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

Minimal resolutions of Iwasawa modules

Takenori Kataoka Department of Mathematics, Faculty of Science Division II, Tokyo University of Science. 1-3 Kagurazaka, Shinjuku-ku, Tokyo 162-8601, Japan [email protected]  and  Masato Kurihara Faculty of Science and Technology, Keio University. 3-14-1 Hiyoshi, Kohoku-ku, Yokohama, Kanagawa 223-8522, Japan [email protected]
Abstract.

In this paper, we study the module-theoretic structure of classical Iwasawa modules. More precisely, for a finite abelian pp-extension K/kK/k of totally real fields and the cyclotomic p\mathbb{Z}_{p}-extension K/KK_{\infty}/K, we consider XK,S=Gal(MK,S/K)X_{K_{\infty},S}=\operatorname{Gal}(M_{K_{\infty},S}/K_{\infty}) where SS is a finite set of places of kk containing all ramifying places in KK_{\infty} and archimedean places, and MK,SM_{K_{\infty},S} is the maximal abelian pro-pp-extension of KK_{\infty} unramified outside SS. We give lower and upper bounds of the minimal numbers of generators and of relations of XK,SX_{K_{\infty},S} as a p[[Gal(K/k)]]\mathbb{Z}_{p}[[\operatorname{Gal}(K_{\infty}/k)]]-module, using the pp-rank of Gal(K/k)\operatorname{Gal}(K/k). This result explains the complexity of XK,SX_{K_{\infty},S} as a p[[Gal(K/k)]]\mathbb{Z}_{p}[[\operatorname{Gal}(K_{\infty}/k)]]-module when the pp-rank of Gal(K/k)\operatorname{Gal}(K/k) is large. Moreover, we prove an analogous theorem in the setting that K/kK/k is non-abelian. We also study the Iwasawa adjoint of XK,SX_{K_{\infty},S}, and the minus part of the unramified Iwasawa module for a CM-extension. In order to prove these theorems, we systematically study the minimal resolutions of XK,SX_{K_{\infty},S}.

1. Introduction

Throughout this paper we fix a prime number pp. We write FF_{\infty} for the cyclotomic p\mathbb{Z}_{p}-extension of FF for any number field FF.

Let K/kK/k be a finite abelian pp-extension of totally real fields (see Theorem 3.3 for the non-abelian case). We consider the abelian extension K/kK_{\infty}/k, whose Galois group we denote by 𝒢=Gal(K/k)\mathcal{G}=\operatorname{Gal}(K_{\infty}/k). Suppose that SS is a finite set of places of kk, containing all archimedean places and all places that ramify in KK_{\infty}. In particular, SS contains all pp-adic places. Let MK,SM_{K_{\infty},S} denote the maximal abelian pro-pp-extension of KK_{\infty} unramified outside SS. Our main purpose in this paper is to study the classical Iwasawa module XK,S=Gal(MK,S/K)X_{K_{\infty},S}=\operatorname{Gal}(M_{K_{\infty},S}/K_{\infty}) over the Iwasawa algebra =p[[𝒢]]\mathcal{R}=\mathbb{Z}_{p}[[\mathcal{G}]].

Define I𝒢I_{\mathcal{G}} to be the augmentation ideal of =p[[𝒢]]\mathcal{R}=\mathbb{Z}_{p}[[\mathcal{G}]], namely I𝒢=Ker(p[[𝒢]]p)I_{\mathcal{G}}=\operatorname{Ker}(\mathbb{Z}_{p}[[\mathcal{G}]]\to\mathbb{Z}_{p}). We write Q()Q(\mathcal{R}) for the total quotient ring of \mathcal{R}. We consider an \mathcal{R}-submodule \mathcal{R}^{\sim} of Q()Q(\mathcal{R}), which consists of elements xQ()x\in Q(\mathcal{R}), satisfying xI𝒢xI_{\mathcal{G}}\subset\mathcal{R}. This is the module of pseudo-measures of 𝒢\mathcal{G} in the sense of Serre. The pp-adic LL-function of Deligne and Ribet is an element gK/k,Sg_{K_{\infty}/k,S} in \mathcal{R}^{\sim}, satisfying the following property. Suppose that κ:𝒢p×\kappa:\mathcal{G}\to\mathbb{Z}_{p}^{\times} is the cyclotomic character. For a character ψ\psi of 𝒢\mathcal{G} of finite order with values in an algebraic closure ¯p\overline{\mathbb{Q}}_{p} of p\mathbb{Q}_{p} and for a positive integer nn, one can extend a character κnψ:𝒢¯p×\kappa^{n}\psi:\mathcal{G}\to\overline{\mathbb{Q}}_{p}^{\times} to a ring homomorphism ¯p\mathcal{R}\to\overline{\mathbb{Q}}_{p}, and also to ¯p\mathcal{R}^{\sim}\to\overline{\mathbb{Q}}_{p}. Then gK/k,Sg_{K_{\infty}/k,S} satisfies

κnψ(gF/k,S)=LS(1n,ψωn)\kappa^{n}\psi(g_{F_{\infty}/k,S})=L_{S}(1-n,\psi\omega^{-n})

for any character ψ\psi of 𝒢\mathcal{G} of finite order and for any positive integer n>0n\in\mathbb{Z}_{>0}, where LS(s,ψωn)L_{S}(s,\psi\omega^{-n}) is the SS-truncated LL-function, and ω\omega is the Teichmüller character.

Put G=Gal(K/k)G=\operatorname{Gal}(K_{\infty}/k_{\infty}). In [5, Theorem 3.3] and [6, Theorem 4.1], as a refinement of the usual main conjecture, Greither and the second author computed the Fitting ideal of XK,SX_{K_{\infty},S} as an \mathcal{R}-module to obtain

(1.1) Fitt(XK,S)=𝔞GI𝒢gK/k,S,\operatorname{Fitt}_{\mathcal{R}}(X_{K_{\infty},S})={\mathfrak{a}}_{G}I_{\mathcal{G}}g_{K_{\infty}/k,S},

where 𝔞G{\mathfrak{a}}_{G} is a certain ideal of \mathcal{R} which is determined only by the group structure of GG. The explicit description of 𝔞G{\mathfrak{a}}_{G} is obtained in [7, §1.2] by Greither, Tokio and the second author. We do not explain this ideal 𝔞G{\mathfrak{a}}_{G} in this paper, but only mention two facts. If ss is the pp-rank of GG (i.e., s=dim𝔽p(𝔽pG)s=\dim_{\mathbb{F}_{p}}(\mathbb{F}_{p}\otimes_{\mathbb{Z}}G)) and 𝔪{\mathfrak{m}}_{\mathcal{R}} is the maximal ideal of \mathcal{R}, then we have 𝔞G𝔪s(s1)/2{\mathfrak{a}}_{G}\subset{\mathfrak{m}}_{\mathcal{R}}^{s(s-1)/2}. Also, if GG is isomorphic to (/pm)s(\mathbb{Z}/p^{m})^{\oplus s}, then 𝔞G=(pm+I𝒢)s(s1)/2{\mathfrak{a}}_{G}=(p^{m}\mathcal{R}+I_{\mathcal{G}})^{s(s-1)/2}. We also note here that the classical main conjecture in Iwasawa theory studies the character component XK,SψX_{K_{\infty},S}^{\psi} for KK which corresponds to the kernel of ψ\psi. In this case, GG is cyclic, and only the case s=1s=1 is studied.

The above computation of Fitt(XK,S)\operatorname{Fitt}_{\mathcal{R}}(X_{K_{\infty},S}) suggests that XK,SX_{K_{\infty},S} is complicated as an \mathcal{R}-module when the pp-rank ss of GG is large. To understand such complicatedness, we study in this paper the minimal numbers of generators and relations of XK,SX_{K_{\infty},S}. Let gen(XK,S)\operatorname{gen}_{\mathcal{R}}(X_{K_{\infty},S}) (resp. r(XK,S)r_{\mathcal{R}}(X_{K_{\infty},S})) be the minimal number of generators (resp. of relations) of XK,SX_{K_{\infty},S} as an \mathcal{R}-module.

In order to state the main result of this paper, we need the maximal abelian pro-pp-extension Mk,SM_{k,S} of kk unramified outside SS. By our choice of SS, we have

kKMk,S.k_{\infty}\subset K_{\infty}\subset M_{k,S}.

Now we state the main result of this paper. For any abelian group AA, we define its pp-rank by rankpA=dim𝔽p(𝔽pA)\operatorname{rank}_{p}A=\dim_{\mathbb{F}_{p}}(\mathbb{F}_{p}\otimes_{\mathbb{Z}}A), which is finite in all cases we consider in this paper.

Theorem 1.1.

Let us write

s=rankpGal(K/k)=rankpG,t=rankpGal(Mk,S/K)s=\operatorname{rank}_{p}\operatorname{Gal}(K_{\infty}/k_{\infty})=\operatorname{rank}_{p}G,\quad t=\operatorname{rank}_{p}\operatorname{Gal}(M_{k,S}/K_{\infty})

for the pp-ranks of the Galois groups. Then we have

max{s(s+1)2,t}gen(XK,S)s(s+1)2+t\max\left\{\frac{s(s+1)}{2},t\right\}\leq\operatorname{gen}_{\mathcal{R}}(X_{K_{\infty},S})\leq\frac{s(s+1)}{2}+t

and

r(XK,S)=s(s+1)(s+2)6+gen(XK,S).r_{\mathcal{R}}(X_{K_{\infty},S})=\frac{s(s+1)(s+2)}{6}+\operatorname{gen}_{\mathcal{R}}(X_{K_{\infty},S}).
Remark 1.2.

We note that Leopoldt’s conjecture for kk is equivalent to that the extension Mk,S/kM_{k,S}/k_{\infty} is finite though we do not assume it in Theorem 1.1. We get the equalities of the minimal numbers of generators in the following special cases.

  • (1)

    If Leopoldt’s conjecture holds and p3p\geq 3, then we may take KK so that K=Mk,SK_{\infty}=M_{k,S}. In this case, we have t=0t=0, so the theorem says

    gen(XMk,S,S)=s(s+1)2,\operatorname{gen}_{\mathcal{R}}(X_{M_{k,S},S})=\frac{s(s+1)}{2},

    where s=rankpGal(Mk,S/k)s=\operatorname{rank}_{p}\operatorname{Gal}(M_{k,S}/k_{\infty}).

  • (2)

    In case K=kK_{\infty}=k_{\infty}, we have s=0s=0, so the theorem says

    gen(Xk,S)=t=rankpGal(Mk,S/k).\operatorname{gen}_{\mathcal{R}}(X_{k_{\infty},S})=t=\operatorname{rank}_{p}\operatorname{Gal}(M_{k,S}/k_{\infty}).

    Indeed, this follows directly from Lemma 6.1.

Except for these cases, we have no theoretical method to determine the exact value of gen(XK,S)\operatorname{gen}_{\mathcal{R}}(X_{K_{\infty},S}) so far.

Remark 1.3.

In §7 we give several numerical examples for p=2p=2, 33. We take k=k=\mathbb{Q} and K/K/\mathbb{Q} which is a real abelian extension such that Gal(K/)(/p)s\operatorname{Gal}(K/\mathbb{Q})\simeq(\mathbb{Z}/p\mathbb{Z})^{\oplus s}. Here, we pick up some typical examples from §7.

  • (1)

    Take p=3p=3. For two primes i\ell_{i} with i=1i=1, 22 such that i1\ell_{i}\equiv 1 (mod 33), let KK be the unique (/3)2(\mathbb{Z}/3\mathbb{Z})^{\oplus 2}-extension over \mathbb{Q} with conductor 12\ell_{1}\ell_{2}. Take 1=7\ell_{1}=7, and 2\ell_{2} satisfying 21\ell_{2}\equiv 1 (mod 99), and consider S={3,1,2,}S=\{3,\ell_{1},\ell_{2},\infty\}. In this case s=2s=2 and t=1t=1. So Theorem 1.1 says that

    3gen(XK,S)4.3\leq\operatorname{gen}_{\mathcal{R}}(X_{K_{\infty},S})\leq 4.

    For 2\ell_{2} less than 200, we have

    gen(XK,S)=3if2=19,37,73,109,163,199\operatorname{gen}_{\mathcal{R}}(X_{K_{\infty},S})=3\ \ \mbox{if}\ \ \ell_{2}=19,37,73,109,163,199

    and

    gen(XK,S)=4otherwise, namely if2=127,181.\operatorname{gen}_{\mathcal{R}}(X_{K_{\infty},S})=4\ \ \mbox{otherwise, namely if}\ \ \ell_{2}=127,181.

    Thus the above inequality is sharp in this case.

  • (2)

    Take p=2p=2. Suppose that 1\ell_{1}, 2\ell_{2}, 3\ell_{3} are three distinct primes such that i1\ell_{i}\equiv 1 (mod 44). We take K=(1,2,3)K=\mathbb{Q}(\sqrt{\ell_{1}},\sqrt{\ell_{2}},\sqrt{\ell_{3}}) and S={2,1,2,3,}S=\{2,\ell_{1},\ell_{2},\ell_{3},\infty\}. Then Gal(K/)(/2)3\operatorname{Gal}(K/\mathbb{Q})\simeq(\mathbb{Z}/2\mathbb{Z})^{\oplus 3}, so s=3s=3. Taking account of the archimedean place, we know t=1+3=4t=1+3=4. Since s(s+1)/2=6s(s+1)/2=6, Theorem 1.1 says in this case

    6gen(XK,S)10.6\leq\operatorname{gen}_{\mathcal{R}}(X_{K_{\infty},S})\leq 10.

    Take 1=5\ell_{1}=5. For any 5<2<31005<\ell_{2}<\ell_{3}\leq 100, we have gen(XK,S)=7\operatorname{gen}_{\mathcal{R}}(X_{K_{\infty},S})=7 except for

    (2,3)=(17,89),(37,41),(41,61),(41,73),(41,89),(53,89),(73,89),(89,97),(\ell_{2},\ell_{3})=(17,89),(37,41),(41,61),(41,73),(41,89),(53,89),(73,89),(89,97),

    for which we have gen(XK,S)=8\operatorname{gen}_{\mathcal{R}}(X_{K_{\infty},S})=8. Also we have gen(XK,S)=9\operatorname{gen}_{\mathcal{R}}(X_{K_{\infty},S})=9 for (1,2,3)=(17,73,89)(\ell_{1},\ell_{2},\ell_{3})=(17,73,89), and gen(XK,S)=10\operatorname{gen}_{\mathcal{R}}(X_{K_{\infty},S})=10 for (1,2,3)=(73,89,97)(\ell_{1},\ell_{2},\ell_{3})=(73,89,97). We do not have gen(XK,S)=6\operatorname{gen}_{\mathcal{R}}(X_{K_{\infty},S})=6 at least in this range.

In this paper, we prove not only the above Theorem 1.1 but also its non-abelian generalization in Theorem 3.3. We also study and determine the minimal numbers of generators and relations of the dual (Iwasawa adjoint) of XK,SX_{K_{\infty},S} (see Theorem 3.4). This is relatively easier than Theorem 3.3. Also, we give in §3.3 some applications to the minus part of certain Iwasawa modules of CM-fields (see Corollary 3.5), using Kummer duality.

A key to the proof of our theorems is the existence of certain exact sequences, called Tate sequences. We remark here that Greither also used a different kind of Tate sequence in [3] to get information on the minimal numbers of generators of class groups of number fields. Our method of using the Tate sequences is totally different from Greither’s.

This paper is organized as follows. After algebraic preliminaries in §2, we will state the main results in §3. The proof is given in §§46. Finally in §7, we will observe numerical examples.

Acknowledgments

The authors would like to thank Yuta Nakamura, who computed gen(XK,S)\operatorname{gen}_{\mathcal{R}}(X_{K_{\infty},S}) for several examples in his master’s thesis in a slightly different situation from ours. They also thank Cornelius Greither heartily for his interest in the subject of this paper and for giving them some valuable comments. The first and the second authors are supported by JSPS KAKENHI Grant Numbers 22K13898 and 22H01119, respectively.

2. Algebraic preliminaries

2.1. Minimal resolutions

Let RR be a Noetherian local ring, which we do not assume to be commutative. Let 𝔪\mathfrak{m} be the Jacobson radical of RR, that is, 𝔪\mathfrak{m} is the maximal left (right) ideal of RR. For simplicity, let us assume that 𝐤:=R/𝔪\mathbf{k}:=R/\mathfrak{m} is a commutative field. We will often consider the case R=p[[𝒢]]R=\mathbb{Z}_{p}[[\mathcal{G}]] for a pro-pp group 𝒢\mathcal{G}, in which case RR is indeed local and we have 𝐤=𝔽p\mathbf{k}=\mathbb{F}_{p} (see [13, Proposition 5.2.16 (iii)]).

Definition 2.1.

For a finitely generated (left) RR-module MM, we write genR(M)\operatorname{gen}_{R}(M) for the minimal number of generators of MM as an RR-module. Also, we write rR(M)r_{R}(M) for the minimal number of relations of MM as an RR-module (see Definition 2.4 below).

Remark 2.2.

The following observations will be often used.

  • (1)

    By Nakayama’s lemma (e.g., [8, Corollary 13.12]), we have

    genR(M)=genR/I((R/I)RM)\operatorname{gen}_{R}(M)=\operatorname{gen}_{R/I}((R/I)\otimes_{R}M)

    for any two-sided ideal I𝔪I\subset\mathfrak{m} of RR. In particular, we have

    genR(M)=gen𝐤(𝐤RM)=dim𝐤(𝐤RM).\operatorname{gen}_{R}(M)=\operatorname{gen}_{\mathbf{k}}(\mathbf{k}\otimes_{R}M)=\dim_{\mathbf{k}}(\mathbf{k}\otimes_{R}M).

    Therefore, for a finitely generated p\mathbb{Z}_{p}-module MM, we have

    genp(M)=dim𝔽p(𝔽ppM)=rankp(M).\operatorname{gen}_{\mathbb{Z}_{p}}(M)=\dim_{\mathbb{F}_{p}}(\mathbb{F}_{p}\otimes_{\mathbb{Z}_{p}}M)=\operatorname{rank}_{p}(M).
  • (2)

    If we have an exact sequence

    0MMM′′00\to M^{\prime}\to M\to M^{\prime\prime}\to 0

    of finitely generated RR-modules, we have

    genR(M′′)genR(M)genR(M)+genR(M′′).\operatorname{gen}_{R}(M^{\prime\prime})\leq\operatorname{gen}_{R}(M)\leq\operatorname{gen}_{R}(M^{\prime})+\operatorname{gen}_{R}(M^{\prime\prime}).

    The proof is standard.

  • (3)

    In item (2) above, if we assume that RR is a discrete valuation ring (DVR), the formula is refined as

    max{genR(M),genR(M′′)}genR(M)genR(M)+genR(M′′).\max\left\{\operatorname{gen}_{R}(M^{\prime}),\operatorname{gen}_{R}(M^{\prime\prime})\right\}\leq\operatorname{gen}_{R}(M)\leq\operatorname{gen}_{R}(M^{\prime})+\operatorname{gen}_{R}(M^{\prime\prime}).

    This follows from the structure theorem for finitely generated modules over principal ideal domains.

Example 2.3.

Let us observe an example for which the formula in item (3) above does not hold when RR is not a DVR. Let R=p[[T]]R=\mathbb{Z}_{p}[[T]]. Consider M=p[[T]]M=\mathbb{Z}_{p}[[T]] and its submodule

M=(p,T)n=(pn,pn1T,,pTn1,Tn)M^{\prime}=(p,T)^{n}=(p^{n},p^{n-1}T,\dots,pT^{n-1},T^{n})

with n1n\geq 1. Then we have genR(M)=n+1\operatorname{gen}_{R}(M^{\prime})=n+1 and genR(M)=1\operatorname{gen}_{R}(M)=1, so genR(M)genR(M)\operatorname{gen}_{R}(M^{\prime})\leq\operatorname{gen}_{R}(M) does not hold.

Next we introduce the minimal resolutions of modules.

Definition 2.4.

Let MM be a finitely generated RR-module. We can construct an exact sequence of RR-modules

Rr2Rr1Rr0M0\cdots\to R^{r_{2}}\to R^{r_{1}}\to R^{r_{0}}\to M\to 0

such that the image of each homomorphism Rrn+1RrnR^{r_{n+1}}\to R^{r_{n}} (n0n\geq 0) is contained in 𝔪rn\mathfrak{m}^{r_{n}}. Such a sequence is called a minimal resolution of MM. In this case, since Rrn+1RrnR^{r_{n+1}}\to R^{r_{n}} induces the zero map on (R/𝔪)rn+1(R/𝔪)rn(R/\mathfrak{m})^{r_{n+1}}\to(R/\mathfrak{m})^{r_{n}}, by the definition of the Tor\operatorname{Tor} functor, the integer rnr_{n} coincides with

rn(M)=rnR(M):=dim𝐤TornR(𝐤,M)r_{n}(M)=r_{n}^{R}(M):=\dim_{\mathbf{k}}\operatorname{Tor}_{n}^{R}(\mathbf{k},M)

for n0n\geq 0. In particular, the integer rnr_{n} is independent of the choice of minimal resolutions. By definition we have

genR(M)=r0R(M),rR(M)=r1R(M).\operatorname{gen}_{R}(M)=r_{0}^{R}(M),\qquad r_{R}(M)=r_{1}^{R}(M).
Lemma 2.5.

If GG is a finite pp-group, we have

rnp[G](p)=dim𝔽pHn(G,𝔽p).r_{n}^{\mathbb{Z}_{p}[G]}(\mathbb{Z}_{p})=\dim_{\mathbb{F}_{p}}H_{n}(G,\mathbb{F}_{p}).
Proof.

This follows from Hn(G,𝔽p)Tornp[G](p,𝔽p)Tornp[G](𝔽p,p)H_{n}(G,\mathbb{F}_{p})\simeq\operatorname{Tor}_{n}^{\mathbb{Z}_{p}[G]}(\mathbb{Z}_{p},\mathbb{F}_{p})\simeq\operatorname{Tor}_{n}^{\mathbb{Z}_{p}[G]}(\mathbb{F}_{p},\mathbb{Z}_{p}) and the formula in Definition 2.4. ∎

2.2. Group homology

In this subsection, we summarize facts about group homology.

Let GG be a finite group. The following lemma is well-known.

Lemma 2.6.

We have

H1(G,)Gab,H_{1}(G,\mathbb{Z})\simeq G^{\operatorname{ab}},

the abelianization of GG, and

H1(G,/M)Gab/MH_{1}(G,\mathbb{Z}/M\mathbb{Z})\simeq G^{\operatorname{ab}}/M

for M1M\in\mathbb{Z}_{\geq 1}.

As for the second homology groups, if GG is abelian, it is known that H2(G,)H_{2}(G,\mathbb{Z}) is isomorphic to 2G\bigwedge^{2}G (see [1, Chap. V, Theorem 6.4 (iii)]). If GG is not abelian, H2(G,)H_{2}(G,\mathbb{Z}) is much harder to study, which is also known as the Schur multiplier of GG (cf. [11]).

For now, we observe a relation between Hn(G,)H_{n}(G,\mathbb{Z}) and Hn(G,/M)H_{n}(G,\mathbb{Z}/M\mathbb{Z}) for a pp-power MM.

Lemma 2.7.

Let n2n\geq 2. For any m1m\geq 1, we have

rankpHn(G,/pm)=rankpHn(G,)+rankpHn1(G,).\operatorname{rank}_{p}H_{n}(G,\mathbb{Z}/p^{m}\mathbb{Z})=\operatorname{rank}_{p}H_{n}(G,\mathbb{Z})+\operatorname{rank}_{p}H_{n-1}(G,\mathbb{Z}).

In particular, as the right hand side is independent from mm, we have

rankpHn(G,/pm)=dim𝔽pHn(G,𝔽p).\operatorname{rank}_{p}H_{n}(G,\mathbb{Z}/p^{m}\mathbb{Z})=\dim_{\mathbb{F}_{p}}H_{n}(G,\mathbb{F}_{p}).
Proof.

This follows from the universal coefficient theorem (see [1, Chap. I, Proposition 0.8], for example), which says in our case that

0Hn(G,)/pmHn(G,/pm)Tor1(Hn1(G,),/pm)00\to H_{n}(G,\mathbb{Z})\otimes\mathbb{Z}/p^{m}\mathbb{Z}\to H_{n}(G,\mathbb{Z}/p^{m}\mathbb{Z})\to{\rm Tor}^{\mathbb{Z}}_{1}(H_{n-1}(G,\mathbb{Z}),\mathbb{Z}/p^{m})\to 0

is split exact. ∎

In case GG is abelian, it is not hard to compute the pp-rank of the nn-th homology group:

Lemma 2.8.

Suppose GG is abelian and put s=rankpGs=\operatorname{rank}_{p}G. Then we have

dim𝔽pHn(G,𝔽p)=s(s+1)(s+n1)n!\dim_{\mathbb{F}_{p}}H_{n}(G,\mathbb{F}_{p})=\frac{s(s+1)\cdots(s+n-1)}{n!}

for n0n\geq 0 (when n=0n=0, the right hand side is understood to be 11).

Proof.

By replacing GG by its pp-Sylow subgroup, we may assume that GG is a pp-group. As in [5, §1.2] or [12, §4.3], we can construct an explicit minimal free resolution of p\mathbb{Z}_{p} as an RR-module

Rs3Rs2Rs1Rs0p0\cdots\to R^{s_{3}}\to R^{s_{2}}\to R^{s_{1}}\to R^{s_{0}}\to\mathbb{Z}_{p}\to 0

with sn=s(s+1)(s+n1)/n!s_{n}=s(s+1)\cdots(s+n-1)/n!. Thus, the lemma follows from Lemma 2.5. ∎

We also need the following duality theorem between the cohomology groups and the homology groups (see [1, Chap VI Proposition 7.1], for example).

Lemma 2.9.

Let GG be a finite group and MM a (discrete) GG-module. We define its Pontryagin dual MM^{\vee} by M=Hom(M,/)M^{\vee}=\operatorname{Hom}(M,\mathbb{Q}/\mathbb{Z}). Then for any n0n\in\mathbb{Z}_{\geq 0}, we have an isomorphism between Hn(G,M)H^{n}(G,M) and Hom(Hn(G,M),/)\operatorname{Hom}(H_{n}(G,M^{\vee}),\mathbb{Q}/\mathbb{Z}).

3. The main results

3.1. Setting

As in §1, let pp be any prime number, kk a totally real field, and kk_{\infty} its cyclotomic p\mathbb{Z}_{p}-extension. For a finite set SS of places of kk such that SS contains all the archimedean places and all pp-adic places, we write Mk,SM_{k,S} for the maximal abelian pro-pp-extension of kk unramified outside SS.

Let K/kK_{\infty}/k be a pro-pp Galois extension of totally real fields such that KK_{\infty} contains kk_{\infty} and the extension K/kK_{\infty}/k_{\infty} is finite. We do not assume that K/kK_{\infty}/k is abelian, but we have to assume the following.

Assumption 3.1.

There exists an intermediate finite Galois extension K/kK/k of K/kK_{\infty}/k such that

K=kK,kK=k.K_{\infty}=k_{\infty}K,\quad k_{\infty}\cap K=k.

In other words, the map induced by the restriction maps

Gal(K/k)Gal(k/k)×Gal(K/k)\operatorname{Gal}(K_{\infty}/k)\to\operatorname{Gal}(k_{\infty}/k)\times\operatorname{Gal}(K/k)

is an isomorphism.

Lemma 3.2.

If K/kK_{\infty}/k is abelian, then Assumption 3.1 holds.

Proof.

We consider the restriction homomorphism f:Gal(K/k)Gal(k/k)f:\operatorname{Gal}(K_{\infty}/k)\twoheadrightarrow\operatorname{Gal}(k_{\infty}/k). Since ff is a homomorphism of p\mathbb{Z}_{p}-modules and the target is free, ff has a section. We take a section and define KK to be the fixed field of the image of the section. A point is that K/kK/k is then automatically Galois as K/kK_{\infty}/k is abelian. ∎

Set 𝒢=Gal(K/k)\mathcal{G}=\operatorname{Gal}(K_{\infty}/k) and G=Gal(K/k)G=\operatorname{Gal}(K_{\infty}/k_{\infty}). We take an SS such that K/kK_{\infty}/k is unramified outside SS. Let MK,S/KM_{K_{\infty},S}/K_{\infty} be the Galois group of the maximal abelian pro-pp-extension of KK_{\infty} that is unramified outside places lying above SS, and XK,S=Gal(MK,S/K)X_{K_{\infty},S}=\operatorname{Gal}(M_{K_{\infty},S}/K_{\infty}) as in the Introduction. Then it is known that XK,SX_{K_{\infty},S} is a finitely generated torsion module over the associated Iwasawa algebra =p[[𝒢]]\mathcal{R}=\mathbb{Z}_{p}[[\mathcal{G}]]. Since K/kK_{\infty}/k is a pro-pp extension, the algebra \mathcal{R} is a local ring whose residue field is 𝔽p\mathbb{F}_{p}.

3.2. The statements

We use the notation in §3.1. To state the result, let us put

sn=dim𝔽pHn(G,𝔽p)s_{n}=\dim_{\mathbb{F}_{p}}H_{n}(G,\mathbb{F}_{p})

for n0n\geq 0 (recall G=Gal(K/k)G=\operatorname{Gal}(K_{\infty}/k_{\infty})). For instance, we have s0=1s_{0}=1 and s1=rankpGabs_{1}=\operatorname{rank}_{p}G^{\operatorname{ab}} by Lemma 2.6. Recall that Lemma 2.8 tells us an explicit formula of sns_{n} in case K/kK_{\infty}/k_{\infty} is abelian; in particular, we have s2=s(s+1)/2s_{2}=s(s+1)/2 and s3=s(s+1)(s+2)/6s_{3}=s(s+1)(s+2)/6 with s=rankpG(=s1)s=\operatorname{rank}_{p}G(=s_{1}).

The following is the main result, which contains a non-abelian generalization of Theorem 1.1.

Theorem 3.3.

When Assumption 3.1 is satisfied, the following inequalities and equalities hold.

  • (1)

    Put t=rankpGal(Mk,S/Mk,SK)t=\operatorname{rank}_{p}\operatorname{Gal}(M_{k,S}/M_{k,S}\cap K_{\infty}). Then we have

    max{s2,t}gen(XK,S)s2+t.\max\left\{s_{2},t\right\}\leq\operatorname{gen}_{\mathcal{R}}(X_{K_{\infty},S})\leq s_{2}+t.
  • (2)

    We have

    rn(XK,S)=sn+2+sn+1r_{n}(X_{K_{\infty},S})=s_{n+2}+s_{n+1}

    for n2n\geq 2 and

    r1(XK,S)r0(XK,S)=r(XK,S)gen(XK,S)=s3.r_{1}(X_{K_{\infty},S})-r_{0}(X_{K_{\infty},S})=r_{\mathcal{R}}(X_{K_{\infty},S})-\operatorname{gen}_{\mathcal{R}}(X_{K_{\infty},S})=s_{3}.

It is easy to see that Theorem 3.3 implies Theorem 1.1, thanks to Lemma 3.2.

We also prove corresponding theorems concerning the dual (Iwasawa adjoint) of XK,SX_{K_{\infty},S}. For a finitely generated torsion \mathcal{R}-module MM, we define the dual (Iwasawa adjoint) of MM by

M=Ext1(M,).M^{*}=\operatorname{Ext}^{1}_{\mathcal{R}}(M,\mathcal{R}).

Put Λ=p[[Gal(K/K)]]\Lambda=\mathbb{Z}_{p}[[\operatorname{Gal}(K_{\infty}/K)]] by using Assumption 3.1, so \mathcal{R} is isomorphic to Λ[G]\Lambda[G]. Then we have

MExtΛ1(M,Λ)M^{*}\simeq\operatorname{Ext}^{1}_{\Lambda}(M,\Lambda)

because Hom(N,)HomΛ(N,Λ)\operatorname{Hom}_{\mathcal{R}}(N,\mathcal{R})\simeq\operatorname{Hom}_{\Lambda}(N,\Lambda) for any \mathcal{R}-module NN. Therefore, our MM^{*} coincides with the Iwasawa adjoint of MM in [13, Definition 5.5.5], [9, §5.1], [10, §1.3].

We are interested in the \mathcal{R}-module XK,SX_{K_{\infty},S}^{*}. It is known that the structure of XK,SX_{K_{\infty},S}^{*} is often simpler than XK,SX_{K_{\infty},S} itself (e.g., when we are concerned with their Fitting ideals). The following theorem implies that we encounter such a phenomenon when we are concerned with the minimal resolutions.

Theorem 3.4.

When Assumption 3.1 is satisfied, the following equalities hold.

  • (1)

    We have

    gen(XK,S)=rankpGal(Mk,S/k).\operatorname{gen}_{\mathcal{R}}(X_{K_{\infty},S}^{*})=\operatorname{rank}_{p}\operatorname{Gal}(M_{k,S}/k_{\infty}).
  • (2)

    If KkK_{\infty}\supsetneqq k_{\infty}, then we have

    rn(XK,S)=sn2+sn3r_{n}(X_{K_{\infty},S}^{*})=s_{n-2}+s_{n-3}

    for n3n\geq 3,

    r2(XK,S)=s0+s0=2,r_{2}(X_{K_{\infty},S}^{*})=s_{0}+s_{0}=2,

    and

    r1(XK,S)r0(XK,S)=r(XK,S)gen(XK,S)=s0=1.r_{1}(X_{K_{\infty},S}^{*})-r_{0}(X_{K_{\infty},S}^{*})=r_{\mathcal{R}}(X_{K_{\infty},S}^{*})-\operatorname{gen}_{\mathcal{R}}(X_{K_{\infty},S}^{*})=s_{0}=1.

    If K=kK_{\infty}=k_{\infty}, then we have rn(XK,S)=0r_{n}(X_{K_{\infty},S}^{*})=0 for n2n\geq 2 and r1(XK,S)r0(XK,S)=0r_{1}(X_{K_{\infty},S}^{*})-r_{0}(X_{K_{\infty},S}^{*})=0.

In §5, we will prove s2gen(XK,S)s_{2}\leq\operatorname{gen}_{\mathcal{R}}(X_{K_{\infty,S}}) in Theorem 3.3(1), Theorem 3.3(2), and Theorem 3.4(2). These parts follow only from the existence of the Tate sequence introduced in §4. The rest of the statements (tgen(XK,S)s2+tt\leq\operatorname{gen}_{\mathcal{R}}(X_{K_{\infty},S})\leq s_{2}+t in Theorem 3.3(1) and Theorem 3.4(1)) will be proved in §6.

3.3. Applications for the minus parts of Iwasawa modules for CM-extensions

In this subsection we apply the main theorems in the previous subsection to CM-extensions. We keep the notation in §3.1, so K/kK_{\infty}/k is an extension of totally real fields satisfying Assumption 3.1. Only in this subsection we assume that pp is odd, which is mainly for making the functor of taking the character component exact for characters of Gal(K(μp)/K)\operatorname{Gal}(K_{\infty}(\mu_{p})/K_{\infty}).

We consider the field 𝒦=K(μp){\mathcal{K}}_{\infty}=K_{\infty}(\mu_{p}) obtained by adjoining all pp-th roots of unity to KK_{\infty}. So 𝒦/k{\mathcal{K}}_{\infty}/k is a CM-extension. We also use an intermediate field 𝒦n{\mathcal{K}}_{n} of the p\mathbb{Z}_{p}-extension 𝒦/K(μp){\mathcal{K}}_{\infty}/K(\mu_{p}) such that [𝒦n:K(μp)]=pn[{\mathcal{K}}_{n}:K(\mu_{p})]=p^{n} for each n0n\geq 0. Let n{\mathcal{L}}_{n} be the maximal abelian pro-pp-extension of 𝒦n{\mathcal{K}}_{n} unramified everywhere. So Gal(n/𝒦n)\operatorname{Gal}({\mathcal{L}}_{n}/{\mathcal{K}}_{n}) is isomorphic to the pp-component A𝒦nA_{{\mathcal{K}}_{n}} of the ideal class group of 𝒦n{\mathcal{K}}_{n} by class field theory. We denote by A𝒦A_{{\mathcal{K}}_{\infty}} the inductive limit of A𝒦nA_{{\mathcal{K}}_{n}}, which is a discrete p[[Gal(𝒦/k)]]\mathbb{Z}_{p}[[\operatorname{Gal}({\mathcal{K}}_{\infty}/k)]]-module. We write 𝒳𝒦{\mathcal{X}}_{{\mathcal{K}}_{\infty}} for the projective limit of A𝒦nA_{{\mathcal{K}}_{n}}. Then 𝒳𝒦{\mathcal{X}}_{{\mathcal{K}}_{\infty}} is a compact p[[Gal(𝒦/k)]]\mathbb{Z}_{p}[[\operatorname{Gal}({\mathcal{K}}_{\infty}/k)]]-module. Defining {\mathcal{L}}_{\infty} to be the maximal abelian pro-pp-extension of 𝒦{\mathcal{K}}_{\infty} unramified everywhere, we know that 𝒳𝒦=Gal(/𝒦){\mathcal{X}}_{{\mathcal{K}}_{\infty}}=\operatorname{Gal}({\mathcal{L}}_{\infty}/{\mathcal{K}}_{\infty}). Let n0n_{0} be the smallest integer such that all pp-adic places are totally ramified in 𝒦/𝒦n0{\mathcal{K}}_{\infty}/{\mathcal{K}}_{n_{0}}. We define a submodule 𝒴𝒦{\mathcal{Y}}_{{\mathcal{K}}_{\infty}} of 𝒳𝒦{\mathcal{X}}_{{\mathcal{K}}_{\infty}} by 𝒴𝒦=Gal(/n0𝒦){\mathcal{Y}}_{{\mathcal{K}}_{\infty}}=\operatorname{Gal}({\mathcal{L}}_{\infty}/{\mathcal{L}}_{n_{0}}{\mathcal{K}}_{\infty}).

Put Δ=Gal(𝒦/K)\Delta=\operatorname{Gal}({\mathcal{K}}_{\infty}/K_{\infty}), which is of order prime to pp by our assumption p2p\neq 2 in this subsection. Therefore, since Gal(𝒦/k)𝒢×Δ\operatorname{Gal}({\mathcal{K}}_{\infty}/k)\simeq\mathcal{G}\times\Delta, any p[[Gal(𝒦/k)]]\mathbb{Z}_{p}[[\operatorname{Gal}({\mathcal{K}}_{\infty}/k)]]-module MM is decomposed into M=χMχM=\bigoplus_{\chi}M^{\chi} where χ\chi runs over all characters of Δ\Delta with values in p×\mathbb{Z}_{p}^{\times}, and MχM^{\chi} is the χ\chi-component of MM defined by

Mχ={xMσ(x)=χ(σ)xfor anyσΔ}.M^{\chi}=\{x\in M\mid\sigma(x)=\chi(\sigma)x\ \mbox{for any}\ \sigma\in\Delta\}.

Note that each MχM^{\chi} is an \mathcal{R}-module. Let ω:Δp×\omega:\Delta\to\mathbb{Z}_{p}^{\times} be the Teichmüller character, giving the action on μp\mu_{p}. Using our main results in §3.2, we study A𝒦ωA_{{\mathcal{K}}_{\infty}}^{\omega} and 𝒴𝒦ω{\mathcal{Y}}_{{\mathcal{K}}_{\infty}}^{\omega}. Note that ω\omega is an odd character, so the complex conjugation acts on these modules as 1-1.

Let SpS_{p} be the set of all pp-adic places and all archimedean places. Recall that we write ()(-)^{\vee} for the Pontryagin dual. By Kummer pairing (see [13, Theorem 11.4.3] or [15, Proposition 13.32]), we have an isomorphism

(A𝒦ω)(1)XK,Sp,(A_{{\mathcal{K}}_{\infty}}^{\omega})^{\vee}(1)\simeq X_{K_{\infty},S_{p}},

where (1)(1) is Tate twist. Also, by [13, Theorem 11.1.8] we have

(A𝒦)𝒴𝒦.(A_{{\mathcal{K}}_{\infty}})^{\vee}\simeq{\mathcal{Y}}_{{\mathcal{K}}_{\infty}}^{*}.

Therefore, we have

XK,Sp(1)(A𝒦ω)(1)(1)((A𝒦ω))((𝒴𝒦ω)).X_{K_{\infty},S_{p}}^{*}(1)\simeq(A_{{\mathcal{K}}_{\infty}}^{\omega})^{\vee}(1)^{*}(1)\simeq\left((A_{{\mathcal{K}}_{\infty}}^{\omega})^{\vee}\right)^{*}\simeq(({\mathcal{Y}}_{{\mathcal{K}}_{\infty}}^{\omega})^{*})^{*}.

For any finitely generated torsion \mathcal{R}-module MM which has no nontrivial finite submodule, we know (M)M(M^{*})^{*}\simeq M (see, for example, [13, Proposition 5.5.8 (iv)]). Since 𝒳𝒦,Sp{\mathcal{X}}_{{\mathcal{K}}_{\infty},S_{p}} has no nontrivial finite submodule, so does 𝒴𝒦ω{\mathcal{Y}}_{{\mathcal{K}}_{\infty}}^{\omega}. Therefore, it follows from the previous isomorphism that

XK,Sp𝒴𝒦ω(1)X_{K_{\infty},S_{p}}^{*}\simeq{\mathcal{Y}}_{{\mathcal{K}}_{\infty}}^{\omega}(-1)

is an isomorphism.

Thus, from Theorems 3.3 and 3.4 we get

Corollary 3.5.

In Theorem 3.3, we further assume that p>2p>2 and S=SpS=S_{p} (so K/kK_{\infty}/k_{\infty} is unramified outside pp).

  • (1)

    Then we have

    max{s2,t}gen((A𝒦ω))s2+t\max\left\{s_{2},t\right\}\leq\operatorname{gen}_{\mathcal{R}}((A_{{\mathcal{K}}_{\infty}}^{\omega})^{\vee})\leq s_{2}+t

    and

    r((A𝒦ω))=gen((A𝒦ω))+s3.r_{\mathcal{R}}((A_{{\mathcal{K}}_{\infty}}^{\omega})^{\vee})=\operatorname{gen}_{\mathcal{R}}((A_{{\mathcal{K}}_{\infty}}^{\omega})^{\vee})+s_{3}.
  • (2)

    Put t=rankpGal(Mk,S/k)t^{\prime}=\operatorname{rank}_{p}\operatorname{Gal}(M_{k,S}/k_{\infty}). Then we have gen(𝒴𝒦ω)=t\operatorname{gen}_{\mathcal{R}}({\mathcal{Y}}_{{\mathcal{K}}_{\infty}}^{\omega})=t^{\prime} and

    r(𝒴𝒦ω)={t+1if Kktif K=k.r_{\mathcal{R}}({\mathcal{Y}}_{{\mathcal{K}}_{\infty}}^{\omega})=\left\{\begin{array}[]{ll}t^{\prime}+1&\ \mbox{if $K_{\infty}\supsetneqq k_{\infty}$}\\ t^{\prime}&\ \mbox{if $K_{\infty}=k_{\infty}$.}\end{array}\right.

4. The Tate sequence

A key ingredient to prove Theorems 3.3 and 3.4 is an exact sequence that XK,SX_{K_{\infty},S} satisfies, which is often called the Tate sequence. Indeed, as noted in the final paragraph of §3.2, parts of main theorems can be deduced from the existence of the Tate sequence only. On the other hand, the other parts require additional arithmetic study that we will do in §6. The Tate sequence also played a key role in computing the Fitting ideal of XK,SX_{K_{\infty},S} in the work [5], [6], and [7] that we mentioned in §1.

In order to prove the main theorems, we need the Tate sequence of the following type.

Theorem 4.1.

There exists an exact sequence of \mathcal{R}-modules

0XK,SPϕQp0,0\to X_{K_{\infty},S}\to P\overset{\phi}{\to}Q\to\mathbb{Z}_{p}\to 0,

where PP and QQ are finitely generated torsion \mathcal{R}-modules whose projective dimensions are 1\leq 1. Moreover, this sequence is functorial when KK_{\infty} varies. More precisely, for a finite normal subgroup HH of Gal(K/k)\operatorname{Gal}(K_{\infty}/k), we have an exact sequence

0XKH,SPHϕHQHp00\to X_{K_{\infty}^{H},S}\to P_{H}\overset{\phi_{H}}{\to}Q_{H}\to\mathbb{Z}_{p}\to 0

over the Iwasawa algebra p[[Gal(KH/k)]]\mathbb{Z}_{p}[[\operatorname{Gal}(K_{\infty}^{H}/k)]], where PHP_{H}, QHQ_{H} denote the HH-coinvariant modules, and ϕH\phi_{H} the homomorphism induced by ϕ\phi.

Proof.

For an intermediate field FF of K/kK_{\infty}/k with [F:k]<[F:k]<\infty, we use a perfect complex 𝖱Γc(𝒪F,S,p(1))\operatorname{\mathsf{R}\Gamma}_{c}({\mathcal{O}}_{F,S},\mathbb{Z}_{p}(1)) in Burns–Flach [2, Proposition 1.20]. Note that this complex works well even for p=2p=2 (here, we use our assumption that SS contains all archimedean places). Taking the project limit, we get a perfect complex C=𝖱Γc(𝒪K,S,p(1))C^{\bullet}=\operatorname{\mathsf{R}\Gamma}_{c}({\mathcal{O}}_{K_{\infty},S},\mathbb{Z}_{p}(1)) which is quasi-isomorphic to a complex of the form [C0d0C1d1C2][C^{0}\overset{d^{0}}{\to}C^{1}\overset{d^{1}}{\to}C^{2}] concentrated on degrees 0, 11, 22 with CiC^{i} finitely generated projective over \mathcal{R} and whose cohomology groups are

H1(C)=H1(OK,S,p/p)=XK,S,H2(C)=H0(OK,S,p/p)=p,H^{1}(C^{\bullet})=H^{1}(O_{K_{\infty},S},\mathbb{Q}_{p}/\mathbb{Z}_{p})^{\vee}=X_{K_{\infty},S},\ \ H^{2}(C^{\bullet})=H^{0}(O_{K_{\infty},S},\mathbb{Q}_{p}/\mathbb{Z}_{p})^{\vee}=\mathbb{Z}_{p},

(see [2, page 86, line 6]) and Hi(C)=0H^{i}(C^{\bullet})=0 for i1i\neq 1, 22, where we used the weak Leopoldt conjecture which is proven in this case (see [13, Theorem 10.3.25]). By the definition of cohomology groups, we have an exact sequence

0XK,SC1/C0ΦC2p0,0\to X_{K_{\infty},S}\to C^{1}/C^{0}\overset{\Phi}{\to}C^{2}\to\mathbb{Z}_{p}\to 0,

where we regard C0C^{0} as a submodule of C1C^{1} via d0d^{0} and Φ\Phi is induced by d1d^{1}. Take a non-zero-divisor ff in the center of \mathcal{R} that annihilates p\mathbb{Z}_{p}. Then the image of Φ\Phi contains fC2fC^{2}, so by the projectivity of C2C^{2} we can construct a commutative diagram of \mathcal{R}-modules

C2\textstyle{C^{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}C2\textstyle{C^{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f×\scriptstyle{f\times}C1/C0\textstyle{C^{1}/C^{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Φ\scriptstyle{\Phi}C2.\textstyle{C^{2}.}

Then defining PP and QQ as the cokernel of these vertical maps respectively, we obtain the Tate sequence as claimed. The functoriality follows from that of 𝖱Γc(𝒪K,S,p(1))\operatorname{\mathsf{R}\Gamma}_{c}({\mathcal{O}}_{K_{\infty},S},\mathbb{Z}_{p}(1)). ∎

5. Abstract Tate sequences

Set Λ=p[[T]]\Lambda=\mathbb{Z}_{p}[[T]]. Let GG be a (not necessarily abelian) finite pp-group.

Motivated by Theorem 4.1, we study a Λ[G]\Lambda[G]-module XX that satisfies an abstract Tate sequence, that is:

Setting 5.1.

There exists an exact sequence of Λ[G]\Lambda[G]-modules

(5.1) 0XPϕQp0,0\to X\to P\overset{\phi}{\to}Q\to\mathbb{Z}_{p}\to 0,

where both PP and QQ are finitely generated torsion Λ[G]\Lambda[G]-modules whose projective dimensions are 1\leq 1.

In this section, we show that the existence of a Tate sequence gives a severe constraint on the integers genΛ[G](X)=r0(X)\operatorname{gen}_{\Lambda[G]}(X)=r_{0}(X), rΛ[G](X)=r1(X)r_{\Lambda[G]}(X)=r_{1}(X), and rn(X)r_{n}(X) (n2n\geq 2).

5.1. The statements

To state the result, let us define

sn=dim𝔽pHn(G,𝔽p)s_{n}=\dim_{\mathbb{F}_{p}}H_{n}(G,\mathbb{F}_{p})

for n0n\geq 0.

The following are the main theorems in this section. As noted in the final paragraph of §3.2, those are enough to show parts of Theorems 3.3 and 3.4.

Theorem 5.2.

Let XX be a Λ[G]\Lambda[G]-module that satisfies a Tate sequence as in Setting 5.1.

  • (1)

    We have genΛ[G](X)s2\operatorname{gen}_{\Lambda[G]}(X)\geq s_{2}.

  • (2)

    We have

    rn(X)=sn+2+sn+1r_{n}(X)=s_{n+2}+s_{n+1}

    for n2n\geq 2 and

    r1(X)r0(X)=s3.r_{1}(X)-r_{0}(X)=s_{3}.

Claims (1) and (2) will be proved respectively in §5.3 and §5.4.

For a Λ[G]\Lambda[G]-module MM, we define its dual (Iwasawa adjoint) by

M=ExtΛ[G]1(M,Λ[G])ExtΛ1(M,Λ).M^{*}=\operatorname{Ext}^{1}_{\Lambda[G]}(M,\Lambda[G])\simeq\operatorname{Ext}^{1}_{\Lambda}(M,\Lambda).

The corresponding theorem for the dual is:

Theorem 5.3.

Let XX be a Λ[G]\Lambda[G]-module that satisfies a Tate sequence as in Setting 5.1.

  • (1)

    We have genΛ[G](X)s1\operatorname{gen}_{\Lambda[G]}(X^{*})\geq s_{1}.

  • (2)

    If GG is non-trivial, then we have

    rn(X)=sn2+sn3r_{n}(X^{*})=s_{n-2}+s_{n-3}

    for n3n\geq 3,

    r2(X)=s0+s0(=2),r_{2}(X^{*})=s_{0}+s_{0}(=2),

    and

    r1(X)r0(X)=s0(=1).r_{1}(X^{*})-r_{0}(X^{*})=s_{0}(=1).

    If GG is trivial, then we have rn(X)=0r_{n}(X^{*})=0 for n2n\geq 2 and r1(X)r0(X)=0r_{1}(X^{*})-r_{0}(X^{*})=0.

This theorem will be proved in §5.6. The idea is basically the same as that of Theorem 5.2. However, we need an additional algebraic proposition shown in §5.5.

5.2. Specialization

We consider modules over Λ=p[[T]]\Lambda=\mathbb{Z}_{p}[[T]]. As explained in Example 2.3, genΛ()\operatorname{gen}_{\Lambda}(-) does not behave very well for short exact sequences. A key idea to prove the main theorems is to apply specialization method to reduce to modules over DVRs.

We define

={p}{fp[T]f is an irreducible monic distinguished polynomial}.\mathcal{F}=\{p\}\cup\{f\in\mathbb{Z}_{p}[T]\mid\text{$f$ is an irreducible monic distinguished polynomial}\}.

Here, a monic distinguished polynomial is by definition a polynomial of the form

Te+a1Te1++ae,T^{e}+a_{1}T^{e-1}+\cdots+a_{e},

where a1,,aeppa_{1},\dots,a_{e}\in p\mathbb{Z}_{p}. By the Weierstrass preparation theorem, any prime element of Λ\Lambda can be written as the product of a unit element and an element of \mathcal{F} in a unique way.

For each ff\in\mathcal{F}, put

𝒪f=Λ/(f),\mathcal{O}_{f}=\Lambda/(f),

which is a domain. We define a subset 0\mathcal{F}_{0}\subset\mathcal{F} by

0={f𝒪f is a DVR}.\mathcal{F}_{0}=\{f\in\mathcal{F}\mid\text{$\mathcal{O}_{f}$ is a DVR}\}.

The following lemma tells us a concrete description of 0\mathcal{F}_{0}. Although the lemma is unnecessary for the proof of the main results, we include it in this paper to clarify the situation.

Lemma 5.4.

We have 0={p}12\mathcal{F}_{0}=\{p\}\cup\mathcal{F}_{1}\cup\mathcal{F}_{2}, where we put

1={Tααpp}\mathcal{F}_{1}=\{T-\alpha\mid\alpha\in p\mathbb{Z}_{p}\}

and

2={Te+a1Te1++aee2,a1,,ae1pp,aeppp2p}.\mathcal{F}_{2}=\{T^{e}+a_{1}T^{e-1}+\dots+a_{e}\mid e\geq 2,a_{1},\dots,a_{e-1}\in p\mathbb{Z}_{p},a_{e}\in p\mathbb{Z}_{p}\setminus p^{2}\mathbb{Z}_{p}\}.
Proof.

It is clear that {p}0\{p\}\subset\mathcal{F}_{0} and 10\mathcal{F}_{1}\subset\mathcal{F}_{0}. Also, 20\mathcal{F}_{2}\subset\mathcal{F}_{0} holds by the Eisenstein irreducibility criterion. Therefore, it remains to only show 0{p}12\mathcal{F}_{0}\setminus\{p\}\subset\mathcal{F}_{1}\cup\mathcal{F}_{2}.

Let f0{p}f\in\mathcal{F}_{0}\setminus\{p\}. Since 𝒪f\mathcal{O}_{f} is a DVR, it is the integral closure of p\mathbb{Z}_{p} in the pp-adic field Kf=Frac(𝒪f)K_{f}=\operatorname{Frac}(\mathcal{O}_{f}). Moreover, since the residue field of 𝒪f\mathcal{O}_{f} is the same as that of Λ\Lambda, namely 𝔽p\mathbb{F}_{p}, we see that the extension Kf/pK_{f}/\mathbb{Q}_{p} is totally ramified. In case the extension Kf/pK_{f}/\mathbb{Q}_{p} is trivial, we have deg(f)=1\deg(f)=1, so we obtain f1f\in\mathcal{F}_{1}. In case Kf/pK_{f}/\mathbb{Q}_{p} is non-trivial, the image of TT in 𝒪f\mathcal{O}_{f} must be a uniformizer of 𝒪f\mathcal{O}_{f}, so its minimal polynomial ff is in 2\mathcal{F}_{2} (see [14, Chap. I, Proposition 18]). This completes the proof. ∎

5.3. Proof of Theorem 5.2(1)

Let us now study a Λ[G]\Lambda[G]-module XX satisfying a Tate sequence as in Setting 5.1. We define a Λ\Lambda-module X(G)X_{(G)} by

(5.2) 0X(G)PGϕGQGp0,0\to X_{(G)}\to P_{G}\overset{\phi_{G}}{\to}Q_{G}\to\mathbb{Z}_{p}\to 0,

where PGP_{G} and QGQ_{G} denote the GG-coinvariant modules and ϕG\phi_{G} denotes the induced homomorphism. Note that X(G)X_{(G)} does not coincide with the coinvariant module XGX_{G} in general; in fact, the difference is what we shall investigate from now on.

The following proposition is a key to prove the main theorem.

Proposition 5.5.

Let ff\in\mathcal{F} be an element that is prime to both charΛ(P)\operatorname{char}_{\Lambda}(P) and charΛ(Q)\operatorname{char}_{\Lambda}(Q), where charΛ()\operatorname{char}_{\Lambda}(-) denotes the characteristic polynomial. We set m=ordp(f(0))1m=\operatorname{ord}_{p}(f(0))\geq 1. Then we have an exact sequence of finitely generated torsion 𝒪f\mathcal{O}_{f}-modules

0H2(G,p/pmp)(X/f)GX(G)/f𝜋H1(G,p/pmp)0.0\to H_{2}(G,\mathbb{Z}_{p}/p^{m}\mathbb{Z}_{p})\to(X/f)_{G}\to X_{(G)}/f\overset{\pi}{\to}H_{1}(G,\mathbb{Z}_{p}/p^{m}\mathbb{Z}_{p})\to 0.
Proof.

Firstly note that f(0)0f(0)\neq 0 since charΛ(Q)\operatorname{char}_{\Lambda}(Q) is divisible by charΛ(p)=(T)\operatorname{char}_{\Lambda}(\mathbb{Z}_{p})=(T). By taking modulo ff of the sequence (5.1), we obtain an exact sequence of finitely generated torsion 𝒪f[G]\mathcal{O}_{f}[G]-modules

0X/fP/fϕ¯Q/fp/pmp0.0\to X/f\to P/f\overset{\overline{\phi}}{\to}Q/f\to\mathbb{Z}_{p}/p^{m}\mathbb{Z}_{p}\to 0.

Let LL denote the image of the map ϕ¯:P/fQ/f\overline{\phi}:P/f\to Q/f. Since both P/fP/f and Q/fQ/f are GG-cohomologically trivial, taking the GG-homology, we obtain exact sequences

0H1(G,L)(X/f)G(P/f)GLG00\to H_{1}(G,L)\to(X/f)_{G}\to(P/f)_{G}\to L_{G}\to 0

and

0H1(G,p/pmp)LGQG/fp/pmp00\to H_{1}(G,\mathbb{Z}_{p}/p^{m}\mathbb{Z}_{p})\to L_{G}\to Q_{G}/f\to\mathbb{Z}_{p}/p^{m}\mathbb{Z}_{p}\to 0

and also an isomorphism H2(G,p/pmp)H1(G,L)H_{2}(G,\mathbb{Z}_{p}/p^{m}\mathbb{Z}_{p})\simeq H_{1}(G,L).

We can combine these observations with the exact sequence obtained by taking modulo ff of sequence (5.2) to construct a diagram

H1(G,p/pmp)\textstyle{H_{1}(G,\mathbb{Z}_{p}/p^{m}\mathbb{Z}_{p})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H2(G,p/pmp)\textstyle{H_{2}(G,\mathbb{Z}_{p}/p^{m}\mathbb{Z}_{p})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(X/f)G\textstyle{(X/f)_{G}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(P/f)G\textstyle{(P/f)_{G}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\simeq}LG\textstyle{L_{G}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}X(G)/f\textstyle{X_{(G)}/f\ignorespaces\ignorespaces\ignorespaces\ignorespaces}PG/f\textstyle{P_{G}/f\ignorespaces\ignorespaces\ignorespaces\ignorespaces}QG/f\textstyle{Q_{G}/f\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p/pmp\textstyle{\mathbb{Z}_{p}/p^{m}\mathbb{Z}_{p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}p/pmp\textstyle{\mathbb{Z}_{p}/p^{m}\mathbb{Z}_{p}}

This is a commutative diagram of finitely generated torsion 𝒪f\mathcal{O}_{f}-modules. By applying the snake lemma, we obtain the proposition. ∎

Proof of Theorem 5.2(1).

In Proposition 5.5, we take ff so that f0f\in\mathcal{F}_{0}, i.e., 𝒪f\mathcal{O}_{f} is a DVR. Then the injective homomorphism from H2(G,p/pmp)H_{2}(G,\mathbb{Z}_{p}/p^{m}\mathbb{Z}_{p}) to (X/f)G(X/f)_{G} in Proposition 5.5 implies

genΛ[G](X)=gen𝒪f((X/f)G)gen𝒪f(H2(G,p/pmp)),\operatorname{gen}_{\Lambda[G]}(X)=\operatorname{gen}_{\mathcal{O}_{f}}((X/f)_{G})\geq\operatorname{gen}_{\mathcal{O}_{f}}(H_{2}(G,\mathbb{Z}_{p}/p^{m}\mathbb{Z}_{p})),

where the first equality follows from Nakayama’s lemma. Since H2(G,p/pmp)H_{2}(G,\mathbb{Z}_{p}/p^{m}\mathbb{Z}_{p}) is annihilated by TT and 𝒪f/(T)p/pmp\mathcal{O}_{f}/(T)\simeq\mathbb{Z}_{p}/p^{m}\mathbb{Z}_{p}, we have

gen𝒪f(H2(G,p/pmp))=genp/pmp(H2(G,p/pmp))=dim𝔽p(H2(G,𝔽p))=s2,\operatorname{gen}_{\mathcal{O}_{f}}(H_{2}(G,\mathbb{Z}_{p}/p^{m}\mathbb{Z}_{p}))=\operatorname{gen}_{\mathbb{Z}_{p}/p^{m}\mathbb{Z}_{p}}(H_{2}(G,\mathbb{Z}_{p}/p^{m}\mathbb{Z}_{p}))=\dim_{\mathbb{F}_{p}}(H_{2}(G,\mathbb{F}_{p}))=s_{2},

where the second equality follows from Lemma 2.7. Combining these formulas, we obtain genΛ[G](X)s2\operatorname{gen}_{\Lambda[G]}(X)\geq s_{2}, as claimed. ∎

Remark 5.6.

This argument also shows

max{s2,gen𝒪f(Ker(π))}genΛ[G](X)s2+gen𝒪f(Ker(π))\max\left\{s_{2},\operatorname{gen}_{\mathcal{O}_{f}}(\operatorname{Ker}(\pi))\right\}\leq\operatorname{gen}_{\Lambda[G]}(X)\leq s_{2}+\operatorname{gen}_{\mathcal{O}_{f}}(\operatorname{Ker}(\pi))

with π\pi as in Proposition 5.5. Then the full statement of Theorem 3.3(1) follows if we can take ff so that gen𝒪f(Ker(π))\operatorname{gen}_{\mathcal{O}_{f}}(\operatorname{Ker}(\pi)) coincides with the tt in Theorem 3.3(1). This is indeed possible, but in §6 we will give a more direct proof for the rest of Theorem 3.3(1) instead.

5.4. Proof of Theorem 5.2(2)

We introduce several lemmas. We abbreviate rnΛ[G]()r_{n}^{\Lambda[G]}(-) as rn()r_{n}(-) and we never omit the coefficient ring otherwise.

Lemma 5.7.

Let PP be a finitely generated torsion Λ[G]\Lambda[G]-module whose projective dimension is 1\leq 1. Then we have rn(P)=0r_{n}(P)=0 for n2n\geq 2 and r1(P)=r0(P)r_{1}(P)=r_{0}(P).

Proof.

By the assumption on PP, there exists a presentation of PP of the form 0Λ[G]aΛ[G]aP00\to\Lambda[G]^{a}\to\Lambda[G]^{a}\to P\to 0. The lemma follows immediately from this. ∎

Lemma 5.8.

Let 0MPM00\to M^{\prime}\to P\to M\to 0 be a short exact sequence of finitely generated torsion Λ[G]\Lambda[G]-modules such that the projective dimension of PP is 1\leq 1. Then we have

rn(M)=rn+1(M)r_{n}(M^{\prime})=r_{n+1}(M)

for n2n\geq 2 and

r1(M)r0(M)=r2(M)r1(M)+r0(M).r_{1}(M^{\prime})-r_{0}(M^{\prime})=r_{2}(M)-r_{1}(M)+r_{0}(M).
Proof.

This follows immediately from the long exact sequence of TorΛ[G](𝔽p,)\operatorname{Tor}_{*}^{\Lambda[G]}(\mathbb{F}_{p},-) applied to the given sequence, taking Lemma 5.7 into account. ∎

Lemma 5.9.

Let MM be a finitely generated p[G]\mathbb{Z}_{p}[G]-module that is free over p\mathbb{Z}_{p}. We regard MM as a Λ[G]\Lambda[G]-module so that TT acts trivially on MM. Then we have

rnΛ[G](M)=rnp[G](M)+rn1p[G](M)r_{n}^{\Lambda[G]}(M)=r_{n}^{\mathbb{Z}_{p}[G]}(M)+r_{n-1}^{\mathbb{Z}_{p}[G]}(M)

for n0n\geq 0. Here, we set r1p[G](M)=0r_{-1}^{\mathbb{Z}_{p}[G]}(M)=0.

Proof.

Let us take a minimal resolution of MM as a p[G]\mathbb{Z}_{p}[G]-module

p[G]ρ2p[G]ρ1p[G]ρ0M0,\cdots\to\mathbb{Z}_{p}[G]^{\rho_{2}}\to\mathbb{Z}_{p}[G]^{\rho_{1}}\to\mathbb{Z}_{p}[G]^{\rho_{0}}\to M\to 0,

where we put ρn=rnp[G](M)\rho_{n}=r_{n}^{\mathbb{Z}_{p}[G]}(M). We have an exact sequence

0Λ×TΛp0,0\to\Lambda\overset{\times T}{\to}\Lambda\to\mathbb{Z}_{p}\to 0,

which may be regarded as a minimal resolution of p\mathbb{Z}_{p} as a Λ\Lambda-module. Then we take the tensor product over p\mathbb{Z}_{p} of the two complex above (omitting MM and p\mathbb{Z}_{p} respectively). As a result, we obtain an exact sequence

Λ[G]ρ2+ρ1Λ[G]ρ1+ρ0Λ[G]ρ0M0.\cdots\to\Lambda[G]^{\rho_{2}+\rho_{1}}\to\Lambda[G]^{\rho_{1}+\rho_{0}}\to\Lambda[G]^{\rho_{0}}\to M\to 0.

By construction, this is a minimal resolution of MM as a Λ[G]\Lambda[G]-module. This completes the proof of the lemma. ∎

Now we are ready to prove Theorem 5.2(2).

Proof of Theorem 5.2(2).

First we recall sn=rnp[G](p)s_{n}=r_{n}^{\mathbb{Z}_{p}[G]}(\mathbb{Z}_{p}) by Lemma 2.5. By applying Lemma 5.9 to M=pM=\mathbb{Z}_{p}, we obtain

rn(p)=rnΛ[G](p)=sn+sn1r_{n}(\mathbb{Z}_{p})=r_{n}^{\Lambda[G]}(\mathbb{Z}_{p})=s_{n}+s_{n-1}

for n0n\geq 0, where we set s1=0s_{-1}=0. We apply Lemma 5.8 to the two short exact sequences obtained by splitting the Tate sequence. As a consequence, we obtain

rn(X)=rn+2(p)=sn+2+sn+1r_{n}(X)=r_{n+2}(\mathbb{Z}_{p})=s_{n+2}+s_{n+1}

for n2n\geq 2 and

(5.3) r1(X)r0(X)\displaystyle r_{1}(X)-r_{0}(X) =r3(p)r2(p)+r1(p)r0(p)\displaystyle=r_{3}(\mathbb{Z}_{p})-r_{2}(\mathbb{Z}_{p})+r_{1}(\mathbb{Z}_{p})-r_{0}(\mathbb{Z}_{p})
(5.4) =(s3+s2)(s2+s1)+(s1+s0)(s0+s1)\displaystyle=(s_{3}+s_{2})-(s_{2}+s_{1})+(s_{1}+s_{0})-(s_{0}+s_{-1})
(5.5) =s3.\displaystyle=s_{3}.

This completes the proof. ∎

5.5. An algebraic proposition

This subsection provides preliminaries to the proof of Theorem 5.3. Let 𝒞\mathcal{C} be the category of finitely generated torsion Λ[G]\Lambda[G]-modules whose projective dimension over Λ\Lambda is 1\leq 1, that is, those that do not have nontrivial finite submodules.

We also write 𝒫\mathcal{P} for the subcategory of 𝒞\mathcal{C} that consists of modules whose projective dimension over Λ[G]\Lambda[G] is 1\leq 1.

For a module M𝒞M\in\mathcal{C}, it is known that the dual

M=ExtΛ[G]1(M,Λ[G])ExtΛ1(M,Λ)M^{*}=\operatorname{Ext}^{1}_{\Lambda[G]}(M,\Lambda[G])\simeq\operatorname{Ext}^{1}_{\Lambda}(M,\Lambda)

is also in 𝒞\mathcal{C} and (M)M(M^{*})^{*}\simeq M ([13, Propositions 5.5.3 (ii) and 5.5.8 (iv)]). Moreover, if P𝒫P\in\mathcal{P}, we have P𝒫P^{*}\in\mathcal{P}. These facts are also explained in [12, §3.1].

In this subsection, we prove the following proposition.

Proposition 5.10.

Let d0d\geq 0 be an integer. Let us consider exact sequences

0NP1PdM00\to N\to P_{1}\to\cdots\to P_{d}\to M\to 0

and

0NP1PdM00\to N^{\prime}\to P_{1}^{\prime}\to\cdots\to P_{d}^{\prime}\to M^{\prime}\to 0

in 𝒞\mathcal{C} such that Pi,Pi𝒫P_{i},P_{i}^{\prime}\in\mathcal{P} for 1id1\leq i\leq d. Then the following hold.

  • (1)

    If MMM\simeq M^{\prime}, then we have

    rn(N)=rn(N)r_{n}(N)=r_{n}(N^{\prime})

    for n2n\geq 2 and

    r1(N)r0(N)=r1(N)r0(N).r_{1}(N)-r_{0}(N)=r_{1}(N^{\prime})-r_{0}(N^{\prime}).
  • (2)

    Similarly, if NNN\simeq N^{\prime}, then we have

    rn(M)=rn(M)r_{n}(M)=r_{n}(M^{\prime})

    for n2n\geq 2 and

    r1(M)r0(M)=r1(M)r0(M).r_{1}(M)-r_{0}(M)=r_{1}(M^{\prime})-r_{0}(M^{\prime}).
Remark 5.11.

It is easy to deduce claim (1) from Lemma 5.8. Indeed, we have

rn(N)=rn+d(M)r_{n}(N)=r_{n+d}(M)

for n2n\geq 2 and

r1(N)r0(N)=i=0d+1(1)ird+1i(M).r_{1}(N)-r_{0}(N)=\sum_{i=0}^{d+1}(-1)^{i}r_{d+1-i}(M).

On the other hand, claim (2) cannot be deduced from Lemma 5.8. Roughly speaking, claims (1) and (2) are respectively what we need to prove Theorems 5.2 and 5.3.

To prove Proposition 5.10, it is convenient to use the concept of axiomatic Fitting invariants introduced by the first author [12]. More concretely, inspired by [4, §3.2], we use the notion of 𝒫\mathcal{P}-trivial Fitting invariant defined as follows.

Definition 5.12.

A 𝒫\mathcal{P}-trivial Fitting invariant is a map :𝒞Ω\mathcal{F}:\mathcal{C}\to\Omega, where Ω\Omega is a commutative monoid, satisfying the following properties:

  • (i)

    If P𝒫P\in\mathcal{P}, we have (P)\mathcal{F}(P) is the identity element of Ω\Omega.

  • (ii)

    For a short exact sequence 0MMP00\to M^{\prime}\to M\to P\to 0 in 𝒞\mathcal{C} with P𝒫P\in\mathcal{P}, we have (M)=(M)\mathcal{F}(M^{\prime})=\mathcal{F}(M).

  • (iii)

    For a short exact sequence 0PMM00\to P\to M\to M^{\prime}\to 0 in 𝒞\mathcal{C} with P𝒫P\in\mathcal{P}, we have (M)=(M)\mathcal{F}(M^{\prime})=\mathcal{F}(M).

It is an important fact [12, Proposition 3.17] that conditions (ii) and (iii) are equivalent to each other (assuming (i)). Note that in this setting we do not have to assume Ω\Omega is a commutative monoid, and instead a pointed set structure suffices.

A fundamental example of a Fitting invariant is of course given by the Fitting ideal; more precisely, the Fitting ideal modulo principal ideals satisfies the axioms of 𝒫\mathcal{P}-trivial Fitting invariants.

The following proposition introduces another kind of Fitting invariants.

Proposition 5.13.

For n2n\geq 2, define n:𝒞\mathcal{F}_{n}:\mathcal{C}\to\mathbb{N} by n(M)=rn(M)\mathcal{F}_{n}(M)=r_{n}(M). We also define 0,1:𝒞\mathcal{F}_{0,1}:\mathcal{C}\to\mathbb{Z} by 0,1(M)=r1(M)r0(M)\mathcal{F}_{0,1}(M)=r_{1}(M)-r_{0}(M). Then these maps n\mathcal{F}_{n} and 0,1\mathcal{F}_{0,1} are all 𝒫\mathcal{P}-trivial Fitting invariants.

Proof.

We check the conditions (i) and (ii). Firstly, (i) is a restatement of Lemma 5.7. Secondly, (ii) follows from the associated long exact sequence, taking Lemma 5.7 into account again. Indeed, the long exact sequence collapses into isomorphisms for degree 2\geq 2 and a 66-term exact sequence for degrees 0,10,1. ∎

Note that (iii) cannot be shown in a similar manner. This is because the lower degree part of the associated long exact sequence becomes an 88-term exact sequence. It is important that (iii) follows from (i) and (ii).

Proof of Proposition 5.10.

By Proposition 5.13, it is enough to show (N)=(N)\mathcal{F}(N)=\mathcal{F}(N^{\prime}) (resp. (M)=(M)\mathcal{F}(M)=\mathcal{F}(M^{\prime})) if MMM\simeq M^{\prime} (resp. NNN\simeq N^{\prime}) for any 𝒫\mathcal{P}-trivial Fitting invariant \mathcal{F}. For this, we apply the theory of shifts d()\mathcal{F}^{\langle d\rangle}(-), d()\mathcal{F}^{\langle-d\rangle}(-) of Fitting invariants of the first author [12, Theorem 3.19]. By the exact sequence involving MM, NN, and PiP_{i}, the definition of the shifts implies

d(M)=(N),d(N)=(M),\mathcal{F}^{\langle d\rangle}(M)=\mathcal{F}(N),\quad\mathcal{F}^{\langle-d\rangle}(N)=\mathcal{F}(M),

and similarly for MM^{\prime}, NN^{\prime}. Then what we have to show is just a reformulation of the well-definedness of the shifts, which is already established by the first author in [12, Theorem 3.19]. ∎

5.6. Proof of Theorem 5.3

Proof of Theorem 5.3.

(1) By taking the dual of the Tate sequence, we obtain an exact sequence

(5.6) 0pQPX0,0\to\mathbb{Z}_{p}\to Q^{*}\to P^{*}\to X^{*}\to 0,

where we used pp\mathbb{Z}_{p}^{*}\simeq\mathbb{Z}_{p}.

As in §5.3, we use the specialization method. Let us take any element f0f\in\mathcal{F}_{0} that is coprime to charΛ(P)\operatorname{char}_{\Lambda}(P) and charΛ(Q)\operatorname{char}_{\Lambda}(Q). Put m=ordp(f(0))1m=\operatorname{ord}_{p}(f(0))\geq 1. Then (5.6) yields an exact sequence

0p/pmpQ/fP/fX/f0.0\to\mathbb{Z}_{p}/p^{m}\mathbb{Z}_{p}\to Q^{*}/f\to P^{*}/f\to X^{*}/f\to 0.

Observe that both P/fP^{*}/f and Q/fQ^{*}/f are GG-cohomologically trivial. So we have

H^1(G,X/f)H1(G,p/pmp),\hat{H}^{-1}(G,X^{*}/f)\simeq H^{1}(G,\mathbb{Z}_{p}/p^{m}\mathbb{Z}_{p}),

where H^1\hat{H}^{-1} denotes the Tate cohomology group. By definition, H^1(G,X/f)\hat{H}^{-1}(G,X^{*}/f) is a submodule of H0(G,X/f)H_{0}(G,X^{*}/f), so the above isomorphism shows

gen𝒪f(H0(G,X/f))gen𝒪f(H1(G,p/pmp))\operatorname{gen}_{\mathcal{O}_{f}}(H_{0}(G,X^{*}/f))\geq\operatorname{gen}_{\mathcal{O}_{f}}(H^{1}(G,\mathbb{Z}_{p}/p^{m}\mathbb{Z}_{p}))

as 𝒪f\mathcal{O}_{f} is a DVR. By Nakayama’s lemma, the left hand side is equal to genΛ[G](X)\operatorname{gen}_{\Lambda[G]}(X^{*}). Also, as in the proof of Theorem 5.2(1) in §5.3,

(5.7) gen𝒪f(H1(G,p/pmp))\displaystyle\operatorname{gen}_{\mathcal{O}_{f}}(H^{1}(G,\mathbb{Z}_{p}/p^{m}\mathbb{Z}_{p})) =genp/pmp(H1(G,p/pmp))\displaystyle=\operatorname{gen}_{\mathbb{Z}_{p}/p^{m}\mathbb{Z}_{p}}(H^{1}(G,\mathbb{Z}_{p}/p^{m}\mathbb{Z}_{p}))
(5.8) =genp/pmp(H1(G,p/pmp))\displaystyle=\operatorname{gen}_{\mathbb{Z}_{p}/p^{m}\mathbb{Z}_{p}}(H_{1}(G,\mathbb{Z}_{p}/p^{m}\mathbb{Z}_{p}))
(5.9) =dim𝔽p(H1(G,𝔽p))\displaystyle=\dim_{\mathbb{F}_{p}}(H_{1}(G,\mathbb{F}_{p}))
(5.10) =s1,\displaystyle=s_{1},

where the second equality follows from Lemma 2.9 and the third from Lemma 2.7. Thus we obtain (1).

(2) In case GG is trivial, since the projective dimension of XX as a Λ\Lambda-module is 1\leq 1, we may apply Lemma 5.7 to obtain the assertion.

From now on, we assume GG is non-trivial. Since sn=rnp[G](p)s_{n}=r_{n}^{\mathbb{Z}_{p}[G]}(\mathbb{Z}_{p}) by Lemma 2.5, a minimal resolution of p\mathbb{Z}_{p} as a p[G]\mathbb{Z}_{p}[G]-module is of the form

(5.11) p[G]s2d2p[G]s1d1p[G]s0𝜀p0.\cdots\to\mathbb{Z}_{p}[G]^{s_{2}}\overset{d_{2}}{\to}\mathbb{Z}_{p}[G]^{s_{1}}\overset{d_{1}}{\to}\mathbb{Z}_{p}[G]^{s_{0}}\overset{\varepsilon}{\to}\mathbb{Z}_{p}\to 0.

We truncate it to an exact sequence

0Ker(d1)p[G]s1d1p[G]s0𝜀p0.0\to\operatorname{Ker}(d_{1})\to\mathbb{Z}_{p}[G]^{s_{1}}\overset{d_{1}}{\to}\mathbb{Z}_{p}[G]^{s_{0}}\overset{\varepsilon}{\to}\mathbb{Z}_{p}\to 0.

Since its dual ()(-)^{*} is also exact and we have pp\mathbb{Z}_{p}^{*}\simeq\mathbb{Z}_{p} and p[G]p[G]\mathbb{Z}_{p}[G]^{*}\simeq\mathbb{Z}_{p}[G], we obtain an exact sequence

(5.12) 0pεp[G]s0d1p[G]s1Ker(d1)0.0\to\mathbb{Z}_{p}\overset{\varepsilon^{*}}{\to}\mathbb{Z}_{p}[G]^{s_{0}}\overset{d_{1}^{*}}{\to}\mathbb{Z}_{p}[G]^{s_{1}}\to\operatorname{Ker}(d_{1})^{*}\to 0.

By comparing (5.6) and (5.12), Proposition 5.10(2) implies

r1(X)r0(X)=r1(Ker(d1))r0(Ker(d1))r_{1}(X^{*})-r_{0}(X^{*})=r_{1}(\operatorname{Ker}(d_{1})^{*})-r_{0}(\operatorname{Ker}(d_{1})^{*})

and

rn(X)=rn(Ker(d1))r_{n}(X^{*})=r_{n}(\operatorname{Ker}(d_{1})^{*})

for n2n\geq 2.

Let us compute rn(Ker(d1))r_{n}(\operatorname{Ker}(d_{1})^{*}) for any n0n\geq 0. We combine (5.12) with (5.11) to an exact sequence

p[G]s2d2p[G]s1d1p[G]s0εεp[G]s0d1p[G]s1Ker(d1)0.\cdots\to\mathbb{Z}_{p}[G]^{s_{2}}\overset{d_{2}}{\to}\mathbb{Z}_{p}[G]^{s_{1}}\overset{d_{1}}{\to}\mathbb{Z}_{p}[G]^{s_{0}}\overset{\varepsilon^{*}\circ\varepsilon}{\to}\mathbb{Z}_{p}[G]^{s_{0}}\overset{d_{1}^{*}}{\to}\mathbb{Z}_{p}[G]^{s_{1}}\to\operatorname{Ker}(d_{1})^{*}\to 0.

By construction, this is a minimal resolution of Ker(d1)\operatorname{Ker}(d_{1})^{*} as a p[G]\mathbb{Z}_{p}[G]-module. For this we need the hypothesis that GG is non-trivial; the map εε\varepsilon^{*}\circ\varepsilon can be identified with the map p[G]p[G]\mathbb{Z}_{p}[G]\to\mathbb{Z}_{p}[G] that sends 11 to the norm element NG=gGgN_{G}=\sum_{g\in G}g, and NGN_{G} is in the Jacobson radical of p[G]\mathbb{Z}_{p}[G] if and only if GG is non-trivial. Therefore, we obtain

rnp[G](Ker(d1))={s1(n=0)s0(n=1,2)sn2(n3).r_{n}^{\mathbb{Z}_{p}[G]}(\operatorname{Ker}(d_{1})^{*})=\begin{cases}s_{1}&(n=0)\\ s_{0}&(n=1,2)\\ s_{n-2}&(n\geq 3).\end{cases}

By applying Lemma 5.9, we obtain

rnΛ[G](Ker(d1))={s1(n=0)s1+s0(n=1)s0+s0(n=2)sn2+sn3(n3).r_{n}^{\Lambda[G]}(\operatorname{Ker}(d_{1})^{*})=\begin{cases}s_{1}&(n=0)\\ s_{1}+s_{0}&(n=1)\\ s_{0}+s_{0}&(n=2)\\ s_{n-2}+s_{n-3}&(n\geq 3).\end{cases}

This completes the proof. ∎

6. Proof of the rest of Theorems 3.3(1) and 3.4(1)

Now we return to the arithmetic situation described in §3.1.

6.1. Proof of Theorem 3.4(1)

First we recall the following well-known fact.

Lemma 6.1.

We have an isomorphism

(Xk,S)Gal(k/k)Gal(Mk,S/k).(X_{k_{\infty},S})_{\operatorname{Gal}(k_{\infty}/k)}\simeq\operatorname{Gal}(M_{k,S}/k_{\infty}).

In particular, by Nakayama’s lemma we have

genΛ(Xk,S)=rankpGal(Mk,S/k).\operatorname{gen}_{\Lambda}(X_{k_{\infty},S})=\operatorname{rank}_{p}\operatorname{Gal}(M_{k,S}/k_{\infty}).

To prove Theorem 3.4(1), we also need the following general lemma.

Lemma 6.2.

As in §5, consider Λ=p[[T]]\Lambda=\mathbb{Z}_{p}[[T]] and a finite pp-group GG. For a finitely generated torsion Λ[G]\Lambda[G]-module MM whose projective dimension is 1\leq 1, we have

genΛ[G](M)=genΛ[G](M).\operatorname{gen}_{\Lambda[G]}(M^{*})=\operatorname{gen}_{\Lambda[G]}(M).
Proof.

As in Lemma 5.7, the minimal resolution of MM is of the form 0Λ[G]aΛ[G]aM00\to\Lambda[G]^{a}\to\Lambda[G]^{a}\to M\to 0. By taking Ext\operatorname{Ext}-functor, this induces an exact sequence 0Λ[G]aΛ[G]aM00\to\Lambda[G]^{a}\to\Lambda[G]^{a}\to M^{*}\to 0, which is again a minimal resolution. Thus we obtain the lemma. ∎

Proof of Theorem 3.4(1).

Applying the dual of the Tate sequence introduced in (5.6) to our setting, we obtain an exact sequence

0pQPXK,S0.0\to\mathbb{Z}_{p}^{*}\to Q^{*}\to P^{*}\to X_{K_{\infty},S}^{*}\to 0.

By the compatibility of the Tate sequences, we obtain an isomorphism

(XK,S)GXk,S.(X_{K_{\infty},S}^{*})_{G}\simeq X_{k_{\infty},S}^{*}.

Therefore, we have

gen(XK,S)=genΛ(Xk,S).\operatorname{gen}_{\mathcal{R}}(X_{K_{\infty},S}^{*})=\operatorname{gen}_{\Lambda}(X_{k_{\infty},S}^{*}).

By Lemma 6.2 (with GG trivial), this is then equal to genΛ(Xk,S)\operatorname{gen}_{\Lambda}(X_{k_{\infty},S}). By Lemma 6.1, this completes the proof. ∎

6.2. Proof of Theorem 3.3(1)

Note that applying Theorem 5.2(1) to the Tate sequence introduced in Theorem 4.1, we already obtained the inequality s2gen(XK,S)s_{2}\leq\operatorname{gen}_{\mathcal{R}}(X_{K_{\infty},S}). Therefore, it remains only to prove tgen(XK,S)s2+tt\leq\operatorname{gen}_{\mathcal{R}}(X_{K_{\infty},S})\leq s_{2}+t.

Lemma 6.3.

We have an exact sequence

0H2(G,)(XK,S)GXk,SGab00\to H_{2}(G,\mathbb{Z})\to(X_{K_{\infty},S})_{G}\to X_{k_{\infty},S}\to G^{\operatorname{ab}}\to 0

of Λ\Lambda-modules.

Proof.

By the Hochschild–Serre spectral sequence we have an exact sequence

0H1(G,p/p)H1(𝒪K,S,p/p)H1(𝒪k,S,p/p)GH2(G,p/p)0,0\to H^{1}(G,\mathbb{Q}_{p}/\mathbb{Z}_{p})\to H^{1}({\mathcal{O}}_{K_{\infty},S},\mathbb{Q}_{p}/\mathbb{Z}_{p})\to H^{1}({\mathcal{O}}_{k_{\infty},S},\mathbb{Q}_{p}/\mathbb{Z}_{p})^{G}\to H^{2}(G,\mathbb{Q}_{p}/\mathbb{Z}_{p})\to 0,

where we used the weak Leopoldt conjecture H2(𝒪K,S,p/p)=0H^{2}({\mathcal{O}}_{K_{\infty},S},\mathbb{Q}_{p}/\mathbb{Z}_{p})=0. For i=1,2i=1,2, the Pontryagin dual of Hi(G,p/p)H^{i}(G,\mathbb{Q}_{p}/\mathbb{Z}_{p}) is the projective limit of Hi(G,/pm)H_{i}(G,\mathbb{Z}/p^{m}) by Lemma 2.9. But since GG is a finite pp-group, it is finite and isomorphic to Hi(G,)H_{i}(G,\mathbb{Z}). Therefore, taking the Pontryagin dual of the above exact sequence and using Lemma 2.6, we get the conclusion. ∎

Note that Gab=Gal(Mk,SK/k)G^{\operatorname{ab}}=\operatorname{Gal}(M_{k_{\infty},S}\cap K_{\infty}/k_{\infty}). Then putting Y=Gal(Mk,S/Mk,SK)Y=\operatorname{Gal}(M_{k_{\infty},S}/M_{k_{\infty},S}\cap K_{\infty}), by Lemma 6.3 we obtain two exact sequences

0H2(G,)(XK,S)GY0,0YXk,SGab0.0\to H_{2}(G,\mathbb{Z})\to(X_{K_{\infty},S})_{G}\to Y\to 0,\quad 0\to Y\to X_{k_{\infty},S}\to G^{\operatorname{ab}}\to 0.

The first sequence implies

(6.1) genΛ(Y)gen(XK,S)genp(H2(G,))+genΛ(Y).\operatorname{gen}_{\Lambda}(Y)\leq\operatorname{gen}_{\mathcal{R}}(X_{K_{\infty},S})\leq\operatorname{gen}_{\mathbb{Z}_{p}}(H_{2}(G,\mathbb{Z}))+\operatorname{gen}_{\Lambda}(Y).

On the other hand, by taking the Γ=Gal(k/k)\Gamma=\operatorname{Gal}(k_{\infty}/k)-coinvariant of the second sequence, we obtain an exact sequence

GabYΓ(Xk,S)ΓGab0.G^{\operatorname{ab}}\to Y_{\Gamma}\to(X_{k_{\infty},S})_{\Gamma}\to G^{\operatorname{ab}}\to 0.

By Lemma 6.1, this is reformulated as

GabYΓGal(Mk,S/Mk,SK)0.G^{\operatorname{ab}}\to Y_{\Gamma}\to\operatorname{Gal}(M_{k,S}/M_{k_{\infty},S}\cap K_{\infty})\to 0.

This sequence, together with Mk,SK=Mk,SKM_{k_{\infty},S}\cap K_{\infty}=M_{k,S}\cap K_{\infty} and the definition of tt, implies that

(6.2) tgenΛ(Y)genp(Gab)+t.t\leq\operatorname{gen}_{\Lambda}(Y)\leq\operatorname{gen}_{\mathbb{Z}_{p}}(G^{\operatorname{ab}})+t.

By combining (6.1) and (6.2), we obtain

tgen(XK,S)genp(H2(G,))+genp(Gab)+t=s2+t,t\leq\operatorname{gen}_{\mathcal{R}}(X_{K_{\infty},S})\leq\operatorname{gen}_{\mathbb{Z}_{p}}(H_{2}(G,\mathbb{Z}))+\operatorname{gen}_{\mathbb{Z}_{p}}(G^{\operatorname{ab}})+t=s_{2}+t,

where we used Lemma 2.7 to get the final equality. Note that if GG is abelian, we have H2(G,)2GH_{2}(G,\mathbb{Z})\simeq\bigwedge^{2}G and genp(2G)=s(s1)/2\operatorname{gen}_{\mathbb{Z}_{p}}(\bigwedge^{2}G)=s(s-1)/2 (with s=rankpGs=\operatorname{rank}_{p}G), which imply explicitly

s+genp(H2(G,))=s+s(s1)/2=s(s+1)/2=s2.s+\operatorname{gen}_{\mathbb{Z}_{p}}(H_{2}(G,\mathbb{Z}))=s+s(s-1)/2=s(s+1)/2=s_{2}.

This completes the proof of Theorem 3.3(1).

7. Numerical Examples

In this section, we numerically check the inequality concerning gen(XK,S)\operatorname{gen}_{\mathcal{R}}(X_{K_{\infty},S}) in Theorem 1.1 by using the computer package PARI/GP. We consider k=k=\mathbb{Q} and its finite abelian pp-extension KK that is totally real. Let SS be a finite set of places of \mathbb{Q} containing pp and the archimedean place \infty, such that K/K/\mathbb{Q} is unramified outside SS.

The basic method is as follows. First, by Nakayama’s lemma and Lemma 6.1, we have

genp[[Gal(K/)]](XK,S)=genp((XK,S)Gal(K/))=genp(Gal(MK,S/K)Gal(K/)).\operatorname{gen}_{\mathbb{Z}_{p}[[\operatorname{Gal}(K_{\infty}/\mathbb{Q})]]}(X_{K_{\infty},S})=\operatorname{gen}_{\mathbb{Z}_{p}}((X_{K_{\infty},S})_{\operatorname{Gal}(K_{\infty}/\mathbb{Q})})=\operatorname{gen}_{\mathbb{Z}_{p}}(\operatorname{Gal}(M_{K,S}/K_{\infty})_{\operatorname{Gal}(K/\mathbb{Q})}).

We observe that MK,SM_{K,S} is the union of the maximal pp-extensions of KK in the ray class fields of modulus pmvS{p}vp^{m}\prod_{v\in S\setminus\{p\}}v for all m0m\geq 0. Note here that, since KK is abelian over \mathbb{Q}, the Leopoldt conjecture is shown to be true for KK by work of Brumer (see [13, Theorem 10.3.16]). Therefore, Gal(MK,S/K)\operatorname{Gal}(M_{K,S}/K_{\infty}) is finite, and we can compute it by computing the ray class groups for finitely many mm. In this way we can determine the quantity genp(Gal(MK,S/K)Gal(K/))\operatorname{gen}_{\mathbb{Z}_{p}}(\operatorname{Gal}(M_{K,S}/K_{\infty})_{\operatorname{Gal}(K/\mathbb{Q})}).

7.1. The case p=3p=3

Let us take p=3p=3, though the discussion is basically valid for any odd prime pp. We write S{p,}={1,,s}S\setminus\{p,\infty\}=\{\ell_{1},\dots,\ell_{s}\}. By the theorem of Kronecker–Weber, the Galois group of the maximal abelian extension of \mathbb{Q} unramified outside SS is isomorphic to p××i=1si×\mathbb{Z}_{p}^{\times}\times\prod_{i=1}^{s}\mathbb{Z}_{\ell_{i}}^{\times}, so we have

Gal(M,S/)p×i=1sp/(i1)p.\operatorname{Gal}(M_{\mathbb{Q},S}/\mathbb{Q})\simeq\mathbb{Z}_{p}\times\prod_{i=1}^{s}\mathbb{Z}_{p}/(\ell_{i}-1)\mathbb{Z}_{p}.

As is well-known, we may assume i1(modp)\ell_{i}\equiv 1(\bmod\ p) for 1is1\leq i\leq s without loss of generality. For such an SS, let us take KK as the unique intermediate field of (μ1,,μs)/\mathbb{Q}(\mu_{\ell_{1}},\dots,\mu_{\ell_{s}})/\mathbb{Q} such that

Gal(K/)(/p)s.\operatorname{Gal}(K/\mathbb{Q})\simeq(\mathbb{Z}/p\mathbb{Z})^{s}.

In this case, the above information on M,SM_{\mathbb{Q},S} implies that

Gal(M,S/K)i=1spp/(i1)p,\operatorname{Gal}(M_{\mathbb{Q},S}/K_{\infty})\simeq\prod_{i=1}^{s}p\mathbb{Z}_{p}/(\ell_{i}-1)\mathbb{Z}_{p},

so the integer tt in Theorem 1.1 is determined as t=#{1isi1(modp2)}t=\#\{1\leq i\leq s\mid\ell_{i}\equiv 1(\bmod\ p^{2})\}. To ease the notation, we write gen(X)=genp[[Gal(K/)]](XK,S)\operatorname{gen}(X)=\operatorname{gen}_{\mathbb{Z}_{p}[[\operatorname{Gal}(K_{\infty}/\mathbb{Q})]]}(X_{K_{\infty},S}). Then Theorem 1.1 asserts

max{s(s+1)2,t}gen(X)s(s+1)2+t.\max\left\{\frac{s(s+1)}{2},t\right\}\leq\operatorname{gen}(X)\leq\frac{s(s+1)}{2}+t.

Let PP be the set of all prime numbers \ell such that 1(modp)\ell\equiv 1(\bmod\ p). We divide PP into two subsets P1P_{1} and P2P_{2} defined by

P1={P1(modp2)}={7,13,31,43,61,67,79,97,.}P_{1}=\{\ell\in P\mid\ell\not\equiv 1(\bmod\ p^{2})\}=\{7,13,31,43,61,67,79,97,\dots.\}

and

P2={P1(modp2)}={19,37,73,109,127,163,181,199,.}.P_{2}=\{\ell\in P\mid\ell\equiv 1(\bmod\ p^{2})\}=\{19,37,73,109,127,163,181,199,\dots.\}.

Then we have t=#{1isiP2}t=\#\{1\leq i\leq s\mid\ell_{i}\in P_{2}\}.

First we consider the case t=0t=0, that is, 1,,s\ell_{1},\dots,\ell_{s} are ss distinct primes in P1P_{1}. Then Theorem 1.1 asserts gen(X)=s(s+1)/2\operatorname{gen}(X)=s(s+1)/2 as in Remark 1.2(1). By using PARI/GP, we numerically checked gen(X)=1\operatorname{gen}(X)=1 if s=1s=1 and 1100\ell_{1}\leq 100, and gen(X)=3\operatorname{gen}(X)=3 if s=2s=2 and 1,2100\ell_{1},\ell_{2}\leq 100, as Theorem 1.1 says.

Suppose that t=1t=1 and s=1s=1, that is, S={3,1,}S=\{3,\ell_{1},\infty\} with 1P2\ell_{1}\in P_{2}. Then Theorem 1.1 asserts

1gen(X)2.1\leq\operatorname{gen}(X)\leq 2.

By numerical computation, we find only gen(X)=2\operatorname{gen}(X)=2 in the range 1200\ell_{1}\leq 200, and did not encounter gen(X)=1\operatorname{gen}(X)=1.

Next consider the case t=1t=1 and s=2s=2, that is, S={3,1,2,}S=\{3,\ell_{1},\ell_{2},\infty\} with 1P1\ell_{1}\in P_{1} and 2P2\ell_{2}\in P_{2}. Then we have

3gen(X)43\leq\operatorname{gen}(X)\leq 4

by Theorem 1.1. In the range 1<100\ell_{1}<100 and 2<200\ell_{2}<200, we find gen(X)=3\operatorname{gen}(X)=3 except for

(7.1) (1,2)=\displaystyle(\ell_{1},\ell_{2})= (7,127),(7,181),(13,73),(13,109),(13,181),(31,109),(31,163),(43,127),\displaystyle(7,127),(7,181),(13,73),(13,109),(13,181),(31,109),(31,163),(43,127),
(7.2) (43,199),(61,37),(61,163),(67,109),(79,199),(97,19),(97,109),(97,127),\displaystyle(43,199),(61,37),(61,163),(67,109),(79,199),(97,19),(97,109),(97,127),

for which gen(X)=4\operatorname{gen}(X)=4. Thus the above inequality on gen(X)\operatorname{gen}(X) is sharp in this case.

Finally we consider the case t=2t=2 and s=2s=2, that is, S={3,1,2,}S=\{3,\ell_{1},\ell_{2},\infty\} with 1\ell_{1}, 2P2\ell_{2}\in P_{2}. Then Theorem 1.1 says

3gen(X)5.3\leq\operatorname{gen}(X)\leq 5.

By numerical computation, in the range 1<2200\ell_{1}<\ell_{2}\leq 200, we find gen(X)=4\operatorname{gen}(X)=4 except for (1,2)=(109,199)(\ell_{1},\ell_{2})=(109,199), for which gen(X)=5\operatorname{gen}(X)=5. We did not encounter gen(X)=3\operatorname{gen}(X)=3 in this situation.

Due to the limitation of the machine power we could not handle s3s\geq 3 when p=3p=3.

7.2. The case p=2p=2

Let us discuss the case p=2p=2. Suppose that S={2,1,,s,}S=\{2,\ell_{1},\dots,\ell_{s},\infty\} where 1,,s\ell_{1},\dots,\ell_{s} are ss distinct odd prime numbers. We have

Gal(M,S/)2×/2×i=1s2/(i1)2.\operatorname{Gal}(M_{\mathbb{Q},S}/\mathbb{Q})\simeq\mathbb{Z}_{2}\times\mathbb{Z}/2\mathbb{Z}\times\prod_{i=1}^{s}\mathbb{Z}_{2}/(\ell_{i}-1)\mathbb{Z}_{2}.

Let us set KK as K=(1,,s)K=\mathbb{Q}(\sqrt{\ell_{1}},\dots,\sqrt{\ell_{s}}), so we have Gal(K/)(/2)s\operatorname{Gal}(K/\mathbb{Q})\simeq(\mathbb{Z}/2\mathbb{Z})^{s} and

Gal(M,S/K)/2×i=1s22/(i1)2.\operatorname{Gal}(M_{\mathbb{Q},S}/K_{\infty})\simeq\mathbb{Z}/2\mathbb{Z}\times\prod_{i=1}^{s}2\mathbb{Z}_{2}/(\ell_{i}-1)\mathbb{Z}_{2}.

As in the previous subsection, we define PP to be the set of all odd prime numbers, P2P_{2} to be the subset of PP consisting of \ell such that 1(mod 4)\ell\equiv 1(\bmod\ 4), and P1=PP2P_{1}=P\setminus P_{2}. The integer tt in Theorem 1.1 is then t=1+#{1isiP2}t=1+\#\{1\leq i\leq s\mid\ell_{i}\in P_{2}\}.

We first consider the case s=1s=1, so K=(1)K=\mathbb{Q}(\sqrt{\ell_{1}}). If 1\ell_{1} is in P1P_{1}, then t=1t=1 and Theorem 1.1 says 1gen(X)21\leq\operatorname{gen}(X)\leq 2. For 1P1\ell_{1}\in P_{1} less than 100100, we always have gen(X)=2\operatorname{gen}(X)=2. If 1\ell_{1} is in P2P_{2}, then t=2t=2 and Theorem 1.1 says 2gen(X)32\leq\operatorname{gen}(X)\leq 3. For 1P2\ell_{1}\in P_{2} less than 100100, we have gen(X)=2\operatorname{gen}(X)=2 except for 1=73\ell_{1}=73, 8989, 9797, for which we have gen(X)=3\operatorname{gen}(X)=3. So in this case the inequality is sharp.

In the following we focus on the case 1,,s\ell_{1},\dots,\ell_{s} are all in

P2={5,13,17,29,37,41,53,61,73,89,97,.}.P_{2}=\{5,13,17,29,37,41,53,61,73,89,97,\dots.\}.

Then t=s+1t=s+1, and the inequality in Theorem 1.1 becomes

s(s+1)2gen(X)s(s+1)2+s+1\frac{s(s+1)}{2}\leq\operatorname{gen}(X)\leq\frac{s(s+1)}{2}+s+1

for s2s\geq 2.

For s=2s=2, we have

3gen(X)63\leq\operatorname{gen}(X)\leq 6

by Theorem 1.1. By numerical computation, we find gen(X)=4,5,6\operatorname{gen}(X)=4,5,6 in the range 1<2100\ell_{1}<\ell_{2}\leq 100. Concretely, we have gen(X)=6\operatorname{gen}(X)=6 if

(1,2)=(17,89),(41,73),(73,89),(73,97),(89,97),(\ell_{1},\ell_{2})=(17,89),(41,73),(73,89),(73,97),(89,97),

gen(X)=5\operatorname{gen}(X)=5 if

(7.3) (1,2)=\displaystyle(\ell_{1},\ell_{2})= (5,41),(5,89),(13,17),(17,41),(17,53),(17,73),(17,97),(37,41),\displaystyle(5,41),(5,89),(13,17),(17,41),(17,53),(17,73),(17,97),(37,41),
(7.4) (37,73),(41,61),(41,89),(41,97),(53,89),(53,97),(61,73),(61,97),\displaystyle(37,73),(41,61),(41,89),(41,97),(53,89),(53,97),(61,73),(61,97),

and gen(X)=4\operatorname{gen}(X)=4 otherwise.

For s=3s=3, Theorem 1.1 says

6gen(X)10.6\leq\operatorname{gen}(X)\leq 10.

For 1=5\ell_{1}=5 and 5<2<31005<\ell_{2}<\ell_{3}\leq 100, we have gen(X)=7\operatorname{gen}(X)=7 except for

(2,3)=(17,89),(37,41),(41,61),(41,73),(41,89),(53,89),(73,89),(89,97),(\ell_{2},\ell_{3})=(17,89),(37,41),(41,61),(41,73),(41,89),(53,89),(73,89),(89,97),

for which we have gen(X)=8\operatorname{gen}(X)=8. Also, we find examples for gen(X)=9,10\operatorname{gen}(X)=9,10 by taking respectively (1,2,3)=(17,73,89),(73,89,97)(\ell_{1},\ell_{2},\ell_{3})=(17,73,89),(73,89,97).

7.3. A variant

As a final remark, let us briefly discuss a variant that matters only when p=2p=2. So far we always assumed that SS contains all archimedean places, so we studied the narrow class groups. Theoretically this assumption is necessary to use the Tate sequence in Theorem 4.1. However, the numerical computation in this section is possible (and simpler) even if we remove the archimedean places from SS.

Suppose that p=2p=2, k=k=\mathbb{Q}, K=(1,,s)K=\mathbb{Q}(\sqrt{\ell_{1}},\dots,\sqrt{\ell_{s}}) as in the previous subsection, and S=S{}S^{\prime}=S\setminus\{\infty\}. We consider genp[[Gal(K/)]](XK,S)\operatorname{gen}_{\mathbb{Z}_{p}[[\operatorname{Gal}(K_{\infty}/\mathbb{Q})]]}(X_{K_{\infty},S^{\prime}}), which we abbreviate as gen(X)\operatorname{gen}(X^{\prime}). We have a natural surjective homomorphism from XK,SX_{K_{\infty},S} to XK,SX_{K_{\infty},S^{\prime}} whose kernel is a cyclic module (since k=k=\mathbb{Q}). Therefore, we have

gen(X)gen(X)gen(X)+1.\operatorname{gen}(X^{\prime})\leq\operatorname{gen}(X)\leq\operatorname{gen}(X^{\prime})+1.

Still assuming i1(mod 4)\ell_{i}\equiv 1(\bmod\ 4) for any 1is1\leq i\leq s, we find the following numerical examples.

  • When s=1s=1, we find examples for gen(X)=1,2\operatorname{gen}(X^{\prime})=1,2.

  • When s=2s=2, we find examples for gen(X)=3,4,5\operatorname{gen}(X^{\prime})=3,4,5.

  • When s=3s=3, we find examples for gen(X)=6,7,8,9\operatorname{gen}(X^{\prime})=6,7,8,9.

These results suggest that s(s+1)/2gen(X)s(s+1)/2+ss(s+1)/2\leq\operatorname{gen}(X^{\prime})\leq s(s+1)/2+s, but this does not follow directly from Theorem 1.1. The above computations suggest that Theorem 1.1 holds true without assumption that SS contains all archimedean places.

References

  • [1] K. S. Brown. Cohomology of groups, volume 87 of Graduate Texts in Mathematics. Springer-Verlag, New York-Berlin, 1982.
  • [2] D. Burns and M. Flach. Motivic LL-functions and Galois module structures. Math. Ann., 305(1):65–102, 1996.
  • [3] C. Greither. Tate sequences and lower bounds for ranks of class groups. Acta Arithmetica, 160(1):55–66, 2013.
  • [4] C. Greither and T. Kataoka. Fitting ideals and various notions of equivalence for modules. Manuscripta Math., 173(1-2):259–291, 2024.
  • [5] C. Greither and M. Kurihara. Tate sequences and Fitting ideals of Iwasawa modules. Algebra i Analiz, 27(6):117–149, 2015.
  • [6] C. Greither and M. Kurihara. Fitting ideals of Iwasawa modules and of the dual of class groups. Tokyo J. Math., 39(3):619–642, 2017.
  • [7] C. Greither, M. Kurihara, and H. Tokio. The second syzygy of the trivial GG-module, and an equivariant main conjecture. In Development of Iwasawa theory—the centennial of K. Iwasawa’s birth, volume 86 of Adv. Stud. Pure Math., pages 317–349. Math. Soc. Japan, Tokyo, 2020.
  • [8] I. M. Isaacs. Algebra: a graduate course, volume 100 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2009. Reprint of the 1994 original.
  • [9] K. Iwasawa. On Γ\Gamma-extensions of algebraic number fields. Bull. Amer. Math. Soc., 65:183–226, 1959.
  • [10] K. Iwasawa. On 𝐙l{\bf Z}_{l}-extensions of algebraic number fields. Ann. of Math. (2), 98:246–326, 1973.
  • [11] G. Karpilovsky. The Schur multiplier, volume 2 of London Mathematical Society Monographs. New Series. The Clarendon Press, Oxford University Press, New York, 1987.
  • [12] 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.
  • [13] 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.
  • [14] J.-P. Serre. Local fields, volume 67 of Graduate Texts in Mathematics. Springer-Verlag, New York-Berlin, 1979. Translated from the French by Marvin Jay Greenberg.
  • [15] L. C. Washington. Introduction to cyclotomic fields, volume 83 of Graduate Texts in Mathematics. Springer-Verlag, New York, second edition, 1997.