Fitting ideals of class groups for CM abelian extensions
Abstract.
Let be a finite abelian CM-extension and a suitable finite set of finite primes of . In this paper, we determine the Fitting ideal of the minus component of the -ray class group of , except for the -component, assuming the validity of the equivariant Tamagawa number conjecture. As an application, we give a necessary and sufficient condition for the Stickelberger element to lie in that Fitting ideal.
Key words and phrases:
class groups, Fitting ideals, CM-fields, equivariant Tamagawa number conjecture2010 Mathematics Subject Classification:
11R291. Introduction
In number theory, the relationship between class groups and special values of -functions is of great importance. In this paper we discuss such a phenomenon for a finite abelian CM-extension , that is, a finite abelian extension such that is a totally real field and is a CM-field. We focus on the minus components of the (ray) class groups of , except for the -components, and study the Fitting ideals of them.
Let denote the ideal class group of . Let denote the minus component after inverting the multiplication by . When , Kurihara and Miura [10] succeeded in proving a conjecture of Kurihara [7] on a description of the Fitting ideal of using the Stickelberger elements. However, for a general totally real field , the problem to determine the Fitting ideal of is still open.
There seems to be an agreement that the Pontryagin duals (denoted by ) of the class groups are easier to deal with (see Greither-Kurihara [4]). In [3], Greither determined the Fitting ideal of , assuming that the equivariant Tamagawa number conjecture (eTNC for short) holds and that the group of roots of unity in is cohomologically trivial. Subsequently, Kurihara [9] generalized the results of Greither on to results on , where denotes the -ray class group, for a finite set of finite primes of . This enables us, by taking suitably large , to remove the assumption that the group of roots of unity is cohomologically trivial, though we still need to assume the validity of eTNC. In recent work [2], Dasgupta and Kakde succeeded in proving unconditionally the same formula as Kurihara on the Fitting ideal of (see (1.2) below for the formula).
In this paper, for a general totally real field , we determine the Fitting ideal of without the Pontryagin dual, assuming eTNC, except for the -component. This problem has been considered to be harder than that on and actually our result is more complicated. Our main tool is the technique of shifts of Fitting ideals, which was established by the second author in [6].
As an application of the description, we will obtain a necessary and sufficient condition for the Stickelberger element to be in the Fitting ideal of (assuming eTNC). Note that the question for the dualized version is called the strong Brumer-Stark conjecture and is answered affirmatively by Dasgupta-Kakde [2] unconditionally.
Though we mainly assume the validity of eTNC in this paper, we also obtain interesting unconditional results. For instance, we will show that the Fitting ideal of is always contained in that of , and that the inclusion is often proper.
In the rest of this section, we give precise statements of the main results.
1.1. Description of the Fitting ideal
Let be a finite abelian CM-extension and put . Let be the set of archimedean places of . Let be the set of places of which are ramified in , including . For each finite prime , let denote the inertia group of in and the arithmetic Frobenius of . We then define elements and by
where we put . These elements are introduced in [3, Lemmas 6.1 and 8.3] and [9, §2.2, equations (7) and (10)] (up to the involution). Moreover, we define a -module by
We write , where is the complex conjugation in . For any -module , we also define the minus component by . Note that we are implicitly inverting the action of . For any , we write for the image of under the natural map .
In general, for a set of places of , we write for the set of places of which lie above places in . We take and fix a finite set of finite primes of satisfying the following.
-
•
.
-
•
is torsion free. Here, denotes the normalized additive valuation.
Note that, if we fix an odd prime number and are concerned with the -components, the last condition can be weakened to that is -torsion-free. We consider the -ray ideal class group of defined by
where runs over the finite primes of which are not in .
For a character of , we write for the primitive -function for ; this function omits exactly the Euler factors of primes dividing the conductor of . For any finite prime of , we put , where is the residue field of . We then define the -modified -function by
We define
(1.1) |
where runs over the characters of and is the idempotent of the -component.
Now the first main theorem of this paper is the following, whose proof will be given in §3.
Theorem 1.1.
Assume that eTNC for holds. Then we have
where is the first shift of the Fitting ideal (see Definition 2.3).
In the second main result below, we will obtain a concrete description of , which completes the description of the Fitting ideal of . We do not review the precise statement of eTNC (see e.g. [1, Conjecture 3.6]).
In order to compare with Theorem 1.1, we recall the result for the dualized version:
(1.2) |
As already mentioned, Kurihara [9, Corollary 3.7] showed this formula under the validity of the eTNC, and recently Dasgupta-Kakde [2, Theorem 1.4] removed the assumption. Here, for a general -module , we equip the Pontryagin dual with the -action by for , , and . This convention is the opposite of [9] and [2], so the right hand side of the formula (1.2) differs from those by the involution.
We now briefly outline the proof of Theorem 1.1. An important ingredient is an exact sequence of -modules of the form
as in Proposition 3.2, where is free of finite rank . Here, is an auxiliary finite set of places of . This sequence was constructed by Kurihara [9], based on preceding work such as Ritter-Weiss [12] and Greither [3], and played a key role in proving (1.2) under eTNC. Our novel idea is to construct an explicit injective homomorphism from to whose cokernel is isomorphic to the direct sum of for . Moreover, assuming eTNC, we will compute the determinant of the composite map . By using these observations, we obtain an exact sequence to which the theory of shifts of Fitting ideals can be applied, and then Theorem 1.1 follows.
1.2. Computation of the shift of Fitting ideal
In order to make the formula of Theorem 1.1 more explicit, in §4, we will compute . This will be accomplished by using a similar method as Greither-Kurihara [5, §1.2], which was actually a motivation for introducing the shifts of Fitting ideals in [6].
As the problem is purely algebraic, we deal with a general situation as follows (it should be clear from the notation how to apply the results below to the arithmetic situation; simply add subscripts appropriately). Let be a finite abelian group. Let and be subgroups of such that and that the quotient is a cyclic group. We choose a generator of and put
which are non-zero-divisors. We define a finite -module by
In order to state the result, we introduce some notations. We choose a decomposition
(1.3) |
as an abelian group such that is a cyclic group for each . Here, we do not assume any minimality on this decomposition, so we allow even the extreme case where is trivial for some .
For each , we put . We also put .
Definition 1.2.
For , we define as the ideal of generated by where runs over all tuples of integers satisfying , that is,
We clearly have . We then define an ideal of by
Note that the definition of does depend on the choice of the decomposition (1.3). On the other hand, it can be shown directly that the ideal is independent from the choice. We omit the direct proof because, at any rate, the independency can be deduced from the discussion in §4.
Example 1.3.
When , we have .
When , we have
When , we have
In this setting, we can describe as follows. It is convenient to state the result after multiplying by .
Theorem 1.4.
We have
as fractional ideals of .
1.3. Stickelberger element and Fitting ideal
As an application of Theorems 1.1 and 1.4, we shall discuss the problem whether or not the Stickelberger element lies in the Fitting ideal of .
We return to the setup in §1.1. Let be a fixed odd prime number and we shall work over . Let denote the maximal subgroup of of order prime to . We put , which is the maximal -extension of contained in . For each character of , we regard as a -module via , and put . For a -module , we put , which is an -module. For an element , we write for the image of by the natural map . We note that is isomorphic to the direct product of if runs over the equivalence classes of characters of .
From now on, we fix an odd character of . We define . Then is a CM-field, , and is a cyclic extension of order prime to .
We put and consider the -component of the Stickelberger element defined by
(1.4) |
where runs over characters of whose restriction to coincides with and we write
Note that, comparing (1.1) and (1.4), we have
(1.5) |
Concerning the dualized version, by Dasgupta-Kakde [2, Theorem 1.3],
is always true. This is called the strong Brumer-Stark conjecture. More precisely, the displayed claim is a bit stronger than [2, Theorem 1.3] as we took instead of in the definition of the Stickelberger element, but in any case it is an immediate consequence of the formula (1.2).
On the other hand, the corresponding claim without dual is known to be false in general (see [4]). However, we had only partial results and an exact condition was unknown. The following theorem is strong as it gives a necessary and sufficient condition.
Theorem 1.5.
Assume that eTNC for holds. Then, for each odd character of , the following are equivalent.
-
(i)
We have .
-
(ii)
We have either or, for any , one of the following holds.
-
(a)
does not split completely in .
-
(b)
The inertia group is cyclic.
-
(a)
This theorem will be proved in §5 as an application of Theorems 1.1 and 1.4. Note that there is an elementary equivalent condition for as in Lemma 5.3.
Theorem 1.5 indicates that the failure of the inertia groups to be cyclic is an obstruction for studying the Fitting ideal of the class group without dual. The same phenomenon will appear again in Theorem 1.6 below. We should say that this kind of phenomenon had been observed in preceding work, such as Greither-Kurihara [4]. It is also remarkable that the obstruction does not occur in the absolutely abelian case (i.e. when ), since in that case the inertia groups are automatically cyclic, apart from the -parts. This seems to fit the fact that Kurihara and Miura [7], [10] succeeded in studying the class groups without dual in the absolutely abelian case.
Let us outline the proof of Theorem 1.5. We assume that is a faithful character of (i.e. ); actually we can deduce the general case from this case. Since is a non-zero-divisor of , by Theorem 1.1 and (1.5), we have if and only if
(1.6) |
holds as fractional ideals of , where on both sides runs over the elements of .
Obviously we may assume that . The proof of (ii) (i) is the easier part. We will show that, under the assumption (ii), the inclusion of (1.6) holds even for every before taking the product. On the other hand, the opposite direction (i) (ii) is the harder part. That is because, roughly speaking, we have to work over the ring , whose ring theoretic properties are not very nice. A key idea to overcome this issue is to reduce to a computation in a discrete valuation ring. More concretely, we make use of a character of which satisfies and a certain additional condition, whose existence is verified by Lemma 5.3, and we consider the -algebra . By investigating the ideals in (1.6) after base change from to , we will show (i) (ii).
1.4. Unconditional consequences
Even if we do not assume the validity of eTNC, our argument shows the following.
Theorem 1.6.
We have an inclusion
Moreover, the inclusion is an equality if is cyclic for every .
This theorem follows immediately from Corollaries 3.6 and 4.2. Furthermore, by similar arguments as the proof of Theorem 1.5, we can observe that the inclusion is often proper.
As already remarked, Dasgupta-Kakde [2] proved the formula (1.2) unconditionally. Therefore, if is cyclic for every , we can also deduce from Theorem 1.6 that also coincides with that ideal, and this removes the assumption on eTNC in Theorem 1.1. However, in Theorem 1.1 we still need to assume eTNC when is not cyclic for some .
2. Definition of Fitting ideals and their shifts
In this section, we fix our notations concerning Fitting ideals.
2.1. Fitting ideals
Let be a noetherian ring.
Definition 2.1 (cf. Northcott [11]).
We define the Fitting ideals as follows.
-
(i)
Let be a matrix over with rows and columns. For each integer , we define as the ideal of generated by the minors of . For each integer , we also define .
-
(ii)
Let be a finitely generated -module. We choose a finite presentation of with rows and columns, that is, an exact sequence
Here and henceforth, as a convention, we deal with row vectors, so we multiply matrices from the right. Then, for each , we define the -th Fitting ideal of by
It is known that this ideal does not depend on the choice of . When , we also write and call it the initial Fitting ideal.
We will later make use of the following elementary lemma. We omit the proof (cf. Kurihara [8, Lemma 3.3]).
Lemma 2.2.
Let be a finitely generated -module and be an ideal of . If is generated by elements over , then
2.2. Shifts of Fitting ideals
In this subsection, we review the definition of shifts of Fitting ideals introduced by the second author [6].
Although we can deal with a more general situation, for simplicity we consider the following. Let be a Dedekind domain (e.g. , , or ). Let be a finite abelian group and consider the ring .
We define as the category of -modules of finite length. We also define a subcategory of by
where denotes the projective dimension over . Note that any module in satisfies .
Definition 2.3.
Let be an -module in and an integer. We take an exact sequence
in with . Then we define
The well-definedness (i.e. the independence from the choice of the -step resolution) is proved in [6, Theorem 2.6 and Proposition 2.7].
We also introduce a variant for the case where is negative.
Definition 2.4.
Let be an -module in and an integer. We take an exact sequence
in with . Then we define
The well-definedness is proved in [6, Theorem 3.19 and Propositions 2.7 and 3.17].
3. Fitting ideals of ideal class groups
In this section, we prove Theorem 1.1, which describes the Fitting ideal of using shifts of Fitting ideals. We keep the notation in §1.1.
3.1. Brief review of work of Kurihara
We first review necessary ingredients from Kurihara [9], which in turn relies on preceding work, in particular Ritter-Weiss [12] and Greither [3].
For each place of , let and denote the decomposition subgroup and the inertia subgroup of in , respectively. These subgroups depend only on the place of which lies below .
Let us introduce local modules . For any finite group , we define as the augmentation ideal in .
Definition 3.1.
For each finite prime of , we define a -module by
(3.1) |
where denotes the image of in . For each finite prime of , we define the -module by taking the direct sum as
where runs over the finite primes of which lie above . Alternatively, can be regarded as the induced module of from to , as long as we choose a place of above .
We take an auxiliary finite set of places of satisfying the following conditions.
-
•
.
-
•
.
-
•
, where .
-
•
is generated by the decomposition groups of for all .
We define a -module by
By using local and global class field theory, Kurihara constructed an exact sequence of the following form.
Proposition 3.2 (Kurihara [9, §2.2, sequence (5)]).
We have an exact sequence
where is a free -module of rank .
3.2. Definition of
Our key ingredient for the proof of Theorem 1.1 is the following homomorphism .
Definition 3.3.
For a finite prime of , we define a -homomorphism
by (recall the definition of in Definition 3.1). For a finite prime of , we then define a -homomorphism by
(3.2) |
where the last isomorphism depends on a choice of .
In §1.1 we introduced a finite -module with . It is actually motivated by the following.
Lemma 3.4.
For any finite prime of , the map is injective and
Proof.
It is enough to show that is injective and for any finite prime of . Put . We define a homomorphism by . Let us consider the following commutative diagram
where the lower sequence is the trivial one, the commutativity of the left square is easy, and the right vertical arrow is the induced one. By the definition of , we have
Since is a cyclic group generated by , the -module is free of rank with a basis . Moreover, sends this basis to . Therefore, is injective with cokernel isomorphic to . Then by the diagram also satisfies the desired properties. ∎
For any , we consider the homomorphism which is induced by . For any , we have by choosing , so we fix this isomorphism and write for it. Using these , we consider the following commutative diagram
where the upper sequence is that in Proposition 3.2 and the map is defined by the commutativity. By Lemma 3.4 and the snake lemma, we get the following proposition.
Proposition 3.5.
We have an exact sequence
Moreover, the -module is finite with .
By Proposition 3.5, the Fitting ideal is a principal ideal of generated by a non-zero-divisor. Then we can describe the Fitting ideals of and of as follows.
Corollary 3.6.
We have
and
3.3. Fitting ideal of
Recall the definitions of and of in §1.1.
Theorem 3.7.
Assume that eTNC for holds. Then we have
Proof.
For each , we define a basis of as in [9, §2.2, equation (9)] (we do not recall the precise definition here). Then we can see that its dual basis of is given by
where is a lift of . Then, by the definition of , this element satisfies , where by abuse of notation denotes the homomorphism induced by . For , as a basis over , we take the element of which is characterized by .
Let us consider the isomorphism induced by the sequence in Proposition 3.2. Then, under eTNC, Kurihara [9, Theorem 3.6] proved
with respect to a certain basis of as a -module and the basis of . Actually this is an easy reformulation of the result of Kurihara, which concerns the determinant of the linear dual of .
Therefore, the determinant of the composite map of and , with respect to the basis of and the standard basis of , also coincides with . This shows the theorem. ∎
We are now ready to prove Theorem 1.1.
Proof of Theorem 1.1.
4. Computation of shifts of Fitting ideals
4.1. Computation of
Before , we determine , which is actually much easier.
We choose a lift of and put
which is again a non-zero-divisor. Obviously, is then the natural image of to .
Proposition 4.1.
We have
Therefore, we also have
Proof.
We have an exact sequence
Since multiplication by is injective on each of these modules, applying the snake lemma, we obtain an exact sequence
By Definition 2.4, we then have
This proves the former formula of the proposition.
Since we have , the former formula implies . Then the latter formula follows from . ∎
Before proving Theorem 1.4, we show a corollary.
Corollary 4.2.
We have an inclusion
Moreover, if is a cyclic group, the inclusion is an equality.
4.2. Computation of
This subsection is devoted to the proof of Theorem 1.4.
We fix the decomposition (1.3) of . For each , we choose a generator of and put . Note that we then have and . As in §4.1, we put after choosing .
We recall and also put . Then we have and .
We begin with a proposition.
Proposition 4.3.
We have
Proof.
Our next task is to determine for . The result will be Proposition 4.9 below. For that purpose, we construct a concrete free resolution of over , using an idea of Greither-Kurihara [5, §1.2] (one may also refer to [6, §4.3]).
For each , we have a homological complex
over , concentrated at degrees . Let be the degree component of , so if and otherwise. Then the homology groups are for and .
We define a complex over by
which is the tensor product of complexes over (we do not specify the sign convention as it does not matter to us; we define it appropriately so that the descriptions of and below are valid). Explicitly, the degree component of is defined as
Clearly the tensor product is zero unless , and in that case
It is convenient to write for the basis of for each , following [5]. Then, for each , the module is a free module on the set of monomials of of degree .
A basic property of tensor products of complexes implies that for and . Therefore, is a free resolution of over .
It will be necessary to investigate some components of of low degrees. Note that is free of rank one with a basis , is a free module on the set
and is a free module on the set where
Moreover, the differential for are described as follows. We have
for each ,
for each , and
for each .
Let denote the presentation matrix of . For clarity, we define formally as follows.
Definition 4.4.
We define a matrix
with the columns (resp. the rows) indexed by (resp. ), by
Here, we do not specify the orders of the sets and . The ambiguity does not matter for our purpose.
For later use, we also define a matrix
as the submatrix of with the rows in removed. More precisely, we define the matrix with the columns (resp. rows) indexed by (resp. ), by
Therefore, by choosing appropriate orders of rows and columns, we have
Example 4.5.
When , we have
Here, we use the order for the set .
Proposition 4.6.
The matrix , over , is a presentation matrix of the module .
Proof.
By the construction, is a presentation matrix of over . Since is flat over , we obtain the proposition by base change. ∎
Proposition 4.7.
For each , we have
Here, for each , in the second summation runs over subsets of of elements, and for each we define by requiring
The matrix is defined as in Definition 4.4 for and instead of and .
Proof.
By the definition of higher Fitting ideals, is generated by for square submatrices of of size . Such a matrix is in one-to-one correspondence with choices of a subset with and a subset with . We only have to deal with satisfying .
For each , we define and by
(so clearly ) and
Recall that the row in the matrix contains a unique non-zero component in the column. Therefore, the assumption forces and
where is the square submatrix of of size , with rows in and columns in .
Let denote the submatrix of obtained by removing the columns. Then it is clear that ’s (for fixed and ) as above generate . The argument so far implies
By the formula , we may remove the components from the matrix in the right hand side. It is easy to check that the resulting matrix is nothing but (with several zero rows added). This completes the proof. ∎
Proposition 4.8.
For and , we have
Proof.
Since has columns, the case for is clear.
We show the vanishing when and . Let be the polynomial ring over . Then we have a ring homomorphism defined by sending to . We define a matrix over in the same way as in Definition 4.4, with replaced by . Then, by the base change via , we have
Hence the left hand side would vanish if we show that .
For each , we consider the complex
over , which satisfies for and . Similarly as previous, by taking the tensor product of the complexes over , we obtain an exact sequence
over . (Alternatively, this exact sequence is obtained from the Koszul complex for the regular sequence .) This implies that is the Fitting ideal of the augmentation ideal of . Since , the augmentation ideal of is generically of rank one, so we obtain as desired.
Finally we show the case where . Since the components of the matrix are either or one of , the inclusion is clear. In order to show the other inclusion, we use the induction on .
For a while we fix an arbitrary . Then, by permuting the rows and columns, the matrix can be transformed into
(The symbol means omitting that term.) Here, the column is placed in the right-most, and the rows are placed in the lower. We also reversed the signs of some rows for readability as that does not matter at all.
This expression implies
By the induction hypothesis (note that ), we have
Now we vary and then obtain
where the last equality follows from . This completes the proof of the proposition. ∎
Proposition 4.9.
For , we define an ideal of by
Then we have
Proof.
When , since by the choice of , Proposition 4.8 implies
Then we obtain
Using the relation , for each , we have
These formulas imply . ∎
We are ready to prove Theorem 1.4.
Proof of Theorem 1.4.
By Propositions 4.3 and 4.9, we have
Then, noting , we can deduce
in the same way as in the proof of Proposition 4.1. Then it is enough to show
(4.4) |
We claim that
(4.5) |
holds for . We first see
We also have . Since as an ideal, these show the claim (4.5).
Using (4.5), we next show
(4.6) |
More generally we actually show
by induction on , for each . The case is trivial. For , we have
(4.7) | ||||
(4.8) |
Here, the second equality follows from the induction hypothesis and expanding the power . By (4.5), for , we have Therefore, we obtain
(4.9) | ||||
(4.10) |
This completes the proof of (4.6).
The right hand side of (4.6) can be computed as
(4.11) | ||||
(4.12) | ||||
(4.13) |
Here, the first equality follows from the definition of , the second by putting , the third by , and the final by the definition of . Then, combining this with (4.6), we obtain the formula (4.4). This completes the proof of Theorem 1.4. ∎
5. Stickelberger element and Fitting ideal
In this section, we prove Theorem 1.5. As explained after the statement, we need to compare the ideals in the both sides of (1.6) for each before taking the product. That task will be done in §5.1, and after that we complete the proof of Theorem 1.5 in §5.2.
In this section we fix an odd prime number and always work over .
5.1. Comparison of ideals
In this subsection, we again consider the general algebraic situation as in §1.2. Our task in this subsection is to compare the two fractional ideals
of . In Lemma 5.1 (resp. Lemma 5.2) below, we deal with the case where is not (resp. is) a -group. We will make use of the concrete description of in Theorem 1.4. As we always work over instead of , by abuse of notation, in this subsection we simply write , , , and for the extensions of those ideals from to . We have no afraid of confusion due to this.
Let denote the maximal subgroup of whose order is prime to .
Lemma 5.1.
Let be a faithful character of . Suppose that is not a -group. Then we have
as fractional ideals of .
Proof.
We write and . We first note that . This is because is non-trivial on by the assumptions. Then we have by Definition 1.2, so Theorem 1.4 implies
We have to show
When is non-trivial, then as is non-trivial on , so this is obvious. Let us suppose that is trivial. Since , we have
The element of is a unit since , , and . This completes the proof. ∎
Lemma 5.2.
Suppose that is non-trivial and that is a -group. Let be the -rank of , that is, the number of minimal generators of (note that ).
-
(1)
We have
as fractional ideals of .
-
(2)
Let be a character of such that is faithful on and that is non-trivial on . Then we have
as ideals of .
Proof.
We may take a decomposition (1.3) of so that coincides with the -rank of as the lemma, and then is non-trivial for each .
(2) We first show . By the claim (1) above, the inclusion holds. For each , we observe since is either or . Moreover, we have since is non-trivial on and we have if is any non-trivial root of unity. These observations imply for . By the definition of , we then have as claimed.
By Theorem 1.4 and the above claim, we have
We have to show . This is obvious if is non-trivial on . If is trivial on , we have
where the last inclusion follows from . This completes the proof of (2). ∎
5.2. Proof of Theorem 1.5
Now we consider the setup in §1.3. In particular, we fix an odd prime number and an odd character of . Recall the -component of the Stickelberger element defined as (1.4)
Lemma 5.3.
We have if and only if there exists a character of such that and that is non-trivial on for any .
Proof.
By (1.5) and the fact that is a non-zero-divisor, we have if and only if there exists a character of such that and, for every , we have . The last condition is equivalent to that is non-trivial on . This proves the lemma. ∎
We begin the proof of Theorem 1.5.
Proof of Theorem 1.5.
As already remarked in the outline of the proof after Theorem 1.5, we may and do assume that is a faithful character of . This is because we have as the degree of is prime to . Moreover, to simplify the notation, we write and . Recall that, by Theorem 1.1, the condition (i) is equivalent to (1.6). As in §5.1, for each , we consider the fractional ideals of
We first suppose (ii) and aim at showing (i). The case where is trivial, so we assume that, for any , either (a) or (b) holds. Then we obtain for any , by applying Lemma 5.1 (resp. Lemma 5.2(1)) if (a) (resp. (b)) holds. Thus (1.6) holds.
We now prove that (i) implies (ii). Suppose that both (i) and the negation of (ii) hold. Since , we may take a character as in Lemma 5.3. By applying to (1.6), we obtain
On the other hand, by Lemmas 5.1 and 5.2(2), for each , we have . Moreover, the inclusion is proper if and only if both the conditions (a) and (b) in (ii) are false. Therefore, by the hypothesis that (ii) fails, we obtain
Thus we get a contradiction. This completes the proof of Theorem 1.5. ∎
Acknowledgments
Both of the authors are sincerely grateful to Masato Kurihara for his continuous support during the research. They also thank Cornelius Greither for extremely encouraging comments. The second author is supported by JSPS KAKENHI Grant Number 19J00763.
References
- [1] D. Burns, M. Kurihara, and T. Sano. On zeta elements for . Doc. Math., 21:555–626, 2016.
- [2] S. Dasgupta and M. Kakde. On the Brumer-Stark conjecture. preprint, arXiv:2010.00657, 2020.
- [3] C. Greither. Determining Fitting ideals of minus class groups via the equivariant Tamagawa number conjecture. Compos. Math., 143(6):1399–1426, 2007.
- [4] C. Greither and M. Kurihara. Stickelberger elements, Fitting ideals of class groups of CM-fields, and dualisation. Math. Z., 260(4):905–930, 2008.
- [5] C. Greither and M. Kurihara. Tate sequences and Fitting ideals of Iwasawa modules. Algebra i Analiz, 27(6):117–149, 2015.
- [6] T. Kataoka. Fitting invariants in equivariant Iwasawa theory. Development of Iwasawa Theory – the Centennial of K. Iwasawa’s Birth, 2020.
- [7] M. Kurihara. Iwasawa theory and Fitting ideals. J. Reine Angew. Math., 561:39–86, 2003.
- [8] M. Kurihara. On the structure of ideal class groups of CM-fields. Number Extra Vol., pages 539–563. 2003. Kazuya Kato’s fiftieth birthday.
- [9] M. Kurihara. Notes on the dual of the ideal class groups of CM-fields. preprint, arXiv:2006.05803, 2020.
- [10] M. Kurihara and T. Miura. Stickelberger ideals and Fitting ideals of class groups for abelian number fields. Math. Ann., 350(3):549–575, 2011.
- [11] D. G. Northcott. Finite free resolutions. Cambridge University Press, Cambridge-New York-Melbourne, 1976. Cambridge Tracts in Mathematics, No. 71.
- [12] J. Ritter and A. Weiss. A Tate sequence for global units. Compositio Math., 102(2):147–178, 1996.