result is false we have an ideal I ≠ 0 maximal relative to the property that I does not contain a product of non-zero prime ideals. Then I is not prime and hence there exists ideals Ij, j = 1, 2, such that Ij I and I1 I2 ⊂ I. By the maximality of I, Ij contains a product of non-zero prime ideals. Since I ⊃ I1 I2, this gives a contradiction.
LEMMA 3. Let D be an integrally closed noetherian domain, F the field of fractions, I a fractional D-ideal. Then
Proof. Let s ∈ S. Since I is finitely generated and faithful as D-module and sI ⊂ I, s is D-integral by the lemma on p. 408. Since D is integrally closed, s ∈ D. Then S ⊂ D and since D ⊂ S follows from the definition of a fractional ideal, we have S = D.
LEMMA 4. Let D be a noetherian domain in which every non-zero prime ideal is maximal and let I be an ideal of D such that I D. Then I − 1 D.
Proof. Let a ≠ 0 be in I. Then D ⊃ I ⊃ aD, and by Lemma 2, aD ⊃ P1 P2 … Pm where the Pi are prime ideals ≠ 0. We assume that m is minimal. By the noetherian condition (or by Zorn’s lemma) there exists a maximal ideal P ⊃ I. Then we have P ⊃ I ⊃ aD ⊃ P1 … Pm. Since P and the Pi are prime, the hypothesis that non-zero prime ideals are maximal implies that P = Pi for some i, and we may assume that P = P1. If m = 1 we have aD = I. Then I − 1 = a− lD and since I D, a − 1 D and I− 1 = a− 1D D. Next assume that m > 1. Then aD P2 … Pm by the minimality of m and so we can choose a b ∈ P2 … Pm, aD. Put c = a− lb. Then c D and cI ⊂ cP = a− 1 bP ⊂ a− lbP ⊂ a− 1 PP2 … Pm ⊂ a − 1(aD) = D. Thus c ∈ I− 1 and c D, so again we have I − 1 D.
We can now prove
THEOREM 10.3. A domain D is Dedekind if and only if (1) D is noetherian. (2)non-zero prime ideal in D is maximal. (Equivalently, D has Krull dimension ≤ 1.)(3)D is integrally closed.
Proof. We have seen that (1)-(3) hold for any Dedekind domain. Now assume that D is a domain satisfying these conditions and let I be a D-fractional ideal. The fractional ideal I − 1 I is integral and II − 1(II− 1)− 1 ⊂ D, so I − 1(II − 1)− 1 ⊂ I– 1. It follows from Lemma 3 that (II − 1)− 1 ⊂ D. Since II − 1 ⊂ D, it follows from Lemma 4 that II − 1 = D. Hence every fractional ideal is invertible and D is Dedekind.
Our next characterization of Dedekind domains will be in terms of localizations. Let D be a domain, S a submonoid of the multiplicative monoid of non-zero elements of D. We consider the subring Ds of the field of fractions F of D consisting of the elements as as − 1, a ∈ D, s ∈ S. It is readily seen that this is isomorphic to the localization of D at S as defined in section 7.2. Hence the results on localization that were developed in sections 7.2 and 7.3 are available for the study of the rings Ds.
Let I be a fractional ideal for D and put Is = {bs− 1|b ∈ I, s ∈ S}. Then Isis a Ds-submodule of F, Is ≠ 0, and if a is a non-zero element of D such that aI ⊂ D, then aIs ⊂ Ds. Since it is evident that F is the field of fractions for Ds, it follows that Is is a Ds-fractional ideal. If I1 and I2 are D-fractional ideals, then
The first two of these are clear. To verify the third we note that (I1 ∩ I2)S ⊂ I1S ∩ I2S is clear. Now let a ∈ I1S ∩ I2S, so a = a1s1−1 = a2s2 −1 where ai ∈ Ii, si ∈ S. Then b = a1s2 = a2s1 ∈ I1 ∩ I2 and a = b(s1s2)−1 ∈ (I1 ∩ I2)s. Hence (6) holds.
We note next that if I is a finitely generated fractional D-ideal, then
For, it follows from the definition of I − 1 that (I1 + I2)−1 = I1−1 ∩ I2 −1. It is clear also that if b ≠ 0 in F, then (Db)s = Dsb. Hence if I = Db1 + … + Dbm, then Is = Dsb1 + … + Dsbm and Is −1 = m1 Dsbi −1. On the other hand, I −1 = m1Dbi −1, and so by (6), (I −1)s = m1Dsbi −1 = Is −1. Hence we have (7).
Now let I′ be a fractional ideal for Ds. Then we have an element as −1, a ≠ 0 in D, s ∈ S, such that as− 1 I′ ⊂ Ds. Then as − 1 I′ is a non-zero ideal contained in the localization Ds. By section 7.2 this has the form Is for a nonzero ideal I contained in D. Then I ′ = a −1sIs = ((a− 1s)I)s. Thus any fractional Ds-ideal has the form Is for some fractional D-ideal I.
We can now prove
THEOREM 10.4. Let D be a domain.
(1) If D is Dedekind, then Ds is Dedekind for every submonoid of the multiplicative monoid of non-zero elements of D.
(2) If D is Dedekind and P is a prime ideal ≠ 0 in D, then DP is a discrete valuation ring.
(3) If D is noetherian and DP is a discrete valuation ring for every maximal ideal P in D, then D is Dedekind.
(We recall that if P is prime ideal, DP ≡ Ds where S is the multiplicative monoid D − P.)
Proof. (1) Suppose that D is Dedekind and let I′ be a Ds-fractional ideal. Then I′ = Is for a fractional D-ideal I. Since D is Dedekind, D is noetherian and hence I is finitely generated. Then (I′)−1 = Is−1 = (I−1)s by (7) and I′I′−1 = Is(I−1)s = (II−1)s (by (5)) = Ds. Hence I′ is invertible and Ds is Dedekind.
(2) The localization DP at the prime ideal P ≠ 0 is a local ring whose only maximal ideal is Pp. Since Dp is Dedekind, Pp is the only prime ideal ≠ 0 in Dp. Hence Dp is a p.i.d. by exercise 14, p. 605. Then Dp is a discrete valuation ring by Proposition 9.2, p. 570.
(3) Now assume that D is noetherian and that Dp is a discrete valuation ring for every maximal ideal P of D. Then Dp is a p.i.d. and hence Dp is Dedekind. Let I be a fractional D-ideal. Since D is noetherian, I is finitely generated. Hence Ip−1 = (I−1)p. Then(II − 1)p = Ip(I−1)p = IpIp−1 = DP forevery maximal ideal P of D. We can now apply the principal of passage from local to global: We have the injection i: II − 1 D, which localizes to the injection (I I′)p Dp. Since the latter is surjective for all P, it follows from Proposition 7.11.2 (p. 402) that II − 1 D is surjective. Thus II − 1 = D for every fractional D-ideal I. Hence D is Dedekind.
Dedekind domains can also be characterized by the factorization property given in Theorem 10.1. In fact, we have the following stronger result:
THEOREM 10.5. Let D be a domain with the property that every proper integral ideal of D is a product of prime ideals. Then D is Dedekind.
Proof(Zariski-Samuel). We show first that any invertible prime ideal P of D is maximal in D. Let a ∈ D, P. Suppose that aD + P ≠ D. Then aD + P = P1 … Pm, a2D + P = Q1 … Qn where the Pi and Qj are prime ideals. Since aD + P P and a2D + P P (since P is prime), we have Pi P, Qj P for all i and j. Passing to the domain = D/P we obtain = 1 … m and = 1 … n where x is the canonical homomorphism of D onto . The principal ideals and are invertible and the i and j are prime. Moreover, 2 = ()2, so 1 … n = 12 … m2. By Proposition 10.4, the sequence of prime ideals {1, …, n} and {1, 1, …, m, m} are the same except for order. The corresponding sequences of ideals of D containing P are {Q1, …, Qn} and {P1, P1, …, Pm Pm}. Hence these coincide except for order. This implies that (aD + P)2 = P12 … Pm2 = Q1 … Qn = a2D + P. Then P ⊂ (aD + P)2 = a2D + aP + P2 ⊂ aD + P2. Then if p ∈ P, p = ax + y where x ∈ D and y ∈ P2. Then ax ∈ P and since a P, x ∈ P. Hence P ⊂ aP + P2 ⊂ P, so P = aP + P2. Multiplication by P− 1 gives D = aD + P, contrary to hypothesis. Thus we have aD + P = D for every a ∈ D, P. This implies that P is a maximal ideal of D.
Now let P be any prime ideal ≠ 0 in D and let b be a non-zero element of P. Then P ⊃ bD and bD = P1P2 … Pm where the Pi are prime and are invertible. Hence the Pi are maximal. Since P is prime, we have P ⊃ Pi for some i. Then P = Pi and so P is invertible. Since any proper integral ideal is a product of prime ideals, it follows that every integral ideal is invertible. Since any fractional ideal I ′ = aI for I integral and a ≠ 0, I ′(I ′)− 1 = II − 1 = D. Hence every fractional ideal is invertible and D is Dedekind.
The following theorem summarizes the characterizations of Dedekind domains that we have given.
THEOREM 10.6. The following conditions on a domain are equivalent:
(i) D is Dedekind.
(ii) D is integrally closed noetherian and every non-zero prime ideal of D is maximal.
(iii) Every proper integral ideal of D can be written in one and only one way as a product of prime ideals.
(iv) D is noetherian and DP is a discrete valuation ring for every maximal ideal P in D.
EXERCISES
1. Prove that any factorial domain is integrally closed.
2. Prove that the following conditions characterize Dedekind domains: D is integrally closed noetherian and D/I is artinian for every integral ideal I.
3. (H. Sah.) Show that a domain D is Dedekind if and only if it has the following property:
If I is any integral ideal of D and a is any non-zero element of I, then there exists a b ∈ I such that I = Da + Db.
(Sketch of proof: The proof that any Dedekind domain has this property was indicated in exercise 11, p. 605. Now suppose that D has the property. Then D is noetherian. Also it is easily seen that the property is inherited by any localization Ds. Now suppose that D is a local domain having the property and let P ≠ 0 be its maximal ideal. Let I be any ideal ≠ 0 in D. Then I = IP + Db for some b ∈ I. It follows by Nakayama’s lemma that I = Db. Thus D is a p.i.d. and hence a discrete valuation ring (p. 571). Now conclude the proof by using Theorem 10.4.3 (p. 628).
Exercises 4-7 were communicated to me by Tsuneo Tamagawa. The first of these is well known. In all of these exercises, D is Dedekind with F as field of fractions and x is an indeterminate.
4. If f(x) ∈ D[x], we define the content c(f) to be the ideal in D generated by the coefficients of f(cf. BAI, p. 151). Prove that c(fg) = c(f)c(g) for f, g ∈ D[x].
5. Let S = {f ∈ D[x]|c(f) = D}. Note that S is a submonoid of the multiplicative monoid. Let D[x]s denote the subring of F(x) of fractions f(x)/g(x) where f(x) ∈ D[x] and g(x) ∈ S. Show that if f(x), g(x) ∈ D[x], then f(x)/g(x)∈ D[x]s if and only if c(f) ⊂ c(g).
6. Show that D[x]s ∩ F = D. Let I ′ be an ideal in D[x]s. Show that I = I ′ ∩ D is an ideal in D such that ID[x]s = I ′. Show that I I ′ = ID[x]s is a bijective map of the set of ideals of D onto the set of ideals of D[x]s.
7. Prove that D[x]s is a p.i.d.
10.3 INTEGRAL EXTENSIONS OF DEDEKIND DOMAINS
The classical examples of Dedekind domains are the rings of -integers of number fields, that is, of finite dimensional extension fields of and of fields of algebraic functions of one variable. In this section we derive the general result that if D is Dedekind with field of fractions F and E is a finite dimensional extension field of F, then the subring D′ of D-integral elements of E is Dedekind. This implies the Dedekind property of the classical domains and gives the fundamental theorem on the unique factorization of integral ideals as products of prime ideals in these domains.
We prove first
PROPOSITION 10.7. Let D be a domain, F its field of fractions, E a finite dimensional extension field of F, and D′ the subring of E of D-integral elements. Then any element of E has the form rb− 1 where r ∈ D′ and b ∈ D.
Proof. Let u ∈ E and let g(x) = xm + a1xm − 1 + … + am ∈ F[x] be a polynomial such that g(u) = 0. We can write ai = cib − 1, ci, b ∈ D. Then
∈ D[x] and h(bu) = 0. Hence r = bu ∈ D′ and u = b − 1r as required.
Evidently this result implies that E is the field of fractions of D′. Since D′ is the ring of D-integers of E, D′ is integrally closed in E. Hence D′ is an integrally closed domain.
We suppose next that D is integrally closed and we prove
PROPOSITION 10.8. Let D, E, F, D ′ be as in Proposition 10.7 and assume that D is integrally closed. Let r ∈ E and let m(x),f(x) be the minimum polynomial and characteristic polynomial respectively of r over F. Then the following conditions on r are equivalent:
(1) r is integral.
(2) m(x) ∈ D[x].
(3) f(x) ∈ D[x].
If the conditions hold, then TE/F(r) and NE/F(r) ∈ D.
Proof. Since f(x) is a power of m(x), it is clear that (2) (3). Also obviously (3) (1). Now assume (1). Let be an algebraic closure of F containing E. We have m(x) = (x − ri) in [x] where r1 = r. For any i we have an automorphism of /F sending r into ri. Since r is D-integral, there exists a monic polynomial g(x) ∈ D [x] such that g(r) = 0. Applying an automorphism of /F sending r into ri, we obtain g(ri) = 0. Hence every root ri of m(x) is a D-integer. Since the coefficients of m(x) are symmetric polynomials in the ri, these elements of F are D-integers and since D is integrally closed, they are contained in D. Thus m(x) ∈ D[x] and (1) (2). The last assertion is clear, since f(x) = xn − TE/F(r)xn − 1 + … +(− l)nNE/F(r).
A key result for the proof of the theorem on integral closures of Dedekind domains is the following
PROPOSITION 10.9. Let D, E, F, D′ be as in Proposition 10.8 and assume that E/F is separable and D is noetherian. Then D′ is a finitely generated D-module.
Proof. Since E/F is separable, there exist n = [E:F] distinct monomorphisms σ1, …, σn of E/F into the algebraic closure /F. Moreover, if (ul, …, un) is a base for E/F, then the matrix
is invertible (BAI, p. 292). We know also that TE/F(u) = ∑ n1 σi (u) for u ∈ E (BAI, p. 430). We have (tA)A = (TE/F(uiuj)). Hence
We have seen in Proposition 10.7 that there exists a bi ≠ 0 in D such that biui ∈ D′. Hence we may assume that the ui ∈ D′. Let r ∈ D′ and write r = ∑n1aiui, ai ∈ F. Then ruj ∈ D′, 1 ≤ j ≤ n, and ruj = ∑iaiuiuj. Hence
The system of equations (10) for ai, 1 ≤ i ≤ n, can be solved by Cramer’s rule. This shows that ai ∈ d−1 D. Since this holds for every r ∈ D ′, we see that D′ is contained in the D-module M = ∑n1D(d− 1 ui). Thus D ′ is contained in a finitely generated D-module. Since D is noetherian, M is a noetherian D-module. Hence the submodule D′ is finitely generated.
We can now prove the main result of this section.
THEOREM 10.7. Let D be a Dedekind domain, F its field of fractions, E a finite dimensional extension field of F, D′ the subring of D-integral elements of E. Then D′ is Dedekind.
Proof. We assume first that E is separable over F. In this case we shall prove the theorem by showing that D′ has the three properties characterizing Dedekind domains given in Theorem 10.3, namely, (1) D′ is noetherian, (2) every prime ideal ≠ 0 in D′ is maximal, and (3) D′ is integrally closed. The first of these follows from Proposition 10.9, since this proposition implies that D′ is a noetherian D-module. A fortiori D′ is noetherian as D′-module. Now let P′ be a non-zero prime ideal in D′. Let r ≠ 0 be in P′. Then we have f(r) = 0 for the characteristic polynomial f(x) of r. This implies that NE/F (r) ∈ P ′ and hence, by Proposition 10.8, NE/F(r) ∈ P = P′ ∩ D. Since NE/F(r) ≠ 0, we see that P ≠ 0. It is clear also that P is a prime ideal in D. Since D is Dedekind, this implies that P is a maximal ideal in D. By the Corollary to Proposition 7.17, p. 410, we conclude that P′ is maximal in D′. Hence (2) holds. The remark following Proposition 10.7 shows that (3) holds. Hence D′ is Dedekind in the separable case.
We assume next that char F = p and that E is purely inseparable over F. Since [E:F] < ∞, there exists a q = pe such that upe ∈ F for all u ∈ E. Let be an algebraic closure of F containing E and let F1/q(Dl/q) be the subfield (subring) of of elements v such that vq ∈ F(D). Then v vq is an isomorphism of F1/q onto F mapping Dl/q onto D. Since D is Dedekind, D1/q is Dedekind. If u ∈ D′, uq ∈ F and uq is D-integral. Hence uq ∈ D and so D′ ⊂ D1/q. We shall now show that D′ is Dedekind by proving that every D′-fractional ideal I of E is invertible. Let I′ = Dl/qI. Then I′ is a D1/q-fractional ideal. Hence there exists a D1/q-fractional ideal J′ such that I′ J′ = D1/q and since I′ = ID1/q and Dl/qJ′ = J′, we have elements bi ∈ I, c′i ∈ J ′ such that ∑bici′ = 1. Then ∑biqc′iq = 1 and ∑bici = 1 where ci = biq−1 c′iq. Now biq−1 c′iq ∈ I′ q−1J′q ⊂ J′. Since c′iq ∈ F and bi ∈ E, ci = bi q−1 c′iq ∈ E. Thus ci ∈ J = J′ ∩ E. Evidently IJ ⊂ D1/q ∩ E and since the elements of D1/q are D-integral, D1/q ∩ E ⊂ D′. Hence IJ ⊂ D′, so J ⊂ I− l. Since the ci ∈ J and bi ∈ I and ∑cibi = 1, we have II− 1 = D′. Thus I is invertible and D′ is Dedekind.
The general case can be obtained by combining the preceding two cases in the obvious way: If E is arbitrary finite dimensional over F, let K be the maximal separable subfield of E/F. Then E is purely inseparable over K. Now D″ = D′ ∩ K is Dedekind by the first case and since D′ is the set of D″-integral elements of E, D′ is Dedekind by the second case. This completes the proof.
The foregoing result is applicable in particular to F = and D = and to F = K(x) where K is an arbitrary field, x an indeterminate, and D = K[x]. This is clear since these D are p.i.d. and hence are Dedekind. If E is a finite dimensional extension field of F, then the ring D′ of D-integers of E is Dedekind. In the case F = , D = , E is a number field and in the other cases E is an algebraic function field in one variable.
EXERCISES
1. Let the notations be as in Theorem 10.7. Assume that E/F is separable and that D is a p.i.d. Show that D′ is a free D-module of rank n = [E: F].
Exercises 3, 4, and 5 of BAI, p. 287, are relevant for this section.
2. (Sah.) Show that if D is a Dedekind domain, then so is the ring D((x)) of Laurent series over D. (Hint: It is easily seen that D((x)) is a domain. If I is an ideal in D((x)), let l(I ) be the set of leading coefficients of the elements of I. Then l(I) is an ideal and if l(I) = Da1 + … + Dam where ai is the leading coefficient of fi ∈ I, then I = D((x))f1 + … + D((x))fm. Use this and exercise 3 on p. 630.
10.4 CONNECTIONS WITH VALUATION THEORY
We give first a valuation theoretic characterization of the integral closure of a subring of a field. This is
THEOREM 10.8 (Krull). Let D be a subring of a field F and let D′ be the integral closure of D in F (= the subring of D-integral elements of F). Then D′ is the intersection of the valuation rings of F containing D.
Proof. Let u ∈ D′ and let R be a valuation ring in F containing D, φ a valuation of F having R as valuation ring (e.g., the canonical valuation defined by R). Suppose that u R. Then φ(u− 1) < 1. On the other hand, we have a relation un + a1un − 1 + … + an = 0, ai ∈ D, which gives the relation
Since the ai ∈ D ⊂ R, φ(ai) ≤ 1. Then φ(aiu− i) < 1. Since φ(1) = 1, we have the contradiction 1 ≤ max{φ(aiu −i)} < 1. Hence u ∈ R and D′ ⊂ R where the intersection is taken over all of the valuation rings R of F containing D. Next, suppose that u D′. Then u−1 is not a unit in the subring D[u−1] of F, since otherwise, its inverse u = a0 + alu−1 + … + an−1u−(n−1), ai ∈ D, and hence un = a0un−1 + a1un − 2 + … + an −1, contrary to u D ′. Since u−1 is not a unit in D[u−1], u−1 D[u−1] is a proper ideal in D[u− 1] and this can be imbedded in a maximal ideal P0 of D[u− 1]. Then = D[u− 1]/P0 is a field and we have the canonical homomorphism of D[u− 1] onto , which can be regarded as a homomorphism 0 of D[u−1] into an algebraic closure of . By Theorem 9.10 (p. 561), there exists a -valued place extending 0. If R is the valuation ring on which is defined and P is its maximal ideal, then P is the kernel of . We have R ⊃ D[u−l] ⊃ D and P ⊃ P0. Then u− 1 D[u− 1]⊂ P0 ⊂ P;, so (u− 1) = 0 and hence u R. Thus we have shown that if u D′ then u R where R ranges over the valuation rings of Fcontaining D. Hence D′ = R.
Now let D be Dedekind, F its field of fractions. Since D is integrally closed, D is the intersection of the valuation rings of F containing D. We claim that these valuation rings are the localizations DP, for the prime ideals P of D. We have seen that every DP is a discrete valuation ring for every prime ideal ≠ 0 in D. This is clear also if P = 0, since D0 = F. Now let R be a valuation ring in F containing D such that R ≠ F. Let M be the maximal ideal of R and let P = M ∩ D. Then P is a prime ideal in D and if P = 0, every nonzero element of D is invertible in R. Since F is the field of fractions of D, this implies that F = R, contrary to hypothesis. Thus P is an integral prime ideal in D. We have the valuation ring DP with maximal ideal Pp = PDp. If a ∈ D − P, then a ∈ R − M and a− l ∈ R. Then R ⊃ DP. Moreover, since M ⊃ P, M = RM ⊃ DpP. Now it is easily seen that if R1 and R2 are valuation rings with maximal ideals M1 and M2 respectively, then R1 ⊃ R2 and M1 ⊃ M2 imply R1 = R2 and M1 = M2. Hence R = Dp and M = PDp. Thus the valuation rings of F containing D (and ≠F) are just the localizations Dp, P a prime ideal ≠ 0 in D. By Theorem 10.8, D = Dp.
A prime ideal P ≠ 0 in a Dedekind domain D determines an exponential valuation vp for the field of fractions F of D in a manner similar to the definition of the exponential valuation vp on defined by a prime integer p. If a ≠ 0 in F, we write Da = Pk P1k1 … Plkl where P, P1, …, Pl are distinct prime ideals and k and the ki ∈ . Then we put vp(a) = k. Moreover, we define vp(0) = ∞. It is readily verified, as in the special case of , that vp is an exponential valuation of F. Hence if we take a real γ, 0 < γ < 1, then |a|p = γ vP (a) is a non-archimedean absolute value on F. We call this a P-adic absolute value on F. The valuation ring of | |p evidently contains the localization DP and the maximal ideal of the valuation ring contains PDp. Hence the valuation ring of | |p is Dp and its maximal ideal is PDp.
Since Dp is Dedekind and PDp is its only prime ideal ≠ 0, the integral ideals of Dp are the ideals PkDp, k ≥0. Evidently, Pk ⊂ PkDp ∩ D. On the other hand, if a ∈ PkDp ∩ D for k ≥ 1, then a = bc− 1 where b ∈ Pk, c ∈ D − P. Then ac ≡ 0 (mod Pk) and c 0 (mod P). The latter condition implies that c + P is a unit in D/P. Then c + Pk is a unit in D/Pk (p. 195). It follows that a ∈ Pk. Hence PkDp ∩ D = Pk for k ≥ 0.
Let | | be a non-archimedean absolute value having Dp as valuation ring. We claim that | | = | | p, a P-adic valuation determined by the prime P. To see this we observe that PDp = π Dp where π ∈ PDp, P2Dp. Then PkDp = πkDp for k ≥ 0 and any element of Dp can be written as uπk, u a unit in Dp and k ≥ 0. Then |uπk| = γk where |π| = γ, 0 < γ < 1. Hence, if a ∈ Pk, Pk + 1, then a ∈ PkDp, Pk + 1DP, and |a| = γk. On the other hand, it is clear that Da = Pk P1k1 … Plkl where Pl, …, Pl are primes distinct from P and hence if | |P is the P-adic valuation determined by γ, then |a|p = |a| for a ∈ D. Then | | = | |p.
We show next that the residue field Dp/Pp of | |P is canonically isomorphic to D/P. Since P = Pp ∩ D, we have the monomorphism a + P a + Pp of D/P into Dp/Pp. If a, b ∈ D and b P, then since D/P is a field there exists a c ∈ D such that cb ≡ a (mod P). Then ab− 1 ≡ c (mod Pp), which implies that a + P a + Pp is surjective on Dp/Pp. Then this map is an isomorphism.
We summarize this collection of results in
PROPOSITION 10.10. Let D be a Dedekind domain with field of fractions F. Then (1) Any valuation ring in F containing D has the form Dp where P is a prime ideal in D. (2) Any non-archimedean absolute value on F having DP, P ≠ 0, as valuation ring is a P-adic valuation. (3) The map a + P a + Pp is an isomorphism of D/P onto Dp/Pp.
Now let E be a finite dimensional extension field of F and let D′ be the ring of D-integers of E. Then D′ is Dedekind. If P is an integral ideal of D, then the ideal in D′ generated by P is D′ p. Now D′ P ≠ D′. For, by the “Lying- over” theorem (p. 411), there exists a prime ideal P′ in D′ such that P′ ∩ D = P. Hence P′ ⊃ PD′ and PD′ ≠ D′. It may happen that the extension ideal PD′ of P is not prime. At any rate, we have a factorization
where the P′i are distinct prime ideals in D′ and the ei > 0.
We have treated the general problem of extension of an absolute value on a field F to a finite dimensional extension field in Chapter 9 (p. 585). We shall now treat the special case of a P-adic absolute value of the field of fractions F of a Dedekind domain in an alternative manner based on the factorization (12). We prove first
THEOREM 10.9. Let D be a Dedekind domain, F the field of fractions of D, E a finite dimensional extension field of E, D′ the subring of E of D-integral elements. Let P be an integral prime ideal in D, | |P a P-adic absolute value on F, (12) the factorization of PD′ into prime ideals in D′. Then for each P′i there is a unique P′i-adic absolute value | |p′i on E extending | |p. Moreover, | |p′i and | |p′j are inequivalent if i ≠ j and the | |p′i; are the only extensions of | |P to absolute values on E.
Proof. Let a ≠ 0 be in F and write Da = Pk Pkl1 … Pkl1 where P and the Pj are distinct prime ideals and k and the kj ∈ . Then vP(d) = k, and by (12),
where the terms after P′gegk come from the factorizations of the D′Pj into prime ideals in D′. Now we observe that if P1 and P2 are distinct prime ideals in D, then P1 + P2 = D and hence D′ P1 + D′ P2 = D′. This implies that the prime ideals in D′ dividing D′P1 are different from those dividing D′P2. It follows that vp′j(a) = ejk. Hence if | |p = γ vp( ) for 0 < γ < 1, then we shall have | |p = | |p′j on F if and only if ||p′j = (γ1/ej)vp′j( ). This proves the first assertion. We have seen that the valuation ring of | |p′j (or of vP′j) is D′p′j. Since these are distinct for distinct choices of j, it follows that | |p′i and | |p′j are inequivalent if i ≠ j. Now let | |′ be any absolute value on E extending | |P and let R′ denote the valuation ring of | |′. Then R′ ⊃ Dp and since D′ is the integral closure of D in E, R′ ⊃ D′ by Theorem 10.8. By Proposition 10.10.1, R′ = D′p′ for P′, a prime ideal in D′. Since R′ ⊃ Dp, it follows that P′ ⊃ P. Hence P′ = P′i for some i. By Proposition 10.10.2, | |′ is a P′i;-adic valuation.
We recall that the ramification index of | | p′i relative to | |p is the index (p. 589). Since | F* | P is the subgroup of * generated by γ and |E*|p′i is the subgroup generated by γ1/ei, the ramification index is ei.
We recall also that the residue degree fi of | |p′i relative to | | p is the dimensionality [D′P′l/P′ipl : Dp/Pp] where DP/Pp is imbedded in D′p/P′iP by the monomorphism x + Pp x + P′ipl (p. 589). We also have a monomorphism a + P a + P′i, a ∈ D, of D/P into D′/P′i. Hence we can define the dimensionality [D′/P′i : D/P]. Now we have the isomorphisms σ : a + P a + Pp and s : x + Pi x + P′ip′l of D/P onto Dp/Pp and of D′Pl onto D′p′l /P′ip′i respectively. We have s((a + P)(x + P′i)) = s(ax + P′i) = ax + P′ip′l and hence(σ(a + P))(s(x + P′i)) = (a + Pp)(x + P′iP′l′ ) = ax + P′ip′l. Hence s is a semi-linear isomorphism of D′/P′i over D/P onto D′p′l/P′ip′l with associated field isomorphism σ. It follows that the residue degree fi coincides with the dimensionality [D′/P′i : D/P].
The results on ramification index and residue degree can be stated as follows:
THEOREM 10.10. Let the notations be as in Theorem 10.9. Let | |p be a P-adic absolute value on F, | |P; a P′i;-adic absolute value on E extending | |P. Then ei is the ramification index and [D′/P′i : D/P] is the residue degree of | | p′l relative to | |p.
An immediate consequence of this result and the main theorem on residue degrees and ramification indices (Theorem 9.15, p. 592) is
THEOREM 10.11. Let the notations be as in Theorem 10.9. Then Moreover, ∑g1eifi = n if E/F is separable.
We remark that this result has the consequence that the number g of prime ideals of D′ containing D′P is bounded by n and it gives considerably more information on the factorization (12).
EXERCISES
1. Let E be the cyclotomic field () when is a primitive lth root of unity, l a prime. Let D = , D′ be the integral closure of in E. Show that lD′ = (( − l)D′)p −1 and ( − 1)D′ is a prime ideal in D′. (Hint: [E : Q] = l − 1 and f(x) = xl − 1 + xl − 2 + … + 1 is the minimum polynomial of . We have f(x) = 1l − 1(x − i) in E[x] and every i, 1 ≤ i ≤ l − 1, is a primitive lth root of 1. (See BAI, p. 154.) Substituting 1 in f(x) gives l = 1l − 1(1 − i). Moreover, (1 − i)/(1 − ) = 1 + + … + i −1 ∈ D′. Show that l + + … + i−1 is a unit in D′ so lD′ = (( − 1)D′)l − 1. Show that l is not a unit in D′ and use Theorem 10.11 to conclude that ( − 1)D′ is prime.)
The next three exercises are designed to show that in the special case of Theorem 10.11 in which E/F is Galois, the factorization (12) of PD′ = (P′1 … P′g)e and all of the residue degrees are equal. Hence n = [E : F] = efg.
2. Let E/F be Galois with Gal E/F = G. Show that any σ ∈ G maps D′ into itself. Show that if P′ is a prime ideal in D′ then σP′ is a prime ideal. Show that : a + P′ σ a + σP′ is an isomorphism of D′/P′ onto D′/σP′.
3. Let P be a non-zero prime ideal of D, PD′ = P′1e1 … P′geg, the factorization of PD′ into prime ideals in D′ where P′i ≠ P′j if i ≠ j. Show that if σ ∈ G, then {σP′1, …, σP′g} = {P′, …, P′g} and that G acts transitively on the set of prime ideals {P′1, …, P′g}. (Hint: To prove transitivity of G on {P′1, …, P′g} one shows that if P′ is any prime ideal of D′ such that P′ ∩ D = P then any prime ideal Q′ of D′ such that Q′ ∩ D = P is one of the ideals σ P′, σ ∈ G. Otherwise, by Proposition 7.1, p. 390, Q σ ∈ G σ P′. Hence there exists an a ∈ Q′ such that σa P′ for any σ ∈ G. Then NE/F(a) = σ ∈ G σ a P′ This is a contradiction since NE/F(a)∈ Q′ ∩ D = P.)
4. Use the results of exercises 2 and 3 to show that ei = ej for every i and j and that the fields D′/P′i and D′/P′j are isomorphic over D/P. Hence conclude that [D′/P′i : D/P] = [D′/P′j : D/P]. Put e = ei, f = fi = [D′Pi : D/P]. Show that n = efg.
10.5 RAMIFIED PRIMES AND THE DISCRIMINANT
As in the previous section, let D be a Dedekind domain, F its field of fractions, E a finite dimensional extension field of F, and D′ the subring of D-integral elements of E. Let P be a non-zero prime ideal in D and let PD′ = P′1e1 … P′geg where the P′i are distinct prime ideals in D, as in (12). We consider the ring = D′/PD′, which we can regard also as a module over D′ and by restriction as a module over D. Since P annihilates , can also be regarded as a vector space over = D/P. The action of a + P, a ∈ D, on x + PD′, x ∈ D′, is (a + P)(x + PD′) = ax + PD′. Taking into account the ring structure as well as the -vector space structure, we obtain the algebra over . We shall now consider the structure of this algebra. The main result on this is the following
THEOREM 10.12. Let D be a Dedekind domain with field of fractions F and let E be a finite dimensional extension field, D′ the subring of E of D-integral elements. Suppose that P is a non-zero prime ideal in D and let PD′ = P′1e1 … P′geq where the P′i are distinct prime ideals in D′ and the ei > 0. Then = D′/PD′ is an algebra over the field = D/P in which (a + P)(x + PD′) = ax + PD′, a ∈ D, x ∈ D′. The dimensionality [ : ] = ∑1geifi where fi is the residue degree [ D′/P′i : ]. We have [ : ] ≤ n = [E : F] and [ : ] = n if E/F is separable. is a direct sum of g ideals isomorphic to the algebras D′/P′iei. The radical of is P′1 … P′g/PD′ and /rad is a direct sum of g ideals isomorphic to the fields /P′i.
Proof. Put . Then
This implies that
and . Hence
We have the chain of ideals
each of which is minimal over the next. Hence
is a composition series for D′/P′iel as D′-module. The composition factors are isomorphic to D′/P′i, P′i/P′2i, …, P′iel− 1/P′iel, each of which is annihilated by P′i and so can be regarded as a vector space over the field D′/P′i. Since these vector spaces have no non-zero subspaces, they are one-dimensional over D′/P′i. Hence the composition factors are isomorphic as D′-modules and hence also as D-modules and as vector spaces over . Then [P′ij − 1/P′ij : ] = [D′/P′i : ] = fi for j = 1, …, ei. It follows from (18) that [D′/P′iel : ] = eifi. Since we have [ : ] = Σ1geifi. We have shown in Theorem 10.11 that Σ eifi ≤ n and Σ eifi = n if E/F is separable. Hence [ : ] ≤ n and = n if E/F is separable.
Let e = max{ei}. Then (P′1 P′2 … P′g)e ⊂ D′P. Hence P′1 … P′g/D′P is a nilpotent ideal in = D′/D′P and . Since every D′/P′i is a field, it follows that P′1 … P′g/D′P = rad and /rad is a direct sum of g ideals isomorphic to the fields D′/P′i.
We now give the following
DEFINITION 10.4. A non-zero prime ideal P of D is said to be unramified in D′ if D′P = P′1 … P′g where the P′i are distinct prime ideals in D′ and every D′/P′i is separable over = D/P. Otherwise, P is ramified in D′.
It is clear from Theorem 10.12 that P is unramified in D′ if and only if = D′/D′P is a direct sum of fields that are separable over = D/P. We shall show that if E/F is separable, then there are only a finite number of prime ideals P of D that are ramified in D′ and in principle we shall determine these prime ideals. For this purpose we give the following
DEFINITION 10.5. Let D, F, E, D′ be as usual and assume that E/F is separable. Then the discriminant (ideal) dD′/D of D′ over D is the ideal generated by all of the elements
where (u1, …, un) is a base for E/F consisting of elements ui ∈ D′.
We have seen in the proof of Proposition 10.9 (p. 612) that det(TE/F(uiuj)) ≠ 0 for every base (u1, …, un) of a separable extension E/F. The main result on ramification is
THEOREM 10.13. Let D be a Dedekind domain, F its field of fractions, E a finite dimensional separable extension of F, D′ the subring of E of D-integral elements. Then a non-zero prime ideal P of D is ramified in D′ if and only if P is a divisor of the discriminant dD′/D.
Of course, this implies the finiteness of the set of ramified primes. To achieve the proof of Theorem 10.13 we need to obtain a trace criterion that a finite dimensional commutative algebra A over a field F be a direct sum of ideals that are separable field extensions of F.
If A is a finite dimensional algebra over a field F, we define the trace bilinear form T(a, b) on A/F by
where TA/F is the trace defined by the regular representation. Since TA/F is a linear function and TA/F(ab) = TA/F(ba), T(a, b) is a symmetric bilinear form (BAI, pp. 424, 426). We recall also that T(a, b) is non-degenerate if and only if det(T(ui, uj)) ≠ 0 for any base (ui, …, un) of A/F.
We require the following
LEMMA. If A is a finite dimensional commutative algebra over a field F, then A/F is a direct sum of ideals that are separable fields over F if and only if the trace bilinear form T(a, b) is non-degenerate on A.
Proof. Suppose first that where Ei is a separable field extension of F. Let (ui1, …, uin1) be a base for Ei. Then (ull, …, u1n1 …, ug1, …, ugng) is a base for A/F and
where Ti denotes the trace bilinear form on Ei. Since we have shown (p. 612) that Ti is non-degenerate, the foregoing formula shows that T is non-degenerate on A.
Conversely, suppose that T is non-degenerate on A. We show first that A is semi-primitive. Let z ∈ rad A. Then za = az is nilpotent for every a ∈ A. Since the trace of a linear transformation is the sum of its characteristic roots, the trace of a nilpotent linear transformation is 0. Hence T(a, z) = 0 for all a. Since T is non-degenerate, z = 0. Thus rad A = 0 and A is semiprimitive. Then where Ei is a field extension of F. By (21), the trace bilinear form Ti is non-degenerate on every Ei. On the other hand, if E/F is a/finite dimensional inseparable extension field of F then the trace function is identically 0 on E. For, E S where S is the maximal separable subfield of E. Then E is purely inseparable over S. It follows that there exists a subfield K of E containing S such that [E : K] = p the characteristic. Then the minimum polynomial over K of any a ∈ E is either of the form xp − α or of the form x − α. In either case the characteristic polynomial is of the form xp − β (β = αp in the second case). Then TE/K(a) = 0 for all a. By the transitivity of the trace (BAI, p. 426) we have TE/F(a) = TK/F(TE/K(a)) = 0. It follows that the trace bilinear form on an inseparable field is identically 0. Hence every Ei is separable.
We can now give the
Proof of Theorem 10.13. We assume first that D is a p.i.d. By Proposition 10.9, D′ is finitely generated as D-module. Since there is no torsion and D is a p.i.d., D′ is a free D-module. Since every element of E has the form r− 1u where r ∈ D and u ∈ D′, the rank of D′ over D = n = [E : F]. Let (vl, …, vn) be a D-base for D′. Then det(T(vi, vj)) generates dD′/D, so we have to show that is a direct sum of separable field extensions of if and only if Pdoes not contain the element det(T(vi, vj)). Since (vl, …, vn) is a D-base for D′, PD′ = Σ Pvi and if i = vi + PD′, then (1, …, n) is a base for over . It follows that if a ρ(a) is the regular matrix representation of E/F obtained by using the base (vl, …, vn), then the regular matrix representation of / using the base (1, …, n) is ( = a + PD′). This implies that det(T(i, j)) = det(T(vi, vj)) + P. Hence by the Lemma, is a direct sum of separable fields over if and only if det (T(vi, Vj)) P. This completes the proof in the case in which D is a p.i.d.
The proof for arbitrary D can be reduced to the p.i.d. case by localizing at the prime P of D. Since we are interested in localizing D′ as well as D and since P is not a prime ideal in D′, we make explicit the localizing monoid as S = D − P. We form Ds and D′s. By Proposition 7.16, D′s is the integral closure of Ds in E. Since P′i ∩ D = P, P′ ∩ S = Ø. Hence P′is is a prime ideal in D′s. We have Ps D′s = P′1se1 … P′gseg and the P′is are distinct. Since P′is ∩ D′ ⊂ P′ip′ ∩ D′ = P′i, we have P′is ∩ D′ = P′i. It follows as in the proof of Theorem 10.10 that we have a semi-linear map of D′s/P′is over Ds/Ps onto D′/P′i over D/P, which is a ring isomorphism. Hence D′s/P′is is separable over Ds/Ps if and only if D′/P′i is separable over D/P. Thus P is ramified in D′ if and only if Ps is ramified in D′s.
We show next that dD′s/Ds = (dD′/D)s. Let(u1, …, un) be a base for E/F such that the ui ∈ D′. Then ui ∈ D′s and det(T(ui, uj))∈ dD′s/Ds. Hence dD′/D ⊂ dD′s/Ds and (dD′/D)s ⊂ dD′s/Ds. Next let (u′1, …, u′n) be a base for E/F such that the u′i ∈ D′s. Then the u′i = ui s− 1, ui ∈ D′, s∈ D − P, and det(T(u′i))= s− 2n det(T(ui, uj)) ∈ (dD′/D)s. Hence dD′s/Ds⊂ (dD′/D)s so dD′s/Ds = dD′s/Ds. It is clear also that P is a divisor of dD′/D if and only if Ps is a divisor of (dD′/D)s = dD′s/Ds. We have therefore achieved a reduction of the theorem to the rings Ds and D′s. Since Ds is a p.i.d., the result follows from the case we considered first.
Let m be a square-free integer, E = (), and let D′ be the integral closure of in E. Then D′ has the base (1, ) over if m ≡ 2 or 3 (mod 4) and D′ has the base (1,(1 + )/2) over if m ≡ 1 (mod 4) (exercises 4 and 5, p. 287 of BAI). Hence in the first case, the discriminant dD′/ is the principal ideal generated by
Hence the ramified primes (which are the primes p in such that p is a square of a prime ideal in D′) are 2 and the prime divisors of m.
In the second case (m ≡ 1 (mod 4)), dD′/ is the principal ideal generated by
Hence the ramified primes are the prime divisors of m.
EXERCISES
1. Notations as in the example. Show that if p is odd and pD unramified in D′, then pD′ is a prime or a product of two distinct prime ideals in D′ according as the Legendre symbol (m/p) = − 1 or = 1. Show that if m ≡ 1 (mod 4), then 2D′ is prime or a product of two prime ideals in D′ according as m ≡ 5 or m ≡ 1 (mod 8). (Hint: Use Theorems 10.9 and 9.14.)
2. Let E be the cyclotomic field (), a primitive lth root of unity, l a prime, as in exercise 1, p. 618, which showed that ( − 1)D′ is prime in D′. Show that [D′/( − 1)D′ : /l] = 1 and hence that |D/( − 1)D′| = l Show that D′ = [] + ( − 1)D′ Use exercise 2, p. 276 of BAI, to show that |det T(i, j| = ll− 2 if 0 ≤ i, j ≤ l − 2. Hence show that D′ ⊂ l−(l + 2)[] and l(l + 2)D′ ⊂ [] ⊂ D′ Use these results to prove that D′ = [] and that l is the only prime ideal of that is ramified in D′.
3. Show that a finite dimensional commutative algebra A over a field F is a direct sum of separable fields if and only if AE is semi-primitive for every field extension E of F. (The latter property was discussed on p. 374 where it was called separability.)
10.6 FINITELY GENERATED MODULES OVER A DEDEKIND DOMAIN
In this section we shall extend the theory of finitely generated modules over a p.i.d. that we treated in BAI (sections 3.6–3.9, pp. 179-194) to finitely generated modules over a Dedekind domain. We recall that if M is a module over a domain D, the subset T of elements x ∈ M for which there exists an a ≠ 0 such that ax = 0 is a submodule called the torsion submodule of M. M is called torsion free if T = 0. Any submodule of a torsion-free module is torsion free. Hence, since any free module is torsion free, any submodule of a free module is torsion free. It is clear also that if T is the torsion submodule of a module M, then M/T is torsion free. We shall derive next a pair of less obvious results on torsion-free modules.
PROPOSITION 10.11. Any finitely generated torsion-free module M over a domain D is isomorphic to a submodule of a free module of finite rank.
Proof. Let {u1, …, ur} be a set of generators for M. We assume M ≠ 0 and u1 ≠ 0. We may assume that we have an s, 1 ≤ s, ≤ r such that {u1, …, us} are linearly independent over D in the sense that ∑1sdiui = 0 holds only if every di = 0, but {u1, …, us, us + j} are linearly dependent over D for all j, 1 ≤ j ≤ r − s. Then the submodule ∑1sDui of M is free and hence the result holds if s = r. On the other hand, if s < r, then for each j there exists a nonzero dj ∈ D such that djus + j ∈ ∑1sDui. Then dus + j ∈ ∑1sDui for d = dj ≠ 0 and so dM ⊂ ∑1s Dui. Hence the D-endomorphism x dx of M maps M into the free submodule ∑1s Dui Since M is torsion free, this map is a monomorphism. Hence M is isomorphic to a submodule of the free module ∑1s Dui.
We shall obtain next a generalization of the theorem that any submodule of a finitely generated free module over a p.i.d. is free (Theorem 3.7 of BAI, p. 179). This is
PROPOSITION 10.12. Let R be a ring, which is not necessarily commutative, with the property that every left ideal of R is a projective R-module. Then any submodule M of the free module R(n) is a direct sum of m ≤ n submodules isomorphic to left ideals of R. Hence M is projective.
Proof. The result is clear if n = 1, since the submodules of R = R(1) are left ideals. Hence assume that n > 1. Let S denote the submodule of R(n) of elements of the form (x1, …, xn − 1, 0). Evidently S ≅ R(n − 1]. Now consider the R-homomorphism p : y = (yl, …, yn) yn of M into R. The image is a left ideal I of R and the kernel is isomorphic to a submodule N of R(n − 1). We have an exact sequence
Since I is projective, this exact sequence splits (p. 150); hence M ≅ N I. Since N ⊂ R(n − 1), we may use induction to conclude that N is isomorphic to a direct sum of ≤ n − 1 left ideals. Then M is isomorphic to a direct sum m ≤ n left ideals of R. Since any left ideal is projective, it follows that M is projective.
We now assume that M is a finitely generated module over a Dedekind domain D. The defining property for D that we used is that any fractional ideal of D is invertible and hence is projective (Proposition 10.3). In particular, any integral ideal of D is projective and since 0 is trivially projective, every ideal in D is a projective module. Thus D satisfies the hypothesis of Proposition 10.12. Let T be the torsion submodule of M and put = M/T. Then is finitely generated and torsion free. Hence, by Proposition 10.11, is isomorphic to a submodule of a free module of finite rank. Then, by Proposition 10.12, is projective. Since we have an exact sequence 0 → T → M → → 0 and is projective, M ≅ T. Then T is a homomorphic image of Mand so it, too, is finitely generated. It is clear that we have reduced the problem of classifying finitely generated D-modules into isomorphism classes to the same problem for torsion-free modules and torsion modules.
We consider first the torsion-free modules. We have seen that any such module is projective. On the other hand, if M is projective, M is a direct summand of a free module, and since D is a domain, free modules and hence projective modules over D are torsion-free. Thus M is finitely generated torsion free over D if and only if M is finitely generated projective.
In determining conditions for isomorphism of two such modules, we may assume that the modules have the form M = I1 … In where Ij is an integral ideal. Although there is no gain in generality in doing so, it is somewhat more natural to assume that the Ij are fractional ideals. Consider the localization MD* determined by the monoid D* of non-zero elements of D. We have and since DD* is the field F of fractions of D and IjD* = F, MD* is an n-dimensional vector space over F. Thus MD* ≅ F(n). The dimensionality n is the rank of M over D as defined in section 7.7 (Corollary, p. 415). The canonical map x x/1 of M into MD* is a monomorphism. It follows that we have an isomorphism of M with the D-submodule of F(n) consisting of the vectors (x1, …, xn) where xj ∈ Ij. Hence we may assume that M consists of this set of vectors.
If η is an isomorphism of M onto a second finitely generated torsion-free module M′, η has a unique extension to a bijective linear transformation of MD*/F onto M′D*/F. It follows that M and M′ have the same rank n and we may assume also that M′ is the D-submodule of F(n) consisting of the vectors (x′, …, x′n) where x′j ∈ I′j, a fractional ideal in F, 1 ≤ j ≤ n. Taking into account the form of a bijective linear transformation of F(n), we see that η has the form
where A = (aij) ∈ GLn(F), the group of n × n invertible matrices with entries in F. We have x′k = ∑ xjajk ∈ I′k and if A− 1 = (bij), then xj = ∑ x′kbkj ∈ Ij if x′k ∈ I′k, 1 ≤ k ≤ n. Conversely, if A ∈ GLn(F) satisfies these conditions, then (23) is an isomorphism of M onto M′.
If n = 1, the foregoing conditions imply that two fractional ideals I and I′ are isomorphic if and only if there exists an a ≠ 0 in F such that I′ = Ia. We can state this result in terms of the class group of the Dedekind domain D, which we define to be the factor group / where is the multiplicative group of fractional ideals of D and is the subgroup consisting of the principal ideals aD, a ≠ 0 in F. The ideal I and aI = (aD)I are said to be in the same class. Thus we have shown that I and I′ are isomorphic D-modules if and only if I and I′ are in the same class. It is clear that we have a 1-1 correspondence between the isomorphism classes of fractional ideals and hence of torsion-free modules of rank one with the elements of the class group /.
We prove next the
LEMMA. Let I1 and I2 be fractional ideals. Then M = I1 I2 ≅ D I1I2.
Proof. We use exercise 10 on p. 605. According to this, if a1 ≠ 0 in I1 and J = a1I1− 1(⊂ D), then there exists an a2 ∈ I2 such that a1I1− 1 + a2I2 − 1 = D. Hence we have bi ∈ Ii − 1 such that a1b1 + a2b2 = 1. Then the matrix
is invertible with A− 1 = . If xj ∈ Ij, then y1 = x1b1 + x2b2 ∈ D and y2 = − x1a2 + x2a1 ∈ I1I2. On the other hand, if y1 ∈ D and y2 = c1c2, Cj ∈ Ij, then x1 = a1y1 − b2c1c2 ∈ I1 and x2 = a2y1 + b1c1c2 ∈ I2. Hence (x1, x2) (x1, x2)A is an isomorphism of I1 I2 onto D I1 I2.
We can now prove the main theorem on finitely generated modules without torsion over a Dedekind domain.
THEOREM 10.14. Any finitely generated torsion-free module M over a Dedekind domain is isomorphic to a direct sum of a finite number of integral ideals. Necessary and sufficient conditions for the isomorphism of M = I1 … In and M′ = I′1 … I′m where the Ij and I′k are fractional ideals are m = n and I1 … In and I′1 … I′m are in the same class.
Proof. The validity of the first statement has been noted before. Now suppose that M ≅ M′ where M and M′ are as in the statement of the theorem. Then m = n and we have a matrix A = (aij) ∈ GLn(F) such that if xj ∈ Ij, 1 ≤ j ≤ n, then x′k = ∑jn = 1 xjajk ∈ I′k. Moreover, if A− 1 = (bij) and x′k ∈ I′k, 1 ≤ k ≤ n, then xj = ∑x′kbkj ∈ Ij. The first condition implies that xjajk ∈ I′k for any xj ∈ Ij. Thus Ijajk ⊂ I′k, so (I′k)− 1Ijajk ⊂ D and ajk ∈ Ij− 1I′k. Hence if (kl, …, kn) is a permutation of 1, 2, …, n, then alk1 … ankn, ∈ ( Ij)− 1 and (I′k). This implies that det A ∈ (Ij)− 1(I′k) and (det A)Ij ⊂ I′j. Similarly, (det A− 1 )I′j ⊂ I′j. Hence I′j = (det A)Ij and so Ij and I′j are in the same class. Conversely, assume that m = n and that Ij and I′j are in the same class.By the lemma, M =I1 I2 … In ≅ D I1I2 I3 … In ≅ D D I1 I2 I3 I4 … In ≅ … ≅ D … D Ij. Similarly, M′ ≅ D … D I′j. Since Ij and I′j in the same class, these are isomorphic and hence M ≅ M′.
We consider next the structure of a finitely generated torsion module M over D. Our program for studying M will be to replace it by a closely related module over a p.i.d. Let I = annDM. If M = Dx1 + Dx2 + … + Dxn, then
Hence I = annDM ≠ 0. Let {P1, …, Pr} be the prime ideal divisors of I and let
Then S is a submonoid of the multiplicative monoid of D and we can form the ring Ds and the Ds-module Ms = Ds DM. Any prime ideal of Ds has the form Ps where P is a prime ideal such that P ∩ S = Ø (p. 401). Then P ⊂ Pi and hence P ⊂ Pi for some i (Proposition 7.1, p. 390), which implies P = Pi. Thus Ds is a Dedekind domain with only a finite number of prime ideals, P′i = Pis, 1 ≤ i ≤ r. Such a domain is a p.i.d. (exercise 15, p. 625). Moreover, Ms is a torsion Ds-module since Is ≠ 0 and Is Ms = 0.
The fundamental theorem on finitely generated modules over a p.i.d. (BAI, p. 187) is applicable to the Ds-module Ms. According to this, Ms = Dsy1 … Dsym where annDsy1 ⊃ annDsy2 ⊃ … ⊃ annDsym. Moreover, annDsyi ⊃ Is ≠ 0 and annDsyi is a proper ideal in Ds. We shall be able to apply this result to determining the structure of M as D-module, since we can recover M from Ms by using the following
LEMMA. Let I be a non-zero ideal in D, P1, …, Pr the prime ideal divisors of I in D, S = (D − Pi). Let N be a D-module such that IN = 0. Then N ≅ (D/I) DNs as D-modules.
Proof. We have (D/I)DNs ≅ (D/I)D(Ds DN)≅ ((D/I) D Ds)DN ≅ (D/I)s DN. It is readily seen that the elements s + I, s ∈ S, are units in D/I, which implies that (D/I)s ≅ D/I. Hence (D/I)s DN ≅ N, since IN = 0.
We can now prove the following structure theorem.
THEOREM 10.15. Let M be a finitely generated torsion module over a Dedekind domain D. Then M is isomorphic to a direct sum of cyclic modules Dz1 … Dzm such that ann z1 ⊃ ann z2 ⊃ … ⊃ann zm 0. Moreover, the ideals ann zi, 1 ≤ i ≤ m, are uniquely determined.
Proof. Let S be as before. Then Ms = Dsy1 … Dsym where annDsy1 ⊃ … ⊃ annDsym. Then if I′j = annDsyj, Dsyj ≅ Ds/I′j. Let Ij = {a ∈ D|a/s ∈ I′j for some s ∈ S}. Then Ijs = I′j (p. 401) and D/Ij is a cyclic D-module, Dzj, such that (D/Ij)s ≅ Ds/Ijs = Ds/I′j. Moreover, annDzj = Ij and I1 ⊃ … ⊃ Im. Now consider the module N = Dz1 … Dzm. We have Ns = (Dz1)s … (Dzm)s ≅ Dsy1 … Dsym ≅ Ms. Hence by the lemma, M ≅ N = Dz1 … Dzm. The uniqueness follows also by using S-localization and the corresponding uniqueness result in the p.i.d. case (BAI, p. 192). We leave the details to the reader.
Theorems 10.14 and 10.15 and the fact that any finitely generated module over a Dedekind domain D is a direct sum of its torsion submodule and a torsion-free module constitute a direct generalization of the fundamental theorem on finitely generated modules over a p.i.d. If D is a p.i.d., then the class group / = 1 and the results show that M is a direct sum of copies of D and cyclic modules Dz1, …, Dzm with non-zero annihilators I1, …, Im such that I1 ⊃ … ⊃ Im. This is the fundamental theorem as given in BAI, p. 187. The Dedekind domain is a p.i.d. if and only if / = 1. The group / gives a measure of the departure of D from being a p.i.d. It is a classical group of algebraic number theory. There it is shown by transcendental methods that the class group is finite. On the other hand, it has been shown by L. Claborn (Pacific J. of Math., vol. 18 (1966), 219–222) that any abelian group is the class group of a suitable Dedekind domain.
We shall now show that the class group coincides with the projective class group, which we defined in section 7.4. The latter, denoted as Pic D, is the set of isomorphism classes of faithful finitely generated projective modules of rank one with multiplication defined by the tensor product. Since D is a domain, projective D-modules are automatically faithful. Hence the elements of Pic D can be represented by the fractional ideals, and as we have seen, two such ideals are isomorphic if and only if they are in the same class. Hence we have the bijective map [I] I of Pic D onto / where [I] denotes the isomorphism class of the fractional ideal I. Now we have the canonical module homomorphism of I1 I2 into I1I2 sending a1 a2 into a1 a2. Localization at all of the maximal ideals show that this is an isomorphism. Thus I1 I2 ≅ I1I2 and hence [I] I is a group isomorphism of Pic D with the class group.
REFERENCES
O. Zariski and P. Samuel, Commutative Algebra vol. I, D. van Nostrand, 1958. New printing, Springer-Verlag, New York, 1975.
S. Lang, Algebraic Number Theory, Addison-Wesley, Reading, Mass., 1970.
______,Introduction to Algebraic and Abelian Functions, Addison-Wesley, Reading, Mass., 1972.
G. J. Janusz, Algebraic Number Fields, Academic Press, New York, 1973.