IDEAL CLASSES AND RELATIVE INTEGERS

Similar documents
STABLY FREE MODULES KEITH CONRAD

x y B =. v u Note that the determinant of B is xu + yv = 1. Thus B is invertible, with inverse u y v x On the other hand, d BA = va + ub 2

(1) A frac = b : a, b A, b 0. We can define addition and multiplication of fractions as we normally would. a b + c d

Math 121 Homework 5: Notes on Selected Problems

5 Dedekind extensions

Finitely Generated Modules over a PID, I

Math 210B: Algebra, Homework 4

UNIVERSAL IDENTITIES. b = det

Honors Algebra 4, MATH 371 Winter 2010 Assignment 4 Due Wednesday, February 17 at 08:35

Algebra Homework, Edition 2 9 September 2010

NOTES ON LINEAR ALGEBRA OVER INTEGRAL DOMAINS. Contents. 1. Introduction 1 2. Rank and basis 1 3. The set of linear maps 4. 1.

18.312: Algebraic Combinatorics Lionel Levine. Lecture 22. Smith normal form of an integer matrix (linear algebra over Z).

THE MINIMAL POLYNOMIAL AND SOME APPLICATIONS

ALGEBRA EXERCISES, PhD EXAMINATION LEVEL

ZORN S LEMMA AND SOME APPLICATIONS

MINKOWSKI THEORY AND THE CLASS NUMBER

ALGEBRAIC GEOMETRY COURSE NOTES, LECTURE 2: HILBERT S NULLSTELLENSATZ.

LINEAR ALGEBRA BOOT CAMP WEEK 1: THE BASICS

MODULES OVER A PID. induces an isomorphism

Algebraic Number Theory

Introduction to modules

Thus, the integral closure A i of A in F i is a finitely generated (and torsion-free) A-module. It is not a priori clear if the A i s are locally

ALGEBRA QUALIFYING EXAM PROBLEMS LINEAR ALGEBRA

Injective Modules and Matlis Duality

Dedekind Domains. Mathematics 601

ALGEBRA HW 3 CLAY SHONKWILER

Math 120 HW 9 Solutions

MATH 8253 ALGEBRAIC GEOMETRY WEEK 12

GALOIS GROUPS AS PERMUTATION GROUPS

11. Finitely-generated modules

Mathematics Department Stanford University Math 61CM/DM Vector spaces and linear maps

5 Dedekind extensions

Algebraic structures I

Math 762 Spring h Y (Z 1 ) (1) h X (Z 2 ) h X (Z 1 ) Φ Z 1. h Y (Z 2 )

Math 110, Summer 2012: Practice Exam 1 SOLUTIONS

ISOMORPHISMS KEITH CONRAD

TENSOR PRODUCTS II KEITH CONRAD

MATH PRACTICE EXAM 1 SOLUTIONS

SPRING 2006 PRELIMINARY EXAMINATION SOLUTIONS

Solution. That ϕ W is a linear map W W follows from the definition of subspace. The map ϕ is ϕ(v + W ) = ϕ(v) + W, which is well-defined since

Integral Extensions. Chapter Integral Elements Definitions and Comments Lemma

Elementary linear algebra

3.1. Derivations. Let A be a commutative k-algebra. Let M be a left A-module. A derivation of A in M is a linear map D : A M such that

Linear Algebra. Min Yan

Algebra Exam Syllabus

Introduction to Arithmetic Geometry Fall 2013 Lecture #18 11/07/2013

D-MATH Algebra I HS 2013 Prof. Brent Doran. Exercise 11. Rings: definitions, units, zero divisors, polynomial rings

Modules Over Principal Ideal Domains

NONCOMMUTATIVE POLYNOMIAL EQUATIONS. Edward S. Letzter. Introduction

ALGEBRA HW 4. M 0 is an exact sequence of R-modules, then M is Noetherian if and only if M and M are.

Tensor Product of modules. MA499 Project II

Math 210B. Artin Rees and completions

GALOIS GROUPS OF CUBICS AND QUARTICS (NOT IN CHARACTERISTIC 2)

Definition 2.3. We define addition and multiplication of matrices as follows.

2.2. Show that U 0 is a vector space. For each α 0 in F, show by example that U α does not satisfy closure.

CYCLICITY OF (Z/(p))

OSTROWSKI S THEOREM FOR Q(i)

φ(xy) = (xy) n = x n y n = φ(x)φ(y)

MATH 326: RINGS AND MODULES STEFAN GILLE

BENJAMIN LEVINE. 2. Principal Ideal Domains We will first investigate the properties of principal ideal domains and unique factorization domains.

Modules over Principal Ideal Domains

Topics in linear algebra

Graduate Preliminary Examination

Computing Invariant Factors

Number Theory Fall 2016 Problem Set #3

1 Linear Algebra Problems

Lecture Notes Math 371: Algebra (Fall 2006) by Nathanael Leedom Ackerman

120A LECTURE OUTLINES

h Q h Ç h - In S.

Eighth Homework Solutions

Bulletin of the Iranian Mathematical Society

DUAL MODULES KEITH CONRAD

and this makes M into an R-module by (1.2). 2

The most important result in this section is undoubtedly the following theorem.

Review of Linear Algebra

Chapter 2 Linear Transformations

GRE Subject test preparation Spring 2016 Topic: Abstract Algebra, Linear Algebra, Number Theory.

TROPICAL SCHEME THEORY

0.2 Vector spaces. J.A.Beachy 1

Answers in blue. If you have questions or spot an error, let me know. 1. Find all matrices that commute with A =. 4 3

THE FUNDAMENTAL THEOREM OF ALGEBRA VIA PROPER MAPS

First we introduce the sets that are going to serve as the generalizations of the scalars.

Chapter 1 Vector Spaces

EXTERIOR POWERS KEITH CONRAD

ADVANCED COMMUTATIVE ALGEBRA: PROBLEM SETS

This is a closed subset of X Y, by Proposition 6.5(b), since it is equal to the inverse image of the diagonal under the regular map:

Solution to Homework 1

Linear Equations in Linear Algebra

Pacific Journal of Mathematics

SPLITTING OF SHORT EXACT SEQUENCES FOR MODULES. 1. Introduction Let R be a commutative ring. A sequence of R-modules and R-linear maps.

(1.) For any subset P S we denote by L(P ) the abelian group of integral relations between elements of P, i.e. L(P ) := ker Z P! span Z P S S : For ea

Lecture 7. This set is the set of equivalence classes of the equivalence relation on M S defined by

Lecture Summaries for Linear Algebra M51A

MATH 101B: ALGEBRA II PART A: HOMOLOGICAL ALGEBRA

JORDAN NORMAL FORM. Contents Introduction 1 Jordan Normal Form 1 Conclusion 5 References 5

Math 4310 Solutions to homework 1 Due 9/1/16

GROUPS OF ORDER p 3 KEITH CONRAD

MATH 320: PRACTICE PROBLEMS FOR THE FINAL AND SOLUTIONS

Linear Algebra. Paul Yiu. Department of Mathematics Florida Atlantic University. Fall 2011

Transcription:

IDEAL CLASSES AND RELATIVE INTEGERS KEITH CONRAD The ring of integers of a number field is free as a Z-module. It is a module not just over Z, but also over any intermediate ring of integers. That is, if E Q we can consider O E as an O -module. Since O E is finitely generated over Z, it is also finitely generated over O (just a larger ring of scalars), but O E may or may not have a basis over O. When we treat O E as a module over O, rather than over Z, we speak about a relative extension of integers. If O is a PID then O E will be a free O -module, so O E will have a basis over O. Such a basis is called a relative integral basis for E over. The next three examples illustrate some possibilities when O is not a PID. Example 1. Let = Q( 5) and E = Q(i, 5) = (i). Although O = Z[ 5] is not a PID, O E is a free O -module with relative integral basis {1, i+ 5 }. Example. Let = Q( 15) and E = Q( 15, 6) = ( 6). Then h( ) =, so O is not a PID, but it turns out that O E = O O 6, so OE is a free O -module. Example 3. Let = Q( 6) and E = Q( 6, 3) = ( 3). Then h( ) =, so O is not a PID, and it turns out that (1) O E = O e 1 pe, where e 1 = 1+ 3, e = 1 3, and p = (3, 6). (Although e is not in O E, there isn t a problem with the direct sum decomposition (1) for O E over O since the coefficients of e run not over O but over the ideal p, which doesn t include 1, so e pe.) Equation (1) says that as an O -module, O E p. The ideal p is not principal and this suggests O E is not a free O -module, although that does require an argument. To reinforce this point, p p does not look like a free O -module, since p is not principal, but p p has a second direct sum decomposition that admits an O -basis, so a direct sum of two non-free modules can be free. We will see how in Example 9 below. What we are after is a classification of finitely generated torsion-free modules over a Dedekind domain, which will then be applied in the number field setting to describe O E as an O -module. The extent to which O E could fail to have an O -basis will be related to ideal classes in. A technical concept we need to describe modules over a Dedekind domain is projective modules. Definition 4. Let A be any commutative ring. An A-module P is called projective if every surjective linear map f : M P from any A-module M onto P looks like a projection out of a direct sum: there is an isomorphism h: M = P N for some A-module N such that h(m) = (f(m), ) for all m M. The isomorphism h is not unique. or example, taking A = Z, P = Z, and M = Z Z with f(a, b) = a b, we can use h: M P Z by h(a, b) = (a b, b) or h(a, b) = 1

KEITH CONRAD (a b, a b). Each of these works since the first coordinate of h(a, b) is f(a, b) and h is obviously invertible. The complementary summand N in the definition of a projective module is isomorphic to the kernel of f. Indeed, the condition h(m) = (f(m), ) means f(m) = 0 if and only if h(m) is in {0} N, which means h restricts to an isomorphism between ker f and {0} N = N. It is easy to give examples of non-projective modules. or instance, if P is a projective A-module with n generators there is a surjective A-linear map A n P, so A n = P Q for some A-module Q. When A is a domain, any submodule of A n is torsion-free, so a finitely generated projective module over a domain is torsion-free. Therefore a finitely generated module over a domain that has torsion is not projective: Z Z/() is not a projective Z-module. More importantly for us, though, is that fractional ideals in a Dedekind domain are projective modules. Lemma 5. or a domain A, any invertible fractional A-ideal is a projective A-module. In particular, when A is a Dedekind domain all fractional A-ideals are projective A-modules. Proof. Let a be an invertible fractional A-ideal. Then k x iy i = 1 for some x i a and y i a 1. or each x a, x = 1 x = x 1 (x 1x) + + x k (x k x) and x i x a 1 a = A, so a k Ax i a, so a = Ax 1 + + Ax k. In a similar way, a 1 = Ay 1 + + Ay k. Suppose f : M a is a surjective A-linear map. Choose m i M such that f(m i ) = x i. Define g : a M by g(x) = k (xy i)m i. Note xy i aa 1 = A for all i, so g(x) makes sense and g is A-linear. Then f(g(x)) = k (xy i )f(m i ) = n (xy i )x i = x n x i y i = x. Check the A-linear map h: M a ker f given by the formula h(m) = (f(m), m g(f(m))) has inverse (x, y) g(x) + y. Here is the main structure theorem. Theorem 6. Every finitely generated torsion-free module over a Dedekind domain A is isomorphic to a direct sum of ideals in A. Proof. Let M be a finitely generated torsion-free A-module. We can assume M 0 and will show there is an embedding M A d for some d 1 such that the image of M intersects each standard coordinate axis of A d. Let be the fraction field of A and x 1,..., x n be a generating set for M as an A-module. We will show n is an upper bound on the size of any A-linearly independent subset of M. Let f : A n M be the linear map where f(e i ) = x i for all i. (By e 1,..., e n we mean the standard basis of A n.) Let y 1,..., y k be linearly independent in M, so their A-span is isomorphic to A k. Write y j = n a ijx i with a ij A. We pull the y j s back to A n by setting v j = (a 1j,..., a nj ), so f(v j ) = y j. A linear dependence relation on the v j s is transformed by f into a linear dependence relation on the y j s, which is a trivial relation by their linear independence. Therefore v 1,..., v k is A-linearly independent in A n, hence -linearly independent in n. By linear algebra over fields, k n. rom the bound k n, there is a linearly independent subset of M with maximal size, say t 1,..., t d. Then d j=1 At j = A d by identifying t j with the jth standard basis

IDEAL CLASSES AND RELATIVE INTEGERS 3 vector in A d. We will find a scalar multiple of M inside d j=1 At j. or any x M, the set {x, t 1,..., t d } is linearly dependent by maximality of d, so there is a nontrivial linear relation a x x + d a it i = 0, necessarily with a x 0 in A. Thus a x x d j=1 At j. Letting x run through the spanning set x 1,..., x n, we have ax i d j=1 At j for all i where a = a x1 a xn 0. Thus am d j=1 At j. Multiplying by a is an isomorphism of M with am, so we have the sequence of A-linear maps M am d At j A d, j=1 where the first and last maps are A-module isomorphisms. In the above composite map, t j M is mapped to ae j in A d, so this composite map is an embedding M A d such that M meets each standard coordinate axis of A d in a nonzero vector. Compose this linear map with projection A d A onto the last coordinate in the standard basis: a 1 e 1 + + a d e d a d. Denote the restriction of this to a map M A as ϕ, so a := ϕ(m) is a nonzero ideal in A. With ϕ we get a surjective map M a, so Lemma 5 (the first time we need A to be a Dedekind domain, not just an integral domain) tells us M = a ker ϕ. Obviously ker ϕ A d 1 0 = A d 1, so ker ϕ is a finitely generated (and torsion-free) A-module with at most d 1 A-linearly independent elements. Using induction on the largest number of linearly independent elements in the module, ker ϕ is a direct sum of ideals in A. Remark 7. Using equations rather than isomorphisms, Theorem 6 says M = M 1 M d where each M i is isomorphic to an ideal in A. Those ideals need not be principal, so M i need not have the form Am i. If M is inside a vector space over the fraction field of A, then M = d a ie i for some linearly independent e i s, but be careful: if a i is a proper ideal in A then e i is not in M since 1 a i. The e i s are not a spanning set for M as a module since their coefficients are not running through A. The decomposition of the integers of Q( 6, 3) as a module over Z[ 6] in Example 3 illustrates this point. How much does a direct sum a 1 a d, as a module, depend on the individual a i s? Lemma 8. Let A be a Dedekind domain. A-module isomorphism a b = A ab. or fractional A-ideals a and b, there is an Proof. Both sides of the isomorphism are unchanged up to A-module isomorphism when we scale a and b, so without loss of generality a and b are nonzero ideals in A. We can further scale so a and b are relatively prime. Indeed, let a 1 a 0 where a 0 A. Using the Chinese remainder theorem in A, there is a nonzero ideal c such that a 0 c is principal and gcd(c, b) = (1). Since c a 1 0 a, we can replace a by c without changing a b or A ab up to A-module isomorphism. The linear map f : a b a + b = A given by f(a, b) = a b is surjective and ker f = {(a, a) : a a b} = a b, which is ab since gcd(a, b) = (1). Applying Lemma 5 to the fractional A-ideal A, a b = A ker f = A ab. Example 9. or A = Z[ 5], let p = (, 1 + 5), so p is not principal but p = A is principal. Then there is an A-module isomorphism p p = A p = A A. That is intriguing: p does not have an A-basis but p p does! Working through the proof of

4 KEITH CONRAD Lemma 8 will show you how to write down a basis of p p explicitly. In a similar way, p p in Example 3 is a free Z[ 6]-module since p is principal. Theorem 10. Let A be a Dedekind domain. or fractional A-ideals a 1,..., a d, there is an A-module isomorphism a 1 a d = A d 1 a 1 a d. Proof. Induct on d and use Lemma 8. Corollary 11. Let E/ be a finite extension of number fields with [E : ] = n. As an O -module, O E n 1 a for some nonzero ideal a in O. Proof. Since O E is a finitely generated Z-module it is a finitely generated O -module and obviously has no torsion, so Theorems 6 and 10 imply O E d 1 a for some d 1 and nonzero ideal a in O. Letting m = [ : Q], both O and a are free of rank m over Z, while O E is free of rank mn over Z. Computing the rank of O E and O d 1 a over Z, mn = m(d 1) + m = md, so d = n. Thus O E is almost a free O -module. If a is principal then O E is free. As an O -module up to isomorphism, O n 1 a only depends on a through its ideal class, since a and any xa (x ) are isomorphic O -modules. Does O n 1 a, as an O -module, depend on a exactly through its ideal class? That is, if O n 1 a n 1 b as O -modules, does [a] = [b] in Cl( )? The next two theorems together say the answer is yes. Theorem 1. Let A be a domain with fraction field. or fractional A-ideals a and b in, a = b as A-modules if and only if a = xb for some x. Here is any field, not necessarily a number field. Proof. ( ): Multiplication by x is an A-module isomorphism from b to a. ( ): Suppose f : a b is an A-module isomorphism. We want an x such that f(t) = xt for all t a. or this to be possible, f(t)/t has to be independent of the choice of nonzero t. Then we could define x to be this common ratio, so f(t) = xt for all t in a (including t = 0). or any nonzero t 1 and t in a, f(t 1 )? f(t ) = t f(t 1 ) =? t 1 f(t ). t 1 t You may be tempted to pull the t and t 1 inside on the right, confirming the equality, but that is bogus because f is A-linear and we don t know if t 1 and t are in A (they are just in ). This is easy to fix. Since a is a fractional A-ideal, it has a denominator: da A for some nonzero d A. Then dt 1, dt A, so t f(t 1 )? = t 1 f(t ) dt f(t 1 )? = dt 1 f(t ) f(dt t 1 ) = f(dt 1 t ). Theorem 13. or nonzero ideals a 1,..., a m and b 1,..., b n in a Dedekind domain A,we have a 1 a m = b1 b n as A-modules if and only if m = n and [a 1 a m ] = [b 1 b n ] in Cl(A). Proof. The if direction follows from Theorems 10 and 1: a 1 a m = A m 1 a 1 a m, so if m = n and [a 1 a m ] = [b 1 b m ] in Cl(A) then A m 1 a 1 a m = A m 1 b 1 b m = b1 b m.

IDEAL CLASSES AND RELATIVE INTEGERS 5 Turning to the only if direction, we show m is determined by the A-module structure of a 1 a m : it is the largest number of A-linearly independent elements in this module. Picking a nonzero a i a i, the m-tuples (..., 0, a i, 0,... ) for 1 i m are easily A-linearly independent, so m a i has m linearly independent members. Since a 1 a m m, where is the fraction field of A, any set of more than m members of a 1 a m has a nontrivial -linear relation in m, which can be scaled to a nontrivial A-linear relation in m a i by clearing a common denominator in the coefficients. Therefore if a 1 a m = b 1 b n as A-modules we must have m = n by computing the maximal number of A-linearly independent elements in both modules. To show m a i = m b i a 1 a m and b 1 b m are scalar multiples, we can collect ideals by multiplication into the last summands: it is enough to show A m 1 a = A m 1 b a and b are scalar multiples. Let ϕ: A m 1 a A m 1 b be an A-module isomorphism. Viewing A m 1 a and A m 1 b as column vectors of length m with the last coordinate in a or b, ϕ can be represented as an m m matrix of A-linear maps (ϕ ij ), where ϕ ij has domain A or a and target A or b. The proof of Theorem 1 shows any A-linear map from one fractional A-ideal to another (not necessarily injective or surjective) is a scaling function. Therefore ϕ is described by an m m matrix of numbers, say M, acting in the usual way on column vectors. or any α a, let D α = diag(1,..., 1, α) be the diagonal m m matrix with α in the lower right entry. Then D α (A m 1 A) A m 1 a, so MD α maps A m 1 A to A m 1 b. This means the bottom row of MD α has all entries in b, so det(md α ) b. Since D α has determinant α and α is arbitrary in a, det(m)a b. In the same way, det(m 1 )b a, so det(m)a = b. Example 14. Let s return to Example 3: = Q( 6), E = ( 3), and O E p, where p = (3, 6). We can show O E is not a free O -module: if it were free then O E, so O p O as O -modules. Then Theorem 13 implies p as O -modules, so p is principal, but p is nonprincipal. This is a contradiction. We can now associate to any finite extension of number fields E/ a canonical ideal class in Cl( ), namely [a] where O E n 1 a as O -modules. Theorem 13 assures us [a] is well-defined. Since the construction of [a] is due to Steinitz (191), [a] is called the Steinitz class of E/. Example 15. By Example 14, when = Q( 6) the nontrivial member of Cl( ) is the Steinitz class of the quadratic extension ( 3)/. Since the ideal class group of a number field is finite, as E varies over all extensions of with a fixed degree n the different O E s have finitely many possible O -module structures, in fact at most h( ) of them. There are infinitely many nonisomorphic extensions of with degree n, so it is natural to ask if each ideal class is realized among them: for any [a] Cl( ) and integer n is there some extension E/ of degree n whose Steinitz class is [a], i.e., O E n 1 a as O -modules? The answer is yes for n = and n = 3, and in fact Kable and Wright [1] have proved an equidistribution of the E s with degree or 3 over in terms of their Steinitz classes in Cl( ). In particular, each ideal class in Cl( ) is a Steinitz class for infinitely many nonisomorphic degree n extensions of when n = or 3. The extension to general degrees is still an open problem in general. We have focused on the description of a single finitely generated torsion-free module over a Dedekind domain. What if we want to compare such a module and a submodule? If A

6 KEITH CONRAD is a PID, M is a finite free A-module and M is a submodule then we can align M and M in the sense that there is a basis e 1,..., e n of M and nonzero scalars a 1,..., a m in A (where m n) such that M = n Ae i and M = m j=1 Ae j. This has an analogue over Dedekind domains, but it doesn t use bases. If A is a Dedekind domain, M is a finitely generated torsion-free A-module, and M is a submodule of M, then we can align M and M in the sense that we can write n m M = M i and M = a j M j where m n, each M i is isomorphic to a nonzero ideal in A, and each a j is a nonzero ideal in A. It is generally false that such an alignment is compatible with an isomorphism M = A n 1 a. That is, such an isomorphism need not restrict to M to give an isomorphism M = A m 1 a with a a, even for A = Z. Consider, for instance, M = Z Z and M = az az for a > 1 (so m = n = ). We can t write M = Ze 1 Ze and M = Ze 1 bze for some integer b since that would imply M/M = Z/bZ is cyclic, whereas M/M = (Z/aZ) is not cyclic. References [1] A. C. Kable and D. J. Wright, Uniform distribution of the Steinitz invariants of quadratic and cubic extensions, Compositio Math. 14 (006), 84 100. j=1