This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Fitting ideals of class groups for CM abelian extensions

Mahiro Atsuta and Takenori Kataoka Faculty of Science and Technology, Keio University. 3-14-1 Hiyoshi, Kohoku-ku, Yokohama, Kanagawa 223-8522, Japan [email protected] [email protected]
Abstract.

Let K/kK/k be a finite abelian CM-extension and TT a suitable finite set of finite primes of kk. In this paper, we determine the Fitting ideal of the minus component of the TT-ray class group of KK, except for the 22-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 conjecture
2010 Mathematics Subject Classification:
11R29

1. Introduction

In number theory, the relationship between class groups and special values of LL-functions is of great importance. In this paper we discuss such a phenomenon for a finite abelian CM-extension K/kK/k, that is, a finite abelian extension such that kk is a totally real field and KK is a CM-field. We focus on the minus components of the (ray) class groups of KK, except for the 22-components, and study the Fitting ideals of them.

Let ClK\operatorname{Cl}_{K} denote the ideal class group of KK. Let ()(-)^{-} denote the minus component after inverting the multiplication by 22. When k=k=\mathbb{Q}, Kurihara and Miura [10] succeeded in proving a conjecture of Kurihara [7] on a description of the Fitting ideal of ClK\operatorname{Cl}_{K}^{-} using the Stickelberger elements. However, for a general totally real field kk, the problem to determine the Fitting ideal of ClK\operatorname{Cl}_{K}^{-} is still open.

There seems to be an agreement that the Pontryagin duals (denoted by ()(-)^{\vee}) of the class groups are easier to deal with (see Greither-Kurihara [4]). In [3], Greither determined the Fitting ideal of ClK,\operatorname{Cl}_{K}^{\vee,-}, assuming that the equivariant Tamagawa number conjecture (eTNC for short) holds and that the group of roots of unity in KK is cohomologically trivial. Subsequently, Kurihara [9] generalized the results of Greither on ClK,\operatorname{Cl}_{K}^{\vee,-} to results on ClKT,,\operatorname{Cl}_{K}^{T,\vee,-}, where ClKT\operatorname{Cl}_{K}^{T} denotes the TT-ray class group, for a finite set TT of finite primes of kk. This enables us, by taking suitably large TT, 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 ClKT,,\operatorname{Cl}_{K}^{T,\vee,-} (see (1.2) below for the formula).

In this paper, for a general totally real field kk, we determine the Fitting ideal of ClKT,\operatorname{Cl}_{K}^{T,-} without the Pontryagin dual, assuming eTNC, except for the 22-component. This problem has been considered to be harder than that on ClKT,,\operatorname{Cl}_{K}^{T,\vee,-} 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 ClKT,\operatorname{Cl}_{K}^{T,-} (assuming eTNC). Note that the question for the dualized version ClKT,,\operatorname{Cl}_{K}^{T,\vee,-} 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 ClKT,\operatorname{Cl}_{K}^{T,-} is always contained in that of ClKT,,\operatorname{Cl}_{K}^{T,\vee,-}, 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 K/kK/k be a finite abelian CM-extension and put G=Gal(K/k)G=\operatorname{Gal}(K/k). Let S(k)S_{\infty}(k) be the set of archimedean places of kk. Let Sram(K/k)S_{\operatorname{ram}}(K/k) be the set of places of kk which are ramified in K/kK/k, including S(k)S_{\infty}(k). For each finite prime vSram(K/k)v\in S_{\operatorname{ram}}(K/k), let IvGI_{v}\subset G denote the inertia group of vv in GG and φvG/Iv\varphi_{v}\in G/I_{v} the arithmetic Frobenius of vv. We then define elements gvg_{v} and hvh_{v} by

gv=1φv1+#Iv[G/Iv],hv=1νIv#Ivφv1+νIv[G],g_{v}=1-\varphi_{v}^{-1}+\#I_{v}\in\mathbb{Z}[G/I_{v}],\;\;\;h_{v}=1-\cfrac{\nu_{I_{v}}}{\#I_{v}}\varphi_{v}^{-1}+\nu_{I_{v}}\in\mathbb{Q}[G],

where we put νIv=τIvτ\nu_{I_{v}}=\sum_{\tau\in I_{v}}\tau. 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 [G]\mathbb{Z}[G]-module AvA_{v} by

Av=[G/Iv]/(gv).A_{v}=\mathbb{Z}[G/I_{v}]/(g_{v}).

We write [G]=[1/2][G]/(1+j)\mathbb{Z}[G]^{-}=\mathbb{Z}[1/2][G]/(1+j), where jj is the complex conjugation in GG. For any [G]\mathbb{Z}[G]-module MM, we also define the minus component by M=M[G][G]M^{-}=M\otimes_{\mathbb{Z}[G]}\mathbb{Z}[G]^{-}. Note that we are implicitly inverting the action of 22. For any xMx\in M, we write xx^{-} for the image of xx under the natural map MMM\to M^{-}.

In general, for a set SS of places of kk, we write SKS_{K} for the set of places of KK which lie above places in SS. We take and fix a finite set TT of finite primes of kk satisfying the following.

  • TSram(K/k)=T\cap S_{\operatorname{ram}}(K/k)=\emptyset.

  • KT×={xK×ordw(x1)>0 for all primes wTK}K_{T}^{\times}=\{x\in K^{\times}\mid\operatorname{ord}_{w}(x-1)>0\mbox{ for all primes }w\in T_{K}\} is torsion free. Here, ordw\operatorname{ord}_{w} denotes the normalized additive valuation.

Note that, if we fix an odd prime number pp and are concerned with the pp-components, the last condition can be weakened to that KT×K_{T}^{\times} is pp-torsion-free. We consider the TT-ray ideal class group of KK defined by

ClKT=Cok(KT×ordwwTK),\operatorname{Cl}_{K}^{T}=\operatorname{Cok}\left(K_{T}^{\times}\overset{\oplus\operatorname{ord}_{w}}{\longrightarrow}\bigoplus_{w\notin T_{K}}\mathbb{Z}\right),

where ww runs over the finite primes of KK which are not in TKT_{K}.

For a character ψ\psi of GG, we write L(s,ψ)L(s,\psi) for the primitive LL-function for ψ\psi; this function omits exactly the Euler factors of primes dividing the conductor of ψ\psi. For any finite prime vv of kk, we put N(v)=#𝔽vN(v)=\#\mathbb{F}_{v}, where 𝔽v\mathbb{F}_{v} is the residue field of vv. We then define the TT-modified LL-function by

LT(s,ψ)=(vT(1ψ(φv)N(v)1s))L(s,ψ).L_{T}(s,\psi)=\left(\prod_{v\in T}(1-\psi(\varphi_{v})N(v)^{1-s})\right)L(s,\psi).

We define

(1.1) ωT=ψLT(0,ψ)eψ1[G],\omega_{T}=\sum_{\psi}L_{T}(0,\psi)e_{\psi^{-1}}\in\mathbb{Q}[G],

where ψ\psi runs over the characters of GG and eψ=1#GσGψ(σ)σ1e_{\psi}=\frac{1}{\#G}\sum_{\sigma\in G}\psi(\sigma)\sigma^{-1} is the idempotent of the ψ\psi-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 K/kK/k holds. Then we have

Fitt[G](ClKT,)=(vSram(K/k)S(k)hvFitt[G][1](Av))ωT,\operatorname{Fitt}_{\mathbb{Z}[G]^{-}}\left(\operatorname{Cl}_{K}^{T,-}\right)=\left(\prod_{v\in S_{\operatorname{ram}}(K/k)\setminus S_{\infty}(k)}h_{v}^{-}\operatorname{Fitt}^{[1]}_{\mathbb{Z}[G]^{-}}\left(A_{v}^{-}\right)\right)\omega_{T}^{-},

where Fitt[G][1]\operatorname{Fitt}^{[1]}_{\mathbb{Z}[G]^{-}} 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 hvFitt[G][1](Av)h_{v}^{-}\operatorname{Fitt}^{[1]}_{\mathbb{Z}[G]^{-}}\left(A_{v}^{-}\right), which completes the description of the Fitting ideal of ClKT,\operatorname{Cl}_{K}^{T,-}. 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) Fitt[G](ClKT,,)=(vSram(K/k)S(k)(νIv,1νIv#Ivφv1))ωT.\operatorname{Fitt}_{\mathbb{Z}[G]^{-}}(\operatorname{Cl}_{K}^{T,\vee,-})=\left(\prod_{v\in S_{\operatorname{ram}}(K/k)\setminus S_{\infty}(k)}\left(\nu_{I_{v}},1-\cfrac{\nu_{I_{v}}}{\#I_{v}}\varphi_{v}^{-1}\right)^{-}\right)\omega_{T}^{-}.

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 GG-module MM, we equip the Pontryagin dual MM^{\vee} with the GG-action by (σf)(x)=f(σx)(\sigma f)(x)=f(\sigma x) for σG\sigma\in G, fMf\in M^{\vee}, and xMx\in M. 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 [G]\mathbb{Z}[G]^{-}-modules of the form

0𝔄WSClKT,00\longrightarrow\mathfrak{A}^{-}\longrightarrow W_{S_{\infty}}^{-}\longrightarrow\operatorname{Cl}_{K}^{T,-}\longrightarrow 0

as in Proposition 3.2, where 𝔄\mathfrak{A}^{-} is free of finite rank #S\#S^{\prime}. Here, SS^{\prime} is an auxiliary finite set of places of kk. 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 WSW_{S_{\infty}}^{-} to ([G])#S(\mathbb{Z}[G]^{-})^{\oplus\#S^{\prime}} whose cokernel is isomorphic to the direct sum of AvA_{v}^{-} for vSS(k)v\in S^{\prime}\setminus S_{\infty}(k). Moreover, assuming eTNC, we will compute the determinant of the composite map 𝔄WS([G])#S\mathfrak{A}^{-}\hookrightarrow W_{S_{\infty}}^{-}\hookrightarrow(\mathbb{Z}[G]^{-})^{\oplus\#S^{\prime}}. 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 Fitt[G][1](Av)\operatorname{Fitt}^{[1]}_{\mathbb{Z}[G]}\left(A_{v}\right). 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 vv appropriately). Let GG be a finite abelian group. Let II and DD be subgroups of GG such that IDGI\subset D\subset G and that the quotient D/ID/I is a cyclic group. We choose a generator φ\varphi of D/ID/I and put

g=1φ1+#I[G/I],h=1νI#Iφ1+νI[G],g=1-\varphi^{-1}+\#I\in\mathbb{Z}[G/I],\qquad h=1-\frac{\nu_{I}}{\#I}\varphi^{-1}+\nu_{I}\in\mathbb{Q}[G],

which are non-zero-divisors. We define a finite [G]\mathbb{Z}[G]-module AA by

A=[G/I]/(g).A=\mathbb{Z}[G/I]/(g).

In order to state the result, we introduce some notations. We choose a decomposition

(1.3) I=I1××IsI=I_{1}\times\cdots\times I_{s}

as an abelian group such that IlI_{l} is a cyclic group for each 1ls1\leq l\leq s. Here, we do not assume any minimality on this decomposition, so we allow even the extreme case where IlI_{l} is trivial for some ll.

For each 1ls1\leq l\leq s, we put νl=νIl=σIlσ[G]\nu_{l}=\nu_{I_{l}}=\sum_{\sigma\in I_{l}}\sigma\in\mathbb{Z}[G]. We also put D=Ker([G][G/D])\mathcal{I}_{D}=\operatorname{Ker}(\mathbb{Z}[G]\to\mathbb{Z}[G/D]).

Definition 1.2.

For 0is0\leq i\leq s, we define ZiZ_{i} as the ideal of [G]\mathbb{Z}[G] generated by νl1νlsi\nu_{l_{1}}\cdots\nu_{l_{s-i}} where (l1,,lsi)(l_{1},\dots,l_{s-i}) runs over all tuples of integers satisfying 1l1<<lsis1\leq l_{1}<\cdots<l_{s-i}\leq s, that is,

Zi=(νl1νlsi1l1<<lsis).Z_{i}=(\nu_{l_{1}}\cdots\nu_{l_{s-i}}\mid 1\leq l_{1}<\cdots<l_{s-i}\leq s).

We clearly have Z0=(νI)Z1Zs=(1)Z_{0}=(\nu_{I})\subset Z_{1}\subset\cdots\subset Z_{s}=(1). We then define an ideal 𝒥\mathcal{J} of [G]\mathbb{Z}[G] by

𝒥=i=1sZiDi1.\mathcal{J}=\sum_{i=1}^{s}Z_{i}\mathcal{I}_{D}^{i-1}.

Note that the definition of ZiZ_{i} does depend on the choice of the decomposition (1.3). On the other hand, it can be shown directly that the ideal 𝒥\mathcal{J} 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 s=1s=1, we have 𝒥=(1)\mathcal{J}=(1).

When s=2s=2, we have

𝒥=(ν1,ν2)+D.\mathcal{J}=(\nu_{1},\nu_{2})+\mathcal{I}_{D}.

When s=3s=3, we have

𝒥=(ν1ν2,ν2ν3,ν3ν1)+(ν1,ν2,ν3)D+D2.\mathcal{J}=(\nu_{1}\nu_{2},\nu_{2}\nu_{3},\nu_{3}\nu_{1})+(\nu_{1},\nu_{2},\nu_{3})\mathcal{I}_{D}+\mathcal{I}_{D}^{2}.

In this setting, we can describe Fitt[G][1](A)\operatorname{Fitt}^{[1]}_{\mathbb{Z}[G]}(A) as follows. It is convenient to state the result after multiplying by hh.

Theorem 1.4.

We have

hFitt[G][1](A)=(νI,(1νI#Iφ1)𝒥)h\operatorname{Fitt}^{[1]}_{\mathbb{Z}[G]}(A)=\left(\nu_{I},\left(1-\frac{\nu_{I}}{\#I}\varphi^{-1}\right)\mathcal{J}\right)

as fractional ideals of [G]\mathbb{Z}[G].

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 ClKT,\operatorname{Cl}_{K}^{T,-}.

We return to the setup in §1.1. Let pp be a fixed odd prime number and we shall work over p\mathbb{Z}_{p}. Let GG^{\prime} denote the maximal subgroup of GG of order prime to pp. We put kp=KGk_{p}=K^{G^{\prime}}, which is the maximal pp-extension of kk contained in KK. For each character χ\chi of GG^{\prime}, we regard 𝒪χ=p[Im(χ)]\mathcal{O}_{\chi}=\mathbb{Z}_{p}[\operatorname{Im}(\chi)] as a p[G]\mathbb{Z}_{p}[G^{\prime}]-module via χ\chi, and put p[G]χ=p[G]p[G]𝒪χ\mathbb{Z}_{p}[G]^{\chi}=\mathbb{Z}_{p}[G]\otimes_{\mathbb{Z}_{p}[G^{\prime}]}\mathcal{O}_{\chi}. For a p[G]\mathbb{Z}_{p}[G]-module MM, we put Mχ=Mp[G]p[G]χM^{\chi}=M\otimes_{\mathbb{Z}_{p}[G]}\mathbb{Z}_{p}[G]^{\chi}, which is an p[G]χ\mathbb{Z}_{p}[G]^{\chi}-module. For an element xMx\in M, we write xχx^{\chi} for the image of xx by the natural map MMχM\to M^{\chi}. We note that p[G]\mathbb{Z}_{p}[G] is isomorphic to the direct product of p[G]χ\mathbb{Z}_{p}[G]^{\chi} if χ\chi runs over the equivalence classes of characters of GG^{\prime}.

From now on, we fix an odd character χ\chi of GG^{\prime}. We define Kχ=KKer(χ)K_{\chi}=K^{\operatorname{Ker}(\chi)}. Then KχK_{\chi} is a CM-field, KχkpK_{\chi}\supset k_{p}, and Kχ/kpK_{\chi}/k_{p} is a cyclic extension of order prime to pp.

We put Sχ=Sram(Kχ/k)S_{\chi}=S_{\operatorname{ram}}(K_{\chi}/k) and consider the χ\chi-component of the Stickelberger element defined by

(1.4) θK/k,Tχ=ψ|G=χLSχ,T(0,ψ)eψ1p[G]χ,\theta_{K/k,T}^{\chi}=\sum_{\psi|_{G^{\prime}}=\chi}L_{S_{\chi},T}(0,\psi)e_{\psi^{-1}}\in\mathbb{Z}_{p}[G]^{\chi},

where ψ\psi runs over characters of GG whose restriction to GG^{\prime} coincides with χ\chi and we write

LSχ,T(s,ψ)=(vSχS(k)(1ψ(φv)))(vT(1ψ(φv)N(v)1s))L(s,ψ).L_{S_{\chi},T}(s,\psi)=\left(\prod_{v\in S_{\chi}\setminus S_{\infty}(k)}(1-\psi(\varphi_{v}))\right)\left(\prod_{v\in T}(1-\psi(\varphi_{v})N(v)^{1-s})\right)L(s,\psi).

Note that, comparing (1.1) and (1.4), we have

(1.5) θK/k,Tχ=(vSχS(k)(1νIv#Ivφv1)χ)ωTχ.\theta_{K/k,T}^{\chi}=\left(\prod_{v\in S_{\chi}\setminus S_{\infty}(k)}\left(1-\frac{\nu_{I_{v}}}{\#I_{v}}\varphi_{v}^{-1}\right)^{\chi}\right)\omega_{T}^{\chi}.

Concerning the dualized version, by Dasgupta-Kakde [2, Theorem 1.3],

θK/k,TχFittp[G]χ((ClKTp),χ)\theta_{K/k,T}^{\chi}\in\operatorname{Fitt}_{\mathbb{Z}_{p}[G]^{\chi}}((\operatorname{Cl}_{K}^{T}\otimes\mathbb{Z}_{p})^{\vee,\chi})

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 SχS_{\chi} instead of Sram(K/k)S_{\operatorname{ram}}(K/k) 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 K/kK/k holds. Then, for each odd character χ\chi of GG^{\prime}, the following are equivalent.

  • (i)

    We have θK/k,TχFittp[G]χ((ClKTp)χ)\theta_{K/k,T}^{\chi}\in\operatorname{Fitt}_{\mathbb{Z}_{p}[G]^{\chi}}((\operatorname{Cl}_{K}^{T}\otimes\mathbb{Z}_{p})^{\chi}).

  • (ii)

    We have either θK/k,Tχ=0\theta_{K/k,T}^{\chi}=0 or, for any vSχS(k)v\in S_{\chi}\setminus S_{\infty}(k), one of the following holds.

    • (a)

      vv does not split completely in Kχ/kpK_{\chi}/k_{p}.

    • (b)

      The inertia group IvI_{v} is cyclic.

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 θK/k,Tχ=0\theta_{K/k,T}^{\chi}=0 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 k=k=\mathbb{Q}), since in that case the inertia groups are automatically cyclic, apart from the 22-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 χ\chi is a faithful character of GG^{\prime} (i.e. Kχ=KK_{\chi}=K); actually we can deduce the general case from this case. Since ωTχ\omega_{T}^{\chi} is a non-zero-divisor of p[G]χ\mathbb{Z}_{p}[G]^{\chi}, by Theorem 1.1 and (1.5), we have θK/k,TχFittp[G]χ((ClKTp)χ)\theta_{K/k,T}^{\chi}\in\operatorname{Fitt}_{\mathbb{Z}_{p}[G]^{\chi}}((\operatorname{Cl}_{K}^{T}\otimes\mathbb{Z}_{p})^{\chi}) if and only if

(1.6) v(1νIv#Ivφv1)χv(hvFittp[G][1](Avp))χ\prod_{v}\left(1-\frac{\nu_{I_{v}}}{\#I_{v}}\varphi_{v}^{-1}\right)^{\chi}\subset\prod_{v}\left(h_{v}\operatorname{Fitt}^{[1]}_{\mathbb{Z}_{p}[G]}(A_{v}\otimes\mathbb{Z}_{p})\right)^{\chi}

holds as fractional ideals of p[G]χ\mathbb{Z}_{p}[G]^{\chi}, where on both sides vv runs over the elements of Sram(K/k)S(k)S_{\operatorname{ram}}(K/k)\setminus S_{\infty}(k).

Obviously we may assume that θK/k,Tχ0\theta_{K/k,T}^{\chi}\neq 0. The proof of (ii) \Rightarrow (i) is the easier part. We will show that, under the assumption (ii), the inclusion of (1.6) holds even for every vv before taking the product. On the other hand, the opposite direction (i) \Rightarrow (ii) is the harder part. That is because, roughly speaking, we have to work over the ring p[G]χ\mathbb{Z}_{p}[G]^{\chi}, 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 ψ\psi of GG which satisfies ψ|G=χ\psi|_{G^{\prime}}=\chi and a certain additional condition, whose existence is verified by Lemma 5.3, and we consider the p[G]χ\mathbb{Z}_{p}[G]^{\chi}-algebra 𝒪ψ=p[Im(ψ)]\mathcal{O}_{\psi}=\mathbb{Z}_{p}[\operatorname{Im}(\psi)]. By investigating the ideals in (1.6) after base change from p[G]χ\mathbb{Z}_{p}[G]^{\chi} to 𝒪ψ\mathcal{O}_{\psi}, we will show (i) \Rightarrow (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

Fitt[G](ClKT,)Fitt[G](ClKT,,).\operatorname{Fitt}_{\mathbb{Z}[G]^{-}}\left(\operatorname{Cl}_{K}^{T,-}\right)\subset\operatorname{Fitt}_{\mathbb{Z}[G]^{-}}\left(\operatorname{Cl}_{K}^{T,\vee,-}\right).

Moreover, the inclusion is an equality if IvI_{v} is cyclic for every vSram(K/k)S(k)v\in S_{\operatorname{ram}}(K/k)\setminus S_{\infty}(k).

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 IvI_{v} is cyclic for every vSram(K/k)S(k)v\in S_{\operatorname{ram}}(K/k)\setminus S_{\infty}(k), we can also deduce from Theorem 1.6 that Fitt[G](ClKT,)\operatorname{Fitt}_{\mathbb{Z}[G]^{-}}\left(\operatorname{Cl}_{K}^{T,-}\right) 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 IvI_{v} is not cyclic for some vv.

2. Definition of Fitting ideals and their shifts

In this section, we fix our notations concerning Fitting ideals.

2.1. Fitting ideals

Let RR be a noetherian ring.

Definition 2.1 (cf. Northcott [11]).

We define the Fitting ideals as follows.

  • (i)

    Let AA be a matrix over RR with mm rows and nn columns. For each integer 0in0\leq i\leq n, we define Fitti,R(A)\operatorname{Fitt}_{i,R}(A) as the ideal of RR generated by the (ni)×(ni)(n-i)\times(n-i) minors of AA. For each integer i>ni>n, we also define Fitti,R(A)=(1)\operatorname{Fitt}_{i,R}(A)=(1).

  • (ii)

    Let XX be a finitely generated RR-module. We choose a finite presentation AA of XX with mm rows and nn columns, that is, an exact sequence

    Rm×ARnX0.R^{m}\overset{\times A}{\to}R^{n}\to X\to 0.

    Here and henceforth, as a convention, we deal with row vectors, so we multiply matrices from the right. Then, for each i0i\geq 0, we define the ii-th Fitting ideal of XX by

    Fitti,R(X)=Fitti,R(A).\operatorname{Fitt}_{i,R}(X)=\operatorname{Fitt}_{i,R}(A).

    It is known that this ideal does not depend on the choice of AA. When i=0i=0, we also write FittR(X)=Fitt0,R(X)\operatorname{Fitt}_{R}(X)=\operatorname{Fitt}_{0,R}(X) 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 XX be a finitely generated RR-module and II be an ideal of RR. If XX is generated by nn elements over RR, then

Fitt0,R(X/IX)=i=0nIiFitti,R(X).\operatorname{Fitt}_{0,R}(X/IX)=\sum_{i=0}^{n}I^{i}\operatorname{Fitt}_{i,R}(X).

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 Λ\Lambda be a Dedekind domain (e.g. Λ=\Lambda=\mathbb{Z}, [1/2]\mathbb{Z}[1/2], or p\mathbb{Z}_{p}). Let Δ\Delta be a finite abelian group and consider the ring R=Λ[Δ]R=\Lambda[\Delta].

We define 𝒞\mathcal{C} as the category of RR-modules of finite length. We also define a subcategory 𝒫\mathcal{P} of 𝒞\mathcal{C} by

𝒫={P𝒞pdR(P)1},\mathcal{P}=\{P\in\mathcal{C}\mid\operatorname{pd}_{R}(P)\leq 1\},

where pdR\operatorname{pd}_{R} denotes the projective dimension over RR. Note that any module MM in 𝒞\mathcal{C} satisfies pdΛ(M)1\operatorname{pd}_{\Lambda}(M)\leq 1.

Definition 2.3.

Let XX be an RR-module in 𝒞\mathcal{C} and d0d\geq 0 an integer. We take an exact sequence

0YP1PdX00\to Y\to P_{1}\to\dots\to P_{d}\to X\to 0

in 𝒞\mathcal{C} with P1,,Pd𝒫P_{1},\dots,P_{d}\in\mathcal{P}. Then we define

FittR[d](X)=(i=1dFittR(Pi)(1)i)FittR(Y).\operatorname{Fitt}^{[d]}_{R}(X)=\left(\prod_{i=1}^{d}\operatorname{Fitt}_{R}(P_{i})^{(-1)^{i}}\right)\operatorname{Fitt}_{R}(Y).

The well-definedness (i.e. the independence from the choice of the nn-step resolution) is proved in [6, Theorem 2.6 and Proposition 2.7].

We also introduce a variant for the case where dd is negative.

Definition 2.4.

Let XX be an RR-module in 𝒞\mathcal{C} and d0d\leq 0 an integer. We take an exact sequence

0XP1PdY00\to X\to P_{1}\to\dots\to P_{-d}\to Y\to 0

in 𝒞\mathcal{C} with P1,,Pd𝒫P_{1},\dots,P_{-d}\in\mathcal{P}. Then we define

FittRd(X)=(i=1dFittR(Pi)(1)i)FittR(Y).\operatorname{Fitt}^{\langle d\rangle}_{R}(X)=\left(\prod_{i=1}^{-d}\operatorname{Fitt}_{R}(P_{i})^{(-1)^{i}}\right)\operatorname{Fitt}_{R}(Y).

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 ClKT,\operatorname{Cl}_{K}^{T,-} 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 ww of KK, let DwD_{w} and IwI_{w} denote the decomposition subgroup and the inertia subgroup of ww in GG, respectively. These subgroups depend only on the place of kk which lies below ww.

Let us introduce local modules WvW_{v}. For any finite group HH, we define ΔH\Delta H as the augmentation ideal in [H]\mathbb{Z}[H].

Definition 3.1.

For each finite prime ww of KK, we define a [Dw]\mathbb{Z}[D_{w}]-module WKwW_{K_{w}} by

(3.1) WKw={(x,y)ΔDw[Dw/Iw]x¯=(1φv1)y},W_{K_{w}}=\{(x,y)\in\Delta D_{w}\oplus\mathbb{Z}[D_{w}/I_{w}]\mid\overline{x}=(1-\varphi_{v}^{-1})y\},

where x¯\bar{x} denotes the image of xx in [Dw/Iw]\mathbb{Z}[D_{w}/I_{w}]. For each finite prime vv of kk, we define the [G]\mathbb{Z}[G]-module WvW_{v} by taking the direct sum as

Wv=w|vWKw,W_{v}=\bigoplus_{w|v}W_{K_{w}},

where ww runs over the finite primes of KK which lie above vv. Alternatively, WvW_{v} can be regarded as the induced module of WKwW_{K_{w}} from DwD_{w} to GG, as long as we choose a place ww of KK above vv.

We take an auxiliary finite set SS^{\prime} of places of kk satisfying the following conditions.

  • SSram(K/k)S^{\prime}\supset S_{\operatorname{ram}}(K/k).

  • ST=S^{\prime}\cap T=\emptyset.

  • ClK,ST=0\operatorname{Cl}_{K,S^{\prime}}^{T}=0, where ClK,ST=Cok(KT×ordwwSKTK)\operatorname{Cl}_{K,S^{\prime}}^{T}=\operatorname{Cok}\left(K_{T}^{\times}\overset{\oplus\operatorname{ord}_{w}}{\longrightarrow}\bigoplus_{w\notin S^{\prime}_{K}\cup T_{K}}\mathbb{Z}\right).

  • GG is generated by the decomposition groups DvD_{v} of vv for all vSv\in S^{\prime}.

We define a [G]\mathbb{Z}[G]-module WSW_{S_{\infty}} by

WS=wS(K)ΔDwvSS(k)Wv.W_{S_{\infty}}=\bigoplus_{w\in S_{\infty}(K)}\Delta D_{w}\oplus\bigoplus_{v\in S^{\prime}\setminus S_{\infty}(k)}W_{v}.

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

0𝔄WSClKT,0,0\longrightarrow\mathfrak{A}^{-}\longrightarrow W_{S_{\infty}}^{-}\longrightarrow\operatorname{Cl}_{K}^{T,-}\longrightarrow 0,

where 𝔄\mathfrak{A}^{-} is a free [G]\mathbb{Z}[G]^{-}-module of rank #S\#S^{\prime}.

In [9], the author took the linear dual of this sequence, and the resulting sequence played an important role to study ClKT,,\operatorname{Cl}_{K}^{T,\vee,-}. In this paper, we do not take the linear dual but instead study the sequence itself for the proof of Theorem 1.1.

3.2. Definition of fvf_{v}

Our key ingredient for the proof of Theorem 1.1 is the following homomorphism fvf_{v}.

Definition 3.3.

For a finite prime ww of KK, we define a [Dw]\mathbb{Z}[D_{w}]-homomorphism

fw:WKw[Dw]f_{w}:W_{K_{w}}\longrightarrow\mathbb{Z}[D_{w}]

by fw(x,y)=x+νIw(y)f_{w}(x,y)=x+\nu_{I_{w}}(y) (recall the definition of WKwW_{K_{w}} in Definition 3.1). For a finite prime vv of kk, we then define a [G]\mathbb{Z}[G]-homomorphism fv:Wv[G]f_{v}:W_{v}\longrightarrow\mathbb{Z}[G] by

(3.2) fv:Wv=w|vWKwfww|v[Dw][G],\displaystyle f_{v}:W_{v}=\bigoplus_{w|v}W_{K_{w}}\overset{\oplus f_{w}}{\longrightarrow}\bigoplus_{w|v}\mathbb{Z}[D_{w}]\simeq\mathbb{Z}[G],

where the last isomorphism depends on a choice of ww.

In §1.1 we introduced a finite [G]\mathbb{Z}[G]-module Av=[G/Iv]/(gv)A_{v}=\mathbb{Z}[G/I_{v}]/(g_{v}) with gv=1φv1+#Ivg_{v}=1-\varphi_{v}^{-1}+\#I_{v}. It is actually motivated by the following.

Lemma 3.4.

For any finite prime vv of kk, the map fvf_{v} is injective and

CokfvAv.\operatorname{Cok}f_{v}\simeq A_{v}.
Proof.

It is enough to show that fwf_{w} is injective and Cokfw[Dw/Iw]/(gv)\operatorname{Cok}f_{w}\simeq\mathbb{Z}[D_{w}/I_{w}]/(g_{v}) for any finite prime ww of KK. Put Jw=Ker([Dw][Dw/Iw])J_{w}=\operatorname{Ker}(\mathbb{Z}[D_{w}]\longrightarrow\mathbb{Z}[D_{w}/I_{w}]). We define a homomorphism αw:JwWKw\alpha_{w}:J_{w}\to W_{K_{w}} by αw(x)=(x,0)\alpha_{w}(x)=(x,0). Let us consider the following commutative diagram

0JwαwWKwCokαw0fwfw0Jw[Dw][Dw/Iw]0,\setcounter{MaxMatrixCols}{11}\begin{CD}0@>{}>{}>J_{w}@>{\alpha_{w}}>{}>W_{K_{w}}@>{}>{}>\operatorname{Cok}\;\alpha_{w}@>{}>{}>0\\ \Big{\|}@V{}V{f_{w}}V@V{}V{f^{\prime}_{w}}V\\ 0@>{}>{}>J_{w}@>{}>{}>\mathbb{Z}[D_{w}]@>{}>{}>\mathbb{Z}[D_{w}/I_{w}]@>{}>{}>0,\\ \end{CD}

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 WKwW_{K_{w}}, we have

Cokαw={(x¯,y)Δ(Dw/Iw)×[Dw/Iw]x¯=(1φv1)y}.\displaystyle\operatorname{Cok}\;\alpha_{w}=\{(\bar{x},y)\in\Delta(D_{w}/I_{w})\times\mathbb{Z}[D_{w}/I_{w}]\mid\overline{x}=(1-\varphi_{v}^{-1})y\}.

Since Dw/IwD_{w}/I_{w} is a cyclic group generated by φv1\varphi_{v}^{-1}, the [Dw/Iw]\mathbb{Z}[D_{w}/I_{w}]-module Cokαw\operatorname{Cok}\;\alpha_{w} is free of rank 11 with a basis (1φv1,1)(1-\varphi_{v}^{-1},1). Moreover, fwf_{w}^{\prime} sends this basis to gv=1φv1+#Ivg_{v}=1-\varphi_{v}^{-1}+\#I_{v}. Therefore, fwf_{w}^{\prime} is injective with cokernel isomorphic to [Dw/Iw]/(gv)\mathbb{Z}[D_{w}/I_{w}]/(g_{v}). Then by the diagram fwf_{w} also satisfies the desired properties. ∎

For any vSS(k)v\in S^{\prime}\setminus S_{\infty}(k), we consider the homomorphism fv:Wv[G]f_{v}^{-}:W_{v}^{-}\longrightarrow\mathbb{Z}[G]^{-} which is induced by fvf_{v}. For any vS(k)v\in S_{\infty}(k), we have (wvΔDw)[G](\oplus_{w\mid v}\Delta D_{w})^{-}\simeq\mathbb{Z}[G]^{-} by choosing ww, so we fix this isomorphism and write fvf_{v}^{-} for it. Using these fvf_{v}^{-}, we consider the following commutative diagram

0𝔄WSClKT,0vSfv0𝔄ψvS[G]Cokψ0,\setcounter{MaxMatrixCols}{11}\begin{CD}0@>{}>{}>\mathfrak{A}^{-}@>{}>{}>W_{S_{\infty}}^{-}@>{}>{}>\operatorname{Cl}_{K}^{T,-}@>{}>{}>0\\ \Big{\|}@V{}V{\oplus_{v\in S^{\prime}}f_{v}^{-}}V@V{}V{}V\\ 0@>{}>{}>\mathfrak{A}^{-}@>{\psi}>{}>\bigoplus_{v\in S^{\prime}}\mathbb{Z}[G]^{-}@>{}>{}>\operatorname{Cok}\psi @>{}>{}>0,\\ \end{CD}

where the upper sequence is that in Proposition 3.2 and the map ψ\psi 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

0ClKT,CokψvSS(k)Av0.0\longrightarrow\operatorname{Cl}_{K}^{T,-}\longrightarrow\operatorname{Cok}\psi\longrightarrow\bigoplus_{v\in S^{\prime}\setminus S_{\infty}(k)}A_{v}^{-}\longrightarrow 0.

Moreover, the [G]\mathbb{Z}[G]^{-}-module Cokψ\operatorname{Cok}\psi is finite with pd[G](Cokψ)1\operatorname{pd}_{\mathbb{Z}[G]^{-}}(\operatorname{Cok}\psi)\leq 1.

By Proposition 3.5, the Fitting ideal Fitt[G](Cokψ)\operatorname{Fitt}_{\mathbb{Z}[G]^{-}}(\operatorname{Cok}\psi) is a principal ideal of [G]\mathbb{Z}[G]^{-} generated by a non-zero-divisor. Then we can describe the Fitting ideals of ClKT,\operatorname{Cl}_{K}^{T,-} and of ClKT,,\operatorname{Cl}_{K}^{T,\vee,-} as follows.

Corollary 3.6.

We have

Fitt[G](ClKT,)=Fitt[G](Cokψ)vSS(k)Fitt[G][1](Av)\operatorname{Fitt}_{\mathbb{Z}[G]^{-}}(\operatorname{Cl}_{K}^{T,-})=\operatorname{Fitt}_{\mathbb{Z}[G]^{-}}(\operatorname{Cok}\psi)\prod_{v\in S^{\prime}\setminus S_{\infty}(k)}\operatorname{Fitt}^{[1]}_{\mathbb{Z}[G]^{-}}\left(A_{v}^{-}\right)

and

Fitt[G](ClKT,,)=Fitt[G](Cokψ)vSS(k)Fitt[G]1(Av).\operatorname{Fitt}_{\mathbb{Z}[G]^{-}}(\operatorname{Cl}_{K}^{T,\vee,-})=\operatorname{Fitt}_{\mathbb{Z}[G]^{-}}(\operatorname{Cok}\psi)\prod_{v\in S^{\prime}\setminus S_{\infty}(k)}\operatorname{Fitt}^{\langle-1\rangle}_{\mathbb{Z}[G]^{-}}\left(A_{v}^{-}\right).
Proof.

The first formula follows directly from Proposition 3.5 and Definition 2.3. For the second formula, by [6, Proposition 4.7], we have

Fitt[G](ClKT,,)=Fitt[G]2(ClKT,).\operatorname{Fitt}_{\mathbb{Z}[G]^{-}}(\operatorname{Cl}_{K}^{T,\vee,-})=\operatorname{Fitt}^{\langle-2\rangle}_{\mathbb{Z}[G]^{-}}(\operatorname{Cl}_{K}^{T,-}).

By Proposition 3.5 and Definition 2.4, we also have

(3.3) Fitt[G]2(ClKT,)\displaystyle\operatorname{Fitt}^{\langle-2\rangle}_{\mathbb{Z}[G]^{-}}(\operatorname{Cl}_{K}^{T,-}) =Fitt[G](Cokψ)vSS(k)Fitt[G]1(Av).\displaystyle=\operatorname{Fitt}_{\mathbb{Z}[G]^{-}}(\operatorname{Cok}\psi)\prod_{v\in S^{\prime}\setminus S_{\infty}(k)}\operatorname{Fitt}^{\langle-1\rangle}_{\mathbb{Z}[G]^{-}}\left(A_{v}^{-}\right).

This completes the proof. ∎

3.3. Fitting ideal of Cokψ\operatorname{Cok}\psi

Recall the definitions of ωT\omega_{T} and of hvh_{v} in §1.1.

Theorem 3.7.

Assume that eTNC for K/kK/k holds. Then we have

Fitt[G](Cokψ)=((vSS(k)hv)ωT).\operatorname{Fitt}_{\mathbb{Z}[G]^{-}}(\operatorname{Cok}\psi)=\left(\left(\prod_{v\in S^{\prime}\setminus S_{\infty}(k)}h_{v}^{-}\right)\omega_{T}^{-}\right).
Proof.

For each vSS(k)v\in S^{\prime}\setminus S_{\infty}(k), we define a basis eve_{v} of Hom[G](Wv,[G])\operatorname{Hom}_{\mathbb{Q}[G]}(W_{v}\otimes\mathbb{Q},\mathbb{Q}[G]) as in [9, §2.2, equation (9)] (we do not recall the precise definition here). Then we can see that its dual basis eve_{v}^{\prime} of WvW_{v}\otimes\mathbb{Q} is given by

ev=11φv~1+NIv(1φv~1,1),e_{v}^{\prime}=\frac{1}{1-\widetilde{\varphi_{v}}^{-1}+N_{I_{v}}}(1-\widetilde{\varphi_{v}}^{-1},1),

where φv~\widetilde{\varphi_{v}} is a lift of φv\varphi_{v}. Then, by the definition of fvf_{v}, this element satisfies fv(ev)=1f_{v}(e_{v}^{\prime})=1, where by abuse of notation fvf_{v} denotes the homomorphism Wv[G]W_{v}\otimes\mathbb{Q}\to\mathbb{Q}[G] induced by fv:Wv[G]f_{v}:W_{v}\to\mathbb{Z}[G]. For vS(k)v\in S_{\infty}(k), as a basis over [G]\mathbb{Z}[G]^{-}, we take the element ev,e_{v}^{\prime,-} of (wvΔDw)\left(\bigoplus_{w\mid v}\Delta D_{w}\right)^{-} which is characterized by fv(ev,)=1f_{v}^{-}(e_{v}^{\prime,-})=1.

Let us consider the isomorphism Ψ:𝔄WS\Psi:\mathfrak{A}^{-}\otimes\mathbb{Q}\to W_{S_{\infty}}^{-}\otimes\mathbb{Q} induced by the sequence in Proposition 3.2. Then, under eTNC, Kurihara [9, Theorem 3.6] proved

det(Ψ)=(vSS(k)hv)ωT\det(\Psi)=\left(\prod_{v\in S^{\prime}\setminus S_{\infty}(k)}h_{v}^{-}\right)\omega_{T}^{-}

with respect to a certain basis of 𝔄\mathfrak{A}^{-} as a [G]\mathbb{Z}[G]^{-}-module and the basis (ev,)vS(e_{v}^{\prime,-})_{v\in S^{\prime}} of WSW_{S_{\infty}}^{-}. Actually this is an easy reformulation of the result of Kurihara, which concerns the determinant of the linear dual of Ψ\Psi.

Therefore, the determinant of the composite map ψ\psi of Ψ\Psi and vSfv\bigoplus_{v\in S^{\prime}}f_{v}^{-}, with respect to the basis of 𝔄\mathfrak{A}^{-} and the standard basis of ([G])#S(\mathbb{Z}[G]^{-})^{\oplus\#S^{\prime}}, also coincides with (vSS(k)hv)ωT\left(\prod_{v\in S^{\prime}\setminus S_{\infty}(k)}h_{v}^{-}\right)\omega_{T}^{-}. This shows the theorem. ∎

We are now ready to prove Theorem 1.1.

Proof of Theorem 1.1.

By Corollary 3.6 and Theorem 3.7, we have

Fitt[G](ClKT,)=(vSS(k)hvFitt[G][1](Av))ωT.\operatorname{Fitt}_{\mathbb{Z}[G]^{-}}(\operatorname{Cl}_{K}^{T,-})=\left(\prod_{v\in S^{\prime}\setminus S_{\infty}(k)}h_{v}^{-}\operatorname{Fitt}^{[1]}_{\mathbb{Z}[G]^{-}}\left(A_{v}^{-}\right)\right)\omega_{T}^{-}.

For vSSram(K/k)v\in S^{\prime}\setminus S_{\operatorname{ram}}(K/k), we have Av=[G]/(hv)A_{v}=\mathbb{Z}[G]/(h_{v}), so

Fitt[G][1](Av)=(hv)1.\operatorname{Fitt}^{[1]}_{\mathbb{Z}[G]^{-}}\big{(}A_{v}^{-}\big{)}=(h_{v}^{-})^{-1}.

Then Theorem 1.1 follows. ∎

Remark 3.8.

Similarly, under the validity of eTNC, Corollary 3.6 and Theorem 3.7 also imply a formula

Fitt[G](ClKT,,)=(vSram(K/k)S(k)hvFitt[G]1(Av))ωT.\operatorname{Fitt}_{\mathbb{Z}[G]^{-}}(\operatorname{Cl}_{K}^{T,\vee,-})=\left(\prod_{v\in S_{\operatorname{ram}}(K/k)\setminus S_{\infty}(k)}h_{v}^{-}\operatorname{Fitt}^{\langle-1\rangle}_{\mathbb{Z}[G]^{-}}\left(A_{v}^{-}\right)\right)\omega_{T}^{-}.

Combining this with Proposition 4.1 below, we can recover the formula (1.2). This argument may be regarded as a reinterpretation of the work [9] by using the shifts of Fitting ideals.

4. Computation of shifts of Fitting ideals

In this section, we prove Theorem 1.4 on the description of Fitt[G][1](A)\operatorname{Fitt}^{[1]}_{\mathbb{Z}[G]}(A). We keep the notations as in §1.2.

4.1. Computation of Fitt[G]1(A)\operatorname{Fitt}^{\langle-1\rangle}_{\mathbb{Z}[G]}(A)

Before Fitt[G][1](A)\operatorname{Fitt}^{[1]}_{\mathbb{Z}[G]}(A), we determine Fitt[G]1(A)\operatorname{Fitt}^{\langle-1\rangle}_{\mathbb{Z}[G]}(A), which is actually much easier.

We choose a lift φ~D\widetilde{\varphi}\in D of φ\varphi and put

g~=1φ~1+#I[G],\widetilde{g}=1-\widetilde{\varphi}^{-1}+\#I\in\mathbb{Z}[G],

which is again a non-zero-divisor. Obviously, gg is then the natural image of g~\widetilde{g} to [G/I]\mathbb{Z}[G/I].

Proposition 4.1.

We have

Fitt[G]1(A)=(1,νIg1).\operatorname{Fitt}^{\langle-1\rangle}_{\mathbb{Z}[G]}(A)=\left(1,\nu_{I}g^{-1}\right).

Therefore, we also have

hFitt[G]1(A)=(νI,1νI#Iφ1).h\operatorname{Fitt}^{\langle-1\rangle}_{\mathbb{Z}[G]}(A)=\left(\nu_{I},1-\frac{\nu_{I}}{\#I}\varphi^{-1}\right).
Proof.

We have an exact sequence

0[G/I]νI[G][G]/(νI)0.0\to\mathbb{Z}[G/I]\overset{\nu_{I}}{\to}\mathbb{Z}[G]\to\mathbb{Z}[G]/(\nu_{I})\to 0.

Since multiplication by g~\widetilde{g} is injective on each of these modules, applying the snake lemma, we obtain an exact sequence

0A[G]/(g~)[G]/(g~,νI)0.0\to A\to\mathbb{Z}[G]/(\widetilde{g})\to\mathbb{Z}[G]/(\widetilde{g},\nu_{I})\to 0.

By Definition 2.4, we then have

Fitt[G]1(A)=(g~)1(g~,νI)=(1,νIg1).\operatorname{Fitt}^{\langle-1\rangle}_{\mathbb{Z}[G]}(A)=(\widetilde{g})^{-1}(\widetilde{g},\nu_{I})=\left(1,\nu_{I}g^{-1}\right).

This proves the former formula of the proposition.

Since we have νIg=νIh\nu_{I}g=\nu_{I}h, the former formula implies hFitt[G]1(A)=(νI,h)h\operatorname{Fitt}^{\langle-1\rangle}_{\mathbb{Z}[G]}(A)=\left(\nu_{I},h\right). Then the latter formula follows from h1νI#Iφ1(mod(νI))h\equiv 1-\frac{\nu_{I}}{\#I}\varphi^{-1}(\bmod(\nu_{I})). ∎

Before proving Theorem 1.4, we show a corollary.

Corollary 4.2.

We have an inclusion

Fitt[G][1](A)Fitt[G]1(A).\operatorname{Fitt}^{[1]}_{\mathbb{Z}[G]}(A)\subset\operatorname{Fitt}^{\langle-1\rangle}_{\mathbb{Z}[G]}(A).

Moreover, if II is a cyclic group, the inclusion is an equality.

Proof.

By Definition 1.2, the ideal 𝒥\mathcal{J} is contained in [G]\mathbb{Z}[G] and we have 𝒥=[G]\mathcal{J}=\mathbb{Z}[G] if II is cyclic. Hence this corollary immediately follows from Theorem 1.4 and Proposition 4.1. ∎

4.2. Computation of Fitt[G][1](A)\operatorname{Fitt}^{[1]}_{\mathbb{Z}[G]}(A)

This subsection is devoted to the proof of Theorem 1.4.

We fix the decomposition (1.3) of II. For each 1ls1\leq l\leq s, we choose a generator σl\sigma_{l} of IlI_{l} and put τl=σl1[G]\tau_{l}=\sigma_{l}-1\in\mathbb{Z}[G]. Note that we then have νl=1+σl+σl2++σl#Il1\nu_{l}=1+\sigma_{l}+\sigma_{l}^{2}+\dots+\sigma_{l}^{\#I_{l}-1} and τlνl=0\tau_{l}\nu_{l}=0. As in §4.1, we put g~=1φ~1+#I\widetilde{g}=1-\widetilde{\varphi}^{-1}+\#I after choosing φ~\widetilde{\varphi}.

We recall D=Ker([G][G/D])\mathcal{I}_{D}=\operatorname{Ker}(\mathbb{Z}[G]\to\mathbb{Z}[G/D]) and also put I=Ker([G][G/I])\mathcal{I}_{I}=\operatorname{Ker}(\mathbb{Z}[G]\to\mathbb{Z}[G/I]). Then we have I=(τ1,,τs)\mathcal{I}_{I}=(\tau_{1},\dots,\tau_{s}) and D=(I,1φ~1)\mathcal{I}_{D}=(\mathcal{I}_{I},1-\widetilde{\varphi}^{-1}).

We begin with a proposition.

Proposition 4.3.

We have

Fitt[G][1](A)=i=0sg~i1Fitti,[G](I).\operatorname{Fitt}^{[1]}_{\mathbb{Z}[G]}(A)=\sum_{i=0}^{s}\widetilde{g}^{i-1}\operatorname{Fitt}_{i,\mathbb{Z}[G]}(\mathcal{I}_{I}).
Proof.

We have the tautological exact sequence

0I[G][G/I]0.0\to\mathcal{I}_{I}\to\mathbb{Z}[G]\to\mathbb{Z}[G/I]\to 0.

Since multiplication by g~\widetilde{g} is injective on each of these modules, by applying snake lemma, we obtain an exact sequence

0I/g~I[G]/(g~)A0.0\to\mathcal{I}_{I}/\widetilde{g}\mathcal{I}_{I}\to\mathbb{Z}[G]/(\widetilde{g})\to A\to 0.

Then Definition 2.3 implies

(4.1) Fitt[G][1](A)\displaystyle\operatorname{Fitt}^{[1]}_{\mathbb{Z}[G]}(A) =g~1Fitt[G](I/g~I).\displaystyle=\widetilde{g}^{-1}\operatorname{Fitt}_{\mathbb{Z}[G]}(\mathcal{I}_{I}/\widetilde{g}\mathcal{I}_{I}).

Since I\mathcal{I}_{I} is generated by the ss elements τ1,,τs\tau_{1},\dots,\tau_{s}, we have

Fitt[G](I/g~I)=i=0sg~iFitti,[G](I)\operatorname{Fitt}_{\mathbb{Z}[G]}(\mathcal{I}_{I}/\widetilde{g}\mathcal{I}_{I})=\sum_{i=0}^{s}\widetilde{g}^{i}\operatorname{Fitt}_{i,\mathbb{Z}[G]}(\mathcal{I}_{I})

by Lemma 2.2. Thus we obtain the proposition. ∎

Our next task is to determine Fitti,[G](I)\operatorname{Fitt}_{i,\mathbb{Z}[G]}(\mathcal{I}_{I}) for 0is0\leq i\leq s. The result will be Proposition 4.9 below. For that purpose, we construct a concrete free resolution of \mathbb{Z} over [I]\mathbb{Z}[I], using an idea of Greither-Kurihara [5, §1.2] (one may also refer to [6, §4.3]).

For each 1ls1\leq l\leq s, we have a homological complex

Cl:τl[Il]νl[Il]τl[Il]0C^{l}:\cdots\overset{\tau_{l}}{\to}\mathbb{Z}[I_{l}]\overset{\nu_{l}}{\to}\mathbb{Z}[I_{l}]\overset{\tau_{l}}{\to}\mathbb{Z}[I_{l}]\to 0

over [Il]\mathbb{Z}[I_{l}], concentrated at degrees 0\geq 0. Let CnlC^{l}_{n} be the degree nn component of ClC^{l}, so Cnl=[Il]C^{l}_{n}=\mathbb{Z}[I_{l}] if n0n\geq 0 and Cnl=0C^{l}_{n}=0 otherwise. Then the homology groups are Hn(Cl)=0H_{n}(C^{l})=0 for n0n\neq 0 and H0(Cl)H_{0}(C^{l})\simeq\mathbb{Z}.

We define a complex CC over [I]\mathbb{Z}[I] by

C=l=1sCl,C=\bigotimes_{l=1}^{s}C^{l},

which is the tensor product of complexes over \mathbb{Z} (we do not specify the sign convention as it does not matter to us; we define it appropriately so that the descriptions of d1d_{1} and d2d_{2} below are valid). Explicitly, the degree nn component CnC_{n} of CC is defined as

Cn=n1++ns=nCn11Cnss.C_{n}=\bigoplus_{n_{1}+\dots+n_{s}=n}C^{1}_{n_{1}}\otimes\dots\otimes C^{s}_{n_{s}}.

Clearly the tensor product is zero unless n1,,ns0n_{1},\dots,n_{s}\geq 0, and in that case

Cn11Cnss=[I1][Is][I].C^{1}_{n_{1}}\otimes\dots\otimes C^{s}_{n_{s}}=\mathbb{Z}[I_{1}]\otimes\dots\otimes\mathbb{Z}[I_{s}]\simeq\mathbb{Z}[I].

It is convenient to write x1n1xsnsx_{1}^{n_{1}}\cdots x_{s}^{n_{s}} for the basis of Cn11CnssC^{1}_{n_{1}}\otimes\dots\otimes C^{s}_{n_{s}} for each n1,,ns0n_{1},\dots,n_{s}\geq 0, following [5]. Then, for each n0n\geq 0, the module CnC_{n} is a free module on the set of monomials of x1,,xsx_{1},\dots,x_{s} of degree nn.

A basic property of tensor products of complexes implies that Hn(C)=0H_{n}(C)=0 for n0n\neq 0 and H0(C)H_{0}(C)\simeq\mathbb{Z}. Therefore, CC is a free resolution of \mathbb{Z} over [I]\mathbb{Z}[I].

It will be necessary to investigate some components of CC of low degrees. Note that C0C_{0} is free of rank one with a basis 1(=x10xs0)1(=x_{1}^{0}\cdots x_{s}^{0}), C1C_{1} is a free module on the set

S1={x1,,xs},S_{1}=\{x_{1},\dots,x_{s}\},

and C2C_{2} is a free module on the set S2S2S_{2}\cup S_{2}^{\prime} where

S2={x12,,xs2},S2={xlxl1l<ls}.S_{2}=\{x_{1}^{2},\dots,x_{s}^{2}\},\qquad S_{2}^{\prime}=\{x_{l}x_{l^{\prime}}\mid 1\leq l<l^{\prime}\leq s\}.

Moreover, the differential dn:CnCn1d_{n}:C_{n}\to C_{n-1} for n=1,2n=1,2 are described as follows. We have

d1(xl)=τl1d_{1}(x_{l})=\tau_{l}\cdot 1

for each 1ls1\leq l\leq s,

d2(xl2)=νlxld_{2}(x_{l}^{2})=\nu_{l}x_{l}

for each 1ls1\leq l\leq s, and

d2(xlxl)=τlxlτlxld_{2}(x_{l}x_{l^{\prime}})=\tau_{l}x_{l^{\prime}}-\tau_{l^{\prime}}x_{l}

for each 1l<ls1\leq l<l^{\prime}\leq s.

Let MM denote the presentation matrix of d2d_{2}. For clarity, we define MM formally as follows.

Definition 4.4.

We define a matrix

M=Ms(ν1,,νs,τ1,,τs)M=M_{s}(\nu_{1},\dots,\nu_{s},\tau_{1},\dots,\tau_{s})

with the columns (resp. the rows) indexed by S1S_{1} (resp. S2S2S_{2}\cup S_{2}^{\prime}), by

{the (xl2,xl) component is νlfor 1ls,the (xlxl,xl) component is τlfor 1l<ls,the (xlxl,xl) component is τlfor 1l<ls,and the other components are zero.\begin{cases}\text{the $(x_{l}^{2},x_{l})$ component is $\nu_{l}$}&\text{for $1\leq l\leq s$,}\\ \text{the $(x_{l}x_{l^{\prime}},x_{l})$ component is $-\tau_{l^{\prime}}$}&\text{for $1\leq l<l^{\prime}\leq s$,}\\ \text{the $(x_{l}x_{l^{\prime}},x_{l^{\prime}})$ component is $\tau_{l}$}&\text{for $1\leq l<l^{\prime}\leq s$,}\\ \text{and the other components are zero.}\end{cases}

Here, we do not specify the orders of the sets S1S_{1} and S2S2S_{2}\cup S_{2}^{\prime}. The ambiguity does not matter for our purpose.

For later use, we also define a matrix

Ns(τ1,,τs)N_{s}(\tau_{1},\dots,\tau_{s})

as the submatrix of MM with the rows in S2S_{2} removed. More precisely, we define the matrix Ns(τ1,,τs)N_{s}(\tau_{1},\dots,\tau_{s}) with the columns (resp. rows) indexed by S1S_{1} (resp. S2S_{2}^{\prime}), by

{the (xlxl,xl) component is τlfor 1l<ls,the (xlxl,xl) component is τlfor 1l<ls,and the other components are zero.\begin{cases}\text{the $(x_{l}x_{l^{\prime}},x_{l})$ component is $-\tau_{l^{\prime}}$}&\text{for $1\leq l<l^{\prime}\leq s$,}\\ \text{the $(x_{l}x_{l^{\prime}},x_{l^{\prime}})$ component is $\tau_{l}$}&\text{for $1\leq l<l^{\prime}\leq s$,}\\ \text{and the other components are zero.}\end{cases}

Therefore, by choosing appropriate orders of rows and columns, we have

Ms(ν1,,νs,τ1,,τs)=(ν1νsNs(τ1,,τs)).M_{s}(\nu_{1},\dots,\nu_{s},\tau_{1},\dots,\tau_{s})=\begin{pmatrix}\begin{matrix}\nu_{1}&&\\ &\ddots&\\ &&\nu_{s}\\ \end{matrix}\\ N_{s}(\tau_{1},\dots,\tau_{s})\end{pmatrix}.
Example 4.5.

When s=3s=3, we have

M=(ν1ν2ν3τ3τ2τ3τ1τ2τ1)M=\begin{pmatrix}\nu_{1}&&\\ &\nu_{2}&\\ &&\nu_{3}\\ &-\tau_{3}&\tau_{2}\\ -\tau_{3}&&\tau_{1}\\ -\tau_{2}&\tau_{1}&\end{pmatrix}

Here, we use the order x2x3,x1x3,x1x2x_{2}x_{3},x_{1}x_{3},x_{1}x_{2} for the set S2S_{2}^{\prime}.

Proposition 4.6.

The matrix Ms(ν1,,νs,τ1,,τs)M_{s}(\nu_{1},\dots,\nu_{s},\tau_{1},\dots,\tau_{s}), over [G]\mathbb{Z}[G], is a presentation matrix of the module I\mathcal{I}_{I}.

Proof.

By the construction, MM is a presentation matrix of Ker([I])\operatorname{Ker}(\mathbb{Z}[I]\to\mathbb{Z}) over [I]\mathbb{Z}[I]. Since [G]\mathbb{Z}[G] is flat over [I]\mathbb{Z}[I], we obtain the proposition by base change. ∎

Proposition 4.7.

For each 0is0\leq i\leq s, we have

Fitti,[G](M)=j=0sia{1,2,,s}#a=jνa1νajFitti,[G](Nsj(τaj+1,,τas)).\operatorname{Fitt}_{i,\mathbb{Z}[G]}(M)=\sum_{j=0}^{s-i}\sum_{\begin{subarray}{c}a\subset\{1,2,\dots,s\}\\ \#a=j\end{subarray}}\nu_{a_{1}}\cdots\nu_{a_{j}}\operatorname{Fitt}_{i,\mathbb{Z}[G]}(N_{s-j}(\tau_{a_{j+1}},\dots,\tau_{a_{s}})).

Here, for each jj, in the second summation aa runs over subsets of {1,2,,s}\{1,2,\dots,s\} of jj elements, and for each aa we define a1,,asa_{1},\dots,a_{s} by requiring

a={a1,,aj},{a1,,as}={1,2,,s},a1<<aj,aj+1<<as.a=\{a_{1},\dots,a_{j}\},\qquad\{a_{1},\dots,a_{s}\}=\{1,2,\dots,s\},\qquad a_{1}<\dots<a_{j},\qquad a_{j+1}<\dots<a_{s}.

The matrix Nsj(τaj+1,,τas)N_{s-j}(\tau_{a_{j+1}},\dots,\tau_{a_{s}}) is defined as in Definition 4.4 for sjs-j and τaj+1,,τas\tau_{a_{j+1}},\dots,\tau_{a_{s}} instead of ss and τ1,,τs\tau_{1},\dots,\tau_{s}.

Proof.

By the definition of higher Fitting ideals, Fitti,[G](M)\operatorname{Fitt}_{i,\mathbb{Z}[G]}(M) is generated by det(H)\det(H) for square submatrices HH of MM of size sis-i. Such a matrix HH is in one-to-one correspondence with choices of a subset AHcolumnS1={x1,,xs}A_{H}^{\operatorname{column}}\subset S_{1}=\{x_{1},\dots,x_{s}\} with #AHcolumn=si\#A_{H}^{\operatorname{column}}=s-i and a subset AHrowS2S2={x12,,xs2,x1x2,,xs1xs}A_{H}^{\operatorname{row}}\subset S_{2}\cup S_{2}^{\prime}=\{x_{1}^{2},\dots,x_{s}^{2},x_{1}x_{2},\dots,x_{s-1}x_{s}\} with #AHrow=si\#A_{H}^{\operatorname{row}}=s-i. We only have to deal with HH satisfying det(H)0\det(H)\neq 0.

For each HH, we define jj and aa by

j=#(AHrowS2)j=\#(A_{H}^{\operatorname{row}}\cap S_{2})

(so clearly 0jsi0\leq j\leq s-i) and

AHrowS2={xa12,,xaj2}.A_{H}^{\operatorname{row}}\cap S_{2}=\{x_{a_{1}}^{2},\dots,x_{a_{j}}^{2}\}.

Recall that the xl2x_{l}^{2} row in the matrix MM contains a unique non-zero component νl\nu_{l} in the xlx_{l} column. Therefore, the assumption det(H)0\det(H)\neq 0 forces xa1,,xajAHcolumnx_{a_{1}},\dots,x_{a_{j}}\in A_{H}^{\operatorname{column}} and

det(H)=±νa1νajdet(H),\det(H)=\pm\nu_{a_{1}}\cdots\nu_{a_{j}}\det(H^{\prime}),

where HH^{\prime} is the square submatrix of HH of size (si)j(s-i)-j, with rows in AHrow=AHrow{xa12,,xaj2}=AHrowS2A_{H^{\prime}}^{\operatorname{row}}=A_{H}^{\operatorname{row}}\setminus\{x_{a_{1}}^{2},\dots,x_{a_{j}}^{2}\}=A_{H}^{\operatorname{row}}\cap S_{2}^{\prime} and columns in AHcolumn=AHcolumn{xa1,,xaj}A_{H^{\prime}}^{\operatorname{column}}=A_{H}^{\operatorname{column}}\setminus\{x_{a_{1}},\dots,x_{a_{j}}\}.

Let NaN_{a} denote the submatrix of Ns(τ1,,τs)N_{s}(\tau_{1},\dots,\tau_{s}) obtained by removing the xa1,,xajx_{a_{1}},\dots,x_{a_{j}} columns. Then it is clear that det(H)\det(H^{\prime})’s (for fixed jj and aa) as above generate Fitti,[G](Na)\operatorname{Fitt}_{i,\mathbb{Z}[G]}(N_{a}). The argument so far implies

Fitti,[G](M)=j=0sia{1,2,,s}#a=jνa1νajFitti,[G](Na).\operatorname{Fitt}_{i,\mathbb{Z}[G]}(M)=\sum_{j=0}^{s-i}\sum_{\begin{subarray}{c}a\subset\{1,2,\dots,s\}\\ \#a=j\end{subarray}}\nu_{a_{1}}\cdots\nu_{a_{j}}\operatorname{Fitt}_{i,\mathbb{Z}[G]}(N_{a}).

By the formula τlνl=0\tau_{l}\nu_{l}=0, we may remove the components ±τa1,,±τaj\pm\tau_{a_{1}},\dots,\pm\tau_{a_{j}} from the matrix NaN_{a} in the right hand side. It is easy to check that the resulting matrix is nothing but Nsj(τaj+1,,τas)N_{s-j}(\tau_{a_{j+1}},\dots,\tau_{a_{s}}) (with several zero rows added). This completes the proof. ∎

Proposition 4.8.

For s0s\geq 0 and i0i\geq 0, we have

Fitti,[G](Ns(τ1,,τs))={(1)(is)0(s1,i=0)(τ1,,τs)si(1i<s)\operatorname{Fitt}_{i,\mathbb{Z}[G]}(N_{s}(\tau_{1},\dots,\tau_{s}))=\begin{cases}(1)&(i\geq s)\\ 0&(s\geq 1,i=0)\\ (\tau_{1},\dots,\tau_{s})^{s-i}&(1\leq i<s)\\ \end{cases}
Proof.

Since Ns(τ1,,τs)N_{s}(\tau_{1},\dots,\tau_{s}) has ss columns, the case for isi\geq s is clear.

We show the vanishing when s1s\geq 1 and i=0i=0. Let R=[T1,,Ts]R=\mathbb{Z}[T_{1},\dots,T_{s}] be the polynomial ring over \mathbb{Z}. Then we have a ring homomorphism f:R[G]f:R\to\mathbb{Z}[G] defined by sending TlT_{l} to τl\tau_{l}. We define a matrix Ns(T1,,Ts)N_{s}(T_{1},\dots,T_{s}) over RR in the same way as in Definition 4.4, with τ\tau_{\bullet} replaced by TT_{\bullet}. Then, by the base change via ff, we have

Fitt[G](Ns(τ1,,τs))=f(FittR(Ns(T1,,Ts)))[G].\operatorname{Fitt}_{\mathbb{Z}[G]}(N_{s}(\tau_{1},\dots,\tau_{s}))=f(\operatorname{Fitt}_{R}(N_{s}(T_{1},\dots,T_{s})))\mathbb{Z}[G].

Hence the left hand side would vanish if we show that FittR(Ns(T1,,Ts))=0\operatorname{Fitt}_{R}(N_{s}(T_{1},\dots,T_{s}))=0.

For each 1ls1\leq l\leq s, we consider the complex

C~l:0[Tl]Tl[Tl]0,\widetilde{C}^{l}:0\to\mathbb{Z}[T_{l}]\overset{T_{l}}{\to}\mathbb{Z}[T_{l}]\to 0,

over [Tl]\mathbb{Z}[T_{l}], which satisfies Hn(C~l)=0H_{n}(\widetilde{C}^{l})=0 for n0n\neq 0 and H0(C~l)H_{0}(\widetilde{C}^{l})\simeq\mathbb{Z}. Similarly as previous, by taking the tensor product of the complexes C~l\widetilde{C}^{l} over \mathbb{Z}, we obtain an exact sequence

C~2Ns(T1,,Ts)C~1(T1Ts)C~00\cdots\to\widetilde{C}_{2}\overset{N_{s}(T_{1},\dots,T_{s})}{\to}\widetilde{C}_{1}\overset{\tiny\begin{pmatrix}T_{1}\\ \vdots\\ T_{s}\end{pmatrix}}{\to}\widetilde{C}_{0}\to\mathbb{Z}\to 0

over RR. (Alternatively, this exact sequence is obtained from the Koszul complex for the regular sequence T1,,TsT_{1},\dots,T_{s}.) This implies that FittR(Ns(T1,,Ts))\operatorname{Fitt}_{R}(N_{s}(T_{1},\dots,T_{s})) is the Fitting ideal of the augmentation ideal of RR. Since s1s\geq 1, the augmentation ideal of RR is generically of rank one, so we obtain FittR(Ns(T1,,Ts))=0,\operatorname{Fitt}_{R}(N_{s}(T_{1},\dots,T_{s}))=0, as desired.

Finally we show the case where 1i<s1\leq i<s. Since the components of the matrix Ns(τ1,,τs)N_{s}(\tau_{1},\dots,\tau_{s}) are either 0 or one of τ1,,τs\tau_{1},\dots,\tau_{s}, the inclusion \subset is clear. In order to show the other inclusion, we use the induction on ss.

For a while we fix an arbitrary 1ls1\leq l\leq s. Then, by permuting the rows and columns, the matrix Ns(τ1,,τs)N_{s}(\tau_{1},\dots,\tau_{s}) can be transformed into

(Ns1(τ1,,τlˇ,,τs)τlτlτlτlτ1τlˇτs).\begin{pmatrix}N_{s-1}(\tau_{1},\dots,\check{\tau_{l}},\dots,\tau_{s})&\\ \begin{array}[]{ccccc}-\tau_{l}&&&&\\ &-\tau_{l}&&&\\ &&\ddots&&\\ &&&-\tau_{l}&\\ &&&&-\tau_{l}\end{array}&\begin{array}[]{c}\tau_{1}\\ \vdots\\ \check{\tau_{l}}\\ \vdots\\ \tau_{s}\end{array}\end{pmatrix}.

(The symbol ()ˇ\check{(-)} means omitting that term.) Here, the xlx_{l} column is placed in the right-most, and the x1xl,,xl1xl,xlxl+1,,xlxsx_{1}x_{l},\dots,x_{l-1}x_{l},x_{l}x_{l+1},\dots,x_{l}x_{s} 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

Fitti,[G](Ns(τ1,,τs))(τ1,,τlˇ,,τs)Fitti,[G](Ns1(τ1,,τlˇ,,τs)).\operatorname{Fitt}_{i,\mathbb{Z}[G]}(N_{s}(\tau_{1},\dots,\tau_{s}))\supset(\tau_{1},\dots,\check{\tau_{l}},\dots,\tau_{s})\operatorname{Fitt}_{i,\mathbb{Z}[G]}(N_{s-1}(\tau_{1},\dots,\check{\tau_{l}},\dots,\tau_{s})).

By the induction hypothesis (note that 1is11\leq i\leq s-1), we have

Fitti,[G](Ns(τ1,,τs))(τ1,,τlˇ,,τs)(τ1,,τlˇ,,τs)s1i=(τ1,,τlˇ,,τs)si.\operatorname{Fitt}_{i,\mathbb{Z}[G]}(N_{s}(\tau_{1},\dots,\tau_{s}))\supset(\tau_{1},\dots,\check{\tau_{l}},\dots,\tau_{s})(\tau_{1},\dots,\check{\tau_{l}},\dots,\tau_{s})^{s-1-i}=(\tau_{1},\dots,\check{\tau_{l}},\dots,\tau_{s})^{s-i}.

Now we vary ll and then obtain

Fitti,[G](Ns(τ1,,τs))l=1s(τ1,,τlˇ,,τs)si=(τ1,,τs)si,\operatorname{Fitt}_{i,\mathbb{Z}[G]}(N_{s}(\tau_{1},\dots,\tau_{s}))\supset\sum_{l=1}^{s}(\tau_{1},\dots,\check{\tau_{l}},\dots,\tau_{s})^{s-i}=(\tau_{1},\dots,\tau_{s})^{s-i},

where the last equality follows from si<ss-i<s. This completes the proof of the proposition. ∎

Now we incorporate Propositions 4.6, 4.7 and 4.8 to prove the following.

Proposition 4.9.

For 0is0\leq i\leq s, we define an ideal JiJ_{i} of [G]\mathbb{Z}[G] by

Ji={(ν1νs)=(νI)(i=0)j=0siZi+jIj=Zi+Zi+1I++ZsIsi(1is).J_{i}=\begin{cases}(\nu_{1}\cdots\nu_{s})=(\nu_{I})&(i=0)\\ \sum_{j=0}^{s-i}Z_{i+j}\mathcal{I}_{I}^{j}=Z_{i}+Z_{i+1}\mathcal{I}_{I}+\cdots+Z_{s}\mathcal{I}_{I}^{s-i}&(1\leq i\leq s).\end{cases}

Then we have

Fitti,[G](I)=Ji.\operatorname{Fitt}_{i,\mathbb{Z}[G]}(\mathcal{I}_{I})=J_{i}.
Proof.

By Propositions 4.6 and 4.7, we have

(4.2) Fitti,[G](I)\displaystyle\operatorname{Fitt}_{i,\mathbb{Z}[G]}(\mathcal{I}_{I}) =Fitti,[G](M)\displaystyle=\operatorname{Fitt}_{i,\mathbb{Z}[G]}(M)
(4.3) =j=0sia{1,2,,s}#a=jνa1νajFitti,[G](Nsj(τaj+1,,τas)).\displaystyle=\sum_{j=0}^{s-i}\sum_{\begin{subarray}{c}a\subset\{1,2,\dots,s\}\\ \#a=j\end{subarray}}\nu_{a_{1}}\cdots\nu_{a_{j}}\operatorname{Fitt}_{i,\mathbb{Z}[G]}(N_{s-j}(\tau_{a_{j+1}},\dots,\tau_{a_{s}})).

When i=0i=0, Proposition 4.8 implies

Fitt0,[G](Nsj(τaj+1,,τas))={(1)(j=s)0(0j<s).\operatorname{Fitt}_{0,\mathbb{Z}[G]}(N_{s-j}(\tau_{a_{j+1}},\dots,\tau_{a_{s}}))=\begin{cases}(1)&(j=s)\\ 0&(0\leq j<s).\\ \end{cases}

Clearly, j=sj=s forces a={1,2,,s}a=\{1,2,\dots,s\}, so we obtain

Fitt0,[G](I)=(ν1νs)=J0.\operatorname{Fitt}_{0,\mathbb{Z}[G]}(\mathcal{I}_{I})=(\nu_{1}\cdots\nu_{s})=J_{0}.

When 1is1\leq i\leq s, since 1isj1\leq i\leq s-j by the choice of jj, Proposition 4.8 implies

Fitti,[G](Nsj(τaj+1,,τas))=(τaj+1,,τas)sij.\operatorname{Fitt}_{i,\mathbb{Z}[G]}(N_{s-j}(\tau_{a_{j+1}},\dots,\tau_{a_{s}}))=(\tau_{a_{j+1}},\dots,\tau_{a_{s}})^{s-i-j}.

Then we obtain

Fitti,[G](I)=j=0sia{1,2,,s}#a=jνa1νaj(τaj+1,,τas)sij.\operatorname{Fitt}_{i,\mathbb{Z}[G]}(\mathcal{I}_{I})=\sum_{j=0}^{s-i}\sum_{\begin{subarray}{c}a\subset\{1,2,\dots,s\}\\ \#a=j\end{subarray}}\nu_{a_{1}}\cdots\nu_{a_{j}}(\tau_{a_{j+1}},\dots,\tau_{a_{s}})^{s-i-j}.

Using the relation νlτl=0\nu_{l}\tau_{l}=0, for each 0jsi0\leq j\leq s-i, we have

a{1,2,,s}#a=jνa1νaj(τaj+1,,τas)sij=a{1,2,,s}#a=jνa1νajIsij=ZsjIsij.\sum_{\begin{subarray}{c}a\subset\{1,2,\dots,s\}\\ \#a=j\end{subarray}}\nu_{a_{1}}\cdots\nu_{a_{j}}(\tau_{a_{j+1}},\dots,\tau_{a_{s}})^{s-i-j}=\sum_{\begin{subarray}{c}a\subset\{1,2,\dots,s\}\\ \#a=j\end{subarray}}\nu_{a_{1}}\cdots\nu_{a_{j}}\mathcal{I}_{I}^{s-i-j}=Z_{s-j}\mathcal{I}_{I}^{s-i-j}.

These formulas imply Fitti,[G](I)=Ji\operatorname{Fitt}_{i,\mathbb{Z}[G]}(\mathcal{I}_{I})=J_{i}. ∎

We are ready to prove Theorem 1.4.

Proof of Theorem 1.4.

By Propositions 4.3 and 4.9, we have

Fitt[G][1](A)=i=0sg~i1Ji.\operatorname{Fitt}^{[1]}_{\mathbb{Z}[G]}(A)=\sum_{i=0}^{s}\widetilde{g}^{i-1}J_{i}.

Then, noting J0=(νI)J_{0}=(\nu_{I}), we can deduce

hFitt[G][1](A)=(νI,(1νI#Iφ1)i=1sg~i1Ji)h\operatorname{Fitt}^{[1]}_{\mathbb{Z}[G]}(A)=\left(\nu_{I},\left(1-\frac{\nu_{I}}{\#I}\varphi^{-1}\right)\sum_{i=1}^{s}\widetilde{g}^{i-1}J_{i}\right)

in the same way as in the proof of Proposition 4.1. Then it is enough to show

(4.4) 𝒥=i=1sg~i1Ji.\mathcal{J}=\sum_{i=1}^{s}\widetilde{g}^{i-1}J_{i}.

We claim that

(4.5) (I,#I)Ji+1Ji(\mathcal{I}_{I},\#I)J_{i+1}\subset J_{i}

holds for 1is11\leq i\leq s-1. We first see

IJi+1=Ij=0si1Zi+1+jIj=j=1siZi+jIjJi.\mathcal{I}_{I}J_{i+1}=\mathcal{I}_{I}\sum_{j=0}^{s-i-1}Z_{i+1+j}\mathcal{I}_{I}^{j}=\sum_{j=1}^{s-i}Z_{i+j}\mathcal{I}_{I}^{j}\subset J_{i}.

We also have νIJi+1(νI)J0Ji\nu_{I}J_{i+1}\subset(\nu_{I})\subset J_{0}\subset J_{i}. Since (I,#I)=(I,νI)(\mathcal{I}_{I},\#I)=(\mathcal{I}_{I},\nu_{I}) as an ideal, these show the claim (4.5).

Using (4.5), we next show

(4.6) i=1sg~i1Ji=i=1s(1φ~1)i1Ji.\sum_{i=1}^{s}\widetilde{g}^{i-1}J_{i}=\sum_{i=1}^{s}(1-\widetilde{\varphi}^{-1})^{i-1}J_{i}.

More generally we actually show

i=1sg~i1Ji=i=1s(1φ~1)i1Ji\sum_{i=1}^{s^{\prime}}\widetilde{g}^{i-1}J_{i}=\sum_{i=1}^{s^{\prime}}(1-\widetilde{\varphi}^{-1})^{i-1}J_{i}

by induction on ss^{\prime}, for each 0ss0\leq s^{\prime}\leq s. The case s=0s^{\prime}=0 is trivial. For 1ss1\leq s^{\prime}\leq s, we have

(4.7) i=1sg~i1Ji\displaystyle\sum_{i=1}^{s^{\prime}}\widetilde{g}^{i-1}J_{i} =g~s1Js+i=1s1g~i1Ji\displaystyle=\widetilde{g}^{{s^{\prime}}-1}J_{s^{\prime}}+\sum_{i=1}^{s^{\prime}-1}\widetilde{g}^{i-1}J_{i}
(4.8) =(i=1s(1φ~1)i1(#I)si)Js+i=1s1(1φ~1)i1Ji.\displaystyle=\left(\sum_{i=1}^{s^{\prime}}(1-\widetilde{\varphi}^{-1})^{i-1}(\#I)^{s^{\prime}-i}\right)J_{s^{\prime}}+\sum_{i=1}^{s^{\prime}-1}(1-\widetilde{\varphi}^{-1})^{i-1}J_{i}.

Here, the second equality follows from the induction hypothesis and expanding the power g~s1\widetilde{g}^{s^{\prime}-1}. By (4.5), for 1is11\leq i\leq s^{\prime}-1, we have (#I)siJsJi.(\#I)^{s^{\prime}-i}J_{s^{\prime}}\subset J_{i}. Therefore, we obtain

(4.9) i=1sg~i1Ji\displaystyle\sum_{i=1}^{s^{\prime}}\widetilde{g}^{i-1}J_{i} =(1φ~1)s1Js+i=1s1(1φ~1)i1Ji\displaystyle=(1-\widetilde{\varphi}^{-1})^{s^{\prime}-1}J_{s^{\prime}}+\sum_{i=1}^{s^{\prime}-1}(1-\widetilde{\varphi}^{-1})^{i-1}J_{i}
(4.10) =i=1s(1φ~1)i1Ji.\displaystyle=\sum_{i=1}^{s^{\prime}}(1-\widetilde{\varphi}^{-1})^{i-1}J_{i}.

This completes the proof of (4.6).

The right hand side of (4.6) can be computed as

(4.11) i=1s(1φ~1)i1Ji\displaystyle\sum_{i=1}^{s}(1-\widetilde{\varphi}^{-1})^{i-1}J_{i} =i=1sj=0siZi+jIj(1φ~1)i1\displaystyle=\sum_{i=1}^{s}\sum_{j=0}^{s-i}Z_{i+j}\mathcal{I}_{I}^{j}(1-\widetilde{\varphi}^{-1})^{i-1}
(4.12) =k=1sj=0kZkIj(1φ~1)kj1\displaystyle=\sum_{k=1}^{s}\sum_{j=0}^{k}Z_{k}\mathcal{I}_{I}^{j}(1-\widetilde{\varphi}^{-1})^{k-j-1}
(4.13) =k=1sZkDk1=𝒥.\displaystyle=\sum_{k=1}^{s}Z_{k}\mathcal{I}_{D}^{k-1}=\mathcal{J}.

Here, the first equality follows from the definition of JiJ_{i}, the second by putting i+j=ki+j=k, the third by D=(I,1φ~1)\mathcal{I}_{D}=(\mathcal{I}_{I},1-\widetilde{\varphi}^{-1}), and the final by the definition of 𝒥\mathcal{J}. 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 vv 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 pp and always work over p\mathbb{Z}_{p}.

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

𝒜=hFittp[G][1](Ap),=(1νI#Iφ1)\mathcal{A}=h\operatorname{Fitt}^{[1]}_{\mathbb{Z}_{p}[G]}(A\otimes\mathbb{Z}_{p}),\qquad\mathcal{B}=\left(1-\frac{\nu_{I}}{\#I}\varphi^{-1}\right)

of p[G]\mathbb{Z}_{p}[G]. In Lemma 5.1 (resp. Lemma 5.2) below, we deal with the case where DD is not (resp. is) a pp-group. We will make use of the concrete description of 𝒜\mathcal{A} in Theorem 1.4. As we always work over p\mathbb{Z}_{p} instead of \mathbb{Z}, by abuse of notation, in this subsection we simply write I\mathcal{I}_{I}, D\mathcal{I}_{D}, ZiZ_{i}, and 𝒥\mathcal{J} for the extensions of those ideals from [G]\mathbb{Z}[G] to p[G]\mathbb{Z}_{p}[G]. We have no afraid of confusion due to this.

Let GG^{\prime} denote the maximal subgroup of GG whose order is prime to pp.

Lemma 5.1.

Let χ\chi be a faithful character of GG^{\prime}. Suppose that DD is not a pp-group. Then we have

𝒜χ=χ\mathcal{A}^{\chi}=\mathcal{B}^{\chi}

as fractional ideals of p[G]χ\mathbb{Z}_{p}[G]^{\chi}.

Proof.

We write D=DGD^{\prime}=D\cap G^{\prime} and I=IGI^{\prime}=I\cap G^{\prime}. We first note that Dχ=(1)\mathcal{I}_{D}^{\chi}=(1). This is because χ\chi is non-trivial on DD^{\prime} by the assumptions. Then we have 𝒥χ=(1)\mathcal{J}^{\chi}=(1) by Definition 1.2, so Theorem 1.4 implies

𝒜χ=(νI,(1νI#Iφ1))χ.\mathcal{A}^{\chi}=\left(\nu_{I},\left(1-\frac{\nu_{I}}{\#I}\varphi^{-1}\right)\right)^{\chi}.

We have to show

νIχ(1νI#Iφ1)χ.\nu_{I}^{\chi}\in\left(1-\frac{\nu_{I}}{\#I}\varphi^{-1}\right)^{\chi}.

When II^{\prime} is non-trivial, then νIχ=0\nu_{I}^{\chi}=0 as χ\chi is non-trivial on II^{\prime}, so this is obvious. Let us suppose that II^{\prime} is trivial. Since νI2=(#I)νI\nu_{I}^{2}=(\#I)\nu_{I}, we have

νI(1νI#Iφ1)=νI(1φ1).\nu_{I}\left(1-\frac{\nu_{I}}{\#I}\varphi^{-1}\right)=\nu_{I}\left(1-\varphi^{-1}\right).

The element (1φ~1)χ(1-\widetilde{\varphi}^{-1})^{\chi} of p[G]χ\mathbb{Z}_{p}[G]^{\chi} is a unit since D=(I,1φ~1)\mathcal{I}_{D}=(\mathcal{I}_{I},1-\widetilde{\varphi}^{-1}), Dχ=(1)\mathcal{I}_{D}^{\chi}=(1), and Iχ(1)\mathcal{I}_{I}^{\chi}\subsetneqq(1). This completes the proof. ∎

Lemma 5.2.

Suppose that II is non-trivial and that DD is a pp-group. Let s=rankp(I)s=\operatorname{rank}_{p}(I) be the pp-rank of II, that is, the number of minimal generators of II (note that s1s\geq 1).

  • (1)

    We have

    𝒜Ds1\mathcal{A}\supset\mathcal{I}_{D}^{s-1}\mathcal{B}

    as fractional ideals of p[G]\mathbb{Z}_{p}[G].

  • (2)

    Let ψ\psi be a character of GG such that ψ|G\psi|_{G^{\prime}} is faithful on GG^{\prime} and that ψ\psi is non-trivial on DD. Then we have

    ψ(𝒜)=ψ(D)s1ψ()\psi(\mathcal{A})=\psi\left(\mathcal{I}_{D}\right)^{s-1}\psi(\mathcal{B})

    as ideals of 𝒪ψ\mathcal{O}_{\psi}.

Proof.

We may take a decomposition (1.3) of II so that ss coincides with the pp-rank of II as the lemma, and then IlI_{l} is non-trivial for each 1ls1\leq l\leq s.

(1) By Definition 1.2, we have Ds1𝒥\mathcal{I}_{D}^{s-1}\subset\mathcal{J} (by the i=si=s term as Zs=(1)Z_{s}=(1)). Then Theorem 1.4 shows the claim (1).

(2) We first show ψ(𝒥)=ψ(D)s1\psi(\mathcal{J})=\psi(\mathcal{I}_{D})^{s-1}. By the claim (1) above, the inclusion ψ(𝒥)ψ(D)s1\psi(\mathcal{J})\supset\psi(\mathcal{I}_{D})^{s-1} holds. For each 1ls1\leq l\leq s, we observe (ψ(νl))(p)(\psi(\nu_{l}))\subset(p) since ψ(νl)\psi(\nu_{l}) is either 0 or #Il\#I_{l}. Moreover, we have (p)ψ(D)(p)\subset\psi(\mathcal{I}_{D}) since ψ\psi is non-trivial on DD and we have (p)(1ζ)(p)\subset(1-\zeta) if ζ\zeta is any non-trivial root of unity. These observations imply ψ(Zi)ψ(D)si\psi(Z_{i})\subset\psi(\mathcal{I}_{D})^{s-i} for 1is1\leq i\leq s. By the definition of 𝒥\mathcal{J}, we then have ψ(𝒥)ψ(D)s1\psi(\mathcal{J})\subset\psi(\mathcal{I}_{D})^{s-1} as claimed.

By Theorem 1.4 and the above claim, we have

ψ(𝒜)=(ψ(νI),ψ()ψ(D)s1).\psi(\mathcal{A})=\left(\psi(\nu_{I}),\psi(\mathcal{B})\psi(\mathcal{I}_{D})^{s-1}\right).

We have to show ψ(νI)ψ()ψ(D)s1\psi(\nu_{I})\in\psi(\mathcal{B})\psi(\mathcal{I}_{D})^{s-1}. This is obvious if ψ\psi is non-trivial on II. If ψ\psi is trivial on II, we have

ψ(νI)=#I(ps)ψ()ψ(D)s1,\psi(\nu_{I})=\#I\in(p^{s})\subset\psi(\mathcal{B})\psi(\mathcal{I}_{D})^{s-1},

where the last inclusion follows from ψ()=ψ(D)=(1ψ(φ)1)(p)\psi(\mathcal{B})=\psi(\mathcal{I}_{D})=(1-\psi(\varphi)^{-1})\supset(p). 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 pp and an odd character χ\chi of GG^{\prime}. Recall the χ\chi-component of the Stickelberger element θK/k,Tχ\theta_{K/k,T}^{\chi} defined as (1.4)

Lemma 5.3.

We have θK/k,Tχ0\theta_{K/k,T}^{\chi}\neq 0 if and only if there exists a character ψ\psi of GG such that ψ|G=χ\psi|_{G^{\prime}}=\chi and that ψ\psi is non-trivial on DvD_{v} for any vSχS(k)v\in S_{\chi}\setminus S_{\infty}(k).

Proof.

By (1.5) and the fact that ωTχ\omega_{T}^{\chi} is a non-zero-divisor, we have θK/k,Tχ0\theta_{K/k,T}^{\chi}\neq 0 if and only if there exists a character ψ\psi of GG such that ψ|G=χ\psi|_{G^{\prime}}=\chi and, for every vSχS(k)v\in S_{\chi}\setminus S_{\infty}(k), we have ψ(1νIv#Ivφv1)0\psi\left(1-\frac{\nu_{I_{v}}}{\#I_{v}}\varphi_{v}^{-1}\right)\neq 0. The last condition is equivalent to that ψ\psi is non-trivial on DvD_{v}. 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 χ\chi is a faithful character of GG^{\prime}. This is because we have (ClKTp)p[G]p[Gal(Kχ/k)]ClKχTp(\operatorname{Cl}_{K}^{T}\otimes\mathbb{Z}_{p})\otimes_{\mathbb{Z}_{p}[G]}\mathbb{Z}_{p}[\operatorname{Gal}(K_{\chi}/k)]\simeq\operatorname{Cl}_{K_{\chi}}^{T}\otimes\mathbb{Z}_{p} as the degree of K/KχK/K_{\chi} is prime to pp. Moreover, to simplify the notation, we write S=Sχ=Sram(K/k)S=S_{\chi}=S_{\operatorname{ram}}(K/k) and Sfin=SS(k)S_{\operatorname{fin}}=S\setminus S_{\infty}(k). Recall that, by Theorem 1.1, the condition (i) is equivalent to (1.6). As in §5.1, for each vSfinv\in S_{\operatorname{fin}}, we consider the fractional ideals of p[G]\mathbb{Z}_{p}[G]

𝒜v=hvFittp[G][1](Avp),v=(1νIv#Ivφv1).\mathcal{A}_{v}=h_{v}\operatorname{Fitt}^{[1]}_{\mathbb{Z}_{p}[G]}(A_{v}\otimes\mathbb{Z}_{p}),\qquad\mathcal{B}_{v}=\left(1-\frac{\nu_{I_{v}}}{\#I_{v}}\varphi_{v}^{-1}\right).

We first suppose (ii) and aim at showing (i). The case where θK/k,Tχ=0\theta_{K/k,T}^{\chi}=0 is trivial, so we assume that, for any vSfinv\in S_{\operatorname{fin}}, either (a) or (b) holds. Then we obtain vχ𝒜vχ\mathcal{B}_{v}^{\chi}\subset\mathcal{A}_{v}^{\chi} for any vSfinv\in S_{\operatorname{fin}}, 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 θK/k,Tχ0\theta_{K/k,T}^{\chi}\neq 0, we may take a character ψ\psi as in Lemma 5.3. By applying ψ\psi to (1.6), we obtain

vSfinψ(v)vSfinψ(𝒜v).\prod_{v\in S_{\operatorname{fin}}}\psi(\mathcal{B}_{v})\subset\prod_{v\in S_{\operatorname{fin}}}\psi(\mathcal{A}_{v}).

On the other hand, by Lemmas 5.1 and 5.2(2), for each vSfinv\in S_{\operatorname{fin}}, we have ψ(𝒜v)ψ(v)\psi(\mathcal{A}_{v})\subset\psi(\mathcal{B}_{v}). 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

vSfinψ(𝒜v)vSfinψ(v).\prod_{v\in S_{\operatorname{fin}}}\psi(\mathcal{A}_{v})\subsetneqq\prod_{v\in S_{\operatorname{fin}}}\psi(\mathcal{B}_{v}).

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 𝔾m\mathbb{G}_{m}. 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.