On the Galois module structure of minus class groups
Abstract.
The main object of this paper is the minus class groups associated to CM-fields as Galois modules. In a previous article of the authors, we introduced a notion of equivalence for modules and determined the equivalence classes of the minus class groups. In this paper, we show a concrete application of this result. We also study how large a proportion of equivalence classes can be realized as classes of minus class groups.
Key words and phrases:
Class groups, integral representations, lattices, syzygies2020 Mathematics Subject Classification:
11R29 (Primary) 11R33 (Secondary)1. Introduction
In some sense this paper is a continuation of a paper of the authors [5]. The setting is given by a CM-field which is an abelian extension of a totally real field , the Galois group being called . Let be the minus part of the group ring, where denotes complex conjugation. For any -module , the minus part is defined as . We consider the minus part of the -modified class group (the technicalities will be explained in §2.4).
The question is, as in previous work: How much can we say about the structure of as a module over , based on equivariant -values and field-theoretic information (ramification and the like) attached to ? The Fitting ideal of the Pontryagin dual had been determined, more and more generally and unconditionally, by the first author [4], Kurihara [10], and Dasgupta–Kakde [3]. Somewhat unexpectedly the Fitting ideal of the non-dualized module was determined later, in the paper [1] by Atsuta and the second author. In particular, as a consequence of lengthy computations, we obtained an inclusion relation
In [5], we introduced a new notion of equivalence for the category of finite -modules, denoted simply by . Moreover, we described the equivalence class of (see §2). A guiding principle in defining is to regard -cohomologically trivial modules as zero. From the view point of Fitting ideals, this implies that ignores the contribution of invertible ideals. Therefore, the analytic factor coming from -functions, which was present in the earlier descriptions of the Fitting ideal, is lost. Nevertheless, the notion is useful enough. For instance, as a first application of the theory, we have reproved the aforementioned inclusion relation.
This paper deals with two problems concerning the structure of from the viewpoint of . The first one is to obtain a more concrete application of the notion of . The second is to obtain some idea how large a proportion of equivalence classes is realized via classes of class groups. It turns out that this proportion tends to be remarkably small. In §1.1 and §1.2 respectively, we will briefly explain these results.
1.1. A concrete application of the equivalence
It is our purpose here to show that our description of up to can actually lead to concrete predictions concerning the structure of . Of course our predictions will fall short of determining the isomorphism class a priori – that would be way too ambitious.
Let be a prime number. We consider the case where is a -extension, where as usual denotes the maximal totally real subfield of . In this case, has a unique intermediate field such that is a quadratic extension. Then must be a CM-field satisfying .
The following is the main result here. Let be the additive -adic valuation normalized by . In Lemma 2.5, we will see that the -modification is unnecessary because of the second assumption.
Theorem 1.1.
Suppose the following:
-
•
is a cyclic -extension of degree for some .
-
•
has no non-trivial -th roots of unity.
-
•
There is a unique prime of that is ramified in and split in .
-
•
is totally ramified in .
-
•
.
Then is in the set
In other words, is nonzero and either divisible by or larger than .
1.2. Realization problem
The second topic in this paper is the question: Which finite -modules can be equivalent to a class group a priori?
To be more precise, let us fix an abstract finite abelian group . For simplicity, we assume that the order of is odd. Let be the category of finite -modules, where we put . In this setting we have the notion of equivalence on . Let be the set of equivalence classes. It is known that can be regarded as a commutative monoid with respect to direct sums.
Let us consider various abelian CM-extensions such that . Since the order of is odd, as in §1.1, there is a unique intermediate CM-field satisfying and . Then we have an identification
induced by the inclusion . Therefore, we may talk about the class of the minus class group in .
We call an element of a realizable class if it is the class of for some extension described above (both and vary). Let be the subset of realizable classes. Now the question is to study the size of .
Our main results involve another subset of , whose elements we call admissible classes. The definition of will be given in §2.6. Here we only mention that is defined in an algebraic way (independent from arithmetic) so that we have . Also, is by definition a submonoid, while it is not clear whether so is a priori.
Now the problem splits naturally into two sub-problems:
-
(a)
Do we have ?
-
(b)
What is the monoid structure of ?
Note that (a) is an arithmetic problem; to prove , we have to construct suitable extensions . On the other hand, (b) is an algebraic problem.
As for (a), we will give the following affirmative answer, which will be proved in §4:
Theorem 1.2.
For any finite abelian group whose order is odd, we have .
In particular, is a submonoid of . Note that we will moreover have a concrete condition on the base field to realize each admissible class. In particular, if is a cyclic group, every admissible class is realized by (see Remark 4.4).
As for (b), we need to introduce a finite set , which depends only on the group structure of . When is a -group, the set is identified with the set of pairs such that
-
•
are subgroups,
-
•
is non-trivial, and
-
•
is cyclic.
The general definition will be given in Definition 5.1. The main result for (b) is the following:
Theorem 1.3.
The following hold:
-
(1)
Suppose is cyclic or is a -group for some prime number . Then is a free monoid of rank .
-
(2)
Otherwise, is not a free monoid.
Let us focus on -groups. As an immediate corollary of Theorems 1.2 and 1.3, we obtain the following:
Corollary 1.4.
Suppose that is a -group for some prime number . Then the subset of is a commutative monoid that is free of rank .
Here is a brief discussion on the relative size of within when is a non-trivial -group for some prime number . Note that the structure of is already discussed in [5].
-
•
If is of order , then both and are free of rank one. Indeed, we may even prove , that is, all finite -modules up to equivalence occur as minus class groups (see Theorem 4.1).
-
•
If is cyclic of order , in [5] we have shown that is not a free monoid and the rank of the abelian group associated to is . This results from the Heller–Reiner classification on the -lattices, given in [2]. On the other hand, it is easy to see that . Therefore, is much smaller than just as is smaller than .
- •
So only a small portion of equivalence classes are realized as minus class groups. It is interesting to observe that the situation around plus components is totally in contrast (see Remark 2.14).
1.3. Organization of this paper
2. Review of the equivalence relation
We begin with a review of the notion of equivalence introduced by the authors [5].
In §§2.1–2.3, we recall the equivalence relation , the re-interpretation of in terms of lattices, and the notion of shift. The theory of shifts is basically known from work of the second author [9], but we add a new aspect, linking it with Heller’s loop operator for lattices.
After fixing the arithmetic setup in §2.4, we obtain the description of the equivalence class of minus class groups in §2.5. In §2.6, we introduce the notion of realizable classes and admissible classes.
2.1. The equivalence relation
Let be a commutative ring that is Gorenstein of Krull dimension one. Typical examples include finite group rings such as and , where
-
•
is a finite abelian group and is considered with respect to a fixed element whose exact order is .
-
•
is a finite abelian group.
Note that in [5] we established the general theory for Gorenstein rings of finite Krull dimension, but in this paper we only need dimension one cases.
Let be the category of finitely generated torsion -modules. Let us write for the subcategory of that consists of modules whose projective dimensions over are at most one. Note that when or as above, consists of all finite -modules and consists of all finite -modules that are -c.t. or -c.t., respectively (“c.t.” is an abbreviation of “cohomologically trivial”).
Definition 2.1.
We define a relation on as follows:
-
(a)
A sandwich is a module in with a three-step filtration by submodules satisfying the following conditions:
-
–
The top quotient and the bottom quotient are both in .
-
–
The middle filtration quotient is in .
The middle filtration quotient is called the filling of the sandwich.
-
–
-
(b)
Two modules and in are equivalent (), if is the filling of some sandwich , is the filling of some other sandwich , and and are isomorphic as -modules. The isomorphism between and is not assumed to respect the filtrations.
For basic properties of this notion and in particular the nontrivial fact that this is an equivalence relation we refer to [5, Remark 2.4, Propositions 2.5–2.6]. For now let us just remark that one easily shows: is equivalent to . Indeed, the definition of arose from the idea of forcing that property.
The set of equivalence classes is equipped with a commutative monoid structure with respect to direct sums. In the following, will always be studied as a commutative monoid. For each module in , we write for the equivalence class of in .
2.2. The equivalence classes via lattices
In [5, §4], we established an interpretation of the equivalence relation via lattices. Let us briefly review the results here.
Let denote the commutative monoid of -lattices up to projective equivalence. Here, an -lattice (which we sometimes simply call a lattice) is by definition a finitely generated torsion-free -module. Two -lattices are projectively equivalent if they become isomorphic after adding finitely generated projective -modules. We write
if and are projectively equivalent lattices. The monoid structure of is defined by direct sums.
Definition 2.2.
We define an injective monoid homomorphism
as follows: To each , choose an epimorphism from a finitely generated projective -module to and define a lattice as its kernel. Though depends on the choice of the epimorphism, it is proved in [5, Theorem 4.2] that this induces a well-defined map that sends the class of to the class of , and that moreover is injective. The image of is also discussed in [5, Lemma 4.3].
2.3. The shift operator
There are shift operators on the monoid , see [5, Definition 3.4] and following discussion. ( in this paper means in [5].) Let us briefly define them.
Definition 2.3.
For any , taking a short exact sequence
where , we define . This is indeed well-defined and an automorphism. We then define the general inductively by for any integer .
On the lattice side, we have the Heller operator , which is an automorphism of . It works as follows: given a lattice , take an exact sequence , again with finitely generated projective over . Then sends the class of to the class of .
We can link our shift operator and the Heller operator . The relation is as simple as possible.
Lemma 2.4.
There is a commutative square as follows:
Proof.
Take and an exact sequence with , so . Take compatible projective resolutions so as to obtain a diagram
in which the modules are finitely generated projective and the upper (lower) vertical arrows are injections (surjections respectively). Then represents and represents . Moreover, is projective over since and is projective. Therefore we also have that represents . These observations imply that both and are represented by , so the lemma follows. ∎
2.4. The arithmetic setup
We review the setting, which will be in force throughout the paper. Assume that is a CM-field which is an abelian extension of a totally real field . Write for . Note that must have even order; it contains a privileged element of order 2 given by complex conjugation.
We also have to discuss -modification, in order to make the results from [1] applicable. Let be any finite set of prime ideals of not containing any ramified prime. Let denote the set of primes of that lie above primes in .
An ideal of is called -principal, if it admits a generator , where the latter group is defined as
The -modified class group is then defined as the group of all fractional ideals of having support disjoint from , modulo the subgroup of -principal ideals. This is a slight enlargement of the usual class group . More precisely, there is a canonical surjection , and its kernel is an epimorphic image of , where denotes the residue field at .
The requirement for in [1] and many other papers is the following: In addition to the conditions already stated, must be such that is -torsion-free. In other words, we must have , where denotes the group of roots of unity in . Trivially this implies that cannot be empty. However, the following lemma implies that in certain cases this modification does not matter. Let and be the -Sylow subgroups of and , respectively.
Lemma 2.5.
If has no non-trivial -th roots of unity, there is a legitimate choice of such that .
Proof.
Let be the order of , which is prime to by the assumption. By Tchebotarev’s density theorem, one can find a prime of such that the order of is divisible by but not by . Set with . Then we have since . On the other hand, the order of is prime to , so in the -part there is no difference between and . ∎
2.5. The equivalence classes of class groups
When one takes [1] and [5] together, one sees that using the notions of equivalence and of shifting one can say a lot on .
We need a little more notation. Let run through the finite primes of ramifying in . For each such , let be the inertia group and the Frobenius at . Define
where denotes the order of .
We will work over the ring and define and accordingly. Then [1, Proposition 3.6] shows the existence of a short exact sequence of -modules
where is a -c.t. module (i.e., in ). Given the good behaviour of shift under equivalence, this gives the following basic result:
Theorem 2.6.
With all the notation introduced so far, we have
where runs over the finite primes of that are ramified in .
Note that the right hand side does not depend on the set , and that the variance under is hidden in the equivariant -value, which is not a part of the statement here. Nevertheless, to keep things technically correct, one has to leave in at least formally.
Lemma 2.7.
If is ramified or inert in , then is in .
Proof.
It is enough to show that is -c.t. for any odd prime . First suppose that is ramified in , that is, . Then acts as on , so the minus part is already trivial. Second suppose that . In this case, is -c.t., so its quotient is also -c.t.
Finally, suppose is inert in and . In this case, the restriction of to is a power of the Frobenius , that is, there is a positive integer such that . Then
is a multiple of in . Since and in the minus component, the displayed element is a -adic unit in . Therefore, is also a unit in , so we obtain . ∎
Definition 2.8.
We define as the set of finite primes of that are ramified in and split in .
Corollary 2.9.
We have
2.6. Realizable classes and admissible classes
Let us fix a finite abelian group whose order is odd. We study the category over . As explained in the introduction, we define the set of realizable classes as follows:
Definition 2.10.
We say that an element of is realizable if it is the class of for some extension such that , where we identify with . The set of realizable classes is denoted by .
Next we define the submonoid of admissible classes. The motivation for the definition will be clear in Corollary 2.12.
Let us employ a useful term from group theory. Given a prime number , a finite group is called -elementary if it is the product of a -group and a cyclic group. A finite group is called elementary if it is -elementary for some prime number .
Definition 2.11.
-
(1)
Associated to , we define as the set of pairs , where
-
–
is a subgroup,
-
–
is non-trivial,
-
–
is an elementary group, and
-
–
is an element.
-
–
-
(2)
For each , we define a finite -module by
-
(3)
We define a submonoid of admissible classes by
which is generated by for various .
Corollary 2.12.
Let be an extension such that . We identify and . Then we have
(1) |
In particular, we have .
Proof.
By local class field theory, for each prime , the inertia group is -elementary for the prime number lying below . Therefore, we have and also . The corollary follows immediately follows from Corollary 2.9. ∎
Remark 2.13.
The following observation will be used to prove Theorem 1.3. For any integer , the monoid is isomorphic to the submonoid generated by for various . This is because the shift automorphisms respect the monoid structure of .
Remark 2.14.
The authors are grateful to Manabu Ozaki, who provided them with the following information.
Let be an odd prime number. Let be any finite -group. Then, for any finite -module , there is a finite abelian extension of totally real fields such that and is isomorphic to as -modules.
This claim can be shown as follows. For the given -module , we consider the semi-direct product of . It is a -group, so we may apply the main theorem of Hajir–Maire–Ramakrishna [6]. As a consequence, there exists a totally real field such that the Galois group of the maximal unramified (not necessarily abelian) -extension over is isomorphic to . Then defining as the intermediate field corresponding to , the requirement is satisfied.
3. Concrete applications
In this section, we prove Theorem 1.1. For this, in §3.1, we compute the lattice associated to the class group explicitly when the Galois group is cyclic. The general case is doable in principle, but it seems to be complicated. Then the proof of Theorem 1.1 will be given in §3.2. In §3.3, we will also observe numerical examples, which suggest that our theoretical result may be sharp. A direct generalization of Theorem 1.1 will be also provided.
3.1. Computation of the lattice
First we compute the lattice .
Theorem 3.1.
Let be a cyclic group of odd order. Let , which simply means that is a subgroup and is an element. Take a lift of . We consider the module over . Then we have
where we define the norm element and the right hand side is the ideal of generated by the two elements inside the brackets.
Proof.
By Lemma 2.4, we have . Let be a generator of . Put . Then we have , so we have to compute
Put . Let be the surjective homomorphism that sends the first basis element to and the second to . Then by definition is projectively equivalent to .
We claim that is generated by and . Indeed, is clear. Suppose that , that is, . Since is a non-zero-divisor, is annihilated by , and so we can write for some . Then is another element in whose second component is zero. The fact that easily gives that . This shows the claim.
It is now easily checked that the first projection gives an isomorphism between and . This completes the proof. ∎
Theorem 3.2.
3.2. Proof of Theorem 1.1
Now we begin the proof of Theorem 1.1. As in §1.1, we consider the case where is a cyclic -extension, where is an odd prime number. Let be the unique quadratic extension of in . Let us write and .
We begin with the following, which implies that is a cyclic -module in the situation of Theorem 1.1.
Proposition 3.3.
Suppose that is a cyclic -extension and . Then the -module is generated by elements and annihilated by .
Proof.
The last statement that annihilates is a direct consequence of the assumption .
For the first statement we use genus theory. By Nakayama’s lemma, it is enough to show that the Galois coinvariant module is generated by elements as a -module. Let be the extension of that is a subfield of the Hilbert class field of and the Artin map gives an isomorphism . Then by Galois theory we find an intermediate field of such that
Since is cyclic, it is known that is the abelianization of , that is, is the maximal abelian extension of in .
Since is Galois, the Galois group acts on , so we have a decomposition
with respect to the action of the complex conjugation. By the construction, we have and .
For any finite prime of , we define a subgroup as , where denotes the (either one or two) primes of lying above and denotes the inertia group of in . Then is stable under the action of , so we also have a decomposition . Note that is identified with the inertia group of in .
Since we assume , the group is generated by for all finite primes of . Therefore, the proposition follows if we show that unless and, moreover, is cyclic when . Since is unramified, for each prime of , the inertia group in is isomorphic to the inertia group of in . This already shows unless ; if does not split in , then , where is the unique prime of lying above . If , there are two primes of lying above . Both and are cyclic since is cyclic, and moreover both are isomorphic to . Combining this with , we conclude that is cyclic, as claimed. ∎
From now on, let us assume the hypotheses of Theorem 1.1. By Proposition 3.3, we can write for a suitable ideal . By Theorem 3.2, taking Lemma 2.5 into account, is projectively equivalent, and even isomorphic, to . Note that this lattice is non-free, so the case follows at once. By Proposition 3.3, must contain .
Therefore, Theorem 1.1 follows from the following algebraic proposition:
Proposition 3.4.
Suppose that is a cyclic group of order with a prime and . Let be an ideal of such that and . Then is in .
Proof.
By the assumption, there is an element such that . Then we have
Claim 3.5.
We have and , where denotes the augmentation.
Proof.
The claim is clear. We have , so also follows. It remains to show by using .
We have
so there are and such that
Since , we have , so this is simplified to
This says , so the claim follows. ∎
Now we have
So
where we put .
We fix a generator of . We also fix a compatible system of -power roots of unity, that is, is a generator of the group of -th roots of unity and we have . For each , let be the character such that . We also write to mean the induced algebra homomorphism . Then gives an injective homomorphism
The cokernel is finite. Then by a standard argument, we obtain
It follows that
(2) | ||||
(3) |
where denotes the additive valuation on , normalized so that we have .
Here is a quick summary: Put . Then we have
We have to investigate . By Claim 3.5, we have , so there exists an element such that
Then we have
and Claim 3.5 implies .
Put .
Claim 3.6.
If one of is less than , then we have .
Proof.
For , since , we have modulo . Therefore, one of the following holds:
-
•
, that is, .
-
•
and , that is, and .
This observation implies the claim (the latter option cannot occur for any by induction). ∎
Now let us complete the proof of the proposition. We put
Case 1. Suppose one of is less than . Then Claim 3.6 implies
so
Then, since
we obtain for . Therefore,
Case 2. Suppose all of are . Then all of are . As in Case 1, we have
so we deduce that all of are . Therefore, we have
This completes the proof of Proposition 3.4. ∎
This also finishes the proof of Theorem 1.1.
3.3. Numerical examples
For numerical examples we are forced to choose , , and . We take the imaginary quadratic field as one of
(Note that is not allowed.) The class numbers of these are respectively, so they are prime to .
Also, we take as the unique subfield of of degree for some prime that is congruent to modulo . The prime must split in , which can be rephrased as a certain congruence condition of (e.g., when , then is congruent to modulo ). We consider the primes in the range .
Our fields will be the compositum of and . The numerical result is that the 3-valuation of the class number of takes values
This does not violate the prediction, of course, and also suggests that our prediction is sharp.
Remark 3.7.
We even did more: even if we remove the condition that consists of a single element (but all are totally ramified in ), a similar reasoning shows that is in the set
where we put . When , this recovers Theorem 1.1. We can of course check this generalized prediction for numerical examples. For instance, for , , and , the possibilities are . This theoretical result can be shown by suitably modifying Proposition 3.4; the details are omitted.
4. The realizability problem
In this section, we prove Theorem 1.2. Before that, in §4.1, we will illustrate the problem in the simplest non-trivial case, i.e., when is the cyclic group whose order is an odd prime number . The proof of Theorem 1.2 will be given in §4.2.
4.1. First case study
Let us show the following, which was stated in the introduction:
Theorem 4.1.
Let be a cyclic group whose order is an odd prime number and we work with the coefficient ring . Then we have
that is, every equivalence class of finite -modules are realized as the class of for some extension with . Moreover, we may restrict the base field to be .
Proof.
Since is a -group, the monoid for can be identified with that for (see Proposition 5.6). Therefore, we may work over instead. By the interpretation via lattices as in §2.2, it is enough to examine
considered in .
In [5, §5.1], we showed that is a free monoid of rank one. This corresponds to the well-known classification of -lattices that every lattice with constant rank is up to a free summand a direct sum of copies of , where is the maximal order in . Therefore, the basis of is the class of .
On the other hand, by Theorem 3.2, consists of the classes of
where varies. Here we used the observation that, for any , we have and is trivial since is a simple group. It is easy to see that is isomorphic to .
As a result, the theorem follows if we show that for any given integer , there is an abelian CM extension with such that . This is a fairly easy exercise. When , take prime numbers that are congruent to modulo , and take as a cyclic extension of of order in in which all are ramified. By taking an imaginary quadratic field in which are split, we find a desired field as . When , we only have to take so that the primes are not split. ∎
4.2. Proof of Theorem 1.2
Now we come back to general . By the description in Corollary 2.12, we obtain Theorem 1.2 from the following:
Theorem 4.2.
Let be an abstract finite abelian group whose order is odd. Suppose that we are given a family . Then there exist a totally real field , a finite abelian CM-extension , and a group isomorphism satisfying the following: We have and we can label so that the inertia group corresponds to and the Frobenius in corresponds to .
To prove this, we make use of the following, which results from global class field theory:
Theorem 4.3 (Grunwald–Wang theorem [11, (9.2.8)]).
Let be a number field, a finite abelian group, and a finite set of primes of . Suppose that for every we are given a finite abelian extension and an embedding . Suppose that we are not in the special case (in the sense of [11, (9.1.5), (9.1.7)]). Then there exist a finite abelian extension and an isomorphism that realizes the designated local extensions for .
Proof of Theorem 4.2.
Step 1. First, we construct a totally real field and distinct primes of (). The required condition is mild: it is enough to choose them so that there is a surjective homomorphism
for each , where denotes the local unit group. This is possible since, by the definition of , for each , there is a prime number such that is -elementary. We may take as a -adic prime.
Step 2. We construct a finite abelian extension for each . Let be a lift of and let be the subgroup generated by and . Let us choose a uniformizer of , which gives an isomorphism . Then, combining the surjective homomorphism in Step 1 with the map that sends to , we obtain a surjective homomorphism . We define a finite abelian extension as the one corresponding to this via local class field theory. Then by construction, we have an isomorphism such that the inertia group corresponds to and the Frobenius corresponds to .
Step 3. Now we apply the Grunwald–Wang theorem to construct a finite abelian extension . We take and the local extension for each is as in Step 2. Because the exponent of is odd (so not divisible by ), we are not in “the special case.” Therefore, by the Grunwald–Wang theorem, we can construct an abelian extension and an isomorphism such that the localizations at are as designated. Note that this is certainly totally real since the order of is odd.
Step 4. We construct a quadratic CM-extension so that the composite field is an extension of with the desired properties.
Let be the set of finite primes of that are ramified in . We shall construct satisfying the following:
-
•
Each is split in for .
-
•
Each does not split in .
We can find such an by again using the Grunwald–Wang theorem. The global Galois group is the cyclic group of order two, so we are not in “the special case.” The local extensions for are as described above. The local extensions for archimedean places are all , so that is a CM extension of .
Here is a sketch of an alternative construction of . Since it should be a Kummer extension, it is enough to find an element of whose square root generates . The element should satisfy suitable congruent conditions at primes in , -adic primes, and archimedean places. Then the existence follows from the approximation theorem.
Now, by the construction of , if we set , we clearly have . The inertia group and the Frobenius at each are as required, because of the construction of in Steps 2–3 ( does not affect them since is split in ). This completes the proof of Theorem 4.2. ∎
Remark 4.4.
In the proof of Theorem 4.2, Step 1 tells us a recipe for the construction of the base field . If is cyclic, then each is also cyclic, so we may take , thanks to the theorem on arithmetic progressions (cf. Theorem 4.1). On the other hand, if is not cyclic, we cannot take a uniform that satisfies Theorem 4.2 for all families .
5. Rephrasing the problem on
In this section, we show Theorem 5.3, which describes the structure of . It will be a key step to prove Theorem 1.3.
5.1. The key theorem
Let be a finite abelian group. In what follows we do not assume that the order of is odd and work over instead of , which simply widens the scope of the argument. Let be the monoid associated to the ring . As in Definition 2.11, we re-define
(so the former one is recovered by the base-change to from ), and then define the submonoid in the same way. We will study the structure of .
Definition 5.1.
We define various sets as follows:
-
(1)
Let be the set of pairs , where
-
–
are subgroups,
-
–
is non-trivial,
-
–
is an elementary group, and
-
–
is cyclic.
-
–
-
(2)
For each prime , we write for the maximal -quotient of . Let be the set of pairs such that are subgroups satisfying
-
–
is non-trivial and
-
–
is cyclic.
In other words, is defined just as , for instead of . Note that unless .
-
–
-
(3)
Let be the set of tuples such that
-
–
is a prime number (necessarily a prime divisor of ),
-
–
is a subgroup such that is cyclic of order prime to , and
-
–
.
-
–
Definition 5.2.
We define a monoid homomorphism
by
for each , where the sum runs over satisfying , , and .
Now we can state the key theorem, whose proof will be given in the rest of this section.
Theorem 5.3.
The monoid is isomorphic to the image of .
The image of will be studied in §6, which results in Theorem 1.3. For now, let us consider the case where is a -group.
Corollary 5.4.
Suppose is a -group for some prime number . Then is a free monoid of rank .
Proof.
By identifying , we have . Moreover, we have by identifying with . The map is then the identity map. As a consequence, we obtain the corollary. ∎
Example 5.5.
Suppose that is cyclic of order . Then the choice of and is
with and . Therefore, we have
5.2. Reduction to consideration over local rings
For each prime , the ring is decomposed as a product of local rings
where runs over a set of representatives of the characters of of order prime to , modulo -conjugacy. Here, recall that denotes the maximal -quotient of . We will also write and for the maximal -quotient of and , respectively. Let be the category of finite -modules.
For each , as we reviewed in §2.2, we have a monoid injective homomorphism
Moreover, [5, Theorem 5.2] implies that is free on the set of indecomposable -lattices that are not projective (i.e., free). Note that this is true since is a henselian local ring, so the theorem of Krull–Remak–Schmidt–Azumaya holds.
Proposition 5.6.
The natural monoid homomorphism
is an isomorphism.
Proof.
For each , since is finite, is identified with the -Sylow subgroup of and we have
Moreover, any -module is a direct sum of its -components . In addition, is -c.t. if and only if so are all its components. These observations imply the proposition. ∎
For a prime number , let us put
Now we are forced to study the relation among (or equivalently among the associated lattices) for various .
5.3. Reduction to two propositions
For , define to be the subgroup generated by and a lift of ; consequently, generates . The proof of the following two propositions will be given later.
Proposition 5.7.
Let be a prime number and a character of whose order is prime to . The following are equivalent:
-
(i)
We have .
-
(ii)
is trivial or is non-trivial on .
For , let us define as the lattice associated to . The reason why we consider (instead of ) will be explained later.
Proposition 5.8.
Let and be two elements of . Let be a prime number and a character of whose order is prime to . Suppose that and . Then the following are equivalent:
-
(i)
.
-
(ii)
and have a common (nonzero) direct summand in .
-
(iii)
We have and .
Let us prove Theorem 5.3, assuming these propositions.
Recall that is defined as the image of the homomorphism
that sends to . As noted in Remark 2.13, we may consider instead.
First, for each , let us consider the image of given by . Thanks to Proposition 5.7 and Proposition 5.8 (i) (ii), the image is a free monoid and its basis is the set
Here, projectively equivalent lattices are counted as the same. Moreover, by Proposition 5.8 (ii) (iii), this set is in one-to-one correspondence with the set by . Consequently, we have a commutative diagram
where the map sends to
Now we vary . By the description of , we obtain the following commutative diagram
The surjective homomorphism is induced by the surjective map that sends to as before. The injective homomorphism is the diagonal one that sends -component to -components with . Finally, the map sends to
Then, identifying with , we may identify the map as . Thus, we obtain Theorem 5.3, assuming Propositions 5.7 and 5.8.
5.4. Tate cohomology groups
In this subsection, we deduce Proposition 5.7 and a part of Proposition 5.8. As observed in [5, Lemma 6.1], the definition of implies that equivalent modules in have isomorphic Tate cohomology groups. So our idea is to compute Tate cohomology groups for various subgroups of (now we consider an arbitrary subgroup in contrast with Definition 5.1).
Lemma 5.9.
For any subgroup , both and are isomorphic to as -modules.
Proof.
First let us show that
For this, we observe that, for any ,
(4) | ||||
(5) | ||||
(6) |
by using Shapiro’s lemma. When , the claim follows from
To show the claim for , we only have to observe that , , and the multiplication by coincides with the multiplication by on .
By the definition of , we have an exact sequence
Then the lemma follows from the resulting long exact sequence. Here, we need to use that is annihilated by . ∎
Proof of Proposition 5.7.
Suppose (ii) is false, i.e., is non-trivial and is trivial on . Then
is nonzero, so (i) is false.
Now suppose (ii) is true. If is trivial, then is -c.t., so is also -c.t. If is non-trivial and is non-trivial on , then . Therefore, (i) is true. ∎
Proof of Proposition 5.8 (i) (ii) and (i) (iii).
(i) (ii) is clear. To show (i) (iii), it is enough to show that the module structure of
allows to recover the groups and . Here, is non-trivial and is trivial on .
For each -subgroup of , the order is determined by the minimum positive integer that annihilates . By varying , we thus determine the subgroup of .
Then, by taking the -Sylow subgroup of as , we know the module . Since is non-trivial, this determines . This is what we wanted. ∎
We will prove (ii) (i) and (iii) (i) in the subsequent subsections.
5.5. The lattice associated to
To do this, we obtain a concrete description of , which was defined as the lattice associated to . It is a key idea here that is much easier than , which we described in §3.1 only when the group is cyclic. We write for the norm element.
Proposition 5.10.
We have
Proof.
We use the computation in [1, §4A], which was used to determine . Let us mention here that the idea here comes from the fact that is easier than , which corresponds to versus .
We have an exact sequence
Let be a lift of . Put , which is of course a lift of . By the snake lemma, we obtain an exact sequence
This implies
Therefore, by the construction of , we see that is the class of the lattice
Let us modify this lattice by multiplying some non-zero-divisors of . First we have
Put . Then , so
This completes the proof. ∎
From now on, when we write , it always means the representative described in this proposition. For each odd prime number and a character of of order prime to , we have
as lattices of , where denotes the image of .
Proof of Proposition 5.8 (iii) (i).
It is enough to show that (iii) implies that and are isomorphic. Since and generate the same subgroup of , the elements and generate the same ideal of . It follows that there is a unit such that
To ease the notation, let us put . Then we have . Indeed,
(7) | ||||
(8) | ||||
(9) | ||||
(10) |
Thus we have proved Proposition 5.8 (iii) (i). ∎
Remark 5.11.
In fact, we have a more natural proof of Proposition 5.8 (i) (iii). Let us sketch it. By Proposition 5.12 below, the lattice is an extension of by . It is possible to directly compute its extension class; we have an isomorphism
and the extension class corresponds to the class of . Therefore, condition (iii) in Proposition 5.8 claims that the extension classes are the same up to a unit, which indicates that the lattices are isomorphic.
5.6. Direct summands of
Let us study the lattice described in Proposition 5.10, as a preparation for the missing equivalence of the proof of Proposition 5.8.
Proposition 5.12.
We have an exact sequence
(11) |
Proof.
Consider the natural exact sequence
Let us show that this induces the claimed exact sequence, by observing the image and the preimage of .
Since and , we see that is generated by over . The natural homomorphism is injective and its image is generated by over . Therefore, the image of is , as claimed.
To determine the preimage, let be any element such that . We want to show . Let us take elements and such that . Then . In particular, this equation implies , so
Then
This implies , so , as claimed. This completes the proof. ∎
Proposition 5.13.
Let be a prime number and a character of of order prime to . Then either is indecomposable over , or the sequence
which is obtained by Proposition 5.12, splits.
Proof.
Note that both and are indecomposable unless zero, since they are cyclic modules over a local ring.
Suppose that there is a decomposition with nonzero and . Since is a cyclic module, by Nakayama’s lemma, we may assume that the map is surjective.
We claim that the map is zero. For this, we may work after base-change from to so that everything is semi-simple. Then and have no common irreducible components as is (generically) free of rank one. Since there is a surjective map from to , we see that and have no common irreducible components. This shows the claim.
Now by the displayed exact sequence, is isomorphic to . Therefore, is isomorphic, so the sequence splits. ∎
Proof of Proposition 5.8 (ii) (i).
Suppose that and have a common direct summand. We want to show that then these two lattices are indeed isomorphic. If one of them is indecomposable, then the claim is clear (notice that the -ranks of and are the same). Suppose that both are decomposable. By Proposition 5.13, we have
and similarly for . The assumption implies that one of the following holds:
Neither the second nor the third isomorphism can hold, because one side contains the trivial character component and the other does not. Therefore, the first or the fourth occurs, which implies and the desired isomorphism of lattices follows. This completes the proof. ∎
6. The structure of
In this section, we prove Theorem 1.3 by using Theorem 5.3. Let be a finite abelian group. The case where is a -group was done in Corollary 5.4. The case where is cyclic will be done in §6.2, and the other cases will be in §6.3.
6.1. Useful observations
To study the image of , the following is useful.
Lemma 6.1.
Let and we suppose is cyclic. Then we have
where denotes the -Sylow subgroup of . Here we have thanks to the assumption that is cyclic.
Proof.
This can be checked directly from the definition of . ∎
Corollary 6.2.
Define a subset by
Then we have .
Proof.
According to this corollary, we only have to study the image of the homomorphism
Let us compare the cardinalities of and .
Proposition 6.3.
We have and the equality holds if and only if is cyclic or is a prime-power.
Proof.
We define
(12) | ||||
(13) |
Then it is clear that . Moreover, it is easy to see that . Therefore, we have . The equality is equivalent to . The equality fails if and only if there is such that is non-cyclic and is non-prime-power. Such a pair exists if and only if is non-cyclic and is non-prime-power. ∎
6.2. The case of cyclic groups
In this subsection, we prove Theorem 1.3(1) for cyclic groups . Thanks to Corollary 6.2, it is enough to show the following:
Proposition 6.5.
When is cyclic, the homomorphism is injective.
Proof.
As in Proposition 6.3, we have
By definition, is also decomposed as a disjoint union
where denotes the set of subgroups such that is cycic of order prime to . Also, the homomorphism respects these decompositions. Therefore, it is enough to check the injectivity for each components. The proposition follows from the next lemma, applied for the prime-to--component of as . ∎
Lemma 6.6.
For a finite abelian group , let
where and run over all subgroups of , be the homomorphism defined by
Then this homomorphism is injective.
Proof.
Indeed, we can recover from the family of values by induction on the size of . ∎
Note that in this lemma, it is important that runs over all subgroups. However, in the definition of , the quotient group must be cyclic. So we need to use the assumption that is cyclic again.
Before closing this subsection, it is worth mentioning the cardinality of and when is a cyclic group.
Lemma 6.7.
Suppose
where are distinct primes and . Then we have
and
Example 6.8.
If , then and .
Proof.
Since is cyclic, the set consists of pairs such that . This corresponds to divisors . The number of such pairs is
which is equal to the claimed formula.
Next we consider . For each , the number of subgroups such that is cyclic of order prime to is equal to
Therefore, we obtain
We also have
By combining these, we obtain the lemma. ∎
6.3. The case of non-cyclic groups
In this subsection we prove Theorem 1.3(2). To show a monoid is non-free, we will count the irreducible elements of the monoid as in [5, §5.2].
Recall that an irreducible element of a commutative monoid is a non-invertible element that cannot be represented as a sum of two non-invertible elements. Then the basis of a free commutative monoid is determined as the set of irreducible elements.
Proposition 6.9.
For , we have is an irreducible element of if and only if . Also, is injective on .
Proof.
The “only if” part follows from the proof of Corollary 6.2. Let us show the “if” part. Let and suppose that
for a family in . We want to show is a singleton and the family coincides with .
For each prime , we have
In the both sides, satisfies is cyclic and . Also, the left side should be understood to be zero unless . This equality can be rephrased as the combination of
-
(a)
For any and , we have either or .
-
(b)
We have
By considering the component, we can divide (b) as
-
(b1)
We have , where denotes the set of prime divisors of . In other words, we have for any and moreover, for each , there exists a unique such that .
-
(b2)
For that , we have
We claim that (b2) is equivalent to (assuming (a)). To show this, we use the following:
Lemma 6.10.
Let be a finite abelian group. Let be two subgroups of . Suppose that for any subgroup such that is cyclic, we have if and only if . Then we have .
Proof.
This is because any subgroup of can be expressed as the intersection of subgroups with is cyclic. ∎
Since we impose an additional condition , we deduce from (b2) only that the prime-to- component of coincides with that of . But by combining this with (a), we obtain , as claimed.
Note that, since any is represented as for some , we obtain for any . As a summary, we have observed:
-
(b1)
We have .
-
(a)’
For each and the , we have .
-
(b2)’
For any , we have .
Now we use the assumption that , that is, either is non-cyclic or is a prime-power. If is a -power, then (b1) implies that is a singleton and then (a)’ and (b2)’ imply , as desired.
Suppose is non-cyclic. Then there is such that the -group is non-cyclic. Then (b2)’ implies that is also non-cyclic for any . Since is cyclic, this implies that is non-zero for any . By (b1), we must have is a singleton. By (b1), (a)’, and (b2)’, we have , as desired.
This completes the proof of Proposition 6.9. ∎
Now we show Theorem 1.3(2). Suppose is a free monoid. As a summary of Corollary 6.2 and Proposition 6.9, we have a surjective homomorphism
which yields a bijection from to the set of irreducible elements of . Therefore, the rank of is equal to . On the other hand, since is a submonoid of , its rank must be . As a consequence, we must have . By Proposition 6.3, we deduce that either is cyclic or is a prime-power. This completes the proof of Theorem 1.3(2).
Acknowledgments
We sincerely thank the anonymous referees for providing valuable comments to improve this paper. The second author is supported by JSPS KAKENHI Grant Number 22K13898.
References
- [1] M. Atsuta and T. Kataoka. Fitting ideals of class groups for CM abelian extensions. Algebra Number Theory, 17(11):1901–1924, 2023.
- [2] C. W. Curtis and I. Reiner. Methods of representation theory. Vol. I. Wiley Classics Library. John Wiley & Sons, Inc., New York, 1990. With applications to finite groups and orders, Reprint of the 1981 original, A Wiley-Interscience Publication.
- [3] S. Dasgupta and M. Kakde. On the Brumer–Stark conjecture. Ann. of Math. (2), 197(1):289–388, 2023.
- [4] C. Greither. Determining Fitting ideals of minus class groups via the equivariant Tamagawa number conjecture. Compos. Math., 143(6):1399–1426, 2007.
- [5] C. Greither and T. Kataoka. Fitting ideals and various notions of equivalence for modules. Manuscripta Math., 173(1-2):259–291, 2024.
- [6] F. Hajir, C. Maire, and R. Ramakrishna. On Ozaki’s theorem realizing prescribed p-groups as p-class tower groups. Algebra Number Theory, 18(4):771–786, 2024.
- [7] A. Heller and I. Reiner. Representations of cyclic groups in rings of integers. I. Ann. of Math. (2), 76:73–92, 1962.
- [8] A. Heller and I. Reiner. Representations of cyclic groups in rings of integers. II. Ann. of Math. (2), 77:318–328, 1963.
- [9] T. Kataoka. Fitting invariants in equivariant Iwasawa theory. In Development of Iwasawa theory—the centennial of K. Iwasawa’s birth, volume 86 of Adv. Stud. Pure Math., pages 413–465. Math. Soc. Japan, Tokyo, 2020.
- [10] M. Kurihara. Notes on the dual of the ideal class groups of CM-fields. J. Théor. Nombres Bordeaux, 33(3, part 2):971–996, 2021.
- [11] J. Neukirch, A. Schmidt, and K. Wingberg. Cohomology of number fields, volume 323 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, second edition, 2008.